Sunteți pe pagina 1din 7

Colloids and Surfaces B: Biointerfaces 153 (2017) 104–110

Contents lists available at ScienceDirect

Colloids and Surfaces B: Biointerfaces


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

Bone-specific poly(ethylene sodium phosphate)-bearing


biodegradable nanoparticles
Yuya Hirano a , Yasuhiko Iwasaki b,c,∗
a
Graduate School of Science and Engineering, Kansai University, 3-3-35 Yamate-cho, Suita-shi, Osaka, 564-8680, Japan
b
Department of Chemistry and Materials, Faculty of Chemistry, Materials and Bioengineering, Kansai University, 3-3-35 Yamate-cho, Suita-shi, Osaka,
564-8680, Japan
c
ORDIST, Kansai University, 3-3-35 Yamate-cho, Suita-shi, Osaka, 564-0836, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Chemotherapy is the most reliable treatment for osteoporosis and osseous metastases. To facilitate better
Received 24 September 2016 drug delivery for bone treatments, a novel preparation of polymeric nanoparticles with high affinity to
Received in revised form 10 January 2017 bone has been prepared. Two-step synthesis of cholesteryl-functionalized poly(ethylene sodium phos-
Accepted 13 February 2017
phate) (Ch-PEPn ·Na) was performed via ring-opening polymerization of cyclic phosphoesters and the
Available online 16 February 2017
demethylation. The molecular weight of Ch-PEPn ·Na could be well controlled by changing the ratio of
cholesterol and cyclic phosphoesters. Because Ch-PEPn ·Na exhibits an amphiphilic nature in aqueous
Keywords:
media, Ch-PEPn ·Na-bearing nanoparticles (PEPn ·Na NPs) were prepared by a solvent evaporation tech-
Polyphosphoester
Nanoparticle nique. The size of the nanoparticles investigated in the current study is approximately 100 nm, which
Drug delivery system was determined by dynamic light scattering (DLS) and transmission electron microscopy (TEM). Due to
Bone therapy the presence of highly water-soluble polymer chains, dispersion of PEPn ·Na NPs in aqueous media was
Ring-opening polymerization stable for at least 1 week. Hemolytic activity of PEPn ·Na NPs was found to be low and PEPn ·Na NPs did not
Biodegradable polymer disintegrate mammalian cell membranes. Additionally, cytotoxicity of PEPn ·Na NPs was not observed at
concentrations below 100 ␮g/mL. The adsorption of PEPn ·Na NPs on hydroxyapatite (HAp) microparticles
was studied in comparison with poly(ethylene glycol) nanoparticles (PEG NPs). Both PEPn ·Na NPs and
PEG NPs adsorbed well onto HAp microparticles in distilled water with binding equilibrium constants
(KHAp ) for PEPn ·Na NPs and PEG NPs of 3.6 × 106 and 7.9 × 106 , respectively. In contrast, only PEPn ·Na NPs
adsorbed onto HAp microparticles in a saline phosphate buffer. Moreover, the adsorption of PEPn ·Na NPs
onto HAp microparticles occurred even in the presence of 1.2 mM calcium ions or low-pH media. The
affinity of the nanoparticles to bovine bone slices was also studied, with the result that large quantities
of adsorbed PEPn ·Na NPs were observed on the slices by scanning electron microscope.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction for metastatic cancer. Present chemotherapy procedures for bone


metastasis use anti-cancer drugs as well as bone-modifying agents,
Bone is the most common metastatic site in epithelial cancers such as bisphosphonates (BPs) [5–9] and denosumab [10,11], in
such as breast, prostate, and lung cancers [1,2]. Bone metastasis clinical practice. Both these agents can inhibit or prolong the start-
often causes symptoms, such as pathological fractures, hypercal- up period of skeletal-related events, such as surgery and radiation
cemia, and spinal cord compression, which significantly decrease therapy, which greatly contribute to improving the QOL of patients
the quality of life (QOL) of patients [3,4]. Treatments for bone [12,13]. Bone-modifying agents are also applied to improve the
metastatic cancers currently involve performing a combination of delivery of anticancer drugs to metastatic sites in bone. A variety
surgery, radiation therapy, and chemotherapy. Chemotherapy in of BP-bearing drug carriers have been proposed for this purpose
particular is believed to be one of the most essential treatments because their ability to form bidentate or tridentate chelates with
calcium ions results in a high affinity to hydroxyapatite (HAp),
the inorganic component of bone [14–19]. For example, BPs have
been immobilized to bind to the side chains of hydrophilic poly-
∗ Corresponding author at: Department of Chemistry and Materials, Faculty of
mers [16], liposomes [17], and polysaccharides such as pullulan
Chemistry, Materials and Bioengineering, Kansai University, 3-3-35 Yamate-cho,
Suita-shi, Osaka, 564-8680, Japan. [18]. However, BPs do not have sufficient functionality to conjugate
E-mail address: yasu.bmt@kansai-u.ac.jp (Y. Iwasaki). biofunctional molecules or anti-cancer drugs. Moreover, BP-related

http://dx.doi.org/10.1016/j.colsurfb.2017.02.015
0927-7765/© 2017 Elsevier B.V. All rights reserved.
Y. Hirano, Y. Iwasaki / Colloids and Surfaces B: Biointerfaces 153 (2017) 104–110 105

osteonecrosis of the jawbone has been reported [20,21]. To achieve previously described method [22,37]. Other chemical reagents
therapeutic bone treatments safely and efficiently, new drug deliv- were purchased from Wako Pure Chemical Inc. (Tokyo, Japan) and
ery materials and systems are required. were used without further purification.
Here, we propose poly(ethylene sodium phosphate) (PEPn ·Na)
as a new polymeric candidate material with an affinity to bone 2.2. Synthesis of cholesteryl-terminated poly(ethylene phosphate)
substrates. PEPn ·Na is an analog of polyphosphoesters and has a
phosphodiester backbone [22]. Polyphosphoesters have recently Scheme 1 shows the synthesis route of cholesteryl-terminated
been focused as new biodegradable polymers that have mimetic poly(ethylene phosphate) (Ch-PEPn ·Na; n: number of repeating
structures of nucleic acids [23–29]. Compared with conven- units), an amphiphilic polyphosphoester. 2.76 g (20 mmol) MP was
tional aliphatic polyesters, the molecular functionalization of placed into a sufficiently dry 20 mL Schlenk tube equipped with
polyphosphoesters is more easily achieved because various cyclic a three-way stopcock followed by drying under reduced pressure
phosphoesters, which function as monomers, can be obtained by a for 2 h. A certain amount of cholesterol, which had previously been
simple condensation reaction between an alcohol and 2-chloro-2- dried under reduced pressure overnight, was dissolved in 2 mL of
oxo-1,3,2-dioxaphosphorane (COP) [30,31]. That is, theoretically, dichloromethane. The cholesterol/dichloromethane solution and
any alcohol can be introduced into polyphosphoesters. Further- DBU (1.0 equivalent to cholesterol) were added under an Ar gas
more, the solubility of polyphosphoesters can be controlled by the atmosphere. The mixture was stirred with a magnetic stirrer in
substituent side chain groups. We have successfully polymerized an ice bath for 1 h and was further stirred at room temperature
cyclic phosphoester monomers using organocatalysts [28], which for 5 h. The reaction was quenched by the addition of acetic acid
have advantages of being able to undergo ring-opening polymer- (1.0 equivalent to DBU). The synthesized choresteryl-terminated
ization in living manner and to generate polyphosphoesters with poly(MP) (Ch-PMPn ; n: number of repeating units) was purified
a narrow molecular weight distribution (Mw /Mn < 1.1). Addition- by reprecipitation from toluene followed by diethyl ether, placed
ally, end-functionalized polyphosphoesters are easily synthesized into a desiccator, and dried under reduced pressure. The Ch-PMPn
because alcohol molecules can be used as polymerization initiators. was dissolved in 10 mL of distilled water and mixed with a 30%
As described above, cyclic monomers for polyphosphoesters can trimethylamine aqueous solution (7.88 g; equivalent to 2.0 MP
be obtained by a simple condensation reaction and various func- units). The mixture was stirred at room temperature for 24 h and
tional groups can be introduced into the side chains of polymers, subsequently stirred with 22.2 g of a cationic ion exchange resin
®
allowing versatile control of polyphosphoester solubility. The bio- (Amberllite IR-120, Merck KGaA, Darmstadt, Germany) at room
compatibility of polyphosphoesters has also been determined in temperature for 1 h. In order to completely transform polymers
previously published works, and several biomedical applications containing quaternized amimonium salts into polyacids (polyethy-
have been proposed such as drug- [24,25,27,32] and gene delivery lene phosphate), this procedure was repeated twice. The polymer
[33,34] and tissue engineering [35,36]. solution was then dialyzed against distilled water for 6 h and freeze-
Polyphosphoester ionomer, consisting of phosphotriester and dried in order to concentrate the solution. After concentrating, the
phosphodiester units, is one of the water-soluble polyphospho- pH of the aqueous solution was adjusted to 7.0 using a sodium
esters we have studied previously [37]. Interestingly, the amount hydroxide aqueous solution and the polymer solution was again
of polyphosphoester ionomer adsorption onto HAp substrates dialyzed against distilled water. Finally, Ch-PEPn ·Na was obtained
increased with more phosphodiester units (ionized units) in the in the solid state.
polyphosphoesters [38]. Moreover, polyphosphoester ionomers The apparent critical micelle concentration (cmc) value for Ch-
clearly exhibited superior physicochemical properties for mineral- PEPn ·Na in aqueous media was determined using a previously
ization and biocompatibility when compared with BPs. According described method using pyrene as a fluorescence probe [39].
to the earlier experiments, fully ionized PEPn ·Na is expected to
show high affinity to HAp. A more recent study evaluated the 2.3. Preparation of Ch-PEPn ·Na-bearing poly(L-lactic acid) (PLLA)
effects of PEPn ·Na on bone cell viability and activity [22], and found nanoparticles
that PEPn ·Na exhibited excellent biocompatibility with osteoblasts.
In contrast, selective inhibition of osteoclast adhesion and func- The Ch-PEPn ·Na-bearing nanoparticles (PEPn ·Na NPs) were pre-
tion was observed following cultivation with PEPn ·Na. PEPn ·Na pared using a solvent evaporation method [40–43]. Briefly, a 40 mL
should thus be considered as a unique polymer with diverse uses aqueous solution containing 10 mg/mL Ch-PEPn ·Na was placed in
for bone treatment; however, no studies have yet addressed the a sample tube and stirred at 450 rpm while cooled in an ice bath.
use of PEPn ·Na for processing drug carriers. In the current study, 1 mg PLLA was dissolved in 1.0 mL chloroform and then added drop-
amphiphilic PEPn ·Na was newly synthesized to prepare biodegrad- wise into the aqueous polymer solution. The mixture was sonicated
able polymer nanoparticles that can interact strongly with bone. using a probe-type generator (Sonifier 250, Branson, CT, USA) for
The characterization, biocompatibility, and mineral affinity of the 30 min and kept under reduced pressure for 30 min to evaporate the
nanoparticles are also described. chloroform. The nanoparticles formed were fractionated by cen-
trifugation at 45000 rpm at 4 ◦ C for 1.5 h (himac 70MX, Hitachi Koki
Co., Ltd., Tokyo, Japan). The nanoparticles were resuspended as a
2. Materials and methods precipitate in distilled water and centrifuged again under the same
conditions. This procedure was repeated three times to completely
2.1. Materials remove any free Ch-PEPn ·Na.
Poly(ethylene glycol)-bearing nanoparticles (PEG NPs) were
2-Chloro-2-oxo-1,3,2-dioxaphospholane (COP) was kindly sup- also prepared by using cholesteryl-terminated PEG (Mw = 6000;
plied by NOF Corporation (Tokyo, Japan) and cholesterol NOF Corporation, Tokyo, Japan) as an emulsifier. The preparation
was purchased from Kanto Chemical Co., Inc. (Tokyo, Japan). process of PEG NPs was identical to that of the PEPn ·Na NPs.
1,8-Diazabicyclo[5,4,0]-undec-7-ene (DBU) was obtained from The particle size, intensity size distribution, and ␨-potential of
Sigma-Aldrich (Saint Louis, USA), and was purified by dis- the nanoparticles were determined using a Malvern Instruments
tillation under reduced pressure, in which a fraction of bp Zetasizer (Worcestershire, UK). Scattering was performed with a
130 ◦ C/3.0 mmHg (lit.: bp 80 ◦ C/0.6 mmHg) was used. 2-Methoxy- vertically polarized incident beam at a wavelength of 633 nm sup-
2-oxo-1,3,2-dioxaphospholane (MP) was synthesized using a plied by a He-Ne ion laser with a scattering angle of 90◦ . The
106 Y. Hirano, Y. Iwasaki / Colloids and Surfaces B: Biointerfaces 153 (2017) 104–110

Scheme 1. Synthesis route for amphiphilic poly(ethylene phosphate).

morphology of the nanoparticles was observed by transmission aqueous media such as distilled water, PBS(−), PBS containing
electron microscopy (TEM; JEM-1400, JEOL, Tokyo, Japan). 1.2 mM calcium ions (PBS(+)), and 50 mM Tris-buffer (pH = 5.7).
®
HAp microparticles (10 mg, 22.82 m2 /g; Macro-Prep ; Ceramic
2.4. Cytotoxicity assay Hydroxyapatite Type II 80 ␮m, Bio-Rad, Hercules, CA, USA) were
added to the suspension and incubated for 3 h with gentle mix-
Mouse osteoblastic cells (MC3T3-E1) were used for cell culture ing. The supernatant was then collected and further centrifuged at
experiments because they are a well-characterized osteoblast-like 6200 rpm for 10 s (PMC-060, TOMY SEIKO Co., Ltd., Tokyo, Japan)
cell line that can serve as a model for endogenous osteoblasts to precipitate nanoparticles adsorbed on the HAp. The centrifuged
[44,45]. The cells were maintained in a culture medium [␣-media supernatant (100 ␮L) was mixed with ethanol and chloroform
modified essential medium (␣-MEM); Thermo Fisher Scientific, (1:2:1 vol ratio) to dissolve the nanoparticles. The fluorescence
Inc., Grand Island, NY, USA] containing 10% fetal bovine serum intensity (ex 570 nm, em 628 nm), which is due to eluted oxonol
and 1% antibiotic-antimycotic. The cells were routinely cultured in VI from the dissolved nanoparticles, was recorded with a fluo-
tissue-culture polystyrene flasks (Asahi Glass Co., Ltd, Tokyo, Japan) rescence spectrophotometer (F-2500, Hitachi, Tokyo, Japan). The
at 37 ◦ C in a humidified atmosphere of air containing 5% CO2 . After amount of nanoparticles adsorbed on the HAp microparticles was
treatment with trypsin- ethylenediaminetetraacetic acid (EDTA; then determined by comparing the concentration of nanoparticles
Gibco, Invitrogen Corporation, USA), the MC3T3-E1 osteoblasts in the supernatant with that in the original nanoparticle solution.
were harvested. The amount of adsorbed nanoparticles was calculated relative to
The suspension of MC3T3-E1 cells (100 ␮L; 5 × 104 cell/mL) in ␣- the unit surface area of the HAp microparticles.
MEM was introduced into a 96-well tissue-culture plate (Thermo
Scientific, Rochester, USA) and incubated in a culture medium at 2.7. Adsorption of PEP106 ·Na NPs on bone slices
37 ◦ C with 5% CO2 . After 24 h of cultivation, Dulbecco’s phosphate-
buffered saline (PBS (−)) (10 ␮L) containing either PEP106 ·Na NPs or Bone slices were prepared as previously described [47]. Briefly, a
PEG NPs was added to each of the wells and the cells were cultivated frozen bovine cortical femur was cut with a diamond saw (Buehler,
for a further 24 h. To determine the cell viability, a Cell Counting Kit- Lake Bluff, IL, USA) into slices of 130–180 ␮m thickness. The bone
8 (CCK-8; DOJINDO LABOLATORIES, Kumamoto, Japan) was used. slices (5 mm × 5 mm) were rinsed by ultrasonication in distilled
water, then soaked in an aqueous suspension containing 0.8 mg/mL
2.5. Hemolysis test of nanoparticles and stored for 3 h with gentle shaking. The spec-
imens were then rinsed with distilled water three times and
Hemolysis testing was carried out according to a previously freeze-dried under vacuum pressure. Subsequently, the specimen
described method [46]. Blood samples (20 mL) were collected from surfaces were coated with Au-Pd by a MSP-S1 magnetron sputter
healthy volunteers with a 30 mL syringe, to which was added 3 mL (Vacuum Device Inc., Ibaraki, Japan) and observed with a scanning
of a 3.8 wt% trisodium citrate aqueous solution. The collected blood electron microscope (SEM; JSM-6700, JEOL, Tokyo, Japan).
samples were centrifuged for 15 min at 1200 rpm (AllegraTM 21R
Centrifuge, Beckman Coulter, Brea, CA, USA) in order to precipitate 3. Results and discussion
the erythrocytes. The supernatant and interface were discarded
and Hank’s balanced salt solution (HBSS) in the amount of 1.5 3.1. Synthesis and characterization of Ch-PEPn ·Na
times that of the precipitated erythrocytes was added and resus-
pended. The solution was then centrifuged for 10 min at 2800 rpm Scheme 1 and Supporting Table S1 illustrate the synthesis route
and the supernatant and interface were discarded followed by dilu- and characterization of Ch-PEPn ·Na, respectively. The number of
tion with HBSS. This suspension was then diluted 10 times with repeating units (n) in the polymers and the molecular weight of
HBSS. 200 ␮L of PEP106 ·Na NPs suspended in HBSS and 800 ␮L of Ch-PEPn ·Na were determined by proton nuclear magnetic reso-
the erythrocytes/HBSS suspension were mixed followed by incu- nance (1 H NMR; ␣-500, JEOL, Tokyo, Japan) end group analysis. The
bation for 3 h at 37◦ (Dry Block Bath EB-603, AS-ONE, Osaka, Japan). ring opening polymerization of cyclic phosphoesters proceeds in a
The final preparation of PEP106 ·Na NPs had a concentration in the controlled/living fashion with organocatalysts [28]. The molecular
range of 1 × 10−4 –1 × 10−1 mg/mL. After the mixture was stored weight of Ch-PEPn ·Na was then controlled by changing the fractions
for 3 h at 37 ◦ C, it was centrifuged for 10 min at 2800 rpm and the of MP and cholesterol. During the transformation of polytriphos-
absorbance of the supernatant at 405 nm was measured using a phoester (PMPn ) into polydiphosphoester (PEPn ), no change in the
microplate reader (Model 680 Microplate reader, Bio-Rad, Her- number of repeating units (n) in the polymers was observed (See
cules, CA, USA). Solutions of pure HBSS and Triton-X 100/HBSS supporting Table S1).
(1 wt%) were used as positive and negative controls, respectively, Ch-PEPn ·Na consists of a hydrophilic chain and a hydropho-
in place of the PEP106 ·Na NPs suspension and were mixed with the bic cholesteryl end. Therefore, Ch-PEPn ·Na exhibits an amphiphilic
erythrocytes/HBSS suspension. nature and form micelles in aqueous solution. In order to deter-
mine the cmc of Ch-PEPn ·Na in a Tris-HCl buffer solution (10 mM,
2.6. Binding affinity of PEP106 ·Na NPs for HAp surfaces pH = 7.4), pyrene was used as a fluorescence probe. The ratios of
the pyrene peak intensities at 383 and 373 nm (I383 /I373 ) were
For fluorescent labeling of nanoparticles, 0.2 mg of oxonol VI plotted as a function of Ch-PEPn ·Na concentration because the flu-
was dissolved with PLLA in chloroform when the nanoparticles orescence emission spectra of pyrene are polarity dependent [39].
were prepared. The nanoparticles were suspended in various In the process of aggregation, the I383 /I373 values of pyrene are
Y. Hirano, Y. Iwasaki / Colloids and Surfaces B: Biointerfaces 153 (2017) 104–110 107

Fig. 1. TEM images of PEP106 ·Na NPs (a) and PEG NPs (b). Scale bar = 200 nm.

Fig. 2. Particle size and PDI of PEP106 ·Na NPs (a) and PEG NPs (b) suspended in water (䊉) and PBS(−) (䊏) during days 1–7. (PEP106 ·Na NPs; n = 3, PEG NPs; n = 3).

lower at concentrations below the cmc value, whereas the ratios water and PBS(−) at a concentration of 0.5 mg/mL and the parti-
are enhanced at higher Ch-PEPn ·Na concentrations when pyrene cle size and PDI were measured every day by DLS for 7 days. As
is encapsulated in the micelles. Therefore, the two fitting lines shown in Fig. 2, both PEP106 ·Na NPs and PEG NPs suspended in
reflecting the two states intersect at the turning point, the horizon- water showed high colloidal stability over one week. The negatively
tal ordinate of which determines the cmc values to be 1.0 × 10−2 charged ␨-potential of nanoparticles in water contributes to their
(Ch-PEP24 ·Na), 2.0 × 10−2 (Ch-PEP46 ·Na), and 4.0 × 10−2 g/dL (Ch- high dispersion ability. When PEP106 ·Na NPs and PEG NPs were
PEP106 ·Na), as shown in Supporting Table S1 and Supporting Fig. suspended in PBS(−), the ␨-potential values of the nanoparticles
S1. The cmc increased with an increase in the molecular weight decreased and the ␨-potential of PEG NPs became neutral due to
of the polymer. The phosphodiester polymer chain of Ch-PEPn ·Na the charge screening effect by the ionic environment (Table 2), indi-
is highly water soluble polymers [37]. Hydrophobic association cating that the dispersion ability of PEG NPs in PBS(−) was inferior
of cholesteryl groups can be then weakened by long hydrophilic to that of PEP106 ·Na NPs. The negative charge of PEP106 ·Na NPs in
polyphosphoesters. PBS(−) resulted in excellent dispersion stability.
The solvent evaporation technique for preparation of nanopar-
3.2. Preparation of PEPn ·Na NPs ticles composed of hydrophobic inner core and an outer shell of
hydrophilic polymers is reliable method to make nanoparticles.
The preparation of nanoparticles was performed by solvent This procedure has been used for preparation of various nanopar-
evaporation technique with amphiphilic polymer concentration ticulate drug carriers incorporating hydrophobic molecules [48].
above the cmc. DLS data for PEPn ·Na NPs and PEG NPs are shown
in Table 1 and Supporting Fig. S2. All nanoparticles prepared in
this study had diameters of approximately 100 nm (polydisper- 3.3. Biocompatibility of PEP106 ·Na NPs
sity index (PDI), 0.1–0.3). The surface ␨-potential of PEPn ·Na NPs
was around −70 mV and was highly negatively charged, while that Fig. 3 shows the viability of MC3T3-E1 cells in contact with vari-
of PEG NPs was around −40 mV. Fig. 1 shows the TEM images of ous concentrations of PEP106 ·Na NPs dissolved in PBS(−). MC3T3-E1
PEPn ·Na106 NPs and PEG NPs, indicating that the NPs are spherical cells in contact with PEP106 ·Na NPs displayed high viability values
in shape and that the diameters coincide well with the DLS data. similar to those with PEG NPs, and the 50% inhibitory concentration
To evaluate the stability of PEP106 ·Na NPs in the dispersion (IC50 ) value of PEP106 ·Na NPs was not confirmed in this concentra-
medium, PEP106 ·Na NPs and PEG NPs were suspended in distilled tion range.
108 Y. Hirano, Y. Iwasaki / Colloids and Surfaces B: Biointerfaces 153 (2017) 104–110

Table 1
Characterization of PEPn ·Na NPs and PEG NPs.

Sample SizeTEM (d nm)a SizeDLS (d nm)b PDIb ␨-potential(mV)b

PEP24 ·Na NPs – 99.2 0.24 −72.7


PEP46 ·Na NPs – 77.5 0.24 −64.8
PEP106 ·Na NPs 88.2 108.0 0.14 −67.2
PEG NPs 94.0 119.8 0.15 −39.2
a
Calculated from TEM images using SemAfore 5.21 (JEOL, Tokyo, Japan).
b
Determined DLS measurement in distilled water.

Table 2
Binding constant (KHAp ) of nanoparticles to HAp.

Nanoparticle Medium pH KHAp ( × 106 M−1 ) R2 ␨-potential (mV)

PEP106 ·Na NPs Water – 3.55 ± 2.02 0.99 −67.2


PBS(−) 7.4 9.75 ± 1.03 0.98 −30.9
PBS(+) 7.4 10.08 ± 1.45 0.99 −34.6
Tris-buffer 5.7 7.91 ± 3.43 0.90 −49.2
PEG NPs Water – 7.93 0.87 −39.2
PBS(−) 7.4 N.D. – −3.5

HBSS for a concentration range of 0.1–100 ␮g/mL, indicating that


there was no hemolysis activity. It was therefore concluded that
PEP106 ·Na NPs do not damage cell membranes.

3.4. Binding affinity of PEP106 ·Na NPs on HAp surfaces

Various tetracyclines, bisphosphonates, acidic oligopeptides,


chelating compounds, and salivary proteins have all been employed
to target bone diseases. These compounds bind to the inorganic
HAp portion of bone and have specificity for a certain HAp crystal
size [50]. Recently, we reported that polyphosphoester ionomers
had high affinity for HAp surfaces, or the inorganic component
of bone [38]. Because of their high affinity to HAp surfaces, it is
therefore expected that PEP106 ·Na NPs can induce extravasation
from blood vessels by enhanced permeability and retention effects
and bind strongly to bone surfaces. In order to determine their
Fig. 3. Viability of MC3T3-E1 cells in contact with nanoparticles (n = 3). binding affinity for HAp surfaces, PEP106 ·Na NPs were placed in
contact with HAp microparticles in various media, such as distilled
water, PBS(−), 1.2 mM of calcium ions containing PBS (PBS(+)), and
Tris-buffer (50 mM, pH = 5.7). According to preferred encapsula-
tion of hydrophobic molecules into nanoparticles, oxonol VI-loaded
PEP106 ·Na NPs were used to quantify the amount of PEP106 ·Na NPs
adsorbed onto the HAp surface. Supporting Fig. S3 shows the bind-
ing isotherms of nanoparticles on HAp surfaces. When distilled
water was used as a medium, the amount of adsorbed PEP106 ·Na
NPs on the HAp surface is similar to that of PEG NPs. In contrast,
the amount of adsorbed PEP106 ·Na NPs to HAp in PBS(−) is much
larger than that of PEG NPs.
In order to quantify the binding affinity on HAp surfaces, the
binding equilibrium constant for HAp (KHAp ) was calculated from
the Langmuir equation as follows:

qm · KHAp · c
q= (1)
1 + KHAp · c

Fig. 4. Hemolytic activity of nanoparticles (n = 3). where q is the quantity of adsorbed nanoparticles per unit amount
of HAp (␮g/mg), qm is the maximum quantity of adsorbed nanopar-
In vivo hemolysis is a serious phenomenon, in which the ery- ticles (␮g/mg), KHAp is the binding equilibrium constant (M−1 ), and
throcyte membrane is destroyed by certain factors, such as viruses, c is the nanoparticle concentration (mg/mL). To calculate KHAp , we
chemicals, and low osmotic pressure, and hemoglobin is released utilized the coverage value (coverage:  = q/qm ) such that the Lang-
from the erythrocyte [49]. In vivo hemolysis results in anemia muir equation can be transformed as follows:
and the released hemoglobin causes serious complications such as 1 1
renal failure, stroke, and heart attack. To investigate the effect of =1+ (2)
 KHAp · c
PEP106 ·Na NPs on erythrocyte membranes, the PEP106 ·Na suspen-
sion and the erythrocyte suspension were mixed. The hemolysis The coverage was calculated using the density of PLLA, and the
rates of PEP106 ·Na NPs, shown in Fig. 4, were similar to that of Avogadro number (NA ) was utilized to express KHAp as M−1 [51].
Y. Hirano, Y. Iwasaki / Colloids and Surfaces B: Biointerfaces 153 (2017) 104–110 109

Fig. 5. SEM images of bovine bone in contact with 0.8 mg/mL of PEP106 ·Na NPs and PEG NPs suspended in PBS(−). Scale bar = 1 ␮m.

Table 2 shows the KHAp values for PEP106 ·Na NPs and PEG NPs hemolysis activity or cytotoxicity against MC3T3-E1 cells. PEPn ·Na
to HAp. PEP106 ·Na NPs suspended in PBS(−) exhibited a high KHAp NPs also showed high in vitro affinity to bone components, includ-
value, while the KHAp value for PEG NPs could not be obtained. ing hydroxyapatite (HAp) and bone slices. Additionally, the affinity
Moreover, we evaluated the binding affinity of PEP106 ·Na NPs of PEPn ·Na NPs to HAp was not disrupted by the presence of cal-
on HAp in the presence of calcium ions [38]. It was predicted cium ions or low-pH conditions, which promote bone resorption by
that 1.2 mM of free calcium ions would inhibit the adsorption of activated osteoclasts. PEPn ·Na NPs therefore have high potential for
PEP106 ·Na NPs onto HAp surfaces due to possible electrostatic inter- delivery of hydrophobic drugs to bone.
action between PEP106 ·Na and calcium ions. However, the KHAp
value of PEP106 ·Na NPs suspended in PBS(+) was equally as high as Acknowledgments
that in PBS(−), indicating that free calcium ions do not influence the
adsorption of PEP106 ·Na NPs onto HAp surfaces. It is assumed that This work was supported by JSPS KAKENHI Grant Number
the interaction of calcium ions with the PBS(+) and PEP106 ·Na NPs (26107719, 15H03520, 15H03526, and 26107719), and also by
can be easily exchanged with other ions. However, the exchange Hitachi Metals Materials Science Foundation 2014.
of interaction between PEP106 ·Na NPs and HAp surface does not
easily occur because both PEP106 ·Na NPs and HAp surfaces contain Appendix A. Supplementary data
many interaction sites. The adsorption of PEP106 ·Na NPs onto HAp
surfaces would therefore not be suppressed even in the presence Supplementary data associated with this article can be found, in
of various other ions found in body fluids. the online version, at http://dx.doi.org/10.1016/j.colsurfb.2017.02.
Osteoclasts play an important role in bone metastasis by cre- 015.
ating space for cancer cells via bone resorption [52]. Osteoclasts
degrade the bone mineral component by secreting protons through
References
vacuolar H+ -ATPase, where the pH in the lacunae drops to nearly
4 [50,53]. The environment surrounding activated osteoclasts is [1] R.E. Coleman, Cancer Treat. Rev. 27 (2001) 165.
also acidic [54]. To deliver anti-cancer agents to bone, a carrier [2] G.R. Mundy, Nat. Rev. Cancer 2 (2002) 584.
[3] M. Kawatani, H. Osada, Cancer Sci. 100 (2009) 1999.
is required to have an affinity to bone under low-pH conditions.
[4] A. Chilla, D. Bianconi, N. Geetha, A. Dorda, M. Poettler, M. Unseld, D. Sykoutri,
Therefore, the affinity of PEP106 ·Na NPs to HAp surfaces was deter- K. Redlich, C.C. Zielinski, G.W. Prager, Exp. Cell Res. 337 (2015) 68.
mined under low-pH conditions (Tris buffer (50 mM, pH = 5.7)). The [5] C.H. Van Poznak, S. Temin, G.C. Yee, N.A. Janjan, W.E. Barlow, J.S. Biermann,
KHAp for this condition was 7.91 ± 3.43 × 106 M−1 and the KHAp L.D. Bosserman, C. Geoghegan, B.E. Hillner, R.L. Theriault, D.S. Zuckerman, J.H.
Von Roem, J. Clin. Oncol. 29 (2011) 1221.
value did not depend (p > 0.1) on the ion content (PBS(−) and [6] M. Aapro, F. Saad, Ther. Adv. Urol. 4 (2012) 85.
PBS(+)). Therefore, PEP106 ·Na NPs have high affinity to HAp surfaces [7] R.G.G. Russell, M.J. Rogers, Bone 25 (1999) 97.
under low-pH conditions similar to those found in bone metastatic [8] J.E. Junford, K. Thompson, F.P. Coxon, S.P. Luckman, F.M. Hahn, C.D. Poulter,
F.H. Ebetino, M.J. Rogers, J. Pharmacol, Exp. Ther. 296 (2001) 235.
sites. [9] A. Hokugo, S. Sun, S. Park, C.E. McKenna, I. Nishimura, Bone 53 (2013) 59.
Fig. 5 shows scanning electron microscopy (SEM) images of a [10] D.W. Dempster, C.L. Lambing, P.J. Kostenuik, A. Grauer, Clin. Ther. 34 (2012)
bovine bone slice in contact with 1 mL of 0.8 mg/mL PEP106 ·Na NPs 521.
[11] A. Dellis, A.G. Papatsoris, Expert. Opin Biol. Ther. 14 (2014) 7.
and PEG NPs suspended in PBS(−) for 3 h at room temperature. [12] M. Froehner, T. Hölscher, O.W. Hakenberg, M.P. Wirth, Urol. Int. 93 (2014) 249.
Similarly to KHAp , large quantities of adsorbed PEP106 ·Na NPs were [13] R.R. Mckay, X. Lin, J.J. Perkins, D.Y.C. Heng, R. Simantov, T.K. Choueiri, Eur.
observed on the bone slice surfaces. In contrast, the amount of Urol. 66 (2014) 502.
[14] E.V. Giger, B. Castagner, J.-C. Leroux, J. Control. Release 167 (2013) 175.
adsorbed PEG NPs on the bone slices was significantly lower. This
[15] S.I. Thamake, S.L. Raut, Z. Gryczynski, A.P. Ranjan, J.K. Vishwanatha,
demonstrates that PEP106 ·Na NPs have a particularly high affinity Biomaterials 33 (2012) 7164.
to bone slices. [16] S.W. Morton, N.J. Shah, M.A. Quadir, Z.J. Deng, Z. Poon, P.T. Hammond, Adv.
Healthcare Mater. 3 (2014) 867.
[17] G. Wang, N.Z. Mostafa, V. Incani, C. Kucharski, J. Biomed, Mater. Res. A 3
4. Conclusion (2012) 684.
[18] G. Bonzi, S. Salmaso, A. Scomparin, A. Eldar-Boock, R. Stachi-Fainaro, P.
Caliceti, Bioconjugate Chem. 26 (2015) 489.
To facilitate safer and more efficient drug delivery to bones
[19] S.-W. Choi, J.-H. Kim, J. Control Release 122 (2007) 24.
in comparison with BP-immobilized drug carriers, poly(ethylene [20] T. Ikeda, J. Kuraguchi, Y. Kogashiwa, H. Yokoi, T. Satomi, N. Kohno, Bone 73
sodium phosphate) (Ch-PEPn ·Na)-immobilized PLLA NPs (PEPn ·Na (2015) 217.
NPs) were prepared by a solvent evaporation technique. The [21] G. Campisi, S. Fedele, V. Fusco, G. Pizzo, O.D. Fede, A. Bedogni, Future Oncol. 10
(2014) 257.
PEPn ·Na NPs had diameters of approximately 100 nm and exhib- [22] S. Kootala, M. Tokunaga, J. Hilborn, Y. Iwasaki, Macromol. Biosci. 15 (2015)
ited high colloidal stability. Additionally, PEPn ·Na NPs did not show 1634.
110 Y. Hirano, Y. Iwasaki / Colloids and Surfaces B: Biointerfaces 153 (2017) 104–110

[23] S. Zhang, H. Wang, Y. Shen, F. Zhang, K. Seetho, J. Zou, J.-S.A. Taylor, A.P. Dove, [38] Y. Iwasaki, K. Katayama, M. Yoshida, M. Yamamoto, Y. Tabata, J. Biomat. Sci.
K.L. Wooley, Macromolecules 46 (2013) 5141. Polym. Ed. 24 (2013) 882.
[24] Q. Zhang, J. He, M. Zhang, P. Ni, J. Mater. Chem. B 3 (2015) 4922. [39] Ö. Topel, B.A. Çakir, L. Budama, N. Hoda, J. Mol. Liq. 177 (2013) 40.
[25] F. Zhang, S. Zhang, S.F. Pollack, R. Li, A.M. Gonzalez, J. Fan, J. Zou, S.E. Leininger, [40] T. Konno, K. Kurita, Y. Iwasaki, N. Nakabayashi, K. Ishihara, Biomaterials 22
A. Pavía-Sanders, R. Johnson, L.D. Nelson, J.E. Raymond, M. Elsabahy, D.M.P. (2001) 1883.
Hughes, M.W. Lenox, T.P. Gustafson, K.L. Wooley, J. Am. Chem. Soc. 137 (2015) [41] J. Watanabe, K. Ishihara, Biomacromolecules 7 (2006) 171.
2056. [42] Y. Iwasaki, Y.H. Maie, K. Akiyoshi, Biomacromolecules 8 (2007) 3162.
[26] Y.H. Lim, K.M. Tiemann, G.S. Heo, P.O. Wagers, Y.H. Rezenom, S. Zhang, F. [43] Y. Goto, R. Matsuno, T. Konno, M. Takai, K. Ishihara, Biomacromolecules 9
Zhang, W.J. Youngs, D.A. Hunstad, K.L. Wooley, ACS Nano 9 (2015) 1995. (2008) 3252.
[27] F. Li, J. He, M. Zhang, P. Ni, Polym. Chem. 6 (2015) 5009. [44] R.T. Franceschi, B.S. Iyer, Y. Cui, J. Bone Miner. Res. 9 (1994) 843.
[28] Y. Iwasaki, E. Yamaguchi, Macromolecules 43 (2010) 2664. [45] D. Wang, K. Christensen, K. Chawla, G. Xiao, P.H. Krebsbach, R.T. Franceschi, J.
[29] Y.-C. Wang, Y.-Y. Yuan, J.-Z. Du, X.-Z. Yang, J. Wang, Macromol. Biosci. 9 Bone Miner. Res. 14 (1999) 893.
(2009) 1154. [46] R. Ikeuchi, Y. Iwasaki, J. Biomed, Mater. Res. A 101 (2013) 318.
[30] Y. Iwasaki, K. Akiyoshi, Macromolecules 37 (2004) 7637. [47] M. Husheem, J.K. Nyman, J. Vääräniemi, H.K. Vaananen, T.A. Hentunen, Calcif.
[31] S. Zhang, J. Zou, F. Zhang, M. Elsabahy, S.E. Felder, J. Zhu, D.J. Pochan, K.L. Tissue Int. 76 (2005) 222.
Wooley, J. Am. Chem. Soc. 134 (2012) 18467. [48] K. Ishiharaa, W. Chena, Y. Liub, Y. Tsukamotoa, Y. Inoue, Sci. Tech. Adv. Mater.
[32] E.M. Alexandrino, S. Ritz, F. Marsico, G. Baier, V. Mailänder, K. Landfestera, F.R. 17 (2016) 300.
Wurm, J. Mater. Chem. B 2 (2014) 1298. [49] R. Mendonca, A.A.A. Silveira, N. Conran, Inflamm. Res. 65 (2016) 665.
[33] Y. Hao, J. He, M. Zhang, Y. Tao, J. Lin, P. Ni, J. Polym. Sci. Part A: Polym. Chem. [50] S.A. Low, J. Kopeček, Adv. Drug Deliv. Rev. 64 (2012) 1189.
51 (2013) 2150. [51] L. de Miguel, M. Noiray, G. Surpateanu, B.I. Iorga, G. Ponchel, Int. J. Pharm. 460
[34] T.-M. Sun, J.-Z. Du, L.-F. Yan, H.-Q. Mao, J. Wang, Biomaterials 29 (2008) 4348. (2014) 73.
[35] J.-Z. Du, T.-M. Sun, S.-Q. Weng, X.-S. Chen, J. Wang, Biomacromolecules 8 [52] J. Aoki, I. Yamamoto, M. Hino, C. Shigeno, N. Kitamura, T. Sone, K. Shiomi, J.
(2007) 3375. Konishi, Cancer 62 (1988) 98.
[36] X.-Z. Yang, T.-M. Sun, S. Dou, J. Wu, Y.-C. Wang, J. Wang, Biomacromolecules [53] S. Georges, C.R. Velasco, V. Trichet, Y. Fortun, D. Heymann, M. Padrines,
10 (2009) 2213. Cytokine Growth F. R. 20 (2009) 29.
[37] Y. Iwasaki, T. Kawakita, S. Yusa, Chem. Lett. 38 (2009) 1054. [54] A.-V. Rousselle, D. Heymann, Bone 30 (2002) 533.

S-ar putea să vă placă și