Biodegradation Profile of Polylactide Composites For Bone Tissue Engineering

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

ARAB JOURNAL OF BASIC AND APPLIED SCIENCES University of Bahrain

2021, VOL. 28, NO. 1, 351–359


https://doi.org/10.1080/25765299.2021.1971865

ORIGINAL ARTICLE

Biodegradation profiles of chitin, chitosan and titanium reinforced


polylactide biocomposites as scaffolds in bone tissue engineering
Abraham K. Aworindea, Oluwaseyi O. Taiwob, Samson O. Adeosunb, Esther T. Akinlabic,
Hassana Jonathand, Odunlami A. Olayemie and Olufunmilayo O. Josepha
a
Mechanical Engineering Department, College of Engineering, Covenant University, Ota, Nigeria; bMetallurgical and Materials
Engineering Department, University of Lagos, Lagos, Nigeria; cPan African University for Life and Earth Sciences Institute
(PAULESI), Ibadan, Nigeria; dChemistry Department, College of Science and Technology, Covenant University, Ota, Nigeria;
e
Chemical Engineering Department, College of Engineering, Covenant University, Ota, Nigeria

ABSTRACT ARTICLE HISTORY


In this study, chitin (Ct), chitosan (Ch) and titanium powder (Ti) were melt-blended with pol- Received 5 May 2021
ylactide (PLA) to produce three composites of PLA. The samples were subjected to multiple Revised 17 July 2021
characterisations, which includes immersion in a phosphate buffer solution (PBS) for ten Accepted 19 August 2021
weeks. The results showed that the three fillers reduced the hydrophobicity of PLA, making
KEYWORDS
the reinforced samples to be more hydrophilic. In addition, the organic reinforced PLA com- Bio-composites scaffold;
posites became more amorphous while the PLA/Ti samples became more crystalline when hydrolytic degradation;
compared with unreinforced PLA. The results of the samples’ immersion in PBS showed that bone internal fixation;
the organic reinforced PLA presented a biodegradation profile that would allow for the hydrophobicity effect;
required gradual transfer of load to the partially healed fractured bone. Except for the chitin formation of apatite
reinforced PLA (PLA/Ct) with 16.67 wt. %, all other composites produced using the organic
fillers gave a swelling percentage which was less than 17%. However, titanium-reinforced
PLA (PLA/Ti) composites were found to continue to swell after ten weeks of the immersion
of the samples. Although all the samples formed apatite on their surfaces (except the unre-
inforced PLA), the biodegradation profile of PLA/Ti was deemed unsuitable as an implant
because of the possibility of postoperative palpability.

1. Introduction osteoconductive surface (Wypych, 2018), which pre-


vents the fixation’s instability over a long period
One of the criteria that limit metallic implants in
(Kokubo et al., 2016). Although a study argues that
osteosynthesis is the problem of biodegradation.
Adequate biodegradation rate has been identified as AFA of a material may not be a predictor of bioactiv-
the solution to stress shielding or off-loading. ity (Pan, Zhao, Darvell, & Lu, 2010), the formation of
Gradual healing of fractured bone requires a balance apatite has been established as a part of the
of load between the partially healed bone and par- implant-bone bonding process (Kazek-KeR sik et al.,
tially biodegraded implant (Chandra & Pandey, 2020; Kokubo et al., 2016).
2020). Biodegradability is also in demand in areas One of the most prominent biopolymers on the
such as drug delivery systems, degradable sutures, list of biodegradable polymers as fixations is the
surgical void fillers, etc., (Lyu & Untereker, 2009). Polylactide (PLA). PLA is a biodegradable thermoplas-
Some biopolymers have been listed as being suitable tic that has been copiously processed for biodegrad-
in these areas, especially as scaffolds for bone frac- able applications (Akpan et al., 2019; Aworinde,
ture fixation (Aworinde, Adeosun, Oyawale, Akinlabi, Adeosun, & Oyawale, 2020a; Aworinde, Adeosun,
& Emagbetere, 2018). Apart from biodegradability, a Oyawale, Akinlabi, & Akinlabi, 2019; Aworinde,
biopolymer that would be used as a bone implant is Kehinde, Emagbetere, Adeosun, & Akinlabi, 2021;
also expected to possess an apatite-forming ability Gbenebor et al., 2018a; 2018b; Wang, Wang, Ito,
(AFA) through which the implant would be able to Zhang, & Chen, 2016). Its biodegradation products
bond to the living bone (Kazek-KeR sik et al., 2020; (CO2 and H2O) are entirely non-toxic, and its biore-
Kokubo & Yamaguchi, 2016; Sadeghzade, Emadi, sorbability neither overburdens the body absorptive
Tavangarian, & Doostmohammadi, 2020). The forma- capacity nor creates foreign material in the body.
tion of an apatitic layer on an implant creates an However, its biodegradation profile is quite slow as a

CONTACT Odunlami A. Olayemi olayemi.odunlami@covenantuniversity.edu.ng Chemical Engineering Department, College of Engineering,


P.M.B 1023, Ota, Ogun State, Nigeria
ß 2021 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group on behalf of the University of Bahrain.
This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits
unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
352 A. K. AWORINDE ET AL.

result of its degree of hydrophobicity (Aworinde, Table 1. The weight per cent of the fillers and initial
Adeosun, Oyawale, Akinlabi, & Akinlabi, 2020b; masses of the solid cylinders.
Salazar-Sanchez, Campo-Erazo, Samuel Villada- Initial mass (g)

Castillo, & Fernando Solanilla-Duque, 2019). Attempts Filler (wt. %) PLA/Ch PLA/Ct PLA/Ti
have been made to improve the hydrophobicity of 0.00 1.2473 1.2473 1.2473
1.04 1.2094 1.3011 1.1973
PLA by ion implantation (Kurzina et al., 2020), melt 2.08 1.0173 1.3290 1.1813
blending (Aworinde et al., 2020b), gamma irradiation 4.17 1.1857 1.1184 1.2092
8.33 1.3278 1.3618 1.1859
(Qi et al., 2019), copolymerisation (Hendrick & Frey, 16.67 1.2272 1.4951 1.3791
2014) and so on.
The biodegradation of PLA is expected to be 2.2. Methods
controlled whenever it is used as an implant for
bone fracture healing. Biodegradation is neither 2.2.1. Composites development
The matrix was melt-blended with each of the fillers
supposed to be excessively accelerated nor unduly
at the weight percentages reported in our previous
slow. Fractured bone, especially with operative
work (Aworinde et al., 2020c). Each of the molten
management, typically requires patients to be non-
composites was mould-pressed at the pouring tem-
weight bearing while healing is in progress (Jones
perature of 290  C to form solid cylinders with
& Waterson, 2020). The minimal weight that should
12.5 mm diameter and 7.0 mm length. The cylinders
be borne by bone during fracture healing is usu-
were weighed, and their masses recorded, as shown
ally transferred to the implants. The load is grad-
in Table 1.
ually transferred back to the bone as healing
begins to take place so as to avoid stress shielding
2.2.2. Morphological examinations
or off-loading. The dynamics of load transfer imply
A Tescan Vega 3LMH scanning electron microscope
that the implant is expected to lose its mechanical
was used for the surface morphological examinations
strength after a period of time. For this to happen, of the reinforcements and the developed compo-
the implant loses mass (i.e., after it has swollen as sites. An accelerating voltage of 20 kV was applied to
a result of fluid intake) and then degrade. all the samples. However, the beam intensity of
Biodegradation commences at the exterior of the 12 W/m2 was applied to titanium powder, while
bio-implant and progresses to its interior (Azevedo 17 W/m2 beam intensity was applied to the rest of
& Reis, 2004). These processes were studied in this the samples. Apart from the titanium powder, all
work by immersing three melt-blended composites other samples were first carbon coated on copper
of PLA in phosphate buffer solution (PBS). Two grids with Agar Turbo Carbon Coater to enhance the
composites had organic fillers, namely chitin and quality of the images generated.
chitosan, while the third had an inorganic (titan-
ium) powder additive. The effects of these fillers 2.2.3. Fourier Transform-Infrared Spectroscopy
on the hydrophobicity of PLA had been reported (FTIR) test on the solid composites
in our previous work (Aworinde et al., 2020b), the The solid composite samples were crushed to get
biodegradation profiles of the composites devel- powder-like samples suitable for the FTIR test. Each
oped using these fillers were compared in sample in its powder form was placed in the FTIR
this study. spectrometer and scanned. Acetone and tissue paper
were used to clean the probe of the machine at the
end of each test sample to avoid contamination.
Spectra were recorded at 32 scans with a resolution
of 2 cm1. The transmittance measurements were
2. Materials and methods carried out in the range of 400 cm1 to 4000 cm1.
MATLAB R2019a (9.6) was used to further analyse
2.1. Materials
the spectra.
PLA with a monomeric molecular weight of 144 g/
mol, having an overall lactide purity of 99.5% was 2.2.4. X-Ray diffraction test
used in this study as the matrix. Chitin (Ct) and chi- XRD was carried out to determine the crystalline
tosan (Ch) served as organic fillers and were separ- nature of the composites and measure the percent-
ately used to reinforce PLA, while titanium (Ti) age crystallinity using Equation (1) (Tufekci, Brantley,
powder (Ti-6Al-2Sn-2Mo-2Cr-0.25Si) was the inor- Mitchell, Foreman, & Georgette, 1999). The XRD ana-
ganic additive. The maximum particle size of the lysis of the composites was carried out using Rigaku
organic fillers was 75 mm. The sizes of the particles of Miniflex 600 powder diffractometer equipped with
Ti powder were 45  90 mm. Cu Ka radiation source generated at 18 kW and
ARAB JOURNAL OF BASIC AND APPLIED SCIENCES 353

Figure 1. SEM micrographs of the additives. (a) Chitosan, (b) chitin, and (c) titanium.

