Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2299 word(s) 2299 2021-02-03 09:07:25 |
2 Format correct Meta information modification 2299 2021-02-25 07:00:19 | |
3 Format correct Meta information modification 2299 2021-02-25 07:00:46 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Duta, L. CaP-Based Coatings Fabricated by PLD. Encyclopedia. Available online: https://encyclopedia.pub/entry/7541 (accessed on 27 July 2024).
Duta L. CaP-Based Coatings Fabricated by PLD. Encyclopedia. Available at: https://encyclopedia.pub/entry/7541. Accessed July 27, 2024.
Duta, Liviu. "CaP-Based Coatings Fabricated by PLD" Encyclopedia, https://encyclopedia.pub/entry/7541 (accessed July 27, 2024).
Duta, L. (2021, February 24). CaP-Based Coatings Fabricated by PLD. In Encyclopedia. https://encyclopedia.pub/entry/7541
Duta, Liviu. "CaP-Based Coatings Fabricated by PLD." Encyclopedia. Web. 24 February, 2021.
CaP-Based Coatings Fabricated by PLD
Edit

Pulsed Laser Deposition is an atractive technique used for coating dental and orthopedic implants with various biomaterials, including calcium phosphate-based ones.

calcium phosphate-based coatings synthetic and biological hydroxyapatite in vivo assessment mechanical tests bone implants biomedical applications Pulsed Laser Deposition technique

1. Introduction

Bioceramic coatings are currently used for various biomedical applications [1] to modify the implant surface (by increasing its surface roughness), to promote osseointegration. Nowadays, plasma spraying (PS) is the only industrial technology used for coating dental and orthopedic implants with calcium phosphate (CaP)-based materials. In this respect, in the dedicated literature there are numerous and very interesting research works that point out and critically evaluate the physical-chemical and biological characteristics of the structures fabricated by this technique, along with their clinical performances [2][3]. The CaP-based coatings fabricated by conventional thermal PS technique onto medical implants function as an intermediate layer between the tissues and the metallic implants [4]. Despite its wide commercial availability, this technique still has some drawbacks, such as: (i) the synthesized coatings generally consist of several phases (i.e., β-TCP forms at 1200 °C, and it transforms into TTCP at T > 1400 °C); (ii) at higher synthesis temperatures, the negative influence of the mismatch between the thermal expansion coefficients of hydroxyapatite (HA) and tricalcium phosphate and the ones corresponding to Titanium-based alloys (11–15 × 10−6 cm/(cm·K) vs. 8–10 × 10−6 cm/(cm·K), respectively), limit the obtaining of good CaP coatings onto metallic substrates [5]; (iii) it supplies very thick structures with low adherence to the substrate (because the coatings’ tensile stresses have a greater tendency to initiate cracks and cause film delamination [5]); (iv) surface morphology, phase composition, or uniformity of crystallization [6][7] are difficult to be controlled; and (v) it is a line-of-sight method [8]. Therefore, no coating technique can be considered perfect and all these drawbacks gradually supported research efforts focused on the introduction of various, alternative coating techniques to PS (i.e., radio frequency magnetron sputtering, pulsed laser deposition - PLD, electrochemical deposition, etc.) [9][8][10][11][12]. Among these, PLD technique is worth mentioning due to some important advantages over PS method, such as: (i) a much faster surface deposition process; (ii) a stoichiometric transfer of the material’s composition from the target in the synthesized coating; (iii) a better morphological and compositional uniformity; (iv) a lower porosity; (v) precise thickness control; (vi) effectiveness for coating small implants, with complex shapes; (vii) a decreased tendency of the synthesized structures to crack or delaminate; and, very important, (viii) a high adherence to the metallic substrate [9][13]. In the biomedical field, for the fabrication of CaP-based coatings for bone implant applications, one of the most applied plasma-assisted methods is PLD [10]. This technique is able to fabricate dense and extremely adherent films. In this respect, synthetic HA and biological HA (BioHA) coatings obtained by PLD previously demonstrated high adherence to metallic substrates [14]. Moreover, the composition of these coatings is consistent with the one corresponding to the raw (base) materials, along with an improved crystallinity [14]. One should note that in PLD, after the ablation of the target by laser pulses, a plasma plume is generated. When the species existing in the plume reach the surface of the substrate, they may deposit onto the surface and form a film [15]. The number of the deposited species depends both on their density in the plume and their energy. In the case of low energies, the species may not deposit on the substrate surface even if they arrive at the surface [16]. If the substrate temperature is high, the energy of the species can be compensated. Consequently, the number of deposited species onto the substrate surface will increase, along with the density of the droplets. High substrate temperatures also contribute to the atomic diffusion, which, in turn, can determine the appearance of two phenomena: the first one is the atomic rearrangements and crystallization of the film and the second one, the improvement of the film−substrate bonding state [17].

The laser sources appropriate for the ablation of a wide range of materials use wavelengths in the UV domain due to some important advantages over IR and/or visible laser sources, such as (i) a higher penetration depth of the laser beam in the target material and (ii) a higher energy of the photons that allows for a more efficient vaporization of the target [18]. In this respect, the laser sources used for PLD experiments are either excimer lasers (i.e., ArF [19], KrF [20], or XeCl [21], emitting at wavelengths of 193, 248, or 308 nm, respectively) or solid-state lasers (i.e., Nd:YAG [22], emitting at 266 nm). It should be also emphasized that to increase the amount of evaporated material from the ablated target to the detriment of expulsed liquid or solid phases, lasers emitting in a pulsed regime with pulse durations in either the nanoseconds or picoseconds range are generally used [23]. In these regimes, the absorption process takes place more quickly than in the case of thermal diffusion processes. More insights on the PLD and PS techniques are well described elsewhere [24][15][13].

It is important to mention that post-deposition treatments are generally applied to transform CaP phases with lower Ca/P ratio to crystalline HA. Thus, there are two commercially used post-treatments: sintering [24][25][26] and soaking in alkaline solutions [27][28]. These treatments are generally applied for several hours, in the range of 600–800 °C. Their aim is to transform the water trapped in the film during the synthesis process in OH ions, to stabilize the crystalline structure [29].

There are reports in the dedicated literature on the in vivo testing of CaP-based coatings (especially HA) fabricated by different physical vapor deposition methods, but, to the best of our knowledge, only few of them addressed the PLD technique. Therefore, the in vivo results pertaining to the studies reported in the last two decades on either synthetic or biological-derived CaP coatings only, synthesized by PLD technique, were gathered, compiled and thoroughly discussed.

2. Pulsed Laser Deposition Technique for Bone Implant Applications

Because the biomaterials’ osteoinduction mechanism is not yet entirely understood, one could not precisely answer the question whether if the sole biomaterial or an interaction between the biomaterial and the relevant proteins present in the living system are responsible for the osteoinduction process. Because most of the implants do not possess the capability to induce bone growth, specific material properties are required to activate the osteoinduction process. To begin the differentiation of the undifferentiated inducible osteoprogenitor cells into bone-forming cells, it was suggested that both the chemistry and the geometry of the biomaterial in contact with these cells represent critical factors to be considered [30].

Metallic implants (including Ti) are generally used for various biomedical applications, mainly due to their resistance to corrosion and favorable mechanical characteristics [31]. Because of its bioinert nature, bulk Ti is not capable to form a biochemical bond with the bone, and this biological inactivity often generates a fibrous tissue that surrounds the implanted device [32]. To improve both osseointegration rates and longevity of Ti implants, the deposition of CaP-based coatings onto their surfaces is envisaged. It was therefore demonstrated that implants’ surface functionalization with CaP-based coatings could promote the formation of real bonds with the surrounding bone, due to their proved chemical similarity with natural bone tissue and their high biocompatibility [33]. This process occurs rapidly along the entire surface of the coating, in comparison to the case of simple Ti implants (used as controls in the experiments) [5].

