Anda di halaman 1dari 12

Nanoscale

View Article Online


PAPER View Journal | View Issue

A smart pH-responsive nano-carrier as a drug


Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

Cite this: Nanoscale, 2017, 9, 9428


delivery system for the targeted delivery of ursolic
acid: suppresses cancer growth and metastasis by
modulating P53/MMP-9/PTEN/CD44 mediated
multiple signaling pathways†
Kai Jiang,‡ Ting Chi,‡ Tao Li, Guirong Zheng, Lulu Fan, Yajun Liu, Xiufen Chen,
Sijia Chen, Lee Jia and Jingwei Shao *

Ursolic acid (UA) has been recently used as a promising anti-tumor and cancer metastatic chemo-preven-
tive agent due to its low toxicity and liver-protecting property. However, the low bioavailability and non-
specific tumor targeting restrict its further clinical application. To address the problem, a silica-based
mesoporous nanosphere (MSN) controlled-release drug delivery system (denoted UA@M-CS-FA) was
designed and successfully synthesized, and was functionalized with folic acid (FA) and pH-sensitive chito-
san (CS) for the targeted delivery of UA to folate receptor (FR) positive tumor cells. UA@M-CS-FA were
spherical with mean diameter below 150 nm, and showed about −20 mV potential. Meanwhile,
UA@M-CS-FA exhibited a pH-sensitive release manner and high cellular uptake in FR over-expressing
HeLa cancer cells. Also, in vitro cellular assays suggested that UA@M-CS-FA inhibited cancer cell growth,
invasion and migration. Mechanistically, UA@M-CS-FA induced cancer cell apoptosis and inhibited
migration via cell cycle arrest in the G0/G1 stage, regulating the PARP/Bcl-2/MMP-9/CD44/PTEN/P53.
Importantly, in vivo experiments further confirmed that UA@M-CS-FA significantly suppressed the tumor
progression and lung metastasis in tumor-bearing nude mice. Immunohistochemical analysis revealed
Received 10th March 2017, that UA@M-CS-FA treatment regulated CD44, a biomarker of cancer metastasis. Overall, our data demon-
Accepted 4th June 2017
strated that a CS and FA modified MSN controlled-release drug delivery system could help broaden the
DOI: 10.1039/c7nr01677h usage of UA and reflect the great application potential of the UA as an anticancer or cancer metastatic
rsc.li/nanoscale chemopreventive agent.

Introduction therapy and metastasis. Various studies have shown that UA


exhibits growth inhibition properties against various human
Ursolic acid is a type of ursane-type pentacyclic triterpenic cancers cell lines.9–14 Our previous studies also provided
acid, ubiquitous in a variety of natural medicinal plants1 with evidence that UA and its derivatives showed potential anti-
multiple pharmacological properties, including anti-inflamma- cancer and especially cancer chemo-preventive activities.15–17
tory,2 liver-protective,3,4 anti-atherosclerotic,5 anti-epileptic, However, the poor water solubility of UA causes its low bio-
anti-cancer,6 anti-epileptic,7 and anti-diabetic activities.8 In availability, and ultimately restricts its potential clinical
recent years, it has garnered major attention in the field of applications.
cancer research due to its good anti-tumor activity and low tox- In recent decades, nano-technology based delivery systems
icity, which especially makes it a candidate for cancer chemo- have been making a significant impact on the development of
new strategies for cancer treatment. Various nano-materials, as
drug-carriers, have been widely used in nano-drugs for cancer
Cancer Metastasis Alert and Prevention Center, Fujian Provincial Key Laboratory of treatment, such as Fe3O4 18,19 and gold nanoparticles,20 nano-
Cancer Metastasis Chemoprevention and Chemotherapy, Fuzhou University, gels,21 liposomes,22 dendrimers,23 polymers,24 micelles,25
Fuzhou 350002, China. E-mail: shaojingwei@fzu.edu.cn, shaojw12@163.com; nanoemulsions,26 and silica nanoparticles.27,28 Among the
Tel: +86-13600802402
above nanocarriers, mesoporous silica nanoparticles (MSNs)
† Electronic supplementary information (ESI) available. See DOI: 10.1039/
c7nr01677h have a very high degree of attention in biomedical fields,
‡ These authors contributed equally to this work. owing to their small size, highly-ordered mesoporous struc-

9428 | Nanoscale, 2017, 9, 9428–9439 This journal is © The Royal Society of Chemistry 2017
View Article Online

Nanoscale Paper

ture, large specific surface area, large pore volumes, low tox- zeta potentials were measured by using a Malvern Zetasizer
icity, excellent biocompatibility, good mechanical stability and Nano-S90 (England). The surface areas and pore-size distri-
convenient chemical modification. Hence, a variety of pharma- butions of the MSNs were examined by Brunauer–Emmett–
ceutical drugs and fluorescent dyes have been loaded into Teller (BET) measurements. An ASAP 2020 Physisorption
MSNs for controlled drug release. Many studies have reported Analyzer (Micromeritics Instruments Corporation, USA) was
that some components were used as the gatekeeper of MSNs used to determine nitrogen-adsorption–desorption isotherms
for controlled drug delivery, for example, inorganic nano- of the M-C. A multipoint BET method was used to measure the
particles (Au, Fe3O4, CdS, and ZnO), dendrimers, antibodies, specific surface areas of the M-C and the pore-size distri-
rotaxanes, poly acrylic acid, pseudorotaxanes, and butions were calculated from desorption isotherms by the
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

