Kayla
Neyra
a,
Sara
Desai
b and
Divita
Mathur
*a
aDepartment of Chemistry, Case Western Reserve University, Cleveland, OH 44106, USA. E-mail: dxm700@case.edu
bDepartment of Biochemistry, Case Western Reserve University, Cleveland, OH 44106, USA
First published on 29th November 2024
Synthetic DNA nanotechnology has emerged as a powerful tool for creating precise nanoscale structures with diverse applications in biotechnology and materials science. Recently, it has evolved to include gene-encoded DNA nanoparticles, which have potentially unique advantages compared to alternative gene delivery platforms. In exciting new developments, we and others have shown how the long single strand within DNA origami nanoparticles, the scaffold strand, can be customized to encode protein-expressing genes and engineer nanoparticles that interface with the transcription–translation machinery for protein production. Remarkably, therefore, DNA nanoparticles – despite their complex three-dimensional shapes – can function as canonical genes. Characteristics such as potentially unlimited gene packing size and low immunogenicity make DNA-based platforms promising for a variety of gene therapy applications. In this review, we first outline various techniques for the isolation of the gene-encoded scaffold strand, a crucial precursor for building protein-expressing DNA nanoparticles. Next, we highlight how features such as sequence design, staple strand optimization, and overall architecture of gene-encoded DNA nanoparticles play a key role in the enhancement of protein expression. Finally, we discuss potential applications of these DNA origami structures to provide a comprehensive overview of the current state of gene-encoded DNA nanoparticles and motivate future directions.
Recent advancements in the biomedical front have seen a convergence in DNA nanoparticle shapes towards biomimetic designs, with three-dimensional wireframe polyhedral nanoparticles accelerating the precise display of proteins and ligands for mimicking viral structures while serving as vaccines.10–13 More excitingly, the evolution of DNA nanotechnology has extended beyond structural programmability and into functional programmability. Functional nucleic acid oligonucleotides, such as DNA aptamers and non-coding RNAs, can be seamlessly integrated into DNA nanostructures through standard base pairing or by appending sequences to staple strand ends, imparting additional functionality.14–16 By integrating functional DNA and RNA with DNA-based delivery vehicles, it is possible to enhance the targeting of therapeutic systems and lower non-specific effects.16
We would like to draw attention to another rapidly growing area within functional nucleic acid nanoparticles that merges DNA's biological function as genetic sequences in the central dogma of life with its architectural capabilities. These functional nucleic acids are gene-encoded DNA origami nanoparticles that are created by encoding the scaffold strand with a protein-expressing gene rather than arbitrary or non-compatible sequences. The resultant gene-encoded nanostructures undergo transcription in bacterial and mammalian systems to produce the corresponding proteins. This enabling technology has been demonstrated for prototypical fluorescent protein expression in cells and cell-free systems as well as therapeutic protein (p53) expression in mice (described below). Protein expression via gene-encoded DNA nanoparticles, combined with the ability of DNA nanoparticles to predictably target cells, underscores its importance and significant potential in transforming gene therapies. Herein, we feature the latest progress in creating gene-encoded DNA nanoparticles as well as their prospective applications. Additionally, we provide insight on the developing field by emphasizing how to encode custom genes into DNA nanoparticles, key considerations identified thus far in design and function, as well as future directions that require attention for translational applications.
In gene-encoded DNA origami nanoparticles, the scaffold strand is neither an indigenous bacteriophage genomic sequence nor an arbitrary sequence. Rather, the scaffold strand encodes a genetic sequence that is transcribable by RNA polymerases to produce a protein of interest.18 Applying established DNA origami design principles and available computational tools, nanoparticles can be created with a unique scaffold – one that is specific to any desired genetic sequence. Experimental assembly of gene-encoded DNA origami nanoparticles is similar to traditional (M13mp18-encoded) DNA origami nanoparticles; 5- to 10-fold excess staple strands are combined with the gene-encoded scaffold strand in appropriate Mg2+ supplemented buffer and incubated under a thermal annealing program.19 Canonical genetic sequences are encoded into double stranded plasmids or chromosomes; therefore, strategies to extract or produce gene-encoding single DNA strands from double stranded sources are required for the assembly of gene-encoded DNA nanoparticles.
Many therapeutic genes (important for gene delivery purposes) are a few hundred to several thousand bases pairs long. Until de novo DNA synthesis can make technological advancements to produce high-throughput long (>3 kilobases) ssDNA, well-established molecular biology techniques leveraging enzymes and microbial engineering for ssDNA production need to be translated for custom scaffold strands in DNA nanotechnology. There are other comprehensive reviews explaining the different methods of scaffold production of arbitrary sequences, and some of these have led to the production of long single-stranded genes.20,21 For these methods, the precursor DNA template is most often a gene-encoding plasmid. Plasmids encoding prototypical genes can be acquired from non-profit repositories such as Addgene but molecular biology techniques, such as Golden Gate assembly, can also be used to create a plasmid with a custom gene insert.22 The plasmid can be amplified using bacterial culturing and plasmid extraction, after which it is ready for use as a template. Fig. 1 summarizes the various enzymatic methods and Fig. 2 represents bacteriophage-based methods adopted to produce gene-encoded scaffold strands.
Fig. 3 Gene-encoded DNA nanoparticles: assembly and current applications. (A) Schematic representation of DNA nanoparticle assembly from a single-stranded gene-encoding scaffold and complementary staple strands. (B) Cylindrical model of 18 hb DNA nanoparticle with extended homology arms for genome targeting. The gene of interest can then be integrated within the genome via CRISPR-Cas9 homology directed repair. Adapted from ref. 18 licensed under CC-BY-NC. (C) Visualization of four DNA nanoparticles with varying aspect ratios. All structures were assembled from the same scaffold strand (sc_EGFP1). eGFP expression levels showed no significant difference between the 20 hb and the 32 hb, but expression as measured by flow cytometry was slightly lower for the 12 hb. Adapted from ref. 22 licensed under CC-BY. (D) Visualization of DNA nanoparticle monomers combining to produce dimer (i), trimer (i and ii), and tetramer (i, ii, iii, and iv) higher-order assemblies. Codelivery of eGFP and mCherry was carried out with a dimeric assembly with a ratio of 1:1. Epifluorescent microscopy allowed for the visualization of mCherry (red), eGFP (green) and coexpression (yellow) in HEK293T cells. Adapted from ref. 22 licensed under CC-BY. (E) 20 hb DNA nanoparticle encoding for the mCherry protein. The bottom schematic is representative of successful nuclear uptake and subsequent protein expression with a DNA targeting sequence (DTS). Adapted from ref. 24 licensed under CC-BY. (F) Hybridization of two halves of the DNA plasmid template (sense and antisense) annealed with corresponding staple strands to make a singular DNA nanoparticle encoding for the p53 anti-tumor gene. The DNA nanoparticle was coated with lipids (DOPE-FA) to facilitate cellular uptake. Western blot analysis showed significant increase in p53 protein production for Gp53-DO@lipid-FA. Adapted with permission from ref. 25. Copyright (2023) American Chemical Society. (G) Depiction of a 12 hb DNA nanoparticle encoding for the luciferase protein. Delayed addition of the NanoGlo substrate allowed for bioluminescence quantification; modifications made to the promoter region domain exhibited a decrease in bioluminescence intensity. Adapted with permission from ref. 26. Copyright (2024) American Chemical Society. Figure created with https://www.biorender.comhttps://www.biorender.com. |
Wu et al. used a similar approach where one of the primers was modified with a biotin tag for streptavidin magnetic bead-based purification after PCR, but they also added terminal DNA loops to the primer 5′ ends.25 The terminal DNA loop is known to enhance stability of linear dsDNA fragments and protect against nucleases.27 This approach was used to generate both green fluorescent protein (GFP) and p53 encoded scaffold strands that were assembled into two different DNA rectangular nanostructures with varying dimensions (42 nm and 74 nm on one side, respectively). After pulldown of the dsDNA with the magnetic beads, NaOH was again used for separating the two constituent strands. In this work, both antisense and sense strands were recovered from the supernatant (two biotinylated primers were used independently) and subsequently used as scaffold strands. The sense and antisense scaffold strands were each mixed with their respective staple strands in the ratio of 1:10 in 1 × TAE/Mg2+ buffer and thermally annealed. The resulting DNA origami monomers were then hybridized with each other through base pairing to create the gene-encoded DNA origami nanoparticle. Structural integrity was confirmed through agarose gel electrophoretic analysis and AFM images. Aside from biotin tags, the scaffold strand could be separated from its complementary strand using specific nucleases (Fig. 1(B)).
Custom scaffold strand generated via aPCR has exponentially increased the impact on DNA nanotechnology. This technique helped realize wireframe polyhedral DNA origami structures in a variety of sizes and shapes, all derived from the same M13mp18 ssDNA template.2 The Veneziano group has demonstrated that use of modified nucleotide triphosphates (dNTPs) in the aPCR reaction in place of canonical dNTPs can introduce different functionalities to the scaffold strand, such as amine-modified dCTP (NH2-dCTP).28 Thus, the programmable landscape of a DNA origami nanostructure is doubled where the staples can be chemically modified as usual but the scaffold can also have programmed functionality such as different densities of thiol groups, fluorescent molecules, and phosphorothioate modifications.
For gene-encoded DNA nanostructures, aPCR has proven useful for synthesizing a gene-encoding scaffold strand from a precursor plasmid or synthetic dsDNA gene fragment.29 The Henderson group used aPCR-generated scaffold strand to encode GFP into a cylindrical helix bundle DNA origami nanoparticle. This study assessed variations in architectures of their structure by tuning the position of the promoter region of the gene as well as changing the number of staples used to fold the nanostructure. The gene cassette included a T7 promoter, a GFP gene, and a poly A signal. It should be noted that a CMV enhancer and promoter were also included in the scaffold makeup, but these elements were not the focus in this study. aPCR followed by gel extraction purification was used to produce two variants of the ssDNA scaffold (encoding the antisense strand) – one that was inclusive of the T7 promoter and one that was not. In total, five nanostructures were designed, all following the same overall shape with a consistent crossover pattern. The designs were each folded with a varying number of staples and are identified as follows: T7GHL PO (promoter only in duplex form), T7GHL HS (folded with half set of staples), T7GHL FS (folded with full set of staples), T7GHL BP (promoter buried on internal portion of structure), and T7GHL-T7 (folded with scaffold that lacks promoter region). Apart from the T7GHL BP, the promoter of each construct is linear, with no crossovers, and is adjacent to the cylindrical core. To assess gene expression efficiency in each of the constructs, in vitro transcription (IVT) analysis was performed using a commercial T7 polymerase kit. Messenger RNA (mRNA) production of each variant was qualitatively assessed via electrophoretic analysis, and it was found that all DNA nanoparticles yielded RNA products that were identical in size, except for the construct that lacked the T7 promoter sequence in the scaffold. Additionally, there was an inverse correlation between crossover density of the structure and RNA production – meaning that T7 polymerase can transcribe through crossovers, however, this was quantitatively reduced as compactness of the structure increased. This study suggests that T7 RNA polymerase can complete transcription of a gene-encoded DNA nanoparticle despite high crossover density and limited access to the promoter.
In a recent work by us, we created a luciferase-encoding DNA nanostructure where the scaffold strand was synthesized using aPCR.26 We used a pET-22 Luc9 plasmid as the aPCR precursor DNA template in which Renilla luciferase protein (Luc9) was expressed under a bacterial T7 promoter.26 First, the transcription coding (sense) and template (antisense) strands within the plasmid were identified, based on which an excess of forward primer was added for selective amplification of the template strand. Primers were designed to flank the promoter and genetic sequence on the plasmid, but primer design was tuned based on the desired length of the scaffold strand. The Luc-encoding aPCR ssDNA product was purified to remove the dsDNA “byproduct” and other aPCR-related contaminants using gel extraction and applied in assembling a 12-helix bundle DNA origami structure (Fig. 3(G)). This nanostructure was fully compatible with in vitro cell-free transcription translation systems and produced a functional luciferase protein which could catalyze light upon the addition of luciferase substrate (vide infra).
As mentioned above, ssDNA is generally purified via agarose gel electrophoresis after aPCR synthesis. Unfortunately, gel extraction leads to low yield (∼16%) and considerable loss of material in the agarose gel.30 To compound the loss of sample, a second round of purification is a requirement after the DNA origami nanoparticle is assembled to remove excess staple strands. Therefore, custom-scaffolded DNA nanoparticle synthesis is currently a two-step process with significant loss of the scaffold strand in the first step. Notable size-exclusion techniques (other than gel extraction), such as fast protein liquid chromatography, are difficult to employ because they tend to significantly dilute the sample during fraction elution and require additional concentration steps.31 Therefore, new ways to improve the scaffold recovery post aPCR are needed. We have shown that photocleavable biotin tethers strategically placed on staple strands can allow the one-step purification of custom-scaffolded DNA origami nanoparticles.30 A photocleavable linker is a photosensitive (to 365 nm light) single bond between the 5′ terminal O and an aliphatic amino group, to which biotin can be attached via a long spacer.32 This photocleavable biotin tether on a staple strand allows the pulldown of assembled DNA origami particles, leaving behind excess staples as well as any aPCR-related contaminants. While agarose gel extracted custom scaffold strand leads to over 85% loss of the scaffold strand, using photocleavable biotin tethers one can directly use the crude aPCR reaction mix as the scaffold solution and excess staple strands (with one photocleavable biotin tether modification) to assemble a helix bundle or wireframe DNA origami nanoparticle and subsequently acquire purified sample with up to 90% yield.30 Photo-induced nucleotide damage at 365 nm is known to be minimal, making this approach viable (but untested) for gene-encoded DNA nanoparticle purification.
Scaffold features | Function | Benefit of incorporation | Ref. |
---|---|---|---|
Gene location: adapted from ref. 22 licensed under CC-BY. Internal crosslinking: adapted from ref. 22 licensed under CC-BY. Aspect ratio: reprinted (adapted) with permission from ref. 42. Copyright (2018) American Chemical Society. Scaffold orientation: used with permission from ref. 22 licensed under CC-BY. Promoter domain design: reprinted with permission from ref. 26. Copyright (2024) American Chemical Society. | |||
Kozak sequence | Initiates translation in eukaryotic cells; affects the probability of a ribosome recognizing the start codon | Ensures that the transcribed mRNA is efficiently translated into protein | 22 and 37 |
Woodchuck hepatitis virus post-transcriptional regulatory element (WPRE) | Enhances mRNA stability and protein expression | Increases transgene expression when placed downstream of the gene and proximal to the polyA signal | 38 and 39 |
Inverted terminal repeats (ITRs) | Single-stranded sequence of nucleotides directly followed by its reverse complement | Improves expression efficiency when placed upstream of expression cassette (mimics viral expression mechanism of AAV) | 22 and 41 |
Simian virus 40 (SV40) | Multiple transcription factors bind to SV40 DTS sequence within the cytoplasm; recruit transcription factors and bind with them in the cytoplasm | Direct transportation of DNA nanostructure to nucleus | 24 |
polyA sequence | Important for stability of mRNA upon nuclear export – prevents degredation | Ensures that the transcribed mRNA is stable enough to undergo translation | 22 |
Origami design feature/assessed variable | Definition | Summarized findings | Ref. | |
---|---|---|---|---|
Gene location | Internal versus external placement of gene on origami structure | Placement of the gene on a certain region of the origami structure does not have an effect on gene expression – it should be noted that in both cases, the promoter sequence was on the exterior of the structure | 22 | |
Internal crosslinking | Extra thymine residues were included on the staple strands to induce crosslinking via UV point welding – when structure is exposed to UV light, thymine dimers are created preventing the dissociation of scaffold and staples | These variants had almost complete suppression of eGFP signal – indicating that the unfolding of the origami must occur in order for the desired gene to be transcribed | 22 | |
Aspect ratio (AR) | Value comparing the length to width dimensions on origami structure | Tested Structures: 20 hb (AR = 5), 32 hb (AR = 2), 12 hb (AR= 15) | 22 and 42 | |
No difference in levels of expression between 20 hb and 32 hb; levels were slightly lower for 12 hb, but should be noted that cell density was also lower for this treatment group | ||||
Scaffold orientation | Varying the scaffold sequence to correspond with the coding (sense) or template (antisense) strand | Use of the template strand exhibited higher electroporation efficiencies when the scaffold was delivered alone and when the scaffold and staples were codelivered (non-annealed) – when either scaffold was folded into an origami structure, there was no significant difference between expression values | 22 | |
Promoter domain design | Analyzed three promoter variants: native double-stranded, scrambled (arbitratry DNA sequence replacing promoter) double-stranded, and single-stranded (no staple bound) | Single-stranded promoter regions significantly decrease gene expression and those with arbitrary sequences completely turn off expression | 26 |
To maximize efficiency of gene delivery, the gene-encoded nanoparticle would require efficient transport to the nucleus since it is the site of transcription, but transport is often a challenge due to the selective permeability of the nuclear membrane. The nuclear membrane consists of a lipid bilayer that is perforated by various nuclear pore complexes; these complexes regulate the passage of molecules between the cytoplasm and the nucleus. Larger molecules (such as DNA nanoparticles) can enter the nucleus more easily through active transport, but often require specific nuclear localization sequences (NLS) or DNA nuclear targeting sequences (DTS) that aid in nuclear import.43 To assist with the nuclear import of their mCherry-encoded DNA nanoparticle, Liedl et al. included multiple Simian virus 40 (SV40) derived DTS.44,45 The mCherry-encoded scaffold strand was bacteriophage-produced and included a CMV promoter, mCherry reporter gene, and a bovine growth hormone polyadenylation (bGH polyA) signal (coding/sense strand). Additionally, zero, one, three, or six repeats of the 72 bp SV40 DTS were inserted in the scaffold, however, the location of these repeats varied between each of the DNA nanoparticle constructs. In all cases, the coding domain (mCherry gene) was always on the outer helix of the nanoparticles (Fig. 3(E)). 24 hours after electroporating 0.75 μg (∼0.545 pmol) of these nanostructures into HEK293T cells, it was found that three repeats of the SV40 DTS downstream of the polyA signal maximized mCherry expression, as quantified by MFI and flow cytometry.
