Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Dissection of Splicing Regulation at an Endogenous Locus by Zinc-Finger Nuclease-Mediated Gene Editing

Abstract

Sequences governing RNA splicing are difficult to study in situ due to the great difficulty of traditional targeted mutagenesis. Zinc-finger nuclease (ZFN) technology allows for the rapid and efficient introduction of site-specific mutations into mammalian chromosomes. Using a ZFN pair along with a donor plasmid to manipulate the outcomes of DNA repair, we introduced several discrete, targeted mutations into the fourth intron of the endogenous BAX gene in Chinese hamster ovary cells. Putative lariat branch points, the polypyrimidine tract, and the splice acceptor site were targeted. We recovered numerous otherwise isogenic clones carrying the intended mutations and analyzed the effect of each on BAX pre-mRNA splicing. Mutation of one of three possible branch points, the polypyrimidine tract, and the splice acceptor site all caused exclusion of exon five from BAX mRNA. Interestingly, these exon-skipping mutations allowed usage of cryptic splice acceptor sites within intron four. These data demonstrate that ZFN-mediated gene editing is a highly effective tool for dissection of pre-mRNA splicing regulatory sequences in their endogenous context.

Introduction

The use of mutagenesis to reveal gene function is a classic technique in biology. The difficulty of achieving targeted mutagenesis in mammalian cells often has necessitated the use of extra-chromosomal or randomly-integrated reporter constructs as a proxy for endogenous gene function. While reporter-based experiments have contributed immensely to our understanding of the cell, loss of correct gene dosage, regulation, and chromatin structure can misrepresent the biology of the endogenous gene. In the case of RNA splicing, use of reporter genes can be unusually problematic as splicing is influenced or regulated by large-scale processes like chromatin modification [1], [2], [3], [4], [5], transcription [6], [7], , and mRNA export [13], [14]. Furthermore, the large size of mammalian genes often simply precludes the use of reporter systems to analyze splicing. Retrospective analysis of splicing in cells with naturally-occurring mutations has been informative but is not compatible with directed experimentation and lacks appropriate isogenic controls [15], [16], [17]. Given the centrality of alternative splicing to metazoan biology and disease, techniques that allow investigation of splicing regulation in its natural context are sorely needed.

A ZFN pair creates a targeted double-strand break in chromosomal DNA. When repaired inaccurately by the non-homologous end joining (NHEJ) DNA repair machinery, gene disruption results [18], [19], [20], [21]. Alternately, the homology-directed DNA repair (HDR) pathway can be manipulated to engineer mutations into endogenous genomic loci [22], [23], [24], [25]. In this application, a plasmid with chromosomal DNA sequence flanking the ZFN cleavage site and containing the desired mutation is co-delivered with the ZFNs. The cell can use this donor molecule as a template for DNA repair, resulting in copying of the mutated region into the chromosome (Figure 1A).

thumbnail
Figure 1. Alteration of BAX pre-mRNA splicing in response to mutagenesis of key cis regulatory splicing sequences.

A) Diagram of homology-directed repair in the BAX exon five region. Cleavage at the end of exon five stimulates strand invasion by the resected single-strand DNA. Once base pairing is established between the chromosome and the donor, new DNA synthesis and repair of the break using the newly synthesized DNA results in incorporation of the mutated sequence into the chromosome. Thick black lines, homology between the chromosome (top) and the donor plasmid (bottom); black arrow, strand invasion and new DNA synthesis; grey patch in donor plasmid and new DNA, mutation. Exon five and the donor sequence are drawn to scale. B) Putative splice site sequences found within the 3′ end of BAX intron four. Potential lariat branch points, the polypyrimidine tract, and the splice acceptor site in BAX intron four are indicated with black bars. The site of ZFN cleavage (shown in grey) is approximately 90 bp from the end of exon five. C) Mutations introduced into BAX intron four and their effect on BAX splicing. The location of specific base changes made in each isogenic cell line are shown in grey. For the branch point C mutation, the terminal C of the consensus sequence shown in part A was not altered. Arrows link the mutation with the gel lane containing RT-PCR products from a cell clone bearing the corresponding mutation. Lane 1; wild-type CHO-K1 cells; lane 2, branch point A mutation; lane 3, branch point B mutation; lane 4, branch point C mutation; lane 5, mutation of branch point C and the polypyrimidine tract (top); lane 6, mutation of the polypyrimidine tract (bottom); lane 7, splice acceptor mutation; lane 8, no template control. The identity and size of bands excised and confirmed by Sanger sequencing is shown to the right of the gel. High-molecular weight bands present in lanes 4–7 are heteroduplex material formed late in the PCR reaction by annealing of normal and exon five-skipped RT-PCR products. Molecular weight markers are in base pairs. The splicing pattern for each mutation was assayed between 4 and 12 times; representative data are shown. D) RT-PCR of the wild-type BAK gene in all eight samples serves as a loading control and is shown below the BAX RT-PCR.

