Next Article in Journal
Ultra-Narrow Bandwidth Microwave Photonic Filter Implemented by Single Longitudinal Mode Parity Time Symmetry Brillouin Fiber Laser
Previous Article in Journal
Velocity and Out-Step Frequencies for a Micro-Swimmer Based on Spiral Carbon Nanotubes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Basic Principles of RNA Interference: Nucleic Acid Types and In Vitro Intracellular Delivery Methods

by
Marie Isenmann
1,2,
Martin James Stoddart
1,2,
Rainer Schmelzeisen
1,
Christian Gross
1,
Elena Della Bella
2,† and
René Marcel Rothweiler
1,2,*,†
1
Department of Oral and Maxillofacial Surgery, Faculty of Medicine, University of Freiburg, Hugstetterstrasse 55, 79106 Freiburg, Germany
2
AO Research Institute Davos, Clavadelerstrasse 8, 7270 Davos, Switzerland
*
Author to whom correspondence should be addressed.
These authors contributed equally to the work.
Micromachines 2023, 14(7), 1321; https://doi.org/10.3390/mi14071321
Submission received: 18 May 2023 / Revised: 23 June 2023 / Accepted: 26 June 2023 / Published: 27 June 2023
(This article belongs to the Section B:Biology and Biomedicine)

Abstract

:
Since its discovery in 1989, RNA interference (RNAi) has become a widely used tool for the in vitro downregulation of specific gene expression in molecular biological research. This basically involves a complementary RNA that binds a target sequence to affect its transcription or translation process. Currently, various small RNAs, such as small interfering RNA (siRNA), micro RNA (miRNA), small hairpin RNA (shRNA), and PIWI interacting RNA (piRNA), are available for application on in vitro cell culture, to regulate the cells’ gene expression by mimicking the endogenous RNAi-machinery. In addition, several biochemical, physical, and viral methods have been established to deliver these RNAs into the cell or nucleus. Since each RNA and each delivery method entail different off-target effects, limitations, and compatibilities, it is crucial to understand their basic mode of action. This review is intended to provide an overview of different nucleic acids and delivery methods for planning, interpreting, and troubleshooting of RNAi experiments.

1. Introduction

In 1928, an astonishing observation was made during experiments with viruses on tobacco plants by Wingard et al., who observed that only the first leaves infected with the ringspot virus developed the full virus disease, while the upper leaves showed a miraculous recovery and resistance. Wingard was not able to explain this on a molecular biological level, but this recovery phenomenon formed the starting point for the discovery of the mechanism of RNA interference (RNAi).
More than 50 years later, Izant et al. showed that injection of complementary transcripts into mouse cells reduced the expression of specific genes [1]. The concept of “pathogen-derived resistance” was developed by Abel et al. in 1986, whereby defective expression of a pathogen’s gene product resulted in protection against that pathogen. However, this concept was still based on the assumption that the interaction occurred at the level of gene products (protein complexes) [2]. In 1989, Powell et al. studied the effect of antisense and satellite RNA and found that nucleic acid interaction was responsible for this protective effect [3].
The phenomenon of complementary nucleic acids inhibiting each other is conserved in most eukaryotes and is an indispensable part of the physiology of many organisms [4]. In plants, it balances the organism’s efforts between pathogen defense and growth [5]. As shown as early as 1928, RNA interference plays a central role in virus defense, as the double stranded RNA (dsRNA) triggers the plant RNAi system to silence complementary genes, thereby generating immunity [6,7]. In mammalian organisms, endogenous RNAi systems are important elements for the control of development, fate, and death of cells in various physiological and pathological states.
In experiments studying the effects of gene expression, targeted knockout has become an indispensable procedure to gain new insights. For a long time, culturing knockout organisms was the gold standard for this purpose, limited by the enormous time and costs, ethical concerns, and limited analysis of individual tissue or cells. RNA interference has overcome these obstacles, and today, there are numerous companies offering a wide range of artificial nucleotide acids for RNAi as well as various carrier systems for intracellular delivery.
In this review, we aim to provide an overview of the basic biological principles of RNA interference, available nucleotide acids, and in vitro delivery systems for use in gene expression experiments.

2. Principles of RNA Interference

The term RNA interference describes the principle of reducing the expression of a particular gene by complementary short RNAs. In general, this effect can be induced after transcription by mRNA cleavage or translation repression and at the transcriptional level by transcriptional silencing [8].
The process starts with RNA-dependent RNA polymerases (RdRPs) that generate long dsRNA from single stranded RNA templates [9,10]. The long dsRNA is transferred by endocytosis to the cytosol and processed by endoribonuclease Dicer or Tar-RNA-binding protein (TRBP) [11]. The resulting nucleic acid is called siRNA [12,13].
In the following steps, the nucleic acid is loaded onto the RNA-induced silencing complex, or RISC. This multiprotein RNA complex plays an essential role in the RNA silencing process. Its function is mainly based on Argonaute proteins that occur in subclades for miRNA, siRNA, or piRNA processes and contain specific domains [8]. The N-domain of Argonautes unwinds the single strand of RNA, while the PAZ-domain binds the 3′-overhang [14,15,16]. The MID-domain binds the 5′ end, while the PIWI-domain can cleave the target sequence [17,18]. In this process, the RISC selects one strand as the guide or antisense strand, while the complementary passenger strand is degraded. This strand selection is influenced by thermodynamic stability and nucleobase of the 5′ end [15]. The entire process is dependent on HSP90 proteins that keep the Argonautes in the correct conformation [19].
After the process of RISC loading, the attached siRNA is 20–27 nt long and directs the RISC to its complementary target sequence, either RNA or DNA. Subsequently, DNA methylation or chromatin modification inhibits the transcription of DNA (TGS), and mRNA cleavage or translation inhibition affects the posttranscriptional processes (PTGS) [20].
MiRNAs have a similar yet different origin, function, and purpose. The polynucleotides are approximately 18–26 nt long, single stranded, and in a stem-loop structure. Their production begins with the transcription of specific genes by RNA polymerase II [21]. The resulting pri-miRNA is capped and polyadenylated and further processed by the Drosha (RNAse III) and DGCR8 protein [22]. This pre-miRNA is hairpin-structured and exported from the nucleus by exportin 5 [23]. Subsequently, Dicer cleavage forms a 21 nt ds miRNA, which is loaded onto RISC similar to siRNAs [24]. However, in some cases both strands of miRNA (passenger and guide strand) bind a target sequence and affect gene expression [25]. Unlike siRNAs, miRNAs are often only partially complementary to their target sequence [26]. More specifically, they target 3′ UTR regions of mRNAs, with nucleotides 2–8 being in most cases fully complementary and referred to as the “seed region” with canonical binding, while the remaining part is only partially complementary [27]. In miRNAs, the target sequence-RISC interaction usually does not lead to cleavage by Dicer, but to recruitment of the GW182 protein [28]. Through the interaction with the cytoplasmic poly(A)-binding protein PABPC, GW182 induces both translation repression and mRNA deadenylation, the latter followed by 5′-cap removal and mRNA degradation by exoribonucleases [29] (Figure 1).

3. Nucleic Acids in RNAi

3.1. siRNA

siRNAs play an important role in gene expression silencing for research and potential therapeutic use. siRNAs are less likely than longer nucleotides to cause immune stimulation. They can be transported between different tissues in some species and are very precise due to their full complementarity. Overall, siRNAs are highly efficient tools for in vitro experiments and pose fewer problems than other interfering RNAs [30,31,32].
Nevertheless, many side effects have been observed with the widespread use of siRNA, which can currently be explained by three mechanisms:

3.1.1. miRNA-like Off-Target Effects

Off-target effects are often caused by siRNA binding to non-target genes that have partial complementarity to the 5′ end of their guide strand [33,34]. The exogenous siRNA essentially takes over the function of an endogenous miRNA, causing unintended effects on cell growth or altering other experimental outcomes [35,36]. This phenomenon is referred to as the “miRNA-like off-target effect”. Since this principle is part of cell physiology, it cannot be completely eliminated; however, there are strategies to reduce the likelihood that this phenomenon occurs.
One option is improving the siRNA sequence design by analyzing the whole genome of target cells and avoiding sequences that could induce miRNA-like effects [37,38].
Second, chemical modification of siRNA reduces off-target-effects by destabilizing the two strands. Since miRNA-like bonds are shorter than the intended siRNA target bond, they are more affected by this destabilization. This modification can be either 2′-O-methylation or locked nucleic acid (LNA) incorporation, which is particularly effective at position 2 of the 5′ end [35,38].
As a third option, lower siRNA-concentration has been shown to reduce these off-target effects [39,40,41]. Since a mere reduction in siRNA-concentration also reduces the target effects [40], the method of siRNA-pooling was developed. The method uses a pool of siRNA sequences that all target the same gene but bind at different sites. There are several strategies for creating such a pool. The least complex pools are produced by combining just a small number of siRNAs that share the same target gene (e.g., smart pools with four different siRNAs). Their dilution effect, which should reduce the miRNA-like off-target effects, is relatively low. This dilution effect is increased with endoribonuclease-produced siRNA pools (esiRNA), produced by digestion of dsRNA using RNAse III [42] which results in hundreds of different siRNAs [43]. Third, so-called siPools with about 30 different sequences, in contrast to esiRNA, are designed in vitro and therefore are well-defined but also more costly to produce. They eliminate sequence-specific off-target effects such as esiRNA, while it is much easier to control and understand their effects on cells [43].

3.1.2. Immunostimulatory Response

SiRNA avoids some immunostimulatory response due to their size of less than 30 nt. Nevertheless, they can still trigger an immune response [44]. The immune activation is concentration-dependent and detectable with each siRNA application [45]. Besides dsRNA, carriers can also be immunostimulatory triggers. Endosomal transfection systems have been shown to be much more likely to cause immune stimulation because the endosome contains many immune-activating receptors [44,46,47]. Many of the effects are also sequence-dependent, which reduces the informative value of nonbinding sequences as negative controls but can be reduced by avoiding immunostimulatory motifs [48]. Cell type also influences appearance and extent of the immune activation [49].
In general, there are three distinct signaling pathways for siRNA-induced immune stimulation.
First, dsRNA can bind dsRNA recognition proteins, which triggers antiviral responses and causes upregulation of Interferon (IFN) and other antiviral proteins. IFN activates IFN-stimulated genes (ISG) such as PKR that inhibit viral replication and protein synthesis [50]. Second, dsRNA activates oligoadenylate synthetases (OAS). They convert ATP to oligoadenylates, thereby activating RNAse L, which is capable of degrading intracellular single-stranded RNA [51]. Third, dsRNA binds to Toll-like receptors (TLR) and to transcription factor IRF3, leading to the induction of IFN, TNF-alpha, and IL6 [44].
Activation of the cell immune system can have many complex effects on the target cells and the experimental outcome. Chemical modification of siRNA, such as 2′-O-methylation or locked nucleic acid incorporation, is one approach to reduce this problem [52].

3.1.3. Saturation of Endogenous RNA Interference

Exogenous siRNA can affect the endogenous RNA interference machinery of cells. High siRNA concentrations lead to intense siRNA-RISC loading, which may reduce the ability to generate RISC for miRNA-induced silencing. The reduced miRNA suppression may lead to undesired gene expression, thereby affecting cell phenotype [53].
Considering all these challenges, choosing an appropriate siRNA sequence is not trivial. This sequence should not only be complementary to the target gene but also very specific and preferably not affecting other genes or signaling pathways in the cell. Sequence design is mainly done in silico applying several rules. A low G/U content is preferred as this reduces immunostimulation [54], as is low internal stability at the 5′ end of antisense strands to facilitate RISC entry. Stable internal repeats are avoided since they cause internal folding that interferes with target binding [55].

3.2. miRNA

MiRNAs, short for micro RNAs, are 21 to 25 nt long, occur ubiquitously in eukaryotic cells, and form a stem-loop-structure [56]. They not only inhibit gene expression but can also induce transcription by mRNA-promoter binding [57]. Most miRNAs are formed by modifying specific DNA-transcripts (pre-miRNAs) that are exported from the nucleus and processed by Dicer enzymes. In addition, there are other, non-canonical pathways for miRNA production. These include so-called “mirtrons”, spliced introns of mRNA, fully Dicer-independent miRNA which is derived from endogenous shRNA processed by Drosha and cleaved by hAgo2 (Human Protein argonaute-2) or m7g (7-methylguanosine)-capped pre-miRNA that can be exported to the cytoplasm without Drosha cleavage [58,59,60]. In many cases, multiple miRNAs are transcribed as one long transcript (cluster) that is subsequently cleaved. These “miRNA families” usually bind similar seed regions [61].
Unlike other small RNAs, miRNAs are able to move between different compartments of an organism and can therefore be detected in extracellular fluids [62,63].
Currently, miRNAs have gained importance due to their expression in various diseases, especially cancer [64].
However, miRNAs are less suitable for in vitro analysis of gene expression and for experiments that require precise gene silencing. Their complementarity is not perfect, resulting in unstable and non-specific mRNA binding that can even be toxic [26]. The main benchside application of miRNA is to analyze and validate their expected effects on gene expression and phenotype of cells to decide on further investigations and possible therapeutic applications [65]. To this end, cell cultures are transfected with a miRNA mimic and a scramble sequence [66]. However, in these experiments, miRNAs show many side effects, such as causing interferon response, strand bias, or unspecific binding to non-target sites [67,68,69]. For this reason, miRNA inhibitors are the preferred approach for miRNA validation, especially miRNA sponges. MiRNA sponges are plasmids that contain many miRNA binding sites [70]. To avoid RNAse H activity, their sequences are not perfectly complementary to miRNAs. To avoid unintended binding, their design is quite complex, and they are mostly planned by using webtools such as miRNAsong, whereas also engineered circular RNA (circRNA) with miRNA-sponging function may be used [71,72,73].

3.3. shRNA

Short hairpin RNAs (shRNAs) are RNA sequences that form a tight hairpin based on their sequence consisting of a target specific part, a spacer, and a reverse complement of the target sequence [74].
To achieve more stable knockdown experiments, researchers have been inspired by the design of endogenous pre-miRNA in the development of shRNAs [75,76].
Usually, shRNA sequences are introduced into the cell by vectors (e.g., plasmids) and must be transcribed in the nucleus to obtain the hairpin-structured shRNA. Based on their transcription pathway, current shRNAs can be divided into first and second generation.
The first generation of shRNAs uses RNA polymerase III promoters in their vectors, in most cases the U6 and H1 promoter [77,78,79]. Transcription produces stem-loop-structured, pre-miRNA-like shRNAs in the cell. These shRNAs can be processed into more potent RNA interference nucleotides than those provided by endogenous mechanisms [80].
However, first generation shRNAs cause many off-target effects that lead to toxicity and disruption of endogenous miRNA [81,82,83].
Second-generation shRNAs mimic pri-miRNAs, a preform of pre-miRNAs that requires an additional processing step [76,84]. Their gene template is transcribed by RNA polymerase II. This transcription process involves capping and poly-A tailing [84,85]. In comparison to the first generation, this approach is more adaptable and offers the possibility of transferring entire shRNA clusters [86,87]. However, second-generation shRNAs are less well understood and more complex.
After transcription, shRNAs are processed into siRNA. This is achieved with the help of the cell’s endogenous RNAi-processing machinery. For the shRNAs to be recognized and processed by the endogenous pathways, specific signals are required in the shRNA. Since these design requirements are quite complex, endogenous miRNAs are currently used as templates for the design of shRNAs [88]. Another challenge is that cleavage sites for the same shRNA sequence have been shown to be inconsistent. Rules for length and loop position may mitigate this disturbance [89,90].
RISC loading of the resulting siRNA can be improved by aiming for hAgo2 cleavage-dependent RISC formation. Strand selection is improved by designing the 5′ end of the guide strand to be less stable than the passengers one [91,92]. For ideal target sequence binding, imperfect complementarity has been shown to result in a weaker outcome and more off-target effects [93]. Perfect matches, whereas complementarity at the 3′ end is negligible, result in more efficient hAgo2-dependent cleavage of the target [94].
Overall, shRNA systems have many advantages over siRNA. Their effect on cell gene expression lasts much longer because the vector often remains in the cell and is transcribed more than once [75]. Moreover, controllable vectors can be designed by inserting selection markers or inducible elements into the sequence and its promoter.
Nevertheless, the entire shRNA system is very complex and still not well understood. Identical shRNA sequences are processed differently in different cell lines, causing miRNA-like off-target effects and immune stimulation, that cannot yet be avoided by improved shRNA design [48,95,96]. Furthermore, shRNA must be transcribed in the nucleus, requiring vectors with precise nuclear delivery [74]. Not least, shRNA utilizes many parts of the cell’s endogenous RNAi system, which can easily lead to saturation of, e.g., exportin 5 or Argonaute proteins and thus severely disrupt the cells’ gene expression regulation [97,98].
Currently, shRNAs are widely used to transduce cells for efficient gene knockdown. They can enable mass production of siRNA in vitro, and their potential future role in treatment of viral diseases should not be underestimated due to numerous ongoing research and trials in different phases [86,99].

3.4. piRNA

PIWI-interacting RNAs (piRNAs) are 21–35 nt long single stranded nucleic acids that carry a 2′-O-methylation at their 3′ end, uridine as a terminal base at the 5′ end, or adenosine at the tenth position [100,101,102]. They do not share a specific common secondary structure [100].
PiRNAs were first identified in animal germ cells. They were found to be produced in a Dicer-independent manner, copied from non-coding genomic regions with repeats, and are an important player in posttranscriptional regulation, particularly in protecting germline cells from transposable elements (TE) [103]. PiRNAs have also been detected in somatic cells, where they are required for epigenetic regulation through methylation, transposon silencing, and chromatin modification. Their importance is particularly evident in various malignant pathogenesis pathways [104].
PiRNAs interact with PIWI proteins. PIWI proteins represent a subfamily of Argonautes and therefore play an important role in the formation and function of RISC. In this context, PIWI proteins have an endonuclease function and can cleave RNAs [105].
PiRNAs affect cell gene expression through various mechanisms. In transcriptional gene silencing (TGS), piRNA/PIWI protein complexes bind the target gene, methylate DNA, and modify histones [106,107,108]. In post-transcriptional gene silencing (PTGS) piRNAs act similarly to miRNAs and form a piRISC on mRNAs to prevent their translation [109,110]. Furthermore, piRNA/PIWI protein complexes modify posttranslational processes (PTM) by interacting with transcription factors, leading to their posttranscriptional phosphorylation [111,112].
Since piRNAs bind nonspecifically to different targets and their effects in cells are not yet well predicted, they are currently not used for gene expression experiments. Nevertheless, their role in controlling Tes could provide an approach for therapy in cancer or other diseases [113].

3.5. ASO as an Alternative to RNAi

The use of Antisense Oligonucleotides (ASOs) is an alternative approach to RNAi for the regulation of gene expression. Considering the common goals and shared challenges of RNAi and antisense approaches, ASOs are herein discussed.
ASOs are short, synthetic, single-stranded oligonucleotides (both DNA- and RNA-based) with antisense function, and they downregulate gene expression via different mechanisms [114]. Because of a DNA:RNA heteroduplex formation, some induce Rnase H-mediated target cleavage [115]. Others induce cleavage by hAgo2 and other Argonaute proteins. In addition, there are ASOs that only occupy their target, thereby either preventing translation and causing cleavage through the resulting arrest or promoting translation through altered splicing [116].
When first developed, ASOs were found to be toxic, rapidly degradable, and difficult to transfer through membranes due to their negative charge [117]. Today, multiple ASO modifications are established to overcome these obstacles [118,119].
Compared to nucleic acids for RNAi, ASOs have been shown to be more flexible. They comprise both hydrophobic and hydrophilic parts, making them amphiphilic [120]. Interestingly, discoveries in siRNA design have improved ASO development and vice versa [121]. ASOs can cross the cell membrane in different ways. Most ASOs are modified with phosphorothioates (PS-ASOs), which allow them to bind surface proteins and enter the cell through endocytosis [122]. After passive diffusion through nuclear pores, ASOs can bind their target sequence and initiate various pathways [123].
As mentioned previously, the most common ASO modification today is PS-ASO, in which the phosphodiester in the backbone is replaced by a phosphorothioate at one or more sites [120]. This modification increases the distance between the charged parts, making the molecule more lipophilic and thus facilitating protein binding [124].
Modification of 2′-C of ribose increases stability, target affinity, avoids DNA:RNA heteroduplex formation, but can also trigger inflammatory processes [116,119,123]. When RNAse H degradation is intended, 2′ modifications in the target-binding core should be avoided, only the extremities can carry modifications to increase stability (“gapmer” structure) [125]. In many cases, the core of this gapmer structure contains deoxynucleotides, with RNA flanking regions. These chimeric DNA-RNA molecules enable the formation of DNA:RNA-duplexes with the target RNA, which are well recognized by Rnase H [123].
For specific drug delivery, ASO can carry specific conjugates. For example, N-Acetylgalactosamine (GalNAc) bound to PS-ASO enhances delivery to hepatocytes, while glucagon-like peptide 1 (GLP-1)-PS-ASOs are specifically delivered to the pancreatic beta cells [123,126].
Challenges in using ASOs for in vitro knockdown include high nonspecific signals by scramble sequences, no significant knockdown, and viability reduction [127,128]. As ASO design is complicated, most researchers purchase them from manufacturers. Lacking knowledge of the exact sequence or chemistry, it is much more difficult to interpret nonspecific signals and optimize design [129,130,131]. Nevertheless, there are already approaches and studies using ASOs in therapeutic contexts to treat viral diseases, genetic alterations, cancer, chronic inflammation, and COVID-19 [130,132,133].

4. Intracellular Delivery

4.1. Biochemical Methods

4.1.1. Lipid-Based Delivery

The first established lipid-based delivery method was lipofection (or lipoplex-based delivery), in which nucleic acids, lipids and polymers form complexes [134]. These complexes are mainly introduced into the cell by endocytosis, while also fusion to the membrane occurs in some cases [135].
The cationic lipids of lipoplexes interact with and neutralize negatively charged nucleic acids [136]. They contain a positively charged polar head, a hydrophobic tail, and a linker bond [137]. The type, length, and orientation of linkers have a critical impact on the efficiency, toxicity, stability and biodegradability of lipoplexes [138]. In addition, linkers can be designed to be environmentally sensitive and can be altered by pH, oxidation, or enzymes [139,140]. The most widely used and best characterized cationic lipid is the ether-linked 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP) [141].
The neutral lipids contain phosphatidylethanolamines, phosphatidylcholines, or cholesterol and very often 1,2-dioleoyl-3-glycero-phosphatidylethanolamine (DOPE) is used. They decrease cytotoxicity and increase transfection efficiency [142,143].
However, lipofection shows to have several side effects and disadvantages.
The delivery of cationic lipids depends mostly on the cellular endocytosis system, whose function varies between different cell types and is very sensitive to factors that disrupt these endocytic pathways [144,145,146] Furthermore, lipofection reduces cell viability, which is mainly caused by the cytotoxicity of their headgroups [147] but also by stimulating pro-inflammatory pathways through binding of pattern recognition factors (PPRs) in endosomes [145,148,149,150,151]. Beyond that, lipofection shows relatively slow and weak effects caused by the high lysosomal degradation during endosomal delivery [152].
An approach to address these issues are multi-component lipoplexes, that increase efficiency 10–100-fold by destabilizing the endosome and adjusting pH through polymers [153,154,155,156]. Another possibility is the protection of nucleic acids by albumin, chitosan, or protamine. In some cases chloroquine is added to improve endosomal release [157].
To bypass endocytosis, fusogenic liposomes were developed as another lipid-based delivery agent [158]. Here DOPE, DOTAP, and an aromatic molecule are used to create a cationic liposome that fuses directly with the negatively charged cell membrane without requiring interaction with a protein [159,160,161]. This is achieved by electrostatic interactions of a delocalized π-electron system [158,162]. As a result, nucleic acids are delivered directly to the cytosol [163].
Neutral lipids can be used as a control element, as smaller head groups increase fusion efficiency [162]. However, the ratios need to be optimized as neutral lipids on the one hand reduce toxicity, but on the other hand can disrupt the interaction between the positive liposome and the negative cell membrane [164]. Overall, fusogenic liposomes are more efficient, cause less cell death and achieve much faster effects than lipoplex-based systems and therefore represent an attractive alternative for in vitro gene expression experiments [145].
Especially in vivo, so-called lipid nanoparticles (LNPs) are used. Unlike standard lipoplexes, they can carry ionizable lipids instead of cationic lipids, which are pH sensitive and can adjust their electrical charge to the environment [165]. In addition, lipids can be modified with PEG residues to be exposed on the surface of the liposome, preventing serum protein uptake, phagocytosis, and aggregation, and can be functionalized to bind specific targets, while often impeding cellular uptake and endosomal release [166].

4.1.2. Polycationic Polymers

Another approach for oligonucleotide delivery is represented by the use of polycationic polymers that form polyplexes with negatively charged nucleic acids through electrostatic interaction to facilitate membrane passage and improve stability [167]. Commonly used polycationic polymers include polyethylenimine (PEI), polyaminoethyl methacrylate, and dendrimers [168]. Polyplexes can be modified to allow active and passive targeting, stimulation of endosomal release, and encapsulation of other drugs [168,169]. Despite numerous modifications and developments in polymer technology, they still have low biodegradability, which often leads to cytotoxicity and limits their application [170]. DNA-inspired nucleic acid vehicles may solve this problem [171].

4.2. Physical Methods

4.2.1. Electroporation

Delivery by electroporation is based on the principle that an electric field applied to a cell increases its membrane permeability. This is achieved by raising the transmembrane potential (TMP) above a certain threshold. For example, in eukaryotic cells normal TMP is at −0.07 V and the threshold at which permeability is increased is around 0.2–0.5 V [31] (Figure 2).
In this process, higher TMP raises the energy level of the membrane, causing formation of random hydrophobic pores. With TMP staying elevated, liquid enters the pores, lipids turn around and form hydrophilic pores, which is called reversible electroporation (RE) [172]. If the TMP is raised above a threshold (around 1 V), cells cannot restore a closed membrane anymore, which is called irreversible electroporation (IRE) [173]
The TMP, which is critical for the efficiency of transmission and cell viability, is determined by several parameters such as membrane diameter, cell shape and radius, the applied electric field, and the angle of the field to the cell [174]. An important factor is also the conductivity of extracellular fluid, membrane, and cytoplasm, which changes dynamically during electroporation due to ion flow [175].
The pulse frequency has an enormous influence on efficiency. In most cases, low frequencies of 1–10 Hz are chosen [176]. They are particularly suitable for longer pulses (around 100 µs) [177]. Higher frequencies may cause side effects especially in vivo [178]. However, for nanosecond pulses, higher frequencies increase efficiency [179]. Very high frequency pulses can accumulate in cells and reduce the threshold of energy required for RE [180,181]. In contrast, very low frequencies (0.1–1 Hz) can increase efficiency by electrosensitization of the membrane [182,183].
There are numerous different electroporation systems used in experimental research. Based on their size, they can be divided in major 3 groups: macro-, micro-, and nanoscale electroporation.
During macroscale electroporation, also called bulk or cuvette electroporation, multiple cells are treated at once in chambers with a diameter of at least 1 mm, providing a straightforward, inexpensive, and high-throughput transfection method [184].
Microscale electroporation is performed in chambers or channels with a diameter of micrometers. It offers many advantages over the bulk approach: smaller electrodes and lower voltages are required, therefore being better at maintaining cell viability; as surface-to-volume ratio of cells increases, there is less heat dissipation; the possibility of real-time monitoring; electrode positions can be adjusted to allow electroporation of individual cells while maintaining high-throughput through flow devices [185]. Microscale systems can have parallel or transverse electrodes, contain channels of varying width for locally enhanced electric fields, or specialized microfluids that enable droplet-based electroporation [186,187,188].
In nanoscale electroporation, charged fluids pass nanostructures (nanochannels, nanopores or nanostraws). This allows electric fields to be applied very precisely to specific membrane regions of a cell [189,190,191]. Nanofountains even exhibit a gun-like structure by applying an electric field generated with an atomic force microscope through a microcatheter with an opening of less than 1 µm [192]. While the precision of nanoscale electroporation cannot be surpassed by other systems, nanoscale electroporation is very complex and expensive to establish and has low throughput.
Overall, electroporation is a relatively inexpensive and safe approach for intracellular delivery. It is feasible for many cell types, especially primary cells where viral transduction is often insufficient [193,194,195,196,197,198,199]. However, the system configuration to achieve an appropriate TMP is challenging, because several factors need to be considered and high TMPs can reduce viability through IRE, while when TMPs are too low, the applied energy is used for heat dissipation, electrophoresis and electrolysis [173,174,175,200,201].

