Effect of Micro - and Nanoparticle Shape On Biological Processes

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Journal of Controlled Release 342 (2022) 93–110

Contents lists available at ScienceDirect

Journal of Controlled Release


journal homepage: www.elsevier.com/locate/jconrel

Review article

Effect of micro- and nanoparticle shape on biological processes


Hicheme Hadji, Kawthar Bouchemal *
Université Paris-Saclay, Institut Galien Paris Saclay, CNRS UMR 8612, 92296 Châtenay-Malabry, France

A R T I C L E I N F O A B S T R A C T

Keywords: In the drug delivery field, there is beyond doubt that the shape of micro- and nanoparticles (M&NPs) critically
Nanoparticle affects their biological fate. Herein, following an introduction describing recent technological advances for
Shape designing nonspherical M&NPs, we highlight the role of particle shape in cell capture, subcellular distribution,
Cytotoxicity
intracellular drug delivery, and cytotoxicity. Then, we discuss theoretical approaches for understanding the
Cell internalization
Biodistribution
effect of particle shape on internalization by the cell membrane. Subsequently, recent advances on shape-
Pharmacokinetics dependent behaviors of M&NPs in the systemic circulation are detailed. In particular, the interaction of
M&NPs with blood proteins, biodistribution, and circulation under flow conditions are analyzed. Finally, the
hurdles and future directions for developing nonspherical M&NPs are underscored.

1. Introduction nonspherical particle shape to Janus properties [6]. Each particle face
can be functionalized differently and independently [6]. In another
Since the 60’, micro- and nanoparticles (M&NPs) have had a embodiment, the self-assembly of hydrophobically modified poly­
powerful impact in biomedical applications such as imaging, diagnosis, saccharide and α-cyclodextrin in water resulted from hexagonal-shaped
and drug delivery. Since then, various classes of spherical M&NPs have M&NPs, referred to as micro- and nanoplatelets [7–14]. It was possible
been designed, mainly by tuning compositions, sizes, and surface to modify surface composition by changing the nature of the poly­
properties. More recently, a new criterion to control biological processes saccharide, while the size was adjusted by acting on the stirring duration
has been added by conceiving M&NPs with specific shapes. Synthetic of the starting materials [12]. Top-down approaches, on the other hand,
nonspherical M&NPs have been generated using different fabrication are based on a controlled shaping of starting materials using different
techniques divided into two categories: bottom-up and top-down ap­ techniques such as lithography [15–18], and physical stretching of
proaches. Bottom-up fabrication relies on the association of specifically spherical particles [19–23].
designed molecules into ordered structures. Chemical synthesis of par­ All the available techniques have offered the possibility for a precise
ticles with complex architectures (e.g., carbon nanotubes, inorganic modification of particle shape, in addition to chemical, physicochem­
particles) is based on bottom-up fabrication. Self-assembly of amphi­ ical, and physical properties. Consequently, the researchers considered
philic molecules constitutes the second group of bottom-up approaches. particle shape as a novel parameter for controlling biological processes.
For example, disk liposomes [1] and filamentous micelles [2] were ob­ Over the last decade, we have learned that particle shape mediates its
tained by self-assembly processes. More recently, polymerization- motion in biological fluids in static conditions [10,24,25], and under a
induced self-assembly (PISA) [3,4] offered the possibility to tune par­ shear flow [26–30], interactions with the cell membrane [31–34],
ticle shape, size, and glass transition temperature [5] by varying the endocytosis pathway [18–21,35–41], and intracellular distribution
composition of starting materials during polymerization. In a pionner [42–45]. Herein, we will first review how M&NP shape impacts endo­
investigation, a novel class of particles was designed by combining cytosis. Both phagocytic and nonphagocytic pathways were

Abbreviations: AFM, Atomic force microscopy; AR, aspect ratio; Au, gold; CARPA, complement activation-related pseudoallergy; CLSM, Confocal laser scanning
microscopy.; CTAB, Cetyl trimethylammonium bromide; DOPE, dioleoylphosphatidylethanolamine; DOTAP, 1,2-dioleoyl-3-trimethylammoniumpropane; ICAM-1,
intracellular adhesion molecule; IPC-AES, Inductively coupled plasma-atomic emission spectrometer; M&NPs, micro- and nanoparticles; MPS, mononuclear
phagocytic system; PEG, polyethylene glycol; PGS, poly(glycerol sebacate); PISA, polymerization-induced self-assembly; PLGA, poly(lactic acid)-co-(glycolic acid);
PRINT, Particle Replication In Nonwetting Templates; ROS, reactive oxygen species; SWNTs, single-walled carbon nanotubes; TEM, Transmission electron
microscopy.
* Corresponding author.
E-mail address: [email protected] (K. Bouchemal).

https://doi.org/10.1016/j.jconrel.2021.12.032
Received 17 September 2021; Received in revised form 23 December 2021; Accepted 24 December 2021
Available online 29 December 2021
0168-3659/© 2021 Elsevier B.V. All rights reserved.
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

underscored. Then, we will give a comprehensive analysis of whether duration in the blood. Finally, we will discuss whether particle shape
particle shape impacts subcellular distribution and cytotoxicity. As an could dictate targeting a specific organ and increase therapeutic efficacy
endocytosis process begins with the interaction of the particle with the by encapsulating a drug.
cell membrane followed by the membrane deformation, a section will be
dedicated to the multiscale modeling of the mechanics and energies 2. Endocytosis
occurring during a wrapping episode. The complexity of the interaction
between a particle and the cell membrane due to the shape effect will be 2.1. General overview
discussed.
In this review, special attention was paid to the events occurring Endocytosis is a general term describing a fundamental cellular
immediately after administering a particle in the body, even before process by which the eukaryotic cells internalize substances from the
interacting with the cell membrane. Immediately after their adminis­ extracellular environment. The substances to be engulfed are sur­
tration, the biomolecules present in the biological media (e.g., proteins, rounded by pieces of the cell membrane and carried into the cell in
lipids, and sugars) will competitively bind to the M&NP surface leading vesicular structures that eventually pinch off the membrane inside the
to a biomolecular corona. Adsorption of blood proteins is most relevant cell (Fig. 1). Endocytosis plays a critical role in nutrient uptake, surface
since blood contains thousands of different proteins, each of which may receptor regulation, cell motility, cell adhesion and orchestrates several
potentially interact with a particle through non-covalent interactions cell signalling pathways. Viruses and bacteria exploit this process to
[46,47]. The protein corona controls pharmacokinetics, biodistribution, enter the cells.
interaction with cells, intracellular trafficking, targeting capabilities Endocytosis is usually subdivided into two internalization mecha­
[48,49], biological activity, immunological reactions [50], and toxicity nisms: ‘phagocytosis’ or cell eating and ‘pinocytosis’ or cell drinking.
[48]. In this review, we will explore how nonspherical shape impacts Pinocytosis is also referred to as a nonphagocytic pathway. Phagocytosis
particle behaviors in the systemic circulation, including interaction with is mainly undertaken by professional phagocytes of the immune system,
blood protein, biodistribution, and circulation under a shear flow. We including neutrophils, macrophages, and mature dendritic cells [51]. In
will analyze the possibility of manipulating the M&NP shape to escape addition to professional phagocytes, the engulfment of cells by
the complement cascade’s activation and prolong the circulation nonprofessional phagocytic cells has been described for decades. In this

Fig. 1. (A) Illustration of the main nonphagocytic pathways in eukaryotic cells.

94
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

process, living cells engulf other living or dying cells. High rates of 2.2. Shape-dependent endocytosis
nonprofessional phagocytic pathways were reported in cancer cells. This
phenomenon was called cell cannibalism [52,53], which is defined as ‘a While it is beyond doubt that particle size and surface properties
cell that is contained within another bigger cell with a crescent-shaped significantly impact their cellular internalization pathway, shape-
nucleus’. dependent cell uptake has been studied more recently. Investigating
Pinocytosis occurs in all cells according to four main engulfment the impact of particle shape on phagocytosis determines whether un­
mechanisms depending on the different proteins and lipids involved: desirable capture by the mononuclear phagocytic system (MPS) can be
macropinocytosis, clathrin-mediated endocytosis, caveolae-mediated limited. In contrast, the cell capture of M&NPs can dictate therapeutic
endocytosis, and clathrin- and caveolae-independent endocytosis. applications for intracellular drug delivery. The mechanism by which
Clathrin-mediated endocytosis is the most elucidated mechanism for the M&NPs are captured dictates intracellular transport and the possi­
nonphagocytic pathways of M&NPs and macromolecules. Unlike bility to target specific intracellular compartments by acting on particle
phagocytosis, all mammalian cells can ensure this uptake pathway [54], shape, as detailed in section 2.3. Although there is evidence that M&NP
which mainly engages receptor-ligand complexes. Endocytic coat pro­ shape affects the rate and the mechanism of particle endocytosis
teins in the cytosol start to cluster on the inner plasma membrane upon [18,62–67], contradictory data were reported on the shape-dependency
binding of a ligand to its receptor on the cell membrane. The process of M&NPs cellular uptake. Previous works showed that the shape of
continues with further recruitment of coat proteins from the cytosol. The M&NPs affects phagocytosis [21], internalization by nonphagocytic
process involves a triskelion-shaped protein called clathrin. The flat cells [18], and intracellular transport [63]. In contrast, other works re­
membrane transforms on a ‘clathrin-coated pit’ following the self- ported that M&NP geometry does not play a dominant role in their
organization of clathrin into cage-like structures (Fig. 1). Subse­ cellular uptake [68]. Hereafter, the influence of M&NPs shape on
quently, the pit becomes deeper until it pinches inside the cytoplasm as cellular uptake was analyzed and discussed.
vesicles (120 nm size) that are typically delivered to the early endo­
somes (See Transmission electron microscopy (TEM) images in Fig. 2). 2.2.1. Phagocytosis of nonspherical particles
For review article on clathrin-mediated endocytosis pathway see refer­ The phagocytosis process is critical in eliminating invading patho­
ence [54]. gens such as viruses, bacteria, yeasts, and dead cells [69]. It is also
Caveolae-mediated endocytosis is a clathrin-independent endocy­ involved in immune regulation, inflammation, and cancer. In the drug
tosis pathway. It is present in different cells, mainly endothelial cells. delivery context, the phagocytic mechanism was extensively studied for
Over this internalization process, a dimeric protein called caveolin exogenous M&NPs as it regulates their biological fate and toxicological
triggers the invagination of the cell membrane in the cytosolic side and behaviors [70]. One crucial step before particle phagocytosis is their
forms small flask-shaped caveolae (50 − 60 nm) (Fig.s 1 and 2). Then, opsonization. This process consists of attaching proteins, called opso­
the cargo containing caveolar vesicle is detached from the cell mem­ nins, on the surface of the particles to be engulfed. Opsonized particles
brane and released into the cytosol after a GTPase dynamin ensures the are thus tagged and will be recognized by the phagocyte membrane
fission from the cell membrane. Although this endocytic pathway is receptors. To date, several opsonins have been reported, including
described to be relatively slow (half-time >20 min), the internalized complement components [71], immunoglobulins (e.g., IgM, IgG), type
materials could bypass the endosomal and/or the lysosomal compart­ one collagen, and C-reactive protein [72]. Conventional receptors for
ments [55]. opsonized particles include the Fc region of immunoglobulins, comple­
Macropinocytosis, also called big drink, is clathrin- and caveolin- ment receptors, and scavenger receptors [73]. Others investigated the
independent endocytosis pathway. It occurs in most mammalian cells, effect of other receptors such as CD44 in the opsonization process [74].
including professional phagocytes [56] and cancer cells under stimuli Importantly, unopsonized particles can also be recognized by phago­
such as growth hormones [57]. Macropinocytosis is mostly initiated by cytes, either as a consequence of the presence of specific patterns in their
actin-driven membrane ruffles (Fig.s 1 and 2). These laters often fuse surface such as carbohydrates residues (Mannose, I-fructose, glucose) or
with the cell membrane to form large vesicles (macropinosomes), whose via the CD11/CD18 integrins receptors that can bind to both opsonized
size varies from 0.2 μm to more than 5 μm (Fig. 1). For review article, see and unopsonized particles [75].
reference [58]. Once inside the cytoplasmic environment, unmatured Once introduced to the phagocyte’s membrane receptors, a signal­
macropinosomes shrink through a series of tubulations and fissions [59]. ling cascade will trigger the assembly of actin [76]. This event provides a
Then, they sequentially fuse with endosomes and lysosomes. force necessary for forming cell membrane extensions that engulf the

Fig. 2. Transmission electron microscopy images of micropinocytosis (1), clathrin-mediated endocytosis (2) (Reproduced from reference [60] with permission), and
caveolin-mediated endocytosis (3) (Reproduced from reference [61] with permission). In (1) and (2), the red arrows indicate filamentous influenza virion undergoing
clathrin-mediated endocytosis. The blue arrowhead in (2) indicates the clathrin lattice. Scale bars in (1) and (2) correspond to 100 nm. The image in (3) represents
the entry of Simian Hemorrhagic Fever Virus (green arrow) into MA-104 cells through caveolin-mediated endocytosis at 5 min. Noteworthy, the image published in
reference [61] without scale bar, but the diameter of ranges from 40 to 50 nm.

95
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

foreign particle and create a phagosome [77], (Fig. 3.A) which will carry 2.2.2. Nonphagocytic pathways of nonspherical particles
the uptaken particles in the cytoplasm. After actin depolymerization, the
phagosome transforms into an intracellular vacuole membrane. This 2.2.2.1. Improved cell uptake of nonspherical particles. Several studies
later fuses with the endosomes and eventually lysosomes forming suggested that nonspherical M&NPs were taken up by cells at faster rates
phagolysosomes [78], which acidic environment is responsible for the and in more significant amounts than spheres. Comparing the cellular
engulfed particle’s degradation [79]. uptake of mesoporous silica M&NPs with different ARs showed that
Although M&NP uptake by macrophages is usually undesirable, it particles having higher ARs were taken up by cells (e.g., HeLa cells,
was also exploited to target reticuloendothelial system for the delivery human melanoma cells, and adenocarcinomic human alveolar basal
of drugs (e.g., cytotoxic, antiparasitic, antifungal, antibacterial, anti­ epithelial cells) in large amounts and had faster internalization rates
viral, and anti-inflammatory drugs). In the drug delivery context, the [66,67]. The particles had similar equivalent diameter, surface compo­
impact of M&NPs size and surface chemistry on phagocytosis was sition, and surface charge but different AR: nanospheres (AR of ~1,
investigated by several research groups. Studies have suggested that diameter ~100 nm), short rods (AR of ~2 and ~240 nm in length), and
macrophage phagocytes can engulf a wide range of particle sizes from long rods (AR of ~4 and 450 nm in length) [66]. The number of long
250 nm to several micrometers [53,80]. It is also known that macro­ rods internalized by cells was almost twice that of short rods.
phage phagocytes can ‘eat bigger than their head’. [81] The engulfment Similar results were reported earlier by Gratton and coworkers [18].
of large particles is essential for host defense against infection and In this investigation, hydrogel particles with variable size, shape, and
removing dead cells. surface charge were designed by a top-down lithographic fabrication
Along with the size, the shape of M&NPs has emerged as a critical method called PRINTTM (Particle Replication In Nonwetting Templates)
attribute of their internalization by macrophages. The Mitragotri group [83,84]. While all the particles were internalized by HeLa cells, high AR
revealed that particle shape plays a crucial role in macrophage phago­ cylindrical particles (150 ×450 nm) were internalized faster than cubic
cytosis [20,21,35,36]. Polystyrene particles were prepared by the film particles and short cylinders. Those cylindrical particles had an AR of 3
stretching method. They had variables sizes and shapes: spheres (radius and were taken up by cells ~4 times faster than spheres. The particles
varied from 1 to 12.5 μm), oblate ellipsoids (1 × 4 μm, aspect ratio (AR) were positively charged and had a similar volume.
4), prolate ellipsoids (major axis 2 − 6 μm, AR 1.3 − 3), elliptical disks In agreement with those findings, other investigations showed that
(major axis 4 − 8 μm, AR 1.5 − 4.5) and UFO-shaped particles (sphere rod-like polymeric micelles were taken up by Caco-2 cells 12 times more
radius 1.5 μm, ring radius 4 μm) [21]. Particle internalization revealed efficiently than spherical particles with similar compositions [85]. The
that the phagocytosis kinetic was controlled by the local geometry of the exact mechanism for faster and higher internalization rates of
particle-cell interface. In this regard, ellipsoidal disks (3 × 14 μm) were nonspherical M&NPs is not well elucidated. However, the increased
internalized within a few minutes (<6 min) when the tangent angle (Ω) surface area of a long rod and disk particles with the cell membrane is
of the particle surface with macrophages was smaller than 45◦ (Fig. 3.A). believed to play a relevant role in cell internalization [18,19,37–39].
In contrast, for Ω > 45◦ (Fig. 3.B), the macrophages, attached to the flat In addition to AR, the sharpness of the edges was found to be a
side of the particle, were able to spread on the particle surface but did relevant parameter for cell internalization. For instance, it was found
not internalize it even after 12 h [21]. Importantly, those results were that microparticles that had sharp edges were faster internalized by cells
related to the differences in the actin structure formed when the particle than round-shaped particles [40]. While the exact mechanism was not
interacted with the macrophage surface. Indeed, when the particle was yet studied; the authors explained those differences by the possible
attached to the cell with its flat side, actin polymerized at points of difficulty to recruit enough actin filaments for round-shaped particles
contact but failed to create the adequate actin structure necessary to compared to particles with sharp edges [40]. Similarly, Hu et al. [41],
engulf the particle (Fig. 3.B). revealed that particles with irregular spiked surfaces exhibited faster
Those findings could be interesting for different applications, cellular uptake than spheres and smooth disks. The shape-controlled
including the possibility of acting on particle shape to inhibit particle nanostructures were formed by the hierarchical self-assembly
phagocytosis by macrophages [82]. It was demonstrated that worm-like approach of amphiphilic block copolymers.
particles with high AR (>20) exhibited negligible phagocytosis
compared to spherical particles with equivalent volume [82]. Video­ 2.2.2.2. Decreased cell uptake of nonspherical particles. Several studies
microscopy and fluorescence microscopy techniques revealed that suggested that cellular internalization of nonspherical M&NPs through
worm-like microparticles interacted with macrophages at the two- nonphagocytic pathways was reduced compared to spherical particles.
particle endpoints (Fig. 3.C). In particular, Au nanospheres were taken up by cells to a larger extent

Fig. 3. Schematic representation of the different orientations of a particle during phagocytosis. The tangent angle Ω between N and T was <90◦ in (A) and >90◦ in
(B). Dash lines in (C) represent the interaction of high AR particles at the two-particle endpoints. Adapted with permission from references.21, 82

96
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

than chemically similar nanorods [65]. In this experiment, uptake effi­ Shape-directed subcellular compartmentalization was already
ciency was inversely proportional to AR. The cellular internalization of demonstrated with iron oxide nanoparticles [44]. The particles had
Au nonspherical particles tends to be reduced compared to spherical similar composition and surface charge but differed by their shape
particles [65,86]. However, it is worth noting that the surface properties (spheres, spindles, biconcave particles, and nanotubes). The intracel­
of spherical and rod-like M&NPs are different. Indeed, the surface of the lular distribution revealed that the biconcave particles were preferably
Au spherical particles is coated with citric acid, while the nanorod localized in the nucleus, whereas spheres were distributed in the cyto­
surface was stabilized with cetyl trimethylammonium bromide (CTAB). plasm and nucleus. The spindle and nanotube geometries were observed
The differences in surface composition of rods and spheres may partially mainly in the cytoplasm. In another investigation, Xu et al. [45],
explain the contradictory results obtained with other particles. revealed that layered double hydroxide nanorods (30 − 60 nm in width
Nevertheless, similar trends were also reported for polystyrene disk and 100 − 200 nm in length) and nanosheets (10 − 20 nm in thickness
particles [19,38]. Indeed, while spherical particles enter the cells and and 50 − 150 nm in lateral wide) were internalized through clathrin-
perturb the cellular function, the disks bind only to the cell membrane mediated endocytosis, and they rapidly escape the endosome. The au­
with a significantly reduced cellular function such as reactive oxygen thors revealed that the intracellular distribution is shape-dependent
species (ROS) generation, apoptosis, and cell cycle progression [38]. The since the nanorods quickly translocated into the nucleus while nano­
nanospheres and the disks had similar diameters (~20 nm) and surface sheets were retained in the cytoplasm. The authors hypothesized that
potentials (ζ ≈ − 24 mV in serum). As the disks had a 2 nm thickness, the width of layered double hydroxide nanorods reduced from 30 − 60
the volume was not constant. Furthermore, the nanoparticles used in nm to 20 − 30 nm after deacidification within the endosome. Indeed,
this investigation are about 10 to 50 times smaller than those used in layered double hydroxides are a family of anionic materials that can
further experiments revealing preferential cell uptake for nonspherical buffer endosome acidification. Consequently, protons were pumped
particles. Lower cell uptake for nonspherical M&NPs agrees with theo­ from the cytosol, leading to particles releasing into the cytosol.
retical models based on free energy minimization. In a model developed Considering that the cylindrical nuclear pores are 20 − 30 nm in width,
by Decuzzi and Ferrari [64], it was predicted that nanoparticles with it is likely that the nanorods enter the nucleus after the endosomal
extremely low or high AR were not endocytosed. The AR should be in an escape. The size-fit of short single-walled carbon nanotubes (SWNTs) to
intermediate range for complete wrapping by the cell membrane. the nuclear pores partly explained their rapid translocation into the
nucleus. SWNTs, characterized by a diameter of ~1 nm and a length of
2.3. Subcellular distribution ~300 − 1000 nm, could cross the lipidic cell membrane through an
energy-dependent nonendocytic pathway (Fig. 1.4) and mainly distrib­
We have detailed how the M&NPs interact with cells and how they utes inside the nucleus [95]. While others suggested that energy-
are internalized depending on their shape. While the correlation be­ dependent endocytosis pathways may occur for SWNTs. For example,
tween the M&NP shape, their interactions with cells, and the internal­ SWNTs (~50 − 200 nm in length) were internalized by cells primarily
ization mechanism has been thoroughly investigated, only a few studies by a clathrin-coated pit pathway [96]. While there is no doubt on the cell
addressed their distribution inside the cells. Beyond the internalization internalization and accumulation of SWNTs in the nucleus, a central
mechanism, understanding the intracellular distribution of M&NPs is question is whether those particles are excreted from the cells by
essential for the subcellular drug delivery to the target compartment, exocytosis. Jin et al. [97], studied SWNT exocytosis by using single-
cell imaging, and understanding cytotoxicity behaviors. The intracel­ particle tracking. This investigation revealed that the exocytosis rate
lular distribution of a drug or its carrier is mediated by the internali­ closely matched the endocytosis rate.
zation route (Fig. 1). After their cell uptake, most M&NPs are confined in From another perspective, several studies revealed that the shape
early endosomes, which would mature into late endosomes and then and AR of M&NPs dictated the cell internalization mechanisms and
lysosomes before exocytosis [87]. Several strategies have been envi­ significantly impacted intracellular trafficking (Table 1). Macro­
sioned for endosome targeting through receptor-mediated internaliza­ pinocytosis and clathrin-mediated endocytosis were the main mecha­
tion. Ligands such as folate, transferrin, and low-density lipoproteins nisms for cell entry of ellipsoidal polymeric particles [98]. The ellipsoids
have been exploited for their ability to bind receptors that are overex­ were formed by the nanoprecipitation of poly(glycerol sebacate)-co-
pressed in malignant tumors for therapy or imaging [88,89]. The pH of polyethylene glycol (PGS-co-PEG) copolymers. The dimensions of the
early endosome (~6 − 6.5) was exploited to design pH-sensitive de­ ellipsoids were changed by acting on the molar ratio of PEG within the
livery systems to treat severe diseases such as cancer or Alzheimer’s copolymer: AR ~4.9 (235 × 48 nm) or AR ~2.4 (195 × 80 nm).
disease [90,91]. However, the internalization of the early endosome In other investigations, the uptake of rod-shaped mesoporous parti­
could represent a hurdle in developing M&NPs for targeting other sub­ cles with an AR of 2.1 − 2.5 (60 − 90 × 160 − 190 nm) was internalized
cellular compartments such as the cytosol or the nucleus. Most of the through a macropinocytosis process [67]. The particles with an inter­
drug-loaded particles that are endocytosed accumulate in late endo­ mediate AR were endocytosed faster and in a higher amount than par­
somes and lysosomes, where they are degraded due to harsh acidic pH or ticles with low AR (1~1.2) and high AR (4.0~4.5). Interestingly the rod
removed from the cell before reaching the target subcellular compart­ particles with an intermediate AR accumulated preferentially in the
ment (Fig. 1) [92]. For example, mRNA or siRNA carriers must escape perinuclear region, while shorter and longer rods were more randomly
the endosome to be released into the cytosol, where they interact with distributed in the cell. Gratton et al. [18], revealed that hydrogel
the RNA machinery. Ionizable lipids were used in designing lipid-based PRINTTM nanoparticles with an AR of 3 (150 × 450 nm) were rapidly
nanoparticles to release the drugs in the cytosol and avoid endosomal internalized into cells and translocated close to the nuclear membrane.
degradation. These lipids are neutral at physiological pH but protonated Those particles were internalized by cells through multiple modes of
in endosomes. In this way, lipid ionization facilitates the fusion of their non-specific endocytosis, most notably by a clathrin-mediated mecha­
lipids with the endosomal membrane and enables cytosolic RNA de­ nism, caveolae-mediated endocytosis, and to a lesser extent,
livery (For reviews on mRNA delivery, see references 93, 94). In another macropinocytosis.
embodiment, it was demonstrated that particles with sharp corners and However, other findings revealed that nanoparticle shape and AR did
edges offered a rapid endosomal escape to the cytosol by breaking the not significantly affect intracellular trafficking. The impact of Au
endosome’s membrane [42,43]. Soft-shaped particles failed to escape nanoparticle shape on intracellular trafficking was investigated by
the endosome, regardless of their surface charge, size, and composition. varying the ARs from 1 to 7 [99]. Both spherical and rod particles were
The AR of the particles used in this study was ~1, the average size was functionalized with DNA oligonucleotides to target class A scavenger
~115 nm, and most of the particles had at least 1 or 2 sharp corners receptors. The particles entered the cells through a caveolae-mediated
[42]. pathway regardless of the AR. Targeted nanorods reached the late

97
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

Table 1
Summary of the effect of the M&NP shape on the cell internalization pathway.
Particle type Shape Endocytic pathway Methods for characterizing Dimensions ζ (mV) Cell line Ref
particle cell uptake (nm)

Hydroxyapatite crystals Plates Macropinocytosis CLSM 221 × 115 -8.45 A7R5a [102]
Spheres Clathrin-mediated endocytosis 63 × 63 -2.57
Needles Multiple pathways 227 × 44 -4.58
Rods Mainly macropinocytosis followed by 129 × 41 -7.02
caveola-mediated endocytosis
Mannose-functionalized Spheres Clathrin- and caveola-mediated CLSM, flow cytometry 46 × 46 -10.9 RAW264.7b [103]
micelles# endocytosis
Short Clathrin-mediated endocytosis 99 × 50 -13.9
cylinders
Long Clathrin-mediated endocytosis 215 × 47 -13.3
cylinders
Silica M&NPs spheres Clathrin-mediated endocytosis CLSM, TEM 178 × 178 87 A549c [68,104]
worms Macropinocytosis 232 × 1348 58 &
RAW 264.7b
Au nanoparticles Stars Clathrin-mediated endocytosis ICP-AES, TEM ~68.55 -2.47 RAW 264.7b [105]
Rods Clathrin- and caveolae-mediated ~70.49 40.53
endocytosis
Triangles Clathrin Dynamin dependent pathway ~61.33 36.07
Au nanoparticles Stars Clathrin and Caveolae ICP-AES, TEM, dark field 15 × 15 -13.87 SMMC [106]
Rods Clathrin microscopy 33 × 10 54.27 7721d
Sphere Clathrin 15 × 15 -25.8 &
(small) GES-1e
Sphere Clathrin 45 × 45 -25.4 &
(medium) 4T1f
Sphere Macropinocytosis 80 × 80 -8.83
(large)
Silica particles rod Macropinocytosis Flow cytometry, 60-90 × 160- 13.0 HeLa [67]
CLSM, TEM 190

a/ aortic smooth muscle cell line, b/ macrophages, c/ human lung tumor epithelial cells, d/ human hepatocarcinoma cell line, e/ human gastric epithelial cell line, f/
mouse breast cancer cell line. #two different polymers were used for the nanoparticle design: poly(DL-lactic acid)-b-poly-(acrylic acid) for spherical micelles and poly
(L-lactic acid)-b-poly(acrylic acid) for cylindrical micelles. CLSM: Confocal laser scanning microscopy. TEM: Transmission electron microscopy. AFM: Atomic force
microscopy. IPC-AES: Inductively coupled plasma-atomic emission spectrometer.

endosome within 4 to 8 h post-incubation with cells, but they did not delivery of nucleic acids, which is challenging because of their negative
progress from the late endosome to the lysosome. In agreement with charge and degradation in physiological conditions. For instance, SWNT/
these findings, the lysosomal compartmentalization was independent of DNA complexes resulted in an efficient gene delivery capacity in cells [108].
the shape of soft hydrogel capsules [100]. In this investigation, the AR SWNTs had a diameter of 20 nm and a length of 200 nm. In another
varied from 1 to 3.8. Shi et al. [101], reported a higher accumulation of investigation, small interfering RNA (si-RNA)-encapsulated needle-shaped
rod-like particles with high AR (~18 × 102 nm) in the endosome and the PLGA particles (80 × 320 nm) were generated by PRINTTM process
lysosome than low AR rods (~25 × 74 nm). Intracellular trafficking [109]. Particle surface was coated with lipids DOTAP:DOPE (1 : 1 wt%) to
experiments were performed on polyplexes of N-(2-hydroxypropyl) increase the transfection efficiency (DOTAP: 1,2-dioleoyl-3-trimethylammo­
methacrylamide-oligolysine brush polymers. When both polyplexes had niumpropane, DOPE: dioleoylphosphatidylethanolamine). The particles
the same molecular weight and chemical composition, they differed in were efficiently internalized by different cell lines, and the knockdown of
the oligolysine length, which confers to them a different shape. therapeutically relevant genes was demonstrated [109]. In a more recent
Furthermore, the preferential accumulation of high AR rod-like particles investigation,[110] the si-RNA loading, release and internalization effi­
in the endosomal/lysosomal compartments delayed nuclear delivery. ciencies, and effectiveness of post-transcriptional gene silencing of hollow
Those findings were attributed to the cationic polyplexes that could Au particles (45 − 50 nm size) with different shapes and dimensions were
interact through multivalent interactions with the endosomal and compared (nanoshells, nanocages, and nanorods). The three particles were
lysosomal membranes. Higher interactions are expected with a higher internalized by cells, but nanoshells and nanocages displayed the highest
surface area offered by rod-shaped polyplexes. knockdown efficiency.

2.4. Intracellular drug delivery 2.5. Cytotoxicity

A key question is how to exploit shape-dependent cell uptake for Among the different available approaches to evaluate the safety of
efficient drug delivery? In this regard, it was demonstrated that rod- M&NPs, in vitro experiments represent the first step before selecting the
shaped mesoporous silica particles with an AR of ~2.1 − 2.5 were most promising preparation to be tested ex vivo and in vivo. A panel of
more efficient in delivering hydrophobic chemotherapeutic agents (e.g., cell viability tests is available to evaluate cytotoxicity, as summarized in
camptothecin or paclitaxel) than spherical ones [67]. The cytotoxic Table 2. Compared with M&NPs size and surface properties, the effect of
potential of the encapsulated drugs was studied in HeLa cells. In a study particle shape on cytotoxicity was more recently studied. Efforts devoted
conducted by Enlow et al. [15], cylindrical poly(lactic acid-co-glycolic to understand the relationship between particle shape and cell toxicity
acid) (PLGA) particles (200 × 200 nm) prepared by PRINTTM process highlighted contradictory results (Table 3). For example, no significant
showed high and efficient loading of docetaxel. Particles containing up cytotoxicity was observed with cylindrical PEGylated PRINTTM micro­
to 40 wt% of the drug exhibited higher cytotoxicity than Taxotere. particles after 4 h or 72 h of incubation with four different cell lines,
Similarly, the cell uptake and the cytotoxicity of tubular polymersome- including HeLa and macrophages [16]. The microparticles had a mean
loaded doxorubicin were higher than for spherical particles [107]. height of 0.68 μm, a mean width of 1 μm, and positive or negative zeta
Nonspherical M&NPs showed promising results for the intracellular potentials. In another investigation, PEGylated nanoimprinted disks

98
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

Table 2
Commonly used methods and assays to determine M&NPs cytotoxicity.
Toxicity test Principle Data interpretation Ref

MTT assay MTT is a colorimetric assay, which determines the mitochondrial The purple color’s intensity (or optical density) is directly [122,123]
activity in living cells. It is based on the reduction of MTT (3-[4,5- proportional to the mitochondrial oxidoreductase activity. The
dimethylthiazol-2-yl]-2,5-diphenyl tetrazolium bromide) (yellow healthy cells exhibit high rates of MTT reduction to formazan
colored) by the action of mitochondrial oxidoreductase enzymes. This and high optical density, corresponding to lower cytotoxicity of
reduction results in insoluble purple formazan crystals that can be the M&NPs at the tested concentration.
quantified after solubilization in dimethyl sulfoxide and measuring
the optical density at 570 nm.
MTS assay MTS is a colorimetric assay based in the reduction of the MTS reagent Similar to MTT assay, the purple’s color intensity is correlated [124,125]
(3-[4,5-dimethylthiazol-2-yl]-5-(3-carboxymethoxyphenyl)-2-(4- to the cell viability. Lower optical density corresponds to higher
sulfophenyl)-2H-tetrazolium). In MTS assay, the formazan dye is cell death.
soluble in cell culture media, and the optical density is measured at
490-500 nm.
LDH assay (or LDH In the lactate dehydrogenase (LDH) assay, the level of plasma The cell death is proportional to the intensity of the red color, [123,126,127]
release assay) membrane damage in a cell population is assessed. LDH is an enzyme which is proportional to the reduction of tetrazolium salt by the
found in the cytoplasm of living cells. When the cell membrane is LDH leaked into the cell culture.
damaged, the LDH is released into the cell culture medium. The assay
is based on the reduction of tetrazolium salt into a red-colored
formazan dye. The optical density is measured at 492 nm.
ROS Reactive oxygen species are produced inside the cells as a response to Elevated ROS (e.g., elevated intensity in the fluorescence or [127,128]
assays environmental stress. There are several assays for the detection of spectrometry methods) is highly correlated with cell death by
ROS: necrosis or apoptosis.

- Fluorescence dependent methods


- Chemiluminescence dependent methods
- Spectrometry methods
- Chromatography methods
- Electrochemical biosensors
- Electron spin resonance
- Fluorescent protein-based methods.
WST-1® assay WST-1 (2-(4-iodophenyl)-3-(4-nitrophenyl)-5-(2,4-disulfophenyl)- Similar to MTT and MTS assays, the cell viability is correlated [129]
2H-tetrazolium) is a water-soluble reagent. It does not penetrate the to the optical density intensity.
cell but is instead cleaved by the succinate-tetrazolium reductase at
the cell surface to form a water-soluble formazan dye. The optical
density of formazan dye is measured at 450 nm.
WST-8® assay (CCK- WST-8® (CCK-8®) assay is based on the same principle as WST-1® or Similar to MTT, MTS, and WST-1® assays, the formazan dye [130]
8® assay) MTS assay but exhibits superior detection sensitivity by the concentration is proportional to the intensity of the optical
tetrazolium salts are highly soluble. The optical density of formazan density and the cell viability.
dye is measured at 450 nm.
Neutral red uptake Neutral red uptake assay is based on the detection of neutral red (NR) Neutral red uptake assay provides a quantitative estimation of [131]
assay dye inside the living cells. Viable cells can internalize the NR dye, the number of viable cells. The cell viability is positively
which is taken up by the lysosome. This later is stained red. On the correlated to the amount of NR dye released and thus the
other hand, non-viable cells are not able to internalize this dye. After optical density intensity.
the lysosome staining, cells are washed, and the NR dye extracted. The
optical density of NR dye is quantified at 540 nm.
Alamar blue® assay Alamar blue® (AB) is a water-soluble dye, which is not toxic to the The magnitude of AB reduction is correlated to the optical [127]
cells. AB enters the cytosol, where it is reduced by mitochondrial density intensity (or the fluorescence), which is proportional to
enzymes in living cells. This reduction is accompanied by a change in the number of viable cells.
the culture medium color from indigo blue to fluorescent pink. The
change in color is easily measured by colorimetric or fluorometric
reading.
PicoGreen® assay In the PicoGreen® fluorescence enhancement assay, the fluorochrome Fluorescence intensity amplitude is a consequence of a high [132]
(PicoGreen) selectively binds double-stranded DNA (dsDNA) in living amount of DNA in the living cell.
cells, leading to a fluorescence enhancement. The samples are excited
at 480 nm, and their emission is set at 520-530 nm.
Annexin V-affinity Annexin V staining assay allows sensitive and non-destructive analysis In flow cytometry, living cells emit in the window of the [133]
assay of apoptosis. Annexin V has a high affinity for phosphatidylserine. In fluorescent conjugate of Annexin V (e.g., dark green for
healthy cells, phosphatidylserine is found in the inner leaflet of the Annexin V-FITC). In contrast, the apoptotic cells emit in a
lipid bilayer of the plasma membrane. During an apoptotic episode, different window (e.g., blue).
the phosphatidylserine becomes exposed to the cell surface and is
subsequently detected by the Annexin V labeled kit. Reading can be
done either by flow cytometry or microscopy techniques.
Caspase 3 The activation of caspase 3, a mammalian cysteine protease, plays a Higher fluorescence emitted after caspase 3 assay corresponds [134]
assay significant role in the apoptotic process. Caspase 3 is detected in the to a high apoptotic process.
cell lysate by means of a fluorogenic substrate (N-acetyl-Asp-Glu-Val-
Asp-7-amino-4-methylcoumarin). This substrate is cleaved by caspase
3, generating a fluorescent component (7-amino-4-methylcoumarin)
that can be detected in a fluorescent reader (λex = 380 nm) and λem =
420 nm).
Cell membrane In addition to the release of LDH and cytokines, it was reported that The high fluorescence toward a targeted surface molecule [135]
surface molecules cells may over-express some molecules on their surface after contact compared to a negative control is interpreted by an over-
assessment with M&NPs. expression of this surface molecule after the cell’s contact with
M&NPs.
- CD11c: corresponds to an enhancement of macrophage maturation.
(continued on next page)

99
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

Table 2 (continued )
Toxicity test Principle Data interpretation Ref

- MHCII, CD1d: corresponds to an enhancement of antigen


presentation.
- CD40 and CD86: which translate a co-stimulation activity of the
antigen-presenting cells.
A way to access these molecules is by targeting them with specific
staining after realizing an Fc-blocking.
RT-PCR Reverse transcript polymerase chain reaction (RT-PCR) is based on the When M&NPs are toxic to a cell, the apoptosis is translated by [136]
detection of bcl-2 and bax, which are specific gene primers associated high values of bax/bcl-2 ratio.
with apoptosis. The calculation of the bax/bcl-2 ratio determines the
susceptibility of a cell to undergo apoptosis. Accordingly, high values
of this ratio are highly correlated with an apoptotic episode.
CLSM Confocal laser scanning microscopy is often used for analyzing the cell CLSM allows the analysis of a cell apoptosis episode, the [137]
damage induced by the M&NPs. Appropriate staining of M&NPs and presence of white vesicles, and particle penetration inside the
cells offers the possibility to visualize the internalization events and/ cell.
or subcellular particle distribution. This technique was also used for
visualizing cell shrinking and asymmetry during apoptosis.
Electron microscopy Electron microscopy techniques are mainly used for the non- Owing to their high resolution, electron microscopy techniques, [138–140]
techniques quantitative determination of cell apoptosis. They lead to the when correctly set up, allow visualizing apoptotic events after
morphological characterization of dead cells and the determination of cells treatment with M&NPs.
hallows regions in the dell membrane. Although very useful, these
techniques are not often used as a first intention because, compared to
other techniques, they are expensive, and they mainly allow a
qualitative analysis.

(325 nm × 100 nm), and rod-shaped particles (400 nm × 100 nm × 100 chronic granulomatous inflammation due to the release of reactive ox­
nm) did not induce cell death on three different cell lines (HeLa, HEK ygen species and hydrolytic enzymes [118,119]. Interestingly, spheri­
293, and HUVEC). A ratio of 105 particles per cell was used [39]. As cally shaped carbon nanotubes induced lower cytotoxicity than high AR
expected from the biocompatible and biodegradable PLGA, M&NPs multi-walled carbon nanotubes [120]. In a recent study, Nahle et al.
were not toxic on cells regardless of their shape [15]. In this investiga­ [121], studied the effect of the morphology of carbon nanotubes on their
tion, cylinders (200 nm × 200 nm, 320 nm × 80 nm, or 600 nm × 200 cytotoxicity on rat alveolar macrophage cell line. Their results showed
nm), 2 μm cubes, and 3 μm particles with center fenestrations were higher toxicity of multi-walled carbon nanotubes than single-walled
tested on ovarian carcinoma cells. In agreement with those findings, carbon nanotubes[108]. The nanotubes were ammonium functional­
ellipsoidal particles composed of PGS-co-PEG copolymers with two ized and had a diameter of about 20 nm and an apparent length of about
different AR (AR ~4.9, 235 nm × 48 nm and AR ~2.4, 195 nm × 80 nm) 200 nm. In addition to the shape and the length, other characteristics
did not show any toxicity toward two cell lines preosteoblasts [98]. could contribute to adverse biological outcomes, including surface
Similarly, changing the shape from spherical to rods, cylinders, disks, or functionalization. Indeed, significant differences in cytotoxic behaviors
worms of silica, Au, or polystyrene nanoparticles did not affect cell were observed with multi-walled carbon nanotubes (10 − 30 μm in
viability [38,68,86,104,111,112]. Collectively, it was found that the length and 20 − 30 nm in diameter), which induced less profibrogenic
cytotoxicity was rather related to particle concentration, surface charge effects when functionalized with carboxylic groups [115].
[109], particle size, and the cell type.
In opposition to these results, other studies showed that AR had a 3. Theoretical approaches for particle endocytosis
significant impact on cell death. The cytotoxicity of high AR silica
nanorods (AR ~4, 450 nm in length) was higher than low AR nanorods 3.1. Mechanical forces and energies over particle endocytosis
(AR ~2, 240 nm in length) [66]. It was found that DNA-coated Au
nanorods that had high AR (~4 or 7) induced higher cytotoxicity than Forces and energies occuring during the interaction of a particle with
low AR nanorods [99]. Similarly, nanorods and star-shaped particles a cell membrane are essential for cellular function and various biological
had higher toxicity compared to spherical particles [106]. In another processes. The creation of a vesicle wrapping the particle requires a
investigation, both diameter and AR of cerium oxide nanorods were mechanical force to deform the cell membrane and deviate from its
varied to determine the critical AR that induced cell death [113]. The AR spontaneous curvature denoted κ0. Spontaneous curvature arises from
varied from 1 to >100 (from 33 to >1000 nm in length and 7 to 9.5 nm combining several mechanisms, including membrane lipid composition
in diameter). While no significant cell death was reported with low AR and asymmetry, partitioning of transmembrane domains and protein
nanorods (up to 31), the highest AR induced significantly higher cell crowding, insertion of hydrophobic protein motifs, the assembly of hy­
death and activated the NALP3 inflammasome. Subsequent activation of drophilic protein domains (e.g., clathrin, dynamin, caveolin), and
the multiprotein complex NALP3 inflammasome activates caspase 1, macroscopic scaffolding by the cytoskeleton. For a review see, reference
leading to pro-inflammatory cytokines including interleukin-1β (IL-1β) [143].
and IL-18. For review, see reference [114]. Similar conclusions were Contacting a particle with the cell membrane is an essential step for
drawn for high AR multi-walled carbon nanotubes, which activated the beginning of the endocytic cascade, which results in particle wrap­
NALP3 inflammasome and led to cytokine release in a human mono­ ping, either through specific or non-specific mechanisms. When a par­
myelocytic leukemia cell line [115]. Some researchers showed that the ticle docks on the cell membrane, various new tensions are created
cytotoxicity of multi-walled carbon nanotubes could be decreased by which, are drained by electrostatic forces, van der Waals forces, hy­
acting on their length [116,117]. Indeed, short carbon nanotubes (600 drophobic forces, and receptor-ligand interactions [144]. All these
nm in length) induced less ROS and inflammatory cytokines than long forces cumulate to form an adhesion strength (w) that creates the
nanotubes (20 μm in length). Several studies showed that long (>15 μm necessary imbalance to the resistance forces and thus starts the wrap­
in length) and more rigid multi-walled carbon tubes were not wrapped ping process.
by macrophages. Frustrated phagocytosis in macrophages induced Let us consider a simple example representing the interaction of a

100
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

Table 3
Cytotoxicity of nonspherical M&NPs prepared from different materials.
M&NPs Particle Particle ζ Incubation C (μg/ Cell line Cytotoxicity Observation Ref
shape size (nm) (mV) time mL) assay

Hydroxyapatite Rod 96 -5.7 24 h 10 − RAW 264.7a PicoGreen assay PicoGreen assay revealed that the necrosis of [141]
Plates 221 × 115 -9.7 300 and BEAS- BEAS-2B cells was tributary to the particle
2Bb shape, which is in the order needles > plates
> spheres > rods.
Based on this assay, RAW264.7 was reported
to be less sensitive to this test than BEAS-2B.
Note that plate-shaped particles showed up to
15% reduction in dsDNA at 10 and 30 μg/mL.
Overall, needles and plates were found to be
more toxic according to this test.
Spheres 92.3 -14.1 ROS assay DCFDA assay (2’,7’-
(DCFDA) dichlorodihydrofluorescein) is a fluorescent-
dependent assay that allows the detection of
ROS. The assay showed that, in BEAS-2B cells,
regardless of the nanoparticle shape, ROS
levels were at the same level as controls
without nanoparticles. However, in
RAW264.7 cells, needle-shaped particles
generated increased ROS levels than the other
shapes.
Needles 74.9 -4.8 Cytokine assay Cytokine test revealed high expression of IL-6
in needle-shaped and plates particles. The IL-
6 expression with the other nanoparticle
shapes was comparable to the negative
control. The same study did not reveal any
difference in the TNF-α expression regardless
of the particle shape.
Silica Long rods 278.1 -30 24 h 50 − Caco2c MTT assay No significant changes in cell viability were [111]
(AR~4) 1000 reported with the cells after incubation with
Short 289.0 the nanoparticles, compared to control cells
rods (AR~2) that were not exposed to nanoparticles. The
Spheres 250.9 authors explained the results by degrading
the nanoparticles into nontoxic silicate ions,
regardless of their shape.
Alumina Rods 20 − 30 in N/A 72 h 15.6 − Astrocytesd MTT assay According to MTT results, nanorods exhibited [142]
width and 8000 higher toxicity for astrocytes than nanoflakes
100 − 200 (LC50 value 3.6-fold higher for nanoflakes
in length than that for nanorods).
Flakes 20 − 30 in N/A ROS assay Both particles (nanorods and nanoflakes)
both width induced ROS production. However, nanorods
and length induced higher ROS generation than
nanoflakes.
Cytokine assay The increase in IL-1β, IL-2, and IL-6
concentration was 1.5 to 2-fold higher after
exposure of the cells to nanorods compared to
nanoflakes.
Silica Worms 232 × 87 72 h [1] 0.0001 RAW 264.7 WST-8 assay, It was found that, at low concentrations [68]
1348 − 1000 and A549e caspase 3, (<100 μg/mL), nanoparticle geometry played
Cylinders 214 × 428 79 24 h [2] annexin V, and little to no role in mitochondrial function,
Spheres 178 58 LDH assay membrane integrity, or cell death.
Au Rods 10 × 33 54.2 24 h 0− SMMC 721f, Cell counting kit CCK-8 assay revealed high cell viability (up to [106]
Stars 15 -13.8 8000 GES-1g, and (CCK-8) assay 80%) after the exposure with the
Spheres 15, 45 or -8.8 4T1h nanoparticles, regardless of their shape at a
80 to high concentration of 3000 μg/mL. However,
-25.4 lower toxicity was noticed for spherical
particles compared to rods and stars.
mPEG coated Rods ~90.81 1.14 24 h 2.5 − 40 RAW 264.7a CCK-8 assay The three shapes of PEGylated Au [105]
Au Stars ~82.88 -3.87 nanoparticles were nontoxic for RAW 264.7
Triangles ~84.57 -1.88 cells regardless of particle concentration.

1/ incubation at 72 h for WST-1 assay. 2/ incubation at 24 h for LDH assay. a/ macrophage; Abelson murine leukaemia virus-transformed cell line. b/Primary and
immortalized human bronchial epithelial cells. c/ Human colon tumor epithelial cells. d/ star-shaped glial cells in the brain and spinal cord. e/ human lung tumor
epithelial cells. f/ human hepatocarcinoma cell line. g/ human gastric epithelial cell line. h/ mouse breast cancer cell line.

spherical particle with the cell membrane in Fig. 4. Upon the interaction energy Γ(η) (Fig. 4). The wrapping extent (or the degree of wrapping) η
of a particle with the cell membrane surface, adhesion energy (ΔE) is (with 0 < η < 1) is calculated from the ratio of the wrapped particle area
generated at the particle-membrane interface. While unspecific ΔE is to the total particle area. Additional deformation energy Λ(η) stored in
rapidly set up after particle interaction with the cell membrane, specific the curved membrane as represented in Fig. 4 [145–147]. Typically, for
ΔE requires a longer time as it introduces a translational entropy. Once a fully wrapped particle, η = 1, the cell membrane comes back to its
the particles are robustly attached to the cell membrane, the wrapping original curvature and Λ = 0. These two energies and the additional
process begins, and the cell membrane deflects from its original curva­ deformation energy are necessary for an effective wrapping of the par­
ture κ0, relying primarily on a bending energy C(η) and a stretching ticle (Fig. 4). The total membrane deformation energy (DE) required for

101
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

an effective particle wrapping is the sum of C(η), Γ(η), and Λ(η), as Before the interaction, it was assumed that the receptors were uni­
illustrated in Fig. 4 and Eq. 1. formly distributed on the flat cell membrane with density ξ0, and that
their total number remains constant over wrapping (Fig. 5). Upon the
DE = C(η) + Γ(η) + Λ(η) (1)
contact of the nanoparticle with the cell membrane, the mobile receptors
Many theoretical studies were dedicated to understand how the cell diffused toward the contact area. The receptor density interacting with
membrane responds to external forces induced by the interaction of a the ligands on the nanoparticle surface (ξb) became higher in the contact
particle [32,33,145–152]. Although various parameters are recurrently area. In contrast, a depletion zone with reduced receptor density (ξ+) is
linked with a wrapping process; particle size undeniably plays a signif­ formed in the noninteracting region of the cell membrane (Fig. 5). ξb
icant role. Based on energetic and kinetic considerations, it was possible reaches a maximum value when each ligand interacts with one receptor
to predict the minimum radius (Rmin) upon which the endocytosis of a (ξb = ξL). In the model developed by Gao et al. [154], the nanoparticle
particle cannot be effective. If we consider a simple case in which a was considered either cylindrical or spherical. Notably, if we assume
spherical particle interacts with a cell membrane through non-specific that the cell size was much larger than nanoparticle size, the initial state
adhesion, the Rmin is expressed by Eq.2 [147]. of the cell membrane was presented as flat. Simulation results revealed
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ that an optimal receptor density ratio (ξ = ξ0 /ξL ) is required for effec­
2B
Rmin = (2) tive wrapping. As the ξ increased, the optimal wrapping radius and
αns − σ
optimal wrapping time decreased. In contrast, for small ξ, the optimal
where B is the membrane bending modulus, αns represents the particle radius and wrapping time increased [154].
adhesion strength, and σ the membrane tension. For biological mem­
branes, membrane bending modulus B, at a temperature of 300 K, is
commonly expressed by Eq.3 [153]. 3.2. Shape effect

B ∼ 15kB T (3)
While the fundamental biophysical behaviors of cell membrane
2 deformation over a wrapping episode remain applicable for nonspher­
Thus, when αns is equal to 1 kBT/nm , the minimum radius below
which a wrapping process cannot occur is Rmin = 5 nm [147]. In this ical particles, they exhibited some differences in the wrapping mecha­
situation, other endocytic pathways for cell entry, such as trans­ nism compared with spherical ones. Understanding the energetics of a
membrane diffusion, are privileged (Fig. 1.4). nonspherical particle is more complicated due to the various orienta­
In the case of specific interactions between ligand-coated spherical or tions possible upon the interaction with the cell membrane. Thanks to
cylindrical particles with mobile receptors in the cell membrane, computational modeling at different length scales, the energetics for a
mathematical modeling revealed that particles which size ranged from nonspherical particle wrapping were better understood. Typically,
10 − 100 nm were endocytosed and exocytosed cells via wrapping coarse-grained models and numerical calculations of membrane me­
events without implicating clathrin or caveolin coats [154]. According chanics based on energy minimization revealed that the endocytosis
to the same authors, the optimal radius of spherical particles at which mechanism of nonspherical particle is defined as a sequence of four
the endocytic time is the shortest was 27 − 30 nm. Those predictions are different steps: [31–34] a nonwrapped state, a binding state, a partially
in good agreement with several experimental data, revealing that the wrapped state, and a complete wrapped state, where the particle is
optimal radius for a maximum cell uptake was ~25 − 30 nm completely engulfed by the cell membrane (Fig. 6). A partially wrapped
[86,155,156]. Thermodynamic models developed for specific in­ state could be further subdivided into two other steps: the shallow
teractions of ligand-coated nanoparticles and receptors in the cell wrapped state, which comes right after an effective particle binding to
membrane also predicted an optimal size for nanoparticle wrapping of the surface area, and a deep-wrapped state mainly characterized by a
~25 − 28 nm [150,157]. In the developed models, the ligands were deeper nanoparticle penetration and a higher surface tension of the cell
assumed to be uniformly distributed on the particle surface with a (Fig. 6).
density ξL and were immobile [154,157]. The influence of ligand dis­ Interestingly, several groups agree that nonspherical particles can
tribution on cellular uptake using coarse-grained molecular dynamins enter the cell through different modes (Fig. 6): [31,34,160–162] a
[158] or statistical dynamics model of endocytosis [159], revealed that submarine mode (or parallel adhering mode), when the particle is pre­
nanoparticles with homogenous ligand distribution are the most effec­ sented to the cell membrane via its long axis and a rocket mode (or
tively wrapped. perpendicular entry mode) when the long axis is perpendicular to the
cell membrane. Another case is also possible. It consists of the compe­
tition between both submarine and rocket modes, characterized by the
rotation of the particle over a wrapping episode (Fig. 6.a and 6.b)
[34,163].
Within the same framework, modeling of particle membrane inter­
action revealed that the wrapping of cube-like and rod-like particles
compared to the wrapping of spheres and ellipsoids is different [31].
Complete wrapping of cube-like particles required increased adhesion
energy in the lateral edges to back up the deformation energy of the
lower edge and increased deformation energy of the upper edge [31].
For cube-like particles, the adhesion strength to reach a deep wrapping
state is ~2 times higher than for a sphere with an equal surface area.
This was accompanied by increased membrane tension and adhesion
strength. For a complete wrapping of Hauser’s cube, the adhesion
strength was found to be ~3 times higher than for a sphere. Moreover,
very high adhesion strength is necessary for the complete wrapping of
cube-like particles. On the other hand, particle shape dictated its
internalization mode. For example, a nanorod with AR ~1.5 switched
Fig. 4. Schematic representation of membrane deformation energies over a from a rocket position to a submarine position, but for sharper edges,
wrapping episode. Adapted with permission from references [145–147]. R is rocket mode predominates [31].
the effective radius of the sphere. In another embodiment, molecular dynamic simulations on a coarse-

102
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

endocytosis process of ligand-coated rod particles was thermodynami­


cally favorable, while cylinders did not undergo complete wrapping, but
they remained attached to the membrane. Notably, in this article, rod
particles were referred to as spherocylinders (cylinders with hemi­
spherical caps at both ends). The cylinders had a diameter D, while the
whole rod’s length was L. The AR was ~1 or ~2, as calculated from the
ratio L/D. The results showed that, upon the interaction with the
membrane, the rods changed their orientation from perpendicular to
parallel (Fig. 6.a). The same investigation revealed that the endocytosis
of rods was more thermodynamically favorable than spheres [161].
However, opposite results were found with PEG-coated nano­
particles with variable shapes and using large-scale dissipative particle
dynamics simulations [164]. The simulations revealed that, at low PEG
density, spherical nanoparticles had the fastest internalization rate,
followed by cubes, then rods and disks [164]. The molecular simulations
showed that the entry pathway of nanoparticles with high AR (rods and
disks) was highly dependent on the contact angle with the membrane.
When the nanoparticles with high AR interacted with a cell membrane
through their major axis, the large membrane deformation and time-
consuming particle orientation delayed the internalization. The same
study showed that star-shaped nanoparticles exhibited fast wrapping by
the cell membrane, which is similar to spherical nanoparticles. It was
noticed that when the nanoparticles were fully covered with PEG, the
effective shapes of all the particles were close to spherical. In these
conditions, the shape of the nanoparticle core is invisible to the cell
membrane (see Fig. 9 in reference [164]).
Fig. 5. Schematic representation of the different steps for receptor-mediated
wrapping of a spherical nanoparticle. Step 1: Nonwrapped state. Step 2: bind­ 4. Shape-dependent behaviors in the systemic circulation
ing state. Step 3: partially wrapped state. ξ0 is the receptor density in the remote
region. ξ+ is the receptor density in the depletion region. ξb is the density of the Protein corona represents the first barrier a particle encounters when
receptor in the binding region. ξL is the ligand density on the nano­ administrated in the blood because thousands of different blood proteins
particle surface. will competitively bind to the particle surface [165,166]. This phe­
nomenon complicates any prediction of cell interactions, bio­
grained model revealed that the local curvature radius and the sharpness distribution, and toxicity because it is the protein corona/M&NP
of the edges of the nanoparticle in contact with the membrane played a complex, rather than the bare M&NPs interacting with biological ma­
significant role in the wrapping process [161]. Indeed, the passive chinery [167]. Furthermore, targeting properties of M&NPs are

Fig. 6. Illustration of the different entry modes of nonspherical nanoparticles according to numerical simulations.

103
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

hindered by protein adsorption on functional ligands, preventing their Similarly, by tuning the shape of polystyrene or Au particles from
interaction with the corresponding cell receptors [49]. After adsorption spheres to rods or elliptical disks, macrophage uptake was reduced,
of opsonins on the M&NP surface and phagocytosis, the particles will be circulation t1/2 in mice blood was increased, while liver uptake was
directed to MPS organs, such as the liver and spleen. Indeed, 30 − 99% decreased [62,182].
of administered M&NPs accumulate and sequester in the liver [168]. To Contrarily to spheres, the disks, targeted to intracellular adhesion
reduce non-specific adsorption of proteins on the M&NPs surface, a molecule (ICAM-1), preferentially accumulated in the lungs. The lung to
plethora of research works were focused on grafting hydrophilic poly­ liver accumulation was 1.7-fold higher for anti-ICAM-coated poly­
mers such as PEG on the particle surface [169]. This strategy reduced styrene rods (125 × 500 nm) than for spherical particles (200 nm
opsonization leading to the escape of immediate sequestration by the diameter) [189]. Shape-specific tissue accumulation was also observed
immune system after administration in the blood. Unfortunately, there is for anti-transferrin-conjugated rod particles for brain targeting.
a piece of evidence that repeated administration of PEG-coated nano­ Transferrin-coated rods exhibited ~7 fold higher accumulation in the
particles in the blood induces hypersensitivity reactions described as brain compared with transferrin-coated spherical particles [189]. In
complement activation-related pseudoallergy (CARPA) [170]. Several vitro experiments performed on a synthetic microvascular network
strategies have been envisioned so far to reduce the adsorption of op­ model showed that rod particles exhibited higher binding to the walls
sonins on the particle surface. Unlike size and surface properties (e.g., under shear flow than spherical particles. This phenomenon, known as
hydrophilic/hydrophobic ratio, steric hindrance), the effect of M&NP margination, was tackled a decade ago for nonspherical M&NPs
shape on opsonization was understudied. Some data were reported on [190–192]. In a theoretical study, Toy et al. [190], concluded that Au
the impact of the radius of curvature of M&NPs in the composition and nanorods had an 8-fold higher margination rate than nanospheres.
conformation of adsorbed proteins by using spherical particles with Another investigation revealed that both size and morphology had a
variable sizes. For example, previous works on spherical gold nano­ consequent impact on the margination and adhesion of M&NPs [193].
particles showed that big particle size, and consequently high radius of Their experimental analysis, conducted on spherical, oblate, prolate,
curvature, affects protein corona structure as proteins may stretch to fit and rod M&NPs, revealed a higher margination and adhesion of micro-
the large particle surface. Smaller dimensions of isotropic nanoparticles scale nonspherical particles than nano-scale spherical counterparts.
resulted in decreased interaction with proteins, causing fewer structural Additionally, Da Silva-Candal et al. [194], found that rod particles have
changes [171–174]. For anisotropic nanoparticles with high AR, one a better targeting efficiency to the endothelium receptors than spherical
expects large proteins aligning along the long axis to fit on the nano­ particles. Spherical particles are more likely to follow the shear forces’
particle surface [175,176]. Large proteins, on the other hand, may not fit direction and continue their way in the streamline. Once they
in small spaces and hence be excluded from the small dimension of the marginate, nonspherical particles display another characteristic which
nanoparticles. Additionally, anisotropic particle geometry does affect is their ability to set up a multi-point contact with the receptors in the
complement activation. The IgM and the classical and lectin pathway vessel wall due to their bigger surface area and lower number of cur­
convertase (C4b2a) have cross-sectional diameters of ~40 nm and ~30 vatures. Several experimental and theoretical studies confirmed the
nm, respectively, making its assembly and deposition onto the surface of ability of elongated M&NPs to marginate to the vascular wall of the
nanoparticles with high curvature rather tricky [177–179]. Therefore, vessel and subsequent extravasation under shear and pressure forces
particle dimensions lower than 30 nm, and depending on their surface [195]. For review articles, see references [196,197].
chemistry, may not efficiently trigger calcium-sensitive pathways of The tendency of rod-like or disk-like particles to align the blood flow
complement activation, while a surface-bound C3b is expected to and to extravasate partly explains higher accumulation in tumoral tis­
occupy an area of 40 nm [180]. sues (Table 4) [182,184,187,198–200]. The biodistribution of porous
Specific interactions of nonspherical M&NPs with blood proteins, silicon discoidal particles (400 × 600 nm and 400 × 1000 nm, height ×
with macrophages and polymorphonuclear neutrophils, as well as the diameter) in a murine bearing cancer model was found to be up to 5
dynamic behaviors of the particles under shear flow, significantly times higher than with spherical particles (600 nm diameter) [200].
impact their biodistribution (Table 4). The most reported tendency is Spherical particles accumulated more prominently in the liver. Rod-like
that nonspherical M&NPs had a prolonged circulation time in the blood particles (80 × 320 nm) loaded with docetaxel showed reduced
[2,41,62,181,182], and had a preferential accumulation in specific or­ phagocytosis by macrophages and prolonged circulation time in the
gans [2,17,39,181,183–187], compared to spherical particles. For blood compared to spherical particles, resulting from higher tumor
example, rod-shaped PEGylated Au particles (10 × 45 nm) had a longer exposure to the drug [181]. Additionally, nonspherical M&NPs showed
circulation time in the blood than spherical particles (50 nm in diam­ improved diffusion through tumoral tissues compared with spheroidal
eter), which preferentially accumulated in the liver [182]. Exceptional particles [201]. Chauhan et al. [199], revealed that nanorods (15 × 54
long circulation times (up to 1 week) were reported with highly flexible nm) diffused in a tumor-mimetic collagen gel ~5 times faster than the
filamentous micelles (~20 − 60 nm in cross-sectional diameter and a nanospheres (35 nm). The same nanorods penetrated tumors ~4 times
tuneable length ~2 − 18 μm) after intravenous injection in rats [2]. faster than nanospheres in mice bearing orthotopic tumors [199]. The
Comparatively, PEGylated spherical vesicles (120 nm) with similar two particles had similar plasma t1/2 and clearance in mice.
surface properties to filomicelles were cleared within 48 h. Several Although it is not possible to make a clear conclusion on the in vivo
mechanisms are involved in the prolonged circulation time of filamen­ fate of nonspherical M&NPs particles compared to spherical ones, there
tous micelles. One possible explanation could be their low phagocytosis is a piece of evidence that particles with a high AR and high flexibility
profiles. Indeed, high AR and high flexibility of the filamentous micelles had prolonged circulation in the blood. The increased accumulation of
reduced their interaction with macrophages, required for their inter­ elongated particles in tumoral tissues is promising. It could lead to an
nalization. Previous research on spherical nanoparticles with variable increase in the efficacy of chemotherapeutics and a reduction of the side
rigidity showed that rigid particles displayed higher uptake by macro­ effects by limiting the accumulation in healthy tissues. Those findings
phages than soft ones [30,188]. Rapid elimination of stiff nanoparticles should trigger further studies to design more effective nanomedicines by
was correlated to their higher macrophage capture [30]. When filomi­ acting on M&NP shape.
celle rigidity was increased by crosslinking, they were cleared within a
few hours. 5. Summary and future directions
On the other hand, the same study revealed that filomicelle size is a
critical attribute for their interaction with macrophages and subsequent While a consensus on the M&NPs shape dependence on biological
internalization [2]. Long worms were less captured by macrophages processes was reached, some hurdles should be pointed out at several
than short filomicelles or spheres, thus reducing their clearance. levels. The major issue is related to a discrepancy in particle

104
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

Table 4
Comparison of the effect of M&NP morphology on their behavior in vivo regarding their biodistribution, clearance, and therapeutic efficiency.
M&NPs Route of Shape and size Studied property Observations Ref
administration
(Animal used)

Mesoporous silica Intravenous (mice) Long rods (720 nm in Biodistribution (2 h, 24 h, - Two hours after injection, both short and long [202]
nanoparticles length, AR ~5) and 7 days after PEGylated nanorods accumulated mainly in the liver,
Short rods (185 nm in injection) spleen, and lung (over 80% of the injected dose). Short
length, AR ~1.5) PEGylated nanorods were more easily trapped in the
liver, while long PEGylated nanorods accumulated
preferentially in the lung. However, 24 h after
injection, both short and long PEGylated nanorods
were trapped in the spleen > lung > liver.
Excretion - Short PEGylated nanorods had a faster clearance than
long PEGylated nanorods in urine and feces.
Biocompatibility - There is no association between the shape and the
toxicity of the nanoparticles.
Micelles Intravenous (mice Filomicelles (~20 − Biodistribution (ap to 6 - The blood circulation of filomicelles was up to 6 days, [2]
and rats) 60 nm) × (2 − 18 μm) days) while spherical vesicles of similar chemistry were
cleared within two days after injection to rats.
- Four days after injection, the filomicelles accumulated
in the liver > spleen > kidney > lung.
Treatment - 7 days after injection to tumor-bearing mice,
paclitaxel-loaded filomicelles can more effectively
reduce tumor size compared to free paclitaxel
Micelles Intravenous Filomicelles Biodistribution (after 24 - Filomicelles remained in the blood for at least 24 h. [203]
(tumor-bearing Spheres h) - 24 h after injection, uptake was higher in the MPS
mice) organs (liver > spleen > lungs > heart > tumor >
kidney > brain).
Treatment - The high accumulation of paclitaxel-loaded filomi­
celles produced a higher mass shrinkage and tumor cell
death regarding the spherical control.
Rod-shaped nucleoprotein Intravenous Short rods (~18 × 60 Biodistribution - The results are represented as total tissue penetration [204]
nanoparticles derived from (tumor-bearing nm, AR ~3.5) (bladder + liver + spleen). PEGylated short rods
tobacco mosaic virus mice) Medium rods showed a peak tissue accumulation at 2 h, while
(Either PEGylated or RGD (~18 × 130 nm, AR PEGylated long rods showed a peak tissue
targeted) ~7) accumulation at 6 h.
Long rods Clearance - PEGylated long rods were cleared by the RE system as
(~18 × 300 nm, AR the nanoparticles were cleared via the liver and spleen.
~16) - PEGylated short rods were mainly cleared via the
bladder suggesting a renal filtration elimination.
Accumulation in the - The accumulation of both stealth and targeted
tumors nanorods in the tumors was AR-dependent. PEGylated
short rods showed the highest accumulation in tumors,
followed by median and long nanorods. RGD-targeted
median nanorods showed the best accumulation in
tumors.
PEGylated tobacco mosaic Intravenous Rods (~18 × 130 nm, Biodistribution - Rods and spherical particles accumulated mainly in the [205]
virus-derived nanoparticles AR ~7) liver and spleen.
Spheres (54 nm Clearance - Spherical particles showed shorter circulation in the
measured by TEM) blood and were more rapidly eliminated than rods.
They were no longer detected after 24 h of injection.
- Rod particles were detectable in the liver and spleen
after 24 h.
Liposomes Intravenous (rats) Spheres Biodistribution - The discoidal shape enhanced by nearly 2 folds the [1]
Disks circulation half-life of liposomes in a rat model.
Polymeric nanoparticles Intravenous Rods (37 nm in Biodistribution - Worms displayed the highest accumulation in the [206]
(polymerization induced (tumor-bearing diameter and 350 − spleen > liver > lungs > kidneys.
self-assembly) mice) 500 nm in length) Clearance - Regardless of their morphology, the particles were
Worms (44.5 nm in cleared from the blood circulation 2 days post-delivery.
diameter and 1 − 2 μm Accumulation in the - The highest tumor accumulation was obtained with
in length) tumors small and spherical micelles, whereas the lowest tumor
Spheres (~21 nm or 33 accumulation was observed for the worms.
nm)

composition, size, shape, surface properties. In several investigations, technological tools for designing nonspherical M&NPs with simulta­
the surface area and volume of M&NPs with different shapes are not neous control of their properties is still needed. Indeed, each technique
always kept constant and the biological process may be rather driven by used for preparing nonspherical particles has its advantages and draw­
other particle parameters than geometry. The isolation of the parame­ backs. For example, mainly micrometer-sized particles with different
ters one by one should lead to a better understanding of the exact role of shapes and AR can be obtained by photolithography [207], and micro­
the shape on biological processes. Such understanding requires robust fluidics [208]. By self-assembly of small amphiphilic molecules, a vari­
technological tools for designing particles with controlled properties in a ety of morphologies are harvested, including rods, nanoribbons,
reproducible manner. Not only by handling the shape but also 3D di­ filomicelles, and worms [209]. Generally speaking, shape modification
mensions, surface properties, and particle mechanical behaviors because is mediated by the composition of the amphiphilic molecules. Ring-
those parameters are interconnected. The development of novel opened benzyl-L-glutamate led to particles with different AR (up to 3.5)

105
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

[210]. However, particles with different AR do not have comparable observed that the height cross-section of flat particles obtained by hi­
surface properties since the shape is controlled by polymer molecular erarchical self-assembly of α-cyclodextrin and hydrophobically modified
weight. In some cases, nonspherical M&NPs did not always show shape hyaluronan in liquid were ~4.5-fold lower than that measured in dry
stability. For example, elliptical particles switch their shape to a sphere conditions [11]. An accurate characterization thus is essential for un­
[211]. The particles were obtained by physical deformation of poly derstanding and the correct interpretation of the biological behaviors.
(lactide-co-glycolide) spheres contained in a film composed of poly In short, despite the interest in particle shape in biological processes,
(vinyl alcohol). In a recent investigation, Ahmed et al. [23], succeeded the issues mentioned above should be addressed to speed up the trans­
in preparing shape-stable elliptical particles by film stretching tech­ lation of nonspherical particles into the clinic. Indeed, despite the efforts
nique. However, the same study highlighted some difficulties in in developing shape-controlled M&NPs, only a few examples are
removing poly(vinyl alcohol) from the particle preparation [23]. In commercially available. Besides inorganic particles, PRINTTM particles
addition to surface modification following the adsorption of poly(vinyl that are uniform in shape and bioresorbable (LIQ001) were tested in the
alcohol), moving from spherical to elongated shape could also modify clinic to improve immune response and efficacy of seasonal influenza
the particle mechanical properties, leading to mixed parameters. The vaccine. The therapeutic and commercial success of disk-like liposomes
exact implication of each parameter in the biological process could be (Amphotec®) and Ribbon-like particles (Abelcet®) should pave the way
complicated. to additional investigations for other applications.
Modeling the interaction of M&NPs with the cell membrane offers
the opportunity to isolate the parameters conferring a better under­ Funding
standing of fundamental events over particle wrapping. Such knowledge
will pave the way toward the correct interpretation of the experimental Author K.B. received funding from the Institut Universitaire de
data of the endocytosis process in different applications, including drug France and the ANR-17-CE09-0038-1.
delivery. Furthermore, the models developed for particle wrapping
could be applicable to a wide range of biological processes such as viral References
entry to cells, intracellular trafficking, and exocytosis. The basic struc­
tures representing only the double bilayer membrane and receptors [1] S. Li, J. Nickels, A.F. Palmer, Liposome-encapsulated actin–hemoglobin (LEAcHb)
artificial blood substitutes, Biomaterials 26 (17) (2005) 3759–3769.
could be further improved by adding complex structures of the cell [2] Y. Geng, P. Dalhaimer, S. Cai, R. Tsai, M. Tewari, T. Minko, D.E. Discher, Shape
membrane. The effect of clathrin and other envelope proteins should effects of filaments versus spherical particles in flow and drug delivery, Nat.
also be considered. Finally, it is necessary to consider physiologically Nanotechnol. 2 (4) (2007) 249.
[3] F. d’Agosto, J. Rieger, M. Lansalot, RAFT-mediated polymerization-induced self-
relevant conditions such as osmotic pressure. assembly, Angew. Chem. Int. Ed. 59 (22) (2020) 8368–8392.
On the other hand, in several reports, it was found that the biological [4] J. Rieger, Guidelines for the synthesis of block copolymer particles of various
effect was correlated to the purification level of the preparations. For morphologies by RAFT dispersion polymerization, Macromol. Rapid Commun. 36
(16) (2015) 1458–1471.
carbon nanotubes, cytotoxicity was rather due to metal traces contam­ [5] G.L. Mellot, Combinaison de la chimie supramoléculaire et de la PISA contrôlée
inants than the carbon nanotube itself [212]. For Au particles, the par RAFT pour synthétiser des nanofibres dans l’eau, Sorbonne Université, 2019.
cytotoxicity was mainly due to extrachemicals present in the prepara­ [6] S. Han, S. Pensec, D. Yilmaz, C. Lorthioir, J. Jestin, J.-M. Guigner, F. Niepceron,
J. Rieger, F. Stoffelbach, E. Nicol, O. Colombani, L. Bouteiller, Straightforward
tion. The free detergent hexadecyltrimethylammonium bromide, due to
preparation of supramolecular Janus nanorods by hydrogen bonding of end-
incomplete purification of the Au nanorods, was shown to be toxic to the functionalized polymers, Nat. Commun. 11 (1) (2020) 1–6.
cell membrane [112,213,214]. Standardization of the properties of the [7] A. Galus, J.-M. Mallet, D. Lembo, V. Cagno, M. Djabourov, H. Lortat-Jacob,
preparations, their purification, and the conditions for cytotoxicity ex­ K. Bouchemal, Hexagonal-shaped chondroitin sulfate self-assemblies have exalted
anti-HSV-2 activity, Carbohydr. Polym. 136 (2016) 113–120.
periments is necessary to investigate the effect of particle shape [8] F. Carn, S. Nowak, I. Chaab, R. Diaz-Salmeron, M. Djabourov, K. Bouchemal,
accurately. Autoassemblies of α-cyclodextrin and grafted polysaccharides: Crystal structure
Furthermore, the experimental conditions for the biological evalua­ and specific properties of the platelets, J. Phys. Chem. B 122 (22) (2018)
6055–6063.
tions, such as the cell lines, the ratio between the cells and the M&NPs, [9] R. Diaz-Salmeron, I. Chaab, F. Carn, M. Djabourov, K. Bouchemal, Pickering
and the animal model used, are of crucial importance. For example, it emulsions with α-cyclodextrin inclusions: Structure and thermal stability,
was shown that the same preparation was more rapidly eliminated from J. Colloid Interface Sci. 482 (2016) 48–57.
[10] R. Diaz-Salmeron, A. Da Costa, J.-P. Michel, G. Ponchel, K. Bouchemal, Real-time
the mice blood than rats. Comparisons between M&NPs can be reason­ visualization of morphology-dependent self-motion of hyaluronic acid
ably drawn between experiments conducted on similar cells and nanomaterials in water, Int. J. Pharm. 121172 (2021).
animals. [11] R. Diaz-Salmeron, J.-P. Michel, H. Hadji, E. Gout, R.R. Vivès, G. Ponchel,
K. Bouchemal, Role of the interactions with CD44 and supported bilayer
Another challenge in the future is related to the lack of standardi­ membranes in the cellular uptake of soft multivalent hyaluronan nanoparticles,
zation of the physicochemical techniques for the characterization of Colloids Surf. B: Biointerfaces (2021) 111916.
M&NPs dimensions. In several studies, the particle size was given for [12] R. Diaz-Salmeron, G. Ponchel, K. Bouchemal, Hierarchically built hyaluronan
nano-platelets have symmetrical hexagonal shape, flattened surfaces and
spherical and nonspherical particles, while nonspherical particles are
controlled size, Eur. J. Pharm. Sci. 133 (2019) 251–263.
typically characterized by at least two dimensions (e.g., length, width, [13] R. Diaz-Salmeron, G. Ponchel, J.-F. Gallard, K. Bouchemal, Hierarchical
thickness). Furthermore, most of the investigations used dynamic light supramolecular platelets from hydrophobically-modified polysaccharides and
scattering techniques to characterize the size of nonspherical particles. α-cyclodextrin: Effect of hydrophobization and α-cyclodextrin concentration on
platelet formation, Int. J. Pharm. 548 (1) (2018) 227–236.
However, this technique, based on the Stocks-Einstein equation, was not [14] D. Lembo, M. Donalisio, C. Laine, V. Cagno, A. Civra, E.P. Bianchini, N. Zeghbib,
adapted to compare the size of particles with different diffusion co­ K. Bouchemal, Auto-associative heparin nanoassemblies: a biomimetic platform
efficients. Indeed, in a recent investigation, Diaz-Salmeron et al. [10], against the heparan sulfate-dependent viruses HSV-1, HSV-2, HPV-16 and RSV,
Eur. J. Pharm. Biopharm. 88 (1) (2014) 275–282.
revealed that, in static conditions, hyaluronan particles with a high AR, [15] E.M. Enlow, J.C. Luft, M.E. Napier, J.M. DeSimone, Potent engineered PLGA
designated nanoplatelets, exhibited more linear trajectories and faster nanoparticles by virtue of exceptionally high chemotherapeutic loadings, Nano
diffusion in water than nanospheres. Electron microscopy techniques Lett. 11 (2) (2011) 808–813.
[16] S.E. Gratton, M.E. Napier, P.A. Ropp, S. Tian, J.M. DeSimone, Microfabricated
could provide information on the particle shape and dimensions. How­ particles for engineered drug therapies: elucidation into the mechanisms of
ever, imaging techniques that require the drying of the sample (e.g., cellular internalization of PRINT particles, Pharm. Res. 25 (12) (2008)
TEM, SEM, and AFM imaging in the air), did not lead to a correct 2845–2852.
[17] S.E. Gratton, P.D. Pohlhaus, J. Lee, J. Guo, M.J. Cho, J.M. DeSimone,
determination of the exact particle dimensions in liquid conditions, and Nanofabricated particles for engineered drug therapies: A preliminary
in turn, it could lead to a misinterpretation of the results obtained in biodistribution study of PRINTTM nanoparticles, J. Control. Release 121 (1–2)
culture media or in vivo. A better analysis of M&NPs dimensions was (2007) 10–18.
provided by cryo-TEM, cryo-SEM, and AFM in liquid conditions. It was

106
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

[18] S.E. Gratton, P.A. Ropp, P.D. Pohlhaus, J.C. Luft, V.J. Madden, M.E. Napier, J. [48] C. Corbo, R. Molinaro, A. Parodi, N.E. Toledano Furman, F. Salvatore,
M. DeSimone, The effect of particle design on cellular internalization pathways, E. Tasciotti, The impact of nanoparticle protein corona on cytotoxicity,
Proc. Natl. Acad. Sci. 105 (33) (2008) 11613–11618. immunotoxicity and target drug delivery, Nanomedicine 11 (1) (2016) 81–100.
[19] L. Florez, C. Herrmann, J.M. Cramer, C.P. Hauser, K. Koynov, K. Landfester, [49] A. Salvati, A.S. Pitek, M.P. Monopoli, K. Prapainop, F.B. Bombelli, D.R. Hristov, P.
D. Crespy, V. Mailänder, How shape influences uptake: interactions of anisotropic M. Kelly, C. Åberg, E. Mahon, K.A. Dawson, Transferrin-functionalized
polymer nanoparticles and human mesenchymal stem cells, Small 8 (14) (2012) nanoparticles lose their targeting capabilities when a biomolecule corona adsorbs
2222–2230. on the surface, Nat. Nanotechnol. 8 (2) (2013) 137–143.
[20] J.A. Champion, Y.K. Katare, S. Mitragotri, Making polymeric micro-and [50] F. Barbero, L. Russo, M. Vitali, J. Piella, I. Salvo, M.L. Borrajo, M. Busquets-Fité,
nanoparticles of complex shapes, Proc. Natl. Acad. Sci. 104 (29) (2007) R. Grandori, N.G. Bastús, E. Casals, Formation of the protein corona: the interface
11901–11904. between nanoparticles and the immune system, in: Seminars in Immunology,
[21] J.A. Champion, S. Mitragotri, Role of target geometry in phagocytosis, Proc. Natl. Elsevier, 2017, pp. 52–60.
Acad. Sci. 103 (13) (2006) 4930–4934. [51] D. Flaherty, Immunology for Pharmacy-E-Book, Elsevier Health Sci. 12 (2014) 97.
[22] C. Palazzo, G. Ponchel, J.J. Vachon, S. Villebrun, F. Agnely, C. Vauthier, [52] S. Fais, Cannibalism: a way to feed on metastatic tumors, Cancer Lett. 258 (2)
Obtaining nonspherical poly (alkylcyanoacrylate) nanoparticles by the stretching (2007) 155–164.
method applied with a marketed water-soluble film, Int. J. Polym. Mater. Polym. [53] Y. Brill-Karniely, D. Dror, T. Duanis-Assaf, Y. Goldstein, O. Schwob, T. Millo,
Biomater. 66 (8) (2017) 416–424. N. Orehov, T. Stern, M. Jaber, N. Loyfer, Triangular correlation (TrC) between
[23] Z. Ahmed, G. Ponchel, K. Bouchemal, Shape stability of ellipsoidal nanomaterials cancer aggressiveness, cell uptake capability, and cell deformability, Sci. Adv. 6
prepared by physical deformation, Int. J. Pharm. 609 (2021), 121178. (3) (2020) eaax2861.
[24] N. Fakhri, F.C. MacKintosh, B. Lounis, L. Cognet, M. Pasquali, Brownian motion [54] M. Kaksonen, A. Roux, Mechanisms of clathrin-mediated endocytosis, Nat. Rev.
of stiff filaments in a crowded environment, Science 330 (6012) (2010) Mol. Cell Biol. 19 (5) (2018) 313–326.
1804–1807. [55] L.M. Bareford, P.W. Swaan, Endocytic mechanisms for targeted drug delivery,
[25] Y. Han, A.M. Alsayed, M. Nobili, J. Zhang, T.C. Lubensky, A.G. Yodh, Brownian Adv. Drug Deliv. Rev. 59 (8) (2007) 748–758.
motion of an ellipsoid, Science 314 (5799) (2006) 626–630. [56] Z. Liu, P.A. Roche, Macropinocytosis in phagocytes: regulation of MHC class-II-
[26] A.C. Anselmo, C.L. Modery-Pawlowski, S. Menegatti, S. Kumar, D.R. Vogus, L. restricted antigen presentation in dendritic cells, Front. Physiol. 6 (2015) 1.
L. Tian, M. Chen, T.M. Squires, A. Sen Gupta, S. Mitragotri, Platelet-like [57] H.T. Haigler, J.A. McKANNA, S. Cohen, Rapid stimulation of pinocytosis in
nanoparticles: mimicking shape, flexibility, and surface biology of platelets to human carcinoma cells A-431 by epidermal growth factor, J. Cell Biol. 83 (1)
target vascular injuries, ACS Nano 8 (11) (2014) 11243–11253. (1979) 82–90.
[27] K. Vahidkhah, P. Bagchi, Microparticle shape effects on margination, near-wall [58] J.A. Swanson, Shaping cups into phagosomes and macropinosomes, Nat. Rev.
dynamics and adhesion in a three-dimensional simulation of red blood cell Mol. Cell Biol. 9 (8) (2008) 639–649.
suspension, Soft Matter 11 (11) (2015) 2097–2109. [59] C.M. Buckley, J.S. King, Drinking problems: mechanisms of macropinosome
[28] F. Gentile, C. Chiappini, D. Fine, R. Bhavane, M. Peluccio, M.M.-C. Cheng, X. Liu, formation and maturation, FEBS J. 284 (22) (2017) 3778–3790.
M. Ferrari, P. Decuzzi, The effect of shape on the margination dynamics of non- [60] J.S. Rossman, G.P. Leser, R.A. Lamb, Filamentous influenza virus enters cells via
neutrally buoyant particles in two-dimensional shear flows, J. Biomech. 41 (10) macropinocytosis, J. Virol. 86 (20) (2012) 10950–10960.
(2008) 2312–2318. [61] Y. Caì, E.N. Postnikova, J.G. Bernbaum, S. Yú, S. Mazur, N.M. Deiuliis, S.
[29] F. Gentile, A. Curcio, C. Indolfi, M. Ferrari, P. Decuzzi, The margination R. Radoshitzky, M.G. Lackemeyer, A. McCluskey, P.J. Robinson, Simian
propensity of spherical particles for vascular targeting in the microcirculation, hemorrhagic fever virus cell entry is dependent on CD163 and uses a clathrin-
J. Nanobiotechnol. 6 (1) (2008) 9. mediated endocytosis-like pathway, J. Virol. 89 (1) (2015) 844–856.
[30] A.C. Anselmo, M. Zhang, S. Kumar, D.R. Vogus, S. Menegatti, M.E. Helgeson, [62] S. Muro, C. Garnacho, J.A. Champion, J. Leferovich, C. Gajewski, E.
S. Mitragotri, Elasticity of nanoparticles influences their blood circulation, H. Schuchman, S. Mitragotri, V.R. Muzykantov, Control of endothelial targeting
phagocytosis, endocytosis, and targeting, ACS Nano 9 (3) (2015) 3169–3177. and intracellular delivery of therapeutic enzymes by modulating the size and
[31] S. Dasgupta, T. Auth, G. Gompper, Shape and orientation matter for the cellular shape of ICAM-1-targeted carriers, Mol. Ther. 16 (8) (2008) 1450–1458.
uptake of nonspherical particles, Nano Lett. 14 (2) (2014) 687–693. [63] J.W. Yoo, N. Doshi, S. Mitragotri, Endocytosis and intracellular distribution of
[32] A.H. Bahrami, R. Lipowsky, T.R. Weikl, The role of membrane curvature for the PLGA particles in endothelial cells: effect of particle geometry, Macromol. Rapid
wrapping of nanoparticles, Soft Matter 12 (2) (2016) 581–587. Commun. 31 (2) (2010) 142–148.
[33] C. Van Der Wel, A. Vahid, A. Šarić, T. Idema, D. Heinrich, D.J. Kraft, Lipid [64] P. Decuzzi, M. Ferrari, The receptor-mediated endocytosis of nonspherical
membrane-mediated attraction between curvature inducing objects, Sci. Rep. 6 particles, Biophys. J. 94 (10) (2008) 3790–3797.
(1) (2016) 1–10. [65] X. Zhao, D. Lu, F. Hao, R. Liu, Exploring the diameter and surface dependent
[34] C. Huang, Y. Zhang, H. Yuan, H. Gao, S. Zhang, Role of nanoparticle geometry in conformational changes in carbon nanotube-protein corona and the related
endocytosis: laying down to stand up, Nano Lett. 13 (9) (2013) 4546–4550. cytotoxicity, J. Hazard. Mater. 292 (2015) 98–107.
[35] N. Doshi, S. Mitragotri, Macrophages recognize size and shape of their targets, [66] X. Huang, X. Teng, D. Chen, F. Tang, J. He, The effect of the shape of mesoporous
PLoS One 5 (4) (2010), e10051. silica nanoparticles on cellular uptake and cell function, Biomaterials 31 (3)
[36] G. Sharma, D.T. Valenta, Y. Altman, S. Harvey, H. Xie, S. Mitragotri, J.W. Smith, (2010) 438–448.
Polymer particle shape independently influences binding and internalization by [67] H. Meng, S. Yang, Z. Li, T. Xia, J. Chen, Z. Ji, H. Zhang, X. Wang, S. Lin, C. Huang,
macrophages, J. Control. Release 147 (3) (2010) 408–412. Aspect ratio determines the quantity of mesoporous silica nanoparticle uptake by
[37] Y. Zhang, T.R. Nayak, H. Hong, W. Cai, Graphene: a versatile nanoplatform for a small GTPase-dependent macropinocytosis mechanism, ACS Nano 5 (6) (2011)
biomedical applications, Nanoscale 4 (13) (2012) 3833–3842. 4434–4447.
[38] Y. Zhang, S. Tekobo, Y. Tu, Q. Zhou, X. Jin, S.A. Dergunov, E. Pinkhassik, B. Yan, [68] H.L. Herd, A. Malugin, H. Ghandehari, Silica nanoconstruct cellular toleration
Permission to enter cell by shape: nanodisk vs nanosphere, ACS Appl. Mater. threshold in vitro, J. Control. Release 153 (1) (2011) 40–48.
Interfaces 4 (8) (2012) 4099–4105. [69] K.W. Yoon, Dead cell phagocytosis and innate immune checkpoint, BMB Rep. 50
[39] R. Agarwal, V. Singh, P. Jurney, L. Shi, S. Sreenivasan, K. Roy, Mammalian cells (10) (2017) 496.
preferentially internalize hydrogel nanodiscs over nanorods and use shape- [70] H.H. Gustafson, D. Holt-Casper, D.W. Grainger, H. Ghandehari, Nanoparticle
specific uptake mechanisms, Proc. Natl. Acad. Sci. 110 (43) (2013) 17247–17252. uptake: the phagocyte problem, Nano Today 10 (4) (2015) 487–510.
[40] Y. He, K. Park, Effects of the microparticle shape on cellular uptake, Mol. Pharm. [71] A. Vonarbourg, C. Passirani, P. Saulnier, J.-P. Benoit, Parameters influencing the
13 (7) (2016) 2164–2171. stealthiness of colloidal drug delivery systems, Biomaterials 27 (24) (2006)
[41] X. Hu, J. Hu, J. Tian, Z. Ge, G. Zhang, K. Luo, S. Liu, Polyprodrug amphiphiles: 4356–4373.
hierarchical assemblies for shape-regulated cellular internalization, trafficking, [72] K.B. Bodman-Smith, A.J. Melendez, I. Campbell, P.T. Harrison, J.M. Allen, J.
and drug delivery, J. Am. Chem. Soc. 135 (46) (2013) 17617–17629. G. Raynes, C-reactive protein-mediated phagocytosis and phospholipase D
[42] Z. Chu, S. Zhang, B. Zhang, C. Zhang, C.-Y. Fang, I. Rehor, P. Cigler, H.-C. Chang, signalling through the high-affinity receptor for immunoglobulin G (FcγRI),
G. Lin, R. Liu, Unambiguous observation of shape effects on cellular fate of Immunology 107 (2) (2002) 252–260.
nanoparticles, Sci. Rep. 4 (2014) 4495. [73] S. Gordon, in: D.J. Weatherall, J.G.G. Ledingham, D.A. Warrell (Eds.),
[43] Z. Chu, K. Miu, P. Lung, S. Zhang, S. Zhao, H.-C. Chang, G. Lin, Q. Li, Rapid Macrophage Activation in: Encyclopaedia of Immunology, 2nd ed., Academic
endosomal escape of prickly nanodiamonds: implications for gene delivery, Sci. Press, London, 1998.
Rep. 5 (1) (2015) 1–8. [74] E. Vachon, R. Martin, V. Kwok, V. Cherepanov, C.-W. Chow, C.M. Doerschuk,
[44] P. Chaturbedy, M. Kumar, K. Salikolimi, S. Das, S.H. Sinha, S. Chatterjee, J. Plumb, S. Grinstein, G.P. Downey, CD44-mediated phagocytosis induces inside-
B. Suma, T.K. Kundu, M. Eswaramoorthy, Shape-directed compartmentalized out activation of complement receptor-3 in murine macrophages, Blood, The J.
delivery of a nanoparticle-conjugated small-molecule activator of an epigenetic Am. Soc. Hematol. 110 (13) (2007) 4492–4502.
enzyme in the brain, J. Control. Release 217 (2015) 151–159. [75] P.S. Hiemstra, M.R. Daha, Opsonization, 1998.
[45] Z.P. Xu, M. Niebert, K. Porazik, T.L. Walker, H.M. Cooper, A.P. Middelberg, P. [76] R.C. May, L.M. Machesky, Phagocytosis and the actin cytoskeleton, J. Cell Sci.
P. Gray, P.F. Bartlett, G.Q.M. Lu, Subcellular compartment targeting of layered 114 (6) (2001) 1061–1077.
double hydroxide nanoparticles, J. Control. Release 130 (1) (2008) 86–94. [77] V. Jaumouillé, S. Grinstein, Molecular mechanisms of phagosome formation.
[46] C.D. Walkey, W.C. Chan, Understanding and controlling the interaction of Myeloid Cells in Health and Disease: A, Synthesis (2017) 507–526.
nanomaterials with proteins in a physiological environment, Chem. Soc. Rev. 41 [78] S.H. Kaufmann, H.L. Collins, U.E. Schaible, Immune responses to intracellular
(7) (2012) 2780–2799. bacteria, in: Clinical Immunology: Principles and Practice, Mosby, 2008,
[47] A.E. Nel, L. Mädler, D. Velegol, T. Xia, E.M. Hoek, P. Somasundaran, F. Klaessig, pp. 389–409.
V. Castranova, M. Thompson, Understanding biophysicochemical interactions at [79] C.-O. Wong, S. Gregory, H. Hu, Y. Chao, V.E. Sepúlveda, Y. He, D. Li-Kroeger, W.
the nano–bio interface, Nat. Mater. 8 (7) (2009) 543. E. Goldman, H.J. Bellen, K. Venkatachalam, Lysosomal degradation is required

107
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

for sustained phagocytosis of bacteria by macrophages, Cell Host Microbe 21 (6) [110] E. Morgan, D. Wupperfeld, D. Morales, N. Reich, Shape matters: Gold
(2017), 719-730. e6. nanoparticle shape impacts the biological activity of siRNA delivery, Bioconjug.
[80] D. Paul, S. Achouri, Y.-Z. Yoon, J. Herre, C.E. Bryant, P. Cicuta, Phagocytosis Chem. 30 (3) (2019) 853–860.
dynamics depends on target shape, Biophys. J. 105 (5) (2013) 1143–1150. [111] Y. Zhao, Y. Wang, F. Ran, Y. Cui, C. Liu, Q. Zhao, Y. Gao, D. Wang, S. Wang,
[81] A. Aderem, How to eat something bigger than your head, Cell 110 (1) (2002) 5–8. A comparison between sphere and rod nanoparticles regarding their in vivo
[82] J.A. Champion, S. Mitragotri, Shape induced inhibition of phagocytosis of biological behavior and pharmacokinetics, Sci. Rep. 7 (1) (2017) 1–11.
polymer particles, Pharm. Res. 26 (1) (2009) 244–249. [112] D. Fernando, S. Sulthana, Y. Vasquez, Cellular uptake and cytotoxicity of varying
[83] J.P. Rolland, B.W. Maynor, L.E. Euliss, A.E. Exner, G.M. Denison, J.M. DeSimone, aspect ratios of gold nanorods in HeLa cells, ACS Appl. Bio Mater. 3 (3) (2020)
Direct fabrication and harvesting of monodisperse, shape-specific 1374–1384.
nanobiomaterials, J. Am. Chem. Soc. 127 (28) (2005) 10096–10100. [113] Z. Ji, X. Wang, H. Zhang, S. Lin, H. Meng, B. Sun, S. George, T. Xia, A.E. Nel, J.
[84] J.Y. Kelly, J.M. DeSimone, Shape-specific, monodisperse nano-molding of protein I. Zink, Designed synthesis of CeO2 nanorods and nanowires for studying
particles, J. Am. Chem. Soc. 130 (16) (2008) 5438–5439. toxicological effects of high aspect ratio nanomaterials, ACS Nano 6 (6) (2012)
[85] F.E. Alemdaroglu, N.C. Alemdaroglu, P. Langguth, A. Herrmann, DNA block 5366–5380.
copolymer micelles–a combinatorial tool for cancer nanotechnology, Adv. Mater. [114] J. Tschopp, K. Schroder, NLRP3 inflammasome activation: The convergence of
20 (5) (2008) 899–902. multiple signalling pathways on ROS production? Nat. Rev. Immunol. 10 (3)
[86] B.D. Chithrani, A.A. Ghazani, W.C. Chan, Determining the size and shape (2010) 210–215.
dependence of gold nanoparticle uptake into mammalian cells, Nano Lett. 6 (4) [115] X. Wang, T. Xia, S. Addo Ntim, Z. Ji, S. Lin, H. Meng, C.-H. Chung, S. George,
(2006) 662–668. H. Zhang, M. Wang, Dispersal state of multiwalled carbon nanotubes elicits
[87] R.E. Yanes, D. Tarn, A.A. Hwang, D.P. Ferris, S.P. Sherman, C.R. Thomas, J. Lu, A. profibrogenic cellular responses that correlate with fibrogenesis biomarkers and
D. Pyle, J.I. Zink, F. Tamanoi, Involvement of lysosomal exocytosis in the fibrosis in the murine lung, ACS Nano 5 (12) (2011) 9772–9787.
excretion of mesoporous silica nanoparticles and enhancement of the drug [116] S. Sweeney, D. Grandolfo, P. Ruenraroengsak, T.D. Tetley, Functional
delivery effect by exocytosis inhibition, Small 9 (5) (2013) 697–704. consequences for primary human alveolar macrophages following treatment with
[88] D. Kim, Z.G. Gao, E.S. Lee, Y.H. Bae, In vivo evaluation of doxorubicin-loaded long, but not short, multiwalled carbon nanotubes, Int. J. Nanomedicine 10
polymeric micelles targeting folate receptors and early endosomal pH in drug- (2015) 3115.
resistant ovarian cancer, Mol. Pharm. 6 (5) (2009) 1353–1362. [117] S. Sweeney, D. Berhanu, S.K. Misra, A.J. Thorley, E. Valsami-Jones, T.D. Tetley,
[89] M. Fernández, F. Javaid, V. Chudasama, Advances in targeting the folate receptor Multi-walled carbon nanotube length as a critical determinant of bioreactivity
in the treatment/imaging of cancers, Chem. Sci. 9 (4) (2018) 790–810. with primary human pulmonary alveolar cells, Carbon 78 (2014) 26–37.
[90] K.E. Marshall, D.M. Vadukul, K. Staras, L.C. Serpell, Misfolded amyloid-β-42 [118] G. Oberdörster, V. Stone, K. Donaldson, Toxicology of nanoparticles: a historical
impairs the endosomal–lysosomal pathway, Cell. Mol. Life Sci. 77 (23) (2020) perspective, Nanotoxicology 1 (1) (2007) 2–25.
5031–5043. [119] C.A. Poland, R. Duffin, I. Kinloch, A. Maynard, W.A. Wallace, A. Seaton, V. Stone,
[91] G. Moku, S.K. Gulla, N.V. Nimmu, S. Khalid, A. Chaudhuri, Delivering anti-cancer S. Brown, W. MacNee, K. Donaldson, Carbon nanotubes introduced into the
drugs with endosomal pH-sensitive anti-cancer liposomes, Biomater. Sci. 4 (4) abdominal cavity of mice show asbestos-like pathogenicity in a pilot study, Nat.
(2016) 627–638. Nanotechnol. 3 (7) (2008) 423–428.
[92] E. Xu, W.M. Saltzman, A.S. Piotrowski-Daspit, Escaping the endosome: Assessing [120] S. Kang, J.-E. Kim, D. Kim, C.G. Woo, P.V. Pikhitsa, M.-H. Cho, M. Choi,
cellular trafficking mechanisms of non-viral vehicles, J. Control. Release 335 Comparison of cellular toxicity between multi-walled carbon nanotubes and
(2021) 465–480. onion-like shell-shaped carbon nanoparticles, J. Nanopart. Res. 17 (9) (2015)
[93] N. Chaudhary, D. Weissman, K.A. Whitehead, mRNA vaccines for infectious 1–11.
diseases: principles, delivery and clinical translation, Nat. Rev. Drug Discov. [121] S. Nahle, R. Safar, S. Grandemange, B. Foliguet, M. Lovera-Leroux, Z. Doumandji,
(2021) 1–22. A. Le Faou, O. Joubert, B. Rihn, L. Ferrari, Single wall and multiwall carbon
[94] K.A. Hajj, K.A. Whitehead, Tools for translation: non-viral materials for nanotubes induce different toxicological responses in rat alveolar macrophages,
therapeutic mRNA delivery, Nat. Rev. Mater. 2 (10) (2017) 1–17. J. Appl. Toxicol. 39 (5) (2019) 764–772.
[95] D. Pantarotto, J.-P. Briand, M. Prato, A. Bianco, Translocation of bioactive [122] A. Almutary, B. Sanderson, The MTT and crystal violet assays: potential
peptides across cell membranes by carbon nanotubes, Chem. Commun. 1 (2004) confounders in nanoparticle toxicity testing, Int. J. Toxicol. 35 (4) (2016)
16–17. 454–462.
[96] N.W.S. Kam, Z. Liu, H. Dai, Carbon nanotubes as intracellular transporters for [123] A. Kroll, M.H. Pillukat, D. Hahn, J. Schnekenburger, Interference of engineered
proteins and DNA: an investigation of the uptake mechanism and pathway, nanoparticles with in vitro toxicity assays, Arch. Toxicol. 86 (7) (2012)
Angew. Chem. Int. Ed. 45 (4) (2006) 577–581. 1123–1136.
[97] H. Jin, D.A. Heller, M.S. Strano, Single-particle tracking of endocytosis and [124] L. Braydich-Stolle, S. Hussain, J.J. Schlager, M.-C. Hofmann, In vitro cytotoxicity
exocytosis of single-walled carbon nanotubes in NIH-3T3 cells, Nano Lett. 8 (6) of nanoparticles in mammalian germline stem cells, Toxicol. Sci. 88 (2) (2005)
(2008) 1577–1585. 412–419.
[98] P. Desai, A. Venkataramanan, R. Schneider, M.K. Jaiswal, J.K. Carrow, [125] M. Rösslein, J.T. Elliott, M. Salit, E.J. Petersen, C. Hirsch, H.F. Krug, P. Wick, Use
A. Purwada, A. Singh, A.K. Gaharwar, Self-assembled, ellipsoidal polymeric of cause-and-effect analysis to design a high-quality nanocytotoxicology assay,
nanoparticles for intracellular delivery of therapeutics, J. Biomed. Mater. Res. A Chem. Res. Toxicol. 28 (1) (2015) 21–30.
106 (7) (2018) 2048–2058. [126] P. Kumar, A. Nagarajan, P.D. Uchil, Analysis of cell viability by the lactate
[99] H. Yang, Z. Chen, L. Zhang, W.Y. Yung, K.C.F. Leung, H.Y.E. Chan, C.H.J. Choi, dehydrogenase assay, Cold Spring Harb Protoc 2018 (6) (2018) pdb. prot095497.
Mechanism for the cellular uptake of targeted gold nanorods of defined aspect [127] V. Tzankova, D. Aluani, Y. Yordanov, M. Valoti, M. Frosini, I. Spassova,
ratios, Small 12 (37) (2016) 5178–5189. D. Kovacheva, B. Tzankov, In vitro toxicity evaluation of lomefloxacin-loaded
[100] O. Shimoni, Y. Yan, Y. Wang, F. Caruso, Shape-dependent cellular processing of MCM-41 mesoporous silica nanoparticles, Drug Chem. Toxicol. 44 (3) (2021)
polyelectrolyte capsules, ACS Nano 7 (1) (2013) 522–530. 238–249.
[101] J. Shi, J.L. Choi, B. Chou, R.N. Johnson, J.G. Schellinger, S.H. Pun, Effect of [128] Y. Zhang, M. Dai, Z. Yuan, Methods for the detection of reactive oxygen species,
polyplex morphology on cellular uptake, intracellular trafficking, and transgene Anal. Methods 10 (38) (2018) 4625–4638.
expression, ACS Nano 7 (12) (2013) 10612–10620. [129] B. Bakan, S. Gülcemal, S. Akgöl, P.H. Hoet, N.Ü.K. Yavaşoğlu, Synthesis,
[102] L.H. Huang, J. Han, J.M. Ouyang, B.S. Gui, Shape-dependent adhesion and characterization and toxicity assessment of a new polymeric nanoparticle, l-
endocytosis of hydroxyapatite nanoparticles on A7R5 aortic smooth muscle cells, glutamic acid-gp (HEMA), Chem. Biol. Interact. 315 (2020), 108870.
J. Cell. Physiol. 235 (1) (2020) 465–479. [130] A.-H. Lutter, J. Scholka, H. Richter, U. Anderer, Applying XTT, WST-1, and WST-8
[103] Z. Li, L. Sun, Y. Zhang, A.P. Dove, R.K. O’Reilly, G. Chen, Shape effect of glyco- to human chondrocytes: A comparison of membrane-impermeable tetrazolium
nanoparticles on macrophage cellular uptake and immune response, ACS Macro salts in 2D and 3D cultures, Clin. Hemorheol. Microcirc. 67 (3-4) (2017) 327–342.
Lett. 5 (9) (2016) 1059–1064. [131] W. Hu, S. Culloty, G. Darmody, S. Lynch, J. Davenport, S. Ramirez-Garcia,
[104] H. Herd, N. Daum, A.T. Jones, H. Huwer, H. Ghandehari, C.-M. Lehr, Nanoparticle K. Dawson, I. Lynch, H. Doyle, D. Sheehan, Neutral red retention time assay in
geometry and surface orientation influence mode of cellular uptake, ACS Nano 7 determination of toxicity of nanoparticles, Mar. Environ. Res. 111 (2015)
(3) (2013) 1961–1973. 158–161.
[105] X. Xie, J. Liao, X. Shao, Q. Li, Y. Lin, The effect of shape on cellular uptake of gold [132] B.C. Heng, G.K. Das, X. Zhao, L.-L. Ma, T.T.-Y. Tan, K.W. Ng, J.S.-C. Loo,
nanoparticles in the forms of stars, rods, and triangles, Sci. Rep. 7 (1) (2017) 1–9. Comparative cytotoxicity evaluation of lanthanide nanomaterials on mouse and
[106] L. Ding, C. Yao, X. Yin, C. Li, Y. Huang, M. Wu, B. Wang, X. Guo, Y. Wang, M. Wu, human cell lines with metabolic and DNA-quantification assays, Biointerphases 5
Size, shape, and protein corona determine cellular uptake and removal (3) (2010) FA88–FA97.
mechanisms of gold nanoparticles, Small 14 (42) (2018) 1801451. [133] A.P. Demchenko, Beyond annexin V: fluorescence response of cellular membranes
[107] E. Scarpa, C. De Pace, A.S. Joseph, S.C. de Souza, A. Poma, E. Liatsi-Douvitsa, to apoptosis, Cytotechnology 65 (2) (2013) 157–172.
C. Contini, V. De Matteis, J.S. Martí, G. Battaglia, Tuning cell behavior with [134] N. Fujita, A. Nagahashi, K. Nagashima, S. Rokudai, T. Tsuruo, Acceleration of
nanoparticle shape, PLoS One 15 (11) (2020), e0240197. apoptotic cell death after the cleavage of Bcl-X L protein by caspase-3-like
[108] D. Pantarotto, R. Singh, D. McCarthy, M. Erhardt, J.P. Briand, M. Prato, proteases, Oncogene 17 (10) (1998) 1295–1304.
K. Kostarelos, A. Bianco, Functionalized carbon nanotubes for plasmid DNA gene [135] J. Palomäki, P. Karisola, L. Pylkkänen, K. Savolainen, H. Alenius, Engineered
delivery, Angew. Chem. Int. Ed. 43 (39) (2004) 5242–5246. nanomaterials cause cytotoxicity and activation on mouse antigen presenting
[109] W. Hasan, K. Chu, A. Gullapalli, S.S. Dunn, E.M. Enlow, J.C. Luft, S. Tian, M. cells, Toxicology 267 (1-3) (2010) 125–131.
E. Napier, P.D. Pohlhaus, J.P. Rolland, Delivery of multiple siRNAs using lipid- [136] X. Wang, Y. Xia, L. Liu, M. Liu, N. Gu, H. Guang, F. Zhang, Comparison of MTT
coated PLGA nanoparticles for treatment of prostate cancer, Nano Lett. 12 (1) assay, flow cytometry, and RT-PCR in the evaluation of cytotoxicity of five
(2012) 287–292.

108
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

prosthodontic materials, J Biomed Mater Res B Appl Biomater 95 (2) (2010) [169] R. Gref, Y. Minamitake, M.T. Peracchia, V. Trubetskoy, V. Torchilin, R. Langer,
357–364. Biodegradable long-circulating polymeric nanospheres, Science 263 (5153)
[137] M. Dekaliuk, K. Pyrshev, A. Demchenko, Visualization and detection of live and (1994) 1600–1603.
apoptotic cells with fluorescent carbon nanoparticles, J. Nanobiotechnol. 13 (1) [170] G.T. Kozma, T. Mészáros, I. Vashegyi, T.S. Fülöp, E. Örfi, L. Dézsi, L. Rosivall,
(2015) 1–8. Y. Bavli, R. Urbanics, T.E. Mollnes, Pseudo-anaphylaxis to polyethylene glycol
[138] D.J. Taatjes, B.E. Sobel, R.C. Budd, Morphological and cytochemical (PEG)-coated liposomes: Roles of Anti-PEG IgM and complement activation in a
determination of cell death by apoptosis, Histochem. Cell Biol. 129 (1) (2008) porcine model of human infusion reactions, ACS Nano 13 (8) (2019) 9315–9324.
33–43. [171] S. Tenzer, D. Docter, S. Rosfa, A. Wlodarski, J.R. Kuharev, A. Rekik, S.K. Knauer,
[139] S. Malli, C. Bories, M. Bourge, P.M. Loiseau, K. Bouchemal, Surface-dependent C. Bantz, T. Nawroth, C. Bier, Nanoparticle size is a critical physicochemical
endocytosis of poly(isobutylcyanoacrylate) nanoparticles by Trichomonas determinant of the human blood plasma corona: a comprehensive quantitative
vaginalis, Int. J. Pharm. 548 (1) (2018) 276–287. proteomic analysis, ACS Nano 5 (9) (2011) 7155–7167.
[140] S. Malli, S. Pomel, Y. Ayadi, C. Delomenie, A. Da Costa, P.M. Loiseau, [172] M. Lundqvist, J. Stigler, G. Elia, I. Lynch, T. Cedervall, K.A. Dawson, Nanoparticle
K. Bouchemal, Topically applied chitosan-coated poly(isobutylcyanoacrylate) size and surface properties determine the protein corona with possible
nanoparticles are active against cutaneous Leishmaniasis by accelerating lesion implications for biological impacts, Proc. Natl. Acad. Sci. 105 (38) (2008)
healing and reducing the parasitic load, ACS Appl. Bio Mater. 2 (6) (2019) 14265–14270.
2573–2586. [173] M.A. Dobrovolskaia, A.K. Patri, J. Zheng, J.D. Clogston, N. Ayub, P. Aggarwal, B.
[141] X. Zhao, S. Ng, B.C. Heng, J. Guo, L. Ma, T.T.Y. Tan, K.W. Ng, S.C.J. Loo, W. Neun, J.B. Hall, S.E. McNeil, Interaction of colloidal gold nanoparticles with
Cytotoxicity of hydroxyapatite nanoparticles is shape and cell dependent, Arch. human blood: effects on particle size and analysis of plasma protein binding
Toxicol. 87 (6) (2013) 1037–1052. profiles, Nanomedicine 5 (2) (2009) 106–117.
[142] L. Dong, S. Tang, F. Deng, Y. Gong, K. Zhao, J. Zhou, D. Liang, J. Fang, M. Hecker, [174] P. Koegler, A. Clayton, H. Thissen, G.N.C. Santos, P. Kingshott, The influence of
J.P. Giesy, Shape-dependent toxicity of alumina nanoparticles in rat astrocytes, nanostructured materials on biointerfacial interactions, Adv. Drug Deliv. Rev. 64
Sci. Total Environ. 690 (2019) 158–166. (15) (2012) 1820–1839.
[143] H.T. McMahon, E. Boucrot, Membrane curvature at a glance, J. Cell Sci. 128 (6) [175] D. Dutta, S.K. Sundaram, J.G. Teeguarden, B.J. Riley, L.S. Fifield, J.M. Jacobs, S.
(2015) 1065–1070. R. Addleman, G.A. Kaysen, B.M. Moudgil, T.J. Weber, Adsorbed proteins
[144] F. Villanueva-Flores, A. Castro-Lugo, O.T. Ramírez, L.A. Palomares, influence the biological activity and molecular targeting of nanomaterials,
Understanding cellular interactions with nanomaterials: towards a rational design Toxicol. Sci. 100 (1) (2007) 303–315.
of medical nanodevices, Nanotechnology 31 (13) (2020), 132002. [176] S. Chakraborty, P. Joshi, V. Shanker, Z. Ansari, S.P. Singh, P. Chakrabarti,
[145] M. Deserno, T. Bickel, Wrapping of a spherical colloid by a fluid membrane, EPL Contrasting effect of gold nanoparticles and nanorods with different surface
(Europhysics Letters) 62 (5) (2003) 767. modifications on the structure and activity of bovine serum albumin, Langmuir 27
[146] M. Deserno, Elastic deformation of a fluid membrane upon colloid binding, Phys. (12) (2011) 7722–7731.
Rev. E 69 (3) (2004), 031903. [177] A.J. Andersen, S.H. Hashemi, T.L. Andresen, A.C. Hunter, S.M. Moghimi,
[147] S. Zhang, H. Gao, G. Bao, Physical principles of nanoparticle cellular endocytosis, Complement: alive and kicking nanomedicines, J. Biomed. Nanotechnol. 5 (4)
ACS Nano 9 (9) (2015) 8655–8671. (2009) 364–372.
[148] R. De, A general model of focal adhesion orientation dynamics in response to [178] M.B. Pedersen, X. Zhou, E.K.U. Larsen, U.S. Sørensen, J. Kjems, J.V. Nygaard, J.
static and cyclic stretch, Commun. Biol. 1 (1) (2018) 1–7. R. Nyengaard, R.L. Meyer, T. Boesen, T. Vorup-Jensen, Curvature of synthetic and
[149] Y. Brill-Karniely, Mechanical Measurements of Cells Using AFM: 3D or 2D natural surfaces is an important target feature in classical pathway complement
Physics? Front. Bioeng. Biotechnol. 8 (2020). activation, J. Immunol. 184 (4) (2010) 1931–1945.
[150] S. Zhang, J. Li, G. Lykotrafitis, G. Bao, S. Suresh, Size-dependent endocytosis of [179] S.M. Moghimi, A.J. Andersen, D. Ahmadvand, P.P. Wibroe, T.L. Andresen, A.
nanoparticles, Adv. Mater. 21 (4) (2009) 419–424. C. Hunter, Material properties in complement activation, Adv. Drug Deliv. Rev.
[151] G. Bao, X.R. Bao, Shedding light on the dynamics of endocytosis and viral 63 (12) (2011) 1000–1007.
budding, Proc. Natl. Acad. Sci. 102 (29) (2005) 9997–9998. [180] B.J. Janssen, A. Christodoulidou, A. McCarthy, J.D. Lambris, P. Gros, Structure of
[152] P. Rangamani, K.K. Mandadap, G. Oster, Protein-induced membrane curvature C3b reveals conformational changes that underlie complement activity, Nature
alters local membrane tension, Biophys. J. 107 (3) (2014) 751–762. 444 (7116) (2006) 213.
[153] G.I. Bell, Models for the specific adhesion of cells to cells, Science 200 (4342) [181] K.S. Chu, W. Hasan, S. Rawal, M.D. Walsh, E.M. Enlow, J.C. Luft, A.S. Bridges, J.
(1978) 618–627. L. Kuijer, M.E. Napier, W.C. Zamboni, Plasma, tumor and tissue pharmacokinetics
[154] H. Gao, W. Shi, L.B. Freund, Mechanics of receptor-mediated endocytosis, Proc. of Docetaxel delivered via nanoparticles of different sizes and shapes in mice
Natl. Acad. Sci. 102 (27) (2005) 9469–9474. bearing SKOV-3 human ovarian carcinoma xenograft, Nanomedicine 9 (5) (2013)
[155] F. Osaki, T. Kanamori, S. Sando, T. Sera, Y. Aoyama, A quantum dot conjugated 686–693.
sugar ball and its cellular uptake. On the size effects of endocytosis in the subviral [182] M. Arnida, A. Ray, C. Peterson, H. Ghandehari, Geometry and surface
region, J. Am. Chem. Soc. 126 (21) (2004) 6520–6521. characteristics of gold nanoparticles influence their biodistribution and uptake by
[156] B.D. Chithrani, W.C. Chan, Elucidating the mechanism of cellular uptake and macrophages, Eur. J. Pharm. Biopharm. 77 (3) (2011) 417.
removal of protein-coated gold nanoparticles of different sizes and shapes, Nano [183] T. Chen, X. Guo, X. Liu, S. Shi, J. Wang, C. Shi, Z. Qian, S. Zhou, A strategy in the
Lett. 7 (6) (2007) 1542–1550. design of micellar shape for cancer therapy, Adv. Healthcare Mater. 1 (2) (2012)
[157] H. Yuan, S. Zhang, Effects of particle size and ligand density on the kinetics of 214–224.
receptor-mediated endocytosis of nanoparticles, Appl. Phys. Lett. 96 (3) (2010), [184] W. Li, X. Zhang, X. Hao, J. Jie, B. Tian, X. Zhang, Shape design of high drug
033704. payload nanoparticles for more effective cancer therapy, Chem. Commun. 49 (93)
[158] V. Schubertová, F.J. Martinez-Veracoechea, R. Vácha, Influence of ligand (2013) 10989–10991.
distribution on uptake efficiency, Soft Matter 11 (14) (2015) 2726–2730. [185] A.L. Van De Ven, P. Kim, J.R. Fakhoury, G. Adriani, J. Schmulen, P. Moloney,
[159] L. Li, Y. Zhang, J. Wang, Effects of ligand distribution on receptor-diffusion- F. Hussain, M. Ferrari, X. Liu, S.-H. Yun, Rapid tumoritropic accumulation of
mediated cellular uptake of nanoparticles, R. Soc. Open Sci. 4 (5) (2017), 170063. systemically injected plateloid particles and their biodistribution, J. Control.
[160] X. Yi, X. Shi, H. Gao, A universal law for cell uptake of one-dimensional Release 158 (1) (2012) 148–155.
nanomaterials, Nano Lett. 14 (2) (2014) 1049–1055. [186] D.L. Jasinski, H. Li, P. Guo, The effect of size and shape of RNA nanoparticles on
[161] R. Vácha, F.J. Martinez-Veracoechea, D. Frenkel, Receptor-mediated endocytosis biodistribution, Mol. Ther. 26 (3) (2018) 784–792.
of nanoparticles of various shapes, Nano Lett. 11 (12) (2011) 5391–5395. [187] Y. Yin, Y. Hao, N. Wang, P. Yang, N. Li, X. Zhang, Y. Song, X. Feng, W. Ma, PPy
[162] L. Chen, S. Xiao, H. Zhu, L. Wang, H. Liang, Shape-dependent internalization nanoneedle based nanoplatform capable of overcoming biological barriers for
kinetics of nanoparticles by membranes, Soft Matter 12 (9) (2016) 2632–2641. synergistic chemo-photothermal therapy, RSC Adv. 10 (13) (2020) 7771–7779.
[163] L. Chen, X. Li, Y. Zhang, T. Chen, S. Xiao, H. Liang, Morphological and mechanical [188] Y. Hui, D. Wibowo, Y. Liu, R. Ran, H.-F. Wang, A. Seth, A.P. Middelberg, C.-
determinants of cellular uptake of deformable nanoparticles, Nanoscale 10 (25) X. Zhao, Understanding the effects of nanocapsular mechanical property on
(2018) 11969–11979. passive and active tumor targeting, ACS Nano 12 (3) (2018) 2846–2857.
[164] Y. Li, M. Kröger, W.K. Liu, Shape effect in cellular uptake of PEGylated [189] P. Kolhar, A.C. Anselmo, V. Gupta, K. Pant, B. Prabhakarpandian, E. Ruoslahti,
nanoparticles: comparison between sphere, rod, cube and disk, Nanoscale 7 (40) S. Mitragotri, Using shape effects to target antibody-coated nanoparticles to lung
(2015) 16631–16646. and brain endothelium, Proc. Natl. Acad. Sci. 110 (26) (2013) 10753–10758.
[165] T. Cedervall, I. Lynch, S. Lindman, T. Berggård, E. Thulin, H. Nilsson, K. [190] R. Toy, E. Hayden, C. Shoup, H. Baskaran, E. Karathanasis, The effects of particle
A. Dawson, S. Linse, Understanding the nanoparticle–protein corona using size, density and shape on margination of nanoparticles in microcirculation,
methods to quantify exchange rates and affinities of proteins for nanoparticles, Nanotechnology 22 (11) (2011), 115101.
Proc. Natl. Acad. Sci. 104 (7) (2007) 2050–2055. [191] S.-Y. Lee, M. Ferrari, P. Decuzzi, Shaping nano-/micro-particles for enhanced
[166] A.L. Barrán-Berdón, D. Pozzi, G. Caracciolo, A.L. Capriotti, G. Caruso, vascular interaction in laminar flows, Nanotechnology 20 (49) (2009), 495101.
C. Cavaliere, A. Riccioli, S. Palchetti, A. Laganà, Time evolution of [192] H.T. Ta, N.P. Truong, A.K. Whittaker, T.P. Davis, K. Peter, The effects of particle
nanoparticle–protein corona in human plasma: relevance for targeted drug size, shape, density and flow characteristics on particle margination to vascular
delivery, Langmuir 29 (21) (2013) 6485–6494. walls in cardiovascular diseases, Expert Opin. Drug Deliv. 15 (1) (2018) 33–45.
[167] D. Walczyk, F.B. Bombelli, M.P. Monopoli, I. Lynch, K.A. Dawson, What the cell [193] M. Cooley, A. Sarode, M. Hoore, D.A. Fedosov, S. Mitragotri, A.S. Gupta, Influence
“sees” in bionanoscience, J. Am. Chem. Soc. 132 (16) (2010) 5761–5768. of particle size and shape on their margination and wall-adhesion: implications in
[168] Y.-N. Zhang, W. Poon, A.J. Tavares, I.D. McGilvray, W.C. Chan, drug delivery vehicle design across nano-to-micro scale, Nanoscale 10 (32) (2018)
Nanoparticle–liver interactions: cellular uptake and hepatobiliary elimination, 15350–15364.
J. Control. Release 240 (2016) 332–348. [194] A. Da Silva-Candal, T. Brown, V. Krishnan, I. Lopez-Loureiro, P. Ávila-Gómez,
A. Pusuluri, A. Pérez-Díaz, C. Correa-Paz, P. Hervella, J. Castillo, Shape effect in

109
H. Hadji and K. Bouchemal Journal of Controlled Release 342 (2022) 93–110

active targeting of nanoparticles to inflamed cerebral endothelium under static the biodistribution and tumor homing of rigid soft-matter nanorods, Adv.
and flow conditions, J. Control. Release 309 (2019) 94–105. Healthcare Mater. 4 (6) (2015) 874–882.
[195] P. Jurney, R. Agarwal, V. Singh, D. Choi, K. Roy, S. Sreenivasan, L. Shi, Unique [205] M.A. Bruckman, L.N. Randolph, A. VanMeter, S. Hern, A.J. Shoffstall, R.
size and shape-dependent uptake behaviors of non-spherical nanoparticles by E. Taurog, N.F. Steinmetz, Biodistribution, pharmacokinetics, and blood
endothelial cells due to a shearing flow, J. Control. Release 245 (2017) 170–176. compatibility of native and PEGylated tobacco mosaic virus nano-rods and-
[196] X. Zhu, C. Vo, M. Taylor, B.R. Smith, Non-spherical micro-and nanoparticles in spheres in mice, Virology 449 (2014) 163–173.
nanomedicine, Mater. Hor. 6 (2019) 1094–1121. [206] S. Kaga, N.P. Truong, L. Esser, D. Senyschyn, A. Sanyal, R. Sanyal, J.F. Quinn, T.
[197] J.W. Myerson, A.C. Anselmo, Y. Liu, S. Mitragotri, D.M. Eckmann, V. P. Davis, L.M. Kaminskas, M.R. Whittaker, Influence of size and shape on the
R. Muzykantov, Non-affinity factors modulating vascular targeting of nano-and biodistribution of nanoparticles prepared by polymerization-induced self-
microcarriers, Adv. Drug Deliv. Rev. 99 (2016) 97–112. assembly, Biomacromolecules 18 (12) (2017) 3963–3970.
[198] B.R. Smith, P. Kempen, D. Bouley, A. Xu, Z. Liu, N. Melosh, H. Dai, R. Sinclair, S. [207] X. Fu, J. Cai, X. Zhang, W.-D. Li, H. Ge, Y. Hu, Top-down fabrication of shape-
S. Gambhir, Shape matters: intravital microscopy reveals surprising geometrical controlled, monodisperse nanoparticles for biomedical applications, Adv. Drug
dependence for nanoparticles in tumor models of extravasation, Nano Lett. 12 (7) Deliv. Rev. 132 (2018) 169–187.
(2012) 3369–3377. [208] M. Maeki, N. Kimura, Y. Sato, H. Harashima, M. Tokeshi, Advances in
[199] V.P. Chauhan, Z. Popović, O. Chen, J. Cui, D. Fukumura, M.G. Bawendi, R.K. Jain, microfluidics for lipid nanoparticles and extracellular vesicles and applications in
Fluorescent nanorods and nanospheres for real-time in vivo probing of drug delivery systems, Adv. Drug Deliv. Rev. 128 (2018) 84–100.
nanoparticle shape-dependent tumor penetration, Angew. Chem. 123 (48) (2011) [209] Y. Mai, A. Eisenberg, Self-assembly of block copolymers, Chem. Soc. Rev. 41 (18)
11619–11622. (2012) 5969–5985.
[200] B. Godin, C. Chiappini, S. Srinivasan, J.F. Alexander, K. Yokoi, M. Ferrari, [210] O. Cauchois, F. Segura-Sanchez, G. Ponchel, Molecular weight controls the
P. Decuzzi, X. Liu, Discoidal porous silicon particles: fabrication and elongation of oblate-shaped degradable poly (γ-benzyl-L-glutamate)
biodistribution in breast cancer bearing mice, Adv. Funct. Mater. 22 (20) (2012) nanoparticles, Int. J. Pharm. 452 (1-2) (2013) 292–299.
4225–4235. [211] J.-W. Yoo, E. Chambers, S. Mitragotri, Factors that control the circulation time of
[201] Q. Sun, T. Ojha, F. Kiessling, T. Lammers, Y. Shi, Enhancing tumor penetration of nanoparticles in blood: challenges, solutions and future prospects, Curr. Pharm.
nanomedicines, Biomacromolecules 18 (5) (2017) 1449–1459. Des. 16 (21) (2010) 2298–2307.
[202] X. Huang, L. Li, T. Liu, N. Hao, H. Liu, D. Chen, F. Tang, The shape effect of [212] K. Pulskamp, S. Diabaté, H.F. Krug, Carbon nanotubes show no sign of acute
mesoporous silica nanoparticles on biodistribution, clearance, and toxicity but induce intracellular reactive oxygen species in dependence on
biocompatibility in vivo, ACS Nano 5 (7) (2011) 5390–5399. contaminants, Toxicol. Lett. 168 (1) (2007) 58–74.
[203] D.A. Christian, S. Cai, O.B. Garbuzenko, T. Harada, A.L. Zajac, T. Minko, D. [213] C.J. Murphy, A.M. Gole, J.W. Stone, P.N. Sisco, A.M. Alkilany, E.C. Goldsmith, S.
E. Discher, Flexible filaments for in vivo imaging and delivery: persistent C. Baxter, Gold nanoparticles in biology: beyond toxicity to cellular imaging, Acc.
circulation of filomicelles opens the dosage window for sustained tumor Chem. Res. 41 (12) (2008) 1721–1730.
shrinkage, Mol. Pharm. 6 (5) (2009) 1343–1352. [214] A.M. Alkilany, C.J. Murphy, Toxicity and cellular uptake of gold nanoparticles:
[204] S. Shukla, F.J. Eber, A.S. Nagarajan, N.A. DiFranco, N. Schmidt, A.M. Wen, what we have learned so far? J. Nanopart. Res. 12 (7) (2010) 2313–2333.
S. Eiben, R.M. Twyman, C. Wege, N.F. Steinmetz, The impact of aspect ratio on

110

You might also like