ArticlePDF Available

In vitro osteogenesis by intracellular uptake of strontium containing bioactive glass nanoparticles

Authors:

Abstract and Figures

Significance: We report, for the first time, that monodispersed bioactive glass nanoparticles (∼90 nm) are internalised into preosteoblast cells by endocytosis but by unspecific mechanisms. The bioactive nanoparticles and their dissolution products (without the particles present) stimulated the expression of osteogenic markers from preosteoblast cells without the addition of other osteogenic supplements. Incorporating Sr into the bioactive glass nanoparticle composition, in addition to Ca, increased the total cation content (and therefore dissolution rate) of the nanoparticles, even though nominal total cation addition was constant, without changing size or morphology. Increasing Sr content in the nanoparticles and in their dissolution products enhanced osteogenesis in vitro. The particles therefore have great potential as an injectable therapeutic for bone regeneration, particularly in patients with osteoporosis, for which Sr is known to be therapeutic agent.
Content may be subject to copyright.
Full length article
In vitro osteogenesis by intracellular uptake of strontium containing
bioactive glass nanoparticles
Parichart Naruphontjirakul, Alexandra E. Porter, Julian R. Jones
Department of Materials, Imperial College London, South Kensington Campus, London SW7 2AZ, UK
article info
Article history:
Received 11 August 2017
Received in revised form 5 October 2017
Accepted 7 November 2017
Available online 10 November 2017
Keywords:
Bioactive glass nanoparticles
Osteogenic
Strontium
Intracellular uptake
Endocytosis
abstract
Monodispersed strontium containing bioactive glass nanoparticles (Sr-BGNPs) with two compositions
were synthesised, through a modified sol-gel Stöber process, wherein silica nanoparticles (SiO
2
-NPs)
were formed prior to incorporation of calcium and strontium, with diameters of 90 ± 10 nm. The osteo-
genic response of a murine preosteoblast cell line, MC3T3-E1, was investigated in vitro for a nanoparticle
concentration of 250 mg/mL with compositions of 87 mol% SiO
2
, 7 mol% CaO, 6 mol% SrO and 83 mol%
SiO
2
, 3 mol% CaO, 14 mol% SrO. Dissolution studies in minimum essential media (
a
-MEM) at pH 7.4
and artificial lysosomal fluid (ALF) at pH 4.5 showed that the particles dissolved and that Sr
2+
ions were
released from Sr-BGNPs in both environments. Both particle compositions and their ionic dissolution
products enhanced the alkaline phosphatase (ALP) activity of the cells and calcium deposition.
Immunohistochemistry (IHC) staining of Col1a1, osteocalcin (OSC) and osteopontin (OSP) showed that
these proteins were expressed in the MC3T3-E1 cells following three weeks of culture. In the basal con-
dition, the late osteogenic differentiation markers, OSC and OSP, were more overtly expressed by cells
cultured with Sr-BGNPs with 14 mol% SrO and their ionic release products than in the control condition.
Col1a1 expression was only slightly enhanced in the basal condition, but was enhanced further by the
osteogenic supplements. These data demonstrate that Sr-BGNPs accelerate mineralisation without osteo-
genic supplements. Sr-BGNPs were internalised into MC3T3-E1 cells by endocytosis and stimulated
osteogenic differentiation of the pre-osteoblast cell line. Sr-BGNPs are likely to be beneficial for bone
regeneration and the observed osteogenic effects of these particles can be attributed to their ionic release
products.
Statement of Significance
We report, for the first time, that monodispersed bioactive glass nanoparticles (90 nm) are internalised
into preosteoblast cells by endocytosis but by unspecific mechanisms. The bioactive nanoparticles and
their dissolution products (without the particles present) stimulated the expression of osteogenic mark-
ers from preosteoblast cells without the addition of other osteogenic supplements.
Incorporating Sr into the bioactive glass nanoparticle composition, in addition to Ca, increased the total
cation content (and therefore dissolution rate) of the nanoparticles, even though nominal total cation
addition was constant, without changing size or morphology.
Increasing Sr content in the nanoparticles and in their dissolution products enhanced osteogenesis
in vitro. The particles therefore have great potential as an injectable therapeutic for bone regeneration,
particularly in patients with osteoporosis, for which Sr is known to be therapeutic agent.
Ó2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
1. Introduction
Bioactive glass-based products are used for several orthopaedic
and dental applications [1] because they form strong bonds with
host bone [1–4] and the ions released from these products stimu-
late osteogenic gene expression, leading to rapid bone regeneration
[5–9]. Recently, the development of new formulations by introduc-
tion of other therapeutic metallic cations, such as copper, zinc,
cobalt into the silica network, and using co-networks of borate,
have widened the potential therapeutic applications to angiogene-
sis, wound healing, antimicrobial and anti-cancer [9–12].
https://doi.org/10.1016/j.actbio.2017.11.008
1742-7061/Ó2017 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Corresponding author.
E-mail address: julian.r.jones@imperial.ac.uk (J.R. Jones).
Acta Biomaterialia 66 (2018) 67–80
Contents lists available at ScienceDirect
Acta Biomaterialia
journal homepage: www.elsevier.com/locate/actabiomat
Strontium (Sr) has been clinically used as pharmaceutical agent for
osteoporosis treatment [13,14] because of its ability to activate
osteoblasts and inhibit osteoclast activities [15–18]. Strontium
ions (Sr
2+
) have a similar charge and ionic radius to calcium ions
(Ca
2+
)[15]. SrO can therefore replace some CaO in the bioactive
glass structure [15,19–21]. Sr is slightly larger and heavier com-
pared with Ca, therefore the substitution decreases the connectiv-
ity of the silica network by further disrupting the network [11,22].
Bioactive glass nanoparticles (BGNPs) are promising injectable
biomaterials for bone regeneration applications [12,23,24]. The
potential benefits of nanoparticles (NPs) over microparticles are
their high surface to volume ratios, which increases their dissolu-
tion rate [17,25,26], and their small size allows them to be inter-
nalised into cells to deliver their therapeutic ions intracellularly
[26,27]. The benefit of the glass is that it can provide a sustained
delivery of the therapeutic cations from their amorphous structure
during dissolution [6,16,26,28].
The size and shape of NPs is known to affect the endocytosis
pathway [29–33]. NPs internalise and localise within the cells
through different uptake pathways, including phagocytosis,
macropinocytosis, clathrin-independent endocytosis, and
clathrin-dependent endocytosis [34,35]. To explore the potential
utility of Sr-BGNPs as therapeutic cation carriers, it is critical to
understand the different mechanisms of Sr-BGNPs uptake by cells.
SiO
2
-CaO NPs with diameters of 215 ± 20 nm were previously
found to be internalised into human bone marrow and adipose
derived stem cells and internalization by the stem cells was not
affected following inhibition of clathrin- or caveolin- mediated
endocytosis [35].
A previous study [36] showed that cation incorporation into the
silica network was not trivial, indicating that previous studies had
overestimated the amount of calcium incorporated into nanoparti-
cles, without checking composition. The most effective method to
incorporate calcium ions into the NPs is by adding calcium nitrate,
after the monodispersed SiO
2
-NPs has already been formed [36].
This modified Stöber method was adapted to produce SiO
2
-CaO-
SrO NPs [37]. In each case, not all of the nominal Ca
2+
and Sr
2+
ions
were incorporated into the silica network [36,37]. There is an
upper limit of network modifier incorporation into the dense
SiO
2
-NPs. A maximum of 10% mol CaO incorporation into the bin-
ary BGNPs (90% mol SiO
2
and 10% mol CaO), using a nominal ratio
of Si:Ca of 1:1.3, was possible using a modified Stöber process [36].
The upper limit of Sr
2+
ion incorporation into the binary BGNPs was
approximately 17 mol% (83% mol SiO
2
and 17% mol SrO), using the
nominal ratio of Si:Sr of 1:1.3 [37]. Substituting Sr
2+
for Ca
2+
ions
on a molar basis had no effect on the size and morphology of the
particles. The in vitro cell viability also showed that the Sr-BGNPs
had low cytotoxicity (>70% viability) for particle concentrations
of up to 250
l
g/mL and the ions released from these particles sig-
nificantly increased the viability of the MC3T3-E1 osteoblast-like
cells at concentrations of 200 and 250
l
g/mL [37]. Only cell viabil-
ity was assessed.
Here, the aim was to investigate the effect of Sr-BGNPs on the
osteogenic response of MC3T3-E1 cells in vitro using Sr-BGNPs,
with diameters within the range of 80–100 nm, and the influ-
ence of mol% of SrO. The mechanism of incorporation of the
nanoparticles into the cells was also investigated. To understand
whether or not the therapeutic effect is dependent on intracellu-
lar release of therapeutic ions from the particles, the cells were
exposed to both the particles and their dissolution products.
Ion release from the Sr-BGNPs was also measured in three differ-
ent conditions: phosphate-buffered saline (PBS), minimum
essential media (
a
-MEM) at pH 7.4 and artificial lysosomal fluid
(ALF) at pH 4.5.
2. Materials and methods
2.1. Sr-BGNPs synthesis
All reagents were from Sigma–Aldrich (Dorset, UK) unless sta-
ted otherwise. SiO
2
-NPs and Sr-BGNPs, with diameters of 90
nm, were prepared using the modified Stöber method described
previously [36,37]. A diameter of 90 nm was chosen because of
previous work on size dependence on internalisation of Stöber-
like particles, where particles of <200 nm were internalised into
cells [38,39] and particles of 60 nm caused higher toxicity than
100 nm particles [40], which did not cause toxicity up to the con-
centration of 500 mg/mL [41].
For their synthesis, first, 32.92 mL of ethanol (99.5%), 4.11 mL of
distilled water, and 0.48 mL of ammonium hydroxide were mixed
in an ultrasonication bath for 10 min. Then, 2.50 mL of tetraethyl
orthosilicate (TEOS) was gently added to the mixed solution and
left in the ultrasonication bath at least 6 h to complete hydrolysis
and polycondensation reactions [37]. Hydrolysis and condensation
reactions of TEOS occurred simultaneously to form the silica net-
work (Si-O-Si). SiO
2
-NPs were centrifuged for collection and then
were washed with ethanol (two times) and distilled water.
