Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Contents lists available at ScienceDirect

Advanced Industrial and Engineering Polymer Research


journal homepage: http://www.keaipublishing.com/aiepr

In vitro osteoconductivity of PMMA-Y2O3 composite resins


Taigi Honma a, **, Elia Marin a, b, c, *, Francesco Boschetto a, b, d,
Muhammad Daniel bin Idrus a, Kai Mizuno a, Nao Miyamoto b, Tetsuya Adachi b,
Toshiro Yamamoto b, Narisato Kanamura b, Wenling Zhu a, Giuseppe Pezzotti a, b, d, e
a
Ceramic Physics Laboratory, Kyoto Institute of Technology, Sakyo-ku, Matsugasaki, 606-8585 Kyoto, Japan
b
Department of Dental Medicine, Graduate School of Medical Science, Kyoto Prefectural University of Medicine, Kamigyo-ku, Kyoto, 602-8566, Japan
c
Polytechnic Department of Engineering and Architecture, University of Udine, 33100, Udine, Italy
d
Department of Immunology, Graduate School of Medical Science, Kyoto Prefectural University of Medicine, Kamigyo-ku, Kyoto, 602-8566, Japan
e
The Center for Advanced Medical Engineering and Informatics, Osaka University, Yamadaoka, Suita, 565-0871, Osaka, Japan

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

Article history: Polymethyl methacrylate (PMMA) bone cements are used as “glue” between orthopedic/orthodontic
Received 4 June 2022 implants and bone, since they can be modeled and easily applied by the surgeon, while being chemically
Received in revised form and biologically stable for many years after surgery. In this research, Y2O3 powder was added to com-
1 August 2022
mercial PMMA bone cement to produce a composite resin, which was then characterized and tested
Accepted 14 August 2022
in vitro to evaluate the cell proliferation and the quality of osteoblastic bone formed in vitro on the
composite. Biological assays showed an increase in cell proliferation on the Y2O3-PMMA composite as
Keywords:
compared to the pristine sample. Alizarin Red staining (ARS) showed the amount of bone formed on a
Osteoconductivity
Composites
composite PMMA resin was about 30% higher than that on the pristine PMMA bone cement reference.
Polymethyl methacrylate The quality of bone tissue was evaluated using Raman spectroscopy, showing the bone tissue formed on
Y2O3 the composite had a better degree of mineralization and a higher maturity as compared to the tissue
Bone grown on the control sample. These preliminary results suggest that Y2O3 plays a biologically active role
in bone growth and Y2O3-composites are affordable, superior candidates as bone cement materials.
© 2022 Kingfa Scientific and Technological Co. Ltd. Publishing services by Elsevier B.V. on behalf of KeAi
Communications Co. Ltd. This is an open access article under the CC BY license (http://creativecommons.
org/licenses/by/4.0/).

1. Introduction ability to withstand mechanical stress even in chemically aggres-


sive environments such as when in contact with biological fluids
PMMA is an amorphous thermoplastic polymer produced from and tissues [1]. In the orthodontic field, PMMA resins are used to fill
methyl methacrylate (MMA) monomers and belongs to the family cavities in the teeth enamel and dentin or to produce dental plates
of acrylic resins. It is used in various applications for it is easy to dye [2].
and work with, in addition to its thermoplastic properties. For Even if PMMA possesses durability and stability, it does not
example, it is commonly used as material for optical devices possess the ability to stimulate cell proliferation or bone formation.
components such as lens for optical scope or light-guiding plates Improving osteoconductivity could drastically reduce the duration
for cellphones. of healing and hospitalization, especially for patients with long-
PMMA is also a bio-compatible (bio-inert) polymer commonly term bone related diseases such as osteoporosis or osteomalacia
used in the fields of clinical orthopedics and orthodontics. In or- [3,4]. Even in the absence of bone diseases, implant fixation might
thopedics, PMMA resins are usually referred to as “bone cement” fail over time if the implant does not bond well with the bones [5].
and used for fixing implants to bone or repairing bone fractures. Moreover, bacteria can proliferate in the gaps between bone and
The main reason for choosing PMMA for these applications is its implants, resulting in post-operative infections [6].
Bioactive PMMA composite might help to strengthen implant
fixation, stimulate bone regeneration and prevent bacterial adhe-
* Corresponding author. Ceramic Physics Laboratory, Kyoto Institute of Technol- sion. For these reasons, modern commercial bone cements nowa-
ogy, Sakyo-ku, Matsugasaki, 606-8585, Kyoto, Japan.elia-marin@kit.ac.jp days contain calcium ions to improve bioactivity or antibiotics to
** Corresponding author.
fight against post-operative infections caused by bacteria [7], but
E-mail address: elia-marin@kit.ac.jp (E. Marin).

https://doi.org/10.1016/j.aiepr.2022.08.003
2542-5048/© 2022 Kingfa Scientific and Technological Co. Ltd. Publishing services by Elsevier B.V. on behalf of KeAi Communications Co. Ltd. This is an open access article
under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Please cite this article as: T. Honma, E. Marin, F. Boschetto et al., In vitro osteoconductivity of PMMA-Y2O3 composite resins, Advanced Industrial
and Engineering Polymer Research, https://doi.org/10.1016/j.aiepr.2022.08.003
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

the widespread use of antibiotics in both animals and humans has


contributed to the rise of so-called “superbugs”, namely, bacteria
that have grown resistant to a wide range of antibiotics [8].
There are two main methods for fixation of orthopedic implants,
either with or without the use of bone cements. For example,
cementless HIP joint prostheses are designed with small grooves
that increase the surface area, providing grip for the bone to adhere
[9]. In this kind of fixation, the surface properties such as macro-
and micro-porosity are important factors to consider [10], but the
adhesion can also be enhanced by using bioactive coatings such as
hydroxyapatite or bio-glass [11,12]. The second method of fixation
is by using the acrylic resins as bone cements, which can also
benefit from the presence of bioactive substances in their formu-
lation [13].
Over the years, many additives were used in order to make
PMMA bio-active. A few recent examples being fluorapatite and
graphene oxide, which improved mechanical properties, increased
cell viability and improved cellular adhesion [14], carbon nano-
tubes and graphene oxide, which increased the overall cellular
activity [15]. Even if other innovative materials are often tested, the
most common additives used in PMMA bone cements to improve
their bioactivity are based on calcium-phosphates [16e18], due to
their superior chemical affinity to the surrounding bone tissue.
Y2O3 is commonly used for phase stabilization in zirconia-based
bio-ceramics, referred to as Yttria-Stabilized-Zirconia (YSZ).
Tetragonal zirconia doped with 3 mol.% of yttria is widely used in
the field of restorative dentistry, but similar compositions have also
been applied in the orthopedic field to produce femoral heads and
knee femoral components [19].
Although Y2O3 has been widely used for applications inside the
human body, few investigations have been performed to assess its Fig. 1. (a) Pristine PMMA substrate and (b) composite containing 4.0 vol % of Y2O3. (c)
effect on mesenchymal stem cell lines and their osteoconductivity. Aggregation of Y2O3 particle observed by SEM under 5000  magnification.
Some research indicated that Y2O3 particles possess a strong anti-
oxidant effect that reduces oxidative stress in cells [20].
In this work, after verifying the effects of Y2O3 to the surface biological testing, 4.0 vol % PMMA- Y2O3 composite was chosen as it
morphology and mechanical properties of the composite, the possessed similar mechanical strength to the pristine PMMA
in vitro effects of the Y2O3-PMMA bone cement composite was sample.
investigated. It was found that the addition of Y2O3 has a positive Before all in vitro testing, samples were rinsed with distilled
effect on cell proliferation and bone tissue formation on the com- water and sterilized under UV.
posite according to ARS, SEM/EDS (Scanning Electron Microscopy/
Energy Dispersive X-ray spectroscopy) and Raman imaging. The
2.2. X-ray diffraction (XRD)
Y2O3-added composite improves both quantity and quality of the
bone formed. Considering the results, Y2O3-added composite may
XRD analyses for Y2O3 powder were conducted using a desktop
have a potential for application in the medical field.
MiniFlex300/600 diffractometer (Rigaku, Kyoto, Japan) equipped
with a Cu source, in a 2q/q configuration. The 2q range was
2. Experimental procedure
comprised between 15 and 70 with a step of 0.02 . Measurements
were performed 3 times (N ¼ 3). The obtained data was analyzed by
2.1. Sample production
commercially available software (PDXL. Rigaku, Kyoto, Japan). The
wavenumber of the source was 1.54 Å. The lattice constant was
Commercially available Y2O3 ceramic powder (Shin-Etsu
obtained following the Bragg's law:
Chemical, Tokyo, Japan) was utilized as an additive.
In a preliminary biocompatibility comparison test, pure Y2O3 2dsinq ¼ l (1)
and Si3N4 powders, as well as a 50%/50% mixture of the two, were
molded into green ceramic discs at a pressure 20 KPa and subse- qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
quently tested in vitro. Si3N4 was used as a reference material due a ¼ d h2 þ k2 þ l2 ð AÞ (2)
to the recently increased scientific interest on its potential appli-
cations as a bioactive material. (d: lattice spacing, q: peak position/2, l ¼ 1.54 Å, a: lattice constant,
For composite bone cements, PMMA powder and MMA liquid (h k l): Miller indices).
were provided by a commercial producer (Shofu quick resin O, As an additional analysis for the powder, grain size was calcu-
Shofu, Kyoto, Japan). Four different types of samples were pre- lated following the Scherrer's formula:
pared: one pristine PMMA sample as a control, and PMMA com-
posites containing 1.2 vol %, 4.0 vol %, and 7.3 vol % of Y2O3 powder D ¼ K l=Bcosq ð
AÞ (3)
respectively. The diameter of the aggregated Y2O3 particle was
approximately 1e2 mm (Fig. 1c). The disc diameter and thickness (D: grain size, K: Scherrer's constant (¼ 0.89), B: FWHM).
were about 13 mm and 3.5 mm, respectively (Fig. 1a and b). For The strongest peak (2 2 2) was used for both calculations.
2
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 2. Dumbbell shaped tensile testing sample.

2.3. Laser microscopy

Micrographs were taken using 3D laser-scanning microscopy


(VK-X series, Keyence Osaka, Japan). This technique was used to
characterize the surface before the cell culture and to measure the
area covered by bone tissue after the cell culture. The analysis was
performed with micrographs taken under the magnifications of
10  , 20  , and 50  . The microscope used an automated x-y stage
and an autofocus function for z range. The surface roughness values
were obtained from the average of 10 measurements. To evaluate
the influence of Y2O3 powders at a micrometric scale, areas of
50  50 mm were cut from the images acquired at 50  and flat-
tened by using 2nd order baseline removals before measuring.

2.4. X-ray photoelectron spectroscopy (XPS)

Tests were performed using a photoelectron spectrometer (JPS-


9010 MC, JEOL, Tokyo, Japan) with an X-ray source of mono-
chromatic MgKa (output 10 kV, 10 mA). The measurements were
conducted in a vacuum chamber at around 1.0  106, upon setting
Fig. 3. Averaged XRD spectrum of Y2O3 powder. (N ¼ 3).
the analyzer pass energy to 10 eV and the voltage step size to 0.1 eV.

Fig. 4. surface morphology of preliminary green ceramic discs before and after in vitro testing with KUSA-A1 cells; pristine (a) Y2O3, (b) 50%/50% and (c) Si3N4 samples and in vitro
tested (d) Y2O3, (e) 50%/50% and (f) Si3N4 samples, same magnifications.

3
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 5. SEM/EDS map of the surface of the sample consisting of 50% Y2O3 powder and 50% of Si3N4 powder, after biological testing in vitro with KUSA-A1.

Fig. 6. Laser micrographs of (a) pristine PMMA sample surface, (b) 1.2 vol% Y2O3, (c) 4.0 vol% Y2O3 and (d) 7.3 vol% Y2O3 composites observed at a magnification of 50  .

4
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

(LabSpec, Horiba/Jobin-Yvon, Kyoto, Japan). All spectra have been


post processed by removing the baseline and reducing the noise
with a moving average filter. After fitting with Gaussian curves, the
intensity, the full width at half maximum (FWHM), and the area of
specific bands (at a range of 1640e1690 cm1 for collagen amide I,
1070 cm1 for carbonate apatite and 960 cm1 for phosphate
apatite) have been used to evaluate the quality of bone tissue, by
calculating the mineral to matrix ratio (RM/M), crystallinity of
apatite, carbonate to apatite ratio and the ratio between hydroxy-
apatite (HAp) and b-tricalcium phosphate apatite (b-TCP).

2.7. Mechanical testing

The tensile testing was performed with dumbbell shaped sam-


ples (Fig. 2) using the MCT 2150 Desktop Tensile-Compression
Tester (AND Discover Precision, Tokyo, Japan) at an elongation
rate of 10 mm/min. Tests were repeated three times for each Y2O3
concentration. The Young's modulus was calculated from the initial
slope of the stress-strain curve, while the maximum stress was
obtained directly from the curves.
Fig. 7. Surface roughness (Ra) of each sample with different concentration of Y2O3 as
measured by Laser Microscopy, red for macroscopic roughness (measured on the
Compressive testing was performed on cylindrical samples
whole 50  images) and blue for the microscopic roughness (acquired on insets of (length 10 mm, diameter 4 mm) at a strain rate of 1.3 mm/min
100 mm  100 mm after flattening). following the standard ASTM F451-21.