https://doi.org/10.1371/journal.pone.0016961.g001

The BAX gene is haploid in CHO cells and its mutation by NHEJ can result in loss of a splice acceptor site resulting in skipping of exon five [26]. We reasoned that engineered mutagenesis of the endogenous gene via HDR would allow us to dissect the sequences required for normal BAX splicing. Here we demonstrate the utility of ZFNs to reveal splice site sequences. We introduce six mutations into intron four of the endogenous BAX gene and assay their effect on splicing by RT-PCR and high-throughput sequencing. Notably, our targeted mutagenesis identifies the branch point nucleotide for intron four lariat formation and uncovers cryptic splice acceptor regions within intron four, providing the first in situ functional annotation of splice site sequences in mammalian cells.

Methods

Donor plasmid design and construction

Donor plasmids were created with BAX homology regions on either side of the targeted mutations using fusion PCR. Each homology arm was synthesized separately by PCR amplification of BAX genomic DNA with a distal primer and a central primer containing the targeted mutation. For the initial round of PCR, the left homology arms were synthesized using GJC 47F (5′-cag aag ttc aag att ggc tta c-3′) and a mutation-specific reverse primer while right homology arms were created with a partially-complementary mutation-specific forward primer and GJC 114R (5′-tga acc agg ctg gga gat tt -3′). Annealing of the complementary ends containing the mutated regions during a second round of PCR created a 882 bp region of homology with a central intronic mutation. This second PCR was done using two primers inside GJC 47F and GJC 114R, namely GJC 250F (5′-cca gcc tgg ttt aca cag at-3′) and GJC 249R (5′-ggt tct gca taa gtg gtt ct-3′). Each resulting amplicon was cloned into the pCR2.1 vector (Invitrogen) and clones confirmed by DNA sequencing. All PCR amplifications were done with Accuprime Taq Hi Fidelity DNA Polymerase (Invitrogen) per the manufacturer's recommendations.

Mutation-specific oligonucleotides are shown below with mutations in bold. For the branch point A donor, the left homology arm was created by PCR of the BAX locus with GJC 47F and SC 12R #3 (5′-agt cct tcg aac act cct gga gaa ggt ggt-3′); the right homology arm by PCR with SC 13F #3 (5′-agt gtt cga agg act cag tcc tcc cgt gt-3′) and GJC 114R. The branch point B donor was made as above but using SC 12R (5′-ggg agg acc gag tcc tgt gaa cac tcc tgg-3′ and SC 13F (5′-gga ctc ggt cct ccc gtg taa cct tcc ca-3′). The branch point C donor was made as above but using SC 12R #2 (5′-agg cta gcc ggg agg act gag tcc tgt gaa-3′) and SC 13F #2 (5′-cct ccc ggc tag cct tcc cac tcc ctg cag-3′). The pY tract donor with AG dinucleotides was made as above but using SC 14R (5′-gct ctc ctt tct ctt ctt aca cgg gag gac tga g-3′) and SC 15F (5′- gaa gag aaa gga gag cag gcc ctg tgt acc aaa g -3′). The pY tract donor with no AG dinucleotides was made as above but with SC 14_2R (5′-tgc cat cat ttt cca tct tac acg gga gga ctg a-3′) and SC 15_2F (5′-gat gga aaa tga tgg cag gcc ctg tgt acc aaa g-3′). The splice acceptor donor was made as above but with SC 16R (5′- cag ggc ccg cag gga gtg gga agg tta-3and SC 17F (5′- tcc ctg cgg gcc ctg tgt acc aaa gtg-3′).

Transfection and cell cloning

CHO-K1 cells were obtained from the American Type Culture Collection (ATCC CCL-61) and grown and transfected with the FokI EL/KK heterodimer mutant BAX ZFNs as described previously (Figure S1) [26]. Twenty micrograms of donor was cotransfected with the ZFNs. Genomic DNA of the transfected pools was extracted from the aforementioned samples using the Masterpure kit (Epicentre). Each sample was PCR-amplified using GJC 47F and GJC 114R, oligonucleotides outside of the donor sequence. This PCR was used to analyze BAX ZFN action via the Cel I assay and to analyze the extent of targeted mutation by assay of restriction fragment length polymorphism (RFLP). For the Cel I assay, a nested PCR was performed with a primer closer to the ZFN sites, GJC 52F (5′- cag agg aat gaa agc aaa gg-3′), and GJC 114R. The percentage of modified cells was calculated as described previously [19], [26]. For the RFLP assay, each donor was designed so that the targeted mutation either deletes or creates a new restriction site relative to the unmodified sequence. The branch point A mutation creates a BstBI site; the branch point B mutation creates a Sau961 site; the branch point C mutation creates an NheI site; the two polyY tract mutations and the splice acceptor mutation delete a PstI site. Digestion with the corresponding diagnostic restriction enzyme indicated the extent of specific modification by HDR.

For cell cloning, each co-transfected pool was diluted to 0.4 cells/well in 96-well plates. After approximately 2 weeks of growth, samples from each occupied well were harvested and genomic DNA isolated using the Epicentre QuickExtract Solution. Each sample was PCR amplified with GJC 47F and GJC 114R and analyzed by the RFLP assay described above. Positive clones were expanded and then genotyped by DNA sequencing.

RNA analysis

Total RNA was prepared from each clone with the High Pure RNA Isolation Kit (Roche). Complementary DNA was synthesized using a polydT oligonucleotide to prime reverse transcription by the SuperScript III RT (Invitrogen). For analysis of BAX, the cDNA from each clone was PCR amplified using GJC 12F (in exon four, 5′-ctt ctt ccg tgt ggc agc tg-3′) and GJC 23R (in exon six, 5′-ccc gaa gta tga gag gag gcc atc -3′), and PCR analyzed by electrophoresis on 10% polyacrylamide gels. The identity of each BAX cDNA band annotated in Figure 1 was confirmed by Sanger DNA sequencing. For analysis of BAK, cDNA was PCR amplified using GJC164F (in exon 1, 5′-cgg gat ctt tgt ctt cag ac-3′) and GJC 25R (in exon 4, 5′-ctg gaa ctc tgt gtc gta tct ccg g-3′).

High-throughput DNA sequencing

Solexa/Illumina-based DNA sequencing was performed on a pool of RT-PCR products from the clone in lane six (the polypyrimidine tract mutation containing AG dinucleotides). cDNA from this clone was PCR amplified with a forward oligonucleotide in exon four (5′- aat gat acg gcg acc acc gag atc tac act ctt tcc cta cac gac gct ctt ccg atc tag caa act ggt gct caa g-3′) and a reverse oligonucleotide in exon 6 (5′-caa gca gaa gac ggc ata cga gct ctt ccg atc taa gta tga gag gag gcc atc-3′) where bold font indicates homology to BAX genomic DNA. The amplicon was purified first with the Qiaquick PCR Purification kit (Qiagen) then with the GeneJET Purification Protocol (Fermentas).

