untitled Reviews Nonviral Gene Delivery with the Sleeping Beauty Transposon System Zoltán Ivics 1–3 and Zsuzsanna Izsvák 1,2 Abstract Effective gene therapy requires robust delivery of therapeutic genes into relevant target cells, long-term gene expression, and minimal risks of secondary effects. Nonviral gene transfer approaches typically result in only short-lived transgene expression in primary cells, because of the lack of nuclear maintenance of the vector over several rounds of cell division. The development of efficient and safe nonviral vectors armed with an integrating feature would thus greatly facilitate clinical gene therapy studies. The latest generation transposon technology based on the Sleeping Beauty (SB) transposon may potentially overcome some of these limitations. SB was shown to provide efficient stable gene transfer and sustained transgene expression in primary cell types, including human hematopoietic progenitors, mesenchymal stem cells, muscle stem/progenitor cells (myoblasts), induced pluripotent stem cells, and T cells. These cells are relevant targets for stem cell biology, regenerative medicine, and gene- and cell-based therapies of complex genetic diseases. Moreover, the first-in-human clinical trial has been launched to use redirected T cells engineered with SB for gene therapy of B cell lymphoma. We discuss aspects of cellular delivery of the SB transposon system, transgene expression provided by integrated transposon vectors, target site selection of the transposon vectors, and potential risks associated with random genomic insertion. Introduction The ability to efficiently deliver foreign genes intocells provides the basis of using gene therapy to cure genetic diseases. The vast majority of gene delivery systems currently tested in clinical trials are based on viral vectors. Because viruses are highly specialized at crossing through cellular membranes by infection, they are efficient at deliv- ering nucleic acids to target cell populations. However, some viral vectors, including those derived from adenoviruses or adeno-associated viruses, remain largely episomal, requiring readministration in order to maintain a desired level of transgene expression over time. However, repeated delivery can provoke immune responses against vector-encoded pro- teins (reviewed in Hartman et al., 2008). In contrast, retro- viruses integrate their therapeutic cargo into the genome, resulting in long-term transgene expression. A concern with using retroviral vectors is genotoxicity associated with mu- tagenic effects elicited by insertion of the vector into or near genes (Hacein-Bey-Abina et al., 2003, 2008; Baum et al., 2004). Such risk is especially pronounced with gammaretroviral vectors based on the murine leukemia virus ( MLV) that preferentially integrate into transcription start sites (Wu et al., 2003). Furthermore, HIV-derived lentiviral vectors are po- tential mutagens due to their biased insertion into transcrip- tion units (Schroder et al., 2002; Bushman, 2003). Indeed, adverse events provoked by insertional mutagenesis of MLV- based vectors and resulting aberrant T cell proliferation have been observed in clinical trials for gene therapy of X-linked severe combined immunodeficiency (SCID-X1) (Hacein-Bey- Abina et al., 2003, 2008; Thrasher et al., 2006). Additional concerns include the possibility that transcriptional silencing may compromise expression of the integrated transgene (re- viewed in Ellis, 2005), and that large transgenes may inhibit viral reverse transcription and packaging, thereby setting limitations to vector design for clinical use of retroviral vec- tors. Even though the second-generation, self-inactivating (SIN) recombinant retroviral vectors that lack strong enhancer elements in their long terminal repeats (LTRs) (Schambach et al., 2006, 2007; Modlich et al., 2009) may be able to address some of the inadvertent side effects such as insertional onco- genesis, the high costs associated with the manufacture of 1Max Delbrück Center for Molecular Medicine, Berlin 13125, Germany. 2University of Debrecen, Debrecen 4032, Hungary. 3Paul Ehrlich Institute, Langen 63225, Germany. HUMAN GENE THERAPY 22:1043–1051 (September 2011) ª Mary Ann Liebert, Inc. DOI: 10.1089/hum.2011.143 1043 clinical-grade retroviral vectors as well as regulatory issues limit their widespread clinical translation. As a result, signif- icant efforts and developments have been made toward crafting gene transfer vector systems that are as efficient but safer than viruses in clinical gene transfer. Nonviral vector systems generally suffer from inefficient cellular delivery and limited duration of transgene expres- sion due to the lack of genomic insertion and resulting degradation and/or dilution of the vector in transfected cell populations. Developments of nonviral delivery techniques, including liposomal formulations, nanoparticles, advanced electroporation methods such as nucleofection, and cell- penetrating peptides can significantly enhance the transfer of nucleic acids into therapeutically relevant cell types, even in vivo. In postmitotic tissues, nonviral vectors may provide long-lasting transgene expression; however, in the absence of long-term nuclear maintenance in dividing cell types such as stem cells, even efficient introduction of nucleic acids into cells does not guarantee long-term transgene expression. One class of nonviral vector systems that could potentially offer long-term expression in dividing cell types is based on scaffold/matrix attachment region (S/MAR)-containing episomal vectors that can promote replication and mainte- nance in mammalian cells (Hagedorn et al., 2011). Yet an- other type of therapeutic vector would ideally unite the advantages of integrating viral vectors (i.e., long-lasting transgene expression) with those of nonviral delivery sys- tems (i.e., lower immunogenicity, enhanced safety profile and reduced costs of manufacture). Transposable elements (transposons) could potentially offer such an alternative. The Sleeping Beauty Transposon as a Nonviral Vector for Gene Therapy DNA transposons are discrete pieces of DNA with the ability to change their positions within the genome via a cut- and-paste mechanism called transposition. In nature, these elements exist as single units containing