Next Article in Journal
Expression of Hypoxia-Inducible Factor 1α (HIF-1α) and Genes of Related Pathways in Altered Gravity
Next Article in Special Issue
Are the Biological and Biomechanical Properties of Meniscal Scaffolds Reflected in Clinical Practice? A Systematic Review of the Literature
Previous Article in Journal
Extracellular Vesicle-Mediated Cell–Cell Communication in the Nervous System: Focus on Neurological Diseases
Previous Article in Special Issue
Osteochondral Tissue Regeneration Using a Tyramine-Modified Bilayered PLGA Scaffold Combined with Articular Chondrocytes in a Porcine Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Current Trends in Fabrication of Biomaterials for Bone and Cartilage Regeneration: Materials Modifications and Biophysical Stimulations

Chair and Department of Biochemistry and Biotechnology, Medical University of Lublin, W. Chodzki 1 Street, 20-093 Lublin, Poland
Int. J. Mol. Sci. 2019, 20(2), 435; https://doi.org/10.3390/ijms20020435
Submission received: 21 December 2018 / Revised: 15 January 2019 / Accepted: 18 January 2019 / Published: 20 January 2019
(This article belongs to the Special Issue Biomaterials for Musculoskeletal System)

Abstract

:
The aim of engineering of biomaterials is to fabricate implantable biocompatible scaffold that would accelerate regeneration of the tissue and ideally protect the wound against biodevice-related infections, which may cause prolonged inflammation and biomaterial failure. To obtain antimicrobial and highly biocompatible scaffolds promoting cell adhesion and growth, materials scientists are still searching for novel modifications of biomaterials. This review presents current trends in the field of engineering of biomaterials concerning application of various modifications and biophysical stimulation of scaffolds to obtain implants allowing for fast regeneration process of bone and cartilage as well as providing long-lasting antimicrobial protection at the site of injury. The article describes metal ion and plasma modifications of biomaterials as well as post-surgery external stimulations of implants with ultrasound and magnetic field, providing accelerated regeneration process. Finally, the review summarizes recent findings concerning the use of piezoelectric biomaterials in regenerative medicine.

Graphical Abstract

1. Introduction: Biomaterials for Bone and Cartilage Regeneration

The aim of the engineering of biomaterials is to fabricate a biocompatible scaffold that would support or ideally enhance regeneration of tissue after biomaterial implantation within the injured area. Optimal scaffold for bone and cartilage regeneration should have good surgical handiness and mechanical properties (especially Young’s modulus value) adjusted to the implantation area, e.g., Young’s modulus (E) value = 0.4–0.83 MPa for articular cartilage [1,2], E = 0.5–3 GPa for trabecular bone [3], and E = 15–23 GPa for cortical bone [4,5]. It should be noted that flexible biomaterials possessing lower stiffness (lower E value) compared to the host bone can be used only in non-load bearing implantation areas. These kinds of materials under exposure to the mechanical load may exert constant physical pressure to the surrounded bone tissue, resulting in excessive ossification. When material reveals superior stiffness compared to the host tissue at the implantation area, implant takes whole mechanical load and shields the bone cells from mechanical stimulus responsible for induction of bone formation process, causing excessive bone resorption and implant loosening [6,7]. Therefore, to achieve biomaterial showing mechanical properties close to natural bone, researchers primarily produce composite materials made of polymers—mimicking flexible organic part of the bone—and bioceramics, e.g., hydroxyapatite (HA) [8,9,10,11,12], α- and β-tricalcium phosphate (TCP) [13,14], low-temperature calcium phosphate cements (CPCs) [15,16,17,18], or bioactive glass (BG) [19,20], which imitate natural mineral of the bone (HA), providing better mechanical properties of the resultant scaffold. Biomaterials for cartilage regeneration should be more elastic compared to the bone scaffolds and are mainly produced using biocompatible polymers (natural and synthetic) often resembling those occurring in natural cartilage tissue, e.g. chitosan, hyaluronic acid, collagen, fibrin, silk, alginate, polylactic acid (PLA), poly(3-caprolactone) (PCL), or poly(l-lactide-co-caprolactone) (PLCL) [21,22]. To obtain better mechanical and biological properties, cartilage scaffolds are also produced as composite materials, using combination of different polymers, e.g., agarose and poly(ethyleneglycol) (PEG) [23], collagen and PEG [24], or chitosan and hyaluronic acid [25].
Biomaterials for regenerative medicine applications should be favorable to cell adhesion, proliferation and differentiation to ensure rapid regeneration process at the site of injury. To obtain highly biocompatible scaffolds promoting cell adhesion and growth, materials scientists are still searching for novel modifications of biomaterials to obtain higher surface roughness, wettability, and surface free energy, allowing for better adsorption of cell adhesive proteins (e.g., laminin, fibronectin, vitronectin) and thereby more effective cell attachment [26,27,28,29,30]. Furthermore, to obtain better regeneration rate, scientists reach for such a solution like external physical stimulation of biomaterial after its implantation with low-intensity pulsed ultrasound (LIPUS) [31], magnetic field [32], or electrical forces [33]. Ideally, biomaterials for bone and cartilage regeneration should not only promote new tissue formation at the site of injury, but also protect the wound against biodevices-related infections, which may cause prolonged inflammation and biomaterial failure [34]. Thus, there is a great tendency in engineering of biomaterials to produce implants with antibacterial activity against Gram-positive and Gram-negative bacteria which are typical of infections related to orthopedic surgery: Staphylococcus aureus, Staphylococcus epidermidis, Escherichia coli, and Pseudomonas aeruginosa [34,35,36].
This review presents current trends in the field of engineering of biomaterials concerning application of various modifications and biophysical stimulation of biomaterials to obtain implants allowing for fast regeneration process and providing long-lasting antimicrobial protection at the site of injury.

2. Biomaterial Modifications with Metal Ions

2.1. Biomaterials with Osteopromotive Properties

Researchers have recently focused on the modifications of biomaterials with various metal ions in order to give them better biological properties. Biomaterials are usually modified by ionic substitution of bioceramics (HA and BG) or by incorporation/deposition of metal ions on the surface of the metal implants. Since the main goal of engineering of biomaterials and tissue engineering is enhancement of regeneration process at the site of implantation, there is a growing trend towards modification of implantable materials with metal ions, which provide osteopromotive properties (Table 1), e.g., magnesium (Mg2+), zinc (Zn2+), copper (Cu2+), strontium (Sr2+), cobalt (Co2+), lithium (Li+), and fluoride (F) [37]. Galli et al. [38] conducted physical deposition of Mg ions within mesoporous titanium (Ti) films coated on titanium threaded screws and demonstrated that local release of Mg ions significantly improved implant osseointegration with tibia of rabbits. Liu et al. [39] produced biodegradable Mg–Cu alloys that thanks to constant Cu2+ and Mg2+ release were proved to enhance bone formation (in vitro mouse calvarial preosteoblasts model—MC3T3-E1 cell line) and angiogenesis (in vitro model of human umbilical vein endothelial cells—HUVEC cell line) and also to possess long-lasting antibacterial properties (studies on S. aureus). Yusa et al. [40,41] modified titanium surfaces with zinc to obtain biomaterials with large potential in orthopedic applications. The release of Zn ions from Zn-modified titanium implant promoted osteogenic differentiation leading to prominent bone formation process (studies on in vitro cellular models: human bone marrow-derived mesenchymal stem cells—BMDSCs and human dental pulp stem cells—DPSCs). Similarly, Thian et al. [42] synthesized Zn-doped HA and demonstrated enhanced proliferation and differentiation potential of human adipose tissue-derived mesenchymal stem cells (ADSCs) on the biomaterial made of Zn-doped HA. In turn, Andersen et al. [43] showed that Sr-modified surface of titanium implant improved osteogenic differentiation of human DPSCs and osseointegration of implant with femur bone of rats. Fluoride ions are also known to promote osteogenic differentiation and thus they are often used for HA substitution [44]. Uysal et al. [45] synthesized HA co-doped with Zn2+ and F ions and showed that modified HA not only possessed better mechanical properties compared to pure HA, but also increased proliferation and bone alkaline phosphatase activity (bALP) of osteosarcoma derived osteoblast-like cells (Saos-2 cell line). Ionic substitution is also often applied for BGs. Khorami et al. [46] used Li+ for substitution of 45S5 BG and proved that modified biomaterial provided higher proliferation rate and bALP activity of primary rat calvarial osteoblasts. Whereas Wu et al. [47,48] produced Co- and Cu-containing mesoporous BGs, which were demonstrated to promote proliferation and osteogenic differentiation of human BMDSCs.

2.2. Biomaterials with Antimicrobial Properties

Because implant-related infections are a severe problem in orthopedic surgery, more often researchers put emphasis on the modification of biomaterials with metal ions revealing antimicrobial properties, such as: Zn2+, Cu2+, and predominantly silver (Ag+) which is commonly known for its broad spectrum of antimicrobial activity against bacteria and fungi (Table 2). Nevertheless modification of biomaterials with antimicrobial metal ions has some limitations since high concentrations of Ag+, Zn2+, and Cu2+ showing satisfactory antimicrobial action may also be lethal to the eukaryotic cells.
Rau et al. [64] fabricated calcium phosphate bone cements containing Ag-doped TCP revealing the inhibitory effect towards pathogenic E. coli, but the paper does not present cytotoxicity studies on eukaryotic cells. Whereas in our recent studies [65] we produced α-TCP-based bone cement containing CaCO3 and Ag-doped HA showing antibacterial effect against S. aureus, S. epidermidis, and E. coli. However, produced bone cements revealed also cytotoxic effect against eukaryotic cells (MC3T3-E1) (Figure 1). In turn, Lim et al. [66] produced Ag-doped HA and demonstrated its high antibacterial activity against S. aureus and non-toxicity against human BMDSCs. Other metal ions that are commonly used for antimicrobial modifications of biomaterials include Zn2+ and Cu2+. Thian et al. [42] produced Zn-doped HA that significantly reduced growth of S. aureus bacteria on its surface and showed non-toxicity against human ADSCs. While Wu et al. [48] used Cu2+ for substitution of BG and obtained biocompatible (studies on human BMDSCs) porous biomaterial with high antibacterial activity against E. coli. Interestingly, apart from typical antimicrobial metal ions, also lithium (Li+) [63] and cerium (Ce3+) [62,67] have been reported to possess potential to be used as an effective antimicrobial agent for biomaterial modification.

