ArticlePDF Available

Abstract and Figures

This review shows the most common and promising strategies to generate colloidally stable silica nanoparticles (NPs) in simulated biological fluids and sheds light on the latest advances in producing degradable silica-based structures. Silica NPs can be synthesized in a wide variety of morphologies, porosity levels, and sizes. This versatility makes silica NPs one of the most promising nano-platforms for imaging and disease treatment. Nonetheless, biological barriers can decrease the success of translating them for therapeutic applications since the media composition can induce their colloidal stability loss. It can, consequently, lead to the NPs aggregation and affect their degradation profile. The interplay between NPs aggregation and degradation has been scarcely explored in the literature when biological fluids are seriously taken into account. Herein we discuss the theory behind the colloidal stability of silica NPs, the processes leading to their aggregation, and some strategies to overcome this issue (mainly focused on NPs surface functionalization). Furthermore, we addressed the main issues that affect the degradability of NPs in biological fluids, and explored some strategies, such as chemical surface modification, which are able to tune these degradation-driven profiles. Thus, the understanding of the silica NPs behavior in body fluids is essential for the approval of nanomedicines and, therefore, more investigations concerning the dynamics, thermodynamics, biological response, and structural parameters of silica-based NPs are of utmost importance.
Content may be subject to copyright.
Journal of Sol-Gel Science and Technology
https://doi.org/10.1007/s10971-021-05695-8
INVITED REVIEW: SOLGEL AND HYBRID MATERIALS FOR BIOLOGICAL
AND HEALTH (MEDICAL) APPLICATIONS
Colloidal stability and degradability of silica nanoparticles in
biological uids: a review
Andressa da Cruz Schneid1Lindomar Jose Calumby Albuquerque1Gabriela Borba Mondo1,2 Marcelo Ceolin3
Agustin Silvio Picco3Mateus Borba Cardoso 1,2
Received: 14 May 2021 / Accepted: 18 November 2021
© The Author(s), under exclusive licence to Springer Science+Business Media, LLC, part of Springer Nature 2022
Abstract
This review shows the most common and promising strategies to generate colloidally stable silica nanoparticles (NPs) in
simulated biological uids and sheds light on the latest advances in producing degradable silica-based structures. Silica NPs
can be synthesized in a wide variety of morphologies, porosity levels, and sizes. This versatility makes silica NPs one of the
most promising nano-platforms for imaging and disease treatment. Nonetheless, biological barriers can decrease the success
of translating them for therapeutic applications since the media composition can induce their colloidal stability loss. It can,
consequently, lead to the NPs aggregation and affect their degradation prole. The interplay between NPs aggregation and
degradation has been scarcely explored in the literature when biological uids are seriously taken into account. Herein we
discuss the theory behind the colloidal stability of silica NPs, the processes leading to their aggregation, and some strategies
to overcome this issue (mainly focused on NPs surface functionalization). Furthermore, we addressed the main issues that
affect the degradability of NPs in biological uids, and explored some strategies, such as chemical surface modication,
which are able to tune these degradation-driven proles. Thus, the understanding of the silica NPs behavior in body uids is
essential for the approval of nanomedicines and, therefore, more investigations concerning the dynamics, thermodynamics,
biological response, and structural parameters of silica-based NPs are of utmost importance.
These authors contributed equally: Andressa da Cruz Schneid,
Lindomar Jose Calumby Albuquerque, Gabriela Borba Mondo
*Agustin Silvio Picco
apicco@inifta.unlp.edu.ar
*Mateus Borba Cardoso
cardosomb@lnls.br
1Laboratório Nacional de Luz Sincrotron (LNLS), Centro Nacional
de Pesquisa em Energia e Materiais (CNPEM), CEP 13083-970,
Campinas, SP, Brasil
2Instituto de Química (IQ), Universidade Estadual de Campinas
(UNICAMP), CEP 13083-970, Caixa Postal 6154, Campinas, SP,
Brasil
3Instituto de Investigaciones Fisicoquímicas Teóricas y Aplicadas
(INIFTA-UNLP-CONICET), 1900 La Plata, Argentina
1234567890();,:
1234567890();,:
Graphical Abstract
Keywords Silica nanoparticles Colloidal stability Aggregation Degradation Biological uids
Highlights
We discuss the challenges faced by silica-based nanoparticles in biological media.
We address relevant strategies to avoid aggregation of silica nanoparticles in biological uids.
We present the impact of media and particle characteristics on silica nanoparticles degradation.
We discuss strategies used to tune silica nanoparticle degradability in the biological environment.
1 Introduction
Silica nanoparticles (NPs) have emerged over the last few
decades as a promising nanoplatform for diseases treatment
and diagnosis [14]. Tailoring silica NP properties allows
the production of a limitless range of nanostructures, mak-
ing them a versatile platform that can be ne-tuned to dis-
tinct applications [5]. They have been engineered for
medical purposes into different structures, including non-
porous [68], mesoporous [911], and hollow silica NPs
(MSNs and HSNs, respectively) [8,12,13], as schematized
in Fig. 1a. Fundamental knowledge on the synthesis,
solgel chemistry, and physical chemistry of silica nanos-
tructures has been extensively documented in classical
books such as the ones authored by Iler [14], and Brinker
and Scherer [15], or edited by Bergna and Roberts [16]. The
ultimate design of these particles is engineered to fulll
specic nanomedicine purposes since each system presents
unique structural properties and [5,17,18] results in dif-
ferent biological media responses [1921]. Non-porous
silica NPs are among the most efcient platforms to eval-
uate the inuence of surface functionalization in biological
environments [22,23] while MSNs and HSNs can entrap
high therapeutic loads for drug delivery applications [5,19].
Silica surfaces can also be found in composite structures as
a shell surrounding cores of metallic NPs (e.g., gold),
magnetic NPs (e.g., SPIONs), or quantum dots, among
other materials [4,18,24]. These core@shell nanostructures
benet from several well-known silica functionalization
routes and the biocompatibility of silica-based materials
[25,26].
The use of silica in the biomedical eld seems evident
since it is already part of the daily human diet throughout
freshwater, seafood, grains, and vegetable consumption
[27]. After ingestion, silicon-based compounds reach the
bloodstream, where a fraction is adsorbed as silicic acid (the
main product of silica degradation), while the remaining
portion (ca. 41%) is excreted in the urine [27,28]. In 2010,
silica was accepted as Generally Recognized As Safeby
the US Food and Drug Administration (FDA) and has been
used in cosmetics and FDA-approved food [5,27]. None-
theless, this safeness cannot be directly extrapolated to the
NP form of silica, since they can trigger different body
responses depending on their structural properties and sur-
face functionalization [3,21]. In this context, the approval
of Cornell dots (C-dots) for clinic imaging trials by the FDA
in 2014 was a milestone for the silica-based NPs research
[5,29], boosting the studies using this platform for treating
diseases (cancer, bacteria, and virus infection) and imaging
diagnostics [5,3032]. Recently, the approval of Sebacia
and AuroLase for the treatment of severe inammatory acne
vulgaris and solid tumors, respectively, highlighted the
potential of inorganic-based silica-containing nanomedi-
cines [33].
Journal of Sol-Gel Science and Technology
The medical potential of silica-based nanomaterials has
been widely demonstrated in simple uids, such as water
and buffers. However, their physicochemical properties and
designed biological activity might considerably change
when exposed to complex media [21,34]. The high ionic
strength combined with the adsorption of biomolecules
(mainly proteins) from biological environments can trigger
silica NPsaggregation and degradation (Fig. 1b) [20,35].
The aggregation process can impair silica NPs functionality
altering their interaction with cells and tissue components
[36]. The loss of colloidal stability has a profound impact
on NPsbiological behavior, which can reduce cellular
uptake, targeting, and drug delivery efciency while pro-
moting early silica NPs clearance from the bloodstream
[37]. Alongside aggregation, silica NPs also face degrada-
tion as a major challenge when exposed to biological media.
Degradation is a spontaneous and uncontrollable process for
bare silica NPs that can enhance their clearance from the
body, impacting their ultimate efciency as a drug-carrier
and bioaccumulation prole [20]. On the other hand, silica
NPsdegradation in biological media can be explored as a
benecial process, which can be used to control drug release
into a desired site of action. For this purpose, the doping of
silica NPsstructure with organosilane precursors, metal
ions, and oxides has been explored to modulate their
degradation rate under a given stimulus (e.g., light, pH, and
enzymatic attack) [2,27].
In addition to the aggregation and degradation processes,
the contact of silica NPs with biouids, such as plasma,
which is rich in proteins and other biomolecules, can also
lead to their adsorption onto NP surface, forming the so-
called protein corona (PC). The PC modies the chemical
identity of the NPssurface, affects their colloidal stability,
changes NPsinteraction with biological components, and,
consequently, alters their fate and ultimate performance
[38]. Thus, different functionalization approaches have been
explored to overcome protein adsorption, preserve colloidal
stability, and tune the degradation of silica NPs [39,40].
Among various methods [8,41,42], silanization using
alkoxysilanes (mainly trialkoxy derivatives; see Fig. 1c) is
the most used strategy for silica NPssurface modication.
Hydrophilic polymers such as polyethylene glycol (PEG)
Fig. 1 Schematic representation of athe silica NPs families with desirable properties for nanomedicine application, bthe possible outcomes of
silica NPs in biological environment, and cthe main strategy for functionalizing silica NPs surface using organo(trialkoxy)silanes
Journal of Sol-Gel Science and Technology
and zwitterionic compounds, as sulfobetaine moieties, for
instance, have been applied to enhance colloidal stability
and prevent unwanted silica-based NPs degradation
[40,43]. Multiple responses (protection against aggrega-
tion, targeting, responsiveness to media conditions, and
degradation rates tuning) often require the use of multiple
functional agents onto NPssurface, each one providing a
specic effect or acting as an anchor point for further
modications. For instance, targeting groups such as folic
acid, boronic acid derivates, and a wide range of antibodies
have been used to direct silica NPs toward tumor cells and
bacteria [4447].
While PC formation has been widely explored and
reviewed in the literature [48,49], the understanding of
colloidal stability and degradation rate in body and complex
uids has been considerably neglected and poorly docu-
mented. It is noteworthy to mention that the discrepancy of
the media complexity in aggregation and degradation tests
still precludes the complete comprehension of the system
behavior, which requires the evaluation of both processes
under the same medium stimuli. While researchers have
designed silica NPs to overcome aggregation in a complex
biological milieu (e.g., cell culture media and blood plasma)
[19,35,50], the degradation mechanism has been primarily
described in much simpler environments, such as simulated
body uids (SBF) [28,38,51]. We here discuss silica-based
nanostructuresfate within this scenario, focusing on the
aggregation and degradation of these NPs in different bio-
logically relevant media. Table 1summarizes the chemical
composition of some selected media mentioned along with
this review. We aim to provide the reader with a clear
picture of some of the main challenges silica-based nano-
medicines face nowadays in their translation to clinics and
potential solutions to overcome them.
2 Colloidal stability of silica nanoparticles in
biological uids
2.1 Basics on the colloidal stability of silica
nanoparticles
Aggregation affects the diffusion, gravitational settling, and
the available NPssurface. Thus, it obviously alters the
interactions of the particles with cellular and tissue com-
ponents, ultimately modifying biological outcomes such as
cellular uptake, immune-biocompatibility, and toxicity of
NPs [39,5254]. Hence, understanding and controlling
NPs colloidal stability are of paramount importance. Tra-
ditionally, NPs colloidal stability in aqueous electrolyte
solutions is interpreted and quantied using the
DerjaguinLandauVerweyOverbeek (DLVO) theory
[5557]. In its original formulation, DLVO theory models
colloidal interactions of likely-charged particles as the sum
of two opposing forces: attractive van der Waals (vdW) and
repulsive electrostatic (ES) forces (see Fig. 2a). vdW forces
are short-ranged interactions that result from the interaction
between permanent dipoles (Keesom Forces), permanent
and induced dipoles (Debye Forces), and induced dipoles
(London Forces). For colloids (diameters above ~0.5 nm),
vdW forces usually decay with the rst or second power of
the interparticle distance depending on the shape of the
interacting particles. A key parameter for quantifying/esti-
mating the vdW forces between NPs in a given medium is
the effective Hamaker constant, which depends on both the
material and solvent nature. Details on the determination of
Hamaker constants and selected values for inorganic
materials, in particular for the silica-based ones, can be
found in references [58] and [59].
In aqueous media, most NPs present surface charge due
to the dissociation/ionization of surface groups and the
adsorption of charged molecules/ions [55]. In the case of
silica surfaces, they are negatively charged due to the dis-
sociation of silanol groups as explained below. The surface
charge is balanced by counter-ions forming a cloud around
the particle. The classical electrical double layer (EDL)
model describes the ionic cloud as being constituted by two
regions: (a) the Stern layer, composed of approximately a
plane of adsorbed counter-ions, and (b) the diffuse layer
where the concentration of counter-ions gradually decreases
until reaching bulk electroneutrality (see Fig. 2b). The plane
separating these two layers is known as the Stern plane.
More important from an experimental point of view is the
shear plane (or slipping plane), at a distance where ions are
no longer bound to the particle. It is challenging to locate
this plane precisely, and it is usually thought to be inside the
diffuse layer but close to the Stern plane. The potential at
this plane is the measurable ζ-potential, an essential indi-
cator of colloidal stability (or the lack of it). A repulsive
force is generated when EDLs of two similar particles
overlap due to the osmotic pressure from the difference in
ion concentration between the mid-plane of the overlapping
EDLs (higher) and the media bulk (lower).
The ES repulsions are long-ranged and, for colloids,
experience an exponential decay with distance. Besides
the nature of particle surface groups, the most critical
(media) parameters inuencing ES repulsions are pH and
electrolyte concentration (or, most precisely, the ionic
strength). The pH determines the degree of ionization of
dissociating surface groups. For acidic groups (e.g.,
COOH, or silanol groups from silica), higher pH leads to
a higher ionization degree and, consequently, a higher
negative surface charge, resulting in a greater ES repul-
sion between similar particles. The opposite trend is
observed for surfaces containing basic groups (e.g.,
NH2), where lower pHs lead to higher positive surface
Journal of Sol-Gel Science and Technology
charges and greater ES repulsions. On the other hand,
electrolytes in solution produce a screening effect on the
ES forces, and at high ionic strength, the particle EDL is
compressed, promoting a reduction of the interparticle ES
repulsion that ultimately may lead to the loss of colloidal
stability. In addition, the greater the electrolytesvalency,
the higher the colloidal destabilization they generate
(Schulze-Hardy rule) [57,60].
