You are on page 1of 17

Materials Science and Engineering C 71 (2017) 1175–1191

Contents lists available at ScienceDirect

Materials Science and Engineering C

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

Review

Biodegradable ceramic-polymer composites for biomedical applications:


A review
Michal Dziadek a,⁎, Ewa Stodolak-Zych b,⁎, Katarzyna Cholewa-Kowalska a,⁎
a
AGH University of Science and Technology, Faculty of Materials Science and Ceramics, Department of Glass Technology and Amorphous Coatings, 30 Mickiewicza Ave., 30-059 Krakow, Poland
b
AGH University of Science and Technology, Faculty of Materials Science and Ceramics, Department of Biomaterials, 30 Mickiewicza Ave., 30-059 Krakow, Poland

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

Article history: The present work focuses on the state-of-the-art of biodegradable ceramic-polymer composites with particular
Received 2 August 2016 emphasis on influence of various types of ceramic fillers on properties of the composites. First, the general
Received in revised form 18 September 2016 needs to create composite materials for medical applications are briefly introduced. Second, various types of
Accepted 13 October 2016
polymeric materials used as matrices of ceramic-containing composites and their properties are reviewed.
Available online 14 October 2016
Third, silica nanocomposites and their material as well as biological characteristics are presented. Fourth, differ-
Keywords:
ent types of glass fillers including silicate, borate and phosphate glasses and their effect on a number of properties
Silica of the composites are described. Fifth, wollastonite as a composite modifier and its effect on composite character-
Bioglasses istics are discussed. Sixth, composites containing calcium phosphate ceramics, namely hydroxyapatite, tricalcium
Wollastonite phosphate and biphasic calcium phosphate are presented. Finally, general possibilities for control of properties of
Calcium phosphate ceramics composite materials are highlighted.
© 2016 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1176
2. Biodegradable polymer matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1176
3. Silica based composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1176
4. Bioglass based composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1178
4.1. Silicate bioactive glasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1178
4.2. Borate and borosilicate bioactive glasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1180
4.3. Phosphate glasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1180
5. Wollastonite based composites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1181
6. Calcium phosphate ceramics based composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1185

Abbreviations: ABTS•+, 2,2′-azinobis(3-ethylbenzothiazoline-6-sulfonic acid) radical cation; ACP, amorphous calcium phosphate; ALP, alkaline phosphatase; BBG, borate-based glass;
BCP, biphasic calcium phosphate; BET SSA, Brunauer-Emmett-Teller specific surface area; BG, bioactive glass/bioglass; BMP-2, bone morphogenetic protein 2; BSP, bone sialoprotein; CAM,
cylinder chorioallantoic membrane; CaP, calcium phosphate; CBF-alpha-1, core-binding factor alpha-1; CMC, carboxymethylcellulose; CS, chitosan; DEX, dexamethasone; DPPH•, 2,2-
diphenyl-1-picrylhydrazyl radical; DSC, differential scanning calorimetry; ECM, extracellular matrix; EDX, energy dispersive X-ray spectroscopy; FN, fibronectin; FTIR, Fourier
transform infrared spectroscopy; FTIR-ATR, attenuated total reflectance Fourier transform infrared spectroscopy; gHAp, grafted hydroxyapatite; HAp, hydroxyapatite; hAT-MSCs,
human adipose tissue-derived mesenchymal stem cells; hBMSCs, human bone marrow mesenchymal stem cells; HCA, carbonated hydroxyapatite; HOC, human osteoblastic cells;
MSCs, mesenchymal stem cells; MSNs, mesoporous silica nanoparticles; mWS, mesoporous wollastonite particles; nHAp, nano-sized hydroxyapatite; NMR, nuclear magnetic
resonance; OC, osteocalcin; ON, osteonectin; OPN, osteopontin; Osx, osteoblast-specific transcription factor Osterix; PBG, phosphate glass; PBS, phosphate buffer saline; PBSu,
poly(butylene succinate); PCL, poly(ε-caprolactone); PDGF, platelet-derived growth factor; PDLLA, poly(D,L-lactide); PEG, polyethylene glycol; PEI, polyethylenimine; PEO,
poly(ethylene oxide); PGA, poly(glycolic acid); PHB, poly(3-hydroxybutyrate); PHBV, poly(3-hydroxybutyrate-co-3-hydroxyvalerate); PIXE, proton-induced X-ray emission; PLA,
poly(lactic acid); PLDLA, poly(L-lactide-co-D,L-lactide); PLGA, poly(lactic-co-glycolic acid); PLLA, poly(L-lactide); PRP, platelet-rich plasma; rBMSCs, rat bone marrow mesenchymal
stem cells; RT-PCR, reverse transcription polymerase chain reaction; RUNX2, runt related transcription factor 2; SBF, simulated body fluid; SBG, silicate bioactive glass; SCPL, solvent-
casting particulate leaching; SF, silk fibroin; TCP, tricalcium phosphate; TGF, transforming growth factor; TIPS, thermally induced phase separation; VEGF, vascular endothelial growth
factor; WAXS, wide-angle X-ray scattering; wHAp, hydroxyapatite whiskers; WS, wollastonite; XRD, X-ray diffraction.
⁎ Corresponding authors.
E-mail addresses: dziadek@agh.edu.pl (M. Dziadek), stodolak@agh.edu.pl (E. Stodolak-Zych), cholewa@agh.edu.pl (K. Cholewa-Kowalska).

http://dx.doi.org/10.1016/j.msec.2016.10.014
0928-4931/© 2016 Elsevier B.V. All rights reserved.
1176 M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191

6.1. Hydroxyapatite based composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186


6.2. Tricalcium phosphate based composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186
6.3. Biphasic calcium phosphate based composites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1187
7. Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1187
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1188
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1188

1. Introduction and overall hydrophilicity [1]. Synthetic polymers possess relatively


good mechanical strength and their properties (e.g. porosity, shape)
On account of rapid development of novel biomedical technologies, can be tailored for specific applications. However, the surfaces of syn-
including tissue engineering, regenerative medicine, gene therapy and thetic polymers are hydrophobic and lacking in cell-recognition se-
controlled drug delivery, new materials are being developed to meet quences [1–2].
specific requirements of these fields. Conventional single-component In addition to the biodegradability, the composites take advantage of
ceramic or polymer materials cannot satisfy them. In addition, in order high formability of polymer matrices. Many methods have been devel-
to fully meet the basic requirements such as biocompatibility, biode- oped to fabricate composite materials for biomedical applications, in-
gradability, appropriate mechanical properties, there is a need to obtain cluding solvent casting [4–6], tape casting [7], particulate leaching [8],
materials fulfilling several advanced functions at once. For example, a freeze drying [9], phase separation [10], thermal processing [11−12],
multifunctional material for bone tissue regeneration should induce for- gas foaming [13], electrospinning [14] and rapid prototyping [15].
mation of new bone tissue without an addition of organic bone growth These methods allow obtaining the required properties, especially the
factors (e.g. BMP-2), degrade progressively at a rate matching the re- form and microstructure of the composites.
generation of new bone, induce new blood vessels formation and exhib-
it antibacterial and anti-inflammatory activity. Therefore, key material
and biological features can be achieved by design and development of 3. Silica based composites
multi-component materials, including selection of matrix and modifier
materials, their parameters (e.g. shape, distribution, content), as well Reports on silica (silicon dioxide, SiO2) as a filler or a nanofiller of
as fabrication techniques of composites. In particular, this work deals polymer matrix composites showed many advantages of composite or
with the introduction of ceramic modifiers into biodegradable polymer nanocomposite materials based on biodegradable polymers [16–55].
matrices to obtain composites with specific properties for biomedical Silica particles, incorporated into polymer matrices, cause higher bio-
applications, especially tissue engineering and regenerative medicine. compatibility and bioactivity of the materials and/or implants. These
The most widely used ceramic modifiers including bioactive glasses compositions stimulate biological properties interesting for bone tissue
and calcium phosphates, as well as less widespread silica and wollas- applications, such as given bioresorption rate and porosity, as well as
tonite, with focus on their beneficial effects on material and biological the ability to induce formation of calcium phosphate similar to the
properties of the composites will be discussed. one present in bone on biomaterials surfaces, and the introduction of bi-
A table placed at the end of the review (Table 3) provides an over- ologically active agents [16–18]. Since silica reinforcements are fre-
view of polymer matrices, ceramic modifiers, methods of fabrication quently used in the form of nanoparticles, understanding of
and forms of composites discussed in this work, as well as their physical mechanisms of dissolution and metabolism of these particles in the or-
properties. ganisms have also deserved considerable attention [17]. Nanometric
particles of SiO2 in polyesters provide many advantages compared to
other nanofillers and exhibit many properties associated with an ideal
2. Biodegradable polymer matrices material for grafting and scaffolding [19]. Due to silanol groups (Si-
OH) present on the surface of silica, covalent bonds can be formed be-
Many types of biodegradable polymeric materials have already been tween macromolecular chains and the fillers [20]. During the process
used as matrices of ceramic-modified composites for tissue engineering of nanocomposite synthesis, silane coupling agents play important
applications. These materials can be classified into two major groups roles in connecting the interfaces of organic (polymer chain) and inor-
based on their origin, namely natural-based polymers, including pro- ganic phases (nanofillers, SiO2 particles). It is because they can be func-
teins (soy, collagen, fibrin gels, silk) or polysaccharides (starch, alginate, tionalized at the interface to create a chemical bridge between the
chitin/chitosan, hyaluronic acid derivatives) and synthetic polymers, reinforcement and the polymer matrix, and thus improve the stability,
such as poly(lactic acid) (PLA), poly(glycolic acid) (PGA), poly(ε- adhesion and mechanical properties such as strength, Young's modulus
caprolactone) (PCL), poly(3-hydroxybutyrate) (PHB) [1–2]. or wear resistance of the nanocomposites [21–24].
Generally, biodegradation of polymeric biomaterials involves cleav- Some investigations on nuclear magnetic resonance spectra (13C
age of enzymatically (natural polymers) or hydrolytically (synthetic NMR) of nanocomposite materials i.e.; PBSu/SiO2 (poly(butylene succi-
polymers) sensitive bonds in the polymer structure leading to the poly- nate)/SiO2) confirmed a reaction between the surface silanol groups
mer erosion [3]. Naturally-derived polymers possess the ability to bio- from SiO2 nanoparticles with hydroxyl end groups of the polymer
logical recognition, including presentation of receptor-binding ligands, (PBSu) leading to formation of covalent bonds [20]. These covalent
that may support cell adhesion, migration, differentiation and prolifera- bonds resulted in a substantial improvement of mechanical properties
tion. However, the rate of their in vivo degradation depends on avail- of PBSu, even in cases where amounts of SiO2 lower than 2.5 wt.%
ability and concentration of the enzymes at the site of implantation, were used [20,25]. Nano-SiO2 in different forms i.e.; nanotubes, nano-
therefore it is difficult to predict. The use of natural polymers alone is particles present in the polymer matrix changes also physicochemical
often restricted because of potential immunogenic reactions, possibility properties of the nanocomposite surface e.g.; some of the hydroxyl
of disease transmission and relatively poor mechanical properties [1–3]. groups are exposed on the surface and increase hydrophilic character
Synthetic polymers have more predictable properties including degra- of the surface of materials. This phenomenon can explain a higher hy-
dation kinetics that can be controlled by chemical composition and con- drolysis rate of nanocomposites based on biodegradable polymers
figurational structure, molecular weight, polydispersity, crystallinity, [26]. A similar accelerating effect of SiO2 nanoparticles on a PLA
material morphology (e.g. porosity, surface area), chain orientation hydrolysis rate was reported by many authors [27–30]. They found
M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191 1177

Fig. 1. Durability of poly(L-lactide-co-D,L-lactide)/SiO2 (PLDLA/SiO2) membrane after 8-week incubation in H2O/37 °C/PBS.

that PLA/SiO2 nanocomposites degrade faster than the pure PLA, which citrate used, which enhanced degradation of the matrix, thus leading
indicate that incorporation of silica enhances biodegradation of PLA in to an increased release rate. A new method of carrying drugs into a dam-
the nanocomposites. The addition of silica may lead to facilitated attack aged tissue is an application of core-shell structures. Formation of firm
of water and enzyme molecules on ester groups of the PLA chains, be- bonds between silica and polymer (PEG, PEI, PLGA) on the silica core
cause of high hydrophilicity of silica. The effect of degradation was improves controlled release of a drug or other molecules i.e. molecules
much more visible (Fig. 1) in case of porous nanocomposite membrane that can regulate biological behavior of osteoblasts/osteoclasts [46–47].
after 8-week in vitro degradation [30]. This experiment also showed a Silica nanoparticles with a drug incorporated inside their pores
decreasing molecular weight of the polymer and disintegration of the could be used as fillers of a polymeric matrix used to prepare a scaffold
materials. or a membrane, which provided an additional degree of control of drug
The incorporation of silica nanoparticles into the polymer matrix release, independent from the polymer material itself [48]. Additionally,
stimulated osteoblast-like cells interaction with natural tissue after con- scaffolds based on mesoporous silica characterized by macro-, meso- as
tact with the surface of material i.e. cell viability increased since silicon well as micropores were identified as important contributors to bone
(at critical concentrations) is able to stimulate proliferation of osteo- formation. The micropores may serve as nucleation sites for mineraliza-
blast-like MG-63 cells [31]. Silicon can also be involved in bone forma- tion, causing local supersaturation and subsequent nucleation of apatite
tion and mineralization [32], whereas orthosilicate acid (Si(OH)4) at forms, the mesopores can be used as reservoirs of nutrients and growth
physiological concentration of 10 μmol was shown to stimulate forma- factors or carry therapeutic agents, and macropores can promote cells
tion of type I collagen in human osteoblastic cells (HOC) and to stimu- infiltration. Such implants could by fabricated by traditional techniques
late cell differentiation [33]. Even though no significant effects of SiO2 such as; salt leaching, freeze drying or new methods like
inclusion on cells behavior could be detected, the results were in agree- electrospinning or rapid prototyping. Polymer scaffolds modified by
ment with reports from the literature showing silica materials used for mesoporous silica nanoparticles (MSNs) were often called multifunc-
bone regeneration. Si-OH groups can induce apatite formation and thus tional scaffolds [48–50]. Porous 2D or 3D structures fabricated from a
strong bone bonding [34]. Furthermore, silica nanotube meshes were polymer modified with nanometric silica using the electrospinning
shown to support the adhesion of murine osteoblast-like MC3T3-E1 technique can be used to obtain a superhydrophobic layer on the im-
cells and expression of alkaline phosphatase (ALP), even without plant surface [51]. This method enables control of nanoroughness of
BMP-2 supplements [35]. the surface by the presence of polymer fibers with nanometric or
Mesoporous silica characterized with regular, parallel pores of 2 to submicrometric diameters containing silica nanoparticles which in
50 nm diameter, high specific surface area, low density, high biocom- turn influence wettability and surface free energy (Table 1).
patibility shows capacity for encapsulation of drugs and biological CS fibers were successfully electrospun with poly(ethylene
agents [36]. These particles are widely used in pharmaceutical applica- oxide) (PEO) as a co-blending polymer and with nanometric silica
tions, in most cases in drug delivery systems. In order to control drug re- (CS(PEO)/SiO2) by Suzuki and Mizusima [52]. These organic-inor-
lease from the mesoporous silica, it was mainly mixed with ganic hybrid biomaterials earlier proved to increase oxygen perme-
biodegradable polymers such as; poly(lactic-co-glycolic acid) (PLGA), ability, biocompatibility and biodegradability. Other works showed
polyethylene glycol (PEG), PCL, chitosan (CS), alginate, gelatin, how to control a process of manufacturing of CS(PEO)/SiO2 nanofi-
polyethylenimine (PEI) etc. [37–40]. Drugs such as vancomicine, bers and the ability of these hybrid nanofibers to favor cell attach-
gentamicine, fluoxetine, lidocaine, morphine, nifedipine, paracetamol, ment of murine osteoblast-like 7F2 cells. A hybrid fibrous mesh has
tetracyclin, ibuprofen etc. were used in these applications [41–43]. The a capability to be modified by incorporation of calcium ions resulting
drug could be incorporated in the silica xerogel during the sol–gel pro- in formation of bioactive carbonated hydroxyapatite (HCA) crystals
cess instead of a adsorption method. In some cases the drug can be in-
corporated into the polymer matrix as well, which may have an effect
on the release properties of the composite [39]. Ahola et al. showed
Table 1
that increase of an initial drug loading increased an amount of
Diameters of the PCL and the PCL/SiO2 nanocomposite fibers with physicochemical prop-
toremifene citrate released from PCL matrix [40]. The release rate was erties of the layer.
found to be directly proportional to the load of toremifene citrate. The
Material Fiber diameter Water contact angle Surface free energy
drug release kinetics was controlled by the slowest process i.e. the re-
(nm) (°) [mN/mm]
lease from the polymer matrix which was a diffusion controlled process
[44–45]. Deviations from linearity were observed after 70% of the drug PCL 152 ± 19 151 44.8
PCL/SiO2 178 ± 53 113 34.7
had been released. This may be due to a large amount of toremifene
1178 M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191