3. Plasma-Modified Biomaterials

3.1. Biomaterials with Improved Biocompatibility

Plasma treatment may be used for surface functionalization with hydrophilic chemical groups and for improvement of surface properties of biomaterials like wettability, roughness or surface free energy. All mentioned features have great impact on cell adhesion and thereby material biocompatibility. The effectiveness of plasma modification highly depends on substrate gas used for the treatment, reactor design, or the type of biomaterial subjected to the modification. In engineering of biomaterials, plasma (recently special attention has been paid to atmospheric pressure plasma) combined with argon, oxygen, air, amonia, or nitrogen gas is most often used for surface modifications of primarily polymeric materials [68,69,70,71]. Kostov et al. [72] modified polyethyleneterephthalate, polyethylene (PE), and polypropylene with cold atmospheric plasma jet using argon gas and obtained polymers with increased roughness and wettability. Sagbas et al. [71] used argon/oxygen atmospheric plasma to increase wettability and wear resistance of ultra-high molecular weight PE. They proved that surface treatment with argon plasma followed by oxygen plasma led to incorporation of polar groups (mainly hydroxyl groups), which are known to promote cell adhesion. Jordá-Vilaplana et al. [73] increased surface free energy as well as improved wettability and roughness of PLA material by its modification with air atmospheric plasma. Griffin [70] et al. compared the effect of argon, oxygen, and nitrogen plasma treatment on biocompatibility of polyurethane (PU) scaffold. They demonstrated that among all tested plasma treatments, oxygen plasma had the greatest effect on surface properties of the scaffold. The oxygen plasma significantly decreased water contact angle (increased wettability) and meaningfully increased roughness of the material. Nevertheless, argon-modified material not only revealed the highest protein adsorption ability and human dermal fibroblast adhesion effectiveness, but also the highest tissue integration and angiogenesis after scaffold subcutaneous implantation in a mouse model, compared to oxygen and nitrogen-plasma treated scaffolds. In another study, Griffin et al. [74] used oxygen plasma for functionalization of PU-based scaffold with amino (NH2) groups (plasma polimerisation with the use of allylamine) and carboxyl (COOH) groups (plasma polymerisation with the use of acrylic acid monomers). They showed that plasma functionalization of the biomaterials significantly improved their biocompatibility since COOH-functionalized material promoted chondrogenic differentiation, whereas NH2-functionalized scaffold enhanced osteogenic differentiation of human ADSCs. Chen et al. [69] used argon and nitrogen plasma for improvement of biocompatibility of polyurethane methacrylate (PUMA) and off-stoichiometry thiol-ene (OSTE-80) polymers. Plasma treatment introduced oxygen and nitrogen moieties on the surface of polymers, increasing surface energy, hydrophilicity, and chemical functionalities. As a consequence, biomaterials exhibited ability to promote HUVEC adhesion and proliferation. In turn, Wang et al. [75] applied helium cold atmospheric plasma for modification of PLA scaffold and they observed that plasma-treated polymer had significantly decreased water contact angle (from 70 ± 2° to 24 ± 2°) and drastically increased nano-scale roughness, resulting in improved biocompatibility (enhanced attachment and proliferation of primary human osteoblast and human BMDSC).

3.2. Biomaterials with Antibacterial Properties

Plasma treatment is widely used not only for improvement of biocompatibility of biomaterials, but also to provide them antimicrobial properties. The antibacterial properties of medical materials may be achieved by immobilization of organic active reagents or introducing metal ions (Zn, Cu, Ag) into the surface as well as by modifications of surface morphologies (topography) and chemical composition (e.g., by introducing specific functional groups) to prevent attachment and proliferation of pathogenic bacteria [76,77]. To produce antibacterial films on the surface of biomaterials, some plasma-assisted modification techniques have been recently more often used, e.g. metal or gas plasma immersion ion implantation (PIII) or magnetron sputtering [76]. Zhang et al. [78] applied PIII to incorporate Ag into PE and obtained biocompatible (studies on normal human fetal osteoblast cell line—hFOB 1.19) and antibacterial surface (studies on E. coli). They also demonstrated that application of Ag PIII in combination with nitrogen PIII provided prolonged antibacterial activity; however nitrogen-containing functional groups (such as C–N and C=N) formed on the surface of PE reduced proliferation of hFOB 1.19 osteoblasts. In other studies by Zhang et al. [79,80], PE and medical poly(vinyl chloride) samples were modified using oxygen plasma, then precoated with antibacterial agents (triclosan or bronpol), followed by argon PIII to ensure that active agents strongly bonded to the surface of materials. Modified PE and poly(vinyl chloride) samples showed excellence antibacterial activity against E. coli and S. aureus, however the papers do not present biocompatibility test on eukaryotic cells. Whereas Wang et al. [81] applied nitrogen PIII for the production of biocompatible (studies on primary rat calvarial osteoblasts) poly(butylene succinate) samples with antibacterial activity against E. coli and S. aureus.

4. External Biophysical Stimulation of the Biomaterials

4.1. Low-Intensity Pulsed Ultrasound Stimulation

Low-intensity pulsed ultrasound (LIPUS) is an external stimulation, which according to basic and clinical studies has ability to improve bone and cartilage healing process by promoting proliferation and differentiation of mesenchymal stem cells (MSCs) [82,83,84,85]. Moonga et al. [86] proved that LIPUS treatment reduced cell proliferation but significantly enhanced ECM mineralization of MC3T3-E1 preosteoblasts cultured directly on bovine trabecular bone scaffold. Carina et al. [31] demonstrated that 20-minute exposure of Mg-HA/collagen scaffold pre-seeded with human MSCs to the LIPUS improved colonization of biomaterial and enhanced osteogenic differentiation of stem cells. Zhou et al. [87] determined optimal LIPUS parameters for bone regeneration process (1.5 MHz, 20% duty cycle with 150 mW/cm2 intensity) and revealed that PE-based 3D printed scaffold containing arginine-glycine-aspartic acid-serene (RGDS) peptide and nanocrystalline HA under LIPUS stimulation greatly promoted proliferation and osteogenic differentiation of BMDSCs. In turn, Aliabouzar et al. [88] investigated the effect of LIPUS treatment on proliferation and chondrogenic differentiation of human MSCs seeded onto 3D printed PEG- diacrylate scaffold. They demonstrated that LIPUS stimulation (optimal parameters were found to be 1.5 MHz, 20% duty cycle with 100 mW/cm2 intensity) significantly increased MSC proliferation as well as enhanced glycosaminoglycan (GAG) and type II collagen synthesis. In another study, Aliabouzar et al. [89] used lipid-coated microbubbles combined with LIPUS to enhance proliferation and chondrogenic differentiation of human MSCs seeded onto 3D printed PEG-diacrylate hydrogel scaffold. They proved that application of LIPUS along with microbubbles significantly increased MSC proliferation up to 40% (up to 18% with LIPUS alone), promoted GAG synthesis by 17% (by 5% with LIPUS alone), and enhanced type II collagen production by 78% (by 44% with LIPUS alone).

4.2. Magnetic Field Stimulation

Magnetic field is another type of external biophysical stimulation proved to substantially improve regeneration process [32,90]. Therefore there is a trend in engineering of biomaterials towards production of biomaterials with magnetic nanoparticles (MNPs) incorporated [91,92]. Novel approach to bone and cartilage regeneration includes the use of MNPs-loaded scaffolds combined with magnetic field stimulation in order to accelerate tissue regeneration and in the case of bone implants also angiogenesis process. Samal et al. [93] fabricated MNPs-loaded silk fibroin scaffold, which exhibited outstanding hyperthermia properties under exposure to magnetic field and revealed capability to enhance adhesion and colonization of MC3T3-E1 preosteoblasts. Cai et al. [94] produced magnetic scaffold made of poly(l-lactide) (PLLA) nanofibers loaded with ferromagnetic (Fe3O4) nanoparticles (NPs), which was demonstrated to promote proliferation and osteogenic differentiation of MC3T3-E1 cells not only after stimulation with static magnetic field, but also in the absence of external stimulation. Interestingly, in our recent studies it was observed that the incorporation of ferromagnetic NPs within the scaffold itself (without external stimulation with magnetic field) may be sufficient to improve biocompatibility of the resultant biomaterial since MC3T3-E1 preosteoblasts cultured on the surface of magnetic chitosan-based scaffold exhibited better spreading and faster proliferation compared to the cells grown on the scaffold without ferromagnetic NPs (Figure 2). In turn, Yun et al. [95] showed enhanced osteogenic differentiation of primary mouse osteoblasts seeded onto magnetic MNPs-loaded PCL scaffold upon exposure to static magnetic field. Moreover magnetic field stimulation of MNPs/PCL scaffold implanted in mouse calvarium defects resulted in significantly improved bone regeneration (as demonstrated by the histological and microcomputed tomography analyses). Yan et al. [96] fabricated novel biocompatible (studies on human MG-63 osteosarcoma-derived cell line) magnetic scaffold by incorporating a spherical core-shell nano-iron oxide-hydroxyapatite (Fe3O4-HA) into PU. Whereas Heidari et al. [97] developed biocompatible (studies on human MSCs) chitosan/HA/nano magnetite (nano-Fe3O4) composite with improved mechanical properties. D’Amora et al. [98] produced magnetic scaffolds by incorporating Fe-HA NPs into a PCL matrix and demonstrated that magnetic stimulation led to improved proliferation of human MSCs, which were also better spread compared to the cells grown on the surface of unstimulated scaffolds. Aliramaji et al. [99] developed magnetic silk fibroin/chitosan/Fe2O3 scaffold and proved its great biocompatibility under a static magnetic field. While Anjaneyulu et al. [100] demonstrated that magnetic HA/Fe3O4 coatings on the Ti-6Al-4V material significantly improved bioactivity of the resultant biomaterial by developing Fe-OH groups on the surface.