For Ca and Sr incorporation, calcium nitrate tetrahydrate (99%)
and strontium nitrate tetrahydrate (99%) were added. Based on
previous work [37], a nominal molar ratio of 1:1.3 ratio of Si:total
cations was selected. Binary BGNPs (SiO
2
-CaO) and ternary BGNPs
(SiO
2
-CaO-SrO) compositions were synthesised, with 25 mol% and
75 mol% calcium nitrate tetrahydrate being replaced with stron-
tium nitrate tetrahydrate, leaving SiO
2
-CaO-SrO compositions con-
taining 6.2 mol% SrO (6%Sr-BGNPs) and 14.2 mol% SrO (14%Sr-
BGNPs) respectively (measured by acid digestion, Section 2.2).
The resulting primary particle suspension was then dried at 60 °C
overnight to remove excess water following with thermal treat-
ment at 680 °C for 3 h at a heating rate 3 °C/min in order to pro-
duce Sr-BGNPs. These particles were then washed with ethanol
two times.
2.2. Acid digestion compositional analysis
Acid digestion compositional analysis was carried out to mea-
sure the composition of the BGNPs by the lithium metaborate
fusion dissolution method. 50 mg of finely ground particles was
mixed carefully with 250 mg of anhydrous lithium metaborate
(80% w/w) and lithium tetraborate (20% w/w) (Spectroflux 100B,
Alfa Aesar, Lancashire, UK) in a clean and dry platinum crucible
using a glass rod [42]. The mixture was fused in a furnace for
20 min at 1050 °C and later dropped to room temperature. The
mixture was subsequently dissolved in 2 M nitric acid [43].
The elemental concentration in the solution was measured using
inductively coupled plasma optical emission spectroscopy (ICP-
OES, Thermo Scientific iCAP 6000 series).
2.3. Particle characterization
Particle size was investigated using Dynamic Light Scattering
(DLS, Malvern instrument 2000) and Transmission Electron Micro-
scopy (TEM, JEOL 2100 Plus microscope operated at 200 kV). To
prepare samples, the dried particles were dissolved in ethanol
and sonicated in the sonication bath for 15 min before conducting
the DLS measurements. Particles were collected on 400 mesh cop-
per transmission electron microscopy (TEM) grids, coated with
holey carbon film (TAAB, Berkshire, UK). TEM images were used
to confirm particles’ size and morphology.
68 P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80
To indicate the stability of particles in solutions, Zeta (f) poten-
tial values were measured in distilled water in three different pH:
3.0, 7.4, and 11.0 using Zeta sizer (Malvern instrument 2000).
X-ray Diffraction (XRD) patterns were collected with a Philips
PW1700 series automated powder diffractometer using Cu K
a
radiation (1.54 Å) at 40 kV/40 mA. Data was collected in the 10–
70°2hrange with a step size of 0.04°and a dwell time of 1.0 s to
identify the crystallised pattern of the particles.
2.4. Dissolution study
To compare the release rate of ions from the Sr-BGNPs with dif-
ferent mol% SrO, the release of Si, Ca and Sr ions from Sr-BGNPs
was evaluated as a function of time in three different solutions:
minimum essential media (
a
-MEM) medium (Thermo Fisher Sci-
entific, Hemel Hempstead, UK) at pH 7.4, artificial lysosomal fluid
(ALF) at pH 4.5 (Supplementary information Table S1) and
phosphate-buffered saline (PBS) at pH 7.4. 75 mg of BGNPs were
suspended with 5 mL of media in dialysis tubing, that had a molec-
ular weight cut-off of 10 kDa, and immersed into 45 mL of media
[42]. All samples were incubated at 37 °C with continuous shaking
at 120 rpm for 1, 2, 3, 4, 5, 6, 7, 24, 48, 72, 96, 168, 240 h in PBS and
ALF and for 1, 2, 3, 4, 5, 6, 7, 24 h in
a
-MEM medium. At each of the
time intervals, 1 mL of the bulk solution was collected and then
immediately replaced with 1 mL of the fresh solution. The pH of
solution was monitored at each specific interval over a period of
240 h in PBS and ALF and for 24 h in
a
-MEM.
The collected solution was diluted in distilled water (for the
a
-
MEM medium) and 2 M nitric acid (for ALF and PBS) with a 10-fold
dilution factor. The elemental concentrations of Si, Ca, and Sr were
measured using ICP-OES (Thermo Scientific iCAP 6000 series).
At the end of incubation period, the Sr-BGNPs were washed
with ethanol and acetone to terminate any reactions [42] and then
collected on 400 mesh copper TEM grids, coated with a holey car-
bon support film and bright field TEM imaging was conducted.
2.5. In vitro cytotoxicity assay
The next step was to evaluate whether the therapeutic effect of
the Sr-BGNPs could be linked directly to an ionic effect of the Sr
and Ca ions which are released from the particles. A murine pre-
osteoblast cell line, MC3T3-E1 cells (ATCC) was incubated with
both the particles and media containing only their dissolution
products. MC3T3-E1 cells were routinely cultured under standard
condition in a humidified atmosphere at 37 °C and 5% CO
2
in basal
a
-MEM media. These media were supplemented with 10% fetal
bovine serum (FBS) (v/v), 100 U/mL penicillin and 100
l
g/mL
streptomycin (Thermo Fisher Scientific, Hemel Hempstead, UK).
Cells were seeded in flat-bottomed 96-well plates, at a seeding
density of 5 10
4
cells/mL, and incubated at 37 °C and 5% CO
2
for 24 h to allow cells to attach in a monolayer.
Based on our previous report [37], 6%Sr-BGNPs and 14%Sr-
BGNPs did not cause toxicity to the MC3T3-E1 cells up to 250 mg/
mL. Here, effects of the of 6%Sr-BGNPs and 14%Sr-BGNPs on cell
viability was investigated with the extended NP concentration
range from 0 to 1000 mg/mL (0.01, 0.1, 1, 10, 100, 150, 200, 250,
500, 1000) using a pulse-chase exposure, where cells were exposed
to the pulse phase for 24 h, followed by chase period of 1, 3 and 7
days. The effect of the ions released from the 6%Sr-BGNPs and 14%
Sr-BGNPs on cell viability was measured after cells were incubated
with the media retrieved from the dissolution experiments for 1, 3,
and 7 days.
Cell viability was determined using the MTT colorimetric assay
(Thermo Fisher Scientific, Hemel Hempstead, UK) based on the
conversion of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazo
lium bromide (MTT) into formazan. The formazan is soluble in
dimethyl sulfoxide (DMSO) and the concentration of soluble for-
mazan was determined at 570 nm using a microplate reader (Spec-
traMax M2
e
, Molecular device).
2.6. Effect of Sr-BGNPs and their ionic release products on MC3T3-E1
differentiation
MC3T3-E1 cells were cultured, using a cell density at 5 10
4
cells/mL, in a flat-bottomed 24-well plate and incubated at 37 °C
and 5% CO
2
for 24 h to allow cells to attach to the plate. Cells were
cultured either in basal
a
-MEM or osteogenic medium (
a
-MEM
supplemented with 100
l
M L-ascorbic acid (Sigma-Aldrich, UK),
10 mM b-glycerophosphate (Sigma-Aldrich, UK) and 10 nM dex-
amethasone (DEX, Sigma-Aldrich, UK)).
Cells were exposed to Sr-BGNPs or the media containing the 6%
Sr-BGNP’s and 14%Sr-BGNP’s ionic release products both in the
basal and osteogenic conditions. Media containing the ionic release
products were made by immersing Sr-BGNPs in the media at con-
centration of 250 mg/mL for 24 h with continuous shaking at 120
rpm. The culture media were changed every three days to ensure
a high nutrient concentration. The NPs were incubated with the
cells with every media change. Cells were fixed with 4%
paraformaldehyde in PBS at time intervals of up to 21 days. Cells
were stained with an alkaline phosphate (ALP) detection kit (Merck
Millipore, Middlesex, UK) according to the manufacturer’s
instructions.
Key osteoblastic differentiations marker staining was carried
out for Collagen type 1 (Col1a1), osteocalcin (OSC), and osteopon-
tin (OSP). After fixation, cells were permeabilised with a perme-
ability buffer for 30 min and then blocked with 1% BSA in PBS for
5 min. The cells were stained with a rabbit IgG primary antibody
(Abcam, Cambridge, UK) at 4 °C overnight. The secondary antibody
used for immunofluorescence was goat anti-rabbit IgG H&L conju-
gated with Alexa Fluor 455 (Abcam, Cambridge, UK). 4
0
,6-Diami
dino-2-Phenylindole, Dihydrochloride (DAPI, Thermo Fisher Scien-
tific, Hemel Hempstead, UK) was used to stain nuclei.
Cells were stained with 1% Alizarin Red S in PBS at pH 4.2 to
detect calcified tissue formation.
2.7. Endocytosis study
2.7.1. Fluorescent labelling of nanoparticles
14%Sr-BGNPs were functionalised using a modified method
from previous work [44]. 50 mg of 14%Sr-BGNPs was re-
dispersed in 10 mL absolute ethanol (5% w/v) followed by careful
addition of 500 mL of 28% NH
4
OH. Then, 2 mL of (3-aminopropyl)
triethoxysilane (APTES) was mixed on a shaker at 200 rpm over-
night to complete reactions. Finally, amine functionalised Sr-
BGNPs were completely washed with absolute ethanol two times
to remove the excess components.
After Sr-BGNPs were functionalised with amine groups, the par-
ticles were labelled with fluorescein coupling using a modified
method from previous research [45]. For 1% w/v fluorecein 5(6)-
isothiocyanate bioreage: FITC, 20 mg of FITC was dissolved in abso-
lute ethanol and gently mixed (200 rpm) in dark conditions. Next,
20 mg of amine functionalised Sr-BGNPs was added and stirred for
16 h to complete the reaction. Lastly, FITC-14%Sr-BGNPs were
washed with absolute ethanol (2 times) and D.I. water (once)
(Scheme 1). To investigate the effect of the FITC conjugated BGNPs
(FITC-14%Sr-BGNPs) on the viability of the MCT3T-E1 cells, an MTT
assay was performed.
2.7.2. Endocytosis mechanism inhibition study
To understand the mechanism by which the NPs were inter-
nalised by the cells, MCT3T-E1 cells (cell density at 5 10
4
cells/
mL) were seeded on 6-well plates. After the cells were cultured
P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80 69
under standard conditions, in a humidified atmosphere at 37 °C
and 5% CO
2
in basal
a
-MEM media, the cells were subsequently
incubated with different endocytosis inhibitors for 2 h.
The MC3T3-E1 cells were then pre-treated at 37 °C for 2 h
with five different endocytosis chemical inhibitors including
wortmannin (wor, an inhibitor of phagocytosis); amiloride
hydrochloride hydrate (ami, an inhibitor of macropinocytosis);
chlorpromazine hydrochloride USP (chlor, an inhibitor of
macropinocytosis); genistein (gen, an inhibitor of clathrin-
independent endocytosis) and cytochalasin D (cytD, an inhibitor
of clathrin-dependent endocytosis). The NPs were replaced at
concentration 250 mg/mL and were then incubated for 24 h. The
concentration of endocytosis inhibitors was: 23 mM of wor; 1
mM of ami; 37 mM of gen (VWR International Ltd, Lutterworth,
UK); 4 mM of cytD; and 30 mM of chlor. These concentrations were
selected using previous work as a guide [46]. After 24 h cells were
fixed with 4% paraformaldehyde in PBS.
2.8. Statistics
Statistical analyses were performed by one-way analysis of
variance (ANOVA) in Minitab with the appropriate post hoc com-
parison test (Tukey’s test). A p-value <.05 was considered signifi-
cant. The graphs shown present the results as the mean value
with the standard deviation (SD) as the error bars.
3. Results
3.1. Material characterization
Table 1 shows the compositions of the Sr-BGNPs. Elemental
analysis confirmed that the binary glass had a composition of 93
mol% SiO
2
, 7 mol% CaO. Interestingly, adding a third component
(Sr), while keeping the nominal ratio of Si:total modifying cations
constant, increased the overall amount of network modifiers in the
stabilised NPs (13 mol% and 17 mol% CaO + SrO), compared with
the binary NPs containing only CaO (7 mol% CaO). This increase
might be due to the larger ionic radius of the Sr compared to Ca,
causing the silicate network to be more open, allowing diffusion
of both cations into the NPs during thermal stabilisation. However,
this does not explain the reduction in Ca content.
Scheme 1. (a) Sr-BGNPs conjugation with FITC and (b) indirect measurement of FITC-Sr-BGNPs endocytosis mechanism.
Table 1
Composition of the synthesised Sr-BGNPs (mol%).
Samples Mol%
SiO
2
CaO SrO
0%Sr-BGNPs 92.8 ± 1.3 7.2 ± 0.2 0.0 ± 0.0
6%Sr-BGNPs 87.1 ± 0.3 6.7 ± 0.2 6.2 ± 0.0
14%Sr-BGNPs 83.3 ± 0.2 2.5 ± 0.2 14.2 ± 0.0
70 P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80
The morphology of the batches and particle diameters agreed
with that found previously [37]. The different compositions of
the BGNPs, for the different Ca and Sr contents, did not influence
the size and morphology of the particles. TEM images of dense
monodisperse Sr and Si-BGNPs (Fig. 1) showed that they all had
spherical morphology. Fig. 2 shows their size distribution mea-
sured using DLS with modal diameters of 90 ± 10 nm.
As an indicator of the stability of particles in water, and the
effect of changing pH, the f-potential of the particles of were mea-
sured by suspending them in DI water at three different pH values:
3.0, 7.4 and 11. (Table 2). The f-potential values ranged from 1.5 ±
0.1 to 4.6 ± 0.3 mV (pH 3.0), 22.6 ± 0.6 to 29.7 ± 1.3 mV (p7.4)
and 32.6 ± 1.6 to 41.6 ± 1.0 mV (pH 11.0). As expected, the f-
potential shifted when the pH changed, being positive in the acidic
condition and negative in the neutral and basic conditions.
XRD patterns (Fig. 3) of the Sr-BGNPs after drying at 130 °C (but
prior to thermal stabilization) showed the presence of the cation
precursors. The XRD patterns showed amorphous halos once the
Sr-NPs were heated to 680 °C, indicating that calcium oxide
(CaO) and strontium oxide (SrO), were incorporated into the silica
networks after thermal stabilisation, agreeing with previous stud-
ies on cation incorporation [36,47–50]. Previous slice and view
studies, using FIB and TEM, on 500 nm SiO
2
-CaO NPs made using
the same method, showed a homogeneous distribution of Ca
throughout the NP [36]. As diffusion distances are smaller here,
it is likely that the Sr-BGNPs also have a homogeneous distribution,
although their small size made it impossible to do slice and view
imaging. Sr-BGNPs are likely to have the potential to generate a
sustained release of therapeutic cations because of the Sr being
part of an amorphous silica network.
3.2. Ion release profiles
Fig. 4 shows the dissolution profiles of each of the Sr-BGNPs in
PBS, ALF, and
a
-MEM media. The results demonstrate that the Si
content of the
a
-MEM media, ALF, and PBS increased as the incu-
bation period increased, confirming that soluble silica was released
from the BGNPs at a sustained rate over the immersion periods. Si
release was slowest in PBS, not reaching 20 mg/mL at 3 days. Si
release in ALF and
a
-MEM was faster and exceeded 50 mg/mL at
3 days. In ALF, the concentration of Sr increased sharply during
the first 24 h of immersion, after which it increased more slowly,
Fig. 1. TEM images of Sr-BGNPs: (a) SiO
2
-NPs (100 mol% SiO
2
); (b) 0%Sr-BGNPs (93 mol% SiO
2
, 7 mol% CaO); (c) 6%Sr-BGNPs (87 mol% SiO
2
, 7 mol% CaO, 6 mol% SrO); and (d)
14%Sr-BGNPs (83 mol% SiO
2
, 3 mol% CaO, 14 mol% SrO).
Fig. 2. Dynamic Light Scattering (DLS) results of 0%Sr-BGNPs, 6%Sr-BGNPs and 14%
Sr-BGNPs.
P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80 71
reaching a plateau at 48 h; this increase was significantly higher
for the 14% Sr-BGNPs. In
a
-MEM, the Sr-BGNPs rapidly released
Sr ions during the first 4 h, followed by a gradual release for a pro-
longed period. In PBS, the concentration of Sr increased slightly as a
function of time for the first 24 h of immersion before levelling off.
The medium type and pH affected Sr release: the total amount of Sr
release was greatest in the neutral
a
-MEM medium and was sim-
ilar for the two NP compositions (Table 3), which could be due to
the amino acids present in the medium chelating the Sr
2+
ions. In
contrast, the amount of Sr released in ALF was significantly higher
for the 14% Sr-BGNPs and similar to that measured in
a
-MEM. The
Ca ion concentrations in all three solutions remained approxi-
mately constant over time, after rising in the first 4 h of immersion.
Interestingly, in ALF solution, there was an increase in the Ca ion
concentration for the 6%Sr-BGNPs which was pronounced at 24
and 48 h. It is well known that low pH accelerates cation exchange
in aqueous media for all glasses. The reduction in the stability of
the NPs in an acidic environment is likely to be beneficial by trig-
gering the release of cations inside the lysosomes following cellular
uptake of the NPs. Table 3 showed that the total amount of Ca and
Table 2
Size and Zeta potential of Sr-BGNPs in water at three different pH values: 3.0, 7.4 and 11.0.
Samples DLS size distribution (nm) TEM size (nm) Zeta potential (mV)
pH 3.0 pH 7.4 pH 11.0
0%Sr-BGNPs 89.3 ± 2.9 75.1 ± 5.5 1.5 ± 0.1 22.6 ± 0.6 36.2 ± 0.9
6%Sr-BGNPs 91.2 ± 10.6 79.4 ± 7.7 3.2 ± 2.1 29.7 ± 1.3 32.6 ± 1.6
14%Sr-BGNPs 98.0 ± 9.2 83.2 ± 7.7 4.6 ± 0.3 29.2 ± 0.6 41.6 ± 1.0
Fig. 3. XRD patterns of (a) SiO
2
-NPs; (b) 0%Sr-BGNPs; (c) 6%Sr-BGNPs; and (d) 14%Sr-BGNPs before and after heat treatment process (heating at 680 °C).
72 P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80
Sr release from Sr-BGNPs in ALF and
a
-MEM were 3-times higher
than the amount released in the PBS. None of the Ca profiles show
a reduction in Ca content of the media within the 24 h that the
cells were exposed to the nanoparticles, indicating that there was
no calcium phosphate deposition.
TEM images (Fig. 5) show the morphological changes to the 14%
Sr-BGNPs following incubation in the three different media. When
the 14%Sr-BGNPs were immersed in PBS for 10 days, salt precipita-
tion on the surface of the particles was considerable in PBS (Fig. 5
(a)). Fig. 5(b) shows that in ALF the surfaces of the particles chan-
ged, developing a mottled morphology, indicating that the silicate
network of Sr-BGNPs became more unstable and degraded in the
acidic environment. In
a
-MEM (Fig. 5(c)), a shell with reduced con-
trast was observed around the particles, the particles had a mottled
appearance and necks were observed between the particles, fol-
lowing 24 h of incubation. These features all indicated that the par-
ticles had undergone degradation.
