Sunteți pe pagina 1din 13

Inuence of crosslinked polyethylene structure on wear of joint replacements

Alastair Kilgour

, Alistair Elck
Institute for Materials and Processes, The University of Edinburgh, Edinburgh EH9 3JL, UK
a r t i c l e i n f o
Article history:
Received 21 July 2008
Received in revised form
21 November 2008
Accepted 28 November 2008
Keywords:
Polymer
Wear
Scanning electron
a b s t r a c t
Crosslinking is known to increase the wear resistance of ultra-high molecular weight polyethylene
(UHMWPE) used as an acetabular cup in total hip replacement. The same wear benet is not afforded
when UHMWPE is used as a tibial component. A programmable multi-directional motion and dynamic
load tribometer has been used to investigate ultra-structural development in both unirradiated (PE)
and highly crosslinked (100kGy) UHMWPE (+PE). To investigate surface anisotropy in UHMWPE, both
linear-reciprocating and elliptical wear paths were applied. Following three million elliptical cycles,
crosslinking reduced wear by up to 92%. Under reciprocating motion, mean steady state wear of PE
and +PE groups was not signicantly different (p 0.652). Raman spectra indicated a de-crystallisation
zone on the near surface of PE and +PE reciprocating pins. This was attributed to large strain
development in conjunction with slow lamellar removal and renewal of new surface material in the
lower wearing specimens. SEM images of fragmented lamellae supported this observation.
& 2008 Elsevier Ltd. All rights reserved.
1. Introduction
The perceived role of pin-on-plate wear testers in orthopaedic
research has been to rank material combinations in terms of their
efcacy as bearing couples [16]. Materials which show good
performance (low friction, low wear, good abrasive resistance), in
these preliminary investigations will then be studied further on
full joint simulators. Simulators closely approximate the physio-
logical conditions found in the natural joint. For the majority of
cases, the bearing material combination in use for total hip
replacement (THR) has changed little conceptually since its
introduction in the early 1960s [7]; a hard wearing metallic/
ceramic ball component articulates against a low wear polymer
socket. The polymer in question is ultra-high molecular weight
polyethylene (UHMWPE) which forms an extremely effective
bearing component. The discovery that UHMWPE micron and
sub-micron particles can induce a pathogenic response in
periprosthetic tissue [810], causing late-stage failure after some
1015 years has stimulated work into understanding the origins of
wear. The failure mode is not wearing-out of the acetabular
component but rather wear-mediated osteolysis causing loss of
bony support around the implant [9]. Pain due to micro-motion
and joint instability results in joint revision surgery [11]. Revision
surgery can be technically challenging and the operational lifetime
of revised components often being shorter than the primary
implant. The increasing number of young, active patients receiving
THR places greater tribological demands on these bearings.
For applications in orthopaedics, crosslinking of UHMWPE
through either gamma irradiation or electron beam radiation in an
inert environment, has been shown to reduce wear in pin-on-plate
tests [1214], hip simulator studies [15], and preliminary short-
term radiograph follow-up studies [16,17]. Irradiation doses used
commercially range from 50 to 105kGy. Exposing UHMWPE to
radiation is not a recent development. As early as 1968, UHMWPE
acetabular components were traditionally sterilised using a
comparatively low 25kGy dose of gamma irradiation in air.
Dumbleton and Shen report early attempts to increase the wear
resistance of UHMWPE by exposing the material to elevated doses
of gamma irradiation (levels well beyond those for sterilisation) to
induce pronounced chain scission, free radical formation and
subsequent crosslinking [18,19]. Wear tests were conducted on a
simple ring-on-disc machine. However, in comparison to uni-
rradiated material, an increase in coefcient of friction and wear
rate was associated with the highly irradiated specimens. Findings
were attributed to newly created crosslinks preventing the
formation of a stable transfer-lm on the metal counterface. These
wear tests were, however, conducted in the absence of lubrication.
It has since been established from pin-on-plate work that wear
tests conducted in the presence of serum have been found to
consistently produce worn surfaces more representative of those
from explanted components [3]. Furthermore, contemporary wear
tests in dilute bovine calf serum have consistently shown cross-
linked material to offer greater wear resistance.
Recent hypotheses for the apparent increase in wear resistance
of polyethylene after crosslinking are based on theories of
retardation in the re-orientation of crystalline lamellar in response
to friction forces. Tie-molecules within the inter-crystalline
ARTICLE IN PRESS
Contents lists available at ScienceDirect
journal homepage: www.elsevier.com/locate/triboint
Tribology International
0301-679X/$ - see front matter & 2008 Elsevier Ltd. All rights reserved.
doi:10.1016/j.triboint.2008.11.011

