Sunteți pe pagina 1din 19

Biomaterials 19 (1998) 1621 — 1639

Review
Titanium alloys in total joint replacement—a materials
science perspective
Marc Long, H.J. Rack*
School of Chemical and Materials Engineering, Clemson University, Clemson, SC 29634, USA
Received 26 May 1996; accepted 6 July 1997

Abstract

Increased use of titanium alloys as biomaterials is occurring due to their lower modulus, superior biocompatibility and enhanced
corrosion resistance when compared to more conventional stainless steels and cobalt-based alloys. These attractive properties were
a driving force for the early introduction of a (cpTi) and a#b (Ti—6Al—4V) alloys as well as for the more recent development of new
Ti-alloy compositions and orthopaedic metastable b titanium alloys. The later possess enhanced biocompatibility, reduced elastic
modulus, and superior strain-controlled and notch fatigue resistance. However, the poor shear strength and wear resistance of
titanium alloys have nevertheless limited their biomedical use. Although the wear resistance of b-Ti alloys has shown some
improvement when compared to a#b alloys, the ultimate utility of orthopaedic titanium alloys as wear components will require
a more complete fundamental understanding of the wear mechanisms involved. This review examines current information on the
physical and mechanical characteristics of titanium alloys used in artifical joint replacement prostheses, with a special focus on those
issues associated with the long-term prosthetic requirements, e.g., fatigue and wear. ( 1998 Published by Elsevier Science Ltd.
All rights reserved

Keywords: Titanium; Titanium alloys; Total joint replacement; Orthopaedics; Fatigue; Wear; Biocompatibility

1. Introduction degenerative joint disease [5]. Premature joint degener-


ation may arise from deficiencies in joint biomaterial
Natural synovial joints, e.g., hip, knee or shoulder properties, from excessive loading conditions, or from
joints, are complex and delicate structures capable of failure of normal repair processes, the explicit degen-
functioning under critical conditions. Their performance erative processes not yet being completely understood.
is due to the optimized combination of articular carti- Degeneration of weight bearing joints often requires
lage, a load-bearing connective tissue covering the bones surgery to relieve pain and increase mobility. Through
involved in the joint, and synovial fluid, a nutrient fluid minimum invasive damage, arthroscopic surgery, most
secreted within the joint area [1, 2] (Fig. 1). Unfortunate- frequently performed on knee joints, provides an efficient
ly, human joints are prone to degenerative and inflam- surgical method for diagnosis and symptomatic relief of
matory diseases that result in pain and joint stiffness. painful joints. Ultimately replacement of diseased joint
Primary or secondary osteoarthritis (osteoarthrosis), and surfaces by metal, plastic, or ceramic artificial materials is
to a lesser extent rheumatoid arthritis (inflammation of accomplished through arthroplastic surgery when the
the synovial membrane) and chondromalacia (softening natural joint can no longer adequately perform.
of cartilage), are, apart from normal ageing of articular Total joint replacement (TJR) arthroplasty is recog-
cartilage, the most common degenerative processes af- nized as a major achievement in orthopaedic surgery.
fecting synovial joints [3, 4]. In fact, 90% of the popula- Successful replacement of the natural joints through ar-
tion over the age of 40 suffers from some degree of throplastic surgery has been the long-time objective of
orthopaedic surgeons. Arthroplasty (Dorland’s Medical
Dictionary definition ‘plastic repair of a joint’) is a surgi-
cal technique which replaces all articulating degenerated
* Corresponding author. natural surfaces with artificial materials, hence achieving

0142-9612/98/$ — See front matter ( 1998 Published by Elsevier Science Ltd. All rights reserved.
PII S 0 1 4 2 - 9 6 1 2 ( 9 7 ) 0 0 1 4 6 - 4
1622 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

implantation success rate to'90%. It was further recog-


nized that the use of identical metals in the tribological
pair, though necessary to avoid galvanic corrosion, was
not an optimized tribology design. A high rate of loosen-
ing was encountered with early metal-on-metal artificial
joints due to non-optimum fit between the articulating
surfaces which produced high frictional moments and
excessive wear of the bearing surfaces [8—10]. These early
concerns limited the application of metal-on-metal ar-
ticulating devices, although follow-up examinations of
metal-on-metal hip protheses have shown very low wear
rates (a few lm per year per component) for protheses
implanted for up to 20 years [11].
Sir John Charnley in the 1960s developed the concept
of low-friction arthroplasty by introducing a new design
consisting of a small-diameter metallic femoral head ar-
ticulating with a polymeric (originally PTFE to be later
replaced by ultra-high-molecular-weight polyethylene
Fig. 1. Representation of human knee and hip joints during loading (UHMWPE)) acetabular cup [12, 13]. The initial success
[44].
of UHMWPE as the cup material [14] has prevailed for
30 yrs, UHMWPE being the dominant orthopaedic ma-
terial in TJRs [6]. Wear of UHMWPE has, however,
relief of pain and improved joint mobility by creation of been invariably observed when rubbing against metal
a new prosthetic joint. From early excision through inter- femoral heads or femoral condylar components of TJRs
position to replacement arthroplasty, great progress has [15].
been achieved over 170 years of orthopaedic surgery and Beside clinical factors and design considerations, the
joint prostheses are now being considered for many latter to minimize contact stresses, choice of counterpart
joints in the human body [6]. Total hip (THR) and total material has been shown to be a critical factor in
knee (TKR) joint replacements, due to population needs UHMWPE wear behavior. Ti—6Al—4V has generally
and their complex behavior, have nevertheless been the been found to have a more detrimental impact on
principal focus of artificial joint studies. TJR is now UHMWPE wear than Co—Cr—Mo alloys [14, 15]. In
a fairly well established orthopaedic technique involving order to achieve minimum wear and maximum success
the replacement of a growing number of hip and knee rate, subsequent studies have considered alternative ma-
joints, 275 000 hip and knee joints having been replaced terials combinations [16—23], Table 1, e.g., ceramic/
during 1995 in the United States [7]. Unfortunately UHMWPE prostheses where the ceramic component
in vivo degradation, primarily as a result of the higher creates minimum damage to the UHMWPE counterpart
wear rates associated with artificial implant materials, when compared to Co—Cr—Mo or Ti—6Al—4V. Wear of
and the consequent adverse biological effect of the gener- joint prostheses materials unavoidably represents a long-
ated wear debris on bone mass/density and implant fix- term limitation to the lifetime of a total joint replacement
ation, typically, however, results in a shorter lifetime for as accumulation of UHMWPE, and to a lesser extent
these artificial joints when compared with natural metal or ceramic wear debris has been associated with
synovial joints. Further, when compared to the initial incidence of non-specific pain and prosthesis loosening.
TJR surgery, revision surgery of an implant is more The former is a result of adverse tissue reaction, while the
difficult, has a lower success rate, may induce additional latter is a result of adverse reaction to wear debris of the
damage to the surrounding tissues and increases health implant/bone fixation [24—36]. There is therefore an in-
care costs by one third [7]. creasing concern about the long-term use of UHMWPE,
Replacement arthroplasty made important advance- underscored by recent recognition of the non-specificity
ments during the 1950s and 1960s, through the out- (variable MW and MW distribution, processing and fab-
standing contributions of G.K. McKee and Sir John rication history) of the material, reports on the possible
Charnley. McKee introduced metal-on-metal hip pros- harmful effects of UHMWPE sterilization, and the inter-
thesis in which components were originally made of action of UHMWPE debris with the body fluids and
stainless steel which rapidly changed to a cobalt—chro- tissues [37—41].
mium—molybdenum alloy (VitalliumTM) to mitigate the Simultaneously development in metal-on-metal tech-
excessive friction and rapid loosening of the stainless steel nology [10], through optimization of CoCrMo alloys,
pair [8, 9]. Moreover, the substitution of methyl-methac- prosthesis geometry, and manufacturing practices, has
rylate cement for fixation screws increased the short-term created a renaissance of metal-on-metal prostheses in
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1623

Table 1
Materials combinations used in TJR prostheses

Femoral component Hip/tibial component Results

Co—Cr—Mo Co—Cr—Mo Early high loosening rate and limited use. New developments show lowest
wear rate (THR only — in clinical use in Europe)
Co—Cr—Mo UHMWPE Widely employed; low wear
Alumina/zirconia UHMWPE Very low wear rate. Zirconia more impact resistant (not used in TKR but in
clinical evaluation in Japan)
Alumina Alumina Minimum wear rate (components matched) Pain—Not in clinical use in the US
Ti—6Al—4V UHMWPE Reports of high UHMW-PE wear due to breakdown of titanium surface
Surface coated Ti—6Al—4V UHMWPE Enhanced wear resistance to abrasion. Only thin treated layer achieved

tions, an excellent resistance to degradation (corrosion)


in the human body environment, acceptable strength to
sustain the cyclic loading endured by the joint, a low
modulus to minimize bone resorption, and a high-wear
resistance to minimize debris generation.
Until recently, the mainstream approach taken for the
introduction of orthopaedic materials has involved ad-
aptation of existing materials, as exemplified by the use of
Ti—6Al—4V ELI, an alloy originally designed for aero-
space applications. Standard metallic orthopaedic mate-
rials include stainless steels, cobalt-base alloys, and
titanium-base alloys [44—46] (Table 2), with an increas-
ing number of devices being made of titanium and tita-
nium alloys. The latter alloys are generally preferred to
stainless steel and Co-alloys because of their lower
modulus, superior biocompatibility and corrosion resist-
ance [6, 44]. Recently, new titanium alloy compositions,
specifically tailored for biomedical applications, have
been developed. These first generation orthopaedic alloys
Fig. 2. Comparison of wear behavior of different material combina- included Ti—6Al—7Nb [47] and Ti—5Al—2.5Fe [48, 49],
tions (adapted from Ref. [10]). two alloys with properties similar to Ti—6Al—4V that
were developed in response to concerns relating V to
potential cytotoxicity [50, 51] and adverse reaction with
body tissues [52]. Further, biocompatibility enhance-
Europe [6, 42, 43]. Metal-on-metal combinations may ment and lower modulus has been achieved through the
now provide wear rates lower than metal-or ceramic- introduction of second generation titanium orthopaedic
on-UHMWPE combinations (Fig. 2), eliminating the alloys including Ti—12Mo—6Zr—2Fe ‘TMZF’ [53, 54]
high-wear debris generation and long-term degradation Ti—15Mo—5Zr—3Al [55], Ti—15Mo—3Nb—3O (21SRx)
associated with UHMWPE. [56], Ti—15Zr—4Nb—2Ta—0.2Pd and Ti—15Sn—4Nb—
While continued development of TJR materials has 2Ta—0.2Pd alloys [57], as well as the ‘completely biocom-
increased the success of total joint replacements (with patible’ Ti—13Nb—13Zr alloy [58, 59]. Finally, minimum
success rates of 95% and higher at 10 years and 85—90% elastic moduli have been achieved by ‘TNZT’ alloys
at 15 years), longer human life expectancy and implanta- based on the Ti—Nb—Ta—Zr system [60], specifically by
tion in younger patients has driven bioengineers from the development of the ‘biocompatible’ Ti—35Nb—5Ta—
original implant concerns, e.g., materials strength, infec- 7Zr alloy.
tion and short-term rejection, to consideration of long- This review presents the advances in the development
term materials limitations, e.g., wear, fatigue strength, of orthopaedic titanium alloys. After a brief summary of
and long-term biocompatibility, Fig. 3. The ‘ideal’ mate- the physical metallurgy of titanium and titanium alloys,
rial or material combination for TJR prostheses should their biocompatibility, corrosion behavior, and mechan-
therefore exhibit the following properties: a ‘biocompat- ical properties will be discussed. Finally, their wear be-
ible’ chemical composition to avoid adverse tissue reac- havior will be examined.
1624 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

Fig. 3. Implant materials requirements in orthopaedic applications (adapted from Ref. [126]).