2.8. Vickers indentation


Table 1
Atomic percentage on the surface of pristine sample and composites, as measured by
XPS. The samples were indented with a Micro Vickers hardness tester
(Micro Hardness tester, SHIMADZU, Kyoto, Japan). Five measure-
Sample Carbon [at. %] Oxygen [at. %] Yttrium [at. %]
ments with different points were taken for each concentration.
PMMA 82.3 ± 1.4 17.7 ± 1.4 N/D Indentations were then observed with Laser microscope and the
PMMA þ1.2 vol% Y2O3 81.0 ± 1.3 18.2 ± 1.2 0.8 ± 0.1
lengths of two diagonals (d1 and d2) from one indentation square
PMMA þ4.0 vol% Y2O3 79.5 ± 1.0 18.9 ± 0.7 1.6 ± 0.5
PMMA þ7.3 vol% Y2O3 78.6 ± 1.4 19.1 ± 1.5 2.4 ± 0.7 were measured, and then, Vickers hardness (HV) was calculated
with following the formula:
a
For both samples, a narrow scan was carried out acquiring 30 scans F ½kgf  2F,sin ½kgf  1:854 h . i
HV ¼   ¼  2 h2 i ¼  2 kgf mm
2
(4)
repetitions on each investigated region and specifically: Carbon S mm 2
(C1s, ~285 eV), Oxygen (O1s, ~530 eV), Yttrium (Y3d, ~160 eV). The
d mm2 d
X-ray angle of incidence and the takeoff angle were 34 and 90 ,

respectively. Measurements were performed on 3 different areas d1 þ d2
for each sample (N ¼ 3). The penetration depth of the XPS probe F ¼ 1; d ¼ ; a ¼ 136
2
was in the order of 10 nm.
2.9. Biological testing
2.5. Scanning electron microscopy and Energy Dispersive X-ray
spectroscopy The KUSA-A1 mouse osteoblastic cell line (JCRB, Osaka, Japan)
was selected to form the bone tissue on the samples. The cells were
A field-emission-gun scanning electron microscope (FE-SEM) cultured and incubated in an osteoblast-inducer medium consist-
equipped with an Energy Dispersive X-ray spectroscopy (EDS) (JSM ing of 4.5 g/L of glucose Dulbecco's modified Eagle medium
7001F Scanning Electron Microscope, JEOL, Tokyo, Japan) was used (DMEM) (D-glucose, L-glutamine, phenol red, and sodium pyruvate)
to acquire high-resolution images and chemical composition maps supplemented with 10% of fetal bovine serum. Cells were allowed
of the surface, before and after the cell culture. All images were to proliferate within a Petri dish for 24 h at 37  C with the initial
collected at an acceleration voltage of 15 kV and magnifications of concentration of KUSA-A1 at 5  105 cell/mL. Then, the cultured
100  and 500  before the culture, and of 200  and 1000  after cells were deposited to the surface of the samples. Bone formation
the culture. All samples were sputter-coated (Gressington, Watford, took place within the osteogenic medium, which consisted of
UK) with a thin (20e30 Å) layer of platinum to improve their DMEM supplemented with the following compounds:
electrical conductivity. 50 mg/mL of ascorbic acid, 10  103 M of b-glycerol phosphate,
100  103 M of hydrocortisone, and 10% of fetal bovine calf serum.
2.6. Raman spectroscopy All samples were incubated for 14 days at 37  C while the medium
was changed 4 times during the incubation term.
Raman images and spectra of the sample surface and the bone
tissue produced by KUSA-A1 cells were collected at room temper- 2.10. Viability assay
ature by Raman-touch (Nanophoton, Osaka, Japan) with a
20  optical lens for surface characterization. The excitation source The number of living cells on the surface of the samples after
was a 785 nm laser with a nominal power of 52 mW. The obtained 24 h of the cell culture was analyzed with colorimetric assay based
spectra were analyzed by commercially available software on water -soluble tetrazolium (Cell Counting Kit-8, Dojindo,
5
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 8. Raman spectrum of (a) Y2O3 particles, (b) PMMA-4.0 vol% Y2O3 composite and (c) pristine PMMA reference. Raman imaging of (d) pristine reference and (b) Y2O3 particles,
(b) PMMA-4.0 vol% Y2O3 composite. Red signal for Y2O3 (band at 378 cm1), green signal for PMMA (band at 1450 cm1).

Kumamoto, Japan). A water-soluble formazan dye is produced upon 0.05 are considered statistically relevant with two asterisks (**)
the reduction of the colorimetric indicator (WST-8) induced by the and statistically relevant with one asterisk (*) respectively. Values
presence of electron mediator. The amount of the formazan dye of p larger than 0.05 were considered as statistically non-significant
generated indicates the number of living microorganisms present. and marked as “n.s.”.
Solution was analyzed using microplate readers (Emax, Molecular
Devices, Sunnyvale, California, USA) by collecting the related OD 3. Results
value.
3.1. Y2O3 powder characterization
2.11. Alizarin Red staining
Fig. 3 shows the average XRD diffraction pattern as obtained on
The area where bone formed was visualized by Alizarin Red the ceramic reinforcement powder. Peaks associated with Y2O3 can
staining (ARS) (Sigma Aldrich, Tokyo, Japan). The cells were rinsed be observed around 2q ¼ 21, 29 , 34 . 48 and 58 . The composi-
with distilled water twice and then fixed with 4% Para- tion of this powder determined with PDXL software based on the
formaldehyde phosphate buffer solution (PFA) (FUJIFILM Wako peak position was (Y 0.95: Eu 0.05)2O3. Moreover, the crystal
Pure Chemical Corporation, Osaka, Japan). The bone tissue was left structure was determined to be cubic by the analytical software. In
stained with ARS for 30 min at room temperature. After stained, the result of calculation using peak at 29.2 (2 2 2), lattice constant was
samples were rinsed 5 times with distilled water to remove excess 10.57 ± 0.01 Å and, the lattice volume was 1.18 nm3. Grain size
ARS. Then, micrographs of the surface were taken with Laser mi- computed using the (2 2 2) peak followed the Scherrer's formula
croscope under magnification of 10  and 20  . For the quantifi- and was determined to be 33.24 ± 1.02 nm.
cation of staining density, the staining was removed with 10%
Formic acid (Wako Pure Chemical Industries, Osaka, Japan) and the 3.2. Preliminary in vitro comparison with Si3N4
concentration was determined by measuring the absorbance at
450 nm. Fig. 4 shows the surface morphologies of the preliminary green
ceramic discs before and after in vitro testing with KUSA-A1
2.12. Statistical analysis mesenchymal cells. On all three compositions, 100% Y2O3, 50%
Y2O3/50% Si3N4 and 100% Si3N4, deposits of mineralized tissue were
All the data were expressed as mean values ± one standard formed, but while for Si3N4 the deposits resulted in a thin layer
deviation and were analyzed for their statistical significance using covering the ceramic crystals, the samples containing Y2O3 showed
ANOVA multi-variance test; p values which smaller than 0.01 and large aggregates with complex, three dimensional morphologies,
6
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 9. Scanning electron microscope (SEM) images of (a) pristine PMMA sample surface, (b) 1.2 vol% Y2O3, (c) 4.0 vol% Y2O3 and (d) 7.3 vol% Y2O3.

in order to identify the locations of the three main elements: sili-


con, yttrium and calcium. Carbon, another major constituent of the
EDS maps and mainly associated with the organic materials, is not
shown in Fig. 5 due to its presence at all locations while phos-
phorous, the second main constituent of the bone apatite, is not
shown as it overlaps perfectly with calcium. Even if a weak carbon
signal was measured on the whole map, it can be observed that the
mineralized tissue is preferentially formed on the surface of Y2O3
particles, where it grows to form large deposits. This result suggests
that Y2O3 might be a more effective reinforcement ceramic for bone
cements when compared to Si3N4.
Fig. 6 shows the pristine morphology of the surface of the
pristine PMMA sample (Fig. 6a) and the PMMA-Y2O3 composite
(Fig. 6b). Fig. 6c shows the surface roughness (Ra) value measured
from the different samples. Both samples showed the characteris-
tics spherical bubbles, which were formed during the polymeri-
zation reaction of PMMA resins. The heat from the exothermic
reaction caused the dissolved air inside the MMA liquid to be
released thus forming the bubbles. These bubbles have an average
diameter of 50 mm. The particles of Y2O3 are clearly visible as a
white, well-dispersed phased on the surface of the composite
material (Fig. 6b, c and d), in particular in comparison with the
Fig. 10. EDXS analysis results for the pristine PMMA polymer and the composites with
different fractions of particulate. smooth surface (Fig. 6a) of the pristine control. The presence of the
particles resulted in a slight change in the morphology of the sur-
face, which appeared more wrinkled. The particle dispersion in the
suggesting that the presence of Y2O3 leaded to the formation of composite was observed to be homogeneous at all concentrations,
much higher amounts of mineralized tissues. when observed at relatively low magnifications.
Fig. 5 shows the morphology of the surface of the sample con- As shown in Fig. 7, there were no significant differences in
taining both ceramic powders after in vitro testing with KUSA-A1. macroscopic surface roughness between the four types of samples,
The optical imaging has been combined with the SEM/EDS maps proving that the addition of Y2O3 particles did not influence the

7
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 11. (a) Stress-strain curve measured by tensile test, (b) Stress of each different concentration of PMMA-Y2O3 composites, and (c) Young's modulus of each sample.

Table 2 near-infrared laser as the excitation source. In the spectrum of the


Ultimate compressive strength for the different samples. Y2O3 particles, shown in Fig. 8a, some relatively intense bands are
Sample Ultimate strength [MPa] Modulus [GPa] identified as good indicators for the presence of Y2O3 in the com-
posites. A total of 9 bands could be identified and in particular two
PMMA 86.4 ± 5.3 3.4 ± 0.4
PMMA þ 1.2 vol% Y2O3 88.1 ± 3.4 4.1 ± 0.4 bands at relatively low Raman shifts, 378 and 470 cm1 (Fg þ Ag).
PMMA þ 4.0 vol% Y2O3 91.4 ± 4.2 4.6 ± 0.6 Bands at higher shifts, 1334, 1363, 1535, 1563, 1598, 1799 and
PMMA þ 7.3 vol% Y2O3 84.6 ± 11.5 5.1 ± 0.8 1981 cm1 are not actually Raman bands but infrared photo-
luminescence bands caused by the presence of small amounts of
contaminants, in particular rare earth elements [21].
overall morphology of the surfaces. When the roughness is For PMMA, the bands appearing in Fig. 8c are caused by
measured on a micrometric-scale, however, the presence and vs(CeCeO), at 601 cm1, vs(CeOeC) at 815 cm1, a-CH3 rocking at
amount of reinforcing particulate plays a clear role, as shown by the 970 cm1, OeCH3 rocking at 990 cm1, ds(CeH) of OeCH3 at
blue bars. A statistically relevant increase in micro-surface roughness 1450 cm1 and finally C]O symmetric stretching at 1730 cm1
is shown between each increment of the amount of particulate. [22].
Table 1 shows the atomic percentage measured by XPS at the The mixture of PMMA and Y2O3, here represented in Fig. 8b by
external surface of each sample. Despite the low reliability of XPS the average spectrum collected on the sample containing 4% vol. of
data when focusing on carbon contents (carbon signals are usually Y2O3 particulate, clearly shows the presence of both molecules. It
overestimated due to surface contamination), it can be observed can be observed that most of the bands maintain the original in-
that the amounts of both oxygen and yttrium increase with the tensity ratio of the pristine spectra (cf. either Fig. 8a or Fig. 8c) with
amount of particulate, confirming that a fraction of the particles is the important exception of the PMMA band located at 990 cm1
exposed. The amount of exposed particles seems to increase with and assigned to OeCH3 rocking. This results indicate that the
the concentration in a linear fashion, suggesting that high contents presence of Y2O3 during the setting of the PMMA resin lead to a
of ceramic reinforcement could potentially lead to more bioactive higher oxidation of the polymer, probably due to the interfering of
composites. the particulate with the polymerization process.
Fig. 8 shows the most significant results of the Raman spec- Fig. 8d and e shows the Raman imaging results on the pristine
troscopy and imaging experiments, as collected using a 785 nm PMMA reference sample (green) and the composite containing 4%

8
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 12. (a) Laser micrographs of the indentation and (b) Vickers Hardness (HV) of each type of samples.

of the otherwise smooth surface of the pristine PMMA (cf. Fig. 9a).
As the fraction of ceramic particulate is increased, the secondary
phase seems to start protruding from the polymeric matrix, con-
firming a progressive increase in the roughness of the surface.
Fig. 10 shows the average chemical composition of the four
types of samples, as a function of particulate fraction. It can be
observed that all materials appear to be highly oxidized, with ox-
ygen content between 16% and 20%. This is probably due to the low
reliability of EDXS for the concentration of very light elements like
carbon and oxygen. Nevertheless, the yttrium concentration in-
crease with the fraction of particulate, further supporting the hy-
pothesis that the powders are well dispersed, at least in the first
few microns below the surface which account for the depth of the
EDXS probe.

3.3. Mechanical properties

Fig. 11 shows the results of tensile tests of the different samples.


Fig. 11a shows the representative stress-strain curves for four
samples containing different fractions of yttria. The strength and
Fig. 13. Optical density (OD) of the WST-8 viability assay for living KUSA-A1 cells
grown on the different samples. (N ¼ 5). Young's modulus were calculated from stress-strain curve and
shown in the histograms (Fig. 11a and b). In Fig. 11b, the pristine
PMMA sample has the highest strength, while strength decreased
vol. of Y2O3 particulate. It can be observed that, at these magnifi- by about 25% for all the composites regardless of the concentration
cations, the distribution of Y2O3 (red) seems to be more or less of Y2O3. Furthermore, the more the ceramic powder added, the
homogeneous on the surface of the composite sample, with the larger the data scattering. This was probably caused by the
exception of the polymeric bubbles formed on PMMA due to the decreasing homogeneity of PMMA due to the addition of Y2O3
presence of residual gas, where the signal of Y2O3 appears to be powder. Young's modulus was calculated from each stress-strain
weaker. These results are in line with the previous images acquired curve (Fig. 11c). For Young's modulus, pristine PMMA sample
by laser microscopy, where the Y2O3 particulate appeared to be showed the highest value. From Fig. 11c, we can deduce that as the
homogeneously distributed on the surface of the samples. amount of Y2O3 powder increases, the stiffness of the composite
Fig. 9 shows SEM maps of the surface of pristine PMMA sample decreases. On the other hand, from the different composites, 4.0 vol
and the various PMMA-Y2O3 composites, going from 1.2 vol% (b), % PMMA-Y2O3 composite had the most similar value of Young's
4.0 vol% (c) and finally 7.3 vol% (d). The relative atomic percentage modulus compared to the pristine PMMA sample. Other than ten-
of elements measured by EDS is shown in Fig. 10. In Fig. 6b, the sile, compressive stresses were also measured following the stan-
presence of Y2O3 particles can be observed mainly as a roughening dard ASTM F451-21. Results, summarized in Table 2, show that the
9
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 14. Images of (a) pristine PMMA sample and (b) PMMA- 4.0 vol% of Y2O3 composite stained with ARS. Laser micrographs of the stained surface of (c), (e) pristine PMMA sample
and (d), (f) PMMA- 4.0 vol% of Y2O3 composite under magnification of 10  and 20  respectively.

ultimate strength of the composite and, in particular, the elastic 3.4. In vitro osteoconductivity test
modulus increase with the fraction of particulate up to the highest
concentration, where despite the high elastic modulus the ultimate Fig. 13 shows results of optical density measured from the
strength drops. viability assay. The optical density is proportional to the number of
Fig. 12 shows the results of the indentation testing performed on living cells present on the surface of the sample. The composites
the surface of the samples. Fig. 12a shows the Vickers hardness ob- showed higher optical densities when compared to the pristine
tained from the images and the formula in (4), expressed in HV PMMA sample and the negative control. The composite containing
(GPa), as a function of the fraction of Y2O3 particulate. The measured 4.0 vol% of Y2O3, showed a 25% increase in optical density, which
hardness seems to be more or less constant, with only the high can be translated as a 100% increase in the number of cells,
fraction composite showing a different behavior and, in particular, an considering the value of the negative control that only contains PBS
higher average hardness but with a larger distribution. solution the other two composites show comparable behaviors,
3D images of the hardness indentations from the reference with a larger statistical dispersion for the composite containing
PMMA and the samples containing 1.2 vol% and 4.0 vol% of Y2O3 lower fractions of particulate.
confirm, shown in Fig. 12 b, c and d, further support the observation Fig. 14 shows images of the pristine PMMA sample and 4.0 vol%
that the hardness of the three samples is similar, as the tree prints of Y2O3 composite after the samples were stained with ARS [23].
have comparable diagonal lengths. For the sample containing The stained areas appeared to be wider on surface of the composite
7.3 vol% Y2O3, the hardness seems to greatly fluctuate depending on than that of pristine PMMA sample. In both cases, the staining did
the location. This is supported by the indentations in Fig. 12e and f, not cover the surfaces of the samples completely and preferential
one of which is much smaller than the other. distributions can be clearly observed in concave areas made by

10
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

pristine PMMA sample. When compared to the sample containing


4.0 vol% Y2O3, the other two composites showed comparable re-
sults (not shown in Fig. 14).
On composite resins, the OD value of ARS (Fig. 15) was between
5 and 40% higher respect to the pristine PMMA reference. As the OD
value is related to the amount of Ca ion which is produced by cells
on the surface, the KUSA-A1 cells on the composite showed more
active osteogenic reaction compared to the pristine PMMA sample.
Fig. 16 shows the Raman spectra and imaging of the surface of
the reference PMMA sample and the sample containing 4 vol% of
Y2O3 after in vitro testing with KUSA-A1 cells. The spectra, collected
in the two regions marked in yellow, were normalized on the main
peak of PMMA (1450 cm1). The regions were divided in two areas:
from 900 cm1 to 1225 cm1 and from 1400 cm1 to 1850 cm1. In
the first region (Fig. 16c), the composite and the pristine PMMA
spectra are dominated by the peak related to phosphate stretching
1
vibration (n1 PO3
4 ), located at 960 cm . Additionally, two shoulder
bands at 946 cm1 and 970 cm1 were observed and identified as
carbonate aptatite and tricalcium phosphate (TCP), respectively.
Two additional peaks at 1030 cm1 and 1070 cm1 were caused by
phosphate asymmetric stretch and the b-type carbonate symmetric
stretch. In the second region, the dominant peak was PMMA
Fig. 15. Optical density (OD) value of the Alizarin red stains (ARS) for each sample.
(1450 cm1, CH2 stretching vibration) and a relatively weak peak at
1660 cm1 which was associated with Amide I as a marker of
scratches and small pores. By comparing Fig. 14e and f, it can be collagen. The Raman imaging shows the intensity of the apatite
observed that, for the composite, the staining also spread outside of mineral with peak at 960 cm1 (phosphate stretching vibration)
the porous area. The presence of calcium ions was measured by ARS labelled as green tones, and of PMMA with the peak at 1450 cm1 as
staining [24] and the results showed that cells with apatite covered red tones. On the other hand, the intensity of Y2O3 is presented as
larger portions of the PMMA-Y2O3 composite compared to the peaks at 1535 cm1 and 1563 cm1 was labelled in blue color.

Fig. 16. Raman imaging maps collected on the surfaces of the (a) PMMA reference sample and (b) composite sample containing 4 vol% of Y2O3 (Green: HAp, Red: PMMA, and Blue:
Y2O3). Normalized Raman spectra after in vitro testing for both samples in the spectra regions between (a) 900 cm1 to 1225 cm1 and (b) 1400 cm1 to 1850 cm1.

11
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 17. Bone quality as evaluated by using Raman spectroscopy, (a) Quality of bone formation on each surface, FWHM of 960 cm1 band, (b) Carbonate-to-phosphate ratio, (c) HAp-
to-TCP ratio, (d) Mineral-to-matrix ratio. (n ¼ 8).

Fig. 12e shows the overlapping of PO3 4 phosphate stretching vi- Fig. 17d shows the result of mineral-to-matrix ratio (RM/M),
bration band attributing to HAp (green) at 960 cm1 and CH2 which provides information of the composition of the bone tissue.
stretching vibration band of PMMA (red) at 1450 cm1. Form the This value was obtained from the ratio of area of the peak of
image it proved that the composite sample was covered with a phosphate band at 960 cm1 and the peak at 1660 cm1 (related to
larger amount of bone formation than the pristine sample. the Amide I vibration in collagen). RM/M value indicated there were
Fig. 17a shows the Full width at half maximum (FWHM) of the no significant differences in the level of mineralization of bone
960 cm1, phosphate PO3 4 band which is related to mineral hy- formations on both samples [31].
droxyapatite. Mineral crystallinity can be estimated using FWHM: a Fig. 18 show SEM images of the surface of pristine PMMA sample
narrow FWHM of the phosphate band (960 cm1) represents a and PMMA-Y2O3 composite after culture of KUSA-A1. In Fig. 18a and
higher degree of crystallinity. The FWHM which was observed on c, it was observed that a layer of bone tissue formed on the pristine
the composite was narrower compared to that on the pristine PMMA sample, but the layer appeared to be relatively thin as the
sample. This indicates that the bone produced on the composite surface of the sample is still clearly visible. On the other hand, the
reached higher bone crystallinity, which is also a considered surface of the composite sample Figs. 18b and d was covered in a
marker for good development and overall health of bone [28]. thicker layer of bone tissue, with some localized deposits pro-
Fig. 17b shows the carbonate-to-phosphate ratio which in- truding from the surface for more than 10 mm. As calcium and
dicates carbonate ion substitution in the apatite mineral, calculated phosphorus are the main elements of bone, Fig. 18e shows that
with the ratio of the area of carbonate apatite band (1070 cm1) and there are much more of these elements on the composite than the
the area of phosphate band (960 cm1). This parameter shows pristine PMMA sample. The amount of carbon of composite is lower
significant changes with different samples. Bone produced on the than of the PMMA, means that there is more bone covered on the
PMMA-Y2O3 composite had smaller carbonate-to-phosphate ratio surface of PMMA-Y2O3 composite.
than the pristine PMMA sample [30]. Fig. 19 shows the ratio of the EDS signals from calcium and
Fig. 17c shows the ratio of area of the main HAp band (960 cm1) phosphorous, as extracted from Fig. 18e. As the amount of Y2O3 is
and the sum of area of b-TCP bands (946 cm1 and 970 cm1). progressively increased, so is the amount of calcium, indicating a
Results showed the amount of HAp on the composite was higher higher maturity of the bone tissue formed, which gets closer to the
than the control sample. theoretical value for healthy bone (Ca/P ¼ 1.6).

12
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Fig. 18. SEM images of the bone formation on surface of (a) pristine PMMA sample and (b) PMMA-Y2O3 composite under magnification of 200  . EDS images of the bone on the
surface of (c) pristine PMMA sample and (d) PMMA-Y2O3 composite under magnification of 1000  (Blue: Calcium, Green: Phosphorus, Red: Carbon) (e) Atomic percentage of bone
formation on surface of each sample after cell culture.

4. Discussion without decreasing the bone quality compared to common PMMA


bone cement.
PMMA-Y2O3 composite promoted proliferation of KUSA-A1 cells Analysis based on SEM, EDS and Laser microscopy before cell
and stimulated much larger amounts of bone tissue formation, culture demonstrated that the composite had Y2O3 ceramic powder

13
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

on the outer surface, but these particles did not lead to a significant components in bone structure [31]. This result should support that
increase in surface roughness. One of the main reasons was prob- the composite does not decrease the relative mineral content and
ably due the aggregation of the particles as the diameter of the maintain the “healthy level” of bone tissue. Moreover, based on the
clusters observed with SEM ranged around 1e2 mm. Suggested by both results of FWHM and carbonate-to-apatite ratio, the bone
M. Lampin et al. [32], cell attachment ability should be affected by crystallinity on the composite increased and the ion substitution
the surface roughness. From the results we can conclude that ratio in the HAp lattice decreased (Fig. 13a and b). Decrease in HAp
addition of Y2O3 powder up to at least 7.3 vol% may maintain a crystallinity should be induced through ion substitution in the
propensity of cell attachment similar to that to pristine PMMA bone crystalline lattice [38] and the integration of carbonate ion causes
cement, only from the viewpoint of surface roughness. distortion in the crystal structure of apatite [39e43]. In this work, it
Analysis for mechanical properties by tensile test and indenta- is predicted that the composite may induce cell to produce more
tion suggested that the addition of Y2O3 powder into PMMA sample stable apatite especially in terms of solubility in water or acidic
caused decrease in strength and stiffness even if there were no environmental [44]. In addition, the result of the ratio between HAp
significant differences. Additionally, for strength, the increase of and b-TCP (Fig. 13c) indicates that the composite should produce
standard deviation in the results shows PMMA lost its homoge- better quality bone tissue compared to the pristine PMMA sample
neities in proportion to the amount of Y2O3 powder added. Simi- because as reported by L. Sun et al. [45], relative biological solubility
larly, the same pattern was observed for indentation. Especially, the of these two substances is: b-TCP > HAp.
sample containing 7.3 vol % of Y2O3 powder showed an exceptional In short, PMMA-Y2O3 composite should have contribute to not
change in mechanical properties compared to the other samples. only improving both the quantity and the quality of the bone.
The results of viability assay proved that PMMA-Y2O3 composite
promoted better cell proliferation if compared to that with the 5. Conclusion
pristine PMMA sample as shown in Fig. 9. Y2O3 powder is likely to
affect the metabolism of the cells. Mitochondria in cells produce In this research, bioactive bone cement consisting of PMMA and
reactive oxygen species (ROS) which have negative effects to cells. Y2O3 was produced by simply mixing a commercially available
Y2O3 are known as free radicals’ scavenger [33] and may act as an PMMA resin with the Y2O3 ceramic powders.
antioxidant, protecting the surrounding cells from the ROS [16]. As
reported by Xiang song et al. [34], Y2O3 particles also contributed to - Results suggest that PMMA-Y2O3 composite has a positive effect
the inhibition the ROS/NF-k-B pathway. It is possible that Y2O3 on KUSA-A1 osteoblastic cells as it stimulates cell proliferation
particle reduced the oxidative stress of KUSA-A1 cells during the and bone production, while increasing bone tissue quality
proliferation thus enable the contribution of higher number of compared to common PMMA bone cement;
living cells. The images of the surface and optical density value (OD - From the viewpoint of mechanical properties, addition of Y2O3
value) after the cell culture and in vitro testing indicated that Y2O3 powder to PMMA affected its strength, Young's modulus, and
had a positive effect on bone production. hardness negatively. Moreover, data scattering in those value
Moreover, RM/M of each sample showed no significant differ- was also influenced by the amount of Y2O3 powder added to the
ences (Fig. 13d), even though Y2O3 promoted the production of composite;
much more bone matrix as shown in Figs. 10e14. This fact indicates - PMMA-Y2O3 composite has potential for application in ortho-
that the amount of type I collagen which works as a scaffold for pedics and orthodontics.
bone formation [35e37] existed plentifully on the composite
compared to the pristine PMMA control sample. Based on biological The authors want to emphasize that this manuscript contains
results, it can be concluded that Y2O3 has an ability to create much preliminary characterizations that should be supported with
better environment for cell by reducing ROS, and then these cells additional in vitro cytotoxicity experiments, as well as long-term
could produce collagen proportionally thus making it possible to performances in simulated body fluids, before moving to in vivo
increase the amount of bone tissue formed on the scaffold. experiments.
For bone quality, PMMA-Y2O3 composite showed an ideal RM/M,
a value that indicates the ratio between organic and inorganic CRediT authorship contribution statement

Taigi Honma: Writing-original draft, Formal analysis, Inves-


tigation.
Elia Marin: Conceptualization, Methodology, Project adminis-
tration, Writing-review & editing, Supervision.
Francesco Boschetto: Methodology, Investigation, Supervision.
Muhammad Daniel bin Idrus: Investigation, Writing-review &
editing.
Kai Mizuno: Investigation.
Nao Miyamoto: Resources.
Tetsuya Adachi: Resources, Formal analysis.
Toshiro Yamamoto: Resources, Supervision.
Narisato Kanamura: Resources, Supervision.
Wenling Zhu: Formal analysis, Writing-review & editing,
Supervision.
Giuseppe Pezzotti: Writing-review & editing, Supervision.

Funding

This work was supported by JSPS KAKENHI Grant Number


Fig. 19. Ca/P ratio as measured by using EDS on the surface of the different samples.
18K17175.
14
T. Honma, E. Marin, F. Boschetto et al. Advanced Industrial and Engineering Polymer Research xxx (xxxx) xxx

Conflicts of interest [19] C.M. Cristache, M. Burlibasa, G. Cristache, S. Drafta, I.A. Popovici, A.A. Iliescu,
L. Burlibasa, Zirconia and its biomedical applications, Metalurgia International
16 (7) (2011) 18 [15] Hosseini, A., Sharifi, A. M., Abdollahi, M., Najafi, R.,
The authors declare the following financial interests/personal Baeeri, M., Rayegan, S., ... & Safa, M. (2015). Cerium and yttrium oxide
relationships which may be considered as potential competing nanoparticles against lead-induced oxidative stress and apoptosis in rat hip-
interests: Elia Marin reports financial support was provided by pocampus. Biological trace element research, 164(1), 80-89.
[20] D. Schubert, R. Dargusch, J. Raitano, S.W. Chan, Cerium and yttrium oxide
Japan Society for the Promotion of Science. nanoparticles are neuroprotective, Biochem. Biophys. Res. Commun. 342 (1)
(2006) 86e91.
References [21] A. Gajovi c, N. Tomasi c, I. Djerdj, D.S. Su, K. Furi
c, Influence of mechano-
chemical processing to luminescence properties in Y2O3 powder, J. Alloys
Compd. 456 (1e2) (2008 May 29) 313e319.
[1] S. Saha, S. Pal, Mechanical properties of bone cement: a review, J. Biomed. [22] H.A. Willis, V.J. Zichy, P.J. Hendra, The laser-Raman and infra-red spectra of
Mater. Res. 18 (4) (1984) 435e462.
poly (methyl methacrylate), Polymer 10 (1969) 737e746.
[2] N. Murakami, N. Wakabayashi, R. Matsushima, A. Kishida, Y. Igarashi, Effect of [23] T. Miyazaki, S. Miyauchi, A. Tawada, T. Anada, S. Matsuzaka, O. Suzuki,
high-pressure polymerization on mechanical properties of PMMA denture Oversulfated chondroitin sulfate-E binds to BMP-4 and enhances osteoblast
base resin, J. Mech. Behav. Biomed. Mater. 20 (2013) 98e104. differentiation, J. Cell. Physiol. 217 (3) (2008) 769e777.
[3] H. Zhang, B.F. Ricciardi, X. Yang, Y. Shi, N.P. Camacho, M.P. Bostrom, Poly- [24] H. Puchtler, S.N. Meloan, M.S. Terry, On the history and mechanism of alizarin
methylmethacrylate particles stimulate bone resorption of mature osteoclasts
and alizarin red S stains for calcium, J. Histochem. Cytochem. 17 (2) (1969)
in vitro, Acta Orthop. 79 (2) (2008) 281e288. 110e124.
[4] J.X. Lu, Z.W. Huang, P. Tropiano, B.C. d'Orval, M. Remusat, J. Dejou, D. Poitout, ,
[28] G. Pezzotti, A. Rondinella, E. Marin, W. Zhu, N.N. Aldini, G. Ulian, G. Valdre
Human biological reactions at the interface between bone tissue and poly- Raman spectroscopic investigation on the molecular structure of apatite and
methylmethacrylate cement, J. Mater. Sci. Mater. Med. 13 (8) (2002) 803e809. collagen in osteoporotic cortical bone, J. Mech. Behav. Biomed. Mater. 65
[5] U.E. Pazzaglia, Pathology of the bone-cement interface in loosening of total
(2017) 264e273.
hip replacement, Arch. Orthop. Trauma Surg. 109 (2) (1990) 83e88. [30] T. Buchwald, K. Niciejewski, M. Kozielski, M. Szybowicz, M. Siatkowski,
[6] M. Ribeiro, F.J. Monteiro, M.P. Ferraz, Infection of orthopedic implants with
H. Krauss, Identifying compositional and structural changes in spongy and
emphasis on bacterial adhesion process and techniques used in studying subchondral bone from the hip joints of patients with osteoarthritis using
bacterial-material interactions, Biomatter 2 (4) (2012) 176e194. Raman spectroscopy, J. Biomed. Opt. 17 (1) (2012), 017007.
[7] G.H. Walenkamp, L.L. Kleijn, M. de Leeuw, Osteomyelitis treated with [31] E. Marin, N. Hiraishi, T. Honma, F. Boschetto, M. Zanocco, W. Zhu, G. Pezzotti,
gentamicin-PMMA beads: 100 patients followed for 1e12 years, Acta Orthop. Raman spectroscopy for early detection and monitoring of dentin deminer-
Scand. 69 (5) (1998) 518e522.
alization, Dent. Mater. 36 (12) (2020) 1635e1644.
[8] E. Tacconelli, E. Carrara, A. Savoldi, S. Harbarth, M. Mendelson, D.L. Monnet, [32] M. Lampin, R. Warocquier-Cle rout, C. Legris, M. Degrange, M.F. Sigot-Luizard,
A. Zorzet, Discovery, research, and development of new antibiotics: the WHO Correlation between substratum roughness and wettability, cell adhesion, and
priority list of antibiotic-resistant bacteria and tuberculosis, Lancet Infect. Dis. cell migration, J. Biomed. Mater. Res. Off. Soc. Biomater. Jpn. Soc. Biomater. 36
18 (3) (2018) 318e327. (1) (1997) 99e108.
[9] H. Yamada, Y. Yoshihara, O. Henmi, M. Morita, Y. Shiromoto, T. Kawano,
[33] S. Parham, D.H. Wicaksono, S. Bagherbaigi, S.L. Lee, H. Nur, Antimicrobial
E. Fukaya, Cementless total hip replacement: past, present, and future, treatment of different metal oxide nanoparticles: a critical review, J. Chin.
J. Orthop. Sci. 14 (2) (2009) 228e241.
Chem. Soc. 63 (4) (2016) 385e393.
[10] E.W. Morscher, Cementless total hip arthroplasty, Clin. Orthop. Relat. Res. [34] X. Song, P. Shang, Z. Sun, M. Lu, G. You, S. Yan, H. Zhou, Therapeutic effect of
(181) (1983) 76e91. yttrium oxide nanoparticles for the treatment of fulminant hepatic failure,
[11] F. Barrere, C.A. van Blitterswijk, K. de Groot, Bone regeneration: molecular and
Nanomedicine 14 (19) (2019) 2519e2533.
cellular interactions with calcium phosphate ceramics, Int. J. Nanomed. 1 (3) [35] W. Suchanek, M. Yoshimura, Processing and properties of hydroxyapatite-
(2006) 317.
based biomaterials for use as hard tissue replacement implants, J. Mater.
[12] M.H. Fathi, A. Doostmohammadi, Bioactive glass nanopowder and bioglass Res. 13 (1) (1998) 94e117.
coating for biocompatibility improvement of metallic implant, J. Mater. Pro- [36] J.D. Currey, The structure and mechanics of bone, J. Mater. Sci. 47 (1) (2012)
cess. Technol. 209 (3) (2009) 1385e1391. 41e54.
[13] E. Marin, F. Boschetto, M. Zanocco, T. Honma, W. Zhu, G. Pezzotti, Explorative [37] A.L. Boskey, Biomineralization: an overview, Connect. Tissue Res. 44 (1)
study on the antibacterial effects of 3D-printed PMMA/nitrides composites,
(2003) 5e9.
Mater. Des. 206 (2021), 109788. [38] J.J. Freeman, B. Wopenka, M.J. Silva, J.D. Pasteris, Raman spectroscopic
[14] F. Pahlevanzadeh, H.R. Bakhsheshi-Rad, E. Hamzah, In-vitro biocompatibility,
detection of changes in bioapatite in mouse femora as a function of age and
bioactivity, and mechanical strength of PMMA-PCL polymer containing fluo- in vitro fluoride treatment, Calcif. Tissue Int. 68 (3) (2001) 156.
rapatite and graphene oxide bone cements, J. Mech. Behav. Biomed. Mater. 82 [39] G. Daculsi, J.M. Bouler, R.Z. LeGeros, Adaptive crystal formation in normal and
(2018 Jun 1) 257e267. pathological calcifications in synthetic calcium phosphate and related bio-
[15] S. Soleymani Eil Bakhtiari, H.R. Bakhsheshi-Rad, S. Karbasi, M. Tavakoli, materials, Int. Rev. Cytol. 172 (1997) 129e191.
M. Razzaghi, A.F. Ismail, S. RamaKrishna, F. Berto, Polymethyl methacrylate-
[40] F.F.M. De Mul, M.H.J. Hottenhuis, P. Bouter, J. Greve, J. Arends, J.J. Ten Bosch,
based bone cements containing carbon nanotubes and graphene oxide: an Micro-Raman line broadening in synthetic carbonated hydroxyapatite, J. Dent.
overview of physical, mechanical, and biological properties, Polymers 12 (7) Res. 65 (3) (1986) 437e440.
(2020 Jun 30) 1469. [41] D.G. Nelson, J.D. Featherstone, Preparation, analysis, and characterization of
[16] R. Fada, N. Farhadi Babadi, R. Azimi, M. Karimian, M. Shahgholi, Mechanical carbonated apatites, Calcif. Tissue Int. 34 (1982) S69eS81.
properties improvement and bone regeneration of calcium phosphate bone
[42] R.Z. LeGeros, J.P. Legeros, Phosphate minerals in human tissues, in: Phosphate
cement, Polymethyl methacrylate and glass ionomer, Journal of Nanoanalysis Minerals, Springer, Berlin, Heidelberg, 1984, pp. 351e385.
8 (1) (2021 Mar 1) 60e79.
[43] H. Catherine, W. Skinner, Low temperature carbonate phosphate materials or
[17] D.Q. Tham, M.D. Huynh, N.T. Linh, D.T. Van, D.V. Cong, N.T. Dung, N.T. Trang, the carbonatedapatite problem: a review, in: Origin, Evolution, and Modern
P.V. Lam, T. Hoang, T.D. Lam, Pmma bone cements modified with silane- Aspects of Biomineralization in Plants and Animals, 1989, pp. 251e264.
treated and pmma-grafted hydroxyapatite nanocrystals: preparation and [44] R.Z. LeGeros, M.S. Tung, Chemical stability of carbonate-and fluoride-con-
characterization, Polymers 13 (22) (2021 Nov 9) 3860. taining apatites, Caries Res. 17 (5) (1983) 419e429.
[18] U. Dubey, S. Kesarwani, P. Kyratsis, R.K. Verma, Development of modified
[45] L. Sun, C.C. Berndt, K.A. Gross, A. Kucuk, Material fundamentals and clinical
polymethyl methacrylate and hydroxyapatite (PMMA/HA) biomaterial com- performance of plasma-sprayed hydroxyapatite coatings: a review, J. Biomed.
posite for orthopaedic products, in: InAdvances in Product Design Engineer- Mater. Res. Appl. Biomater. 58 (2001) 570e592.
ing, Springer, Cham, 2022, pp. 159e178.

15

You might also like