Corresponding author.
E-mail address: Alastair.Kilgour@googlemail.com (A. Kilgour).
Tribology International ] (]]]]) ]]]]]]
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
amorphous domains give an increased resistance to reorientation
[12,20]. Wear of an UHMWPE acetabular cup occurs mostly on the
surface, a result of micro-adhesion and abrasive wear mechanisms
[21]. Under repeated cyclic loading and frictional stresses at the
articulating surface, an anisotropic and orientated layer develops
in response to strain accumulation [22]. The discovery of
microstructural alignment in response to quasi-linear motion of
early wear screening devices [23] and the consequent strain
hardening [24] offered an explanation for the divergent wear rates
produced in early pin-on-plate studies when compared to
considerable in vivo wear rates [25]. Microstructural orientation
in one axis is thought to precipitate accelerated wear from a
perpendicular motion; the cross-shear direction [24,26]. The
complex, cross-shearing motion paths found in the hip joint are
a result of abduction/adduction and internal/external rotation
about the exion/extension path. Debris produced in response to
multi-directional motion on a strain hardened surface is released
through fracture [22] and rupture [26] of material drawn out of
the surface as particles are incrementally enlarged and released
[27]. As a result of crosslinking, the addition of increased
carboncarbon covalent bonds between the amorphous molecules
is thought to retard chain movement and produce a worn surface
more resistant to bular pull-out [15,28]. Frictional strain has also
been proposed to produce a near-surface plasticity zone [20]. This
sub-surface zone, a precursor to wear, has been shown to occur on
crosslinked specimens albeit to a smaller depth beneath the
surface. This suggests the transfer of strain through inter-lamellae
communication is more conned in crosslinked material. How-
ever, multi-axial loading tests revealed more pronounced orienta-
tion in highly crosslinked polyethylenes over their conventional
counterparts, especially in higher molecular weight samples,
adding to uncertainty as to the role of tie-molecules and
interlamellar communication [29]. It has been reported that
crosslinking does not prevent microstructural mobility [30], but
the contribution of such a polymer network to the wear resistance
under cyclic loading and shear stresses is still not clear.
To date there have been few pin-on-plate studies on ortho-
paedic grade crosslinked UHMWPE [1214,31,32], and fewer still
on microstructural characterisation [32]. Zhou et al. reported on
microstructural disparity between unmodied and 100kGy
gamma irradiated UHMWPE [32]. Wear tests were conducted on
a reciprocating pin-on-disc machine, the polymer pin loaded in a
constant manner onto a reciprocating cobalt chrome disc. Under
average nominal contact pressures between 20 and 30MPa they
attributed a fourfold increase in wear resistance of the crosslinked
material to greater resistance to plastic ow. Tests conducted on
motion machines with limited cross-shear angles, more repre-
sentative of knee motion have reported marginal differences in
wear rates as function of radiation dose [30]. Further studies on
this material have been conducted on machines, which produce
limited translation and rotation [14,31]; polymer test pins are
loaded under constant force onto a reciprocating test plate, or
counter-bearing. In such cases the pin rotates via drive gears to
operate in synchronisation with the oscillating plate. The nature
of the rotating pin inherently creates non-uniform tribological
conditions across the wear surface. Each point on an increasing
radius from the pin centre will experience a different wear path,
creating a complex worn surface. This tribological setup is more
suitable for wear ranking bearing couples under constant load and
sinusoidal velocity, than used to probe microstructural evolution.
A well designed pin-on-plate tribometer has the potential to offer
in-sight into fundamental wear mechanisms under controlled
conditions.
Advances in pin-on-plate design have led to a number of
contemporary desktop machines capable of producing conven-
tional UHMWPE wear data in what is considered the clinical
range, assessed through linear penetration rates and wear factor
comparisons [31,33,34]. Our hypothesis was to use a custom built,
novel, pin-on-plate wear machine capable of programmable
dynamic load and motion to compare wear, topography, and
morphology in PE and +PE materials. This novel multi-
directional pin-on-plate tribometer was used to investigate
whether crosslinks aid plasticity mechanisms by improving
interlamellar communication, aiding reorientation whilst resist-
ing chain fracture. Or, alternatively, whether cross-links anchor
the resultant microstructure, alleviating alignment in the pre-
ferential sliding direction reducing susceptibility to wear as a
result of cross-shear.
2. Materials and methods
2.1. Materials
Two groups of materials were supplied by Smith and Nephew
Inc. The virgin material and control group were ram extruded
GUR1050 bar stock. The second group underwent a subsequent
gamma irradiation dose of 100kGy at room temperature. A post-
irradiation above melt stabilisation process at 1501C quenched
remaining residual free radicals from the bulk. The rod stock was
nally machined into pins of 5mm diameter and parted-off in
20mm lengths. Elevated crosslink density levels were characterised
through swelling experiments according to the ASTM standard for
the determination of swell ratio [35]. The swelling ratio, described
as the volume of a swollen polymer network divided by the volume
of the original unswollen network [36] indicates the maximum
amount of liquid that such a network can hold. The ability to
dissolve or absorb the solvent and swell is dependant on the
crosslink density, the structural integrity under equilibrium swel-
ling conditions and the afnity (interaction parameter) between
the polymer and the swelling solvent. It is difcult to physically
measure the volume of the swollen network. Instead, masses were
recorded and converted to volume using a predened ratio relating
the density of polymer to that of the solvent at the equilibrium
swelling temperature. Swelling was performed on three specimens
from each group. Crosslink density was then calculated according
to the method of Flory and Rehner [37].
2.2. Wear tests
In accordance with ASTM F732 [38], a six-station wear-tester
(Fig. 1) was designed and built capable of programmable dynamic
loading and motion proles [39]. The ASTM standard denotes the
clinical relevance of wear methods implemented in conjunction
with machines designed to evaluate simplied specimen geome-
tries, offering guidelines for the type of motion, pin and plate
conditions and appropriate test protocols.
The machine is a biaxial device with two degrees of freedom.
Two programmable stepper controlled linear slides provide the
necessary mechanics, allowing the carriage to translate along the
x and, y axes and interpolate between the two. Test pins are
loaded independently in the z-axis by six electric coil actuators.
The arrangement is such that the actuators are free to follow and
consistently load the polymer test pins as wear occurs. Test pins
did not rotate relative to the plate, ensuring a uniform wear path
across the complete pin contact area. The six test plates were xed
in individual lubricant chambers, four on the motion carriage to
provide constrained motion in the xy plane (Fig. 2) and two
stationary loaded soak controls.
A schematic showing key components and a single station are
shown in Fig. 3. Four polymer test pins from each material group
were subjected to linear reciprocating and multi-directional
ARTICLE IN PRESS
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 2
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
elliptical motion, articulating against medical grade cobalt
chrome alloy plates, polished (Buehler, Alpha) to an average
femoral implant nish [40] of 0.01mm Ra (Zygo, NV100). Wear
paths were chosen to approximate the length of those found from
gait analysis [4143]. Linear reciprocating (stroke length 25mm)
and elliptical (20mm major and 10mm minor axes, respectively)
wear paths were applied (Table 1). The elliptical path was applied
to better approximate the more open quasi-elliptical wear paths
observed in vivo. A low 2:1 aspect ratio (AR) of the ellipse was
used to provide this open wear path.
Lower aspect ratios have been found to restrict lamellae
alignment [44] and at these ratios, discrepancies in the orienta-
tion softening model [26] suggest some of the underlying
assumptions maybe overly simplied [27,34]. A double peak
dynamic load prole [45,46] with peaks 1 and 2 of 94.5 and 48.3N
(corresponding to nominal contact stresses of 4.7 and 2.4MPa
during the stance phase) was synchronised to each motion cycle,
the heel-strike occurring at the point of maximum exion, as
shown in Fig. 4. The swing phase was programmed to account for
50% of each cycle. Load was not recorded continuously but
instead, data collected at the beginning and end of each test
period throughout the duration of the wear test. Wear was
assessed through gravimetric analysis every 250,000 cycles using
a Mettler balance with a resolution of 70.001mg. Each pin was
cleaned and dried in a method described in ASTM F732.
Subsequently, pins were weighed in order, each four times, giving
a resultant repeatability of 70.05mg. All tests were run in
newborn calf serum (Harlan Sera lab) diluted 1:1.5 with Ringers
solution to a physiological protein concentration of 22g/l (10ml of
serum:15ml of Ringers solution per chamber). To inhibit bacterial
growth, 0.2wt% sodium azide was added. All pin-on-plate stations
were lubricated independently. The serum was kept at 37721C
through PID control.
ARTICLE IN PRESS
Fig. 1. The six-station wear test device; shown with the dust cover lid open
exposing the four load and motion stations and two load only stations.
Fig. 2. Detailed view of the four load and motion stations. Each chamber is
supported by four columns providing future capacity to mount strain gauges or
install a 3-axis load cell beneath each chamber for friction and continuous load
measurement. Heating resistors, mounted on the underside of each stainless steel
base provide necessary energy to heat chamber lubricant.
Fig. 3. Simplied schematic of the design with one station shown: (1) electric coil actuator; (2) loading arm; (3) pivot housing; (4) test pin holder; (5) acrylic chamber wall;
(6) axis 1; (7) axis 0; (8) stainless steel collar; (9) CoCrMo test plate; (10) o-ring seal; (11) Tufnol base; and (12) stainless steel base.
Table 1
Test parameters.
Material GUR 10500kGy GUR 1050100kGy
Test Linear Elliptical Linear Elliptical
Duration (km) 150 150 150 150
Cycles (10
6
) 3 3 3 3
Stroke length (mm) 25.00 25.00
Cycle circumference (mm) 49.67 49.67
Av sliding speed (mms
1
) 50.65 50.17 50.65 50.17
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 3
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
2.3. Vibrational spectroscopy
Raman spectroscopy was used to assess near-surface crystal-
linity before and after each wear test. This non-invasive, non-
destructive analysis technique has previously been used to
calculate polyethylene crystallinity [4749] and UHMWPE acet-
abular cup crystallinity [5053]. Crystallinity is often used to
summarise the morphological constituents that form the bulk
material. Changes in the ratio of crystalline to amorphous
domains can signify changes in mechanical behaviour [54] hence,
wear resistance of the polymer [20]. The internal Raman
vibrational band structure of polyethylene is well studied
[47,49]. The 1416cm
1
band is associated with CH
2
bending in
the orthorhombic crystal phase and the 1296cm
1
and broader
second peak around 1303cm
1
is associated with amorphous CH
2
twisting phase. According to Strobl [49] polyethylene crystallinity
can be calculated from the ratio
%Crystallinity I
1414
=I
1293
I
1305
0:46
A Renishaw InVia system was used to measure Raman spectra at
room temperature. Excitation of Raman bands was achieved by a
785nm laser ca 30mW, focused through a 50 objective. Spectra
were gathered in confocal mode to minimise scattering volume.
The depth resolution of the system was calculated at less than
10mm. Depth proling was performed by focusing the laser beam
at increasing depth beneath the surface. Ten scans were taken at
each depth. Scattering volume (V) was approximated as a cylinder
of length (l)10mm, with a cross-section equal to the focused
laser spot (d) 1mm,
V p d
2
=4 10 7:85mm
3
On each pin and at each scanning depth (surface, sub 12.5, 25 and
37.5mm), 10 spectra were recorded. Therefore, the volume of
material fromeach pin contributing to the crystallinity calculation
per depth was 78mm
3
. Hence, the average scattering volume
from four worn pins was 314mm
3
per depth. UHMWPE lamellae
are typically on the order of 50nm 1mm, depth and length,
respectively. The ten scans performed will therefore, include a
substantial population of crystals in an effort to statistically
represent the bulk material.
2.4. Scanning electron microscopy
Field emission scanning electron microscopy (Hitachi S-5000)
was used to qualitatively evaluate the extent of wear-induced
lamellar re-orientation and topographical wear features following
the completion of three million wear cycles. Oxidising acids
have historically been used to etch chemically resistant semi-
crystalline polymers to reveal microstructural detail [55]. The acid
preferentially diffuses into and attacks the disordered less tightly
packed amorphous domains [56], leaving the crystalline tightly
packed chains within the lamellae proud of the surface. A less
oxidizing acid, permanganic etchant has been shown to reveal
lamellar structures with more detail [57,58]. This technique has
previously been applied to orthopaedic UHMWPE to assess the
friction oriented response of lamellae after simple pin-on-plate
wear studies [32,59] and in hip simulator studies [20]. Worn pins
were sectioned in half to permit separate imaging of topography
and microstructure, Fig. 5.
3. Results
3.1. Swelling
Swelling experiments clearly differentiated PE from +PE as
shown in Fig. 6. Unirradiated material (PE) was found to swell
over 10 times more than crosslinked. This was attributed to
elevated crosslink levels reducing the number of possible chain
conformations limiting the swell-ability, and the improved
structural integrity making the network less prone to chain
stretch and solvent absorption. From FloryRehner network
swelling model, the average molecular weight between crosslinks
was 238,080 and 5663g/mol for PE and +PE, respectively.
3.2. Wear tests
Unirradiated and crosslinked wear volume with respect to
linear reciprocating sliding distance is shown in Fig. 7. Both
groups exhibit similar wear rates, with two exceptions. The
irregular accelerated wear rates of conventional pins at (i) and (ii)
ARTICLE IN PRESS
100
80
60
40
20
0
L
o
a
d

