What is the role of lipopolysaccharide on the tribocorrosive behavior of titanium?

15
JOURNAL OF THE MECHANICAL BEHAVIOR OF BIOMEDICAL MATERIALS 8 (2012) 71–85 Available online at www.sciencedirect.com journal homepage: www.elsevier.com/locate/jmbbm Research paper What is the role of lipopolysaccharide on the tribocorrosive behavior of titanium? Mathew T. Mathew a,,1 , Valentim A. Barão b,c,1 , Judy Chia-Chun Yuan c , Wirley G. Assunção b , Cortino Sukotjo c , Markus A. Wimmer a a Department of Orthopedic Surgery, Section of Tribology, Rush University Medical Center, 1653 W. Congress Parkway, Chicago, IL, 60612, USA b Department of Dental Materials and Prosthodontics, Univ Estadual Paulista (UNESP), Araçatuba Dental School, Jose Bonifacio, 1193, Araçatuba, São Paulo, 16015-050, Brazil c Department of Restorative Dentistry, University of Illinois at Chicago – College of Dentistry, 801 S Paulina, Chicago, IL, 60612, USA ARTICLE INFO Article history: Received 21 September 2011 Accepted 11 November 2011 Published online 20 November 2011 Keywords: Titanium Lipopolysaccharide Electrochemistry Tribocorrosion Synergism ABSTRACT In an oral environment, titanium dental implants are exposed to a complex degradation process which is predominantly influenced by the intermittent mechanical events (masti- cation), continuous exposure to varying chemical solutions (saliva and food) and formation of microbiological (biofilm). Several studies have investigated the chemical corrosion and mechanical resistance of titanium; however, very few attempted to report on the effects of combined chemical, mechanical and microbiological interactions, which simulates the oral environment. A new multi-disciplinary research area, “tribocorrosion” (a combined study of wear and corrosion), was used to address such issues. The tribocorrosive nature of titanium in artificial saliva (pH 6.5) with lipopolysaccharide (LPS) was investigated. Twenty-four tita- nium discs (12 mm diameter, 7 mm thickness), were divided into 8 groups (n = 3) as a func- tion of material (commercially pure titanium (cpTi) and titanium–aluminum–vanadium (TiAlV) alloy) and LPS concentrations (0, 0.15, 15 and 150 µg/ml). Sliding duration (2000 cy- cles), frequency (1.2 Hz) and load (20 N) parameters mimicked the daily mastication process. Electrochemical impedance spectroscopy was conducted before and after tribocorrosion to comprehend the changes in corrosion kinetics. Worn surfaces were examined using white- light-interferometry and scanning electron microscopy. Total weight loss and roughness values were calculated. LPS affected the tribocorrosive behavior of both titanium types. LPS statistically accelerated the ion exchange between titanium and saliva, and reduced the resistance of the titanium surface against corrosion (p <.05). Sliding events decreased the protectiveness of the titanium surface. In general, TiAlV exhibited better corrosion behav- ior, but both titanium types showed similar in total weight loss (p >.05). LPS significantly increased the cpTi weight loss (p = .041), and the roughness of the surface (p <.001). In summary, LPS negatively affected the corrosion/wear behavior of titanium, which may con- tribute to the failure of dental implants. c 2011 Elsevier Ltd. All rights reserved. This study was performed at Rush University Medical Center, Chicago, USA. This study was the finalist of the Arthur R. Frechette 2011 New Investigator Award, International Association for Dental Research (IADR), San Diego, USA. This study also received a Silver Medal Award, International Conference in Metallurgical Coatings & Thin Films, Advanced Surface Engineering Division, 2011, San Diego, USA. Corresponding author. Tel.: +1 312 942 8310. E-mail address: [email protected] (M.T. Mathew). 1 These authors contributed equally to this work. 1751-6161/$ - see front matter c 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.jmbbm.2011.11.004

Transcript of What is the role of lipopolysaccharide on the tribocorrosive behavior of titanium?

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

Available online at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/jmbbm

Research paper

What is the role of lipopolysaccharide on the tribocorrosivebehavior of titanium?✩

Mathew T. Mathewa,∗,1, Valentim A. Barãob,c,1, Judy Chia-Chun Yuanc,Wirley G. Assunçãob, Cortino Sukotjoc, Markus A. WimmeraaDepartment of Orthopedic Surgery, Section of Tribology, Rush University Medical Center, 1653 W. Congress Parkway,Chicago, IL, 60612, USAbDepartment of Dental Materials and Prosthodontics, Univ Estadual Paulista (UNESP), Araçatuba Dental School, Jose Bonifacio, 1193,Araçatuba, São Paulo, 16015-050, BrazilcDepartment of Restorative Dentistry, University of Illinois at Chicago – College of Dentistry, 801 S Paulina, Chicago, IL, 60612, USA

A R T I C L E I N F O

Article history:

Received 21 September 2011

Accepted 11 November 2011

Published online 20 November 2011

Keywords:

Titanium

Lipopolysaccharide

Electrochemistry

Tribocorrosion

Synergism

A B S T R A C T

In an oral environment, titanium dental implants are exposed to a complex degradation

process which is predominantly influenced by the intermittent mechanical events (masti-

cation), continuous exposure to varying chemical solutions (saliva and food) and formation

of microbiological (biofilm). Several studies have investigated the chemical corrosion and

mechanical resistance of titanium; however, very few attempted to report on the effects of

combined chemical, mechanical and microbiological interactions, which simulates the oral

environment. A new multi-disciplinary research area, “tribocorrosion” (a combined study of

wear and corrosion), was used to address such issues. The tribocorrosive nature of titanium

in artificial saliva (pH 6.5) with lipopolysaccharide (LPS) was investigated. Twenty-four tita-

nium discs (12 mm diameter, 7 mm thickness), were divided into 8 groups (n = 3) as a func-

tion of material (commercially pure titanium (cpTi) and titanium–aluminum–vanadium

(TiAlV) alloy) and LPS concentrations (0, 0.15, 15 and 150 µg/ml). Sliding duration (2000 cy-

cles), frequency (1.2 Hz) and load (20 N) parametersmimicked the dailymastication process.

Electrochemical impedance spectroscopy was conducted before and after tribocorrosion to

comprehend the changes in corrosion kinetics. Worn surfaces were examined using white-

light-interferometry and scanning electron microscopy. Total weight loss and roughness

values were calculated. LPS affected the tribocorrosive behavior of both titanium types. LPS

statistically accelerated the ion exchange between titanium and saliva, and reduced the

resistance of the titanium surface against corrosion (p < .05). Sliding events decreased the

protectiveness of the titanium surface. In general, TiAlV exhibited better corrosion behav-

ior, but both titanium types showed similar in total weight loss (p > .05). LPS significantly

increased the cpTi weight loss (p = .041), and the roughness of the surface (p < .001). In

summary, LPS negatively affected the corrosion/wear behavior of titanium, which may con-

tribute to the failure of dental implants. c⃝ 2011 Elsevier Ltd. All rights reserved.

✩ This study was performed at Rush University Medical Center, Chicago, USA. This study was the finalist of the Arthur R. Frechette 2011

l

d

New Investigator Award, International Association for Dental ReseaAward, International Conference in Metallurgical Coatings & Thin Fi

∗ Corresponding author. Tel.: +1 312 942 8310.E-mail address: [email protected] (M.T. Mathew).

1 These authors contributed equally to this work.

1751-6161/$ - see front matter c⃝ 2011 Elsevier Ltd. All rights reservedoi:10.1016/j.jmbbm.2011.11.004

rch (IADR), San Diego, USA. This study also received a Silver Medalms, Advanced Surface Engineering Division, 2011, San Diego, USA.

.