To account for the fact that gene placement did not have a significant impact on expression, it was hypothesized that the DNA nanostructure unfolded for the gene to be successfully transcribed by RNA polymerase in the nucleus. To test this, extra thymine residues were included on the staple strands to induce internal crosslinking via UV point welding (Table 1). When the nanostructures were exposed to UV light (310 nm, 2 hours), thymine dimers were created between neighboring staple strands, preventing the dissociation of the scaffold and staple strands upon delivery. In electroporating these crosslinked variants into the HEK293T cell line, there was almost complete suppression of the eGFP signal, leading to the interpretation that DNA nanoparticle unfolding into constituent staple and scaffold strands occurs during transcription. This is an exciting observation on the processing of DNA nanoparticles through transcription, but it is unclear whether there was any loss in biological activity in the gene-encoded DNA nanoparticle owing nucleic acid damage from prolonged UV exposure.
The next translation, thus, could be combining the spatial control afforded by modified staple strands to display targeting proteins and ligands with a gene-encoded scaffold strand to create gene therapy platforms. Recently, Wu et al. investigated the targeted delivery of a p53 gene in HeLa (human cervical epithelial) cells and in tumor xenograft BALB/c (Bagg Albino immunodeficient) nude mice through a gene-encoded DNA nanoparticle (Fig. 3(F)).25 The p53 gene is known to be an important tumor suppressing gene as it induces apoptosis and cell-cycle arrests.47 The DNA nanoparticle delivery system was created by joining two gene-encoded DNA origami monomer nanoparticles. Each monomer consisted of a custom scaffold strand generated via PCR. Interestingly, both scaffold strands were produced from the same precursor p53 encoding plasmid, but one scaffold encoded the sense strand while the other scaffold encoded the antisense strand. Each strand was folded into a pre-designed shape (using complementary staples) to form DNA origami monomers. The two monomers hybridized with each other to form the gene-encoded DNA origami (DO) construct.
The final DO construct (made of the two monomers) was structurally designed with short 8 nucleotide oligo extensions as tethers. These oligo extensions were incorporated to accelerate the disassembly of the gene-encoded DO construct during the transcription process, supporting efficient gene expression. To improve cellular uptake, the DO was coated with lipid via a templated growth method. A lipid–DNA conjugate was synthesized and precisely organized onto the DO's surface. Phospholipid with folate (DOPE-FA) was included in the lipid growth around the DO to enhance the targeted cellular uptake of the nanoparticles since folate receptors are often overexpressed in most tumor cells.48 The lipid-coated gene-encoded DO (FA-DO) was characterized by transmission electron microscopy. The DO and FA-DO were labeled with Cyanine5 (Cy5) and incubated with HeLa cells for 6 hours. Cy5 allowed for successful tracking of the DOs, and it was found that FA-DO had a stronger signal when compared to DO. This was confirmed by the MFI recorded through flow cytometry. Quantitative real time PCR (qRT-PCR) determined that the FA-DO had nearly 8 times higher p53 mRNA levels than DO only, and western blot analysis confirmed protein expression. Flow cytometry showed that 81.4% of HeLa cell apoptosis was induced by FA-DO. A cell-viability assay established a dose-dependent inhibition of viability. Remarkably, 80% of tumor inhibition was achieved under the dosage of 3.2 nM with no noticeable cytotoxicity.
