In vitro effect of magnesium inclusion in sol gel … · In vitro effect of magnesium inclusion in...

5
In vitro effect of magnesium inclusion in solgel derived apatite Guochao Qi a,b , Sam Zhang a, , K.A. Khor a , Chunming Liu b , Xianting Zeng c , Wenjian Weng d , Min Qian c a School of Mechanical and Aerospace Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore b School of Materials and Metallurgy, Northeastern University, Shenyang 110004, PR China c Singapore Institute of Manufacturing Technology, 71 Nanyang Drive, Singapore 638075, Singapore d Department of Materials Science and Engineering, Zhejiang University, Hangzhou, Zhejiang, 310027, PR China Available online 12 July 2007 Abstract Magnesium-containing apatite coatings were prepared on Ti6Al4V substrates by solgel dip coating method. Standard simulated body fluid (SBF) was used to evaluate the bioactivity of the coatings. A series of the coatings according to the composition (Ca 10x Mg x )(PO 4 ) 6 (OH) 2 , where x = 0 to 2, is synthesized and immersed in the standard SBF for periods of 7 to 35 days for direct deposition of apatite layer from the SBF solution. Scanning electron microscopy (SEM) was used to examine the morphology changes of the SBF apatite layer that occurred during in vitro immersion. X-ray diffractometry, Fourier Transformation Infra-Red Spectroscopy and X-ray Photoelectron Spectroscopy were used to analyse the phases, chemical groups and composition of the solgel coating. Results show that as the solgel coating contains magnesium, this promotes deposition of apatite layer from SBF. As x 1, SBF immersion gives rise to a dense apatite layer. However, as x 1, selected dissolution of the deposited layer takes place, which results in serious pitting on the surface. Also, Mg ions from the dissolution of the solgel coating during immersion in the SBF apparently played a role in the subsequent deposition of apatite o the coating, evidence of Mg was found in the apatite layer. © 2007 Elsevier B.V. All rights reserved. Keywords: Magnesium; Hydroxyapatite; Titanium alloy; Simulated body fluid (SBF) 1. Introduction The application of apatite coatings on titanium alloy sub- strates has been widely studied for clinical applications due to their superb osteointegration properties[13]. Hydorxyapatite (Ca 10 (PO 4 ) 6 (OH) 2 , HA) and β-tricalcium phosphate (β-Ca 3 (PO 4 ) 2 , β-TCP) are most popular because they can form chemical bonds directly with bone tissues without the inter- vention of soft tissues[4,5]. However, coatings like pure HA can dissolve while β-TCP has high dissolution rate in body fluid which affects the long- term stability. Addition of fluorine has been found effective in retardation of the coating dissolution in SBF[6,7]. Substitution of cations like K + , Na + , Zn 2+ , Mn 2+ and Mg 2+ in the apatite structure to mimic natural bone has received much attention [811] to take advantage of their roles in bone metabolism, promoting catalytic reaction and controlling biological func- tions. Undoubtedly, magnesium is one of the most important bivalent ions associated with biological apatites [12]. Okazaki et al. [13] studied the crystallographic behavior of fluoridated hydroxyapatite (FHA) containing Mg 2+ and CO 3 2ions and found that both ions caused decrease in crystallinity of the FHA. Fanovich et al. [14] sintered hydroxyapatite with low concentrations of Mg 2+ at 1200 °C and established that addition of 1% magnesium improves the microstructure and the microhardness of the samples. Too much Mg addition causes MgO agglomeration, resulting in a diminution in microhard- ness. Kannan et al. [15] prepared magnesium whitlockite (β- TCMP) via aqueous precipitation and found that the β-TCMP powders had higher density and contracted lattice constants and unit cell volumes with respect to that of pure β-TCP powders. Suchanek et al. [16] synthesized magnesium-substituted hydroxyapatite powders by the mechanochemicalhydrother- mal method with Mg substitution levels as high as 28.4 wt%. The essential requirement for a bioactive material to bond to living bone is the formation of bonelike apatite on the material surface in the living body [4]. The in vivo apatite formation can be Available online at www.sciencedirect.com Thin Solid Films 516 (2008) 5176 5180 www.elsevier.com/locate/tsf Corresponding author. Tel.: +65 6790 4400; fax: +65 6791 1859. E-mail address: [email protected] (S. Zhang). 0040-6090/$ - see front matter © 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.tsf.2007.07.011

Transcript of In vitro effect of magnesium inclusion in sol gel … · In vitro effect of magnesium inclusion in...

Available online at www.sciencedirect.com

008) 5176–5180www.elsevier.com/locate/tsf

Thin Solid Films 516 (2

In vitro effect of magnesium inclusion in sol–gel derived apatite

Guochao Qi a,b, Sam Zhang a,⁎, K.A. Khor a, Chunming Liu b, Xianting Zeng c,Wenjian Weng d, Min Qian c

a School of Mechanical and Aerospace Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singaporeb School of Materials and Metallurgy, Northeastern University, Shenyang 110004, PR China

c Singapore Institute of Manufacturing Technology, 71 Nanyang Drive, Singapore 638075, Singapored Department of Materials Science and Engineering, Zhejiang University, Hangzhou, Zhejiang, 310027, PR China

Available online 12 July 2007

Abstract

Magnesium-containing apatite coatings were prepared on Ti6Al4V substrates by sol–gel dip coating method. Standard simulated body fluid(SBF) was used to evaluate the bioactivity of the coatings. A series of the coatings according to the composition (Ca10−xMgx)(PO4)6(OH)2, wherex=0 to 2, is synthesized and immersed in the standard SBF for periods of 7 to 35 days for direct deposition of apatite layer from the SBF solution.Scanning electron microscopy (SEM) was used to examine the morphology changes of the SBF apatite layer that occurred during in vitroimmersion. X-ray diffractometry, Fourier Transformation Infra-Red Spectroscopy and X-ray Photoelectron Spectroscopy were used to analyse thephases, chemical groups and composition of the sol–gel coating. Results show that as the sol–gel coating contains magnesium, this promotesdeposition of apatite layer from SBF. As x≤1, SBF immersion gives rise to a dense apatite layer. However, as x≫1, selected dissolution of thedeposited layer takes place, which results in serious pitting on the surface. Also, Mg ions from the dissolution of the sol–gel coating duringimmersion in the SBF apparently played a role in the subsequent deposition of apatite o the coating, evidence of Mg was found in the apatite layer.© 2007 Elsevier B.V. All rights reserved.

Keywords: Magnesium; Hydroxyapatite; Titanium alloy; Simulated body fluid (SBF)

1. Introduction

The application of apatite coatings on titanium alloy sub-strates has been widely studied for clinical applications due totheir superb osteointegration properties[1–3]. Hydorxyapatite(Ca10(PO4)6(OH)2, HA) and β-tricalcium phosphate (β-Ca3(PO4)2, β-TCP) are most popular because they can formchemical bonds directly with bone tissues without the inter-vention of soft tissues[4,5].

However, coatings like pure HA can dissolve while β-TCPhas high dissolution rate in body fluid which affects the long-term stability. Addition of fluorine has been found effective inretardation of the coating dissolution in SBF[6,7]. Substitutionof cations like K+, Na+, Zn2+, Mn2+ and Mg2+ in the apatitestructure to mimic natural bone has received much attention [8–11] to take advantage of their roles in bone metabolism,promoting catalytic reaction and controlling biological func-

⁎ Corresponding author. Tel.: +65 6790 4400; fax: +65 6791 1859.E-mail address: [email protected] (S. Zhang).

0040-6090/$ - see front matter © 2007 Elsevier B.V. All rights reserved.doi:10.1016/j.tsf.2007.07.011

tions. Undoubtedly, magnesium is one of the most importantbivalent ions associated with biological apatites [12]. Okazaki etal. [13] studied the crystallographic behavior of fluoridatedhydroxyapatite (FHA) containing Mg2+ and CO3

2− ions andfound that both ions caused decrease in crystallinity of the FHA.Fanovich et al. [14] sintered hydroxyapatite with lowconcentrations of Mg2+ at 1200 °C and established that additionof 1% magnesium improves the microstructure and themicrohardness of the samples. Too much Mg addition causesMgO agglomeration, resulting in a diminution in microhard-ness. Kannan et al. [15] prepared magnesium whitlockite (β-TCMP) via aqueous precipitation and found that the β-TCMPpowders had higher density and contracted lattice constants andunit cell volumes with respect to that of pure β-TCP powders.Suchanek et al. [16] synthesized magnesium-substitutedhydroxyapatite powders by the mechanochemical–hydrother-mal method with Mg substitution levels as high as 28.4 wt%.

The essential requirement for a bioactive material to bond toliving bone is the formation of bonelike apatite on the materialsurface in the living body [4]. The in vivo apatite formation can be

Table 1Reagents for preparing SBF (pH 7.40, 1 L)

Reagent Amount Order

NaCl 7.996 g 1NaHCO3 0.350 g 2KCl 0.224 g 3K2HPO4·3H2O 0.228 g 4MgCl2·6H2O 0.305 g 51 M–HCl 35–40 ml 6CaCl2 0.278 g 7Na2SO4 0.071 g 8NH2C(CH2OH)3 6.057 g 9

5177G. Qi et al. / Thin Solid Films 516 (2008) 5176–5180

reproduced in a simulated body fluid (SBF) [17,18]. The bio-activity of sol–gel derived magnesium apatite coatings onTi6Al4V substrate in simulated body fluid has not been sys-tematically assessed. Apatite coatings of different magnesiumcontents were soaked in SBF solution in this study to elucidate theprecipitation of the bone-like apatite on the coatings.

2. Experimental procedures

2.1. Coating preparation

Calcium nitrate tetrahydrate (Ca(NO3)2·4H2O, Sigma-Aldrich, AR), phosphorous pentoxide (P2O5, Merck, GR) andmagnesium nitrate hexhydrate (Mg(NO3)2·6H2O, Sigma-Aldrich, AR) were selected to prepare Ca-precursor, P-precursorand Mg-precursor. A designed amount of Mg-precursor and Ca-precursor was mixed to form the Ca–Mg mixture. This mixturewas added drop-wise into the P-precursor to obtain a solutionwith(Ca, Mg)/P ratio of 1.67. The designed degree of substitution ofCa2+ by Mg2+ in the mixture was indicated by the x value in thegeneral formula of (Ca10−xMgx(PO4)6(OH)2), where x=0/2, 1/2,2/2, 3/2, 4/2. The subsequent coatings were labeled as M0, M1,M2, M3 and M4 respectively.

Titanium alloy (Ti6Al4V) substrate (20×30×1.2 mm3) wasdipped vertically into the sol and withdrawn at a speed of 3 cm/min at the first dip to allow formation of a thin but good bond

Fig. 1. SEM micrographs of the sol–gel derived pure H

layer. The sol-coated substrate was then immediately transferredto an oven at 150 °C to dry for 15 min followed by firing at600 °C for 15 min. To reach a coating thickness of about 2 μm,the dipping–drawing–drying–firing process was repeated 4times at a new drawing speed of 4.5 cm/min to allow thickercoating at each dipping.

2.2. In vitro test in standard SBF solutions

The standard SBF solution was prepared according toKokubo's protocol [19] by dissolving appropriate quantities ofthe relevant reagent-grade chemicals in deionized water: NaCl,NaHCO3, KCl, K2HPO4.3H2O, MgCl2·6H2O, HCl (1 M),CaCl2, Na2SO4 and NH2C(CH2OH)3. Reagents were added,one by one after each reagent was completely dissolved in thedeionized water according to the order and amount given inTable 1. After all the reagents were added and dissolved, thesolution was then raised to 37 °C and maintained at this tem-perature while titrating the solution to a pH of 7.4 with 1 M HClsolution. The inorganic ion concentrations in the standard SBFsolution are close to human blood plasma.

The coatings were placed in sterilized bottle containingsolution with a liquid/area ratio of 50 ml/cm2. Before soaking inthe solution, the samples were washed ultrasonically in acetonefor 10 min and then sterilized in ethanol. The soaking bottleswere kept for various periods of 7, 10, 15, 20, 25, 30, and35 days at 37±0.1 °C. After immersion, the samples were takenout, gently washed with deionized water and dried at roomtemperature before characterization.

2.3. Coating Characterization

The surface morphology of the coating was observed usingScanning ElectronMicroscopy (SEM, JEOL JSM-5600LV).Thephase characterization was conducted by X-ray diffraction anal-ysis (XRD, Philips, PW1830) using monochromatic CuKαradiation at a step size of 0.02°. Analysis of chemical groups wasconducted using Fourier Transformation Infra-Red Spectrosco-py (FTIR) at a Bio-Rad FTS-3000 spectrometer in the range of

A coating soaked in SBF for a) 10 days b) 35 days.

Fig. 2. Magnesium apatite coatings soaked in SBF for 10 days.

5178 G. Qi et al. / Thin Solid Films 516 (2008) 5176–5180

4000 – 400 cm−1 in transmission mode with a resolution of4 cm−1. The composition of the apatite precipitated from theSBF was determined by X-ray Photoelectron Spectroscopy(XPS, Kratos-Axis Ultra System) using monochromatic Al KαX-ray source (1486.7 eV).

3. Results and discussion

3.1. The morphologies of the apatite layer precipitated from theSBF

Fig. 1 shows the characteristics of HA surface after soakingin SBF for 10 and 35 days (Fig. 1a and b respectively). After10 days, a new apatite layer (the “secondary coating”) is formedon top of the HA coating surface. However the apatite layerdoes not covered the original HA surface completely. After35 days, the apatite layer densify (Fig. 1b). Morphologies ofmagnesium apatite coatings after 10 days in SBF are shown in

Table 2Molar concentration of Ca, P and Mg in the sol–gel derived coatings

M0 M1 M2 M3 M4

x in sol 0 0.5 1.00 1.5 2.0x in coating 0 0.55 1.01 3.93 4.98Ca/P 1.67 1.57 1.50 1.46 1.60(Ca+Mg)/P 1.67 1.66 1.67 2.41 3.20

Fig. 2. The amount of magnesium incorporation was found tohave an impact on the secondary coating formation. For samplesM1 and M2, the apatite layer is dense and compact, and withmany nuclear boundaries. On M3, small holes formed on thesurface and on M4, the holes are obvious and much larger thanthat on M3.

Compared with the pure HA primary coating (c.f., Fig. 1) orwith increasing level of Mg incorporation, increase in “nuclear”sites is observed. In atomic structure, hydroxyapatite builds

Fig. 3. Cross-section morphology of M2 in SBF after 35 days.

Fig. 4. The newly formed apatite on magnesium apatite coatings in SBF after 35 days soaking.

5179G. Qi et al. / Thin Solid Films 516 (2008) 5176–5180

up with a central Ca(OH)2 and three surrounding Ca3(PO4)2groups. When magnesium is incorporated, the central calciumatom is substitute by magnesium [11]. Since the ionic radius ofMg2+(0.69 Å) is considerably shorter than that of Ca2+(0.99 Å)[20], substitution of Mg2+ into Ca2+position distorts the HAstructure and the structure more resembles β-TCP. To denote theexistence of Mg, it is often written as β-TCMP (where M standsfor Mg), which is also a kind of whitlockite [20–22].

As x=0.5 (sample M1), both HA and β-TCMP phases co-exist in the primary coating. The secondary coating on M1exhibits characteristics of both M0 and M2. The secondarycoating is smooth, dense and compact. As x≫1, however, onthe surfaces of M3 as shown in Fig. 2, serious “pitting” indicatessevere dissolution. This becomes even more severe at x=2.0(sample M4). The pores are distributed uniformly over thewhole surface. The increase of dissolution rate on M3 and M4 isconsistent with the higher magnesium content (Table 2) (ascompared with that designed in the sols).

Prolonged soaking of the β-TCMP in SBF promotes thedeposition of apatite layer. The typical cross-section morphol-

Fig. 5. XRD spectra of the apatite coatings before and after soaked in SBF for35 days.

ogy of M2 in SBF for 35 days (M2) is shown in Fig. 3. It showsthat about 2 μm thick secondary coating is formed on the sol–gel derived primary apatite layer. The secondary coating isdense and uniform, no pores and voids are observed. Fig. 4shows the surface morphology of the secondary coating formedon M1 and M2 after 35 days in SBF. The nucleation of the newapatite on the former formed surface can be clearly seen. Thephenomenon of the growth of secondary coating indicates thatthe incorporation of magnesium in the apatite structure to formβ-TCMP has good biological response in SBF.

3.2. Phase, group and composition of the apatite coatings afterSBF soaking

Typical XRD patterns of the apatite coatings before and aftersoaked in SBF are shown in Fig. 5. There are three phasesobserved in the patterns: the substrate (open circles), the HA(solid circles) and the β-TCMP (solid star). Before SBFimmersion, the HA peaks on M0 and β-TCMP peaks on M2 areprominent. After 35 days immersion, the similar crystal apatitepeaks can be observed on the samples. These results indicate

Fig. 6. Typical FTIR profiles of the secondary apatite coatings in SBF for35 days.

Fig. 7. Typical XPS survey scan spectra of the SBF-derived apatite layer on M1(soaked in SBF for 35 days). (Inset: narrow scan spectra of magnesium).

5180 G. Qi et al. / Thin Solid Films 516 (2008) 5176–5180

that the newly formed “secondary coating” keeps the samecrystal phase as the “primary coating”.

Fig. 6 shows typical FTIR profile of apatite coatings after35 days soaking in SBF. The coatings have similar chemicalgroups: PO4

3−, CO32− and OH−, typical chemical groups of

carbonate apatite structure. The FTIR profile of M0 has a largeH2O peak that can not be observed on other patterns. Thedepletion of H2O in the magnesium apatite samples indicatesthat the magnesium in the primary apatite coatings can affect thegrowing process of new apatite in the SBF solution. It followsthat the secondary coating is denser (as is observed in Fig. 3).

The compositions of the apatite deposition after 35 dayssoaking in SBF are analyzed with XPS and the typical surveyscan spectra on M1 are shown in Fig. 7. Since XPS affects onlyabout 10 nm deep from the surface, while the secondary coatingis about 2 μm thick Fig. 3, therefore the XPS results give thecompositions of the secondary coatings: Ca, P, O and C togetherwith a small peak for Mg. (This does not happen in thesecondary coating on M0). There are three peaks in the surveyscan spectra: Mg 2p3/2 at about 50 eV, Mg 2 s at about 88 eVandthe Mg Auger peak at about 301 eV [23]. The narrow scananalysis of Mg 2p is shown as inset of Fig. 7. The peak at∼50.2 eV belonging to Mg 2p is the fingerprint for magnesiumin apatite structure. In other words, as the primary coating (i.e.,the sol–gel derived apatite) contains Mg, the secondary coating(i.e., the apatite deposition directly form SBF) will then containMg. Therefore, incorporation of magnesium in the sol–gelapatite coating promotes inclusion of magnesium in the SBFprecipitated apatite. Since the natural bone contains ∼0.6 wt.%magnesium [20], this finding is very encouraging.

4. Conclusion

The incorporation of magnesium in the sol–gel derivedapatite (Ca10−xMgx(PO4)6(OH)2) promotes deposition of apatite

layer directly from simulated body fluid (SBF). As x≤1,deposition from SBF gives rise to dense apatite layer. However,as x≫1, selected dissolution of the deposited layer takes place,which results in serious pitting on the surface.

The incorporation of magnesium in the sol–gel derivedapatite promotes inclusion of magnesium in the SBF precipi-tated apatite.

Acknowledgements

This work is supported by the Agency for ScienceTechnology and Research, Singapore (A⁎Star) through project032101 0005 and SIMTech-NTU collaboration project U03-S-389B.

References

[1] S.W.K. Kweh, K.A. Khor, P. Cheang, Biomaterials 23/3 (2002) 775.[2] A. Stoch, W. Jastrzebski, E. Długoń, W. Lejda, B. Trybalska, G.J. Stoch,

A. Adamczyk, J. Mol. Struct. 744–747 (2005) 633.[3] M. Metiko-Hukovi, E. Tkalec, A. Kwokal, J. Piljac, Surf. Coat. Technol.

165/1 (2003) 40.[4] T. Kokubo, H. Takadama, Biomaterials 27/15 (2006) 2907.[5] C. Ergun, T.J. Webster, R. Bizios, R.H. Doremus, J. Biomed. Mater. Res.

59/2 (2001) 305.[6] K. Cheng, W. Weng, H. Wang, S. Zhang, Biomaterials 26/32 (2005) 6288.[7] Y. Wang, S. Zhang, X. Zeng, K. Cheng, M. Qian, W. Weng, Mater. Sci.

Eng., C 27/2 (2007) 244.[8] A. Ito, H. Kawamura, S. Miyakawa, P. Layrolle, N. Kanzaki, G. Treboux,

K. Onuma, S. Tsutsumi, J. Biomed. Mater. Res. 60/2 (2001) 224.[9] S. Miao, W. Weng, K. Cheng, P. Du, G. Shen, G. Han, S. Zhang, Surf.

Coat. Technol. 198/1–3 (2005) 223.[10] F. Miyaji, Y. Kono, Y. Suyama, Mater. Res. Bull. 40/2 (2005) 209.[11] G. Izabela, M. Zygmunt, M. Bogustaw, J. Biomed. Mater. Res., Part A 75/4

(2005) 788.[12] S. Kim, J. Lee, Y. Kim, D.H. Riu, S. Jung, Y. Lee, S. Chung, Y. Kim,

Biomaterials 24/8 (2003) 1389.[13] M. Okazaki, Biomaterials 12/9 (1991) 831.[14] M.A. Fanovich, M.S. Castro, J.M.P. López, Ceram. Int. 25/6 (1999) 517.[15] S. Kannan, J.M. Ventura, J.M.F. Ferreira, Ceram. Int. 33/4 (2007) 637.[16] W.L. Suchanek, K. Byrappa, P. Shuk, R.E. Riman, V.F. Janas, K.S.

TenHuisen, Biomaterials 25/19 (2004) 4647.[17] F. Barrere, C.A.V. Blitterswijk, K.D. Groot, P. Layrolle, Biomaterials 23/9

(2002) 1921.[18] J. Park, D. Lee, K. Oh, Y. Lee, K. Kim, K. Kim, Mater. Lett. 60/21–22

(2006) 2573.[19] A. Oyane, K. Nakanishi, H.-M. Kim, F. Miyaji, T. Kokubo, N. Soga, T.

Nakamura, Biomaterials 20/1 (1999) 79.[20] I. Mayer, R. Schlam, J.D.B. Featherstone, J. Inorg. Biochem. 66/1 (1997) 1.[21] A. Bigi, G. Falini, E. Foresti, A. Ripamonti, M. Gazzano, N. Roveri,

J. Inorg. Biochem. 49/1 (1993) 69.[22] R. Lagier, C.A. Baud, Pathol. Res. Pract. 199/5 (2003) 329.[23] F. John Moulder, Handbook of X-ray Photoelectron Spectroscopy,

Physical Electronics, Eden Prairie, Minn., 1995, p. 257.