ArticlePDF Available

Functionalized Fluorescent Silica Nanoparticles for Bioimaging of Cancer Cells

Authors:

Abstract and Figures

Functionalized fluorescent silica nanoparticles were designed and synthesized to selectively target cancer cells for bioimaging analysis. The synthesis method and characterization of functionalized fluorescent silica nanoparticles (50–60 nm), as well as internalization and subcellular localization in HeLa cells is reported here. The dye, rhodamine 101 (R101) was physically embedded during the sol–gel synthesis. The dye loading was optimized by varying the synthesis conditions (temperature and dye concentration added to the gel) and by the use of different organotriethoxysilanes as a second silica precursor. Additionally, R101, was also covalently bound to the functionalized external surface of the silica nanoparticles. The quantum yields of the dye-doped silica nanoparticles range from 0.25 to 0.50 and demonstrated an enhanced brightness of 230–260 fold respect to the free dye in solution. The shell of the nanoparticles was further decorated with PEG of 2000 Da and folic acid (FA) to ensure good stability in water and to enhance selectivity to cancer cells, respectively. In vitro assays with HeLa cells showed that fluorescent nanoparticles were internalized by cells accumulating exclusively into lysosomes. Quantitative analysis showed a significantly higher accumulation of FA functionalized fluorescent silica nanoparticles compared to nanoparticles without FA, proving that the former may represent good candidates for targeting cancer cells.
Content may be subject to copyright.
Sensors 2020, 20, 5590; doi:10.3390/s20195590 www.mdpi.com/journal/sensors
Article
Functionalized Fluorescent Silica Nanoparticles
for Bioimaging of Cancer Cells
Ruth Prieto-Montero 1, Alberto Katsumiti 2, Miren Pilare Cajaraville 2, Iñigo López-Arbeloa 1
and Virginia Martínez-Martínez 1,*
1 Departamento de Química Física, Universidad del País Vasco/Euskal Herriko Unibertsitatea (UPV/EHU),
48080 Bilbao, Spain; ruth.prieto@ehu.eus (R.P.-M.); inigo.lopezarbeloa@ehu.eus (I.L.-A.)
2 Departamento de Zoología y Biología Celular Animal, Universidad del País Vasco/Euskal Herriko
Unibertsitatea (UPV/EHU), 48080 Bilbao, Spain; alberto.katsumiti@ehu.eus (A.K.);
mirenp.cajaraville@ehu.eus (M.P.C.)
* Correspondence: virginia.martinez@ehu.eus; Tel.: +34-946-015-969
Received: 13 August 2020; Accepted: 26 September 2020; Published: 29 September 2020
Abstract: Functionalized fluorescent silica nanoparticles were designed and synthesized to
selectively target cancer cells for bioimaging analysis. The synthesis method and characterization of
functionalized fluorescent silica nanoparticles (50–60 nm), as well as internalization and subcellular
localization in HeLa cells is reported here. The dye, rhodamine 101 (R101) was physically embedded
during the sol–gel synthesis. The dye loading was optimized by varying the synthesis conditions
(temperature and dye concentration added to the gel) and by the use of different
organotriethoxysilanes as a second silica precursor. Additionally, R101, was also covalently bound
to the functionalized external surface of the silica nanoparticles. The quantum yields of the dye-
doped silica nanoparticles range from 0.25 to 0.50 and demonstrated an enhanced brightness of 230–
260 fold respect to the free dye in solution. The shell of the nanoparticles was further decorated with
PEG of 2000 Da and folic acid (FA) to ensure good stability in water and to enhance selectivity to
cancer cells, respectively. In vitro assays with HeLa cells showed that fluorescent nanoparticles were
internalized by cells accumulating exclusively into lysosomes. Quantitative analysis showed a
significantly higher accumulation of FA functionalized fluorescent silica nanoparticles compared to
nanoparticles without FA, proving that the former may represent good candidates for targeting cancer
cells.
Keywords: targeting; functionalized fluorescent silica nanoparticles; rhodamine 101; polyethylene
glycol; folic acid; HeLa cells
1. Introduction
Cancer is the second cause of human death worldwide [1–3]. Its early diagnosis is the key for an
effective treatment to ensure patient survival and cure. Currently, the most used imaging techniques to
detect cancer are based on X-ray sources (e.g., computed tomography scan), high magnetic fields (e.g.,
magnetic resonance imaging), or radioactive substances as tracers (e.g., positron emission tomography
scan); however, these techniques can cause side effects in the patients. In fact, to improve the contrast
in the images sometimes several scans are required, increasing the chances to suffer side effects [4–9].
In the last decade an alternative complementary detection technique, "fluorescence microscopy", is in
expansion since it is less invasive and offers a safe detection with high sensitivity, specificity, and
resolution. This versatile technique enables direct imaging of biological structures both in in vitro and
in vivo experiments by the use of suitable fluorophores [10–12].
Sensors 2020, 20, 5590 2 of 15
Generally speaking, a good fluorophore should fulfill several requirements; (i) high molar
extinction coefficients (ε), (ii) high fluorescence quantum yields (Φ) and narrow emission spectra, (iii)
low tendency to aggregate, (iv) high Stokes shift in order to prevent reabsorption–reemission
processes, and (v) high chemical-, thermal-, and photostability [13]. Many commercial fluorophores
meet those conditions but they have some drawbacks to be used as sensors, such as the lack of
selectivity for a specific tissue or organelle, low photostability, and poor solubility in physiological
media [14].
To overcome these limitations, one strategy is to chemically modify commercial fluorescent dyes
to increase their specificity as chemosensors for bioimaging. However, the required multistep
chemistry increases the cost of production making these modified dyes an unviable alternative.
Another alternative is to associate the fluorescent dyes to a carrier that confers the missing properties
of fluorochromes alone. Nanoparticles have been increasingly used as nanocarriers of different
molecules because their large surface area serves as platform for the attachment of several molecules.
Additionally, nanoparticles help to stabilize hydrophobic components in aqueous media and prevent
degradation and inactivation of active compounds [15–18]. In cancer diagnosis and treatment,
targeted nanoparticles can be designed and synthesized to enhance their selective uptake and
retention inside tumoral cells [4,19–27].
Targeted nanoparticles should be carefully designed in terms of size, structure, composition,
and functionalization to balance their stability, diffusion, specificity, and biocompatibility. Currently,
there are several nanosystems based on liposomes-, polymeric-, micellar-, metallic-, or protein-based
nanoparticles that are approved by the Food and Drug Administration (FDA) for medical
applications [4,19–24,28–30]. Among them, silica nanoparticles (SN), have been receiving special
attention due to their wide spectrum of applications. In the last years, SNs have emerged as potential
nanocarriers for selective imaging (diagnosis) and targeted drug delivery (therapy) due to their high
surface area of easy functionalization, good biocompatibility, optically transparent properties, and
low cost [31,32]. Dye-loaded silica nanoparticles have been reported as very promising fluorescent
biocompatible nanoplatforms with enhanced photostability and brightness compared to free dyes,
thus allowing long-term tracking and higher signal-to-noise ratio fluorescent signals [14,33–46]. Dye
molecules can be physically encapsulated or covalently attached to the silica external or internal
surface of the nanoparticles [47–50]. Fluorophores such as rhodamines represent good candidates for
labelling nanocarriers because they can be easily associated to silica nanoparticles and show excellent
photophysical properties, such as intense absorption and emission bands, in the green-red visible
spectra [51].
A part of using fluorescent dyes to label nanoparticles, the external surface of nanoparticles can
be further functionalized with molecules that confer stability in aqueous solution. Polyethylene
glycol (PEG) is a molecule that improves water stability, minimizes non-specific interactions with
other molecules in the extracellular matrix, and does not activate the immune response [26,27,52–55].
Thus, PEG ensures nanoparticles dispersion and high bioavailability to cells.
Finally, the selectivity of nanocarriers to cancer cells can be further increased by functionalizing
with molecules known to have specific interactions with plasma membrane receptors, which are
overexpressed on tumor cells but not on healthy cells. For instance, folate receptors (FRs) exhibit limited
expression on healthy cells, but are overexpressed on cancer cells in ovary, mammary gland, colon,
lung, prostate, nose, throat, and brain [56,57]. Therefore functionalization of silica nanoparticles with
folic acid (FA) turns them into highly selective sensors of cancer cells [43,53,54,57–67].
In this context, in the present work functionalized fluorescent silica nanoparticles were designed
and synthesized to target cancer cells. All functionalized silica nanoparticles were (photo)physically
characterized (diameter, size distribution, stability, and fluorescent efficiency) in aqueous media.
Then, in vitro assays with HeLa cells were used to assess nanoparticles cytotoxicity and, for the most
promisor nanoparticles, internalization, and intracellular localization was studied. Finally,
internalization of silica nanoparticles with and without FA was quantified in HeLa cells in order to
evaluate if functionalization with FA enhances nanoparticles uptake by cancer cells.
Sensors 2020, 20, 5590 3 of 15
2. Materials and Methods
2.1. Synthesis of the Core-Shell Nanoparticles
All starting materials and reagents were commercially obtained and used without any further
modification. Tetraethoxysilane (TEOS) ( 99%), ammonium hydroxide solution (NH4OH) (25%
NH3 basis), hexadecyltrimethylammonium bromide (CTBA) (98%), 3-aminopropyltrimethoxysilane
(APTMS) (97%), triethoxymehylsilane (MTES) (99%), triethoxyvinylsilane (VTES) (97%),
phenyltriethoxysilane (PTES) (98%), triethoxy(octyl)silane (OTES) (98%), rhodamine 101 (R101)
(99%), N-hydroxysuccimide (NHS) (98%), and N-(3-(dimethylaminopropyl)-N’-ethylcarboiimide
(EDC) (97%) and folic acid (FA) (97%) were purchased from Sigma-Aldrich (Darmstadt, Germany)
and polyethylene glycol (PEG) (>95%) from Iris BIOTECH GMBH (Maktredwitz, Germany).
Mesoporous silica nanoparticles (MSNs) were synthesized as it has been described previously [33].
Ormosil nanoparticles (ONPs) were synthesized modifying the MSN synthesis. First, 0.1 g of CTBA
was dissolved in 50 mL of NH4OH at 60 °C. When CTBA was dissolved, TEOS was added together
with a second silica source (MTES or VTES or PTES or OTES) in different ratios from 1:0.1 to 1:1,
respectively. After 5 h under vigorous stirring at 60 °C, 0.8 mL of TEOS (1 M in EtOH, 0.8 mmol) and
0.8 mL of an APTMS solution (12% v/v in EtOH, 0.007 mmol) was added and kept stirring for 24 h at
60 °C. Then, the temperature was decreased to 25 °C and the mixture was left with vigorous stirring for
other 12 h. The NPs were collected by centrifugation at 19,000 rpm at room temperature for 15 min. The
collected solid was washed three times with a mixture of Milli Q water/EtOH and a fourth time with
EtOH. The surfactant was removed by stirring the NPs with concentrated HCl (0.2 g of HCl in EtOH)
for 24 h. The NPs were collected by filtration.
2.2. Dye Encapsulation within the NP Core
Rhodamine 101 (R101) dye was directly added to the synthesis gel before silica source addition.
The concentration of dye in the synthesis gel and the temperature were varied (5·103 M–5·104 M and
T = 60–80 °C) to optimize the size of the nanoparticles and the dye loading. When the R101 was
completely dissolved in the mixture, the silica source was added; TEOS for MSNs or TEOS and the
second silica source for ONPs. After stirring the mixture for 5 h, the shell functionalization with amine
groups was carried out as it is explained previously. The corresponding nanoparticles will be denoted
MSN-C-R101-T and X-ONP-R101-T (being X the second silica source and T the temperature used
during the synthesis).
2.3. Grafting of Molecules on NP Surface
R101 and/or folic acid (FA) were grafted to the amine groups of nanoparticles in the external
surface by carbodiimide method, following the synthesis described previously [33]. In contrast,
silylated-PEG chain (2000 Da) was condensed to the hydroxyl groups in the shell of nanoparticles as
it has just been described [33]. The corresponding nanoparticles will be denoted as MSN-S-R101-60,
MSN-S-R101-60 –PEG, and MSN-S-R101-60-PEG-FA.
2.4. Characterization
The size, shape and morphology of the silica nanoparticles were characterized by electron
microscopes, scanning electron microscopy (SEM) and transmission electron microscopy (TEM). SEM
images were obtained in a JEOL JSM-6400 (JEOL, Tokyo, Japan) and TEM images were obtained in a
Philips SuperTwin CM200 (Thermo Fisher Scientific, Eindhoven, Netherlands) at 200 kV. The
nanoparticles size distribution was analyzed by Images-J software (1.52u, National Institute of
Health, Bethesda, MD, USA). Dynamic light scattering (DLS) and Zeta potential (Zpot) measurements
to analyze the NP size and their stability in suspension were carried out using a Malvern Zetasizer
Nano ZS (Malvern Products, Madrid, Spain), which has a Helio-Neon (λ = 633 nm) laser. FTIR spectra
were obtained from neat samples in powder using ATR technique in Affinity-1S Shimadzu
spectrometer (Izasa Scientific, Barcelona, Spain) (4000–400 cm1 range). The silica nanoparticles
Sensors 2020, 20, 5590 4 of 15
absorption spectra were recorded by UV-Vis-NIR Spectroscopy (model Cary 7000, Agilent
Technologies, Spain) equipped with two lamps (halogen lamp for Vis-IR region and deuterium lamp
for UV region) and an integrating sphere (model Internal DRA 900, Livingston, UK). The fluorescence
measurements were recorded with an Edinburgh Instruments Spectrofluorimeter (FLSP920 model,
Livingston, UK) equipped with a xenon flash lamp 450 W as the excitation source. The fluorescence
spectra were corrected from the wavelength dependence on the detector sensibility. The absolute
photoluminescence quantum yields of the dye-containing nanoparticles were measured in an
integrated sphere coupled to this spectrofluorimeter. The absorbance at excitation wavelength was
obtained by comparing the scatter signal of the dye-loaded hybrid material with a Teflon disk, used
as a reference (with a diffuse reflectance of 100%).
The amount of dye uptake into the MSNs or ONPs was estimated photometrically, by dissolving
the silica matrix with KOH [33,68,69].
2.5. In Vitro Assays
Cells culture: Human cervix adenocarcinoma HeLa cells obtained from ATCC were grown in
Dulbecco´s modified Eagle´s medium (DMEM) supplemented with 10% (v/v) fetal bovine serum (FBS)
and 50 U/mL penicillin and 50 mg/mL streptomycin, in a humidified 5% CO2 cells incubator at 37 °C.
For the cell viability assays, cells were grown to monolayer confluency in 96-well microplates. For
internalization and subcellular localization studies, cells were seeded in glass-bottom 35 mm petri
dishes and subconfluent monolayers were used.
Sample preparation: Samples used for in vitro experiments were prepared by suspending MSN
samples directly in PBS buffer (1·104M). Suspensions were stirred for at least 24 h before the exposures.
Cell viability (MTT) assay: Cytotoxicity of MSN samples (MSN-C-R101-70-PEG and MSN-S-
R101-60-PEG) was assessed in HeLa cells using the thiazolyl blue tetrazolium bromide (MTT) assay
following manufacturer’s instructions. After exposures, cells were incubated with a 50 mg/mL MTT
solution for 3 h at 37 °C. Then, reduced formazan product was extracted from cells with pure DMSO
and the absorbance was measured at 570 nm in a Biotek EL 312 microplate spectrophotometer reader
(Biotek instruments, Winooski, VT, USA). Cell viability was expressed as the percentage respect to
control cells. Differences between control and treated cells were analyzed through the Kruskal–Wallis
test followed by Dunn’s post hoc test using the SPSS 23.0 software (IBM, Chicago, CA, USA). Significance
level was established at 5% (p < 0.05). Four replicates per treatment were performed for all tests and tests
were repeated three times each.
Internalization and subcellular localization: To evaluate internalization and subcellular
localization of MSN samples through confocal microscopy, cells were incubated for 24 h with 1, 10, and
100 μg/mL of MSN-C-R101-70-PEG and MSN-S-R101-60-PEG-FA in 10% FBS supplemented DMEM
culture medium. Unexposed cells were used as control. After exposures, cells were washed three times
with culture medium and incubated for 30 min with 50 nM LysoTrackerTM Deep Red (Invitrogen,
Paisley, UK) to label cell’s lysosomes, and fixed with 0.4% paraformaldehyde for 10 min at 4 °C. Cells
were then washed three times with culture medium and observed under an Olympus Fluorview FV500
confocal microscope (Olympus, Hamburg, Germany). Images were edited using Fiji
software (ImageJ 1.49a, National Institutes of Health, Bethesda, MD, USA). To quantify the
internalization of MSN samples, cells were incubated for 24 h with 0.1 and 1 μg/mL of MSN samples
with and without folic acid (MSN-S-R101-60-PEG-FA and MSN-S-R101-60-PEG, respectively) in 10%
FBS supplemented DMEM culture medium. Unexposed cells were used as control. After exposures,
cells were washed three times with culture medium and fluorescence of MSN samples was measured
at λex = 530 nm/λem = 590 nm in a Cytation 5 Cell Imaging Multi-mode reader (Biotek instruments,
Winooski, VT, USA). Fluorescence of MSN samples at 0, 0.0001, 0.001, 0.01, 0.1, and 1 μg/mL was used
to normalize internalization data. Differences between control and treated cells were analyzed through
Kruskal–Wallis test followed by Dunn’s post hoc test. Differences between the MSN-S-R101-60-PEG-
FA and the MSN-S-R101-60-PEG treatments were analyzed through Kruskal–Wallis test followed by
Mann–Whitney test. Significance level was established at 5% (p < 0.05). All tests were performed using
Sensors 2020, 20, 5590 5 of 15
the SPSS software. Four replicates per treatment were performed for all tests and tests were repeated
three times each.
3. Results and Discussion
The synthesis of MSNs and ONPs, using the modified Stöber method [70], was directed to obtain
silica nanoparticles of around 50 nm, which is considered a suitable size for biomedical
applications [52,67,71]. In the case of ONPs different organophilic silica sources (XTES, Figure 1) and
ratios respect the main silica source (TEOS), TEOS:XTES 1:0.1, 1:0.5 and 1:1, were studied. The external
surface of all MSNs and ONPs was functionalized with amine groups by adding
aminopropyltrimethoxysilane (APTMS) after the core nanoparticle formation (Figure 1).
Figure 1. Synthesis of the Ormosil nanoparticles by modified Stöber method with different silica
sources (MTES: triethoxymehylsilane, VTES: triethoxyvinylsilane, PTES: phenyltriethoxysilane, and
OTES: triethoxy(octyl)silane).
The MSN size and distribution were analyzed by TEM and SEM (Figures 2 and S1). Electron
microscopy images show spherical nanoparticles with a narrow size distribution of 47 ± 10 nm.
Regarding the synthesis of ormosil nanoparticles, ONPs, several organophilic silica sources, acting as
co-precursors of the silica, at different concentrations, were used with the aim of modulating the
hydrophilicity of the porous environment for efficient confinement of the rhodamine fluorescent
dye (Table 1). The morphology of these nanoparticles and the size distribution, analyzed by TEM, are
depicted in Figure S2 and Table 1. Except for the synthesis with the mixture TEOS:OTES which did not
form nanoparticles at any ratio of silica sources, the rest of the samples rendered spherical nanoparticles
with a size distribution of around 40–50 nm (Table 1), although in some cases the size distribution is
broader than that previously described for MSNs.
Figure 2. Transmission electron microscopy (TEM) images (A,B), scanning electron microscopy (SEM)
image (C) and size distribution (D) of mesoporous silica nanoparticles (MSNs).
Sensors 2020, 20, 5590 6 of 15
Table 1. Silica nanoparticles synthesized and their average size by TEM.
Name Silica Source Size (nm)
MSN TEOS 47 ± 10
M1-ONP TEOS:MTES (1:0.1) 44 ± 16
M2-ONP TEOS:MTES (1:0.5) 38 ± 7
M3-ONP TEOS:MTES (1:1) 39 ± 18
V1-ONP TEOS:VTES (1:0.1) 42 ± 7
V2-ONP TEOS:VTES (1:1) 40 ± 9
P1-ONP TEOS:PTES (1:0.1) 47 ± 10
P2-ONP TEOS:PTES (1:1) 49 ± 18
O1-ONP TEOS:OTES (1:0.1) -
O2-ONP TEOS:OTES (1:1) -
Rhodamine 101 (R101), with intense absorption and fluorescence bands (λab = 560 nm, ɛ = 8.4·104
M1 cm1, λfl = 597 nm and Φ = 0.77 in water) was chosen as fluorophore to be loaded into the silica
nanoparticles by two different methods: i) physically embedded within the porous core of MSNs and
ONPs and ii) covalently tethered at their outside surface [25].
In the first approach, to encapsulate R101 within the MSNs, the dye was added to the mixture,
before the silica source, at a concentration of 5·104 M (Table 2). Generally, the dye loaded into silica
reached 0.5–1 μmol/g being in the same range as other fluorescent silica nanoparticles, with diameter
sizes between 20 and 50 nm, previously optimized with rhodamine 6G [33]. Nevertheless, it is
considered low, and with the aim of increasing the dye uptake, R101 was occluded into the different
ormosil silica nanoparticles with varied hydrophobic porous environment following the same
procedure. As a result, the final dye amount within the ormosil silica nanoparticles was slightly
increased (1.5–1.7 fold). However, much higher dye incorporation was found by the rise of temperature
of the gel from 60 °C to 70 °C and by augmenting the concentration in the gel to 2.5·103 M (Table 2).
Under these synthesis conditions, particles of around 60 nm diameter and a considerable dye amount
embedded were reached with higher dye loaded (>4 μmol/g). Nonetheless, a further increase in the
temperature and/or dye concentration in the gel led to a drastic increase in the size of the nanoparticles
reaching a diameter of around 500 nm (Table 2 and Figure S3). According to the results obtained in the
present study, the sample MSN-C-R101-70 was considered the best fluorescent nanoplatform in this
series and was selected for further studies in HeLa cells.
Table 2. Synthesis conditions of MSNs and Ormosil Nanoparticles (ONPs): TEOS:XTES ratio,
temperature, and initial concentration of R101 in the sol-gel mixture. Average size of nanoparticles (by
TEM) and the final amount of loaded dye inside the nanoparticles are given.
Sample Mixture Ratio
T
(°C)
[Dye]0
(M)
Size
(nm)
Dye
(μmol/g)
MSN-C-R101-60 TEOS 1:0 60 5·104 47 ± 9 0.56
M-ONP-R101-60 TEOS:MTES 1:1 60 5·104 54 ± 8 0.81
V-ONP-R101-60 TEOS:VTES 1:1 60 5·104 29 ± 5 0.96
P-ONP-R101-60 TEOS:PTES 1:1 60 5·104 39 ± 7 0.94
MSN-C-R101-70 TEOS 1:0 70 2.5·103 60 ± 9 9.98
M-ONP-R101-70 TEOS:MTES 1:1 70 2.5·103 58 ± 11 7.54
V-ONP-R101-70 TEOS:VTES 1:1 70 2.5·103 63 ± 14 4.21
MSN-C-R101-80 TEOS 1:0 80 5·103 541 ± 73 11.4
In the second approach, rhodamine 101 was covalently anchored to the external amine function
of MSNs through its carboxylic group by common peptide reaction (sample named as MSN-S-R101-
60). Note here that this particular rhodamine, R101, allows this grafting since, after the depronotation
process, the formation of spiro-lactone is avoided by the rigidity of the alkyls on N atoms and the
zwitterionic form is favored, whereas only lactone species is present in aprotic solvents for the rest of
Sensors 2020, 20, 5590 7 of 15
rhodamines and consequently the peptide coupling does not take place (Figure 3) [72–74]. The
estimated amount of the R101 dye tethered outside, of 22 μmol/g, implied 2-fold increase respect the
sample MSN-C-R101-70 with the largest amount of dye occluded inside the core (Table 2).
Figure 3. (top) General molecular structures of rhodamines in equilibrium and (bottom) structure of
Rhodamine 101 (R101): cationic (A), zwitterionic (B), and lactone (C) forms.
Note that the presence of the R101 dye at the external surface in the sample MSN-S-R101-60
should render a more hydrophobic shell and consequently, these nanoparticles were not stable in
water and only a relatively stabilized suspension was found for a less polar solvent, such as
chloroform. Regarding the sample MSN-C-R101-70, with the dye embedded in the silica core,
although the hydrodynamic size in aqueous media, registered by DLS, did not point to a particle
aggregation process in water, the low Zpot value indicated poor stability in such media (Table 3). In
fact, the photophysical properties of sample MSN-C-R101-70 could not be studied in water due to the
particle flocculation while recording the absorption and emission spectra. Thus, it is of crucial
importance to improve the stability of these fluorescent nanosystems in aqueous media for their
future implementation as bioimaging agents.
Table 3. Hydrodynamic diameter (in nm), zeta potential (in mV), absorption peak (λab in nm),
fluorescence peak (λfl in nm), and fluorescence quantum yield (Φfl) in water of the dye-loaded silica
nanoparticles without and with PEG-coated at their external surface.
Sample DLS
(nm) Zpot (mV) λab λfl Φa Brightness
MSN-C-R101-70 60 4.0 - - - -
MSN-C-R101-70-PEG 69 21.0 572.0 594.0 0.51 230
MSN-S-R101-60-PEG 64 23.0 571.0 595.0 0.25 260
R101 in water - - 560.0 597.0 0.77 1
To improve the stability in water of samples MSN-C-R101-70 and MSN-S-R101-60, polyethylene
glycol (PEG) chains of 2000 Da with a silylated end was anchored to the inherent hydroxyl groups of
the external surface of silica nanoparticles (samples denoted as MSN-C-R101-70-PEG and MSN-S-R101-
60-PEG in Table 3). The presence of PEG at the silica nanoparticles was checked by FTIR
(Figure S4) [26]. After PEGylation of the outside surface of MSN, a drastic increase in the Zpot
values (Table 3), from 4 mV up to 23 mV was reached, ensuring good stability in water.
The photophysical properties of the PEGylated nanoparticles MSN-C-R101-70-PEG and MSN-
S-R101-60-PEG characterized in water are shown in Table 3 and Figure 4. The absorption bands were
broader with a more pronounced shoulder, centered at around 525 nm, respect to the R101 in diluted
Sensors 2020, 20, 5590 8 of 15
aqueous solution. This could indicate the presence of dye aggregates, which according to exciton
splitting bring new absorption bands depending on their geometry. However, as it was previously
stated that scattering effects also introduce spectral distortions in the absorption spectra, inducing a
“new’’ shoulder in near position as the current weak vibronic band [75]. Although molecular
aggregation cannot be ruled out, the scattering caused by the silica nanoparticles was depicted by the
increase of the baseline in the absorption spectra at shorter wavelengths for the samples MSN-C-
R101-70-PEG and MSN-S-R101-60-PEG in suspension respect to the dye in solution (Figure 4).
Figure 4. Normalized absorption and emission spectra for MSN-C-R101-70-PEG (black), MSN-S-
R101-60-PEG (blue), and rhodamine 101 (brown) in diluted aqueous solution.
Conversely, the recorded fluorescence spectra for the dye-loaded silica nanoparticles are similar
to those obtained for the free dye. The fluorescent nanoparticles render a fluorescence quantum yield
of Φ = 0.51 and Φ = 0.25 for MSN-C-R101-70-PEG and MSN-S-R101-60-PEG, respectively (Table 3).
The lower fluorescence quantum yield registered in sample MSN-S-R101-60-PEG could be assigned
to reabsorption–reemission processes and/or molecular aggregation as a consequence of a higher
amount of dye molecules which were distributed mainly at the surface of the nanoparticle.
Although the quantum yield of those fluorescent silica nanoparticles was lower than that of the
free dye in water solution (Table 3), the relative brightness of nanoparticles was usually much higher,
about tens of times, due to a greater number of fluorophores per particle, enhancing the signal and
consequently the sensitivity in the fluorescence imaging detection. Taking into account the area of
fluorescence spectra of each suspension and the dye solution, the diameter of nanoparticles, the
number of nanoparticles and R101 molecules in the cuvette [76] (see more details in ESI), MSN-C-
R101-70-PEG and MSN-S-R101-60-PEG showed a relative brightness of 230 and 260 times respect
R101 in water, which made them at least one order of magnitude brighter than the most used
quantum dots [41,54,76–78].
Finally, internalization capacity and cytotoxicity of MSN-C-R101-70-PEG and MSN-S-R101-60-
PEG were studied through in vitro experiments in HeLa cells.
HeLa cells were exposed to a wide range of concentrations of MSN-C-R101-70-PEG and MSN-
S-R101-60-PEG (0.1–1000 μg/mL) and cytotoxicity was assessed through the MTT assay (Figure 5).
MSN-C-R101-70-PEG was more cytotoxic than MSN-S-R101-60-PEG, reducing cell viability to less
than 40% in HeLa cells exposed to 10 μg/mL, and to less than 20% in those exposed to 100 and 1000
μg/mL. MSN-S-R101-60-PEG reduced cell viability to up to 72.4% in cells exposed to the highest
concentration (1000 μg/mL). This difference could be attributed to the possible release of the
surfactant CTAB from the core of sample MSN-C-R101-70-PEG, used as the template in the synthesis
of the mesoporous nanoparticles, which has been proved to be toxic to many cell types [79–81]. In the
case of MSN-S-R101-60-PEG the surfactant was previously removed before the R101 grafting,
whereas for MSN-C-R101-70-PEG this process cannot be undertaken because it would bleach the
R101 dye inside the particles mesopores.
Sensors 2020, 20, 5590 9 of 15
Figure 5. Results of MTT assay for MSN-C-R101-70-PEG (A) and MSN-S-R101-60-PEG (B). Stars
indicate significant differences with respect to controls according to the Kruskal
Wallis test followed
by the Dunn´s test (p < 0.05).
Based on the confocal microscopy analysis, MSN-C-R101-70-PEG were internalized in the cells and
specifically accumulated into the lysosomes as shown by the subcellular localization
experiments (Figure 6), where the lysotracker (green) and the nanoparticles (red) were co-localized
(yellow). The fact that these NPs are localized into lysosomes indicates that they are possibly taken up
by endocytosis, as widely reported for silica NPs at similar size range [82,83]. Note that the
nanoparticles offer a sharper quality in the bioimaging of lysosomes with respect to commercial
lysotrackers. Thus, results indicate that MSN-C-R101-70-PEG could be potentially employed as an
alternative lysotracker in cancer cells or other cells that express FR on their plasma membrane.
Nevertheless, further studies are needed to confirm it.
Figure 6. Fluorescence images of MSN-C-R101-70-PEG internalized into lysosomes of HeLa cells.
Images show lysosomes (green-left), rhodamine 101 from MSN-C-R101-70-PEG (red-middle), and
merged fluorescence of lysosomes and rhodamine 101 (yellow-right). Scale bars 100 μm.
As a step forward to enhance the selectivity of the hybrid nanosystem to cervix adenocarcinoma
cells, FA was tethered to the shell of MSN-S-R101-60-PEG nanoparticles. The presence of FA at the
surface of silica nanoparticles was confirmed by absorption spectroscopy where the characteristic
absorption band of FA, centered at 365 nm, was detected together with the main band of R101 at
575 nm, as well as its emission band at 455 nm under 355 nm excitation (Figure S5) [84].
0
20
40
60
80
100
120
control 0.1 1 10 100 1000
% control
MSN-C-R101-70-PEG (μg/ml)
0
20
40
60
80
100
120
control 0.1 1 10 100 1000
% control
MSN-S-R101-60-PEG (μg/ml)
AB
Sensors 2020, 20, 5590 10 of 15
Similar to the results obtained in MSN-C-R101-70-PEG exposures, MSN-S-R101-60-PEG
functionalized with FA (MSN-S-R101-60-PEG-FA) were internalized into lysosomes of HeLa
cells (videos 1 and 2, supplementary material).
The internalization of the nanosystems MSN-S-R101-60-PEG and MSN-S-R101-60-PEG-FA was
quantified in HeLa cells exposed to 0.1 and 1 μg/mL of each MSN sample (Table 4). In accordance
with previous studies [85], results showed that functionalization with FA significantly increased the
internalization of MSNs into HeLa cells, being 13% and 20% higher at 0.1 μg/mL and 1 μg/mL
nanoparticle concentrations, respectively (Table 4).
Table 4. Quantification (μg/mL) of MSN-S-R101-60-PEG and MSN-S-R101-60-PEG-FA internalized
into HeLa cells after 24 h exposure to 0.1 and 1 μg/mL of each MSN sample (mean ± SD). Different
letters indicate significant (p < 0.05) differences among groups.
MSN Samples Control 0.1 1
(μg/mL) (μg/mL) (μg/mL)
MSN-S-R101-60-PEG 0 ± 0.0008 a 0.355 ± 0.029 b 0.491 ± 0.017 d
MSN-S-R101-60-PEG-FA 0 ± 0.0005 a 0.406 ± 0.033 c 0.616 ± 0.023 e
FA is a manufactured form of folate which is required for the synthesis, repair, and methylation of
DNA, as well as for the metabolism of amino acids and RNA [56]. Cancer cells are known to require high
levels of folate for cell growth and proliferation; thus, overexpress folic acid receptors on their surface.
Folate receptors are a cell surface glycosyl phosphatidylinositol-anchored glycopolypeptides [86], which
recognize and internalize FA via endocytosis [87,88]. Folate receptors exhibit limited expression on
healthy cells, but are often present in a large number of cancer cells [89]. Thus, as found in the present
work, functionalization of nanoparticles with FA increased their uptake by cancer cells.
4. Conclusions
Functionalized silica nanoparticles with PEG chains and FA at the external surface, and with
rhodamine 101 as fluorescent label embedded into silica nanoparticle’s porous core or covalently
linked outside of nanoparticles have demonstrated to be the most suitable fluorescent nanoplatforms
for bioimaging of cancer cells. These nanoplatforms showed a suitable dye loading (5–10 mg dye/g
nanoparticle), high brightness (230–260 fold increase respect to the dye in solution), improved
stability in water (Zpot ~ 23 mV), low cytotoxicity (at concentration 1 μg/ml), high internalization
into HeLa cells and great specificity to cells lysosomes. Functionalization with FA enhanced the
internalization of the functionalized silica nanoparticles. These nanosystems offer sharper
fluorescence imaging with greater signal-to-noise ratio with respect to commercial lysotrackers,
making them promising nanoplatforms for bioimaging of cancer cells.
Supplementary Materials: The following are available online at www.mdpi.com/1424-8220/20/19/5588/s1,
Figure S1: SEM image of MSNs; Figure S2: TEM images for ORMOSIL nanoparticles using different second silica
source and proportions; Figure S3: TEM images for MSN-C-R101-80 sample; Figure S4: FTIR spectra of MSN and
MSN-PEG nanoparticles; Figure S5: Normalized absorption and emission spectra for MSN-C-R101-70-PEG,
MSN-S-R101-60-PEG, and rhodamine 101 in water, Video 1 and Video 2: Fluorescence images of MSN-S-R101-
60-PEG-FA internalized into lysosomes of HeLa cells.
Author Contributions: conceptualization, V.M.-M.; methodology, R.P.-M., A.K. and V.M.-M.; writing—original
draft preparation, R.P.-M., A.K. and V.M.-M.; writing—review and editing, R.P.-M., A.K., M.P.C., I. L.-A. and
V.M.-M.; supervision, M.P.C., I. L.-A. and V.M.-M.; funding acquisition, M.P.C., I. L.-A. and V.M.-M..; All
authors have read and agreed to the published version of the manuscript
Funding: This research was funded by the Basque Government, grant numbers IT912-16 and IT-1302-19;
Ministry of Economy and Competitiveness (MINECO), grant numbers MAT2017-83856-C3-3-P and CTM2016-
81130-R; and the University of the Basque Country (UPV/EHU), grant number COLAB19/01.
Acknowledgments: We would like to thank the Advanced Research Facilities (SGIker) of the University of the
Basque Country for the support on the image analysis.
Sensors 2020, 20, 5590 11 of 15
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of the
study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to
publish the results.
References
1. Shewach, D.S.; Kuchta, R.D. Introduction to cancer chemotherapeutics. Chem. Rev. 2009, 109, 2859–2861,
doi:10.1021/cr900208x.
2. Pérez-Herrero, E.; Fernández-Medarde, A. Advanced targeted therapies in cancer: Drug nanocarriers, the
future of chemotherapy. Eur. J. Pharm. Biopharm. 2015, 93, 52–79, doi:10.1016/j.ejpb.2015.03.018.
3. Hassanpour, S.H.; Dehghani, M. Review of cancer from perspective of molecular. J. Cancer Res. Pract. 2017,
4, 127–129, doi:10.1016/j.jcrpr.2017.07.001.
4. Chaturvedi, V.K.; Singh, A.; Singh, V.K.; Singh, M.P. Cancer Nanotechnology: A New Revolution for
Cancer Diagnosis and Therapy. Curr. Drug Metab. 2018, 20, 416–429,
doi:10.2174/1389200219666180918111528.
5. Eftekhari, A.; Hasanzadeh, M.; Sharifi, S.; Dizaj, S.M.; Khalilov, R.; Ahmadian, E. Bioassay of saliva
proteins: The best alternative for conventional methods in non-invasive diagnosis of cancer. Int. J. Biol.
Macromol. 2019, 124, 1246–1255, doi:10.1016/j.ijbiomac.2018.11.277.
6. Majer, M.; Knežević, Ž.; Popić, J.; Hršak, H.; Miljanić, S. Organ doses and associated cancer risks for
computed tomography examinations of the thoracic region. Nucl. Technol. Radiat. Prot. 2018, 33, 100–105,
doi:10.2298/NTRP1801100M.
7. Singh, J.; Daftary, A. Iodinated contrast media and their adverse reactions. J. Nucl. Med. Technol. 2008, 36,
69–74, doi:10.2967/jnmt.107.047621.
8. Gauden, A.J.; Phal, P.M.; Drummond, K.J. MRI safety; Nephrogenic systemic fibrosis and other risks. J.
Clin. Neurosci. 2010, 17, 1097–1104, doi:10.1016/j.jocn.2010.01.016.
9. Marshall, J.; Martin, T.; Downie, J.; Malisza, K. A comprehensive analysis of MRI research risks: In support
of full disclosure. Can. J. Neurol. Sci. 2007, 34, 11–17, doi:10.1017/S0317167100005734.
10. Robertson, T.; Bunel, F.; Roberts, M. Fluorescein Derivatives in Intravital Fluorescence Imaging. Cells 2013,
2, 591–606, doi:10.3390/cells2030591.
11. Singh, K.; Rotaru, A.M.; Beharry, A.A. Fluorescent Chemosensors as Future Tools for Cancer Biology. ACS
Chem. Biol. 2018, 13, 1785–1798, doi:10.1021/acschembio.8b00014.
12. Zhu, S.; Tian, R.; Antaris, A.L.; Chen, X.; Dai, H. Near-Infrared-II Molecular Dyes for Cancer Imaging and
Surgery. Adv. Mater. 2019, 31, 1–25, doi:10.1002/adma.201900321.
13. Bui, H.T.; Mai, D.K.; Kim, B.; Choi, K.-H.; Park, B.J.; Kim, H.-J.; Cho, S. Effect of Substituents on the
Photophysical Properties and Bioimaging Application of BODIPY Derivatives with Triphenylamine
Substituents. J. Phys. Chem. B 2019, 123, 5601–5607, doi:10.1021/acs.jpcb.9b04782.
14. Kolmakov, K.; Belov, V.N.; Bierwagen, J.; Ringemann, C.; Müller, V.; Eggeling, C.; Hell, S.W. Red-emitting
rhodamine dyes for fluorescence microscopy and nanoscopy. Chem. A Eur. J. 2010, 16, 158–166,
doi:10.1002/chem.200902309.
15. Kalepu, S.; Nekkanti, V. Insoluble drug delivery strategies: Review of recent advances and business
prospects. Acta Pharm. Sin. B 2015, 5, 442–453, doi:10.1016/j.apsb.2015.07.003.
16. Mariam, J.; Sivakami, S.; Dongre, P.M. Albumin corona on nanoparticles–a strategic approach in drug
delivery. Drug Deliv. 2016, 23, 2668–2676, doi:10.3109/10717544.2015.1048488.
17. Manzano, M.; Vallet-Regí, M. Mesoporous silica nanoparticles in nanomedicine applications. J. Mater. Sci.
Mater. Med. 2018, 29, doi:10.1007/s10856-018-6069-x.
18. Blanco, E.; Shen, H.; Ferrari, M. Principles of nanoparticle design for overcoming biological barriers to drug
delivery. Nat. Biotechnol. 2015, 33, 941–951, doi:10.1038/nbt.3330.
19. Mirabello, V.; Calatayud, D.G.; Arrowsmith, R.L.; Ge, H.; Pascu, S.I. Metallic nanoparticles as synthetic
building blocks for cancer diagnostics: From materials design to molecular imaging applications. J. Mater.
Chem. B 2015, 3, 5657–5672, doi:10.1039/C5TB00841G.
20. Bobo, D.; Robinson, K.J.; Islam, J.; Thurecht, K.J.; Corrie, S.R. Nanoparticle-Based Medicines: A Review of
FDA-Approved Materials and Clinical Trials to Date. Pharm. Res. 2016, 33, 2373–2387, doi:10.1007/s11095-
016-1958-5.
Sensors 2020, 20, 5590 12 of 15
21. Zhuang, Y.; Zhao, L.; Zheng, L.; Hu, Y.; Ding, L.; Li, X.; Liu, C.; Zhao, J.; Shi, X.; Guo, R. LAPONITE-
Polyethylenimine Based Theranostic Nanoplatform for Tumor-Targeting CT Imaging and Chemotherapy.
ACS Biomater. Sci. Eng. 2017, 3, 431–442, doi:10.1021/acsbiomaterials.6b00528.
22. Alexis, F.; Pridgen, E.; Molnar, L.K.; Farokhzad, O.C. Factors affecting the clearance and biodistribution of
polymeric nanoparticles. Mol. Pharm. 2008, 5, 505–515, doi:10.1021/mp800051m.
23. Mousa, M.; Evans, N.D.; Oreffo, R.O.C.; Dawson, J.I. Clay nanoparticles for regenerative medicine and
biomaterial design: A review of clay bioactivity. Biomaterials 2018, 159, 204–214,
doi:10.1016/j.biomaterials.2017.12.024.
24. Rudramurthy, G.R.; Swamy, M.K. Potential applications of engineered nanoparticles in medicine and
biology: An update. JBIC J. Biol. Inorg. Chem. 2018, 23, 1185–1204, doi:10.1007/s00775-018-1600-6.
25. Conde, J.; Dias, J.T.; Grazú, V.; Moros, M.; Baptista, P.V.; de la Fuente, J.M. Revisiting 30 years of
biofunctionalization and surface chemistry of inorganic nanoparticles for nanomedicine. Front. Chem. 2014,
2, 48, doi:10.3389/fchem.2014.00048.
26. Adumeau, L.; Genevois, C.; Roudier, L.; Schatz, C.; Couillaud, F.; Mornet, S. Impact of surface grafting
density of PEG macromolecules on dually fluorescent silica nanoparticles used for the in vivo imaging of
subcutaneous tumors. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 1587–1596,
doi:10.1016/j.bbagen.2017.01.036.
27. Dawidczyk, C.M.; Kim, C.; Park, J.H.; Russell, L.M.; Lee, K.H.; Pomper, M.G.; Searson, P.C. State-of-the-art
in design rules for drug delivery platforms: Lessons learned from FDA-approved nanomedicines. J. Control.
Release 2014, 187, 133–144, doi:10.1016/j.jconrel.2014.05.036.
28. Ribeiro, T.; Prazeres, T.J. V.; Moffitt, M.; Farinha, J.P.S. Enhanced Photoluminescence from Micellar
Assemblies of Cadmium Sulfide Quantum Dots and Gold Nanoparticles. J. Phys. Chem. C 2013, 117, 3122–
3133, doi:10.1021/jp311200r.
29. Pérez, N.; Ruiz-Rubio, L.; Vilas, J.L.; Rodríguez, M.; Martinez-Martinez, V.; León, L.M. Synthesis and
characterization of near-infrared fluorescent and magnetic iron zero-valent nanoparticles. J. Photochem.
Photobiol. A Chem. 2016, 315, 1–7, doi:10.1016/j.jphotochem.2015.09.004.
30. Jeevanandam, J.; Barhoum, A.; Chan, Y.S.; Dufresne, A.; Danquah, M.K. Review on nanoparticles and
nanostructured materials: History, sources, toxicity and regulations. Beilstein J. Nanotechnol. 2018, 9,
1050–1074, doi:10.3762/bjnano.9.98.
31. Malvindi, M.A.; Brunetti, V.; Vecchio, G.; Galeone, A.; Cingolani, R.; Pompa, P.P. SiO2 nanoparticles
biocompatibility and their potential for gene delivery and silencing. Nanoscale 2012, 4, 486,
doi:10.1039/c1nr11269d.
32. Wheeler, P.A.; Wang, J.; Baker, J.; Mathias, L.J. Synthesis and characterization of covalently functionalized
laponite clay. Chem. Mater. 2005, 17, 3012–3018, doi:10.1021/cm050306a.
33. Epelde-Elezcano, N.; Prieto-Montero, R.; Martínez-Martínez, V.; Ortiz, M.J.; Prieto-Castañeda, A.; Peña-
Cabrera, E.; Belmonte-Vázquez, J.L.; López-Arbeloa, I.; Brown, R.; Lacombe, S. Adapting BODIPYs to
singlet oxygen production on silica nanoparticles. Phys. Chem. Chem. Phys. 2017, 19, 13746–13755,
doi:10.1039/C7CP01333G.
34. Martins Estevão, B.; Miletto, I.; Marchese, L.; Gianotti, E. Optimized Rhodamine B labeled mesoporous
silica nanoparticles as fluorescent scaffolds for the immobilization of photosensitizers: A theranostic
platform for optical imaging and photodynamic therapy. Phys. Chem. Chem. Phys. 2016, 18, 9042–9052,
doi:10.1039/C6CP00906A.
35. Ribeiro, T.; Raja, S.; Rodrigues, A.S.; Fernandes, F.; Farinha, J.P.S.; Baleizão, C. High performance NIR
fluorescent silica nanoparticles for bioimaging. RSC Adv. 2013, 3, 9171, doi:10.1039/c3ra41199k.
36. Burns, A.; Ow, H.; Wiesner, U. Fluorescent core-shell silica nanoparticles: Towards “Lab on a Particle”
architectures for nanobiotechnology. Chem. Soc. Rev. 2006, 35, 1028–1042, doi:10.1039/b600562b.
37. Burns, A.A.; Vider, J.; Ow, H.; Here, E.; Oula, P.M.; Baumgart, M.; Larson, S.M.; Wiesner, U.; Bradbury, M.
Fluorescent silica nanoparticles with efficient urinary excretion for nanomedicine. Nano Lett. 2009, 9, 442–
448, doi:10.1021/nl803405h.
38. Tapec, R.; Zhao, X.J.; Tan, W. Development of Organic Dye-Doped Silica Nanoparticles for Bioanalysis and
Biosensors. J. Nanosci. Nanotechnol. 2002, 2, 405–409, doi:10.1166/jnn.2002.114.
39. Tavernaro, I.; Cavelius, C.; Peuschel, H.; Kraegeloh, A. Bright fluorescent silica-nanoparticle probes for
high-resolution STED and confocal microscopy. Beilstein J. Nanotechnol. 2017, 8, 1283–1296,
doi:10.3762/bjnano.8.130.
Sensors 2020, 20, 5590 13 of 15
40. Gubala, V.; Giovannini, G.; Kunc, F.; Monopoli, M.P.; Moore, C.J. Cancer Nanotechnology; Springer Vienna:
Viena, Austria, 2020; pp. 1–43.
41. Sokolov, I.; Naik, S. Novel fluorescent silica nanoparticles: Towards ultrabright silica nanoparticles. Small
2008, 4, 934–939, doi:10.1002/smll.200700236.
42. Bonacchi, S.; Genovese, D.; Juris, R.; Montalti, M.; Prodi, L.; Rampazzo, E.; Zaccheroni, N. Luminescent
silica nanoparticles: Extending the frontiers of brightness. Angew. Chemie Int. Ed. 2011, 50, 4056–4066,
doi:10.1002/anie.201004996.
43. Zhang, W.H.; Hu, X.X.; Zhang, X.B. Dye-doped fluorescent silica nanoparticles for live cell and in vivo
bioimaging. Nanomaterials 2016, 6, 81, doi:10.3390/nano6050081.
44. Rampazzo, E.; Voltan, R.; Petrizza, L.; Zaccheroni, N.; Prodi, L.; Casciano, F.; Zauli, G.; Secchiero, P. Proper
design of silica nanoparticles combines high brightness, lack of cytotoxicity and efficient cell endocytosis.
Nanoscale 2013, 5, 7897–7905, doi:10.1039/c3nr02563b.
45. Jiao, L.; Song, F.; Zhang, B.; Ning, H.; Cui, J.; Peng, X. Improving the brightness and photostability of NIR
fluorescent silica nanoparticles through rational fine-tuning of the covalent encapsulation methods. J.
Mater. Chem. B 2017, 5, 5278–5283, doi:10.1039/C7TB00856B.
46. Ow, H.; Larson, D.R.; Srivastava, M.; Baird, B. a.; Webb, W.W.; Wiesnert, U. Bright and stable core-shell
fluorescent silica nanoparticles. Nano Lett. 2005, 5, 113–117, doi:10.1021/nl0482478.
47. Gianotti, E.; MartinsEstevão, B.; Cucinotta, F.; Hioka, N.; Rizzi, M.; Renò, F.; Marchese, L. An efficient rose
bengal based nanoplatform for photodynamic therapy. Chem. A Eur. J. 2014, 10921–10925,
doi:10.1002/chem.201404296.
48. Deng, T.; Li, J.S.; Jiang, J.H.; Shen, G.L.; Yu, R.Q. Preparation of near-IR fluorescent nanoparticles for
fluorescence- anisotropy-based immunoagglutination assay in whole blood. Adv. Funct. Mater. 2006, 16,
2147–2155, doi:10.1002/adfm.200600149.
49. Shahabi, S.; Treccani, L.; Rezwan, K. A comparative study of three different synthesis routes for hydrophilic
fluorophore-doped silica nanoparticles. J. Nanoparticle Res. 2016, 18, 1–13, doi:10.1007/s11051-016-3334-0.
50. Auger, A.; Samuel, J.; Poncelet, O.; Raccurt, O. A comparative study of non-covalent encapsulation
methods for organic dyes into silica nanoparticles. Nanoscale Res. Lett. 2011, 6, 328, doi:10.1186/1556-276X-
6-328.
51. Wang, L.; Du, W.; Hu, Z.; Uvdal, K.; Li, L.; Huang, W. Hybrid Rhodamine Fluorophores in the Visible/NIR
Region for Biological Imaging. Angew. Chemie Int. Ed. 2019, 58, 14026–14043, doi:10.1002/anie.201901061.
52. Manzano, M.; Vallet-Regí, M. Mesoporous Silica Nanoparticles for Drug Delivery. Adv. Funct. Mater. 2020,
30, 1902634, doi:10.1002/adfm.201902634.
53. Santiago, A.M.; Ribeiro, T.; Rodrigues, A.S.; Ribeiro, B.; Frade, R.F.M.; Baleizão, C.; Farinha, J.P.S.
Multifunctional Hybrid Silica Nanoparticles with a Fluorescent Core and Active Targeting Shell for
Fluorescence Imaging Biodiagnostic Applications. Eur. J. Inorg. Chem. 2015, 2015, 4579–4587,
doi:10.1002/ejic.201500580.
54. Peerzade, S.A.M.A.; Qin, X.; Laroche, F.J.F.; Palantavida, S.; Dokukin, M.; Peng, B.; Feng, H.; Sokolov, I.
Ultrabright fluorescent silica nanoparticles for: In vivo targeting of xenografted human tumors and cancer
cells in zebrafish. Nanoscale 2019, 11, 22316–22327, doi:10.1039/c9nr06371d.
55. Ma, K.; Zhang, D.; Cong, Y.; Wiesner, U. Elucidating the Mechanism of Silica Nanoparticle PEGylation
Processes Using Fluorescence Correlation Spectroscopies. Chem. Mater. 2016, 28, 1537–1545,
doi:10.1021/acs.chemmater.6b00030.
56. Vijay Fernández, M.; Javaida, F.; Chudasama; Vijay Advances in targeting the folate receptor in
thetreatment/imaging of cancers. Chem. Sci. 2018, 9, 790–810, doi:10.1039/c7sc04004k.
57. Stallivieri, A.; Baros, F.; Jetpisbayeva, G.; Frochot, B.M. and C. The Interest of Folic Acid in Targeted
Photodynamic Therapy. Curr. Med. Chem. 2015, 22, 3185–3207, doi:10.2174/0929867322666150729113912.
58. Yang, H.; Lou, C.; Xu, M.; Wu, C.; Miyoshi, H.; Liu, Y. Investigation of folate-conjugated fluorescent silica
nanoparticles for targeting delivery to folate receptor-positive tumors and their internalization mechanism.
Int. J. Nanomedicine 2011, 6, 2023–2032, doi:10.2147/ijn.s24792.
59. Mamaeva, V.; Sahlgren, C.; Lindén, M. Mesoporous silica nanoparticles in medicine-Recent advances. Adv.
Drug Deliv. Rev. 2013, 65, 689–702, doi:10.1016/j.addr.2012.07.018.
60. Yoo, J.; Park, C.; Yi, G.; Lee, D.; Koo, H. Active targeting strategies using biological ligands for nanoparticle
drug delivery systems. Cancers (Basel) 2019, 11, 640, doi:10.3390/cancers11050640.
Sensors 2020, 20, 5590 14 of 15
61. Qian, J.; Li, X.; Wei, M.; Gao, X.; Xu, Z.; He, S. Bio-molecule-conjugated fluorescent organically modified
silica nanoparticles as optical probes for cancer cell imaging. Opt. Express 2008, 16, 19568–19578.
62. Lu, J.; Liong, M.; Li, Z.; Zink, J.I.; Tamanoi, F. Biocompatibility, biodistribution, and drug-delivery
efficiency of mesoporous silica nanoparticles for cancer therapy in animals. Small 2010, 6, 1794–1805,
doi:10.1002/smll.201000538.
63. Liong, M.; Lu, J.; Kovochich, M.; Xia, T.; Ruehm, S.G.; Nel, A.E.; Tamanoi, F.; Zink, J.I. Multifunctional
inorganic nanoparticles for imaging, targeting, and drug delivery. ACS Nano 2008, 2, 889–896,
doi:10.1021/nn800072t.
64. Cotí, K.K.; Belowich, M.E.; Liong, M.; Ambrogio, M.W.; Lau, Y. a; Khatib, H. a; Zink, J.I.; Khashab, N.M.;
Stoddart, J.F. Mechanised nanoparticles for drug delivery. Nanoscale 2009, 1, 16–39, doi:10.1039/b9nr00162j.
65. Castillo, R.R.; Vallet-Regí, M. Functional mesoporous silica nanocomposites: Biomedical applications and
biosafety. Int. J. Mol. Sci. 2019, 20, 929, doi:10.3390/ijms20040929.
66. Son, J.; Yang, S.M.; Yi, G.; Roh, Y.J.; Park, H.; Park, J.M.; Choi, M.; Koo, H. Folate-modified PLGA
nanoparticles for tumor-targeted delivery of pheophorbide a in vivo. Biochem. Biophys. Res. Commun. 2018,
498, 523–528, doi:10.1016/j.bbrc.2018.03.013.
67. Vallet-Regí, M.; Colilla, M.; Izquierdo-Barba, I.; Manzano, M. Mesoporous silica nanoparticles for drug
delivery: Current insights. Molecules 2018, 23, 1–19, doi:10.3390/molecules23010047.
68. Cheng, S.-H.; Lee, C.-H.; Yang, C.-S.; Tseng, F.-G.; Mou, C.-Y.; Lo, L.-W. Mesoporous silica nanoparticles
functionalized with an oxygen-sensing probe for cell photodynamic therapy: Potential cancer theranostics.
J. Mater. Chem. 2009, 19, 1252, doi:10.1039/b816636f.
69. Lin, Y.; Haynes, C.L. Impacts of Mesoporous Silica Nanoparticle Size, Pore Ordering, and Pore Integrity on
Hemolytic Activity. J. Am. Chem. Soc. 2010, 132, 4834–4842.
70. Stöber, W.; Fink, A.; Bohn, E. Controlled growth of monodisperse silica spheres in the micron size range. J.
Colloid Interface Sci. 1968, 26, 62–69, doi:10.1016/0021-9797(68)90272-5.
71. Mai, W.X.; Meng, H. Mesoporous silica nanoparticles: A multifunctional nano therapeutic system. Integr.
Biol. 2013, 5, 19–28, doi:10.1039/c2ib20137b.
72. Zhang, X.F.; Zhang, Y.; Liu, L. Fluorescence lifetimes and quantum yields of ten rhodamine derivatives:
Structural effect on emission mechanism in different solvents. J. Lumin. 2014, 145, 448–453,
doi:10.1016/j.jlumin.2013.07.066.
73. Magde, D.; Rojas, G.E.; Seybold, P.G. Solvent dependence of the fluorescence lifetimes of xanthene dyes.
Photochem. Photobiol. 1999, 70, 737–744, doi:10.1111/j.1751-1097.1999.tb08277.x.
74. López Arbeloa, T.; López Arbeloa, F.; Hernández Bartolomé, P.; López Arbeloa, I. On the mechanism of
radiationless deactivation of rhodamines. Chem. Phys. 1992, 160, 123–130, doi:10.1016/0301-0104(92)87095-Q.
75. Cazade, P.A.; Bordat, P.; Blanc, S.; Baraille, I.; Brown, R. Possible artifacts of fluorescence of dyes in
concentrated materials: Comment on “highly fluorescent rhodamine B nanoparticles entrapped in hybrid
glasses.” Langmuir 2008, 24, 2252–2257, doi:10.1021/la702291h.
76. Cho, E.B.; Volkov, D.O.; Sokolov, I. Ultrabright fluorescent silica mesoporous silica nanoparticles: Control
of particle size and dye loading. Adv. Funct. Mater. 2011, 21, 3129–3135, doi:10.1002/adfm.201100311.
77. Medintz, I.L.; Uyeda, H.T.; Goldman, E.R.; Mattoussi, H. Quantum dot bioconjugates for imaging, labelling
and sensing. Nat. Mater. 2005, 4, 435–446, doi:10.1038/nmat1390.
78. Resch-Genger, U.; Grabolle, M.; Cavaliere-Jaricot, S.; Nitschke, R.; Nann, T. Quantum dots versus organic
dyes as fluorescent labels. Nat. Methods 2008, 5, 763–775, doi:10.1038/nmeth.1248.
79. Vlachy, N.; Touraud, D.; Heilmann, J.; Kunz, W. Determining the cytotoxicity of catanionic surfactant
mixtures on HeLa cells. Colloids Surfaces B Biointerfaces 2009, 70, 278–280, doi:10.1016/j.colsurfb.2008.12.038.
80. Hwang, T.L.; Sung, C.T.; Aljuffali, I.A.; Chang, Y.T.; Fang, J.Y. Cationic surfactants in the form of
nanoparticles and micelles elicit different human neutrophil responses: A toxicological study. Colloids
Surfaces B Biointerfaces 2014, 114, 334–341, doi:10.1016/j.colsurfb.2013.10.021.
81. Connor, E.E.; Mwamuka, J.; Gole, A.; Murphy, C.J.; Wyatt, M.D. Gold nanoparticles are taken up by human
cells but do not cause acute cytotoxicity. Small 2005, 1, 325–327, doi:10.1002/smll.200400093.
82. Jafari, S.; Derakhshankhah, H.; Alaei, L.; Fattahi, A.; Varnamkhasti, B.S.; Saboury, A.A. Mesoporous silica
nanoparticles for therapeutic/diagnostic applications. Biomed. Pharmacother. 2019, 109, 1100–1111,
doi:10.1016/j.biopha.2018.10.167.
Sensors 2020, 20, 5590 15 of 15
83. Bakhshian Nik, A.; Zare, H.; Razavi, S.; Mohammadi, H.; Torab Ahmadi, P.; Yazdani, N.; Bayandori, M.;
Rabiee, N.; Izadi Mobarakeh, J. Smart drug delivery: Capping strategies for mesoporous silica
nanoparticles. Microporous Mesoporous Mater. 2020, 299, 110115, doi:10.1016/j.micromeso.2020.110115.
84. Off, M.K.; Steindal, A.E.; Porojnicu, A.C.; Juzeniene, A.; Vorobey, A.; Johnsson, A.; Moan, J. Ultraviolet
photodegradation of folic acid. J. Photochem. Photobiol. B Biol. 2005, 80, 47–55,
doi:10.1016/j.jphotobiol.2005.03.001.
85. Kruger, C.A.; Abrahamse, H. Utilisation of targeted nanoparticle photosensitiser drug delivery systems for
the enhancement of photodynamic therapy. Molecules 2018, 23, doi:10.3390/molecules23102628.
86. Salazar, M.D.A.; Ratnam, M. The folate receptor: What does it promise in tissue-targeted therapeutics?
Cancer Metastasis Rev. 2007, 26, 141–152, doi:10.1007/s10555-007-9048-0.
87. Kamen, B.A.; Capdevila, A. Receptor-mediated folate accumulation is regulated by the cellular folate
content. Proc. Natl. Acad. Sci. USA 1986, 83, 5983–5987, doi:10.1073/pnas.83.16.5983.
88. Luhrs, C.A.; Slomiany, B.L. A human membrane-associated folate binding protein is anchored by a
glycosyl-phosphatidylinositol tail. J. Biol. Chem. 1989, 264, 21446–21449.
89. Hilgenbrink, A.R.; Low, P.S. Folate receptor-mediated drug targeting: From therapeutics to diagnostics. J.
Pharm. Sci. 2005, 94, 2135–2146, doi:10.1002/jps.20457.