Table 1 Chemical composition
of selected biologically relevant
media used for the testing of
colloidal stability and
degradation of silica NPs
PBS SBF FaSSGF FaSSIF SLF ALF DMEMa
10% FBS
RPMIa
10% FBS
pH 7.4 7.4 1.6 6.5 7.4 4.5 7.4 7.4
Na+b(mM) 157 142 34.2 134.3 143.5 208.9 155.3 124.3
K+(mM) 4.5 5 4 5.33 5.33
Ca2+(mM) 2.5 2.5 0.87 1.8 0.42
Mg2+(mM) 1.5 1 0.53 0.81 0.41
Fe3+(mM) 0.25
Cl-b (mM) 139.7 148.8 59.2 105.7 114 57.7 117.5 100.2
HCO3
(mM) 4.2 31 44.0 23.8
SO4
(mM) 0.5 0.5 0.27 0.81 0.41
HPO42(mM) 11.8 1 28.6 1 0.5 0.92 5.63
NO3
(mM) 0.74 0.85
TRIS (mM) 50
Acetate (mM) 7
Citrate (mM) 0.33 108.5
Tartrate (mM) 0.39
Lactate (mM) 0.76
Pyruvate (mM) 0.78
Glycine (mM) 0.79
Amino acidsc(mM) 10.6 6.44
Vitaminsc(mM) 0.15 0.24
Proteinc(g/L) 3.55.5d3.55.5d
Albumin (g/L) 1.73.5d1.73.5d
Glucose (mM) 25e11.1
Taurocholate (μM) 80 3000
Lecithin (μM) 20 750
Pepsin (g/L) 0.1
DPPC (mM) 0.27
The values informed where collected from references [180] (SBF, FaSSGF, FaSSIF, and ALF) [181] (SLF),
and [35] (DMEM +10% FBS and RPMI +10% FBS). In addition to FaSSGF and FaSSIF, simpler versions
(known as simulated gastric and intestinal uids, SGF and SIF, respectively) and versions emulated fed-state
are used to study the fate of silica NPs. Information on these media can be found in the references previously
mentioned
PBS phosphate buffer saline, SBF simulated body uid, FaSSGF fasted state simulated gastric uid, FaSSIF
fasted state simulated intestinal uid, SLF simulated lung uid, ALF articial lysosomal uid, DMEM
Dulbeccos Modied Eagle Medium, RPMI Roswell Park Memorial Institute medium, TRIS tris
(hydroxymethyl)aminomethane, DPPC dipalmitoylphosphatidylcholine
aThe versions of DMEM and RPMI supplemented with FBS are shown in this table. Non-supplemented
DMEM and RPMI are also used in the testing of NP colloidal stability and degradation. In these latter cases,
the composition is almost the same but without the constituents coming from FBS (e.g., proteins among
others)
bSince the pH in most of these media is adjusted using HCl or NaOH, the concentrations of Na+and Cl-may
slightly vary after pH adjustment
cRefers to the total content of vitamins, free amino acids, and proteins
dThese values depend on FBS batch and distributor
eThis concentration is used in the high-glucoseversion of DMEM. Moreover, depending on the version,
DMEM can also contain, pyruvate, HEPES buffer, among other constituents
Journal of Sol-Gel Science and Technology
In NPs having acidbase titratable surface groups, like
those exemplied in the previous paragraph, the surface
charge depends on media pH. Therefore, the point of zero
charge (PZC), the pH at which the colloid surface has zero
net charge, is of fundamental importance [61]. The PZC of a
NP is usually measured through potentiometric (acidbase)
titrations. On the other hand, the isoelectric point (IEP) is
the pH at which the particle ζ-potential is zero and can be
interpreted as the PZC at the slipping plane [62]. At the IEP,
particle diffusion is no longer inuenced by the presence of
an electric eld. Different from PZC, the IEP of a NP is
derived from electrokinetic methods. Although these terms
are frequently used interchangeably, strictly speaking, they
are equivalent only when specic ion adsorption is absent
[63]. Values of IEP and PZC of different solid hydroxides
and oxides (including silica) can be found in the seminal
and pioneering review by Parks in the 60s [64]. More
recently, Kosmulski has revisited and updated Parkswork
[65].
Although the original DLVO theory is a milestone for
colloidal physical chemistry, it is frequently unable to
explain the colloidal behavior of NPs with more complex
surfaces and surrounding media. For instance, decoration of
NPs surfaces with synthetic polymers may enhance NPs
resistance toward aggregation due to steric repulsion [66].
On the other hand, cell culture media and plasma present
complex ion composition (including divalent ones) and a
plethora of large (e.g., proteins) and small (e.g., carbohy-
drates, lipids, vitamins) molecules that may alter NP col-
loidal stability. Taking proteins as an example, their
adsorption onto NPs may lead to stabilization due to steric
or electrosteric repulsions or induce aggregation due to
surface neutralization or bridging interactions [35,67].
Then, various extensions of the DLVO theory have been
proposed to account for these other repulsive (e.g., steric,
hydration) and attractive (e.g., hydrophobic, depletion)
interactions [56,68,69]. As discussed below, the silica NPs
colloidal behavior, even in aqueous suspensions, cannot be
fully described by DLVO theory.
Bare silica NPs exhibit IEPs around pH ~24[65], and
at pH > IEP are negatively charged when in aqueous
media. The negative charges arise from the proton dis-
sociation of surface acidic silanol groups (Si-OH
SI-O-;seeFig.3a), and the pKa reported for this reaction
rangesfrom4to7.5(mostfrequentlyaround67) [70
73]. Differences in both IEP and pKa values are attributed
to the silica NP nature and the evaluation method [73].
Silanol groups at the silica surface can be classied
depending on the proximity to other silanols and the
number of -OH groups linked to the silicon atom (see Fig.
3a). Experimentally, from titrations at varying pH, two
silanol groups have been detected, with pKa values
around 4.55.5 (constituting 1519% of the total) and
8.59.9 (8185%) [74,75]. These observations are fur-
ther conrmed by ab initio molecular dynamic simula-
tions that show the presence of isolated silanol and
interacting silanol groups with pKa values ca.910 and
23, respectively [76]. In addition, another acidbase
equilibrium takes place on silanol groups involving the
protonation of neutral ones (Si-OH Si-OH2+;see
Fig. 3)[71,73], and the pKa for this reaction is estimated
between 1.8 and 1. Finally, the silica NPs surface also
presents siloxane groups (Si-O-Si, see Fig. 3b). Unlike
silanols, which can act as hydrogen bonding acceptors or
donors [77], siloxanes are weak hydrogen acceptors [78].
Thus, higher silanol contents induce higher silica surface
Fig. 2 aSchematic representation of interaction potential energy vs
surface distance proles according to the classic DLVO theory. Net-
DLVOstands for the overall potential energy (summation of elec-
trostatic repulsion and van der Waals attraction terms). bSchematic
representation of electrical double layer (EDL) model presenting the
different planes and potentials involved. The inset depicts the origin of
the (negative) charges in silica surfaces (silanol groups)
Journal of Sol-Gel Science and Technology
hydrophilicity. Contrarily, high siloxane content makes
silica surfaces more hydrophobic.
The colloidal behavior of bare silica NPs suspended in
aqueous electrolyte solutions is frequently described as
anomaloussince the regular DLVO theory cannot explain
the different regimes observed. Figure 4is adapted from
experimental published data and presents the inuence of
pH and sodium chloride on the silica NPs colloidal behavior
(in this case, the tendency to coagulate) [79,80]. At high
salt concentrations (Fig. 4a), silica NPs are stable at low pH
and coagulate at pH above 8. Also, the critical coagulation
concentration (CCC) in NaCl solution decreases with the
pH increasing in the range of 612 (Fig. 4b). Following
DLVO theory predictions, the opposite trend is expected in
the data shown in Fig. 4since the ionization of silica surface
is likely larger at higher pH and increases the NP stability
due to ES repulsion. Moreover, at low pH (around the IEP),
where no ES repulsion is present, silica NPs are very stable
and tolerate high salt concentrations without any aggrega-
tion signal.
Different hypotheses have been postulated to explain the
anomalous colloidal stability of silica sols considering
short-range repulsive interactions. The formation of a hairy
layerof polysilicilic acid chains (see Fig. 5a) at the silica/
water interface has been proposed as a source of stability in
silica NPs [8183]. In this case, a steric repulsion is gen-
erated when such layers overlap and, thus, counterbalance
the attractive vdW forces. On the other hand, the formation
of a hydration layer (see Fig. 5a) over silica surfaces (as
thick as 45 nm) leads to strong hydration repulsive forces
and has been considered as the possible mechanism behind
the stability of silica colloids [84,85]. Regardless of the
mechanism, surface silanols likely play a pivotal role in this
silica NPs anomalous stability since the reduction of silanol
surface density (by dehydroxylation and formation of sur-
face siloxanes or by surface silanization) [86,87] makes
silica NPs behave more following DLVO theory.
2.2 Colloidal stability of silica nanoparticles in
biologically relevant uids
Colloidal stability of silica-based NPs has been inten-
sively studied in a plethora of different biologically
relevant media such as cell culture media (e.g.,DMEM
and RPMI) with or without fetal bovine serum (FBS),
buffer solutions containing selected proteins (e.g.,bovine
serum albumin, BSA) and diluted human plasma or
serum. The composition of selected media can be found
Fig. 4 aTurbidity measurements (at λ=280 nm) of silica nanoparticle
(diameter ~54 nm) suspensions in the presence of 1.5 M NaCl at
varying pH. bCritical coagulation concentration of NaCl for silica
nanoparticles at different pHs (Ludox AM, diameter ~15 nm). In a,γ
refers to the extinction coefcient (roughly, higher γindicates a more
coagulated sample). In b, the CCC at different pHs was determined by
following the suspension Rayleigh ratio at 45° (at λ=546 nm) and it is
expressed in the graph as the logarithm of the salt concentration
required to coagulate the sample. In both cases, the values obtained in
basic condition (pH > 910) should be taken with caution since
degradation also takes place at this pH range, as discussed in the next
section. aand bwere adapted from Depasse et al. [79] and Allen et al.
[80], respectively, with permission from Elsevier [199]
Fig. 3 Schematic representation of aacidbase equilibrium of surface
silanols and bmolecular structures of different silanols and
siloxane groups
Journal of Sol-Gel Science and Technology
in Table 1presented in the previous section. The majority
of the studies have shown that bare non-porous silica NPs
tend to be relatively stable in non-supplemented DMEM
[37,52,8892]andRPMI[37,39,93], but are prone to
aggregate in the presence of FBS. In the latter case, large
NPs seem to be slightly more stable than the smaller ones.
For instance, bare silica NPs of 20 and 30 nm aggregated
in FBS-supplemented DMEM, while those of 110 nm
remained stable with only a slight increase in their
hydrodynamic diameters, which is attributed to PC for-
mation [91]. Similar results were observed when the
colloidal stability of 50 and 500 nm silica NPs was
investigated in RPMI with 10% FBS. Moreover, the silica
NPs colloidal stability has also been compromised in
supplemented media in rarely used cell culture media,
like EMEM [94]orHamsF-12[92].
In protein solutions, silica NPs exhibit different colloidal
behaviors depending on the protein identity. For instance,
severe aggregation is observed after silica NPs incubation
with lysozyme (pI =11.4; positively charged at physiolo-
gical pH) [67,95,96]. This phenomenon has been attributed
to NP surface neutralization after lysozyme adsorption onto
silica NPs surfaces, thus, eliminating ES repulsion and
bridging interactions between NPs. In PBS solution con-
taining albumin (e.g., BSA), silica NPs tend to aggregate
with time regardless of size [26,97]. However, colloidally
stable mixtures of silica NPs and BSA were obtained when
buffers of lower ionic strengths than PBS were used (e.g.,
phosphate buffer without added salt) [98]. Conversely,
different trends have been observed when silica NPs were
exposed to plasma, depending on NP size, plasma con-
centration, and incubation time [26,36,99,100]. For
example, silica NPs of 200 nm were considerably stable in
solutions containing plasma from 3 to 80% v/v [99]. These
NPs exhibited a decreased ζ-potential after plasma incuba-
tion that was attributed to an adsorbed protein layer. Simi-
larly, good colloidal stability of 100 nm silica NPs was
observed in high serum contents (above 30% v/v), although
aggregation was observed when these particles were incu-
bated in low serum contents (10% v/v) [100]. Aggregation
of 35, 120, and 140 nm silica NPs has also been reported in
plasma after 30 min of incubation, although they remained
stable in shorter times [36]. Rapid aggregation was found
when smaller silica NPs (~23 nm) are incubated in PBS
with 25% v/v human serum [26].
MSNs have been shown to be less colloidally stable than
their non-porous counterparts [93,101]. They tend to
aggregate in buffers under physiological ionic strength
conditions like PBS [102] and HEPES saline (HBS, con-
taining NaCl 150 mM) [103] or SBF [104,105]. Contrast-
ing results have been seen in FBS-supplemented media
(DMEM, RPMI, HBS), which is somehow similar to what
has been discussed above for non-porous silica NPs. Thus,
while some of the MSNs aggregated in the presence of FBS
[102], others remained stable [103], likely due to protein
adsorption. Table 2compiles the main results from selected
works devoted to the colloidal stability of non-
functionalized and functionalized silica NPs. The inu-
ence of surface functionalization on the colloidal behavior is
treated in the next subsection.
Several factors should be considered to rationalize the
colloidal behavior of NPs having bare silica surfaces in
complex media. As discussed before, the dissociation of
silanols provides silica surfaces with a negative charge and,
consequently, with an ES repulsion that operates as a source
of colloidal stability. In this regard, the relatively high ionic
strength presented in media like PBS, SBF, or cell culture
Fig. 5 aPossible mechanisms involved in the surface stabilization of
silica nanoparticles accounting for their anomalouscolloidal beha-
vior. bRepresentation of the most used strategies applied for enhan-
cing the colloidal stability of silica-based NPs. At the bottom of the
gure, the typical positive or negative groups used in the zwitterionic
strategies can be found
Journal of Sol-Gel Science and Technology
Table 2 Summary of selected works assessing colloidal stability of silica nanoparticles in biologically relevant uids
Nanoparticles Coating Fluids Techniques Observations Ref.
MSN (50 nm) Bare, PEG SBF TEM / DLS PEGylated NP more stable than bare ones [40]
MSN (80 nm) Bare, PEG PBS DLS PEGylated NPs colloidally stable [117]
MSN (4060 nm) Bare, PEG, PEG +hydrophobic silane (HS) PBS, DMEM (FBS 10%), SBF DLS Colloidal stability roughly ranks as NP-PEG +HS > NP-PEG bare >
bare NP
NP-PEG and NP-PEG +HS stable for days in different media
[104]
MSN (500 nm)
Non-porous 50 and 500 nm) Bare RPMI and RPMI +FBS 10% DLS / ζ-potential Colloidal stability roughly ranks as large non-porous NP > small non-porous
NP > MSN [93]
Non-porous (uorescent; 30 and
80 nm) Bare PBS, DMEM, and Ham F-12 (with and
without FBS 10%) DLS / DCS / TEM NPs colloidally stable in FBS-free media
Smaller NPs less stable than Larger NP in FBS supp. Media [92]
Non-porous (50 nm) Bare MEM with FBS 10% DLS Gradual increase of NP size. Stable dispersion up to 300min [89]
Non-porous (50 and 200 nm) Bare Human plasma in PBS (0-80%) DLS / DCS Protein adsorption promotes NP colloidal Stability [99]
Non-porous (uorescent 100 nm) Bare Human serum (0-100%) in PBS DCS NPs colloidally stable above 10% of human serum [100]
Non-porous (50 nm) Bare, Amine, Amino acids, PEG PBS, TRIS, DMEM, RPMI (with or without
FBS 10%) DLS / ζ-potential Colloidal stability roughly ranks as NP-PEG > bare > amines [37]
Non-porous (uorescent 58 nm) Bare, Amine-Sulfonate mixtures BSA solution, DMEM (with and without
FBS 10%) DLS Amine:sulfonate 50:50 mixture shows the worst colloidal stability [138]
Non-porous (100 nm) Bare, Amine, Phosphonate (Ph.), Amine +Ph,
SBS, PEG PBS, DMEM and RPMI (with and without
FBS 10%) DLS / DCS All NPs stable in PBS
NP-Amine +pH. not stable in FBS supp. DMEM or RPMI [140]
Non-porous (100 nm) Bare, Amine, Thiol, PVP PBS, DMEM and DMEM +FBS 10% DLS / TEM Colloidal stability roughly ranks as NP-PVP > bare NP> NP-thiol > NP-
Thiol [90]
Non-porous (20, 30, and 100 nm) Bare, Carboxilic acid Buffer A, DMEM and DMEM +FBS 10% DLS NPs are colloidally instable in DMEM +FBS 10% [91]
Non-porous (20 nm) Bare, Cysteine (Cys) BSA, lysozyme and human serum
(25%) in PBS DLS / Turbidimetry NP-Cys more colloidally stable than bare NP [26]
Non-porous (10220 nm) Bare, PEG BSA solution in PBS DLS / ζ-potential / TEM PEGylated NP more stable than bare NP
Explores inuence of incubation time, concentration, and curvature among
other topics
[97]
Non-porous (110 nm) Bare, PEG PBS AFFF Colloidal stability and dispersibility of NP-PEG > bare NP [113]
Non-porous (40 nm) Bare, PEG, Sulfobetaine (SBS) Lysozyme and FBS PBS solutions
NaCl 0.5 and 3 M DLS / Turbidimetry NP-SBS and NP-PEG stable in protein solutions
NP-SBS tolerates high salt concentration (3 M) [43,128]
Non-porous (100 nm) Bare, PEG, SBS Lysozyme and BSA in PB (different
NaCl conc.) DLS / ζ-potential / ITC NP-SBS and NP-PEG stable in protein solutions (bare only in BSA solution)
NP-SBS aggregates in lysozyme solution at a low salt concentration [98]
Non-porous (120140 nm) Bare, PEI-PEG PBS TEM / DLS NP-PEI-PEG colloidally stable [114]
Non-porous (60120 nm) PBSMA Lysozyme and BSA in PBS DLS Polyzwitterion promotes colloidal stability [182]
Non-porous (60220 nm) p-CBMA BSA and lysozyme in PBS DLS Coated NPs more colloidally stable than Bare NPs [146]
Non-porous (5080 nm) p-DMAEMA-co-CBMA (different
quaternization degrees, QD) BSA in PB lysozyme in CHES DLS / Turbidimetry Colloidal stability improves with increasing QD [145]
Non-porous (50500 nm)
MSN (500 nm) Bare RPMI and RPMI +FBS 10% DLS / ζ-potential Colloidal stability roughly ranks as large non-porous NP > small non-porous
NP > MSN [93]
MSN (50 nm) Bare, PEG SBF TEM / DLS PEGylated NP more stable than bare ones [40]
MSN (80 nm) Bare, PEG PBS DLS PEGylated NPs colloidally stable [117]
MSN (4060 nm) Bare, PEG, PEG +hydrophobic silane (HS) PBS, DMEM (FBS 10%), SBF DLS Colloidal stability roughly ranks as NP-PEG +HS > NP-PEG bare >
Bare NP
NP-PEG and NP-PEG +HS stable for days in different media
[104]
Nanoparticles and coatings: MSN mesoporous nanoparticles, PVP poly(vinylpyrrolidone), PEG poly(ethyleneglycol), p-CBMA poly(carboxybetaine), PEI poly(ethyleneimine), PBSMA poly
(sulfobetaine methacrylate), DMAEMA 2-(dimethylamino)ethyl methacrylate, CBMA carboxybetaine methacrylate
Media: PB phosphate buffer, PBS phosphate buffer saline, MEM minimum essential medium, FBS fetal bovine serum, SBF simulated body uid, DMEM Dulbeccos Modied Eagle Medium,
RPMI Rose Park Memorial Institute medium, CHES N-Cyclohexyl-2-aminoethanesulfonic acid buffer, MEM minimum essential medium
Techniques: DLS dynamic light scattering, DCS differential centrifugal sedimentation, TEM transmission electron microscopy, AFFF asymmetric eld ow fractionation
Journal of Sol-Gel Science and Technology
media (DMEM, RPMI) or biological uids like plasma,
screens this ES repulsion, thus decreasing colloidal stability
[35,37]. On the other hand, adsorption of biomolecules, in
particular proteins, from complete cell culture media (sup-
plemented with FBS), diluted mammals plasma/sera, or
even single protein buffered solutions alters NPscolloidal
behavior considerably. Protein adsorption onto silica NPs
usually decreases the NP ζ-potential, which tends to reach
the isolated protein (or proteome) ζ-potential value at
saturation, hence compromising colloidal stability
[91,93,97]. However, protein adsorption can be benecial
for preventing NP aggregation owing to steric hindrance
depending on protein nature [35,97,99]. For this latter to
happen, it seems to be essential that a homogeneous protein
coating is formed over the NP. Otherwise, if the adsorbed
proteins are not uniformly distributed over the NP surface
(i.e., forming patches), bridging interaction may arise and,
combined with the already mentioned decrease in NP ζ-
potential, can promote aggregation [35,67,97]. Further-
more, the temporal evolution of the PC composition, when
NPs are dispersed in complex media containing several
different proteins (Vromans effect) [106108], implies that
NP surface properties also evolve with time accordingly.
Then, the colloidal properties of the NPs might evidence
changes during the incubation time [36,49,100].
2.3 Surface modication as a strategy to enhance
silica nanoparticles colloidal stability
Silica NP surface functionalization is used to anchor tar-
geting moieties and biomolecules, prevent fouling and
aggregation, and regulate degradation, among other pur-
poses [8,109,110]. Every surface modication, regardless
of its ultimate goal, impacts NPscolloidal stability. For
instance, silica NPs functionalization with amine groups,
using 3-aminopropyl(triethoxy)silane (APTES) or 3-ami-
nopropyl(trimethoxy)silane (APTMS), is frequently used as
arst step for conjugating biomolecules in grafting from
approaches [109]. However, functionalization with APTES
or APTMS usually triggers silica NPs aggregation due to
positive surface charges (-NH3+) that counterbalance the
negative charges of pristine silica (Si-O), and reduces the
absolute value of NP ζ-potential and ES repulsion [37].
Alternatively, Bagwe et al. proposed one of the rst stra-
tegies for stabilizing amino-containing silica NPs through
simultaneous NP functionalization using APTES and 3-
(trihydroxysilyl)-propylmethylphosphonate [111].
Surface modication using PEG (see Fig. 5b) is perhaps
the most traditional strategy to maintain NPs colloidal sta-
bility on biological uids [112] and has been widely applied
to different silica-based nanomaterials such as non-porous
[43,98,113115], mesoporous [40,116], hollow-
mesoporous [117] and silica-coated NPs [118,119]. PEG
chains are strongly hydrated due to their hydrophilic nature
derived from the formation of hydrogen bonds between
water and PEG ether groups [120,121]. At appropriate
grafting densities, a layer of PEG provides steric repulsive
forces that prevent fouling (e.g., protein adsorption) and
aggregation in complex media [66,122]. Despite the
widespread use of PEGylation, PEG coatings present some
drawbacks like not preserving NP colloidal stability at high
ionic strengths and experiencing oxidative degradation in
environments containing transition metals and oxygen
[43,120,123].
Besides PEG, other neutral and hydrophilic polymers
have been used to functionalize NPs exposedsilica surfaces
to prevent aggregation. For instance, surface modication
with hyperbranched polyglycerol [124] and copolymer-
containing poly(N-isopropylacrylamide) (PNIPAM) [125]
have proven to improve the colloidal stability of silica-
coated NPs (gold, iron oxide, and quantum dots) and Stöber
silica NPs, respectively. In the latter case, the colloidal
stability also depends on temperature due to the thermo-
responsiveness of PNIPAM [126].
On the other hand, surface functionalization based on
zwitterions (neutral compounds having formal electrical
charges of opposite sign) [127] has gained attention in the
last few years as a powerful strategy for preventing fouling
and aggregation of NPs in biological media [121,122,128].
The use of zwitterions, mainly for antifouling purposes, was
inspired by the exposed headgroups of mammalian cell
membranes (composed primarily by phosphatidylcholine
zwitterionic molecule) that were demonstrated to prevent
biofouling on surfaces decorated with polymers [129131].
Zwitterions electrostatically bind and structure water
molecules more efciently than PEG and other hydrophilic
polymers. Hence, a NP surface coated with these com-
pounds must overcome a high dehydration energy barrier
for interacting with other nanometric objects nearby (e.g.,
other NPs, proteins) [41]. Moreover, the binding of mac-
romolecules containing distributed charges on their sur-
faces, like proteins, is relatively hindered for surfaces
exhibiting simultaneously negative and positive charges.
Small zwitterionic moieties (see Fig. 5b), such as sulfo-
betaine [43,98,128,132134], carboxybetaine [135,136],
and phosphorylcholine, have been extensively employed to
maintain the colloidal stability of silica-based nanomater-
ials. Moreover, amino acids like cysteine or arginine have
also been used to produce zwitterionic coatings that effec-
tively avoid silica NPs aggregation in complex media
[26,137]. Alternatively, the NP surface can be simulta-
neously tailored with moieties bearing positive (e.g.,NH
3+,
-NH(CH3)2+) and negative (e.g.,SO
3
, -PO3
, -PCH3O2
)
groups and give rise to pseudo-zwitterionicsurfaces that
result in good colloidal stabilization [138140]. In this case,
the ratio between positive and negative moieties must be
Journal of Sol-Gel Science and Technology
carefully engineered to optimize surface passivation against
aggregation and protein adsorption.
Polymeric zwitterions (see Fig. 5b) containing pendant
sulfobetaines [141143], carboxybetaine [144146], or
phosphorylcholine [147,148], among others moieties [149],
have also been applied for coating and preserving silica NPs
colloidal stability in biological environments. In addition to
the already mentioned mechanisms accounting for small
zwitterionssignicant inuence on NP interactions and
colloidal stability, polymeric zwitterions also take advan-
tage of the steric effects related to their larger volume [122].
It is also worth mentioning that NPs purication proce-
dure and storage (liquid suspension vs.dry powder) before
the transference to the testing media can profoundly inu-
ence the colloidal behavior. In particular, silica NP dried
powders are difcult to redisperse due to the formation of
interparticle siloxane bonds that result in irreversible
aggregates [150,151]. In this case, any further evaluation of
the NP colloidal behavior will be heavily biased by this
original aggregation. For instance, this situation is fre-
quently observed in MSNs produced following procedures
implying calcination for organic template removal or
pyrogenic non-porous silica NPs, but not restricted to these
examples [104]. Despite this, silica NP powders with good
redispersability in relevant media and showing reduced to
negligible aggregation attributed to the dry state can be
produced by freeze-drying in the presence of proper pro-
tectants such as simple carbohydrates [151] or albumin
[22,50].
3 Degradation of silica nanoparticles in vitro
and in vivo
3.1 Basics on silica nanoparticles degradation
The complete silica-based nanomedicines clearance from
the body after their therapeutic and diagnostic effects must
be guaranteed before their translation from bench to clinics
[2,152]. Some silica materials have been considered safe by
the FDA, mainly due to the low toxicity of silicic acid (the
main product of silica degradation) and the presence of
silicates on several body tissues [27]. However, it has been
reported that silica NPs can accumulate in organs such as
the liver and spleen after administration leading to silicosis
[20,27]. Thus, tuning silica NPs degradation is of utmost
importance since this process can improve the excretion of
ultrasmall silica fragments by the urinary and intestinal tract
[20,27].
Silica surface is naturally prone to nucleophilic attack in
aqueous solutions, causing the erosion of their external
structure and inducing premature damage, which can be
favored or hampered depending on the groups disposed onto
its surface [105,153]. On the other hand, the degradation of
silica can be regulated by doping its structure using
degradable sites, such as di- and tetrasulde species, Schiff
base, oxamide-phenylene groups, and metallic ions [3,154
157]. These sites can undergo degradation when triggered by
a specic stimulus, such as pH [154], enzymatic attack
[155], and redox reactions [3].
In aqueous media, the dissolution of silica framework
results from sequential surface hydration and hydrolysis
(see Fig. 6a). After water molecules adsorb onto siloxane
(Si-O-Si) and silanol (Si-OH) species, an electropositive
silicon atom is attacked by a hydroxyl group (OH-) from the
medium, resulting in an unstable pentavalent intermediate
complex. Then, this structure is further attacked by another
hydroxyl group, opening the silica structure by converting a
siloxane into a silanol group [20]. As a consequence, silicic
acid (Si(OH)4) is released. Fortunately, silicic acid and other
polysilicic acids (which are also found after hydrolysis of
silica surfaces) [158] are nontoxic [159], able to diffuse
through blood and lymphatic vessels and, eventually, are
eliminated in the urine [160]. The degradation kinetics of
silica surfaces has been modeled through Eq. 1, taking into
account the surface concentration of three silanol species:
protonated (Si-OH2+), neutral (Si-OH), and deprotonated
(Si-O-)[28].
Rdiss ¼kHþ
ðÞ
SiOHþ
2

mþkH2OðÞ
SiOH½þkOH
ðÞ
SiO
½
p
ð1Þ
In the equation, Rdiss is the rate of dissolution, [Si-OH2+],
[Si-OH], [Si-O] are the surface concentration of the three
mentioned species, k(H+),k(H2O), and k(OH)refer to dis-
solution rate constants, and mand pstand for reaction
orders.
The degradation rate of silica NPs strongly depends on
the characteristics of both the media (pH and composition)
and the NPs (composition, porosity, colloidal stability, and
condensation rate) [27]. The pH effect on silica dissolution
has been widely reported in the last few decades to under-
stand the chemical stability, catalytic, and adsorption
activity of silica-based materials in different aqueous solu-
tions [161163]. It denes the concentration of protonated
and deprotonated species onto silicas surface and, hence,
deeply impacts the silica degradation rate. Ideally, the IEP
of bare colloidal silica ranges from pH 24 in water, and the
concentrations of protonated and deprotonated are more
prominent in pH < 1 and pH > 5, respectively. Thus, the
hydrolysis and degradation of the silica surface would be
more likely to happen in pH conditions below or above
silicas IEP, since the energy barrier for this process in both
protonated and deprotonated sites is lower [164,165].
However, in a more complex scenario, the silica NPsIEP is
Journal of Sol-Gel Science and Technology
shifted toward the IEP of the adsorbed species, which can
be impuritiesfrom the medium (e.g., ions and proteins) or
functional groups attached to the particle surface, changing
the pH in which the degradation of its structure would occur
[61,64]. Therefore, the IEPs determination under these
conditions would be a challenge once the high concentra-
tion of ionic groups and their interactions, for instance, can
lead to experimental errors, such as imprecision in the pH
measurement [166,167].
3.2 Inuence of media on the degradation of silica
nanoparticles
In the last few years, the increasing interest in silica NPs in
nanomedicine has propelled studies about their degradation
behavior in SBF under different pHs and compositions (a
compilation of studies is found in Table 3)[20,27]. For
instance, it has been demonstrated that silica-based NPs
with different structural features present higher degradation
rates in simulated lysosomal uids (pH 4.5), simulated
intestinal uids (pH 6.5), and SBF (pH 7.4) than in simu-
lated gastric one (pH 1.2) [20].
Mono- and bivalent ions, organic salts, amino acids, and
proteins are examples of components in biological envir-
onments that can inuence silica NPs degradation [168].
Table 1summarizes the most popular different media
compositions used along with the degradation studies.
These media can potentially enhance silica degradation due
to their ionic rich content since the adsorption of cations
onto silica surface can weaken the siloxane bonds (Fig. 6b),
while anion adsorption can enhance silanol deprotonation. It
stimulates siloxane hydrolysis and further silica fragments
leaching from the particle surface [20,51]. As reported
previously, simulated lung uids (SLF) and SBF promote a
higher degradation of MSNs than simulated gastric uid
(Table 3)[51]. This result reects the impact of differences
in the composition and pH of the used media (Table 1).
Salts of organic acids and proteins, which are particularly
important in complex physiologic uids, have signicantly
impacted silica degradation. Organic acid salts, mainly di-
and tri-carboxylates, have been described as the main factor
of the higher silica degradation rates observed in SLF
[28,51]. This behavior has been attributed to the formation
of complexes between carboxylates and silica surface by
nucleophilic attack of the former toward electropositive
silicon atoms of the latter. Alternatively, the nucleophilic
attack of the carboxylic acid from proteins to silicon atoms
is facilitated by the interaction of aminated protein moieties
with the Si-Ogroups on silica surfaces (Fig. 6b) [28]. For
instance, MSNs degradation enhancement has been reported
in DMEM supplemented with FBS (10%). Signs of degra-
dation have been detected for 87 nm MSNs after 2 h of
incubation and, after 60 days, 54% of MSNs had been
degraded in FBS-enriched DMEM. Conversely, no sig-
nicant degradation has been observed in FBS-free DMEM
[38]. Complementary, silica-based NP degradation tests in
the intracellular environment have shown, after 72 h of
incubation, that only ~3.8 and ~0.3% of MSN (100 nm) and
Fig. 6 Schematic representation of the silica degradation mechanisms in awater (at IEP) and bin biological media, considering the inuence of
different factorssurface functionalization and adsorption of ion and proteins
Journal of Sol-Gel Science and Technology
non-porous silica NP (100 nm) structures degraded,
respectively [20]. The degradation mechanism in intracel-
lular environments and throughout the in vivo tests is still to
be revealed. The results and conditions of degradation
experiments reported in several studies in vitro and in vivo
are presented in Tables 4and 5.
3.3 Inuence of silica nanoparticles structure on
their degradation
Silica NPs structural characteristics are as important as the
media along their degradation process [2,27,169]. For
instance, while non-porous particles do not signicantly
degrade in different media compositions, porous particles
are highly susceptible to environmental conditions [38].
Porous particles allow better diffusion of the medium
through its structure and, hence, provide enhanced interac-
tion of silanol and siloxane groups with ions, proteins, and
other media [105,170]. Studies have shown that particles
with larger pore sizes degrade faster than the ones with
smaller pores and similar surface area [171,172]. Alongside
porosity, the condensation degree also plays a signicant
role in silica NPs degradation rate [27,105,171]. A highly
condensed silica particle, which is usually obtained by the
sintering process, presents a majority of Q4sites ((SiO)4-Si)
and a minority of Q3species ((SiO)3-Si-OH) [27,51,105].
Conversely, non-calcined silica NPs show higher Q3con-
tents while the overall Q4amount is minor. The porosity
also impacts on silica NPs condensation degree, for
instance, MSNs present a lower condensation degree (Q4/Q3
Table 3 Degradation conditions and outcomes of silica-based nanoparticles in biological uids
Nanoparticles Doping or/and coating Fluids Techniques Tested time Observations Ref.
HSN Disulde doping PBS +GSH TEM 48 h Full degradation (48 h) [183]
MSN PEG coating
-Cl coating
-NH2coating
SBF BET / TEM 28 d PEGpartial degradation (28 d)
Clfull degradation (28 d)
NH2full degradation (28 d)
[105]
MSN SGF/SIF/SBF +GSH ICP-MS /
TEM
28 d SGF6% degraded (28 d)
SIF100% degraded (28 d)
SBF100% degraded (6 h)
[20]
MSN Disulde doping SGF/SIF/SBF +GSH ICP-MS /
TEM
28 d SGF19% degraded (28 d)
SIF55% degraded (28 d)
SBF13% degraded (6 h)
[20]
MSN SLF/SBF/SGF/PBS ICP-MS 8 h SLFfull degradation (6 h)
SBFfull degradation (6 h)
SGFpartial degradation (14 d)
PBSfull degradation (6 h)
[51]
MSN SBF TEM 96 h Full degradation (24 h) [171]
MSN PEG and FA coating PBS +GSH ICP-OES /
TEM
72 h pH 590% degraded (14 d)
pH 7.415% degraded (14 d)
[184]
MSN Lysine-silane doping DI Water +trypsin TEM 72 h Full degradation (72 h) [185]
Organosilica Oxamide-phenylene doping PBS +trypsin TEM 48 h Partial degradation (24 h)
Full degradation (48 h)
[155]
Organosilica Ethylene doping PBS +ME TEM / DLS 48 h Full degradation (48 h) [186]
Organosilica Oxamidedoping PBS +trypsin TEM / FTIR 48 h Full degradation (48 h) [187]
Organosilica Disuldedoping PBS +ME TEM 48 h Full degradation (48 h) [188]
Stöber SGF/SIF/SBF ICP-MS /
TEM
28 d SGF8% degraded (28 d)
SIF46% degraded (28 d)
SBF64% degraded (28 d)
[20]
Stöber SLF/SBF/SGF/PBS ICP-MS 8 h SLFfull degradation (8 d)
SBFfull degradation (10 d)
SGFno degradation (14 d)
PBSfull degradation (14 d)
[51]
Stöber Disulde doping PBS +GSH TEM 5 d Full degradation (5 d) [189]
Mn-MSN Mn doping PBS +GSH ICP-MS /
TEM
15 d Full degradation (15 d) [190]
Mn-MSN Mn doping SBF +GSH ICP-OES 12 h Full degradation (12 h) [191]
Zr-MSN Zr doping PBS TEM 24 h Slight degradation (24 h) [192]
ME mercaptoethanol, DI distilled, FA folic acid, GSH gluthatione
Journal of Sol-Gel Science and Technology
around 22.5) compared to Stöber non-porous particles (Q4/
Q3around 510). Accordingly, only 30% of a calcined and
non-porous silica NP degrade after 15 days, while a non-
calcined MSN can fully degrade in less than 48 h [27].
3.4 Strategies to tune silica nanoparticle
degradation
The silica NPs degradation rate is highly dependent on how
extensively the media can access the silica surface
[105,170]. Then, good colloidal stability allows proper
interaction between media and individual particles, resulting
in faster particle degradation [27,172]. In contrast, the
media componentsaccess to the inner region of aggregated
particles is limited and leads to lower degradation rates
[172]. As aforementioned, silica NPs functionalization can
prevent particles aggregation, improve colloidal stability
and enhance surface exposure. However, depending on the
surface modication, this approach can also delay the par-
ticles degradation process[105,153]. The functionalization
of porous silica NPs with -phenyl, -NH2, and -Cl groups has
been demonstrated to lead to fast degradation of the particle
(<24 h), which is attributed to the low condensation degree
of the silica structure and insufcient protection provided by
these groups (Table 4). In contrast, when PEG coats silica
NPs, the sample has shown enhanced structural stability in
biologically relevant media and presented signs of degra-
dation only after 4 days of incubation in SBF [105]. In
addition, several studies have shown that the degradation
behavior of MSNs is modied after their surface functio-
nalization. While non-functionalized mesoporous particles
are degraded by a combination of surface erosion and pore
degradation [105,173175], particles functionalized with
polymers start degrading from the interior of the pores (see
Fig. 7)[38,169,176].
In the nanomedicine scenario, silica NPs degradation
before reaching the target is undesirable since premature
drug leakage leads to efciency loss. Unfortunately, pure
silica NPs degradation cannot be designed to be triggered
by any given condition [152]. For this reason, scientists
were driven to introduce responsive sites into the silica
framework to control the degradation rate of silica-based
NPs (Fig. 8)[27]. These sites are usually composed of
organic moieties that respond to a specic exterior stimulus
and, when triggered, lead to NP structure fragmentation
[2,27,152]. Then, a wide range of organosilane groups has
been used to trigger silica NPs degradation under specic
conditions. Precursors containing di- and tetrasulde, amino
Table 5 Degradation conditions
and outcomes of silica-based
nanoparticles in ex vivo and
in vivo
Nanoparticles Doping or/and
functionalization
Particle
concentration
Tested time Observations Ref.
MSN 10200 mg/kg 10 d Full excretion (4 d) [196]
89Zr-MSN 89Zr labeling 100 uL of 1 mg/mL 48 h Partial excretion
(48 h)
[197]
MSN 25 mg/kg 7 d 53% excreted (7 d) [20]
MSN disulde doping 25 mg/kg 7 d 39% excreted (7 d) [20]
MSN 20 mg/kg 4 d Full excretion (4 d) [198]
Stöber 25 mg/kg 7 d 27% excreted (7 d) [20]
Table 4 Degradation conditions and outcomes of silica-based nanoparticles in cell culture media and intracellular environment
Nanoparticles Doping or/and coating Culture media/cell line Technique Tested time Observations Ref.
MSN Lonza Mesenchymal Stem Cell
Growth Medium
TEM/ICP 21 d Full degradation (18 d) [193]
MSN Containing AuNPs IMDM or RPMI +FBS TEM 72 h Full degradation (4 h) [194]
MSN IMDM or RPMI +FBS TEM 72 h No degradation (72 h) [194]
MSN PEG and FA coating HepG2 cells TEM 72 h Full degradation (72 h) [184]
MSN DMEM TEM 60 d No degradation (60 d) [38]
MSN DMEM +FBS TEM 60 d Partial degradation
(60 d)
[38]
Organosilica Disulde doping DMEM +FBS TEM 48 h Full degradation (48 h) [195]
Organosilica Disulde doping RAW 264.7 macrophages ICP-MS 72 h Slight degradation
(72 h)
[20]
Stöber RAW 264.7 macrophages ICP-MS 72 h Slight degradation
(72 h)
[20]
Journal of Sol-Gel Science and Technology
acids, and Schiff base can be degraded by redox reactions,
enzymatic attacks, and pH variations, respectively. Thus,
silica NPs containing redox responsive groups, for example,
can be triggered by reduced glutathione (GSH) in the cell
cytosol, making these engineered systems a promising
alternative to promote drug release into targeted cells
[3,20]. Therefore, the combination of surface functionali-
zation (using polymers or zwitterionic groups) and
responsive organosilanes is a promising strategy to suc-
cessfully control silica-based NPs while guaranteeing the
NP degradation in a pre-dened environment.
Even though organosilane groups on silica NPs struc-
tures can be employed to tune their controlled degradation,
several studies report that their use delays the complete
clearance from the body [20,177]. This phenomenon is
likely explained through the hydrophobicity induced by the
organosilanes structure after the NP collapse and hinders
further degradation of the silica moiety to small fragments
(<5 nm). This trend was reported in the literature, where
only 1.2% of the structure of MSNs containing disulde
organosilane was degraded after 72 h in an intracellular
environment, while its pure silica counterpart degraded
3.8% in the same period [20]. Conversely, the same study
showed that silica fragments concentration (<5.5 nm) in
mice kidneys was higher for MSN containing disulde
Fig. 8 Timeline of the complete degradation in aqueous media of different silica nanoparticles based on their porosity and composition. This gure
was adapted from Croissant et al. [27] and reproduced with permission from Wiley
Fig. 7 Schematic representation of the degradation course, in biolo-
gical media, of mesoporous (top), functionalized mesoporous (middle),
and non-porous (bottom) silica nanoparticles
Journal of Sol-Gel Science and Technology
groups (ca. 26%) than the amount detected for pure silica
MSN and non-porous silica NP, which was around 10%
[20]. This disparity clearly indicates the relevant impact of
biological mechanisms on silica-based NP degradation and,
yet, this process remains unclear.
In an attempt to enhance silica-based NP biocompatibility,
researchers have been also doping silica NPs with metal or
metal oxides to either accelerate (iron, manganese, and cal-
cium) or delay (zirconia and aluminum) silica NPs degrada-
tion. The advantages of silica doping with iron, calcium, and
manganeseincludeafasterNPsclearanceduetothewide
range of interactions of these ions with biological species
promotinganefcient degradation. In addition, several studies
have shown that these particles can also be responsive to
media stimuli. For instance, Mn-doped MSNs are triggered by
mild acidic and reducing environments [156,178,179], fol-
lowed by Ca-doped MSN that also degrades under acidic
conditions [157]. Conversely, in zirconia and aluminum-
dopped systems, the degradation rate is lagged due to the
higher chemical stability of the framework [27]. Some
examples of degradation conditions and outcomes for dopped
silica NP are also presented in Tables 35.
4 Conclusions and perspectives
Efforts have been made to understand the bio-related aggre-
gation and degradability of silica NPs in biological uids.
Consequently, this knowledge has built up evidence of the
role of different parameters in this complex puzzle. Silica NPs
morphology, porosity, and density have an evident impact on
their colloidal stability and degradation rates, which are
usually enhanced by environmental features (composition,
pH, and ionic strength). Silica NPs surface functionalization
with protectant agents has successfully prevented these
undesirable processes in SBF and cell culture media. Never-
theless, the reported studies are still performed in highly
controlled media conditions, where the NP concentration is
pre-dened, and the environment composition is well known.
Thus, the direct translation of these systems to medical
treatments is not trivial. These studies are limited to experi-
mental conditions far from grasping the real response when
applied in vivo. Therefore, further advances in this eld still
rely on the comprehension of silica NPs behavior in whole
blood and living tissues. Then, it is imperative to correctly
answer how silica NPs engage with each major blood com-
ponent, what path is followed after cell uptake, and how they
achieve complete clearance from the body. The advances
made in the last decade represent a solid ground to boost the
following steps, but we still have a long path ahead. We
envisage that breakthrough understanding of the silica NPs
behavior in the biological environment will be only possible
when high-resolution techniques (such as synchrotron and
electron microscopy techniques) are fully designed and opti-
mized to probe colloidal stability in complex media, NP-cell
interaction, and NP degradation.
Acknowledgements The authors also acknowledge the nancial sup-
port of the Fundação de Amparo à Pesquisa do Estado de São Paulo
(processes 2018/00763-8, 2019/24894-7, 2016/16905-0, 2020/00767-
3, and 2015/25406-5). ASP and MBC acknowledge UNLP and
CONICET for their support. ASP and MC staff members of CONICET.
Compliance with ethical standards
Conict of interest The authors declare no competing interests.
Publishers note Springer Nature remains neutral with regard to
jurisdictional claims in published maps and institutional afliations.
References
1. Huang P, Chen Y, Lin H et al. (2017) Molecularly organic/
inorganic hybrid hollow mesoporous organosilica nanocapsules
with tumor-specic biodegradability and enhanced chemother-
apeutic functionality. Biomaterials 125:2337. https://doi.org/10.
1016/j.biomaterials.2017.02.018
2. Croissant JG, Fatieiev Y, Almalik A, Khashab NM (2018)
Mesoporous silica and organosilica nanoparticles: physical
chemistry, biosafety, delivery strategies, and biomedical appli-
cations. Adv Health Mater 7:1700831. https://doi.org/10.1002/a
dhm.201700831
3. Maggini L, Cabrera I, Ruiz-Carretero A et al. (2016) Breakable
mesoporous silica nanoparticles for targeted drug delivery.
Nanoscale 8:72407247. https://doi.org/10.1039/C5NR09112H
4. Tayama M, Inose T, Yamauchi N et al. (2019) Fabrication and
dual-modal imaging properties of quantum dot/silica core-shell
particles with immobilized single-nanometer-sized gold nano-
particles. Colloids Surf A Physicochem Eng Asp 574:162170.
https://doi.org/10.1016/j.colsurfa.2019.04.055
5. Lei Q, Guo J, Noureddine A et al. (2020) Solgelbased advanced
porous silica materials for biomedical applications. Adv Funct Mater
30:1909539. https://doi.org/10.1002/adfm.201909539
6. Stöber W, Fink A, Bohn E (1968) Controlled growth of mono-
disperse silica spheres in the micron size range. J Colloid Interface
Sci 26:6269. https://doi.org/10.1016/0021-9797(68)90272-5
7. Bogush GH, Tracy MA, Zukoski CF (1988) Preparation of
monodisperse silica particles: control of size and mass fraction. J
Non Cryst Solids 104:95106. https://doi.org/10.1016/0022-
3093(88)90187-1
8. Liberman A, Mendez N, Trogler WC, Kummel AC (2014)
Synthesis and surface functionalization of silica nanoparticles for
nanomedicine. Surf Sci Rep 69:132158. https://doi.org/10.
1016/j.surfrep.2014.07.001
9. Beck JS, Vartuli JC, Roth WJ et al. (1992) A new family of
mesoporous molecular sieves prepared with liquid crystal tem-
plates. J Am Chem Soc 114:1083410843. https://doi.org/10.
1021/ja00053a020
10. Kresge CT, Leonowicz ME, Roth WJ et al. (1992) Ordered
mesoporous molecular sieves synthesized by a liquid-crystal
template mechanism. Nature 359:710712. https://doi.org/10.
1038/359710a0
11. Tarn D, Ashley CE, Xue M et al. (2013) Mesoporous silica
nanoparticle nanocarriers: biofunctionality and biocompatibility.
Acc Chem Res 46:792801. https://doi.org/10.1021/ar3000986
Journal of Sol-Gel Science and Technology
12. Tissot I, Reymond JP, Lefebvre F, Bourgeat-Lami E (2002)
SiOH-functionalized polystyrene latexes. A step toward the
synthesis of hollow silica nanoparticles. Chem Mater
14:13251331. https://doi.org/10.1021/cm0112441
13. Chen JF, Ding HM, Wang JX, Shao L (2004) Preparation and
characterization of porous hollow silica nanoparticles for drug
delivery application. Biomaterials 25:723727. https://doi.org/
10.1016/S0142-9612(03)00566-0
14. Iler RK (1979) The chemistry of silica: solubility, polymeriza-
tion, colloid and surface properties, and biochemistry. John
Wiley & Sons
15. Brinker CJ, Scherer GW (1990) Sol-gel science. Elsevier
16. Bergna HE, Roberts WO (2006) Colloidal silica: fundamentals
and applications. surfactant science series. CRC Press, Taylor
and Francis Group, LLC
17. Zou Y, Huang B, Cao L et al. (2021) Tailored mesoporous
inorganic biomaterials: assembly, functionalization, and drug
delivery engineering. Adv Mater 33:2005215. https://doi.org/10.
