Next Article in Journal
Synthesis and Characterization of Novel Epoxy Geopolymer Hybrid Composites
Next Article in Special Issue
Phosphorus Effects of Mesoporous Bioactive Glass on Occlude Exposed Dentin
Previous Article in Journal
Dopamine-Mediated Sclerotization of Regenerated Chitin in Ionic Liquid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Calcium Orthophosphate-Based Bioceramics

Kudrinskaja sq. 1-155, Moscow 123242, Russia
Materials 2013, 6(9), 3840-3942; https://doi.org/10.3390/ma6093840
Submission received: 20 June 2013 / Revised: 7 August 2013 / Accepted: 19 August 2013 / Published: 6 September 2013
(This article belongs to the Special Issue Ceramics for Healthcare 2013)

Abstract

:
Various types of grafts have been traditionally used to restore damaged bones. In the late 1960s, a strong interest was raised in studying ceramics as potential bone grafts due to their biomechanical properties. A bit later, such synthetic biomaterials were called bioceramics. In principle, bioceramics can be prepared from diverse materials but this review is limited to calcium orthophosphate-based formulations only, which possess the specific advantages due to the chemical similarity to mammalian bones and teeth. During the past 40 years, there have been a number of important achievements in this field. Namely, after the initial development of bioceramics that was just tolerated in the physiological environment, an emphasis was shifted towards the formulations able to form direct chemical bonds with the adjacent bones. Afterwards, by the structural and compositional controls, it became possible to choose whether the calcium orthophosphate-based implants remain biologically stable once incorporated into the skeletal structure or whether they were resorbed over time. At the turn of the millennium, a new concept of regenerative bioceramics was developed and such formulations became an integrated part of the tissue engineering approach. Now calcium orthophosphate scaffolds are designed to induce bone formation and vascularization. These scaffolds are often porous and harbor different biomolecules and/or cells. Therefore, current biomedical applications of calcium orthophosphate bioceramics include bone augmentations, artificial bone grafts, maxillofacial reconstruction, spinal fusion, periodontal disease repairs and bone fillers after tumor surgery. Perspective future applications comprise drug delivery and tissue engineering purposes because calcium orthophosphates appear to be promising carriers of growth factors, bioactive peptides and various types of cells.

1. Introduction

One of the most exciting and rewarding areas of the engineering discipline involves development of various devises for health care. Some of them are implantable. Examples comprise sutures, catheters, heart valves, pacemakers, breast implants, fracture fixation plates, nails and screws in orthopedics, various filling formulations, orthodontic wires, total joint replacement prostheses, etc. However, in order to be accepted by the living body without any unwanted side effects, all implantable items must be prepared from a special class of tolerable materials, called biomedical materials or biomaterials, in short. The physical character of the majority of the available biomaterials is solids [1,2].
Since all types of solids are divided into four major groups of materials: metals, polymers ceramics and various blends thereof, called composites, similarly, all types of solid biomaterials are also divided into the same groups: biometals, biopolymers, bioceramics and biocomposites. All of them play very important roles in both replacement and regeneration of human tissues; however, setting biometals, biopolymers and biocomposites aside, this review is focused on bioceramics only. In general, the modern bioceramics comprise polycrystalline materials, amorphous materials (glasses) and blends thereof (glass-ceramics). However, the chemical elements used to manufacture bioceramics form just a small set of the Periodic Table. Namely, bioceramics might be prepared from alumina, zirconia, magnesia, carbon, silica-contained and calcium-contained compounds, as well as from a limited number of other chemicals. All these formulations might be manufactured in both dense and porous forms in bulk, as well as in the forms of crystals, powders, particles, granules, scaffolds and/or coatings [1,2,3].
As seen from the above, the entire subject of bioceramics is still rather broad. To specify it further, let me limit myself by a description of calcium orthophosphate-based bioceramics only. Due to the chemical similarity to mammalian bones and teeth, this type of bioceramics is used in a number of different applications throughout the body, covering all areas of the skeleton. The examples include healing of bone defects, fracture treatment, total joint replacement, bone augmentation, orthopedics, cranio-maxillofacial reconstruction, spinal surgery, otolaryngology, ophthalmology and percutaneous devices [1,2,3], as well as dental fillings and periodontal treatments [4]. Depending upon the required properties, different calcium orthophosphates might be used. For example, Figure 1 shows some randomly chosen samples of the commercially available calcium orthophosphate bioceramics for bone graft applications. One should note that, in 2010, only in the USA the sales of bone graft substitutes were valued at ~$1.3 billion with a forecast of ~$2.3 billion by 2017 [5]. This clearly demonstrates an importance of calcium orthophosphate-based bioceramics.
A list of the available calcium orthophosphates, including their standard abbreviations and major properties, is available in Table 1 [6]. To narrow the subject further, with a few important exceptions, bioceramics prepared from undoped and un-substituted calcium orthophosphates will be considered and discussed only. Due to this reason, calcium orthophosphate-based bioceramics prepared from biological resources, such as bones, teeth, corals, etc. [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25], as well as the ion-substituted ones [26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57] is not considered. The readers interested in both topics are advised to study the original publications.
Figure 1. Several examples of the commercial calcium orthophosphate-based bioceramics.
Figure 1. Several examples of the commercial calcium orthophosphate-based bioceramics.
Materials 06 03840 g001

2. General Knowledge and Definitions

A number of definitions have been developed for the term “biomaterials”. For example, by the end of the 20th century, the consensus developed by the experts was the following: biomaterials were defined as synthetic or natural materials to be used to replace parts of a living system or to function in intimate contact with living tissues [58]. However, in September 2009, a more advanced definition was introduced: “A biomaterial is a substance that has been engineered to take a form which, alone or as part of a complex system, is used to direct, by control of interactions with components of living systems, the course of any therapeutic or diagnostic procedure, in human or veterinary medicine” [59]. The definition alterations were accompanied by a shift in both the conceptual ideas and the expectations of biological performance, which mutually changed in time [60].
In general, the biomaterials discipline is founded in the knowledge of the synergistic interaction of material science, biology, chemistry, medicine and mechanical science and it requires the input of comprehension from all these areas so that potential implants perform adequately in a living body and interrupt normal body functions as little as possible [61]. As biomaterials deal with all aspects of the material synthesis and processing, the knowledge in chemistry, material science and engineering is essential. On the other hand, as clinical implantology is the main purposes of biomaterials, biomedical sciences become the key part of the research. These include cell and molecular biology, histology, anatomy and physiology. The final aim is to achieve the correct biological interaction of the artificial grafts with living tissues of a host. In order to achieve the goals, several stages have to be performed, such as: material synthesis, design and manufacturing of prostheses, followed by various types of tests. Furthermore, any potential biomaterial must also pass all regulatory requirements before its clinical application [62].
Table 1. Existing calcium orthophosphates and their major properties [6].
Table 1. Existing calcium orthophosphates and their major properties [6].
Ca/P molar ratioCompounds and their typical abbreviationsChemical formulaSolubility at 25 °C, −log KsSolubility at 25 °C, g/LpH stability range in aqueous solutions at 25 °C
0.5Monocalcium phosphate monohydrate (MCPM)Ca(H2PO4)2·H2O1.14~180.0–2.0
0.5Monocalcium phosphate anhydrous (MCPA or MCP)Ca(H2PO4)21.14~17[c]
1.0Dicalcium phosphate dihydrate (DCPD), mineral brushiteCaHPO4·2H2O6.59~0.0882.0–6.0
1.0Dicalcium phosphate anhydrous (DCPA or DCP), mineral monetiteCaHPO46.90~0.048[c]
1.33Octacalcium phosphate (OCP)Ca8(HPO4)2(PO4)4·5H2O96.6~0.00815.5–7.0
1.5α-Tricalcium phosphate (α-TCP)α-Ca3(PO4)225.5~0.0025[a]
1.5β-Tricalcium phosphate (β-TCP)β-Ca3(PO4)228.9~0.0005[a]
1.2–2.2Amorphous calcium phosphates (ACP)CaxHy(PO4)z·nH2O, n = 3 − 4.5; 15%–20% H2O[b][b]~5–12 [d]
1.5–1.67Calcium-deficient hydroxyapatite (CDHA or Ca-def HA) [e]Ca10−x(HPO4)x(PO4)6−x(OH)2−x (0 < x < 1)~85~0.00946.5–9.5
1.67Hydroxyapatite (HA, HAp or OHAp)Ca10(PO4)6(OH)2116.8~0.00039.5–12
1.67Fluorapatite (FA or FAp)Ca10(PO4)6F2120.0~0.00027–12
1.67Oxyapatite (OA, OAp or OXA) [f]Ca10(PO4)6O~69~0.087[a]
2.0Tetracalcium phosphate (TTCP or TetCP), mineral hilgenstockiteCa4(PO4)2O38–44~0.0007[a]
[a] These compounds cannot be precipitated from aqueous solutions; [b] Cannot be measured precisely. However, the following values were found: 25.7 ± 0.1 (pH = 7.40), 29.9 ± 0.1 (pH = 6.00), 32.7 ± 0.1 (pH = 5.28). The comparative extent of dissolution in acidic buffer is: ACP >> α-TCP >> β-TCP > CDHA >> HA > FA; [c] Stable at temperatures above 100 °C; [d] Always metastable; [e] Occasionally, it is called “precipitated HA (PHA)”; and [f] Existence of OA remains questionable.
In any case, biomaterials are intended to interface with biological systems in vivo to evaluate, treat, augment or replace any tissue, organ or function of the body and are now used in a number of different applications throughout the body. Thus, biomaterials are solely associated with the health care domain and must have an interface with tissues or tissue components. One should stress, that any artificial materials those simply are in contact with skin, such as hearing aids and wearable artificial limbs, are not included in the definition of biomaterials since the skin acts as a protective barrier between the body and the external world [1,2,63].
The major difference of biomaterials from other classes of materials lays in their ability to remain in a biological environment with neither damaging the surroundings nor being damaged in that process. Therefore, biomaterials must be distinguished from biological materials because the former are the materials that are accepted by living tissues and, therefore, they might be used for tissue replacements, while the latter are just the materials being produced by various biological systems (wood, cotton, bones, chitin, etc.) [64]. Furthermore, there are biomimetic materials, which are not made by living organisms but have the composition, structure and properties similar to those of biological materials. Concerning the subject of current review, bioceramics (or biomedical ceramics) is defined as biomaterials having the ceramic origin. Now it is important to define the meaning of ceramics. According to Wikipedia, the free encyclopedia: “The word ceramic comes from the Greek word κεραμικός (keramikos) meaning pottery, which is said to derive from the Indo-European word ker, meaning heat. A ceramic is an inorganic, non-metallic solid prepared by the action of heat and subsequent cooling. Ceramic materials may have a crystalline or partly crystalline structure, or may be amorphous (e.g., a glass). Because most common ceramics are crystalline, the definition of ceramic is often restricted to inorganic crystalline materials, as opposed to the non-crystalline glasses. Ceramic may be used as an adjective describing a material, product or process; or as a singular noun, or, more commonly, as a plural noun, ceramics.” [65]. As any other type of biomaterials, bioceramics can have structural functions as joint or tissue replacements, be used as coatings to improve the biocompatibility, as well as function as resorbable lattices, providing temporary structures and frameworks those are dissolved and/or replaced as the body rebuilds the damaged tissues [66,67,68,69,70,71]. Some types of bioceramics even feature a drug-delivery capability [72,73,74,75].
A progressive deterioration of all tissues with age is the major contributor to the need for spare parts for the body. Bones are especially vulnerable to fracture in older people due to a loss of density and strength with age. This effect is especially severe in women due to the hormonal changes associated with menopause. A graphical representation of the effect of time on bone strength and density from the age of 30 years onward is available in literature (Reference [68], Figure 1). Bone density decreases because bone-growing cells (osteoblasts) become progressively less productive in making new bone and repairing micro-fractures. The lower density greatly deteriorates the strength of bones and an unfortunate consequence is that many old people fracture their hips or have collapsed vertebrae and spinal problems [68].
In medicine, calcium orthophosphate bioceramics is needed to alleviate pain and restore functions of diseased or damaged calcified tissues (bones and teeth) of the body. A great challenge facing its medical application is, first, to replace and, second, to regenerate old and deteriorating bones with a biomaterial that can be replaced by a new mature bone without transient loss of a mechanical support [1,2]. Surface bioactivity is the major feature of calcium orthophosphate bioceramics. It contributes to a bone bonding ability and enhances new bone formation. After implantation, various interactions occur at the bioceramic/tissue interfaces leading to time-dependent changes in the surface characteristics of the implanted bioceramics and the surrounding tissues [76]. Because the average life span of humans is now 80+ years and the major need for spare parts begins at about 60 years of age, the after-effects of the implanted calcium orthophosphate bioceramics need to last, at least, for 20+ years. This demanding requirement of survivability is under conditions of use that are especially harsh to implanted biomaterials: corrosive saline solutions at 37 °C under variable, multiaxial and cyclical mechanical loads. The excellent performance of the specially designed calcium orthophosphate bioceramics that have survived these clinical conditions represented one of the most remarkable accomplishments of research, development, production and quality assurance by the end of the past century [68].

3. Bioceramics of Calcium Orthophosphates

3.1. History

The detailed history of hydroxyapatite and other calcium orthophosphates, including the subject of calcium orthophosphate bioceramics, as well as description of their past biomedical applications might be found elsewhere [77,78], where the interested readers are referred.

3.2. Chemical Composition and Preparation

Currently, calcium orthophosphate bioceramics can be prepared from various sources [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25]. Nevertheless, up to now, all attempts to synthesize bone replacement materials for clinical applications featuring the physiological tolerance, biocompatibility and a long-term stability have had only a relative success; this clearly demonstrates both the superiority and a complexity of the natural structures [79].
In general, a characterization of calcium orthophosphate bioceramics should be performed from various viewpoints such as the chemical composition (including stoichiometry and purity), homogeneity, phase distribution, morphology, grain sizes and shape, grain boundaries, crystallite size, crystallinity, pores, cracks, surface roughness, etc. From the chemical point of view, the vast majority of calcium orthophosphate bioceramics is based on HA [80,81,82,83,84,85], both types of TCP [80,86,87,88,89,90,91,92,93,94,95] and/or multiphase formulations thereof. Biphasic formulations (commonly abbreviated as BCP—biphasic calcium phosphate) are the simplest among the latter ones. They include β-TCP + HA [96,97,98,99,100,101,102,103,104,105,106,107,108], α-TCP + HA [26,27,28,109,110] and biphasic TCP (commonly abbreviated as BTCP) consisting of α-TCP and β-TCP [111,112,113,114,115,116]. In addition, triphasic formulations (HA + α-TCP + β-TCP) have been prepared as well [117,118,119,120]. Further details on this topic might be found in a special review [121]. Leaving aside a big subject of DCPD-forming self-setting formulations [122,123], one should note that just a few publications on bioceramics, prepared from other types of calcium orthophosphates, are available [124,125,126,127,128,129,130].
The preparation techniques of various calcium orthophosphates have been extensively reviewed in literature [6,131,132,133,134,135,136] and references therein] where the interested readers are referred to. Briefly, when compared to both α- and β-TCP, HA is a more stable phase under the physiological conditions, as it has a lower solubility (Table 1) and, thus, a slower resorption kinetics [137,138,139,140]. Therefore, the BCP concept is determined by the optimum balance of a more stable phase of HA and a more soluble TCP. Due to a higher biodegradability of the α- or β-TCP component, the reactivity of BCP increases with the TCP/HA ratio increasing. Thus, in vivo bioresorbability of BCP can be controlled through the phase composition [101]. Similar conclusions are also valid for the biphasic TCP (in which α-TCP is a more soluble phase), as well as for both triphasic (HA, α-TCP and β-TCP) and yet more complex formulations.
As implants made of sintered HA are found in bone defects for many years after implantation (Figure 2, bottom), bioceramics made of more soluble calcium orthophosphates [26,27,28,80,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,141,142,143] are preferable for the biomedical purposes (Figure 2, top). Furthermore, the experimental results showed that BCP had a higher ability to adsorb fibrinogen, insulin or type I collagen than HA [144]. Thus, according to both observed and measured bone formation parameters, calcium orthophosphate bioceramics have been ranked as follows: low sintering temperature BCP (rough and smooth) ≈ medium sintering temperature BCP ≈ TCP > calcined low sintering temperature HA > non-calcined low sintering temperature HA > high sintering temperature BCP (rough and smooth) > high sintering temperature HA [145]. This sequence has been developed in year 2000 and, thus, neither multiphase formulations, nor other calcium orthophosphates have been included.
Figure 2. Soft X-ray photographs of the operated portion of the rabbit femur. (a) Four weeks; (b) 12 weeks; (c) 24 weeks; and (d) 72 weeks after implantation of CDHA. (e) Four weeks; (f) 12 weeks; (g)24 weeks; and (h) 72 weeks after implantation of sintered HA. Reprinted from Reference [141] with permission.
Figure 2. Soft X-ray photographs of the operated portion of the rabbit femur. (a) Four weeks; (b) 12 weeks; (c) 24 weeks; and (d) 72 weeks after implantation of CDHA. (e) Four weeks; (f) 12 weeks; (g)24 weeks; and (h) 72 weeks after implantation of sintered HA. Reprinted from Reference [141] with permission.
Materials 06 03840 g002

3.3. Forming and Shaping

In order to fabricate bioceramics in progressively complex shapes, scientists are investigating the use of both old and new manufacturing techniques. These techniques range from an adaptation of the age-old pottery techniques to the newest manufacturing methods for high-temperature ceramic parts for airplane engines. Namely, reverse engineering [146,147] and rapid prototyping [148,149,150] technologies have revolutionized a generation of physical models, allowing the engineers to efficiently and accurately produce physical models and customized implants with high levels of geometric intricacy. Combined with the computer-aided design and manufacturing (CAD/CAM), complex physical objects of the anatomical structure can be fabricated in a variety of shapes and sizes. In a typical application, an image of a bone defect in a patient can be taken and used to develop a three-dimensional (3D) CAD computer model [151,152,153,154]. Then a computer can reduce the model to slices or layers. Afterwards, 3D objects and coatings are constructed layer-by-layer using rapid prototyping techniques. The examples comprise fused deposition modeling [155,156], selective laser sintering [157,158,159,160,161], laser cladding [162,163,164,165], 3D printing [91,166,167,168,169,170,171,172,173,174,175,176], solid freeform fabrication [177,178,179,180,181,182,183,184,185] and stereo lithography [186,187,188,189]. Furthermore, a thermal printing process of melted calcium orthophosphates has been proposed [190], while, in some cases, laser processing might be applied as well [191]. A schematic of 3D printing technique, as well as some 3D printed items are shown in Figure 3 [74]. A custom-made implant of actual dimensions would reduce the time it takes to perform the medical implantation procedure and subsequently lower the risk to the patient. Another advantage of a pre-fabricated, exact-fitting implant is that it can be used more effectively and applied directly to the damaged site rather than a replacement, which is formulated during surgery from a paste or granular material [178,191,192,193].
Figure 3. A schematic of 3D printing and some 3D printed parts (fabricated at Washington State University) showing the versatility of 3D printing technology for ceramic scaffolds fabrication with complex architectural features. Reprinted from Reference [74] with permission.
Figure 3. A schematic of 3D printing and some 3D printed parts (fabricated at Washington State University) showing the versatility of 3D printing technology for ceramic scaffolds fabrication with complex architectural features. Reprinted from Reference [74] with permission.
Materials 06 03840 g003
In addition to the aforementioned modern techniques, classical forming and shaping approaches are still widely used. The selection of the desired technique depends greatly on the ultimate application of the bioceramic device, e.g., whether it is for a hard-tissue replacement or an integration of the device within the surrounding tissues. In general, three types of the processing technologies might be used: (1) employment of a lubricant and a liquid binder with ceramic powders for shaping and subsequent firing; (2) application of self-setting and self-hardening properties of water-wet molded powders; (3) materials are melted to form a liquid and are shaped during cooling and solidification [194,195,196,197]. Since calcium orthophosphates are either thermally unstable (MCPM, MCPA, DCPA, DCPD, OCP, ACP, CDHA) or have a melting point at temperatures exceeding ~1400 °C with a partial decomposition (α-TCP, β-TCP, HA, FA, TTCP), only the first and the second consolidation approaches are used to prepare bulk bioceramics and scaffolds. The methods include uniaxial compaction [198,199,200], isostatic pressing (cold or hot) [108,201,202,203,204,205,206,207], granulation [208,209,210,211,212,213], loose packing [214], slip casting [93,215,216,217,218,219,220], gel casting [188,189,221,222,223,224,225], pressure mold forming [226], injection molding [227,228,229], polymer replication [230,231,232,233,234,235,236,237], extrusion [238,239,240,241,242], slurry dipping and spraying [243]. In addition, to form ceramic sheets from slurries, tape casting [105,223,244,245,246], doctor blade [247] and colander methods might be employed [194,195,196,197,248]. Various combinations of several techniques are also possible [95,223,249,250,251]. Furthermore, some of these processes might be performed under the electromagnetic field, which helps crystal aligning [216,220,252,253,254,255].
For example, powders are usually pressed damp in metal dies or dry in lubricated dies at pressures high enough to form sufficiently strong structures to hold together until they are sintered [256]. An organic binder such as polyvinyl alcohol helps to bind the powder together. The binder is removed by heating in air to oxidize the organic phases to carbon dioxide and water. Since many binders contain water, drying at ~100 °C is a critical step in preparing damp-formed pieces for firing. Too much or too little water in the compacts can lead to blowing apart the ware on heating or crumbling, respectively [194,195,196,197,202]. Furthermore, removal of water during drying often results in subsequent shrinkage of the product. In addition, due to local variations in water content, warping and even cracks may be developed during drying. Dry pressing and hydrostatic molding can minimize these problems [197]. Afterwards, the manufactured green samples are sintered.
Furthermore, forming and shaping of any ceramic products require a proper selection of the raw materials in terms of particle sizes and size distribution. Namely, tough and strong bioceramics consist of pure, fine and homogeneous microstructures. To attain this, pure powders with small average size and high surface area must be used as the starting sources. However, for maximum packing and least shrinkage after firing, mixing of ~70% coarse and ~30% fine powders have been suggested [197]. Mixing is usually carried out in a ball mill for uniformity of properties and reaction during subsequent firing. Mechanical die forming or sometimes extrusion through a die orifice can be used to produce a fixed cross-section.
Finally, to produce the accurate shaping, necessary for the fine design of bioceramics, machine finishing might be essential [153,194,257]. Unfortunately, cutting tools developed for metals are usually useless for bioceramics due to their fragility; therefore, grinding and polishing appear to be the convenient finishing techniques [153,194]. In addition, the surface of bioceramics might be modified by various supplementary treatments [258].

3.4. Sintering and Firing

A sintering (or firing) procedure appears to be of a great importance to manufacture bulk bioceramics with the required mechanical properties. Usually, this stage is carried out according to controlled temperature programs of electric furnaces in adjusted ambience of air with necessary additional gasses; however, always at temperatures below the melting points of the materials. The firing step can include temporary holds at intermediate temperatures to burn out organic binders [194,195,196,197]. The heating rate, sintering temperature and holding time depend on the starting materials. For example, in the case of HA, these values are in the ranges of 0.5–3 °C/min, 1000–1250 °C and 2–5 h, respectively [259]. In the majority cases, sintering allows a structure to retain its shape. However, this process might be accompanied by a considerable degree of shrinkage [260,261,262], which must be accommodated in the fabrication process. For instance, in the case of FA sintering, a linear shrinkage was found to occur at ~715 °C and the material reached its final density at ~890 °C. Above this value, grain growth became important and induced an intra-granular porosity, which was responsible for density decrease. At ~1180 °C, a liquid phase was formed due to formation of a binary eutectic between FA and fluorite contained in the powder as impurity. This liquid phase further promoted the coarsening process and induced formation of large pores at high temperatures [263].
In general, sintering occurs only when the driving force is sufficiently high, while the latter relates to the decrease in surface and interfacial energies of the system by matter (molecules, atoms or ions) transport, which can proceed by solid, liquid or gaseous phase diffusion. Namely, when solids are heated to high temperatures, their constituents are driven to move to fill up pores and open channels between the grains of powders, as well as to compensate for the surface energy differences among their convex and concave surfaces (matter moves from convex to concave). At the initial stages, bottlenecks are formed and grow among the particles (Figure 4). Existing vacancies tend to flow away from the surfaces of sharply curved necks; this is an equivalent of a material flow towards the necks, which grow as the voids shrink. Small contact areas among the particles expand and, at the same time, a density of the compact increases and the total void volume decreases. As the pores and open channels are closed during a heat treatment, the particles become tightly bonded together and density, strength and fatigue resistance of the sintered object improve greatly. Grain-boundary diffusion was identified as the dominant mechanism for densification [264]. Furthermore, strong chemical bonds are formed among the particles and loosely compacted green bodies are hardened to denser materials [194,195,196,197]. Further knowledge on the ceramic sintering process might be found elsewhere [265].
Figure 4. A schematic diagram representing the changes occurring with particles under sintering. Shrinkage is noticeable.
Figure 4. A schematic diagram representing the changes occurring with particles under sintering. Shrinkage is noticeable.
Materials 06 03840 g004
In the case of calcium orthophosphates, the earliest paper on their sintering was published in 1971 [266]. Since then, numerous papers on this subject were published and several specific processes were found to occur during calcium orthophosphate sintering. Firstly, moisture, carbonates and all other volatile chemicals remaining from the synthesis stage, such as ammonia, nitrates and any organic compounds, are removed as gaseous products. Secondly, unless powders are sintered, the removal of these gases facilitates production of denser ceramics with subsequent shrinkage of the samples (Figure 5). Thirdly, all chemical changes are accompanied by a concurrent increase in crystal size and a decrease in the specific surface area. Fourthly, a chemical decomposition of all acidic orthophosphates and their transformation into other phosphates (e.g., 2HPO42− → P2O74− + H2O↑) takes place. Besides, sintering causes toughening [84], densification [85,267], partial dehydroxylation (in the case of HA) [85], grain growth [264,268], as well as it increases the mechanical strength [269,270,271]. The latter events are due to presence of air and other gases filling gaps among the particles of un-sintered powders. At sintering, the gases move towards the outside of powders and green bodies shrink owing to decrease of distances among the particles. For example, sintering of a biologically formed apatites was investigated [272,273] and the obtained products were characterized [274,275]. In all cases, the numerical value of Ca/P ratio in sintered apatites of biological origin was higher than that of the stoichiometric HA. One should mention that in the vast majority cases, calcium orthophosphates with Ca/P ratio <1.5 (Table 1) are not sintered, since these compounds are thermally unstable, while sintering of non-stoichiometric calcium orthophosphates (CDHA and ACP) always leads to their transformation into various types of biphasic, triphasic and multiphase formulations [121].
Figure 5. Linear shrinkage of the compacted ACP powders that were converted into β-TCP, BCP (50% HA + 50% β-TCP) and HA upon heating. According to the authors: “At 1300 °C, the shrinkage reached a maximum of approximately ~25%, ~30% and ~35% for the compacted ACP powders that converted into HA, BCP 50/50 and β-TCP, respectively” [261]. Reprinted from Reference [261] with permission.
Figure 5. Linear shrinkage of the compacted ACP powders that were converted into β-TCP, BCP (50% HA + 50% β-TCP) and HA upon heating. According to the authors: “At 1300 °C, the shrinkage reached a maximum of approximately ~25%, ~30% and ~35% for the compacted ACP powders that converted into HA, BCP 50/50 and β-TCP, respectively” [261]. Reprinted from Reference [261] with permission.
Materials 06 03840 g005
An extensive study on the effects of sintering temperature and time on the properties of HA bioceramics revealed a correlation between these parameters and density, porosity, grain size, chemical composition and strength of the scaffolds [276]. Namely, sintering below ~1000 °C was found to result in initial particle coalescence, with little or no densification and a significant loss of the surface area and porosity. The degree of densification appeared to depend on the sintering temperature whereas the degree of ionic diffusion was governed by the period of sintering [276]. To enhance sinterability of calcium orthophosphates, a variety of sintering additives might be added [277,278,279,280].
Solid-state pressureless sintering is the simplest procedure. For example, HA bioceramics can be pressurelessly sintered up to the theoretical density at 1000–1200 °C. Processing at even higher temperatures usually lead to exaggerated grain growth and decomposition because HA becomes unstable at temperatures exceeding ~1300 °C [6,131,132,133,134,135,136,281,282,283,284]. The decomposition temperature of HA bioceramics is a function of the partial pressure of water vapor. Moreover, processing under vacuum leads to an earlier decomposition of HA, while processing under high partial pressure of water prevents from the decomposition. On the other hand, a presence of water in the sintering atmosphere was reported to inhibit densification of HA and accelerated grain growth [248,285]. Unexpectedly, an application of a magnetic field during sintering was found to influence the growth of HA grains [268]. A definite correlation between hardness, density and a grain size in sintered HA bioceramics was found: despite exhibiting high bulk density, hardness started to decrease at a certain critical grain size limit [286,287,288].
Since grain growth occurs mainly during the final stage of sintering, to avoid this, a new method called ‘‘two-step sintering’’ (TSS) was proposed [289]. The method consists of suppressing grain boundary migration responsible for grain growth, while keeping grain boundary diffusion that promotes densification. The TSS approach was successfully applied to calcium orthophosphate bioceramics [107,290,291,292,293]. For example, HA compacts prepared from nanodimensional powders were two-step sintered. The average grain size of near full dense (>98%) HA bioceramics made via conventional sintering was found to be ~1.7 μm, while that for TSS HA bioceramics was ~190 nm (i.e., ~9 times less) with simultaneous increasing the fracture toughness of samples from 0.98 ± 0.12 to 1.92 ± 0.20 MPa m1/2. In addition, due to the lower second step sintering temperature, no HA phase decomposition was detected in TSS method [290].
Hot pressing [288,294,295,296,297,298,299], hot isostatic pressing [108,201,206] or hot pressing with post-sintering [300,301] processes make it possible to decrease a temperature of the densification process, diminish the grain size, as well as achieve higher densities. This leads to finer microstructures, higher thermal stability and subsequently better mechanical properties of calcium orthophosphate bioceramics. Microwave [18,302,303,304,305,306,307,308,309,310] and spark plasma [87,124,311,312,313,314,315,316,317,318,319,320,321] sintering techniques are alternative methods to the conventional sintering, hot pressing and hot isostatic pressing. Both alternative methods were found to be time and energy efficient densification techniques. Further developments are still possible. For example, a hydrothermal hot pressing method has been developed to fabricate OCP [125], CDHA [322], HA/β-TCP [296] and HA [297,298,300,323] bioceramics with neither thermal dehydration nor thermal decomposition. Further details on the sintering and firing processes of calcium orthophosphate bioceramics are available in literature [67,133,248,324,325].
To conclude this part, one should mention that the sintering stage is not always necessary. For example, calcium orthophosphate-based bulk bioceramics with the reasonable mechanical properties might be prepared by means of self-setting (self-hardening) formulations (see Section 5.1 below). Furthermore, the reader’s attention is paid on an excellent review on various ceramic manufacturing techniques [326], in which various ceramic processing techniques are well described.

4. The Major Properties

4.1. Mechanical Properties

The modern generation of biomedical materials should stimulate the body’s own self-repairing abilities [327]. Therefore, during healing, a mature bone should replace the modern grafts and this process must occur without transient loss of the mechanical support. Unluckily for material scientists, a human body provides one of the most inhospitable environments for the implanted biomaterials. It is warm, wet and both chemically and biologically active. For example, a diversity of body fluids in various tissues might have a solution pH varying from 1 to 9. In addition, a body is capable of generating quite massive force concentrations and the variance in such characteristics among individuals might be enormous. Typically, bones are subjected to ~4 MPa, whereas tendons and ligaments experience peak stresses in the range of 40–80 MPa. The hip joints are subjected to an average load up to three times body weight (3000 N) and peak loads experienced during jumping can be as high as 10 times body weight. These stresses are repetitive and fluctuating depending on the nature of the activities, which can include standing, sitting, jogging, stretching and climbing. Therefore, all types of implants must sustain attacks of a great variety of aggressive conditions [328]. Regrettably, there is presently no artificial material fulfilling all these requirements.
Now it is important to mention, that the mechanical behavior of any ceramics is rather specific. Namely, ceramics is brittle, which is attributed to high strength ionic bonds. Thus, it is not possible for plastic deformation to happen prior to failure, as a slip cannot occur. Therefore, ceramics fail in a dramatic manner. Namely, if a crack is initiated, its progress will not be hindered by the deformation of material ahead of the crack, as would be the case in a ductile material (e.g., a metal). In ceramics, the crack will continue to propagate, rapidly resulting in a catastrophic breakdown [195]. All of these are applicable to calcium orthophosphate bioceramics.
For dense bioceramics, the strength is a function of the grain sizes. It appears to be very sensitive to a slow crack growth [329]. Finer grain size bioceramics have smaller flaws at the grain boundaries and thus are stronger than one with larger grain sizes. In general, the mechanical properties decrease significantly with increasing content of an amorphous phase, microporosity and grain sizes, while a high crystallinity, a low porosity and small grain sizes tend to give a higher stiffness, a higher compressive and tensile strength and a greater fracture toughness. Accordingly, from the mechanical point of view, calcium orthophosphate bioceramics appear to be brittle polycrystalline materials for which the mechanical properties are governed by crystallinity, grain size, grain boundaries, porosity and composition [207]. Thus, it possesses poor mechanical properties (for instance, a low impact and fracture resistances) that do not allow calcium orthophosphate bioceramics to be used in load-bearing areas, such as artificial teeth or bones [66,67,68,69,70,71,330]. For example, fracture toughness (this is a property, which describes the ability of a material containing a crack to resist fracture and is one of the most important properties of any material for virtually all design applications) of HA bioceramics does not exceed the value of ~1.2 MPa·m1/2 [331] (human bone: 2–12 MPa·m1/2). It decreases almost linearly with a porosity increasing [248]. Generally, fracture toughness increases with grain size decreasing. However, in some materials, especially non-cubic ceramics, fracture toughness reaches the maximum and rapidly drops with decreasing grain size. For example, a fracture toughness of pure hot pressed HA with grain sizes between 0.2–1.2 µm was investigated. The authors found two distinct trends, where fracture toughness decreased with increasing grain size above ~0.4 µm and subsequently decreased with decreasing grain size. The maximum fracture toughness measured was 1.20 ± 0.05 MPa·m1/2 at ~0.4 µm [299]. Fracture energy of HA bioceramics is in the range of 2.3–20 J/m2, while the Weibull modulus (it is a measure of the spread or scatter in fracture strength) is low (~5–12) in wet environments, which means that HA behaves as a typical brittle ceramics and indicates to a low reliability of HA implants [248]. Porosity has a great influence on the Weibull modulus [332,333]. Interestingly that 3 peaks of internal friction were found at temperatures about −40, 80 and 130 °C for HA but no internal friction peaks were obtained for FA in the measured temperature range; this effect was attributed to the differences of F and OH positions in FA and HA, respectively [334]. The differences in internal friction values were also found between HA and TCP [335].
Bending, compressive and tensile strengths of dense HA bioceramics are in the ranges of 38–250 MPa, 120–900 MPa and 38–300 MPa, respectively. Similar values for porous HA bioceramics are substantially lower: 2–11 MPa, 2–100 MPa and ~3 MPa, respectively [248]. These wide variations in the properties are due to both structural variations (e.g., an influence of remaining microporosity, grain sizes, presence of impurities, etc.) and manufacturing processes, as well as they are caused by a statistical nature of the strength distribution. Strength was found to increase with Ca/P ratio increasing, reaching the maximum value around Ca/P ~1.67 (stoichiometric HA) and decreases suddenly when Ca/P > 1.67 [248]. Furthermore, strength decreases almost exponentially with porosity increasing [96,97]. However, by changing the pore geometry, it is possible to influence the strength of porous bioceramics. It is also worth mentioning that porous HA bioceramics is considerably less fatigue resistant than dense ones (in materials science, fatigue is the progressive and localized structural damage that occurs when a material is subjected to cyclic loading). Both grain sizes and porosity are reported to influence the fracture path, which itself has a little effect on the fracture toughness of calcium orthophosphate bioceramics [207,336]. However, no obvious decrease in mechanical properties was found after calcium orthophosphate bioceramics had been aged in the various solutions during the different periods of time [337].
Young’s (or elastic) modulus of dense HA bioceramics is in the range of 35–120 GPa [338,339], which is more or less similar to those of the most resistant components of the natural calcified tissues (dental enamel: ~74 GPa, dentine: ~21 GPa, compact bone: ~18–22 GPa). This value depends on porosity [340]. Nevertheless, dense bulk compacts of HA have mechanical resistances of the order of 100 MPa vs. ~300 MPa of human bones, diminishing drastically their resistances in the case of porous bulk compacts [341]. Young’s modulus measured in bending is between 44 and 88 GPa. To investigate the subject in more details, various types of modeling and calculations are increasingly used [342,343,344,345,346]. For example, the elastic properties of HA appeared to be significantly affected by the presence of vacancies, which softened HA via reducing its elastic modules [346]. In addition, a considerable anisotropy in the stress-strain behavior of the perfect HA crystals was found by ab initio calculations [343]. The crystals appeared to be brittle for tension along the z-axis with the maximum stress of ~9.6 GPa at 10% strain. Furthermore, the structural analysis of the HA crystal under various stages of tensile strain revealed that the deformation behavior manifested itself mainly in the rotation of PO4 tetrahedrons with concomitant movements of both the columnar and axial Ca ions [343]. Data for single crystals are also available [347]. Vickers hardness (that is a measure of the resistance to permanent indentation) of dense HA bioceramics is within 3–7 GPa, while the Poisson’s ratio (that is the ratio of the contraction or transverse strain to the extension or axial strain) for HA is about 0.27, which is close to that of bones (~0.3). At temperatures within 1000–1100 °C, dense HA bioceramics was found to exhibit superplasticity with a deformation mechanism based on grain boundary sliding [321,348,349]. Furthermore, both wear resistance and friction coefficient of dense HA bioceramics are comparable to those of dental enamel [248].
Due to a high brittleness (associated to a low crack resistance), the biomedical applications of calcium orthophosphate bioceramics are focused on production of non-load-bearing implants, such as pieces for middle ear surgery, filling of bone defects in oral or orthopedic surgery, as well as coating of dental implants and metallic prosthesis (see below) [79,350,351]. Therefore, ways are continuously sought to improve the reliability of calcium orthophosphate bioceramics. Namely, the mechanical properties of sintered bioceramics might be improved by changing the morphology of the initial calcium orthophosphates [352]. In addition, diverse reinforcements (ceramics, metals or polymers) have been applied to manufacture various biocomposites and hybrid biomaterials [353], but that is another story. However, successful hybrid formulations consisted of calcium orthophosphates only [354,355,356,357,358,359,360,361] are within the scope of this review. Namely, bulk HA bioceramics might be reinforced by HA whiskers [355,356,357,358,359]. Furthermore, various biphasic apatite/TCP formulations were tested [354,360,361] and, for example, a superior superplasticity of HA/β-TCP biocomposites to HA bioceramics was detected [360].
Another approach to improve the mechanical properties of calcium orthophosphate bioceramics is to cover the items by polymeric coatings [362,363,364] or infiltrate porous structures by polymers [365]; however, this is still other story. Further details on the mechanical properties of calcium orthophosphate bioceramics are available elsewhere [248,366], where the interested readers are referred to.

4.2. Electrical Properties

Occasionally, an interest to the electrical properties of calcium orthophosphates is expressed [306,367,368,369,370]. For example, a surface ionic conductivity of both porous and dense HA bioceramics was examined for humidity sensor applications, since the room temperature conductivity was influenced by relative humidity [371]. Namely, the ionic conductivity of HA has been a subject of research for its possible use as a gas sensor for alcohol [372], carbon dioxide [372,373] or carbon monoxide [374]. Electrical measurements have also been used as a characterization tool to study the evolution of microstructure in HA bioceramics [375]. More to the point, Valdes et al. examined the dielectric properties of HA to understand its decomposition to β-TCP [376]. In the case of CDHA, the electrical properties, in terms of ionic conductivity, were found to increase after compression of the samples at 15 t/cm2, which was attributed to establishment of some order within the apatitic network [377]. The conductivity mechanism of CDHA appeared to be multiple [378]. Furthermore, there was an attempt to develop CDHA whisker electrets for biomedical utilization [379].
Interestingly that the electrical properties of calcium orthophosphate bioceramics appear to influence their biomedical applications. For example, there is an interest in polarization of HA bioceramics to generate a surface charge by the application of electric fields at elevated temperatures [380,381]. The presence of surface charges on HA was shown to have a significant effect on both in vitro and in vivo crystallization of biological apatite [382,383,384,385,386,387,388]. Furthermore, a growth of both biomimetic calcium orthophosphates and bones was found to be accelerated on negatively charged surfaces and decelerated at positively charged surfaces [386,387,388,389,390,391,392,393,394,395,396,397,398,399]. Similar effect was found for adsorption of bovine serum albumin [400]. In addition, the electrical polarization of calcium orthophosphates was found to accelerate a cytoskeleton reorganization of osteoblast-like cells [401,402,403,404], extend bioactivity [405] and enhance bone ingrowth through the pores of porous implants [406]. The positive effect of electric polarization was found for carbonated apatite as well [407]. There is an interesting study on the interaction of a blood coagulation factor on electrically polarized HA surfaces [408]. Further details on the electrical properties of calcium orthophosphate-based bioceramics are available in literature [306,409,410,411,412,413].