250 mA. The XRD spectra were obtained at room in mass after weeks of immersion, MD, was calcu-
temperature. High angle measurement was per- lated for every sample using Equation (1).

formed at a range of 2h ¼ 10  90 at a step rate of Mf  Mi
0.40/min. The spectra obtained were processed using MD ¼  100% (2)
Mi
OriginLab 2019b (9.75).
where
Ac
%Crystallinity ¼  100 (1) MD ¼ percentagechangeinmass,
AT
Mi ¼ initialmassbeforeimmersion, and
Ac is the area of crystalline peaks, while AT is the Mf ¼ finalmassafterimmersion
area of all peaks (that is, the area of crystalline peaks
and amorphous peaks).

2.2.5. Biodegradation test 3. Results and discussion


In a beaker (1 L), phosphate buffer solution (PBS) 3.1. Micrographs of fillers from scanning
was prepared by mixing sodium phosphate dibasic electron microscopy
heptahydrate (20.209 g) and sodium phosphate
monobasic monohydrate (3.394 g) in distilled water Figure 1 shows the micrographs of the fillers
(800 mL). Each of the solutes was allowed to dissolve obtained from the scanning electron microscope
before adding the other. A few drops of sodium test. Visual inspection reveals the flaky structure of
hydroxide (NaOH) were added to adjust the final pH Ch (Figure 1a), the sheet-like property of Ct (Figure
to 7.4 (that is, the pH of human blood). Distilled 1b) and the hard sphere of Ti (Figure 1c). The micro-
water was thereafter added until the volume was graphs of Ch and Ct are different from those
1 L. Test tubes were each filled with 20 mL of PBS. obtained in some studies because morphological
The test tubes containing PBS were put in the structures of crustacean shells-based derivations dif-
racks and kept in an oven with a preset temperature fer with species (Gbenebor, Adeosun, Lawal, & Jun,
of 36.5  C. The test tubes were left in the oven for 2016) and type of chemical treatment (Gbenebor,
about 30 minutes to ensure the conditioning of the Philips Akpan, & Adeosun, 2017). The flaky structure
PBS to 36.5  C (Kokubo & Takadama, 2006). The of Ch presents an increased surface area for reaction.
weighed samples were then immersed in 20 mL of The morphology of Ch is very close to the sheet-like
PBS. The test tubes were returned to the oven, and structure of Ct. The structure of Ct also contributes
the temperature maintained at 36.5  C. to an increased rate of reaction, although not as
The biodegradation test was left on for ten (10) much as it would be obtained in Ch. However, the
weeks. Changes in mass, which is considered as the spherical structure of Ti presents a reduced surface
progress of biodegradation (Azevedo et al., 2004), area leading to a much lower rate of reaction than
were measured after the first four (4) weeks and at the flaky Ch and the sheety Ct (Ollila, 1997). The dif-
the end of the tenth week. The percentage change ference in the fillers’ shapes is possibly one of the
354 A. K. AWORINDE ET AL.

Figure 2. SEM Micrographs of PLA and its composites (a) PLA, (b) PLA/Ch, (c) PLA/Ct, and (d) PLA/Ti.

reasons for the varying degree of the biodegradation al., 2018). The least hydrophilic unreinforced PLA had
of their respective composites. the least ability to form apatite (Figure 6(a)).