[53]. The incorporation of silica particles into natural hydrogels such resorbable phosphate glasses and their application in tissue
as CS or alginate creates another possibilities for cells e.g.; bone-like engineering.
cells such as human osteoblast-like Saos-2 cells and RAW 264.7
mouse macrophage cells differentiated into osteoclast-like cells 4.1. Silicate bioactive glasses
were embedded in silica-containing Na-alginate-based hydrogel.
During the test biocompatible hydrogels with Saos-2 cells retained Silicate bioactive glasses (SBGs) are the most extensively researched
their capacity to synthesize the hydroxyapatite (HAp) crystals [54] glasses for biomedical applications and also the most frequently consid-
and increasingly express the gene encoding for osteoprotegrin [55]. ered as fillers of polymer-based composites for bone tissue engineering.
In another study when CS was used as a matrix a significantly in- The first melt-derived silicate SBG of the quaternary SiO2–Na2O–CaO–
creased protein adsorption and improved apatite deposition were P2O5 system – Bioglass® 45S5, synthesized by Hench et al., is widely
observed by the addition of nano-silica into the composite scaffold reported as a modifier of biodegradable synthetic polymers and copoly-
[54]. mers: PLA [65–66], PGA [67], PCL [4,8,10,68], PHB [69], PLGA [65,70–71],
To conclude, nanosilica present in the polymer matrix composite as well as natural polymers: collagen [72] and CS [73]. Moreover, other
works not only as a filler enhancing its biological properties and SBGs based on both binary SiO2–CaO [74] and ternary SiO2–CaO–P2O5
guaranteeing increase of its biomineralization capability but it also in- [4–6,8,71,75] systems are also used for modifying a polymer matrix.
creases stiffness of the polymer composite without decreasing its me- Bioglass® 45S5 and other SBGs exhibit an ability to bond to soft and
chanical strength. hard tissues which is attributed to the formation of a HCA layer on the
glass surface in contact with the body fluids [57–58]. Apart from the bio-
activity, it was observed that ionic dissolution products (e.g. Si, Ca, P)
4. Bioglass based composites from some of BGs stimulate expression of several genes of osteoblastic
[76] and stem cells [71,77] involved in bone growth, show ability to
Bioactive glasses (BGs, bioglasses) are very attractive materials for stimulate angiogenesis [78], while possibly exhibit antibacterial [79]
modifying a polymer matrix of composites dedicated for tissue engi- and anti-inflammatory actions [80]. Furthermore, release of calcium
neering applications. One of the key reason that makes these materials (Clotting factor IV) being considered a reason for SBG haemostatic prop-
relevant composite modifiers is the possibility of controlling a wide erties [81]. All these specific properties and advantages of BGs can be
range of biological and chemical properties. The structure and chemistry conferred to a polymer matrix by the glass filler incorporation.
of glasses can be tailored at a molecular level by varying composition, a Probably the most important reason for developing polymer-glass
preparation method (melt-quenching or sol-gel), and manufacturing composites for bone tissue engineering is the possibility of achieving a
conditions [1,56–57]. It provides a simple way to control properties of bioactive properties of the polymer matrix [1]. The degree of bioactivity
glass-containing composites. of a composite can be controlled mainly by the chemical composition [8,
BGs are commonly produced by a melt-quench route or a sol-gel 71,82], volume fraction [9–11,65–66,83], size [5,11,68–69,73] and dis-
method [56,58]. Gel-derived BGs usually exhibit improved bioactivity tribution [5,68–69] of a bioactive phase. It was shown that the rate
and cellular response related to a larger volume fraction of mesopores and intensity of apatite-like layer formation on the surface of PCL and
[59–61], a larger concentration of silanols on the surface [59–60] and PLGA-based composites containing SBG particles increased with an in-
also larger specific surface area [58,60–61] than the melt-derived creasing CaO:SiO2 molar ratio of the glass [8,71] (Fig. 2.). Other results
glasses. Furthermore, an advantage of the sol-gel process is the indicated that lower P2O5 content in SiO2–CaO–P2O5 SBG nanoparticles
ability to control the phase composition, structure, texture, micro- enhanced the bioactivity of (poly(L-lactide)-based (PLLA) scaffolds [82].
structure and surface chemistry of the glasses by the synthesis condi- Furthermore, an increased volume fraction [9–11,65–66,83] and higher
tions (e.g. type of a catalyst) and thermal treatment [61–62]. The sol– surface area to volume ratio of the glass modifiers (e.g. the incorpora-
gel process allows obtaining mesoporous micro- and nanoparticles, tion of nano-sized particles instead of microparticles) [5,11,68–69,73]
simply by changing pH of the process, which can be directly used improved the composite bioactivity.
as composite modifiers [61]. To improve or induce desired biological The incorporation of SBG particles in a biodegradable polymer ma-
response and adjust the surface reactivity, as well as solubility in trix can be advantageous as it imparts osteoinductivity to the composite.
biological environment, BGs can be doped with trace elements and It was demonstrated that the addition of gel-derived SBG particles into a
other therapeutic oxides [17,58,63,64]. The research works mainly PLGA matrix promoted the expression of early osteogenesis-related
focus on developing silicate and borate/borosilicate BGs, as well as transcription factors (runt related transcription factor 2 (RUNX2) and

Fig. 2. In vitro bioactivity of PCL/SBG films after 7-day incubation in simulated body fluid (SBF). No morphological changes on pure PCL film (a) after immersion in SBF are observed. On the
surface of a composite film containing silica-rich bioglass (S2 (mol%): 80SiO2–16CaO–4P2O5) (b) clusters of calcium phosphate precipitates occur, while the surface of a film with calcium-
rich glass (A2 (mol%): 40SiO2–54CaO–6P2O5) (c) is fully covered with a thick, continuous layer of HAp.
M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191 1179

Fig. 3. The real-time reverse transcription polymerase chain reaction (RT-PCR) analyses of RUNX2 and Osx mRNA levels in; (a) hBMSCs cultured for two days on PLGA, PLGA/S2 and PLGA/
A2 without osteogenic inducers and BMP-2, osteopontin and collagen I mRNA levels (b-d) in hBMSCs cultured for ten days without osteogenic inducers (NONE) or in the presence of
recombinant human BMP-2 or dexamethasone (DEX). Fold changes in mRNA levels (mean ± SD) are expressed relative to PLGA [77]. © IOP Publishing.
Reproduced with permission. All rights reserved.

osteoblast-specific transcription factor Osterix (Osx)) and osteogenic size of SBG are, the more rapid degradation rate of the poly(ε-
genes (BMP-2, osteopontin (OPN), osteocalcin (OC) and collagen I) in caprolactone-co-D,L-lactide)/SBG [12] and the higher water absorption
human bone marrow mesenchymal stem cells (hBMSCs). As with bioac- of the PHB/SBG composites are [69]. On the other hand, a delayed deg-
tive properties, osteogenic potential depends on chemical composition radation rate in the composite can be achieved by the dissolution of al-
of glass modifiers [77] (Fig. 3.) and their content in composites [71]. kaline ions (e.g. Ca2+, Na+) from the bioactive glass structure. It results
The experiment during which rat mesenchymal stem cells (MSCs) in neutralisation of the released acidic by-products of the polymer deg-
were cultured in the presence of PLGA/SBG porous scaffolds (but phys- radation and thus prevents its autocatalytic degradation process [83].
ically separated from them) showed that the dissolution products from Furthermore, a buffering effect of the glass modifiers can be considered
the materials stimulated osteogenesis alone [84]. Apart from the effect as an advantage, because it helps to avoid a possible inflammatory re-
of solution-mediated factors, a three-dimensional structure and surface sponse, due to acidic degradation of the polymer [4,8–9]. It was
properties of the scaffolds also strongly influenced the osteogenic re- shown that the degradation behavior of PLGA, PLLA and PCL composites
sponse [8,84]. A recent trend is the incorporation of different elements can be controlled not only by tailoring the concentration of BG [9,70–71]
into the composition of SBGs to enhance their biological benefits, espe- but also SBG composition, namely the amount of alkaline glass network
cially osteogenic and angiogenic potential [63]. Therefore, Ren et al. [14] modifiers (e.g. CaO) [4,8,71].
and Poh et al. [15] demonstrated that the addition of strontium- The introduction of SBG particles into a polymer matrix contributed
substituted SBG into a PCL matrix enhance osteoblast differentiation, to increase stiffness of the composites [4,10,71,75,86]. In particular,
collagen deposition and also matrix mineralisation compared to pure nano-sized SBG particles had a significant enhancing effect on the
PCL scaffolds. In turn, Oh et al. showed that PLDLA-based membranes Young's modulus of CS/SBG [73], PHB/SBG [69] and PCL/SBG [68] com-
with zinc-substituted SBG possessed significantly improved osteogenic posites when compared with SBG microparticles. Mechanical properties
and matrix mineralization potential of the rat bone marrow mesenchy- of a composite can be controlled mainly by the amount [4,10,70–71,75,
mal stem cells (rBMSCs) comparing to zinc-free glass-containing mem- 86] textural properties [4,90], and phase composition [4,68] of glass
branes and pure polymer membranes [85]. fillers. It was shown that porous microstructure of gel-derived 58S
The addition of 45S5 SBG particles into poly(D,L-lactide) (PDLLA) and SBG particles compared to dense melt-derived 45S5 glass [90] and
PLGA matrices, as well as a PGA mesh significantly enhanced vascular also larger mesopores of gel-derived calcium-rich SBG (~ 13.3 nm)
endothelial growth factor (VEGF) in vitro secretion of fibroblasts [67, with respect to smaller pores of silica-rich BG (~3.0 nm) [4–5] resulted
86–87] and also in vivo blood vessel formation [67,86]. These results in- in an improvement of mechanical strength of PDLLA and PCL-based
dicated a beneficial effect of bioglass inclusion into a polymer matrix on films, respectively. In turn, partially crystallized SBG particles could
induction of neo-vascularisation and rapid vascular ingrowth necessary have different interfacial bonding strength with PCL as compared with
for delivery of nutrients and oxygen into the developing tissues, and the fully amorphous particles, which would affect strength of the com-
thus the possibility of using glass-containing composites, not only for posites [4–5,68]. The presence of SBG particles in a polymer porous scaf-
hard tissue regeneration, but also for the soft one [67], and soft-hard tis- folds generally induces an increase of their compressive strength [10,77,
sue interfaces [86]. Other experimental study showed that the addition 86]. On the other hand, decrease of tensile strength is a frequent effect of
of a limited concentration (10 wt.%) of 45S5 SBGs nanoparticles to col- modification of polymer films with SBG fillers [4,71,75]. In order to im-
lagen films induced an early angiogenic response evaluated using the prove the interface compatibility between SBG particles and a polymer
cylinder chorioallantoic membrane (CAM) model [72]. Quinlan et al. de- matrix, surface modification of the SBG particles is used. The modifica-
veloped collagen-glycosaminoglycan-based porous scaffolds containing tion would make the particles disperse within the matrix more homo-
cobalt-substituted SBG which show ability to stimulate both angiogen- geneously [91]. Furthermore, molecules grafted onto the particles
esis and vascularisation through the release of Co2 + ions which are could interact with the polymer matrix, thus the final mechanical prop-
known for their potential to mimic hypoxia [88]. erties could be improved. It was shown that surface modification of SBG
SBGs present in a bioresorbable polymer matrix can also affect the particles by esterification reaction [90] or surface-grafting with
polymer degradation behavior by modification of its surface and bulk diisocyanate [91] improved the tensile strength of PDLLA and PLLA-
properties. On one hand, the degradation process of the composite can based composites, respectively.
be accelerated because hydrophilic SBG particles improve surface wet- In case of composites consisting of semicrystalline polymers (e.g.
tability of the composite, allowing the scaffold to absorb more water PCL, PLLA), their properties strongly depends on crystallinity of the
[9,12,83,89] and also increase the surface area of a hydrolytic attack polymer matrix. It is well known that a degree of crystallinity, size and
[4]. Furthermore, water can more easily diffuse through the interface distribution of crystallites in semicrystalline polymers have a large ef-
between the two phases by capillarity and microcrack transport mech- fect on their mechanical properties [92]. Furthermore, amorphous re-
anisms [12]. In particular, the greater amount and the smaller particle gions of a semicrystalline polymer degrade prior to crystalline
1180 M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191

Table 2
Degree of crystallinity (χc), melting temperature (Tm) (mean ± SD) and crystallite size (L110 and L200) of pure PCL and SBG-PCL composite films. The composite materials contained 12 and
21 vol.% of gel-derived silica-based bioglass microparticles with different compositions - S2 and A2.

PCL 12S2-PCL 21S2-PCL 12A2-PCL 21A2-PCL

χc (%)a 42.69 ± 1.16 36.22 ± 0.68 20.79 ± 0.79 37.25 ± 0.72 20.84 ± 0.85
Tm (°C)a 62.33 ± 0.09 60.17 ± 0.14 59.00 ± 0.14 60.67 ± 0.13 59.20 ± 0.10
L110 (nm)b 30.4 26.5 22.1 24.7 19.1
L200 (nm)b 23.6 19.1 17.2 20.5 15.5
a
Evaluated by the differential scanning calorimetry (DSC).
b
Evaluated by the wide-angle X-ray scattering (WAXS).

