Next Article in Journal
VDR Polymorphic Variants Are Related to Improvements in CRP and Disease Activity in Patients with Axial Spondyloarthritis That Undergo Anti-TNF Treatment
Next Article in Special Issue
Non-Viral Episomal Vector Mediates Efficient Gene Transfer of the β-Globin Gene into K562 and Human Haematopoietic Progenitor Cells
Previous Article in Journal
Expression and Variations in EPAS1 Associated with Oxygen Metabolism in Sheep
Previous Article in Special Issue
Appraisal for the Potential of Viral and Nonviral Vectors in Gene Therapy: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Episomes and Transposases—Utilities to Maintain Transgene Expression from Nonviral Vectors

Chair for Biochemistry and Molecular Medicine, Center for Biomedical Education and Research, School of Life Sciences (ZBAF), Faculty of Health, Witten/Herdecke University, Stockumer Strasse 10, 58453 Witten, Germany
*
Author to whom correspondence should be addressed.
Genes 2022, 13(10), 1872; https://doi.org/10.3390/genes13101872
Submission received: 22 August 2022 / Revised: 7 October 2022 / Accepted: 14 October 2022 / Published: 16 October 2022
(This article belongs to the Special Issue Advances in Non-viral Gene Transfer for Gene Therapy Applications)

Abstract

:
The efficient delivery and stable transgene expression are critical for applications in gene therapy. While carefully selected and engineered viral vectors allowed for remarkable clinical successes, they still bear significant safety risks. Thus, nonviral vectors are a sound alternative and avoid genotoxicity and adverse immunological reactions. Nonviral vector systems have been extensively studied and refined during the last decades. Emerging knowledge of the epigenetic regulation of replication and spatial chromatin organisation, as well as new technologies, such as Crispr/Cas, were employed to enhance the performance of different nonviral vector systems. Thus, nonviral vectors are in focus and hold some promising perspectives for future applications in gene therapy. This review addresses three prominent nonviral vector systems: the Sleeping Beauty transposase, S/MAR-based episomes, and viral plasmid replicon-based EBV vectors. Exemplarily, we review different utilities, modifications, and new concepts that were pursued to overcome limitations regarding stable transgene expression and mitotic stability. New insights into the nuclear localisation of nonviral vector molecules and the potential consequences thereof are highlighted. Finally, we discuss the remaining limitations and provide an outlook on possible future developments in nonviral vector technology.

1. Introduction

Gene therapy is a treatment based on the delivery of genetic material into one person’s cells to prevent or cure diseases, including cancer, genetic diseases, or infectious diseases. As such, gene therapy products are used to either (I) replace a dysfunctional gene, (II) inactivate a disease-causing gene, or (III) introduce a new or modified gene. The number of gene therapy clinical trials increased 8-fold within the last decade (https://www.clinicaltrials.gov/ct2/home, accessed on 22 June 2022). The key step in gene therapy is the safe and efficient delivery of the genetic material to the target tissue/cells. The vehicles used for gene delivery, called vectors, can be either of viral or nonviral origin. Both viral and nonviral vectors can be classified into two groups according to their ability to either integrate into the host genome (e.g., lentiviral vectors [1], sleeping beauty transposon system [2]) or persist as extrachromosomal episomes (e.g., adenoviral vectors [3], S/MAR-based vectors [4], Figure 1).
Viral vectors are, to date, the most efficient delivery systems relying on the natural ability of viruses to deliver their genetic material to recipient cells. Furthermore, vector capsids can be modified to direct vectors to specific target cells or tissues (reviewed in [5]). As such, the majority of ongoing clinical trials are based on viral vectors (www.genetherapynet.com; accessed on 22 June 2022). Nonintegrating viral vector systems, such as adenoviral and adeno-associated virus (AAV) vectors, lack an intrinsic ability to integrate into the host’s genome. Instead, vector genomes remain episomally, thus reducing the risk of insertional mutagenesis. However, these genomes are gradually lost in proliferating cells, as these vectors are usually replication-deficient and, therefore, suitable for disease treatment in post-mitotic tissues [6,7]. Still, gene therapy treatment with these vectors has led to severe adverse effects in clinical studies due to high vector doses of AAV (>1014 vg/kg) [8] or the high immunogenicity of adenoviral vectors [9]. The seeming disadvantage of being highly immunogenic is increasingly employed in developing adenoviral vectors as vaccines [10] and oncolytic vectors [11]. A recent study determined integration frequencies of AAV vectors in hepatocytes and detected chromosomal integrations at surprisingly high frequencies (1–3%) in vivo and in vitro. Notably, most of the inserted AAV sequences were rearranged and accompanied by deletions of genomic sequences at the integration site [12]. These results indicate that integration frequencies and, thus, the genotoxicity of AAV and possibly other viral nonintegrating vectors have been underestimated thus far.
Integrating viral vectors, such as retroviral vectors, are equipped for chromosomal integration and are thus mitotically stable in dividing cells resulting in prolonged expression of the therapeutic genes. However, vector insertions occur preferably in actively transcribed genes and have led in the past to genotoxic side effects resulting in the development of leukaemia [13,14,15]. As a consequence, self-inactivating (SIN) lentiviral vectors have been developed that show decreased genotoxicity [16,17,18].
So-called viral plasmid replicon-based vectors represent an intermediate between viral and nonviral vectors. Viral plasmid replicon vectors are typically delivered as naked DNA but possess the ability to replicate in the target cell’s nucleus. In contrast to conventional expression plasmids, viral replicon-based vectors exhibit highly efficient strategies to replicate within the nucleus facilitating stable maintenance and prolonged transgene expression in proliferating cells. Of this vector group, Epstein-Barr virus (EBV) derived plasmid replicons have been extensively studied (Table 1). EBV coordinates replication from different origins of replication in lytic and latent infection, respectively. In lytic infections, binding of the viral origin binding protein (OBP) to the lytic origin of replication (ori-Lyt) initiates a rolling circle amplification of the EBV genome. In latent infections, the EBV genome is maintained in infected cells as an episome with approximately 2–20 copies per cell. Here, viral genome replication relies on the binding of the EBV nuclear antigen 1 (EBNA1) to the oriP [19,20,21]. The oriP is a bipartite DNA sequence consisting of a family of repeats (FR) element, harbouring 20 high-affinity binding sites for EBNA1, and a dyad symmetry (DS) element, harbouring four low-affinity binding sites for EBNA1 [22]. The replication of EBV replicons is integrated into the host’s cell cycle by the DS element, whereas the FR element facilitates stable segregation (Figure 2A). Plasmids carrying the DS element and lacking the FR replicate in an EBNA1-dependent fashion but are not stably maintained [23,24]. Smaller versions of the EBV genome lacking the lytic genes have been constructed for gene therapy approaches. However, these mini-EBVs can still transform human B-cells due to the expression of latent genes [25,26] and the EBNA1 itself [27]. Since successful replication and maintenance of EBV episomes is always dependent on oriP/EBNA1, the development of such oriP/EBNA1-only plasmids was a promising attempt at a safe EBV-derived gene vector design [28]. However, the above-mentioned safety concerns, along with observations that EBNA1 might mimic the functions of the cellular high mobility group protein HMG1a [29,30,31], led subsequently to the substitution of viral components of the oriP/EBNA1 system with cellular elements. In principle, these substitutions can be introduced in cis and in trans and will be discussed in the paragraph Mitotic Stability.
Nonviral vectors are considered advantageous due to their safety profile, simplicity, and large-scale producibility. Hence, 20% of all ongoing clinical trials employ nonviral gene transfer. The family of nonviral vectors can be grouped into integrating and episomal vectors. The sleeping beauty (SB) transposase system represents an extensively studied integrating, nonviral vector system (Table 1). Belonging to the Tc1/mariner superfamily of class II transposable elements (TE), SB has been reconstructed from fossil DNA sequences of fish genomes [32]. Tc1/mariner TE transposition does not rely on an RNA intermediate but occurs directly through DNA instead [33]. The SB transposon encodes for a single gene, the transposase, which enzymatically catalyses the transposition reaction, flanked by transposon-specific terminal inverted repeats (TIR). Expression and binding of the transposase to the TIR induces DNA double-strand breaks (DSBs) at both ends, thus releasing the transposon. Mediated by DNA-repair mechanisms, the DNA-transposase complex then integrates at suitable target sites [34,35]. For gene therapy applications, this system has been modified such that both components of the SB transposase system, the transposase and TIRs, are delivered in trans. The transposase gene was replaced with a sequence of interest, e.g., a therapeutic gene, flanked by the TIRs. The transposase itself is provided on a separate expression plasmid [36], as mRNA [37] or as recombinant protein [38] (Figure 2B). Thus, the SB system facilitates transgene insertion and sustained gene expression in the recipient cell.
However, in their simplest form, nonviral vectors are DNA plasmids, often referred to as nonviral episomes, and are among the three most frequently used vectors in clinical trials (www.genetherapynet.com; accessed on 22 June 2022). Nonetheless, these plasmid-based vector technologies usually suffer from inefficient and untargeted delivery (Table 1). Further, most nonintegrating nonviral vectors are not mitotically stable, resulting in the loss of vector molecules as a consequence of cell division and/or degradation. Additionally, transgene silencing, a phenomenon linked to CpG methylation, impedes stable transgene expression [39,40,41]. Therefore, plasmid-based vectors must be equipped with cis-acting sequences to ensure mitotic stability and, thus, prolonged transgene expression.
The characterisation of mapped mammalian origins of replications revealed that so-called scaffold/matrix attachment regions (S/MAR) were often found in close vicinity to replication origins. The insertion of a S/MAR sequence, derived from the human interferon-β gene cluster, into a conventional expression plasmid resulted in the first nonviral vectors shown to replicate autonomously in a variety of different cell lines, including primary cells [42,43]. For episomal replication and maintenance of S/MAR-based episomes, transcription of the S/MAR element linked to an expression cassette is mandatory (Figure 2C). Deleting promoter and/or transcription units or inserting transcription termination signals between transgene and S/MAR led to integration or episome loss after transfection [44]. Transfected in mammalian cells, these S/MAR-based episomes replicate at low copy numbers (1–10 copies per cell) in a once-per-cell cycle manner [45,46]. They are stably maintained in the absence of selection pressure for basically unlimited time [42]. Since these first observations, S/MAR-based episomes have been extensively studied and optimised for their use in gene therapeutic interventions.
An obvious vector for stable large genomic gene loci could be an artificial minichromosome. With the first construction in 1997 [47], human artificial chromosomes (HACs) offered an alternative approach to address the complex requirements of an ideal vector (Table 1). HACs are mitotically and meiotically stable, nonessential, additional chromosomes. They offer a large cloning capacity allowing the insertion of entire genomic loci harbouring all regulatory elements that enable mimicking of the regular pattern of endogenous gene expression. However, for correct replication and segregation during mitosis, the presence of an origin of replication and a centromere are mandatory. Being crucial for the attachment of chromosomes to the mitotic spindle and their subsequent segregation, the centromere is the most important structural feature of HACs. Normal human centromeres contain large arrays of alphoid satellite DNA (α-DNA) that consists of tandem repeats of a 170 bp monomer arranged in highly ordered repeats [48,49]. While only α-DNA efficiently forms centromeric structures after being delivered to mammalian cells, not every α-DNA can form active centromeres [47,50]. It further appeared that functional centromeres require a minimum size of approx. 100 kb [51], with the exception of α-DNA from chromosome 21, which could vary between 50 and 100 kb [52]. As exogenous minichromosomes, HACs are generated using either a “top-down” (engineered chromosomes) or a “bottom-up” (de novo artificial chromosomes) approach. De novo HACs are generated in the recipient cells by transfecting a vector equipped with the essential centromeric chromosomal sequences [47] and subsequent recombination [53]. Using this approach, HACs establish as circular chromosomes and thus circumvent the need for telomeric sequences. The delivery of HACs usually relies on polyamines and lipofection, followed by microcell-mediated chromosome transfer (MMCT). However, these techniques share a limited efficiency. As such, different delivery strategies were explored, utilising either the fusion machinery of viruses [54] or virus-based amplicons [55]. While the development and improvement of HACs experienced significant progress within the last decade [56,57,58,59], their design and construction require particular considerations and utilities that differ from those needed for nonviral vectors. As these have been comprehensively reviewed elsewhere [60,61,62], we will focus in this review on the three nonviral vector systems EBV-based replicons, Sleeping Beauty, and S/MAR-based episomes.
As outlined above, long-term nuclear maintenance and stable transgene expression in therapeutically relevant cell types and tissues is a major challenge in nonviral vector development. Accordingly, significant efforts have been made to address these hurdles. In this review, we will discuss recent progress in nonviral vector development regarding (I) transgene expression, (II) mitotic stability, and (III) nuclear localization.

2. Transgene Expression

The stable and controlled expression of a therapeutic gene independent of the potency of the target cell to proliferate is critical for lasting therapeutic success, particularly in gene replacement approaches. Thus, one major challenge is to achieve stable long-term transgene expression. The choice of the transgene promoter type has the most fundamental impact on transgene expression levels. Most promoters used in viral and non-viral vectors are variants of viral or cellular constitutive promoters and exhibit distinct advantages. The cytomegalovirus (CMV) early enhancer/promoter facilitates a high level of transgene expression in various cell types [63,64,65]. Thus, this promoter is most commonly used in nonviral vectors [42,43,44,66]. Linking the CMV promoter with an intron, placed between promoter and transgene, was shown to additionally increase transgene expression levels in S/MAR-based episomes [67]. However, the CMV promoter is highly prone to silencing over time, resulting in the loss of transgene expression. Although this effect may vary between different cell types [68,69], other constitutive promoters such as the cytomegalovirus early enhancer/human elongation factor-1α (CMV/hEF-1α) and cytomegalovirus early enhancer/chicken β-actin (CAG) promoters have been proven to mediate stable transgene expression for long periods of time. Replacing the CMV promoter with a CMV/hEF-1α promoter significantly increased and prolonged transgene expression in vitro and in vivo [70]. In contrast to the CMV promoter, the CAG promoter is less prone to cytosine methylation [71]. It has been successfully used for the generation of induced pluripotent stem cells (iPSCs) [72] and the genetic engineering of stem cells [73].
High-level transgene expression may have toxic effects depending on the nature of the transgene and the recipient cell or tissues [74]. Here, low-yield endogenous promoters such as the human 3-phosphoglycerate kinase (PGK) or hEF-1α promoters have been successfully used to stably express transgenes in patient-derived primary cells [75] and stem cells [76]. However, some gene therapy applications require a tissue-restricted expression of the transgene. While viral vectors naturally have a limited tropism or can be modified to target specific tissues [5], nonviral vectors are usually delivered as naked DNA molecules. Here, a restricted transgene expression can be achieved by using tissue-specific promoters. For example, replacing the constitutive CMV promoter with either the liver-specific human α anti-trypsin (hAAT) promoter, the muscle-specific smooth muscle 22 (SM22) promoter, or the tumour-specific α-fetoprotein (AFP) promoter led to tissue-specific transgene expression in liver, muscle cells, and AFP-positive carcinoma cells, respectively [77,78].
Compared to strong viral promoters such as CMV, endogenous or tissue-specific promoters are often characterised by lower expression levels. Thus, combining a weak or tissue-specific promoter with enhancer elements mediated strong transgene expression and sustained tissue-specificity [78]. However, due to safety concerns, considerable efforts have been made to substitute viral enhancer sequences with cellular elements that avoid the spread of heterochromatin and hypermethylation. Several different of such protective elements, such as the 5′HS chicken β-globin (cHS4) insulator [79,80,81,82], the D4Z4 insulator [82], matrix attachment regions (MAR) [83,84], and Ubiquitous Chromatin Opening Elements (UCOE) [81,82], have already been exploited for transgene protection in both, viral and non-viral vector systems. The UCOE elements have been identified within the TBP-PSMB1 and HNRPA2B1-CBX3 gene loci. These loci’s dual divergently transcribed promoter regions could confer stable gene expression even from within a centromeric heterochromatic environment [85]. The UCOE derived from the HNRPA2B1-CBX3 locus (A2UCOE) significantly improved transgene expression concerning expression levels and duration [86]. In viral vectors, UCOEs have been shown to mediate both enrichment of permissive histone marks (e.g., H3K4me3) and hypomethylation [87,88]. Inserted upstream of the CMV promoter in the context of a S/MAR-based nonviral episome, the 4 kb-A2UCOE element led to increased and homogeneous transgene expression within a cell population [81]. Likewise, in the context of SB transposon vectors, introducing a 1.5 kb-UCOE element immediately upstream of a CMV promoter led to reduced transgene silencing. Moreover, a core fragment within the 1.5 kb-UCOE was identified using bioinformatic approaches. This core fragment, only 0.86 kb in size, offered effective protection from transcriptional silencing [82].
Other cis-acting elements, such as insulators, have also been successfully used to shield transgene cassettes from silencing. The DNAseI hypersensitive site of the chicken β-globin locus control region (cHS4) exhibits classical insulator properties. The cHS4 insulator has been shown to block enhancer activity and functions as a barrier. The cHS4 recruits several DNA binding proteins, of which the CCCTC-binding factor (CTCF) has been linked to its enhancer-blocking function [89]. Other proteins, such as the Poly(ADP-ribose) Polymerase-1 (PARP-1) and the upstream stimulatory factors 1 and 2 (USF1 and 2), are presumably involved in the insulator’s barrier activity [90,91]. Besides its enhancer-blocking and barrier activities, evidence suggests that the cHS4 insulator is also involved in nuclear organisation and can anchor active domains to specific subnuclear compartments [92,93]. In SB transposon vector context, cHS4 insulator sequences flanking the transgene expression cassette increased expression level and reduced interclonal variation in CHO cells [82] and murine embryonal carcinoma cells [79].
In summary, these results clearly show that a stable and prolonged transgene expression significantly depends on the choice of promoter and subsidiary cis-acting elements. Here, the promoter and, if necessary, a respective cis-acting element should be carefully chosen depending on the desired application, targeted tissue and desired transgene expression level.

3. Mitotic Stability

3.1. Viral Elements That Confer Mitotic Stability

Stable maintenance of nonintegrating vectors requires both replication and segregation of the vector genomes synchronous to the cell cycle. While for integrating nonviral vector systems, such as SB transposons, stable vector genome maintenance is facilitated via integration, episomal vectors rely on certain genomic elements facilitating replication and segregation. As outlined above, the EBV-derived oriP/EBNA1 vector system relies on viral elements that facilitate vector genome replication (DS element) and segregation (FR element). However, safety concerns led to attempts to substitute some of these viral components with cellular elements, both in cis and in trans. It has been shown that replication of the EBV genome can be initiated at sites other than the DS-element [94]. Consequently, it is dispensable for stable maintenance of the whole EBV genome [95]. For Mini-EBV genomes, Ott et al. demonstrated that replication is initiated primarily in the close vicinity of the DS-element when its position has been altered. Moreover, when completely deleted, replication starts throughout the mini-EBV genome at weaker replication origins [96]. However, for oriP-only plasmids, the DS element is mandatory for stable episome maintenance and thus needs to be replaced. Cellular sequences with replication competence, termed origins of replications, are thus favourable for substitution. Yet, analyses of mapped mammalian origins of replications have not revealed any consensus sequences. Instead, AT-rich sequences, CpG-islands, bent DNA and the presence of S/MAR sequences are common structural features within origins [97]. Using a so-called origin-trapping assay, Gerhardt et al. isolated potential origins bound by Orc2 by ChIP-assay. Orc2 is a protein of the replication machinery involved in assembling the pre-replication complex (preRC). Subsequently, these sequences were cloned in oriP plasmids lacking the DS element and analysed for their replication competence. They demonstrated that the DS element could be replaced with cellular sequences, and, remarkably, these sequences were all located at a distance of 7 to 10 kb from matrix attachment regions. Moreover, some identified sequences could facilitate replication independent of EBNA1 [98].
Structural analyses of the EBNA1 protein identified the N-terminus being critical for chromosome binding while the C-terminus enables the accumulation of replicated oriP-plasmids in the nucleus. High mobility group (HMG) proteins contain AT-hook motifs. Especially one member, HMGA1, binds highly specific to the minor groove of AT stretches, thus inducing conformational changes [99]. Such AT-hooks have also been identified within the N-terminus of EBNA1 [30]. Replacing the EBNA1 N-terminus with cellular proteins revealed that amino acids 1–90 of high mobility group-I protein and histone 1 could substitute the EBNA1 N-terminus and mediate association with mitotic chromosomes, nuclear retention, and long-term maintenance [29]. In a subsequent study, an HMGA1:EBNA1 fusion protein was employed to analyse the effects of HMGA1 on episome replication. This fusion protein mediated replication of an oriP-plasmid in an Orc-dependent manner. In fact, HMGA1:EBNA1 binds specifically to Orc, thus generating a functional origin of replication. Fusing HMGA1 to tetracycline repressor (TetR) could facilitate plasmid replication when the DS-element was replaced by four tet-operator (tetO) sites [31]. Using a similar approach, Pich et al. constructed a conditionally replicating oriP-vector. The 20 EBNA1 binding sites of FR were replaced with 20 TetR binding sites. Accordingly, the EBNA1 N-terminus was fused with TetR. When doxycycline is added to the system, the TetR domains undergo conformational changes and can no longer bind to the tetO sites. In the absence of doxycycline, binding of EBNA1:TetR to the modified oriP facilitated DNA replication and episome maintenance [100]. In a combinatorial approach [31], the FR and DS elements were replaced with tetO sites. This artificial oriP-vector was maintained episomally and became almost undetectable when doxycycline was added to the cells [100].

3.2. Chromosomal Elements That Confer Mitotic Stability

S/MAR-based vectors were the first vectors lacking viral elements. They were first introduced in the late 1990s and rely on a S/MAR-sequence, derived from the β-interferon (IFNβ) gene locus, linked to a transgene expression cassette [42]. These episomes establish with low copy numbers (1–10 copies per cell) in the nucleus [46] and are mitotically stable for basically unlimited time [42]. The S/MAR-episomes are associated with the nuclear matrix via the SAF-A protein [101,102]. The replication of S/MAR-based episomes occurs in a once-per-cell cycle manner during the early S-phase, and proteins of the preRC have been shown to assemble at various sites throughout the vector [45]. Since the mitotic stability of S/MAR-based episomes relies on an ongoing transcription into the S/MAR [44], switching off transgene expression would lead to episome loss. Using a doxycycline-inducible promoter (TetOn), Rupprecht et al. demonstrated in a proof-of-concept study that the removal of doxycycline resulted in a gradual loss of vector molecules in vitro and in vivo [103]. Since the efficiency with which S/MAR-based episomes are established in the nucleus is comparably low [46], various attempts focused on identifying cis-acting elements supporting establishment efficiency. While the original INFβ-S/MAR spans 2 kb, Jenke et al. designed a minimal S/MAR module (155 bp) that comprises the core unwinding element of the upstream IFNβ-S/MAR. Inserted as a dimer or tetramer downstream of the transgene, only the tetramer facilitated episomal maintenance, while the dimer integrated into the host’s genome [102]. Another approach is based on the observation that the cHS4 insulator anchors active domains to specific subnuclear compartments [92,93]. Thus, the 1.2 kb cHS4 was inserted downstream and in a reverse orientation of the S/MAR. These S/MAR-cHS4 episomes showed up to 2.5-fold increased establishment efficiencies compared to S/MAR-based episomes without the cHS4. In the episome context, the binding of CTCF was mapped exclusively to the 5′-region of cHS4 and is thought to act as an additional anchor for these episomes [81].
Although the original S/MAR-based episome [42] was shown to establish in primary CD34+ cells, only 1% of progeny cells stably maintained the S/MAR-based episome [43]. To overcome this limitation, a bona fide mammalian origin of replication, the β-globin replicator, was employed. Being the replication initiation region (IR) of the β-globin gene locus, this element was inserted upstream of the promoter. The IR-S/MAR vectors were established episomally in CD34+ cells with significantly increased frequencies [104]. Additionally equipped with an artificial transcription activator that binds specifically to the Aγ-globin promoter, Stavrou et al. demonstrated the efficient induction of γ-globin transcription. This development may contribute to effective gene therapy for treating β-thalassaemia and sickle cell disease [105]. While S/MAR sequences linked to an expression cassette facilitate episomal maintenance of plasmids, they have also been studied for their effect on SB transposition activity. Inserting a S/MAR sequence derived from the human β-globin locus upstream of the transposase expression cassette improved transposition frequency and, thus, long-term transgene expression [83].
Another structural feature tightly linked with DNA replication is the so-called G-quadruplex (G4) structure. These secondary DNA structures are constructed from guanine-quartet-building blocks forming square-planar assemblies of four Hoogsteen-bonded guanine bases [106]. Potential G4-forming sequences (pG4) are not distributed randomly throughout the genome but tend to accumulate in promoter regions and 5′-and 3′-untranslated regions of mRNA [107]. Origin G-rich Repeated Elements (OGRE) are considered pG4s and are present in more than 60% of origins in fly, mouse, and human cells [108,109]. They are located upstream of the replication initiation site, compatible with the position of the preRC [110]. As such, plasmids equipped with an OGRE element could replicate in HEK293 cells that express EBNA1 almost as efficiently as plasmids containing the oriP. Accordingly, the deletion of the OGRE element strongly reduces its replication efficiency [111]. This is particularly interesting since the EBNA1-mediated recruitment of Orc to the DS element has been linked to an RNA-dependent mechanism [112]. The EBNA1 linking regions 1 and 2 (LR1 and 2) have been shown to bind G-rich RNA that is predicted to form G4 structures. G4-interacting drugs, such as BRACO-19, consequently disrupted the interaction between EBNA1 and Orc, inhibited EBNA1-mediated replication, and interfered with the ability of EBNA1 to tether to metaphase chromosomes [113]. The introduction of a pG4 oligonucleotide into S/MAR-based vectors immediately upstream of the promoter led to stable maintenance even when the S/MAR element was deleted (Hagedorn and Lipps, unpublished results). These observations suggest that episomal replication and maintenance might also be mediated through G4 structures.

