Next Article in Journal
Experimental and Numerical Study of Adhesively and Bolted Connections of Pultruded GFRP I-Shape Profiles
Next Article in Special Issue
Design and Characterization of a Bioinspired Polyvinyl Alcohol Matrix with Structural Foam-Wall Microarchitectures for Potential Tissue Engineering Applications
Previous Article in Journal
Experimental Investigation of the Solid Conveying Behavior of Smooth and Grooved Single-Screw Extruders at High Screw Speeds
Previous Article in Special Issue
Fabrication, Characterization and In Vitro Assessment of Laevistrombus canarium-Derived Hydroxyapatite Particulate-Filled Polymer Composite for Implant Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Novel Trends into the Development of Natural Hydroxyapatite-Based Polymeric Composites for Bone Tissue Engineering

by
Diana-Elena Radulescu
1,
Ionela Andreea Neacsu
1,2,3,*,
Alexandru-Mihai Grumezescu
1,2,4 and
Ecaterina Andronescu
1,2,3
1
Department of Science and Engineering of Oxide Materials and Nanomaterials, Faculty of Applied Chemistry and Materials Science, University Politehnica of Bucharest, 011061 Bucharest, Romania
2
Academy of Romanian Scientists, 54 Independentei, 050094 Bucharest, Romania
3
National Research Center for Micro and Nanomaterials, Faculty of Applied Chemistry and Materials Science, University Politehnica of Bucharest, 060042 Bucharest, Romania
4
Research Institute of the University of Bucharest (ICUB), University of Bucharest, 050657 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(5), 899; https://doi.org/10.3390/polym14050899
Submission received: 25 January 2022 / Revised: 16 February 2022 / Accepted: 22 February 2022 / Published: 24 February 2022
(This article belongs to the Special Issue Advanced Biodegradable Polymer Scaffolds for Tissue Engineering)

Abstract

:
In recent years, the number of people needing bone replacements for the treatment of defects caused by chronic diseases or accidents has continuously increased. To solve these problems, tissue engineering has gained significant attention in the biomedical field, by focusing on the development of suitable materials that improve osseointegration and biologic activity. In this direction, the development of an ideal material that provides good osseointegration, increased antimicrobial activity and preserves good mechanical properties has been the main challenge. Currently, bone tissue engineering focuses on the development of materials with tailorable properties, by combining polymers and ceramics to meet the necessary complex requirements. This study presents the main polymers applied in tissue engineering, considering their advantages and drawbacks. Considering the potential disadvantages of polymers, improving the applicability of the material and the combination with a ceramic material is the optimum pathway to increase the mechanical stability and mineralization process. Thus, ceramic materials obtained from natural sources (e.g., hydroxyapatite) are preferred to improve bioactivity, due to their similarity to the native hydroxyapatite found in the composition of human bone.

Graphical Abstract

1. Introduction

Bone tissue engineering (BTE) has gained great interest in the last decade, as the frequency of degenerative diseases or tissue-damaging has increased [1]. Hence, the applications of BTE focus on the restoration of the native bone’s functionality by augmenting its regeneration rate with the aid of cells, osteogenic factors, or biomaterials. The perfect solution in improving bone recovery is represented by developing a biomaterial with good biological activity and mechanical properties that can maintain tissue functions [2,3].
Scaffolds are three-dimensional (3D) structures that need to meet specific criteria for BTE: biocompatibility, high surface area, appropriate mechanical properties for bone, enhance cell adhesion and lead to new tissue formation [3,4]. For the moment, the use of a single material as a scaffold is difficult to be further applied, as it cannot fully meet the requirements for BTE applications, as shown in (Figure 1). Therefore, the latest studies are currently concentrated on the development of composite materials, based on polymers and ceramics, that can mimic the composition of osseous tissue [5]. In this regard, biopolymers are known for their superior biocompatibility and good processability. Further, the mechanical strength is reduced, and the degradation rate is according to the osteogenesis rate, leading to pH modifications of the adjacent environment, generated by degradation products. On the other hand, even though bioceramics have good bone conductivity and biocompatibility, there is still a problem regarding the low toughness and significant brittleness that leads to damage and poor use reliability [5,6]. The limitations mentioned above have been addressed by developing biomaterials characterized by biomimicry and tissue regeneration capacity. The bioresorbability and bioactivity can be merged into composite materials to overcome the drawbacks of bioceramics and biopolymers. The obtained scaffold should be able to meet the complex requirements of tissue engineering, progressing from bone substituting materials to bone regeneration. These bioactive materials must have good biomechanical stability, biocompatibility, non-toxicity, and good processability to stimulate natural bone regeneration, by encapsulating inductive factors that promote bone healing [7,8,9].
Osseous tissue is a complex composite that can be reinforced with the aid of metallic, ceramic, composite, or polymeric materials. In this regard, bioceramic materials are the most promising materials for implantation due to their resemblance with bone composition. Moreover, calcium phosphates (CaPs) have been widely used as implantable materials due to their outstanding biocompatibility and similar chemical properties with hard tissues [10,11]. Additionally, CaP materials are essential for hard tissue reconstruction applications due to their outstanding biocompatibility, bioactivity, non-immunogenicity, non-toxicity, and non-inflammatory behavior [12]. Until this moment, numerous types of hydroxyapatite-based materials (HA-based materials), such as pure HA, HA/polymer composites, and ion-doped HA, have been investigated and designed, although some disadvantages still exist. Therefore, there is still a long way to go before obtaining satisfying HA-based materials [13,14,15].
Nowadays, already chemically synthesized HA has been used for BTE, but its poor durability and stability have reduced its feasibility into biomedical applications. These drawbacks of physically and chemically synthesized HA led to the use of natural biowastes to synthesize HA. Recently, a unique and efficient strategy for preparing functionalized biomaterials with complex structures obtained from natural origins has been demonstrated. Thus, organic food waste, such as bovine/fish bones, seashells, and eggshells, could have perfect potential in obtaining HA, with very high availability that could enhance orthopedic applications [16,17]. Further, to improve the bioactivity of the developed materials, scaffolds can be obtained by combining biodegradable polymers (for adaptive degradation and biocompatibility) and bioceramics (for strength and bioactivity) successfully [18,19].
This study aims to present the most used polymers in BTE, that can be further combined with ceramic materials. The main highlight of this research is represented by the use of HA obtained from natural sources to increase the bioactivity of composite scaffolds. Further, the required properties of composites for the proper functionality of scaffolds, and recent applications, will be briefly described to select the ideal material for orthopedic applications.

2. Polymers Applied in BTE

Over the years, it was concluded that the properties of biomaterials must be thoroughly investigated before considering them for BTE applications. These include chemical composition, structural and biological characteristics, degradation rate, and processability. Thus, polymeric scaffolds presented great interest due to their distinguishing features, such as the capacity to provide osteoinductive support to the transplanted or native cells at the injured zone, harbor multiple spatio-temporal cues, degradation rate similar to osteogenesis rate, and capacity to integrate with native tissue throughout wide chemical modifications. Considering their source of provenience, polymers applied in BTE can be classified as synthetic or natural polymers [20].
Several natural biopolymers, such as collagen (Col), chitosan (CS), silk fibroin (SF), cellulose, alginate, and hyaluronic acid, have been intensively researched, considering their present incorporation into other nanomaterials due to their increased bioactivity. On the other hand, synthetic polymers have also been presented in this review, including poly(ε-caprolactone) (PCL), polyethylene glycol (PEG), poly (lactic-co-glycolic) acid (PLGA), poly (lactic acid) (PLA), and poly(vinyl alcohol) (PVA), due to their enhanced mechanical properties. Both synthetic and natural polymers are either coated or combined as nanomaterials to form nanocomposite, nano surfaces, nanofibers, and other nanostructures used in BTE [21,22,23,24,25].

2.1. Natural Polymers

Commonly, natural polymers are obtained from plants, animals, and microorganisms. These polymers are classified into three main categories: polysaccharides, polypeptides, and polyesters. Therefore, various exclusive characteristics of these polymers (Table 1) for tissue regeneration applications are represented by their superior stability, structural design, tailorable solubility, 3D geometry, excellent biocompatibility, low immunogenicity, antigenicity, cytocompatibility, and frequently specific cell targeting.
CS is known as one of the most favored materials in BTE due to its biodegradability and non-toxic nature [38]. There are numerous similarities between CS and the native glycosaminoglycans’ structure, which improves the rate of bone regeneration. Thus, Cunha et al. noted that amino and hydroxyl functional groups enable chemical modifications, increasing the potential of CS-based nanocomposites, which have gained significant attention in this domain [39,40]. Such an application can be seen in Figure 2, illustrating the inhibition phenomenon of bacterial attachment.
Additionally, CS is incorporated without difficulty into films, gels, membranes, metallic or non-metallic nanoparticles, as well as composites. This material interacts with negatively charged microbial membranes and, at the same time, exhibits antimicrobial properties against pathogenic microbes. Considering all these aspects, we must also consider the low mechanical strength, unable to support load-bearing applications [41]. Further, Lavanya et al. mentioned that CS can be easily applied as a bone substitute material due to its thermo-responsive capacity and non-toxic abilities. At physiological temperature, injectable CS hydrogels undergo a sol-gel transition, providing improved cellular activity and bone regeneration. Also, these hydrogels facilitate the incorporation of biomolecules, nanoparticles, or polymers, expanding the hydrogel’s properties [42].
Another good example is represented by collagen (Col), an essential component of the bone native extracellular matrix (ECM). Col can be easily obtained from food waste, such as bones, scales, and skins. Further, Col is a natural bioactive polymer with good biocompatibility, facilitating the proliferation and adhesion of osseous cells and reducing antigenicity. In addition, bones’ ECM is composed mainly of Col that nourishes osseous cells and aids the proper transport of nutrients. Even though Col’s biocompatibility enables cell proliferation and adhesion, it also exhibits increased swelling and poor mechanical strength [21]. These limitations are the main reasons to further study the use of Col with various composite biomaterials. Rethinam et al. presented the possibility and applicability of this polymer in the biomedical field by combining it with synthetic polymers or ceramic materials. This research aimed to improve its mechanical properties and bioactivity [43]. Moreover, Marques et Al. presented the applicability of Col-based bioinks for 3D printing in BTE (such as cartilage, bone, and osteochondral regeneration). They have shown that Col is maintained in a liquid state at low temperatures, forming a fibrous structure by increasing the temperature. The solution for this drawback is to combine Col with other materials, thus, improving the rheological properties of bioinks [44].
Gelatin is known as another natural polymer, a processed form of Col, also applied in bone regeneration, including cell attachment of arginine glycine-aspartic acid (RGD) sequences. Even though it is easily biodegradable, this polymer has poor mechanical properties, which are necessary for this type of application [45]. Nevertheless, gelatin can form a thermally reversible network in water, ensuring the formation of a cross-linked network with satisfactory thermal and mechanical stability [46]. Considering these aspects, the mechanical properties of gelatin can be improved by combining it with other polymeric materials or ceramics. To prove this, Luetchford et al. developed scaffolds based on SF and gelatin. After thorough investigation, the obtained scaffolds exhibited improved mechanical strength, and good adhesion of rat mesenchymal stem cells after culturing in an osteogenic differentiation medium [47]. Moreover, Mishra et al. obtained scaffolds based on gelatin and polyvinylpyrrolidone (PVP). The material presented high porosity and interconnected architecture, ensuring a favorable environment to support and aid proliferation and migration using C3H10t1/2 cells [48].
Another preferred natural polymer for BTE applications is hyaluronic acid, which has been intensively used for healing applications due to its good anti-inflammatory capacity. This polymer is well known for its non-immunogenicity, biocompatibility, and biodegradability, promoting cell interaction and proliferation [49,50]. Furthermore, Sionkowska et al. presented the applicability of hyaluronic acid in BTE, drug delivery systems, or coatings [51]. Further, Fang et al. mentioned that hyaluronic acid scaffolds influence several behaviors of stem cells, such as proliferation, adhesion, migration, and differentiation, binding them to specific signaling pathways and cellular receptors. Therefore, researchers focused on developing stem cell- hyaluronic acid systems to improve the efficacy of the composite material [52]. The incorporation of growth factors has been also demonstrated by Park et al. The researchers showed that bone morphogenetic protein-2 (BMP-2) induced the osteogenic differentiation of human dental pulp stem cells and displayed prolonged viability of growth tissue [53].

2.2. Synthetic Polymers

Compared with natural polymers, synthetic polymers do not possess the required biological characteristics, which, despite avoiding the generation of an immune response, lead to low capacity, causing detailed cellular phenotypes. Therefore, by incorporating biologically active molecules into synthetic materials, the materials could be designed to fabricate biomimetic scaffolds with well-defined composition, as shown in Table 2. Alternatively, synthetic polymers can effortlessly modify their degradation time, mechanical properties, and molecular weight for BTE applications [23,54].
Furthermore, synthetic polymers do not always generate cell adhesive surfaces and require surface adjustments for suitable growth in 3D architectures. Therefore, the inclusion and encapsulation of cells in coarse conditions reduces cell–biomaterial interactions, limiting its use in bone tissue regeneration [42]. A synthetic polymer appreciated in BTE is PLA, a linear aliphatic thermoplastic material derived from the fermentation of renewable sources rich in starch (e.g., potatoes, corn, or agricultural by-products), making it accessible and available for biomedical applications. Similar to other polymers mentioned above, it has poor mechanical properties, limiting its medical field applicability. In this direction, inorganic particles or other polymers are often incorporated to improve organic polymers’ mechanical properties for bone regeneration [67]. For example, Bal et al. developed scaffolds based on PLA-PEG to properly sustain the incorporation of BMP-2. Still, as the mechanical properties were not sufficient, HA has also been considered to be introduced into the system. The results show the success of its application in BTE [68].
By observing Table 2, it can also be established that PEG is an effective polymer, considering cell adhesion, growth, and proliferation, with good biocompatibility and non-immunogenicity [69,70]. This synthetic polymer has been intensively used in various tissue engineering applications and drug delivery systems. In this regard, Bose et al. mentioned that PEG is preferred in osseous regeneration, as the material can alleviate poor tissue adsorption and rapid metabolism. Further, even at low concentrations, PEG forms a dense, protective, hydrophilic coating, with long flexible chains, leading to a decrease in surface hydrophobicity to the drug molecule [71]. This polymer can be combined with other materials to improve its efficacy in orthopedic applications. Dethe et al. demonstrated the performance of PCL-PEG scaffolds to form fast and reversible physical gelation performance with good bioactivity [72]. Moreover, Bai et al. introduced roxithromycin into a similar system, to prevent inflammatory response at the implanted site. The biological assessments proved good viability and cell growth capacity [73]. On the other hand, Etminanfar et al. fabricated scaffolds containing both polymer PCL-PEG and HA. The obtained material exhibited good biocompatibility and porosity [74].
Another polymer, widely applied in the biomedical field, is represented by PVA, due to its inherent low toxicity, high hydrophilicity, and biocompatibility. Hence, Chen et al. proved that scaffolds based on PVA with β-tricalcium phosphate (β-TCP) provided good thermal processability, while the material presented interconnected channels. Further, in vitro studies demonstrated good biocompatibility, cell proliferation, and adhesion [75,76]. HA is one more ceramic often combined with PVA. For example, Salim et al. obtained composites based on PVA and HA, which significantly enhanced the mechanical and thermal stability, delaying the swelling index. Moreover, the addition of a natural polymer (CS) into the system enhanced the antimicrobial activity and swelling index [77].
Another alternative of synthetic polymer is represented by PCL, a semi-crystalline aliphatic polyester, exhibiting good biodegradability and mechanical properties. Compared with other materials, PCL exhibits great flexibility at a low melting temperature and a slow biodegradation rate for long-term application [78]. Even with these advantages, and it being recommended in tissue regeneration applications, this polymer exhibits poor cell adhesion, proliferation, migration, and low hydrophilicity [79]. To address these issues, PCL can be combined with numerous organic or inorganic materials to surpass the abovementioned limitations. In this regard, Wang et al. obtained a scaffold based on PCL and β-TCP to be further applied for 3D printing. The complex structure was ideal for cell growth promotion, permitting cells to enter through the pores, grow, and progressively cover the pores, improving the osteogenic performance of the scaffold [80]. The same results were also supported by Gatto et al., by developing a scaffold based on PCL that also contained a ceramic material (HA) [81]. Moreover, Cestari et al. demonstrated that by replacing the commercial HA with one derived from seashells, the composite presented enhanced the bioactivity [82].
PMMA is a well-known polymer applied in bone cement, widely used to fixate implants in trauma treatments or orthopedic surgeries [83]. This thermoplastic polymer presents numerous advantages, such as quick polymerization, fast patient recovery, and facile application, but it also presents significant drawbacks. The main disadvantages are the poor mechanical properties and low bioactivity [84,85]. Furthermore, PMMA is a bioinert material and prevents osseointegration and chemical bonding at implanted sites. To surpass these disadvantages, this polymer has been combined with HA, increasing the interfacial shear strength and bioactivity of the developed material [86]. Moreover, Al-Sherify et al. demonstrated that the composite material had a decreased corrosion rate compared with simple polymeric scaffolds [87]. Other researchers focused on other routes, by combining PMMA with other polymers (natural or synthetic). Thus, Atila et al. presented the fabrication of SF/PMMA membranes. The obtained composite showed improvement, considering the mechanical properties and suitable porous architecture. Moreover, biological investigations exhibited good cell attachment and proliferation [88].

