Вы находитесь на странице: 1из 12

Hard Tissues and Materials

Journal of Biomaterials Applications


0(0) 1–12
Preparation and characterization of ! The Author(s) 2018
Reprints and permissions:
titanium—segmented polyurethane sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/0885328218772708
composites for bone tissue engineering journals.sagepub.com/home/jba

Fernando Javier Aguilar-Perez1 , Rossana Vargas-Coronado1,


Jose Manuel Cervantes-Uc1, Juan Valerio Cauich-Rodriguez1,
nez2, Juan Jose Pavon-Palacio†,3,
Raul Rosales-Iba~
Yadir Torres-Hernandez4 and Jose Antonio Rodriguez-Ortiz4

Abstract
Segmented polyurethanes were prepared with polycaprolactone diol as soft segment and 4,4-methylene-bis cyclohexyl
diisocyanate and L-glutamine as the rigid segment. These polyurethanes were filled with 1 wt.% to 5 wt.% titanium
particles (Ti), physicochemically characterized and their biocompatibility assessed using human dental pulp stem cells and
mice osteoblasts. Physicochemical characterization showed that composites retained the properties of the semicrys-
talline polyurethane as they exhibited a glass transition temperature (Tg) between 35 C and 45 C, melting temper-
ature (Tm) at 52 C and crystallinity close to 40% as determined by differential scanning calorimetry. In agreement with
this, X-ray diffraction showed reflections at 21.3 and 23.6 for polycaprolactone diol and reflections at 35.1 , 38.4 , and
40.2 for Ti particles suggesting that these particles are not acting as nucleating sites. The addition of up to 5 wt.% of Ti
reduced both, tensile strength and maximum strain from 1.9 MPa to 1.2 MPa, and from 670% to 172% for pristine and
filled polyurethane, respectively. Although there were differences between composites at low strain rates, no significant
differences in mechanical behavior were observed at higher strain rate where a tensile stress of 8.5 MPa and strain of
223% were observed for 5 wt.% composites. The addition to titanium particles had a beneficial effect on both human
dental pulp stem cells and osteoblasts viability, as it increased with the amount of titanium in composites up to 10 days
of incubation.

Keywords
Segmented polyurethane, titanium particles, bone regeneration, dental pulp stem cells, osteoblasts

Introduction dental pulp stem cells expressed high levels of RunX2


In the quest of improving properties of segmented pol- osteogenic markers. In an effort to improve the
yurethanes (SPUs) for biomedical use, several mechanical and biological properties of glutamine-
approaches have been pursued. Some of them, made
use of biodegradable soft and rigid segment in the PU,
while others introduce an inorganic phase of proved 1
Centro de Investigacion Cientifica de Yucatan, Merida, Yucatan, Mexico
biocompatibility to obtain a composite based on 2
Universidad Nacional Autonoma de Mexico, Facultad de Estudios
PU.1,2 In previous studies, our group prepared SPU Superiores Iztacala, Ciudad de Mexico, Mexico
3
Universidad de Antioquia, Medellin, Colombia
with polycaprolactone diol (PCL) as soft segment and 4
Universidad de Sevilla, Escuela Politecnica Superior, Sevilla, Espa~
na
4,4-methylene-bis cyclohexyl diisocyanate (HMDI) and †
Deceased
L-glutamine (GLU) as the rigid segment and improved
their biocompatibility by the addition of either Corresponding author:
Juan Valerio Cauich-Rodriguez, Centro de Investigacion Cientifica de
hydroxyapatite or silica-based bioactive glasses.3,4 Yucatan A.C., Calle 43 No. 130, Colonia Chuburna de Hidalgo, C.P.
In the first case, alveolar odontoblast were able to 97205, Merida, Yucatan, Mexico.
adhere to these composites while in the second, Email: jvcr@cicy.mx
2 Journal of Biomaterials Applications 0(0)

based PUs, titanium particles (Ti) are used in Materials and methods
this study.
Titanium is a suitable metallic biomaterial for Characterization of titanium particles (Ti)
bone tissue regeneration because of its balance of
The titanium (Ti) particles used in this study were grade
mechanical, physical–chemical, and biofunctional
4 commercially pure (c.p.) Ti powder, synthesized by a
properties.5 It is considered inert in the human body
hydrogenation/dehydrogenation process as reported by
since it does not present either surface corrosion or
the supplier, SE-JONG Materials Co. Ltd. (Seoul,
other reactions in physiological medium. When tita-
Korea). In order to establish the chemical nature, struc-
nium is implanted, good bone healing is observed
since it does not lead to the usual foreign body reac- ture, and morphology of the particles they were charac-
tions, typical of implanted materials such as encapsu- terized by several physical-chemical techniques. Particle
lation or chronic inflammation.6 The titanium surface size was measured by suspending Ti particles in distilled
has the ability to adsorb various proteins such as water using a Beckman Coulter LS100Q laser diffrac-
glycosaminoglycans, fibronectin, fibrinogen, and tion particle size analyzer. Particle morphology was
therefore, some authors suggest the possibility of a observed by scanning electron microscopy (SEM)
titanium–protein complex formation, which enhances using a JEOL JSM 6360LV microscope. Elemental com-
its biocompatibility.7–9 Despite these good properties, position was obtained by energy-dispersive X-ray spec-
improvements in titanium biocompatibility have been troscopy (EDX) by means of INCA X-sight Model
achieved through various types of surface treat- 7582, Oxford Instruments coupled to the SEM. In addi-
ments.10–13 For example, pure titanium does not tion, X-ray photoelectron spectroscopy (XPS) was con-
adhere to the bone, since a biological layer between ducted with a Thermo Scientific K-Alpha X-ray
5 nm and 10 nm thick separating the titanium photoelectron spectrometer (with a monochromatic
from the bone is formed.6 However, when an source of Al Ka with an energy of 1486.6 eV) on both
apatitic layer is deposited on titanium, the bone etched (Ar ions for 30 s) and non-etched samples. X-ray
heals without encapsulation, and allows good bone diffractograms (XRDs) were obtained with a Siemen
interlocking. D5000 diffractometer with radiation CuKa (k ¼ 1.5416
Composites of PU and titanium-based particles have Å), in the 2h range from 5 to 60 , with a step count of 3
been reported with various applications including anti- s and a step size of 0.02 (2h).
microbial coatings and substrates for bone regenera-
tion. Zhang et al.14 describe antimicrobial properties PU synthesis
of composites of PU and titanium, under ultraviolet
radiation. Hieda et al.15 reported to increase shear SPUs were synthesized by two-step polymerization
bonding strength of Ti and SPU interface by modifying method with 1:2.05:1.05 (PCL:HMDI:GLU)
terminal groups with a silane. A study by Pareta et al.16 molar ratios. For this, PCL (molecular weight of
report good adhesion of osteoblasts to titanium struc- 2000 g mol1 from Aldrich) was dissolved in anhy-
tures coated with PU. Similarly, the study by drous dimethylformamide (DMF) (Sigma-Aldrich) and
Sakamoto et al.17 report good adhesion strength of then, an excess of 4,4-methylene-bis ciclohexyl diiso-
PU-coated titanium structures. On the other hand, da cyanate (Aldrich) was slowly added in the presence of
Silva et al.18 reported composites of PU and titania and 0.3% of stannous octoate (Sigma) during 4 h at 60 C.
concluded that the composites showed better mechan- During the second step, a nonessential amino acid,
ical and thermal properties than pristine PU without GLU (Sigma), was added and left to react during 2 h
mentioning their biocompatibility. more. The PU was precipitated in water, and washed
Although good properties are expected for PU/Ti several times before drying at 60 C during 24 h.
composite biomaterials, no reports on their biocompat- The molecular weight of the synthetized pristine
ibility with either HDPSC or osteoblasts have been PU was Mw ¼ 77,649 g mol1, as obtained by gel
found in the literature. Therefore, the aim of this permeation chromatography.
study was to obtain composites made of segmented
PU and titanium particles by a low temperature
Composite preparation by solvent casting
method, and to perform a physical–chemical and
mechanical characterization of them. Furthermore, as Composites with 1, 3 and 5 wt.% of c.p. Ti powder were
there is little knowledge on the suitable scaffold prop- prepared by dispersing the proper amount of titanium
erties to support HDPSC and osteoblast adhesion and particles in chloroform, and then adding this suspension
proliferation, we expect to determine their suitability to a PU chloroform solution. The suspension was soni-
for dental tissue regeneration. cated during 5 min and then poured in a Teflon mold.
Aguilar-Perez et al. 3

Films of the composites were obtained after slow solvent X-ray diffraction (XRD). Crystallinity of the PCL and Ti
evaporation at 25 C for a minimum of 24 h. was observed by XRD using a Siemens D5000 diffrac-
tometer as mentioned above while Xc was calculated
Physical-chemical and mechanical characterization using equation (2):
of PU/Ti composites Ac
%Xc XRD ¼  100 (2)
Fourier-transform infrared spectroscopy (FTIR). FTIR spec- Aa þ Ac
tra were obtained using attenuated total reflectance
(ATR) with ZnSe crystal in the 4000 and 650 cm1 where Aa is the area under the amorphous hump, and
spectral range averaging 100 scans with a resolution Ac is the area under the crystalline peak.
of 4 cm1. For this, a Thermo Scientific Nicolet, 8700
spectrometer was used. Scanning electron microscopy (SEM). Surface morphology
and elemental distribution Ti on the composites was
Raman spectroscopy. Spectra were acquired using an studied with a JEOL, JSM 6360LV with acceleration
inVia Renishaw Raman spectrometer. The system voltage of 20 kV. Microanalysis mapping was con-
was capable of collecting spectra over a Raman shift ducted with EDX (INCA X-sight Model 7582,
spectral range of approximately 3200–100 cm1. A 633 Oxford Instruments) during 300 s.
nm argon laser was used as the excitation radia-
tion source. Tensile mechanical test. Elastic modulus (E), maximum
tensile strength (rmax), tensile strength at break (rbreak),
Differential scanning calorimetry (DSC). Melting point (Tm) deformation at break (ebreak), and strength (roy) and
of the soft segment was determined with a Perkin strain (eoy) to an offset yield at 2% of deformation
Elmer DSC-7, using 5–7 mg of the composite, after were obtained after tensile test according to ASTM
heating from 5 C to 150 C with a heating rate of D882. For this, a Shimadzu AGSX with a cross head
10 C/min under nitrogen atmosphere. In addition, the speed of 250 mm/min was used. Films of 40 mm
long (effective gauge length of 25 mm), 5 mm wide, and
relative percent crystallinity (Xc) of the PCL in the PUs
0.1 mm thickness, were used. Data were compared using
was determined from the enthalpy of fusion using the
one-way analysis of variance (ANOVA) with post hoc
equation (1):
multiple comparison performed using Tukey’s test.
DHf A minimum of n ¼ 5 were used and p < 0.05 was consid-
%Xc DSC ¼  100 (1) ered significant. Additionally, a low cross head speed test
Wss  DH f
(50 mm/min) was performed in order to compare
mechanical behavior of materials at different strain rates.
where DHf is the enthalpy of melting of PUs obtained
experimentally by DSC during the first trace, Wss is the Biological studies
theoretical mass fraction of the flexible segment, and

DH f is the enthalpy of 100% crystalline PCL reported Isolation and characterization of human dental pulp
as 136 J g1.19 stem cells (HDPSCs)
Human dental pulp was obtained, with informed
and written consent, from the third molar of 16–20
Thermogravimetric analysis (TGA). Mass loss was obtained
years old individuals after routine tooth extraction.
with 25 mg of the sample after heating from 50 C to HDPSC were isolated from dental pulp as follows.
650 C at 10 C/min under nitrogen atmosphere by Dental pulp was cut into small pieces, then washed
means of a TGA-7 from Perkin-Elmer. From the with phosphate buffered saline (PBS) (Biowest),
mass loss first derivative, the decomposition tempera- digested with 1% collagenase solution (Sigma-
ture (Td) was obtained. Aldrich) for 30 min at 37 C and centrifuged at 1000
r/min to obtain a cell pellet. Then, cells were resus-
Dynamic mechanical analysis (DMA). Storage modulus (E0 ) pended in low glucose Dulbecco’s modified Eagle
and Tan d were obtained with a Perkin Elmer DMA-7. medium (DMEM) (Biowest), supplemented with 10%
Rectangular samples of 20 mm long, 4 mm wide, and fetal calf serum (Biowest) and 1% penicillin–strepto-
0.1 mm thickness were deformed in the extension mode mycin (Sigma-Aldrich). Cells were cultured in 25 cm2
with a static load of 80 mN and a dynamic load of 65 Nunc flasks (NunclonTM) at 37 C and 5% CO2 in a
mN at frequency of 1 Hz while heated from 100 C to humidified atmosphere. Medium was changed twice a
75 C at 5 C/min. week during expansion. When cells reached 80–90%
4 Journal of Biomaterials Applications 0(0)

confluency, they were separated by magnetic cell sort- OPTIMA (Ortenberg, Germany) at 550 nm. A compar-
ing (MACS), using CD44 micro beads, human ison between treatments at day 1 and 10 was made by
(Miltenyi, Biotec) as described elsewhere.20 The cellular ANOVA, and followed by a Mann–Whitney test, where
populations isolated were identified as mesenchymal p  0.05 was considered significant.
stem cells based on their ability to adhere to the plastic
culture plates, fusiform morphology, and expression of
surface markers like antihuman CD44. HDPSCs from Results and discussion
third passage, seeded at a density of 5  103 per well,
were used for cell viability and proliferation assays. Properties of titanium particles
Ti particle size showed an average particle size of
Isolation of mice osteoblasts. Osteoblasts were isolated
23.5 lm, median of 23.2 lm, and mode of 28.7 lm
from the femurs of five female mice at postnatal fifth
within the range from 1.5 lm to 43.6 lm in agreement
day, after CO2 sacrifices in accordance to NOM-062-
with manufacturer data. By SEM (micrographs not
ZOO-1999 standard under approval by the Committee
for the Use of Experimental Animals at FES-I. Tissue shown), it was observed that Ti particles were of irreg-
was cut in small pieces of approximately of 2 mm2 in ular shape and with sizes in close approximation of
preparation for enzymatic digestion in type I collage- what is reported by manufacturer (average particle
nase (Sigma-Aldrich) for 10 min, and this solution was size of 18 lm 3, max. of 45 lm).
then filtered. After this, DMEM was added and cells XPS analysis (see Figure 1(a)) showed after 30 s
were incubated in a type “t” Falcon flask at 37 C with argon erosion that the amount of titanium (Ti2p at
5% of CO2 and 95% of humidity. Culture medium was 459 eV) was 24.5 at%, 20.1 at% for carbon (C1s at
replaced twice a week until cells reached confluence at 285 eV), and 55.3 at% for oxygen (O1s at 531 eV)
day 5. Finally, cells of the cell passage number 2 were in agreement with Cai et al.22 who reported a similar
harvested and seeded to a density of 5  103 per well for elemental composition of Ti particles by XPS.
each assay. Deconvolution of the peaks for Ti2p (Figure 1(b))
was associated to Ti4þ (90.8%) and Ti2þ (9.2%)
Cell viability. Pristine PU and its composites (PU/1% Ti, while deconvolution of O1s (Figure 1(c)) showed the
PU/3% Ti, PU/5% Ti) were sterilized by ultraviolet presence of O2 (66.7%) and other oxygen containing
exposure for 20 min on each side. The viability species (33.3%). Overall, these results suggest that par-
was assessed with a live/dead assay (Live/Dead, ticles are from titanium with an oxidized surface layer,
InvitrogenTM) according to the manufacturer’s proto- TiO2. The presence of carbon can be explained by par-
col. In this case, calcein AM (0.5 ml) and ethidium bro- ticle contamination or residuals from synthesis method.
mide (2.0 ml) was dissolved in PBS (997.5 ml), then 100
ll of this solution were added to each sample and incu-
bated for 30 min while protected from light at 37  C in Ti distribution in PU/Ti composites
a humidified atmosphere with 5% CO2. The cells were Physical inspection of Ti composites showed that they
counted using a fluorescence microscope (Zeiss become opaque and darker as the concentration of Ti
HXP 120C). increased in the translucent PU. SEM micrographs did
V
R not reveal the distribution and/or location of Ti par-
Cell proliferation. The MTT test (Roche Life Science)
ticles as shown in Figure 2 (top). However, EDX map-
was used to assess cell proliferation in the composites
ping (see Figure 2 (bottom)), revealed that Ti particles
seeded with 5  103 of HDPSCs after 24 h and 10 days
are well dispersed in composites, although rich areas of
of culture. After the incubation time, 100 mL of dimeth-
yl sulfoxide was added to solubilize the formazan salts. Ti were detected, especially at Ti concentrations of
Afterward, the mixture was analyzed in a microplate 5 wt.%.
reader at 570 nm. In addition to the distribution of Ti particles, EDX
Osteoblasts proliferation was assessed by Alamar microanalysis (see Table 1) showed that as the wt.% of
blue assay (Sigma-Aldrich) as this assay does not lead Ti is increased in the composite, this element increased
to cell death.21 For this, 20 mL of rezasurin salt were very close to the expected amount. The maximum con-
diluted to 5.98 mL with DMEM without phenol red. centration of Ti detected at the surface was 0.8 at.% for
Cells were seeded at a density of 1  103 per well, in a 1% composites, while it was 3.1 at.% and 4.4 at.% for
96-well plate and incubated 24 h at 37 C with 5% of 3% and 5% composites, respectively. This composi-
CO2 and 95% humidity. Then, 100 mL of rezasurin solu- tional drift clearly reflects either the heterogeneous sur-
tion were added and incubated for 2 h. Absorbance was face distribution of the Ti particles or its possible
measured using an ELISA BMG LAbtech POLARstar coating by the PU.
Aguilar-Perez et al. 5

(a) 7E+04 Titanium Particles (b) 16000


O1s Ti2p3/2
Surface
30 s etching 14000
6E+04

12000
5E+04
Intensity (Counts/s)

Intensity (Counts/s)
4+
Ti2p 10000 Ti2p 1/2 Ti
4E+04
8000
Satellite 2+
Ti
3E+04 6000
C1s
2E+04 4000

2000
1E+04
0
0E+00
800 600 400 200 474 472 470 468 466 464 462 460 458 456 454 452 450
Binding energy (eV) Binding energy (eV)

(c) O1s
25000

O2-
20000
Intensity (Counts/s)

15000

10000

5000

544 542 540 538 536 534 532 530 528 526
Binding energy (eV)

Figure 1. XPS survey spectra (a) of non-etched Ti particles and 30 s argon etched Ti particles. Deconvoluted peaks of Ti2p (b), and
O1s (c).

Figure 2. SEM micrographs (top) of the cross section of PU/Ti composites, PU/1% Ti (a), PU/3% Ti (b), and PU/5% Ti (c). EDX
mapping for Ti on the corresponding composites (bottom).

Infrared/Raman spectroscopy. FTIR spectra of PU neat stretching vibration of CH2 groups from PCL,
films and composites are shown in Figure 3(a). HMDI and glutamine. The carbonyl stretching
The unfilled polymer showed absorption bands at, (C¼O) appeared in the interval between, 1680 and,
3373 cm1 corresponding to the N–H bond in the ure- 1760 cm1 which includes the ester group from the
thane/urea and at, 2932 cm1 and, 2860 cm1 those PCL, and the urethane group (NHCOO). Amide II
corresponding to the asymmetrical and symmetrical absorption (urethane N–H bending þ C–N stretching)
6 Journal of Biomaterials Applications 0(0)

was located at, 1530 cm1 and, 1240 cm1, while the Thermal properties
peak at, 1165 cm1 was attributed to the C–O–C
DSC thermograms showed that the PU and their com-
stretching vibration in the soft segment.23,24 At
posites are semicrystalline as the exhibited a melting
1636 cm1, urea absorptions were detected in agree-
temperature (Tm) of 52 C, while DMA thermograms
ment with previous works.4,10,25 In composites, infra-
(data not shown) showed a significant drop in the stor-
red absorptions from the PU remain and bands from Ti
age modulus of the composite between 40 C and
were not observed since metal–metal bonds cannot be
20 C, suggesting their alpha transition (Ta) at this
detected by conventional FTIR spectrometers using
temperature range.3,4 As the Tg of PCL is located at
ATR technique or because they were masked due to
60 C, their shift to higher temperatures is suggesting
their low concentration. For pristine titanium only
polymer movement restrictions when incorporated into
adsorbed water (3432 cm1), carbon-containing spe- the PU. However, as this transition was not affected by
cies (2921 and 2852 cm1), and oxidized titanium Ti particles, it is suggested that Ti particles are proba-
(705 cm1) were observed.26 bly located either in the interface between soft and rigid
Raman spectroscopy can be used for the detection segment or in the rigid segment bulk.
of titanium (see Figure 3(b)). It is expected that rutile As the melting temperature of the composites did
phase of TiO2 shows a Raman shift at 141.1, 466.3, not change with Ti addition, this suggests that Ti par-
614.5, and 819.1 cm1, while anatase phase at 141.1, ticles are not acting as nucleating agents (see first DSC
154.4, 381.8, 503.7, 526.9, and 653.6 cm1. In this trace on Figure 4(a)). In agreement with these observa-
study, peaks at 417 and 610 cm1 were detected and tions, XRDs of the composites showed typical PCL
assigned to TiO2, but not detected in composites prob- reflections at 23.6 and 21.3 (see Figure 4(b)).2,3,23 In
ably because of their low concentration. Furthermore, composites, PCL peaks become less intense while the
PU showed strong Raman absorptions at 1108 (C–O), typical peaks of titanium at 35.1 , 38.4 , and 40.2
1305 (CH2 or C–N), 1440 (CH2 or C–N), 2867 (CH2 increased.12 The percentage of crystallinity of PUs, as
sym), and 2919 cm1 (CH2 asym) and medium intensity measured by DSC, was not to be affected by the inclu-
absorption at 1723 cm1 (C¼O in ester or urethane). sion of Ti particles in the composites with values of
40%, 39%, 43%, and 37% for PU pure, PU/1% Ti,
Table 1. Elemental composition (at.%) of PU/Ti composites PU/3% Ti, and PU/5% Ti, respectively. Crystallinity
obtained by energy-dispersive X-ray spectroscopy (EDX). percentage, measured by XRD showed values of 40%,
PU PU/1% Ti PU/3% Ti PU/5% Ti 39%, 42%, and 36%, respectively in agreement with
those obtained by DSC and literature.3
C 66.1  1.3 65.3  0.8 67.7  3.1 67.9  2.3 TGA thermograms of composites showed a main
N 7.6  2.9 8.9  0.3 3.9  2.3 3.9  2.9
decomposition temperature (Td) between 300 C and
O 26.3  0.9 25.0  1.4 25.3  1.4 23.8  2.6
350 C, which was related to degradation of urethane
Ti 00 0.8  0.7 3.1  0.5 4.4  0.3
bonds in the soft and rigid segments.27,28 A second
PU: polyurethane. residual mass drop was observed near 450 C and it

PU PU/1% Ti PU/3% Ti PU/5% Ti Ti


(a) 240 (b) 160000
PU
220
140000
PU/1% Ti
200 PU/3% Ti
180 120000 PU/5% Ti
Transmittance (a.u.)

2919
Ti
160
Intensity (a.u.)

100000
140 417 610

2867
120 80000
100
60000 1108 1440
80 1305 1723

60 40000

40
20000
20

0 0
4000 3500 3000 2500 2000 1500 1000 500 500 1000 1500 2000 2500 3000

Wavenumber (cm–1) Raman shift (cm–1)

Figure 3. Vibrational spectra of PU/Ti composites. IR (a), and Raman (b).


Aguilar-Perez et al. 7

(a) 52°C (b) 3000 21.3°


32 PU
PU
PU/1% Ti
PU/1% Ti
28 PU/3% Ti 2500
PU/3% Ti
PU/5% Ti PU/5% Ti
24 2000
Ti

Intensity (a.u.)
Heat flow (W/g)

20
1500

23.6°
16
40.2°
1000

12
500 38.4°
35.1°
8

0
–20 0 20 40 60 80 100 120 140 160 0 10 20 30 40 50 60
Temperature (°C) 2q

Figure 4. Effect of titanium on crystallinity of composites. DSC thermograms (a) and XRD patterns (b).

(a) (b)
100
6 0

First derivative (residual mass % / °C)


–5
4
80
–10
Residual mass (%)

2
60
–15

0
40 –20
PU
PU/1% Ti 600 650 PU
–25
PU/3% Ti PU/1% Ti
20
PU/5% Ti PU/3% Ti
–30
PU/5% Ti
0
–35

100 200 300 400 500 600 100 200 300 400 500 600
Temperature (°C) Temperature (°C)

Figure 5. TGA thermograms of PU/Ti composites. Residual mass (a) and first derivative of residual mass (b).

Table 2. Properties of polyurethane (PU)/Ti composites obtained by thermal analysis.

DMA DSC TGA

Ta Tm DHf Crystallinity Td1 Td2 Residual



C 
C J g1 % 
C 
C mass %

PU 40 52 40.9 40 351 461 0.9


PU/1% Ti 39 52 39.8 39 346 442 2.1
PU/3% Ti 43 52 43.5 44 344 438 2.5
PU/5% Ti 37 52 37.7 39 346 436 5.9
DSC: differential scanning calorimetry; DMA: dynamic mechanical analysis; TGA: thermogravimetric analysis.

was related to decomposition of carbon chains of rigid maximum temperature used during TGA was 650 C,
segment. Both decomposition temperatures seem not to which is far from the Tm of titanium reported at
be affected by the increase of Ti concentration in con- 1668 C. As observed in the inset of Figure 5(a), the
trast to some reported catalytic activity that accelerates final mass is, as expected, in close agreement with the
decomposition.26 Figure 5 summarizes these findings, original weight composition of Ti particles in compo-
where the residual mass (Figure 5(a)) and first deriva- sites i.e. 2.1% of residual mass for PU/1% Ti, 2.5% for
tive (Figure 5(b)) are plotted against temperature. The PU/3% Ti, and 5.9% of residual mass for PU/5% Ti.
8 Journal of Biomaterials Applications 0(0)

Table 2 summarizes the thermal properties of the property studied. In general, the mechanical properties
PU/Ti composites. of composites of PU/Ti were lower than PU/bioactive
glass composites in spite of their higher rigidity, except
Tensile mechanical properties for the Young’s modulus that was significantly higher at
low strain rates.4 However, the mechanical properties of
Figure 6 shows typical stress–strain curves obtained
these composites are far from those exhibited by trabec-
during the mechanical test of composites at high defor-
ular bone from human femur (17.5  1.12 GPa) or cor-
mation rate, while the corresponding mechanical prop-
tical bone (22.1 GPa) measured by ultrasonic
erties are summarized in Table 3. In general, at low
technique.29
strain rate, tensile strength decreased from 1.9 MPa
to 1.2 MPa with Ti addition suggesting a poor inter-
Biocompatibility of composites
face, while deformation followed the same trend, i.e., a
reduction from 670% for pristine PU to 172% for PU/ Primary cultures can provide information more rele-
5% Ti. However, as shown in Table 3, Young’s mod- vant to clinical outcomes when compared to immortal-
ulus did not show a significant difference between the ized cell lines such as Saos-2 or MG63, and are also
pristine PU and most Ti composites being the highest increasingly used in the biomaterials field with tissue
(19 MPa) when 1% of titanium was used. In agreement engineering applications, and for these reasons were
with this, the storage modulus (E0 ) obtained by DMA chosen for this study.30 Although several reports exist
showed that both PU/1% Ti and PU/3% Ti composites on stem cell/osteoblast biocompatibility of either tita-
showed higher modulus at 25 C that unfilled PU (see nium31–35 or PU,36–39 no studies were found on PU/Ti
Table 3). Results of tensile tests at high deformation composites. Therefore, the potential cytotoxicity of
rates showed a higher Young’s modulus (77–83 MPa), those used in this study was assessed with both
higher tensile strength (7.1–8.5 MPa) but lower strain to HDPSC and osteoblasts.
break (185–223%). Nevertheless, no statistical differen- HDPSC adhesion to PU and composites was good
ces were observed between materials in any mechanical at day 1, as shown by the live/dead assay in Figure 7(a).
This figure also shows that there are a higher number of
9
live cells when the content of Ti was increased.
However, at day 10, there was a significant reduction
8
in cell viability at the surface for any composition.
7 HDPSC proliferation on composites increased from
6 day 1 to day 10 as measured by the MTT assay (see
Figure 8(a)). However, at day 1, there was no difference
Stress (MPa)

5
between pristine PU and Ti composites. However,
4 PU
PU/ 1% Ti
there was a statistical difference between composites
3 PU/ 3% Ti at day 10, increasing with Ti content in composites,
2
PU/ 5% Ti except on PU/3% Ti, which is higher than PU, but
lower than PU/1% Ti.
1
In contrast to HDPSC behavior, osteoblast viability
0 at both 1 and 10 days improved with Ti concentration
0 20 40 60 80 100 120 140 160 180 200 220
(see Figure 7(b)), cells count at 24 h of incubation were
Strain (%)
297, 452, 656, and 480 cells per field, and at 10 days
Figure 6. Tensile stress-strain curves of PU/Ti composites were 28, 255, 370, and 440 cells per field for PU, PU/
obtained at 250 mm/min. 1% Ti, PU/3% Ti, and PU/5% Ti, respectively.

Table 3. Tensile mechanical properties of polyurethane (PU)/Ti composites.


DMA Cross head speed (250 mm/min) Cross head speed (50 mm/min)

E0 25 C E rmax roy rbreak eoy ebreak E rmax roy rbreak eoy ebreak
Material MPa MPa MPa MPa MPa % % MPa MPa MPa MPa % %

PU 112.6 78.5  6.0 8.2  0.9 7.4  0.1 7.2  0.8 11.5  0.3 223  34 17.3  1.2 1.9  0.3 1.3  0.1 1.9  0.3 8.7  0.5 671  77
PU/1% Ti 170.8 77.2  4.1 7.8  0.6 7.4  0.6 7.1  0.6 11.0  0.3 185  56 19.9  3.9 1.8  0.1 1.3  0.2 1.8  0.1 8  1.9 537  40
PU/3% Ti 172.4 83.6  4.4 8.1  0.5 7.9  0.5 7.4  0.4 10.5  0.6 221  53 16.1  2.5 1.7  0.1 1.3  0.1 1.7  0.1 9.3  0.5 499  29
PU/5% Ti 116.1 80.7  3.5 8.5  0.8 6.3  1.1 8.5  0.8 10.2  2.7 223  42 15.0  0.6 1.2  0.1 0.9  0.2 1.2  0.8 6.9  1.2 172  38

DMA: dynamic mechanical analysis.


Aguilar-Perez et al. 9

(a) (b)
24 hours 10 days 24 hours 10 days

PU

PU
1%
Ti

PU
3%
Ti

PU
5%
Ti

Figure 7. Fluorescent microscopy images of (a) HDPSCs at 20 and (b) osteoblasts at 4, after 24 h and 10 days on
PU/Ti composites.

(a) 0.6 (b) PU ∗


∗ PU/1% Ti
PU 0.10 PU/3% Ti
0.5 PU/1% Ti PU/5% Ti
PU/3% Ti
PU/5% Ti 0.08

Absorbance (a.u.)


Absorbance (a.u.)

0.4

0.3
0.06 ∗

0.2 0.04

0.1 0.02

0.0 0.00
24 hours 10 days 24 hours 10 days

Figure 8. HDPSC proliferation on composites by MTT (a) and osteoblast proliferation by Alamar Blue on PU/Ti composites (b).
(*p  0.05).

However, when comparing the same composition after At 24 h, osteoblast proliferation on pristine PU was
10 days, osteoblast number decreased with exception of significantly lower than PU/1% Ti, and in this compos-
PU/5% Ti, which showed similar amount of live cells ite, osteoblast proliferation was significantly higher than
present at both times. PU/3% Ti, while PU/5% Ti was significantly higher
Osteoblasts proliferation increased from day 1 to day than pristine PU and PU/3% Ti. At 10 days, composites
10 as determined by Alamar Blue (see Figure 8(b)). showed an increase in cell proliferation when higher
10 Journal of Biomaterials Applications 0(0)

amounts of Ti were used in the composites, where PU/ no significant difference in tensile mechanical behavior
3% Ti and PU/5% Ti were significantly higher than PU between materials but a higher modulus was achieved.
and PU/1% Ti. This indicates that Ti particles could be acting as stress
It is not clear why HDPSC detached after 10 days concentrators due to their large particle size and that
from the PU/Ti composites while osteoblast remained they have a slight reinforcing effect on PU matrix, espe-
as shown in Figure 7. A possible explanation is that cially when used at 1%. Both, viability and prolifera-
they almost reached confluency after 1 day, leading to tion were good in composites containing 1%, 3%, and
contact inhibition and detaching at longer times, as 5% of Ti up to 10 days, for mice osteoblasts, and
seen in Figure 7(a). This is in contrast with previous HDPSCs, and therefore, biocompatibility tends to
reports where either HDPSC or osteoblast were cul- improve at higher content of Ti particles. Based on
tured, in the absence of a polymer, on laser sintered these results, composites of PU and Ti in the form of
or acid etched titanium alloys and where it was films may be suitable for membrane-guided bone
observed no difference in cell adhesion.31 regeneration.
Surface roughness,32 surface chemistry,33,34 zeta
potential, and surface energy22,30 have been proposed Acknowledgements
as the main factors contributing to osteoblast attach-
The authors want to thank Patricia Quintana and Daniel
ment to titanium implants, but to our knowledge, there
Aguilar Trevi~
no for XRD and XPS experiments at
is no similar study on HDPSC.35 Laboratorio Nacional de Nano y Biomateriales (LANBIO),
Although it has been reported that committed cells Cinvestav-IPN, Unidad Mérida (Projects FOMIX-Yucatan
(human osteoblasts) display better proliferation, distri- 2008–108160 and CONACYT LAB-2009–01 no. 123913).
bution, and adhesion in 2D cultures on roughened tita-
nium surface compared to HDPSC,40 it is possible that Declaration of Conflicting Interests
HDPSC exhibited higher proliferation at shorter times
The author(s) declared no potential conflicts of interest with
either due to the higher proliferative index as many
respect to the research, authorship, and/or publication of
multipotent cells or due to the stimulating effect
this article.
of titanium.
In has been reported that collagen sponges, porous
hydroxyapatite/tricalcium phosphate ceramics, and Funding
titanium meshes support the attachment, growth, and The author(s) disclosed receipt of the following financial sup-
differentiation of human DPSC equally well in vitro port for the research, authorship, and/or publication of this
but with little amounts of hard tissue formed in article: This research was supported by Consejo Nacional de
vivo.41 A similar result was obtained during the Ciencia y Tecnologia through grants 248378, 1360 and
course of this study where HDPSC adhere well at 268595. The authors Y. Torres and J.A. Rodrı́guez-Ortiz
short times to a PU/Ti composite but also it was would like to thanks the support of the Ministry of
found that osteoblast proliferate better in a 2D cell Economy and Competitiveness of the State General
culture model. Overall, these composites were not cyto- Administration of Spain under the grant MAT2015-71284-P.
toxic but depending on the final application the use of
differentiated or undifferentiated cells should be used. ORCID iD
For tissue engineering purposes (3D cultures), HDPSC Fernando Javier Aguilar-Perez http://orcid.org/0000-
should be used due to their higher proliferation rate 0003-4266-0464
while for biomaterials development or screening, 2D
osteoblasts cultures are recommended.
References
1. Guelcher SA. Biodegradable polyurethanes: synthesis
Conclusions and applications in regenerative medicine. Tissue Eng
Part B Rev 2008; 14: 3–17.
Semicrystalline polymeric composites were obtained by
2. Chan-Chan LH, Tkaczyk C, Vargas-Coronado RF, et al.
mixing metallic titanium powder and a segmented PU
Characterization and biocompatibility studies of new
based on glutamine. In these composites, Ti particles
degradable poly(urea)urethanes prepared with arginine,
did not act as nucleating agents as their crystallinity, glycine or aspartic acid as chain extenders. J Mater Sci
demonstrated by DSC and XRD, remained the same. Mater Med 2013; 24: 1733–1744.
Titanium particles, however, had an effect on mechan- 3. Cetina-Diaz SM, Chan-Chan LH, Vargas-Coronado RF,
ical properties. At low strain rates, both tensile strength et al. Physicochemical characterization of segmented pol-
and maximum strain of composites tend to decrease yurethanes prepared with glutamine or ascorbic acid as
when higher Ti concentration while a low Young’s chain extenders and their hydroxyapatite composites.
modulus was observed; at higher strain rates, there is J Mater Chem B 2014; 2: 1966–1976.
Aguilar-Perez et al. 11

4. Aguilar-Pérez FJ, Vargas-Coronado RF, Cervantes-Uc 20. Gronthos S, Brahim J, Li W, et al. Stem cell properties of
JM, et al. Preparation and bioactive properties of nano human dental pulp stem cells. J Dent Res 2002; 81: 531–535.
bioactive glass and segmented polyurethane composites. 21. Gloeckner H, Jonuleit T and Lemke HD. Monitoring of
J Biomater Appl 2016; 30: 1362–1372. cell viability and cell growth in a hollow-fiber bioreactor
5. Steinemann SG. Titanium—the material of choice? by use of the dye Alamar Blue. J Immunol Methods 2001;
Periodontol 2000 1998; 17: 7–21. 252: 131–138.
6. Brunette DM, Tengvall P, Textor M, et al. Titanium in 22. Cai K, Frant M, Bossert J, et al. Surface functionalized
medicine: material science, surface science, engineering, titanium thin films: Zeta-potential, protein adsorption
biological responses and medical applications. Berlin and cell proliferation. Colloids Surf B Biointerfaces
Heidelberg: Springer, 2012, p. 1019. 2006; 50: 1–8.
7. Schliephake H, Aref A, Scharnweber D, et al. Effect of 23. De Oliveira AAR, de Carvalho SM, Leite MDF, et al.
immobilized bone morphogenic protein 2 coating of tita- Development of biodegradable polyurethane and bioac-
nium implants on peri-implant bone formation. Clin Oral tive glass nanoparticles scaffolds for bone tissue engineer-
Implants Res 2005; 16: 563–569. ing applications. J Biomed Mater Res Part Res 2012;
8. Singh R, Lee PD, Jones JR, et al. Hierarchically struc- 100: 1387–1396.
tured titanium foams for tissue scaffold applications. 24. Ryszkowska JL, Auguscik M, Sheikh A, et al.
Acta Biomater 2010; 6: 4596–4604. Biodegradable polyurethane composite scaffolds contain-
9. Jimi E, Hirata S, Osawa K, et al. The current and future ing BioglassVR for bone tissue engineering. Compos Sci
therapies of bone regeneration to repair bone defects. Int Technol 2010; 70: 1894–1908.
J Dent 2012; 2012: 148261. 25. Asefnejad A, Behnamghader A, Khorasani MT, et al.
10. Shi Z, Neoh KG, Kang ET, et al. Surface functionaliza- Polyurethane/fluor-hydroxyapatite nanocomposite scaf-
tion of titanium with carboxymethyl chitosan and immo- folds for bone tissue engineering. Part I: morphological,
bilized bone morphogenetic protein-2 for enhanced physical, and mechanical characterization. Int J
osseointegration. Biomacromolecules 2009; 10: 1603–1611.
Nanomed. 2011; 6: 93–100.
11. Jonasova L, Muller FA, Helebrant A, et al. Biomimetic
26. Avolio R, D’albore M, Guarino V, et al. Pure titanium
apatite formation on chemically treated titanium.
particle loaded nanocomposites: study on the polymer/
Biomaterials 2004; 25: 1187–1194.
filler interface and hMSC biocompatibility. J Mater
12. Wang XX, Yan W, Hayakawa S, et al. Apatite deposi-
Sci: Mater Med 2016; 27: 153.
tion on thermally and anodically oxidized titanium sur-
27. Mohd-Rus AZ, Kemp TJ and Clark AJ. Degradation
faces in a simulated body fluid. Biomaterials 2003;
studies of polyurethanes based on vegetable oils. Part 2.
24: 4631–4637.
Thermal degradation and materials properties. Prog Rct
13. Spoerke ED, Murray NG, Li H, et al. A bioactive tita-
Kin Mech 2009; 34: 1–41.
nium foam scaffold for bone repair. Acta Biomater 2005;
28. Cervantes-Uc JM, Moo-Espinosa JI, Cauich-Rodrıguez
1: 523–533.
JV, et al. TGA/FTIR studies of segmented aliphatic pol-
14. Zhang X, Su H, Zhao Y, et al. Antimicrobial activities of
hydrophilic polyurethane/titanium dioxide complex film yurethanes and their nanocomposites prepared with com-
under visible light irradiation. J Photochem Photobiol A mercial montmorillonites. Polym Degrad Stab 2009;
Chem 2008; 199: 123–129. 94: 1666–1677.
15. Hieda J, et al. Effect of terminal functional groups of 29. Grimal Q, Haupert S, Mitton D, et al. Assessment of
silane layers on adhesive strength between biomedical cortical bone elasticity and strength: mechanical testing
Ti-29Nb-13Ta-4.6Zr alloy and segment polyurethanes. and ultrasound provide complementary data. Med Eng
Surf Coat Technol 2012; 206: 3137–3141. Phys 2009; 31: 1140–1147.
16. Pareta RA, Reising AB, Miller T, et al. An understanding 30. Hempel U, Hefti T, Dieter P,et al. Response of human
of enhanced osteoblast adhesion on various nanostruc- bone marrow stromal cells, MG-63, and SaOS-2 to
tured polymeric and metallic materials prepared by ionic titanium-based dental implant surfaces with different
plasma deposition. J Biomed Mater Res Part A 2010; 92 topography and surface energy. Clin Oral Impl Res
(3): 1190–1201. 2013; 24: 174–182.
17. Sakamoto H, Doi H, Kobayashi E, et al. Structure and 31. Mangano C, De Rosa A, Desiderio V, et al. The osteo-
strength at the bonding interface of a titanium-segmented blastic differentiation of dental pulp stem cells and bone
polyurethane composite through 3-(trimethoxysilyl) formation on different titanium surface textures.
propyl methacrylate for artificial organs. J Biomed Biomaterials 2010; 31: 3543–3551.
Mater Res Part Res 2007; 82: 52–61. 32. Shen X, Ma P, Hu Y, et al. Mesenchymal stem
18. Demétrio da Silva V, dos Santos LM, Subda SM, et al. cell growth behavior on micro/nano hierarchical surfaces
Synthesis and characterization of polyurethane/titanium of titanium substrates. Colloids Surf B 2015;
dioxide nanocomposites obtained by in situ polymeriza- 127: 221–232.
tion. Polym Bull 2013; 70: 1819–1833. 33. Do Nacimento RM, De Carvalho VR, Govone JS, et al.
19. Jiang S, Ji X, An L, et al. Crystallization behavior of PCL Effects of negatively and positively charged Ti metal sur-
in hybrid confined environment. Polymer 2001; faces on ceramic coating adhesion and cell response.
42: 3901–3907. J Mater Sci: Mater Med 2017; 28: 33.
12 Journal of Biomaterials Applications 0(0)

34. Felgueiras HP, Decambron A, Manassero M, et al. Bone hyperbranched polyurethane for bone tissue regenera-
tissue response induced by bioactive polymer functional- tion. Macromol Biosci 2017; 17: 1600271.
ized Ti6Al4V surfaces: in vitro and in vivo study. Colloid 39. Xing J, Ma Y, Lin M, et al. Stretching-induced nano-
Interface Sci 2017; 491: 44–54. structures on shape memory polyurethane films and
35. Yavari SA, Van der Stok J, Chai YC, et al. Bone regen- their regulation to osteoblasts morphology. Colloids
eration performance of surface-treated porous titanium. Surf B Biointerfaces 2016; 146: 431–441.
Biomaterials 2014; 35: 6172–6181. 40. Palumbo C, Baldini A, Cavani F, et al.
36. Wang L, Li Y, Zuo Y, et al. Porous bioactive scaffold of Immunocytochemical and structural comparative study
aliphatic polyurethane and hydroxyapatite for tissue of committed versus multipotent stem cells cultured
regeneration. Biomed Mater 2009; 4: 25003. with different biomaterials. Micron 2013; 47: 1–9.
37. Guo R, Lu S, Merkel AR, et al. Substrate modulus reg- 41. Zhang W, Walboomers XF, Van Kuppevelt TH, et al.
ulates osteogenic differentiation of rat mesenchymal stem The performance of human dental pulp stem cells on dif-
cells through integrin b1 and BMP receptor type IA. ferent three-dimensional scaffold materials. Biomaterials
J Mater Chem B 2016; 4: 3584–3593. 2006; 27: 5658–5668.
38. Gogoi S, Maji S, Mishra D, et al. Nano-bio engineered
carbon dot-peptide functionalized water dispersible

Вам также может понравиться