the transposase gene flanked by terminal inverted repeats (TIRs) that carry transposase-binding sites (Fig. 1A). However, under labora- tory conditions, it is possible to use transposons as bicom- ponent systems, in which virtually any DNA sequence of interest can be placed between the transposon TIRs and mobilized by trans-supplementing the transposase in the form of an expression plasmid (Fig. 1B) or mRNA synthe- sized in vitro. In the transposition process, the transposase enzyme mediates the excision of the element from its donor plasmid, followed by reintegration of the transposon into a chromosomal locus (Fig. 1C). This feature makes transposons natural and easily controllable DNA delivery vehicles that can be used as tools for versatile applications, including gene therapy. On the basis of transposon fossils that are presumed to have been active > 10 million years ago in fish genomes, an ancient transposon was resurrected, and named Sleeping Beauty (SB) after the famous fairy tale collected by the Brothers Grimm, because it was literally awakened after a long evolutionary ‘‘sleep’’ (Ivics et al., 1997). SB was the first transposon ever shown capable of efficient transposition in vertebrate cells, thereby enabling new avenues for genetic engineering in animal model species (reviewed in Ivics et al., 2009) as well as for human gene therapy (Yant et al., 2000; Ivics and Izsvak, 2006). As a nonviral alternative to viral vectors, the potential of the SB system has been thoroughly probed (Izsvák and Ivics, 2004; Fernando and Fletcher, 2006; Ivics and Izsvak, 2006; VandenDriessche et al., 2009; Hackett et al., 2010; Izsvák et al., 2010). The advantage of SB transposon-based gene delivery is that it combines the favorable features of viral vectors with those of naked DNA molecules. Namely, owing to perma- nent genomic insertion of transgene constructs (Fig. 1C), transposition-mediated gene delivery can lead to sustained and efficient transgene expression in preclinical animal models (Hackett et al., 2010). However, in contrast to viral FIG. 1. General organiza- tion and use of transposable elements as gene vectors. (A) Autonomous transpos- able elements consist of ter- minal inverted repeats (TIRs; black arrows) that flank the transposase gene. (B) Bicomponent transposon vector system for delivering transgenes that are main- tained in plasmids. One component contains a DNA of interest between the transposon TIRs carried by a plasmid vector, whereas the other component is a trans- posase expression plasmid, in which the black arrow represents the promoter driving expression of the transposase. (C) The trans- poson carrying a DNA of interest is excised from the donor plasmid and is integrated at a chromosomal site by the transposase. Color images available online at www.liebertonline.com/hum 1044 IVICS AND IZSVÁK vectors, transposon vectors can be maintained and propa- gated as plasmid DNA (Fig. 1B and C), which makes them simple and inexpensive to manufacture, an important con- sideration for the implementation of future clinical trials. Further advantages of the SB system include its reduced immunogenicity (Yant et al., 2000), no strict limitation of the size of expression cassettes (Zayed et al., 2004), and improved safety and toxicity profiles (Ivics et al., 2007; Moldt et al., 2007; Walisko et al., 2008; VandenDriessche et al., 2009). Because transposition proceeds through a cut-and-paste mechanism that involves only DNA, transposon vectors are not prone to incorporating mutations by reverse transcrip- tion (that are generated in retroviral stocks at reasonable frequencies), and can tolerate larger and more complex transgenes. Unlike the LTRs of retroviruses, the TIRs of SB vectors have low enhancer/promoter activity ( Moldt et al., 2007; Walisko et al., 2008). The insertion of chromatin boundary elements (insulators) flanking the transposon- contained expression cassettes (Fig. 2A) to prevent accidental trans-activation of cellular promoters further improved the safety profile of the SB system (Walisko et al., 2008). Further- more, synthetically produced mRNA can serve as a source of the transposase, thereby limiting the duration of transposase expression and lowering the risk of ‘‘rehopping’’ of the already integrated transposon-based vector (Wilber et al., 2006). Chromosomal integration of SB transposons is random, which is potentially mutagenic (see below), but no SB-associated adverse effects have been observed in preclinical animal studies (Fernando and Fletcher, 2006; Ivics and Izsvak, 2006; Hackett et al., 2010; Izsvák et al., 2010). Last, SB transposons can be harnessed to integrate plasmid-based small hairpin RNA (shRNA) expression cassettes into chromosomes to ob- tain stable knockdown cell lines by RNA interference (Fig. 2B) (Kaufman et al., 2005). Such technologies have been evaluated as a potential approach to the therapy of acquired immuno- deficiency syndrome by stable RNA interference with SB vectors knocking down the CCR5 and CXCR4 cell surface coreceptors that are required for viral entry as a first step to confer resistance to HIV (Tamhane and Akkina, 2008). Delivery of the Sleeping Beauty vector system In ex vivo gene delivery, the therapeutic gene vector is introduced into a selected cell population that had been isolated from the patient, and the treated cells are trans- planted back into the same patient. Because transposons are not infectious, it is necessary to combine the plasmid-based transposon vectors with technologies capable of efficient delivery of these nonviral vectors into cells. Because the ef- ficiency of transposition is dependent on the efficiency of uptake of the introduced plasmids into the cell nuclei, de- livery is a rate-limiting factor in transposition, and is thus of paramount importance. In principle, any technology devel- oped for transferring nucleic acids into cells can be combined with transposon vectors. Unfortunately, there is no generally applicable method, and procedures must be established for each cell type. In difficult-to-transfect cells, including pri- mary stem cells, delivery of transposon-based vectors can be significantly facilitated by nucleofection, a procedure based on electroporation that transfers nucleic acids directly into the nucleus. Indeed, nucleofection facilitated transposition in various stem cells including CD34 + hematopoietic progeni- tors (Hollis et al., 2006; Izsvák et al., 2009; Mates et al., 2009; Sumiyoshi et al., 2009; Xue et al., 2009), primary T cells (Huang et al., 2008, 2009; Singh et al., 2008) and human em- bryonic stem cells (Wilber et al., 2007; Orban et al., 2009). Importantly, in the context of the hematopoietic system, this ex vivo gene delivery procedure apparently did not com- promise the potential of transposon-marked CD34 + cells to differentiate normally into the erythroid, megakaryocytic, granulocyte/monocyte/macrophage lineage ( Mates et al., 2009) as well as into the CD4 + CD8 + T, CD19 + B, CD56 + CD3 - natural killer (NK), and CD33 + myeloid lineages (Xue et al., 2009). The robustness and feasibility of this nonviral, transposon-based procedure may significantly facilitate clinical realization of ex vivo stem cell therapy for the treat- ment of hematopoietic disorders and cancer, which has led to its application in humans (Williams, 2008). In in vivo gene delivery, the therapeutic gene vector is directly introduced into an organ, where expression of the therapeutic gene construct is desired. Complexing naked DNA with polyethylenimine (PEI) followed by intravenous injection is one of the most effective methods to deliver therapeutic genes to the lung (Belur et al., 2007). This ap- proach resulted in long-term expression of luciferase after delivery of SB transposon vectors to the lungs of mice (Belur et al., 2003, 2007). Complexing of transposon vectors with PEI allowed expression of therapeutic levels (10% of normal) of the blood coagulation factor VIII (FVIII) as well as pheno- typic correction of the bleeding disorder in a mouse model of hemophilia A (L. Liu et al., 2006b). In a rat model of pul- monary hypertension, systemic administration of an SB vector harboring an endothelial nitric oxide synthase (eNOS) gene resulted in inhibition of induced pulmonary FIG. 2. Expression cassettes de- livered by Sleeping Beauty trans- poson vectors. (A) A typical therapeutic expression cassette contains a ubiquitous or tissue- specific enhancer/PolII promoter that drives expression of a thera- peutic gene. To enhance the safety of such a vector, the expression cassette might be flanked by in- sulator elements that will block trans-activation of endogenous promoters by the transposon insertion, and simultaneously protect the expression of the therapeutic gene from position effects. (B) Knockdown expression cassette including a PolII promoter that drives expression of a marker gene and a PolIII promoter that drives expression of a short hairpin RNA (shRNA). Color images available online at www.liebertonline.com/hum NONVIRAL GENE DELIVERY WITH SB 1045 hypertension (L. Liu et al., 2006a). Furthermore, PEI-based systemic delivery of the SB transposon system encoding the human indoleamine-2,3-dioxygenase (hIDO) gene was eval- uated in the context of lung transplantation-associated chronic complications, and found to elicit a remarkable therapeutic response, as evidenced by near normal pulmo- nary function in lung allografts (H. Liu et al., 2006). Last, SB- mediated gene transfer significantly increased survival of mice bearing human glioblastoma xenografts by expressing antiangiogenic gene products (Ohlfest et al., 2005a), and improved the efficacy of immunotherapy by facilitating sustained cytokine expression after local, intratumoral in- jections of vector–PEI complexes (Wu et al., 2007). As an alternative to complexing with a chemical com- pound, the use of physical delivery systems, such as hy- drodynamic injection, can facilitate cellular uptake of naked DNA molecules (Liu et al., 1999; Zhang et al., 1999). This procedure involves rapid injection of a large volume of DNA solution through the tail vein. Delivery of therapeutic SB transposon constructs by hydrodynamic injection (Bell et al., 2007) was first demonstrated to be applicable to mediate stable, long-term expression in mouse liver by Yant and colleagues, who reported sustained expression of a1- antitrypsin in C57BL/6 mice (Yant et al., 2000). This proce- dure has been successfully applied to confer a therapeutic benefit in several animal models of human diseases, in- cluding hemophilia B after delivery of human clotting factor IX (FIX) as a therapeutic gene product in FIX-deficient mice (Yant et al., 2000) and tyrosinemia after hydrodynamic de- livery of a fumarylacetoacetate hydrolase (FAH)-encoding SB transposon vector into the livers of FAH-deficient mice that showed selective outgrowth of genetically corrected hepatocytes ( Montini et al., 2002). Additional, successful preclinical testing of the SB system has been established in disease models for hemophilia A (Ohlfest et al., 2005b) and mucopolysaccharidosis (Aronovich et al., 2007, 2009). Cell type-specific expression of the transgene is often desirable in order to exclude or limit ectopic transgene ex- pression. One way to achieve this is to equip the transpo- son-contained expression cassette with tissue-specific promoters; this was shown to confer erythroid-specific ex- pression of b-globin in a sickle cell disease model (Zhu et al., 2007). Another strategy for targeted expression of therapeutic gene constructs is their targeted delivery into tissues of interest. Hyaluronan- and asialoorosomucoid- coated nanocapsules, were found to target SB-based vectors carrying an FVIII transgene to hepatocytes in vivo and to improve the phenotype of hemophilia A mice, after intra- venous injection (Kren et al., 2009). Similarly, packaging an SB vector expressing chUGT1A1 into proteoliposomes in- corporating a fusogenic glycoprotein that promoted asia- loglycoprotein receptor (ASGPR)-mediated endocytosis allowed successful in vivo delivery and sustainable, thera- peutic gene expression in the hepatocytes of a rat model of Crigler-Najjar syndrome type 1 (Wang et al., 2009). Im- portantly, in neither study was a significant host immune response toward the gene delivery vehicle or the transgene product observed (Kren et al., 2009; Wang et al., 2009). In summary, when combined with nonviral delivery ap- proaches, the SB transposon system shows considerable efficacy in providing sustained levels of therapeutic gene expression both ex vivo and in vivo. Target site