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... Thus, tethering FA on the nanoparticle's surface promotes a higher cellular uptake via endocytosis [101][102][103][104]. In fact, our recently published study quantitatively demonstrated a significantly higher accumulation of FAfunctionalized fluorescent MSNs compared to nanoparticles without FA in HeLa cells [105]. ...
... Mesoporous silica nanoparticles with a suitable size for medical applications and particularly for PDT [98,106] were synthesized by the modified Stöber method [107] as described elsewhere [105]. The external surface of mesoporous nanoparticles surface was functionalized with amino group (NH-MSN) or carboxylic group (COOH-MSN). ...
... For the samples RB-PEG 2000 -NP-a, RB-PEG 5000 -NP-a, and RB-PEG-NP-c (RB grafted at the hydroxyl groups of MSNs, Table 2) the estimated RB loading was equal, and consequently, the observed dye aggregation in these samples should be assigned to interparticle processes, as supported by Zpot values and previously attributed to a lower presence of PEG molecules at the external surface. Sample RB-PEG-NP-b, with RB loading at the amine groups half of that obtained for the RB at the hydroxyl groups (sample RB-PEG-NP-b vs. sample RB-PEG-NP-a in Table 2), showed a narrower absorption band, not much different from that recorded for RB in diluted solution [105]. However, reducing the cargo of PS per nanoparticle would compel a higher concentration of nanoparticles per volume to reach effective PS doses for PDT in the cells, which would also promote particle-particle agglomeration. ...
Article
Full-text available
BODIPY dyes have recently attracted attention as potential photosensitizers. In this work, commercial and novel photosensitizers (PSs) based on BODIPY chromophores (haloBODIPYs and orthogonal dimers strategically designed with intense bands in the blue, green or red region of the visible spectra and high singlet oxygen production) were covalently linked to mesoporous silica nanoparticles (MSNs) further functionalized with PEG and folic acid (FA). MSNs approximately 50 nm in size with different functional groups were synthesized to allow multiple alternatives of PS-PEG-FA decoration of their external surface. Different combinations varying the type of PS (commercial Rose Bengal, Thionine and Chlorine e6 or custom-made BODIPY-based), the linkage design, and the length of PEG are detailed. All the nanosystems were physicochemically characterized (morphology, diameter, size distribution and PS loaded amount) and photophysically studied (absorption capacity, fluorescence efficiency, and singlet oxygen production) in suspension. For the most promising PS-PEG-FA silica nanoplatforms, the biocompatibility in dark conditions and the phototoxicity under suitable irradiation wavelengths (blue, green, or red) at regulated light doses (10–15 J/cm2) were compared with PSs free in solution in HeLa cells in vitro.
... Currently, nanoparticles are widely employed for the efficient transportation of fluorescent probes in various applications (10). In comparison to conventional biomarker techniques, nanoprobes offer several advantages including uniform particle size distribution, stable optical properties, minimal cytotoxicity, excellent water solubility, and remarkable biocompatibility (11). Therefore, they exhibit enhanced resolution, heightened fluorescence intensity, and superior detection capabilities. ...
Article
Full-text available
Tumor-specific fluorescent probes must fulfill the dual requirements of targeted accumulation within tumors and high-resolution imaging capabilities. To achieve both tumor-targeted accumulation and high-resolution imaging performance, we developed a composite comprising an acid-responsive bodipy conjugated to amphiphilic PEG-b-PLA polymer, along with folic acid (FA)-modified PEG-b-PLA as a targeting moiety for active tumor-specific accumulation. Finally, a novel assembly of hybrid fluorescent nanoparticles was successfully synthesized by integrating these two components, demonstrating exceptional responsiveness to acidic conditions for fluorescence excitation and remarkable tumor-targeted accumulation capabilities. We conducted comprehensive in vitro and in vivo investigations employing techniques such as analysis of physicochemical properties, fluorescence-based probes detection at varying pH levels, assessment of in vitro cytotoxicity, evaluation of cellular uptake capacity, analysis of lysosomal co-localization imaging, examination of tumor fluorescence images in vivo, and investigation of biological distribution patterns. The results demonstrated that the acid-responsive nanofluorescence probe we designed and synthesized possesses desirable physical and chemical characteristics, including a small particle size and low cytotoxicity. Moreover, it exhibits rapid real-time response to acidic environments and displays enhanced fluorescence intensity, enabling the real-time tracking of probe entry into tumor cells as well as intracellular lysozyme accumulation. We achieved highly specific in vivo tumor visualization by combining nanoprobes targeting folate receptor. Through imaging cervical tumor mice, we demonstrated the precise imaging performance and high targeted accumulation of FA-targeted nanofluorescence probes in tumor tissue. Furthermore, we confirmed the in vivo safety of the FA-targeted nanofluorescence probe through biological distribution analysis. These findings highlight the potential widespread application of FA-targeted acid-responsive nanofluorescence probes for selective imaging of tumor cells and tissues.
... Apart from the detection of Hg 2+ ions in the environmental water sample, the fluorescent Si-NPs have been widely used in the bio-imaging of cancer cells, [47] they have the potential to be used as electroluminescent nanosensors for detection of antigens. [22] The other applications of fluorescent Si-NPs are-pathogens [14], local environment [32], [46,66]. ...
Article
Full-text available
The good “natural” surroundings and health of animals and plants on the earth will be maintained by the early detection and removal of potentially hazardous heavy metal (Hg²⁺) ions. Several conventional and advanced techniques such as GC–MS, ICP–MS, AAS, HPLC, Surface-Enhanced Raman Spectroscopy (SERS), Colorimetric, Electrochemical and fluorescence-based sensors were employed for this ecological remediation. The colorimetric and fluorescent detection methods are quickly responsive, easy to probe synthesis/availability, cost-effective and highly sensitive. The SiO2-containing hybrid nanomaterials are fluorescent, biocompatible, cost-effective, easily multi-functionalizeable, highly sensitive and selective to the analytes. Considering these advantages and the flexibility of modifications, the hybrid SiO2 nanomaterial developed. In the present short review, we have revised and evaluated the recent modifications (2017–2023) in the silica nanoparticle relative to their functionality, methods of synthesis, range of detections, and limit of detections. In addition, the general fluorescence sensing mechanisms of this studied hybrid silicon nanoparticles to the Hg²⁺ ions have been discussed. Further, the future perspective of research for modified silica/silicon nanomaterials is discussed. The distance-dependant change in the fluorescence of these modified Si-NPs demonstrate that the doping/mixing of Si-NPs with fluorophore leads to the fluorescence ‘Turn-On’ response with few exceptions while covalently linked fluorophores with Si-NPs show ‘Turn-off’ response.
... Dye-loaded nanoparticles demonstrate superior efficacy for bioimaging applications due to their increased photostability and enhanced fluorescence [19,20]. Yet, despite the promise, studies reported to date show a number of challenges, such as dye leakage that contributes to decreased fluorescence intensity and lower signal to noise ratios in vivo [21]. ...
Article
Full-text available
Modified fluorescent nanoparticles continue to emerge as promising candidates for drug delivery, bioimaging, and labeling tools for various biomedical applications. The ability of nanomaterials to fluorescently label cells allow for the enhanced detection and understanding of diseases. Silica nanoparticles have a variety of unique properties that can be harnessed for many different applications, causing their increased popularity. In combination with an organic dye, fluorescent nanoparticles demonstrate a vast range of advantageous properties including long photostability, surface modification, and signal amplification, thus allowing ease of manipulation to best suit bioimaging purposes. In this study, the Stöber method with tetraethyl orthosilicate (TEOS) and a fluorescent dye sulfo-Cy5-amine was used to synthesize fluorescent silica nanoparticles. The fluorescence spectra, zeta potential, quantum yield, cytotoxicity, and photostability were evaluated. The increased intracellular uptake and photostability of the dye-silica nanoparticles show their potential for bioimaging.
... Also, fluorescent silica nanoparticles with specific functionalization were developed and prepared to exclusively target cancer cells for bioimaging assessment. The surface of the nanoparticles was additionally modified with a 2000 Da PEG layer and folic acid to ensure excellent water stability and improve the selectivity towards cancer cells, respectively [45]. During cancer cell metastasis, tumor cells interact with bone components, including inorganic minerals, which may lead to bone mineralization (Chiou et al. [46]). ...
... Wang et al. [142] grafted PhENH2, an aggregation-induced emission (AIE) dye, onto SiNPs via chemical bonds with FA and molybdenum disulfide (MoS 2 ) nanosheets was embedded into SiNPs, which exhibited stable fluorescence for photothermal therapy in breast cancer. Prieto-Montero et al. reported functionalized SiNPs with both covalently bound and physically embedded rhodamine 101 for the visual targeting of HeLa cells [143]. FA-modified SiNPs embedded with 5,10,15,20-tetrakis(1-methyl-4-pyridinio)porphyrin tetra(p-toluenesulfonate), a photosensitizer, achieved targeted two-photon excitation fluorescence imaging in human breast cancer cells [144]. ...
Article
Full-text available
Silica nanoparticles (SiNPs) are composed of silicon dioxide, the most abundant compound on Earth, and are used widely in many applications including the food industry, synthetic processes, medical diagnosis, and drug delivery due to their controllable particle size, large surface area, and great biocompatibility. Building on basic synthetic methods, convenient and economical strategies have been developed for the synthesis of SiNPs. Numerous studies have assessed the biomedical applications of SiNPs, including the surface and structural modification of SiNPs to target various cancers and diagnose diseases. However, studies on the in vitro and in vivo toxicity of SiNPs remain in the exploratory stage, and the toxicity mechanisms of SiNPs are poorly understood. This review covers recent studies on the biomedical applications of SiNPs, including their uses in drug delivery systems to diagnose and treat various diseases in the human body. SiNP toxicity is discussed in terms of the different systems of the human body and the individual organs in those systems. This comprehensive review includes both fundamental discoveries and exploratory progress in SiNP research that may lead to practical developments in the future.
Article
Multiplexed bead assays for solution‐phase biosensing often encounter cross‐over reactions during signal amplification, leading to false positives and high background signals. Current solutions involve complex, custom‐designed equipment or costly setups, limiting their application in simple laboratory environments. In this study, we introduce a straightforward protocol to adapt a multiplexed single‐bead assay to standard fluorescence imaging plates, enabling the simultaneous analysis of thousands of reactions per plate. Our approach focuses on the design and synthesis of bright fluorescent and magnetic microspheres (MagSiGlow) with multiple fluorescent wavelengths serving as detection markers. This imaging‐based single‐bead assay, combined with a scripted algorithm, allows the detection, segmentation, and co‐localization on average of 7500 microspheres per field of view across five imaging channels in less than one second. We demonstrate the effectiveness of this method with remarkable sensitivity (detection limits of 100 pg/mL). This technique showed over 85% reduction in signal cross‐over to the solution‐based approaches after the concurrent detection of tumor‐associated protein biomarkers. This approach promises to significantly advance high throughput biosensing in diverse laboratory settings.