β-cyclodextrin.28–30 In our previous work, we designed a silica- Barrett–Joyner–Halenda method. The chemical structures of
based mesoporous nanosphere (MSN) controlled-release the chitosan–folic acid conjugate and the nanoparticles (M-C,
prodrug delivery system (UA@MSN-UA), which could improve M-CS and M-CS-FA) were confirmed by Fourier transform
the bioavailability and extend the drug release of UA.31 infrared spectroscopy (FT/IR-480). The spectra were recorded
Chitosan is typically used in the field of pH-responsive drug from 4000 to 500 cm−1. In addition, CS-FA was also character-
delivery owing to its strong sensitivity in near-neutral or ized by 1H-NMR on a Bruker Avance-400 MHz NMR spectro-
slightly acidic pH environments.32,33 In order to further meter (Bruker BioSpin, MA, USA) using an acetic acid/deuter-
enhance drug molecule specific recognition by the tumor ium oxide solution (1/3 v/v) as a solvent. Thermogravimetric
tissues, nanocarriers generally require further modification analysis (TGA) of nanoparticles was performed using a thermo-
with targeting molecules, such as antibodies, peptides, nucleic gravimetric analyser (TGS-II, PerkinElmer).
acids, polysaccharides, or the small molecular compound folic
acid (FA).34 According to the literature we consulted, there is Drug loading into nanoparticles
no report that aimed at realizing targeted delivery of UA by the
Typically, UA (30 mg) and M-C (10 mg) were added to 30 mL
CS and FA functional modification of MSN.
methanol and stirred for 24 h at room temperature. Unbound
In the present study, a smart nanocarrier consisting of
UA, adsorbed on the surface, was removed by washing twice
mesoporous silica nanoparticles and folic acid conjugated
with a PBS solution, after centrifuging the UA-loaded M-C
chitosan (M-CS-FA) was designed and synthesized as a smart
from the suspension and removing the supernatant comple-
drug delivery system for the targeted delivery of UA to FR over-
tely. The resulting UA@M-C was freeze-dried. UV–vis spec-
expressing cancer cells for the first time. In our strategy, UA
troscopy was used to measure the UA concentration in the
was encapsulated into the MSN through non-covalent inter-
original solution and the supernatant solution at a wavelength
actions, and then the surface of MSN was covalently conju-
of 210 nm. The UA calibration curve is shown in Fig. S3.† The
gated with CS-FA through an acid-labile amide bond.
encapsulation and drug-loading efficiencies were calculated
Following the nanoparticle preparation, the physicochemical
according to the following equations:
characteristics were investigated, such as the particle size, zeta
potentials, morphology and in vitro drug release. Then, the Encapsulation ð%Þ
biological properties of UA@M-CS-FA nanoparticles were ¼ ðThe total amount of UA  the total amount of free UAÞ=
further investigated by cellular uptake, MTT assay, invasion the total amount of UA  100%:
assay, wound healing assay, cell cycle, cell apoptosis analysis
and western blot, respectively. In addition, its anti-tumor and Drug loading ð%Þ
anti-metastasis effects were further investigated in vivo.
¼ ðThe total amount of UA  the total amount of free UAÞ=
ðthe total amount of M‐C þ the total amount of UAÞ  100%:

Experimental section
Characterization of the nanoparticles In vitro drug release response
For the characterization of M-C and M-CS-FA, one drop of The in vitro release experiments were carried out under simu-
nanoparticles was dripped on a copper grid covered with a lated physiological conditions (Phosphate Buffered Saline,
nitrocellulose membrane, and transmission electron PBS) at different acidic environments ( pH 5.0, 6.5 and 7.4) at
microscopy (TEM; a JEOL JEM-1200EX microscope, Japan) was 37 °C to evaluate the feasibility of using M-CS-FA as a drug
used to observe the morphology and structure of M-C and delivery carrier. In brief, 5 mg of UA@M-CS-FA were suspended
M-CS-FA. An Atomic Force Microscope (AFM; Bruker in 5 mL of a freshly made phosphate buffer solution at various
MultiMode8, Germany) was used to observe the morphology of pH values (5.0, 6.5 and 7.4, respectively). The solution was put
M-C and M-CS-FA. Samples for AFM were prepared following a into a dialysis tube (MWCO: 12 kDa) which was directly
standard procedure. Briefly, 8 μL of a nanoparticle diluent was immersed in 50 mL incubation medium and kept at 37 °C,
dripped on pretreated new cleaved mica, and dried at room with gentle shaking (100 rpm). The release medium (2 ml) was
temperature for detection. M-C, M-CS and M-CS-FA were dis- withdrawn at predetermined time intervals (0.5, 1, 2, 4, 8, 12,
persed in deionized water and then the size distributions and 24, 48 and 72 h). Afterwards, the same volume (2 mL) of fresh

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 9428–9439 | 9429
View Article Online

Paper Nanoscale

release medium was added to the system. The amount of free three times with PBS ( pH = 7.4). The collected cells were cul-
UA in the supernatant was quantified by UV–vis spectroscopy tured on 96-cell plates in 100 μL medium and incubated for
at a wavelength of 210 nm as above. The release rate was calcu- another 24 h before being determined by MTT assay.
lated with the formula: Drug release = (Wi/Wtotal ) × 100%,
where Wi is the amount of accumulative released UA, and Western blot
Wtotal is the total amount of UA. HeLa cells were incubated for 24 h with 20 μg mL−1 UA,
UA@M-C, UA@M-CS, and UA@M-CS-FA for exploring the
Cell uptake
expression of P53, Bcl-2 and for 24 h with 1 μg mL−1 UA,
HeLa and HepG2 cells were placed into a 24-well plate (3 × 104 UA@M-C, UA@M-CS, UA@M-CS-FA for exploring the
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