3.2. Micrographs from scanning electron 3.4. Percentage crystallinity from XRD
microscopy test of PLA and its composites
The percentages of crystallinity of the developed
In Figure 2, the surface morphologies of PLA and its composites were calculated from the XRD data ana-
composites are presented. Similar surface morph- lysed using OriginLab 2019b by Equation (1) and
ology of PLA composites have been reported graphically presented in Figure 4 using MATLAB
(Murariu et al., 2010). Since biodegradation usually R2019a (9.6). The percentages of crystallinity of unre-
starts from the external surface of the bio-implant inforced PLA, PLA/Ch, PLA/Ch and PLA/Ti (Table 2)
and proceeds to its inner part (Azevedo & Reis, were 13.36%, 0%, 11.03% and 41.23%, respectively.
2004), the cracks on the surfaces of unreinforced The percentage of crystallinity of PLA obtained here
PLA (Figure 2(a)), PLA/Ch (Figure 2(b)) and PLA/Ct must have been due to the application of heat
(Figure 2(c)) seemed to be a catalyst to biodegrad- (Vadas, Nagy, Csontos, Marosi, & Bocz, 2020) via the
ation. These cracks were not conspicuous on the sur- fabrication method employed in the development of
face of PLA/Ti. These surface cracks made these the composite, considering the fact that no peaks
composites more percolative, hence the possible were seen when the solvent volatilising method was
gradual filtering of PBS through the surface cracks. used (Chu, Zhao, Li, Fan, & Qin, 2017).
Additionally, the rough surfaces of PLA, PLA/Ch and The addition of chitosan gave an approximately
PLA/Ct have the potentials to support cell adhesion 100% amorphous composite, while chitin loading
and possibly cell proliferation more than PLA/Ti. resulted in  11.03% crystalline composite (Figure 4
and Table 2). A trend similar to this was reported
when PLA was loaded with 5% chitin (Nasrin et al.,
3.3. FTIR spectra of PLA and its composites
2017). Besides, the reduction in crystallinity of PLA
Figure 3 shows the spectra of PLA and the three appears to be the norm whenever it is reinforced
developed composites. As discussed in our previous with biomaterials (Nizamuddin et al., 2019). However,
work, apart from the peaks at 1450 and 1360 cm1 reinforcing PLA with titanium gave the PLA compos-
other peaks were indicative of hydrophilic contents ite with the highest percentage crystallinity (Table 2).
(Aworinde et al., 2020b). We had shown that PLA/Ch The increase in the percentage of crystallinity seems
was the most hydrophilic sample. The study also to be the characteristic behaviour of PLA whenever
showed that the unreinforced PLA was the least it is loaded with inorganic filler (Chu et al., 2017).
hydrophilic sample. The effect of hydrophilicity was The implication of the amorphous nature of the
pronounced in the AFA of the samples, hydrophil- developed composites is the fast degradation profile
icity being one of the crucial factors for AFA (Liu et since the denser structure of a crystalline material
ARAB JOURNAL OF BASIC AND APPLIED SCIENCES 355

Figure 3. FTIR spectra of PLA and its composites.

Figure 4. XRD Results of PLA and Its Composites.

could make it, to some extent, impervious to fluid in Table 3. Figure 5 illustrates the quantitative meas-
attack (Pantani & Sorrentino, 2013; Tokiwa & Calabia, ure of the mass gained as a result of fluid absorption
2006). Hence, the most amorphous sample satisfac- (especially during the first four weeks) or lost in
torily responded to fluid imbibition within the weeks mass as seen after ten weeks of immersion. At the
of immersion. PLA/Ti, the most crystalline sample, end of the first four (4) weeks, it was observed that
presented a somewhat erratic swelling index. all the composites responded to fluid uptake by a
noticeable change in mass. They all absorb the fluid
into which they were immersed, causing them to
3.5. Fluid imbibition properties of
increase in mass. The organic reinforced composites
the composites
were on the lead. This fluid absorption property (or
Table 3 shows the masses of the samples after 4 and hydrophilicity) is an essential initial stage in the bio-
10 weeks of immersion of PLA and its composites in degradation of any biopolymer and biopolymer com-
PBS due to fluid uptake. Figure 1 displays the per- posites (Adeosun, Aworinde, Diwe, & Olaleye, 2016;
centage change in mass after 4 and 10 weeks of Azevedo et al., 2004; Huang, Xiong, Liu, Zhu, &
immersion in PBS. It details information on the fluid Wang, 2013). Generally, fluid absorption is important
intake properties of the solids. where the application requires the material to be in
The biodegradation (in vitro) results of the immer- contact with a fluid, such as in tissue engineering
sion of all the composites produced in the phos- and drug delivery systems (Adeosun et al., 2016;
phate buffer solution for ten (10) weeks are shown Azevedo et al., 2004). The fluid absorption property,
356 A. K. AWORINDE ET AL.

Figure 5. Fluid imbibition propensity of PLA and its composites.

Figure 6. Formation of apatite on (a) unreinforced PLA and (b) reinforced PLA.

Table 2. The degree of crystallinity and amorphousness of


as displayed by these composites, is an indication of PLA and its composites.
the possibility of biodegradation of the composites. Sample  % Crystallinity  % Amorphousness
In Figure 1, the percentage change in mass dur- PLA 13.36 86.64
ing biodegradation, using Equation 1, is presented. PLA/Ch 0.00 100.00
PLA/Ct 11.03 88.97
As noted earlier, both the unreinforced PLA and all PLA/Ti 41.23 58.77
reinforced PLA absorbed the PBS in which they were
immersed. This fluid uptake is a sign of degradation
(Azevedo et al., 2004). after the tenth week of immersion exceeded their
During the first four weeks of immersion, the masses after the fourth week of immersion. PLA/Ct
composites with organic fillers demonstrated a (16.67 wt. %), for instance, may not be proper as an
higher rate of fluid absorption than the titanium- implant because of the excessive increase in mass.
reinforced PLA. This may be due to the nature of the Disproportionate change in mass can lead to bulki-
fillers (Figure 1(a) and (b)) and the surface cracks ness, thereby causing postoperative palpability
(Figure 2(a)–(c)) on the organic reinforced samples. (Leonhardt et al., 2008) and create an absorption
PLA/Ch showed a higher rate of fluid absorption capacity problem for the patient.
than PLA/Ct. This can be attributed to the flakier Similarly, all the PLA/Ti composites continued to
structure of chitosan (Figure 1(a)) than chitin struc- increase in mass after four weeks. They all show the
ture (Figure 1b). PLA/Ti also absorbed fluid to a rea- tendency to create the same problem that was
sonable extent. Its fluid absorption property must observed in PLA/Ct at 16.67 wt% of the filler, namely:
have been due to the onset of hydrolytic degrad- bulkiness and postoperative palpability. Another
ation (Azevedo et al., 2004). problem envisaged with PLA/Ti is the possible health
The mass changes observed after ten (10) weeks issue that can be raised by particulate titanium
of immersion showed that some of the composites either close to the implant site or in some organs
exhibit excessive swelling propensity. Their masses away from the implants (Frisken, Dandie, Lugowski,
ARAB JOURNAL OF BASIC AND APPLIED SCIENCES 357

Table 3. Changes in the masses of composites during biodegradation test.


Mass (g) after 4 weeks Mass (g) after 10 weeks
Filler (wt. %) PLA/Ch PLA/Ct PLA/Ti PLA/Ch PLA/Ct PLA/Ti
0.00 1.3935 1.3935 1.3935 1.3251 1.3251 1.3251
1.04 1.3607 1.4064 1.2164 1.2928 1.4163 1.2776
2.08 1.1878 1.4636 1.2716 1.1709 1.411 1.3653
4.17 1.2473 1.2592 1.2991 1.2585 1.2248 1.3696
8.33 1.4026 1.4878 1.2553 1.4511 1.4790 1.3485
16.67 1.3857 1.6863 1.4607 1.3722 1.8330 1.5073

& Jordan, 2002) as the composites degrade. Titanium implant. Findings from this study showed that unre-
particles are not expected to biodegrade; these par- inforced PLA does not combine these two important
ticles will eventually accumulate in the body and factors. Although the biodegradation of PLA was not
possibly affect the body organs. outrageous, the formation of apatitic layer on its sur-
All PLA/Ch composites and all PLA/Ct composites face was unduly scanty. The fillers used influenced
except PLA/Ct at 16.67 wt% appear to give satisfac- biodegradation and apatite formation in varying
tory results in terms of biodegradation. Variation in degrees. Organic reinforced PLA samples (i.e., PLA/
mass has been a valid sign of biodegradation Ch and PLA/Ct), for instance, increased the rate of
(Rudnik, 2013). The mass variation pattern of all PLA/ biodegradation and apatite formation while the inor-
Ti and PLA/Ct at 16.67 wt%, however, is worrisome. ganic reinforced PLA samples (i.e., PLA/Ti) enhanced
There was a proportionate change in the percentage apatite formation but put biodegradation at risk as a
in mass of other composites, especially after the first result of their erratic hydrolytic behaviours.
four weeks, which was not the case with all PLA/Ti This study equally identified hydrophilicity as an
and PLA/Ct at 16.67 wt. %. essential factor in the formation of apatitic layer on
the surface of a scaffold. Therefore, the reinforced
3.6. Formation of apatitic layer on the surface PLA that showed reduction in hydrophobicity also
of the composites showed the formation of apatite more than the
unreinforced PLA. In the same vein, biodegradation
Figure 2 presents the physical changes observed was linked to the degree of crystallinity of an
when the solids were immersed in PBS. It was implant. Hence, the more amorphous samples
observed that at the end of the fourth week, a thin showed a more promising biodegradation profile.
film of white deposits confirming the formation of Findings showed that organic reinforced PLA would
apatite (Jose & Alagar, 2015; Ramadas, El Mabrouk, & make better scaffolds in term of biodegradability
Ballamurugan, 2020; Swe, Shariff, Noor, Ishikawa, & than the PLA composites produced using inorganic
Mohamad, 2019) was seen. These white deposits filler. Besides biodegradation, postoperative palpabil-
continued to grow until it completely covered the ity seemed inevitable with the use of PLA/Ti compo-
entire surfaces of the reinforced composites (Figure sites due to their pattern of swelling within the
2b). The neat PLA did not show much of apatite for- duration of in vitro study. AFA of a material is, there-
mation as recorded in reinforced PLA. The white fore, not sufficient to qualify a biodegradable poly-
deposits could be scarcely seen on the unreinforced mer as an ideal scaffold.
PLA, as seen in Figure 2a. Studies have established
that a material that forms apatite on its surface has
the tendency to bond to living bone through the Disclosure statement
layer of the apatite formed on its surface in the liv- No potential conflict of interest was reported by
ing body (Kokubo et al., 2006). This implies that the the author(s).
fillers used in this work aided the formation of an
apatitic layer on the composites. By inference, the
Funding
reinforced composites possessed AFA needed to
facilitate the bonding of an implant with the liv- This work was supported by the Covenant University,
Ota, Nigeria.
ing bone.

4. Conclusion
References
This study focused on the two critical factors that
Adeosun, S. O., Aworinde, A. K., Diwe, I. V., & Olaleye, S. A.
dictate the quality of an ideal scaffold. These factors (2016). Mechanical and microstructural characteristics of
(i.e., biodegradability and apatite formation) have rice husk reinforced polylactide nanocomposite. The
been identified as strong indications of a good West Indian Journal of Engineering, 39(2), 63–71.
358 A. K. AWORINDE ET AL.

Akpan, E. I., Gbenebor, O. P., Igogori, E. A., Aworinde, A. K., Gbenebor, O. P., Atoba, R. A., Akpan, E. I., Aworinde, A. K.,
Adeosun, S. O., & Olaleye, S. A. (2019). Electrospun por- Adeosun, S. O., & Olaleye, S. A. (2018a). Study on poly-
ous bio-fibre mat based on polylactide/natural fibre par- lactide-coconut fibre for biomedical applications. In The
ticles. Arab Journal of Basic and Applied Sciences, 26(1), Minerals, Metals & Materials Society (Eds.), Minerals,
225–235. doi:10.1080/25765299.2019.1607995 Metals and Materials Series (pp. 263–273). Switzerland:
Aworinde, A. K., Adeosun, S. O., & Oyawale, F. A. (2020a). Springer.
Mechanical properties of poly (L-lactide)-based compo- Gbenebor, O. P., Akpan, E. I., Atoba, R. A., Adeosun, S. O.,
sites for hard tissue repairs. International Journal of Olaleye, S. A., Taiwo, O. O., Igogori, E. A., Alamu, O. B.,
Innovative Technology and Exploring Engineering, 9(5), Aworinde, A. K. (2018b). Development and performance
2152–2155. analysis of high voltage generator for electrospinning of
Aworinde, A. K., Adeosun, S. O., Oyawale, F. A., Akinlabi, nano fibres. Unilag Journal of Medicine, Science and
E. T., & Akinlabi, S. A. (2019). The Strength characteristics Technology, 6(2), 45–58.
of chitosan- and titanium- poly (L-lactic) acid based Hendrick, E., & Frey, M. (2014). Increasing surface hydrophil-
composites. Journal of Physics: Conference Series, 1378(2), icity in poly(lactic acid) electrospun fibers by addition of
022061. Pla-B-Peg co-polymers. Journal of Engineered Fibers and
Aworinde, A. K., Adeosun, S. O., Oyawale, F. A., Akinlabi, Fabrics, 9(2), doi:10.1177/155892501400900219
E. T., & Akinlabi, S. A. (2020b). Comparative effects of Huang, J., Xiong, J., Liu, J., Zhu, W., & Wang, D. (2013).
organic and inorganic bio-fillers on the hydrophobicity Investigation of the in vitro degradation of a novel poly-
of polylactic acid. Results in Engineering, 5, lactide/nanohydroxyapatite composite for artificial bone.
100098–100093. doi:10.1016/j.rineng.2020.100098 Journal of Nanomaterials, 2013, 1–10.
Aworinde, A. K., Adeosun, S. O., Oyawale, F. A., Akinlabi, Jones, M. S., & Waterson, B. (2020). Principles of manage-
E. T., & Emagbetere, E. (2018). Mechanical strength and ment of long bone fractures and fracture healing.
biocompatibility properties of materials for bone Surgery (Oxford), 38(2), 91–99. doi:10.1016/j.mpsur.2019.
internal fixation: A brief overview. Proceedings of the 12.010
International Conference on Industrial Engineering and Jose, A. J., & Alagar, M. (2015). Preparation and character-
ization of polysulfone-based nanocomposites. In Vikas
Operations Management, October 28 – November 1,
Mittal (Ed.), Manufacturing of nanocomposites with engin-
2018, Pretoria, South Africa, pp. 2115–2126.
eering plastics (pp. 31–59).
Aworinde, A. K., Adeosun, S. O., Oyawale, F. A.,
Kazek-KeR sik, A., Nosol, A., Płonka, J., Smiga-Matuszowicz,
Emagbetere, E., Ishola, F. A., Olatunji, O., … Akinlabi,
M., Student, S., Brzychczy-Włoch, M., … Simka, W.
E. T. (2020c). Comprehensive data on the mechanical
(2020). Physico-chemical and biological evaluation of