3.3. In vitro cytotoxicity
Our previous work showed that submicron BGNPs can be inter-
nalised by cells and degrade inside the cells [35,43]. Therefore, the
effect of NPs themselves (direct method) was evaluated, in which
the cell monolayer was exposed to NPs directly. A reduction in cell
viability by more than 30% is considered a cytotoxic effect (cell via-
bilities less than 70%) and this was used as a cut-off value to eval-
uate cytotoxicity of these particles (ISO 10993-5). To evaluate the
effect of the 6%Sr-BGNPs and 14%Sr-BGNPs on the viability of the
MC3T3-E1 cells, both NP compositions were introduced to
MC3T3-E1 cells for 24 h (pulse period), i.e. the media containing
non-internalised particles was removed and the cell culture was
continued for 0, 1, 3, and 7 days (chase period). Cells cultured on
tissue culture plates (TCP) served as controls. The results were con-
sistent with our previous report, which showed no toxicity at all
concentrations up to 250 mg/mL, but also little difference between
samples (Fig. 6 (a)). A significant reduction in cell viability was
observed with NP concentrations greater than 500 mg/mL for both
compositions of Sr-BGNPs. When cells were exposed to the Sr-
BGNPs with concentrations ranging from 0.01–250
l
g/mL, the cell
viability did not decrease significantly compared to the control
(TPC). The Stöber SiO
2
-NPs without Ca or Sr did reduce the cell via-
bility after 24 h at a concentration of 200
l
g/mL.
The effect of the BGNP ionic release products were also evalu-
ated (Fig. 6 (b)). The cell viability of the MC3T3-E1 cells increased
Fig. 4. Dissolution profiles of 6%Sr-BGNPs and 14%Sr-BGNPs in three different solutions: (a)
a
-MEM media at pH 7.4; (b) ALF at pH 4.5; (c) PBS at pH 7.4.
Table 3
Percentage elemental release from 6%Sr-BGNPs and 14%Sr-BGNPs after immersion in different media for 1 (
a
-MEM) and 10 days (PBS and ALF).
Samples PBS ALF
a
-MEM
% release
Si Ca Sr Si Ca Sr Si Ca Sr
6%Sr-BGNPs 2.32 1.50 8.17 4.62 7.44 16.13 5.46 3.47 22.94
14%Sr-BGNPs 3.20 0.97 3.68 4.71 3.52 20.38 5.95 4.85 23.56
P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80 73
Fig. 5. Bright-field TEM images of 14%Sr-BGNPs following immersion of the NPs in: (a) PBS for 10 days, (b) ALF for 10 days, and (c)
a
-MEM for 24 h (all scale bars 50 nm).
Fig. 6. Effect of (a) Sr-BGNPs and (b) their ionic release products, on the metabolic cell activity of the pre-osteoblastic cell line (MC3T3-E1) based on an MTT assay after 24 h
pulse followed by chase period in culture of 0, 1, 3, and 7 days (n = 6 per group). The metabolism of the cells treated with different concentrations (0.01–1000
l
g/mL).
*
(6%Sr-
BGNPs) and +(14%Sr-BGNPs) were statistically different from the control (TCP), p < 0.05.
**
(6%Sr-BGNPs) and ++(14%Sr-BGNPs) were statistically decreased from the control
(TCP), p<0.05.
74 P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80
significantly (p<.05) when treated with the Sr-BGNP ion release
products with 87 mol% SiO
2
, 7 mol% CaO, 6 mol% SrO
2
(6% Sr-
BGNPs) and 83 mol% SiO
2
, 3 mol% CaO, 14 mol% SrO
2
(14% Sr-
BGNPs) at 200 and 250 mg/mL after 3 and 7 days, compared to
the cells in the other groups and compared to cells under the basal
condition. The concentration of Si, Ca and Sr ions released into the
media is shown in Supplementary information (Table S2). The con-
centration of Sr required to enhance the activity of the MC3T3-E1
cells was in the range of 2 to 6 mg/mL.
3.4. Osteogenic differentiation
To evaluate therapeutic efficacy of Sr-BGNPs and the ion release
products for bone regeneration, several early and late bone forma-
tion markers were investigated. Fig.7 shows the expression of ALP
activity in the extracellular matrix of the MC3T3-E1 cells treated
with 6%Sr-BGNPs and 14%Sr-BGNPs, and their ionic release prod-
ucts, following 21 days of culture (media contained particles or dis-
solution ions at every media change), compared to cells cultured
under the basal and osteogenic media (basal media plus osteogenic
supplements). There was a significant increase in the ALP activity
of the MCT3T-E1 cells treated with 14%Sr-BGNPs compared to
the 6%Sr-BGNPs in the basal medium. No difference was seen for
the different Sr contents when osteogenic supplements were used.
The ionic release products from both the 6%Sr-BGNPs and 14%Sr-
BGNPs also stimulated the ALP activity of the cultured MC3T3-E1
cells in the basal condition, suggesting that the ALP simulation
resulted from an effect of the ions released from the BGNPs as they
dissolved.
To illustrate the effect of the NPs on osteogenic differentiation
of the MC3T3-E1 cells in vitro, markers associated with early and
late osteogenic differentiation, including Col1a1, OSC and OSP
were investigated using immunohistochemistry (IHC) staining.
Fig. 8 shows that Col1a1, OSC, and OSP were expressed on the
MC3T3-E1 cells following 21 days of culture. In the basal condi-
tion, the late osteogenic differentiation markers, OSC and OSP,
were more overtly expressed by the 14%Sr-BGNPs and their ionic
release products. In contrast, Col1a1 expression was only slightly
enhanced in the basal condition, but was enhanced by the osteo-
genic supplements. Both OSP and Colla1 are proteins associated
with extracellular matrix formation [51,52]. Thus, the data show
that the Sr-BGNPs and their dissolution products could affect both
mineralisation and extracellular matrix formation, but highlight
that osteogenic supplements are also essential for early extracel-
lular matrix formation.
Calcified nodule formation was observed using Alizarin Red S
staining at 7, 14 and 21 days. MC3T3-E1 cells were periodically
incubated with 6% Sr-BGNPs and 14%Sr-BGNPs and their ionic
release products. After 7 days, no difference was observed between
the basal and osteogenic conditions, with either the 6%Sr-BGNPs or
14%Sr-BGNPs or the ionic release products (Fig. 9). After two and
three weeks in culture, mineralization of the MC3T3-E1 cells trea-
ted with 6%Sr-BGNPs and 14%Sr-BGNPs and also their ionic release
products, increased significantly compared with the cells cultured
in the control conditions (basal without BGNPs (Fig. 9(A)) and
osteogenic condition without BGNPs (Fig. 9(B)). This effect of the
BGNPs on mineralisation was significantly more pronounced in
the cells cultured with osteogenic medium. Quantitative analysis
of the Alizarin Red S staining confirmed an increase in calcium for-
mation of the treated MC3T3-E1 cells exposed to both the 6%Sr-
BGNPs and 14%Sr-BGNPs compared to the control cells (untreated
cells in both the basal and osteogenic conditions) (Fig. S3, Supple-
mentary information). This increase in calcium formation was also
seen for the Sr-BGNP ionic release products. These results indicate
that both the 6%Sr-BGNPs and 14%Sr-BGNPs enhanced mineralisa-
tion without the osteogenic supplements, and that this effect can
be attributed to their ionic release products.
The final aim was to track whether BGNPs had been internalised
by the cells and to understand the mechanism by which the BGNPs
were internalised. Cell viability was not affected, remaining above
70% of the positive control, by functionalization of the Sr-BGNPs
with APTES and fluorescently labelling with FITC (Fig. S4, Supple-
mentary information).
Confocal microscopy images (2D in Fig. 10 and z-stacks shown
in the video in Supplementary information) showed that the FITC-
14%Sr-BGNPs were internalised by the cells and were present in
the cytoplasm of the MCT3T-E1 after 24 h of incubation. Fig. 11
shows the fluorescence intensity of the FITC-14%Sr-BGNPs in the
MC3T3-E1 cells, confirming that the NPs were internalised by the
cells. There was a significant decrease in uptake of the FITC-14%
Sr-BGNPs by the cells treated with CytD, suggesting that a
clathrin-dependent endocytosis pathway was responsible for
uptake of the FITC-14%Sr-BGNPs by the MC3T3-E1 cells. However,
some particles were still internalised in the cells treated by CytD
and uptake of FITC-14%Sr-BGNPs was nearly completely sup-
pressed when the MC3T3-E1 cells were treated with all five endo-
cytosis inhibitors, indicating that the MCT3T3-E1 cells also used
alternative pathways to uptake particles. These data indicate that
the FITC-14%Sr-BGNPs enter the cells via a number of endocytosis
pathways.
Fig. 7. Staining for ALP activity, which indicates the differentiation of MC3T3-E1 cells grown in media containing 6%Sr-BGNPs and 14%Sr-BGNPs or their ionic release
products (Sr-BGNP concentration at 250 mg/mL). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80 75
Fig. 8. Fluorescence image, DAPI (blue), Col1a1, OSC and OSP staining (all green) of MC3T3-E1 cells exposure to 14%Sr-BGNPs and their ionic release products (Sr-BGNP
concentration at 250 mg/mL) in basal and osteogenic conditions (3-week culture period). Scale bar 150 mm. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
76 P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80
4. Discussion
Monodispersed Sr-BGNPs were successfully synthesised using
the modified Stӧber process [36,37]. These nanoparticles were
amorphous and spherical with diameters of 90 ± 10 nm. To
improve the osteogenic response to the BGNPs, additional cations,
such as Sr
2+
were introduced during the sol-gel process [16,53,54].
An important step is to incorporate cations, by calcination above
450 °C, while maintaining particle morphology and diameter
[36,49]. Here, Ca(NO
3
)
2
and Sr(NO
3
)
2
were used as the cation pre-
cursors to incorporate Ca
2+
and Sr
2+
into the silica networks. Dur-
ing drying, Ca(NO
3
)
2
and Sr(NO
3
)
2
deposits onto the secondary
particle surface of the silica [49]. During thermal stabilisation, Ca
and Sr were incorporated into the silica network as network mod-
ifiers, as confirmed by XRD (Fig. 3), without affecting particle size,
distribution and morphology (Table 1, Figs. 1and 2). When the per-
centage of Sr substitution increased from 0 to 14 mol%, the total
amount of network modifiers increased from approximately 7%
to 17%, even though the nominal concentration of potential net-
work modifiers was constant. The reason behind this is that the
ionic radii of Sr (1.12 A°) is larger than Ca (0.99 A°), leading
to a more open silica network, which also leads to more rapid dis-
solution (Fig. 4).