selection of Sleeping Beauty and its experimental manipulation The insertion pattern of most transposons is nonrandom, showing characteristic preferences for insertion sites at the primary DNA sequence level, and ‘‘hotspots’’ and ‘‘cold re- gions’’ on a genome-wide scale. SB displays considerable specificity in target site selection at the primary DNA se- quence level in that TA dinucleotides are obligate target sites (Vigdal et al., 2002). A palindromic AT-repeat consensus se- quence with bendability and hydrogen-bonding potential was found to constitute the preferred target site (Vigdal et al., 2002). It was later shown that a characteristic deformation of the DNA sequence may be a recognition signal for target selection (Liu et al., 2005). This deformation, and the likeli- hood that a particular TA will be targeted by SB, can be computationally predicted (Liu et al., 2005), which may allow a theoretical assessment of risks associated with transposon insertions in particular genomic regions (Geurts et al., 2006). Nevertheless, a systematic assessment of potential genotoxic effects associated with genomic integration of transposon vectors will need to be performed either in cell-based assays and/or in animal models to provide clinically relevant data. In contrast to the considerable specificity at the primary DNA sequence level, SB integration can be considered fairly random at the genomic level (Vigdal et al., 2002; Yant et al., 2005; Moldt et al., 2011). Roughly one-third of SB insertions in mouse and human cells occur in transcribed regions, and because genes cover about one-third of the genome, such frequency suggests neither preference for nor disfavoring insertion into genes. The vast majority of those insertions that occur in genes are located in introns (Vigdal et al., 2002), and, in contrast to most integrating viral vectors, the tran- scriptional status of targeted genes apparently does not in- fluence the integration profile of SB (Yant et al., 2005). Target site selection properties suggest that SB might be safer for therapeutic gene delivery than the integrating viral vectors that are currently used in clinical trials. However, with any vector that integrates into chromosomes in a nearly random manner (theoretically, the SB transposon could in- sert into any of the *108 TA sites in the human genome) comes the potential risk of insertional mutagenesis leading to transcriptional activation or inactivation of cellular genes (Baum et al., 2004). Integration of the vector into a gene or its regulatory elements can knock out the gene, overexpress the gene, or alter its spatio/temporal expression pattern. Such genotoxic effects can have devastating consequences for the cell and the whole organism, including the development of cancer, as discussed previously (Baum et al., 2004). As a possible strategy, introducing an imposed bias into the in- sertion profile, ideally, targeted integration of the therapeu- tic gene into a ‘‘safe’’ site in the human genome, could lower or eliminate possible hazards to the host cell. For tar- geted transposon insertion at least one component of the transposon system, either the transposon vector DNA or the transposase (or factors interacting with either of these com- ponents), needs to be engineered to be physically linked or interact with a heterologous DNA-binding domain (DBD), which is to tether the transposase/transposon complex to defined sites in the human genome, and to facilitate inte- gration of the transposon into adjacent DNA (Voigt et al., 2008). Fusions of the SB transposase with the GAL4 DBD 1046 IVICS AND IZSVÁK showed an enrichment of transposon insertions in a *400-bp window around GAL4-binding sites in plasmid targets in human cells (Yant et al., 2007). Target-selected SB insertion was also assessed by employing a molecular strategy based on engineering a LexA operator sequence into an SB trans- poson vector. Targeted transposition events into chromo- somal S/MAR sequences as well as a chromosomally integrated tetracycline response element were recovered by coexpressing targeting fusion proteins containing LexA and either the SAF box, a protein domain that binds to S/MARs, or the tetracycline repressor (TetR) (Ivics et al., 2007). Pre- sumably, these targeting fusion proteins ‘‘sandwiched’’ the transposon and target DNA sites, allowing local insertion events. Last, targeted SB transposition within a 2.5-kb win- dow around a chromosomally located tetracycline response element was observed at a frequency of > 10% (Ivics et al., 2007) by coexpressing the SB transposase with a targeting fusion protein consisting of TetR and a subdomain of the SB transposase spanning the N-terminal helix–turn–helix do- main (N57) that mediates protein–protein interactions be- tween transposase subunits (Izsvák et al., 2002). Ongoing work in the authors’ laboratory is dedicated to determining whether SB transposition can potentially be directed to physiologically relevant, endogenous sites in the human genome, by using both naturally occurring as well as syn- thetic DBDs. Transgene expression Any transgene vector system should ideally provide long- term expression of transgenes. By using classical, plasmid- based, nonviral delivery approaches, expression from the extrachromosomal plasmid rapidly declines after delivery. Transgenes delivered by nonviral approaches often form long, repeated arrays (concatemers) that are targets for transcriptional silencing by heterochromatin formation. In addition, long-term expression of transgenes delivered by retroviruses has been shown to be compromised by tran- scriptional silencing ( Jahner et al., 1982). It was shown that the zinc finger protein ZFP809 bridges the integrated pro- viral DNA of the murine leukemia virus and the tripartite motif-containing 28 transcriptional corepressor in embryonic stem cells (Wolf and Goff, 2009). Thus, sequence elements in the vector itself can predispose the cargo for silencing. The cut-and-paste mechanism of DNA transposition results in a single copy of the transgene per insertion locus, and thus concatemer-induced gene silencing is unlikely to be an issue with transposition-mediated gene transfer. Indeed, Gra- bundzija and colleagues found that transposon insertions delivered by the SB system only rarely (< 2% of all