Table 2
Some characteristics of orthopaedic metallic implant materials

Stainless steels Cobalt-base alloys Ti & Ti-base alloys

Designation ASTM F-138 ASTM F-75 ASTM F-67 (ISO 5832/II)


(‘316 LDVM’) ASTM F-799 ASTM F-136 (ISO 5832/II)
ASTM F-1537 ASTM F-1295
(Cast and wrought) (Cast and wrought)
Principal alloying Fe(bal.) Co(bal.) Ti(bal.)
elements (wt%) Cr(17—20) Cr(19—30) Al(6)
Ni(12—14) Mo(0—10) V(4)
Mo(2—4) Ni(0—37) Nb(7)
Advantages f cost, availability f wear resistance f biocompatibility
f processing f corrosion resistance f corrosion
f fatigue strength f minimum modulus
f fatigue strength
Disadvantages f long term behavior f high modulus f power wear resistance
f high modulus f biocompatibility f low shear strength
Primary utilisations Temporary devices Dentistry castings Used in THRs with modular
(fracture plates, screws, hip nails) Prostheses stems (CoCrMo or ceramic) femoral heads
Used for THRs stems Load-bearing components Long-term, permanent
in UK (high Nitrogen) in TJR (wrought alloys) devices (nails, pacemakers)

2. Physical metallurgy of titanium alloys—a brief alloying elements for titanium fall into three categories:
overview a-stabilizers, such as Al, O, N, C, b-stabilizers, such as
Mo, V, Nb, Ta, (isomorphous), Fe, W, Cr, Si, Ni, Co, Mn,
Titanium is a transition metal with an incomplete shell H (eutectoid), and neutral, such as Zr. a and near-a
in its electronic structure enables it to form solid solu- titanium alloys exhibit superior corrosion resistance with
tions with most substitutional elements having a size their utility as biomedical materials being principally
factor within $20%. In its elemental form titanium has limited by their low ambient temperature strength. In
a high melting point (1678°C), exhibiting an hexagonal contrast, a#b alloys exhibit higher strength due to the
close packed crystal structure (hcp) a up to the beta presence of both a and b phases. Their properties depend
transus (882.5°C), transforming to a body centered cubic upon composition, the relative proportions of the a/b
structure (bcc) b above this temperature [61]. phases, and the alloy’s prior thermal treatment and
Titanium alloys may be classified as either a, near-a, thermo-mechanical processing conditions. b alloys (meta-
a#b, metastable b or stable b depending upon their stable or stable) are titanium alloys with high strength,
room temperature microstructure [61, 62]. In this regard good formability and high hardenability. b alloys also
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1625

3. Biocompatibility and corrosion behavior of orthopaedic


titanium alloys

Studies of the biological behavior of metallic elements


have shown that the composition of biomaterials should
be carefully tailored to minimize adverse body reactions
[50, 51]. Local adverse tissue reactions or elicit allergy
reactions caused by metallic implants originate from the
release of metal ions from the implant. This release of
ions depends upon the corrosion rate of the alloy and on
the solubility of the first formed corrosion products. In
an in vivo corrosion study, Steinemann [51] concluded
that V, Ni, and Co were toxic elements while Ti and its
Fig. 4. Pseudo-binary phase diagram of Ti-b stabilizer [63].
alloys, stainless steels and CoCrMoNi alloys, and Ta, Zr,
Nb, and Pt composed the class of ‘resistant metallic
biomaterials’ based on corrosion rates. Consideration of
corrosion product stability in tissue further limits this
offer the unique possibility of combined low elastic choice [79, 84], Ti and some of its alloys, Ta, Nb, and Zr,
modulus and superior corrosion resistance [63, 64]. producing essentially insoluble oxides, Table 3.
A b-alloy is operationally defined as an alloy whose Response to these observations initially resulted in the
chemical composition lies above b , (Fig. 4) that is, it development of two alloys, Ti—6Al—7Nb [47] and
#
contains sufficient total b stabilizer content to retain Ti—5Al—2.5Fe [67], where Nb and Fe were substituted for
100% b upon quenching from above the b transus [63]. V in Ti—6Al—4V, V having been reported to be toxic
Alloys lying above this critical minimum level of b-stabi- [50, 51] and to show adverse tissue effects [52]. These
lizer content may still lie within a two-phase region, with alloys still, however, contained Al which has been sugges-
the resulting as-quenched b-phase being metastable with ted to be causal in osteolysis and neural disorders
the potential of precipitating a second phase upon aging. [68, 69]. Subsequent b-titanium alloys based on the
Alloys with increasing alloying content ultimately ex- Ti—Mo system were then developed: ‘TMZF’ Ti—12Mo—
ceeding a critical b value are considered stable b alloys, 6Zr—2Fe [53, 54], Ti—15Mo—5Zr—3Al [55], and Ti—
4
in which no precipitation takes place during practical 15Mo—3Nb—3O (21SRx) [56], although the large per-
long-time thermal exposure. centage of Mo may still be potentially detrimental, Mo
Process variations are traditionally used to control the having been associated with severe tissue reactions in
alloy microstructure and therefore to optimize titanium animal studies [52]. Elimination of Mo was preferred in
alloys properties, i.e. ductility, strength, fatigue resistance Ti—15Zr—4Nb—2Ta—0.2Pd and Ti—15Sh—4Nb—2Ta—
or fracture toughness. The effects of various microstruc- 0.2Pd alloys [57], although here again elemental Sn and
tures are then correlated with engineering properties, Pd do not show complete biocompatibility. Ultimately,
with the most common microstructural features studied development of Ti—13Nb—13Zr [58, 59] may have an-
in metastable b alloys being b grain size and the size and swered the issue of biocompatibility with the exclusive
distribution of aged a [65]. Apart from a phase, precipi- addition of biocompatible elements, i.e. Zr and Nb. An-
tation of transient b@ or u phases and/or intermetallic other group [70] has investigated the possible use of
compounds may be observed in metastable b alloys de- titanium—zirconium binary alloys. Finally, recently
pending upon alloy composition, heat treatment, pro- synthesized Ti—Nb—Zr—Ta ‘TNZT’ alloys [60] offer
cessing history and service conditions [65, 66]. the opportunity of minimizing potentially adverse tissue

Table 3
Tissue reaction around metallic implants (adapted from [50, 51])

Classification by thicknesses of Minor reaction Severe reaction


pseudomembrane around implant
Titanium alloys Fe, Co, Cr,Ni,
Stainless steels, CoCr alloys Mo, V, Mn, Incoloy

Classification by type of reaction ‘Vital’ ‘Capsule’ ‘Toxic’


(cellular inflammatory, fibrous) Ti, Zr, Nb, Ta, Al, Fe, Mo, Ag, Au, Co, Ni, Cu, V
Pt, Ti alloys Stainless steels,
CoCr alloys
1626 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

Table 4 develop highly protective passive layers, resulting in


Nature of oxides formed on titanium and its alloys (adapted from [71]) a much lower potential electrochemical interaction than
Ti—6Al—4V. Finally, Nb and Zr exhibit ideal passivity
Material Oxide
and are not prone to chemical breakdown of the passive
TiO Al O Nb O MoO /MoO ZrO layer, exhibiting minimum passive dissolution rates. In
2 2 3 2 5 3 2 2
fact, Nb and Zr contribute to the formation of a spontan-
cp Ti x eous highly protective passive film on titanium alloys and
Ti—6Al—4V x x are not, as are Al and V, released into the environment as
Ti—5Al—2.5Fe x x
Ti—6Al—7Nb x x x
dissolved metal ions, but are rather incorporated into the
Ti—15Mo—5Zr—3Al x x x x passive layer [58]. The latter report again emphasizes the
relationship between bulk alloy composition and the
nature of the surface oxides. Electropolishing studies
confirm this suggestion, Table 4, the authors noting that
reaction through the restricted use of ‘biocompatible’ Nb, when the alloying elements (except V and Fe) form an
Ta, and Zr. oxide, those oxides occur as discrete clusters embedded in
As previously mentioned, the biocompatibility perfor- a titanium oxide matrix [71]. Surface oxides composition
mance of a metallic alloy is closely associated with and/or distribution should be expected to affect the cor-
its corrosion resistance and the biocompatibility of its rosion behavior of orthopaedic alloys and detailed char-
corrosion products. Corrosion data show excellent resis- acterizations of these surfaces are required in order to
tance for titanium and its alloys though some adequately interpret and understand corrosion data to
precautions should be taken in order to optimize their optimize the ‘biocompatibility’ of titanium alloys.
composition [50, 51]. b-titanium alloys generally show
attractive corrosion behavior, their corrosion resistance
again depending on alloy composition and environment 4. Mechanical properties of orthopaedic titanium alloys
[64]. For example, anodic polarization tests [53] in-
dicated that Ti—12Mo—6Zr—2Fe (TMZF)’s protective ox- Alloy design and thermo-mechanical processing con-
ide has a breakdown resistance equal to Ti—6Al—4V, trol of titanium alloys has allowed the production of
while corrosion current densities lower than that of cp implant materials with enhanced properties. As shown in
titanium were found for Ti—5Mo—5Zr—3Al alloy [55]. Table 5, strength levels for orthopaedic alloys are gener-
Electrochemical measurements of Ti—13Nb—13Zr ally acceptable with adequate ductility, as defined by
[58, 59] also confirmed the potency of Ti, Nb, and Zr to either the percent elongation or the percent reduction of

Table 5
Orthopaedic alloys developed and/or utilized as orthopaedic implants and their mechanical properties (E"elastic modulus, YS"yield strength,
UTS"ultimate strength)

Alloy designation Microstructure E (GPa) YS (MPa) UTS (MPa)

cpTi MaN 105 692 785


Ti—6Al—4V Ma/bN 110 850—900 960—970
Ti—6Al—7Nb (protasul-100) Ma/bN 105 921 1024
Ti—5Al—2.5Fe Ma/bN 110 914 1033
Ti—12Mo—6Zr—2Fe (TMZF) MMetastable bN 74—85 1000—1060 1060—1100
Ti—15Mo—5Zr—3Al MMetastable bN 75 870—968 882—975
MAged b#aN 88—113 1087—1284 1099—1312
Ti—15Mo—2.8Nb—3Al MMetastable bN 82 771 812
MAged b#aN 100 1215 1310
Ti—0/20Zr—0/20Sn-4/8Nb-2/4Ta#(Pd, N, O) Ma/bN N/A 726—990 750—1200
Ti—Zr Cast Ma@/bN N/A N/A 900
Ti—13Nb—13Zr Ma@/bN 79 900 1030
Ti—15Mo—3Nb—0.3O (21SRx) MMetastable bN#silicides 82 1020 1020
Ti—35Nb—5Ta—7Zr (TNZT) MMetastable bN 55 530 590
Ti—35Nb—5Ta—7Zr—0.4O (TNZTO) MMetastable bN 66 976 1010
CoCrMo MAustenite(fcc)#hcpN 200—230 275—1585 600—1795
Stainless Steel 316 L MAusteniteN 200 170—750 465—950
Bone Viscoelastic composite 10—40 — 90—140
MOHAp#collagenN 150—400!

! Compressive strength.
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1627

Table 6 materials. For example, carbon—carbon and carbon—


Typical modulus of elasticity of joints materials is use (adapted from polymer composites, because of the ability to tailor their
[2]) elastic modulus closer to bone than metals [44, 47], have
Joint material Elastic modulus (GPa)
been investigated as candidates for a new generation of
implants. However, they are far from being totally effec-
Articular cartilage 0.001—0.17 tive due to potential environmental degradation and
Natural rubber 0.0025—0.1 poor tribological behavior.
Silicone rubber 0.01 Alternatively, a first attempt at reducing the elastic
PTFE 0.5
UHMW-PE 0.5
modulus of orthopaedic alloys was made by the intro-
Bone cement (PMMA) 3.0 duction of a/b titanium alloys having elastic modulus
Bone 10—30 values approximately half that of stainless steels or
TNZT alloys 55—66 CoCrMo alloys (Fig. 5). However, the modulus of
‘New generation’ Ti-alloys 74—85 Ti—6Al—4V and related a/b alloys is still high (110 GPa),
Ti—6Al—4V alloy 110
Zirconia 200
approximately 4—10 times that of bone. Recent attempts
Stainless steel 205 at further minimizing orthopaedic alloys moduli have led
Co—Cr—Mo alloy 230 to the introduction of metastable b-titanium alloys,
Alumina 350 Ti—15Mo—5Zr—3Al, Ti—12Mo—6Zr—2Fe (TMZF),
Ti—15Mo—3Nb—0.3O (21SRx) and Ti—13Nb—13Zr, hav-
ing minimum elastic modulus values ranging from 74 to
88 GPa (Fig. 5 and Table 5). The elastic modulus values
area in a standard tensile test, being retained at room of these second generation b-alloys are still 2—7 times
temperature. However there has been, and is still, con- higher than E . Continued synthesis of minimum
"0/%
cern about the high elastic modulus of the alloys as modulus Ti—Nb—Zr—Ta alloys (TNZT) intended for or-
compared to bone, and the variable fatigue resistance of thopaedic applications has recently been demonstrated
the metallic implant. Both properties, if not optimized, [60], these alloys exhibiting moduli 20—25% lower than
may eventually lead to prosthesis failure through loosen- other available alloys (Fig. 5).
ing or fracture. Cyclic loading is applied to orthopaedic implants
Long-term experience indicates that insufficient load during body motion, resulting in alternating plastic de-
transfer from the artificial implant to the adjacent re- formation of microscopically small zones of stress con-
modeling bone may result in bone resorption and event- centration produced by notches or microstructural
ual loosening of the prosthetic device [72, 73]. ‘¼olff ’s inhomogeneities. The interdependency between factors
¸aw (‘The form being given, tissue adapts to best fulfill its such as implant shape, material, processing and type of
mechanical function’) suggests that the coupling of an cyclic loading, makes the determination of the fatigue
implant with a previously load bearing natural structure resistance of a component an intricate, but critical, task.
may result in tissue loss. Indeed, it has been shown that Since testing an actual implant under simulated im-
when the tension/compression load or bending moment plantation and load conditions is a difficult and expen-
to which living bone is exposed is reduced, decreased sive process, standardized fatigue tests have been selected
bone thickness, bone mass loss, and increased osteoporo- for initial screening of orthopaedic material candidates,
sis ensue [74—76]. This phenomenon, termed ‘stress joint simulator trials being generally reserved for a later
shielding’, has been related to the difference in flexibility stage in the implant development process. ‘Standard’
or stiffness, dependent in part on elastic moduli, between fatigue tests include tension/compression, bending, tor-
natural bone and the implant material [77]. Dowson [2] sion, and rotating bending fatigue (RBF) testing, the
appropriately pointed out that, as improvements in the latter, a relatively simple test, being widely used to evalu-
combinations of TJR sliding material pairs have been ate orthopaedic metallic materials. Unfortunately, no
recorded, the elastic modulus of the prosthetic materials standard for fatigue evaluation of biomaterials testing
has been moving further away from those of the natural has yet been established, a variety of testing conditions
joint they were intended to replace (Table 6). Any reduc- being encountered in reported fatigue studies of ortho-
tion in the stiffness of the implant, for example, through paedic materials.
substitution of present orthopaedic alloys with newer, Nonetheless, Ti—6Al—4V is generally considered as
lower modulus materials, is expected to enhance stress a ‘standard material’ when evaluating the fatigue resis-
redistribution to the adjacent bone tissues, therefore min- tance of new orthopaedic titanium alloys. The mechan-
imizing stress shielding and eventually prolonging device ical response of Ti—6Al—4V alloy is, however, extremely
lifetime. sensitive to prior thermo-mechanical processing history,
The problems related to implant stiffness-related-stress e.g., prior b grain size, the ratio of primary a to trans-
shielding of bone have resulted in a number of proposed formed b, the a grain size and the a/b morphologies, all
solutions for more flexible designs and low modulus impacting performance, particularly high-cycle fatigue
1628 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

Fig. 5. Elastic modulus values of orthopaedic alloys.

lifetime (HCF) [78—80]. For example, maximum fracture polishing or surface preparation) may induce early crack
toughness and fatigue crack growth resistance is achieved initiation causing a reduction in the fatigue limit. Indeed,
with Widmanstätten microstructures resulting from a b electropolishing after shot-peening results in the highest
recrystallization anneal. However, this microstructure fatigue limit achievable (30% higher than unshot-peened)
results in inferior HCF performance, the development of [85].
a bi-modal primary a plus transformed b microstructure The sensitivity of Ti-alloy fatigue properties to surface
being preferred [81] to prevent fatigue crack initiation. condition is associated with their high notch sensitivity,
Indeed, the transition to fine equiaxed, fine lamellar, as exemplified by Ti—6Al—4V whose smooth RBF
coarse equiaxed, and coarse lamellar leads to progressive strength is reduced by 40% with notched samples [87].
reductions in lifetime [82]. Various surface preparation techniques and treatments
Further enhancement of the HCF resistance of may result in even larger reductions (up to 80%) in
Ti—6Al—4V may be achieved, under careful control, by fatigue strength [88]. This is illustrated by the effect of
shot peening. Shot peening is a cold working process in surface finishing techniques on the fatigue strength of
which the surface is bombarded with small, typically a and a/b titanium alloys, where a reduction in fatigue
spherical media plastically deforming the surface. The limit of as much as 80% may be observed (Table 7). For
resulting compressive residual stresses may provide in- biomedical applications the notch sensitivity of a/b tita-
creased part life when surface-related failure mechanisms, nium alloys is a critical factor in the performance of
such as fatigue or corrosion, are involved. While shot porous-coated implant for cementless prostheses, where
peening may increase the fatigue limit, a balance between the application of a bead- or wire-coating produces pref-
the high compressive surface residual stresses and the erential crack initiation sites at the porous-coating/sub-
increased surface roughness produced during shot peen- strate interface. Porous-coated Ti—5Al—2.5Fe [89] and
ing is required for optimal fatigue performance. Wagner Ti—6Al—4V [90] hip stems both show a large reduction in
et al. [83—85], have suggested that since the high-cycle the fatigue limit as compared to the smooth condition,
fatigue strength for a smooth surface is determined by the resulting in an unacceptable low fatigue resistance, i.e.,
resistance to fatigue crack nucleation, shot-peening im- below the suggested 425 MPa minimum required for
proves HCF mainly though the beneficial influence of prostheses [91]. An FEM model of Ti—6Al—4V implants,
residual compressive stresses on microcrack initiation correlating with actual results, showed that the porous-
and propagation in the surface region. For instance, coated condition exhibits a HCF strength approximately
shot-peening prior to grit blasting can increase by 10% one-third the strength of the uncoated condition [92] due
the fatigue strength of Ti—6Al—4V over grit blasting alone to the poor fatigue crack initiation resistance of the
[86]. In contrast, the increase of surface damage and Ti—6Al—4V substrate. The latter substrate has, because of
surface roughness due to shot peening (as well as poor the coating sintering treatment, been transformed to
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1629

Table 7
Effect of surface preparation on the fatigue properties of a and a/b titanium alloys

Titanium alloy Test conditions Fatigue limit K`"


&
(MPa)!

cpTi Mechanically polished [88] 234 —


Electrolytically polished 200 0.9
Ti—6Al—4V Gentle surface grinding [88] 427 —
Gentle chemical machining 90 0.8
Abusive chemical machining 352 0.8
Abusive surface grinding 310 0.2
Polished (320—600 alumina grit) [86] 596 —
Belted and glass bead blasted 610 1.0
Belted, beaded, shot-peened, and grit blasted 505 0.9
Belted, beaded, and grit blasted 555 0.8
Ti—5Al—2.5Sn Ultrasonic machined [88] 676 —
Shot peened 531 0.8
Ground 359 0.5
Electrical-discharge machined 145 0.2

! At 107 cycles.
" K is a fatigue strength reduction factor defined as fatigue limit (surface treatment)/fatigue limit (smooth-control) under same test conditions.
&

Table 8
Smooth fatigue strength of orthopaedic titanium alloys

Alloy designation Test conditions Fatigue limit! Fatigue limit/


(MPa) yield strength

cpTi RBF(R"!1/100 Hz) [55] 430 0.6


Ti—6Al—4V axial (R"!1/292 Hz) [87] 500 0.6
axial (R"0.1/292 Hz) [87] 330 0.4
RBF(R"!1/60 Hz) 610 0.7
Ti—6Al—7Nb RBF(R"!1) [47] 500—600 0.7
Ti—5Al—2.5Fe RBF(R"!1) [48] 580 0.8
Ti—15Mo—5Zr—3Al RBF(R"!1/100 Hz) [55] 560—640 0.5
(aged b#a condition)
Ti—13Nb—13Zr axial (R"0.1/60 Hz) [59] 500 0.6
Ti—12Mo—6Zr—2Fe (TMZF) RBF(R"!1/167 Hz) [53] 525 0.5
Ti—15Mo—3Nb—0.3O (21SRx) RBF(R"!1/60 Hz) 490 0.5
TNZT RBF(R"!1/60 Hz) 265 0.5
TNZTO RBF(R"!1/60 Hz) 450 0.5
SS 316L RBF(R"!1/100 Hz) 440 0.6
CoCrMo RBF(R"!1) [91] 400—500 0.4—0.5
RBF(R"!1/100 Hz) [55] 500—580 —

! Fatigue limit at 107 cycles.

a low-crack-initiation-resistance b-transformed coarse the b-transus will increase the fatigue resistance of meta-
lamellar microstructure, with associated high surface stable b-alloys by the transformation of the b-phase to
stress concentrations at the coating/substrate interface two-phase a/b microstructures. For instance, the fatigue
[92]. limit of beta-C may be increased from 390 MPa in the
The introduction of new low modulus orthopaedic SHT condition to 650 MPa after a 16 h/530°C aging
titanium alloys has been accompanied with renewed real- treatment [94]. However, aging increases the elastic
ization that the smooth fatigue resistance of b-titanium modulus, therefore eliminating the benefit of modulus
alloys is generally low [93, 94] when compared to a/b reduction associated with b-alloys. For instance, the elas-
titanium alloys on an equivalent yield strength basis tic modulus of Ti—15Mo—5Zr—3Al increases from 75 GPa
(Table 8). Aging the SHT (Solution Heat Treated: heated in SHT condition to 88—113 GPa after various aging
above b-transus followed by rapid cooling) alloy below treatments [55].
1630 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

Fig. 6. Stress-controlled (RBF, R"!1, 60 Hz) fatigue response of new Fig. 7. Total strain-controlled (RBF, R"!1, 60 Hz) fatigue response
metastable-b Ti-alloys as compared to a#b Ti—6Al—4V. of TNZT alloys and 21SRx as compared to Ti—6Al—4V.

Fatigue properties may also be improved by altering behavior, more closely associated with strain-controlled
the interstitial content (O, C, N, H) as illustrated by the fatigue, may be more representative of in vivo conditions.
TNZT alloys (Fig. 6) [60]. Increasing the oxygen level in For example, hip stems rarely have a smooth surface but
TNZT resulted in an increase in strength and fatigue are typically structured with wedges and coatings cre-
limit for TNZTO, with some increase in modulus, the ating stress concentration sites. When considering fatigue
latter value still, however, remaining below the presently strain, i.e., the ratio between fatigue stress and elastic
available orthopaedic alloys. A similar approach has modulus (Fig. 7), the strain-controlled fatigue behavior of
been undertaken in the development of TIMETAL' TNZT alloys is comparable to that of a/b Ti—6Al—4V
21SRx (21SRx), an orthopaedic grade of the TIME- alloy. Indeed a smaller reduction in fatigue limit occa-
TAL'21S commercial alloy, where ‘toxic’ Al present in sioned by the introduction of notches is typically ob-
the latter was eliminated in the former and compensated served in b-alloys when compared to Ti—6Al—4V, the
by an increase in O content to 0.3 wt % in order to confer former exhibiting a comparable or higher notch fatigue
additional strength to the Rx grade [56]. SHT-21SRx resistance than Ti—6Al—4V in all cases (Table 9).
show typical strength values for b-titanium alloys with
a good fatigue behavior (only 15% lower than
Ti—6Al—4V) (Fig. 6). 5. Wear behavior of orthopaedic titanium alloys
Finally, the lower stress-controlled smooth fatigue
limit of b-alloys may not be an appropriate characteriza- Tribology, defined as the science and technology of
tion for orthopaedic applications, where notch fatigue interacting surfaces in relative motion, and embracing

Table 9
Notch fatigue strength of orthopaedic titanium alloys

Alloy designation Smooth fatigue limit Notch fatigue limit K" KTi6Al4V#
&
(MPa)! (MPa)!

a/b alloys
Ti—6Al—4V 500 290 (K "3.3) 0.6 —
5
290 (K "3.3)
5
Ti—5Al—2.5Fe 580 300 (K "3.6) 0.5 —
5
Ti—15Mo—5Zr—3Al 560—640 190 (K "2.8) 0.3 1.0
5
(aged)
Martensitic a@/b alloy
Ti—13Nb—13Zr 500 335 (K "1.6) 0.7 1.0
5
215 (K "3.0) 0.4 1.3
5
Metastable-b alloy
Ti—12Mo—6Zr—2Fe 525 410 (K "1.6) 0.8 1.4
5
! At 107 cycles.
" K is a fatigue strength reduction factor defined as fatigue limit (notch)/fatigue limit (smooth control) under same test conditions.
&
# KTi6Al4V is a fatigue strength factor relative to Ti—6Al—4V defined as fatigue limit (alloy)/fatigue limit (Ti6Al4V) under same test conditions.
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1631

Fig. 8. Diagram illustrating the sequence of events during oxidative/abrasive/adhesive wear of the Ti—6Al—4V/UHMW-PE tribological pair (adapted
from [16]).

the study of friction, wear and lubrication [95, 96], has have generally shown directional scratching and pitting/
emerged as a primary field in bioengineering. While delamination of bearing surfaces, those features being
healthy natural joints exhibit remarkable tribological non-uniformly distributed over the femoral head area
characteristics, the latter being attributed to the intrinsic [27, 101, 102]. The high UHMWPE wear rates asso-
properties of articular cartilage (high compliance) and ciated with titanium alloy counterparts has been related
synovial fluid and the subsequent optimized lubrication to the mechanical instability of the metal oxide layer
modes [1, 2, 97, 98], total replacement joints based upon [16, 22, 103] (Fig. 8). It has been proposed that when
current available materials experience mixed/boundary normal or shear stresses are high enough to induce
lubrication [6, 97]. This lower lubrication performance is breakdown of the surface passive layer, the oxide will be
generally attributed to the high rigidity (low compliance) disrupted. The exposed metal surface may then either
of artificial materials. As some surface contact takes reform a passive layer or adhesively bond to the polymer
place, friction between artificial materials is much higher surface. The latter situation leads to continuous removal
than in natural joints (k"0.005—0.02) and non-recover- (material disruption) and reformation (oxidation) of the
able wear of the artificial joint materials takes place. passivating layer and results in gradual consumption of
Clinical studies and retrieval examinations have shown alloy material. Concurrently, the surface roughness of the
evidence that excessive wear of UHMWPE and/or metal metal surface will increase which results in yet higher
appears to be the principal mode of failure for the long UHMWPE wear [104, 105]. Ultimately, the breakdown
term use of TJRs [24—29, 30, 33]. Failure generally oc- of the oxide layer creates the potential for abrasive wear,
curs due to excessive wear of the components [99], wear where the hard oxide debris act as third body abrasive
debris accumulation producing an adverse cellular re- components (Fig. 9). Finally, it has also been observed
sponse leading to inflammation, release of damaging that excessive Ti—6Al—4V wear may be caused by the
enzymes, bone cell lysis, osteolysis, infection and pain, presence of foreign bodies in the UHMWPE counterpart
implant loosening eventually ensuing [44, 45]. component leading to severe abrasive wear of the
Early studies of Ti—6Al—4V wear performance in labo- Ti—6Al—4V femoral head [27].
ratory tests has resulted in contradictory conclusions While wrought Co—Cr—Mo and ceramic (alumina and
[100]. Although the Ti—6Al—4V/UHMWPE combina- zirconia) have been preferred to titanium alloys for bear-
tion seemed acceptable for use in total joint replacement ing surface UHMWPE counterpart implant materials,
prosthesis, care should be taken as UHMWPE wear UHMWPE wear and long-term degradation have gener-
rates for Ti—6Al—4V have been reported to 35% greater ated renewed interest in metal-on-metal prostheses. In-
than that for Co—Cr—Mo in hip simulator testing. Re- deed optimum friction and wear conditions can be
trieval of implanted Ti—6Al—4V femoral components achieved and retained with metal-on-metal Co—Cr—Mo
1632 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

Table 10
Oxides (dominant species) to be considered in aqueous solutions [113]

Alloying element Oxide(s)

Ti TiO, TiO , Ti O , TiO , Ti O


2 2 3 3 3 5
Nb NbO, NbO , Nb O
2 2 5
Ta Ta O
2 5
Zr ZrO
2

Fig. 9. Schematic illustration of the suggested oxidative/abrasive wear


process during articulation of metal on UHMWPE (adapted from Ref. mechanical integrity of the surface layer during rubbing
[16]). contact.
As the surface features of titanium alloys are of prime
importance of friction and wear resistance, the nature
and properties of the oxides present in the near-surface
pairs [10]. Notwithstanding this behavior the superior region deserve special attention. Tribo-chemical reac-
biocompatibility and mechanical properties of titanium tions during use, or even detailed surface treatments
alloys make them superior candidates for joint implants before implantation, modify the surface characteristics of
should a better understanding of their tribological behav- the alloy [111, 112]. Further, bulk composition has been
ior and an enhancement in their wear resistance be shown to alter the composition of surface oxide layer, as
achieved. illustrated in Table 4, suggesting it may be possible to
Detailed studies of titanium alloys friction and wear optimize the mechanical response of this outermost layer
performance are however sparse [106]. An original study and improve its properties and integrity to the bulk
on friction and wear properties of a-titanium has shown material through bulk chemical modification. For
poor wear characteristics for unalloyed Ti as well as example, the various ‘biocompatible’ oxides which could
common titanium alloys [107]. This performance was possibly exist in a stable state in aqueous solution are
related to the properties of the oxide layer and the defor- listed in Table 10 [113], where bold formulas correspond
mation behavior of the subsurface regions. a-titanium, to the surface oxide stable in potential-pH conditions
a relatively low shear strength hcp material, exhibited similar to human body fluids surrounding orthopaedic
higher k values, but also greater material transfer, due to implants [45].
its high reactivity, to non-metallic counterfaces, than However, care must be taken regarding the presence of
higher strength materials. Hence, titanium and titanium oxides at the surface and compositional effects cannot be
alloys were considered to have poor oxidative wear res- regarded as the only factor influencing the surface prop-
istance when ‘tribo-chemical’ reactions occur at the con- erties of oxides. The kinetics of repassivation (material-
tact area. In a fundamental tribological study of titanium electrolyte property) and the shear resistance of the oxide
sliding against Al O [108], the static formation of TiO layer (material property only) are two important para-
2 3 2
was reported to decrease the wear and friction coefficient meters that will influence the behavior of the oxide layer
of titanium. However, relatively high friction coefficients [114]. For instance, Ta tends to repassivate more rapidly
(0.4—0.75) were observed at room temperature, these (96 ms) than Ti (172 ms) but more slowly than CoCrMo
values being in contradiction with the reported low fric- (77 ms); Ta’s shear resistance (1.09 N mm~2) is weaker
tion coefficient of TiO (0.1—0.15) [109]. XRD analysis of than Ti (3.7 N mm~2) and CoCrMo (4.9 N mm~2), but
2
wear debris showed that TiO was a dominant oxide, higher than Ti—6Al—4V (0.67 N mm~2). Another study
suggesting that the formation of TiO during tribo-oxida- [115] demonstrated that the presence of Ta and Nb
tion destroys the protective oxide layer and therefore layers at the interface between Ti—30Ta and Al O before
2 3
increases friction. It was proposed that the scaling layer diffusion welding resulted in a reduction of the brittleness
due to tribo-oxidation is composed, from surface to bulk of the interface and a decrease in the O and Al uptake of
material, of a thin TiO layer, a thicker TiO layer, and the metal, and consequently a decrease in the brittleness
2
the Ti matrix. These findings were confirmed in another of the coupling. Oxide reaction at the interface might be
study by analysis of wear debris revealing the cubic TiO the cause of better bonding, reflecting the effect of com-
structure [110], this debris originating from regions position on surface/interface properties.
where critical wear was observed (‘smeared’ regions). In addition to the surface characteristics of titanium
A non-continuous discrete layer of compacted wear frag- alloys, a basic understanding of the mechanisms involved
ment was revealed, suggesting that the mechanical insta- in the friction and wear of titanium alloys is required with
bility of this layer was responsible for the erratic and particular attention on the subsurface deformation in-
high-friction behavior. Sufficient resistance of the under- duced during wear. High-strain deformation occurring in
lying base material to plastic deformation is required for the near-surface zone of a material undergoing wear is
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1633

a fundamental part of the wear process [116, 117] and


has been studied extensively in high stacking fault energy
materials such as copper alloys and aluminium [116—
119], but to a very limited extent in titanium alloys.
Sub-surface wear damage of cp-Ti has been investigated,
cross sections of wear samples revealing a layer of com-
pacted debris at the worn surface, with plastic flow be-
neath. Subsurface plastic flow was characterized by
a highly deformed cellular microstructure with a prefer-
red crystallographic orientation, validating dislocation
theories of friction and wear [120, 121]. While limited
data is available for a-Ti, none exists for b-Ti, as typified
by the new orthopaedic alloys. However, if the low strain
hardening characteristics of a/b titanium alloys [122] are
responsible for their low resistance to deformation within
the subsurface were region, then b-titanium alloys, which
exhibit some strain hardening [123], may offer the poten-
tial for enhanced wear resistance [124].
Cyclic deformation behavior, due to the reciprocating
motion/loading of implants, is a critical element of ortho-
paedic alloys. While in-depth reciprocating-sliding stud-
ies of titanium and titanium alloys are not available,
fretting wear [131—137], resulting from low amplitude
((:300 lm)-reciprocating sliding motion between two
materials, has been studied in titanium and titanium
alloys [137—138]. As the displacement amplitude in-
creases, reciprocating sliding must be considered, that is
when wear mechanisms and wear rates may be related to
those in unidirectional sliding [96]. An important dis-
tinction between fretting and reciprocating-sliding wear
rises from the ease with which the wear debris can escape
from the contact region. After only a few strokes or
passes, the contact situation changes from a two to
a three-body contact formed by the two rubbing speci-
mens and the interface [135]. The wear process, resulting
from velocity accommodation both in the bulk and at the
interface, may be divided into three steps, i.e., (i) particle
detachment (by adhesion, abrasion, corrosion, fatigue,
etc.), (ii) third body behavior (trapped particle in the
interface region, changing in both morphology and com-
position), and (ii) particle ejection.
Waterhouse and Taylor [125] concluded from a fret- Fig. 10. ‘Friction log’ curves of titanium alloys: (a) definition of the
ting study on pure titanium that fretting debris were three axes, (b) aged ‘a’-Ti—15V—3Cr—3Sn, (c) a#b-Ti—6Al—4V, (d) solu-
produced by the spreading of sub-surface cracks leading tion treated b-Ti—15V—3Al—3Cr—3Sn [134].
to the detachment of phatelike particles of oxide coated
metal, this observation being consistent with the de-
lamination theory of wear during sliding. Further studies
of fretting behavior in titanium alloys have focused on elastic deformation of the samples and the device, (2)
defining the particle detachment process and the evolu- tangential load nearly constant with increasing displace-
tion of the superficial layers during wear (gross slip con- ment, actual sliding taking place at the interface. Particle
dition) [133, 134]. Fretting maps or friction logs were detachment was observed in every case, and compacted
obtained (Fig. 10) in titanium—titanium friction condi- debris particle were observed on the wear track. The
tions such that gross slip at the interface was achieved. friction coefficient was very high ('1) and slightly higher
These curves were divided into two parts: (1) tangential for the b alloy (1.2) than for the other a/a#b alloys (1.1).
load linearly increasing with the displacement, corres- In all alloys, a superficial layer with a tribologically
ponding to the elastic response of the system and the transformed structure, named ‘TTS’ by the authors,
1634 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

was observed. This TTS layer was formed of ultra fine ments producing a harder layer composed of various
non-oriented grains of a-titanium with no b-phase. The oxides improve lubrication, no long term data are yet
thickness of this layer decreased with increasing b con- available and the limitation of such surface treatments to
tent, from 100 lm (aged-Ti—15V—3Al—3Cr—3Sn) to 40 lm the modification of only a thin layer ((10 lm in best
(a#b-Ti—6Al—4V) to 15 lm (b-Ti—15V—3Al—3Cr—3Sn). cases) may promote catastrophic wear as the treated
Continuity between the TTS layer and the bulk alloy was surface wears away or become discontinuous.
also reported, with cracks being observed in the TTS Surface modification by oxygen diffusion hardening
region. The authors concluded that the thickness of the (ODH) has been considered to enhance the wear resis-
debris layer, and thus that of the TTS, is a very critical tance of Ti—6Al—7Nb [144]. This treatment provides
parameter, directly controlling the stress transmitted to a gradual increase in hardness through a relatively thick
the surface and near-surface layers. The detachment of 50 lm transformed layer and a friction coefficient for
wear particles was then associated with propagation of ODH—Ti—6Al—7Nb against UHMWPE lower than other
the cracks in the TTS or at the interface between the TTS low wear materials (Table 11). A similar approach was
and the bulk material. Here intense plastic deformation taken by Zwicker et al. [67] for enhanced friction behav-
occurred, wear debris particles, characterized as TiO and ior of Ti—5Al—2.5Fe against UHMWPE (Table 11), using
TiO oxides, originating from the TTS layer. Formation oxide films formation by thermal oxidation. Properly
2
of this TTS was ascribed to deformation-induced trans- oxidized and oil quenched Ti—5Al—2.5Fe balls displayed
formation, the transformation of the near-surface zones friction properties comparable to alumina balls based
leading to the formation of the more stable phase, with on topography measurements made before and after
the major controlling parameters being strains ampli- testing.
tudes and rates. Two types of microstructural changes Sliding wear tests have also been conducted in order to
were observed in the superficial layers of the titanium assess the wear properties of the newer titanium alloys
alloys after friction: (i) transformation of b-phase, and (ii) (Table 11). In general, improved friction and wear behav-
formation of the ultra-fine grained a-phase. While the ior has been observed, with or without surface treatment,
authors concluded through limited evidence that the relative to Ti—6Al—4V. In a pin-on-disk study against
metastable b-phase transformed to a, their X-ray data PMMA cement in deionized water, the friction coeffic-
can also be interpreted as the formation of stress-induced ient of TMZF was found to be less than half that of
orthorhombic martensite, aA [66, 138]. Stress-induced Ti—6Al—4V [53, 54]. At low load and after 105 cycles,
martensite and twinning around wear scratches were also TMZF exhibited no change in surface roughness and no
observed in Ti—6Al—4V. It was finally suggested that the surface scratching. The ‘self-perpetuating’ wear asso-
mechanisms of particle detachment are related to plastic ciated with Ti—6Al—4V, where the formation of third
deformation of the superficial and subsurface layers, but body metallic and bone cement particles results in high
correlations with reciprocating/fretting sliding or con- weight loss of both parts and black debris from the
tinuous sliding observations were not fully established. titanium alloy, was not observed with the TMZF alloy.
However, it can be agreed that the mechanisms of par- When tested against UHMWPE, the friction coefficient
ticle detachment during fretting wear are closely related of the TMZF alloy was again half the value of Ti—6Al—4V
to those observed in continuous/reciprocating sliding, against UHMWPE.
i.e., formation of a highly deformed layer, transfer layer, Diffusion/oxidation surface hardening (DH) was very
particle detachment/delamination, and third body (de- beneficial in improving the abrasive wear of TI—13Nb—
bris, lubricant) contact. Introduction of the influence of 13Zr to levels comparable to Co—Cr—Mo alloy and much
cyclic loading and consequent fatigue behavior still need superior to TiN-coated Ti—6Al—4V [59] (Table 11). This
to be addressed. diffusion hardening process produced a hardened surface
The poor tribological properties of Ti—6Al—4V for im- by diffusion of oxygen into the substrate, and not by
plant articulation surfaces has resulted in the develop- deposition of an overlay coating on the substrate, as in
ment of surface treatments to enhance the hardness and the case of N-implantation. A blue ceramic surface layer,
the abrasive wear resistance of the alloy and thereby to 0.2 lm thick, composed of TiO , TiO, and ZrO , was
2 2
minimize UHMWPE wear debris generation [23, 103, formed on the alloy, the depth of the diffusion hardened
139, 140]. Various procedures including PVD coating layer being 2—3 lm. The presence of ZrO oxides in the
2
(TiN, TiC), ion implantation (N`), thermal treatments ‘ceramic’ surface of diffusion hardened Ti—13Nb—13Zr
(nitriding, diffusion hardening) [141], or laser alloying resulted in improved wear resistance to abrasion sugges-
with TiC [142] have been examined. Ion-implantation ting again that the composition of the oxide layer can be
has been the most common treatment employed tailored through composition control of bulk composi-
[22, 140, 143], resulting in either little or substantial im- tions in order to optimize the surface properties of ortho-
provement in the sliding wear resistance of Ti—6Al—4V, paedic alloys and improve their wear resistance.
though there has been reported consistent improvement Future improvements in the wear resistance of ortho-
in wear resistance to abrasion [139]. While surface treat- paedic titanium alloys will eventually develop from a
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1635

Table 11
Summarized tribological studies of orthopaedic titanium alloys

Titanium alloy Tribo-test conditins! Friction/wear data [‘Reference’ material]


[load-stress/sliding velocity] (same conditions)

Ti—6Al—7Nb [47, 48] Pin-on-disk, Ringer’s#30% serum k k-value (]10~7) k k-value


Abrasive PMMA pin [3.45 MPa/25 mm s~1] 0.100 3.32 (N`) (]10~7)
0.078 2.11 (TiN) 0.094 2.72 [CoCrMo]
0.051 1.35 (ODH") 0.079 2.13 [Al O ]
2 3
Ti—5Al—2.5Fe [67] Pin-on-disk, 0.9% NaCl-solution Depth (km) of wear track on disk
UHMWPE disk [50N] 18—44 (polished) N/A#
12—24 (oxidized)
Ball-in-socket, 0.9% NaCl-solution Friction moment (Nm) Friction moment (Nm)
UHMWPE socket [100—2500 N] 0.5—1 (oxidized, induction hardened) 1.5 [Al O ]
2 3
Ti—15Mo—5Zr—3Al [55] NaCl-solution k wear k wear
Other parameters N/A# 0.82(SHTb) (SHTb)"3](SS316L) 0.43—0.53 cpTi"2]SS
0.43—0.53 (b#a)$"3](SS316L) [SS, cpTi]
(b#a)$
Ti—12Mo—6Zr—2Fe [53] Pin-on-disk, deionized water k k
abrasive PMMA pin [100 g/74 mm s~1] 0.30—0.44 :0.80 [Ti—6Al—4V]
UHMWPE pin [500 g/73 mm s~1] 0.04 0.08 [Ti—6Al—4V]
Ti—13Nb—13Zr [59] Reciprocating pin-on-disk, Ringer’s Depth (km) of metal wear track Depth (km)
PMMA pin [107 MPa/74 mm s~1
(2.5 Hz/15 mm)]
Ti—13Nb—13Zr 0.21 21 [Ti—6Al—4V]
Ti—13Nb—13Zr(ODH") 0.15 7.8 [Ti—6Al—4V/TiN coated]

! Tribo-part in italics is titanium alloy.


" ODH: Oxygen Diffusion Hardening.
# Not Available.
$ Aged condition.

systematic approach based on achieving a basic under- 2. Enhanced biocompatibility and reduced elastic
standing of their tribological properties. Though suc- modulus in titanium alloys have been achieved by the
cessful in many cases, the ‘trial and comparison’ ap- recent development of biomedical alloys baed on b-solu-
proach exemplified in Table 11 by the inconsistent proto- tion treatment (metastable b or martensite a@#b) micro-
cols followed, has limited progress in improving bi- structures.
omaterials properties [145]. Because of the complexity of 3. The wear resistance of b-Ti alloys show some im-
tribology and wear problems, a systematic approach provement when compared to a#b alloys.
aiming at understanding basic mechanisms, suitable to 4. Overall alloy composition, which controls surface
a large number of non-specific applications, should oxide composition and subsurface deformation behavior,
be implemented. More specifically, the separate invest- is a critical factor in the wear behavior of b-alloys.
igation of surface, subsurface, and third body (debris) 5. Ultimately the use of orthopaedic titanium alloys as
behaviors, the three wear ‘precursors’ as described wear components will require a more detailed under-
by Sannino and Rack [146] could ultimately identify standing of the basic wear mechanisms involved.
basic wear mechanisms while avoiding misleading extra-
polation when different experimental parameters are se-
lected. Acknowledgements

The authors would like to acknowledge the partial


6. Conclusions support of Osteonics, Inc. (Allendale, NJ), Teledyne-AL-
LVAC (Monróe, NC), and TIMET (Henderson, NV).
1. Titanium alloys are generally preferred mateials for The authors wish to thank Jonathan Black, Ph.D., FBSE,
orthopaedic applications due to their relatively low for his review of this manuscript and his valuable com-
modulus vis-à-vis Co—Cr—Mo alloys, superior biocom- ments. Special thanks are due to Martine LaBerge,
patibility and corrosion resistance. Ph.D., for numerous beneficial discussions.
1636 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

References mechanisms and influence of fixed-stress magnitude. J Biomed


Mater Res 1987;21:367—77.
[1] Mow VC, Soslowsky LJ. Friction, lubrication, and wear of [23] Peterson CD, Hillberry BM, Heck DA. Component wear of total
dirthridial joints. In: Basic orthopaedic biomechanics. New knee prostheses using Ti—6Al—4V, titanium nitride coated
York: Raven Press Ltd., 1991:245—92. Ti—6Al—4V, and cobalt-chromium-molybdenum femoral com-
[2] Dowson D. Bio-tribology of natural and replacement synovial ponents. J Biomed Mater Res 1988:22.
joints. In: Mow VC, Ratcliffe A, Woo SL-Y, editors. Bio- [24] Agins HJ, Alcock NC, Bansal M, Salvati EA, Wilson PD, Pel-
mechanics of diarthrodial joints, vol. II. Chap. 29. New York: licci PM, Bullough PG. Metallic wear in failed titanium-alloy
Springer, 1992:305—45. total hip replacements. J Bone J Surg 1988;70-A(3):347—56.
[3] Ardill J. What is orthopaedics? PhD Thesis. The Department of [25] Argenson J-N, O’Connor JJ. Polyethylene wear in meniscal knee
Orthopaedic Surgery, The Queen’s University of Belfast, 1995. replacement. J Bone J Surg 1992;74-B(2):228—32.
[4] Schumacher HR. In: Schumacher HR, Klippel JH, Robinson [26] Berry DJ, Wold LE, Rand JA. Extensive osteolysis around an
DR, editors. Primer on the Rheumatic Diseases, 9th edn. At- aseptic, stable, uncemented total knee replacement. Clin Orthop
lanta, GA: The Arthritis Foundation, 1988. Relat Res 1993;293:204—7.
[5] Lowman EW. Osteoarthritis. J Amer Med Acad 1955;157:487. [27] Black J, Sherk H, Bonini J, Rostoker WR, Schajowicz F, Galante
[6] Dowson D. Friction and wear of medical implants and pros- JO. Metallosis associated with a stable titanium-alloy femoral
thetic devices. ASM Handbook, vol. 18. Gereland: Materials component in total hip replacement. J Bone J Surg 1990;72-A(1):
Park, OH: ASM International, 1992:656—64. 126—30.
[7] American Academy of Orthopaedic Surgeons. Survey Presented [28] Cates HE, Faris PM, Keating EM, Ritter MA. Polyethylene
at the Orthopaedic Research Society Annual Meeting. Orlando, wear in cemented metal-backed acetabular cups. J Bone J Surg
FL, February 1995. 1993;75-B(2):249—53.
[8] McKee GK, Watson-Farrar J. Replacement of arthritic hips by [29] Engh GA, Dwyer KA, Hanes CK. Polyethylene wear of metal-
the McKee-Farrar prosthesis. J Bone J Surg 1996;48-B(2): backed tibial components in total and uncompartimental knee
245—59. prostheses. J Bone J Surg 1992;74-B(1):9—17.
[9] Walker PS, Gold BL. The tribology (friction lubrication and [30] Winter M, Griss P, Scheller G, Moser T. 10—14 years results of
wear) of all-metal artificial hip joints. Wear 1971;17:285—99. ceramic-metal-composite hip prostheses with a cementless
[10] Protek AG. METASULTM metal-on metal articulation. Tech- socket. In: Heimke G, editor. Bioceramics, vol. 2. Cologne,
nical brochure, Protek AG, Switzerland, Edition 2/94, 1994. Germany: Deutsche Keramische Gesellschaft eV, 1990:
[11] Streicher RM, Schön R, Semlitsch M. Investigation of the 436—44.
tribological behavior of metal-on-metal combinations for artifi- [31] Jones SMG, Pinder IM, Moran CG, Malcom AJ. Polyethylene
cial hip joints. Biomedizinische Technik 1990;35(5):3—7. wear in uncemented knee replacements. J Bone J Surg 1992:
[12] Charnley J. Arthroplasty of the hip. Lancet 27 May 1961:1129—32. 74-B:18—22.
[13] Charnley J. Total hip replacement by low-friction arthroplasty. [32] Kabo JM, Gebhard JS, Loren G, Amstutz HC. In vivo wear of
Clin Orthop Relat Res 1970;72:7—21. polyethylene acetabular components. J Bone J Surg 1993;75-B:
[14] Galante JO, Rostoker W. Wear in total hip protheses. Acta 254—58.
Orthp Scand Suppl 1973;145:1—46. [33] Nolan JF, Bucknill TM. Aggressive granulomatosis from poly-
[15] Dumbleton JH. Wear and prosthetic joints. In: Morrey BF, ethylene failure in an uncemented knee replacement. J Bone
editor. Joint replacement arthroplasty, Chap 6. New York: J Surg 1992;74-B(1):23—4.
Churchill Livingstone, 1991:47—9. [34] Santavirta S, Nordström D, Metsärinne K, Kontinnen YT. Bio-
[16] Poggie RA, Wert JJ, Mishra AK, Davidson JA. Friction and compatibility of polyethylene and host response to loosening of
wear characterization of UHMWPE in reciprocating sliding cementless total hip replacement. Clin Orthop Relat Res 1993;
contact with Co—Cr, Ti—6Al—4V, and zirconia implant bearing 297:100—10.
surfaces In: Denton R, Keshavan MK, editors. In: Wear and [35] Shanbhag AS, Jacobs JJ, Glant TT, Gilbert JL, Black J, Galante
friction of elastomers. ASTM STP 1145, Philadelphia, PA: JO. Composition and morphology of wear debris in failed un-
American Society for Testing and Materials, 1992:65—81. cemented total hip replacement. J Bone J Surg 1994;76-B(1):
[17] McKellop H, Clarke IC, Markolf KL, Amstutz HC, Friction and 60—7.
wear characteristics of polymer, metal, and ceramic prosthetic [36] Jacobs JJ, Shanbag A, Glant TT, Black J, Galante JO. Wear
joint materials evaluated on a multichannel screening device. debris in total joint replacements. J Amer Acad Orthop Surg
J Biomed Res 1981;15:619—53. 1994;2(4):212—20.
[18] Oonishi H, Tsuji E, Mizukoshi T, Fujisawa A, Murata N, [37] Dumbleton JH. The clinical significance of wear in total hip anf
Kushitani S, Aono M, Meguro Y. Wear of polyethylene and knee prostheses. J Biomater Appl 1988;3:3—32.
alumina in clinical cases of alumina total knee prosthesis. In: [38] Ries MD, Rose RM, Greer J, Weaver KD, Sauer WL. Steriliza-
Hulbert J, Hulbert SF, editors. Bioceramics 3. Terra Haute, IN: tion-induced effects on UHMWPE performance properties. Sci-
Rose-Hulman Institute of Technology, 1992:137—45. entific Exhibition Presented at the 62nd Annual Meeting of the
[19] Oonishi H, Aono M. Clinical results of total knee arthroplasty in American Academy of Orthopaedic Surgeons. Orlando, FL,
combination with alumina against polyethylene—a five to eight 16—21 February 1995.
year follow up study. In: Hulbert J, Hulbert SF, editors. Bio- [39] Sutula LC, Sperling DK, Collier JP, Saum KA, Williams IR.
ceramics 3. Terra Haute, IN: Rose-Hulman Institute of Techno- Clinical wear of UHMWPE acetabular and tibial compo-
logy, 1992:147—56. nents—impact of gamma sterilization in air and material con-
[20] Saikko V. Wear and friction properties of prosthetic joint mate- solidation. Scientific exhibition Presented at the 62nd Annual
rials evaluated on a reciprocating pin-on-flat apparatus. Wear Meeting of the American Academy of Orthopaedic Surgeons.
1993;166:169—78. Orlando, FL, 16—21 Februay 1995.
[21] Saikko V. Wear of polyethylene acetabular cups against zirconia [40] White SE, Whiteside LA, Poggie RA, Farrar D, Tanner MG.
femoral heads studied with a hip joint simulator. Wear In-vivo wear and material properties of UHMWPE tribial com-
1994;176:207—12. ponents. Scientific Exhibition Presented at the 62nd Annual
[22] Buchanan RA, Rigney ED Jr, Williams JM. Wear-accelerated Meeting of the American Academy of Orthopaedic Surgeons.
corrosion of Ti—6Al—4V and nitrogen-ion-implanted Ti—6Al—4V: Orlando, FL, 16—21 February 1995.
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1637

[41] Wright Medical Technology Inc. Effects of sterilization methods [60] Ahmed T, Long M, Silvestri J, Ruiz C, Rack HJ. A new low
on UHMWPE. Technical Monograph Presented at the 62nd modulus, biocompatible titanium alloy. Presented at the
Annual Meeting of the American Academy of Orthopaedic 8th World Titanium Conference. Birmingham, UK, October
Surgeon. Orlando, FL, 16—21 February 1995. 1995.
[42] Streicher RM, Semlitsch MF, Weber H, Schon R. Metal-on- [61] Collings EW. The physical metallurgy of titanium alloys, ASM
metal articulation: a new generation of wear resistant implants. Series in Metal Processing. Gegel HL, editor. Cleveland, Metals
Transactions of the 20th Annual Meeting of the Society for Park, OH: American Society for Metals, 1984.
Biomaterials. Boston, MA, April 1994:323. [62] Polmear JJ. Titanium alloys. In: Light alloys, Chapter 6. Lon-
[43] Medley JB, Krygier JJ, Bobyn JD, Chan FW, Tanzer M. don, UK: Edward Arnold Publ, 1981.
Metal—metal bearing surfaces in the hip: early wear results from a [63] Bania PJ. Beta titanium alloys and their role in the titanium
simulator apparatus. Transactions of the 21st Annual Meeting of industry In: Eylon D, Boyer RR, Koss DA, editors. Titanium
the Society for Biomaterials. San Francisco, CA, April 1995:47. Alloys in the 1990’s. Warrendale: The Mineral, Metals & Mate-
[44] Park JB, Lakes RS. Biomaterials—An introduction, 2nd edn. rials Society, 1993:3—14.
New York, London: Plenum Press, 1992. [64] Schutz RW. An overview of beta titanium alloy environmental
[45] Black J. Biological performance of materials—fundamentals of behavior. In: Eylon D, Boyer RR, Koss DA, editors. Beta Tita-
biocompatibility, 2nd edn. New York: Marcel Dekker Inc., 1992. nium Alloys in the 1990’s. Warrendale, PA: The Mineral, Metals
[46] Freese H, Teledyne-Allvac, Monroe NC. Private communica- & Materials Society, 1993:75—91.
tion, 1995. [65] Boyer RR, Hall JA. Microstructure-property relationships in
[47] Semlitsch MF, Weber H, Streicher RM, Schön R. Joint replace- titanium alloys (critical review). In: Froes FH, Caplan I, editors.
ment components made of hot-forged and surface-treated Titanium’92 Science and Technology, Warrendale, PA: The
Ti—6Al—7Nb alloy. Biomater 1992;13(11):781—8. Mineral, Metals & Materials Society, 1993:77—88.
[48] Borowy K-H, Kramer K-H. On the properties of a new titanium [66] Vassel A. Microstructural instabilities in beta titanium alloys. In:
alloy (TiAl5Fe2.5) as implant material. In: Titanium’84 Science Eylon D, Boyer RR, Koss DA, editors. Beta Titanium Alloys in
and Technology, vol. 2. Munich, Deutsche Gesellschaft Für the 1990’s. The Mineral, Metals & Materials Society,
Metallkunde EV, 1995:1381—6. 1993:173—85.
[49] Zwicker R, Bühler K, Müller R, Beck H, Schmid H-J, Ferstl J. In: [67] Zwicker J, Etzold U, Moser Th. Abrasive properties of oxide
Kimura H, Izumi O, editors. Titanium’80 Science and Techno- layers on TiAl15Fe2.5 in contact with high density polyethylene.
logy. Warrendale, PA, TMS-AIME 1980:505—14. In: Titanium’84 Science and Technology, vol. 2. Munich,
[50] Steinemann SG. Corrosion of surgical implants-in vivo and in Deutsche Gesellschaft Für Metallkunde EV 1985;2:1343—50.
vitro tests In: Winter GD, Leray JL, de Groot K, editors. [68] Perl DP, Brody AR. Alzeihmer’s disease: X-ray spectrometric
Evaluation of Biomaterials. New York: Wiley, 1980:1. evidence of aluminum accumulation in neurofibrillary tangle-
[51] Steinemann SG. Corrosion of titanium and titanium alloys bearing neurons. Science 1980;208:297—99.
for surgical implants. Titanium’84 Science and Technology, [69] Crapper DR, McLachlan DR, Farnell B, Galin H, Karlik S,
vol. 2. Munich, Deutsche Gesellschaft Für Metallkunde EV Eichhorn G, De Boni U. Aluminum in human brain disease. In:
1985;2:1373—9. Sarkar B, editor. Biological Aspects of Metals and Metals-Re-
[52] Laing PG, Fergosun AB Jr, Hodge ES. Tissue reaction in rabbit lated Diseases. New York: Raven Press, 1993:209—18.
muscle exposed to metallic implants. J Biomed Mater Res [70] Kobayashi E, Matsumoto S, Doi H, Yoneyama T, Hamanaka
1967;1:135—49. H. Mechanical properties of the binary titanium-zirconium
[53] Wang K, Gustavson L, Dumbleton J. The characterization of alloys and their potential for biomedical materials. J Biomed
Ti—12Mo—6Zr—2Fe. A new biocompatible titanium alloy de- Mater Res 1995;29:943—50.
veloped for surgical implants. In: Beta Titanium in the 1990’s. [71] Mäusli P-A, Steineman, SG, Simpson JP. Properties of surface
Warrendale: The Minerals, Metals & Materials Society, 1993: oxides on titanium and some titanium alloys. 6th World Confer-
49—60. ence on Titanium, les éditions de physique, Les Ulis Cedex,
[54] Wang K, Gustavson L, Dumbleton J. Low modulus, high France, part IV, 1988:1759—64.
strength, biocompatible titanium alloy for medical implants. In: [72] Dujovne AR, Bobyn JD, Krygier JJ, Miller JE, Brooks CE.
Titanium’92 Science and Technology. Warrendale: The Min- Mechanical compatibility of noncemented hip prostheses with
erals, Metals & Materials Society, 1993:2697—704. the human femur. J Arthroplasty 1993;8(1):7—22.
[55] Steinemann SG, Mäusli P-A, Szmukler-Moncler S, Semlitsch M, [73] Engh CA, Bobyn JD. The influence of stem size and extent of
Pohler O, Hintermann H-E, Perren SM. Beta-titanium alloy for porous coating on femoral bone resorption after primary cement-
surgical implants. In: Titanium’92 Science and Technology. less hip arthroplasty. Clin Orthop Relat Res 1988;231:7—28.
Warrendale: The Minerals, Metals & Materials Society, [74] Cheal EJ, Spector M, Hayes WC. Role of loads and prosthesis
1993:2689—96. material properties on the mechanics of the proximal femur after
[56] Fanning JC. TIMET Technical Laboratories. Henderson NV, total hip arthroplasty. J Orthop Res 1992;10:405—22.
Private communication, 1995. [75] Prendergast P, Taylor D. J Biomed Eng 1990;12:379.
[57] Okazaki Y, Ito Y, Ito A, Tateishi T. Effect of alloying elements [76] Huiskes R, Weinans H, van Rietbergen B. The relationship
on mechanical properties of titanium alloys for medical im- between stress shielding and bone resorption around total hip
plants. Mater Trans JIM 34(12):1217—22 (1993); J Japan Inst stems and the effects of flexible materials. Clin Orthop Relat Res
Metals 1993;57(3):332—7. 1992;274:124—34.
[58] Kovacs P, Davidson JA. The electrochemical behavior of a new [77] Sumner DR, Galante JO. Determinants of stress shielding: de-
titanium alloy with superior biocompatibility. In: Titanium’92 sign versus materials versus interface. Clin Orthop Relat Res
Science and Technology, Warrendale: The Minerals, Metals & 1992:274:202—12.
Materials Society, 1993:2705—12. [78] Chesnutt JC, Thompson AW, Williams JC. Influence of metal-
[59] Mishra AK, Davidson JA, Kovacs P, Poggie RA. lurgical factors on the fatigue crack growth rate in alpha—beta
Ti—13Nb—13Zr: a new low modulus, high strength, corrosion titanium alloys. Rockwell International Science Center, Thou-
resistant near-beta alloy for orthopaedic implants. In: Beta Tita- sand Oaks, CA. Report d AFML-TR-78-68, Air Force Ma-
nium in the 1990’s. Warrendale: The Minerals, Metals & Mate- terials Laboratory, Air Force Systems Command, Wright-Pat-
rials Society, 1993:61—72. terson Air Force Base, OH, May 1978.
1638 M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639

[79] Peters M, Gysler A, Lütjering G. Influence of microstructure on [97] Fisher J, Dowson D. Tribology of total artificial joints. Proc
the fatigue behavior of Ti—6Al—4V. In: Kimura H, Izumi O, Instn Mechanical Engineers—IMechE 1991;205:73—9.
editors. Titanium’80—Science and technology, 1981:1777—86. [98] Unsworth A. Tribology of human and artifical joints. Proc Inst
[80] Rüdinger K, Fischer D. Relationship between primary alpha Mechnical Engineers—IMechE 1991;205:163—72.
content, tensile properties and high cycle fatigue behavior of [99] Jasty M. Clinical reviews: particulate debris and failure of total
Ti—6Al—4V. In: Lütjering G, Zwicker U, Bunk W, editors. Tita- hip replacements. J Appl Biomater 1993;4:273—6.
nium’84—Science and technology, vol. 4. Munich, Deutsche [100] Clarke IC, McKellop HA, McGuire P, Okuda R, Sarmiento A.
Gesellschaft Fur Metallkunde EV, 1985:2123—30. Wear of Ti—6Al—4V implant alloy and ultrahigh molecular
[81] Boyer RR, Bajoraitis R, Spurr WF. The effects of thermal pro- weight polyethylene combinations. In: Luckey HA, Kubli F,
cessing variations on the properties of Ti—6Al—4V. In: Micro- editors. Titanium alloys in surgical implants. ASTM STP 796,
structure, fracture toughness and fatigue crack growth rate in Philadelphia, PA: American Society for Testing and Materials,
titanium alloys. The Minerals, Metals & Materials Society, 1983:136—47.
Warrendale, PA, 1987:149—70. [101] Dobbs HS, Scales T. Behavior of commercially pure titanium
[82] Lütjering G, Gysler A. Fatigue. In: Lütjering G, Zwicker U, and Ti-318 (Ti—6Al—4V) in orthopedic implants. In: Luckey HA,
Bunk W editors. Titanium’84—Science and Technology, vol. 4. Kubli F, editors. Titanium alloys in surgical implants. ASTM
Munich, Deutsche Gesellschaft Fur Metallkunde EV 1985: STP 796. Philadelphia, PA: American Society for Testing and
2065—83. Materials, 1983:173—86.
[83] Wagner L, Gerdes C, Lütjering G. Influence of surface treatment [102] Easter TL, Graham RM, Jacobs JJ, Black J, LaBerge M. Clinical
on fatigue strength of Ti—6Al—4V. In: Lütjering G, Zwicker U, performance of Ti6Al4V femoral components: wear mechanisms
Bunk W, editors. Titanium’84 — Science and Technology, vol. 4. vs. surface profile. Proc 20th Annual meeting of the Society For
Munich, Deutsche Gesellschaft für Metallkunde EV, Biomaterials, Boston, MA, 5—9 April 1994:185.
1985:2147—54. [103] Rostoker W, Galante JO. The influence of titanium surface
[84] Wagner L, Lütjering G, Jaffee RI. Optimization of bi-modal treatments on the wear of medical grade polyethylene. Biomater
microstructure and texture in Ti—6Al—4V. In: Kim Y-W, Boyer 1981;2(10):221—4.
RR, editors. Microstructure/property relationships in titanium [104] Dowson D, El-Hady Diab MM, Gillis BJ, Atkinson R. Influence
aluminides and alloys. Warrendale: The Minerals, Metals & of counterface topography on the wear of ultra high molecular
Materials Society, 1991:521—31. weight polyethylene under wet or dry conditions. In: Lee L-H,
[85] Wagner L, Lütjering G. Influence of surface condition on fatigue editor. Polymer wear and its control. Washington DC: Ameri-
strength. In: Kitagawa J, Tanaka T, editors. Fatigue 90. Birmin- can Chemical Society, 1985:171—87.
gham, UK: Materials and Component Engineering Publications [105] Fisher J, Dowson D, Hamdzah H, Lee HL. The effect of sliding
Ltd, 1991:323—8. velocity on the friction and wear of UHMW-PE for use in total
[86] Eberhardt AW, Kim BS, Rigney ED, Kutner GL, Harte CR. artificial joints. Wear 1994;175:219—25.
Effects of precoating surface temperatures on fatigue of [106] Budinski KG. Tribological properties of titanium alloys. Wear
Ti—6Al—4V. J Appl Biomat 1995;6:171—4. 1991;151:203—17.
[87] Military Handbook—Metallic materials and elements for aero- [107] Miller PD, Holladay JW. Friction and wear properties of tita-
space vehicle structures. US Department of Defense, Washing- nium. Wear 1958/59;2:133—40.
ton DC, MIL-HDBK-5C, 1976:5/78. [108] Hong H, Winer WO. A fundamental tribological study of Ti/Al O
2 3
[88] Williams DN, Wood RA. Effects of surface condition on the contact in sliding wear. J Tribol—Trans ASME 1989;111:504—9.
mechanical properties of titanium and its alloys. MCIC-71-01 [109] Steijn RP. Friction and wear of rutile single crystals. ASLE
Report. Metals and Ceramics Information Center, Battelle Col- Trans 1969;12:21—33.
umbus Laboratories, OH, August 1971. [110] Nutt SR. A study of the friction and wear behavior of titanium
[89] Merget M, Aldinger F. Influence of technological parameters on under dry sliding conditions. In: Wear of Materials 1983. New
the fatigue strength of Ti5Al2.5Fe — a new material for endop- York: The American Society of Mechanical Engineers,
rostheses. In: Lütjering G, Zwicker U, Bunk W, editors. Tita- 1983:426—33.
nium’84—Science and Technology, vol. 4. Munich, Deutsche [111] Hernández de Gatica NL, Jones GL, Gardella JA Jr. Surface
Gesellschaft Fur Metallkunde EV 1985:1393—400. characterization of titanium alloys sterilized for biomedical ap-
[90] Cook SD, Georgette FS, Skinner HB, Haddad RJ Jr, Fatigue plications. Appl Surface Sci 1993;68:107—21.
properties of carbon- and porous-coated Ti—6Al—4V alloy. [112] Raikar NG, Gregory J, Ong JL, Lucas LC, Lemons JE,
J Biomed Mater Res 1984;18:497—512. Kawahara D, Nakamura M. Surface characterization of tita-
[91] Lorenz M, Semlitsch M, Panic B, Weber H, Willert HG. Fatigue nium implants. J Vac Sci Technol 1995;A 13(5):2633—7.
strength of cobalt-bae alloys with high corrosion resistance for [113] Pourbaix M. Atlas of electrochemical equilibria in aqueous
artificial hip joints. Eng Med 1978;7(4):241—50. solutions. New York: Pergamon Press, 1966.
[92] Wolfarth D, Filiaggi M, Ducheyne P. Parametric analysis of [114] Tümmler HP, Thull R. Surface properties of titanium and its
interfacial stress concentrations in porous coated implants. alloys — mechanical and electrochemical investigation. In: Tita-
J Appl Biomat 1990;1:3—12. nium’84 Science and Technology, vol. 2. Munich, Deutsche
[93] Fanning JC. In: Eylon D, Boyer RR, Koss DA, editors. Beta Gesellschaft Für Metallkunde EV 1985;2:1335—42.
Titanium Alloys in the 1990’s Warrendale, PA: The Minerals, [115] Gibbesh B, Elssner G, Petzow G. Investigation of Ti/Al O
2 3
Metals & Materials Society, 1993:411—62. joints with intermediate tantalum and niobium layers. Biomater
[94] Gregory JK, Wagner L. Heat treatment and mechanical behav- 1992;13(7):455—61.
ior in Beta CTM. Presented at the VII International Meeting on [116] Rice SL, Nowotny H, Wayne SF. A survey of the development of
Titanium, 15 November 1991, Torino, Italy. Organised by GTT, subsurface zones in the wear of materials. Key Eng Mater
Ginatta Torino Titanium S.p.a. Torino. 1989;33:77—100.
[95] Dumbleton JH. Tribology of natural and artificial joints. Tribol- [117] Wert JJ. The role of microstructure in subsurface damage in-
ogy Series 3. Amsterdam: Elsevier Scientific Publishing Com- duced by sliding contact. Key Eng Mater 1989;33:101—34.
pany, 1981. [118] Heilmann P, Clark WAT, Rigney DA. Orientation determina-
[96] Hutchings IM. Tribology: friction and wear of engineering ma- tion of subsurface cells generated by sliding. Acta Metall
terials. Boca Raton, FL: CRC Press, 1992. 1983;31(8):1293—305.
M. Long, H.J. Rack / Biomaterials 19 (1998) 1621—1639 1639

[119] Kuo SM, Rigney DA. Sliding behavior of aluminum. Mater Sci [133] Blanchard P, Colombie C, Pellerin V, Fayeulle S, Vincent L.
Engng 1992;A157:131—43. Material effects in fretting wear: application to iron, titanium,
[120] Kulhman-Wilsdorf D. Dislocation concepts in friction and wear. and aluminum alloys. Metall Trans A 1991;2A:1535—44.
In: Rigney DA, editors. Fundamentals of friction and wear of [134] Fayeulle S, Blanchard P, Vincent L. Fretting behavior of tita-
materials. Metals Park, OH: American Society for Metals, nium alloys. Tribol Trans 1993;36(2):267—75.
1980:119—86. [135] Berthier Y, Vincent L, Godet M. Fretting fatigue and fretting
[121] Hirth JP, Rigney DA. Microstructural models for friction and wear. Tribol Int 1989;22(4):235—42.
wear. In: Haasen P, Gerold V, Kostorz G, editors. Strength of [136] Colombie C, Berthier Y, Vincent L, Godet M. Fretting wear and
Metals and Alloys, vol. 3. Oxford: Pergamon Press, 1980: fretting fatigue damage. Fatigue’87, Warley, England, 1987:567—75.
1483—502. [137] Vincent L. Usure et frottement: une approach pluri-disciplinaire
[122] Kuhlman GW. A critical appraisal of thermomechanical pro- ‘mécanique-matériaux’. In: Traitements de surface et protection
cessing of structural titanium alloys. In: Kim Y-W, Boyer RD, contre la corrosion. Audisio S, Caillet M, Galerie A, Mazille H,
editors. Microstructure/property relationships in titanium editors. CNRS, Paris, Chap. 2.6, 1987.
alumindes and alloys. Warrendale, PA: The Minerals, Metals & [138] Flower HM. The plastic deformation of metastable b titanium
Materials Society, 1991:465—91. alloys. In: Lütjering G, Zwicker U, Bunk W, editors. Tita-
[123] Ling F-W, Starke EA, Lafevre BG. Deformation behavior and nium’84—Science and Technology, vol. 4. Munich, Deutsche
texture development in beta Ti—V alloys. Metall Trans Gesellschaft Fur Metallkunde EV 1985:1651—8.
1974;15:179—87. [139] McKellop HA, Röstlund TV. The wear behavior of ino-om-
[124] Moore MA, Douthwaite RM. Plastic deformation below worn planted Ti—6Al—4V against UHMW polyethylene. J Biomed
surfaces. Metall Trans A 1976;7A:1833—9. Mater Res 1990;24:1413—25.
[125] Waterhouse RB, Taylor DE. Fretting debris and the delamina- [140] Rieu et al. J. Ion implantation effects on friction and wear of joint
tion theory of wear. Wear 1974;29:337—44. prosthesis materials. Biomater 1991;12:139—43.
[126] Hoeppner DW, Chandrasekarn V. Fretting in orthopaedic im- [141] Morton PH, Bell T. Surface engineering of titanium. Mémoires
plants: a review. Wear 1994;173:189—97. et Ëtudes Scientifiques Revue de Métallurgie 1989; 86(10):639—46.
[127] Waterhouse RB, Dutta MK. The fretting fatigue of titanium and [142] Chengwei C, Zhiming Z, Xitang T, Wang Y, Sun XT. Influence
some titanium alloys in a corrosive environment. Wear of rapidly solidified structures on wear behavior of Ti—6Al—4V
1973;25:171—5. laser alloyed with TiC. Tribol Trans 1995;38(4):875—8.
[128] Waterhouse RB, Lamb M. Fretting corrosion of orthopaedic [143] Tateishi T, Terui A, Yunoki H. Friction and wear properties of
implant materials by bone cement. Wear 1980;60:357—68. biomaterials for artificial joint. In: Heimke G, editor. Bio-
[129] Fricker DC, Shivanath R. Fretting corrosion studies of universal ceramics, vol. 2. Cologne, Germany: German Ceramic Society,
femoral head protheses and cone taper ingots. Biomater 1992:145—51.
1990;11:495—500. [144] Streicher RM, Weber H, Schön R, Semlitsch M. New surface
[130] Sandstrom PW, Sridharan K, Conrad J. A machine for fretting modification for Ti—6Al—7Nb alloy: oxygen diffusion hardening
wear testing of plasma surface modified materials. Wear 1993; (ODH). Biomater 1991;12:125—9.
166:163—8. [145] Berthier Y. Tribologie: science carrefour. Collection of Papers
[131] Bill RC. Selected fretting-wear-resistant coatings for and dabates, European Conference on Braking. Lille, France,
Ti—6%Al—4%V alloy. Wear 1985;106:283—301. 3—4 December 1992:1—26.
[132] Cronin J, Warburton J. Amplitude-controlled transitions in fret- [146] Sannino AP, Rack HJ. Dry sliding wear of discontinuously
ting: the comparative behavior of six materials. Tribol Int reinforced aluminum composites: review and discussion. Wear
1988;21(6):309—15. 1995;189:1—19.

S-ar putea să vă placă și