cells per well) and incubated at 37 °C with 5% CO2 for 24 h. expression of MMP-2, and CD44 protein. Cells were washed
After being treated with FITC labeled M-CS-FA (0.1 mg ml−1) with cold PBS and lysed with RIPA lysis buffer for ten
and M-CS (0.1 mg ml−1) for 2 h, the cells were rinsed twice minutes at an ice-cold bath, and then the supernatants of cell
with PBS buffer (2 × 100 μL) to remove the remaining nano- lysates were recovered by centrifugation at 12 000g for 5 min
particles and dead cells. After treatment with 200 μL of a paraf- under 4 °C. The concentrations of proteins were measured by
ormaldehyde (PA) solution for 30 min at 4 °C, their nuclei the BCA method. Sodium dodecyl sulfate polyacrylamide
were stained with DAPI for 10 min. Finally, the cells were gel electrophoresis (SDS-PAGE) was performed to separate
observed under a confocal microscope (Zeiss LSM 780, Zeiss the samples, as described previously.35 The target protein
Co., Germany). expression was quantified by the use of the Image Lab analysis
To confirm whether the uptake was mediated by FR, a fluo- software (Bio-Rad).
rescence-activated cell sorter (FACS) was used. Briefly, HeLa
cells were seeded into 12-well plates at a density of 5 × 104 cells Statistical analysis
per well. After 24 h, cells were incubated with various nano- All experiments were performed in triplicate and the acquired
particles (FITC labeled M-CS-FA and FITC labeled M-CS) data are presented as the mean ± SD. Statistical significance
in RPMI 1640 medium. In the competition experiments, was determined using a Student’s t-test. The differences were
2 mg mL−1 FA was applied and incubated with the HeLa cells considered statistically significant if *p < 0.05 and very signifi-
before the nanoparticles were added. After 2 h incubation, the cant if **p < 0.01.
cells were washed three times with PBS, trypsinized and resus-
pended in 500 μL PBS. The mean fluorescence intensity (MFI)
was measured by FACS (BD FACSAria III, USA). Results and discussion
Cell viability assay Preparation and characterization of UA@M-CS-FA
In vitro cytotoxicities of the blank nanocarrier, UA and UA The procedure for the synthesis and effects of UA@M-CS-FA is
loaded nanoparticles against HeLa and HepG2 cell lines were shown in Fig. 1. Firstly, MSN with the functionalization of car-
assessed by the standard MTT assay. Briefly, cells were seeded boxyl (M-C) was synthesized by co-condensation. Then the
into 96-well plates at an initial density of 80 000 cells per mL surface of M-C was conjugated with chitosan-folic acid (CS-FA)
for 0.1 mL (8 × 103 cells per well) and cultured overnight. Cells by the amide bond.
were treated with various concentrations of nanodrugs in From the TEM analysis (Fig. 2A), it could be seen that
RPMI 1640 with 10% FBS. After 24 h of incubation, the M-C NPs were uniform spherical nanoparticles with an average
medium was removed, and a medium containing 0.5 mg mL−1 diameter of about 100 nm. In addition, the mean particle size of
MTT reagent was added for continuous incubation for 4 h. M-C was measured by using dynamic light scattering (DLS). As
Then, after removing the medium, 150 μL of dimethyl sulfox- revealed in Fig. 2B, it presented a relatively larger diameter
ide (DMSO) was added per well. The absorbance at 570 nm (120 nm) due to the hydrated layer surrounding the particles
was measured by using a Microquant plate reader (Thermo with a narrow particle size distribution. The consecutive modifi-
Fisher). Meanwhile, we explored the effect of the blank nano- cation processes were measured by zeta potential measure-
carrier, UA and UA loaded nanoparticles against HeLa and ments. As summarized in Fig. 2C, the zeta potential of M-C,
HepG2 cell lines in blood. Blood samples were drawn from cer- M-CS and M-CS-FA was −29.3 mV, +17.5 mV, and −25 mV,
vical cancer patients when they were under operation. The respectively, and it was consistent with the modification of chit-
study was reviewed and approved by the hospital institutional osan (M-CS) on the negatively charged silica surface (M-C) and
review board (IRB). Blood was collected into Vacutainer tubes the subsequent conjugation of FA on the positively charged
containing the anticoagulant EDTA. In blood, HeLa and M-CS. In addition, Fig. S1† shows the z-average diameter, zeta-
HepG2 cells were harvested with trypsin and resuspended in potential and AFM images of different nanoparticles. As shown
PBS. After washing with PBS, the cells were resuspended in in Fig. 2D, the results of the nitrogen adsorption–desorption
50 μL PBS at a concentration of 8000 cells per μL, and 0.5 mL isotherms showed that the gas adsorption isotherm is in a typical
of the diluted blood specimens was mixed with the cells. Then type IV isotherm loop, which indicated that M-C was character-
the above samples were incubated with MSN-CS-FA, UA, and istic of a porous structure. The surface areas and pore sizes of
UA loaded nanodrugs for 4 h at 37 °C, followed by washing the M-C were evaluated by BET and BJH analyses, and the

9430 | Nanoscale, 2017, 9, 9428–9439 This journal is © The Royal Society of Chemistry 2017
View Article Online

Nanoscale Paper
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

Fig. 1 Schematic illustration of UA@M-CS-FA for targeted delivery of UA guided pH-triggered chemotherapy. (A) Schematic diagram of the prepa-
ration of UA@M-CS-FA nanoparticles. (B) UA@M-CS-FA exerted anti-metastatic effect via the inhibition of cell adhesion, invasion and migration. (C)
The tumor targeting was achieved by specific interaction between FA and tumor cells, and the pH sensitive manner was triggered by the swelling of
chitosan chains ( pH < 5.5 in the endosome).