domains, thus initial crystallinity of the polymer determines its degra- melted ones (BET SSA 0.3 m2 g− 1) and also to the presence of –OH
dation behavior [89]. Recent studies showed that different polymer groups on the surface of the gel-derived glass that can bind polyphe-
crystallinity strongly influences adhesion and proliferation of osteo- nol molecules.
blasts and fibroblasts [93]. Some studies indicated a possibility of mod-
ulation of crystalline properties of a polymer matrix by the addition of 4.2. Borate and borosilicate bioactive glasses
glass particles. On one hand, microparticles of SBG introduced into PCL
and PLLA matrix caused a reduction of the polymer crystallinity, melting Borate-based glasses (BBGs) are known for their high reactivity and
temperature [4–5,8] and crystallite size [4,94], which were affected by bioactivity, however their biocompatibility still remains debatable [58,
the filler amount (Table 2.). On the other hand, the incorporation of 96]. It is due to cytotoxicity of high concentration of boron released
nano-sized SBG into PLLA resulted in an increase of a nucleation rate into the solution as borate ions (BO3)3− related to high solubility of bo-
and crystallinity of the polymer. In addition, PLLA grafting onto the sur- rate glasses [97–99]. Probably for this reason, borate-based glasses are
face of the nanoparticles with diisocyanate enhanced this effect [91]. not as common choice for tissue engineering applications as the silicate
Microstructure and porosity of a material play a critical role in tissue ones. Under static in vitro culture conditions, some BBGs showed signif-
engineering because they determine cell migration, cell-cell interac- icant reduction in cell proliferation, in turn, the cytotoxicity was
tions and molecular transport of nutrients, wastes, and biological medi- inhibited under dynamic culture conditions [100]. In vitro response
ators within a scaffold. There are various methods for controlling study of scaffolds consisting of borate glass 13-93B3 (B2O3–CaO–K2O–
porosity and pore architecture of polymer and composite scaffolds. In Na2O–MgO–P2O5 system) exhibited a cytotoxic effect, whereas the
particular, it is possible by varying the amount and size of porogen par- same scaffolds showed the ability to support in vivo tissue infiltration
ticles in the solvent-casting particulate leaching (SCPL) method [8] or by [101]. The conversion mechanism of BBG to HAp is similar to that of sil-
using various solvents [10], polymer concentrations [10,65] and cooling icate glass, with the formation of a borate-rich layer, similar to the silica-
temperatures [65] in the thermally induced phase separation (TIPS) rich layer of the SBGs [58]. Partial replacement of B2O3 in BBG with
process. It turned out that when TIPS method was used to obtain com- Al2O3 [102] or SrO [96] was shown to reduce the dissolution rate of
posite scaffolds based on PCL [10], PDLLA [65–66,83], PLGA [65,70] the glass. Furthermore, the presence of strontium oxide induce the ad-
and PLLA [9], porosity of the materials decreased and the pores became hesion of osteoblast-like cells, thus significantly increasing biocompati-
irregular in shape with increasing content of SBG particles. According to bility [96,102].
the authors, the probable reason for these phenomena is a hindering ef- Based on the available literature, there are only a few reports on the
fect of randomly distributed solid inorganic fillers in the polymer/sol- use of borate-based glasses as polymer matrix modifiers [7,103–105].
vent solution on the growth of solvent crystals during the phase Because of high degradation rate, BBGs are considered as modifiers of
separation process. On the other hand, the enlargement of pore size in a biodegradable polymer matrix to obtain bioactive and drug-releasing
PLLA-based scaffold upon the addition of SBG particles can be attributed scaffolds. Zhang and Jia et al. developed teicoplanin-loaded CS/BBG
to the effect of released ions from the glass surface on the TIPS thermo- composites for the chronic osteomyelitis (bone infection) treatment.
dynamics [94]. The materials did not cause any toxic injury to local or systemic tissues,
Surface properties of biomaterials, such as wettability, roughness, while they showed ability to cure chronic bone infection within
surface energy, surface charge determine cell adhesion, proliferation 12 weeks of implantation in a rabbit tibia osteomyelitis model. Further-
and differentiation [95], protein adsorption, bioactivity [69], as well as more, the composites exhibited a multifunctional role including an
degradation behavior. It was shown that SBG particles, as a hydrophilic antibiotic release function, sufficient load-bearing capacity, biodegrad-
material, remarkably improve hydrophilicity of PCL [4,68,75] and PHB ability and the ability to support bone regeneration [103–105].
[69] matrices. Furthermore, the decrease in water contact angle was A significantly higher degradation rate of borate-based glasses com-
more visible for PHB-based composites containing SBG nanoparticles pared to the silicate ones makes them more suitable for soft tissue appli-
than for microparticles due to the fact that more of the nano-sized cations. Furthermore, by altering composition of the glass fillers, namely
SBG particles were exposed on the surface [69]. Some authors observed SiO2 and B2O3 content, physical and chemical properties of a polymer-
that the excess over the certain content of BG in a composite did not re- based composite can be easily tailored to meet tissue regenerative re-
sult in a further reduction of the contact angle. It can be attributed to the quirements. Mohammadkhah et al. showed that composites containing
notable increase of surface roughness [4,69,75]. 50 wt.% PCL and (1) 50 wt.% 13-93B3 BBG or (2) 50 wt.% 45S5 SBG or (3)
Our recent study indicated that PCL/SBG composite films are capable a blend of 25 wt.% 13-93B3 and 25 wt.% 45S5 glass particles showed a
of serving as a carrier of natural polyphenols extracted from sage (Salvia modified degradation rate. It was also demonstrated that faster conver-
officinalis L.) [6]. Polyphenols-loaded composites showed potential anti- sion of 13–93B3 fillers to HAp may improve the effectiveness of the PCL-
oxidant activity against the ABTS•+ and DPPH• radicals, and antiprolifer- based material as a nerve guide conduit when compared to silicate
ative activity against human malignant melanoma WM266–4 cells. In glass-containing material [7].
contrast to PCL film, composites containing gel-derived and melt-de-
rived SBG particles exhibited a reduced initial burst release of drug in 4.3. Phosphate glasses
the first day and also excellent in vitro bioactivity. Reduction in the
initial fast release was particularly pronounced for a film with gel- Phosphate glasses (PBGs), containing P2O5 as a network-forming
derived fillers which can be attributed to much larger specific surface oxide, have a wide range of solubility which can be controlled by mod-
area of these particles (BET SSA 101.6 m2 g− 1 ) compared to the ifying their composition [106]. In the basic P2O5-CaO-Na2O system, a
M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191 1181

lower CaO content causes higher dissolution rate [107]. Furthermore, the presence of an amorphous phase. The Ca/P ratio obtained from the
the dissolution rate of PBGs can be tailored by the addition of several ox- energy dispersive X-ray spectroscopy (EDX) confirmed the presence
ides such as; Al2O3, B2O3, Fe2O3, Ga2O3, TiO2, ZnO, MgO, SrO, NiO, MnO of amorphous calcium phosphate (ACP) [112].
and CuO [17,58,106,108]. Therefore, PBGs may be used as an inorganic To conclude, besides calcium phosphate ceramics, bioactive and re-
filler of a bioresorbable polymer matrix (e.g. PCL, PLA) to obtain fully re- sorbable glasses are the most widely used and researched materials
sorbable composite materials with a controlled resorption profile [106, for modifying biodegradable polymer matrices. It is due to the possibil-
109]. Moreover, PBGs are considered as materials for bone tissue engi- ity of modulating a wide range of BG properties and thus, by its incorpo-
neering due to their chemical similarity to the inorganic phase of ration into a polymer matrix, possibility of imparting and controlling,
human bone [17]. It was also observed that dissolution products of not only physicochemical or mechanical properties, but also a number
PBGs from P2O5-CaO-Na2O system with a lower dissolution rate (28 of biological effects of the obtained composites.
and 40 mol% CaO) enhanced proliferation of osteoblastic cells and also
expression of the bone-associated proteins (bone sialoprotein (BSP), 5. Wollastonite based composites
osteonectin (ON) and fibronectin (FN)) in contrast to highly soluble
glasses (8 and 16 mol% CaO) which showed inhibitory action [107]. An- Wollastonite (WS, CaO-SiO2, CaSiO3) is a typical example of calcium-
other work showed that the incorporation of TiO2 (5 mol%) alone, as silicate-based ceramics. WS is a naturally occurring calcium silicate,
well as a mixture of TiO 2 (5 mol%) and ZnO (1, 3 and 5 mol%) in which has been widely used as a filler in polymers and cements to fab-
PBG of the basic system enhanced proliferation of osteoblastic cells ricate composites with improved mechanical properties [113–115]. Be-
compared with the basic glass and PBG containing ZnO only, which cause of its excellent bioactivity and degradability, WS has been
can be attributed to the significant reduction in a degradation rate proposed as a potential material for bone tissue regeneration [116–
of these glasses and also beneficial effect of Zn2 + ions. Moreover, 118] with the same or better possibilities than classic ceramic particles
the transcription level of osteogenesis-related genes (ALP, ON and such as TCP or HAp. Nevertheless, the extensive use of WS, TCP, HAp
core-binding factor alpha-1 (CBF-alpha-1)) of osteoblastic cells or BG is still limited by their brittle nature. Most of literature positions
seeding on glasses containing a mixture of TiO2 and 1, 3 mol% ZnO showed composite materials based on synthetic or natural polymers
was comparable or even significantly higher compared to a positive in which WS was used as a filler to produce composites with improved
control (Thermanox®) [108]. The presented studies showed that, mechanical and bioactive properties [1,119–120]. Synthetic biodegrad-
by tailoring composition of PBGs, it is possible to modulate their bio- able matrices such as; PLGA, PLDLA, PCL or poly(3-hydroxybutyrate-co-
logical properties. 3-hydroxyvalerate (PHBV) guarantee also short-term durability
Most of the works concerning the modification of a polymer ma- under in vitro/in vivo conditions, this property of materials can be
trix with phosphate bioglass are devoted to degradation studies. It controlled by modification with WS. Results of studies of Li and Chang
was shown that the dissolution of glass fillers directly influences [121−122] showed influence of WS particles on PHBV matrix. Thermo-
the degradation behavior of the polymer-based composites, as with plastic polyester such as PHBV gains bioactivity in similar way as poly-
pure phosphate glasses, the degradation rate of the composites in- mer composites modified with BG particles (with CaO-SiO2 formula).
creased with decreasing CaO content in PBG fillers [110−111]. According to this mechanism, silicate ions released from WS particles
Mohammadi et al. obtained two kinds of PCL-based composites con- are attached to Si–OH groups inducing nucleation of apatite on the sur-
taining PBGs modified with SiO2 and/or Fe2O3: (1) composite films face. At the same time, release of calcium ions from WS particles in-
containing PBG fibers of the (mol%) 50P 2O 5 –40CaO– (10-x)SiO 2 – creases the ionic activity product of SBF with respect to apatite and
xFe2O3 system, where x = 0, 5 and 10 [109], as well as (2) composite accelerates the apatite nucleation. Finally, the apatite nuclei formed on
films containing 40 vol.% glass particles of the (mol%) 50P2O 5 – the surface grow spontaneously by consuming Ca and P ions from the
40CaO–10SiO 2 or 50P2O 5 –40CaO–10Fe2 O 3 systems or blends of fluid [123–124]. A layer formed by apatite crystals was observed on
these glasses (0/40, 10/30, 20/20, 30/10, 40/0) [106]. The authors the surface of PHBV/20 or 40 wt.% WS composite after soaking in SBF
showed that the increasing SiO2 content enhanced the degradation for 14 days. Additionally, pH of the immersion medium (SBF) was stable
rate of the composites. It was manifested by higher weight loss, in- until 21 days of soaking which means that the WS incorporated into the
crease in the release of PO4 3 − , Ca2 +, Fe3 + and Si4 + ions, formation PHBV could neutralize acidic by-products and stabilize pH of the SBF so-
of pores, decrease in mechanical properties and more rapid reduc- lution [121–122]. It suggested that WS/PHBV composites could gradual-
tion in pH during the degradation test [106,109]. In turn, Georgiou ly release Ca and Si ions, and the release rate depended on content of WS
et al. showed that the degradation behavior of PLLA-based compos- in the composites (20 wt.% or 40 wt.%). The released Ca and Si ions were
ites can be also controlled by tailoring the concentration of glass par- able to form basic hydrates and this can possibly explain the ability of
ticles: the higher content of the (mol%) 46CaO– 4Na2O–50P2O5 PBG pH-stabilization showed by WS. A rapid degradation process of WS par-
particles in the composite is, the higher is the degradation rate of ticles was confirmed in long-term observations and in the other immer-
the materials [13]. sion medium i.e.; phosphate buffer saline (PBS) where pH was also
Polymer-based composites containing phosphate glasses are consid- stable after 14 days of incubation [122]. The alkaline ions neutralized
ered as drug delivery systems. It was shown that the higher the CaO acidic products of degradation of PHBV and had a buffering effect on
content in glass is, the lower a vancomycin release from the PCL-based the acidification caused by degradation of PHBV. This effect could also
composites is. Furthermore, the addition of glass with higher CaO con- be identified in the other aliphatic polyesters matrices such as PGLA or
tent into PCL matrix reduced the initial burst effect and prolongs release PLLA [125–126]. Both properties i.e.; bioactivity and biodegradability
of a drug. It is probably related to high affinity of hydrophilic drugs to of composite materials based on WS are results of high hydrophilicity
the hydrophilic surface of glass particles and lower solubility of high- of WS particles. The suitable amount of inorganic material i.e. N
CaO glasses. These results indicated that it is possible to control degra- 20 wt.% reduces hydrophobic character of polymers surface and facili-
dation rate and drug release of the PCL/PBG composites depending on tates local supersaturation, which are necessary for apatite nucleation,
the glass composition while maintaining excellent biocompatibility and hydrolysis of a polymer chain, which is a beginning of its degrada-
[110]. tion process. A good example of possibility of the water-uptake capacity
Navarro et al. showed that the incorporation of PBG particles from was characterized by a composite scaffold based on silk fibroin (SF)
the (mol%) 44.5P2O5–44.5CaO–6Na2O–5TiO2 system in a PLA matrix in- formed by the freeze-drying technique [124]. The first reason of high
duced formation of calcium phosphate precipitates at the composite water absorption is high porosity (80–90%) and then wettability of
surface after immersion in SBF. The X-ray diffraction (XRD) and Fourier composite materials e.g. water contact angle decreases from 65o to
transform infrared spectroscopy (FTIR) of the CaP precipitation showed 40o with an increasing amount of WS.
1182 M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191

Fig. 4. The dispersion of WS in the PCL/nanoWS composite determined with the PIXE technique. Pure PCL – reference sample. FTIR-ATR confirming chemical interaction between a
polymer chain and nanofillers; a) nano-WS, b) pure PCL, c), d), e) the nanocomposite with different contents of nano-WS particles; 1%, 4%, 6%wt.

Another aspect of modification of a biodegradable polymer matrix also could release antibiotics to minimize infection problems [123–124].
by WS is its influence on mechanical properties of composite materials The most popular model of this study is gentamicin – an antibiotic
i.e.; porous scaffolds. The addition of WS to the SF matrix resulted in a used in inflammatory therapeutic. It was proved that WS in poly-
structure with higher compressive strength than the pure SF e.g. com- mer-based composites such as PGLA/WS/gentamicin or PBVH/WS/
pressive strength of SF/40 wt.% WS composite was two times higher gentamicin retarded release of gentamicin and this might be due to
than strength of the pure SF [124]. Compressive strength of PHBV/WS formation of an apatite layer on the surface of the composite. This ef-
composite scaffolds was higher and reached 0.28 MPa when weight fect strongly depends on composition of an ions-reached immersion
ratio of WS in the composite was 40% [122]. The same effect was ob- medium such as SBF or PBS which could stimulate bioactivity process
served in the case of collagen/WS composites; where tensile strength [127–128].
and bioactivity of the composite scaffolds were improved by increasing Mesoporous particles of wollastonite (mWS), similarly to the meso-
WS content [126]. porous silica, possess high specific surface area and high pore volume.
The same author showed that composite systems based on polymer These properties significantly improve the kinetics of process of apatite
and wollastonite (PGLA/WS, PBVH/WS) might be used as bioactive drug formation on their surface thereby exhibiting better bioactivity and pos-
delivery systems which were not only able to enhance bone growth, but sibility of being a drug carrier [129–130]. Wie et al. confirmed that an

Fig. 5. Morphology of the nanocomposite fibers; PLA modified with 1% wt. TCP nanoparticles (a), morphology of the PLA/TCP nanocomposite fibers after 7 days of incubation in SBF and
EDX analysis confirming apatite composition (b). Republished with permission of Trans Tech Publications Inc., from [184]; permission conveyed through Copyright Clearance Center, Inc.
Table 3
Review of biodegradable ceramic-polymer composites, their form, preparation method, potential application and mechanical properties.