Results

The splicing machinery recognizes several key splice site sequences in pre-mRNA to effect intron removal. In brief, binding of U2AF to the polypyrimidine tract promotes formation of a 5′-2′ lariat linkage between the splice donor site and the branch point nucleotide. The resulting lariat molecule is excised by nucleophilic attack of the splice acceptor site by the upstream exon. Potential lariat branch points, polypyrimidine tracts, and splice acceptor sites in BAX intron four were identified computationally by the Sroogle algorithm and by visual inspection (Figure 1B) [27], [28], [29], [30]. Six donor plasmids were assembled, each containing 882 bp of flanking BAX genomic DNA along with six separate mutated regions designed to disrupt each of three potential lariat branch points, the polypyrimidine tract (two mutations), and the splice acceptor site, respectively. BAX ZFNs were transfected into CHO-K1 cells with each individual donor plasmid. Mutation at BAX three days post-transfection was 6.3+/−2.0% (n = 6) of alleles as measured by the Cel I assay [19]. The transfected pool was cloned by limiting dilution and BAX PCR products from each clone screened for the gain or loss of a restriction site generated by the intended mutation in BAX intron four. The genotype of positive clones was confirmed by DNA sequencing. Across all six transfections, correct clones were recovered at a frequency of 0.3+/−0.2% (n = 6; see Discussion).

RNA was prepared from wild-type CHO-K1 cells and from one representative clone for each of the six mutations. RT-PCR was performed using primers in exons four and six. Analysis of PCR products for each mutation revealed the contribution of each sequence to the correct splicing of exon five (Figure 1C). Firstly, mutation of putative branch points A and B did not affect splicing. Both mutants resulted in a single RT-PCR band indistinguishable from that in wild-type cells, demonstrating that genome editing per se does not alter RNA splicing. In contrast, loss of branch point C caused an approximate 50% loss of exon five from the BAX transcript (lane 4). Both mutations of the polypyrimidine tract resulted in very significant loss of correct splicing (∼90%, lanes 5,6). As expected, a single nucleotide change in the highly-conserved splice acceptor site completely prevented splicing of exon five (lane 7). The level of BAK mRNA (an unmodified, wild-type gene) was unaffected by any of the mutations introduced into BAX (Figure 1D).

Certain mutations caused the appearance of or an increase in of minor splice forms (lanes 4–7). To sample splice form diversity, we subjected RT-PCR products from lane six to high-throughput DNA sequencing using a forward primer placed at the very end of exon four. Five different transcripts containing varying lengths of the 3′ end of intron four were found: inclusion of 153, 50, 38, 142, and 31 bp, in order of approximate frequency. All such cryptic splice sites are immediately preceded by an YAG splice acceptor sequence. Clearly visible in lanes 4–7, the most abundant cryptic splice site (379 bp band) has a well-positioned YUNAY branch point consensus and polypyrimidine tract (cttac-ctcttccctgccttcc-ag). None of the mRNAs derived from cryptic splicing result in a 5′ extension of the normal BAX exon 5 open reading frame; all transcripts except the rarest 31 bp extension should be degraded by nonsense-mediated mRNA decay.

Taken together, these data demonstrate the successful creation of isogenic CHO cells carrying an allelic series of mutations implicated in the regulation of splicing at the endogenous BAX gene. Our data reveal the critical importance of just one of three potential branch points, the polypyrimidine tract and the splice acceptor sequence in the maintenance of normal exon five splicing. Moreover, these data reveal cryptic splice sites in intron four active only when the preferred natural site is mutated.