1002/adma.202005215
18. Li Z, Mu Y, Peng C et al. (2021) Understanding the mechanisms
of silica nanoparticles for nanomedicine. Wiley Interdiscip Rev
Nanomed Nanobiotechnol 13:123. https://doi.org/10.1002/wna
n.1658
19. Jiang S, Prozeller D, Pereira J et al. (2020) Controlling protein
interactions in blood for effective liver immunosuppressive
therapy by silica nanocapsules. Nanoscale 12:26262637.
https://doi.org/10.1039/c9nr09879h
20. Hadipour Moghaddam SP, Mohammadpour R, Ghandehari H
(2019) In vitro and in vivo evaluation of degradation, toxicity,
biodistribution, and clearance of silica nanoparticles as a function
of size, porosity, density, and composition. J Control Release
311312:115. https://doi.org/10.1016/j.jconrel.2019.08.028
21. Zhu X, Vo C, Taylor M, Smith BR (2019) Non-spherical micro-
and nanoparticles in nanomedicine. Mater Horiz 6:10941121.
https://doi.org/10.1039/C8MH01527A
22. Picco AS, Mondo GB, Ferreira LF et al. (2021) Protein corona
meets freeze-drying: overcoming the challenges of colloidal
stability, toxicity, and opsonin adsorption. Nanoscale
13:753762. https://doi.org/10.1039/d0nr06040b
23. Capeletti LB, Oliveira JFA, Loiola LMD et al. (2019) Gram
negative bacteria targeting mediated by carbohydratecarbohydrate
interactions induced by surfacemodied nanoparticles. Adv Funct
Mater 29:1904216. https://doi.org/10.1002/adfm.201904216
24. Lee JE, Lee N, Kim T et al. (2011) Multifunctional meso-
porous silica nanocomposite nanoparticles for theranostic
applications. Acc Chem Res 44:893902. https://doi.org/10.
1021/ar2000259
25. Guerrero-Martínez A, Pérez-Juste J, Liz-Marzán LM (2010)
Recent progress on silica coating of nanoparticles and related
nanomaterials. Adv Mater 22:11821195. https://doi.org/10.
1002/adma.200901263
26. Rosen JE, Gu FX (2011) Surface functionalization of silica
nanoparticles with cysteine: a low-fouling zwitterionic surface.
Langmuir 27:1050710513. https://doi.org/10.1021/la201940r
27. Croissant JG, Fatieiev Y, Khashab NM (2017) Degradability and
clearance of silicon, organosilica, silsesquioxane, silica mixed
oxide, and mesoporous silica nanoparticles. Adv Mater 29:151.
https://doi.org/10.1002/adma.201604634
28. Ehrlich H, Demadis KD, Pokrovsky OS, Koutsoukos PG (2010)
Modern views on desilicication: biosilica and abiotic silica
dissolution in natural and articial environments. Chem Rev
110:46564689. https://doi.org/10.1021/cr900334y
29. Bobo D, Robinson KJ, Islam J et al. (2016) Nanoparticle-based
medicines: a review of FDA-approved materials and clinical
trials to date. Pharm Res 33:23732387. https://doi.org/10.1007/
s11095-016-1958-5
30. Lo TH, Wu ZY, Chen SY et al. (2021) Curcumin-loaded
mesoporous silica nanoparticles with dual-imaging and tem-
perature control inhibits the infection of Zika virus. Microporous
Mesoporous Mater 314:110886. https://doi.org/10.1016/j.
micromeso.2021.110886
31. Carvalho GC, Sábio RM, de Cássia Ribeiro T, et al. (2020)
Highlights in mesoporous silica nanoparticles as a multi-
functional controlled drug delivery nanoplatform for infectious
diseases treatment. Pharm Res 37:191. https://doi.org/10.1007/
s11095-020-02917-6
32. Ni D, Bu W, Ehlerding EB et al. (2017) Engineering of inorganic
nanoparticles as magnetic resonance imaging contrast agents.
Chem Soc Rev 46:74387468. https://doi.org/10.1039/
c7cs00316a
33. Anselmo AC, Mitragotri S (2015) A review of clinical translation
of inorganic nanoparticles. AAPS J 17:10411054. https://doi.
org/10.1208/s12248-015-9780-2
34. Giri K, Kuschnerus I, Ruan J, GarciaBennett AE (2019) Inu-
ence of a protein corona on the oral pharmacokinetics of tes-
tosterone released from mesoporous silica. Adv Ther 3:1900110.
https://doi.org/10.1002/adtp.201900110
35. Moore TL, Rodriguez-Lorenzo L, Hirsch V et al. (2015)
Nanoparticle colloidal stability in cell culture media and impact
on cellular interactions. Chem Soc Rev 44:62876305. https://
doi.org/10.1039/c4cs00487f
36. Tenzer S, Docter D, Kuharev J et al. (2013) Rapid formation of
plasma protein corona critically affects nanoparticle pathophy-
siology. Nat Nanotechnol 8:772781. https://doi.org/10.1038/
nnano.2013.181
37. Graf C, Gao Q, Schütz I et al. (2012) Surface functionalization of
silica nanoparticles supports colloidal stability in physiological
media and facilitates internalization in cells. Langmuir
28:7598613. https://doi.org/10.1021/la204913t
38. Hao N, Liu H, Li L et al. (2012) In vitro degradation behavior of
silica nanoparticles under physiological conditions. J Nanosci
Nanotechnol 12:63466354. https://doi.org/10.1166/jnn.2012.
6199
39. Mortensen NP, Hurst GB, Wang W et al. (2013) Dynamic
development of the protein corona on silica nanoparticles:
composition and role in toxicity. Nanoscale 5:63726380.
https://doi.org/10.1039/c3nr33280b
40. Cauda V, Argyo C, Bein T (2010) Impact of different PEGyla-
tion patterns on the long-term bio-stability of colloidal meso-
porous silica nanoparticles. J Mater Chem 20:86938699. https://
doi.org/10.1039/c0jm01390k
41. Khung YL, Narducci D (2015) Surface modication strategies
on mesoporous silica nanoparticles for anti-biofouling zwitter-
ionic lm grafting. Adv Colloid Interface Sci 226:166186.
https://doi.org/10.1016/j.cis.2015.10.009
42. Tiemblo P, García N, Hoyos M et al. (2015) Organic mod-
ication of hydroxylated nanoparticles: silica, sepiolite, and
polysaccharides. In: Aliofkhazraei M (ed) Handbook of nano-
particles. Springer International Publishing, Cham, p 135
43. Estephan ZG, Schlenoff PS, Schlenoff JB (2011) Zwitteration as
an alternative to PEGylation. Langmuir 27:67946800. https://
doi.org/10.1021/la200227b
44. Li Y, Hu X, Ding D et al. (2017) In situ targeted MRI detection
of Helicobacter pylori with stable magnetic graphitic nano-
capsules. Nat Commun 8:110. https://doi.org/10.1038/
ncomms15653
45. de Oliveira JFA, da Silva RF, Ribeiro IRS et al. (2020) Selective
targeting of lymphoma cells by monoclonal antibody grafted
onto Zwitterionic-functionalized nanoparticles. Part Part Syst
Charact 37:15. https://doi.org/10.1002/ppsc.201900446
46. Cheng W, Nie J, Xu L et al. (2017) pH-sensitive delivery vehicle
based on folic acid-conjugated polydopamine-modied
Journal of Sol-Gel Science and Technology
mesoporous silica nanoparticles for targeted cancer therapy. ACS
Appl Mater Interfaces 9:1846218473. https://doi.org/10.1021/a
csami.7b02457
47. de Oliveira LF, Bouchmella K, Gonçalves KDA et al. (2016)
Functionalized silica nanoparticles as an alternative platform for
targeted drug-delivery of water insoluble drugs. Langmuir
32:32173225. https://doi.org/10.1021/acs.langmuir.6b00214
48. Del Pino P, Pelaz B, Zhang Q et al. (2014) Protein corona for-
mation around nanoparticles from the past to the future. Mater
Horiz 1:301313. https://doi.org/10.1039/C3MH00106G
49. Nel AE, Mädler L, Velegol D et al. (2009) Understanding bio-
physicochemical interactions at the nanobio interface. Nat
Mater 8:543557. https://doi.org/10.1038/nmat2442
50. Schneid AC, Silveira CP, Galdino FE et al. (2020) Colloidal
stability and redispersibility of mesoporous silica nanoparticles
in biological media. Langmuir 36:1144211449. https://doi.org/
10.1021/acs.langmuir.0c01571
51. Braun K, Pochert A, Beck M et al. (2016) Dissolution kinetics of
mesoporous silica nanoparticles in different simulated body
uids. J Sol-Gel Sci Technol 79:319327. https://doi.org/10.
1007/s10971-016-4053-9
52. Wohlleben W, Driessen MD, Raesch S et al. (2016) Inuence of
agglomeration and specic lung lining lipid/protein interaction
on short-term inhalation toxicity. Nanotoxicology 10:970980.
https://doi.org/10.3109/17435390.2016.1155671
53. Francia V, Yang K, Deville S et al. (2019) Corona composition
can affect the mechanisms cells use to internalize nanoparticles.
ACS Nano 13:1110711121. https://doi.org/10.1021/acsnano.
9b03824
54. Auría-Soro C, Nesma T, Juanes-Velasco P, et al. (2019) Inter-
actions of nanoparticles and biosystems: Microenvironment of
nanoparticles and biomolecules in nanomedicine. Nanomaterials
9:1365. https://doi.org/10.3390/nano9101365
55. Israelachvili J (2013) Intermolecular and surface forces, 3rd edn.
Academic Press
56. Boström M, Deniz V, Franks GV, Ninham BW (2006) Extended
DLVO theory: electrostatic and non-electrostatic forces in oxide
suspensions. Adv Colloid Interface Sci 123126:515. https://
doi.org/10.1016/j.cis.2006.05.001
57. Tadros T (2011) General principles of colloid stability and the
role of surface forces. In: Tadros T (ed.) Colloid stability. Wiley-
VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, p 122
58. Parsegian VA (2005) Van Der Waals forces: a handbook for
biologists, chemists, engineers, and physicists. Cambridge Uni-
versity Press
59. Bergström L (1997) Hamaker constants of inorganic materials.
Adv Colloid Interface Sci 70:125169. https://doi.org/10.1007/
978-3-642-72624-8_11
60. Kontogeorgis GM, Kiil S (2016) Introduction to applied colloid
and surface chemistry. John Wiley & Sons, Ltd, Chichester, UK
61. Parks GA, de Bruyn PL (1962) The zero point of charge of
oxides. J Phys Chem 66:967973. https://doi.org/10.1021/
j100812a002
62. Adamson A, Gast A (1997) Physical chemistry of surfaces, 6th
edn. John Wiley & Sons, Inc.
63. Kosmulski M (2009) Compilation of PZC and IEP of sparingly
soluble metal oxides and hydroxides from literature. Adv Colloid
Interface Sci 152:1425. https://doi.org/10.1016/j.cis.2009.08.
003
64. Parks GA (1965) The isoelectric points of solid oxides, solid
hydroxides, and aqueous hydroxo complex systems. Chem Rev
65:177198. https://doi.org/10.1021/cr60234a002
65. Kosmulski M (2016) Isoelectric points and points of zero charge
of metal (hydr) oxides: 50 years after Parksreview. Adv Colloid
Interface Sci 238:161. https://doi.org/10.1016/j.cis.2016.10.005
66. Currie EPK, Norde W, Cohen Stuart MAC (2003) Tethered poly-
mer chains: surface chemistry and their impact on colloidal and
surface properties. Adv Colloid Interface Sci 100102:205265.
https://doi.org/10.1016/S0001-8686(02)00061-1
67. Bharti B, Meissner J, Klapp SHL, Findenegg GH (2014) Brid-
ging interactions of proteins with silica nanoparticles: the inu-
ence of pH, ionic strength and protein concentration. Soft Matter
10:71828. https://doi.org/10.1039/c3sm52401a
68. Grasso D, Subramaniam K, Butkus M et al. (2002) A review of
non-DLVO interactions in environmental colloidal systems. Rev
Environ Sci Bio/Technol 1:1738. https://doi.org/10.1023/A:
1015146710500
69. Ninham BW (1999) On progress in forces since the DLVO
theory. Adv Colloid Interface Sci 83:117. https://doi.org/10.
1016/S0001-8686(99)00008-1
70. Sonnefeld J, Löbbus M, Vogelsberger W (2001) Determina-
tion of electric double layer parameters for spherical silica
particles under application of the triple layer model using
surface charge density data and results of electrokinetic sonic
amplitude measurements. Colloids Surf A Physicochem Eng
Asp 195:215225. https://doi.org/10.1016/S0927-7757(01)
00845-7
71. Hiemstra T, De Wit JC, Van Riemsdijk W (1989) Multisite
proton adsorption modeling at the solid/solution interface of
(hydr)oxides: a new approach. J Colloid Interface Sci
133:105117. https://doi.org/10.1016/0021-9797(89)90285-3
72. Sverjensky DA (2005) Prediction of surface charge on oxides in
salt solutions: revisions for 1:1 (M+L-) electrolytes. Geochim
Cosmochim Acta 69:225257. https://doi.org/10.1016/j.gca.
2004.05.040
73. Leroy P, Devau N, Revil A, Bizi M (2013) Inuence of surface
conductivity on the apparent zeta potential of amorphous silica
nanoparticles. J Colloid Interface Sci 410:8193. https://doi.org/
10.1016/j.jcis.2013.08.012
74. Allen LH, Matijevíc E, Meites L (1971) Exchange of Na+for
the silanolic protons of silica. J Inorg Nucl Chem 33:12931299.
https://doi.org/10.1016/0022-1902(71)80423-2
75. Ong S, Zhao X, Eisenthal KB (1992) Polarization of water
molecules at a charged interface: second harmonic studies of the
silica/water interface. Chem Phys Lett 191:327335. https://doi.
org/10.1016/0009-2614(92)85309-X
76. Pfeiffer-Laplaud M, Costa D, Tielens F et al. (2015) Bimodal
acidity at the amorphous silica/water interface. J Phys Chem C
119:2735427362. https://doi.org/10.1021/acs.jpcc.5b02854
77. I-Ssuer C, Maciel GE (1996) Probing hydrogen bonding and the
local environment of silanols on silica surfaces via nuclear spin
cross polarization dynamics. J Am Chem Soc 118:401406.
https://doi.org/10.1021/ja951550d
78. West R, Whatley LS, Lake KJ (1961) Hydrogen bonding studies.
V. The relative basicities of ethers, alkoxysilanes and siloxanes
and the nature of the silicon-oxygen bond. J Am Chem Soc
83:761764. https://doi.org/10.1021/ja01465a001
79. Depasse J, Watillon A (1970) The stability of amorphous col-
loidal silica. J Colloid Interface Sci 33:430438. https://doi.org/
10.1016/0021-9797(70)90235-3
80. Allen LH, MatijevićE (1969) Stability of colloidal silica. J
Colloid Interface Sci 31:287296. https://doi.org/10.1016/0021-
9797(69)90172-6
81. Vigil G, Xu Z, Steinberg S, Israelachvili J (1994) Interactions of
silica surfaces. J Colloid Interface Sci 165:367385. https://doi.
org/10.1006/jcis.1994.1242
82. Adler JJ, Rabinovich YI, Moudgil BM (2001) Origins of the
Non-DLVO force between glass surfaces in aqueous solution. J
Colloid Interface Sci 237:249258. https://doi.org/10.1006/jcis.
2001.7466
Journal of Sol-Gel Science and Technology
83. Zhmud BV, Meurk A, Bergström L (1998) Evaluation of surface
ionization parameters from AFM data. J Colloid Interface Sci
207:332343. https://doi.org/10.1006/jcis.1998.5783
84. Schrader AM, Monroe JI, Sheil R et al. (2018) Surface chemical
heterogeneity modulates silica surface hydration. Proc Natl Acad
Sci USA 115:28902895. https://doi.org/10.1073/pnas.
1722263115
85. He L, Hu Y, Wang M, Yin Y (2012) Determination of solvation
layer thickness by a magnetophotonic approach. ACS Nano
6:41964202. https://doi.org/10.1021/nn3007288
86. Kobayashi M, Skarba M, Galletto P et al. (2005) Effects of heat
treatment on the aggregation and charging of Stöber-type silica. J
Colloid Interface Sci 292:139147. https://doi.org/10.1016/j.jcis.
2005.05.093
87. Harding RD (1971) Stability of silica dispersions. J Colloid
Interface Sci 35:172174. https://doi.org/10.1016/0021-9797(71)
90203-7
88. Drescher D, Orts-Gil G, Laube G et al. (2011) Toxicity of
amorphous silica nanoparticles on eukaryotic cell model is
determined by particle agglomeration and serum protein
adsorption effects. Anal Bioanal Chem 400:13671373. https://
doi.org/10.1007/s00216-011-4893-7
89. Lesniak A, Fenaroli F, Monopoli MP et al. (2012) Effects of the
presence or absence of a protein corona on silica nanoparticle
uptake and impact on cells. ACS Nano 6:58455857. https://doi.
org/10.1021/nn300223w
90. Izak-Nau E, Voetz M, Eiden S et al. (2013) Altered character-
istics of silica nanoparticles in bovine and human serum: the
importance of nanomaterial characterization prior to its tox-
icological evaluation. Part Fibre Toxicol 10:56. https://doi.org/
10.1186/1743-8977-10-56
91. Docter D, Bantz C, Westmeier D et al. (2014) The protein corona
protects against size- and dose-dependent toxicity of amorphous
silica nanoparticles. Beilstein J Nanotechnol 5:13801392.
https://doi.org/10.3762/bjnano.5.151
92. Halamoda-kenzaoui B, Ceridono M, Colpo P et al. (2015) Dis-
persion behaviour of silica nanoparticles in biological media and
its inuence on cellular uptake. PLoS One 10:e0141593. https://
doi.org/10.1371/journal.pone.0141593
93. Saikia J, Yazdimamaghani M, Hadipour Moghaddam SP,
Ghandehari H (2016) Differential protein adsorption and cellular
uptake of silica nanoparticles based on size and porosity. ACS
Appl Mater Interfaces 8:3482034832. https://doi.org/10.1021/a
csami.6b09950
94. Gollwitzer C, Bartczak D, Goenaga-Infante H et al. (2016) A
comparison of techniques for size measurement of nanoparticles
in cell culture medium. Anal Methods 8:52725282. https://doi.
org/10.1039/C6AY00419A
95. Bharti B, Meissner J, Findenegg GH (2011) Aggregation of
silica nanoparticles directed by adsorption of lysozyme. Lang-
muir 27:98239833. https://doi.org/10.1021/la201898v
96. Kumar S, Aswal VK, Kohlbrecher J (2011) SANS and UV-vis
spectroscopy studies of resultant structure from lysozyme
adsorption on silica nanoparticles. Langmuir 27:1016710173.
https://doi.org/10.1021/la201291k
97. Orts-Gil G, Natte K, Thiermann R et al. (2013) On the role of
surface composition and curvature on biointerface formation and
colloidal stability of nanoparticles in a protein-rich model sys-
tem. Colloids Surf B Biointerfaces 108:110119. https://doi.org/
10.1016/j.colsurfb.2013.02.027
98. Galdino FE, Picco AS, Sforca ML et al. (2020) Effect of particle
functionalization and solution properties on the adsorption of
bovine serum albumin and lysozyme onto silica nanoparticles.
Colloids Surf B Biointerfaces 186:110677. https://doi.org/10.
1016/j.colsurfb.2019.110677
99. Monopoli MP, Walczyk D, Campbell A et al. (2011) Physical-
chemical aspects of protein corona: relevance to in vitro and
in vivo biological impacts of nanoparticles. J Am Chem Soc
133:25252534. https://doi.org/10.1021/ja107583h
100. Lara S, Perez-Potti A, Herda LM et al. (2018) Differential
recognition of nanoparticle protein corona and modied low-
density lipoprotein by macrophage receptor with collagenous
structure. ACS Nano 12:49304937. https://doi.org/10.1021/a
csnano.8b02014
101. Yamamoto E, Kuroda K (2016) Colloidal mesoporous silica
nanoparticles. Bull Chem Soc Jpn 89:501539. https://doi.org/
10.1246/bcsj.20150420
102. Lin Y, Abadeer N, Haynes CL (2011) Stability of small meso-
porous silica nanoparticles in biological media. Chem Commun
47:532534. https://doi.org/10.1039/C0CC02923H
103. Rascol E, Daurat M, Da Silva A et al. (2017) Biological fate of
Fe3O4 core-shell mesoporous silica nanoparticles depending on
particle surface chemistry. Nanomaterials 7:162. https://doi.org/
10.3390/nano7070162
104. Lin Y, Abadeer N, Hurley KR, Haynes CL (2011) Ultrastable,
redispersible, small, and highly organomodied mesoporous
silica nanotherapeutics. J Am Chem Soc 133:2044420457
105. Cauda V, Schlossbauer A, Bein T (2010) Bio-degradation study
of colloidal mesoporous silica nanoparticles: effect of surface
functionalization with organo-silanes and poly(ethylene glycol).
Microporous Mesoporous Mater 132:6071. https://doi.org/10.
1016/j.micromeso.2009.11.015
106. Monopoli MP, Åberg C, Salvati A, Dawson KA (2012) Bio-
molecular coronas provide the biological identity of nanosized
materials. Nat Nanotechnol 7:779786. https://doi.org/10.1038/
nnano.2012.207
107. Walkey CD, Chan WCW (2012) Understanding and controlling
the interaction of nanomaterials with proteins in a physiological
environment. Chem Soc Rev 41:27802799. https://doi.org/10.
1039/c1cs15233e
108. Vroman L, Adams A, Fischer G, Munoz P (1980) Interaction of
high molecular weight kininogen, factor XII, and brinogen in
plasma at interfaces. Blood 55:156159. https://doi.org/10.1182/
blood.V55.1.156.156
109. Wang L, Wang K, Santra S et al. (2006) Watching silica nano-
particles glow in the biological world. Anal Chem 78:646654.
https://doi.org/10.1021/ac0693619
110. Nakamura M (2010) Approaches to the biofunctionalization of
spherical silica nanomaterials. In: Kumar CSSR (ed) Nano-
technologies for the life sciences. Wiley-VCH Verlag GmbH &
Co. KGaA, Weinheim, Germany
111. Bagwe RP, Hilliard LR, Tan W (2006) Surface modication of
silica nanoparticles to reduce aggregation and nonspecic bind-
ing. Langmuir 22:43574362. https://doi.org/10.1021/la052797j
112. Jokerst JV, Lobovkina T, Zare RN, Gambhir SS (2011) Nano-
particle PEGylation for imaging and therapy. Nanomedicine
6:715728. https://doi.org/10.2217/nnm.11.19
113. Xu H, Yan F, Monson EE, Kopelman R (2003) Room-temperature
preparation and characterization of poly (ethylene glycol)-coated
silica nanoparticles for biomedical applications. J Biomed Mater Res
A 66:870879. https://doi.org/10.1002/jbm.a.10057
114. Thierry B, Zimmer L, McNiven S et al. (2008) Electrostatic self-
assembly of PEG copolymers onto porous silica nanoparticles.
Langmuir 24:81438150. https://doi.org/10.1021/la8007206
115. Rio-Echevarria IM, Selvestrel F, Segat D et al. (2010) Highly
PEGylated silica nanoparticles: ready to usestealth functional
nanocarriers. J Mater Chem 20:27802787. https://doi.org/10.
1039/b921735e
116. Díaz B, Sánchez-Espinel C, Arruebo M et al. (2008) Assessing
methods for blood cell cytotoxic responses to inorganic
Journal of Sol-Gel Science and Technology
nanoparticles and nanoparticle aggregates. Small 4:20252034.
https://doi.org/10.1002/smll.200800199
117. Zhu Y, Fang Y, Borchardt L, Kaskel S (2011) PEGylated hollow
mesoporous silica nanoparticles as potential drug delivery vehi-
cles. Microporous Mesoporous Mater 141:199206. https://doi.
org/10.1016/j.micromeso.2010.11.013
118. Mosquera J, García I, Henriksen-Lacey M et al. (2019) Reducing
protein corona formation and enhancing colloidal stability of
gold nanoparticles by capping with silica monolayers. Chem
Mater 31:5761. https://doi.org/10.1021/acs.chemmater.8b04647
119. Perdoor SS, Dubois F, Barbara A et al. (2020) Ultrabright silica-
coated organic nanocrystals for two-photon in vivo imaging.
ACS Appl Nano Mater 3:1193311944. https://doi.org/10.
1021/acsanm.0c02499
120. Keefe AJ, Jiang S (2012) Poly(zwitterionic)protein conjugates
offer increased stability without sacricing binding afnity or
bioactivity. Nat Chem 4:5963. https://doi.org/10.1038/nchem.
1213
121. García KP, Zarschler K, Barbaro L et al. (2014) Zwitterionic-
coated stealthnanoparticles for biomedical applications:
Recent advances in countering biomolecular corona formation
and uptake by the mononuclear phagocyte system. Small
10:25162529. https://doi.org/10.1002/smll.201303540
122. Schlenoff JB (2014) Zwitteration: coating surfaces with zwit-
terionic functionality to reduce non specic adsorption. Lang-
muir 30:96259636. https://doi.org/10.1021/la500057j
123. Knop K, Hoogenboom R, Fischer D, Schubert US (2010) Poly
(ethylene glycol) in drug delivery: pros and cons as well as
potential alternatives. Angew Chem Int Ed 49:62886308.
https://doi.org/10.1002/anie.200902672
124. Panja P, Das P, Mandal K, Jana NR (2017) Hyperbranched
polyglycerol grafting on the surface of silica-coated nano-
particles for high colloidal stability and low nonspecic inter-
action. ACS Sustain Chem Engeneering 5:48794889. https://
doi.org/10.1021/acssuschemeng.7b00292
125. Liu J, Pelton R, Hrymak AN (2000) Properties of poly(N-iso-
propylacrylamide)-grafted colloidal silica. J Colloid Interface Sci
227:408411. https://doi.org/10.1006/jcis.2000.6871
126. Schild HG (1992) Poly(N-isopropylacrylamide): experiment,
theory and application. Prog Polym Sci 17:163249. https://doi.
org/10.1016/0079-6700(92)90023-R
127. Muller P (1994) Glossary of terms used in physical organic
chemistry (IUPAC Recommendations 1994). Pure Appl Chem
66:10771184. https://doi.org/10.1351/pac199466051077
128. Estephan ZG, Jaber JA, Schlenoff JB (2010) Zwitterion-
stabilized silica nanoparticles: toward nonstick nano. Langmuir
26:1688416889. https://doi.org/10.1021/la103095d
129. Hayward JA, Chapman D (1984) Biomembrane surfaces as
models for polymer design: the potential for haemocompatibility.
Biomaterials 5:135142. https://doi.org/10.1016/0142-9612(84)
90047-4
130. Ishihara K, Aragaki R, Ueda T et al. (1990) Reduced thrombo-
genicity of polymers having phospholipid polar groups. J
Biomed Mater Res 24:10691077. https://doi.org/10.1002/jbm.
820240809
131. Ishihara K, Ziats NP, Tierney BP et al. (1991) Protein adsorption
from human plasma is reduced on phospholipid polymers. J
Biomed Mater Res 25:13971407. https://doi.org/10.1002/jbm.
820251107
132. Knowles BR, Wagner P, Maclaughlin S et al. (2017) Silica
nanoparticles functionalized with Zwitterionic sulfobetaine
siloxane for application as a versatile antifouling coating system.
ACS Appl Mater Interfaces 9:1858418594. https://doi.org/10.
1021/acsami.7b04840
133. Knowles BR, Yang D, Wagner P et al. (2019) Zwitterion func-
tionalized silica nanoparticle coatings: the effect of particle size
on protein, bacteria, and fungal spore adhesion. Langmuir
35:13351345. https://doi.org/10.1021/acs.langmuir.8b01550
134. Loiola LMD, Batista M, Capeletti LB et al. (2019) Shielding and
stealth effects of zwitterion moieties in double-functionalized
silica nanoparticles. J Colloid Interface Sci 553:540548. https://
doi.org/10.1016/j.jcis.2019.06.044
135. Hu F, Chen K, Xu H, Gu H (2015) Functional short-chain
zwitterion coated silica nanoparticles with antifouling property in
protein solutions. Colloids Surf B Biointerfaces 126:251256.
https://doi.org/10.1016/j.colsurfb.2014.12.036
136. Knowles BR, Wagner P, Maclaughlin S et al. (2020) Carbox-
ybetaine functionalized nanosilicas as protein resistant surface
coatings. Biointerphases 15:011001. https://doi.org/10.1063/1.
5126467
137. Wang H, Cheng F, Shen W et al. (2016) Amino acid-based anti-
fouling functionalization of silica nanoparticles using divinyl
sulfone. Acta Biomater 40:273281. https://doi.org/10.1016/j.
actbio.2016.03.035
138. Shahabi S, Treccani L, Dringen R, Rezwan K (2015) Modulation
of silica nanoparticle uptake into human osteoblast cells by
variation of the ratio of amino and sulfonate surface groups:
effects of serum. ACS Appl Mater Interfaces 7:1382113833.
https://doi.org/10.1021/acsami.5b01900
139. Encinas N, Angulo M, Astorga C et al. (2019) Mixed-charge
pseudo-zwitterionic mesoporous silica nanoparticles with low-
fouling and reduced cell uptake properties. Acta Biomater
84:317327. https://doi.org/10.1016/j.actbio.2018.12.012
140. Scheffer FR, Silveira CP, Morais J et al. (2020) Tailoring
pseudo-zwitterionic bifunctionalized silica nanoparticles: from
colloidal stability to biological interactions. Langmuir
36:1075610763. https://doi.org/10.1021/acs.langmuir.0c01545
141. Sun JT, Yu ZQ, Hong CY, Pan CY (2012) Biocompatible
zwitterionic sulfobetaine copolymer-coated mesoporous silica
nanoparticles for temperature-responsive drug release. Macromol
Rapid Commun 33:811818. https://doi.org/10.1002/marc.
201100876
142. Runser A, Dujardin D, Ernst P et al. (2020) Zwitterionic stealth
dye-loaded polymer nanoparticles for intracellular imaging. ACS
Appl Mater Interfaces 12:117125. https://doi.org/10.1021/acsa
mi.9b15396
143. Dong Z, Mao J, Yang M et al. (2011) Phase behavior of poly
(sulfobetaine methacrylate)-grafted silica nanoparticles and their
stability in protein solutions. Langmuir 27:1528215291. https://
doi.org/10.1021/la2038558
144. Zhu Y, Sundaram HS, Liu S et al. (2014) A robust graft-to
strategy to form multifunctional and stealth zwitterionic
polymer-coated mesoporous silica nanoparticles. Biomacromo-
lecules 15:18451851. https://doi.org/10.1021/bm500209a
145. Chen K, Hu F, Gu H, Xu H (2017) Tuning of surface protein
adsorption by spherical mixed charged silica brushes (MCB)
with zwitterionic carboxybetaine component. J Mater Chem B
5:435443. https://doi.org/10.1039/c6tb02817a
146. Jia G, Cao Z, Xue H et al. (2009) Novel zwitterionic-polymer-
coated silica nanoparticles. Langmuir 25:31963199. https://doi.
org/10.1021/la803737c
147. Matsuda Y, Kobayashi M, Annaka M et al. (2008) Dimensions
of a free linear polymer and polymer immobilized on silica
nanoparticles of a zwitterionic polymer in aqueous solutions with
various ionic strengths. Langmuir 24:87728778. https://doi.org/
10.1021/la8005647
148. Sanchez-Salcedo S, Vallet-Regí M, Shahin SA et al. (2018)
Mesoporous core-shell silica nanoparticles with anti-fouling
properties for ovarian cancer therapy. Chem Eng J 340:114124.
https://doi.org/10.1016/j.cej.2017.12.116
149. Safavi-Sohi R, Maghari S, RaouM et al. (2016) Bypassing
protein corona issue on active targeting: Zwitterionic coatings
Journal of Sol-Gel Science and Technology
dictate specic interactions of targeting moieties and cell
receptors. ACS Appl Mater Interfaces 8:2280822818. https://
doi.org/10.1021/acsami.6b05099
150. Pálmai M, Nagy LN, Mihály J et al. (2013) Preparation, pur-
ication, and characterization of aminopropyl-functionalized
silica sol. J Colloid Interface Sci 390:3440. https://doi.org/10.
1016/j.jcis.2012.09.025
151. Picco AS, Ferreira LF, Liberato MS et al. (2018) Freeze-drying
of silica nanoparticles: Redispersibility toward nanomedicine
applications. Nanomedicine 13:1618. https://doi.org/10.2217/
nnm-2017-0280
152. Du X, Kleitz F, Li X et al. (2018) Disulde-bridged organosilica
frameworks: designed, synthesis, redox-triggered biodegrada-
tion, and nanobiomedical applications. Adv Funct Mater
28:1707325. https://doi.org/10.1002/adfm.201707325
153. Fuentes C, Ruiz-Rico M, Fuentes A et al. (2020) Degradation of
silica particles functionalised with essential oil components
under simulated physiological conditions. J Hazard Mater
399:123120. https://doi.org/10.1016/j.jhazmat.2020.123120
154. Liu L, Kong C, Huo M et al. (2018) Schiff base interaction tuned
mesoporous organosilica nanoplatforms with pH-responsive
degradability for efcient anti-cancer drug delivery in vivo.
Chem Commun 54:91909193. https://doi.org/10.1039/
C8CC05043K
155. Croissant JG, Fatieiev Y, Julfakyan K et al. (2016) Biodegrad-
able oxamide-phenylene-based mesoporous organosilica nano-
particles with unprecedented drug payloads for delivery in cells.
Chemistry 22:1480614811. https://doi.org/10.1002/chem.
201601714
156. Fei W, Chen D, Tang H et al. (2020) Targeted GSH-exhausting
and hydroxyl radical self-producing manganesesilica nano-
missiles for MRI guided ferroptotic cancer therapy. Nanoscale
12:1673816754. https://doi.org/10.1039/D0NR02396E
157. Yang G, Phua SZF, Bindra AK, Zhao Y (2019) Degradability
and clearance of inorganic nanoparticles for biomedical appli-
cations. Adv Mater 31:e1805730. https://doi.org/10.1002/adma.
201805730
158. Dietzel M (2000) Dissolution of silicates and the stability of
polysilicic acid. Geochim Cosmochim Acta 64:32753281.
https://doi.org/10.1016/S0016-7037(00)00426-9
159. He Q, Zhang Z, Gao Y et al. (2009) Intracellular localization and
cytotoxicity of spherical mesoporous silica nano- and micro-
particles. Small 5:27222729. https://doi.org/10.1002/smll.
200900923
160. Popplewell J, King S, Day J et al. (1998) Kinetics of uptake and
elimination of silicic acid by a human subject: a novel applica-
tion of 32Si and accelerator mass spectrometry. J Inorg Biochem
69:177180. https://doi.org/10.1016/S0162-0134(97)10016-2
161. Ogura M, Shinomiya S, Tateno J et al. (2001) Alkali-treatment
techniquenew method for modication of structural and acid-
catalytic properties of ZSM-5 zeolites. Appl Catal A Gen
219:3343. https://doi.org/10.1016/S0926-860X(01)00645-7
162. Rimer JD, Trofymluk O, Navrotsky A et al. (2007) Kinetic and
thermodynamic studies of silica nanoparticle dissolution. Chem
Mater 19:41894197. https://doi.org/10.1021/cm070708d
163. Etienne M, Walcarius A (2003) Analytical investigation of the
chemical reactivity and stability of aminopropyl-grafted silica in
aqueous medium. Talanta 59:11731188. https://doi.org/10.
1016/S0039-9140(03)00024-9
164. Nangia S, Garrison BJ (2008) Reaction rates and dissolution
mechanisms of quartz as a function of pH. J Phys Chem A
112:20272033. https://doi.org/10.1021/jp076243w
165. Nangia S, Garrison BJ (2009) Ab initio study of dissolution and
precipitation reactions from the edge, kink, and terrace sites of
quartz as a function of pH. Mol Phys 107:831843. https://doi.
org/10.1080/00268970802665621
166. Kosmulski M (2011) The pH-dependent surface charging and
points of zero charge. J Colloid Interface Sci 353:115. https://
doi.org/10.1016/j.jcis.2010.08.023
167. Kosmulski M (2018) The pH dependent surface charging and
points of zero charge. VII. Update. Adv Colloid Interface Sci
251:115138. https://doi.org/10.1016/j.cis.2017.10.005
168. Ratirotjanakul W, Suteewong T, Polpanich D, Tangboriboonrat
P (2019) Amino acid as a biodegradation accelerator of meso-
porous silica nanoparticles. Microporous Mesoporous Mater
282:243251. https://doi.org/10.1016/j.micromeso.2019.02.033
169. Choi E, Kim S (2019) Surface pH buffering to promote degra-
dation of mesoporous silica nanoparticles under a physiological
condition. J Colloid Interface Sci 533:463470. https://doi.org/
10.1016/j.jcis.2018.08.088
170. Yang Y, Wan J, Niu Y et al. (2016) Structure-dependent and
glutathione-responsive biodegradable dendritic mesoporous
organosilica nanoparticles for safe protein delivery. Chem Mater
28:90089016. https://doi.org/10.1021/acs.chemmater.6b03896
171. Shen D, Yang J, Li X et al. (2014) Biphase stratication
approach to three-dimensional dendritic biodegradable meso-
porous silica nanospheres. Nano Lett 14:923932. https://doi.
org/10.1021/nl404316v
172. Yamada H, Urata C, Aoyama Y et al. (2012) Preparation of
colloidal mesoporous silica nanoparticles with different dia-
meters and their unique degradation behavior in static aqueous
systems. Chem Mater 24:14621471. https://doi.org/10.1021/
cm3001688
173. Hadipour Moghaddam SP, Saikia J, Yazdimamaghani M,
Ghandehari H (2017) Redox-responsive polysulde-based bio-
degradable organosilica nanoparticles for delivery of bioactive
agents. ACS Appl Mater Interfaces 9:2113321146. https://doi.
org/10.1021/acsami.7b04351
174. Möller K, Bein T (2019) Degradable drug carriers: vanishing
mesoporous silica nanoparticles. Chem Mater 31:43644378.
https://doi.org/10.1021/acs.chemmater.9b00221
175. Choi E, Lim D-K, Kim S (2020) Hydrolytic surface erosion of
mesoporous silica nanoparticles for efcient intracellular deliv-
ery of cytochrome c. J Colloid Interface Sci 560:416425.
https://doi.org/10.1016/j.jcis.2019.10.100
176. Yang SA, Choi S, Jeon SM, Yu J (2018) Silica nanoparticle
stability in biological media revisited. Sci Rep 8:19. https://doi.
org/10.1038/s41598-017-18502-8
177. Quignard S, Masse S, Laurent G, Coradin T (2013) Introduction
of disulde bridges within silica nanoparticles to control their
intra-cellular degradation. Chem Commun 49:34103412.
https://doi.org/10.1039/c3cc41062e
178. Yu L, Chen Y, Wu M et al. (2016) Manganese extraction
strategy enables tumor-sensitive biodegradability and ther-
anostics of nanoparticles. J Am Chem Soc 138:98819894.
https://doi.org/10.1021/jacs.6b04299
179. Sha Z, Yang S, Fu L et al. (2021) Manganese-doped gold core
mesoporous silica particles as a nanoplatform for dual-modality
imaging and chemo-chemodynamic combination osteosarcoma
therapy. Nanoscale 13:50775093. https://doi.org/10.1039/
d0nr09220g
180. Marques MRC, Loebenberg R, Almukainzi M (2011) Simulated
biological uids with possible application in dissolution testing.
Dissolution Technol 18:1528. https://doi.org/10.14227/
DT180311P15
181. Yang W, Tam J, Miller DA et al. (2008) High bioavailability
from nebulized itraconazole nanoparticle dispersions with bio-
compatible stabilizers. Int J Pharm 361:177188. https://doi.org/
10.1016/j.ijpharm.2008.05.003
182. Dong R, Liu Y, Zhou Y et al. (2011) Photo-reversible supra-
molecular hyperbranched polymer based on hostguest interac-
tions. Polym Chem 2:2771. https://doi.org/10.1039/c1py00426c
Journal of Sol-Gel Science and Technology
183. Wang D, Xu Z, Chen Z et al. (2014) Fabrication of single-hole
glutathione-responsive degradable hollow silica nanoparticles for
drug delivery. ACS Appl Mater Interfaces 6:1260012608.
https://doi.org/10.1021/am502585x
184. Tang H, Li C, Zhang Y et al. (2020) Targeted manganese doped
silica nano GSH-cleaner for treatment of liver cancer by
destroying the intracellular redox homeostasis. Theranostics
10:98659887. https://doi.org/10.7150/thno.46771
185. Maggini L, Travaglini L, Cabrera I et al. (2016) Biodegradable
peptide-silica nanodonuts. Chemistry 22:36973703. https://doi.
org/10.1002/chem.201504605
186. Croissant J, Cattoën X, Man MWC et al. (2014) Biodegradable
ethylene-bis(propyl)disulde-based periodic mesoporous orga-
nosilica nanorods and nanospheres for efcient in-vitro drug
delivery. Adv Mater 26:61746180. https://doi.org/10.1002/a
dma.201401931
187. Fatieiev Y, Croissant JG, Julfakyan K et al. (2015) Enzymati-
cally degradable hybrid organicinorganic bridged silsesquiox-
ane nanoparticles for in vitro imaging. Nanoscale
7:1504615050. https://doi.org/10.1039/C5NR03065J
188. Croissant JG, Mauriello-Jimenez C, Maynadier M et al. (2015)
Synthesis of disulde-based biodegradable bridged silsesquiox-
ane nanoparticles for two-photon imaging and therapy of cancer
cells. Chem Commun 51:1232412327. https://doi.org/10.1039/
C5CC03736K
189. Xu Z, Zhang K, Liu X, Zhang H (2013) A new strategy to
prepare glutathione responsive silica nanoparticles. RSC Adv
3:17700. https://doi.org/10.1039/c3ra43098g
190. Tang H, Chen D, Li C et al. (2019) Dual GSH-exhausting sor-
afenib loaded manganese-silica nanodrugs for inducing the fer-
roptosis of hepatocellular carcinoma cells. Int J Pharm
572:118782. https://doi.org/10.1016/j.ijpharm.2019.118782
191. Li XW, Zhao WR, Liu YJ et al. (2016) Facile synthesis of
manganese silicate nanoparticles for pH/GSH-responsive
T1-weighted magnetic resonance imaging. J Mater Chem B
4:43134321. https://doi.org/10.1039/C6TB00718J
192. Fontecave T, Sanchez C, Azaïs T, Boissière C (2012) Chemical
modication as a versatile tool for tuning stability of silica based
mesoporous carriers in biologically relevant conditions. Chem
Mater 24:43264336. https://doi.org/10.1021/cm302142k
193. Kempen PJ, Greasley S, Parker KA et al. (2015) Theranostic
mesoporous silica nanoparticles biodegrade after pro-survival drug
delivery and ultrasound/magnetic resonance imaging of stem cells.
Theranostics 5:631642. https://doi.org/10.7150/thno.11389
194. Ryabchikova EI, Chelobanov BP, Parkhomenko RG et al. (2017)
Degradation of core-shell Au@SiO 2 nanoparticles in biological
media. Microporous Mesoporous Mater 248:4653. https://doi.
org/10.1016/j.micromeso.2017.04.006
195. Prasetyanto EA, Bertucci A, Septiadi D et al. (2016) Breakable
hybrid organosilica nanocapsules for protein delivery. Angew
Chem Int Ed 55:33233327. https://doi.org/10.1002/anie.
201508288
196. Lu J, Liong M, Li Z et al. (2010) Biocompatibility, biodis-
tribution, and drug-delivery efciency of mesoporous silica
nanoparticles for cancer therapy in animals. Small 6:17941805.
https://doi.org/10.1002/smll.201000538
197. Bindini E, Ramirez M, de los A, Rios X et al. (2021) In vivo
tracking of the degradation of mesoporous silica through 89Zr
radio-labeled coreshell nanoparticles. Small 2101519:110.
https://doi.org/10.1002/smll.202101519
198. Bhavsar D, Patel V, Sawant K (2019) Systemic investigation of
in vitro and in vivo safety, toxicity and degradation of meso-
porous silica nanoparticles synthesized using commercial sodium
silicate. Microporous Mesoporous Mater 284:343352. https://
doi.org/10.1016/j.micromeso.2019.04.050
199. Allen LH, MatijevićE (1970) Stability of colloidal silica II. Ion
exchange. J Colloid Interface Sci 33:420429. https://doi.org/10.
1016/0021-9797(70)90234-1
Journal of Sol-Gel Science and Technology
... Gold nanoparticles may degrade through oxidative processes, with renal clearance depending on their size and surface chemistry [72][73][74]. Silica nanoparticles may break down into silicic acid, which is excreted via the kidneys [75][76][77]. Quantum dots demonstrate how surface charge affects clearance, with neutral coatings improving renal filtration [78,79]. ...
Article
Full-text available
Eggshells are regular domestic, agricultural waste, and the primary composition of eggshells is majorly of calcium, as well as trace amounts of magnesium and phosphorous. These two elements (calcium and magnesium) are also present in living organisms and play an important role in many biological processes (cell growth, muscle contraction, glycolysis, angiogenesis, and vasculogenesis). Considering their role in different biological processes, especially in angiogenesis (formation of new blood vessels from pre-existing vasculature), we hypothesized the involvement of calcium and magnesium (present in eggshells) in the nanoform may induce angiogenesis. To this context, the present manuscript attempts to design calcium-rich nanoparticles derived from both unfertilized and fertilized eggshells (U-ES and F-ES), and investigate their pro-angiogenic properties. Both U-ES nanoparticles (U-ES-NP) and F-ES nanoparticles (F-ES-NP) were developed by the calcination of raw eggshells. These nanoparticles (U-ES-NP and F-ES-NP) are characterized using various analytical techniques. These nanoparticles exhibit pro-angiogenic properties, as validated by in vitro assays (cell proliferation assay, tube formation assay, etc), ex vivo (chick aorta assay) and in vivo (chick choriallantoic membrane assay) experiments. The hemolysis experiment (ex vivo) was performed by incubating mouse RBCs with nanoparticles, which further validates the biocompatibility of these nanomaterials. Taking these results altogether, the current study demonstrates pro-angiogenic properties of biocompatible ES-NP, that could be further utilized for the treatment of several diseases and other biomedical applications after proper biosafety evaluation.
Article
Full-text available
NANOG transcription factor, a molecule associated with cancer cell resistance, was targeted using decoy oligodeoxynucleotides (ODNs) loaded nanoparticles. We designed decoy ODN to mimic a human gene promoter, aiming to disrupt the function of NANOG. Silica (SiO2) nanoparticles were coated with polydopamine (PDOPA) and loaded with decoy ODNs. These nanoparticles were then functionalized with a targeting agent, the NL2 peptide, to achieve specific binding to HER2‐positive cells. Following this, their physicochemical properties were characterized using Fourier‐transform infrared spectroscopy (FT‐IR), dynamic light scattering (DLS), scanning electron microscopy (SEM), UV–visible spectroscopy (UV–vis), and drug release assays. The effectiveness of these nanoparticles was then tested on SKBR3 (HER2 positive) and MDA‐MB‐468 (HER2 negative) breast cancer cells. The results were encouraging the nanocomposites were successfully absorbed by the cancer cells. SiO2@PDOPA/DEC‐NL2 significantly inhibits cell growth and increased cell death (apoptosis). These findings suggest that the synthesized nanosystem targeting the HER2 receptor can potentially suppress the cancerous properties of SKBR3 cells in comparison with MDA‐MB‐468 cells. In conclusion, this study presents a promising approach for targeted drug delivery in breast cancer treatment. The developed nanoparticles loaded with NANOG decoy ODNs effectively targeted HER2‐positive cancer cells and demonstrated significant anticancer properties.
Article
Full-text available
Mesoporous silica nanoparticles (MSNs) offer new opportunities for the oral delivery of peptide drugs and biologics, owing to their biocompatibility and high drug-loading capacity. This perspective article explores strategies for...
Article
Conspectus The heating of matter upon interacting with light is a fundamental process ubiquitous in the natural world. With the rise of nanotechnology over the past decades, a variety of nanomaterials capable of converting light into heat have been discovered and their physicochemical properties investigated. Perhaps the most exotic is the photothermal heating of metallic nanocrystals via surface plasmons. Here an incoming electromagnetic wave triggers the oscillation of the nanoparticle’s electron cloud. When in resonance, this generates an enormous increase to the absorption coefficient, enabling more energy to dissipate as heat. The plasmonic phenomenon has an incredibly diverse range of functions, from the vibrant coloration of medieval stained-glass windows to the localization and enhancement of light at the nanoscale level. Plasmonic heating or thermoplasmonics is a relatively new addition that has gained popularity mainly through applications in therapeutics and biotechnology. With this Account, we aim to put a spotlight on the use of thermoplasmonics to drive chemical synthesis, a rapidly expanding area of research with immense potential. Throughout the long tradition of chemical synthesis, chemists have rarely deviated from the typical oven or hot plate to set and maintain a homogeneous temperature within the reaction vessel. In contrast, the use of thermoplasmonic nanomaterials can introduce heterogeneity to the heating profile of a reaction by forming steep temperature gradients near the surface of nanoparticles. Additionally, photothermal conversion enables heat activated processes to benefit from the advantages of light initiation, e.g., contactless activation and spatial control. Thus, thermoplasmonics offers an attractive alternative to the long-standing norm. Several early studies demonstrated the power of this method, taking advantage of the localized heating to carry out reactions with minimal change to the bulk temperature of the surrounding medium. However, tapping into this potential can be very challenging as colloidal solutions tend to aggregate even with small changes to the environment. Different strategies have been utilized to overcome this obstacle, for example embedding particles into glass or other heterogeneous substrates. Our group has experimented with coating gold nanostructures with a silica shell. This ensures the structural and colloidal stability that is critical for thermoplasmonic chemistry. Recently, we applied this methodology to advance olefin metathesis, the synthesis of iron oxide (IO), palladium (Pd) and silver (Ag) nanoparticles, and the formation of various metal–organic frameworks (MOFs). In addition, highly stable hybrid materials could be isolated as composites of plasmonic particles with polymers, MOFs, and other nanostructures. The large variety of reaction conditions and the different precursors, additives, and catalysts that our method proved to be compatible with highlight the versatility that silica encapsulation provides. The unique properties of plasmonic heating coupled with the added stability can open a wide range of opportunities for more efficient reactions and altogether new reactivity along with the formation of novel composite materials.
Article
Full-text available
Biophotonics has gained significant interest in recent years due to its potential in medical theranostic applications, with nano‐materials emerging as key enablers for advancing optical and electronic functionalities in biological environments. In this study, conjugated polymer nanoparticles (CP‐NPs), namely regio‐regular poly(3‐hexylthiophene) (P3HT), [6,6]‐phenyl C61‐butyric acid methyl ester (PCBM), and their blend (P3HT:PCBM), are exploited as nano‐materials for biophotonic applications. The CP‐NPs, obtained via a nanoprecipitation method, showed an average size of ca. 180 nm. Their optoelectrical properties indicate visible absorbance (350–600 nm) and red/near infra‐red (NIR, 650–900 nm) emission, demonstrating their suitability for biophotonic applications, in particular in biohybrid interfaces where effective light absorption and emission in biological environments are crucial. Interestingly, under light stimulation, the photocurrent response of the CP‐NPs in electrolyte solution (phosphate‐buffered saline, PBS) showed a stable and reproducible signal (current density ranging from 0.18 to 7 nA cm⁻²) thereby enhancing their potential for bio‐sensing/stimulation. Simulations of CP‐NPs interactions with biological fluids (i.e., PBS) under light stimulation showed distinct carrier generation and transport behaviors, with P3HT‐NPs exhibiting consistent charge generation (up to 3 × 10²⁰ nA cm⁻³). These findings demonstrate that CP‐NPs are promising for biophotonic applications, such as photothermal therapy, due to their efficient charge transport, UV‐vis absorption, NIR emission, and controlled interactions with biological environments.
Article
Full-text available
Silica-binding peptides (SBPs) are increasingly recognized as versatile tools for various applications spanning biosensing, biocatalysis, and environmental remediation. This review explores the interaction between these peptides and silica surfaces, offering insights into how variables such as surface silanol density, peptide sequence and composition, and solution conditions influence binding affinity. Key advancements in SBP applications are discussed, including their roles in protein purification, biocatalysis, biosensing, and biomedical engineering. By examining the underlying binding mechanisms and exploring their practical potential, this work provides a comprehensive understanding of how SBPs can drive innovations in materials science and biotechnology.
Article
Full-text available
Electrification has become increasingly common in aerospace due to climate change concerns. After successful applications in general aviation aircraft, electrification is now addressing subregional (up to 19 passengers) and regional aircraft (around 80 passengers). Megawatt-class electric motors are needed both to drive propellers and to act as high-power generators in hybrid–electric propulsion systems. Power levels for this class of aircraft require a proper design of heat management systems capable of dissipating a much higher quantity of heat than that dissipated by traditional cooling systems. The technical solution here explored is based on the addition into a diathermic base liquid of nanoparticles, which can increase (by up to 30%) the thermal conductivity of the refrigerant, also providing large surface area enhancing the heat transfer capacity of base liquids. The Italian Aerospace Research Centre (CIRA), as part of the European research initiative Optimised Electric Network Architectures and Systems for More-Electric Aircraft (ORCHESTRA), developed a thermal management system (TMS) based on impinging jets technology for a 1 MW electric motor. In this work, a numerical verification of the possibility for nanofluids to improve the heat exchange efficiency of a submerged oil impinging jets TMS designed to directly cool the inner components of a 1 MW motor is conducted. Investigations aimed to analyse two nanoparticle types (alumina and graphite) added to diathermic oil with concentrations between 1% and 5% by volume. The application of nanofluids significantly increases final thermal conductivity with respect to conventional coolants, a 60% improvement in heat transfer at a fixed mass flow rate is achieved. Electric motor maximum temperatures are approximately 10% lower than those achieved with solely diathermic oil. This result is significant as a safety margin is needed in all cases where a sudden increase in power occurs.
Preprint
Full-text available
The safety of inorganic nanoparticles (NPs) remains a critical challenge for their clinical translation. To address this, we developed a machine-learning (ML) framework that predicts NP toxicity both in vitro and in vivo, leveraging physicochemical properties and experimental conditions. A curated in vitro cytotoxicity dataset was used to train and validate binary classification models, with top-performing models undergoing explainability analysis to identify key determinants of toxicity and establish structure-toxicity relationships. External testing with diverse mesoporous silica NPs validated the framework’s predictive accuracy for in vitro settings. To enable organ-specific toxicity predictions in vivo, we integrated a physiologically-based pharmacokinetic (PBPK) model into the ML pipeline to quantify NP exposure across organs. Retraining the ML models with PBPK-derived exposure metrics yielded robust predictions of organ-specific nanotoxicity, further validating the framework. This PBPK-informed ML approach can thus serve as a potential Novel Alternative Method (NAM) to streamline NP safety assessment, enabling the rational design of safer NPs and expediting their clinical translation.
Article
Full-text available
While mesoporous silica nanoparticles (MSNs) are extensively studied as high‐potential drug delivery platforms, the successful clinical translation of these nanocarriers strongly depends on their biodistribution, biodegradation, and elimination patterns in vivo. Here, a novel method is reported to follow the in vivo degradation of MSNs by tracking a radioactive label embedded in the silica structure. Core–shell silica nanoparticles (NPs) with a dense core and a mesoporous shell are labeled with low quantities of the positron emitter ⁸⁹Zr, either in the dense core or in the mesoporous shell. In vivo positron emission tomography imaging and ex vivo organ measurements reveal a remarkable difference in the ⁸⁹Zr biodistribution between the shell‐labeled and the core‐labeled NPs. Release of the radiotracer from shell‐labeled NPs is used as a probe of the extent of silica dissolution, and a prompt release of the radioisotope is observed, with partial excretion already in the first 2 h post injection, and a slower accumulation in bones over time. On the other hand, when ⁸⁹Zr is embedded in the nanoparticle core, the biodistribution remains largely unchanged during the first 6 h. These findings indicate that MSNs have fast, hour‐scale, degradation kinetics in vivo.
Article
Full-text available
Infectious or immune diseases have caused serious threat to human health due to their complexity and specificity, and emerging drug delivery systems (DDSs) have evolved into the most promising therapeutic strategy for drug‐targeted therapy. Various mesoporous biomaterials are exploited and applied as efficient nanocarriers to loading drugs by virtue of their large surface area, high porosity, and prominent biocompatibility. Nanosized mesoporous nanocarriers show great potential in biomedical research, and it has become the research hotspot in the interdisciplinary field. Herein, recent progress and assembly mechanisms on mesoporous inorganic biomaterials (e.g., silica, carbon, metal oxide) are summarized systematically, and typical functionalization methods (i.e., hybridization, polymerization, and doping) for nanocarriers are also discussed in depth. Particularly, structure–activity relationship and the effect of physicochemical parameters of mesoporous biomaterials, including morphologies (e.g., hollow, core–shell), pore textures (e.g., pore size, pore volume), and surface features (e.g., roughness and hydrophilic/hydrophobic) in DDS application are overviewed and elucidated in detail. As one of the important development directions, advanced stimuli‐responsive DDSs (e.g., pH, temperature, redox, ultrasound, light, magnetic field) are highlighted. Finally, the prospect of mesoporous biomaterials in disease therapeutics is stated, and it will open a new spring for the development of mesoporous nanocarriers.
Article
Full-text available
Infectious diseases are a major global concern being responsible for high morbidity and mortality mainly due to the development and enhancement of multidrug-resistant microorganisms exposing the fragility of medicines and vaccines commonly used to these treatments. Taking into account the scarcity of effective formulation to treat infectious diseases, nanotechnology offers a vast possibility of ground-breaking platforms to design new treatment through smart nanostructures for drug delivery purposes. Among the available nanosystems, mesoporous silica nanoparticles (MSNs) stand out due their multifunctionality, biocompatibility and tunable properties make them emerging and actual nanocarriers for specific and controlled drug release. Considering the high demand for diseases prevention and treatment, this review exploits the MSNs fabrication and their behavior in biological media besides highlighting the most of strategies to explore the wide MSNs functionality as engineered, smart and effective controlled drug release nanovehicles for infectious diseases treatment. Graphical AbstractSchematic representation of multifunctional MSNs-based nanoplatforms for infectious diseases treatment
Article
Full-text available
Background: Glutathione (GSH), the primary antioxidant in cells, could fight against oxidative stress. Tumor cells display a higher GSH level than normal cells for coping with the hyperoxidative state, which meets the requirements of enhanced metabolism and vicious proliferation. Therefore, the consumption of GSH will lead to cell redox imbalance and impede life activities. Herein, targeted sorafenib (SFB) loaded manganese doped silica nanoparticle (FaPEG-MnMSN@SFB) was constructed, which could destroy the intracellular redox homeostasis by consuming GSH. Methods: In this study, MnMSN was prepared by an optimized one-pot Stober's method for loading SFB, and FaPEG chain was modified on the surface of MnMSN to achieve long circulation and targeted delivery. The anticancer efficacy and mechanism of the designed FaPEG-MnMSN@SFB were assessed both in vitro and in vivo. Results: FaPEG-MnMSN@SFB exhibited efficient antitumor activity by dual depleting intracellular GSH (the degradation of MnMSN would consume intracellular GSH and the SFB would inhibit the effect of Xc⁻ transport system to inhibit GSH synthesis). Moreover, disruption of redox balance would lead to apoptosis and reactive oxygen species (ROS)-dependent ferroptosis of tumor cells. Conclusion: Such a GSH-starvation therapeutic strategy would cause multi-path programmed cell death and could be a promising strategy for cancer therapy.
Article
In this study, an effective and facile strategy is reported to construct a multifunctional nanoplatform by in situ doping metal manganese on gold core mesoporous silica nanoparticles (Au@MMSN). After further modification of alendronate (Ald) on Au@MMSN, the obtained Au@MMSN-Ald efficiently integrates bone targeted chem-chemodynamic combination therapy and dual-modailty computed tomography/magnetic resonance (CT/MR) imaging into a single platform. Specifically, Au@MMSN-Ald exhibits the excellent tumor microenvironment responsive drug release efficiency. The doxorubicin hydrochloride (DOX) loaded Au@MMSN-Ald (DOX@Au@MMSN-Ald) is demonstrated with excellent targeted ability toward osteosarcoma. Accordingly, in specific tumor microenvironment, DOX@Au@MMSN-Ald also displays outstanding combined efficiency for killing the cancer cell in vitro and suppressing the osteosarcoma growth in vivo. Benefiting from the Au nanoparticles confined in core and manganese ions released from shell, CT and MR dual-modality imaging were performed to verify the effective accumulation of Au@MMSN-Ald at the tumor site. Overall, the constructed DOX@Au@MMSN-Ald nanoparticles integrated imaging guide, responsive drug release and combination therapy, which may provide some insight for further biomedical applications in efficient osteosarcoma therapy.
Article
This study synthesized a temperature-controlled drug carrier with a dual-imaging function which when loaded in curcumin, inhibits the infection of Zika virus. The infection of Zika will lead to high fever, and the rise temperature in body is used to design a temperature-controlled release carrier, so that the drug can be released when symptoms occur. Curcumin is the drug we chose to inhibit the virus. To increase the solubility and biocompatibility of the drug, we loaded the drug into the pores of the drug carrier, mesoporous silica nanoparticles (MSN), which had bonded to temperature -responding polymers(PEGMA) and doped magnetic and phosphorescence imaging metal ions (Eu³⁺ and Gd³⁺). The experiment will be divided into two parts. In the first part, mesoporous silica nanoparticles with dual-imaging function and temperature-controlled release and loaded curcumin ([email protected]) were synthesized. First, Eu³⁺ and Gd³⁺ were doped into silica nanoparticles for imaging function. After the assembly formed, mesoporous silica nanoparticles with dual-imaging function (MSN-EuGd) can be obtained by calcining to removing the template. The synthesized MSN-EuGd will be modified with PEGMA on the surface and loaded with curcumin. The second part is comparing curcumin and [email protected] for the ability of anti-virus. Finally, the temperature controlled drug carrier with dual-imaging function to inhibit Zika virus was completely synthesized.
Article
Freeze-drying of nanoparticle suspensions is capable of generating stable nanoformulations with improved storage times and easier transportation. Nonetheless, nanoparticle aggregation is likely induced during freeze-drying, which reduces its redispersibility upon reconstitution and leads to undesirable effects such as non-specific toxicity and impaired efficacy. In this work, bovine serum albumin (BSA) is described as a suitable protectant for silica nanoparticles (SNPs), which result in solid structures with excellent redispersibility and negligible signs of aggregation even when longer storage times are considered. We experimentally demonstrated that massive system aggregation can be prevented when a saturated BSA corona around the nanoparticle is formed before the lyophilization process. Furthermore, the BSA corona is able to suppress non-specific interactions between these nanoparticles and biological systems, as evidenced by the lack of residual cytotoxicity, hemolytic activity and opsonin adsorption. Hence, BSA can be seriously considered for industry as an additive for nanoparticle freeze-drying since it generates solid and redispersible nanoformulations with improved biocompatibility.
Article
Herein, we demonstrate the potential of ultrabright fluorescent silica-coated organic nanocrystals for two-photon in vivo imaging. These unique nanoparticles (NPs) containing a crystalline core of small push–pull dipolar dye are specifically designed to exhibit two-photon absorption for fluorescence imaging. The NPs can be easily functionalized using click chemistry in pure water, with preservation of the organic core. A novel small-volume DLS technique was used to evaluate the effect of PEGylation on the colloidal stability of the NPs in complex media containing salts and proteins, mimicking the composition of blood serum. The potential of these bright red-emitting NPs for two-photon fluorescence imaging is demonstrated both in vitro and in vivo.
Article
The outreach of nanoparticle-based medical treatments has been severely hampered due to the imbalance between the efforts in designing extremely complex materials and the general lack of studies devoted to understanding their colloidal stability in biological environments. Over the years, the scientific community has neglected the relevance related to the nanoparticles' colloidal state, which consequently resulted in very poor bench-to-clinic translation. In this work, we show how mesoporous silica nanoparticles (MSNs, one of the most promising and tested drug delivery platforms) can be efficiently synthesized and prepared, resulting in a colloidally stable system. We first compared three distinct methods of template removal of MSNs and evaluated their ultimate colloidal stability. Then, we also proposed a simple way to prevent aggregation during the drying step by adsorbing BSA onto MSNs. The surface modification resulted in colloidally stable particles that are successfully redispersed in biologically relevant medium while retaining high hemocompatibility and low cytotoxicity.