3. HA-Based on Natural Sources Used in BTE

Materials based on CaP attracted the attention of researchers for their effective use in hard tissue engineering. In this direction, HA is a major component for developing coatings, presenting properties such as non-toxicity, osteoconductivity, and increased biocompatibility. Further, HA can be synthesized from chemical solutions containing calcium and phosphate ions. In numerous studies, chemically synthesized HA has been used for bone replacement. Their nanostructures (e.g., nanoparticles, nanotubes) have been obtained through a simple precipitation method. Still, synthetic HA does not show a substantial similarity to native HA, considering its chemical composition, resistance, and bioactivity. In this regard, researchers have been trying to discover the use of natural materials as a substitute source for HA. Several research papers have demonstrated the successful preparation of HA using natural sources, as shown in Figure 3, such as corals, fish scales, eggshells, fish bones, seaweed, animal bones [89,90,91].
Biomineralization is a complex process, widespread in nature, that leads to the formation of a variety of inorganic minerals from living organisms [9]. The minerals produced from biological pathways, with stimulating properties and precise hierarchical structures, are organic–inorganic composites, designed by a self-assembled process in mild conditions (such as temperatures below 100 °C or aqueous solutions) [92]. Numerous studies have shown the potential of converting food waste into extremely valuable bioceramics, using simple and efficient approaches. Different synthesis techniques have been extensively established using animal bones, eggshells, fish bones, oyster shells, or corals. These types of ceramics are presented in Table 3. Further, Rocha et al. demonstrated the conversion of natural aragonite from cuttlefish bone in HA, through the hydrothermal method [93,94]. HA obtained from biological sources generally retains various properties from the precursor materials, such as chemical composition and pore structure. In this direction, it possesses a rich calcium concentration and beneficial trace elements, such as Mg and Na, for bone growth [95].
In addition, bio-waste, including eggshells, animal bones, and marine shells, have shown great potential in this path. The use of bio-wastes is economical, helps sustainability, and adds value [114,115]. In recent years, the possibility of preparing HA from fish bones has been demonstrated by simple calcination. The HA obtained through this process has a structure and morphology very similar to human bone. Furthermore, by subjecting it to a high-temperature treatment, all pathogens and organic components from the source are removed. Furthermore, HA extracted from natural sources generally contains traces of ions, such as Mg2+, Na+, Zn2+, and K+. These ions enhance the usefulness of natural HA, promoting the regeneration processes and bone formation [116].
Cestari et al. obtained nano-HA from different sources (eggshells, cuttlefish bone, and mussel shells). HA was synthesized via wet mechano-synthesis and drying in an oven. The powders were further mixed with phosphoric acid (~85% H3PO4), or ammonium phosphate dibasic ((NH4)2HPO4), to achieve a Ca/P ratio of 1.67. The results of their study are presented in Figure 4. After a thorough investigation, researchers concluded that HA obtained from food wastes exhibited good cell adhesion and proliferation after seeding through confocal microscopy. SEM micrographs of the obtained materials also show that most pores have dimensions ~1 µm; thus, it can be considered that they are not perceived as holes by the cells, dimensions of which are up to100 µm. Nevertheless, the micrometric pores are beneficial for cell–bioceramic interaction [117].
All prepared samples proved good cellular adhesion with no cytotoxic effect. Moreover, HA derived from eggshells and sintered at 900 °C encouraged the greatest cell adhesion pattern, comparable with pure HA, without forming clusters. At the same time, those from mussel shells and cuttlebones also have the capacity to support cell adhesion, while amorphous calcium carbonate-based HA displayed lower cell-to-cell interconnection and round-shaped cells.

3.1. Recent Strategies for Obtaining HA from Natural Sources

Recently, new, sustainable methods for obtaining HA with particular properties have been extensively studied. Although numerous synthesis methods were developed, HA’s preparation with specific properties remains a challenge, due to the risk of appearing toxic intermediates during synthesis [118]. Hence, several chemical methods have been chosen to obtain HA in this field: hydrothermal techniques, precipitation, wet chemical precipitation, microwave processing, and hydrolysis. HA obtained from marine species has great economic influence over commercial HA due to the low-cost synthesis technique. Moreover, natural sources possess the advantage of taking over certain natural material properties. Nowadays, numerous compounds are isolated from marine organisms and are applied in bioactive ingredients on medical devices [119,120,121,122].
Different synthetic approaches have been applied to prepare HA-based composites, varying their properties according to the application’s necessities. Solid-state synthesis is a well-known technique, used to obtain inorganic oxide materials, because it allows excellent control of the chemical stoichiometry. However, challenges arise when the formation of the phase composition of the anticipated product is not compatible with the thermal treatment and high temperatures to obtain a complete solid-state reaction [123]. Depending on ionic strength, precipitation conditions, pH, and temperature, HA is obtained with a Ca/P molar ratio similar to natural bone, making it an ideal candidate for clinical applications [124]. However, these methods can be very complex and time-consuming. HA synthesized from animal bones presents the advantage of retaining certain raw material properties, such as composition and chemical structure [125,126]. The ball milling method was also preferred to mill the fish bones before performing the calcination process. The resulting powder was produced within a well-established time frame, which can influence the purity of the ceramics obtained [127].
In this direction, Yanhong et al. suggested an ecological and green method to extract HA from fish scales with a deep eutectic solvent (e.g., choline chloride/1,4-butanediol). The ideal extraction conditions were as follows: the extraction temperature of 65 °C, in 2 h, and a specific solid:liquid ratio of 1:15 (g:g). All these conditions lead to a HA extraction of 40.58%. The results showed that the extracted HA presented a heterogeneous morphology, a mixture of phosphates of phosphoric acid and carbonic acid, with excellent thermal stability and high purity. The obtained HA reported a good adsorption capacity for metal ions, such as copper, lead, zinc, and silver [128]. Wu et al. prepared three synthesis methods from three different fruits and eggshells. After performing the hydrothermal reactions at 150 °C for 24 h, aggregate particles with a rod-like or needle-like nanostructure were obtained. Needle-like nanostructures were transformed into rods as the hydrothermal reaction time increased by 72 h.
Moreover, Wu et al. concluded that HA obtained with pomelo peel extract showed the best structure, having the most similar crystalline structure of HA, present in natural bone [129]. Further, Karunakaran et al. reported the influence of PVP. This polymer is used as an organic modifier for HA synthesis. Seashells have been used for HA synthesis as a source of calcium. The microwave-assisted synthesis method was applied to obtain various sizes and complex mesoporous morphologies that can be subsequently applied in implantation. Also, the presence of PVP improved the microwave-stimulated approach, which effectively regulated pore volume, surface area, size, and dimensions by regulating the formation of HA crystals [130].

3.2. HA Obtained from Eggshells

This natural method of obtaining HA nanorods is not only an optimal and energy-saving process, but it also excludes high-energy calcination and uses exogenous surfactants to assist the synthesis of HA from shells or oysters as a source of calcium. This obtainment method improved CO32- ions’ loading capacities and enhanced the poor conversion performance of calcite into HA, to some extent [131,132]. The use of shells proved advantageous through the impurities derived from the organic matrix in CaCO3. The residual organic matrix of CaCO3, derived from mollusk shells, offers biocompatibility and bioactivity to the composite, promoting bone regeneration [133,134]. The quantity of HA improved significantly, and the diffraction peaks’ intensity increased through the ball grinding process. A longer grinding time leads to the improvement of these two properties. Studies have shown that stoichiometric HA powder can be effortlessly obtained through ball milling without further heat treatment. In this regard, Lala et al. showed that 10 h grinding is not enough to obtain pure HA without sintering aids [135].
Eggshells are also a natural mineral source (e.g., calcium carbonate) that can be extracted and used to obtain HA. From these resources, it offers bioactivity and biocompatibility that allow osteoblast cells the capacity to adhere and proliferate in different periods [136,137,138]. Francis et al. showed the applicability of eggshells. They are generally subjected to grinding to facilitate membrane removal, either by heat treatment or by bleaching [139]. Wu et al. described the development of HA powder using dicalcium phosphate dihydrate (DCPD) and eggshell powders through ball milling and following heat treatment. HA phase formation can be started by sintering the sample at 1000 °C for 1 h, while the pure HA phase is obtained by sintering the ground sample after 10 h [140]. Also, Ho et al. proposed eggshells as a possible technology for recycling waste management of materials. Eggshell is also a possible material in CaP ceramics synthesis for biomedical applications. Some trace elements do not modify the elementary crystallographic properties of HA but may improve the general biological performance of the material used for implants [141]. Alhussary et al. applied nano-HA from seashells and eggshells to the damaged mandibular bone of rabbits and reasoned that HA obtained from eggshells has an improved osteoinductivity and porosity than the one obtained from seashells [142]. Furthermore, Waheed recommended that eggshell scaffolds have superior compressive strength, increased mineralization, and improved osteogenic differentiation potential compared to commercial HA [143].

3.3. HA Obtained from Seashells

Nanometer-sized ceramics of marine origin provide an abundant source of new materials in BTE but are also a source of stimulation for developing novel biomimetic composites. Natural corals are composed of an organic matrix that presents calcified nodes. The shells consist mostly of calcium carbonate, which forms multilayered microstructures and small organic components (1–5 wt.%), located mainly in the inter-crystalline boundaries. Despite its composition and special composite microstructure, shells show an improvement in strength by three orders of magnitude over synthetic calcium carbonate. [144].
The oyster shells represent another natural source, composed of pure aragonite, which is crystallized in an organic matrix. Vecchio et al. performed a classic hydrothermal conversion process from sea urchin spines into TCP. The obtained products exhibited good bioactivity and osteoconductivity [145]. Researchers demonstrated that rod-shaped HA crystals are favored in the attachment of proteins, especially in the binding of vitamin D, applied in BTE, due to its strong adhesion to osteoclasts and similarity with the native HA structure. Moreover, HA synthesized from microwave-assisted treatments is different from conventional methods. Heat is initiated uniformly into the material for this production method, with a higher heat transfer rate and spatial temperature distribution. Once the sample nears the heating source, the size, shape, and chemical structure develop high sensitivity to heating [146].
Mitran et al. studied the bioactivity of CaP derived from Mytilus galloprovincialis seashells and dolomitic marble. The obtained materials were investigated on MC3T3-E1 pre-osteoblasts, regarding morphology, viability, cell adhesion, proliferation, and differentiation, to investigate their potential for BTE (as shown in Figure 5). Results exhibited suitable cell adhesion and high viability, with no distinguishable differences in the morphological features. Moreover, the pre-osteoblast proliferation on naturally derived CaPs showed a substantial increase compared with commercial HA. Nevertheless, seashell-derived CaP displayed a higher efficiency in promoting pre-osteoblast differentiation and further exhibited its potential in BTE [147].

3.4. HA Obtained from Fish Bone

While synthetic materials have been known for their use in medical applications, natural materials are nowadays acknowledged as a potential candidate for new biomedical applications. Given that fish bones are abundant materials, rich in CaP, their exploitation has certainly been beneficial in reducing disposal costs and reducing the risk of environmental pollution, in addition to their biomedical benefits [12,148]. Numerous species, such as codfish (Gadus morhua), barramundi (Lates calcarifer), carp (Cyprinus carpio), cuttlefish (Sepia Officinalis), golden (Sparus aurata), conger (Conger conger), perch (Dicentrarchus labrax), flatfish (Heterosomota pleuronectiformes), sardines (Sardina pilchardus), anchovies (Engraulis encrasicolus), sharks, swordfish (Xiphias gladius), tuna (Thunnus albacares), have been applied as raw materials to produce HA and β-TCP [12]. For example, Ozawa and Suzuki prepared HA from Japanese sea bream by calcination at 1300 °C [149]. Calcination time, calcination temperature, extraction method, and bone nature are aspects that influence the final properties of CaPs, such as morphology, Ca:P ratio, size distribution, phase purity, and specific surface area [150].
Moreover, CaPs derived from fishbone have been shown to accomplish functions similar to commercial ones. Fish bones would make a significant input to developing biomaterials for BTE. They possess great potential in developing several products that could be used in many applications, including tissue engineering, environmental remediation, and drug delivery [12,151]. Buraiki et al. studied the synthesis of pure HA from fish scales. The results confirm the synthesis of HA powder, which could be further used for bone regeneration [152]. Another good example is represented by the study performed by Piccirillo et al. The research group examined the annealing of codfish bones at different temperatures of 900, 950, 1000, 1100, and 1200 °C. The final materials obtained were composed of β-TCP and HA. Hence, the temperature is another property that greatly influences the process of obtaining ceramics. The β-TCP content increased by increasing annealing temperature [153]. Goto and Sasaki also examined the influence of the temperature (400–1000 °C) on bone calcination. It was demonstrated that the diameters of the crystallite samples increased, whereas their specific surface area decreased as the calcination temperature increased [154].
HA powder obtained using heat treatment on fish scales (Catla catla) retained, in its composition, natural trace elements, such as magnesium (Mg) and strontium (Sr), with a concentration of ~3.5 times more Mg and ~10 times more Sr, related to chemically synthesized HA powder [112]. Fish-derived HA crystallite of nanometric dimensions was investigated compared to chemically synthesized HA. In this regard, the average crystallite size and HA crystallinity derived from fish increased with the calcination temperature. From the point of view of biological analyses, in vitro cellular materials’ interactions with osteoblast cells showed no toxicity [112,155]. The porosity of HA from marine resources is higher compared to ceramics obtained by chemical methods. The increased porosity is considered a great advantage of its application in biomedicine. The values of synthesized HA hardness are in the range of human cortical bone [156].
Popescu-Pelin et al. presented the applicability of fish bone-derived BCP for obtaining thin films via pulsed laser deposition (PLD). The targets were prepared from sea bream and salmon fish bones for this experiment. In the obtained coating, a significant fraction of β-TCP was observed in their composition besides HA, which contributes to improved biocompatibility and solubility for BTE. The deposited structures were biocompatible and presented no signs of cytotoxicity in human gingival fibroblast cells, demonstrating their increased osseointegration (Figure 6). Further, the composite inhibited microbial adhesion and biofilm development. Moreover, protection against bacterial colonization of Escherichia coli was established for at least 72 h, due to the release of their natural elements (i.e., Mg, Na, S, Si) from fish bones [157].
Moreover, to support the use of fish bones, Bas et al. investigated the in vitro biocompatibility of Salmon-derived BCP, by performing the characterization on osteosarcoma (Saos-2) cells. All results have shown that the obtained material had no cytotoxicity and promoted their applicability [158]. In vitro biocompatibility was also examined in the behavior of osteosarcoma (Saos-2) cells, indicating that the CaP bioceramics had no cytotoxicity effect. Salmon-derived biphasic calcium phosphates (BCP) have the potential to contribute to the development of bone-substituted materials.

