ArticlePDF Available

The Role of Flow-Independent Viscoelasticity in the Biphasic Tensile and Compressive Responses of Articular Cartilage

Authors:

Abstract and Figures

A long-standing challenge in the biomechanics of connective tissues (e.g., articular cartilage, ligament, tendon) has been the reported disparities between their tensile and compressive properties. In general, the intrinsic tensile properties of the solid matrices of these tissues are dictated by the collagen content and microstructural architecture, and the intrinsic compressive properties are dictated by their proteoglycan content and molecular organization as well as water content. These distinct materials give rise to a pronounced and experimentally well-documented nonlinear tension-compression stress-strain responses, as well as biphasic or intrinsic extracellular matrix viscoelastic responses. While many constitutive models of articular cartilage have captured one or more of these experimental responses, no single constitutive law has successfully described the uniaxial tensile and compressive responses of cartilage within the same framework. The objective of this study was to combine two previously proposed extensions of the biphasic theory of Mow et al. [1980, ASME J. Biomech. Eng., 102, pp. 73-84] to incorporate tension-compression nonlinearity as well as intrinsic viscoelasticity of the solid matrix of cartilage. The biphasic-conewise linear elastic model proposed by Soltz and Ateshian [2000, ASME J. Biomech. Eng., 122, pp. 576-586] and based on the bimodular stress-strain constitutive law introduced by Curnier et al. [1995, J. Elasticity, 37, pp. 1-38], as well as the biphasic poroviscoelastic model of Mak [1986, ASME J. Biomech. Eng., 108, pp. 123-130], which employs the quasi-linear viscoelastic model of Fung [1981, Biomechanics: Mechanical Properties of Living Tissues, Springer-Verlag, New York], were combined in a single model to analyze the response of cartilage to standard testing configurations. Results were compared to experimental data from the literature and it was found that a simultaneous prediction of compression and tension experiments of articular cartilage, under stress-relaxation and dynamic loading, can be achieved when properly taking into account both flow-dependent and flow-independent viscoelasticity effects, as well as tension-compression nonlinearity.
Content may be subject to copyright.
Chun-Yuh Huang
VanC.Mow
Gerard A. Ateshian
Departments of Mechanical Engineering and
Biomedical Engineering,
Columbia University,
New York, NY 10027
The Role of Flow-Independent
Viscoelasticity in the Biphasic
Tensile and Compressive
Responses of Articular Cartilage
A long-standing challenge in the biomechanics of connective tissues (e.g., articular car-
tilage, ligament, tendon) has been the reported disparities between their tensile and com-
pressive properties. In general, the intrinsic tensile properties of the solid matrices of
these tissues are dictated by the collagen content and microstructural architecture, and
the intrinsic compressive properties are dictated by their proteoglycan content and mo-
lecular organization as well as water content. These distinct materials give rise to a
pronounced and experimentally well-documented nonlinear tensioncompression stress
strain responses, as well as biphasic or intrinsic extracellular matrix viscoelastic re-
sponses. While many constitutive models of articular cartilage have captured one or more
of these experimental responses, no single constitutive law has successfully described the
uniaxial tensile and compressive responses of cartilage within the same framework. The
objective of this study was to combine two previously proposed extensions of the biphasic
theory of Mow et al. [1980, ASME J. Biomech. Eng., 102, pp. 7384] to incorporate
tensioncompression nonlinearity as well as intrinsic viscoelasticity of the solid matrix of
cartilage. The biphasic-conewise linear elastic model proposed by Soltz and Ateshian
[2000, ASME J. Biomech. Eng., 122, pp. 576586] and based on the bimodular stress-
strain constitutive law introduced by Curnier et al. [1995, J. Elasticity, 37, pp. 138], as
well as the biphasic poroviscoelastic model of Mak [1986, ASME J. Biomech. Eng., 108,
pp. 123130], which employs the quasi-linear viscoelastic model of Fung [1981, Biome-
chanics: Mechanical Properties of Living Tissues, Springer-Verlag, New York], were com-
bined in a single model to analyze the response of cartilage to standard testing configu-
rations. Results were compared to experimental data from the literature and it was found
that a simultaneous prediction of compression and tension experiments of articular car-
tilage, under stress-relaxation and dynamic loading, can be achieved when properly tak-
ing into account both flow-dependent and flow-independent viscoelasticity effects, as well
as tensioncompression nonlinearity. DOI: 10.1115/1.1392316
Introduction
Over the past two decades, several studies have established that
the viscous drag induced by interstitial fluid flowing within the
porous-permeable collagenproteoglycan matrix of cartilage im-
parts viscoelasticity to the mechanical response of this tissue. This
flow-dependent viscoelastic phenomenon has been the basis of
successful porous media models 1–7, which can describe the
response of articular cartilage under various compressive loading
conditions, including confined 1,2,4 and unconfined 7,8 com-
pression of cylindrical cartilage discs, as well as indentation of
cartilage layers with a flat or spherical indenter 9,10. In addition
to this mechanism of flow-dependent viscoelasticity, some inves-
tigators have proposed that there also exists an intrinsic, flow-
independent viscoelasticity in the solid matrix 11,12, leading to
the formulation and application of porous media models with a
viscoelastic solid phase 13–15. Incorporation of intrinsic vis-
coelasticity into porous media models has often produced better
agreement between theory and experiments than in the absence of
modeling such effects 16,17, though not always 18.
However, it has been difficult to assess whether this improved
agreement has indeed resulted from the existence of intrinsic
solidmatrix viscoelasticity, or was caused by the increased math-
ematical flexibility of the governing equations and the number of
material parameters in the model. A long-standing argument in
favor of the former interpretation has been the experimental ob-
servation of a frequency-dependent response in the dynamic shear
loading of cartilage 11,12,19,20; indeed, porous media models
of isotropic materials undergoing infinitesimal deformation pre-
dict an isochoric deformation under torsional shear, which would
preclude interstitial fluid pressurization and flow. Thus, the obser-
vation of a viscoelastic response in torsional shear should support
the premise of intrinsic viscoelasticity of the solid matrix 11,19.
However, it is important to recognize the limiting assumptions of
this analysis, namely that cartilage is not necessarily isotropic and
that for certain classes of anisotropy torsional shear can produce
nonzero dilatation, and that the prediction of isochoric deforma-
tion under infinitesimal strain is a mathematical idealization that
neglects higher order deformation effects that remain present ex-
perimentally i.e., higher order changes in dilatation may produce
non-negligible interstitial fluid pressurization and flow.
Another confounding factor in the assessment of intrinsic solid
matrix viscoelasticity has been the observation that other model-
ing assumptions of cartilage can equally improve agreement be-
tween theoretical predictions and experimental data. For example,
in confined and unconfined compression and indentation, good
agreement with experiments has been found not only with linear
isotropic poroviscoelastic biphasic models 14,17,21, but also
when using a linear transversely isotropic biphasic model 8,9,10,
or nonlinear bimodular biphasic models 7,22. Modeling the in-
Contributed by the Bioengineering Division for publication in the JOURNAL OF
BIOMECHANICAL ENGINEERING. Manuscript received by the Bioengineering Divi-
sion December 13, 2000; revised manuscript received May 16, 2001. Associate Edi-
tor: L. A. Setton.
410 Õ Vol. 123, OCTOBER 2001 Copyright © 2001 by ASME Transactions of the ASME
homogeneity of cartilage observed in compression 23–25 simi-
larly appears to have a potential for improving agreement between
the isotropic biphasic theory and experimental data in confined
and unconfined compression 26,27.
In our own studies, we have been able to demonstrate good
agreement between the infinitesimal or finite deformation isotro-
pic biphasic models 1,28 and experiments in confined compres-
sion creep and stress-relaxation 4,29, and dynamic loading 18,
not only by curve-fitting the corresponding experimental load or
deformation response but also by predicting the measured inter-
stitial fluid pressurization. For unconfined compression of cylin-
drical samples of cartilage, where the tissue is simultaneously
subjected to axial compression and radial and circumferential ten-
sion 30, we have recently proposed to incorporate the Conewise
Linear Elasticity CLE model of Curnier et al. 31 into the bi-
phasic theory of Mow et al. 1 to account for the disparity in the
tensile and compressive moduli of articular cartilage observed in
various studies e.g., 1,4,3239兴兲. This model was shown to pro-
duce good agreement between theory and experiments in uncon-
fined compression, including in its ability to predict the cartilage
interstitial fluid response 22. However, an interesting prediction
of this biphasic-CLE model, which is presented below, is that the
response of cartilage under uniaxial tension an experimental con-
figuration often investigated in the literature 32,33,3538,40but
not yet predicted successfully with a porous media model exhib-
its almost no transient response, unlike experimental observations
35,37,40. This finding suggests that uniaxial tension of articular
cartilage may indeed be a discriminating testing configuration for
investigating the intrinsic viscoelasticity of the solid matrix of
articular cartilage, in the context of a porous media model that
also describes the tension-compression nonlinearity of articular
cartilage.
Therefore, the short-term objective of this study is to under-
stand the role of flow-independent viscoelasticity in the context of
a porous media model of cartilage which accounts for its tension
compression nonlinearity. The long-term objective is to develop a
more comprehensive framework for understanding the mechanical
behavior of cartilage than the currently available theoretical mod-
els, which can better interpret the diverse experimental outcomes
reported in the literature; this framework should further help ex-
plain how cartilage is able to sustain the high compressive stresses
typical of in vivo loading conditions that far exceed its equilibrium
compressive modulus 41,42. The specific aims are to combine
existing theories of cartilage that incorporate intrinsic viscoelas-
ticity and tensioncompression nonlinearity of the solid matrix
within a biphasic model; and to investigate the response of such a
model to standard testing configurations and compare the out-
comes to experimental data reported in the literature.
Model Formulation
Mak 13 developed a formulation for an isotropic biphasic
model of cartilage whose solid phase is described by the quasi-
linear viscoelasticity QLV theory of Fung 43, with the tissue
modeled as a binary mixture of an intrinsically incompressible
solid phase, representing primarily the collagen fibers, proteogly-
cans, and chondrocytes, and an intrinsically incompressible fluid
phase representing the interstitial water 1. The governing equa-
tions for this model, known as the biphasic poroviscoelastic
BPVE theory, are the momentum equation for the mixture ne-
glecting inertia and in the absence of body forces,
⫽⫺p
v
e
0, (1)
where
represents the total stress tensor, which is the sum of the
interstitial fluid pressure p and the viscoelastic or effective stress
v
e
resulting from deformation of the solid matrix (
⫽⫺pI
v
e
), denotes the divergence operator, and is the gradient
operator; and the continuity equation for the mixture,
v
s
w
0, (2)
where w
f
(v
f
v
s
) is the flux of fluid relative to the solid,
f
is
the fluid volume fraction tissue porosity and v
s
, v
f
are the solid
and fluid phase velocities, respectively. In one of the simpler em-
bodiments of the QLV theory,
1
the viscoelastic stress tensor
v
e
can be related to the stress tensor under equilibrium conditions,
e
, through
v
e
t
g
t
e
E
0
0
t
g
t
e
E
d
. (3)
The infinitesimal strain tensor E, which appears above, is related
to the solid displacement u through E (1/2)(u u
T
), and the
displacement to the solid velocity through v
s
D
s
u/Dt material
derivative following the solid phase. The reduced relaxation
function, g(t), is given by
g
t
1 c
E
i
t
2
E
i
t
1
(4)
where E
i
() represents the exponential integral function, and c,
1
,
2
are material properties of the QLV theory. Physically,
1/
2
,1/
1
represents the frequency range over which most of the
intrinsic viscoelastic energy dissipation occurs under dynamic
loading, whereas (1 c ln
2
/
1
) is the ratio of instantaneous to
equilibrium moduli resulting from the intrinsic viscoelasticity
alone. Note that unlike the classical formulation of Fung 43,we
adopt the trivial modification of the function g(t) such that
g(0) 1 c ln(
2
/
1
) and g() 1, so that the viscoelastic stress
v
e
reduces to the elastic stress
e
at equilibrium.The remain-
ing constitutive relations adopted in the current formulation are
Darcy’s law,
w⫽⫺kp, (5)
which relates the fluid flux to the pressure gradient, with k repre-
senting the hydraulic permeability assumed isotropic and con-
stant here, though it is generally recognized to be strain-dependent
44兴兲, and the Conewise Linear Elasticity model of Curnier et al.
31, in its cubic symmetry embodiment 22,
e
E
a1
3
1
A
a
:E
tr
A
a
E
A
a
b1
ba
3
2
tr
A
a
E
A
b
2
E, (6)
which describes a bimodular response, or tensioncompression
nonlinearity, of the solid matrix. Here, tr() is the trace operator
that yields the first invariant of its tensorial argument, and A
a
:E
tr(A
a
T
E). A
a
is a texture tensor corresponding to each of three
preferred material directions defined by the unit vectors a
a
a
a
a
a
1, no sum over a, denoting the dot product of vectors,
with A
a
a
a
a
a
denoting the dyadic product of vectors, no
sum over a. For cubic material symmetry, a
a
a
b
0 when b
a, and the three directions are generally taken to be: a
1
parallel
to the split line direction,
2
a
2
perpendicular to the split line direc-
tion, and a
3
normal to the articular cartilage surface. The term
A
a
:E represents the component of normal strain along the pre-
ferred direction a
a
. Tensioncompression nonlinearity stems
from the conditional statement,
1
The function g(t) in Eq. 3is taken to be a scalar function for simplicity in this
analysis. A most general formulation could employ a fourth-order tensor instead; for
cartilage, it has sometimes been proposed to separate the viscoelastic response in
bulk deformation from that in shear deformation when using an isotropic model
13,15, though this is not done here.
2
Split lines have been used as indicators of predominant collagen fibril directions
on the articular surface Hulkrantz, W., 1898, ‘‘Ueber die Spaltrichtungen der Gelen-
kknorpel,’’ Verh. D. Anat. Ges., 12, pp. 248256.
Journal of Biomechanical Engineering OCTOBER 2001, Vol. 123 Õ 411
1
A
a
:E
1
, A
a
:E 0
1
, A
a
:E 0
. (7)
This signifies that the material properties
1
differ whether the
normal strain component along the direction a
a
is compressive or
tensile. The physical meaning of these elastic constants is as fol-
lows: H
A
1
2
is the equilibrium confined compression
modulus of the tissue the ‘aggregate’ modulus and H
A
1
2
is the equivalent modulus in tension;
2
is the ‘off-
diagonal’’ modulus, which could be determined from the equilib-
rium ratio of radial stress to axial strain in confined compression
the radial stress being measurable on the side wall, e.g., see the
experiments of Khalsa and Eisenberg 45兴兲. Note that the choice
of an isotropic permeability is consistent with the choice of cubic
symmetry for the stressstrain law, since isotropic and cubic sym-
metry are identical in second-order tensors such as the permeabil-
ity tensor; however, it is also possible to adopt a more general
orthotropic model for permeability, as in our earlier study 22.
In summary, the model presented above, which is valid for
infinitesimal strains and can describe tensioncompression non-
linearity Eqs. 6and 7兲兴 as well as intrinsic viscoelasticity Eqs.
3 and 4兲兴 of the solid matrix of a solid-fluid biphasic mixture,
has eight material constants:
1
,
1
,
2
,
,c,
1
,
2
,k. This
model can be reduced to the isotropic biphasic poroviscoelastic
model of Mak 13 by letting
1
1
2
⬅␭ and noting that
A
1
A
2
A
3
I. It can be reduced to our recently proposed
biphasic-CLE model 22 by letting c 0. It can also be reduced
to the linear isotropic biphasic theory of Mow et al.1 by imple-
menting both of the reductions described above.
Uniaxial Tension and Unconfined Compression
Typical experiments of unconfined compression of articular car-
tilage are performed on cylindrical samples e.g., 1 mm thick, 6
mm in diameter, which can be easily harvested with a circular
core punch cutting perpendicularly to the articular surface e.g.,
1兴兲. In contrast, uniaxial tensile tests are performed on long strips
of cartilage, either dumbbell shaped or prismatic e.g, 10 mm
long, with a rectangular cross section e.g., 0.21.5 mm, which
are typically harvested with a pair of blades perpendicularly to the
surface e.g., 33,35兴兲. The solution of Armstrong et al. 30 for
unconfined compression of a linear isotropic biphasic material, as
well as many subsequent solutions e.g., 7,8兴兲 assumed friction-
less conditions at the loading platens, which is reasonable in view
of the typically low friction coefficient of articular cartilage. This
simplification, along with the assumption of axisymmetric condi-
tions facilitated by the specimen geometry, leads to equations
amenable to a closed-form analytical solution. In contrast, the
biphasic or biphasic poroviscoelastic analysis of the uniaxial ten-
sile response of cartilage has generally been performed on pris-
matic bars with rectangular cross sections, requiring either a sim-
plification of the boundary conditions 46or a numerical scheme
such as finite element analysis 47; the reason is that the analysis
of the biphasic response of a prismatic bar with rectangular cross
section, whose lateral boundaries are free draining, is fully three
dimensional and does not lend itself to an analytical closed-form
solution because of the complexity of the transient interstitial fluid
flow fields that would result from loading. Therefore, in order to
achieve a closed-form solution for the biphasic-CLE-QLV analy-
sis of uniaxial tension for the purpose of examining the different
responses of cartilage in tension and compression, we assume in
this study that the cartilage specimen is a prismatic bar with a
circular cross section, while recognizing that such a specimen
geometry cannot be easily obtained in practice by typical speci-
men preparation. It will be demonstrated below that this assump-
tion is not as restrictive as it first may seem. By St. Venant’s
principle, the effects of the clamping conditions at the two ends of
the prismatic bar are also neglected, since the specimen length is
typically much greater than its characteristic cross-sectional
width, so that the analyses of uniaxial tension and unconfined
compression are treated in a similar fashion, with the only differ-
ence in those two configurations arising from the bimodular con-
stitutive assumptions of the CLE theory.