Discussion

Experiments done in vitro and in yeast have informed our knowledge of splicing biology enormously, yet understanding has been hampered in most organisms by a simple inability to perform targeted mutagenesis. Such a facility would be particularly useful for understanding the integration of splicing with chromatin remodeling and transcription and for studying the alternative splicing that accounts for a large part of biological diversity in higher eukaryotes. We performed ZFN-mediated targeted mutagenesis of the C. griseus BAX gene to dissect the importance of various splice site sequences in their normal, endogenous chromosomal context. Traditional recombination is also capable of creating site-specific mutations in endogenous genes but is laborious and time-consuming. In contrast, the mutations described here were created in two overlapping tranches, each of ∼10 weeks in duration, from donor creation to isolation of the final clones. Most of this period involved clone growth and expansion and therefore little investigator effort. The inefficiency of traditional homologous recombination necessitates co-integration of a selectable marker gene. The ability to obtain the desired mutations without the use of a selectable marker obviates concerns about the confounding impact of such an exogenous, independent transcription unit.

The conversion efficiency of double-strand breaks into correctly mutated alleles in this experiment was much lower than the amount of ZFN cleavage (6.3% cleavage, 0.3% conversion) and lower than the ∼33% of breaks typically converted into targeted integrations [25], (data not shown). This is likely the result of the ∼110 bp between the site of the ZFN-induced break and the site of the intended mutations. Because of this distance, the break can be repaired successfully without single-strand extension long enough to reach the mutated site in the donor plasmid. ZFNs can be targeted to essentially any DNA sequence. Use of a custom ZFN pair positioned more closely to the target area should substantially improve the relative frequency of mutated clones. Furthermore, only 3′ end invasion from the right side of the break will result in the introduction of the mutation. (Invasion from the left will copy only wild-type sequence from the donor plasmid.) Despite the modest overall efficiency of mutagenesis, we were still able to isolate the desired mutants in a single round of experimentation.

BAX is naturally haploid in CHO cells. Exon five is in frame with exon four and the terminal exon six; transcripts lacking exon five will therefore not be subject to nonsense-mediated mRNA decay and will provide an unbiased readout of BAX splicing. Interrogation of most other genes will require a similar choice of an in-frame exon and the generation of homozygous mutant cells. Homozygotes are typically formed at a approximately one-third the frequency of heterozygous mutants or can be made via a second round of ZFN mutagenesis [23]. Alternately, the complications of diploidy can be avoided by use of a near-haploid cell line such as the human leukemia line P1-55 [31].

Mutation of branch point consensus C resulted in an incompletely penetrant splicing phenotype. We suspect that an second lariat branch point was used to produce the correctly-spliced fraction of BAX mRNA, perhaps sites A or B. Similar usage of a back-up, cryptic branch point sequence has been seen in the case of the β-globin gene [32]. Alternately, perhaps the lariat is simply formed at the mutated site with lower efficiency [33]. Creation of a doubly-mutant allele will be needed to identify the back-up branch point conclusively.

Candidate branch point C was not identified in the initial screen of intron four for splicing-related sequences. As a result, the mutations made to the polypyrimidine tract overlap the branch point C consensus sequence by one nucleotide. The splicing defect seen in lanes five and six and might therefore reflect loss of both polypyrimidine tract and branch point C function. As loss of branch point C function alone (lane 4) results in only partial exclusion of exon five from BAX mRNA, the near total skipping of exon five seen with these potentially compound mutants indicates a significant contribution of the polypyrimidine tract to correct BAX splicing.

We made two mutations to the polypyrimidine tract. The second mutation attempted to alter splice acceptor site usage by inclusion of multiple AG dinucleotides (lane six). In this context, the lariat branch point sequence could be provided by the A or B sites and the polypyrimidine tract by the run of seven Ys between branch points B and C. (Seven pyrimidines are sufficient for U2AF binding [34].) We did not observe usage of these potential splice acceptor sites even by high-throughput DNA sequencing. As the primary purpose of this mutation was to disrupt the polypyrimidine tract, we did not include the pyrimidine typically present just 3′ of a strong splice acceptor site. This omission likely accounts for the failure of these AG dinucleotides to be used as cryptic splice sites. All five of the cryptic splice sites found by high-throughput DNA sequencing have a YAG splice acceptor site, whereas only the most abundant has a readily-identifiable lariat branch point and polypyrimidine tract.