A nowadays growing research interest in the field of biomaterials is related to the use of biological-derived CaP materials as viable, safe, and low-cost alternatives to synthetic CaP-ones [24]. It should be emphasized that, unfortunately, the Earth’s available mineral resources are threatened to become limited in the near future because of the rapid demographic increase and economic growth. The access to sustainable resources is therefore critical. Consequently, this will generate a beneficial economic and environmental impact over the society, allowing for an intelligent use of these renewable resources.

The mechanical properties of CaP-based coatings are responsible with the overall success rate of an implant [34]. Thus, the optimal functioning of an endosseous implant is directly influenced by the biomaterial’s mechanical stability, which can be easily evaluated by extraction tests. To obtain information on the force that occurs between the bone tissue and implanted materials, various experimental study models have been developed, each of them with their own particularities [35][36][37]. In this respect, the investigation of the coating bond strength is typically performed by scratch [38][39][40], pull-off [41], tensile adhesion [42][43][44], or shear strength tests [45], respectively. It is important to emphasize upon that the ISO 13779-2:2008 standard requirement for tensile adhesion strength of CaP-based coatings, used for load-bearing applications, is of 15 MPa [46]. It was reported that, in general, CaP-based coatings synthesized by the PLD technique easily surpass this imposed value [24][25][47]. There are some studies in the literature concentrated on tensile strength measurements [48][49], which, very importantly, can provide a direct measurement of the attachment between the bone and the implant surface, being therefore influenced only by the chemical bonding between those two [50][51][52]. This way, the effects of friction and of mechanical forces introduced by surface roughness can be minimized [53]. In the case of animal trials, the implant’s increased bone retention is considered a clinically relevant indicator of improved stability and capacity of the implants to carry loads without detaching. Unfortunately, this type of information cannot be acquired by histological or SEM investigations, which provide only limited information on the functional performance of an implant. One should note that in none of the studies included in this review, related to mechanical testing of CaP-based coatings, alteration or disruption of the implants were present. In general, the inferred values for all functionalized Ti implants demonstrated significantly improved bone attachment in comparison to Ti ones. Next to the PLD surface functionalization of metallic implants, a longer implantation time period was demonstrated to induce a positive influence on the overall bone bonding strength characteristics of the investigated medical devices. One should note that the fabrication by PLD of novel BioHA implant coatings derived from sustainable and inexpensive CaP-based resources, with improved mechanical properties, correlated with an increased bone fixation in vivo, could stand for a pioneering contribution to the progress of advanced medical devices.

In general, the results of the studies included in this review, obtained using standardized radiography and microradiography [54], computed tomography [55][56][57], histomorphometry [58][59][60][61], and tomodensitometry [71], have confirmed the osseointegration process (pointing to an increase in the osseous density), along with a strong bone–implant connection and no inflammatory process of the soft tissues. The histological investigations, performed with various microscopy techniques [6][17][54][62][58][60][63][55][56][64][57], have indicated also new active bone formation and demonstrated no adverse inflammatory reactions or gaps around the sites of implantation or at the bone–implant interface, for any of the investigated structures. In addition, the bone and the implant were shown to tightly adhere to each other along the full length of the interface. One interesting observation was that the temperature applied during the deposition process seemed to play an important role in the bone growth of the synthesized structures, the osteoblasts in the newly formed bone tissue being shown clear and homogenous [17]. Moreover, the osteointegration rate was demonstrated to be slightly superior in the case of annealed samples, rather than for the non-annealed ones [62]. It seemed also that no matter what the laser source used (i.e., KrF or CO2), between the synthesized coatings, no significant statistical differences in the osteogenesis process were inferred [62]. Significantly higher values of the bone area between the implant surface and the boundaries and bone adherence ratios were inferred at various implantation time periods in the case of functionalized implants in comparison to control ones [56][57]. In this respect, in the case of control samples, a fibrous connective tissue between the metallic implants and the newly formed bone was shown, while in the case of the synthesized coatings, this layer could be observed only seldom [54].

Even though it is generally accepted that CaP-based coatings deposited by PLD improve bone strength and the initial osseointegration rate, the coatings’ properties necessary to achieve an optimum bone response are yet to be determined. This is mainly because of the limited number of in vivo studies available in the dedicated literature. It should be emphasized upon that the in vivo testing should demonstrate stability in biological environment for up to 1 month, which corresponds to the initial healing phase of the implants [65]. The limitations on such experiments can be related to (i) the difficulty to select a suitable animal model in order to properly simulate the actual mechanical loading and unloading conditions in which an implant should function inside a living system; (ii) the need to sacrifice a large number of animals to reach a significant statistical relevance, able to validate the obtained results; (iii) the demands for high costs and long time frame in the case of clinical trials; (iv) the lack of coordination among material scientists and biologists and thus an insufficient understanding of this interdisciplinary subject; and (v) the serious ethical concerns related to the used animals (including also the choice of their correct number), as they might be sometimes subjected to painful procedures or toxic exposures during the experimental trials [66].

Taking into consideration all these aspects, a future important progress of CaP-based materials might be linked to a shift of the focus from osteoconduction to osteoinduction, e.g., by additive manufacturing of scaffolds with complex, controlled three-dimensional porous structures and development of novel ion-substituted CaPs with increased biological activity. Moreover, new strategies, possibly based on self-assembling and/or nanofabrication might be developed for the successful fabrication of load-bearing bone graft substitutes. In the future, the composition, microstructure, and molecular surface chemistry of various types of CaPs might be tailored in such a way to match the specific biological and metabolic requirements of tissues or disease states. The multilayer composite coating systems, fabricated by the PLD technique, should represent also a future trend, able to provide multifunctional properties for the biomedical implants.