specific surface area of the M-C was 887.2 m2 g−1. Fig. 2E are the characteristic absorption peaks of the carbon skeleton
revealed that M-C had a narrow pore size distribution centered of the aromatic ring. In the pure FA the absorption bands were
at 2.8 nm and the pore volume was 1.32 cm−3 g−1. Moreover, observed at 1600 cm−1 and 1480 cm−1; meanwhile, in the FA-CS
Fig. 2F depicts thermogravimetric analysis (TGA) curves of the absorption bands were observed at 1590 cm−1 and
M-C, M-CS and M-CS-FA under constant N2 flow conditions. 1560 cm−1. Moreover, FA and FA-CS all included the –CONH–
Upon heating to 800 °C, the total weight loss of the materials group. The carboxylic absorption of CvO was at around
was obtained: 37.27% (H3) for M-C, 40.99% (H2) for M-CS, and 1740 cm−1, while its position could shift to lower frequencies
52.98% (H1) for M-CS-FA, respectively. On heating to 100 °C, due to the conjugative effect when amidated. Meanwhile, it was
M-C, M-CS and M-CS-FA showed a weight loss of 31.83% (L3), found that there was an absorption band at 1700 cm−1 in FA-CS
20.59% (L2) and 15.59% (L1), respectively, which was the conjugates, which demonstrated the CvO group of FA and the
volatilization of residual moisture from the pores of the nano- –NH2 group of CS conjugated into the amide. In addition, it
particles. The weight loss between 100 and 800 was attributed was observed that the absorption position of N–H shifted to
to the decomposition of the functionalized groups, so the higher frequencies in CS due to an inductive effect aroused by
grafting rate of CS and CS-FA on M-C was calculated according the CvO group. The absorption position of N–H in CS was
to the formula: (H3 − L3)/(100 − H3) = (H2 − L2 − W2)/ 1600 cm−1, and in CS-FA conjugates, it shifted to 1630 cm−1.
(100 − H2); (H3 − L3)/(100 − H3) = (H1 − L1 − W1)/(100 − H1). CS-FA was also characterized by 1H NMR spectroscopy. From
Thus, the grafting rate of CS and CS-FA on M-C was 15.34% the results in Fig. S2,† it could be seen that the signals were at
and 33.31%, respectively. δ 1.57, 2.63 and 3.23–3.41 ppm, which were attributed to the res-
Chitosan–folic acid conjugate (CS-FA) was prepared by onance of the monosaccharide residue protons, –COCH3, –CH–
carbodiimide (EDC) and N-hydroxysuccinimide (NHS) mediated NH–, and –CH2–O–, respectively. The signals at δ 6.27–8.30 ppm
polymerization. Fig. 2G shows the results of the FTIR spectra of in the 1H NMR spectra of CS-FA were attributed to the reson-
FA, CS, and CS-FA. CS-FA was investigated through IR spectra. It ance of the folate aromatic protons.36,37 The above results indi-
is known that there are several peaks at 1650–1450 cm−1 which cated that CS-FA was synthesized successfully. M-CS-FA was syn-

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 9428–9439 | 9431
View Article Online

Paper Nanoscale
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

Fig. 2 Characterization of the nanoparticles and conjugation, and in vitro release curves of UA. (A) TEM image of M-C. (B) Size distribution of M-C.
(C) Average zeta potential of different nanoparticles. (D) Nitrogen adsorption–desorption isotherms of M-C. (E) Pore size distribution of M-C. (F) TGA
curves of different nanoparticles. (G) Fourier-transform IR spectra of FA, CS, and CS-FA. (H) Fourier-transform IR spectra of M-C, M-CS and
M-CS-FA. (I) Cumulative UA (%) from the M-CS-FA drug carrier at 37 °C in a phosphate buffer solution ( pH = 5.0, 6.5 and 7.4).

thesized from the reaction between M-C and CS-FA using EDC/ drug loading and encapsulation efficiency, whose efficiencies
NHS between the amino group of CS-FA and the carboxyl group were 21.8% and 56.7%, respectively.
of M-C. The FTIR spectra of M-C, M-CS, and M-CS-FA are pre- The release properties of UA from M-CS-FA nanoparticles
sented in Fig. 2H. There was a visible characteristic absorption were measured under different pH conditions ( pH 7.4, pH 5.0
peak of CvO of the –COOH group in M-C at 1715 cm−1. The and 6.5) at 37 °C. As shown in Fig. 2I, it is clearly seen that
absorption position of CvO of the –COOH group could shift to different pH media had a strong influence on the release of UA
lower frequencies when there appear a conjugative effect, hence from M-CS-FA nanoparticles. At neutral pH 7.4, UA-loaded
it would show a lower frequency absorption peak when the M-CS-FA exhibited a very slow release rate and only 20% of the
–COOH group and –NH2 conjugated. Interestingly, there was a UA molecules were released after 72 h. Relatively,
characteristic absorption peak at 1643 cm−1 in M-CS, which UA@M-CS-FA showed a significantly enhanced release rate at
showed that the –COOH group in M-C and –NH2 in CS ami- acidic pH 5.0 and almost 75% of the accumulative release of
dated. Meanwhile, it was observed that there was an absorption UA was found for UA@M-CS-FA at acidic pH 5.0.
peak at 1700 cm−1 in M-CS-FA caused by the –COOH group con- The reasons for the rapid UA release of UA@M-CS-FA under
tained in FA. All data displayed MSN, CS and FA conjugated. pH 5.0 are as follows. On the one hand, CS-FA was conjugated
on the surface of MSN through an acid-labile amide bond.31
On the other hand, chitosan, as a smart response molecule, is
In vitro drug loading and release behavior responsive to external pH-stimuli.38,39 Acidic pH 5.0 makes
Ursolic acid, as a model anticancer drug, was loaded into these chitosan chains swell and the mesopores of the
M-CS-FA. UV-Vis spectroscopy was performed to measure the UA@M-CS-FA open, so that UA could diffuse out of the chan-

9432 | Nanoscale, 2017, 9, 9428–9439 This journal is © The Royal Society of Chemistry 2017
View Article Online

Nanoscale Paper

nels of the UA@M-CS-FA. While the pH value of the drug- by HepG2 cells than by HeLa cells, resulting in far smaller
loaded CS-MSN medium was adjusted to 7.4, the shielding extent of receptor-mediated endocytosis. As previously
layers on the mesoporous surface of the MSNs were formed reported, FR-positive cells have a much higher FA-functional
due to the deprotonation and collapse of these chitosan nanoparticle uptake than FR-negative cells.43
polymer chains, which could effectively block the loaded UA To explore whether the entry of M-CS-FA into the cells is via
from escaping from the channels of the UA@M-CS-FA. In view receptor-mediated endocytosis, competition experiments were
of the fact that the tumor microenvironment is weakly performed by FACS. Before being incubated with FITC labeled
acidic,40–42 UA@M-CS-FA exhibited a fast release in pH 5.0, M-CS-FA, HeLa cells were pretreated with free FA. The results
which may be conducive to tumor therapy. Thus, suggested that the MFI of the pretreated group decreased to
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

UA@M-CS-FA may be a better platform for the delivery of UA one third of the untreated group (In Fig. S4†), similar to the
triggered by pH. MFI of M-CS, which showed that the cellular uptake of
M-CS-FA was markedly inhibited. All these results revealed that
Cell targeting ability of FITC labeled M-CS-FA FA ligands on UA@MSN particles play an important role in
To verify the ability of M-CS-FA nanoparticles for FA-mediated increasing cellular uptake via FA-receptor-mediated pathway.
targeted delivery of UA, FITC labeled M-CS-FA nanoparticles
were prepared. The green fluorescence emitted by FITC and
the blue fluorescence emitted by DAPI were captured by Effects of UA@M-CS-FA on viability
LCFM. In Fig. 3, the images are shown from top to bottom as The cytotoxicities of nanoparticles with and without loading of
follows: DAPI (blue), FITC labeled M-CS-FA (green) and the UA in culture medium and in blood were evaluated by the
overlay of the two channels. After 2 h of incubation with MTT assay with the FR-positive HeLa cells and FR-negative
samples, only the nuclei are observed in blue color in control HepG2 cells. As depicted in Fig. 4A and B, free UA and nano-
groups untreated with FITC labeled M-CS-FA. Obviously, in the particles with loading of UA showed cytotoxicity towards both
overlay images, the green fluorescence in the cytoplasm in cell lines in a concentration-dependent manner. It was worth
HeLa cells, produced by FITC labeled M-CS-FA, was greater noting that, in HepG2 cells, the cytotoxicity of free UA was
and more closely situated around the nuclei (blue) than that in slightly lower than that of UA loaded nanoparticles, and
the HepG2 resulting from FITC labeled M-CS-FA. The main UA@M-CS-FA has a similar cytotoxicity to UA@M-C, UA@M-CS
reason for this was that far fewer folate receptors are expressed and UA@M-FA. However, UA@M-CS-FA showed higher cyto-

Fig. 3 Confocal fluorescence images of HeLa and HepG2 cells after incubation with free medium and FITC labeled M-CS-FA (0.2 mg mL−1).

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 9428–9439 | 9433
View Article Online

Paper Nanoscale
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

Fig. 4 Cell viability of FR-overexpressing HeLa (A and C) and FR-negative HepG2 (B and D) incubated with nanoparticles and UA loaded
nanoparticles.

toxicity than UA@M-C, UA@M-CS, UA@M-FA and free UA in induced different cell apoptosis through different mecha-
all the doses tested in HeLa cells. nisms.10,45,46 In the present study, the anti-metastatic effect of
Fig. 4C and D showed that M-CS-FA and UA loaded nano- UA@M-CS-FA was confirmed by cell adhesion, invasion and
particles could downregulate the cell viability in blood, while wound healing assays. Fig. 5A showed representative experi-
UA did not show significant effect of cytotoxicity, compared to mental images of HeLa cells. From the data of the adhesion
HeLa cells and HepG2 cells cultured in the medium. assay in Fig. 5B, it was clearly seen that there was no signifi-
UA@M-CS-FA had a lower effect on the viability of HepG2 cant difference in HepG2 cells between groups, while the
cells, which should be attributed to the very low expression of adhesion rate of HeLa cells treated with UA@M-CS-FA to
folate receptors on HepG2 cells. In Fig. 4, it was clearly seen HUVECs was lower than that of other groups. Invasion assays
that the cytotoxicity of UA@M-CS was similar to that of showed that UA, UA@M-C, UA@M-CS and UA@M-CS-FA had
UA@M-FA. In addition, M-CS-FA had little influence on the different inhibition of invasion on HeLa cells, compared to the
viability of HeLa and HepG2 cells under 1 μg mL−1–200 control group. It was observed that UA@M-CS-FA nanoparticles
μg mL−1 concentration. This result again confirmed the safety significantly inhibit invasion. A wound healing assay was con-
of M-CS-FA. These results indicated that M-CS-FA particles can ducted to explore the effect of UA@M-CS-FA nanoparticles on
effectively perform targeted delivery of UA to cancer cells and HeLa cell mobility. The wound of the control group gradually
enhance the anticancer activity. healed after 24 hours, while a significant wound remained
uncovered in the UA@M-CS-FA treated group, which showed
Anti-metastatic effect of UA@M-CS-FA that the mobility of HeLa cells was effectively inhibited by
Cancer cell metastasis is a very complex dynamic continuous UA@M-CS-FA nanoparticles, compared to cells treated with UA
biological process consisting of multiple relatively indepen- and other nanodrugs. However, it is notable that the wound of
dent steps. Cell adhesion, invasion and migration are the key the group treated with UA@M-CS-FA was similar to that of the
points of metastasis.44 Many studies have shown that UA group treated with UA@M-CS in HepG2 cells. It is also seen

9434 | Nanoscale, 2017, 9, 9428–9439 This journal is © The Royal Society of Chemistry 2017
View Article Online

Nanoscale Paper
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

Fig. 5 Effect of UA@M-CS-FA on the mobility of HeLa and HepG2 cells. (A) HeLa cell representative images of adhesion, invasion, and scratch
assays. (B) Quantitative analysis of the inhibition of the adhesion, invasion, migration of HeLa and HepG2 cells.