72 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

1. Introduction

Dental implants are widely used and they are consideredas an ideal solution to substitute missing teeth (Heueret al., 2007). Commercially pure titanium (cpTi) is preferredin dentistry (Mabilleau et al., 2006), whereas titaniumalloys, such as titanium–aluminum–vanadium (TiAlV), arefavored in orthopedics because of their superior mechanicalstrength. In general, titanium (Ti) has favorable mechanicalproperties, excellent resistance to corrosion and acceptablebiocompatibility (Cortada et al., 2000; de Assis et al., 2006;Vieira et al., 2006). The spontaneous ability of Ti to forma stable and dense oxide layer (TiO2) is responsible for itscorrosion resistance and the reason for wide acceptance,particularly in biomedical application (Huang, 2003).

Although Ti materials exhibit superior corrosion resis-tance, they are not inert to corrosive attack under spec-ified conditions, such as pH variation and presence offluorides (Chaturvedi, 2009). Inferior wear resistance of Tiis a critical factor in such conditions (Barril et al., 2004).Further, removal of Ti oxide layer (totally or partially, thick-ness varies from 10 to 20 nm) under mechanical or chemi-cal incidents possibly activate corrosion mechanisms on theexposed Ti surface as observed with many other metal al-loys (Chaturvedi, 2009; Henry, 1999). When inserted in oralcavity, dental implants are exposed to several adverse me-chanical, chemical and microbiological events, leading intocomplex degradation processes (Nakagawa et al., 2002).During mastication, which is a unique mechanical ac-tion, implants are subjected to both axial and obliqueforces which create micromovements at several inter-faces, i.e. implant/prosthetic crown, implant/abutment, abut-ment/prosthetic crown and implant/bone interfaces (Vieiraet al., 2006). In such situations fretting-corrosion, a surfacedamage mode caused by low-amplitude oscillatory slidingmotion between two contacting surfaces (Dalmiglio et al.,2008), can induce wear and corrosion. Particles created fromwear (surface damage) of implants promote the release of in-flammatory mediators from various cells (Azzi and Szpunar,2007) which contribute to bone resorption (Nikolopoulou,2006), and the encapsulation of the implant by fibrous tis-sue (Jones, 2001) with subsequent failure of the implant(Nikolopoulou, 2006; Sinnett-Jones et al., 2005).

In addition, dental implants are exposed to an aggressiveenvironment including agents such as bacterial biofilm andsaliva (Correa et al., 2009). The thermal, ionic, microbiologic,and enzymatic properties of this environment are ideal toinduce the biodegradation of metals (Fathi et al., 2003).Lipopolysaccharide (LPS), a membrane constituent of gram-negative bacteria, is a highly bioactive molecule (Simon et al.,1969) and has marked effects on cells like macrophages,lymphocytes, fibroblasts, and osteoblasts (Wilson, 1995).LPS is presented in the gingival crevicular fluid (Casarinet al., 2010; Simon et al., 1969), mainly on the surfacesof periodontally diseased teeth (Wilson, 1995), contributingto inflammation, osteoclasis, inhibition of cell growth, anddelayed healing (Knoernschild et al., 1996). In the peri-implant region, LPS influences peri-implant inflammationand implant prognosis (Nelson et al., 1997).

Several studies have attempted to evaluate the chemicalcorrosion of Ti (Barao et al., 2011, in press; Huang, 2003;Joska and Fojt, 2010; Joska et al., 2010; Mabilleau et al.,2006; Mareci et al., 2009; Messer et al., 2010, 2009; Nakagawaet al., 1999, 2002; Popa et al., 2008; Schiff et al., 2002; Souzaet al., 2009); however, very few have reported about theeffect of combined chemical, mechanical and microbiologicalinteractions, which resemble the oral environment. In thisinvestigation, a multi-disciplinary research area, entitled“tribocorrosion”, is employed to address this issue in acomprehensive manner.

A simple definition of tribocorrosion research can bestated as “a degradation phenomenon that is subjected to thecombined action of tribological (wear and fretting) and corro-sive (chemical and/or electrochemical) events influenced bythe variation in mechanical contact conditions (load and rel-ative velocity) and in nature of the environment (pH, humid-ity and biochemistry)” (Mathew et al., 2009). The combina-tion of such processes can result in various forms of deteri-oration, including surface wear and fatigue. Moreover, whenthe tribocorrosion occurs in a biological environment, suchas in oral system, the presence of biological species createsfurther complexity in understanding the materials’ degrada-tion pathways and dominant driving mechanisms generatedby the interplay of wear and corrosion (Landolt et al., 2001).

Previous study (Souza et al., 2010b) investigated thetribocorrosion behavior of Ti in the presence of mixedoral biofilms. Although biofilms induced low friction onTi during tribocorrosion test, its corrosion resistance wasreduced probably due to the reduction of pH caused by thepresence of micro-organisms and their secreted products.The reduction of friction could affect the mechanical integrityand stability (elastic deformation, rupture and irreversibleshear) of internal connections of dental implants (Souza et al.,2010a).

However, to the authors’ best knowledge, the contributionof LPS on the corrosion/wear of Ti material has notbeen evaluated yet. In this paper, the role of LPSon the tribocorrosive nature of cpTi and TiAlV in asimulated oral environment has been investigated. Thesynergistic effect of electrochemical corrosion and wearunder mechanical sliding motion has been quantitativelyevaluated. It was hypothesized that LPS induces highwear/corrosion irrespective of the type of Ti.

2. Materials and methods

2.1. Surface preparation

In the present study, cpTi and TiAlV alloy discs, 12 mmdiameter and 7 mm thickness, were milled from titaniumrods (Mac-Master Carr, Elmhurst, IL, USA). The nominalchemical composition, density and molecular weight of cpTiand TiAlV alloy are shown in Table 1. A total of 24 discs wereused. The samples were divided into eight groups (n = 3)

as a function of Ti type (cpTi and TiAlV alloy) andEscherichia coli LPS (055:B5, Sigma-Aldrich, USA) concentration(0 - control, 0.15,15 and 150 µg/ml). Ti discs were wet-ground with 320–800 grit silicone carbide paper (Carbimet

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5 73

Table 1 – Nominal composition, density and molecular weight of cpTi and TiAlV alloy.

Ti typeComposition (wt%) Density

(cm3)

MolecularweightTi Al V C Fe O2 N2 H2

cpTi 99.7 – – 0.006 0.12 0.16 0.004 0.0019 4.50 47.85TiAlV 89.62 6.1 4.0 0.004 0.16 0.106 0.008 0.0022 4.44 46.72

2, Buehler, Lake Bluff, IL, USA), and further polished withpolishing cloth (TextMet Polishing Cloth, Buehler), diamondpaste (MetaDi 9-micron, Buehler) and lubricant (MetaDiFluid, Buehler). The surface was then mirror finishedwith chemomet polishing cloth (Chemomet I, Buheler) andcolloidal silica polishing suspension (MasterMed, Buehler)(final average roughness (Ra) of 8.68 ± 1.53 nm and10.63 ± 1.44 nm for cpTi and TiAlV alloy, respectively). Thediscs were ultrasonically cleaned (FS 20, Fisher Scientific,Pittsburgh, PA, USA) in deionized water and 70% propanol,and finally dried with hot air of 250 ◦C. Samples wererandomly allotted in each group. Artificial saliva was preparedaccording to Liu et al. (2007) and its composition wasKCl (0.4 g/L), NaCl (0.4 g/L), CaCl2·2H2O (0.906 g/L), NaH2PO4 ·2H2O (0.690 g/L),Na2S·9H2O (0.005 g/L), and urea (1 g/L).

2.2. Corrosion/wear test

Tribocorrosion tests can be conducted under differentelectrochemical conditions, such as free potential (a regularwear test under corrosive condition where the potential ismonitored), potentiostatic (a fixed potential is applied andthe current is monitored) and potentiodynamic (where thepotential is varied). In this study, potentiostatic test waschosen as it facilitates the estimation of the corrosion lossduring the tribocorrosion exposure. This will assist in theunderstanding of the influence of LPS and mechanical stresson the tribocorrosion behavior of Ti and the synergisticinteractions between wear and corrosion. For this reason, itis necessary to choose a potential based on the polarizationcurves of Ti. These are found through basic electrochemicaltests as follows.

2.2.1. Basic electrochemical testA basic electrochemical test was conducted in a custommade acrylic electrochemical cell (Fig. 1(a)) with a standard3-electrode setting according to the American Society forTesting for Materials (ASTM) guidelines (G61 and G31-72).A saturated calomel electrode (SCE) was used as thereference electrode, graphite rod as the auxiliary electrodeand the exposed surface of the Ti as the working electrode(the exposed area was 1.77 cm2). A potentiostat (G300,Gamry Inc., Warminster, PA, USA), linked to a computerfor data acquisition, was used to perform the corrosionmeasurements. A total volume of 150 ml of electrolytewas used for each corrosion experiment. In this trial salivawith no LPS (control) was used as an electrolyte. Similarpolarization curves of Ti were observed in environmentwith and without LPS (Barao et al., 2011). After mountingthe sample into the cell, approximately 4 h were allowedto achieve electrochemical stabilization of the system. Thetemperature of the test solution was maintained at 37 ± 1 ◦Cto mimic the oral environment.

Initially, open circuit potential (OCP) wasmonitored duringa period of 3600 s to evaluate the potential and to stabilize thesystem. Further, the samples were cyclically polarized from−0.8 to 1.8 V at a scan rate of 2 mV/s. Fig. 2 illustrated thecyclic polarization curves of cpTi and TiAlV alloy in artificialsaliva at pH 6.5 (normal oral saliva pH). Both Ti types exhibitedsimilar potentiodynamic curves, and a potential value of−0.3 V vs. SCE close to the free corrosion potential (Eoc)within the anodic region of the curve was chosen for thetribocorrosion experiment. It is important to clarify that theselected potential lies in the corresponds to an anodic regionwhen doing polarization curves beginning in a far cathodicpotential of −0.8 V, when the oxide film is partially reduced.

2.2.2. Tribocorrosion testThe tribocorrosion test apparatus is shown in Fig. 1(b). Thetribological system consisted of an alumina ceramic ball(28 mm in diameter) rubbing against the Ti sample (pinwith flat surface) in an electrolyte chamber. The standard3-electrode model as shown on the basic electrochemicaltest was employed. The exposed area of the sample to theelectrolyte was 1.13 cm2. A total of 150 ml of artificial salivaat pH 6.5 was used in each test. To evaluate the effect of LPSconcentration on the tribocorrosion behavior of Ti, differentLPS concentrations were added to the artificial saliva toprepare 0.15, 15 and 150 µg/ml of LPS. In the oral crevicularfluid the LPS concentration varied from 0.8 to 9.6 µg/ml in caseof healthy gingival sulcus and inflammation, respectively(Tzamouranis et al., 1979). Herein, higher LPS concentrationswere used owing to potentiate the possible effect of LPS onthe corrosion/wear behavior of Ti, similar as reported in aprevious study (Barao et al., 2011). Tests were performed ata specific anodic potentiostatic condition (−0.3 V vs. SCE)chosen based on the potentiodynamic curves establishedduring basic electrochemical test.

A standard protocol was employed for the tribocorrosiontest (Fig. 1(c)). Initially, samples were electrochemicallycleaned at cathodic potential of −0.9 V vs. SCE. This is donein an effort to achieve identical starting conditions. OCP wasthen monitored during a period of 600 s to evaluate thepotential and to stabilize the system (with a newly formedTiO2 film). Next, an electrochemical impedance spectroscopy(EIS) test was conducted to investigate the properties of theoxide film formed on the Ti surface (corrosion kinetics). TheEIS measurements were performed in the frequency rangefrom 100 kHz to 5 mHz (millihertz), with an AC sine waveamplitude of 10 mV oscillating about its corrosion potential.The measurement of potential and current were used todetermine the real (Z′) and imaginary (Z′′) components ofthe impedance, which were plotted with the Nyquist plot, orthe total impedance (|Z|) and phase angle. EIS results wereused to model the corrosion kinetics and to understand theproperties of the oxide film formed on the Ti surface. Data

74 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

Fig. 1 – (a) Schematic electrochemical set-up (a standard 3-electrode cell) used during basic corrosion test. (b) Schematictribocorrosion set-up. (c) Standard protocol used during tribocorrosion test (OCP — open circuit potential; PS —potentiostatic test; EIS — electrochemical impedance spectroscopy).

Fig. 2 – Representative cyclic polarization curves for cpTi and TiAlV alloy in artificial saliva at pH 6.5.

were adjusted using the Zview2 software (Scribner AssociatesInc., Southern Pines, NC, USA) by adopting two well-knownequivalent circuits (Bozzini et al., 2008; Messer et al., 2009;Moisel et al., 2008; Zhang et al., 2011) as further discussed.Double layer capacitance (Cdl) and polarization resistance(Rp) were determined. Impedance experimental data were

fitted with chi-square error < 0.001 (Barao et al., in press;Vidal and Munoz, 2008).

During the tribocorrosion test, evolution of current wasmonitored through a potentiostatic test, according to theprotocol presented in Fig. 1(c). The sliding duration (2000cycles), frequency (1.2 Hz), load (20 N, Initial Hertzian

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5 75

pressure estimation of 372 MPa, Final Hertzian pressureestimation of 16 MPa), and oscillatory amplitude (±5◦)

parameters were selected to simulate the oral environmentand mastication process. The coefficient of friction wasmeasured throughout the sliding process. After sliding, EISmeasurements were performed to investigate the changein the surface chemistry/condition after the tribocorrosionexposure. Finally, open circuit potential was monitored as apart of final stabilization of the test. After the test, sampleswere cleaned with deionized water as per the standardprotocol.

2.3. Surface characterization

A white light interferometry microscope (Zygo New View6300, Zygo Corporation, Middlefield, CT, USA) was used tocapture three-dimensional (3D) surfaces of the untouched(baseline) and worn/corroded Ti discs in order to understandthe changes on the damaged surface. Surface roughnesswas characterized using the same equipment. The averageroughness Ra was evaluated. Two areas were chosen forinvestigation: inside the wear scar and the area adjacentto the wear scar. Five measurements were performed ineach area for every sample and calculated the average. Ascanning electron microscope (SEM) (Jeol JSM-6490 LV, OxfordInstruments, Oxford, UK) was also used to characterize thesurfaces.

2.4. Total wear–corrosion weight loss

The weight loss during tribocorrosion was estimated and thesynergistic interaction between wear and corrosion can bestudied by using the following method according to Stackand Abdulrahman (2010). The total weight loss (Kwc) can beexpressed as

Kwc = Kw + Kc, (1)

where, Kw is the total weight loss due to (mechanical) wearand Kc is the total weight loss due to corrosion. The Kw andKc variables can be divided as follow:

Kw = Kwo + ∆Kw (2)

and

Kc = Kco + ∆Kc, (3)

where Kwo is the weight loss in the absence of corrosion, Kcois the corrosion weight loss in the absence of wear, ∆Kw is theeffect of corrosion on the weight loss due to wear and ∆Kc isthe effect of wear on the corrosion rate.

Total weight loss of the Ti samples (Kwc) was calculatedby profilometric analysis of the wear scar as described above.The weight loss due to corrosion (Kc) was estimated usingFaraday’s Law (Mathew et al., 2010):

Kc =M × i × tn × F

(4)

and

i × t = Q, (5)

where ‘M’ is the atomic mass of the material or equivalentweight in g/mol, ‘i’ is the total current in A/cm2, ‘t’ is the

total exposure time in seconds, ‘n’ is the number of electronsinvolved in the corrosion process which can be 2+, and 3+

(for the simplicity we assumed n = 2, for this study), ‘F’is Faraday’s constant (96500 C/mol−1), and ‘Q’ is the chargepassed through the working electrode in coulombs.

In order to classify the involved wear and corrosionmechanisms, a criteria was established by Stack andAbdulrahman (2010) based on the ratio of weight loss dueto corrosion (Kc) and weight loss due to wear (Kw). Themechanistic transitions were classified as follows:

Kc/Kw < 0.1 Wear (6)

0.1 ≤ Kc/Kw < 1 Wear–corrosion (7)

1 ≤ Kc/Kw < 10 Corrosion-wear (8)

Kc/Kw ≥ 10 Corrosion. (9)

2.5. Synergistic study

As per Eqs. (2) and (3), to calculate the weight loss inthe absence of corrosion (Kwo), a specific tribocorrosionexperiment was conducted in a cathodic protection potential(−0.9 V vs. SCE, the selection of this potential was based onthe potentiodynamic curve), in which corrosion is negligibleand material loss is only due to the tribological interactionbetween the two sliding bodies (Mischler, 2008).

Then the effect of corrosion on the weight loss due to wear,∆Kw is estimated by

∆Kw = Kw − Kwo. (10)

The corrosion rate in the absence of wear (Kco) wascalculated by measuring the current under potentiostaticcondition (without sliding) and based on Faraday’s equation(Eq. (4)). In addition, the effect of wear on the corrosion rate∆Kc is calculated by

∆Kc = Kc − Kco. (11)

Based on suchwear/corrosion regimes and to elaborate thesynergistic interactions, another ratio ∆Kw/∆Kc (enhance-ment of wear due to corrosion to enhancement of corrosiondue to wear) was estimated and classified into three groups(antagonistic, antagonistic–synergistic and synergistic) (Stackand Abdulrahman, 2010) as presented below:

∆Kw/∆Kc < 0.1 Antagonistic (12)

1 > ∆Kw/∆Kc ≥ 0.1 Antagonistic–synergistic (13)

∆Kw/∆Kc > 1 Synergistic. (14)

Generally, an antagonistic behavior is defined when theinterplay of corrosion and wear reduces the total materialweight loss (beneficial effect). On the other hand, synergisticbehavior is defined when corrosion enhances the wear; hencethe total weight loss (Stack and Abdulrahman, 2010).

2.6. Statistical analysis

Electrochemical parameters such as Cdl and Rp werestatistically analyzed using one-way ANOVA for analyzingeach factor separately (Ti type and LPS concentration) either

76 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

Fig. 3 – Representative curves of evolution of current density during the wear/corrosion test and the evolution of frictioncoefficient during sliding period for cpTi and TiAlV alloy in artificial saliva as a function of different LPS concentrations.

before or after sliding. Total weight loss, and roughnessdata were also analyzed by one-way ANOVA. Tukey HSDtest was chosen as the multiple-comparison technique whennecessary. T-test was used to compare cpTi and TiAlV. Paired-T test was used to compare the electrochemical data beforeand after sliding for each Ti type. A significant level of0.05 was used for all tests (Statistical Package for the SocialSciences, version 17.0; SPSS Inc, Chicago, IL, USA).

3. Results and discussion

3.1. Evolution of current and friction coefficient

Fig. 3 illustrates the evolution of current and frictioncoefficient during sliding period for cpTi and TiAlV alloy. Itis clearly noticeable that when sliding started, the currentsuddenly increased and during the whole period of sliding itshows consistent and noticeable fluctuations. This indicatesa destruction of passive oxide film on the Ti surfaceunder sliding action (Fernandes et al., 2006), which resultsin the exposure of the bare surface to the solution andreformation of oxide films, shifting from passive to activeregime (Mischler, 2008). TiO2 films in their native formpresent poor tribological properties and fractured under wearregimes (i.e. fretting and sliding) (Fernandes et al., 2006).Constant depassivation and repassivation processes on theworn surface during sliding were observed (Mischler et al.,1998). The current converged to the original value whensliding was stopped. A new and stable protective film formed(passivation) on the Ti surface, covering the electrochemicallyactive area of exposed specimen. Apparently, it indicates thatthe presence of LPS did not change the evolution of currentfor both Ti types.

The friction coefficient was in the range of 0.3–0.4 duringthe sliding test. The slight fluctuations may be attributed

to build-up and accumulation of third-body particles in thecontact region (Barril et al., 2004; Vieira et al., 2006). Inaddition, LPS groups showed similar coefficient of frictionwhen compared with the control.

3.2. Electrochemical impedance spectroscopy (EIS) mea-surements

The results from the EIS before and after sliding arestudied through Nyquist (impedance real vs. impedanceimaginary, Fig. 4) and Bode plots (Bode phase plot andBode impedance plot, Fig. 5). Nyquist plots demonstrate theelectrochemical resistance of the Ti surface. It is clearly visiblethat the presence of LPS reduces the semicircular diameterof capacitance loop, indicating poor corrosion resistance forboth cpTi and TiAlV alloy. After sliding, the diameter ofcapacitance loop suffered a higher decrease (Fig. 4), indicatingsome modification of the characteristics of the passive filmformed on the Ti surface (Fernandes et al., 2006).

Bode impedance plots provide the variation of impedanceas a function of frequency of the electrochemical double layer(with or without the presence of passive film) formed at theinterface of Ti and solution during corrosion process. Just onetime constant was observed for all groups, except for cpTiafter sliding which presented two time constants as indicatedin the Bode phase plot (Fig. 5). One time constant indicatesthe presence of a compact and homogeneous passive filmwhile two time constants indicate the formation of a porousfilm on the Ti surface (Souza et al., 2009). At high frequency,the phase angle and impedance exhibited low values for LPSgroups (mainly at 150 µg/ml of LPS) both before and aftersliding, corroborating with the Nyquist results.

3.3. EIS model

Two equivalent circuits were employed to estimate theformation and characteristics of the oxide film on the Ti

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5 77

Fig. 4 – Representative Nyquist plots from EIS recorded for cpTi and TiAlV alloy in artificial saliva as a function of differentLPS concentration, before and after sliding.

surface as a function of LPS presence, sliding period andTi type. In Randle’s equivalent circuit (Fig. 6(a)) the oxidelayer consists of a compact layer with a constant phaseelement (CPE − Cdl) and a polarization resistance (Rp). In thesecond circuit (Fig. 6(b)) we assumed that the oxide layer isformed by two layers: the porous outer layer represented bythe components Cdl out (capacity of the metal/film/electrolyteinterface) and Rp out (outer layer polarization resistance), andthe inner (barrier) layer, for lower frequencies, representedby the components Cdl in and Rp in (inner layer polarizationresistance). Rsol represents the solution resistance for bothcircuits. The impedance of a CPE is defined by ZCPE =

1/((jw)nC), where j =√

(−1) and w = 2πf , and the exponent ‘n’of the CPE is related to non-equilibrium current distributiondue to surface roughness (Ariza and Rocha, 2005; Hsuet al., 2004). The parameter ‘C’ is a constant, representingtrue capacitance of the oxide barrier layer. All results ofEIS were modeled by Randle’s circuit except for cpTi aftertribocorrosion test that was represented by the second circuit.

The presence of LPS influenced significantly on thecorrosion kinetics and resistance of both Ti types (Cdl cpTi p =

.018 before sliding, p = .431 after sliding; Cdl TiAlV p = .127before sliding, p = .026 after sliding; Rp cpTi p < 0.001 before

sliding, p = .058 after sliding; Rp TiAlV p = .003 before sliding,p = .028 after sliding). Fig. 6(c)–(d) shows the evolution of Rpand Cdl and the post-hoc comparisons (Tukey HSD test, T testand Paired-T test) for both Ti types, before and after sliding.Cdl increased with the increase of LPS concentration. Thesevalues were statistically significant for cpTi before sliding, andfor TiAlV alloy after sliding. Cdl of cpTi was statistically higherwhen compared to TiAlV alloy after sliding. For both Ti typesRp was statistically reduced in presence of LPS. TiAlV alloyexhibited higher Rp in comparison to cpTi regardless of thepresence or absence of LPS. As observed by others, the TiO2formed on the surface of TiAlV alloys is an efficient barrierto corrosion, and increases charge transfer resistance at thecorrosion interface (Hsu et al., 2004); therefore these samplesexhibited high polarization resistance (Fernandes et al., 2006).In addition the lower corrosion resistance of cpTi after slidingcompared to TiAlV may be due to the presence of a porouslayer in the previous case. The electrolyte can be penetratedthrough such pores in the passive film to the underlyingsurface of cpTi (Joska et al., 2010).

Most of the studies of corrosion resistance of Ti areconducted in saline or artificial saliva. Little is known aboutcorrosion kinetics and variation in the mechanisms induced

78 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

Fig. 5 – Representative Bode plots from EIS recorded for cpTi and TiAlV alloy in artificial saliva as a function of different LPSconcentration, before and after sliding. (Rsol—resistance of solution and Rp—polarization resistance).

by released-cell products such as LPS. Mabilleau et al. (2006)found that LPS-treated cells increased corrosion of Ti in salivaas evidenced by the increased Ti surface roughness. Messeret al. (2010) showed an increase of corrosion rate of machinedimplants when exposed to simulated inflammatory (THP1cells and LPS) and hyperglycemic (dextrose) conditions.

LPS are composed by lipid and polysaccharide joined by acovalent bond. According to Messer et al. (2009) the presenceof dextrose may carry away metal ions from the alloy surface,encouraging further corrosion. Therefore, the polysaccharidepart of the LPS molecule may also induce high metal releasefrom the Ti to the media which may explain the reductionof the corrosion inhibiting characteristics of the passive oxidefilm formed on the Ti surface. In addition, the passive film canbe formed with same defects in environment containing LPS,explaining the Cdl and Rp results in the present study.

After the wear test Cdl increased and Rp decreased forboth Ti types for all groups (Fig. 6). Some authors (Lucasand Lemons, 1992) stated that under static conditions theoxide layer on the Ti surface is stable and protective, butunder sliding and loading the film stability is compromised.This may explain the findings of the current study. Komotoriet al. (2007) observed a reduction of corrosion resistance ofTiAlV alloy under the simultaneous action of corrosion andwear in Ringer’s solution. Authors found that the reduction

is intensified by increasing the scratching rate and appliedpotential.

It is important to highlight that both the Cdl and Rpvalues may be affected by changing specimen’s dimensionsas promoted by the wear and rough worn surfaces, whichincreases during test. By estimating those topographicalchanges, an increase of 0.08% to 0.12% of the original area ofthe specimens was observed. Due to the small area changes(from unworn area to worn area) such effects were neglectedfor Cdl and Rp calculation.

3.4. Surface characterization

Fig. 7 illustrates the white light interferometry images of cpTiand TiAlV alloy before and after tribocorrosion (around thewear scar) as a function of LPS concentration. An increase insurface damage was observed after electrochemical attack.After tribocorrosion test a wear scar with well definedboundaries was observed on the Ti surface for all groups(Fig. 8(a)). The surface profile of a representative wear scar isshown (Fig. 8(b)).

Fig. 9 shows the surface roughness (Ra) of the samplesbefore and after tribocorrosion as a function of differentLPS concentrations. LPS did not affect the Ra values of Tiaround the wear scar area (p > .05; ANOVA); however, a

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5 79

Fig. 6 – (a)–(b) Equivalent circuit models selected for impedance spectra analysis. Evolution of (c) capacitance of double layer(Cdl), and (d) polarization resistance (Rp) for cpTi and TiAlV alloy in artificial saliva as a function of different LPSconcentrations. Lower case letters were used to compare the different LPS concentrations at the same Ti type and slidingperiod. Different letters indicate significant difference among the groups (α = 0.05). For each variable, (*) shows significantdifference between sliding periods at the same Ti type and LPS concentration (α = 0.05).

tendency of increasing surface roughness has been observedwhen the samples were exposed to saliva containing LPS. Theroughness inside the wear scar was affected by LPS for bothTi types (p < .0001; ANOVA): the higher the LPS concentration,the higher the surface roughness (p < .05; Tukey HSD test).Both Ti types exhibited similar surface roughness values (p ≥

.068; T-test) except for 15 µg/ml LPS groups at peripheral area(p = .027; T-test) and for 150 µg/ml LPS groups inside the wearscar (p = .040; T-test).

It is known that the corrosion process can increase theroughness of Ti (Correa et al., 2009; Mabilleau et al., 2006)based on the corrosion mechanisms. Our results showedthat LPS accelerate the wear/corrosion of cpTi and TiAlValloy, which may explain the greater surface roughness ofLPS groups. Surface roughness is an important property ofadhesion and colonization of bacteria (Morgan and Wilson,2001); therefore, a greater biofilm accumulation might beexpected on the Ti surface after wear/corrosion in presence ofLPS, which could lead to peri-implantitis (Bollen et al., 1996).

SEM images of the wear scars of both Ti types arepresented in Fig. 10. The worn surface is characterized bywear marks aligned in the direction of sliding. Under thetribocorrosion test conditions, two areas can be observedin the wear track. Boundary wear scar area exhibited somesmearing as a result of wear debris spread during slidingtest which removes the protective passive film and activelycorrode the Ti surface. The area inside of the wear scar isnoticeable with severe material damage and presence of weardebris (third body). Particles are observed in the central part ofthe wear scar probably due to the plastic surface deformation.

Those particles are likely to delaminate and detach from thesurface, inducing oscillations in the friction coefficient (Vieiraet al., 2006). Such effects also cause the transitions in the wearmechanisms from two body to three body. Delamination andcracking were observed mainly in the 150 µg/ml LPS group.The backscattering images of the 150 µg/ml LPS groups clearlyshow fatigue (arrows) of the Ti surface.

3.5. Wear–corrosion weight loss distribution

As explained in Section 2.4 the total weight loss couldbe separated into individual contributions from wear andcorrosion and listed in Table 2. The variation of weight lossas a function of LPS concentration is shown in Fig. 11. Theindividual contributions of corrosion (Kc) andwear (Kw) to theweight loss are presented. The combined effect of corrosionand wear (Kwc) is also illustrated in Fig. 11. In general, LPSsignificantly influenced the total weight loss due to corrosion(Kcp = .316 for cpTi and p = .029 for TiAlV alloy; ANOVA),the total weight loss due to wear (Kwp = .300 for cpTi andp = .014 for TiAlV alloy; ANOVA) and the total weight loss dueto wear and corrosion (Kwcp = .05 for cpTi and p = .130 forTiAlV alloy; ANOVA). For both Ti types, LPS increased the totalweight loss (Fig. 11). In LPS groups the passive oxide film isless protective (as observed in Cdl and Rp results) which mayinduce higher release of metals from Ti, and, consequently,high weight loss. Fernandes et al. (2006) observed correlationbetween degradation of the passive film with high weight lossof plasma nitride and plasma nitride + oxidized TiAlV alloyafter wear. It is also believed that in the presence of LPS the

80 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

Fig. 7 – White interferometry microscopy 3D images of cpTi and TiAlV alloy before and after tribocorrosion (around thewear scar) as a function of LPS concentration. X and Y scales are in mm, and Z scale is in µm. The Z scale covers 0.4 µm inamplitude.

Fig. 8 – White interferometry microscopy 3D images and surface profile of a representative wear scar of Ti. X and Y scalesare in mm, and Z scale is in µm. The Z scale covers 50 µm in amplitude.

passive film is harder to reform, leading to high total weight

loss. No significant difference in the total weight loss (Kc,Kw

and Kwc) was observed between cpTi and TiAlV alloy (p > .05,

T-test).

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5 81

Fig. 9 – Roughness average (Ra) of cpTi and TiAlV alloy (a) around the wear scar; (b) inside the wear scar. In the same Titype, means with different lower case letters are significant different (p > .05). In the same LPS concentration, differentupper case letters show significant difference between cpTi and TiAlV alloy (p > .05).

Fig. 10 – SEM photographs of corroded wear surface of cpTi and TiAlV alloy in artificial saliva as a function of different LPSconcentrations. Arrows indicate the cracking on the titanium surfaces through backscattering image technique.

3.6. Mechanistic and synergistic interaction between wearand corrosion

Identification of wear/corrosion regime transitions is made byusing the ratio of weight loss due to corrosion and weight

loss due to wear (Kc/Kw ratio). Kc/Kw values ranging from0.11 to 0.16 were obtained (Table 2). According to theseresults a wear dominated tribocorrosion mechanisms is thedriving degradation process (Stack and Abdulrahman, 2010).Those regimes of wear–corrosion were defined as additive,

82 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

Table 2 – Weight loss distribution as a function of LPS concentration for cpTi and TiAlV alloy in artificial saliva.

Ti type Kwc (µg) Kc (µg) Kw (µg) Kc/Kw Kwo (µg) ∆Kw (µg) Kco (µg) ∆Kc (µg) ∆Kw/∆Kc

cpTi .Control 96.31 ± 15.84 13.21 ± 1.25 83.09 ± 1.71 0.16 ± 0.05 14.17 68.92 ± 17.09 3.60E-3 ± 3.43E-3 13.21 ± 0.88 5.27 ± 1.30.15 µg/ml 111.27 ± 9.95 15.68 ± 1.25 95.58 ± 8.71 0.16 ± 0.01 12.69 82.90 ± 8.71 1.21E-2 ± 3.76E-3 15.67 ± 0.87 5.29 ± 0.115 µ g/ml 125.98 ± 24.91 15.64 ± 2.16 110.33 ± 27.06 0.15 ± 0.06 17.54 92.78 ± 27.06 3.32E-3 ± 1.42E-2 15.64 ± 1.52 6.1 ± 1.8150 µ g/ml 152.33 ± 64.10 16.69 ± 4.16 135.35 ± 59.98 0.12 ± 0.02 11.88 146.07 ± 59.98 7.48E-3 ± 7.49E-3 16.97 ± 4.16 7.0 ± 1.7

TiAlV .Control 106.01 ± 2.23 12.95 ± 2.49 93.05 ± 2.48 0.14 ± 0.01 36.84 56.21 ± 40.56 −4.84E-3 ± 7.06E-3 12.96 ± 18.37 4.3 ± 0.20.15 µg/ml 114.99 ± 10.40 14.96 ± 1.28 100.00 ± 10.40 0.15 ± 0.02 6.09 93.93 ± 38.84 7.25E-3 ± 1.34E-2 14.95 ± 0.01 6.3 ± 0.515 µg/ml 153.95 ± 8.78 15.07 ± 0.24 138.87 ± 8.64 0.11 ± 0.01 0.013 4.53 ± 54.03 1.91E-3 ± 9.51E-3 15.07 ± 0.24 0.3 ± 0.6150 µg/ml 129.81 ± 2.58 16.05 ± 1.94 113.76 ± 24.62 0.15 ± 0.03 0.015 −41.28 ± 70.35 −7.36E-3 ± 1.25E-2 16.05 ± 1.95 −2.7 ± 1.6

Notes:Kwc: total weight loss due to wear and corrosion (tribocorrosion)Kc: total weight loss due to corrosionKw: total weight loss due to (mechanical) wearKwo: weight loss in the absence of corrosion∆Kw: effect of corrosion on the weight loss due to wearKco: corrosion weight loss in the absence of wear.∆Kc: effect of mechanical wear on the corrosion rate.

Fig. 11 – Variation of weight loss of cpTi and TiAlV alloy asa function of different LPS concentration. The individualcontribution of corrosion (Kc) and wear (Kw) to the weightloss are expressed. The combined effect of corrosion andwear (Kwc) is illustrated by the whole bar. In the same Titype, means with different lower case letters are significantdifferent (p > .05) regarding Kwc values.

synergistic or antagonistic based on the ∆Kw/∆Kc ratiovalues (Stack and Abdulrahman, 2010). All groups exhibited asynergistic effect in which corrosion and wear are enhancingeach other i.e. a detrimental effect. However, for the TiAlValloy at 150 µg/ml LPS concentration, antagonistic effectwas observed in which corrosion possibly reduced the wear(Table 2). This later result can be explained by the decrease oftotal weight loss of TiAlV alloy at 150 µg/ml LPS concentration(Fig. 11).

These results indicate the complexity of the wear and/orcorrosion mechanisms during the tribocorrosion process. Ingeneral, material deterioration results from twomechanisms:mechanical wear and wear-accelerated corrosion (Mischler,2008). Mechanical wear (i.e. sliding) is related to the metaldetached from the material surface. During sliding, the oxidefilm on the Ti surface was removed and the new surfaceof Ti was exposed and more likely to be corroded thanthe TiO2 surface. Therefore, the wear accelerates corrosion

by removing the Ti protection until another passive filmis formed (Mischler, 2008). In the present study, the wear-accelerated corrosion mechanism was more pronounced inthe presence of LPS. The new Ti surface after sliding sufferedfrom increased attack by the LPS, which induced a highcorrosion rate and consequently more weight loss.

A simple schematic diagram has been developed todemonstrate the influence of LPS on the wear/corrosionmechanism of Ti (Fig. 12). Initially, Ti surface was protectedby the oxide film. Before sliding, the polysaccharide part ofLPS may attack the oxide film, inducing some defects. Part ofthe Ti surface was exposed, and an exchange of ions betweenTi (metal ions — M+) and saliva (electrons — e−) occurred.During the sliding, the rest of passive filmwas removed. Somedebris was observed and the new surface of Ti was attackedagain by the LPS, which increased the total wear weight loss.A new passive filmwas formedwhen sliding was stopped. Butaccording to our results, this film is less protective than thenative film (as observed by the Cdl and Rp results). In addition,the influence of LPS on the newly formed film may be basedon an adsorption tendency of LPS on Ti. However, all thesespeculations are not yet fully supported.

3.7. Clinical relevance, limitations and future scope

When Ti dental implants are inserted in oral cavity, astable passive film is formed. However, due to severalunfavorable conditions this passive film can break downand corrosion may take place. The present study showedthat the damage due to sliding wear, performed at 1.2 Hzresembling the mastication frequency, reduced the corrosionresistance of cpTi and TiAlV alloy. In addition, the presenceof components of membrane of gram negative bacteria,such as LPS, accelerated the wear/corrosion process of Ti.Therefore, patients with oral infections (i.e. periodontitis)may be more likely to have their dental implants corroded.Also, the surface roughness of Ti was increased by thepresence of LPS, which may induce greater bacterial biofilmaccumulation and consequently peri-implantitis, and further

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5 83

Fig. 12 – Schematic diagram of the tribological contact zone which exhibit the influence of LPS on the wear/corrosionmechanism of Ti. (a) before sliding; (b) during sliding; (c) after sliding. (e−—electrons and M+—metal ions).

implant therapy failure. Previous study has shown thatbacteria increased the wear/corrosion of Ti (Souza et al.,2010b). Therefore, this whole process will be transformed ina “vicious circle” in which LPS induces corrosion, corrosionincreases surface roughness, surface roughness promotemore bacteria adherence, bacteria induces corrosion and soon. For this reason it is very important that patients withoral Ti implants are assisted/instructed to practice a good oralhygiene.

The employed tribocorrosion set-up did not completelyresemble the actual adverse factors/conditions that im-plants are subjected in oral cavity. The simulated force wasuni-directional and constant in magnitude. Implants areexposed to vertical, horizontal and oblique forces, and high-intensity forces created by parafunctional habits (i.e. brux-ism). Furthermore, the motion amplitude was larger thanwhat is expected in-vivo and uni-directional rather thanmulti-directional. Smaller multi-directional amplitudes maygenerate a fretting-type tribological contact conditions. Inoral cavity implants are inserted into the bone, which mayassist to distribute the stress generated during loading. How-ever, in this study Ti was firmly held. The microbiologicalenvironment created in this study was based on LPS only(i.e. bacteria were missing). Further investigations with actualbacteria enriched with LPS are necessary. Specially focused

investigations are required to understand the effect of corro-sion products on the cell proliferation and viability, inflam-matory reaction, and bone level.

4. Conclusions

The tribocorrosion behavior of commercially pure titanium(cp-Ti) and titanium–aluminum–vanadium (TiAlV) alloy wasinvestigated in the presence of artificial saliva as a functionof lipopolysaccharide presence. Based on the outcomes of thepresent in-vitro study the following conclusions can be drawn:

• Lipopolysaccharide affected negatively the wear/corrosionbehavior of cpTi and TiAlV alloy by increasing thecapacitance and reducing the resistance of the titaniumoxide film.

• Both titanium types exhibited inferior corrosion behaviorafter sliding as observed in the capacitance and resistanceof double layer data and Nyquist plots.

• The total wear–corrosion weight loss was high whenlipopolysaccharide was added regardless of titanium type.Wear–corrosion was the dominant mechanistic model.

• The presence of lipopolysaccharide increased the rough-ness of the worn surface of both titanium types.

84 J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5

• Clinically, the results of this study showed that patientswith oral infections (presence of lipopolysaccharide) maybe more likely to have their dental implants corroded andwith greater bacterial biofilm accumulation which couldnegatively affect the implant prognosis.

Acknowledgments

Authors would like to express gratitude to CAPES (BEX:4129/09-1) and FAPESP (2007/58926-5) for the scholarshipsprovided to the second author (Barão, V.A.R). Authors wouldlike to thank Prof. R Urban and Ms. D Hall (Departmentof Orthopedic Surgery, Rush University Medical Center) fortheir assistance in SEM images. Acknowledgments also go toDr. Thomas Pandorf of Ceramtech, Plochingen, Germany forproviding ceramic heads.

R E F E R E N C E S

Ariza, E., Rocha, L.A., 2005. Evaluation of corrosion resistance ofmulti-layered Ti/glass-ceramic interfaces by electrochemicalimpedance spectroscopy. Functionally Graded Materials Viii492-493, 189-194.

Azzi, M., Szpunar, J.A., 2007. Tribo-electrochemical technique forstudying tribocorrosion behavior of biomaterials. Biomol. Eng.24, 443–446.

Barao, V.A., Mathew, M.T., Assuncao, W.G., Yuan, J.C., Wimmer,M.A., Sukotjo, C., 2011. The role of lipopolysaccharide on theelectrochemical behavior of titanium. J. Dent. Res. 90, 613–618.

Barao, V.A., Mathew, M.T., Assuncao, W.G., Yuan, J.C., Wimmer,M.A., Sukotjo, C., 2011. Stability of cp-Ti and Ti–6Al-4 Valloy for dental implants as a function of saliva pH —an electrochemical study. Clin. Oral Implants Res., in press(doi:10.1111/j.1600-0501.2011.02265.x).

Barril, S., Mischler, S., Landolt, D., 2004. Influence of frettingregimes on the tribocorrosion behaviour of Ti6Al4 V in 0.9 wt%sodium chloride solution. Wear 256, 963–972.

Bollen, C.M., Papaioanno,W., Van Eldere, J., Schepers, E., Quirynen,M., van Steenberghe, D., 1996. The influence of abutmentsurface roughness on plaque accumulation and peri-implantmucositis. Clin. Oral Implants Res. 7, 201–211.

Bozzini, B., Carlino, P., D’Urzo, L., Pepe, V., Mele, C., Venturo,F., 2008. An electrochemical impedance investigation of thebehaviour of anodically oxidised titanium in human plasmaand cognate fluids, relevant to dental applications. J. Mater.Sci. Mater. Med. 19, 3443–3453.

Casarin, R.C., Ribeiro Edel, P., Mariano, F.S., Nociti Jr., F.H., Casati,M.Z., Goncalves, R.B., 2010. Levels of aggregatibacter actino-mycetemcomitans, porphyromonas gingivalis, inflammatorycytokines and species-specific immunoglobulin G in general-ized aggressive and chronic periodontitis. J. Periodontal. Res.45, 635–642.

Chaturvedi, T.P., 2009. An overview of the corrosion aspect ofdental implants (titanium and its alloys). Indian J. Dent. Res.20, 91–98.

Correa, C.B., Pires, J.R., Fernandes-Filho, R.B., Sartori, R., Vaz, L.G.,2009. Fatigue and fluoride corrosion on Streptococcus mutansadherence to titanium-based implant/component surfaces. J.Prosthodont. 18, 382–387.

Cortada, M., Giner, L.L., Costa, S., Gil, F.J., Rodriguez, D., Planell,J.A., 2000. Galvanic corrosion behavior of titanium implantscoupled to dental alloys. J. Mater. Sci. Mater. Med. 11, 287–293.

Dalmiglio, M., Schaaff, P., Holzwarth, U., Chiesa, R., Rondelli, G.,2008. The effect of surface treatments on the fretting behaviorof Ti–6Al-4 V alloy. J. Biomed. Mater. Res. B Appl. Biomater. 86B,407–416.

de Assis, S.L., Wolynec, S., Costa, I., 2006. Corrosion characteri-zation of titanium alloys by electrochemical techniques. Elec-trochim. Acta 51, 1815–1819.

Fathi, M.H., Salehi, M., Saatchi, A., Mortazavi, V., Moosavi, S.B.,2003. In vitro corrosion behavior of bioceramic, metallic, andbioceramic–metallic coated stainless steel dental implants.Dent. Mater. 19, 188–198.

Fernandes, A.C., Vaz, F., Ariza, E., Rocha, L.A., Ribeiro, A.R.L.,Vieira, A.C., Riviere, J.P., Pichon, L., 2006. Tribocorrosionbehaviour of plasma nitrided and plasma nitrided plusoxidised Ti6A14 V alloy. Surf. Coat. Technol. 200, 6218–6224.

Henry, P.J., 1999. Clinical experiences with dental implants. Adv.Dent. Res. 13, 147–152.

Heuer, W., Elter, C., Demling, A., Neumann, A., Suerbaum, S.,Hannig, M., Heidenblut, T., Bach, F.W., Stiesch-Scholz, M., 2007.Analysis of early biofilm formation on oral implants in man. J.Oral Rehabil. 34, 377–382.

Hsu, R.W.W., Yang, C.C., Huang, C.A., Chen, Y.S., 2004. Electro-chemical corrosion properties of Ti–6Al-4V implant alloy in thebiological environment. Mat. Sci. Eng. A-Struct. 380, 100–109.

Huang, H.H., 2003. Effect of fluoride and albumin concentrationon the corrosion behavior of Ti–6Al-4 V alloy. Biomaterials 24,275–282.

Jones, F.H., 2001. Teeth and bones: applications of surface scienceto dental materials and related biomaterials. Surf. Sci. Rep. 42,79–205.

Joska, L., Fojt, J., 2010. Corrosion behaviour of titanium after short-term exposure to an acidic environment containing fluorideions. J. Mater. Sci. Mater. Med. 21, 481–488.

Joska, L., Fojt, J., Hradilova, M., Hnilica, F., Cvrcek, L., 2010.Corrosion behaviour of TiN and ZrN in the environmentcontaining fluoride ions. Biomed. Mater. 5, 054108.

Knoernschild, K.L., Tompkins, G.R., Lefebvre, C.A., Griffiths, L.L.,Schuster, G.S., 1996. Effect of pH on Porphyromonas gingivalisendotoxin affinity for resins. Int. J. Prosthodont. 9, 239–247.

Komotori, J., Hisamori, N., Ohmori, Y., 2007. The corrosion/wearmechanisms of Ti–6Al-4 V alloy for different scratching rates.Wear 263, 412–418.

Landolt, D., Mischler, S., Stemp, M., 2001. Electrochemicalmethods in tribocorrosion: a critical appraisal. Electrochim.Acta 46, 3913–3929.

Liu, I.H., Lee, T.M., Chang, C.Y., Liu, C.K., 2007. Effect of loaddeflection on corrosion behavior of NiTi wire. J. Dent. Res. 86,539–543.

Lucas, L.C., Lemons, J.E., 1992. Biodegradation of restorativemetallic systems. Adv. Dent. Res. 6, 32–37.

Mabilleau, G., Bourdon, S., Joly-Guillou, M.L., Filmon, R., Basle,M.F., Chappard, D., 2006. Influence of fluoride, hydrogenperoxide and lactic acid on the corrosion resistance ofcommercially pure titanium. Acta Biomater. 2, 121–129.

Mareci, D., Chelariu, R., Gordin, D.M., Ungureanu, G., Gloriant, T.,2009. Comparative corrosion study of Ti-Ta alloys for dentalapplications. Acta Biomater. 5, 3625–3639.

Mathew, M.T., Ariza, E., Rocha, L.A., Vaz, F., Fernandes, A.C., Stack,M.M., 2010. Tribocorrosion behaviour of TiCxOy thin films inbio-fluids. Electrochim. Acta 56, 929–937.

Mathew, M.T., Srinivasa Pai, P., Pourzal, R., Fischer, A., Wimmer,M.A., 2009. Significance of tribocorrosion in biomedicalapplications: overview and current status. Adv. in Tribology2009, Article ID 250986, 250912 pages.

Messer, R.L., Seta, F., Mickalonis, J., Brown, Y., Lewis, J.B., Wataha,J.C., 2010. Corrosion of phosphate-enriched titanium oxidesurface dental implants (TiUnite) under in vitro inflammatoryand hyperglycemic conditions. J. Biomed. Mater. Res. B Appl.Biomater. 92, 525–534.

J O U R N A L O F T H E M E C H A N I C A L B E H AV I O R O F B I O M E D I C A L M A T E R I A L S 8 ( 2 0 1 2 ) 7 1 – 8 5 85

Messer, R.L., Tackas, G., Mickalonis, J., Brown, Y., Lewis, J.B.,Wataha, J.C., 2009. Corrosion of machined titanium dentalimplants under inflammatory conditions. J. Biomed. Mater.Res. B Appl. Biomater. 88, 474–481.

Mischler, S., 2008. Triboelectrochemical techniques and interpre-tation methods in tribocorrosion: a comparative evaluation.Tribol. Int. 41, 573–583.

Mischler, S., Debaud, S., Landolt, D., 1998. Wear-acceleratedcorrosion of passive metals in tribocorrosion systems. J.Electrochem. Soc. 145, 750–758.

Moisel, M., de Mele, M.A.F.L., Muller, W.D., 2008. Biomaterial inter-face investigated by electrochemical impedance spectroscopy.Adv. Eng. Mater. 10, B33–B46.

Morgan, T.D., Wilson, M., 2001. The effects of surface roughnessand type of denture acrylic on biofilm formation byStreptococcus oralis in a constant depth film fermentor. J.Appl. Microbiol. 91, 47–53.

Nakagawa, M., Matsuya, S., Shiraishi, T., Ohta, M., 1999. Effectof fluoride concentration and pH on corrosion behavior oftitanium for dental use. J. Dent. Res. 78, 1568–1572.

Nakagawa, M., Matsuya, S., Udoh, K., 2002. Effects of fluoride anddissolved oxygen concentrations on the corrosion behavior ofpure titanium and titanium alloys. Dent. Mater. J. 21, 83–92.

Nelson, S.K., Knoernschild, K.L., Robinson, F.G., Schuster,G.S., 1997. Lipopolysaccharide affinity for titanium implantbiomaterials. J. Prosthet. Dent. 77, 76–82.

Nikolopoulou, F., 2006. Saliva and dental implants. Implant Dent.15, 372–376.

Popa, M.V., Vasilescu, E., Drob, P., Vasilescu, C., Demetrescu, I.,Ionita, D., 2008. Long-term assessment of the implant titaniummaterial–artificial saliva interface. J. Mater. Sci. Mater. Med. 19,1–9.

Schiff, N., Grosgogeat, B., Lissac, M., Dalard, F., 2002. Influenceof fluoride content and pH on the corrosion resistance oftitanium and its alloys. Biomaterials 23, 1995–2002.

Simon, B.I., Goldman, H.M., Ruben, M.P., Baker, E., 1969. Therole of endotoxin in periodontal disease. I. A reproducible,quantitative method for determining the amount of endotoxinin human gingival exudate. J. Periodontol. 40, 695–701.

Sinnett-Jones, P.E., Wharton, J.A., Wood, R.J.K., 2005. Micro-abrasion-corrosion of a CoCrMo alloy in simulated artificial hipjoint environments. Wear 259, 898–909.

Souza, J.C., Henriques, M., Oliveira, R., Teughels, W., Celis, J.P.,Rocha, L.A., 2010a. Biofilms inducing ultra-low friction ontitanium. J. Dent. Res. 89, 1470–1475.

Souza, J.C.M., Henriques, M., Oliveira, R., Teughels, W., Celis, J.P.,Rocha, L.A., 2010b. Do oral biofilms influence the wear andcorrosion behavior of titanium? Biofouling 26, 471–478.

Souza, M.E., Lima, L., Lima, C.R., Zavaglia, C.A., Freire, C.M., 2009.Effects of pH on the electrochemical behaviour of titaniumalloys for implant applications. J. Mater. Sci. Mater. Med. 20,549–552.

Stack, M.M., Abdulrahman, G.H., 2010. Mapping erosion-corrosionof carbon steel in oil exploration conditions: Some newapproaches to characterizing mechanisms and synergies.Tribol. Int. 43, 1268–1277.

Tzamouranis, A., Matthys, J., Ishikawa, I., Cimasoni, G., 1979.Increase of endotoxin concentration in gingival washingsduring experimental gingivitis in man. J. Periodontol. 50,175–177.

Vidal, C.V., Munoz, A.I., 2008. Electrochemical characterisationof biomedical alloys for surgical implants in simulated bodyfluids. Corros. Sci. 50, 1954–1961.

Vieira, A.C., Ribeiro, A.R., Rocha, L.A., Celis, J.P., 2006. Influence ofpH and corrosion inhibitors on the tribocorrosion of titaniumin artificial saliva. Wear 261, 994–1001.

Wilson, M., 1995. Biological activities of lipopolysaccharides fromoral bacteria and their relevance to the pathogenesis of chronicperiodontitis. Sci. Prog. 78, 19–34.

Zhang, B.B., Wang, B.L., Li, L., Zheng, Y.F., 2011. Corrosion behaviorof Ti–5Ag alloy with and without thermal oxidation in artificialsaliva solution. Dent. Mater. 27, 214–220.