(
N
)
L
o
a
d

(
N
)
F
l
e
x
i
o
n
/
E
x
t
e
n
s
i
o
n

(
m
m
)

F
l
e
x
i
o
n
/
E
x
t
e
n
s
i
o
n

(
m
m
)

A
d
d
u
c
t
i
o
n
/
A
b
d
u
c
t
i
o
n

(
m
m
)
A
d
d
u
c
t
i
o
n
/
A
b
d
u
c
t
i
o
n

(
m
m
) -2
0
2
4
6
8
10
12
14
16
18
20
22
24
26
100
80
60
40
20
0
12
10
8
6
4
2
0
-2
-4
-6
-8
-10
-12
12
8
6
4
2
0
-2
-4
-8
-10
-12
Fig. 4. Synchronised load and motion proles for (a) linear reciprocating wear and (b) elliptical wear.
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 4
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
are a result of low lubricant levels during those periods. Never-
theless, the difference in mean steady-state wear between PE
and +PE under reciprocating motion was not signicant (Student
T-Test p 0.652), Fig. 8.
Fig. 9 shows the volume of PE and +PE material removed with
respect to sliding distance under multi-directional motion. The
elliptical wear path produced signicantly higher mean wear rates
in unirradiated pins (Welshs T-Test po0.05); 100kGy irradiation
was shown to reduce wear at the completion of three million
elliptical cycles by 92%.
Because all groups exhibited an approximately linear relation-
ship with sliding distance, a steady-state wear coefcient; a
measure of wear volume produced with respect to normal load
and total sliding distance was calculated according to Archard
[60]. The coefcient, originally developed for metal-on-metal dry
sliding and modelled on adhesive wear is not without burden
when applied to soft-on-hard rubbing in the presence of proteins.
However, it is often adopted as a way of communicating the
probability that UHMWPE debris will be released from the
articulating surface. In these calculations we applied the average
force per cycle, 42.872.8N (S.D.), calculated from integrating
under the load prole. The steady-state wear factors for both
material groups subjected to reciprocating motion were two orders
of magnitude lower than wear data calculated from retrieved or
radiographic studies of conventional UHMWPE (Table 2). Under
ARTICLE IN PRESS
35
30
25
20
15
10
5
0
-PE +PE
S
w
e
l
l
i
n
g

r
a
t
i
o
Fig. 6. Effect of radiation dose on swelling ratio.
Worn surface
Worn surface
Worn surface
Etch to reveal extent of near-surface
lamellae alignment
SEM for topographic information
(ripples, folds, cracks, debris)
Fig. 5. Worn pin surfaces were sectioned longitudinally to permit topography and morphology studies.
2.5
2
1.5
1
0.5
0
0 20 40 60 80 100 120 140 160
sliding distance (km)
W
e
a
r

v
o
l
u
m
e

(
m
m
3
)
(i)
(ii)
(1, 2)
Fig. 7. Wear graphs under linear reciprocating motion of: (1) PE pins, note at (I)
chamber 1 ran dry and (II) chamber 4 lubricant level ran low, both causing
elevated wear rates and (2) +PE pins.
4.00
3.50
3.00
2.50
2.00
1.50
1.00
0.50
0.00
-PE linear +PE linear +PE elliptical -PE elliptical
w
e
a
r

r
a
t
e

(
m
g
/
m
i
l
l
i
o
n
)
Fig. 8. Comparison of the average wear rates of unirradiated and irradiated
material under both linear reciprocating and elliptical sliding.
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 5
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
elliptical motion, the wear coefcient and linear penetration rates
for the unirradiated group were in good agreement with clinical
data, supporting the importance of open wear paths in pin-on-
plate studies. The elliptical +PE wear factor was an order of
magnitude lower than that of the unirradiated group subjected to
the same motion, and an order of magnitude higher than either
group under reciprocating wear.
3.3. Raman spectroscopy
Fig. 10 plots the changes in crystallinity of unirradiated
UHMWPE with respect to depth of Raman scattering. The group
subjected to elliptical motion showed no statistically signicant
change in surface and sub-surface crystallinity when compared to
the unworn control, Fig. 10a and b, respectively. The mean load-
soak crystallinity was consistently around 50%, Fig. 10b. A drop in
crystallinity caused by linear reciprocating motion was apparent.
The average surface crystallinity of 49.8% is 5% lower than either
the unworn control or those surfaces worn in a multi-directional
manner. Lower crystallinity in this group was consistently found
up to the maximum depth recorded of sub-37.5mm, Fig. 10d.
+PE control crystallinity was 53.6171.80% within the rst
37.5mm of material, Fig. 11a. A gradual increase in +PE crystallinity
was observed for groups; Fig. 11bd as the scattering volume
moved deeper beneath the surface. Under elliptical motion, a
mean surface crystallinity of 51.1171.59 was similar to that of the
unworn control surface, 53.1671.92% (ANOVA p 0.112). +PE pins
followed a similar trend to that of the unirradiated material;
linear motion resulted in the largest reduction in surface crystal-
linity, 47.7874.371%. This was statistically lower than the +PE
control surface crystallinity (ANOVA po0.05). Lower crystallinity
values were consistently found in all crosslinked groups.
3.4. Scanning electron microscopy
3.4.1. Unirradiated
Parallel ripples were imaged on the surface of the more rapidly
wearing unirradiated pins under elliptical motion, Fig. 12a.
Tearing, folding and cracking were all imaged at higher magni-
cation, Fig. 12b. Linear reciprocating produced light scratches,
the edges of which were folded, plastically deformed producing
brils running perpendicular to the motion path. Areas of PE
were re-organised into domains running parallel to the surface in
the direction of wear, Fig. 12c. In areas where this formation was
highly developed, elongated brils were observed extending from
the surface, Fig. 12d.
Following etching, distinct microstructural developments
were imaged in response to wear path. Worn surfaces subjected
to multi-directional motion exhibited random lamellar orienta-
tion and similar lamellar size and breadth to that of the unworn
control surface (Fig. 13a and b), the white arrow indicates the
primary elliptical wear axis. Higher magnication conrmed
similarities between the two groups, Fig. 13d and e. Alignment
and lamellar break-up mechanisms were less obvious on the worn
elliptical surfaces. Evidence of texture development was more
prominent under linear reciprocating motion. Large domains
of lamellae were imaged orientated in the direction of wear
(Fig. 13c). At higher magnication, evidence of lamellar break-up
and fragmentation (Fig. 13f) was revealed. This was substantiated
with Raman data which conrmed lower crystallinity in the linear
reciprocating group.
ARTICLE IN PRESS
14
12
10
8
6
4
2
0
0 20 40 60 80 100 120 140 160
(2)
(i)
(1)
W
e
a
r

v
o
l
u
m
e

(
m
m
3
)
Sliding distance (km)
Fig. 9. Wear graphs under elliptical motion of: (1) PE pins, note at (i) chamber 3
lubricant level ran low causing an increase in wear rate and (2) +PE pins.
Table 2
Comparison of wear factors and linear penetration rates.
Material Wear path Wear factor
( 10
6
mm
3
/Nm)
Penetration
(mm/yr)
PE Reciprocating 0.03 0.003
+PE Reciprocating 0.04 0.0026
PE Elliptical 1.6 0.18
+PE Elliptical 0.13 0.01
Conventional Clinical [56] 2 0.15
Fig. 10. PE crystallinity for (a) three million elliptical cycles; (b) control surface;
(c) three million load only cycles; and (d) three million linear reciprocating cycles.
Fig. 11. Highly crosslinked UHMWPE crystallinity for: (a) unworn control (b) three
million elliptical cycles; (c) three million load only cycles; and (d) three million
linear reciprocating cycles.
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 6
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
3.4.2. Irradiated
The micrographs of +PE surfaces worn under elliptical motion
were highly smooth, indicated by the lack of contrast picked on
the FE-SEM. Sub-micron debris was also imaged on the surface,
Fig. 14a. In other areas a granular texture was observed; the beads
appeared on the order of 500nm in width, Fig. 14b. This texture
was found highly developed in some regions. Beads were drawn
out into folded elongated strands running across the primary
elliptical wear axis, Fig. 14c. Finally, small parallel ripples were
consistently imaged on +PE linear reciprocating surfaces, Fig. 14d.
The wavelengths of these features were approximately 1mm, and
were found transverse to the direction of sliding.
Fig. 15 shows microstructure images of +PE subjected to three
million linear and elliptical cycles in comparison to the unworn
control, Fig. 15a. Both Fig. 15b and c shows a degree of surface
lamellae orientation and mobility in response to the different
motion paths, less so under the elliptical path. Lamellar were imaged
folded, bent and broken in response to friction forces. Fig. 15d shows,
at higher magnication the post-irradiated unworn lamellar size,
shape and distinct lack of preferential alignment. In contrast, the low
wearing surfaces of Fig. 15e and f (linear and elliptical wear,
respectively) showed increasing signs of lamellar break-up and
micro-domains of stack rotation and close packing. The distinction
between linear and elliptical motion on microstructural develop-
ment was found to be less clear in crosslinked material.
4. Discussion
To date, few studies have attempted to link microstructural
evolution and wear rate of unirradiated and irradiated orthopae-
dic polyethylene. Such information could improve understanding
of the role of crosslinking on wear reduction and debris size and
shape. The wear performance of both highly crosslinked and
conventional polyethylene is intimately related to the sliding
conditions. The experimental results presented herein conrm the
notion that crosslinking reduces volumetric wear in UHMWPE,
when sliding takes the form of open motion tracks possessing a
cross-shearing action. Linear-reciprocating motion, dened as
having zero cross-shear [26], did not exhibit a statistically
signicant difference in the mean steady-state wear rates
between the two material groups. These results support the
ndings of Wang et al. [30] who previously investigated this
motion dependant behaviour by wear testing UHMWPE exposed
to 0100kGy on both knee and hip simulators. The tibial
components, tested on a knee simulator with limited cross-shear,
were found to wear similarly for all radiation dosages. In contrast,
acetabular components tested on a hip simulator were shown to
have exponentially decreasing wear rate with increasing radiation
dose. In this study, +PE worn under elliptical motion was found to
wear signicantly less than PE, supporting the motion depen-
dant wear response of crosslinked polyethylene.
The use of dynamic loading in pin-on-plate studies has been
adopted by few investigators. Muratoglu et al. used a dynamically
loaded (Paul-type, 445N peak 1, 290N peak 2), bi-directional
(10mm5mm rectangular wear path) tribometer to investigate
wear rate reduction for UHMWPE with a radiation dose of 100kGy
in comparison to unirradiated material. They report wear rates
for unirradiated material as 9.8mg/million cycles reducing to
1.6mg/million cycles for irradiated; an 84% reduction in wear rate
[12]. The wear rate of +PE subjected to elliptical motion in the
ARTICLE IN PRESS
Fig. 12. Scanning electron micrographs of surface topography after three million cycles for: (a) PE elliptical wear showing regular ripples across surface. Mag. 800; (b)
PE elliptical wear with tearing, folding and opening of surface cracks running 451 across primary wear axis, Mag. 8000; (c) PE linear wear showing light scratching
and bril tearing, Mag. 2000; and (d) PE linear wear showing origin of bular debris fromhighly textured surface (see black arrow), Mag. 4000. White arrows indicate
primary wear direction.
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 7
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
current study was found to be 0.26mg/million cycles substantially
lower than that of Muratoglu et al. Wear factors were not reported
for the bi-directional tests so a direct comparison between the
tribological conditions, with respect to sliding distance and
nominal load is not possible. However, the sliding distance per
cycle of the ellipse used in this work is 50 mm in comparison to
the total rectangular sliding path length of 30 mm. This equates to
the pins used in this study sliding an additional 20km per million
cycles. From Archards equation, volumetric wear is proportional
to contact force and sliding distance, therefore the increased
ARTICLE IN PRESS
Fig. 13. Scanning electron micrographs after three million cycles for: (a) PE control surface; (b) PE elliptical wearno preferential alignment observed; (c) PE linear
reciprocating weartexture development and alignment; (d) magnied control surface; (e) PE lamellae after three million elliptical cycles; and (f) PEreciprocating
motion produced fragmented lamellae. White arrows indicate axis of primary wear path.
Fig. 14. Scanning electron micrographs of surface topography after three million cycles for: (a) +PE elliptical wear showing micron and sub-micron debris on a highly
smooth worn surface Mag. 2500; (b) +PE elliptical wear with granular texture, Mag. 10000 ; (c) +PE elliptical wear, adhesive drawing out from initial granular texture,
Mag. 10000; and (d) +PE linear wear showing ripple formation perpendicular to sliding direction, Mag. 8000. White arrows indicate primary axis of wear path.
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 8
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
sliding distance used in the present study would be expected to
produce higher not lower wear rates per million cycles. In
addition, Muratoglu et al. employed 9mm diameter pins onto
which a heel-strike peak contact stress of 7MPa was applied
equating to a peak force of 445N; higher than the peak force of
94.3N used in the present study. The effect of dynamic loading on
the wear rate is likely to be complex as additional modes of
lubrication may be experienced between different load proles
and over statically loaded instruments. Detailed comparison
between the static and two exemplars of dynamically loaded
pin-on-plate testing is beyond the scope of this discussion.
Recognition of the importance of cross-shear in the wear of
cross-linked UHMWPE resulted in multi-directional pin-on-plate
studies on retrotted linear-reciprocating machines. Often a
statically loaded pin is rotated; either at a constant angular
velocity or oscillated over a given arc, both synchronised with a
reciprocating plate. Such instruments have shown further reduc-
tions in wear rate of crosslinked material, albeit to a lesser extent.
Galvin et al. report a 73% reduction in 100kGy material rubbing
against a smooth counterface in comparison to unirradiated
material [14]. The reduced benet in wear resistance for cross-
linking found by Galvin et al. may result from the variable
tribological conditions experienced at the wear surface. The
rotating pin-on-oscillating plate design results in a continuously
variable wear path along the radius of the face of the pin; each
path possessing different degrees of cross-shear. Those points on
the pin surface nearer the centre will travel in a more linear
fashion, be subjected to limited multi-directionality and trace
paths with higher aspect ratios. High AR wear paths have been
found to cause lower wear rates in conventional UHMWPE [33,34]
which may explain the reduced difference between wear of non-
crosslinked and crosslinked material on this machine. Galvin et al.
report a wear factor of 610
8
mm
3
/Nm for 100kGy material
experiencing 601 of pin rotation against a smooth counterface.
This is an order of magnitude lower than the elliptical wear factor
of 1.310
7
mm
3
/Nm found in this paper. However, they also
report a wear factor of 2.210
7
mm
3
/Nm for their unirradiated
material which is an order of magnitude lower than the value
reported herein and that reported from clinical studies. From this
study, a wear factor of 1.610
6
mm
3
/Nm for unirradiated
polyethylene under elliptical motion is in good agreement with
published clinical wear rates of 2.110
6
mm
3
/Nm [61]. Conven-
tional UHMWPE will typically be sterilised to 30kGy. The modest
level of crosslinking achieved at these radiation doses do not
provide the dramatic wear resistance offered by heavily irradiated
UHMWPE. Our non-crosslinked material would, therefore be
expected to show similar wear behaviour to clinical data, and is in
good agreement with the wear of conventional material (Table 2).
Hip simulator studies have also shown signicant wear savings
in crosslinked material. Mckellop et al. found an 87% reduction in
wear rate of UHMWPE gamma irradiated at 95kGy compared to
cups machined from bulk material irradiated at 33kGy [15]. This
is in good agreement with the 93% decrease in wear rate of +PE
reported in this study for samples subjected to elliptical sliding.
This level of agreement between the advanced pin-on-plate
tribometer described in this study and multi-directional joint
simulators conrms the value of the further developments in
tribometer design during the past decade.
A contemporary pin-on-disc tribometer, designed by Saikko
et al. has been used to investigate the effects of surface roughness
on wear of conventional and highly crosslinked UHMWPE.
Conventional UHMWPE was gamma-sterilised in nitrogen to
2540kGy and the highly crosslinked material, electron beam
irradiated to 95kGy [13]. Wear tests were conducted using a
circulatory translating pin-on-disc device (CTPOD). Wear data
were correlated to counterface surface roughness using a power
regression expression. The wear factor of conventional and
crosslinked material was expressed as k 5.8710
5
(Ra)
0.91
and k 7.8710
5
(Ra)
2.49
, respectively. On polished surfaces
equivalent to surfaces used in the current study (mean Ra
0.01mm), wear factors for conventional material of
k 1.210
6
mm
3
/Nm and for crosslinked, 1.910
9
mm
3
/Nm
can be calculated. The machine reported herein produces a
constantly changing shear vector throughout the elliptical cycle,
uniform across the pin surface. The CTPOD device was one of the
rst desktop machines to exert these tribological conditions in a
pin-on-disc device, doing so through a circular motion at the pin-
plate interface where, the pin remains stationary relative to the
ARTICLE IN PRESS
Fig. 15. Scanning electron micrographs after three million cycles for: (a) +PE control surface; (b) +PE linear weardomains of preferential alignment observed; (c) +PE
elliptical weardomains of alignment observed; (d) +PE control surface; (e) +PE linear weartexture development and lamellar fragmentation; and (f) +PEelliptical
motiondomains of fragmented lamellae. White arrows indicate axis of primary wear path.
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 9
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
plate. Therefore, both contemporary machines account for
the strain hardening behaviour of UHMWPE, producing multi-
directional friction forces on the polyethylene pin surface.
However, the crosslinked wear factor from Saikkos study is three
orders of magnitude lower than that for their conventional
material and two orders of magnitude lower than the crosslinked
wear factor reported in this study; 1.910
9
mm
3
/Nm vs.
1.310
7
mm
3
/Nm. This difference may be accounted for by
the weaker roughness and wear factor correlation found for
crosslinked material, particularly with a surface nish below a
value of around Ra 0.03mm. Under this surface roughness they
report a wear factor similar in magnitude to our study. They
further found a mean wear debris size generated from crosslinked
material articulating on smooth counterfaces smaller than those
particles generated from conventional material, 0.170.04 and
0.2170.09mm, respectively. At surfaces articially roughened to
approximate damaged femoral components, the mean difference
in particle size was not signicant; both materials produced mean
particle sizes approaching 0.38mm. The roughened surfaces, with
scratches in no preferential direction, may promote an abrasive
wear regime, which would dominate over an otherwise adhesive/
fatigue regime, shown by higher polymer wear rates. The effects
of surface roughness on wear were found to be more pronounced
in the crosslinked polyethylene, illustrated by a steeper regression
gradient. This is not surprising if we consider wear behaviour in
light of bulk polymer mechanical properties. The toughness of
crosslinked material has been shown to be less than that of
conventional material; ultimate tensile strength and strain are
degraded in highly crosslinked material [54]. A well known
relationship between polymer bulk mechanical properties and
resistance to abrasive wear is called the RatnerLancaster
relationship [62]. This correlates polymer wear to be inversely
proportional to the toughness of the polymer (the product of the
ultimate stress and strain at tensile break), supporting the greater
dependence of crosslinked material on surface roughness found in
the Saikko study.
Under rougher counter-surface conditions which result in
higher crosslinked wear rates, the polymer surface is worn more
quickly; a culmination of ploughing and fracture producing wear
debris before large plasticity levels and fatigue become the
dominant wear mechanism. However, on polished counterfaces
promoting an adhesive type wear regime, crosslinked material
may accumulate larger levels of plastic strain in the near surface
with respect to conventional material, due to a reduction in
surface turn-over. The reduction can be attributed to additional
cross-linked bonds increasing the cohesive strength of the
material. As a result, larger strain accumulation may increase
the occurrence of deformation mechanisms in the crystalline
domains causing lamellae fragmentation. De-crystallisation may
result in the generation and release of smaller wear particles than
those released from conventional material.
The evolution of microstructural change in the polyethylene is
of great inuence to the wear properties of this material. Edidin
et al. [20] presented transmission electron micrographs taken
from stained virgin and cross-linked acetabular cups after three
million cycles on a hip simulator. Sections taken to map the micro-
structure beneath the surface revealed lamellae re-orientation to a
sub-surface depth of 9 and 4mm for non-crosslinked and cross-
linked material, respectively. This modied sub-surface zone was
labelled the plasticity induced layer. In order to characterise the
depth at which the sub-surface plasticity layer extends, we
conducted depth proling using a confocal Raman microscope.
The Raman results presented here represent the percentage of
crystalline material at the surface and sub-surface after three
million linear and elliptical wear cycles. Unirradiated material,
worn under elliptical motion was shown to retain its original
unworn crystallinity up to the investigated depth of 25mm. This
indicates that deformation mechanisms for PE, in response to
multi-directional motion, may be constrained to the amorphous
domains, leaving the majority of lamellae undamaged (Fig. 13b or
e) and, hence, percentage crystallinity unchanged. In contrast,
under linear reciprocating motion the lower wearing PE was
found to experience de-crystallisation. FE-SEM images yield
evidence of crystalline fragmentation, with the appearance of
smaller crystals on the worn microstructures (Fig. 13c or f). The
accumulation of strain at, and beneath, the surface may result in a
transfer of the deformation mechanism from acting preferentially
within the amorphous domains to occurring, in addition, within
the crystalline lamellae. Although a drop in unirradiated crystal-
linity was found in response to reciprocating motion (in compar-
ison to the unworn control and elliptical motion), we observed
no statistical change in mean crystallinity between the surface
and sub-surface within this group. This suggests that the de-
crystallisation zone extends beyond the maximum sampled depth
of 37.5mm. Had the extent of sub-surface damage been less than
this, an increase in crystallinity with increasing depth should have
been observed as the microstructure returns to its native
conguration being no longer affected by the frictional forces at
the surface.
Edidin et al. suggest that cross-linking may constrain the
ability of the crystalline lamellae to re-orient resulting in a
smaller plasticity induced damage layer and lower wear rates. In
this study crosslinking has been shown to produce noticeable
reductions in wear rate under elliptical motion. Raman spectro-
scopy indicates that in response to sliding, both surface and sub-
surface crystallinity is reduced for +PE specimens. Under elliptical
motion, the nal surface crystallinity was lower than the unworn
control. This was supported by micrographs showing evidence of
areas consisting of fragmented lamellae. Linear motion caused the
largest drop in +PE crystallinity. Crosslinked material worn with
either a linear or elliptical motion exhibits a trend of increasing
crystallinity with depth beneath the surface. This suggests the
extent of surface stresses and strains diminishes as the distance
beneath the surface increases in agreement with contact
mechanics theorems e.g. Hertz. However, neither microstructure
approaches the crystallinity ratio of the unworn crosslinked
control material. This further indicates that the de-crystallised
damage layer may extend further than the maximum depth of
37.5mm explored in this study. The results presented here indicate
that crosslinked UHMWPE experiences a sub-surface damage
layer which contains re-oriented and fragmented lamellae
regardless of motion path. Under both types of motion, a
de-crystallisation regime was found to extend more than
37.5mm beneath the surface. More wear resistant surfaces may
develop higher levels of strain; the increased period taken for a
given amount of material to be removed results in a greater
chance for strain to accumulate. It was apparent that the slower
wearing surfaces, from either material group, exhibited the
greatest de-crystallisation at the completion of each wear test.
Broken lamellae, particularly in the highly wear resistance
crosslinked material may be nally released as sub-micron debris.
Further studies, indicate a larger sub-surface zone which may
extend up to and exceed 200mm beneath the surface [63]. Dryzek
et al. conducted pin-on-disc studies rubbing 10mm diameter
pellets of UHMWPE onto a stainless steel disc with nominal loads
of 100 and 150N. Sliding was in the absence of lubricant. X-ray
diffraction indicated 100m of sliding was enough to cause a
de-crystallised surface and sub-surface layer, well beyond 200mm
into the material. Clearly, the likely transfer-lm formation
and heat build-up experienced in Dryzek et al.s study disallow
a direct comparison. However, it does serve to illustrate that
the uni-directional nature of the pin-on-disc will, just like the
ARTICLE IN PRESS
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 10
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
linear-reciprocating motion applied here, cause strain hardening
and preferential lamellar alignment at substantial depths beneath
the surface beyond those expected from a consideration of the
contact mechanics. Under these simplied tribological conditions
we also nd a de-crystallised worn surface and sub-surface layer
in the unirradiated UHMWPE.
Davey et al. [44] conducted polarised Fourier-transform
infrared spectroscopy (FTIR) on retrieved UHMWPE acetabular
components. Cylindrical cores were taken perpendicular to the
worn surface and then sectioned on a microtome into 200mm
thick slices. They found the greatest amount of lamellar orienta-
tion occurring in the rst 200mm beneath the surface, although
microstructural re-arrangement was also apparent up to 400mm.
The Raman technique employed within this study indicates a sub-
surface damage layer consisting of a de-crystallisation zone
extending to a depth greater than 37mm. The only material and
wear path combination that showed no changes in surface
crystallinity under sliding was the unirradiated material worn
under elliptical motion.
Parallel ripples on the surface of high wear areas of explanted
UHMWPE cups were rst observed by Dowling et al. [64] using
SEM. The limitation of the ripples to occur in areas of the cup
exhibiting high wear suggests a contact area subjected to a high
degree of multi-directionality and the presence of adhesive or
fatigue wear mechanisms. Ripples from Dowlings study, this
study, and many more consistently report the feature with
wavelengths between 1 and 2mm. We imaged rippling on the
worn surface of unirradiated polyethylene under elliptical rubbing
(Fig. 12a), a tribological setup resulting in the highest wear. This
may suggest that rippling is a consequence of multi-directional
motion under dynamic loading, however, surface ripples have also
been recorded by Elck et al. [27] who suggests motion
approaching uni-directionality may be a requirement for this
topographical feature. In agreement, Bragdon et al. [65] found
only uniaxial reciprocating motion produced signicant ripples of
wavelength 1.5mm on the worn surface of non-irradiated 4150
UHMWPE pins. We also imaged rippling on the worn surface of
+PE subjected to linear motion (Fig. 14d). Due to the inconsistent
tribological conditions under which sightings of rippling have
been observed, the suggestion by Elck et al. that rippling may be
a transient phenomenon seems appropriate.
Uni-directional motion has been found to cause the greatest
degree of molecular orientation under a range of sliding motions
[66]. Sambasivan et al. also found signicant alignment in
response to linear reciprocating wear during a 5000 cycle test
period. In the current study microstructural re-orientation was
conrmed by eld-emission scanning electron microscopy. Micro-
structure evolution in response to motion path was highly distinct
in the unirradiated group. Signicant texture development
occurred under linear reciprocating motion. Lamellae were
orientated in a preferential manner which coincided with the
sliding direction (Fig. 13c). Surfaces worn under elliptical motion
showed considerably less re-arrangement and negligible crystal-
line degradation. Both these material characteristics could be
attributed to the higher wear rate caused by elliptical motion. A
faster wearing surface will leave less time for large strains to
develop in the surface. This may result in a worn surface showing
no preferential orientation or fragmented lamellae. Likewise, the
same unirradiated material subjected to linear motion was found
to wear more slowly, hence, would have more time to accumulate
higher levels of surface and sub-surface strain, resulting in greater
amounts of orientation and crystalline degradation. This is
supported by SEM images showing more frequently occurring
lamellar tilting, bending, rotation and ultimately fragmentation in
ultra-structures subjected to linear motion. Crosslinked material,
the more wear resistant material under both types of motion may
experience a similar effect. The slower wear rates could explain
the drop in crosslinked crystallinity, and the similarities of
microstructure deformation imaged on both worn linear and
elliptical microstructures.
SEM images of PE and +PE ultrastructures after linear motion
were similar. Both showed degrees of lamellar alignment and
crystalline deformation mechanisms. The difference in the two
polymer networks was revealed by swelling results. These
indicated elevated crosslinks occurring in the amorphous domains
of the +PE material. The presence of crosslinks did not signi-
cantly inhibit lamellae re-orientation, and did not increase the
wear resistance of +PE under simplied reciprocating motion.
Therefore, we postulate the increased wear resistance of cross-
linked material comes from the elevated tie-molecule densities in
amorphous domains, adding increased resistance to amorphous
deformation mechanisms such as interlamellar stack rotation,
interlamellar shear and interlamellar separation [67] in response
to crossing motion paths. The greater resistance to plastic
deformation and debris release from the surface creates a more
wear resistant surface. The increased structural integrity of the
amorphous domain may more effectively transfer strain to the
crystalline domains, supporting the de-crystallised surface and
sub-surface zones found in this study. Under physiological hip
stresses, accumulated strain may not be relieved from the surface
as debris, as would be in conventional material, but rather
transferred from amorphous to crystalline domains increasing the
time it takes to generate debris particles. Fragmented lamellae
may then provide the basis for sub-micron debris generation.
5. Conclusions
The use of a novel six-station, programmable load and motion
wear machine to further the tribological investigation into wear
and sub-surface plasticity of ultra-high molecular weight poly-
ethylene has conrmed the importance of multi-directional
motion in pin-on-plate studies. Further, it has provided a basis
for non-destructive analysis techniques to characterise micro-
structure development. The signicance of uniform tribological
conditions across the pin contact area has allowed the wear
resistance of highly crosslinked polyethylene to be investigated.
Raman spectroscopy revealed a de-crystallised damage zone
beneath the worn surface extending to a distance greater than
37.5mm. This suggests the strain-induced damage zone is more
extensive than initially proposed.
Signicant lamellar alignment was only observed under linear
reciprocating motion, questioning the basis of orientation soft-
ening and accelerated wear under cross-shear. However, the
hypothesis that crosslinking does not inhibit lamellar mobility is
supported as SEM images revealed re-orientation under sliding.
The elevated number of carboncarbon covalent bonds may
increase the resistance to plastic deformation and brillar pull-
out when the microstructure shows no preferential alignment.
De-crystallised surface and sub-surface zones were found in the
lowest wearing groups i.e. unirradiated and crosslinked material
subjected to reciprocating motion. This was attributed to large
strain accumulation over a longer period of time providing a
mechanism for crystalline fragmentation. Lamellae fragments
may form the basis of sub-micron debris from the more wear
resistant crosslinked surfaces.
Acknowledgements
The authors would like to thank the Engineering and Physical
Sciences Research Council for funding the project, Smith and
ARTICLE IN PRESS
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 11
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
Nephew Inc. for supplying the samples, The Royal Society of
Edinburgh, Prof. Lisa Pruitt at the University of California, Berkeley
and Dr. Chris Jeffree for FESEM work at the University of
Edinburgh.
References
[1] Galante JO, Rostoker W. Wear in total hip prosthesesexperimental
evaluation of candidate materials. Acta Orthopaedica Scandinavica Supple-
mentum 1973(145):146.
[2] McKellop H, et al. Wear characteristics of UHMW polyethylene: a method for
accurately measuring extremely low wear rates. Journal of Biomedical
Materials Research 1978;12(6):895927.
[3] McKellop H, et al. Friction and wear properties of polymer, metal, and ceramic
prosthetic joint materials evaluated on a multichannel screening device.
Journal of Biomedical Materials Research 1981;15(5):61953.
[4] Rose RM, Cimino WR. Exploratory investigations on the structure dependence
of the wear resistance of polyethylene. Wear 1982;77(1):89104.
[5] Rostoker W, Galante JO. Contact pressure dependence of wear rates of ultra
high molecular weight polyethylene. Journal of Biomedical Materials
Research 1979;13(6):95764.
[6] Wright KWJ, Dobbs HS, Scales JT. Wear studies on prosthetic materials using
the pin-on-disc machine. Biomaterials 1982;3(1):418.
[7] Charnley J. Total hip replacement by low-friction arthroplasty. Clinical
Orthopaedics and Related Research 1970(72):721.
[8] Cooper RA, et al. Polyethylene debris-induced osteolysis and loosening in
uncemented total hip arthroplasty: a cause of late failure. The Journal of
Arthroplasty 1992;7(3):28590.
[9] Harris WH. The problem is osteolysis. Clinical Orthopaedics and Related
Research 1995(311):4653.
[10] Willert HG, Bertram H, Buchhorn GH. Osteolysis in alloarthroplasty of the
hipthe role of ultra-high-molecular-weight polyethylene wear particles.
Clinical Orthopaedics and Related Research 1990(258):95107.
[11] Bourne RB, et al. Pain in the thigh following total hip-replacement with a
porous-coated anatomic prosthesis for osteoarthrosisa 5-year follow-up
study. Journal of Bone and Joint SurgeryAmerican Volume 1994;76A(10):
146470.
[12] Muratoglu OK, et al. Unied wear model for highly crosslinked ultra-high
molecular weight polyethylenes (UHMWPE). Biomaterials 1999;20:146370.
[13] Saikko V, Calonius O, Kera nen J. Effect of counterface roughness on the wear
of conventional and crosslinked ultrahigh molecular weight polyethylene
studied with a multi-directional motion pin-on-disk device. Journal of
Biomedical Materials Research 2001;57(4):50612.
[14] Galvin A, et al. Wear of crosslinked polyethylene under different tribological
conditions. Journal of Materials Science: Materials in Medicine 2006;
17(3):23543.
[15] McKellop H, et al. Development of an extremely wear-resistant ultra high
molecular weight polythylene for total hip replacements. Journal of
Orthopaedic Research 1999;17(2):15767.
[16] Dorr LD, et al. Clinical performance of a durasul highly cross-linked
polyethylene acetabular liner for total hip arthroplasty at ve years. Journal
of Bone and Joint SurgeryAmerican Volume 2005;87A(8):181621.
[17] Heisel C, et al. Short-term in vivo wear of cross-linked polyethylene. Journal of
Bone and Joint SurgeryAmerican Volume 2004;86A(4):74851.
[18] Dumbleton JH, Shen C, Miller EH. Study of wear of some materials in
connection with total hip-replacement. Wear 1974;29(2):16371.
[19] Shen C, Dumbleton JH. Friction and wear behavior of irradiated very high
molecular-weight polyethylene. Wear 1974;30(3):34964.
[20] Edidin AA, et al. Plasticity-induced damage layer is a precursor to wear in
radiation-cross-linked UHMWPE acetabular components for total hip
replacement. The Journal of Arthroplasty 1999;14(5):61627.
[21] Mckellop HA, et al. The origin of submicron polyethylene wear debris in total
hip-arthroplasty. Clinical Orthopaedics and Related Research 1995(311):320.
[22] Jasty M, et al. Wear of polyethylene acetabular components in total hip
arthroplastyan analysis of one hundred and twenty-eight components
retrieved at autopsy or revision operations. Journal of Bone and Joint
SurgeryAmerican Volume 1997;79A(3):34958.
[23] Bragdon CR, et al. The importance of multidirectional motion on the wear of
polyethylene. Proceedings of the Institution of Mechanical Engineers [H]
1996;210(3):15765.
[24] Wang A, et al. Orientation softening in the deformation and wear of ultra-
high molecular weight polyethylene. Wear 1997;203204:23041.
[25] Charnley J. Wear of plastics materials in hip-joint. Plastics and Rubber 1976;
1(2):5963.
[26] Wang A. A unied theory of wear for ultra-high molecular weight
polyethylene in multi-directional sliding. Wear 2001;248(12):3847.
[27] Elck APD, et al. A re-appraisal of wear features of acetabular sockets using
atomic force microscopy. Wear 2002;253(78):83947.
[28] Deboer J, Vandenberg HJ, Pennings AJ. Crosslinking of ultrahigh molecular-
weight polyethylene in the oriented state with dicumylperoxide. Polymer
1984;25(4):5139.
[29] Kurtz SM, et al. Radiation and chemical crosslinking promote strain
hardening behavior and molecular alignment in ultra high molecular weight
polyethylene during multi-axial loading conditions. Biomaterials 1999;
20(16):144962.
[30] Wang A, et al. Lubrication and wear of ultra-high molecular weight
polyethylene in total joint replacements. Tribology International 1998;
31(13):1733.
[31] Joyce TJ, et al. A comparison of the wear of cross-linked polyethylene against
itself under reciprocating and multi-directional motion with different
lubricants. Wear 2001;250(112):20611.
[32] Zhou J, et al. Tribological and nanomechanical properties of unmodied and
crosslinked ultra-high molecular weight polyethylene for total joint replace-
ments. Journal of Tribology 2004;126(2):38694.
[33] Saikko V, Calonius O, Kera nen J. Effect of slide track shape on the wear of
ultra-high molecular weight polyethylene in a pin-on-disk wear simulation of
total hip prosthesis. Journal of Biomedical Materials Research 2004;
69B(2):1418.
[34] Turell M, Wang A, Bellare A. Quantication of the effect of cross-path motion
on the wear rate of ultra-high molecular weight polyethylene. Wear
2003;255(712):10349.
[35] ASTM D2765-01. Standard test methods for determination of gel content and
swell ratio of crosslinked ethylene plastics. ASTM International, 2006.
[36] Rodriguez F, et al. Principles of polymer systems. 5th ed. London: Taylor &
Francis; 2003.
[37] Flory PJ. Principles of polymer chemistry. Ithaca, NY: Cornell University Press;
1953.
[38] ASTM F732-00. Standard test method for wear testing of polymeric materials
for use in total joint prostheses. ASTM International, 2006.
[39] Kilgour A, Elck AP. Investigation of polyethylene structure and its bearing on
wear of joint replacments. In: World biomaterials congress, Amsterdam, 2008.
[40] Hall RM, et al. Wear in retrieved Charnley acetabular sockets. Proceedings of
the Institution of Mechanical Engineers 1996;210(3):197207. [0954-4119
(Print)].
[41] Ramamurti BS, et al. Loci of movement of selected points on the femoral head
during normal gait: three-dimensional computer simulation. The Journal of
Arthroplasty 1996;11(7):84552.
[42] Bennett DB, Orr JF, Baker R. Movement loci of selected points on the femoral
head for individual total hip arthroplasty patients using three-dimensional
computer simulation. The Journal of Arthroplasty 2000;15(7):90915.
[43] Saikko V, Calonius O. Slide track analysis of the relative motion between
femoral head and acetabular cup in walking and in hip simulators. Journal of
Biomechanics 2002;35(4):45564.
[44] Davey SM, et al. Measurement of molecular orientation in retrieved ultra-
high-molecular-weight polyethylene (UHMWPE) hip sockets using Fourier-
transform infrared spectroscopy. Strain 2004;40(4):20310.
[45] Bergmann G, Graichen F, Rohlmann A. Hip joint loading during walking and
running, measured in two patients. Journal of Biomechanics 1993;26(8):
96990.
[46] Paul JP. Forces transmitted by joints in the human body. Proceedings of the
Institution of Mechanical Engineers 1966;181:815.
[47] Glotin M, Mandelkern L. A Raman spectroscopic study of the morphological
structure of the polyethylenes. Colloid & Polymer Science 1982;260(2):
18292.
[48] Mandelkern L, Alamo RG, Kennedy MA. Interphase thickness of linear
polyethylene. Macromolecules 1990;23(21):47213.
[49] Strobl GR, Hagedorn W. Raman spectroscopic method for determining the
crystallinity of polyethylene. Journal of Polymer Science: Polymer Physics
Edition 1978; 118193.
[50] Affatato S, et al. The performance of gamma- and EtO-sterilised UHMWPE
acetabular cups tested under severe simulator conditions. Part 2: wear particle
characteristics with isolation protocols. Biomaterials 2003;24(22):404555.
[51] Affatato S, et al. Wear behaviour of cross-linked polyethylene assessed in vitro
under severe conditions. Biomaterials 2005;26(16):325967.
[52] Affatato S, et al. Effects of the sterilisation method on the wear of UHMWPE
acetabular cups tested in a hip joint simulator. Biomaterials 2002;23(6):
143946.
[53] Bertoluzza A, et al. Micro-Raman spectroscopy for the crystallinity character-
ization of UHMWPE hip cups run on joint simulators. Journal of Molecular
Structure 2000;521(13):8995.
[54] Pruitt LA. Deformation, yielding, fracture and fatigue behavior of conven-
tional and highly cross-linked ultra high molecular weight polyethylene.
Biomaterials 2005;26(8):90515.
[55] Olley RH, Hodge AM, Bassett DC. A permanganic etchant for polyolenes.
Journal of Polymer Science: Polymer Physics Edition 1979;17(4):62743.
[56] Sawyer LC, Grubb DT. Polymer microscopy. 2nd ed. New York: Chapman &
Hall Ltd; 1987.
[57] Olley RH, Bassett DC. An improved permanganic etchant for polyolenes.
Polymer 1982;23(12):170710.
[58] Shahin MM, Olley RH, Blissett MJ. Renement of etching techniques to reveal
lamellar proles in polyethylene banded spherulites. Journal of Polymer
Science Part B: Polymer Physics 1999;37(16):227986.
[59] Klapperich C, Komvopouolos K, Pruitt L. Tribological properties and micro-
structure evolution of ultra-high molecular weight polyethylene. Journal of
Tribology, Transactions of the ASME 1999;121(2):394403.
[60] Archard JF. Contact and rubbing of at surfaces. Journal of Applied Physics
1953;24(8):9818.
[61] Elck APD, et al. The effect of socket design, materials and liner thickness on
the wear of the porous coated anatomic total hip replacement. Proceedings of
ARTICLE IN PRESS
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 12
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011
the Institution of Mechanical Engineers Part HJournal of Engineering in
Medicine 2001;215(H5):44757.
[62] Lancaster JK. Relationships between wear of polymers and their mechanical
properties. Industrial Lubrication and Tribology 1969;21(7):214.
[63] Dryzek E, Dryzek J. Measurement of subsurface zone in UHMWPE after
sliding against stainless steel using the new experimental method DSIP.
Radiation Physics and Chemistry 2007;76(2):1579.
[64] Dowling JM, et al. Characteristics of acetabular cups worn in human-
body. Journal of Bone and Joint SurgeryBritish Volume 1978;60(3):37582.
[65] Bragdon CR, et al. A new pin-on-disk wear testing method for simulating
wear of polyethylene on cobaltchrome alloy in total hip arthroplasty. The
Journal of Arthroplasty 2001;16(5):65865.
[66] Sambasivan S, et al. Molecular orientation of ultrahigh molecular weight
polyethylene induced by various sliding motions. Journal of Biomedical
Materials Research 2004;70B(2):27885.
[67] Lin L, Argon AS. Structure and plastic-deformation of polyethylene. Journal of
Materials Science 1994;29(2):294323.
ARTICLE IN PRESS
A. Kilgour, A. Elck / Tribology International ] (]]]]) ]]]]]] 13
Please cite this article as: Kilgour A, Elck A. Inuence of crosslinked polyethylene structure on wear of joint replacements. Tribol Int
(2009), doi:10.1016/j.triboint.2008.11.011

S-ar putea să vă placă și