insertions) undergo silencing in HeLa cells (Grabundzija et al., 2010). Furthermore, stable transgene expression observed in *300 independent insertions in this study suggests that SB rarely targets heterochromatic chromosomal regions for in- sertion, and that it is unlikely that certain sequence motifs in the transposon vector are recognized by mediators of si- lencing in the cell. An additional factor that may provoke transgene silencing is the cargo DNA, particularly the type of promoter used to drive expression of the gene of interest. Indeed, it was previously shown that transgene constructs delivered into mouse cells by SB transposition can be subject to epigenetic regulation by CpG methylation and that a de- terminant of epigenetic modifications of the integrating transposon vector is the cargo transgene construct, with the promoter playing a major role (Garrison et al., 2007). How- ever, with careful promoter choice, several studies have es- tablished that SB-mediated transposition provides long-term expression in vivo, as discussed previously. Notably, stable transgene expression from SB vectors was seen in mice after gene delivery in the liver (Yant et al., 2000; Ohlfest et al., 2005b; Aronovich et al., 2009; Kren et al., 2009), lung (Belur et al., 2003; L. Liu et al., 2006b), brain (Ohlfest et al., 2005a), and blood after hematopoietic reconstitution in vivo ( Mates et al., 2009; Xue et al., 2009). Thus, although our under- standing of all the factors that will ultimately determine the expressional fate of an integrated transposon is still rudi- mentary, it appears that transposon vectors have the capacity to provide long-term expression of transgenes both in vitro and in vivo. Gene transfer into stem cells by the SB100X hyperactive transposase The hyperactive variant of the SB transposase, SB100X, yields efficient stable gene transfer after nonviral gene de- livery into therapeutically relevant primary cell types, in- cluding stem or progenitor cells. For example, the use of the SB100X system yielded robust gene transfer efficiencies into human hematopoietic progenitors ( Mates et al., 2009; Xue et al., 2009), mesenchymal stem cells, muscle stem/progeni- tor cells (myoblasts), induced pluripotent stem cells (iPSCs) (Belay et al., 2010), and T cells ( Jin et al., 2011). These cells are relevant targets for stem cell biology and for regenerative medicine and gene- and cell-based therapies of complex ge- netic diseases. Importantly, expression of the SB100X hy- peractive transposase did not adversely influence the differentiation or function of these adult stem/progenitor cells, nor was there any evidence of any cytogenetic abnor- malities (Belay et al., 2010). In the context of iPSC technology, the ability to coax the differentiation of pluripotent stem cells into clinically relevant, transplantable cell types is a key step toward their ultimate use in clinical applications, especially because undifferentiated iPSCs pose an intrinsic tumori- genic risk (Yamanaka, 2009). It was demonstrated that SB transposon-mediated delivery of the myogenic PAX3 tran- scription factor into iPSCs coaxed their differentiation into MyoD + myogenic progenitors and multinucleated myofibers (Belay et al., 2010), suggesting that PAX3 may serve as a myogenic ‘‘molecular switch’’ in iPSCs, a finding that has implications for cell therapy of congenital degenerative muscle diseases, including Duchenne muscular dystrophy. Thus, the SB100X hyperactive transposase holds great promise for ex vivo and in vivo gene therapies. Future Considerations/Outlook There has been steadily growing interest in applying the SB system for gene therapy (Izsvák and Ivics, 2004; Essner et al., 2005; Hackett et al., 2005, 2010; Ivics and Izsvak, 2006; Izsvák et al., 2010) in the context of several conditions in- cluding hemophilia A and B (Yant et al., 2000; Ohlfest et al., 2005b; L. Liu et al., 2006b; Kren et al., 2009; Hausl et al., 2010), junctional epidermolysis bullosa (Ortiz-Urda et al., 2002), tyrosinemia I ( Montini et al., 2002), Huntington disease (Chen et al., 2005), sickle cell disease (Zhu et al., 2007), NONVIRAL GENE DELIVERY WITH SB 1047 mucopolysaccharidosis (Aronovich et al., 2007, 2009), cancer (Ohlfest et al., 2005a; Peng et al., 2009; Jin et al., 2011), and type 1 diabetes (He et al., 2004). In addition, important steps have been made toward SB-mediated gene transfer in the lung for potential therapy of a1-antitrypsin deficiency, cystic fibrosis, and a variety of cardiovascular diseases (Belur et al., 2003; Liu et al., 2004). The first clinical application of the SB system is currently ongoing, using autologous T cells gene-modified with SB vectors (Williams, 2008) carrying a chimeric antigen receptor (CAR) to render the T cells cytotoxic specifically toward CD19-positive lymphoid tumors (Huang et al., 2008; Singh et al., 2008). Lymphocytes represent a suitable initial platform for testing new gene transfer systems, as T cells can be ge- netically modified by viral and nonviral approaches without apparent resulting genotoxicity (Bonini et al., 2003). Ad- vantages of using the SB system for genetic modification of T cells include the ease and reduced cost associated with the manufacturing of clinical-grade, plasmid-based vectors compared with recombinant viral vectors. Sleeping Beauty faces a number of challenges before widespread clinical translation. For in vivo delivery into the liver, the hydrodynamic technology needs to be calibrated first to large-animal models and then eventually to humans. This is not a simple task to solve, but important steps have been made toward this goal by developing computer- assisted protocols (Suda et al., 2008). Another complicating issue with DNA delivery to the liver is the resulting tissue damage and the associated (at least transiently) elevated values of liver enzymes that may compromise clinical ap- plication of these procedures. A major hurdle in ex vivo de- livery of the transposon components into relevant primary cell types is the toxicity of the transfection/electroporation protocols that is typically observed. In situations in which target cells are scarce and/or culturing and expansion of the transfected cells are impossible or cannot be solved without compromising cell identity and grafting potential, cytotox- icity of the transfection procedures is a serious issue that may undermine clinical applications. Development of nonviral chemical reagents promoting plasmid DNA delivery with reduced toxicity may provide a solution. Alternatively, the development of hybrid vector systems combining the natural ability of viruses to traverse cell membranes with efficient genomic insertion mediated by the SB system is a promising strategy. Indeed, components of the SB transposon have been incorporated into integrase-defective lentiviral particles that showed efficient gene transfer in a range of human cell types and an insertion profile favorable to conventional lentiviral vectors (Staunstrup et al., 2009; Vink et al., 2009; Moldt et al., 2011). Hybrid adenovirus–SB vectors (Yant et al., 2002) have been used to efficiently deliver SB transposon vectors expressing FIX into the liver in a hemophilic dog model (Hausl et al., 2010). Retroviral vectors disabled in generating a cDNA copy of the retroviral vector have been shown to deliver the SB transposase mRNA into target cells with impressive efficiency (Galla et al., 2011). Last, herpes simplex virus vectors with a tropism toward infecting neural progenitor cells have been used to target SB insertions in the central nervous system in an in utero gene delivery system in the mouse (Bowers et al., 2006; de Silva et al., 2010). Fur- thermore, not only the procedures that are used to deliver the transposon vector components into cells, but those components themselves, may elicit unwanted effects in the cell. For example, it was found that overexpression of the SB transposase can have cytotoxic effects (Galla et al., 2011). The molecular mechanism of transposase cytotoxicity is currently not understood, but it is unlikely to be due to uncontrolled cleavage of genomic DNA by the transposase (Galla et al., 2011). At any rate, careful dosing of the transposase as well as the transposon donor plasmids in gene delivery experi- ments appears to be of fundamental importance; luckily, with plasmid-based vectors such as the SB transposon sys- tem, this can easily be achieved. Last, potential genotoxic effects elicited by transcriptional upregulation of proto- oncogenes and other signaling genes on random transposon insertion is a relatively unexplored area of research. In- vestigations into these questions will be required to docu- ment the safety of the SB system for prospective clinical trials. The next phase of research will undoubtedly focus on these issues to introduce the SB transposon system as a nonviral gene delivery tool for gene therapy. Acknowledgments Work in the authors’ laboratories was supported by EU FP6 (INTHER) and EU FP7 (PERSIST and InduStem), grants from the Deutsche Forschungsgemeinschaft SPP1230 ‘‘Me- chanisms of Gene Vector Entry and Persistence,’’ and from the Bundesministerium für Bildung und Forschung (NGFN- 2, NGFNplus, iGene, InTherGD, and ReGene). Author Disclosure Statement No competing financial interests exist. References Aronovich, E.L., Bell, J.B., Belur, L.R., et al. (2007). Prolonged expression of a lysosomal enzyme in mouse liver after Sleeping Beauty transposon-mediated gene delivery: implications for non-viral gene therapy of mucopolysaccharidoses. J. Gene Med. 9, 403–415. Aronovich, E.L., Bell, J.B., Khan, S.A., et al. (2009). Systemic correction of storage disease in MPS I NOD/SCID mice using the Sleeping Beauty transposon system. Mol. Ther. 17, 1136–1144. Baum, C., Von Kalle, C., Staal, F.J., et al. (2004). Chance or ne- cessity? Insertional mutagenesis in gene therapy and its con- sequences. Mol. Ther. 9, 5–13. Belay, E., Matrai, J., Acosta-Sanchez, A., et al. (2010). Novel hy- peractive transposons for genetic modification of induced pluripotent and adult stem cells: A nonviral paradigm for coaxed differentiation. Stem Cells 28, 1760–1771. Bell, J.B., Podetz-Pedersen, K.M., Aronovich, E.L., et al. (2007). Preferential delivery of the Sleeping Beauty transposon system to livers of mice by hydrodynamic injection. Nat. Protoc. 2, 3153–3165. Belur, L.R., Frandsen, J.L., Dupuy, A.J., et al. (2003). Gene in- sertion and long-term expression in lung mediated by the Sleeping Beauty transposon system. Mol. Ther. 8, 501–507. Belur, L.R., Podetz-Pedersen, K., Frandsen, J., and McIvor, R.S. (2007). Lung-directed gene therapy in mice using the nonviral Sleeping Beauty transposon system. Nat. Protoc. 2, 3146–3152. Bonini, C., Grez, M., Traversari, C., et al. (2003). Safety of ret- roviral gene marking with a truncated NGF receptor. Nat. Med. 9, 367–369. 1048 IVICS AND IZSVÁK Bowers, W.J., Mastrangelo, M.A., Howard, D.F., et al. (2006). Neuronal precursor-restricted transduction via in utero CNS gene delivery of a novel bipartite HSV amplicon/transposase hybrid vector. Mol. Ther. 13, 580–588. Bushman, F.D. (2003). Targeting survival: integration site selec- tion by retroviruses and LTR-retrotransposons. Cell 115, 135–138. Chen, Z.J., Kren, B.T., Wong, P.Y., et al. (2005). Sleeping Beauty- mediated down-regulation of huntingtin expression by RNA interference. Biochem. Biophys. Res. Commun. 329, 646–652. de Silva, S., Mastrangelo, M.A., Lotta, L.T., Jr., et al. (2010). Herpes simplex virus/Sleeping Beauty vector-based embryonic gene transfer using the HSB5 mutant: Loss of appar- ent transposition hyperactivity in vivo. Hum. Gene Ther. 21, 1603–1613. Ellis, J. (2005). Silencing and variegation of gammaretrovirus and lentivirus vectors. Hum. Gene Ther. 16, 1241–1246. Essner, J.J., McIvor, R.S., and Hackett, P.B. (2005). Awakening gene therapy with Sleeping Beauty transposons. Curr. Opin. Pharmacol. 5, 513–519. Fernando, S., and Fletcher, B.S. (2006). Sleeping Beauty transposon- mediated nonviral gene therapy. BioDrugs 20, 219–229. Galla, M., Schambach, A., Falk, C.S., et al. (2011). Avoiding cy- totoxicity of transposases by dose-controlled mRNA delivery. Nucleic Acids Res. 39, 7147–7160. Garrison, B.S., Yant, S.R., Mikkelsen, J.G., and Kay, M.A. (2007). Postintegrative gene silencing within the Sleeping Beauty transposition system. Mol. Cell. Biol. 27, 8824–8833. Geurts, A.M., Hackett, C.S., Bell, J.B., et al. (2006). Structure- based prediction of insertion-site preferences of transposons into chromosomes. Nucleic Acids Res. 34, 2803–2811. Grabundzija, I., Irgang, M., Mates, L., et al. (2010). Comparative analysis of transposable element vector systems in human cells. Mol. Ther. 18, 1200–1209. Hacein-Bey-Abina, S., Von Kalle, C., Schmidt, M., et al. (2003). LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science 302, 415–419. Hacein-Bey-Abina, S., Garrigue, A., Wang, G.P., et al. (2008). Insertional oncogenesis in 4 patients after retrovirus-mediated gene therapy of SCID-X1. J. Clin. Invest. 118, 3132–3142. Hackett, P.B., Ekker, S.C., Largaespada, D.A., and McIvor, R.S. (2005). Sleeping Beauty transposon-mediated gene therapy for prolonged expression. Adv. Genet. 54, 189–232. Hackett, P.B., Largaespada, D.A., and Cooper, L.J. (2010). A transposon and transposase system for human application. Mol. Ther. 18, 674–683. Hagedorn, C., Wong, S.P., Harbottle, R., and Lipps, H.J. (2011). Scaffold/matrix attached region-based nonviral episomal vectors. Hum. Gene Ther. 22, 915–923. Hartman, Z.C., Appledorn, D.M., and Amalfitano, A. (2008). Adenovirus vector induced innate immune responses: impact upon efficacy and toxicity in gene therapy and vaccine ap- plications. Virus Res. 132, 1–14. Hausl, M.A., Zhang, W., Muther, N., et al. (2010). Hyperactive Sleeping Beauty transposase enables persistent phenotypic correction in mice and a canine model for hemophilia B. Mol. Ther. 18, 1896–1906. He, C.X., Shi, D., Wu, W.J., et al. (2004). Insulin expression in livers of diabetic mice mediated by hydrodynamics-based administration. World J. Gastroenterol. 10, 567–572. Hollis, R.P., Nightingale, S.J., Wang, X., et al. (2006). Stable gene transfer to human CD34 + hematopoietic cells using the Sleeping Beauty transposon. Exp. Hematol. 34, 1333–1343. Huang, X., Guo, H., Kang, J., et al. (2008). Sleeping Beauty transposon-mediated engineering of human primary T cells for therapy of CD19 + lymphoid malignancies. Mol. Ther. 16, 580–589. Huang, X., Wilber, A., McIvor, R.S., and Zhou, X. (2009). DNA transposons for modification of human primary T lympho- cytes. Methods Mol. Biol. 506, 115–126. Ivics, Z., and Izsvak, Z. (2006). Transposons for gene therapy! Curr. Gene Ther. 6, 593–607. Ivics, Z., Hackett, P.B., Plasterk, R.H., and Izsvak, Z. (1997). Molecular reconstruction of Sleeping Beauty, a Tc1-like trans- poson from fish, and its transposition in human cells. Cell 91, 501–510. Ivics, Z., Katzer, A., Stuwe, E.E., et al. (2007). Targeted Sleeping Beauty transposition in human cells. Mol. Ther. 15, 1137–1144. Ivics, Z., Li, M.A., Mates, L., et al. (2009). Transposon- mediated genome manipulation in vertebrates. Nat. Methods 6, 415–422. Izsvák, Z., and Ivics, Z. (2004). Sleeping Beauty transposition: Biology and applications for molecular therapy. Mol. Ther. 9, 147–156. Izsvák, Z., Khare, D., Behlke, J., et al. (2002). Involvement of a bifunctional, paired-like DNA-binding domain and a trans- positional enhancer in Sleeping Beauty transposition. J. Biol. Chem. 277, 34581–34588. Izsvák, Z., Chuah, M.K., VandenDriessche, T., and Ivics, Z. (2009). Efficient stable gene transfer into human cells by the Sleeping Beauty transposon vectors. Methods 49, 287–297. Izsvák, Z., Hackett, P.B., Cooper, L.J., and Ivics, Z. (2010). Translating Sleeping Beauty transposition into cellular thera- pies: Victories and challenges. Bioessays 32, 756–767. Jahner, D., Stuhlmann, H., Stewart, C.L., et al. (1982). De novo methylation and expression of retroviral genomes during mouse embryogenesis. Nature 298, 623–628. Jin, Z., Maiti, S., Huls, H., et al. (2011). The hyperactive Sleeping Beauty transposase SB100X improves the genetic modification of T cells to express a chimeric antigen receptor. Gene Ther. [Epub ahead of print] Kaufman, C.D., Izsvak, Z., Katzer, A., and Ivics, Z. (2005). Frog Prince transposon-based RNAi vectors mediate efficient gene knockdown in human cells. J. RNAi Gene Silenc. 1, 97–104. Kren, B.T., Unger, G.M., Sjeklocha, L., et al. (2009). Nanocapsule- delivered Sleeping Beauty mediates therapeutic factor VIII ex- pression in liver sinusoidal endothelial cells of hemophilia A mice. J. Clin. Invest. 119, 2086–2099. Liu, F., Song, Y., and Liu, D. (1999). Hydrodynamics-based transfection in animals by systemic administration of plasmid DNA. Gene Ther. 6, 1258–1266. Liu, G., Geurts, A.M., Yae, K., et al. (2005). Target-site pref- erences of Sleeping Beauty transposons. J. Mol. Biol. 346, 161–173. Liu, H., Liu, L., Fletcher, B.S., and Visner, G.A. (2006). Sleeping Beauty-based gene therapy with indoleamine 2,3-dioxygenase inhibits lung allograft fibrosis. FASEB J. 20, 2384–2386. Liu, L., Sanz, S., Heggestad, A.D., et al. (2004). Endothelial tar- geting of the Sleeping Beauty transposon within lung. Mol. Ther. 10, 97–105. Liu, L., Liu, H., Visner, G., and Fletcher, B.S. (2006a). Sleeping Beauty-mediated eNOS gene therapy attenuates monocrota- line-induced pulmonary hypertension in rats. FASEB J. 20, 2594–2596. NONVIRAL GENE DELIVERY WITH SB 1049 Liu, L., Mah, C., and Fletcher, B.S. (2006b). Sustained FVIII expression and phenotypic correction of hemophilia A in neonatal mice using an endothelial-targeted Sleeping Beauty transposon. Mol. Ther. 13, 1006–1015. Mates, L., Chuah, M.K., Belay, E., et al. (2009). Molecular evo- lution of a novel hyperactive Sleeping Beauty transposase en- ables robust stable gene transfer in vertebrates. Nat. Genet. 41, 753–761. Modlich, U., Navarro, S., Zychlinski, D., et al. (2009). Insertional transformation of hematopoietic cells by self-inactivating lentiviral and gammaretroviral vectors. Mol. Ther. 17, 1919– 1928. Moldt, B., Yant, S.R., Andersen, P.R., et al. (2007). Cis-acting gene regulatory activities in the terminal regions of Sleeping Beauty DNA transposon-based vectors. Hum. Gene Ther. 18, 1193–1204. Moldt, B., Miskey, C., Staunstrup, N.H., et al. (2011). Compara- tive genomic integration profiling of Sleeping Beauty transpo- sons mobilized with high efficacy from integrase-defective lentiviral vectors in primary human cells. Mol. Ther. 19, 1499–1510. Montini, E., Held, P.K., Noll, M., et al. (2002). In vivo correction of murine tyrosinemia type I by DNA-mediated transposition. Mol. Ther. 6, 759–769. Ohlfest, J.R., Demorest, Z.L., Motooka, Y., et al. (2005a). Com- binatorial antiangiogenic gene therapy by nonviral gene transfer using the Sleeping Beauty transposon causes tumor regression and improves