The reduction of the general biphasic equations to the configu-
ration of unconfined compression with frictionless platens and
axisymmetric conditions has been described previously for the
linear isotropic biphasic model 30, the poroviscoelastic biphasic
model 21, and the biphasic-CLE model 22. In these analyses,
the shear traction at the interface between cartilage and the load-
ing platens is set to zero and the axial normal strain is homoge-
neous; the interstitial fluid pressure and normal traction are also
set to zero on the lateral boundary. The same approach can be
followed with the constitutive equations of Eqs. 36, hence
only a summary of the results is presented here. As is typical for
this type of problems, the governing equations and closed-form
solution is given in Laplace transform space. The differential
equation for the radial displacement is given by
2
u
¯
r
r
2
1
r
u
¯
r
r
u
¯
r
r
2
f
u
¯
r
rf
2
¯
s
, (8)
whereas the boundary conditions reduce to
u
¯
r
r0
0, H
⫿ A
u
¯
r
r
2
u
¯
r
r
¯
s
rr
0
0. (9)
The fluid pressure can be determined from the radial displacement
using
p
¯
r,s
⫽⫺
1
k
r
r
0
su
¯
r
,s
s
¯
s
/2
d
. (10)
The solution then reduces to
u
¯
r
r,s
r
0
2
1
2
H
⫿ A
I
1
f
r
r
0
f
I
0
f
1
2
H
⫿ A
I
1
f
r
r
0
¯
s
,
(11)
p
¯
r,s
H
⫿ A
2
1 c ln
1
2
s
1
1
s
f
1
2
H
⫿ A
I
0
f
r
r
0
I
0
f
f
I
0
f
1
2
H
⫿ A
I
1
f
¯
s
,
(12)
F
¯
s
r
0
2
H
⫿ A
1 c ln
1
2
s
1
1
s
2H
A
3
2
H
⫿ A
2H
⫿ A
f
I
0
f
1
2
H
⫿ A
冊冉
2
2
H
A
H
⫿ A
H
⫿ A
I
1
f
f
I
0
f
1
2
H
⫿ A
I
1
f
¯
s
,
(13)
412 Õ Vol. 123, OCTOBER 2001 Transactions of the ASME
F
¯
p
s
r
0
2
H
⫿ A
1 c ln
1
2
s
1
1
s
1
2
H
⫿ A
I
1
f
1
2
f
I
0
f
f
I
0
f
1
2
H
⫿ A
I
1
f
¯
s
,
(14)
where
f
r
0
2
s
H
⫿ A
k
1 c ln
1
2
s
1
1
s
. (15)
In these equations, r is the radial coordinate, u
r
refers to the radial
displacement, p is the interstitial fluid pressure, is the axial
strain, F is the total axial force across the specimen, and F
p
is that
component of the force supported by interstitial fluid pressure,
i.e., F
p
2
0
r
0
rpdr, where r
0
is the specimen radius. I
0
() and
I
1
() are modified Bessel functions of the first kind, of order 0
and 1, respectively. Overbars indicate Laplace transformation
from the time domain and s is the Laplace transform variable.
Superscripted on these parameters refer to the solution for ten-
sion or compression , as it can be observed that these
solutions differ by the interchange of the material constants H
A
1
2
and H
A
1
2
. Note that the solution does not
depend on the axial dimension thickness or length of the cylin-
drical specimen.
These solutions are valid for a variety of loading conditions.
For example, for a step application of strain with a magnitude
0
,
use
¯
(s)⫽⫾
0
/s; if the strain is ramped over a ramp time of t
0
and subsequently kept constant at a magnitude of
0
, use
¯
(s)
⫽⫾(1 e
st
0
)
0
/s
2
t
0
. Alternatively, the total axial load F
¯
(s)
may be prescribed in a similar way, and a solution for
¯
(s)
obtained from Eq. 13 then substituted into the remaining
expressions.
For step or ramp strain application, inverse Laplace transforma-
tion into the time domain can be performed numerically, e.g.,
using the INLAP routine from the IMSL library Visual Numerics,
Inc., Houston, TX.
3
For the steady-state solution to a sinusoidal
axial load or strain, it suffices to substitute s i
into the solu-
tions of Eqs. 1115, where
is the angular frequency of the
applied load or strain and i
1, to derive the frequency-
dependent amplitude and phase response of the corresponding
parameter.
The dynamic modulus G
¯
(s) of this biphasic-CLE-QLV mate-
rial can be derived from the ratio of F
¯
(s)/
r
0
2
and
¯
(s) in Eq.
13. To determine the material’s ‘‘instantaneous’ modulus or the
modulus in the limit of loading at high frequency, denoted by
E
Y
0
, it suffices to take the limit of the resulting expression as s
, which corresponds to the real time limit of t 0
:
E
Y
0
1 c ln
2
1
冊冉
H
A
3
2
2
H
⫿ A
2
. (16)
Similarly, the modulus at equilibrium or in the limit of loading at
very low frequency, E
Y
Young’s modulus in classical elastic-
ity, can be obtained by taking the limit of the dynamic modulus
as s 0,
E
Y
H
A
2
2
2
H
⫿ A
2
. (17)
It can be observed that, as expected, the equilibrium modulus is
independent of the QLV parameters. Finally, the instantaneous or
high-frequency fluid load support can be obtained by taking the
limit, as s , of the ratio of the expressions in Eqs. 14 and
13,
F
¯
p
F
¯
1
1 2
H
A
2
H
⫿ A
2
. (18)
It is noteworthy that this expression is independent of the QLV
parameters, even though it represents an instantaneous response.
Results
A variety of testing configurations in tension and compression
can be simulated from the solutions described above, a subset of
which are presented here to provide sufficient insight into the
behavior of this biphasic-CLE-QLV model. Since this compound
model has not been employed previously in the literature, repre-
sentative material properties used in the simulations here derive
from two separate sources: For the biphasic-CLE properties, we
employ the results of our recent analysis 22, derived from con-
fined and unconfined compression stress-relaxation and torsional
shear experiments on bovine articular cartilage; H
A
13.2 MPa, H
A
0.64 MPa,
2
0.48 MPa,
0.17 MPa, k
6.1 10
16
m
4
/N.s; these properties were obtained under a total
strain of approximately 18 percent including tare loading. For
the QLV parameters, we use the results of Setton et al. 14 who
curve-fit the biphasic poroviscoelastic theory of Mak 13 to con-
fined compression creep data, also on bovine articular cartilage:
c 0.16,
1
0.06 s,
2
201 s.
In uniaxial tension, a representative radius of r
0
0.345 mm is
employed, which produces a surface area equivalent to a rectan-
gular cross section of dimensions 1.5 mm 0.25 mm. In the first
analysis, a step tensile strain of magnitude
0
0.10 10 percent
is applied to the sample and the time-dependent stress-relaxation
response of the biphasic-CLE-QLV model is presented in Fig. 1,
together with the specialized case of the biphasic-CLE model
with c 0. It can be noted that, unlike the biphasic-CLE-QLV
response, the biphasic-CLE response exhibits almost no transient
relaxation under this testing configuration. Substitution of these
material constants into Eq. 16 confirms that the instantaneous
tensile modulus of the biphasic-CLE-QLV material, E
Y
0
3
Numerical inverse Laplace transformation is best achieved by nondimensional-
izing the expressions in Eqs. 612 to avoid numerical overflow or underflow.