Supporting Information

Figure S1.

The full nucleotide sequence of the BAX ZFN expression plasmid.

https://doi.org/10.1371/journal.pone.0016961.s001

(TXT)

Acknowledgments

We thank Ed Rebar, Jeff Miller, Dave Paschon, Lei Zhang, Sarah Hinkley, George Katibah, Gladys Dulay, Anna Vincent, Stephen Lam, and Rainier Amora for ZFN construction and cloning, and Michael Holmes, Christine Guthrie, and Alex Plocik for productive discussion.

Author Contributions

Conceived and designed the experiments: PDG FDU GJC. Performed the experiments: SC. Analyzed the data: SC FDU GJC. Wrote the manuscript: PDG FDU GJC.

References

  1. 1. Schwartz S, Ast G (2010) Chromatin density and splicing destiny: on the cross-talk between chromatin structure and splicing. Embo J 29: 1629–1636.
  2. 2. Schwartz S, Meshorer E, Ast G (2009) Chromatin organization marks exon-intron structure. Nat Struct Mol Biol 16: 990–995.
  3. 3. Tilgner H, Nikolaou C, Althammer S, Sammeth M, Beato M, et al. (2009) Nucleosome positioning as a determinant of exon recognition. Nat Struct Mol Biol 16: 996–1001.
  4. 4. Kolasinska-Zwierz P, Down T, Latorre I, Liu T, Liu XS, et al. (2009) Differential chromatin marking of introns and expressed exons by H3K36me3. Nat Genet 41: 376–381.
  5. 5. Luco RF, Pan Q, Tominaga K, Blencowe BJ, Pereira-Smith OM, et al. (2010) Regulation of alternative splicing by histone modifications. Science 327: 996–1000.
  6. 6. Maniatis T, Reed R (2002) An extensive network of coupling among gene expression machines. Nature 416: 499–506.
  7. 7. Proudfoot NJ, Furger A, Dye MJ (2002) Integrating mRNA processing with transcription. Cell 108: 501–512.
  8. 8. Neugebauer KM (2002) On the importance of being co-transcriptional. J Cell Sci 115: 3865–3871.
  9. 9. Bentley D (2002) The mRNA assembly line: transcription and processing machines in the same factory. Curr Opin Cell Biol 14: 336–342.
  10. 10. Kornblihtt AR, de la Mata M, Fededa JP, Munoz MJ, Nogues G (2004) Multiple links between transcription and splicing. Rna 10: 1489–1498.
  11. 11. Hicks MJ, Yang CR, Kotlajich MV, Hertel KJ (2006) Linking splicing to Pol II transcription stabilizes pre-mRNAs and influences splicing patterns. PLoS Biol 4: e147.
  12. 12. Das R, Yu J, Zhang Z, Gygi MP, Krainer AR, et al. (2007) SR proteins function in coupling RNAP II transcription to pre-mRNA splicing. Mol Cell 26: 867–881.
  13. 13. Luo MJ, Reed R (1999) Splicing is required for rapid and efficient mRNA export in metazoans. Proc Natl Acad Sci U S A 96: 14937–14942.
  14. 14. Reed R (2003) Coupling transcription, splicing and mRNA export. Curr Opin Cell Biol 15: 326–331.
  15. 15. Carothers AM, Urlaub G, Grunberger D, Chasin LA (1993) Splicing mutants and their second-site suppressors at the dihydrofolate reductase locus in Chinese hamster ovary cells. Mol Cell Biol 13: 5085–5098.
  16. 16. Chen IT, Chasin LA (1993) Direct selection for mutations affecting specific splice sites in a hamster dihydrofolate reductase minigene. Mol Cell Biol 13: 289–300.
  17. 17. O'Neill JP, Rogan PK, Cariello N, Nicklas JA (1998) Mutations that alter RNA splicing of the human HPRT gene: a review of the spectrum. Mutat Res 411: 179–214.