The tumor inhibition effects of the FA-DO were investigated in vivo as well through a HeLa tumor xenograft model in BALB/c nude mice. The HeLa tumor bearing mice were treated with equal doses of DO or FA-DO (DNA scaffolds: 1.2 mg kg−1) via tail vein injections every three days for four treatments. By monitoring the tumor volumes and weight, as well as protein, mRNA expression and apoptosis levels, it was deduced that the FA-DO group was the best platform for tumor inhibition. The mice organs also did not demonstrate any observable system toxicity. These results demonstrate the validity of using DNA nanoparticles as a platform for gene delivery, particularly in gene therapy-based tumor inhibition. Wu et al. established that the complementary DNA strands of a functional gene can be directly used to form genetically-encoded DNA origami nanoparticle which can act as a template for lipid growth. The lipid-coating enables the DNA nanoparticle to penetrate the cell membrane in doses that do not elicit dangerous cytotoxicity levels.
One of the challenges of current gene delivery systems made from lipid and viral nanoparticles is the limited DNA loading capacity. For genes larger than 5 kilobases, it becomes necessary to cleave the gene into two parts and co-deliver using multiple delivery vehicles.49,50 DNA nanoparticles are advantageous in overcoming the loading capacity limits since one can engineer DNA nanoparticles of any size. Moreover, it is possible to co-deliver multiple genes simultaneously. To that end, the Dietz group developed multiplexed assemblies for the codelivery of eGFP and mCherry (each being 4816 nt long) genes. DNA nanoparticles encoding for these respective genes into the scaffold strands were created in stoichiometric ratios of 1:1, 1:2, and 1:3 mCherry to eGFP (Fig. 3(D)). The overall design concept involved creating individual origami blocks for each genetic sequence (mCherry and eGFP), which could then be linked together to form a single multifunctional origami structure. This linkage was achieved through complementary base pairing of extended staple sequences, and their modular approach allowed for precise control over the ratio of delivered genes within a single nanostructure. Here, they found that the expression level of eGFP was directly proportional to the number of monomers present within the nanostructure as confirmed by fluorescence microscopy and flow cytometry.22
On the design level, we also now have a preliminary understanding that the secondary and tertiary structure of DNA nanoparticles (crossover density and promoter location) can influence the overall protein expression efficiency. Surveying the studies that have contributed to this understanding, some studies were designed in in vitro bacterial IVT systems while others used mammalian cell electroporation setup. It is important to note that T7 polymerases are not present in mammalian cells, and therefore the behavior of these constructs may vary when comparing efficiencies in IVT systems versus a mammalian system. T7 polymerase originates from the T7 bacteriophage, a virus known for infecting bacteria; this polymerase is very specific to its promoter, and because of that, T7 systems can be tightly controlled in an experimental setting.52 For gene expression in mammalian cells, RNA polymerase II recognizes a CMV promoter. This promoter interacts with the general eukaryotic transcription machinery, allowing for expression in a wide range of mammalian cell types, but expression can ultimately be affected by discrepancies between cellular environments. Additionally, the complex nuclear environment of mammalian cells as well as accompanying post-transcriptional processes are not replicated in IVT systems. Despite these fundamental differences, DNA nanoparticles that retain a T7 promoter for expression in cell-free systems are able to provide insight on design features that affect gene expression in a more time efficient manner. Similarly, electroporation of gene-encoded DNA nanoparticles were among the first experiments demonstrating transgene expression in mammalian cells. The translational potential of electroporation is limited, as it is primarily suitable for ex vivo cell therapies or transdermal applications. Looking forward, the invasive nature and unpredictability of potential cell damage may need to be addressed for its use in many clinical scenarios. Thus, while both experimental systems (IVT and electroporation) are excellent tools for fundamental research, there is a noticeable need for investigations focused on mammalian cell transfections via gene-encoded DNA nanoparticles. Furthermore, several traditional methods used to augment translation of DNA in vivo are yet to be explored and implemented, such as codon usage and additional regulatory elements.
Despite significant advances in ssDNA production, strategies to produce gene-encoded scaffold strands are key to broadening the scope of the lengths and multiplexity of genes that can be encoded into these nanoparticles and accelerating their use in clinical applications. The bacteriophage-based scaffold synthesis has delivered milligram-scale quantities of ssDNA, and with proper standardization of quality control (and removal of endotoxins) it could be translated for production of gene-encoded scaffolds.53
Studies using DNA origami nanocarriers have consistently demonstrated low cytotoxicity in different cell lines.54–56 Several studies have also established the basic biosafety of DNA nanoparticles in vivo through body weight measurements and histopathological examinations of major organs (heart, liver, kidney, lung, etc.) after treatment with the DNA nanoparticle.57–59 Compared to carriers containing inorganic materials that the body cannot decompose or clear, DNA nanoparticles are a promising alternative that consist of organic monomers that can be processed by the body, eliminating the risk of unwanted accumulation. There remain a few aspects that must be thoroughly investigated prior to using DNA nanoparticles in gene therapy, but in most studies, DNA nanoparticles have not stimulated an immune response.57,59–61 In 2022, an M13mp18-encoded DNA nanoparticle was administered to ICR mice at a dosage of 12 mg kg−1 to maximize the potential immunostimulatory response. Here, the CD11b+ cell levels were monitored, as high cell populations tend to indicate inflammation or tissue damage, however, there was no indication that these DNA nanoparticles were the cause for adverse effects.60 Similarly in 2023, the Bathe group administered DNA nanoparticles (M13mp18 scaffold) intravenously to BALB/c mice at a dosage of 4 mg kg−1 to stay within the standard dose range for nucleic acid therapeutics (1–10 mg kg−1).62 Toxicity readouts were assessed, and it was found that there was no evident liver or kidney damage as indicated by histology.62 Despite these findings, Perrault et al. found that their DNA nano-octahedron induced an inflammatory cytokine response similar to that produced by foreign bacterial or viral nucleic acids.63 They combated this by encapsulating the nano-octahedron in lipid bilayers resembling viral membranes. The immune response and the pharmacokinetics of DNA nanoparticles are not yet fully understood, and continued research is crucial for the successful translation of these structures as a biosafe delivery platform.
The above-mentioned studies utilized DNA nanoparticles as a nanocarrier as opposed to a gene-encoded system. Gene-encoded DNA nanoparticles are yet to be scrutinized for concerns associated with currently applied gene therapy vectors since gene therapy research is heavily regulated due to safety concerns. Predominant gene therapy vectors such as AAVs have shown success in different trials but have also been associated with immune responses and off-target effects.64,65 In some cases, the immune responses can result in effects as severe as death.66 Wild types of Adenoviruses can also infect humans and so there remains a possibility that antibodies targeting these viruses can reduce the efficiency of AAV vectors. Additionally, the risk of unintended genomic integration can result in the inactivation and dysregulation of genes.67 Each of these risks must be fully investigated in gene-encoded DNA nanoparticles before they can be extensively used in gene therapy.
As a basic science tool, gene-encoded DNA nanoparticles have the potential to increase resolution of characterization of DNA nanoparticles to the base-pair level. They could be combined with an optical readout for probing at the base pair level the stability of DNA nanoparticles under different physical and chemical stressors since transcription (and concomitant protein expression) depends on having a contiguous gene sequence and a double stranded promoter region. Any damage to the scaffold strand would result in loss of protein expression, something that is not resolved by traditional means of studying DNA nanoparticle stability.17 For example, using a cell-free transcription–translation system we evaluated how promoter domain configuration within a gene-encoded DNA nanostructure affected its protein expression, as described above. In this case, protein luminescence was the optical readout which allowed time-resolved determination of the physical state of the DNA nanoparticle. UV crosslinking of thymine dimers is emerging as a way to increase DNA nanoparticle stability.68–70 In case of linear gene cassettes, considerable loss in biological activity can be seen post UV exposure. Perhaps a quick in vitro protein expression study can confirm the extent of photo-induced loss in biological activity in a DNA nanoparticle. The optical readout can be made multi-faceted by integrating chromophores engaged in multistep Förster resonance energy transfer (FRET) into gene-encoded nanoparticles. Previously, a strong correlation between the shape of small DNA nanoparticles and its cytosolic stability was observed using multi-step FRET and single cell manipulation, suggesting that future gene-encoded DNA nanoparticles would require additional modifications for resistance to degradation.71 Notwithstanding all the challenges that need to be addressed through further research,72 the promise of gene-encoded nanoparticles is increasingly positive and an area worth keeping an eye on.
This journal is © The Royal Society of Chemistry 2025 |