Though IMSL has a built-in routine to evaluate Bessel functions with a complex
argument, a custom-written routine implementing asymptotic expansions for large
arguments was employed instead.
Fig. 1 Uniaxial tensile response of the biphasic-CLE-QLV and
biphasic-CLE models, to a step strain of
0
Ä0.10 10 percent,
as derived from Eq. 13 by numerical inverse Laplace transfor-
mation
Journal of Biomechanical Engineering OCTOBER 2001, Vol. 123 Õ 413
29.4 MPa, is considerably greater than that of the biphasic-CLE
material, E
Y
0
12.8 MPa. In contrast, for both models, the equi-
librium modulus in tension is E
Y
12.3 MPa. The dynamic
modulus in tension is displayed for both models in Fig. 2, which
displays the amplitude and phase angle as a function of frequency
over the range f
/2
10
6
10
2
Hz. For reference, the three
characteristic frequencies for the material are f
H
A
k/r
0
2
0.0033 Hz, 1/
2
0.005 Hz, 1/
1
16.7 Hz. As already evi-
denced by the values of E
Y
0
and E
Y
, there is virtually no fre-
quency dependence of the dynamic tensile modulus in the
biphasic-CLE model, whereas the inclusion of QLV produces a
characteristic flow-independent viscoelastic response.
In unconfined compression, a radius of r
0
2.39 mm is
assumed.
4
In the first of these analyses, a ramped-strain stress-
relaxation test is employed, with
0
0.10 and t
0
3 s, 150 s, and
300 s. The resulting stress-relaxation responses are presented in
Fig. 3, with and without QLV effects. Unlike the case for tension,
a very significant transient response is observed in unconfined
compression in both models. Differences in the two models are
most evident only in the fastest of the ramp rates employed (t
0
3 s), with the peak stress at the end of the ramp achieving a
larger value for the biphasic-CLE-QLV model. The amplitude and
phase angle of the unconfined compression dynamic modulus is
presented in Fig. 4, for both models; for this testing configuration,
the three characteristic frequencies for the material are f
H
A
k/r
0
2
0.0014 Hz, 1/
2
0.005 Hz, 1/
1
16.7 Hz. From
this figure, as from Eq. 16, it can be observed that E
Y
0
15.0 MPa for the biphasic-CLE-QLV model, and E
Y
0
6.5 MPa for the biphasic-CLE model, whereas the equilibrium
compressive modulus is only E
Y
0.57 MPa for both cases.
Discussion
The short-term objective of this study was to investigate the
role of intrinsic viscoelasticity of the solid matrix of cartilage
4
Representative radii are not required when performing these analyses with the
nondimensional form of the equations; they are used here for illustration.
Fig. 2 Dynamic modulus versus frequency under uniaxial ten-
sile loading, for the biphasic-CLE-QLV and biphasic-CLE mod-
els:
a
magnitude;
b
phase angle. In the limit of high frequen-
cies, the magnitude of the dynamic modulus is given by
E
¿
Y
0
¿
in
Eq. 16, whereas at low frequencies it is given by
E
¿
Y
in Eq.
17
Fig. 3 Unconfined compression stress-relaxation response of
the biphasic-CLE-QLV and biphasic-CLE models, to a ramp
strain of magnitude
0
Ä0.10 10 percent, and for three differ-
ent ramp times
t
0
Ä3 s, 150 s, and 300 s, as derived from Eq.
13 by numerical inverse Laplace transformation
Fig. 4 Dynamic modulus versus frequency under unconfined
compression loading, for the biphasic-CLE-QLV and biphasic-
CLE models:
a
magnitude;
b
phase angle. In the limit of high
frequencies the magnitude of the dynamic modulus is given by
E
À
Y
0
¿
in Eq. 16, whereas at low frequencies it is given by
E
À
Y
in
Eq. 17
414 Õ Vol. 123, OCTOBER 2001 Transactions of the ASME
within the context of a biphasic mixture theory that accounts for
tensioncompression nonlinearity. The long-term objectives are to
help elucidate the mechanism by which articular cartilage can
sustain typical physiological stresses in the range of 26 MPa
e.g., 41,42兴兲, with occasional peak values as high as 18 MPa
48, while exhibiting a compressive equilibrium modulus on the
order of 0.5 MPa only; and to provide a theoretical framework
that can simultaneously predict the tensile and compressive re-
sponses of articular cartilage, under transient and dynamic condi-
tions. As discussed below, the results of this study provide a more
comprehensive agreement between theory and experiments in the
literature, than achieved from the linear biphasic, the QLV, the
biphasic-QLV biphasic poroviscoelastic, or the biphasic-CLE
theories alone.
First, because the biphasic-CLE-QLV subsumes the linear bi-
phasic theory, it can provide the same successful agreement with
confined compression experiments as demonstrated previously
e.g., 1,2,18,29,49兴兲. Second, because the biphasic-QLV and the
biphasic-CLE are subsumed by the more general model, good
agreement can also be expected in the curvefitting of unconfined
compression stress-relaxation results 17,22. However, under dy-
namic unconfined compression, cartilage has been shown to ex-
hibit a dynamic compressive modulus on the order of 1220 MPa
50,51at the highest tested frequencies. Looking at Eq. 16with
c 0, the highest achievable dynamic modulus with the biphasic-
CLE model is E
Y
0
H
A
/2 given that H
A
,
2
H
A
, which
puts it on the order of 6.5 MPa when using the representative
value of H
A
employed in the simulations given above see the
high-frequency range in Fig. 4. With the biphasic-QLV i.e., let-
ting H
A
H
A
in Eq. 16兲兲, the highest predicted dynamic
modulus in compression would be E
Y
0
(3/2)(1 c ln
2
/
1
)(H
A
2
), which is on the order of 0.5 MPa for the typical
material constants used above. Clearly, while the biphasic-CLE
model is a better predictor of the dynamic unconfined compres-
sion modulus than the biphasic-QLV, it still falls somewhat short
of the values observed experimentally. However, when using the
combined biphasic-CLE-QLV model, Eq. 16 produces a dy-
namic unconfined compression modulus E
Y
0
on the order of 15
MPa Fig. 4, in better agreement with experiments in the
literature.
Looking at Eq. 17, it can be observed that the biphasic-CLE
theory can account for the differences observed in the literature
between the compressive and tensile equilibrium moduli of carti-
lage e.g., E
Y
0.55 MPa and E
Y
12.3 MPa, using the above-
given typical material constants, represented for typical samples
of humeral head cartilage in Fig. 5; clearly, neither the linear
biphasic nor the biphasic-QLV models account for tension
compression nonlinearity, so they are unable to model both ten-
sion and compression using consistent material constants. How-
ever, as observed in Fig. 1, the biphasic-CLE is unable to produce
a substantial transient response under a step strain in tension, even
though such responses have always been observed in the literature
e.g., 35,37,40, also see the typical experimental response in Fig.
6, whereas the biphasic-CLE-QLV can account for this effect. It
is important to appreciate that in tension, the flow-dependent vis-
coelasticity is curtailed by the large difference between the tensile
and compressive moduli in the biphasic-CLE and biphasic-CLE-
QLV theories, whereas in compression, the fluid pressurization is
enhanced by the tensioncompression nonlinearity. This is evi-
dent from the fluid load support as given in Eq. 18; whereas in
unconfined compression the fluid load support for the
biphasic-CLE or biphasic-CLE-QLV models is very high
(
F
¯
p
/F
¯
97.6 percent), it becomes virtually negligible in
tension (
F
¯
p
/F
¯
0.6 percent). For the linear biphasic or the
biphasic-QLV, the fluid load support in tension and compression
remains the same,
F
¯
p
/F
¯
33.3 percent; indeed, the latter
theories would predict the same transient relaxation in tension as
in unconfined compression, though they otherwise appear to be
unsuitable for predicting the tensioncompression nonlinearity as
discussed above.
Since flow-dependent viscoelasticity is negligible when
H
A
/H
A
1, uniaxial tension experiments may conceivably be
used to extract only the QLV parameters, c,
1
,
2
, and the equi-
librium tensile modulus, E
Y
, even when using the framework of
the biphasic-CLE-QLV theory. This simplification in the analysis
of experimental data in uniaxial tension also implies that the pre-
cise geometry of the specimen cross section would not impact the
experimental outcome since fluid flow effects, which otherwise
depend on geometry, are negligible. The results of this analysis
also suggest that for biological hydrated soft tissues, which gen-
erally sustain very little compressive loads and thus exhibit very
significant tensioncompression nonlinearity, such as tendons and
ligaments, flow-dependent viscoelasticity may not be a significant
effect.
The results of this study are primarily based on a theoretical
analysis at this time, though favorable qualitative comparisons are
observed relative to literature findings. Experiments that directly
Fig. 5 Typical experimental equilibrium tensile and compres-
sive responses of human humeral head articular cartilage from
the superficial and middle zones 55 y.o. male. The tensile re-
sponse is measured from long prismatic samples with rectan-
gular cross section, harvested parallel to the surface, along the
split line direction; the compressive response is measured
from cylindrical plugs harvested normal to the articular surface
34.
Fig. 6 Experimental uniaxial tensile response of the superfi-
cial zone specimen of Fig. 5 to a ramp strain of
0
Ä0.02 2
percent and ramp time of
t
0
Ä1 s, as a function of time
Journal of Biomechanical Engineering OCTOBER 2001, Vol. 123 Õ 415
test the predictions of the biphasic-CLE-QLV model represent the
next step in this effort to establish a more comprehensive theoret-
ical framework for articular cartilage. Recently, models that have
accounted for the tensioncompression nonlinearity of cartilage
7,22, such as the biphasic-CLE model reviewed above, have
demonstrated that interstitial fluid load support is considerably
enhanced in compression by the disparity in tensile and compres-
sive moduli of cartilage, producing a greater dynamic modulus.
The current study demonstrates that the addition of intrinsic vis-
coelasticity provides the necessary boost to the theoretical predic-
tion of the dynamic modulus to match experimental findings at
higher frequencies. In the biphasic-CLE-QLV model adopted in
this study, the dynamic modulus is greater than that of the
biphasic-CLE model by a factor of (1 c ln
2
/
1
) according to
Eq. 16. Interestingly, in this model, the incorporation of solid
matrix intrinsic viscoelasticity neither enhances nor defeats the
fluid load support, as indicated by Eq. 18. As noted by others
17, and evidenced by comparing the biphasic-CLE and biphasic-
CLE-QLV results of Fig. 3 or 4, it appears that the effect of
intrinsic solid matrix viscoelasticity is most evident at relatively
fast strain rates e.g., with a ramp time of t
0
3 s in Fig. 3 in
stress-relaxation, or at dynamic frequencies above 10
3
Hz ac-
cording to Fig. 4; the stress-relaxation results of Fig. 3 suggest
that a combination of stress-relaxation experiments with a slow
ramp time and a fast ramp time could also be used to discriminate
between flow-dependent and flow-independent effects see a typi-
cal experimental response in Fig. 7, though not as dramatically as
uniaxial tensile tests Fig. 1 or Fig. 6.
There are a number of limitations to the current study. For
example, the assumption of a linear response in tension is known
to be a simplification, since cartilage has been shown to behave
nonlinearly even at small tensile strains; furthermore, the current
model does not adequately describe the anisotropy of cartilage,
such as the known disparity in tensile moduli measured parallel
and perpendicular to the split lines, as well as the observation of
tensile Poisson’s ratios in excess of 0.5, though these could be
potentially addressed using an orthotropic model. The known in-
homogeneity of cartilage properties from the surface to the deep
zone would need to be taken into account to produce more accu-
rate results. A finite strain formulation of the constitutive equa-
tions would also be more appropriate to investigate higher physi-
ological loads. These issues, which typically increase the
complexity of cartilage modeling, can be addressed in future
analyses.
Conclusion
This study attempts to explain various observed experimental
findings in the testing of articular cartilage, by combining several
constitutive models previously described in the literature, each of
which had demonstrated success in one or more testing configu-
rations. By comparing theoretical predictions with experimental
findings in the literature, it is found that a simultaneous prediction
of compression and tension experiments, under stress-relaxation
and dynamic loading, can be achieved when taking into account
flow-dependent and flow-independent viscoelasticity effects, as
well as tensioncompression nonlinearity. While a guiding prin-
ciple of constitutive modeling of biological tissues is to adopt the
simplest possible formulation that can describe experimental data,
it is becoming increasingly clear that the complexity of articular
cartilage mechanics requires more elaborate models than those
already presented in the literature. The biphasic-CLE-QLV model
adopted in this study has the advantage that it employs as building
blocks theories that have proved popular and are thus familiar to
the concerned research community. In the embodiment adopted
here, the model has eight material constants; clearly, this signifies
that an experimental characterization of the material properties of
a particular tissue sample could not come from a single test but
would rather require multiple experiments, either on the same
sample or on adjoining samples from the same tissue source. The
combination of tests that might provide sufficient data to deter-
mine these constants uniquely is the subject of ongoing studies.
Acknowledgments
This study was supported in part by the National Institute for
Arthritis, and Musculoskeletal and Skin Diseases of the National
Institutes of Health AR46532, AR43628, AR42850, and
AR41913.
References
1 Mow, V. C., Kuei, S. C., Lai, W. M., and Armstrong, C. G., 1980, ‘Biphasic
Creep and Stress Relaxation of Articular Cartilage in Compression: Theory
and Experiments,’’ ASME J. Biomech. Eng., 102, pp. 7384.
2 Frank, E. H., and Grodzinsky, A. J., 1987, ‘‘Cartilage Electromechanics—II. A
Continuum Model of Cartilage Electrokinetics and Correlation With Experi-
ments,’’ J. Biomech., 20, pp. 629639.
3 Lai, W. M., Hou, J. S., and Mow, V. C., 1991, ‘A Triphasic Theory for the
Swelling and Deformation Behaviors of Articular Cartilage,’’ ASME J. Bio-
mech. Eng., 113, pp. 245258.
4 Ateshian, G. A., Warden, W. H., Kim, J. J., Grelsamer, R. P., and Mow, V. C.,
1997, ‘Finite Deformation Biphasic Material Properties of Bovine Articular
Cartilage From Confined Compression Experiments,’ J. Biomech., 30, pp.
11571164.
5 Huyghe, J. M., and Janssen, J. D., 1997, ‘‘Quadriphasic Mechanics of Swelling
Incompressible Porous Media,’’ Int. J. Eng. Sci., 35, pp. 793802.
6 Gu, W. Y., Lai, W. M., and Mow, V. C., 1998, ‘A Mixture Theory for Charged
Hydrated Soft Tissues Containing Multi-electrolytes: Passive Transport and
Swelling Behaviors,’’ ASME J. Biomech. Eng., 102, pp. 169180.
7 Soulhat, J., Buschmann, M. D., and Shirazi-Adl, A., 1999, ‘A Fibril-Network
Reinforced Model of Cartilage in Unconfined Compression,’’ ASME J. Bio-
mech. Eng., 121, pp. 340347.
8 Cohen, B., Lai, W. M., and Mow, V. C., 1998, ‘A Transversely Isotropic
Biphasic Model for Unconfined Compression of Growth Plate and Chon-
droepiphysis,’’ ASME J. Biomech. Eng., 120, pp. 491496.
9 Cohen, B., Gardner, T. R., and Ateshian, G. A., 1993, ‘The Influence of
Transverse Isotropy on Cartilage Indentation Behavior—A Study of the Hu-
man Humeral Head,’’ Trans. Orthop. Res. Soc., 18, p. 185.
10 Mow, V. C., Good, P. M., and Gardner, T. R., 2000, ‘A New Method to
Determine the Tensile Properties of Articular Cartilage Using the Indentation
Test,’’ Trans. Orthop. Res. Soc., 25, p. 103.
11 Hayes, W. C., and Mockros, L. F., 1971, ‘Viscoelastic Properties of Human
Articular Cartilage,’’ J. Appl. Physiol., 31, pp. 562568.
12 Hayes, W. C., and Bodine, A. J., 1978, ‘Flow-Independent Viscoelastic Prop-
erties of Articular Cartilage Matrix,’’ J. Biomech., 11, pp. 407419.
13 Mak, A. F., 1986, ‘The Apparent Viscoelastic Behavior of Articular
Cartilage—The Contributions From the Intrinsic Matrix Viscoelasticity and
Interstitial Fluid Flows,’’ ASME J. Biomech. Eng., 108, pp. 123130.
14 Setton, L. A., Zhu, W., and Mow, V. C., 1993, ‘The Biphasic Poroviscoelastic
Behavior of Articular Cartilage: Role of the Surface Zone in Governing the
Compressive Behavior,’’ J. Biomech., 26, pp. 581592.
15 Suh, J.-K., and DiSilvestro, M. R., 1999, ‘Biphasic Poroviscoelastic Behavior
of Hydrated Biological Soft Tissue,’ ASME J. Appl. Mech., 66, pp. 528535.
Fig. 7 Typical experimental unconfined compression stress-
relaxation response of bovine glenohumeral articular cartilage,
to a ramp strain of magnitude
0
Ä0.05 5 percent, and for two
different ramp times
t
0
Ä0.126 s and
t
0
Ä400 s
416 Õ Vol. 123, OCTOBER 2001 Transactions of the ASME
16 Suh, J.-K., and Bai, S., 1997, ‘‘Biphasic Poroviscoelastic Behavior of Articular
Cartilage in Creep Indentation Test,’’ Trans. Orthop. Res. Soc., 22, p. 823.
17 DiSilvestro, M. R., Zhu, Q., and Suh, J.-K., 1999, ‘Biphasic Poroviscoelastic
Theory Predicts the Strain Rate Dependent Viscoelastic Behavior of Articular
Cartilage,’Proc. 1999 Bioeng. Conf., ASME BED-Vol. 42, pp. 105106.
18 Soltz, M. A., and Ateshian, G. A., 2000, ‘Interstitial Fluid Pressurization Dur-
ing Confined Compression Cyclical Loading of Articular Cartilage,’ Ann.
Biomed. Eng., 28, pp. 150159.
19 Zhu, W., Mow, V. C., Koob, T. J., and Eyre, D. R., 1993, ‘Viscoelastic Shear
Properties of Articular Cartilage and the Effects of Glycosidase Treatments,’ J.
Orthop. Res., 11, pp. 771781.
20 Setton, L. A., Mow, V. C., and Howell, D. S., 1995, ‘Mechanical Behavior of
Articular Cartilage in Shear Is Altered by Transection of the Anterior Cruciate
Ligament,’’ J. Orthop. Res., 13, pp. 473482.
21 Mak, A. F., 1986, ‘‘Unconfined Compression of Hydrated Viscoelastic Tissues:
A Biphasic Poroviscoelastic Analysis,’’ Biorheology, 23, pp. 371383.
22 Soltz, M. A., and Ateshian, G. A., 2000, ‘A Conewise Linear Elasticity Mix-
ture Model for the Analysis of Tension-Compression Nonlinearity in Articular
Cartilage,’’ ASME J. Biomech. Eng., 122, pp. 576586.
23 Guilak, F., Ratcliffe, A., and Mow, V. C., 1995, ‘Chondrocyte Deformation
and Local Tissue Strain in Articular Cartilage: A Confocal Microscopy Study,’
J. Orthop. Res., 13, pp. 410421.
24 Schinagl, R. M., Gurskis, D., Chen, A. C., and Sah, R. L., 1997, ‘Depth-
Dependent Confined Compression Modulus of Full-Thickness Bovine Articu-
lar Cartilage,’’ J. Orthop. Res., 15, pp. 499506.
25 Wang, C. C.-B., Soltz, M. A., Mauck, R. L., Valhmu, W. B., Ateshian, G. A.,
and Hung, C. T., 2000, ‘Comparison of Equilibrium Axial Strain Distribution
in Articular Cartilage Explants and Cell-Seeded Alginate Disks Under Uncon-
fined Compression,’’ Trans. Orthop. Res. Soc., 25, p. 131.
26 Wang, C. C.-B., Hung, C. T., and Mow, V. C., 2001, ‘An Analysis of the
Effects of Depth-Dependent Aggregate Modulus on Articular Cartilage Stress-
Relaxation Behavior in Compression,’’ J. Biomech., 34, pp. 7584.
27 Li, L. P., Buschmann, M. D., and Shirazi-Adl, A., 2000, ‘A Fibril Reinforced
Nonhomogeneous Poroelastic Model for Articular Cartilage: Inhomogeneous
Response in Unconfined Compression,’’ J. Biomech., 33, pp. 15331541.
28 Holmes, M. H., and Mow, V. C., 1990, ‘The Nonlinear Characteristics of Soft
Gels and Hydrated Connective Tissues in Ultrafiltration,’ J. Biomech., 23,
pp. 11451156.
29 Soltz, M. A., and Ateshian, G. A., 1998, ‘Experimental Verification and The-
oretical Prediction of Cartilage Interstitial Fluid Pressurization at an Imperme-
able Contact Interface in Confined Compression,’ J. Biomech., 31, pp. 927
934.
30 Armstrong, C. G., Mow, V. C., and Lai, W. M., 1984, ‘An Analysis of Un-
confined Compression of Articular Cartilage,’’ ASME J. Biomech. Eng., 106,
pp. 165173.
31 Curnier, A., He, Q.-C., and Zysset, P., 1995, ‘Conewise Linear Elastic Mate-
rials,’’ J. Elast., 37, pp. 138.
32 Kempson, G. E., Freeman, M. A., and Swanson, S. A., 1968, ‘Tensile Prop-
erties of Articular Cartilage,’ Nature London, 220, pp. 1127–1128.
33 Woo, S. L.-Y., Akeson, W. H., and Jemmott, G. F., 1976, ‘Measurements of
Nonhomogeneous, Directional Mechanical Properties of Articular Cartilage in
Tension,’’ J. Biomech., 9, pp. 785791.
34 Armstrong, C. G., and Mow, V. C., 1982, ‘‘Variations in the Intrinsic Mechani-
cal Properties of Human Articular Cartilage With Age, Degeneration, and Wa-
ter Content,’’ J. Bone Jt. Surg., Am. Vol., 64A, pp. 8894.
35 Akizuki, S., Mow, V. C., Muller, F., Pita, J. C., Howell, D. S., and Manicourt,
D. H., 1986, ‘Tensile Properties of Human Knee Joint Cartilage: I. Influence
of Ionic Conditions, Weight Bearing, and Fibrillation on the Tensile Modulus,’
J. Orthop. Res., 4, pp. 379-392.
36 Akizuki, S., Mow, V. C., Muller, F., Pita, J. C., and Howell, D. S., 1987,
‘Tensile Properties of Human Knee Joint Cartilage. II. Correlations Between
Weight Bearing and Tissue Pathology and the Kinetics of Swelling,’ J. Orthop.
Res., 5, pp. 173-186.
37 Schmidt, M. B., Mow, V. C., Chun, L. E., and Eyre, D. R., 1990, ‘Effects of
Proteoglycan Extraction on the Tensile Behavior of Articular Cartilage,’ J.
Orthop. Res., 8, pp. 353363.
38 Huang, C.-Y., Stankiewicz, A., Ateshian, G. A., Flatow, E. L., Bigliani, L. U.,
and Mow, V. C., 1999, ‘Tensile and Compressive Stiffness of Human Gleno-
humeral Cartilage Under Finite Deformation,’ Proc. 1999 Bioeng. Conf.,
ASME BED-Vol. 42, pp. 469470.
39 Soltz, M. A., Palma, C., Barsoumian, S., Wang, C. C.-B., Hung, C. T., and
Ateshian, G. A., 2000, ‘Multi-Axial Loading of Bovine Articular Cartilage in
Unconfined Compression,’’ Trans. Orthop. Res. Soc., 25, p. 888.
40 Woo, S. L.-Y, Simon, B. R., Kuei, S. C., and Akeson, W. H., 1980, ‘Quasi-
Linear Viscoelastic Properties of Normal Articular Cartilage,’’ ASME J. Bio-
mech. Eng., 102, pp. 8590.
41 Ahmed, A. M., and Burke, D. L., 1983, ‘In-Vitro Measurement of Static
Pressure Distribution in Synovial Joints—Part I: Tibial Surface of the Knee,’
ASME J. Biomech. Eng., 105, pp. 216-225.
42 Huberti, H. H., and Hayes, W. C., 1984, ‘Patellofemoral Contact Pressures.
The Influence of Q-Angle and Tendofemoral Contact,’’ J. Bone Jt. Surg., Am.
Vol., 66A, pp. 715724.
43 Fung, Y. C., 1981, Biomechanics: Mechanical Properties of Living Tissues,
Springer-Verlag, New York.
44 Mansour, J. M., and Mow, V. C., 1976, ‘The Permeability of Articular Carti-
lage Under Compressive Strain and at High Pressures,’’ J. Bone Jt. Surg., Am.
Vol., 58A, pp. 509516.
45 Khalsa, P. S., and Eisenberg, S. R., 1997, ‘Compressive Behavior of Articular
Cartilage Is Not Completely Explained by Proteoglycan Osmotic Pressure,’’ J.
Biomech., 30, pp. 589594.
46 Mak, A. F., 1985, ‘Uniaxial Tension of Hydrated Viscoelastic Tissues,’ASME
Adv. Bioengng, N. A. Langrana, ed., pp. 1819.
47 LeRoux, M. A., Ateshian, G. A., Vail, T. P., and Setton, L. A., 2001, ‘Effects
of Collagen Fiber Anisotropy on the Hydraulic Permeability of the Meniscus,’
Trans. Orthop. Res. Soc., 26,p.45.
48 Hodge, W. A., Carlson, K. L., Fijan, R. S., Burgess, R. G., Riley, P. O., Harris,
W. H., and Mann, R. W., 1989, ‘Contact Pressures From an Instrumented Hip
Endoprosthesis,’’ J. Bone Jt. Surg., Am. Vol., 71A, pp. 13781386.
49 Lee, R. C., Frank, E. H., Grodzinsky, A. J., and Roylance, D. K., 1981, ‘Os-
cillatory Compressional Behavior of Articular Cartilage and Its Associated
Electromechanical Properties,’’ ASME J. Biomech. Eng., 103, pp. 280292.
50 Kim, Y. J., Bonassar, L. J., and Grodzinsky, A. J., 1995, ‘The Role of Carti-
lage Streaming Potential, Fluid Flow and Pressure in the Stimulation of Chon-
drocyte Biosynthesis During Dynamic Compression,’ J. Biomech., 28, pp.
10551066.
51 Buschmann, M. D., Kim, Y. J., Wong, M., Frank, E., Hunziker, E. B., and
Grodzinsky, A. J., 1999, ‘Stimulation of Aggrecan Synthesis in Cartilage Ex-
plants by Cyclic Loading Is Localized to Regions of High Interstitial Fluid
Flow,’’ Arch. Biochem. Biophys., 366, pp. 17.
Journal of Biomechanical Engineering OCTOBER 2001, Vol. 123 Õ 417
... However, for some tissues and in particular contexts, the relative fluid transport in the tissue with respect to the solid matrix is key for understanding the response under compression, dynamic loading, volume changes, and drug transport [23]. For instance in cartilage, multi-phasic theories are needed to explain the performance of this tissue in compression at our joins [24,25]. Bi-phasic theories are also needed for subcutaneous tissue in the context of drug delivery [26,27]. ...
... In accordance with the tension-compression nonlinearity observed in healthy articular cartilage, the repair cartilage in all of the experimental groups exhibited tensile moduli that were appreciably larger than their respective compressive moduli [45] ; however, acellular composites that were implanted with fibrin glue possessed significantly higher tensile moduli than pinned acellular composites and glued precultured composites. Further work would be needed to understand the reason for this. ...
Article
Full-text available
The surgical repair of articular cartilage remains an ongoing challenge in orthopedics. Tissue engineering is a promising approach to treat cartilage defects; however, scaffolds must (i) possess the requisite material properties to support neocartilage formation, (ii) exhibit sufficient mechanical integrity for handling during implantation, and (iii) be reliably fixed within cartilage defects during surgery. In this study, we demonstrate the reinforcement of soft norbornene-modified hyaluronic acid (NorHA) hydrogels via the melt electrowriting (MEW) of polycaprolactone to fabricate composite scaffolds that support encapsulated porcine mesenchymal stromal cell (pMSC, three donors) chondrogenesis and cartilage formation and exhibit mechanical properties suitable for handling during implantation. Thereafter, acellular MEW-NorHA composites or MEW-NorHA composites with encapsulated pMSCs and precultured for 28 days were implanted in full-thickness cartilage defects in porcine knees using either bioresorbable pins or fibrin glue to assess surgical fixation methods. Fixation of composites with either biodegradable pins or fibrin glue ensured implant retention in most cases (80%); however, defects treated with pinned composites exhibited more subchondral bone remodeling and inferior cartilage repair, as evidenced by micro-computed tomography (micro-CT) and safranin O/fast green staining, respectively, when compared to defects treated with glued composites. Interestingly, no differences in repair tissue were observed between acellular and cellularized implants. Additional work is required to assess the full potential of these scaffolds for cartilage repair. However, these results suggest that future approaches for cartilage repair with MEW-reinforced hydrogels should be carefully evaluated with regard to their fixation approach for construct retention and surrounding cartilage tissue damage.
... Further generalization is proposed in poroviscoelastic models, that include flowindependent viscoelasticity in the solid-phase description, and are frequently adopted in the analysis of cartilage in confined and unconfined compression, indentation, pure shear and uniaxial tension [119][120][121]. ...
Article
Full-text available
Articular cartilage is a complex connective tissue with the fundamental functions of load bearing, shock absorption and lubrication in joints. However, traumatic events, aging and degenerative pathologies may affect its structural integrity and function, causing pain and long-term disability. Osteoarthritis represents a health issue, which concerns an increasing number of people worldwide. Moreover, it has been observed that this pathology also affects the mechanical behavior of the articular cartilage. To better understand this correlation, the here proposed review analyzes the physiological aspects that influence cartilage microstructure and biomechanics, with a special focus on the pathological changes caused by osteoarthritis. Particularly, the experimental data on human articular cartilage are presented with reference to different techniques adopted for mechanical testing and the related theoretical mechanical models usually applied to articular cartilage are briefly discussed.
... The apparent viscoelastic response of cartilage is modulated through a combination of fluid flow-dependent (fluid-solid interactions, permeability i.e., poroelasticity) and flow-independent (inherent viscoelasticity of PGs and collagen) mechanisms [102,191,192]. To elucidate the mechanism of viscosity in our measurement range, we simulated FE models with median FRPE material parameter values of normal and severe OA femoral cartilage and extracted phase differences and fluid pressure as a function of frequency (Figure 10.1). ...
Thesis
Full-text available
Osteoarthritis (OA), which is the most common form of degenerative joint disease, causes deterioration in the articular cartilage, resulting in a loss of proteoglycans (PGs) and collagen fibers, and it also impairs the normal arcade-like architecture of the collagen network. The disease ultimately progresses to the advanced stage, where invasive treatments such as joint replacement are the eventual choice. The OA-induced alterations in articular cartilage weaken the normal mechanical function of the tissue, reduce the amount of tissue moduli in equilibrium and instantaneous response and increase tissue permeability. The onset and development of OA and the structure-function relationships of cartilage have been widely studied using animal tissues. Although animal models are crucial in understanding causes for OA in controlled environments, translation of this knowledge to human tissue is lacking. More importantly, the constituent-specific mechanical properties and the structure-function relationships of human femoral and tibial cartilage have not been investigated yet. Therefore, a comprehensive characterization of human knee joint articular cartilage is crucial. To characterize the material properties of each constituent, fibril-reinforced poroelastic (FRPE) or corresponding material modeling is required. The material parameters are obtained through finding the best match between the experimental and the model-generated data. The primary purpose of this thesis was to characterize the elastic, viscoelastic, and constituent-specific FRPE material properties of the human femoral condyle and tibial plateau cartilage during different stages of OA. The secondary aim was to establish structure-function relationships in healthy and osteoarthritic tissue and to compare the mechanical and compositional properties between different stages of OA as well as different cartilage sites. To achieve these goals, the cartilage samples were tested mechanically in the indentation device and the FRPE material parameters were obtained from the model fit. Using microscopy and infrared spectroscopy, the compositional and structural parameters namely PG content, collagen content, collagen fiber orientation angle and parallelism index were obtained. The samples were also histopathologically graded using the OARSI grading system and pooled to normal (OARSI 0-1), moderate OA (OARSI 2-3), and severe OA (OARSI ≥ 4). The results of this thesis indicate that OA progression in tibial cartilage is associated with the weakening of the mechanical function of fibrillar and non-fibrillar matrices in both moderate and severe OA. However, the femoral condyle cartilage experienced no essential deterioration in moderate OA. In addition, the constituent-specific material parameters related to the collagen network and proteoglycan matrix were greater in femoral condyle cartilage compared to those of tibial cartilage. This finding highlights the adaptation of cartilage to specific physiological functions, loading regimes, and the site-specific inherent differences. In the moderate OA stage tibial cartilage samples, we found that the loss of collagen fibrils pretension (represented by the initial fibril network modulus) was related to the loss of PGs (loss of tissue swelling pressure). At the severe OA stage, the loss of PGs together with the disorganization of the collagen network were related to the loss of collagen fibers pretension. The higher tissue viscosity (represented by the phase difference) characterized at low loading frequency was associated with a loss of PG content (for tibial and femoral cartilage) and collagen content (for tibial cartilage only). At the high loading frequency, the higher tissue viscosity was associated with a disorganization of collagen fibers (higher collagen orientation angle). Therefore, the results highlight that proteoglycan degradation and collagen disorganization increase the viscosity of cartilage, but their contribution to increased viscosity occurs at completely different loading frequencies. Moreover, collagen content seems to modulate tissue viscosity at low frequency only in tibial cartilage. Put together, the knowledge established from native human cartilage in this thesis helps us to better understand OA. The results also provide a reference for tissue engineering efforts and improve the computational biomechanical models of the knee joint by providing an experimentally derived cartilage material representation.