  18. 18. Santiago Y, Chan E, Liu PQ, Orlando S, Zhang L, et al. (2008) Targeted gene knockout in mammalian cells by using engineered zinc-finger nucleases. Proc Natl Acad Sci U S A 105: 5809–5814.
  19. 19. Perez EE, Wang J, Miller JC, Jouvenot Y, Kim KA, et al. (2008) Establishment of HIV-1 resistance in CD4+ T cells by genome editing using zinc-finger nucleases. Nat Biotechnol 26: 808–816.
  20. 20. Liu PQ, Chan EM, Cost GJ, Zhang L, Wang J, et al. (2009) Generation of a triple-gene knockout mammalian cell line using engineered zinc-finger nucleases. Biotechnol Bioeng 106: 97–105.
  21. 21. Malphettes L, Freyvert Y, Chang J, Liu P-Q, Chan E, et al. (2010) Highly-efficient deletion of FUT8 in CHO cell lines using zinc-finger nucleases yields cells which produce completely nonfucosylated antibodies. Biotechnology and Bioengineering 106: 774–783.
  22. 22. Urnov FD, Miller JC, Lee YL, Beausejour CM, Rock JM, et al. (2005) Highly efficient endogenous human gene correction using designed zinc-finger nucleases. Nature 435: 646–651.
  23. 23. Hockemeyer D, Soldner F, Beard C, Gao Q, Mitalipova M, et al. (2009) Efficient targeting of expressed and silent genes in human ESCs and iPSCs using zinc-finger nucleases. Nat Biotechnol 27: 851–857.
  24. 24. Dekelver RC, Choi VM, Moehle EA, Paschon DE, Hockemeyer D, et al. (2010) Functional genomics, proteomics, and regulatory DNA analysis in isogenic settings using zinc finger nuclease-driven transgenesis into a safe harbor locus in the human genome. Genome Res.
  25. 25. Orlando SJ, Santiago Y, Dekelver RC, Freyvert Y, Boydston EA, et al. (2010) Zinc-finger nuclease-driven targeted integration into mammalian genomes using donors with limited chromosomal homology. Nucleic Acids Res.
  26. 26. Cost GJ, Freyvert Y, Vafiadis A, Santiago Y, Miller JC, et al. (2009) BAK and BAX deletion using zinc-finger nucleases yields apoptosis-resistant CHO cells. Biotechnol Bioeng 105: 330–340.
  27. 27. Kol G, Lev-Maor G, Ast G (2005) Human-mouse comparative analysis reveals that branch-site plasticity contributes to splicing regulation. Hum Mol Genet 14: 1559–1568.
  28. 28. Gao K, Masuda A, Matsuura T, Ohno K (2008) Human branch point consensus sequence is yUnAy. Nucleic Acids Res 36: 2257–2267.
  29. 29. Schwartz SH, Silva J, Burstein D, Pupko T, Eyras E, et al. (2008) Large-scale comparative analysis of splicing signals and their corresponding splicing factors in eukaryotes. Genome Res 18: 88–103.
  30. 30. Schwartz S, Hall E, Ast G (2009) SROOGLE: webserver for integrative, user-friendly visualization of splicing signals. Nucleic Acids Res 37: W189–192.
  31. 31. Kotecki M, Reddy PS, Cochran BH (1999) Isolation and characterization of a near-haploid human cell line. Exp Cell Res 252: 273–280.
  32. 32. Ruskin B, Greene JM, Green MR (1985) Cryptic branch point activation allows accurate in vitro splicing of human beta-globin intron mutants. Cell 41: 833–844.
  33. 33. Hornig H, Aebi M, Weissmann C (1986) Effect of mutations at the lariat branch acceptor site on beta-globin pre-mRNA splicing in vitro. Nature 324: 589–591.
  34. 34. Sickmier EA, Frato KE, Shen H, Paranawithana SR, Green MR, et al. (2006) Structural basis for polypyrimidine tract recognition by the essential pre-mRNA splicing factor U2AF65. Mol Cell 23: 49–59.