GET THE APP

pEPI for Gene Therapy Non viral episomes and their Application in
Journal of Cell Science & Therapy

Journal of Cell Science & Therapy
Open Access

ISSN: 2157-7013

+44 1300 500008

Research Article - (2011) Volume 2, Issue 3

pEPI for Gene Therapy Non viral episomes and their Application in Somatic Gene Therapy

Claudia Hagedorn* and Hans J. Lipps
Centre for Biomedical Education and Research, Institute of Cell Biology, Witten/Herdecke University, Germany
*Corresponding Author: Claudia Hagedorn, Centre for Biomedical Education and Research, Institute of Cell Biology, Witten-Herdecke University, Stockumer Str. 10, 58453 Witten, Germany, Tel: +49 2302 926144 Email:

Abstract

Nowadays, in most gene therapy trials virus based vectors are used because of their high efficiency. Nevertheless, safety risks like transformation of the cell by viral proteins, insertional mutagenesis, or innate immune reactions cannot be excluded. Basing on the idea of an ideal vector for gene therapy that is highly efficient but lacks these safety risks, non viral vectors systems represent an attractive alternative. With the construction of the non viral, S/ MAR based vector pEPI at the end of the last century, a first step towards non viral gene therapy has been made. S/MAR based vectors do not contain any viral elements, do not integrate and show stable transgene expression in the targeted cell or organism. Within the last decade, S/MAR based vectors were further improved and modified, and find now broad application in basic research and also become more and more recognized in gene therapeutic and clinical trials.

Introduction

The development of an episomal vector

During the last decade many clinical gene therapy trials have been taken place. Besides beneficial results for the patients, also severe problems became obvious regarding safety aspects of the currently used technologies. Currently, a variety of episomal vectors of different origins find broad application in gene therapeutic approaches. Episomal vectors of viral origin benefit from their high transduction efficiency of most tissues and cells. Nevertheless, their successful application is limited by several safety issues. For example, adenoviral vectors were successfully used for gene therapeutic correction of Crigler Najjar (CN) disease in rats for more than 2 years [1]. These vectors can be produced easily, but bear the potential risk of random integration and toxicity [2,3]. Adeno-associated virus (AAV) based vectors have also been used in several clinical studies regarding cystic fibrosis, Parkinson’s disease, and haemophilia B [4-6]. AAV vectors display low toxicity, but low cloning capacity and rare integration events are limiting factors [7]. Episomal vectors that are of herpes simplex viral origin have a large cloning capacity and are capable of transducing neurons, myoblasts, retinal cells, and hepatocytes - but also bear the potential to integrate into the cellular genome [8-11]. Efficient and sustained transgene expression was also achieved with non-integrating lentiviral vectors. Even though the integration risk of these vectors is greatly reduced, loss of expression in non dividing cells and a remaining risk of insertional mutagenesis were observed [12]. Since episomal vectors of non viral origin lack the ability to transduce the target cell, they represent an attractive alternative to viral vectors. The naked DNA (plasmid DNA) needs to be transfected into target cells, and once delivered into nuclei these vectors exhibit different strategies for replication and maintenance. The Epstein Barr virus (EBV) derived plasmid replicon replicates once per cell cycle but requires expression of viral proteins (EBNA) bearing the risk of cellular transformation [13,14]. With the construction of the first human artificial chromosome (HAC) in 1997 a break through for gene therapeutic approaches was expected. The key advantages of HAC include their mitotic stability, non integrating maintenance, and large cloning capacity [15], while specialized delivery techniques are required [16]. Recently, HACs were firstly applied in a gene therapy approach using the mdx mouse model for Duchenne muscular dystrophy (DMD). In the two step approach, initially mesoangioblasts from dystrophic mdx mice were genetically corrected with a HAC containing the human dystrophin genetic locus, and subsequently engrafted robustly into mice. A morphological and functional amelioration of the phenotype that lasted for up to 8 months after transplantation was observed and indicates a potential role of HACs in pre-clinical studies [17].

As outlined above, viral based vectors bear potential safety risks like insertional mutagenesis and may trigger innate immune reactions [2]. In consequence, viral vectors are used in gene therapy solely after a careful risk benefit analysis. Accordingly, non viral, autonomously replicating vectors are considered to be a safe alternative, but their application in gene therapy is nowadays strongly limited due to their low establishment efficiency [18]. Attempts to construct non-viral episomal vectors date back to the 1980s when ARS (autonomously replicating sequences) have been described in yeast [19]. The basic was to identify replication promoting sequences in mammalian cells and insert them into plasmids, very similar to the ARS assay in yeast. Thus, these “new” vectors would undergo the safety problems of viral vectors and should not lead to transformation of the recipient cell [18]. Yeast ARS are short sequences and represent binding sites for the ORC (origin recognition complex) [20]. The ARS isolated from yeast were found to promote replication when inserted into a plasmid [19]. When mammalian DNA was restriction digested and cloned into yeast plasmids, a multitude of sequences was found to promote replication in yeast. Transfected in mammalian cells however, the plasmids did not replicate autonomously and rather got lost from cells or integrated into the genome. These results suggested that some sequences of the mammalian genome somehow substitute ARS in yeast but do not act as origins of replication in mammalian cells. Insertion of putative mammalian origins of replication into plasmids resulted in a few rare cases in episomal maintenance [21,22]. Subsequent sequence analyses of various mapped mammalian origins of replication revealed no sequence homologies but rather a number of structural characteristics, such as long AT-rich regions, CpG islands, bent DNA and the presence of S/MAR (scaffold/matrix attached region) sequences [23,24]. S/MAR sequences are involved in a wide variety of biological processes in which compatibility to the nuclear matrix is of importance, including origin of replication function [25,26], modulation of gene expression [27], insulator function [28], and long-term maintenance of high transcription levels by counteracting DNA methylation [29,30]. In the late 1990 a non viral, episomal, and autonomously replicating plasmid was constructed in our lab: The S/MAR based vector pEPI replicates episomally and is mitotically stable over hundreds of generations in the absence of selection [31]. During the following years, pEPI attracted more and more attention. Based on the first prototype, inducible vector derivates were constructed, pEPI minicircle plasmids were developed, and facing to its potential application in gene therapy approaches, pEPI was improved for In vitro and In vivo applications. These issues will be discussed in detail in the following section, pointing out an exciting role for pEPI in basic research and future gene therapeutic applications (Table 1).