Article
Full-text available
The utilization of materials in medical implants, serving as substitutes for non-functional biological structures, supporting damaged tissues, or reinforcing active organs, holds significant importance in modern healthcare, positively impacting the quality of life for millions of individuals worldwide. However, certain implants may only be required temporarily to aid in the healing process of diseased or injured tissues and tissue expansion. Biodegradable metals, including zinc (Zn), magnesium (Mg), iron, and others, present a new paradigm in the realm of implant materials. Ongoing research focuses on developing optimized materials that meet medical standards, encompassing controllable corrosion rates, sustained mechanical stability, and favorable biocompatibility. Achieving these objectives involves refining alloy compositions and tailoring processing techniques to carefully control microstructures and mechanical properties. Among the materials under investigation, Mg- and Zn-based biodegradable materials and their alloys demonstrate the ability to provide necessary support during tissue regeneration while gradually degrading over time. Furthermore, as essential elements in the human body, Mg and Zn offer additional benefits, including promoting wound healing, facilitating cell growth, and participating in gene generation while interacting with various vital biological functions. This review provides an overview of the physiological function and significance for human health of Mg and Zn and their usage as implants in tissue regeneration using tissue scaffolds. The scaffold qualities, such as biodegradation, mechanical characteristics, and biocompatibility, are also discussed.
Article
Full-text available
Tissue loss through injury, surgery, and disease motivates the development of new biomaterials to enable tissue repair and regeneration. Emerging at the interface between bioadhesives and regenerative medicine, a new generation of regenerative bioadhesives is created to possess dual functions of seamless tissue adhesion and effective tissue repair. This bioadhesive innovation has wide clinical applications, ranging from wound management to the regeneration of musculoskeletal tissues such as tendons and intervertebral discs. This perspective covers the design principles of regenerative bioadhesives in considering both mechanical and biological elements. Case studies of regenerative bioadhesives for load‐bearing organs such as skin, tendon, and intervertebral discs are presented here. Finally, immediate opportunities and future perspectives are outlined to further advance the field of regenerative bioadhesives.
Article
Full-text available
Purpose: The sclera is believed to biomechanically influence eye size, facilitating the excessive axial elongation that occurs during myopigenesis. Here, we test the hypothesis that the sclera will be remodeled and exhibit altered biomechanics in the mouse model of form-deprivation (FD) myopia, accompanied by altered retinoid concentrations, a potential signaling molecule involved in the process. Methods: Male C57 Bl/6J mice were subjected to unilateral FD (n = 44 eyes), leaving the contralateral eye untreated (contra; n = 44). Refractive error and ocular biometry were measured in vivo prior to and after 1 or 3 weeks of FD. Ex vivo measurements were made of scleral biomechanical properties (unconfined compression: n = 24), scleral sulfated glycosaminoglycan (sGAG) content (dimethylmethylene blue: n = 18, and immunohistochemistry: n = 22), and ocular all-trans retinoic acid (atRA) concentrations (retina and RPE + choroid + sclera, n = 24). Age-matched naïve controls were included for some outcomes (n = 32 eyes). Results: Significant myopia developed after 1 (-2.4 ± 1.1 diopters [D], P < 0.001) and 3 weeks of FD (-4.1 ± 0.7 D, P = 0.025; mean ± standard deviation). Scleral tensile stiffness and permeability were significantly altered during myopigenesis (stiffness = -31.4 ± 12.7%, P < 0.001, and permeability = 224.4 ± 205.5%, P < 0.001). Total scleral sGAG content was not measurably altered; however, immunohistochemistry indicated a sustained decrease in chondroitin-4-sulfate and a slower decline in dermatan sulfate. The atRA increased in the retinas of eyes form-deprived for 1 week. Conclusions: We report that biomechanics and GAG content of the mouse sclera are altered during myopigenesis. All scleral outcomes generally follow the trends found in other species and support a retina-to-sclera signaling cascade underlying mouse myopigenesis.
Article
Biphasic poro-viscoelastic constitutive material model (BPVE) captures both the fluid flow dependent and independent behavior of cartilage under stress relaxation type indentation. A finite element model based on BPVE formulation was developed to explore the sensitivity of the model to Young's modulus, Poisson's ratio, permeability and viscoelastic constitutive parameters expressed in terms of Prony series coefficients. Then we fit the numerical model to experimental force versus time curves from stress relaxation indents on bovine tibial plateaus to extract the material properties. Measurements were made over the period of two days to capture the material property changes that resulted from trypsin-induced degradation. We measured spatial and temporal changes in mechanical properties in the cartilage. The areas of degradation were characterized by an increase in both permeability and summation of Prony series shear relaxation amplitude constants. These findings suggest that cartilage degradation reduces the intrinsic viscoelastic properties of the tissue's solid phase in addition to impairing its capacity to offer frictional drag to the interstitial fluid flow (permeability). The changes in material properties are measurable well before structural degradation occurs.
Chapter
Blood vessels belong to the class of soft tissues discussed in Chapter 7. They do not obey Hooke’s law. Figure 7.5:1 in Chapter 7, Sec. 7.5 demonstrates the nonlinearity of the stress-strain relationship and the existence of hysteresis. They also creep under constant stress and relax under constant strain. These mechanical properties must have a structural basis. In Sec. 8.2 we shall consider the structure of the blood vessel wall and its correlation with the mechanical properties. From Sec. 8.3 on, however, our attention will be concentrated on the mathematical description of the mechanical properties. In Secs. 8.3–8.5 we formulate a quasi-linear viscoelastic theory for blood vessels, using the pseudo-elasticity concept introduced in Chapter 7. In Sec. 8.6 we discuss the use of arterial pulse waves as a means to determine the mechanical properties of arteries. In Secs. 8.7–8.9 we consider the mechanical properties of arterioles, capillary blood vessels, venules, and veins. Finally, in Sec. 8.10, we discuss the long-term response of blood vessels to stresses: their reaction to hypertension, growth, regeneration, and resorption.
Article
Hydrated biological soft tissue consists of a porous extracellular matrix (ECM) and an interstitial fluid. The poroelastic theory (Biot, 1962), which was originally developed for soil mechanics, has been widely used for mathematical modeling of such hydrated biological tissue. This theory assumes that the ECM is incompressible and purely elastic, and that the interstitial fluid is incompressible and inviscid. The overall viscoelasticity of the tissue is expressed as a result of the frictional interaction between the elastic porous matrix and the interstitial fluid. The poroelastic theory, also known as the biphasic theory (Mow et al., 1980) in the biomechanics field, has served well over the past 20 years as an excellent modeling tool for the interstitial fluid flow-dependent viscoelastic response of hydrated soft tissue. It has been demonstrated that hydrated soft tissue also possesses a significant intrinsic viscoelasticity, independent of the interstitial fluid flow. The biphasic poroviscoelastic (BPVE) theory, which was first introduced by Mak (1986a and 1986b), incorporates a viscoelastic relaxation function into the effective solid stress of the poroelastic theory thus accounting for both intrinsic fluid flow-independent and fluid flow-dependent viscoelasticity. The objective of the present study is to investigate the biphasic poroviscoelastic characteristics of hydrated soft tissue, with an emphasis on the relative contribution of fluid flow-independent viscoelasticity to the overall viscoelastic behavior of soft tissue.
Article
A combined experimental and analytical approach was used to determine the history-dependent viscoelastic properties of normal articular cartilage in tension. Specimens along the surface split line direction, taken from the middle zone of articular cartilage were subjected to relaxation and cyclic tests. A quasi-linear viscoelastic theory proposed by Fung [31] was used in combination with the experimental results to determine the nonlinear viscoelastic properties and the elastic stress-strain relationship of normal articular cartilage.
Article
In this study, the nonlinearity of the tensile and compressive stiffness of glenohumeral joint (GHJ) cartilage and its variation with depth, direction and joint surface were examined. One hundred and twenty tensile specimens and 70 compressive plugs were prepared from five fresh-frozen human GHJs. Overall, the results demonstrate the complexity of the mechanical properties of human GHJ articular cartilage.