that FA modification of nanodrugs improves the therapeutic that FA-modified nanoparticles had the potential for targeted
effect of UA. The above results showed that treatment of delivery of UA into FR-positive cancer cells.47
UA@M-CS-FA at low concentrations inhibited HeLa cell
migration. Analysis of apoptosis by DAPI staining
To verify whether the growth inhibitory activity of UA loaded
Effects of UA@M-CS-FA on the cell cycle nanoparticles was related to the induction of apoptosis,
To further study the cell death mechanisms of nano-drugs, changes in the nuclear morphological character of HeLa
a cell cycle analysis was performed by flow cytometry. The data cells and HepG2 cells were assayed by DAPI staining. As shown
in Fig. 6 indicated that free UA and nano-drugs (UA@M-C, in Fig. S5,† there was no sign of apoptosis in the control
UA@M-CS, UA@M-CS-FA) blocked cell cycle progression in the group. Cell death features, such as cell shrinking and
G0/G1 phase and the G2/M phase in HeLa and HepG2 cells detachment, were clearly detectable in HeLa cells treated
under 20 μg mL−1 concentration. UA, UA@M-C and UA@M-CS with free UA and nano-drugs, while the nuclear changes were
showed no obvious difference between HeLa and HepG2 cells, more obvious in UA@M-CS-FA treated HeLa cells, resulting
consistent with the MTT results. While UA@M-CS-FA showed in insignificant nuclear condensation and cell shrinking
obvious effect on the cell cycle in HeLa cells, the cell could be accompanied by a reduction in the number of nuclei.
significantly arrested in the G0/G1 and G2/M phases in the However, in comparison with free UA, UA@M-C, UA@M-CS
group, compared to the other drugs (UA, UA@M-C and and UA@M-CS-FA, the nuclei had no obvious changes in
UA@M-CS). However, in HepG2 cells, the cell cycle distri- UA@M-CS-FA treated HepG2 cells. The results indicated that
bution showed no significant difference between groups. the capability of UA@M-CS-FA to induce FR over-expressing
These results further indicated that CS-modified nanoparticles HeLa cancer cell apoptosis was significantly stronger than the
could achieve less release before reaching the tumor site. effect of UA, UA@M-C and UA@M-CS. The above results were
Meanwhile, these results were in accordance with the result consistent with cellular uptake studies and corroborated

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 9428–9439 | 9435
View Article Online

Paper Nanoscale
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

Fig. 6 Cell cycle analysis of HeLa and HepG2 cells after being treated with UA, UA@M-C, UA@M-CS, and UA@M-CS-FA.

Fig. 7 Western blot analysis of PARP, PTEN, P53 and Bcl-2 expressions after HeLa cells were treated with free UA, UA loaded nanoparticles at high con-
centrations for 24 h and MMP-9, CD44 expression after HeLa cells were treated with free UA, UA loaded nanoparticles at low concentrations for 24 h.

9436 | Nanoscale, 2017, 9, 9428–9439 This journal is © The Royal Society of Chemistry 2017
View Article Online

Nanoscale Paper

the MTT assay, further demonstrating the role played by expression was observed. Notably, UA@M-CS-FA could signifi-
FA-mediated targeting. cantly down-regulate the levels of PARP and Bcl-2 compared
with UA and the other drugs, while the expression of PTEN
Effects of UA@M-CS-FA on apoptosis and metastasis related and P53 was up-regulated by the free UA and UA loaded nano-
protein expression particles. The effects of UA@M-CS on the levels of PTEN and
To verify the molecular mechanism of anti-tumor and anti- P53 were equivalent to UA, but better than UA@M-C.
migration induced by UA@M-CS-FA, we assessed the four Meanwhile, UA@M-CS-FA could significantly induce the levels
apoptosis-related protein and two migration-related protein of PTEN and P53 compared with free UA and other UA loaded
expressions in HeLa cells by western blot, including PARP, nanoparticles, whereas treatment with UA@M-CS-FA at low
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

PTEN, P53, Bcl-2 and MMP-9, CD44. As shown in Fig. 7, when concentrations may influence the expression of migration-
HeLa cells were treated at high concentrations of UA and UA related proteins in HeLa cells, which was consistent with our
loaded nanoparticles, a decrease in PARP and Bcl-2 protein previous study.48 Similarly, Huang et al. also found that UA

Fig. 8 In vivo experiments. (A–B) Changes of tumor volume and the body weight change curve of nude mice treated with PBS, UA, UA@M-CS, and
UA@M-CS-FA. (C) Histopathological analysis of the heart, liver, spleen and lung stained with hematoxylin and eosin after treatment with PBS, UA,
UA@M-CS, and UA@M-CS-FA. (D–E) Optical images of lungs and quantitative analysis of pulmonary metastatic nodules separated from mice receiv-
ing PBS, UA, UA@M-CS, and UA@M-CS-FA. Circles indicate the lung metastases. **p < 0.01 versus the UA@M-CS group. Immunohistochemistry ana-
lysis of CD-44 expressions of lungs after treatment with PBS, UA, UA@M-CS, and UA@M-CS-FA.

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 9428–9439 | 9437
View Article Online

Paper Nanoscale

inhibits metastasis through the down-regulation of MMP-9 in Conclusion


a dose-dependent manner.49
In this study, UA@M-CS-FA has been successfully developed
for targeted delivery of UA to tumor cells and guided by pH-
In vivo anticancer activity of UA@M-CS-FA triggered drug release. The CS-FA on the surface of MSNs
could swell triggered by pH and especially identify FR over-
The encouraging in vitro results of UA@M-CS-FA nanoparticles
expressing cancer cells. The nanoparticles had the attractive
on HeLa cell mediated FR warranted further in vivo assays to
feature of a high drug-loading capacity. Furthermore, M-CS-FA
evaluate its antitumor. The ability of UA@M-CS-FA to inhibit
was biocompatible and UA@M-CS-FA exhibited more signifi-
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

the growth of the primary tumor was evaluated by using a


cant cell proliferation inhibitory action on HeLa cells com-
HeLa bearing tumor model. As shown in Fig. 8A, PBS or UA
pared with UA@M-C, UA@M-CS and free UA alone.
showed negligible influence on the tumor growth, while the
Meanwhile, cell adhesion, invasion and scratch assays indi-
UA@M-CS-FA group alone significantly inhibited the HeLa
cated that UA@M-CS-FA at non-cytotoxic concentrations sig-
tumor growth. Notably, during the administration, there was
nificantly inhibited HeLa cells adhesion, invasion and immi-
no significant fluctuation in the body weights of mice among
gration. A western blot assay indicated that UA@M-CS-FA
the groups (in Fig. 8B). In addition, the data in Fig. 8C indi-
induced cancer cell apoptosis and inhibited migration via reg-
cated that there were no significant differences in the histo-
ulating the P53/MMP-9/PTEN/CD44 protein expression. What’s
logical findings between the control and various treatment
more, in vivo experiment further confirmed that FA mediated
groups, indicating that M-CS-FA showed good biocompatibility
delivery of UA exhibited effective anti-tumor and anti-meta-
and UA@M-CS-FA treatment was safe in mice at therapeutic
stasis activities. These findings indicated that M-CS-FA, as a
doses, which was essential for their further clinical translation.
carrier, can efficiently delivery UA to folate receptor positive
The cell apoptosis of tumor sections was evaluated through
cancer cells to improve tumor therapy of UA.
TUNEL staining to investigate the anti-tumor activity from
the cellular level (Fig. 8C). The PBS group had almost no cell
apoptosis with nuclei, while the UA and UA@M-CS groups
exhibited slight apoptosis with a very small part of the nuclei
Conflict of interest
stained brown by TUNEL. For the UA@M-CS-FA group, most The authors declare that they have no conflict of interest.
nuclei of tumors showed apoptosis. The selective uptake in
tumor cells of UA is a very important plus factor for improving
the anti-tumor efficacy. Given the above results, UA@M-CS-FA Acknowledgements
has the potential to be an antitumor drug for further clinical
studies. This project was supported by the National Science
Foundation of China (no. 81472767, 81673698, and 81201709).

In vivo anti-metastasis activity of UA@M-CS-FA


References
Finally, we evaluated the in vivo anti-metastasis activity of
UA@M-CS-FA on lung metastatic nodules and the expression 1 M. K. Shanmugam, X. Dai, A. P. Kumar, B. K. Tan, G. Sethi
of CD44 in a mouse model of established pulmonary meta- and A. Bishayee, Biochem. Pharmacol., 2013, 85, 1579–1587.
stasis. It is worth noting that the images of the lung tissues in 2 D. Chattopadhyay, G. Arunachalam, A. B. Mandal,
Fig. 8D and E exhibit that treatment with PBS displayed severe T. K. Sur, S. C. Mandal and S. K. Bhattacharya,
lung metastasis, while UA showed a little anti-metastasis. J. Ethnopharmacol., 2002, 82, 229–237.
However, UA@M-CS-FA significantly suppressed the pulmon- 3 S. Prasad, V. R. Yadav, R. Kannappan and B. B. Aggarwal,
ary metastasis. Furthermore, we checked the expression of J. Biol. Chem., 2011, 286, 5546–5557.
CD44, and IHC staining findings indicated that tumors treated 4 M. K. Shanmugam, P. Rajendran, F. Li, T. Nema, S. Vali,
with UA@M-CS-FA significantly inhibited the expression of T. Abbasi, S. Kapoor, A. Sharma, A. P. Kumar, P. C. Ho,
CD44 in HeLa tumor cells, superior to other groups, which K. M. Hui and G. Sethi, J. Mol. Med., 2011, 89, 713–727.
was consistent with the western-blot results. Surprisingly, the 5 K. A. Kim, J. S. Lee, H. J. Park, J. W. Kim, C. J. Kim,
effects of UA@M-CS in vivo were better than those in vitro; it I. S. Shim, N. J. Kim, S. M. Han and S. Lim, Life Sci., 2004,
may be delivered to the tumor tissue based on the enhanced 74, 2769–2779.
permeability and retention (EPR) effect. So, it could inhibit 6 L. R. Tannock, Atherosclerosis, 2011, 219, 397–398.
the growth and metastasis of tumors to some extent. All 7 I. Kazmi, M. Afzal, G. Gupta and F. Anwar, CNS Neurosci.
these results suggest that the high loading efficiency of Ther., 2012, 18, 799–800.
UA in M-CS-FA and FA mediated UA delivery enhanced 8 S. M. Jang, M. J. Kim, M. S. Choi, E. Y. Kwon and M. K. Lee,
anti-metastasis activities. Hence, these nano-drugs are found Metab., Clin. Exp., 2010, 59, 512–519.
to be safe and effective therapeutic modalities for tumor 9 D. Andersson, J. J. Liu, A. Nilsson and R. D. Duan,
treatment. Anticancer Res., 2003, 23, 3317–3322.

9438 | Nanoscale, 2017, 9, 9428–9439 This journal is © The Royal Society of Chemistry 2017
View Article Online

Nanoscale Paper

10 D. K. Kim, J. H. Baek, C. M. Kang, M. A. Yoo, J. W. Sung, 29 Y. Wang, S. Song, J. Liu, D. Liu and H. Zhang, Angew.
H. Y. Chung, N. D. Kim, Y. H. Choi, S. H. Lee and Chem., 2015, 54, 536–540.
K. W. Kim, Int. J. Cancer, 2000, 87, 629–636. 30 Q. Gan, X. Lu, Y. Yuan, J. Qian, H. Zhou, X. Lu, J. Shi and
11 J. H. Baek, Y. S. Lee, C. M. Kang, J. A. Kim, K. S. Kwon, C. Liu, Biomaterials, 2011, 32, 1932–1942.
H. C. Son and K. W. Kim, Int. J. Cancer, 1997, 73, 725–728. 31 T. Li, X. Chen, Y. Liu, L. Fan, L. Lin, Y. Xu, S. Chen and
12 Y. H. Choi, J. H. Baek, M. A. Yoo, H. Y. Chung, N. D. Kim J. Shao, Eur. J. Pharm. Sci., 2017, 96, 456–463.
and K. W. Kim, Int. J. Oncol., 2000, 17, 565–571. 32 S. A. Agnihotri, N. N. Mallikarjuna and T. M. Aminabhavi,
13 M. T. Huang, C. T. Ho, Z. Y. Wang, T. Ferraro, Y. R. Lou, J. Controlled Release, 2004, 100, 5–28.
K. Stauber, W. Ma, C. Georgiadis, J. D. Laskin and 33 R. Vivek, V. N. Babu, R. Thangam, K. S. Subramanian and
Published on 07 June 2017. Downloaded by National Tsing Hua University on 08/02/2018 18:22:44.

A. H. Conney, Cancer Res., 1994, 54, 701–708. S. Kannan, Colloids Surf., B, 2013, 111, 117–123.
14 D. Es-Saady, A. Simon, C. Jayat-Vignoles, A. J. Chulia and 34 S. D. Weitman, R. H. Lark, L. R. Coney, D. W. Fort,
C. Delage, Anticancer Res., 1996, 16, 481–486. V. Frasca, V. R. Zurawski Jr. and B. A. Kamen, Cancer Res.,
15 L. Xiang, T. Chi, Q. Tang, X. Yang, M. Ou, X. Chen, X. Yu, 1992, 52, 3396–3401.
J. Chen, R. J. Ho, J. Shao and L. Jia, Oncotarget, 2015, 6, 35 S. C. Hsu, C. L. Kuo, J. P. Lin, J. H. Lee, C. C. Lin, C. C. Su,
9295–9312. M. D. Yang and J. G. Chung, Anticancer Res., 2007, 27,
16 X. Yang, Y. Li, W. Jiang, M. Ou, Y. Chen, Y. Xu, Q. Wu, 2377–2384.
Q. Zheng, F. Wu, L. Wang, W. Zou, Y. J. Zhang and J. Shao, 36 S. J. Yang, F. H. Lin, K. C. Tsai, M. F. Wei, H. M. Tsai,
Chem. Biol. Drug Des., 2015, 86, 1397–1404. J. M. Wong and M. J. Shieh, Bioconjugate Chem., 2010, 21,
17 H. Dong, X. Yang, J. Xie, L. Xiang, Y. Li, M. Ou, T. Chi, 679–689.
Z. Liu, S. Yu, Y. Gao, J. Chen, J. Shao and L. Jia, Biochem. 37 H. M. Li, Z. Li, J. Zhao, B. Q. Tang, Y. H. Chen, Y. K. Hu,
Pharmacol., 2015, 93, 151–162. Z. D. He and Y. Wang, Nanoscale Res. Lett., 2014, 9, 146.
18 Y. X. Wang, S. M. Hussain and G. P. Krestin, Eur. J. Radiol., 38 W. Feng, W. Nie, C. He, X. Zhou, L. Chen, K. Qiu, W. Wang
2001, 11, 2319–2331. and Z. Yin, ACS Appl. Mater. Interfaces, 2014, 6, 8447–8460.
19 J. H. Lee, J. T. Jang, J. S. Choi, S. H. Moon, S. H. Noh, 39 C. Y. Chang, M. He, J. P. Zhou and L. N. Zhang,
J. W. Kim, J. G. Kim, I. S. Kim, K. I. Park and J. Cheon, Nat. Macromolecules, 2011, 44, 1642–1648.
Nanotechnol., 2011, 6, 418–422. 40 J. Liu, Y. Huang, A. Kumar, A. Tan, S. Jin, A. Mozhi and
20 I. H. El-Sayed, X. Huang and M. A. El-Sayed, Cancer Lett., X. J. Liang, Biotechnol. Adv., 2014, 32, 693–710.
2006, 239, 129–135. 41 K. Engin, D. B. Leeper, J. R. Cater, A. J. Thistlethwaite,
21 N. V. Nukolova, H. S. Oberoi, S. M. Cohen, A. V. Kabanov L. Tupchong and J. D. McFarlane, Int. J. Hyperthermia,
and T. K. Bronich, Biomaterials, 2011, 32, 5417–5426. 1995, 11, 211–216.
22 P. P. Deshpande, S. Biswas and V. P. Torchilin, 42 W. Gao, J. M. Chan and O. C. Farokhzad, Mol. Pharm.,
Nanomedicine, 2013, 8, 1509–1528. 2010, 7, 1913–1920.
23 D. J. Bharali, M. Khalil, M. Gurbuz, T. M. Simone and 43 Y. Wu, R. Guo, S. Wen, M. Shen, M. Zhu, J. Wang and
S. A. Mousa, Int. J. Nanomed., 2009, 4, 1–7. X. Shi, J. Mater. Chem. B, 2014, 2, 7410–7418.
24 T. Yoshida, T. C. Lai, G. S. Kwon and K. Sako, Expert Opin. 44 P. Friedl and K. Wolf, Nat. Rev. Cancer, 2003, 3, 362–374.
Drug Delivery, 2013, 10, 1497–1513. 45 Y. Q. Meng, D. Liu, L. L. Cai, H. Chen, B. Cao and
25 Y. Li, K. Xiao, W. Zhu, W. Deng and K. S. Lam, Adv. Drug Y. Z. Wang, Bioorg. Med. Chem., 2009, 17, 848–854.
Delivery Rev., 2014, 66, 58–73. 46 L. Yang, X. Liu, Z. Lu, J. Yuet-Wa Chan, L. Zhou, K. P. Fung,
26 S. Ganta, M. Talekar, A. Singh, T. P. Coleman and P. Wu and S. Wu, Cancer Lett., 2010, 298, 128–138.
M. M. Amiji, AAPS PharmSciTech, 2014, 15, 694–708. 47 Y. Gao, Z. Li, X. Xie, C. Wang, J. You, F. Mo, B. Jin, J. Chen,
27 W. Fan, B. Shen, W. Bu, F. Chen, Q. He, K. Zhao, S. Zhang, J. Shao, H. Chen and L. Jia, Eur. J. Pharm. Sci., 2015, 70,
L. Zhou, W. Peng, Q. Xiao, D. Ni, J. Liu and J. Shi, 55–63.
Biomaterials, 2014, 35, 8992–9002. 48 Q. Tang, Y. Liu, T. Li, X. Yang, G. Zheng, H. Chen, L. Jia
28 C. Y. Lai, B. G. Trewyn, D. M. Jeftinija, K. Jeftinija, S. Xu, and J. Shao, Oncotarget, 2016, 7, 73114–73129.
S. Jeftinija and V. S. Lin, J. Am. Chem. Soc., 2003, 125, 49 H. C. Huang, C. Y. Huang, S. Y. Lin-Shiau and J. K. Lin,
4451–4459. Mol. Carcinog., 2009, 48, 517–531.

This journal is © The Royal Society of Chemistry 2017 Nanoscale, 2017, 9, 9428–9439 | 9439

Anda mungkin juga menyukai