References

  1. Ballini, A.; Mastrangelo, F.; Gastaldi, G.; Tettamanti, L.; Bukvic, N.; Cantore, S.; Cocco, T.; Saini, R.; Desiate, A.; Gherlone, E.; et al. Osteogenic differentiation and gene expression of dental pulp stem cells under low-level laser irradiation: A good promise for tissue engineering. J. Biol. Regul. Homeost. Agents 2015, 29, 813–822.
  2. Cheang, P.; Khor, K.A. Addressing processing problems associated with plasma spraying of hydroxyapatite coatings. Biomaterials 1996, 17, 537–544.
  3. Epinette, J.A.; Manley, M.T. Fifteen Years of Clinical Experience with Hydroxyapatite Coatings in Joint Arthroplasty, 1st ed.; Springer: Paris, France, 2004; p. 452.
  4. Bose, S.; Tarafder, S.; Bandyopadhyay, A. Hydroxyapatite coatings for metallic implants. In Hydroxyapatite (Hap) for Biomedical Applications; Mucalo, M., Ed.; Woodhead Publishing Series in Biomaterials; Elsevier: Amsterdam, The Netherlands, 2015; pp. 143–157.
  5. Yingchao, S.; Irsalan, C.; Yufeng, Z.; Liping, T.; Yi-Xian, Q.; Donghui, Z. Biofunctionalization of metallic implants by calcium phosphate coatings. Bioact. Mat. 2019, 4, 196–206.
  6. Hayami, T.; Hontsu, S.; Higuchi, Y.; Nishikawa, H.; Kusunoki, M. Osteoconduction of a stoichiometric and bovine hydroxyapatite bilayer-coated implant. Clin. Oral Implants Res. 2011, 22, 774–776.
  7. Yoshinari, M.; Hayakawa, T.; Wolke, J.G.; Nemoto, K.; Jansen, J.A. Influence of rapid heating with infrared radiation on RF magnetron-sputtered calcium phosphate coatings. J. Biomed. Mater. Res. 1997, 37, 60–67.
  8. Eliaz, N.; Metoki, N. Calcium Phosphate Bioceramics: A Review of Their History, Structure, Properties, Coating Technologies and Biomedical Applications. Materials 2017, 10, 334.
  9. Surmenev, R.A.; Surmeneva, M.A.; Ivanova, A.A. Significance of calcium phosphate coatings for the enhancement of new bone osteogenesis—A review. Acta Biomater. 2014, 2, 557–579.
  10. Graziani, G.; Boi, M.; Bianchi, M. A Review on Ionic Substitutions in Hydroxyapatite Thin Films: Towards Complete Biomimetism. Coatings 2018, 8, 269.
  11. Tite, T.; Popa, A.C.; Balescu, L.M.; Bogdan, I.M.; Pasuk, I.; Ferreira, J.M.F.; Stan, G.E. Cationic substitutions in hydroxyapatite: Current status of the derived biofunctional effects and their in vitro interrogation methods. Materials 2018, 11, 2081.
  12. Surmenev, R.A.; Surmeneva, M.A. A critical review of decades of research on calcium phosphate–based coatings: How far are we from their widespread clinical application? Curr. Opin. Biomed. Eng. 2019, 10, 35–44.
  13. Surmenev, R.A. A review of plasma-assisted methods for calcium phosphate-based coatings fabrication. Surf. Coat. Technol. 2012, 206, 2035–2056.
  14. Duta, L.; Oktar, F.N.; Stan, G.E.; Popescu-Pelin, G.; Serban, N.; Luculescu, C.; Mihailescu, I.N. Novel doped hydroxyapatite thin films obtained by pulsed laser deposition. Appl. Surf. Sci. 2013, 265, 41–49.
  15. Eason, R. Pulsed Laser Deposition of Thin Films-Applications-Led Growth of Functional Materials; Wiley-Interscience: Hoboken, NJ, USA, 2006; pp. 1–682.
  16. Zeng, H. Evaluation of Bioceramic Coatings Produced by Pulsed Laser Deposition and Ion Beam Sputtering; University of Alabama at Birmingham: Birmingham, AL, USA, 1997.
  17. Wang, D.G.; Chena, C.Z.; Ma, Q.S.; Jin, Q.P.; Li, H.C. A study on in vitro and in vivo bioactivity of HA/45S5 composite films by pulsed laser deposition. Appl. Surf. Sci. 2013, 270, 667–674.