3.3. Modifications to Minimize Effects of Innate Immune Responses

In eukaryotes, CpG dinucleotides appear with a frequency below their statistical expectations and are often methylated to 5-methyl-cytosine (mCpG). However, some genomic regions, termed CpG islands, contain atypical high frequencies of CpG dinucleotides that generally lack methylation [41]. In bacterial genomes, CpG dinucleotides are represented with a frequency reflecting their statistical expectation and are usually unmethylated [114]. Thus, mechanisms of the human innate immune response to differentiate between bacterial and own DNA rely on CpG dinucleotide occurrence and methylation. The toll-like receptor 9 (TLR9) interacts with unmethylated CpG dinucleotides in DNA molecules and subsequently activates inflammatory downstream signalling involving MyD88. This increases NFkB and AP-1 expression, resulting in the production of inflammatory cytokines [115,116] and might lead to vector loss and/or transgene silencing. Remarkably, one single CpG dinucleotide in a DNA vector triggered inflammatory responses in vivo [117]. These understandings led to the development of CpG-depleted DNA plasmids, thus minimising silencing phenomena and undesired stimulation of the innate immune system [117,118]. Consistently, S/MAR-based episomes with reduced CpG content in the backbone have been constructed. Further, the CpG-depleted vector backbone was combined with a hCMVe/EF1a promoter that is less prone to silencing [117]. This novel class of S/MAR-based episomes, referred to as pEPito, showed enhanced transgene expression and increased colony formation efficiencies, indicating improved mitotic stability [70]. Accordingly, episomal vectors lacking any residual bacterial sequences, so-called minicircles, should be able to avoid innate immune responses [119]. The first S/MAR-based minicircles were generated using the Flp-recombinase technique producing two circular units: a miniplasmid containing the bacterial vector elements and a promoter-transgene-S/MAR minicircle. Like the parental S/MAR-based episomes, these minicircles are mitotically stable and showed enhanced transgene expression and establishment efficiencies in vitro and in vivo [77,120].
Nevertheless, the production of minicircles is a time-consuming process that involves intramolecular recombination. Antibiotic-free (AF) minimally sized vectors have been developed as an alternative. The AF selectable vectors either rely on RNA-interference [121] or amber-suppressor tRNAs [122]. Equipped with the IFNβ-S/MAR sequence, these AF S/MAR vectors produced more robust transgene expression and showed increased establishment efficiencies in dividing cells [123]. Combined with the liver-specific HCRHPi promoter that contains the apolipoprotein hepatic control region (ApoE-HCR), increased and stable transgene expression restricted to liver cells was shown [124]. Yet, aside from a proof-of-concept, AF S/MAR vectors have been reported for their application in the genetic modification of primary pancreatic cancer cells. Besides the sustained supplementation of a tumour suppressor, in terms of genotoxicity, only minimal impact on the target cell genome was detected [123].
Similar modifications have also been applied to SB vectors. The success of SB-mediated gene therapy strongly depends on the efficient delivery of the plasmid DNAs and subsequent transgene integration rates. Therefore, applying the minicircle technique to the SB transposase vector system is cogent. Converting the SB transposon system into a minicircle combined with the hyperactive SB100x enhanced integration properties in human haematopoietic stem cells (HPSCs) [36]. Further, delivery of both SB100x transposase and transgene using an AF minimally sized plasmid significantly increased transgene integration efficiencies [125]. This is particularly interesting since the number of integrated transposon copies correlates linearly with transgene expression [126].
Recently, the group of Harbottle fundamentally modified S/MAR-based episomes, developing a new generation of S/MAR vectors termed nano-S/MARs (nS/MAR) [73,75]. The prokaryotic vector backbone was depleted for CpG sites, and inserting an RNA OUT element [121,123] enabled antibiotic-free selection. Further, the IFNβ-S/MAR was replaced with a more compact version isolated from the apolipoprotein B (ApoB) gene cluster. Stretches of ATTA-ATTTA are the predominant motif of matrix attachment sites and are recognised by homeodomain proteins [127]. The ApoB S/MAR comprises a stretch of 555 bp, almost entirely composed of these motifs. Since the mitotic stability of S/MAR-based vectors relies on transcription running into the S/MAR sequence, a long mRNA containing the S/MAR sequence is transcribed [46]. However, AU-rich elements within the untranslated regions of mRNAs serve as signals for rapid degradation mediated by the human antigen R [128]. This would impair transgene expression while the vectors are still episomally maintained. In a very sophisticated approach, Bozza et al. addressed this problem by introducing splice sites flanking the S/MAR [75]. The nS/MAR were maintained episomally, displayed highly stable transgene expression, and were established most efficiently compared to previous vector generations. While this approach was successful for intron-less transgenes, it should be noted that the use of intron-containing transgenes (e.g., whole genomic loci) might yield unfavourable splicing variants and, thus, might not be feasible. The exact mechanism underlying the transcription-coupled mitotic stability of S/MAR vectors is still not fully understood. However, these results indicate that an RNA-dependent mechanism might be involved, as described for EBNA1 (90,91).
Moreover, nS/MAR vectors were employed to generate CAR-T cells. In a side-by-side comparison with a lentiviral vector, nS/MAR vectors had only a minimal impact on human T-cell gene expression. They were shown to be suitable for large-scale GMP-compatible production of CAR-T cells. Finally, nS/MAR engineered T-Cells mediated tumour killing in vivo with an efficacy comparable to CAR-T cells modified with a lentiviral vector [75]. In the same group, nS/MAR vectors were exploited for the genetic modification of pluripotent stem cells (PSCs). Remarkably, the nS/MAR vectors did not impair the pluripotency of the modified PSCs and persisted during reprogramming and differentiation, both in vitro and in vivo [73].
Interestingly, all nonintegrating, nonviral vectors and minicircles are maintained with a similar copy number in the nucleus [46,75,81,104], suggesting stringent copy number control in the recipient cells. Thus, epigenetic features, such as chromatin structure and nuclear localisation, seem to be essential in regulating the establishment of nonviral episomes.

4. Nuclear Localisation

Within a eukaryotic nucleus, the genome is compartmentalised. Chromosomal regions with similar functional properties (e.g., heterochromatin and euchromatin) often cluster together, forming distinct compartments referred to as the A- and B-compartments [129]. Further, during interphase, chromosomes are partitioned into topologically associated domains (TADs). These domains are characterised by intensive internal chromatin interactions and fewer contacts to neighbouring regions [130]. Today it is widely accepted that the spatial organisation of the genome and transcriptional activity are tightly liked [131,132].
The stable establishment of S/MAR-based vectors in the nucleus is a rare event. Only 1–5% of all initially transfected cells stably establish episomal vectors as autonomous replicons [81,102]. The episomal establishment is considered a stochastic process that presumably depends on the nuclear compartment a vector molecule reaches after transfection. However, the underlying mechanisms are still poorly understood. The nuclear localisation of S/MAR-based episomes has been extensively studied using fluorescence in situ hybridisation and immunofluorescence. Established episomes co-localised in the nucleus with chromosomal domains that harboured active histone marks, such as acetylated H3K9 and K14, associated with active promoters and enhancers [46,133]. Further, episomes were found in close proximity to nuclear speckles and early replication foci. Neither an association with repressive chromatin marks nor specific chromosomes was detected [46]. Using chromatin-immunoprecipitation (ChIP) assay, Rupprecht et al. monitored the association of S/MAR-based vectors with certain histone marks throughout the cell cycle. In S-phase, the chromatin structure of S/MAR episomes was marked by H3K4me1 and H3K4me3. Both histone modifications are associated with transcriptionally competent chromatin [134,135] and were predominantly enriched in the 3′-region of the S/MAR. Interestingly, during mitoses, most histone modifications analysed in this study were depleted [136].
As different as EBV-derived and S/MAR-based vectors are regarding the mechanisms facilitating their mitotic stability, they have much in common regarding their nuclear localisation. The establishment process of EBV-derived vectors relies on the efficacy of initiating DNA replication [137]; a prerequisite for efficient replication is the sufficient tethering of oriP-plasmids to chromosomes. However, tethering to specific chromosomes has not been described [138]. Within the nucleus, EBV-genomes were described to co-localise with chromatin domains enriched for active histone modifications (H3K4me3 and H3K9ac) but not with nuclear speckles. Strikingly, enrichment of H3K4me3 histone modifications was detected within the oriP of the EBV genomes [137]. This is particularly interesting as a similar enrichment was detected for the S/MAR sequence [136], being the functional unit of S/MAR-based vectors.
More recently, using circular chromosome conformation capturing (4C), the genomic contact sites of three different S/MAR-based vectors were mapped. As previous results indicated, all S/MAR-based vectors interacted with genomic loci enriched for histone marks of open chromatin (e.g., H3K36me3, H3K27ac, H3K4me2) and markers associated with replication initiation (H3K79me2). Further, S/MAR-based vectors preferentially interacted with promoters and transcription start sites. Accordingly, within the genomic contact sites, repressive chromatin marks and transcription end sites were diminished [67]. Even though the epigenetic signature of the chromosomal contact sites was coherent, S/MAR-based vectors displayed an individual chromosomal contact pattern depending on inserted cis-acting elements. While the parental S/MAR-based vector exhibited few and non-clustered contact sites, those of an episome harbouring the cHS4 insulator downstream of the S/MAR [81] appeared clustered to distinct chromosomal loci. In contrast, genomic contact sites of an intron-carrying S/MAR-based episome were more frequent and evenly scattered throughout the genome [67]. It is conceivable that the distinct pattern of genomic contact sites reflects co-transcribed sequences in specialised transcription factories [139,140,141]. Thus, these results indicate that the genetic composition of an episome influences its nuclear localisation and the contact sites it preferentially associates with. This knowledge might lead in the future to the design of directed nonviral episomes, a perspective that has already been realised for SB vector systems.
High-resolution, genome-wide mapping of SB integration sites revealed a close-to-random insertion profile concerning genes and chromosomes [126,142]. A side-by-side comparison of SB, MLV retrovirus integration sites, piggyBac (PB) transposase, and HIV lentivirus revealed unequal biases across the four systems concerning integration in genes previously linked to genotoxicity. While PB, MLV and HIV were preferentially inserted into expressed genes, SB revealed a close-to-random insertion profile, supporting the relative safety of this vector system among integrating vectors [142]. A similar insertion profile has been detected for a sandwich version of SB (SA) that carries two complete transposon elements in head-to-head orientation, flanking the expression cassettes [143]. Again, the SA vector system displayed a close-to-random insertion profile with a slight overrepresentation of repetitive elements, such as satellites, LINS and small RNA genes [126]. Of note, the relatively safe SB insertion profile is independent of its delivery. The insertion profile remained close-to-random when delivered with adenoviral or lentiviral vector systems [144,145,146,147]. However, even with a close-to-random insertion profile, the interactions between integrated transgenes and genomic insertion loci are hard to predict. The position of the transgene within the genome affects not only transgene expression but also, owing to circumstances, endogenous gene expression [14,148,149,150]. Thus, the possibility of a directed insertion would be most desirable. First attempts employed custom DNA binding domains (DBD) engineered to specifically bind to a genomic DNA sequence of choice. Mainly two strategies were applied: direct fusion of such engineered DNA binding protein (DBPs) to the SB transposase or fusing the DBPs to adaptor proteins that interact with the transposase or transposon [151,152,153]. A recent study combined a catalytically inactive Cas9 protein (dCas9) with the SB100x transposase or an N-terminal fragment (N57) encompassing the DNA-binding and dimerisation functions. By generating guide RNAs (sgRNAs) to target the HPRT gene and AluY, directed SB-mediated insertion could be analysed. While enrichment for dCas9-N57 targeting AluY was relatively weak, enrichment dCas9-SB100x was more pronounced and occurred in the vicinity of sites specified by the sgRNA. However, no insertion at the HPRT locus was detected within 5 kb in either direction. It is speculated that either targeting a single-copy locus is not possible with this system or the number of insertion sites was too low to provide the necessary resolution for mapping [154]. However, despite these promising attempts, targeted insertion approaches still face some hurdles. The direct fusion approach often led to reduced transposase activity, and targeting efficiency, in general, was relatively low compared to untargeted background integration (reviewed in [155]).
As mentioned above, a significant safety concern of integrating vector systems is the possibility of genotoxicity. However, the close vicinity of episomal vectors to actively transcribed chromatin regions raises the question of whether the pure presence of a transcriptionally active yet episomal unit may impact host gene expression. This concern is supported by the observations that transcription is a major force in shaping the spatial organisation of the human genome (reviewed in [131,132]). Consequently, careful transcriptome analyses of cells modified with S/MAR-based vectors were performed. In established cell lines and primary cells, only minimal impact on the host cell’s endogenous gene expression was monitored [67,73,75]. This minimal impact was further reduced when next-generation nS/MAR vectors were used [73], which are depleted for CpG sites and dispensable prokaryotic sequences. Gene ontology analyses did not reveal specific genes deregulated by S/MAR-based vectors, suggesting that the gene groups being up- or down-regulated are cell-type- or transgene-specific [67,73,75]. However, as an analogue to the genotoxicity of integrating vectors, it is yet unknown whether episomes tethered to chromosomes impact the genomic stability of the cell. As such, monitoring the incidence of chromosomal rearrangements and aneuploidy in cells stably maintaining episomal vectors compared to unmodified cells would give first hints on the existence of episomal mediated “chromotoxicity” and should be addressed in future studies.

5. Nucleic Acid Delivery

One common theme for all the diverse systems mentioned above is the roadblock of specific and efficient delivery to target cells. Delivery vectors that are derived from virus capsids and/or incorporate viral functions for, e.g., endosomal escape and nuclear import, typically exhibit reasonable gene transfer efficiencies in vitro but often suffer from immunogenicity, toxicity, unwanted sequestration, and mistargeting in vivo. Nevertheless, several technologies have been developed during the last decades to improve specific targeting and/or shielding of such vectors. This research area is out of the focus of this review, and we would like to (subjectively and exemplarily) redirect the reader to [156,157,158].
One delivery vector, however, deserves explicit mention here since it has been used in presumably hundreds of millions of humans world wide with distinct success: the so-called lipid nanoparticle (LNP). It has been used in slightly different forms as the delivery vehicle for mRNA-based vaccines against COVID-19 [159,160]. Lipid nanoparticles form lipid envelopes that compact and protect nucleic acids for delivery [161]. The use of different lipid components influences the stability and fusion properties of these lipid nanoparticles. Importantly, by using shielding reagents, protein and non-protein targeting ligands, and different bioresponsive chemical bonds that enable environment-specific dynamics, very specific delivery vectors can be tailored [162]. Thus, it appears likely that once the genetic material can be successfully delivered to the nucleus, the systems discussed in this review might be delivered both in vitro and in vivo by optimized lipid nanoparticles.

6. Discussion and Outlook

Gene therapy treatments are complex procedures and are still associated with some risks. Given the predominance of delivery efficiency, most approved gene therapy products are based on viral vectors. Although several challenges remain to be solved, viral vectors bear the intrinsic capability of introducing genetic material into cells with high specificity and efficiency. Thus, viral vectors are suitable for genetic modifications in vitro and in vivo. Nonviral vectors and episomes have also been substantially improved within the last decade. However, of the three vector systems discussed in this review, only the SB vector system is currently being tested in several clinical trials (reviewed in [163]), all focusing on T-cell modifications. Despite many advances and significant improvements, the efficient and specific delivery of these vectors is still a major challenge. Usually, nonviral vectors are delivered by different transfection methods, like lipofection and electroporation, that are comparably inefficient and not applicable for targeted in vivo delivery. Thus, the development of hybrid viral vectors combining the specificity and efficiency of the delivery of viral vectors with the relative safe long-term maintenance of nonviral vectors was a huge step forward in the field. In fact, S/MAR-based episomes, EBV-based replicons and SB have been adopted for viral vectors [144,145,146,147,164,165,166,167,168,169,170]. Combining lentiviral delivery with transposase-mediated transgene integration only slightly reduced the transduction efficiency of lentiviral vectors but resulted in safer integration profiles compared with lentiviral-mediated integration [144,145,146]. The adoption of SB for adenoviral or baculoviral delivery enabled in vivo delivery of SB transposases in the murine and canine liver [169] and mouse eye [168]. Again the integration profile was similar to that observed for nonviral SB vectors [147]. Likewise, the quantitative delivery of oriP/EBNA1 episomes was improved when delivered with adenoviral vectors, resulting in stable long-term transgene expression, both in vivo and in vitro [164,165,166]. An adenoviral-S/MAR hybrid vector was developed to synergise the efficient delivery of adenoviral vectors with the episomal maintenance of S/MAR-based episomes. This hybrid vector system was delivered highly efficient and facilitated episomal maintenance and long-term transgene expression both in vitro and in vivo [170]. Equipping an adeno-associated viral genome with a S/MAR sequence facilitated long-term maintenance of the vector-hybrid genome in mitotic cells [167].
Despite these promising approaches, a literature search (August 2022) revealed that the development of hybrid viral vectors peaked between 1999 and 2016, with approximately 86 publications per year. Still, no significant contributions to the field were found thereafter. The reasons for this stagnation remain speculative. Viral hybrid vectors are rather complex compared to their current nonhybrid counterparts. Thus, the process of adopting nonviral systems to viral delivery and the subsequent production of those hybrid vectors can easily become a highly demanding endeavour. Equipping an AAV vector genome with a S/MAR sequence resulted in episomal maintenance in dividing cells. Yet, the initial establishment efficiency of AAV-S/MAR episomes was approximately ten times lower compared to standard S/MAR-based episomes [167], indicating that additional and system-specific modifications are necessary when adopting nonviral vectors for viral delivery. Another factor that might contribute to the stagnation in the development of hybrid vectors is the lack of comprehensive safety profiles. Nonhybrid vectors have been successfully used in clinical trials, and their safety profiles are extensively studied. However, thus far, only little is known regarding the safety profile of hybrid vectors. For example, adopting S/MAR-based episomes for AAV- or Ad-mediated delivery involves the transformation of an initially circular episome to a linear viral genome. S/MAR sequences confer episomal maintenance when inserted downstream of an expression cassette [44]. However, positioned upstream of an expression cassette, these sequences have also been shown to increase chromosomal integration [83,171,172]. These observations indicate that the safety profile of S/MAR-based vectors is context-specific. Other hybrid systems rely on recombination machinery, often co-delivered using a second vector, to release the nonviral episomes from the viral vector backbone [165,170]. The high recombinase concentrations needed for efficient release and the high vector doses required within a two-vector system bear the risk of recombinase-dependent genotoxicity and vector-dependent immunogenicity (reviewed in [173,174]). Thus, detailed analyses for each particular viral hybrid system are inevitable since the transferability of safety profiles of their nonhybrid counterparts is limited.
Presumably, developing new production systems for viral vectors might foster the engineering of novel vectors combining ideal features from both worlds. As discussed in this review, recent developments in nonviral vector designs rendered S/MAR-based episomes and SB transposases able for their use in ex vivo gene therapy applications. Both vector systems were successfully used for the genetic modification of patient-derived T-cells, and SB is currently involved in respective clinical trials [75,163]. However, for a broader field of applications, hybrid viral vectors may represent a reemerging perspective regarding efficiency and safety.
Although currently approved viral vectors are highly efficient in terms of delivery per se, the lack of specific targeting strategies requires the administration of high vector doses. Moreover, the fact that only a fraction of applied vector particles reach the intended tissue or organ demands the use of strong promoters facilitating high transgene expression level that often exceeds the expression level of the endogenous gene. Especially the administration of high vector doses led in the past and recently to severe immunologic adverse effects and fatalities [8,175,176]. Thus, for future gene therapy approaches, it is imperative to consider targeting strategies to reduce vector doses and the need for high-level transgene expression. Within the last decade, an enduring toolbox for modifying viral capsids has been developed. In addition to genetic approaches, the chemical and genetic-chemical modification of viral capsids allows different de- and re-targeting strategies, enabling not only the targeted delivery of viral vectors but also reduced immunological side and off-target effects [177] (and reviewed in [5]). Thus, complementing the recent progress in the design of nonviral vectors with the versatile toolbox of viral delivery strategies is a valuable contribution to future developments in vector design.

Author Contributions

Conceptualization, F.K. and C.H.; writing—original draft preparation, C.H.; writing—review and editing, F.K. and C.H.; visualization, C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sakuma, T.; Barry, M.A.; Ikeda, Y. Lentiviral Vectors: Basic to Translational. Biochem. J. 2012, 443, 603–618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kebriaei, P.; Izsvák, Z.; Narayanavari, S.A.; Singh, H.; Ivics, Z. Gene Therapy with the Sleeping Beauty Transposon System. Trends Genet. 2017, 33, 852–870. [Google Scholar] [CrossRef] [PubMed]
  3. Bulcha, J.T.; Wang, Y.; Ma, H.; Tai, P.W.L.; Gao, G. Viral Vector Platforms within the Gene Therapy Landscape. Sig. Transduct. Target Ther. 2021, 6, 1–24. [Google Scholar] [CrossRef] [PubMed]
  4. Hagedorn, C.; Wong, S.-P.; Harbottle, R.; Lipps, H.J. Scaffold/Matrix Attached Region-Based Nonviral Episomal Vectors. Hum. Gene Ther. 2011, 22, 915–923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Weklak, D.; Pembaur, D.; Koukou, G.; Jönsson, F.; Hagedorn, C.; Kreppel, F. Genetic and Chemical Capsid Modifications of Adenovirus Vectors to Modulate Vector–Host Interactions. Viruses 2021, 13, 1300. [Google Scholar] [CrossRef] [PubMed]
  6. Russell, S.; Bennett, J.; Wellman, J.A.; Chung, D.C.; Yu, Z.-F.; Tillman, A.; Wittes, J.; Pappas, J.; Elci, O.; McCague, S.; et al. Efficacy and Safety of Voretigene Neparvovec (AAV2-HRPE65v2) in Patients with RPE65-Mediated Inherited Retinal Dystrophy: A Randomised, Controlled, Open-Label, Phase 3 Trial. Lancet 2017, 390, 849–860. [Google Scholar] [CrossRef]
  7. Day, J.W.; Finkel, R.S.; Chiriboga, C.A.; Connolly, A.M.; Crawford, T.O.; Darras, B.T.; Iannaccone, S.T.; Kuntz, N.L.; Peña, L.D.M.; Shieh, P.B.; et al. Onasemnogene Abeparvovec Gene Therapy for Symptomatic Infantile-Onset Spinal Muscular Atrophy in Patients with Two Copies of SMN2 (STR1VE): An Open-Label, Single-Arm, Multicentre, Phase 3 Trial. Lancet Neurol. 2021, 20, 284–293. [Google Scholar] [CrossRef]
  8. High-Dose AAV Gene Therapy Deaths. Nat. Biotechnol. 2020, 38, 910. [CrossRef] [PubMed]
  9. Somanathan, S.; Calcedo, R.; Wilson, J.M. Adenovirus-Antibody Complexes Contributed to Lethal Systemic Inflammation in a Gene Therapy Trial. Mol. Ther. 2020, 28, 784–793. [Google Scholar] [CrossRef] [PubMed]
  10. Folegatti, P.M.; Ewer, K.J.; Aley, P.K.; Angus, B.; Becker, S.; Belij-Rammerstorfer, S.; Bellamy, D.; Bibi, S.; Bittaye, M.; Clutterbuck, E.A.; et al. Safety and Immunogenicity of the ChAdOx1 NCoV-19 Vaccine against SARS-CoV-2: A Preliminary Report of a Phase 1/2, Single-Blind, Randomised Controlled Trial. Lancet 2020, 396, 467–478. [Google Scholar] [CrossRef]
  11. Khuri, F.R.; Nemunaitis, J.; Ganly, I.; Arseneau, J.; Tannock, I.F.; Romel, L.; Gore, M.; Ironside, J.; MacDougall, R.H.; Heise, C.; et al. A Controlled Trial of Intratumoral ONYX-015, a Selectively-Replicating Adenovirus, in Combination with Cisplatin and 5-Fluorouracil in Patients with Recurrent Head and Neck Cancer. Nat. Med. 2000, 6, 879–885. [Google Scholar] [CrossRef] [PubMed]
  12. Dalwadi, D.A.; Calabria, A.; Tiyaboonchai, A.; Posey, J.; Naugler, W.E.; Montini, E.; Grompe, M. AAV Integration in Human Hepatocytes. Mol. Ther. 2021, 29, 2898–2909. [Google Scholar] [CrossRef] [PubMed]
  13. Hacein-Bey-Abina, S.; Garrigue, A.; Wang, G.P.; Soulier, J.; Lim, A.; Morillon, E.; Clappier, E.; Caccavelli, L.; Delabesse, E.; Beldjord, K.; et al. Insertional Oncogenesis in 4 Patients after Retrovirus-Mediated Gene Therapy of SCID-X1. J. Clin. Investig. 2008, 118, 3132–3142. [Google Scholar] [CrossRef] [PubMed]
  14. Hacein-Bey-Abina, S.; Von Kalle, C.; Schmidt, M.; McCormack, M.P.; Wulffraat, N.; Leboulch, P.; Lim, A.; Osborne, C.S.; Pawliuk, R.; Morillon, E.; et al. LMO2-Associated Clonal T Cell Proliferation in Two Patients after Gene Therapy for SCID-X1. Science 2003, 302, 415–419. [Google Scholar] [CrossRef] [PubMed]
  15. Howe, S.J.; Mansour, M.R.; Schwarzwaelder, K.; Bartholomae, C.; Hubank, M.; Kempski, H.; Brugman, M.H.; Pike-Overzet, K.; Chatters, S.J.; De Ridder, D.; et al. Insertional Mutagenesis Combined with Acquired Somatic Mutations Causes Leukemogenesis Following Gene Therapy of SCID-X1 Patients. J. Clin. Investig. 2008, 118, 3143–3150. [Google Scholar] [CrossRef] [PubMed]
  16. Dougherty, J.P.; Temin, H.M. A Promoterless Retroviral Vector Indicates That There Are Sequences in U3 Required for 3’ RNA Processing. Proc. Natl. Acad. Sci. USA 1987, 84, 1197–1201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Hawley, R.G.; Covarrubias, L.; Hawley, T.; Mintz, B. Handicapped Retroviral Vectors Efficiently Transduce Foreign Genes into Hematopoietic Stem Cells. Proc. Natl. Acad. Sci. USA 1987, 84, 2406–2410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Yu, S.F.; von Rüden, T.; Kantoff, P.W.; Garber, C.; Seiberg, M.; Rüther, U.; Anderson, W.F.; Wagner, E.F.; Gilboa, E. Self-Inactivating Retroviral Vectors Designed for Transfer of Whole Genes into Mammalian Cells. Proc. Natl. Acad. Sci. USA 1986, 83, 3194–3198. [Google Scholar] [CrossRef] [Green Version]
  19. Lin, C.L.; Li, H.; Wang, Y.; Zhu, F.X.; Kudchodkar, S.; Yuan, Y. Kaposi’s Sarcoma-Associated Herpesvirus Lytic Origin (Ori-Lyt)-Dependent DNA Replication: Identification of the Ori-Lyt and Association of K8 BZip Protein with the Origin. J. Virol. 2003, 77, 5578–5588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Tsurumi, T.; Fujita, M.; Kudoh, A. Latent and Lytic Epstein-Barr Virus Replication Strategies. Rev. Med. Virol. 2005, 15, 3–15. [Google Scholar] [CrossRef] [PubMed]
  21. Rose, C.; Green, M.; Webber, S.; Kingsley, L.; Day, R.; Watkins, S.; Reyes, J.; Rowe, D. Detection of Epstein-Barr Virus Genomes in Peripheral Blood B Cells from Solid-Organ Transplant Recipients by Fluorescence in Situ Hybridization. J. Clin. Microbiol. 2002, 40, 2533–2544. [Google Scholar] [CrossRef] [Green Version]
  22. Reisman, D.; Yates, J.; Sugden, B. A Putative Origin of Replication of Plasmids Derived from Epstein-Barr Virus Is Composed of Two Cis-Acting Components. Mol. Cell. Biol. 1985, 5, 1822–1832. [Google Scholar]
  23. Shirakata, M.; Hirai, K. Identification of Minimal OriP of Epstein-Barr Virus Required for DNA Replication. J. Biochem. 1998, 123, 175–181. [Google Scholar] [CrossRef] [PubMed]
  24. Yates, J.L.; Camiolo, S.M.; Bashaw, J.M. The Minimal Replicator of Epstein-Barr Virus OriP. J. Virol. 2000, 74, 4512–4522. [Google Scholar] [CrossRef] [PubMed]
  25. Kempkes, B.; Pich, D.; Zeidler, R.; Hammerschmidt, W. Immortalization of Human Primary B Lymphocytes in Vitro with DNA. Proc. Natl. Acad. Sci. USA 1995, 92, 5875–5879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kaye, K.M.; Izumi, K.M.; Li, H.; Johannsen, E.; Davidson, D.; Longnecker, R.; Kieff, E. An Epstein-Barr Virus That Expresses Only the First 231 LMP1 Amino Acids Efficiently Initiates Primary B-Lymphocyte Growth Transformation. J. Virol. 1999, 73, 10525–10530. [Google Scholar] [CrossRef] [Green Version]
  27. Humme, S.; Reisbach, G.; Feederle, R.; Delecluse, H.J.; Bousset, K.; Hammerschmidt, W.; Schepers, A. The EBV Nuclear Antigen 1 (EBNA1) Enhances B Cell Immortalization Several Thousandfold. Proc. Natl. Acad. Sci. USA 2003, 100, 10989–10994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Black, J.; Vos, J.-M. Establishment of an OriP/EBNA1-Based Episomal Vector Transcribing Human Genomic Beta-Globin in Cultured Murine Fibroblasts. Gene Ther. 2002, 9, 1447–1454. [Google Scholar] [CrossRef] [PubMed]
  29. Hung, S.C.; Kang, M.S.; Kieff, E. Maintenance of Epstein-Barr Virus (EBV) OriP-Based Episomes Requires EBV-Encoded Nuclear Antigen-1 Chromosome-Binding Domains, Which Can Be Replaced by High-Mobility Group-I or Histone H1. Proc. Natl. Acad. Sci. USA 2001, 98, 1865–1870. [Google Scholar] [CrossRef] [PubMed]
  30. Sears, J.; Ujihara, M.; Wong, S.; Ott, C.; Middeldorp, J.; Aiyar, A. The Amino Terminus of Epstein-Barr Virus (EBV) Nuclear Antigen 1 Contains AT Hooks That Facilitate the Replication and Partitioning of Latent EBV Genomes by Tethering Them to Cellular Chromosomes. J. Virol. 2004, 78, 11487–11505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Thomae, A.W.; Pich, D.; Brocher, J.; Spindler, M.-P.; Berens, C.; Hock, R.; Hammerschmidt, W.; Schepers, A. Interaction between HMGA1a and the Origin Recognition Complex Creates Site-Specific Replication Origins. Proc. Natl. Acad. Sci. USA 2008, 105, 1692–1697. [Google Scholar] [CrossRef] [PubMed]
  32. Ivics, Z.; Hackett, P.B.; Plasterk, R.H.; Izsvák, Z. Molecular Reconstruction of Sleeping Beauty, a Tc1-like Transposon from Fish, and Its Transposition in Human Cells. Cell 1997, 91, 501–510. [Google Scholar] [CrossRef] [Green Version]
  33. Finnegan, D.J. Eukaryotic Transposable Elements and Genome Evolution. Trends Genet. 1989, 5, 103–107. [Google Scholar] [CrossRef]
  34. Van Luenen, H.G.; Colloms, S.D.; Plasterk, R.H. The Mechanism of Transposition of Tc3 in C. Elegans. Cell 1994, 79, 293–301. [Google Scholar] [CrossRef]
  35. Plasterk, R.H.; Izsvák, Z.; Ivics, Z. Resident Aliens: The Tc1/Mariner Superfamily of Transposable Elements. Trends Genet. 1999, 15, 326–332. [Google Scholar] [CrossRef]
  36. Holstein, M.; Mesa-Nuñez, C.; Miskey, C.; Almarza, E.; Poletti, V.; Schmeer, M.; Grueso, E.; Ordóñez Flores, J.C.; Kobelt, D.; Walther, W.; et al. Efficient Non-Viral Gene Delivery into Human Hematopoietic Stem Cells by Minicircle Sleeping Beauty Transposon Vectors. Mol. Ther. 2018, 26, 1137–1153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Wilber, A.; Frandsen, J.L.; Geurts, J.L.; Largaespada, D.A.; Hackett, P.B.; McIvor, R.S. RNA as a Source of Transposase for Sleeping Beauty-Mediated Gene Insertion and Expression in Somatic Cells and Tissues. Mol. Ther. 2006, 13, 625–630. [Google Scholar] [CrossRef] [PubMed]
  38. Querques, I.; Mades, A.; Zuliani, C.; Miskey, C.; Alb, M.; Grueso, E.; Machwirth, M.; Rausch, T.; Einsele, H.; Ivics, Z.; et al. A Highly Soluble Sleeping Beauty Transposase Improves Control of Gene Insertion. Nat. Biotechnol. 2019, 37, 1502–1512. [Google Scholar] [CrossRef] [PubMed]
  39. Alhaji, S.Y.; Ngai, S.C.; Abdullah, S. Silencing of Transgene Expression in Mammalian Cells by DNA Methylation and Histone Modifications in Gene Therapy Perspective. Biotechnol. Genet. Eng. Rev. 2019, 35, 1–25. [Google Scholar] [CrossRef]
  40. Boyes, J.; Bird, A. Repression of Genes by DNA Methylation Depends on CpG Density and Promoter Strength: Evidence for Involvement of a Methyl-CpG Binding Protein. EMBO J. 1992, 11, 327–333. [Google Scholar] [CrossRef] [PubMed]
  41. Illingworth, R.S.; Bird, A.P. CpG Islands—’A Rough Guide’. FEBS Lett. 2009, 583, 1713–1720. [Google Scholar] [CrossRef] [PubMed]
  42. Piechaczek, C.; Fetzer, C.; Baiker, A.; Bode, J.; Lipps, H.J. A Vector Based on the SV40 Origin of Replication and Chromosomal S/MARs Replicates Episomally in CHO Cells. Nucleic Acids Res. 1999, 27, 426–428. [Google Scholar] [CrossRef] [PubMed]
  43. Papapetrou, E.P.; Ziros, P.G.; Micheva, I.D.; Zoumbos, N.C.; Athanassiadou, A. Gene Transfer into Human Hematopoietic Progenitor Cells with an Episomal Vector Carrying an S/MAR Element. Gene Ther. 2006, 13, 40–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Stehle, I.M.; Scinteie, M.F.; Baiker, A.; Jenke, A.C.; Lipps, H.J. Exploiting a Minimal System to Study the Epigenetic Control of DNA Replication: The Interplay between Transcription and Replication. Chromosome Res. 2003, 11, 413–421. [Google Scholar] [CrossRef] [PubMed]
  45. Schaarschmidt, D.; Baltin, J.; Stehle, I.M.; Lipps, H.J.; Knippers, R. An Episomal Mammalian Replicon: Sequence-Independent Binding of the Origin Recognition Complex. EMBO J. 2004, 23, 191–201. [Google Scholar] [CrossRef] [Green Version]
  46. Stehle, I.M.; Postberg, J.; Rupprecht, S.; Cremer, T.; Jackson, D.A.; Lipps, H.J. Establishment and Mitotic Stability of an Extra-Chromosomal Mammalian Replicon. BMC Cell Biol. 2007, 8, 33. [Google Scholar] [CrossRef] [Green Version]
  47. Harrington, J.J.; Van Bokkelen, G.; Mays, R.W.; Gustashaw, K.; Willard, H.F. Formation of de Novo Centromeres and Construction of First-Generation Human Artificial Microchromosomes. Nat. Genet. 1997, 15, 345–355. [Google Scholar] [CrossRef] [PubMed]
  48. Rosenberg, H.; Singer, M.; Rosenberg, M. Highly Reiterated Sequences of SIMIANSIMIANSIMIANSIMIANSIMIAN. Science 1978, 200, 394–402. [Google Scholar] [CrossRef] [PubMed]
  49. Thayer, R.E.; Singer, M.F.; McCutchan, T.F. Sequence Relationships between Single Repeat Units of Highly Reiterated African Green Monkey DNA. Nucleic Acids Res. 1981, 9, 169–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Mejia, J.E.; Willmott, A.; Levy, E.; Earnshaw, W.C.; Larin, Z. Functional Complementation of a Genetic Deficiency with Human Artificial Chromosomes. Am. J. Hum. Genet. 2001, 69, 315–326. [Google Scholar] [CrossRef] [Green Version]
  51. Yang, J.W.; Pendon, C.; Yang, J.; Haywood, N.; Chand, A.; Brown, W.R. Human Mini-Chromosomes with Minimal Centromeres. Hum. Mol. Genet. 2000, 9, 1891–1902. [Google Scholar] [CrossRef] [PubMed]
  52. Lo, A.W.; Liao, G.C.; Rocchi, M.; Choo, K.H. Extreme Reduction of Chromosome-Specific Alpha-Satellite Array Is Unusually Common in Human Chromosome 21. Genome Res. 1999, 9, 895–908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kotzamanis, G.; Cheung, W.; Abdulrazzak, H.; Perez-Luz, S.; Howe, S.; Cooke, H.; Huxley, C. Construction of Human Artificial Chromosome Vectors by Recombineering. Gene 2005, 351, 29–38. [Google Scholar] [CrossRef] [PubMed]
  54. Navaratnarajah, C.K.; Leonard, V.H.J.; Cattaneo, R. Measles Virus Glycoprotein Complex Assembly, Receptor Attachment, and Cell Entry. Curr. Top. Microbiol. Immunol. 2009, 329, 59–76. [Google Scholar] [CrossRef] [PubMed]
  55. Frenkel, N. The History of the HSV Amplicon: From Naturally Occurring Defective Genomes to Engineered Amplicon Vectors. Curr. Gene Ther. 2006, 6, 277–301. [Google Scholar] [CrossRef] [PubMed]
  56. Ponomartsev, S.V.; Sinenko, S.A.; Skvortsova, E.V.; Liskovykh, M.A.; Voropaev, I.N.; Savina, M.M.; Kuzmin, A.A.; Kuzmina, E.Y.; Kondrashkina, A.M.; Larionov, V.; et al. Human AlphoidtetO Artificial Chromosome as a Gene Therapy Vector for the Developing Hemophilia A Model in Mice. Cells 2020, 9, 879. [Google Scholar] [CrossRef] [Green Version]
  57. Chan, D.Y.; Moralli, D.; Wheatley, L.; Jankowska, J.D.; Monaco, Z.L. Multigene Human Artificial Chromosome Vector Delivery with Herpes Simplex Virus 1 Amplicons. Exp. Cell Res. 2020, 388, 111840. [Google Scholar] [CrossRef]
  58. Kouprina, N.; Earnshaw, W.C.; Masumoto, H.; Larionov, V. A New Generation of Human Artificial Chromosomes for Functional Genomics and Gene Therapy. Cell Mol. Life Sci. 2013, 70, 1135–1148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Kouprina, N.; Petrov, N.; Molina, O.; Liskovykh, M.; Pesenti, E.; Ohzeki, J.-I.; Masumoto, H.; Earnshaw, W.C.; Larionov, V. Human Artificial Chromosome with Regulated Centromere: A Tool for Genome and Cancer Studies. ACS Synth. Biol. 2018, 7, 1974–1989. [Google Scholar] [CrossRef]
  60. Ohzeki, J.-I.; Otake, K.; Masumoto, H. Human Artificial Chromosome: Chromatin Assembly Mechanisms and CENP-B. Exp. Cell Res. 2020, 389, 111900. [Google Scholar] [CrossRef]
  61. Moralli, D.; Monaco, Z.L. Gene Expressing Human Artificial Chromosome Vectors: Advantages and Challenges for Gene Therapy. Exp. Cell Res. 2020, 390, 111931. [Google Scholar] [CrossRef] [PubMed]
  62. Ikeno, M.; Hasegawa, Y. Applications of Bottom-up Human Artificial Chromosomes in Cell Research and Cell Engineering. Exp. Cell Res. 2020, 390, 111793. [Google Scholar] [CrossRef] [PubMed]
  63. Foecking, M.K.; Hofstetter, H. Powerful and Versatile Enhancer-Promoter Unit for Mammalian Expression Vectors. Gene 1986, 45, 101–105. [Google Scholar] [CrossRef]
  64. Xia, W.; Bringmann, P.; McClary, J.; Jones, P.P.; Manzana, W.; Zhu, Y.; Wang, S.; Liu, Y.; Harvey, S.; Madlansacay, M.R.; et al. High Levels of Protein Expression Using Different Mammalian CMV Promoters in Several Cell Lines. Protein Expr. Purif. 2006, 45, 115–124. [Google Scholar] [CrossRef] [PubMed]
  65. Sakurai, F.; Kawabata, K.; Yamaguchi, T.; Hayakawa, T.; Mizuguchi, H. Optimization of Adenovirus Serotype 35 Vectors for Efficient Transduction in Human Hematopoietic Progenitors: Comparison of Promoter Activities. Gene Ther. 2005, 12, 1424–1433. [Google Scholar] [CrossRef] [Green Version]
  66. Manzini, S.; Vargiolu, A.; Stehle, I.M.; Bacci, M.L.; Cerrito, M.G.; Giovannoni, R.; Zannoni, A.; Bianco, M.R.; Forni, M.; Donini, P.; et al. Genetically Modified Pigs Produced with a Nonviral Episomal Vector. Proc. Natl. Acad. Sci. USA 2006, 103, 17672–17677. [Google Scholar] [CrossRef] [Green Version]
  67. Hagedorn, C.; Gogol-Döring, A.; Schreiber, S.; Epplen, J.T.; Lipps, H.J. Genome-Wide Profiling of S/MAR-Based Replicon Contact Sites. Nucleic Acids Res. 2017, 45, 7841–7854. [Google Scholar] [CrossRef] [Green Version]
  68. Qin, J.Y.; Zhang, L.; Clift, K.L.; Hulur, I.; Xiang, A.P.; Ren, B.-Z.; Lahn, B.T. Systematic Comparison of Constitutive Promoters and the Doxycycline-Inducible Promoter. PLoS ONE 2010, 5, e10611. [Google Scholar] [CrossRef] [Green Version]
  69. Chen, C.; Krohn, J.; Bhattacharya, S.; Davies, B. A Comparison of Exogenous Promoter Activity at the ROSA26 Locus Using a ΦiC31 Integrase Mediated Cassette Exchange Approach in Mouse ES Cells. PLoS ONE 2011, 6, e23376. [Google Scholar] [CrossRef] [Green Version]
  70. Haase, R.; Argyros, O.; Wong, S.-P.; Harbottle, R.P.; Lipps, H.J.; Ogris, M.; Magnusson, T.; Vizoso Pinto, M.G.; Haas, J.; Baiker, A. PEPito: A Significantly Improved Non-Viral Episomal Expression Vector for Mammalian Cells. BMC Biotechnol. 2010, 10, 20. [Google Scholar] [CrossRef] [Green Version]
  71. Sawicki, J.A.; Morris, R.J.; Monks, B.; Sakai, K.; Miyazaki, J. A Composite CMV-IE Enhancer/Beta-Actin Promoter Is Ubiquitously Expressed in Mouse Cutaneous Epithelium. Exp. Cell Res. 1998, 244, 367–369. [Google Scholar] [CrossRef] [PubMed]
  72. Drozd, A.M.; Walczak, M.P.; Piaskowski, S.; Stoczynska-Fidelus, E.; Rieske, P.; Grzela, D.P. Generation of Human IPSCs from Cells of Fibroblastic and Epithelial Origin by Means of the OriP/EBNA-1 Episomal Reprogramming System. Stem Cell Res. Ther. 2015, 6, 122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Roig-Merino, A.; Urban, M.; Bozza, M.; Peterson, J.D.; Bullen, L.; Büchler-Schäff, M.; Stäble, S.; van der Hoeven, F.; Müller-Decker, K.; McKay, T.R.; et al. An Episomal DNA Vector Platform for the Persistent Genetic Modification of Pluripotent Stem Cells and Their Differentiated Progeny. Stem Cell Rep. 2022, 17, 143–158. [Google Scholar] [CrossRef] [PubMed]
  74. Khabou, H.; Cordeau, C.; Pacot, L.; Fisson, S.; Dalkara, D. Dosage Thresholds and Influence of Transgene Cassette in Adeno-Associated Virus-Related Toxicity. Hum. Gene Ther. 2018, 29, 1235–1241. [Google Scholar] [CrossRef]
  75. Bozza, M.; Roia, A.D.; Correia, M.P.; Berger, A.; Tuch, A.; Schmidt, A.; Zörnig, I.; Jäger, D.; Schmidt, P.; Harbottle, R.P. A Nonviral, Nonintegrating DNA Nanovector Platform for the Safe, Rapid, and Persistent Manufacture of Recombinant T Cells. Sci. Adv. 2021, 7, eabf1333. [Google Scholar] [CrossRef]
  76. Thyagarajan, B.; Scheyhing, K.; Xue, H.; Fontes, A.; Chesnut, J.; Rao, M.; Lakshmipathy, U. A Single EBV-Based Vector for Stable Episomal Maintenance and Expression of GFP in Human Embryonic Stem Cells. Regen. Med. 2009, 4, 239–250. [Google Scholar] [CrossRef]
  77. Argyros, O.; Wong, S.P.; Fedonidis, C.; Tolmachov, O.; Waddington, S.N.; Howe, S.J.; Niceta, M.; Coutelle, C.; Harbottle, R.P. Development of S/MAR Minicircles for Enhanced and Persistent Transgene Expression in the Mouse Liver. J. Mol. Med. 2011, 89, 515–529. [Google Scholar] [CrossRef]
  78. Haase, R.; Magnusson, T.; Su, B.; Kopp, F.; Wagner, E.; Lipps, H.; Baiker, A.; Ogris, M. Generation of a Tumor- and Tissue-Specific Episomal Non-Viral Vector System. BMC Biotechnol. 2013, 13, 49. [Google Scholar] [CrossRef] [Green Version]
  79. Dalsgaard, T.; Moldt, B.; Sharma, N.; Wolf, G.; Schmitz, A.; Pedersen, F.S.; Mikkelsen, J.G. Shielding of Sleeping Beauty DNA Transposon-Delivered Transgene Cassettes by Heterologous Insulators in Early Embryonal Cells. Mol. Ther. 2009, 17, 121–130. [Google Scholar] [CrossRef]
  80. Arumugam, P.I.; Scholes, J.; Perelman, N.; Xia, P.; Yee, J.-K.; Malik, P. Improved Human Beta-Globin Expression from Self-Inactivating Lentiviral Vectors Carrying the Chicken Hypersensitive Site-4 (CHS4) Insulator Element. Mol. Ther. 2007, 15, 1863–1871. [Google Scholar] [CrossRef]
  81. Hagedorn, C.; Antoniou, M.N.; Lipps, H.J. Genomic Cis-Acting Sequences Improve Expression and Establishment of a Nonviral Vector. Mol. Ther. Nucleic Acids 2013, 2, e118. [Google Scholar] [CrossRef] [PubMed]
  82. Skipper, K.A.; Hollensen, A.K.; Antoniou, M.N.; Mikkelsen, J.G. Sustained Transgene Expression from Sleeping Beauty DNA Transposons Containing a Core Fragment of the HNRPA2B1-CBX3 Ubiquitous Chromatin Opening Element (UCOE). BMC Biotechnol. 2019, 19, 75. [Google Scholar] [CrossRef] [PubMed]
  83. Sjeklocha, L.; Chen, Y.; Daly, M.C.; Steer, C.J.; Kren, B.T. β-Globin Matrix Attachment Region Improves Stable Genomic Expression of the Sleeping Beauty Transposon. J. Cell Biochem. 2011, 112, 2361–2375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Saunders, F.; Sweeney, B.; Antoniou, M.N.; Stephens, P.; Cain, K. Chromatin Function Modifying Elements in an Industrial Antibody Production Platform - Comparison of UCOE, MAR, STAR and CHS4 Elements. PLoS ONE 2015, 10, e0120096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Antoniou, M.; Harland, L.; Mustoe, T.; Williams, S.; Holdstock, J.; Yague, E.; Mulcahy, T.; Griffiths, M.; Edwards, S.; Ioannou, P.A.; et al. Transgenes Encompassing Dual-Promoter CpG Islands from the Human TBP and HNRPA2B1 Loci Are Resistant to Heterochromatin-Mediated Silencing. Genomics 2003, 82, 269–279. [Google Scholar] [CrossRef]
  86. Williams, S.; Mustoe, T.; Mulcahy, T.; Griffiths, M.; Simpson, D.; Antoniou, M.; Irvine, A.; Mountain, A.; Crombie, R. CpG-Island Fragments from the HNRPA2B1/CBX3 Genomic Locus Reduce Silencing and Enhance Transgene Expression from the HCMV Promoter/Enhancer in Mammalian Cells. BMC Biotechnol. 2005, 5, 17. [Google Scholar] [CrossRef] [Green Version]
  87. Müller-Kuller, U.; Ackermann, M.; Kolodziej, S.; Brendel, C.; Fritsch, J.; Lachmann, N.; Kunkel, H.; Lausen, J.; Schambach, A.; Moritz, T.; et al. A Minimal Ubiquitous Chromatin Opening Element (UCOE) Effectively Prevents Silencing of Juxtaposed Heterologous Promoters by Epigenetic Remodeling in Multipotent and Pluripotent Stem Cells. Nucleic Acids Res. 2015, 43, 1577–1592. [Google Scholar] [CrossRef] [Green Version]
  88. Zhang, F.; Santilli, G.; Thrasher, A.J. Characterization of a Core Region in the A2UCOE That Confers Effective Anti-Silencing Activity. Sci. Rep. 2017, 7, 10213. [Google Scholar] [CrossRef] [Green Version]
  89. Bell, A.C.; West, A.G.; Felsenfeld, G. The Protein CTCF Is Required for the Enhancer Blocking Activity of Vertebrate Insulators. Cell 1999, 98, 387–396. [Google Scholar] [CrossRef] [Green Version]
  90. Aker, M.; Bomsztyk, K.; Emery, D.W. Poly(ADP-Ribose) Polymerase-1 (PARP-1) Contributes to the Barrier Function of a Vertebrate Chromatin Insulator. J. Biol. Chem. 2010, 285, 37589–37597. [Google Scholar] [CrossRef] [Green Version]
  91. West, A.G.; Huang, S.; Gaszner, M.; Litt, M.D.; Felsenfeld, G. Recruitment of Histone Modifications by USF Proteins at a Vertebrate Barrier Element. Mol. Cell 2004, 16, 453–463. [Google Scholar] [CrossRef] [PubMed]
  92. Gerasimova, T.I.; Byrd, K.; Corces, V.G. A Chromatin Insulator Determines the Nuclear Localization of DNA. Mol. Cell 2000, 6, 1025–1035. [Google Scholar] [CrossRef]
  93. Yusufzai, T.M.; Tagami, H.; Nakatani, Y.; Felsenfeld, G. CTCF Tethers an Insulator to Subnuclear Sites, Suggesting Shared Insulator Mechanisms across Species. Mol. Cell 2004, 13, 291–298. [Google Scholar] [CrossRef]
  94. Little, R.D.; Schildkraut, C.L. Initiation of Latent DNA Replication in the Epstein-Barr Virus Genome Can Occur at Sites Other than the Genetically Defined Origin. Mol. Cell Biol. 1995, 15, 2893–2903. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Norio, P.; Schildkraut, C.L.; Yates, J.L. Initiation of DNA Replication within OriP Is Dispensable for Stable Replication of the Latent Epstein-Barr Virus Chromosome after Infection of Established Cell Lines. J. Virol. 2000, 74, 8563–8574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ott, E.; Norio, P.; Ritzi, M.; Schildkraut, C.; Schepers, A. The Dyad Symmetry Element of Epstein-Barr Virus Is a Dominant but Dispensable Replication Origin. PLoS ONE 2011, 6, e18609. [Google Scholar] [CrossRef] [PubMed]
  97. Boulikas, T. Common Structural Features of Replication Origins in All Life Forms. J. Cell Biochem. 1996, 60, 297–316. [Google Scholar] [CrossRef]
  98. Gerhardt, J.; Jafar, S.; Spindler, M.-P.; Ott, E.; Schepers, A. Identification of New Human Origins of DNA Replication by an Origin-Trapping Assay. Mol. Cell Biol. 2006, 26, 7731–7746. [Google Scholar] [CrossRef] [Green Version]
  99. Reeves, R.; Nissen, M.S. The A.T-DNA-Binding Domain of Mammalian High Mobility Group I Chromosomal Proteins. A Novel Peptide Motif for Recognizing DNA Structure. J. Biol. Chem. 1990, 265, 8573–8582. [Google Scholar] [CrossRef]
  100. Pich, D.; Humme, S.; Spindler, M.-P.; Schepers, A.; Hammerschmidt, W. Conditional Gene Vectors Regulated in Cis. Nucleic Acids Res. 2008, 36, e83. [Google Scholar] [CrossRef] [Green Version]
  101. Baiker, A.; Maercker, C.; Piechaczek, C.; Schmidt, S.B.; Bode, J.; Benham, C.; Lipps, H.J. Mitotic Stability of an Episomal Vector Containing a Human Scaffold/Matrix-Attached Region Is Provided by Association with Nuclear Matrix. Nat. Cell Biol. 2000, 2, 182–184. [Google Scholar] [CrossRef] [PubMed]
  102. Jenke, A.C.; Stehle, I.M.; Herrmann, F.; Eisenberger, T.; Baiker, A.; Bode, J.; Fackelmayer, F.O.; Lipps, H.J. 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. USA 2004, 101, 11322–11327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Rupprecht, S.; Hagedorn, C.; Seruggia, D.; Magnusson, T.; Wagner, E.; Ogris, M.; Lipps, H.J. Controlled Removal of a Nonviral Episomal Vector from Transfected Cells. Gene 2010, 466, 36–42. [Google Scholar] [CrossRef] [PubMed]
  104. Stavrou, E.F.; Lazaris, V.M.; Giannakopoulos, A.; Papapetrou, E.; Spyridonidis, A.; Zoumbos, N.C.; Gkountis, A.; Athanassiadou, A. The β-Globin Replicator Greatly Enhances the Potential of S/MAR Based Episomal Vectors for Gene Transfer into Human Haematopoietic Progenitor Cells. Sci. Rep. 2017, 7, 40673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Stavrou, E.F.; Simantirakis, E.; Verras, M.; Barbas, C.; Vassilopoulos, G.; Peterson, K.R.; Athanassiadou, A. Episomal Vectors Based on S/MAR and the β-Globin Replicator, Encoding a Synthetic Transcriptional Activator, Mediate Efficient γ-Globin Activation in Haematopoietic Cells. Sci. Rep. 2019, 9, 19765. [Google Scholar] [CrossRef] [Green Version]
  106. Valton, A.-L.; Prioleau, M.-N. G-Quadruplexes in DNA Replication: A Problem or a Necessity? Trends Genet. 2016, 32, 697–706. [Google Scholar] [CrossRef]
  107. Rhodes, D.; Lipps, H.J. G-Quadruplexes and Their Regulatory Roles in Biology. Nucleic Acids Res. 2015, 43, 8627–8637. [Google Scholar] [CrossRef] [Green Version]
  108. Cayrou, C.; Coulombe, P.; Vigneron, A.; Stanojcic, S.; Ganier, O.; Peiffer, I.; Rivals, E.; Puy, A.; Laurent-Chabalier, S.; Desprat, R.; et al. Genome-Scale Analysis of Metazoan Replication Origins Reveals Their Organization in Specific but Flexible Sites Defined by Conserved Features. Genome Res. 2011, 21, 1438–1449. [Google Scholar] [CrossRef] [Green Version]
  109. Comoglio, F.; Schlumpf, T.; Schmid, V.; Rohs, R.; Beisel, C.; Paro, R. High-Resolution Profiling of Drosophila Replication Start Sites Reveals a DNA Shape and Chromatin Signature of Metazoan Origins. Cell Rep. 2015, 11, 821–834. [Google Scholar] [CrossRef] [Green Version]
  110. Cayrou, C.; Ballester, B.; Peiffer, I.; Fenouil, R.; Coulombe, P.; Andrau, J.-C.; van Helden, J.; Méchali, M. The Chromatin Environment Shapes DNA Replication Origin Organization and Defines Origin Classes. Genome Res. 2015, 25, 1873–1885. [Google Scholar] [CrossRef] [Green Version]
  111. Prorok, P.; Artufel, M.; Aze, A.; Coulombe, P.; Peiffer, I.; Lacroix, L.; Guédin, A.; Mergny, J.-L.; Damaschke, J.; Schepers, A.; et al. Involvement of G-Quadruplex Regions in Mammalian Replication Origin Activity. Nat. Commun. 2019, 10, 3274. [Google Scholar] [CrossRef] [PubMed]
  112. Norseen, J.; Thomae, A.; Sridharan, V.; Aiyar, A.; Schepers, A.; Lieberman, P.M. RNA-Dependent Recruitment of the Origin Recognition Complex. EMBO J. 2008, 27, 3024–3035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Norseen, J.; Johnson, F.B.; Lieberman, P.M. Role for G-Quadruplex RNA Binding by Epstein-Barr Virus Nuclear Antigen 1 in DNA Replication and Metaphase Chromosome Attachment. J. Virol. 2009, 83, 10336–10346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Shimizu, T.S.; Takahashi, K.; Tomita, M. CpG Distribution Patterns in Methylated and Non-Methylated Species. Gene 1997, 205, 103–107. [Google Scholar] [CrossRef]
  115. Ishii, K.J.; Akira, S. Innate Immune Recognition of, and Regulation by, DNA. Trends Immunol. 2006, 27, 525–532. [Google Scholar] [CrossRef] [PubMed]
  116. Bauer, S.; Kirschning, C.J.; Häcker, H.; Redecke, V.; Hausmann, S.; Akira, S.; Wagner, H.; Lipford, G.B. Human TLR9 Confers Responsiveness to Bacterial DNA via Species-Specific CpG Motif Recognition. Proc. Natl. Acad. Sci. USA 2001, 98, 9237–9242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Hyde, S.C.; Pringle, I.A.; Abdullah, S.; Lawton, A.E.; Davies, L.A.; Varathalingam, A.; Nunez-Alonso, G.; Green, A.-M.; Bazzani, R.P.; Sumner-Jones, S.G.; et al. CpG-Free Plasmids Confer Reduced Inflammation and Sustained Pulmonary Gene Expression. Nat. Biotechnol. 2008, 26, 549–551. [Google Scholar] [CrossRef] [Green Version]
  118. Yew, N.S.; Zhao, H.; Przybylska, M.; Wu, I.-H.; Tousignant, J.D.; Scheule, R.K.; Cheng, S.H. CpG-Depleted Plasmid DNA Vectors with Enhanced Safety and Long-Term Gene Expression in Vivo. Mol. Ther. 2002, 5, 731–738. [Google Scholar] [CrossRef]
  119. Darquet, A.M.; Rangara, R.; Kreiss, P.; Schwartz, B.; Naimi, S.; Delaère, P.; Crouzet, J.; Scherman, D. Minicircle: An Improved DNA Molecule for in Vitro and in Vivo Gene Transfer. Gene Ther. 1999, 6, 209–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Broll, S.; Oumard, A.; Hahn, K.; Schambach, A.; Bode, J. Minicircle Performance Depending on S/MAR-Nuclear Matrix Interactions. J. Mol. Biol. 2010, 395, 950–965. [Google Scholar] [CrossRef]
  121. Luke, J.M.; Vincent, J.M.; Du, S.X.; Gerdemann, U.; Leen, A.M.; Whalen, R.G.; Hodgson, C.P.; Williams, J.A. Improved Antibiotic-Free Plasmid Vector Design by Incorporation of Transient Expression Enhancers. Gene Ther. 2011, 18, 334–343. [Google Scholar] [CrossRef] [PubMed]
  122. Marie, C.; Vandermeulen, G.; Quiviger, M.; Richard, M.; Préat, V.; Scherman, D. PFARs, Plasmids Free of Antibiotic Resistance Markers, Display High-Level Transgene Expression in Muscle, Skin and Tumour Cells. J. Gene Med. 2010, 12, 323–332. [Google Scholar] [CrossRef] [PubMed]
  123. Bozza, M.; Green, E.W.; Espinet, E.; De Roia, A.; Klein, C.; Vogel, V.; Offringa, R.; Williams, J.A.; Sprick, M.; Harbottle, R.P. Novel Non-Integrating DNA Nano-S/MAR Vectors Restore Gene Function in Isogenic Patient-Derived Pancreatic Tumor Models. Mol. Ther. Methods Clin. Dev. 2020, 17, 957–968. [Google Scholar] [CrossRef] [PubMed]
  124. Giannakopoulos, A.; Quiviger, M.; Stavrou, E.; Verras, M.; Marie, C.; Scherman, D.; Athanassiadou, A. Efficient Episomal Gene Transfer to Human Hepatic Cells Using the PFAR4-S/MAR Vector. Mol. Biol. Rep. 2019, 46, 3203–3211. [Google Scholar] [CrossRef] [PubMed]
  125. Pastor, M.; Johnen, S.; Harmening, N.; Quiviger, M.; Pailloux, J.; Kropp, M.; Walter, P.; Ivics, Z.; Izsvák, Z.; Thumann, G.; et al. The Antibiotic-Free PFAR4 Vector Paired with the Sleeping Beauty Transposon System Mediates Efficient Transgene Delivery in Human Cells. Mol.Ther. Nucleic Acids 2018, 11, 57–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Turchiano, G.; Latella, M.C.; Gogol-Döring, A.; Cattoglio, C.; Mavilio, F.; Izsvák, Z.; Ivics, Z.; Recchia, A. Genomic Analysis of Sleeping Beauty Transposon Integration in Human Somatic Cells. PLoS ONE 2014, 9, e112712. [Google Scholar] [CrossRef] [PubMed]
  127. Boulikas, T. Homeotic Protein Binding Sites, Origins of Replication, and Nuclear Matrix Anchorage Sites Share the ATTA and ATTTA Motifs. J. Cell Biochem. 1992, 50, 111–123. [Google Scholar] [CrossRef] [PubMed]
  128. Ripin, N.; Boudet, J.; Duszczyk, M.M.; Hinniger, A.; Faller, M.; Krepl, M.; Gadi, A.; Schneider, R.J.; Šponer, J.; Meisner-Kober, N.C.; et al. Molecular Basis for AU-Rich Element Recognition and Dimerization by the HuR C-Terminal RRM. Proc. Natl. Acad. Sci. USA 2019, 116, 2935–2944. [Google Scholar] [CrossRef] [Green Version]
  129. Lieberman-Aiden, E.; van Berkum, N.L.; Williams, L.; Imakaev, M.; Ragoczy, T.; Telling, A.; Amit, I.; Lajoie, B.R.; Sabo, P.J.; Dorschner, M.O.; et al. Comprehensive Mapping of Long-Range Interactions Reveals Folding Principles of the Human Genome. Science 2009, 326, 289–293. [Google Scholar] [CrossRef] [Green Version]
  130. Dixon, J.R.; Selvaraj, S.; Yue, F.; Kim, A.; Li, Y.; Shen, Y.; Hu, M.; Liu, J.S.; Ren, B. Topological Domains in Mammalian Genomes Identified by Analysis of Chromatin Interactions. Nature 2012, 485, 376–380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Zirkel, A.; Papantonis, A. Transcription as a Force Partitioning the Eukaryotic Genome. Biol. Chem. 2014, 395, 1301–1305. [Google Scholar] [CrossRef] [PubMed]
  132. Van Steensel, B.; Furlong, E.E.M. The Role of Transcription in Shaping the Spatial Organization of the Genome. Nat. Rev. Mol. Cell Biol. 2019, 20, 327–337. [Google Scholar] [CrossRef] [PubMed]
  133. Karmodiya, K.; Krebs, A.R.; Oulad-Abdelghani, M.; Kimura, H.; Tora, L. H3K9 and H3K14 Acetylation Co-Occur at Many Gene Regulatory Elements, While H3K14ac Marks a Subset of Inactive Inducible Promoters in Mouse Embryonic Stem Cells. BMC Genom. 2012, 13, 424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Santos-Rosa, H.; Schneider, R.; Bannister, A.J.; Sherriff, J.; Bernstein, B.E.; Emre, N.C.; Schreiber, S.L.; Mellor, J.; Kouzarides, T. Active Genes Are Tri-Methylated at K4 of Histone H3. Nature 2002, 419, 407–411. [Google Scholar] [CrossRef] [PubMed]
  135. Koch, C.M.; Andrews, R.M.; Flicek, P.; Dillon, S.C.; Karaoz, U.; Clelland, G.K.; Wilcox, S.; Beare, D.M.; Fowler, J.C.; Couttet, P.; et al. The Landscape of Histone Modifications across 1% of the Human Genome in Five Human Cell Lines. Genome Res. 2007, 17, 691–707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Rupprecht, S.; Lipps, H.J. Cell Cycle Dependent Histone Dynamics of an Episomal Non-Viral Vector. Gene 2009, 439, 95–101. [Google Scholar] [CrossRef] [PubMed]
  137. Deutsch, M.J.; Ott, E.; Papior, P.; Schepers, A. The Latent Origin of Replication of Epstein-Barr Virus Directs Viral Genomes to Active Regions of the Nucleus. J. Virol. 2009. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Hodin, T.L.; Najrana, T.; Yates, J.L. Efficient Replication of Epstein-Barr Virus-Derived Plasmids Requires Tethering by EBNA1 to Host Chromosomes. J. Virol. 2013, 87, 13020–13028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Papantonis, A.; Kohro, T.; Baboo, S.; Larkin, J.D.; Deng, B.; Short, P.; Tsutsumi, S.; Taylor, S.; Kanki, Y.; Kobayashi, M.; et al. TNFalpha Signals through Specialized Factories Where Responsive Coding and MiRNA Genes Are Transcribed. EMBO J. 2012, 31, 4404–4414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Xu, M.; Cook, P.R. Similar Active Genes Cluster in Specialized Transcription Factories. J. Cell Biol. 2008, 181, 615–623. [Google Scholar] [CrossRef] [Green Version]
  141. Papantonis, A.; Cook, P.R. Fixing the Model for Transcription: The DNA Moves, Not the Polymerase. Transcription 2011, 2, 41–44. [Google Scholar] [CrossRef]
  142. Gogol-Döring, A.; Ammar, I.; Gupta, S.; Bunse, M.; Miskey, C.; Chen, W.; Uckert, W.; Schulz, T.F.; Izsvák, Z.; Ivics, Z. Genome-Wide Profiling Reveals Remarkable Parallels Between Insertion Site Selection Properties of the MLV Retrovirus and the PiggyBac Transposon in Primary Human CD4(+) T Cells. Mol. Ther. 2016, 24, 592–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Zayed, H.; Izsvák, Z.; Walisko, O.; Ivics, Z. Development of Hyperactive Sleeping Beauty Transposon Vectors by Mutational Analysis. Mol. Ther. 2004, 9, 292–304. [Google Scholar] [CrossRef] [PubMed]
  144. Moldt, B.; Miskey, C.; Staunstrup, N.H.; Gogol-Döring, A.; Bak, R.O.; Sharma, N.; Mátés, L.; Izsvák, Z.; Chen, W.; Ivics, Z.; et al. Comparative Genomic Integration Profiling of Sleeping Beauty Transposons Mobilized with High Efficacy from Integrase-Defective Lentiviral Vectors in Primary Human Cells. Mol. Ther. 2011, 19, 1499–1510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Staunstrup, N.H.; Moldt, B.; Mátés, L.; Villesen, P.; Jakobsen, M.; Ivics, Z.; Izsvák, Z.; Mikkelsen, J.G. Hybrid Lentivirus-Transposon Vectors with a Random Integration Profile in Human Cells. Mol. Ther. 2009, 17, 1205–1214. [Google Scholar] [CrossRef] [PubMed]
  146. Vink, C.A.; Gaspar, H.B.; Gabriel, R.; Schmidt, M.; McIvor, R.S.; Thrasher, A.J.; Qasim, W. Sleeping Beauty Transposition from Nonintegrating Lentivirus. Mol. Ther. 2009, 17, 1197–1204. [Google Scholar] [CrossRef] [PubMed]
  147. Zhang, W.; Muck-Hausl, M.; Wang, J.; Sun, C.; Gebbing, M.; Miskey, C.; Ivics, Z.; Izsvak, Z.; Ehrhardt, A. Integration Profile and Safety of an Adenovirus Hybrid-Vector Utilizing Hyperactive Sleeping Beauty Transposase for Somatic Integration. PLoS ONE 2013, 8, e75344. [Google Scholar] [CrossRef] [Green Version]
  148. Bestor, T.H. Gene Silencing as a Threat to the Success of Gene Therapy. J. Clin. Investig. 2000, 105, 409–411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Stein, S.; Ott, M.G.; Schultze-Strasser, S.; Jauch, A.; Burwinkel, B.; Kinner, A.; Schmidt, M.; Krämer, A.; Schwäble, J.; Glimm, H.; et al. Genomic Instability and Myelodysplasia with Monosomy 7 Consequent to EVI1 Activation after Gene Therapy for Chronic Granulomatous Disease. Nat. Med. 2010, 16, 198–204. [Google Scholar] [CrossRef] [Green Version]
  150. Cavazzana-Calvo, M.; Payen, E.; Negre, O.; Wang, G.; Hehir, K.; Fusil, F.; Down, J.; Denaro, M.; Brady, T.; Westerman, K.; et al. Transfusion Independence and HMGA2 Activation after Gene Therapy of Human β-Thalassaemia. Nature 2010, 467, 318–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Yant, S.R.; Huang, Y.; Akache, B.; Kay, M.A. Site-Directed Transposon Integration in Human Cells. Nucleic Acids Res. 2007, 35, e50. [Google Scholar] [CrossRef] [PubMed]
  152. Voigt, K.; Gogol-Döring, A.; Miskey, C.; Chen, W.; Cathomen, T.; Izsvák, Z.; Ivics, Z. Retargeting Sleeping Beauty Transposon Insertions by Engineered Zinc Finger DNA-Binding Domains. Mol. Ther. 2012, 20, 1852–1862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Ammar, I.; Gogol-Döring, A.; Miskey, C.; Chen, W.; Cathomen, T.; Izsvák, Z.; Ivics, Z. Retargeting Transposon Insertions by the Adeno-Associated Virus Rep Protein. Nucleic Acids Res. 2012, 40, 6693–6712. [Google Scholar] [CrossRef] [Green Version]
  154. Kovač, A.; Miskey, C.; Menzel, M.; Grueso, E.; Gogol-Döring, A.; Ivics, Z. RNA-Guided Retargeting of Sleeping Beauty Transposition in Human Cells. Elife 2020, 9, e53868. [Google Scholar] [CrossRef] [PubMed]
  155. Kovač, A.; Ivics, Z. Specifically Integrating Vectors for Targeted Gene Delivery: Progress and Prospects. Cell Gene Ther. Insights 2017, 3, 103–123. [Google Scholar] [CrossRef]
  156. Waehler, R.; Russell, S.J.; Curiel, D.T. Engineering Targeted Viral Vectors for Gene Therapy. Nat. Rev. Genet. 2007, 8, 573–587. [Google Scholar] [CrossRef] [PubMed]
  157. Kreppel, F.; Hagedorn, C. Capsid and Genome Modification Strategies to Reduce the Immunogenicity of Adenoviral Vectors. Int. J. Mol. Sci. 2021, 22, 2417. [Google Scholar] [CrossRef] [PubMed]
  158. Kreppel, F.; Kochanek, S. Modification of Adenovirus Gene Transfer Vectors With Synthetic Polymers: A Scientific Review and Technical Guide. Mol. Ther. 2008, 16, 16–29. [Google Scholar] [CrossRef]
  159. Baden, L.R.; El Sahly, H.M.; Essink, B.; Kotloff, K.; Frey, S.; Novak, R.; Diemert, D.; Spector, S.A.; Rouphael, N.; Creech, C.B.; et al. Efficacy and Safety of the MRNA-1273 SARS-CoV-2 Vaccine. N. Engl. J. Med. 2021, 384, 403–416. [Google Scholar] [CrossRef] [PubMed]
  160. Thomas, S.J.; Moreira, E.D.; Kitchin, N.; Absalon, J.; Gurtman, A.; Lockhart, S.; Perez, J.L.; Pérez Marc, G.; Polack, F.P.; Zerbini, C.; et al. Safety and Efficacy of the BNT162b2 MRNA Covid-19 Vaccine through 6 Months. N. Engl. J. Med. 2021, 385, 1761–1773. [Google Scholar] [CrossRef]
  161. Hou, X.; Zaks, T.; Langer, R.; Dong, Y. Lipid Nanoparticles for MRNA Delivery. Nat. Rev. Mater. 2021, 6, 1078–1094. [Google Scholar] [CrossRef] [PubMed]
  162. Steffens, R.C.; Wagner, E. Directing the Way—Receptor and Chemical Targeting Strategies for Nucleic Acid Delivery. Pharm. Res. 2022, 1–30. [Google Scholar] [CrossRef] [PubMed]
  163. Amberger, M.; Ivics, Z. Latest Advances for the Sleeping Beauty Transposon System: 23 Years of Insomnia but Prettier than Ever: Refinement and Recent Innovations of the Sleeping Beauty Transposon System Enabling Novel, Nonviral Genetic Engineering Applications. Bioessays 2020, 42, e2000136. [Google Scholar] [CrossRef] [PubMed]
  164. Dorigo, O.; Gil, J.S.; Gallaher, S.D.; Tan, B.T.; Castro, M.G.; Lowenstein, P.R.; Calos, M.P.; Berk, A.J. Development of a Novel Helper-Dependent Adenovirus-Epstein-Barr Virus Hybrid System for the Stable Transformation of Mammalian Cells. J. Virol. 2004, 78, 6556–6566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Kreppel, F.; Kochanek, S. Long-Term Transgene Expression in Proliferating Cells Mediated by Episomally Maintained High-Capacity Adenovirus Vectors. J. Virol. 2004, 78, 9–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Gil, J.S.; Gallaher, S.D.; Berk, A.J. Delivery of an EBV Episome by a Self-Circularizing Helper-Dependent Adenovirus: Long-Term Transgene Expression in Immunocompetent Mice. Gene Ther. 2010, 17, 1288–1293. [Google Scholar] [CrossRef] [PubMed]
  167. Hagedorn, C.; Schnödt-Fuchs, M.; Boehme, P.; Abdelrazik, H.; Lipps, H.J.; Büning, H. S/MAR Element Facilitates Episomal Long-Term Persistence of Adeno-Associated Virus Vector Genomes in Proliferating Cells. Hum. Gene Ther. 2017, 28, 1169–1179. [Google Scholar] [CrossRef] [PubMed]
  168. Turunen, T.A.K.; Laakkonen, J.P.; Alasaarela, L.; Airenne, K.J.; Ylä-Herttuala, S. Sleeping Beauty-Baculovirus Hybrid Vectors for Long-Term Gene Expression in the Eye. J. Gene Med. 2014, 16, 40–53. [Google Scholar] [CrossRef] [PubMed]
  169. Hausl, M.A.; Zhang, W.; Müther, N.; Rauschhuber, C.; Franck, H.G.; Merricks, E.P.; Nichols, T.C.; Kay, M.A.; Ehrhardt, A. Hyperactive Sleeping Beauty Transposase Enables Persistent Phenotypic Correction in Mice and a Canine Model for Hemophilia B. Mol. Ther. 2010, 18, 1896–1906. [Google Scholar] [CrossRef] [PubMed]
  170. Voigtlander, R.; Haase, R.; Muck-Hausl, M.; Zhang, W.; Boehme, P.; Lipps, H.J.; Schulz, E.; Baiker, A.; Ehrhardt, A. A Novel Adenoviral Hybrid-Vector System Carrying a Plasmid Replicon for Safe and Efficient Cell and Gene Therapeutic Applications. Mol. Ther. Nucleic Acids 2013, 2, e83. [Google Scholar] [CrossRef]
  171. Harraghy, N.; Gaussin, A.; Mermod, N. Sustained Transgene Expression Using MAR Elements. Curr. Gene Ther. 2008, 8, 353–366. [Google Scholar] [CrossRef] [PubMed]
  172. Grandjean, M.; Girod, P.-A.; Calabrese, D.; Kostyrko, K.; Wicht, M.; Yerly, F.; Mazza, C.; Beckmann, J.S.; Martinet, D.; Mermod, N. High-Level Transgene Expression by Homologous Recombination-Mediated Gene Transfer. Nucleic Acids Res. 2011, 39, e104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Zhang, W.; Hagedorn, C.; Schulz, E.; Lipps, H.-J.; Ehrhardt, A. Viral Hybrid-Vectors for Delivery of Autonomous Replicons. Curr. Gene Ther. 2014, 14, 10–23. [Google Scholar] [CrossRef] [PubMed]
  174. Huang, S.; Kamihira, M. Development of Hybrid Viral Vectors for Gene Therapy. Biotechnol. Adv. 2013, 31, 208–223. [Google Scholar] [CrossRef] [PubMed]
  175. Raper, S.E.; Chirmule, N.; Lee, F.S.; Wivel, N.A.; Bagg, A.; Gao, G.; Wilson, J.M.; Batshaw, M.L. Fatal Systemic Inflammatory Response Syndrome in a Ornithine Transcarbamylase Deficient Patient Following Adenoviral Gene Transfer. Mol. Genet. Metab. 2003, 80, 148–158. [Google Scholar] [CrossRef] [PubMed]
  176. Zolgensma Acute Liver Failure Update. Available online: https://www.novartis.com/news/zolgensma-acute-liver-failure-update (accessed on 17 August 2022).
  177. Kreppel, F.; Gackowski, J.; Schmidt, E.; Kochanek, S. Combined Genetic and Chemical Capsid Modifications Enable Flexible and Efficient De- and Retargeting of Adenovirus Vectors. Mol. Ther. 2005, 12, 107–117. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Vector families and vector groups for gene therapy. Vector systems can be classified into families of viral and non-viral vectors. Viral vectors deliver nucleic acids to a variety of cells via transduction. Nonviral vectors are usually delivered using different transfection methods, such as lipofection or electroporation. Both viral and nonviral vectors can be grouped into integrating and nonintegrating vectors. Examples for each vector group are given without any claim of completeness. AAV, Adeno-associated Virus; HSV, Herpes Simplex Virus; S/MAR, scaffold/matrix attachment region; EBV, Eppstein-Barr virus.
Figure 1. Vector families and vector groups for gene therapy. Vector systems can be classified into families of viral and non-viral vectors. Viral vectors deliver nucleic acids to a variety of cells via transduction. Nonviral vectors are usually delivered using different transfection methods, such as lipofection or electroporation. Both viral and nonviral vectors can be grouped into integrating and nonintegrating vectors. Examples for each vector group are given without any claim of completeness. AAV, Adeno-associated Virus; HSV, Herpes Simplex Virus; S/MAR, scaffold/matrix attachment region; EBV, Eppstein-Barr virus.
Genes 13 01872 g001
Figure 2. Mechanisms of DNA replication and nuclear retention of different nonviral vector systems. (A) EBV-based replicons replicate EBNA1 and MCM2-7 dependent. The binding of EBNA1 to the DS element recruits MCM2-7, thus mediating DNA replication, while binding of EBNA1 to both the FR element and metaphase chromosomes provides a “piggy-back” mechanism for nuclear retention and segregation. (B) SB transposases are delivered to the target cells either directly as recombinant protein or as a transposes-encoding plasmid, or mRNA, along with transposon-carrying plasmids (transgene flanked by TIR sites). Once both components (transposon and transposase) are available in the cell, the SB transposase binds to the TIR, excises the transposon and mediates its integration in the cellular genome. Subsequently, the inserted transposon is replicated and segregated during the cell cycle. (C) Episomal maintenance of S/MAR-based vectors relies on the presence of active transcription running into the S/MAR sequence. Replication of S/MAR-based episomes is initiated during the early S-phase at various sites of the vector, depending on the host’s cell replication machinery. Segregation and nuclear retention rely on the interaction of the episomes with metaphase chromosomes in a “piggy-back” manner, although the proteins involved are still unknown. EBV, Eppstein-Barr virus; FR, family of repeats element; DS, dyad symmetry element; TIR, terminal inverted repeats; EBNA1, EBV nuclear antigen 1; MCM2, minichromosome maintenance complex component 2; S/MAR, scaffold/matrix attachment region.
Figure 2. Mechanisms of DNA replication and nuclear retention of different nonviral vector systems. (A) EBV-based replicons replicate EBNA1 and MCM2-7 dependent. The binding of EBNA1 to the DS element recruits MCM2-7, thus mediating DNA replication, while binding of EBNA1 to both the FR element and metaphase chromosomes provides a “piggy-back” mechanism for nuclear retention and segregation. (B) SB transposases are delivered to the target cells either directly as recombinant protein or as a transposes-encoding plasmid, or mRNA, along with transposon-carrying plasmids (transgene flanked by TIR sites). Once both components (transposon and transposase) are available in the cell, the SB transposase binds to the TIR, excises the transposon and mediates its integration in the cellular genome. Subsequently, the inserted transposon is replicated and segregated during the cell cycle. (C) Episomal maintenance of S/MAR-based vectors relies on the presence of active transcription running into the S/MAR sequence. Replication of S/MAR-based episomes is initiated during the early S-phase at various sites of the vector, depending on the host’s cell replication machinery. Segregation and nuclear retention rely on the interaction of the episomes with metaphase chromosomes in a “piggy-back” manner, although the proteins involved are still unknown. EBV, Eppstein-Barr virus; FR, family of repeats element; DS, dyad symmetry element; TIR, terminal inverted repeats; EBNA1, EBV nuclear antigen 1; MCM2, minichromosome maintenance complex component 2; S/MAR, scaffold/matrix attachment region.
Genes 13 01872 g002
Table 1. Advantages and disadvantages of nonviral vector systems. The major limitation of each vector system is marked in bold. The term delivery efficiency, as used here, refers to the ratio of the copy number of genetic material that becomes established/integrated into target cells to the copy number of genetic material used for initial transfection/transduction/injection.
Table 1. Advantages and disadvantages of nonviral vector systems. The major limitation of each vector system is marked in bold. The term delivery efficiency, as used here, refers to the ratio of the copy number of genetic material that becomes established/integrated into target cells to the copy number of genetic material used for initial transfection/transduction/injection.
VECTOR SYSTEMADVANTAGESDISADVANTAGES
EBV-based replicons- stable transgene expression
- high cloning capacity (>100 kb)
- episomal maintenance
- efficient modification of target cells
- easy engineering, low-cost production
- contains viral elements
- low delivery efficiencies (in vitro and in vivo)
- limited targeting options
- EBNA-1-related toxicity
Sleeping Beauty- stable transgene expression
- comprehensively studied safety profile
- efficient modification of target cells
- risk of insertional mutagenesis
- requires co-delivery of transposases
- low delivery efficiencies (in vitro and in vivo)
S/MAR-based replicons- lack of viral elements
- episomal maintenance
- easy engineering, low-cost production
- low immunogenicity
- low establishment efficiencies
- limited cloning capacity
- low delivery efficiencies (in vitro and in vivo)
- limited targeting options
Artificial Chromosomes- unlimited cloning capacity
- copy number control
- nonintegrating
- nonessential, additional chromosome
- complex construction
- low delivery efficiencies (in vitro and in vivo)
- construct-dependent inefficiencies in segregation to daughter cells possible
- potential risks imposed by homologous recombination need to be analysed in depth
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kreppel, F.; Hagedorn, C. Episomes and Transposases—Utilities to Maintain Transgene Expression from Nonviral Vectors. Genes 2022, 13, 1872. https://doi.org/10.3390/genes13101872

AMA Style

Kreppel F, Hagedorn C. Episomes and Transposases—Utilities to Maintain Transgene Expression from Nonviral Vectors. Genes. 2022; 13(10):1872. https://doi.org/10.3390/genes13101872

Chicago/Turabian Style

Kreppel, Florian, and Claudia Hagedorn. 2022. "Episomes and Transposases—Utilities to Maintain Transgene Expression from Nonviral Vectors" Genes 13, no. 10: 1872. https://doi.org/10.3390/genes13101872

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop