Sunteți pe pagina 1din 47

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/275727417

Biomaterials Corrosion

Article · January 2009


DOI: 10.1515/CORRREV.2009.27.S1.287

CITATIONS READS

4 1,720

2 authors:

Sridhar tm Sivasankari SARAVANAN Rajeswari


University of Madras Ramco Institute of Technology
20 PUBLICATIONS 32 CITATIONS 120 PUBLICATIONS 2,328 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Nanobioceramic coatings View project

Corrosion of implants View project

All content following this page was uploaded by Sridhar tm on 06 February 2016.

The user has requested enhancement of the downloaded file.


BIOMATERIALS CORROSION

T.M. Sridhar1 and S. Rajeswari2

1
Department of Biomedical Engineering,
SMK FOMRA Institute of Technology, Old Mahabalipuram Road,
Thaiyur Village, Kelambakkam, Chennai 603 103, India

2
Department of Analytical Chemistry
University of Madras, Guindy Campus, Chennai 600 025.
Fax No: +91 44 22352494, Tel. No. +91 44 22351137.
E-mail : tmsridhar@gmail.com, profsrajeswari@yahoo.co.uk

ABSTRACT:

The ability to replace or augment diseased body parts totally or partially


has improved both the quality and life span of human population. The decline
in surgical risks during recent decades has encouraged the development of
more complex procedures for prosthetic implantation. Additionally, a variety
of extracorporeal devices, such as heart, lung and blood dialysis machines,
are used routinely, but these prosthetic elements have several limitations.
Hence, research projects are currently underway to overcome the limitations
of synthetic materials by developing formulations with varying properties,
such as asymptomatic, long-term function in the human physiological
environment etc., to meet the needs of biomedical surgeons. This review
focuses on biomaterials corrosion and its measures to prevent corrosion.

Keywords: Biomaterials, metals, corrosion, orthopedic and dental.

INTRODUCTION

Many types of materials are available and used in various medical


devices. These materials are used in conjunction with medical intervention.
The generally accepted term for such materials is “biomaterials”. A
Biomaterials Corrosion

biomaterial is defined as “any synthetic material that is used to replace or


restore function of a body tissue and is continuously or intermittently in
contact with body fluids”. The ability to replace or augment diseased body
parts totally or partially, has improved the quality of life for millions of
people over the quarter of a century. In the course of a single year, thousands
of patients throughout the world receive some kind of implant device. Such
devices include artificial hips, knees, elbows, pacemakers, heart-valves,
intraocular lenses etc. The materials from which these devices are constructed
range from the sophisticated and highly specialized to the simple and low
technology designs. Orthopaedic implant materials and devices are needed
for a number of reasons, such as replacing tissue that has become damaged or
destroyed by disease or trauma /1,2/.
Today, the field of biomaterials has evolved such that more than 50
different materials are used in more than 40 types of complex prosthetic
devices. One of the most important developments in clinical medicine has
been the replacement of diseased joints with artificial implants. Since the
early cemented hip replacements of the 1960’s there has been a constant flow
of new materials and designs for implantable devices and the number is
steadily increasing as technological progress and medical practice allow for
increased functionality and safety.

HISTORICAL PERSPECTIVES

The history of biomaterials dates back to antiquity. Many of the initial


thrusts were attempted by man to correct deformities, since, in the years
before anesthesia and asepsis, surgical procedures were limited to the body
surface. Earlier surgeries performed were those by Hindu surgeons for the
restoration of missing parts. Susrutha, in 600 B.C, repaired an injured nose
with a patch of living flesh from the cheek region. This technique for nose
reconstruction migrated from East to West. Around 1430, the Brancas, a
family of Sicilian Laymen, perfected the method now referred to as the
Italian method for nose construction by using skin flap taken from the arm. In
the nineteenth century, Von Graefe and Gillies in England, Davis, Ivy and
Kazanjian in the United States and Filator in Russia were stimulated by
World War I tragedies to pioneer newer methods of wound closure and tissue
transfer /3/.

2
T.M. Sridhar and S. Rajeswari Corrosion Reviews

The earliest written record of an application of metal in surgical


procedure is from the year 1565. However, until Lister’s aseptic surgical
technique was developed in the 1860s, various metal devices such as wires
and pins which were constructed of iron, gold, silver, platinum etc. and tissue
transplantations were not largely successful mainly due to infection after
implantation. The use of alloys in surgical implants is a relatively modern
development, dating back about a century. Early steel formulations were
found to degrade rapidly in the physiological environment and also to
produce adverse effect. Lane in England (1893 – 1912) designed a fracture
plate using steel. During the twentieth century (in the mid 1920s) type 316
stainless steel was developed and was found to be stronger than previous
alloys and to have higher resistance to corrosion in body fluids. In the 1930s,
cobalt-chromium (CoCr) alloys were developed for the aerospace industry
and were found to have superior characteristics to steel for many prosthetic
applications. Around the middle of the 20th century, pure titanium (Ti)
implants, which have excellent biocompatibility, were used and over the next
few decades, Ti alloys with superior mechanical characteristics to the pure
metal were developed for orthopedic surgery. In recent years, a new material
– the ‘smart’ nickel-titanium (NiTi) alloy often referred to as “Nitinol” (the
shape-memory alloy) and the high-strength alloy MP35N have been
attracting considerable interest for prosthetic implants. Nowadays, CoCr and
Ti alloys are the most widely used alloys for orthopedic surgery, with 316
and 316L (low carbon) steel and nitinol being used in specific situations and
for specific applications /4/. A brief summary of historical developments of
biomaterials is described in Table 1.
Other novel metals and alloys are being investigated for possible
prosthetic applications but have not yet been fully assessed. All metallic
implantable alloys are susceptible to corrosion to a certain degree, depending
on the metallurgical condition, residual or service stresses, thermal history
and final surface treatment applied prior to implantation.

RECONSTRUCTION MATERIALS

Tissue replacements with synthetic materials are achieved by selecting the


material, which has physical properties most similar to those of natural tissue.
Table 2 illustrate surgical applications of a wide range of materials namely
metals and alloys, ceramics, polymers and composites.

3
Biomaterials Corrosion

Table 1
Major historical developments of biomaterials /5/

Stainless steels, titanium alloys and cobalt chromium alloys are used
universally for most of the high load bearing applications in skeletal system.
Conducting metals like platinum-iridium alloys are used for electrical
stimulation of the heart and nervous tissues. Nitinol, an alloy of nickel and
titanium finds applications in orthodontics /5/.

4
T.M. Sridhar and S. Rajeswari Corrosion Reviews

Table 2
Materials used as implants

The stable and inert nature of alumina, zirconia and titania ceramics
enhances their potential usage in orthopedic joint replacements /6/. The
chemical inertness and abrasive resistance provide improvements over the
hitherto widely used metals. The degradable ceramics, which are almost

5
Biomaterials Corrosion

invariably based on calcium phosphates, find applications in hard tissue


regeneration.
The resemblance of physical nature of polymers with soft tissues of
humans plays a dominant role in the replacement of skin, tendons, cartilage,
vessel walls, lens, breast and bladder. A number of synthetic polymers find
applications as biomaterials /7/.

REQUIREMENTS FOR A PERFECT BIOMATERIAL

Materials to be implanted into the human body are subjected to exposure


in body fluids, which are hostile and extremely sensitive, that restricts the
applications of a perfect biomaterial. Additionally, there is the crucial fact
that human body tissues are extremely sensitive to foreign materials and are
very easily stimulated into showing signs of poisoning and rejection. Hence a
biomaterial must be biocompatible - the ability of a material to perform with
an appropriate host response in a specific application (i.e. it should not elicit
an adverse response from the body, and vice versa). A perfect biomaterial
should not cause chronic inflammation, impairment of cellular functions and
should be non-toxic and non-carcinogenic. Adequate physical and
mechanical properties are necessitated to meet the required demands of the
body. Fig. 1 lists the requirements for a perfect biomaterial for successful
application in total joint replacement /8/. The ideal material or material
combination should exhibit the following properties:

Fig. 1: Implant material requirements in orthopedic applications

6
T.M. Sridhar and S. Rajeswari Corrosion Reviews

™ A biocompatible chemical composition to avoid adverse tissue reactions


™ Excellent resistance to degradation (e.g., corrosion resistance for metals
or resistance to biological degradation in polymers)
™ Acceptable strength to sustain cyclic loading endured by the joint
™ A low modulus to minimize bone resorption
™ High wear resistance to minimize wear debris generation

NATURE OF THE PHYSIOLOGICAL ENVIRONMENT

The nature of physiological environment is extremely hostile to all


foreign materials and hence the effect of environment on metallic implant and
the effect of implant on its host tissue are of primary concern (Fig. 2). The
concentrations of chloride ions in serum and interstitial fluid are 113 and 117
mEq l-1, respectively, which is about 1/3 the concentration of brine and a
seriously corrosive environment for metallic materials /9/. The aqueous layer
at the implant site will naturally contain numerous hydrated ionic species
(Na+, Ca2+ etc.). The type and concentration of ions in solution is likely to
change over a period of time as the cells surrounding the implant react and
adapt to the presence of foreign material. Other electrolytes present in the
body fluid include Ca2+, Mg2+, PO43-, SO42- and organic acid anions. Some
complex compounds present in smaller amounts are phospholipids,
cholesterols, natural fats, proteins, glucose and aminoacids /10/. In addition,
the concentration of dissolved oxygen is 1/4 that of air in venous blood and
1/8–1/4 that of air in intercellular spaces, which accelerates the corrosion of
metallic materials. Changes in the pH of body fluids are small because the
fluids are buffered solutions and the pH usually remains between 7.0 and
7.35 /11/. The change in the pH leads to the inflammation around the implant
/12/, producing highly reactive oxygen species /13/. This oxidizing species
(e.g. H2O2) will interact with the implant surface. The pH of the hard tissue
into which a material is implanted decreases to approximately 5.2 and then
recovers to 7.4 within two weeks /14/. The cell is also a kind of charging
body that may influence the corrosion of metallic materials. Metallic
materials themselves do not show any toxicity, but some dissolved metal
ions, corrosion products, and wear debris may show toxicity when they
combine with biomolecules and cells. Corrosion and electrochemical

7
Biomaterials Corrosion

properties of metallic materials in biological environments was reviewed by


Hanawa /15/.

Fig. 2: The effect of physiological environment on metallic implant

INTERACTIONS OF IMPLANT MATERIAL AND LIVING TISSUE -


HOST RESPONSE – INFLAMMATION

Fixation of an implant in the human body is a dynamic process that


remodels the interface zone between the implant and living tissue at all
dimensional levels from the molecular upto the cell and tissue morphology
level and at all time scales from the first year to several years after
implantation /16/. This is represented in Fig. 3, in which the logarithmic
length and time scales indicate this complex dynamic process. After
implantation, a space filled with biofluid exists next to the implant surface
and as the time progresses the proteins will be adsorbed (Fig. 4) at the oxide
surface (nm thickness) on the implant’s surface. This will in turn result in
osteoinduction which occurs by proliferation of cells, revascularisation and
eventual gap closing. Ideally, a strong bond will be formed between implant
and tissue. However, sometimes connective tissue is formed at the interface,
resulting in a fibrous tissue capsule that prevents osteointegration (see inset)
and will cause implant loosening /17/.
There are various adverse reactions that may occur at the implantation of
medical devices:

8
T.M. Sridhar and S. Rajeswari Corrosion Reviews

Fig. 3: Dynamic behaviour of the interface between implant (left) and bony
tissue (right)

Fig. 4: Schematic picture of cells close to material surface illustrating that the
cells with the dynamic hydration (water and ions) and protein layers,
which cover the material surface in the biological environment /17/

9
Biomaterials Corrosion

™ Adsorption of proteins
™ Allergic foreign body response and hypersensitivity
™ Coagulation and haemolysis
™ Cytotoxicity
™ Mutagenicity and carcinogenesis

Implant designs and biomaterials must be easily accepted by the body. All
implants cause some sort of response in the surrounding – no material can be
considered to be completely inert. Any surgical procedure results in a
disruption of blood supply and damage to tissue. Complications of medical
devices are largely based on both the effects of the implant on the host tissue
and the effects of the host on the implant. Placing a biomaterial in the in vivo
environment involves injection, insertion, or surgical implantation, all of
which injure the tissues or organs involved. This initiates a response by the
body and mechanisms are activated to maintain homeostasis and to heal the
wound. The degree to which this condition is created and resolved is a
measure of the host reaction to the biomaterial that may ultimately determine
its biocompatibility /18/.
Implantation may lead to acute inflammation which is of relatively short
duration, lasting from minutes to days depending on the extent of the injury.
The main characteristics are exudation of fluid and plasma proteins (edema)
I think this should be and the recruitment of white blood cells (leukocytes), such as ploymorho-
“polymorphonuclear”, not nuclear granulocytes (PMS) (more commonly called neutrophils), monocytes
as spelt here. and platelets. However motion at the implant site, extensive surgical injury,
Doreen bacterial infection or host factors such as poor blood supply or nutrition can
also contribute to this prolonged inflammatory state. Hence, the different
types of material variables that effect biomaterials are: bulk material
composition, microstructure, morphology, crystallinity and crystallography,
elastic constants, compliance, surface chemical composition, chemical
gradient, molecular mobility, surface topography and porosity, as well as
water content, hydrophobic–hydrophilic balance, surface energy, corrosion
parameters, ion release profile, metal ion toxicity, polymer degradation
profile, degradation product toxicity, leachables, catalysts, additives,
contaminants, ceramic dissolution profile, wear debris release profile, particle
size, sterility and endotoxins /19,20/.

10
T.M. Sridhar and S. Rajeswari Corrosion Reviews

BIOMATERIALS EMPLOYED IN ORTHOPEDICS

Many synthetic biomaterials used for implants are common materials


familiar to the average materials engineer or scientist. In general these
materials can be divided into the following categories: metals, polymers,
ceramics and composites. A short section on each of these groups is
presented below with emphasis on hard tissue materials.

Metals /21-25/
The high modulus and yield point coupled with the ductility of metals
make them suitable for bearing high loads. Metallic implants are therefore
mostly used to replace hard tissue. A number of authors have reviewed the
use of metals in implants
Metallic implants are used for two primary purposes, i). Implant devices
used as prostheses serve to replace a portion of the body and include devices
such as total joint replacements and skull plates, and ii) As fixation devices
that are used to stabilize broken bones and other tissues while normal
healing. Three of the most commonly used metals and alloys are; titanium,
stainless steel and cobalt-chromium. During the initial days, stainless steel
and cobalt-chrome alloys were preferred for bone replacement applications.
They were primarily used for their good mechanical properties. However, the
high mechanical strength of such metallic implants resulted in stress-
shielding and bone resorption due to the elastic modulus mismatch with the
surrounding bone. This drawback combined with findings such as corrosion
leading to a reduced mechanical strength and toxic by-products, directed the
attention to titanium and its alloys. Titanium and titanium alloys have the
advantage of possessing relatively lower modulus of elasticity and a higher
resistance against corrosion. The oxide layer of titanium has also been
proposed by some authors to have significant influence on the integration of
this metal with bone tissue.

Polymers /22-25/
Polymers are long-chain high molecular weight materials consisting of
repeating monomer units. Besides the chemical composition, other variables
such as molecular weight distribution and extent of cross-linking influence

11
Biomaterials Corrosion

their physiochemical properties. The physical properties of polymers have a


close resemblance to soft tissue and are hence useful in replacing skin,
tendon, cartilage and vessel walls and drug delivery /26/ etc. The enormous
variation in production possibilities has contributed to a huge number of
polymers available for implantation purposes, e.g. ultra high molecular
weight polyethylene (UHMWPE) is used for one of the articulating surface
components in joint prosthesis, and polycaprolactone has been used as
resorbable sutures or resorbable screws and plates for fracture fixation.
Polymeric materials also possess several advantages over metals. For
example, the isotonic saline solution that comprises the extracellular fluid has
little or no effect on polymeric components whereas it is extremely hostile
towards metals. Elastic modulus mismatch between the polymers and bone is
reduced, which decreases stress shielding and bone resorption. In addition
most polymers can be made degradable which means that through the
degradation process they are gradually replaced by the host tissue, thus
eliminating the need for additional surgery. This approach can however in
some cases lead to the deterioration of mechanical properties and in addition
adverse tissue reactions might occur due to released degradation products,
thus restricting the use of these materials.

Ceramics /22-27/
Ceramics used for the repair and reconstruction of diseased or damaged
parts of the body are known as bioceramics. Though a large number of
ceramics are known, only a few are suitably biocompatible. These ceramics
can be grouped according to their relative reactivity in physiological
environment (Fig. 5). They include a broad range of inorganic/non-metallic
compositions, which may be bioinert (alumina, zirconia), resorbable
(tricalcium phosphate), bioactive (hydroxyapatite, bioactive glasses and glass
ceramics), or porous for tissue in growth (hydroxyapatite coated metals,
alumina). Ceramics are stiff, hard and chemically stable and are often used in
situations where wear resistance is vital. Applications include replacements
for hips, knees, teeth, tendons, repair for periodontal disease, maxiofacial
reconstruction etc. Implants of bioceramic origin have in the last couple of
years played an increasingly important role. These materials provide an
interface of such biological compatibility with bone forming cells that these

12
T.M. Sridhar and S. Rajeswari Corrosion Reviews

lay down bone in direct apposition to the material. Their main drawbacks are
poor mechanical properties and strength.

Fig. 5: Relative reactivity of bioceramics /27/

Composites /13, 25/


Composite materials are a mixture of two or more phases bonded together
so that stress transfer occurs across the phase boundary. Consequently,
because stress is not transferred to voids, porous materials are not usually
considered composites even though they contain two phases, solid and void.
Typically, composite materials are designed to provide a combination of
properties that cannot be achieved with a single-phase material. It is clear,
when comparing human tissue with the various metals, polymers and
ceramics, that there is a considerable mismatch among their individual
properties. Hence, much attention has been paid to the development of
composites, thus trying to combine the good properties of different types of
materials while avoiding some of their drawbacks. An example of this is the
coating of a bioinert material such as alumina with a bioactive material such
as hydroxyapatite (HAP) or bioglass to promote direct bone attachment for
e.g. in hip replacements.
Each class of biomaterial has its own merits and applications. However,
they also suffer from certain drawbacks. The poor mechanical properties of

13
Biomaterials Corrosion

ceramics restrict their use in the bulk loading areas. The polymers, usually
being organic in nature, are found to be vulnerable to attack by the range of
physiological constituents. In this perspective, metallic materials have gained
considerable attention. The excellent mechanical properties, high corrosion
resistance in body fluids and biocompatibility of the metals make them
suitable for implantation purposes.

ALLOYS EMPLOYED IN PROSTHETIC APPLICATIONS

An implant material to be suitable for use in prostheses must have the


appropriate functional properties – mainly mechanical – for utilization and it
must be biocompatible. The applications of metallic implants in the modern
era are usually classified into three types: stainless steels, cobalt-chromium
alloy and titanium alloys /28-30/.

Stainless steels
Stainless steel is biocompatible and has been used for many decades as a
permanent surgical implant material. The type of stainless steel that is
normally used for implants is 316L. It achieves its biocompatibility by being
highly corrosion-resistant due to the formation of a thin protective chromium
oxide layer on its surface. The environment with which stainless steel must
deal within the body is, however, rather complex and, if corrosion occurs,
release of potentially harmful material could ensue. Due to the fact that
stronger and more corrosion-resistant materials are available, they are
suitable for permanent prosthetic devices. Stainless steels contain enough
chromium to confer corrosion resistance by virtue of passive chromium oxide
layer /31,32/.

Cobalt-chromium alloys
CoCr alloys, developed several decades ago for the aerospace industry,
also achieve their inertia through the formation of a chromium oxide surface
layer. They have excellent mechanical properties and are widely used in
orthopaedic implants. The alloys are generally CoCrMo or CoNiCrMo, and
may also include other elements such as tungsten or iron (Fe). Apart from the

14
T.M. Sridhar and S. Rajeswari Corrosion Reviews

fact that Ni can be avoided in the formulation, CoCr alloys have advantages
over stainless steel in terms of better corrosion resistance and better
mechanical properties for certain applications. Both wrought and cast CoCr
alloys are used in prosthetic devices, each version having distinct properties.
They are often used as components in modular prosthetic devices such as hip
or knee joints, being the most suitable for bearing surfaces (often against
ultra-high-molecular-weight polyethylene). They may also be used for joint
stems and in various other prosthetic devices. There is renewed interest in
metal-on-metal bearing surface for hip joints (in both total hip replacements
and surface replacements) since it appears that strict manufacturing
tolerances and the use of appropriate CoCr alloys can lead to very low wear
rates similar to ceramic-on-ceramic surfaces. This may offer the advantage of
avoiding the production of polyethylene wear particles associated with
polyethylene acetabular cup systems, these being implicated with tissue
reactions and eventual loosening of the hip stems.

Titanium and titanium alloys


In the second half of the 20th century, titanium began to find widespread
application in many industrial sectors. Its high inertia, due to the formation of
a thin surface titanium oxide layer, light weight and excellent
biocompatibility, seemed to make it the ideal material for surgical implants.
Its strength-to-weight ratio, moreover, is better than any other surgical
implant material. Both pure Ti and Ti6Al4V are widely used in prosthetic
devices; the choice depends on the functional requirements of material. An
additional advantage of these materials is their compatibility with imaging
techniques such as computed tomography (CT) scanning and magnetic
resonance imaging (MRI). Due to their poor tribological properties, Ti and Ti
alloys are not normally used for the bearing surfaces of articulating joints.
The good strength-to weight ratio, fatigue resistance and low modulus of
elasticity of Ti6Al4V make it one of the best alloys for implanting into bone
and it is often used for the bone stems of modular artificial joints, as well as
for other prosthetic devices. Recently, some concern has been expressed over
the use of Ti6Al4V since it appears that small amounts of both Al and V can
be dissolved out of the alloy, and both of these elements may have negative
effects on the human body. For this reason, related alloys where V is
substituted by niobium (Nb) or tantalum (Ta), and Al is substituted by

15
Biomaterials Corrosion

zirconium (Zr), are under investigation. Studies on substitution with other


alloying elements are also underway. It should be noted that the amounts of
Al and V released from Ti6Al4V may be so small that the alloy is perfectly
acceptable as an implant material. It is generally regarded as one of the most
biocompatible of all the available prosthetic alloys.
Other Ti alloys that are under consideration are so called β or near-β
phase alloys, these having advantages in terms of better formability,
toughness and lower modulus. The latter property is important for bone
implants and hence Ti alloys scores over SS/Co-Cr alloys. Stress shielding is
a well-known phenomenon caused by the large difference in elastic modulus
between bone and implant materials, which can lead to bone loss and
eventual implant failure. β or near-βalloys include binary TiMo alloys and
various, more complex systems with Nb, Zr, Fe, Ta, palladium and/or tin as
alloying components.

NICKEL-TITANIUM ALLOYS

In recent years, a novel material has stimulated considerable interest for


its potential biomedical applications. NiTi, or ‘nitinol’, is a so-called ‘smart
material’, with shape-memory and superelastic properties. It also has an
elastic modulus closer to that of bone than other metal implant materials.
These may be used to great advantage in orthodontics, in the treatment of
bone fractures and as bone suture anchors for attaching soft tissues such as
tendons and ligaments to bone. A further important application of NiTi is in
the use of stents employed for reinforcement of blood vessels. Due to their
shape memory property, they can be introduced into a patient through a thin
catheter before expanding into their functional form. Such minimally
invasive procedures are very much in line with the development of modern
surgical techniques. NiTi displays excellent biocompatibility due to the
formation of a thin surface titanium oxide. It does, however, contain Ni, and
its use in subjects with high Ni sensitivity is questionable, though most
studies have shown that NiTi is highly stable in the body with minimal Ni
release /30 /. However, the long-term stability and biocompatibility of this
material in the various environments encountered in the human body needs
further investigation. Recently, the related ternary system TiNiCu, with its
atomic percentage of copper, has attracted interest for biomedical and other

16
T.M. Sridhar and S. Rajeswari Corrosion Reviews

applications due to better fatigue and modified shape-memory properties.


While some studies appear to indicate that the corrosion resistance and
biocompatibility of this system are inferior to the binary material, further
work needs to be undertaken to assess this system.
The present review focuses on biological corrosion phenomena in alloys
used for orthopaedic implant applications and the contributions of researchers
in this regard in India.

CORROSION OF METALLIC IMPLANT MATERIALS

From an engineering standpoint, in situ degradation of metal-alloy


implants is undesirable: the degradation process may decrease the structural
integrity of the implant, and the release of degradation products may elicit an
adverse biological reaction in the host. Degradation may result from
electrochemical dissolution phenomena, wear, or a synergistic combination
of the two. Electrochemical processes may include generalized corrosion,
uniformly affecting the entire surface of the implant, and localized corrosion,
affecting either regions of the device that are shielded from the tissue fluids
(crevice corrosion) or seemingly random sites on the surface (pitting
corrosion). Electrochemical and mechanical processes (for example, stress
corrosion cracking, corrosion fatigue, and fretting corrosion) may interact,
causing premature structural failure and accelerated release of metal particles
and ions.
The clinical importances of degradation of metal implants is witnessed by
particulate corrosion and wear products in tissue surrounding the implant,
which may ultimately result in a cascade of events leading to periprosthetic
bone loss. Furthermore, many authors have reported increased concentrations
of local and systemic trace metal in association with metal implants /33-34/.
There is a low but finite prevalence of corrosion-related fracture of the
implant. A common occurrence with early implants was a chronic
inflammatory reaction due to grossly corroded ferrous alloys /35/. Mild
corrosion in many cases can also produce symptoms that necessitate the
removal of the implant. The symptoms range from a local tenderness at the
site of the corroded area to acute pain, reddening and swelling over the whole
general area around the device /36/. These symptoms are an indication that
the tissue is reacting to the implant. The corrosion of metallic implants can

17
Biomaterials Corrosion

affect the surrounding tissues in three ways /37/. Electrical currents may
affect the behaviour of cells; the corrosion process may alter the chemical
environment (pH, pO2); and the metallic ions may affect cellular metabolism.
Of the three, the last is usually the mostly severe. Changes in the surrounding
bone and fibrosis are often the result of implant corrosion products.
The process of corrosion can be described as metallurgy in reverse. When
most pure metals are placed in solution they tend to revert to soluble ionic
species, oxides, or hydroxides. X-ray analysis of the corrosion products of a
29-year-old low-alloy-steel bone plate indicates that Fe2O3 and β-FeOOH are
the principal corrosion products /38/.

GENERAL CONCEPT RELATED TO IMPLANT CORROSION

Corrosion, the gradual degradation of materials by electrochemical or


chemical attack, is a concern particularly when a metallic implant is placed in
the hostile electrolytic environment provided by the human body /39/. Even
though the freely corroding implant materials used in the past have been
replaced with modern corrosion-resistant superalloys, deleterious corrosion
processes have been observed in certain clinical settings.
Corrosion of biomaterials is a complex multifactorial phenomenon that
depends on geometric, metallurgical, mechanical, and solution-chemistry
parameters, and a firm understanding of these factors and their interactions is
required in order to comprehend how and why implant materials corrode.
Two essential features determine corrosion of metal at the implant site. The
first characteristic involves thermodynamic driving forces, which cause
corrosion (oxidation and reduction) reactions, and the second involves kinetic
barriers, which limit the rate of these reactions. The thermodynamic driving
forces that cause corrosion correspond to the energy required or released
during a reaction. The kinetic barriers to corrosion are related to factors that
impede or prevent corrosion reactions. The basic underlying reaction that
occurs during corrosion is the increase of the valence state that is, the loss of
electrons of the metal atom to form an ion, as expressed by the equation:

M ⇔ Mn+ + ne–.

This oxidation event may result in free ions in solution, which then can
migrate away from the metal surface or can lead to the formation of metal

18
T.M. Sridhar and S. Rajeswari Corrosion Reviews

oxides, metal chlorides, organometallic compounds, or other chemical


species. These latter forms may be soluble or may precipitate out to form
solid phases. The solid oxidation products may be subdivided into those that
form adherent compact oxide films and those that form non-adherent oxide,
chloride, phosphate, or other particles that can migrate away from the metal
surface. In all of these possible reactions there is a thermodynamic driving
force for the oxidation of metal atoms to their ionic form.
Most alloys used for orthopaedic appliances rely on the formation of a
passive film to prevent oxidation from taking place. These films consist of
metal oxides, which form spontaneously on the surface of the metal in such a
way that they prevent transport of metal ions and electrons across the film. In
order to limit oxidation, passive films must have certain characteristics. They
must be non-porous and must fully cover the metal surface; they must have
an atomic structure that limits the migration of ions and electrons across the
metal oxide/solution interface; and they must be able to remain on the surface
of these alloys even with mechanical stress or abrasion, which can be
expected in association with orthopaedic devices /40,41/.

Interactions with water


A surgically implanted surface initially comes in contact with an
environment that is dominated by water molecules. The adsorption of water
and hydrated ions is likely to be the first event that occurs. The initial
hydration layer, or Helmholtz layer, extends a few angstroms to nanometres
from the surface. Although the detailed structure will be dynamic, the local
structure of the monolayer or bilayer closest to the surface is likely to be very
different to that of liquid water. A strong interaction with the surface may
cause dissociation of the water molecules, resulting in bound hydroxyl
groups. Alternatively, water may chemisorb molecularly to the surface,
remaining undissociated but relatively strongly bound. On highly non-
wetting or hydrophobic surfaces, the interactions with the surface may be
very weak (equivalent or physisorption). At the molecular level, the
interactions may therefore be covalent, ionic or weak van der Waals’ forces.
The importance of the interaction with water has been addressed in several
discussions of biological interactions /41/.

19
Biomaterials Corrosion

VARIOUS TYPES OF IMPLANT CORROSION

Uniform attack
Uniform attack refers to the inevitable corrosion encountered in all metals
immersed in electrolytic solutions. Titanium-base alloys have lower overall
corrosion rates compared to stainless steel and cobalt chromium alloys /44/.

Galvanic corrosion
Dissolution of metals driven by macroscopic differences in
electrochemical potentials, usually as a result of dissimilar metals in
proximity, is termed as galvanic corrosion. Inappropriate use of metals, e.g.,
a stainless steel wire in contact with a cobalt or titanium-alloy femoral stem,
a cobalt-alloy femoral head in contact with a titanium-alloy femoral stem,
and a titanium-alloy screw in contact with a stainless steel plate may result in
galvanic corrosion /45, 46/. Compositional differences, either between parts
because of manufacture from different master ingots within the same
specification limits or because of deliberate mixing of metals, are the most
likely causes of such effects.

Fretting corrosion
Whenever two metal surfaces are in contact, micromotion of the surfaces
disrupts the passivation film and permits the area of contact to corrode
rapidly. This type of damage is called fretting corrosion /47/. The corrosion
occurring at contact areas between materials under load subjected to vibration
and slip tears out small particles of metal from the surfaces. Fretting may
occur in metal-on-metal joint prostheses, producing particles of metal from
0.1 to 1 μm in diameter. Repeated oscillatory motion is required, such as
when multicomponent implanted devices are placed in weight bearing limbs
or when the fixation achieved by a screw and plate construct is unstable.
Cohen /48/ subjected plate and screw assembles to cyclic stresses in saline
solutions and found that the corrosion occurred in the screw assemblies due
to disruption of the passivation layer. Similar assemblies, which were not
subjected to the cyclical stresses, did not show this marked effect. Weinstein
et al /49/ examined multiple component implants and found that 27 to 28
stainless steel implants demonstrated fretting corrosion.

20
T.M. Sridhar and S. Rajeswari Corrosion Reviews

Crevice corrosion
Crevice corrosion is undoubtedly the most prominent form of corrosion.
This is a form of local corrosion due to differences in oxygen tension or
concentration of electrolytes or changes in pH in a confined space, such as in
the crevices between a screw and a plate /50/. Crevice corrosion is commonly
associated with stainless steel multicomponent devices; it is often
accompanied by severe tissue reaction to the corrosion products, making
removal of the device necessary. The narrower and deeper is the crack, the
more likely crevice corrosion is to start /44/. The principal cause of crevice
corrosion is differential aeration of the stagnant solution /47/. The oxygen-
deficient regions within the crevice become anodic in relation to the material
as a whole, and corrosion proceeds more rapidly. The low oxygen tension in
wounds probably accelerates this effect in vivo. Retrieval studies have shown
that 16 to 35% of modular total hip implants demonstrated moderate-to-
severe corrosion in the conical head-neck taper connections /51/. Studies of
retrieved stainless steel multipart internal fixation devices show visible
corrosion at the junction between screw head and the plate in 50-75% of all
devices /44/. Other typical crevices are scratches on the surface of an implant,
the interface between bone and an implant, the cement-metal interface, and
any other sharp interface likely to be depleted of oxygen relative to another
oxygenated area.

Pitting corrosion
Another common form of corrosion that occurs with metallic implants is
pitting. It is a form of localised, symmetric corrosion in which pits form on
the metal surface. Metals are particularly susceptible to pitting in
environments containing chloride ions, as in tissues, and it is also enhanced
when the oxygen content of the solution is low. Pitting is probably associated
also with the stability of the passive film and with crevice corrosion. Once
the passive film is broken, the crevice, with its low oxygen content and the
presence of chloride ions greatly hinder the self-healing of the passivation
film. A high current density at the pit results, causing them to pinhole into the
metal surface while most of the surface remains unaffected.
Stainless steel is particularly predisposed to pitting corrosion due to
inclusions of dissimilar material trapped in the metal during a manufacturing
process. These impurities may initiate pitting corrosion in relation to a grain

21
Biomaterials Corrosion

boundary and thus can lead to component failure /52/. It can also be initiated
by scratches or handling damage. Pitting was frequently observed in older
stainless steel fracture fixation hardware, e.g., on the underside of screw
heads. It also occurs infrequently on the neck or the underside of the flange of
proximal femoral endoprostheses /44/. Sivakumar et al /54-56/ have
investigated failures of stainless steel orthopaedic implant devices. Their
diagnostic study described that the failure of the implant is typically due to
pit-induced fatigue corrosion. They reported that the pitting attack on the
prosthesis have been initiated owing to the low molybdenum content and
excess of sulphide inclusion. Fig. 6 shows that the edges were severely pitted
and most cracks were associated with pits. A typical crack origin from the pit
on the implant surface was exhibited and severe damage was also observed at
the proximal end of the prosthesis.

Fig. 6: Scanning electron micrograph view of the (a) crack origin (b) pit on
the failed implant surface.

Intergranular corrosion
Another form of corrosion sometimes encountered with implanted
metallic devices is intergranular corrosion /55/. This is a form of galvanic
corrosion due to impurities and inclusions in an alloy. Intergranular corrosion
is associated with the sensitization of austenitic stainless steels. Stainless
steels, if improperly heat-treated after fabrication, may corrode by this
mechanism owing to a relative depletion of chromium from the regions near
the grain boundaries. This phenomenon is called sensitization. Welding of

22
T.M. Sridhar and S. Rajeswari Corrosion Reviews

metals, which produce local melting and solidification, can also lead to a
variant of this process, called knife-edge attack.

Leaching
This form of corrosion results from chemical differences not within grain
boundaries but within the grains themselves /44/. Leaching occurs in alloy,
which contains more than one phase (multiphasic), e.g., 35% Ni containing
cobalt-base alloy, F582.

Stress - corrosion cracking


Stress-corrosion cracking is another corrosion-related cause of failure for
some implants. It is due to residual tensile stresses resulting from
deformation (bending) of an implant. It is a phenomenon in which a metal in
a certain environment, especially those rich in chlorides, is subjected to stress
and falls at a much lower level than usual as a result of corrosion /56/.
Stress-corrosion cracking is not frequent, but it has been reported.
Lisagor /58/ described a clinical case in which an intramedullary nail had
undergone sufficient loading to cause permanent deformation of the nail. This
resulted in high residual surface tensile stresses, making the nail sensitive to
stress corrosion. A straight fracture plate, when flexed, will experience a
tensile stress on its convex surface and a compressive stress on its concave
surface. This produces a difference in electrochemical potential, which
renders the convex surface anodic with respect to the rest of the plate.
Corrosion, as an acceleration of uniform attack, or perhaps secondary tensile
rupture of the passive film, will then attack the convex surface preferentially.
The same process will occur at stress risers in loaded devices, such as screw
holes in fracture fixation plates or kinks in cerclage wire. In this case, the
regions of higher stress in the immediate vicinity of the stress risers will
corrode at the expense of the surrounding less stressed material /44/.
Sivakumar et al /59/ have reported that failure of stainless steel
orthopaedic implant devices was a result of stress corrosion cracking (SCC)
due to propagation of cracks radiating from a pit. The crack was aggravated
by the high inclusion content and large grain size of the implant (Fig. 7).

23
Biomaterials Corrosion

Fig. 7: Scanning electron micrograph view of (a) the crack associated with
pits (b) the crack morphology was transgranular and branched.

SURVEY OF FAILED IMPLANT DEVICES

Orthopaedic implants have improved the quality of life for millions of


people over the last quarter of a century. Over 450,000 total knee and total
hip replacement operations were performed in the United States in 1994, with
an anticipated increase of 10% per year /60/. The typical useful life for a
replaced joint is between 10 and 15 years. The problems are due to loosening
of the implant because of its bioinertness and /or stress concentration related
to higher stiffness of the implant than the natural bone. It is important to
understand that the corrosion of orthopaedic biomaterials is not just an
exercise in physics and chemistry. There are real problems related to the
corrosion of implants, its failures and remedies have been reported by
Kamachi et al. /60/.

Corrosion at Modular Interfaces of Joint-Replacement


Components
A current problem related to orthopaedic alloys is corrosion at the taper
connections of modular joint replacement components. With the large and
growing number of total joint designs that include metal-on-metal conical
taper connections, the effects of crevices, stress, and motion assume
increasing importance. Retrieval studies have shown that severe corrosive
attack can occur in the crevices formed by these tapers in vivo /60, 62-64/.
This attack was observed in components consisting of a Ti- 6Al-4V-alloy

24
T.M. Sridhar and S. Rajeswari Corrosion Reviews

femoral stem and a cobalt-alloy femoral head as well as in those consisting of


a cobalt-alloy stem and a cobalt-alloy head. It has been postulated that this
corrosion process is the result of a combination of stress and motion at the
taper connection and the crevice geometry of the taper /65/. The stresses
resulting from the use of prostheses cause fracture and abrasion of the oxide
film and thus covering these passive metal surfaces. This in turn causes
changes in the metal surface potential, making it more negative, and in the
chemistry of the crevice solution as the oxides continuously fracture and
repassivate. Such changes may result in deaeration (loss of oxygen) of the
crevice solution and in a decrease of the pH in the crevice, as is expected in
crevice corrosive attack /51,66/. The ultimate result of this process is a loss of
the oxide film and its kinetic barrier effect and an increase in the rate of
corrosive attack in the taper region. Severe and extensive corrosive attack has
been seen primarily in cobalt-alloy systems with modular taper connections.
However, corrosive attack in titanium-alloy stems also has been reported
/67/. The corrosion processes in cobalt alloy consisted of intergranular
corrosion due to impurities and inclusions in an alloy /44/, etching, selective
dissolution of cobalt, and formation of chromium-rich particles including
oxides, oxychlorides, and phosphates. A key factor that may contribute to
relative motion (fretting) at modular connections and ultimately to abrasive
loss of the passivating oxide layer is angular mismatch between the taper on
the male aspect of the connection and the bore on the female aspect. These
studies point to the effect of combined stresses and motion and to the
electrochemical processes that occur at metal oxide-solution interfaces. The
mechanical integrity of the oxide films that form on these alloys is essential
for the long-term stability and survival of the implant.

Corrosion at Internal Fixation Devices


The development of corrosion-resistant, biocompatible metal alloys was
one of several essential factors in the evolution of internal fixation as a
treatment for closed fractures. The history of this process, until the precursors
of modern-day alloys were introduced, was described in detail by Venable
and Stuck /68/ and, more recently, by Peltier /69/. Two case reports in which
Lane plates were removed after more than fifty years provide additional
indication of the corrosion of these devices and the deposition of large
amounts of corrosion products in the surrounding tissue /70, 71/. Hundreds of
internal fixation devices and early joint-replacement prostheses were

25
Biomaterials Corrosion

examined in retrieval studies in the 1960s and 1970s. Many of the stainless
steel and cobalt-chromium-alloy devices that were used during this period
were prone to accelerated corrosion because of improper selection of
materials, faulty fabrication techniques, and use of mixed metals. These
deficiencies have largely been eliminated in modern implants through sound
metallurgical practice and fabrication processes. Although modern, single-
part devices used as permanent implants rarely show visible signs of
corrosion, Cook et al /72/ found some degree of interface crevice corrosion in
89 per cent of the plates and 88 per cent of the screws of 250 multiple-part
stainless steel internal fixation devices removed between 1977 and 1985.

Corrosion in Dentistry
Alloys that are used in dentistry are permanently exposed to changeable
conditions of the oral environment, which is practically ideal for corrosion
and chemical disintegration of often used materials /73/. Implant materials
must be wear resistant, chemically inert in many basic and acid food
components, and also in the oral fluids. If the materials used in dental
practice are not resistant to dissolution in the oral cavity, the developed
products are harmful for the tissue (ions, etc.) and can cause pathological
changes. The influence of such metal prosthesis reflects on the remaining
teeth, on the mucus of the oral cavity and even on distant organs. Other teeth
can be damaged when prosthesis creates conditions beneficial to caries and
parodontosis. For prosthetic practice, alloys of silver, gold, chromium, cobalt,
nickel, molybdenum, iron and carbon are mostly used in various
combinations. Dental alloys should have an optimal ratio of hardness and
ductility, and consequentially a high hardness value is not always desirable.
For example, alloy of iron and carbon, i.e. steel, is inappropriate for fixed
prostheses because its high value of hardness leads to abrasion of the natural
teeth in the opposite jaw. Co-Cr alloys, because of their high strength,
hardness, corrosion resistance and biocompatibility, have wide use for
various implants in dentistry and medicine. However, their drawbacks are
low ductility and possible cancerogenic influence. Namely, there is a
possibility that the corrosion products of Co-Cr alloys can cause health
problems inside the body. It is known that cobalt inhibits the absorption of
iron in the blood and causes anemia, while chromium species lead to
disturbances in the central nervous system. As that all metals corrode more or

26
T.M. Sridhar and S. Rajeswari Corrosion Reviews

less in the oral cavity environment, it follows that their corrosion is almost
impossible to prevent. The solution to this problem is the reduction of
corrosion by applying materials of better quality, such as titanium which does
not corrode noticeably and does not create harmful effects in the body. The
chemical properties of the oxide layer formed on the surface of titanium play
an important role in the biocompatibility of the titanium implants and
surrounding tissues / 74,75/.
Kumar and Sankaranarayanan /76 / have studied the corrosion behaviour
of Ti–15Mo alloy in 0.15 M NaCl solution containing varying concentrations
of fluoride ions (190, 570, 1140 and 9500 ppm), which is evaluated using
potentiodynamic polarization, electrochemical impedance spectroscopy (EIS)
and chronoamperometric/current–time transient (CTT) studies to ascertain its
suitability for dental implant applications. The study reveals that there is a
strong dependence of the corrosion resistance of Ti–15Mo alloy on the
concentration of fluoride ions in the electrolyte medium.
Three-dimensional printing (3DPTM) /77/ is a rapid part-fabrication
process that can produce complex parts with high precision. Hong et al / 78/
have designed, synthesized by 3DPTM, and characterized a new Ti-5Ag
(wt%) alloy. Silver nitrate was found to be an appropriate inorganic binder
for the Ti powder-based skeleton, and the optimum sintering parameters for
full densification were determined. The hardness of the Ti-5Ag alloy was
shown to be much higher than that of a pure titanium sample.
Potentiodynamic measurements, carried out in saline solution at body
temperature, showed that the Ti-5Ag alloy had good passivation behavior,
similar to that of pure titanium. It is concluded that the Ti- Ag system may be
suitable for fabrication of customized prostheses by 3DPTM.
Corrosion resistance of dental materials could be evaluated according to
distinct criterions. Thus a great deal of research has been performed on the
orthodontic materials and has provided warnings on the adverse effects of
corrosion.

CORROSION PREVENTION METHODS

a) Surface treatment:
When a synthetic material is placed within the human body, tissue reacts
towards the implant in a variety of ways depending on the type of the

27
Biomaterials Corrosion

materials. The mechanism of tissue interaction depends on the tissue


response to the implant surface. Under these conditions it is necessary to
search an alternative to rectify the defects arising out of the failure of metallic
implants. Surface modification of materials has received much attention as it
permits the independent optimization of bulk and surface properties. A
substrate material that possesses some combination of desirable bulk
properties e.g. strength, toughness, density but is lacking in some specific
surface property e.g. corrosion or wear resistance, may be altered by a surface
modification that implies the required properties.

1. Ion implantation
Implantation of ions helps to harden the surface and improve the
resistance to wear - accelerated corrosion phenomenon. Ion implantation, a
process that is widely used to modify the electronic properties of
semiconductor devices had become well established commercially by the
early to mid 1970s. After its successful applications in the semiconductor
industry, ion implantation process has been extended for the treatment of
biomaterials to improve their corrosion and wear resistance. The concept of
using ion-implantation as a surface modification technique in improving the
wear accelerated corrosion of orthopaedic implant materials was first
introduced by Buchanan et al /79/. Ion implantation involves the introduction
of a small, economical amount of the atoms of any element to the surface of
the material by means of high-velocity ions, without modifying the surface
finish or the bulk properties of the underlying material and independent of
thermodynamic constraints. In orthopaedic applications, titanium and its
alloys showed poor wear properties that impede the use of the alloy. Ion
implantation process has been shown to be extremely effective in enhancing
the wear performance of titanium surfaces. Sundararajan /80-85/ reported a
very significant reduction in the corrosive wear of titanium–based alloys as a
result of nitrogen ion implantation. Nitrogen ion implantation on titanium and
its alloys and also titanium modified 316L SS (at different doses ranging
form 1x1015 to 2.5x1017 ions/cm2) exhibited high corrosion resistance to
wear. The implanted specimens showed variations in the corrosion resistance
with varying doses and the specimen implanted at 1 x 1017 ions/cm2 showed
an optimum corrosion resistance. The detrimental effect of the specimen
implanted at the dose of 2.5 x 1017 ions/cm2 was attributed to the formation of
oxynitrides during implantation, which are present as islands in the passive

28
T.M. Sridhar and S. Rajeswari Corrosion Reviews

film. The stability of the passive film at higher potential was assessed by
potentiotransient technique after impressing a constant potential of 1.5 V for
three hours. This improvement arises from the formation of precipitates of
TiN and Ti2N, which screen underlying titanium atoms, avoiding their
migration and stabilizing the growth of the oxide film. The results of the
investigation indicated that nitrogen ion implantation can be used as a viable
method to improve the corrosion resistance of the orthopaedic implant
devices made of Ti6Al4V alloy.
Nitrogen ion implantation on Co-Cr-based alloys has proven to be
extremely effective in enhancing the corrosion resistance. Considerable work
has been done in the improvement of corrosion resistance of nitrogen ion
implanted materials namely Type 316L SS and Ti6Al4V alloy by
Veerabadran et al /86 / and Sundararajan et al /80- 85/. The enhanced
corrosion resistance was attributed to the protective oxynitride formation in
the passive film and this inturn widens the passive range. Studies were
undertaken to evaluate the corrosion resistance behaviour of type 316L SS
and advanced type 316L SS on surface modification by nitrogen ion
implantation in simulated body fluid conditions by electrochemical methods.
Nitrogen ion was implanted at different doses at fixed energy, and implanted
samples were subjected to electrochemical study to get the optimum dose that
can evince good corrosion resistance at simulated body fluid conditions.
Secondary Ion Mass Spectroscopy (SIMS) and X-ray Photoelectron
Spectroscopy (XPS) have been used for characterization of passive films of
implanted and unimplanted specimens to find out the elemental depth profile
and chemical state of the surface in order to understand the role of nitrogen in
improving the passivity of nitrogen ion implanted specimens. A two-fold
increase in pitting and crevice corrosion potentials was observed for the
nitrogen ion implanted specimen compared to the unimplanted type 316L SS.
This is attributed to the formation of protective passive film by the implanted
nitrogen /87,88/. The contribution made by the Indian groups led by Kamachi
Mudali and Subbaiyan in the development of nitrogen ion implanted
materials is commendable.
Rieu et al /89/ discussed in detail the implantations with different ion
species such as carbon, nitrogen, oxygen, boron, argon etc. on Ti-6Al-4V
knee, hip joints and showed the best resistance to corrosion and wear as a
function of nitrogen ion implantation in human body environment. Karim
Bordji et al /90/ have conducted tests to improve the wear and corrosion

29
Biomaterials Corrosion

resistance as well as the hardness of 316L SS implants. Three surface


treatments, such as (1) glow discharge nitrogen implantation, (2) carbon-
doped stainless steel sputtering and (3) low temperature plasma nitriding
were performed on 316L SS. Surface characterization according to the
different treatments showed that corrosion and wear resistance were strongly
improved, especially by ion implantation or carbon doped SS coating
sputtering. The effect of such treatments on the biocompatibility of 316L SS
was studied with human osteoblast and fibroblast cultures. Such surface
treatments may have relevance for increasing the lifetime of 316L SS
biomedical devices.

2. Passivation
The metals or alloys owe their good corrosion resistant properties to a
thin and corrosion resistant layer of reaction products, which is formed on the
exposed surface and drastically lowers the dissolution rate of metal ions. This
phenomenon is called passivation. The primary aim of the passivation is to
enhance the protective passive film by changing its composition, structure
and thickness, and or by reducing weak points such as non-metallic
inclusions. The mechanistic and electrochemical characterization have been
reported in detail by Kamachi Mudali and group /31,32,40/.
Several authors have studied the effect of alkali and acid treatment on
metals. Alkali treatment of titanium with subsequent heat treatment has been
adapted as an important pre-treatment procedure for hydroxyapatite
formation in orthopaedic applications. Raman et al /91/ have carried out
electrochemical impedance spectroscopic (EIS) studies have been employed
to analyse the electrochemical behaviour of titanium during the alkali
treatment. The open circuit potential and potentiodynamic polarisation
measurements were carried out in simulated body fluid (SBF) solution. An
optimum growth of the passive film was found to occur at the end of 17th
hour of treatment by alkali treatment. The alkali treated titanium immersed in
SBF solution for various durations exhibited the formation of a duplex layer
structure due to an inner barrier layer and an outer gel layer during the initial
periods of immersion. However, with increase in immersion time to 10 days,
a stable apatite layer was formed over the barrier layer which was confirmed
from the equivalent circuit fitted for the impedance data.
Miyazaki et al. /92/ have investigated the effect of thermal treatments on
mechanical properties and apatite-forming ability of the surface of the NaOH

30
T.M. Sridhar and S. Rajeswari Corrosion Reviews

treated tantalum metal. Stainless steel forms a chromium oxide, a process that
can be enhanced by chemical treatment with hot, concentrated nitric acid
(“passivation”), boiling in distilled water or by electrochemical method
(anodisation). Noh et al. /93/ have reported that both the enrichment of
chromium oxide on the surface and removal of MnS from the surface of
316L SS take place during nitric acid passivation. The examination of H2SO4
passivation of stainless steels by Peled et al. /94/ also ensures a passive film,
which resists pit initiation. Kannan et al /95, 96/ have studied the
electrochemical behaviour of hydroxyapatite coatings on both nitric acid and
sulphuric acid treated stainless steels, respectively. The results have indicated
that the HAP coatings on metal surfaces treated with acid delay the onset of
pitting and thus promote the tendency to resist the metallic corrosion. Thus
the literature survey regarding the surface treatment of metals ensures the
importance of surface modification.

3. Electropolishing and thermal methods


Various cost-effective and simple treatments like electropolishing and
thermal methods have had a profound effect on the corrosion properties of
metals. Eliaz and Nissan developed an electropolishing process effective for
complex metallic implants such as artificial heart valve frames and miniature
glaucoma implants /97/. Polishing in an ultrasonic bath and pulsed voltage
polishing processes were studied and compared to the standard ASTM
process. Current voltage curves were constructed for different solutions and
bath temperatures. The polished parts were evaluated by stereomicroscopy,
optical microscopy, atomic force microscopy, noncontact surface
profilometry, and X-ray diffraction. Pulse polishing was found useful in
eliminating the erosion effects of gas bubbles in solution. Electropolishing in
an ultrasonic bath was found useful when a rough, patterned surface is
needed, e.g. for osseointegration purposes. Preliminary animal studies
followed by histopathology indicated that the polished surfaces stimulated
only a moderate body reaction, as desired in such applications /97/.
Electrodeposited anodic oxide coatings produced on Ti–6Al–4V from
phosphoric acid electrolyte are reported /98/. Different coatings were
produced by varying the time periods. The coatings were exposed to
simulated body fluid for a period of 1 month and the weight loss is used to
calculate the corrosion rate. Electrochemical polarization and ac impedance
studies were performed on the coatings in simulated body fluid and indicated

31
Biomaterials Corrosion

that coatings produced from shorter times showed very good resistance to the
attack by this medium. Wasielewski and Lindblad /99/ suggested that the
tensile strength of cast Co-Cr-Mo-C alloy was found to be improved by
specific treatment in oxygen atmosphere and hot isostatic pressing. This
process may heal the micro voids that arise in castings during solidification.
316L SS and Ti-6Al-4V samples heated to the optimal temperature at 280°C
for 20 minutes and 3 hours, respectively resulted in the formation of oxide
layer. This oxide layer is reported to have the maximum osseointegration of
bone. Further it is speculated that it would be possible to create a condition
where bony ingrowth to various metal implants would be predetermined
according to specific demands by altering the heating temperature in various
gaseous environments. It was also proved that the electropolishing of the
implant specimens played a major role in enhancing the corrosion resistance
of metal specimens.

4. Bulk Alloying
New modified alloys of titanium are now available for implant
applications to overcome the toxicity of vanandium and aluminium. The
electrochemical behaviour of β titanium alloys, namely Ti–15Mo (TiMo) and
Ti–29Nb–13Ta–4.6Zr (TNTZ), were studied by Karthega et al /100/, The
OCP data for TNTZ alloy indicated a noble behaviour compared to TiMo
alloy. The current density value for TNTZ alloy calculated from polarization
measurement was found to be comparable to that of TiMo. The EIS spectra
obtained for TiMo alloy exhibited a single time constant for all potentials,
indicating a highly compact passive layer over the surface. The TNTZ alloy,
however, exhibited a single time constant at lower potentials and two time
constants at higher potentials, indicating a bilayer structure at higher
potentials.
Studies on the corrosion behaviour of Ti–6Al–7Nb and Ti–6Al–4V ELI
(extra low interstitial) investigated as a function of immersion hours in
simulated body fluid (SBF) condition, utilizing potentiodynamic polarisation
and electrochemical impedance spectroscopy (EIS) techniques was reported
/101/,. From the polarisation curves, very low current densities were
obtained for Ti–6Al–7Nb alloy compared to Ti–6Al–4V ELI, indicating a
formation of stable passive layer. Impedance spectra plots exhibited a two
time constant system suggesting the formation of two layers. Further, in vitro

32
T.M. Sridhar and S. Rajeswari Corrosion Reviews

and in vivo biocompatibility studies would decide the future applications of


these alloys.

5. Bioceramic coatings on implants


At present, for all those clinical applications where load-bearing
properties are required, most of the implants used are metallic, with
subsequent and serious problems. An alternative option is to coat the metallic
implants with bioceramics. This technique is being used nowadays, both for
dental implants and hip joint prosthesis. There is still a long way to go, but
several metallic implants with bioceramics coatings are commercially
available already, and the research in problem solving is underway. The
bioceramics coating process on a metallic substrate is quite complicated, and
several methods are available in this sense. A great deal of the clinical
success depends on this coating, since the quality and durability of the
interface attachment greatly depend on the purity, particle size, chemical
composition of the coating, layer thickness and surface morphology of the
substrate. An additional advantage of bioceramics coating is the reduction of
release of ions from the metal alloy. The bioceramics coating represents a
truly effective barrier that hinders the metallic ion kinetics of release towards
the living body. Hydroxyapatite [Ca10(PO4)6(OH)2] [HAP] is being
specifically used for this purpose, in order to improve the bone ingrowth and
to minimize the motion at the implant-bone junction. HAP possesses
excellent biological properties such as non-toxicity and lack of inflammatory
response and fibrous and/or immunitary reactions.
Hydroxyapatite, tricalcium phosphate (TCP) and their biphasic
combinations are important ceramic materials in the replacement of hard
tissues, because they can form a strong bond with the bone and favour bone
formation. However, the poor mechanical properties of calcium phosphates
limit the use of the bulk material to non-load bearing implants. For this
reason, one of most important uses of these calcium phosphates is to coat
inert or biotolerable implants with mechanical properties adequate for
orthopedic substitutions. In this way, the coated implants will not only have
the good mechanical properties of the substrate but also an enhanced
osseointegration and bioactivity due to the calcium phosphate layer /102,
103/. Surface modification of metals and alloys with HAP coatings plays a
dual role, both in preventing the release of metal ions and making the metal
surface bioactive. Sridhar et al /104-109/ have prepared the electrophoretic

33
Biomaterials Corrosion

HAP coatings on 316L SS substrates. EPD of HAP on metal substrates has


been used to achieve the uniform distribution of fine HAP deposits. The
advantages of this process include high purity of layers, ease of obtaining the
desired thickness, and high layer adhesion to the substrate. HAP coatings
obtained by EPD improve the corrosion resistance of implant materials, and
thus improve their biocompatibility. A substantial contribution is made by
Kamachi Mudali group on the electrochemical characterization of HAP
coated stainless steel implants. In vitro corrosion tests showed that coatings
have a tendency to resist pitting attack on the metal substrate. This indicates
that the HAP coatings offer better corrosion resistance. Generally, higher
values of impedance and polarization resistance and lower capacitance values
were obtained for the HAP-coated samples in comparison with the uncoated
ones. Immersion studies carried out for a period of 30 days indicated the
stable nature of the HAP coatings. These studies indicated that HAP coatings
obtained by EPD should be considered as a viable alternative for improving
the corrosion resistance of type 316L SS, thus enhancing the biocompatibility
of the medical device.
Fathi and co-workers /111/ have evaluated the significant effects on the
clinical success, bone tissue response and histopathological results of HAP
coated/uncoated metallic implants (316L SS and Ti) in animals. The results
showed that different substrates had pronounced effects on the
histopathological response to HAP coated on different implants with
beneficial corrosion resistance of the coatings.
Much interest in electrodeposition has evolved due to (1) the low
temperatures involved, which enable formation of highly crystalline deposits
with low solubility in body fluids and low residual stresses, (2) the ability to
coat porous, geometrically complex, or non-line-of-sight surfaces, (3) the
ability to control the thickness, composition, and microstructure of the
deposit, (4) the possible improvement of the substrate/coating bond strength,
and (5) the availability and low cost of equipment /112/.
Electrocrystallization of hydroxyapatite (HAP) on titanium was achieved
by cathodic polarization in solution containing calcium nitrate and
ammonium dihydrogen phosphate. The composition and pH of the bath were
found to significantly affect the nature and surface morphology of the
deposit. The effect of bath temperature was also studied. X-ray diffraction
tests and microscopic inspections confirmed the formation of well-

34
T.M. Sridhar and S. Rajeswari Corrosion Reviews

Fig 8: AFM images of the top surface of coatings produced at: (a) pH =
6.0, T = 90 °C, t = 5 min; (b) pH = 6.0, T = 85 °C, t = 2 h; and (c)
pH = 6.0, T = 90 °C, t = 2 h, in the presence of 0.10 M KCl. A 3D
view of the latter sample is presented in (d)./114/

crystallized HAP at pH0 = 6.0 at any temperature between 70 and 95 °C,


whereas, at pH0 = 4.2, less-crystallized, thicker, and more porous coatings
that contained traces of octacalcium phosphate were observed /113/. The
influence of potassium chloride and sodium nitrite on the composition and
surface morphology of the deposit was also evaluated. A speciation-
precipitation model was applied to better understand the effect of bath
conditions. The corrosion resistance of the coatings was determined by open-
circuit potential and cyclic potentiodynamic polarization measurements in a
simulated body fluid. The samples coated at pH = 6.0 exhibited nobler

35
Biomaterials Corrosion

behavior. The topography of different coatings was also evaluated by means


of ex situ AFM imaging. Figure 8 shows some typical AFM images. The
deflection image is more sensitive than topography images to delegate spatial
information, such as sharp edges. This suggests that synthetic HAP coatings
formed by electrocrystallization are more biomimetic, at least with respectto
their structure and morphology /113, 114/.
The ability to modify the chemistry and surface morphology of the
coating by fine control of bath composition, pH, and temperature makes
electrochemical deposition a versatile process for deposition of coatings on
implants, with a tailored body response /115, 116/. The ability to modify the
chemistry and surface morphology of the coating by fine control of bath
composition, pH and temperature makes electrochemical deposition a
versatile process for deposition of coatings on implants, with a tailored body
response.
Electrochemically deposited nano-grained calcium phosphate coatings
were produced on titanium alloy substrates using aqueous electrolyte
maintained at acidic pH by Narayanan et al. /117/. Different coatings were
produced by using different cathodic current densities and/or ultrasonic
agitation of the electrolytic bath. Ultrasonated bath produced coatings
containing dicalcium phosphate dehydrate and the grain sizes were in the
range of 50–100 nm. An electrochemical method of producing
nanocrystalline hydroxyapatite coatings on titanium surface is reported by
Narayanan et al. /118/. The bath contained Ca(NO3)2 and NH4H2PO4 in the
molar ratio 1.67:1. The electrolyte was maintained at physiological pH and
was ultrasonically agitated throughout the time of electrolysis. Coatings were
deposited for 30 min at 10 and 15 mA/cm2 and contained mono
hydroxyapatite phase whose crystal sizes were lower than 30 nm. These sizes
are comparable to the size of the bone hydroxyapatite crystals. Small
globules of hydroxyapatite covered the coating surface completely.
Ultrasonic agitation promoted the formation of nanocrystalline structure,
which will help in better attachment of bone tissues to the implant surface.
TiO2 coating offers excellent corrosion resistance and the bone mineral
hydroxyapatite (HA) offers very good biocompatibility. A suitable
combination of these coatings is expected to produce good corrosion
resistance as well as biocompatibility of an implant. Functionally graded
coatings containing TiO2 and HA were produced on Ti–6Al–4V /119/. It was

36
T.M. Sridhar and S. Rajeswari Corrosion Reviews

found that coating containing the oxide alone had the highest corrosion
resistance and showed better implant-bone bonding.
Plasma spray technique is currently the only method commercially
available for coating metallic substrates /115,120/. The significant
deficiencies found in the plasma sprayed HAP coatings have promoted the
search for new deposition methods, such as ion beam assisted deposition,
magnetron sputtering, sol-gel, pulsed laser deposition, laser sputtering, /121-
125/, etc. Although different deposition methods have been applied in the last
years, the sol-gel method offers a good alternative since the synthesis
temperatures are low and it can be applied to a great number of substrates,
including those which would oxidize at higher temperatures. Sol-gel
technology offers a chemically homogenous and pure product and has been
used for HAP production since 1988 /126/. Several authors have prepared
HAP via sol-gel technique using different precursors. Hijon et al /127/ have
deposited single-phase HAP coatings on Ti6Al4V by the sol-gel dipping
technique from aqueous solutions containing triethyl phosphite and calcium
nitrate. Balamurugan et al /128/ have reported that the coating thickness
alters the shear strength and corrosion resistance of sol-gel derived apatite
films of 316L SS.

b) Quality control
1. Improved standards and quality control: The manufacturer should adopt
the recommended metallurgical standards, fabricate the implants with
care, and maintain adequate testing facilities.
2. Improvements in design to minimise pits, crevices, large grain size,
inclusions and porosity /53/. Improved alloy ’cleanliness’, especially the
use of vacuum melting, and remelting, has largely eliminated pitting in
such hardware /44/.
3. The reduction of carbon to less than 0.03% has virtually eliminated the
risk of intergranular corrosion, which can occur when there is
precipitation of chromium carbide at the grain boundary in stainless steel
with a carbon content above this value /31,32/. Unfortunately, lowering
the carbon content results in lowering the ultimate tensile strength of
stainless steel /129/.
4. Proper heat treatment after welding will restore the appropriate
compositional distribution and prevent intergranular attack /44/.

37
Biomaterials Corrosion

5. Avoiding implantation of different types of metal in the same region – In


the manufacturing process, matched parts from the same batch of the
same variant of a given alloy must be provided. It must be ensured that
instruments are made from the same material as the implant /129/.

c) Research and development


1. Development of alloy with good wear resistance and ability to repassivate
at a high rate (to prevent fretting)
2. Coupling of two metals widely separated in the galvanic series, e.g.
Titanium and chromium, may not result in Galvanic corrosion but in
enhancement of protection. The risk, however, is a potential increase of
pitting or crevice corrosion.
3. Using an alloy whose open circuit or rest potential lies below the critical
potential for pitting /53/.

FUTURE DIRECTIONS

Biomaterials corrosion remains a serious clinical concern. Even though


the freely corroding implant materials used in the past have been replaced
with modern corrosion-resistant superalloys, deleterious corrosion processes
have been observed in certain clinical settings. There is reason to believe that
attention to variables related to metallurgical processing, tolerances of
modular connections, surface-processing modalities and appropriate selection
of materials could decrease the rate of corrosion and minimize the potential
for adverse clinical outcomes. The mechanical-electrochemical interactions
of passive metal-oxide surfaces must be investigated further. The stresses and
motion that are needed to fracture passivating oxide films as well as the
effects of repeated oxide abrasion on the electrochemical behaviour of the
interface and ultimately the implant are areas of active investigation. The role
of particulate corrosion products in adverse local tissue reactions also needs
to be investigated further. The clinical ramifications of increase in metal
content in body fluids and remote organs of patients who have a metal
implant need to be elucidated. Considerable work is required to discern the
chemical form of the metal, the nature of its ligands, and ultimately the

38
T.M. Sridhar and S. Rajeswari Corrosion Reviews

potential toxicity. Surface engineering, bioceramic and functionally graded


coatings are the promising techniques to battle corrosion of biomaterials.

ACKNOWLEDGEMENT:

The authors are very grateful to Mr. K. Prabakaran and Mrs. U.


Vijayalakshmi for their help in the preparation of this manuscript.

REFERENCES

1. D. Hill, Design Engineering of Biomaterials for Medical Devices,


John Wiley and Sons, Chichester, 1998.
2. M. Karlsson, “Nano-Porous Alumina, a Potential Bone Implant
Coating”, Ph.D. Thesis, Acta Universitatis Upsaliensis, Uppsala,
September 2004.
3. S. F. Hulbert, J.C. Bockros, L.L. Hench, J. Wilson, and G. Heimke,
High Tech Ceramics, Elsevier, Amsterdam, 1987, p.189.
4. N. Eliaz, “Biomaterials and Corrosion,” Chapter 12, in: Corrosion
Science and Technology: Mechanism, Mitigation and Monitoring, eds.
U. Kamachi Mudali and Baldev Raj, Narosa Publishing House, New
Delhi (2008), 356-397.
5. J. Black, Orthopedic Biomaterials Research and Practice, Churchill
Livingstone, 1988.
6. J. Black, Biological Performance of Materials, Dekkar, New York,
1980.
7. E. Doelker, Biopolymers, Springer Verlag, Berlin, 1993, p.200.
8. M. Long, and H.J. Rack, Biomaterials, 19 (1998), 1621-1639.
9. T. Hanawa, Science and Technology of Advanced Materials, 3 (2002),
289-295.
10. D. I. Bardos, “Stainless Steels in Medical Devices”, In: Hand Book of
Stainless Steels, McGraw-Hill, New York, 1977, p.1.
11. J. Black, Biological Performance of Materials, Plenum Press, New
York, 1984.
12. C.A. Van Blitterswijk, D. Bakker, S. S. Hesseling, and H. K. Koerten,
Biomaterials., 12 (1991), 187.

39
Biomaterials Corrosion

13. B. A. Freeman, and J. D. Crapo, Lab. Invest., 47 (1982), 412.


14. L.L. Hench, and E.C. Ethridge, Adv. Biomed. Eng., 5 (1975), 35-150.
15. T. Hanawa, Corros. Eng., 49 (2000), 687-699.
16. B. Kasemo, and J. Lausmaa, “The Biomaterial-Tissue Interface and its
Analogoues in Surface Science and Technology”, In: The Bone-
Biomaterial Interface, University of Toronto Press, 1991, p. 19-32
17. J.E. Sundgren, P. Bodo, and I. Lundstorm, J. Colloid Interface Sci.,
110 (1986)
18. J.M. Anderson, Biomaterials Science: An Introduction to Materials in
Medicine, Academic Press, San Diego, 1996, p.165-173.
19. R.J. Racicot, C.D. Crouch and M.E. Rauch, Corros.
Rev. 25 (2007) 97-106.
20. D.M. Vieira and F.P. de Franca, Corros. Rev. 26 (2008) 73-85.
21. L. Thair, U. Kamachi Mudali, S. Rajagopalan, R. Asokamani, and
Baldev Raj, Corr. Sci., 45 (2003) 1951-1967.
22. F.W. Cooke, J.E. Lemons, and B.D. Ratner, Biomaterials Science: An
Introduction to Materials in Science, Academic Press, San Diego,
1996, p. 11-35.
23. S. Virtanen, Corros. Rev. 26 (2008) 147-171.
24. A. Zielinski and S. Sobieszczyk, Corros. Rev., 26 (2008) 1-22.
25. H. Alexander, J.B. Brunski, S.L. Cooper, L.L. Hench, R.W.
Hergenrother, A.S. Hoffman, J. Kohn, R. Langer, N.A. Peppas, B.D.
Ratner, S.W. Shalaby, S.A. Visser, and I.V. Yannas, Classes of
Materials Used in Medicine, in Biomaterials Science: An Introduction
to Materials in Science, Academic Press, San Diego, 1996.
26. B. Sun, Balu Ranganthan and Si-Shen Feng, Biomaterials, 29 (2008)
475 – 486.
27. L.L. Hench, J. Am. Ceram. Soc., 74 (1991), 1487-1510.
28. P. Li, C. Ohtsuki, T. Kokubo, K. Nakanishi, N. Soga, and K. de Groot,
J. Biomed. Mater. Res., 28 (1994) 7-15.
29. D.F. Williams, Titanium for Medical Applications, in Titanium in
Medicine, Springer, Berlin, 2001, p. 561-585.
30. N. Gibson and H. Stamm, The Use of Alloys in Prosthesis, In:
Business Briefing-Medical Device Manufacturing and Technology,
2002, p. 48-51.
31. U. Kamachi Mudali, R.K. Dayal T.P.S. Gill and J.B. Gnanamoorthy,
Werks und korro., 37 (1986) 637-643

40
T.M. Sridhar and S. Rajeswari Corrosion Reviews

32. U. Kamachi Mudali, S. Ningshen and R.K. Dayal, Bull. Electrochem.


15 (1999) 74-78.
33. A. Bartolozzi, and J. Black, Biomaterials., 6 (1985), 2-8.
34. L.D. Dorr, R. Bloebaum, J. Emmanual, and R. Meldrum, Clin.
Orthop., 261 (1990), 82-95.
35. J.T. Scales, G.D. Winter, and H.T. Shirley, J. Bone Jt. Surg., Br. 41B
(1959), 810-820.
36. J.H. Hicks, and W.H. Cater, J. Bone Jt. Surg., Bone, 44B (1962), 122-
128.
37. D.C. Mears, J. Biomed. Mater. Res., 6 (1975), 138-148.
38. J.R. Cahoon, J. Biomed. Mater. Res., 7 (1973), 375-383.
39. A. Choubey, B.Basu, and R. Balasubramaniam, Intermetallics., 12
(2004) 679- 682.
40. U. Kamachi Mudali, R.K Dayal, J.B. Gnanamoorthy and P.
Rodriguez, Iron and Steel Inst. of Japan Inter. 36 (1986) 799-806
41. U. Kamachi Mudali and Y. Katada, Electrochim Acta 46 (2001) 3735-
3742.
42. F.H. Jones, Surface Science Reports., 42 (2001), 75-205.
43. B. Raton, Handbook of Chemistry and Physics, CRC Press, Florida,
1982, D133-D135.
44. J. Black, Corrosion and Degradation, In: Orthopaedic Biomaterials in
Research and Practice, Churchchill Livingstone, New York, 1988, p.
235-266.
45. C.D. Griffin, R.A. Buchanan, and J.E. Lemons, J. Biomed. Mater.
Res., 17 (1983), 489-500.
46. F.P. Bowden, J.B.P. Williamson, and P.G. Laing, J. Bone Joint, 37B
(1955), 676-690.
47. M.G. Fontana, and N.D. Green, Corrosion Engineering, McGraw-
Hill, New York, 1967.
48. J. Cohen, J. Bone Joint Surg., 44A (1962), 302-316.
49. A. Weinstein, E. Horowitz, and A.W. Ruff, Retrieval and Analysis of
Orthopaedic Implants, NBS Spec. Publ., U.S., 1973, p. 472.
50. C.O. Bechtol, A.B. Ferguson, and P.G. Laing, Metals and Engineering
in Bone and Joint Surgery, Williams and Wilkins, Baltimore,
Maryland, 1959.
51. T.L. Gilbert, C.A. Buckley, J.J. Jacobs, J. Biomed. Mater. Res., 27
(1993) 1533-1544.

41
Biomaterials Corrosion

52. U. Kamachi Mudali, R.K. Dayal T.P.S. Gill and J.B. Gnanamoorthy,
Corrosion, 37 (1990) 454-460
53. J.B. Park and R.S. Lakes, Metallic Implant Materials, In: Biomaterials
– an Introduction, 1992, p. 75-115.
54. M. Sivakumar, U. Kamachi Mudali, and S. Rajeswari, J. Mater. Sci.
Lett., 14 (1995), 148-151.
55. M. Sivakumar, U. Kamachi Mudali, and S. Rajeswari, J Mater. Eng.
Perform. 3 (1994) 744-753.
56. M. Sivakumar, U. Kamachi Mudali, and S. Rajeswari,. J Mater. Eng.
Perform. 3 (1994) 111-114
57. A.N. Hughes, B.A. Jordan, J. Biomed. Mater. Res., 6 (1972), 33-48.
58. B. Lisagor, Corrosion and Fatigue of Surgical Implants, Stan. News,
ASTM 3, 1975, p. 20-24, 43, 44.
59. M. Sivakumar, and S. Rajeswari, J. Mater. Sci. Lett., 11 (1992), 1039-
1042.
60. U. Kamachi Mudali, T.M. Sridhar, N. Eliaz and Baldev Raj, Corros.
Rev. 21 (2003) 231-267.
61. National Center for Health Statistics, Report (1994).
62. J.P. Collier, V.A. Surprenant, R.E. Jensen, and M.B. Mayor, Clin.
Orthop., 271 (1991), 305-312.
63. J.P. Collier, V.A. Surprenant, R.E. Jensen, M.B. Mayor, and H.P.
Surprenant, J. Bone and Joint Surg., 74B (4) (1992), 511-517.
64. E. B. Mathlesen, J.U. Lindgren, G.G.A. Blomgren, and F.P. Reinholt,
J. Bone and Joint Surg., 74B (1991) 569-575.
65. J.L. Gilbert, and J.J. Jacobs, The Mechanical and Electrochemical
Processes Associated with Taper Fretting Crevice Corrosion: a
review, In: Modularity of Orthopedic Implants, American Society for
Testing and Materials Special Technical Publication 1301, American
Society for Testing and Materials, West Conshohocken, Pennsylvania,
1997, p. 45-59.
66. J.L. Gilbert, and C.A. Buckley, Mechanical-electrochemical
Interactions during in vitro fretting corrosion tests of Modular Taper
Connections, In: Total Hip Revision Surgery, Raven Press, New York,
1995, p. 41-50.
67. R. Michel, J. Hofmann, F. Loer, and J. Zilkens, Arch. Orthop. and
Trauma Surg., 103 (1984), 85-95.

42
T.M. Sridhar and S. Rajeswari Corrosion Reviews

68. C.S. Venable, and W.G. Stuck, The Internal Fixation of Fractures,
Springfield, Illinots, 1947.
69. L.E. Peltier, Fractures: A History and Iconography of Their
Treatment, Norman Publishing, San Francisco, 1998, p.114-167.
70. L.A. Danzig, S.L.Y. Woo, W.H. Akeson, G.F. Jemmott, and M.G.
Wickham, Clin. Orthop., 149 (1980), 201-206.
71. R.R. Tarr, R. Jorge, L.L. Latta, and L. Ghandur-Mnaymneh, J.
Biomed. Mater. Res., 17 (1983), 785-792.
72. S.D. Cook, K.A. Thomson, A.F. Harding, C.L. Collins, R.J., Jr.
Haddad, M. Milicic, and W.L. Fischer, Biomaterials., 7 (1987), 177-
184.
73. K. Suresh Kumar Danadurai, T.M. Sridhar, T.K. Arumugam, S.
Rajeswari and M. Subbaiyan, Electrochemical characterisation of
orthodontic materials in the oral environment, In Biomedical
Materials and Devices - New Frontiers, Ed. M. Jayabalan, SCTIMST,
Trivandrum, India (1998) 57-60.
74. M. Karthega, S. Tamilselvi and N. Rajendran, Trends Biomater Artif.
Organs, 20 (2006) 31-34
75. S. Tamilselvi and N. Rajendran, Trends Biomater Artif. Organs,
20(2006) 49-52
76. S. Kumar and T S.N. Sankarnarayanan, J. Dentistry, 36 (2008) 500 –
507.
77. S.-B. Hong, N. Eliaz, E.M. Sachs, S.M. Allen and R.M. Latanision,
Corros. Sci., 43 (2001) 1781-1791.
78. S.-B. Hong, N. Eliaz, G.G. Leisk, E.M. Sachs, R.M. Latanision and
S.M. Allen, J. Dental Res. 80 (2001) 860-863.
79. R.A. Buchanan, R.K. Bacon, J.M. Williams, J. Biomed. Mater. Res.,
21 (1987), 155-166.
80. T. Sundararajan, and U. Kamachi Mudali, K.G.M. Nair, S. Rajeswari
and M. Subbaiyan, Materials and Corrosion., 50 (1999), 344-349.
81. T. Sundararajan, and U. Kamachi Mudali, K.G.M. Nair, S. Rajeswari
and M. Subbaiyan, J. Mater. Eng. Perf., 8 (2) (1999), 252-260.
82. T. Sundararajan, U. Kamachi Mudali, K.G.M. Nair, S. Rajeswari and
M. Subbaiyan, Anti. Corr Methods and Mater., 45(3) (1998) 162-166.
83. T. Sundararajan, and U. Kamachi Mudali, K.G.M. Nair, S. Rajeswari
and M. Subbaiyan, Pro. Discus. Meet. on Surf. Sci. Eng. (SURE 96),
(1996) IIM Kalpakkam Chapter, 234-242.

43
Biomaterials Corrosion

84. T. Sundararajan, and U. Kamachi Mudali, K.G.M. Nair, S. Rajeswari


and M. Subbaiyan, Mater. Trans. JIM 39 (1998) 759-764
85. T. Sundararajan, and U. Kamachi Mudali, K.G.M. Nair, S. Rajeswari
and M. Subbaiyan, Trans. IIM, 52 (2000) 413-421
86. K.M. Veerabadran, U. Kamachi Mudali, K.G.M. Nair and M.
Subbaiyan, Mater. Sci. Forum., 318-320 (1999) 561-568.
87. K.M. Veerabadran, U. Kamachi Mudali, K.G.M. Nair and M.
Subbaiyan, Surface Modification of Surgical Grade Stainless Steel
Orthopaedic Implants by Ion Implantation and Investigation of
Localised Corrosion. Pro. Discus. Meet. on Surf. Sci. Eng. (SURE 96),
(1996) IIM Kalpakkam Chapter, 226-233.
88. K.M. Veerabadran, Ph.D Thesis, 1999 University of Madras
89. J. Rieu, A. Pichat, L.M. Rabbe, and M. Robelet, Mater. Sci. Forum.,
102-104 (1992), 505.
90. Karim Bordji, Jeans-Yves Jouzeau, Didier Mainard, Elisabeth Payan,
Jeans-Pierre Delagoutte and Patrick Netter, Biomaterials., 17 (1996)
491-500.
91. V. Raman, S. Tamilselvi and N. Rajendran, Electrochim Acta 52
(2007) 7418–7424.
92. T. Miyazaki, H.M. Kim, T. Kokubo, F. Miyaji, H. Kato, and T.
Nakamura, J. Mater. Sci. Mater. Med., 12 (2001), 683-687.
93. J.S. Noh, N.J. Laycock, W. Gao, and D.B. Wells, Corros. Sci., 42
(2000), 2069.
94. P. Peled, and D. Itzhak, Corros. Sci., 32 (1991), 83.
95. S. Kannan, A. Balamurugan, and S. Rajeswari, Electrochim. Acta, 50
(2005), 2065.
96. S. Kannan, A. Balamurugan, and S. Rajeswari, Mater. Lett., 57
(2003), 2382.
97. N. Eliaz and O. Nissan, J. Biomed. Mater. Res. A 83 (2007) 546–557.
98. R. Narayanan and S.K. Seshadri, Corros. Sci. 49 (2007) 542-558.
99. G.E. Waisielewski, and N.R. Lindblad, Proc. of Second Inter. Conf.
On Super Alloys Processing, AIME, 1972.
100. M. Karthega, V. Raman and N. Rajendran, Acta Biomaterialia 3
(2007) 1019–1023
101. S. Tamilselvi, V. Raman and N. Rajendran, Electrochim Acta 52
(2006) 839–846

44
T.M. Sridhar and S. Rajeswari Corrosion Reviews

102. N. Eliaz and U. Kamachi Mudali (Guest Editors), Biomaterials


Corrosion (Special Issue), Corros. Rev. 21 (2003).
103. M. Vallet-Regi, and J.M. Gonzalez-Calbet, Progress in Solid State
Chemistry, 32 (2004), 1-31.
104. T.M. Sridhar, S. Rajeswari and M. Subbaiyan, Bull. Electro.Chem., 15
(3-4) (1999) 139.
105. T.M. Sridhar, T.K. Arumugam, S. Rajeswari and M. Subbaiyan, J.
Mater. Sci. Lett., 16 (1997) 1964.
106. T.M. Sridhar, U. Kamachi Mudali, M. Subbaiyan and S. Rajeswari, 7th
Intl. Symp. on Electrochemical Meth in Corros. Research, EMCR,
BUDAPEST, Hungary, May –June 2000.
107. T.M. Sridhar, N. Eliaz, U. Kamachi Mudali and Baldev Raj, Corros.
Rev. 20 (2002) 255-293.
108. T.M. Sridhar, U. Kamachi Mudali and M. Subbaiyan, Corr Sci, 45
(2003) 237-252.
109. T.M. Sridhar, U. Kamachi Mudali and M. Subbaiyan Corr. Sci., 45
(2003) 2337-2359.
110. U. Kamachi Mudali, T.M. Sridhar and Baldev Raj, Sādhanā, 28, Parts
3&4 (2003), 601-637.
111. M.H. Fathi, M. Salehi, M. Saatchi, V. Mortazavi, S.B. Moosavi,
Dental Mater., 19 (2003), 188-198.
112. N. Eliaz, T.M. Sridhar, U. Kamachi Mudali and Baldev Raj, Surf.
Eng. 21 (2005) 238-242.
113. N. Eliaz and T.M. Sridhar, Cryst. Growth & Des. 8 (2008) 3965-3977.
114. N. Eliaz and M. Eliyahu, J. Biomed. Mater. Res. A 80 (2007) 621-634.
115. H. Wang, N. Eliaz, Z. Xiang, H.-P. Hsu, M. Spector and L.W. Hobbs,
Biomaterials 27 (2006) 4192-4203.
116. N. Eliaz, W. Kopelovitch, L. Burstein, E. Kobayashi and T. Hanawa,
J. Biomed. Mater. Res. A, DOI 10.1002/jbm.a.32129
117. R . Narayanan , S . Seshadri , T . Kwon , K . Kim, Scripta Materialia,
56, (2007) 229 - 232
118. R. Narayanan, Tae-Yub Kwon and Kyo-Han Kim, Mater. Sci. and
Eng.:C (DOI:10.1016/j.msec.2007.11.009 )
119. R. Narayanan and S.K. Seshadri, Mater. Chem. Phy., 106 (2007) 406-
411.
120. K. de Groot, R. Gessink, C.P.A.T. Klein, and P. Serekain, J. Biomed.
Mater. Res., 21 (1987), 1375.

45
Biomaterials Corrosion

121. M. Geetha, U. Kamachi Mudali, N.D. Pandey R. Asokamani and B.


Raj, Surf. Eng., 20, (2004) 68-74.
122. M. Geetha, U. Kamachi Mudali, N.D. Pandey, R. Asokamani and B.
Raj, In Vitro Corrosion Behaviour of Laser Nitrided Ti-13Nb-13Zr
Alloy. Proc. of the First Asian Pacific Conference and 6th National
Convention on Corrosion on CD, Nov 28-30, 2001
123. L. Thair, U. Kamachi Mudali, S. Rajagopalan, K.G. M. Nair, R.
Asokamani and Baldev Raj, Role of Alloying Elements on the Passive
Films of Nitrogen Ion Implanted Ti-6Al 4V and Ti-6Al-7Nb Alloys.
Proc. of the First Asian-Pacific Conference and 6th National
Convention on Corrosion on CD, Nov 28-30, 2001
124. L. Thair, U. Kamachi Mudali, S. Rajagopalan, K.G. M. Nair, R.
Asokamani and Baldev Raj, Corros. Sci. 2002, 2439-57.
125. L. Thair, U. Kamachi Mudali, S. Rajagopalan, K.G. M. Nair, R.
Asokamani and Baldev Raj, Corros. Sci. 2003, 1951-67.
126. Y. Masuda, K. Matubara, and S. Sakka, J. Ceram. Soc. Japan, 98
(1990), 84-95.
127. N. Hijon, M.V. Cabanas, I. Izquierdo-Barba, and M. Vallet-Regi, Key
Eng. Mater., 254-256 (2004), 363.
128. A. Balamurugan, G. Balossier, S. Kannan, and S. Rajeswari,
Materials Letters., 60 (2006) 2288-2293.
129. J.R. Atkinson, and B. Jobbins, Properties of Engineering Materials for
Use in the Body, In: Introduction to the Biomechanics of Joints and
Joint Replacement, Mechanical Engineering Publications Ltd.,
London, 1981.

46

View publication stats

S-ar putea să vă placă și