Acessibilidade / Reportar erro

Electrochemical behavior of three CP titanium dental implants in artificial saliva

Abstract

The electrochemical behavior of three titanium dental implants purchased on the Brazilian market was evaluated in artificial saliva using open-circuit potential measurements, electrochemical impedance spectroscopy and potentiodynamic polarization. The three implants are passive in this medium but their corrosion resistance and the stability of their passive oxide films are quite different. Although the three implants were made of commercially pure grade titanium, their corrosion current densities and passive current densities were shown to differ from one to two orders of magnitude.

dental implants; titanium; corrosion; electrochemical techniques


REGULAR ARTICLES

Electrochemical behavior of three CP titanium dental implants in artificial saliva

Norma Terumi Kadowaki; Gustavo Aristides Santana Martinez* * e-mail: martinez@demar.eel.usp.br ; Alain Robin

Department of Materials Engineering - LOM - EEL-USP Pólo Urbo Industrial, Gleba AI-06, P.O. BOX 116, 12600-970 Lorena - SP, Brazil

ABSTRACT

The electrochemical behavior of three titanium dental implants purchased on the Brazilian market was evaluated in artificial saliva using open-circuit potential measurements, electrochemical impedance spectroscopy and potentiodynamic polarization. The three implants are passive in this medium but their corrosion resistance and the stability of their passive oxide films are quite different. Although the three implants were made of commercially pure grade titanium, their corrosion current densities and passive current densities were shown to differ from one to two orders of magnitude.

Keywords: dental implants, titanium, corrosion, electrochemical techniques

1. Introduction

Titanium has been widely used for dental implants and restorations due to its good osseointegration, excellent corrosion resistance and biocompatibility in biological fluids and high resistance/weight ratio1. The influence of contaminants and surface treatments of titanium implants (alumina-blasting, acid etching, anodization, hydroxyapatite coating, for example) on osseous integration has been extensively studied. Many works have been also published on the corrosion behavior of commercially pure titanium in artificial saliva, Ringer, Hanks solutions and others2-6. All the authors agree that the titanium corrosion resistance in these media is due to the formation of an adherent and highly protective oxide film on its surface which is mainly formed of TiO2[7]. Nevertheless, this oxide can lose its chemical stability under certain conditions. In fluoride medium for example, titanium can be severely attacked depending on the pH due to the formation of stable titanium fluoride and oxy-fluoride compounds8. In spite of the abundant bibliography on titanium corrosion properties, no study was found about the corrosion resistance of commercial titanium implants. Although titanium implants are considered as highly corrosion resistant materials for applications in body fluids, the determination of their corrosion rate is of great importance. Indeed, some studies indicated that titanium release in the organism, even at low levels, can lead from local discoloration of the tissues to serious inflammatory reactions9-10.

The aim of the present work was to study the corrosion behavior of three titanium dental implants purchased on the Brazilian market in artificial saliva, using open-circuit potential measurements, electrochemical impedance spectroscopy and potentiodynamic polarization.

2. Experimental

The dental implants (namely A, B and C) investigated in the present work were obtained on the Brazilian market from three different suppliers and are made of grade 2, 3 and 4 CP-titanium, respectively. Their chemical composition according to ASTM F67-95 standard11 is reported in Table 1.

Samples were cut from the implants for microstructure analysis, mounted in resin, grounded until 2400 grit, polished using an aqueous suspension of silica powder and finally etched in a modified Kroll solution (3 mL HF + 8 mL HNO3 + 100 mL H20). The three implants present a similar coarse-grained structure but differ in the mean Ti grain size (Figure 1). The Ti grain sizes of A, B and C implants are in the 50-100, 20-40 and 10-30 µm ranges, respectively.


The remaining materials were machined to nearly 3 mm-diameter cylinders. A copper wire was welded on one side of each cylinder in order to assure the electric contact. Then each cylinder was embedded in resin. The cross-section of the electrode (nearly 7 mm2 exposed area) was mechanically ground with emery-paper up to 600 grit, rinsed with distilled water and dried.

The electrolytic solution used for the electrochemical study was the Fusayama artificial12 saliva with the following composition: 0.4 g NaCl + 0.4 g KCl + 0.795 g CaCl2-2H20 + 0.690g NaH2PO4-H2O + 0.005 g Na2S-9H2O + 1.0 g urea in 1,000 mL deionized water. The Fusayama saliva has a natural pH of 5.5. The saliva was naturally aerated and the experiments were conducted without stirring. The temperature was maintained at 37 ± 1 ºC using a thermostatic bath. The counter electrode was a square-shaped platinum sheet of 18 cm2 area. All potentials were referred to the saturated calomel electrode (SCE) potential (= +0.242V/SHE).

Open-circuit potential, electrochemical impedance spectroscopy (EIS) and polarization measurements were performed using the Electrochemical Interface SOLARTRON® mod. 1287A and the Frequency Response Analyzer SOLARTRON® mod. 1260 A, controlled by the Ecorr/Zplot SOLARTRON® mod. 125587S software.

Prior to polarization experiments, the working electrodes were immersed in the Fusayama saliva for 3 hours, taking the moment of immersion as zero time. Then, impedance measurements at open-circuit potential were made using a sinusoidal signal of 10 mV amplitude and frequencies in the 0.1 Hz-100 kHz range. Cathodic and anodic polarization was then carried out potentiodynamically with 0.16 mV/s sweep rate. After each run, the samples were reground with emery papers to a 600-grit finish in order to remove any product formed on the metal surface which could affect the following tests, rinsed with distilled water and dried. Each experiment was repeated three times.

3. Results and Discussion

Figure 2 presents the variation of the open-circuit potential (OCP) of the three Ti implants as a function of exposure time in artificial saliva. All OCPs shift in the noble direction with time which depicts the growth of a passive oxide film on the three implants in this medium. The potential of both A and B implants presents some oscillations, which can be related to some instability of their oxide films. These oscillations are not observed for C implant.


The open-circuit potentials tend to stabilize with time. Their values after 3h exposure, -0.235, -0.046 and +0.075 V/SCE for A, B and C implants, respectively (or 0.007, 0.196 and 0.317 V/SHE), show that C implant presents the noblest character, following by B implant and finally A implant. Nevertheless, all these potentials are found in the stability region of TiO2 oxide in the Ti-H2O Pourbaix diagram for solutions of pH 5.5[13]. A comparison of the implants corrosion potentials in artificial saliva with published data is a difficult task as the OCP of CP-Ti is highly dependent on heat- and mechanical treatments, surface preparation of the samples and exposure time. For example Schiff14 and Al-Mayouf15 reported OCPs of +0.025 V/SCE and -0.331 V/SCE for CP-Ti in Fusayama artificial saliva, respectively.

Figure 3 shows the polarization curves of the three Ti implants in artificial saliva. The anodic polarization curves have the same shape and are typical of passive behavior. Nevertheless, great differences in the values of anodic current density are evident and this suggests that the passive films formed on the three implants do not have the same stability and that these implants present difference in corrosion resistance.


The corrosion current density obtained by Tafel extrapolation method and the passive current density measured at 1 V/SCE are reported in Table 2. The results show that the increasing order of corrosion resistance of the Ti implants is: A < B < C.

There is difference of two orders of magnitude between corrosion current densities of A and C implants. Schiff14 and Al-Mayouf15 also reported very different corrosion current density for CP-Ti in Fusayama artificial saliva, 10-6 and 3.10-8 A cm-2, respectively.

Considering the corrosion current densities reported in Table 2 and using the Faraday's law for a four-electron transfer, the calculated corrosion rates of A, B and C implants are 43.0, 5.0 and 0.5 µm/year, respectively. These values depict the excellent corrosion resistance of the three implants, since it is usually admitted that a metal or alloy presents an excellent corrosion resistance when its corrosion rate is in the 25-100 µm/year range16. Nevertheless, metallic implants generally do not fail due to corrosion damage but due to tissue irritation or infection caused by the corrosion products. Qualitative estimates suggest that implants must corrode at rates of 0.25 µm/year or less to avoid the possibility of tissue reaction16. Only C implant corrosion rate is close to this limiting value.

EIS experiments were carried out at OCP in artificial saliva. The Bode diagrams presented in Figures 4 a and b show that the metal/solution interface for the three Ti implants has a predominantly capacitive behavior, which is characteristic of passive materials. Nevertheless the capacitive behavior is limited to a narrower frequency range for A and B implants, 101-102 and 10-1-102 Hz, respectively. The C Ti implant presents a capacitive behavior in a larger frequency range,10-1-103 Hz. This is related to the higher stability of the passive film formed on C implant, which has already been observed from its polarization curve (Figure 3) that shows the lowest values of the anodic current density. The impedance values measured at low frequencies (Figure 4a) indicate that the increasing order of corrosion resistance of the Ti implants in artificial saliva is: A < B < C.


The difference in chemical composition of the three Ti implants (grade 2, 3 and 4 for A, B and C, respectively) cannot explain the superior resistance of C implant and the lowest resistance of A implant. Indeed, the general corrosion resistance of commercially pure Ti grades is quite similar and our results showed differences of one to two orders of magnitude for corrosion current densities of the implants. ASTM grade 2 is only more resistant than ASTM grades 3 and 4 to stress corrosion cracking in chloride media. The difference in Ti grain size (Figure 1) cannot also explain the difference in corrosion resistance of the three Ti implants. Indeed, for most metals and alloys a decrease in corrosion resistance is usually observed when the grain size diminishes17-18. From the grain size standpoint, the increasing order of corrosion resistance should be C < B < A. An improvement in corrosion resistance of CP-Ti grade 2 in simulated body fluid was already observed but for ultra fine grained microstructure (sub-micron grain size)19. Then, other reasons such as heat- and mechanical treatments during manufacture of the implants may be raised to explain the significant difference in the corrosion resistance of the three implants.

4. Conclusions

The corrosion behavior of three commercially pure Ti dental implants purchased on the Brazilian market from three different suppliers was evaluated in artificial saliva using electrochemical techniques.

The three implants present a passive behavior in this medium and high corrosion resistance. Nevertheless their corrosion resistance and the stability of their passive oxide films are quite different.

The corrosion current densities of the three implants were shown to differ from one to two orders of magnitude.

Nor the chemical composition neither the Ti grain size of the implants could explain this difference in corrosion resistance.

Received: May 15, 2009

Revised: August 18, 2009

  • 1. Jones FH. Teeth and bones: applications of surface science to dental materials and related biomaterials. Surface Science Reports. 2001; 42(3-5):75-205.
  • 2. Pan J, Thierry D and Leygraf C. Electrochemical impedance spectroscopy study of the passive oxide film on titanium for implant application. Electrochimica Acta. 1996; 41(7-8):1143-1153.
  • 3. González JEG and Mirza-Rosca JC. Study of the corrosion behavior of titanium and some of its alloys for biomedical and dental implant applications. Journal of Electroanalytical Chemistry. 1999; 471(2):109-115.
  • 4. Koike M and Fujii H. In vitro assessment of corrosive properties of titanium as a biomaterial. Journal of Oral Rehabilitation. 2001; 28(6):540-548.
  • 5. Ibris N and Mirza-Rosca JC. EIS study of Ti and its alloys in biological media. Journal of Electroanalytical Chemistry. 2002; 526(1-2):53-62.
  • 6. Shukla AK, Balasubramaniam R and Bhargava S. Properties of passive film formed on CP titanium, Ti-6Al-4V and Ti-13.4Al-29Nb alloys in simulated human body conditions. Intermetallics. 2005; 13(6):631-637.
  • 7. American Society for Metals Handbook. Corrosion. Materials Park: ASM International; 1993.
  • 8. Boere G. Influence of fluoride on titanium in an acidic environment measured by polarization resistance technique. Journal of Applied Biomaterials. 1995; 6(4):283-288.
  • 9. Black J, Sherk H, Bonini J, Rostoker WR and Galante JO. Metallosis associated with a stable titanium-alloy femoral component in total hip-replacement. A case report. Journal of Bone and Joint Surgery 1990; 72(1):126-130.
  • 10. Amstutz HC, Campbell P, Kossovsky N and Clarke IC. Mechanism and clinical significance of wear debris-induced osteolysis. Clinical Orthopaedics e Related Research 1992; 276:7-18.
  • 11
    American Society for Testing and Materials. Medical Devices and Services Standard Specification for Unalloyed Titanium for Sirurgical Implant Applications. West Conshohocken, PA; 1997. p. 1-3. [Annual Book, v.13]
  • 12. Baboian R. Corrosion Tests and Standards: Application and Interpretation. USA: ASTM; 1995.
  • 13. Pourbaix M. Atlas of Electrochemical Equilibria in Aqueous Solutions New York: Pergamon Press; 1966.
  • 14. Schiff N, Grosgogeat B, Lissac M and Dalard F. Influence of fluoride content and pH on the corrosion resistance of titanium and its alloys. Biomaterials. 2002; 23(9):1995-2002.
  • 15. Al-Mayouf AM, Al-Swayih AA, Al-Mobarak NA and Al-Jabab AS. Corrosion behavior of a new titanium alloy for dental implant applications in fluoride media. Materials Chemistry and Physics. 2004; 86(2-3):320-329.
  • 16. Fontana MG. Corrosion Engineering. New York: McGraw-Hill Book Company; 1986.
  • 17. Di Schino A and Kenny JM. Effects of the grain size on the corrosion behavior of refined AISI 304 austenitic stainless steels . Journal of Materials Science Letters. 2002; 21(20):1631-1634.
  • 18. Osório WR, Siqueira CA, Freire CMA and Garcia A. Efeito das diferentes morfologias estruturais e tamanho de grão na resistência à corrosão em amostras de alumínio e zinco puros. In: 6o Conferência sobre Tecnologia de Equipamentos; Salvador-Brazil; 2002.
  • 19. Balakrishnan A, Lee BC, Kim TN and Panigrahi BB. Corrosion behavior of ultra fine grained titanium in simulated body fluid for implant application. Trends in Biomaterials and Artificial Organs. 2008; 22(1):54-60.
  • *
    e-mail:
  • Publication Dates

    • Publication in this collection
      06 Nov 2009
    • Date of issue
      Sept 2009

    History

    • Accepted
      18 Aug 2009
    • Received
      15 May 2009
    ABM, ABC, ABPol UFSCar - Dep. de Engenharia de Materiais, Rod. Washington Luiz, km 235, 13565-905 - São Carlos - SP- Brasil. Tel (55 16) 3351-9487 - São Carlos - SP - Brazil
    E-mail: pessan@ufscar.br