4.2.2. Sonoporation

During sonoporation, acoustic waves are applied to cells or the fluid surrounding them to disrupt the cell membrane [202]. Most sonoporation systems rely on bubbles in the surrounding fluid (bubble-based), whereas newer approaches can disrupt the membrane without bubbles (non-bubble-based).
For bubble-based sonoporation, 3 main mechanisms are currently in use.
“Inertial cavitation” uses the jet flow generated by the bursting of bubbles due to sound waves. This jet flow leads to rupture of the cell membrane, and the fluid stream generated by the collapse also leads to membrane perforation [203,204]. However, irreversible pores lead to cell death and unstable byproducts, such as temperature dissipation and reactive oxygen species, which decrease viability [203,205,206].
“Stable cavitation” uses the shear stress generated by the fluid stream of oscillating bubbles to disrupt membranes [207]. The approach has fewer side effects than inertial cavitation, and the bubbles can also adhere directly to the cell membrane and open micropores [208,209,210]. However, this method requires precise bubble size and bubble-to-cell distance, which is often difficult to maintain even under experimental conditions [209,211].
“Acoustic radiation force” as a third mechanism pushes bubbles through the cell membrane, creating holes in it [212,213]. Factors such as bubble size, acoustic impedance, and acoustic energy density must be adjusted to achieve satisfactory results [214].
The challenges of bubble-based sonoporation are the need for a contrast agent, a specific bubble distance and bubble-to-cell ratio [209,211,215].
In non-bubble-based mechanisms, three main forces are applied to the cell: acoustic radiation force, shear force due to acoustic streaming and energy applied by an adherent substance stimulated by acoustic waves [216,217,218]. These forces stress the cell membrane, leading to pore formation. The radiation force is used to push the cells to a pressure node where they can be observed, or to push them through constricting nozzles or against walls to increase membrane stress [219,220,221]. Cells attached to acoustically stimulated substrates are directly exposed by their attachment [222]. High frequency acoustic waves as concentrated acoustic radiation can precisely target one single cell [223,224]. Hyper-frequency acoustic waves or focused transducers can even achieve membrane disruption by the stream of acoustic waves [225,226].
In summary, sonoporation is a promising tool for intracellular delivery that is suitable for various cell types and cargoes. It can be combined with other delivery methods [227,228]. Nevertheless, there are still many challenges: thermal dissipation can affect cell viability, reactive oxygen species can cause apoptosis and necrosis, and genotoxicity has also been observed [229,230,231,232].

4.2.3. Microinjection

The oldest method to transfer genetic material into a cell is microinjection. Using a glass pipette with a diameter of 0.5–15 µm, fluids can be injected into floating and adherent cells [233]. This allows for targeted delivery into single cells, such as zygotes, to generate transgenic organisms [234]. However, this method has a particularly low throughput and requires an experienced researcher for cell holding, injection site selection, and volume [235]. Automated microinjection systems are currently being developed to address this challenge [235].

4.3. Viral Transduction

In 1967, it was discovered that adenoviruses can transiently regulate the gene expression of a cell [236]. The adenovirus genome contains so-called Early genes (E genes), that control the viral life cycle. Of these, E1A is required to initiate viral replication, while E3 does not play a critical role for viral survival or replication [237].
Adenoviruses used for intracellular delivery include replication defective and conditionally replicating adenoviruses. Replication defective adenoviral vectors lack E1 and E3. Therefore, they cannot replicate but provide space for insertion of external genes [238]. They are commonly used for gene delivery in in vitro research. For construction, the gene of interest (GOI) is cloned into a plasmid vector. The final plasmids contain at least the GOI (usually shRNA in case of RNAi) in an open reading frame (ORF), a promoter and a marker gene. The adenovirus is then transfected into packaging cells that express E1A and allow adenoviral reproduction [239,240]. The replicated adenovirus, particles lacking DNA and cellular debris are separated by ultracentrifugation so that the final adenovirus contains the GOI and lacks E1 and E3, preventing it from replicating in humans. Adenoviral vectors release their genome into the nucleus, where it is not integrated into the genome but remains in the episomal state for transcription, is retained much longer than non-virally delivered nucleic acids, but considerably reduced by cell division [241,242] (Figure 3).
Conditionally replicating adenoviruses (CRA) have the E1A promoter replaced by a cancer-specific one [243]. This modification restricts viral replication to cancer cells, while benign cells are unaffected. Nevertheless, its clinical application has not been successful so far. Target specific CRAs carry promoters that depend on several factors, and viral replication is possible only when these factors are present [244].
Adeno-associated viruses (AAV are non-enveloped, single-stranded DNA viruses belonging to the Parvoviridae family [245]. Because they have low pathogenicity and immunotoxicity, a high safety profile in clinical trials, long-lasting transgene expression, and a simple genome that is easy to be modified, AAV are promising candidates for in vivo drug delivery [245,246,247]. However, AAV are dependent parvoviruses as their replication is dependent on other viruses [248]. Despite intensive research on stable production lines in recent decades, the production of high quantities of AAV is very time-consuming and costly [248]. In addition, AAV can cause damage to insertion sites and have limited capacity for transgene cargo [249,250]. Overall, viral delivery is still the most efficient and durable method for gene transfer into most cell types [251]. Nevertheless, its price and especially its higher risk profile for insertional mutagenesis and immune responses dampen enthusiasm about its use and potential, especially for in vivo therapies [252].
Another common virus family for gene transfer are lentiviruses. In lentiviruses, genes encoding viral structural proteins can be replaced by GOIs, and unlike adenoviruses, they integrate their genes into the genome of infected cells [253,254]. For this reason, modulation of gene expression with lentiviruses is exceptionally long and stable, but also entails more oncogenic risks depending on the insertion site. Lentiviruses are well suited for gene transfer as they elicit little immune response while inducing stable transgene expression [255].

5. Conclusions

Numerous discoveries in the field of RNA interference and intracellular delivery have been reported in the past few decades. Today, researchers can choose from a vast array of methods to perform their gene expression experiments. However, knowledge of background processes, pitfalls, and compatibilities with cells and cargo is indispensable for appropriate method selection, correct application, and meaningful interpretation. For this reason, the review provides an overview and orientation for all those approaching RNA interference and relative in vitro application.

Author Contributions

M.I.: Conceptualization, Resources, Writing—Original Draft, and Project Administration. M.J.S.: Supervision, Writing—Original Draft and Project Administration. R.S.: Supervision and Writing—Review and Editing. C.G.: Writing—Review and Editing. E.D.B.: Supervision, Writing—Original Draft, and Project Administration. R.M.R.: Supervision, Writing—Original Draft, and Project Administration. All authors have read and agreed to the published version of the manuscript.

Funding

Funding was available for the project from AO Foundation; AOCMF, Grabenstrasse 15, 7000 Chur (Switzerland) [AOCMFS-18-25R]. We furthermore acknowledge support by the Open Access Publication Fund of the University of Freiburg.

Data Availability Statement

All data can be requested from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AbbreviationDefinition
hAgo2Human Protein argonaute-2
ASOAntisense Oligonucleotides
ATPAdenosine triphosphate
COVID-19Coronavirus disease 2019
CRAConditionally replicating adenovirus
CsClCesium chloride
DGCR8 proteinDiGeorge syndrome critical region 8 protein
DOPE1,2-dioleoyl-3-glycero-phosphatidylethanolamine
DOTAP1,2-dioleoyl-3-trimethylammonium-propane
dsDNADouble stranded DNA
E1–4Early-transcribed regions 1–4 in adenoviruses
esiRNAEndoribonuclease-produced siRNA pools
GalNAcN-Acetylgalactosamine
GLP-1glucagon-like peptide 1
GOIgene of interest
GW182Protein Gawky
HSP90Heat shock protein 90
IFNInterferon
IL6Interleukin-6 gene
IREirreversible electroporation
IRFInterferon regulatory factor
ISGIFN-stimulated genes
LNAlocked nucleic acid
m7g7-methylguanosine
miRNAmicroRNA
mRNAmessengerRNA
OASoligoadenylate synthetases
ORFopen reading frame
PABPC proteinpoly(A)-binding protein cytoplasmatic
piRNAPIWI-interacting RNA
PIWIP-element induced wimpy testis
PKRProtein kinase R
PPRpattern recognition factor
pre-miRNAPrecursor-miRNA
pri-miRNAPrimary-miRNA
PS-ASOPhosphorothioate-modified ASO
PTGSpost-transcriptional gene silencing
PTMPosttranscriptional modification
rAdVrecombinant adenoviruses
RdRPRNA-dependent RNA polymerase
REreversible electroporation
RISCRNA-induced silencing complex
RNAiRNA interference
RNAPolRNA polymerase
RNAseRibonuclease
shRNAShort hairpin RNA
siRNASmall interfering RNA
sRNASmall RNA
TEtransposable element
TGStranscriptional gene silencing
TLRToll-like receptor
TMPtransmembrane potential
TNF-alphaTumor necrosis factor alpha
TRBPTar-RNA-binding protein