properties and biodegradation profile of polylactide
doxycycline loaded into hybrid oxide-polymer layer on
composites developed for hard tissue repairs. Data in
Ti-Mo alloy. Bioactive Materials, 5(3), 553–563. doi:10.
Brief, 32, 106107. doi:10.1016/j.dib.2020.106107
1016/j.bioactmat.2020.04.009
Aworinde, A., Kehinde, Emagbetere, E., Adeosun, S. O., &
Kokubo, T., & Takadama, H. (2006). How useful is SBF in
Akinlabi, E. T. (2021). Polylactide and its composites on
predicting in vivo bone bioactivity? Biomaterials, 27(15),
various scales of hardness. Pertanika J. Sci. & Technol,
2907–2915. doi:10.1016/j.biomaterials.2006.01.017
29(2), 1–10. Kokubo, T., & Yamaguchi, S. (2016). Biomineralization of
Azevedo, H., & Reis, R. (2004). Understanding the enzym-
metals using chemical and heat treatments. In Conrado
atic degradation of biodegradable polymers and strat- Aparicio & Maria-Pau Ginebra (Ed.), Biomineralization and
egies to control their degradation rate. In Reis, R. L., & biomaterials: Fundamentals and applications. Woodhead
Roman, J. S. (Eds.), Biodegradable systems in tissue engin- Publishing.
eering and regenerative medicine (pp. 177–202). Taylor & Kurzina, I.A., Laput, O.A., Zuza, D.A., Vasenina, I.V.,
Francis. Salvadori, M.C., Savkin, K.P., … Kalashnikov, M.P. (2020).
Chandra, G., & Pandey, A. (2020). Biodegradable bone Surface & coatings technology surface property modifi-
implants in orthopedic applications: A review. cation of biocompatible material based on polylactic
Biocybernetics and Biomedical Engineering, 40(2), acid by ion implantation. Surface and Coatings
596–610. doi:10.1016/j.bbe.2020.02.003 Technology, 388, 125529. doi:10.1016/j.surfcoat.2020.
Chu, Z., Zhao, T., Li, L., Fan, J., & Qin, Y. (2017). 125529
Characterization of antimicrobial poly (lactic acid)/nano- Leonhardt, H., Demmrich, A., Mueller, A., Mai, R., Loukota,
composite films with silver and zinc oxide nanoparticles. R., & Eckelt, U. (2008). INION V R compared with titanium
Materials, 10(6)doi:10.3390/ma10060659 osteosynthesis: A prospective investigation of the treat-
Frisken, K. W., Dandie, G. W., Lugowski, S., & Jordan, G. ment of mandibular fractures. British Journal of Oral and
(2002). A study of titanium release into body organs fol- Maxillofacial Surgery, 46(8), 631–634. doi:10.1016/j.bjoms.
lowing the insertion of single threaded screw implants 2008.04.021
into the mandibles of sheep. Australian Dental Journal, Liu, C., Li, Y., Wang, J., Liu, C., Liu, W., & Jian, X. (2018).
47(3), 214–217. doi:10.1111/j.1834-7819.2002.tb00331.x Improving hydrophilicity and inducing bone-like apatite
Gbenebor, O., Philips Akpan, E. I., & Adeosun, O. S. (2017). formation on PPBES by polydopamine coating for bio-
Thermal, structural and acetylation behavior of snail and medical application. Molecules, 23(7), 1–11.
periwinkle shells chitin. Progress in Biomaterials, 6(3), Lyu, S. P., & Untereker, D. (2009). Degradability of polymers
97–111. doi:10.1007/s40204-017-0070-1 for implantable biomedical devices. International Journal
Gbenebor, O. P., Adeosun, S. O., Lawal, G. I., & Jun, S. of Molecular Sciences, 10(9), 4033–4065. doi:10.3390/
(2016). Role of CaCO3 in the physicochemical properties ijms10094033
of crustacean-sourced structural polysaccharides. Murariu, M., Dechief, A. L., Bonnaud, L., Paint, Y., Gallos, A.,
Materials Chemistry and Physics, 184, 203–209. doi:10. Fontaine, G., … Dubois, P. (2010). The production and
1016/j.matchemphys.2016.09.043 properties of polylactide composites filled with
ARAB JOURNAL OF BASIC AND APPLIED SCIENCES 359

