Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Building Bones

Young Hwa Kim, Wilson Tang, Meghan Tighe


December 16, 2015

Introduction

When damaged or diseased bone cannot support the body, artificial bones may be used to replace
the damaged bone. Hip surgery is an especially common procedure that is performed by orthopedic
surgeons to replace diseased or damaged hips. This is particularly relevant now as the nation has a
huge aging population that often have osteoarthritis or other conditions.
Damaged bone can be replaced with bone from other parts of the body, from cadavers, or with
various ceramics or metallic alloys. The replacement must satisfy some conditions: no immune
reaction, easily combined with existing tissues, less corrosion, nontoxic products and have similar
properties and functionality of real bones.
The characteristics of the specific materials used in these replacements, such as tensile strength,
compressive strength, elasticity, fatigue, structure, and hardness, directly determine the safety and
usefulness of the replacement.
We seek to compare the relevant mechanical properties of three materials frequently used in hip
replacements with the mechanical properties of the bones that they replace[20]. The three most
commonly used metals in these replacements are stainless steels, titanium and titanium alloys, and
cobalt-base alloys. Our selected materials include 304 stainless steel and a titanium alloy (Ti6Al4V),
as well as ultra high molecular weight polyethylene (UHMWPE), typically used as a polymeric socket
that interfaces with the metal [11]. We will also test the mechanical properties of bone, specifically
a femur. For each of these we will look at experimentally derived values as well as those reported
in literature. We will also briefly discuss other factors that we have been unable to test ourselves,
such as biocompatibility.

1.1

Bone As a Material

Bone is a composite made of approximately 70% mineral, 5% to 8% water and 22%-25% organic
matrix. These components form a three-dimensional network that provides mechanical integrity for
locomotion and is associated with mineral homeostasis. Bones mechanical properties are composed
of its various levels of hierarchical structural organization: macrostructure, microstructure, submicrostructure, and nanostructure. Within the femur specifically, the epiphysis is mainly composed
of trabecular bone while the diaphysis is mainly composed of cortical bone (Figure 2). Both trabecular and cortical bone are made of the same material but differ in porosity, mineralization, and
organization of the solid matrix[22]. Structurally, trabecular bone acts as a reinforcing cancellous
network within the femur and is softer, weaker, and more flexible than cortical bone. Cortical bone
is much more compact and is harder, stronger, and stiffer than trabecular bone.
Throughout life, bone is constantly renewed; old bone is removed as new bone is formed. As
shown in Figure 1, bone resorption is the process in which the osseous tissue releases its mineral and
1

Figure 1: Bone Resorption Cycle [16]


goes through the process of breaking down, resulting in loses. Osseous tissue, one of the supporting
tissues, is involved in movement, maintaining the homeostasis of calcium, and maintaining the cells
function. To maintain the function of osseous tissue, there needs to be a cycle of destruction and
renaturation.

1.2

The Hip Joint

The hip is a structurally important ball and socket joint in the human leg system that experiences
both tensile and torsional forces during usage. Figure 2 details the maximum tensile test lines for a
femur.

Figure 2: A)Maximum Tensile Stress Lines of a Human Femur[1] B)Reference Human Femur Diagram [2]
2

The weight of the body exerts force(s) on the upper epiphysis which transmits the force(s) down
through the diaphysis to the lower epiphysis and to the rest of the leg. The femoral head fits into
the acetabulum as a ball and socket joint and experiences compression and rotational grinding as it
transfers load. The diaphysis experiences tensile and compressive forces transmitted from the upper
epihysis.

1.3

Hip Replacements

Joint replacements must be able to mirror the behavior of the tissue they are replacing as closely as
possible. Hip replacements are generally consist of replacing the upper femur and the mating pelvis
area. As shown in Figure 3, typical hip replacements consist of the femoral stem, a femoral ball, a
polymeric socket with or without a metallic backing, and a metal acetabular shell [11].

