Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Author’s Accepted Manuscript

Structure and mechanical properties of


hydroxyapatite coatings produced on titanium using
plasma spraying with induction preheating

Aleksandr Fomin, Marina Fomina, Vladimir


Koshuro, Igor Rodionov, Andrey Zakharevich,
Aleksandr Skaptsov
www.elsevier.com/locate/ceri

PII: S0272-8842(17)30981-1
DOI: http://dx.doi.org/10.1016/j.ceramint.2017.05.168
Reference: CERI15328
To appear in: Ceramics International
Received date: 9 May 2017
Revised date: 22 May 2017
Accepted date: 24 May 2017
Cite this article as: Aleksandr Fomin, Marina Fomina, Vladimir Koshuro, Igor
Rodionov, Andrey Zakharevich and Aleksandr Skaptsov, Structure and
mechanical properties of hydroxyapatite coatings produced on titanium using
plasma spraying with induction preheating, Ceramics International,
http://dx.doi.org/10.1016/j.ceramint.2017.05.168
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Structure and mechanical properties of hydroxyapatite coatings produced on

titanium using plasma spraying with induction preheating

Aleksandr Fomina, Marina Fominaa, Vladimir Koshuroa, Igor Rodionova, Andrey

Zakharevichb, Aleksandr Skaptsovb


a
Yuri Gagarin State Technical University of Saratov, Institute of Electronic Technology

and Engineering, 77 Politechnicheskaya str., Saratov, Russia, 410054

b
Saratov State University, Institute of Nanostructures and Biosystems, 83 Astrakhanskaya

str., Saratov, Russia, 410012

Abstract

Coatings of hydroxyapatite (HAp) were prepared by plasma spraying with induction

preheating of titanium substrate from 200 to 1000 ºC. The combination of conventional

plasma spraying and induction preheating ensured high mechnical properties of HAp

coatings. The coatings produced in the temperature range 400–600 ºC were characterized

by homogeneous nanostructure of splats with an average grain size of 12–31 nm.

1
According to the results of nanoindentation HAp coatings with high hardness 0.9–

1.2 GPa and elastic modulus 7–16 GPa were formed on the titanium.

Keywords: Hydroxyapatite coatings, Induction preheating, Plasma spraying, Titanium.

2
1. Introduction

Modern intraosseous implants and endoprostheses are produced from different

metals and alloys varying in their properties [1,2]. These devices are characterized by a

biocompatible layer (a film or coating) having the following distinctive features:

1). Necessary structural parameters, in particular porosity [3], roughness [2,4],

presence of micro- and nano-sized grains and their agglomerates [6-10].

2). Biocompatible chemical and phase composition determining the presence of

bioinert [9,11] or bioactive [12-17] interaction with the bone tissue, sufficient corrosion

resistance [18], as well as bactericidal [19,20] and other properties of the coating

[8,10,21,22].

3). High mechanical properties, particularly adhesive strength [18,23,24],

hardness H and elastic modulus E [14,25,26], fracture toughness, resistance to plastic

deformation [27], and wear resistance [28].

This combination of properties required for the implant surface and design in

general is practically impossible to provide by a single method of production of a

biocompatible interface layer. Conventionally, to obtain the required surface

characteristics various chemical [29-32], mechanical [3], physical [33-35], and vacuum-

condensation methods [36] are applied. For the surface treatment of intraosseous implants

complex methods are widely used. They include preparatory mechanical microtexturing

and/or etching [3,4], functional layer deposition [14,17,21,32,35,36], and in some cases

subsequent thermal [3,37] or chemical treatment, particularly biomineralization with the

deposition of bioactive ceramics [9,38]. The latter treatment can also be applied as a test

3
method for the evaluation of the potential for osteoinduction [16]. On the other hand, a

number of methods for the production of coatings can meet the criteria of bioactivity in

terms of biochemistry, morphological heterogeneity of micro- and

nanostructures [7,39,40] but their mechanical properties are not sufficient for the implants

exposed to extremely high loads of hard bone pieces, i.e. crystals of hydroxyapatite

(HAp), when installed into the bone bed. Under these conditions elements of the porous

coating are subjected to friction, which leads to the inevitable wear and, in some cases,

they are completely separated from the metal substrate of the implant [1]. Some

approaches enumerated below can prevent this process:

1). Lowering of the parameters of surface morphology and porosity, which is

especially well manifested when PVD and CVD methods are used [36]. These methods

are widely used in microelectronics, however, their appication is appropriate in the

nanostructuring of the surfaces combined with other methods of microtexturing;

2). Strengthening of porous materials (blocks and coatings) with special fillers of

submicron and nano-sized bioceramic powders of various oxides (Al2O3, ZrO2, TiO2)

[6,41,42], carbides [43], nitrides (TiN) [23], multi-walled carbon nanotubes [44], and

whiskers [11];

3). Sublayer deposition enhancing the adhesion of the main functional layer

[23,24,36];

4). Production of high strength coatings of matrix type for the subsequent

introduction of the bioactive filler [32,45];

5). Application of subsequent heat treatment, which also enhances the crystallinity

of HAp coatings [35,46];

4
6). Thermal activation during production of the coating, the use of which is

limited due to a number of technological peculiarities [47].