Delivery of Sr inside cells may form part of a therapeutic treat-
ment for osteoporosis, so it is important to understand how the
particles are internalised and the behaviour of the particles in
physiological solutions outside and inside cells. One hypothesis
was that when cells uptake Sr-BGNPs, they form lysosomal vesicles
that encapsulate the particles. Sr-BGNPs would therefore be sub-
jected to acidic conditions inside the cell lysosome. Here, in vitro
dissolution showed that more cations dissolved from the Sr-
BGNPs in ALF solution at pH 4.5 than in the media at pH 7.4,
because cation exchange is promoted under the low pH environ-
ment (Fig. 4). Dissolution in
a
-MEM was more rapid than in PBS
at the same pH, which is likely due to organic molecules having
an affinity (chelation) to cations in the glass.
The in vitro cell viability of MC3T3-E1 cells treated with Sr-
BGNPs was not significantly changed when the particle concentra-
tion increased up to 250
l
g/mL, but Sr-BGNPs concentrations at or
above 500
l
g/mL caused cells death after 3 days in the direct
method. The results suggested BGNPs containing cations had
higher cellular biocompatibility compared to the control of the
much less degradable SiO
2
-NPs (Fig. 6). The relative cell viability
increased when the Sr concentration in the media was in a range
of 2 to 6 mg/mL. Following on from previous research on 45S5 Bio-
glass ionic release products, which induce osteogenic differentia-
tion of osteoblasts without adding osteogenic supplements [8,55–
57], we hypothesised that the media containing the ionic release
products of the Sr-BGNPs would stimulate osteogenic differentia-
Fig. 9. Alizarin red staining for calcified nodule formation from MC3T3-E1 cells
treated with 6% Sr-BGNPs and 14% Sr-BGNPs or their ionic release products (Sr-
BGNP concentration at 250 mg/mL) under: (A) the basal condition and (B) the
osteogenic condition. Scale bar = 400
l
m. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.)
Video
P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80 77
tion of MC3T3-E1 with no added osteogenic supplements. MC3T3-
E1 cells treated under basal conditions with particles or their ionic
release products did indeed show an increase in ALP activity
(Fig. 7) and calcium deposition compared to the control (Fig. 9).
Osteogenesis associated markers, including Col1a1, OSC, and OSP,
were also evaluated (Fig. 8). MC3T3-E1 cells treated with the Sr-
BGNPs and their ionic release products under basal conditions
showed a strong green fluorescent intensity of Col1a1, OSC, and
OSP. The results indicate that Sr-BGNPs stimulate osteogenic dif-
ferentiation and that this effect arises due to the ionic release prod-
ucts of the Sr-BGNPs.
To identify the uptake route used by the Sr-BGNPs, uptake by
MC3T3-E1 cells was investigated under specific inhibitors using
microplate reader and confocal fluorescent microscope, as sug-
gested in previous studies [58,59]. The 14%Sr-BGNPs were proba-
bly predominantly taken up via the clathrin-dependent
endocytosis pathway in the MC3T3-E1 cells (Fig. 11), which is con-
sistent with previous findings that clathrin-dependent endocytosis
is the main pathway for particle size up to 200 nm [59]. However,
the 14%Sr-BGNPs were also internalised into cells treated with all
of the inhibitors. The possible explanation for this might be
because either the concentrations of inhibitors were not suited
Fig. 10. Confocal microscopy images showing internationalisation of the FITC-14%Sr-BGNPs (concentration at 250 mg/mL) by the MC3T3-E1 after pre-treatment with
endocytosis inhibitors; Wor (wortmannin), Ami (amiloride hydrochloride hydrate), Chlor (chlorpromazine), Gen (genistein), CytD (cytochalasin D). Scale bar = 20 mm. Nuclei
were stained with DAPI. F-actin filaments were stained with rhodamine phalloidin. (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)
Fig. 11. Effect of endocytosis inhibitors on FITC-14%Sr-BGNPs (concentration at 250 mg/mL) internalisation by MCT3T3-E1 cells; Wor (wortmannin), Ami (amiloride
hydrochloride hydrate), Chlor (chlorpromazine), Gen (genistein), CytD (cytochalasin D).
78 P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80
for MC3T3-E1 cells, leading to not completely blocked pathways, or
that the uptake of these particles by the cells occurred through
mixed pathways. The results agree with previous studies on
SiO
2
-CaO NPs in stem cells [35], and it can be speculated that
14%Sr-BGNPs entered into the cells through endocytosis pathways.
The green fluorescent intensity of the confocal microscopy images
(Fig. 10 and the 3D video in Supplementary information) also
showed internalisation of 14%Sr-BGNPs by the MC3T3-E1 cells,
confirming intracellular ion delivery inside the cells.
5. Conclusions
This study reports the osteogenic response of MC3T3-E1 cells
treated with monodispersed Sr-BGNPs SiO
2
-NPs synthesised using
the modified Stöber process and incorporating Ca and Sr. Incorpo-
ration of Ca and Sr into the silica network did not affect the size
and shape of the particles. Culture of the Sr-BGNPs with MC3T3-
E1 cells did not alter the viability of the cells up to concentration
250
l
g/mL. The dissolution products of the Sr-BGNPs were non-
toxic at all concentrations. Ion release products from the 6%Sr-
BGNPs (87 mol% SiO
2
, 7 mol% CaO, 6 mol% SrO) and 14%Sr-BGNPs
(83 mol% SiO
2
, 3 mol% CaO, 14 mol% SrO) at BGNP concentrations
of 200 and 250
l
g/mL enhanced MC3T3-E1 proliferation up to 7
days in vitro. Sr-BGNPs concentrations at or above 500
l
g/mL had
adverse effects on MC3T3-E1 cell viability after 3 days in the direct
method. The 6%Sr-BGNPs and 14%Sr-BGNPs and their ionic release
products had the ability to stimulate an osteogenic response with-
out adding osteogenic supplements in the culture system and this
effect could be attributed to their ionic release products. Crucially,
the data show that the Sr-BGNPs could affect both mineralisation
and extracellular matrix formation, but highlight that osteogenic
supplements are also essential for early extracellular matrix forma-
tion. The results show both an increase in mineralisation and
expression of proteins associated with collagen production. As Sr
content in the NPs and in the dissolution products increased,
ALP, OSC and OSP expression increased. The exact mechanisms
for the cells to uptake Sr-BGNPs remained indefinable, but most
likely based on the clathrin-dependent endocytosis pathway.
These in vitro results revealed that 6%Sr-BGNPs and 14%Sr-BGNPs
at a concentration of 250
l
g/mL promoted osteogenic response
whist maintaining cell proliferation, which is likely to be beneficial
to use as an inorganic drug delivery for bone regeneration
applications.
Acknowledgements
The research work is supported by the Royal Thai Government
and EPSRC (EP/M004414/1) and an Elsie Widdowson Fellowship to
AEP. Raw data is available from rdm-enquiries@imperial.ac.uk.
Appendix A. Supplementary data
Supplementary data associated with this article can be found, in
the online version, at https://doi.org/10.1016/j.actbio.2017.11.008.
References
[1] J.R. Jones, D.S. Brauer, L. Hupa, D.C. Greenspan, Bioglass and bioactive glasses
and their impact on healthcare, Int. J. Appl. Glass Sci. 7 (2016) 423–434.
[2] L.L. Hench, R.J. Splinter, W.C. Allen, T.K. Greenlee, Bonding mechanisms at the
interface of ceramic prosthetic materials, J. Biomed. Mater. Res. Symp. 2 (1971)
117–141.
[3] L.L. Hench, The story of Bioglass
Ò
, J. Mater. Sci. Mater. Med. 17 (2006) 967–978.
[4] L.L. Hench, Opening paper 2015-some comments on bioglass: four eras of
discovery and development, Biomed. Glasses 1 (2015) 1–11.
[5] I.D. Xynos, M.V. Hukkanen, J.J. Batten, L.D. Buttery, L.L. Hench, J.M. Polak,
Bioglass 45S5 stimulates osteoblast turnover and enhances bone formation In
vitro: implications and applications for bone tissue engineering, Calcified
Tissue Int. 67 (2000) 321–329.
[6] A. Hoppe, N.S. Guldal, A.R. Boccaccini, A review of the biological response to
ionic dissolution products from bioactive glasses and glass-ceramics,
Biomaterials 32 (2011) 2757–2774.
[7] J.R. Jones, Review of bioactive glass: from hench to hybrids, Acta Biomater. 9
(2013) 4457–4486.
[8] L.L. Hench, Genetic design of bioactive glass, J. Eur. Ceram. Soc. 29 (2009)
1257–1265.
[9] A. Hoppe, A.R. Boccaccini, Biological impact of bioactive glasses and their
dissolution products, Front. Oral Biol. 17 (2015) 22–32.
[10] M.N. Rahaman, D.E. Day, B.S. Bal, Q. Fu, S.B. Jung, L.F. Bonewald, A.P. Tomsia,
Bioactive glass in tissue engineering, Acta Biomat. 7 (2011) 2355–2373.
[11] S.M. Rabiee, N. Nazparvar, M. Azizian, D. Vashaee, L. Tayebi, Effect of ion
substitution on properties of bioactive glasses: a review, Ceram. Int. 41 (2015)
7241–7251.
[12] Y. Zhou, M. Shi, J.R. Jones, Z. Chen, J. Chang, C. Wu, Y. Xiao, Strategies to direct
vascularisation using mesoporous bioactive glass-based biomaterials for bone
regeneration, Int. Mater. Rev. (2017) 1–23.
[13] P.J. Marie, P. Ammann, G. Boivin, C. Rey, Mechanisms of action and therapeutic
potential of strontium in bone, Calcified Tissue Int. 69 (2001) 121–129.
[14] P.J. Meunier, C. Roux, E. Seeman, S. Ortolani, J.E. Badurski, T.D. Spector, J.
Cannata, A. Balogh, E.M. Lemmel, S. Pors-Nielsen, R. Rizzoli, H.K. Genant, J.Y.
Reginster, The effects of strontium ranelate on the risk of vertebral fracture in
women with postmenopausal osteoporosis, N. Engl. J. Med. 350 (2004) 459–
468.
[15] J. Liu, S.C.F. Rawlinson, R.G. Hill, F. Fortune, Strontium-substituted bioactive
glasses in vitro osteogenic and antibacterial effects, Dent. Mater. 32 (2016)
412–422.
[16] J. Lao, J.M. Nedelec, E. Jallot, New strontium-based bioactive glasses:
physicochemical reactivity and delivering capability of biologically active
dissolution products, J. Mater. Chem. 19 (2009) 2940–2949.
[17] A. Hoppe, B. Sarker, R. Detsch, N. Hild, D. Mohn, W.J. Stark, A.R. Boccaccini, In
vitro reactivity of Sr-containing bioactive glass (type 1393) nanoparticles, J.
Non-Cryst. Solids 387 (2014) 41–46.
[18] H. Autefage, E. Gentleman, E. Littmann, M.A.B. Hedegaard, T. Von Erlach, M.
O’Donnell, F.R. Burden, D.A. Winkler, M.M. Stevens, Sparse feature selection
methods identify unexpected global cellular response to strontium-containing
materials, Proc. Natl. Acad. Sci. 112 (2015) 4280–4285.
[19] M.D. O’Donnell, R.G. Hill, Influence of strontium and the importance of glass
chemistry and structure when designing bioactive glasses for bone
regeneration, Acta Biomater. 6 (2010) 2382–2385.
[20] J. Massera, L. Hupa, Influence of SrO substitution for CaO on the properties of
bioactive glass S53P4, J. Mater. Sci. – Mater. Med. 25 (2014) 657–668.
[21] Y.C. Fredholm, N. Karpukhina, D.S. Brauer, J.R. Jones, R.V. Law, R.G. Hill,
Influence of strontium for calcium substitution in bioactive glasses on
degradation, ion release and apatite formation, J. R. Soc. Interface 9 (2012)
880–889.
[22] M.E. Santocildes-Romero, A. Crawford, P.V. Hatton, R.L. Goodchild, I.M. Reaney,
C.A. Miller, The osteogenic response of mesenchymal stromal cells to
strontium-substituted bioactive glasses, J. Tissue Eng. Regen. Med. 9 (2015)
619–631.
[23] P. Mora-Raimundo, M. Manzano, M. Vallet-Regí, Nanoparticles for the
treatment of osteoporosis, AIMS Bioeng. 4 (2017) 259–274.
[24] M. Vallet-Regi, A.J. Salinas, Chapter 17 mesoporous bioactive glasses in tissue
engineering and drug delivery, in: Bioactive Glasses: Fundamentals,
Technology and Applications, The Royal Society of Chemistry, 2017, pp. 393–
419.
[25] J.P. Fan, P. Kalia, L. Di Silvio, J. Huang, In vitro response of human osteoblasts to
multi-step sol–gel derived bioactive glass nanoparticles for bone tissue
engineering, Mater. Sci. Eng., C 36 (2014) 206–214.
[26] C. Vichery, J.-M. Nedelec, Bioactive glass nanoparticles: from synthesis to
materials design for biomedical applications, Materials 9 (2016) 288.
[27] A.M. El-Kady, M.M. Farag, A.M.I. El-Rashedi, Bioactive glass nanoparticles
designed for multiple deliveries of lithium ions and drugs: curative and
restorative bone treatment, Eur. J. Pharm. Sci. 91 (2016) 243–250.
[28] C. Wu, W. Fan, M. Gelinsky, Y. Xiao, P. Simon, R. Schulze, T. Doert, Y. Luo, G.
Cuniberti, Bioactive SrO-SiO2 glass with well-ordered mesopores:
characterization, physiochemistry and biological properties, Acta Biomater. 7
(2011) 1797–1806.
[29] M. Chen, A. von Mikecz, Formation of nucleoplasmic protein aggregates
impairs nuclear function in response to SiO2 nanoparticles, Exp. Cell Res. 305
(2005) 51–62.
[30] M.V. Park, D.P. Lankveld, H. Van Loveren, W.H. De Jong, The status of in vitro
toxicity studies in the risk assessment of nanomaterials, Nanomedicine
(London, England) 4 (2009) 669–685.
[31] D.M. Huang, T.H. Chung, Y. Hung, F. Lu, S.H. Wu, C.Y. Mou, M. Yao, Y.C. Chen,
Internalization of mesoporous silica nanoparticles induces transient but not
sufficient osteogenic signals in human mesenchymal stem cells, Toxicol. Appl.
Pharmacol. 231 (2008) 208–215.
[32] X. Lu, J. Qian, H. Zhou, Q. Gan, W. Tang, J. Lu, Y. Yuan, C. Liu, In vitro cytotoxicity
and induction of apoptosis by silica nanoparticles in human HepG2 hepatoma
cells, Int. J. Nanomed. 6 (2011) 1889–1901.
[33] Y. Wu, W. Tang, P. Wang, C. Liu, Y. Yuan, J. Qian, Cytotoxicity and cellular
uptake of amorphous silica nanoparticles in human cancer cells, Part. Part.
Syst. Charact. 32 (2015) 779–787.
P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80 79
[34] T.G. Iversen, T. Skotland, K. Sandvig, Endocytosis and intracellular transport of
nanoparticles: present knowledge and need for future studies, Nano Today 6
(2011) 176–185.
[35] O. Tsigkou, S. Labbaf, M.M. Stevens, A.E. Porter, J.R. Jones, Monodispersed
bioactive glass submicron particles and their effect on bone marrow and
sdipose tissue-derived stem cells, Adv. Healthc. Mater. 3 (2014) 115–125.
[36] S.L. Greasley, S.J. Page, S. Sirovica, S. Chen, R.A. Martin, A. Riveiro, J.V. Hanna, A.
E. Porter, J.R. Jones, Controlling particle size in the Stöber process and
incorporation of calcium, J. Colloid Interface Sci. 469 (2016) 213–223.
[37] P. Naruphontjirakul, S.L. Greasley, S. Chen, A.E. Porter, J.R. Jones,
Monodispersed strontium containing bioactive glass nanoparticles and
MC3T3-E1 cellular response, Biomed. Glasses 2 (2016) 72–81.
[38] J. Zhu, L. Liao, L. Zhu, P. Zhang, K. Guo, J. Kong, C. Ji, B. Liu, Size-dependent
cellular uptake efficiency, mechanism, and cytotoxicity of silica nanoparticles
toward HeLa cells, Talanta 107 (2013) 408–415.
[39] W.-K. Oh, S. Kim, M. Choi, C. Kim, Y.S. Jeong, B.-R. Cho, J.-S. Hahn, J. Jang,
Cellular uptake, cytotoxicity, and innate immune response of silicatitania
hollow nanoparticles based on size and surface functionality, ACS Nano 4
(2010) 5301–5313.
[40] D. Napierska, L.C.J. Thomassen, V. Rabolli, D. Lison, L. Gonzalez, M. Kirsch-
Volders, J.A. Martens, P.H. Hoet, Size-dependent cytotoxicity of monodisperse
silica nanoparticles in human endothelial cells, Small 5 (2009) 846–853.
[41] T. Yu, A. Malugin, H. Ghandehari, Impact of silica nanoparticle design on
cellular toxicity and hemolytic activity, ACS Nano 5 (2011) 5717–5728.
[42] A.L.B. Maçon, T.B. Kim, E.M. Valliant, K. Goetschius, R.K. Brow, D.E. Day, A.
Hoppe, A.R. Boccaccini, I.Y. Kim, C. Ohtsuki, T. Kokubo, A. Osaka, M. Vallet-Regí,
D. Arcos, L. Fraile, A.J. Salinas, A.V. Teixeira, Y. Vueva, R.M. Almeida, M. Miola, C.
Vitale-Brovarone, E. Verné, W. Höland, J.R. Jones, A unified in vitro evaluation
for apatite-forming ability of bioactive glasses and their variants, J. Mater. Sci.
– Mater. Med. 26 (2015) 115.
[43] S. Labbaf, O. Tsigkou, K.H. Muller, M.M. Stevens, A.E. Porter, J.R. Jones, Spherical
bioactive glass particles and their interaction with human mesenchymal stem
cells in vitro, Biomaterials 32 (2011) 1010–1018.
[44] A.N. Zelikin, Q. Li, F. Caruso, Degradable polyelectrolyte capsules filled with
oligonucleotide sequences, Angew. Chem. Int. Ed. 45 (2006) 7743–7745.
[45] W. Xu, J. Riikonen, T. Nissinen, M. Suvanto, K. Rilla, B. Li, Q. Wang, F. Deng, V.-P.
Lehto, Amine surface modifications and fluorescent labeling of thermally
stabilized mesoporous silicon nanoparticles, J. Phys. Chem. C 116 (2012)
22307–22314.
[46] J. Linares, M.C. Matesanz, M. Vila, M.J. Feito, G. Gonçalves, M. Vallet-Regí, P.A.A.
P. Marques, M.T. Portolés, Endocytic mechanisms of graphene oxide
nanosheets in osteoblasts, hepatocytes and macrophages, ACS Appl. Mater.
Interf. 6 (2014) 13697–13706.
[47] Z. Lin, J.R. Jones, J.V. Hanna, M.E. Smith, A multinuclear solid state NMR
spectroscopic study of the structural evolution of disordered calcium silicate
sol-gel biomaterials, Phys. Chem. Chem. Phys. 17 (2015) 2540–2549.
[48] B. Yu, C.A. Turdean-Ionescu, R.A. Martin, R.J. Newport, J.V. Hanna, M.E. Smith, J.
R. Jones, Effect of calcium source on structure and properties of sol-gel derived
bioactive glasses, Langmuir 28 (2012) 17465–17476.
[49] S. Lin, C. Ionescu, K.J. Pike, M.E. Smith, J.R. Jones, Nanostructure evolution and
calcium distribution in sol-gel derived bioactive glass, J. Mater. Chem. 19
(2009) 1276–1282.
[50] L.J. Skipper, F.E. Sowrey, D.M. Pickup, K.O. Drake, M.E. Smith, P. Saravanapavan,
L.L. Hench, R.J. Newport, The structure of a bioactive calcia-silica sol-gel glass,
J. Mater. Chem. 15 (2005) 2369–2374.
[51] M.D. McKee, A. Nanci, Osteopontin: an interfacial extracellular matrix protein
in mineralized tissues, Connect. Tissue Res. 35 (1996) 197–205.
[52] L.A. Strobel, N. Hild, D. Mohn, W.J. Stark, A. Hoppe, U. Gbureck, R.E. Horch, U.
Kneser, A.R. Boccaccini, Novel strontium-doped bioactive glass nanoparticles
enhance proliferation and osteogenic differentiation of human bone marrow
stromal cells, J. Nanopart. Res. 15 (2013) 1780.
[53] J. Lao, E. Jallot, J.-M. Nedelec, Strontium-delivering glasses with enhanced
bioactivity: a new biomaterial for antiosteoporotic applications?, Chem Mater.
20 (2008) 4969–4973.
[54] J. Isaac, J. Nohra, J. Lao, E. Jallot, J.M. Nedelec, A. Berdal, J.M. Sautier, Effects of
strontium-doped bioactive glass on the differentiation of cultured osteogenic
cells, Eur. Cell Mater. 21 (2011) 130–143.
[55] O. Tsigkou, J.R. Jones, J.M. Polak, M.M. Stevens, Differentiation of fetal
osteoblasts and formation of mineralized bone nodules by 45S5 Bioglass
Ò
conditioned medium in the absence of osteogenic supplements, Biomaterials
30 (2009) 3542–3550.
[56] G. Jell, I. Notingher, O. Tsigkou, P. Notingher, J.M. Polak, L.L. Hench, M.M.
Stevens, Bioactive glass-induced osteoblast differentiation: a noninvasive
spectroscopic study, J. Biomed. Mater. Res. A 86 (2008) 31–40.
[57] I.D. Xynos, A.J. Edgar, L.D.K. Buttery, L.L. Hench, J.M. Polak, Ionic products of
bioactive glass dissolution increase proliferation of human osteoblasts and
induce insulin-like growth factor II mRNA expression and protein synthesis,
Biochem. Biophys. Res. Commun. 276 (2000) 461–465.
[58] J. Blechinger, A.T. Bauer, A.A. Torrano, C. Gorzelanny, C. Brauchle, S.W.
Schneider, Uptake kinetics and nanotoxicity of silica nanoparticles are cell
type dependent, Small 9 (3970–80) (2013) 06.
[59] D. Vercauteren, R.E. Vandenbroucke, A.T. Jones, J. Rejman, J. Demeester, S.C. De
Smedt, N.N. Sanders, K. Braeckmans, The use of inhibitors to study endocytic
pathways of gene carriers: optimization and pitfalls, Mol. Ther. 18 (2010) 561–
569.
80 P. Naruphontjirakul et al. / Acta Biomaterialia 66 (2018) 67–80
... Their findings on ALP activity in the MC3T3-E1 cell line were in line with the results of this study, showing a similar ALP activity for the control group and lower concentrations of nanoparticles. Nanoparticles are known to act as cell signaling cues through therapeutic ions such as silicon, calcium, and gallium to induce osteogenic differentiation, promoting bone regeneration [23,24,28,76]. To summarize, the addition of higher concentrations of NPs-Ca and NPs-Ca-Ga to the osMEM initiated an earlier osteoblast differentiation, leading to a decrease in cell proliferation. ...
... Due to their adaptability to the human body and physiochemical characteristics, they have the potential for use in treatment of periodontal and endodontic infections, antimicrobial tissue-engineering, orthopedic implants, in vivo gene transfer, protein encapsulation, and antibacterial coating for metal implants (Sadalage et al. 2021;Alayed et al. 2022;Lina et al. 2023). CaO-NPs have also been explored as valuable drug-delivery vehicles that are capable of delivering anticancer, anti-leishmanial, and anti-diabetic drugs (Naruphontjirakul et al. 2018). CaO-NPs have also been used to reduce heavy metal toxicity (Cd 2+ , Pb 2+ , Cr 3+ , Fe 3+ , and As 3+ ) in soil, hence protecting plant roots from oxidative damage (Cai et al. 2010;Jalu et al. 2021;Nazir et al. 2022a, b;Raza et al. 2024). ...
Article
Full-text available
Herein, we have investigated and compared biocompatible calcium oxide nanoparticles (CaO-NPs) with different surface chemistries for biological properties. In the study, green synthesis is achieved using an aqueous extract of Potentilla bifurca while chemical synthesized polyethylene glycol (PEG) and cetyltrimethylammonium bromide (CTAB)-coated CaO-NPs are prepared using a facile coprecipitation approach. The nanoparticles are characterized using different techniques including, X-ray diffraction (XRD), Fourier-transformed infrared spectroscopy (FTIR), UV-visible spectroscopy (UV–Vis), scanning electron microscopy (SEM), and energy-dispersive X-ray analysis (EDX). Moreover, yield and pH-responsive dispersion studies are also carried out. After comprehensive characterization, the NPs are explored for potential biological properties such as antibacterial, antifungal, antileshmanial, antioxidant, and biocompatibility. Our study reveals that green synthesis results in CaO-NPs with high antioxidant and dispersion properties while chemical synthesis in particular CTAB-coated CaO-NPs exhibits remarkably high antibacterial, antifungal, and leishmancidal properties. For instance, the NPs resulted in 25 ± 1.4 mm and 16 ± 0.4 mm zone of inhibition (ZOI) against Bacillus subtilis and Aspergillus flavus,, respectively, while 81.10% and 79.34% inhibition against the promastigote and amastigote forms of the leishmania tropica. However, none of the NPs displayed hemolytic behavior, affirming the biocompatible nature of all the NPs. Our study thus concludes that surface characteristics play a vital role in defining the biological properties of CaO-NPs and the NPs could be tailored to harness the required biological properties.
... Sr enhanced the metabolic activity of osteoblasts, promoted cell growth, and increased alkaline phosphatase (ALP) activity [42]. Sr possess the capability to boost bone formation by promoting osteoblast function and to impede bone resorption by suppressing osteoclast activity [43,44]. Zn was doped into MBGNs to stimulate osteoblast activity, enhance bone formation, and enhance antimicrobial activity. ...
Article
Full-text available
Mesoporous bioactive glass nanoparticles (MBGNs) have attracted significant attention as multifunctional nanocarriers for various applications in both hard and soft tissue engineering. In this study, multifunctional strontium (Sr)- and zinc (Zn)-containing MBGNs were successfully synthesized via the microemulsion-assisted sol–gel method combined with a cationic surfactant (cetyltrimethylammonium bromide, CTAB). Sr-MBGNs, Zn-MBGNs, and Sr-Zn-MBGNs exhibited spherical shapes in the nanoscale range of 100 ± 20 nm with a mesoporous structure. Sr and Zn were co-substituted in MBGNs (60SiO2-40CaO) to induce osteogenic potential and antibacterial properties without altering their size, morphology, negative surface charge, amorphous nature, mesoporous structure, and pore size. The synthesized MBGNs facilitated bioactivity by promoting the formation of an apatite-like layer on the surface of the particles after immersion in Simulated Body Fluid (SBF). The effect of the particles on the metabolic activity of human mesenchymal stem cells was concentration-dependent. The hMSCs exposed to Sr-MBGNs, Zn-MBGNs, and Sr-Zn-MBGNs at 200 μg/mL enhanced calcium deposition and osteogenic differentiation without osteogenic supplements. Moreover, the cellular uptake and internalization of Sr-MBGNs, Zn-MBGNs, and Sr-Zn-MBGNs in hMSCs were observed. These novel particles, which exhibited multiple functionalities, including promoting bone regeneration, delivering therapeutic ions intracellularly, and inhibiting the growth of Staphylococcus aureus and Escherichia coli, are potential nanocarriers for bone regeneration applications.
Article
Articular cartilage defects are a global challenge, causing substantial disability. Repairing large defects is problematic, often exceeding cartilage's self‐healing capacity and damaging bone structures. To tackle this problem, we develop a scaffold‐mediated therapeutic ion delivery system. These scaffolds are constructed from poly(ε‐caprolactone) and strontium (Sr)‐doped bioactive nanoglasses (SrBGn), creating a unique hierarchical structure featuring macro‐pores from 3D printing, micro‐pores, and nano‐topologies due to SrBGn integration. The SrBGn‐embedded scaffolds (SrBGn‐μCh) release Sr, silicon (Si), and calcium (Ca) ions, which improve chondrocyte activation, adhesion, proliferation, and maturation‐related gene expression. This multiple ion delivery significantly affects metabolic activity and maturation of chondrocytes. Importantly, Sr ions may play a role in chondrocyte regulation through the Notch signaling pathway. Notably, the scaffold's structure and topological cues expedite the recruitment, adhesion, spreading, and proliferation of chondrocytes and bone marrow‐derived mesenchymal stem cells. Si and Ca ions accelerate osteogenic differentiation and blood vessel formation, while Sr ions enhance the polarization of M2 macrophages. Our findings show that SrBGn‐μCh scaffolds accelerate osteochondral defect repair by delivering multiple ions and providing structural/topological cues, ultimately supporting host cell functions and defect healing. This scaffold holds great promise for osteochondral repair applications. This article is protected by copyright. All rights reserved
Article
Estrogen deficiency is one of the main causes for postmenopausal osteoporosis. Current osteoporotic therapies are of high cost and associated with serious side effects. So there is an urgent need for cost effective anti-osteoporotic agents. Anti-osteoporotic activity of Litsea glutinosa extract (LGE) is less explored. Moreover, its role in fracture healing and mechanism of action is still unknown. In the present study we explore the osteoprotective potential of LGE in osteoblast cells and fractured and ovariectomized (Ovx) mice models. ALP, MTT and mineralization assays revealed that LGE treatment increased osteoblast cell differentiation, viability and mineralization. LGE treatment at 0.01µg increased the expression of BMP2, PSMAD, RUNX2 and Type 1 col. LGE also mitigated RANKL induced osteoclastogenesis. Next, drill hole injury Balb/C mice model was treated with LGE for 12 days. Micro-CT analysis and calcein labeling at the fracture site showed that LGE (20mg/kg) enhanced new bone formation and bone regeneration, also increased expression of BMP2/SMAD1 signaling genes at fracture site. Ovx mice were treated with LGE for 1 month. µCT analysis indicated that, treatment of LGE at 20mg/kg dose prevented the alteration in bone microarchitecture, maintained bone mineral density and bone mineral content. Treatment also increased bone strength and restored the bone turnover markers. Furthermore, in bone samples LGE increased osteogenesis by enhancing the expression of BMP2/SMAD1 signaling components and decreased osteoclast number and surface. We conclude that LGE promotes osteogenesis via modulating the BMP2/SMAD1 signaling pathway. The study advocates the therapeutic potential of LGE in osteoporosis treatment.
