ArticlePDF Available

Abstract and Figures

Lipid‐based nanoparticles have emerged as a clinically viable platform technology to deliver nucleic acids for a wide range of healthcare applications. Within this scope, one of the most exciting areas of recent progress and future innovation potential lies in the material science of lipid‐based nanoparticles, both to refine existing nanoparticle strategies and to develop new ones. Herein, the latest efforts to develop next‐generation lipid‐based nanoparticles are covered by taking a nanoarchitectonics perspective and the design, nucleic acid encapsulation methods, scalable production, and application prospects are critically analyzed for three classes of lipid‐based nanoparticles: 1) traditional lipid nanoparticles (LNPs); 2) lipoplexes; and 3) bicelles. Particular focus is placed on rationalizing how molecular self‐assembly principles enable advanced functionalities along with comparing and contrasting the different nanoarchitectures. The current development status of each class of lipid‐based nanoparticle is also evaluated and possible future directions in terms of overcoming clinical translation challenges and realizing new application opportunities are suggested.
Self‐assembly process of LNPs within microfluidic systems. A) The self‐assembly process of nucleic acid–LNPs in a microfluidic mixer begins when lipid excipients (i.e., ionizable lipid, PEG lipid, helper lipid, and cholesterol) dissolved in ethanol are rapidly mixed with the nucleic acid payload (e.g., siRNA, mRNA, etc.) dissolved in an aqueous buffer. This initially occurs at a low buffer pH whereby the ionizable lipid will assume a protonated form and bind, via electrostatic interactions, to the negatively charged backbone of the nucleic acid. Concurrently, the increasingly polar environment drives the formation of vesicles and the encapsulation of the nucleic acid. As the mixing progresses, and with increasing buffer pH, the ionizable lipid becomes increasingly neutral, leading to the fusion of adjacent vesicles to form the interior core of the LNP. The extent of vesicle fusion is influenced by the amount of PEG lipid added due to its hydrophilicity as well as steric effects. Adapted with permission.[⁸¹] Copyright 2021, MDPI Publishing. B) Major microfluidic mixing architectures for LNP formulation (top to bottom): T‐junction mixer, staggered herringbone micromixer, bifurcating mixer, and baffle mixer. C) The effect of varying flow rate on the mixing profile of ethanol (red) and water (blue) within a baffle mixer, visualized through a computational fluid dynamics simulation. Adapted with permission.[⁸⁹] Copyright 2018, American Chemical Society. D) A diagrammatic representation of an example of microfluidics parallelization, showing (from top to bottom) the arrangement of mixing channels, each mixing unit, and each mixing cycle within a parallelized microfluidics device, which adopts the staggered herringbone architecture. Evaluation of the in vivo delivery performance of E) siRNA–LNPs and F) mRNA–LNPs obtained using the parallelized device compared to bulk mixing. The evaluation was performed by means of administering to mice. The mRNA encodes for luciferase and the in vivo delivery performance was evaluated by directly visualizing the luminescence from selected regions of the mice. The results clearly show that the in vivo delivery performance of nucleic acid–LNPs obtained via microfluidic mixing is superior to those obtained via bulk mixing. D–F) Adapted with permission.[⁸⁵] Copyright 2021, American Chemical Society.
… 
This content is subject to copyright. Terms and conditions apply.
www.afm-journal.de
© 2022 Wiley-VCH GmbH
2203669 (1 of 38)
Review
Lipid Nanoparticle Technologies for Nucleic Acid Delivery:
A Nanoarchitectonics Perspective
Abdul Rahim Ferhan, Soohyun Park, Hyeonjin Park, Hyunhyuk Tae, Joshua A. Jackman,*
and Nam-Joon Cho*
Lipid-based nanoparticles have emerged as a clinically viable platform
technology to deliver nucleic acids for a wide range of healthcare applica-
tions. Within this scope, one of the most exciting areas of recent progress
and future innovation potential lies in the material science of lipid-based
nanoparticles, both to refine existing nanoparticle strategies and to develop
new ones. Herein, the latest eorts to develop next-generation lipid-based
nanoparticles are covered by taking a nanoarchitectonics perspective and the
design, nucleic acid encapsulation methods, scalable production, and applica-
tion prospects are critically analyzed for three classes of lipid-based nanopar-
ticles: 1) traditional lipid nanoparticles (LNPs); 2) lipoplexes; and 3) bicelles.
Particular focus is placed on rationalizing how molecular self-assembly prin-
ciples enable advanced functionalities along with comparing and contrasting
the dierent nanoarchitectures. The current development status of each class
of lipid-based nanoparticle is also evaluated and possible future directions in
terms of overcoming clinical translation challenges and realizing new applica-
tion opportunities are suggested.
DOI: 10.1002/adfm.202203669
A. R. Ferhan, S. Park, H. Park, H. Tae, N.-J. Cho
School of Materials Science and Engineering
Nanyang Technological University
Singapore 639798, Singapore
E-mail: njcho@ntu.edu.sg
H. Park, J. A. Jackman
School of Chemical Engineering and Translational Nanobioscience
Research Center
Sungkyunkwan University
Suwon 16419, Republic of Korea
E-mail: jjackman@skku.edu
possibilities have enabled rapid develop-
ment and deployment of mRNA vaccines
for coronavirus disease- (COVID-)
prevention, for example.[] Ongoing
advances in the field related to issues like
nucleic acid engineering, nanoparticle sta-
bility improvement, and immune safety
considerations along with application
trends have been extensively covered in
the past two years.[] However, the archi-
tectural design principles of lipid-based
nanoparticles have been relatively less
covered, especially from a material science
perspective that takes into account the
latest concepts and trends.
One emerging concept to rationally con-
trol the structure and function of nanoscale
objects is called nanoarchitectonics and has
been used to design advanced nanomate-
rials for biomedical, energy, and environ-
mental applications.[] Nanoarchitectonics
refers to the fusion of nanotechnology with organic chemistry,
supramolecular chemistry, and biology and involves combining
molecular-level manipulation with self-assembly and nanofabri-
cation.[] It is thus ideally suited to serve as a conceptual frame-
work to evaluate design trends and to suggest future directions
for the development of lipid-based nanoparticles for nucleic acid
delivery applications, especially as the field grows to include not
only traditional lipid nanoparticles (LNPs) but also lipoplexes and
bicelles among various possibilities (Scheme1). The molecular
design and self-assembly properties of natural and synthetic
lipids underlies the core nanoparticle technology and under-
standing how to modulate the coassembly of lipids by themselves
and in combination with nucleic acids can lead to new types of
lipid nanoarchitectures and support future innovation.
In this review, we cover the latest eorts to develop next-
generation lipid-based nanoparticles for nucleic acid delivery
applications by taking a nanoarchitectonics perspective and criti-
cally analyze the design, production, and application prospects for
three classes of lipid-based nanoparticles as follows: ) traditional
LNPs; ) lipoplexes; and ) bicelles. The nucleic acid encapsula-
tion strategies for each nanoparticle type are introduced and help
to rationalize how dierent lipid nanoarchitectures can be useful
depending on the delivery objective and application scope. While
the main focus is on distilling molecular self-assembly principles
to build advanced lipid-based nanoparticle systems, we also dis-
cuss pertinent issues related to overcoming clinical translation
challenges and enabling scaled-up manufacture that can help to
realize new application opportunities.
The ORCID identification number(s) for the author(s) of this article
can be found under https://doi.org/10.1002/adfm.202203669.
1. Introduction
Lipid-based nanoparticles are enabling new possibilities for
nucleic acid medicine across various dimensions such as infec-
tious disease vaccines and gene editing.[] These capabilities are
broadly applicable to various types of nucleic acids, including
deoxyribonucleic acid (DNA)[] and ribonucleic acid (RNA),[]
and some of the most promising types are messenger RNA
(mRNA),[] small interfering RNA (siRNA),[] transfer RNA
(tRNA),[] and self-amplifying RNA (saRNA).[] One of the
greatest benefits of lipid-based nanoparticle technology is the
platform aspect whereby one nucleic acid molecule can be
swapped out with another one, while utilizing the same nan-
oparticle design principles and manufacturing route.[] Such
Adv. Funct. Mater. 2022, 32, 2203669
www.afm-journal.dewww.advancedsciencenews.com
2203669 (2 of 38) © 2022 Wiley-VCH GmbH
2. Lipid Nanoparticles
Lipid nanoparticles (LNPs) represent the most clinically
advanced non-viral vector for the delivery of ribonucleic acids
(RNAs).[] LNPs have been widely used for encapsulating and
delivering small interfering RNA (siRNA) in RNA interference
(RNAi) therapy (see ref. []) and, more recently, have gained
widespread attention by enabling the rapid development of
messenger RNA (mRNA)-based vaccines to prevent COVID-
infection.[] The remarkable success of LNPs in the clinical
setting can be attributed to a few major technological mile-
stones with respect to lipid design and scalable methods of
production. In particular, the utilization of ionizable lipids has
led to tremendous improvements in LNP functional perfor-
mance, especially transfection eciency.[]
At the same time, the shift from bulk preparation methods
such as batch extrusion to microfluidic-based production strat-
egies has allowed more precise control over the physicochem-
ical properties and overall quality of LNPs along with scalable
manufacturing capabilities.[] This enabled systematic studies
to be conducted to understand the relationship between the
physicochemical properties of LNPs, which depend on the
nanoarchitectonic design strategy and production methods,
and in vivo delivery performance. Finally, advances in microflu-
idic mixing technologies have facilitated scale-up through par-
allelization, which paved the way for large-scale manufacturing
Scheme 1. Applying the nanoarchitectonics concept to support innovation of lipid-based nanoparticle technology for nucleic acid delivery. A) Timeline
of lipid-based nanoparticle technology progress and conceptual advances in the nanoarchitectonics field. B) Schematic representations of dierent
lipid-based nanoparticle types, including (1) traditional lipid nanoparticles (LNPs), (2) lipoplexes, and (3) bicelles. The nanoparticles are enabled by
the combination of controlling molecular self-assembly with various-shape lipids and using dierent nanofabriation strategies, highlighting how a
nanoarchitectonics perspective can be useful to support future innovation. FDA, United States Food and Drug Administration; EMA, European Medi-
cines Agency; DOPE, 1,2-dioleoyl-sn-glycero-3-phosphoethanolamine; DOPC, 1,2-dioleoyl-sn-glycero-3-phosphocholine; DSPC, 1,2-distearoyl-sn-glycero-
3-phosphocholine; DHPC, 1,2-dihexanoyl-sn-glycero-3-phosphocholine; GML, glycerol monolaurate.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (3 of 38) © 2022 Wiley-VCH GmbH
of high-quality RNA-encapsulated LNPs.[] This ultimately
accelerated the translation of RNA-based LNP therapeutic
solutions from the benchtop to the clinical setting. Neverthe-
less, eorts to continually improve the performance of LNPs
in delivering dierent types of nucleic acid payloads are still
ongoing. Along this line, it is crucial to understand the influ-
ence of dierent physicochemical parameters on the perfor-
mance of LNPs, and how specific design characteristics can
be incorporated to overcome limitations particularly associ-
ated with the transport of nucleic acids across biological bar-
riers. In this section, we will review the evolution of LNPs as
delivery systems for nucleic acids, especially RNAs, based on
their design principles and production platforms and shed
light on how nanoarchitectonic concepts are relevant to the
design of LNPs.
2.1. General Design Rules from Biological Inspiration
Within the context of nucleic acid delivery using lipid-based sys-
tems, a series of works have taken inspiration from the delivery
of native nucleic acids by lipoproteins.[] In biological systems,
native nucleic acids such as microRNAs are delivered to target
cells by lipoproteins such as high-density lipoproteins (HDLs).[]
This occurs via binding of apolipoprotein A-I (ApoA-I),
which is a major structural protein of HDLs, to specific scav-
enger receptors, such as SR-B that are present on target cell
membranes.[a] While HDLs are most popularly known for
their role in reverse cholesterol transport (RCT)[] in which
case they remove excess cholesterol from peripheral tissues and
transport it to the liver for biliary excretion, they have also been
found to play critical roles in the transport of other biomolecules
(e.g., signaling lipids, proteins, and nucleic acids).[] Endoge-
nous HDL particles are relatively small (–nm in diameter)
and are highly dynamic in terms of physicochemical properties.
Notably, they transition from a discoidal shape (i.e.,in their nas-
cent form) to a spherical shape (i.e., in their mature form) and
undergo changes in lipid composition during RCT (i.e.,due to
the adsorption, accumulation, and subsequent release of cho-
lesterol). Such morphological transitions oer insights into
key design criteria related to multiple aspects of nucleic acid
delivery, namely loading eciency, physiological stability, and
cell targeting and uptake eciency.[] Consequently, HDL par-
ticles represent excellent natural model systems for the formu-
lation of lipid-based nanoparticle drug delivery vehicles[] and
significant work has been performed to develop synthetic lipid
and lipoprotein nucleic acid delivery vehicles that are inspired
by these systems and follow a few general design principles as
described below.[,]
With regard to nucleic acid loading, the shape and size
transition of HDL particles upon cholesterol adsorption
during RCT demonstrate how shape and size constraints
may limit the number of nucleic acid molecules that can
be loaded and transported per particle, with spherical parti-
cles oering higher loading capacities than discoidal particles
of the same diameter. As such, the diameter of the particles
needs to be carefully considered to maximize loading without
compromising particle distribution, particle stability, or cell
uptake eciency. In addition, the amount of loading is also
influenced by the loading method. While passive loading is
generally more convenient, the loaded molecules can diuse
or leak out during circulation even before the delivery vehicle
reaches target cells. On the other hand, loaded molecules
that are covalently bound to the delivery particle are retained
until reaching target cells. However, an active release mecha-
nism needs to be implemented to ensure that the molecules
in that case can be eectively unloaded once the delivery
vehicle reaches a target cell. Practically, this issue highlights
how careful consideration of the noncovalent intermolecular
forces in the delivery system is warranted in order to ensure
that the delivery is optimally suited to enable both nucleic acid
encapsulation and eventual release. In some cases, stimuli-
responsive systems that are sensitive to environmental condi-
tions such as solution pH have proven useful to achieve these
performance objectives.
To further enable systemic circulation and attain a long
circulation half-life, the particles should first be physically
robust under flow conditions at physiological tempera-
ture (i.e., maintain structural integrity and not degrade or
destabilize under shear stress).[] In addition, the particles
should avoid binding with blood components, which could
lead to interparticle aggregation and uptake by the mononu-
clear phagocyte system (MPS). MPS evasion is traditionally
achieved through the introduction of hydrophilic moieties or
the attachment of non-fouling polymers (e.g., poly ethylene
glycol or PEG) to the outer surface of the particles.[] In
terms of size, the particles should not be too small (i.e.,
not less than  nm) to avoid kidney clearance.[] Along
this line, lipid conjugation strategies have been explored to
obtain lipidated therapeutic nucleic acids that boast higher
circulation times and are less prone to degradation.[] Ulti-
mately, such protection improves the overall biodistribution
of the nucleic acids and ensures that the nucleic acids effec-
tively reach target cells.[] Besides providing protection from
nuclease degradation, the conjugation of lipids to nucleic
acids has also led to new functionalities. For example, lipi-
dated nucleic acids can be designed to assemble into pre-
programmed shapes with precise control over geometrical
dimensions.[] This offers greater flexibility in terms of
architectural design and more faithfully mimics natural
delivery systems.
Finally, cell targeting and nucleic acid uptake in the cyto-
plasm represent two closely related critical considerations in
the development of targeted delivery vehicles. Inspired by the
ability of HDL particles to achieve targeted delivery via binding
of ApoA-I to specific scavenger receptors, cell targeting in syn-
thetic delivery vehicles can likewise be accomplished either
through opsonization by apolipoprotein A or E (ApoA or ApoE)
when the delivery vehicles are injected into the systemic circula-
tion or by functionalizing the delivery vehicles with natural tar-
geting molecules (e.g., ApoA or ApoE) or synthetic ligands.[]
Once the delivery vehicle reaches the target cell, the uptake
mechanism is also influenced by the targeting strategy. While
most lipid-based delivery systems rely on endocytosis, delivery
vehicles that undergo opsonization are internalized by cells
through phagocytosis.[] In addition, the mechanism of uptake
also depends on the surface charge, shape, size, and surface
chemistry of the delivery vehicle.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (4 of 38) © 2022 Wiley-VCH GmbH
2.2. LNP Design Principles
In addition to the general design rules inspired by biological
nanoparticle examples such as HDL particles described above,
the design principles behind the formulation of LNPs also
expand on the early designs of synthetic liposome-based car-
riers, which consist of a neutral unilamellar lipid bilayer with
an aqueous core, encapsulating short-stranded nucleic acid
molecules (i.e., oligonucleotides).[] While liposomes can serve
as carriers for a variety of therapeutic molecules, their simple
structure and the use of largely neutral lipids for the delivery of
oligonucleotides posed several challenges.[] Firstly, weak inter-
actions between the lipids and negatively charged nucleic acids
hampered ecient entrapment during production. Secondly,
when introduced into biological systems, liposomes are prone
to clearance by the reticuloendothelial system (RES), opsoniza-
tion and destabilization, leading to reduced circulation time.
Over the years, the design has significantly evolved to achieve
lipid-based delivery systems that overcome these limitations
through employing a combination of dierent types of phos-
pholipids. Within the context of nucleic acid delivery, it is vital
to ) ensure that the phospholipid carrier can eciently interact
with, condense, and entrap the nucleic acids, in a process col-
lectively known as encapsulation and ) eectively deliver the
encapsulated nucleic acids to the cells with minimal biological
interference during circulation. Encapsulation serves not only
to spatially entrap the nucleic acids within the delivery vehicle,
but to also isolate them from the external environment while
protecting their structure and function. Initially, permanently
cationic lipids were largely used to improve the interaction
between positively charged lipid membrane interfaces and neg-
atively charged nucleic acids.
While it was successful in increasing the encapsulation e-
ciency, the use of permanently cationic lipids attracted nonspe-
cific biological interactions and suers from problems associ-
ated with triggering the innate immune response (e.g., arising
from the activation of the complement system) and cytotoxicity
(i.e., arising from the activation of proapoptotic and proinflam-
matory cascades), thus failing to provide a sustainable solution
to issues related to reduced systemic circulation times. Early
eorts to mitigate these issues focused on utilizing cationic
lipids with delocalized charge while the issues were eventually
tackled through the design of a unique class of lipids with
switchable charge known as ionizable lipids, which achieved
significant improvements in encapsulation eciency by being
positively charged at low pH during production and avoided
nonspecific biological interactions by remaining neutral at
physiological pH during circulation. A careful balance of head-
group and tail properties further enabled control over mem-
brane biophysics to guide the development of tailored LNPs
with heightened functionalities.
When combined with PEG lipid, helper lipid, and choles-
terol, the delivery system becomes distinct from a traditional
liposomal carrier in terms of structure, morphology, and lipid
composition, and is therefore referred to as an LNP. Archi-
tecturally, LNPs are characterized by a phospholipid bilayer
surrounding an electron-dense core in which ionizable lipid-
encapsulated nucleic acid payloads are densely packed in the
interior[] (Scheme2). Compared to liposomes, which consist
of phospholipid bilayers and cholesterol with distinct phase
transitions from ordered gel phase to liquid crystalline phase,
Larson et al. reported that cationic mRNA–LNPs transition
from an inverse hexagonal phase at pH values below the pKa
of the cationic lipid, to a lamellar phase above the pKa.[] Like-
wise, such phase changes also occur when the temperature is
increased, indicating that the inverted hexagonal phase is more
thermodynamically favorable. Based on these developments, it
becomes clear that the most critical aspects of LNP formulation
are nanoarchitecture design and lipid composition. By varying
the lipid composition, the interaction with the nucleic acid pay-
load can be carefully tuned. This directly impacts the encapsu-
lation eciency and influences the particle morphology and
overall charge, which taken together will determine the overall
performance of the delivery system. Conversely, the choice of
lipids also depends on the nature of the nucleic acid payload. In
the following subsections, we elaborate on the rationale behind
the choice of lipid compositions, as well as the considerations
for dierent payloads, in designing RNA–LNP delivery systems.
2.2.1. Lipid Composition
The formulation of LNPs is uniquely characterized by the uti-
lization of four dierent lipid excipients namely: ) ionizable
Scheme 2. Lipid nanoparticle architecture. Structural configuration and key components of a lipid nanoparticle with nucleic acid payload, highlighting
how each component contributes to the overall structure and function.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (5 of 38) © 2022 Wiley-VCH GmbH
lipids, ) helper phospholipids, ) cholesterol, and ) PEG lipid.
Each one of these excipients plays specific roles in ensuring the
overall eectiveness of LNP as a nucleic acid delivery vehicle.
Recently, Siegwart Daniel and co-workers developed selec-
tive organ targeting (SORT) nanoparticles, which enable the
delivery of mRNA and gene editing systems to non-liver tis-
sues.[] SORT nanoparticles include a supplemental molecule
whereby the chemical structure determines the LNP’s tissue-
specific activity. Their work demonstrates how a “fifth excip-
ient” can further enhance the physicochemical properties and
targeted delivery performance of LNPs. In all cases, the key to
designing LNPs with high delivery ecacy lies in achieving an
optimal balance between particle stability, sustained systemic
circulation, and the ability to release the payload in a controlled
manner inside the target cell, all of which rely on the proper
selection of the dierent lipid excipients as described below.
Ionizable Lipids: Ionizable lipids are positively charged at
acidic pH and neutral at physiological pH. During LNP produc-
tion, they specifically serve to condense and entrap negatively
charged nucleic acids (e.g., siRNA, mRNA, pDNA) through elec-
trostatic interactions under acidic conditions. However, when
the LNPs are introduced to the physiological environment, the
ionizable lipids will switch to being neutral[b] (Figure1A).
Compared to permanently charged cationic lipids, the use of
ionizable lipids with neutral surface charge in the physiological
system is especially beneficial in preventing the adsorption of
negatively charged biomolecules onto the LNP and preventing
rapid sequestration by immune cells during systemic circulation.
Once the LNP reaches the endosome, which typically presents
an acidic environment, the ionizable lipids will revert to being
positively charged. This leads to electrostatic interactions with
the anionic endosomal membrane, resulting in the formation of
a nonbilayer hexagonal phase that momentarily destabilizes the
endosomal membrane and thereby facilitates the release of the
nucleic acid to the cytosol[] (FigureB).
There are three major classes of ionizable lipids depending
on the tail structure of lipids, namely conventional two-
tail, multi-tail, and branched-tail ones (Figure2). While the
number/structure of the tail region largely aects the overall
geometry of the lipid self-assembly (i.e., a bulkier tail group
promotes a more cone-shaped structure), another important
factor to consider is tail saturation. Unsaturated ionizable lipids
generally increase fluidity and a tendency to form a nonbilayer
phase, which facilitates membrane disruption and payload
release.[] One of the most widely used two-tail, unsaturated
ionizable lipids is DLin-MC-DMA, which is incorporated in
the FDA-approved siRNA–LNP drug (ONPATTRO). Multi-tail
ionizable lipids have three or more tails, which can result in
a more cone-shaped geometry to enhance endosome disrup-
tion.[] Along this line, branched-tail ionizable lipids have also
been developed, which may present methacrylate, acrylate, or
ester chains that extend either as a C branch near the head or
a C branch at the end. This can promote endosomal escape
arising from stronger protonation of spaced ionizable lipids at
endosomal pH and the increased cross-section of lipid tails,
allowing for the formation of a more cone-shaped geometry.[]
Figure 1. Illustration of the pH-dependent behavior of lipid nanoparticles due to ionizable lipids. A) Principle of the mechanism of ionizable lipids.
During formulation, the ionizable lipids present a positive charge at low pH (below its pKa) to promote favorable interactions with negatively charged
nucleic acids and facilitate nucleic acid encapsulation. During storage and in circulation under physiological pH (above its pKa), the ionizable lipids
will switch to being neutral and remain stable. Once the LNP reaches the endosome, which typically presents an acidic environment (below its pKa),
the ionizable lipids will become protonated again and resume a positive charge. B) Schematic representation of LNP interaction with the endosome,
leading to a change in lipid geometry due to the formation of ion pairs. Protonated ionizable lipids interact with anionic lipids adopting an inverted
cone geometry, which promotes the formation of the inverted hexagonal (HII) phase associated with membrane fusion.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (6 of 38) © 2022 Wiley-VCH GmbH
Biodegradability can be improved by incorporating ester bonds
to promote degradation into nontoxic metabolites after intra-
cellular delivery.[] This is critical to reduce systemic accu-
mulation and eliminate potential side eects. Finally, since
chain branching increases the cross-section of phospholipid
tails (i.e., reduces chain packing) and results in the formation
of phospholipids with cone-shaped geometry that promote
spontaneous negative curvature to the membrane (i.e., due to
the inversion of cone-shaped structures), it generally leads to
better fusogenicity.
As a critical factor in the nucleic acid delivery process, it
is important to understand the structure-function relation-
ship between the molecular structure of ionizable lipids to the
transfection eciency of the LNPs, and how this in turn will
determine the therapeutic eciencies of LNPs.[] For example,
it has been found that lipid tail saturation, as well as the type
of linker between the lipid tail and the headgroup can signifi-
cantly influence transfection eciency.[c,] More interestingly,
small structural variations such as those arising from dier-
ences in the double bond positions of unsaturated lipid tails
Figure 2. Structural classes of ionizable lipids. Three major classes of ionizable lipids depending on the tail structure, namely two-tail, multi-tail, and
branched-tail.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (7 of 38) © 2022 Wiley-VCH GmbH
or branching positions, are sucient to produce significant
variations in delivery performance.[,] Following this line, it
is anticipated that ionizable lipids with novel structures will
continue to emerge, which will facilitate the rational optimiza-
tion of LNP lipid composition. Such developments may also be
guided by the need for improvements in biodegradability, which
is a key feature for clinical translation, as well as the promising
opportunity of fine tuning the lipid geometry through adjust-
ments in tail branching (i.e., in terms of extent and position),
tail length, and number of tails.
Helper Phospholipids: The roles of helper phospholipids are
various, ranging from facilitating membrane fusion, which
contributes to higher transfection eciencies, and increasing
bilayer stability.[] Helper phospholipids can have dierent
molecular structures and geometry. For example, ,-dio-
leoyl-sn-glycero--phosphoethanolamine (DOPE), which is
widely regarded as a fusogenic lipid, has a small phosphoe-
thanolamine head (i.e., comprising of a primary amine and
a phosphoric acid moiety), and two bulky and unsaturated
oleoyl chains, resulting in a cone-shaped geometry. It is sug-
gested that such geometry has the capability to stabilize the
non-bilayer inverted hexagonal (HII) phase, which transiently
forms during membrane fusion. Hence, the addition of DOPE
as helper lipids enhances fusogenicity, eectively increasing the
amount of LNP accumulation in target cells leading to higher
transfection eciencies. On the other hand, phosphotidylcho-
lines (PCs), such as ,-distearoyl-sn-glycero--phosphocholine
(DPSC) and ,-dioleoyl-sn-glycero--phosphocholine (DOPC)
form a cylindrical geometry and provide greater bilayer stability
relative to DOPE. In particular, the high melting temperature
(Tm) of DSPC, which is a saturated PC, leads to highly stable
bilayers at physiological temperatures. However, such high
stabilities will hamper endosomal release and reduce the e-
ciency of payload delivery. In comparison, DOPC, which is an
unsaturated PC, has a lower Tm and exists in the fluid phase
at physiological temperatures. As a result, they contribute to
lower bilayer stability while allowing better eciencies in pay-
load delivery compared to DSPC. Taken together, DSPC oers
the highest bilayer stability, followed by DOPC and DOPE. In
other words, DOPC represents a good intermediate choice as a
helper lipid and the balance between bilayer stability and trans-
fection eciency is often achieved through the introduction of
cholesterol.
Cholesterol: The introduction of cholesterol can aect the
degree of lipid packing, which influences membrane fluidity
and permeability of the bilayer of the LNP.[] Structurally, cho-
lesterol comprises of three domains namely the head, body, and
tail.[] The head is a hydroxyl group positioned in a tilted con-
figuration above one of the cyclohexane or cyclohexene rings
of the body. The body comprises of a fused four-ring system
containing three cyclohexane or cyclohexene rings and one
cyclopentane ring, while the tail is a saturated alkyl side chain
of the cyclopentane ring. It is suggested that the flexibility of
cholesterol arises from the lack of functional groups in the tail
and saturated nature of the alkyl chain. As a result, cholesterol
molecules can fill gaps between other phospholipid molecules,
through the interaction of the hydroxyl head groups with the
aqueous phase of the phospholipid membrane (i.e., via polar
interactions and hydrogen bonding). Together, the body and
tail of the cholesterol molecule promote orderly arrangement
of the hydrophobic portion of the bilayer, leading to tighter lipid
packing of the membrane. Consequently, this reduces mem-
brane permeability and increases the overall structural integrity
of the LNPs. In addition, cholesterol can traverse lipid bilayers
and oer the possibility of equilibration across a concentration
gradient should the need arise. During circulation, cholesterol
also helps to prevent the diusion of the other lipid excipients
of the LNPs into high-density lipoproteins (HDL)-endogenous
lipids thereby improving the LNP stability in vivo.
Poly(ethylene glycol) PEG Lipid: The introduction of PEG lipid
serves to protect the LNP surface from opsonization, reticu-
loendothelial clearance, and destabilization during systemic
circulation.[a,] This is achieved through a combination of
steric repulsion and the formation of a hydration layer arising
from the extension of hydrophilic PEG chains from the LNP
surface. Based on the same mechanism, PEG-lipids also pre-
vent the aggregation of LNPs during production, storage, and
in circulation. Taken together, the structure of PEG extending
from the LNP surface increases the overall stability of LNPs,
which is largely advantageous. However, the prevention of non-
specific molecular adsorption on the LNP surface oered by
PEG lipids comes with a caveat as it also implies a significant
reduction in the adsorption of apolipoprotein E (ApoE), ApoE
is an extracellular protein that plays a key role in influencing
the systemic circulation time of LNPs. The binding of ApoE has
been found to trigger a redistribution of lipids at the shell and
the core and plays a critical role in fusogenicity. In particular, it
facilitates the cellular uptake of LNPs through receptor-medi-
ated cellular entry.[] Hence, it is vital to strike an optimal bal-
ance when deciding on the appropriate amount of PEG to be
introduced to LNPs to ensure good protection against opsoni-
zation and reticuloendothelial clearance without compromising
fusogenicity and cellular uptake. The adjustment of PEG den-
sity also allows the tuning of LNP pharmacokinetics for dif-
ferent applications.
While there are obvious benefits of including PEG lipids in
LNPs, it is worthy to note that PEG is widely known to trigger
immune responses and hypersensitivity reactions by activating
the complement system. In addition, it can also suer from
accelerated blood clearance (ABC) (i.e., following the binding
of anti-PEG antibodies). This has been intensively discussed,
most recently during the development, and ongoing evalu-
ation of COVID- vaccines, in which the potential of PEG
lipids to cause anaphylaxis especially among patients with
PEG allergy represents a significant concern.[] In fact, there
is evidence that highlights the importance of the form of PEG;
specifically suggesting that PEG conjugated with lipids, not
PEG alone, is more likely to cause allergic reactions.[] Along
this line, it is also important to note that PEG-lipids have the
tendency to desorb and therefore escape the LNP architec-
ture to circulate freely, increasing the likelihood of triggering
immune responses. Considering these issues, several natural
(e.g., polyaminoacids, glucosaminoglycans) and synthetic (e.g.,
polyacrylamide, polysulfobetaine methacrylates) polymers have
been investigated as alternatives to PEG.
Cell-Specific Targeting: In addition to their primary roles in
encapsulation, maintaining particle stability, and supporting
systemic circulation, the lipid excipients can also play a critical
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (8 of 38) © 2022 Wiley-VCH GmbH
role in achieving cell-specific targeting as well as in modulating
biodistribution and uptake mechanisms. For example, it has
been established that ApoE adsorption is largely responsible for
LNP accumulation in the liver since lipid-associated ApoE binds
to low-density lipoprotein (LDL) receptors in the liver.[] Along
this line, several works have explored the possibility of varying
the helper lipid composition to modulate the LNP interaction
with ApoE.[,c,] It has been observed that LNPs containing
DOPE helper lipid interact more favorably with ApoE, therefore
resulting in increased levels of accumulation in the liver.[c]
On the other hand, LNPs containing DSPC helper lipid exhib-
ited weaker interactions with ApoE, leading to accumulation
in the spleen. This was attributed to the unsaturated nature of
DOPE, which leads to higher membrane fluidity; greater mem-
brane fluidity has been previously observed to increase protein
adsorption.[]
Based on this rationale, lipid compositions can therefore
be tuned to modulate the uptake mechanism for LNPs in line
with controlling protein adsorption. In the most common sce-
nario in which the cell uptake of LNPs occurs via the endocytic
pathway,[] the collective role of helper lipids and PEG lipids is to
reduce protein adsorption and prevent corona formation. In the
case of receptor-mediated endocytosis of ligand-targeting LNPs,
lipids can be conjugated with antibodies or aptamers. It is worthy
to note that for uptake via endocytosis, the ionizability of the
lipids represents another crucial factor that enables the LNPs to
withstand pH changes throughout the process, achieve endolyso-
somal escape, and eventual release of the nucleic acid cargo.
Alternatively, the cell uptake of LNPs can also occur via phago-
cytosis.[] In this case, lipid compositions should preferably be
tuned to induce protein adsorption and trigger opsonization
involving apolipoproteins (e.g., ApoA-I and ApoE), which would
lead to phagocytosis. Furthermore, the concept of decorating
pre-synthesized LNPs with native or synthetic biomolecules for
the purpose of modulating biodistribution and cell uptake can be
extended to improve cell targeting. Recently, SORT nanoparticles
developed by Siegwart Daniel and co-workers demonstrated the
ability to deliver mRNA with increased cell-targeting capabilities
(i.e., to non-liver tissues) by endowing LNPs with an additional
molecule that exhibits tissue-specific activity.[]
2.2.2. Payload Considerations
LNPs were initially developed and optimized for the delivery of
siRNAs. Although similar lipid compositions can theoretically
encapsulate other types of nucleic acids, it may be challenging
to apply the same formulation across the board considering the
dierences in molecular size, net charge, and conformational
structure.[] For example, small activating RNAs (saRNAs), and
microRNAs (miRNAs), both of which are also involved in the
RNA interference pathway like siRNAs, are double-stranded
and with lengths between  and  base pairs. On the other
hand, single-stranded mRNAs can and double-stranded plasmid
DNAs (pDNAs), can vary from several hundred to several hun-
dred thousand nucleotides. Since the negative charge of nucleic
acids arises from the phosphate backbone, their net charge is
directly correlated to the number of nucleotides and whether
they are single- or double stranded. In addition, nucleic acids
such as pDNAs and transfer RNAs (tRNAs) have well-defined
conformational structures; pDNAs are circular and tRNAs have
a distinctive three-leafed clover structure comprising of three
hairpin loops. As such, adjustments to the lipid composition
are often required to accommodate dierent nucleic acids. Spe-
cifically, the ratio between the four lipid excipients, as well as
the ratio of ionizable lipid to RNA has been found to signifi-
cantly aect the delivery ecacy both in vitro and in vivo.[]
siRNA versus mRNA: In order to exemplify the importance
of considering the dierent characteristics of nucleic acid pay-
loads on the ecacy of LNP delivery systems, we examine two
types of RNAs, namely siRNA and mRNA, that are most com-
monly encapsulated by LNPs. It is well known that siRNA and
mRNA structurally dier based on size and eective charge.
While siRNA has a short and well-defined structure (i.e.,
double-stranded with a typical length of between – base
pairs[]), mRNA has a relatively less defined structure with
variable lengths (i.e., threadlike, single-stranded with lengths
in the range of several hundred to a thousand bases[])—as
a reference, the SARS-CoV- single-stranded genomic RNA
is approximately  kilobases in length.[] These dierences
potentially lead to variations in lipid packing and the overall
LNP configuration and size. For example, it was found that
while siRNA molecules in siRNA–LNPs are largely confined
within multilamellar nanostructures (i.e., consisting of siRNA
molecules sandwiched between tightly packed concentric lipid
bilayers of ionizable lipids with PEG lipids on the outermost
layer), mRNAs, due to their larger size, can rearrange the
compartmental organization within the internal core of the
mRNA–LNPs and form inverted hexagonal nanostructures
instead.[] In this configuration, mRNA molecules are encapsu-
lated within an aqueous environment surrounded by a mono-
layer of ionizable lipids with their headgroups facing inwards.
These individual compartments are in turn densely packed and
contained within a nonpolar environment bounded by a phos-
pholipid monolayer containing a mixture of ionizable lipids,
helper lipids, and PEG lipids with their headgroups facing the
external aqueous environment. Based on these arguments,
it becomes clear that the chemical, as well as architectural,
requirements of the lipid formulation would greatly depend
on size and structure of the intended encapsulated material.
For example, a smaller payload (e.g., siRNA) with well-defined
structure would result in a denser, well-packed arrangement
of encapsulated nucleic acids in the interior of the LNPs and
potentially a smaller overall LNP. Conversely, a larger payload
(e.g., mRNA) with a less ordered structure would result in a
less dense arrangement within the LNPs and possibly a larger
overall LNP. Indeed, previous works have shown that siRNA
and mRNA are most eectively delivered in dierentiated LNP
formulations.[a]
Besides adjusting the ratio between the four lipid excipients,
various proprietary ionizable lipids with new structures and
functionalities have emerged driven by the need to enhance
the ecacy of mRNA–LNPs. In a noteworthy development,
Ball et al. explored the possibility of codelivery of siRNA and
mRNA using a single LNP formulation.[] To achieve this, they
employed a formulation comprising of an ionizable and biode-
gradable amine-containing lipidoid, cholesterol, DSPC, DOPE,
and PEG lipid. They began with two previously reported LNP
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (9 of 38) © 2022 Wiley-VCH GmbH
formulations that were found to be potent for siRNA and mRNA
delivery and systematically varied the lipidoid-to-RNA ratio, as
well as the percentage of lipidoid in relation to the helper lipid,
cholesterol, and PEG lipid. They also gradually changed the
composition of helper lipids from a blend containing majority
DSPC to DOPE, and introduced a negatively charged helper
polymer in the form of poly(sodium -styrenesulfonate) (PSS) to
enhance the ecacy. Surprisingly, they found that coformulation
of siRNA and mRNA in the same LNP was not only possible but
enhanced the ecacy of both drugs in vitro and in vivo. It was
suggested that since DSPC has two saturated aliphatic tails while
DOPE has a cis-double bond in each of its two aliphatic tails, the
packing of nucleic acids within the internal core of the LNP can
be carefully optimized by varying the ratio of DSPC to DOPE
to accommodate the significant structural dierences of both
siRNA and mRNA. Furthermore, the additional negative charge
of PSS promotes electrostatic attraction in the particle and
increased overall particle stability. Separately, Hejdankovaet al.
recently developed a novel structural class of ionizable adaman-
tane-based lipidoids named XMaNs, which overcomes issues
related to tedious optimization of LNP components for a wide
range of nucleic acids.[] Briefly, XMaN LNPs were formu-
lated with DMG-PEG, cholesterol, the helper lipid DOPE,
XMaN lipidoids (Figure3A), and combined with diverse types of
nucleic acids (mRNA, siRNA, pDNA, ,-cGAMP) (FigureB)
in a microfluidic device. Nucleic acid-loaded XMaN LNPs have
an intricate architecture (Figure C) and the microfluidic pro-
duction has allowed the fabrication of XMaN LNPs with good
sizes and excellent size distribution (FigureD). Using the best
performing member of the XMaN family, they were successful
in entrapping and delivering siRNA, mRNA, plasmid DNA as
well as a cyclic dinucleotide. Taken together, these works high-
light that the delivery of dierent nucleic acid payloads, whether
isolated or in combination, is highly achievable and can be sig-
nificantly enhanced through a series of systematic and ecient
formulation optimization.
Figure 3. General overview of the approach to obtain LNPs incorporating XMaN lipidoids to achieve the delivery of dierent types of nucleic acids.
A) Four lipid excipients, including XMaN lipidoids, are first mixed in ethanol, before they are combined, using a microfluidic platform, with B) a wide
range of nucleic acid payloads that are dissolved in an aqueous buer. C) A representation of a possible architectural configuration of nucleic acid-
encapsulating XMaN LNPs. D) Cryo-TEM images of mRNA-loaded XMaN LNPs, their size distribution and evaluation of their functional stability after
prolonged storage (i.e., up to five weeks) at 4°C. Adapted with permission.[63] Copyright 2021, Wiley-VCH.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (10 of 38) © 2022 Wiley-VCH GmbH
Formulation Optimization: The evolution of LNPs from
encapsulating siRNA to mRNA highlights the need to optimize
the lipid formulation for each type of nucleic acid payload since
the formulations are not interchangeable. To achieve the most
ideal LNP formulation for the delivery of dierent nucleic acid
payloads, most works rely either on the one-factor-at-a-time
(OFAT) approach or the design-of-experiment (DoE) approach.
In the OFAT approach, one lipid is varied systematically while
keeping all other factors constant. While this method is advan-
tageous to reveal the direct correlation between the varied
parameter to the overall physicochemical property of the LNP,
it does not consider higher order interactions between the dif-
ferent lipid excipients. Consequently, a possible optimal formu-
lation arising from second or third order interactions between
the lipids could be missed. The method is also time consuming
and could lead to wastage of precious reagents, especially when
it involves optimizing lipid parameters for dierent nucleic
acids. The DoE approach, on the other hand, gains maximum
insights about the eect of varying several parameters to the
overall property of the LNPs through minimal number of
experiments.
Generally, it involves a screening design paired with
response surface methodology (RSM). Screening design first
identifies potential significant variables (e.g., choice of ioniz-
able lipid, helper lipid, amount of cholesterol, ratio of lipid to
nucleic acid payload, etc.) while RSM determines the critical
degree of adjustment that should be applied to the most impor-
tant variable.[] The DoE approach provides information on
the interaction between several parameters, leading to more
accurate conclusions compared to the OFAT approach. More
importantly, it streamlines the formulation optimization pro-
cess. This was clearly demonstrated by Love et al., where the
formulation of LNPs containing erythropoietin (EPO)-mRNA
and C- phospholipid was optimized from a previous for-
mulation designed for siRNA–LNP.[] The study identified
the choice of helper phospholipid as the most important para-
meter for eective delivery EPO–mRNA; the use of DOPE was
favorable over DSPC, which was initially used in the formula-
tion of siRNA–LNP. The weight ratio of C- phospholipid
to mRNA was also identified as another important parameter.
Notably, second-order interactions (e.g., between C- mol%
and C--to-mRNA weight ratio) were detected, highlighting
the advantage of the DoE approach over the OFAT approach.
Such capability to detect higher-order interactions is especially
vital when performing formulation optimization involving ion-
izable lipids with novel functionalities, of which multiple layers
of interdependencies with other parameters could be dicult
to resolve.
In a separate demonstration, Dahlman and co-workers
described a cluster-based screening approach, which was
inspired by statistical DoE methodologies, to optimize the for-
mulation of mRNA–LNPs for the delivery of nebulized thera-
peutic mRNA.[b] The cluster-based approach involves, firstly,
the identification of LNP chemical traits that influence the in
vivo delivery of nebulized LNPs (e.g., the amount of PEG added
to the LNP, the structure of the lipid-PEG, the charge of the
phospholipid and the presence or absence of cholesterol). A
hypothesis relating to the eect of varying each of these traits
is assigned as an axis in an N-dimensional chemical space and
asmall group of LNPs (e.g., around  to ) are then formulated
at the extreme of each axis. Within these “extreme groups,” a
selection process is further applied to determine whether )the
LNPs meet size and stability requirements, ) the stability
remains unchanged when LNPs are pooled together, and
) the pooled LNPs functionally deliver mRNA following nebu-
lization. Promising groups that pass this selection process are
then expanded and combined by formulating LNPs near the
intersection of these groups, with newly formulated LNPs also
subject to the same selection process. All these data, including
from groups that initially failed, were considered to design
subsequent LNPs. Through this regimented process, they
were able to eciently scan a diverse LNP chemical space, and
manipulate several pertinent parameters that influence in vivo
delivery, iteratively.
2.2.3. Physicochemical Parameters
The size and surface charge of the resultant LNPs have been
identified as critical parameters that aect not only the amount
of nucleic acid payload that can be packaged within the internal
core but also the in vivo behavior of LNPs.[] These parameters
influence the pharmacokinetic profile of LNPs. For example,
smaller particles are observed to exhibit longer systemic circula-
tion times and slower clearance from the blood stream. More
importantly, it has been observed that particle size aects cell
uptake since the endocytic pathway varies with particle size,
which has implications on the intracellular processing and
overall ecacy of the LNPs.[] For instance, it was found that
the endocytosis of nm diameter LNPs is based on a dynamin-
dependent pathway while the endocytosis of LNPs larger than
nm in diameter is based on a clathrin-dependent pathway.[]
More recently, Hassetet al. evaluated the eect of LNP size on
mRNA vaccine immunogenicity on mice. They systematically
changed the LNP diameter at a fixed lipid composition and
found that LNPs with smaller diameter did not lead to a signifi-
cant immune response in mice. However, significant immune
responses were observed in non-human primates regardless of
the LNP diameter.[] In a separate study, Chenetal. also found
that the hepatic gene silencing of FVII in mice was significantly
influenced by the size of the siRNA–LNPs.[] Specifically, gene
silencing was far more ecient using LNPs with an optimal
size of –nm in diameter compared to smaller (i.e.,  nm
in diameter) and larger (i.e., nm in diameter) particles. The
observation was attributed to the inability of large LNPs to infil-
trate the fenestrations in the liver vasculature and the instability
of smaller LNPs in serum. In addition, the ionizable lipids also
tend to dissociate more rapidly from smaller LNPs leading to
lower transfection eciencies. Aside from demonstrating the
eect of size on the ecacy of LNPs, these studies also high-
light the importance of producing LNPs with high monodis-
persity without overlapping size distributions. Likewise, the
phospholipid charge not only aects the interaction between
the lipid excipients and the nucleic acid payload, which deter-
mines the encapsulation eciency but also determines the
overall surface charge and zeta potential of the LNPs, which
strongly influences interactions with biological components in
vivo. While the uptake pathway is influenced by particle size,
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (11 of 38) © 2022 Wiley-VCH GmbH
the degree of initial adsorption of the LNP on the cell mem-
brane is governed by its zeta potential. It is therefore essential
to achieve precise control over these parameters.
2.3. Platform Development
As evidenced by the previous section, the key to the success
of LNPs as nucleic acid delivery systems lies in a solid under-
standing of the structure–function relationship. Clearly, the
molecular structures define the resultant geometry of the lipid
components, aect their interaction with the nucleic acids, and
influence the overall morphology and physicochemical proper-
ties of the LNP (i.e., the concept of nanoarchitectonics[]), all
of which in turn determine the behavior of LNPs in vivo, and
ultimately the performance of the LNPs within the context of a
nucleic acid delivery system.[]
While the design principles are built upon these correla-
tions, they would not have been realized without proper and
consistent control over the LNP fabrication process, which has
allowed systematic studies to be conducted in a reproducible
manner. At the same time, reproducibility and scalability rep-
resent key prerequisites for the clinical development of LNPs.
Along this line, eorts have been continually made to improve
the production methods of LNPs. The development of produc-
tion platforms serves two overarching objectives. Firstly, it aims
to achieve controlled and reproducible production of LNPs so
that fundamental investigations looking into the structure-func-
tion relationship can be performed reliably thereby facilitating
formulation optimization of LNPs. Second, it promotes scalable
manufacturing, which paves the way for the entry of LNPs to
the clinic. In the following subsections, we will elaborate on
how the production platforms have evolved over recent years,
particularly through the advent of microfluidics technologies.
2.3.1. Bulk Mixing Methods
Early production of liposome- and LNP-based delivery systems
relies on bulk mixing methods. Liposome-based carriers, par-
ticularly those containing siRNA, have been commonly pre-
pared through lipid film hydration followed by extrusion, or
using the sonication and homogenization methods. Extrusion
involves passing the solution containing the lipid mixtures
through a series of filters to attain liposomes with the desired
size. The intended payload is encapsulated through the sonica-
tion or homogenization steps.
Besides extrusion, liposomes can be fabricated following the
ethanol injection method. In this method, lipid mixtures in
ethanolic solution are injected into a large volume of aqueous
solution under continuous stirring. The ethanol is then
removed via evaporation leading to the spontaneous formation
of vesicles. This method has several advantages over the film
hydration method. Firstly, the method is relatively straightfor-
ward as it does not involve any intermediate processing steps.
Secondly, the method does not involve any sonication steps,
avoiding unnecessary damage to the phospholipid molecules.
Most importantly, it produces unilamellar vesicles with smaller
diameters and better monodispersity. While the film hydration
and ethanol injection methods can be easily adopted and per-
formed at the laboratory scale, they are labor-intensive and lack
scalability. The crossflow injection method was then developed
by Wagneretal. to streamline the process and execution of the
ethanol injection method.[] In the crossflow injection method,
the setup comprises of two stainless steel tubes welded per-
pendicular to one another with a small injection hole at the
intersection between the tubes. Through this injection hole,
lipid mixtures in ethanol are injected into a stream of aqueous
buer leading to the spontaneous self-assembly of lipids to
form liposomes. It was found that the introduction of high
lipid concentrations under high injection pressures leads to
liposomes with a narrow size distribution. It was suggested that
unlike the lipid film hydration and ethanol injection methods,
the crossflow injection method was suitable for continuous and
scalable manufacturing of liposomes. Overall, these bulk prep-
aration methods establish the guiding principles behind LNP
production and provide the basis for the development of more
advanced production platforms that are aimed to improve the
physicochemical properties of the LNPs.
2.3.2. Advent of Microfluidics
Microfluidics refers to the control and manipulation of fluids
within channels, cavities, and structural features with dimen-
sions in the micrometer range.[] Microfluidics have found rel-
evance in a wide variety of biomedical applications, including in
the preparation of gene and drug delivery vehicles.[] It allows
greater consistency in the preparation process since the flow
is well controlled by means of optimizing and fixing the flow
channel dimensions and flow rates of the introduced fluids,
thus leading to improved reproducibility. Besides, microfluidics
also increases throughput and opens the possibility of integra-
tion with digital automation.[] Within the context of LNP pro-
duction, microfluidics overcome challenges related to manual
experimental execution and batch processes. For example,
microfluidics allows controlled mixing of lipids in organic sol-
vent and nucleic acids in aqueous buer under continuous
flow, with minimal operator intervention. This contrasts with
the lipid injection method, which, although principally similar
in terms of solvent mixing, is a batch process that requires con-
tinuous stirring and subsequent removal of the ethanol via an
evaporation step.
Motivated by the growing promise of nanoparticles as an
alternative to viral vectors for the delivery of nucleic acids into
specific types of cells for gene therapy, Wang et al. designed
a rapid developmental pathway for generating nanoparticle-
based vectors for highly ecient delivery of a variety of nucleic
acid payloads for gene therapy applications.[] It relied on a
combinatorial synthetic approach based on supramolecular
assembly executed within a custom-designed digital microre-
actor to achieve a convenient, flexible, and modular method
for generating DNA-encapsulated supramolecular nanoparti-
cles. The employment of controlled microfluidic formulation
for the rapid discovery of potent siRNA-containing LNPs was
later reported by Anderson and co-workers.[] The method
involved the production of large amounts of siRNA–LNPs on
a microliter scale through stepwise ethanol dilution. Briefly,
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (12 of 38) © 2022 Wiley-VCH GmbH
an ethanolic solution containing the lipid excipients was first
rapidly mixed with an equal volume of aqueous solution con-
taining siRNA. This leads to lipid self-assembly into LNPs while
encapsulating the siRNA through electrostatic interactions. The
siRNA–LNPs were then further diluted with aqueous buer to
prevent LNP aggregation. These steps were performed within a
polydimethylsiloxane (PDMS) microfluidic channel, with peri-
odic trenches and ridges on the floor of the channel to promote
mixing. Around the same time, the group of Langer developed
a pattern-tunable microvortex platform for the mass produc-
tion and size control of lipid-polymer hybrid nanoparticles.[]
The platform involves the formation of symmetric microvor-
tices at the intersection of three inlets (i.e., two flanking inlets
introducing lipid and lipid-PEG dissolved in % ethanol/water,
and one central inlet introducing PLGA dissolved in acetoni-
trile). These microvortices promote the rapid mixing of the
lipid excipients with PLGA, followed by nanoprecipitation and
the formation of the nanoparticles. The microvortex formation
was regulated through careful prediction and manipulation of
the D fluid flow patterns. Using this approach, they reported
that the resultant nanoparticle size can be controlled with high
productivity and low polydispersity by adjusting the flow rates,
particularly by varying the Reynolds number between  and
. In a subsequent development, Valenciaet al. developed a
fully integrated microfluidic platform with a multi-inlet micro-
mixer for programmable and systematic mixing of large quanti-
ties of precursors to make nanoparticles.[] The platform was
successful in synthesizing nanoparticles with a wide range of
properties by combining  dierent precursors in dierent
ratios, achieving  dierent formulations with dierent sizes
and surface compositions. Taken together, the obvious advan-
tages of flow-based platforms oered by microfluidics within
the context of preparing nanoparticle-based delivery systems
have led to the realization of the importance of rapid mixing,
and further development of microfluidic mixing technologies
thus followed suit to expand its capabilities.[]
2.3.3. Microfluidic Mixing Architectures
Due to their nonpolar tails, lipids readily dissolve in ethanol
but not in aqueous solvent. The addition of lipids dissolved in
ethanol to an aqueous buer containing nucleic acids results
in the formation lipid bilayer structures upon hydration, which
subsequently interact with the nucleic acids to form small mul-
tilamellar liposomes with nucleic acids sandwiched between
the bilayers. Based on this principle, LNPs are formed when
there is a rapid increase in polarity within the bulk environ-
ment when two miscible phases combine (e.g., when lipids
dissolved in an organic solvent are rapidly introduced to an
aqueous buer). This occurs through the supersaturation of
lipid molecules, which leads to their self-assembly to form
LNPs[] (Figure4A). Along this line, it is important to note
that in order for the multilamellar bilayers to eciently trap
the nucleic acids, the ethanol concentration must be near the
region of breakdown of the liposomal structure.[c] In other
words, while increasing the concentration of ethanol leads to
higher nucleic acid entrapment, it also leads to a higher degree
of membrane destabilization. Hence it has been found that
dilution of the ethanol/buer mixture to around % ethanol is
typically ideal for the formation of LNPs with good encapsula-
tion eciency and sucient stability.
In microfluidic channels, the Reynolds (Re) number is
usually smaller than , which means that the concentration
of lipids in the organic solvent is dominated by molecular
diusion. Based on this principle, microfluidic mixing archi-
tectures have been developed with the aim of enhancing the
mixing eciency and attain a uniform population of LNPs
while achieving excellent entrapment of the nucleic acid
payload.[c,] Microfluidic mixing can be broadly categorized
into passive and active mixing; passive mixers solely rely on
the geometry of the microfluidic channels while active mixers
control the flow of sample solutions by electrohydrodynamic
disturbances. Within the context of LNP fabrication, most
micromixers can be categorized as passive mixers. Some
examples of advanced passive micromixers used for LNP
production include the T-junction mixer, staggered herring-
bone micromixer (SHM), hydrodynamic flow focuser (HFF),
bifurcating and bae mixers (FigureB). These micromixers
dier in their D structure and induce rapid mixing between
the organic and aqueous phases in a controlled environment
through dierent mechanisms.
T-Junction Mixer: T-junction mixing is a rapid mixing
method where input streams are introduced directly opposite
to each other at very high flow rates (i.e., – mL min),
with a perpendicular output.[] When the aqueous phase con-
taining nucleic acids and organic phase containing lipids are
introduced as inputs, the collision of the input flows results in
turbulent mixing, resulting in the formation of nucleic acid–
LNPs. It has been found that the particle size can be reduced
by increasing the flow rates. Likewise, increasing the flow rates
also resulted in significant improvements in the polydispersity
index (PDI), which is an indication of the size distribution.[a]
However, as T-junction mixing is operated at high flow rates, it
is not a preferred method especially in the laboratory where the
reagents for high throughput screening need to be conserved.
Hence, T-junction mixing is favored for large-scale production
of siRNA–LNPs.
Staggered Herringbone Micromixer (SHM): The staggered
herringbone micromixer architecture was introduced by
Stroocketal. as a passive mixing method for steady pressure-
driven flows in microchannels at low Reynolds number.[] It
comprises a series of asymmetric protrusions, which mix sol-
vents based on chaotic advection. Within the context of LNP
production, the characteristic diusion length between the
organic lipid and nucleic acid–aqueous streams is significantly
reduced. This allows controllable mixing to occur faster (i.e.,
in the millisecond scale) than the characteristic timescale for
lipids to aggregate, hence producing LNPs with uniform sizes.
Unlike T-junction mixing, microfluidic devices with SHM are
suitable for mixing small quantities of input solutions, hence
making them ideal for screening applications in the laboratory.
Nevertheless, scale-up can be achieved through parallelization
of microfluidic chips.[] For example, high production rates of
LNPs up to mL min have been achieved by incorporating
six sets of SHM in a single microfluidic device.[] Most notably,
SHM devices were commercialized by Precision NanoSystems
in their first-generation NanoAssemblr platform.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (13 of 38) © 2022 Wiley-VCH GmbH
Hydrodynamic Flow Focuser (HFF): Hydrodynamic flow
focusing (HFF) was first reported by Jahn et al. and involves
the continuous-flow hydrodynamic focusing of the organic
phase (i.e., lipids dissolved in organic solvent) by the aqueous
phase.[] Typically, a central stream introduces the organic
phase, which is pinched by two aqueous streams. These flows
are laminar, which results in a well-defined interface between
the two phases where interfacial forces can be carefully tuned
by adjusting the operating parameters. Two main operating
parameters are the total flow rate (TFR) of the two phases and
Figure 4. Self-assembly process of LNPs within microfluidic systems. A) The self-assembly process of nucleic acid–LNPs in a microfluidic mixer begins
when lipid excipients (i.e., ionizable lipid, PEG lipid, helper lipid, and cholesterol) dissolved in ethanol are rapidly mixed with the nucleic acid payload
(e.g., siRNA, mRNA, etc.) dissolved in an aqueous buer. This initially occurs at a low buer pH whereby the ionizable lipid will assume a protonated
form and bind, via electrostatic interactions, to the negatively charged backbone of the nucleic acid. Concurrently, the increasingly polar environment
drives the formation of vesicles and the encapsulation of the nucleic acid. As the mixing progresses, and with increasing buer pH, the ionizable lipid
becomes increasingly neutral, leading to the fusion of adjacent vesicles to form the interior core of the LNP. The extent of vesicle fusion is influenced
by the amount of PEG lipid added due to its hydrophilicity as well as steric eects. Adapted with permission.[81] Copyright 2021, MDPI Publishing.
B)Major microfluidic mixing architectures for LNP formulation (top to bottom): T-junction mixer, staggered herringbone micromixer, bifurcating mixer,
and bae mixer. C) The eect of varying flow rate on the mixing profile of ethanol (red) and water (blue) within a bae mixer, visualized through a
computational fluid dynamics simulation. Adapted with permission.[89] Copyright 2018, American Chemical Society. D) A diagrammatic representation
of an example of microfluidics parallelization, showing (from top to bottom) the arrangement of mixing channels, each mixing unit, and each mixing
cycle within a parallelized microfluidics device, which adopts the staggered herringbone architecture. Evaluation of the in vivo delivery performance
of E) siRNA–LNPs and F) mRNA–LNPs obtained using the parallelized device compared to bulk mixing. The evaluation was performed by means of
administering to mice. The mRNA encodes for luciferase and the in vivo delivery performance was evaluated by directly visualizing the luminescence
from selected regions of the mice. The results clearly show that the in vivo delivery performance of nucleic acid–LNPs obtained via microfluidic mixing
is superior to those obtained via bulk mixing. D–F) Adapted with permission.[85] Copyright 2021, American Chemical Society.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (14 of 38) © 2022 Wiley-VCH GmbH
the flow rate ratio (FRR) between the two phases. Both param-
eters influence the width of the central stream, which is repre-
sentative of the degree of hydrodynamic focusing. Since LNP
formation is based on a reduction in lipid solubility at the inter-
face between the two phases, the size and size distribution of
the resultant LNP are dependent on diusion, which is influ-
enced by the degree of hydrodynamic focusing. Specifically, it
was found that increasing the FRR can reduce the LNP size
while attaining narrower size distribution, and higher TFRs
resulted in larger LNPs only at low FRRs.
Bifurcating and Bae Mixers: Motivated by the commercial
success of SHM devices for the production of LNPs, Precision
NanoSystems later introduced the NxGen, which is a micro-
mixing architecture consisting of a sequence of bifurcating
mixers for scalable, nonturbulent mixing.[] Briefly, the organic
and aqueous phases containing lipids and nucleic acids, respec-
tively, are introduced from two pronged inlets that merge into
a single flow. Downstream, a series of toroidal mixers promote
chaotic advection to the merged stream. The stream initially
splits upon entry into the first toroidal mixer and travels dif-
ferent path lengths before merging again. This process, which
induces rapid mixing in a single layer by centrifugal force, is
repeated a few more times leading to the formation of nucleic
acid–LNPs with higher encapsulation eciency and reproduci-
bility compared to SHM. More importantly, it allows production
rates to be increased by up to -fold. A similar architecture
that operates based on a similar principle is the bae mixer.
In bae mixers, the organic and aqueous phases are also intro-
duced from two pronged inlets that merge into a single flow.
However, instead of having a series circular, toroidal mixtures,
it comprises sharp perpendicular turns to induce chaotic advec-
tion. Bae mixers can produce LNPs with controllable sizes in
the range of –nm, by varying the TFR, FRR, and device
dimensions (FigureC).[] Taken together, both bifurcating and
bae mixers are single-layer mixing architectures, which serve
as excellent alternatives to SHM and HFF for LNP production.
Each of the abovementioned approaches has their advan-
tages and limitations. For example, while the conventional
method of lipid film hydration followed by extrusion is easy to
perform, it is a time-consuming multistep production process
with relatively low encapsulation eciencies (–%). The
crossflow injection method is highly suitable and is already in
use for large-scale production in the industry but is less suited
for laboratory-scale research. On the other hand, microfluidic-
based approaches allow controlled mixing and generally result
in relatively high encapsulation eciencies. For example, the
T-junction and HFF methods can produce uniform LNPs,
has broad solvent compatibility and can achieve high encap-
sulation eciencies (%). The SHM method can produce
highly uniformed particles with a polydispersity index of less
than .. In addition, the encapsulation eciency of the SHM
method is very high (>%). However, due to the architecture
of the mixture, SHM is prone to clogging within the channel.
All microfluidic approaches require parallelization for scale-up
(FigureD). Nevertheless, parallelization has shown to improve
in vivo delivery for siRNA (FigureE) and mRNA (FigureF)
In conclusion, compared to conventional preparation methods,
the advent of microfluidics especially with the development
of microfluidic mixing architectures has oered better control
over the production process of LNPs leading to LNPs with
greatly improved physicochemical properties and encapsula-
tion eciencies. More importantly, microfluidics has paved the
way for more rapid screening of LNP formulations and scale-
up manufacturing, as well as the prospect of automated LNP
synthesis.[]
2.4. Applications
A better understanding of the structure–function relationship,
as well as improvements in the design principles and platform
development has enabled the exploration of LNPs for use in dif-
ferent clinical applications. Conversely, the promise of LNPs as
delivery systems in certain applications has driven the refine-
ment of their design principles and platform development. In
the following subsections, we will review how two main appli-
cations, namely RNAi therapy and mRNA-based vaccines for
COVID-, have led to the progress of LNPs.
2.4.1. RNAi Therapy
The development of LNPs was originally motivated by RNA
interference (RNAi) therapy.[] RNAi refers to a fundamental
cellular mechanism for silencing gene expression, which can
be harnessed for reducing the expression of pathological pro-
teins and therefore exploited as a novel therapeutic approach.[]
In brief, it relies on the specific binding and silencing of thera-
peutic targets by means of introducing siRNA molecules.[]
While the therapeutic mechanism is very promising since the
approach can reversibly silence any gene, the eective delivery
of siRNA proved to be challenging.[b,]
LNPs were developed as an alternative to viral vectors
and represent the most advanced non-viral vector for gene
delivery.[c,] Nevertheless, formulation optimization had to
be conducted rigorously to achieve maximum siRNA encap-
sulation and transfection eciencies.[b,] These studies had
been largely facilitated by the advent of microfluidics.[b,,,]
For example, Chenetal. demonstrated the utilization of con-
trolled microfluidic formulation leading to the rapid discovery
of potent siRNA–LNPs.[] In the industry, the success of RNAi
can be attributed to T-junction mixing, which enabled the pro-
duction of stable siRNA–LNPs that are capable of potent knock-
down of Apolipoprotein B (ApoB) in non-human primates.[b]
This is a prominent study as it represents one of the earliest
reports highlighting the eectiveness of RNAi in a large animal
study, revealing the significant reduction of ApoB mRNA and
protein even beyond  days. Separately, Merck & Co. also
adopted a similar approach by employing T-junction mixing
to produce siRNA–LNPs targeting the murine gene Ssb.[a]
Likewise, they were successful in reducing the target mRNA
levels in vivo by over %. Undoubtedly, microfluidic devices
have proven beneficial for formulation optimization within the
context of siRNA–LNPs, as they allow mixing of small amounts
of reagents, which is ideal for screening lipid structures with
minimal wastage. In addition, LNPs produced using micro-
fluidic devices are typically smaller, and more uniform, than
those produced via bulk preparation techniques. This has led
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (15 of 38) © 2022 Wiley-VCH GmbH
to the discovery of novel lipid structures for potent hepatic
gene silencing in vivo. In addition, with the ability to carefully
tune the mixing rates between the organic phase containing
the lipids and the aqueous phase containing the nucleic acids,
the size of the LNPs can be better controlled to understand the
eect of LNP size to gene silencing performance. For example,
the use of high flow rates to produce small LNPs corresponded
to better gene silencing capability compared to the use of low
flow rates, which produced large LNPs. All these advances are
critical in the eort to overcome outstanding challenges in the
delivery of LNPs. For example, while an siRNA–LNP formula-
tion has been established and developed into an FDA-approved
prescription medicine (i.e., known as patisiran and marketed as
ONPATTRO, NCT) to treat polyneuropathy in people
with hereditary transthyretin-mediated amyloidosis, clinical
trials led by Roche/Ionis Pharmaceuticals (NCT) and
WAVE Life Sciences/Takeda (NCT and NCT)
to explore similar approaches with LNP candidates for treating
Huntington’s disease are ongoing.
2.4.2. mRNA Vaccine for COVID-19
The translation of mRNA–LNPs to the clinic had been recently
realized in the form of mRNA vaccines against COVID-
.[a,a,] This development not only highlights the benefits of
mRNA-based vaccines but also establishes the biocompatibility
of LNPs as a delivery vehicle.[] While the success accelerated
by the COVID- pandemic was unprecedented considering
there had been only one FDA-approved LNP-based siRNA drug
(i.e., ONPATTRO) prior to that,[] it was not entirely surprising
considering an extensive amount of eort has been put in to
understand the structure-function relationship and improve
the performance of LNPs over the last decade. Such collective
eorts have facilitated the quick adoption of LNPs, which was
initially largely developed for siRNA, to incorporate mRNA pay-
loads instead. As discussed in the previous sections, the evolu-
tion of design principles and platform development play critical
roles in incrementally improving the design and manufacturing
of potent LNPs, culminating in the maturity of the LNP-based
delivery systems that meet the requirements of good manufac-
turing practice and safety. Key milestones that contributed to
the realization of LNP-based mRNA vaccine include the intro-
duction of ionizable lipids as well as the development of micro-
fluidics and advanced micromixing architectures. In particular,
microfluidic platforms that oer high degrees of parallelization
for large-scale manufacturing (e.g., Precision NanoSystems
NanoAssemblr platform) have played a pivotal role in ensuring
the clinical success of mRNA–LNPs. In addition to the avail-
able COVID- vaccines (NCT, NCT, etc.),
other LNP–mRNA vaccines against influenza (NCT
and NCT), Zika virus (NCT), rabies virus
(NCT), and multiple other viruses are currently under-
going human clinical trials.
With the establishment of such platforms, LNPs containing
mRNA as well as other types of nucleic acids continue to be
explored for other therapeutic applications and following
other routes of administration.[] For example, Dahlman and
co-workers recently optimized the formulation of LNPs so that
mRNA can be delivered to the lungs in the form of nebulized
aerosols.[b] Based on all the data gathered via a cluster-based
screening approach, they found that PEG lipids are critical in
attaining stable LNPs. Furthermore, nebulized mRNA delivery
can be improved with the use of cationic helper lipids and a
higher molar fraction of PEG, and more PEG is needed for for-
mulations with cationic helper lipids than those with neutral
lipids. The resultant LNPs optimized following their approach
exhibited greater structural stability after nebulization, as
opposed to LNPs designed for systemic delivery. In essence,
their work demonstrated the possibility of adopting nucleic
acid-based LNP delivery systems for developing therapeutic
application following dierent routes of administration, at the
same time revealed that formulations developed for systemic
delivery cannot be simply adopted for inhalation, which under-
scores that the importance of formulation optimization.
3. Lipoplex
While LNPs are fabricated by direct coassembly of ionizable
lipids, helper lipids, PEGylated lipids, sterols, and nucleic acids
during solvent mixing, another intriguing lipid nanoarchi-
tecture for delivery applications involves the aqueous-phase
mixing of preformed liposomes with nucleic acids, termed
“lipoplexes.” The lipoplex strategy evolved from early eorts[]
to encapsulate nucleic acids into liposomes composed of ani-
onic and zwitterionic lipids through passive loading, which
faced challenges such as low eciency in terms of nucleic acid
entrapment and delivery.[a,] Although negatively charged
liposomes may reduce the leakage of contents with increased
stability and protection, inecient entrapment (ranging
%–% with the highest eciency obtained by the freeze/
thawing procedure[a]) due to the repulsion between the mem-
brane and nucleic acid was the major concern of encapsula-
tion method.[] To enhance interactions between liposomes
and nucleic acids, cationic lipids have been utilized in order to
facilitate electrostatic attraction with negatively charged nucleic
acid molecules.[] As a result, passive loading of nucleic acids
into liposomal interiors is not required, and instead electro-
static interactions between liposome surfaces and nucleic acids
spontaneously drive lipoplex self-assembly. The first lipoplex
demonstrations utilized liposomes composed of cationic and
zwitterionic lipids that were mixed with pDNA[] or mRNA[]
and exhibited superior transfection eciency (especially with
RNA transfection yielding - to -fold more) with less
cytotoxicity compared to a polymer-based delivery vehicle, a
diethylaminoethyl (DEAE)-dextran complex, that was tested in
parallel.
Motivated by these pioneering studies, lipoplexes have
emerged as a powerful platform technology that can potentially
take a single liposome formulation and mix it with dierent,
customizable nucleic acids to create distinct lipoplex options for
various applications. Indeed, while LNPs incorporating ioniz-
able lipids have demonstrated high practical utility for mRNA
vaccine applications, lipoplexes have reached human clinical
trials for applications such as cancer therapy.[] Even so, there
continue to be key issues related to the cationic lipid design
along with paired neutral lipid choice, processing parameters,
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (16 of 38) © 2022 Wiley-VCH GmbH
and the corresponding influence on the nanoparticle’s architec-
ture and delivery functionality. In this section, we cover these
design points and also discuss selected cases that directly com-
pare the performance of LNPs and lipoplexes.
3.1. Nanoarchitecture Design Principles
3.1.1. Lipid Composition
The composition of a lipoplex is defined by two lipid types:
) cationic lipid and ) neutral (helper) lipid. As with LNPs, in
lipoplexes, cationic lipids mainly interact with nucleic acids while
neutral lipids influence self-assembly properties and assist with
conferring particle stability. A noteworthy distinction is that the
majority of cationic lipids used in lipoplex formation are perma-
nently charged, meaning that pKa is larger than  (or without
pKa), while a few cases feature slightly ionizable lipid-like charac-
teristics (e.g., DC-Chol, DODMA, multivalent cationic lipids, etc.
that are discussed below). Hence, the stability and eciency of
the complex are highly dependent on the neutral lipid choice and
molar charge ratio (N/P, cationic lipid-to-nucleic acid ratio). It
is also worthy to note that the transfection eciency might vary
depending on the cell types, and there are several comprehensive
studies comparing commercially available transfection reagents
such as Lipofectamine, RNAiMAX, and Lipofectin.[]
Cationic Lipid: Cationic lipids incorporated in lipoplexes
should be able to capture the nucleic acid, maintain stability
before entry, facilitate cellular entry (via fusion or destabili-
zation), and subsequently release nucleic acids in the target
cell.[] Lipid structures can be divided into three functional
units consisting of a hydrophilic headgroup, hydrophobic tail,
and linker, connecting the former two moieties. Modifying
each element can aect the stability (compaction with nucleic
acid), disassembly, and phase behavior of the lipoplex, which
have been extensively discussed.[] Depending on the tail
group, cationic lipids can be categorized into aliphatic chains or
cholesterol-based derivatives while the majority of headgroups
are composed of permanently positively charged, quaternary
ammonium or amines and their derivatives (Figure5).
The first cationic lipid used to form a DNA lipoplex
was ,-di-O-octadecenyl--trimethylammonium propane
(DOTMA), which is a double-chain quaternary ammonium
with ether bonds. One of the major commercial transfection
reagents, Lipofectin, is based on DOTMA/DOPE (/ mol%)
composition and is recommended to deliver nucleic acids to
endothelial cells.[] Since then, several other related com-
pounds with aliphatic chains were developed to improve trans-
fection eciency and reduce cytotoxicity. ,-dioleoyl--trimeth-
ylammonium-propane (DOTAP) with ester bonds,[] dimethyl-
dioctadecyl ammonium bromide (DDAB), and single-chain
lipid, cetyltrimethyl ammonium bromide (CTAB), were subse-
quently utilized to fabricate lipoplexes while the most widely
used cationic lipids have remained DOTMA and DOTAP.
Based on these lipids, various derivatives were also developed
by replacing the methyl group with the hydroxyethyl group,
leading to lipids such as ,-dioleoyl--dimethyl-hydroxyethyl
ammonium bromide (DORIE) and ,-dioleoyloxypropyl--di-
methyl-hydroxyethyl ammonium chloride (DORI) that demon-
strated enhanced compaction of nucleic acids with lipids.[b,b]
Aside from dioleoyl derivatives, dimyristoyl compounds were
recognized later to display higher transfection eciency, leading
to the development of another commonly used cationic lipid,
,-dimyristyloxypropyl--dimethyl-hydroxyethyl ammonium bro-
mide (DMRIE), which is currently used in clinical-stage lipoplex
formulations together with DOTAP and DOTMA.[]
Figure 5. Lipoplex components and nanoarchitecture organization. Schematic illustration of the lipoplex and molecular structures of the cationic lipids
and neutral lipids.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (17 of 38) © 2022 Wiley-VCH GmbH
Furthermore, multivalent cationic lipids have gained atten-
tion by showing higher transfection eciencies due to their
ability to condense nucleic acids more tightly and to par-
tially ionize to act as a buer to better protect the nucleic
acid from degradation compared to their monovalent coun-
terparts.[] For example, another widely used transfection
agent, Lipofectamine, is based on polycationic ,-dioleyloxy-
N-[(sperminecarboxamido)ethyl]-N,N-dimethyl-l-propana-
minium trifluoroacetate (DOSPA) and DOPE at a .: molar
ratio and is often considered a benchmark composition for
evaluating transfection eciency.[] Another advantage mul-
tivalent lipids might confer is their high eciency to capture
nucleic acid with a lower amount of lipid (lower N/P) compared
to the monovalent ones.[]
On the other hand, cholesterol derivatives such as
β-[N-(N,N-dimethylaminoethane)carbamoyl]-cholesterol
hydrochloride (DC-Chol) are another popular class of cationic
lipids used in lipoplexes.[] As cholesterol itself is commonly
used as a neutral, helper lipid in lipoplex formation to increase
membrane packing by embedding its bulky steroid moiety
between hydrophobic tails of neighboring lipids, the addi-
tion of a cationic moiety in cholesterol derivatives was found
to exhibit superior performance, enabling delivery of nucleic
acids into various cell types in vitro and in vivo.[b,] Also,
DC-Chol contains a carbamate linker and a tertiary amine with
a relatively low pKa value (% charged in buer[]) compared
to other permanently charged lipids, providing higher biodegra-
dability and aiding the ecient release of nucleic acid.[] Due
to these advantages, a DC-Chol and pDNA lipoplex formulation
has been shown to exhibit low toxicity upon local or systemic
administration and has entered clinical trials for the treatment
of genetic disorders and cancer along with its multivalent deriv-
ative with spermine, N-cholesteryl-Spermine (Genzyme lipid
, GL).[b,]
Typically, ionizable lipids used in lipid nanoparticle formula-
tion possess apparent pKa less than  in order to exhibit posi-
tive surface charges at acidic (endosomal) pH while neutral
charges at physiological pH. Abovementioned DC-Chol and
-dioleyloxy-N,N-dimethyl--aminopropane (DODMA) possess
pKa value around , oering pH-dependent nucleic acid and
lipid interactions.[,] For example, DODMA molecules rear-
range toward the hydrophobic region of the bilayer at high pH,
oering an increased flip-flop rate that might aect the fusion
process.[] As ionizable lipids may oer more ecient endo-
somal escape by membrane rearrangement upon contact with
anionic lipid membrane, there were several eorts to discover
ecient ionizable lipids that can be used without a neutral lipid,
leading to a development of a series of aminoglycoside deriva-
tives that exhibit higher stability and smaller mean diameters
compared to cationic lipid-based lipoplexes.[] Other eorts
include an ionizable lipid based on vitamin E-scaolds[] and
α-amino-lipophosphonates to target dendritic cells.[]
Neutral Lipid: Neutral lipids have many useful functions as
part of lipoplex compositions, including assisting with forma-
tion-related phase changes and reducing interparticle aggrega-
tion. While certain cationic lipids forming lipoplexes without
neutral lipids can be functional to deliver nucleic acids, appro-
priate inclusion of neutral lipids such as DOPE and/or choles-
terol can significantly improve the stability and the transfection
activity of lipoplex.[] Thus, neutral lipids in lipoplex are also
called “co-lipids,” functioning similarly to “helper lipids” in
LNP excipients. Interestingly, the choice of neutral lipids might
significantly enhance or hamper the transfection eciency.
For example, DOTMA,[,b] DOTAP,[] and DC-Chol[]
have shown that the incorporation of DOPE largely augmented
transfection eciency whereas DOPC hampered the nucleic
acid delivery both in vitro and in vivo.[] These propensities are
closely related to the geometry of lipids and in turn, the archi-
tectural configuration of lipoplex that will be more discussed in
the next Section...
Cholesterol is another widely used neutral lipid[] aiding col-
loidal stability of lipoplex and participating in membrane fusion
by promoting contact sites for lipid mixing and expanding the
pore induced by fusion.[] Also, its inclusion in lipoplex has
shown higher resistance to serum-induced aggregation and
protein binding, which may explain enhanced transfection.[]
Indeed, a quantitative analysis performed by replacing DOTAP
lipids with cholesterol (or DOPE) showed reduced anity
of negatively charged plasma proteins that might be respon-
sible for protein corona formation.[] Interestingly, the choice
between cholesterol and DOPE may vary depending on the
tail group of cationic lipid.[] When various chain lengths of
lipids were tested with an equimolar ratio of either cholesterol
or DOPE inclusion, the highest eects were observed with sat-
urated  carbon or  carbon chains, respectively, while both
neutral lipids boosted the transfection eciency compared to
cationic lipid alone.[] Therefore, it should be noted that the
selection and optimum mole fraction of cationic and neutral
lipid depend on the lipid types, possibly N/P charge ratio, types
of nucleic acid, and cells.
3.1.2. Architectural Configuration
The architectural configuration of lipoplex is closely related to
the phase behavior governed by several factors such as types of
lipids, the composition of cationic lipids, and N/P charge ratio.
Despite recent clinical achievements, the structure–activity
relationship of lipoplex continues to be unraveled and func-
tional optimization is often based on empirical approaches.
To improve clinical performances, it is vital to understand the
physicochemical properties aected by dierent parameters.
The supramolecular structure of lipoplexes depends on the
participating components’ intrinsic properties such as their
geometry and other environmental (extrinsic) factors as well as
the balance between electrostatic interaction (headgroup) and
repulsive force driven by hydrophobic domain upon the addi-
tion of nucleic acid.[]
The most common structures are lamellar phase (
α
C
L
) and
inverted hexagonal phase ( II
C
H
), which can also be referred to
external and internal model, respectively, depending on the
relative orientations of nucleic acid (Figure6A).[] Safinya pio-
neered early studies (and continue to explore more) of supra-
molecular structures of lipoplexes.[] Initiated by applying
the small-angle X-ray scattering (SAXS) technique to a mix-
ture of two cylindrical-shaped lipids, DOTAP/DOPC lipoplex,
a lamellar phase composed of multiple sheets of bilayers was
discovered (Figure B).[] SAXS measurements show that
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (18 of 38) © 2022 Wiley-VCH GmbH
the DNA monolayers are sandwiched between cationic mem-
branes with an interlayer spacing of .nm (nm of bilayer
and . nm of water spacing with intercalated DNA) in an
equimolar fraction case, and the spacing varied with the molar
fraction of lipid molecules as the bilayer thickness changed
due to the dierent length between DOPC and DOTAP.[]
Interestingly, the spacing between DNA was relevant to the
membrane charge density as indicated by the arrows: the
spacing increased as the membrane charge density decreased
(increasing DOPC fraction). A cryo-EM image in the absence
of DNA shows enclosed liposomes with an equimolar composi-
tion of DOTAP/DOPC (Figure C).[] When the nucleic acid
is added, these liposomes transform into a layered structure to
accommodate nucleic acid molecules.[]
Soon after verifying lamellar phase, the same group also
identified inverted hexagonal phase (
II
C
H
) when the inverted
cone-shaped DOPE was included (Figure D).[] With
increasing DOPE mole fraction in DOTAP/DOPE lipoplex from
. to ., SAXS scans showed that the internal structure of
the DOTAP/DOPE lipoplex changes from
α
C
L
to something else,
which was defined as II
C
H
with a diameter of .nm embed-
ding DNA within a lipid monolayer arranged in a hexagonal lat-
tice (FigureE).[] Of note, when much smaller-sized siRNA is
entrapped instead of DNA, the peaks in SAXS may appear less
obvious (meaning less orientation) because of the higher disor-
dered phase.[] Cryo-EM analysis also showed that the distinct
feature of equimolar DOTAP/DOPE with the coexistence of
α
C
L
and
II
C
H
(FigureF).[]
Although it is well known that particular
II
C
H
complexes
(DOTAP/DOPE = .) successfully delivered DNA, whereas
similarly high fraction of DOPC-based lipoplex failed to deliver
nucleic acids,[] it should be noted that the architectural con-
figuration of lipoplex and biological functionality cannot be
directly correlated.[] In other words, lipoplexes with
do
Figure 6. Architectural configuration of lipoplexes. A) Schematic of the lamellar phase (
C
α
L
) of DOTAP/DOPC lipoplex. B) SAXS scans of lipoplex at
constant N/P charge ratio of 2.2 with increasing DOPC ratio. The arrows indicate the DNA peaks, whereby the spacing between DNA increase as the
membrane charge density decreased. C) Cryo-EM images of DOTAP/DOPC lipoplex without/with nucleic acids. D) Schematic of the inverted hexagonal
phase ( II
C
H
) of DOTAP/DOPE lipoplex. E) SAXS patterns of the C
α
L
and II
C
H
of lipoplexes with increasing DOPE ratio. F) Cryo-EM images of DOTAP/DOPE
lipoplex without/with nucleic acids. G) Coarse-grained molecular dynamics simulations to visualize the fusion mechanism between lipoplexes and
endosomal membrane: i) free lipoplex complexed with dsDNAs approaches the endosomal bilayer; ii) stalk formation upon initial contact; iii) hemifu-
sion diaphragm; iv) membrane fusion with successful delivery of nucleic acids. H) Schematic of the hexagonal phase ( I
C
H
) of lipoplex. B,E) Adapted with
permission.[139] Copyright 2001, Elsevier. C,F) Adapted with permission.[130] Copyright 2001, American Society for Biochemistry and Molecular Biology.
G) Adapted with permission.[146] Copyright 2020, eLife Sciences Publications Ltd.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (19 of 38) © 2022 Wiley-VCH GmbH
not always show higher transfection eciency compared to the
ones with
α
C
L
, thus the phase is not a universal parameter to
predict transfection eciencies.[] Also, the existence of serum
can aect the outcome, in which membrane fusion is not a
dominant factor, and lipoplexes with
α
C
L
phase outperformed the
others.[] Some studies have reported that the morphological
changes or phase transitions from
α
C
L
to II
C
H
upon lipoplex–cell
membrane contact trigger better transfection eciency, and a
series of studies have attempted to optimize the cationic lipids
by tuning their molecular structures.[]
More recently, the interaction of lipoplex and negatively
charged endosomal membrane has been computationally simu-
lated, interrogating key parameters such as lipid types (dierent
headgroups and the extent of chain length and saturation),
endosomal membrane composition, and sizes of lipoplex.[]
By using this approach, two alternative fusion pathways were
suggested (Figure G, perpendicular and parallel) with a pos-
sible endosomal membrane composition dependency. Also, an
unsaturated chain in a cationic lipid was found to be necessary
for eective membrane fusion. It was also simulated that larger
lipoplexes ( nm) confer more thermodynamically stable self-
assembled structures than smaller ones ( nm). For example,
the highest amount of nucleic acid delivery was achieved for
the unsaturated DOTAP/DOPE lipoplex compared to its satu-
rated counterparts, which can be explained by the decreased
bending modulus of unsaturated lipids to favor the fusion.
About a decade after the
II
C
H
report, continuous eorts
revealed another representative phase called hexagonal ( I
C
H
)
structure based on dendritic cationic lipid/DOPC lipoplex that
was highly ecient in in vitro transfection (FigureH).[] Due
to highly charged (+) bulky headgroups of dendritic lipid,
the complex forms long micelle-like structure to surround the
nucleic acid. The improved eciency was ascribed to the e-
cient packing of nucleic acid within lipoplex that can facilitate
easier release of nucleic acid in cells. More recently, a concept
of cuboplex has emerged, utilizing bicontinuous cubic phases
(cubosome) made of glycerol monooleate (GMO), DOTAP, and
GMO-PEG (// mol%).[] As marginal amount of cationic
lipid inclusion suggests, the endosomal membrane interac-
tions of cuboplex is not governed by electrostatics but by elas-
ticity energetics, and possess intrinsic fusogenic properties to
aid endosomal escape. Although cuboplex is largely dierent
from conventional lipoplex composition, the development and
structural characterization of this novel vehicle provide valuable
insights how nanoarchitectonic approach can contribute to the
advancement of lipid-based gene delivery technologies.
3.2. Platform Development
3.2.1. Lipoplex Manufacturing Considerations
One of the biggest advantages of lipoplex comes from the
simplicity and flexibility of its fabrication, which takes two
steps: ) formation of cationic liposomes and ) complexation
with nucleic acid (Figure7A). Liposomes can be prepared via
traditionally known liposome production methods, ranging
from simple rehydration of lipids, sonication, or ethanol injec-
tion; sophisticated monodisperse ones can be formulated by
extrusion and microfluidics. In this section, a few considera-
tions while fabricating lipoplex (cationic liposome fabrication
followed by nucleic acid complexation), their eects on func-
tional activities, and the importance of N/P charge ratio are
critically discussed.
3.2.2. Liposome Fabrication
Liposomes are sphere-shaped lipid vesicles consisting of one
(unilamellar) or more (multilamellar) lipid bilayers. When
amphipathic lipids are hydrated, hydrophilic headgroups face
aqueous phase whereas hydrophobic tail groups orient or aggre-
gate towards each other to form thermodynamically favored
self-assembly structure. Therefore, the simplest liposomes can
be fabricated by mechanically dispersing the lipid suspension in
aqueous medium, which can be done by shaking or vortexing.
Largely, the preparation of methods can be divided into
mechanical and organic solvent-assisted methods. Most of
mechanical preparations start from lipid thin-film hydration
as mentioned above, then subjected to further processing
including sonication, freeze-thawing, and/or extrusion. Among
them, sonication is the most used method to prepare liposomes
for lipoplex formation even though its size distribution can be
relatively larger than other processing methods.[] From the
first reported lipoplex fabrication to recent optimization studies,
sonication has been widely used and proven eective even com-
pared to other methods.[,b,] For more homogenous and
smaller unilamellar liposome production, freeze-thaw and/or
extrusion can be applied after lipid hydration or after mild soni-
cation. Freeze-thawing is adopted commonly to improve the
encapsulation eciency of molecules by expanding the inner
water phase by ice crystal formation, which result in high pop-
ulation of unilamellar vesicles while physically disrupting the
lamellar structure.[] Extrusion can be accompanied to yield
more homogenous particles, often using nm filter to pro-
duce cationic liposomes. Ethanol injection is another widely
used method to fabricate liposomes, which can be versatile and
easy to scale up. In this method, the lipid–ethanol solution is
injected into the aqueous phase (or vice versa), then ethanol
is removed by rotary evaporation or dialysis to produce con-
centrated liposome suspensions.[] The microfluidic method
is a sophisticated improvement of ethanol injection that the
mixing of organic and aqueous phases can be controlled by the
interfacial diusion of narrow microchannel, which has been
described in the previous section. In lipoplex formation, micro-
fluidics can be used in either fabricating initial liposomes or
mixing nucleic acid with liposomes, which will be discussed in
the next section.
Multiple studies have compared the transfection eciency
of lipoplex formulated from dierent methods, whereby the
reason of one’s performance is debatable whether it is due
to the particle size and/or lamellarity, or other factors. Early
studies have reported that the transfection eciency is higher
in large multilamellar liposomes (– nm) formed by
vortexing rather than in relatively smaller sized liposomes
(< nm) formed by sonication (FigureB).[b,,] Similar
conclusion was drawn when the transfection eciency of lipo-
plex was tested with or without serum as well.[] On the other
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (20 of 38) © 2022 Wiley-VCH GmbH
Figure 7. Platform development of lipoplex. A) Schematic illustrations of lipoplex formulation from cationic liposome formation to addition of nucleic
acids. B) Comparison of transfection eciency depending on the liposome preparation method. Vortexing and sonication were used to produce dominantly
large multilamellar (red) and smaller (blue) liposomes, respectively. C) One-step formation of lipoplex assisted by a microfluidic device. The graph (right)
shows the hydrodynamic diameter and polydispersity index of microfluidic-obtained and bulk mixing-prepared lipoplexes. D) Schematic representation to
show that the N/P charge ratio largely aects the colloidal stability of lipoplexes, which can be presumed by their particle size (green line) and net surface
charge (orange line). E) Lipoplexes formulated at various N/P ratios display a distinct instable range around N/P = 1, whereby the zeta potential is close
to neutral. F) Bioluminescence imaging of BALB/c mice after intravenous administration of Luc-RNA-lipoplexes from panel E) at various N/P charge
ratios. B) Adapted with permission.[135] Copyright 2010, Springer Nature. Panel C schematic is adapted with permission.[173] Copyright 2017 Elsevier. Panel
C graph is adapted with permission.[165b] Copyright 2016 American Chemical Society. E,F) Adapted with permission.[169] Copyright 2016, Springer Nature.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (21 of 38) © 2022 Wiley-VCH GmbH
hand, ethanol injection method was reported to be equally or
more eective than the dry-film method[] and its eciency
proven in human clinical studies.[d,]
3.2.3. Nucleic Acid Complexation
During the formation of lipid-nucleic acid complex, the
nucleic acid electrostatically binds to the cationic lipid by fast
exothermic process, then subsequently rearrange/fuse into
lipoplex by slower endothermic reaction.[] While the first fast
regime occurs instantly when the lipid and nucleic acid con-
tact, the kinetics of slow rearrangement and fusion-induced
growth of lipoplexes largely depend on the mixing protocol.
The mode of mixing can vary by the sequence of adding com-
ponents and whether the addition is bulk (one-step mixing by
pipetting or vortexing) or titration (more than two steps). In
permanently charged cationic lipid, the complexation solution
conditions are generally maintained at physiological pH or in
deionized water to facilitate strong electrostatic interactions
between nucleic acid and lipids. Whereas excess electrolytes
can shield and decrease the driving force of self-assembly of
lipoplex when there was significant dierences (. and
. ),[] a direct link between transfection eciency and
solution conditions during the formation of lipoplex (mild salt
concentrations or deionized water) remains controversial.[b]
However, in the cases of pH-responsive lipids such as DC-Chol
and DODMA, slightly acidic or neutral buer solution is pre-
ferred over deionized water to increase the payload and trans-
fection eciency.[]
One of the first studies comparing the order of addition of
nucleic acid and lipids was reported by Zelphatiet al. that the
small, stable lipoplexes could be formed at high N/P charge
ratio (which will be explained in the next section) by bulk-
adding the nucleic acids into the excess of liposomes while the
opposite order produces large aggregates.[] At low N/P charge
ratio, the trend was reversed such that the addition of liposomes
to the excess of nucleic acid could produce stable lipoplexes.
This observation could be explained by titration approach.[]
During the nucleic acid addition into liposomes (in excess),
nucleic acids come into contact with multiple liposomes, asso-
ciating at the interface of two (or more) liposomes at once. This
interaction may generate asymmetrical stresses, leading to rup-
ture and aggregation of the lipoplex. In the opposite case when
the liposome is added, the excess nucleic acids rapidly bind and
wrap the surface of the cationic liposomes, forming a relatively
stable complex until a critical point where liposomes bind to
other ones that are already coated with nucleic acid. Then, the
aggregates start to form, and the generated stress drives the
rupture and growth of the particles. The former case might pro-
duce too large particle that can limit the transfection eciency,
but the latter case could inhibit the intimate lipid mixing and
necessary structural rearrangements.[a,] Likewise, small
variations in each step can result in largely dierent self-assem-
bled structure, thus care should be taken during formulation
and optimization of lipoplex preparation. Of note, an optimal
mixing method might vary depending on the nucleic acid types
and concentration as well. For example, vortex-mixing resulted
in better transfection eciency than spontaneous mixing
when the siRNA concentration was higher than a certain point
whereas the opposite was true when N/P charge ratio was
high.[]
After presenting that the order of mixing can be a critical
parameter influencing the complex behavior, Zelphatietal. from
the same study proposed a controlled mixing method to mix
both components simultaneously at a fixed ratio/speed with the
aid of a syringe pump.[] The optimal mixing rate was higher
for the smaller liposomes, whereby the results were reproduc-
ible and more homogenous lipoplexes could be formed. This
controlled mixing was further developed recently via micro-
fluidics to formulate the liposome as well as to complex with
nucleic acids.[] This method has demonstrated potential for
controlling physical properties of liposomes including size dis-
tribution and lamellarity in precisely confined microenviron-
ment. Compared to the bulk mixing method, microfluidics
generated lipoplexes with more homogenous and smaller sizes
(Figure C).[a,] This feature might explain higher transfec-
tion eciency in some cases, whereby about half number of
stacked bilayers were needed (reduced amount of cationic lipid)
to incorporate the same amount of nucleic acid.[b]
Lastly, it is worthy to note that generalizing the eect of lipo-
plex size by comparing dierent literature should be avoided
because multiple parameters involve in the final resulting
aggregates such as lipid compositions, N/P charge ratio,
nucleic acid, and reported transfection eciency might vary a
lot depending on the cell lines as mentioned above.[c] Nev-
ertheless, large lipoplexes (up to a certain extent) feature wide
range of contact area with cells and tend to form large intra-
cellular vesicles after endocytosis, which are more susceptible
to release the nucleic acid.[,] Consequently, the transfection
eciency also depends on whether a particular cell dominantly
uptake particles by endocytosis.[] While most of the men-
tioned studies were conducted in vitro, the physicochemical
properties of lipoplex might be more pronounced in the in vivo
experiments because cellular uptake of nanoparticles is strongly
dependent on their size and surface charge.[] Generally, the
mean diameter of lipoplex particles that have entered clinical
trials ranged –nm.[b]
3.2.4. Charge Ratio (N/P)
The charge ratio (N/P) denotes the molar ratio between cati-
onic moieties, such as protonatable nitrogen (N) atoms of
the lipid headgroup, and the anionic charges brought by
the phosphate (P) groups of the nucleic acids. Theoretically,
N/P = / will result in the net charge of  but usually this point
can be achieved empirically by a slight excess of either negative
(nucleic acid) or positive charges (cationic lipid) due to steric
hindrance and/or other geometrical constraints.[] As N/P is
the most easily tunable and dominant factor that directly aects
the colloidal stability of the complex, examining its eect on
colloidal stability is a prerequisite to optimize lipoplex formula.
From a functional perspective, too low N/P increases uncom-
plexed nucleic acid,[] but too high N/P leads to increased cyto-
toxicity due to excess cationic lipids as mentioned above. With
respect to stability, there are three main regimes depending on
the N/P (Figure D,E). Starting from higher N/P (low amount
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (22 of 38) © 2022 Wiley-VCH GmbH
of nucleic acid), most particles are positively charged and repel
each other because the population is dominantly cationic with
a small part of lipoplexes. Upon adding more nucleic acid, lipo-
plexes become near to neutral surface charge (see zeta poten-
tial measurement) such that they start to aggregate because of
colloidal instability.[] With increasing nucleic acid, the lipo-
plex reaches again colloidally stable structure that is negatively
charged with some uncomplexed or not suciently condensed
nucleic acid molecules surrounding the particle.
The optimal N/P charge ratio might vary depending on the
lipid choices and respective applications, ranging from >
for monovalent cationic lipids such as DOTMA to less than 
for multivalent GL or several DNA-complexed cases.[b,]
Although negatively charged regime is often regarded as less
eective considering the anionic cell membrane,[a] in vivo
studies have reported that Luc expression was predominant
in the lungs of mice when the particle is positively charged
whereas the gradual increase of nucleic acid can shift the tar-
geted organ to the spleens (FigureF).[,] Interestingly, not
only the lead composition of DOTMA/DOPE, but also other
combinations of lipids including cholesterol and DOTAP have
led to exclusive targeting of spleen when the lipoplex was nega-
tively charged, prompting the selection of N/P = . condition
for clinical testing.
To conclude the manufacturing considerations with per-
spective of scalable production, lipoplex-based gene delivery
features a simple preparation step for liposome and a flexible
choice of nucleic acid that can be complexed without aids of
sophisticated devices or specific skills. All liposome produc-
tion methods have been established in depth, and fabrication
methods except for microfluidics are quite straightforward and
widely available, allowing us to scale-up the liposome manu-
facturing and to easily combine with commercially established
protocols for nucleic acid production. Therefore, the flexibility
regarding nucleic acid choice comes as a major advantage in
utilizing lipoplex in “on demand” situations such as imminent
viral outbreaks or personalized production of gene delivery
vehicle.[a] Furthermore, the qualities of liposome and nucleic
acid can be separately screened prior to complex formation,
which can minimize the contamination and enhance the batch-
to-batch reproducibility with lower cost.[]
3.3. Current Applications and Future Possibilities
3.3.1. Cancer Vaccines
There has been intense interest in developing lipoplex as a vac-
cine that stimulates targeted immune response to treat cancer.
Toward this goal, antigen-presenting cells such as dendritic
cells (DCs) are ideal targets of lipoplexes to boost the immune
response. Among the pioneering works, intravenously admin-
istered nonfunctionalized DOTAP/DOPE lipoplex has been
developed for systematic targeting of DCs.[a,] The admin-
istration of lipoplex induced both adaptive and innate immune
responses via antigen-encoding RNA.[] Importantly, this
study highlights that surface charge-dependent organ-targeting
is achievable by simply tuning the N/P charge ratio. The site-
specific RNA expression from lungs to spleen was achieved
when the N/P was – and .–., respectively (FigureF),[]
which is distinguishable from targeting function of neutral
LNP mainly to hepatocytes.[]
Additionally, lipoplex has been used as a vaccine specifically
for treating melanoma with various nucleic acids such as pDNA
or RNA.[d,b,] For example, multiple nonmutated, tumor-
associated antigens were targeted by intravenously administered
melanoma FixVac (BNT-mRNA)-DOTMA/DOPE lipoplex,
leading to the elevated cytokines such as IFN-α, γ, IL- with
ex vivo CD+ T-cell responses across dierent cancer types.[]
On the other hand, OVA-encoding mRNA–DOTAP/DOPE
lipoplex induced a potent type I IFN response upon subcu-
taneous, intradermal, and intranodal injection, interfering
with the generation of potent cytolytic T-cell responses.[]
When the antitumor eect of mRNA–DOTAP/DOPE lipo-
plex was compared with pDNA–lipoplex, mRNA–lipoplex
showed higher lipofection eciency than pDNA one.[]
Thus, mRNA–DOTAP/DOPE lipoplexes demonstrate great
potential to treat cancer including melanoma in the context
of cancer immunotherapy, gene, and vaccine therapy. Indeed,
several human clinical studies are ongoing with lipoplex vac-
cines to treat melanoma (NCT, NCT, and
NCT), breast cancer (NCT), advanced esopha-
geal cancer and (non-small cell) lung cancer (NCT),
or multiple tumors including pancreatic, colorectal and lung
cancers (NCT), whereby the lipoplex carriers include
DOTAP/DOPE or DOTMA/DOPE compositions or a proprietary
RNAiMAX composition. Both intravenous and subcutaneous
administration routes have been utilized for lipoplex admin-
istration while emerging evidence shows that the administra-
tion route can critically determine whether mRNA–lipoplexes
stimulate or inhibit type I interferon (IFN) immune responses
based on a recent in vivo mouse study.[] Promoting strong
immune responses should also be balanced with safety con-
siderations, highlighting the potential of lipoplexes along with
important translational needs for refining vaccination regimens
and advancing mechanistic understanding of how administra-
tion route and dosing relate to immune responses.
3.3.2. Infectious Disease Vaccines
Likewise, lipoplexes can be utilized as a vaccine for infectious
disease such as immunodeficiency syndrome caused by human
immunodeficiency virus (HIV). Similar to the immunotherapy
for cancer, DOTAP/DOPE lipoplex carrying mRNA encoding
Gag protein modulated DCs to stimulate HIV-specific immune
responses.[] The same lipoplex system with HIV- antigen
Gag-encoding mRNA produced antigen-specific functional
T cells in spleens and lymph nodes, together with the immune-
activation characterized by the release of type I IFN.[] Another
inspiring example includes the parallel evaluation of lipoplex
and LNP entrapping HIV- Env-encoding saRNA to compare
the transfection eciency in vitro (Figure8A) and in vivo.[]
Strikingly, immunogenicity comparison of both formulations
resulted in a distinctive trend that the cationic lipid-based lipo-
plexes (DDAB rather than DOTAP) achieved the maximum
response of HIV IgG titers after a single dose whereas all LNPs
showed the enhanced antibody response only after the second
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (23 of 38) © 2022 Wiley-VCH GmbH
Figure 8. Possibilities of tuning the lipid compositions in conventional lipoplex or LNP compositions. A) Transfection eciency of Luc-saRNA delivered
by LNP/lipoplex under standard conditions without FCS (() FCS), with FCS ((+) 50% FCS), or with RNAse ((+) RNAse). B) Immunogenicity measured
by antibody titers after IM injection of HIV-1 Env saRNA complexed to the LNP/lipoplex. C) Zeta potential measurements of mRNA lipoplexes with
varying complexing lipids. D) Total protein expression mediated by C) lipoplexes within dierent cell subtype. E) Schematic representation showing
selective organ targeting (SORT) strategy for tissue-specific nucleic acid delivery by adding dierent net charged lipid molecules to aim specific biodis-
tribution. A,B) Adapted with permission.[172] Copyright 2019, Springer Nature. C,D) Adapted with permission.[185] Copyright 2021, Elsevier. E) Adapted
with permission.[38] Copyright 2021, National Academy of Sciences.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (24 of 38) © 2022 Wiley-VCH GmbH
injection (FigureB). This proof-of-concept study highlights the
potential usage of cationic lipids with adjuvant properties such
as enhancing humoral immune response and immunostimula-
tory eects in vivo.[]
Recently, DNA-lipoplex vaccine was also developed for the
treatment of SARS-CoV- utilizing saturated neutral lipid
(DOTAP/DOPE/DPPC = // mol%) and low N/P charge
ratio of . with highly negative net charge.[a] The study
selected intramuscular (IM) injection method and has proven
its eective spike protein expression, neutralizing antibody
responses, and T-cell activation comparable to the parallelly
tested gold standard, intramuscular electroporation method.
Similar trend was reported in the eorts to develop malaria
surface protein-encoded pDNA-lipoplex vaccine (equimolar
DOTAP/cholesterol), whereby the N/P of . (anionic) lipoplex
performed higher immunogenicity in mice compared to the
N/P of . (cationic) one when it was administered IM.[b] In
both cases with DNA-lipoplex, higher N/P triggered increased
cytotoxicity while lower N/P displayed higher transfection, pre-
sumably because of the more ecient release of DNA from
negatively charged lipoplexes (lower N/P) compared to the
higher ones. Considering that DNA-based drugs can be stored
at – °C for a few years and room temperature for  year,[]
ecient entrapment of DNA using lipoplex can improve the
availability of the vaccine, avoiding costly frozen-state supply
chain for existing mRNA vaccines.[]
3.3.3. Lipid Nanoparticles versus Lipoplexes: Which One Is Better?
As the formulation and concept of LNP and lipoplex are dis-
tinctive, it is an immediate point of inquiry to consider which
one works better and under what conditions. However, there
have been a few direct comparison studies to address this
point. In one example, Kubota et al. evaluated cationic lipid-
based LNP and lipoplex in conventional LNP composition
(DODMA/DSPC/cholesterol/PEG-lipid = //./.), using
ethanol mixing and extrusion ( nm-filtered), respectively, to
compare their physicochemical properties as well as biological
functionalities including particle characterization, transfection
eciency, and inflammatory cytokine response.[] While both
particles showed the maximum size at N/P = , lipoplexes were
only slightly larger than LNP with  nm margin ( and
 nm, respectively), which was probably attributed to the
addition of PEG lipid and DODMA with low pKa that could be
verified again by unilamellar structure of lipoplexes observed by
cryo-EM images. While cellular uptake eciency and inflam-
matory cytokine responses indicated LNP is more compatible
with the given cell line, it should be noted that the selected
composition was ideal for LNP formulation and not for lipo-
plexes as the entrapment mechanism of the latter one is mainly
governed by electrostatic attraction on surface.
On the other hand, Blakney et al. have fabricated both
ionizable and cationic lipid-based LNP and lipoplex with
a conventional lipoplex formula (complexing/neutral
lipid = / mol%) to characterize physicochemical proper-
ties and biological activities.[] Although both liposomes and
LNPs were produced by microfluidics, all resulted particle sizes
ranged –nm, which are quite larger than typical LNPs,
postulating that the self-assembled structure of LNPs formed
here is dierent from the one from classical composition with
PEG lipids. Only ionizable lipid-based lipoplex showed negative
surface charge, reflecting incompletely coupled saRNA due to
the ionization of the lipid at physiological pH on surface, which
later significantly aected the susceptibility to degradation, and
lower transfection in vitro as well as in vivo (FigureA). Conclu-
sively, the overall functionality was the highest from ionizable
lipid (C-) and cationic lipid (DDAB rather than DOTAP
in this case) in LNP and lipoplex, respectively, confirming that
the optimization strategy of entrapping nucleic acid should be
distinguished between LNP and lipoplex platform.
For further study, the same group has formulated lipo-
plexes by varying complexing lipids from cationic to zwitteri-
onic or ionizable lipids (complexing/neutral lipid = / mol%)
to intradermally deliver mRNA in human skin explants and
evaluated the eects of complexing lipids on cellular uptake
and expression.[] Intriguingly, the combinations with zwit-
terionic (cephalin) and ionizable lipids (C- and MC)
could also successfully entrap nucleic acid (complexation e-
ciency > %) with varied extent of negative zeta potentials
(FigureC). Surprisingly, the dierences in the surface charge
after forming lipoplex were strongly correlated with the pat-
tern of nucleic acid uptake and expression among cell types
(Figure D), not the inherent lipid types classified (i.e., cati-
onic/zwitterionic/ionizable lipids). For instance, cellular uptake
depending on the cell types shared similarities among moder-
ately charged cephalin and DDAB lipoplexes, whereas highly
charged DOTAP and MC lipoplexes were expressed in dif-
ferent distributions of cells within skin explants.
Another example includes recent eorts on developing selec-
tive organ targeting (SORT) lipid nanoparticles designed to
deliver nucleic acid to specific organs such as lungs, spleens,
and livers by tuning the membrane charge of the composi-
tion.[,] Based on conventional LNP composition (ionizable
lipid/DOPE/cholesterol/PEG-lipid = /// mol%), ioniz-
able (cationic) lipids, anionic lipids, and permanently charged
cationic lipids were supplemented, and the result clearly indi-
cated that the respective inclusion of particles delivered mRNA
to the liver, spleen, and lungs after IV injection (FigureE).[]
This distinctive tissue-targeting property was derived from the
recognition, anity, and adsorption of specific serum proteins
toward particle’s surface charge/composition.[] These findings
strongly support the nanoarchitectonic design strategy of lipid
self-assemblies that can be adopted to engineer optimized lipid-
based nucleic acid delivery vehicles.
4. Lipid Bicelles
While lipid bicelles were originally invented as suitable mem-
brane-mimetic environments to reconstitute membrane pro-
teins for structural biology studies,[] they have also emerged
as promising tools for biotechnology applications such as drug
delivery,[] antibacterial medicine,[] and lipid bilayer coat-
ings.[] To date, LNPs and lipoplexes are the most advanced
types of lipid-based nanoparticles for nucleic acid delivery appli-
cations, but there is growing attention to lipid bicelles. Com-
pared to the typical spherical geometry of LNPs and lipoplexes,
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (25 of 38) © 2022 Wiley-VCH GmbH
bicelles have a wider range of possible geometries, including
a disk-like shape in some cases that can be advantageous for
tissue penetration and cellular uptake. As such, lipid bicelles
represent a promising frontier for lipid-based nanoparticles for
nucleic acid delivery applications and are covered in this part.
4.1. Nanoarchitecture Design Principles
Bicelle morphology depends on a wide range of factors and
controlling pertinent parameters such as lipid concentration,
molar ratios, and temperature can enable rational control over
nanoarchitecture features. In this section, we begin by intro-
ducing the self-assembly principles of bicelles and relevant con-
siderations related to morphology and lipid geometry, before
expanding our coverage to more advanced functionalization
concepts that enhance bicelle stability.
4.1.1. Self-Assembly Principles
It is widely appreciated that long-chain phospholipids self-
assemble to form liposomes in aqueous solution whereas
detergents self-assemble into micelles.[] Interestingly, new
morphological phases emerge when long-chain phospholipids
and detergents are mixed together and this original discovery
gave rise to the “bicelle” concept.[] Practical utilization of
bicelles was further advanced by recognizing that short-
chain phospholipids can behave as detergent-like molecules
and hence it is possible to form bicelles using phospholipids
only.[] The molar ratio of long-chain phospholipids to short-
chain phospholipids is defined as the q-ratio and strongly influ-
ences bicellar phase properties.[] Depending on the lipid
composition and q-ratio, bicelles can form a wide range of mor-
phologies, including multilamellar vesicles, toroidal pore com-
plexes, extended lamellae, chiral nematic ribbons, disk-shaped
nanostructures, and spherical micelles[] (Figure9A).
While the phase properties of bicelles are complex and
depend on various parameters such as lipid concentration,
ionic strength, and temperature, it is possible to control them,
and we turn our attention to the case of disk-like bicelles that
form in certain conditions. The self-assembly of bicelle disks
can be rationalized by considering the geometrical properties
of the dierent lipid components, as depicted in FigureB,C.
In general, lipid molecules have a hydrophilic headgroup and
one or more hydrophobic alkyl chains. For short-chain phos-
pholipids, the headgroup is relatively large compared to the two
short chains and hence the geometry is cone-like. A similar con-
ical geometry is also found for typical detergent molecules. As
such, short-chain phospholipid and detergent molecules tend
to pack in high-curvature regions. On the other hand, long-
chain phospholipids usually have a more cylindrical appear-
ance because the headgroup and two long chains are more
similar shape wise. It is therefore more energetically favorable
for long-chain phospholipids to reside in planar regions. For
these reasons, upon mixing of lipid molecules in bicelle com-
positions, the overall self-assembly process is driven by hydro-
phobic interactions and the dierent types of lipids are prefer-
entially organized into distinct bicelle regions. Indeed, due to
the curvature-related energetic preferences, short-chain phos-
pholipids and detergents have low miscibility with long-chain
phospholipids[] and the cone-like and cylinder-like molecules
tend to self-assemble into the disk edges and central planar
region, respectively.[] An interesting case concerns the use
of headgroup-functionalized, long-chain phospholipids such as
PEGylated lipids. Due to the increased size of the headgroup
Figure 9. Lipid bicelle concept and molecular building blocks. A) Schematic illustrations of dierent bicelle morphologies that can form depending on
molecular components and experimental conditions as follows: (i) multilamellar vesicles; (ii) toroidal pore complexes; (iii) extended lamellae; (iv) chiral
nematic ribbons; (v) disk-shaped nanostructures; and (vi) spherical micelles. B) Cross-sectional view of a bicelle disk that is composed of long-chain
and short-chain phospholipids. C) Molecular geometry of dierent molecular building blocks, including (i) detergents and short-chain phospholipids
with cone-like geometries, (ii) long-chain phospholipids with cylinder-like geometries, and (iii) PEGylated phospholipids with cone-like geometries.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (26 of 38) © 2022 Wiley-VCH GmbH
in such cases, PEGylated lipids can assume a more cone-like
appearance than typical long-chain phospholipids and may
locate in the edge and/or planar regions of a bicelle disk.[]
There are also a few practical points to highlight: ) the
q-ratio must be within a certain range to form bicelle disks
while other morphologies can form in dierent q-ratio
ranges;[] ) the planar region of bicelle disks can present a
membrane-mimicking environment, especially in cases where
biologically relevant phospholipids are used. In certain cases,
phospholipids and cholesterol can both be incorporated in
the planar region and can form phase-separated domains in
some cases;[] ) bicelle preparation should take place at a suf-
ficiently high temperature so that all lipid components, espe-
cially the long-chain phospholipids, in the mixture are in the
fluid phase.[] The fabricated sample can then be cooled down
afterward; and ) the specific morphology of bicelles is highly
sensitive to the specific phospholipids used in a mixture and
the corresponding phase diagram must be carefully considered
in each particular case. For example, including saturated or
unsaturated lipids in bicelle compositions can cause wide vari-
ance in the phase diagram and possible morphologies[] (for
some compositions, the disk morphology is not possible), and
similar attention must be placed on charged lipids which are
sometimes added to enhance bicelle stability but can also aect
possible morphologies.[]
4.1.2. Advanced Functionalization Strategies
There have been various functionalization strategies introduced
into bicelle nanoarchitecture designs in order to improve perfor-
mance and enable new application opportunities. One direction
has focused on improving bicelle stability to prevent coalescence
and to inhibit the structural transformation of bicellar disks into
vesicles over time, especially at higher temperatures. Charged
lipids such as negatively charged phosphatidylglycerols (PG)
have often been incorporated into bicelles and prevent bicelle
coalescence due to charge repulsion.[] Inspired by other
classes of lipid-based nanomedicines, researchers have also
added PEGylated lipids such as distearoyl phosphoethanolamine-
[methoxy (polyethyleneglycol)-] (DSPE-PEG) to further
improve bicelle stability and reduce uptake by the reticuloen-
dothelial system, potentially enabling increased circulation time
in vivo. For example, it is possible to form disk-like bicelles in
the presence of  mol% PEG-DSPE, which can enhance
bicelle stability due to steric hindrance of the PEG chains.[]
While PEGylated lipids have a mainly cone-like geometry, more
detailed structural studies have shown that a neutral PEGylated
lipid (C-PEG-ceramide) can distribute across both the
planar region and rims of bicellar disk. Of note, there is indeed
an approximately, two-fold greater density of PEGylated lipids in
the rims,[] however, it should also be noted that the broader
distribution of the PEGylated lipid is advantageous for poten-
tial in vivo applications. In molecular dynamics simulations, it
has also been shown that bicelles containing ligand-modified
PEGylated lipids are more easily internalized by cells than simi-
larly functionalized liposomes due to more favorable membrane
energetics[] (see also other examples of bicelles with improved
cellular uptake properties over liposomes[]).
Other eorts have explored how to create inorganic–organic
hybrid bicelles based on mixing short-chain phospholipids
and long-chain organoalkoxysilane lipids (termed cerasome-
forming lipids, CFL).[] After lipid hydration, the bicelles were
incubated overnight at room temperature, during which time
a sol-gel reaction occurs between the alkoxysilyl headgroups
of the long-chain lipid molecules and results in a crosslinked
siloxane network (a silica-like coating). The hybrid bicelles had
typical diameters and heights of around  and nm, respec-
tively, and retained their structural morphology in the air envi-
ronment upon drying. Whereas conventional bicelles trans-
formed into vesicles at high temperature, the hybrid bicelles
also maintained their disk-like nanostructures at high tem-
perature.[] Based on these advantageous properties, there
have been eorts to develop partially silica-coated bicelles by
mixing short-chain phospholipids with CFL and PEGylated
lipids (DSPE-PEG). With increasing DSPE-PEG lipid
fraction, there was improved release of a small-molecule drug
and the PEGylated bicelles also had improved biocompatibility
and cellular uptake properties in vitro.[] As such, there is
increasing control over the functional properties of bicelles and
these eorts have coincided with attempts to utilize bicelles for
nucleic acid delivery.
4.2. Nucleic Acid Complexation
Currently, there are two main strategies to encapsulate nucleic
acids within bicelles. The first strategy involves fabricating
bicelles in the presence of nucleic acids while the second
strategy focuses on bicelle fabrication initially, followed by
mixing bicelles and nucleic acids in a lipoplex-type format. The
merits and latest progress of both strategies, including struc-
tural and functional insights, are critically discussed below.
4.2.1. Coassembly of Lipids and Nucleic Acids
In the earliest example, the stacking of disk-like, cationic bicelles
and negatively charged DNA was explored in order to form a new
lipid nanoarchitecture for potential nucleic acid delivery applica-
tions.[] The bicelles were composed of long-chain and short-
chain lipids in a : molar ratio; the short-chain phospholipid
was ,-diheptanoyl-sn-glycero--phosphocholine (diCPC) while
the long-chain phospholipid consisted of a mixture of ,-dihexa-
decanoyl-sn-glycero--phosphocholine (DPPC) lipid and a cationic
lipid termed β-[N-(N,N-dimethylaminoethane)-carbamoyl]
cholesterol hydrochloride (DC-Chol). The DNA strands had
 or  base pairs and the lipid-to-DNA molar ratio was
from  to  (– mol% DC-Chol in the long-chain lipid
population). An interesting point in the study was the simplicity
of the fabrication approach. The dry lipid film was directly
hydrated in an aqueous solution containing DNA molecules, fol-
lowed by bath sonication at  °C. As such, the bicelles were
directly formed in the presence of DNA, which is distinct from
the lipoplex approach in which case liposomes are first formed
before mixing with nucleic acids. It was determined that the
DNA strands intercalate between the bicellar disk layers, and the
number of layers increased at greater cationic lipid fractions in
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (27 of 38) © 2022 Wiley-VCH GmbH
the bicelles (Figure10A). Accordingly, the size of the lipid-DNA
nanostructures could be varied between – nm and the
ability to fabricate sub- nm nanostructures was highlighted
as a potential merit compared to lipoplexes, which typically have
> nm diameter. It was further noted that the DNA–bicelle
interactions were stable in acidic pH conditions and the fabri-
cation approach was also applicable to other commonly used
cationic lipids and surfactants such as ,-dioleoyl--trimethyl-
ammonium-propane (DOTAP) and hexadecyl-trimethyl-ammo-
nium bromide (CTAB).
The bicelle–DNA stacking concept was further extended
to anionic lipid bicelles by using the same basic design prin-
ciples and fabrication approach as in the previous example
and replacing the cationic lipid with an anionic lipid termed
,-dipalmitoyl-sn-glycero--phospho-(’-rac-glycerol) (sodium
salt) (DPPG) ( mol% in the long-chain lipid population).[]
While cationic lipids can readily form complexes with nucleic
acids and are eective transfection agents, they are cytotoxic,
which prompted the exploration of anionic lipids as an alter-
native option. In order to promote bicelle–DNA interactions,
divalent Ca+ ions were incorporated into the system and a min-
imum Ca+ concentration was needed to form stacked bicelle–
DNA complexes (Figure B). Complex formation did not
occur at  ×  Ca+ concentration, whereas bicelle–DNA
complexes did form at  ×  Ca+ concentration and
the interlayer spacing within the complexes did not depend on
the Ca+ concentration. At ×  Ca+ concentration, an
additional lamellar phase of ion-lipid complexes also formed in
cases of low DNA concentration, indicating that all three com-
ponents—lipids, DNA, and ions—play an important role in
driving the self-assembly process with controlled stoichiometry.
The resulting bicelle–DNA complexes had typical dimensions
around –nm diameter and –nm length.
In addition to DNA, there has also been exploration of uti-
lizing bicelles to encapsulate peptide nucleic acids (PNAs)
that target microRNA for potential gene editing and thera-
peutic purposes.[] The bicelles had a lipid composition of
around  mol% DPPC,  mol% DHPC,  mol% charged
lipid (anionic DPPG or cationic DOTAP), and  mol% DSPE-
PEG, and were prepared using a combination of vortexing,
temperature cycling (between °C and °C), and subsequent
room-temperature dilution—all steps were conducted in the
presence of PNAs at a PNA:lipid molar ratio between :
and :. The control bicelles without PNAs had an nm
diameter radius, while PNA-loaded anionic bicelles had an
 min diameter and disk-like appearance. In marked con-
trast, PNA-loaded cationic bicelles underwent extensive aggre-
gation due to electrostatic attraction, resulting in large vesicles.
By optimizing the PNA:lipid molar ratio, encapsulation e-
ciencies as high as % could be achieved with anionic bicelles,
which also demonstrated superior cellular uptake properties in
vitro compared to cationic bicelles and a polymeric nanoparticle
control. The PNA-loaded anionic bicelles also had negligible
cell cytotoxicity in vitro. Mechanistic studies further revealed
that the PNA-loaded anionic bicelles were internalized by HeLa
cells by an endocytic mechanism and the delivered PNAs were
functionally active to inhibit a target microRNA.
4.2.2. Mixing Prefabricated Bicelles with Nucleic Acids
Anecdotally, there have also been reports of bicelle–nucleic acid
complexes that were developed using other fabrication proto-
cols and were evaluated in terms of structural properties and
in vitro transfection eciency. For example, lipid mixtures
consisting of gel-phase DPPC helper lipid and one of two cati-
onic lipids (malonic acid diamides) were prepared using thin-
film hydration followed by shaker incubation at °C and bath
sonication, and were noted to form heterogenous, disk-like
bicelle aggregates with  nm height and  nm diameter
(: molar ratio of cationic lipid to helper lipid).[] The cationic
lipid bicelles were then mixed with negatively charged plasmid
DNA (pDNA: . to  N/P ratios) using a lipoplex-type protocol
at room temperature before attempting to transfect human
lung carcinoma (A) cells. It was noted that the bicelle–pDNA
complexes had low transfection eciency, which was attributed
to poor mixing and phase separation, whereas liposome–pDNA
complexes formed using other helper lipids such as fluid-phase
DOPC had eective lipoplex-type transfection. A follow-up
study characterized the interactions between DPPC/cationic
lipid mixtures (: molar ratio) and mainly double-stranded
DNA ( N/P ratio), and noted that the bicelle–DNA complexes
Figure 10. Nucleic acid interactions with lipid bicelles. A) Schematic illustration of a cationic bicelle–DNA complex. The interaction between cationic
lipid bicelles and DNA molecules induced self-assembly of a multilamellar complex consisting of alternating layers of bicelles and DNA molecules. The
complex formation was mediated by attractive electrostatic interactions between cationic lipids and negatively charged DNA. B) Schematic illustration
of an anionic bicelle–DNA complex. The interaction between anionic lipid bicelles and DNA molecules is mediated by divalent Ca2+ ions and induces
self-assembly of a multilamellar complex that consists of alternating layers of bicelles and DNA molecules.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (28 of 38) © 2022 Wiley-VCH GmbH
retained disk-like nanostructures with a condensed lamellar
phase.[] In addition, using a similar preparation protocol,
lipid mixtures consisting of DMPC helper lipid and a three-
chain cationic lipid (DiTT; another malonic acid diamide)
were reported to form disk-like bicelles when the DiTT molar
fraction was around – mol% and had –nm diame-
ters.[] In that case, a lipoplex-type protocol was again followed
to mix the DiTT/DMPC bicelles with pDNA (. to  N/P
ratio) and the resulting bicelle–pDNA complexes demonstrated
ecient transfection of A cells, including in the presence
of % serum. There is also evidence that the DiTT lipid by
itself can self-assemble into disk-like nanostructures that can be
useful for transfection.[]
In another study, bicelles composed of CFL, DOTAP, and
DHPC along with ligand-modified DSPE-PEG or DHPE-
PEG were prepared for gene delivery applications.[] The
base bicelle composition was CFL, DOTAP, and DHPC in an
approximately :: molar ratio and, in some cases, ligand-
modified DSPE-PEG or DHPE-PEG was included too.
The conceptual idea was that ligand-modified DSPE-PEG
would present the ligand (i.e., cRGD peptide that binds to inte-
grin receptors) on the main plane of the bicellar disk, whereas
the ligand-modified DHPE-PEG would present cRGD pep-
tide on the bicellar edge instead. In general, the bicelles were
initially prepared by hydrating a dry lipid film at °C, followed
by ultrasonication for min and then overnight incubation at
room temperature to induce sol–gel formation on the bicellar
surface due to the presence of the CFL lipid. Afterward, using
a lipoplex-type protocol, the bicelles were incubated for min
at °C with small interfering RNA (siRNA) at a bicelle:siRNA
molar ratio of :. The resulting bicelles had typical height
and diameter values of  and  nm, respectively, while it
was noted that the siRNA-loaded bicelles tended to be slightly
larger. Notably, when bicelles with edge-located cRGD had
superior siRNA loading capacity compared to bicelles with
plane-located cRGD, in which case it was suggested that the
ligand location diminishes bicelle–siRNA interactions. More-
over, the siRNA-loaded bicelles with edge-located cRGD showed
greater performance to inhibit a cellular signaling pathway in
vivo along with improved antitumor activities. The performance
dierences were attributed to the tailored, orthogonal function-
alities aorded by the bicelle nanoarchitecture, whereby the
targeting ligands were located on the bicelle edges while the
cationic lipids (DOTAP) were located in the plane region to
facilitate bicelle–siRNA interactions.
Together, these findings highlight how multiple strategies
are being implemented to encapsulate and deliver nucleic acids
using bicelles while the field evolves from fundamental struc-
tural studies to early application contexts where feasibility and
utility are being demonstrated. Concurrently, eorts are being
made to explore new ways of producing bicelles that can sup-
port the translation of these innovations from laboratory con-
cepts to real-life applications.
4.3. Bicelle Manufacturing Considerations
While lipid bicelles are typically made by the thin-film hydra-
tion method as described above, the method has practical
limitations to advance beyond the laboratory scale. Here, we
begin by briefly introducing the method in more detail before
critically discussing the latest progress in developing more
advanced bicelle production strategies.
For the thin-film hydration method, lipids in organic sol-
vent are first mixed together and the solvent is evaporated in
order to form a dry lipid film.[] The film is then hydrated and
the aqueous lipid suspension is subjected to multiple rounds
of freeze-thaw-vortex cycling. The freezing step often involves
immersing the sample in liquid nitrogen (°C or below) and
the heating step is usually done in a water bath with –°C
temperature. Other variations with dierent temperature cycles
are also possible to some extent. However, in all cases, the
method is limited to batch production and the harsh processing
conditions can potentially denature biological components such
as proteins and nucleic acids. As such, to enhance translational
prospects, there have been ongoing eorts to develop more suit-
able bicelle fabrication methods that are ecient and scalable.
4.3.1. Semispontaneous Method
One pioneering example is adoption of the semispontaneous
method to easily prepare bicelles from aordable lipids and sur-
factants.[] The method was first demonstrated using soybean
lecithin and poly(oxyethylene) sorbitan monooleate (Tween ),
which were initially dissolved in a water-soluble organic sol-
vent, ,-butanediol in this case (and dipropyleneglycol in some
other cases), followed by water dilution; the temperature during
the processing steps was fixed at °C (other studies have used
up to °C). The lipid/surfactant mixture was then subjected
to tip ultrasonication as the final processing step. Dynamic
light scattering (DLS) and transmission electron microscopy
(TEM) characterization of the aqueous suspension identified
the formation of disk-like bicelles at a : molar ratio of lecithin
to Tween . The typical diameter and thickness values of the
bicelles were around  and nm, respectively, and remained
physically stable for at least  d. One recent mechanistic
study discussed how two-tail lecithin molecules constitute the
disk body while one-tail Tween  molecules form the disk
rim.[] The semi-spontaneous method has also been utilized to
fabricate disk-like bicelles from other compositions, including
using a polyglycerol-type nonionic surfactant with two asym-
metrical tails[] and mixtures of lecithin and poly(oxyethylene)
cholesteryl ethers by themselves[] or in combination with up
to  mol% cholesterol in the liquid-ordered phase.[]
4.3.2. Microfluidic Production
In addition to the semispontaneous method that focuses on
batch processing, there have been recent eorts to develop a
continuous bicelle production method by utilizing microflu-
idic-based hydrodynamic flow focusing, which additionally
does not require heating or cooling.[] A schematic out-
line of the microfluidic chip is presented in Figure11A and
involves mixing tangential streams of long-chain phospho-
lipids dispersed in an organic solvent such as methanol or
ethanol with short-chain phospholipids dispersed in aqueous
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (29 of 38) © 2022 Wiley-VCH GmbH
buer (pH .). When the two streams merged, the long-chain
and short-chain phospholipids mixed and self-assembled to
form bicelles in a continuous manner while it was important
to evaporate residual solvent in a low vacuum environment
after sample collection from the microfluidic channel. Using
,-dimyristoyl-sn-glycero--phosphocholine (DMPC) and
,-dihexanoyl-sn-glycero--phosphocholine (DHPC) as model
long-chain and short-chain phospholipids, respectively, it was
found that the molar ratio (q-ratio) of DMPC to DHPC was an
important determinant of the resulting lipid nanoparticle prop-
erties. When the DHPC concentration was fixed at  × 
, it was identified that heterogenous vesicles formed at low
DMPC concentrations (<× ; equivalent to q-ratio of less
than .) (Figure B). Conversely, bicelles formed at higher
DMPC concentrations ( ×  ; equivalent to q-ratio of
equal to or greater than .). The morphological structures
in the dierent cases were confirmed by TEM imaging and it
was further verified that the disk-like bicelle structures exhib-
ited interfacial membrane properties that are characteristic of
ordered bicelles (as opposed to those of mixed micelles[])
(FigureC,D). Systematic experiments identified that hetero-
geneous vesicles, homogenous vesicles, or bicelles could be
formed depending on the total lipid concentration and q-ratio,
and that a suciently high total lipid concentration was needed
to form bicelles (FigureE). The diusional specifics of the sol-
vent mixing process also influenced the self-assembly outcome
while controlling the flow conditions along with the q-ratio ena-
bled fine-tuning of the bicelle size (FigureF).
A follow-up study explored the general utility of the micro-
fluidic chip based on hydrodynamic focusing to produce
bicelles composed of DHPC short-chain phospholipid with
other long-chain phospholipids that have dierent gel-to-fluid
phase transition temperatures (Tm).[] Interestingly, the eect
of long-chain phospholipid concentration on lipid nanoparticle
morphology varied depending on the specific type of long-
chain phospholipid that was used in the microfluidic produc-
tion process. While heterogeneous vesicles and bicelles were
formed at low and high DMPC concentrations, respectively,
replacing DMPC (Tm  °C) with -palmitoyl--oleoyl-sn-
glycero--phosphocholine (POPC; Tm  °C) lipid led to a
nearly opposite trend. In that case, bicelles were formed at low
POPC concentrations, which was attributed to strong hydro-
phobic interactions between POPC molecules, whereas large
vesicles were formed at high POPC concentrations. Additional
experiments with ,-dipalmitoyl-sn-glycero--phosphocholine
(DPPC; Tm°C) lipid further revealed how lipid phase prop-
erties can aect self-assembly outcomes. When the microflu-
idic production process was carried out with DPPC at  °C,
large vesicles were formed at all tested DPPC concentrations
due to strong interactions between DPPC molecules, which is
consistent with the gel-phase properties of DPPC at that tem-
perature. In marked contrast, when the microfluidic production
process was carried out with DPPC at °C, DPPC was in the
fluid-phase state during production and bicelles were formed
at all tested DPPC concentrations. Of note, it was mentioned
that the resulting DPPC/DHPC bicelles had high membrane
Figure 11. Microfluidic production of lipid bicelles. A) Schematic illustration of microfluidic setup whereby long-chain and short-chain phospholipids
are dispersed in organic solvent and aqueous buer, respectively, and mix to form lipid bicelles according to the hydrodynamic focusing concept. In
this example, DMPC and DHPC are model long-chain and short-chain phospholipids, respectively. B) Eect of long-chain phospholipid (DMPC) con-
centration on the size of lipid nanoparticles. At low DMPC concentration, the size distribution indicates formation of heterogeneous vesicles (indicated
by label ii), whereas bicelle formation (indicated by label i) occurred at higher DMPC concentrations. C) Transmission electron microscopy (TEM)
image of microfluidic-fabricated vesicles and inset corresponds to a cryo-TEM image of the vesicles. D) TEM image of microfluidic-fabricated bicelles.
E) Morphology of microfluidic-fabricated lipid nanoparticles depending on long-chain phospholipid (DMPC) and short-chain phospholipid (DHPC)
concentrations. i–iii) Bicelles, heterogeneous vesicles, and homogenous vesicles, respectively. F) Size of microfluidic-fabricated bicelles depending
on the eective q-ratio of long-chain phospholipid to short-chain phospholipid concentrations. All panels are adapted with permission.[223] Copyright
2021, American Chemical Society.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (30 of 38) © 2022 Wiley-VCH GmbH
ordering after cooling down to ambient conditions, which is
indicative of the gel-phase properties of DPPC in that tempera-
ture range. Together, these findings establish that microfluidic
systems are capable of producing bicelles with diverse lipid
compositions, and it is possible to rationally control bicelle size
and lipid composition.
From a broader perspective, continued progress on the man-
ufacturing side highlights how it is possible to produce bicelles
using low-cost lipid and lipid-like materials and to shift from
batch to continuous production methods. However, there is an
outstanding need to expand eorts in this direction. Until now,
such eorts have focused solely on producing lipid bicelles
by considering short-chain and long-chain phospholipids and
related surfactants, and it would be beneficial to also focus on
how to fabricate bicelles containing nucleic acids based on the
strategies described above along with adding other important
components such as PEGylated lipids.
4.4. Current Applications and Future Possibilities
Until now, there have been only a few cases that utilize nucleic-
acid-loaded bicelles for applications. However, bicelles have
emerged as a promising type of lipid nanoparticle for various
biomedical applications due to their much smaller size than
LNPs, liposomes, and lipoplexes so that they can eectively
penetrate intercellular spaces (-nm) and due to their unique
morphological diversity depending on the lipid composition
and molar ratio of components. Among numerous potential
applications, bicelles are already known to reinforce the skin’s
biophysical properties and complement drug penetration. We
critically review the latest progress in those fields, whereby par-
ticular focus is placed on advantageous features that could be
useful for nucleic acid delivery applications in future work.
4.4.1. Skin Repair and Delivery
Disk-like bicelles initially received attention for topical skin
applications because they are smaller than liposomes (which
cannot penetrate intracellular spaces of the stratum corneum
(SC),[] which is the outermost layer of skin) and contain less
irritating molecular components than micelles. Early studies
verified that disk-like bicelles are useful for their permeability
enhancing eects on skin while not causing irritation,[] and
can also penetrate the SC[] and influence the lipid composi-
tion/phase there.[] Follow-up studies revealed that adding
specific molecular components such as cholesterol sulfate could
improve the distribution and penetration of bicelles throughout
the SC,[] and such findings heightened interest in utilizing
bicelles as lipid nanoparticles to improve skin properties and to
deliver other medically useful molecules.[] Indeed, continued
research delineated how tuning bicelle morphology could
expand the range of skin-related biological functionalities,[]
leading to the use of bicelles as dermal carriers and enhancers
for treating impaired skin.
In terms of applying bicelles as carriers, it was reported
that bicelle-encapsulated diclofenac diethylamine (DDEA) had
lower skin penetration than aqueous DDEA suspensions.[]
However, when skin was pretreated with empty bicelles prior
to adding aqueous DDEA, the degree of skin penetration
was enhanced. Similar retarding eects were observed when
flufenamic acid was encapsulated in bicelles, which was sug-
gested to be a potentially advantageous feature for safely deliv-
ering drugs that have otherwise rapid percutaneous absorp-
tion[] (see also related structural studies[]). Further investi-
gation supported that bicelle-encapsulated DDEA can be deliv-
ered to damaged skin and help to repair SC barrier function
while minimizing systemic side eects.[]
While relevant application studies have been limited to small
molecule drugs so far, there continues to be high interest in uti-
lizing bicelles for dermal applications.[] Several eorts have
been made to evaluate the biophysical properties of healthy
as well as irritated human skin in vivo before and after bicelle
treatment.[,] While conventional bicelles composed of
DMPC/DHPC increased transepidermal water loss, skin elas-
ticity, and skin dehydration without aecting the phase behavior
of SC lipidic microstructures in human subjects,[] incorpora-
tion of long-chain lipids with longer alkyl chain lengths might
tune the extent of skin reinforcement rather than penetrating
into the SC lipid structures.[b] Another study utilizing encap-
sulated bicelles in liposomes (bicosomes) to mimic epidermal
lamellar bodies showed that treatment with bicosomes rein-
forced the skin barrier function in both healthy and irritated
human skin, probably by reestablishing the lamellar lipid struc-
ture of theSC.[]
Looking forward, there is an outstanding need to investi-
gate how bicelle carriers might be useful carriers for delivering
nucleic acids to the skin, especially in intradermal applications.
Recent discussions have suggested that intradermal adminis-
tration of RNA-based vaccines may reduce dose requirements
and be possible with needle-free approaches.[] In these
respects, bicelles stand out among possible lipid nanoparticle
carrier options because they have well-validated properties for
skin delivery applications.
4.4.2. Cancer Therapy
There has also been extensive interest in utilizing disk-like
bicelles to deliver small-molecule and peptide drugs for anti-
cancer therapy. For example, PEGylated lipid bicelles have been
used to encapsulate the anticancer drug doxorubicin (DOX),
with up to % encapsulation eciency and >-times longer
circulation time in mice.[] Compared to free DOX, the bicelle-
encapsulated DOX was more eective at accumulating in
tumors and the bicelles could be internalized by breast cancer
cells. Anionic lipid bicelles have also been developed to coen-
capsulate two anticancer drugs, paclitaxel (PTX) and parthe-
nolide (PTL), for in vivo therapeutic treatment.[] To improve
tumor targeting, bicelles have been further functionalized with
RGD peptides and this strategy was able to increase tumor pen-
etration of bicelle-encapsulated melittin peptide compared to
non-functionalized bicelles, which in turn improved antitumor
activity in vivo.[] Motivated by these early results, the field has
continued to progress toward clinical applications, especially
in terms of utilizing PEGylated lipids to enhance functional
performance. In one case, the addition of a PEGylated lipid to
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (31 of 38) © 2022 Wiley-VCH GmbH
adisk-like bicelle composition triggered the formation of small,
spherical micelles that, when loaded with the anticancer drug
docetaxel (DTX), not only had superior stability and cell and
tumor uptake properties compared to PEGylated liposomes,
but also demonstrated improved antitumor activity in vivo.[]
There have also been other reports that show PEGylated lipids
can help to sterically stabilize bicelles in order to enhance
uptake and anticancer drug delivery to human cancer cells.[]
Expanding on these concepts, there have also been eorts to
utilize more durable, organic–inorganic hybrid bicelles for anti-
cancer applications.[] In one interesting example, DOX was
loaded together with indocyanine green (ICG)—a fluorescent
dye used for photothermal therapy—for temperature-regulated
DOX release and photothermal-triggered cancer cell cytotox-
icity, which could be achieved in the presence of near-infrared
fluorescence light.[] The bicelles exhibited high levels of
tumor targeting and exhibited anticancer activity in vivo, which
was superior to either the chemotherapy or photothermal
therapy alone. Together with advances in developing partially
silica-coated hybrid bicelles as described above,[] such pos-
sibilities open the door to advanced cancer theragnostics and
it has been suggested the hybrid bicelles could also be useful
for gene delivery applications as well.[] Indeed, one study uti-
lizing hybrid bicelles for siRNA delivery has been reported for
anticancer treatment in vivo[] while highlighting the future
opportunities that lie ahead to use various types of bicelles,
including inorganic-organic hybrid bicelles, for nucleic acid
delivery applications.
5. Conclusion and Outlook
The development of lipid-based nucleic acid delivery vehi-
cles has increasingly gained traction over the past decade
and shifted from innovative laboratory concepts to practically
eective and scalable clinical technologies, most recently cul-
minating in the large-scale manufacturing and worldwide
distribution of mRNA–LNP vaccines to prevent COVID-
infection. These advances have been largely enabled by ration-
ally designing the LNP structure to achieve maximum perfor-
mance based on the nanoarchitectonics concept along with
utilizing microfluidic production technologies, which not only
paved the way for production scale-up, but more importantly
allowed precise control over production parameters to tune the
nanoarchitecture features of manufactured LNPs in a precise
and reproducible manner.
Along this line, there exist numerous opportunities to explore
numerous combinations of lipid excipients and nucleic acid
payloads, and the role of nanoarchitectonics-based research will
become increasingly crucial to spur future innovation. We iden-
tify several key areas in which nanoarchitectonics is expected to
play a central role. First, there is an opportunity to explore dif-
ferent routes of administration, especially by taking advantage
of dierent lipid nanoarchitectures such as bicelles that can be
potentially useful for intradermal applications. Secondly, there
is an urgent demand to overcome challenges related to the in
vivo behavior of LNPs, for example, due to issues related to the
immune response. Specifically, recent eorts have been made
to identify appropriate alternatives to PEG lipids. In doing so,
it is important to apply the concept of nanoarchitectonics to
modulate interactions of the PEG lipid (or alternatives) with
other lipid excipients and select appropriate candidates that will
retain favorable interactions with the rest of the components.
Another area of interest is the expansion of microfluidic
capabilities, especially to emerging classes of lipid-based nano-
particles such as bicelles for not only proof-of-concept dem-
onstration as done so far but also for practical application (as
in the LNP case). To date, improvements to microfluidic plat-
forms have been achieved incrementally and are mostly aimed
at obtaining better control over the physical properties of the
resultant lipid-based vehicle (i.e., mainly size and size distri-
bution). Enhancing the capability of microfluidic platforms
to precisely control the morphology and architecture of the
lipid-based delivery vehicle needs guidance from nanoarchitec-
tonics and could help to improve nucleic acid encapsulation in
tandem as well.
Finally, while clinically approved LNP-based drug and vac-
cines are composed of a single formulation (i.e., delivering a
single type of nucleic acid payload) there is growing interest in
multiple- or coformulations. For example, a recent study dem-
onstrated successful in vivo CRISPR/Cas gene editing for
the transthyretin gene in mice and rats using LNPs that were
coformulated to include single guide RNA (sgRNA) and Cas
mRNA.[] Formulation optimization was performed for in vitro
and in vivo gene editing with necessary modifications to incor-
porate sgRNA. The LNPs were formulated using microfluidics
and they achieved >% transthyretin protein knockdown for
months. This study highlights the need for nucleic acid LNPs
to be carefully engineered to match their therapeutic applica-
tions; specifically, how optimized, co-formulated LNPs can be
applied for cell engineering and genome editing. More notably,
it emphasizes the vast potential of LNPs as nucleic acid delivery
systems, which allows coformulations to be optimized with high
eciencies, paving the way for novel nucleic acid-based thera-
peutics for a wide range of diseases in the near future.
As a closing remark, we may also add that the three types
of lipid-based nanoparticles covered herein were selected due
to their distinct nanoarchitectures and breadth of development
stages, spanning conceptual innovations to full-fledged com-
mercial products. Across this spectrum, the nanoarchitectonics
concept is a useful approach to analyze the progress and trends
and to build a rational understanding of how to eectively
design lipid-based nanoparticles for various biomedical appli-
cations. Until now, the impact of nanoarchitectonics on lipid-
based nanoparticles for nucleic acid delivery applications had
been largely implicit, but we foresee that nanoarchitectonics
will play an increasingly prominent and visible role to drive for-
ward future innovations both in terms of scientific discussion
and research progress.
Acknowledgements
A.R.F and S.P. contributed equally to this work. This work was supported
by the Ministry of Education (MOE) in Singapore under grant AcRF
TIER1-2020-T1-002-032 (RG111/20) and by a sponsored research
agreement from LUCA AICell Inc. (RCA- LUCA AICell REQ0239282 –
LUCA), and by the National Research Foundation of Korea (NRF) grants
funded by the Korean government (MSIT) (Nos. 2020R1C1C1004385
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (32 of 38) © 2022 Wiley-VCH GmbH
and 2021R1A4A1032782). In addition, this work was supported by
the International Research & Development Program of the National
Research Foundation of Korea (NRF) funded by the Ministry of Science
and ICT (2020K1A3A1A39112724). This work was also supported by a
grant of the Korea Health Technology R&D Project through the Korea
Health Industry Development Institute (KHIDI), funded by the Ministry
of Health & Welfare, Republic of Korea (Grant Number: HI19C1328).
H.T. was supported by an SINGA graduate scholarship from the A*STAR
Graduate Academy, Singapore. Schematic illustrations were created with
BioRender.com under an academic lab subscription.
Conflict of Interest
N.-J.C. and J.A.J. are listed as coinventors on patents and patent
applications that are related to lipid bilayer technologies. In addition,
N.-J.C. is a founder of and J.A.J. is a scientific advisor to LUCA AICell Inc,
which is developing lipid-related technologies. The other authors declare
no conflict of interest.
Keywords
drug delivery, lipid nanoparticle, microfluidics, nanoarchitectonics,
vaccine
Received: March 31, 2022
Revised: May 9, 2022
Published online: June 16, 2022
[1] a) J. A. Kulkarni, D. Witzigmann, S. B. Thomson, S. Chen,
B. R. Leavitt, P. R. Cullis, R. van der Meel, Nat. Nanotechnol.
2021, 16, 630; b) H. Yin, R. L.Kanasty, A. A.Eltoukhy, A. J.Vegas,
J. R.Dorkin, D. G.Anderson, Nat. Rev. Genet. 2014, 15, 541.
[2] S.Chesnoy, L.Huang, Annu. Rev. Biophys. Biomol. Struct. 2000, 29, 27.
[3] P.Guo, J. Nanosci. Nanotechnol. 2005, 5, 1964.
[4] Y.Eygeris, S.Patel, A.Jozic, G.Sahay, Nano Lett. 2020, 20, 4543.
[5] Y. Zhang, H. Li, J. Sun, J. Gao, W. Liu, B. Li, Y. Guo, J. Chen,
Int. J. Pharm. 2010, 390, 198.
[6] E.Dolgin, Nat. Biotechnol. 2022, 40, 283.
[7] P. F. McKay, K. Hu, A. K. Blakney, K. Samnuan, J. C. Brown,
R.Penn, J.Zhou, C. R.Bouton, P.Rogers, K.Polra, Nat. Commun.
2020, 11, 3523.
[8] C. Webb, S. Ip, N. V. Bathula, P. Popova, S. K. Soriano,
H. H.Ly, B.Eryilmaz, V. A.Nguyen Huu, R.Broadhead, M.Rabel,
Mol. Pharmaceutics 2022, 19, 1047.
[9] a) K. S.Corbett, D. K.Edwards, S. R.Leist, O. M.Abiona, S.Boyoglu-
Barnum, R. A. Gillespie, S. Himansu, A. Schäfer, C. T. Ziwawo,
A. T. DiPiazza, Nature 2020, 586, 567; b) L. Schoenmaker,
D. Witzigmann, J. A. Kulkarni, R. Verbeke, G.Kersten, W. Jiskoot,
D. J.Crommelin, Int. J. Pharm. 2021, 601, 120586.
[10] a) X. Hou, T. Zaks, R. Langer, Y. Dong, Nat. Rev. Mater.
2021, 6, 1078; b) R. Tenchov, R. Bird, A. E. Curtze, Q. Zhou,
ACS Nano 2021, 15, 16982; c) B. Z.Igyártó, S.Jacobsen, S.Ndeupen,
Curr. Opin. Virol. 2021, 48, 65; d) E. Kon, U. Elia, D. Peer,
Curr. Opin. Biotechnol. 2022, 73, 329.
[11] a) M. Aono, K. Ariga, Adv. Mater. 2016, 28, 989; b) K. Ariga,
J. A. Jackman, N. J. Cho, S. h. Hsu, L. K. Shrestha, T. Mori,
J. Takeya, Chem. Rec. 2019, 19, 1891; c) K. Ariga, Y. Yamauchi,
Chem.- Asian J. 2020, 15, 718.
[12] a) K.Ariga, S.Watanabe, T.Mori, J.Takeya, NPG Asia Mater. 2018,
10, 90; b) K.Ariga, M.Nishikawa, T.Mori, J.Takeya, L. K.Shrestha,
J. P. Hill, Sci. Technol. Adv. Mater. 2019, 20, 51; c) K.Ariga, X. Jia,
J.Song, J. P.Hill, D. T. Leong, Y. Jia, J.Li, Angew. Chem., Int. Ed.
2020, 59, 15424; d) K.Ariga, Nanoscale Horiz. 2021, 6, 364.
[13] a) J. Geall Andrew, A. Verma, R.Otten Gillis, A.Shaw Christine,
A. Hekele, K. Banerjee, Y. Cu, W. Beard Clayton, A. Brito Luis,
T. Krucker, T. O’Hagan Derek, M. Singh, W. Mason Peter,
M. Valiante Nicholas, R. Dormitzer Philip, W. Barnett Susan,
R.Rappuoli, B.Ulmer Jerey, W.Mandl Christian, Proc. Natl. Acad.
Sci. USA 2012, 109, 14604; b) E.Samaridou, J.Heyes, P.Lutwyche,
Adv. Drug Delivery Rev. 2020, 154-155, 37; c) Y. Zhang, C. Sun,
C.Wang, K. E.Jankovic, Y.Dong, Chem. Rev. 2021, 121, 12181.
[14] a) A. C. Anselmo, S. Mitragotri, Bioeng. Transl. Med. 2016, 1, 10;
b) A. C. Anselmo, S. Mitragotri, Bioeng. Transl. Med. 2019, 4,
e10143; c) S.Yonezawa, H.Koide, T. Asai, Adv. Drug Delivery Rev.
2020, 154–155, 64.
[15] a) T. T.Thi, E. J. A. Suys, J. S. Lee, D. H. Nguyen, K. D. Park,
N. P.Truong, Vaccines 2021, 9, 359; b) B. N.Aldosari, I. M.Alfagih,
A. S.Almurshedi, Pharmaceutics 2021, 13, 206; c) A. C. Anselmo,
S.Mitragotri, Bioeng. Transl. Med. 2021, 6, e10246.
[16] a) M.Cornebise, E.Narayanan, Y.Xia, E. Acosta, L.Ci, H.Koch,
J.Milton, S. Sabnis, T.Salerno, K. E.Benenato, Adv. Funct. Mater.
2022, 32, 2106727; b) G. Basha, T. I. Novobrantseva, N. Rosin,
Y. Y. C. Tam, I. M. Hafez, M. K. Wong, T. Sugo, V. M. Ruda,
J.Qin, B. Klebanov, M.Ciufolini, A.Akinc, Y. K. Tam, M. J.Hope,
P. R. Cullis, Mol. Ther. 2011, 19, 2186; c) N.Maurer, K. F.Wong,
H.Stark, L.Louie, D.McIntosh, T.Wong, P.Scherrer, S. C.Semple,
P. R.Cullis, Biophys. J. 2001, 80, 2310; d) S. C.Semple, S. K.Klimuk,
T. O. Harasym, N. Dos Santos, S. M. Ansell, K. F. Wong,
N.Maurer, H.Stark, P. R.Cullis, M. J.Hope, P.Scherrer, Biochim.
Biophys. Acta, Biomembr. 2001, 1510, 152.
[17] a) S. J. Shepherd, D. Issadore, M. J. Mitchell, Biomaterials 2021,
274, 120826; b) M.Maeki, S.Uno, A.Niwa, Y.Okada, M.Tokeshi,
J. Controlled Release 2022, 344, 80.
[18] a) K. Ward, Z. H.Fan, J. Micromech. Microeng. 2015, 25, 094001;
b) C.-Y. Lee, C.-L. Chang, Y.-N.Wang, L.-M. Fu, Int. J. Mol. Sci.
2011, 12, 3263; c) M. J. W.Evers, J. A.Kulkarni, R. van der Meel,
P. R. Cullis, P. Vader, R. M. Schielers, Small Methods 2018, 2,
1700375.
[19] a) K. M. McMahon, M. P. Plebanek, C. S. Thaxton, Adv. Funct.
Mater. 2016, 26, 7824; b) K. M. McMahon, R. K. Mutharasan,
S.Tripathy, D.Veliceasa, M.Bobeica, D. K.Shumaker, A. J.Luthi,
B. T.Helfand, H. Ardehali, C. A.Mirkin, O.Volpert, C. S.Thaxton,
Nano Lett. 2011, 11, 1208; c) K. M. McMahon, C. S. Thaxton,
Expert Opin. Drug Delivery 2014, 11, 231; d) J.Wang, A. E.Calvert,
N. Kaplan, K. M. McMahon, W. Yang, K. Q. Lu, H. Peng,
C. S.Thaxton, R. M.Lavker, Adv. Ther. 2020, 3, 2000138.
[20] a) K. C.Vickers, B. T.Palmisano, B. M.Shoucri, R. D.Shamburek,
A. T.Remaley, Nat. Cell Biol. 2011, 13, 423; b) F.Tabet, K. C.Vickers,
L. F. Cuesta Torres, C. B. Wiese, B. M. Shoucri, G. Lambert,
C. Catherinet, L. Prado-Lourenco, M. G. Levin, S. Thacker,
P.Sethupathy, P. J.Barter, A. T.Remaley, K.-A.Rye, Nat. Commun.
2014, 5, 3292.
[21] M.Ouimet, T. J.Barrett, E. A.Fisher, Circ. Res. 2019, 124, 1505.
[22] a) K. C. Vickers, A. T. Remaley, Curr. Opin. Lipidol. 2012, 23, 91;
b) K. C.Vickers, A. T.Remaley, J. Lipid Res. 2014, 55, 4; c) S.Nazir,
V. Jankowski, G. Bender, S. Zewinger, K.-A. Rye, E. P. C. van
derVorst, Adv. Drug Delivery Rev. 2020, 159, 94.
[23] X. Chen, L. S. Mangala, L. Mooberry, E. Bayraktar, S. K. Dasari,
S. Ma, C. Ivan, K. A. Court, C. Rodriguez-Aguayo, R. Bayraktar,
S. Raut, N. Sabnis, X. Kong, X. Yang, G. Lopez-Berestein,
A. G.Lacko, A. K.Sood, Oncogene 2019, 38, 6095.
[24] A. G.Lacko, N. A.Sabnis, B. Nagarajan, W. J.McConathy, Front.
Pharmacol. 2015, 6, 247.
[25] M. M. K. Shahzad, L. S. Mangala, H. D. Han, C. Lu,
J. Bottsford-Miller, M. Nishimura, E. M. Mora, J.-W. Lee,
R. L. Stone, C. V. Pecot, D. Thanapprapasr, J.-W. Roh, P. Gaur,
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (33 of 38) © 2022 Wiley-VCH GmbH
M. P.Nair, Y.-Y.Park, N.Sabnis, M. T.Deavers, J.-S.Lee, L. M.Ellis,
G. Lopez-Berestein, W. J. McConathy, L. Prokai, A. G. Lacko,
A. K.Sood, Neoplasia 2011, 13, 309.
[26] T. M.Allen, P. R.Cullis, Adv. Drug Delivery Rev. 2013, 65, 36.
[27] a) J. S. Suk, Q. Xu, N. Kim, J. Hanes, L. M. Ensign, Adv. Drug
Delivery Rev. 2016, 99, 28; b) Z.Wang, E. D.Hood, J.Nong, J.Ding,
O. A.Marcos-Contreras, P. M.Glassman, K. M.Rubey, M.Zaleski,
C. L. Espy, D.Gullipali, T. Miwa, V. R. Muzykantov, W.-C. Song,
J. W.Myerson, J. S.Brenner, Adv. Mater. 2022, 34, 2107070.
[28] M.Longmire, P. L.Choyke, H.Kobayashi, Nanomedicine 2008, 3,
703.
[29] a) J. Lu, Y. Zhao, X. Zhou, J. H. He, Y. Yang, C. Jiang, Z. Qi,
W.Zhang, J.Liu, Biomacromolecules 2017, 18, 2286; b) A. Orellana
Esteban, S. Tenneti, L. Rangasamy, L. T. Lyle, S. Low Philip,
L.Kasinski Andrea, Sci. Transl. Med. 2017, 9, eaam9327.
[30] a) C.Wolfrum, S.Shi, K. N.Jayaprakash, M.Jayaraman, G.Wang,
R. K.Pandey, K. G.Rajeev, T.Nakayama, K.Charrise, E. M.Ndungo,
T. Zimmermann, V. Koteliansky, M. Manoharan, M. Stoel,
Nat. Biotechnol. 2007, 25, 1149; b) M. Egli, M. Manoharan,
Acc. Chem. Res. 2019, 52, 1036.
[31] a) Y. Yang, J. Wang, H. Shigematsu, W. Xu, W. M. Shih,
J. E. Rothman, C. Lin, Nat. Chem. 2016, 8, 476; b) Z. Zhang,
Y.Yang, F.Pincet, M. C.Llaguno, C.Lin, Nat. Chem. 2017, 9, 653.
[32] a) J.Guan, Q. Shen, Z.Zhang, Z.Jiang, Y.Yang, M.Lou, J.Qian,
W. Lu, C. Zhan, Nat. Commun. 2018, 9, 2982; b) N.AlSawaftah,
W. G. Pitt, G. A. Husseini, ACS Pharmacol. Transl. Sci. 2021, 4,
1028.
[33] C.Kelly, C.Jeeries, S.-A.Cryan, J. Drug Delivery 2011, 2011, 727241.
[34] a) P.-A. Monnard, T.Oberholzer, P. Luisi, Biochim. Biophys. Acta,
Biomembr. 1997, 1329, 39; b) S.Dow, Expert Opin. Drug Delivery
2008, 5, 11; c) C. R.Safinya, K. K. Ewert, R. N.Majzoub, C.Leal,
New J. Chem. 2014, 38, 5164.
[35] A. A. Barba, G. Lamberti, C. Sardo, B. Dapas, M. Abrami,
M. Grassi, R.Farra, F.Tonon, G. Forte, F. Musiani, M. Licciardi,
G. Pozzato, F. Zanconati, B.Scaggiante, G.Grassi, G.Cavallaro,
Curr. Drug Metab. 2015, 16, 427.
[36] a) E. H.Pilkington, E. J. A.Suys, N. L. Trevaskis, A. K.Wheatley,
D.Zukancic, A.Algarni, H.Al-Wassiti, T. P.Davis, C. W.Pouton,
S. J.Kent, N. P.Truong, Acta Biomater. 2021, 131, 16; b) Y.Eygeris,
M.Gupta, J.Kim, G.Sahay, Acc. Chem. Res. 2022, 55, 2.
[37] N. R. Larson, G. Hu, Y. Wei, A. D. Tuesca, M. L. Forrest,
C. R.Middaugh, J. Pharm. Sci. 2022, 111, 690.
[38] A. Dilliard Sean, Q. Cheng, J. Siegwart Daniel, Proc. Natl. Acad.
Sci. USA 2021, 118, e2109256118.
[39] M. Schlich, R. Palomba, G. Costabile, S.Mizrahy, M. Pannuzzo,
D.Peer, P.Decuzzi, Bioeng. Transl. Med. 2021, 6, e10213.
[40] a) S. M.Lee, Q.Cheng, X.Yu, S.Liu, L. T.Johnson, D. J.Siegwart,
Angew. Chem., Int. Ed. 2021, 60, 5848; b) M. Jayaraman,
S. M. Ansell, B. L. Mui, Y. K. Tam, J. Chen, X. Du, D. Butler,
L. Eltepu, S. Matsuda, J. K. Narayanannair, K. G. Rajeev,
I. M. Hafez, A. Akinc, M. A. Maier, M. A. Tracy, P. R. Cullis,
T. D.Madden, M.Manoharan, M. J.Hope, Angew. Chem., Int. Ed.
2012, 51, 8529; c) O. S.Fenton, K. J.Kauman, R. L. McClellan,
E. A.Appel, J. R.Dorkin, M. W.Tibbitt, M. W.Heartlein, F.DeRosa,
R.Langer, D. G.Anderson, Adv. Mater. 2016, 28, 2939.
[41] a) S. Liu, Q. Cheng, T. Wei, X. Yu, L. T. Johnson, L. Farbiak,
D. J.Siegwart, Nat. Mater. 2021, 20, 701; b) N.Meng, D. Grimm,
Signal Transduction Targeted Ther. 2021, 6, 206.
[42] K. A. Hajj, R. L. Ball, S. B. Deluty, S. R. Singh, D. Strelkova,
C. M.Knapp, K. A.Whitehead, Small 2019, 15, 1805097.
[43] a) M. A. Maier, M. Jayaraman, S. Matsuda, J. Liu, S. Barros,
W. Querbes, Y. K.Tam, S. M.Ansell, V.Kumar, J.Qin, X.Zhang,
Q. Wang, S. Panesar, R. Hutabarat, M. Carioto, J. Hettinger,
P. Kandasamy, D. Butler, K. G. Rajeev, B. Pang, K. Charisse,
K.Fitzgerald, B. L.Mui, X.Du, P.Cullis, T. D.Madden, M. J.Hope,
M.Manoharan, A.Akinc, Mol. Ther. 2013, 21, 1570; b) Y. Suzuki,
K. Hyodo, T. Suzuki, Y. Tanaka, H. Kikuchi, H. Ishihara, Int. J.
Pharm. 2017, 519, 34; c) H. Tanaka, T. Takahashi, M. Konishi,
N. Takata, M. Gomi, D. Shirane, R. Miyama, S. Hagiwara,
Y.Yamasaki, Y.Sakurai, K.Ueda, K.Higashi, K.Moribe, E.Shinsho,
R.Nishida, K.Fukuzawa, E.Yonemochi, K.Okuwaki, Y.Mochizuki,
Y.Nakai, K.Tange, H.Yoshioka, S.Tamagawa, H.Akita, Adv. Funct.
Mater. 2020, 30, 1910575; d) H. Tanaka, T.Takahashi, M.Konishi,
N. Takata, M. Gomi, D. Shirane, R. Miyama, S. Hagiwara,
Y.Yamasaki, Y.Sakurai, Adv. Funct. Mater. 2020, 30, 1910575.
[44] a) O. S.Fenton, K. J.Kauman, R. L.McClellan, J. C. Kaczmarek,
M. D. Zeng, J. L. Andresen, L. H. Rhym, M. W. Heartlein,
F.DeRosa, D. G.Anderson, Angew. Chem., Int. Ed. 2018, 57, 13582;
b) M. J.Carrasco, S.Alishetty, M.-G.Alameh, H. Said, L. Wright,
M.Paige, O. Soliman, D.Weissman, T. E.Cleveland, A.Grishaev,
M. D.Buschmann, Commun. Biol. 2021, 4, 956.
[45] F. Ding, H. Zhang, J. Cui, Q. Li, C. Yang, Biomater. Sci. 2021, 9,
7534.
[46] M. Kim, M. Jeong, S. Hur, Y. Cho, J. Park, H. Jung, Y. Seo,
H. A.Woo, K. T.Nam, K.Lee, H.Lee, Sci. Adv. 2021, 7, eabf4398.
[47] a) S.Mochizuki, N. Kanegae, K.Nishina, Y.Kamikawa, K.Koiwai,
H.Masunaga, K.Sakurai, Biochim. Biophys. Acta, Biomembr. 2013,
1828, 412; b) J. A.Kulkarni, D.Witzigmann, J.Leung, Y. Y. C.Tam,
P. R. Cullis, Nanoscale 2019, 11, 21733; c) R.Zhang, R. El-Mayta,
T. J. Murdoch, C. C.Warzecha, M. M. Billingsley, S. J.Shepherd,
N.Gong, L.Wang, J. M.Wilson, D. Lee, M. J.Mitchell, Biomater.
Sci. 2021, 9, 1449; d) J.Kim, A. Jozic, G. Sahay, Cell. Mol. Bioeng.
2020, 13, 463.
[48] S. Patel, N. Ashwanikumar, E. Robinson, Y. Xia, C. Mihai,
J. P. Grith, S. Hou, A. A. Esposito, T. Ketova, K. Welsher,
J. L.Joyal, Ö.Almarsson, G.Sahay, Nat. Commun. 2020, 11, 983.
[49] H.Tanaka, R.Miyama, Y.Sakurai, S.Tamagawa, Y.Nakai, K.Tange,
H.Yoshioka, H.Akita, Pharmaceutics 2021, 13, 2097.
[50] F. Sebastiani, M. Yanez Arteta, M. Lerche, L. Porcar, C. Lang,
R. A. Bragg, C. S. Elmore, V. R. Krishnamurthy, R. A. Russell,
T. Darwish, H. Pichler, S. Waldie, M. Moulin, M. Haertlein,
V. T.Forsyth, L.Lindfors, M.Cárdenas, ACS Nano 2021, 15, 6709.
[51] a) P. Sellaturay, S. Nasser, S. Islam, P.Gurugama, P. W. Ewan,
Clin. Exp. Allergy 2021, 51, 861; b) A. Troelnikov, G. Perkins,
C. Yuson, A. Ahamdie, S. Balouch, P. R. Hurtado, P. Hissaria,
J. Allergy Clin. Immunol. 2021, 148, 91; c) M.Picard, J.-P. Drolet,
M.-S. Masse, C. A. Filion, F. Almuhizi, M. Fein, A. Copaescu,
G. A. C. Isabwe, M. Blaquière, M.-N. Primeau, J. Allergy Clin.
Immunol.: Pract. 2022, 10, 620; d) J. A. Cahill, M. Kan, Allergy
2022, 77, 337; e) P.Sellaturay, P.Gurugama, V.Harper, T.Dymond,
P.Ewan, S.Nasser, Clin. Exp. Allergy 2022, 52, 12.
[52] B.Cabanillas, N.Novak, C. A.Akdis, Allergy 2021, 76, 1938.
[53] a) A. Akinc, W. Querbes, S. De, J. Qin, M. Frank-Kamenetsky,
K. N. Jayaprakash, M. Jayaraman, K. G. Rajeev, W. L. Cantley,
J. R. Dorkin, J. S. Butler, L. Qin, T. Racie, A. Sprague, E. Fava,
A. Zeigerer, M. J. Hope, M. Zerial, D. W. Y. Sah, K. Fitzgerald,
M. A. Tracy, M. Manoharan, V. Koteliansky, A. d. Fougerolles,
M. A. Maier, Mol. Ther. 2010, 18, 1357; b) S. Sabnis,
E. S. Kumarasinghe, T. Salerno, C. Mihai, T.Ketova, J. J. Senn,
A. Lynn, A. Bulychev, I. McFadyen, J. Chan, Ö. Almarsson,
M. G.Stanton, K. E.Benenato, Mol. Ther. 2018, 26, 1509.
[54] a) D. W.Lee, X.Banquy, K.Kristiansen, Y.Kaufman, J. M. Boggs,
J. N. Israelachvili, Proc. Natl. Acad. Sci. USA 2014, 111, E768;
b) A. H.Pande, S.Qin, S. A.Tatulian, Biophys. J. 2005, 88, 4084.
[55] a) A. Wittrup, A. Ai, X. Liu, P. Hamar, R.Trifonova, K. Charisse,
M. Manoharan, T. Kirchhausen, J. Lieberman, Nat. Bio-
technol. 2015, 33, 870; b) J. Gilleron, W. Querbes, A. Zeigerer,
A.Borodovsky, G.Marsico, U. Schubert, K.Manygoats, S. Seifert,
C.Andree, M. Stöter, H.Epstein-Barash, L.Zhang, V. Koteliansky,
K.Fitzgerald, E.Fava, M.Bickle, Y.Kalaidzidis, A.Akinc, M.Maier,
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (34 of 38) © 2022 Wiley-VCH GmbH
M.Zerial, Nat. Biotechnol. 2013, 31, 638; c) G.Sahay, W.Querbes,
C. Alabi, A. Eltoukhy, S. Sarkar, C. Zurenko, E. Karagiannis,
K.Love, D. Chen, R.Zoncu, Y. Buganim, A.Schroeder, R.Langer,
D. G.Anderson, Nat. Biotechnol. 2013, 31, 653.
[56] a) N. Hartl, F. Adams, O. M. Merkel, Adv. Ther. 2021, 4,
2000092; b) P. Sabourian, G. Yazdani, S. S. Ashraf, M. Frounchi,
S.Mashayekhan, S.Kiani, A.Kakkar, Int. J. Mol. Sci. 2020, 21, 8019.
[57] R.Gaspar, F.Coelho, B. F. B.Silva, Molecules 2020, 25, 5006.
[58] R. L. Ball, K. A. Hajj, J. Vizelman, P. Bajaj, K. A. Whitehead,
Nano Lett. 2018, 18, 3814.
[59] a) J. K.Watts, G. F.Deleavey, M. J.Damha, Drug Discovery Today
2008, 13, 842; b) N. J. Caplen, E. W. Alton, P. G. Mddleton,
J. R. Dorin, B. J. Stevenson, X. Gao, S. R. Durham, P. K. Jeery,
M. E.Hodson, C.Coutelle, Nat. Med. 1995, 1, 39.
[60] a) B. Lewin, Genes 9, Jones & Bartlett Learning, Burlington, MA
2008; b) B.Alberts, Molecular Biology of the Cell, Garland Science,
New York 2004.
[61] A. Wu, Y. Peng, B. Huang, X. Ding, X. Wang, P. Niu, J. Meng,
Z. Zhu, Z. Zhang, J. Wang, J. Sheng, L. Quan, Z. Xia, W. Tan,
G.Cheng, T.Jiang, Cell Host Microbe 2020, 27, 325.
[62] a) M. Yanez Arteta, T. Kjellman, S. Bartesaghi, S. Wallin, X. Wu,
J. Kvist Alexander, A. Dabkowska, N. Székely, A. Radulescu,
J. Bergenholtz, L. Lindfors, Proc. Natl. Acad. Sci. 2018, 115, E3351;
b) J. A.Kulkarni, M. M.Darjuan, J. E.Mercer, S.Chen, R.van derMeel,
J. L.Thewalt, Y. Y. C.Tam, P. R.Cullis, ACS Nano 2018, 12, 4787.
[63] Z. Hejdankova, V. Vanek, F. Sedlak, J. Prochazka, A. Diederichs,
S.Kereïche, B.Novotna, M.Budesinsky, G.Birkus, K.Grantz Saskova,
P.Cigler, Adv. Funct. Mater. 2021, 31, 2101391.
[64] a) K. J. Kauman, J. R. Dorkin, J. H. Yang, M. W. Heartlein,
F. DeRosa, F. F. Mir, O. S. Fenton, D. G. Anderson, Nano Lett.
2015, 15, 7300; b) M. P. Lokugamage, D. Vanover, J.Beyersdorf,
M. Z. C. Hatit, L. Rotolo, E. S. Echeverri, H. E. Peck, H. Ni,
J.-K. Yoon, Y.Kim, P. J.Santangelo, J. E.Dahlman, Nat. Biomed.
Eng. 2021, 5, 1059; c) S. C.Semple, A.Akinc, J.Chen, A. P.Sandhu,
B. L. Mui, C. K. Cho, D. W. Y. Sah, D. Stebbing, E. J. Crosley,
E.Yaworski, I. M.Hafez, J. R.Dorkin, J.Qin, K.Lam, K. G.Rajeev,
K. F.Wong, L. B.Jes, L.Nechev, M. L.Eisenhardt, M.Jayaraman,
M.Kazem, M. A.Maier, M.Srinivasulu, M. J. Weinstein, Q.Chen,
R. Alvarez, S. A. Barros, S. De, S. K. Klimuk, T.Borland, etal.,
Nat. Biotechnol. 2010, 28, 172; d) M. M. Billingsley, N. Singh,
P. Ravikumar, R. Zhang, C. H. June, M. J. Mitchell, Nano Lett.
2020, 20, 1578; e) T. Terada, J. A.Kulkarni, A. Huynh, S. Chen,
R.van derMeel, Y. Y. C.Tam, P. R.Cullis, Langmuir 2021, 37, 1120.
[65] K. T. Love, K. P.Mahon, C. G. Levins, K. A. Whitehead, W.Querbes,
J. R.Dorkin, J.Qin, W.Cantley, L. L.Qin, T.Racie, M.Frank-Kamenetsky,
K. N. Yip, R. Alvarez, D. W. Y. Sah, A. de Fougerolles, K. Fitzgerald,
V.Koteliansky, A.Akinc, R.Langer, D. G.Anderson, Proc. Natl. Acad. Sci.
USA 2010, 107, 1864.
[66] a) R. A. Petros, J. M. DeSimone, Nat. Rev. Drug Discovery 2010,
9, 615; b) T.Nakamura, M.Kawai, Y.Sato, M.Maeki, M. Tokeshi,
H.Harashima, Mol. Pharmaceutics 2020, 17, 944.
[67] S. Chen, Y. Y. C. Tam, P. J. C. Lin, M. M. H. Sung, Y. K.Tam,
P. R.Cullis, J. Controlled Release 2016, 235, 236.
[68] A. U.Andar, R. R.Hood, W. N.Vreeland, D. L.DeVoe, P. W.Swaan,
Pharm. Res. 2014, 31, 401.
[69] K. J. Hassett, J. Higgins, A. Woods, B. Levy, Y. Xia, C. J. Hsiao,
E. Acosta, Ö. Almarsson, M. J. Moore, L. A. Brito, J. Controlled
Release 2021, 335, 237.
[70] D.Chen, K. T.Love, Y.Chen, A. A.Eltoukhy, C.Kastrup, G.Sahay,
A.Jeon, Y.Dong, K. A. Whitehead, D. G.Anderson, J. Am. Chem.
Soc. 2012, 134, 6948.
[71] K.Ariga, Molecules 2021, 26, 1621.
[72] a) A.Tan, L. Hong, J. D. Du, B. J.Boyd, Adv. Sci. 2019, 6, 1801223;
b) M. R. Molla, P. A. Levkin, Adv. Mater. 2016, 28, 1159;
c) S.Banerjee, J.Pillai, Expert Opin. Drug Metab. Toxicol. 2019, 15, 499.
[73] A.Wagner, K.Vorauer-Uhl, G.Kreismayr, H.Katinger, J. Liposome
Res. 2002, 12, 259.
[74] a) G. M.Whitesides, Nature 2006, 442, 368; b) J. Zhang, S. Yan,
D.Yuan, G.Alici, N.-T. Nguyen, M. EbrahimiWarkiani, W.Li, Lab
Chip 2016, 16, 10.
[75] a) T. A. Duncombe, A. M.Tentori, A. E.Herr, Nat. Rev. Mol. Cell
Biol. 2015, 16, 554; b) S.Damiati, U. B. Kompella, S. A. Damiati,
R.Kodzius, Genes 2018, 9, 103; c) Y.Bai, M.Gao, L. Wen, C.He,
Y.Chen, C.Liu, X.Fu, S. Huang, Biotechnol. J. 2018, 13, 1700170;
d) L.Zhang, Q.Chen, Y.Ma, J.Sun, ACS Appl. Bio Mater. 2020, 3,
107; e) M.Maeki, N.Kimura, Y.Sato, H.Harashima, M.Tokeshi,
Adv. Drug Delivery Rev. 2018, 128, 84.
[76] a) E. C.Jensen, B. P.Bhat, R. A. Mathies, Lab Chip 2010, 10, 685;
b) B.Schuster, M.Junkin, S. S.Kashaf, I.Romero-Calvo, K.Kirby,
J. Matthews, C. R. Weber, A.Rzhetsky, K. P. White, S. Tay, Nat.
Commun. 2020, 11, 5271.
[77] H. Wang, K. Liu, K.-J. Chen, Y. Lu, S. Wang, W.-Y. Lin, F. Guo,
K.-i. Kamei, Y.-C. Chen, M. Ohashi, M. Wang, M. A. Garcia,
X.-Z.Zhao, C. K. F.Shen, H.-R.Tseng, ACS Nano 2010, 4, 6235.
[78] Y. Kim, B. Lee Chung, M. Ma, W. J. M. Mulder, Z. A. Fayad,
O. C.Farokhzad, R.Langer, Nano Lett. 2012, 12, 3587.
[79] P. M.Valencia, E. M.Pridgen, M.Rhee, R.Langer, O. C.Farokhzad,
R.Karnik, ACS Nano 2013, 7, 10671.
[80] P. M.Valencia, O. C.Farokhzad, R.Karnik, R.Langer, Nat. Nanotechnol.
2012, 7, 623.
[81] M. D. Buschmann, M. J. Carrasco, S. Alishetty, M. Paige,
M. G.Alameh, D.Weissman, Vaccines 2021, 9, 65.
[82] C. M. Bailey-Hytholt, P. Ghosh, J. Dugas, I. E. Zarraga,
A.Bandekar, J. Visualized Exp. 2021, 168, e62226.
[83] a) M. T.Abrams, M. L.Koser, J.Seitzer, S. C.Williams, M. A.DiPietro,
W. Wang, A. W. Shaw, X. Mao, V. Jadhav, J. P. Davide, P. A. Burke,
A. B.Sachs, S. M.Stirdivant, L.Sepp-Lorenzino, Mol. Ther. 2010, 18, 171;
b) T. S.Zimmermann, A. C. H.Lee, A.Akinc, B.Bramlage, D.Bumcrot,
M. N.Fedoruk, J.Harborth, J. A.Heyes, L. B.Jes, M.John, A. D.Judge,
K. Lam, K. McClintock, L. V.Nechev, L. R. Palmer, T. Racie, I. Röhl,
S. Seiert, S. Shanmugam, V. Sood, J. Soutschek, I. Toudjarska,
A. J. Wheat, E. Yaworski, W. Zedalis, V. Koteliansky, M. Manoharan,
H.-P.Vornlocher, I.MacLachlan, Nature 2006, 441, 111.
[84] D.Stroock Abraham, K. W.Dertinger Stephan, A.Ajdari, I.Mezić,
A.Stone Howard, M.Whitesides George, Science 2002, 295, 647.
[85] S. J. Shepherd, C. C. Warzecha, S. Yadavali, R. El-Mayta,
M.-G.Alameh, L.Wang, D.Weissman, J. M.Wilson, D. Issadore,
M. J.Mitchell, Nano Lett. 2021, 21, 5671.
[86] N. M. Belliveau, J. Huft, P. J. C. Lin, S. Chen, A. K. K. Leung,
T. J.Leaver, A. W.Wild, J. B.Lee, R. J.Taylor, Y. K.Tam, C. L.Hansen,
P. R.Cullis, Mol. Ther.–Nucleic Acids 2012, 1, e37.
[87] A.Jahn, W. N.Vreeland, M.Gaitan, L. E. Locascio, J. Am. Chem.
Soc. 2004, 126, 2674.
[88] C. Webb, N. Forbes, C. B. Roces, G. Anderluzzi, G. Lou,
S. Abraham, L. Ingalls, K. Marshall, T. J. Leaver, J. A. Watts,
J. W.Aylott, Y.Perrie, Int. J. Pharm. 2020, 582, 119266.
[89] N. Kimura, M. Maeki, Y. Sato, Y. Note, A. Ishida, H. Tani,
H.Harashima, M.Tokeshi, ACS Omega 2018, 3, 5044.
[90] K.Amreen, S.Goel, ECS J. Solid State Sci. Technol. 2021, 10, 017002.
[91] a) J. B. Lee, K. Zhang, Y. Y. C. Tam, Y. K. Tam, N. M. Belliveau,
V. Y. C. Sung, P. J. C. Lin, E. LeBlanc, M. A. Ciufolini, P. S.Rennie,
P. R. Cullis, Int. J. Cancer 2012, 131, E781; b) S. Rietwyk, D. Peer,
ACS Nano 2017, 11, 7572; c) Z.Liu, S. Wang, C.Tapeinos, G. Torrieri,
V. Känkänen, N. El-Sayed, A. Python, J. T. Hirvonen, H. A. Santos,
Adv. Drug Delivery Rev. 2021, 174, 576; d) A. del Pozo-Rodríguez,
M. Á.Solinís, A.Rodríguez-Gascón, Eur. J. Pharm. Biopharm. 2016, 109,
184.
[92] a) D.Bumcrot, M.Manoharan, V.Koteliansky, D. W. Y.Sah, Nat.
Chem. Biol. 2006, 2, 711; b) R. L.Setten, J. J.Rossi, S.-p.Han, Nat.
Rev. Drug Discovery 2019, 18, 421.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (35 of 38) © 2022 Wiley-VCH GmbH
[93] a) A. Fire, S. Xu, M. K. Montgomery, S. A.Kostas, S. E.Driver,
C. C.Mello, Nature 1998, 391, 806; b) D. M.Dykxhoorn, D.Palliser,
J.Lieberman, Gene Ther. 2006, 13, 541.
[94] D. M.Dykxhoorn, J.Lieberman, Annu. Rev. Biomed. Eng. 2006, 8, 377.
[95] J. A. Kulkarni, D. Witzigmann, S. Chen, P. R. Cullis, R. van
derMeel, Acc. Chem. Res. 2019, 52, 2435.
[96] V. Kumar, J. Qin, Y. Jiang, R. G. Duncan, B. Brigham, S. Fishman,
J. K.Nair, A.Akinc, S. A.Barros, P. V.Kasperkovitz, Mol. Ther.–Nucleic
Acids 2014, 3, e210.
[97] C. B. Roces, G. Lou, N. Jain, S. Abraham, A. Thomas,
G. W.Halbert, Y.Perrie, Pharmaceutics 2020, 12, 1095.
[98] a) L. Schoenmaker, D. Witzigmann, J. A. Kulkarni, R. Verbeke,
G.Kersten, W.Jiskoot, D. J. A.Crommelin, Int. J. Pharm. 2021, 601,
120586; b) M.Gaviria, B.Kilic, Nat. Biotechnol. 2021, 39, 546.
[99] a) C. Wang, Y. Zhang, Y. Dong, Acc. Chem. Res. 2021, 54, 4283;
b) M.Qiu, Y. Li, H.Bloomer, Q.Xu, Acc. Chem. Res. 2021, 54, 4001;
c) L. R.Baden, H. M.El Sahly, B.Essink, K.Kotlo, S.Frey, R. Novak,
D.Diemert, S. A. Spector, N.Rouphael, C. B.Creech, J.McGettigan,
S. Khetan, N. Segall, J. Solis, A. Brosz, C. Fierro, H. Schwartz,
K.Neuzil, L.Corey, P.Gilbert, H.Janes, D.Follmann, M. Marovich,
J.Mascola, L. Polakowski, J.Ledgerwood, B. S.Graham, H.Bennett,
R. Pajon, C. Knightly, et al., New Engl. J. Med. 2020, 384, 403;
d) F. P. Polack, S. J. Thomas, N. Kitchin, J. Absalon, A. Gurtman,
S. Lockhart, J. L. Perez, G. Pérez Marc, E. D. Moreira, C. Zerbini,
R.Bailey, K. A.Swanson, S.Roychoudhury, K.Koury, P.Li, W. V.Kalina,
D.Cooper, R. W.Frenck, L. L.Hammitt, Ö.Türeci, H.Nell, A.Schaefer,
S. Ünal, D. B. Tresnan, S. Mather, P. R. Dormitzer, U. Şahin,
K. U.Jansen, W. C.Gruber, New Engl. J. Med. 2020, 383, 2603.
[100] a) D. Adams, A. Gonzalez-Duarte, W. D.O’Riordan, C.-C.Yang,
M. Ueda, A. V. Kristen, I. Tournev, H. H. Schmidt, T. Coelho,
J. L. Berk, K.-P. Lin, G. Vita, S. Attarian, V. Planté-Bordeneuve,
M. M. Mezei, J. M. Campistol, J. Buades, T. H. Brannagan,
B. J. Kim, J. Oh, Y. Parman, Y. Sekijima, P. N. Hawkins,
S. D. Solomon, M. Polydefkis, P. J. Dyck, P. J.Gandhi, S. Goyal,
J. Chen, A. L. Strahs, et al., New Engl. J. Med. 2018, 379, 11;
b) A.Akinc, M. A.Maier, M.Manoharan, K.Fitzgerald, M.Jayaraman,
S. Barros, S. Ansell, X. Du, M. J. Hope, T. D. Madden, B. L. Mui,
S. C.Semple, Y. K. Tam, M.Ciufolini, D.Witzigmann, J. A.Kulkarni,
R.van derMeel, P. R.Cullis, Nat. Nanotechnol. 2019, 14, 1084.
[101] M. L.Guevara, F.Persano, S.Persano, Front. Chem. 2020, 8, 963.
[102] a) R.Homan, L.Margolis, L.Bergelson, FEBS Lett. 1978, 93, 365;
b) R. Fraley, S. Subramani, P.Berg, D. Papahadjopoulos, J. Biol.
Chem. 1980, 255, 10431.
[103] F. Martinon, S. Krishnan, G. Lenzen, R. Magné, E. Gomard,
J. G.Guillet, J. P.Lévy, P.Meulien, Eur. J. Immunol. 1993, 23, 1719.
[104] R.Fraley, D.Papahadjopoulos, Trends Biochem. Sci. 1981, 6, 77.
[105] P.Felgner, Y.Barenholz, J.Behr, S.Cheng, P.Cullis, L.Huang, J.Jessee,
L.Seymour, F.Szoka, A.Thierry, Hum. Gene Ther. 1997, 8, 511.
[106] P. L. Felgner, T. R. Gadek, M. Holm, R. Roman, H. W. Chan,
M.Wenz, J. P.Northrop, G. M.Ringold, M.Danielsen, Proc. Natl.
Acad. Sci. USA 1987, 84, 7413.
[107] R. W.Malone, P. L.Felgner, I. M.Verma, Proc. Natl. Acad. Sci. USA
1989, 86, 6077.
[108] U. Sahin, P. Oehm, E. Derhovanessian, R. A. Jabulowsky,
M. Vormehr, M. Gold, D. Maurus, D. Schwarck-Kokarakis,
A. N.Kuhn, T.Omokoko, Nature 2020, 585, 107.
[109] a) T.Wang, L. M.Larcher, L.Ma, R. N.Veedu, Molecules 2018, 23,
2564; b) X.Yu, X.Liang, H.Xie, S. Kumar, N.Ravinder, J.Potter,
X.deMollerat du Jeu, J. D.Chesnut, Biotechnol. Lett. 2016, 38, 919;
c) A.Masotti, G.Mossa, C.Cametti, G.Ortaggi, A.Bianco, N.Del
Grosso, D.Malizia, C.Esposito, Colloids Surf., B 2009, 68, 136.
[110] a) F. Barthel, J.-S. Remy, J.-P. Loeer, J.-P. Behr, DNA Cell Biol.
1993, 12, 553; b) J. H.Felgner, R.Kumar, C.Sridhar, C. J.Wheeler,
Y. J. Tsai, R.Border, P. Ramsey, M. Martin, P. L. Felgner, J. Biol.
Chem. 1994, 269, 2550.
[111] a) F. Ponti, M.Campolungo, C. Melchiori, N.Bono, G.Candiani,
Chem. Phys. Lipids 2021, 235, 105032; b) D. Zhi, S.Zhang, S.Cui,
Y. Zhao, Y. Wang, D. Zhao, Bioconjugate Chem. 2013, 24, 487;
c) D. Zhi, Y. Bai, J. Yang, S. Cui, Y. Zhao, H. Chen, S. Zhang,
Adv. Colloid Interface Sci. 2018, 253, 117.
[112] M. L.Tilkins, P.Hawley-Nelson, P.Battista, Focus 1994, 16, 117.
[113] S.Guan, J.Rosenecker, Gene Ther. 2017, 24, 133.
[114] a) E.Galanis, E.Hersh, A.Stopeck, R.Gonzalez, P.Burch, C.Spier,
E. Akporiaye, J. Rinehart, J. Edmonson, R. Sobol, J. Clin. Oncol.
1999, 17, 3313; b) J. Buck, P. Grossen, P. R. Cullis, J. Huwyler,
D.Witzigmann, ACS Nano 2019, 13, 3754.
[115] a) J.-S. Remy, C. Sirlin, P.Vierling, J.-P.Behr, Bioconjugate Chem.
1994, 5, 647; b) D. A.Balazs, W.Godbey, J. Drug Delivery 2011, 2011,
326497.
[116] a) C.Srinivasan, D. J. Burgess, J. Controlled Release 2009, 136, 62;
b) E.Dodds, M.Dunckley, K.Naujoks, U. Michaelis, G.Dickson,
Gene Ther. 1998, 5, 542; c) B. Maiti, M. Kamra, A. A. Karande,
S. Bhattacharya, Org. Biomol. Chem. 2018, 16, 1983; d) J. Yang,
L.Huang, Gene Ther. 1998, 5, 380.
[117] N. F. Bouxsein, C. S. McAllister, K. K. Ewert, C. E. Samuel,
C. R.Safinya, Biochemistry 2007, 46, 4785.
[118] X.Gao, L.Huang, Biochem. Biophys. Res. Commun. 1991, 179, 280.
[119] a) S. Li, X.Gao, K.Son, F. Sorgi, H.Hofland, L.Huang, J. Con-
trolled Release 1996, 39, 373; b) C.Esposito, J.Generosi, G.Mossa,
A.Masotti, A. C.Castellano, Colloids Surf., B. 2006, 53, 187.
[120] N. J. Zuidam, Y. Barenholz, Biochim. Biophys. Acta, Biomembr.
1997, 1329, 211.
[121] D. Pozzi, C. Marchini, F. Cardarelli, F. Salomone, S. Coppola,
M. Montani, M. E. Zabaleta, M. Digman, E. Gratton,
V.Colapicchioni, Biochim. Biophys. Acta, Biomembr. 2014, 1838, 957.
[122] a) Y.Tan, L.Huang, J. Drug Targeting 2002, 10, 153; b) G. J.Nabel,
E. G.Nabel, Z.-Y.Yang, B. A.Fox, G. E.Plautz, X.Gao, L.Huang,
S.Shu, D.Gordon, A. E. Chang, Proc. Natl. Acad. Sci. USA 1993,
90, 11307.
[123] G. Settanni, W. Brill, H. Haas, F. Schmid, Macromol. Rapid
Commun. 2021, 2100683.
[124] a) D. Habrant, P. Peuziat, T. Colombani, L. Dallet, J. Gehin,
E. Goudeau, B. Evrard, O. Lambert, T. Haudebourg, B. Pitard,
J. Med. Chem. 2016, 59, 3046; b) L. Desigaux, M. Sainlos,
O.Lambert, R.Chevre, E.Letrou-Bonneval, J.-P.Vigneron, P.Lehn,
J.-M.Lehn, B.Pitard, Proc. Natl. Acad. Sci. USA 2007, 104, 16534.
[125] N. Tateshita, N. Miura, H. Tanaka, T. Masuda, S. Ohtsuki,
K. Tange, Y. Nakai, H. Yoshioka, H. Akita, J. Controlled Release
2019, 310, 36.
[126] S.Akhter, M.Berchel, P.-A.Jarès, P.Midoux, C.Pichon, Pharma-
ceutics 2022, 14, 581.
[127] D. P. Vangasseri, Z. Cui, W. Chen, D. A. Hokey, L. D. FaloJr,
L.Huang, Mol. Membr. Biol. 2006, 23, 385.
[128] Z.Du, M. M. Munye, A. D.Tagalakis, M. D.Manunta, S. L. Hart,
Sci. Rep. 2014, 4, 7107.
[129] H. Farhood, N. Serbina, L. Huang, Biochim. Biophys. Acta,
Biomembr. 1995, 1235, 289.
[130] D. Simberg, D. Danino, Y. Talmon, A. Minsky, M. E. Ferrari,
C. J.Wheeler, Y.Barenholz, J. Biol. Chem. 2001, 276, 47453.
[131] a) Y. Liu, L. C. Mounkes, H. D. Liggitt, C. S. Brown, I. Solodin,
T. D. Heath, R. J. Debs, Nat. Biotechnol. 1997, 15, 167;
b) B. Sternberg, K. Hong, W. Zheng, D. Papahadjopoulos,
Biochim. Biophys. Acta, Biomembr. 1998, 1375, 23.
[132] D. Pozzi, C. Marchini, F. Cardarelli, H. Amenitsch, C. Garulli,
A.Bifone, G.Caracciolo, Biochim. Biophys. Acta, Biomembr. 2012,
1818, 2335.
[133] L.Xu, T. J.Anchordoquy, Biochim. Biophys. Acta, Biomembr. 2008,
1778, 2177.
[134] G. Caracciolo, D. Pozzi, A. L. Capriotti, C. Cavaliere, A. Laganà,
J. Nanopart. Res. 2013, 15, 1498.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (36 of 38) © 2022 Wiley-VCH GmbH
[135] R. Koynova, B. Tenchov, in Nucleic Acid Transfection, Vol. 296
(Eds: W. Bielke, C. Erbacher), Springer, Berlin 2010, p. 51.
[136] J.Wang, X.Guo, Y.Xu, L.Barron, F. C.Szoka, J. Med. Chem. 1998,
41, 2207.
[137] M.Kang, H.Kim, C.Leal, Curr. Opin. Colloid Interface Sci. 2016, 26,
58.
[138] S. Eastman, C. Siegel, J. Tousignant, A. Smith, S. Cheng,
R.Scheule, Biochim. Biophys. Acta, Biomembr. 1997, 1325, 41.
[139] C. R.Safinya, Curr. Opin. Struct. Biol. 2001, 11, 440.
[140] J. O.Rädler, I.Koltover, T.Salditt, C. R. Safinya, Science 1997, 275,
810.
[141] I.Koltover, T.Salditt, J. O.Rädler, C. R.Safinya, Science 1998, 281, 78.
[142] A. J. Lin, N. L. Slack, A. Ahmad, C. X. George, C. E. Samuel,
C. R.Safinya, Biophys. J. 2003, 84, 3307.
[143] a) K. Ewert, A.Ahmad, H. M.Evans, C. R.Safinya, Expert Opin.
Biol. Ther. 2005, 5, 33; b) G. Caracciolo, D. Pozzi, R. Caminiti,
A.Congiu Castellano, Eur. Phys. J. 2003, 10, 331.
[144] B. Ma, S. Zhang, H. Jiang, B. Zhao, H. Lv, J. Controlled Release
2007, 123, 184.
[145] a) R. Koynova, L. Wang, R. C. MacDonald, Proc. Natl. Acad.
Sci. USA 2006, 103, 14373; b) R.Hulst, I. Muizebelt, P.Oosting,
C.van derPol, A.Wagenaar, J. Šmisterová, E.Bulten, C.Driessen,
D.Hoekstra, J. B.Engberts, Eur. J. Org. Chem. 2004, 2004, 835.
[146] B. M. Bruininks, P. C.Souza, H.Ingolfsson, S. J.Marrink, eLife
2020, 9, e52012.
[147] K. K.Ewert, H. M.Evans, A.Zidovska, N. F.Bouxsein, A.Ahmad,
C. R.Safinya, J. Am. Chem. Soc. 2006, 128, 3998.
[148] a) H.Kim, C.Leal, ACS Nano 2015, 9, 10214; b) H.Kim, J. Sung,
Y.Chang, A.Alfeche, C.Leal, ACS Nano 2018, 12, 9196.
[149] M.Riaz, Pak. J. Pharm. Sci. 1996, 9, 65.
[150] J. W.Meisel, G. W.Gokel, Sci. Rep. 2016, 6, 27662.
[151] S.Sriwongsitanont, M.Ueno, Open Colloid Sci. J. 2010, 4, 1.
[152] S.Batzri, E. D.Korn, Biochim. Biophys. Acta, Biomembr. 1973, 298,
1015.
[153] P.Ross, S.Hui, Gene Ther. 1999, 6, 651.
[154] M. R. Almofti, H. Harashima, Y. Shinohara, A. Almofti, W. Li,
H.Kiwada, Mol. Membr. Biol. 2003, 20, 35.
[155] a) Y.Maitani, S.Igarashi, M.Sato, Y. Hattori, Int. J. Pharm. 2007,
342, 33; b) S. Yang, J. Chen, D. Zhao, D. Han, X. Chen, Int. J.
Pharm. 2012, 434, 155; c) I.Tranchant, B.Thompson, C.Nicolazzi,
N.Mignet, D.Scherman, J. Gene Med. 2004, 6, S24.
[156] C.Krienke, L.Kolb, E. Diken, M. Streuber, S.Kirchho, T.Bukur,
Ö. Akilli-Öztürk, L. M. Kranz, H. Berger, J. Petschenka, Science
2021, 371, 145.
[157] a) V. Pector, J. Backmann, D. Maes, M. Vandenbranden,
J.-M.Ruysschaert, J. Biol. Chem. 2000, 275, 29533; b) E.Pozharski,
R. C.MacDonald, Biophys. J. 2003, 85, 3969.
[158] S. Even-Chen, R. Cohen, Y. Barenholz, Chem. Phys. Lipids 2012,
165, 414.
[159] J. Kim, J. Y. Kim, H. Kim, E. Kim, S. Park, K.-H. Ryu, E. G. Lee,
Int. J. Mol. Sci. 2021, 22, 12344.
[160] O. Zelphati, C.Nguyen, M.Ferrari, J. Felgner, Y.Tsai, P.Felgner,
Gene Ther. 1998, 5, 1272.
[161] M. T. Kennedy, E. V. Pozharski, V. A. Rakhmanova,
R. C.MacDonald, Biophys. J. 2000, 78, 1620.
[162] L.Wasungu, D.Hoekstra, J. Controlled Release 2006, 116, 255.
[163] J. M.Barichello, T.Ishida, H.Kiwada, in Liposomes: Methods and
Protocols (Ed: V. Weissig), Springer, Berlin 2010, p. 461.
[164] a) T. A. Balbino, N. T. Aoki, A. A. Gasperini, C. L. Oliveira,
A. R.Azzoni, L. P.Cavalcanti, G.Lucimara, Chem. Eng. J. 2013, 226,
423; b) T. A.Balbino, A. R.Azzoni, L. G.deLa Torre, Colloids Surf.,
B 2013, 111, 203; c) X.Liu, B.Bahloul, R. L. Kuen, K. Andrieux,
C.Roques, D.Scherman, Int. J. Pharm. 2021, 605, 120772.
[165] a) D. Llères, J. M. Weibel, D. Heissler, G. Zuber, G. Duportail,
Y.Mély, J. Gene Med. 2004, 6, 415; b) T. A.Balbino, J. M. Serafin,
A. A. Malfatti-Gasperini, C. L. de Oliveira, L. P. Cavalcanti,
M. B. de Jesus, L. G. de La Torre, Langmuir 2016, 32, 1799;
c) A. T.-H. Hsieh, N. Hori, R.Massoudi, P. J.-H.Pan, H.Sasaki,
Y. A.Lin, A. P.Lee, Lab Chip 2009, 9, 2638.
[166] V. Escriou, C. Ciolina, F. Lacroix, G. Byk, D. Scherman, P. Wils,
Biochim. Biophys. Acta, Biomembr. 1998, 1368, 276.
[167] P.Foroozandeh, A. A.Aziz, Nanoscale Res. Lett. 2018, 13, 339.
[168] a) S. Grabbe, H. Haas, M. Diken, L. M. Kranz, P. Langguth,
U. Sahin, Nanomedicine 2016, 11, 2723; b) J. C. Birchall,
I. W.Kellaway, S. N.Mills, Int. J. Pharm. 1999, 183, 195.
[169] L. M.Kranz, M.Diken, H.Haas, S.Kreiter, C.Loquai, K. C.Reuter,
M.Meng, D.Fritz, F.Vascotto, H.Hefesha, Nature 2016, 534, 396.
[170] a) A. Peletta, E. Prompetchara, K. Tharakhet, P. Kaewpang,
S. Buranapraditkun, T. Techawiwattanaboon, T. Jbilou,
P. Krangvichian, S. Sirivichayakul, S. Manopwisedjaroen,
Vaccines 2021, 9, 874; b) Z.Heidari, J. S.Arora, D.Datta, V. T.John,
N.Kumar, G. P.Bansal, Pharm. Res. 2017, 34, 1796.
[171] Q. Cheng, T. Wei, L. Farbiak, L. T. Johnson, S. A. Dilliard,
D. J.Siegwart, Nat. Nanotechnol. 2020, 15, 313.
[172] A. K. Blakney, P. F. McKay, B. I. Yus, Y. Aldon, R. J. Shattock,
Gene Ther. 2019, 26, 363.
[173] T. A.Balbino, J. M.Serafin, A.Radaic, M. B.deJesus, G.Lucimara,
Colloids Surf., B 2017, 152, 406.
[174] D. Simberg, S. Weisman, Y. Talmon, A. Faerman, T. Shoshani,
Y.Barenholz, J. Biol. Chem. 2003, 278, 39858.
[175] W.-Z. Zhou, D. Hoon, S. Huang, S. Fujii, K. Hashimoto,
R.Morishita, Y.Kaneda, Hum. Gene Ther. 1999, 10, 2719.
[176] A. De Beuckelaer, C. Pollard, S. Van Lint, K. Roose, L. Van
Hoecke, T. Naessens, V. K.Udhayakumar, M.Smet, N.Sanders,
S.Lienenklaus, Mol. Ther. 2016, 24, 2012.
[177] K. Okumura, M. Nakase, M. Inui, S. Nakamura, Y. Watanabe,
T.Tagawa, J. Gene Med. 2008, 10, 910.
[178] L.Van Hoecke, K.Roose, M.Ballegeer, Z.Zhong, N. N. Sanders,
S.De Koker, X.Saelens, S.Van Lint, Mol. Ther.– Nucleic Acids 2020,
22, 373.
[179] W. De Haes, J. Rejman, C. Pollard, C. Merlin, M. Vekemans,
E.Florence, S. C. De Smedt, J.Grooten, G. Vanham, S.De Koker,
Nanomedicine 2013, 8, 77.
[180] C. Pollard, J. Rejman, W. De Haes, B. Verrier, E. Van Gulck,
T. Naessens, S. De Smedt, P. Bogaert, J. Grooten, G. Vanham,
Mol. Ther. 2013, 21, 251.
[181] D. Christensen, K. S.Korsholm, I.Rosenkrands, T.Lindenstrøm,
P.Andersen, E. M.Agger, Expert Rev. Vaccines 2007, 6, 785.
[182] T. R. Smith, A. Patel, S. Ramos, D. Elwood, X. Zhu, J. Yan,
E. N.Gary, S. N.Walker, K.Schultheis, M.Purwar, Z.Xu, J.Walters,
P. Bhojnagarwala, M. Yang, N. Chokkalingam, P. Pezzoli,
E.Parzych, E. L.Reuschel, A.Doan, N.Tursi, M.Vasquez, J.Choi,
E.Tello-Ruiz, I.Maricic, M. A.Bah, Y.Wu, D.Amante, D. H.Park,
Y.Dia, A. R.Ali, etal., Nat. Commun. 2020, 11, 2601.
[183] D. J. Crommelin, T. J. Anchordoquy, D. B. Volkin, W. Jiskoot,
E.Mastrobattista, J. Pharm. Sci. 2021, 110, 997.
[184] K. Kubota, K. Onishi, K. Sawaki, T. Li, K. Mitsuoka, T. Sato,
S.Takeoka, Int. J. Nanomed. 2017, 12, 5121.
[185] A. K. Blakney, P. Deletic, P. F. McKay, C. R. Bouton, M.Ashford,
R. J.Shattock, A.Sabirsh, J. Controlled Release 2021, 330, 1250.
[186] A. A.De Angelis, S. J.Opella, Nat. Protoc. 2007, 2, 2332.
[187] L.Lin, X.Wang, Y. Guo, K.Ren, X.Li, L. Jing, X. Yue, Q.Zhang,
Z.Dai, RSC Adv. 2016, 6, 79811.
[188] T. N.Sut, E. R.Valle-González, B. K.Yoon, S.Park, J. A.Jackman,
N.-J.Cho, Appl. Mater. Today 2021, 22, 100947.
[189] a) K. Kolahdouzan, J. A. Jackman, B. K. Yoon, M. C. Kim,
M. S.Johal, N.-J. Cho, Langmuir 2017, 33, 5052; b) J. A.Jackman,
N.-J.Cho, Langmuir 2020, 36, 1387.
[190] E. J. Dufourc, Biochim. Biophys. Acta, Biomembr. 2021, 1863,
183478.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (37 of 38) © 2022 Wiley-VCH GmbH
[191] a) N. E. Gabriel, M. F. Roberts, Biochemistry 1984, 23, 4011;
b) C. R. Sanders, G. C. Landis, Biochemistry 1995, 34, 4030;
c) R. R.Vold, R. S.Prosser, J. Magn. Reson. 1996, 113, 267.
[192] C. R.Sanders, J. P.Schwonek, Biochemistry 1992, 31, 8898.
[193] T. A. Harroun, M. Koslowsky, M.-P. Nieh, C.-F. de Lannoy,
V.Raghunathan, J.Katsaras, Langmuir 2005, 21, 5356.
[194] U. H.Dürr, R.Soong, A.Ramamoorthy, Prog. Nucl. Magn. Reson.
Spectrosc. 2013, 69, 1.
[195] a) N. L. Yamada, M. Hishida, N. Torikai, Phys. Rev. E 2009,
79, 032902; b) U. H. Dürr, M. Gildenberg, A. Ramamoorthy,
Chem. Rev. 2012, 112, 6054.
[196] S.Ghosh, A.Ray, N.Pramanik, Biophys. Chem. 2020, 265, 106429.
[197] C.-H.Yang, T.-L.Lin, U.-S.Jeng, Langmuir 2019, 35, 9483.
[198] L. van Dam, G. Karlsson, K. Edwards, Biochim. Biophys. Acta,
Biomembr. 2004, 1664, 241.
[199] H. S.Cho, J. L.Dominick, M. M. Spence, J. Phys. Chem. B 2010,
114, 9238.
[200] M. N.Triba, P. F.Devaux, D. E.Warschawski, Biophys. J. 2006, 91, 1357.
[201] M.Li, W. T.Heller, C.-H.Liu, C. Y.Gao, Y.Cai, Y.Hou, M.-P.Nieh,
Biochim. Biophys. Acta, Biomembr. 2020, 1862, 183315.
[202] J.Struppe, J. A.Whiles, R. R.Vold, Biophys. J. 2000, 78, 281.
[203] Y. Liu, M. Li, Y. Yang, Y. Xia, M.-P.Nieh, Biochim. Biophys. Acta,
Biomembr. 2014, 1838, 1871.
[204] Z.Shen, H.Ye, M.Kröger, S.Tang, Y.Li, Nanoscale 2019, 11, 15971.
[205] a) W.Aresh, Y. Liu, J.Sine, D. Thayer, A.Puri, Y.Huang, Y.Wang,
M.-P. Nieh, J. Biomed. Nanotechnol. 2016, 12, 1852; b) X.Wang,
L.Lin, R.Liu, M. Chen, B.Chen, B.He, B. He, X. Liang, W. Dai,
H.Zhang, Adv. Funct. Mater. 2017, 27, 1700406.
[206] K. Yasuhara, S. Miki, H. Nakazono, A. Ohta, J.-i. Kikuchi, Chem.
Commun. 2011, 47, 4691.
[207] K.Yasuhara, H.Hayashi, J.-i.Kikuchi, Chem. Lett. 2012, 41, 1223.
[208] L. Lin, X. Wang, X. Li, Y. Yang, X. Yue, Q. Zhang, Z. Dai,
Bioconjugate Chem. 2017, 28, 53.
[209] P.-W. Yang, T.-L. Lin, T.-Y. Lin, C.-H. Yang, Y. Hu, U.-S. Jeng,
Soft Matter 2013, 9, 11542.
[210] P.-W.Yang, T.-L.Lin, Y.Hu, U.-S.Jeng, Soft Matter 2014, 10, 2313.
[211] A. T. Rad, S. Malik, L. Yang, T. K. Oberoi-Khanuja, M.-P. Nieh,
R.Bahal, Nanoscale 2019, 11, 12517.
[212] C. Janich, C. Wölk, S. Taßler, S. Drescher, A. Meister,
G. Brezesinski, B. Dobner, A. Langner, Eur. J. Lipid Sci. Technol.
2014, 116, 1184.
[213] C.Janich, S.Taßler, A.Meister, G.Hause, J.Schäfer, U.Bakowsky,
G.Brezesinski, C.Wölk, Soft Matter 2016, 12, 5854.
[214] C. Wölk, C. Janich, A. Meister, S. Drescher, A. Langner,
G.Brezesinski, U.Bakowsky, Bioconjugate Chem. 2015, 26, 2461.
[215] C. Wölk, C. Janich, D. Pawlowska, S. Drescher, A. Meister,
G.Hause, B. Dobner, A.Langner, G.Brezesinski, ChemPhysChem
2015, 16, 2115.
[216] Q. Chen, G. Guan, F. Deng, D. Yang, P. Wu, S. Kang, R. Sun,
X.Wang, D.Zhou, W.Dai, Biomaterials 2020, 251, 120008.
[217] H. Zhang, in Liposomes: Methods and Protocols (Ed: V. Weissig),
Springer, Berlin 2017, p. 17.
[218] Y. Watanabe, K. Aramaki, Y. Kadomatsu, K. Tanaka, Y. Konno,
Chem. Lett. 2016, 45, 558.
[219] I. K. Mkam Tsengam, M. Omarova, E. G. Kelley, A. McCormick,
G. D.Bothun, S. R.Raghavan, V. T.John, J. Phys. Chem. B 2022,
126, 2208.
[220] K. Aramaki, C. Iwata, J. Mata, T. Maehara, D. Aburano,
Y.Sakanishi, K.Kitao, Phys. Chem. Chem. Phys. 2017, 19, 23802.
[221] K. Aramaki, K. Adachi, M. Maeda, J. Mata, J. Kamimoto-Kuroki,
D.Tsukamoto, Y.Konno, Materials 2020, 13, 3066.
[222] J. Kamimoto-Kuroki, M. Yamashita, K. Tanaka, Y. Kadomatsu,
D. Tsukamoto, K. Aramaki, K. Adachi, Y. Konno, Colloids Surf., A
2020, 606, 125418.
[223] S.Choi, B.Kang, S.Taguchi, H.Umakoshi, K.Kim, M. K.Kwak,
H.-S.Jung, Langmuir 2021, 37, 12255.
[224] S. Taguchi, K. Suga, K. Hayashi, Y. Okamoto, H.-S. Jung,
H.Nakamura, H.Umakoshi, Colloids Interfaces 2018, 2, 73.
[225] S. Choi, B. Kang, T. Shimanouchi, K. Kim, H. Jung, Micro Nano
Syst. Lett. 2021, 9, 7.
[226] H. A.Benson, Curr. Drug Delivery 2005, 2, 23.
[227] L.Barbosa-Barros, C.Barba, M.Cócera, L.Coderch, C.López-Iglesias,
A.deLa Maza, O.López, Int. J. Pharm. 2008, 352, 263.
[228] L.Barbosa-Barros, A.De La Maza, J.Estelrich, A.Linares, M.Feliz,
P.Walther, R.Pons, O.López, Langmuir 2008, 24, 5700.
[229] a) G.Rodriguez, L. Barbosa-Barros, L.Rubio, M. Cocera, A.Diez,
J.Estelrich, R. Pons, J.Caelles, A. D. l.Maza, O.Lopez, Langmuir
2009, 25, 10595; b) G.Rodriguez, L.Rubio, M.Cócera, J.Estelrich,
R.Pons, A.de laMaza, O.López, Langmuir 2010, 26, 10578.
[230] G.Rodríguez, M.Cócera, L.Rubio, C.Alonso, R. Pons, C.Sandt,
P.Dumas, C.López-Iglesias, A.de laMaza, O.López, Phys. Chem.
Chem. Phys. 2012, 14, 14523.
[231] L. Barbosa-Barros, G. Rodríguez, C. Barba, M.Cócera, L.Rubio,
J.Estelrich, C.López-Iglesias, A.de laMaza, O.López, Small 2012,
8, 807.
[232] a) G.Rodriguez, M.Cocera, L.Rubio, C.Lopez-Iglesias, R. Pons,
A. de la Maza, O. López, Mol. Pharmaceutics 2012, 9, 482;
b) G. Rodríguez, L. Rubio, C. Barba, C. López-Iglesias,
A.de laMaza, O.López, M.Cócera, Eur. Biophys. J. 2013, 42, 333.
[233] L.Rubio, C. Alonso, G.Rodríguez, L. Barbosa-Barros, L.Coderch,
A.De la Maza, J.Parra, O.López, Int. J. Pharm. 2010, 386, 108.
[234] L. Rubio, C.Alonso, G.Rodríguez, M.Cócera, C. López-Iglesias,
L.Coderch, A.De la Maza, J.Parra, O.López, Int. J. Pharm. 2013,
444, 60.
[235] L. Rubio, G.Rodriguez, C.Alonso, C.Lopez-Iglesias, M. Cocera,
L.Coderch, A.De la Maza, J.Parra, O.Lopez, Soft Matter 2011, 7,
8488.
[236] L.Rubio, C.Alonso, G. Rodríguez, M.Cócera, L. Barbosa-Barros,
L.Coderch, A.de laMaza, J.Parra, O.López, Eur. J. Pharm. Biop-
harm. 2014, 86, 212.
[237] a) G. Rodríguez, L. Barbosa-Barros, L. Rubio, M. Cócera,
F. Fernández-Campos, A.Calpena, E. Fernández, A.De La Maza,
O.López, J. Biomed. Nanotechnol. 2015, 11, 282; b) M. S.Kapoor,
S.GuhaSarkar, R.Banerjee, Ther. Delivery 2017, 8, 701.
[238] V. Moner, E. Fernández, A. Del Pozo, G. Rodríguez, M. Cócera,
A.de laMaza, O.López, Contact Dermatitis 2017, 77, 25.
[239] G. Anderluzzi, G. Lou, S. Woods, S. T. Schmidt, S. Gallorini,
M. Brazzoli, R. Johnson, C. W. Roberts, D. T. O’Hagan,
B. C.Baudner, J. Controlled Release 2022, 342, 388.
[240] W. Zhang, J. Sun, Y. Liu, M. Tao, X. Ai, X. Su, C. Cai, Y.Tang,
Z.Feng, X.Yan, Mol. Pharmaceutics 2014, 11, 3279.
[241] A. T. Rad, D. Hargrove, L. Daneshmandi, A. Ramsdell, X. Lu,
M.-P.Nieh, Adv. NanoBiomed Res. 2022, 2, 2100080.
[242] J.Gao, C.Xie, M. Zhang, X.Wei, Z.Yan, Y. Ren, M.Ying, W. Lu,
Nanoscale 2016, 8, 7209.
[243] P. E.Saw, M.Yu, M.Choi, E.Lee, S.Jon, O. C.Farokhzad, Bioma-
terials 2017, 123, 118.
[244] V. M. Stees, Z. Zhang, S. MacDonald, J. Crowe, K. K. Ewert,
B.Carragher, C. S.Potter, C. R.Safinya, ACS Appl. Mater. Interfaces
2019, 12, 151.
[245] L.Lin, X.Liang, Y. Xu, Y. Yang, X.Li, Z. Dai, Bioconjugate Chem.
2017, 28, 2410.
[246] S.Hameed, P.Bhattarai, Z.Dai, Front. Chem. 2018, 6, 127.
[247] J. D. Finn, A. R. Smith, M. C. Patel, L. Shaw, M. R. Youniss,
J. van Heteren, T. Dirstine, C. Ciullo, R. Lescarbeau, J. Seitzer,
R. R. Shah, A. Shah, D. Ling, J. Growe, M. Pink, E. Rohde,
K. M. Wood, W. E. Salomon, W. F.Harrington, C.Dombrowski,
W. R.Strapps, Y.Chang, D. V.Morrissey, Cell Rep. 2018, 22, 2227.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.afm-journal.dewww.advancedsciencenews.com
2203669 (38 of 38) © 2022 Wiley-VCH GmbH
Abdul Rahim Ferhan, Ph.D. is a research fellow in the Translational Materials Innovation Group in
the School of Materials Science and Engineering at Nanyang Technological University. He received
his B.Eng. degree in Chemical and Biomolecular Engineering and Ph.D. degree in Biomedical
Engineering from the School of Chemical and Biomedical Engineering at Nanyang Technological
University. His research focuses on lipid membrane interactions and the incorporation of lipid
membrane models within novel nanoplasmonic sensing platforms for fundamental investigations
into biomacromolecular interaction processes.
Soohyun Park, Ph.D. is a research fellow in the School of Materials Science and Engineering,
Nanyang Technological University. She earned her B.S. in Biomaterials Engineering from Seoul
National University in 2016 and Ph.D. in Materials Science and Engineering from Nanyang
Technological University in 2021. Her research interests include lipid membrane-targeting antiviral
peptides, lipid nanoparticles, model lipid membranes, and biomaterials.
Joshua A. Jackman, Ph.D. is an assistant professor in the School of Chemical Engineering and
Director of the Translational Nanobioscience Research Center at Sungkyunkwan University. He
earned his B.S. degree in Chemistry from the University of Florida in 2010 and his Ph.D. degree in
Materials Science and Engineering from Nanyang Technological University in 2015, and completed
postdoctoral studies at the Stanford University School of Medicine. His research focuses on lipid
membrane biotechnology, membrane biophysics, and the development of membrane-targeting
strategies for infectious disease and cancer applications.
Nam-Joon Cho, Ph.D. is the MRS-Singapore Chair Professor in the School of Materials Science
and Engineering at Nanyang Technological University. He earned his B.S. degree in Civil
Engineering from the University of California, Berkeley in 1996 and his M.S. degree in Materials
Science and Engineering and Ph.D. degree in Chemical Engineering from Stanford University
in 2003 and 2007, respectively, and completed postdoctoral studies at the Stanford University
School of Medicine. His research focuses on biomaterial strategies to develop new classes of
anti-infective drugs, to engineer advanced lipid nanoparticle technology, and to create natural
materials to replace plastics.
Adv. Funct. Mater. 2022, 32, 2203669
16163028, 2022, 37, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adfm.202203669 by Sungkyunkwan University, Wiley Online Library on [25/12/2023]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
... For example, recent papers advocating nanoarchitectonics can be found in a wide range of fields. In addition to application-oriented areas such as catalysis [131][132][133][134][135], sensors [136][137][138][139][140], devices [141][142][143][144][145], energy production [146][147][148][149][150], energy storage [151][152][153][154][155], environmental response [156][157][158][159][160], drug delivery [161][162][163][164][165], and biomedical applications [166][167][168][169][170], there are also fundamental areas such as material synthesis [171][172][173][174][175][176], structural control [177][178][179][180][181], the exploration of physical phenomena [182][183][184][185][186], relatively basic biochemical studies [187][188][189][190][191], and research on cellular interactions [192][193][194][195][196]. Since all matter is principally composed of atoms and molecules, the methodology of building matter from atoms and molecules is applicable to all material synthesis. It could be likened to the ultimate theory of everything in physics [197], and nanoarchitectonics could be called a method for everything in materials science [198,199]. ...
Article
Full-text available
The characteristic feature of a biofunctional system is that components with various functions work together. These multi-components are not simply mixed together, but are rationally arranged. The fundamental technologies to do this in an artificial system include the synthetic chemistry of the substances that make the component unit, the science and techniques for assembling them, and the technology for analyzing their nanoostructures. A new concept, nanoarchitectonics, can play this role. Nanoarchitectonics is a post-nanotechnology concept that involves building functional materials that reflect the nanostructures. In particular, the approach of combining and building multiple types of components to create composite materials is an area where nanoarchitectonics can be a powerful tool. This review summarizes such examples and related composite studies. In particular, examples are presented in the areas of catalyst & photocatalyst, energy, sensing & environment, bio & medical, and various other functions and applications to illustrate the potential for a wide range of applications. In order to show the various stages of development, the examples are not only state-of-the-art, but also include those that are successful developments of existing research. Finally, a summary of the examples and a brief discussion of future challenges in nanoarchitectonics will be given. Nanoarchitectonics is applicable to all materials and aims to establish the ultimate methodology of materials science.
... This partly agrees with published literature which argues that LNP size is primarily influenced by concentration, flow rate (for particles synthesized using a microfluidic system), and PEG-lipid concentration. However, there has been speculation that differences in cargo size and morphology should affect LNP size 8 . This is because compartmental organization of lipids and cargo within an LNP differs based on cargo size 39,40 . ...
Article
Full-text available
With the recent success of lipid nanoparticle (LNP) based SARS-CoV-2 mRNA vaccines, the potential for RNA therapeutics has gained widespread attention. LNPs are promising non-viral delivery vectors to protect and deliver delicate RNA therapeutics, which are ineffective and susceptible to degradation alone. While food and drug administration (FDA) approved formulations have shown significant promise, benchmark lipid formulations still require optimization and improvement. In addition, the translatability of these formulations for several different RNA cargo sizes has not been compared under the same conditions. Herein we analyze “gold standard” lipid formulations for encapsulation efficiency of various non-specific RNA cargo lengths representing antisense oligonucleotides (ASO), small interfering RNA (siRNA), RNA aptamers, and messenger RNA (mRNA), with lengths of 10 bases, 21 base pairs, 96 bases, 996 bases, and 1929 bases, respectively. We evaluate encapsulation efficiency as the percentage of input RNA encapsulated in the final LNP product (EEinput%), which shows discrepancy with the traditional calculation of encapsulation efficiency (EE%). EEinput% is shown to be < 50% for all formulations tested, when EE% is consistently > 85%. We also compared formulations for LNP size (Z-average) and polydispersity index (PDI). LNP size does not appear to be strongly influenced by cargo size, which is a counterintuitive finding. Thoughtful characterization of LNPs, in parallel with consideration of in vitro or in vivo behavior, will guide design and optimization for better understanding and improvement of future RNA therapeutics.
Article
Full-text available
Harnessing surface engineering strategies to functionalize nucleic acid-lipid nanoparticles (LNPs) for improved performance has been a hot research topic since the approval of the first siRNA drug, patisiran, and two mRNA-based COVID-19 vaccines, BNT162b2 and mRNA-1273. Currently, efforts have been mainly made to construct targeted LNPs for organ- or cell-type-specific delivery of nucleic acid drugs by conjugation with various types of ligands. In this review, we describe the surface engineering strategies for nucleic acid-LNPs, considering ligand types, conjugation chemistries, and incorporation methods. We then outline the general purification and characterization techniques that are frequently used following the engineering step and emphasize the specific techniques for certain types of ligands. Next, we comprehensively summarize the currently accessible organs and cell types, as well as the other applications of the engineered LNPs. Finally, we provide considerations for formulating targeted LNPs and discuss the challenges of successfully translating the “proof of concept” from the laboratory into the clinic. We believe that addressing these challenges could accelerate the development of surface-engineered LNPs for targeted nucleic acid delivery and beyond.
Article
Full-text available
It has become clear that superior material functions are derived from precisely controlled nanostructures. This has been greatly accelerated by the development of nanotechnology. The next step is to assemble materials with knowledge of their nano-level structures. This task is assigned to the post-nanotechnology concept of nanoarchitectonics. However, nanoarchitectonics, which creates intricate three-dimensional functional structures, is not always easy. Two-dimensional nanoarchitectonics based on reactions and arrangements at the surface may be an easier target to tackle. A better methodology would be to define a two-dimensional structure and then develop it into a three-dimensional structure and function. According to these backgrounds, this review paper is organized as follows. The introduction is followed by a summary of the three issues; (i) 2D to 3D dynamic structure control: liquid crystal commanded by the surface, (ii) 2D to 3D rational construction: a metal–organic framework (MOF) and a covalent organic framework (COF); (iii) 2D to 3D functional amplification: cells regulated by the surface. In addition, this review summarizes the important aspects of the ultimate three-dimensional nanoarchitectonics as a perspective. The goal of this paper is to establish an integrated concept of functional material creation by reconsidering various reported cases from the viewpoint of nanoarchitectonics, where nanoarchitectonics can be regarded as a method for everything in materials science.
Article
Full-text available
Nanotechnology has advanced the techniques for elucidating phenomena at the atomic, molecular, and nano-level. As a post nanotechnology concept, nanoarchitectonics has emerged to create functional materials from unit structures. Consider the material function when nanoarchitectonics enables the design of materials whose internal structure is controlled at the nanometer level. Material function is determined by two elements. These are the functional unit that forms the core of the function and the environment (matrix) that surrounds it. This review paper discusses the nanoarchitectonics of confined space, which is a field for controlling functional materials and molecular machines. The first few sections introduce some of the various dynamic functions in confined spaces, considering molecular space, materials space, and biospace. In the latter two sections, examples of research on the behavior of molecular machines, such as molecular motors, in confined spaces are discussed. In particular, surface space and internal nanospace are taken up as typical examples of confined space. What these examples show is that not only the central functional unit, but also the surrounding spatial configuration is necessary for higher functional expression. Nanoarchitectonics will play important roles in the architecture of such a total system.
Article
Full-text available
Recently, mRNA‐based therapeutics, including vaccines, have gained significant attention in the field of gene therapy for treating various diseases. Among the various mRNA delivery vehicles, lipid nanoparticles (LNPs) have emerged as promising vehicles for packaging and delivering mRNA with low immunogenicity. However, while mRNA delivery has several advantages, the delivery efficiency and stability of LNPs remain challenging for mRNA therapy. In this study, an ionizable helper cholesterol analog, 3β[L‐histidinamide‐carbamoyl] cholesterol (Hchol) lipid is developed and incorporated into LNPs instead of cholesterol to enhance the LNP potency. The pKa values of the Hchol‐LNPs are ≈6.03 and 6.61 in MC3‐ and SM102‐based lipid formulations. Notably, the Hchol‐LNPs significantly improve the delivery efficiency by enhancing the endosomal escape of mRNA. Additionally, the Hchol‐LNPs are more effective in a red blood cell hemolysis at pH 5.5, indicating a synergistic effect of the protonated imidazole groups of Hchol and cholesterol on endosomal membrane destabilization. Furthermore, mRNA delivery is substantially enhanced in mice treated with Hchol‐LNPs. Importantly, LNP‐encapsulated SARS‐CoV‐2 spike mRNA vaccinations induce potent antigen‐specific antibodies against SARS‐CoV‐2. Overall, incorporating Hchol into LNP formulations enables efficient endosomal escape and stability, leading to an mRNA delivery vehicle with a higher delivery efficiency.
Article
Full-text available
mRNA, as one of the foci of biomedical research in the past decade, has become a candidate vaccine solution for various infectious diseases and tumors and for regenerative medicine and immunotherapy due to its high efficiency, safety, and effectiveness. A stable and effective delivery system is needed to protect mRNAs from nuclease degradation while also enhancing immunogenicity. The success of mRNA lipid nanoparticles in treating COVID-19, to a certain extent, marks a milestone for mRNA vaccines and also promotes further research on mRNA delivery systems. Here, we explore mRNA vaccine delivery systems, especially lipid nanoparticles (LNPs), considering the current research status, prospects, and challenges of lipid nanoparticles, and explore other mRNA delivery systems.
Article
Full-text available
Liposomes are widely used in the biological field due to their good biocompatibility and surface modification properties. With the development of biochemistry and material science, many liposome structures and their surface functional components have been modified and optimized one by one, pushing the liposome platform from traditional to functionalized and intelligent, which will better satisfy and expand the needs of scientific research. However, a main limiting factor effecting the efficiency of liposomes is the complicated environmental conditions in the living body. Currently, in order to overcome the above problem, functionalized liposomes have become a very promising strategy. In this paper, binding strategies of liposomes with four main functional elements, namely nucleic acids, antibodies, peptides, and stimuli‐responsive motif have been summarized for the first time. In addition, based on the construction characteristics of functionalized liposomes, such as drug‐carrying, targeting, long‐circulating, and stimulus‐responsive properties, a comprehensive overview of their features and respective research progress are presented. Finally, the paper critically presents the limitations of these functionalized liposomes in the current applications and also prospectively suggests the future development directions, aiming to accelerate realization of their industrialization.
Article
Full-text available
The phospholipid lecithin (L) and the nonionic surfactant Tween 80 (T) are used together in various contexts, including in drug delivery and oil spill remediation. There is hence a need to elucidate the nanostructures in LT mixtures, which is the focus of this paper. We study these mixtures using cryogenic transmission electron microscopy (cryo-TEM), coupled with dynamic light scattering and small-angle neutron scattering. As the concentration of Tween 80 is increased, the vesicles formed by lecithin are transformed into spherical micelles. We identify bicelles (i.e., disc-like micelles) as well as cylindrical micelles as the key stable nanostructures formed at intermediate L/T ratios. The bicelles have diameters ∼13-26 nm, and the bicelle size decreases as the Tween 80 content increases. We propose that the lecithin lipids form the body of the discs, while the Tween 80 surfactants occupy the rims. This hypothesis is consistent with geometric arguments because lecithin is double-tailed and favors minimal curvature, whereas the single-tailed Tween 80 molecules prefer curved interfaces. In the case of cylindrical micelles, cryo-TEM reveals that the micelles are short (length < 22 nm) and flexible. We are able to directly visualize the microstructure of the aggregates formed by lecithin-Tween 80 mixtures, thereby enhancing the understanding of morphological changes in the lecithin-Tween 80 system.
Article
Full-text available
Cationic liposomes are attractive carriers for mRNA delivery. Here, mRNA lipoplexes (LX) were prepared with the cationic lipids α-aminolipophosphonate (3b) or imidazolium lipophosphoramidate (2) associated with various α-aminolipophosphonates co-lipids comprising protonable groups (imidazole or pyridine) and DOPE. Physicochemical parameters of liposomes and their membrane fusion activity were measured. LXs comprising either 3b- or 2- allowed transfection of ~25% and 40% of dendritic cells with low cytotoxicity, respectively; the efficiency increased up to 80% when 2 was combined with the imidazole-based co-lipid 1. The transfections were high with 3b/1, 3b/DOPE, 2/1 and 2/DOPE LXs. We observed that the transfection level was not well correlated with the acid-mediated membrane fusion activity of liposomes supposed to destabilize endosomes. The mRNA release from LXs and its translation capacity after release were studied for the most efficient LXs. The results showed that the more mRNA was condensed, the poorer the translation efficiency after release was. In contrast to DNA, circular dichroism performed on mRNA complexed with 2/DOPE revealed the presence of denatured mRNA in LXs explaining this lack of translation efficiency. This is an important parameter that should be stressed for the preparation of mRNA LXs with a conserved mRNA translation activity.
Article
Full-text available
The coronavirus disease of 2019 (COVID-19) pandemic launched an unprecedented global effort to rapidly develop vaccines to stem the spread of the novel severe acute respiratory syndrome coronavirus (SARS-CoV-2). Messenger ribonucleic acid (mRNA) vaccines were developed quickly by companies that were actively developing mRNA therapeutics and vaccines for other indications, leading to two mRNA vaccines being not only the first SARS-CoV-2 vaccines to be approved for emergency use but also the first mRNA drugs to gain emergency use authorization and to eventually gain full approval. This was possible partly because mRNA sequences can be altered to encode nearly any protein without significantly altering its chemical properties, allowing the drug substance to be a modular component of the drug product. Lipid nanoparticle (LNP) technology required to protect the ribonucleic acid (RNA) and mediate delivery into the cytoplasm of cells is likewise modular, as are technologies and infrastructure required to encapsulate the RNA into the LNP. This enabled the rapid adaptation of the technology to a new target. Upon the coattails of the clinical success of mRNA vaccines, this modularity will pave the way for future RNA medicines for cancer, gene therapy, and RNA engineered cell therapies. In this review, trends in the publication records and clinical trial registrations are tallied to show the sharp intensification in preclinical and clinical research for RNA medicines. Demand for the manufacturing of both the RNA drug substance (DS) and the LNP drug product (DP) has already been strained, causing shortages of the vaccine, and the rise in development and translation of other mRNA drugs in the coming years will exacerbate this strain. To estimate demand for DP manufacturing, the dosing requirements for the preclinical and clinical studies of the two approved mRNA vaccines were examined. To understand the current state of mRNA-LNP production, current methods and technologies are reviewed, as are current and announced global capacities for commercial manufacturing. Finally, a vision is rationalized for how emerging technologies such as self-amplifying mRNA, microfluidic production, and trends toward integrated and distributed manufacturing will shape the future of RNA manufacturing and unlock the potential for an RNA medicine revolution.
Article
Full-text available
Complement opsonization is among the biggest challenges facing nanomedicine. Nearly instantly after injection into blood, nanoparticles are opsonized by the complement protein C3, leading to clearance by phagocytes, fouling of targeting moieties, and release of anaphylatoxins. While surface polymers such as polyethylene glycol (PEG) partially decrease complement opsonization, most nanoparticles still suffer from extensive complement opsonization, especially when linked to targeting moieties. To ameliorate the deleterious effects of complement, we have conjugated two of mammals’ natural regulators of complement activation (RCAs), Factors H & I, to the surface of nanoparticles. In vitro, Factor H or I conjugation to PEG-coated nanoparticles decrease their C3 opsonization, and markedly reduce nanoparticle uptake by phagocytes. In an in vivo mouse model of sepsis-induced lung injury, Factor I-conjugation abrogates nanoparticle uptake by intravascular phagocytes in the lungs, allowing the blood concentration of the nanoparticle to remain elevated much longer. For nanoparticles targeted to the lung's endothelium by conjugation to anti-ICAM antibodies, Factor I-conjugation shifts the cell type distribution away from phagocytes and towards endothelial cells. Finally, Factor I-conjugation abrogates the severe anaphylactoid responses common to many nanoparticles, preventing systemic capillary leak and preserving blood flow to visceral organs and the brain. Thus, conjugation of RCAs, like Factor I, to nanoparticles is likely to help in nanomedicine's long battle against complement, improving several key parameters critical for clinical success. This article is protected by copyright. All rights reserved
Article
Companies advance tRNA therapeutics to overcome mutant stoppages in protein synthesis shared by thousands of genetic diseases and cancers.
Article
In 2021, mRNA vaccines against COVID-19 were approved by the Food and Drug Administration. mRNA vaccines are important for preventing severe COVID-19 and returning to normal life. The development of RNA-delivery technology, including mRNA vaccines, has been investigated worldwide for ~30 years. Lipid nanoparticles (LNPs) are a breakthrough technology that stably delivers RNA to target organs, and RNA-loaded LNP-based nanomedicines have been studied for the development of vaccines and nanomedicines for RNA-, gene-, and cell-based therapies. Recently, microfluidic devices and technologies have attracted attention for the production of LNPs, particularly RNA-loaded LNPs. Microfluidics provides many advantages for RNA-loaded LNP production, including precise LNP size controllability, high reproducibility, high-throughput optimization of LNP formulation, and continuous LNP-production processes. In this review, we summarize microfluidic-based RNA-loaded LNP production and its applications in RNA-based therapy and genome editing.
Article
Significance Liver accumulation represents a significant barrier in the development of therapeutically efficacious nanoparticle drug delivery systems. Using a series of lipid nanoparticles with distinct organ-targeting properties, we provide evidence for a plausible mechanism of action for nanoparticle delivery to non-liver tissues. Following intravenous injection, specific proteins in the blood are recruited to the nanoparticle’s surface based on its molecular composition and they endow it with a unique biological identity that governs its ultimate fate in the body. An innovative paradigm emerges from this mechanistic understanding of nanoparticle delivery—endogenous targeting—wherein the molecular composition of a nanoparticle is rationally engineered to interact with specific proteins in the blood to overcome liver accumulation and to target specific organs.
Article
The efficacy of RNA-based vaccines has been recently demonstrated, leading to the use of mRNA-based COVID-19 vaccines. The application of self-amplifying mRNA within these formulations may offer further enhancement to these vaccines, as self-amplifying mRNA replicons enable longer expression kinetics and more potent immune responses compared to non-amplifying mRNAs. To investigate the impact of administration route on RNA-vaccine potency, we investigated the immunogenicity of a self-amplifying mRNA encoding the rabies virus glycoprotein encapsulated in different nanoparticle platforms (solid lipid nanoparticles (SLNs), polymeric nanoparticles (PNPs) and lipid nanoparticles (LNPs)). These were administered via three different routes: intramuscular, intradermal and intranasal. Our studies in a mouse model show that the immunogenicity of our 4 different saRNA vaccine formulations after intramuscular or intradermal administration was initially comparable; however, ionizable LNPs gave higher long-term IgG responses. The clearance of all 4 of the nanoparticle formulations from the intramuscular or intradermal administration site was similar. In contrast, immune responses generated after intranasal was low and coupled with rapid clearance for the administration site, irrespective of the formulation. These results demonstrate that both the administration route and delivery system format dictate self-amplifying RNA vaccine efficacy.