survival in mice bearing intracranial human glioblastoma. Mol. Ther. 12, 778–788. Ohlfest, J.R., Frandsen, J.L., Fritz, S., et al. (2005b). Phenotypic correction and long-term expression of factor VIII in hemo- philic mice by immunotolerization and nonviral gene trans- fer using the Sleeping Beauty transposon system. Blood 105, 2691–2698. Orban, T.I., Apati, A., Nemeth, A., et al. (2009). Applying a ‘‘double-feature’’ promoter to identify cardiomyocytes differentiated from human embryonic stem cells follow- ing transposon-based gene delivery. Stem Cells 27, 1077– 1087. Ortiz-Urda, S., Thyagarajan, B., Keene, D.R., et al. (2002). Stable nonviral genetic correction of inherited human skin disease. Nat. Med. 8, 1166–1170. Peng, P.D., Cohen, C.J., Yang, S., et al. (2009). Efficient nonviral Sleeping Beauty transposon-based TCR gene transfer to pe- ripheral blood lymphocytes confers antigen-specific antitumor reactivity. Gene Ther. 16, 1042–1049. Schambach, A., Mueller, D., Galla, M., et al. (2006). Overcoming promoter competition in packaging cells improves produc- tion of self-inactivating retroviral vectors. Gene Ther. 13, 1524–1533. Schambach, A., Galla, M., Maetzig, T., et al. (2007). Improving transcriptional termination of self-inactivating gamma- retroviral and lentiviral vectors. Mol. Ther. 15, 1167–1173. Schroder, A.R., Shinn, P., Chen, H., et al. (2002). HIV-1 integra- tion in the human genome favors active genes and local hot- spots. Cell 110, 521–529. Singh, H., Manuri, P.R., Olivares, S., et al. (2008). Redirecting specificity of T-cell populations for CD19 using the Sleeping Beauty system. Cancer Res. 68, 2961–2971. Staunstrup, N.H., Moldt, B., Mates, L., et al. (2009). Hybrid lentivirus–transposon vectors with a random integration profile in human cells. Mol. Ther. 17, 1205–1214. Suda, T., Suda, K., and Liu, D. (2008). Computer-assisted hy- drodynamic gene delivery. Mol. Ther. 16, 1098–1104. Sumiyoshi, T., Holt, N.G., Hollis, R.P., et al. (2009). Stable transgene expression in primitive human CD34 + hematopoi- etic stem/progenitor cells, using the Sleeping Beauty transpo- son system. Hum. Gene Ther. 20, 1607–1626. Tamhane, M., and Akkina, R. (2008). Stable gene transfer of CCR5 and CXCR4 siRNAs by Sleeping Beauty transposon system to confer HIV-1 resistance. AIDS Res. Ther. 5, 16. Thrasher, A.J., Gaspar, H.B., Baum, C., et al. (2006). Gene ther- apy: X-SCID transgene leukaemogenicity. Nature 443, E5-E6; discussion E6-E7. VandenDriessche, T., Ivics, Z., Izsvak, Z., and Chuah, M.K. (2009). Emerging potential of transposons for gene therapy and generation of induced pluripotent stem cells. Blood 114, 1461–1468. Vigdal, T.J., Kaufman, C.D., Izsvak, Z., et al. (2002). Common physical properties of DNA affecting target site selection of Sleeping Beauty and other Tc1/mariner transposable elements. J. Mol. Biol. 323, 441–452. Vink, C.A., Gaspar, H.B., Gabriel, R., et al. (2009). Sleeping Beauty transposition from nonintegrating lentivirus. Mol. Ther. 17, 1197–1204. Voigt, K., Izsvak, Z., and Ivics, Z. (2008). Targeted gene insertion for molecular medicine. J. Mol. Med. 86, 1205–1219. Walisko, O., Schorn, A., Rolfs, F., et al. (2008). Transcriptional activities of the Sleeping Beauty transposon and shielding its genetic cargo with insulators. Mol. Ther. 16, 359–369. Wang, X., Sarkar, D.P., Mani, P., et al. (2009). Long-term re- duction of jaundice in Gunn rats by nonviral liver-targeted delivery of Sleeping Beauty transposon. Hepatology 50, 815–824. Wilber, A., Frandsen, J.L., Geurts, J.L., et al. (2006). RNA as a source of transposase for Sleeping Beauty-mediated gene in- sertion and expression in somatic cells and tissues. Mol. Ther. 13, 625–630. Wilber, A., Linehan, J.L., Tian, X., et al. (2007). Efficient and stable transgene expression in human embryonic stem cells using transposon-mediated gene transfer. Stem Cells 25, 2919–2927. Williams, D.A. (2008). Sleeping Beauty vector system moves toward human trials in the United States. Mol. Ther. 16, 1515– 1516. Wolf, D., and Goff, S.P. (2009). Embryonic stem cells use ZFP809 to silence retroviral DNAs. Nature 458, 1201–1204. Wu, A., Oh, S., Ericson, K., et al. (2007). Transposon-based in- terferon c gene transfer overcomes limitations of episomal plasmid for immunogene therapy of glioblastoma. Cancer Gene Ther. 14, 550–560. Wu, X., Li, Y., Crise, B., and Burgess, S.M. (2003). Transcription start regions in the human genome are favored targets for MLV integration. Science 300, 1749–1751. Xue, X., Huang, X., Nodland, S.E., et al. (2009). Stable gene transfer and expression in cord blood-derived CD34 + hema- topoietic stem and progenitor cells by a hyperactive Sleeping Beauty transposon system. Blood 114, 1319–1330. Yamanaka, S. (2009). A fresh look at iPS cells. Cell 137, 13–17. Yant, S.R., Meuse, L., Chiu, W., et al. (2000). Somatic integration and long-term transgene expression in normal and haemo- philic mice using a DNA transposon system. Nat. Genet. 25, 35–41. Yant, S.R., Ehrhardt, A., Mikkelsen, J.G., et al. (2002). Trans- position from a gutless adeno-transposon vector stabi- lizes transgene expression in vivo. Nat. Biotechnol. 20, 999–1005. 1050 IVICS AND IZSVÁK Yant, S.R., Wu, X., Huang, Y., et al. (2005). High-resolution genome-wide mapping of transposon integration in mam- mals. Mol. Cell. Biol. 25, 2085–2094. Yant, S.R., Huang, Y., Akache, B., and Kay, M.A. (2007). Site- directed transposon integration in human cells. Nucleic Acids Res. 35, e50. Zayed, H., Izsvak, Z., Walisko, O., and Ivics, Z. (2004). Devel- opment of hyperactive Sleeping Beauty transposon vectors by mutational analysis. Mol. Ther. 9, 292–304. Zhang, G., Budker, V., and Wolff, J.A. (1999). High levels of foreign gene expression in hepatocytes after tail vein injections of naked plasmid DNA. Hum. Gene Ther. 10, 1735–1737. Zhu, J., Kren, B.T., Park, C.W., et al. (2007). Erythroid-specific expression of b-globin by the Sleeping Beauty transposon for sickle cell disease. Biochemistry 46, 6844–6858. Address correspondence to: Dr. Zoltan Ivics Max Delbruck Center for Molecular Medicine Robert Rössle Strasse 10 D-13125 Berlin Germany E-mail: zivics@mdc-berlin.de Received for publication August 5, 2011; accepted after revision August 22, 2011. Published online: August 25, 2011. NONVIRAL GENE DELIVERY WITH SB 1051