4.3. Possible Transparency

Single crystals of all calcium orthophosphates are optically transparent for the visible light. As bioceramics of calcium orthophosphates have a polycrystalline nature with a random orientation of big amounts of small crystals, it is opaque and of white color, unless colored dopants have been added. However, in some cases, a transparency is convenient to provide some essential advantages (e.g., to enable direct viewing of living cells, their attachment, spreading, proliferation, and osteogenic differentiation cascade in a transmitted light). Thus, transparent calcium orthophosphate bioceramics (Figure 6) [414] have been prepared and investigated [87,108,201,203,317,319,320,321,414,415,416,417,418,419,420,421,422,423,424]. They can exhibit an optical transmittance of ~66% at a wavelength of 645 nm [423]. The preparation techniques include a hot isostatic pressing [108,201,203], an ambient-pressure sintering [415], gel casting coupled with a low-temperature sintering [418,421], a pulse electric current sintering [419], as well as a spark plasma sintering [87,317,318,319,320,321]. Fully dense, transparent calcium orthophosphate bioceramics are obtained at temperatures above ~800 °C. Depending on the preparation technique, the transparent bioceramics has a uniform grain sizes ranging from ~67 nm [108] to ~250 μm [418] and always is pore-free. Furthermore, a translucent calcium orthophosphate bioceramics is also known [108,266,425,426,427]. However, due to a lack of both porosity and the necessity to have see-through implants inside the body, the transparent and translucent forms of calcium orthophosphate bioceramics will hardly be ever used in medicine with the specific eye implants as the only reasonable exception.
Figure 6. Transparent HA bioceramics prepared by spark plasma sintering at 900 °C from nano-sized HA single crystals. Reprinted from Reference [414] with permission.
Figure 6. Transparent HA bioceramics prepared by spark plasma sintering at 900 °C from nano-sized HA single crystals. Reprinted from Reference [414] with permission.
Materials 06 03840 g006

4.4. Porosity

Porosity is defined as a percentage of voids in solids and this morphological property is independent of the material. The surface area of porous bodies is much higher, which guarantees a good mechanical fixation in addition to providing sites on the surface that allow chemical bonding between the bioceramics and bones [428]. Furthermore, a porous material may have both closed (isolated) pores and open (interconnected) pores. The latter look like tunnels and are accessible by gases, liquids and particulate suspensions [429]. The open-cell nature of porous materials (also known as reticulated materials) is a unique characteristic essential in many applications. In addition, pore dimensions are also important. Namely, the dimensions of open pores are directly related to bone formation, since such pores grant both the surface and space for cell adhesion and bone ingrowth [430,431]. On the other hand, pore interconnection provides the ways for cell distribution and migration, as well as it allows an efficient in vivo blood vessel formation suitable for sustaining bone tissue neo-formation and possibly remodeling [144,432,433,434,435,436,437,438,439,440,441,442]. Namely, porous calcium orthophosphate bioceramics can be colonized by bone tissues [437,442,443,444,445,446,447,448,449,450,451,452,453,454]. Therefore, interconnecting macroporosity (pore size >100 μm) [104,428,437,455,456], which is defined by its capacity to be colonized by cells, is intentionally introduced in solid bioceramics (Figure 7).
Figure 7. Photographs of a commercially available porous calcium orthophosphate bioceramics with (a,b) different porosity; and (c) a method of their production. For photos, the horizontal field width is 20 mm. The picture (c) is reprinted from Reference [457] with permission.
Figure 7. Photographs of a commercially available porous calcium orthophosphate bioceramics with (a,b) different porosity; and (c) a method of their production. For photos, the horizontal field width is 20 mm. The picture (c) is reprinted from Reference [457] with permission.
Materials 06 03840 g007
Macroporosity is usually formed due to a release of various easily removable compounds and, for that reason, incorporation of pore-creating additives (porogens) is the most popular technique to create macroporosity. The porogens are crystals, particles or fibers of either volatile (they evolve gases at elevated temperatures) or soluble substances, such as paraffin [458,459,460], naphthalene [207,461,462,463], sucrose [464,465], NaHCO3 [466,467,468], NaCl [469,470], polymethylmethacrylate [92,471,472,473], hydrogen peroxide [474,475,476,477,478] and cellulose derivatives [82]. Several other compounds [96,181,324,479,480,481,482,483,484,485,486,487,488,489,490] might be used as porogens, as well. Obviously, the ideal porogen should be nontoxic and be removed at ambient temperature, thereby allowing the bioceramic/porogen mixture to be injected directly into a defect site and allowing the scaffold to fit the defect [491]. Sintering particles, preferably spheres of equal size, is a similar way to generate porous 3D bioceramics of calcium orthophosphates (Figure 8). However, pores resulting from this method are often irregular in size and shape and not fully interconnected with one another.
Figure 8. β-TCP porous ceramics with different pore sizes prepared using polymethylmethacrylate balls with diameter equal to: (a) 100–200 μm; (b) 300–400 μm; (c) 500–600 μm; and (d) 700–800 μm. Horizontal field width is 45 mm. Reprinted from Reference [92] with permission.
Figure 8. β-TCP porous ceramics with different pore sizes prepared using polymethylmethacrylate balls with diameter equal to: (a) 100–200 μm; (b) 300–400 μm; (c) 500–600 μm; and (d) 700–800 μm. Horizontal field width is 45 mm. Reprinted from Reference [92] with permission.
Materials 06 03840 g008
Many other techniques, such as replication of polymer foams by impregnation [230,231,232,235,492,493,494,495] (Figure 7), various types of casting [217,219,223,225,496,497,498,499,500,501,502,503,504,505], suspension foaming [120], surfactant washing [506], as well as several other approaches [86,89,92,161,507,508,509,510,511,512,513,514,515,516,517,518,519,520,521,522,523,524,525,526,527,528,529,530,531,532,533,534,535,536,537,538,539,540,541,542,543,544,545,546,547,548] have been applied to fabricate porous calcium orthophosphate bioceramics. Some of them have been summarized in Table 2 [491]. In addition, natural porous materials, as coral skeletons [549,550,551] or shells [552] made of CaCO3, can be converted into porous calcium orthophosphates under the hydrothermal conditions (250 °C, 24–48 h) with the microstructure undamaged. Porous HA bioceramics can also be obtained by hydrothermal hot pressing. This technique allows solidification of the HA powder at 100–300 °C (30 MPa, 2 h) [323]. In another approach, bi-continuous water-filled microemulsions have been used as pre-organized systems for the fabrication of needle-like frameworks of crystalline HA (2 °C, 3 weeks) [510,511]. Besides, porous calcium orthophosphates might be prepared by a combination of gel casting and foam burn out methods [249,251], as well as by setting of the self-setting formulations [459,460,467,468,470,479,480,543]. Lithography was used to print a polymeric material, followed by packing with HA and sintering [514]. Both hot pressing [294,295] and ice templating [515,516,517,518] techniques were applied as well. More to the point, a HA suspension can be cast into a porous CaCO3 skeleton, which is then dissolved, leaving a porous network [508]. 3D periodic macroporous frame of HA has been fabricated via a template-assisted colloidal processing technique [519,520]. More to the point, porous HA bioceramics might be prepared by using different starting HA powders and sintering at various temperatures by a pressureless sintering [523]. Porous bioceramics with an improved strength might be fabricated from calcium orthophosphate fibers or whiskers. In general, fibrous porous materials are known to exhibit an improved strength due to fiber interlocking, crack deflection and/or pullout [553]. Namely, porous bioceramics with well-controlled open pores was processed by sintering of fibrous HA particles [512]. In another approach, porosity was achieved by firing apatite-fiber compacts mixed with carbon beads and agar. By varying the compaction pressure, firing temperature and carbon/HA ratio, the total porosity was controlled in the ranges from ~40% to ~85% [82]. Finally, a superporous (~85% porosity) HA bioceramics was developed as well [540,541,542]. Additional information on the processing routes to produce porous ceramics might be found in literature [554].
Bioceramic microporosity (pore size <10 μm), which is defined by its capacity to be impregnated by biological fluids [555], results from the sintering process, while the pore dimensions mainly depend on the material composition, thermal cycle and sintering time. The microporosity provides both a greater surface area for protein adsorption and increased ionic solubility. For example, embedded osteocytes distributed throughout microporous rods might form a mechanosensory network, which would not be possible in scaffolds without microporosity [556,557]. Calcium orthophosphate bioceramics with nanodimensional (<100 nm) pores might be fabricated as well [196,558,559,560,561,562]. It is important to stress, that differences in porogens usually influence the bioceramics’ macroporosity, while differences in sintering temperatures and conditions affect the percentage of microporosity. Usually, the higher the sintering temperature, the lower both the microporosity content and the specific surface area of bioceramics. Namely, HA bioceramics sintered at ~1200 °C shows significantly less microporosity and a dramatic change in crystal sizes, if compared with that sintered at ~1050 °C (Figure 9) [563]. Furthermore, the average shape of pores was found to transform from strongly oblate to round at higher sintering temperatures [564]. The total porosity (macroporosity + microporosity) of calcium orthophosphate bioceramics was reported to be ~70% [565] or even ~85% [540,541,542] of the entire volume. In the case of coralline HA or bovine-derived apatites, the porosity of the original biologic material (coral or bovine bone) is usually preserved during processing [566]. To finalize the production topic, creation of the desired porosity in calcium orthophosphate bioceramics is a rather complicated engineering task and the interested readers are referred to additional publications on the subject [96,433,567,568,569,570,571,572,573,574,575,576,577,578,579].
Table 2. The procedures used to manufacture porous calcium orthophosphate scaffolds for tissue engineering [491].
Table 2. The procedures used to manufacture porous calcium orthophosphate scaffolds for tissue engineering [491].
YearLocationProcessApatite from:SinteringCompressive strengthPore sizePorosityMethod of porosity control
2006Deville et al. Berkeley, CAHA + ammonium methacrylate in PTFE mold, freeze dried and sinteredHA #30Yes: 1300 °C16 MPa
65 MPa
145 MPa
open unidirectional 50–150 μm>60%
56%
47%
Porosity control: slurry conc. Structure controlled by physics of ice front formation.
2006Saiz et al. Berkeley, CAPolymer foams coated, compressed after infiltration, then calcined.HA powderYes: 700–1300 °C100–200 μmPorosity control: extent of compression, HA loading
2006Murugan et al. Singapore + USABovine bone cleaned, calcinedbovine boneYes: 500 °Cretention of nano-sized poresPorosity control: native porosity of bovine bone
2006Xu et al. Gaithersburg, MDDirectly injectable calcium orthophosphate cement, self hardens, mannitol as porogennanocrystalline HANo2.2–4.2 MPa (flexural)0%–50% macroporous65%–82%Porosity control: mannitol mass fraction in mixture
2004Landi et al. Italy + IndonesiaSponge impregnation, isotactic pressing, sintering of HA in simulated body fluidCaO + H3PO4Yes: 1250 °C for 1 h23 ± 3.8 MPaclosed 6% open 60%66%Porosity control: possibly by controlling HA particle size. Not suggested by authors
2003Charriere et al. EPFL, SwitzerlandThermoplastic negative porosity by Ink jet printing, slip casting process for HADCPA + calciteNo: 90 °C for 1 day12.5 ± 4.6 MPa44%Porosity control: negative printing
2003Almirall et al. Barcelona, Spainα-TCP foamed with hydrogen peroxide at different conc., liq. ratios, poured in PTFE moldsα-TCP + (10% and 20% H2O2)No: 60 °C for 2 h1.41 ± 0.27 MPa
2.69 ± 0.91 MPa
35.7% macro
29.7% micro
26.8% macro
33.8% micro
65.5%
60.7%
Porosity control: different concentration, α-TCP particle sizes
2003Ramay et al. Seattle, WASlurries of HA prepared: gel-casting + polymer sponge technique, sintered.HA powderYes: 600 °C for 1 h 1350 °C for 2 h0.5–5 MPa200–400 μm70%–77%Porosity control: replicate of polymer sponge template
2003Miao et al. SingaporeTTCP to calcium orthophosphate cement cement. Slurry cast on polymer foam, sintered.TTCPYes: 1200 °C for 2 h1 mm macro
5 μm micro
~70%Porosity control: Recoating time, polyurethane foam
2003Uemura et al. China + JapanSlurry of HA with polyoxyethylenelaurylether (cross-linked) and sinteredHA powdersYes: 1200 °C for 3 h2.25 MPa (0 wk) 4.92 MPa (12 wks)
11.2 MPa (24 wks)
500 μm
200 μm interconnects
~77%Porosity control: polymer interconnects cross-linking
2003Ma et al. Singapore + USAElectrophoretic deposition of HA, sintering.HA powdersYes: 1200 °C for 2 h860 MPa0.5 μm
130 μm
~20%Porosity control: electrophoresis field
2002Barralet et al. Birmingham, London, UKCalcium orthophosphate cement + sodium phosphate ice, evaporatedCaCO3 + DCPD1st step: 1400 °C for 1 day0.6 ± 0.27 MPa2 μm62% ± 9%Porosity control: porogen shape.
Figure 9. SEM pictures of HA bioceramics sintered at (A) 1050 °C; and (B) 1200 °C. Note the presence of microporosity in (A) and not in (B). Scale bar is 1 μm. Reprinted from Reference [563] with permission.
Figure 9. SEM pictures of HA bioceramics sintered at (A) 1050 °C; and (B) 1200 °C. Note the presence of microporosity in (A) and not in (B). Scale bar is 1 μm. Reprinted from Reference [563] with permission.
Materials 06 03840 g009
Concerning the biomedical importance of porosity, studies revealed that increasing of both the specific surface area and pore volume of bioceramics might greatly accelerate the in vivo process of apatite deposition and, therefore, enhance the bone-forming bioactivity. More importantly, a precise control over the porosity, pore dimensions and internal pore architecture of bioceramics on different length scales is essential for understanding of the structure-bioactivity relationship and the rational design of better bone-forming biomaterials [576,580,581]. Namely, in antibiotic charging experiments, a calcium orthophosphate bioceramics with nanodimensional (<100 nm) pores showed a much higher charging capacity (1621 μg/g) than that of commercially available calcium orthophosphate (100 μg/g), which did not contain nanodimensional porosity [572]. In other experiments, porous blocks of HA were found to be viable carriers with sustained release profiles for drugs [582] and antibiotics over 12 days [583] and 12 weeks [584], respectively. Unfortunately, porosity significantly decreases the strength of implants [248,336,366]. Thus, porous calcium orthophosphate implants cannot be loaded and are used to fill only small bone defects. However, their strength increases gradually when bones ingrow into the porous network of calcium orthophosphate implants [139,585,586,587,588]. For example, Martin et al. reported bending strengths of 40–60 MPa for a porous HA implant filled with 50%–60% of cortical bone [585], while in another study an ingrown bone increased strength of porous HA bioceramics by a factor of 3 to 4 [587].
Unfortunately, the biomedical effects of bioceramics’ porosity are not straightforward. For example, the in vivo response of calcium orthophosphates of different porosity was investigated and a hardly any effect of macropore dimensions (~150, ~260, ~510 and ~1220 μm) was observed [589]. In another study, a greater differentiation of mesenchymal stem cells was observed when cultured on ~200 μm pore size HA scaffolds when compared to those on ~500 μm pore size HA [590]. The latter finding was attributed to the fact that a higher pore volume in ~500 μm macropore scaffolds might contribute to a lack of cell confluency leading to the cells proliferating before beginning differentiation. Besides, the authors hypothesized that bioceramics having a less than the optimal pore dimensions induced quiescence in differentiated osteoblasts due to reduced cell confluency [590]. In still another study, the use of BCP (HA/TCP = 65/35 wt. %) scaffolds with cubic pores of ~500 μm resulted in the highest bone formation compared with the scaffold with lower (~100 μm) or higher (~1000 μm) pore sizes [591]. Furthermore, calcium orthophosphate bioceramics with greater strut porosity appeared to be more osteoinductive [592]. Already in 1979, Holmes suggested that the optimal pore range was 200–400 μm with the average human osteon size of ~223 μm [550]. In 1997, Tsurga and coworkers implied that the optimal pore size of bioceramics that supported ectopic bone formation was 300–400 μm [593]. Thus, there is no need to create calcium orthophosphate bioceramics with very big pores; however, the pores must be interconnected [440,455,456,594]. Interconnectivity governs a depth of cells or tissue penetration into the porous bioceramics, as well as it allows development of blood vessels required for new bone nourishing and wastes removal [555,595]. Nevertheless, the total porosity of implanted bioceramics appears to be important. For example, 60% porous β-TCP granules achieved a higher bone fusion rate than 75% porous β-TCP granules in lumbar posterolateral fusion [556].

5. Biomedical Applications

Since Levitt et al. described a method of preparing a FA bioceramics and suggested its possible use in medical applications in 1969 [596], calcium orthophosphate bioceramics have been widely tested for clinical applications. Namely, a great number of forms, compositions and trade-marks (Table 3) currently are either in use or under a consideration in many areas of orthopedics and dentistry, with even more in development. For example, bulk materials, available in dense and porous forms, are used for alveolar ridge augmentation, immediate tooth replacement and maxillofacial reconstruction [4]. Other examples comprise burr-hole buttons [597,598], orbital implants (including Bio-Eye®) [599,600,601,602,603,604,605,606], increment of the hearing ossicles [607,608,609], spine fusion [610,611,612,613] and repair of bone defects [138,614,615]. In order to permit growth of new bone into defects, a suitable bioresorbable material should fill these defects. Otherwise, ingrowth of fibrous tissue might prevent bone formation within the defects.
Table 3. Registered commercial trademarks (current and past) of calcium orthophosphate-based bioceramics and biomaterials.
Table 3. Registered commercial trademarks (current and past) of calcium orthophosphate-based bioceramics and biomaterials.
Calcium orthophosphateTrade name and producer (when available)
CDHACalcibon (Biomet, IN, USA)
Cementek (Teknimed, France)
Osteogen (Impladent, NY, USA)
HAActifuse (ApaTech, UK)
Alveograf (Cooke-Waite Laboratories, USA)
Apaceram (HOYA Corp., PENTAX New Ceramics Division, Japan)
ApaPore (ApaTech, UK)
Bio-Eye (Integrated Orbital Implants, CA, USA)
BioGraft (IFGL BIO CERAMICS, India)
Bioroc (Depuy-Bioland, France)
Bonefil (Pentax, Japan)
Bonetite (Pentax, Japan)
Boneceram (Sumitomo Osaka Cement, Japan)
BoneSource (Stryker Orthopaedics, NJ, USA)
Calcitite (Zimmer, IN, USA)
Cerapatite (Ceraver, France)
Durapatite (unknown producer)
HA BIOCER (CHEMA–ELEKTROMET, Poland)
HAnan° Surface (Promimic, Sweden)
nanoXIM (Fluidinova, Portugal)
Neobone (Covalent Materials, Japan)
OssaBase-HA (Lasak, Czech Republic)
Ostegraf (Ceramed, CO, USA)
Ostim (Heraeus Kulzer, Germany)
Periograf (Cooke-Waite Laboratories, USA)
PermaOS (Mathys, Switzerland)
PurAtite (PremierBiomaterials, Ireland)
Synatite (SBM, France)
Synthacer (KARL STORZ Recon, Germany)
without trade name (Cam Bioceramics, Netherlands)
without trade name (CaP Biomaterials, WI, USA)
HA embedded in silica gelNanoBone (Artoss, Germany)
HA/collagenBioimplant (Connectbiopharm, Russia)
Bonject (Koken, Japan)
Collagraft (Zimmer and Collagen Corporation, USA)
CollapAn (Intermedapatite, Russia)
HAPCOL (Polystom, Russia)
LitAr (LitAr, Russia)
HA/sodium alginateBialgin (Biomed, Russia)
HA/poly-L-Lactic AcidSuperFIXSORB30 (Takiron, Japan)
HA/polyethyleneHAPEX (Gyrus, TN, USA)
HA/CaSO4Hapset (LifeCore, MIN, USA)
PerOssal (aap Implantate, Germany)
coralline HAInterpore (Interpore, CA, USA)
ProOsteon (Interpore, CA, USA)
algae-derived HAAlgipore (Dentsply Friadent, Germany)
bovine bone apatite (unsintered)BioOss (Geitslich, Switzerland)
Laddec (Ost-Developpement, France)
Lubboc (Ost-Developpement, France)
Oxbone (Bioland biomateriaux, France)
Tutoplast (Tutogen Medical, Germany)
bovine bone apatite (sintered)BonAP (unknown producer)
Cerabone (aap Implantate, Germany)
Endobon (Merck, Germany)
Navigraft (Zimmer Dental, USA)
Osteograf (Ceramed, CO, USA)
PepGen P-15 (Dentsply Friadent, Germany)
Pyrost (Osteo AG, Germany)
hyman bone allograftmaxgraft (botiss, Germany)
Osnatal (aap Implantate, Germany)
equineBioGen (unknown producer)
α-TCPBioBase (Zimmer, IN, USA)
without trade name (Cam Bioceramics, Netherlands)
without trade name (PremierBiomaterials, Ireland)
β-TCPadboneTCP (Medbone Medical Devices, Portugal)
Antartik TCP (MedicalBiomat, France)
Augment Bone Graft (BioMimetic Therapeutics, TN, USA)
BioGraft (IFGL BIO CERAMICS, India)
Bioresorb (Sybron Implant Solutions, Germany)
Biosorb (SBM S.A., France)
Bi-Ostetic (Berkeley Advanced Biomaterials, CA, USA)
Calc-i-oss classic (Degradable Solutions, Switzerland)
Calciresorb (Ceraver, France)
CELLPLEX (Wright Medical Technology, TN, USA)
Cerasorb (Curasan, Germany)
Ceros (Thommen Medical, Switzerland)
ChronOS (Synthes, PA, USA)
Conduit (DePuy Spine, USA)
GenerOs (Berkeley Advanced Biomaterials, CA, USA)
HT BIOCER (CHEMA–ELEKTROMET, Poland)
JAX (Smith and Nephew Orthopaedics, USA)
Osferion (Olympus Terumo Biomaterials, Japan)
OsSatura TCP (Integra Orthobiologics, CA, USA)
PORESORB-TCP (Lasak, Czech Republic)
SynthoGraft (Bicon, MA, USA)
Synthos (unknown producer)
Syntricer (KARL STORZ Recon, Germany)
Vitoss (Orthovita, PA, USA)
without trade name (Cam Bioceramics, Netherlands)
without trade name (CaP Biomaterials, WI, USA)
without trade name (Shanghai Bio-lu Biomaterials, China)
β-TCP/collagenIntegra Mozaik (Integra Orthobiologics, CA, USA)
BCP (HA + β-TCP)4Bone (MIS, Israel)
adboneBCP (Medbone Medical Devices, Portugal)
Antartik Genta (MedicalBiomat, France)
Artosal (aap Implantate, Germany)
BCP BiCalPhos (Medtronic, MN, USA)
BioGraft (IFGL BIO CERAMICS, India)
Biosel (Depuy Bioland, France)
BonaGraft (Biotech One, Taiwan)
BoneCeramic (Straumann, Switzerland)
BoneSave (Stryker Orthopaedics, NJ, USA)
Calcicoat (Zimmer, IN, USA)
Calciresorb (Ceraver, France)
Calc-i-oss crystal (Degradable Solutions, Switzerland)
CellCeram (Scaffdex, Finland)
Ceraform (Teknimed, France)
Ceratite (NGK Spark Plug, Japan)
CuriOs (Progentix Orthobiology BV, Netherlands)
Eurocer (FH Orthopedics, France)
GenPhos HA TCP (Baumer, Brazil)
Graftys BCP (Graftys, France)
Hatric (Arthrex, Naples, FL, USA)
Hydros (Biomatlante, France)
Indost (Polystom, Russia)
Kainos (Signus, Germany)
MasterGraft Granules (Medtronic Sofamor Danek, TN, USA)
MBCP (Biomatlante, France)
OrthoCer HA TCP (Baumer, Brazil)
OpteMX (Exactech, FL, USA)
OsSatura BCP (Integra Orthobiologics, CA, USA)
ossceram nano (bredent medical, Germany)
Osteosynt (Einco, Brazil)
Ostilit (Stryker Orthopaedics, NJ, USA)
ReproBone (Ceramisys, UK)
SBS (Expanscience, France)
Scaffdex (Scaffdex Oy, Finland)
TCH (Kasios, France)
Triosite (Zimmer, IN, USA)
Tribone (Stryker, Europe)
without trade name (Cam Bioceramics, Netherlands)
without trade name (CaP Biomaterials, WI, USA)
BCP (HA + α-TCP)Skelite (Millennium Biologix, ON, Canada)
BCP (HA + β-TCP)/collagenAllograft (Zimmer, IN, USA)
Collagraft (Zimmer, IN, USA)
BCP/fibrinTricOS (Baxter BioScience, France)
BCP/siliconFlexHA (Xomed, FL, USA)
FAwithout trade name (CaP Biomaterials, WI, USA)
FA + BCP (HA + β-TCP)FtAP (Polystom, Russia)
carbonateapatiteHealos (Orquest, CA, USA)
SRS (Norian, CA, USA)
In spite of the aforementioned serious mechanical limitations (see Section 4.1), bioceramics of calcium orthophosphates is available in various physical forms: powders, particles, granules (or granulates [12]), dense blocks, porous scaffolds, self-setting formulations, implant coatings and composite component of different origin (natural, biological or synthetic) often with the specific shapes, such as implants, prostheses or prosthetic devices. Furthermore, bone grafts are also proposed as non-hardening injectable formulations [99,616,617,618,619] and pastes [620,621,622]. Generally, they consist of a mixture of calcium orthophosphate powders or granules and a “glue”, which can be a highly viscous hydrogel [99,619]. More to the point, custom-designed shapes like wedges for tibial opening osteotomy, cones for spine and knee and inserts for vertebral cage fusion are also available [565]. Various trademarks of the commercially available types of calcium orthophosphate-based bioceramics and biomaterials have been summarized in Table 3, while their surgical applications are schematically shown in Figure 10 [623]. A long list of both trademarks and producers clearly demonstrates that calcium orthophosphate bioceramics is easy to produce and not very difficult to register for the biomedical applications.
Figure 10. Different types of biomedical applications of calcium orthophosphate bioceramics. Reprinted from Reference [623] with permission.
Figure 10. Different types of biomedical applications of calcium orthophosphate bioceramics. Reprinted from Reference [623] with permission.
Materials 06 03840 g010
One should note, that among the existing calcium orthophosphates (Table 1), only certain compounds are useful for biomedical applications, because those having the Ca/P ionic ratio less than 1 are not suitable for implantation due to their high solubility and acidity. Furthermore, due to its basicity, TTCP alone is not suitable either. However, to be used in medicine, these “unsuitable” calcium orthophosphates might be successfully combined with either other calcium orthophosphates or other chemicals.

5.1. Self-Setting (Self-Hardening) Formulations

The need for bioceramics for minimal invasive surgery has induced a concept of self-setting (or self-hardening) formulations consisting of calcium orthophosphates only to be applied as injectable and/or mouldable bone substitutes [122,123,145,480,514,624,625,626,627,628]. In addition, there are reinforced formulations, which, in a certain sense, might be defined as calcium orthophosphate concretes [122]. Furthermore, self-setting formulations able to form porous bioceramics are also available [459,460,467,468,470,479,480,514,543,625,626,627,628].
All types of the self-setting calcium orthophosphate formulations belong to a low temperature bioceramics. They are divided into two major groups. The first one is a dry mixture of two different calcium orthophosphates (a basic one and an acidic one), in which, after being wetted, the setting reaction occurs according to an acid-base reaction. The second group contains only one calcium orthophosphate, such as ACP with Ca/P molar ratio within 1.50–1.67 or α-TCP: both of them form CDHA upon contact with an aqueous solution [122,145]. Chemically, setting (=hardening, curing) is due to the succession of dissolution and precipitation reactions. Mechanically, it results from crystal entanglement and intergrowth (Figure 11) [629]. Despite a large number of the initial compositions, all types of self-setting calcium orthophosphate formulations can form two different products only: CDHA and DCPD [122,123,145,480,514,624,625,626,627,628]. Special reviews on the subject are available [122], where the interested readers are referred for further details.
Figure 11. A typical microstructure of calcium orthophosphate cement after hardening. The mechanical stability is provided by the physical entanglement of crystals. Reprinted from Reference [629] with permission.
Figure 11. A typical microstructure of calcium orthophosphate cement after hardening. The mechanical stability is provided by the physical entanglement of crystals. Reprinted from Reference [629] with permission.
Materials 06 03840 g011

5.2. Coatings, Films and Layers

For many years, the clinical application of calcium orthophosphate-based bioceramics has been largely limited to non-load bearing parts of the skeleton due to their inferior mechanical properties. Therefore, materials with better mechanical properties appear to be necessary. For example, metallic implants are encountered in endoprostheses (total hip joint replacements) and artificial teeth sockets. As metals do not undergo bone bonding, i.e., do not form a mechanically stable link between the implant and bone tissue, ways have been sought to improve contacts at the interface. The major way is to coat metals with calcium orthophosphates those exhibit a bone-bonding ability between the metal and bone [194,206,385,630,631,632,633,634].
A number of factors influence the properties of calcium orthophosphate coatings, films and layers. They include coating thickness (this will influence coating adhesion and fixation—the agreed optimum now seems to be within 50–100 µm), crystallinity (this affects the dissolution and biological behavior), phase and chemical purity, porosity and adhesion. The coated implants combine the surface biocompatibility and bioactivity of calcium orthophosphates with the core strength of strong substrates (Figure 12). Moreover, calcium orthophosphate coatings, films and layers decrease a release of potentially hazardous chemicals from the core implant and shield the substrate surface from environmental attack. In the case of porous implants, the coated by calcium orthophosphates surface enhances bone ingrowth into the pores [248]. The production techniques of the coatings, films and layers and their properties are now largely standardized. The clinical results for calcium orthophosphate-coated implants reveal that they have much longer life times after implantation than uncoated devices and they have been found to be particularly beneficial for younger patients. Further details on this topic might be found in special reviews [529,630,631,632,633,634].
Figure 12. Shows how a plasma-sprayed HA coating on a porous titanium (dark bars) dependent on the implantation time will improve the interfacial bond strength compared to uncoated porous titanium (light bars). Reprinted from Reference [66] with permission.
Figure 12. Shows how a plasma-sprayed HA coating on a porous titanium (dark bars) dependent on the implantation time will improve the interfacial bond strength compared to uncoated porous titanium (light bars). Reprinted from Reference [66] with permission.
Materials 06 03840 g012

5.3. Functionally Graded Bioceramics

In general, functionally gradient materials (FGMs) are defined as the materials, having gradient either compositional or structural changes from their surface to the interior of the materials. The idea of FGMs allows one device to possess two different properties. One of the most important combinations for the biomedical field is that of a mechanical strength and biocompatibility. Namely, only surface properties govern a biocompatibility of the entire device. In contrast, the strongest material determines the mechanical strength of the entire device. Although, this subject belongs to the previous section on coatings, films and layers, in a certain sense, all types of implants covered by calcium orthophosphates might be also considered as a FGM.
Within the scope of this review, functionally graded bioceramics consisting of calcium orthophosphates solely is considered and discussed. Such formulations have been developed [92,318,502,505,575,635,636,637,638,639,640,641,642,643,644]. For example, dense sintered bodies with gradual compositional changes from α-TCP to HA were prepared by sintering a diamond-coated HA compacts at 1280 °C under a reduced pressure, followed by heating under the atmospheric conditions [635]. The content of α-TCP gradually decreased, while the content of HA increased with increasing depth from the surface. This functionally gradient bioceramics consisting of HA core and α-TCP surface showed a potential value as bone-substituting biomaterials [635]. Two types of functionally gradient FA/β-TCP biocomposites were prepared in another study [636]. As shown in Figure 13, one of the graded biocomposites was in the shape of a disk and contained four different layers of about 1 mm thick. The other graded biocomposite was also in the shape of a disk but contained two sets of the four layers, each layer being 0.5 mm thick controlled by using a certain amount of the mixed powders. The final FA/β-TCP graded structures were formed at 100 MPa and sintered at 1300 °C for 2 h [636]. Calcium orthophosphates coatings with graded crystallinity were prepared as well [642].
Figure 13. A schematic diagram showing the arrangement of the FA/β-TCP biocomposite layers. (a) A non-symmetric functionally gradient material (FGM); and (b) symmetric FGM. Reprinted from Reference [636] with permission.
Figure 13. A schematic diagram showing the arrangement of the FA/β-TCP biocomposite layers. (a) A non-symmetric functionally gradient material (FGM); and (b) symmetric FGM. Reprinted from Reference [636] with permission.
Materials 06 03840 g013
Besides, it is well known that a bone cross-section from cancellous to cortical bone is non-uniform in porosity and pore dimensions. Thus, in various attempts to mimic the porous structure of bones, calcium orthophosphate bioceramics with graded porosity have been fabricated [92,429,502,505,575,635,636,637,638,639,640,641]. For example, graded porous calcium orthophosphate bioceramics can be produced by means of tape casting and lamination (Figure 14a). Other manufacturing techniques, such as a compression molding process (Figure 14b) followed by impregnation and firing, are known as well [429]. In the first method, a HA slurry was mixed with a pore former. The mixed slurry was then cast into a tape. Using the same method, different tapes with different pore former sizes were prepared individually. The different tape layers were then laminated together. Firing was then done to remove the pore formers and sinter the HA particle compacts, resulting in graded porous bioceramics [639]. This method was also used to prepare graded porous HA with a dense part (core or layer) in order to improve the mechanical strength, as dense ceramics are much stronger than porous ceramics. However, as in the pressure infiltration of mixed particles, this multiple tape casting also has the problem of poor connectivity of pores, although the pore size and the porosity are relatively easy to control. Furthermore, the lamination step also introduces additional discontinuity of the porosity on the interfaces between the stacked layers.
Figure 14. Schematic illustrations of fabrication of pore-graded bioceramics: (a) lamination of individual tapes, manufactured by tape casting; and (b) a compression molding process. Reprinted from Reference [429] with permission.
Figure 14. Schematic illustrations of fabrication of pore-graded bioceramics: (a) lamination of individual tapes, manufactured by tape casting; and (b) a compression molding process. Reprinted from Reference [429] with permission.
Materials 06 03840 g014
Since diverse biomedical applications require different configurations and shapes, the graded (or gradient) porous bioceramics can be grouped according to both the overall shape and the structural configuration [429]. The basic shapes include rectangular blocks and cylinders (or disks). For the cylindrical shape, there are configurations of dense core—porous layer, less porous core—more porous layer, dense layer—porous core and less porous layer—more porous core. For the rectangular shape, in the gradient direction i.e., the direction with varying porosity, pore size or composition, there are configurations of porous top—dense bottom (same as porous bottom—dense top), porous top—dense center—porous bottom, dense top—porous center—dense bottom, etc. Concerning biomedical applications, a dense core—porous layer structure is suitable for implants of a high mechanical strength and with bone ingrowth for stabilization, whereas a less porous layer—more porous core configuration can be used for drug delivery systems. Furthermore, a porous top—dense bottom structure can be shaped into implants of articulate surfaces for wear resistance and with porous ends for bone ingrowth fixation; while a dense top—porous center—dense bottom arrangement mimics the structure of head skull. Further details on bioceramics with graded porosity might be found in literature [429].

6. Biological Properties and in Vivo Behavior

The most important differences between bioactive bioceramics and all other implanted materials comprise inclusion in the metabolic processes of the organism, adaptation of either surface or the entire material to the biomedium, integration of a bioactive implant with bone tissues at the molecular level or the complete replacement of a resorbable bioceramics by healthy bone tissues. All of the enumerated processes are related to the effect of an organism on the implant. Nevertheless, another aspect of implantation is also important—the effect of the implant on the organism. For example, using of bone implants from corpses or animals, even after they have been treated in various ways, provokes a substantially negative immune reactions in the organism, which substantially limits the application of such implants. In this connection, it is useful to dwell on the biological properties of bioceramic implants, particularly those of calcium orthophosphates, which in the course of time may be resorbed completely [645].

6.1. Interactions with Surrounding Tissues and the Host Responses

All interactions between implants and the surrounding tissues are dynamic processes. Water, dissolved ions, various biomolecules and cells surround the implant surface within initial few seconds after the implantation. It has been accepted that no foreign material placed inside a living body is completely compatible. The only substances that conform completely are those manufactured by the body itself (autogenous), while any other substance, which is recognized as foreign, initiates some types of reactions (a host-tissue response). The reactions occurring at the biomaterial/tissue interfaces lead to time-dependent changes in the surface characteristics of both the implanted biomaterials and the surrounding tissues [76,646].
In order to develop new biomaterials, it is necessary to understand the in vivo host responses. Like any other species, biomaterials and bioceramics react chemically with their environment and, ideally, they should neither induce any changes nor provoke undesired reactions in the neighboring or distant tissues. In general, living organisms can treat artificial implants as biotoxic (or bioincompatible [71]), bioinert (or biostable [62]), biotolerant (or biocompatible [71]), bioactive and bioresorbable materials [1,2,3,58,59,63,66,67,68,69,70,71,645,646,647]. Biotoxic (e.g., alloys containing cadmium, vanadium, lead and other toxic elements) materials release to the body substances in toxic concentrations and/or trigger the formation of antigens that may cause immune reactions ranging from simple allergies to inflammation to septic rejection with the associated severe health consequences. They cause atrophy, pathological change or rejection of living tissue near the material as a result of chemical, galvanic or other processes. Bioinert (this term should be used with care, since it is clear that any material introduced into the physiological environment will induce a response. However, for the purposes of biomedical implants, the term can be defined as a minimal level of response from the host tissue), such as zirconia, alumina, carbon and titanium, as well as biotolerant (e.g., polymethylmethacrylate, titanium and Co-Cr alloy) materials do not release any toxic constituents but also do not show positive interaction with living tissue. They evoke a physiological response to form a fibrous capsule, thus, isolating the material from the body. In such cases, thickness of the layer of fibrous tissue separating the material from other tissues of an organism can serve as a measure of bioinertness. Generally, both bioactivity and bioresorbability phenomena are fine examples of chemical reactivity and calcium orthophosphates (both non-substituted and ion-substituted ones) fall into these two categories of bioceramics [1,2,3,58,59,63,66,67,68,69,70,71,645,646,647]. A bioactive material will dissolve slightly but promote formation of a surface layer of biological apatite before interfacing directly with the tissue at the atomic level, that result in formation of a direct chemical bonds to bones (see Section 6.4. for details). Such implants provide a good stabilization for materials that are subject to mechanical loading. A bioresorbable material will dissolve over time (regardless of the mechanism leading to the material removal) and allow a newly formed tissue to grow into any surface irregularities but may not necessarily interface directly with the material. Consequently, the functions of bioresorbable materials are to participate in dynamic processes of formation and re-absorption occurring in bone tissues; thus, bioresorbable materials are used as scaffolds or filling spacers allowing to the tissues their infiltration and substitution [67,194,324,648,649,650].
It is important to stress, that a distinction between the bioactive and bioresorbable bioceramics might be associated with structural factors only. Namely, bioceramics made from non-porous, dense and highly crystalline HA behaves as a bioinert (but a bioactive) material and is retained in an organism for at least 5–7 years without noticeable changes (Figure 2 bottom), while a highly porous bioceramics of the same composition can be resorbed approximately within a year. Furthermore, submicron-sized HA powders are biodegraded even faster than the highly porous HA scaffolds. Other examples of bioresorbable materials include porous bioceramic scaffolds made of biphasic, triphasic or multiphasic formulations (see Section 3.2) or bone grafts (dense or porous) made of CDHA [141], TCP [92,651,652] and/or ACP [481,653]. One must stress that recently the concepts of bioactive and bioresorbable materials have converged and bioactive materials are made bioresorbable, while bioresorbable materials are made bioactive [654].
Although in certain in vivo experiments inflammatory reactions were observed after implantation of calcium orthophosphate bioceramics [655,656,657,658,659,660,661,662], the general conclusion on using calcium orthophosphates with Ca/P ionic ratio within 1.0–1.7 is that all types of implants (bioceramics of various porosities and structures, powders or granules) are not only nontoxic but also induce neither inflammatory nor foreign-body reactions [128,663,664]. The biological response to implanted calcium orthophosphates follows a similar cascade observed in fracture healing. This cascade includes a hematoma formation, inflammation, neovascularization, osteoclastic resorption and a new bone formation. An intermediate layer of fibrous tissue between the implants and bones has been never detected. Furthermore, calcium orthophosphate implants display the ability to directly bond to bones [1,2,3,58,59,63,66,67,68,69,70,71,645,646,647]. For further details, the interested readers are referred to a good review on cellular perspectives of bioceramic scaffolds for bone tissue engineering [491].
One should note that the aforementioned rare cases of the inflammatory reactions to calcium orthophosphate bioceramics were often caused by “other” reasons. For example, a high rate of wound inflammation occurred when highly porous HA was used. In that particular case, the inflammation was explained by sharp implant edges, which irritated surrounding soft tissues [656]. To avoid this, only rounded material should be used for implantation (Figure 15) [665]. Another reason for inflammation produced by porous HA could be due to micro movements of the implants, leading to simultaneous disruption of a large number of micro-vessels, which grow into the pores of the bioceramics. This would immediately produce an inflammatory reaction. Additionally, problems could arise in clinical tests connected with migration of granules used for alveolar ridge augmentation, because it might be difficult to achieve a mechanical stability of implants at the implantation sites [656]. Besides, presence of calcium pyrophosphate impurity might be the reason of inflammation [659]. Additional details on inflammatory cell responses to calcium orthophosphates might be found in a special review on this topic [660].
Figure 15. Rounded β-TCP granules of 2.6–4.8 mm in size, providing no sharp edges for combination with bone cement. Reprinted from Reference [665] with permission.
Figure 15. Rounded β-TCP granules of 2.6–4.8 mm in size, providing no sharp edges for combination with bone cement. Reprinted from Reference [665] with permission.
Materials 06 03840 g015

6.2. Osteoinduction

Before recently, it was generally considered, that alone, any type of synthetic bioceramics possessed neither osteogenic (osteogenesis is the process of laying down new bone material by osteoblasts) nor osteoinductive (is the property of the material to induce bone formation de novo or ectopically (i.e., in non-bone forming sites)) properties and demonstrated a minimal immediate structural support. However, a number of reports have already shown the osteoinductive properties of certain types of calcium orthophosphate bioceramics [171,563,592,666,667,668,669,670,671,672,673,674,675,676,677,678,679,680,681,682,683,684,685] and the amount of such publications rapidly increases. For example, bone formation was found to occur in dog muscle inside porous calcium orthophosphates with surface microporosity, while bone was not observed on the surface of dense bioceramics [667]. Furthermore, implantation of porous β-TCP bioceramics appeared to induce bone formation in soft tissues of dogs, while no bone formation was detected in any α-TCP implants [668]. More to the point, titanium implants coated by a microporous layer of OCP were found to induce ectopic bone formation in goat muscles, while a smooth layer of carbonated apatite on the same implants was not able to induce bone formation there [674,675]. In another study, β-TCP powder, biphasic (HA + β-TCP) powder and intact biphasic (HA + β-TCP) rods were implanted into leg muscles of mice and dorsal muscles of rabbits [683]. One month and three months after implantation, samples were harvested for biological and histological analysis. New bone tissues were observed in 10 of 10 samples for β-TCP powder, 3 of 10 samples biphasic powder and 9 of 10 samples for intact biphasic rods at 3rd month in mice, but not in rabbits. The authors concluded that the chemical composition was the prerequisite in osteoinduction, while porosity contributed to more bone formation [683]. Therefore, researchers have already discovered the ways to prepare osteoinductive calcium orthophosphate bioceramics.
Unfortunately, the underlying mechanism(s) leading to bone induction by synthetic materials remains largely unknown. Nevertheless, besides the specific genetic factors [681] and chosen animals [683], the dissolution/precipitation behavior of calcium orthophosphates [630], their microporosity [679,683,686,687], physicochemical properties [677,679], composition [683], the specific surface area [687], nanostructure [682], as well as the surface topography and geometry [673,688,689,690] have been pointed out as the relevant parameters. A positive effect of increased microporosity on the ectopic bone formation could be both direct and indirect. Firstly, an increased microporosity is directly related to the changes in surface topography, i.e., increases a surface roughness, which might affect the cellular differentiation. Secondly, an increased microporosity indirectly means a larger surface that is exposed to the body fluids leading to elevated dissolution/precipitation phenomena as compared to non-microporous surfaces. In addition, other hypotheses are also available. Namely, Reddi explained the apparent osteoinductive properties as an ability of particular bioceramics to concentrate bone growth factors, which are circulating in biological fluids, and those growth factors induce bone formation [688]. Other researchers proposed a similar hypothesis that the intrinsic osteoinduction by calcium orthophosphate bioceramics is a result of adsorption of osteoinductive substances on their surface [673]. Moreover, Ripamonti [689] and Kuboki et al. [690] independently postulated that the geometry of calcium orthophosphate bioceramics is a critical parameter in bone induction. Specifically, bone induction by calcium orthophosphates was never observed on flat bioceramic surfaces. All osteoinductive cases were observed on either porous structures or structures contained well-defined concavities. What’s more, bone formation was never observed on the peripheries of porous implants and was always found inside the pores or concavities, aligning the surface [194]. Some researchers speculated that a low oxygen tension in the central region of implants might provoke a dedifferentiation of pericytes from blood micro-vessels into osteoblasts [691]. Finally but yet importantly, both nano-structured rough surfaces and a surface charge on implants were found to cause an asymmetrical division of the stem cells into osteoblasts, which is important for osteoinduction [686].
Nevertheless, to finalize this topic, it is worth citing a conclusion made by Boyan and Schwartz [692] (p. 9): “Synthetic materials are presently used routinely as osteoconductive bone graft substitutes, but before purely synthetic materials can be used to treat bone defects in humans where an osteoinductive agent is required, a more complete appreciation of the biology of bone regeneration is needed. An understanding is needed of how synthetic materials modulate the migration, attachment, proliferation and differentiation of mesenchymal stem cells, how cells on the surface of a material affect other progenitor cells in the peri-implant tissue, how vascular progenitors can be recruited and a neovasculature maintained, and how remodeling of newly formed bone can be controlled.”

6.3. Biodegradation

Shortly after implantation, a healing process is initiated by compositional changes of the surrounding bio-fluids and adsorption of biomolecules. Following this, various types of cells reach the bioceramic surface and the adsorbed layer dictates the ways the cells respond. Further, a biodegradation of the implanted bioceramics begins. This process can occur by either physicochemical dissolution with a possibility of phase transformations or cellular activity (so called, bioresorption). More likely, a combination of both processes takes place in vivo. Since the existing calcium orthophosphates are differentiated by Ca/P ratio, basicity/acidity and solubility (Table 1), their degradation kinetics and mechanism depend on the chosen type of calcium orthophosphate [693,694]. Since dissolution is a physical chemistry process, it is controlled by some factors, such as solubility, surface area to volume ratio, local acidity, fluid convection and temperature. For HA and FA, the dissolution mechanism in acids has been described by a sequence of four successive chemical equations, in which several other calcium orthophosphates, such as TCP, DCPD/DCPA and MCPM/MCPA, appear as virtual intermediate phases [695,696].
With a few exceptions, dissolution rates of calcium orthophosphates are inversely proportional to the Ca/P ratio (except of TTCP), phase purity and crystalline size, as well as it is directly related to both the porosity and the surface area. In addition, phase transformations might occur with OCP, DCPA, DCPD, α-TCP, β-TCP and ACP because they are unstable in aqueous environment under the physiological conditions. Bioresorption is a biological process mediated by cells (mainly, osteoclasts and, in a lesser extent, macrophages) [697,698]. It depends on the response of cells to their environment. Osteoclasts attach firmly to the implant and dissolve calcium orthophosphates by secreting an enzyme carbonic anhydrase or any other acid, leading to a local pH drop to ~4–5 [699]. Furthermore, nanodimensional particles of calcium orthophosphate can also be phagocytosed by cells, i.e., they are incorporated into cytoplasm and thereafter dissolved by acid attack and/or enzymatic processes [700]. In any case, in vivo biodegradation of calcium orthophosphates is a complicated combination of various non-equilibrium processes, occurring simultaneously and/or in competition with each other.
Usually, an in vitro biodegradation of calcium orthophosphate bioceramics is simulated by suspending the material in a slightly acidic (pH ~4) buffer and monitoring the release of major ions with time [181,694,701,702,703]. The acidic buffer, to some extent, mimics the acidic environment during osteoclastic activity. In one study, an in vivo behavior of porous β-TCP bioceramics prepared from rod-shaped particles and that prepared from non-rod-shaped particles in the rabbit femur was compared. Although the porosities of both types of β-TCP bioceramics were almost the same, a more active osteogenesis was preserved in the region where rod-shaped bioceramics was implanted [704]. This result implied that the microstructure affected the activity of bone cells and subsequent bone replacement.
The experimental results demonstrated that both the dissolution kinetics and in vivo biodegradation of biologically relevant calcium orthophosphates proceed in the following decreasing order: β-TCP > bovine bone apatite (unsintered) > bovine bone apatite (sintered) > coralline HA > HA. In the case of biphasic (HA + TCP), triphasic and multiphasic calcium orthophosphates, the biodegradation kinetics depends on the HA/TCP ratio: the higher the ratio, the lower the degradation rate. Similarly, in vivo degradation rate of biphasic TCP (α-TCP + β-TCP) bioceramics appeared to be lower than that of α-TCP and higher than that of β-TCP bioceramics, respectively [112]. Furthermore, incorporation of doping ions can either increase (e.g., CO32−, Mg2+ or Sr2+) or decrease (e.g., F) the solubility (therefore, biodegradability) of CDHA and HA. Contrarily to apatites, solubility of β-TCP is decreased by incorporation of either Mg2+ or Zn2+ ions [563]. Here, one should remind that ion-substituted calcium orthophosphates are not considered in this review; the interested readers are advised to read the original publications [26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57].

6.4. Bioactivity

Generally, bioactive materials interact with surrounding bone resulting in formation of a chemical bond to this tissue (bone bonding). The bioactivity phenomenon is determined by both chemical factors, such as crystal phases and molecular structures of a biomaterial, and physical factors, such as surface roughness and porosity. Currently, it is agreed that the newly formed bone bonds directly to biomaterials through a carbonated CDHA layer precipitating at the bone/biomaterial interface. Strange enough but a careful seeking in the literature resulted in just a few publications [563,705,706,707], where the bioactivity mechanism of calcium orthophosphates was briefly described. For example, the chemical changes occurring after exposure of a synthetic HA bioceramics to both in vivo (implantation in human) and in vitro (cell culture) conditions were studied. A small amount of HA was phagocytozed but the major remaining part behaved as a secondary nucleator as evidenced by the appearance of a newly formed mineral [705]. In vivo, a cellular activity (e.g., of macrophages or osteoclasts) associated with an acidic environment were found to result in partial dissolution of calcium orthophosphates, causing liberation of calcium and orthophosphate ions to the microenvironment. The liberated ions increased a local supersaturation degree of the surrounding biologic fluids, causing precipitation of nano-sized crystals of biological apatite with simultaneous incorporating of various ions presented in the fluids. Infrared spectroscopic analyses demonstrated that these nanodimensional crystals were intimately associated with bioorganic components (probably proteins), which might also have originated from the biologic fluids, such as serum [563].
Therefore, one should consider the bioactivity mechanism of other biomaterials, particularly of bioactive glasses—the concept introduced by Prof. Larry L. Hench [66,67,68,69]. The bonding mechanism of bioactive glasses to living tissues involves a sequence of 11 successive reaction steps (Figure 16), some of which comprise calcium orthophosphates. The initial five steps occurred on the surface of bioactive glasses are “chemistry” only, while the remaining six steps belong to “biology” because the latter include colonization by osteoblasts, followed by proliferation and differentiation of the cells to form a new bone that had a mechanically strong bond to the implant surface. Therefore, in the case of bioactive glasses the border between “dead” and “alive” is postulated between stages 5 and 6. According to Hench, all bioactive materials “form a bone-like apatite layer on their surfaces in the living body and bond to bone through this apatite layer. The formation of bone-like apatite on artificial material is induced by functional groups, such as Si–OH (in the case of biological glasses), Ti–OH, Zr–OH, Nb–OH, Ta–OH, –COOH and –H2PO4 (in the case of other materials). These groups have specific structures revealing negatively charge and induce apatite formation via formations of an amorphous calcium compound, e.g., calcium silicate, calcium titanate and ACP” [66,67,68,69].
Figure 16. A sequence of interfacial reactions involved in forming a bond between tissue and bioactive ceramics. Reprinted from References [66,67,68,69] with permission.
Figure 16. A sequence of interfacial reactions involved in forming a bond between tissue and bioactive ceramics. Reprinted from References [66,67,68,69] with permission.
Materials 06 03840 g016
In addition, one should mention another set of 11 successive reaction steps for bonding mechanism of unspecified bioceramics, developed by Prof. Paul Ducheyne (Figure 17) [76]. One can see that the Ducheyne’s model is rather similar to that proposed by Hench; however, there are noticeable differences between them. For example, Ducheyne mentions on ion exchange and structural rearrangement at the bioceramic/tissue interface (stage 3), as well as on interdiffusion from the surface boundary layer into bioceramics (stage 4) and deposition with integration into the bioceramics (stage 7), which are absent in the Hench’s model. On the other hand, Hench describes six biological stages (stages 6–11), while Ducheyne describes only four ones (stages 8–11). Both models have been developed almost 2 decades ago and, to the best of my knowledge, remain unchanged since then. Presumably, both approaches have pro et contra of their own and, obviously, should be updated and/or revised. Furthermore, in literature there are at least two other descriptions of the biological and cellular events occurring at the bone/implant interface [708,709]. Unfortunately, both of them comprise lesser number of stages. In 2010, one more hypothesis has been proposed (Figure 18). For the first time, it describes reasonable surface transformations, happening with calcium orthophosphate bioceramics (in that case, HA) shortly after the implantation [707].
Figure 17. A schematic diagram representing the events, which take place at the interface between bioceramics and the surrounding biological environment: (1) dissolution of bioceramics; (2) precipitation from solution onto bioceramics; (3) ion exchange and structural rearrangement at the bioceramic/tissue interface; (4) interdiffusion from the surface boundary layer into the bioceramics; (5) solution-mediated effects on cellular activity; (6) deposition of either the mineral phase (a) or the organic phase (b) without integration into the bioceramic surface; (7) deposition with integration into the bioceramics; (8) chemotaxis to the bioceramic surface; (9) cell attachment and proliferation; (10) cell differentiation; and (11) extracellular matrix formation. All phenomena, collectively, lead to the gradual incorporation of a bioceramic implant into developing bone tissue. Reprinted from Reference [76] with permission.
Figure 17. A schematic diagram representing the events, which take place at the interface between bioceramics and the surrounding biological environment: (1) dissolution of bioceramics; (2) precipitation from solution onto bioceramics; (3) ion exchange and structural rearrangement at the bioceramic/tissue interface; (4) interdiffusion from the surface boundary layer into the bioceramics; (5) solution-mediated effects on cellular activity; (6) deposition of either the mineral phase (a) or the organic phase (b) without integration into the bioceramic surface; (7) deposition with integration into the bioceramics; (8) chemotaxis to the bioceramic surface; (9) cell attachment and proliferation; (10) cell differentiation; and (11) extracellular matrix formation. All phenomena, collectively, lead to the gradual incorporation of a bioceramic implant into developing bone tissue. Reprinted from Reference [76] with permission.
Materials 06 03840 g017
Figure 18. A schematic diagram representing the phenomena that occur on HA surface after implantation: (1) beginning of the implant procedure, where a solubilization of the HA surface starts; (2) continuation of the solubilization of the HA surface; (3) the equilibrium between the physiological solutions and the modified surface of HA has been achieved (changes in the surface composition of HA does not mean that a new phase of DCPA or DCPD forms on the surface); (4) adsorption of proteins and/or other bioorganic compounds; (5) cell adhesion; (6) cell proliferation; (7) beginning of a new bone formation; and (8) new bone has been formed. Reprinted from Reference [707] with permission.
Figure 18. A schematic diagram representing the phenomena that occur on HA surface after implantation: (1) beginning of the implant procedure, where a solubilization of the HA surface starts; (2) continuation of the solubilization of the HA surface; (3) the equilibrium between the physiological solutions and the modified surface of HA has been achieved (changes in the surface composition of HA does not mean that a new phase of DCPA or DCPD forms on the surface); (4) adsorption of proteins and/or other bioorganic compounds; (5) cell adhesion; (6) cell proliferation; (7) beginning of a new bone formation; and (8) new bone has been formed. Reprinted from Reference [707] with permission.
Materials 06 03840 g018
An important study on formation of calcium orthophosphate precipitates on various types of bioceramic surfaces in both simulated body fluid (SBF) and rabbit muscle sites was performed [710]. The bioceramics were sintered porous solids, including bioglass, glass-ceramics, α-TCP, β-TCP and HA. An ability to induce calcium orthophosphate precipitation was compared among these types of bioceramics. The following conclusions were made: (1) OCP formation ubiquitously occurred on all types of bioceramic surfaces both in vitro and in vivo, except on β-TCP; (2) Apatite formation did not occur on every type of bioceramic surface; it was less likely to occur on the surfaces of HA and α-TCP; (3) Precipitation of calcium orthophosphates on the bioceramic surfaces was more difficult in vivo than in vitro; (4) Differences in calcium orthophosphate precipitation among the bioceramic surfaces were less noticeable in vitro than that in vivo; and (5) β-TCP bioceramics showed a poor ability of calcium orthophosphate precipitation both in vitro and in vivo [710]. These findings clearly revealed that apatite formation in the physiological environments could not be confirmed as the common feature of bioceramics. Nevertheless, for want of anything better, currently the bioactivity mechanism of calcium orthophosphate bioceramics should be described by a reasonable combination of Figure 16, Figure 17 and Figure 18, e.g., by updating the Ducheyne’s and Hench’s models by three initial stages taken from Figure 18.
Interestingly that the bioactivity of HA bioceramics might be enhanced by a high-energy ion irradiation [711]. The effect was attributed to formation of a unique 3D macroporous apatite layer of decreased crystallinity and crystal size on the irradiated surfaces. Obviously, to get further insights into the bioactivity phenomenon, the atomic and molecular processes occurring at the bioceramic surface in aqueous solutions and their effects on the relevant reaction pathways of cells and tissues must be elucidated in more details.

6.5. Cellular Response

Fixation of any implants in the body is a complex dynamic process that remodels the interface between the implants and living tissues at all dimensional levels, from the molecular up to the cell and tissue morphology level, and at all time scales, from the first second up to several years after implantation. Immediately following the implantation, a space filled with biological fluids appears next to the implant surface. With time, cells are adsorbed at the implant surface that will give rise to their proliferation and differentiation towards bone cells, followed by revascularisation and eventual gap closing. Ideally, a strong bond is formed between the implants and surrounding tissues [71]. An interesting study on the interfacial interactions between calcined HA and substrates has been performed [712], where the interested readers are referred for further details.
The aforementioned paragraph clearly demonstrates an importance of studies on cellular responses to calcium orthophosphate bioceramics. Therefore, such investigations have been performed extensively for several decades [660,713,714,715,716,717,718,719,720,721,722,723,724]. For example, bioceramic discs made of 7 different calcium orthophosphates (TTCP, HA, carbonate apatite, β-TCP, α-TCP, OCP and DCPD) were incubated in osteoclastic cell cultures for 2 days. In all cases, similar cell morphologies and good cell viability were observed; hoverer, different levels of resorbability of various calcium orthophosphates were detected [716]. Similar results were found for fluoridated HA coatings [718]. Experiments performed with human osteoblasts revealed that nanostructured bioceramics prepared from nano-sized HA showed significant enhancement in mineralization compared to microstructured HA bioceramics [717]. In addition, the influence of lengths and surface areas of rod-shaped HA on cellular response was studied. Again, similar cell morphologies and good cell viability were observed; however, it was concluded that high surface area could increase cell-particle interaction [721]. Nevertheless, another study with cellular response to rod-shaped HA bioceramics, revealed that some types of crystals might trigger a severe inflammatory response [722]. In addition, calcium orthophosphate-based sealers appeared to show less cytotoxicity and inflammatory mediators compared with other sealers [719]. More examples are available in literature.
Cellular biodegradation of calcium orthophosphate bioceramics is known to depend on its phases. For example, a higher solubility of β-TCP was shown to prevent L-929 fibroblast cell adhesion, thereby leading to damage and rupture of the cells [725]. A mouse ectopic model study indicated the maximal bone growth for the 80:20 β-TCP:HA biphasic formulations preloaded with human mesenchymal stem cells when compared to other calcium orthophosphates [726]. The effects of substrate microstructure and crystallinity have been corroborated with an in vivo rabbit femur model, where rod-like crystalline β-TCP was reported to enhance osteogenesis when compared to non-rod like crystalline β-TCP [704]. Additionally, using a dog mandibular defect model, a higher bone formation on a scaffold surface coated by nanodimensional HA was observed when compared to that coated by a micro-dimensional HA [727]. Furthermore, studies revealed a stronger stress signaling response by osteoblast precursor cells in 3D scaffolds when compared to 2D surfaces [728].
Mesenchymal stem cells are one of the most attractive cellular lines for application as bone grafts [729,730]. Early investigations by Okumura et al., indicated an adhesion, proliferation and differentiation, which ultimately became new bone and integrated with porous HA bioceramics [714]. Later, a sustained co-culture of endothelial cells and osteoblasts on HA scaffolds for up to 6 weeks was demonstrated [731]. Furthermore, a release of factors by endothelial and osteoblast cells in co-culture supported proliferation and differentiation was suggested to ultimately result in microcapillary-like vessel formation and supported a neo-tissue growth within the scaffold [491]. More to the point, investigation of rat calvaria osteoblasts cultured on transparent HA bioceramics, as well as the analysis of osteogenic-induced human bone marrow stromal cells at different time points of culturing indicated to a good cytocompatibility of HA bioceramics and revealed favorable cell proliferation [421]. The positive results for other types of cells have been obtained in other studies [203,416,417,420,452,453,454,732,733,734].
Interestingly that HA scaffolds with marrow stromal cells in a perfused environment were reported to result in ~85% increase in mean core strength, a ~130% increase in failure energy and a ~355% increase in post-failure strength. The increase in mineral quantity and promotion of the uniform mineral distribution in that study was suggested to attribute to the perfusion effect [586]. Additionally, other investigators indicated to mechanical properties increasing for other calcium orthophosphate scaffolds after induced osteogenesis [585,588].
To finalize this section, one should mention on the newest developments to influence the cellular response. First, to facilitate interactions with cells, the calcium orthophosphate surface might be functionalized [735,736,737,738]. Second, it appears that crystals of biological apatite of calcified tissues exhibit different orientations depending on the tissue. Namely, in vertebrate bones and tooth enamel surfaces, the respective a,b-planes and c-planes of the apatite crystals are preferentially exposed. Therefore, ideally, this should be taken into account in artificial bone grafts. Recently, a novel process to fabricate dense HA bioceramics with highly preferred orientation to the a,b-plane was developed [739,740]. The results revealed that increasing the a,b-plane orientation degree shifted the surface charge from negative to positive and decreased the surface wettability with simultaneous decreasing of cell attachment efficiency [740].

7. Non-Biomedical Applications

Due to their strong adsorption ability, surface acidity or basicity and ion exchange abilities, both hydroxyapatite and other calcium orthophosphates appear to possess a catalytic activity [25,741,742,743,744,745,746,747,748,749,750,751,752]. As seen from the references, calcium orthophosphates are able to catalyze oxidation and reduction reactions, as well as formation of C–C bonds. Namely, the application in oxidation reactions mainly includes oxidation of alcohol and dehydrogenation of hydrocarbons, while the reduction reactions include hydrogenolysis and hydrogenation. The formation of C–C bonds mainly comprises Claisen-Schmidt and Knoevenagel condensation reactions, Michael addition reaction, as well as Friedel-Crafts, Heck, Diels-Alder and adol reactions [749].
In addition, due to the chemical similarity to the inorganic part of mammalian calcified tissues, both hydroxyapatite and other calcium orthophosphates appear to be good solid carriers for chromatography of biological substances. Namely, such high-value biological materials, as recombinant proteins, therapeutic antibodies and nucleic acids are separated and purified [753,754,755,756,757,758]. However, since both subjects are almost irrelevant to bioceramics, they are not detailed further.

8. Calcium Orthophosphate Bioceramics in Tissue Engineering

8.1. Tissue Engineering

Tissue/organ repair has been the ultimate goal of surgery from ancient times to nowadays [74,75]. The repair has traditionally taken two major forms: tissue grafting followed by organ transplantation and alloplastic or synthetic material replacement. Both approaches, however, have limitations. Grafting requires second surgical sites with associated morbidity and is restricted by limited amounts of material, especially for organ replacement. Synthetic materials often integrate poorly with host tissue and fail over time due to wear and fatigue or adverse body response [759]. In addition, all modern orthopedic implants lack three of the most critical abilities of living tissues: (i) self-repairing; (ii) maintaining of blood supply; and (iii) self-modifying their structure and properties in response to external aspects such as a mechanical load [760]. Needless to mention, that bones not only possess all of these properties but, in addition, they are self-generating, hierarchical, multifunctional, nonlinear, composite and biodegradable; therefore, the ideal artificial bone grafts must possess similar properties [79].
The last decades have seen a surge in creative ideas and technologies developed to tackle the problem of repairing or replacing diseased and damaged tissues, leading to the emergence of a new field in healthcare technology now referred to as tissue engineering. This is an interdisciplinary field that exploits a combination of living cells, engineering materials and suitable biochemical factors in a variety of ways to improve, replace, restore, maintain or enhance living tissues and whole organs [761,762,763]. However, as two of three major components (namely, cells and biochemical factors) of the tissue engineering subject appear to be far beyond the scope of this review, the topic of tissue engineering is narrowed down to the engineering materials prepared from calcium orthophosphate bioceramics only.
Regeneration, rather than a repair, is the central goal of any tissue engineering strategy. Thus, tissue engineering has a potential to create tissues and organs de novo [762]. This field of science started more than two decades ago [764,765] and the famous publication by Langer and Vacanti [766] has greatly contributed to the promotion of tissue engineering research worldwide. The field of tissue engineering, particularly when applied to bone substitutes where tissues often function in a mechanically demanding environment [767,768,769], requires a collaboration of excellence in cell and molecular biology, biochemistry, material sciences, bioengineering and clinical research. For the success, it is necessary that researchers with expertise in one area have an appreciation of the knowledge and challenges of the other areas. However, since the technical, regulatory and commercial challenges might be substantial, the introduction of new products is likely to be slow [762].
Nowadays, tissue engineering is at full research potential due to the following key advantages: (i) the solutions it provides are long-term, much safer than other options and cost-effective as well; (ii) the need for a donor tissue is minimal, which eliminates the immuno-suppression problems; and(iii) the presence of residual foreign material is eliminated as well [770,771].

8.2. Scaffolds and Their Properties

It would be very convenient to both patients and physicians if devastated tissues or organs of patients can be regenerated by simple cell injections to the target sites but such cases are rare. The majority of large-sized tissues and organs with distinct 3D form require a support for their formation from cells. The support is called scaffold, template and/or artificial extracellular matrix [148,149,553,764,767,768,769,770,771,772,773,774,775]. The major function of scaffolds is similar to that of the natural extracellular matrix that assists proliferation, differentiation and biosynthesis of cells. In addition, scaffolds placed at the regeneration sites will prevent disturbing cells from invasion into the sites of action [776,777]. The role of scaffolds has been perfectly described by Andrés Segovia (1893–1987), a Spanish classical guitarist: “When one puts up a building one makes an elaborate scaffold to get everything into its proper place. But when one takes the scaffold down, the building must stand by itself with no trace of the means by which it was erected. That is how a musician should work.”
Therefore, the idea behind tissue engineering is to create or engineer autografts by either expanding autologous cells in vitro guided by a scaffold or implanting an acellular template in vivo and allowing the patient’s cells to repair the tissue guided by the scaffold. The first phase is the in vitro formation of a tissue construct by placing the chosen cells and scaffolds in a metabolically and mechanically supportive environment with growth media (in a bioreactor), in which the cells proliferate and elaborate extracellular matrix. It is expected that cells infiltrate into the porous matrix and consequently proliferate and differentiate therein [778,779]. In the second phase, the construct is implanted in the appropriate anatomic location, where remodeling in vivo is intended to recapitulate the normal functional architecture of an organ or a tissue [780,781]. The key processes occurring during both in vitro and in vivo phases of the tissue formation and maturation are: (1) cell proliferation, sorting and differentiation; (2) extracellular matrix production and organization; (3) biodegradation of the scaffold; and (4) remodeling and potentially growth of the tissue [782].
To achieve the goal of tissue reconstruction, the scaffolds must meet several specific requirements [148,149,772]. A reasonable surface roughness is necessary to facilitate cell seeding and fixation [783,784,785,786,787]. A sufficient mechanical strength and stiffness are mandatory to oppose contraction forces and later for the remodeling of damaged tissues [788,789]. A high porosity and an adequate pore dimensions (Table 2 and Table 4) are very important to allow cell migration, vascularization, as well as a diffusion of nutrients [440]. A French architect Robert le Ricolais (1894–1977) stated: “The art of structure is where to put the holes”. Therefore, to enable proper tissue ingrowth, vascularization and nutrient delivery, scaffolds should have a highly interconnected porous network, formed by a combination of macro- and micropores, in which more than ~60% of the pores should have a size ranging from ~150 μm to ~400 μm and at least ~20% should be smaller than ~20 μm [28,440,451,452,456,549,550,551,555,557,563,589,590,591,592,593,594,595,669,759,790,791,792,793,794,795,796,797,798,799,800]. In addition, scaffolds must be manufactured from the materials with controlled biodegradability and/or bioresorbability, such as calcium orthophosphates, so that a new bone will eventually replace the scaffold [767,794,801]. Furthermore, the degradation by-products of scaffolds must be non-cytotoxic. More to the point, the resorption rate has to coincide as much as possible with the rate of bone formation (i.e., between a few months and about 2 years) [802]. This means that while cells are fabricating their own natural matrix structure around themselves, the scaffold is able to provide a structural integrity within the body and eventually it will break down leaving the newly formed tissue that will take over the mechanical load. However, one should bear in mind that the scaffold’s architecture changes with the degradation process and the degradation by-products affect the biological response. Besides, scaffolds should be easily fabricated into a variety of shapes and sizes [803] and be malleable to fit irregularly shaped defects. In many cases, ease of processability, as well as easiness of conformation and injectability, such as self-setting calcium orthophosphate formulations possess (see Section 5.1), can determine the choice of a certain biomaterial. Finally, sterilization with no loss of properties is a crucial step in scaffold production at both a laboratory and an industrial level [767,768,769]. Thus, each scaffold should fulfill many functions before, during and after implantation.
Table 4. A hierarchical pore size distribution that an ideal scaffold should exhibit [28].
Table 4. A hierarchical pore size distribution that an ideal scaffold should exhibit [28].
NumberPore sizes of a 3D scaffoldA biochemical effect or function
1<1 μmInteraction with proteins
Responsible for bioactivity
21–20 μmType of cells attracted
Cellular development
Orientation and directionality of cellular ingrowth
3100–1,000 μmCellular growth
Bone ingrowth
Predominant function in the mechanical strength
4>1,000 μmImplant functionality
Implant shape
Implant esthetics
Many fabrication techniques are available to produce porous calcium orthophosphate scaffolds (Table 2) with varying architectural features (for details, see Section 3.3). In order to achieve the desired properties at minimum expenses, the production process should be optimized [804]. With the advent of tissue engineering, the search is on for the ultimate option—a “tissue engineered bone substitute”, consisting of a synthetic calcium orthophosphate scaffold impregnated with cells and growth factors. Figure 19 schematically depicts a possible fabrication process of such item that, afterwards, will be implanted into a living organism to induce bone regeneration [62].
Figure 19. A schematic view of a third generation biomaterial, in which porous calcium orthophosphate bioceramics acts as a scaffold or a template for cells, growth factors, etc. Reprinted from Reference [62] with permission.
Figure 19. A schematic view of a third generation biomaterial, in which porous calcium orthophosphate bioceramics acts as a scaffold or a template for cells, growth factors, etc. Reprinted from Reference [62] with permission.
Materials 06 03840 g019
To finalize this topic, one should mention on fundamental unfeasibility to create so-called “ideal scaffold” for bone grafting. For instance, bones of human skeleton have very different dimensions, shapes and structures depending on their functions and locations. Therefore, synthetic bone grafts of various sizes, shapes, porosity, mechanical strength, composition and resorbability appear to be necessary. Thus, HA bioceramics of 0% to 15% porosity is used as both ilium and intervertebral spacers, where a high strength is required, HA bioceramics of 30% to 40% porosity is useful as spinous process spacer for laminoplasty, where both bone formation and middle strength are necessary, while HA bioceramics of 40% to 60% porosity is useful for the calvarias plate, where a fast bone formation is needed (Figure 20) [542]. Furthermore, defining the optimum parameters for artificial scaffolds is in fact an attempt to find a reasonable compromise between various conflicting functional requirements. Namely, an increased mechanical strength of bone substitutes requires solid and dense structures, while colonization of their surfaces by cells requires interconnected porosity. Additional details and arguments on this subject are well described in a recent publication [805] (p. 478), in which the authors concluded: “there is enough evidence to postulate that ideal scaffold architecture does not exist.”
Figure 20. A schematic drawing presenting the potential usage of HA with various degrees of porosity. Reprinted from Reference [542] with permission.
Figure 20. A schematic drawing presenting the potential usage of HA with various degrees of porosity. Reprinted from Reference [542] with permission.
Materials 06 03840 g020

8.3. Bioceramic Scaffolds from Calcium Orthophosphates

Philosophically, the increase in life expectancy requires biological solutions to all biomedical problems, including orthopedic ones, which were previously managed with mechanical solutions. Therefore, since the end of 1990s, the biomaterials research focuses on tissue regeneration instead of tissue replacement [806]. The alternatives include use hierarchical bioactive scaffolds to engineer in vitro living cellular constructs for transplantation or use bioresorbable bioactive particulates or porous networks to activate in vivo the mechanisms of tissue regeneration [807,808]. Thus, the aim of calcium orthophosphates is to prepare artificial porous bioceramic scaffolds able to provide the physical and chemical cues to guide cell seeding, differentiation and assembly into 3D tissues of a newly formed bone [463,727,809,810,811,812,813,814,815,816,817]. Particle sizes, shape and surface roughness of the scaffolds are known to affect cellular adhesion, proliferation and phenotype [783,784,785,786,787]. Additionally, the surface energy might play a role in attracting particular proteins to the bioceramic surface and, in turn, this will affect the cells affinity to the material. More to the point, cells are exceedingly sensitive to the chemical composition and their bone-forming functions can be dependent on grain morphology of the scaffolds. For example, osteoblast functions were found to increase on nanodimensional fibers if compared to nanodimensional spheres because the former more closely approximated the shape of biological apatite in bones [818]. Besides, a significantly higher osteoblast proliferation on HA bioceramics sintered at 1200 °C as compared to that on HA bioceramics sintered at 800 °C and 1000 °C was reported [819]. Furthermore, since ions of calcium and orthophosphate are known to regulate bone metabolism, calcium orthophosphates appear to be among the few bone graft substitute materials, which can be considered as a drug.
Thus, to meet the tissue engineering requirements, much attention is devoted to further improvements of calcium orthophosphate bioceramics [820,821]. From the chemical point of view, the developments include synthesis of novel ion-substituted calcium orthophosphates [26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57]. From the material point of view, the major research topics include nanodimensional and nanocrystalline structures [822,823,824], amorphous compounds [825], organic-inorganic biocomposites and hybrid biomaterials [353], biphasic, triphasic and multiphasic formulations [121], as well as various types of structures, forms and shapes. The latter comprise fibers, whiskers and filaments [241,826,827,828,829,830,831,832,833,834,835,836,837,838,839], macro-, micro- and nano-sized spheres, beads and granules [839,840,841,842,843,844,845,846,847,848,849,850,851,852,853,854,855,856], micro- and nano-sized tubes [857,858,859,860], porous 3D scaffolds made of ACP [481,653,861], TCP [86,89,161,528,529,862], HA [167,455,456,498,530,531,532,804,863,864,865,866,867,868,869] and biphasic formulations [251,489,502,552,814,850,870,871,872,873,874], structures with graded porosity [92,429,502,505,575,635,636,637,638,639,640,641] and hierarchically organized ones [875,876]. Furthermore, an addition of defects through an intensive milling [877,878] or their removal by a thermal treatment [879] can be used to modify a chemical reactivity of calcium orthophosphates. Besides, more attention should be paid to a crystallographically aligned calcium orthophosphate bioceramics [880].
There are three principal therapeutic strategies for treating diseased or injured tissues in patients: (i) implantation of freshly isolated or cultured cells; (ii) implantation of tissues assembled in vitro from cells and scaffolds; and (iii) in situ tissue regeneration. For cellular implantation, individual cells or small cellular aggregates from the patient or a donor are either injected into the damaged tissue directly or are combined with a degradable scaffold in vitro and then implanted. For tissue implantation, a complete 3D tissue is grown in vitro using patient or donor cells and a bioresorbable scaffold and then is implanted into the patients to replace diseased or damaged tissues. For in situ regeneration, a scaffold implanted directly into the injured tissue stimulates the body’s own cells to promote local tissue repair [327,761]. In any case, simply trapping cells at the particular point on a surface is not enough: the cells must be encouraged to differentiate, which is impossible without the presence of suitable biochemical factors [881]. All previously mentioned clearly indicates that for the purposes of tissue engineering, calcium orthophosphate bioceramics plays an auxiliary role; namely, it acts as a suitable material to manufacture the appropriate 3D templates, substrates or scaffolds to be colonized by living cells before the successive implantation [882,883,884]. The in vitro evaluation of potential calcium orthophosphate scaffolds for tissue engineering has been described elsewhere [885], while the data on the mechanical properties of calcium orthophosphate bioceramics for use in tissue engineering are also available [815,886,887]. The effect of a HA-based biomaterial on gene expression in osteoblast-like cells was reported as well [888]. To conclude this part, the excellent biocompatibility of calcium orthophosphate bioceramics, its possible osteoinductivity [171,563,592,666,667,668,669,670,671,672,673,674,675,676,677,678,679,680,681,682,683,684,685] and a high affinity for drugs [72,73,74,75,889,890], proteins and cells [891] make them very functional for the tissue engineering applications. The feasible production of scaffolds with tailored structures and properties opens up a spectacular future for calcium orthophosphates [888,889,890,891,892,893,894,895,896,897].

8.4. A Clinical Experience

To date, there are just a few publications on clinical application of cell-seeded calcium orthophosphate bioceramics for bone tissue engineering of humans. Namely, Quarto et al. [898] were the first to report a treatment of large (4–7 cm) bone defects of the tibia, ulna and humerus in three patients from 16 to 41 years old, where the conventional surgical therapies had failed. The authors implanted a custom-made unresorbable porous HA scaffolds seeded with in vitro expanded autologous bone marrow stromal cells. In all three patients, radiographs and computed tomographic scans revealed abundant callus formation along the implants and good integration at the interfaces with the host bones by the second month after surgery [898]. In the same year, Vacanti et al. [899] reported the case of a man who had a traumatic avulsion of the distal phalanx of a thumb. The phalanx was replaced with a specially treated natural coral (porous HA; 500-pore ProOsteon (see Table 3)) implant that was previously seeded with in vitro expanded autologous periosteal cells. The procedure resulted in the functional restoration of a stable and biomechanically sound thumb of normal length, without the pain and complications that are usually associated with harvesting a bone graft.
Morishita et al. [900] treated a defect resulting from surgery of benign bone tumors in three patients using HA scaffolds seeded with in vitro expanded autologous bone marrow stromal cells after osteogenic differentiation of the cells. Two bone defects in a tibia and one defect in a femur were treated. Although ectopic implants in nude mice were mentioned to show the osteogenicity of the cells, details such as the percentage of the implants containing bone and at what quantities were not reported. Furthermore, cell-seeded calcium orthophosphate scaffolds were found to be superior to autograft, allograft or cell-seeded allograft in terms of bone formation at ectopic implantation sites [901]. Besides, it has been hypothesized that dental follicle cells combined with β-TCP bioceramics might become a novel therapeutic strategy to restore periodontal defects [902]. In still another study, the behavior of human periodontal ligament stem cells on a HA-coated genipin-chitosan scaffold in vitro was studied followed by evaluation on bone repair in vivo [903]. The study demonstrated the potential of this formulation for bone regeneration.

9. Conclusions and Outlook

The available chronology of seeking for a suitable bioceramics for bone substitutes is as follows: since the 1950s, the first aim was to use bioinert bioceramics, which had no reaction with living tissues. They included inert and tolerant compounds, which were designed to withstand physiological stress without, however, stimulating any specific cellular responses. Later on, in the 1980s, the trend changed towards exactly the opposite: the idea was to implant bioceramics that reacted with the surrounding tissues by producing newly formed bone (a “responsive” bioceramics because it was able to elicit biological responses). These two stages have been referred to as the first and the second generations of bioceramics, respectively [904] and, currently, both of them have been extensively commercialized. Thus, the majority of the marketable products listed in Table 3 belong to the first and the second generations of bone substitute biomaterials. However, the progress keeps going and, in current century, scientists search for the third generation of bioceramics [327], which will be able to “instruct” the physiological environment toward desired biological responses (i.e., bioceramics will be able to regenerate bone tissues by stimulating specific responses at the molecular level) [60,62]. Since each generation represents an evolution on the requirements and properties of the biomaterials involved, one should stress that these three generations should not be interpreted as the chronological but the conceptual ones. This means that at present, research and development is still devoted to biomaterials and bioceramics that, according to their properties, could be considered to be of the first or the second generations, because the second generation of bioceramics with added porosity is one of the initial approaches in developing of the third generation of bioceramics [905]. Furthermore, there is another classification of the history of biomaterials introduced by Prof. James M. Anderson. According to Anderson, within 1950–1975 the researchers studied bioMATERIALS, within 1975–2000 they studied BIOMATERIALS and since 2000 the time for BIOmaterials has been coming [906]. Here, the capital letters emphasis the major direction of the research efforts in the complex subject of biomaterials. As bioceramics are biomaterials of the ceramic origin (see Section 2), the Anderson’s historical classification appears to be applicable to the bioceramics field as well.
The history development of bioceramics, one category of biomaterials, informs that the widespread use of biomaterials, however, experiences two major difficulties. The first is an incomplete understanding of the physical and chemical functioning of biomaterials and of the human response to these materials. Recent advances in material characterization and computer science, as well as in cell and molecular biology are expected to play a significant role in studies of biomaterials. A second difficulty is that many biomaterials do not perform as desirably as we would like. This is not surprising, since many materials used in medicine were not designed for medical purposes. It needs to be mentioned here that biomaterials are expected to perform in our body’s internal environment, which is very aggressive. For example, solution pH of body fluids in various tissues varies in the range from 1 to 9. During daily activities, bones are subjected to a stress of ~4 MPa, whereas the tendons and ligaments experience peak stresses in the range of 40–80 MPa. The mean load on a hip joint is up to three times body weight (3000 N) and peak load during jumping can be as high as ~10 times body weight. More importantly, these stresses are repetitive and fluctuating, depending on the activities, such as standing, sitting, jogging, stretching and climbing. All of these require careful designing of biomaterials in terms of composition, shape, physical and biocompatibility properties. Therefore, a significant challenge is the rational design of human biomaterials based on a systematic evaluation of desired biological, chemical and engineering requirements.
Nevertheless, the field of biomaterials is in the midst of a revolutionary change in which the life sciences are becoming equal in importance to materials science and engineering as the foundation of the field. Simultaneously, advances in engineering (for example nanotechnology) are greatly increasing the sophistication with which biomaterials are designed and have allowed fabrication of biomaterials with increasingly complex functions [907]. Specifically, during last ~40 years, calcium orthophosphate bioceramics has become an integral and vital segment of our modern health care delivery system. In the modern fields of the third generation bioceramics (Hench) or BIOceramics (Anderson), the full potential of calcium orthophosphates has only begun to be recognized. Namely, calcium orthophosphates, which were intended as osteoconductive bioceramics in the past, stand for materials to fabricate osteoinductive implants nowadays [171,563,592,666,667,668,669,670,671,672,673,674,675,676,677,678,679,680,681,682,683,684,685]. Some steps in this direction have been already made by fabricating scaffolds for bone tissue engineering through the design of controlled 3D-porous structures and increasing the biological activity through development of novel ion-substituted calcium orthophosphate bioceramics [28,565]. The future of biosynthetic bone implants will point to better mimicking the autologous bone grafts. Therefore, the composition, structure and molecular surface chemistry of various types of calcium orthophosphates will be tailored to match the specific biological and metabolic requirements of tissues or disease states [908,909]. This new generation of calcium orthophosphate bioceramics should enhance the quality of life of millions of people, as they grow older.
In spite of the great progress, there is still a great potential for major advances to be made in the field of calcium orthophosphate bioceramics [910]. This includes requirements for:
  • Improvement of the mechanical performance of existing types of bioceramics.
  • Enhanced bioactivity in terms of gene activation.
  • Improvement in the performance of biomedical coatings in terms of their mechanical stability and ability to deliver biological agents.
  • Development of smart biomaterials capable of combining sensing with bioactivity.
  • Development of improved biomimetic composites.
Furthermore, still there are needs for a better understanding of the biological systems. For example, the bonding mechanism between the bone mineral and collagen remains unclear. It is also unclear whether a rapid repair that is elicited by the new generation of bioceramics is a result of the enhancement of mineralization per se or whether there is a more complex signaling process involving proteins in collagen. If we were able to understand the fundamentals of bone response to specific ions and the signals they activate, then we would be able to design better bioceramics for the future [910].
To finalize this review, it is completely obvious that the present status of research and development in the field of calcium orthophosphate bioceramics is still at the starting point for the solution of new problems at the confluence of materials science, biology and medicine, concerned with the restoration of damaged functions in the human organisms. A large increase in active elderly people has dramatically raised the need for load-bearing bone graft substitutes, for example, for bone reconstruction during revision arthroplasty or for the reinforcement of osteoporotic bones. Strategies applied in the last four decades towards this goal have failed. So new strategies, possibly based on self-assembling and/or nanofabrication, will have to be proposed and developed [911]. Furthermore, in future, it should be feasible to design a new generation of gene-activating calcium orthophosphate based scaffolds tailored for specific patients and disease states. Perhaps, sometime bioactive stimuli will be used to activate genes in a preventative treatment to maintain the health of aging tissues. Currently this concept seems impossible. However, we need to remember that only ~40 years ago the concept of a material that would not be rejected by living tissues also seemed impossible [654].

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Comprehensive Biomaterials; Ducheyne, P.; Healy, K.; Hutmacher, D.E.; Grainger, D.W.; Kirkpatrick, C.J. (Eds.) Elsevier: Amsterdam, The Netherlands, 2011; p. 3672.
  2. Biomaterials Science: An Introduction to Materials in Medicine, 3rd ed.; Ratner, B.D.; Hoffman, A.S.; Schoen, F.J.; Lemons, J.E. (Eds.) Academic Press: Oxford, UK, 2013; p. 1573.
  3. Vallet-Regí, M. Ceramics for medical applications. J. Chem. Soc. Dalton Trans. 2001, 2, 97–108. [Google Scholar] [CrossRef]
  4. Dorozhkin, S.V. Calcium orthophosphates in dentistry. J. Mater. Sci. Mater. Med. 2013, 24, 1335–1363. [Google Scholar] [CrossRef] [PubMed]
  5. US Bone Grafts Market to Reach US$2.3 Billion by 2017, According to New Report by Global Industry Analysts, Inc. Available online: http://www.prweb.com/releases/bone_grafts/standard_bone_allografts/prweb8953883.htm (accessed on 3 September, 2013).
  6. Dorozhkin, S.V. Calcium Orthophosphates: Applications in Nature, Biology, and Medicine; Pan Stanford: 8 Temasek Boulevard, Singapore, 2012; p. 850. [Google Scholar]
  7. Oktar, F.N.; Kesenci, K.; Pişkin, E. Characterization of processed tooth hydroxyapatite for potential biomedical implant applications. Artif. Cells Blood Substit. Immobil. Biotechnol. 1999, 27, 367–379. [Google Scholar] [CrossRef] [PubMed]
  8. Lee, S.J.; Oh, S.H. Fabrication of calcium phosphate bioceramics by using eggshell and phosphoric acid. Mater. Lett. 2003, 57, 4570–4574. [Google Scholar] [CrossRef]
  9. Murugan, R.; Ramakrishna, S. Crystallographic study of hydroxyapatite bioceramics derived from various sources. Cryst. Growth Des. 2005, 5, 111–112. [Google Scholar] [CrossRef]
  10. Balazsi, C.; Weber, F.; Kover, Z.; Horvath, E.; Nemeth, C. Preparation of calcium-phosphate bioceramics from natural resources. J. Eur. Ceram. Soc. 2007, 27, 1601–1606. [Google Scholar] [CrossRef]
  11. Ooi, C.Y.; Hamdi, M.; Ramesh, S. Properties of hydroxyapatite produced by annealing of bovine bone. Ceram. Int. 2007, 33, 1171–1177. [Google Scholar] [CrossRef]
  12. Lee, S.J.; Lee, Y.C.; Yoon, Y.S. Characteristics of calcium phosphate powders synthesized from cuttlefish bone and phosphoric acid. J. Ceram. Process. Res. 2007, 8, 427–430. [Google Scholar]
  13. Oktar, F.N. Microstructure and mechanical properties of sintered enamel hydroxyapatite. Ceram. Int. 2007, 33, 1309–1314. [Google Scholar] [CrossRef]
  14. Ruksudjarit, A.; Pengpat, K.; Rujijanagul, G.; Tunkasiri, T. Synthesis and characterization of nanocrystalline hydroxyapatite from natural bovine bone. Curr. Appl. Phys. 2008, 8, 270–272. [Google Scholar] [CrossRef]
  15. Figueiredo, M.; Fernando, A.; Martins, G.; Freitas, J.; Judas, F.; Figueiredo, H. Effect of the calcination temperature on the composition and microstructure of hydroxyapatite derived from human and animal bone. Ceram. Int. 2010, 36, 2383–2393. [Google Scholar] [CrossRef]
  16. Han, F.; Wu, L. Preparing and characterizing natural hydroxyapatite ceramics. Ceram. Int. 2010, 220, 281–285. [Google Scholar]
  17. Gergely, G.; Wéber, F.; Lukács, I.; Illés, L.; Tóth, A.L.; Horváth, Z.E.; Mihály, J.; Balázsi, C. Nano-hydroxyapatite preparation from biogenic raw materials. Cent. Eur. J. Chem. 2010, 8, 375–381. [Google Scholar] [CrossRef]
  18. Harabi, A.; Belamri, D.; Karboua, N.; Mezahi, F.Z. Sintering of bioceramics using a modified domestic microwave oven—Natural hydroxyapatite sintering. J. Therm. Anal. Calorim. 2011, 104, 383–388. [Google Scholar] [CrossRef]
  19. Mondal, S.; Mahata, S.; Kundu, S.; Mondal, B. Processing of natural resourced hydroxyapatite ceramics from fish scale. Adv. Appl. Ceram. 2010, 109, 234–239. [Google Scholar] [CrossRef]
  20. Xuebin, Z.; Yunfei, D.; Songlin, W.; Jie, X.; Yi, F. Sintering behavior and kinetic evaluation of hydroxyapatite bio-ceramics from bovine bone. Ceram. Silik. 2010, 54, 248–252. [Google Scholar]
  21. Lim, K.T.; Suh, J.D.; Kim, J.; Choung, P.H.; Chung, J.H. Calcium phosphate bioceramics fabricated from extracted human teeth for tooth tissue engineering. J. Biomed. Mater. Res. B 2011, 99, 399–411. [Google Scholar] [CrossRef]
  22. Rajesh, R.; Hariharasubramanian, A.; Ravichandran, Y.D. Chicken bone as a bioresource for the bioceramic (hydroxyapatite). Phosphorus Sulfur Silicon Relat. Elem. 2012, 187, 914–925. [Google Scholar] [CrossRef]
  23. Seo, D.S.; Hwang, K.H.; Yoon, S.Y.; Lee, J.K. Fabrication of hydroxyapatite bioceramics from the recycling of pig bone. J. Ceram. Proc. Res. 2012, 13, 586–589. [Google Scholar]
  24. Ho, W.F.; Hsu, H.C.; Hsu, S.K.; Hung, C.W.; Wu, S.C. Calcium phosphate bioceramics synthesized from eggshell powders through a solid state reaction. Ceram. Int. 2013, 39, 6467–6473. [Google Scholar] [CrossRef]
  25. Piccirillo, C.; Dunnill, C.W.; Pullar, R.C.; Tobaldi, D.M.; Labrincha, J.A.; Parkin, I.P.; Pintado, M.M.; Castro, P.M.L. Calcium phosphate-based materials of natural origin showing photocatalytic activity. J. Mater. Chem. A 2013, 1, 6452–6461. [Google Scholar] [CrossRef]
  26. Sayer, M.; Stratilatov, A.D.; Reid, J.W.; Calderin, L.; Stott, M.J.; Yin, X.; MacKenzie, M.; Smith, T.J.N.; Hendry, J.A.; Langstaff, S.D. Structure and composition of silicon-stabilized tricalcium phosphate. Biomaterials 2003, 24, 369–382. [Google Scholar] [CrossRef] [PubMed]
  27. Reid, J.W.; Pietak, A.M.; Sayer, M.; Dunfield, D.; Smith, T.J.N. Phase formation and evolution in the silicon substituted tricalcium phosphate/apatite system. Biomaterials 2005, 26, 2887–2897. [Google Scholar] [CrossRef] [PubMed]
  28. Sanchez-Sálcedo, S.; Arcos, D.; Vallet-Regí, M. Upgrading calcium phosphate scaffolds for tissue engineering applications. Key Eng. Mater. 2008, 377, 19–42. [Google Scholar] [CrossRef]
  29. Ergun, C.; Webster, T.J.; Bizios, R.; Doremus, R.H. Hydroxylapatite with substituted magnesium, zinc, cadmium, and yttrium. I. Structure and microstructure. J. Biomed. Mater. Res. 2002, 59, 305–311. [Google Scholar] [CrossRef] [PubMed]
  30. Webster, T.J.; Ergun, C.; Doremus, R.H.; Bizios, R. Hydroxylapatite with substituted magnesium, zinc, cadmium, and yttrium. II. Mechanisms of osteoblast adhesion. J. Biomed. Mater. Res. 2002, 59, 312–317. [Google Scholar] [CrossRef] [PubMed]
  31. Kim, S.R.; Lee, J.H.; Kim, Y.T.; Riu, D.H.; Jung, S.J.; Lee, Y.J.; Chung, S.C.; Kim, Y.H. Synthesis of Si, Mg substituted hydroxyapatites and their sintering behaviors. Biomaterials 2003, 24, 1389–1398. [Google Scholar] [CrossRef] [PubMed]
  32. Landi, E.; Celotti, G.; Logroscino, G.; Tampieri, A. Carbonated hydroxyapatite as bone substitute. J. Eur. Ceram. Soc. 2003, 23, 2931–2937. [Google Scholar] [CrossRef]
  33. Vallet-Regí, M.; Arcos, D. Silicon substituted hydroxyapatites. A method to upgrade calcium phosphate based implants. J. Mater. Chem. 2005, 15, 1509–1516. [Google Scholar] [CrossRef]
  34. Gbureck, U.; Thull, R.; Barralet, J.E. Alkali ion substituted calcium phosphate cement formation from mechanically activated reactants. J. Mater. Sci. Mater. Med. 2005, 16, 423–427. [Google Scholar] [CrossRef] [PubMed]
  35. Gbureck, U.; Knappe, O.; Grover, L.M.; Barralet, J.E. Antimicrobial potency of alkali ion substituted calcium phosphate cements. Biomaterials 2005, 26, 6880–6886. [Google Scholar] [CrossRef] [PubMed]
  36. Kannan, S.; Ventura, J.M.; Ferreira, J.M.F. In situ formation and characterization of flourine-substituted biphasic calcium phosphate ceramics of varied F-HAP/β-TCP ratios. Chem. Mater. 2005, 17, 3065–3068. [Google Scholar] [CrossRef]
  37. Landi, E.; Tampieri, A.; Mattioli-Belmonte, M.; Celotti, G.; Sandri, M.; Gigante, A.; Fava, P.; Biagini, G. Biomimetic Mg- and MgCO3-substituted hydroxyapatites: Synthesis characterization and in vitro behaviour. J. Eur. Ceram. Soc. 2006, 26, 2593–2601. [Google Scholar] [CrossRef]
  38. Reid, J.W.; Tuck, L.; Sayer, M.; Fargo, K.; Hendry, J.A. Synthesis and characterization of single-phase silicon substituted α-tricalcium phosphate. Biomaterials 2006, 27, 2916–2925. [Google Scholar] [CrossRef] [PubMed]
  39. Tas, A.C.; Bhaduri, S.B.; Jalota, S. Preparation of Zn-doped β-tricalcium phosphate (β-Ca3(PO4)2) bioceramics. Mater. Sci. Eng. C 2007, 27, 394–401. [Google Scholar] [CrossRef]
  40. Pietak, A.M.; Reid, J.W.; Stott, M.J.; Sayer, M. Silicon substitution in the calcium phosphate bioceramics. Biomaterials 2007, 28, 4023–4032. [Google Scholar] [CrossRef] [PubMed]
  41. Landi, E.; Tampieri, A.; Celotti, G.; Sprio, S.; Sandri, M.; Logroscino, G. Sr-substituted hydroxyapatites for osteoporotic bone replacement. Acta Biomater. 2007, 3, 961–969. [Google Scholar] [CrossRef] [PubMed]
  42. Kannan, S.; Ventura, J.M.G.; Ferreira, J.M.F. Synthesis and thermal stability of potassium substituted hydroxyapatites and hydroxyapatite/β-tricalcium phosphate mixtures. Ceram. Int. 2007, 33, 1489–1494. [Google Scholar] [CrossRef]
  43. Kannan, S.; Rebelo, A.; Lemos, A.F.; Barba, A.; Ferreira, J.M.F. Synthesis and mechanical behaviour of chlorapatite and chlorapatite/β-TCP composites. J. Eur. Ceram. Soc. 2007, 27, 2287–2294. [Google Scholar] [CrossRef]
  44. Ito, A.; LeGeros, R.Z. Magnesium- and zinc-substituted beta-tricalcium phosphates as potential bone substitute biomaterials. Key Eng. Mater. 2008, 377, 85–98. [Google Scholar] [CrossRef]
  45. Yao, Z.P.; Liu, W.G.; Ni, G.X. Biology characteristics and clinical application of strontium substituted hydroxyapatite bone cement. J. Clin. Rehabil. Tissue Eng. Res. 2008, 12, 7151–7154. [Google Scholar]
  46. Kannan, S.; Goetz-Neunhoeffer, F.; Neubauer, J.; Ferreira, J.M.F. Ionic substitutions in biphasic hydroxyapatite and β-tricalcium phosphate mixtures: Structural analysis by rietveld refinement. J. Am. Ceram. Soc. 2008, 91, 1–12. [Google Scholar] [CrossRef]
  47. Lafon, J.P.; Champion, E.; Bernache-Assollant, D. Processing of AB-type carbonated hydroxyapatite Ca10−x(PO4)6−x(CO3)x(OH)2−x−2y(CO3)y ceramics with controlled composition. J. Eur. Ceram. Soc. 2008, 28, 139–147. [Google Scholar] [CrossRef]
  48. Ren, F.; Xin, R.; Ge, X.; Leng, Y. An experimental and computational study on Zn-substituted hydroxyapatite. Adv. Mater. Res. 2008, 47–50, 1379–1382. [Google Scholar] [CrossRef]
  49. Meejoo, S.; Pon-On, W.; Charnchai, S.; Amornsakchai, T. Substitution of iron in preparation of enhanced thermal property and bioactivity of hydroxyapatite. Adv. Mater. Res. 2008, 55–57, 689–692. [Google Scholar] [CrossRef]
  50. Kannan, S.; Goetz-Neunhoeffer, F.; Neubauer, J.; Ferreira, J.M.F. Synthesis and structure refinement of zinc-doped β-tricalcium phosphate powders. J. Am. Ceram. Soc. 2009, 92, 1592–1595. [Google Scholar] [CrossRef]
  51. Matsumoto, N.; Yoshida, K.; Hashimoto, K.; Toda, Y. Thermal stability of β-tricalcium phosphate doped with monovalent metal ions. Mater. Res. Bull. 2009, 44, 1889–1894. [Google Scholar] [CrossRef]
  52. Boanini, E.; Gazzano, M.; Bigi, A. Ionic substitutions in calcium phosphates synthesized at low temperature. Acta Biomater. 2010, 6, 1882–1894. [Google Scholar] [CrossRef] [PubMed]
  53. Habibovic, P.; Barralet, J.E. Bioinorganics and biomaterials: Bone repair. Acta Biomater. 2011, 7, 3013–3026. [Google Scholar] [CrossRef] [PubMed]
  54. Mellier, C.; Fayon, F.; Schnitzler, V.; Deniard, P.; Allix, M.; Quillard, S.; Massiot, D.; Bouler, J.M.; Bujoli, B.; Janvier, P. Characterization and properties of novel gallium-doped calcium phosphate ceramics. Inorg. Chem. 2011, 50, 8252–8260. [Google Scholar] [CrossRef] [PubMed]
  55. Ansar, E.B.; Ajeesh, M.; Yokogawa, Y.; Wunderlich, W.; Varma, H. Synthesis and characterization of iron oxide embedded hydroxyapatite bioceramics. J. Am. Ceram. Soc. 2012, 95, 2695–2699. [Google Scholar] [CrossRef]
  56. Zhang, M.; Wu, C.; Li, H.; Yuen, J.; Chang, J.; Xiao, Y. Preparation, characterization and in vitro angiogenic capacity of cobalt substituted β-tricalcium phosphate ceramics. J. Mater. Chem. 2012, 22, 21686–21694. [Google Scholar] [CrossRef]
  57. Shepherd, J.H.; Shepherd, D.V.; Best, S.M. Substituted hydroxyapatites for bone repair. J. Mater. Sci. Mater. Med. 2012, 23, 2335–2347. [Google Scholar] [CrossRef] [PubMed]
  58. Williams, D.F. The Williams Dictionary of Biomaterials; Liverpool University Press: Liverpool, UK, 1999; p. 368. [Google Scholar]
  59. Williams, D.F. On the nature of biomaterials. Biomaterials 2009, 30, 5897–5909. [Google Scholar] [CrossRef] [PubMed]
  60. Bongio, M.; van den Beucken, J.J.J.P.; Leeuwenburgh, S.C.G.; Jansen, J.A. Development of bone substitute materials: From ‘biocompatible’ to ‘instructive’. J. Mater. Chem. 2010, 20, 8747–8759. [Google Scholar] [CrossRef]
  61. Biomimetic Materials Chemistry; Mann, S. (Ed.) Wiley-VCH: Oxford, UK, 1996; p. 400.
  62. Vallet-Regí, M. Bioceramics: Where do we come from and which are the future expectations. Key Eng. Mater. 2008, 377, 1–18. [Google Scholar] [CrossRef]
  63. Jandt, K.D. Evolutions, revolutions and trends in biomaterials science—A perspective. Adv. Eng. Mater. 2007, 9, 1035–1050. [Google Scholar] [CrossRef]
  64. Meyers, M.A.; Chen, P.Y.; Lin, A.Y.M.; Seki, Y. Biological materials: Structure and mechanical properties. Prog. Mater. Sci. 2008, 53, 1–206. [Google Scholar] [CrossRef]
  65. Ceramics. Available online: http://en.wikipedia.org/wiki/Ceramics (accessed on 14 August 2013).
  66. Hench, L.L. Bioceramics: From concept to clinic. J. Am. Ceram. Soc. 1991, 74, 1487–1510. [Google Scholar] [CrossRef]
  67. Cao, W.; Hench, L.L. Bioactive materials. Ceram. Int. 1996, 22, 493–507. [Google Scholar] [CrossRef]
  68. Hench, L.L. Bioceramics. J. Am. Ceram. Soc. 1998, 81, 1705–1728. [Google Scholar] [CrossRef]
  69. Hench, L.L.; Day, D.E.; Höland, W.; Rheinberger, V.M. Glass and medicine. Int. J. Appl. Glass Sci. 2010, 1, 104–117. [Google Scholar] [CrossRef]
  70. Pinchuk, N.D.; Ivanchenko, L.A. Making calcium phosphate biomaterials. Powder Metall. Metal Ceram. 2003, 42, 357–371. [Google Scholar] [CrossRef]
  71. Heimann, R.B. Materials science of crystalline bioceramics: A review of basic properties and applications. CMU J. 2002, 1, 23–46. [Google Scholar]
  72. Tomoda, K.; Ariizumi, H.; Nakaji, T.; Makino, K. Hydroxyapatite particles as drug carriers for proteins. Colloids Surf. B 2010, 76, 226–235. [Google Scholar] [CrossRef]
  73. Zamoume, O.; Thibault, S.; Regnié, G.; Mecherri, M.O.; Fiallo, M.; Sharrock, P. Macroporous calcium phosphate ceramic implants for sustained drug delivery. Mater. Sci. Eng. C 2011, 31, 1352–1356. [Google Scholar] [CrossRef]
  74. Bose, S.; Tarafder, S. Calcium phosphate ceramic systems in growth factor and drug delivery for bone tissue engineering: A review. Acta Biomater. 2012, 8, 1401–1421. [Google Scholar] [CrossRef] [PubMed]
  75. Arcos, D.; Vallet-Regí, M. Bioceramics for drug delivery. Acta Mater. 2013, 61, 890–911. [Google Scholar] [CrossRef]
  76. Ducheyne, P.; Qiu, Q. Bioactive ceramics: The effect of surface reactivity on bone formation and bone cell function. Biomaterials 1999, 20, 2287–2303. [Google Scholar] [CrossRef] [PubMed]
  77. Dorozhkin, S.V. Calcium orthophosphates and human beings. A historical perspective from the 1770s until 1940. Biomatter 2012, 2, 53–70. [Google Scholar] [CrossRef] [PubMed]
  78. Dorozhkin, S.V. A detailed history of calcium orthophosphates from 1770s till 1950. Mater. Sci. Eng. C 2013, 33, 3085–3110. [Google Scholar] [CrossRef]
  79. Vallet-Regí, M.; González-Calbet, J.M. Calcium phosphates as substitution of bone tissues. Progr. Solid State Chem. 2004, 32, 1–31. [Google Scholar] [CrossRef]
  80. Taş, A.C.; Korkusuz, F.; Timuçin, M.; Akkaş, N. An investigation of the chemical synthesis and high-temperature sintering behaviour of calcium hydroxyapatite (HA) and tricalcium phosphate (TCP) bioceramics. J. Mater. Sci. Mater. Med. 1997, 8, 91–96. [Google Scholar] [PubMed]
  81. Layrolle, P.; Ito, A.; Tateishi, T. Sol-gel synthesis of amorphous calcium phosphate and sintering into microporous hydroxyapatite bioceramics. J. Am. Ceram. Soc. 1998, 81, 1421–1428. [Google Scholar] [CrossRef]
  82. Engin, N.O.; Tas, A.C. Manufacture of macroporous calcium hydroxyapatite bioceramics. J. Eur. Ceram. Soc. 1999, 19, 2569–2572. [Google Scholar] [CrossRef]
  83. Ahn, E.S.; Gleason, N.J.; Nakahira, A.; Ying, J.Y. Nanostructure processing of hydroxyapatite-based bioceramics. Nano Lett. 2001, 1, 149–153. [Google Scholar] [CrossRef]
  84. Khalil, K.A.; Kim, S.W.; Dharmaraj, N.; Kim, K.W.; Kim, H.Y. Novel mechanism to improve toughness of the hydroxyapatite bioceramics using high-frequency induction heat sintering. J. Mater. Process. Technol. 2007, 187–188, 417–420. [Google Scholar] [CrossRef]
  85. Laasri, S.; Taha, M.; Laghzizil, A.; Hlil, E.K.; Chevalier, J. The affect of densification and dehydroxylation on the mechanical properties of stoichiometric hydroxyapatite bioceramics. Mater. Res. Bull. 2010, 45, 1433–1437. [Google Scholar] [CrossRef]
  86. Kitamura, M.; Ohtsuki, C.; Ogata, S.; Kamitakahara, M.; Tanihara, M. Microstructure and bioresorbable properties of α-TCP ceramic porous body fabricated by direct casting method. Mater. Trans. 2004, 45, 983–988. [Google Scholar] [CrossRef]
  87. Kawagoe, D.; Ioku, K.; Fujimori, H.; Goto, S. Transparent β-tricalcium phosphate ceramics prepared by spark plasma sintering. J. Ceram. Soc. Jpn. 2004, 112, 462–463. [Google Scholar] [CrossRef]
  88. Wang, C.X.; Zhou, X.; Wang, M. Influence of sintering temperatures on hardness and Young’s modulus of tricalcium phosphate bioceramic by nanoindentation technique. Mater. Charact. 2004, 52, 301–307. [Google Scholar] [CrossRef]
  89. Ioku, K.; Kawachi, G.; Nakahara, K.; Ishida, E.H.; Minagi, H.; Okuda, T.; Yonezawa, I.; Kurosawa, H.; Ikeda, T. Porous granules of β-tricalcium phosphate composed of rod-shaped particles. Key Eng. Mater. 2006, 309–311, 1059–1062. [Google Scholar] [CrossRef]
  90. Kamitakahara, M.; Ohtsuki, C.; Miyazaki, T. Review paper: Behavior of ceramic biomaterials derived from tricalcium phosphate in physiological condition. J. Biomater. Appl. 2008, 23, 197–212. [Google Scholar] [CrossRef] [PubMed]
  91. Vorndran, E.; Klarner, M.; Klammert, U.; Grover, L.M.; Patel, S.; Barralet, J.E.; Gbureck, U. 3D powder printing of β-tricalcium phosphate ceramics using different strategies. Adv. Eng. Mater. 2008, 10, B67–B71. [Google Scholar] [CrossRef]
  92. Descamps, M.; Duhoo, T.; Monchau, F.; Lu, J.; Hardouin, P.; Hornez, J.C.; Leriche, A. Manufacture of macroporous β-tricalcium phosphate bioceramics. J. Eur. Ceram. Soc. 2008, 28, 149–157. [Google Scholar] [CrossRef]
  93. Liu, Y.; Kim, J.H.; Young, D.; Kim, S.; Nishimoto, S.K.; Yang, Y. Novel template-casting technique for fabricating β-tricalcium phosphate scaffolds with high interconnectivity and mechanical strength and in vitro cell responses. J. Biomed. Mater. Res. A 2010, 92, 997–1006. [Google Scholar] [PubMed]
  94. Carrodeguas, R.G.; de Aza, S. α-tricalcium phosphate: Synthesis, properties and biomedical applications. Acta Biomater. 2011, 7, 3536–3546. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, Y.; Kong, D.; Feng, X. Fabrication and properties of porous β-tricalcium phosphate ceramics prepared using a double slip-casting method using slips with different viscosities. Ceram. Int. 2012, 38, 2991–2996. [Google Scholar] [CrossRef]
  96. Tancret, F.; Bouler, J.M.; Chamousset, J.; Minois, L.M. Modelling the mechanical properties of microporous and macroporous biphasic calcium phosphate bioceramics. J. Eur. Ceram. Soc. 2006, 26, 3647–3656. [Google Scholar] [CrossRef]
  97. Bouler, J.M.; Trecant, M.; Delecrin, J.; Royer, J.; Passuti, N.; Daculsi, G. Macroporous biphasic calcium phosphate ceramics: Influence of five synthesis parameters on compressive strength. J. Biomed. Mater. Res. 1996, 32, 603–609. [Google Scholar] [CrossRef] [PubMed]
  98. Gauthier, O.; Bouler, J.M.; Aguado, E.; Pilet, P.; Daculsi, G. Macroporous biphasic calcium phosphate ceramics: Influence of macropore diameter and macroporosity percentage on bone ingrowth. Biomaterials 1998, 19, 133–139. [Google Scholar] [CrossRef] [PubMed]
  99. Daculsi, G. Biphasic calcium phosphate concept applied to artificial bone, implant coating and injectable bone substitute. Biomaterials 1998, 19, 1473–1478. [Google Scholar] [CrossRef] [PubMed]
  100. LeGeros, R.Z.; Lin, S.; Rohanizadeh, R.; Mijares, D.; LeGeros, J.P. Biphasic calcium phosphate bioceramics: Preparation, properties and applications. J. Mater. Sci. Mater. Med. 2003, 14, 201–209. [Google Scholar] [CrossRef] [PubMed]
  101. Daculsi, G.; Laboux, O.; Malard, O.; Weiss, P. Current state of the art of biphasic calcium phosphate bioceramics. J. Mater. Sci. Mater. Med. 2003, 14, 195–200. [Google Scholar] [CrossRef] [PubMed]
  102. Dorozhkina, E.I.; Dorozhkin, S.V. Mechanism of the solid-state transformation of a calcium-deficient hydroxyapatite (CDHA) into biphasic calcium phosphate (BCP) at elevated temperatures. Chem. Mater. 2002, 14, 4267–4272. [Google Scholar] [CrossRef]
  103. Daculsi, G. Biphasic calcium phosphate granules concept for injectable and mouldable bone substitute. Adv. Sci. Technol. 2006, 49, 9–13. [Google Scholar] [CrossRef]
  104. Lecomte, A.; Gautier, H.; Bouler, J.M.; Gouyette, A.; Pegon, Y.; Daculsi, G.; Merle, C. Biphasic calcium phosphate: A comparative study of interconnected porosity in two ceramics. J. Biomed. Mater. Res. B 2008, 84, 1–6. [Google Scholar] [CrossRef]
  105. Tanimoto, Y.; Shibata, Y.; Murakami, A.; Miyazaki, T.; Nishiyama, N. Effect of varying HAP/TCP ratios in tape-cast biphasic calcium phosphate ceramics on responcce in vitro. J. Hard Tiss. Biol. 2009, 18, 71–76. [Google Scholar] [CrossRef]
  106. Daculsi, G.; Baroth, S.; LeGeros, R.Z. 20 years of biphasic calcium phosphate bioceramics development and applications. Ceram. Eng. Sci. Proc. 2010, 30, 45–58. [Google Scholar]
  107. Lukić, M.; Stojanović, Z.; Škapin, S.D.; Maček-Kržmanc, M.; Mitrić, M.; Marković, S.; Uskoković, D. Dense fine-grained biphasic calcium phosphate (BCP) bioceramics designed by two-step sintering. J. Eur. Ceram. Soc. 2011, 31, 19–27. [Google Scholar] [CrossRef]
  108. Descamps, M.; Boilet, L.; Moreau, G.; Tricoteaux, A.; Lu, J.; Leriche, A.; Lardot, V.; Cambier, F. Processing and properties of biphasic calcium phosphates bioceramics obtained by pressureless sintering and hot isostatic pressing. J. Eur. Ceram. Soc. 2013, 33, 1263–1270. [Google Scholar] [CrossRef]
  109. Li, Y.; Kong, F.; Weng, W. Preparation and characterization of novel biphasic calcium phosphate powders (α-TCP/HA) derived from carbonated amorphous calcium phosphates. J. Biomed. Mater. Res. B 2009, 89, 508–517. [Google Scholar] [CrossRef]
  110. Sureshbabu, S.; Komath, M.; Varma, H.K. In situ formation of hydroxyapatite–alpha tricalcium phosphate biphasic ceramics with higher strength and bioactivity. J. Am. Ceram. Soc. 2012, 95, 915–924. [Google Scholar]
  111. Oishi, M.; Ohtsuki, C.; Kitamura, M.; Kamitakahara, M.; Ogata, S.; Miyazaki, T.; Tanihara, M. Fabrication and chemical durability of porous bodies consisting of biphasic tricalcium phosphates. Phosphorus Res. Bull. 2004, 17, 95–100. [Google Scholar] [CrossRef]
  112. Kamitakahara, M.; Ohtsuki, C.; Oishi, M.; Ogata, S.; Miyazaki, T.; Tanihara, M. Preparation of porous biphasic tricalcium phosphate and its in vivo behavior. Key Eng. Mater. 2005, 284–286, 281–284. [Google Scholar] [CrossRef]
  113. Wang, R.; Weng, W.; Deng, X.; Cheng, K.; Liu, X.; Du, P.; Shen, G.; Han, G. Dissolution behavior of submicron biphasic tricalcium phosphate powders. Key Eng. Mater. 2006, 309–311, 223–226. [Google Scholar] [CrossRef]
  114. Li, Y.; Weng, W.; Tam, K.C. Novel highly biodegradable biphasic tricalcium phosphates composed of α-tricalcium phosphate and β-tricalcium phosphate. Acta Biomater. 2007, 3, 251–254. [Google Scholar] [CrossRef] [PubMed]
  115. Li, Y.; Li, D.; Weng, W. In vitro dissolution behavior of biphasic tricalcium phosphate composite powders composed of α-tricalcium phosphate and β-tricalcium phosphate. Key Eng. Mater. 2008, 368–372, 1206–1208. [Google Scholar] [CrossRef]
  116. Zou, C.; Cheng, K.; Weng, W.; Song, C.; Du, P.; Shen, G.; Han, G. Characterization and dissolution–reprecipitation behavior of biphasic tricalcium phosphate powders. J. Alloys Compd. 2011, 509, 6852–6858. [Google Scholar] [CrossRef]
  117. Albuquerque, J.S.V.; Nogueira, R.E.F.Q.; da Silva, T.D.P.; Lima, D.O.; da Silva, M.H.P. Porous triphasic calcium phosphate bioceramics. Key Eng. Mater. 2004, 254–256, 1021–1024. [Google Scholar] [CrossRef]
  118. Mendonça, F.; Lourom, L.H.L.; de Campos, J.B.; da Silva, M.H.P. Porous biphasic and triphasic bioceramics scaffolds produced by gelcasting. Key Eng. Mater. 2008, 361–363, 27–30. [Google Scholar] [CrossRef]
  119. Vani, R.; Girija, E.K.; Elayaraja, K.; Parthiban, P.S.; Kesavamoorthy, R.; Narayana Kalkura, S. Hydrothermal synthesis of porous triphasic hydroxyapatite/(α and β) tricalcium phosphate. J. Mater. Sci. Mater. Med. 2009, 20, S43–S48. [Google Scholar] [CrossRef] [PubMed]
  120. Ahn, M.K.; Moon, Y.W.; Koh, Y.H.; Kim, H.E. Production of highly porous triphasic calcium phosphate scaffolds with excellent in vitro bioactivity using vacuum-assisted foaming of ceramic suspension (VFC) technique. Ceram. Int. 2013, 39, 5879–5885. [Google Scholar] [CrossRef]
  121. Dorozhkin, S.V. Biphasic, triphasic and multiphasic calcium orthophosphates. Acta Biomater. 2012, 8, 963–977. [Google Scholar] [CrossRef] [PubMed]
  122. Dorozhkin, S.V. Self-setting calcium orthophosphate formulations: Cements, concretes, pastes and putties. Int. J. Mater. Chem. 2011, 1, 1–48. [Google Scholar]
  123. Tamimi, F.; Sheikh, Z.; Barralet, J. Dicalcium phosphate cements: Brushite and monetite. Acta Biomater. 2012, 8, 474–487. [Google Scholar] [CrossRef] [PubMed]
  124. Drouet, C.; Largeot, C.; Raimbeaux, G.; Estournès, C.; Dechambre, G.; Combes, C.; Rey, C. Bioceramics: Spark plasma sintering (SPS) of calcium phosphates. Adv. Sci. Technol. 2006, 49, 45–50. [Google Scholar] [CrossRef]
  125. Ishihara, S.; Matsumoto, T.; Onoki, T.; Sohmura, T.; Nakahira, A. New concept bioceramics composed of octacalcium phosphate (OCP) and dicarboxylic acid-intercalated OCP via hydrothermal hot-pressing. Mater. Sci. Eng. C 2009, 29, 1885–1888. [Google Scholar] [CrossRef]
  126. Barinov, S.M.; Komlev, V.S. Osteoinductive ceramic materials for bone tissue restoration: Octacalcium phosphate (review). Inorg. Mater. Appl. Res. 2010, 1, 175–181. [Google Scholar] [CrossRef]
  127. Moseke, C.; Gbureck, U. Tetracalcium phosphate: Synthesis, properties and biomedical applications. Acta Biomater. 2010, 6, 3815–3823. [Google Scholar] [CrossRef] [PubMed]
  128. Morimoto, S.; Anada, T.; Honda, Y.; Suzuki, O. Comparative study on in vitro biocompatibility of synthetic octacalcium phosphate and calcium phosphate ceramics used clinically. Biomed. Mater. 2012, 7, 045020. [Google Scholar] [CrossRef] [PubMed]
  129. Tamimi, F.; Nihouannen, D.L.; Eimar, H.; Sheikh, Z.; Komarova, S.; Barralet, J. The effect of autoclaving on the physical and biological properties of dicalcium phosphate dihydrate bioceramics: Brushite vs. monetite. Acta Biomater. 2012, 8, 3161–3169. [Google Scholar] [CrossRef] [PubMed]
  130. Suzuki, O. Octacalcium phosphate (OCP)-based bone substitute materials. Jpn. Dent. Sci. Rev. 2013, 49, 58–71. [Google Scholar] [CrossRef]
  131. LeGeros, R.Z. Calcium Phosphates in Oral Biology and Medicine; Monographs in Oral Science; Karger: Basel, Switzerland, 1991; Volume 15, p. 201. [Google Scholar]
  132. Narasaraju, T.S.B.; Phebe, D.E. Some physico-chemical aspects of hydroxylapatite. J. Mater. Sci. 1996, 31, 1–21. [Google Scholar] [CrossRef]
  133. Elliott, J.C. Structure and Chemistry of the Apatites and Other Calcium Orthophosphates; Studies in Inorganic Chemistry; Elsevier: Amsterdam, The Netherlands, 1994; Volume 18, p. 389. [Google Scholar]
  134. Hydroxyapatite and Related Materials; Brown, P.W.; Constantz, B. (Eds.) CRC Press: Boca Raton, FL, USA, 1994; p. 343.
  135. Calcium Phosphates in Biological and Industrial Systems; Amjad, Z. (Ed.) Kluwer Academic Publishers: Boston, MA, USA, 1997; p. 529.
  136. Phosphates: Geochemical, Geobiological and Materials Importance; Series: Reviews in Mineralogy and Geochemistry; Hughes, J.M.; Kohn, M.; Rakovan, J. (Eds.) Mineralogical Society of America: Washington, DC, USA, 2002; Volume 48, p. 742.
  137. O’Neill, W.C. The fallacy of the calcium–phosphorus product. Kidney Int. 2007, 72, 792–796. [Google Scholar] [CrossRef] [PubMed]
  138. Da Silva, R.V.; Bertran, C.A.; Kawachi, E.Y.; Camilli, J.A. Repair of cranial bone defects with calcium phosphate ceramic implant or autogenous bone graft. J. Craniofac. Surg. 2007, 18, 281–286. [Google Scholar] [CrossRef] [PubMed]
  139. Okanoue, Y.; Ikeuchi, M.; Takemasa, R.; Tani, T.; Matsumoto, T.; Sakamoto, M.; Nakasu, M. Comparison of in vivo bioactivity and compressive strength of a novel superporous hydroxyapatite with beta-tricalcium phosphates. Arch. Orthop. Trauma Surg. 2012, 132, 1603–1610. [Google Scholar] [CrossRef] [PubMed]
  140. Draenert, M.; Draenert, A.; Draenert, K. Osseointegration of hydroxyapatite and remodeling-resorption of tricalciumphosphate ceramics. Microsc. Res. Tech. 2013, 76, 370–380. [Google Scholar] [CrossRef] [PubMed]
  141. Okuda, T.; Ioku, K.; Yonezawa, I.; Minagi, H.; Gonda, Y.; Kawachi, G.; Kamitakahara, M.; Shibata, Y.; Murayama, H.; Kurosawa, H.; et al. The slow resorption with replacement by bone of a hydrothermally synthesized pure calcium-deficient hydroxyapatite. Biomaterials 2008, 29, 2719–2728. [Google Scholar] [CrossRef] [PubMed]
  142. Daculsi, G.; Bouler, J.M.; LeGeros, R.Z. Adaptive crystal formation in normal and pathological calcifications in synthetic calcium phosphate and related biomaterials. Int. Rev. Cytol. 1997, 172, 129–191. [Google Scholar] [PubMed]
  143. Astala, R.; Calderin, L.; Yin, X.; Stott, M.J. Ab initio simulation of Si-doped hydroxyapatite. Chem. Mater. 2006, 18, 413–422. [Google Scholar] [CrossRef]
  144. Zhu, X.D.; Zhang, H.J.; Fan, H.S.; Li, W.; Zhang, X.D. Effect of phase composition and microstructure of calcium phosphate ceramic particles on protein adsorption. Acta Biomater. 2010, 6, 1536–1541. [Google Scholar] [CrossRef] [PubMed]
  145. Bohner, M. Calcium orthophosphates in medicine: From ceramics to calcium phosphate cements. Injury 2000, 31, D37–D47. [Google Scholar] [CrossRef]
  146. Ahato, I. Reverse engineering the ceramic art of algae. Science 1999, 286, 1059–1061. [Google Scholar] [PubMed]
  147. Popişter, F.; Popescu, D.; Hurgoiu, D. A new method for using reverse engineering in case of ceramic tiles. Qual. Access Success 2012, 13, 409–412. [Google Scholar]
  148. Yang, S.; Leong, K.F.; Du, Z.; Chua, C.K. The design of scaffolds for use in tissue engineering. Part II. Rapid prototyping techniques. Tissue Eng. 2002, 8, 1–11. [Google Scholar] [CrossRef] [PubMed]
  149. Yeong, W.Y.; Chua, C.K.; Leong, K.F.; Chandrasekaran, M. Rapid prototyping in tissue engineering: Challenges and potential. Trends Biotechnol. 2004, 22, 643–652. [Google Scholar] [CrossRef] [PubMed]
  150. Hieu, L.C.; Zlatov, N.; Sloten, J.V.; Bohez, E.; Khanh, L.; Binh, P.H.; Oris, P.; Toshev, Y. Medical rapid prototyping applications and methods. Assem. Autom. 2005, 25, 284–292. [Google Scholar] [CrossRef]
  151. Eufinger, H.; Wehniöller, M.; Machtens, E.; Heuser, L.; Harders, A.; Kruse, D. Reconstruction of craniofacial bone defects with individual alloplastic implants based on CAD/CAM-manipulated CT-data. J. Cranio Maxillofac. Surg. 1995, 23, 175–181. [Google Scholar] [CrossRef]
  152. Klein, M.; Glatzer, C. Individual CAD/CAM fabricated glass-bioceramic implants in reconstructive surgery of the bony orbital floor. Plast. Reconstruct. Surg. 2006, 117, 565–570. [Google Scholar] [CrossRef]
  153. Yin, L.; Song, X.F.; Song, Y.L.; Huang, T.; Li, J. An overview of in vitro abrasive finishing & CAD/CAM of bioceramics in restorative dentistry. Int. J. Mach. Tools Manuf. 2006, 46, 1013–1026. [Google Scholar] [CrossRef]
  154. Li, J.; Hsu, Y.; Luo, E.; Khadka, A.; Hu, J. Computer-aided design and manufacturing and rapid prototyped nanoscale hydroxyapatite/polyamide (n-HA/PA) construction for condylar defect caused by mandibular angle ostectomy. Aesthet. Plast. Surg. 2011, 35, 636–640. [Google Scholar] [CrossRef]
  155. Yardimci, M.A.; Guceri, S.I.; Danforth, S.C. Process modeling for fused deposition of ceramics. Ceram. Eng. Sci. Proc. 1996, 17, 78–82. [Google Scholar]
  156. Bellini, A.; Shor, L.; Guceri, S.I. New developments in fused deposition modeling of ceramics. Rapid Prototyp. J. 2005, 11, 214–220. [Google Scholar] [CrossRef]
  157. Tan, K.H.; Chua, C.K.; Leong, K.F.; Cheah, C.M.; Cheang, P.; Abu Bakar, M.S.; Cha, S.W. Scaffold development using selective laser sintering of polyetheretherketone-hydroxyapatite biocomposite blends. Biomaterials 2003, 24, 3115–3123. [Google Scholar] [CrossRef] [PubMed]
  158. Wiria, F.E.; Leong, K.F.; Chua, C.K.; Liu, Y. Poly-ε-caprolactone/hydroxyapatite for tissue engineering scaffold fabrication via selective laser sintering. Acta Biomater. 2007, 3, 1–12. [Google Scholar] [CrossRef] [PubMed]
  159. Xiao, K.; Dalgarno, K.W.; Wood, D.J.; Goodridge, R.D.; Ohtsuki, C. Indirect selective laser sintering of apatite-wollostonite glass-ceramic. Proc. Inst. Mech. Eng. H 2008, 222, 1107–1114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Zhou, W.Y.; Lee, S.H.; Wang, M.; Cheung, W.L.; Ip, W.Y. Selective laser sintering of porous tissue engineering scaffolds from poly(L-lactide)/carbonated hydroxyapatite nanocomposite microspheres. J. Mater. Sci. Mater. Med. 2008, 19, 2535–2540. [Google Scholar] [CrossRef] [PubMed]
  161. Shuai, C.; Zhuang, J.; Hu, H.; Peng, S.; Liu, D.; Liu, J. In vitro bioactivity and degradability of β-tricalcium phosphate porous scaffold fabricated via selective laser sintering. Biotechnol. Appl. Biochem. 2013, 60, 266–273. [Google Scholar] [CrossRef] [PubMed]
  162. Lusquiños, F.; Pou, J.; Boutinguiza, M.; Quintero, F.; Soto, R.; León, B.; Pérez-Amor, M. Main characteristics of calcium phosphate coatings obtained by laser cladding. Appl. Surf. Sci. 2005, 247, 486–492. [Google Scholar] [CrossRef]
  163. Wang, D.G.; Chen, C.Z.; Ma, J.; Zhang, G. In situ synthesis of hydroxyapatite coating by laser cladding. Colloids Surf. B 2008, 66, 155–162. [Google Scholar] [CrossRef]
  164. Comesaña, R.; Lusquiños, F.; del Val, J.; Malot, T.; López-Álvarez, M.; Riveiro, A.; Quintero, F.; Boutinguiza, M.; Aubry, P.; de Carlos, A.; et al. Calcium phosphate grafts produced by rapid prototyping based on laser cladding. J. Eur. Ceram. Soc. 2011, 31, 29–41. [Google Scholar] [CrossRef]
  165. Lv, X.; Lin, X.; Hu, J.; Gao, B.; Huang, W. Phase evolution in calcium phosphate coatings obtained by in situ laser cladding. Mater. Sci. Eng. C 2012, 32, 872–877. [Google Scholar] [CrossRef]
  166. Seitz, H.; Rieder, W.; Irsen, S.; Leukers, B.; Tille, C. Three-dimensional printing of porous ceramic scaffolds for bone tissue engineering. J. Biomed. Mater. Res. B 2005, 74, 782–788. [Google Scholar] [CrossRef]
  167. Leukers, B.; Gülkan, H.; Irsen, S.H.; Milz, S.; Tille, C.; Schieker, M.; Seitz, H. Hydroxyapatite scaffolds for bone tissue engineering made by 3D printing. J. Mater. Sci. Mater. Med. 2005, 16, 1121–1124. [Google Scholar] [CrossRef] [PubMed]
  168. Gbureck, U.; Hölzel, T.; Klammert, U.; Würzler, K.; Müller, F.A.; Barralet, J.E. Resorbable dicalcium phosphate bone substitutes prepared by 3D powder printing. Adv. Funct. Mater. 2007, 17, 3940–3945. [Google Scholar] [CrossRef]
  169. Gbureck, U.; Hölzel, T.; Doillon, C.J.; Müller, F.A.; Barralet, J.E. Direct printing of bioceramic implants with spatially localized angiogenic factors. Adv. Mater. 2007, 19, 795–800. [Google Scholar] [CrossRef]
  170. Khalyfa, A.; Vogt, S.; Weisser, J.; Grimm, G.; Rechtenbach, A.; Meyer, W.; Schnabelrauch, M. Development of a new calcium phosphate powder-binder system for the 3D printing of patient specific implants. J. Mater. Sci. Mater. Med. 2007, 18, 909–916. [Google Scholar] [CrossRef] [PubMed]
  171. Habibovic, P.; Gbureck, U.; Doillon, C.J.; Bassett, D.C.; van Blitterswijk, C.A.; Barralet, J.E. Osteoconduction and osteoinduction of low-temperature 3D printed bioceramic implants. Biomaterials 2008, 29, 944–953. [Google Scholar] [CrossRef] [PubMed]
  172. Fierz, F.C.; Beckmann, F.; Huser, M.; Irsen, S.H.; Leukers, B.; Witte, F.; Degistirici, O.; Andronache, A.; Thie, M.; Müller, B. The morphology of anisotropic 3D-printed hydroxyapatite scaffolds. Biomaterials 2008, 29, 3799–3806. [Google Scholar] [CrossRef] [PubMed]
  173. Seitz, H.; Deisinger, U.; Leukers, B.; Detsch, R.; Ziegler, G. Different calcium phosphate granules for 3-D printing of bone tissue engineering scaffolds. Adv. Eng. Mater. 2009, 11, B41–B46. [Google Scholar] [CrossRef]
  174. Suwanprateeb, J.; Sanngam, R.; Panyathanmaporn, T. Influence of raw powder preparation routes on properties of hydroxyapatite fabricated by 3D printing technique. Mater. Sci. Eng. C 2010, 30, 610–617. [Google Scholar] [CrossRef]
  175. Butscher, A.; Bohner, M.; Roth, C.; Ernstberger, A.; Heuberger, R.; Doebelin, N.; von Rohr, R.P.; Müller, R. Printability of calcium phosphate powders for three-dimensional printing of tissue engineering scaffolds. Acta Biomater. 2012, 8, 373–385. [Google Scholar] [CrossRef] [PubMed]
  176. Butscher, A.; Bohner, M.; Doebelin, N.; Galea, L.; Loeffel, O.; Müller, R. Moisture based three-dimensional printing of calcium phosphate structures for scaffold engineering. Acta Biomater. 2013, 9, 5369–5378. [Google Scholar] [CrossRef] [PubMed]
  177. Porter, N.L.; Pilliar, R.M.; Grynpas, M.D. Fabrication of porous calcium polyphosphate implants by solid freeform fabrication: A study of processing parameters and in vitro degradation characteristics. J. Biomed. Mater. Res. 2001, 56, 504–515. [Google Scholar] [CrossRef] [PubMed]
  178. Leong, K.F.; Cheah, C.M.; Chua, C.K. Solid freeform fabrication of three-dimensional scaffolds for engineering replacement tissues and organs. Biomaterials 2003, 24, 2363–2378. [Google Scholar] [CrossRef] [PubMed]
  179. Calvert, J.W.; Brenner, K.A.; Mooney, M.P.; Kumta, P.; Weiss, L.E. Cellular fusion of hydroxyapatite layers: Solid freeform fabrication of synthetic bone grafts. Riv. Ital. Chir. Plast. 2004, 36, 145–150. [Google Scholar]
  180. Jongpaiboonkit, L.; Lin, C.Y.; Krebsbach, P.H.; Hollister, S.J.; Halloran, J.W. Mechanical behavior of complex 3D calcium phosphate cement scaffolds fabricated by indirect solid freeform fabrication in vivo. Key Eng. Mater. 2006, 309–311, 957–960. [Google Scholar] [CrossRef]
  181. Dellinger, J.G.; Wojtowicz, A.M.; Jamison, R.D. Effects of degradation and porosity on the load bearing properties of model hydroxyapatite bone scaffolds. J. Biomed. Mater. Res. A 2006, 77, 563–571. [Google Scholar] [CrossRef] [PubMed]
  182. Dellinger, J.G.; Cesarano, J., 3rd; Jamison, R.D. Robotic deposition of model hydroxyapatite scaffolds with multiple architectures and multiscale porosity for bone tissue engineering. J. Biomed. Mater. Res. A 2007, 82, 383–394. [Google Scholar] [CrossRef] [PubMed]
  183. Shanjani, Y.; de Croos, J.N.A.; Pilliar, R.M.; Kandel, R.A.; Toyserkani, E. Solid freeform fabrication and characterization of porous calcium polyphosphate structures for tissue engineering purposes. J. Biomed. Mater. Res. B 2010, 93, 510–519. [Google Scholar] [CrossRef]
  184. Kim, J.; Lim, D.; Kim, Y.H.; Koh, Y.H.; Lee, M.H.; Han, I.; Lee, S.J.; Yoo, O.S.; Kim, H.S.; Park, J.C. A comparative study of the physical and mechanical properties of porous hydroxyapatite scaffolds fabricated by solid freeform fabrication and polymer replication method. Int. J. Precis. Eng. Manuf. 2011, 12, 695–701. [Google Scholar] [CrossRef]
  185. Shanjani, Y.; Hu, Y.; Toyserkani, E.; Grynpas, M.; Kandel, R.A.; Pilliar, R.M. Solid freeform fabrication of porous calcium polyphosphate structures for bone substitute applications: in vivo studies. J. Biomed. Mater. Res. B 2013, 101, 972–980. [Google Scholar] [CrossRef]
  186. Chu, T.M.G.; Halloran, J.W.; Hollister, S.J.; Feinberg, S.E. Hydroxyapatite implants with designed internal architecture. J. Mater. Sci. Mater. Med. 2001, 12, 471–478. [Google Scholar] [CrossRef] [PubMed]
  187. Li, X.; Li, D.; Lu, B.; Tang, Y.; Wang, L.; Wang, Z. Design and fabrication of CAP scaffolds by indirect solid free form fabrication. Rapid Prototyp. J. 2005, 11, 312–318. [Google Scholar] [CrossRef]
  188. Woesz, A.; Rumpler, M.; Stampfl, J.; Varga, F.; Fratzl-Zelman, N.; Roschger, P.; Klaushofer, K.; Fratzl, P. Towards bone replacement materials from calcium phosphates via rapid prototyping and ceramic gel casting. Mater. Sci. Eng. C 2005, 25, 181–186. [Google Scholar] [CrossRef]
  189. Li, X.; Li, D.; Lu, B.; Wang, C. Fabrication of bioceramic scaffolds with pre-designed internal architecture by gel casting and indirect stereolithography techniques. J. Porous Mater. 2008, 15, 667–671. [Google Scholar] [CrossRef]
  190. Saber-Samandari, S.; Gross, K.A. The use of thermal printing to control the properties of calcium phosphate deposits. Biomaterials 2010, 31, 6386–6393. [Google Scholar] [CrossRef] [PubMed]
  191. Narayan, R.J.; Jin, C.; Doraiswamy, A.; Mihailescu, I.N.; Jelinek, M.; Ovsianikov, A.; Chichkov, B.; Chrisey, D.B. Laser processing of advanced bioceramics. Adv. Eng. Mater. 2005, 7, 1083–1098. [Google Scholar] [CrossRef]
  192. Bone Grafts and Bone Substitutes: Basic Science and Clinical Applications; Nather, A. (Ed.) World Scientific: Singapore, 2005; p. 592.
  193. Bio-Materials and Prototyping Applications in Medicine; Bártolo, P.; Bidanda, B. (Eds.) Springer: New York, NY, USA, 2008; p. 216.
  194. Bioceramics and Their Clinical Applications; Kokubo, T. (Ed.) Woodhead Publishing: Cambridge, UK, 2008; p. 784.
  195. Biomedical Materials; Narayan, R. (Ed.) Springer: New York, NY, USA, 2009; p. 566.
  196. Raksujarit, A.; Pengpat, K.; Rujijanagul, G.; Tunkasiri, T. Processing and properties of nanoporous hydroxyapatite ceramics. Mater. Des. 2010, 31, 1658–1660. [Google Scholar] [CrossRef]
  197. Park, J. Bioceramics: Properties, Characterizations, and Applications; Springer: New York, NY, USA, 2008; p. 364. [Google Scholar]
  198. Rodríguez-Lorenzo, L.M.; Vallet-Regí, M.; Ferreira, J.M.F. Fabrication of hydroxyapatite bodies by uniaxial pressing from a precipitated powder. Biomaterials 2001, 22, 583–588. [Google Scholar] [CrossRef] [PubMed]
  199. Miranda, P.; Pajares, A.; Saiz, E.; Tomsia, A.P.; Guiberteau, F. Mechanical behaviour under uniaxial compression of robocast calcium phosphate scaffolds. Eur. Cells Mater. 2007, 14, 79. [Google Scholar]
  200. Nazarpak, M.H.; Solati-Hashjin, M.; Moztarzadeh, F. Preparation of hydroxyapatite ceramics for biomedical applications. J. Ceram. Process. Res. 2009, 10, 54–57. [Google Scholar]
  201. Uematsu, K.; Takagi, M.; Honda, T.; Uchida, N.; Saito, K. Transparent hydroxyapatite prepared by hot isostatic pressing of filter cake. J. Am. Ceram. Soc. 1989, 72, 1476–1478. [Google Scholar] [CrossRef]
  202. Itoh, H.; Wakisaka, Y.; Ohnuma, Y.; Kuboki, Y. A new porous hydroxyapatite ceramic prepared by cold isostatic pressing and sintering synthesized flaky powder. Dent. Mater. 1994, 13, 25–35. [Google Scholar] [CrossRef]
  203. Takikawa, K.; Akao, M. Fabrication of transparent hydroxyapatite and application to bone marrow derived cell/hydroxyapatite interaction observation in vivo. J. Mater. Sci. Mater. Med. 1996, 7, 439–445. [Google Scholar] [CrossRef]
  204. Gautier, H.; Merle, C.; Auget, J.L.; Daculsi, G. Isostatic compression, a new process for incorporating vancomycin into biphasic calcium phosphate: Comparison with a classical method. Biomaterials 2000, 21, 243–249. [Google Scholar] [CrossRef] [PubMed]
  205. Tadic, D.; Epple, M. Mechanically stable implants of synthetic bone mineral by cold isostatic pressing. Biomaterials 2003, 24, 4565–4571. [Google Scholar] [CrossRef] [PubMed]
  206. Onoki, T.; Hashida, T. New method for hydroxyapatite coating of titanium by the hydrothermal hot isostatic pressing technique. Surf. Coat. Technol. 2006, 200, 6801–6807. [Google Scholar] [CrossRef]
  207. Pecqueux, F.; Tancret, F.; Payraudeau, N.; Bouler, J.M. Influence of microporosity and macroporosity on the mechanical properties of biphasic calcium phosphate bioceramics: Modelling and experiment. J. Eur. Ceram. Soc. 2010, 30, 819–829. [Google Scholar] [CrossRef]
  208. Irsen, S.H.; Leukers, B.; Höckling, Chr.; Tille, C.; Seitz, H. Bioceramic granulates for use in 3D printing: Process engineering aspects. Materialwissenschaft Werkstofftechnik 2006, 37, 533–537. [Google Scholar] [CrossRef]
  209. Hsu, Y.H.; Turner, I.G.; Miles, A.W. Fabrication and mechanical testing of porous calcium phosphate bioceramic granules. J. Mater. Sci. Mater. Med. 2007, 18, 1931–1937. [Google Scholar] [CrossRef] [PubMed]
  210. Zyman, Z.Z.; Glushko, V.; Dedukh, N.; Malyshkina, S.; Ashukina, N. Porous calcium phosphate ceramic granules and their behaviour in differently loaded areas of skeleton. J. Mater. Sci. Mater. Med. 2008, 19, 2197–2205. [Google Scholar] [CrossRef] [PubMed]
  211. Viana, M.; Désiré, A.; Chevalier, E.; Champion, E.; Chotard, R.; Chulia, D. Interest of high shear wet granulation to produce drug loaded porous calcium phosphate pellets for bone filling. Key Eng. Mater. 2009, 396–398, 535–538. [Google Scholar] [CrossRef]
  212. Chevalier, E.; Viana, M.; Cazalbou, S.; Chulia, D. Comparison of low-shear and high-shear granulation processes: Effect on implantable calcium phosphate granule properties. Drug Dev. Ind. Pharm. 2009, 35, 1255–1263. [Google Scholar] [CrossRef] [PubMed]
  213. Lakevics, V.; Locs, J.; Loca, D.; Stepanova, V.; Berzina-Cimdina, L.; Pelss, J. Bioceramic hydroxyapatite granules for purification of biotechnological products. Adv. Mater. Res. 2011, 284–286, 1764–1769. [Google Scholar] [CrossRef]
  214. Reikerås, O.; Johansson, C.B.; Sundfeldt, M. Bone ingrowths to press-fit and loose-fit implants: Comparisons between titanium and hydroxyapatite. J. Long-Term Eff. Med. Implant. 2006, 16, 157–164. [Google Scholar] [CrossRef]
  215. Rao, R.R.; Kannan, T.S. Dispersion and slip casting of hydroxyapatite. J. Am. Ceram. Soc. 2001, 84, 1710–1716. [Google Scholar] [CrossRef]
  216. Sakka, Y.; Takahashi, K.; Matsuda, N.; Suzuki, T.S. Effect of milling treatment on texture development of hydroxyapatite ceramics by slip casting in high magnetic field. Mater. Trans. 2007, 48, 2861–2866. [Google Scholar] [CrossRef]
  217. Zhang, Y.; Yokogawa, Y.; Feng, X.; Tao, Y.; Li, Y. Preparation and properties of bimodal porous apatite ceramics through slip casting using different hydroxyapatite powders. Ceram. Int. 2010, 36, 107–113. [Google Scholar] [CrossRef]
  218. Albano, M.P.; Garrido, L.B. Processing of concentrated aqueous fluorapatite suspensions by slip casting. J. Mater. Sci. 2011, 46, 5117–5128. [Google Scholar] [CrossRef]
  219. Zhang, Y.; Kong, D.; Yokogawa, Y.; Feng, X.; Tao, Y.; Qiu, T. Fabrication of porous hydroxyapatite ceramic scaffolds with high flexural strength through the double slip-casting method using fine powders. J. Am. Ceram. Soc. 2012, 95, 147–152. [Google Scholar] [CrossRef]
  220. Hagio, T.; Yamauchi, K.; Kohama, T.; Matsuzaki, T.; Iwai, K. Beta tricalcium phosphate ceramics with controlled crystal orientation fabricated by application of external magnetic field during the slip casting process. Mater. Sci. Eng. C 2013, 33, 2967–2970. [Google Scholar] [CrossRef]
  221. Sepulveda, P.; Ortega, F.S.; Innocentini, M.D.M.; Pandolfelli, V.C. Properties of highly porous hydroxyapatite obtained by the gel casting of foams. J. Am. Ceram. Soc. 2000, 83, 3021–3024. [Google Scholar] [CrossRef]
  222. Padilla, S.; Vallet-Regí, M.; Ginebra, M.P.; Gil, F.J. Processing and mechanical properties of hydroxyapatite pieces obtained by the gel-casting method. J. Eur. Ceram. Soc. 2005, 25, 375–383. [Google Scholar] [CrossRef]
  223. Sánchez-Salcedo, S.; Werner, J.; Vallet-Regí, M. Hierarchical pore structure of calcium phosphate scaffolds by a combination of gel-casting and multiple tape-casting methods. Acta Biomater. 2008, 4, 913–922. [Google Scholar] [CrossRef] [PubMed]
  224. Chen, B.; Zhang, T.; Zhang, J.; Lin, Q.; Jiang, D. Microstructure and mechanical properties of hydroxyapatite obtained by gel-casting process. Ceram. Int. 2008, 34, 359–364. [Google Scholar] [CrossRef]
  225. Marcassoli, P.; Cabrini, M.; Tirillò, J.; Bartuli, C.; Palmero, P.; Montanaro, L. Mechanical characterization of hydroxiapatite micro/macro-porous ceramics obtained by means of innovative gel-casting process. Key Eng. Mater. 2010, 417–418, 565–568. [Google Scholar]
  226. Fomin, A.S.; Barinov, S.M.; Ievlev, V.M.; Smirnov, V.V.; Mikhailov, B.P.; Belonogov, E.K.; Drozdova, N.A. Nanocrystalline hydroxyapatite ceramics produced by low-temperature sintering after high-pressure treatment. Dokl. Chem. 2008, 418, 22–25. [Google Scholar] [CrossRef]
  227. Kankawa, Y.; Kaneko, Y.; Saitou, K. Injection molding of highly-purified hydroxylapatite and TCP utilizing solid phase reaction method. J. Ceram. Soc. Jpn. 1991, 99, 438–442. [Google Scholar] [CrossRef]
  228. Cihlář, J.; Trunec, M. Injection moulded hydroxyapatite ceramics. Biomaterials 1996, 17, 1905–1911. [Google Scholar] [CrossRef] [PubMed]
  229. Jewad, R.; Bentham, C.; Hancock, B.; Bonfield, W.; Best, S.M. Dispersant selection for aqueous medium pressure injection moulding of anhydrous dicalcium phosphate. J. Eur. Ceram. Soc. 2008, 28, 547–553. [Google Scholar] [CrossRef]
  230. Kwon, S.H.; Jun, Y.K.; Hong, S.H.; Lee, I.S.; Kim, H.E.; Won, Y.Y. Calcium phosphate bioceramics with various porosities and dissolution rates. J. Am. Ceram. Soc. 2002, 85, 3129–3131. [Google Scholar] [CrossRef]
  231. Fooki, A.C.B.M.; Aparecida, A.H.; Fideles, T.B.; Costa, R.C.; Fook, M.V.L. Porous hydroxyapatite scaffolds by polymer sponge method. Key Eng. Mater. 2009, 396–398, 703–706. [Google Scholar] [CrossRef]
  232. Sopyan, I.; Kaur, J. Preparation and characterization of porous hydroxyapatite through polymeric sponge method. Ceram. Int. 2009, 35, 3161–3168. [Google Scholar] [CrossRef]
  233. Bellucci, D.; Cannillo, V.; Sola, A. Shell scaffolds: A new approach towards high strength bioceramic scaffolds for bone regeneration. Mater. Lett. 2010, 64, 203–206. [Google Scholar] [CrossRef]
  234. Cunningham, E.; Dunne, N.; Walker, G.; Maggs, C.; Wilcox, R.; Buchanan, F. Hydroxyapatite bone substitutes developed via replication of natural marine sponges. J. Mater. Sci. Mater. Med. 2010, 21, 2255–2261. [Google Scholar] [CrossRef] [PubMed]
  235. Sung, J.H.; Shin, K.H.; Koh, Y.H.; Choi, W.Y.; Jin, Y.; Kim, H.E. Preparation of the reticulated hydroxyapatite ceramics using carbon-coated polymeric sponge with elongated pores as a novel template. Ceram. Int. 2011, 37, 2591–2596. [Google Scholar] [CrossRef]
  236. Mishima, F.D.; Louro, L.H.L.; Moura, F.N.; Gobbo, L.A.; da Silva, M.H.P. Hydroxyapatite scaffolds produced by hydrothermal deposition of monetite on polyurethane sponges substrates. Key Eng. Mater. 2012, 493–494, 820–825. [Google Scholar]
  237. Sopyan, I.; Mardziah, C.M.; Ramesh, S. Fabrication of porous ceramic scaffolds via polymeric sponge method using sol-gel derived strontium doped hydroxyapatite. Appl. Mech. Mater. 2012, 117–119, 829–832. [Google Scholar]
  238. Velayudhan, S.; Ramesh, P.; Sunny, M.C.; Varma, H.K. Extrusion of hydroxyapatite to clinically significant shapes. Mater. Lett. 2000, 46, 142–146. [Google Scholar] [CrossRef]
  239. Yang, H.Y.; Thompson, I.; Yang, S.F.; Chi, X.P.; Evans, J.R.G.; Cook, R.J. Dissolution characteristics of extrusion freeformed hydroxyapatite–tricalcium phosphate scaffolds. J. Mater. Sci. Mater. Med. 2008, 19, 3345–3353. [Google Scholar] [CrossRef] [PubMed]
  240. Yang, S.; Yang, H.; Chi, X.; Evans, J.R.G.; Thompson, I.; Cook, R.J.; Robinson, P. Rapid prototyping of ceramic lattices for hard tissue scaffolds. Mater. Des. 2008, 29, 1802–1809. [Google Scholar] [CrossRef]
  241. Yang, H.Y.; Chi, X.P.; Yang, S.; Evans, J.R.G. Mechanical strength of extrusion freeformed calcium phosphate filaments. J. Mater. Sci. Mater. Med. 2010, 21, 1503–1510. [Google Scholar] [CrossRef] [PubMed]
  242. Cortez, P.P.; Atayde, L.M.; Silva, M.A.; da Silva, P.A.; Fernandes, M.H.; Afonso, A.; Lopes, M.A.; Maurício, A.C.; Santos, J.D. Characterization and preliminary in vivo evaluation of a novel modified hydroxyapatite produced by extrusion and spheronization techniques. J. Biomed. Mater. Res. B 2011, 99, 170–179. [Google Scholar] [CrossRef]
  243. Muthutantri, A.I.; Huang, J.; Edirisinghe, M.J.; Bretcanu, O.; Boccaccini, A.R. Dipping and electrospraying for the preparation of hydroxyapatite foams for bone tissue engineering. Biomed. Mater. 2008, 3, 25009. [Google Scholar] [CrossRef]
  244. Roncari, E.; Galassi, C.; Pinasco, P. Tape casting of porous hydroxyapatite ceramics. J. Mater. Sci. Lett. 2000, 19, 33–35. [Google Scholar] [CrossRef]
  245. Tian, T.; Jiang, D.; Zhang, J.; Lin, Q. Aqueous tape casting process for hydroxyapatite. J. Eur. Ceram. Soc. 2007, 27, 2671–2677. [Google Scholar] [CrossRef]
  246. Tanimoto, Y.; Teshima, M.; Nishiyama, N.; Yamaguchi, M.; Hirayama, S.; Shibata, Y.; Miyazaki, T. Tape-cast and sintered β-tricalcium phosphate laminates for biomedical applications: Effect of milled Al2O3 fiber additives on microstructural and mechanical properties. J. Biomed. Mater. Res. B 2012, 100, 2261–2268. [Google Scholar] [CrossRef]
  247. Suzuki, S.; Itoh, K.; Ohgaki, M.; Ohtani, M.; Ozawa, M. Preparation of sintered filter for ion exchange by a doctor blade method with aqueous slurries of needlelike hydroxyapatite. Ceram. Int. 1999, 25, 287–291. [Google Scholar] [CrossRef]
  248. Suchanek, W.L.; Yoshimura, M. Processing and properties of hydroxyapatite-based biomaterials for use as hard tissue replacement implants. J. Mater. Res. 1998, 13, 94–117. [Google Scholar] [CrossRef]
  249. Padilla, S.; Roman, J.; Vallet-Regí, M. Synthesis of porous hydroxyapatites by combination of gel casting and foams burn out methods. J. Mater. Sci. Mater. Med. 2002, 13, 1193–1197. [Google Scholar] [CrossRef] [PubMed]
  250. Yang, T.Y.; Lee, J.M.; Yoon, S.Y.; Park, H.C. Hydroxyapatite scaffolds processed using a TBA-based freeze-gel casting/polymer sponge technique. J. Mater. Sci. Mater. Med. 2010, 21, 1495–1502. [Google Scholar] [CrossRef] [PubMed]
  251. Baradararan, S.; Hamdi, M.; Metselaar, I.H. Biphasic calcium phosphate (BCP) macroporous scaffold with different ratios of HA/β-TCP by combination of gel casting and polymer sponge methods. Adv. Appl. Ceram. 2012, 111, 367–373. [Google Scholar] [CrossRef]
  252. Inoue, K.; Sassa, K.; Yokogawa, Y.; Sakka, Y.; Okido, M.; Asai, S. Control of crystal orientation of hydroxyapatite by imposition of a high magnetic field. Mater. Trans. 2003, 44, 1133–1137. [Google Scholar] [CrossRef]
  253. Iwai, K.; Akiyama, J.; Tanase, T.; Asai, S. Alignment of HAp crystal using a sample rotation in a static magnetic field. Mater. Sci. Forum 2007, 539–543, 716–719. [Google Scholar] [CrossRef]
  254. Iwai, K.; Akiyama, J.; Asai, S. Structure control of hydroxyapatite using a magnetic field. Mater. Sci. Forum 2007, 561–565, 1565–1568. [Google Scholar] [CrossRef]
  255. Sakka, Y.; Takahashi, K.; Suzuki, T.S.; Ito, S.; Matsuda, N. Texture development of hydroxyapatite ceramics by colloidal processing in a high magnetic field followed by sintering. Mater. Sci. Eng. A 2008, 475, 27–33. [Google Scholar] [CrossRef]
  256. Fleck, N.A. On the cold compaction of powders. J. Mech. Phys. Solids 1995, 43, 1409–1431. [Google Scholar] [CrossRef]
  257. Kang, J.; Hadfield, M. Parameter optimization by Taguchi methods for finishing advanced ceramic balls using a novel eccentric lapping machine. Proc. Inst. Mech. Eng. B 2001, 215, 69–78. [Google Scholar] [CrossRef]
  258. Kurella, A.; Dahotre, N.B. Surface modification for bioimplants: The role of laser surface engineering. J. Biomater. Appl. 2005, 20, 5–50. [Google Scholar] [CrossRef] [PubMed]
  259. Oktar, F.N.; Genc, Y.; Goller, G.; Erkmen, E.Z.; Ozyegin, L.S.; Toykan, D.; Demirkiran, H.; Haybat, H. Sintering of synthetic hydroxyapatite compacts. Key Eng. Mater. 2004, 264–268, 2087–2090. [Google Scholar] [CrossRef]
  260. Georgiou, G.; Knowles, J.C.; Barralet, J.E. Dynamic shrinkage behavior of hydroxyapatite and glass-reinforced hydroxyapatites. J. Mater. Sci. 2004, 39, 2205–2208. [Google Scholar] [CrossRef]
  261. Fellah, B.H.; Layrolle, P. Sol-gel synthesis and characterization of macroporous calcium phosphate bioceramics containing microporosity. Acta Biomater. 2009, 5, 735–742. [Google Scholar] [CrossRef] [PubMed]
  262. Dudek, A.; Kolan, C. Assessments of shrinkage degree in bioceramic sinters HA + ZrO2. Diffusion Defect Data B 2010, 165, 25–30. [Google Scholar]
  263. Ben Ayed, F.; Bouaziz, J.; Bouzouita, K. Pressureless sintering of fluorapatite under oxygen atmosphere. J. Eur. Ceram. Soc. 2000, 20, 1069–1076. [Google Scholar] [CrossRef]
  264. He, Z.; Ma, J.; Wang, C. Constitutive modeling of the densification and the grain growth of hydroxyapatite ceramics. Biomaterials 2005, 26, 1613–1621. [Google Scholar] [CrossRef] [PubMed]
  265. Rahaman, M.N. Sintering of Ceramics; CRC Press: Boca Raton, FL, USA, 2007; p. 388. [Google Scholar]
  266. Monroe, E.A.; Votava, W.; Bass, D.B.; McMullen, J. New calcium phosphate ceramic material for bone and tooth implants. J. Dent. Res. 1971, 50, 860–861. [Google Scholar] [CrossRef] [PubMed]
  267. Landi, E.; Tampieri, A.; Celotti, G.; Sprio, S. Densification behaviour and mechanisms of synthetic hydroxyapatites. J. Eur. Ceram. Soc. 2000, 20, 2377–2387. [Google Scholar] [CrossRef]
  268. Chen, S.; Wang, W.; Kono, H.; Sassa, K.; Asai, S. Abnormal grain growth of hydroxyapatite ceramic sintered in a high magnetic field. J. Cryst. Growth 2010, 312, 323–326. [Google Scholar] [CrossRef]
  269. Ruys, A.J.; Wei, M.; Sorrell, C.C.; Dickson, M.R.; Brandwood, A.; Milthorpe, B.K. Sintering effect on the strength of hydroxyapatite. Biomaterials 1995, 16, 409–415. [Google Scholar] [CrossRef] [PubMed]
  270. Van Landuyt, P.; Li, F.; Keustermans, J.P.; Streydio, J.M.; Delannay, F.; Munting, E. The influence of high sintering temperatures on the mechanical properties of hydroxylapatite. J. Mater. Sci. Mater. Med. 1995, 6, 8–13. [Google Scholar] [CrossRef]
  271. Pramanik, S.; Agarwal, A.K.; Rai, K.N.; Garg, A. Development of high strength hydroxyapatite by solid-state-sintering process. Ceram. Int. 2007, 33, 419–426. [Google Scholar] [CrossRef]
  272. Haberko, K.; Bućko, M.M.; Brzezińska-Miecznik, J.; Haberko, M.; Mozgawa, W.; Panz, T.; Pyda, A.; Zarebski, J. Natural hydroxyapatite—Its behaviour during heat treatment. J. Eur. Ceram. Soc. 2006, 26, 537–542. [Google Scholar] [CrossRef]
  273. Haberko, K.; Bućko, M.M.; Mozgawa, W.; Pyda, A.; Brzezińska-Miecznik, J.; Carpentier, J. Behaviour of bone origin hydroxyapatite at elevated temperatures and in O2 and CO2 atmospheres. Ceram. Int. 2009, 35, 2537–2540. [Google Scholar] [CrossRef]
  274. Janus, A.M.; Faryna, M.; Haberko, K.; Rakowska, A.; Panz, T. Chemical and microstructural characterization of natural hydroxyapatite derived from pig bones. Mikrochim. Acta 2008, 161, 349–353. [Google Scholar] [CrossRef]
  275. Bahrololoom, M.E.; Javidi, M.; Javadpour, S.; Ma, J. Characterisation of natural hydroxyapatite extracted from bovine cortical bone ash. J. Ceram. Process. Res. 2009, 10, 129–138. [Google Scholar]
  276. Mostafa, N.Y. Characterization, thermal stability and sintering of hydroxyapatite powders prepared by different routes. Mater. Chem. Phys. 2005, 94, 333–341. [Google Scholar] [CrossRef]
  277. Suchanek, W.; Yashima, M.; Kakihana, M.; Yoshimura, M. Hydroxyapatite ceramics with selected sintering additives. Biomaterials 1997, 18, 923–933. [Google Scholar] [CrossRef] [PubMed]
  278. Kalita, S.J.; Bose, S.; Bandyopadhyay, A.; Hosick, H.L. Oxide based sintering additives for HAp ceramics. Ceram. Trans. 2003, 147, 63–72. [Google Scholar]
  279. Kalita, S.J.; Bose, S.; Hosick, H.L.; Bandyopadhyay, A. CaO–P2O5–Na2O-based sintering additives for hydroxyapatite (HAp) ceramics. Biomaterials 2004, 25, 2331–2339. [Google Scholar] [CrossRef] [PubMed]
  280. Safronova, T.V.; Putlyaev, V.I.; Shekhirev, M.A.; Tretyakov, Y.D.; Kuznetsov, A.V.; Belyakov, A.V. Densification additives for hydroxyapatite ceramics. J. Eur. Ceram. Soc. 2009, 29, 1925–1932. [Google Scholar] [CrossRef]
  281. Muralithran, G.; Ramesh, S. Effects of sintering temperature on the properties of hydroxyapatite. Ceram. Int. 2000, 26, 221–230. [Google Scholar] [CrossRef]
  282. Eskandari, A.; Aminzare, M.; Hassani, H.; Barounian, H.; Hesaraki, S.; Sadrnezhaad, S.K. Densification behavior and mechanical properties of biomimetic apatite nanocrystals. Curr. Nanosci. 2011, 7, 776–780. [Google Scholar] [CrossRef]
  283. Ramesh, S.; Tolouei, R.; Tan, C.Y.; Aw, K.L.; Yeo, W.H.; Sopyan, I.; Teng, W.D. Sintering of hydroxyapatite ceramic produced by wet chemical method. Adv. Mater. Res. 2011, 264–265, 1856–1861. [Google Scholar]
  284. Ou, S.F.; Chiou, S.Y.; Ou, K.L. Phase transformation on hydroxyapatite decomposition. Ceram. Int. 2013, 39, 3809–3816. [Google Scholar] [CrossRef]
  285. Bernache-Assollant, D.; Ababou, A.; Champion, E.; Heughebaert, M. Sintering of calcium phosphate hydroxyapatite Ca10(PO4)6(OH)2 I. Calcination and particle growth. J. Eur. Ceram. Soc. 2003, 23, 229–241. [Google Scholar] [CrossRef]
  286. Ramesh, S.; Tan, C.Y.; Bhaduri, S.B.; Teng, W.D.; Sopyan, I. Densification behaviour of nanocrystalline hydroxyapatite bioceramics. J. Mater. Process. Technol. 2008, 206, 221–230. [Google Scholar] [CrossRef]
  287. Wang, J.; Shaw, L.L. Grain-size dependence of the hardness of submicrometer and nanometer hydroxyapatite. J. Am. Ceram. Soc. 2010, 93, 601–604. [Google Scholar] [CrossRef]
  288. Kobayashi, S.; Kawai, W.; Wakayama, S. The effect of pressure during sintering on the strength and the fracture toughness of hydroxyapatite ceramics. J. Mater. Sci. Mater. Med. 2006, 17, 1089–1093. [Google Scholar] [CrossRef] [PubMed]
  289. Chen, I.W.; Wang, X.H. Sintering dense nanocrystalline ceramics without final-stage grain growth. Nature 2000, 404, 168–170. [Google Scholar] [CrossRef] [PubMed]
  290. Mazaheri, M.; Haghighatzadeh, M.; Zahedi, A.M.; Sadrnezhaad, S.K. Effect of a novel sintering process on mechanical properties of hydroxyapatite ceramics. J. Alloys Compd. 2009, 471, 180–184. [Google Scholar] [CrossRef]
  291. Lin, K.; Chen, L.; Chang, J. Fabrication of dense hydroxyapatite nanobioceramics with enhanced mechanical properties via two-step sintering process. Int. J. Appl. Ceram. Technol. 2012, 9, 479–485. [Google Scholar] [CrossRef]
  292. Panyata, S.; Eitssayeam, S.; Rujijanagul, G.; Tunkasiri, T.; Pengpat, K. Property development of hydroxyapatite ceramics by two-step sintering. Adv. Mater. Res. 2012, 506, 190–193. [Google Scholar] [CrossRef]
  293. Esnaashary, M.; Fathi, M.; Ahmadian, M. The effect of the two-step sintering process on consolidation of fluoridated hydroxyapatite and its mechanical properties and bioactivity. Int. J. Appl. Ceram. Technol. 2013. [Google Scholar] [CrossRef]
  294. Kasuga, T.; Ota, Y.; Tsuji, K.; Abe, Y. Preparation of high-strength calcium phosphate ceramics with low modulus of elasticity containing β-Ca(PO3)2 fibers. J. Am. Ceram. Soc. 1996, 79, 1821–1824. [Google Scholar] [CrossRef]
  295. Suchanek, W.L.; Yoshimura, M. Preparation of fibrous, porous hydroxyapatite ceramics from hydroxyapatite whiskers. J. Am. Ceram. Soc. 1998, 81, 765–767. [Google Scholar] [CrossRef]
  296. Kim, Y.; Kim, S.R.; Song, H.; Yoon, H. Preparation of porous hydroxyapatite/TCP composite block using a hydrothermal hot pressing method. Mater. Sci. Forum 2005, 486–487, 117–120. [Google Scholar] [CrossRef]
  297. Li, J.G.; Hashida, T. Preparation of hydroxyapatite ceramics by hydrothermal hot-pressing method at 300 °C. J. Mater. Sci. 2007, 42, 5013–5019. [Google Scholar] [CrossRef]
  298. Li, J.G.; Hashida, T. In situ formation of hydroxyapatite-whisker ceramics by hydrothermal hot-pressing method. J. Am. Ceram. Soc. 2006, 89, 3544–3546. [Google Scholar] [CrossRef]
  299. Halouani, R.; Bernache-Assolant, D.; Champion, E.; Ababou, A. Microstructure and related mechanical properties of hot pressed hydroxyapatite ceramics. J. Mater. Sci. Mater. Med. 1994, 5, 563–568. [Google Scholar] [CrossRef]
  300. Nakahira, A.; Murakami, T.; Onoki, T.; Hashida, T.; Hosoi, K. Fabrication of porous hydroxyapatite using hydrothermal hot pressing and post-sintering. J. Am. Ceram. Soc. 2005, 88, 1334–1336. [Google Scholar] [CrossRef]
  301. Auger, M.A.; Savoini, B.; Muñoz, A.; Leguey, T.; Monge, M.A.; Pareja, R.; Victoria, J. Mechanical characteristics of porous hydroxyapatite/oxide composites produced by post-sintering hot isostatic pressing. Ceram. Int. 2009, 35, 2373–2380. [Google Scholar] [CrossRef]
  302. Fang, Y.; Agrawal, D.K.; Roy, D.M.; Roy, R. Microwave sintering of hydroxyapatite ceramics. J. Mater. Res. 1994, 9, 180–187. [Google Scholar] [CrossRef]
  303. Yang, Y.; Ong, J.L.; Tian, J. Rapid sintering of hydroxyapatite by microwave processing. J. Mater. Sci. Lett. 2002, 21, 67–69. [Google Scholar] [CrossRef]
  304. Nath, S.; Basu, B.; Sinha, A. A comparative study of conventional sintering with microwave sintering of hydroxyapatite synthesized by chemical route. Trends Biomater. Artif. Organs 2006, 19, 93–98. [Google Scholar]
  305. Ramesh, S.; Tan, C.Y.; Bhaduri, S.B.; Teng, W.D. Rapid densification of nanocrystalline hydroxyapatite for biomedical applications. Ceram. Int. 2007, 33, 1363–1367. [Google Scholar] [CrossRef]
  306. Silva, C.C.; Graça, M.P.F.; Sombra, A.S.B.; Valente, M.A. Structural and electrical study of calcium phosphate obtained by a microwave radiation assisted procedure. Phys. Rev. B Condens. Matter 2009, 404, 1503–1508. [Google Scholar] [CrossRef]
  307. Chanda, A.; Dasgupta, S.; Bose, S.; Bandyopadhyay, A. Microwave sintering of calcium phosphate ceramics. Mater. Sci. Eng. C 2009, 29, 1144–1149. [Google Scholar] [CrossRef]
  308. Veljović, D.; Zalite, I.; Palcevskis, E.; Smiciklas, I.; Petrović, R.; Janaćković, D. Microwave sintering of fine grained HAP and HAP/TCP bioceramics. Ceram. Int. 2010, 36, 595–603. [Google Scholar] [CrossRef]
  309. Kalita, S.J.; Verma, S. Nanocrystalline hydroxyapatite bioceramic using microwave radiation: Synthesis and characterization. Mater. Sci. Eng. C 2010, 30, 295–303. [Google Scholar] [CrossRef]
  310. Veljović, D.; Palcevskis, E.; Dindune, A.; Putić, S.; Balać, I.; Petrović, R.; Janaćković, D. Microwave sintering improves the mechanical properties of biphasic calcium phosphates from hydroxyapatite microspheres produced from hydrothermal processing. J. Mater. Sci. 2010, 45, 3175–3183. [Google Scholar] [CrossRef]
  311. Gu, Y.W.; Loh, N.H.; Khor, K.A.; Tor, S.B.; Cheang, P. Spark plasma sintering of hydroxyapatite powders. Biomaterials 2002, 23, 37–43. [Google Scholar] [CrossRef] [PubMed]
  312. Guo, X.; Xiao, P.; Liu, J.; Shen, Z. Fabrication of nanostructured hydroxyapatite via hydrothermal synthesis and spark plasma sintering. J. Am. Ceram. Soc. 2005, 88, 1026–1029. [Google Scholar] [CrossRef]
  313. Li, H.; Khor, K.A.; Chow, V.; Cheang, P. Nanostructural characteristics, mechanical properties and osteoblast response of spark plasma sintered hydroxyapatite. J. Biomed. Mater. Res. A 2007, 82, 296–303. [Google Scholar] [CrossRef] [PubMed]
  314. Nakamura, T.; Fukuhara, T.; Izui, H. Mechanical properties of hydroxyapatites sintered by spark plasma sintering. Ceram. Trans. 2006, 194, 265–272. [Google Scholar]
  315. Zhang, F.; Lin, K.; Chang, J.; Lu, J.; Ning, C. Spark plasma sintering of macroporous calcium phosphate scaffolds from nanocrystalline powders. J. Eur. Ceram. Soc. 2008, 28, 539–545. [Google Scholar] [CrossRef]
  316. Grossin, D.; Rollin-Martinet, S.; Estournès, C.; Rossignol, F.; Champion, E.; Combes, C.; Rey, C.; Geoffroy, C.; Drouet, C. Biomimetic apatite sintered at very low temperature by spark plasma sintering: Physico-chemistry and microstructure aspects. Acta Biomater. 2010, 6, 577–585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Chesnaud, A.; Bogicevic, C.; Karolak, F.; Estournès, C.; Dezanneau, G. Preparation of transparent oxyapatite ceramics by combined use of freeze-drying and spark-plasma sintering. Chem. Commun. 2007, 1550–1552. [Google Scholar] [CrossRef]
  318. Watanabe, T.; Fukuhara, T.; Izui, H.; Fukase, Y.; Okano, M. Properties of HAp/β-TCP functionally graded material by spark plasma sintering. Trans. Jpn. Soc. Mech. Eng. A 2009, 75, 612–618. [Google Scholar]
  319. Eriksson, M.; Liu, Y.; Hu, J.; Gao, L.; Nygren, M.; Shen, Z. Transparent hydroxyapatite ceramics with nanograin structure prepared by high pressure spark plasma sintering at the minimized sintering temperature. J. Eur. Ceram. Soc. 2011, 31, 1533–1540. [Google Scholar] [CrossRef]
  320. Liu, Y.; Shen, Z. Dehydroxylation of hydroxyapatite in dense bulk ceramics sintered by spark plasma sintering. J. Eur. Ceram. Soc. 2012, 32, 2691–2696. [Google Scholar] [CrossRef]
  321. Yoshida, H.; Kim, B.N.; Son, H.W.; Han, Y.H.; Kim, S. Superplastic deformation of transparent hydroxyapatite. Scripta Mater. 2013, 69, 155–158. [Google Scholar] [CrossRef]
  322. Yanagisawa, K.; Kim, J.H.; Sakata, C.; Onda, A.; Sasabe, E.; Yamamoto, T.; Matamoros-Veloza, Z.; Rendón-Angeles, J.C. Hydrothermal sintering under mild temperature conditions: Preparation of calcium-deficient hydroxyapatite compacts. Z. Naturforsch. B 2010, 65, 1038–1044. [Google Scholar] [CrossRef]
  323. Hosoi, K.; Hashida, T.; Takahashi, H.; Yamasaki, N.; Korenaga, T. New processing technique for hydroxyapatite ceramics by the hydrothermal hot-pressing method. J. Am. Ceram. Soc. 1996, 79, 2771–2774. [Google Scholar] [CrossRef]
  324. Gross, K.A.; Berndt, C.C. Biomedical Application of Apatites. In Phosphates: Geochemical, Geobiological and Materials Importance; Series: Reviews in Mineralogy and Geochemistry; Hughes, J.M., Kohn, M., Rakovan, J., Eds.; Mineralogical Society of America: Washington, DC, USA, 2002; Volume 48, pp. 631–672. [Google Scholar]
  325. Champion, E. Sintering of calcium phosphate bioceramics. Acta Biomater. 2013, 9, 5855–5875. [Google Scholar] [CrossRef] [PubMed]
  326. Evans, J.R.G. Seventy ways to make ceramics. J. Eur. Ceram. Soc. 2008, 28, 1421–1432. [Google Scholar] [CrossRef]
  327. Hench, L.L.; Polak, J.M. Third-generation biomedical materials. Science 2002, 295, 1014–1017. [Google Scholar] [CrossRef] [PubMed]
  328. Black, J. Biological Performance of Materials: Fundamentals of Biocompatibility, 4th ed.; CRC Press: Boca Raton, FL, USA, 2005; p. 520. [Google Scholar]
  329. Benaqqa, C.; Chevalier, J.; Saädaoui, M.; Fantozzi, G. Slow crack growth behaviour of hydroxyapatite ceramics. Biomaterials 2005, 26, 6106–6112. [Google Scholar] [CrossRef] [PubMed]
  330. Linhart, W.; Briem, D.; Amling, M.; Rueger, J.M.; Windolf, J. Mechanisches versagen einer porösen hydroxylapatitkeramik 7,5 jahre nach implantation an der proximalen tibia (in German). Unfallchirurg 2004, 107, 154–157. [Google Scholar] [CrossRef] [PubMed]
  331. Ramesh, S.; Tan, C.Y.; Sopyan, I.; Hamdi, M.; Teng, W.D. Consolidation of nanocrystalline hydroxyapatite powder. Sci. Technol. Adv. Mater. 2007, 8, 124–130. [Google Scholar] [CrossRef]
  332. Fan, X.; Case, E.D.; Ren, F.; Shu, Y.; Baumann, M.J. Part I: Porosity dependence of the Weibull modulus for hydroxyapatite and other brittle materials. J. Mech. Behav. Biomed. Mater. 2012, 8, 21–36. [Google Scholar] [CrossRef] [PubMed]
  333. Fan, X.; Case, E.D.; Gheorghita, I.; Baumann, M.J. Weibull modulus and fracture strength of highly porous hydroxyapatite. J. Mech. Behav. Biomed. Mater. 2013, 20, 283–295. [Google Scholar] [CrossRef] [PubMed]
  334. Suzuki, S.; Sakamura, M.; Ichiyanagi, M.; Ozawa, M. Internal friction of hydroxyapatite and fluorapatite. Ceram. Int. 2004, 30, 625–627. [Google Scholar] [CrossRef]
  335. Suzuki, S.; Takahiro, K.; Ozawa, M. Internal friction and dynamic modulus of polycrystalline ceramics prepared from stoichiometric and Ca-deficient hydroxyapatites. Mater. Sci. Eng. B 1998, 55, 68–70. [Google Scholar] [CrossRef]
  336. le Huec, J.C.; Schaeverbeke, T.; Clement, D.; Faber, J.; le Rebeller, A. Influence of porosity on the mechanical resistance of hydroxyapatite ceramics under compressive stress. Biomaterials 1995, 16, 113–118. [Google Scholar] [CrossRef] [PubMed]
  337. Hsu, Y.H.; Turner, I.G.; Miles, A.W. Mechanical properties of three different compositions of calcium phosphate bioceramic following immersion in Ringer’s solution and distilled water. J. Mater. Sci. Mater. Med. 2009, 20, 2367–2374. [Google Scholar] [CrossRef] [PubMed]
  338. Torgalkar, A.M. Resonance frequency technique to determine elastic modulus of hydroxyapatite. J. Biomed. Mater. Res. 1979, 13, 907–920. [Google Scholar] [CrossRef] [PubMed]
  339. Gilmore, R.S.; Katz, J.L. Elastic properties of apatites. J. Mater. Sci. 1982, 17, 1131–1141. [Google Scholar] [CrossRef]
  340. Fan, X.; Case, E.D.; Ren, F.; Shu, Y.; Baumann, M.J. Part II: Fracture strength and elastic modulus as a function of porosity for hydroxyapatite and other brittle materials. J. Mech. Behav. Biomed. Mater. 2012, 8, 99–110. [Google Scholar] [CrossRef] [PubMed]
  341. De Aza, P.N.; de Aza, A.H.; de Aza, S. Crystalline bioceramic materials. Bol. Soc. Esp. Ceram. V. 2005, 44, 135–145. [Google Scholar] [CrossRef]
  342. Fritsch, A.; Dormieux, L.; Hellmich, C.; Sanahuja, J. Mechanical behavior of hydroxyapatite biomaterials: An experimentally validated micromechanical model for elasticity and strength. J. Biomed. Mater. Res. A 2009, 88, 149–161. [Google Scholar] [CrossRef] [PubMed]
  343. Ching, W.Y.; Rulis, P.; Misra, A. Ab initio elastic properties and tensile strength of crystalline hydroxyapatite. Acta Biomater. 2009, 5, 3067–3075. [Google Scholar] [CrossRef] [PubMed]
  344. Fritsch, A.; Hellmich, C.; Dormieux, L. The role of disc-type crystal shape for micromechanical predictions of elasticity and strength of hydroxyapatite biomaterials. Philos. Trans. R. Soc. Lond. A 2010, 368, 1913–1935. [Google Scholar] [CrossRef]
  345. Menéndez-Proupin, E.; Cervantes-Rodríguez, S.; Osorio-Pulgar, R.; Franco-Cisterna, M.; Camacho-Montes, H.; Fuentes, M.E. Computer simulation of elastic constants of hydroxyapatite and fluorapatite. J. Mech. Behav. Biomed. Mater. 2011, 4, 1011–1120. [Google Scholar] [CrossRef] [PubMed]
  346. Sun, J.P.; Song, Y.; Wen, G.W.; Wang, Y.; Yang, R. Softening of hydroxyapatite by vacancies: A first principles investigation. Mater. Sci. Eng. C 2013, 33, 1109–1115. [Google Scholar] [CrossRef]
  347. Sha, M.C.; Li, Z.; Bradt, R.C. Single-crystal elastic constants of fluorapatite, Ca5F(PO4)3. J. Appl. Phys. 1994, 75, 7784–7787. [Google Scholar] [CrossRef]
  348. Wakai, F.; Kodama, Y.; Sakaguchi, S.; Nonami, T. Superplasticity of hot isostatically pressed hydroxyapatite. J. Am. Ceram. Soc. 1990, 73, 457–460. [Google Scholar] [CrossRef]
  349. Tago, K.; Itatani, K.; Suzuki, T.S.; Sakka, Y.; Koda, S. Densification and superplasticity of hydroxyapatite ceramics. J. Ceram. Soc. Jpn. 2005, 113, 669–673. [Google Scholar] [CrossRef]
  350. Burger, E.L.; Patel, V. Calcium phosphates as bone graft extenders. Orthopedics 2007, 30, 939–942. [Google Scholar] [PubMed]
  351. Rodriguez-Lorenzo, L.M.; Vallet-Regí, M.; Ferreira, J.M.F.; Ginebra, M.P.; Aparicio, C.; Planell, J. A hydroxyapatite ceramic bodies with tailored mechanical properties for different applications. J. Biomed. Mater. Res. 2002, 60, 159–166. [Google Scholar] [CrossRef] [PubMed]
  352. Song, J.; Liu, Y.; Zhang, Y.; Jiao, L. Mechanical properties of hydroxyapatite ceramics sintered from powders with different morphologies. Mater. Sci. Eng. A 2011, 528, 5421–5427. [Google Scholar] [CrossRef]
  353. Dorozhkin, S.V. Biocomposites and hybrid biomaterials based on calcium orthophosphates. Biomatter 2011, 1, 3–56. [Google Scholar] [CrossRef] [PubMed]
  354. Bouslama, N.; Ben Ayed, F.; Bouaziz, J. Sintering and mechanical properties of tricalcium phosphate–fluorapatite composites. Ceram. Int. 2009, 35, 1909–1917. [Google Scholar] [CrossRef]
  355. Suchanek, W.; Yashima, M.; Kakihana, M.; Yoshimura, M. Processing and mechanical properties of hydroxyapatite reinforced with hydroxyapatite whiskers. Biomaterials 1996, 17, 1715–1723. [Google Scholar] [CrossRef] [PubMed]
  356. Suchanek, W.; Yashima, M.; Kakihana, M.; Yoshimura, M. Hydroxyapatite/hydroxyapatite-whisker composites without sintering additives: Mechanical properties and microstructural evolution. J. Am. Ceram. Soc. 1997, 80, 2805–2813. [Google Scholar] [CrossRef]
  357. Simsek, D.; Ciftcioglu, R.; Guden, M.; Ciftcioglu, M.; Harsa, S. Mechanical properties of hydroxyapatite composites reinforced with hydroxyapatite whiskers. Key Eng. Mater. 2004, 264–268, 1985–1988. [Google Scholar] [CrossRef]
  358. Bose, S.; Banerjee, A.; Dasgupta, S.; Bandyopadhyay, A. Synthesis, processing, mechanical, and biological property characterization of hydroxyapatite whisker-reinforced hydroxyapatite composites. J. Am. Ceram. Soc. 2009, 92, 323–330. [Google Scholar] [CrossRef]
  359. Lie-Feng, L.; Xiao-Yi, H.; Cai, Y.X.; Weng, J. Reinforcing of porous hydroxyapatite ceramics with hydroxyapatite fibres for enhanced bone tissue engineering. J. Biomim. Biomater. Tissue Eng. 2011, 1314, 67–73. [Google Scholar]
  360. Shiota, T.; Shibata, M.; Yasuda, K.; Matsuo, Y. Influence of β-tricalcium phosphate dispersion on mechanical properties of hydroxyapatite ceramics. J. Ceram. Soc. Jpn. 2009, 116, 1002–1005. [Google Scholar] [CrossRef]
  361. Shuai, C.; Feng, P.; Nie, Y.; Hu, H.; Liu, J.; Peng, S. Nano-hydroxyapatite improves the properties of β-tricalcium phosphate bone scaffolds. Int. J. Appl. Ceram. Technol. 2013. [Google Scholar] [CrossRef]
  362. Dorozhkin, S.V.; Ajaal, T. Toughening of porous bioceramic scaffolds by bioresorbable polymeric coatings. Proc. Inst. Mech. Eng. Η 2009, 223, 459–470. [Google Scholar] [CrossRef]
  363. Dorozhkin, S.V.; Ajaal, T. Strengthening of dense bioceramic samples using bioresorbable polymers— A statistical approach. J. Biomim. Biomater. Tissue Eng. 2009, 4, 27–39. [Google Scholar] [CrossRef]
  364. Dressler, M.; Dombrowski, F.; Simon, U.; Börnstein, J.; Hodoroaba, V.D.; Feigl, M.; Grunow, S.; Gildenhaar, R.; Neumann, M. Influence of gelatin coatings on compressive strength of porous hydroxyapatite ceramics. J. Eur. Ceram. Soc. 2011, 31, 523–529. [Google Scholar] [CrossRef]
  365. Martinez-Vazquez, F.J.; Perera, F.H.; Miranda, P.; Pajares, A.; Guiberteau, F. Improving the compressive strength of bioceramic robocast scaffolds by polymer infiltration. Acta Biomater. 2010, 6, 4361–4368. [Google Scholar] [CrossRef] [PubMed]
  366. He, L.H.; Standard, O.C.; Huang, T.T.; Latella, B.A.; Swain, M.V. Mechanical behaviour of porous hydroxyapatite. Acta Biomater. 2008, 4, 577–586. [Google Scholar] [CrossRef] [PubMed]
  367. Kasuga, T.; Nakano, M.; Nogami, M. Fast proton conductors derived from calcium phosphate hydrogels. Adv. Mater. 2002, 14, 1490–1492. [Google Scholar] [CrossRef]
  368. Wang, W.; Itoh, S.; Yamamoto, N.; Okawa, A.; Nagai, A.; Yamashita, K. Electrical polarization of β-tricalcium phosphate ceramics. J. Am. Ceram. Soc. 2010, 93, 2175–2177. [Google Scholar] [CrossRef]
  369. Suresh, S. Theoretical studies of solid state dielectric parameters of hydroxyapatite. Mater. Phys. Mech. 2012, 14, 145–151. [Google Scholar]
  370. Singh, B.; Kumar, S.; Basu, B.; Gupta, R. Enhanced ionic conduction in hydroxyapatites. Mater. Lett. 2013, 95, 100–102. [Google Scholar] [CrossRef]
  371. Nagai, M.; Nishino, T. Surface conduction of porous hydroxyapatite ceramics at elevated temperatures. Solid State Ionics 1988, 28–30, 1456–1461. [Google Scholar] [CrossRef]
  372. Yamashita, K.; Owada, H.; Umegaki, T.; Kanazawa, T.; Futagamu, T. Ionic conduction in apatite solid solutions. Solid State Ionics 1988, 28–30, 660–663. [Google Scholar] [CrossRef]
  373. Mahabole, M.P.; Mene, R.U.; Khairnar, R.S. Gas sensing and dielectric studies on cobalt doped hydroxyapatite thick films. Adv. Mater. Lett. 2013, 4, 46–52. [Google Scholar]
  374. Mahabole, M.P.; Aiyer, R.C.; Ramakrishna, C.V.; Sreedhar, B.; Khairnar, R.S. Synthesis, characterization and gas sensing property of hydroxyapatite ceramic. Bull. Mater. Sci. 2005, 28, 535–545. [Google Scholar] [CrossRef]
  375. Fanovich, M.A.; Castro, M.S.; Lopez, J.M.P. Analysis of the microstructural evolution in hydroxyapatite ceramics by electrical characterisation. Ceram. Int. 1999, 25, 517–522. [Google Scholar] [CrossRef]
  376. Valdes, J.J.P.; Rodriguez, A.V.; Carrio, J.G. Dielectric properties and structure of hydroxyapatite ceramics sintered by different conditions. J. Mater. Res. 1995, 10, 2174–2177. [Google Scholar] [CrossRef]
  377. Bensaoud, A.; Bouhaouss, A.; Ferhat, M. Electrical properties in compressed poorly crystalline apatite. J. Solid State Electrochem. 2001, 5, 362–365. [Google Scholar] [CrossRef]
  378. Tanaka, Y.; Nakamura, M.; Nagai, A.; Toyama, T.; Yamashita, K. Ionic conduction mechanism in Ca-deficient hydroxyapatite whiskers. Mater. Sci. Eng. B 2009, 161, 115–119. [Google Scholar] [CrossRef]
  379. Tanaka, Y.; Takata, S.; Shimoe, K.; Nakamura, M.; Nagai, A.; Toyama, T.; Yamashita, K. Conduction properties of non-stoichiometric hydroxyapatite whiskers for biomedical use. J. Ceram. Soc. Jpn. 2008, 116, 815–821. [Google Scholar] [CrossRef]
  380. Nakamura, S.; Takeda, H.; Yamashita, K. Proton transport polarization and depolarization of hydroxyapatite ceramics. J. Appl. Phys. 2001, 89, 5386–5392. [Google Scholar] [CrossRef]
  381. Gittings, J.P.; Bowen, C.R.; Turner, I.G.; Baxter, F.R.; Chaudhuri, J.B. Polarisation behaviour of calcium phosphate based ceramics. Mater. Sci. Forum 2008, 587–588, 91–95. [Google Scholar] [CrossRef]
  382. Itoh, S.; Nakamura, S.; Kobayashi, T.; Shinomiya, K.; Yamashita, K.; Itoh, S. Effect of electrical polarization of hydroxyapatite ceramics on new bone formation. Calcif. Tissue Int. 2006, 78, 133–142. [Google Scholar] [CrossRef] [PubMed]
  383. Iwasaki, T.; Tanaka, Y.; Nakamura, M.; Nagai, A.; Hashimoto, K.; Toda, Y.; Katayama, K.; Yamashita, K. Rate of bonelike apatite formation accelerated on polarized porous hydroxyapatite. J. Am. Ceram. Soc. 2008, 91, 3943–3949. [Google Scholar] [CrossRef]
  384. Itoh, S.; Nakamura, S.; Kobayashi, T.; Shinomiya, K.; Yamashita, K. Enhanced bone ingrowth into hydroxyapatite with interconnected pores by electrical polarization. Biomaterials 2006, 27, 5572–5579. [Google Scholar] [CrossRef] [PubMed]
  385. Kobayashi, T.; Itoh, S.; Nakamura, S.; Nakamura, M.; Shinomiya, K.; Yamashita, K. Enhanced bone bonding of hydroxyapatite-coated titanium implants by electrical polarization. J. Biomed. Mater. Res. A 2007, 82, 145–151. [Google Scholar] [CrossRef] [PubMed]
  386. Bodhak, S.; Bose, S.; Bandyopadhyay, A. Role of surface charge and wettability on early stage mineralization and bone cell-materials interactions of polarized hydroxyapatite. Acta Biomater. 2009, 5, 2178–2188. [Google Scholar]
  387. Sagawa, H.; Itoh, S.; Wang, W.; Yamashita, K. Enhanced bone bonding of the hydroxyapatite/β-tricalcium phosphate composite by electrical polarization in rabbit long bone. Artif. Organs 2010, 34, 491–497. [Google Scholar] [CrossRef] [PubMed]
  388. Ohba, S.; Wang, W.; Itoh, S.; Nagai, A.; Yamashita, K. Enhanced effects of new bone formation by an electrically polarized hydroxyapatite microgranule/platelet-rich plasma composite gel. Key Eng. Mater. 2013, 529–530, 82–87. [Google Scholar]
  389. Yamashita, K.; Oikawa, N.; Umegaki, T. Acceleration and deceleration of bone-like crystal growth on ceramic hydroxyapatite by electric poling. Chem. Mater. 1996, 8, 2697–2700. [Google Scholar] [CrossRef]
  390. Teng, N.C.; Nakamura, S.; Takagi, Y.; Yamashita, Y.; Ohgaki, M.; Yamashita, K. A new approach to enhancement of bone formation by electrically polarized hydroxyapatite. J. Dent. Res. 2001, 80, 1925–1929. [Google Scholar] [CrossRef] [PubMed]
  391. Kobayashi, T.; Nakamura, S.; Yamashita, K. Enhanced osteobonding by negative surface charges of electrically polarized hydroxyapatite. J. Biomed. Mater. Res. 2001, 57, 477–484. [Google Scholar] [CrossRef] [PubMed]
  392. Park, Y.J.; Hwang, K.S.; Song, J.E.; Ong, J.L.; Rawls, H.R. Growth of calcium phosphate on poling treated ferroelectric BaTiO3 ceramics. Biomaterials 2002, 23, 3859–3864. [Google Scholar] [CrossRef] [PubMed]
  393. Hwang, K.S.; Song, J.E.; Jo, J.W.; Yang, H.S.; Park, Y.J.; Ong, J.L.; Rawls, H.R. Effect of poling conditions on growth of calcium phosphate crystal in ferroelectric BaTiO3 ceramics. J. Mater. Sci. Mater. Med. 2002, 13, 133–138. [Google Scholar] [CrossRef] [PubMed]
  394. Yamashita, K. Enhanced bioactivity of electrically poled hydroxyapatita ceramics and coatings. Mater. Sci. Forum 2003, 426–432, 3237–3242. [Google Scholar] [CrossRef]
  395. Nakamura, S.; Kobayashi, T.; Yamashita, K. Highly orientated calcification in newly formed bones on negatively charged hydroxyapatite electrets. Key Eng. Mater. 2005, 284–286, 897–900. [Google Scholar] [CrossRef]
  396. Kato, R.; Nakamura, S.; Katayama, K.; Yamashita, K. Electrical polarization of plasma-spray-hydroxyapatite coatings for improvement of osteoconduction of implants. J. Biomed. Mater. Res. A 2005, 74, 652–658. [Google Scholar] [CrossRef] [PubMed]
  397. Nakamura, S.; Kobayashi, T.; Nakamura, M.; Itoh, S.; Yamashita, K. Electrostatic surface charge acceleration of bone ingrowth of porous hydroxyapatite/β-tricalcium phosphate ceramics. J. Biomed. Mater. Res. A 2010, 92, 267–275. [Google Scholar] [CrossRef] [PubMed]
  398. Tarafder, S.; Bodhak, S.; Bandyopadhyay, A.; Bose, S. Effect of electrical polarization and composition of biphasic calcium phosphates on early stage osteoblast interactions. J. Biomed. Mater. Res. B 2011, 97, 306–314. [Google Scholar] [CrossRef]
  399. Ohba, S.; Wang, W.; Itoh, S.; Takagi, Y.; Nagai, A.; Yamashita, K. Acceleration of new bone formation by an electrically polarized hydroxyapatite microgranule/platelet-rich plasma composite. Acta Biomater. 2012, 8, 2778–2787. [Google Scholar] [CrossRef] [PubMed]
  400. Tarafder, S.; Banerjee, S.; Bandyopadhyay, A.; Bose, S. Electrically polarized biphasic calcium phosphates: Adsorption and release of bovine serum albumin. Langmuir 2010, 26, 16625–16629. [Google Scholar] [CrossRef] [PubMed]
  401. Itoh, S.; Nakamura, S.; Nakamura, M.; Shinomiya, K.; Yamashita, K. Enhanced bone regeneration by electrical polarization of hydroxyapatite. Artif. Organs 2006, 30, 863–869. [Google Scholar] [CrossRef] [PubMed]
  402. Nakamura, M.; Nagai, A.; Ohashi, N.; Tanaka, Y.; Sekilima, Y.; Nakamura, S. Regulation of osteoblast-like cell behaviors on hydroxyapatite by electrical polarization. Key Eng. Mater. 2008, 361–363, 1055–1058. [Google Scholar] [CrossRef]
  403. Nakamura, M.; Nagai, A.; Tanaka, Y.; Sekilima, Y.; Yamashita, K. Polarized hydroxyapatite promotes spread and motility of osteoblastic cells. J. Biomed. Mater. Res. A 2010, 92, 783–790. [Google Scholar] [PubMed]
  404. Nakamura, M.; Nagai, A.; Yamashita, K. Surface electric fields of apatite electret promote osteoblastic responses. Key Eng. Mater. 2013, 529–530, 357–360. [Google Scholar]
  405. Nakamura, S.; Kobayashi, T.; Yamashita, K. Extended bioactivity in the proximity of hydroxyapatite ceramic surfaces induced by polarization charges. J. Biomed. Mater. Res. 2002, 61, 593–599. [Google Scholar] [CrossRef] [PubMed]
  406. Wang, W.; Itoh, S.; Tanaka, Y.; Nagai, A.; Yamashita, K. Comparison of enhancement of bone ingrowth into hydroxyapatite ceramics with highly and poorly interconnected pores by electrical polarization. Acta Biomater. 2009, 5, 3132–3140. [Google Scholar] [CrossRef] [PubMed]
  407. Nagai, A.; Tanaka, K.; Tanaka, Y.; Nakamura, M.; Hashimoto, K.; Yamashita, K. Electric polarization and mechanism of B-type carbonated apatite ceramics. J. Biomed. Mater. Res. A 2011, 99, 116–124. [Google Scholar] [CrossRef] [PubMed]
  408. Nakamura, M.; Niwa, K.; Nakamura, S.; Sekijima, Y.; Yamashita, K. Interaction of a blood coagulation factor on electrically polarized hydroxyapatite surfaces. J. Biomed. Mater. Res. B 2007, 82, 29–36. [Google Scholar] [CrossRef]
  409. Nagai, M.; Shibuya, Y.; Nishino, T.; Saeki, T.; Owada, H.; Yamashita, K.; Umegaki, T. Electrical conductivity of calcium phosphate ceramics with various Ca/P ratios. J. Mater. Sci. 1991, 26, 2949–2953. [Google Scholar] [CrossRef]
  410. Laghzizil, A.; Elherch, N.; Bouhaouss, A.; Lorente, G.; Coradin, T.; Livage, J. Electrical behavior of hydroxyapatites M10(PO4)6(OH)2 (M = Ca, Pb, Ba). Mater. Res. Bull. 2001, 36, 953–962. [Google Scholar] [CrossRef]
  411. Louati, B.; Guidara, K.; Gargouri, M. Dielectric and ac ionic conductivity investigations in the monetite. J. Alloys Compd. 2009, 472, 347–351. [Google Scholar] [CrossRef]
  412. Gittings, J.P.; Bowen, C.R.; Dent, A.C.; Turner, I.G.; Baxter, F.R.; Chaudhuri, J.B. Electrical characterization of hydroxyapatite-based bioceramics. Acta Biomater. 2009, 5, 743–754. [Google Scholar] [CrossRef] [PubMed]
  413. Tofail, S.A.M.; Baldisserri, C.; Haverty, D.; McMonagle, J.B.; Erhart, J. Pyroelectric surface charge in hydroxyapatite ceramics. J. Appl. Phys. 2009, 106, 106104. [Google Scholar] [CrossRef]
  414. Ioku, K. Tailored bioceramics of calcium phosphates for regenerative medicine. J. Ceram. Soc. Jpn. 2010, 118, 775–783. [Google Scholar] [CrossRef]
  415. Fang, Y.; Agrawal, D.K.; Roy, D.M.; Roy, R. Fabrication of transparent hydroxyapatite ceramics by ambient-pressure sintering. Mater. Lett. 1995, 23, 147–151. [Google Scholar] [CrossRef]
  416. Kotobuki, N.; Kawagoe, D.; Fujimori, H.; Goto, S.; Loku, K.; Ohgushi, H. In vitro osteogenic activity of rat bone marrow derived mesenchymal stem cells cultured on transparent hydroxyapatite ceramics. Key Eng. Mater. 2004, 254–256, 1055–1058. [Google Scholar] [CrossRef]
  417. Kotobuki, N.; Ioku, K.; Kawagoe, D.; Nomura, D.; Fujimori, H.; Goto, S.; Ohgushi, H. In vitro osteogenic activity of rat mesenchymal cells cultured on transparent β-tricalcium phosphate ceramics. Key Eng. Mater. 2005, 284–286, 663–666. [Google Scholar] [CrossRef]
  418. Varma, H.; Vijayan, S.P.; Babu, S.S. Transparent hydroxyapatite ceramics through gel-casting and low-temperature sintering. J. Am. Ceram. Soc. 2002, 85, 493–495. [Google Scholar] [CrossRef]
  419. Watanabe, Y.; Ikoma, T.; Monkawa, A.; Suetsugu, Y.; Yamada, H.; Tanaka, J.; Moriyoshi, Y. Fabrication of transparent hydroxyapatite sintered body with high crystal orientation by pulse electric current sintering. J. Am. Ceram. Soc. 2005, 88, 243–245. [Google Scholar] [CrossRef]
  420. Kotobuki, N.; Ioku, K.; Kawagoe, D.; Fujimori, H.; Goto, S.; Ohgushi, H. Observation of osteogenic differentiation cascade of living mesenchymal stem cells on transparent hydroxyapatite ceramics. Biomaterials 2005, 26, 779–785. [Google Scholar] [CrossRef] [PubMed]
  421. John, A.; Varma, H.K.; Vijayan, S.; Bernhardt, A.; Lode, A.; Vogel, A.; Burmeister, B.; Hanke, T.; Domaschke, H.; Gelinsky, M. In vitro investigations of bone remodeling on a transparent hydroxyapatite ceramic. Biomed. Mater. 2009, 4. [Google Scholar] [CrossRef]
  422. Kotobuki, N.; Kawagoe, D.; Nomura, D.; Katou, Y.; Muraki, K.; Fujimori, H.; Goto, S.; Ioku, K.; Ohgushi, H. Observation and quantitative analysis of rat bone marrow stromal cells cultured in vitro on newly formed transparent β-tricalcium phosphate. J. Mater. Sci. Mater. Med. 2006, 17, 33–41. [Google Scholar] [CrossRef] [PubMed]
  423. Wang, J.; Shaw, L.L. Transparent nanocrystalline hydroxyapatite by pressure-assisted sintering. Scripta Mater. 2010, 63, 593–596. [Google Scholar] [CrossRef]
  424. Tan, N.; Kou, Z.; Ding, Y.; Leng, Y.; Liu, C.; He, D. Novel substantial reductions in sintering temperatures for preparation of transparent hydroxyapatite bioceramics under ultrahigh pressure. Scr. Mater. 2011, 65, 819–822. [Google Scholar] [CrossRef]
  425. Kobune, M.; Mineshige, A.; Fujii, S.; Iida, H. Preparation of translucent hydroxyapatite ceramics by HIP and their physical properties. J. Ceram. Soc. Jpn. 1997, 105, 210–213. [Google Scholar] [CrossRef]
  426. Barralet, J.E.; Fleming, G.J.P.; Campion, C.; Harris, J.J.; Wright, A.J. Formation of translucent hydroxyapatite ceramics by sintering in carbon dioxide atmospheres. J. Mater. Sci. 2003, 38, 3979–3993. [Google Scholar] [CrossRef]
  427. Chaudhry, A.A.; Yan, H.; Gong, K.; Inam, F.; Viola, G.; Reece, M.J.; Goodall, J.B.M.; ur Rehman, I.; McNeil-Watson, F.K.; Corbett, J.C.W.; Knowles, J.C; Darr, J.A. High-strength nanograined and translucent hydroxyapatite monoliths via continuous hydrothermal synthesis and optimized spark plasma sintering. Acta Biomater. 2011, 7, 791–799. [Google Scholar] [CrossRef] [PubMed]
  428. Tancred, D.C.; McCormack, B.A.; Carr, A.J. A synthetic bone implant macroscopically identical to cancellous bone. Biomaterials 1998, 19, 2303–2311. [Google Scholar] [CrossRef] [PubMed]
  429. Miao, X.; Sun, D. Graded/gradient porous biomaterials. Materials 2010, 3, 26–47. [Google Scholar] [CrossRef]
  430. Schliephake, H.; Neukam, F.W.; Klosa, D. Influence of pore dimensions on bone ingrowth into porous hydroxylapatite blocks used as bone graft substitutes. A histometric study. Int. J. Oral Maxillofac. Surg. 1991, 20, 53–58. [Google Scholar] [CrossRef] [PubMed]
  431. Otsuki, B.; Takemoto, M.; Fujibayashi, S.; Neo, M.; Kokubo, T.; Nakamura, T. Pore throat size and connectivity determine bone and tissue ingrowth into porous implants: Three-dimensional micro-CT based structural analyses of porous bioactive titanium implants. Biomaterials 2006, 27, 5892–5900. [Google Scholar] [CrossRef] [PubMed]
  432. Gauthier, O.; Bouler, J.M.; Weiss, P.; Bosco, J.; Daculsi, G.; Aguado, E. Kinetic study of bone ingrowth and ceramic resorption associated with the implantation of different injectable calcium-phosphate bone substitutes. J. Biomed. Mater. Res. 1999, 47, 28–35. [Google Scholar] [CrossRef] [PubMed]
  433. Hing, K.A.; Best, S.M.; Bonfield, W. Characterization of porous hydroxyapatite. J. Mater. Sci. Mater. Med. 1999, 10, 135–145. [Google Scholar] [CrossRef] [PubMed]
  434. Carotenuto, G.; Spagnuolo, G.; Ambrosio, L.; Nicolais, L. Macroporous hydroxyapatite as alloplastic material for dental applications. J. Mater. Sci. Mater. Med. 1999, 10, 671–676. [Google Scholar] [CrossRef] [PubMed]
  435. Bucholz, R.W.; Carlton, A.; Holmes, R. Interporous hydroxyapatite as a bone graft substitute in tibial plateau fractures. Clin. Orthop. 1989, 240, 53–62. [Google Scholar] [PubMed]
  436. Cavagna, R.; Daculsi, G.; Bouler, J.M. Macroporous calcium phosphate ceramic: A prospective study of 106 cases in lumbar spinal fusion. J. Long Term Eff. Med. Implant. 1999, 9, 403–412. [Google Scholar]
  437. Lu, J.X.; Flautre, B.; Anselme, K.; Hardouin, P.; Gallur, A.; Descamps, M.; Thierry, B. Role of interconnections in porous bioceramics on bone recolonization in vitro and in vivo. J. Mater. Sci. Mater. Med. 1999, 10, 111–120. [Google Scholar] [CrossRef] [PubMed]
  438. Ayers, R.A.; Simske, S.J.; Nunes, C.R.; Wolford, L.M. Long-term bone ingrowth and residual microhardness of porous block hydroxyapatite implants in humans. J. Oral Maxillof. Surg. 1998, 56, 1297–1302. [Google Scholar] [CrossRef]
  439. Jones, A.C.; Arns, C.H.; Sheppard, A.P.; Hutmacher, D.W.; Milthorpe, B.K.; Knackstedt, M.A. Assessment of bone ingrowth into porous biomaterials using MICRO-CT. Biomaterials 2007, 28, 2491–2504. [Google Scholar] [CrossRef] [PubMed]
  440. Karageorgiou, V.; Kaplan, D. Porosity of 3D biomaterial scaffolds and osteogenesis. Biomaterials 2005, 26, 5474–5491. [Google Scholar] [CrossRef] [PubMed]
  441. Tamai, N.; Myoui, A.; Kudawara, I.; Ueda, T.; Yoshikawa, H. Novel fully interconnected porous hydroxyapatite ceramic in surgical treatment of benign bone tumor. J. Orthop. Sci. 2010, 15, 560–568. [Google Scholar] [CrossRef] [PubMed]
  442. Panzavolta, S.; Torricelli, P.; Amadori, S.; Parrilli, A.; Rubini, K.; Della Bella, E.; Fini, M.; Bigi, A. 3D interconnected porous biomimetic scaffolds: in vitro cell response. J. Biomed. Mater. Res. A 2013. [Google Scholar] [CrossRef]
  443. Ohgushi, H.; Goldberg, V.M.; Caplan, A.I. Heterotopic osteogenesis in porous ceramics induced by marrow cells. J. Orthop. Res. 1989, 7, 568–578. [Google Scholar] [CrossRef] [PubMed]
  444. Cheung, H.S.; Haak, M.H. Growth of osteoblasts on porous calcium phosphate ceramic: An in vitro model for biocompatibility study. Biomaterials 1989, 10, 63–67. [Google Scholar] [CrossRef] [PubMed]
  445. Zyman, Z.; Ivanov, I.; Glushko, V.; Dedukh, N.; Malyshkina, S. Inorganic phase composition of remineralisation in porous CaP ceramics. Biomaterials 1998, 19, 1269–1273. [Google Scholar] [CrossRef] [PubMed]
  446. Yoshikawa, T.; Ohgushi, H.; Tamai, S. Immediate bone forming capability of prefabricated osteogenic hydroxyapatite. J. Biomed. Mater. Res. 1996, 32, 481–492. [Google Scholar] [CrossRef] [PubMed]
  447. Chang, B.S.; Lee, C.K.; Hong, K.S.; Youn, H.J.; Ryu, H.S.; Chung, S.S.; Park, K. Osteoconduction at porous hydroxyapatite with various pore configurations. Biomaterials 2000, 21, 1291–1298. [Google Scholar] [CrossRef] [PubMed]
  448. Flautre, B.; Descamps, M.; Delecourt, C.; Blary, M.C.; Hardouin, P. Porous HA ceramic for bone replacement: Role of the pores and interconnections—Experimental study in the rabbits. J. Mater. Sci. Mater. Med. 2001, 12, 679–682. [Google Scholar] [CrossRef] [PubMed]
  449. McAfee, P.C.; Cunningham, B.W.; Orbegoso, D.O.; Sefter, J.C.; Dmitriev, A.E.; Fedder, I.L. Analysis of porous ingrowth in intervertebral disc prostheses. Spine 2003, 28, 332–340. [Google Scholar] [PubMed]
  450. Tamai, N.; Myoui, A.; Tomita, T.; Nakase, T.; Tanaka, J.; Ochi, T.; Yoshikawa, H. Novel hydroxyapatite ceramics with an interconnective porous structure exhibit superior osteoconduction in vivo. J. Biomed. Mater. Res. 2002, 59, 110–117. [Google Scholar] [CrossRef] [PubMed]
  451. Mastrogiacomo, M.; Scaglione, S.; Martinetti, R.; Dolcini, L.; Beltrame, F.; Cancedda, R.; Quarto, R. Role of scaffold internal structure on in vivo bone formation in macroporous calcium phosphate bioceramics. Biomaterials 2006, 27, 3230–3237. [Google Scholar] [CrossRef] [PubMed]
  452. Okamoto, M.; Dohi, Y.; Ohgushi, H.; Shimaoka, H.; Ikeuchi, M.; Matsushima, A.; Yonemasu, K.; Hosoi, H. Influence of the porosity of hydroxyapatite ceramics on in vitro and in vivo bone formation by cultured rat bone marrow stromal cells. J. Mater. Sci. Mater. Med. 2006, 17, 327–336. [Google Scholar] [CrossRef] [PubMed]
  453. Zhang, L.; Hanagata, N.; Maeda, M.; Minowa, T.; Ikoma, T.; Fan, H.; Zhang, X. Porous hydroxyapatite and biphasic calcium phosphate ceramics promote ectopic osteoblast differentiation from mesenchymal stem cells. Sci. Technol. Adv. Mater. 2009, 10. [Google Scholar] [CrossRef]
  454. Li, X.; Liu, H.; Niu, X.; Fan, Y.; Feng, Q.; Cui, F.Z.; Watari, F. Osteogenic differentiation of human adipose-derived stem cells induced by osteoinductive calcium phosphate ceramics. J. Biomed. Mater. Res. B 2011, 97, 10–19. [Google Scholar] [CrossRef]
  455. Omae, H.; Mochizuki, Y.; Yokoya, S.; Adachi, N.; Ochi, M. Effects of interconnecting porous structure of hydroxyapatite ceramics on interface between grafted tendon and ceramics. J. Biomed. Mater. Res. A 2006, 79, 329–337. [Google Scholar] [CrossRef] [PubMed]
  456. Yoshikawa, H.; Tamai, N.; Murase, T.; Myoui, A. Interconnected porous hydroxyapatite ceramics for bone tissue engineering. J. R. Soc. Interface 2009, 6, S341–S348. [Google Scholar] [CrossRef] [PubMed]
  457. Ishikawa, K. Bone substitute fabrication based on dissolution-precipitation reactions. Materials 2010, 3, 1138–1155. [Google Scholar] [CrossRef]
  458. Ribeiro, G.B.M.; Trommer, R.M.; dos Santos, L.A.; Bergmann, C.P. Novel method to produce β-TCP scaffolds. Mater. Lett. 2011, 65, 275–277. [Google Scholar] [CrossRef]
  459. Silva, T.S.N.; Primo, B.T.; Silva, A.N., Jr.; Machado, D.C.; Viezzer, C.; Santos, L.A. Use of calcium phosphate cement scaffolds for bone tissue engineering: in vitro study. Acta Cir. Bras. 2011, 26, 7–11. [Google Scholar] [CrossRef] [PubMed]
  460. de Moraes MacHado, J.L.; Giehl, I.C.; Nardi, N.B.; dos Santos, L.A. Evaluation of scaffolds based on α-tricalcium phosphate cements for tissue engineering applications. IEEE Trans. Biomed. Eng. 2011, 58, 1814–1819. [Google Scholar] [CrossRef] [PubMed]
  461. Li, S.H.; de Wijn, J.R.; Layrolle, P.; de Groot, K. Novel method to manufacture porous hydroxyapatite by dual-phase mixing. J. Am. Ceram. Soc. 2003, 86, 65–72. [Google Scholar] [CrossRef]
  462. de Oliveira, J.F.; de Aguiar, P.F.; Rossi, A.M.; Soares, G.D.A. Effect of process parameters on the characteristics of porous calcium phosphate ceramics for bone tissue scaffolds. Artif. Organs 2003, 27, 406–411. [Google Scholar] [CrossRef] [PubMed]
  463. Swain, S.K.; Bhattacharyya, S. Preparation of high strength macroporous hydroxyapatite scaffold. Mater. Sci. Eng. C 2013, 33, 67–71. [Google Scholar] [CrossRef]
  464. Maeda, H.; Kasuga, T.; Nogami, M.; Kagami, H.; Hata, K.; Ueda, M. Preparation of bonelike apatite composite sponge. Key Eng. Mater. 2004, 254–256, 497–500. [Google Scholar] [CrossRef]
  465. Herliansyah, M.K.; Suyitno, Dewo, P.; bin Abdul Shukor, M.H.; Ide-Ektessabi, A. Development and characterization of bovine hydroxyapatite porous bone graft for biomedical applications. Adv. Mater. Res. 2011, 277, 59–65. [Google Scholar] [CrossRef]
  466. Li, S.H.; de Wijn, J.R.; Layrolle, P.; de Groot, K. Synthesis of macroporous hydroxyapatite scaffolds for bone tissue engineering. J. Biomed. Mater. Res. 2002, 61, 109–120. [Google Scholar] [CrossRef] [PubMed]
  467. Hesaraki, S.; Sharifi, D. Investigation of an effervescent additive as porogenic agent for bone cement macroporosity. Biomed. Mater. Eng. 2007, 17, 29–38. [Google Scholar] [PubMed]
  468. Hesaraki, S.; Moztarzadeh, F.; Sharifi, D. Formation of interconnected macropores in apatitic calcium phosphate bone cement with the use of an effervescent additive. J. Biomed. Mater. Res. A 2007, 83, 80–87. [Google Scholar] [CrossRef] [PubMed]
  469. Pal, K.; Pal, S. Development of porous hydroxyapatite scaffolds. Mater. Manuf. Process. 2006, 21, 325–328. [Google Scholar] [CrossRef]
  470. Tas, A.C. Preparation of porous apatite granules from calcium phosphate cement. J. Mater. Sci. Mater. Med. 2008, 19, 2231–2239. [Google Scholar] [CrossRef] [PubMed]
  471. Yao, X.; Tan, S.; Jiang, D. Improving the properties of porous hydroxyapatite ceramics by fabricating methods. J. Mater. Sci. 2005, 40, 4939–4942. [Google Scholar] [CrossRef]
  472. Song, H.Y.; Youn, M.H.; Kim, Y.H.; Min, Y.K.; Yang, H.M.; Lee, B.T. Fabrication of porous β-TCP bone graft substitutes using PMMA powder and their biocompatibility study. Korean J. Mater. Res. 2007, 17, 318–322. [Google Scholar] [CrossRef]
  473. Youn, M.H.; Paul, R.K.; Song, H.Y.; Lee, B.T. Fabrication of porous structure of BCP sintered bodies using microwave assisted synthesized HAp nano powder. Mater. Sci. Forum 2007, 534–536, 49–52. [Google Scholar] [CrossRef]
  474. Almirall, A.; Larrecq, G.; Delgado, J.A.; Martínez, S.; Ginebra, M.P.; Planell, J.A. Fabrication of low temperature hydroxyapatite foams. Key Eng. Mater. 2004, 254–256, 1001–1004. [Google Scholar] [CrossRef]
  475. Almirall, A.; Larrecq, G.; Delgado, J.A.; Martínez, S.; Planell, J.A.; Ginebra, M.P. Fabrication of low temperature macroporous hydroxyapatite scaffolds by foaming and hydrolysis of an α-TCP paste. Biomaterials 2004, 25, 3671–3680. [Google Scholar] [CrossRef] [PubMed]
  476. Huang, X.; Miao, X. Novel porous hydroxyapatite prepared by combining H2O2 foaming with PU sponge and modified with PLGA and bioactive glass. J. Biomater. Appl. 2007, 21, 351–374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  477. Strnadova, M.; Protivinsky, J.; Strnad, J.; Vejsicka, Z. Preparation of porous synthetic nanostructured HA scaffold. Key Eng. Mater. 2008, 361–363, 211–214. [Google Scholar] [CrossRef]
  478. Li, B.; Chen, X.; Guo, B.; Wang, X.; Fan, H.; Zhang, X. Fabrication and cellular biocompatibility of porous carbonated biphasic calcium phosphate ceramics with a nanostructure. Acta Biomater. 2009, 5, 134–143. [Google Scholar] [CrossRef] [PubMed]
  479. Takagi, S.; Chow, L.C. Formation of macropores in calcium phosphate cement implants. J. Biomed. Mater. Res. 2001, 12, 135–139. [Google Scholar]
  480. Walsh, D.; Tanaka, J. Preparation of a bone-like apatite foam cement. J. Mater. Sci. Mater. Med. 2001, 12, 339–344. [Google Scholar] [CrossRef] [PubMed]
  481. Tadic, D.; Beckmann, F.; Schwarz, K.; Epple, M. A novel method to produce hydroxylapatite objects with interconnecting porosity that avoids sintering. Biomaterials 2004, 25, 3335–3340. [Google Scholar] [CrossRef] [PubMed]
  482. Komlev, V.S.; Barinov, S.M. Porous hydroxyapatite ceramics of bi-modal pore size distribution. J. Mater. Sci. Mater. Med. 2002, 13, 295–299. [Google Scholar] [CrossRef] [PubMed]
  483. Sepulveda, P.; Binner, J.G.; Rogero, S.O.; Higa, O.Z.; Bressiani, J.C. Production of porous hydroxyapatite by the gel-casting of foams and cytotoxic evaluation. J. Biomed. Mater. Res. 2000, 50, 27–34. [Google Scholar] [CrossRef] [PubMed]
  484. Hsu, Y.H.; Turner, I.G.; Miles, A.W. Mechanical characterization of dense calcium phosphate bioceramics with interconnected porosity. J. Mater. Sci. Mater. Med. 2007, 18, 2319–2329. [Google Scholar] [CrossRef] [PubMed]
  485. Zhang, H.G.; Zhu, Q. Preparation of porous hydroxyapatite with interconnected pore architecture. J. Mater. Sci. Mater. Med. 2007, 18, 1825–1829. [Google Scholar] [CrossRef] [PubMed]
  486. Chevalier, E.; Chulia, D.; Pouget, C.; Viana, M. Fabrication of porous substrates: A review of processes using pore forming agents in the biomaterial field. J. Pharm. Sci. 2008, 97, 1135–1154. [Google Scholar] [CrossRef] [PubMed]
  487. Tang, Y.J.; Tang, Y.F.; Lv, C.T.; Zhou, Z.H. Preparation of uniform porous hydroxyapatite biomaterials by a new method. Appl. Surf. Sci. 2008, 254, 5359–5362. [Google Scholar] [CrossRef]
  488. le Ray, A.M.; Gautier, H.; Bouler, J.M.; Weiss, P.; Merle, C. A new technological procedure using sucrose as porogen compound to manufacture porous biphasic calcium phosphate ceramics of appropriate micro- and macrostructure. Ceram. Int. 2010, 36, 93–101. [Google Scholar] [CrossRef] [Green Version]
  489. Abdulqader, S.T.; Rahman, I.A.; Ismail, H.; Ponnuraj Kannan, T.; Mahmood, Z. A simple pathway in preparation of controlled porosity of biphasic calcium phosphate scaffold for dentin regeneration. Ceram. Int. 2013, 39, 2375–2381. [Google Scholar] [CrossRef]
  490. Stares, S.L.; Fredel, M.C.; Greil, P.; Travitzky, N. Paper-derived hydroxyapatite. Ceram. Int. 2013, 39, 7179–7183. [Google Scholar] [CrossRef]
  491. Guda, T.; Appleford, M.; Oh, S.; Ong, J.L. A cellular perspective to bioceramic scaffolds for bone tissue engineering: The state of the art. Curr. Top. Med. Chem. 2008, 8, 290–299. [Google Scholar] [CrossRef] [PubMed]
  492. Tian, J.; Tian, J. Preparation of porous hydroxyapatite. J. Mater. Sci. 2001, 36, 3061–3066. [Google Scholar] [CrossRef]
  493. Swain, S.K.; Bhattacharyya, S.; Sarkar, D. Preparation of porous scaffold from hydroxyapatite powders. Mater. Sci. Eng. C 2011, 31, 1240–1244. [Google Scholar] [CrossRef]
  494. Zhao, K.; Tang, Y.F.; Qin, Y.S.; Luo, D.F. Polymer template fabrication of porous hydroxyapatite scaffolds with interconnected spherical pores. J. Eur. Ceram. Soc. 2011, 31, 225–229. [Google Scholar] [CrossRef]
  495. Sung, J.H.; Shin, K.H.; Moon, Y.W.; Koh, Y.H.; Choi, W.Y.; Kim, H.E. Production of porous calcium phosphate (CaP) ceramics with highly elongated pores using carbon-coated polymeric templates. Ceram. Int. 2012, 38, 93–97. [Google Scholar] [CrossRef]
  496. Rivera-Munoz, E.; Diaz, J.R.; Rodriguez, J.R.; Brostow, W.; Castano, V.M. Hydroxyapatite spheres with controlled porosity for eye ball prosthesis: Processing and characterization. J. Mater. Sci. Mater. Med. 2001, 12, 305–311. [Google Scholar] [CrossRef] [PubMed]
  497. Cyster, L.A.; Grant, D.M.; Howdle, S.M.; Rose, F.R.A.J.; Irvine, D.J.; Freeman, D.; Scotchford, C.A.; Shakesheff, K.M. The influence of dispersant concentration on the pore morphology of hydroxyapatite ceramics for bone tissue engineering. Biomaterials 2005, 26, 697–702. [Google Scholar] [CrossRef] [PubMed]
  498. Deville, S.; Saiz, E.; Tomsia, A.P. Freeze casting of hydroxyapatite scaffolds for bone tissue engineering. Biomaterials 2006, 27, 5480–5489. [Google Scholar] [CrossRef] [PubMed]
  499. Lee, E.J.; Koh, Y.H.; Yoon, B.H.; Kim, H.E.; Kim, H.W. Highly porous hydroxyapatite bioceramics with interconnected pore channels using camphene-based freeze casting. Mater. Lett. 2007, 61, 2270–2273. [Google Scholar] [CrossRef]
  500. Fu, Q.; Rahaman, M.N.; Dogan, F.; Bal, B.S. Freeze casting of porous hydroxyapatite scaffolds. I. Processing and general microstructure. J. Biomed. Mater. Res. B 2008, 86, 125–135. [Google Scholar] [CrossRef]
  501. Impens, S.; Schelstraete, R.; Luyten, J.; Mullens, S.; Thijs, I.; van Humbeeck, J.; Schrooten, J. Production and characterisation of porous calcium phosphate structures with controllable hydroxyapatite/β-tricalcium phosphate ratios. Adv. Appl. Ceram. 2009, 108, 494–500. [Google Scholar] [CrossRef]
  502. Macchetta, A.; Turner, I.G.; Bowen, C.R. Fabrication of HA/TCP scaffolds with a graded and porous structure using a camphene-based freeze-casting method. Acta Biomater. 2009, 5, 1319–1327. [Google Scholar] [CrossRef] [PubMed]
  503. Potoczek, M.; Zima, A.; Paszkiewicz, Z.; Ślósarczyk, A. Manufacturing of highly porous calcium phosphate bioceramics via gel-casting using agarose. Ceram. Int. 2009, 35, 2249–2254. [Google Scholar] [CrossRef]
  504. Zuo, K.H.; Zeng, Y.P.; Jiang, D. Effect of polyvinyl alcohol additive on the pore structure and morphology of the freeze-cast hydroxyapatite ceramics. Mater. Sci. Eng. C 2010, 30, 283–287. [Google Scholar] [CrossRef]
  505. Soon, Y.M.; Shin, K.H.; Koh, Y.H.; Lee, J.H.; Choi, W.Y.; Kim, H.E. Fabrication and compressive strength of porous hydroxyapatite scaffolds with a functionally graded core/shell structure. J. Eur. Ceram. Soc. 2011, 31, 13–18. [Google Scholar] [CrossRef]
  506. Ng, S.; Guo, J.; Ma, J.; Loo, S.C.J. Synthesis of high surface area mesostructured calcium phosphate particles. Acta Biomater. 2010, 6, 3772–3781. [Google Scholar] [CrossRef] [PubMed]
  507. Ota, Y.; Kasuga, T.; Abe, Y. Preparation and compressive strength behaviour of porous ceramics with β-Ca3(PO3)2 fiber skeletons. J. Am. Ceram. Soc. 1997, 80, 225–231. [Google Scholar] [CrossRef]
  508. White, E.; Shors, E.C. Biomaterial aspects of Interpore-200 porous hydroxyapatite. Dent. Clin. North Am. 1986, 30, 49–67. [Google Scholar] [PubMed]
  509. Liu, D.M. Fabrication of hydroxyapatite ceramic with controlled porosity. J. Mater. Sci. Mater. Med. 1997, 8, 227–232. [Google Scholar] [CrossRef] [PubMed]
  510. Walsh, D.; Hopwood, J.D.; Mann, S. Crystal tectonics: Construction of reticulated calcium phosphate frameworks in bicontinuous reverse microemulsions. Science 1994, 264, 1576–1578. [Google Scholar] [CrossRef] [PubMed]
  511. Walsh, D.; Mann, S. Chemical synthesis of microskeletal calcium phosphate in bicontinuous microemulsions. Chem. Mater. 1996, 8, 1944–1953. [Google Scholar] [CrossRef]
  512. Aizawa, M.; Howell, S.F.; Itatani, K.; Yokogawa, Y.; Nishizawa, K.; Toriyama, M.; Kameyama, T. Fabrication of porous ceramics with well-controlled open pores by sintering of fibrous hydroxyapatite particles. J. Ceram. Soc. Jpn. 2000, 108, 249–253. [Google Scholar] [CrossRef]
  513. Rodriguez-Lorenzo, L.M.; Vallet-Regí, M.; Ferreira, J.M.F. Fabrication of porous hydroxyapatite bodies by a new direct consolidation method: Starch consolidation. J. Biomed. Mater. Res. 2002, 60, 232–240. [Google Scholar] [CrossRef] [PubMed]
  514. Charriere, E.; Lemaitre, J.; Zysset, P. Hydroxyapatite cement scaffolds with controlled macroporosity: Fabrication protocol and mechanical properties. Biomaterials 2003, 24, 809–817. [Google Scholar] [CrossRef] [PubMed]
  515. Zhao, K.; Tang, Y.F.; Qin, Y.S.; Wei, J.Q. Porous hydroxyapatite ceramics by ice templating: Freezing characteristics and mechanical properties. Ceram. Int. 2011, 37, 635–639. [Google Scholar] [CrossRef]
  516. Zhou, K.; Zhang, Y.; Zhang, D.; Zhang, X.; Li, Z.; Liu, G.; Button, T.W. Porous hydroxyapatite ceramics fabricated by an ice-templating method. Scripta Mater. 2011, 64, 426–429. [Google Scholar] [CrossRef]
  517. Flauder, S.; Gbureck, U.; Muller, F.A. TCP scaffolds with an interconnected and aligned porosity fabricated via ice-templating. Key Eng. Mater. 2013, 529–530, 129–132. [Google Scholar]
  518. Zhang, Y.; Zhou, K.; Bao, Y.; Zhang, D. Effects of rheological properties on ice-templated porous hydroxyapatite ceramics. Mater. Sci. Eng. C 2013, 33, 340–346. [Google Scholar] [CrossRef]
  519. Eichenseer, C.; Will, J.; Rampf, M.; Wend, S.; Greil, P. Biomorphous porous hydroxyapatite-ceramics from rattan (Calamus Rotang). J. Mater. Sci. Mater. Med. 2010, 21, 131–137. [Google Scholar] [CrossRef] [PubMed]
  520. Zhou, L.; Wang, D.; Huang, W.; Yao, A.; Kamitakahara, M.; Ioku, K. Preparation and characterization of periodic porous frame of hydroxyapatite. J. Ceram. Soc. Jpn. 2009, 117, 521–524. [Google Scholar]
  521. Kawata, M.; Uchida, H.; Itatani, K.; Okada, I.; Koda, S.; Aizawa, M. Development of porous ceramics with well-controlled porosities and pore sizes from apatite fibers and their evaluations. J. Mater. Sci. Mater. Med. 2004, 15, 817–823. [Google Scholar] [CrossRef] [PubMed]
  522. Koh, Y.H.; Kim, H.W.; Kim, H.E.; Halloran, J.W. Fabrication of macrochannelled-hydroxyapatite bioceramic by a coextrusion process. J. Am. Ceram. Soc. 2002, 85, 2578–2580. [Google Scholar] [CrossRef]
  523. Nakahira, A.; Tamai, M.; Sakamoto, K.; Yamaguchi, S. Sintering and microstructure of porous hydroxyapatite. J. Ceram. Soc. Jpn. 2000, 108, 99–104. [Google Scholar] [CrossRef]
  524. de Sousa, F.C.G.; Evans, J.R.G. Tubular hydroxyapatite scaffolds. Adv. Appl. Ceram. 2005, 104, 30–34. [Google Scholar] [CrossRef]
  525. Kitamura, M.; Ohtsuki, C.; Ogata, S.I.; Kamitakahara, M.; Tanihara, M.; Miyazaki, T. Mesoporous calcium phosphate via post-treatment of α-TCP. J. Am. Ceram. Soc. 2005, 88, 822–826. [Google Scholar] [CrossRef]
  526. Walsh, D.; Boanini, E.; Tanaka, J.; Mann, S. Synthesis of tri-calcium phosphate sponges by interfacial deposition and thermal transformation of self-supporting calcium phosphate films. J. Mater. Chem. 2005, 15, 1043–1048. [Google Scholar] [CrossRef]
  527. Gonzalez-McQuire, R.; Green, D.; Walsh, D.; Hall, S.; Chane-Ching, J.Y.; Oreffo, R.O.C.; Mann, S. Fabrication of hydroxyapatite sponges by dextran sulphate/amino acid templating. Biomaterials 2005, 26, 6652–6656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  528. Sohier, J.; Daculsi, G.; Sourice, S.; de Groot, K.; Layrolle, P. Porous beta tricalcium phosphate scaffolds used as a BMP-2 delivery system for bone tissue engineering. J. Biomed. Mater. Res. A 2010, 92, 1105–1114. [Google Scholar] [PubMed]
  529. Stähli, C.; Bohner, M.; Bashoor-Zadeh, M.; Doebelin, N.; Baroud, G. Aqueous impregnation of porous β-tricalcium phosphate scaffolds. Acta Biomater. 2010, 6, 2760–2772. [Google Scholar] [CrossRef] [PubMed]
  530. Ramay, H.R.; Zhang, M. Preparation of porous hydroxyapatite scaffolds by combination of the gel-casting and polymer sponge methods. Biomaterials 2003, 24, 3293–3302. [Google Scholar] [CrossRef] [PubMed]
  531. Xu, S.; Li, D.; Lu, B.; Tang, Y.; Wang, C.; Wang, Z. Fabrication of a calcium phosphate scaffold with a three dimensional channel network and its application to perfusion culture of stem cells. Rapid Prototyp. J. 2007, 13, 99–106. [Google Scholar] [CrossRef]
  532. Saiz, E.; Gremillard, L.; Menendez, G.; Miranda, P.; Gryn, K.; Tomsia, A.P. Preparation of porous hydroxyapatite scaffolds. Mater. Sci. Eng. C 2007, 27, 546–550. [Google Scholar] [CrossRef]
  533. Kamitakahara, M.; Ohtsuki, C.; Kawachi, G.; Wang, D.; Ioku, K. Preparation of hydroxyapatite porous ceramics with different porous structures using a hydrothermal treatment with different aqueous solutions. J. Ceram. Soc. Jpn. 2008, 116, 6–9. [Google Scholar] [CrossRef]
  534. Abdurrahim, T.; Sopyan, I. Recent progress on the development of porous bioactive calcium phosphate for biomedical applications. Recent Pat. Biomed. Eng. 2008, 1, 213–229. [Google Scholar] [CrossRef]
  535. Peña, J.; Román, J.; Cabañas, M.V.; Vallet-Regí, M. An alternative technique to shape scaffolds with hierarchical porosity at physiological temperature. Acta Biomater. 2010, 6, 1288–1296. [Google Scholar] [CrossRef] [PubMed]
  536. Nakamura, S.; Nakahira, A. Synthesis and evaluation of porous hydroxyapatite prepared by hydrothermal treatment and subsequent sintering method. J. Ceram. Soc. Jpn. 2008, 116, 42–45. [Google Scholar] [CrossRef]
  537. Zhang, J.; Fujiwara, M.; Xu, Q.; Zhu, Y.; Iwasa, M.; Jiang, D. Synthesis of mesoporous calcium phosphate using hybrid templates. Micropor. Mesopor. Mater. 2008, 111, 411–416. [Google Scholar] [CrossRef]
  538. Song, H.Y.; Islam, S.; Lee, B.T. A novel method to fabricate unidirectional porous hydroxyapatite body using ethanol bubbles in a viscous slurry. J. Am. Ceram. Soc. 2008, 91, 3125–3127. [Google Scholar] [CrossRef]
  539. Kawachi, G.; Misumi, H.; Fujimori, H.; Goto, S.; Ohtsuki, C.; Kamitakahara, M.; Ioku, K. Fabrication of porous blocks of calcium phosphate through hydrothermal processing under glycine coexistence. J. Ceram. Soc. Jpn. 2010, 118, 559–563. [Google Scholar] [CrossRef]
  540. Sakamoto, M.; Nakasu, M.; Matsumoto, T.; Okihana, H. Development of superporous hydroxyapatites and their examination with a culture of primary rat osteoblasts. J. Biomed. Mater. Res. A 2007, 82A, 238–242. [Google Scholar] [CrossRef] [PubMed]
  541. Sakamoto, M. Development and evaluation of superporous hydroxyapatite ceramics with triple pore structure as bone tissue scaffold. J. Ceram. Soc. Jpn. 2010, 118, 753–757. [Google Scholar] [CrossRef]
  542. Sakamoto, M.; Matsumoto, T. Development and Evaluation of Superporous Ceramics Bone Tissue Scaffold Materials with Triple Pore Structure a) Hydroxyapatite, b) Beta-Tricalcium Phosphate. In Bone Regeneration; Tal, H., Ed.; InTech Europe: Rijeka, Croatia, 2012; pp. 301–320. [Google Scholar]
  543. Ishikawa, K.; Tsuru, K.; Pham, T.K.; Maruta, M.; Matsuya, S. Fully-interconnected pore forming calcium phosphate cement. Key Eng. Mater. 2012, 493–494, 832–835. [Google Scholar]
  544. Yoon, H.J.; Kim, U.C.; Kim, J.H.; Koh, Y.H.; Choi, W.Y.; Kim, H.E. Fabrication and characterization of highly porous calcium phosphate (CaP) ceramics by freezing foamed aqueous CaP suspensions. J. Ceram. Soc. Jpn. 2011, 119, 573–576. [Google Scholar] [CrossRef]
  545. Ahn, M.K.; Shin, K.H.; Moon, Y.W.; Koh, Y.H.; Choi, W.Y.; Kim, H.E. Highly porous biphasic calcium phosphate (BCP) ceramics with large interconnected pores by freezing vigorously foamed BCP suspensions under reduced pressure. J. Am. Ceram. Soc. 2011, 94, 4154–4156. [Google Scholar] [CrossRef]
  546. Ji, L.; Jell, G.; Dong, Y.; Jones, J.R.; Stevens, M.M. Template synthesis of ordered macroporous hydroxyapatite bioceramics. Chem. Commun. 2011, 47, 9048–9050. [Google Scholar] [CrossRef]
  547. Schlosser, M.; Kleebe, H.J. Vapor transport sintering of porous calcium phosphate ceramics. J. Am. Ceram. Soc. 2012, 95, 1581–1587. [Google Scholar] [CrossRef]
  548. Tang, Y.; Zhao, K.; Hu, L. Pore size gradient hydroxyapatite scaffolds with interconnected pores fabricated by a template method. Adv. Mater. Res. 2013, 602–604, 1219–1222. [Google Scholar]
  549. Roy, D.M.; Linnehan, S.K. Hydroxyapatite formed from coral skeletal carbonate by hydrothermal exchange. Nature 1974, 247, 220–222. [Google Scholar] [CrossRef] [PubMed]
  550. Holmes, R.E. Bone regeneration within a coralline hydroxyapatite implant. Plast. Reconstr. Surg. 1979, 63, 626–633. [Google Scholar] [CrossRef] [PubMed]
  551. Elsinger, E.C.; Leal, L. Coralline hydroxyapatite bone graft substitutes. J. Foot Ankle Surg. 1996, 35, 396–399. [Google Scholar] [CrossRef] [PubMed]
  552. Yang, Y.; Yao, Q.; Pu, X.; Hou, Z.; Zhang, Q. Biphasic calcium phosphate macroporous scaffolds derived from oyster shells for bone tissue engineering. Chem. Eng. J. 2011, 173, 837–845. [Google Scholar] [CrossRef]
  553. Moroni, L.; de Wijn, J.R.; van Blitterswijk, C.A. Integrating novel technologies to fabricate smart scaffolds. J. Biomater. Sci. Polym. Edition. 2008, 19, 543–572. [Google Scholar] [CrossRef]
  554. Studart, A.R.; Gonzenbach, U.T.; Tervoort, E.; Gauckler, L.J. Processing routes to macroporous ceramics: A review. J. Am. Ceram. Soc. 2006, 89, 1771–1789. [Google Scholar] [CrossRef]
  555. Hing, K.; Annaz, B.; Saeed, S.; Revell, P.; Buckland, T. Microporosity enhances bioactivity of synthetic bone graft substitutes. J. Mater. Sci. Mater. Med. 2005, 16, 467–475. [Google Scholar] [CrossRef] [PubMed]
  556. Wang, Z.; Sakakibara, T.; Sudo, A.; Kasai, Y. Porosity of β-tricalcium phosphate affects the results of lumbar posterolateral fusion. J. Spinal Disord. Tech. 2013, 26, E40–E45. [Google Scholar] [CrossRef] [PubMed]
  557. Lan Levengood, S.K.; Polak, S.J.; Wheeler, M.B.; Maki, A.J.; Clark, S.G.; Jamison, R.D.; Wagoner Johnson, A.J. Multiscale osteointegration as a new paradigm for the design of calcium phosphate scaffolds for bone regeneration. Biomaterials 2010, 31, 3552–3563. [Google Scholar] [CrossRef] [PubMed]
  558. Ruksudjarit, A.; Pengpat, K.; Rujijanagul, G.; Tunkasiri, T. The fabrication of nanoporous hydroxyapatite ceramics. Adv. Mater. Res. 2008, 47–50, 797–800. [Google Scholar] [CrossRef]
  559. Murugan, R.; Ramakrishna, S.; Rao, K.P. Nanoporous hydroxy-carbonate apatite scaffold made of natural bone. Mater. Lett. 2006, 60, 2844–2847. [Google Scholar] [CrossRef]
  560. Li, Y.; Tjandra, W.; Tam, K.C. Synthesis and characterization of nanoporous hydroxyapatite using cationic surfactants as templates. Mater. Res. Bull. 2008, 43, 2318–2326. [Google Scholar] [CrossRef]
  561. El Asri, S.; Laghzizil, A.; Saoiabi, A.; Alaoui, A.; El Abassi, K.; M’hamdi, R.; Coradin, T. A novel process for the fabrication of nanoporous apatites from Moroccan phosphate rock. Colloids Surf. A 2009, 350, 73–78. [Google Scholar] [CrossRef]
  562. Ramli, R.A.; Adnan, R.; Bakar, M.A.; Masudi, S.M. Synthesis and characterisation of pure nanoporous hydroxyapatite. J. Phys. Sci. 2011, 22, 25–37. [Google Scholar]
  563. LeGeros, R.Z. Calcium phosphate-based osteoinductive materials. Chem. Rev. 2008, 108, 4742–4753. [Google Scholar] [CrossRef] [PubMed]
  564. Prokopiev, O.; Sevostianov, I. Dependence of the mechanical properties of sintered hydroxyapatite on the sintering temperature. Mater. Sci. Eng. A 2006, 431, 218–227. [Google Scholar] [CrossRef]
  565. Daculsi, G.; Jegoux, F.; Layrolle, P. The Micro Macroporous Biphasic Calcium Phosphate Concept for Bone Reconstruction and Tissue Engineering. In Advanced Biomaterials: Fundamentals, Processing and Applications; Basu, B., Katti, D.S., Kumar, A., Eds.; American Ceramic Society, Wiley: Hoboken, NJ, USA, 2009; p. 768. [Google Scholar]
  566. Shipman, P.; Foster, G.; Schoeninger, M. Burnt bones and teeth: An experimental study of color, morphology, crystal structure and shrinkage. J. Archaeol. Sci. 1984, 11, 307–325. [Google Scholar] [CrossRef]
  567. Simske, S.J.; Ayers, R.A.; Bateman, T.A. Porous materials for bone engineering. Mater. Sci. Forum 1997, 250, 151–182. [Google Scholar] [CrossRef]
  568. Vaz, L.; Lopes, A.B.; Almeida, M. Porosity control of hydroxyapatite implants. J. Mater. Sci. Mater. Med. 1999, 10, 239–242. [Google Scholar] [CrossRef] [PubMed]
  569. Rice, R.W. Porosity of Ceramics; Marcel Dekker: New York, NY, USA, 1998; p. 560. [Google Scholar]
  570. Wang, H.; Zhai, L.; Li, Y.; Shi, T. Preparation of irregular mesoporous hydroxyapatite. Mater. Res. Bull. 2008, 43, 1607–1614. [Google Scholar] [CrossRef]
  571. Ota, T.; Eitsuka, T.; Yoshida, H.; Adachi, N. Porous apatite ceramics derived from woods. Adv. Mater. Res. 2006, 11–12, 247–250. [Google Scholar] [CrossRef]
  572. Fan, J.; Lei, J.; Yu, C.; Tu, B.; Zhao, D. Hard-templating synthesis of a novel rod-like nanoporous calcium phosphate bioceramics and their capacity as antibiotic carriers. Mater. Chem. Phys. 2007, 103, 489–493. [Google Scholar] [CrossRef]
  573. Sopyan, I.; Mel, M.; Ramesh, S.; Khalid, K.A. Porous hydroxyapatite for artificial bone applications. Sci. Technol. Adv. Mater. 2007, 8, 116–123. [Google Scholar] [CrossRef]
  574. Fu, Y.C.; Ho, M.L.; Wu, S.C.; Hsieh, H.S.; Wang, C.K. Porous bioceramic bead prepared by calcium phosphate with sodium alginate gel and PE powder. Mater. Sci. Eng. C 2008, 28, 1149–1158. [Google Scholar] [CrossRef]
  575. Hsu, Y.H.; Turner, I.G.; Miles, A.W. Fabrication of porous bioceramics with porosity gradients similar to the bimodal structure of cortical and cancellous bone. J. Mater. Sci. Mater. Med. 2007, 18, 2251–2256. [Google Scholar] [CrossRef] [PubMed]
  576. Munch, E.; Franco, J.; Deville, S.; Hunger, P.; Saiz, E.; Tomsia, A.P. Porous ceramic scaffolds with complex architectures. JOM 2008, 60, 54–59. [Google Scholar] [CrossRef]
  577. Yun, H.S. Design of hierarchically porous materials for bone tissue regeneration. Key Eng. Mater. 2010, 441, 139–153. [Google Scholar] [CrossRef]
  578. Deisinger, U. Generating porous ceramic scaffolds: Processing and properties. Key Eng. Mater. 2010, 441, 155–179. [Google Scholar] [CrossRef]
  579. Ohji, T.; Fukushima, M. Macro-porous ceramics: Processing and properties. Int. Mater. Rev. 2012, 57, 115–131. [Google Scholar] [CrossRef]
  580. Yan, X.; Yu, C.; Zhou, X.; Tang, J.; Zhao, D. Highly ordered mesoporous bioactive glasses with superior in vitro bone-forming bioactivities. Angew. Chem. 2004, 43, 5980–5984. [Google Scholar] [CrossRef]
  581. Izquierdo-Barba, I.; Ruiz-González, L.; Doadrio, J.C.; González-Calbet, J.M.; Vallet-Regí, M. Tissue regeneration: A new property of mesoporous materials. Solid State Sci. 2005, 7, 983–989. [Google Scholar] [CrossRef]
  582. Cosijns, A.; Vervaet, C.; Luyten, J.; Mullens, S.; Siepmann, F.; van Hoorebeke, L.; Masschaele, B.; Cnudde, V.; Remon, J.P. Porous hydroxyapatite tablets as carriers for low-dosed drugs. Eur. J. Pharm. Biopharm. 2007, 67, 498–506. [Google Scholar] [CrossRef] [PubMed]
  583. Uchida, A.; Shinto, Y.; Araki, N.; Ono, K. Slow release of anticancer drugs from porous calcium hydroxyapatite ceramic. J. Orthop. Res. 1992, 10, 440–445. [Google Scholar] [CrossRef] [PubMed]
  584. Shinto, Y.; Uchida, A.; Korkusuz, F.; Araki, N.; Ono, K. Calcium hydroxyapatite ceramic used as a delivery system for antibiotics. J. Bone. Joint. Surg. Br. 1992, 74, 600–604. [Google Scholar] [PubMed]
  585. Martin, R.B.; Chapman, M.W.; Sharkey, N.A.; Zissimos, S.L.; Bay, B.; Shors, E.C. Bone ingrowth and mechanical properties of coralline hydroxyapatite 1 yr after implantation. Biomaterials 1993, 14, 341–348. [Google Scholar] [CrossRef] [PubMed]
  586. Kazakia, G.J.; Nauman, E.A.; Ebenstein, D.M.; Halloran, B.P.; Keaveny, T.M. Effects of in vitro bone formation on the mechanical properties of a trabeculated hydroxyapatite bone substitute. J. Biomed. Mater. Res. A 2006, 77, 688–699. [Google Scholar] [CrossRef] [PubMed]
  587. Hing, K.A.; Best, S.M.; Tanner, K.E.; Bonfield, W.; Revell, P.A. Mediation of bone ingrowth in porous hydroxyapatite bone graft substitutes. J. Biomed. Mater. Res. A 2004, 68, 187–200. [Google Scholar] [CrossRef] [PubMed]
  588. Vuola, J.; Taurio, R.; Goransson, H.; Asko-Seljavaara, S. Compressive strength of calcium carbonate and hydroxyapatite implants after bone marrow induced osteogenesis. Biomaterials 1998, 19, 223–227. [Google Scholar] [CrossRef] [PubMed]
  589. von Doernberg, M.C.; von Rechenberg, B.; Bohner, M.; Grünenfelder, S.; van Lenthe, G.H.; Müller, R.; Gasser, B.; Mathys, R.; Baroud, G.; Auer, J. In vivo behavior of calcium phosphate scaffolds with four different pore sizes. Biomaterials 2006, 27, 5186–5198. [Google Scholar] [CrossRef] [PubMed]
  590. Mygind, T.; Stiehler, M.; Baatrup, A.; Li, H.; Zou, X.; Flyvbjerg, A.; Kassem, M.; Bunger, C. Mesenchymal stem cell ingrowth and differentiation on coralline hydroxyapatite scaffolds. Biomaterials 2007, 28, 1036–1047. [Google Scholar] [CrossRef] [PubMed]
  591. Mankani, M.H.; Afghani, S.; Franco, J.; Launey, M.; Marshall, S.; Marshall, G.W.; Nissenson, R.; Lee, J.; Tomsia, A.P.; Saiz, E. Lamellar spacing in cuboid hydroxyapatite scaffolds regulates bone formation by human bone marrow stromal cells. Tissue Eng. A 2011, 17, 1615–1623. [Google Scholar] [CrossRef]
  592. Chan, O.; Coathup, M.J.; Nesbitt, A.; Ho, C.Y.; Hing, K.A.; Buckland, T.; Campion, C.; Blunn, G.W. The effects of microporosity on osteoinduction of calcium phosphate bone graft substitute biomaterials. Acta Biomater. 2012, 8, 2788–2794. [Google Scholar] [CrossRef] [PubMed]
  593. Tsuruga, E.; Takita, H.; Wakisaka, Y.; Kuboki, Y. Pore size of porous hydoxyapatite as the cell-substratum controls BMP-induced osteogenesis. J. Biochem. 1997, 121, 317–324. [Google Scholar] [CrossRef] [PubMed]
  594. LeGeros, R.Z.; LeGeros, J.P. Calcium phosphate bioceramics: Past, present, future. Key Eng. Mater. 2003, 240–242, 3–10. [Google Scholar] [CrossRef]
  595. Woodard, J.R.; Hilldore, A.J.; Lan, S.K.; Park, C.J.; Morgan, A. W.; Eurell, J.A.C.; Clark, S.G.; Wheeler, M.B.; Jamison, R.D.; Wagoner, J.A.J. The mechanical properties and osteoconductivity of hydroxyapatite bone scaffolds with multi-scale porosity. Biomaterials 2007, 28, 45–54. [Google Scholar] [CrossRef] [PubMed]
  596. Levitt, G.E; Crayton, P.H.; Monroe, E.A.; Condrate, R.A. Forming methods for apatite prosthesis. J. Biomed. Mater. Res. 1969, 3, 683–685. [Google Scholar] [CrossRef] [PubMed]
  597. Easwer, H.V.; Rajeev, A.; Varma, H.K.; Vijayan, S.; Bhattacharya, R.N. Cosmetic and radiological outcome following the use of synthetic hydroxyapatite porous-dense bilayer burr-hole buttons. Acta Neurochir. 2007, 149, 481–485. [Google Scholar] [CrossRef] [PubMed]
  598. Kashimura, H.; Ogasawara, K.; Kubo, Y.; Yoshida, K.; Sugawara, A.; Ogawa, A. A newly designed hydroxyapatite ceramic burr-hole button. Vasc. Health Risk Manag. 2010, 6, 105–108. [Google Scholar] [CrossRef] [PubMed]
  599. Jordan, D.R.; Gilberg, S.; Bawazeer, A. Coralline hydroxyapatite orbital implant (Bio-Eye): Experience with 158 patients. Ophthalmic Plast. Reconstr. Surg. 2004, 20, 69–74. [Google Scholar] [CrossRef]
  600. Liao, H.F.; Xiao, W.; Chen, Q.J. Ophthalmic applications of hydroxyapatite and its polymer composites. J. Clin. Rehabil. Tissue Eng. Res. 2008, 12, 8905–8908. [Google Scholar]
  601. Yoon, J.S.; Lew, H.; Kim, S.J.; Lee, S.Y. Exposure rate of hydroxyapatite orbital implants. A 15-year experience of 802 cases. Ophthalmology 2008, 115, 566–572. [Google Scholar] [CrossRef] [PubMed]
  602. Chai, G.R.; Chen, M. Clinical effect of hydroxyapatite orbital implantation. Int. J. Ophthalmic 2010, 10, 999–1000. [Google Scholar]
  603. Tabatabaee, Z.; Mazloumi, M.; Rajabi, T.M.; Khalilzadeh, O.; Kassaee, A.; Moghimi, S.; Eftekhar, H.; Goldberg, R.A. Comparison of the exposure rate of wrapped hydroxyapatite (Bio-Eye) vs. unwrapped porous polyethylene (Medpor) orbital implants in enucleated patients. Ophthalmic Plast. Reconstr. Surg. 2011, 27, 114–118. [Google Scholar] [CrossRef]
  604. Ma, X.Z.; Bi, H.S.; Zhang, X. Effect of hydroxyapatite orbital implant for plastic surgery of eye in 52 cases. Int. Eye Sci. 2012, 12, 988–990. [Google Scholar]
  605. Wang, L. Simple fixation of hydroxyapatite artificial eye mount of auto sclera. Int. Eye Sci. 2012, 12, 1394–1395. [Google Scholar]
  606. Kundu, B.; Sanyal, D.; Basu, D. Physiological and elastic properties of highly porous hydroxyapatite potential for integrated eye implants: Effects of SIRC and L-929 cell lines. Ceram. Int. 2013, 39, 2651–2664. [Google Scholar] [CrossRef]
  607. Wehrs, R.E. Hearing results with incus and incus stapes prostheses of hydroxylapatite. Laryngoscope 1991, 101, 555–556. [Google Scholar] [CrossRef] [PubMed]
  608. Smith, J.; Gardner, E.; Dornhoffer, J.L. Hearing results with a hydroxylapatite/titanium bell partial ossicular replacement prosthesis. Laryngoscope 2002, 112, 1796–1799. [Google Scholar] [CrossRef] [PubMed]
  609. Doi, T.; Hosoda, Y.; Kaneko, T.; Munemoto, Y.; Kaneko, A.; Komeda, M.; Furukawa, M.; Kuriyama, H.; Kitajiri, M.; Tomoda, K.; et al. Hearing results for ossicular reconstruction using a cartilage-connecting hydroxyapatite prosthesis with a spearhead. Otol. Neurotol. 2007, 28, 1041–1044. [Google Scholar] [CrossRef] [PubMed]
  610. Thalgott, J.S.; Fritts, K.; Giuffre, J.M.; Timlin, M. Anterior interbody fusion of the cervical spine with coralline hydroxyapatite. Spine 1999, 24, 1295–1299. [Google Scholar] [CrossRef] [PubMed]
  611. Mashoof, A.A.; Siddiqui, S.A.; Otero, M.; Tucci, J.J. Supplementation of autogenous bone graft with coralline hydroxyapatite in posterior spine fusion for idiopathic adolescent scoliosis. Orthopedics 2002, 25, 1073–1076. [Google Scholar] [PubMed]
  612. Minamide, A.; Yoshida, M.; Kawakami, M.; Yamasaki, S.; Kojima, H.; Hashizume, H.; Boden, S.D. The use of cultured bone marrow cells in type I collagen gel and porous hydroxyapatite for posterolateral lumbar spine fusion. Spine 2005, 30, 1134–1138. [Google Scholar] [CrossRef] [PubMed]
  613. Liu, W.Y.; Mo, J.W.; Gao, H.; Liu, H.L.; Wang, M.Y.; He, C.L.; Tang, W.; Ye, Y.J. Nano-hydroxyapatite artificial bone serves as a spacer for fusion with the cervical spine after bone grafting. Chin. J. Tissue Eng. Res. 2012, 16, 5327–5330. [Google Scholar]
  614. Silva, R.V.; Camilli, J.A.; Bertran, C.A.; Moreira, N.H. The use of hydroxyapatite and autogenous cancellous bone grafts to repair bone defects in rats. Int. J. Oral Maxillofac. Surg. 2005, 34, 178–184. [Google Scholar] [CrossRef] [PubMed]
  615. Damron, T.A. Use of 3D β-tricalcium phosphate (Vitoss®) scaffolds in repairing bone defects. Nanomedicine 2007, 2, 763–775. [Google Scholar] [CrossRef] [PubMed]
  616. Bai, B.; Jazrawi, L.M.; Kummer, F.J.; Spivak, J.M. The use of an injectable, biodegradable calcium phosphate bone substitute for the prophylactic augmentation of osteoporotic vertebrae and the management of vertebral compression fractures. Spine 1999, 24, 1521–1526. [Google Scholar] [CrossRef] [PubMed]
  617. Busso, M.; Karlsberg, P.L. Cheek augmentation and rejuvenation using injectable calcium hydroxylapatite (Radiesse®). Cosmet. Dermatol. 2006, 19, 583–588. [Google Scholar]
  618. Low, K.L.; Tan, S.H.; Zein, S.H.S.; Roether, J.A.; Mouriño, V.; Boccaccini, A.R. Calcium phosphate-based composites as injectable bone substitute materials. J. Biomed. Mater. Res. B 2010, 94, 273–286. [Google Scholar]
  619. Daculsi, G.; Uzel, A.P.; Weiss, P.; Goyenvalle, E.; Aguado, E. Developments in injectable multiphasic biomaterials. The performance of microporous biphasic calcium phosphate granules and hydrogels. J. Mater. Sci. Mater. Med. 2010, 21, 855–861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  620. Bohner, M.; Baroud, G. Injectability of calcium phosphate pastes. Biomaterials 2005, 26, 1553–1563. [Google Scholar] [CrossRef] [PubMed]
  621. Laschke, M.W.; Witt, K.; Pohlemann, T.; Menger, M.D. Injectable nanocrystalline hydroxyapatite paste for bone substitution: in vivo analysis of biocompatibility and vascularization. J. Biomed. Mater. Res. B 2007, 82, 494–505. [Google Scholar] [CrossRef]
  622. Klesing, J.; Chernousova, S.; Kovtun, A.; Neumann, S.; Ruiz, L.; Gonzalez-Calbet, J.M.; Vallet-Regi, M.; Heumann, R.; Epple, M. An injectable paste of calcium phosphate nanorods, functionalized with nucleic acids, for cell transfection and gene silencing. J. Mater. Chem. 2010, 20, 6144–6148. [Google Scholar] [CrossRef]
  623. Salinas, A.J.; Esbrit, P.; Vallet-Regí, M. A tissue engineering approach based on the use of bioceramics for bone repair. Biomater. Sci. 2013, 1, 40–51. [Google Scholar] [CrossRef]
  624. Constantz, B.R.; Ison, I.C.; Fulmer, M.T.; Poser, R.D.; Smith, S.T.; Vanwagoner, M.; Ross, J.; Goldstein, S.A.; Jupiter, J.B.; Rosenthal, D.I. Skeletal repair by in situ formation of the mineral phase of bone. Science 1995, 267, 1796–1799. [Google Scholar] [CrossRef] [PubMed]
  625. Chow, L.C. Next generation calcium phosphate-based biomaterials. Dent. Mater. J. 2009, 28, 1–10. [Google Scholar] [CrossRef] [PubMed]
  626. Del Real, R.P.; Wolke, J.G.C.; Vallet-Regí, M.; Jansen, J.A. A new method to produce macropores in calcium phosphate cements. Biomaterials 2002, 23, 3673–3680. [Google Scholar] [CrossRef] [PubMed]
  627. Victor, S.P.; Kumar, T.S.S. Processing and properties of injectable porous apatitic cements. J. Ceram. Soc. Jpn. 2008, 116, 105–107. [Google Scholar] [CrossRef]
  628. Hesaraki, S.; Nemati, R.; Nosoudi, N. Preparation and characterisation of porous calcium phosphate bone cement as antibiotic carrier. Adv. Appl. Ceram. 2009, 108, 231–240. [Google Scholar] [CrossRef]
  629. Bohner, M. Resorbable biomaterials as bone graft substitutes. Mater. Today 2010, 13, 24–30. [Google Scholar] [CrossRef]
  630. Habibovic, P.; Li, J.; van der Valk, C. M.; Meijer, G.; Layrolle, P.; van Blitterswijk, C. A.; de Groot, K. Biological performance of uncoated and octacalcium phosphate-coated Ti6Al4V. Biomaterials 2005, 26, 23–36. [Google Scholar] [CrossRef] [PubMed]
  631. Narayanan, R.; Seshadri, S.K.; Kwon, T.Y.; Kim, K.H. Calcium phosphate-based coatings on titanium and its alloys. J. Biomed. Mater. Res. B 2008, 85, 279–299. [Google Scholar] [CrossRef]
  632. Paital, S.R.; Dahotre, N.B. Calcium phosphate coatings for bio-implant applications: Materials, performance factors, and methodologies. Mater. Sci. Eng. R 2009, 66, 1–70. [Google Scholar] [CrossRef]
  633. Thin Calcium Phosphate Coatings for Medical Implants; León, B.; Jansen, J.A. (Eds.) Springer: New York, NY, USA, 2009; p. 326.
  634. Dorozhkin, S.V. Calcium orthophosphate coatings, films and layers. Prog. Biomater. 2012, 1, 1–40. [Google Scholar] [CrossRef]
  635. Kon, M.; Ishikawa, K.; Miyamoto, Y.; Asaoka, K. Development of calcium phosphate based functional gradient bioceramics. Biomaterials 1995, 16, 709–714. [Google Scholar] [CrossRef] [PubMed]
  636. Wong, L.H.; Tio, B.; Miao, X. Functionally graded tricalcium phosphate/fluoroapatite composites. Mater. Sci. Eng. C 2002, 20, 111–115. [Google Scholar] [CrossRef]
  637. Tampieri, A.; Celotti, G.; Sprio, S.; Delcogliano, A.; Franzese, S. Porosity-graded hydroxyapatite ceramics to replace natural bone. Biomaterials 2001, 22, 1365–1370. [Google Scholar] [CrossRef] [PubMed]
  638. Lu, W.W.; Zhao, F.; Luk, K.D.K.; Yin, Y.J.; Cheung, K.M.C.; Cheng, G.X.; Yao, K.D.; Leong, J.C.Y. Controllable porosity hydroxyapatite ceramics as spine cage: Fabrication and properties evaluation. J. Mater. Sci. Mater. Med. 2003, 14, 1039–1046. [Google Scholar] [CrossRef] [PubMed]
  639. Werner, J.; Linner-Krcmar, B.; Friess, W.; Greil, P. Mechanical properties and in vitro cell compatibility of hydroxyapatite ceramics with graded pore structure. Biomaterials 2002, 23, 4285–4294. [Google Scholar] [CrossRef] [PubMed]
  640. Rodriguez-Lorenzo, L.M.; Ferreira, J.M.F. Development of porous ceramic bodies for applications in tissue engineering and drug delivery systems. Mater. Res. Bull. 2004, 39, 83–91. [Google Scholar] [CrossRef]
  641. Kaito, T.; Mukai, Y.; Nishikawa, M.; Ando, W.; Yoshikawa, H.; Myoui, A. Dual hydroxyapatite composite with porous and solid parts: Experimental study using canine lumbar interbody fusion model. J. Biomed. Mater. Res. B 2006, 78, 378–384. [Google Scholar] [CrossRef]
  642. Bai, X.; Sandukas, S.; Appleford, M.R.; Ong, J.L.; Rabiei, A. Deposition and investigation of functionally graded calcium phosphate coatings on titanium. Acta Biomater. 2009, 5, 3563–3572. [Google Scholar] [CrossRef] [PubMed]
  643. Roy, M.; Balla, V.K.; Bandyopadhyay, A.; Bose, S. Compositionally graded hydroxyapatite/tricalcium phosphate coating on Ti by laser and induction plasma. Acta Biomater. 2011, 7, 866–873. [Google Scholar] [CrossRef] [PubMed]
  644. Tazaki, J.; Yodogawa, S.; Murata, M.; Ito, K.; Akazawa, T.; Hino, J.; Kabir, M.A.; Nagayasu, H.; Arisue, M.; Shibata, T.; et al. Behavior of human blood adsorption to biomimetic functionally graded hydroxyapatite. Key Eng. Mater. 2013, 529–530, 44–49. [Google Scholar]
  645. Dubok, V.A. Bioceramics—Yesterday, today, tomorrow. Powder Metall. Met. Ceram. 2000, 39, 381–394. [Google Scholar] [CrossRef]
  646. Heness, G.; Ben-Nissan, B. Innovative bioceramics. Mater. Forum 2004, 27, 104–114. [Google Scholar]
  647. Ohtsuki, C.; Kamitakahara, M.; Miyazaki, T. Bioactive ceramic-based materials with designed reactivity for bone tissue regeneration. J. R. Soc. Interface 2009, 6, S349–S360. [Google Scholar] [CrossRef] [PubMed]
  648. Greenspan, D.C. Bioactive ceramic implant materials. Curr. Opin. Solid State Mater. Sci. 1999, 4, 389–393. [Google Scholar] [CrossRef]
  649. Blokhuis, T.J.; Termaat, M.F.; den Boer, F.C.; Patka, P.; Bakker, F.C.; Haarman, H.J.T.M. Properties of calcium phosphate ceramics in relation to their in vivo behavior. J. Trauma 2000, 48, 179–189. [Google Scholar] [CrossRef] [PubMed]
  650. Kim, H.M. Bioactive ceramics: Challenges and perspectives. J. Ceram. Soc. Jpn. 2001, 109, S49–S57. [Google Scholar] [CrossRef]
  651. Seeley, Z.; Bandyopadhyay, A.; Bose, S. Tricalcium phosphate based resorbable ceramics: Influence of NaF and CaO addition. Mater. Sci. Eng. C 2008, 28, 11–17. [Google Scholar] [CrossRef]
  652. Descamps, M.; Richart, O.; Hardouin, P.; Hornez, J.C.; Leriche, A. Synthesis of macroporous β-tricalcium phosphate with controlled porous architectural. Ceram. Int. 2008, 34, 1131–1137. [Google Scholar] [CrossRef]
  653. Cushnie, E.K.; Khan, Y.M.; Laurencin, C.T. Amorphous hydroxyapatite-sintered polymeric scaffolds for bone tissue regeneration: Physical characterization studies. J. Biomed. Mater. Res. A 2008, 84, 54–62. [Google Scholar] [CrossRef] [PubMed]
  654. Hench, L.L. Challenges for bioceramics in the 21st century. Am. Ceram. Soc. Bull. 2005, 84, 18–21. [Google Scholar]
  655. Nagase, M.; Baker, D.G.; Schumacher, H.R. Prolonged inflammatory reactions induced by artificial ceramics in the rat pouch model. J. Rheumatol. 1988, 15, 1334–1338. [Google Scholar] [PubMed]
  656. Rooney, T.; Berman, S.; Indersano, A.T. Evaluation of porous block hydroxylapatite for augmentation of alveolar ridges. J. Oral Maxillofac. Surg. 1988, 46, 15–18. [Google Scholar] [CrossRef] [PubMed]
  657. Prudhommeaux, F.; Schiltz, C.; Lioté, F.; Hina, A.; Champy, R.; Bucki, B.; Ortiz-Bravo, E.; Meunier, A.; Rey, C.; Bardin, T. Variation in the inflammatory properties of basic calcium phosphate crystals according to crystal type. Arthritis Rheum. 1996, 39, 1319–1326. [Google Scholar] [CrossRef] [PubMed]
  658. Ghanaati, S.; Barbeck, M.; Orth, C.; Willershausen, I.; Thimm, B.W.; Hoffmann, C.; Rasic, A.; Sader, R.A.; Unger, R.E.; Peters, F.; Kirkpatrick, C.J. Influence of β-tricalcium phosphate granule size and morphology on tissue reaction in vivo. Acta Biomater. 2010, 6, 4476–4487. [Google Scholar] [CrossRef] [PubMed]
  659. Lin, K.; Yuan, W.; Wang, L.; Lu, J.; Chen, L.; Wang, Z.; Chang, J. Evaluation of host inflammatory responses of β-tricalcium phosphate bioceramics caused by calcium pyrophosphate impurity using a subcutaneous model. J. Biomed. Mater. Res. B 2011, 99, 350–358. [Google Scholar] [CrossRef]
  660. Velard, F.; Braux, J.; Amedee, J.; Laquerriere, P. Inflammatory cell response to calcium phosphate biomaterial particles: An overview. Acta Biomater. 2013, 9, 4956–4963. [Google Scholar] [CrossRef] [PubMed]
  661. Rydén, L.; Molnar, D.; Esposito, M.; Johansson, A.; Suska, F.; Palmquist, A.; Thomsen, P. Early inflammatory response in soft tissues induced by thin calcium phosphates. J. Biomed. Mater. Res. A 2013, 101, 2712–2717. [Google Scholar] [CrossRef] [PubMed]
  662. Chatterjea, A.; van der Stok, J.; Danoux, C.B.; Yuan, H.; Habibovic, P.; van Blitterswijk, C.A.; Weinans, H.; de Boer, J. Inflammatory response and bone healing capacity of two porous calcium phosphate ceramics in a critical size cortical bone defects. J. Biomed. Mater. Res. A 2013. [Google Scholar] [CrossRef]
  663. Uchida, A.; Araki, N.; Shinto, Y.; Yoshikawa, H.; Kurisaki, E.; Ono, K. The use of calcium hydroxyapatite ceramic in bone tumour surgery. J. Bone Joint Surg. Br. 1990, 72, 298–302. [Google Scholar] [PubMed]
  664. Ghanaati, S.; Barbeck, M.; Detsch, R.; Deisinger, U.; Hilbig, U.; Rausch, V.; Sader, R.; Unger, R.E.; Ziegler, G.; Kirkpatrick, C.J. The chemical composition of synthetic bone substitutes influences tissue reactions in vivo: Histological and histomorphometrical analysis of the cellular inflammatory response to hydroxyapatite, beta-tricalcium phosphate and biphasic calcium phosphate ceramics. Biomed. Mater. 2012, 7. [Google Scholar] [CrossRef]
  665. Draenert, K.; Draenert, M.; Erler, M.; Draenert, A.; Draenert, Y. How bone forms in large cancellous defects: Critical analysis based on experimental work and literature. Injury 2011, 42, S47–S55. [Google Scholar] [CrossRef] [PubMed]
  666. Yuan, H.; Yang, Z.; Li, Y.; Zhang, Z.; de Bruijn, J.D.; de Groot, K. Osteoinduction by calcium phosphate biomaterials. J. Mater. Sci. Mater. Med. 1998, 9, 723–726. [Google Scholar] [CrossRef] [PubMed]
  667. Yuan, H.; Kurashina, K.; de Bruijn, D.J.; Li, Y.; de Groot, K.; Zhang, X. A preliminary study of osteoinduction of two kinds of calcium phosphate bioceramics. Biomaterials 1999, 20, 1799–1806. [Google Scholar] [CrossRef] [PubMed]
  668. Yuan, H.P.; de Bruijn, J.D.; Li, Y.B.; Feng, J.Q.; Yang, Z.J.; de Groot, K.; Zhang, X.D. Bone formation induced by calcium phosphate ceramics in soft tissue of dogs: A comparative study between porous α-TCP and β-TCP. J. Mater. Sci. Mater. Med. 2001, 12, 7–13. [Google Scholar] [CrossRef] [PubMed]
  669. LeGeros, R.Z. Properties of osteoconductive biomaterials: Calcium phosphates. Clin. Orthop. Rel. Res. 2002, 395, 81–98. [Google Scholar] [CrossRef]
  670. Gosain, A.K.; Song, L.; Riordan, P.; Amarante, M.T.; Nagy, P.G.; Wilson, C.R.; Toth, J.M.; Ricci, J.L. A 1-year study of osteoinduction in hydroxyapatite-derived biomaterials in an adult sheep model: Part I. Plast. Reconstr. Surg. 2002, 109, 619–630. [Google Scholar] [CrossRef] [PubMed]
  671. Matsushita, N.; Terai, H.; Okada, T.; Nozaki, K.; Inoue, H.; Miyamoto, S.; Takaoka, K. A new bone-inducing biodegradable porous β-tricalcium phosphate. J. Biomed. Mater. Res. A 2004, 70, 450–458. [Google Scholar] [CrossRef] [PubMed]
  672. Le Nihouannen, D.; Daculsi, G.; Saffarzadeh, A.; Gauthier, O.; Delplace, S.; Pilet, P.; Layrolle, P. Ectopic bone formation by microporous calcium phosphate ceramic particles in sheep muscles. Bone 2005, 36, 1086–1093. [Google Scholar] [CrossRef] [PubMed]
  673. Cheng, L.; Ye, F.; Yang, R.; Lu, X.; Shi, Y.; Li, L.; Fan, H.; Bu, H. Osteoinduction of hydroxyapatite/β-tricalcium phosphate bioceramics in mice with a fractured fibula. Acta Biomater. 2010, 6, 1569–1574. [Google Scholar] [CrossRef] [PubMed]
  674. Barrere, F.; van der Valk, C.M.; Dalmeijer, R.A.; Meijer, G.; van Blitterswijk, C.A.; de Groot, K.; Layrolle, P. Osteogenecity of octacalcium phosphate coatings applied on porous titanium. J. Biomed. Mater. Res. A 2003, 66, 779–788. [Google Scholar] [CrossRef] [PubMed]
  675. Habibovic, P.; van der Valk, C. M.; van Blitterswijk, C. A.; de Groot, K.; Meijer, G. Influence of octacalcium phosphate coating on osteoinductive properties of biomaterials. J. Mater. Sci. Mater. Med. 2004, 15, 373–380. [Google Scholar] [CrossRef] [PubMed]
  676. Yuan, H.; van Blitterswijk, C.A.; de Groot, K.; de Bruijn, J.D. Cross-species comparison of ectopic bone formation in biphasic calcium phosphate (BCP) and hydroxyapatite (HA) scaffolds. Tissue Eng. 2006, 12, 1607–1615. [Google Scholar] [CrossRef] [PubMed]
  677. Ripamonti, U.; Richter, P.W.; Nilen, R.W.; Renton, L. The induction of bone formation by smart biphasic hydroxyapatite tricalcium phosphate biomimetic matrices in the non human primate Papio ursinus. J. Cell. Mol. Med. 2008, 12, 2609–2621. [Google Scholar] [CrossRef] [PubMed]
  678. Ripamonti, U.; Crooks, J.; Khoali, L.; Roden, L. The induction of bone formation by coral-derived calcium carbonate/hydroxyapatite constructs. Biomaterials 2009, 30, 1428–1439. [Google Scholar] [CrossRef] [PubMed]
  679. Yuan, H.; Fernandes, H.; Habibovic, P.; de Boer, J.; Barradas, A.M.C.; de Ruiter, A.; Walsh, W.R.; van Blitterswijk, C.A.; de Bruijn, J.D. Osteoinductive ceramics as a synthetic alternative to autologous bone grafting. Proc. Natl. Acad. Sci. USA 2010, 107, 13614–13619. [Google Scholar] [CrossRef] [PubMed]
  680. Yao, J.F.; Li, X.Y.; Wang, A.J.; Liang, R.; Bao, C.Y.; Chen, Z.Q. Osteoinductive calcium phosphate ceramics for in vivo construction of tissue engineered bone in adipose tissue. J. Clin. Rehabil. Tissue Eng. Res. 2011, 15, 2109–2112. [Google Scholar]
  681. Barradas, A.M.; Yuan, H.; van der Stok, J.; le Quang, B.; Fernandes, H.; Chaterjea, A.; Hogenes, M.C.; Shultz, K.; Donahue, L.R.; van Blitterswijk, C.; de Boer, J. The influence of genetic factors on the osteoinductive potential of calcium phosphate ceramics in mice. Biomaterials 2012, 33, 5696–5705. [Google Scholar] [CrossRef] [PubMed]
  682. Li, B.; Liao, X.; Zheng, L.; Zhu, X.; Wang, Z.; Fan, H.; Zhang, X. Effect of nanostructure on osteoinduction of porous biphasic calcium phosphate ceramics. Acta Biomater. 2012, 8, 3794–3804. [Google Scholar] [CrossRef] [PubMed]
  683. Cheng, L.; Shi, Y.; Ye, F.; Bu, H. Osteoinduction of calcium phosphate biomaterials in small animals. Mater. Sci. Eng. C 2013, 33, 1254–1260. [Google Scholar] [CrossRef]
  684. Song, G.; Habibovic, P.; Bao, C.; Hu, J.; van Blitterswijk, C.A.; Yuan, H.; Chen, W.; Xu, H.H.K. The homing of bone marrow MSCs to non-osseous sites for ectopic bone formation induced by osteoinductive calcium phosphate. Biomaterials 2013, 34, 2167–2176. [Google Scholar] [CrossRef] [PubMed]
  685. He, P.; Sahoo, S.; Ng, K.S.; Chen, K.; Toh, S.L.; Goh, J.C.H. Enhanced osteoinductivity and osteoconductivity through hydroxyapatite coating of silk-based tissue-engineered ligament scaffold. J. Biomed. Mater. Res. A 2013, 101, 555–566. [Google Scholar] [CrossRef] [PubMed]
  686. Habibovic, P.; Yuan, H.; van der Valk, C.M.; Meijer, G.; van Blitterswijk, C.A.; de Groot, K. 3D microenvironment as essential element for osteoinduction by biomaterials. Biomaterials 2005, 26, 3565–3575. [Google Scholar] [CrossRef] [PubMed]
  687. Habibovic, P.; Sees, T.M.; van den Doel, M.A.; van Blitterswijk, C.A.; de Groot, K. Osteoinduction by biomaterials—physicochemical and structural influences. J. Biomed. Mater. Res. A 2006, 77, 747–762. [Google Scholar] [CrossRef] [PubMed]
  688. Reddi, A.H. Morphogenesis and tissue engineering of bone and cartilage: Inductive signals, stem cells and biomimetic biomaterials. Tissue Eng. 2000, 6, 351–359. [Google Scholar] [CrossRef] [PubMed]
  689. Ripamonti, U. The morphogenesis of bone in replicas of porous hydroxyapatite obtained by conversion of calcium carbonate exoskeletons of coral. J. Bone Joint Surg. A 1991, 73, 692–703. [Google Scholar]
  690. Kuboki, Y.; Takita, H.; Kobayashi, D. BMP-induced osteogenesis on the surface of hydroxyapatite with geometrically feasible and nonfeasible structures: Topology of osteogenesis. J. Biomed. Mater. Res. 1998, 39, 190–199. [Google Scholar] [CrossRef] [PubMed]
  691. Diaz-Flores, L.; Gutierrez, R.; Lopez-Alonso, A.; Gonzalez, R.; Varela, H. Pericytes as a supplementary source of osteoblasts in periosteal osteogenesis. Clin. Orthop. Relat. Res. 1992, 275, 280–286. [Google Scholar] [PubMed]
  692. Boyan, B.D.; Schwartz, Z. Are calcium phosphate ceramics ‘smart’ biomaterials? Nat. Rev. Rheumatol. 2011, 7, 8–9. [Google Scholar] [CrossRef] [PubMed]
  693. Lu, J.; Descamps, M.; Dejou, J.; Koubi, G.; Hardouin, P.; Lemaitre, J.; Proust, J.P. The biodegradation mechanism of calcium phosphate biomaterials in bone. J. Biomed. Mater. Res. 2002, 63, 408–412. [Google Scholar] [CrossRef] [PubMed]
  694. Wang, H.; Lee, J.K.; Moursi, A.; Lannutti, J.J. Ca/P ratio effects on the degradation of hydroxyapatite in vitro. J. Biomed. Mater. Res. A 2003, 67, 599–608. [Google Scholar] [CrossRef] [PubMed]
  695. Dorozhkin, S.V. Inorganic chemistry of the dissolution phenomenon, the dissolution mechanism of calcium apatites at the atomic (ionic) level. Comment Inorg. Chem. 1999, 20, 285–299. [Google Scholar] [CrossRef]
  696. Dorozhkin, S.V. Dissolution mechanism of calcium apatites in acids: A review of literature. World J. Methodol. 2012, 2, 1–17. [Google Scholar] [CrossRef] [PubMed]
  697. Wenisch, S.; Stahl, J.P.; Horas, U.; Heiss, C.; Kilian, O.; Trinkaus, K.; Hild, A.; Schnettler, R. In vivo mechanisms of hydroxyapatite ceramic degradation by osteoclasts: Fine structural microscopy. J. Biomed. Mater. Res. A 2003, 67, 713–718. [Google Scholar] [CrossRef] [PubMed]
  698. Riihonen, R.; Nielsen, S.; Väänänen, H.K.; Laitala-Leinonen, T.; Kwon, T.H. Degradation of hydroxyapatite in vivo and in vitro requires osteoclastic sodium-bicarbonate co-transporter NBCn1. Matrix Biol. 2010, 29, 287–294. [Google Scholar] [CrossRef] [PubMed]
  699. Teitelbaum, S.L. Bone resorption by osteoclasts. Science 2000, 289, 1504–1508. [Google Scholar] [CrossRef] [PubMed]
  700. Narducci, P.; Nicolin, V. Differentiation of activated monocytes into osteoclast-like cells on a hydroxyapatite substrate: An in vitro study. Ann. Anat. 2009, 191, 349–355. [Google Scholar] [CrossRef] [PubMed]
  701. Raynaud, S.; Champion, E.; Lafon, J.P.; Bernache-Assollant, D. Calcium phosphate apatites with variable Ca/P atomic ratio. III. Mechanical properties and degradation in solution of hot pressed ceramics. Biomaterials 2002, 23, 1081–1089. [Google Scholar] [CrossRef] [PubMed]
  702. Barrère, F.; van der Valk, C.M.; Dalmeijer, R.A.J.; van Blitterswijk, C.A.; de Groot, K.; Layrolle, P. In vitro and in vivo degradation of biomimetic octacalcium phosphate and carbonate apatite coatings on titanium implants. J. Biomed. Mater. Res. A 2003, 64, 378–387. [Google Scholar] [CrossRef] [PubMed]
  703. Souto, R.M.; Laz, M.M.; Reis, R.L. Degradation characteristics of hydroxyapatite coatings on orthopaedic TiAlV in simulated physiological media investigated by electrochemical impedance spectroscopy. Biomaterials 2003, 24, 4213–4221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  704. Okuda, T.; Ioku, K.; Yonezawa, I.; Minagi, H.; Kawachi, G.; Gonda, Y.; Murayama, H.; Shibata, Y.; Minami, S.; Kamihara, S.; Kurosawa, H.; Ikeda, T. The effect of the microstructure of β-tricalcium phosphate on the metabolism of subsequently formed bone tissue. Biomaterials 2007, 28, 2612–2621. [Google Scholar] [CrossRef] [PubMed]
  705. Orly, I.; Gregoire, M.; Menanteau, J.; Heughebaert, M.; Kerebel, B. Chemical changes in hydroxyapatite biomaterial under in vivo and in vitro biological conditions. Calcif. Tissue Int. 1989, 45, 20–26. [Google Scholar] [CrossRef] [PubMed]
  706. Sun, L.; Berndt, C.C.; Gross, K.A.; Kucuk, A. Review: Material fundamentals and clinical performance of plasma sprayed hydroxyapatite coatings. J. Biomed. Mater. Res. 2001, 58, 570–592. [Google Scholar] [CrossRef]
  707. Bertazzo, S.; Zambuzzi, W.F.; Campos, D.D.P.; Ogeda, T.L.; Ferreira, C.V.; Bertran, C.A. Hydroxyapatite surface solubility and effect on cell adhesion. Colloids Surf. B 2010, 78, 177–184. [Google Scholar] [CrossRef]
  708. Schwartz, Z.; Boyan, B.D. Underlying mechanisms at the bone-biomaterial interface. J. Cell. Biochem. 1994, 56, 340–347. [Google Scholar] [CrossRef] [PubMed]
  709. Puleo, D.A.; Nanci, A. Understanding and controlling the bone-implant interface. Biomaterials 1999, 20, 2311–2321. [Google Scholar] [CrossRef] [PubMed]
  710. Xin, R.; Leng, Y.; Chen, J.; Zhang, Q. A comparative study of calcium phosphate formation on bioceramics in vitro and in vivo. Biomaterials 2005, 26, 6477–6486. [Google Scholar] [CrossRef] [PubMed]
  711. Girija, E.K.; Parthiban, S.P.; Suganthi, R.V.; Elayaraja, K.; Joshy, M.I.A.; Vani, R.; Kularia, P.; Asokan, K.; Kanjilal, D.; Yokogawa, Y.; et al. High energy irradiation—A tool for enhancing the bioactivity of hydroxyapatite. J. Ceram. Soc. Jpn. 2008, 116, 320–324. [Google Scholar] [CrossRef]
  712. Okada, M.; Furukawa, K.; Serizawa, T.; Yanagisawa, Y.; Tanaka, H.; Kawai, T.; Furuzono, T. Interfacial interactions between calcined hydroxyapatite nanocrystals and substrates. Langmuir 2009, 25, 6300–6306. [Google Scholar] [CrossRef] [PubMed]
  713. Callis, P.D.; Donaldson, K.; McCord, J.F. Early cellular responses to calcium phosphate ceramics. Clin. Mater. 1988, 3, 183–190. [Google Scholar] [CrossRef]
  714. Okumura, M.; Ohgushi, H.; Tamai, S. Bonding osteogenesis in coralline hydroxyapatite combined with bone marrow cells. Biomaterials 1990, 12, 28–37. [Google Scholar]
  715. Holtgrave, E.A.; Donath, K. Response of odontoblast-like cells to hydroxyapatite ceramic granules. Biomaterials 1995, 16, 155–159. [Google Scholar] [CrossRef] [PubMed]
  716. Doi, Y.; Iwanaga, H.; Shibutani, T.; Moriwaki, Y.; Iwayama, Y. Osteoclastic responses to various calcium phosphates in cell cultures. J. Biomed. Mater. Res. 1999, 47, 424–433. [Google Scholar] [CrossRef] [PubMed]
  717. Guo, X.; Gough, J.E.; Xiao, P.; Liu, J.; Shen, Z. Fabrication of nanostructured hydroxyapatite and analysis of human osteoblastic cellular response. J. Biomed. Mater. Res. A 2007, 82, 1022–1032. [Google Scholar] [CrossRef] [PubMed]
  718. Wang, Y.; Zhang, S.; Zeng, X.; Ma, L.L.; Weng, W.; Yan, W.; Qian, M. Osteoblastic cell response on fluoridated hydroxyapatite coatings. Acta Biomater. 2007, 3, 191–197. [Google Scholar] [CrossRef] [PubMed]
  719. Bae, W.J.; Chang, S.W.; Lee, S.I.; Kum, K.Y.; Bae, K.S.; Kim, E.C. Human periodontal ligament cell response to a newly developed calcium phosphate-based root canal sealer. J. Endod. 2010, 36, 1658–1663. [Google Scholar] [CrossRef] [PubMed]
  720. Li, J.; Song, Y.; Zhang, S.; Zhao, C.; Zhang, F.; Zhang, X.; Cao, L.; Fan, Q.; Tang, T. In vitro responses of human bone marrow stromal cells to a fluoridated hydroxyapatite coated biodegradable Mg-Zn alloy. Biomaterials 2010, 31, 5782–5788. [Google Scholar] [CrossRef] [PubMed]
  721. Zhao, X.; Heng, B.C.; Xiong, S.; Guo, J.; Tan, T.T.-Y.; Boey, F.Y.C.; Ng, K.W.; Loo, J.S.C. In vitro assessment of cellular responses to rod-shaped hydroxyapatite nanoparticles of varying lengths and surface areas. Nanotoxicology 2011, 5, 182–194. [Google Scholar] [CrossRef] [PubMed]
  722. Liu, X.; Zhao, M.; Lu, J.; Ma, J.; Wei, J.; Wei, S. Cell responses to two kinds of nanohydroxyapatite with different sizes and crystallinities. Int. J. Nanomed. 2012, 7, 1239–1250. [Google Scholar] [CrossRef]
  723. Perez, R.A.; Kim, T.H.; Kim, M.; Jang, J.H.; Ginebra, M.P.; Kim, H.W. Calcium phosphate cements loaded with basic fibroblast growth factor: Delivery and in vitro cell response. J. Biomed. Mater. Res. A 2013, 101, 923–931. [Google Scholar] [CrossRef] [PubMed]
  724. Yin, P.; Feng, F.F.; Lei, T.; Zhong, X.H.; Jian, X.C. Osteoblastic cell response on biphasic fluorhydroxyapatite/strontium-substituted hydroxyapatite coatings. J. Biomed. Mater. Res. A 2013. [Google Scholar] [CrossRef]
  725. Suzuki, T.; Ohashi, R.; Yokogawa, Y.; Nishizawa, K.; Nagata, F.; Kawamoto, Y.; Kameyama, T.; Toriyama, M. Initial anchoring and proliferation of fibroblast L-929 cells on unstable surface of calcium phosphate ceramics. J. Biosci. Bioeng. 1999, 87, 320–327. [Google Scholar] [CrossRef] [PubMed]
  726. Arinzeh, T.L.; Tran, T.; McAlary, J.; Daculsi, G. A comparative study of biphasic calcium phosphate ceramics for human mesenchymal stem-cell-induced bone formation. Biomaterials 2005, 26, 3631–3638. [Google Scholar] [CrossRef] [PubMed]
  727. Oh, S.; Oh, N.; Appleford, M.; Ong, J.L. Bioceramics for tissue engineering applications—A review. Am. J. Biochem. Biotechnol. 2006, 2, 49–56. [Google Scholar] [CrossRef]
  728. Appleford, M.; Oh, S.; Cole, J.A.; Carnes, D.L.; Lee, M.; Bumgardner, J.D.; Haggard, W.O.; Ong, J.L. Effects of trabecular calcium phosphate scaffolds on stress signaling in osteoblast precursor cells. Biomaterials 2007, 28, 2747–2753. [Google Scholar] [CrossRef] [PubMed]
  729. Gamie, Z.; Tran, G.T.; Vyzas, G.; Korres, N.; Heliotis, M.; Mantalaris, A.; Tsiridis, E. Stem cells combined with bone graft substitutes in skeletal tissue engineering. Expert Opin. Biol. Ther. 2012, 12, 713–729. [Google Scholar] [CrossRef] [PubMed]
  730. Manfrini, M.; di Bona, C.; Canella, A.; Lucarelli, E.; Pellati, A.; d’Agostino, A.; Barbanti-Bròdano, G.; Tognon, M. Mesenchymal stem cells from patients to assay bone graft substitutes. J. Cell. Physiol. 2013, 228, 1229–1237. [Google Scholar] [CrossRef] [PubMed]
  731. Unger, R.E.; Sartoris, A.; Peters, K.; Motta, A.; Migliaresi, C.; Kunkel, M.; Bulnheim, U.; Rychly, J.; Kirkpatrick, C.J. Tissue-like self-assembly in cocultures of endothelial cells and osteoblasts and the formation of microcapillary like structures on three-dimensional porous biomaterials. Biomaterials 2007, 28, 3965–3976. [Google Scholar] [CrossRef] [PubMed]
  732. Nazir, N.M.; Dasmawati, M.; Azman, S.M.; Omar, N.S.; Othman, R. Biocompatibility of in house β-tricalcium phosphate ceramics with normal human osteoblast cell. J. Eng. Sci. Technol. 2012, 7, 169–176. [Google Scholar]
  733. Tan, F.; O’Neill, F.; Naciri, M.; Dowling, D.; Al-Rubeai, M. Cellular and transcriptomic analysis of human mesenchymal stem cell response to plasma-activated hydroxyapatite coating. Acta Biomater. 2012, 8, 1627–1638. [Google Scholar] [CrossRef] [PubMed]
  734. Li, B.; Liao, X.; Zheng, L.; He, H.; Wang, H.; Fan, H.; Zhang, X. Preparation and cellular response of porous A-type carbonated hydroxyapatite nanoceramics. Mater. Sci. Eng. C 2012, 32, 929–936. [Google Scholar] [CrossRef]
  735. Teixeira, S.; Fernandes, M.H.; Ferraz, M.P.; Monteiro, F.J. Proliferation and mineralization of bone marrow cells cultured on macroporous hydroxyapatite scaffolds functionalized with collagen type I for bone tissue regeneration. J. Biomed. Mater. Res. A 2010, 95, 1–8. [Google Scholar] [CrossRef] [PubMed]
  736. Yan-Zhong, Z.; Yan-Yan, H.; Jun, Z.; Shai-Hong, Z.; Zhi-You, L.; Ke-Chao, Z. Characteristics of functionalized nano-hydroxyapatite and internalization by human epithelial cell. Nanoscale Res. Lett. 2011, 6. [Google Scholar] [CrossRef]
  737. Borcard, F.; Staedler, D.; Comas, H.; Juillerat, F.K.; Sturzenegger, P.N.; Heuberger, R.; Gonzenbach, U.T.; Juillerat-Jeanneret, L.; Gerber-Lemaire, S. Chemical functionalization of bioceramics to enhance endothelial cells adhesion for tissue engineering. J. Med. Chem. 2012, 27, 7988–7997. [Google Scholar] [CrossRef]
  738. Treccani, L.; Klein, T.Y.; Meder, F.; Pardun, K.; Rezwan, K. Functionalized ceramics for biomedical, biotechnological and environmental applications. Acta Biomater. 2013, 9, 7115–7150. [Google Scholar] [CrossRef] [PubMed]
  739. Zhuang, Z.; Yoshimura, H.; Aizawa, M. Synthesis and ultrastructure of plate-like apatite single crystals as a model for tooth enamel. Mater. Sci. Eng. C 2013, 33, 2534–2540. [Google Scholar] [CrossRef]
  740. Zhuang, Z.; Fujimi, T.J.; Nakamura, M.; Konishi, T.; Yoshimura, H.; Aizawa, M. Development of a,b-plane-oriented hydroxyapatite ceramics as models for living bones and their cell adhesion behavior. Acta Biomater. 2013, 9, 6732–6740. [Google Scholar] [CrossRef] [PubMed]
  741. Freidlin, L.K.; Sharf, V.Z. Two paths for the dehydration of 1,4-butandiol to divinyl with a tricalcium phosphate catalyst. Bull. Acad. Sci. USSR Div. Chem. Sci. 1960, 9, 1577–1579. [Google Scholar] [CrossRef]
  742. Bett, J.A.S.; Christner, L.G.; Hall, W.K. Studies of the hydrogen held by solids. XII. Hydroxyapatite catalysts. J. Am. Chem. Soc. 1967, 89, 5535–5541. [Google Scholar] [CrossRef]
  743. Monma, H. Catalytic behavior of calcium phosphates for decompositions of 2-propanol and ethanol. J. Catal. 1982, 75, 200–203. [Google Scholar] [CrossRef]
  744. Tsuchida, T.; Yoshioka, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Synthesis of biogasoline from ethanol over hydroxyapatite catalyst. Ind. Eng. Chem. Res. 2008, 47, 1443–1452. [Google Scholar] [CrossRef]
  745. Tsuchida, T.; Kubo, J.; Yoshioka, T.; Sakuma, S.; Takeguchi, T.; Ueda, W. Reaction of ethanol over hydroxyapatite affected by Ca/P ratio of catalyst. J. Catal. 2008, 259, 183–189. [Google Scholar] [CrossRef]
  746. Xu, J.; White, T.; Li, P.; He, C.; Han, Y.F. Hydroxyapatite foam as a catalyst for formaldehyde combustion at room temperature. J. Am. Chem. Soc. 2010, 132, 13172–13173. [Google Scholar] [CrossRef] [PubMed]
  747. Hakim, L.; Yaakob, Z.; Ismail, M.; Daud, W.R.W. Nickel-hydroxyapatite as biomaterial catalysts for hydrogen production via glycerol steam reforming. Key Eng. Mater. 2010, 447–448, 770–774. [Google Scholar]
  748. Hatano, M.; Moriyama, K.; Maki, T.; Ishihara, K. Which is the actual catalyst: Chiral phosphoric acid or chiral calcium phosphate? Angew. Chem. Int. Ed. Engl. 2010, 49, 3823–3826. [Google Scholar] [CrossRef] [PubMed]
  749. Zhang, D.; Zhao, H.; Zhao, X.; Liu, Y.; Chen, H.; Li, X. Application of hydroxyapatite as catalyst and catalyst carrier. Prog. Chem. 2011, 23, 687–694. [Google Scholar]
  750. Gruselle, M.; Kanger, T.; Thouvenot, R.; Flambard, A.; Kriis, K.; Mikli, V.; Traksmaa, R.; Maaten, B.; Tõnsuaadu, K. Calcium hydroxyapatites as efficient catalysts for the michael C–C bond formation. ACS Catal. 2011, 1, 1729–1733. [Google Scholar] [CrossRef]
  751. Stošić, D.; Bennici, S.; Sirotin, S.; Calais, C.; Couturier, J.-L.; Dubois, J.-L.; Travert, A.; Auroux, A. Glycerol dehydration over calcium phosphate catalysts: Effect of acidic-basic features on catalytic performance. Appl. Catal. A 2012, 447–448, 124–134. [Google Scholar] [CrossRef]
  752. Ghantani, V.C.; Lomate, S.T.; Dongare, M.K.; Umbarkar, S.B. Catalytic dehydration of lactic acid to acrylic acid using calcium hydroxyapatite catalysts. Green Chem. 2013, 15, 1211–1217. [Google Scholar] [CrossRef]
  753. Urist, M.R.; Huo, Y.K.; Brownell, A.G.; Hohl, W.M.; Buyske, J.; Lietze, A.; Tempst, P.; Hunkapiller, M.; de Lange, R.J. Purification of bovine bone morphogenetic protein by hydroxyapatite chromatography. Proc. Natl. Acad. Sci. USA 1984, 81, 371–375. [Google Scholar] [CrossRef] [PubMed]
  754. Kawasaki, T. Hydroxyapatite as a liquid chromatographic packing. J. Chromatogr. 1991, 544, 147–184. [Google Scholar] [CrossRef]
  755. Kuiper, M.; Sanches, R.M.; Walford, J.A.; Slater, N.K.H. Purification of a functional gene therapy vector derived from moloney murine leukaemia virus using membrane filtration and ceramic hydroxyapatite chromatography. Biotechnol. Bioeng. 2002, 80, 445–453. [Google Scholar] [CrossRef] [PubMed]
  756. Jungbauer, A.; Hahn, R.; Deinhofer, K.; Luo, P. Performance and characterization of a nanophased porous hydroxyapatite for protein chromatography. Biotechnol. Bioeng. 2004, 87, 364–375. [Google Scholar] [CrossRef] [PubMed]
  757. Wensel, D.L.; Kelley, B.D.; Coffman, J.L. High-throughput screening of chromatographic separations: III. Monoclonal antibodies on ceramic hydroxyapatite. Biotechnol. Bioeng. 2008, 100, 839–854. [Google Scholar] [CrossRef] [PubMed]
  758. Hilbrig, F.; Freitag, R. Isolation and purification of recombinant proteins, antibodies and plasmid DNA with hydroxyapatite chromatography. Biotechnol. J. 2012, 7, 90–102. [Google Scholar] [CrossRef] [PubMed]
  759. Hollister, S.J. Porous scaffold design for tissue engineering. Nat. Mater. 2005, 4, 518–524. [Google Scholar] [CrossRef] [PubMed]
  760. Jones, J.R.; Hench, L.L. Regeneration of trabecular bone using porous ceramics. Curr. Opin. Solid State Mater. Sci. 2003, 7, 301–307. [Google Scholar] [CrossRef]
  761. Griffith, L.G.; Naughton, G. Tissue engineering—Current challenges and expanding opportunities. Science 2002, 295, 1009–1014. [Google Scholar] [CrossRef] [PubMed]
  762. Goldberg, V.M.; Caplan, A.I. Orthopedic Tissue Engineering Basic Science and Practice; Marcel Dekker: New York, NY, USA, 2004; p. 338. [Google Scholar]
  763. Tissue Engineering; van Blitterswijk, C.A.; Thomsen, P.; Hubbell, J.; Cancedda, R.; de Bruijn, J.D.; Lindahl, A.; Sohier, J.; Williams, D.F. (Eds.) Academic Press: Burlington, MA, USA, 2008; p. 760.
  764. Ikada, Y. Challenges in tissue engineering. J. R. Soc. Interface 2006, 3, 589–601. [Google Scholar] [CrossRef] [PubMed]
  765. Cima, L.G.; Langer, R. Engineering human tissue. Chem. Eng. Prog. 1993, 89, 46–54. [Google Scholar]
  766. Langer, R.; Vacanti, J.P. Tissue engineering. Science 1993, 260, 920–926. [Google Scholar] [CrossRef] [PubMed]
  767. El-Ghannam, A. Bone reconstruction: From bioceramics to tissue engineering. Expert Rev. Med. Dev. 2005, 2, 87–101. [Google Scholar] [CrossRef]
  768. Kneser, U.; Schaefer, D.J.; Polykandriotis, E.; Horch, R.E. Tissue engineering of bone: The reconstructive surgeon’s point of view. J. Cell. Mol. Med. 2006, 10, 7–19. [Google Scholar] [CrossRef] [PubMed]
  769. Scott, T.G.; Blackburn, G.; Ashley, M.; Bayer, I.S.; Ghosh, A.; Biris, A.S.; Biswas, A. Advances in bionanomaterials for bone tissue engineering. J. Nanosci. Nanotechnol. 2013, 13, 1–22. [Google Scholar] [CrossRef] [PubMed]
  770. Lutolf, M.P.; Hubbell, J.A. Synthetic biomaterials as instructive extracellular microenvironments for morphogenesis in tissue engineering. Nat. Biotechnol. 2005, 23, 47–55. [Google Scholar] [CrossRef] [PubMed]
  771. Ma, P.X. Biomimetic materials for tissue engineering. Adv. Drug. Deliv. Rev. 2008, 60, 184–198. [Google Scholar] [CrossRef] [PubMed]
  772. Yang, S.; Leong, K.F.; Du, Z.; Chua, C.K. The design of scaffolds for use in tissue engineering. Part I. Traditional factors. Tissue Eng. 2001, 7, 679–689. [Google Scholar] [CrossRef] [PubMed]
  773. Hutmacher, D.W. Scaffolds in tissue engineering bone and cartilage. Biomaterials 2000, 21, 2529–2543. [Google Scholar] [CrossRef] [PubMed]
  774. Ma, P.X. Scaffolds for tissue fabrication. Mater. Today 2004, 7, 30–40. [Google Scholar] [CrossRef]
  775. Yasuhiko, T. Biomaterial technology for tissue engineering applications. J. R. Soc. Interface 2009, 6, S311–S324. [Google Scholar] [CrossRef] [PubMed]
  776. Scaffolding in Tissue Engineering; Ma, P.X.; Elisseeff, J. (Eds.) CRC Press: Boca Raton, FL, USA, 2006; p. 638.
  777. Schieker, M.; Seitz, H.; Drosse, I.; Seitz, S.; Mutschler, W. Biomaterials as scaffold for bone tissue engineering. Eur. J. Trauma 2006, 32, 114–124. [Google Scholar] [CrossRef]
  778. Freed, L.E.; Guilak, F.; Guo, X.E.; Gray, M.L.; Tranquillo, R.; Holmes, J.W.; Radisic, M.; Sefton, M.V.; Kaplan, D.; Vunjak-Novakovic, G. Advanced tools for tissue engineering: Scaffolds, bioreactors, and signaling. Tissue Eng. 2006, 12, 3285–3305. [Google Scholar] [CrossRef] [PubMed]
  779. Gandaglia, A.; Bagno, A.; Naso, F.; Spina, M.; Gerosa, G. Cells, scaffolds and bioreactors for tissue-engineered heart valves: A journey from basic concepts to contemporary developmental innovations. Eur. J. Cardio-Thorac. Surg. 2011, 39, 523–531. [Google Scholar] [CrossRef]
  780. Hui, J.H.P.; Buhary, K.S.; Chowdhary, A. Implantation of orthobiologic, biodegradable scaffolds in osteochondral repair. Orthop. Clin. North Am. 2012, 43, 255–261. [Google Scholar] [CrossRef] [PubMed]
  781. Vanderleyden, E.; Mullens, S.; Luyten, J.; Dubruel, P. Implantable (bio)polymer coated titanium scaffolds: A review. Curr. Pharm. Des. 2012, 18, 2576–2590. [Google Scholar] [CrossRef] [PubMed]
  782. Service, R.F. Tissue engineers build new bone. Science 2000, 289, 1498–1500. [Google Scholar] [CrossRef] [PubMed]
  783. Deligianni, D.D.; Katsala, N.D.; Koutsoukos, P.G.; Missirlis, Y.F. Effect of surface roughness of hydroxyapatite on human bone marrow cell adhesion, proliferation, differentiation and detachment strength. Biomaterials 2001, 22, 87–96. [Google Scholar] [CrossRef] [PubMed]
  784. Fini, M.; Giardino, R.; Borsari, V.; Torricelli, P.; Rimondini, L.; Giavaresi, G.; Aldini, N.N. In vitro behaviour of osteoblasts cultured on orthopaedic biomaterials with different surface roughness, uncoated and fluorohydroxyapatite-coated, relative to the in vivo osteointegration rate. Int. J. Artif. Organs 2003, 26, 520–528. [Google Scholar] [PubMed]
  785. Sato, M.; Webster, T.J. Designing orthopedic implant surfaces: Harmonization of nanotopographical and chemical aspects. Nanomedicine 2006, 1, 351–354. [Google Scholar] [CrossRef] [PubMed]
  786. Li, X.; van Blitterswijk, C.A.; Feng, Q.; Cui, F.; Watari, F. The effect of calcium phosphate microstructure on bone-related cells in vitro. Biomaterials 2008, 29, 3306–3316. [Google Scholar] [CrossRef] [PubMed]
  787. Kumar, G.; Waters, M.S.; Farooque, T.M.; Young, M.F.; Simon, C.G. Freeform fabricated scaffolds with roughened struts that enhance both stem cell proliferation and differentiation by controlling cell shape. Biomaterials 2012, 33, 4022–4030. [Google Scholar] [CrossRef] [PubMed]
  788. Zhou, Y.; Chen, F.; Ho, S.T.; Woodruff, M.A.; Lim, T.M.; Hutmacher, D.W. Combined marrow stromal cell-sheet techniques and high-strength biodegradable composite scaffolds for engineered functional bone grafts. Biomaterials 2007, 28, 814–824. [Google Scholar] [CrossRef] [PubMed]
  789. Vitale-Brovarone, C.; Baino, F.; Verné, E. High strength bioactive glass-ceramic scaffolds for bone regeneration. J. Mater. Sci. Mater. Med. 2009, 20, 643–653. [Google Scholar] [CrossRef] [PubMed]
  790. Annaz, B.; Hing, K.A.; Kayser, M.; Buckland, T.; di Silvio, L. An ultrastructural study of cellular response to variation in porosity in phase-pure hydroxyapatite. J. Microscopy 2004, 216, 97–109. [Google Scholar] [CrossRef]
  791. Li, S.; de Wijn, J.R.; Li, J.; Layrolle, P.; de Groot, K. Macroporous biphasic calcium phosphate scaffold with high permeability/porosity ratio. Tissue Eng. 2003, 9, 535–548. [Google Scholar] [CrossRef] [PubMed]
  792. Ebaretonbofa, E.; Evans, J.R. High porosity hydroxyapatite foam scaffolds for bone substitute. J. Porous Mater. 2002, 9, 257–263. [Google Scholar] [CrossRef]
  793. Specchia, N.; Pagnotta, A.; Cappella, M.; Tampieri, A.; Greco, F. Effect of hydroxyapatite porosity on growth and differentiation of human osteoblast-like cells. J. Mater. Sci. 2002, 37, 577–584. [Google Scholar] [CrossRef]
  794. Hing, K.A. Bioceramic bone graft substitutes: Influence of porosity and chemistry. Int. J. Appl. Ceram. Technol. 2005, 2, 184–199. [Google Scholar] [CrossRef]
  795. Malmström, J.; Adolfsson, E.; Arvidsson, A.; Thomsen, P. Bone response inside free-form fabricated macroporous hydroxyapatite scaffolds with and without an open microporosity. Clin. Implant Dent. Relat. Res. 2007, 9, 79–88. [Google Scholar] [CrossRef] [PubMed]
  796. Teixeira, C.C.; Nemelivsky, Y.; Karkia, C.; LeGeros, R.Z. Biphasic calcium phosphate: A scaffold for growth plate chondrocyte maturation. Tissue Eng. 2006, 12, 2283–2289. [Google Scholar] [CrossRef] [PubMed]
  797. Teixeira, S.; Oliveira, S.; Ferraz, M.P.; Monteiro, F.J. Three dimensional macroporous calcium phosphate scaffolds for bone tissue engineering. Key Eng. Mater. 2008, 361–363, 947–950. [Google Scholar] [CrossRef]
  798. Peng, Q.; Jiang, F.; Huang, P.; Zhou, S.; Weng, J.; Bao, C.; Zhang, C.; Yu, H. A novel porous bioceramics scaffold by accumulating hydroxyapatite spherules for large bone tissue engineering in vivo. I. Preparation and characterization of scaffold. J. Biomed. Mater. Res. A 2010, 93, 920–929. [Google Scholar] [PubMed]
  799. Lew, K.S.; Othman, R.; Ishikawa, K.; Yeoh, F.Y. Macroporous bioceramics: A remarkable material for bone regeneration. J. Biomater. Appl. 2012, 27, 345–358. [Google Scholar] [CrossRef] [PubMed]
  800. Ren, L.M.; Todo, M.; Arahira, T.; Yoshikawa, H.; Myoui, A. A comparative biomechanical study of bone ingrowth in two porous hydroxyapatite bioceramics. Appl. Surf. Sci. 2012, 262, 81–88. [Google Scholar] [CrossRef]
  801. Stevens, M.M. Biomaterials for bone tissue engineering. Mater. Today 2008, 11, 18–25. [Google Scholar] [CrossRef]
  802. Artzi, Z.; Weinreb, M.; Givol, N.; Rohrer, M.D.; Nemcovsky, C.E.; Prasad, H.S.; Tal, H. Biomaterial resorbability and healing site morphology of inorganic bovine bone and beta tricalcium phosphate in the canine: A 24-month longitudinal histologic study and morphometric analysis. Int. J. Oral Maxillofic. Implant. 2004, 19, 357–368. [Google Scholar]
  803. Burg, K.J.L.; Porter, S.; Kellam, J.F. Biomaterial developments for bone tissue engineering. Biomaterials 2000, 21, 2347–2359. [Google Scholar] [CrossRef] [PubMed]
  804. Ajaal, T.T.; Smith, R.W. Employing the Taguchi method in optimizing the scaffold production process for artificial bone grafts. J. Mater. Process. Technol. 2009, 209, 1521–1532. [Google Scholar] [CrossRef]
  805. Bohner, M.; Loosli, Y.; Baroud, G.; Lacroix, D. Commentary: Deciphering the link between architecture and biological response of a bone graft substitute. Acta Biomater. 2011, 7, 478–484. [Google Scholar] [CrossRef] [PubMed]
  806. Peppas, N.A.; Langer, R. New challenges in biomaterials. Science 1994, 263, 1715–1720. [Google Scholar] [CrossRef] [PubMed]
  807. Hench, L.L. Biomaterials: A forecast for the future. Biomaterials 1998, 19, 1419–1423. [Google Scholar] [CrossRef] [PubMed]
  808. Barrère, F.; Mahmood, T.A.; de Groot, K.; van Blitterswijk, C.A. Advanced biomaterials for skeletal tissue regeneration: Instructive and smart functions. Mater. Sci. Eng. R 2008, 59, 38–71. [Google Scholar] [CrossRef]
  809. Dellinger, J.G.; Eurell, J.A.C.; Jamison, R.D. Bone response to 3D periodic hydroxyapatite scaffolds with and without tailored microporosity to deliver bone morphogenetic protein 2. J. Biomed. Mater. Res. A 2006, 76, 366–376. [Google Scholar] [CrossRef] [PubMed]
  810. Bae, C.J.; Kim, H.W.; Koh, Y.H.; Kim, H.E. Hydroxyapatite (HA) bone scaffolds with controlled macrochannel pores. J. Mater. Sci. Mater. Med. 2006, 17, 517–521. [Google Scholar] [CrossRef] [PubMed]
  811. Guo, H.; Wei, J.; Kong, H.; Liu, C.; Pan, K. Biocompatibility and osteogenesis of calcium phosphate cement scaffolds for bone tissue engineering. Adv. Mater. Res. 2008, 47–50, 1383–1386. [Google Scholar] [CrossRef]
  812. Guo, H.; Su, J.; Wei, J.; Kong, H.; Liu, C. Biocompatibility and osteogenicity of degradable Ca-deficient hydroxyapatite scaffolds from calcium phosphate cement for bone tissue engineering. Acta Biomater. 2009, 5, 268–278. [Google Scholar] [CrossRef] [PubMed]
  813. Wilson, C.E.; van Blitterswijk, C.A.; Verbout, A.J.; Dhert, W.J.A.; de Bruijn, J.D. Scaffolds with a standardized macro-architecture fabricated from several calcium phosphate ceramics using an indirect rapid prototyping technique. J. Mater. Sci. Mater. Med. 2011, 22, 97–105. [Google Scholar] [CrossRef] [PubMed]
  814. Schumacher, M.; Deisinger, U.; Detsch, R.; Ziegler, G. Indirect rapid prototyping of biphasic calcium phosphate scaffolds as bone substitutes: Influence of phase composition, macroporosity and pore geometry on mechanical properties. J. Mater. Sci. Mater. Med. 2010, 21, 3119–3127. [Google Scholar] [CrossRef] [PubMed]
  815. Tripathi, G.; Basu, B. A porous hydroxyapatite scaffold for bone tissue engineering: Physico-mechanical and biological evaluations. Ceram. Int. 2012, 38, 341–349. [Google Scholar] [CrossRef]
  816. Mayr, H.O.; Klehm, J.; Schwan, S.; Hube, R.; Südkamp, N.P.; Niemeyer, P.; Salzmann, G.; von Eisenhardt-Rothe, R.; Heilmann, A.; Bohner, M.; Bernstein, A. Microporous calcium phosphate ceramics as tissue engineering scaffolds for the repair of osteochondral defects: Biomechanical results. Acta Biomater. 2013, 9, 4845–4855. [Google Scholar] [CrossRef] [PubMed]
  817. Bernstein, A.; Niemeyer, P.; Salzmann, G.; Südkamp, N.P.; Hube, R.; Klehm, J.; Menzel, M.; von Eisenhart-Rothe, R.; Bohner, M.; Görz, L.; Mayr, H.O. Microporous calcium phosphate ceramics as tissue engineering scaffolds for the repair of osteochondral defects: Histological results. Acta Biomater. 2013, 9, 7490–7505. [Google Scholar] [CrossRef] [PubMed]
  818. Liu, H.; Webster, T.J. Nanomedicine for implants: A review of studies and necessary experimental tools. Biomaterials 2007, 28, 354–369. [Google Scholar] [CrossRef] [PubMed]
  819. Wang, C.; Duan, Y.; Markovic, B.; Barbara, J.; Howlett, C.R.; Zhang, X.; Zreiqat, H. Proliferation and bone-related gene expression of osteoblasts grown on hydroxyapatite ceramics sintered at different temperature. Biomaterials 2004, 25, 2949–2956. [Google Scholar] [CrossRef] [PubMed]
  820. Matsumoto, T.; Okazaki, M.; Nakahira, A.; Sasaki, J.; Egusa, H.; Sohmura, T. Modification of apatite materials for bone tissue engineering and drug delivery carriers. Curr. Med. Chem. 2007, 14, 2726–2733. [Google Scholar] [CrossRef] [PubMed]
  821. Chai, Y.C.; Carlier, A.; Bolander, J.; Roberts, S.J.; Geris, L.; Schrooten, J.; van Oosterwyck, H.; Luyten, F.P. Current views on calcium phosphate osteogenicity and the translation into effective bone regeneration strategies. Acta Biomater. 2012, 8, 3876–3887. [Google Scholar] [CrossRef] [PubMed]
  822. Traykova, T.; Aparicio, C.; Ginebra, M.P.; Planell, J.A. Bioceramics as nanomaterials. Nanomedicine 2006, 1, 91–106. [Google Scholar] [CrossRef] [PubMed]
  823. Kalita, S.J.; Bhardwaj, A.; Bhatt, H.A. Nanocrystalline calcium phosphate ceramics in biomedical engineering. Mater. Sci. Eng. C 2007, 27, 441–449. [Google Scholar] [CrossRef]
  824. Dorozhkin, S.V. Nanodimensional and nanocrystalline calcium orthophosphates. Int. J. Chem. Mater. Sci. 2013, 1, 105–174. [Google Scholar]
  825. Dorozhkin, S.V. Amorphous calcium orthophosphates: Nature, chemistry and biomedical applications. Int. J. Mater. Chem. 2012, 2, 19–46. [Google Scholar] [CrossRef]
  826. Aizawa, M.; Porter, A.E.; Best, S.M.; Bonfield, W. Ultrastructural observation of single-crystal apatite fibres. Biomaterials 2005, 26, 3427–3433. [Google Scholar] [CrossRef] [PubMed]
  827. Park, Y.M.; Ryu, S.C.; Yoon, S.Y.; Stevens, R.; Park, H.C. Preparation of whisker-shaped hydroxyapatite/β-tricalcium phosphate composite. Mater. Chem. Phys. 2008, 109, 440–447. [Google Scholar] [CrossRef]
  828. Morisue, H.; Matsumoto, M.; Chiba, K.; Matsumoto, H.; Toyama, Y.; Aizawa, M.; Kanzawa, N.; Fujimi, T.J.; Uchida, H.; Okada, I. A novel hydroxyapatite fiber mesh as a carrier for recombinant human bone morphogenetic protein-2 enhances bone union in rat posterolateral fusion model. Spine 2006, 31, 1194–1200. [Google Scholar] [CrossRef] [PubMed]
  829. Matsuda, A.; Ikoma, T.; Kobayashi, H.; Tanaka, J. Preparation and mechanical property of core-shell type chitosan/calcium phosphate composite fiber. Mater. Sci. Eng. C 2004, 24, 723–728. [Google Scholar] [CrossRef]
  830. Tas, A.C. Molten salt synthesis of calcium hydroxyapatite whiskers. J. Am. Ceram. Soc. 2001, 84, 295–300. [Google Scholar] [CrossRef]
  831. Aizawa, M.; Ueno, H.; Itatani, K.; Okada, I. Syntheses of calcium-deficient apatite fibres by a homogeneous precipitation method and their characterizations. J. Eur. Ceram. Soc. 2006, 26, 501–507. [Google Scholar] [CrossRef]
  832. Wu, Y.; Hench, L.L.; Du, J.; Choy, K.L.; Guo, J. Preparation of hydroxyapatite fibers by electrospinning technique. J. Am. Ceram. Soc. 2004, 87, 1988–1991. [Google Scholar] [CrossRef]
  833. Ramanan, S.R.; Venkatesh, R. A study of hydroxyapatite fibers prepared via sol-gel route. Mater. Lett. 2004, 58, 3320–3323. [Google Scholar] [CrossRef]
  834. Seo, D.S.; Lee, J.K. Synthesis of hydroxyapatite whiskers through dissolution-reprecipitation process using EDTA. J. Cryst. Growth 2008, 310, 2162–2167. [Google Scholar] [CrossRef]
  835. Tas, A.C. Formation of calcium phosphate whiskers in hydrogen peroxide (H2O2) solutions at 90 °C. J. Am. Ceram. Soc. 2007, 90, 2358–2362. [Google Scholar] [CrossRef]
  836. Neira, I.S.; Guitián, F.; Taniguchi, T.; Watanabe, T.; Yoshimura, M. Hydrothermal synthesis of hydroxyapatite whiskers with sharp faceted hexagonal morphology. J. Mater. Sci. 2008, 43, 2171–2178. [Google Scholar] [CrossRef]
  837. Yang, H.Y.; Yang, S.F.; Chi, X.P.; Evans, J.R.G.; Thompson, I.; Cook, R.J.; Robinson, P. Sintering behaviour of calcium phosphate filaments for use as hard tissue scaffolds. J. Eur. Ceram. Soc. 2008, 28, 159–167. [Google Scholar] [CrossRef]
  838. Cardoso, G.B.C.; Ramos, S.L.F.; Rodas, A.C.D.; Higa, O.Z.; Zavaglia, C.A.C.; Arruda, A.C.F. Scaffolds of poly (ε-caprolactone) with whiskers of hydroxyapatite. J. Mater. Sci. 2010, 45, 4990–4993. [Google Scholar] [CrossRef]
  839. Zhang, H.; Zhu, Q. Synthesis of nanospherical and ultralong fibrous hydroxyapatite and reinforcement of biodegradable chitosan/hydroxyapatite composite. Mod. Phys. Lett. B 2009, 23, 3967–3976. [Google Scholar] [CrossRef]
  840. Ribeiro, C.C.; Barrias, C.C.; Barbosa, M.A. Preparation and characterisation of calcium-phosphate porous microspheres with a uniform size for biomedical applications. J. Mater. Sci. Mater. Med. 2006, 17, 455–463. [Google Scholar] [CrossRef] [PubMed]
  841. Ribeiro, C.C.; Barrias, C.C.; Barbosa, M.A. Calcium phosphate-alginate microspheres as enzyme delivery matrices. Biomaterials 2004, 25, 4363–4373. [Google Scholar] [CrossRef] [PubMed]
  842. Zhou, W.Y.; Wang, M.; Cheung, W.L.; Guo, B.C.; Jia, D.M. Synthesis of carbonated hydroxyapatite nanospheres through nanoemulsion. J. Mater. Sci. Mater. Med. 2008, 19, 103–110. [Google Scholar] [CrossRef] [PubMed]
  843. Kim, H.W.; Gu, H.J.; Lee, H.H. Microspheres of collagen-apatite nanocomposites with osteogenic potential for tissue engineering. Tissue Eng. 2007, 13, 965–973. [Google Scholar] [CrossRef] [PubMed]
  844. Kawai, T.; Sekikawa, H.; Unuma, H. Preparation of hollow hydroxyapatite microspheres utilizing poly(divinylbenzene) as a template. J. Ceram. Soc. Jpn. 2009, 117, 340–343. [Google Scholar] [CrossRef]
  845. Descamps, M.; Hornez, J.C.; Leriche, A. Manufacture of hydroxyapatite beads for medical applications. J. Eur. Ceram. Soc. 2009, 29, 369–375. [Google Scholar] [CrossRef]
  846. Cho, J.S.; Jung, D.S.; Han, J.M.; Kang, Y.C. Spherical shape hydroxyapatite powders prepared by flame spray pyrolysis. J. Ceram. Process. Res. 2008, 9, 348–352. [Google Scholar]
  847. Kimura, I.; Honma, T.; Riman, R.E. Preparation of hydroxyapatite microspheres by interfacial reaction in a multiple emulsion. J. Ceram. Soc. Jpn. 2007, 115, 888–893. [Google Scholar] [CrossRef]
  848. Yao, A.; Ai, F.; Liu, X.; Wang, D.; Huang, W.; Xu, W. Preparation of hollow hydroxyapatite microspheres by the conversion of borate glass at near room temperature. Mater. Res. Bull. 2010, 45, 25–28. [Google Scholar] [CrossRef]
  849. Jiang, D.; Chen, M.; Li, D.; Zhu, J.; Lü, X.; Xie, J. One-pot synthesis of hierarchical fluorapatite hollow microparticles. Mater. Lett. 2009, 63, 2639–2642. [Google Scholar] [CrossRef]
  850. Lim, J.H.; Park, J.H.; Park, E.K.; Kim, H.J.; Park, I.K.; Shin, H.Y.; Shin, H.I. Fully interconnected globular porous biphasic calcium phosphate ceramic scaffold facilitates osteogenic repair. Key Eng. Mater. 2008, 361–363, 119–122. [Google Scholar] [CrossRef]
  851. Cho, J.S.; Ko, Y.N.; Koo, H.Y.; Kang, Y.C. Synthesis of nano-sized biphasic calcium phosphate ceramics with spherical shape by flame spray pyrolysis. J. Mater. Sci. Mater. Med. 2010, 21, 1143–1149. [Google Scholar] [CrossRef] [PubMed]
  852. Ye, F.; Guo, H.; Zhang, H.; He, X. Polymeric micelle-templated synthesis of hydroxyapatite hollow nanoparticles for a drug delivery system. Acta Biomater. 2010, 6, 2212–2218. [Google Scholar] [CrossRef] [PubMed]
  853. He, W.; Tao, J.; Pan, H.; Xu, R.; Tang, R. A size-controlled synthesis of hollow apatite nanospheres at water-oil interfaces. Chem. Lett. 2010, 39, 674–675. [Google Scholar] [CrossRef]
  854. Itatani, K.; Tsugawa, T.; Umeda, T.; Musha, Y.; Davies, I.J.; Koda, S. Preparation of submicrometer-sized porous spherical hydroxyapatite agglomerates by ultrasonic spray pyrolysis technique. J. Ceram. Soc. Jpn. 2010, 118, 462–466. [Google Scholar] [CrossRef]
  855. Xiao, W.; Fu, H.; Rahaman, M.N.; Liu, Y.; Bal, B.S. Hollow hydroxyapatite microspheres: A novel bioactive and osteoconductive carrier for controlled release of bone morphogenetic protein-2 in bone regeneration. Acta Biomater. 2013, 9, 8374–8383. [Google Scholar] [CrossRef] [PubMed]
  856. Bohner, M.; Tadier, S.; van Garderen, N.; de Gasparo, A.; Döbelin, N.; Baroud, G. Synthesis of spherical calcium phosphate particles for dental and orthopedic applications. Biomatter 2013, 3, e25103:1–e25103:15. [Google Scholar] [CrossRef]
  857. Kamitakahara, M.; Murakami, S.; Takahashi, H.; Watanabe, N.; Ioku, K. Formation of hydroxyapatite microtubes assisted with anatase under hydrothermal conditions. Chem. Lett. 2010, 39, 854–855. [Google Scholar] [CrossRef]
  858. Chandanshive, B.; Dyondi, D.; Ajgaonkar, V.R.; Banerjee, R.; Khushalani, D. Biocompatible calcium phosphate based tubes. J. Mater. Chem. 2010, 20, 6923–6928. [Google Scholar] [CrossRef]
  859. Kamitakahara, M.; Takahashi, H.; Ioku, K. Tubular hydroxyapatite formation through a hydrothermal process from α-tricalcium phosphate with anatase. J. Mater. Sci. 2012, 47, 4194–4199. [Google Scholar] [CrossRef]
  860. Ustundag, C.B.; Kaya, F.; Kamitakahara, M.; Kaya, C.; Ioku, K. Production of tubular porous hydroxyapatite using electrophoretic deposition. J. Ceram. Soc. Jpn. 2012, 120, 569–573. [Google Scholar] [CrossRef]
  861. Nonoyama, T.; Kinoshita, T.; Higuchi, M.; Nagata, K.; Tanaka, M.; Kamada, M.; Sato, K.; Kato, K. Arrangement techniques of proteins and cells using amorphous calcium phosphate nanofiber scaffolds. Appl. Surf. Sci. 2012, 262, 8–12. [Google Scholar] [CrossRef]
  862. Lin, K.; Chen, L.; Qu, H.; Lu, J.; Chang, J. Improvement of mechanical properties of macroporous β-tricalcium phosphate bioceramic scaffolds with uniform and interconnected pore structures. Ceram. Int. 2011, 37, 2397–2403. [Google Scholar] [CrossRef]
  863. Simon, J.L.; Michna, S.; Lewis, J.A.; Rekow, E.D.; Thompson, V.P.; Smay, J.E.; Yampolsky, A.; Parsons, J.R.; Ricci, J.L. In vivo bone response to 3D periodic hydroxyapatite scaffolds assembled by direct ink writing. J. Biomed. Mater. Res. A 2007, 83, 747–758. [Google Scholar] [CrossRef] [PubMed]
  864. Chu, T.M.; Orton, D.G.; Hollister, S.J.; Feinberg, S.E.; Halloran, J.W. Mechanical and in vivo performance of hydroxyapatite implants with controlled architectures. Biomaterials 2002, 23, 1283–1293. [Google Scholar] [CrossRef] [PubMed]
  865. Yoshikawa, H.; Myoui, A. Bone tissue engineering with porous hydroxyapatite ceramics. J. Artif. Organs 2005, 8, 131–136. [Google Scholar] [CrossRef] [PubMed]
  866. Detsch, R.; Uhl, F.; Deisinger, U.; Ziegler, G. In vitro studies of cell growth on three differently fabricated hydroxyapatite ceramic scaffolds for bone tissue engineering. Key Eng. Mater. 2008, 361–363, 1181–1184. [Google Scholar] [CrossRef]
  867. Min, S.H.; Jin, H.H.; Park, H.Y.; Park, I.M.; Park, H.C.; Yoon, S.Y. Preparation of porous hydroxyapatite scaffolds for bone tissue engineering. Mater. Sci. Forum 2006, 510–511, 754–757. [Google Scholar] [CrossRef]
  868. Deville, S.; Saiz, E.; Nalla, R.K.; Tomsia, A.P. Strong biomimetic hydroxyapatite scaffolds. Adv. Sci. Technol. 2006, 49, 148–152. [Google Scholar] [CrossRef]
  869. Buckley, C.T.; O’Kelly, K.U. Fabrication and characterization of a porous multidomain hydroxyapatite scaffold for bone tissue engineering investigations. J. Biomed. Mater. Res. B 2010, 93, 459–467. [Google Scholar] [CrossRef]
  870. Yang, W.; Zhou, D.; Yin, G.; Chen, H. Progress of biphasic calcium phosphate bioceramic as scaffold materials of bone tissue engineering. J. Chin. Ceram. Soc. 2004, 32, 1143–1149. [Google Scholar]
  871. Ramay, H.R.R.; Zhang, M. Biphasic calcium phosphate nanocomposite porous scaffolds for load-bearing bone tissue engineering. Biomaterials 2004, 25, 5171–5180. [Google Scholar] [CrossRef] [PubMed]
  872. Chen, G.; Li, W.; Zhao, B.; Sun, K. A novel biphasic bone scaffold: β-calcium phosphate and amorphous calcium polyphosphate. J. Am. Ceram. Soc. 2009, 92, 945–948. [Google Scholar] [CrossRef]
  873. Guo, D.; Xu, K.; Han, Y. The in situ synthesis of biphasic calcium phosphate scaffolds with controllable compositions, structures, and adjustable properties. J. Biomed. Mater. Res. A 2009, 88, 43–52. [Google Scholar] [CrossRef] [PubMed]
  874. Sarin, P.; Lee, S.J.; Apostolov, Z.D.; Kriven, W.M. Porous biphasic calcium phosphate scaffolds from cuttlefish bone. J. Am. Ceram. Soc. 2011, 94, 2362–2370. [Google Scholar] [CrossRef]
  875. Furuichi, K.; Oaki, Y.; Ichimiya, H.; Komotori, J.; Imai, H. Preparation of hierarchically organized calcium phosphate-organic polymer composites by calcification of hydrogel. Sci. Technol. Adv. Mater. 2006, 7, 219–225. [Google Scholar] [CrossRef]
  876. Wei, J.; Jia, J.; Wu, F.; Wei, S.; Zhou, H.; Zhang, H.; Shin, J.W.; Liu, C. Hierarchically microporous/macroporous scaffold of magnesium-calcium phosphate for bone tissue regeneration. Biomaterials 2010, 31, 1260–1269. [Google Scholar] [CrossRef] [PubMed]
  877. Gbureck, U.; Grolms, O.; Barralet, J.E.; Grover, L.M.; Thull, R. Mechanical activation and cement formation of β-tricalcium phosphate. Biomaterials 2003, 24, 4123–4131. [Google Scholar] [CrossRef] [PubMed]
  878. Gbureck, U.; Barralet, J.E.; Hofmann, M.; Thull, R. Mechanical activation of tetracalcium phosphate. J. Am. Ceram. Soc. 2004, 87, 311–313. [Google Scholar] [CrossRef]
  879. Bohner, M.; Luginbühl, R.; Reber, C.; Doebelin, N.; Baroud, G.; Conforto, E. A physical approach to modify the hydraulic reactivity of α-tricalcium phosphate powder. Acta Biomater. 2009, 5, 3524–3535. [Google Scholar] [CrossRef] [PubMed]
  880. Hagio, T.; Tanase, T.; Akiyama, J.; Iwai, K.; Asai, S. Formation and biological affinity evaluation of crystallographically aligned hydroxyapatite. J. Ceram. Soc. Jpn. 2008, 116, 79–82. [Google Scholar] [CrossRef]
  881. Blawas, A.S.; Reichert, W.M. Protein patterning. Biomaterials 1998, 19, 595–609. [Google Scholar] [CrossRef] [PubMed]
  882. Kasai, T.; Sato, K.; Kanematsu, Y.; Shikimori, M.; Kanematsu, N.; Doi, Y. Bone tissue engineering using porous carbonate apatite and bone marrow cells. J. Craniofac. Surg. 2010, 21, 473–478. [Google Scholar] [CrossRef] [PubMed]
  883. Wang, L.; Fan, H.; Zhang, Z.Y.; Lou, A.J.; Pei, G.X.; Jiang, S.; Mu, T.W.; Qin, J.J.; Chen, S.Y.; Jin, D. Osteogenesis and angiogenesis of tissue-engineered bone constructed by prevascularized β-tricalcium phosphate scaffold and mesenchymal stem cells. Biomaterials 2010, 31, 9452–9461. [Google Scholar] [CrossRef] [PubMed]
  884. Daculsi, G.; Miramond, T.; Borget, P.; Baroth, S. Smart calcium phosphate bioceramic scaffold for bone tissue engineering. Key Eng. Mater. 2013, 529–530, 19–23. [Google Scholar]
  885. Sánchez-Salcedo, S.; Izquierdo-Barba, I.; Arcos, D.; Vallet-Regí, M. In vitro evaluation of potential calcium phosphate scaffolds for tissue engineering. Tissue Eng. 2006, 12, 279–290. [Google Scholar] [CrossRef] [PubMed]
  886. Meganck, J.A.; Baumann, M.J.; Case, E.D.; McCabe, L.R.; Allar, J.N. Biaxial flexure testing of calcium phosphate bioceramics for use in tissue engineering. J. Biomed. Mater. Res. A 2005, 72, 115–126. [Google Scholar] [CrossRef] [PubMed]
  887. Case, E.D.; Smith, I.O.; Baumann, M.J. Microcracking and porosity in calcium phosphates and the implications for bone tissue engineering. Mater. Sci. Eng. A 2005, 390, 246–254. [Google Scholar] [CrossRef]
  888. Sibilla, P.; Sereni, A.; Aguiari, G.; Banzi, M.; Manzati, E.; Mischiati, C.; Trombelli, L.; del Senno, L. Effects of a hydroxyapatite-based biomaterial on gene expression in osteoblast-like cells. J. Dent. Res. 2006, 85, 354–358. [Google Scholar] [CrossRef] [PubMed]
  889. Verron, E.; Bouler, J.M. Calcium phosphate ceramics as bone drug-combined devices. Key Eng. Mater. 2010, 441, 181–201. [Google Scholar] [CrossRef]
  890. Zhou, T.H.; Su, M.; Shang, B.C.; Ma, T.; Xu, G.L.; Li, H.L.; Chen, Q.H.; Sun, W.; Xu, Y.Q. Nano-hydroxyapatite/β-tricalcium phosphate ceramics scaffolds loaded with cationic liposomal ceftazidime: Preparation, release characteristics in vitro and inhibition to Staphylococcus aureus biofilms. Drug Dev. Ind. Pharm. 2012, 38, 1298–1304. [Google Scholar] [CrossRef] [PubMed]
  891. Rapoport, A.; Borovikova, D.; Kokina, A.; Patmalnieks, A.; Polyak, N.; Pavlovska, I.; Mezinskis, G.; Dekhtyar, Y. Immobilisation of yeast cells on the surface of hydroxyapatite ceramics. Process Biochem. 2011, 46, 665–670. [Google Scholar] [CrossRef]
  892. Ono, I.; Ohura, T.; Murata, M.; Yamaguchi, H.; Ohnuma, Y.; Kuboki, Y. A study on bone induction in hydroxyapatite combined with bone morphogenetic protein. Plast. Reconstr. Surg. 1992, 90, 870–879. [Google Scholar] [CrossRef] [PubMed]
  893. Ono, I.; Yamashita, T.; Jin, H.Y.; Ito, Y.; Hamada, H.; Akasaka, Y.; Nakasu, M.; Ogawa, T.; Jimbow, K. Combination of porous hydroxyapatite and cationic liposomes as a vector for BMP-2 gene therapy. Biomaterials 2004, 25, 4709–4718. [Google Scholar] [CrossRef] [PubMed]
  894. Sawyer, A.A.; Hennessy, K.M.; Bellis, S.L. The effect of adsorbed serum proteins, RGD and proteoglycan-binding peptides on the adhesion of mesenchymal stem cells to hydroxyapatite. Biomaterials 2007, 28, 383–392. [Google Scholar] [CrossRef] [PubMed]
  895. Mastrogiacomo, M.; Muraglia, A.; Komlev, V.; Peyrin, F.; Rustichelli, F.; Crovace, A.; Cancedda, R. Tissue engineering of bone: Search for a better scaffold. Orthod. Craniofac. Res. 2005, 8, 277–284. [Google Scholar] [CrossRef] [PubMed]
  896. Schek, R.M.; Taboas, J.M.; Hollister, S.J.; Krebsbach, P.H. Tissue engineering osteochondral implants for temporomandibular joint repair. Orthod. Craniofac. Res. 2005, 8, 313–319. [Google Scholar] [CrossRef] [PubMed]
  897. Nishikawa, M.; Myoui, A.; Ohgushi, H.; Ikeuchi, M.; Tamai, N.; Yoshikawa, H. Bone tissue engineering using novel interconnected porous hydroxyapatite ceramics combined with marrow mesenchymal cells: Quantitative and three-dimensional image analysis. Cell Transplant. 2004, 3, 367–376. [Google Scholar] [CrossRef]
  898. Quarto, R.; Mastrogiacomo, M.; Cancedda, R.; Kutepov, S.M.; Mukhachev, V; Lavroukov, A.; Kon, E.; Marcacci, M. Repair of large bone defects with the use of autologous bone marrow stromal cells. N. Engl. J. Med. 2001, 344, 385–386. [Google Scholar] [CrossRef] [PubMed]
  899. Vacanti, C.A.; Bonassar, L.J.; Vacanti, M.P.; Shufflebarger, J. Replacement of an avulsed phalanx with tissue-engineered bone. N. Engl. J. Med. 2001, 344, 1511–1514. [Google Scholar] [CrossRef] [PubMed]
  900. Morishita, T.; Honoki, K.; Ohgushi, H.; Kotobuki, N.; Matsushima, A.; Takakura, Y. Tissue engineering approach to the treatment of bone tumors: Three cases of cultured bone grafts derived from patients’ mesenchymal stem cells. Artif. Organs 2006, 30, 115–118. [Google Scholar] [CrossRef] [PubMed]
  901. Eniwumide, J.O.; Yuan, H.; Cartmell, S.H.; Meijer, G.J.; de Bruijn, J.D. Ectopic bone formation in bone marrow stem cell seeded calcium phosphate scaffolds as compared to autograft and (cell seeded) allograft. Eur. Cell Mater. 2007, 14, 30–39. [Google Scholar] [PubMed]
  902. Zuolin, J.; Hong, Q.; Jiali, T. Dental follicle cells combined with beta-tricalcium phosphate ceramic: A novel available therapeutic strategy to restore periodontal defects. Med. Hypotheses 2010, 75, 669–670. [Google Scholar] [CrossRef] [PubMed]
  903. Ge, S.; Zhao, N.; Wang, L.; Yu, M.; Liu, H.; Song, A.; Huang, J.; Wang, G.; Yang, P. Bone repair by periodontal ligament stem cell-seeded nanohydroxyapatite-chitosan scaffold. Int. J. Nanomed. 2012, 7, 5405–5414. [Google Scholar] [CrossRef]
  904. Hench, L.L.; Wilson, J. Surface-active biomaterials. Science 1984, 226, 630–636. [Google Scholar] [CrossRef] [PubMed]
  905. Navarro, M.; Michiardi, A.; Castano, O.; Planell, J.A. Biomaterials in orthopaedics. J. R. Soc. Interface 2008, 5, 1137–1158. [Google Scholar] [CrossRef] [PubMed]
  906. Anderson, J.M. The future of biomedical materials. J. Mater. Sci. Mater. Med. 2006, 17, 1025–1028. [Google Scholar] [CrossRef] [PubMed]
  907. Huebsch, N.; Mooney, D.J. Inspiration and application in the evolution of biomaterials. Nature 2009, 462, 426–432. [Google Scholar] [CrossRef] [PubMed]
  908. Chevalier, J.; Gremillard, L. Ceramics for medical applications: A picture for the next 20 years. J. Eur. Ceram. Soc. 2009, 29, 1245–1255. [Google Scholar] [CrossRef]
  909. Salgado, P.C.; Sathler, P.C.; Castro, H.C.; Alves, G.G.; de Oliveira, A.M.; de Oliveira, R.C.; Maia, M.D.C.; Rodrigues, C.R.; Coelh, P.G.; Fuly, A.; et al. Bone remodeling, biomaterials and technological applications: Revisiting basic concepts. J. Biomater. Nanobiotechnol. 2011, 2, 318–328. [Google Scholar] [CrossRef]
  910. Vallet-Regí, M. Evolution of bioceramics within the field of biomaterials. Comptes Rendus Chimie 2010, 13, 174–185. [Google Scholar] [CrossRef]
  911. Hartgerink, J.D.; Beniash, E.; Stupp, S.I. Self-assembly and mineralization of peptide-amphiphile nanofibers. Science 2001, 294, 1684–1688. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Dorozhkin, S.V. Calcium Orthophosphate-Based Bioceramics. Materials 2013, 6, 3840-3942. https://doi.org/10.3390/ma6093840

AMA Style

Dorozhkin SV. Calcium Orthophosphate-Based Bioceramics. Materials. 2013; 6(9):3840-3942. https://doi.org/10.3390/ma6093840

Chicago/Turabian Style

Dorozhkin, Sergey V. 2013. "Calcium Orthophosphate-Based Bioceramics" Materials 6, no. 9: 3840-3942. https://doi.org/10.3390/ma6093840

Article Metrics

Back to TopTop