References

  1. Izant, J.G.; Weintraub, H. Inhibition of thymidine kinase gene expression by anti-sense RNA: A molecular approach to genetic analysis. Cell 1984, 36, 1007–1015. [Google Scholar] [CrossRef] [PubMed]
  2. Abel, P.P.; Nelson, R.S.; De, B.; Hoffmann, N.; Rogers, S.G.; Fraley, R.T.; Beachy, R.N. Delay of disease development in transgenic plants that express the tobacco mosaic virus coat protein gene. Science 1986, 232, 738–743. [Google Scholar] [CrossRef] [PubMed]
  3. Powell, P.A.; Stark, D.M.; Sanders, P.R.; Beachy, R.N. Protection against tobacco mosaic virus in transgenic plants that express tobacco mosaic virus antisense RNA. Proc. Natl. Acad. Sci. USA 1989, 86, 6949–6952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhang, Y.; Yun, Z.; Gong, L.; Qu, H.; Duan, X.; Jiang, Y.; Zhu, H. Comparison of miRNA Evolution and Function in Plants and Animals. Microrna 2018, 7, 4–10. [Google Scholar] [CrossRef]
  5. Qiao, Y.; Xia, R.; Zhai, J.; Hou, Y.; Feng, L.; Zhai, Y.; Ma, W. Small RNAs in Plant Immunity and Virulence of Filamentous Pathogens. Annu. Rev. Phytopathol. 2021, 59, 265–288. [Google Scholar] [CrossRef]
  6. Ratcliff, F.; Harrison, B.D.; Baulcombe, D.C. A similarity between viral defense and gene silencing in plants. Science 1997, 276, 1558–1560. [Google Scholar] [CrossRef]
  7. Wingard, S.A. Hosts and Symptoms of Ring Spot, a Virus Disease of Plants; USDA: Washington, DC, USA, 1928; pp. 127–153. [Google Scholar]
  8. Shabalina, S.A.; Koonin, E.V. Origins and evolution of eukaryotic RNA interference. Trends Ecol. Evol. 2008, 23, 578–587. [Google Scholar] [CrossRef] [Green Version]
  9. Mello, C.C.; Conte, D. Revealing the world of RNA interference. Nature 2004, 431, 338–342. [Google Scholar] [CrossRef]
  10. Kim, D.H.; Behlke, M.A.; Rose, S.D.; Chang, M.S.; Choi, S.; Rossi, J.J. Synthetic dsRNA Dicer substrates enhance RNAi potency and efficacy. Nat. Biotechnol. 2005, 23, 222–226. [Google Scholar] [CrossRef] [Green Version]
  11. Zuckerman, J.E.; Davis, M.E. Clinical experiences with systemically administered siRNA-based therapeutics in cancer. Nat. Rev. Drug Discov. 2015, 14, 843–856. [Google Scholar] [CrossRef]
  12. Filipowicz, W. RNAi: The nuts and bolts of the RISC machine. Cell 2005, 122, 17–20. [Google Scholar] [CrossRef] [Green Version]
  13. Treiber, T.; Treiber, N.; Meister, G. Regulation of microRNA biogenesis and its crosstalk with other cellular pathways. Nat. Rev. Mol. Cell Biol. 2019, 20, 5–20. [Google Scholar] [CrossRef]
  14. Kwak, P.B.; Tomari, Y. The N domain of Argonaute drives duplex unwinding during RISC assembly. Nat. Struct. Mol. Biol. 2012, 19, 145–151. [Google Scholar] [CrossRef]
  15. Sheu-Gruttadauria, J.; MacRae, I.J. Structural Foundations of RNA Silencing by Argonaute. J. Mol. Biol. 2017, 429, 2619–2639. [Google Scholar] [CrossRef]
  16. Ma, J.B.; Ye, K.; Patel, D.J. Structural basis for overhang-specific small interfering RNA recognition by the PAZ domain. Nature 2004, 429, 318–322. [Google Scholar] [CrossRef] [Green Version]
  17. Parker, J.S.; Roe, S.M.; Barford, D. Crystal structure of a PIWI protein suggests mechanisms for siRNA recognition and slicer activity. EMBO J. 2004, 23, 4727–4737. [Google Scholar] [CrossRef] [Green Version]
  18. Nakanishi, K. Anatomy of four human Argonaute proteins. Nucleic Acids Res. 2022, 50, 6618–6638. [Google Scholar] [CrossRef]
  19. Dueck, A.; Meister, G. Assembly and function of small RNA—Argonaute protein complexes. Biol. Chem. 2014, 395, 611–629. [Google Scholar] [CrossRef]
  20. Ghildiyal, M.; Zamore, P.D. Small silencing RNAs: An expanding universe. Nat. Rev. Genet. 2009, 10, 94–108. [Google Scholar] [CrossRef] [Green Version]
  21. Bartel, D.P. MicroRNAs: Target recognition and regulatory functions. Cell 2009, 136, 215–233. [Google Scholar] [CrossRef] [Green Version]
  22. Cai, X.; Hagedorn, C.H.; Cullen, B.R. Human microRNAs are processed from capped, polyadenylated transcripts that can also function as mRNAs. RNA 2004, 10, 1957–1966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Bohnsack, M.T.; Czaplinski, K.; Gorlich, D. Exportin 5 is a RanGTP-dependent dsRNA-binding protein that mediates nuclear export of pre-miRNAs. RNA 2004, 10, 185–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Zhang, H.; Kolb, F.A.; Brondani, V.; Billy, E.; Filipowicz, W. Human Dicer preferentially cleaves dsRNAs at their termini without a requirement for ATP. EMBO J. 2002, 21, 5875–5885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ro, S.; Park, C.; Young, D.; Sanders, K.M.; Yan, W. Tissue-dependent paired expression of miRNAs. Nucleic Acids Res. 2007, 35, 5944–5953. [Google Scholar] [CrossRef] [PubMed]
  26. Esau, C.C.; Monia, B.P. Therapeutic potential for microRNAs. Adv. Drug Deliv. Rev. 2007, 59, 101–114. [Google Scholar] [CrossRef]
  27. Rajewsky, N.; Socci, N.D. Computational identification of microRNA targets. Dev. Biol. 2004, 267, 529–535. [Google Scholar] [CrossRef]
  28. Behm-Ansmant, I.; Rehwinkel, J.; Doerks, T.; Stark, A.; Bork, P.; Izaurralde, E. mRNA degradation by miRNAs and GW182 requires both CCR4:NOT deadenylase and DCP1:DCP2 decapping complexes. Genes Dev. 2006, 20, 1885–1898. [Google Scholar] [CrossRef] [Green Version]
  29. Jonas, S.; Izaurralde, E. Towards a molecular understanding of microRNA-mediated gene silencing. Nat. Rev. Genet. 2015, 16, 421–433. [Google Scholar] [CrossRef]
  30. Elbashir, S.M.; Harborth, J.; Lendeckel, W.; Yalcin, A.; Weber, K.; Tuschl, T. Duplexes of 21-nucleotide RNAs mediate RNA interference in cultured mammalian cells. Nature 2001, 411, 494–498. [Google Scholar] [CrossRef]
  31. Das, S.; Ansel, K.M.; Bitzer, M.; Breakefield, X.O.; Charest, A.; Galas, D.J.; Gerstein, M.B.; Gupta, M.; Milosavljevic, A.; McManus, M.T.; et al. The Extracellular RNA Communication Consortium: Establishing Foundational Knowledge and Technologies for Extracellular RNA Research. Cell 2019, 177, 231–242. [Google Scholar] [CrossRef] [Green Version]
  32. Takahashi, Y.; Yamaoka, K.; Nishikawa, M.; Takakura, Y. Quantitative and temporal analysis of gene silencing in tumor cells induced by small interfering RNA or short hairpin RNA expressed from plasmid vectors. J. Pharm. Sci. 2009, 98, 74–80. [Google Scholar] [CrossRef]
  33. Birmingham, A.; Anderson, E.M.; Reynolds, A.; Ilsley-Tyree, D.; Leake, D.; Fedorov, Y.; Baskerville, S.; Maksimova, E.; Robinson, K.; Karpilow, J.; et al. 3′ UTR seed matches, but not overall identity, are associated with RNAi off-targets. Nat. Methods 2006, 3, 199–204. [Google Scholar] [CrossRef]
  34. Jackson, A.L.; Bartz, S.R.; Schelter, J.; Kobayashi, S.V.; Burchard, J.; Mao, M.; Li, B.; Cavet, G.; Linsley, P.S. Expression profiling reveals off-target gene regulation by RNAi. Nat. Biotechnol. 2003, 21, 635–637. [Google Scholar] [CrossRef]
  35. Jackson, A.L.; Burchard, J.; Leake, D.; Reynolds, A.; Schelter, J.; Guo, J.; Johnson, J.M.; Lim, L.; Karpilow, J.; Nichols, K.; et al. Position-specific chemical modification of siRNAs reduces “off-target” transcript silencing. RNA 2006, 12, 1197–1205. [Google Scholar] [CrossRef] [Green Version]
  36. Lin, X.; Ruan, X.; Anderson, M.G.; McDowell, J.A.; Kroeger, P.E.; Fesik, S.W.; Shen, Y. siRNA-mediated off-target gene silencing triggered by a 7 nt complementation. Nucleic Acids Res. 2005, 33, 4527–4535. [Google Scholar] [CrossRef] [Green Version]
  37. Jackson, A.L.; Burchard, J.; Schelter, J.; Chau, B.N.; Cleary, M.; Lim, L.; Linsley, P.S. Widespread siRNA “off-target” transcript silencing mediated by seed region sequence complementarity. RNA 2006, 12, 1179–1187. [Google Scholar] [CrossRef] [Green Version]
  38. Seok, H.; Lee, H.; Jang, E.S.; Chi, S.W. Evaluation and control of miRNA-like off-target repression for RNA interference. Cell. Mol. Life Sci. 2018, 75, 797–814. [Google Scholar] [CrossRef]
  39. Amarzguioui, M.; Rossi, J.J. Principles of Dicer substrate (D-siRNA) design and function. Methods Mol. Biol. 2008, 442, 3–10. [Google Scholar] [CrossRef]
  40. Persengiev, S.P.; Zhu, X.; Green, M.R. Nonspecific, concentration-dependent stimulation and repression of mammalian gene expression by small interfering RNAs (siRNAs). RNA 2004, 10, 12–18. [Google Scholar] [CrossRef] [Green Version]
  41. Semizarov, D.; Frost, L.; Sarthy, A.; Kroeger, P.; Halbert, D.N.; Fesik, S.W. Specificity of short interfering RNA determined through gene expression signatures. Proc. Natl. Acad. Sci. USA 2003, 100, 6347–6352. [Google Scholar] [CrossRef] [Green Version]
  42. Kittler, R.; Heninger, A.K.; Franke, K.; Habermann, B.; Buchholz, F. Production of endoribonuclease-prepared short interfering RNAs for gene silencing in mammalian cells. Nat. Methods 2005, 2, 779–784. [Google Scholar] [CrossRef] [PubMed]
  43. Hannus, M.; Beitzinger, M.; Engelmann, J.C.; Weickert, M.T.; Spang, R.; Hannus, S.; Meister, G. siPools: Highly complex but accurately defined siRNA pools eliminate off-target effects. Nucleic Acids Res. 2014, 42, 8049–8061. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Karikó, K.; Bhuyan, P.; Capodici, J.; Weissman, D. Small interfering RNAs mediate sequence-independent gene suppression and induce immune activation by signaling through toll-like receptor 3. J. Immunol. 2004, 172, 6545–6549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Sledz, C.A.; Holko, M.; de Veer, M.J.; Silverman, R.H.; Williams, B.R. Activation of the interferon system by short-interfering RNAs. Nat. Cell Biol. 2003, 5, 834–839. [Google Scholar] [CrossRef] [PubMed]
  46. Sioud, M. Induction of inflammatory cytokines and interferon responses by double-stranded and single-stranded siRNAs is sequence-dependent and requires endosomal localization. J. Mol. Biol. 2005, 348, 1079–1090. [Google Scholar] [CrossRef] [PubMed]
  47. Sioud, M.; Sørensen, D.R. Cationic liposome-mediated delivery of siRNAs in adult mice. Biochem. Biophys. Res. Commun. 2003, 312, 1220–1225. [Google Scholar] [CrossRef]
  48. Judge, A.D.; Sood, V.; Shaw, J.R.; Fang, D.; McClintock, K.; MacLachlan, I. Sequence-dependent stimulation of the mammalian innate immune response by synthetic siRNA. Nat. Biotechnol. 2005, 23, 457–462. [Google Scholar] [CrossRef]
  49. Hornung, V.; Guenthner-Biller, M.; Bourquin, C.; Ablasser, A.; Schlee, M.; Uematsu, S.; Noronha, A.; Manoharan, M.; Akira, S.; de Fougerolles, A.; et al. Sequence-specific potent induction of IFN-alpha by short interfering RNA in plasmacytoid dendritic cells through TLR7. Nat. Med. 2005, 11, 263–270. [Google Scholar] [CrossRef]
  50. Chong, K.L.; Feng, L.; Schappert, K.; Meurs, E.; Donahue, T.F.; Friesen, J.D.; Hovanessian, A.G.; Williams, B.R. Human p68 kinase exhibits growth suppression in yeast and homology to the translational regulator GCN2. EMBO J. 1992, 11, 1553–1562. [Google Scholar] [CrossRef]
  51. Li, G.; Xiang, Y.; Sabapathy, K.; Silverman, R.H. An apoptotic signaling pathway in the interferon antiviral response mediated by RNase L and c-Jun NH2-terminal kinase. J. Biol. Chem. 2004, 279, 1123–1131. [Google Scholar] [CrossRef]
  52. Braasch, D.A.; Corey, D.R. Locked nucleic acid (LNA): Fine-tuning the recognition of DNA and RNA. Chem. Biol. 2001, 8, 1–7. [Google Scholar] [CrossRef] [Green Version]
  53. Khan, A.A.; Betel, D.; Miller, M.L.; Sander, C.; Leslie, C.S.; Marks, D.S. Transfection of small RNAs globally perturbs gene regulation by endogenous microRNAs. Nat. Biotechnol. 2009, 27, 549–555. [Google Scholar] [CrossRef]
  54. Holen, T.; Amarzguioui, M.; Wiiger, M.T.; Babaie, E.; Prydz, H. Positional effects of short interfering RNAs targeting the human coagulation trigger Tissue Factor. Nucleic Acids Res. 2002, 30, 1757–1766. [Google Scholar] [CrossRef] [Green Version]
  55. Reynolds, A.; Leake, D.; Boese, Q.; Scaringe, S.; Marshall, W.S.; Khvorova, A. Rational siRNA design for RNA interference. Nat. Biotechnol. 2004, 22, 326–330. [Google Scholar] [CrossRef]
  56. Knight, S.W.; Bass, B.L. A role for the RNase III enzyme DCR-1 in RNA interference and germ line development in Caenorhabditis elegans. Science 2001, 293, 2269–2271. [Google Scholar] [CrossRef] [Green Version]
  57. Dharap, A.; Pokrzywa, C.; Murali, S.; Pandi, G.; Vemuganti, R. MicroRNA miR-324-3p induces promoter-mediated expression of RelA gene. PLoS ONE 2013, 8, e79467. [Google Scholar] [CrossRef] [Green Version]
  58. Babiarz, J.E.; Ruby, J.G.; Wang, Y.; Bartel, D.P.; Blelloch, R. Mouse ES cells express endogenous shRNAs, siRNAs, and other Microprocessor-independent, Dicer-dependent small RNAs. Genes Dev. 2008, 22, 2773–2785. [Google Scholar] [CrossRef] [Green Version]
  59. Yang, J.S.; Lai, E.C. Dicer-independent, Ago2-mediated microRNA biogenesis in vertebrates. Cell Cycle 2010, 9, 4455–4460. [Google Scholar] [CrossRef] [Green Version]
  60. Xie, J.; He, C.; Su, Y.; Ding, Y.; Zhu, X.; Xu, Y.; Ding, J.; Zhou, H.; Wang, H. Research progress on microRNA in gout. Front. Pharmacol. 2022, 13, 981799. [Google Scholar] [CrossRef]
  61. Tanzer, A.; Stadler, P.F. Molecular evolution of a microRNA cluster. J. Mol. Biol. 2004, 339, 327–335. [Google Scholar] [CrossRef]
  62. Makarova, J.A.; Shkurnikov, M.U.; Wicklein, D.; Lange, T.; Samatov, T.R.; Turchinovich, A.A.; Tonevitsky, A.G. Intracellular and extracellular microRNA: An update on localization and biological role. Prog. Histochem. Cytochem. 2016, 51, 33–49. [Google Scholar] [CrossRef] [PubMed]
  63. Huang, W. MicroRNAs: Biomarkers, Diagnostics, and Therapeutics. Methods Mol. Biol. 2017, 1617, 57–67. [Google Scholar] [CrossRef] [PubMed]
  64. Matulić, M.; Gršković, P.; Petrović, A.; Begić, V.; Harabajsa, S.; Korać, P. miRNA in Molecular Diagnostics. Bioengineering 2022, 9, 459. [Google Scholar] [CrossRef] [PubMed]
  65. Quévillon Huberdeau, M.; Simard, M.J. A guide to microRNA-mediated gene silencing. FEBS J. 2019, 286, 642–652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Goldgraben, M.A.; Russell, R.; Rueda, O.M.; Caldas, C.; Git, A. Double-stranded microRNA mimics can induce length- and passenger strand-dependent effects in a cell type-specific manner. RNA 2016, 22, 193–203. [Google Scholar] [CrossRef] [Green Version]
  67. Marques, J.T.; Devosse, T.; Wang, D.; Zamanian-Daryoush, M.; Serbinowski, P.; Hartmann, R.; Fujita, T.; Behlke, M.A.; Williams, B.R. A structural basis for discriminating between self and nonself double-stranded RNAs in mammalian cells. Nat. Biotechnol. 2006, 24, 559–565. [Google Scholar] [CrossRef]
  68. Søkilde, R.; Newie, I.; Persson, H.; Borg, Å.; Rovira, C. Passenger strand loading in overexpression experiments using microRNA mimics. RNA Biol. 2015, 12, 787–791. [Google Scholar] [CrossRef] [Green Version]
  69. Jin, H.Y.; Gonzalez-Martin, A.; Miletic, A.V.; Lai, M.; Knight, S.; Sabouri-Ghomi, M.; Head, S.R.; Macauley, M.S.; Rickert, R.C.; Xiao, C. Transfection of microRNA Mimics Should Be Used with Caution. Front. Genet. 2015, 6, 340. [Google Scholar] [CrossRef] [Green Version]
  70. Ebert, M.S.; Sharp, P.A. MicroRNA sponges: Progress and possibilities. RNA 2010, 16, 2043–2050. [Google Scholar] [CrossRef] [Green Version]
  71. Barta, T.; Peskova, L.; Hampl, A. miRNAsong: A web-based tool for generation and testing of miRNA sponge constructs in silico. Sci. Rep. 2016, 6, 36625. [Google Scholar] [CrossRef]
  72. Zhang, W.; Zhang, C.; Hu, C.; Luo, C.; Zhong, B.; Yu, X. Circular RNA-CDR1as acts as the sponge of microRNA-641 to promote osteoarthritis progression. J. Inflamm. 2020, 17, 8. [Google Scholar] [CrossRef] [Green Version]
  73. He, A.T.; Liu, J.; Li, F.; Yang, B.B. Targeting circular RNAs as a therapeutic approach: Current strategies and challenges. Signal Transduct. Target. Ther. 2021, 6, 185. [Google Scholar] [CrossRef]
  74. McAnuff, M.A.; Rettig, G.R.; Rice, K.G. Potency of siRNA versus shRNA mediated knockdown in vivo. J. Pharm. Sci. 2007, 96, 2922–2930. [Google Scholar] [CrossRef]
  75. Paddison, P.J.; Caudy, A.A.; Bernstein, E.; Hannon, G.J.; Conklin, D.S. Short hairpin RNAs (shRNAs) induce sequence-specific silencing in mammalian cells. Genes Dev. 2002, 16, 948–958. [Google Scholar] [CrossRef] [Green Version]
  76. Zeng, Y.; Wagner, E.J.; Cullen, B.R. Both natural and designed micro RNAs can inhibit the expression of cognate mRNAs when expressed in human cells. Mol. Cell 2002, 9, 1327–1333. [Google Scholar] [CrossRef]
  77. Brummelkamp, T.R.; Bernards, R.; Agami, R. A system for stable expression of short interfering RNAs in mammalian cells. Science 2002, 296, 550–553. [Google Scholar] [CrossRef] [Green Version]
  78. Sui, G.; Soohoo, C.; Affar, E.B.; Gay, F.; Shi, Y.; Forrester, W.C. A DNA vector-based RNAi technology to suppress gene expression in mammalian cells. Proc. Natl. Acad. Sci. USA 2002, 99, 5515–5520. [Google Scholar] [CrossRef] [Green Version]
  79. Ma, H.; Wu, Y.; Dang, Y.; Choi, J.G.; Zhang, J.; Wu, H. Pol III Promoters to Express Small RNAs: Delineation of Transcription Initiation. Mol. Ther. Nucleic Acids 2014, 3, e161. [Google Scholar] [CrossRef]
  80. Lebbink, R.J.; Lowe, M.; Chan, T.; Khine, H.; Wang, X.; McManus, M.T. Polymerase II promoter strength determines efficacy of microRNA adapted shRNAs. PLoS ONE 2011, 6, e26213. [Google Scholar] [CrossRef]
  81. Bridge, A.J.; Pebernard, S.; Ducraux, A.; Nicoulaz, A.L.; Iggo, R. Induction of an interferon response by RNAi vectors in mammalian cells. Nat. Genet. 2003, 34, 263–264. [Google Scholar] [CrossRef]
  82. Pebernard, S.; Iggo, R.D. Determinants of interferon-stimulated gene induction by RNAi vectors. Differentiation 2004, 72, 103–111. [Google Scholar] [CrossRef]
  83. Grimm, D.; Streetz, K.L.; Jopling, C.L.; Storm, T.A.; Pandey, K.; Davis, C.R.; Marion, P.; Salazar, F.; Kay, M.A. Fatality in mice due to oversaturation of cellular microRNA/short hairpin RNA pathways. Nature 2006, 441, 537–541. [Google Scholar] [CrossRef] [PubMed]
  84. Silva, J.M.; Li, M.Z.; Chang, K.; Ge, W.; Golding, M.C.; Rickles, R.J.; Siolas, D.; Hu, G.; Paddison, P.J.; Schlabach, M.R.; et al. Second-generation shRNA libraries covering the mouse and human genomes. Nat. Genet. 2005, 37, 1281–1288. [Google Scholar] [CrossRef] [PubMed]
  85. Stegmeier, F.; Hu, G.; Rickles, R.J.; Hannon, G.J.; Elledge, S.J. A lentiviral microRNA-based system for single-copy polymerase II-regulated RNA interference in mammalian cells. Proc. Natl. Acad. Sci. USA 2005, 102, 13212–13217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Choi, J.G.; Bharaj, P.; Abraham, S.; Ma, H.; Yi, G.; Ye, C.; Dang, Y.; Manjunath, N.; Wu, H.; Shankar, P. Multiplexing seven miRNA-Based shRNAs to suppress HIV replication. Mol. Ther. 2015, 23, 310–320. [Google Scholar] [CrossRef] [Green Version]
  87. Truscott, M.; Islam, A.B.; Frolov, M.V. Novel regulation and functional interaction of polycistronic miRNAs. RNA 2016, 22, 129–138. [Google Scholar] [CrossRef] [Green Version]
  88. Fellmann, C.; Hoffmann, T.; Sridhar, V.; Hopfgartner, B.; Muhar, M.; Roth, M.; Lai, D.Y.; Barbosa, I.A.; Kwon, J.S.; Guan, Y.; et al. An optimized microRNA backbone for effective single-copy RNAi. Cell Rep. 2013, 5, 1704–1713. [Google Scholar] [CrossRef] [Green Version]
  89. Ma, Y.; Lin, H.; Qiu, C. High-efficiency transfection and siRNA-mediated gene knockdown in human pluripotent stem cells. Curr. Protoc. Stem Cell Biol. 2012, 21, 5C.2.1–5C.2.9. [Google Scholar] [CrossRef]
  90. Nguyen, T.A.; Jo, M.H.; Choi, Y.G.; Park, J.; Kwon, S.C.; Hohng, S.; Kim, V.N.; Woo, J.S. Functional Anatomy of the Human Microprocessor. Cell 2015, 161, 1374–1387. [Google Scholar] [CrossRef] [Green Version]
  91. Matranga, C.; Tomari, Y.; Shin, C.; Bartel, D.P.; Zamore, P.D. Passenger-strand cleavage facilitates assembly of siRNA into Ago2-containing RNAi enzyme complexes. Cell 2005, 123, 607–620. [Google Scholar] [CrossRef] [Green Version]
  92. Frank, F.; Sonenberg, N.; Nagar, B. Structural basis for 5′-nucleotide base-specific recognition of guide RNA by human AGO2. Nature 2010, 465, 818–822. [Google Scholar] [CrossRef]
  93. Gu, S.; Kay, M.A. How do miRNAs mediate translational repression? Silence 2010, 1, 11. [Google Scholar] [CrossRef] [Green Version]
  94. De, N.; Young, L.; Lau, P.W.; Meisner, N.C.; Morrissey, D.V.; MacRae, I.J. Highly complementary target RNAs promote release of guide RNAs from human Argonaute2. Mol. Cell 2013, 50, 344–355. [Google Scholar] [CrossRef] [Green Version]
  95. Conrad, T.; Marsico, A.; Gehre, M.; Ørom, U.A. Microprocessor Activity Controls Differential miRNA Biogenesis In Vivo. Cell Rep. 2015, 10, 1020. [Google Scholar] [CrossRef] [Green Version]
  96. Burchard, J.; Jackson, A.L.; Malkov, V.; Needham, R.H.; Tan, Y.; Bartz, S.R.; Dai, H.; Sachs, A.B.; Linsley, P.S. MicroRNA-like off-target transcript regulation by siRNAs is species specific. RNA 2009, 15, 308–315. [Google Scholar] [CrossRef] [Green Version]
  97. Yi, R.; Doehle, B.P.; Qin, Y.; Macara, I.G.; Cullen, B.R. Overexpression of exportin 5 enhances RNA interference mediated by short hairpin RNAs and microRNAs. RNA 2005, 11, 220–226. [Google Scholar] [CrossRef] [Green Version]
  98. Grimm, D.; Wang, L.; Lee, J.S.; Schürmann, N.; Gu, S.; Börner, K.; Storm, T.A.; Kay, M.A. Argonaute proteins are key determinants of RNAi efficacy, toxicity, and persistence in the adult mouse liver. J. Clin. Investig. 2010, 120, 3106–3119. [Google Scholar] [CrossRef] [Green Version]
  99. Mohr, S.E.; Smith, J.A.; Shamu, C.E.; Neumüller, R.A.; Perrimon, N. RNAi screening comes of age: Improved techniques and complementary approaches. Nat. Rev. Mol. Cell Biol. 2014, 15, 591–600. [Google Scholar] [CrossRef] [Green Version]
  100. Ozata, D.M.; Gainetdinov, I.; Zoch, A.; O’Carroll, D.; Zamore, P.D. PIWI-interacting RNAs: Small RNAs with big functions. Nat. Rev. Genet. 2019, 20, 89–108. [Google Scholar] [CrossRef] [Green Version]
  101. Aravin, A.; Gaidatzis, D.; Pfeffer, S.; Lagos-Quintana, M.; Landgraf, P.; Iovino, N.; Morris, P.; Brownstein, M.J.; Kuramochi-Miyagawa, S.; Nakano, T.; et al. A novel class of small RNAs bind to MILI protein in mouse testes. Nature 2006, 442, 203–207. [Google Scholar] [CrossRef]
  102. Girard, A.; Sachidanandam, R.; Hannon, G.J.; Carmell, M.A. A germline-specific class of small RNAs binds mammalian Piwi proteins. Nature 2006, 442, 199–202. [Google Scholar] [CrossRef]
  103. Huang, X.; Fejes Tóth, K.; Aravin, A.A. piRNA Biogenesis in Drosophila melanogaster. Trends Genet. 2017, 33, 882–894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Siddiqi, S.; Matushansky, I. Piwis and piwi-interacting RNAs in the epigenetics of cancer. J. Cell. Biochem. 2012, 113, 373–380. [Google Scholar] [CrossRef] [PubMed]
  105. Sasaki, T.; Shiohama, A.; Minoshima, S.; Shimizu, N. Identification of eight members of the Argonaute family in the human genome. Genomics 2003, 82, 323–330. [Google Scholar] [CrossRef] [PubMed]
  106. Carmell, M.A.; Girard, A.; van de Kant, H.J.; Bourc’his, D.; Bestor, T.H.; de Rooij, D.G.; Hannon, G.J. MIWI2 is essential for spermatogenesis and repression of transposons in the mouse male germline. Dev. Cell 2007, 12, 503–514. [Google Scholar] [CrossRef] [Green Version]
  107. Klenov, M.S.; Lavrov, S.A.; Korbut, A.P.; Stolyarenko, A.D.; Yakushev, E.Y.; Reuter, M.; Pillai, R.S.; Gvozdev, V.A. Impact of nuclear Piwi elimination on chromatin state in Drosophila melanogaster ovaries. Nucleic Acids Res. 2014, 42, 6208–6218. [Google Scholar] [CrossRef] [Green Version]
  108. Post, C.; Clark, J.P.; Sytnikova, Y.A.; Chirn, G.W.; Lau, N.C. The capacity of target silencing by Drosophila PIWI and piRNAs. RNA 2014, 20, 1977–1986. [Google Scholar] [CrossRef] [Green Version]
  109. Goh, W.S.; Falciatori, I.; Tam, O.H.; Burgess, R.; Meikar, O.; Kotaja, N.; Hammell, M.; Hannon, G.J. piRNA-directed cleavage of meiotic transcripts regulates spermatogenesis. Genes Dev. 2015, 29, 1032–1044. [Google Scholar] [CrossRef] [Green Version]
  110. Gou, L.T.; Dai, P.; Yang, J.H.; Xue, Y.; Hu, Y.P.; Zhou, Y.; Kang, J.Y.; Wang, X.; Li, H.; Hua, M.M.; et al. Pachytene piRNAs instruct massive mRNA elimination during late spermiogenesis. Cell Res. 2015, 25, 266. [Google Scholar] [CrossRef] [Green Version]
  111. Mai, D.; Ding, P.; Tan, L.; Zhang, J.; Pan, Z.; Bai, R.; Li, C.; Li, M.; Zhou, Y.; Tan, W.; et al. PIWI-interacting RNA-54265 is oncogenic and a potential therapeutic target in colorectal adenocarcinoma. Theranostics 2018, 8, 5213–5230. [Google Scholar] [CrossRef]
  112. Yin, J.; Jiang, X.Y.; Qi, W.; Ji, C.G.; Xie, X.L.; Zhang, D.X.; Cui, Z.J.; Wang, C.K.; Bai, Y.; Wang, J.; et al. piR-823 contributes to colorectal tumorigenesis by enhancing the transcriptional activity of HSF1. Cancer Sci. 2017, 108, 1746–1756. [Google Scholar] [CrossRef] [Green Version]
  113. Jia, D.D.; Jiang, H.; Zhang, Y.F.; Zhang, Y.; Qian, L.L. The regulatory function of piRNA/PIWI complex in cancer and other human diseases: The role of DNA methylation. Int. J. Biol. Sci. 2022, 18, 3358–3373. [Google Scholar] [CrossRef]