Article
Bone disorders and conditions have been increasing at an alarming rate all over the world, especially in niches where increased obesity and poor physical activity have been prevailing. Synthetic nanohydroxyapatite (HAp) is one of the remedies to reconstruct bone formation. Its rate of dissolution and compatibility is in the moderately acceptable range. The doping of HAp with bone-forming ions can make them highly biologically compatible materials. In the present work, we formulated HAp doped with essential micronutrients of strontium and copper. Nanoglobular Sr and Cu doped HAp (SC-HAp) with an average size of 30 nm was prepared. The SC-HAp was partially crystalline and amorphous, which could influence the dissolution rate of the material. The biomineralization ability of the SC-HAp seemed to be effective in apatite formation. The calcium, collagen and alkaline phosphatase secretion levels after the addition of SC-HAp on MG63 cells indicate the bone-forming capacity of the material. Further, the cell proliferation rate was enhanced compared to the control with SC-HAp.
Article
Full-text available
While youth violence reduction program is a necessity to prevent long-term criminal and violent offending, its effectiveness in youth violent offenders is not well researched. This study investigated the effectiveness of the Violence Prevention Program (VPP) in addressing the aggression, anger, self-control, and empathy of youth violent offenders. One hundred and seventy youths (mean age 15.8 years) who completed VPP from 2008 to 2014 completed self-report measures on study outcomes both before and after the intervention. Repeated measures analyses revealed significant improvement in youths’ anger, aggression, and self-control at post-treatment, but changes in youths’ empathy were not significant. Subsequent analysis found that only youths with lower empathy scores at pre-treatment showed significant increase in empathy post-treatment. Overall, the results suggest that VPP can reduce aggression and mitigate the criminogenic needs of youth offenders. But its effect on empathy may be contingent on youths’ pre-treatment profiles. Limitations and implications for future studies are discussed.
Article
Full-text available
Osteoporosis is by far the most frequent metabolic disease affecting bone. Current clinical therapeutic treatments are not able to offer long-term solutions. Most of the clinically used anti-osteoporotic drugs are administered systemically, which might lead to side effects in non-skeletal tissues. Therefore, to solve these disadvantages, researchers have turned to nanotechnologies and nanomaterials to create innovative and alternative treatments. One of the innovative approaches to enhance osteoporosis therapy and prevent potential adverse effects is the development of bone-targeting drug delivery technologies. It minimizes the systemic toxicity and also improves the pharmacokinetic profile and therapeutic efficacy of chemical drugs. This paper reviews the current available bone targeting drug delivery systems, focusing on nanoparticles, proposed for osteoporosis treatment. Bone targeting delivery systems is still in its infancy, thus, challenges are ahead of us, including the stability and the toxicity issues. Newly developed biomaterials and technologies with potential for safer and more effective drug delivery, require multidisciplinary collaboration between scientists from many different areas, such as chemistry, biology, engineering, medicine, etc, in order to facilitate their clinical applications.
Article
Full-text available
Non-porous monodispersed strontium containing bioactive glass (Si2O-CaO-SrO) nanoparticles (Sr- BGNPs), were synthesised using a modified Stöber process. Silica nanoparticles (Si-NPs) with diameters 90 ± 10 nm were produced through hydrolysis and polycondensation reactions of the silicon alkoxide precursor, tetraethyl orthosilicate (TEOS), prior to the incorporation of cations; calcium (Ca) and strontium (Sr), into the silica networks through heat treatment (calcination). Sr was substituted for Ca on a mole basis from non- (0SrBGNPs) to fullsubstitution (100SrBGNPs) in order to increase the amount of network modifiers in the Si-NPs. The different ratios of Si: Ca; 1:1.3 and 1:8.0, presented various elemental compositions (i.e. 77–92 mol% of SiO
Article
Full-text available
Thanks to their high biocompatibility and bioactivity, bioactive glasses are very promising materials for soft and hard tissue repair and engineering. Because bioactivity and specific surface area intrinsically linked, the last decade has seen a focus on the development of highly porous and/or nano-sized materials. This review emphasizes the synthesis of bioactive glass nanoparticles and materials design strategies. The first part comprehensively covers mainly soft chemistry processes, which aim to obtain dispersible and monodispersed nanoparticles. The second part discusses the use of bioactive glass nanoparticles for medical applications, highlighting the design of materials. Mesoporous nanoparticles for drug delivery, injectable systems and scaffolds consisting of bioactive glass nanoparticles dispersed in a polymer, implant coatings and particle dispersions will be presented.
Article
Blood vessel formation, which encompasses sprouting of capillaries from pre-existing ones (angiogenesis) and the de novo assembly of endothelial progenitor cells to capillaries (vasculogenesis), is vital for biological processes such as organ development, tissue repair and regeneration, and wound healing. The biggest challenge in the regeneration of large bone defects remains the lack of adequate vascularisation within a scaffold/tissue construct to support cell viability and tissue growth. Thus, enhancing the angiogenic potential of biomaterial scaffolds after implantation is pivotal for the success of guided tissue regeneration. Bone is naturally a well vascularised tissue, therefore, for a bone substitute biomaterial to function, a vascular network within the scaffold is a prerequisite. Mesoporous bioactive glasses (MBG) have gained significant attention in the field of bone tissue engineering over the past decade due to their distinct structure and composition. While the ordered mesopores are too small for blood vessel ingrowth, mesopores can increase specific surface area, thus enhancing osteogenesis through controlled ion release and possibly angiogenesis by delivering pro-angiogenic drugs. Engineering the mesoporous structures is a prime example of applying nanotechnology to regenerative medicine. Large macro-pores can be incorporated into mesoporous glasses to produce a highly functional template for tissue regeneration. Various modification strategies for MBG scaffolds have been developed to stimulate angiogenesis, including the addition/delivery of inorganic ionic components, growth factors and drugs, manipulation of angiogenic growth factors such as fibroblast growth factor-1 and vascular endothelial growth factor, and mimicking hypoxic conditions. This review summarises the application of MBG-based biomaterials for bone regeneration with emphasis given to blood vessel formation.
Chapter
Mesoporous glasses exhibit the quickest in vitro bioactive response observed to date for a synthetic material. Furthermore, they present mesopore arrangements and very high surface areas and pore volumes. All these characteristics make them promising candidates as implants in the very near future. In this chapter, the state-of-the-art and the development of current research in mesoporous glasses as scaffolds in bone tissue engineering applications, and as matrixes in drug delivery systems, are reviewed.
Article
Glass caused a revolution in health care when Bioglass was discovered by Larry Hench. It was the first material to bond with bone, rather than be encapsulated by fibrous tissue, launching the field of bioactive ceramics. Bioglass is also biodegradable. Almost 50 years on from its discovery that revolution continues. Bioactive glasses stimulate more bone regeneration than other bioactive ceramics, which is attributed to their dissolution products stimulating cells at the genetic level. This second discovery has changed the way clinicians, scientists, and regulatory bodies think about medical devices and the concept of bioactivity. The original 45S5 Bioglass has only recently found really widespread use in orthopedics, having regenerated the bones of more than 1.5 million patients. Its full potential is still yet to be fulfilled. This article takes the reader from Hench's Bioglass 45S5 to its clinical uses and products, before giving examples of nonsurgical products that now use Bioglass, from consumer products, such as toothpaste, to cosmetics. Other glasses have also found important health care applications, such as borate-based glasses that heal chronic wounds. The revolution looks set to continue as new health care applications are being found for bioactive glasses, contributing to extending the glass age.
Article
Lithium modified bioactive glass nanoparticles were prepared for multiple deliveries of lithium ions and drugs. The particle size, structure and thermal behavior of nanoparticles were analyzed using TEM, FTIR and DSC respectively. The porosity% and specific surface area of glass nanoparticles were about 68.6% and 224.92 (m2/g), respectively. The in vitro bioactivity evaluation in SBF revealed that glass nanoparticles were capable of inducing apatite layer over their surfaces. This could be considered as a good indicator for their future abilities to regenerate bone tissue in vivo. Also, lithium ions were released from glass nanoparticles via diffusion controlled process which could activate Wnt signaling pathway and enhance osteogenesis. As a final point, the possibility of utilizing the glass nanoparticles as a controlled delivery device for vancomycin or 5-FU was verified. Fitting vancomycin or 5-FU release profiles to various mathematical models pointed out that both drugs were released by a diffusion-controlled mode.
Article
The Stӧber process is commonly used for synthesising spherical silica particles. This article reports the first comprehensive study of how the process variables can be used to obtain monodispersed particles of specific size. The modal particle size could be selected within in the range 20 – 500 nm. There is great therapeutic potential for bioactive glass nanoparticles, as they can be internalised within cells and perform sustained delivery of active ions. Biodegradable bioactive glass nanoparticles are also used in nanocomposites. Modification of the Stӧber process so that the particles can contain cations such as calcium, while maintaining monodispersity, is desirable. Here, while calcium incorporation is achieved, with a homogenous distribution, careful characterisation shows that much of the calcium is not incorporated. A maximum of 10 mol% CaO can be achieved and previous reports are likely to have overestimated the amount of calcium incorporated.