promoter
  modification application cell line/ organism transfection method reference
pEPI-1 CMV promoter in vitro CHO-K1, HeLa, IMEF, lipofection Piechaczek et al. [31]
      HCT116, CD34+ cells,   Jenke et al. [34]
  (prototype)       Papapetrou et al. [43]
          Tessadori et al. [39]
    in vivo pig embryos sperm-mediated gene Manzini et al. [44]
        transfer  
pEPI-CAG CAG promoter in vitro CHO-K1 lipofection Manzini et al. [48]
           
pEPI-AAT aplpa1-antitrypsin in vivo mouse liver, neonatal hydrodynamic injection, Argyros et al. [45]
      mouse liver, lungs, heart, systemic injection of PEI  
  (AAT) promoter   brain, kidney DNA Wong et al. [64]
           
pEPI-Ubc human Ubc promoter in vivo mouse liver, neonatal hydrodynamic injection, Argyros et al. [45]
      mouse liver, lungs, heart, systemic injection of PEI  
      brain, kidney DNA Wong et al. [64]
           
pEPito hCMV/EF1P in vitro HEK293, NIH 3T3    
lipofection  
          Haase et al. [49]
    in vivo mouse liver hydrodynamic injection  
           
pEPI-TetON TRE tight module in vitro CHO-K1 lipofection  
          Rupprecht et al. [42]
    in vivo mouse liver hydrodynamic injection  
           
backbone
  modification application cell line/ organism transfection method reference
pEPito CpG depleted in vitro HEK293, NIH 3T3    
lipofection  
  cbackbone       Haase et al. [49]
    in vivo mouse liver hydrodynamic injection  
           
S/MAR based lacking bacterial in vitro CHO-K1 lipofection Nehlsen et al. [54]
minicircle elements       Broll et al. [57]
    in vivo mouse liver hydrodynamic injection Argyros et al. [58]

Table 1: Modifications of the prototype pEPI applied to the promoter and plasmids backbone.

Episomal S/MAR vectors

As mentioned above, the first vector shown to replicate autonomously in a variety of cell lines and retained in the absence of selection was the vector pEPI-1. The large T antigen of SV40 was replaced by a S/MAR sequence derived from the human β-interferon gene cluster [31]. Multiple AATATATTTA elements in this sequence serve as DNA unwinding elements (DUE) and allow stress-induced DNA duplex destabilisation of dsDNA [32]. In CHO cells, pEPI replicates at low copy numbers with 5-10 copies per cell [33] and is stably retained in the absence of selection for basically unlimited time [31]. Episomal maintenance of pEPI was observed in several other cell lines including HeLa, HEK293, and even human primary cells. In all tested cell lines the vector was mitotically stable in the absence of selection and occurred in an average copy number below 10 per cell [18], indicating a highly efficient replication and segregation. In established cells pEPI is associated with the nuclear matrix via the matrix protein SAFA and segregation occurs with the host chromosomes via hitchhiking (Figure 1) [34]. Although hitherto data are unsuggestive of preferential binding sites of pEPI to certain mitotic chromosomes, it may be possible that binding of pEPI is restricted to specific sequence elements [33,35]. Like the cellular genome, pEPI replicates once per cell cycle during early S phase. The origin recognition complex was shown to assemble at various regions of the episome [36], thus behaving as the initiation zone of genomic origins of replication [23] (Table 2).

cell-science-therapy-illustration-episomal

Figure 1: (a) Simplified illustration of episomal maintenance. S/MAR sequences (blue) mediate binding of chromosomal DNA (loops) to the nuclear scaffold (brown). pEPI is bound to the nuclear scaffold via a SAF A – S/MAR interaction. Presumably, additional proteins are involved. (b) FISH staining of pEPI molecules (purple) attached to metaphase chromosomes (blue), using the hitchhiking principle to be segregated during mitosis.