  114. Barresi, V.; Musmeci, C.; Rinaldi, A.; Condorelli, D.F. Transcript-Targeted Therapy Based on RNA Interference and Antisense Oligonucleotides: Current Applications and Novel Molecular Targets. Int. J. Mol. Sci. 2022, 23, 8875. [Google Scholar] [CrossRef]
  115. Vickers, T.A.; Crooke, S.T. The rates of the major steps in the molecular mechanism of RNase H1-dependent antisense oligonucleotide induced degradation of RNA. Nucleic Acids Res. 2015, 43, 8955–8963. [Google Scholar] [CrossRef] [Green Version]
  116. Sazani, P.; Graziewicz, M.A.; Kole, R. Splice switching oligonucleotides as potential therapeutics. In Antisense Drug Technology—Principles, Strategies, and Applications; Stanley, C., Ed.; CRC Press: Boca Raton, FL, USA, 2008; pp. 89–114. [Google Scholar]
  117. Field, A.K.; Young, C.W.; Krakoff, I.H.; Tytell, A.A.; Lampson, G.P.; Nemes, M.M.; Hilleman, M.R. Induction of interferon in human subjects by poly I:C. Proc. Soc. Exp. Biol. Med. 1971, 136, 1180–1186. [Google Scholar] [CrossRef]
  118. Crooke, S.T.; Witztum, J.L.; Bennett, C.F.; Baker, B.F. RNA-Targeted Therapeutics. Cell Metab. 2019, 29, 501. [Google Scholar] [CrossRef] [Green Version]
  119. Swayze, E.; Bhat, B. The medicinal chemistry of oligonucleotides. In Antisense Drug Technology—Principles, Strategies, and Applications; Crooke, S., Ed.; CRC Press: Boca Raton, FL, USA, 2008; pp. 143–182. [Google Scholar]
  120. Crooke, S.T.; Vickers, T.A.; Liang, X.H. Phosphorothioate modified oligonucleotide-protein interactions. Nucleic Acids Res. 2020, 48, 5235–5253. [Google Scholar] [CrossRef]
  121. Chernikov, I.V.; Vlassov, V.V.; Chernolovskaya, E.L. Current Development of siRNA Bioconjugates: From Research to the Clinic. Front. Pharmacol. 2019, 10, 444. [Google Scholar] [CrossRef] [Green Version]
  122. Crooke, S.T.; Wang, S.; Vickers, T.A.; Shen, W.; Liang, X.H. Cellular uptake and trafficking of antisense oligonucleotides. Nat. Biotechnol. 2017, 35, 230–237. [Google Scholar] [CrossRef]
  123. Crooke, S.T.; Baker, B.F.; Crooke, R.M.; Liang, X.H. Antisense technology: An overview and prospectus. Nat. Rev. Drug Discov. 2021, 20, 427–453. [Google Scholar] [CrossRef]
  124. Steinke, C.A.; Reeves, K.K.; Powell, J.W.; Lee, S.A.; Chen, Y.Z.; Wyrzykiewicz, T.; Griffey, R.H.; Mohan, V. Vibrational analysis of phosphorothioate DNA: II. The POS group in the model compound dimethyl phosphorothioate [(CH3O)2(POS)]. J. Biomol. Struct. Dyn. 1997, 14, 509–516. [Google Scholar] [CrossRef] [PubMed]
  125. Shen, W.; De Hoyos, C.L.; Migawa, M.T.; Vickers, T.A.; Sun, H.; Low, A.; Bell, T.A.; Rahdar, M.; Mukhopadhyay, S.; Hart, C.E.; et al. Chemical modification of PS-ASO therapeutics reduces cellular protein-binding and improves the therapeutic index. Nat. Biotechnol. 2019, 37, 640–650. [Google Scholar] [CrossRef] [PubMed]
  126. Ämmälä, C.; Drury, W.J.; Knerr, L.; Ahlstedt, I.; Stillemark-Billton, P.; Wennberg-Huldt, C.; Andersson, E.M.; Valeur, E.; Jansson-Löfmark, R.; Janzén, D.; et al. Targeted delivery of antisense oligonucleotides to pancreatic β-cells. Sci. Adv. 2018, 4, eaat3386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Sandoval-Mojica, A.F.; Hunter, W.B.; Aishwarya, V.; Bonilla, S.; Pelz-Stelinski, K.S. Antibacterial FANA oligonucleotides as a novel approach for managing the Huanglongbing pathosystem. Sci. Rep. 2021, 11, 2760. [Google Scholar] [CrossRef] [PubMed]
  128. Takahashi, M.; Li, H.; Zhou, J.; Chomchan, P.; Aishwarya, V.; Damha, M.J.; Rossi, J.J. Dual Mechanisms of Action of Self-Delivering, Anti-HIV-1 FANA Oligonucleotides as a Potential New Approach to HIV Therapy. Mol. Ther. Nucleic Acids 2019, 17, 615–625. [Google Scholar] [CrossRef] [Green Version]
  129. Margiotta, J.F.; Howard, M.J. Cryptochromes Mediate Intrinsic Photomechanical Transduction in Avian Iris and Somatic Striated Muscle. Front. Physiol. 2020, 11, 128. [Google Scholar] [CrossRef]
  130. Pelisch, N.; Rosas Almanza, J.; Stehlik, K.E.; Aperi, B.V.; Kroner, A. Use of a Self-Delivering Anti-CCL3 FANA Oligonucleotide as an Innovative Approach to Target Inflammation after Spinal Cord Injury. eNeuro 2021, 8, ENEURO.0338-20.2021. [Google Scholar] [CrossRef]
  131. Shytaj, I.L.; Lucic, B.; Forcato, M.; Penzo, C.; Billingsley, J.; Laketa, V.; Bosinger, S.; Stanic, M.; Gregoretti, F.; Antonelli, L.; et al. Alterations of redox and iron metabolism accompany the development of HIV latency. EMBO J. 2020, 39, e102209. [Google Scholar] [CrossRef]
  132. Quemener, A.M.; Bachelot, L.; Forestier, A.; Donnou-Fournet, E.; Gilot, D.; Galibert, M.D. The powerful world of antisense oligonucleotides: From bench to bedside. Wiley Interdiscip. Rev. RNA 2020, 11, e1594. [Google Scholar] [CrossRef]
  133. Quemener, A.M.; Galibert, M.D. Antisense oligonucleotide: A promising therapeutic option to beat COVID-19. Wiley Interdiscip. Rev. RNA 2021, 13, e1703. [Google Scholar] [CrossRef]
  134. Lazebnik, M.; Pack, D.W. Rapid and facile quantitation of polyplex endocytic trafficking. J. Control. Release 2017, 247, 19–27. [Google Scholar] [CrossRef] [Green Version]
  135. Khalil, I.A.; Kogure, K.; Akita, H.; Harashima, H. Uptake pathways and subsequent intracellular trafficking in nonviral gene delivery. Pharmacol. Rev. 2006, 58, 32–45. [Google Scholar] [CrossRef] [Green Version]
  136. Rädler, J.O.; Koltover, I.; Salditt, T.; Safinya, C.R. Structure of DNA-cationic liposome complexes: DNA intercalation in multilamellar membranes in distinct interhelical packing regimes. Science 1997, 275, 810–814. [Google Scholar] [CrossRef] [Green Version]
  137. Niculescu-Duvaz, D.; Heyes, J.; Springer, C.J. Structure-activity relationship in cationic lipid mediated gene transfection. Curr. Med. Chem. 2003, 10, 1233–1261. [Google Scholar] [CrossRef]
  138. Srinivas, R.; Samanta, S.; Chaudhuri, A. Cationic amphiphiles: Promising carriers of genetic materials in gene therapy. Chem. Soc. Rev. 2009, 38, 3326–3338. [Google Scholar] [CrossRef]
  139. Byk, G.; Wetzer, B.; Frederic, M.; Dubertret, C.; Pitard, B.; Jaslin, G.; Scherman, D. Reduction-sensitive lipopolyamines as a novel nonviral gene delivery system for modulated release of DNA with improved transgene expression. J. Med. Chem. 2000, 43, 4377–4387. [Google Scholar] [CrossRef]
  140. Sato, Y.; Matsui, H.; Sato, R.; Harashima, H. Neutralization of negative charges of siRNA results in improved safety and efficient gene silencing activity of lipid nanoparticles loaded with high levels of siRNA. J. Control. Release 2018, 284, 179–187. [Google Scholar] [CrossRef] [Green Version]
  141. Simberg, D.; Weisman, S.; Talmon, Y.; Barenholz, Y. DOTAP (and other cationic lipids): Chemistry, biophysics, and transfection. Crit. Rev. Ther. Drug Carrier Syst. 2004, 21, 257–317. [Google Scholar] [CrossRef]
  142. Koltover, I.; Salditt, T.; Rädler, J.O.; Safinya, C.R. An inverted hexagonal phase of cationic liposome-DNA complexes related to DNA release and delivery. Science 1998, 281, 78–81. [Google Scholar] [CrossRef] [Green Version]
  143. Cheng, X.; Lee, R.J. The role of helper lipids in lipid nanoparticles (LNPs) designed for oligonucleotide delivery. Adv. Drug Deliv. Rev. 2016, 99, 129–137. [Google Scholar] [CrossRef]
  144. Zuhorn, I.S.; Kalicharan, R.; Hoekstra, D. Lipoplex-mediated transfection of mammalian cells occurs through the cholesterol-dependent clathrin-mediated pathway of endocytosis. J. Biol. Chem. 2002, 277, 18021–18028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Hoffmann, M.; Hersch, N.; Merkel, R.; Csiszar, A.; Hoffmann, B. Changing the Way of Entrance: Highly Efficient Transfer of mRNA and siRNA via Fusogenic Nano-Carriers. J. Biomed. Nanotechnol. 2019, 15, 170–183. [Google Scholar] [CrossRef] [PubMed]
  146. Xia, T.; Kovochich, M.; Liong, M.; Zink, J.I.; Nel, A.E. Cationic polystyrene nanosphere toxicity depends on cell-specific endocytic and mitochondrial injury pathways. ACS Nano 2008, 2, 85–96. [Google Scholar] [CrossRef] [PubMed]
  147. Cui, S.; Wang, Y.; Gong, Y.; Lin, X.; Zhao, Y.; Zhi, D.; Zhou, Q.; Zhang, S. Correlation of the cytotoxic effects of cationic lipids with their headgroups. Toxicol. Res. 2018, 7, 473–479. [Google Scholar] [CrossRef] [PubMed]
  148. Jain, S.; Kumar, S.; Agrawal, A.K.; Thanki, K.; Banerjee, U.C. Enhanced transfection efficiency and reduced cytotoxicity of novel lipid-polymer hybrid nanoplexes. Mol. Pharm. 2013, 10, 2416–2425. [Google Scholar] [CrossRef]
  149. Romøren, K.; Fjeld, X.T.; Poléo, A.B.; Smistad, G.; Thu, B.J.; Evensen, Ø. Transfection efficiency and cytotoxicity of cationic liposomes in primary cultures of rainbow trout (Oncorhynchus mykiss) gill cells. Biochim. Biophys. Acta 2005, 1717, 50–57. [Google Scholar] [CrossRef] [Green Version]
  150. Brencicova, E.; Diebold, S.S. Nucleic acids and endosomal pattern recognition: How to tell friend from foe? Front. Cell. Infect. Microbiol. 2013, 3, 37. [Google Scholar] [CrossRef] [Green Version]
  151. Kawai, T.; Akira, S. The roles of TLRs, RLRs and NLRs in pathogen recognition. Int. Immunol. 2009, 21, 317–337. [Google Scholar] [CrossRef] [Green Version]
  152. Leonhardt, C.; Schwake, G.; Stögbauer, T.R.; Rappl, S.; Kuhr, J.T.; Ligon, T.S.; Rädler, J.O. Single-cell mRNA transfection studies: Delivery, kinetics and statistics by numbers. Nanomedicine 2014, 10, 679–688. [Google Scholar] [CrossRef] [Green Version]
  153. Selby, L.I.; Cortez-Jugo, C.M.; Such, G.K.; Johnston, A.P.R. Nanoescapology: Progress toward understanding the endosomal escape of polymeric nanoparticles. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2017, 9, e1452. [Google Scholar] [CrossRef]
  154. Varkouhi, A.K.; Scholte, M.; Storm, G.; Haisma, H.J. Endosomal escape pathways for delivery of biologicals. J. Control. Release 2011, 151, 220–228. [Google Scholar] [CrossRef]
  155. Lönn, P.; Kacsinta, A.D.; Cui, X.S.; Hamil, A.S.; Kaulich, M.; Gogoi, K.; Dowdy, S.F. Enhancing Endosomal Escape for Intracellular Delivery of Macromolecular Biologic Therapeutics. Sci. Rep. 2016, 6, 32301. [Google Scholar] [CrossRef]
  156. Nakamura, T.; Yamada, K.; Fujiwara, Y.; Sato, Y.; Harashima, H. Reducing the Cytotoxicity of Lipid Nanoparticles Associated with a Fusogenic Cationic Lipid in a Natural Killer Cell Line by Introducing a Polycation-Based siRNA Core. Mol. Pharm. 2018, 15, 2142–2150. [Google Scholar] [CrossRef]
  157. Wong-Baeza, C.; Bustos, I.; Serna, M.; Tescucano, A.; Alcántara-Farfán, V.; Ibáñez, M.; Montañez, C.; Wong, C.; Baeza, I. Membrane fusion inducers, chloroquine and spermidine increase lipoplex-mediated gene transfection. Biochem. Biophys. Res. Commun. 2010, 396, 549–554. [Google Scholar] [CrossRef]
  158. Kadonosono, T.; Yamano, A.; Goto, T.; Tsubaki, T.; Niibori, M.; Kuchimaru, T.; Kizaka-Kondoh, S. Cell penetrating peptides improve tumor delivery of cargos through neuropilin-1-dependent extravasation. J. Control. Release 2015, 201, 14–21. [Google Scholar] [CrossRef]
  159. Csiszár, A.; Hersch, N.; Dieluweit, S.; Biehl, R.; Merkel, R.; Hoffmann, B. Novel fusogenic liposomes for fluorescent cell labeling and membrane modification. Bioconjug. Chem. 2010, 21, 537–543. [Google Scholar] [CrossRef]
  160. Kleusch, C.; Hersch, N.; Hoffmann, B.; Merkel, R.; Csiszár, A. Fluorescent lipids: Functional parts of fusogenic liposomes and tools for cell membrane labeling and visualization. Molecules 2012, 17, 1055–1073. [Google Scholar] [CrossRef] [Green Version]
  161. Braun, T.; Kleusch, C.; Naumovska, E.; Merkel, R.; Csiszár, A. A bioanalytical assay to distinguish cellular uptake routes for liposomes. Cytom. A 2016, 89, 301–308. [Google Scholar] [CrossRef]
  162. Kolašinac, R.; Kleusch, C.; Braun, T.; Merkel, R.; Csiszár, A. Deciphering the Functional Composition of Fusogenic Liposomes. Int. J. Mol. Sci. 2018, 19, 346. [Google Scholar] [CrossRef] [Green Version]
  163. Kube, S.; Hersch, N.; Naumovska, E.; Gensch, T.; Hendriks, J.; Franzen, A.; Landvogt, L.; Siebrasse, J.P.; Kubitscheck, U.; Hoffmann, B.; et al. Fusogenic Liposomes as Nanocarriers for the Delivery of Intracellular Proteins. Langmuir 2017, 33, 1051–1059. [Google Scholar] [CrossRef]
  164. Hoffmann, M.; Hersch, N.; Gerlach, S.; Dreissen, G.; Springer, R.; Merkel, R.; Csiszár, A.; Hoffmann, B. Complex Size and Surface Charge Determine Nucleic Acid Transfer by Fusogenic Liposomes. Int. J. Mol. Sci. 2020, 21, 2244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Hou, X.; Zaks, T.; Langer, R.; Dong, Y. Lipid nanoparticles for mRNA delivery. Nat. Rev. Mater. 2021, 6, 1078–1094. [Google Scholar] [CrossRef] [PubMed]
  166. Eygeris, Y.; Gupta, M.; Kim, J.; Sahay, G. Chemistry of Lipid Nanoparticles for RNA Delivery. Acc. Chem. Res. 2022, 55, 2–12. [Google Scholar] [CrossRef] [PubMed]
  167. Saw, P.E.; Yao, H.; Lin, C.; Tao, W.; Farokhzad, O.C.; Xu, X. Stimuli-Responsive Polymer-Prodrug Hybrid Nanoplatform for Multistage siRNA Delivery and Combination Cancer Therapy. Nano Lett. 2019, 19, 5967–5974. [Google Scholar] [CrossRef] [PubMed]
  168. Bholakant, R.; Qian, H.; Zhang, J.; Huang, X.; Huang, D.; Feijen, J.; Zhong, Y.; Chen, W. Recent Advances of Polycationic siRNA Vectors for Cancer Therapy. Biomacromolecules 2020, 21, 2966–2982. [Google Scholar] [CrossRef] [PubMed]
  169. Pan, J.; Mendes, L.P.; Yao, M.; Filipczak, N.; Garai, S.; Thakur, G.A.; Sarisozen, C.; Torchilin, V.P. Polyamidoamine dendrimers-based nanomedicine for combination therapy with siRNA and chemotherapeutics to overcome multidrug resistance. Eur. J. Pharm. Biopharm. 2019, 136, 18–28. [Google Scholar] [CrossRef]
  170. Su, D.; Coste, M.; Diaconu, A.; Barboiu, M.; Ulrich, S. Cationic dynamic covalent polymers for gene transfection. J. Mater. Chem. B 2020, 8, 9385–9403. [Google Scholar] [CrossRef]
  171. Lee, J.; Sands, I.; Zhang, W.; Zhou, L.; Chen, Y. DNA-inspired nanomaterials for enhanced endosomal escape. Proc. Natl. Acad. Sci. USA 2021, 118, e2104511118. [Google Scholar] [CrossRef]
  172. Glaser, R.W.; Leikin, S.L.; Chernomordik, L.V.; Pastushenko, V.F.; Sokirko, A.I. Reversible electrical breakdown of lipid bilayers: Formation and evolution of pores. Biochim. Biophys. Acta 1988, 940, 275–287. [Google Scholar] [CrossRef]
  173. Rubinsky, B. Irreversible electroporation in medicine. Technol. Cancer Res. Treat. 2007, 6, 255–260. [Google Scholar] [CrossRef]
  174. Schwan, H.P. Electrical properties of tissue and cell suspensions. Adv. Biol. Med. Phys. 1957, 5, 147–209. [Google Scholar] [CrossRef]
  175. Ivorra, A.; Villemejane, J.; Mir, L.M. Electrical modeling of the influence of medium conductivity on electroporation. Phys. Chem. Chem. Phys. 2010, 12, 10055–10064. [Google Scholar] [CrossRef] [Green Version]
  176. Pucihar, G.; Mir, L.M.; Miklavcic, D. The effect of pulse repetition frequency on the uptake into electropermeabilized cells in vitro with possible applications in electrochemotherapy. Bioelectrochemistry 2002, 57, 167–172. [Google Scholar] [CrossRef]
  177. Shankayi, Z.; Firoozabadi, S.M. Antitumor efficiency of electrochemotherapy by high and low frequencies and repetitive therapy in the treatment of invasive ductal carcinoma in BALB/c mice. Cell J. 2012, 14, 110–115. [Google Scholar]
  178. Miklavcic, D.; Pucihar, G.; Pavlovec, M.; Ribaric, S.; Mali, M.; Macek-Lebar, A.; Petkovsek, M.; Nastran, J.; Kranjc, S.; Cemazar, M.; et al. The effect of high frequency electric pulses on muscle contractions and antitumor efficiency in vivo for a potential use in clinical electrochemotherapy. Bioelectrochemistry 2005, 65, 121–128. [Google Scholar] [CrossRef]
  179. Vernier, P.T.; Sun, Y.; Gundersen, M.A. Nanoelectropulse-driven membrane perturbation and small molecule permeabilization. BMC Cell Biol. 2006, 7, 37. [Google Scholar] [CrossRef] [Green Version]
  180. Murauskas, A.; Staigvila, G.; Girkontaitė, I.; Zinkevičienė, A.; Ruzgys, P.; Šatkauskas, S.; Novickij, J.; Novickij, V. Predicting electrotransfer in ultra-high frequency sub-microsecond square wave electric fields. Electromagn. Biol. Med. 2020, 39, 1–8. [Google Scholar] [CrossRef]
  181. Novickij, V.; Ruzgys, P.; Grainys, A.; Šatkauskas, S. High frequency electroporation efficiency is under control of membrane capacitive charging and voltage potential relaxation. Bioelectrochemistry 2018, 119, 92–97. [Google Scholar] [CrossRef]
  182. Dermol, J.; Pakhomova, O.N.; Pakhomov, A.G.; Miklavčič, D. Cell Electrosensitization Exists Only in Certain Electroporation Buffers. PLoS ONE 2016, 11, e0159434. [Google Scholar] [CrossRef] [Green Version]
  183. Pakhomova, O.N.; Gregory, B.W.; Khorokhorina, V.A.; Bowman, A.M.; Xiao, S.; Pakhomov, A.G. Electroporation-induced electrosensitization. PLoS ONE 2011, 6, e17100. [Google Scholar] [CrossRef] [Green Version]
  184. Potočnik, T.; Maček Lebar, A.; Kos, Š.; Reberšek, M.; Pirc, E.; Serša, G.; Miklavčič, D. Effect of Experimental Electrical and Biological Parameters on Gene Transfer by Electroporation: A Systematic Review and Meta-Analysis. Pharmaceutics 2022, 14, 2700. [Google Scholar] [CrossRef] [PubMed]
  185. Garcia, P.A.; Ge, Z.; Kelley, L.E.; Holcomb, S.J.; Buie, C.R. High efficiency hydrodynamic bacterial electrotransformation. Lab Chip 2017, 17, 490–500. [Google Scholar] [CrossRef] [PubMed]
  186. Choi, S.E.; Khoo, H.; Hur, S.C. Recent Advances in Microscale Electroporation. Chem. Rev. 2022, 122, 11247–11286. [Google Scholar] [CrossRef] [PubMed]
  187. Geng, T.; Zhan, Y.; Wang, J.; Lu, C. Transfection of cells using flow-through electroporation based on constant voltage. Nat. Protoc. 2011, 6, 1192–1208. [Google Scholar] [CrossRef]
  188. Gach, P.C.; Iwai, K.; Kim, P.W.; Hillson, N.J.; Singh, A.K. Droplet microfluidics for synthetic biology. Lab Chip 2017, 17, 3388–3400. [Google Scholar] [CrossRef] [Green Version]
  189. Boukany, P.E.; Morss, A.; Liao, W.C.; Henslee, B.; Jung, H.; Zhang, X.; Yu, B.; Wang, X.; Wu, Y.; Li, L.; et al. Nanochannel electroporation delivers precise amounts of biomolecules into living cells. Nat. Nanotechnol. 2011, 6, 747–754. [Google Scholar] [CrossRef]
  190. Chen, Z.; Cao, Y.; Lin, C.W.; Alvarez, S.; Oh, D.; Yang, P.; Groves, J.T. Nanopore-mediated protein delivery enabling three-color single-molecule tracking in living cells. Proc. Natl. Acad. Sci. USA 2021, 118, e201222911. [Google Scholar] [CrossRef]
  191. Chang, L.; Bertani, P.; Gallego-Perez, D.; Yang, Z.; Chen, F.; Chiang, C.; Malkoc, V.; Kuang, T.; Gao, K.; Lee, L.J.; et al. 3D nanochannel electroporation for high-throughput cell transfection with high uniformity and dosage control. Nanoscale 2016, 8, 243–252. [Google Scholar] [CrossRef]
  192. Nathamgari, S.S.P.; Pathak, N.; Lemaitre, V.; Mukherjee, P.; Muldoon, J.J.; Peng, C.Y.; McGuire, T.; Leonard, J.N.; Kessler, J.A.; Espinosa, H.D. Nanofountain Probe Electroporation Enables Versatile Single-Cell Intracellular Delivery and Investigation of Postpulse Electropore Dynamics. Small 2020, 16, e2002616. [Google Scholar] [CrossRef]
  193. Roth, T.L.; Puig-Saus, C.; Yu, R.; Shifrut, E.; Carnevale, J.; Li, P.J.; Hiatt, J.; Saco, J.; Krystofinski, P.; Li, H.; et al. Reprogramming human T cell function and specificity with non-viral genome targeting. Nature 2018, 559, 405–409. [Google Scholar] [CrossRef]
  194. Kim, J.Y.; Choi, J.H.; Kim, S.H.; Park, H.; Lee, D.; Kim, G.J. Efficacy of Gene Modification in Placenta-Derived Mesenchymal Stem Cells Based on Nonviral Electroporation. Int. J. Stem Cells 2021, 14, 112–118. [Google Scholar] [CrossRef]
  195. Keller, A.A.; Maeß, M.B.; Schnoor, M.; Scheiding, B.; Lorkowski, S. Transfecting Macrophages. Methods Mol. Biol. 2018, 1784, 187–195. [Google Scholar] [CrossRef]
  196. Scherer, O.; Maeß, M.B.; Lindner, S.; Garscha, U.; Weinigel, C.; Rummler, S.; Werz, O.; Lorkowski, S. A procedure for efficient non-viral siRNA transfection of primary human monocytes using nucleofection. J. Immunol. Methods 2015, 422, 118–124. [Google Scholar] [CrossRef]
  197. Takayama, K.; Igai, K.; Hagihara, Y.; Hashimoto, R.; Hanawa, M.; Sakuma, T.; Tachibana, M.; Sakurai, F.; Yamamoto, T.; Mizuguchi, H. Highly efficient biallelic genome editing of human ES/iPS cells using a CRISPR/Cas9 or TALEN system. Nucleic Acids Res. 2017, 45, 5198–5207. [Google Scholar] [CrossRef]
  198. Marchenko, S.; Flanagan, L. Transfecting human neural stem cells with the Amaxa Nucleofector. J. Vis. Exp. 2007, 6, 240. [Google Scholar] [CrossRef] [Green Version]
  199. Bak, R.O.; Dever, D.P.; Porteus, M.H. CRISPR/Cas9 genome editing in human hematopoietic stem cells. Nat. Protoc. 2018, 13, 358–376. [Google Scholar] [CrossRef] [Green Version]
  200. Novickij, V.; Rembiałkowska, N.; Szlasa, W.; Kulbacka, J. Does the shape of the electric pulse matter in electroporation? Front. Oncol. 2022, 12, 958128. [Google Scholar] [CrossRef]
  201. Takahashi, M.; Furukawa, T.; Saitoh, H.; Aoki, A.; Koike, T.; Moriyama, Y.; Shibata, A. Gene transfer into human leukemia cell lines by electroporation: Experience with exponentially decaying and square wave pulse. Leuk. Res. 1991, 15, 507–513. [Google Scholar] [CrossRef]
  202. Miller, D.L.; Bao, S.; Morris, J.E. Sonoporation of cultured cells in the rotating tube exposure system. Ultrasound Med. Biol. 1999, 25, 143–149. [Google Scholar] [CrossRef]
  203. Liu, Y.; Yan, J.; Prausnitz, M.R. Can ultrasound enable efficient intracellular uptake of molecules? A retrospective literature review and analysis. Ultrasound Med. Biol. 2012, 38, 876–888. [Google Scholar] [CrossRef] [Green Version]
  204. Ohl, C.D.; Arora, M.; Ikink, R.; de Jong, N.; Versluis, M.; Delius, M.; Lohse, D. Sonoporation from jetting cavitation bubbles. Biophys. J. 2006, 91, 4285–4295. [Google Scholar] [CrossRef] [Green Version]
  205. Zupanc, M.; Pandur, Ž.; Stepišnik Perdih, T.; Stopar, D.; Petkovšek, M.; Dular, M. Effects of cavitation on different microorganisms: The current understanding of the mechanisms taking place behind the phenomenon. A review and proposals for further research. Ultrason. Sonochem. 2019, 57, 147–165. [Google Scholar] [CrossRef] [PubMed]
  206. Wei, T.; Gu, L.; Zhou, M.; Zhou, Y.; Yang, H.; Li, M. Impact of Shock-Induced Cavitation Bubble Collapse on the Damage of Cell Membranes with Different Lipid Peroxidation Levels. J. Phys. Chem. B 2021, 125, 6912–6920. [Google Scholar] [CrossRef] [PubMed]
  207. Wu, J. Theoretical study on shear stress generated by microstreaming surrounding contrast agents attached to living cells. Ultrasound Med. Biol. 2002, 28, 125–129. [Google Scholar] [CrossRef] [PubMed]
  208. Marmottant, P.; Hilgenfeldt, S. Controlled vesicle deformation and lysis by single oscillating bubbles. Nature 2003, 423, 153–156. [Google Scholar] [CrossRef]
  209. Fan, Z.; Kumon, R.E.; Deng, C.X. Mechanisms of microbubble-facilitated sonoporation for drug and gene delivery. Ther. Deliv. 2014, 5, 467–486. [Google Scholar] [CrossRef] [Green Version]
  210. Nejad, S.M.; Hosseini, H.; Akiyama, H.; Tachibana, K. Reparable Cell Sonoporation in Suspension: Theranostic Potential of Microbubble. Theranostics 2016, 6, 446–455. [Google Scholar] [CrossRef]
  211. Qin, P.; Xu, L.; Han, T.; Du, L.; Yu, A.C. Effect of non-acoustic parameters on heterogeneous sonoporation mediated by single-pulse ultrasound and microbubbles. Ultrason. Sonochem. 2016, 31, 107–115. [Google Scholar] [CrossRef]
  212. Garbin, V.; Overvelde, M.; Dollet, B.; de Jong, N.; Lohse, D.; Versluis, M. Unbinding of targeted ultrasound contrast agent microbubbles by secondary acoustic forces. Phys. Med. Biol. 2011, 56, 6161–6177. [Google Scholar] [CrossRef]
  213. Postema, M.; Marmottant, P.; Lancée, C.T.; Hilgenfeldt, S.; de Jong, N. Ultrasound-induced microbubble coalescence. Ultrasound Med. Biol. 2004, 30, 1337–1344. [Google Scholar] [CrossRef]
  214. Huang, Y.; Das, P.K.; Bhethanabotla, V.R. Surface acoustic waves in biosensing applications. Sens. Actuators Rep. 2021, 3, 100041. [Google Scholar] [CrossRef]
  215. Frinking, P.; Segers, T.; Luan, Y.; Tranquart, F. Three Decades of Ultrasound Contrast Agents: A Review of the Past, Present and Future Improvements. Ultrasound Med. Biol. 2020, 46, 892–908. [Google Scholar] [CrossRef] [Green Version]
  216. Barnkob, R.; Augustsson, P.; Laurell, T.; Bruus, H. Acoustic radiation- and streaming-induced microparticle velocities determined by microparticle image velocimetry in an ultrasound symmetry plane. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 2012, 86, 056307. [Google Scholar] [CrossRef] [Green Version]
  217. Ahmed, D.; Ozcelik, A.; Bojanala, N.; Nama, N.; Upadhyay, A.; Chen, Y.; Hanna-Rose, W.; Huang, T.J. Rotational manipulation of single cells and organisms using acoustic waves. Nat. Commun. 2016, 7, 11085. [Google Scholar] [CrossRef]
  218. Liu, S.; Yang, Y.; Ni, Z.; Guo, X.; Luo, L.; Tu, J.; Zhang, D.; Zhang, A.J. Investigation into the Effect of Acoustic Radiation Force and Acoustic Streaming on Particle Patterning in Acoustic Standing Wave Fields. Sensors 2017, 17, 1664. [Google Scholar] [CrossRef] [Green Version]
  219. Zarnitsyn, V.G.; Meacham, J.M.; Varady, M.J.; Hao, C.; Degertekin, F.L.; Fedorov, A.G. Electrosonic ejector microarray for drug and gene delivery. Biomed. Microdevices 2008, 10, 299–308. [Google Scholar] [CrossRef] [PubMed]
  220. Belling, J.N.; Heidenreich, L.K.; Tian, Z.; Mendoza, A.M.; Chiou, T.T.; Gong, Y.; Chen, N.Y.; Young, T.D.; Wattanatorn, N.; Park, J.H.; et al. Acoustofluidic sonoporation for gene delivery to human hematopoietic stem and progenitor cells. Proc. Natl. Acad. Sci. USA 2020, 117, 10976–10982. [Google Scholar] [CrossRef] [PubMed]
  221. Carugo, D.; Ankrett, D.N.; Glynne-Jones, P.; Capretto, L.; Boltryk, R.J.; Zhang, X.; Townsend, P.A.; Hill, M. Contrast agent-free sonoporation: The use of an ultrasonic standing wave microfluidic system for the delivery of pharmaceutical agents. Biomicrofluidics 2011, 5, 44108–4410815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Salari, A.; Appak-Baskoy, S.; Coe, I.R.; Abousawan, J.; Antonescu, C.N.; Tsai, S.S.H.; Kolios, M.C. Dosage-controlled intracellular delivery mediated by acoustofluidics for lab on a chip applications. Lab Chip 2021, 21, 1788–1797. [Google Scholar] [CrossRef]
  223. Yoon, S.; Kim, M.G.; Chiu, C.T.; Hwang, J.Y.; Kim, H.H.; Wang, Y.; Shung, K.K. Direct and sustained intracellular delivery of exogenous molecules using acoustic-transfection with high frequency ultrasound. Sci. Rep. 2016, 6, 20477. [Google Scholar] [CrossRef] [Green Version]
  224. Yoon, S.; Wang, P.; Peng, Q.; Wang, Y.; Shung, K.K. Acoustic-transfection for genomic manipulation of single-cells using high frequency ultrasound. Sci. Rep. 2017, 7, 5275. [Google Scholar] [CrossRef] [Green Version]
  225. Guo, X.; Sun, M.; Yang, Y.; Xu, H.; Liu, J.; He, S.; Wang, Y.; Xu, L.; Pang, W.; Duan, X. Controllable Cell Deformation Using Acoustic Streaming for Membrane Permeability Modulation. Adv. Sci. 2021, 8, 2002489. [Google Scholar] [CrossRef]
  226. Ramesan, S.; Rezk, A.R.; Cevaal, P.M.; Cortez-Jugo, C.; Symons, J.; Yeo, L.Y. Acoustofection: High-Frequency Vibrational Membrane Permeabilization for Intracellular siRNA Delivery into Nonadherent Cells. ACS Appl. Bio Mater. 2021, 4, 2781–2789. [Google Scholar] [CrossRef]
  227. Yamashita, Y.; Shimada, M.; Tachibana, K.; Harimoto, N.; Tsujita, E.; Shirabe, K.; Miyazaki, J.; Sugimachi, K. In vivo gene transfer into muscle via electro-sonoporation. Hum. Gene Ther. 2002, 13, 2079–2084. [Google Scholar] [CrossRef]
  228. Blackman, B.R.; Barbee, K.A.; Thibault, L.E. In vitro cell shearing device to investigate the dynamic response of cells in a controlled hydrodynamic environment. Ann. Biomed. Eng. 2000, 28, 363–372. [Google Scholar] [CrossRef]
  229. Stewart, M.P.; Langer, R.; Jensen, K.F. Intracellular Delivery by Membrane Disruption: Mechanisms, Strategies, and Concepts. Chem. Rev. 2018, 118, 7409–7531. [Google Scholar] [CrossRef]
  230. Roti Roti, J.L. Cellular responses to hyperthermia (40–46 degrees C): Cell killing and molecular events. Int. J. Hyperthermia 2008, 24, 3–15. [Google Scholar] [CrossRef]
  231. Miller, A.D.; Subramanian, A.; Viljoen, H.J. Theoretically proposed optimal frequency for ultrasound induced cartilage restoration. Theor. Biol. Med. Model. 2017, 14, 21. [Google Scholar] [CrossRef] [Green Version]
  232. Furusawa, Y.; Fujiwara, Y.; Campbell, P.; Zhao, Q.L.; Ogawa, R.; Hassan, M.A.; Tabuchi, Y.; Takasaki, I.; Takahashi, A.; Kondo, T. DNA double-strand breaks induced by cavitational mechanical effects of ultrasound in cancer cell lines. PLoS ONE 2012, 7, e29012. [Google Scholar] [CrossRef]
  233. Xu, W. Microinjection and Micromanipulation: A Historical Perspective. Methods Mol. Biol. 2019, 1874, 1–16. [Google Scholar] [CrossRef]
  234. Dean, D.A. Gene delivery by direct injection (microinjection) using a pulsed-flow system. CSH Protoc. 2006, 2006, pdb-prot4653. [Google Scholar] [CrossRef] [PubMed]
  235. Eto, T.; Ueda, H.; Ito, R.; Takahashi, T.; Watanabe, T.; Goto, M.; Sotomaru, Y.; Tanaka, N.; Takahashi, R. Establishment of an integrated automated embryonic manipulation system for producing genetically modified mice. Sci. Rep. 2021, 11, 11770. [Google Scholar] [CrossRef] [PubMed]
  236. Green, M.; Wold, W.S. Oncogenic DNA viruses-replication, tumor gene expression, and role in human cancer. Semin. Oncol. 1976, 3, 65–79. [Google Scholar] [PubMed]
  237. Radko, S.; Koleva, M.; James, K.M.; Jung, R.; Mymryk, J.S.; Pelka, P. Adenovirus E1A targets the DREF nuclear factor to regulate virus gene expression, DNA replication, and growth. J. Virol. 2014, 88, 13469–13481. [Google Scholar] [CrossRef] [Green Version]
  238. Li, S.; Ou, M.; Wang, G.; Tang, L. Application of conditionally replicating adenoviruses in tumor early diagnosis technology, gene-radiation therapy and chemotherapy. Appl. Microbiol. Biotechnol. 2016, 100, 8325–8335. [Google Scholar] [CrossRef]
  239. Bett, A.J.; Haddara, W.; Prevec, L.; Graham, F.L. An efficient and flexible system for construction of adenovirus vectors with insertions or deletions in early regions 1 and 3. Proc. Natl. Acad. Sci. USA 1994, 91, 8802–8806. [Google Scholar] [CrossRef] [Green Version]
  240. Duigou, G.J.; Young, C.S. Replication-competent adenovirus formation in 293 cells: The recombination-based rate is influenced by structure and location of the transgene cassette and not increased by overproduction of HsRad51, Rad51-interacting, or E2F family proteins. J. Virol. 2005, 79, 5437–5444. [Google Scholar] [CrossRef] [Green Version]
  241. Lee, C.S.; Bishop, E.S.; Zhang, R.; Yu, X.; Farina, E.M.; Yan, S.; Zhao, C.; Zheng, Z.; Shu, Y.; Wu, X.; et al. Adenovirus-Mediated Gene Delivery: Potential Applications for Gene and Cell-Based Therapies in the New Era of Personalized Medicine. Genes Dis. 2017, 4, 43–63. [Google Scholar] [CrossRef]
  242. Ehrhardt, A.; Xu, H.; Kay, M.A. Episomal persistence of recombinant adenoviral vector genomes during the cell cycle in vivo. J. Virol. 2003, 77, 7689–7695. [Google Scholar] [CrossRef] [Green Version]
  243. 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]
  244. Nagano, S.; Oshika, H.; Fujiwara, H.; Komiya, S.; Kosai, K. An efficient construction of conditionally replicating adenoviruses that target tumor cells with multiple factors. Gene Ther. 2005, 12, 1385–1393. [Google Scholar] [CrossRef]
  245. Santiago-Ortiz, J.L.; Schaffer, D.V. Adeno-associated virus (AAV) vectors in cancer gene therapy. J. Control. Release 2016, 240, 287–301. [Google Scholar] [CrossRef] [Green Version]
  246. Kuzmin, D.A.; Shutova, M.V.; Johnston, N.R.; Smith, O.P.; Fedorin, V.V.; Kukushkin, Y.S.; van der Loo, J.C.M.; Johnstone, E.C. The clinical landscape for AAV gene therapies. Nat. Rev. Drug Discov. 2021, 20, 173–174. [Google Scholar] [CrossRef]
  247. Patel, A.; Zhao, J.; Duan, D.; Lai, Y. Design of AAV Vectors for Delivery of Large or Multiple Transgenes. Methods Mol. Biol. 2019, 1950, 19–33. [Google Scholar] [CrossRef]
  248. Fu, Q.; Polanco, A.; Lee, Y.S.; Yoon, S. Critical challenges and advances in recombinant adeno-associated virus (rAAV) biomanufacturing. Biotechnol. Bioeng. 2023, 1–21. [Google Scholar] [CrossRef]
  249. Marrone, L.; Marchi, P.M.; Azzouz, M. Circumventing the packaging limit of AAV-mediated gene replacement therapy for neurological disorders. Expert Opin. Biol. Ther. 2022, 22, 1163–1176. [Google Scholar] [CrossRef]
  250. Hanlon, K.S.; Kleinstiver, B.P.; Garcia, S.P.; Zaborowski, M.P.; Volak, A.; Spirig, S.E.; Muller, A.; Sousa, A.A.; Tsai, S.Q.; Bengtsson, N.E.; et al. High levels of AAV vector integration into CRISPR-induced DNA breaks. Nat. Commun. 2019, 10, 4439. [Google Scholar] [CrossRef] [Green Version]
  251. Hirai, H.; Satoh, E.; Osawa, M.; Inaba, T.; Shimazaki, C.; Kinoshita, S.; Nakagawa, M.; Mazda, O.; Imanishi, J. Use of EBV-based Vector/HVJ-liposome complex vector for targeted gene therapy of EBV-associated neoplasms. Biochem. Biophys. Res. Commun. 1997, 241, 112–118. [Google Scholar] [CrossRef]
  252. Gardlík, R.; Pálffy, R.; Hodosy, J.; Lukács, J.; Turna, J.; Celec, P. Vectors and delivery systems in gene therapy. Med. Sci. Monit 2005, 11, RA110–RA121. [Google Scholar]
  253. Yi, Y.; Noh, M.J.; Lee, K.H. Current advances in retroviral gene therapy. Curr. Gene Ther. 2011, 11, 218–228. [Google Scholar] [CrossRef]
  254. Palù, G.; Parolin, C.; Takeuchi, Y.; Pizzato, M. Progress with retroviral gene vectors. Rev. Med. Virol. 2000, 10, 185–202. [Google Scholar] [CrossRef]
  255. Escors, D.; Breckpot, K. Lentiviral vectors in gene therapy: Their current status and future potential. Arch. Immunol. Ther. Exp. 2010, 58, 107–119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Biological principle of RNA interference: 1. siRNA-pathway: RdRPs generate long dsRNA from single-stranded RNA templates, that are taken up by endocytosis and processed into siRNA by Dicer or TRBP which is loaded onto RISC. 2. miRNA-pathway: RNAPol II transcribes pri-miRNA, which is processed by RNAse III and DGCR8 protein to pre-miRNA. Pre-miRNA is exported by exportin 5 and processed by Dicer to dsmiRNA, which is loaded onto RISC. The passenger strand is degraded, and the guide strand can bind the target sequence and alter gene expression, by cleavage, methylation, translation inhibition, etc.
Figure 1. Biological principle of RNA interference: 1. siRNA-pathway: RdRPs generate long dsRNA from single-stranded RNA templates, that are taken up by endocytosis and processed into siRNA by Dicer or TRBP which is loaded onto RISC. 2. miRNA-pathway: RNAPol II transcribes pri-miRNA, which is processed by RNAse III and DGCR8 protein to pre-miRNA. Pre-miRNA is exported by exportin 5 and processed by Dicer to dsmiRNA, which is loaded onto RISC. The passenger strand is degraded, and the guide strand can bind the target sequence and alter gene expression, by cleavage, methylation, translation inhibition, etc.
Micromachines 14 01321 g001
Figure 2. Simplified principle of electroporation: Higher TMP causes the formation of random hydrophobic pores. Liquid penetrates, making it more stable for lipids to rotate and form hydrophilic pores.
Figure 2. Simplified principle of electroporation: Higher TMP causes the formation of random hydrophobic pores. Liquid penetrates, making it more stable for lipids to rotate and form hydrophilic pores.
Micromachines 14 01321 g002
Figure 3. Production of recombinant adenoviruses (rAdV): A shuttle vector containing the GOI is recombined with a plasmid containing adenoviral genes but lacking E1 and E3 (pAd). The resulting pAd-GOI is transfected into packaging cells that express E1A and allow replication of rAdV. Ultracentrifugation with CsCl (cesium chloride) is used to separate rAdV from cellular debris and finally dialyzed to obtain purified rAdV.
Figure 3. Production of recombinant adenoviruses (rAdV): A shuttle vector containing the GOI is recombined with a plasmid containing adenoviral genes but lacking E1 and E3 (pAd). The resulting pAd-GOI is transfected into packaging cells that express E1A and allow replication of rAdV. Ultracentrifugation with CsCl (cesium chloride) is used to separate rAdV from cellular debris and finally dialyzed to obtain purified rAdV.
Micromachines 14 01321 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Isenmann, M.; Stoddart, M.J.; Schmelzeisen, R.; Gross, C.; Della Bella, E.; Rothweiler, R.M. Basic Principles of RNA Interference: Nucleic Acid Types and In Vitro Intracellular Delivery Methods. Micromachines 2023, 14, 1321. https://doi.org/10.3390/mi14071321

AMA Style

Isenmann M, Stoddart MJ, Schmelzeisen R, Gross C, Della Bella E, Rothweiler RM. Basic Principles of RNA Interference: Nucleic Acid Types and In Vitro Intracellular Delivery Methods. Micromachines. 2023; 14(7):1321. https://doi.org/10.3390/mi14071321

Chicago/Turabian Style

Isenmann, Marie, Martin James Stoddart, Rainer Schmelzeisen, Christian Gross, Elena Della Bella, and René Marcel Rothweiler. 2023. "Basic Principles of RNA Interference: Nucleic Acid Types and In Vitro Intracellular Delivery Methods" Micromachines 14, no. 7: 1321. https://doi.org/10.3390/mi14071321

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