Article
Full-text available
Multiplexed bead assays for solution‐phase biosensing often encounter cross‐over reactions during signal amplification steps, leading to unwanted false positive and high background signals. Current solutions involve complex custom‐designed and costly equipment, limiting their application in simple laboratory setup. In this study, we introduce a straightforward protocol to adapt a multiplexed single‐bead assay to standard fluorescence imaging plates, enabling the simultaneous analysis of thousands of reactions per plate. This approach focuses on the design and synthesis of bright fluorescent and magnetic microspheres (MagSiGlow) with multiple fluorescent wavelengths serving as unique detection markers. The imaging‐based, single‐bead assay, combined with a scripted algorithm, allows the detection, segmentation, and co‐localization on average of 7500 microspheres per field of view across five imaging channels in less than one second. We demonstrate the effectiveness of this method with remarkable sensitivity at low protein detection limits (100 pg/mL). This technique showed over 85 % reduction in signal cross‐over to the solution‐based method after the concurrent detection of tumor‐associated protein biomarkers. This approach holds the promise of substantially enhancing high throughput biosensing for multiple targets, seamlessly integrating with rapid image analysis algorithms.
Article
Carbon dots (CDs) are nanomaterials with excellent properties, including good biocompatibility, small size, ideal photoluminescence and surface modification, and are becoming one of the most attractive nanomaterials for the imaging, detection and treatment of tumors. Based on these advantages, CDs can be combined other materials to obtain composite particles with improved, even new, performance, mainly in photothermal and photodynamic therapies. This paper reviews the research progress of CDs and their composites in targeted tumor imaging, detection, diagnosis, drug delivery and tumor killing. It also discusses and proposes the challenges and perspectives of their future applications in these fields. This review provides ideas for future applications of novel CD-based materials in the diagnosis and treatment of cancer.
Article
Full-text available
In recent years, nanomedicine has emerged at the forefront of nanotechnology, generating great expectations in the biomedical field. Researchers are developing novel nanoparticles for both diagnostic applications using imaging technology and treatment purposes through drug delivery technologies. Among all the available nanoparticles, inorganic mesoporous silica nanoparticles are the newcomers to the field, contributing with their unique and superlative properties. A brief overview of the most recent progress in the synthesis of mesoporous silica nanoparticles and their use as drug delivery nanocarriers is provided. The latest trends in this type of nanoparticles and their use in modern medicine are discussed, highlighting the significant impact that this technology might have in the near future. A brief overview of the most recent progress in the synthesis of mesoporous silica nanoparticles and their use as drug delivery nanocarriers is presented here. An update on the latest trends in this type of nanoparticles and their use in modern medicine is provided, highlighting the significant impact that this technology might have in the near future.
Article
Full-text available
Targeting nanoparticle (NP) carriers to sites of disease is critical for their successful use as drug delivery systems. Along with optimization of physicochemical properties, researchers have focused on surface modification of NPs with biological ligands. Such ligands can bind specific receptors on the surface of target cells. Furthermore, biological ligands can facilitate uptake of modified NPs, which is referred to as ‘active targeting’ of NPs. In this review, we discuss recent applications of biological ligands including proteins, polysaccharides, aptamers, peptides, and small molecules for NP-mediated drug delivery. We prioritized studies that have demonstrated targeting in animals over in vitro studies. We expect that this review will assist biomedical researchers working with NPs for drug delivery and imaging.
Article
Full-text available
Fluorescence bioimaging affords a vital tool for both researchers and surgeons to molecularly target a variety of biological tissues and processes. This review focuses on summarizing organic dyes emitting at a biological transparency window termed the near‐infrared‐II (NIR‐II) window, where minimal light interaction with the surrounding tissues allows photons to travel nearly unperturbed throughout the body. NIR‐II fluorescence imaging overcomes the penetration/contrast bottleneck of imaging in the visible region, making it a remarkable modality for early diagnosis of cancer and highly sensitive tumor surgery. Due to their convenient bioconjugation with peptides/antibodies, NIR‐II molecular dyes are desirable candidates for targeted cancer imaging, significantly overcoming the autofluorescence/scattering issues for deep tissue molecular imaging. To promote the clinical translation of NIR‐II bioimaging, advancements in the high‐performance small molecule–derived probes are critically important. Here, molecules with clinical potential for NIR‐II imaging are discussed, summarizing the synthesis and chemical structures of NIR‐II dyes, chemical and optical properties of NIR‐II dyes, bioconjugation and biological behavior of NIR‐II dyes, whole body imaging with NIR‐II dyes for cancer detection and surgery, as well as NIR‐II fluorescence microscopy imaging. A key perspective on the direction of NIR‐II molecular dyes for cancer imaging and surgery is also discussed. Among all existing near‐infrared (NIR)‐II fluorophores, the NIR‐II molecular dyes are the most remarkable in translating this imaging window into the clinical setting. Advanced NIR‐II dye‐derived bioconjugates will give doctors an unparalleled view into tissues for tumor detection at greater depths and contrast, allowing early detection during cancer screenings and solid tumor resection by delineation of the boundaries between healthy and cancerous tissues.
Article
Full-text available
Fluorophores and probes are invaluable for the visualization of the location and dynamics of gene expression, protein expression, and molecular interactions in complex living systems. Rhodamine dyes are often used as scaffolds in biological labeling and turn‐on fluorescence imaging. To date, their absorption and emission spectra have been expanded to cover the entire near‐infrared region (650–950 nm), which provides a more suitable optical window for monitoring biomolecular production, trafficking, and localization in real time. This review summarizes the development of rhodamine fluorophores since their discovery and provides strategies for modulating their absorption and emission spectra to generate specific bathochromic‐shifts. We also explain how larger Stokes shifts and dual‐emissions can be obtained from hybrid rhodamine dyes. These hybrid fluorophores can be classified into various categories based on structural features including the alkylation of amidogens, the substitution of the O atom of xanthene, and hybridization with other fluorophores.
Article
Full-text available
The rise and development of nanotechnology has enabled the creation of a wide number of systems with new and advantageous features to treat cancer. However, in many cases, the lone application of these new nanotherapeutics has proven not to be enough to achieve acceptable therapeutic efficacies. Hence, to avoid these limitations, the scientific community has embarked on the development of single formulations capable of combining functionalities. Among all possible components, silica—either solid or mesoporous—has become of importance as connecting and coating material for these new-generation therapeutic nanodevices. In the present review, the most recent examples of fully inorganic silica-based functional composites are visited, paying particular attention to those with potential biomedical applicability. Additionally, some highlights will be given with respect to their possible biosafety issues based on their chemical composition.
Article
Full-text available
Based on unique intrinsic properties of mesoporous silica nanoparticles (MSNs) such as high surface area, large pore size, good biocompatibility and biodegradability, stable aqueous dispersion, they have received much attention in the recent decades for their applications as a promising platform in the biomedicine field. These porous structures possess a pore size ranging from 2 to 50 nm which make them excellent candidates for various biomedical applications. Herein, at first we described the common approaches of cargo loading and release processes from MSNs. Then, the intracellular uptake, safety and cytotoxicity aspects of MSNs are discussed as well. This review also highlights the most recent advances in the biomedical applications of MSNs, including 1) MSNs-based carriers, 2) MSNs as bioimaging agents, 3) MSNs-based biosensors, 4) MSNs as therapeutic agents (photodynamic therapy), 5) MSN based quantum dots, 6) MSNs as platforms for upconverting nanoparticles, and 6) MSNs in tissue engineering.
Article
Systematic delivery of therapeutic agents to specific sites, with a stimulus-responsive drug release profile is currently a rapidly growing area. Mesoporous silica nanoparticles (MSNs) are the useful platforms as drug/gene delivery systems due to their unique properties including the ability to control the pore size, high porosity, and morphology, which can directly affect the mechanism and profile of drug release. The appropriate fabrication strategy can tailor the particle shape and size, leading to enhanced delivery and release mechanisms. The MSN surface can be modified by using either organic or inorganic molecules to induce smart and site-specific drug delivery and release. Furthermore, application of molecules that function as pore gatekeepers with the ability to uncap via physiochemical stimuli can enhance the efficiency of drug delivery and release. This report aims to highlight the recent efforts and developments of strategies applied to render MSNs smarter and more effective for drug delivery applications.
Article
New ultrabright fluorescent silica nanoparticles capable of fast targeting of epithelial tumors in vivo are presented. The synthesized folate-functionalized ultrabright particles of 30-40 nm are 230 times brighter than quantum dots (QD450) and 50% brighter than polymer dots of similar spectra (excitation 365 nm / emission 486 nm). To decrease non-specific targeting, particles are coated with polyethylene glycol (PEG). We demonstrate the targeting of xenographic human cervical epithelial tumors (HeLa cells) in-vivo using zebrafish as a model system. The particles show targeting tumors (and probably even individual HeLa cells) as small as 10-20 microns within 20-30 minutes after blood injection. To demonstrate the advantages of ultrabrightness, we repeated the experiments with similar but 200x less bright particles. Compared to those, ultrabright particles showed ~3x faster tumor detection and ~2x higher relative fluorescent contrast of tumors/cancer cells.
Article
Non-invasive diagnosis of cancer is often the key to effective treatment and patient survival. Saliva as a multi-constituent oral fluid comprises various disease signaling biomarkers, holds great potential for early-stage cancer diagnostics with cost-effective and easy collection, storage, transport and processing. Therefore, detection of biomarkers and proteins in the saliva samples is highly demand. The current review was performed using reliable internet database (mainly PubMed) to provide an overview of the most recent developments on non-invasive diagnosis of cancers in saliva and highlights main challenges and future prospects in sensing of the salivary biomarkers. The conventional detection methods of cancer biomarkers in saliva is discussed in the paper, however, the main focus is on non-invasive diagnosis of cancers in saliva using immunosensing (electrochemical, optical, piezoelectric), DNA based sensors, aptasensors and peptide based bio-assays The reviewed literature revealed that non-invasive cancer detection methods using the mentioned biosensors and without any processing of saliva sample offers a quick, sensitive, specific and cost effective analytical tool. Besides, salivary based detection methods can be used for simultaneous detection of panels of disease specific biomarkers in a real time manner or as home testing kits in near future.