Chapter 17

Corrosion Protective Coatings for Ti and Ti Alloys Used for Biomedical Implants

Liana Maria Muresan    “Babes-Bolyai” University, Faculty of Chemistry and Chemical Engineering, Cluj-Napoca, Romania

Abstract

Due to extensive use of Ti and Ti alloys as biomedical implants, considerable efforts have been made to enhance their corrosion resistance as well as their biological response to surrounding tissues, by using a range of mechanical, chemical, and physical methods.

In this context, recent advances in preparation and characterization of corrosion resistant coatings on Ti substrates are reviewed with emphasis on passive films (oxides, ceramics, hybrid layers, composites, etc.). The methods most frequently used to produce the protective coatings (electrooxidation, laser oxidation, sol-gel spin coating technique, etc.) are reviewed and the challenges for future research are briefly discussed.

Keywords

Titanium

Titanium alloys

Biomaterials

Surface modification

Corrosion

17.1 Introduction

Titanium and its alloys are biologically inert materials widely used in the production of biomedical implants due to their high tensile and fatigue strength, low Young modulus, and superior corrosion and wear resistance.1 Moreover, they are biocompatible, providing direct bonding with bone surface in orthopedic and dental surgery, and are inert in physiological environments.2

Titanium is a material with a high superficial energy and after implantation it provides a favorable body reaction that leads to direct deposition of minerals on the bone-titanium interface and titanium osseointegration.3 This is due mainly to the thin TiO2 film naturally formed on its surface, which is very stable and chemically inert. Nevertheless, bulk titanium used in implants presents three main problems4:

 high processing energy due to melting and casting difficulties

 slow reaction with the human tissues resulting in difficulties in bone attachment

 higher elastic modulus compared to bone

One of the ways to reduce the problems associated with stress generated by the incompatibility between the elastic modulus of Ti implant and that of the bone is to fabricate titanium implants with controlled porosity.5 Recently, porous Ti scaffolds with good mechanical properties (particularly, elastic modulus and stiffness) that provide a favorable environment for bone growth and a significant enhancement in the osteoblastic activity were produced.6

Although “commercially pure” titanium has acceptable mechanical properties and has been used for different implants, presently, for most biomedical applications, titanium is alloyed with nickel or with small amounts of aluminum, vanadium, or niobium. These elements improve some of the Ti characteristics such as strength, mechanical properties, corrosion resistance, and so on. Titanium alloys most frequently used in dental and orthopedic implantations are NiTi (Nitinol), Ti-6Al-4V, and Ti-6Al-7Nb.

NiTi is a material with excellent shape memory effect and super elasticity and damping capacity suitable for use in orthopedic implants. In dentistry, the material is used in orthodontics for brackets and wires connecting the teeth. However, it should be mentioned that its high Ni content constitutes a potential threat to its safe use in vivo.7

Ti-6Al-4V is the material used for ~ 20-30% of commercially available medical devices. It contains 6% Al and 4% V by weight and is by far the most common Ti alloy. As in the case of NiTi, during the Ti-6Al-4V use for implants the release of metallic ions in the physiological environment could be harmful to the health. Thus, Al and V may cause neurological disorders and bone diseases when they are released in the human body.8

Ti-6Al-7Nb is another titanium alloy with excellent strength and biocompatibility dedicated for surgical implants, used especially for replacement hip joints. It is a typical α + β alloy, with Al as stabilizer for the α phase and Nb as stabilizer for the β phase.2 It is usually stable in neutral halide solutions, but in acidic media is prone to accelerated dissolution.9

Irrespective of the material nature, the most important requirement for the long-term success of implants is the stable interface between the biomaterial and the surrounding tissue.10 Hence, an improvement of all Ti-based materials corrosion resistance is of particular importance. In this regard, TiO2 naturally formed on the surface is one of the main candidates for protective coatings on Ti and Ti alloys substrates due to its thermodynamic stability, chemical inertia, and low solubility in body fluids.11 Titania with specific structures of anatase and rutile was found to stimulate apatite formation12 and were proved to be much more beneficial for bone growth than the amorphous TiO2.13 However, the native TiO2 layer is very thin, irregular, and porous and there is also the risk of viruses being carried by this layer. Moreover, the TiO2 naturally formed on Ti surface is not sufficient to ensure a strong chemical bond with bone tissue and a safe, long lifetime of the implants. Consequently efforts are made for thickening the oxide layer along with an improvement of its surface ability to promote osteosynthesis and to increase its corrosion and wear resistance.

The development of nanostructures has allowed unprecedented access to an intimate interaction with the biological environment. For bone regeneration and repair, re-creating substrates on a nanoscale allows scientists to probe the first level of the bone structural hierarchy.14 Studies have shown that TiO2 films consisting of nanostructures (nanotubes, nanorods, nanosheets, etc.) exhibit high hydrophilicity, improving thus the bioactivity and bone-bonding behavior of materials for implants.15,16 However, there is still a lot to do in the direction of improving the behavior of implant biomaterials. Some of the most widely used methods to modify the characteristics of materials surface are presented below.

17.2 Surface Modification Methods

As already mentioned, optimization of the biomaterial surface is necessary in order to trigger osseointegration, to minimize metallic ion release, and to ensure a longer lifetime of an implant. There are several possible ways to form a surface layer on the implant materials17:

(i) materials deposition on the substrate (e.g., by sol-gel method, spray method, or electrolytic deposition)

(ii) material removal from the substrate (e.g., by etching)

(iii) substrate surface transformation (e.g., by anodic oxidation)

Some of the most used methods to modify the surface of Ti and Ti alloys are presented below.

17.3 Sol-Gel Method

Sol-gel process is a method for producing solid materials from small molecules that is suitable for preparing different coatings (e.g., silicium and titanium oxides) on the surface of Ti-based materials. It involves conversion of small molecules (precursors) into a colloidal solution (sol) and then into an integrated network (gel) consisting of either discrete particles or network polymers.

The main advantages of the sol-gel method are18: (1) low processing temperature (avoiding volatilization of entrapped species), (2) the possibility to cast coatings in complex shapes, and (3) the use of compounds that do not introduce impurities into the end product. Among the disadvantages of this method one can count the relatively long period of time for processing flow and the difficulties related to phase separation occurring especially in hybrid coating synthesis.

The sol-gel process involves four stages: (1) hydrolysis, (2) condensation/polymerization of monomers, (3) growth of particles, and (4) gel formation.18 These processes are influenced by several experimental parameters such as pH, temperature, concentration of the reactants, and presence of additives.

Traditional precursors for sol-gel coatings are alkoxysilanes such as tetraethyl orthosilicate (TEOS) and tetramethyl orthosilicate (TMOS), but efforts are made in order to find less toxic and more environmental friendly precursors.

TiO2 can be prepared, for example, starting from precursor solutions containing an alkoxide (e.g., tetrabutylorthotitanate) and diethanolamine dissolved in ethanol16 and mixing them with water (with or without additives) in a certain ratio.

SiO2 coatings can be prepared starting from a precursor solution consisting of tetraethylorthosilicate Si(OC2H5)4, H2O, C2H5OH, and HCl, mixed at a certain molar ratio.19 Different compounds (inhibitors, pigments, etc.) can be added in order to improve the physicochemical properties of the coating.

Irrespective of the nature of the film, the two main techniques used to apply a sol-gel coating on the surface of a metallic substrate are dip-coating and spin-coating.

17.3.1 Dip-coating

By this technique, the material from which the film is produced is put into solution, and then the substrate is progressively dipped into and is extracted from the solution at a controlled rate (Figure 17.1). After the solvent evaporates, a thin and homogeneous film is produced. The thickness of deposited liquid film coatings depends on the coating solution properties such as density, viscosity, and surface tension, as well as surface withdrawal speed from the coating solution. The thickness of the film is generally bigger than that prepared by spin-coating with the same solutions.

f17-01-9780124114678
Figure 17.1 Schematics of a dip-coating process.20

Dip-coating was successfully used, for example, to prepare sol-gel-derived Al2O3 films on γ-TiAl-based alloys,21 porous TiO2 films,16 hydroxyapatite (HA) coatings,22 and SrO-SiO2-TiO2 on NiTi,23 and so on.

17.3.2 Spin-coating

In the case of spin-coating, an amount of solution is placed on the substrate that is rotated at high speed in order to spread the fluid by centrifugal force (Figure 17.2). After the evaporation of the solvent, a thin, homogeneous film is formed. As in the case of dip-coating, final film thickness and other properties will depend on the nature of the sol-gel coating (viscosity, drying rate, surface tension, etc.) and on the parameters chosen for the spin process. Higher spin speeds and longer spin times give birth to thinner films.25 Generally, a moderate spinning speed is recommended. The drying rate should also be slow in order to ensure film uniformity. A supplementary drying step is sometimes necessary after the high-speed spin step to further dry the film without substantially thinning it.

f17-02-9780124114678
Figure 17.2 Schematics of a spin-coating process.24

By using spin-coating, SiO2 films were deposited on Ti-48Al alloy,19 sphene (CaTiSiO3) ceramics on Ti-6Al-4V26 alkoxide-based HA nanocoatings on Ti and Ti-6Al-4V substrates,27 HA28 and HA/polymer coatings on Ti,29 TiO2 on Ti-6Al-4V,30 and many others.

17.4 Laser Oxidation

Laser surface treatment is a thermal process that has an advantage over conventional furnace treatment. It is based on the heating caused by the light adsorption of the surface layer and the cooling ensured by the high conductivity of the material.31 The adsorbed laser energy results in a thin surface layer with desired properties while the bulk of the material is unaffected.

By using laser treatments followed by AFM investigations, it was proved that the morphological modifications of Ti-based alloys play a key role in cell adherence and proliferation by influencing cell-material interactions.32

Porous Ti samples were fabricated by using laser power in an attempt to improve the osteoconductivity and reduce the problems associated with stress shielding.5

NiTi were treated with a pulsed Nd:YAG laser in air, by using selected parameters aiming to obtain uniform oxide films of controlled thickness.7 The corrosion resistance of the so-treated NiTi samples increased about 15 times, while the surface Ni/Ti ratio was reduced. The last feature is important as the high Ni content in NiTi constitutes a potential threat to its safe use in vivo.

17.5 Anodic Oxidation

Anodic oxidation is a traditional surface modification method of Ti and its alloys that allows the obtaining of the desired properties of oxide layer by controlling the electrochemical parameters, such as electrolyte composition and concentration, applied potential or current, temperature, and so on.

Anodic oxidation can be performed potentiostatically33 or galvanostatically34 in different electrolytes. The most common electrolyte solutions used for anodization are H2SO4,12,33 Na2SO4,33 CH3COOH,33,35,36 H3PO4,33 and HF.13,14 Mixture of inorganic substances or hybrid inorganic-organic solutions could be also used.2,37,38 It was reported that porous anodic films consisting of and/or anatase are produced by anodizing in H2SO4 and Na2SO4 solutions, while amorphous TiO2 films are produced in CH3COOH and H3PO4.33 Self-organized TiO2 nanotubes can be generated from viscous electrolytes consisting of glycerol or ethylene glycol with addition of ammonium fluoride.38 Viscosity of the electrolyte plays a key role in the nanotubes growth process.

The oxidation potentials are usually determined before the anodic oxidation step from potentiodynamic polarization experiments.35

Oxide layers formed by anodization differ according to the substrate, for example, oxidation in chromic acid produced a TiO2 layer on titanium substrate, but layers consisting of TiO2 and Al2O3 oxides on a Ti-6Al-4V substrate.39 The minor oxides formed by anodic oxidation in the TiO2 layer on the Ti-6Al-7Nb alloy are Al2O3 and Nb2O5.35

In order to prepare layers with high apatite-forming ability, anodic oxidation can be carried out in complex solutions containing calcium and/or phosphate ions.40

17.6 Plasma Electrolytic Oxidation

Plasma electrolytic oxidation is a relatively inexpensive and simple method used for increasing surface roughness of the materials, allowing also the synthesis of ceramiclike oxide films on some metals.41 It is similar to anodizing, but it employs higher potentials so that the resulting plasma modifies the structure of the oxide layer. Thus, the method combines electrochemical oxidation with a spark treatment in the plating bath. If this contains aluminate, phosphate, silicate, and sulfate anions or some of their combinations, bioactive layers enriched in these elements are formed.42 The incorporation of Ca and P into an oxide film formed via anodic oxidation on Ti-15Mo alloy41 or Ti-6Al-7Nb,43 as well as the formation of a bioactive film on Ti-6Al-4V8 after the application of a voltage higher than the breakdown voltage of the oxide layer, were also reported. Plasma electrolytic oxide coatings are generally recognized for high hardness, wear, and corrosion resistance.

17.7 Electrolytic Deposition

A promising route for deposition of thin oxide films from aqueous solutions is the electrolytic deposition method using TiCl4 or TiOSO4 as starting materials in the presence of hydrogen peroxide.44 The procedure consists in galvanostatic electrodeposition at 0 °C from a solution obtained by slow addition of TiCl4 and H2O2 and subsequent adjustment of the solution volume. The resulting films are uniform, of high purity, present excellent adhesion to the substrate, and have good mechanical properties. The electrolytic deposition can be followed by annealing.

A possible mechanism for TiO2 electrodeposition is based on hydrolysis of the peroxo-complex of Ti in the presence of OH ions generated at the cathode.45

TiCl4Ti4++4Cl

si1_e

Ti4++H2O2+n2H2OTiO2OHn24n++nH+

si2_e

TiO2OHn24n++mOH+kH2OTiO3H2Ox

si3_e

2TiO3H2Ox2TiO2+O2+2xH2O

si4_e

A fluorine-doped HA/ZrO2 two-layer coating was prepared by electrodeposition on titanium in ZrO(NO3)2 aqueous solution and subsequently in a mixed solution of Ca(NO3)2, NH4H2PO4, and NaF. The adherence of the two-layer coating was found to be significantly higher than that of the pure HA coating.46

17.8 Combined Methods

Sometimes, the combination of two or more methods is used in order to improve surface layer properties and to attain the desired characteristics. Thus, the combined laser/sol-gel synthesis of calcium silicate coating was produced on Ti-6Al-4V substrates for improved cell integration.47

Anodic oxidation treatment applied on laser-processed porous Ti samples was used to produce bioactive TiO2 nanotube arrays that significantly changed the surface properties and enhanced the apatite-forming ability of porous TiO2 samples in simulating body fluid.5 By creating the nanotube arrays, the healing time due to increased biocompatibility can be significantly reduced.

The polymeric sponge replication method followed by microarc oxidation (MAO) was used to prepare HA/TiO2 hybrid coatings on highly porous Ti scaffolds.6 For this purpose, a polyurethane sponge was covered with titanium hydride slurry (consisting of TiH2 dispersed in ethanol containing triethyl phosphate and polyvinylbutyl), and then heated at 800 °C to convert TiH2 to Ti. The surface was then coated with a hybrid, bioactive, microporous HA/TiO2 layer using MAO.

17.9 Protective Films

The protective layers produced on the surface of Ti and Ti-based alloys for enhancing their corrosion resistance, as well as their biocompatibility, are of different natures: oxides, composites, HA, hybrid organic-anorganic, ceramic, and so on.

17.9.1 Oxides

Oxide-based films are the most frequently produced on the Ti substrates. The most important factors deciding their corrosion resistance are the oxide thickness, and its evenness and compactness. The main oxide formed on the surface of Ti alloys is TiO2, but minor oxides of the alloying elements (Al2O3, Nb2O5, NiO, etc.) are also produced.

TiO2 can be produced in different crystallographic forms (anatase, rutile, brookite; see Figure 17.3), depending on the substrate nature and the preparation conditions.

f17-03-9780124114678
Figure 17.3 Crystallographic forms of TiO2.

From the point of view of osseointegration abilities, crystalline forms are superior to amorphous ones and nanostructured layers are superior to microstructured ones. For example, rutile can induce apatite formation, especially when it has a crystal orientation in the (101) plane, because the matching structure with apatite (002) promotes the nuclei formation for epitaxial crystal growth. On the contrary, amorphous titania layers are not able to induce apatite formation.33

Osteoblasts are very sensitive to the different materials’ topographies and respond differently on submicron- and nanometer-sized surfaces. This is why nanostructured TiO2 are preferred and are often generated on the top of the thin TiO2 naturally formed on the Ti surface, aiming to accelerate osteogenesis. Besides, the systems based on nanotubes are very hydrophilic and promote greater cell adhesion and increased channeling for fluid exchange that contributes to bone remodeling.14

The adhesion/propagation of the osteoblasts is substantially improved on vertically aligned and laterally spaced TiO2 nanotubes.15 This type of nanotube can be obtained by anodic oxidation of Ti in diluted HF acid13,14 and confer to the substrate increased surface area, porous structure, and superior osteoblast cell growth.

Anodization of the same material in different electrolytes leads to the formation of different oxidic nanostructures. For example, oxide layers obtained by anodic oxidation of Ti6Al7Nb alloy in (NH4)2SO4 + NH4F solution are more homogeneous than those obtained in a solution containing glycerol and NH4F and water.2 The concentration of F in the electrolytes ions is also important: the larger its concentration, the higher the diameters of pores and the surface roughness. It should be also noticed that, as the electrolytes used for anodization are fluoride-based, the nanotubes always contain a certain amount of incorporated fluorine,48 which could be useful for the enhancement of antibacterial properties of the surface.

17.9.2 Hydroxyapatite

Hydroxyapatite, with the formula Ca10(PO4)6(OH)2, is the major inorganic component of bones that is able to accelerate bone ingrowths onto the surface of a biomedical implant. HA has excellent biocompatibility and bioactivity, but bad mechanical properties; therefore, HA is usually deposited on the Ti-based implants in order to obtain materials that possess both bioactivity and mechanical resistance.

HA can be deposited on titanium-based surfaces by using different methods, such as electrochemical deposition,49 plasma spraying,50,51 and sol-gel method.52 Sometimes, an intermediate TiO2 layer is produced between the Ti and HA layer for improving the bonding capacity of HA on the Ti substrate and, thus, improving the integrity of the coating.27,53 Composite materials with improved bioactivity and corrosion resistance can be obtained by embedding HA particles in titania gel, depositing this mixture by spin-coating onto substrates and then calcinating it.52

17.9.3 Composites

Different composite layers can also be deposited on Ti-based biomaterials in order to improve the physicochemical properties and the bioactivity of the surface.

Knowing the beneficial effect of strontium on bone formation, SrO-SiO2-TiO2 composite coatings were produced by sol-gel-method on NiTi substrates aimed at increasing its corrosion resistance and cytocompatibility.23 The coating significantly decreased the release of nickel ions and exhibited enhanced adhesion of osteoblastlike cells. Biphasic fluorine-doped HA/Sr-substituted HA (FHA/SrHA) were prepared on Ti substrates by sol-gel and dip-coating methods with the intention to induce rapid cellular responses after implantation and long lifetime.54 The dispersion of SrHA in FHA has promoted growth and attachment of osteoblast cells. In the same time, an increase of SrHA content determined an increase of surface roughness.

To accelerate the formation of HA on TiO2 modified surfaces, a simulated body fluid with high ionic strength can be used for a biomimetic apatite deposition process.55

A class of composites of particular importance is represented by Ag/TiO2 coatings. Starting from the fact that silver is one of the most interesting antibacterial materials and that TiO2 is an excellent support material, different Ag/TiO2 composites were prepared by using various methods. Some of them are summarized in Table 17.1.

Table 17.1

Composite Ag/TiO2 Coatings

SubstrateComposite CoatingsPreparation MethodPropertiesReference
Rutile-structured TiO2 particlesAg/TiO2Micellar layer-by-layer depositionAntibacterian activity against gram-negative E. coli56
TiAg/TiO2Electrochemical depositionCatalytic activity towards oxygen reduction57
TiAg/TiO2Sol-gel spin coatingAntibacterian activity against gram-negative E. coli58
Ti foilAg/TiO2-nanotubesAnodization of Ti followed by sputter-deposition of AgEnhanced SERS activity59
TiAgHA/TiO2Plasma electrolytic processingCorrosion resistance60

t0010

17.9.4 Hybrid coatings

Hybrid inorganic-organic coatings have drawn a lot of attention due to their unique properties determined by the synergetic combination of inorganic and organic components. When hybrid coatings are deposited on the surface of biomaterials, they significantly improve their characteristics.

HA/TiO2/poly(lactide-co-glycolide) coatings were prepared by sol-gel method on titanium substrates29 and compared with HA layers obtained by plasma spraying. The composite layers possessed no cracks, demonstrated increased adhesion of osteoblasts, and presented micron surface roughness.

Apatite/collagen hybrid coatings were prepared on NiTi alloy through electrochemical deposition61,62 from an electrolyte containing 0.1 mg/ml collagen dissolved in a double-strength simulated body fluid. Apatite/collagen-coated NiTi samples possessed higher wettability and corrosion resistance than either uncoated or apatite- coated samples.

17.9.5 Ceramic coatings

Glass-ceramics, such as hardystonite (Ca2ZnSi2O7) and sphene (CaTiSiO5), have attracted much attention in the biomedical field because they provide great possibilities to manipulate their properties by posttreatments, including strength, degradation rate, and coefficient of thermal expansion.

Nanostructured glass-ceramics prepared by a plasma spray technique on Ti-6Al-4V alloys using conventional powders exhibited enhanced capacity for osteoblasts attachment, which has been ascribed to the Ca and Si ions released from the coatings.26,63 Ceramic coatings consisting of TiO2 rutile and TiAl2O5 were produced on Ti-6Al-4V alloy by an alternating-current, microarc oxidation in aluminate solution.64 Calcium phosphate ceramic (CPC) coatings were electrophoretically deposited and sintered on titanium or its alloys.65 Other ceramic coatings such as CaO-P2O5-TiO2-Na2O exhibiting excellent adhesion to the substrate during the tensile and fatigue tests were deposited on Ti-29Nb-13Ta-4.6Zr.66 Ceramic layers consisting of Al2TiO5, − α Al2O, and rutile TiO2 were prepared on Ti-6Al-4V alloy by pulsed bipolar microplasma oxidation (MPO) in NaAlO2 solution.67 Extending the MPO time, the surface roughness of the coatings increased whereas the thickness and the compactness of the coatings were improved.

17.10 Corrosion Studies

The most important requirement for a biomaterial is its compatibility with the human body. The bioimplants should not cause adverse effects such as allergies or inflammations and should not be toxic. Other important requirements for a successful implant material are very high mechanical, corrosion, and wear resistance in order to ensure a long service period (over 15 years).

In the physiological environment, biomedical implants are subjected to corrosion, especially to mechanically accelerated electrochemical processes such as stress corrosion, corrosion fatigue, and fretting corrosion. Fretting corrosion, alone or in combination with crevice corrosion, has been identified as one of the most important modes of corrosion of Ti-based implants such as hip, knee, and shoulder replacements.9,68

Due to corrosion processes, even the most resistant materials undergo electrochemical exchange and release metal ions in the physiological environment, producing different inconvenient or even implant failure.69 For Ti-based alloys, mostly Ti is released, but also Al and V were detected in the case of Ti-Al-V alloy.9 It is accepted that the tolerable corrosion rate for metallic implants should be about 2.5 × 10− 4 mm/year.70 In this context, the failure due to corrosion remains one of the challenging clinical problems and all the implantable materials should be tested from a corrosion behavior point of view, apart from other properties. Moreover, different approaches should be used in order to enhance the corrosion resistance.

A way to stabilize the passive film formed on Ti-based surfaces is Ti alloying with selected elements that stabilize either the α-phase (Al) or the β-phase (Mo, V, Fe) of the alloy and enlarge its passive range.71,72 Other methods consist, as already mentioned, in modifying the surface of the implant materials in a convenient way.

In order to investigate the corrosion behavior of Ti-based biomaterials, different artificial physiological solutions can be used simulating plasmatic serum (Ringer’s and Hank’s solution), saliva, or sweat. The composition of some of these artificial physiological solutions is presented in Tables 17.217.4.

Table 17.2

Chemical Composition of Simulated Physiological Solutions73

Physiological SolutionSubstanceConcentration (g/l)pH
HankNaCl8.06.9
KCl0.4
NaHCO30.35
NaH2PO4·H2O0.25
Na2HPO4·2H2O0.06
CaCl2·2H2O0.19
MgCl20.19
MgSO4·7H2O0.06
Glucose1.0
RingerNaCl8.696.4
KCl0.30
CaCl20.48

t0015

Table 17.3

Chemical Composition of Artificial Saliva (pH 6.75)74

CompoundConcentration (g/l)
Methyl-p-hydroxybenzoate2.00
Na Carboxymethylecellulose10.0
MgCl2·6H2O0.059
CaCl2·2H2O0.166
K2HPO40.804
KCl0.625
KH2PO40.326

Table 17.4

Chemical Composition of Artificial Sweat75

Chemical CompositionConcentration (% (w/v))
AATCCISOISOEN
pH 4.3pH 5.5pH 8.0pH 6.5
l-histidine monohydrochloride monohydrate (C6H9O2N3·HCl·H2O)0.0250.050.05
Sodium chloride (NaCl)1.000.500.501.08
Disodium hydrogen orthophosphate dodecahydrate (Na2HPO4·12H2O)0.50
Sodium dihydrogen orthophosphate dihydrate (NaH2PO4·2H2O)0.22
Disodium hydrogen orthophosphate anhydrous (Na2HPO4)0.10
Lactic acid (88%)0.0970.12
Urea0.13

t0025

The in vitro evaluation of the corrosion behavior is an important step in the elaboration of new biomaterials. The corrosion resistance of biomaterials used for implants is usually tested by electrochemical methods because they are accelerated tests that allow evaluating rapidly the performance of materials and predicting their long-time behavior. The most common electrochemical methods are open circuit potential measurements, cyclic voltammetry, potentiodynamic polarization measurements, and electrochemical impedance spectroscopy.

By comparing three titanium-based materials (Ti, NiTi, and Ti-6Al-7Nb), it was reported that Ti-6Al-7Nb alloy possesses the greatest corrosion resistance in simulated Hanks physiological solution at pH 7.4 and 37 °C, this behavior being valid for the untreated samples as well as for the samples anodically oxidized at 3.0V in acetic acid.35 The polarization resistance values, determined from EIS measurements for the oxidized Ti-6Al-7Nb alloy during the later stages of immersion in Hank’s solution, are one order of magnitude higher than those determined for the untreated alloy. Thus, the anodized samples possess a much higher corrosion resistance than the untreated ones.

A two-layer model was proposed to accurately describe the corrosion behavior of the oxidized Ti alloys. Thus, it is generally accepted that the surface film consists of a two-layer oxide, composed of a dense inner layer and a porous outer layer76 (Figure 17.4).

f17-04-9780124114678
Figure 17.4 Scheme of the two-layer model proposed for oxidized Ti alloys.

The high corrosion resistance is due to the thin barrier-type inner layer, while its osseointegration ability is attributed to the porous outer layer. The significant enhancement of the polarization resistance values at longer immersion time in the physiological solution could be due to the stabilization of the barrier layer (e.g., by some sealing processes inside pores of the oxide film77), which might diminish the tendency of the alloy to corrosion. The nature of the porous layer depends on the nature of the alloy and the anions present in the solution.78 The two-layer model accurately describes also the electrochemical behavior of HA coatings on Ti-6Al-4V in Ringer’s physiological solution.79

17.11 Conclusions

Surface modification is one of the ways to significantly improve the corrosion and wear resistance of Ti-based materials along with their surface texture and biocompatibility. Protective layers of different natures (oxides, HA, composites, hybrid layers, etc.) can be generated on the materials’ surface by various methods such as electrooxidation, laser oxidation, and sol-gel spin coating technique.

The performance of biomaterials used for implants should be evaluated by determining their physical, chemical, and biological properties. As the materials behavior in vivo is a complex one, the understanding of the implant materials behavior requires joint efforts of materials science, medicine, biology, and engineering. The in vitro experiments should be considered only as screening tests.

References

1 Milosev I. Metallic materials for biomedical applications: laboratory and clinical studies. Pure Appl Chem. 2011;83(2):309–324.

2 Mindroiu M, Pirvu C, Ion R, et al. Comparing performance of nanoarchitectures fabricated by Ti6Al7Nb anodizing in two kinds of electrolytes. Electrochim Acta. 2010;56:193–202.

3 Acero J, Calderón J, Salmeron J, et al. The behaviour of titanium as a biomaterial: microscopy study of plates and surrounding tissues in facial osteosynthesis. J Cranio Maxill Surg. 1999;27:117–123.

4 Oldani C, Dominguez Al. Titanium as a biomaterial for implants. In: Fokter S, ed. Recent advances in arthroplasty. Rijeka, Croatia: InTech; 2012. Available from: http://www.intechopen.com/books/recent-advances-in-arthroplasty/titanium-as-a-biomaterial-for-implants.

5 Das K, Balla VK, Bandyopahyay A, et al. Surface modification of laser-processed porous titanium for load-bearing implants. Scr Mater. 2008;59:822–825.

6 Lee J-H, Kim H-E, Koh Y-H. Highly porous titanium (Ti) scaffolds with bioactive microporous hydroxyapatite TiO2 hybrid coating layer. Mater Lett. 2009;63:1995–1998.

7 Wong MH, Cheng FT, Man HC. Laser oxidation of NiTi for improving corrosion resistance in Hank’s solution. Mater Lett. 2007;61:3391–3394.

8 Krzakala A, Sluzalska K, Dercz G, et al. Characterization of bioactive films on a Ti-6Al-7Nb alloy. Electrochim Acta. 2013;104:425–438.

9 Virtanen S, Milosev I, Gomez-Barrena E, et al. Special modes of corrosion under physiological and simulated physiological conditions. Acta Biomater. 2008;4:468–476.

10 Ochsenbein A, Chai F, Winter S, et al. Osteoblast responses to different oxide coatings produced by the sol-gel process on titanium substrates. Acta Biomater. 2008;4:1506–1517.

11 Breme J. Titanium and titanium alloys, biomaterials of preference. Rev Metall. 1989;10:625–637.

12 Yang B, Uchida M, Kim H-M, et al. Preparation of bioactive titanium metal via anodic oxidation treatment. Biomaterials. 2004;25:1003–1010.

13 W-q Yu., Qiu J, Xu L, et al. Corrosion behaviors of TiO2 nanotube layers on titanium in Hank’s solution. Biomed Mater. 2009;4:1–6.

14 Brammer KS, Frandsen CJ, Jin S. TiO2 nanotubes for bone regeneration. Trends Biotechnol. 2012;30(5):315–322.

15 Oh S, Daraio C, Chen L-H, et al. Significantly accelerated osteoblast cell growth on aligned TiO2 nanotubes. J Biomed Mater Res A. 2006;78:97–103.

16 Fu T, Wu X-ming., Wu F, et al. Surface modification of NiTi alloy by sol-gel derived porous TiO2 film. Trans Nonferrous Met Soc China. 2012;22:1661–1666.

17 Toth C, Szabo G, Kovacs L, et al. Titanium implants with oxidized surfaces: the background and long-term results. Smart Mater Struct. 2002;11:813–818.

18 Wang D, Bierwagen GP. Sol-gel coatings for corrosion protection. Prog Org Coat. 2009;64:327–338.

19 Teng S, Liang W, Li Zh., et al. Improvement of high-temperature oxidation resistance of Ti-Al-based alloy by sol-gel method. J Alloys Compd. 2008;464:452–456.

20 http://hosting.umons.ac.be/php/lpsi/Dip.

21 Zhang XJ, Li Q, Zhao SY, et al. Improvement in the oxidation of γ-TiAl-based alloy by sol-gel derived Al2O3 film. Appl Surf Sci. 2008;255:1860–1864.

22 Zhang JX, Guan RF, Zhang XP. Synthesis and characterization of sol-gel hydroxyapatite coatings deposited on porous NiTi alloys. J Alloys Compd. 2011;509:4643–4648.

23 Zheng CY, Nie FL, Zheng YF, et al. Enhanced corrosion resistance and cellular behavior of ultrafine-grained biomedical NiTi alloy with a novel SrO-SiO2-TiO2 sol-gel coating. Appl Surf Sci. 2011;257:5913–5918.

24 http://hosting.umons.ac.be/php/lpsi/Spin.

25 Tredwin CJ, Georgiou G, Kim H-W, et al. Hydroxyapatite, fluor-hydroxyapatite and fluorapatite produced via the sol-gel method: bonding to titanium and scanning electron microscopy. Dent Mater. 2013;29(5):521–529.

26 Wu C, Ramaswamy Y, Gale D, et al. Novel sphene coatings on Ti-6Al-4V for orthopedic implants using sol-gel method. Acta Biomater. 2008;4:569–576.

27 Roest R, Latella BA, Heness G, et al. Adhesion of sol-gel derived hydroxyapatite nanocoatings on anodized pure titanium amd titanium (Ti6Al4V) alloy substrates. Surf Coat Technol. 2011;205:3520–3529.

28 Carrado A, Viart N. Nanocrystalline spin coated sol-gel hydroxyapatite thin films on Ti substrate: towards potential applications for implants. Solid State Sci. 2010;12:1047–1050.

29 Sato M, Slamovich EB, Webster TJ. Enhanced osteoblast adhesion on hydrothermally treated hydroxyapatite/titania/polya(lactide-co-glycolide) sol-gel titanium coatings. Biomaterials. 2005;26:1349–1357.

30 Zaveri N, McEwen GD, Karpagavalli R, et al. Biocorrosion studies of TiO2 nanopartickle-coated Ti-6Al-4V implant in simulated biofluids. J Nanopart Res. 2010;12:1609–1623.

31 Montealegre MA, Castro G, Rey P, et al. Surface treatments by laser technology. Contemp Mater. 2010;I–1:19–30.

32 Guillemot F, Prima F, Tokarev V, et al. Sur la biofonctionnalisation d’alliages de titane par traitement laser: Résultats et perspectives. Appl Phys A. 2003;77:899–904.

33 Cui X, Kim H-M, Kawashita M, et al. Preparation of bioactive titania films on titanium metal via anodic oxidation. Dent Mater. 2009;25:80–86.

34 Bayat N, Sanjabi S, Barber ZH. Improvement of corrosion resistance of NiTi sputtered thin films by anodization. Appl Surf Sci. 2011;257:8493–8499.

35 Milosev I, Blejan D, Varvara S, et al. Effect of anodic oxidation on the corrosion behavior of Ti-based materials in simulated physiological solution. J Appl Electrochem. 2013;43:645–658.

36 Shi P, Cheng FT, Man HC. Improvement in corrosion resistance of NiTi by anodization in acetic acid. Mater Lett. 2007;61:2385–2388.

37 Dey T, Roy P, Fabry B, et al. Anodic mesoporous TiO2 layer on Ti for enhanced formation of biomimetic hydroxyapatite. Acta Biomater. 2011;7:1873–1879.

38 Macak JM, Schmuki P. Anodic growth of self-organized anodic TiO2 nanotubes in viscous electrolytes. Electrochim Acta. 2006;52:1258–1264.

39 Zwilling V, Darque-Ceretti E, Boutry-Forveille A, et al. Structure and physicochemistry of anodic oxide films on TA6V alloy. Surf Interface Anal. 1999;27:629–637.

40 Ishizawa H, Ogino M. Formation and characterization of anodic titanium oxide films containing Ca and P. J Biomed Mater Res. 1995;29:65–72.

41 Simka W, Krzakala A, Korotin DM, et al. Modification of a Ti-Mo alloy surface via plasma electrolytic oxidation in a solution containing calcium and phosphorus. Electrochim Acta. 2013;96:180–190.

42 Yerokhin AL, Nie X, Leyland A, et al. Characterisation of oxide films produced by plasma electrolytic oxidation of a Ti–6Al–4V alloy. Surf Coat Technol. 2000;130:195–206.

43 Krzakala A, Sluzalska K, Widziolek M, et al. Formation of bioactive coatings on a Ti-6Al-7Nb alloy by plasma electrolytic oxidation. Electrochim Acta. 2013;104:407–424.

44 Kern P, Schwaller P, Michler J. Electrolytic deposition of titania films as interference coatings on biomedical implants: microstructure, chemistry and nano-mechanical properties. Thin Solid Films. 2006;494:279–286.

45 Zhitomirski I, Gal-Or L, Kohn A, et al. Electrodeposition of Ceramic Films from Non-Aqueous and Mixed Solutions. J Mater Sci. 1995;30:5307–5312.

46 Huang Y, Yan Y, Pang X. Electrolytic deposition of fluorine-doped hydroxyapatite/ZrO2 films on titanium for biomedical applications. Ceram Int. 2013;39:245–253.

47 Mirhosseini N, Crouse PL, Li L, et al. Combined laser/sol-gel synthesis of calcium silicate coating were produced on Ti-6Al-4V substrates for improved cell integration. Appl Surf Sci. 2007;253:7998–8002.

48 Matykina E, Hernandez-Lopez JM, Conde A, et al. Morphologies of nanostructured TiO2 doped with F on Ti–6Al–4V alloy. Electrochim Acta. 2011;56:2221–2229.

49 Y-y Zhang. Tao J, Pang Y-c, Wang W, Wang T. Electrochemical deposition of hydroxyapatite coatings on titanium. Trans Nonferrous Met Soc China. 2006;16:633–637.

50 Demnati I, Parco M, Grossin D, et al. Hydroxyapatite coating on titanium by a low energy plasma spraying mini-gun. Surf Coat Technol. 2012;206:2346–2353.

51 Quek CH, Khor KA, Cheang P. Influence of processing parameters in the plasma spraying of hydroxyapatite/Ti–6Al–4V composite coatings. J Mater Process Tech. 1999;89–90:550–555.

52 Su B, Yu Zhang GX, Wang C. Sol-gel derived bioactive hydroxyapatite/titania composite films on Ti6A14V. J Univ Sci Technol Beijing. 2006;13:469–475.

53 Kim HW, Koh YH, Li LH, et al. Hydroxyapatyte film on titanium substrate with titania buffer layer processed by sol-gel method. Biomaterials. 2004;25:2533–2538.

54 Yin P, Feng FF, Lei T, et al. Colloidal-sol gel derived biphasic FHA/SrHA coatings. Surf Coat Tech. 2012;207:608–613.

55 Sun T, Wang M. Low temperature biomimetic formation of apatite/TiO2 composite coatings on Ti and NiTi shape memory alloy and their characterization. Appl Surf Sci. 2008;255:396–400.

56 Lin W-C, Chen C-N, Tseng T-T, et al. Micellar layer-by-layer synthesis of TiO2/Ag hybrid particles for bactericidal and photocatalytic activities. J Eur Ceram Soc. 2010;30:2849–2857.

57 Mentus SV, Boskovic I, Pjesic JM, et al. Tailoring the morphology and electrocatalytic properties of electrochemically formed Ag/TiO2 composite deposits on titanium surfaces. J Serb Chem Soc. 2007;72:1403–1418.

58 Mai L, Wang D, Zhang S, et al. Synthesis and bactericidal ability of Ag/TiO2 composite films deposited on titanium plate. Appl Surf Sci. 2010;257:974–978.

59 Roguska A, Kudelski A, Pisare MK, et al. In situ spectroelectrochemical SERS investigations on composite Ag/TiO2 nanotubes/Ti substrates. Surf Sci. 2009;603:2820–2824.

60 Kotharu V, Ngumothu R, Arumugam CB, et al. Fabrication of corrosion resistant, bioactive and antibacterial silver subtituted hydroxyapatite/titania composite coating on Cp Ti. Ceram Int. 2012;38:731–740.

61 Sun T, Lee W-C, Wang M. A comparative study of apatite coating and apatite/collagen composite coating fabricated on NiTi shape memory alloy through electrochemical deposition. Mater Lett. 2011;65:2575–2577.

62 Hu K, Yang X-J, Cai YL, et al. Prepartion of bone-like composite coating using a modified simulated body fluid with high Ca and P concentrations. Surf Coat Technol. 2006;201:1902–1906.

63 Wang G, Lu Z, Liu X, et al. Nanostructured glass–ceramic coatings for orthopaedic applications. J R Soc Interface. 2011;8(61):1192–1204.

64 Wang C, Chen R, Deng Z, et al. Structure and properties characterization of ceramic coatings produced on Ti–6Al–4V alloy by microarc oxidation in aluminate solution. Mater Lett. 2002;52:435–444.

65 Kim CS, Ducheyne P. Compositional variations in the surface and interface of calcium phosphate ceramic coatings on Ti and Ti-6Al-4V due to sintering and immersion. Biomaterials. 1991;12:461–469.

66 Li SJ, Niinomi M, Akahori T, et al. Fatigue characteristics of bioactive glass-ceramic-coated Ti-29Nb-13Ta-4.6Zr for biomedical application. Biomaterials. 2004;25(17):3369–3378.

67 Yao Z, Jiang Z, Xin S, et al. Electrochemical impedance spectroscopy of ceramic coatings on Ti–6Al–4V by micro-plasma oxidation. Electrochim Acta. 2005;50:3273–3279.

68 Khan MA, Williams RL, Williams DF. Conjoint corrosion and wear in titanium alloys. Biomaterials. 1999;20:765–772.

69 Kohn DH. Metals in medical applications. Curr Opin Solid State Mater Sci. 1998;3:309–316.

70 Manivasagam G, Dhinasekaran D, Rajamanickam A. Biomedical implants: corrosion and its prevention-A review. Recent Patents Corr Sci. 2010;2:40–54.

71 Gonzalez JEG, Mirza-Rosca JC. Study of the corrosion behavior of titanium and some of its alloys for biomedical and dental implant applications. J Electroanal Chem. 1999;471:109–115.

72 Gad MMA, Mohamed KE, El-Sayed AA. Effect of molybdate ions on the corrosion behavior of Ti alloys. J Mater Sci Tech. 2000;16:45–49.

73 Bundy KJ. Corrosion and other electrochemical aspects of biomaterials. Crit Rev Biomed Eng. 1994;22:139–251.

74 McKnight-Hanes C, Whitford GM. Fluoride release from three glass ionomer materials and the effects of varnishing with or without finishing. Caries Res. 1992;26:345–350.

75 Kulthong K, Srisung S, Boonpavanitchakul K, et al. Determination of silver nanoparticle release from antibacterial fabrics into artificial sweat. Part Fibre Toxicol. 2010;7:1–8.

76 Lavos-Valereto IC, Wolynec S, Ramires I, et al. Electrochemical impedance spectroscopy characterization of passive film formed on implant Ti-6Al-7Nb alloy in Hank’s solution. J Mater Sci Mater Med. 2004;15:55–59.

77 Pan J, Thierry D, Leygraf C. Electrochemical impedance spectroscopy study of the passive oxide film on titanium for implant application. Electrochim Acta. 1996;41:1143–1153.

78 Aziz-Kerzo M, Conroy KG, Fenelon AM, et al. Electrochemical studies on the stability and corrosion resistance of titanium-based implant materials. Biomaterials. 2001;22:1531–1539.

79 Souto RM, Laz MM, Reis RL. Degradation characteristics of hydroxyapatite coatings on orthopaedic TiAlV in simulated physiological media investigated by electrochemical impedance spectroscopy. Biomaterials. 2003;24:4213–4221.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset