Next Article in Journal
Electrospun Fibers Loaded with Natural Bioactive Compounds as a Biomedical System for Skin Burn Treatment. A Review
Next Article in Special Issue
Novel Fluorinated Spermine and Small Molecule PEI to Deliver Anti-PD-L1 and Anti-VEGF siRNA for Highly Efficient Tumor Therapy
Previous Article in Journal
Monitoring the Effect of Transdermal Drug Delivery Patches on the Skin Using Terahertz Sensing
Previous Article in Special Issue
Novel PEGylated Lipid Nanoparticles Have a High Encapsulation Efficiency and Effectively Deliver MRTF-B siRNA in Conjunctival Fibroblasts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

New Applications of Lipid and Polymer-Based Nanoparticles for Nucleic Acids Delivery

by
Adelina-Gabriela Niculescu
1,
Alexandra Cătălina Bîrcă
1 and
Alexandru Mihai Grumezescu
1,2,3,*
1
Department of Science and Engineering of Oxide Materials and Nanomaterials, University Politehnica of Bucharest, 011061 Bucharest, Romania
2
Research Institute of the University of Bucharest—ICUB, University of Bucharest, 050657 Bucharest, Romania
3
Academy of Romanian Scientists, Ilfov No. 3, 50044 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Pharmaceutics 2021, 13(12), 2053; https://doi.org/10.3390/pharmaceutics13122053
Submission received: 29 October 2021 / Revised: 26 November 2021 / Accepted: 29 November 2021 / Published: 1 December 2021

Abstract

:
Nucleic acids represent a promising lead for engineering the immune system. However, naked DNA, mRNA, siRNA, and other nucleic acids are prone to enzymatic degradation and face challenges crossing the cell membrane. Therefore, increasing research has been recently focused on developing novel delivery systems that are able to overcome these drawbacks. Particular attention has been drawn to designing lipid and polymer-based nanoparticles that protect nucleic acids and ensure their targeted delivery, controlled release, and enhanced cellular uptake. In this respect, this review aims to present the recent advances in the field, highlighting the possibility of using these nanosystems for therapeutic and prophylactic purposes towards combatting a broad range of infectious, chronic, and genetic disorders.

1. Introduction

The recent investigation of a multitude of therapeutic nucleic acids (TNAs) has paved the way for developing new treatment strategies for various wounds, infectious diseases, and chronic conditions. TNAs, such as plasmid DNA (pDNA), small interfering (or silencing) RNA (siRNA), self-amplifying RNA (saRNA), microRNA mimics, anti-microRNA oligonucleotides, messenger RNA (mRNA), and antisense oligonucleotides (ASOs), can be employed for creating therapeutics and vaccines that provide a cell-mediated immune response [1,2,3,4,5,6,7].
TNAs have gained increasing scientific interest, especially due to the long-lasting effects they offer in contrast to conventional treatments. Specifically, traditional medication induces temporary effects as it targets proteins instead of the underlying causes. In comparison, TNAs have the potential to produce long-term and even curative effects by gene inhibition, addition, replacement, or editing [2].
However, the direct delivery of nucleic acids has several drawbacks as naked nucleic acids are prone to enzymatic degradation, renal clearance, and poor cellular uptake as they face difficulties in crossing the cell membrane [4,8,9,10]. These challenges are mainly due to TNAs’ large molecular weight, negatively charged backbone, and more fragile and immunogenic potential than their oligonucleotide counterparts [6,11]. Thus, the clinical translation of such treatments is highly dependent on the delivery technologies that must enhance TNAs’ stability, protect against extracellular degradation, facilitate cell internalization, and improve target affinity [2,3,8].
In this respect, a broad range of delivery systems has been studied. Delivery vehicles such as lipid nanoparticles, polymeric micelles, dendrimers, polymer-based nanoparticles, hydrogels, polyplexes, proteins, and inorganic nanomaterials are under investigation [1,12]. Out of this plethora of possibilities, lipid nanoparticles are the most clinically advanced [11,13,14]; however, polymer and polymer–lipid hybrid particles have also recently become important categories of carrier platforms [15,16,17,18]. Moreover, such delivery systems have become highly attractive due to their targeting potential as they can be surface functionalized to allow accumulation and payload release in specific tissues [11,19,20].
In this context, this paper aims to provide an overview of the newly developed lipid and polymer-based nanosystems for nucleic acid delivery, emphasizing their importance in combatting various infectious, chronic, and genetic disorders.

2. Lipid-Based Delivery Systems

Lipid nanoparticles (LNPs) have attracted considerable research interest for encapsulating TNAs, especially due to their ionizable lipid presence, which is cationic at a low pH, thus, allowing complexation with negatively charged RNA or DNA [21,22]. Various generations of lipid nanocarriers have been tackled for TNAs delivery, including liposomes, solid lipid nanoparticles, nanostructured lipid carriers, and cationic lipid–nucleic acid complexes [23]. A brief history of how lipid-based RNA delivery systems evolved until the date is represented in Figure 1.
LNPs have also been extensively studied due to the relatively easy and scalable manufacturing processes they can be obtained through [3]. Specialized manufacturing and control techniques ensure the complex morphology and tailored lipid components and proportions required to create functional and efficient LNP-based delivery systems [11]. Examples of conventional synthesis methods include high-pressure homogenization [25,26], solvent emulsification–evaporation [27], ethanol injection nanoprecipitation [28], freeze-drying [29], and preformed vesicle method [30] (Figure 2).
For instance, Gomez-Aguado et al. [27] used the solvent emulsification-evaporation method to develop different solid lipid nanoparticles (SLNs) that combine cationic and ionizable lipids for the delivery of mRNA and pDNA. Their comparative study showed a higher percentage of transfected cells in mRNA formulation than for particles containing pDNA, especially in human retinal pigment epithelial cells (ARPE-19). Nonetheless, pDNA delivery resulted in greater protein production per cell in the mentioned cell line. Among the tested lipid formulations, SLNs containing only 1,2-dioleoyl-3-trimethylammonium-propane are considered the most promising for TNAs delivery.
Other synthesis methods were used by Wang et al. [29], who have created aerosolizable dry powder of lipid nanoparticles by thin-film freeze-drying (TFFD), spray drying, and conventional shelf freeze-drying. The researchers comparatively evaluated the obtained powders and tested the feasibility of engineering SLNs using the TFFD method, concluding that this synthesis strategy leads to better aerosol performance properties than the powders obtained otherwise.
However, some of the newest approaches for synthesizing LNPs are based on the direct mix of the organic phase (containing the lipids) with the aqueous phase (containing the TNA) using a microfluidic device [32]. Microfluidics represents a robust, scalable, and reproducible synthesis method [33]. This strategy has several advantages compared to conventional techniques as it allows the strict and easy control of flow rates, rapid mixing of the phases, and decreased synthesis time, resulting in nanoparticles with a small size, high monodispersity, and high encapsulation efficiency [28,34].
For instance, Bailet-Hytholt et al. [21] have synthesized LNPs encapsulated with mRNA or pDNA by means of a precise microfluidic mixing platform. Their experimental setup allowed the self-assembly of nanoparticles under laminar flow, benefiting from advantages such as reproducibility, speed, and low volume screening. Microfluidic methods were also employed by Roces et al. [33], who have prepared PolyA, ssDNA, and mRNA encapsulated LNPs in a Y-shape staggered herringbone micromixer (Figure 3). In this manner, the researchers obtained a consistent range of particle sizes, with dimensions below 100 nm, narrow size distribution, spherical shape, good stability, and high encapsulation efficiencies.
In what concerns their applicability, numerous lipid-based TNA delivery platforms have been recently constructed as potent novel platforms for cancer immunotherapies [35]. In particular, the use of mRNA formulations for different types of solid tumors and hematological malignancies is currently examined by various studies and even clinical trials [28]. Targeted LNPs can deliver RNA-based therapeutics to leukocytes to allow the precise and modular manipulation of gene expression, thus being a highly promising approach for many immune-related disorders, such as cancer, autoimmunity, and susceptibility to infectious diseases [36].
TNAs can also be employed in formulations for the treatment and prevention of cardiovascular diseases [37,38]. Clinical trials have investigated ASOs for lowering low-density lipoprotein cholesterol (LDL-C) [39], lowering transthyretin (TTR) blood levels [40,41,42], reducing atrial fibrillation burden [43], reducing serum lipoprotein (a) levels [44], and preventing neovascular glaucoma [45]. mRNA has also been evaluated for cardiovascular conditions, with several LNP-based formulations being tested for enhancing cardiac function and regeneration [46], reducing LDL-C levels [47], and treating endotheliopathy associated with vascular senescence [48,49]. Similarly, siRNA lipid-based formulations have been clinically tested for TTR-mediated amyloidosis [50] and hypercholesterolemia [51].
Another highly researched application of LNPs is for the development of COVID-19 vaccines, out of which several formulations have already entered clinical use for mRNA delivery [16,52,53,54,55]. Generally, these vaccines work on the principle of encoding the spike glycoprotein of SARS-CoV-2 (Figure 4) and directing the immune system against these antigens [56,57]. Clinical trials demonstrated the efficacy of LNP-mRNA formulations and an acceptable safety profile, leading to their approval for mass immunization [58,59].
COVID-19 vaccines are the most known current examples that make use of LNPs encapsulated with TNA, but this approach can be used for other infectious diseases as well [37,61,62,63,64,65,66,67]. TNA-LNP formulations have been evaluated against Zika virus [68], influenza [69,70], cytomegalovirus [71,72], chikungunya virus [73], hepatitis B [74,75,76], hepatitis C [77,78,79], human immunodeficiency virus (HIV) [80,81], and genital herpes [82].
The development of nonviral platforms for prenatal TNA delivery has recently emerged among the scientific research hot topics as a strategy to treat diseases before the onset of irreversible pathology and a non-invasive alternative to difficult-to-perform surgeries. What is more, the small size of a fetus compared to a postnatal recipient is a key factor for maximizing the delivery vector titer per weight of the recipient, thus, facilitating gene transduction [37,83]. In this regard, Riley et al. [84] have recently developed a library of ionizable LNPs for in utero mRNA delivery to mouse fetuses. Their LNPs successfully accumulated within fetal livers, lungs, and intestines, with higher therapeutic efficacy and safety than reference delivery systems (i.e., DLin-MC3-DMA and jetPEI), being promising candidates for protein replacement and gene editing platforms.
A series of examples of potential applications of lipid-based TNA delivery systems have been synthesized in Table 1 to summarize the above discussion and emphasize the versatility of such formulations.

3. Polymer-Based Delivery Systems

Cationic polymers have attracted interest for the development of innovative TNA carriers. The main reasoning is that they form electrostatic nanocomplexes with nucleic acids, which are highly negative, to facilitate their uptake by targeted cells. Alternatively, other hydrophobic polymers can physically entrap TNAs within nanoparticles [4].
The first polymer tested as a vehicle for in vitro-transcribed mRNA was diethylaminoethyl (DEAE) dextran. However, after it was proven that lipid-mediated mRNA transfection is 100 to 1000 times more efficient than DEAE-dextran, the evolution of polymeric carriers has been stalled, the attention shifting towards creating perpetually improved lipid-based nanosystems [111].
Nonetheless, after some time, research started being oriented towards developing natural and synthetic polymer-based nanosystems for TNAs delivery as an alternative for cancer treatments and non-viral vaccine formulations. Polymers such as hyaluronic acid [94,112,113,114], chitosan [94,113], DEAE dextran [111], poly-L-lysine (PLL) [94], poly-beta-amino-esters (PBAE) [111], polyethylenimine (PEI) [94,111], dimethylaminoethyl methacrylate (DEAMA) [111], poly(ethylene glycol) methacrylate (PEGMA) [111], and DEAEMA-co-n-butyl methacrylate [111] represent promising candidates for various TNAs delivery platforms.
Another interesting possibility is to combine two or more polymers in the same delivery system. For instance, Lü et al. [115] have developed antibody (Ab)-conjugated lactic-co-glycolic-PEI nanoparticles (LGA-PEI NPs) for the active-targeting delivery of TNAs. The research group obtained promising results for pancreatic cancer cells delivery both in vitro and in vivo, concluding that the synthesized carriers can be employed for other types of cancers as well if other specific biomarkers are targeted.
Polymers have also been tested for microneedle-based TNA delivery [111]. Koh et al. [116] have fabricated an mRNA-loaded dissolving microneedle patch from polyvinylpyrrolidone (PVP) for cutaneous TNA administration. This RNA-patch can mediate in vivo transgene expression for up to 72 h, the transfection efficiency and kinetics being comparable to subcutaneous injection. Moreover, this innovative delivery system induced higher cellular and humoral immune responses than subcutaneous injection. A similar administration strategy was used by Pan et al. [117], who have delivered signal transducer and activity transcription 3 (STAT3)-targeting PEI-encapsulated siRNA through dextran-hyaluronic acid dissolving microneedles. The researchers obtained promising results as the microneedles effectively penetrated and rapidly dissolved into the skin. In addition, the delivered complex successfully suppressed the development of melanoma by silencing the STAT3 gene.
To briefly summarize the current research progress, several examples of polymer-based TNA delivery systems and their applications have been gathered in Table 2.
Although various types of polymers and copolymers have been tested for TNAs delivery, the poorly understood correlation between their structure and their biological response hinders their development. Moreover, the polydispersity and difficulty in metabolizing large molecular weight polymers make them less appealing than LNPs. Their clinical application is also impeded by potential toxicity, colloidal instability, and relatively poor transfection efficiency [64,111]. Therefore, more research is required for understanding the influence of polymers’ chemical structure on their biological activity and optimally modulating their properties to overcome current challenges.

4. Lipid-Polymer Hybrid-Based Delivery Systems

Hybrid delivery systems have recently appeared as an alternative solution for overcoming the limitations of each individual lipid or polymer component. Specifically, the adverse effects of ionizable lipids (e.g., potential immune activation, cytotoxicity) are commonly mediated by shielding the LNPs with polyethylene glycol (PEG) [11].
For example, Sanghani et al. [123] have PEGylated LNPs made of pH-sensitive cationic lipid CL4H6 to safely and efficiently deliver myocardin-related transcription factor B (MRTF-B) siRNA into human conjunctival fibroblasts. A similar hybrid approach on TNA delivery was taken by Scmendel et al. [20], who have developed folate-containing lipoconjugates with PEG spacers incorporated into a liposome. These delivery systems exhibited higher transfection efficiency compared to non-targeted liposomal formulations and the commercial agent Lipofectamine 2000. Improved transfection efficiency was also obtained using the hybrid nanoparticles created by Siewert et al. [124]. The researchers have developed vehicles for mRNA delivery using different proportions of the cationic lipid DOTAP and the cationic biopolymer protamine, reporting significantly increased transfection in comparison to lipid/mRNA and polymer/mRNA particles alone. Another mRNA hybrid carrier was created by Xiong et al. [125]. This research group developed theranostic dendrimer-based LNPs containing PEGylated BODIPY dyes that combine mRNA delivery and NIR imaging, holding great promise for cancer’s simultaneous detection and treatment.
Zhou et al. [126] have synthesized a series of linear-dendritic PEG lipids (PEG-GnCm) to investigate the effect of modulating their hydrophobic domain for siRNA delivery as the surface component of dendrimer lipid-based nanoparticles (DLNPs). The researchers used different lipid alkyl lengths (C8, C12, C16) and different dendrimer generations (G1, G2, G3), creating vehicles with different physical properties and anchoring potential. These alterations did not affect particle size, RNA encapsulation, and stability but had a huge impact on delivery efficacy. PEG-G1C8, PEG-G1C12, PEG-G1C16, and PEG-G2C8 could effectively deliver siRNA in vitro and in vivo; however, all the other tested formulations lost their delivery ability, as the escape of DLNPs from endosomes at early cell incubation times was affected.
Furthermore, by tailoring the surface charge of LNPs, they can be endowed with targeting ability. According to the comparative study performed by Gabal et al. [127], anionic nanostructured lipid carriers exhibited 1.2 times higher targeting efficiency in the brain than their cationic counterparts. Thus, such anionic particles hold promise as carriers for brain disorders therapeutics. Nonetheless, anionic formulations have not faced the same development because of difficulties encountered in nucleic acid packaging and poor transfection efficiency. To overcome these challenges, Tagalakis et al. [128] have used cationic targeting peptides as a bridge between the PEGylated anionic liposomes and the pDNA freight. The newly developed structures demonstrated improved tissue penetration, enhanced dispersal, and more widespread cellular transfection than cationic systems. Similar anionic targeting hybrid nanocarriers have also shown promising results for siRNA delivery to neuroblastoma tumors with reduced systemic and cellular toxicity [129].
Another polymer that can be used in combination with LNPs is poly(lactic-co-glycolic acid) (PLGA). In this respect, Yang et al. [130] have developed a PLGA-LNP hybrid delivery system loaded with CRISPR/Cas9 plasmids targeting the MGMT gene and modified with the cRGD peptide. This system was constructed to open the blood–brain barrier (BBB) and ensure targeted gene delivery in vivo under focused ultrasound (FUS) irradiation. The obtained results emphasized a synergic targeting ability of the physically site-specific characteristics of FUS and the biologically active targeting ability of cRGD peptide, recommending this innovative carrier as a candidate for central nervous system TNA delivery.
A recently approached strategy for designing performant hybrid carriers is the creation of lipopolyplexes (LPPs), complexes combining nucleic acids with lipids and polymers (Figure 5). Such structures are promising delivery systems for gene therapy, especially due to their compositional, physical, and functional versatility [131,132,133]. For instance, Wang et al. [134] have synthesized a tumor-selective LPP consisting of a PEI/p21-saRNA-322 core and a hyaluronan-modulated lipid shell. The system was tested against colorectal cancer, and it was reported that it accumulated preferentially at the tumor site, leading to superior antitumor efficacy in vitro and in vivo. Alternatively, Shah et al. [135] have proposed the incorporation of ultrasound contrast agents in the liposomal cavity, followed by polyplexes addition. Thus, the scientists obtained ultrasound-activated LPPs that promote cancer cell uptake, elevate transfection efficiency, and reduce carrier cytotoxicity.
One more emerging delivery possibility is TNA transport via lipodendriplexes which are complex structures formed by non-covalent hybridization of dendriplexes with the liposomal membrane. In this respect, Tariq et al. [136] have incorporated pDNA-PAMAM-based dendriplexes into an optimized liposomal formulation (Figure 6). The researchers reported significantly improved pDNA transfection, lower LDH release, lower ROS generation, higher cellular protein content, and higher cell viability compared to dendriplexes without the lipid components. Thus, these new carriers can be considered promising candidates for efficient and biocompatible gene delivery systems.
To summarize the current status of lipid polymer-hybrid-based TNA carriers, Table 3 presents several examples of such delivery systems together with their potential applications.

5. Challenges and Emerging Solutions to Overcome Them

LNPs are undeniably promising innovative nonviral vectors for gene delivery. Nonetheless, several challenges still exist, limiting their potential. One of the main unresolved issues is represented by the poor endosomal escape after LNP cell entry. Attempting to understand this phenomenon, Herrera et al. [140] have created a highly sensitive and robust galectin 8-GFP (Gal8-GFP) cell reporter system to visualize the endosomal escape capabilities of LNP-encapsulated mRNA. This sensor system allows the rapid and efficient distinction of endosomal membrane integrity as an indicator of cytosolic availability of mRNA. Moreover, it helped the researchers identify differences in endosomal escape capabilities elicited by the varying sterol composition of mRNA LNP-based delivery systems.
Alternatively, Mihaila et al. [141] propose the optimization of siRNA LNP-mediated delivery by use of an ordinary differential equation (ODE)-based model. By means of mathematical modeling, the researchers designed and validated a predictive model that can compare the relative kinetics of different classes of LNPs towards choosing the best option.
Another problem associated with LNPs is that their intravenous administration results in liver accumulation where the reticuloendothelial system takes them up. To avoid this situation, Saunders et al. [142] propose the administration of a liposome (i.e., Nanoprimer) that can temporarily occupy liver cells as a pretreatment before LNPs delivery. The researchers obtained promising results as the Nanoprimer improved the bioavailability of tested RNA-encapsulated LNPs, increased protein production, and enhanced FVII silencing.
One more aspect that must be considered regarding lipid and polymer-based TNA delivery systems is the development of a biomolecular corona around these nanoparticles (Figure 7). This “crown” is formed by electrolytes, proteins, and lipids adsorbed on nanomaterials’ surfaces when in contact with a biological environment. As the association process is almost irreversible, the biomolecular corona defines the biological identity of the nanosystem, impacting in vivo characteristics, such as circulation time, biodistribution, uptake kinetics, macrophage recognition, release profile, or targeting [143,144,145].
Thus, when moving from in vitro models to in vivo studies or clinical trials, it may be expected to encounter a weaker delivery efficiency and therapeutic effect of carried nucleic acids. To overcome this issue, Yang et al. [147] have proposed the design of a corona made from cyclic RGDyK peptide-modified bovine serum albumin to be used as a precoat on a redox-responsive chitosan-based nanocarrier for siRNA delivery. The synthesized corona remained steady around the delivery vehicle, improved the system’s stability, biocompatibility, and targeting ability, reduced serum proteins adsorption, increased intracellular uptake, facilitated the lysosomal escape while maintaining the redox-sensitive responsiveness of the nanocarrier.

6. Conclusions

To conclude, TNAs represent a promising therapeutic strategy for a broad range of diseases. Benefiting from tremendous scientific interest in the recent years, numerous LNP-based delivery systems for pDNA, mRNA, siRNA, and other nucleic acids have appeared that can help fight against various cancers, infectious diseases, cardiovascular diseases, and inherited disorders. Several such formulations have reached the clinical trials testing stage or have even been approved for use in the general population in record time, as is the case of LNP-mRNA COVID-19 vaccines.
Other TNA delivery possibilities started to emerge as well, including polymer and lipid–polymer hybrid carriers. Nonetheless, these delivery systems are still in their infancy, most of them requiring further thorough research before moving from in vitro and in vivo tests to clinical studies.
Overall, the evolution in developing targeted and controlled release vehicles for efficient cellular uptake faces exponentially increasing progress. Given the recent research growth in the field and the precedent it has been created with the approval of the COVID-19 vaccines, it can be expected that other TNA delivery systems will soon enter the market.

Author Contributions

Conceptualization, A.-G.N., A.C.B. and A.M.G.; methodology, A.-G.N., A.C.B. and A.M.G.; writing—original draft preparation, A.-G.N., A.C.B. and A.M.G.; writing—review and editing, A.-G.N., A.C.B. and A.M.G. All authors have read and agreed to the published version of the manuscript.

Funding

The present work was possible due to the project 36355/23.05.2019 POCU/380/6/13.

Institutional Review Board Statement

Not applicable.

Acknowledgments

The present work was possible due to the project “Excelența academică și valori antreprenoriale—sistem de burse pentru asigurarea oportunităților de formare și dezvoltare a competențelor antreprenoriale ale doctoranzilor și postdoctoranzilor”—ANTREPRENORDOC (36355/23.05.2019 POCU/380/6/13).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berger, A.G.; Chou, J.J.; Hammond, P.T. Approaches to Modulate the Chronic Wound Environment Using Localized Nucleic Acid Delivery. Adv. Wound Care 2021, 10, 503–528. [Google Scholar] [CrossRef] [PubMed]
  2. Kulkarni, J.A.; Witzigmann, D.; Thomson, S.B.; Chen, S.; Leavitt, B.R.; Cullis, P.R.; van der Meel, R. The current landscape of nucleic acid therapeutics. Nat. Nanotechnol. 2021, 16, 630–643. [Google Scholar] [CrossRef] [PubMed]
  3. Aldosari, B.; Alfagih, I.; Almurshedi, A. Lipid nanoparticles as delivery systems for RNA-based vaccines. Pharmaceutics 2021, 13, 206. [Google Scholar] [CrossRef] [PubMed]
  4. Tzeng, S.Y.; Green, J.J. Polymeric nucleic acid delivery for immunoengineering. Curr. Opin. Biomed. Eng. 2018, 7, 42–50. [Google Scholar] [CrossRef]
  5. Blakney, A.K.; McKay, P.F.; Hu, K.; Samnuan, K.; Jain, N.; Brown, A.; Thomas, A.; Rogers, P.; Polra, K.; Sallah, H.; et al. Polymeric and lipid nanoparticles for delivery of self-amplifying RNA vaccines. J. Control. Release 2021, 338, 201–210. [Google Scholar] [CrossRef] [PubMed]
  6. Dammes, N.; Peer, D. Paving the Road for RNA Therapeutics. Trends Pharmacol. Sci. 2020, 41, 755–775. [Google Scholar] [CrossRef] [PubMed]
  7. Massaro, C.; Sgueglia, G.; Frattolillo, V.; Baglio, S.R.; Altucci, L.; Dell’Aversana, C. Extracellular Vesicle-Based Nucleic Acid Delivery: Current Advances and Future Perspectives in Cancer Therapeutic Strategies. Pharmaceutics 2020, 12, 980. [Google Scholar] [CrossRef]
  8. Li, B.; Zhang, X.; Dong, Y. Nanoscale platforms for messenger RNA delivery. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2019, 11, e1530. [Google Scholar] [CrossRef]
  9. Meng, Z.; O’Keeffe-Ahern, J.; Lyu, J.; Pierucci, L.; Zhou, D.; Wang, W. A new developing class of gene delivery: Messenger RNA-based therapeutics. Biomater. Sci. 2017, 5, 2381–2392. [Google Scholar] [CrossRef]
  10. Ramasamy, T.; Munusamy, S.; Ruttala, H.B.; Kim, J.O. Smart Nanocarriers for the Delivery of Nucleic Acid-Based Therapeutics: A Comprehensive Review. Biotechnol. J. 2021, 16, 1900408. [Google Scholar] [CrossRef] [PubMed]
  11. Samaridou, E.; Heyes, J.; Lutwyche, P. Lipid nanoparticles for nucleic acid delivery: Current perspectives. Adv. Drug Deliv. Rev. 2020, 154, 37–63. [Google Scholar] [CrossRef]
  12. Kim, D.; Wu, Y.; Kim, Y.B.; Oh, Y.-K. Advances in vaccine delivery systems against viral infectious diseases. Drug Deliv. Transl. Res. 2021, 11, 1401–1419. [Google Scholar] [CrossRef] [PubMed]
  13. Ramishetti, S.; Hazan-Halevy, I.; Palakuri, R.; Chatterjee, S.; Naidu Gonna, S.; Dammes, N.; Freilich, I.; Kolik Shmuel, L.; Danino, D.; Peer, D. A combinatorial library of lipid nanoparticles for RNA delivery to leukocytes. Adv. Mater. 2020, 32, 1906128. [Google Scholar] [CrossRef]
  14. Schlich, M.; Palomba, R.; Costabile, G.; Mizrahy, S.; Pannuzzo, M.; Peer, D.; Decuzzi, P. Cytosolic delivery of nucleic acids: The case of ionizable lipid nanoparticles. Bioeng. Transl. Med. 2021, 6, e10213. [Google Scholar] [CrossRef] [PubMed]
  15. Park, K.S.; Sun, X.; Aikins, M.E.; Moon, J.J. Non-viral COVID-19 vaccine delivery systems. Adv. Drug Deliv. Rev. 2021, 169, 137–151. [Google Scholar] [CrossRef]
  16. Hou, X.; Zaks, T.; Langer, R.; Dong, Y. Lipid nanoparticles for mRNA delivery. Nat. Rev. Mater. 2021, 6, 1078–1095. [Google Scholar] [CrossRef]
  17. Thi, T.; Suys, E.; Lee, J.; Nguyen, D.; Park, K.; Truong, N. Lipid-Based Nanoparticles in the Clinic and Clinical Trials: From Cancer Nanomedicine to COVID-19 Vaccines. Vaccines 2021, 9, 359. [Google Scholar] [CrossRef] [PubMed]
  18. Milane, L.; Amiji, M. Clinical approval of nanotechnology-based SARS-CoV-2 mRNA vaccines: Impact on translational nanomedicine. Drug Deliv. Transl. Res. 2021, 11, 1309–1315. [Google Scholar] [CrossRef] [PubMed]
  19. Scheideler, M.; Vidakovic, I.; Prassl, R. Lipid nanocarriers for microRNA delivery. Chem. Phys. Lipids 2020, 226, 104837. [Google Scholar] [CrossRef]
  20. Shmendel, E.; Kabilova, T.; Morozova, N.; Zenkova, M.; Maslov, M. Effects of spacers within a series of novel folate-containing lipoconjugates on the targeted delivery of nucleic acids. J. Drug Deliv. Sci. Technol. 2020, 57, 101609. [Google Scholar] [CrossRef]
  21. Bailey-Hytholt, C.M.; Ghosh, P.; Dugas, J.; Zarraga, I.E.; Bandekar, A. Formulating and Characterizing Lipid Nanoparticles for Gene Delivery using a Microfluidic Mixing Platform. J. Vis. Exp. 2021, 168, e62226. [Google Scholar] [CrossRef] [PubMed]
  22. Gaur, P.K.; Pal, H.; Puri, D.; Kumar, N.; Shanmugam, S.K. Formulation and development of hesperidin loaded solid lipid nanoparticles for diabetes. Biointerface Res. Appl. Chem. 2020, 10, 4728–4733. [Google Scholar] [CrossRef]
  23. Tenchov, R.; Bird, R.; Curtze, A.E.; Zhou, Q. Lipid Nanoparticles─From Liposomes to mRNA Vaccine Delivery, a Landscape of Research Diversity and Advancement. ACS Nano 2021, 15, 16982–17015. [Google Scholar] [CrossRef] [PubMed]
  24. Mantri, S. Strategies for Successful Formulation Development of Lipid-Based RNA Delivery and Vaccines. Available online: https://www.sigmaaldrich.com/deepweb/assets/sigmaaldrich/product/documents/376/366/safc-lipids-rna-wp7004en-ms.pdf (accessed on 10 October 2021).
  25. Corrias, F.; Lai, F. New methods for lipid nanoparticles preparation. Recent Patents Drug Deliv. Formul. 2011, 5, 201–213. [Google Scholar] [CrossRef] [PubMed]
  26. Torrecilla, J.; Rodríguez-Gascón, A.; Solinís, M.Á.; del Pozo-Rodríguez, A. Lipid Nanoparticles as Carriers for RNAi against Viral Infections: Current Status and Future Perspectives. BioMed Res. Int. 2014, 2014, 1–17. [Google Scholar] [CrossRef]
  27. Gómez-Aguado, I.; Rodríguez-Castejón, J.; Vicente-Pascual, M.; Rodríguez-Gascón, A.; del Pozo-Rodríguez, A.; Solinís Aspiazu, M.Á. Nucleic Acid Delivery by Solid Lipid Nanoparticles Containing Switchable Lipids: Plasmid DNA vs. Messenger RNA. Molecules 2020, 25, 5995. [Google Scholar] [CrossRef] [PubMed]
  28. Guevara, M.L.; Persano, F.; Persano, S. Advances in lipid nanoparticles for mRNA-based cancer immunotherapy. Front. Chem. 2020, 8, 963. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, J.-L.; Hanafy, M.S.; Xu, H.; Leal, J.; Zhai, Y.; Ghosh, D.; Williams Iii, R.O.; Smyth, H.D.C.; Cui, Z. Aerosolizable siRNA-encapsulated solid lipid nanoparticles prepared by thin-film freeze-drying for potential pulmonary delivery. Int. J. Pharm. 2021, 596, 120215. [Google Scholar] [CrossRef]
  30. Evers, M.J.W.; Kulkarni, J.; Van Der Meel, R.; Cullis, P.R.; Vader, P.; Schiffelers, R.M. State-of-the-Art Design and Rapid-Mixing Production Techniques of Lipid Nanoparticles for Nucleic Acid Delivery. Small Methods 2018, 2, 1700375. [Google Scholar] [CrossRef]
  31. Lundberg, S.; Karlsson, E.; Dahlberg, H.; Glansk, M.; Larsson, S.; Larsson, S.; Carlsson, K. Exosomes and Lipid Nanoparticles-the Future of Targeted Drug Delivery. Bachelor’s Degree, Uppsala University, Uppsala, Sweden, 2020. [Google Scholar]
  32. Patel, S.; Ashwanikumar, N.; Robinson, E.; Xia, Y.; Mihai, C.; Griffith, J.P.; Hou, S.; Esposito, A.A.; Ketova, T.; Welsher, K. Naturally-occurring cholesterol analogues in lipid nanoparticles induce polymorphic shape and enhance intracellular delivery of mRNA. Nat. Commun. 2020, 11, 1–13. [Google Scholar]
  33. Roces, C.B.; Lou, G.; Jain, N.; Abraham, S.; Thomas, A.; Halbert, G.W.; Perrie, Y. Manufacturing Considerations for the Development of Lipid Nanoparticles Using Microfluidics. Pharmaceutics 2020, 12, 1095. [Google Scholar] [CrossRef] [PubMed]
  34. Niculescu, A.-G.; Chircov, C.; Bîrcă, A.; Grumezescu, A. Nanomaterials Synthesis through Microfluidic Methods: An Updated Overview. Nanomaterials 2021, 11, 864. [Google Scholar] [CrossRef] [PubMed]
  35. Obeid, M.A.; Aljabali, A.A.A.; Rezigue, M.; Amawi, H.; Alyamani, H.; Abdeljaber, S.N.; Ferro, V.A. Use of nanoparticles in delivery of nucleic acids for melanoma treatment. Methods Mol. Biol. 2021, 2265, 591–620. [Google Scholar] [PubMed]
  36. Veiga, N.; Diesendruck, Y.; Peer, D. Targeted lipid nanoparticles for RNA therapeutics and immunomodulation in leukocytes. Adv. Drug Deliv. Rev. 2020, 159, 364–376. [Google Scholar] [CrossRef] [PubMed]
  37. Li, M.; Li, Y.; Li, S.; Jia, L.; Wang, H.; Li, M.; Deng, J.; Zhu, A.; Ma, L.; Li, W.; et al. The nano delivery systems and applications of mRNA. Eur. J. Med. Chem. 2021, 227, 113910. [Google Scholar] [CrossRef]
  38. Magadum, A.; Kaur, K.; Zangi, L. mRNA-Based Protein Replacement Therapy for the Heart. Mol. Ther. 2019, 27, 785–793. [Google Scholar] [CrossRef] [Green Version]
  39. Effect of Mipomersen on LDL-Cholesterol Levels in Patients Treated by Regular Apheresis (MICA). Available online: https://clinicaltrials.gov/ct2/show/NCT01598948 (accessed on 12 October 2021).
  40. The Effect of an Antisense Oligonucleotide to Lower Transthyretin (TTR) Levels on the Progression of -Wild-type TTR Involving the Heart. Available online: https://clinicaltrials.gov/ct2/show/NCT02627820 (accessed on 12 October 2021).
  41. 24 Month Open Label Study of the Tolerability and Efficacy of Inotersen in TTR Amyloid Cardiomyopathy Patients. Available online: https://clinicaltrials.gov/ct2/show/NCT03702829 (accessed on 12 October 2021).
  42. ION-682884 in Patients With TTR Amyloid Cardiomyopathy. Available online: https://clinicaltrials.gov/ct2/show/NCT04843020 (accessed on 12 October 2021).
  43. Antiarrhythmic and Symptomatic Effect of ISIS CRP Rx Targeting CRP in Paroxysmal Atrial Fibrillation (ASET). Available online: https://clinicaltrials.gov/ct2/show/NCT01710852 (accessed on 12 October 2021).
  44. Influence of Pelacarsen on Patients After Myocardial Infarction With High Lp(a) Values (PEMILA). Available online: https://clinicaltrials.gov/ct2/show/NCT04993664 (accessed on 12 October 2021).
  45. Trial of Aganirsen in iCRVO Patients at Risk of Developing NVG (STRONG). Available online: https://clinicaltrials.gov/ct2/show/NCT02947867 (accessed on 12 October 2021).
  46. Turnbull, I.C.; Eltoukhy, A.A.; Fish, K.M.; Hajjar, R.J.; Anderson, D.G.; Costa, K.D. Abstract 18088: Myocardial Delivery of Lipidoid Nanoparticle mRNA Designed for Tailored Expression of Cardiogenic Factors. Circulation 2013, 128, A18088. [Google Scholar] [CrossRef]
  47. Conway, A.; Mendel, M.; Kim, K.; McGovern, K.; Boyko, A.; Zhang, L.; Miller, J.C.; DeKelver, R.C.; Paschon, D.E.; Mui, B.L.; et al. Non-viral Delivery of Zinc Finger Nuclease mRNA Enables Highly Efficient In Vivo Genome Editing of Multiple Therapeutic Gene Targets. Mol. Ther. 2019, 27, 866–877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Li, Y.; Zhou, G.; Bruno, I.G.; Cooke, J.P. Telomerase mRNA Reverses Senescence in Progeria Cells. J. Am. Coll. Cardiol. 2017, 70, 804–805. [Google Scholar] [CrossRef] [PubMed]
  49. Chanda, P.; Sukhovershin, R.; Cooke, J. mRNA-Enhanced Cell Therapy and Cardiovascular Regeneration. Cells 2021, 10, 187. [Google Scholar] [CrossRef] [PubMed]
  50. Safety and Tolerability of Patisiran (ALN-TTR02) in Transthyretin (TTR) Amyloidosis. Available online: https://clinicaltrials.gov/ct2/show/NCT01617967 (accessed on 12 October 2021).
  51. Trial to Evaluate the Effect of ALN-PCSSC Treatment on Low Density Lipoprotein Cholesterol (LDL-C) (ORION-1). Available online: https://clinicaltrials.gov/ct2/show/NCT02597127 (accessed on 12 October 2021).
  52. Sohi, A.N.; Kiani, J.; Arefian, E.; Khosrojerdi, A.; Fekrirad, Z.; Ghaemi, S.; Zim, M.K.; Jalili, A.; Bostanshirin, N.; Soleimani, M. Development of an mRNA-LNP Vaccine against SARS-CoV-2: Evaluation of Immune Response in Mouse and Rhesus Macaque. Vaccines 2021, 9, 1007. [Google Scholar] [CrossRef]
  53. Atiroğlu, V.; Atiroğlu, A.; Özsoy, M.; Özacar, M. Coronavirus Disease (COVID-19), Chemical Structure, Therapeutics, Drugs and Vaccines. Biointerface Res. Appl. Chem. 2021, 12, 547–566. [Google Scholar]
  54. Tyagi, P.K.; Tyagi, S.; Kumar, A.; Ahuja, A.; Gola, D. Contribution of nanotechnology in the fight against COVID-19. Biointerface Res. Appl. Chem. 2020, 11, 8233–8241. [Google Scholar]
  55. Le, T.K.; Paris, C.; Khan, K.S.; Robson, F.; Ng, W.-L.; Rocchi, P. Nucleic Acid-Based Technologies Targeting Coronaviruses. Trends Biochem. Sci. 2021, 46, 351–365. [Google Scholar] [CrossRef] [PubMed]
  56. Monajjemi, M.; Mollaamin, F.; Shekarabi, A.S.; Ghadami, A. Variety of spike protein in covid-19 mutation: Stability, effectiveness and outbreak rate as a target for vaccine and therapeutic development. Biointerface Res. Appl. 2021, 11, 10016–10026. [Google Scholar]
  57. Bouyahya, A.; El Omari, N.; Elmenyiy, N.; Hakkour, M.; Balahbib, A.; Guaouguaou, E.; Benali, T.; El Baaboua, A.; Belmehdi, O. Therapeutic Strategies of COVID-19: From Natural Compounds to Vaccine Trials. Biointerface Res. Appl. Chem. 2020, 11, 1. [Google Scholar]
  58. Barda, N.; Dagan, N.; Ben-Shlomo, Y.; Kepten, E.; Waxman, J.; Ohana, R.; Hernán, M.A.; Lipsitch, M.; Kohane, I.; Netzer, D.; et al. Safety of the BNT162b2 mRNA Covid-19 Vaccine in a Nationwide Setting. N. Engl. J. Med. 2021, 385, 1078–1090. [Google Scholar] [CrossRef]
  59. Buschmann, M.; Carrasco, M.; Alishetty, S.; Paige, M.; Alameh, M.; Weissman, D. Nanomaterial Delivery Systems for mRNA Vaccines. Vaccines 2021, 9, 65. [Google Scholar] [CrossRef] [PubMed]
  60. ElBagoury, M.; Tolba, M.M.; Nasser, H.A.; Jabbar, A.; Elagouz, A.M.; Aktham, Y.; Hutchinson, A. The find of COVID-19 vaccine: Challenges and opportunities. J. Infect. Public Health 2021, 14, 389–416. [Google Scholar] [CrossRef]
  61. Chaudhary, N.; Weissman, D.; Whitehead, K.A. mRNA vaccines for infectious diseases: Principles, delivery and clinical translation. Nat. Rev. Drug Discov. 2021, 20, 817–838. [Google Scholar] [CrossRef]
  62. Kowalski, P.; Rudra, A.; Miao, L.; Anderson, D.G. Delivering the Messenger: Advances in Technologies for Therapeutic mRNA Delivery. Mol. Ther. 2019, 27, 710–728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Kulkarni, J.; Cullis, P.R.; van der Meel, R. Lipid Nanoparticles Enabling Gene Therapies: From Concepts to Clinical Utility. Nucleic Acid Ther. 2018, 28, 146–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Alameh, M.-G.; Weissman, D.; Pardi, N. Messenger RNA-Based Vaccines against Infectious Diseases. In Current Topics in Microbiology and Immunology; Springer: Berlin/Heidelberg, Germany, 2020; pp. 1–35. [Google Scholar]
  65. Hadi, A.G.; Kadhom, M.; Hairunisa, N.; Yousif, E.; Mohammed, S.A. A Review on COVID-19: Origin, Spread, Symptoms, Treatment, and Prevention. Biointerface Res. Appl. Chem. 2020, 10, 7234–7242. [Google Scholar] [CrossRef]
  66. Monajjemi, M.; Shahriari, S.; Mollaamin, F. Evaluation of Coronavirus Families & Covid-19 Proteins: Molecular Modeling Study. Biointerface Res. Appl. Chem. 2020, 10, 6039–6057. [Google Scholar] [CrossRef]
  67. Romani, D.; Noureddine, O.; Issaoui, N.; Brandan, S.A. Properties and Reactivities of Niclosamide in Different Media, a Potential Antiviral to Treatment of COVID-19 by Using DFT Calculations and Molecular Docking. Biointerface Res. Appl. Chem. 2020, 10, 7295–7328. [Google Scholar] [CrossRef]
  68. Safety, Tolerability, and Immunogenicity of mRNA-1325 in Healthy Adult Subjects. Available online: https://clinicaltrials.gov/ct2/show/NCT03014089 (accessed on 15 October 2021).
  69. ModernaTx, I. Safety, Tolerability, and Immunogenicity of VAL-506440 in Healthy Adult Subjects. Available online: https://clinicaltrials.gov/ct2/show/NCT03076385 (accessed on 15 October 2021).
  70. ModernaTx, I. Safety, Tolerability, and Immunogenicity of VAL-339851 in Healthy Adult Subjects. Available online: https://clinicaltrials.gov/ct2/show/NCT03345043 (accessed on 15 October 2021).
  71. Scarpini, S.; Morigi, F.; Betti, L.; Dondi, A.; Biagi, C.; Lanari, M. Development of a Vaccine against Human Cytomegalovirus: Advances, Barriers, and Implications for the Clinical Practice. Vaccines 2021, 9, 551. [Google Scholar] [CrossRef] [PubMed]
  72. Safety, Reactogenicity, and Immunogenicity of Cytomegalovirus Vaccines mRNA-1647 and mRNA-1443 in Healthy Adults. Available online: https://clinicaltrials.gov/ct2/show/NCT03382405 (accessed on 15 October 2021).
  73. Safety, Tolerability, and Immunogenicity of VAL-181388 in Healthy Subjects. Available online: https://clinicaltrials.gov/ct2/show/NCT03325075 (accessed on 15 October 2021).
  74. Flisiak, R.; Jaroszewicz, J.; Łucejko, M. siRNA drug development against hepatitis B virus infection. Expert Opin. Biol. Ther. 2018, 18, 609–617. [Google Scholar] [CrossRef] [PubMed]
  75. Thi, E.P.; Dhillon, A.P.; Ardzinski, A.; Bidirici-Ertekin, L.; Cobarrubias, K.D.; Cuconati, A.; Kondratowicz, A.S.; Kwak, K.; Li, A.H.L.; Miller, A.; et al. ARB-1740, a RNA Interference Therapeutic for Chronic Hepatitis B Infection. ACS Infect. Dis. 2019, 5, 725–737. [Google Scholar] [CrossRef]
  76. Ye, X.; Tateno, C.; Thi, E.P.; Kakuni, M.; Snead, N.M.; Ishida, Y.; Barnard, T.R.; Sofia, M.J.; Shimada, T.; Lee, A.C.H. Hepatitis B Virus Therapeutic Agent ARB-1740 Has Inhibitory Effect on Hepatitis Delta Virus in a New Dually-Infected Humanized Mouse Model. ACS Infect. Dis. 2019, 5, 738–749. [Google Scholar] [CrossRef] [PubMed]
  77. Dallas, A.; Ilves, H.; Shorenstein, J.; Judge, A.; Spitler, R.; Contag, C.; Wong, S.P.; Harbottle, R.P.; MacLachlan, I.; Johnston, B.H. Minimal-length Synthetic shRNAs Formulated with Lipid Nanoparticles are Potent Inhibitors of Hepatitis C Virus IRES-linked Gene Expression in Mice. Mol. Ther. Nucleic Acids 2013, 2, e123. [Google Scholar] [CrossRef] [PubMed]
  78. Chandra, P.K.; Kundu, A.K.; Hazari, S.; Chandra, S.; Bao, L.; Ooms, T.; Morris, G.F.; Wu, T.; Mandal, T.; Dash, S. Inhibition of Hepatitis C Virus Replication by Intracellular Delivery of Multiple siRNAs by Nanosomes. Mol. Ther. 2012, 20, 1724–1736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Moon, J.-S.; Lee, S.-H.; Kim, E.-J.; Cho, H.; Lee, W.; Kim, G.-W.; Park, H.-J.; Cho, S.-W.; Lee, C.; Oh, J.-W. Inhibition of Hepatitis C Virus in Mice by a Small Interfering RNA Targeting a Highly Conserved Sequence in Viral IRES Pseudoknot. PLoS ONE 2016, 11, e0146710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Melo, M.; Porter, E.; Zhang, Y.; Silva, M.; Li, N.; Dobosh, B.; Liguori, A.; Skog, P.; Landais, E.; Menis, S.; et al. Immunogenicity of RNA Replicons Encoding HIV Env Immunogens Designed for Self-Assembly into Nanoparticles. Mol. Ther. 2019, 27, 2080–2090. [Google Scholar] [CrossRef] [PubMed]
  81. Saunders, K.O.; Pardi, N.; Parks, R.; Santra, S.; Mu, Z.; Sutherland, L.; Scearce, R.; Barr, M.; Eaton, A.; Hernandez, G.; et al. Lipid nanoparticle encapsulated nucleoside-modified mRNA vaccines elicit polyfunctional HIV-1 antibodies comparable to proteins in nonhuman primates. NPJ Vaccines 2021, 6, 1–14. [Google Scholar] [CrossRef] [PubMed]
  82. Egan, K.P.; Hook, L.M.; Naughton, A.; Pardi, N.; Awasthi, S.; Cohen, G.H.; Weissman, D.; Friedman, H.M. An HSV-2 nucleoside-modified mRNA genital herpes vaccine containing glycoproteins gC, gD, and gE protects mice against HSV-1 genital lesions and latent infection. PLoS Pathog. 2020, 16, e1008795. [Google Scholar] [CrossRef] [PubMed]
  83. Palanki, R.; Peranteau, W.H.; Mitchell, M.J. Delivery technologies for in utero gene therapy. Adv. Drug Deliv. Rev. 2021, 169, 51–62. [Google Scholar] [CrossRef]
  84. Riley, R.S.; Kashyap, M.V.; Billingsley, M.M.; White, B.; Alameh, M.-G.; Bose, S.K.; Zoltick, P.W.; Li, H.; Zhang, R.; Cheng, A.Y.; et al. Ionizable lipid nanoparticles for in utero mRNA delivery. Sci. Adv. 2021, 7, eaba1028. [Google Scholar] [CrossRef] [PubMed]
  85. Yu, X.; Liu, S.; Cheng, Q.; Wei, T.; Lee, S.; Zhang, D.; Siegwart, D.J. Lipid-Modified Aminoglycosides for mRNA Delivery to the Liver. Adv. Healthc. Mater. 2020, 9, 1901487. [Google Scholar] [CrossRef]
  86. Huang, X.; Chau, Y. Enhanced Delivery of siRNA to Retinal Ganglion Cells by Intravitreal Lipid Nanoparticles of Positive Charge. Mol. Pharm. 2020, 18, 377–385. [Google Scholar] [CrossRef] [PubMed]
  87. Hagino, Y.; Khalil, I.A.; Kimura, S.; Kusumoto, K.; Harashima, H. GALA-Modified Lipid Nanoparticles for the Targeted Delivery of Plasmid DNA to the Lungs. Mol. Pharm. 2021, 18, 878–888. [Google Scholar] [CrossRef]
  88. Turnbull, I.C.; A Eltoukhy, A.; Fish, K.M.; Nonnenmacher, M.; Ishikawa, K.; Chen, J.; Hajjar, R.J.; Anderson, D.G.; Costa, K.D. Myocardial Delivery of Lipidoid Nanoparticle Carrying modRNA Induces Rapid and Transient Expression. Mol. Ther. 2016, 24, 66–75. [Google Scholar] [CrossRef]
  89. Chen, C.-Y.; Tran, D.M.; Cavedon, A.; Cai, X.; Rajendran, R.; Lyle, M.J.; Martini, P.G.; Miao, C.H. Treatment of Hemophilia A Using Factor VIII Messenger RNA Lipid Nanoparticles. Mol. Ther. Nucleic Acids 2020, 20, 534–544. [Google Scholar] [CrossRef] [PubMed]
  90. Younis, M.A.; Khalil, I.A.; Elewa, Y.H.A.; Kon, Y.; Harashima, H. Ultra-small lipid nanoparticles encapsulating sorafenib and midkine-siRNA selectively-eradicate sorafenib-resistant hepatocellular carcinoma in vivo. J. Control. Release 2021, 331, 335–349. [Google Scholar] [CrossRef] [PubMed]
  91. Elia, U.; Ramishetti, S.; Rosenfeld, R.; Dammes, N.; Bar-Haim, E.; Naidu, G.S.; Makdasi, E.; Yahalom-Ronen, Y.; Tamir, H.; Paran, N. Design of SARS-CoV-2 hFc-conjugated receptor-binding domain mRNA vaccine delivered via lipid nanoparticles. ACS Nano 2021, 15, 9627–9637. [Google Scholar] [CrossRef] [PubMed]
  92. Atu027 Plus Gemcitabine in Advanced or Metastatic Pancreatic Cancer (Atu027-I-02) (Atu027-I-02). Available online: https://clinicaltrials.gov/ct2/show/record/NCT01808638 (accessed on 23 October 2021).
  93. Study With Atu027 in Patients With Advanced Solid Cancer. Available online: https://clinicaltrials.gov/ct2/show/record/NCT00938574 (accessed on 23 October 2021).
  94. Mainini, F.; Eccles, M.R. Lipid and Polymer-Based Nanoparticle siRNA Delivery Systems for Cancer Therapy. Molecules 2020, 25, 25. [Google Scholar] [CrossRef]
  95. Sarker, D.; Plummer, R.; Meyer, T.; Sodergren, M.H.; Basu, B.; Chee, C.E.; Huang, K.-W.; Palmer, D.H.; Ma, Y.T.; Evans, T.J.; et al. MTL-CEBPA, a Small Activating RNA Therapeutic Upregulating C/EBP-α, in Patients with Advanced Liver Cancer: A First-in-Human, Multicenter, Open-Label, Phase I Trial. Clin. Cancer Res. 2020, 26, 3936–3946. [Google Scholar] [CrossRef] [PubMed]
  96. First-in-Human Safety, Tolerability and Antitumour Activity Study of MTL-CEBPA in Patients With Advanced Liver Cancer (OUTREACH). Available online: https://www.clinicaltrials.gov/ct2/show/NCT02716012 (accessed on 23 October 2021).
  97. EphA2 siRNA in Treating Patients With Advanced or Recurrent Solid Tumors. Available online: https://clinicaltrials.gov/ct2/show/NCT01591356 (accessed on 23 October 2021).
  98. Phase, I. Multicenter, Dose Escalation Study of DCR-MYC in Patients with Solid Tumors, Multiple Myeloma, or Lymphoma. Available online: https://clinicaltrials.gov/ct2/show/NCT02110563 (accessed on 20 October 2021).
  99. Dose Escalation and Efficacy Study of mRNA-2416 for Intratumoral Injection Alone and in Combination with Durvalumab for Participants With Advanced Malignancies. Available online: https://clinicaltrials.gov/ct2/show/NCT03323398 (accessed on 20 October 2021).
  100. A Study of SGT-53 in Children With Refractory or Recurrent Solid Tumors. Available online: https://clinicaltrials.gov/ct2/show/NCT02354547 (accessed on 20 October 2021).
  101. Tang, W.; Lalezari, J.; June, C.; Tebas, P.; Lee, G.; Giedlin, M.; Wang, S.; Stein, D.; Hatano, H.; Sinclair, E. Results of a Phase I Trial of SGT-53: A Systemically Administered, Tumor-Targeting Immunoliposome Nanocomplex Incorporating a Plasmid Encoding wtp53. Mol. Ther. 2012, 20, 1. [Google Scholar]
  102. Study of Combined SGT-53 Plus Gemcitabine/Nab-Paclitaxel for Metastatic Pancreatic Cancer. Available online: https://clinicaltrials.gov/ct2/show/NCT02340117 (accessed on 20 October 2021).
  103. JVRS-100 for the Treatment of Patients With Relapsed or Refractory Leukemia. Available online: https://clinicaltrials.gov/ct2/show/NCT00860522 (accessed on 20 October 2021).
  104. Safety and Immunogenicity Study of 2019-nCoV Vaccine (mRNA-1273) for Prophylaxis of SARS-CoV-2 Infection (COVID-19). Available online: https://clinicaltrials.gov/ct2/show/NCT04283461 (accessed on 10 October 2021).
  105. Schoenmaker, L.; Witzigmann, D.; Kulkarni, J.A.; Verbeke, R.; Kersten, G.; Jiskoot, W.; Crommelin, D.J. mRNA-lipid nanoparticle COVID-19 vaccines: Structure and stability. Int. J. Pharm. 2021, 601, 120586. [Google Scholar] [CrossRef]
  106. Rodríguez, M.J. Study About the Response to the Administration of a Third Dose of mRNA-1273 Vaccine (COVID-19 Vaccine Moderna) in Renal Transplants with Immunological Failure Initial to Vaccination. 2021. Available online: https://clinicaltrials.gov/ct2/show/NCT04930770 (accessed on 10 October 2021).
  107. A Phase III Clinical Study of a SARS-CoV-2 Messenger Ribonucleic Acid (mRNA) Vaccine Candidate Against COVID-19 in Population Aged 18 Years and Above. Available online: https://clinicaltrials.gov/ct2/show/study/NCT04847102 (accessed on 10 October 2021).
  108. COVID-19 Vaccination of Immunodeficient Persons (COVAXID). Available online: https://clinicaltrials.gov/ct2/show/NCT04780659 (accessed on 10 October 2021).
  109. Third Dose of mRNA Vaccination to Boost COVID-19 Immunity. Available online: https://clinicaltrials.gov/ct2/show/NCT05057182 (accessed on 10 October 2021).
  110. Safety and Immunogenicity Study of a SARS-CoV-2 (COVID-19) Variant Vaccine (mRNA-1273.351) in Naïve and Previously Vaccinated Adults. Available online: https://clinicaltrials.gov/ct2/show/NCT04785144 (accessed on 10 October 2021).
  111. Wadhwa, A.; Aljabbari, A.; Lokras, A.; Foged, C.; Thakur, A. Opportunities and challenges in the delivery of mRNA-based vaccines. Pharmaceutics 2020, 12, 102. [Google Scholar] [CrossRef] [Green Version]
  112. Kuche, K.; Pandey, P.K.; Patharkar, A.; Maheshwari, R.; Tekade, R.K. Chapter 12-Hyaluronic Acid as an Emerging Technology Platform for Silencing RNA Delivery. In Biomaterials and Bionanotechnology; Tekade, R.K., Ed.; Academic Press: Cambridge, MA, USA, 2019; pp. 415–458. [Google Scholar] [CrossRef]
  113. Serrano-Sevilla, I.; Artiga, Á.; Mitchell, S.G.; De Matteis, L.; de la Fuente, J.M. Natural Polysaccharides for siRNA Delivery: Nanocarriers Based on Chitosan, Hyaluronic Acid, and Their Derivatives. Molecules 2019, 24, 2570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Paidikondala, M.; Rangasami, V.K.; Nawale, G.N.; Casalini, T.; Perale, G.; Kadekar, S.; Mohanty, G.; Salminen, T.; Oommen, O.P.; Varghese, O.P. An Unexpected Role of Hyaluronic Acid in Trafficking siRNA Across the Cellular Barrier: The First Biomimetic, Anionic, Non-Viral Transfection Method. Angew. Chem. Int. Ed. 2019, 58, 2815–2819. [Google Scholar] [CrossRef] [PubMed]
  115. Lü, J.-M.; Liang, Z.; Liu, D.; Zhan, B.; Yao, Q.; Chen, C. Two Antibody-Guided Lactic-co-Glycolic Acid-Polyethylenimine (LGA-PEI) Nanoparticle Delivery Systems for Therapeutic Nucleic Acids. Pharmaceuticals 2021, 14, 841. [Google Scholar] [CrossRef] [PubMed]
  116. Koh, K.J.; Liu, Y.; Lim, S.H.; Loh, X.J.; Kang, L.; Lim, C.Y.; Phua, K.K.L. Formulation, characterization and evaluation of mRNA-loaded dissolvable polymeric microneedles (RNApatch). Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Pan, J.; Ruan, W.; Qin, M.; Long, Y.; Wan, T.; Yu, K.; Zhai, Y.; Wu, C.; Xu, Y. Intradermal delivery of STAT3 siRNA to treat melanoma via dissolving microneedles. Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Chahal, J.S.; Fang, T.; Woodham, A.W.; Khan, O.; Ling, J.; Anderson, D.G.; Ploegh, H.L. An RNA nanoparticle vaccine against Zika virus elicits antibody and CD8+ T cell responses in a mouse model. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef]
  119. Chahal, J.S.; Khan, O.F.; Cooper, C.L.; McPartlan, J.S.; Tsosie, J.K.; Tilley, L.D.; Sidik, S.M.; Lourido, S.; Langer, R.; Bavari, S.; et al. Dendrimer-RNA nanoparticles generate protective immunity against lethal Ebola, H1N1 influenza, and Toxoplasma gondii challenges with a single dose. Proc. Natl. Acad. Sci. USA 2016, 113, E4133–E4142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Huang, C.-L.; Leblond, A.-L.; Turner, E.C.; Kumar, A.; Martin, K.; Whelan, D.; O’Sullivan, D.M.; Caplice, N.M. Synthetic Chemically Modified mRNA-Based Delivery of Cytoprotective Factor Promotes Early Cardiomyocyte Survival Post-Acute Myocardial Infarction. Mol. Pharm. 2015, 12, 991–996. [Google Scholar] [CrossRef] [PubMed]
  121. Ribas, A.; Tolcher, A.W.; Yen, Y. Safety study of CALAA-01 to treat solid tumor cancers. NIH USA 2008. Available online: https://www.clinicaltrials.gov/ct2/show/NCT00689065 (accessed on 20 October 2021).
  122. Varghese, A.M.; Ang, C.; DiMaio, C.J.; Javle, M.M.; Gutierrez, M.; Yarom, N.; Stemmer, S.M.; Golan, T.; Geva, R.; Semenisty, V.; et al. A phase II study of siG12D-LODER in combination with chemotherapy in patients with locally advanced pancreatic cancer (PROTACT). J. Clin. Oncol. 2020, 38, TPS4672. [Google Scholar] [CrossRef]
  123. Sanghani, A.; Kafetzis, K.N.; Sato, Y.; Elboraie, S.; Fajardo-Sanchez, J.; Harashima, H.; Tagalakis, A.D.; Yu-Wai-Man, C. Novel PEGylated lipid nanoparticles have a high encapsulation efficiency and effectively deliver MRTF-B siRNA in conjunctival fibroblasts. Pharmaceutics 2021, 13, 382. [Google Scholar] [CrossRef]
  124. Siewert, C.D.; Haas, H.; Cornet, V.; Nogueira, S.S.; Nawroth, T.; Uebbing, L.; Ziller, A.; Al-Gousous, J.; Radulescu, A.; Schroer, M.A.; et al. Hybrid Biopolymer and Lipid Nanoparticles with Improved Transfection Efficacy for mRNA. Cells 2020, 9, 2034. [Google Scholar] [CrossRef] [PubMed]
  125. Xiong, H.; Liu, S.; Wei, T.; Cheng, Q.; Siegwart, D.J. Theranostic dendrimer-based lipid nanoparticles containing PEGylated BODIPY dyes for tumor imaging and systemic mRNA delivery in vivo. J. Control. Release 2020, 325, 198–205. [Google Scholar] [CrossRef]
  126. Zhou, K.; Johnson, L.; Xiong, H.; Barrios, S.; Minnig, J.T.; Yan, Y.; Abram, B.; Yu, X.; Siegwart, D.J. Hydrophobic Domain Structure of Linear-Dendritic Poly(ethylene glycol) Lipids Affects RNA Delivery of Lipid Nanoparticles. Mol. Pharm. 2020, 17, 1575–1585. [Google Scholar] [CrossRef] [PubMed]
  127. Gabal, Y.M.; Kamel, A.O.; Sammour, O.A.; Elshafeey, A.H. Effect of surface charge on the brain delivery of nanostructured lipid carriers in situ gels via the nasal route. Int. J. Pharm. 2014, 473, 442–457. [Google Scholar] [CrossRef] [PubMed]
  128. Tagalakis, A.D.; Kenny, G.D.; Bienemann, A.S.; McCarthy, D.; Munye, M.M.; Taylor, H.; Wyatt, M.J.; Lythgoe, M.F.; White, E.A.; Hart, S.L. PEGylation improves the receptor-mediated transfection efficiency of peptide-targeted, self-assembling, anionic nanocomplexes. J. Control. Release 2014, 174, 177–187. [Google Scholar] [CrossRef] [Green Version]
  129. Tagalakis, A.D.; Jayarajan, V.; Maeshima, R.; Ho, K.H.; Syed, F.; Wu, L.; Aldossary, A.M.; Munye, M.M.; Mistry, T.; Ogunbiyi, O.K.; et al. Integrin-Targeted, Short Interfering RNA Nanocomplexes for Neuroblastoma Tumor-Specific Delivery Achieve MYCN Silencing with Improved Survival. Adv. Funct. Mater. 2021, 31, 2104843. [Google Scholar] [CrossRef]
  130. Yang, Q.; Zhou, Y.; Chen, J.; Huang, N.; Wang, Z.; Cheng, Y. Gene Therapy for Drug-Resistant Glioblastoma via Lipid-Polymer Hybrid Nanoparticles Combined with Focused Ultrasound. Int. J. Nanomed. 2021, 16, 185–199. [Google Scholar] [CrossRef] [PubMed]
  131. Perche, F.; Clemençon, R.; Schulze, K.; Ebensen, T.; Guzmán, C.A.; Pichon, C. Neutral Lipopolyplexes for In Vivo Delivery of Conventional and Replicative RNA Vaccine. Mol. Ther. Nucleic Acids 2019, 17, 767–775. [Google Scholar] [CrossRef] [PubMed]
  132. Paris, J.L.; Coelho, F.; Teixeira, A.; Diéguez, L.; Silva, B.F.B.; Abalde-Cela, S. In Vitro Evaluation of Lipopolyplexes for Gene Transfection: Comparing 2D, 3D and Microdroplet-Enabled Cell Culture. Molecules 2020, 25, 3277. [Google Scholar] [CrossRef] [PubMed]
  133. Klein, P.M.; Wagner, E. Click-Shielded and Targeted Lipopolyplexes. Methods Mol. Biol. 2019, 2036, 141–164. [Google Scholar] [CrossRef]
  134. Wang, L.-L.; Feng, C.-L.; Zheng, W.-S.; Huang, S.; Zhang, W.; Wu, H.-N.; Zhan, Y.; Han, Y.-X.; Wu, S.; Jiang, J.-D. Tumor-selective lipopolyplex encapsulated small active RNA hampers colorectal cancer growth in vitro and in orthotopic murine. Biomaterials 2017, 141, 13–28. [Google Scholar] [CrossRef]
  135. Shah, H.; Tariq, I.; Engelhardt, K.; Bakowsky, U.; Pinnapireddy, S.R. Development and Characterization of Ultrasound Activated Lipopolyplexes for Enhanced Transfection by Low Frequency Ultrasound in In Vitro Tumor Model. Macromol. Biosci. 2020, 20, e2000173. [Google Scholar] [CrossRef] [PubMed]
  136. Tariq, I.; Ali, M.Y.; Sohail, M.F.; Amin, M.U.; Ali, S.; Bukhari, N.I.; Raza, A.; Pinnapireddy, S.R.; Schäfer, J.; Bakowsky, U. Lipodendriplexes mediated enhanced gene delivery: A cellular to pre-clinical investigation. Sci. Rep. 2020, 10, 1–15. [Google Scholar] [CrossRef] [PubMed]
  137. Kim, M.; Jeong, M.; Hur, S.; Cho, Y.; Park, J.; Jung, H.; Seo, Y.; Woo, H.A.; Nam, K.T.; Lee, K. Engineered ionizable lipid nanoparticles for targeted delivery of RNA therapeutics into different types of cells in the liver. Sci. Adv. 2021, 7, eabf4398. [Google Scholar] [CrossRef] [PubMed]
  138. Sato, Y.; Hatakeyama, H.; Sakurai, Y.; Hyodo, M.; Akita, H.; Harashima, H. A pH-sensitive cationic lipid facilitates the delivery of liposomal siRNA and gene silencing activity in vitro and in vivo. J. Control. Release 2012, 163, 267–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Sakurai, Y.; Mizumura, W.; Ito, K.; Iwasaki, K.; Katoh, T.; Goto, Y.; Suga, H.; Harashima, H. Improved Stability of siRNA-Loaded Lipid Nanoparticles Prepared with a PEG-Monoacyl Fatty Acid Facilitates Ligand-Mediated siRNA Delivery. Mol. Pharm. 2020, 17, 1397–1404. [Google Scholar] [CrossRef]
  140. Herrera, M.; Kim, J.; Eygeris, Y.; Jozic, A.; Sahay, G. Illuminating endosomal escape of polymorphic lipid nanoparticles that boost mRNA delivery. Biomater. Sci. 2021, 9, 4289–4300. [Google Scholar] [CrossRef] [PubMed]
  141. Mihaila, R.; Ruhela, D.; Keough, E.; Cherkaev, E.; Chang, S.; Galinski, B.; Bartz, R.; Brown, D.; Howell, B.; Cunningham, J.J. Mathematical Modeling: A Tool for Optimization of Lipid Nanoparticle-Mediated Delivery of siRNA. Mol. Ther. Nucleic Acids 2017, 7, 246–255. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Saunders, N.R.M.; Paolini, M.S.; Fenton, O.S.; Poul, L.; Devalliere, J.; Mpambani, F.; Darmon, A.; Bergere, M.; Jibault, O.; Germain, M.; et al. A Nanoprimer To Improve the Systemic Delivery of siRNA and mRNA. Nano Lett. 2020, 20, 4264–4269. [Google Scholar] [CrossRef]
  143. Francia, V.; Schiffelers, R.M.; Cullis, P.R.; Witzigmann, D. The Biomolecular Corona of Lipid Nanoparticles for Gene Therapy. Bioconjug. Chem. 2020, 31, 2046–2059. [Google Scholar] [CrossRef] [PubMed]
  144. Chen, D.; Ganesh, S.; Wang, W.; Amiji, M. The role of surface chemistry in serum protein corona-mediated cellular delivery and gene silencing with lipid nanoparticles. Nanoscale 2019, 11, 8760–8775. [Google Scholar] [CrossRef]
  145. Lorents, A.; Maloverjan, M.; Padari, K.; Pooga, M. Internalisation and Biological Activity of Nucleic Acids Delivering Cell-Penetrating Peptide Nanoparticles Is Controlled by the Biomolecular Corona. Pharmaceuticals 2021, 14, 667. [Google Scholar] [CrossRef] [PubMed]
  146. Moraru, C.; Mincea, M.; Menghiu, G.; Ostafe, V. Understanding the Factors Influencing Chitosan-Based Nanoparticles-Protein Corona Interaction and Drug Delivery Applications. Molecules 2020, 25, 4758. [Google Scholar] [CrossRef] [PubMed]
  147. Yang, H.; Liu, T.; Xu, Y.; Su, G.; Liu, T.; Yu, Y.; Xu, B. Protein corona precoating on redox-responsive chitosan-based nano-carriers for improving the therapeutic effect of nucleic acid drugs. Carbohydr. Polym. 2021, 265, 118071. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Most important benchmarks in the development of lipid-based delivery systems for RNA. Adapted from [2,11,16,24], published by Nature Publishing Group, 2021; Elsevier, 2020; Nature Publishing Group, 2021; https://www.sigmaaldrich.com/deepweb/assets/sigmaaldrich/product/documents/376/366/safc-lipids-rna-wp7004en-ms.pdf, accessed on 10 October 2021.
Figure 1. Most important benchmarks in the development of lipid-based delivery systems for RNA. Adapted from [2,11,16,24], published by Nature Publishing Group, 2021; Elsevier, 2020; Nature Publishing Group, 2021; https://www.sigmaaldrich.com/deepweb/assets/sigmaaldrich/product/documents/376/366/safc-lipids-rna-wp7004en-ms.pdf, accessed on 10 October 2021.
Pharmaceutics 13 02053 g001
Figure 2. Visual representation of several conventional synthesis methods for LNP-based delivery systems. (a) hot high-pressure homogenization method; (b) cold high-pressure homogenization method; (c) solvent evaporation method; (d) microemulsion method. Adapted from [31], published by Uppsala University, 2020.
Figure 2. Visual representation of several conventional synthesis methods for LNP-based delivery systems. (a) hot high-pressure homogenization method; (b) cold high-pressure homogenization method; (c) solvent evaporation method; (d) microemulsion method. Adapted from [31], published by Uppsala University, 2020.
Pharmaceutics 13 02053 g002
Figure 3. Schematic representation of LNP-based delivery systems manufacturing through microfluidic methods. Adapted from [33], published by MDPI, 2020.
Figure 3. Schematic representation of LNP-based delivery systems manufacturing through microfluidic methods. Adapted from [33], published by MDPI, 2020.
Pharmaceutics 13 02053 g003
Figure 4. Cross-sectional model of SARS-CoV-2. Adapted from [60], published by AMER PUBLIC HEALTH ASSOC INC, 2021.
Figure 4. Cross-sectional model of SARS-CoV-2. Adapted from [60], published by AMER PUBLIC HEALTH ASSOC INC, 2021.
Pharmaceutics 13 02053 g004
Figure 5. Schematic representation of lipopolyplexes formation. Adapted from [131], published by Cell Press, 2019.
Figure 5. Schematic representation of lipopolyplexes formation. Adapted from [131], published by Cell Press, 2019.
Pharmaceutics 13 02053 g005
Figure 6. Schematic representation of lipodendriplexes formation and cellular internalization. (1) Endocytosis and pDNA release into the cytoplasm. (2) Transcription of gene-encoded DNA into mRNA. (3) mRNA export from the nucleus to the cytoplasm. (4) Protein expression. Adapted from [136], published by Nature, 2020.
Figure 6. Schematic representation of lipodendriplexes formation and cellular internalization. (1) Endocytosis and pDNA release into the cytoplasm. (2) Transcription of gene-encoded DNA into mRNA. (3) mRNA export from the nucleus to the cytoplasm. (4) Protein expression. Adapted from [136], published by Nature, 2020.
Pharmaceutics 13 02053 g006
Figure 7. (a) Initiation of formation of protein corona (PC) (seconds after the NP reaches the biological fluid); (b) beginning of exchange from the PC of proteins with low affinity with proteins that have higher affinity (seconds to minutes); (c) stabilized PC, with proteins with high affinity occupying the first layer of PC (hard PC) and the majority of the second layer (soft PC) where proteins with low affinity are still present. Adapted from [146], published by MDPI, 2020.
Figure 7. (a) Initiation of formation of protein corona (PC) (seconds after the NP reaches the biological fluid); (b) beginning of exchange from the PC of proteins with low affinity with proteins that have higher affinity (seconds to minutes); (c) stabilized PC, with proteins with high affinity occupying the first layer of PC (hard PC) and the majority of the second layer (soft PC) where proteins with low affinity are still present. Adapted from [146], published by MDPI, 2020.
Pharmaceutics 13 02053 g007
Table 1. A summary of lipid-based nucleic acids delivery systems applications.
Table 1. A summary of lipid-based nucleic acids delivery systems applications.
Delivery SystemPhysicochemical PropertiesDisease/ConditionTesting StageAdministration RouteObservationsRef.
LNP-encapsulated C-24 alkyl phytosterols carrying mRNAShape: polyhedral
Hydrodynamic diameter: ~100 nm
Encapsulation efficiency: >90%
Lipid composition: ionizable lipid (DLin-MC3-DMA, Lipid 9, or DODMA):sterol:DSPC:DMG-PEG2k at molar ratios of 50:38.5:10:1.5
CancerIn vitro-High encapsulation efficiency
High cellular uptake and retention
Enhanced intracellular diffusivity
Enhanced gene transfection
[32]
Aerosolizable siRNA-encapsulated SLNsSize: 164.5 ± 28.3 nm
Polydispersity index: 0.31 ± 0.09
Zeta potential: −29.1 ± 8.6 mV
Lipid composition: lecithin, cholesterol, and lipid-PEG conjugate
Lung diseases
(e.g., cystic fibrosis)
In vitro-SLNs diffused through the simulated mucus layer
Possibility of down-regulating the expression of the gene of interest (e.g., TNF-α siRNA) by alveolar macrophages of lung epithelial cells
Promising for lung delivery by inhalation
[29]
mRNA encapsulated cationic lipid-modified aminoglycosides-based LNPsSize: 100–200 nm
Surface charge: ~0 mV
Encapsulation efficiency: >95%
Lipid composition: GT-EP10, DOPE, cholesterol, DMG-PEG2k
Liver diseasesIn vitro
and
In vivo (tested on mice)
IntravenousEfficient mRNA delivery to the liver
Safe delivery system (no obvious toxicity, changes in ALT or AST, or liver injury in histological sections)
Potential for application in gene editing and delivery of TNA
[85]
siRNA encapsulated LNPs with different surface chargesShape: spheres with filled cores
Size: ~7- nm
Zeta potential: ranged between 0 and 30 mV
Encapsulation efficiency: >99%
Lipid composition: lipidoid, DOTAP, DOPE, cholesterol, DSPE-PEG2k
Retinal diseases (e.g., age-related macular degeneration, glaucoma)In vitro
and
In vivo (tested on mice)
IntravitrealSuccessfully managed gene knockdown in mammalian cell line and primary neurons
No gene silencing facilitation in the retina pigmented epithelium layer
48 h after injection, neutral and mildly positive LNPs mediated limited retinal gene suppression (<10%), whereas more positive LNPs led to approx. 25% gene suppression in the retinal ganglion cell layer
[86]
GALA-modified LNP-encapsulated pDNASize: 125–155 nm
Zeta potential: ~15 mV (DOTMA-shell particles); ~−15 mV (YSK05-shell particles)
Composition: inner coat—DOPE; STR-R8; outer coat—DOTMA/YSK05, cholesterol
Lung diseasesIn vivo
(tested on mice)
IntravenousEfficient lung-selective delivery systemHigh gene expression level and significantly improved transfection activity in the lungs
Improved efficiency of gene expression at the intracellular level due to the double-coating
[87]
Formulated lipidoid nanoparticles (FLNP)-encapsulated modified mRNASize: ~155 nm
Lipid composition: epoxide-derived lipidoid complex (C14-113)
Cardiovascular diseaseIn vivo (tested on rats and pigs)Intramyocardial injectionHighly efficient, rapid, and short-term mRNA expression in the heart
Over 60-fold higher mRNA levels in heart tissue than for naked TNA
Limited off-target biodistribution
[88]
LNP-encapsulated mRNA encoding short-lived factor VIII (FVIII) proteinSize: <100 nm
RNA encapsulation: >80%
Lipid composition: ionizable lipid:DSPC:cholesterol:PEG lipid, at molar ratios of 50:10:38.5:1.5
Hemophilia AIn vivo
(tested on mice)
IntravenousSafe and effective delivery platform
Produced rapid and prolonged duration of FVIII expression
Easy to improve FVIII function by modification of mRNA sequence encoding newly developed FVIII protein with higher activity or longer half-lifeIt could be applied to prophylactic treatment and potentially various other treatment options
[89]
Ultra-small LNPs encapsulating sorafenib and midkine-siRNASize: 60.47 ± 6.9 nm
Zeta potential: −17.4 ± 5 mV
Sorafenib encapsulation efficiency: 96.5 ± 4.8%
RNA encapsulation efficiency: 94.5 ± 6.5%
Lipid composition: YSK05:DOPE:cholesterol:PEG-SP94:PEG, at molar ratios of 5:2:3:1:0.3
Sorafenib-resistant hepatocellular carcinomaIn vivo
(tested on mice)
IntravenousEnhanced tumor accumulation, selectivity, and in vivo gene silencing
Biosafe treatment approach
Synergistic action allowed the eradication of hepatocellular carcinoma at a low drug dose (2.5 mg/kg), which is unattainable with individual monotherapy
[90]
LNP-encapsulated SARS-CoV-2 human Fc-conjugated receptor-binding domain mRNASize: ~100 nm
Polydispersity index: <0.3
Lipid composition: ionizable lipid, DSPC, DMG-PEG, cholesterol
COVID-19In vivo
(tested on mice)
IntramuscularCell-free, simple, and rapid vaccine platform
Produced robust humoral response
Lead to a high level of neutralizing antibodies and a Th1-biased cellular response
It could be used for LNP-based mRNA vaccines in general and for a COVID19 vaccine in particular
[91]
HIV-1 Env-encoded as nucleoside-modified mRNA-LNPSize: ~80 nm
Lipid composition: ionizable cationic lipid, phosphatidylcholine, cholesterol, PEG-lipid
HIV infectionIn vivo (tested on rhesus macaques)IntramuscularElicited durable neutralizing antibodies that were stable for at least 41 weeks
Equal or better immune response when compared to adjuvanted recombinant protein vaccines
[81]
LNP-encapsulated HSV-2 nucleoside-modified mRNAn.r.HSV-2 infectionHSV-1 infectionIn vivo (tested on mice)IntramuscularPotent protection against HSV-1 and HSV-2 genital infections
Better at limiting virus replication in the genital tract than protein-based vaccine alternative
Prevented HSV invasion of the dorsal root ganglia in 97.5% of infected mice
[82]
Atu027 (liposome-encapsulated siRNA)Composition: positively charged AtuFect01, a neutral, fusogenic DPhyPE helperlipid and the PEGylated lipid MPEG-2000-DSPE, at molar ratios of 50:49:1Advanced or metastatic pancreatic adenocarcinomaClinical trial—Phase 1/2IntravenousUsed in combination with gentamicin-based chemotherapeutic treatment
Enhanced antitumor activity
Disease stabilization for 41% of patients
[92,93,94]
MTL-CEBPA
(double-stranded RNA formulated into a SMARTICLES® liposomal nanoparticle)
Lipid composition: liposomal formulationAdvanced hepatocellular carcinoma
Cirrhosis
Clinical trial—Phase 1IntravenousAdministered as monotherapy or in combination with sorafenib
Acceptable safety profile (comparable to drugs such as sorafenib, regorafenib, and nivolumab)Antitumor activity with a mean progression-free survival of 4.6 months
[95,96]
EphA2-targeting DOPC-encapsulated siRNALipid composition: DOPCAdvanced/recurrent malignant solid neoplasmClinical trial—Phase 1IntravenousDose-escalation study
Potential of slowing tumor cells growth by shutting down the causative gene
[97]
DCR-MYC (synthetic double-stranded RNA in a stable LNP suspension)n.r.Solid tumors
Multiple myeloma
Non-Hodgkins Lymphoma
Pancreatic neuroendocrine tumors
Primitive neuro-ectodermal tumors
Clinical trial—Phase 1IntravenousDose-escalation study
Antitumor potential by inhibiting MYC oncogene and thereby inhibiting cancer cell growth
[98]
mRNA-2416 (LNP-encapsulated mRNA encoding human OX40L)n.r.Relapsed/refractory solid tumor malignancies or lymphoma
Ovarian cancer
Clinical trial—Phase 1/2IntratumoralDose-escalation study
Administered alone or in combination with fixed doses of durvalumab
Potential immunomodulatory and antitumor activities
[99]
SGT-53 (cationic liposome encapsulating a normal human wild type p53 DNA sequence in a plasmid backbone)Lipid composition: complex DOTAP:DOPE cationic liposomeNeoplasm
Recurrent/refractory solid tumors in pediatric patients
Clinical trial—Phase 1IntravenousDose-escalating study
Administered alone or in combination with topotecan and cyclophosphamide
Potential to significantly sensitize pediatric cancer cell lines to the killing effect of the standard chemotherapeutic agents
[100,101]
SGT-53Lipid composition: complex DOTAP:DOPE cationic liposomeMetastatic pancreatic cancerClinical trial—Phase 2IntravenousAdministered in combination with gemcitabine/nab-paclitaxel
Efficient and specific delivery of p53 cDNA to tumor cells
Expected to restore wtp53 function in the apoptotic pathway
[101,102]
JVRS-100
(cationic liposome-plasmid DNA complex)
Lipid composition: DOTIM/cholesterol cationic liposomeRelapsed/refractory leukemiaClinical trial—Phase 1IntravenousDose-escalation study
Accelerated titration schema followed with one patient at each dose levelPotential immunostimulant activity
[103]
mRNA-1273 (LNP-encapsulated mRNA encoding for full-length perfusion stabilized spike protein of SARS-CoV-2)Lipid composition: ionizable lipid SM-102, DSPC, cholesterol, DMG-PEG2k, at molar ratios of 50:10:38.5:1.5COVID-19Clinical trial—Phase 1IntramuscularDose-ranging studyEvaluation of the safety and reactogenicity of a second dose vaccination schedule[104,105]
mRNA-1273Lipid composition: ionizable lipid SM-102, DSPC, cholesterol, DMG-PEG2k, at molar ratios of 50:10:38.5:1.5COVID-19Clinical trial—Phase 2IntramuscularAnalyze the development of cellular and humoral immunity against SARS-CoV-2 after administration of the third dose of vaccine in renal or renopancreatic transplant patients who have remained seronegative after the standard two-dose regimen[105,106]
LNP-encapsulated mRNA encoding the receptor-binding domain (RBD) of spike glycoprotein of SARS-CoV-2Lipid composition: lipid 9001, cholesterol, DSPC, DMG-PEG2kCOVID-19Clinical trial—Phase 3IntramuscularEvaluation of humoral immunity induced by the investigational vaccine and solicited adverse effects observed within 7 days post-immunization
Evaluation of protective efficacy
Adverse events collection over 0–28 days after each vaccination
Serious adverse events collection from Dose 1 through 12 months post complete series
[107]
Comirnaty (COVID-19 mRNA-embedded in LNPs)Lipid composition: ALC-0315, DSPC, cholesterol, ALC-0159, at molar ratios of 46.3:9.4:42.7:1.6COVID-19Clinical trial—Phase 4IntramuscularVaccine given in two doses in patients with primary or secondary immunosuppressive disorders
Results compared to a group of healthy control individuals
Investigation of safety and immune responses under 6 months of time for each immunized participant
[105,108]
BNT163b2 (LNP-formulated nucleoside-modified RNA encoding for full-length perfusion stabilized spike protein of SARS-CoV-2)Lipid composition: ALC-0315, DSPC, cholesterol, ALC-0159, at molar ratios of 46.3:9.4:42.7:1.6COVID-19Clinical trials—Phase 4IntramuscularAdministration of a third vaccine dose in adults who received two doses of an inactivated COVID-19 vaccine at least 3 months prior to the study
Evaluation of humoral immunogenicity, reactogenicity, and safety of a heterologous third vaccination dose
[109]
mRNA-1273.351 (LNP-encapsulated mRNA encoding for full-length perfusion stabilized spike protein of SARS-CoV-2 B.1.351 variant)Lipid composition: ionizable lipid SM-102, DSPC, cholesterol, DMG-PEG2k, at molar ratios of 50:10:38.5:1.5COVID-19Clinical trial—Phase 1IntramuscularEvaluation of the safety, reactogenicity, and immunogenicity of the new vaccine variant
Investigations on both naïve and previously vaccinated individuals
[110]
n.r.—not reported.
Table 2. A summary of polymer-based nucleic acids delivery systems applications.
Table 2. A summary of polymer-based nucleic acids delivery systems applications.
Delivery SystemPhysicochemical PropertiesDisease/ConditionTesting StageAdministration RouteObservationsRef.
Modified dendrimer nanoparticle (MDNP)-based RNA repliconComposition: modified PAMAM dendrimer:DMPE-PEG2k:RNA at a mass ratio of 11.5:1:2.3Zika virus infectionEx vivo (tested on mice)-Tool to generate ZIKV vaccine in the absence of reference virus stocks
Elicited ZIKV E protein-specific IgG responses
Assessment of immune responses without the need for recombinant production of native glycoprotein
[118,119]
Ab-conjugated LGA-PEI NPs for the delivery of TNAsSize: ~100–200 nm
Composition: LGA, PEI, TNAs in various mass ratios
Pancreatic cancerIn vitro
and
In vivo (tested on mice)
IntravenousEffective loading of TNAs, such as pDNA, mRNA, and miRNA, resulting in stable and functionalized nucleic acids
Improved binding, internalization, and gene expression in pancreatic cancer cells with high expression of targeted cell surface markers
The developed systems can be extended to other types of cancers that express specific biomarkers
[115]
STAT3-targeting PEI-encapsulated siRNASize: 135.1 ± 5.2 nm
Zeta potential: 34.6 ± 2.2. mV
N/P ratio: 10
Skin melanomaIn vitro
and
In vivo (tested on mice)
IntradermalMinimally invasive administration
Enhanced cellular uptake and transfection of siRNA
Enhanced gene silencing efficiency
Inhibited tumor cells growth in a dose-dependent manner
[117]
PEI-based nanoparticle encapsulated with IGF1 modified mRNAN/P ratio: 6Hypoxia
Myocardial infarction
In vivo (tested on mice)Intramyocardial injectionPotential for an extended cytoprotective effect of transient IGF1
Promoted cardiomyocyte survival and abrogated cell apoptosis under hypoxia-induced apoptosis conditions
Induced downstream increases in the levels of Akt and Erk phosphorylation
[120]
CALAA-01 (cyclodextrin containing polymer encapsulating anti-R2 siRNA)Composition: duplex of synthetic, non-chemically-modified siRNA (C05C), cyclodextrin-containing polymer (CAL101), stabilizing agent (AD-PEG), targeting agent (AD-PEG-Tf)Relapsed or refractory cancer
Solid tumors
Clical trial—Phase 1IntravenoussiRNA-containing nanocomplexes are targeted to cells that overexpress the transferrin receptor (TfR)
transferrin binds to TfRs on the cell surface, and the siRNA-containing nanocomplex enters the cell by endocytosis
inside the cell, the polymer unpacks the siRNA, allowing it to function via RNA interference
significant intratumoral downregulating of the target protein
[121]
siG12D-LODERComposition: miniature biodegradable biopolymeric matrix loaded with siRNAPancreatic ductal adenocarcinomaPancreatic cancerClinical trial—Phase 2ImplantationHighly effective and safe device implantation
High safety and tolerability profiles
Progression-free survival in the study population
[122]
Table 3. A summary of lipid-polymer hybrid-based nucleic acids delivery systems applications.
Table 3. A summary of lipid-polymer hybrid-based nucleic acids delivery systems applications.
Delivery SystemPhysicochemical PropertiesDisease/ConditionTesting StageAdministration RouteObservationsRef.
PEGylated CL4H6-MRTF-B siRNA-loaded LNPsSize: ~200 nm
Shape: spherical
N/P ratio: 7.5
Composition: CL4HS:DOPE: PEG-DMG. At molar ratio of 50:50:1
Conjunctival fibrosisIn vitro-Non-toxic at a concentration of 50 nM
Effectively silenced the MRTF-B gene
Enhanced encapsulation efficiency
Decreased fibroblast contraction after a single transfection
Remarkable efficiency even after long periods of refrigeration
[123]
Ultrasound-activated LPPsSize: ~170–250 nmComposition: polyplexes—PEI, pDNA; lipid formulation—DPPC, cholesterol, DPPG, PEG40SOvarian cancerIn vitro-Safe method for gene delivery
Enhanced transfection efficiency and low cytotoxicity when exposed to low-frequency ultrasound
Ultrasound exposure at a specific post-transfection time interval promotes carrier’s entry through the ECM of the cells into an intracellular environment
Intracellular uptake is reported even in the presence of chlorpromazine, which seems to be the carrier’s endosomal escape
[135]
DLNPs containing PEGylated BODIPY dyesSize: ~138 nm
Zeta potential: ~−1.0 mV
Composition: dendrimer-based lipid nanoparticle with BODIPY core, indole linker, and PEG-lipid of lengths between 1000 and 2000 g/mol
CancerIn vitro
and
In vivo
(tested on mice)
IntravenousParticles formulated with a pH-responsive PBD-lipid produced 5—to 35-fold more functional protein than control ones formulated with traditional PEG-lipid in vitro
Enhanced mRNA delivery potency in vivo
Efficient functional protein expression
Successfully mediated mRNA expression in tumors and simultaneously illuminated tumors via pH-responsive NIR imaging
[125]
PEGylated ionizable LNPs formulated with mRNASize: 60–100 nm
Composition: DOPE, cholesterol, C16-PEG2000 ceramide
Ionizable lipids:mRNA ratio: 10:1
Chronic liver diseases (e.g., liver fibrosis, cirrhosis)In vitro
and
In vivo (tested on mice)
IntravenousParticles obtained through microfluidic synthesis
Particles were mostly distributed to the liver
The highest transfection efficiency among different types of liver cells was reported for hepatocytes
[137]
siRNA-loaded LNPs conjugated with a PEG-monacyl fatty acidSize: ~50–90 nm
Zeta potential: ~ -4.3 - + 4.3 mV
Composition: YSK05, DSPC, cholesterol, PEG-DMG, PEG-MO
CancerIn vitro
and
In vivo (tested on mice)
Intraperitoneal injectionSignificantly improved siRNA delivery efficiency as compared to originally developed LNPs (i.e., [138])
siRNA was stably retained in mouse serum, leading to a gene knockdown effect
Significant gene silencing in cancer cells, ascites, and solid tumor of the mesentery
[139]
PLGA-LNPs loaded with CRISPR/Cas9 plasmidsSize: 179.6 ± 44.82 nm
Zeta potential: 29.6 ± 4.33 mV
Encapsulation efficiency: 76.5 ± 7.2%
Composition: PLGA, lecithin, DSPE-PEG-cRGD, DSPE-PEG-biotin (lipids at a molar ratio of 7:1.5:1.5)
GlioblastomaIn vitro
and
In vivo (tested on mice)
IntravenousEffective gene delivery
Restored sensitivity of glioblastoma cells to temozolomide (TMZ)
Highly safe and biocompatible multi-functional system
Potential treatment for TMZ-resistant glioblastoma and TNA delivery to the central nervous system
[130]
Tumor-selective LPP-p21-saRNA-322Size: 230.2 ± 10.3 nm
Polydispersity index: 0.15 ± 0.02
Zeta potential: −17.3 ± 0.4 mV
Composition: HA-PE tumor-selective liposomes, PEI-p21-saRNA-322 polyplexes
Colorectal cancerIn vitro
and
In vivo
(tested on mice)
RectalHigh drug accumulation in the tumor site
Efficient intracellular delivery and lysosome escapement
Significant inhibition of orthotopic colorectal tumor growth
Biocompatible delivery system; growth inhibition effect attributed to p21-saRNA-322 rather than the carrier
[134]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Niculescu, A.-G.; Bîrcă, A.C.; Grumezescu, A.M. New Applications of Lipid and Polymer-Based Nanoparticles for Nucleic Acids Delivery. Pharmaceutics 2021, 13, 2053. https://doi.org/10.3390/pharmaceutics13122053

AMA Style

Niculescu A-G, Bîrcă AC, Grumezescu AM. New Applications of Lipid and Polymer-Based Nanoparticles for Nucleic Acids Delivery. Pharmaceutics. 2021; 13(12):2053. https://doi.org/10.3390/pharmaceutics13122053

Chicago/Turabian Style

Niculescu, Adelina-Gabriela, Alexandra Cătălina Bîrcă, and Alexandru Mihai Grumezescu. 2021. "New Applications of Lipid and Polymer-Based Nanoparticles for Nucleic Acids Delivery" Pharmaceutics 13, no. 12: 2053. https://doi.org/10.3390/pharmaceutics13122053

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