5. Piezoelectric Biomaterials

Piezoelectric biomaterials possess the ability to generate a bioelectrical signal upon exposure to mechanical stress (without an external power source), which subsequently promotes tissue regeneration at the site of implantation [101]. Since piezoelectric scaffolds resemble sensitive mechano-electrical transducers, they are usually applied to the implantation areas, which are exposed to mechanical loads [101,102]. Nevertheless apart from physiological mechanical load, piezoelectric implants may be also subjected to ultrasound [103] or electrical stimulation [102,104]. The piezoelectric biomaterials produce electric signal, which is similar to that occurring in natural extracellular matrix (ECM) during remodeling process of cartilage and bone tissue. In physiological environment, the compressive force on collagen fibers within ECM leads to the re-organization of dipole moment, resulting in negative charges. As a consequence, electrical signal reaches the cell membrane, leading to the opening of voltage-gated calcium channels. Increased intracellular calcium concentration results in activation of calmodulin, followed by the activation of calcineurin, which causes dephosphorylation of Nuclear Factor of Activated Cells (NF-AT) and its translocation into the nucleus, where along with other transcription factors NF-AT regulates expression of some growth factors, inter alia Transforming Growth Factor-β (TGF-β) and Bone Morphogenetic Proteins (BMPs), which are known to play crucial role in the formation of cartilage and bone [103].
Since the discovery of piezoelectricity in natural cartilage and bone, many piezoelectric biomaterials for regenerative medicine applications have been developed (Table 3). Chen et al. [105] fabricated biocompatible potassium sodium niobate (KNN) ceramics with piezoelectric constant of ~93 pC/N, which promoted growth of MC3T3-E1 preosteoblasts and revealed better protein adsorption capacity compared to non-polarized ceramics. Lv et al. [106] developed 0–3-type niobate-based lead-free composite ceramics with ZnO inclusions showing excellent piezoelectric properties and superior temperature stability. Tang et al. [107] produced HA/barium titanate (BaTiO3) composite material with piezoelectric constant of 1.3 pC/N to 6.8 pC/N dependent on the BaTiO3 content, which promoted osteoblast growth. Augustine et al. [108] modified widely known piezoelectric polymer-poly (vinylidene fluoride-trifluoroethylene) (PVDF-TrFE) by incorporation of zinc oxide (ZnO) nanoparticles and demonstrated that human MSCs and HUVECs cultured on the resultant scaffold showed higher cell viability, enhanced adhesion and proliferation compared to the cells grown on PVDF-TrFE material without ZnO. Damaraju et al. [109] investigated the effect of 3D fibrous PVDF-TrFE scaffold on osteogenic differentiation of human MSCs. They proved that piezoelectric scaffold exhibiting low voltage output promoted chondrogenic differentiation of MSC, whereas PVDF-TrFE scaffold with a high voltage output induced osteogenic differentiation of stem cells. Moreover, they revealed that electromechanical stimulus was more effective in promotion of cell differentiation compared to mechanical load.

6. Concluding Remarks

Modern strategy to bone and cartilage regeneration includes the use of biomaterials to support new tissue formation and to accelerate healing process at the implantation area. However, biomaterials (especially made of natural biopolymers and containing active biomolecules) may induce biomaterial-related immune response. It should be noted that prolonged inflammation may result in oxidative damage of the implant and its loosening. Moreover, the surgery of biomaterial implantation is an invasive procedure, which is always associated with the injury of the tissue, carrying a risk of post-surgery infections. Therefore, scientists with the expertise in the field of engineering of biomaterials are still searching for new modifications of biomaterials to improve their biocompatibility, provide accelerated regeneration process as well as antimicrobial protection at the implantation site. To achieve this goal, biomaterials are often modified with the use of active biomolecules (e.g., growth factors, cytokines, Arg-Gly-Asp (RGD) sequences, antibiotics) or metal ions revealing antimicrobial or osteopromotive properties. However, implantable materials made of natural biopolymers or containing biological components (e.g., cells or cellular products) carry a risk of elicitation of immunogenic immune response, which is induced by dendritic cells followed by activation of lymphocytes [119]. Furthermore, clinical application of biomaterials containing active biomolecules may disturb physiological balance between growth factors and cytokines naturally occurring in a living organism. Regeneration of bone and cartilage tissue is controlled by local release of a number of biological molecules (growth factors and cytokines) followed by crosstalk between different cell types, inter alia osteoblasts, osteoclasts, macrophages, stem cells, neutrophils, and natural killers [120,121,122,123]. Therefore, the changes in local concentrations of some cytokines/growth factors caused by modified with biomolecules material may exert negative effect on subsequent regeneration process after implantation. Thus, there is a clear need to understand the kind of the modes of action of biomolecules used for biomaterial modification in order to predict potential effect of modified implant on response of immune and precursor cells [124]. In this context, post-surgery biophysical stimulation of regeneration process appears to be safer solution in modern regenerative medicine. Consequently, magnetic field, ultrasound or electrical forces are more often applied to provide better healing process after biomaterial implantation. Taking into account great technological development, it is very likely that in the nearest future, routine treatment of cartilage and bone defects will involve implantation of smart antimicrobial biomaterial and remote stimulation of healing process by the patient at home by simply wearing a special orthopedic brace generating e.g. LIPUS.

Funding

The APC was funded by National Science Centre (NCN) within M-Era.Net 2 research project no. UMO-2016/22/Z/ST8/00694 entitled “Nanovectors engineered for plasma enhanced theranostics in regenerative medicine”.

Conflicts of Interest

The author declares no conflict of interest.

Abbreviations

ADSCsAdipose tissue-derived mesenchymal stem cells
bALPBone alkaline phosphatase
BGBioactive glass
BMDSCsBone marrow-derived mesenchymal stem cells
BMPsBone Morphogenetic Proteins
CPCCalcium phosphate cement
DPSCsDental pulp stem cells
EYoung’s modulus
ECMExtracellular matrix
GAGGlycosaminoglycan
HAHydroxyapatite
HUVECsHuman umbilical vein endothelial cells
KNNPotassium sodium niobate
LIPUSLow-intensity pulsed ultrasound
MNPsMagnetic nanoparticles
MSCsMesenchymal stem cells
NF-ATNuclear Factor of Activated Cells
NPsNanoparticles
PCLPoly(3-caprolactone)
PEPolyethylene
PEGPoly(ethylene glycol)
PIIIPlasma immersion ion implantation
PLAPolylactic acid
PLCLPoly(l-lactide-co-caprolactone)
PUPolyurethane
PVDFPoly (vinylidene fluoride)
PVDF-TrFEPoly (vinylidene fluoride-trifluoroethylene)
TCPTricalcium phosphate

References

  1. Sakai, N.; Hashimoto, C.; Yarimitsu, S.; Sawae, Y.; Komori, M.; Murakami, T. A functional effect of the superficial mechanical properties of articular cartilage as a load bearing system in a sliding condition. Biosurf. Biotribol. 2016, 2, 26–39. [Google Scholar] [CrossRef] [Green Version]
  2. Töyräs, J.; Lyyra-Laitinen, T.; Niinimäki, M.; Lindgren, R.; Nieminen, M.T.; Kiviranta, I.; Jurvelin, J.S. Estimation of the Young’s modulus of articular cartilage using an arthroscopic indentation instrument and ultrasonic measurement of tissue thickness. J. Biomech. 2001, 34, 251–256. [Google Scholar] [CrossRef]
  3. Wang, J.; Zhou, B.; Liu, X.S.; Fields, A.J.; Sanyal, A.; Shi, X.; Adams, M.; Keaveny, T.M.; Guo, X.E. Trabecular plates and rods determine elastic modulus and yield strength of human trabecular bone. Bone 2015, 72, 71–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Turner, C.H.; Rho, J.; Takano, Y.; Tsui, T.Y.; Pharr, G.M. The elastic properties of trabecular and cortical bone tissues are similar: Results from two microscopic measurement techniques. J. Biomech. 1999, 32, 437–441. [Google Scholar] [CrossRef]
  5. Mirzaali, M.J.; Schwiedrzik, J.J.; Thaiwichai, S.; Best, J.P.; Michler, J.; Zysset, P.K.; Wolfram, U. Mechanical properties of cortical bone and their relationships with age, gender, composition and microindentation properties in the elderly. Bone 2016, 93, 196–211. [Google Scholar] [CrossRef] [PubMed]
  6. Przekora, A.; Palka, K.; Ginalska, G. Biomedical potential of chitosan/HA and chitosan/β-1,3-glucan/HA biomaterials as scaffolds for bone regeneration—A comparative study. Mater. Sci. Eng. C 2016, 58. [Google Scholar] [CrossRef] [PubMed]
  7. Alvarez, K.; Nakajima, H. Metallic scaffolds for bone regeneration. Materials 2009, 2, 790–832. [Google Scholar] [CrossRef]
  8. Przekora, A.; Vandrovcova, M.; Travnickova, M.; Pajorova, J.; Molitor, M.; Ginalska, G.; Bacakova, L. Evaluation of the potential of chitosan/β-1,3-glucan/hydroxyapatite material as a scaffold for living bone graft production in vitro by comparison of ADSC and BMDSC behaviour on its surface. Biomed. Mater. 2017, 12. [Google Scholar] [CrossRef]
  9. Yang, C.; Huan, Z.; Wang, X.; Wu, C.; Chang, J. 3D printed Fe scaffolds with HA nanocoating for bone regeneration. ACS Biomater. Sci. Eng. 2018, 4, 608–616. [Google Scholar] [CrossRef]
  10. Dong, Y.; Liang, J.; Cui, Y.; Xu, S.; Zhao, N. Fabrication of novel bioactive hydroxyapatite-chitosan-silica hybrid scaffolds: Combined the sol-gel method with 3D plotting technique. Carbohydr. Polym. 2018, 197, 183–193. [Google Scholar] [CrossRef]
  11. Kim, B.S.; Yang, S.S.; Kim, C.S. Incorporation of BMP-2 nanoparticles on the surface of a 3D-printed hydroxyapatite scaffold using an ε-polycaprolactone polymer emulsion coating method for bone tissue engineering. Colloids Surf. B Biointerfaces 2018, 170, 421–429. [Google Scholar] [CrossRef] [PubMed]
  12. Mohammadi, M.; Alibolandi, M.; Abnous, K.; Salmasi, Z.; Jaafari, M.R.; Ramezani, M. Fabrication of hybrid scaffold based on hydroxyapatite-biodegradable nanofibers incorporated with liposomal formulation of BMP-2 peptide for bone tissue engineering. Nanomed. Nanotechnol. Biol. Med. 2018, 14, 1987–1997. [Google Scholar] [CrossRef] [PubMed]
  13. Przekora, A.; Palka, K.; Ginalska, G. Chitosan/β-1,3-glucan/calcium phosphate ceramics composites-Novel cell scaffolds for bone tissue engineering application. J. Biotechnol. 2014, 182–183. [Google Scholar] [CrossRef] [PubMed]
  14. Xu, Z.; Zhao, R.; Huang, X.; Wang, X.; Tang, S. Fabrication and biocompatibility of agarose acetate nanofibrous membrane by electrospinning. Carbohydr. Polym. 2018, 197, 237–245. [Google Scholar] [CrossRef] [PubMed]
  15. Przekora, A.; Czechowska, J.; Pijocha, D.; Slosarczyk, A.; Ginalska, G. Do novel cement-type biomaterials reveal ion reactivity that affects cell viability in vitro? Cent. Eur. J. Biol. 2014, 9. [Google Scholar] [CrossRef]
  16. An, J.; Liao, H.; Kucko, N.W.; Herber, R.P.; Wolke, J.G.C.; Van Den Beucken, J.J.J.P.; Jansen, J.A.; Leeuwenburgh, S.C.G. Long-term evaluation of the degradation behavior of three apatite-forming calcium phosphate cements. J. Biomed. Mater. Res. Part A 2016, 104, 1072–1081. [Google Scholar] [CrossRef]
  17. Akkineni, A.R.; Luo, Y.; Schumacher, M.; Nies, B.; Lode, A.; Gelinsky, M. 3D plotting of growth factor loaded calcium phosphate cement scaffolds. Acta Biomater. 2015, 27, 264–274. [Google Scholar] [CrossRef]
  18. Xu, H.H.; Wang, P.; Wang, L.; Bao, C.; Chen, Q.; Weir, M.D.; Chow, L.C.; Zhao, L.; Zhou, X.; Reynolds, M.A. Calcium phosphate cements for bone engineering and their biological properties. Bone Res. 2017, 5, 17056. [Google Scholar] [CrossRef] [Green Version]
  19. Frajkorová, F.; Bodišová, K.; Boháč, M.; Bartoníčková, E.; Sedláček, J. Preparation and characterisation of porous composite biomaterials based on silicon nitride and bioglass. Ceram. Int. 2015, 41, 9770–9778. [Google Scholar] [CrossRef]
  20. Zhou, Z.; Ruan, J.; Zou, J.; Zhou, Z.; Shen, X. Bioactivity of bioresorbable composite based on bioactive glass and poly-L-lactide. Trans. Nonferrous Met. Soc. China 2007, 17, 394–399. [Google Scholar] [CrossRef]
  21. Duarte Campos, D.F.; Drescher, W.; Rath, B.; Tingart, M.; Fischer, H. Supporting biomaterials for articular cartilage repair. Cartilage 2012, 3, 205–221. [Google Scholar] [CrossRef] [PubMed]
  22. Vinatier, C.; Guicheux, J. Cartilage tissue engineering: From biomaterials and stem cells to osteoarthritis treatments. Ann. Phys. Rehabil. Med. 2016, 59, 139–144. [Google Scholar] [CrossRef] [PubMed]
  23. DeKosky, B.J.; Dormer, N.H.; Ingavle, G.C.; Roatch, C.H.; Lomakin, J.; Detamore, M.S.; Gehrke, S.H. Hierarchically designed agarose and poly(ethylene glycol) interpenetrating network hydrogels for cartilage tissue engineering. Tissue Eng. Part C Methods 2010, 16, 1533–1542. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, H.; Liu, Y.; Jiang, Z.; Chen, W.; Yu, Y.; Hu, Q. Cell-scaffold interaction within engineered tissue. Exp. Cell Res. 2014, 323, 346–351. [Google Scholar] [CrossRef] [PubMed]
  25. Tan, H.; Chu, C.R.; Payne, K.; Marra, K.G. Injectable in situ forming biodegradable chitosan-hyaluronic acid based hydrogels for cartilage tissue engineering. Biomaterials 2009, 30, 2499–2506. [Google Scholar] [CrossRef] [PubMed]
  26. Przekora, A.; Benko, A.; Blazewicz, M.; Ginalska, G. Hybrid chitosan/β-1,3-glucan matrix of bone scaffold enhances osteoblast adhesion, spreading and proliferation via promotion of serum protein adsorption. Biomed. Mater. 2016, 11. [Google Scholar] [CrossRef] [PubMed]
  27. Ercan, B.; Webster, T. Cell response to nanoscale features and its implications in tissue regeneration: An orthopedic perspective. In Nanotechnology and Tissue Engineering: The Scaffold; Laurencin, C., Nair, L., Eds.; CRC Press: Boca Raton, FL, USA, 2017; pp. 151–155. [Google Scholar]
  28. Chang, H.-I.; Wang, Y. Cell response to surface and architecture of tissue engineering scaffolds. In Regenerative Medicine and Tissue Engineering—Cells and Biomaterials; Eberli, D., Ed.; InTech Open Access Publisher: Rijeka, Croatia, 2011. [Google Scholar] [CrossRef]
  29. Le, X.; Poinern, G.E.J.; Ali, N.; Berry, C.M.; Fawcett, D. Engineering a biocompatible scaffold with either micrometre or nanometre scale surface topography for promoting protein adsorption and cellular response. Int. J. Biomater. 2013, 2013, 782549. [Google Scholar] [CrossRef]
  30. Chen, S.; Lee, C.Y.; Li, R.W.; Smith, P.N.; Qin, Q.H. Modelling osteoblast adhesion on surface-engineered biomaterials: Optimisation of nanophase grain size. Comput. Methods Biomech. Biomed. Eng. 2017, 20, 905–914. [Google Scholar] [CrossRef]
  31. Carina, V.; Costa, V.; Raimondi, L.; Pagani, S.; Sartori, M.; Figallo, E.; Setti, S.; Alessandro, R.; Fini, M.; Giavaresi, G. Effect of low-intensity pulsed ultrasound on osteogenic human mesenchymal stem cells commitment in a new bone scaffold. J. Appl. Biomater. Funct. Mater. 2017, 15, e215–e222. [Google Scholar] [CrossRef]
  32. Zhang, J.; Ding, C.; Ren, L.; Zhou, Y.; Shang, P. The effects of static magnetic fields on bone. Prog. Biophys. Mol. Biol. 2014, 114, 146–152. [Google Scholar] [CrossRef]
  33. Leppik, L.; Zhihua, H.; Mobini, S.; Thottakkattumana Parameswaran, V.; Eischen-Loges, M.; Slavici, A.; Helbing, J.; Pindur, L.; Oliveira, K.M.C.; Bhavsar, M.B.; et al. Combining electrical stimulation and tissue engineering to treat large bone defects in a rat model. Sci. Rep. 2018, 8, 1–14. [Google Scholar] [CrossRef] [PubMed]
  34. Griffith, M.; Islam, M.M.; Edin, J.; Papapavlou, G.; Buznyk, O.; Patra, H.K. The quest for anti-inflammatory and anti-infective biomaterials in clinical translation. Front. Bioeng. Biotechnol. 2016, 4, 1–9. [Google Scholar] [CrossRef] [PubMed]
  35. Ribeiro, M.; Monteiro, F.J.; Ferraz, M.P. Infection of orthopedic implants with emphasis on bacterial adhesion process and techniques used in studying bacterial-material interactions. Biomatter 2012, 2, 176–194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Song, Z.; Borgwardt, L.; Høiby, N.; Wu, H.; Sørensen, T.S.; Borgwardt, A. Prosthesis infections after orthopedic joint replacement: The possible role of bacterial biofilms. Orthop. Rev. 2013, 5, 65–71. [Google Scholar] [CrossRef] [PubMed]
  37. O’Neill, E.; Awale, G.; Daneshmandi, L.; Umerah, O.; Lo, K.W.H. The roles of ions on bone regeneration. Drug Discov. Today 2018, 23, 879–890. [Google Scholar] [CrossRef]
  38. Galli, S.; Naito, Y.; Karlsson, J.; He, W.; Miyamoto, I.; Xue, Y.; Andersson, M.; Mustafa, K.; Wennerberg, A.; Jimbo, R. Local release of magnesium from mesoporous TiO2 coatings stimulates the peri-implant expression of osteogenic markers and improves osteoconductivity in vivo. Acta Biomater. 2014, 10, 5193–5201. [Google Scholar] [CrossRef]
  39. Liu, C.; Fu, X.; Pan, H.; Wan, P.; Wang, L.; Tan, L.; Wang, K.; Zhao, Y.; Yang, K.; Chu, P.K. Biodegradable Mg-Cu alloys with enhanced osteogenesis, angiogenesis, and long-lasting antibacterial effects. Sci. Rep. 2016, 6, 1–17. [Google Scholar] [CrossRef]
  40. Yusa, K.; Yamamoto, O.; Takano, H.; Fukuda, M.; Iino, M. Zinc-modified titanium surface enhances osteoblast differentiation of dental pulp stem cells in vitro. Sci. Rep. 2016, 6, 1–11. [Google Scholar] [CrossRef]
  41. Yusa, K.; Yamamoto, O.; Fukuda, M.; Koyota, S.; Koizumi, Y.; Sugiyama, T. In vitro prominent bone regeneration by release zinc ion from Zn-modified implant. Biochem. Biophys. Res. Commun. 2011, 412, 273–278. [Google Scholar] [CrossRef]
  42. Thian, E.S.; Konishi, T.; Kawanobe, Y.; Lim, P.N.; Choong, C.; Ho, B.; Aizawa, M. Zinc-substituted hydroxyapatite: A biomaterial with enhanced bioactivity and antibacterial properties. J. Mater. Sci. Mater. Med. 2013, 24, 437–445. [Google Scholar] [CrossRef]
  43. Andersen, O.Z.; Offermanns, V.; Sillassen, M.; Almtoft, K.P.; Andersen, I.H.; Sørensen, S.; Jeppesen, C.S.; Kraft, D.C.E.; Bøttiger, J.; Rasse, M.; et al. Accelerated bone ingrowth by local delivery of strontium from surface functionalized titanium implants. Biomaterials 2013, 34, 5883–5890. [Google Scholar] [CrossRef] [PubMed]
  44. Lee, M.; Arikawa, K.; Nagahama, F. Micromolar levels of sodium fluoride promote osteoblast differentiation through Runx2 signaling. Biol. Trace Elem. Res. 2017, 178, 283–291. [Google Scholar] [CrossRef] [PubMed]
  45. Uysal, I.; Severcan, F.; Tezcaner, A.; Evis, Z. Co-doping of hydroxyapatite with zinc and fluoride improves mechanical and biological properties of hydroxyapatite. Prog. Nat. Sci. Mater. Int. 2014, 24, 340–349. [Google Scholar] [CrossRef] [Green Version]
  46. Khorami, M.; Hesaraki, S.; Behnamghader, A.; Nazarian, H.; Shahrabi, S. In vitro bioactivity and biocompatibility of lithium substituted 45S5 bioglass. Mater. Sci. Eng. C 2011, 31, 1584–1592. [Google Scholar] [CrossRef]
  47. Wu, C.; Zhou, Y.; Fan, W.; Han, P.; Chang, J.; Yuen, J.; Zhang, M.; Xiao, Y. Hypoxia-mimicking mesoporous bioactive glass scaffolds with controllable cobalt ion release for bone tissue engineering. Biomaterials 2012, 33, 2076–2085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Wu, C.; Zhou, Y.; Xu, M.; Han, P.; Chen, L.; Chang, J.; Xiao, Y. Copper-containing mesoporous bioactive glass scaffolds with multifunctional properties of angiogenesis capacity, osteostimulation and antibacterial activity. Biomaterials 2013, 34, 422–433. [Google Scholar] [CrossRef]
  49. Okuzu, Y.; Fujibayashi, S.; Yamaguchi, S.; Yamamoto, K.; Shimizu, T.; Sono, T.; Goto, K.; Otsuki, B.; Matsushita, T.; Kokubo, T.; et al. Strontium and magnesium ions released from bioactive titanium metal promote early bone bonding in a rabbit implant model. Acta Biomater. 2017, 63, 383–392. [Google Scholar] [CrossRef] [PubMed]
  50. Tao, Z.S.; Zhou, W.-S.; He, X.W.; Liu, W.; Bai, B.L.; Zhou, Q.; Li, H.; Huang, Z.L.; Tu, K.; Hang, L.; et al. A comparative study of zinc, magnesium, strontium-incorporated hydroxyapatite-coated titanium implants for osseointegration of osteopenic rats. Mater. Sci. Eng. C 2016, 62, 226–232. [Google Scholar] [CrossRef]
  51. Yusa, K.; Yamamoto, O.; Iino, M.; Takano, H.; Fukuda, M.; Qiao, Z.; Sugiyama, T. Eluted zinc ions stimulate osteoblast differentiation and mineralization in human dental pulp stem cells for bone tissue engineering. Arch. Oral Biol. 2016, 71, 162–169. [Google Scholar] [CrossRef]
  52. Qiao, Y.; Zhang, W.; Tian, P.; Meng, F.; Zhu, H.; Jiang, X.; Liu, X.; Chu, P.K. Stimulation of bone growth following zinc incorporation into biomaterials. Biomaterials 2014, 35, 6882–6897. [Google Scholar] [CrossRef]
  53. Offermanns, V.; Andersen, O.Z.; Riede, G.; Sillassen, M.; Jeppesen, C.S.; Almtoft, K.P.; Talasz, H.; Öhman-Mägi, C.; Lethaus, B.; Tolba, R.; et al. Effect of strontium surface-functionalized implants on early and late osseointegration: A histological, spectrometric and tomographic evaluation. Acta Biomater. 2018, 69, 385–394. [Google Scholar] [CrossRef] [PubMed]
  54. Perez, R.A.; Kim, J.H.; Buitrago, J.O.; Wall, I.B.; Kim, H.W. Novel therapeutic core-shell hydrogel scaffolds with sequential delivery of cobalt and bone morphogenetic protein-2 for synergistic bone regeneration. Acta Biomater. 2015, 23, 295–308. [Google Scholar] [CrossRef] [PubMed]
  55. Kargozar, S.; Lotfibakhshaiesh, N.; Ai, J.; Mozafari, M.; Brouki Milan, P.; Hamzehlou, S.; Barati, M.; Baino, F.; Hill, R.G.; Joghataei, M.T. Strontium- and cobalt-substituted bioactive glasses seeded with human umbilical cord perivascular cells to promote bone regeneration via enhanced osteogenic and angiogenic activities. Acta Biomater. 2017, 58, 502–514. [Google Scholar] [CrossRef]
  56. Zhu, H.; Hu, C.; Zhang, F.; Feng, X.; Li, J.; Liu, T.; Chen, J.; Zhang, J. Preparation and antibacterial property of silver-containing mesoporous 58S bioactive glass. Mater. Sci. Eng. C 2014, 42, 22–30. [Google Scholar] [CrossRef] [PubMed]
  57. Chung, R.J.; Hsieh, M.F.; Huang, C.W.; Perng, L.H.; Wen, H.W.; Chin, T.S. Antimicrobial effects and human gingival biocompatibility of hydroxyapatite sol-gel coatings. J. Biomed. Mater. Res. Part B Appl. Biomater. 2006, 76, 169–178. [Google Scholar] [CrossRef] [PubMed]
  58. Radovanović, Ž.; Jokić, B.; Veljović, D.; Dimitrijević, S.; Kojić, V.; Petrović, R.; Janaćković, D. Antimicrobial activity and biocompatibility of Ag+- and Cu2+-doped biphasic hydroxyapatite/α-tricalcium phosphate obtained from hydrothermally synthesized Ag+- and Cu2+-doped hydroxyapatite. Appl. Surf. Sci. 2014, 307, 513–519. [Google Scholar] [CrossRef]
  59. Stanić, V.; Dimitrijević, S.; Antić-Stanković, J.; Mitrić, M.; Jokić, B.; Plećaš, I.B.; Raičević, S. Synthesis, characterization and antimicrobial activity of copper and zinc-doped hydroxyapatite nanopowders. Appl. Surf. Sci. 2010, 256, 6083–6089. [Google Scholar] [CrossRef]
  60. Gritsch, L.; Lovell, C.; Goldmann, W.H.; Boccaccini, A.R. Fabrication and characterization of copper(II)-chitosan complexes as antibiotic-free antibacterial biomaterial. Carbohydr. Polym. 2018, 179, 370–378. [Google Scholar] [CrossRef]
  61. Bari, A.; Bloise, N.; Fiorilli, S.; Novajra, G.; Vallet-Regí, M.; Bruni, G.; Torres-Pardo, A.; González-Calbet, J.M.; Visai, L.; Vitale-Brovarone, C. Copper-containing mesoporous bioactive glass nanoparticles as multifunctional agent for bone regeneration. Acta Biomater. 2017, 55, 493–504. [Google Scholar] [CrossRef]
  62. Lin, Y.; Yang, Z.; Cheng, J. Preparation, characterization and antibacterial property of cerium substituted hydroxyapatite nanoparticles. J. Rare Earths 2007, 25, 452–456. [Google Scholar] [CrossRef]
  63. Moghanian, A.; Firoozi, S.; Tahriri, M. Synthesis and in vitro studies of sol-gel derived lithium substituted 58S bioactive glass. Ceram. Int. 2017, 43, 12835–12843. [Google Scholar] [CrossRef]
  64. Rau, J.V.; Fosca, M.; Graziani, V.; Egorov, A.A.; Zobkov, Y.V.; Fedotov, A.Y.; Ortenzi, M.; Caminiti, R.; Baranchikov, A.E.; Komlev, V.S. Silver-doped calcium phosphate bone cements with antibacterial properties. J. Funct. Biomater. 2016, 7, 10. [Google Scholar] [CrossRef] [PubMed]
  65. Siek, D.; Ślósarczyk, A.; Przekora, A.; Belcarz, A.; Zima, A.; Ginalska, G.; Czechowska, J. Evaluation of antibacterial activity and cytocompatibility of α-TCP based bone cements with silver-doped hydroxyapatite and CaCO3. Ceram. Int. 2017, 43. [Google Scholar] [CrossRef]
  66. Lim, P.N.; Teo, E.Y.; Ho, B.; Tay, B.Y.; Thian, E.S. Effect of silver content on the antibacterial and bioactive properties of silver-substituted hydroxyapatite. J. Biomed. Mater. Res. Part A 2013, 101 A, 2456–2464. [Google Scholar] [CrossRef]
  67. Pelletier, D.A.; Suresh, A.K.; Holton, G.A.; McKeown, C.K.; Wang, W.; Gu, B.; Mortensen, N.P.; Allison, D.P.; Joy, D.C.; Allison, M.R.; et al. Effects of engineered cerium oxide nanoparticles on bacterial growth and viability. Appl. Environ. Microbiol. 2010, 76, 7981–7989. [Google Scholar] [CrossRef] [PubMed]
  68. Cheng, Q.; Lee, B.L.-P.; Komvopoulos, K.; Yan, Z.; Li, S. Plasma surface chemical treatment of electrospun poly(l-Lactide) microfibrous scaffolds for enhanced cell adhesion, growth, and infiltration. Tissue Eng. Part A 2013, 19, 1188–1198. [Google Scholar] [CrossRef] [PubMed]
  69. Chen, T.F.; Siow, K.S.; Ng, P.Y.; Majlis, B.Y. Enhancing the biocompatibility of the polyurethane methacrylate and off-stoichiometry thiol-ene polymers by argon and nitrogen plasma treatment. Mater. Sci. Eng. C 2017, 79, 613–621. [Google Scholar] [CrossRef]
  70. Griffin, M.; Palgrave, R.; Baldovino-, V.G.; Butler, P.E.; Kalaskar, D.M. Argon plasma improves the tissue integration and angiogenesis of subcutaneous implants by modifying surface chemistry and topography. Int. J. Nanomed. 2018, 13, 6123–6141. [Google Scholar] [CrossRef]
  71. Sagbas, B. Argon/oxygen plasma surface modification of biopolymers for improvement of wettability and wear resistance. Int. J. Chem. Mol. Nucl. Mater. Metall. Eng. 2016, 10, 889–894. [Google Scholar]
  72. Kostov, K.G.; Nishime, T.M.C.; Castro, A.H.R.; Toth, A.; Hein, L.R.O. Surface modification of polymeric materials by cold atmospheric plasma jet. Appl. Surf. Sci. 2014, 314, 367–375. [Google Scholar] [CrossRef] [Green Version]
  73. Jordá-Vilaplana, A.; Fombuena, V.; García-García, D.; Samper, M.D.; Sánchez-Nácher, L. Surface modification of polylactic acid (PLA) by air atmospheric plasma treatment. Eur. Polym. J. 2014, 58, 23–33. [Google Scholar] [CrossRef] [Green Version]
  74. Griffin, M.F.; Ibrahim, A.; Seifalian, A.M.; Butler, P.E.M.; Kalaskar, D.M.; Ferretti, P. Chemical group-dependent plasma polymerisation preferentially directs adipose stem cell differentiation towards osteogenic or chondrogenic lineages. Acta Biomater. 2017, 50, 450–461. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, M.; Favi, P.; Cheng, X.; Golshan, N.H.; Ziemer, K.S.; Keidar, M.; Webster, T.J. Cold atmospheric plasma (CAP) surface nanomodified 3D printed polylactic acid (PLA) scaffolds for bone regeneration. Acta Biomater. 2016, 46, 256–265. [Google Scholar] [CrossRef] [PubMed]
  76. Wu, S.; Liu, X.; Yeung, A.; Yeung, K.W.K.; Kao, R.Y.T.; Wu, G.; Hu, T.; Xu, Z.; Chu, P.K. Plasma-modified biomaterials for self-antimicrobial applications. ACS Appl. Mater. Interfaces 2011, 3, 2851–2860. [Google Scholar] [CrossRef] [PubMed]
  77. Bazaka, K.; Jacob, M.V.; Crawford, R.J.; Ivanova, E.P. Plasma-assisted surface modification of organic biopolymers to prevent bacterial attachment. Acta Biomater. 2011, 7, 2015–2028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Zhang, W.; Luo, Y.; Wang, H.; Jiang, J.; Pu, S.; Chu, P.K. Ag and Ag/N2 plasma modification of polyethylene for the enhancement of antibacterial properties and cell growth/proliferation. Acta Biomater. 2008, 4, 2028–2036. [Google Scholar] [CrossRef] [PubMed]
  79. Zhang, W.; Chu, P.K.; Ji, J.; Zhang, Y.; Ng, S.C.; Yan, Q. Surface antibacterial characteristics of plasma-modified polyethylene. Biopolymers 2006, 83, 62–68. [Google Scholar] [CrossRef]
  80. Zhang, W.; Chu, P.K.; Ji, J.; Zhang, Y.; Liu, X.; Fu, R.K.Y.; Ha, P.C.T.; Yan, Q. Plasma surface modification of poly vinyl chloride for improvement of antibacterial properties. Biomaterials 2006, 27, 44–51. [Google Scholar] [CrossRef]
  81. Wang, H.; Ji, J.; Zhang, W.; Wang, W.; Zhang, Y.; Wu, Z.; Zhang, Y.; Chu, P.K. Rat calvaria osteoblast behavior and antibacterial properties of O2 and N2 plasma-implanted biodegradable poly(butylene succinate). Acta Biomater. 2010, 6, 154–159. [Google Scholar] [CrossRef]
  82. Costa, V.; Carina, V.; Fontana, S.; De Luca, A.; Monteleone, F.; Pagani, S.; Sartori, M.; Setti, S.; Faldini, C.; Alessandro, R.; et al. Osteogenic commitment and differentiation of human mesenchymal stem cells by low-intensity pulsed ultrasound stimulation. J. Cell. Physiol. 2018, 233, 1558–1573. [Google Scholar] [CrossRef]
  83. Uddin, S.M.Z.; Qin, Y.X. Enhancement of osteogenic differentiation and proliferation in human mesenchymal stem cells by a modified low intensity ultrasound stimulation under simulated microgravity. PLoS ONE 2013, 8, e73914. [Google Scholar] [CrossRef] [PubMed]
  84. Zhang, N.; Chow, S.K.H.; Leung, K.S.; Cheung, W.H. Ultrasound as a stimulus for musculoskeletal disorders. J. Orthop. Transl. 2017, 9, 52–59. [Google Scholar] [CrossRef] [PubMed]
  85. Kusuyama, J.; Bandow, K.; Shamoto, M.; Kakimoto, K.; Ohnishi, T.; Matsuguchi, T. Low intensity pulsed ultrasound (LIPUS) influences the multilineage differentiation of mesenchymal stem and progenitor cell lines through ROCK-Cot/Tpl2-MEK-ERK signaling pathway. J. Biol. Chem. 2014, 289, 10330–10344. [Google Scholar] [CrossRef] [PubMed]
  86. Moonga, S.S.; Qin, Y.X. MC3T3 infiltration and proliferation in bovine trabecular scaffold regulated by dynamic flow bioreactor and augmented by low-intensity pulsed ultrasound. J. Orthop. Transl. 2018, 14, 16–22. [Google Scholar] [CrossRef] [PubMed]
  87. Zhou, X.; Castro, N.J.; Zhu, W.; Cui, H.; Aliabouzar, M.; Sarkar, K.; Zhang, L.G. Improved human bone marrow mesenchymal stem cell osteogenesis in 3D bioprinted tissue scaffolds with low intensity pulsed ultrasound stimulation. Sci. Rep. 2016, 6, 32876. [Google Scholar] [CrossRef] [PubMed]
  88. Aliabouzar, M.; Lee, S.J.; Zhou, X.; Zhang, G.L.; Sarkar, K. Effects of scaffold microstructure and low intensity pulsed ultrasound on chondrogenic differentiation of human mesenchymal stem cells. Biotechnol. Bioeng. 2018, 115, 495–506. [Google Scholar] [CrossRef] [PubMed]
  89. Aliabouzar, M.; Zhang, L.G.; Sarkar, K. Lipid coated microbubbles and low intensity pulsed ultrasound enhance chondrogenesis of human mesenchymal stem cells in 3D printed scaffolds. Sci. Rep. 2016, 6, 37728. [Google Scholar] [CrossRef]
  90. Gujjalapudi, M.; Anam, C.; Mamidi, P.; Chiluka, R.; Kumar, A.; Bibinagar, R. Effect of magnetic field on bone healing around endosseous implants – An in-vivo study. J. Clin. Diagn. Res. 2016, 10, ZF01–ZF04. [Google Scholar] [CrossRef]
  91. Xia, Y.; Sun, J.; Zhao, L.; Zhang, F.; Liang, X.-J.; Guo, Y.; Weir, M.D.; Reynolds, M.A.; Gu, N.; Xu, H.H.K. Magnetic field and nano-scaffolds with stem cells to enhance bone regeneration. Biomaterials 2018, 183, 151–170. [Google Scholar] [CrossRef]
  92. Galli, C.; Pedrazzi, G.; Mattioli-Belmonte, M.; Guizzardi, S. The Use of Pulsed Electromagnetic fields to promote bone responses to biomaterials in vitro and in vivo. Int. J. Biomater. 2018, 2018, 8935750. [Google Scholar] [CrossRef]
  93. Samal, S.K.; Dash, M.; Shelyakova, T.; Declercq, H.A.; Uhlarz, M.; Bañobre-López, M.; Dubruel, P.; Cornelissen, M.; Herrmannsdörfer, T.; Rivas, J.; et al. Biomimetic magnetic silk scaffolds. ACS Appl. Mater. Interfaces 2015, 7, 6282–6292. [Google Scholar] [CrossRef] [PubMed]
  94. Cai, Q.; Shi, Y.; Shan, D.; Jia, W.; Duan, S.; Deng, X.; Yang, X. Osteogenic differentiation of MC3T3-E1 cells on poly(l-lactide)/Fe3O4 nanofibers with static magnetic field exposure. Mater. Sci. Eng. C 2015, 55, 166–173. [Google Scholar] [CrossRef] [PubMed]
  95. Yun, H.M.; Ahn, S.J.; Park, K.R.; Kim, M.J.; Kim, J.J.; Jin, G.Z.; Kim, H.W.; Kim, E.C. Magnetic nanocomposite scaffolds combined with static magnetic field in the stimulation of osteoblastic differentiation and bone formation. Biomaterials 2016, 85, 88–98. [Google Scholar] [CrossRef] [PubMed]
  96. Yan, Y.; Zhang, Y.; Zuo, Y.; Zou, Q.; Li, J.; Li, Y. Development of Fe3O4–HA/PU superparamagnetic composite porous scaffolds for bone repair application. Mater. Lett. 2018, 212, 303–306. [Google Scholar] [CrossRef]
  97. Heidari, F.; Razavi, M.; Bahrololoom, M.E.; Yazdimamaghani, M.; Tahriri, M.; Kotturi, H.; Tayebi, L. Evaluation of the mechanical properties, in vitro biodegradability and cytocompatibility of natural chitosan/hydroxyapatite/nano-Fe3O4 composite. Ceram. Int. 2018, 44, 275–281. [Google Scholar] [CrossRef]
  98. D’Amora, U.; Russo, T.; Gloria, A.; Rivieccio, V.; D’Antò, V.; Negri, G.; Ambrosio, L.; De Santis, R. 3D additive-manufactured nanocomposite magnetic scaffolds: Effect of the application mode of a time-dependent magnetic field on hMSCs behavior. Bioact. Mater. 2017, 2, 138–145. [Google Scholar] [CrossRef]
  99. Aliramaji, S.; Zamanian, A.; Mozafari, M. Super-paramagnetic responsive silk fibroin/chitosan/magnetite scaffolds with tunable pore structures for bone tissue engineering applications. Mater. Sci. Eng. C 2017, 70, 736–744. [Google Scholar] [CrossRef]
  100. Anjaneyulu, U.; Vijayalakshmi, U. Preparation and characterization of novel sol-gel derived hydroxyapatite/Fe3O4 composites coatings on Ti-6Al-4V for biomedical applications. Mater. Lett. 2017, 189, 118–121. [Google Scholar] [CrossRef]
  101. Zhang, K.; Wang, S.; Zhou, C.; Cheng, L.; Gao, X.; Xie, X.; Sun, J.; Wang, H.; Weir, M.D.; Reynolds, M.A.; et al. Advanced smart biomaterials and constructs for hard tissue engineering and regeneration. Bone Res. 2018, 6, 31. [Google Scholar] [CrossRef]
  102. Chorsi, M.T.; Curry, E.J.; Chorsi, H.T.; Das, R.; Baroody, J.; Purohit, P.K.; Ilies, H.; Nguyen, T.D. Piezoelectric biomaterials for sensors and actuators. Adv. Mater. 2018, 1802084, 1802084. [Google Scholar] [CrossRef]
  103. Jacob, J.; More, N.; Kalia, K.; Kapusetti, G. Piezoelectric smart biomaterials for bone and cartilage tissue engineering. Inflamm. Regen. 2018, 38, 2. [Google Scholar] [CrossRef] [PubMed]
  104. Tandon, B.; Blaker, J.J.; Cartmell, S.H. Piezoelectric materials as stimulatory biomedical materials and scaffolds for bone repair. Acta Biomater. 2018, 73, 1–20. [Google Scholar] [CrossRef] [PubMed]
  105. Chen, W.; Yu, Z.; Pang, J.; Yu, P.; Tan, G.; Ning, C. Fabrication of biocompatible potassium sodium niobate piezoelectric ceramic as an electroactive implant. Materials 2017, 10, 345. [Google Scholar] [CrossRef] [PubMed]
  106. Lv, X.; Li, J.; Men, T.-L.; Wu, J.; Zhang, X.; Wang, K.; Li, J.-F.; Xiao, D.; Zhu, J. High-performance 0-3 type niobate-based lead-free piezoelectric composite ceramics with ZnO inclusions. ACS Appl. Mater. Interfaces 2018, 10, 30566–30573. [Google Scholar] [CrossRef] [PubMed]
  107. Tang, Y.; Wu, C.; Wu, Z.; Hu, L.; Zhang, W.; Zhao, K. Fabrication and in vitro biological properties of piezoelectric bioceramics for bone regeneration. Sci. Rep. 2017, 7, 43360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Augustine, R.; Dan, P.; Sosnik, A.; Kalarikkal, N.; Tran, N.; Vincent, B.; Thomas, S.; Menu, P.; Rouxel, D. Electrospun poly(vinylidene fluoride-trifluoroethylene)/zinc oxide nanocomposite tissue engineering scaffolds with enhanced cell adhesion and blood vessel formation. Nano Res. 2017, 10, 3358–3376. [Google Scholar] [CrossRef]
  109. Damaraju, S.M.; Shen, Y.; Elele, E.; Khusid, B.; Eshghinejad, A.; Li, J.; Jaffe, M.; Arinzeh, T.L. Three-dimensional piezoelectric fibrous scaffolds selectively promote mesenchymal stem cell differentiation. Biomaterials 2017, 149, 51–62. [Google Scholar] [CrossRef]
  110. Nunes-Pereira, J.; Ribeiro, S.; Ribeiro, C.; Gombek, C.J.; Gama, F.M.; Gomes, A.C.; Patterson, D.A.; Lanceros-Méndez, S. Poly(vinylidene fluoride) and copolymers as porous membranes for tissue engineering applications. Polym. Test. 2015, 44, 234–241. [Google Scholar] [CrossRef] [Green Version]
  111. Zhou, Z.; Li, W.; He, T.; Qian, L.; Tan, G.; Ning, C. Polarization of an electroactive functional film on titanium for inducing osteogenic differentiation. Sci. Rep. 2016, 6, 1–8. [Google Scholar] [CrossRef]
  112. Ribeiro, C.; Moreira, S.; Correia, V.; Sencadas, V.; Rocha, J.G.; Gama, F.M.; Gómez Ribelles, J.L.; Lanceros-Méndez, S. Enhanced proliferation of pre-osteoblastic cells by dynamic piezoelectric stimulation. RSC Adv. 2012, 2, 11504–11509. [Google Scholar] [CrossRef]
  113. Ribeiro, C.; Parssinen, J.; Sencadas, V.V.; Correia, V.V.; Miettinen, S.; Hytonen, V.P.; Lanceros-Mendez, S.; Pärssinen, J.; Sencadas, V.V.; Correia, V.V.; et al. Dynamic piezoelectric stimulation enhances osteogenic differentiation of human adipose stem cells. J. Biomed. Mater. Res. Part A 2015, 103, 2172–2175. [Google Scholar] [CrossRef] [PubMed]
  114. Cartmell, S.H.; Thurstan, S.; Gittings, J.P.; Griffiths, S.; Bowen, C.R.; Turner, I.G. Polarization of porous hydroxyapatite scaffolds: Influence on osteoblast cell proliferation and extracellular matrix production. J. Biomed. Mater. Res. Part A 2014, 102, 1047–1052. [Google Scholar] [CrossRef] [PubMed]
  115. Kumar, D.; Gittings, J.P.; Turner, I.G.; Bowen, C.R.; Bastida-Hidalgo, A.; Cartmell, S.H. Polarization of hydroxyapatite: Influence on osteoblast cell proliferation. Acta Biomater. 2010, 6, 1549–1554. [Google Scholar] [CrossRef] [PubMed]
  116. Dubey, A.K.; Basu, B. Pulsed electrical stimulation and surface charge induced cell growth on multistage spark plasma sintered hydroxyapatite-barium titanate piezobiocomposite. J. Am. Ceram. Soc. 2014, 97, 481–489. [Google Scholar] [CrossRef]
  117. Liu, B.; Chen, L.; Shao, C.; Zhang, F.; Zhou, K.; Cao, J.; Zhang, D. Improved osteoblasts growth on osteomimetic hydroxyapatite/BaTiO3composites with aligned lamellar porous structure. Mater. Sci. Eng. C 2016, 61, 8–14. [Google Scholar] [CrossRef] [PubMed]
  118. Dubey, A.; Kakimoto, K.; Obata, A.; Kasuga, T. Enhanced polarization of hydroxyapatite using the design concept of functionally graded materials with sodium potassium niobate. RSC Adv. 2014, 4, 24601–24611. [Google Scholar] [CrossRef]
  119. Kou, P.M.; Babensee, J.E. Macrophage and dendritic cell phenotypic diversity in the context of biomaterials. J. Biomed. Mater. Res. Part A 2011, 96 A, 239–260. [Google Scholar] [CrossRef]
  120. Pajarinen, J.; Lin, T.; Gibon, E.; Kohno, Y.; Maruyama, M.; Nathan, K.; Lu, L.; Yao, Z.; Goodman, S.B. Mesenchymal stem cell-macrophage crosstalk and bone healing. Biomaterials 2018. [Google Scholar] [CrossRef]
  121. Selders, G.S.; Fetz, A.E.; Radic, M.Z.; Bowlin, G.L. An overview of the role of neutrophils in innate immunity, inflammation and host-biomaterial integration. Regen. Biomater. 2017, 4, 55–68. [Google Scholar] [CrossRef] [Green Version]
  122. El-Jawhari, J.J.; Jones, E.; Giannoudis, P.V. The roles of immune cells in bone healing; what we know, do not know and future perspectives. Injury 2016, 47, 2399–2406. [Google Scholar] [CrossRef] [Green Version]
  123. Crop, M.J.; Korevaar, S.S.; De Kuiper, R.; Ijzermans, J.N.M.; Van Besouw, N.M.; Baan, C.C.; Weimar, W.; Hoogduijn, M.J. Human mesenchymal stem cells are susceptible to lysis by CD8 + T Cells and NK cells. Cell Transplant. 2011, 20, 1547–1559. [Google Scholar] [CrossRef] [PubMed]
  124. Ho-Shui-Ling, A.; Bolander, J.; Rustom, L.E.; Johnson, A.W.; Luyten, F.P.; Picart, C. Bone regeneration strategies: Engineered scaffolds, bioactive molecules and stem cells current stage and future perspectives. Biomaterials 2018, 180, 143–162. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Cytotoxicity evaluation of antibacterial bone cements containing Ag-doped hydroxyapatite (HA): (a) shows fabricated bone cements; (b) shows control healthy MC3T3-E1 preosteoblasts cultured on polystyrene after live/dead staining; (c) and (d) show live/dead staining of MC3T3-E1 cells with reduced viability grown directly on the antibacterial bone cements with low (c) and high (d) concentration of Ag-doped HA [65] (green fluorescence—viable cells stained with calcein-AM, red fluorescence—nuclei of dead cells stained with propidium iodide).
Figure 1. Cytotoxicity evaluation of antibacterial bone cements containing Ag-doped hydroxyapatite (HA): (a) shows fabricated bone cements; (b) shows control healthy MC3T3-E1 preosteoblasts cultured on polystyrene after live/dead staining; (c) and (d) show live/dead staining of MC3T3-E1 cells with reduced viability grown directly on the antibacterial bone cements with low (c) and high (d) concentration of Ag-doped HA [65] (green fluorescence—viable cells stained with calcein-AM, red fluorescence—nuclei of dead cells stained with propidium iodide).
Ijms 20 00435 g001
Figure 2. Confocal laser scanning microscope images of MC3T3-E1 preosteoblasts cultured on the surface of biomaterial without and with ferromagnetic nanoparticles (NPs): (a) shows MC3T3-E1 cells on chitosan-based scaffold without ferromagnetic NPs; (b) shows greater number of better spread MC3T3-E1 preosteoblasts on chitosan-based scaffold with ferromagnetic NPs incorporated within polysaccharide matrix (cell cytoskeleton was stained with AlexaFluor635phalloidin).
Figure 2. Confocal laser scanning microscope images of MC3T3-E1 preosteoblasts cultured on the surface of biomaterial without and with ferromagnetic nanoparticles (NPs): (a) shows MC3T3-E1 cells on chitosan-based scaffold without ferromagnetic NPs; (b) shows greater number of better spread MC3T3-E1 preosteoblasts on chitosan-based scaffold with ferromagnetic NPs incorporated within polysaccharide matrix (cell cytoskeleton was stained with AlexaFluor635phalloidin).
Ijms 20 00435 g002
Table 1. Examples of osteopromotive biomaterials produced by modification with metal ions.
Table 1. Examples of osteopromotive biomaterials produced by modification with metal ions.
Metal IonType of BiomaterialExperimental ModelBiomaterial Effect on Bone RegenerationRef.
Mg2+Ti threaded screws with Mg-incorporated mesoporous TiO2 coatingIn vivo rabbit modelImproved biomaterial osteoconductivity, enhanced expression of genes related to bone regeneration process[38]
Mg2+ Sr2+Pure Ti samples with Mg or Sr ions deposited on the surfaceIn vivo rabbit model; In vitro model: MC3T3-E1 cell lineEnhanced proliferation and osteogenic differentiation in vitro, improved biomechanical strength and osseointegration in vivo[49]
Mg2+ Zn2+ Sr2+Ti implants with Mg- or Zn- or Sr-doped HA coatingIn vivo rat modelIncreased new bone formation, improved implant osseointegration[50]
Zn2+Zn-modified Ti spongeIn vitro model: human DPSCsIncreased osteogenic differentiation and mineralization[51]
Zn2+Ti rods/plates with Zn-incorporated TiO2 coatingIn vivo rat model; In vitro model: rat BMDSCsEnhanced expression of genes related to bone regeneration process, improved bone formation process in vitro and in vivo[52]
Sr2+Ti implant with Ti-Sr-O coatingIn vivo rabbit modelImproved early implant osseointegration[53]
Co2+Alginate/collagen/α-TCP scaffold with Co incorporatedIn vitro model: rat BMDSCsEnhanced angiogenic properties of cells and osteogenic differentiation[54]
Sr2+ Co2+BG co-substituted with Sr and CoIn vivo rabbit model;
In vitro model: human umbilical cord perivascular cells (HUCPVCs), Saos-2 cell line
Increased expression of genes related to osteogenesis and angiogenesis processes in Saos-2 and HUCPVC cells, respectively, improved bone healing process in vivo[55]
Table 2. Examples of antimicrobial biomaterials produced by modification with metal ions.
Table 2. Examples of antimicrobial biomaterials produced by modification with metal ions.
Metal IonType of BiomaterialDemonstrated Antimicrobial Activity (Microbial Strain)Effect on Eukaryotic CellsRef.
Ag+Mesoporous
58S BG containing Ag
Escherichia coli; Staphylococcus aureusBiocompatible (studies on primary rat calvarial osteoblasts)[56]
Ag+ Zn2+Ti-6Al-4V alloy with Ag- or Zn-doped HA coatingStreptococcus mutansReduced cell attachment (studies on human gingival fibroblast cell line—HGF-1)[57]
Ag+ Cu2+Ag- or Cu-doped HA/α-TCP
Ag- or Cu-doped HA
Staphylococcus aureus; Escherichia coli; Pseudomonas aeruginosa; Candida albicansNon-toxic, slightly reduced cell proliferation rate (studies on human lung fibroblast cell line—MRC-5)[58]
Zn2+ Cu2+Zn- or Cu-doped nanoHAStaphylococcus aureus; Escherichia coli;
Candida albicans
Not tested[59]
Cu2+ Mg2+Mg-Cu alloyStaphylococcus aureusBiocompatible (studies on MC3T3-E1 and HUVEC lines)[39]
Cu2+Chitosan biomaterial with Cu incorporatedEscherichia coli; Staphylococcus carnosusNon-toxic at low Cu concentrations, toxic at high Cu concentrations (studies on mouse embryonic fibroblast cell line—MEF)[60]
Cu2+Mesoporous BG containing CuStaphylococcus epidermidis; Staphylococcus aureus; Escherichia coliNot tested[61]
Ce3+Ce-doped nanoHAEscherichia coli; Staphylococcus aureusNot tested[62]
Li+Li-doped 58S BGStaphylococcus aureusBiocompatible (studies on MC3T3-E1 cell line)[63]
Table 3. Examples of piezoelectric biomaterials and their effect on eukaryotic cells.
Table 3. Examples of piezoelectric biomaterials and their effect on eukaryotic cells.
Type of BiomaterialLoading RegimeIn Vitro Cellular ModelEffect on Eukaryotic CellsRef.
Porous PVDF and PVDF-TrFE membranesStaticMC3T3-E1 cell line mouse myoblast cell line (C2C12)Enhanced cell proliferation[110]
PVDF film on Ti substrateStaticBMDSCsEnhanced cell proliferation and osteogenic differentiation[111]
PVDF film with Ti layerDynamic (mechanical stimulation in a bioreactor)MC3T3-E1 cell lineEnhanced cell proliferation[112]
PVDF filmDynamic (mechanical stimulation in a bioreactor)Human ADSCsEnhanced osteogenic differentiation[113]
Porous HA scaffoldStaticMC3T3-E1 cell lineEnhanced proliferation and matrix mineralization[114]
HA discStaticMC3T3-E1 cell lineIncreased cell adhesion, proliferation, and metabolic activity[115]
KNN ceramicsStaticMC3T3-E1 cell lineEnhanced cell proliferation[105]
HA/BaTiO3 compositeDynamic (mechanical stimulation with loading device: 3 Hz and 60 N)OsteoblastsEnhanced osteoblast growth and bone-inducing activity[107]
HA/BaTiO3 compositeDynamic (electrical stimulation)Primary culture of human osteoblastsEnhanced cell proliferation[116]
HA/BaTiO3 compositeStaticMG-63 cell line (osteosarcoma)Improved cell adhesion, proliferation, and osteogenic differentiation[117]
HA biomaterial with KNN layersStaticSaos-2 cell line (osteosarcoma)Enhanced cell proliferation[118]

Share and Cite

MDPI and ACS Style

Przekora, A. Current Trends in Fabrication of Biomaterials for Bone and Cartilage Regeneration: Materials Modifications and Biophysical Stimulations. Int. J. Mol. Sci. 2019, 20, 435. https://doi.org/10.3390/ijms20020435

AMA Style

Przekora A. Current Trends in Fabrication of Biomaterials for Bone and Cartilage Regeneration: Materials Modifications and Biophysical Stimulations. International Journal of Molecular Sciences. 2019; 20(2):435. https://doi.org/10.3390/ijms20020435

Chicago/Turabian Style

Przekora, Agata. 2019. "Current Trends in Fabrication of Biomaterials for Bone and Cartilage Regeneration: Materials Modifications and Biophysical Stimulations" International Journal of Molecular Sciences 20, no. 2: 435. https://doi.org/10.3390/ijms20020435

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