  modification application cell line/ organism transfection method reference
pEPI-EGFP EGFP in vitro CHO-K1, HeLa, IMEF, lipofection Piechaczek et al. [31]
      HCT116, CD34+ cells,   Jenke et al. [34]
          Papapetrou et al. [43]
          Tessadori et al. [39]
    in vivo mouse liver hydrodynamic injection Rupprecht et al. [40]
pEPI-Luc luciferase in vitro U251 lipofection Argyros et al [58]
    in vivo mouse liver, neonatal hydrodynamic injection, Argyros et al. [45]
      mouse liver, lungs, heart, systemic injection of PEI  
      brain, kidney DNA Wong et al. [64]
      mouse left hind muscle ultrasound-mediated Li et al. [70]
      RIF-1 tumors gene transfer with  
        microbubbles  
iBAC-S/MAR-LDLR LDLR in vitro CHO ldlr(-/-) lipofection Lufino et al. [59]
S/MAR-βLCR-HBB beta -globin in vitro MEL, K562 lipofection Sgourou et al. [60]
  microlocus cassette        
pBcLucA1 bcl-2 in vivo mouse liver hydrodynamic injection Wong et al. [64]
pEPI-1-PDX1 pancreatic in vivo rat liver hydrodynamic injection Cim et al. [66]
  transcription factor        
  pdx1        
S/MAR-UbC-GDNF rat glial cell line-        
-S/MAR derived neurotrophic in vivo rat brain stereotactical injection Yurek et al. [69]
  factor (GDNF)        
pEPI-b3a2 hU6-bcr abl shRNA in vitro K562 lipofection Jenke et al. [61]
pEPI-RNAi hU6-HBV shRNA in vitro HepG2-2.15 lipofection Jenke et al. [62]

Table 2: Modifications of the prototype pEPI applied to the transgene.

After transfection and an initial selection phase, only a small percentage of vector molecules (1-5%) establishes stable as an episome, even though pEPI carries all cis-acting elements required for episomal replication and maintenance. This low establishment rate is one main limiting factor of pEPI for its use in gene therapy approaches. Therefore, it is most likely that epigenetic mechanisms are involved in the establishing process; a phenomenon that has also been reported for EBV-based vectors recently [37,38]. Once established pEPI localises in the interchromatin space (regions of less condensed chromatin) and is found to associate with nuclear speckles, a nuclear region involved in RNA processing. An association with early replicating foci that is stably retained over mitosis may explain the high mitotic stability of the vector. The establishment efficiency strongly depends on the nuclear compartment the vector reaches after transfection and the chromatin structure it thereby adopts. But when established pEPI behaves surprisingly non-dynamic throughout the cell cycle [33,39] and is associated with histone modifications typically found in active chromatin. Especially within the S/MAR region an accumulation of active modifications like H3K4me1 and H3K4me3 is detectable [40]. But regardless of successful establishment, only 30 70% cells containing pEPI as an episome display transgene (GFP) expression to a level detectable by FACS, suggesting that minimal transcriptional activity is sufficient to assure vector replication and maintenance. However, the expression profile of a single cell clone remains constant over numerous cell generations.

Modifications, Improvements and Next Vector Generations

The underlying mechanism of episomal replication and maintenance of pEPI is an active transcription running into the S/ MAR. It has been demonstrated that whenever transcription running into the S/MAR was abrogated, i.e. by deletion of transgene, deletion of the promoter, or a termination signal between transgene and S/MAR, establishment failed and plasmids got lost from the cells or in some rare cases even integrated into the host genome [41]. Recently, in a proof of principle experiment an inducible vector was constructed in which transcription unit was under control of a tetracycline responsive promoter (TetON). Cells were transfected in the presence of doxycycline to ensure establishment. Subsequently, removing of doxycycline after establishment resulted in constant loss of vector molecules from cells. Furthermore, the pEPI-TetON system also works In vivo. Hydrodynamic injection of pEPI-TetON in mice and simultaneous administration of doxycycline resulted in a 5 fold increased transgene expression in mice liver. Despite these exciting results, the system demands further improvements since background expression of the transgene was also detectable in the absence of doxycycline [42]. These results provide evidence of an active transcription running into the S/MAR being essential for the episomal behaviour of the vector. Accordingly, modifications solely apply to vector backbone, insertion of additional, functional sequences, and insertion of diverse promoters (Figure 2).

cell-science-therapy-functional-elements

Figure 2: (a) Schematic illustration of the functional elements of pEPI: Expression of the respective reporter- or transgene (green), driven from an active promoter (turquoise) has to run into the S/MAR element (blue). When these preconditions are not given, S/MAR based vectors fail to establish episomally. In consequence they got lost from the cell or, in rare cases, integrate into the cellular genome. (b) Modifications of pEPI vectors solely apply to promoter, transgene and backbone.

Basically, the prototype vector pEPI-1 represents an almost ideal vector for gene therapy: Its episomal replication and maintenance has been shown in several cell lines [43] and has successfully been used for animal transgenesis [44]. Moreover, pEPI is completely based on chromosomal elements [31] and does not integrate into the cellular genome [41]. But pointing towards gene therapeutic applications, long-term transgene expression and episomal maintenance in every cell type are of immense importance. Subsequently, several endeavours were made to improve pEPI for In vitro and In vivo applications. It has been reported that early silencing of the CMV promoter In vivo is due to cytosine methylation within the CMV [45]. To circumvent early silencing In vivo, the CMV promoter of pEPI was replaced by a CAG promoter. The CAG promoter represents a synthetic hybrid promoter consisting of CMV enhancer element, chicken beta actin promoter sequences and rabbit beta globin 3’ UTR sequences [46]. This promoter is described to be less prone to cytosine methylation and displays stable transcription in a variety of tissues In vivo [47]. Inserted into pEPI 1, increased eGFP expression levels for pEPI-CAG (pEPI 1 25%, pEPI-CAG 55%) were detected [48]. Likewise, a combination of CMV enhancer element and elongation factor-1 promoter (hCMV/EF1P) gave long term gene expression In vitro and In vivo [49]. Moreover, to avoid innate immune reactions, a reduction of the CpG motifs in pEPI backbone has proven to be promising. The CpG reduced vector pEPito showed improved establishment rates In vitro, and transgene expression driven from pEPito was stable up to 32 days post injection In vivo [49].

Preventing innate immune responses is one of the main challenges towards safe gene therapy. Thus, the design of a vector lacking any bacterial elements was a step forward. The so called minicircles were originally constructed with the use of recombinases [50-55], but are nowadays also commercially available [56]. Like pEPito, S/MAR based minicircles display enhanced transgene expression and sustained transgene expression In vitro and In vivo. Introduction of two Flp recombination sites into pEPI at either side of the promoter transgene S/MAR cassette led to the first S/MAR based minicircle. The subsequent induction of a Flp recombinase resulted in a recombination event and the emerging of two circular units: one containing the bacterial elements and one containing exclusively the promoter transgene S/ MAR cassette [54]. S/MAR minicircles are maintained episomally with increased expression levels and establishment rates In vitro [57] as well as In vivo [58]. A new generation of the S/MAR minicircles with a truncated S/MAR element (~700 bp) displayed further improved stability and transgene expression In vitro [57]. An overview of the S/ MAR based vector generations is given in Figure 3. Interestingly, the prototype pEPI 1 as well as next vector generations establish within a certain range of plasmid copies per cell, suggesting a stringent copy number control in the recipient cells. The contradicting observation of similar copy numbers and broad variations in expression levels leads to the suggestion that epigenetic features, chromatin structure, and nuclear localisation strongly influence the regulation of transcription from non-viral episomes.

cell-science-therapy-vector-generations

Figure 3: Schematic depiction of pEPI vector generations. The prototype pEPI (a) was constructed in 1999 [31] consisting of a CMV GFP S/MAR expression cassette, while (b) pEPI used today harbours EGFP [41]. The construction of the pEPI minicircle (c) resulted in a vector lacking any bacterial residual elements [54]. (d) pEPito represents a derivate of the prototype pEPI with a CpG depleted backbone [49]. In our group, an inducible pEPI vector was constructed (e) in which transgene expression is under control of a tetracycline responsive promoter [42].

Applications

In vitro

As discussed above, S/MAR based episomes have been demonstrated to function successfully In vitro and In vivo. For example, episomal maintenance in primary human fibroblast-like cells and at low levels even in human CD34+ cells has been demonstrated [43]. To ensure natural regulation an expression of the therapeutic gene, insertion of whole genomic DNA loci including all regulatory elements into the respective vector is of uttermost importance for gene therapeutic applications. Recently, is has been shown that pEPI is not only capable of expressing one single transgene. In two approaches, cloning of genomic loci into S/MAR based vectors was impressively demonstrated. The genomic locus of the human low density lipoprotein receptor (LDLR) with 135 kb in size was cloned into a S/MAR based vector [59]. The iBAC-S/MAR-LDLR was maintained episomally for at least 11 weeks and excitingly was also capable to restore LDLR function completely when transfected into CHO ldlr(-/-) cells [59]. In a similar approach, Sgourou and colleagues were able to express beta globin at physiological levels by combining an S/MAR element with the beta globin microlocus cassette, including the locus control region of beta globin (betaLCR-HBB) [60]. However, besides the utilisation of S/MAR based vectors to express therapeutic genes, they are also suitable for expressing gene regulatory sequences. Following the idea of silencing viral genes, the bcr abl fusion gene was used as a first target for shRNA expressed by pEPI. Indeed, a significant decrease of bcr-abl gene expressing in K562 cells was detectable, while again the vector behaves like an episome in the absence of selection for at least four month [61]. In a next approach, shRNA targeting HBV replication associated gene expression was designed and cloned into pEPI. A significant reduction of HBV DNA content in HepG2-2.15 cells was detectable eight month post transfection [62].

In vivo

Five years ago, pEPI was firstly utilised in a fascinating In vivo approach. Vector DNA was delivered into female pigs using sperm mediated gene transfer (SMGT) with surprising results: Transgene (GFP) expression was ascertained in nine out of 12 modified pig foetuses and if expression was shown, it was detectable in all tissues of the foetus with up to 80% of GFP expressing cells [44]. The high efficiency in this first In vivo approach indicated that pEPI is capable of adopting the respective epigenetic features needed for efficient establishment and propagation in the developing organism. But researchers were confronted with a much more complex situation when pEPI was firstly delivered into differentiated tissues. These approaches seem to be more challenging, as one of the main imitating factors is rapid silencing of transgene expression. As mentioned above, this silencing process was described to be due to cytosine methylation within the promoter sequence and often occurs within the first week post transfection [45,63]. The insertion of tissue specific promoters was a crucial step to overcome this phenomenon. Cloning of a liver-specific promoter (alpha1-antitrypsin; AAT) into pEPI resulted in prolonged transgene expression in mice liver for up to 24 weeks [45]. Since it is known from In vitro experiments that S/MAR based vectors solely establish in mitotically active cells, this situation was mimicked in the mice liver by a 70% partial hepatectomy. Surprisingly, both, the AAT as well as the CMV based vectors failed to establish in the regenerating liver and subsequently got lost [45]. Again, a similar behaviour was observed In vitro when establishment efficiency of pEPI was determined without an initial selection phase after transfection (own observations). An elegant strategy for In vivo selection was published recently [64]. The authors provided a survival advantage for transfected cells over non transfected cells by introducing the bcl 2 gene in a liver specific pEPI vector (pEPI AAT). Briefly, the physiological ligand FasL initiates cell death in hepatocytes when bound to the liver transmembrane protein Fas. This pathway may also be activated by binding of the agonistic antibody Jo2, while BCL2 in turn acts like decoy receptor for Jo2 and thereby inhibits apoptosis. The AAT-S/MAR-bcl-2 plasmid was delivered to mice liver using hydrodynamic injection. Despite constant Jo2 challenges, luciferase expression was detectable in mice liver for at last 12 weeks indicating maintained bcl-2 expression [64].

Recently, a further step towards pre-clinical application was undertaken when S/MAR based vectors were injected into neonatal mice [3]. Vectors were administered via the superior temporal vein of mice at age 1.5 days. Transcription was driven either from an UbC or AAT promoter and plasmid DNA was injected in combination with a polyethylene (PEI) complex. Using this approach an efficient transfection of lung, brain, heart, brain, spleen, liver, and kidney was achieved. The expression profile for the UbC driven vector gave a gradual increase of luciferase expression peaking at days 11 12 and decreased by day 25 reaching baseline levels. When transcription was driven from the liver specific AAT promoter, a corresponding expression profile was observed, highest luciferase expression levels were detected at days 11 12, and dropped thereafter. As expected, the AAT-S/MAR vector gave tissue specific expression in the midabdomen region [3]. The attenuated expression of the transgene around day 20 is believed to coincide with the rapid proliferation of hepatocytes and the steep increase of liver weight [42], a phenomenon that has also been described after a partial hepatectomy [45]. Likewise, also minicircles harbouring liver specific promoters (AAT or UbC) were examined In vivo. Both, mini AAT-S/MAR and mini UbC S/MAR showed significantly increased transgene expression 24 h post delivery when compared to mini CMV-S/MAR. Expression levels remained elevated over a detected time slot of 92 days [58].

Pre clinical applications

Towards gene therapeutic clinical applications of S/MAR based vectors, first studies have been published recently. Ferber and colleagues mediated ectopic expression of a pancreatic transcription factor (PDX1) in mice liver by introducing PDX1 into the livers. PDX1 was not only expressed in liver, but also capable of reversing streptozotocin mediated hyperglycaemia [65]. Basing on these results, the pdx1 gene was cloned into different S/MAR based vectors and delivered to rat liver, determining level as well as duration of transgene expression [66]. The original pEP 1 PDX1 gave an expression peak at day 1 post injection but dropped significantly at day 3. In comparison, expression driven from AAT-PDX1-S/MAR lasted up to day three and dropped gradually afterwards. Additionally, a vector similar to the above described pEPito (reduced CpG motifs in backbone) was used. The pEPito AAT not only expressed PDX1 at a high level, but also mediated a pancreatic phenotype in the rat livers. Insulin 2 expression was detected in 66%, elevated up to 70 fold on day 14 in comparison to day 1 [66].

Various studies focused on the application of S/MAR vectors in other differentiated tissues than liver, like adult central nervous system, muscle, and tumours. For CNS studies plasmids were constructed in which the transgene is flanked by two S/MAR elements [67-69]. Yurek et al. [69] for example examined plasmid DNA delivery of a S/ MAR UbC GDNF S/MAR vector coding for rat glial cell line-derived neurotrophic factor (GDNF) and found broad GDNF expression up to 400 600% over endogenous expression levels for at least 3 week post-delivery. An analogous vector, S/MAR UbC LUC S/MAR, was delivered in a complex with lysine 30mer peptides substituted with PEG. One single intrathecal injection resulted in stably expressed luciferase for up to one year in the striatum [68]. Supported delivery strategies into muscle were similar successful: Microbubble mediated ultrasound delivery resulted in strong expression levels in muscle tissue, while expression from control vectors lacking a S/MAR element was silenced within a month [70]. The outcome of these studies points towards an utilisation of optimised S/MAR vectors in combination with certain gene delivery strategies, being highly beneficial in future gene therapy applications. Additionally, S/MAR vectors may find broad application in the field of cancer research. As shown previously, delivery of a luciferase expressing S/MAR vector into RIF 1 tumours via ultrasound has been used successfully to monitor tumour growth in mice [70]. In future, the functional luciferase expression encoded by S/MAR vectors may, besides monitoring, play an important in tumour trafficking or even cancer treatment.

Since S/MAR based vectors exclusively consist of chromosomal elements, their episomal maintenance gives an advantage over the potential safety risks of viral vectors. Moreover, their chromosomal maintenance, long term expression, and high insert cloning capacity make them an attractive tool for gene therapeutic applications.The progressions made in minicircle development and the construction of inducible vectors highlight the potential of S/MAR based vectors for a broad range of applications, although further improvements are necessary for future clinical applications.

Acknowledgements

Work in the Lipps lab was supported by the Deutsche Forschungsgemeinschaft (SPP 1230). Claudia Hagedorn is supported by the Peter and Traudl Engelhorn foundation.

Author Disclosure Statement

The authors declare no competing financial interest.

References

  1. Toietta G, Mane VP, Norona WS, Finegold MJ, Ng P, et al. (2005) Lifelong elimination of hyperbilirubinemia in the Gunn rat with a single injection of helper-dependent adenoviral vector. Proc Natl Acad Sci U S A 102: 3930-3935.
  2. Ehrhardt A, Haase R, Schepers A, Deutsch MJ, Lipps HJ, et al. (2008) Episomal vectors for gene therapy. Curr Gene Ther 8: 147-161.
  3. Wong SP, Argyros O, Howe SJ, Harbottle RP (2011) Systemic gene transfer of polyethylenimine (PEI)-plasmid DNA complexes to neonatal mice. J Control Release 150: 298-306.
  4. During MJ, Kaplitt MG, Stern MB, Eidelberg D (2001) Subthalamic GAD gene transfer in Parkinson disease patients who are candidates for deep brain stimulation. Hum Gene Ther 12: 1589-1591.
  5. Flotte TR, Zeitlin PL, Reynolds TC, Heald AE, Pedersen P, et al. (2003) Phase I trial of intranasal and endobronchial administration of a recombinant adeno-associated virus serotype 2 (rAAV2)-CFTR vector in adult cystic fibrosis patients: a two-part clinical study. Hum Gene Ther 14: 1079-1088.
  6. Manno CS, Chew AJ, Hutchison S, Larson PJ, Herzog RW, et al. (2003) AAV-mediated factor IX gene transfer to skeletal muscle in patients with severe hemophilia B. Blood 101: 2963-2972.
  7. Nakai H, Yant SR, Storm TA, Fuess S, Meuse L, et al. (2001) Extrachromosomal recombinant adeno-associated virus vector genomes are primarily responsible for stable liver transduction in vivo. J Virol 75: 6969-6976.
  8. Wang Y, Mukherjee S, Fraefel C, Breakefield XO, Allen PD (2002) Herpes simplex virus type 1 amplicon vector-mediated gene transfer to muscle. Hum Gene Ther 13: 261-273.
  9. Fraefel C, Mendes-Madeira A, Mabon O, Lefebvre A, Le Meur G, et al. (2005) In vivo gene transfer to the rat retina using herpes simplex virus type 1 (HSV-1)-based amplicon vectors. Gene Ther 12: 1283-1288.
  10. Muller L, Saydam O, Saeki Y, Heid I, Fraefel C (2005) Gene transfer into hepatocytes mediated by herpes simplex virus-Epstein-Barr virus hybrid amplicons. J Virol Methods 123: 65-72.
  11. Berges BK, Wolfe JH, Fraser NW (2007) Transduction of brain by herpes simplex virus vectors. Mol Ther 15: 20-29.
  12. Yanez-Munoz RJ, Balaggan KS, MacNeil A, Howe SJ, Schmidt M, et al. (2006) Effective gene therapy with nonintegrating lentiviral vectors. Nat Med 12: 348-353.
  13. Humme S, Reisbach G, Feederle R, Delecluse HJ, Bousset K, et al. (2003) The EBV nuclear antigen 1 (EBNA1) enhances B cell immortalization several thousandfold. Proc Natl Acad Sci U S A 100: 10989-10994.
  14. Taylor GS, Haigh TA, Gudgeon NH, Phelps RJ, Lee SP, et al. (2004) Dual stimulation of Epstein-Barr Virus (EBV)-specific CD4+- and CD8+-T-cell responses by a chimeric antigen construct: potential therapeutic vaccine for EBV-positive nasopharyngeal carcinoma. J Virol 78: 768-778.
  15. Larin Z, Mejia JE (2002) Advances in human artificial chromosome technology. Trends Genet 18: 313-319.
  16. Basu J, Willard HF (2005) Artificial and engineered chromosomes: non-integrating vectors for gene therapy. Trends Mol Med 11: 251-258.
  17. Tedesco FS, Hoshiya H, D'Antona G, Gerli MF, Messina G, et al. (2011) Stem cell-mediated transfer of a human artificial chromosome ameliorates muscular dystrophy. Sci Transl Med 3: 96ra78.
  18. Jackson DA, Juranek S, Lipps HJ (2006) Designing nonviral vectors for efficient gene transfer and long-term gene expression. Mol Ther 14: 613-626.
  19. Stinchcomb DT, Thomas M, Kelly J, Selker E, Davis RW (1980) Eukaryotic DNA segments capable of autonomous replication in yeast. Proc Natl Acad Sci U S A 77: 4559-4563.
  20. Duncker BP, Chesnokov IN, McConkey BJ (2009) The origin recognition complex protein family. Genome Biol 10: 214
  21. Nielsen TO, Cossons NH, Zannis-Hadjopoulos M, Price GB (2000) Circular YAC vectors containing short mammalian origin sequences are maintained under selection as HeLa episomes. J Cell Biochem 76: 674-685.
  22. Price GB, Allarakhia M, Cossons N, Nielsen T, Diaz-Perez M, et al. (2003) Identification of a cis-element that determines autonomous DNA replication in eukaryotic cells. J Biol Chem 278: 19649-19659.
  23. DePamphilis ML (1999) Replication origins in metazoan chromosomes: fact or fiction? Bioessays 21: 5-16.
  24. Gilbert DM (2004) In search of the holy replicator. Nat Rev Mol Cell Biol 5: 848-855.
  25. Jenke BH, Fetzer CP, Stehle IM, Jonsson F, Fackelmayer FO, et al. (2002) An episomally replicating vector binds to the nuclear matrix protein SAF-A in vivo. EMBO rep 3: 349-354.
  26. Lipps HJ, Bode J (2001) Exploiting chromosomal and viral strategies: the design of safe and efficient non-viral gene transfer systems. Curr Opin Mol Ther 3: 133-141.
  27. Bode J, Benham C, Knopp A, Mielke C (2000) Transcriptional augmentation: modulation of gene expression by scaffold/matrix-attached regions (S/MAR elements). Crit Rev Eukaryot Gene Expr 10: 73-90.
  28. Goetze S, Baer A, Winkelmann S, Nehlsen K, Seibler J, et al. (2005) Performance of genomic bordering elements at predefined genomic loci. Mol Cell Biol 25: 2260-2272.
  29. Dang Q, Auten J, Plavec I (2000) Human beta interferon scaffold attachment region inhibits de novo methylation and confers long-term, copy number-dependent expression to a retroviral vector. J Virol 74: 2671-2678.
  30. Kurre P, Morris J, Thomasson B, Kohn DB, Kiem HP (2003) Scaffold attachment region-containing retrovirus vectors improve long-term proviral expression after transplantation of GFP-modified CD34+ baboon repopulating cells. Blood 102: 3117-3119.
  31. Piechaczek C, Fetzer C, Baiker A, Bode J, Lipps HJ (1999) A vector based on the SV40 origin of replication and chromosomal S/MARs replicates episomally in CHO cells. Nucleic Acids Res 27: 426-428.
  32. Bode J, Kohwi Y, Dickinson L, Joh T, Klehr D, et al. (1992) Biological significance of unwinding capability of nuclear matrix-associating DNAs. Science 255: 195-197.
  33. Stehle IM, Postberg J, Rupprecht S, Cremer T, Jackson DA, et al. (2007) Establishment and mitotic stability of an extra-chromosomal mammalian replicon. BMC cell biology 8: 33.
  34. Jenke AC, Stehle IM, Herrmann F, Eisenberger T, Baiker A, et al. (2004) Nuclear scaffold/matrix attached region modules linked to a transcription unit are sufficient for replication and maintenance of a mammalian episome. Proc Natl Acad Sci U S A 101: 11322-11327.
  35. Baiker A, Maercker C, Piechaczek C, Schmidt SB, Bode J, et al. (2000) Mitotic stability of an episomal vector containing a human scaffold/matrix-attached region is provided by association with nuclear matrix. Nat Cell Biol 2: 182-184.
  36. Schaarschmidt D, Baltin J, Stehle IM, Lipps HJ, Knippers R (2004) An episomal mammalian replicon: sequence-independent binding of the origin recognition complex. EMBO J 23: 191-201.
  37. Leight ER, Sugden B (2001) Establishment of an oriP replicon is dependent upon an infrequent, epigenetic event. Mol Cell Biol 21: 4149-4161.
  38. Leight ER, Sugden B (2001) The cis-acting family of repeats can inhibit as well as stimulate establishment of an oriP replicon. J Virol 75: 10709-10720.
  39. Tessadori F, Zeng K, Manders E, Riool M, Jackson D, et al. (2010) Stable S/MAR-based episomal vectors are regulated at the chromatin level. Chromosome Res 18: 757-775.
  40. Rupprecht S, Lipps HJ (2009) Cell cycle dependent histone dynamics of an episomal non-viral vector. Gene 439: 95-101.
  41. Stehle IM, Scinteie MF, Baiker A, Jenke AC, Lipps HJ (2003) Exploiting a minimal system to study the epigenetic control of DNA replication: the interplay between transcription and replication. Chromosome Res 11: 413-421.
  42. Rupprecht S, Hagedorn C, Seruggia D, Magnusson T, Wagner E, et al. (2010) Controlled removal of a nonviral episomal vector from transfected cells. Gene 466: 36-42.
  43. Papapetrou EP, Ziros PG, Micheva ID, Zoumbos NC, Athanassiadou A (2006) Gene transfer into human hematopoietic progenitor cells with an episomal vector carrying an S/MAR element. Gene Ther 13: 40-51.
  44. Manzini S, Vargiolu A, Stehle IM, Bacci ML, Cerrito MG, et al. (2006) Genetically modified pigs produced with a nonviral episomal vector. Proc Natl Acad Sci U S A 103: 17672-17677.
  45. Argyros O, Wong SP, Niceta M, Waddington SN, Howe SJ, et al. (2008) Persistent episomal transgene expression in liver following delivery of a scaffold/matrix attachment region containing non-viral vector. Gene Ther 15: 1593-1605.
  46. Niwa H, Yamamura K, Miyazaki J (1991) Efficient selection for high-expression transfectants with a novel eukaryotic vector. Gene 108: 193-199.
  47. Sawicki JA, Morris RJ, Monks B, Sakai K, Miyazaki J (1998) A composite CMV-IE enhancer/beta-actin promoter is ubiquitously expressed in mouse cutaneous epithelium. Exp Cell Res 244: 367-369.
  48. Manzini S, Vargiolu A, Seruggia D, Busnelli M, Cerrito MG, et al. (2010) S/MAR trek: A new generation of non-viral episomal vectors to be used in gene transfer experiments. Transgenic Research 19: 307-355.
  49. Haase R, Argyros O, Wong SP, Harbottle RP, Lipps HJ, et al. (2010) pEPito: a significantly improved non-viral episomal expression vector for mammalian cells. BMC Biotechnol 10: 20.
  50. Bigger BW, Tolmachov O, Collombet JM, Fragkos M, Palaszewski I, et al. (2001) An araC-controlled bacterial cre expression system to produce DNA minicircle vectors for nuclear and mitochondrial gene therapy. J Biol Chem 276: 23018-23027.
  51. Chen ZY, He CY, Ehrhardt A, Kay MA (2003) Minicircle DNA vectors devoid of bacterial DNA result in persistent and high-level transgene expression in vivo. Mol Ther 8: 495-500.
  52. Darquet AM, Rangara R, Kreiss P, Schwartz B, Naimi S, et al. (1999) Minicircle: an improved DNA molecule for in vitro and in vivo gene transfer. Gene ther 6: 209-218.
  53. Jacobs F, Snoeys J, Feng Y, Van Craeyveld E, Lievens J, et al. (2008) Direct comparison of hepatocyte-specific expression cassettes following adenoviral and nonviral hydrodynamic gene transfer. Gene ther 15: 594-603.
  54. Nehlsen K, Broll S, Bode J (2006) Replicating minicircles: Generation of nonviral episomes for the efficient modification of dividing cells. Gene Therapy and Molecular Biology 10: 233-244.
  55. Vaysse L, Gregory LG, Harbottle RP, Perouzel E, Tolmachov O, et al. (2006) Nuclear-targeted minicircle to enhance gene transfer with non-viral vectors in vitro and in vivo. J Gene Med 8: 754-763.
  56. Mayrhofer P, Blaesen M, Schleef M, Jechlinger W Minicircle-DNA production by site specific recombination and protein-DNA interaction chromatography. J gene med 10: 1253-1269.
  57. Broll S, Oumard A, Hahn K, Schambach A, Bode J (2010) Minicircle performance depending on S/MAR-nuclear matrix interactions. J mol biol 395: 950-965.
  58. Argyros O, Wong SP, Fedonidis C, Tolmachov O, Waddington SN, et al. (2011) Development of S/MAR minicircles for enhanced and persistent transgene expression in the mouse liver. J mol med (Berl) 89: 515-529.
  59. Lufino MM, Manservigi R, Wade-Martins R (2007) An S/MAR-based infectious episomal genomic DNA expression vector provides long-term regulated functional complementation of LDLR deficiency. Nucleic Acids Res 35: e98.
  60. Sgourou A, Routledge S, Spathas D, Athanassiadou A, Antoniou MN (2009) Physiological levels of HBB transgene expression from S/MAR element-based replicating episomal vectors. J Biotechnol 143: 85-94.
  61. Jenke AC, Eisenberger T, Baiker A, Stehle IM, Wirth S, et al. (2005) The nonviral episomal replicating vector pEPI-1 allows long-term inhibition of bcr-abl expression by shRNA. Hum gene ther 16: 533-539.
  62. Jenke AC, Wilhelm AD, Orth V, Lipps HJ, Protzer U, et al. (2008) Long-term suppression of hepatitis B virus replication by short hairpin RNA expression using the scaffold/matrix attachment region-based replicating vector system pEPI-1. Antimicrob Agents Chemother 52: 2355-2359.
  63. Conese M, Auriche C, Ascenzioni F (2004) Gene therapy progress and prospects: episomally maintained self-replicating systems. Gene Ther 11: 1735-1741.
  64. Wong SP, Argyros O, Coutelle C, Harbottle RP (2011) Non-viral S/MAR vectors replicate episomally in vivo when provided with a selective advantage. Gene Ther 18: 82-87.
  65. Ferber S, Halkin A, Cohen H, Ber I, Einav Y, et al. (2000) Pancreatic and duodenal homeobox gene 1 induces expression of insulin genes in liver and ameliorates streptozotocin-induced hyperglycemia. Nat med 6: 568-572.
  66. Cim A, Swayer G, Zhang X, Su H, Collins L, et al. (2010) Nonviral delivery of the rat PDX1 gene to rat liver for the in vivo transdifferentiation of liver cells to pancreatic beta-cells. Hum Gene Ther 21.
  67. Hughes TS, Langer SJ, Johnson KW, Chavez RA, Watkins LR, et al. (2009) Intrathecal injection of naked plasmid DNA provides long-term expression of secreted proteins. Mol Ther 17: 88-94.
  68. Kaytor MD, Weatherspoon MR, Green KJ, Kowalczyk TH, Padegimas P (2009) In vivo delivery of nucleic acid to the brain using DNA nanoparticles. Mol Ther 17: S519.
  69. Yurek DM, Fletcher AM, Smith GM, Seroogy KB, Ziady AG, et al. (2009) Long-term transgene expression in the central nervous system using DNA nanoparticles. Mol Ther 17: 641-650.
  70. Li YS, Davidson E, Reid CN, McHale AP (2009) Optimising ultrasound-mediated gene transfer (sonoporation) in vitro and prolonged expression of a transgene in vivo: potential applications for gene therapy of cancer. Cancer lett 273: 62-69.
Citation: Hagedorn C, Lipps HJ (2011) pEPI for Gene Therapy: Non-viral episomes and their Application in Somatic Gene Therapy. J Cell Sci Ther S6:001.

Copyright: © 2011 Hagedorn C, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top