Figure 3: Hip Replacement Components [11]


The femoral component of the replacement must mimic the properties and behavior of the femur.
In a typical hip replacement (Figure 3), most of upper epiphysis is replaced and then anchored to the
femur inside the diaphysis (Figure 3). This configuration attempts to mimic the weight distribution
of femurs. Weight will be applied from the joint at an angle, resulting in a bending force which
is then transmitted along the shaft of the metal implant to the diaphysis and eventually the lower
epiphysis. Here the mechanical properties of the specific replacement materials play an important
role in the functionality of the implant.

1.4

Biocompatibility

A critical factor to be considered when choosing materials for medical implants is biocompatibility.
This term encompasses the way that a device interacts with the body of the patient and its ability
to elicit an appropriate host response [11]. Biocompatibility is inherently situational because different responses can be expected from the same material in different use cases; a material may be
biocompatible for one use in the body but not for another.
For a permanent implant, necessary tests include cytotoxicity, sensitization, irritation or intracutaneous reactivity, systemic toxicity, subacute and subchronic toxicity, genotoxicity, implantation,
chronic toxicity, and carcinogenicity [10].
3

2
2.1

Materials and Methods


Origin of Materials

For the bone sample, a cow femur was purchased from the Blood Farm in West Groton, MA. The
femur was kept frozen at -23o C for up to three weeks. Over this time, test specimens from the femur
were removed using a diamond bandsaw. After being defrosted, the bone samples were kept in a
cold room at 1.6o C, for up to five days, in paper towels soaked in physiological saline.
The other three materials were purchased from McMaster.The titanium alloy sample was a bar of
high-strength grade 5 titanium (Ti6Al4V). The UHMWPE sample was an impact-resistant ultrahigh
molecular weight polyethylene sheet. The stainless steel sample was a multipurpose 304 stainless
steel sheet.

2.2

Composition of Materials

The composition of each material was confirmed experimentally. Fourier-transform infrared spectroscopy (FTIR) was used for the UHMWPE and X-ray fluorescence (XRF) was used for the metals.
We performed the FTIR with a Bruker Vertex 70 Fourier Transform Infrared Spectrometer in attenuated total reflection (ATR) mode. For the XRF, we used an Innov-X Systems Energy Dispersive
X-Ray Fluorescence.
To compare the experimental results with the literature values, we referenced ASM Handbook.

2.3

Mechanical Testing

To measure the mechanical properties of each material, three-point bending and tensile testing were
utilized.
We tensile testing of multiple reduced cross-section tensile test specimens of steel and titanium
alloy with an Instron 5582 Universal Testing Machine. Each sample was pulled until fracture with
a transducer capable of producing force up to 100,000N. UHMWPE was tested in the same fashion
using a different testing machine: 3345 Universal Testing Machine with an Instron 2519-107 force
transducer, which was capable of producing force up to 5000 N. The dimensions (Table 1) and
testing rates for each material were scaled from the standards given by the American Section of the
International Association for Testing and Materials (ASTM) for tensile testing of metals [8]. The
sample pieces for these materials were manufactured using a water jet cutter.

Figure 4: Dimensions used in tensile testing

L
C
T
B
W
R

Average Tensile Testing Dimensions (mm)


ASTM Standard Steel Titanium UHMWPE
200
76.4
50.8
101.6
20
7.6
5.1
10.2
12.5
0.04
3.175
6.35
50
0.05
12.7
25.4
12.5
0.74
3.175
6.35
12.5
0.74
3.175
6.35