  18. Schneider, C.W.; Lippert, T. Laser Ablation and Thin Film Deposition. In Laser Processing of Materials; Schaaf, P., Ed.; Springer Series in Materials Science; Springer: Berlin/Heidelberg, Germany, 2010; Volume 139, pp. 89–112.
  19. Caricato, A.P.; Martino, M.; Romano, F.; Mirchin, N.; Peled, A. Pulsed laser photodeposition of a-Se nanofilms by ArF laser. Appl. Surf. Sci. 2007, 253, 6517–6521.
  20. Hashimoto, Y.; Ueda, M.; Kohiga, Y.; Imura, K.; Hontsu, S. Application of fluoridated hydroxyapatite thin film coatings using KrF pulsed laser deposition. Dent. Mater. J. 2018, 37, 408–413.
  21. Zaki, A.M.; Blythe, H.J.; Heald, S.M.; Fox, A.M.; Gehring, G.A. Growth of high quality yttrium iron garnet films using standard pulsed laser deposition technique. J. Magn. Magn. Mater. 2018, 453, 254–257.
  22. Novotný, M.; Vondráček, M.; Marešová, E.; Fitl, P.; Bulíř, J.; Pokornýa, P.; Havlová, Š.; Abdellaoui, N.; Pereira, A.; Hubík, P.; et al. Optical and structural properties of ZnO:Eu thin films grown by pulsed laser deposition. Appl. Surf. Sci. 2019, 476, 271–275.
  23. Gyorgy, E.; Ristoscu, C.; Mihailescu, I.N. Role of laser pulse duration and gas pressure in deposition of AlN thin films. J. Appl. Phys. 2001, 90, 456.
  24. Duta, L.; Popescu, A.C. Current Status on Pulsed Laser Deposition of Coatings from Animal-Origin Calcium Phosphate Sources. Coatings 2019, 9, 335.
  25. Zhang, Z.; Dunn, M.F.; Xiao, T.; Tomsia, A.P.; Saiz, E. Nanostructured Hydroxyapatite Coatings for Improved Adhesion and Corrosion Resistance for Medical Implants. Mat. Res. Soc. Symp. Proc. 2002, 703, 291–296.
  26. Jain, P.; Mandal, T.; Prakash, P.; Garg, A.; Balani, K. Electrophoretic deposition of nanocrystalline hydroxyapatite on Ti6Al4V/TiO2 substrate. J. Coat. Technol. Res. 2013, 10, 263–275.
  27. Wen, C.; Guan, S.; Peng, L.; Ren, C.; Wang, X.; Hu, Z. Characterization and degradation behavior of AZ31 alloy surface modified by bone-like hydroxyapatite for implant applications. Appl. Surf. Sci. 2009, 255, 6433–6438.
  28. Bakhsheshi-Rad, H.; Idris, M.; Abdul-Kadir, M. Synthesis and in vitro degradation evaluation of the nano-HA/MgF2 and DCPD/MgF2 composite coating on biodegradable Mg–Ca–Zn alloy. Surf. Coat. Technol. 2013, 222, 79–89.
  29. León, B.; John, J. Thin Calcium Phosphate Coatings for Medical Implants; Springer: New York, NY, USA, 2009; pp. 1–328.
  30. Fujibayashi, S.; Neo, M.; Kim, H.M.; Kokubo, T.; Nakamura, T. Osteoinduction of porous bioactive titanium metal. Biomaterials 2004, 25, 443–450.
  31. Yeo, I.-S.L. Modifications of dental implant surfaces at the microand nano-level for enhanced osseointegration. Materials 2020, 13, 89.
  32. Katta, P.P.K.; Nalliyan, R. Corrosion resistance with self-healing behavior and biocompatibility of Ce incorporated niobium oxide coated 316L SS for orthopedic applications. Surf. Coat. Technol. 2019, 375, 715–726.
  33. Dorozhkin, S.V. Multiphasic calcium orthophosphate (CaPO4) bioceramics and their biomedical applications. Ceram. Int. 2016, 42, 6529–6554.
  34. Pichugin, V.F.; Surmenev, R.A.; Shesterikov, E.V.; Ryabtseva, M.A.; Eshenko, E.V.; Tverdokhlebov, S.I.; Prymak, O.; Epple, M. The preparation of calcium phosphate coatings on titanium and nickel-titanium by rf-magnetron-sputtered deposition: Composition, structure and micromechanical properties. Surf. Coat. Technol. 2008, 202, 3913–3920.
  35. Pearce, A.I.; Richards, R.G.; Milz, S.; Schneider, E.; Pearce, S.G. Animal models for implant biomaterial research in bone: A review. Eur. Cells Mater. 2007, 13, 1–10.
  36. Salou, L.; Hoornaert, A.; Louarn, G.; Layrolle, P. Enhanced osseointegration of titanium implants with nanostructured surfaces: An experimental study in rabbits. Acta Biomater. 2015, 11, 494–502.
  37. Sul, Y.-T.; Johansson, C.; Albrektsson, T. A novel in vivo method for quantifying the interfacial biochemical bond strength of bone implants. J. R. Soc. Interface 2009, 7, 81–90.
  38. Dinda, G.P.; Shin, J.; Mazumder, J. Pulsed laser deposition of hydroxyapatite thin films on Ti-6Al-4V: Effect of heat treatment on structure and properties. Acta Biomater. 2009, 5, 1821–1830.
  39. Łatka, L.; Pawlowski, L.; Chicot, D.; Pierlot, C.; Petit, F. Mechanical properties of suspension plasma sprayed hydroxyapatite coatings submitted to simulated body fluid. Surf. Coat. Technol. 2010, 205, 954–960.
  40. ISO 20502:2005(E). Fine Ceramics (Advanced Ceramics, Advanced Technical Ceramics)—Determination of Adhesion of Ceramic Coatings by Scratch Testing; ISO: Geneva, Switzerland, 2005; Available online: www.iso.org (accessed on 11 December 2020).
  41. Gadow, R.; Killinger, A.; Stiegler, N. Hydroxyapatite coatings for biomedical applications deposited by different thermal spray techniques. Surf. Coat. Technol. 2010, 205, 1157–1164.
  42. Yi, J.; Song, L.; Liu, X.; Xiao, Y.; Wu, Y.; Chen, J.; Wu, F.; Gu, Z. Hydroxyapatite Coatings Deposited by Liquid Precursor Plasma Spraying: Controlled Dense and Porous Microstructures and Osteoblastic Cell Responses. Biofabrication 2010, 2, 045003.
  43. Dey, A.; Mukhopadhyay, A.K.; Gangadharan, S.; Sinha, M.K.; Basu, D.; Bandyopadhyay, N.R. Nanoindentation study of microplasma sprayed hydroxyapatite coating. Ceram. Int. 2009, 35, 2295–2304.
  44. Singh, G.; Singh, S.; Prakash, S. Surface characterization of plasma sprayed pure and reinforced hydroxyapatite coating on Ti6Al4V alloy. Surf. Coat. Technol. 2011, 205, 4814–4820.
  45. Gomes, P.S.; Botelho, C.; Lopes, M.A.; Santos, J.D.; Fernandes, M.H. Evaluation of human osteoblastic cell response to plasma-sprayed silicon-substituted hydroxyapatite coatings over titanium substrates. J. Biomed. Mater. Res. Appl. Biomater. 2010, 94B, 337–346.
  46. ISO 13779-2:2008. Implants for Surgery—Hydroxyapatite—Part 2: Coatings of Hydroxyapatite. Available online: http://www.iso.org/ (accessed on 11 December 2020).
  47. Duta, L.; Mihailescu, N.; Popescu, A.C.; Luculescu, C.R.; Mihailescu, I.N.; Çetin, G.; Gunduz, O.; Oktar, F.N.; Popa, A.C.; Kuncser, A.; et al. Comparative physical, chemical and biological assessment of simple and titanium-doped ovine dentine-derived hydroxyapatite coatings fabricated by pulsed laser deposition. Appl. Surf. Sci. 2017, 413, 129–139.
  48. Edwards, J.T.; Brunski, J.B.; Higuchi, H.W. Mechanical and morphologic investigation of the tensile strength of a bone hydroxyapatite interface. J. Biomed. Mater. Res. 1997, 36, 454–468.
  49. Skripitz, R.; Aspenberg, P. Tensile bond between bone and titanium: A reappraisal of osseointegration. Acta Orthop. Scand. 1998, 69, 315–319.
  50. Shannon, F.J.; Cottrell, J.M.; Deng, X.-H.; Crowder, K.N.; Doty, S.B.; Avaltroni, M.J.; Warren, R.F.; Wright, T.M.; Schwartz, J. A novel surface treatment for porous metallic implants that improves the rate of bony ongrowth. J. Biomed. Mater. Res. A 2008, 86, 857–864.
  51. Schumacher, T.C.; Tushtev, K.; Wagner, U.; Becker, C.; Große Holthaus, M.; Hein, S.B.; Haack, J.; Engelhardt, C.H.M.; Khassawna, T.E.; Rezwan, K. A novel, hydroxyapatite-based screw-like device for anterior cruciate ligament (ACL) reconstructions. Knee 2017, 24, 933–939.
  52. Aparicioa, C.; Padrósb, A.; Gil, F.-J. In vivo evaluation of micro-rough and bioactive titanium dental implants using histometry and pull-out tests. J. Mech. Behav. Biomed. 2011, 4, 1672–1682.
  53. Nakamura, T.; Yamamuro, T.; Higashi, S.; Kokubo, T.; Itoo, S. A new glass-ceramic for bone replacement: Evaluation of its bonding to bone tissue. J. Biomed. Mater. Res. 1985, 19, 685–698.
  54. Dostálová, T.; Jelínek, M.; Himmlová, L.; Grivas, C. Laser–Deposited Hydroxyapatite Films on Dental Implants—Biological Evaluation in vivo. Laser Phys. 1998, 8, 182–186.
  55. Mroz, W.; Budner, B.; Syroka, R.; Niedzielski, K.; Golanski, G.; Slosarczyk, A.; Schwarze, D.; Douglas, T.E.L. In vivo implantation of porous titanium alloy implants coated with magnesium-doped octacalcium phosphate and hydroxyapatite thin films using pulsed laser deposition. J. Biomed. Mater. Res. B Appl. Biomater. 2015, 103, 151–158.
  56. Chen, L.; Komasa, S.; Hashimoto, Y.; Hontsu, S.; Okazaki, J. In Vitro and In Vivo Osteogenic Activity of Titanium Implants Coated by Pulsed Laser Deposition with a Thin Film of Fluoridated Hydroxyapatite. Int. J. Mol. Sci. 2018, 19, 1127.
  57. Duta, L.; Neamtu, J.; Melinte, R.P.; Zureigat, O.A.; Popescu-Pelin, G.; Chioibasu, D.; Oktar, F.N.; Popescu, A.C. In Vivo Assessment of Bone Enhancement in the Case of 3D-Printed Implants Functionalized with Lithium-Doped Biological-Derived Hydroxyapatite Coatings: A Preliminary Study on Rabbits. Coatings 2020, 10, 992.
  58. Dostálová, T.; Himmlová, L.; Jelínek, M.; Grivas, C. Osseointegration of loaded dental implant with KrF laser hydroxylapatite films on Ti6Al4V alloy by minipigs. J. Biomed. Opt. 2001, 6, 239–243.
  59. Kim, H.; Vohra, Y.K.; Louis, P.J.; Lacefield, W.R.; Lemons, J.E.; Camata, R.P. Biphasic and Preferentially Oriented Microcrystalline Calcium Phosphate Coatings: In-vitro and In-vivo Studies. Key Eng. Mat. 2005, 284–286, 207–210.
  60. Peraire, C.; Arias, J.L.; Bernal, D.; Pou, J.; Leon, B.; Arano, A.; Roth, W. Biological stability and osteoconductivity in rabbit tibia of pulsed laser deposited hydroxylapatite coatings. J. Biomed. Mater. Res. Part A 2006, 77, 370–379.
  61. Paz, M.D.; Chiussi, S.; González, P.; Serra, J.; León, B.; Alava, J.I.; Güemes, I.; Sánchez-Margallo, F.M. Osseointegration of Calcium Phosphate Nanofilms on Titanium Alloy Implants. Key Eng. Mater. 2008, 361–363, 645–648.
  62. Antonov, E.N.; Bagratashvili, V.N.; Popov, V.K.; Sobol, E.N.; Howdle, S.M.; Joiner, C.; Parker, K.G.; Parker, T.L.; Doktorov, A.A.; Likhanov, V.B.; et al. Biocompatibility of laser-deposited hydroxyapatite coatings on titanium and polymer implant materials. J. Biomed. Opt. 1998, 3, 423–428.
  63. Hontsu, S.; Hashimoto, Y.; Yoshikawa, Y.; Kusunoki, M.; Nishikawa, H.; Ametani, A. Fabrication of Hydroxyl Apatite Coating Titanium Web Scaffold Using Pulsed Laser Deposition Method. J. Hard. Tissue Biol. 2012, 21, 181–188.
  64. Wang, D.G.; Chen, C.Z.; Yang, X.X.; Ming, X.C.; Zhang, W.L. Effect of bioglass addition on the properties of HA/BG composite films fabricated by pulsed laser deposition. Ceram. Int. 2018, 44, 14528–14533.
  65. Pezeshki, P.; Lugowski, S.; Davies, J.E. Dissolution behavior of calcium phosphate nanocrystals deposited on titanium alloy surfaces. J. Biomed. Mater. Res. 2010, 94A, 660–666.
  66. Paital, S.R.; Dahotre, N.B. Calcium phosphate coatings for bio-implant applications: Materials, performancefactors, and methodologies. Mater. Sci. Eng. 2009, R66, 1–70.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 412
Revisions: 3 times (View History)
Update Date: 25 Feb 2021
1000/1000
Video Production Service