expanded graphite. Polymer Degradation and Stability, Sadeghzade, S., Emadi, R., Tavangarian, F., &
95(5), 889–900. doi:10.1016/j.polymdegradstab.2009.12. Doostmohammadi, A. (2020). In vitro evaluation of diop-
019 side/baghdadite bioceramic scaffolds modified by poly-
Nasrin, R., Biswas, S., Rashid, T. U., Afrin, S., Jahan, R. A., caprolactone fumarate polymer coating. Materials
Haque, P., & Rahman, M. M. (2017). Preparation of Science & Engineering. C, Materials for Biological
Chitin-PLA laminated composite for implantable applica- Applications, 106, 110176. doi:10.1016/j.msec.2019.
tion. Bioactive Materials, 2(4), 199–207. doi:10.1016/j.bio- 110176
actmat.2017.09.003 Salazar-Sanchez, M. d R., Campo-Erazo, S. D., Samuel
Nizamuddin, S., Jadhav, A., Qureshi, S. S., Baloch, H. A., Villada-Castillo, H., & Fernando Solanilla-Duque, J. (2019).
Siddiqui, M. T. H., Mubarak, N. M., … Ahamed, M. I. Structural changes of cassava starch and polylactic acid
(2019). Synthesis and characterization of polylactide/rice films submitted to biodegradation process. International
husk hydrochar composite. Scientific Reports, 9(1), Journal of Biological Macromolecules, 129, 442–447. doi:
5445–5411. doi:10.1038/s41598-019-41960-1 10.1016/j.ijbiomac.2019.01.187
Ollila, R. G. (1997). Shape makes a difference. American Swe, T. T., Shariff, K. A., Noor, A. F. M., Ishikawa, K., &
Ceramic Society Bulletin, 76, 77–78. Mohamad, H. (2019). Effect of silver on the apatite for-
Pan, H., Zhao, X., Darvell, B. W., & Lu, W. W. (2010). mation of bioactive silicate glass. Materials Today:
Apatite-formation ability-predictor of "bioactivity"?. Acta Proceedings, 17, 884–888.
Biomaterialia, 6(11), 4181–4188. doi:10.1016/j.actbio. Tokiwa, Y., & Calabia, B. P. (2006). Biodegradability and bio-
2010.05.013 degradation of poly(lactide). Applied Microbiology and
Pantani, R., & Sorrentino, A. (2013). Influence of crystallinity Biotechnology, 72(2), 244–251. doi:10.1007/s00253-006-
on the biodegradation rate of injection-moulded poly(- 0488-1
lactic acid) samples in controlled composting conditions. Tufekci, E., Brantley, W. A., Mitchell, J. C., Foreman, D. W., &
Polymer Degradation and Stability, 98(5), 1089–1096. doi: Georgette, F. S. (1999). Crystallographic characteristics of
10.1016/j.polymdegradstab.2013.01.005 plasma-sprayed calcium phosphate coatings on Ti-6Al-
Qi, Y., Ma, H. L., Du, Z. H., Yang, B., Wu, J., Wang, R., & 4V. The International Journal of Oral & Maxillofacial
Zhang, X. Q. (2019). Hydrophilic and antibacterial modi- Implants, 14(5), 661–672.
fication of poly(lactic acid) films by c-ray irradiation. ACS Vadas, D., Nagy, Z. K., Csontos, I., Marosi, G., & Bocz, K.
Omega, 4(25), 21439–21445. doi:10.1021/acsomega. (2020). Effects of thermal annealing and solvent-induced
9b03132 crystallization on the structure and properties of poly(-
Ramadas, M., El Mabrouk, K., & Ballamurugan, A. M. (2020). lactic acid) microfibres produced by high-speed electro-
Apatite derived three dimensional (3D) porous scaffolds spinning. Journal of Thermal Analysis and Calorimetry,
for tissue engineering applications. Materials Chemistry 142(2), 581–594. doi:10.1007/s10973-019-09191-8
and Physics, 242(1–7), 122456. doi:10.1016/j.matchem- Wang, Y., Ito, Y., Zhang, P., & Chen, X. (2016). A compara-
phys.2019.122456 tive study on the in vivo degradation of poly(L-lactide)
Rudnik, E. (2013). Biodegradability testing of compostable based composite implants for bone fracture fixation.
polymer materials. In Sina Ebnesajjad (Ed.), Handbook of Scientific Report, 6, 1–12.
Biopolymers and biodegradable plastics (pp. 213–263). Wypych, G. (2018). Functional fillers – applications. In
William Andrew. George Wypych (Ed.), Functional Fillers (pp. 153–179).

You might also like