Polymer matrix Ceramic modifier Composite form Composite preparation Potential Compressive Young's Reference
method application (C), modulus
Type Form/preparation
tensile (T) (MPa)
method/size/amount
flexural
(F) strength
(MPa)

PCL SBG Particles/sol-gel/b50 μm/21 vol.% Scaffold (porosity Solvent casting particulate Bone tissue [8]
80SiO2–16CaO–4P2O5 55–92%, pore size b leaching; Solid-liquid phase engineering
40SiO2–54CaO–6P2O5 (mol%) 300 μm) separation; Phase inversion
SBG Particles/sol-gel/b40 μm/12, 21 Film (thickness ~ Solvent casting Bone tissue 10.5–14.5 (T) 0.6–1.3·103 [4]
80SiO2–16CaO–4P2O5 vol.% 90 μm) engineering
40SiO2–54CaO–6P2O5 (mol%)
45S5 Bioglasss® Particles/melting/b45 μm/25, 50 Scaffold (porosity Solid–liquid phase Bone tissue 92–214·10−3 132–251·10−3 [10]
wt.% 87–92%, pore size separation engineering (C)

M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191


10- some hundreds
μm)
Sr-SBG Particles/melting/b38 μm/10 wt.% Scaffold (fiber Melt-electrospinning Bone graft [14]
46SiO2-26Na2O–17SrO–6CaO–5P2O5 (mol%) diameter 46.1 μm, substitutes
fiber spacing
222.1 μm)
45S5 Bioglasss® Particles/ b38 μm/10 wt.% Scaffold (porosity 3D printing, melt Bone tissue 48–59 [15]
Sr-SBG ~75%, pore size b extrusion-based additive engineering
46SiO2-24Na2O–20SrO–7CaO–3P2O5 (mol%) 2.05 mm) manufacturing technology
SBG Particles/sol-gel template Film (thickness Solvent casting Bone tissue 15–19 (T) 200–383 [75]
60SiO2–63CaO–4P2O5 (mol%) method/50–150 nm/10, 20, 30 wt.% 65 μm) engineering
BBG Particles/b20 μm/50 wt.% Film (thickness Tape casting Neural tissue 40–50 (T) 150–190 [7]
53B2O3–20CaO–12K2O–6Na2O–5MgO–4P2O5 60 μm) engineering
(wt.%)
45S5 Bioglasss®
PBG Particles/melting/10 wt.% Film Solvent extraction Tissue [110]
45P2O5–xCaO–(55-x)Na2O, x = 20, 30, 40, and thermal pressing regeneration
50 and
wound-healing
PBG Fibers/melt–draw Film (thickness ~ Thermal pressing Bone fracture 35–45 (F) ~1.8·103 [109]
50P2O5–40CaO–(10- x)SiO2–xFe2O3, x = 0, spinning/diameter 10–20 μm/~18 200 μm) fixation
5, 10 vol.% devices
(mol%)
BCP Particles/microwave-hydrothermal Scaffold (porosity Solvent casting particulate Bone tissue 1.0–2.2 (C) 6.5–20.5·103 [189]
method/80-100 nm/50–70 wt.% 71–82%, pore size leaching engineering
100–500 μm,)
SiO2 Particles/ 2.3–2.7 nm/10, 30 wt.% Scaffolds (porosity Supercritical carbon dioxide Hard tissue 10–175 [48]
11–55%, pore size (scCO2)-assisted engineering
4.1 nm-39.5 μm) foaming/mixing
WS Particles/sol-gel/20, 40, 50 wt.% Scaffold (porosity Solvent casting particulate Hard tissue 5–12 (C) [129]
57–74%, pore size leaching repair
1–500 μm)
PLGA SBG Particles/sol-gel/b50 μm/12, 21, 33 Film (thickness ~ Solvent casting /solvent Bone tissue 24–55 (T) - 2–4·103 - [71,77]
80SiO2–16CaO–4P2O5 vol.% 110 μm /scaffolds casting particulate leaching engineering films films
40SiO2–54CaO–6P2O5 (mol%) (porosity 88–95%,
pore size
90–300 μm)
45S5 Bioglasss® Particles/melting/b5 μm/10, 25, 50 Scaffold (porosity Thermal induced phase Hard and soft 22–27 [65,70]
wt.% N90%, pore size separation tissue
10–100 μm) engineering

1183
(continued on next page)
1184
Table 3 (continued)

Polymer matrix Ceramic modifier Composite form Composite preparation Potential Compressive Young's Reference
method application (C), modulus
Type Form/preparation
tensile (T) (MPa)
method/size/amount
flexural
(F) strength
(MPa)

BCP Particles/ heating bovine Scaffold (porosity Thermal induced phase Bone tissue 0.16–0.55 3.2–15.0 [192]
63/37 HAp/β-TCP bone/b100 nm/ 10–50 wt.% N89%, pore size separation engineering (yield
60–150 μm) strength) (C)
HAp Particles/~100 nm/50 wt.% Scaffold (porosity Gas foaming, Particulate Bone tissue 2.3–4.5 [146]
86–91%) leaching engineering
PLLA SBG Particles/sol-gel template Scaffold (pore size Thermal induced phase Bone tissue [82]
55SiO2–40CaO–5P2O5 method/20 nm/25 wt.% 10–300 μm) separation engineering
30SiO2–60CaO–10P2O5
68SiO2–28CaO–4P2O5 (mol%)

M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191


45S5 Bioglasss® Particles/melting/50–63 μm/25, 50 Scaffold (porosity Solid–liquid phase Bone tissue [9]
wt.% 48–52%, pore size b separation/freeze-extraction engineering
40 μm)
SBG Particles/melting/10 μm/10, 30, 50 Scaffold (pore size Thermal induced phase Bone tissue 0.62–0.75 (C) 6.61–10.88 [94]
33CaO–28P2O5–16MgO–23SiO2 (wt.%) wt.% 27–150 μm) separation engineering
PBG Particles/melting/32–1000 μm/5, Scaffold (porosity Melt extrusion and Bone tissue 80.0–116.5 [13]
46CaO–4Na2O–50P2O5 (mol%) 10, 20 wt.% ~81%, pore size supercritical CO2-foaming engineering
200–400 μm)
HAp Particles/0–50 μm, 5 μm, b200 Cylindrical profile Melt extrusion Various 100–136 (F) 3.3–4.3·103 [151]
nm/10, 20, 30 wt.% medical
application
β-TCP Particles/microwave accelerated Scaffold (porosity Particulate leaching Bone tissue 1–5 (C) [168]
wet method/100–300 nm/60 vol.% 50%, pore size engineering
200–450 μm)
β-TCP Particles/in the presence of Scaffold (porosity Thermal induced phase Bone tissue 0.3–0.8 [171]
PEG/290 nm/10–30 wt.% N93%, pore size separation + particulate engineering
0.5–300 μm, fiber leaching
diameter 70–300
nm)
PDLLA 45S5 Bioglasss® Particles/melting/b5 μm/5, 10, 25, Scaffold (porosity Thermal induced phase Hard and soft 21 [65–66,83]
40, 50 wt.% N90%, pore size separation tissue
10–100 μm) engineering
45S5 Bioglasss® Particles/melting/0.1–25 μm/5, 10, Film (thickness Solvent casting/solvent Regeneration 0.38–1.59 (C) 0.36–1.80 [86]
20 wt.% - films, 20 wt.%- scaffolds 25–40 μm)/scaffolds casting particulate leaching of hard-soft
Particles/flame spray (porosity 85–93%, tissue defects
synthesis/35–40 nm/5, 10, 20 wt.%
- films, 20 wt.%- scaffolds
PLDLA SBG Particles/sol-gel/b45 μm/30 wt.% Membrane Solvent casting Guided bone 54.2–56.3 (T) [85]
70SiO2–30CaO (thickness ~ regeneration
Zn-SBG 150 μm)
70SiO2–25CaO–5ZnO
PBG Particles/melting/b80 μm/50 wt.% Disk (thickness ~ Solvent casting Bone tissue [112]
44.5P2O5–44.5CaO–6Na2O–5TiO2 (mol%) 1.5 mm) and thermal pressing engineering
Poly(e-caprolactone-co-DL-lactide) SBG Particles/b45 μm, 90–315 μm/40, Disk (thickness 2 Compression moulding Bone [11–12]
53SiO2-23Na2O–20CaO–4P2O5 (wt.%) 60, 70 wt.% mm) substitution
PHB 45S5 Bioglasss® Particles/melting/b5 μm/10, 20, 30 Film (thickness Solvent casting Bone tissue 0.8–1.6·103 [69]
wt.% 120–140 μm) engineering
Particles/flame spray synthesis/29
nm/10, 20, 30 wt.%
PHBV WS Particles/chemical Scaffold (porosity Compression moulding Bone tissue 0.20–0.28 [121]
coprecipitation/98–154 μm/20, 40 73–78%, pore size particulate leaching engineering
wt.% 30–300 μm)
M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191 1185

[103–105]
apatite layer on the surface of a composite with mWS was observed
after 7-days soaking in the SBF medium. In the case of a composite

[141]

[124]
[73]

[72]

[88]
with the same amount of classic WS it took 14 days [129,131].
Higher specific surface area provided by mWS particles present in a
polymer matrix promotes cells interaction (osteoblast, chondrocyte)
5.5–9·10−3

4–6·10−3
modulus)

[76,132–133]. Further studies on the application of mWS showed that


(Storage
17–20

1.6–2
even 0.5% concentration of mWS in chitosan/carboxymethylcellulose
(CS/CMC) scaffolds significantly improves their physicochemical and bi-
ological properties [131]. Inclusion of mWS particles had no effect on al-

0.15–0.2 (C)
tering the porous architecture of the CS/CMC scaffolds but influenced
23–33 (C)

protein adsorption and biomineralization ability. The CS/CMC/mWS


scaffolds were biocompatible and enhanced differentiation of MSCs to
osteoblasts [132].
Hard and soft
osteomyelitis

A small amount of additives present in a degradable polymer matrix


Treatment of
engineering

engineering

engineering

engineering

engineering
Bone tissue

Bone tissue

Bone tissue

Bone tissue

characterizes nanocomposite materials. When PCL was used as a matrix


chronic

with nanometric WS (0.5–1 wt.%.) as a filler a homogenous dispersion


tissue

was obtained. A result of a good distribution of the nanofiller within


the polymer matrix was better mechanical properties (Young's modu-
lus, tensile strength and work-of-break) [134]. It suggests that there is
Thermal induced phase

Thermal induced phase


Compression moulding

some interaction between WS nanoparticles and the polymer matrix


(hydrogen bonding, van der Waals interaction) indicating that the ce-
ramic filler contributes to the overall elastic properties of the composite
Solvent casting

Solvent casting

Freeze-drying

samples. Enhancement of the modulus of nanocomposites at such a low


separation

separation

concentration of ceramic particles cannot be attributed to the introduc-


tion of a ceramic filler with higher modulus, which is observed in tradi-
tional composite materials i.e. with 40 wt.% of WS. The same interaction
between a polymer chain and nanometric WS could be confirmed by
Scaffold (porosity N

81–85%, pore size ~


Scaffold (pore size

Scaffold (porosity
Film (thickness ~

Film (thickness ~

spectroscopic studies such as FTIR in the transmission mode, the FTIR-


ATR technique and the proton-induced X-ray emission (PIXE) method
60–125 μm)

[135]. Dispersion of WS nanoadditives is well visible in PIXE where an-


100 μm)
50 μm)

40 μm)
Particles/melting/b50 μm/~70 wt.% Pellet

alytical elements for WS (i.e. Ca and Si) could be observed (Fig. 4).
98%)

To conclude, the biodegradable polymer/WS composite possess


many interesting properties (biocompatibility, bioactivity, biodegrad-
ability and combining with them carrying therapeutic agents) which
Particles/20–30 nm/10, 20 wt.%
Particles/~50 nm/0.5, 1, 2 wt.%

Particles/melting/38, 100 μm/


synthesis/30–50 nm/30 wt.%

were guaranteed by WS particles. Some authors pointed out that


precipitation/10, 20 wt.%

these materials can be used as an alternative for conventional bioactive


available/5 μm/30 wt.%
Particles/commercially

ceramic fillers such as HAp, TCP, BCP and BG used to modify the polymer
Particles/flame spray

Particles/chemical

matrix.

6. Calcium phosphate ceramics based composites

Over the past few decades, calcium phosphate (CaP) ceramics alone
have been widely used as bone graft substitutes, especially because of
49SiO2-26Na2O–19CaO–1P2O5-4CoO (mol%)
6Na2O–8K2O–8MgO–22CaO–54B2O3–2P2O5

similarity of their chemical composition to the mineral phase of bone


[136–137]. Due to their nature, calcium phosphate ceramics show also
46SiO2-23Na2O–27CaO–4P2O5 (wt.%)

high biocompatibility and ability to bond with bone tissue under certain
conditions, however, because of their brittleness, their clinical applica-
49SiO2-26Na2O–23CaO–1P2O5

tions have been limited to the non- or low-load bearing parts of the
skeleton.
Many types of calcium phosphates have been considered as bioma-
terials for bone reconstruction in dental, orthopaedics and maxillofacial
45S5 Bioglasss®

45S5 Bioglasss®

application due to different behavior in the living organism, including


bioactivity, biodegradability and biological response. Bioactivity, degra-
Co-SBG
(mol%)

dation behavior and osteoconductivity/osteoinductivity of CaP ceramics


HAp
BBG

SBG

WS

generally depend on the Ca/P ratio, crystallinity and phase composition


[138–139]. Synthetic HAp (Ca10(PO4)6(OH)2) shows good stability in
the body, while tricalcium phosphates (α-TCP, β-TCP, Ca3(PO4)2) are
more soluble, whereas biphasic calcium phosphate (BCP; an intimate
Collagen-glycosaminoglycan

mixture of HAp and β-TCP) exhibits intermediate properties depending


on the weight ratio of stable/degradable phases. Thus, the dissolution
rate decreases in the following order: α-TCP N β-TCP N BCP N HAp [1,
138]. Because the natural bone hydroxyapatite is nonstoichiometric
and contains, beside the main components i.e.; Ca, (PO4)3 −, (OH)−,
Collagen


some other groups and trace elements (e.g. CO2– 3 , F , Mg
2+
, Na+, K+,
2+ 2+
Sr , Zn ), new trends in calcium phosphates preparation, especially
CS

SF

HAp, consist in obtaining substituted CaP ceramics. They do not only


1186 M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191

have a different solubility and bioactivity than the parent materials, but strength, and impact energy. Furthermore, a composite with a low con-
also cause modified biological response due to the release of biologically tent (4 wt.%) of gHAp possessed higher mechanical properties than the
active ions during dissolution [136]. CaP ceramics exhibit different bio- pure PLLA. The implantation study of repairing critical-sized defects in
logical effects in vivo, while most of them are osteoconductive, only cer- the radius of a rabbit forelimb showed that the gHAp/PLGA scaffold ex-
tain types are osteoinductive. Samavedi et al. proposed that a range of hibited rapid and strong mineralization and osteoconductivity [143].
osteoinductive potentials of CaPs decrease in the following order Degradation studies showed that the incorporation of HAp [144] and
BCP N TCP N HAp [138]. nHAp particles [145] into PCL and PLGA matrix, respectively, accelerat-
Currently HAp, TCP and BCP ceramics are quite common types of ed degradation of the composites. Furthermore, a degradation rate of
materials used for various biomedical applications and they are avail- composites can be adjusted by varying the HAp content [144–145].
able on the market. In the past few years, many efforts have been direct- Kim et al. obtained nHAp/PLGA composite scaffolds using the gas
ed to develop calcium phosphate-containing composite materials, forming particulate leaching method. The scaffolds exhibited signifi-
especially polymer-matrix composites. cantly higher rat calvarial osteoblast growth, alkaline phosphatase ac-
tivity, and extracellular matrix (ECM) mineralization in vitro
6.1. Hydroxyapatite based composites compared to the pure PLGA material. Furthermore, the presence of
nHAp fillers enhanced hydrophilicity and in vivo osteoconductivity
Stoichiometric HAp (Ca/P molar ratio of 1.67) is considered to be (higher bone formation area and more extensive calcium deposition)
osteoconductive but not osteoinductive. Furthermore, because HAp is of the material [146–147].
the most stable among CaP ceramics, its surfaces provide highly effec- Another study showed that incorporation of Ag-doped HAp into
tive nucleating sites for the precipitation of an apatite crystal in contact PHBV nanofibers imparted antibacterial activity against E. coli and S. au-
with culture medium and body fluids [138]. Solubility, bioactivity and reus bacteria and enhanced in vitro bioactivity when compared to the
biological response of HAp can be modified by anionic and cationic sub- material containing unmodified HAp, while did not show any cytotoxic
stitution [136]. Mainly for these reasons, HAp is widely used to prepare effect [152].
polymer-ceramic composite materials usually with the aim to impart
the bioactivity and osteoconductivity and also improve mechanical 6.2. Tricalcium phosphate based composites
properties [140–147].
It is believed that one of the most promising bone graft material is Tricalcium phosphate (β-TCP, TCP) is a well-known CaP-based
collagen/nano-sized HAp (nHAp) composite because of its ability to bioceramics. β-tricalcium phosphate has been widely investigated and
mimic structure and composition of natural bone. New technics such successfully used in clinical application as a biomaterial for bone repair
as self-assembly and biomineralization have been recently used to ob- applications due to its remarkable biocompatibility, in vivo
tain collagen/nHAp composites with oriented and hierarchical structure resorbability, bioactivity and good osteoconductivity [153–183]. More-
considering the spatial relationship between organic and inorganic over, some results indicated that TCP is considered to be osteoinductive
phases [148–149]. Using the self-assembling method, a highly oriented [138]. For many years TCP has been combined with other ceramic mate-
and with morphology similar to compact bones, collagen/nHAp com- rials such as; HAp, Al2O3, ZrO2 to produce small dental or orthopedic im-
posite was obtained [148]. A pre-crystallization method allowed plants [154–156] in the form of composites or composite layers on the
obtaining a composite in which HAp nanocrystals were situated implants [156–159]. Last decades showed that the most perspective
among the collagen fibers creating a suitable structure for bone defect form of TCP is small particles or nanoparticles which modify a biode-
repair [149]. gradable polymer matrix [157,160–162]. Incorporation of the ceramic
Several studies were conducted on the use another natural polymer, particles into polymer-based composite materials presented significant
namely CS, as a component of HAp-containing composites. The pres- osteogenic benefits and encouraged formation of a new bone on im-
ence of nHAp in a CS matrix increased compressive strength [140] and plant surfaces [163–164]. In addition, during degradation process of
Young's modulus of composites [141]. It can be related to hydrogen- composite implants/scaffolds based on biodegradable aliphatic polyes-
bonding interactions between NH2 and OH groups of nHAp and chela- ters, particles of TCP which modified the polymer matrix, improved
tion between NH2 and Ca2+ when the co-precipitation method of com- not only osteoconductivity but also neutralized acidic pH of the environ-
posite preparation was used [140]. The incorporation of nHAp affected a ment typical for degradation of a polymer matrix [163–165]. These phe-
degradation rate of composites [140,141], induced bioactivity [140] and nomena decrease the risk of local complications (inflammatory
favored murine osteoblast-like MC3T3-E1 cells attachment and prolifer- reactions) and increase the degradation rate comparing to pure poly-
ation [141]. mer implants [166]. Moreover, employment of a room temperature
In the literature, HAp fillers have also been shown to modify numer- preparation technique of such composites may allow incorporating of
ous properties of synthetic polymer matrices, especially poly(α-hy- thermally unstable antibiotics, as well as growth factors e.g.; BMP-2 as
droxy esters). The comparison of hydroxyapatite with different stimulants of bone formation. Osteoconductive composite materials
morphologies, namely nanoparticles (nHAp) and whiskers (wHAp), as such as; PLLA/β-TCP, PCL/β-TCP comprising a large volume fraction of
PLGA matrix modifiers, showed that composites with nHAp had higher β-TCP (≥ 60 vol.%) and a smaller amount of polymer (PLLA, PCL) were
bending strength in comparison with wHAp-containing materials. It can studied as degradable antibiotic carrier materials for the orthopedic
be attributed to a more homogeneous distribution of nHAp in the poly- treatment [166–167]. A study on PCL/TCP beads loaded with vancomy-
mer matrix and also enhanced crystallization of the PLGA matrix. It was cin (1–4 wt.%) showed that a gradual drug release observed over the pe-
also shown that morphology of HAp fillers influenced in vitro degrada- riod of 4–11 weeks had diffusional character and depended on the
tion behavior of the composites [150]. In turn, the use of various con- composite matrix homogeneity and porosity [167].
tents (0–30 wt.%) and also sizes (0–50 μm, 5 μm and N 200 nm) of Methods of fabrication of PLA/β-TCP and PCL/β-TCP composites i.e.;
HAp particles led to changes in thermal properties and/or crystallinity, cold sintering and/or salt leaching allowed obtaining materials suitable
as well as the mechanical strength of the PLLA and PLGA-based compos- for load-bearing bone scaffolds [167–169]. These materials were charac-
ites [151]. terized by high strength and Young's modulus required as biomechani-
In order to improve the adhesive strength between HAp nanoparti- cal properties of such implants. It is, therefore, believed that the first
cles and PLLA and PLGA matrix, the hydroxyl groups on the surface of resorbable and a strong bone graft substitute will consist of a poly-
the ceramic nanoparticles were grafted with PLLA by chemical bonding mer-calcium phosphate composite and will contain a large ceramic frac-
[142–143]. It was shown, that composites containing grafted HAp tion [170]. A small fraction (≤40 vol.%) of the particulate ceramic phase
(gHAp) exhibited significantly improved tensile strength, bending decreased mechanical properties of composite implants and scaffolds
M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191 1187

but they were still osteoconductive and promoted osteoblasts adhesion, concentration of β-TCP phase in the BCP [187]. In turn, Mg substitution
proliferation, penetration, and ECM deposition and influenced their of BCP ceramics enhanced their biodegradation and bioactivity, and also
degradation rate [171]. Kobayashi et al. showed that small amount of improved biocompatibility, as well as in vitro and in vivo
TCP (contents of 5–15 wt.%) in PLLA agglomerated and these regions osteoconductivity tested with human adipose tissue-derived mesen-
acted as water diffusion paths and started hydrolytic degradation of chymal stem cells (hAT-MSCs) [188]. Above-mentioned features make
the polymer matrix near the TCP-PLLA interfaces [172]. BCP ceramics useful as a modifier of biodegradable synthetic polymers
Results of some investigations pointed out that nanometric par- and copolymers; PCL [189], PLLA [190], PLGA [191–192] as well as nat-
ticles of β-TCP with high specific surface area and chemical groups ural polymers and their blends; collagen [193–194], gelatin-pectin
capable of interaction with a polymer chain can affect both bioactiv- [195].
ity/osteoconductivity and mechanical properties [173]. The fibrous Ebrahimian-Hosseinabadi et al. demonstrated that the modification
PLA/TCP (1 wt.%) nanocomposites produced by the electrospinning of PLGA with various amounts (10–50 wt.%) of BCP nanoparticles affects
method belong to a group of biomimetic materials because they in vitro degradation behavior of the composite scaffolds. On one hand,
chemically and structurally mimic a natural bone tissue. The nano- with increasing amount of BCP nanoparticles in the composites, reduc-
composite fibers with nanometric or submicrometric diameters tion of Young's modulus increased as a result of higher weight loss of the
have particle size and composition similar to the natural collagen fi- scaffolds. It could be related to an increase in hydrophilicity and the
bers present in a bone. A three dimensional (3D) form of the mate- water uptake of the composite scaffolds. On the other hand, a buffering
rial facilitates settling and adhering of bone cells and inducting effect of basic degradation products of BCP on a degradation process of
apatite formation on the surface of nanofibers (Fig. 5) [184]. PLGA was observed. Namely, more pronounced decrease in the molec-
Despite of this, implants/scaffolds made of PLGA and nanometric ular weight of the polymer with reducing amount of BCP in the matrix
particles of β-TCP would not induce stronger osteointegration than im- occurred during the degradation process [192].
plants made of PLGA and micrometric-size β-TCP particles e.g. a push Another work showed that BCP particles, homogeneously dispersed
out test showed similar results in both materials. in PLGA nanofibers, enhanced expression of osteogenic differentiation
To overcome these limitations, the reinforcement of a polymer ma- markers (ALP and BSP) and also ECM mineralisation as a function of in-
trix by TCP nanoparticles was proved when crosslinking CS was used creasing content of the ceramic particles in the composites. These find-
as the matrix polymer [174–176]. As a crosslinking agent could be ings demonstrated a beneficial effect of BCP inclusion in PLGA
used; genipine, gluatre aldehyd or tripolyphosphate. It was shown nanofibers on the osteogenic differentiation of osteoblast-like MC3T3-
that a composite scaffold obtained with genipine-crosslinked CS and E1 cells and confirmed osteoinductive character of the composites
nanometric β-TCP improved mechanical, bioactivity and cell supportive [191]. The incorporation of BCP nanoparticles in another polymer ma-
properties such as viability and osteogenic differentiation of human trix, namely PCL, resulted in an increase in ALP activity of human
MSCs [177–178]. Such scaffolds had slightly higher extent of mineraliza- MSCs and also significantly improved compressive strength and Young's
tion after 21 days of incubation in an osteogenic medium. By changing modulus of the materials [189]. In order to enhance mechanical proper-
the crosslinking agent e.g. to tripolyphosphate it was possible to reduce ties of PLLA-based composites, the surface of BCP microparticles was
its degradation rate without harmful side effects [179–180], pores size modified by direct grafting with L-lactide. The results showed improved
and wettability, swelling and compressive strength of scaffolds [176– interfacial interaction between the BCP filler and the polymer matrix,
177]. and thus higher compressive strength in comparison to composites con-
In order to enhance osteoconductive process supported by a taining the unmodified particles [190]. It was also shown that when TIPS
polymer composite with TCP a new strategy was proposed. To im- [192] and SCPL [189] methods were used to obtain PLGA and PCL-based
prove and accelerate healing of bone grafts for orthopedic and dental composite scaffolds the porosity of the materials decreased with in-
use a platelet-rich plasma (PRP) was applied on 3D scaffolds made of creasing BCP nanoparticles content.
PCL/TCP composite materials. The PRP is an autologous source of A commercially available collagen/BCP composite (Osteon™ II colla-
concentrated platelets that contains several prepackaged growth gen, Genoss. Co. Ltd) was studied using a rabbit calvarial defect model
factors including transforming growth factor (TGF), platelet-derived by Lee et al. [193–194]. The composite material showed good
growth factor (PDGF), and vascular endothelial growth factor osteoconductive properties and also relatively slow resorption and
(VEGF). Immersion of a PCL/TCP scaffold in PRP fluid and its thus, in contrast to collagen sponge, it maintained the space needed
implanting into a critical-size femoral defects of a rat led after 3 for bone tissue regeneration [193]. It was also shown, that the colla-
months of the implantation to neovascularization and bone bridging gen/BCP composite is a promising candidate for a carrier of recombinant
[181–182]. In another study, when PCL/TCP scaffolds in combination human BMP-2 providing a constant release profile [194]. In turn, a gel-
with PRP were used in critical-sized defects healing of a canine man- atin-pectin/BCP nanocomposite scaffold was fabricated for delivering
dible the results were confirmed in long-term observations: a bone growth factors (BMP-2 and VEGF). The release studies exhibited that
volume fraction after both control times i.e. 6 and 9 months was bet- the scaffold had constant release properties. Furthermore, in vitro and
ter than in the control group. The regular deposition of osteoid with in vivo tests conducted with osteoblast-like MC3T3-E1 and rat models,
alveolar bone trabeculae was shown for all treated defects [183]. respectively, showed that the composites enhanced cell proliferation
and new bone formation [195].
6.3. Biphasic calcium phosphate based composites To conclude, calcium phosphates, due to their wide range of solubil-
ity, bioactivity, osteoconductivity/osteoinductivity and various mechan-
Biphasic calcium phosphate ceramics are a family of two-phase ma- ical behaviours, can be successfully applied as fillers of biodegradable
terials that combine the low solubility and osteoconductive character of polymer matrices to modulate composite properties.
HAp with the osteoinductivity of a more soluble phase, namely β-TCP.
The main advantage is the possibility of controlling bioactivity, resorp- 7. Summary
tion rate and also mechanical properties of CaP ceramic by manipulating
the HAp/β-TCP ratio [138,185–186]. BCP properties, especially biologi- As pointed out above, ceramic materials discussed in this work,
cal effects, can be modified also by ion substitution. Kim et al. showed namely silica, bioactive and resorbable glasses, wollastonite and calcium
that that Sr incorporation into CaP ceramics increased a proliferation phosphate ceramics have a number of unique and beneficial properties
rate and differentiation (ALP activity) of the MG63 and HOS osteo- for medical applications, especially for tissue engineering and regenera-
blast-like cells. Such behaviours can be attributed to the combined ef- tive medicine. Depending on the type of ceramics, these properties in-
fects, that is, not only to the Sr ions release but also to higher clude; bioactivity, osteoconductivity, osteoinductivity, resorbability,
1188 M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191

antibacterial and anti-inflammatory activity, ability to induce vascular- Biomedical Materials Research - Part B Applied Biomaterials (2014), http://dx.
doi.org/10.1002/jbm.b.33350.
isation and many others. As it has been shown in this review, the cur- [9] A.M. El-Kady, E.A. Saad, B.M.A. El-Hady, M.M. Farag, Synthesis of silicate glass/
rent biomaterials age belongs to composite materials which are poly(L-lactide) composite scaffolds by freeze-extraction technique: characteriza-
classified as the fourth generation of biomaterials. They combine not tion and in vitro bioactivity evaluation, Ceram. Int. 36 (3) (2010) 995–1009.
[10] P. Fabbri, V. Cannillo, A. Sola, A. Dorigato, F. Chiellini, Highly porous
less than two different phases i.e.; polymer and ceramic particles or polycaprolactone-45S5 bioglass® scaffolds for bone tissue engineering, Compos.
nanoparticles. These new materials are mostly characterized by many Sci. Technol. 70 (13) (2010) 1869–1878.
advantageous physicochemical, mechanical and biological properties [11] T. Jaakkola, J. Rich, T. Tirri, T. Närhi, M. Jokinen, J. Seppälä, A. Yli-Urpo, In vitro ca-P
precipitation on biodegradable thermoplastic composite of poly(ε-caprolactone-
and what is even more important, they give scientists the opportunity co-DL-lactide) and bioactive glass (S53P4), Biomaterials 25 (4) (2004) 575–581.
of a better control of such properties as; mechanical behavior, degrada- [12] J. Rich, T. Jaakkola, T. Tirri, T. Närhi, A. Yli-Urpo, J. Seppälä, In vitro evaluation of
tion rate, surface topography, surface charge and wettability and inter- poly(ε-caprolactone-co-DL-lactide)/bioactive glass composites, Biomaterials 23
(10) (2002) 2143–2150.
action with cells. Many of the new composite materials modified with
[13] G. Georgiou, L. Mathieu, D.P. Pioletti, P. Bourban, J.E. Månson, J.C. Knowles, S.N.
the ceramic fillers gain unique futures as antibacterial properties or in- Nazhat, Polylactic acid-phosphate glass composite foams as scaffolds for bone tis-
duct specific cells reactions e.g.; differentiation or activation of secretion sue engineering, Journal of Biomedical Materials Research - Part B Applied Bioma-
of growth factors. Apart from the ceramic filler amount, other parame- terials 80 (2) (2007) 322–331.
[14] J. Ren, K.A. Blackwood, A. Doustgani, P.P. Poh, R. Steck, M.M. Stevens, M.A.
ters such as its size (nanometric or micrometric), its shape (particles, Woodruff, Melt-electrospun polycaprolactone strontium-substituted bioactive
whiskers, nanotubes, fibers), distribution and a type of surface function- glass scaffolds for bone regeneration, Journal of Biomedical Materials Research -
al groups influence final properties of the composites. Type of matrix Part A 102 (9) (2014) 3140–3153.
[15] P.S.P. Poh, D.W. Hutmacher, M.M. Stevens, M.A. Woodruff, Fabrication and in vitro
polymer and its parameters, such as the molecular weight, polydispersi- characterization of bioactive glass composite scaffolds for bone regeneration,
ty, crystallinity, chain orientation, functional groups and overall hydro- Biofabrication 5 (4) (2013).
philicity also strongly affect the composites properties. Of a great [16] K. Shin, Y. Koh, W. Choi, H. Kim, Production of porous poly(ε-caprolactone)/silica
hybrid membranes with patterned surface pores, Mater. Lett. 65 (12) (2011)
importance and impact for applications in the medicine is that natural 1903–1906.
and synthetic polymers have good formability. As presented in Table [17] A. Hoppe, N.S. Güldal, A.R. Boccaccini, A review of the biological response to ionic
3, numerous methods of fabrication of polymer-based composites com- dissolution products from bioactive glasses and glass-ceramics, Biomaterials 32
(11) (2011) 2757–2774.
bined with many types of ceramic fillers allow obtaining the composites [18] J. Korventausta, M. Jokinen, A. Rosling, T. Peltola, A. Yli-Urpo, Calcium phosphate
in various forms (porous scaffolds, films, membranes, fibers) and with formation and ion dissolution rates in silica gel-PDLLA composites, Biomaterials
different morphology and mechanical properties. Moreover, as de- 24 (28) (2003) 5173–5182.
[19] Q.Z. Chen, I.D. Thompson, A.R. Boccaccini, 45S5 bioglass®-derived glass-ceramic
scribed in the subsequent sections of this work, these composites also
scaffolds for bone tissue engineering, Biomaterials 27 (11) (2006) 2414–2425.
vary in terms of biological effects and other relevant properties. The [20] A.A. Vassiliou, D. Bikiaris, K. El Mabrouk, M. Kontopoulou, Effect of evolved interac-
present work shows that through the selection of materials of the ce- tions in poly(butylene succinate)/fumed silica biodegradable in situ prepared
ramic filler and the matrix, their properties and relationships, it is pos- nanocomposites on molecular weight, material properties, and biodegradability,
J. Appl. Polym. Sci. 119 (4) (2011) 2010–2024.
sible to design and obtain materials with a wide range of properties [21] A.K. Gupta, P. Mani, S. Krishnamoorthy, Interfacial adhesion in polyester resin con-
for various applications in the medicine, especially in the bone defects crete, Int. J. Adhes. Adhes. 3 (3) (1983) 149–154.
treatment. [22] V. Kominar, M. Narkis, A. Sigemann, A. Vaxman, Compressive strength of unidirec-
tional and crossply carbon fiber/PEEK composites, J. Mater. Sci. 30 (10) (1995)
To conclude, the composite materials provide a multistage design 2620–2627.
approach and therefore greater possibilities to control their material [23] T. Nihei, A. Dabanoglu, T. Teranaka, S. Kurata, K. Ohashi, Y. Kondo, K.
and biological properties than the ceramics and the polymers alone. Kunzelmann, Three-body-wear resistance of the experimental composites
containing filler treated with hydrophobic silane coupling agents, Dent.
Mater. 24 (6) (2008) 760–764.
Acknowledgments [24] Y.L. Wu, A.I.Y. Tok, F.Y.C. Boey, X.T. Zeng, X.H. Zhang, Surface modification of ZnO
nanocrystals, Appl. Surf. Sci. 253 (12) (2007) 5473–5479.
[25] D. G. Papageorgiou, K. Chrissafis, E. Pavlidou, E. A. Deliyanni, G. Z. Papageorgiou, Z.
The authors would like to acknowledge the financial support from Terzopoulou and D. N. Bikiaris, “Effect of nanofiller's size and shape on the solid
the National Science Centre, Poland Grant Nos. 2015/17/N/ST8/00226 state microstructure and thermal properties of poly(butylene succinate) nanocom-
posites,” Thermochim. Acta, vol. 590, pp. 181–190, y.
(MD), 2014/13/B/ST8/02973 (KCK), 2012/07/B/ST8/03378 (ESZ) and
[26] K. Fukushima, C. Abbate, D. Tabuani, M. Gennari, P. Rizzarelli, G. Camino, Biodegra-
from the Polish Ministry of Science and Higher Education Grant No. N dation trend of poly(ε-caprolactone) and nanocomposites, Mater. Sci. Eng. C 30 (4)
N507 401 939 (ESZ). (2010) 566–574.
[27] Q. Zhou, M. Xanthos, Nanosize and microsize clay effects on the kinetics of the
thermal degradation of polylactides, Polym. Degrad. Stab. 94 (3) (2009) 327–338.
References [28] Y. Li, C. Han, J. Bian, L. Han, L. Dong, G. Gao, Rheology and biodegradation of
polylactide/silica nanocomposites, Polym. Compos. 33 (10) (2012) 1719–1727.
[1] K. Rezwan, Q.Z. Chen, J.J. Blaker, A.R. Boccaccini, Biodegradable and bioactive po- [29] A. Rapacz-Kmita, E. Stodolak-Zych, B. Szaraniec, M. Gajek, P. Dudek, Effect of clay
rous polymer/inorganic composite scaffolds for bone tissue engineering, Biomate- mineral on the accelerated hydrolytic degradation of polylactide in the polymer/
rials 27 (18) (2006) 3413–3431. clay nanocomposites, Mater. Lett. 146 (2015) 73–76.
[2] I. Armentano, M. Dottori, E. Fortunati, S. Mattioli, J.M. Kenny, Biodegradable poly- [30] E. Stodolak-Zych, M. Szumera, M. Blazewicz, Osteoconductive nanocomposite ma-
mer matrix nanocomposites for tissue engineering: a review, Polym. Degrad. terials for bone regeneration, Mater. Sci. Forum 730–732 (2013) 38–43.
Stab. 95 (11) (2010) 2126–2146. [31] M. Shie, S. Ding, H. Chang, The role of silicon in osteoblast-like cell proliferation and
[3] L.S. Nair, C.T. Laurencin, Biodegradable polymers as biomaterials, Progress in Poly- apoptosis, Acta Biomater. 7 (6) (2011) 2604–2614.
mer Science (Oxford) 32 (8–9) (2007) 762–798. [32] E.M. Carlisle, Silicon: a requirement in bone formation independent of vitamin D1,
[4] M. Dziadek, E. Menaszek, B. Zagrajczuk, J. Pawlik, K. Cholewa-Kowalska, New gen- Calcif. Tissue Int. 33 (1) (1981) 27–34.
eration poly(ε-caprolactone)/gel-derived bioactive glass composites for bone tis- [33] D.M. Reffitt, N. Ogston, R. Jugdaohsingh, H.F.J. Cheung, B.A.J. Evans, R.P.H.
sue engineering: part I. material properties, Mater. Sci. Eng. C 56 (2015) 9–21. Thompson, G.N. Hampson, Orthosilicic acid stimulates collagen type 1 synthesis
[5] M. Dziadek, B. Zagrajczuk, M. Ziabka, K. Dziadek, K. Cholewa-Kowalska, The role of and osteoblastic differentiation in human osteoblast-like cells in vitro, Bone 32
solvent type, size and chemical composition of bioactive glass particles in modulat- (2) (2003) 127–135.
ing material properties of poly(ε-caprolactone) based composites, Compos. A: [34] S. Chen, A. Osaka, N. Hanagata, Collagen-templated sol-gel fabrication, microstruc-
Appl. Sci. Manuf. 90 (2016) 90–99. ture, in vitro apatite deposition, and osteoblastic cell MC3T3-E1 compatibility of
[6] M. Dziadek, K. Dziadek, B. Zagajczuk, E. Menaszek, K. Cholewa-Kowalska, Poly(ε- novel silica nanotube compacts, J. Mater. Chem. 21 (12) (2011) 4332–4338.
caprolactone)/bioactive glass composites enriched with polyphenols extracted [35] S. Chen, X. Shi, H. Morita, J. Li, N. Ogawa, T. Ikoma, N. Hanagata, BMP-2-loaded silica
from sage (Salvia officinalis L.), Mater. Lett. (2016). nanotube fibrous meshes for bone generation, Sci. Technol. Adv. Mater. 12 (6)
[7] A. Mohammadkhah, L.M. Marquardt, S.E. Sakiyama-Elbert, D.E. Day, A.B. Harkins, (2011).
Fabrication and characterization of poly-(ε)-caprolactone and bioactive glass com- [36] N. Shadjou, M. Hasanzadeh, Bone tissue engineering using silica-based mesopo-
posites for tissue engineering applications, Mater. Sci. Eng. C 49 (2015) 632–639. rous nanobiomaterials: recent progress, Mater. Sci. Eng. C 55 (2015) 401–409.
[8] M. Dziadek, J. Pawlik, E. Menaszek, E. Stodolak-Zych, K. Cholewa-Kowalska, Effect [37] Z. Wei, C. Wang, H. Liu, S. Zou, Z. Tong, Facile fabrication of biocompatible PLGA
of the preparation methods on architecture, crystallinity, hydrolytic degradation, drug-carrying microspheres by O/W pickering emulsions, Colloids Surf. B:
bioactivity, and biocompatibility of PCL/bioglass composite scaffolds, Journal of Biointerfaces 91 (1) (2012) 97–105.
M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191 1189

[38] R. Gref, Y. Minamitake, M.T. Peracchia, V. Trubetskoy, V. Torchilin, R. Langer, Biode- [69] S.K. Misra, D. Mohn, T.J. Brunner, W.J. Stark, S.E. Philip, I. Roy, A.R. Boccaccini, Com-
gradable long-circulating polymeric nanospheres, Science 263 (5153) (1994) parison of nanoscale and microscale bioactive glass on the properties of P(3HB)/
1600–1603. Bioglass® composites, Biomaterials 29 (12) (2008) 1750–1761.
[39] J. Venkatesan, I. Bhatnagar, P. Manivasagan, K. Kang, S. Kim, Alginate composites [70] A.R. Boccaccini, V. Maquet, Bioresorbable and bioactive polymer/Bioglass® com-
for bone tissue engineering: a review, Int. J. Biol. Macromol. 72 (2015) 269–281. posites with tailored pore structure for tissue engineering applications, Compos.
[40] M. Ahola, J. Rich, P. Kortesuo, J. Kiesvaara, J. Seppälä, A. Yli-Urpo, In vitro evaluation Sci. Technol. 63 (16) (2003) 2417–2429.
of biodegradable ε-caprolactone-co-D,L-lactide/silica xerogel composites contain- [71] E. Pamula, J. Kokoszka, K. Cholewa-Kowalska, M. Laczka, L. Kantor, L. Niedzwiedzki,
ing toremifene citrate, Int. J. Pharm. 181 (2) (1999) 181–191. A.M. Osyczka, Degradation, bioactivity, and osteogenic potential of composites
[41] Y. Wang, Q. Zhao, N. Han, L. Bai, J. Li, J. Liu, S. Wang, Mesoporous silica nanoparticles made of PLGA and two different sol-gel bioactive glasses, Ann. Biomed. Eng. 39
in drug delivery and biomedical applications, Nanomedicine: Nanotechnology, Bi- (8) (2011) 2114–2129.
ology, and Medicine 11 (2) (2015) 313–327. [72] G.E. Vargas, L.A.H. Durand, V. Cadena, M. Romero, R.V. Mesones, M. Mačković, A.A.
[42] G. Abdelbary, N. El-Gendy, Niosome-encapsulated gentamicin for ophthalmic con- Gorustovich, Effect of nano-sized bioactive glass particles on the angiogenic prop-
trolled delivery, AAPS PharmSciTech 9 (3) (2008) 740–747. erties of collagen based composites, J. Mater. Sci. Mater. Med. 24 (5) (2013)
[43] P. Horcajada, A. Rámila, J. Pérez-Pariente, M. Vallet-Regí, Influence of pore size of 1261–1269.
MCM-41 matrices on drug delivery rate, Microporous Mesoporous Mater. 68 (1– [73] S.G. Caridade, E.G. Merino, N.M. Alves, V.D.Z. Bermudez, A.R. Boccaccini, J.F. Mano,
3) (2004) 105–109. Chitosan membranes containing micro or nano-size bioactive glass particles: evo-
[44] H. Lapidus, N.G. Lordi, Drug release from compressed hydrophilic matrices, J. lution of biomineralization followed by in situ dynamic mechanical analysis, J.
Pharm. Sci. 57 (8) (1968) 1292–1301. Mech. Behav. Biomed. Mater. 20 (2013) 173–183.
[45] N. Lordi, Sustained release dosage forms, Theory and Practice of Industrial Pharma- [74] I.Y. Kim, A. Sugino, K. Kikuta, C. Ohtsuki, S.B. Cho, Bioactive composites
cy, Lea and Febiger, Philadelphia 1986, pp. 430–456. consisting of PEEK and calcium silicate powders, J. Biomater. Appl. 24 (2)
[46] I.I. Slowing, J.L. Vivero-Escoto, C. Wu, V.S. Lin, Mesoporous silica nanoparticles as (2009) 105–118.
controlled release drug delivery and gene transfection carriers, Adv. Drug Deliv. [75] B. Lei, K. Shin, D. Noh, I. Jo, Y. Koh, H. Kim, S.E. Kim, Sol-gel derived nanoscale bio-
Rev. 60 (11) (2008) 1278–1288. active glass (NBG) particles reinforced poly(ε-caprolactone) composites for bone
[47] H. Qu, S. Bhattacharyya, P. Ducheyne, Silicon oxide based materials for controlled tissue engineering, Mater. Sci. Eng. C 33 (3) (2013) 1102–1108.
release in orthopedic procedures, Adv. Drug Deliv. Rev. (2015). [76] I.D. Xynos, A.J. Edgar, L.D.K. Buttery, L.L. Hench, J.M. Polak, Gene-expression profil-
[48] M.B.C. De Matos, A.P. Piedade, C. Alvarez-Lorenzo, A. Concheiro, M.E.M. Braga, H.C. ing of human osteoblasts following treatment with the ionic products of bioglass®
De Sousa, Dexamethasone-loaded poly(e-caprolactone)/silica nanoparticles com- 45S5 dissolution, J. Biomed. Mater. Res. 55 (2) (2001) 151–157.
posites prepared by supercritical CO2 foaming/mixing and deposition, Int. J. [77] J. Filipowska, J. Pawlik, K. Cholewa-Kowalska, G. Tylko, E. Pamula, L. Niedzwiedzki,
Pharm. 456 (2) (2013) 269–281. A.M. Osyczka, Incorporation of sol-gel bioactive glass into PLGA improves mechan-
[49] Z. Guo, X. Liu, L. Ma, J. Li, H. Zhang, Y. Gao, Y. Yuan, Effects of particle morphology, ical properties and bioactivity of composite scaffolds and results in their
pore size and surface coating of mesoporous silica on naproxen dissolution rate en- osteoinductive properties, Biomedical Materials (Bristol) 9 (6) (2014).
hancement, Colloids Surf. B: Biointerfaces 101 (2013) 228–235. [78] R.M. Day, Bioactive glass stimulates the secretion of angiogenic growth factors and
[50] R. Ravichandran, S. Gandhi, D. Sundaramurthi, S. Sethuraman, U.M. Krishnan, Hier- angiogenesis in vitro, Tissue Eng. 11 (5–6) (2005) 768–777.
archical mesoporous silica nanofibers as multifunctional scaffolds for bone tissue [79] D. Zhang, O. Leppäranta, E. Munukka, H. Ylänen, M.K. Viljanen, E. Eerola, L. Hupa,
regeneration, J. Biomater. Sci. Polym. Ed. 24 (17) (2013) 1988–2005. Antibacterial effects and dissolution behavior of six bioactive glasses, Journal of
[51] E. Stodolak-Zych, A. Rapacz-Kmita, M. Dudek, Potential of superhydrophobic layer Biomedical Materials Research - Part A 93 (2) (2010) 475–483.
on the implant surface, Corrosion and Surface Engineering 227 (2015) 511–514. [80] R.M. Day, A.R. Boccaccini, Effect of particulate bioactive glasses on human macro-
[52] T. Suzuki, Y. Mizushima, Characteristics of silica-chitosan complex membrane and phages and monocytes in vitro, Journal of Biomedical Materials Research - Part A
their relationships to the characteristics of growth and adhesiveness of L-929 cells 73 (1) (2005) 73–79.
cultured on the biomembrane, J. Ferment. Bioeng. 84 (2) (1997) 128–132. [81] T.A. Ostomel, Q. Shi, C. Tsung, H. Liang, G.D. Stucky, Spherical bioactive glass with
[53] G. Toskas, C. Cherif, R. Hund, E. Laourine, B. Mahltig, A. Fahmi, T. Hanke, enhanced rates of hydroxyapatite deposition and hemostatic activity, Small 2
Chitosan(PEO)/silica hybrid nanofibers as a potential biomaterial for bone regen- (11) (2006) 1261–1265.
eration, Carbohydr. Polym. 94 (2) (2013) 713–722. [82] Z. Hong, R.L. Reis, J.F. Mano, Preparation and in vitro characterization of novel bio-
[54] J.A. Sowjanya, J. Singh, T. Mohita, S. Sarvanan, A. Moorthi, N. Srinivasan, N. active glass ceramic nanoparticles, Journal of Biomedical Materials Research - Part
Selvamurugan, Biocomposite scaffolds containing chitosan/alginate/nano-silica A 88 (2) (2009) 304–313.
for bone tissue engineering, Colloids Surf. B: Biointerfaces 109 (2013) 294–300. [83] V. Maquet, A.R. Boccaccini, L. Pravata, I. Notingher, R. Jérôme, Preparation, charac-
[55] S. Liao, C.K. Chan, S. Ramakrishna, Stem cells and biomimetic materials strategies terization, and in vitro degradation of bioresorbable and bioactive composites
for tissue engineering, Mater. Sci. Eng. C 28 (8) (2008) 1189–1202. based on bioglass®-filled polylactide foams, Journal of Biomedical Materials Re-
[56] M. Dziadek, B. Zagrajczuk, P. Jelen, Z. Olejniczak, K. Cholewa-Kowalska, Structural search - Part A 66 (2) (2003) 335–346.
variations of bioactive glasses obtained by different synthesis routes, Ceram. Int. [84] J. Yao, S. Radin, G. Reilly, P.S. Leboy, P. Ducheyne, Solution-mediated effect of bio-
42 (13) (2016) 14700–14709. active glass in poly (lactic-co-glycolic acid)-bioactive glass composites on osteo-
[57] Q. Fu, E. Saiz, M.N. Rahaman, A.P. Tomsia, Bioactive glass scaffolds for bone tissue genesis of marrow stromal cells, Journal of Biomedical Materials Research - Part
engineering: state of the art and future perspectives, Mater. Sci. Eng. C 31 (7) A 75 (4) (2005) 794–801.
(2011) 1245–1256. [85] S. Oh, J. Won, H. Kim, Composite membranes of poly(lactic acid) with zinc-added
[58] G. Kaur, O.P. Pandey, K. Singh, D. Homa, B. Scott, G. Pickrell, A review of bioactive bioactive glass as a guiding matrix for osteogenic differentiation of bone marrow
glasses: their structure, properties, fabrication and apatite formation, Journal of mesenchymal stem cells, J. Biomater. Appl. 27 (4) (2012) 413–422.
Biomedical Materials Research - Part A 102 (1) (2014) 254–274. [86] L. Gerhardt, K.L. Widdows, M.M. Erol, C.W. Burch, J.A. Sanz-Herrera, I. Ochoa, A.R.
[59] W. Cao and L. L. Hench, “Bioactive materials,” Ceram. Int., vol. 22, no. 6, pp. 493– Boccaccini, The pro-angiogenic properties of multi-functional bioactive glass com-
507, 1996. posite scaffolds, Biomaterials 32 (17) (2011) 4096–4108.
[60] M. Mami, A. Lucas-Girot, H. Oudadesse, R. Dorbez-Sridi, F. Mezahi, E. Dietrich, In- [87] H. Keshaw, G. Georgiou, J.J. Blaker, A. Forbes, J.C. Knowles, R.M. Day, Assessment of
vestigation of the surface reactivity of a sol-gel derived glass in the ternary system polymer/bioactive glass-composite microporous spheres for tissue regeneration
SiO2-CaO-P2O5, Appl. Surf. Sci. 254 (22) (2008) 7386–7393. applications, Tissue Engineering - Part A 15 (7) (2009) 1451–1461.
[61] J.R. Jones, Review of bioactive glass: from hench to hybrids, Acta Biomater. 9 (1) [88] E. Quinlan, S. Partap, M.M. Azevedo, G. Jell, M.M. Stevens, F.J. O'Brien, Hypoxia-
(2013) 4457–4486. mimicking bioactive glass/collagen glycosaminoglycan composite scaffolds to en-
[62] P. Sepulveda, J.R. Jones, L.L. Hench, Characterization of melt-derived 45S5 and sol- hance angiogenesis and bone repair, Biomaterials 52 (1) (2015) 358–366.
gel-derived 58S bioactive glasses, J. Biomed. Mater. Res. 58 (6) (2001) 734–740. [89] C.X.F. Lam, M.M. Savalani, S. Teoh, D.W. Hutmacher, Dynamics of in vitro polymer
[63] S.M. Rabiee, N. Nazparvar, M. Azizian, D. Vashaee, L. Tayebi, Effect of ion substitu- degradation of polycaprolactone-based scaffolds: accelerated versus simulated
tion on properties of bioactive glasses: a review, Ceram. Int. 41 (6) (2015) physiological conditions, Biomed. Mater. 3 (3) (2008).
7241–7251. [90] Y. Gao, J. Chang, Surface modification of bioactive glasses and preparation of
[64] M. Dziadek, B. Zagrajczuk, E. Menaszek, A. Wegrzynowicz, J. Pawlik, K. Cholewa- PDLLA/bioactive glass composite films, J. Biomater. Appl. 24 (2) (2009) 119–138.
Kowalska, Gel-derived SiO2–CaO–P2O5 bioactive glasses and glass-ceramics mod- [91] A. Liu, Z. Hong, X. Zhuang, X. Chen, Y. Cui, Y. Liu, X. Jing, Surface modification of bio-
ified by SrO addition, Ceram. Int. 42 (5) (2016) 5842–5857. active glass nanoparticles and the mechanical and biological properties of poly(L-
[65] V. Maquet, A.R. Boccaccini, L. Pravata, I. Notingher, R. Jérôme, Porous poly(α- lactide) composites, Acta Biomater. 4 (4) (2008) 1005–1015.
hydroxyacid)/Bioglass® composite scaffolds for bone tissue engineering. I: prepa- [92] N. Dusunceli, O.U. Colak, Modelling effects of degree of crystallinity on mechanical
ration and in vitro characterization, Biomaterials 25 (18) (2004) 4185–4194. behavior of semicrystalline polymers, Int. J. Plast. 24 (7) (2008) 1224–1242.
[66] J.J. Blaker, J.E. Gough, V. Maquet, I. Notingher, A.R. Boccaccini, In vitro evaluation of [93] H. Cui, P.J. Sinko, The role of crystallinity on differential attachment/proliferation of
novel bioactive composites based on bioglass®-filled polylactide foams for bone osteoblasts and fibroblasts on poly (caprolactone-co-glycolide) polymeric surfaces,
tissue engineering scaffolds, Journal of Biomedical Materials Research - Part A 67 Front. Mater. Sci. 6 (1) (2012) 47–59.
(4) (2003) 1401–1411. [94] N. Barroca, A.L. Daniel-Da-Silva, P.M. Vilarinho, M.H.V. Fernandes, Tailoring the
[67] R.M. Day, A.R. Boccaccini, S. Shurey, J.A. Roether, A. Forbes, L.L. Hench, S.M. Gabe, morphology of high molecular weight PLLA scaffolds through bioglass addition,
Assessment of polyglycolic acid mesh and bioactive glass for soft-tissue engineer- Acta Biomater. 6 (9) (2010) 3611–3620.
ing scaffolds, Biomaterials 25 (27) (2004) 5857–5866. [95] M. Padial-Molina, P. Galindo-Moreno, J.E. Fernández-Barbero, F. O'Valle, A.B. Jódar-
[68] E. Tamjid, R. Bagheri, M. Vossoughi, A. Simchi, Effect of particle size on the in vitro Reyes, J.L. Ortega-Vinuesa, P.J. Ramón-Torregrosa, Role of wettability and
bioactivity, hydrophilicity and mechanical properties of bioactive glass-reinforced nanoroughness on interactions between osteoblast and modified silicon surfaces,
polycaprolactone composites, Mater. Sci. Eng. C 31 (7) (2011) 1526–1533. Acta Biomater. 7 (2) (2011) 771–778.
1190 M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191

[96] H.B. Pan, X.L. Zhao, X. Zhang, K.B. Zhang, L.C. Li, Z.Y. Li, J. Chang, Strontium borate [123] P. Li, C. Ohtsuki, T. Kokubo, K. Nakanishi, N. Soga, T. Nakamura, T. Yamamuro, Ef-
glass: potential biomaterial for bone regeneration, J. R. Soc. Interface 7 (48) fects of ions in aqueous media on hydroxyapatite induction by silica gel and its rel-
(2010) 1025–1031. evance to bioactivity of bioactive glasses and glass-ceramics, Journal of Applied
[97] W. Huang, D.E. Day, K. Kittiratanapiboon, M.N. Rahaman, Kinetics and mechanisms Biomaterials: An Official Journal of the Society for Biomaterials 4 (3) (1993) 221–229.
of the conversion of silicate (45S5), borate, and borosilicate glasses to hydroxyap- [124] H. Zhu, J. Shen, X. Feng, H. Zhang, Y. Guo, J. Chen, Fabrication and characterization
atite in dilute phosphate solutions, J. Mater. Sci. Mater. Med. 17 (7) (2006) of bioactive silk fibroin/wollastonite composite scaffolds, Mater. Sci. Eng. C 30 (1)
583–596. (2010) 132–140.
[98] A. Yao, D. Wang, W. Huang, Q. Fu, M.N. Rahaman, D.E. Day, In vitro bioactive char- [125] H. Li, J. Chang, pH-compensation effect of bioactive inorganic fillers on the degra-
acteristics of borate-based glasses with controllable degradation behavior, J. Am. dation of PLGA, Compos. Sci. Technol. 65 (14) (2005) 2226–2232.
Ceram. Soc. 90 (1) (2007) 303–306. [126] Y. Chen, A.F.T. Mak, M. Wang, J. Li, M.S. Wong, PLLA scaffolds with biomimetic ap-
[99] M. Brink, T. Turunen, R. Happonen, A. Yli-Urpo, Compositional dependence of bio- atite coating and biomimetic apatite/collagen composite coating to enhance oste-
activity of glasses in the system Na2O- K2O-MgO-cao-B2O3-P2O5-SiO2, J. Biomed. oblast-like cells attachment and activity, Surf. Coat. Technol. 201 (3–4) (2006)
Mater. Res. 37 (1) (1997) 114–121. 575–580.
[100] R.F. Brown, M.N. Rahaman, A.B. Dwilewicz, W. Huang, D.E. Day, Y. Li, B.S. Bal, Effect [127] F. Barrera-Méndez, J.C. Escobedo-Bocardo, D.A. Cortés-Hernández, J.M. Almanza-
of borate glass composition on its conversion to hydroxyapatite and on the prolif- Robles, E.M. Múzquiz-Ramos, Gentamicin sulphate release from lost foam wollas-
eration of MC3T3-E1 cells, Journal of Biomedical Materials Research - Part A 88 (2) tonite scaffolds using poly(DL-lactide-co-glycolide) acid, Ceram. Int. 37 (7) (2011)
(2009) 392–400. 2445–2451.
[101] Q. Fu, M.N. Rahaman, B.S. Bal, L.F. Bonewald, K. Kuroki, R.F. Brown, Silicate, borosil- [128] H. Li, J. Chang, Preparation, characterization and in vitro release of gentamicin from
icate, and borate bioactive glass scaffolds with controllable degradation rate for PHBV/wollastonite composite microspheres, J. Control. Release 107 (3) (2005)
bone tissue engineering applications. II. in vitro and in vivo biological evaluation, 463–473.
Journal of Biomedical Materials Research - Part A 95 (1) (2010) 172–179. [129] J. Wei, F. Chen, J. Shin, H. Hong, C. Dai, J. Su, C. Liu, Preparation and characterization
[102] D. Zhao, W. Huang, M.N. Rahaman, D.E. Day, D. Wang, Mechanism for converting of bioactive mesoporous wollastonite - polycaprolactone composite scaffold, Bio-
Al2O3-containing borate glass to hydroxyapatite in aqueous phosphate solution, materials 30 (6) (2009) 1080–1088.
Acta Biomater. 5 (4) (2009) 1265–1273. [130] X. Li, J. Shi, Y. Zhu, W. Shen, H. Li, J. Liang, J. Gao, A template route to the prepara-
[103] W. Jia, X. Zhang, C. Zhang, X. Liu, W. Huang, M.N. Rahaman, D.E. Day, Elution char- tion of mesoporous amorphous calcium silicate with high in vitro bone-forming
acteristics of teicoplanin-loaded biodegradable borate glass/chitosan composite, bioactivity, Journal of Biomedical Materials Research - Part B Applied Biomaterials
Int. J. Pharm. 387 (1–2) (2010) 184–186. 83 (2) (2007) 431–439.
[104] W. Jia, X. Zhang, S. Luo, X. Liu, W. Huang, M.N. Rahaman, J. Wang, Novel borate [131] R. Sainitya, M. Sriram, V. Kalyanaraman, S. Dhivya, S. Saravanan, M. Vairamani, N.
glass/chitosan composite as a delivery vehicle for teicoplanin in the treatment of Selvamurugan, Scaffolds containing chitosan/carboxymethyl cellulose/mesopo-
chronic osteomyelitis, Acta Biomater. 6 (3) (2010) 812–819. rous wollastonite for bone tissue engineering, Int. J. Biol. Macromol. 80 (2015)
[105] X. Zhang, W. Jia, Y. Gu, W. Xiao, X. Liu, D. Wang, N. Zhou, Teicoplanin-loaded borate 481–488.
bioactive glass implants for treating chronic bone infection in a rabbit tibia osteo- [132] P. Valerio, M.M. Pereira, A.M. Goes, M.F. Leite, The effect of ionic products from bio-
myelitis model, Biomaterials 31 (22) (2010) 5865–5874. active glass dissolution on osteoblast proliferation and collagen production, Bioma-
[106] M.S. Mohammadi, I. Ahmed, B. Marelli, C. Rudd, M.N. Bureau, S.N. Nazhat, Modula- terials 25 (15) (2004) 2941–2948.
tion of polycaprolactone composite properties through incorporation of mixed [133] H. Li, W. Zhai, J. Chang, Effects of wollastonite on proliferation and differentiation of
phosphate glass formulations, Acta Biomater. 6 (8) (2010) 3157–3168. human bone marrow-derived stromal cells in PHBV/Wollastonite composite scaf-
[107] V. Salih, K. Franks, M. James, G.W. Hastings, J.C. Knowles, I. Olsen, Development of folds, J. Biomater. Appl. 24 (3) (2009) 231–246.
soluble glasses for biomedical use part II: the biological response of human osteo- [134] I. Kotela, J. Podporska, E. Soltysiak, K.J. Konsztowicz, M. Blazewicz, Polymer nano-
blast cell lines to phosphate-based soluble glasses, J. Mater. Sci. Mater. Med. 11 composites for bone tissue substitutes, Ceram. Int. 35 (6) (2009) 2475–2480.
(10) (2000) 615–620. [135] C. Paluszkiewicz, E. Stodolak, M. Blazewicz, J.W.M. Kwiatek, FT-IR investigations of
[108] E.A.A. Neel, L.A. O'Dell, W. Chrzanowski, M.E. Smith, J.C. Knowles, Control of surface poly-ε-caprolactone/wollastonite nanocompositesXth International Conference on
free energy in titanium doped phosphate based glasses by co-doping with zinc, Molecular Spectroscopy: from molecules to molecular materials and biological systems
Journal of Biomedical Materials Research - Part B Applied Biomaterials 89 (2) 6–10 September, Kraków–Białka Tatrzańska, 2009.
(2009) 392–407. [136] M. Šupová, Substituted hydroxyapatites for biomedical applications: a review,
[109] M.S. Mohammadi, I. Ahmed, N. Muja, S. Almeida, C.D. Rudd, M.N. Bureau, S.N. Ceram. Int. 41 (8) (2015) 9203–9231.
Nazhat, Effect of si and fe doping on calcium phosphate glass fiber reinforced [137] M. Sadat-Shojai, M. Khorasani, E. Dinpanah-Khoshdargi, A. Jamshidi, Synthesis
polycaprolactone bone analogous composites, Acta Biomater. 8 (4) (2012) methods for nanosized hydroxyapatite with diverse structures, Acta Biomater. 9
1616–1626. (8) (2013) 7591–7621.
[110] H. Kim, E. Lee, I. Jun, H. Kim, J.C. Knowles, Degradation and drug release of phos- [138] S. Samavedi, A.R. Whittington, A.S. Goldstein, Calcium phosphate ceramics in bone
phate glass/polycaprolactone biological composites for hard-tissue regeneration, tissue engineering: a review of properties and their influence on cell behavior, Acta
Journal of Biomedical Materials Research - Part B Applied Biomaterials 75 (1) Biomater. 9 (9) (2013) 8037–8045.
(2005) 34–41. [139] S.V. Dorozhkin, Biphasic, triphasic and multiphasic calcium orthophosphates, Acta
[111] R.L. Prabhakar, S. Brocchini, J.C. Knowles, Effect of glass composition on the degra- Biomater. 8 (3) (2012) 963–977.
dation properties and ion release characteristics of phosphate glass - [140] Z. Li, L. Yubao, Y. Aiping, P. Xuelin, W. Xuejiang, Z. Xiang, Preparation and in vitro
polycaprolactone composites, Biomaterials 26 (15) (2005) 2209–2218. investigation of chitosan/nano-hydroxyapatite composite used as bone substitute
[112] M. Navarro, M.P. Ginebra, J.A. Planell, C.C. Barrias, M.A. Barbosa, In vitro degrada- materials, J. Mater. Sci. Mater. Med. 16 (3) (2005) 213–219.
tion behavior of a novel bioresorbable composite material based on PLA and a sol- [141] W.W. Thein-Han, R.D.K. Misra, Biomimetic chitosan-nanohydroxyapatite compos-
uble CaP glass, Acta Biomater. 1 (4) (2005) 411–419. ite scaffolds for bone tissue engineering, Acta Biomater. 5 (4) (2009) 1182–1197.
[113] J. Tong, Y. Ma, M. Jiang, Effects of the wollastonite fiber modification on the sliding [142] Z. Hong, P. Zhang, C. He, X. Qiu, A. Liu, L. Chen, X. Jing, Nano-composite of poly(L-
wear behavior of the UHMWPE composites, Wear 255 (1–6) (2003) 734–741. lactide) and surface grafted hydroxyapatite: mechanical properties and biocom-
[114] N.M.P. Low, J.J. Beaudoin, Mechanical properties and microstructure of high alumi- patibility, Biomaterials 26 (32) (2005) 6296–6304.
na cement-based binders reinforced with natural wollastonite micro-fibers, Cem. [143] P. Zhang, Z. Hong, T. Yu, X. Chen, X. Jing, In vivo mineralization and osteogenesis of
Concr. Res. 24 (4) (1994) 650–660. nanocomposite scaffold of poly(lactide-co-glycolide) and hydroxyapatite surface-
[115] D.K. Pattanayak, Apatite wollastonite-poly methyl methacrylate bio-composites, grafted with poly(L-lactide), Biomaterials 30 (1) (2009) 58–70.
Mater. Sci. Eng. C 29 (5) (2009) 1709–1714. [144] K.C. Ang, K.F. Leong, C.K. Chua, M. Chandrasekaran, Compressive properties and de-
[116] P. Siriphannon, Y. Kameshima, A. Yasumori, K. Okada, S. Hayashi, Formation of hy- gradability of poly(ε-caprolatone)/ hydroxyapatite composites under accelerated
droxyapatite on CaSiO3 powders in simulated body fluid, J. Eur. Ceram. Soc. 22 (4) hydrolytic degradation, Journal of Biomedical Materials Research - Part A 80 (3)
(2002) 511–520. (2007) 655–660.
[117] X. Liu, C. Ding, Z. Wang, Apatite formed on the surface of plasma-sprayed wollas- [145] J. Liuyun, X. Chengdong, J. Lixin, X. Lijuan, Degradation behavior of hydroxyapatite/
tonite coating immersed in simulated body fluid, Biomaterials 22 (14) (2001) poly(lactic-co-glycolic) acid nanocomposite in simulated body fluid, Mater. Res.
2007–2012. Bull. 48 (10) (2013) 4186–4190.
[118] S.A. Saadaldin, A.S. Rizkalla, Synthesis and characterization of wollastonite [146] S. Kim, K. Ahn, M.S. Park, J. Lee, C.Y. Choi, B. Kim, A poly(lactide-co-glycolide)/hy-
glass-ceramics for dental implant applications, Dent. Mater. 30 (3) (2014) droxyapatite composite scaffold with enhanced osteoconductivity, Journal of Bio-
364–371. medical Materials Research - Part A 80 (1) (2007) 206–215.
[119] S. Kunjalukkal Padmanabhan, F. Gervaso, M. Carrozzo, F. Scalera, A. Sannino, A. [147] S. Kim, M.S. Park, O. Jeon, C.Y. Choi, B. Kim, Poly(lactide-co-glycolide)/hydroxyapa-
Licciulli, Wollastonite/hydroxyapatite scaffolds with improved mechanical, bioac- tite composite scaffolds for bone tissue engineering, Biomaterials 27 (8) (2006)
tive and biodegradable properties for bone tissue engineering, Ceram. Int. 39 (1) 1399–1409.
(2013) 619–627. [148] A. Ficai, E. Andronescu, G. Voicu, C. Ghitulica, B.S. Vasile, D. Ficai, V. Trandafir, Self-
[120] S.M. Rea, R.A. Brooks, S.M. Best, T. Kokubo, W. Bonfield, Proliferation and differen- assembled collagen/hydroxyapatite composite materials, Chem. Eng. J. 160 (2)
tiation of osteoblast-like cells on apatite-wollastonite/polyethylene composites, (2010) 794–800.
Biomaterials 25 (18) (2004) 4503–4512. [149] L. Zhang, P. Tang, M. Xu, W. Zhang, W. Chai, Y. Wang, Effects of crystalline phase on
[121] H. Li, J. Chang, Fabrication and characterization of bioactive wollastonite/PHBV the biological properties of collagen-hydroxyapatite composites, Acta Biomater. 6
composite scaffolds, Biomaterials 25 (24) (2004) 5473–5480. (6) (2010) 2189–2199.
[122] H. Li, J. Chang, In vitro degradation of porous degradable and bioactive PHBV/wol- [150] J. Liuyun, X. Chengdong, J. Lixin, X. Lijuan, Effect of hydroxyapatite with different
lastonite composite scaffolds, Polym. Degrad. Stab. 87 (2) (2005) 301–307. morphology on the crystallization behavior, mechanical property and in vitro
M. Dziadek et al. / Materials Science and Engineering C 71 (2017) 1175–1191 1191

degradation of hydroxyapatite/poly(lactic-co-glycolic) composite, Compos. Sci. [175] S. Rodrigues, A.M.R.D. Costa, A. Grenha, Chitosan/carrageenan nanoparticles: effect
Technol. 93 (2014) 61–67. of cross-linking with tripolyphosphate and charge ratios, Carbohydr. Polym. 89 (1)
[151] B. Damadzadeh, H. Jabari, M. Skrifvars, K. Airola, N. Moritz, P.K. Vallittu, Effect of ce- (2012) 282–289.
ramic filler content on the mechanical and thermal behavior of poly-L-lactic acid [176] C. Ji, N. Annabi, A. Khademhosseini, F. Dehghani, Fabrication of porous chitosan
and poly-L-lactic-co-glycolic acid composites for medical applications, J. Mater. scaffolds for soft tissue engineering using dense gas CO2, Acta Biomater. 7 (4)
Sci. Mater. Med. 21 (9) (2010) 2523–2531. (2011) 1653–1664.
[152] A. Suslu, A.Z. Albayrak, E. Bayir, A. Sendemir Urkmez, U. Cocen, In vitro biocompat- [177] N. Siddiqui, K. Pramanik, Effects of micro and nano β-TCP fillers in freeze-gelled
ibility and antibacterial activity of electrospun ag doped HAp/PHBV composite chitosan scaffolds for bone tissue engineering, J. Appl. Polym. Sci. 131 (21) (2014).
nanofibers, International Journal of Polymeric Materials and Polymeric Biomate- [178] J. Qiu, J. Li, G. Wang, L. Zheng, N. Ren, H. Liu, Y. Wang, In vitro investigation on the
rials 64 (9) (2015) 465–470. biodegradability and biocompatibility of genipin cross-linked porcine acellular
[153] S.V. Dorozhkin, Bioceramics of calcium orthophosphates, Biomaterials 31 (7) dermal matrix with intrinsic fluorescence, ACS Appl. Mater. Interfaces 5 (2)
(2010) 1465–1485. (2013) 344–350.
[154] M. Gasik, A. Keski-Honkola, Y. Bilotsky, M. Friman, Development and optimisation [179] F. Pati, B. Adhikari, S. Dhara, Collagen intermingled chitosan-tripolyphosphate
of hydroxyapatite-β-TCP functionally gradated biomaterial, J. Mech. Behav. nano/micro fibrous scaffolds for tissue-engineering application, J. Biomater. Sci.
Biomed. Mater. 30 (2014) 266–273. Polym. Ed. 23 (15) (2012) 1923–1938.
[155] L. Zheng, F. Yang, H. Shen, X. Hu, C. Mochizuki, M. Sato, Y. Zhang, The effect of com- [180] F. Pati, H. Kalita, B. Adhikari, S. Dhara, Osteoblastic cellular responses on ionically
position of calcium phosphate composite scaffolds on the formation of tooth tissue crosslinked chitosan-tripolyphosphate fibrous 3-D mesh scaffolds, Journal of Bio-
from human dental pulp stem cells, Biomaterials 32 (29) (2011) 7053–7059. medical Materials Research - Part A 101 (9) (2013) 2526–2537.
[156] D. Jang, Y. Kim, B. Lee, Microstructure control of TCP/TCP-(t-ZrO2)/t-ZrO2 compos- [181] B. Rai, M.E. Oest, K.M. Dupont, K.H. Ho, S.H. Teoh, R.E. Guldberg, Combination of
ites for artificial cortical bone, Mater. Sci. Eng. C 31 (8) (2011) 1660–1666. platelet-rich plasma with polycaprolactone-tricalcium phosphate scaffolds for seg-
[157] J. Kulkova, N. Moritz, E.O. Suokas, N. Strandberg, K.A. Leino, T.T. Laitio, H.T. Aro, mental bone defect repair, Journal of Biomedical Materials Research - Part A 81 (4)
Osteointegration of PLGA implants with nanostructured or microsized β-TCP par- (2007) 888–899.
ticles in a minipig model, J. Mech. Behav. Biomed. Mater. 40 (2014) 190–200. [182] J.P.M. Fennis, P.J.W. Stoelinga, J.A. Jansen, Reconstruction of the mandible with an
[158] A. Mina, A. Castaño, J.C. Caicedo, H.H. Caicedo, Y. Aguilar, Determination of physical autogenous irradiated cortical scaffold, autogenous corticocancellous bone-graft
properties for β-TCP + chitosan biomaterial obtained on metallic 316L substrates, and autogenous platelet-rich-plasma: an animal experiment, Int. J. Oral Maxillofac.
Mater. Chem. Phys. 160 (2015) 296–307. Surg. 34 (2) (2005) 158–166.
[159] H. Hu, X. Liu, C. Ding, Preparation and in vitro evaluation of nanostructured TiO2/ [183] B. Rai, K.H. Ho, Y. Lei, K. Si-Hoe, C. Jeremy Teo, K. Yacob, S.H. Teoh,
TCP composite coating by plasma electrolytic oxidation, J. Alloys Compd. 498 (2) Polycaprolactone-20% tricalcium phosphate scaffolds in combination with plate-
(2010) 172–178. let-rich plasma for the treatment of critical-sized defects of the mandible: a pilot
[160] H. Cao, N. Kuboyama, A biodegradable porous composite scaffold of PGA/β-TCP for study, J. Oral Maxillofac. Surg. 65 (11) (2007) 2195–2205.
bone tissue engineering, Bone 46 (2) (2010) 386–395. [184] E. Stodolak, A. Góra, Ł. Zych, M. Szumera, Bioactivity of fibrous polymer based
[161] S.S. Homaeigohar, A.Y. Sadi, J. Javadpour, A. Khavandi, The effect of reinforcement nanocomposites for application in regenerative medicine, Mater. Sci. Forum 714
volume fraction and particle size on the mechanical properties of β-tricalcium (2012) 229–236.
phosphate-high density polyethylene composites, J. Eur. Ceram. Soc. 26 (3) [185] R.Z. Legeros, C.A. Garrido, S.E. Lobo, F.M. Turíbio, Biphasic calcium phosphate
(2006) 273–278. bioceramics for orthopedic reconstructions: clinical outcomes, International Journal
[162] C. Kunze, T. Freier, E. Helwig, B. Sandner, D. Reif, A. Wutzler, H. Radusch, Surface of Biomaterials, vol. 2011 (2011).
modification of tricalcium phosphate for improvement of the interfacial compati- [186] T.L. Arinzeh, T. Tran, J. Mcalary, G. Daculsi, A comparative study of biphasic calcium
bility with biodegradable polymers, Biomaterials 24 (6) (2003) 967–974. phosphate ceramics for human mesenchymal stem-cell-induced bone formation,
[163] S. Bhumiratana, W.L. Grayson, A. Castaneda, D.N. Rockwood, E.S. Gil, D.L. Kaplan, G. Biomaterials 26 (17) (2005) 3631–3638.
Vunjak-Novakovic, Nucleation and growth of mineralized bone matrix on silk-hy- [187] H. Kim, Y. Koh, Y. Kong, J. Kang, H. Kim, Strontium substituted calcium phosphate
droxyapatite composite scaffolds, Biomaterials 32 (11) (2011) 2812–2820. biphasic ceramics obtained by a powder precipitation method, J. Mater. Sci. Mater.
[164] Z. Lu, H. Zreiqat, Beta-tricalcium phosphate exerts osteoconductivity through α2β1 Med. 15 (10) (2004) 1129–1134.
integrin and down-stream MAPK/ERK signaling pathway, Biochem. Biophys. Res. [188] D. Kim, T. Kim, J.D. Lee, K. Shin, J.S. Jung, K. Hwang, S. Yoon, Preparation and in vitro
Commun. 394 (2) (2010) 323–329. and in vivo performance of magnesium ion substituted biphasic calcium phos-
[165] J.M. Kim, T.S. Han, M.H. Kim, D.S. Oh, S.S. Kang, G. Kim, S.H. Choi, Osteogenic eval- phate spherical microscaffolds as human adipose tissue-derived mesenchymal
uation of calcium phosphate scaffold with drug-loaded poly (lactic-co-glycolic stem cell microcarriers, Journal of Nanomaterials 2013 (2013).
acid) microspheres in beagle dogs, Tissue Engineering and Regenerative Medicine [189] Y.M. Shin, J. Park, S.I. Jeong, S. An, H. Gwon, Y. Lim, C. Kim, Promotion of human
9 (3) (2012) 175–183. mesenchymal stem cell differentiation on bioresorbable polycaprolactone/biphasic
[166] M.T. Arafat, C.X.F. Lam, A.K. Ekaputra, S.Y. Wong, X. Li, I. Gibson, Biomimetic com- calcium phosphate composite scaffolds for bone tissue engineering, Biotechnol.
posite coating on rapid prototyped scaffolds for bone tissue engineering, Acta Bioprocess Eng. 19 (2) (2014) 341–349.
Biomater. 7 (2) (2011) 809–820. [190] Y. Weizhong, Y. Guangfu, Z. Dali, Y. Lijun, C. Linhong, Surface-modified biphasic
[167] C. Makarov, V. Cohen, A. Raz-Pasteur, I. Gotman, In vitro elution of vancomycin calcium phosphate/poly(L-lactide) biocomposite, Journal of Wuhan University of
from biodegradable osteoconductive calcium phosphate-polycaprolactone com- Technology-Mater 24 (1) (2007) 81–86.
posite beads for treatment of osteomyelitis, Eur. J. Pharm. Sci. 62 (2014) 49–56. [191] J. Lee, Y. Lee, N. Rim, S. Jo, Y. Lim, H. Shin, Development and characterization of
[168] A. Rakovsky, I. Gotman, E. Rabkin, E.Y. Gutmanas, β-TCP-polylactide composite nanofibrous poly(lactic-co-glycolic acid)/biphasic calcium phosphate composite
scaffolds with high strength and enhanced permeability prepared by a modified scaffolds for enhanced osteogenic differentiation, Macromol. Res. 19 (2) (2011)
salt leaching method, J. Mech. Behav. Biomed. Mater. 32 (2014) 89–98. 172–179.
[169] M. Bernstein, I. Gotman, C. Makarov, A. Phadke, S. Radin, P. Ducheyne, E.Y. [192] M. Ebrahimian-Hosseinabadi, F. Ashrafizadeh, M. Etemadifar, S.S. Venkatraman,
Gutmanas, Low temperature fabrication of β-TCP-PCL nanocomposites for bone Preparation and mechanical behavior of PLGA/nano-BCP composite scaffolds dur-
implants, Adv. Eng. Mater. 2 (8) (2010) B341–B347. ing in-vitro degradation for bone tissue engineering, Polym. Degrad. Stab. 96 (10)
[170] M. Bohner, L. Galea, N. Doebelin, Calcium phosphate bone graft substitutes: failures (2011) 1940–1946.
and hopes, J. Eur. Ceram. Soc. 32 (11) (2012) 2663–2671. [193] E. Lee, D. Kim, H. Lim, J. Lee, U. Jung, S. Choi, Comparative evaluation of biphasic cal-
[171] T. Lou, X. Wang, G. Song, Z. Gu, Z. Yang, Fabrication of PLLA/β-TCP nanocomposite cium phosphate and biphasic calcium phosphate collagen composite on
scaffolds with hierarchical porosity for bone tissue engineering, Int. J. Biol. osteoconductive potency in rabbit calvarial defect, Biomaterials Research 19 (1)
Macromol. 69 (2014) 464–470. (2015).
[172] S. Kobayashi, S. Yamadi, Strain rate dependency of mechanical properties of TCP/ [194] E. Lee, H. Lim, J. Hong, J. Lee, U. Jung, S. Choi, Bone regenerative efficacy of biphasic
PLLA composites after immersion in simulated body environments, Compos. Sci. calcium phosphate collagen composite as a carrier of rhBMP-2, Clinical Oral Im-
Technol. 70 (13) (2010) 1820–1825. plants Research, 2015.
[173] M. Vallet-Regí, J.M. González-Calbet, Calcium phosphates as substitution of bone [195] J. Amirian, N.T.B. Linh, Y.K. Min, B. Lee, Bone formation of a porous gelatin-pectin-
tissues, Progress in Solid State Chemistry 32 (1–2) (2004) 1–31. biphasic calcium phosphate composite in presence of BMP-2 and VEGF, Int. J. Biol.
[174] N. Siddiqui, K. Pramanik, E. Jabbari, Osteogenic differentiation of human mesen- Macromol. 76 (2015) 10–24.
chymal stem cells in freeze-gelled chitosan/nano β-tricalcium phosphate porous
scaffolds crosslinked with genipin, Mater. Sci. Eng. C 54 (2015) 76–83.

You might also like