Orthopaedic Biomechanics J. Michael Kabo, PHD
Orthopaedic Biomechanics J. Michael Kabo, PHD
Force or load (pounds, Newtons) is an applied quantity that causes a material to change its
shape. It is a vector quantity which means that both the magnitude and direction have
significance in determining its effects. For example, a weight, W, sitting on a table top exerts a
force of magnitude |W| on the table top and it acts in the direction of gravity, downward. The
table will experience a change in shape as a result of this load being placed on it which may or
may not be visible to the naked eye owing to the properties of the table and the magnitude of
the load applied. The material to which the force is applied undergoes a deformation in the
direction of the applied force causing it to change its shape. Applied forces which tend to
squeeze or compress (inches, millimeters, centimeters) a material tend to shorten it in
dimension while those in the opposite direction are tensile and cause the material to stretch or
elongate. For the most part we will be dealing with materials that are linear and elastic. Linear
materials exhibit a straight line relationship between the applied load and the resulting
deformation. An elastic material is one that completely returns to its original shape once the
applied load has been removed. The relationship between the applied force and the resulting
deformation is a characteristic of both the material and its shape. This relationship is known as a
structural diagram (Figure 1). It is visualized as a plot of force (load, torque, moment) versus
deformation (change in length, angle of twist, change in curvature). The load versus deformation
diagram characterizes the structural behavior of a particular specimen. If this relationship is
linear over a particular range then we can describe this relationship by a single parameter called
structural stiffness. By way of definition, stiffness (k) = force (F) divided by deformation (ΔL)
(pounds/inch, Newtons/meter, Newton-meters/degree), which in symbolic notation is:
By analogy, one can imagine a simple coil spring made of a particular material that is well
defined by a spring stiffness, k. Obviously, for a coiled spring, the size, shape, and number of
coils as well as the material from which it is made has a direct influence on the stiffness. Springs
which look and are shaped differently will have different stiffness values. This is also true for
other structures as well where the shape and material(s) both have a direct influence on the
behavior of the structure under load.
{strain; stress; shear; Poisson's ratio}
In order to determine what effect the applied load has on the inside of the structure it is
necessary to identify those properties that are independently associated with either the material
itself or the shape of the object. To account for the different shapes that a material may come in
re define additional terms to obtain "normalized" values for the force and deformation. The form
of the structural diagram shown in Figure 1 by itself does not permit that determination. This
diagram merely tells us that when a load is applied to a structure the object changes shape. The
simplest concept is that of a bar that is stretched or compressed causing it to lengthen or
shorten by an amount ΔL. If the original length of the bar was L then we can define normal
strain, ε, as the change in length, ΔL, divided by the original length, L. Our symbolic notation
yields:
Strains for biological materials intended for use in structural roles (metals, ceramics, plastics,
bone) are dimensionless (μ−strain, μ −inch/inch, microns/meter) since it is a length divided by
a length of the same units). These values are quite small since the overall length is much larger
than any deformations induced.
Shear strain , γ, occurs when the object is distorted rather than compressed or stretched and
hence is defined as a change in angle divided by the original angle. Figure 2 displays the
conceptual differences between the two different types of strain.
Usual values for Poisson's ratio are between 0.25 and 0.5 and they are always positive as
defined. In the special case where ν = 0.5 the material is incompressible which means that
although the shape changes there is no change in volume of the specimen.
{Elastic modulus; yield point; elastic limit; proportional limit; ultimate strength; work
hardening; cold rolling; plastic deformation; ductile; brittle; failure strength; strain
energy; hardness}
For a given material fabricated into a well defined shape we can construct a special diagram
(Figure 4) using the same data that was used to produce the structural diagram (Figure 1). If we
divide the force by the cross sectional area to calculate stress, σ, and divide the change in
length by the original length to calculate strain, ε, we can then construct the stress versus strain
diagram. This diagram is used to determine the mechanical properties of the material which are
now independent of the shape. In contrast to the structural diagram this representation absorbs
the geometrical properties of the specimen and therefore we have a relationship that depends
only on the material properties. The material properties that result still, however, depend on the
manner in which the specimen is loaded.
Figure 4
The above figure displays all of the relevant information that can be derived from the stress
versus strain diagram. This is the idealized material response for a material that exhibits linear
elastic behavior. It is useful for understanding the behavior of bone, PMMA, metals, ceramics,
plastics, and under limited circumstances soft tissues as well). If there is a well defined initial
straight line section for the relationship obtained during a compression or tension test then the
slope of this line is called the elastic modulus, E. This parameter is also known as Young's
modulus and the material is said to obey Hooke's Law. The elastic modulus can easily be
determined from this relationship as:
A modulus determined in this fashion is similar to the stiffness defined in Figure 1, however, it
used the normalized values of stress and strain. It has dimensions of Force per unit Area
(pounds/inch2, Newtons/meter2 = 1 Pascal, or other similar familiar pressure units). The way in
which a specimen is loaded has a direct bearing on the meaning of the modulus obtained. For
example, if we tested the specimen by applying a twisting load we would obtain a shear
modulus. The point at which the relationship begins to deviate from a straight line is known as
the proportional limit (shown as P-limit in Figure 4) As one continues to load the specimen you
will next reach the elastic limit and the yield point. The elastic limit is important since this the
point where, if you continue to load the specimen, permanent deformation will result. As the
exact location of this point is difficult to predict a standard has been established to clearly define
when the specimen first undergoes yielding. This is known as the yield point and is determined
as that load which results in permanent deformation after removal which equals a 0.2% change
in the original length (0.2% offset). Up to this point if the specimen is unloaded it will always
return to its original shape. For all intents and purposes of biomechanics the proportional limit,
elastic limit, and yield point are indistinguishable. Once the yield point has been exceeded
there will always be some residual deformation that will always be locked into the material
specimen. Unloading from any part of the curve (even after the yield point has been reached)
will always result in a straight line relationship which will follow a path that is parallel to the
section that determines the elastic modulus. If loading continues every point represents a new
yield point for the material. If the material is then unloaded and reloaded it will exhibit linear
elastic behavior up to the last yield point from which additional loading will follow the path of the
remainder of the curve with additional permanent deformation induced. Each subsequent yield
point is higher than the previous one. This phenomenon is known as work hardening. This is
precisely what happens during the fabrication processes known as cold rolling and forging.
The net result is that the new strength for the structure (as measured by the new yield point) is
higher than it was for the basic material originally. The ultimate strength is the highest stress
observed on the stress versus strain diagram while the failure strength is the stress value at
which the material eventually fails. The reason for the drop off in magnitude is due to a
phenomena known as necking. The area beneath the stress versus strain diagram represents
the strain energy and is a measure of the ductility of the material. The larger the area, the
more ductile the material. Brittle materials do not permit significant plastic deformation before
failure and the ultimate strength is identical to the failure strength.
A higher modulus of elasticity (more vertical slope) does not mean that the material is more
elastic, but rather, a higher stress is required to produce the same amount of strain. The
material is in effect stiffer.
Failure of the material is basically whatever criteria are appropriate for a given situation. Abrupt
fracture is clearly a failure of the material. On the other hand, a nail plate device that has bent
so as to prevent apposition of the bones is also a failure. Although the effect on the material is
not as dramatic (i.e., abrupt fracture caused by exceeding the failure strength of the material)
the result is nonetheless a failure since the yield point of the material has been exceeded
resulting in permanent deformation of the device and preventing bony union. Strain energy is
the area formed beneath the stress vs. strain diagram. This is analogous to the energy that is
the area beneath the force vs. deflection curve.
Hardness is the ability of a material to resist scratching and indentation on the surface. While its
value is not determined from the stress vs. strain diagram it does influence the durability and
machinability of the material. Several scales have been developed for materials of different
hardness ranges. The more common ones include Rockwell, Brinell, and Shore.
{Forging; casting (molding); HIP; material designations}
Forging involves heating the material until red hot and then reshaping it using a series of
hammer blows. This is the same process by which horse shoes are made. Casting involves
pouring molten metal into a pre-shaped receptacle. It is prone to trapping impurities and
shrinkage voids and internal cracks may develop as the material hardens during cooling.
Casting generally results in a material that is weaker than if fabricated by other means. Its
usefulness arises in the ability to form intricate shapes and in the mass production of items. In a
similar fashion, UHMWPE may be molded into desired shapes as the result of subjecting
powder to elevated temperatures and pressures. Hot lsostatic Pressing (HIP) is a powder
metallurgy technique which fuses the material at high temperatures and pressures into the
desired shape. This process is also used to post process finished implants to fill microscopic
defects which may have formed on the surface of the implant..
The American Society for Testing Materials has established a convention for the naming of
metal alloys. The mnemonic involves indicating the significant elements that are present by
weight. For example Ti-Al6-V4 contains 6 parts of aluminum and 4 parts of Vanadium by weight
with the bulk of the material Titanium. The particular designation of ELI grade stands for extra
low interstitial. Co-Cr-Mo alloy contains mostly cobalt with 28% Chromium and 6% Molybdenum
by weight. For other grades of Co-Cr alloy the Molybdenum content is less than 1%. Stainless
steels are ferrous alloys (containing mostly iron) and have a different designation. The stainless
steel grade most commonly used for implants is 316L where the L designates low interstitial
carbon content. Because of the low carbon content this grade is less susceptible to corrosion in
vivo. Other grades of SS are used for surgical tools and instruments. Stainless steel contains
substantial Chromium (17%) and Nickel (14%). Its use for patients with allergies to nickel is
cautioned.
{Heat treatments; annealing; scintering; hot pressing; passivation; ion implantation}
Heat treatments are processes applied to metals and plastics which involve heating the
material for a prescribed period of time. When materials are fabricated in such a way as to
create stresses within the bulk of the material (extruding, bending, etc.) it is desirable to relieve
these stresses which otherwise may jeopardize the finished part if it is subjected to loads.
Annealing is used to relieve internal stresses developed in the material during its fabrication. it
involves beating the material at relatively low temperatures for extended periods of time to allow
the grains or molecules of the material to realign themselves to a stress free state. Scintering is
the process used in fining porous layers onto a bulk substrate. It generally involves very high
temperatures (> 10000C) and high pressures. In the particular case of titanium (and its alloys)
these temperatures are near the melting point and cause the microstructure to become so
altered that what results is a much weaker material. Hot Pressing is a heat treatment that has
been used to improve the surface finish of tibial plateau components. It involves "ironing" the
surface which removes machine tool marks and any minor imperfections. The effects of this
process are relatively shallow and cause embrittlement of the surface layer. Catastrophic wear
and delaminations of this surface layer have been reported.
Passivation is a chemical treatment applied to the metallic implant surfaces to develop a very
thin oxide layer to improve the biocompatibility of the surface. Ion implantation is a technique
available for hardening the surfaces of metals and plastics to improve the durability and wear
characteristics. It involves fining gaseous atoms (usually nitrogen) at very high speeds onto the
surfaces of the material. This forces the atoms between the existing atoms of the material
resulting in a toughened skin. Its effects extend less than 0.lmm into the surface.
{Bending; torsion; force couple; rigidity; neutral axis; moment of inertia}
Up to now we have only considered compression or tension loading, i.e., those loadings that
produce a deformation along the direction in which the force is applied. Deformations may also
occur by loadings that bend a specimen, bending, or those that lead to torsion (or twisting) of
the specimen. Bending may be caused by single force acting at a distance, L, along a beam
fixed at the opposite end, by a pure moment, M, or by a combination of loading and supports
that act transverse (perpendicular) to the long axis of the beam. Torsion results from a force
couple, shear stresses, or torque that acts in a manner so as to twist the specimen about its
long axis. The resistance to deformation is defined as flexural stiffness for bending and
torsional rigidity for cases of torsion. In both cases the resistance to deformation is very
strongly influenced by the distribution of material about the axis of bending or torsion. This axis
(which is actually a line in torsion and a plane in bending) is called the neutral axis (Figure 5).
The term neutral axis arises because the stress has been
"neutralized" along this portion of the specimen. Here there is no stress present. Maximum
stress occurs along the outermost fibers of the specimen for both bending and torsion.
How the material is distributed away from the neutral axis is represented by the moment of
inertia. For bending it is called the area moment of inertia while polar moment of inertia is
used for cases of torsion. In simple terms the more material can be distributed away from the
neutral axis, the higher the moment of inertia and the greater the structural load bearing
capability (lower stress) of the specimen (Figure 6).
In the case of bending the orientation of the specimen has a direct impact on the stiffness of the
member. More material in the plane of bending effectively makes a stiffer member. Figure 6
shows the cross section of a rectangular beam of height, h, and width, b. In this figure the
bending occurs about the "h" dimension with the section in the right hand side of the figure
producing a structure which has much greater structural stiffness simply by virtue of its
orientation with respect to the applied loading. The area moment of inertia for a rectangular
beam is:
where h is the dimension in the plane of curvature. It is readily seen that the moment of inertia is
directly proportional to the cube of the height. If h = 2 and b = 1 then by rotating the beam so
that the thickest dimension is presented to the bending action then we have an 4X increase in
the moment of inertia. This is the same principle used in construction with the I-beam
configuration.
Similarly for torsion we see that by distributing the material further away from the neutral axis
the polar moment of inertia is greatly increased. For a thin tube of thickness, t, and radius, R,
the polar moment of inertia is:
If the radius, R, is doubled then the increase in the polar moment of inertia increases
dramatically to (2R)3/(R)3 = 8X.
It must be remembered that a solid rectangle has a higher area moment of inertia than an
I-beam of the same outer dimensions. A solid circular bar has a higher polar moment of inertia
than a thin tube with the same outer diameter, however, the advantage lies in the ability to
maximize the moment of inertia per unit weight. Long bones that are essentially cortical tubular
constructs epitomize this concept. The protected intramedullary cavity is then available for other
functions.
The relationships between stress, strength, and stiffness for the four modes of loading are
presented and summarized in Table 1. This table lists the factors of proportionality for each of
the given modes. Stiffness is defined in the usual sense as force per unit deflection. Strength is
the load that causes failure and is based on a critical stress factor, ac, which is obtained from
the stress vs. strain diagram as yield stress, ultimate strength, or failure strength. Bear in mind
that these relationships between stress strength, and stiffness are not independent as they are
simply manipulations of the single stress relationship for the specified mode of loading. The
constant terms are defined in the table legend.
Table 1 - Summary Table of Structural Properties
Loading Stress(N/m) Stiffness(N/A) Strength(N)
c = distance from neutral axis - bending c0 = distance from neutral axis to the outer most
fiber
ρ = distance from neutral axis - torsion ρ 0 = distance from neutral axis to the outer most
fiber
{Viscoelastic materials; strain rate; creep; stress relaxation; hyperelastic materials;
hysteresis}
Viscoelastic materials form a special class of materials that are sensitive to the speed at which
the load is applied. In general the faster the strain rate (rate of loading) the higher the stress at
a given level of strain (Figure 7). In general the following shifts occur in the stress vs. strain
diagram as the strain rate is increased: the material becomes stiffer (modulus increases), the
ultimate strength increases, and the total elongation decreases as depicted in the figure. "Silly
Putty" is a good example for observing these phenomenon. As you pull it slowly it can be easily
stretched to arms length. With a very quick and strong pull however, it will break after a relatively
short extension. All soft tissues including ligaments, tendons, and cartilage exhibit viscoelastic
behavior to varying degrees. Even bone at low levels of load has been found to behave as a
viscoelastic material. This explains why under low loading rates (in the knee, for example) there
is often bony tubercle avulsion at the cruciate ligament attachment site while the ligament itself
will usually fail in mid substance while under very dynamic conditions (impact loadings). In effect
the high rate of loading has caused the bone to be effectively stronger than the ligament that is
attached to it. The material is linear viscoelastic if the stress has a linear dependence on the
strain rate.
Creep is a property of some materials by which they continue to deform under a constant load
(Figure 9). When a load is suddenly applied there occurs an instantaneous elastic component of
strain followed by a gradual increase in additional strain. The reciprocal of creep is termed
stress relaxation (Figure 8). The distinction really relates to how the load is applied and the
response measured. Either technique may be used to determine the viscoelastic behavior of the
material. In the latter case a material is stretched (or compressed) and held in place. This is
followed by a gradual reduction in internal stress.
Plastics (including UHMWPE and PMMA) as well as cartilage demonstrate creep phenomenon.
Creep and stress relaxation are distinguishable from permanent (plastic) deformation since
once the loading state has been removed the deformation is completely recovered. The
recovery process may take a considerable amount of time, even hours, to restore the specimen
completely to its original shape.
Until now we have been discussing materials that deform by only a small amount (usually
imperceptible without special measurement tools). In contrast hyperelastic materials can be
deformed significantly without failure (20% and higher) (Figure 10). In the elastic sense the
material deforms under load and then returns to its undeformed shape when the load is
removed. During this loading and unloading process the return path does not follow the loading
and a hysteresis loop is formed. During loading energy is stored in the specimen and as the
load is removed the specimen releases this stored energy. The difference between energy
stored and energy released is the energy lost that accounts for the hysteresis. During the
loading and unloading process this energy is converted into heat and dissipated throughout the
substance of the material. This effect is most pronounced for soft tissues. Determination of a
modulus value for such materials is much more difficult since there is no well defined linear
region of the curve. Typically the modulus will be specified for a given level of strain. Sometimes
it may be appropriate to fit the data empirically to a suitable mathematical relationship. Since
there is no accepted norm for selecting a strain value wide variability exists for modulus values
for UHMWPE and PMMA that behave in this fashion.
{Fatigue strength; S-N diagram; endurance limit}
When a material is subject to multiple loading and unloading cycles we become concerned with
how well the material will endure with this type of loading. Just as bones are subject to stress
fractures, so too metals may fail at a stress level that is far below the yield point. The
Endurance Limit (or fatigue limit) is a measure of the safe stress that can be applied to the
material in a cyclic fashion for ever and ever. Since the stress within the material is maintained
below this level then the material is not expected to fail. A "ball park" estimate of its value can be
obtained as 50% of the ultimate strength of the material. Its value is determined from the S-N
diagram (Figure 11) that is a plot of the stress (S) versus number of cycles (N) that have been
imposed on the specimen. Multiple specimens are required to determine this relationship
completely since each specimen is tested until failure. The fatigue strength (or endurance
strength) is that value of stress at which the material failed. When specified the corresponding
number of cycles also must be given.
{Stress concentration; stress shielding; notching; corrosion; EMF, carcinogenesis)
Forces and moments act to provide motion, stability, and balance. They cause stresses to be
developed in bones, implants, cartilage and other structures in vivo. Forces and moments can
be applied externally as well as resulting from the dynamic action of the muscles. Forces and
moments are vector quantities meaning that they are described by a magnitude and a direction.
A free body diagram is a means of estimating forces and moments which occur within the body
as a result of external loads applied. A region of interest is isolated and all important contact
regions and muscle attachments are replaced with arrows. The direction and length of the arrow
symbolize the direction and magnitude of the respective forces. These can be solved by a
detailed mathematical procedure or in simple cases graphically. If the free body is not moving all
of the forces that are acting are in complete balance and the system is in static equilibrium.
This is important for using the graphical method to solve for the forces involved since all of the
drawn arrows when connected "tip-to-tail" must form a closed geometric figure. (Remember that
the direction and length must be preserved as they are translated). Figure 13 illustrates this
principle for the lower limb pictured during single leg stance. (5/6)W is the weight exerted on the
hip joint, J, is the hip joint reaction force, and M, is the abductor muscle force. We can readily
see that the largest force present (longest arrow) is that of the hip joint force, J. This diagram
also illustrated the concept of a moment arm. A moment arm is the perpendicular distance
from a point of rotation to the line of action of a force. In the case of Figure 13 the body weight
moment arm is designated as distance c and the abductor muscle moment arm is designated
as distance b. The moment produced by the abductor muscle must be sufficient to balance the
moment produced by the body weight to prevent rotation about the hip joint. When this condition
is satisfied the system is in static equilibrium. Reconstruction of the moment arm is perhaps
the single most important biomechanical action that needs to be address during reconstructive
procedures. Failulre to do so will result in an internal imbalance of forces, causing muscles to
work harder, increase energy expenditure, alter posture, and over time lead to additional
mechanical problems both locally and in regions remote from the site of the reconstruction. It
has also been associated with the acceleration of arthritic conditions.
{Fracture; mass; velocity; Energy; Conservation of Energy}
Fracture of bone involves excessive loading to cause failure of the structure. The configuration
of the fracture as well as the underlying circumstances give rise to the mechanism of loading
which produced the fracture. Bone is weaker in tension than in compression and in all but a few
instances the fracture will be initiated where the local state of stress is tensile in nature. Simple
loadings result in simple fracture patterns. Avulsion fractures are tensile in nature and are
situated in close proximity to a tendon or ligament insertion point. The applied force causing the
fracture was likely perpendicular to the fracture line. Bending forces also result in a fracture
which is transverse to the long axis of the bone and may be distinguished from tensile fractures
by the presence of butterfly fragments. Fractures resulting from twisting injuries have the
characteristic spiral appearance. The angle of this spiral makes an approximate 45o angle with
the long axis of the bone. Although it is beyond the scope of this presentation, it can be shown
that the torsional loading which produced this fracture can be resolved through a pure rotation to
reveal that tensile forces are acting perpendicular to the fracture surface. Combinations of
loadings will result in more complex fracture patterns.
Complex fracture patterns and extensive comminution are associated with higher levels of
energy involved in the trauma. Kinetic Energy can be expressed in the form:
where m is mass and v indicates velocity. Significant damage results from high energy trauma.
High energy trauma results from vehicular collisions (large vehicle mass), projectile impact
(bullets or baseballs with high velocity but relatively low mass), or falls (high mass, high or low
velocity). One of the fundamental laws of physics states that energy must be conserved
(conservation of energy). Thus the energy involved during the impact is transferred to the
body. During the collision phase energy is stored in the soft and hard tissues until the strength of
the respective material is exceeded. At the point of failure or fracture this stored energy is
spontaneously released. Excess energy associated with the impact then imparts additional
velocity to all of the members (including bone fragments) involved. High energy impacts are
especially damaging to the surrounding soft tissues since the fragments that are formed during
fracture of the bone may acquire significant velocities propelling them into the surrounding soft
tissues resulting in significant additional local traumatic effects. Potential energy is another form
of energy and is more passive in nature. It is associated with an elevation with respect to a
particular reference level or represents the work done during the process of deformation of the
tissues.
Bibliography
1. Addison, W: Elements of Material Science and Engineering. 3rd Edition, 1985.
2. Annual Book of ASTM Standards: Medical Devices. Section 13, Volume 13.01, 1985 and
1988 series.
3. Basic orthopaedic biomechanics. Editors: Van C. Mow, Wilson C. Hayes. 2nd ed.
Philadelphia, Lippincott-Raven, 1997.
4. Brown, KJ, Atkinson, JR, Dowson, D, and Wright, V: The wear of ultra high molecular
weight polyethylene and a preliminary study of its relation to the in vivo behavior of
replacement thigh joints. Wear 40:255-264, 1976.
5. Brown, SA and Merrit, K: Fretting corrosion in saline and serum. JBMR 15:479-488, 1981.
6. Burke, DW, Gates, EI, and Harris, WH: Centrifugation as a method of improving tensile
and fatigue properties of acrylic bone cement. JBJS 66A:1265-, 1984.
7. Burstein AH and Wright T: Orthopaedic Biomechanics.
8. Burstein, AH, Wright, TM: Fundamentals of orthopaedic biomechanics. Baltimore,
Williams & Wilkins, 1994.
9. Chao, E and Lin, O: International Symposium on Biomaterials: Materials Science
Monographs. Vol 33, Perspectives on Biomaterials, Elsevier Science, 1986.
10. Charnley, J: Acrylic cement in orthopaedic surgery. Longman Group, Edinburgh, 1970.
11. Christel, P and Meunier, A: A histomorphometric comparison of the muscular tissue
reactions to high-density polyethylene in rats and rabbits. JBMR 23:1169-1182, 1989.
12. Colangelo, VJ, Greene, ND, Kettlekamp, DB, Alexander, H and Campbell, CJ: Corrosion
rate measurements in vivo. JBMR 1:405-414, 1967.
13. Cornell, CN and Ranawat, CS: The impact of modern cement techniques on acetabular
fixation in cemented total hip replacement. J. Arthroplasty 1(3):197-202, 1986.
14. Dagleish, BJ and Rawlings, RD: A comparison of the mechanical behavior of aluminas in
air and simulated body environments. JBMR 15:527-542, 1981.
15. Dorlot, J-M, Christel, P and Meunier, A: Wear analysis of retrieved alumina heads and
sockets of hip prostheses. JBMR 23:299-310, 1989.
16. Ducheyne, P: Functional Behavior of Orthopedic Biomaterials. Volume 1, CRC Press,
Boca Raton, FL, 1984.
17. Ducheyne, P: Functional Behavior of Orthopedic Biomaterials. Volume 2, CRC Press,
Boca Raton, FL, 1984.
18. Emmanual, J, Emmanual, JG and Koeneman, JB: Alteration of retrieved implants in vitro
by processing and infiltrating fluids. JBMR 23:337-347, 1989.
19. Fuchs, MD, Salvati, EA, Wilson, PD, Sculco, TP and PelliccL PM: Results of acetabular
revisions with newer cement techniques. Orthop. Clin. North. Am. 19(3):649-655, 1988.
20. Gibbons, DF: Materials for orthopedic joint prostheses. In: Biocompatibility of Orthopedic
Implants, Vol 1, CRC Press, Boca Raton, pp. 111-139, 1982.
21. Harris, WH and Davies, JP: Modern use of modem cement for total hip replacement.
Orthop. Clin. North Am. 19(3):581-589,1988.
22. Hayashi K, Matsuguchi, N, Uenoyama, K, Kanemaru, T and Sugioka, Y: Evaluation of
metal implants coated with several types of ceramics as biomaterials. JBMR
23:1247-1259, 1989.
23. Isaac, GH, Atkinson, JR, Dowson, D, and Wroblewski, BM: The role of cement in the long
term performance and premature failure of Charnley low friction arthroplasties. Eng. Med.
15(1):19-22, 1986.
24. Jones, LC, and Hungerford, DC: Cement disease. Clin. Orthop. 225:192-206, 1987.
25. Kawauchi, K, Kuroki, Y, Saito, S, Ohgiya, H, Sato, S, Kondo, S, Hirose, I, Obara, S, Aso,
H, Yamano, K, Sasada, S and Norose, S: Total hip endoprostheses with ceramic head
and H. D. P. Socket. Clinical wear rate. In:(Oonishi, H, and Ooi, Y, eds.) Orthopedic
Ceramic Implants, Japanese Socieiy of Ceramic Implants, 4:253-257, 1984.
26. Lee, AJ: The effect of mixing technique and surgical technique on the properties of bone
cement. Aktuel. Probl. Chir. Orthop. 31:145-150, 1987.
27. McKellop, H, Clarke, 1, Markolf, K and Amstutz, H: Friction and wear properties of
polymer, metal, and ceramic prosthetic joint materials evaluated on a multichannel
screening device. JBMR 15:619-653, 1981.
28. Meyer, PR, Lautenschlager, EP and Moore, GM: On the setting properties of acrylic bone
cement. JBJS 55A:149-156, 1973.
29. Orthopaedic basic science : biology and biomechanics of the musculoskeletal system.
Editors: Joseph A. Buckwalter, Thomas A. Einhorn, Sheldon R. Simon. 2nd ed.,
Rosemont, IL, American Academy of Orthopaedic Surgeons, 2000.
30. Park, J.: Biomaterials Science and Engineering, Plenum Press, New York and London,
1984.
31. Paul, HA and Bargar, WL. Histologic changes in the dog femur following total hip
replacement with current cementing techniques. J. Arthroplasty 1(1):5-9, 1986.
32. Plenk, HJ: Biocompatibility of ceramics in joint prostheses. In: Biocompatibiliiy of
Orthopedic Implants, Vol 1, CRC Press, Boca Raton, pp. 269-295, 1982.
33. Poss, R, Brick, GW, Wright, RJ, Roberts, DW, and Sledge, CB: The effects of modern
cementing techniques on the longevity of total hip arthroplasty. Orthop. Clin. North Am.
19(3):591-598, 1988.
34. Practical biomechanics for the orthopedic surgeon. Editor: Eric L. Radin, 2nd ed. New
York , Churchill Livingstone, 1992.
35. Ranawat, CS, Rawlins, BA and Harju VJ: Effect of modem cement technique on
acetabular fixation total hip arthroplasty. A retrospective study in matched pairs. Orthop.
Clin. North Am. 19(3):599-603, 1988.
36. Roberts, DW, Poss, R and Kelley, K: Radiographic comparison of cementing techniques
in total hip arthroplasty. J. Arthroplasty 1(4):241-247, 1986.
37. Ross, R: Metallic Materials Handbook. 3rd edition, E&F publishers, London, New York,
1980.
38. Russotti, GM, Coventry, MB and Stauffer, RN: Cemented total hip arthroplasty with
contemporary techniques. A five-year minimum follow-up study. Clin. Orthop.
235:141-147,1988.
39. Ryu, RKN, Bovill, EG, Skinner, HB and Murray, WR: Soft tissue sarcoma associated with
aluminum oxide ceramic total hip arthroplasty. Clin. Orthop. 216:207-212, 1987.
40. Shelley, P and Wroblewski, BM: Socket design and pressurization in the Chamley
low-friction arthroplasty. J. Bone Joint Surg. 70B(3):358-363, 1988.
41. Soballe, K and Christensen, F: Improved cementation in total hip replacement. Arch.
Orthop. Trauma Surg. 107(1):50-53, 1988.
42. Swann, M: Malignant soft-tissue tumor at the site of a total hip replacement. JBJS
66B:629-, 1984.
43. Tencer, AF. Biomechanics in orthopedic trauma: bone fracture and fixation. London: M.
Dunitz ; Philadelphia : Lippincott, 1994.
44. Weber, BG: Pressurized cement fixation in total hip arthroplasty. Clin. Orthop. 231:29-34,
1988.
45. Weightman, B, Freeman, MA, Revell, PA, Braden, M, Albrektsson, BE and Carlson, LV:
The mechanical properties of cement and loosening of the femoral component of hip
replacements. J. Bone Joint Surg. 69B(4):558-564, 1987.
46. Williams, DF: Corrosion of orthopedic implants. In: Biocompatibility of Orthopedic
Implants, Vol 1, CRC Press, Boca Raton, pp. 197-229, 1982.
47. Zitter, H and Plenk, H: The electrochemical behavior of metallic implant materials as an
indicator of their biocompatibility. JBMR 21:881-896, 1987.