In most cases, additional treatment complicates the technology of production of

the coatings, increases labor costs and the overall cost of biomedical products. High-

efficient methods of coating production are gas-thermal methods, such as plasma

spraying [48,49] and high velocity oxygen fuel (HVOF) spraying [50]. The main

disadvantage of gas-thermal methods is significant change of phase composition in the

deposited powder particles. It affects the biomechanical compatibility of the coating and

bone.

During the interaction between the molten particles of bioceramics and the metal

substrate the cooling rate reaches 105-107 K/s, and for some materials it is even higher

[13,51]. Under these conditions, splats solidify almost immediately before their

crystallisation. Numerous crystal nuclei are formed, however their subsequent growth is

limited. During plasma heating of HAp particles their thermal destruction occurs with the

formation of the following phases: oxyhydroxyapatite (OHAp), oxyapatite (OAp),

tricalcium phosphate (TCP), tetracalcium phosphate (TTCP), and amorphous calcium

phosphate (ACP) [1,52]. Resorption rate of the constituent phases of such bioceramic

coating in biological fluids is represented by the following increasing sequence: HAp <<

β-TCP < OHAp / OAp < α-TCP < CaO < TTCP << ACP. The proportion of these phases

should not exceed a certain amount, since their presence accelerates coating resorption

and might cause the loss of the implant [1].

Amorphization of the HAp coating caused by ultra-high cooling rate is one of the

reasons for the accelerated resorption of bioceramic coatings, reduction of mechanical

5
properties, and subsequent rejection of the whole implant [2,3]. In some works, e.g. [53],

the results of production of HAp coatings (75 μm thick) with high hardness (~300 HV)

have been described, however there is no explanation of not changing hardness in the

interface and beyond in the presented graphs of hardness-depth dependency (where depth

is within 0.7 mm). The application of preheating of the sample substrate to 200 ºC has

been also described in the studies [54] but there is no sufficient data redarding the

temperature influence on the improvement of mechanical properties, hardness in

particular. Therefore, this research describes the results of the study of structure, phase

and chemical composition, and mechanical properties of HAp coatings produced by

plasma spraying with preheating at the temperatures within 1000 ºC. The activation of

titanium samples comprises the application of high velocity and non-contact inductive

heating by generating high-frequency electromagnetic waves [55].

2. Materials and methods

2.1. Preparation of coatings

The samples were 2 mm thick plates made of commercially pure (CP) titanium

VT1-00 (analogue of CP Titanium Grade 2). The main preparatory operations comprised

abrasive sandblasting with an average corundum particle size of 200–400 μm, which

provided the necessary parameters of the surface microrelief and further removal of

various impurities in detergent solutions and in ethanol when titanium samples were put

into an ultrasonic bath. Plasma spraying of HAp coatings was performed at the spraying

distance of 90 mm, arc current of 540 A, and voltage of 30 V (plasma spraying

6
equipment "VRES 744.3227.001"). The technological parameters were presented in

Table 1. HAp was prepared by an aqueous precipitation method from the addition of

calcium nitrate Ca(NO3)2 and diammonium phosphate (NH4)2HPO4 in ammonia solution,

which was described in detail, e.g. in [56]. After that HAp block was crushed and sieved

into powder with an average size of 50±20 μm.

Induction preheating was performed immediately before plasma spraying,

according to the scheme shown in Fig. 1. The heating rate of the samples at the maximum

power of induction heating device ranged from 30 to 70 °C/s. The deposition duration

was 3–5 s per sample, which corresponds to the resulting coating thickness 30–40 μm.

The laboratory apparatus for induction preheating included a power supply unit, a

generator unit, and an LC-contour unit comprising an inductor and a quartz chamber for a

sample [57]. The heating temperature was measured with an infrared pyrometer "DT-

8828" [58].

2.2. Coating characterization

Phase composition of HAp coatings was determined by X-ray diffraction (XRD)

on "Gemini/Xcalibur" diffractometer using an X-ray tube with a copper anode (CuKα,

λ = 1,541874 Å, 2Θ = 25-90º). The crystalline structure of the coatings was analyzed

using the software processing of XRD patterns in order to remove the amorphous phase.

Test sample preparation for XRD comprised the obtaining of a conglomerate of sprayed

coating fragments. Delamination from the titanium substrate was performed by

scratching, and then the microparticles were bound into the globular conglomerate using

7
epoxy resin with an amorphous structure. Attachment of the obtained sample was

performed on glass fiber.

Morphology of the coatings was studied on micro- and nanoscale by scanning

electron microscopy (SEM) to identify structure formation patterns. Morphological

parameters of the porous crystalline structure were defined, including shape and

dimensions of particles, as well as pore sizes. Processing of SEM images was performed

using software for the analysis of geometric parameters of micro-objects "AGPM-6M"

and program "Metallograph" (Fig. 2). As a result of the morphological analysis of the

coating surface images the following parameters were determined: average linear

dimension of microrelief elements (grains/pores, protrusions/cavities), their dispersion in

size, and the number of microrelief elements in the view field. SEM and energy

dispersive X-ray fluorescent analysis (EDX) of the deposits were performed on "MIRA II

LMU" with "INCA PentaFETx3" detector. SEM and EDX of the coatings were

performed at 20 kV.

Elastic modulus E and hardness H of HAp coatings were evaluated by

nanoindentation using mechanical properties tester "NANOVEA Ergonomic

Workstation". The selected load applied to Berkovich indenter equalled 100 mN and

provided its penetration into the coating within the thickness of 1–3 splats (about 1.5–

2.5 μm).

3. Results and discussion

3.1. Phase and chemical composition

8
Phase composition of HAp coatings significantly depended on the preheating

temperature (Fig. 3). The initial composition of the sprayed bioceramic powder

corresponded to HAp [59]. Analysis of diffractograms (without correction of amorpous

phase) of the bioceramic coatings showed that the amount of initial component HAp

remained sufficient at the temperatures of 400–800 ºC. Its proportion ranged from 40 to

60 % and the maximum of the crystalline phase equalling 70 % was observed at the

preheating temperature of about 600 °C. A slightly lower amount of the crystalline phase

of calcium phosphate compounds (HAp with 10 % addition of TCP and 3–7 % of CaO

impurity) was produced at the temperature of 600–1000 °C. The samples of coatings

produced at the temperatures of 20 °C (without preheating) and 200 °C were

characterized by lower crystallinity (about 10–20 %), HAp amount – 10–12 % with a

slight addition of TCP and CaO, not more than 5 % each. TCP, CaO, and ACP are the

phases of HAp decomposition, they also have increased resorption rate and lower

mechanical strength compared to the initial HAp [1].

Classification of splats in size and chemical composition was performed

considering the characteristic ratio of [Ca/P], which for the high chemical stability should

be in the range of [Ca/P] = 1.94–2.16 [wt.% / wt.%], corresponding to the phase

composition of HAp with a addition of TCP (Table 2, Fig. 4). Chemical composition

analysis showed that splats were characterized by almost complete retention of the initial

sprayed HAp powder.

Noticeable changes occurred with the preheating temperature of about 800 ºC,

when the amount of Ca was high. XRD data also showed an increased proportion of CaO.

According to SEM results at this preheating temperature individual small splats (less than

9
5–10 μm) almost completely disappeared in the coating, as they were alloyed into the

surface layer (Fig. 5).

Analysis of the influence of induction preheating followed by plasma spraying of

HAp coatings on their phase and chemical composition allows the development of

recommendations on the preheating temperature of titanium samples. When comparing

XRD and EDX data the following was established:

1. Much of the crystalline HAp transformed into the amorphous phase without

preheating and at low temperature of 200 °C. Preheating temperature from 400 to 800 °C

and to a minor extent at 1000 °C contributed to amorphization decrease and maintenance

of sufficient HAp phase.

2. When preheating temperature of titanium substrate increased to 600 °C the dust

fragments of particles almost disappeared. Their chemical composition was characterized

by [Ca/P] ratio of more than 2.00.

3.2. Surface morphology

Morphological heterogeneity MH of HAp coatings is a quantitative characteristic

of the size distribution of protrusions and pores on the surface of splats and their

structural components. The microstructure of HAp coatings had significant differences

depending on the preheating temperature (Fig. 6). Average diameter DS dependency for

microprotrusions of the coating on the temperature of titanium substrate was parabolic

(Table 3). Without preheating size DS took a minimal value. Maximum was observed at

the preheating temperature of 400–600 ºC. With further increase of preheating

10
temperature T to 800–1000 ºC size DS decreased. Abrasive blasting of titanium substrate

ensured high MH value (Fig. 6a).

Analysis of MH parameter was based on the density of splat protrusions in the

studied area, as MH varies according to the periodic law. Minimal value of MH

corresponded to the temperature of 400 and 800 ºC. The same values were observed

without preheating, which corresponded to the conventional plasma spraying (Fig. 6b).

Small spherical particles sprayed at this spraying distance interacted weakly with the

surface of titanium substrate, thus large splats formed the coating. MH grew from 2 to 4

times at 200, 600, and 1000 ºC. At the preheating temperature of 200 ºC the cohesion of

fine spherical particles of HAp powder increased (Fig. 6c). High fluidity of splats lead to

the smoothing of the surface microrelief at the preheating temperature of 400 ºC

(Fig. 6d). However, at the impact of liquid droplets the emerging surges influenced the

microgeometry of HAp coatings at the preheating temperature of 600 ºC. With further

increase of the substrate temperature to 800 ºC MH was reduced due to the lengthy stay

of sprayed particles of all sizes in viscoplastic or liquid state. Splat thickness decreased

from 5–10 μm to the minimum value of about 1–2 μm, which lead to a sharp increase in

the cooling rate up to the supercritical value, and as a result, to the noticeable crack

formation.

At the upper limit of the studied preheating temperature of 1000 ºC the formation

of a TiO2 film was observed, which was characterized by high morphological

heterogeneity. The presence of intermediate oxide sublayer reduced the cracking of

bioceramic HAp layer, enhanced the attachment of small spherical splats, and limited the

spreading of large splats.

11
The structural condition of the coating on the nanoscale was assessed by the

parameter of grain size D. The samples of coatings with morphology parameters close to

those of the bone trabeculae with mineral plates (crystal grains) of HAp were considered

the most appropriate. The resulting HAp coating without preheating was characterized by

considerable heterogeneity of nanostructure and formation of numerous dust particles and

their agglomerates (Fig. 7a). This structure had low mechanical characteristics, in

particular hardness, which normally lead to wear and coating delamination during the

installation into the bone. Similar morphology of HAp coating and bone tissue is required

for activation of the geometrical factor of bioactivity on the nanoscale.

When the substrate was preheated to 200 °C the structure of plasma coatings

varied slightly. The average grain size D increased to 26±9 nm, however, numerous

agglomerates of 50 to 300 nm were observed on the surface (Fig. 7b). With further

increase of the preheating temperature to 400 °C the structure of coatings became

practically homogeneous (Fig. 7c). Agglomerates were sporadic and consisted of a small

amount of nanoparticles. The average diameter reached the maximum of about 31±8 nm.

It should be noted that in the area of heating from 20 to 400 °C there was a linear increase

in the size of nanograins (Fig. 8; line D1).

Viscosity decrease caused by heating to 600 °C contributed to better spreading of

splats, so the reduction of the average grain size D to the initial value of 11±5 nm was

explicable and it was described by the parabolic dependency (Fig. 8; curve D2). The

structure was characterized by high homogeneity, the agglomerates were sporadic. Under

these conditions, the smallest nanocrystals with the size of 3–7 nm making about 50–

60 % were distinguished. Nano-grains formed agglomerates with the size of 14–40 nm

12
(Fig. 7d).

When the preheating temperature was between 800 and 1000 ºC separate

agglomerated nanoparticles of 30 to 90 nm were arranged on the "smooth" surface of the

HAp coating. There was a marked decrease in morphological heterogeneity and the

average size of nanograins equalled D = 6–12 nm.

Analysis of the structures on the nanoscale showed that the preheating of titanium

substrate ensured quite homogeneous nanocrystalline structure of the coatings. As a result

the necessary convergence between morphology types of HAp coatings and bone

trabeculae was established.

3.3. Mechanical properties

In order to produce HAp coatings with high mechanical properties the

comprehensive studies of the elastic modulus and hardness were performed. It should be

noted that the hardness of HAp coating exceeding that of the cortical bone provides the

improved biomechanical compatibility and it is expected to eliminate the danger of its

destruction during the installation of the implant into the bone bed. The suggested

induction preheating of titanium substrate before the deposition of the coating enhanced

mechanical properties, such as adhesion and cohesion. However, when the implant is

installed its local area of the surface is exposed to the concentrated force. Thus, high

hardness and elastic modulus of porous HAp coatings are necessary for their further

reliable performance in the tissues of the human body (Fig. 9).

13
Low preheating temperature of titanium substrate did not significantly increase

the hardness. Elastic modulus of the coatings at the temperature about 200 ºC increased

2–2.5 times. The highest values of the elastic modulus for cortical bone reached 25–

30 GPa with the hardness of about 0.5–0.8 GPa. If this preheating temperature is used

low performance of intraosseous implants may occur.

In the middle range of preheating temperature (400–600 ºC) the hardness of HAp

coatings reached the values typical for the solid fragments of cortical bone. The elastic

modulus remained around 8–12 GPa. Preheating to 800–1000 ºC ensured a 2–3-fold

margin of hardness at the maximum value of the elastic modulus of the HAp coating.

Conclusions

The surface of titanium VT1-00 after plasma spraying with induction preheating

of substrate was characterized by the formation of bioceramic coatings with the phase and

chemical composition close to HAp (which was confirmed by XRD and EDX), high

morphological heterogeneity of microstructure, homogeneous nanostructure (confirmed

by SEM), and high mechanical properties, hardness in particular. HAp coatings with the

required qualities were formed of micro-sized splats, the surface of which had nanograins

(D = 12–31 nm). The best morphological parameters of HAp coatings combined with

high hardness of about H = 0.9–1.2 GPa and elastic modulus of at least E = 7–16 GPa

were determined. These samples of coatings corresponded to the preheating temperature

T = 400–600 ºC.

Acknowledgments

14
The research was supported by the Ministry of Education and Science of the

Russian Federation in the framework of the Program of Scientific Research in

Universities (project No. 11.1943.2017/PP).

References

[1] S.R. Paital, N.B. Dahotre, Calcium phosphate coatings for bio-implant applications:

Materials, performance factors, and methodologies, Mater. Sci. Eng.: R. 66 (2009) 1–

70.

[2] L. Le Guéhennec, A. Soueidan, P. Layrolle, Y. Amouriq, Surface treatments of

titanium dental implants for rapid osseointegration, Dent. Mater. 23 (2007) 844–854.

[3] S.A. Shabalovskaya, D. Siegismund, E. Heurich, M. Rettenmayr, Evaluation of

wettability and surface energy of native Nitinol surfaces in relation to

hemocompatibility, Mater. Sci. Eng.: C 33 (2013) 127–132.

[4] G. Mendonça, D.B.S. Mendonça, F.J.L. Aragão, L.F. Cooper, Advancing dental

implant surface technology – From micron- to nanotopography, Biomaterials 29

(2008) 3822–3835.

[5] S.K. Swain, S. Bhattacharyya, Preparation of high strength macroporous

hydroxyapatite scaffold, Mater. Sci. Eng.: C 33 (2013) 67–71.

[6] L. Pawlowski, Finely grained nanometric and submicrometric coatings by thermal

spraying: A review, Surf. Coat. Technol. 202 (2008) 4318–4328.

[7] N. Wang, H. Li, W. Lü, J. Li, J. Wang, Z. Zhang, Y. Liu, Effects of TiO2 nanotubes

with different diameters on gene expression and osseointegration of implants in

15
minipigs, Biomaterials 32 (2011) 6900–6911.

[8] S.D. Puckett, E. Taylor, T. Raimondo, T.J. Webster, The relationship between the

nanostructure of titanium surfaces and bacterial attachment, Biomaterials 31 (2010)

706–713.

[9] X. Liu, X. Zhao, R.K.Y. Fu, J.P.Y. Ho, C. Ding, P.K. Chu, Plasma-treated

nanostructured TiO2 surface supporting biomimetic growth of apatite, Biomaterials

26 (2005) 6143–6150.

[10] P. Decuzzi, M. Ferrari, Modulating cellular adhesion through nanotopography,

Biomaterials 31 (2010) 173–179.

[11] W. Pon-On, N. Charoenphandhu, I.M. Tang, J. Teerapornpuntakit, J. Thongbunchoo,

N. Krishnamra, Biocomposite of hydroxyapatite-titania rods (HApTiR): Physical

properties and in vitro study, Mater. Sci. Eng.: C 33 (2013) 251–258.

[12] N. Ribeiro, S.R. Sousa, F.J. Monteiro, Influence of crystallite size of nanophased

hydroxyapatite on fibronectin and osteonectin adsorption and on MC3T3-E1

osteoblast adhesion and morphology, J. Colloid Interfaces Sci. 351 (2010) 398–406.

[13] S. Dyshlovenko, L. Pawlowski, B. Pateyron, I. Smurov, J.H. Harding, Modelling of

plasma particle interactions and coating growth for plasma spraying of

hydroxyapatite, Surf. Coat. Technol. 200 (2006) 3757–3769.

[14] M.F. Morks, A. Kobayashi, Influence of gas flow rate on the microstructure and

mechanical properties of hydroxyapatite coatings fabricated by gas tunnel type

plasma spraying, Surf. Coat. Technol. 201 (2006) 2560–2566.

[15] Y. Yan, J. Sun, Y. Han, D. Li, K. Cui, Microstructure and bioactivity of Ca, P and Sr

doped TiO2 coating formed on porous titanium by micro-arc oxidation, Surf. Coat.

16
Technol. 205 (2010) 1702–1713.

[16] S.V. Dorozhkin, Bioceramics of calcium orthophosphates, Biomaterials 31 (2010)

1465–1485.

[17] J.M. Gomez-Vega, E. Saiz, A.P. Tomsia, G.W. Marshall, S.J. Marshall, Bioactive

glass coatings with hydroxyapatite and bioglass particles on Ti-based implants. 1.

Processing, Biomaterials 21 (2000) 105–111.

[18] S. González, E. Pellicer, J. Fornell, A. Blanquer, L. Barrios, E. Ibáñez, P. Solsona, S.

Suriñach, M.D. Baró, C. Nogués, J. Sort, Improved mechanical performance and

delayed corrosion phenomena in biodegradable Mg–Zn–Ca alloys through Pd-

alloying, J. Mech. Behav. Biomed. Mater. 6 (2012) 53–62.

[19] J.P. Yuan, W. Li, C. Wang, Effect of the La alloying addition on the antibacterial

capability of 316L stainless steel, Mater. Sci. Eng.: C 33 (2013) 446–452.

[20] M. Gao, L. Sun, Z. Wang, Y. Zhao, Controlled synthesis of Ag nanoparticles with

different morphologies and their antibacterial properties, Mater. Sci. Eng.: C 33

(2013) 397–404.

[21] S. Qian, X. Liu, C. Ding, Effect of Si-incorporation on hydrophilicity and bioactivity

of titania film, Surf. Coat. Technol. 229 (2013) 156–161.

[22] J. Zhou, B. Li, S. Lu, L. Zhang, Y. Han, Regulation of osteoblast proliferation and

differentiation by interrod spacing of Sr-HA nanorods on microporous titania

coatings, ACS Appl. Mater. Interfaces 5 (2013) 5358–5365.

[23] S. Yang, H.C. Man, W. Xing, X. Zheng, Adhesion strength of plasma-sprayed

hydroxyapatite coatings on laser gas-nitrided pure titanium, Surf. Coat. Technol. 203

(2009) 3116–3122.

17
[24] G. Singh, S. Singh, S. Prakash, Surface characterization of plasma sprayed pure and

reinforced hydroxyapatite coating on Ti6Al4V alloy, Surf. Coat. Technol. 205 (2011)

4814–4820.

[25] S. Saber-Samandari, K.A. Gross, Nanoindentation reveals mechanical properties

within thermally sprayed hydroxyapatite coatings, Surf. Coat. Technol. 203 (2009)

1660–1664.

[26] B. Viswanath, R. Raghavan, U. Ramamurty, N. Ravishankar, Mechanical properties

and anisotropy in hydroxyapatite single crystals, Scr. Mater. 57 (2007) 361–364.

[27] A.D. Pogrebnyak, A.A. Drobyshevskaya, V.M. Beresnev, M.K. Kylyshkanov, T.V.

Kirik, S.N. Dub, F.F. Komarov, A.P. Shipilenko, Yu.Zh. Tuleushev, Micro- and

nanocomposite Ti-Al-N/Ni-Cr-B-Si-Fe-based protective coatings: Structure and

properties, Tech. Phys. 56 (2011) 1023–1030.

[28] L. Lapaj, J. Markuszewski, T. Rybak, A.A. Britsko, V.V. Anosov, M. Wierusz-

Kozlowska, Wear analysis of a ceramic on ceramic hip endoprosthesis, J. Frict. Wear

34 (2013) 32–37.

[29] S. Heinemann, C. Heinemann, S. Wenisch, V. Alt, H. Worch, T. Hanke, Calcium

phosphate phases integrated in silica/collagen nanocomposite xerogels enhance the

bioactivity and ultimately manipulate the osteoblast/osteoclast ratio in a human co-

culture model, Acta Biomater. 9 (2013) 4878–4888.

[30] J. Ni, K. Noh, C.J. Frandsen, S.D. Kong, G. He, T. Tang, S. Jin, Preparation of near

micrometer-sized TiO2 nanotube arrays by high voltage anodization, Mater. Sci.

Eng.: C 33 (2013) 259–264.

[31] H.-T. Chen, C.-H. Hsiao, H.-Y. Long, C.-J. Chung, C.-H. Tang, K.-C. Chen, J.L. He,

18
Micro-arc oxidation of β-titanium alloy: Structural characterization and osteoblast

compatibility, Surf. Coat. Technol. 204 (2009) 1126–1131.

[32] A. Kar, K.S. Raja, M. Misra, Electrodeposition of hydroxyapatite onto nanotubular

TiO2 for implant applications, Surf. Coat. Technol. 201 (2006) 3723–3731.

[33] S. Nag, S.R. Paital, P. Nandawana, K. Mahdak, Y.H. Ho, H.D. Vora, R. Banerjee,

N.B. Dahotre, Laser deposited biocompatible Ca–P coatings on Ti–6Al–4V:

Microstructural evolution and thermal modeling, Mater. Sci. Eng.: C 33 (2013) 165–

173.

[34] A. Siddharthan, T.S. Sampath Kumar, S.K. Seshadri, In situ composite coating of

titania–hydroxyapatite on commercially pure titanium by microwave processing,

Surf. Coat. Technol. 204 (2010) 1755–1763.

[35] S.C.G. Leeuwenburgh, J.G.C. Wolke, J. Schoonman, J.A. Jansen, Deposition of

calcium phosphate coatings with defined chemical properties using the electrostatic

spray deposition technique, J. Eur. Ceram. Soc. 26 (2006) 487–493.

[36] F. Maury, F.-D. Duminica, TiOxNy coatings grown by atmospheric pressure metal

organic chemical vapor deposition, Surf. Coat. Technol. 205 (2010) 1287–1293.

[36] P.K. Chu, Applications of plasma-based technology to microelectronics and

biomedical engineering, Surf. Coat. Technol. 203 (2009) 2793–2798.

[37] J. Cizek, K.A. Khor, Z. Prochazka, Influence of spraying conditions on thermal and

velocity properties of plasma sprayed hydroxyapatite, Mater. Sci. Eng.: C 27 (2007)

340–344.

[38] X. Liu, P.K. Chu, C. Ding, Surface modification of titanium, titanium alloys, and

related materials for biomedical applications, Mater. Sci. Eng.: R 47 (2004) 49–121.

19
[39] C. Kim, M.R. Kendall, M.A. Miller, C.L. Long, P.R. Larson, M.B. Humphrey, A.S.

Madden, A.C. Tas, Comparison of titanium soaked in 5 M NaOH or 5 M KOH

solutions, Mater. Sci. Eng.: C 33 (2013) 327–339.

[40] P.A. Ramires, A. Romito, F. Cosentino, E. Milella, The influence of

titania/hydroxyapatite composite coatings on in vitro osteoblasts behaviour,

Biomaterials 22 (2001) 1467–1474.

[41] C. Arnould, J. Denayer, M. Planckaert, J. Delhalle, Z. Mekhalif, Bilayers coating on

titanium surface: The impact on the hydroxyapatite initiation, J. Colloid Interface Sci.

341 (2010) 75–82.

[42] L. Xie, X. Ma, A. Ozturk, E.H. Jordan, N.P. Padture, B.M. Cetegen, D.T. Xiao, M.

Gell, Processing parameter effects on solution precursor plasma spray process spray

patterns, Surf. Coat. Technol. 183 (2004) 51–61.

[43] D. Caschera, F. Federici, L. Pandolfi, S. Kaciulis, M. Sebastiani, E. Bemporad, G.

Padeletti, Effect of composition on mechanical behaviour of diamond-like carbon

coatings modified with titanium, Thin Solid Films 519 (2011) 3061–3067.

[44] K. Balani, Y. Chen, S.P. Harimkar, N.B. Dahotre, A. Agarwal, Tribological behavior

of plasma-sprayed carbon nanotube-reinforced hydroxyapatite coating in

physiological solution, Acta Biomater. 3 (2007) 944–951.

[45] M.Y. Zhang, C. Ye, U.J. Erasquin, T. Huynh, C. Cai, G.J. Cheng, Laser engineered

multilayer coating of biphasic calcium phosphate/titanium nanocomposite on metal

substrates, ACS Appl. Mater. Interfaces 3 (2011) 339–350.

20
[46] F.-H. Lin, L. Chun-Jen, C. Ko-Shao, S. Jui-Sheng, Thermal reconstruction behavior

of the quenched hydroxyapatite powder during reheating in air, Mater. Sci. Eng.: C 13

(2000) 97–104.

[47] M. Bertagnolli, M. Marchese, G. Jacucci, Modelling of particles impacting on a rigid

substrate under plasma spraying conditions, J. Therm. Spray Technol. 4 (1995) 41–

49.

[48] R. Jaworski, C. Pierlot, L. Pawlowski, M. Bigan, M. Martel, Design of the synthesis

of fine HA powder for suspension plasma spraying, Surf. Coat. Technol. 203 (2009)

2092–2097.

[49] S. Kozerski, F.-L. Toma, L. Pawlowski, B. Leupolt, L. Latka, L.-M. Berger,

Suspension plasma sprayed TiO2 coatings using different injectors and their

photocatalytic properties, Surf. Coat. Technol. 205 (2010) 980–986.

[50] K.A. Khor, H. Li, P. Cheang, Processing–microstructure–property relations in

HVOF sprayed calcium phosphate based bioceramic coatings, Biomaterials 24 (2003)

2233–2243.

[51] A.Yu. Ivannikov, V.I. Kalita, D.I. Komlev, A.A. Radyuk, V.P. Bagmutov, I.N.

Zakharov, S.N. Parshev, The effect of electromechanical treatment on structure and

properties of plasma sprayed Ni–20Cr coating, J. Alloys Compd. 655 (2016) 11–20.

[52] S. Dyshlovenko, B. Pateyron, L. Pawlowski, D. Murano, Numerical simulation of

hydroxyapatite powder behavior in plasma jet, Surf. Coat. Technol. 179 (2004) 110–

117.

21
[53] R. Palanivelu, S. Kalainathan, A.R. Kumarn, Characterization studies on plasma

sprayed (AT/HA) bi-layered nano ceramics coating on biomedical commercially pure

titanium dental implant, Ceram. Int. 40 (2014) 7745–7751.

[54] B.R. Gligorijević, M. Vilotijević, M. Šćepanović, N.S. Vuković, N.A. Radović,

Substrate preheating and structural properties of power plasma sprayed

hydroxyapatite coatings, Ceram. Int. 42 (2016) 411–420.

[55] A. Fomin, M. Fomina, V. Koshuro, A. Voyko, I. Rodionov, Experimental data and

numerical modelling of heating of titanium implants using high-frequency currents, J.

Phys. Conf. Ser. 741 (2016) 012170.

[56] A. Destainville, E. Champion, D. Bernache-Assollant, E. Laborde, Synthesis,

characterization and thermal behavior of apatitic tricalcium phosphate, Mater. Chem.

Phys. 80 (2003) 269–277.

[57] A. Fomin, M. Fomina, A. Steinhauer, N. Petrova, E. Poshivalova, I. Rodionov,

Induction heat treatment device and technique of bioceramic coatings production on

titanium implants, in: RTUCON 2014. Proceedings of the 55th International

Scientific Conference on Power and Electrical Engineering of Riga Technical

University; 2014 Oct 14; Riga (Latvia). Riga: IEEE; 2014. p 111-115.

[58] A. Fomin, S. Dorozhkin, M. Fomina, V. Koshuro, I. Rodionov, A. Zakharevich, N.

Petrova, A. Skaptsov, Composition, structure and mechanical properties of the

titanium surface after induction heat treatment followed by modification with

hydroxyapatite nanoparticles, Ceram. Int. 42 (2016) 10838-10846.

[59] J.M. Hughes, M. Cameron, K.D. Crowley, Structural variations in natural F, OH,

and Cl apatites, Am. Mineral. 74 (1989) 870–876.

22
Figure captions

Fig. 1. Scheme of plasma spraying with induction preheating. Internal circuit: U, I –

voltage and current of plasma torch 1 from the power supply (PS); MPG and MCG – flow

rate of plasma and carrier gas (argon); MP, dP, TPPP – flow rate, particle size, and

thermophysical properties of the powder. Outer circuit: L – spraying distance; TS, TPPS,

CS – temperature, thermophysical properties, and structural features of substrate 5; 3 –

protective screen against the plasma jet 2 of the inductor 4 supplied by voltage Ui and

current Ii with frequency f of the induction heating device comprising the main power

supply and a generator unit (GU); F – function depending on the treated item geometry.

Fig. 2. Morphological analysis of the nanostructure of HAp coating: a – SEM of the

coating (mark 200 nm) and binarized fragment (image obtained by processing software);

b – distribution graph in linear dimensions of nanograins (x-axis – size, nm; y-axis –

percentage, %)

Fig. 3. XRD of HAp powder and coating samples produced at the spraying distance of 90

mm: ● – HAp; □ – TCP; ■ – CaO.

Fig. 4. SEM of the HAp coating produced at 200 °C (points "1–9" show the areas of EDX

analysis)

23
Fig. 5. Ca/P ratio for splats of different sizes: 1 – large splats; 2 – small splats with the

size of not more than 5–10 μm

Fig. 6. Surface morphology of titanium samples after treatment: a – abrasive blasting (a)

and plasma spraying (b–d) at different preheating temperatures T: b – T = 20 °C

(conventional plasma spraying); c – T = 200 °C; d – T = 400 °C.

Fig. 7. Nanostructure of HAp coatings produced at different preheating temperatures: a –

20 °C (conventional plasma spraying); b – 200 °C; c – 400 °C; d – 600 °C

Fig. 8. Dependency of the average HAp grain size D on the preheating temperature

Fig. 9. Dependency of elastic modulus (a) and hardness (b) on the preheating temperature

24
Table 1. Modes of plasma spraying with induction preheating

Parameter Value

Current frequency of inductor, kHz 90

Inductor current, kA – 4,4 7,6

Preheating temperature, °C 20* 200 400

Heating time, s – 2–3 3–4 600 800 1000

Spraying distance, mm 90 7–8 14–15 12–13

Arc current of plasma torch, A 540

Voltage, V 30

Size of powder, μm 50±20

Plasma gas flow rate (Ar), l/min 55–60

Carrier gas flow rate (Ar), l/min 5–7

* – conventional plasma spraying without preheating.

25
Table 2. EDX of the HAp coating (preheating temperature 200 °C) and calculated values

[Ca/P] for bioceramic materials (see Fig. 4)

Spectrum O P Ca Total [Ca/P],

[wt.% / wt.%]

1 45.65 18.73 35.62 100 1.91

2 37.86 21.81 40.33 100 1.85

3 48.05 18.93 33.01 100 1.74

4 54.55 14.01 31.43 100 2.24

5 45.98 18.80 35.23 100 1.87

6 38.32 21.29 40.38 100 1.90

7 32.52 17.83 49.65 100 2.78

8 58.84 11.31 29.86 100 2.64

9 55.07 10.00 34.93 100 3.49

HAp – – – – 2.16

TCP – – – – 1.94

TTCP – – – – 2.59

26
Table 3. Morphological parameters of microstructure of HAp coatings

Preheating Morphological Average splat size

temperature, °C heterogeneity D and confidence

MH, 1/mm2 interval ΔD, μm

20 5821 5.7±3.3

200 21826 8.8±4.5

400 6447 12.4±5.3

600 11245 12.9±5.1

800 4210 10.2±6.5

1000 18918 6.4±4.1

Ti substrate after 99251 4.9±2.8

sandblasting

27
1

28
5

29
7

30
8

31

You might also like