4. Composite Materials: Polymers and HA Based from Natural Sources

It is well known that human bone consists mainly of HA crystals and collagen fibers. Synthetic CaPs have attracted considerable attention due to their functional and structural similarity to natural HA. Their superior biocompatibility and excellent osteoconductivity have increased their popularity in recent decades, but they did not exhibit osteoinductivity. Some approaches implemented to bestow osteoinductivity include ion substitution, combining them with various polymers to obtain a composite, or adding bioactive substances [159]. BTE applications are designed to repair defects by introducing scaffolds into the human body, which are consequently replaced by novel tissue. For these applications, nanocomposite biomaterials are preferred to incorporate a matrix structure with good bioactivity, easily resorbable [160,161].
In this regard, hydrogels enable the mineralization throughout bone regeneration processes with bioactive inorganic materials. Nanohydroxyapatite (nHA), CaPs, and bioglasses are among the most extensively used bioactive inorganic materials to be doped into hydrogels [162]. An example of such applications is presented by Deb et al., in which the team developed scaffolds composed of HA from fishbone into a PMMA matrix. After thorough investigation, the obtained scaffolds proved to meet the physiological requirements but also necessary in vitro bioactivity [163]. Another composite is presented by Pon-On et al., by obtaining PLA/CS composite scaffolds with HA from fish scales, with good bioactivity and cell adhesion [164].

4.1. Properties of Composite Materials for BTE

The mechanical properties of materials (Figure 7) depend on their specific surface, total pore volume, pore size, pore distribution, and porosity. A ceramic with high porosity, inhomogeneous pore size distribution, large pore size, superior total pore volume, and large specific surface area will exhibit low strength. Pores decrease the strength of the ceramic, reducing the cross-sectional area when a load is applied. Thus, reducing the porosity, specific surface area, pore size, and total pore volume will improve ceramic strength [165,166].
As observed in Figure 7, the implant’s surface influences the bacteria’s ability to adhere to the substrate. The application of surface coatings is the optimal technique to prevent bacterial attachment and increase the bioactivity of the implant. Moreover, numerous studies have also demonstrated that the efficiency of the applied coatings is influenced by the materials’ properties used in their development. One of these properties is surface morphology. The roughness of the surface particularly affects the behavior of bacterial adhesion. Even though a polished surface presents the capacity to reduce bacterial adhesion, roughened surfaces offer encouraging biological properties with good biomechanical stability. Surface hydrophilicity also has a major impact on bacterial adhesion. As the roughness is reduced from sub-micron to nano-scale dimensions, the surface properties are progressively modified from hydrophobic to hydrophilic, becoming unfavorable to the adhesion of S. aureus, which is hydrophobic. Moreover, the bacterial adhesion, respectively, the bond between the bacteria and surface deteriorates [168,169].
A promising method to prevent biofilm formation from an early stage is the use of antimicrobial surfaces. Surfaces doped with antimicrobial elements (such as copper, silver, and zinc) are developed to surpass antibiotic delivery deficiencies [170,171,172]. The latest tendency in biomaterial science is the combination of polymeric and ceramic materials. The main disadvantage of ceramics is their fragility. Flexibility can be improved by incorporating inorganic compounds into polymeric matrices. Current research involves the development of different composite materials by incorporating HA into the system. It is well known that HA has great compatibility with osteosarcoma cells. However, it is still necessary to design polymer matrices that also show improved biocompatibility in contact with human cells [173]. To induce the preferred mechanical strength, HA synthesis often requires a high temperature for the sintering process, so ceramic materials, developed by processing at low temperatures, are very fragile and unsuitable for tissue engineering applications. The use of ceramics (e.g., HA) in scaffold development has two major disadvantages: the deficiency of degradability in biological systems and limited processing techniques for manufacturing scaffolds (Table 4).
In general, for biomedical applications, the benefits of ceramic materials have many more advantages than limitations [179]. The ideal property of implants is to exhibit antimicrobial activity, to reduce the treatment period, showing a local antibacterial effect and improving its effectiveness, thus, reducing the need for systemic therapies and their side effects [180].

4.2. Applications of HA-Based Composites in BTE

Biomaterials applied in the biomedical field have many chemical and physical properties. Biomedical applications in the surgical and medical domain require a thorough understanding of their mechanical properties. An exhaustive investigation of the mechano-tribological properties of implanted materials is necessary to determine the reliability in the long term [181]. With the progress of research in recent years, researchers focused on providing solutions to combat biofilm and, at the same time, to maintain osseointegration [182]. There are various methods to address this issue, especially the fabrication of 3D scaffolds, micro or nanoparticles, and the in situ formation of hydrogels that are applicated to design substitutes of the natural tissue, encouraging regeneration [175]. Nevertheless, it is well known that composite materials containing ceramic and polymer constituents have combined properties, such as bioactivity, biodegradability, and flexibility from polymers, but also mechanical strength and toughness provided by the ceramic phase. Except for their biocompatibility, materials applied in BTE also need to support cell viability, proliferation, and differentiation. Moreover, the porous structure ensures satisfactory space to promote bone ingrowth. The macropores facilitate the osteoid formation and mineralization, increasing the osteoblasts and osteoprogenitors migration into the scaffold, while the interconnected microporosity enhances vascularization and nutrient distribution throughout bone reconstruction [183].
Until now, little evidence has been presented regarding the positive properties of HA from natural sources on human health, because of the absence of comprehensive investigation and experimental research. Few researchers have effectively obtained HA-based biomaterials from various food wastes, such as scales, bovine and fish bones, eggshells, and seashells, and performed the corresponding characterization analyses [184]. Recent advances in biomaterial synthesis, especially those applied in the development of medical applications, to prevent bacterial infections and improve osteogenesis, have reawakened natural HA’s research interest. [185,186]. Mahmoud reported alginate/nano-HA composites’ preparation, using fishbone as a natural source for HA extraction. The 3D porous scaffold was manufactured to improve biodegradability and osteoconductivity. The animals studied to assess biological effects survived without extensive or local complications, which means that the implanted materials did not generate any histopathological limitations or non-compatibility with bone tissue [187]. Moreover, Yadong et al. found that combining HA and Col improves the differentiation of osteoblasts, due to their excellent bioactivity, biocompatibility and osteoconductivity. Therefore, the Col/HA structures showed a microstructure and composition similar to natural osseous tissues, attracting great potential candidates [188].
In another study, the twin-screw blending solvent casting process obtained biofilms. The polymer chosen for this application was PLA, a biodegradable polymer, while HA is obtained from fish bones. The obtained PLA/HA biofilms were characterized by several characterization methods, such as XRD, FESEM, DSC/TGA, FTIR, and contact angle. The results obtained were very encouraging, offering the possibility to fixate medical devices to bone tissue. Moreover, XRD confirms that the presence of HA generates a higher crystallization of the films [189].
Further, novel CS/CaP-based composites have been developed to regenerate bone tissue ad were studied by Kara et al. Initially, fish residues are physically fragmented and effectively decellularized into microparticles, and further incorporated into the matrix of CS. The results presented extremely porous composites, in which the ceramic microparticles were incorporated as fibrillar structures, with diverse morphologies into the CS surface. After 14 days, it could be observed that the addition of natural ceramics did not modify the composite’s mechanical properties. Moreover, their morphology changed the polymeric matrix structure and surface roughness, enhancing cell attachment [190]. In another study, Humayun et al. developed an antimicrobial coating, based on CS, combined with halloysite nanotubes, as an ideal solution to avoid biofilm formation, because of the innate biocompatibility of halloysite and ability to offer controlled drug release. As a bioactive principle, gentamicin is used to enhance its further biocompatibility [191].
Another innovative method is presented by Kang et al. The researchers obtained calcium-binding polymer-coated poly(lactide-co-glycolide) microparticles, which incorporate a quorum sensing (QS) reagent (Furanone C-30), and attached them to the HA surface to prevent biofilm formation. After thorough analysis, they showed a good adhesion of the microparticles on the surface of the material and demonstrated that the microparticles containing furanone C-30 can inhibit biofilm formation, generated by S. mutans, up to 18 h [192]. HA-coated titanium implants can induce the complete osseointegration of bone tissue [193]. These coatings on implant surfaces can be obtained through numerous techniques, such as electrophoretic deposition, plasma spraying, sol-gel deposition, sputtering deposition, pulsed laser deposition, spin-coating, ion beam assisted deposition, and many other techniques [183,193].
Another innovative application is represented by 3D printing technology, which recently engrossed extensive attention in BTE. In this direction, HA particles and multipotent mesenchymal stromal cells are directly added to the Col solution and then applied to different molding methods. Li et al. mentioned the success of combining mineral ions with Col, followed by mineral formation through an in-situ deposition. This technique can be applied to control the regular arrangement of fibers in scaffolds [194]. Further, Lowe also studied the possibility of forming a new composite based on salmon bone HA, functionalized with fucoidan and CS as 3D scaffolds. Nano-HA was isolated from the salmon fishbone through alkaline treatment and further uniformly dispersed in the polymer matrix. The composite pore size was 23–354 μm, suitable for accompanying growth factors and nutrients [195].
All these studies presented the possibility to use composite materials in tissue engineering applications to improve osseous regeneration, without exhibiting any harmful effects.

5. Future Perspectives and Challenges in BTE

BTE has progressed extraordinarily considering materials synthesis, application techniques, and cell engineering, due to the increased demand in the medical field. The main goal in this domain was to improve the treatment of bone defects without applying autologous bone grafts. Until this moment, the ideal material to simulate natural bone regeneration has not yet been obtained. The development of composite scaffolds based on natural sources appeared to be the perfect solution to be further applied and improve the biological activity, without the appearance of side effects. In this regard, the fabrication of artificial bone grafts, composed of bioceramics and biopolymers, represents the main solution in bone tissue regeneration [196,197,198]. Even though researchers came up with solutions of synthesis, the obtained biomaterials still showed poor mechanical properties compared with natural tissues. Moreover, synthetic tissues exhibited a lower integration and degradation rate in clinical applications. Numerous domains, including biology, materials science, and biophysics, combined their work in BTE, to interwind polymer synthesis and cell treatment to develop the ideal methods to produce novel materials [199,200]. The interaction between scaffolds and cells is a crucial aspect in designing scaffolds for tissue regeneration applications [201].
Considering bioceramics, HA materials can be improved through doping with several nonmetallic or metallic dopants to enhance their properties for further use. For example, Mg prevents HA crystallization, necessary for bone formation, and decreases bone fragility. This dopant can replace ~1% Ca+2 ions, presenting similar properties to bone tissue. Further, Zn disturbs the stability of the solution, inhibiting bacterial colonization, but also regulates inflammatory mediators. Another dopant, Sr, inhibits bone resorption, promotes osteogenesis, activates signaling pathways, and regulates cell behavior [202,203]. The doped ceramics can be further applied in combination with polymers to avoid the problems concerning biocompatibility and biodegradability; thus, the increase in biological activity. To explain the biological effects, scaffolds used in BTE are investigating angiogenesis, immunoregulation, biomineralization, and osteogenesis as the osseous tissue regeneration process. Although these materials were thoroughly investigated for biomedical applications, there are still many concerns for in vivo suitability and cytocompatibility. Nevertheless, the main challenge is to optimize biomaterials to further use them in manufacturing processes [204].
On the other hand, biofabrication techniques led to the formation of improved scaffolds, with better properties, but still insufficient to recreate the complex structural organization of organs or tissues. This challenge is crucial for the clinical translation of BTE applications and approval in the medical field. Most concerns have been focused, particularly, on the process and applied technology. Additive manufacturing methods (e.g., stereolithography, fused deposition modeling, selective laser sintering, 3D bioprinting) are the most applied techniques in bone regeneration. The main advantage, when one of these methods is used, is that scaffolds are dimensioned directly, reproducing the required 3D space. Moreover, this enables biologists to process precisely inorganic and organic materials [204,205].

6. Conclusions

In this review, the latest materials applied in bone tissue regeneration have been investigated to select the suitable candidate, designed as scaffolds with ideal properties. Firstly, natural, synthetic polymers and their applicability have been discussed, considering their advantages and drawbacks. Moreover, the possibility to obtain HA from natural sources was also studied, due to its similar structure with the native HA. The combination with polymeric materials (e.g., Col, CS, PLA, gelatin) can lead to an increase in the scaffold’s bioactivity and applicability in BTE.
Taking into consideration the possibility to combine polymers and ceramics, to tailor the properties of the composites, the characteristics of the ideal material have been researched to be subsequently applied in the medical field. In this direction, the possibility of using composite materials containing HA derived from natural sources has gained great attention for their application as scaffolds, surface coatings, or bio-inks, for 3D bioprinting. All the presented studies offered a general background in selecting the optimum material to be further applied in BTE, while the biologic activity has not yet been fully discovered.
This study’s final goal is to identify the natural source for HA and the suitable polymer, which can be combined with the selected ceramic. Currently, the main challenge in hard tissue engineering is optimizing the material synthesis method by augmenting the biologic activity and their physicochemical properties, without causing any negative effect.

Author Contributions

Conceptualization, D.-E.R., A.-M.G., I.A.N. and E.A.; methodology, A.-M.G. and I.A.N.; formal analysis, D.-E.R.; writing—original draft preparation, D.-E.R., A.-M.G. and I.A.N.; writing—review and editing, D.-E.R., A.-M.G. and I.A.N.; investigation, D.-E.R.; supervision, E.A., A.-M.G. and I.A.N.; project administration, E.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author Diana-Elena Radulescu is an employee of MDPI; however, she does not work for the journal Polymers at the time of submission and publication.