Table 1: Dimensions of samples used in tensile testing for stainless steel, titanium alloy and
UHMWPE. All measurements are in mm. Descriptions of each dimension can be seen in Figure 4.
The rates used for tensile testing of the metals were estimated from within the appropriate range
given by ASTM [8]: 0.2%-2%. The rates for titanium alloy Ti6Al4V and stainless steel 304 were 1
mm/min and 1.5 mm/min respectively. For tensile testing of the UHMWPE, a rate of 20mm/min
was used.
Three-point bend testing was conducted on the bone. Samples were prepared by sanding a piece
of bone to produce edges that were as even and smooth as possible. The exact dimensions varied
but each rectangular sample was near a size scaled from the ASTM guidelines: 55mm long, 5 mm
wide and 3.5mm thick [8]. The three-point bend tests were then run for each of three samples at a
rate of 3 N/min.
Stress and strain values were calculated from the collected raw tensile and flexural data, which
was in units of displacement (mm) and force (N). The dimensions of the tested samples were used to
convert force and deformation data into stress and strain curves. From the stress-strain curves, the
modulus and fracture stress were calculated for each sample and then averaged to get experimental
values for each material.
= F/A
l
=
lo

E=

3F L
f =
2wd2
6ld
f =
L2

L3 ff
Ef =
4wd3

(1)
(2)
(3)
(4)
(5)
(6)

The hardnesses of the metals and the UHMWPE were measured using the Rockwell Instron
hardness tester. Scales C, D, and R were used for the titanium, steel and UHMWPE respectively.
For each of the three materials, three tests were done and the values were averaged to find a mean
value.
The hardness of the bone was tested with a Buehler Micromet 5103 Microindentation Hardness
Tester. A sample was prepared by sanding and polishing the testing surface. A force of 25 gf was
used. Ten hardness measurements were taken and the values were averaged to find a mean hardness
value.
5

2.4

Imaging

Each materials fracture surface was imaged using scanning electron microscopy (SEM) to get a
greatly magnified image of the fracture surface. All SEM images were taken on a 6060LV Scanning
Electron Microscope (JEOL Ltd, Japan). Samples were prepared for SEM by cutting the entire
fracture surface from the fracture ends of tested samples then mounted.
A cross-sectional sample (6mm x 6mm x 0.1mm) of bone was also imaged using SEM in Back
Scattered Electron (BSE) mode to investigate the mineralization of bone through observation of
microsctructure. This was also done using the 6060LV scanning electron microscope (JEOL Ltd,
Japan).

3
3.1
3.1.1

Results
Composition
Titanium

Results from the XRF test confirmed that the material in question was likely Ti6Al4V. The exact
composition is shown in Table 2. The literature values are from the ASM Handbook. The amounts of
titanium and vanadium are near the reported literature values. Trace amounts of iron and zirconium
were detected as well. Aluminium was not found in the sample.
Elements
Ti
V
Fe
Zr
Al

Experimental Percentage [%]


93.52 1.50
5.86 0.50
0.59 0.14
0.03 0.01
0

Literature Percentage[%]
90
4
0
0
6

Table 2: Measured composition for the titanium alloy by XRF compared to literature values [3]

3.1.2

Stainless Steel

The XRF test also revealed the composition of the 304 stainless steel. Data is shown in Table 3. The
literature values are from the ASM Handbook. All elements which were reported by ASM to account
for more than 1% of the composition are shown in the XRF results. Copper and the molybdenum
were detected, but the carbon, silicon, phosphorus and sulfur were not found in this sample.
3.1.3

UHMWPE

R AnyWareTM was used to analyze the FTIR results. It


The spectral analysis website KnowItAll
matches the absorbance graph of the sample with stored graphs of known materials. From these
comparisons, the sample of UHMWPE shows a 96% match to petrothene NL 409-00, a medium
density polyethylene.

Elements
Cr
Mn
Fe
Ni
Cu
Mo
C
Si
P
S

Experimental Percentage [%]


15.98 0.28
2.20 0.17
71.35 0.69
8.80 0.36
1.09 0.12
0.57 0.02
0
0
0
0

Literature Percentage[%]
18.0-20.0
2.00
66.35-70.85
8.0-10.5
0
0
0.08
1.00
0.045
0.03

Table 3: Measured composition for the stainless steel by XRF compared to literature values [4]

Figure 5: FTIR results of UHMWPE material show a 96% match to petrothene NL 409-00, a medium
density polyethylene.

3.2
3.2.1

Structural Tests
Tensile Strength and Youngs Modulus

Graphs indicating results of the tensile testing of titanium, stainless steel, and UHMWPE are shown
in Figures 6, 7, and 8 respectively.

Material
Titanium
Steel
UHMWPE
Bone

Comparing Youngs Modulus


Measured Value error[GPa] Literature Value[GPa]
17.710 0.564
113.8
16.8355 3.07
205
.36297 0.24
0.5
9.28 22
13.5

Table 4: Measured and literature values [3] [7] of elastic modulus for various materials

Comparing Ultimate Tensile Strength


Measured Value error[MPa] Literature Value[MPa]
970 34
993
689 34
515
22.8 4.8
39
397 50
247

Material
Titanium
Steel
UHMWPE
Bone

Table 5: Measured and literature values [3] [4] [6] of elastic modulus for various materials

Tensile Testing Data for Titanium Samples


1000
Run 1, Max 968.52 MPa
Run 2, Max 982.11 MPa
Run 3, Max 960.35 MPa

900
800

Stress (MPa)

700
600
500
400
300
200
100
0

0.05

0.1

0.15
Strain (mm/mm)

0.2

0.25

Figure 6: Tension test data of Ti6Al4 samples.

Tensile Testing Data for Steel 301 Samples


800

Run 1, Max 6.7576*10 GPa


Run 2, Max 7.0348*105 GPa

700

600

Stress (GPa)

500

400

300

200

100

100

0.1

0.2

0.3

0.4
0.5
Strain (mm/mm)

0.6

0.7

0.8

0.9

Figure 7: Tension test data of 304 stainless steel samples.

Tensile Testing Data for Ultra High Molecular Weight Polyethylene


25

Stress (MPa)

20

15

Run 1, Max 23.081 MPa


Run 2, Max 23.420 MPa
Run 3, Max 21.996 MPa

10

0.5

1.5

2.5

Strain (mm/mm)

Figure 8: Tension Test Data of UHMWPE samples .

3 Point Bend Testing Data for Bone


600
Run 1, Max 403.25 MPa
Run 2, Max 418.41 MPa
Run 3, Max 534.66 MPa
Run 4, Max 343.08 MPa
Run 5, Max 289.97 MPa

500

Flexural Stress MPa)

400

300

200

100

100

0.5

1
1.5
Flexural Strain (mm/mm)

2.5
3

x 10

Figure 9: 3-Point bend test data of cow femur samples.

3.2.2

Imaging

Figure 10 shows the microstructure of the bone captured with SEM. The black holes and the grain
are shown. The straight, near vertical lines are an artifact of our polishing, not the structure of
the bone. Since the bone was cut longitudinally, this image provides a cross section of the collagen
fibers in the bone.

Figure 10: Microstructure of bone captured with scanning electron microscopy in backscattered
electron mode.

Figure 11: A. Fractured stainless steel after tensile testing B. SEM image of fracture surface at 19
times magnification C. SEM image of other fracture surface at 20 times magnification.
Stainless steel fractured with very little necking into two pieces: a convex fracture surface (image
B in Figure 11) and a concave fracture surface (image C in Figure 11). As can be seen in Figure 11,
steel formed a cup-and-cone fracture [17].

10

Figure 12: Fractured Ti6Al4 after tensile testing.


Titanium experienced a moderate amount of necking before fracture. When looking at the
fracture surface in Figure 12, it can be seen that titanium formed a cup-and-cone fracture [17].

Figure 13: Tensile testing of UHMWPE included high amounts of elastic deformation. SEM images
of the fracture surfaces indicate failure was not a clean break.
As can be seen in Figure 8, UHMWPE experienced a large amount of necking. As seen in the
stretched sample Figure 13, the long, cross-linked polyethylene chains tore in a less organized manner
(fibers transverse to the tension broke before non-transverse orientated fibers)

11

Figure 14: SEM images of the two fracture sites of the bone after three point bend testing.
Tensile testing of bone samples was largely unsuccessful and consistently led to fracture within
the grips. Three point bend testing of bone, on the other hand, led to useful results, as demonstrated
in the stress-strain curve (Figure 9). Resultant fractures are shown in Figure 14.
3.2.3

Hardness

The average Vickers hardness value of the bone sample was 38.5 6.0. Rockwell hardness testing
for the titanium on a C scale indicated a hardness of C38.0 0.4. For stainless steel, the hardness
testing on a B scale indicated B93.7 2.9. The hardness testing result for the UHMWPE on a scale
R was R54.0 1.0.

Material
Titanium
Steel
UHMWPE
Bone

Comparing Hardness Values


Measured Value error Literature Value
C38.03 0.8
C30
B93.7 2.94
B79
R54.03 1.026
R40
Vickers 38.55 12.034
N/A

Table 6: Measured and literature values [3] [4] [6] of elastic modulus for Titanium Alloy (Ti6Al4V),
Stainless Steel 304, Ultra High Molecular Weight Polyethylene, and Bone

Discussion

When orthopaedic surgeons and medical implant companies choose materials for joint replacements
such as hip implants, they have to take many factors into consideration. These include mechanical
properties such as hardness, elastic modulus, and tensile strength as well as bio-compatibility. Ideally,
the implants would be composed of materials that mimic bones functional behavior with similar
mechanical properties. However, when choosing materials they have to make trade-offs between
properties in an attempt to create the best combination.
12

4.1

Composition

From the experiments, the composition of each material was confirmed.


UHMWPE was a 96% match for medium density polyethylene, confirming that the UHMWPE
is polyethylene.
The titanium alloy was composed of 93.52% titanium. The biggest inconsistency was a lack of
aluminum in the XRF results. We suspect that the XRF was unable to detect the aluminum since
it is a light element and we therefore conclude that the alloy we tested was likely Ti6Al4V.
Our stainless steel was mainly composed of iron. All elements which were reported by ASM to
account for more than 1% of the composition of stainless steel 304 are shown in the XRF results
(Table 3). We conclude from these results that our sample is most likely stainless steel 304.

4.2

Structural Imaging

When pulled to fracture, each material had different amounts of necking after reaching the plastic
deformation region on its stress-strain curve. The materials in order of most to least necking were:
UHMWPE (Figure 13), titanium (Figure 12), steel (Figure 11), and bone (Figure 14). From the
amount of necking and the fracture surfaces, the ductility of the materials can be characterized.
Titanium and steel are relatively ductile when compared to bone, which experienced very little
plastic deformation. UHMWPE was extremely ductile in comparison.

4.3

Relevant Mechanical Properties

Three relevant mechanical properties of each material were measured and compared to literature
values. The hardness is an important property of artificial bones because replacement materials
should be capable of resisting enough mechanical stress to sustain body weight and movement
without breaking. The tensile strength characterizes the maximum force a material can undergo
before breaking. This is important because lower extremeties need to withstand high amounts of
force. The elastic modulus characterizes the ratio of tensile stress to strain or the tendency of a
material to deform plastically. This characteristic is also very important because it contributes to
the liklihood of bone resorption.
4.3.1

Hardness

Indentation hardness is not a single fundamental property but a combination of properties, and varies
with the type of test. The modulus of elasticity and the depth of indentation influence conversions.
Therefore separate conversion tables are necessary for different materials.
The ASM Handbook suggests a hardness value of C36 for Ti6Al4V [3]. C38 is converted to
173 Kpsi, and C39 is converted to 177 Kpsi [12]. Assuming this small range of the scale C is
proportional, then C38.03 equates to 173.12 Kpsi or 121.715 kgf /mm2 .
The ASM Handbook reports a maximum hardness of B92 for 304 stainless steel [4]. B93 is
converted to 98 Kpsi, and B94 is converted to 99 Kpsi [12]. Assuming this small range of the scale
B is proportional, then B93.7 can be converted to 98.7 Kpsi or 69.393 kgf /mm2 .
The literature hardness of UHMWPE is R40 as reported by ASM [6]. This value was so soft that
it cannot fairly be compared to the hardness values of the other materials.
The mean hardness of the bone was 38.55 kgf /mm2 , titanium was 121.715kgf /mm2 , stainless
steel was 69.393kgf /mm2 , and the UHMWPE was R40 (Table 6). The hardness of each material
was within a range of magnitude of the expected values. The discrepancy for stainless steel is likely
due to an inexact match to the sample tested by ASM. The composition of our sample is shown to be

13

slightly different than that of pure 304 stainless steel (Table 3). The difference between experimental
and literature values of UHMWPE is because the literature value is actually that of high density
polyethylene, a similar but not identical material. In addition, any of these discrepancies could be
due to the fact that the Rockwell machine was not always powered on for 15 minutes before use, as
instructed in the manual.
4.3.2

Modulus and Tensile Strength

The elastic modulus and tensile strength were calculated using Equation (1) on page 5. Titanium,
steel, UHMWPE, and bone had elastic moduli of 17.710 0.564, 16.8355 3.07, .36297 0.24,
and 9.28 22 respectively (Table 4). The titanium and steel sample raw data were flawed due
to not using an extensometer to measure extension of the sample so literature values will be used
for comparison to the other materials. Both UHMWPE and bone samples had elastic moduli that
matched literature values within error. The high amount of variance within the bone elastic modulus
values arises from the high variability due to a high likelihood of imperfections within bone samples.
Results show that bone has a higher modulus than UHMWPE and that titanium and steel both
have moduli even larger than that of bone.
The tensile strength for titanium, steel, UHMWPE, and bone were measured to be 970 34, 689
34, 22.8 4.8, 397 50 respectively (Table 5). These values are all on the same order of magnitude
and are fairly similar to literature values. UHMWPE has a lower tensile strength than bone while
both titanium and steel have higher tensile strengths.
The strength of these components, which can be characterized using hardness and tensile strength,
is especially important because load-bearing implants at the lower extremities must be able to support up to four times the body weight[11]. An ideal material would have the exact same values as
bone. Lower values indicate that the material would not be able to handle the forces that bone
bears.
Looking at Tables 4 and 5, the elastic modulus and tensile strength of each material can be
compared. The ideal implant material needs to be as strong (as measured by hardness and tensile
strength) as bone and also have as close of an elastic modulus to bone as possible. Elastic moduli
values much higher than those of bone are undesirable because they lead to bone resorption. As a
living structure, the femur will react to the pressure and loads exerted on it by the replacement, as
explained by Wolffs Law [21]. If the bone doesnt experience enough pressure, it will enter bone
resorption and become weaker. The ratio of the elastic modulus of the implant component to the
bone determines the extent to which stress shielding occurs and thus how much of the weight is
borne by the bone [11]. The titanium alloy (hardness: 121.715 kgf /mm2 , tensile strength: 970 34
MPa) and steel (hardness: 69.393 kgf /mm2 , tensile strength: 689 34 MPa) are stronger than the
bone (hardness: 38.55 kgf /mm2 , tensile strength: 397 50 MPa) while UHMWPE is much weaker
(hardness: R40, tensile strength: 22.8 4.8 MPa). The ratios of the elastic moduli of titanium
alloy and stainless steel to the modulus of bone are extremely high which will result in significant
stress shielding.
The combination of high strength and a relatively low modulus make titanium alloys one of the
best suited materials for hip replacements. Titanium alloys are in fact one of the most commonly
used implant materials for orthopaedic applications.
Stainless steel is weaker than titanium alloy and has a much larger elastic modulus than titanium,
making it a less suitable choice for implant materials in orthopaedic applications.
Clearly UHMWPE is not fit for use in core components because the tensile strength is much too
low. However, UHMWPE is an excellent material to be used in the socket and replace cartilage.
UHMWPE is softer but durable, suited for load bearing and movement. For this reason it is used

14

for the joints of many parts, not only for artificial hips, but also artificial knees and shoulders [15].
In hip replacements, UHMWPE is used for the polymeric socket (Figure 3).

4.4

Biocompatibility

Whenever any foreign material is added into the body environment, rejection reactions occur. The
scale of the rejection may range from mild irritation or inflammation to death. Any replacement
material must produce a minimum degree of rejection; this is called biocompatibility. Products
resulting from reactions with body fluids must be tolerated by the surrounding body tissues so that
normal tissue function is unimpaired. [18]
As explained in the introduction, biocompatibility is a characteristic defined by many contributing factors. In this section, we will discuss some of the factors that have not yet been mentioned.
This information does not give a comprehensive image of the biocompatibility of the materials, only
an insight into some aspects of it.
4.4.1

Corrosion Resistance

One component of biocompatibility is corrosion resistance. This is critical because these materials
need to be able to go through many use cycles without corroding. The high corrosion resistance
of chromium is a reason for the popularity of cobalt chromium as a metal in implant devices. The
importance of corrosion resistance led to the discontinued use of polytetrafluoroethylene (Teflon)
and its replacement with UHMWPE, which had less ideal frictional properties but much better
corrosion resistance [11]. Stainless steel has a limited ability to withstand corrosion in the human
body in the long term, so it is used more often as short-term implants such as fracture plates and
screws [9]. Stainless steel 304 in particular, along with stainless steel 316 and some other stainless
steels used for implants, has high amounts of chromium which causes it to have a higher resistance
to corrosion than other stainless steels [11].
4.4.2

Foreign Particles Response

Another factor of biocompatibility is the effect of the accumulation of foreign particles in the surrounding tissues. Over time, all components of the bone replacements will degrade due to wear
and release trace particles into the body. These particles can cause multiple issues for the patients
such as inflammation [18] or an adverse immune response (osteolysis), which may result in having
to replace the implant [20].
As a result many materials chosen to be used in components subjected to wear possess extremely
high abrasion resistance and low coefficients of friction such as UHMWPE, which is often used in
the acetabular component (socket).

Conclusion

Material selection is central to the design of medical implants. For this reason, considerations must
include both ability of the material to serve its function as well as its ability to be integrated into the
body without causing harm. Thus, by analyzing the relevant properties of commonly used implant
materials, the utility of each material becomes apparent.
When comparing the titanium alloys, stainless steel and UHMWPE to bone, the mechanical
properties reveal the potential of each material. Titanium alloy (Ti6Al4V) and stainless steel 304,
in addition to being biocompatible, are even stronger and harder, making them viable choices for core

15

components of an implant. The trade-off for titanium alloy and steel is that the Youngs Modulus
of each material is also higher than that of bone, leading to stress shielding. On the other hand,
UHMWPEs tensile strength and hardness are too low to be used to directly replace bone but is a
excellent candidate to replace cartilage due to its excellent abrasion resistance, corrosion resistance,
and low coefficient of friction.
However, the data clearly demonstrates that these materials (Ti6Al4V and stainless steel 304),
despite being some of the most commonly used, are still not ideal for implants. Specifically, the
Youngs Modulus of commonly used metallic alloys is still too high to prevent bone resorption due
to stress shielding. The other problem is the inability of these materials to repair themselves. As a
result, joint replacements are often replaced due to failure, most commonly due to stress shielding
[19]. This lack of materials with ideal properties points to a serious need to develop new materials
with more optimized properties to be used in bone replacement.

References
[1] II. Osteology. 6c. 3. The Femur. Gray, Henry. 1918. Anatomy of the Human Body. Retrieved
December 16, 2015, from http://www.bartleby.com/107/59.html
[2] Long bone. Retrieved December 16, 2015, from http://www.daviddarling.info/encyclopedia/L/long bone.html
[3] J. Hoffman, T. Xu, and S. Donthu, Medical Polymers, Materials for Medical Devices. Vol 23,
ASM Handbook, ASM International, 2012, p 281289
[4] S.D. Washko and G. Aggen, Wrought Stainless Steels, Properties and Selection: Irons, Steels,
and High-Performance Alloys, Vol 1, ASM Handbook, ASM International, 1990, p 841907
[5] R. Pilliar and S.D. Ramsay, Cobalt-Base Alloys, Materials for Medical Devices. Vol 23, ASM
Handbook, ASM International, 2012, p 211222
[6] Engineering Tables: Polymeric Materials, Engineered Materials Handbook Desk Edition, ASM
International, 1995, p. 9299
[7] S. Lampman, Titanium and Its Alloys for Biomedical Implants, Materials for Medical Devices.
Vol 23, ASM Handbook, ASM International, 2012, p 223236
[8] ASTM Standard E8M, 2015, Standard Test Methods for Tension Testing of Metallic Materials, ASTM International, West Conshohocken, PA, 2003, DOI: 10.1520/E0008 E0008M-15,
www.astm.org.
[9] Knee
Replacement
Implant
Materials.
BoneSmart.
http://bonesmart.org/knee/knee-replacement-implant-materials.

Web.

07

Dec.

2015.

[10] Kucklick, Theodore R. Introduction to Biocompatibility Testing. The Medical Device R&D
Handbook. Boca Raton: Taylor & Francis, 2006. 267-94. Print.
[11] Davis, J. R. Handbook of Materials for Medical Devices. Materials Park, OH: ASM International, 2003. Print.
[12] ASTM Standard E140, 2013, Standard Hardness Conversion Tables for Metals Relationship
Among Brinell Hardness, Vickers Hardness, Rockwell Hardness, Superficial Hardness, Knoop
Hardness, Scleroscope Hardness, and Leeb Hardness, ASTM International, West Conshohocken,
PA, 2003, DOI: 10.1520/E0140-12B, www.astm.org.
16

[13] Niinomi, Mitsuo. Mechanical Properties of Biomedical Titanium Alloys. Materials Science
and Engineering 243.1-2 (1998): 231-36. ScienceDirect. Web.
[14] Turner, C.h., and D.b. Burr. Basic Biomechanical Measurements of Bone: A Tutorial. Bone
14.4 (1993): 595-608. Web.
[15] Kurtz, Steven M., Harvey L. Stein, and Gunther Redeker. Meeting the Joint Replacement
Challenge with UHMWPE. MDDI RSS. Medical Device and Diagnostic Industry, 1 Mar. 2005.
Web. 07 Dec. 2015.
[16] Vesper, Hubert W. REVIEW: Analytical and Preanalytical Issues in Measurement of Biochemical Bone Markers. Laboratory Medicine 36.7 (2005): 424-29. Web.
[17] Callister, William D., and David G. Rethwisch. Materials Science and Engineering: An Introduction. Vol. 8. New York: John Wiley Sons, 2010. Print.
[18] Callister, William D., and David G. Rethwisch. Case Study 4 Artificial Total Hip Replacement. Fundamentals of Materials Science and Engineering: An Integrated Approach. 4th ed.
Hoboken, NJ: John Wiley Sons, 2011. CS4.1-S4.7. Print.
[19] Rath, L. Metal Hips Fail Faster, Raise Other Health Concerns. Retrieved December 16, 2015,
from
http://www.arthritis.org/living-with-arthritis/treatments/joint-surgery/types/hip/hipreplacement-failure-rate.php
[20] Hip
Replacement
Implant
Materials.
BoneSmart.
http://bonesmart.org/hip/hip-replacement-implant-materials.

Web.

16

Dec.

2015.

[21] Teichtahl, Andrew J., Anita E. Wluka, Pushpika Wijethilake, Yuanyuan Wang, Ali GhasenZadeh, and Flavia M. Cicuttini.Wolffs Law in Action: A Mechanism for Early Knee Osteoarthritis. Arthritis Research & Therapy 17 (2015): Web
[22] Histology
of
Bone.
Retrieved
December
http://emedicine.medscape.com/article/1254517-overview

17

16,

2015,

from

You might also like