References

  1. Sharma, A.; Sharma, P.K.; Malviya, R. Stimuli-responsive supramolecules for bone tissue engineering. Biointerface Res. Appl. Chem. 2020, 10, 5122–5127. [Google Scholar] [CrossRef]
  2. Oladapo, B.I.; Zahedi, S.A.; Ismail, S.O.; Olawade, D.B. Recent advances in biopolymeric composite materials: Future sustainability of bone-implant. Renew. Sustain. Energy Rev. 2021, 150, 111505. [Google Scholar] [CrossRef]
  3. Adithya, S.P.; Sidharthan, D.S.; Abhinandan, R.; Balagangadharan, K.; Selvamurugan, N. Nanosheets-incorporated bio-composites containing natural and synthetic polymers/ceramics for bone tissue engineering. Int. J. Biol. Macromol. 2020, 164, 1960–1972. [Google Scholar] [CrossRef]
  4. Adzila, A.R.N.M.I.S. Synthesis of Calcium Phosphate Extracted from Eggshell Waste through Precipitation Method. Biointerface Res. Appl. Chem. 2021, 11, 15058–15067. [Google Scholar] [CrossRef]
  5. Shuai, C.; Yu, L.; Feng, P.; Gao, C.; Peng, S. Interfacial reinforcement in bioceramic/biopolymer composite bone scaffold: The role of coupling agent. Colloids Surf. B Biointerfaces 2020, 193, 111083. [Google Scholar] [CrossRef] [PubMed]
  6. Mohammadi Zerankeshi, M.; Bakhshi, R.; Alizadeh, R. Polymer/metal composite 3D porous bone tissue engineering scaffolds fabricated by additive manufacturing techniques: A review. Bioprinting 2022, 25, e00191. [Google Scholar] [CrossRef]
  7. Cao, S.; Zhao, Y.; Hu, Y.; Zou, L.; Chen, J. New perspectives: In-situ tissue engineering for bone repair scaffold. Compos. Part B Eng. 2020, 202, 108445. [Google Scholar] [CrossRef]
  8. Jain, S.; Gujjala, R.; Abdul Azeem, P.; Ojha, S.; Samudrala, R.K. A review on mechanical and In-vitro studies of polymer reinforced bioactive glass-scaffolds and their fabrication techniques. Ceram. Int. 2021, 48, 5908–5921. [Google Scholar] [CrossRef]
  9. Pu’ad, N.A.S.M.; Chuan, L.T.; Salman, N.S.; Selimin, M.A.; Latif, A.F.A.; Muhamad, M.S.; Abdullah, H.Z.; Idris, M.I.; Mustapha, S.N.H. Synthesis of bioactive calcium phosphate from cockle shell for biomedical applications. Biointerface Res. Appl. Chem. 2020, 10, 5787–5791. [Google Scholar] [CrossRef]
  10. Goh, K.W.; Wong, Y.H.; Ramesh, S.; Chandran, H.; Krishnasamy, S.; Ramesh, S.; Sidhu, A.; Teng, W.D. Effect of pH on the properties of eggshell-derived hydroxyapatite bioceramic synthesized by wet chemical method assisted by microwave irradiation. Ceram. Int. 2020, 47, 8879–8887. [Google Scholar] [CrossRef]
  11. Sobczak-Kupiec, A.; Pluta, K.; Drabczyk, A.; Włoś, M.; Tyliszczak, B. Synthesis and characterization of ceramic—Polymer composites containing bioactive synthetic hydroxyapatite for biomedical applications. Ceram. Int. 2018, 44, 13630–13638. [Google Scholar] [CrossRef]
  12. Terzioğlu, P.; Öğüt, H.; Kalemtaş, A. Natural calcium phosphates from fish bones and their potential biomedical applications. Mater. Sci. Eng. C 2018, 91, 899–911. [Google Scholar] [CrossRef]
  13. Shi, H.; Zhou, Z.; Li, W.; Fan, Y.; Li, Z.; Wei, J. Hydroxyapatite Based Materials for Bone Tissue Engineering: A Brief and Comprehensive Introduction. Crystals 2021, 11, 149. [Google Scholar] [CrossRef]
  14. Han, K.-S.; Sathiyaseelan, A.; Saravanakumar, K.; Wang, M.-H. Wound healing efficacy of biocompatible hydroxyapatite from bovine bone waste for bone tissue engineering application. J. Environ. Chem. Eng. 2022, 10, 106888. [Google Scholar] [CrossRef]
  15. Sanchez, A.G.; Prokhorov, E.; Luna-Barcenas, G.; Hernández-Vargas, J.; Román-Doval, R.; Mendoza, S.; Rojas-Chávez, H. Chitosan-hydroxyapatite-MWCNTs nanocomposite patch for bone tissue engineering applications. Mater. Today Commun. 2021, 28, 102615. [Google Scholar] [CrossRef]
  16. Sridevi, S.; Sutha, S.; Kavitha, L.; Gopi, D. Physicochemical and biological behaviour of biogenic derived hydroxyapatite and carboxymethyl cellulose/sodium alginate biocomposite coating on Ti6Al4V alloy for biomedical applications. Mater. Chem. Phys. 2020, 254, 123455. [Google Scholar] [CrossRef]
  17. Sathiyavimal, S.; Vasantharaj, S.; LewisOscar, F.; Selvaraj, R.; Brindhadevi, K.; Pugazhendhi, A. Natural organic and inorganic–hydroxyapatite biopolymer composite for biomedical applications. Prog. Org. Coat. 2020, 147, 105858. [Google Scholar] [CrossRef]
  18. Oladapo, B.I.; Zahedi, S.A.; Omigbodun, F.T. A systematic review of polymer composite in biomedical engineering. Eur. Polym. J. 2021, 154, 110534. [Google Scholar] [CrossRef]
  19. Gritsch, L.; Perrin, E.; Chenal, J.-M.; Fredholm, Y.; Maçon, A.L.B.; Chevalier, J.; Boccaccini, A.R. Combining bioresorbable polyesters and bioactive glasses: Orthopedic applications of composite implants and bone tissue engineering scaffolds. Appl. Mater. Today 2021, 22, 100923. [Google Scholar] [CrossRef]
  20. Dwivedi, R.; Kumar, S.; Pandey, R.; Mahajan, A.; Nandana, D.; Katti, D.S.; Mehrotra, D. Polycaprolactone as biomaterial for bone scaffolds: Review of literature. J. Oral Biol. Craniofacial Res. 2020, 10, 381–388. [Google Scholar] [CrossRef]
  21. Bharadwaz, A.; Jayasuriya, A.C. Recent trends in the application of widely used natural and synthetic polymer nanocomposites in bone tissue regeneration. Mater. Sci. Eng. C 2020, 110, 110698. [Google Scholar] [CrossRef] [PubMed]
  22. ul Hassan, N.; Chaudhery, I.; ur.Rehman, A.; Ahmed, N. Chapter 7—Polymeric nanoparticles used in tissue engineering. In Advances in Polymeric Nanomaterials for Biomedical Applications; Bajpai, A.K., Saini, R.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2021; pp. 191–224. [Google Scholar]
  23. Guo, Z.; Poot, A.A.; Grijpma, D.W. Advanced polymer-based composites and structures for biomedical applications. Eur. Polym. J. 2021, 149, 110388. [Google Scholar] [CrossRef]
  24. Ciocca, L.; Lesci, I.G.; Ragazzini, S.; Gioria, S.; Valsesia, A.; Parrilli, A.; Spadari, A.; Dozza, B.; Mora, P.; Piattelli, A.; et al. Nanostructured surface bioactive composite scaffold for filling of bone defects. Biointerface Res. Appl. Chem. 2020, 10, 5038–5047. [Google Scholar] [CrossRef]
  25. Blidi, O.; El Omari, N.; Balahbib, A.; Ghchime, R.; Ibrahimi, A.; Bouyahya, A.; Chokairi, O.; Barkiyou, M. Extraction Methods, Characterization and Biomedical Applications of Collagen: A Review. Biointerface Res. Appl. Chem. 2021, 11, 13587–13613. [Google Scholar]
  26. Hogan, K.J.; Mikos, A.G. Biodegradable thermoresponsive polymers: Applications in drug delivery and tissue engineering. Polymer 2020, 211, 123063. [Google Scholar] [CrossRef]
  27. Gupta, P.S.; Wasnik, K.; Patra, S.; Pareek, D.; Singh, M.; Maity, S.; Pandey, M.; Paik, P. A Review on Biodegradable Polymeric Materials for Bone Tissue Engineering (BTE) Applications. In Reference Module in Materials Science and Materials Engineering; Elsevier: Amsterdam, The Netherlands, 2021. [Google Scholar]
  28. Iglesias-Mejuto, A.; García-González, C.A. 3D-printed alginate-hydroxyapatite aerogel scaffolds for bone tissue engineering. Mater. Sci. Eng. C 2021, 131, 112525. [Google Scholar] [CrossRef]
  29. Asadi, N.; Del Bakhshayesh, A.R.; Davaran, S.; Akbarzadeh, A. Common biocompatible polymeric materials for tissue engineering and regenerative medicine. Mater. Chem. Phys. 2020, 242, 122528. [Google Scholar] [CrossRef]
  30. Zhai, P.; Peng, X.; Li, B.; Liu, Y.; Sun, H.; Li, X. The application of hyaluronic acid in bone regeneration. Int. J. Biol. Macromol. 2020, 151, 1224–1239. [Google Scholar] [CrossRef]
  31. Li, S.; Panda, A.K.; Liu, X.; Lin, Y.-C.; Huang, W.-Y.; Lin, C.; Zhao, G.; Chung, R.-J. Preparation and biocompatibility studies of Collagen/Hyaluronic Acid/Oligomeric proanthocyanidins composites. Mater. Chem. Phys. 2021, 272, 124959. [Google Scholar] [CrossRef]
  32. Dovedytis, M.; Liu, Z.J.; Bartlett, S. Hyaluronic acid and its biomedical applications: A review. Eng. Regen. 2020, 1, 102–113. [Google Scholar] [CrossRef]
  33. Hickey, R.J.; Pelling, A.E. Cellulose Biomaterials for Tissue Engineering. Front. Bioeng. Biotechnol. 2019, 7, 45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. del Valle, L.J.; Puiggalí, J. Hydroxyapatite Based Polymer Composites for Regenerative Medicine Applications. In Encyclopedia of Materials: Composites; Brabazon, D., Ed.; Elsevier: Oxford, UK, 2021; pp. 785–803. [Google Scholar]
  35. Guo, L.; Liang, Z.; Yang, L.; Du, W.; Yu, T.; Tang, H.; Li, C.; Qiu, H. The role of natural polymers in bone tissue engineering. J. Control. Release 2021, 338, 571–582. [Google Scholar] [CrossRef] [PubMed]
  36. Pina, S.; Reis, R.L.; Oliveira, J.M. Chapter 3—Natural polymeric biomaterials for tissue engineering. In Tissue Engineering Using Ceramics and Polymers, 3rd ed.; Boccaccini, A.R., Ma, P.X., Liverani, L., Eds.; Woodhead Publishing: Cambridge, UK, 2022; pp. 75–110. [Google Scholar]
  37. Kakarla, A.B.; Kong, I.; Turek, I.; Kong, C.; Irving, H. Printable gelatin, alginate and boron nitride nanotubes hydrogel-based ink for 3D bioprinting and tissue engineering applications. Mater. Des. 2022, 213, 110362. [Google Scholar] [CrossRef]
  38. El-Zawahry, M.M.; Hassabo, A.G.; Abdelghaffar, F.; Abdelghaffar, R.A.; Hakeim, O.A. Preparation and Use of Aqueous Solutions Magnetic Chitosan/Nanocellulose Aerogels for the Sorption of Reactive Black 5. Biointerface Res. Appl. Chem. 2021, 11, 12380–12402. [Google Scholar] [CrossRef]
  39. Shakir, M.; Zia, I.; Rehman, A.; Ullah, R. Fabrication and characterization of nanoengineered biocompatible n-HA/chitosan-tamarind seed polysaccharide: Bio-inspired nanocomposites for bone tissue engineering. Int. J. Biol. Macromol. 2018, 111, 903–916. [Google Scholar] [CrossRef] [PubMed]
  40. Cunha, C.S.; Castro, P.J.; Sousa, S.C.; Pullar, R.C.; Tobaldi, D.M.; Piccirillo, C.; Pintado, M.M. Films of chitosan and natural modified hydroxyapatite as effective UV-protecting, biocompatible and antibacterial wound dressings. Int. J. Biol. Macromol. 2020, 159, 1177–1185. [Google Scholar] [CrossRef]
  41. Olmo, J.A.; Ruiz-Rubio, L.; Pérez-Alvarez, L.; Sáez-Martínez, V.; Vilas-Vilela, J.L. Antibacterial Coatings for Improving the Performance of Biomaterials. Coatings 2020, 10, 139. [Google Scholar] [CrossRef] [Green Version]
  42. Lavanya, K.; Chandran, S.V.; Balagangadharan, K.; Selvamurugan, N. Temperature- and pH-responsive chitosan-based injectable hydrogels for bone tissue engineering. Mater. Sci. Eng. C 2020, 111, 110862. [Google Scholar] [CrossRef]
  43. Rethinam, S.; Wilson Aruni, A.; Vijayan, S.; Munusamy, C.; Gobi, N. Enhanced bone regeneration using an electrospun nanofibrous membrane—A novel approach. J. Drug Deliv. Sci. Technol. 2019, 53, 101163. [Google Scholar] [CrossRef]
  44. Marques, C.F.; Diogo, G.S.; Pina, S.; Oliveira, J.M.; Silva, T.H.; Reis, R.L. Collagen-based bioinks for hard tissue engineering applications: A comprehensive review. J. Mater. Sci. Mater. Med. 2019, 30, 32. [Google Scholar] [CrossRef]
  45. Elkhoury, K.; Morsink, M.; Sanchez-Gonzalez, L.; Kahn, C.; Tamayol, A.; Arab-Tehrany, E. Biofabrication of natural hydrogels for cardiac, neural, and bone Tissue engineering Applications. Bioact. Mater. 2021, 6, 3904–3923. [Google Scholar] [CrossRef] [PubMed]
  46. Echave, M.C.; Pimenta-Lopes, C.; Pedraz, J.L.; Mehrali, M.; Dolatshahi-Pirouz, A.; Ventura, F.; Orive, G. Enzymatic crosslinked gelatin 3D scaffolds for bone tissue engineering. Int. J. Pharm. 2019, 562, 151–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Luetchford, K.A.; Chaudhuri, J.B.; De Bank, P.A. Silk fibroin/gelatin microcarriers as scaffolds for bone tissue engineering. Mater. Sci. Eng. C 2020, 106, 110116. [Google Scholar] [CrossRef] [PubMed]
  48. Mishra, R.; Varshney, R.; Das, N.; Sircar, D.; Roy, P. Synthesis and characterization of gelatin-PVP polymer composite scaffold for potential application in bone tissue engineering. Eur. Polym. J. 2019, 119, 155–168. [Google Scholar] [CrossRef]
  49. Rajabnejad keleshteri, A.; Moztarzadeh, F.; Farokhi, M.; Mehrizi, A.A.; Basiri, H.; Mohseni, S.S. Preparation of microfluidic-based pectin microparticles loaded carbon dots conjugated with BMP-2 embedded in gelatin-elastin-hyaluronic acid hydrogel scaffold for bone tissue engineering application. Int. J. Biol. Macromol. 2021, 184, 29–41. [Google Scholar] [CrossRef]
  50. Cui, N.; Qian, J.; Liu, T.; Zhao, N.; Wang, H. Hyaluronic acid hydrogel scaffolds with a triple degradation behavior for bone tissue engineering. Carbohydr. Polym. 2015, 126, 192–198. [Google Scholar] [CrossRef]
  51. Sionkowska, A. Collagen blended with natural polymers: Recent advances and trends. Prog. Polym. Sci. 2021, 122, 101452. [Google Scholar] [CrossRef]
  52. Fang, Y.; Shi, L.; Duan, Z.; Rohani, S. Hyaluronic acid hydrogels, as a biological macromolecule-based platform for stem cells delivery and their fate control: A review. Int. J. Biol. Macromol. 2021, 189, 554–566. [Google Scholar] [CrossRef]
  53. Wang, L.; Zhang, D.; Ren, Y.; Guo, S.; Li, J.; Ma, S.; Yao, M.; Guan, F. Injectable hyaluronic acid hydrogel loaded with BMSC and NGF for traumatic brain injury treatment. Mater. Today Bio 2022, 13, 100201. [Google Scholar] [CrossRef]
  54. Abdollahiyan, P.; Oroojalian, F.; Hejazi, M.; de la Guardia, M.; Mokhtarzadeh, A. Nanotechnology, and scaffold implantation for the effective repair of injured organs: An overview on hard tissue engineering. J. Control. Release 2021, 333, 391–417. [Google Scholar] [CrossRef]
  55. Ferreira, F.V.; Otoni, C.G.; Lopes, J.H.; de Souza, L.P.; Mei, L.H.I.; Lona, L.M.F.; Lozano, K.; Lobo, A.O.; Mattoso, L.H.C. Ultrathin polymer fibers hybridized with bioactive ceramics: A review on fundamental pathways of electrospinning towards bone regeneration. Mater. Sci. Eng. C 2021, 123, 111853. [Google Scholar] [CrossRef] [PubMed]
  56. Jiang, D.; Ning, F.; Wang, Y. Additive manufacturing of biodegradable iron-based particle reinforced polylactic acid composite scaffolds for tissue engineering. J. Mater. Processing Technol. 2021, 289, 116952. [Google Scholar] [CrossRef]
  57. Arslan, A.; Steiger, W.; Roose, P.; Van den Bergen, H.; Gruber, P.; Zerobin, E.; Gantner, F.; Guillaume, O.; Ovsianikov, A.; Van Vlierberghe, S.; et al. Polymer architecture as key to unprecedented high-resolution 3D-printing performance: The case of biodegradable hexa-functional telechelic urethane-based poly-ε-caprolactone. Mater. Today 2021, 44, 25–39. [Google Scholar] [CrossRef]
  58. Wang, J.-Z.; You, M.-L.; Ding, Z.-Q.; Ye, W.-B. A review of emerging bone tissue engineering via PEG conjugated biodegradable amphiphilic copolymers. Mater. Sci. Eng. C 2019, 97, 1021–1035. [Google Scholar] [CrossRef] [PubMed]
  59. Yeo, T.; Ko, Y.-G.; Kim, E.J.; Kwon, O.K.; Chung, H.Y.; Kwon, O.H. Promoting bone regeneration by 3D-printed poly(glycolic acid)/hydroxyapatite composite scaffolds. J. Ind. Eng. Chem. 2021, 94, 343–351. [Google Scholar] [CrossRef]
  60. Thangprasert, A.; Tansakul, C.; Thuaksubun, N.; Meesane, J. Mimicked hybrid hydrogel based on gelatin/PVA for tissue engineering in subchondral bone interface for osteoarthritis surgery. Mater. Des. 2019, 183, 108113. [Google Scholar] [CrossRef]
  61. Kim, H.; Yang, G.H.; Kim, G. Three-dimensional gelatin/PVA scaffold with nanofibrillated collagen surface for applications in hard-tissue regeneration. Int. J. Biol. Macromol. 2019, 135, 21–28. [Google Scholar] [CrossRef]
  62. Bi, S.; Pang, J.; Huang, L.; Sun, M.; Cheng, X.; Chen, X. The toughness chitosan-PVA double network hydrogel based on alkali solution system and hydrogen bonding for tissue engineering applications. Int. J. Biol. Macromol. 2020, 146, 99–109. [Google Scholar] [CrossRef]
  63. Fu, Y.; Ding, Y.; Zhang, L.; Zhang, Y.; Liu, J.; Yu, P. Poly ethylene glycol (PEG)-Related controllable and sustainable antidiabetic drug delivery systems. Eur. J. Med. Chem. 2021, 217, 113372. [Google Scholar] [CrossRef]
  64. Xue, X.; Hu, Y.; Wang, S.; Chen, X.; Jiang, Y.; Su, J. Fabrication of physical and chemical crosslinked hydrogels for bone tissue engineering. Bioact. Mater. 2021. [Google Scholar] [CrossRef]
  65. Zhang, L.; Dou, S.; Li, Y.; Yuan, Y.; Ji, Y.; Wang, Y.; Yang, Y. Degradation and compatibility behaviors of poly(glycolic acid) grafted chitosan. Mater. Sci. Eng. C 2013, 33, 2626–2631. [Google Scholar] [CrossRef] [PubMed]
  66. Zhao, D.; Zhu, T.; Li, J.; Cui, L.; Zhang, Z.; Zhuang, X.; Ding, J. Poly(lactic-co-glycolic acid)-based composite bone-substitute materials. Bioact. Mater. 2021, 6, 346–360. [Google Scholar] [CrossRef] [PubMed]
  67. Ng, H.-M.; Bee, S.-T.; Sin, L.T.; Ratnam, C.T.; Rahmat, A.R. Effect of electron beam irradiation sterilization on biomedical polylactic acid composite filled with Scomberomorus Guttatus-derived hydroxyapatite. Compos. Part B Eng. 2019, 176, 107273. [Google Scholar] [CrossRef]
  68. Bal, Z.; Korkusuz, F.; Ishiguro, H.; Okada, R.; Kushioka, J.; Chijimatsu, R.; Kodama, J.; Tateiwa, D.; Ukon, Y.; Nakagawa, S.; et al. A novel nano-hydroxyapatite/synthetic polymer/bone morphogenetic protein-2 composite for efficient bone regeneration. Spine J. 2021, 21, 865–873. [Google Scholar] [CrossRef]
  69. Kumar, P.; Dehiya, B.S.; Sindhu, A. Synthesis and characterization of nHA-PEG and nBG-PEG scaffolds for hard tissue engineering applications. Ceram. Int. 2019, 45, 8370–8379. [Google Scholar] [CrossRef]
  70. Zhang, Y.; Gao, H.; Luo, H.; Chen, D.; Zhou, Z.; Cao, X. High strength HA-PEG/NAGA-Gelma double network hydrogel for annulus fibrosus rupture repair. Smart Mater. Med. 2022, 3, 128–138. [Google Scholar] [CrossRef]
  71. Bose, S.; Sarkar, N.; Banerjee, D. Effects of PCL, PEG and PLGA polymers on curcumin release from calcium phosphate matrix for in vitro and in vivo bone regeneration. Mater. Today Chem. 2018, 8, 110–120. [Google Scholar] [CrossRef]
  72. Dethe, M.R.; Prabakaran, A.; Ahmed, H.; Agrawal, M.; Roy, U.; Alexander, A. PCL-PEG copolymer based injectable thermosensitive hydrogels. J. Control. Release 2022, 343, 217–236. [Google Scholar] [CrossRef]
  73. Bai, J.; Wang, H.; Gao, W.; Liang, F.; Wang, Z.; Zhou, Y.; Lan, X.; Chen, X.; Cai, N.; Huang, W.; et al. Melt electrohydrodynamic 3D printed poly (ε-caprolactone)/polyethylene glycol/roxithromycin scaffold as a potential anti-infective implant in bone repair. Int. J. Pharm. 2020, 576, 118941. [Google Scholar] [CrossRef]
  74. Etminanfar, M.R.; Sheykholeslami, S.O.R.; Khalili, V.; Mahdavi, S. Biocompatibility and drug delivery efficiency of PEG-b-PCL/hydroxyapatite bilayer coatings on Nitinol superelastic alloy. Ceram. Int. 2020, 46, 12711–12717. [Google Scholar] [CrossRef]
  75. Kumar Saini, R.; Prasad Bagri, L.; Bajpai, A.K. Nano-silver hydroxyapatite based antibacterial 3D scaffolds of gelatin/alginate/poly (vinyl alcohol) for bone tissue engineering applications. Colloids Surf. B Biointerfaces 2019, 177, 211–218. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, G.; Chen, N.; Wang, Q. Fabrication and properties of poly(vinyl alcohol)/β-tricalcium phosphate composite scaffolds via fused deposition modeling for bone tissue engineering. Compos. Sci. Technol. 2019, 172, 17–28. [Google Scholar] [CrossRef]
  77. Salim, S.A.; Loutfy, S.A.; El-Fakharany, E.M.; Taha, T.H.; Hussien, Y.; Kamoun, E.A. Influence of chitosan and hydroxyapatite incorporation on properties of electrospun PVA/HA nanofibrous mats for bone tissue regeneration: Nanofibers optimization and in-vitro assessment. J. Drug Deliv. Sci. Technol. 2021, 62, 102417. [Google Scholar] [CrossRef]
  78. Dong, Q.; Zhang, M.; Zhou, X.; Shao, Y.; Li, J.; Wang, L.; Chu, C.; Xue, F.; Yao, Q.; Bai, J. 3D-printed Mg-incorporated PCL-based scaffolds: A promising approach for bone healing. Mater. Sci. Eng. C 2021, 129, 112372. [Google Scholar] [CrossRef] [PubMed]
  79. Maharjan, B.; Kaliannagounder, V.K.; Jang, S.R.; Awasthi, G.P.; Bhattarai, D.P.; Choukrani, G.; Park, C.H.; Kim, C.S. In-situ polymerized polypyrrole nanoparticles immobilized poly(ε-caprolactone) electrospun conductive scaffolds for bone tissue engineering. Mater. Sci. Eng. C 2020, 114, 111056. [Google Scholar] [CrossRef]
  80. Wang, Q.; Ye, W.; Ma, Z.; Xie, W.; Zhong, L.; Wang, Y.; Rong, Q. 3D printed PCL/β-TCP cross-scale scaffold with high-precision fiber for providing cell growth and forming bones in the pores. Mater. Sci. Eng. C 2021, 127, 112197. [Google Scholar] [CrossRef]
  81. Gatto, M.L.; Furlani, M.; Giuliani, A.; Bloise, N.; Fassina, L.; Visai, L.; Mengucci, P. Biomechanical performances of PCL/HA micro- and macro-porous lattice scaffolds fabricated via laser powder bed fusion for bone tissue engineering. Mater. Sci. Eng. C 2021, 128, 112300. [Google Scholar] [CrossRef]
  82. Cestari, F.; Petretta, M.; Yang, Y.; Motta, A.; Grigolo, B.; Sglavo, V.M. 3D printing of PCL/nano-hydroxyapatite scaffolds derived from biogenic sources for bone tissue engineering. Sustain. Mater. Technol. 2021, 29, e00318. [Google Scholar] [CrossRef]
  83. Soleymani Eil Bakhtiari, S.; Bakhsheshi-Rad, H.R.; Karbasi, S.; Fauzi Ismail, A.; Sharif, S.; Seifalian, A.; Savoji, H.; Berto, F. Poly(methyl methacrylate)-Based Composite Bone Cements With Different Types of Reinforcement Agents. In Encyclopedia of Materials: Composites; Brabazon, D., Ed.; Elsevier: Oxford, UK, 2021; pp. 867–886. [Google Scholar]
  84. Pahlevanzadeh, F.; Bakhsheshi-Rad, H.R.; Kharaziha, M.; Kasiri-Asgarani, M.; Omidi, M.; Razzaghi, M.; Ismail, A.F.; Sharif, S.; RamaKrishna, S.; Berto, F. CNT and rGO reinforced PMMA based bone cement for fixation of load bearing implants: Mechanical property and biological response. J. Mech. Behav. Biomed. Mater. 2021, 116, 104320. [Google Scholar] [CrossRef]
  85. Sas, A.; Helgason, B.; Ferguson, S.J.; van Lenthe, G.H. Mechanical and morphological characterization of PMMA/bone composites in human femoral heads. J. Mech. Behav. Biomed. Mater. 2021, 115, 104247. [Google Scholar] [CrossRef]
  86. Bardelcik, A.; Yang, S.; Alderson, F.; Gadsden, A. The effect of wash treatment on the mechanical properties and energy absorption potential of a 3D printed polymethyl methacrylate (PMMA). Mater. Today Commun. 2021, 26, 101728. [Google Scholar] [CrossRef]
  87. Al-Sherify, Z.F.; Dawood, N.M.; Khulief, Z.T. Corrosion behavior of AZ31 magnesium alloys coated with PMMA/HA as biodegradable implants. Mater. Today Proc. 2021, in press. [Google Scholar] [CrossRef]
  88. Atila, D.; Hasirci, V.; Tezcaner, A. Coaxial electrospinning of composite mats comprised of core/shell poly(methyl methacrylate)/silk fibroin fibers for tissue engineering applications. J. Mech. Behav. Biomed. Mater. 2022, 128, 105105. [Google Scholar] [CrossRef] [PubMed]
  89. Deb, P.; Barua, E.; Das Lala, S.; Deoghare, A.B. Synthesis of hydroxyapatite from Labeo rohita fish scale for biomedical application. Mater. Today Proc. 2019, 15, 277–283. [Google Scholar] [CrossRef]
  90. Pal, A.; Paul, S.; Choudhury, A.R.; Balla, V.K.; Das, M.; Sinha, A. Synthesis of hydroxyapatite from Lates calcarifer fish bone for biomedical applications. Mater. Lett. 2017, 203, 89–92. [Google Scholar] [CrossRef]
  91. Sathiyavimal, S.; Vasantharaj, S.; Shanmugavel, M.; Manikandan, E.; Nguyen-Tri, P.; Brindhadevi, K.; Pugazhendhi, A. Facile synthesis and characterization of hydroxyapatite from fish bones: Photocatalytic degradation of industrial dyes (crystal violet and Congo red). Prog. Org. Coat. 2020, 148, 105890. [Google Scholar] [CrossRef]
  92. Liu, K.S.; Tian, D.L.; Jiang, L. Chapter 24—Frontier of Inorganic Synthesis and Preparative Chemistry (I) Biomimetic Synthesis. In Modern Inorganic Synthetic Chemistry, 2nd ed.; Xu, R., Xu, Y., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 687–721. [Google Scholar]
  93. Wu, S.-C.; Hsu, H.-C.; Hsu, S.-K.; Tseng, C.-P.; Ho, W.-F. Preparation and characterization of hydroxyapatite synthesized from oyster shell powders. Adv. Powder Technol. 2017, 28, 1154–1158. [Google Scholar] [CrossRef]
  94. Vinoth Kumar, K.C.; Jani Subha, T.; Ahila, K.G.; Ravindran, B.; Chang, S.W.; Mahmoud, A.H.; Mohammed, O.B.; Rathi, M.A. Spectral characterization of hydroxyapatite extracted from Black Sumatra and Fighting cock bone samples: A comparative analysis. Saudi J. Biol. Sci. 2020, 28, 840–846. [Google Scholar] [CrossRef]
  95. Nga, N.K.; Thuy Chau, N.T.; Viet, P.H. Facile synthesis of hydroxyapatite nanoparticles mimicking biological apatite from eggshells for bone-tissue engineering. Colloids Surf. B Biointerfaces 2018, 172, 769–778. [Google Scholar] [CrossRef]
  96. Behera, R.R.; Das, A.; Pamu, D.; Pandey, L.M.; Sankar, M.R. Mechano-tribological properties and in vitro bioactivity of biphasic calcium phosphate coating on Ti-6Al-4V. J. Mech. Behav. Biomed. Mater. 2018, 86, 143–157. [Google Scholar] [CrossRef]
  97. Behera, R.R.; Das, A.; Hasan, A.; Pamu, D.; Pandey, L.M.; Sankar, M.R. Deposition of biphasic calcium phosphate film on laser surface textured Ti–6Al–4V and its effect on different biological properties for orthopedic applications. J. Alloy. Compd. 2020, 842, 155683. [Google Scholar] [CrossRef]
  98. Nawaz, A.; Li, E.; Irshad, S.; Hhm, H.; Liu, J.; Shahbaz, H.M.; Ahmed, W.; Regenstein, J.M. Improved effect of autoclave processing on size reduction, chemical structure, nutritional, mechanical and in vitro digestibility properties of fish bone powder. Adv. Powder Technol. 2020, 31, 2513–2520. [Google Scholar] [CrossRef]
  99. Hernández-Cocoletzi, H.; Salinas, R.A.; Águila-Almanza, E.; Rubio-Rosas, E.; Chai, W.S.; Chew, K.W.; Mariscal-Hernández, C.; Show, P.L. Natural hydroxyapatite from fishbone waste for the rapid adsorption of heavy metals of aqueous effluent. Environ. Technol. Innov. 2020, 20, 101109. [Google Scholar] [CrossRef]
  100. Zhu, Q.; Ablikim, Z.; Chen, T.; Cai, Q.; Xia, J.; Jiang, D.; Wang, S. The preparation and characterization of HA/β-TCP biphasic ceramics from fish bones. Ceram. Int. 2017, 43, 12213–12220. [Google Scholar] [CrossRef]
  101. Zanelato, C.B.; Pires, A.F.; da Silva, S.N.; Galdino, A.G.S. Development of biphasic bone cement obtained from chicken eggshell. J. Mater. Res. Technol. 2020, 9, 7297–7304. [Google Scholar] [CrossRef]
  102. Shafiei, S.; Omidi, M.; Nasehi, F.; Golzar, H.; Mohammadrezaei, D.; Rezai Rad, M.; Khojasteh, A. Egg shell-derived calcium phosphate/carbon dot nanofibrous scaffolds for bone tissue engineering: Fabrication and characterization. Mater. Sci. Eng. C 2019, 100, 564–575. [Google Scholar] [CrossRef]
  103. Kolmas, J.; Kaflak, A.; Zima, A.; Ślósarczyk, A. Alpha-tricalcium phosphate synthesized by two different routes: Structural and spectroscopic characterization. Ceram. Int. 2015, 41, 5727–5733. [Google Scholar] [CrossRef]
  104. Roopavath, U.K.; Sah, M.K.; Panigrahi, B.B.; Rath, S.N. Mechanochemically synthesized phase stable and biocompatible β-tricalcium phosphate from avian eggshell for the development of tissue ingrowth system. Ceram. Int. 2019, 45, 12910–12919. [Google Scholar] [CrossRef]
  105. Oladele, I.O.; Agbabiaka, O.G.; Adediran, A.A.; Akinwekomi, A.D.; Balogun, A.O. Structural performance of poultry eggshell derived hydroxyapatite based high density polyethylene bio-composites. Heliyon 2019, 5, e02552. [Google Scholar] [CrossRef] [Green Version]
  106. Horta, M.; Aguilar, M.; Moura, F.; Campos, J.; Ramos, V.; Quizunda, A. Synthesis and characterization of green nanohydroxyapatite from hen eggshell by precipitation method. Mater. Today Proc. 2019, 14, 716–721. [Google Scholar] [CrossRef]
  107. Vidhya, G.; Kumar, G.S.; Kattimani, V.S.; Girija, E.K. Comparative study of hydroxyapatite prepared from eggshells and synthetic precursors by microwave irradiation method for medical applications. Mater. Today Proc. 2019, 15, 344–352. [Google Scholar] [CrossRef]
  108. Núñez, D.; Elgueta, E.; Varaprasad, K.; Oyarzún, P. Hydroxyapatite nanocrystals synthesized from calcium rich bio-wastes. Mater. Lett. 2018, 230, 64–68. [Google Scholar] [CrossRef]
  109. Alif, M.F.; Aprillia, W.; Arief, S. A hydrothermal synthesis of natural hydroxyapatite obtained from Corbicula moltkiana freshwater clams shell biowaste. Mater. Lett. 2018, 230, 40–43. [Google Scholar] [CrossRef]
  110. Zainol, I.; Adenan, N.H.; Rahim, N.A.; Jaafar, C.N.A. Extraction of natural hydroxyapatite from tilapia fish scales using alkaline treatment. Mater. Today Proc. 2019, 16, 1942–1948. [Google Scholar] [CrossRef]
  111. Jyotsna; Vijayakumar, P. Synthesis and characterization of hydroxyapatite nanoparticles and their cytotoxic effect on a fish vertebra derived cell line. Biocatal. Agric. Biotechnol. 2020, 25, 101612. [Google Scholar] [CrossRef]
  112. Paul, S.; Pal, A.; Choudhury, A.R.; Bodhak, S.; Balla, V.K.; Sinha, A.; Das, M. Effect of trace elements on the sintering effect of fish scale derived hydroxyapatite and its bioactivity. Ceram. Int. 2017, 43, 15678–15684. [Google Scholar] [CrossRef]
  113. Pon-On, W.; Suntornsaratoon, P.; Charoenphandhu, N.; Thongbunchoo, J.; Krishnamra, N.; Tang, I.M. Hydroxyapatite from fish scale for potential use as bone scaffold or regenerative material. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 62, 183–189. [Google Scholar] [CrossRef]
  114. Varadavenkatesan, T.; Vinayagam, R.; Pai, S.; Kathirvel, B.; Pugazhendhi, A.; Selvaraj, R. Synthesis, biological and environmental applications of hydroxyapatite and its composites with organic and inorganic coatings. Prog. Org. Coat. 2021, 151, 106056. [Google Scholar] [CrossRef]
  115. Luna-Domínguez, J.H.; Téllez-Jiménez, H.; Hernández-Cocoletzi, H.; García-Hernández, M.; Melo-Banda, J.A.; Nygren, H. Development and in vivo response of hydroxyapatite/whitlockite from chicken bones as bone substitute using a chitosan membrane for guided bone regeneration. Ceram. Int. 2018, 44, 22583–22591. [Google Scholar] [CrossRef]
  116. Nam, P.V.; Hoa, N.V.; Trung, T.S. Properties of hydroxyapatites prepared from different fish bones: A comparative study. Ceram. Int. 2019, 45, 20141–20147. [Google Scholar] [CrossRef]
  117. Cestari, F.; Agostinacchio, F.; Galotta, A.; Chemello, G.; Motta, A.; Sglavo, V.M. Nano-Hydroxyapatite Derived from Biogenic and Bioinspired Calcium Carbonates: Synthesis and In Vitro Bioactivity. Nanomaterials 2021, 11, 264. [Google Scholar] [CrossRef] [PubMed]
  118. Mohd Pu’ad, N.A.S.; Koshy, P.; Abdullah, H.Z.; Idris, M.I.; Lee, T.C. Syntheses of hydroxyapatite from natural sources. Heliyon 2019, 5, e01588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Komalakrishna, H.; Shine Jyoth, T.G.; Kundu, B.; Mandal, S. Low Temperature Development of Nano-Hydroxyapatite from Austromegabalanus psittacus, Star fish and Sea urchin. Mater. Today Proc. 2017, 4, 11933–11938. [Google Scholar] [CrossRef]
  120. Mohd Pu’ad, N.A.S.; Abdul Haq, R.H.; Mohd Noh, H.; Abdullah, H.Z.; Idris, M.I.; Lee, T.C. Nano-size hydroxyapatite extracted from tilapia scale using alkaline heat treatment method. Mater. Today Proc. 2020, 29, 218–222. [Google Scholar] [CrossRef]
  121. Rahavi, S.; Ghaderi, O.; Monshi, A.; Fathi, M. A comparative study on physicochemical properties of hydroxyapatite powders derived from natural and synthetic sources. Russ. J. Non-Ferr. Met. 2017, 58, 276–286. [Google Scholar] [CrossRef]
  122. Khalid, H.; Chaudhry, A.A. 4—Basics of hydroxyapatite—Structure, synthesis, properties, and clinical applications. In Handbook of Ionic Substituted Hydroxyapatites; Khan, A.S., Chaudhry, A.A., Eds.; Woodhead Publishing: Cambridge, UK, 2020; pp. 85–115. [Google Scholar]
  123. Gibson, I.R. 1.3.4A—Natural and Synthetic Hydroxyapatites. In Biomaterials Science, 4th ed.; Wagner, W.R., Sakiyama-Elbert, S.E., Zhang, G., Yaszemski, M.J., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 307–317. [Google Scholar]
  124. Martinez, M.; Bayne, C.; Aiello, D.; Julian, M.; Gaume, R.; Baudelet, M. Multi-elemental matrix-matched calcium hydroxyapatite reference materials for laser ablation: Evaluation on teeth by laser-induced breakdown spectroscopy. Spectrochim. Acta Part B At. Spectrosc. 2019, 159, 105650. [Google Scholar] [CrossRef]
  125. Boutinguiza, M.; Pou, J.; Comesaña, R.; Lusquiños, F.; de Carlos, A.; León, B. Biological hydroxyapatite obtained from fish bones. Mater. Sci. Eng. C 2012, 32, 478–486. [Google Scholar] [CrossRef]
  126. Obada, D.O.; Dauda, E.T.; Abifarin, J.K.; Dodoo-Arhin, D.; Bansod, N.D. Mechanical properties of natural hydroxyapatite using low cold compaction pressure: Effect of sintering temperature. Mater. Chem. Phys. 2020, 239, 122099. [Google Scholar] [CrossRef]
  127. Shi, P.; Liu, M.; Fan, F.; Yu, C.; Lu, W.; Du, M. Characterization of natural hydroxyapatite originated from fish bone and its biocompatibility with osteoblasts. Mater. Sci. Eng. C 2018, 90, 706–712. [Google Scholar] [CrossRef]
  128. Liu, Y.; Liu, M.; Ji, S.; Zhang, L.; Cao, W.; Wang, H.; Wang, S. Preparation and application of hydroxyapatite extracted from fish scale waste using deep eutectic solvents. Ceram. Int. 2020, 47, 9366–9372. [Google Scholar] [CrossRef]
  129. Wu, S.-C.; Tsou, H.-K.; Hsu, H.-C.; Hsu, S.-K.; Liou, S.-P.; Ho, W.-F. A hydrothermal synthesis of eggshell and fruit waste extract to produce nanosized hydroxyapatite. Ceram. Int. 2013, 39, 8183–8188. [Google Scholar] [CrossRef]
  130. Karunakaran, G.; Cho, E.-B.; Kumar, G.S.; Kolesnikov, E.; Janarthanan, G.; Pillai, M.M.; Rajendran, S.; Boobalan, S.; Sudha, K.G.; Rajeshkumar, M.P. Mesoporous Mg-doped hydroxyapatite nanorods prepared from bio-waste blue mussel shells for implant applications. Ceram. Int. 2020, 46, 28514–28527. [Google Scholar] [CrossRef]
  131. Wang, Z.; Jiang, S.; Zhao, Y.; Zeng, M. Synthesis and characterization of hydroxyapatite nano-rods from oyster shell with exogenous surfactants. Mater. Sci. Eng. C 2019, 105, 110102. [Google Scholar] [CrossRef]
  132. Yang, Y.; Yao, Q.; Pu, X.; Hou, Z.; Zhang, Q. Biphasic calcium phosphate macroporous scaffolds derived from oyster shells for bone tissue engineering. Chem. Eng. J. 2011, 173, 837–845. [Google Scholar] [CrossRef]
  133. Morris, J.P.; Wang, Y.; Backeljau, T.; Chapelle, G. Biomimetic and bio-inspired uses of mollusc shells. Mar. Genom. 2016, 27, 85–90. [Google Scholar] [CrossRef]
  134. Haslinda Shariffuddin, J.; Chee Yean, W.; Shariah Ghazali, S. Investigating the catalytic properties of calcium compounds derived from marine based shell waste for wastewater treatment. Mater. Today Proc. 2018, 5, 21718–21727. [Google Scholar] [CrossRef]
  135. Lala, S.; Satpati, B.; Kar, T.; Pradhan, S.K. Structural and microstructural characterizations of nanocrystalline hydroxyapatite synthesized by mechanical alloying. Mater. Sci. Eng. C Mater. Biol. Appl. 2013, 33, 2891–2898. [Google Scholar] [CrossRef]
  136. Hembrick-Holloman, V.; Samuel, T.; Mohammed, Z.; Jeelani, S.; Rangari, V.K. Ecofriendly production of bioactive tissue engineering scaffolds derived from egg- and sea-shells. J. Mater. Res. Technol. 2020, 9, 13729–13739. [Google Scholar] [CrossRef]
  137. Oliveira, D.A.; Benelli, P.; Amante, E.R. A literature review on adding value to solid residues: Egg shells. J. Clean. Prod. 2013, 46, 42–47. [Google Scholar] [CrossRef]
  138. Ramesh, S.; Natasha, A.N.; Tan, C.Y.; Bang, L.T.; Ramesh, S.; Ching, C.Y.; Chandran, H. Direct conversion of eggshell to hydroxyapatite ceramic by a sintering method. Ceram. Int. 2016, 42, 7824–7829. [Google Scholar] [CrossRef]
  139. Baláž, M. Ball milling of eggshell waste as a green and sustainable approach: A review. Adv. Colloid Interface Sci. 2018, 256, 256–275. [Google Scholar] [CrossRef] [PubMed]
  140. Wu, S.-C.; Hsu, H.-C.; Hsu, S.-K.; Chang, Y.-C.; Ho, W.-F. Synthesis of hydroxyapatite from eggshell powders through ball milling and heat treatment. J. Asian Ceram. Soc. 2016, 4, 85–90. [Google Scholar] [CrossRef]
  141. Ho, W.-F.; Hsu, H.-C.; Hsu, S.-K.; Hung, C.-W.; Wu, S.-C. Calcium phosphate bioceramics synthesized from eggshell powders through a solid state reaction. Ceram. Int. 2013, 39, 6467–6473. [Google Scholar] [CrossRef]
  142. Alhussary, B.N.; Taqa, G.A.; Taqa, A.A.A. Preparation And Characterization Of Natural Nano Hydroxyapatite From Eggshell And Seashell And Its Effect On Bone Healing. J. Appl. Vet. Sci. 2020, 5, 25–32. [Google Scholar] [CrossRef] [Green Version]
  143. Waheed, M.; Yousaf, M.; Shehzad, A.; Inam-Ur-Raheem, M.; Khan, M.K.I.; Khan, M.R.; Ahmad, N.; Abdullah; Aadil, R.M. Channelling eggshell waste to valuable and utilizable products: A comprehensive review. Trends Food Sci. Technol. 2020, 106, 78–90. [Google Scholar] [CrossRef]
  144. Correlo, V.M.; Oliveira, J.M.; Mano, J.F.; Neves, N.M.; Reis, R.L. CHAPTER 32—Natural Origin Materials for Bone Tissue Engineering—Properties, Processing, and Performance. In Principles of Regenerative Medicine, 4th ed.; Atala, A., Lanza, R., Thomson, J.A., Nerem, R., Eds.; Academic Press: San Diego, CA, USA, 2011; pp. 557–586. [Google Scholar]
  145. Ağaoğulları, D.; Kel, D.; Gökçe, H.; Duman, I.; Öveçoğlu, L.; Akarsubasi, A.; Bilgiç Alkaya, D.; Oktar, F. Bioceramic Production from Sea Urchins. Acta Phys. Pol. A 2012, 121, 23–26. [Google Scholar] [CrossRef]
  146. Shavandi, A.; Bekhit, A.E.-D.A.; Ali, A.; Sun, Z. Synthesis of nano-hydroxyapatite (nHA) from waste mussel shells using a rapid microwave method. Mater. Chem. Phys. 2015, 149–150, 607–616. [Google Scholar] [CrossRef]
  147. Mitran, V.; Ion, R.; Miculescu, F.; Necula, M.G.; Mocanu, A.-C.; Stan, G.E.; Antoniac, I.V.; Cimpean, A. Osteoblast Cell Response to Naturally Derived Calcium Phosphate-Based Materials. Materials 2018, 11, 1097. [Google Scholar] [CrossRef] [Green Version]
  148. Maidaniuc, A.; Miculescu, F.; Ciocoiu, R.C.; Butte, T.M.; Pasuk, I.; Stan, G.E.; Voicu, S.I.; Ciocan, L.T. Effect of the processing parameters on surface, physico-chemical and mechanical features of bioceramics synthesized from abundant carp fish bones. Ceram. Int. 2020, 46, 10159–10171. [Google Scholar] [CrossRef]
  149. Lalzawmliana, V.; Anand, A.; Mukherjee, P.; Chaudhuri, S.; Kundu, B.; Nandi, S.K.; Thakur, N.L. Marine organisms as a source of natural matrix for bone tissue engineering. Ceram. Int. 2019, 45, 1469–1481. [Google Scholar] [CrossRef]
  150. Venkatesan, J.; Qian, Z.J.; Ryu, B.; Thomas, N.V.; Kim, S.K. A comparative study of thermal calcination and an alkaline hydrolysis method in the isolation of hydroxyapatite from Thunnus obesus bone. Biomed. Mater. 2011, 6, 035003. [Google Scholar] [CrossRef] [PubMed]
  151. Chai, Y.; Nishikawa, M.; Tagaya, M. Preparation of gold/hydroxyapatite hybrids using natural fish scale template and their effective albumin interactions. Adv. Powder Technol. 2018, 29, 1198–1203. [Google Scholar] [CrossRef]
  152. Buraiki, N.S.S.A.; Ali Albadri, B.; Alsheriqi, S.; Alshabibi, B.; Al-Mammari, S.; Premkumar, S.; Sah, M.K.; Sudhakar, M.S. Characterization of Catla catla and Oreochromis niloticus fish scales derived hydroxyapatite scaffolds for regenerative medicine. Mater. Today Proc. 2020, 27, 2609–2616. [Google Scholar] [CrossRef]
  153. Piccirillo, C.; Silva, M.F.; Pullar, R.C.; Braga da Cruz, I.; Jorge, R.; Pintado, M.M.E.; Castro, P.M.L. Extraction and characterisation of apatite- and tricalcium phosphate-based materials from cod fish bones. Mater. Sci. Eng. C 2013, 33, 103–110. [Google Scholar] [CrossRef] [PubMed]
  154. Goto, T.; Sasaki, K. Synthesis of morphologically controlled hydroxyapatite from fish bone by urea-assisted hydrothermal treatment and its Sr2+ sorption capacity. Powder Technol. 2016, 292, 314–322. [Google Scholar] [CrossRef]
  155. Zhang, L.; Zhang, C.; Zhang, R.; Jiang, D.; Zhu, Q.; Wang, S. Extraction and characterization of HA/β-TCP biphasic calcium phosphate from marine fish. Mater. Lett. 2019, 236, 680–682. [Google Scholar] [CrossRef]
  156. Akpan, E.S.; Dauda, M.; Kuburi, L.S.; Obada, D.O.; Dodoo-Arhin, D. A comparative study of the mechanical integrity of natural hydroxyapatite scaffolds prepared from two biogenic sources using a low compaction pressure method. Results Phys. 2020, 17, 103051. [Google Scholar] [CrossRef]
  157. Popescu-Pelin, G.; Ristoscu, C.; Duta, L.; Pasuk, I.; Stan, G.E.; Stan, M.S.; Popa, M.; Chifiriuc, M.C.; Hapenciuc, C.; Oktar, F.N.; et al. Fish Bone Derived Bi-Phasic Calcium Phosphate Coatings Fabricated by Pulsed Laser Deposition for Biomedical Applications. Mar. Drugs 2020, 18, 623. [Google Scholar] [CrossRef]
  158. Bas, M.; Daglilar, S.; Kuskonmaz, N.; Kalkandelen, C.; Erdemir, G.; Kuruca, S.E.; Tulyaganov, D.; Yoshioka, T.; Gunduz, O.; Ficai, D.; et al. Mechanical and Biocompatibility Properties of Calcium Phosphate Bioceramics Derived from Salmon Fish Bone Wastes. Int. J. Mol. Sci. 2020, 21, 8082. [Google Scholar] [CrossRef]
  159. Begam, H.; Nandi, S.K.; Chanda, A.; Kundu, B. Effect of bone morphogenetic protein on Zn-HAp and Zn-HAp/collagen composite: A systematic in vivo study. Res. Vet. Sci. 2017, 115, 1–9. [Google Scholar] [CrossRef]
  160. Wang, N.; Thameem Dheen, S.; Fuh, J.Y.H.; Senthil Kumar, A. A review of multi-functional ceramic nanoparticles in 3D printed bone tissue engineering. Bioprinting 2021, 23, e00146. [Google Scholar] [CrossRef]
  161. Su, X.; Wang, T.; Guo, S. Applications of 3D printed bone tissue engineering scaffolds in the stem cell field. Regen. Ther. 2021, 16, 63–72. [Google Scholar] [CrossRef] [PubMed]
  162. Tran, H.D.N.; Park, K.D.; Ching, Y.C.; Huynh, C.; Nguyen, D.H. A comprehensive review on polymeric hydrogel and its composite: Matrices of choice for bone and cartilage tissue engineering. J. Ind. Eng. Chem. 2020, 89, 58–82. [Google Scholar] [CrossRef]
  163. Deb, P.; Barua, E.; Deoghare, A.B.; Lala, S.D. Development of bone scaffold using Puntius conchonius fish scale derived hydroxyapatite: Physico-mechanical and bioactivity evaluations. Ceram. Int. 2019, 45, 10004–10012. [Google Scholar] [CrossRef]
  164. Pon-On, W.; Suntornsaratoon, P.; Charoenphandhu, N.; Thongbunchoo, J.; Krishnamra, N.; Tang, I.M. Synthesis and investigations of mineral ions-loaded apatite from fish scale and PLA/chitosan composite for bone scaffolds. Mater. Lett. 2018, 221, 143–146. [Google Scholar] [CrossRef]
  165. Nawang, R.; Hussein, M.Z.; Matori, K.A.; Che Abdullah, C.A.; Hashim, M. Physicochemical properties of hydroxyapatite/montmorillonite nanocomposite prepared by powder sintering. Results Phys. 2019, 15, 102540. [Google Scholar] [CrossRef]
  166. Hakobyan, S.; Roohpour, N.; Gautrot, J.E. Modes of adsorption of polyelectrolytes to model substrates of hydroxyapatite. J. Colloid Interface Sci. 2019, 543, 237–246. [Google Scholar] [CrossRef]
  167. Stewart, C.; Akhavan, B.; Wise, S.G.; Bilek, M.M.M. A review of biomimetic surface functionalization for bone-integrating orthopedic implants: Mechanisms, current approaches, and future directions. Prog. Mater. Sci. 2019, 106, 100588. [Google Scholar] [CrossRef]
  168. Morakul, S.; Otsuka, Y.; Ohnuma, K.; Tagaya, M.; Motozuka, S.; Miyashita, Y.; Mutoh, Y. Enhancement effect on antibacterial property of gray titania coating by plasma-sprayed hydroxyapatite-amino acid complexes during irradiation with visible light. Heliyon 2019, 5, e02207. [Google Scholar] [CrossRef] [Green Version]
  169. Lu, A.; Gao, Y.; Jin, T.; Luo, X.; Zeng, Q.; Shang, Z. Effects of surface roughness and texture on the bacterial adhesion on the bearing surface of bio-ceramic joint implants: An in vitro study. Ceram. Int. 2020, 46, 6550–6559. [Google Scholar] [CrossRef]
  170. Sopchenski, L.; Cogo, S.; Dias-Ntipanyj, M.F.; Elifio-Espósito, S.; Popat, K.C.; Soares, P. Bioactive and antibacterial boron doped TiO2 coating obtained by PEO. Appl. Surf. Sci. 2018, 458, 49–58. [Google Scholar] [CrossRef]
  171. Fordham, W.R.; Redmond, S.; Westerland, A.; Cortes, E.G.; Walker, C.; Gallagher, C.; Medina, C.J.; Waecther, F.; Lunk, C.; Ostrum, R.F.; et al. Silver as a Bactericidal Coating for Biomedical Implants. Surf. Coat. Technol. 2014, 253, 52–57. [Google Scholar] [CrossRef]
  172. Barberia-Roque, L.; Obidi, O.F.; Gámez-Espinosa, E.; Viera, M.; Bellotti, N. Hygienic coatings with bioactive nano-additives from Senna occidentalis-mediated green synthesis. NanoImpact 2019, 16, 100184. [Google Scholar] [CrossRef]
  173. Kaczmarek, B.; Sionkowska, A.; Osyczka, A.M. Physicochemical properties of scaffolds based on mixtures of chitosan, collagen and glycosaminoglycans with nano-hydroxyapatite addition. Int. J. Biol. Macromol. 2018, 118, 1880–1883. [Google Scholar] [CrossRef]
  174. Anita Lett, J.; Sagadevan, S.; Fatimah, I.; Hoque, M.E.; Lokanathan, Y.; Léonard, E.; Alshahateet, S.F.; Schirhagl, R.; Oh, W.C. Recent advances in natural polymer-based hydroxyapatite scaffolds: Properties and applications. Eur. Polym. J. 2021, 148, 110360. [Google Scholar] [CrossRef]
  175. Echave, M.C.; Sánchez, P.; Pedraz, J.L.; Orive, G. Progress of gelatin-based 3D approaches for bone regeneration. J. Drug Deliv. Sci. Technol. 2017, 42, 63–74. [Google Scholar] [CrossRef]
  176. Zhu, G.; Zhang, T.; Chen, M.; Yao, K.; Huang, X.; Zhang, B.; Li, Y.; Liu, J.; Wang, Y.; Zhao, Z. Bone physiological microenvironment and healing mechanism: Basis for future bone-tissue engineering scaffolds. Bioact. Mater. 2021, 6, 4110–4140. [Google Scholar] [CrossRef]
  177. Vukajlovic, D.; Parker, J.; Bretcanu, O.; Novakovic, K. Chitosan based polymer/bioglass composites for tissue engineering applications. Mater. Sci. Eng. C 2019, 96, 955–967. [Google Scholar] [CrossRef]
  178. Jodati, H.; Yılmaz, B.; Evis, Z. A review of bioceramic porous scaffolds for hard tissue applications: Effects of structural features. Ceram. Int. 2020, 46, 15725–15739. [Google Scholar] [CrossRef]
  179. Mondal, S.; Pal, U. 3D hydroxyapatite scaffold for bone regeneration and local drug delivery applications. J. Drug Deliv. Sci. Technol. 2019, 53, 101131. [Google Scholar] [CrossRef]
  180. Khurshid, Z.; Zafar, M.S.; Hussain, S.; Fareed, A.; Yousaf, S.; Sefat, F. 10—Silver-substituted hydroxyapatite. In Handbook of Ionic Substituted Hydroxyapatites; Khan, A.S., Chaudhry, A.A., Eds.; Woodhead Publishing: Cambridge, UK, 2020; pp. 237–257. [Google Scholar]
  181. Das, A.; Pamu, D. A comprehensive review on electrical properties of hydroxyapatite based ceramic composites. Mater. Sci. Eng. C 2019, 101, 539–563. [Google Scholar] [CrossRef] [PubMed]
  182. Agbeboh, N.I.; Oladele, I.O.; Daramola, O.O.; Adediran, A.A.; Olasukanmi, O.O.; Tanimola, M.O. Environmentally sustainable processes for the synthesis of hydroxyapatite. Heliyon 2020, 6, e03765. [Google Scholar] [CrossRef]
  183. Szcześ, A.; Hołysz, L.; Chibowski, E. Synthesis of hydroxyapatite for biomedical applications. Adv. Colloid Interface Sci. 2017, 249, 321–330. [Google Scholar] [CrossRef] [PubMed]
  184. Mondal, S.; Hoang, G.; Manivasagan, P.; Moorthy, M.S.; Kim, H.H.; Vy Phan, T.T.; Oh, J. Comparative characterization of biogenic and chemical synthesized hydroxyapatite biomaterials for potential biomedical application. Mater. Chem. Phys. 2019, 228, 344–356. [Google Scholar] [CrossRef]
  185. Mondal, S.; Pal, U.; Dey, A. Natural origin hydroxyapatite scaffold as potential bone tissue engineering substitute. Ceram. Int. 2016, 42, 18338–18346. [Google Scholar] [CrossRef]
  186. Athinarayanan, J.; Periasamy, V.S.; Alshatwi, A.A. Simultaneous fabrication of carbon nanodots and hydroxyapatite nanoparticles from fish scale for biomedical applications. Mater. Sci. Eng. C 2020, 117, 111313. [Google Scholar] [CrossRef]
  187. Mahmoud, E.M.; Sayed, M.; El-Kady, A.M.; Elsayed, H.; Naga, S.M. In vitro and in vivo study of naturally derived alginate/hydroxyapatite bio composite scaffolds. Int. J. Biol. Macromol. 2020, 165, 1346–1360. [Google Scholar] [CrossRef]
  188. Chai, Y.; Okuda, M.; Otsuka, Y.; Ohnuma, K.; Tagaya, M. Comparison of two fabrication processes for biomimetic collagen/hydroxyapatite hybrids. Adv. Powder Technol. 2019, 30, 1419–1423. [Google Scholar] [CrossRef]
  189. Prasad, A.; Mohan Bhasney, S.; Sankar, M.R.; Katiyar, V. Fish Scale Derived Hydroxyapatite reinforced Poly (Lactic acid) Polymeric Bio-films: Possibilities for Sealing/locking the Internal Fixation Devices. Mater. Today Proc. 2017, 4, 1340–1349. [Google Scholar] [CrossRef]
  190. Kara, A.; Tamburaci, S.; Tihminlioglu, F.; Havitcioglu, H. Bioactive fish scale incorporated chitosan biocomposite scaffolds for bone tissue engineering. Int. J. Biol. Macromol. 2019, 130, 266–279. [Google Scholar] [CrossRef]
  191. Humayun, A.; Luo, Y.; Mills, D.K. Electrophoretic Deposition of Gentamicin-Loaded ZnHNTs-Chitosan on Titanium. Coatings 2020, 10, 944. [Google Scholar] [CrossRef]
  192. Kang, M.; Kim, S.; Kim, H.; Song, Y.; Jung, D.; Kang, S.; Seo, J.-H.; Nam, S.; Lee, Y. Calcium-Binding Polymer-Coated Poly(lactide-co-glycolide) Microparticles for Sustained Release of Quorum Sensing Inhibitors to Prevent Biofilm Formation on Hydroxyapatite Surfaces. ACS Appl. Mater. Interfaces 2019, 11, 7686–7694. [Google Scholar] [CrossRef] [PubMed]
  193. Zhang, X.; Song, G.; Qiao, H.; Lan, J.; Wang, B.; Yang, H.; Ma, L.; Wang, S.; Wang, Z.; Lin, H.; et al. Novel ternary vancomycin/strontium doped hydroxyapatite/graphene oxide bioactive composite coatings electrodeposited on titanium substrate for orthopedic applications. Colloids Surf. A Physicochem. Eng. Asp. 2020, 603, 125223. [Google Scholar] [CrossRef]
  194. Li, Z.; Du, T.; Ruan, C.; Niu, X. Bioinspired mineralized collagen scaffolds for bone tissue engineering. Bioact. Mater. 2021, 6, 1491–1511. [Google Scholar] [CrossRef]
  195. Lowe, B.; Venkatesan, J.; Anil, S.; Shim, M.S.; Kim, S.-K. Preparation and characterization of chitosan-natural nano hydroxyapatite-fucoidan nanocomposites for bone tissue engineering. Int. J. Biol. Macromol. 2016, 93, 1479–1487. [Google Scholar] [CrossRef] [PubMed]
  196. Li, Y.; Liu, Y.; Li, R.; Bai, H.; Zhu, Z.; Zhu, L.; Zhu, C.; Che, Z.; Liu, H.; Wang, J.; et al. Collagen-based biomaterials for bone tissue engineering. Mater. Des. 2021, 210, 110049. [Google Scholar] [CrossRef]
  197. Zhang, Y.; Hu, M.; Zhang, W.; Zhang, X. Homology of selenium (Se) and tellurium (Te) endow the functional similarity of Se-doped and Te-doped mesoporous bioactive glass nanoparticles in bone tissue engineering. Ceram. Int. 2022, 48, 3729–3739. [Google Scholar] [CrossRef]
  198. Murugan, S.S.; Dalavi, P.A.; Devi, G.V.Y.; Chatterjee, K.; Venkatesan, J. Natural and Synthetic Biopolymeric Biomaterials for Bone Tissue Engineering Applications. In Reference Module in Materials Science and Materials Engineering; Elsevier: Amsterdam, The Netherlands, 2022. [Google Scholar]
  199. Eftekhari, A.; Maleki Dizaj, S.; Sharifi, S.; Salatin, S.; Rahbar Saadat, Y.; Zununi Vahed, S.; Samiei, M.; Ardalan, M.; Rameshrad, M.; Ahmadian, E.; et al. The Use of Nanomaterials in Tissue Engineering for Cartilage Regeneration; Current Approaches and Future Perspectives. Int. J. Mol. Sci. 2020, 21, 536. [Google Scholar] [CrossRef] [Green Version]
  200. Fang, J.; Wang, D.; Hu, F.; Li, X.; Zou, X.; Xie, J.; Zhou, Z. Strontium mineralized silk fibroin porous microcarriers with enhanced osteogenesis as injectable bone tissue engineering vehicles. Mater. Sci. Eng. C 2021, 128, 112354. [Google Scholar] [CrossRef]
  201. Shirzaei Sani, I.; Rezaei, M.; Baradar Khoshfetrat, A.; Razzaghi, D. Preparation and characterization of polycaprolactone/chitosan-g-polycaprolactone/hydroxyapatite electrospun nanocomposite scaffolds for bone tissue engineering. Int. J. Biol. Macromol. 2021, 182, 1638–1649. [Google Scholar] [CrossRef]
  202. Panda, S.; Biswas, C.K.; Paul, S. A comprehensive review on the preparation and application of calcium hydroxyapatite: A special focus on atomic doping methods for bone tissue engineering. Ceram. Int. 2021, 47, 28122–28144. [Google Scholar] [CrossRef]
  203. Yang, L.; Jin, S.; Shi, L.; Ullah, I.; Yu, K.; Zhang, W.; Bo, L.; Zhang, X.; Guo, X. Cryogenically 3D printed biomimetic scaffolds containing decellularized small intestinal submucosa and Sr2+/Fe3+ co-substituted hydroxyapatite for bone tissue engineering. Chem. Eng. J. 2022, 431, 133459. [Google Scholar] [CrossRef]
  204. Wan, Z.; Zhang, P.; Liu, Y.; Lv, L.; Zhou, Y. Four-dimensional bioprinting: Current developments and applications in bone tissue engineering. Acta Biomater. 2020, 101, 26–42. [Google Scholar] [CrossRef] [PubMed]
  205. Meng, J.; Boschetto, F.; Yagi, S.; Marin, E.; Adachi, T.; Chen, X.; Pezzotti, G.; Sakurai, S.; Yamane, H.; Xu, H. Design and manufacturing of 3D high-precision micro-fibrous poly (l-lactic acid) scaffold using melt electrowriting technique for bone tissue engineering. Mater. Des. 2021, 210, 110063. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the relationship between materials and necessary properties for BTE scaffolds.
Figure 1. Schematic representation of the relationship between materials and necessary properties for BTE scaffolds.
Polymers 14 00899 g001
Figure 2. Mechanism of preventing bacterial attachment by using surface coatings [41].
Figure 2. Mechanism of preventing bacterial attachment by using surface coatings [41].
Polymers 14 00899 g002
Figure 3. Natural sources of CaPs.
Figure 3. Natural sources of CaPs.
Polymers 14 00899 g003
Figure 4. (A) SEM images of fracture and external surfaces, (B) Confocal images of adhered cells on the sintered pellets of obtained materials. Notation corresponds to the provenience of the natural source and sintering temperature: eggshells (ES), cuttlebones (CB), mussel shells (MS), and amorphous calcium carbonate (ACC) [117].
Figure 4. (A) SEM images of fracture and external surfaces, (B) Confocal images of adhered cells on the sintered pellets of obtained materials. Notation corresponds to the provenience of the natural source and sintering temperature: eggshells (ES), cuttlebones (CB), mussel shells (MS), and amorphous calcium carbonate (ACC) [117].
Polymers 14 00899 g004
Figure 5. (A) SEM images of reference, seashell, and marble materials (on the magnification of 10 and 100 µm); (B) Fluorescence microscopy images of the MC3T3-E1 cells [147].
Figure 5. (A) SEM images of reference, seashell, and marble materials (on the magnification of 10 and 100 µm); (B) Fluorescence microscopy images of the MC3T3-E1 cells [147].
Polymers 14 00899 g005
Figure 6. Fluorescence microscopy images of human gingival fibroblast cells morphology on the surface of (a) control (polystyrene surface (b) simple Ti, (c) commercial HA, (d) sea bream-BCP, and (e) salmon-BCP coatings [157].
Figure 6. Fluorescence microscopy images of human gingival fibroblast cells morphology on the surface of (a) control (polystyrene surface (b) simple Ti, (c) commercial HA, (d) sea bream-BCP, and (e) salmon-BCP coatings [157].
Polymers 14 00899 g006
Figure 7. The connection between the properties of the material applied as a surface coating [167].
Figure 7. The connection between the properties of the material applied as a surface coating [167].
Polymers 14 00899 g007
Table 1. General characteristics of natural polymers and their applications in BTE.
Table 1. General characteristics of natural polymers and their applications in BTE.
PolymerPropertiesApplicationsReferences
AdvantagesDisadvantages
CollagenGreat biocompatibility, biodegradability, cytocompatibility, non-toxicityPoor mechanical strengthScaffolds, drug delivery systems, 3D printing[26]
AlginateBiodegradability, biocompatibility, bioresorbability, non-toxicity, presenting synergic effects with bioactive componentsPoor mechanical strength and bioactivityBone tissue applications[27,28]
ChitosanSuperior biocompatibility; biodegradability, anti-inflammatoryPoor stability, mechanical strengthHydrogels, scaffolds, microspheres[29]
Hyaluronic AcidGreat biocompatibility, biodegradability, cell adhesion, proliferation, and differentiationPoor mechanical properties, high degradation rateScaffolds, hydrogel[30,31,32]
Bacterial celluloseGood water absorption, mechanical strength and structural properties, good cell adhesion and biocompatibility, continuous structural supportLow biodegradability in the human body and biological activity3D scaffolds, bone tissue replacements[33,34]
Silk fibroinIncreased flexibility, biocompatibility, with good mechanical strengthReduced biodegradation rateScaffolds[35]
GelatinGreat biocompatibility, biodegradability, non-toxicity, improved cell adhesion, and proliferationPoor mechanical properties, high biodegradation rateScaffolds for hard tissue engineering[26,36,37]
Table 2. General characteristics of synthetic polymers and their applications in BTE.
Table 2. General characteristics of synthetic polymers and their applications in BTE.
PolymerPropertiesApplicationsReferences
AdvantagesDisadvantages
Polylactic acid (PLA)Superior tensile strength, elongation, and modulus, biodegradability, and minimal inflammatory responseLow toughness, mechanical support, insufficient biocompatibilityLoad bearing applications, orthopedic repair, suture anchors, scaffolds[27,55,56]
Poly(ε-caprolactone) (PCL) Good biodegradability, biocompatibility, low Young’s modulus, tailorable physical properties, reduced degradation ratePoor cell adhesion, hydrophobic natureScaffolds, BTE, 3D bioprinting[55,57,58]
Poly(glycolic acid) (PGA)High crystallinity; great mechanical strength, good cell adhesion, proliferation, and differentiationHydrophobic natureScaffolds, BTE[58,59]
Poly(vinyl alcohol) (PVA)Biocompatibility, biodegradability, good compressive mechanical and elastic strengthLow bioactivity, decreased cell attachmentScaffolds, drug delivery systems[60,61,62]
Poly(ethylene glycol) (PEG)Biocompatibility, hydrophilicity, able to improve degradation, non-toxicity, and non-immunogenicity combined with different polymers, enhanced enzymatic stabilityLimited tailorable mechanical property and rheological behavior, reduced bioactivityScaffolds, BTE, 3D bioprinting, orthopedic implant[58,63,64]
Poly(lactic-co-glycolic acid) (PLGA)Excellent biocompatibility, processability, good mechanical strength, adjustable degradation rate, and minimal inflammatory responsePossible inflammatory response, low bioactivityScaffolds, orthopedic implants, drug delivery systems[65,66]
Poly(methyl methacrylate) (PMMA)Processability, durabilityNon-degradabilityScaffolds[66]
Table 3. Examples of CaPs obtained from natural sources.
Table 3. Examples of CaPs obtained from natural sources.
Natural SourceCrystalline PhaseMorphologyApplicationReferences
FishboneHydroxyapatiteLaminar and irregular structure, 149–325 nmSurface coating; nutrition[96,97,98,99]
Biphasic calcium phosphate30–100 nm, as nanorodsScaffolds[100]
EggshellsBiphasic calcium phosphateSpherical structureOrthopedic and dental applications[101]
α-Tricalcium phosphateCompact and agglomerated structureScaffolds; dental reconstruction[102,103]
ꞵ-Tricalcium phosphateRound shape, with dimensions between 150 nm–2 µmScaffold in dental and orthopedic reconstruction[104]
HydroxyapatiteIrregularly shaped, with sizes between 10–18 µmReinforcing filler; biomedical devices[105,106]
HydroxyapatiteFlower-like, with the aspect of hexagonal rods and dimensions between 200–300 nmBiomedical applications[107]
SeashellsHydroxyapatiteNano-rods, with sizes between 20–90 nmBTE; drug delivery; dentalApplications; coating[108,109]
Fish scalesHydroxyapatiteDimensions between 20–60 nm, in the form of agglomerations or nano-rodsCoating; dental applications; bone graft; filler[110,111,112,113]
Table 4. Required properties of scaffolds in BTE applications.
Table 4. Required properties of scaffolds in BTE applications.
PropertyRequired CharacteristicsReferences
BiodegradabilityThe material should possess a prearranged biodegradability to improve the composition of different tissue. In this manner, the biodegradable matrices offer temporary scaffolds within defects into the bone tissues to improve their regeneration.[174]
BiocompatibilityThe composite material must perform with a suitable host response in the regeneration of bone tissue. This ability must be in synchronization with osseous tissue without producing damaging changes.[175]
Mechanical PropertiesSurface roughness enhances cell attachment, differentiation, and maturation. Moreover, scaffolds’ mechanical stability supports their adhesion to the neighboring tissue. These properties enhance the adsorption of adhesive proteins (e.g., fibrin), leading to an improved osteogenic cell attachment, proliferation, and differentiation into osteoblasts, to further bone production integrated within the scaffold.[55,176]
PorosityNeeds be tuned, as the initial porosity must be low or else the scaffold resorbs very fast, incapacitating the mechanical support to further affect novel tissue growth. On the other hand, materials with a low degradation rate can possess high porosity, optimizing the degradation of the scaffold.[177]
BioactivityThis characteristic is essential to improve ECM development through the stimulation of cellular behavior andcan contribute to the cells the molecular signals.[174,176]
ProcessabilityThe composite material should be easily processed to design various formulations and configurations such as nanometric, 3D scaffolds, micro-metric particles, and/or injectable formulations.[175]
Immune response and toxicityThe obtained materials must be non-cytotoxic and allow cell attachment to function properly, proliferate and differentiate. Moreover, they must possess non-inflammatory properties and induce a minimal immune response.[178]
Controlled DeliveryTo deliver biomolecules in BTE applications, it is mandatory to develop scaffolds as a drug delivery system. Additionally, the biological activities of these biomolecules and interaction among surrounding cells in the bone-healing process are the foundation for the fabrication of BTE scaffolds.[176]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Radulescu, D.-E.; Neacsu, I.A.; Grumezescu, A.-M.; Andronescu, E. Novel Trends into the Development of Natural Hydroxyapatite-Based Polymeric Composites for Bone Tissue Engineering. Polymers 2022, 14, 899. https://doi.org/10.3390/polym14050899

AMA Style

Radulescu D-E, Neacsu IA, Grumezescu A-M, Andronescu E. Novel Trends into the Development of Natural Hydroxyapatite-Based Polymeric Composites for Bone Tissue Engineering. Polymers. 2022; 14(5):899. https://doi.org/10.3390/polym14050899

Chicago/Turabian Style

Radulescu, Diana-Elena, Ionela Andreea Neacsu, Alexandru-Mihai Grumezescu, and Ecaterina Andronescu. 2022. "Novel Trends into the Development of Natural Hydroxyapatite-Based Polymeric Composites for Bone Tissue Engineering" Polymers 14, no. 5: 899. https://doi.org/10.3390/polym14050899

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop