You are on page 1of 18

SURVEY ARTICLE

-.

FLUID TRANSPORT AND MECHANICAL PROPERTIES OF


ARTICULAR CARTILAGE: A REVIEW

VAN C. Mow,* MARK H. HOLMES* and W. MICH.AEL LN*


Rensselaer Polytechnic Institute. Troy. N1’ 12181. C.S.A.

.Abstract--This review is aimed at unifying our understanding of cartilage liscoelastic propertics in


compression. in particular the role of compression-dependent permeability in controllinr interstitial fluid
Ho\\ and its contribution to the observed viscoelastic etfects. During the pre\ious decade. Tt was shown that
compression causes the Frmeability ofcartilage to drop in a functional manner described by i, = !q, cxpic.~f~
u here X, and .\I were defined as intrinsic permeability parameters and E is the dilatation of the solid matrix (F
= crVu).Since permeability is inversely related to the diffusive drag coefftcient of relattve Iluid motion utth
respect to the porous solid matrix, the measured load-deformation response of the tissue must thereforealso
depend on the non-linearly permeable nature of the tissue. We have summarized in this review our
understanding of this non-linear phenomenon. This understanding of these flow-dependent viscoelastic
etTectsare put rnto the historical perspectiveofacomprehensive litcraturc rcvica ofcarlierattcmpts to model
the compressive viscoelastic properties of articular cartilage.

ISTRODKCTION mechanics, and advanced mathematics. We intend in


this article to review some of the experimental and
Diarthrodial joints, such as the knee and shoulder, are theoretical research projects we have pursued over the
one of the three types of joints in the human body; the past decade and to discuss some of the unresolved
others are fibrous joints or synarthroses, which have questions in the study of the structure and function of
no relative motion (e.g. the coronal suture of the skull), just one of the major connective tissues of the joint:
and cartilaginous joints or amphiarthroses, which have articular cartilage. In particular. we intend to focus on
little or no relative motion (e.g. the vertebral body- the role of the non-linear permeability function in
intervertebral disk-vertebral body joints). A diart- controlling the rate of interstitial fluid flow. which in
hrodial joint is characterized by its large degree of turn governs the biphasic compressive viscoelastic
motion; consequently. one of its primary functions is properties of the tissue. Reviews of work on synovial
to facilitate body movement and locomotion (see for fluid can be found in Lai cr al. (1978) and some of the
example Warwick and Williams, 1973; Sokoloff, 1978, dynamic interaction problems that may arise within
1980). Under normal conditions, it is an amazingly diarthrodial joints between cartilage and synovial fluid
efficient bearing system, capable of providing a nearly can be found in Mow and Lai (1979, 1980).
frictionless performance with little, if any, wear for the Articular cartilage is a white, dense, connective
entire lifespan of the individual. Although their indi- tissue, from 1 to 5 mm thick, that covers the bony
vidual anatomical forms and material properties vary articulating ends inside the joint. In biomechanical
considerably. there are two components common to all terms, it is a multi-phasic, non-linearly permeable,
diarthrodial joints: synovial fluid and connective viscoelastic material, consisting of two principal
tissue. In the latter category are articular cartilage, phases: a solid organic matrix, which is composed
meniscus, ligaments and tendons. The biomechanical predominately of collagen fibrils and proteoglycan
characterizations of these soft connective tissues (for macromolecules, and a movable interstitial fluid phase,
example, Hayes and Bodine, 1978; Kempson ct al., which is predominately water (Linn and Sokoloff,
1971; Kempson, 1980; Wooer al., 1979,198l; MOWet 1965; Venn and Maroudas, 1977; Muir. 1980;
al., 1980; Roth and Mow, 1980; Lanir, 1983) and Armstrong and Mow, 1982a.b). The importance of the
attempts to develop dynamic interaction models integrity of the tissue in maintaining the health of the
(Higginson et al., 1976; Higginson, 1978; Dowson et joint and of the deformational characteristics of the
al.. 1981; Mow and Lai, 1979, 1980; Collins, 1982) are tissue as a whole has been appreciated for over at least
currently active areas of interdisciplinary research, half a century (Hirsch, 1944; Sokoloff, 1969; Mankin.
incorporating concepts from biochemistry, continuum 1974; Muir, 1977; Schofield and Weightman, 1978;
Moskowitz rt al., 1979). However, only recently has the
Re~riwcl IS MOJ 1983: in rehwf jorm 7 fkwmhrr 1983.
importance of the role of the movement of interstitial
*Department of Mechanical Engineering. Aeronautical
Engineering and Mechanics. fluid through the porous matrix and the subsequent
+Depnrtment oi Mathematical Sciences. interaction of the interstitial fluid with the porous solid
377
378 VAN C. Mow, MARK H. HOLMES and W. MICHAEL LAI

matrix in controlling the deformational characteristics


of the tissue as a whole been established (Maroudas,
1975a.b; IMOWrt ul., 1980; Holmes et n/., in press).
Principally, the fluid movement plays a fundamental
role in: (1) the biological processes by augmenting the
transport of nutrients into, and of waste products out
of, the tissue (Ekholm and Norback. 1951; Maroudas
er al., 1968; McKibbin and Maroudas, 1979; Honner
and Thompson, 1971; Salter et al., 1980); (2) the
deformational processes by controlling the mechan- Inters*i:tY 1’ ‘Proteogiycan
ism, through a non-linear interaction process, of the Water - Asregates

rate of fluid transport through the deforming tissue


Fig. 1. Schematic representation of articularcartilage depict-
(Maroudas, 1975b; Mansour and Mow, 1976; Lai and ing a porous, composite organic solid matrix swollen with
Mow, 1980; Lai et al., 1981; Mow YI al., 1982); and water. The major organic components of the matrix are
(3) the functional processes by providing the lubricant collagen fibrils and proteoglycrtn aggregates (Mow and Lai,
for the conformational gap between the articulating 19801.
surfaces of the joint, through exudation and imbibition
caused by the deformation of the tissue during joint
articulation (McCutchen, 1962, 1978; Malcom. 1976;
Torzilli and Mow, 1976a,b; Armstrong and Mow,
1980; Dowson et al., 1981; Collins. 1982; Mow and
Mak, in press). The portion of cartilage mechanics that
we will review here will be those derived from com- 30 I- .:~:0,,5+022* -I
35

pressive theory describing the behavior of articular


cartilage.
Recent advances from these investigations have
codified the various roles of interstitial fluid movement
on the various compressive responses of the tissue.
These fundamental and simple concepts may now be
summarized in a coherent manner to promote an
understanding of the role of interstitial fluid flow
through articular cartilage, or any other biphasic tissue
(meniscus, nasal cartilage, ligaments, etc.), under a
more complex deformational field. Thus the objective
of this review is to provide a comprehensive elucid-
ation of the role of interstitial fluid flow in controlling Fig. 2. Variation of porosity C;,’YTand solid content V,/ V,
the compressive deformational characteristics of carti- with depth from the surface. Note that the top 204; of the
laginous tissues. tissue has the most fluid (Lipshitz er at.. 1976).

COMPOSITION AND ULTRASTRUCTURE OF ARTICULAR


CARTILAGE Pathological conditions can also have a significant
effect on the water content ofcartilage. For example, in
As a biomaterial, articular cartilage may be consid- osteoarthritic human cartilage, the water content is
ered as a porous composite organic solid matrix above normal (Ballet and Nance, 1966; Maroudas,
swollen by water, Fig. 1. In the water, a variety of 1976,1979; McDevitt and Muir, 1976; Armstrong and
mobile electrolytes, which maintain the charge neut- Mow, 1982a.b)and the water concentration is higher in
rality condition of the ionized proteoglycan aggregates the central region of the tissue rather than in the
of this matrix, are distributed (Sokoloff, 1963; superficial zone (Venn and Maroudas, 1977; Venn.
Schubert and Hamerman, 1968; Maroudas, 1970). The 1978). This increase of water content has been found to
interstitial concentrations of the counterions and ions be the earliest observable change in tissue in animal
are governed by the Donnan equilibrium concept models of osteoarthritis (McDevitt and Muir, 1976)
(Maroudas, 1970, Urban et al., 1979). The specific and this increase has been correlated to the observed
composition of cartilage varies with joint age and the decrease of the compressive equilibrium modulus in
topographical and depth location in the joint, as well as animal models as well as in human autopsy materials
with the specific type ofjoint (Venn, 1978; Muir, 1980). from aging joints (Armstrong and Mow, 1982b).
In normal adult cartilage, the fluid content makes up The largest component of the organic solid matrix is
approximately 85 o/0of the total mass by wet weight in collagen (Muir, 1980). Cartilage collagen, which ac-
the most superficial 25% of the cartilage, and then counts for about half of the tissue mass by dry weight,
decreases in a nearly linear manner to about 70 7: at the occurs in a highly specific ultrastructural arrangement,
subchondral bone (Lipshitz et al., 1976a), Fig. 2. forming four principal layers (McCall, 1968; Clarke,
Properties 0i articular cartilage: a rsbleu 3-9

minoglycans are relatively short chains of repeatmg


disaccharide units of sulphated hexosamines. In car-
tilage. proteoglycans are bound to a linear chain of
hyaluronic acid to form an enormous aggregate with a
molecular weight of up to 2 x 10R and a length of
approximately 2 pm (Hardingham and Muir, 1974:
Rosenberg t’l (II.. 1975). It appears that most of the
proteoglycans in normal cartilage are in these aggre-
gated forms. However, in aging and diseased tissues.
the molecular architecture of this aggregate is altered
in a number of ways (Brandt and Palmoski. 1976:
Fig. 3. Representation of collagen tibril ultrastructure
Muir, 1977).
throughout the depth of the cartilage depicting the distinct
idwlized zones: (a) a superticial tangential zone composed of
Jen\i‘j packed collagen ribrils in sheets parallel to the surface. PHYSICOCHE~IICAL BASIS OF FLUID FLOQ ASD
tb)n randomly arrayed middle zone with abundant inter- C.~RTIL.\GE DEFORSIATIOS
tibrlllar space to accommodats the proteoglycan and water.
and ICI~ radial deep zone where collagen tibrils appear to
In normal adult cartilagc.collagen is woven together
snactomose into larger bundles as they insert into thecalcified
zoner. The cartilage. i.e. the non-calcified and calcified zones, to form a fibrous network in which the huge pro-
rests upon a stitl’ subchondral plats supported by the soft teoglycan aggregates are trapped. Together they form
wnceilous-marrow composite structure (Mow er [I/., 1974). a cohesive porous composite organic solid matrix. The
collagen network is characterized by its great tensile
197-l: Mcachim and Stockwell, 1979). The superficial stiffness (Kempson et (II., 1973; Woo er rll.. 1976, 1979.
tangential zone, which is near the articular surface. 1980: Roth and Mow, l980), but due to the high
consists of sheets of tightly woven collagen fibrils slenderness ratio of each segment of the collagen fibril.
(Mow ef ~1.. 1974; Clarke, 1971, 1974; Ghadially et al., the network is relatively weak in compression. It also
19761. This region accounts for the highest concentra- has a negligible net charge density at neutral pH
tion of collagen. Based on recent X-ray diffraction (Bowes and Kenten. 1948). On the other hand, the
studies (Aspden and Hukins, 1951) and polarized light proteoglycans contain a high concentration of fixed.
and electron microscopy studies (Bullough and negatively charged groups, giving them a propensity to
Goodfellow. 1968; Sprerand Dahners, 1979), the fibers expand and occupy a large solution domain. Since
on the surface seem to be aligned to the split-line these macromolecules are enmeshed in the collagen
patterns.* The tibers of the middle zone, on the other matrix, which prevents them from reaching their fully
hand. are randomly oriented and homogeneously extended solution volume, their closely spaced charged
dispersed. In the deep zone the fibers come together to groups, 5-15 A, together with the high concentration
form larger. radially oriented fiber bundles. These of freely mobile counterions in the interstitial fluid
bundles enter the calcified zone. crossing the tidemark. (Donnan osmotic pressure), endow the tissue with a
to form an interlocking network that anchors the high capacity to swell and gain or lose water when the
tissue to the bony substrate, Fig. 3. external ionic or mechanical environment is altered
The major noncollagenous components of the solid (Schubert and Hamerman, 1968; Pasternack rf ul..
phase of articular cartilage are proteoglycan macromo- 1974; Maroudas, 1976.1979; Hascall, 1977; \low et al..
leculss, which in normal adult cartilage make up 1981; Grodzinsky YItrl., I98 I; Myers et (II.. in press a.b).
approximately 2&30”, of the total dry weight of the In human femoral heads, it has been estimated that the
tissue. However, unlike the collagen and water content, internal osmotic pressure can range up to 0.2 \IN m _ ’
the percentage of proteoglycans is lowest near the (Maroudas, 1979). At equilibrium, this swelling force is
articular surface and increases with depth. These balanced by the tensile strength of the surrounding
macromolecules consist of a protein core on which collagen matrix. Consequently, even in an unloaded
50-100 glycosaminoglycan chains are bonded (Muir, state, the in sifu collagen network of articular cartilage
1950: Buckwalter and Rosenberg, 1982). The glycosa- is under a tensile prestress (Maroudas, 1976).
The hydrophilic nature of the proteoglycans and the
*The oreanization of the collagen fiber network of ar- architecture of the collagen network give the tissue its
titular cartilnee has been of interest to scientists for a long micro-porous characteristic. Very little of the water in
time. Early rn-icroanatomists believed that the extracellular cartilage is intracellular, and a sizable amount of this
substance of the tissue was homogeneous. Hultkrantz (1898)
water, about 30 00, appears to be associated with the
demonstrated the regular split-line patterns on the articular
buriaces by puncturing them with an awl. The appearance of
collagen fibrils (Torzilli rl (II., 1992). Therefore. most of
this regular array of’split lines’on the surfaceofcartilage was the water is in the intermolecular space and is free to
taken as ebidrnce of the surface collagen fiber ultrastructure. move through the tissue. The ‘pores’ through which it
Tncnty-seven years later. BenninghotT(l925)used these split Rows have been estimated to have a diameter of
lincsand pol,irired light techniques to investigate thearrange-
- 60A (McCutchen, 1962); the derivation of this
ment of the collagen fibers and the whole tissue. From this he
postulated the well-known ‘arcade’ theory of collagen fiber estimate will be concisely described below. The hy-
architecture. draulic permeability of the tissue, as well as the
380 VANC. Mow. MARK H. HOLMES and W. MICHAEL LM

aggegate equilibrium compression modulus, are

.
highly dependent on both the water content and the
uranic acid content, hence providing the physicoche-
mica1 basis for the observed decrease of the tissue
permeability with increased compression (Maroudas,
1975a.b; Mansour and Mow, 1976; Armstrong and
Mow. 1982a), Figs 4a-d.
During compression, as the mmhanical strain in-
creases, the concentration of the organic material and
the charge density both increase, since the interstitial
fluid is forced to flow from the matrix. A new
equilibrium state is reached when the charge density,
collagen tension and applied load are in balance. It is
generally believed that this equilibrium compressive

O70 75 80 95 90 95 100

WATER CGNTENT (Z’

Fig. 4(c). Variation of apparent permeability in m’:N’s


units as a function of native uater content of aging human
patellar specimens (Armstrong and Liow, 1982a).

0 01 02 OS
Fixed-charge density (m equiv./cm5)

Fig. 4(a). Variation of permeability in cm’sg-’ units as a


function of tixcd-charge density in m equiv. cm-” of the tissue o,
(Maroudas. 197Sb).

0
.o 75 80 @5 90 35 100
‘NAT’R CiVEFIT (7.)

Fig. 4(d). Variation of intrinsicequilibrium modulus in MPa


units as a function of native water content of aging human
patellar specimens (Armstrong and Mow, 1982a).

stiffness of the tissue is derived from the Donnan


osmotic effect associated with the closely spaced,
negatively charged sulphate and carboxyl groups on
the glycosaminoglycans. as well as from the resistance
to bulk compression of the neutral composite
proteoglycan-collagen solid matrix (Maroudas,
1975b; Parsons and Black, 1979; Armstrong and Mow,
1982b; Myers and Mow, 1953). The ‘flow-independent
shear rigidity’ of cartilage is derived from the collagen
fibriis that apparently carry the tensile stress rendered
3 90 80 m 60 55 by the spatial nature of the collagen ultrastructure
Cortiloge hydration expressed within the tissue when it is sheared, Fig. 5 (Hayes and
OS % of initiol weight Bodine, 1978; Mow et al., 1982, 1983). Because the
Fig. 4(b). Variation of permeability in cm3 sg-’ units as a
proteoglycan aggregates alone do not provide signili-
function of hydration as percentage of initial weight of the cant shear resistance, they must interact with the
tissue (Maroudas, 1975b). collagen network to produce a solid matrix with
Properties of articular cartilage: a review 381

shrinkage in tissue size and a decrease m tis,ue mass


caused by the decrease in tissue hydration, %loreover.
recent experimental and theoretical evidence shows
that the modulation of intra- and inter-molecular
electrostatic repulsion forces among the proteoglycan
charge groups (Mow and Lai, 1980), and or disruption
of collagen-proteoglycan electrostatic interactions
(Muir, 1980) by changes in the counterion environ-
ment in the interstitial fluid. can atfcct compressive
(Sokoloff, 1963; Maroudas. 1979; Parsons and Black.
1979) tensile (Kempson. 1980; Myers t’r LJ.. in prrsh
Fig. 5. Schematic depiction of functional role of collagen a-b), and shear (Hayes and Bodine, 1978; Mow t’t tr!.,
tibrils in resisting the shear deformation of a unit block of
1983) properties of the tissue. It is clear from these
articularcartilage. The existence ofa composite solid matrix is
nccessar); for the collagen fibrils to play an active role during studies that both the movement of interstitial fluid and
shear. the rates of ion transport in cartilage arc of predomi-
nant importance in the functioning of the tissue.
sufhcient shear strength (Armstrong et al., 1981; Mow
er LI/.,in press 1983). Similarly, the tensile stitfness of
cartilage is derived primarily from the intrinsic col- FLL’ID FLOH: SLTRITIO\
lagen fibril properties. the strength of collagen cross-
linking. and, secondarily, from collagen-proteoglycan The importance ofinterstitial fluid flow. particularly
interactions. These equilibrium behaviors are ‘tlow- as it affects the nutrition of cartilage, was appreciated
independent’. When interstitial fluid flow becomes as early as the nineteenth century. Later the role of
significant, the frictional drag or diffusive resistance of fluid flow was more popularly expounded by
relative fluid flow causes the observed flow-dependent Benninghoff (1924.1925)and by Ekholm and Norback
viscoelastic effects, i.e. creep and stress relaxation. In (1951). However. even though there was a great deal of
compression, it has been shown that this biphasiceffect speculation that the deformation of cartilage was
is the dominant mechanism responsible for the ob- strongly influenced by the exudation and imbibition of
served compressive viscoelastic behavior of cartilagin- interstitial fluid, its exact role was not understood in
ous tissues (Mow er al., 1980, Lai et ~1..198 I; Holmes rr the early studies on the elasticity of cartilage (Bsr,
LJ., in press). 1926; Giicke, 1927: Hirsch, 1944). For example. Hirsch
From the discussion of these rather simple notions, conjectured that the ‘circulation of tissue juices’ de-
it is apparent that the general deformation of articular creased as the cartilage lost its elasticity. thereby
cartilage is very complex, involving the mechanical reducing the mechanism for its nutrition. At the same
entanglement of collagen and proteoglycan, the tensile time, he could not explain why he obtained a creep-like
resistance of the collagen network, the compressive response and an incomplete recovery in his indentation
resistance of the proteoglycans, and the diffuse resist- tests of cartilage, giving rise to a phenomenon which
ance generated as the fluid flows through the matrix. became known as ‘imperfect elasticity’. He apparently
The details of the internal equilibration mechanisms did not realize the importance of the exuded fluid that
between the collagen network, proteoglycan swelling he observed on the articular surface around the
pressure, and fluid flow are presently largely un- indenting probe. Nevertheless, Hirsch hypothesized
resolved, although some progress has been made in the that there should be a strong connection between
last few years (Maroudas, 1979; Mow et al., 1980; mechanical strain and tluid Row, being aware of the
Grodzinsky er ol., 1981). In this review we intend to ideas of Hildebrand (1921) and Policard (1936) that
present some of the findings concerning the role of fluid movement can be induced from a ‘suction and
interstitial fluid flow in cartilage deformation and, in pressure effect’ of repeated loading of the cartilage
particular, the non-linear interaction between cartilage surface.
deformation and interstitial fluid flow as it is man- The role of a pumping mechanism for the transport
ifested in the non-linear, intrinsic permeability of nutrients through cartilage has been questioned by
function. Maroudas et a/. (1968) and by McKibbin and
The physicochemical characteristics of the counter- Maroudas (1979). During their intermittent compres-
ion within the interstitial fluid, e.g. valency and concen- sion experiments, they found that the penetration of
tration, also have a significant effect on the response of methylene blue into cartilage is no more than they
articular cartilage to imposed stresses and strains. For believed would be expected from simple diffusion with
example, an increase of interstitial Na’ concentration no intermittent compression. Consequently. they con-
causes a charge-shielding effect on the closely spaced cluded that for small solutes, such as glucose. diffusion
( - 5-I 5 A) anionic groups on the glycosaminoglycans. is the controlling mechanism, whereas a mechanical
This, in turn, can cause the proteoglycans to contract pumping action probably governs the transport of
from their extended solution volume (Pasternack et al., solutes of a larger molecular weight. such as serum
1974). Macroscopically, this results in a measurable albumin. One of the arguments used in support of a
382 VAN C. Mow. MARK H. HOLMES and W. MICHAEL LAr

pumping mechanism is that when the joint is in- incompressible, so v = 1,‘2and E = 311.Inserting these
capacitated, so there is no rhythmic stressing, the values into equation (2). one obtains equation (1) used
cartilage shows signs of degeneration (Honner and by Sokoloff for E. From this. for human patella.
Thompson. 1971: Salter et al., 1980; Brandt. 1981). Sokoloff found the ‘instantaneous Young’s modulus’
However. there is also the possibility that the synovial (which uses w,, at 0.S s after application of the load) to
fluid ‘stagnates’ at the articular surfaces in such a be 2.28 MPa. Also. the’equilibrium Young’s modulus’,
situation, thereby reducing the nutrients available for which uses we at one hour, was found to be 0.69 MPa.
transport by diffusion. The experimental evidence to Moreover, he observed that the deformation of ar-
date supports the fact that low molecular weight ticular cartilage was non-linear, even for strains of
solutes are transported by simple diffusion, but there IO 9,. which is contrary to the basic assumptions
has not yet been a complete coherent theoretical employed in deriving the expressions for E and /I, i.e.
analysis to assess the relative importance of the equations (1) and (2).
mechanical pumping effect vs the ditfusive transport The analysis for a plane-ended, as well as for a
mechanism for cartilage (Lai and Mow, 1978). spherical-ended, indenter on a layer of linearly elastic
material attached to a rigid foundation was carried out
by Hayes et al. (1972) and by Hori and Mockros (1976).
CONTISL’USl ~IODELISC OF ARTICULAR CARTILAGE Using the approach of Lebedev and Ufliand (1958) of
reducing the problem to the inversion of a Fredholm
Single phase: elastic
integral equation of the second kind. Hayes it al. were
In studying the deformational characteristics of able to determine the displacement field of the elastic
articular cartilage under mechanical loading, one of layer at equilibrium. For the case of a plane-ended
the central concerns has been the determination of the indenter, the Young’s modulus was found to be given
‘elastic modulus’ of this thin layer of tissue at the ends as
of the bone in a diarthrodial joint. Because of its
anatomical form and its thinness, the indentation E _ P(l -G)
experiment was the choice of many investigators 2w,ai(a, h, v) ’
(Hirsch, 1944; Sokoloff, 1966; Kempson et al., 1971; where the function K comes from the solution of the
Colctti et al., 1972; Hayes et al., 1972; Hori and
integral equation. From their analysis. Hayes et al.
Mockros, 1976). The first attempts to find the Young’s
concluded that the depth of indentation is very sensit-
modulus usually involved the application of the Hertz
ive to the aspect ratio, a/h, since the radius of the
solution for the contact between two elastic bodies of
indenter is less than the thickness in most indentation
infinite depth. Although the justification for this is tests, Fig. 6. To illustrate, with the assumption that
usually vague, Hirsch (1944), who was aware of the
cartilage is incompressible, it can be shown that for
possible influence of the multiphasic nature of car- cartilage on a rigid foundation, the instantaneous
tilage on its deformational characteristics, justified his Young’s modulus was 30 :,, less than that calculated by
analysis on Gildemeister’s (1914) work on the contact Sokoloff. This shows that in Sokoloff‘s calculations,
between gels. the stitfness of the underlying bone was lumped into
Another approach was taken by Sokoloff (1966), the cartilage stiffness, giving an erroneously high
who considered cartilage to be comparable, in terms of Young’s modulus value (Mow er al., 19S2).
its mechanical response, to medium-hard rubber. For
this latter material, the ‘mean instantaneous deforma- Singlr phase: rkcoelasric
tion’ is apparently not very sensitive to the thickness if The assumption that cartilage is purely elastic
it is thick enough, in this case more than about 2 mm applies, at best, only at equilibrium because there
(Livingston rr a[., 1961). Since the cartilage specimens
used by Sokoloff were approximately 3 mm thick, he
assumed that the Young’s modulus E for cartilage
could be determined from the relation
P
E (1)
=2.67w,a1

where P is the constant applied load, we is the depth of


penetration, and a is the radius of the plane-ended,
Subchondral
cylindrical indenter. This comes from the solution of
Bone
the elastic punch problem for a linearly elastic medium
of infinite depth at equilibrium and with a shear Fig. 6. Representation ofa plane-ended indenter applied to a
modulus of thin layer ofarticular cartilage. Cartilage has been modeled as
a single-phase elastic medium (Sokoloff. 1966). a layered,
P(1 -v) single-phase elastic medium (Hayes et (II.. 1972: Kempson et
/L=p. (2)
4W, a NI., 1979). a single-phase viscoelastic medium (Coletti er ul..
1972; Parsons and Black, 1977). and a layered, biphasic
In his analysis, Sokoloff further assumed cartilage is medium (Mow rr al.. 1991t.
Properties of articular cartilage: a review x3

would bc no dissipative et%ct due to the movement of however. inconsistent with the equilibrium condition
the interstitial fluid. However, it was not until the early used to derive equation (4). and it clearly violates the
1960’s that the systematic study of the role of fluid flow assumption that the material is purely elastic. Further,
on the function of cartilage actually began. Elmore e’t the expression is quite sensitive to the Poisson’s ratio.
al. ( 1963). in one of the Hurststudies, showed that the see equation (3). and had to be guessed at in order to
creep response observed in the indentation testing of calculate E from equation (4). It is not surprising that
cartilage is largely due to the eRlux of interstitial fluid Simon (1971). from his indentation tests on canine
from the tissue. Their obsenation resolved the prob- tibia1 cartilage, obtained inconsistent results using the
lem of the ‘imperfect’ elasticity of cartilage that had ‘two-second creep modulus’ concept. A similar diff-
been observed 3Oyr earlier by Hirsch. They also culty was encountered by Hori and Mockros (1976) in
observed that after removal of the load, complete using equation (3); they found there was considerable
recovery occurs only if the fluid is present to enter the dispersion in their values for E. which they suggested
cartilage. This study was extended by Linn and was due to the non-linear, anisotropic and inhomoge-
Sokoloff (1965), who correlated the magnitude of the neous nature of articular cartilage. This type of
creep response and the amount of fluid exuded from analysis has been extended by Parsons and Black
the tissue. and by Sokoloff (1966). who studied the (1977) to include linear viscoelasticity, using a genera-
topographical variations of these properties. This lized Kelvin model to describe the transient portion of
latter problem has also been examined by Hirsch the creep curve under a constant indenting load. The
(1944). Kempson er ol. (1971). and Cameron t’r 01. concepts of ‘unrelaxed and relaxed modulus’ were
(1975). While most investigators had by then realized introduced to determine the instantaneous and
that interstitial fluid flow is intrinsically linked to equilibrium values of the modulus using equation (3)
cartilage deformation, no theoretical attempts were where the Poisson’s ratio was assumed to be 0.4. With
made to consider these effects in a continuum model. appropriate reinterpretation (Mak PI al., 1987; Mow YL
Most investigators, except Torzilli and Mow (1976a.b) al.. 1982). these unrelaxed and relaxed moduli can be
continued to use single-phase viscoelastic models to used to estimate the intrinsic elastic moduli of the solid
describe the compressive creep response of cartilage. organic matrix of the biphasic model for cartilage of
In the indentation tests of cartilage, after a sudden Mow et al. (1980) without an a priori assumption on
application of static load, a rapid compression takes the value of the Poisson’s ratio. The retardation
place which is followed by a slow creep process toward spectrum also can be reinterpreted to yield the perme-
equilibrium over the next 60 min or so. One method ability of the solid matrix. Thus we see that the
used to account for this is to use a lumped-parameter, evolution of our understanding of cartilage deforma-
single-phase, spring-and-dashpot, viscoelastic model tion, experimental and theoretical, leads us naturally
without regard for the interstitial fluid flow and into a multiphasic model for the tissue.
internal redistribution of the organic matrix and the A comprehensive application of viscoelastic theory
compaction within the cartilage specimen (Camosso to cartilage in tension has been made by Woo er al.
and Marotti, 1962; Hayes and Mockros. 1971; Coletti (1979, 1980). From their experiments on the stretching
ZI al., 1972; Parsons and Black, 1977, 1979). properties of articular cartilage, they developed a
Kempson (‘Ial. (1971,19SO), in examining the results quasi-linear viscoelastic model, which assumes that the
of their indentation experiments, used an expression kernel of the stress-strain-history integral is a function
for what they refer to as a ‘two-second creep modulus’ of strain and time. This is similar to biomechanical
for cartilage, based on a study by Waters (1965) on the models for other types of soft tissues (Fung, 1981).
indentation of thin sheets ofvulcanized natural rubber. From it Woo er al. found that they could accurately
The latter assumes that the classical Hertzian solution predict the relaxation and cyclic behavior, but at the
for a semi-intinite elastic layer can simply be multiplied same time, they were unable to obtain a reasonable
by an appropriate dimensionless function to account value for the elastic stress if the strain rates were low.
for a tit-rite depth. For a plane-ended indenter, he At higher strain rates. the biphasic viscoelastic effects
concluded that would begin to dominate, and the movement of
interstitial fluid must therefore be taken into account.
(4) Recently, Li er al. (1983) produced experimental
evidence, which correlated well with a biphasic analysis
where 4 was determined experimentally. This result is of cartilage strips in tension, that interstitial fluid flow
functionally equivalent to equation (3), but this empi- could be important if the strain rates of the tensile
rical approach says nothing about the displacement experiment are sufficiently high, Fig. 7.
tield of the elastic layer and little about physical
constraints, such as boundary conditions. In any case, Mulliphnsic theories
from the measurement of the indentation, two seconds It is apparent from experimental results. such as
after application of the load, Kempson used equation those obtained from indentation tests, that the move-
(4) to determine E. The resulting value, the’two-second ment of the interstitial fluid plays a fundamental role in
creep modulus’, is supposed to include the initial elastic the dynamic deformational behavior of the tissue. In
response as well as a small portion of the creep. This is, other words, to obtain a realistic rheological model for
3Y4 VAN C. i%tow.MARK H. HOLMES and W. MICHAEL LAI

c fluid phase and the solid phase, and ps is the apparent


density of the solid. The coeRicient z is defined as the
solid content, which is given by the ratio of solidity
r’,/ C; to porosity C;/ C; of the tissue. Here I’,. C’,.and
V, are the solid, fluid and total volume of the tissue,
respectively. The momentum balance equations for the
two components of the mixture. omitting inertial
forces, are
divd-k’(v’-vyl) = 0, @a)
and
div 0’ + K (9 - vJ ) = 0. (6b)
where d and u/are the apparent stress tensors for the
solid and fluid phases, respectively. The second term in
equations (6) represents the diffusive drag arising from
the relative velocities between the fluid and solid
LONGITUDINAL STRAIN (%I
components. Even in the unloaded state, the diffusive
Fig. 7. The stress-strain data for superficial-zone articular coefIicient K has values on the order of 10L5N .s rnmS,
cartilage at three difkrent strain rates. The strain rate and the which helps explain why the inertial terms are omitted
number of specimens tested at each strain are shown on the
in equations (6). Under slow flow conditions, the
figure. Those data are for the small-strain region. For small
strains, the biphasic model predicts that the load per unit area
diffusive coefficient is related to the permeability k of
on the specimen is the tissue by the inverse relation (Lai and MOW, 1980)
2 = h2F:
A
irE


+

-4
k,H,
where E, and II, are the Young’s
A0 - c’ A, exp( -2,2”2 ~~k,,r/h*)
“=I
modulus and aggregate
k=----
1
(1 +a)*K’

modulus of the solid matrix and k, and h are the specimen In the simple version of the biphasic theory used to
permeability and thickness. respectively. and A,, model articular cartilage, known as the KLM model
A ,, . A,. and sI,are constantsdependent on the moduli for cartilage, the solid phase is assumed to be isotropic
of the solid matrix (Li et al.. 1983). and linearly elastic, and the interstitial fluid is inviscid
(Mow and Lai, 1980). Accordingly, assuming small
biomechanical studies on articuiar cartilage. it is strains, the isotropic stress-strain relationship for the
necessary to account for the fluid component as a first order theory for the solid phase is
distinct phase of the system within the tissue. This
d = - apl + i.,eI + Q,e, @a)
means that, at the very least, cartilage should be
modeled as a biphasic (two-phase) material, with the and for the fluid phase it is
solid matrix and the interstitial fluid as the two phases.
flf = -PI, (W
This point of view has been taken by a number of
investigators such as Fessler (1960). McCutchen where p is the apparent fluid pressure, A,, pr are the
(1962), Zarek and Edwards (1964), Torzilli and Mow intrinsic elastic moduli of the solid matrix in the
(1976a.b). and Higginson cl al. (1976). It should be mixture, and e is the infinitesimal strain tensor describ-
pointed out that even though a two-phase model is ing the deformation of the solid matrix. It is also
adequate for the description of the motion found in assumed that the tissue is spatially homogeneous, SO i.,
most mechanical testing done so far, a more involved and y, are constants.
model of articular cartilage would be multiphasic, Because of the generality of the formulation of this
taking into account the influence of the mobile elec- biphasic model, other terms can be included in the
trolytes and the collagen-proteoglycan solid phase as a linearized constitutive laws, equations (8). which can
fiber-reinforced composite porous matrix (Myers et incorporate, for example, a viscoelastic solid matrix,
(II., in press). diffusive couples and capillary forces. Also, the iso-
A biphasic model for articular cartilage was begun tropic case is presented here, but formulae for the more
by Torzilli and Mow (1976a.b) and extended by MOW general case of biphasic anisotropic stress-strain laws
and Lai (1979) and Mow et al. (1980). based on the can be found in Mow and Lai (1979). These extensions
mixture theory of Craine et al. (1970) and Bowen are important to remember because the assumptions of
(1976). In essence, this model depicts articular cartilage the linear KLM model are only approximately correct.
as a soft, porous and permeable, elastic solid filled with For example, cartilage appears to have a significant
water. Assuming that both phases are intrinsically anisotropic structure, as manifested by the split-line
incompressible, the continuity equation for this binary patterns and anisotropic tensile properties (WOO ef a!..
mixture is 1976; Kempson, 1980; Roth and Mow, 1980). Also, it
has been found that in shear the organic solid matrix is
divv’+cxdivv”+rl(v’-v/)-grad In p’= 0, (5)
slightly viscoelastic (Hayes and Bodine, 1978; MOW et
where v’ and v” are, respectively, the velocities of the al., 1982). So a more accurate model of articular
Properties of articuhr cartilage: a review $85

cartilage would have to include terms to describe the compression. the fixed-charge density of the tissue
mtrinsic viscoelastic behavior of the solid matrix. increased while the permeability decreased.
However, in compression the predominant mechanism The approach in all these studies of cartilage
giving rise to the observed viscoelastic behavior of permeability is basically the same. Fluid is forced to
cartilage appears to arise primarily from the diffusive flow through a cylindrical disc of cartilage that has
drag caused by the interstitial fluid flow through the been removed from the bone by applying a direct fluid
solid matrix, and depends only secondarily on the pressure P,, across the tissue. For a one-dimensional
v iscorlastic properties of the solid matrix itself. In fact. flow through a sample of thickness h, the apparent
we have found that the linear KLM theory can permeability k0 measured in these permeation experi-
adequately describe the observed creep and stress- ments is determined from the empirical Darcy’s law,
rslaxation behavior of articular cartilage and nasal which states that
cartilage, as well 3s meniscus. under isothermal and
constant electrolytic conditions (Mow and Lai. 1980; (9)
Mow er al.. 1980; Mow and Schoonbeck. 1982;
Favenesi t’r (11.. 1983).
where Q is the volume flux of permeated fluid through
a permeating area A of the specimen. The difficulty in
PER%lE.ABILITY determining the permeability of soft biological tissues
with this procedure is that the permeation process
The permeability of the tissue is a macroscopic gives rise to a drag force (i.e. the force exerted by the
measure of the ease with which fluid can flow through fluid on the solid as it flows through the specimen) of
the matrix. One of the first studies of permeability was significant magnitude to compact the soft, permeable
made by McCutchen (1962). who examined the uni- solid matrix in a non-uniform manner. This compac-
axial compressive properties of cartilage by creep- tion decreases the permeability within the tissue, and
testing cylindrical plugs of bovine shoulder cartilage. the decrease varies with the distance from the surface.
For the surface layer, he found the average value of the making the measured value of k,, an average. lumped-
permeability to be 5.g x IO-” mJ/N.s and he ob- parameter value. This effect is particularly important
served a decrease in the permeability with depth from for soft tissues such as articular cartilage, where
the articular surface. This dependence on depth was substantial compaction of the matrix can easily occur.
also examined by Maroudas t’t al. (1968). who found Experimentally, this is manifested by the dependence
the permeability for human femoral condyle cartilage of Lo on the driving pressure P, that was first observed
to increase from the supcrticial region to the middle for articular cartilage by Mansour and Mow (1976).
region by about 35 “,,. then to decrease from the middle Note that for porous, permeable elastic materials
region to the deep region of the tissue by about X0”,. where the solid matrix is very stiff, e.g. soils or sintered
This inhomogeneity has been attributed to the dense metallic materials, this effect is negligible or non-
network ofcollagen fibers in the superficial zone (Muir existent. Thus this flow-limiting effect appears to be
cr a/., 1970) and to the increase in charge density in the particularly important only in soft biological multip-
deep zone (Maroudas rr al., 1969). In fact, Maroudas basic materials.
(1975b) found that the decrease of permeability below To separate the effects of the clamping strain E, used
the surface layer correlates significantly well with the to hold the specimen discs in the permeation exper-
observed increase of the fixed-charge density, Fig. 4a. iment and the applied pressure PA, Mow and Lai
This dependence of the permeability on the ionic (1980) obtained a family of permeability curves L,(E,:
content of the bathing solution was first demonstrated PA) with PA as the parameter. In doing this. it was
by Edwards (1967) on cartilage from the hips of dogs. empirically demonstrated that there is an exponential
He found that by increasing the ionicconcentration. by decrease of the permeability function with sc. Figure 8
changing from normal saline to normal Ringer’s, the shows how the permeability changes as a function of E,
permeability increased by a factor of about three. The for various parametric values of the constant pressure
dependence of the permeability of articular cartilage drop PA used to maintain steady permeation. The
c-n the compressive strain in the tissue was first shown empirical exponential law used to curve-fit the data
by Mansour and Mow (1976). More recently, ad- was
ditional data were obtained, and this compression-
strain-dependent permeability was shown empirically k, = A(P,)exp[-W,)s,)]. ( 10)
to depend on the applied compressive strain and the
where A(P,) and a(P,) are also shown in Fig. 8.
applied pressure gradient in an exponential manner,
To incorporate this result into the biphasic theory,
Fig. 8. It was found that there is a non-linear decrease
Lai and Mow (1980) introduced the concept of in-
in the permeability with increased compressive strain,
trinsic permeability, k, defined as
which leads to a significant reduction in the ability of
the fluid to flow through the matrix. This latter X-= lim k,(s,; P,4) = li,exp(Ms), (II)
P,-0
measurement was a simple mechanical method to
verify the correlation of permeability to fixed-charge where k, = A(0). .V = a(0). and E = -E, is the dila-
density obtained earlier by Maroudas er al., since with tation field. It is important to note that this function is
386 VAN C. Mow, MARK H. HOLMES and W. IMICHAEL LAI

PERMEABlLlTY vs. APPLIED STRAIN

p,: . 0 fJF,q MN/,,?

l OS?2 MN/m2

0 I6 24 32 40 l
C
Applied Compressive SlroYn (%I

Fig. 8(a). Experimental curves of permeability versusapplied compressive strain at various levels of applied
pressure for a sample of bovine articular cartilage. The solid curves are least square best fits of the indicated
exponential decay law, equation (10) (Mow er 01.. 1980).

1 I I I )
05 IO I5 20 P
1
Pressvre (MN/m21

Fig. 8(c). Curve of exponent z(PA) vs P, for the same bovine


specimen (Mow et aI., 1980).

through a circular cylinder is used


05 IO I5 20 P naf PA
PressuretMN/m*l
A
Q=-gy
Fig. 8(b). Curve of the amplitude function A(P,) and exper- Here /Ais the viscosity of the permeating fluid, ai and Ii
imental data points vs P, for the same bovine specimen (Mow are the radius and length of each pore, Fig. 9a, and P,, is
er al., 1980). the applied pressure at the upstream end of the tube.
For a specimen whose permeated area is A, and
thickness h. and with n number of pores, the area1
defined for the theoretical limit of PA -+ 0, which can be porosity and tortuosity factor would be given by flA
obtained parametrically from Fig. 8 by extrapolation. =c;_, A,/A, and di = ii/h, respectively. For a un-
For normal bovine cartilage immersed in normal iform Poiseuille model whose pore structure is homo-
Ringer’s solution, we find from the permeation exper- geneous and isotropic,
iment that k, = 1.7 x lO_” m4/N*s and M = 4.3.
Consequently, even for small compressive strains,
there is a significant nonlinear interaction between the and
fluid phase and the solid matrix as the tissue deforms.
It is of interest to note how the estimates of the ,=!!=’
h ii’
‘uniform pore size’ of these cartilage specimens, which
appear in the literature, have been obtained where a and I are the assumed constant radius and
(McCutchen, 1962; Maroudas, 1973). Usually the length of the ‘circular cylindrical pore’. Inserting these
Poiseuille formula for steady-volume rate of flow, Q, model assumptions into equation (12) above and

-’ .
Properties of articular cartilage: a review

lations show that the pore size decreases practically


linearly with the bulk compression. This leads one to
conclude that bulk compression of cartilage is really
achieved by expelling the fluid from the intrrstitium
and that the organic matrix is essentially incom-
pressible. One must warn. however. that this type of
heuristiccalculation may be grossly inaccurate by itself
and should not be used strictly by itself but rather
should be used to supplement other. more rigorously
derived results such as those presented in this review.

Fig. 9(a). A uniform tube Poireuille flow model for assessing


SOS-LINE.aK FLLID SOLID ISTERACTIO\ DL-RISC
the ‘average pore size’ ofsrticular cartilage. In this model, all
circular cylindrical tubes are uniform in size (I, = I and ,-ii CO\lPREsSIOS
= .41. It is assumed that the porosity is isotropic so that the
areal porosity is equal to the volume porosity. The above discussions pertain to steady-state be-
havior ofcartilage during compression or permeation.
They explain the need to consider a nonlinearly
permeable biphasic model for articular cartilage under
compression. This concept of nonlinear, strain-
dependent permeability must be tested in other de-
formational conditions to verify its general validity
(Lai ef al., 1981; Holmes c~ ul., in press). A number of
different testing procedures have been used to ac-
complish this. They generally involve the nonsteady or
X x 0.669 MN/m* dynamic deformational behavior of cartilage; one
method is the in sifu stress-relaxation indentation test
described earlier (Mak er nl., 1982). and another is the
PERMEABILITY
uniaxial confined-compression experiment. A par-
ticular example of thcsc latter experiments arises when
cylindrical ostrochondral plugs arc compressed
03 against a free-draining, rigid. porous filter at the
articular surface. The pores of the filter are small
APPLIED COMPRESSIVE STRAIN (%)
enough so that the filter may be used to compress the
Fig. 9(b). Variation of average pore size as assessed by the solid matrix of the tissue, yet large enough to allow
uniform tube model. The method allows assessment of ‘free’ exudation of the Huid across the surface. At the
change of interstitial pore size with compression and applied same time, the plug of material is inserted into a snugly
pressure. Average solid content z and average porosity /I were
fitting confining chamber to prevent significant
assumed IO be 0.22 and 0.81 respectively. and d = tortuosity
Sactor. amounts of lateral expansion as the tissue is being
compressed normally across the articulating surface of
the specimen, Fig. 10. This experimental protocol was
equating this volume rate of flow with that defined by imposed in an attempt to achieve the theoretical
Darcy’s law, equation (9). the&average pore radius’may condition of one-dimensional confined compression -
be shown to be

(13)

Here k, is the experimentally determined apparent


permeability, equation (10) or (I 1). With known values
of tissue porosity p-O.8 and water viscosity, Fig. 9b
shows the range of average pore radii for normal
bovine articular cartilage in terms of the tortuosity
factor 6. There is no estimate for the tortuosity
factor-the path length actually traveled by a fluid
particle as it traverses the specimen of thickness h. It
would not be unreasonable to assume that this factor
ranges between one and two. For example, for d = 2, Fig. 10. Schematic depiction of the confined-compression
we see from Fig. 9b that for uncompressed tissues biphasic experiments. The rigid porous loading (with either
prescribed tractions or displacements) allows exudate to flow
(from the intrinsic permeability curve), a - 56 A and
freely into the pores of the tilter. The confining chamber is
with 40”” (clamping or bulk) compression a - 30A used to approximate the one-dimensional theoretical require-
(assuming S = 2 for both cases). These simple calcu- ment assumed in the analysis.
388 VANC. Mow. MARK H. HOLMESand W. MICHAELLAI

and it has been found by Mow et al. (1980) and tidemark, i.e. the noncalcilied~lcified juncture. In
Armstrong and Mow (1982a.b) to be a significantly other words, prior to this time, the compression is
accurate method to assess the intrinsic material pro- confined near the surface, so the creep response of the
perties of the solid matrix, its aggregate modulus H, tissue in confined compression is unaffected by the
-_i,+ 2~,, and its linear permeability coefficient k,. boundary at the tidemark. For the long-time behavior,
In an extension of the linear KLM biphasic theory the displacement exponentially approaches its equilib-
of cartilage to incorporate the nonlinear permeation rium value of IIF,,/H,~, which is determined by the
process, equation (11). the equation for the one- aggregate modulus of the linearly elastic solid matrix.
dimensional deformation of the solid matrix along the One of the most important aspects of this asymptotic
axis of the plug was shown to follow the non-linear result is that the early time response of the tissue, as
diffusion equation (Mow and Lai, 1980; Holmes et a!., observed by a creep experiment, is proportional to J;,
in press). irrespective of the load Fe. This behavior can be easily
verified from the creep data. Figure 11 shows some of
H, 2 =k, exp (14) our recent experimental data on normal bovine car-
tilage verifying the asymptotic solution given by
where the intrinsic permeability function, equation equation (17). The other important aspect is that, for
(11). has been used to derive equation (14). Here u(z, t) small strains, the equilibrium stress and the equilib-
is the axial component of the displacement vector rium strain in compression are linearly related; this has
describing the deformation of the solid matrix, Fig. 10. been repeatedly verified experimentally (Mow et al.,
The axial component of velocity c(z, r) of the intersti- 1980; Armstrong and Mow, 1982a).
tial fluid phase is found from the continuity relation- The second confined compression test measures the
ship, equation (5), with the boundary condition u = u uniaxial stress-relaxation behavior of cartilage. This
=Oatz=h experiment was first used by Lipshitz et al. (1976b) to
assess the influence of fluid Row on the stress-
a relaxation behavior of cartilage, Fig. 12. In these
I+. t) = -a z ufz,1). (15)
experiments, a ramp-displacement function is imposed
A common experiment for the uniaxial compression on the articular surface given by
configuration involves the determination of the creep for 0 < t < to (compression phase)
Vet
behavior of the material. For the creep test, a constant %3(t) =
1 V&l for t, < c (relaxation phase).
load is suddenly applied across the porous filter, and
(19)
the subsequent displacement is recorded. In the bip-
basic theory, the appropriate boundary condition at As with the creep problem, the resultant inhomoge-
the articular surface (z = 0) interfaced with the free- neous displacement field must be determined by
draining rigid porous filter is solving the nonlinear diffusion equation, equation (14).
An accurate approximation of the solution to this
0 for tG0
H,gL (16) problem can be obtained for the case of a slow rate of
i -F. for t> 0,
where Fe is the magnitude of the applied compressive
traction. The resulting mathematical problem, which t
includes equation (14), is non-linear, and at present no IS-

analytic solution is known. Therefore either numerical


or asymptotic methods are necessary to obtain a
60-
solution to the non-linear creep problem. We have
done this by examining the short and long-time
behavior of the solution (I-Iolmes, in press). It was 45-
found that, in the initial moments after applying the
load, the creep displacement of the articular surface,
30-
denoted by ue(t), is given as
h2
uo(t)=j30J;,0s t e<-- (17)
k,H,’

where for small compressive strains

/I,, =?exp(-$$($)“‘. (18)


Fig. 11. The measured displacement of the articular surface
This describes the surface displacement for cartilage during the first 120 s of a confined uniaxial creep experiment
on bovine articular cartilage with an applied load of I.2
with a thickness of 1.5 x 10m3m during the first 120s x IO’ N m -I. The linearity of these data compares well with
or so, which corresponds approximately to the time it the predicted behavior from the biphasic theory as given by
takes the deformation, u(z, t), to diffuse down to the the asymptotic solution in equation (17).
Properties of articular cartilage: a review

Control led Romoed Dtsolocement

Stress History During Romp Dispiocement


-w t UP

Timt

Schematic of Fluid Flow During Romp Displacement


EfllUlt Efflux
dir n Equilibrium
Arliculor
Surioct
Articulor
Cortilagt Tidtmorh
HI
E

Fig. I?. The top graph shows a controlled ramp displacement for a confined-compression experiment. The
middlecurve showsa typical stress rise during thecompression phaseand a typical stress-relaxation response
in the relaxation phase. The bottom schematics show that the movement of interstitial fluid predominates the
stress-history response. During the compression phase, fluid exudation gives rise to a peak stress (I,,, and
during the relaxation phase, fluid redistribution gives rise to the relief of the compacted region at the surface,
hence stress relaxation. (For more quantitative details, see Fig. 13.)

compression. For a slow rate, by which we mean representing the strain imposed in the experiment.
Based on the known typical permeation values for
(20) normal bovine cartilage, this expression is valid for
experiments in which the compressive phase lasts
a nearly uniform state of compression exists in the longer than loo0 s and, of course, for strains of no
tissue (Holmes, 1983.1984). In thiscase, theasymptotic more than about 20 y<. Higher rates of compression or
expansion of the solution for the stress history at the higher frequency excitations in compression would

h2
articular surface (-_= 0) is given by induce compressive strains beyond the validity of

U;_(I)
-&o
linear infinitesimal strain theory.
H, ~+c,exp(crMt)
1, ----<arcs
Lo4
(214
For this slow rate it is also relatively easy to
determine the solution during the relaxation phase
(Holmes et al., in press). We find that immediately
subsequent to the start of the relaxation process, at the
where
articular surface,
h*
(Zlb) L(t) - a,-.V,J;_ 0 <r-lo < 1 (23a)
co =3koH,1,
where
v,
c’=T. 2ha,
M, = (2W
t,jnH,k,exp(--EoM)’
Here the equilibrium compressive strain .eois given by
I and
pot0
Eo = ~ (22)
h up = -coH,[l +coexp(eoM)]. (23~)
390 VAN C. Mow. MARK H. HOLMES and W. MICHAEL L,u

Here CT, denotes the equilibrium compressive stress typically the maximum peak load is s 4-5 N and the
attained in the limit I - x given by equilibrium load is about 2-3 N for a 6.35 mm dia-
meter plug ofcartilage. Because of the simplicity of the
0, = c,,H,,. (231 analytical asymptotic solutions for the predicted stress
This shows that at the start of the relaxation phase, the history in these experiments, it is relatively easy to
compressive stress decreases as the square root of time determine the material constants of the model. For the
from the peak compressive stress a, achieved at the end case at hand, we used a non-linear regression analysis
of the compressive phase. This approximation applies, with equation (2la)as the object function to determine
roughly, to the first 100 s of the relaxation phase. As k, and ?lf. The aggregate modulus H,4is obtained from
time increases, the compressive stress continues to the equilibrium stress measured after complete stress
decrease, and in fact it approaches exponentially its relaxation has occurred. Figure I4 shows an actual set
equilibrium value of EdH 4determined by the linearly of data from our experiment where i = 0.0021 y,;s-’
elastic response of the solid matrix. These results show and co = 5OCQs. In this case. we find that k, = 2.51
that stress relaxation occurs because of an internal x 10-1s rn’1N.s. ‘41= 7.83, and H, = 0.41 MPa.
fluid redistribution process, which in turn is governed For the typical slow strain-rate stress-history curve
by the lower permeability within the compressed shown in Fig. 14. we see that, as in Fig. 13, immediately
tissue, k, exp( - E,,M), giving rise to a more rapid rate following the onset of the ramp compression, a
of relaxation relative to the peak stress op. A summary parabolic load-vs-time curve is observed. After this, an
of the results obtained from the asymptotic solutions is almost linear time dependence is seen during the
shown in Fig. 13. These results illustrate the fundamen- compressive portion of the experiment. If we ex-
tal role of the flow-limiting effect on the deformational trapolate this curve back to r = 0, and denote the
behavior of the non-linearly permeable tissue. It tends intercept with the a-axis. by ue, we have from equation
to make the tissue appear stiffer-as with other (21a) u0 = cc,. The contribution of the non-linear
viscoelastic materials-with increasing rates of com- permeability can be seen in Fig. I3 if one compares era
pression by virtue of the increased frictional drag = (cp-c,)/c,7 the normalized difference between
required to extrude the fluid from the tissue during the the peak stress up and equilibrium stress uz, with ue. If
compressive phase of the experiment. Thus the non- the permeability were constant, so M = 0 in equation
linear, strain-dependent permeability effect appears to (11). uR and u,, would be equal. However, in Fig. 14, it
be ideal in the diarthrodial joint loading situation-it can be seen that they are not the same; uR is about
exudes fluid when required, but only with ever- 100 y,, larger than ue. These stress-rise characteristics
increasing resistance to flow. are governed by the rate of compression I’,, thickness
We have performed numerous stress-relaxation h, aggregate modulus H,, and the intrinsic perme-
studies following a very strict experimental protocol to ability parameters k, and 121.The non-linear, flow-
insure the repeatability of the sensitive measurements limiting parameter M, as defined by Lai et al. (1981). is
(Holmes er a[., in press; McCormack, 1983). The responsible for the rise of the stress above and beyond
maximum peak load in all these experiments occurs at the values that would be obtained with a constant
the end of the compressive phase, and it increases with permeability. For the strain rates required by the slow-
increasing strain rate t. For the ‘slow’ rate experiment, rate experiment, the deviations are relatively small

10
t

1.
Time (secl

Fig. 14. Non-linear regression analysis, using equation (Zla)


Fig. 13. Depictions of the asymptotic normalized stress- and the experimental data from 3000s to 5000s. is used to
history prrdiction of a slow-rate compression experiment if determine k, and M. These values, together with H, de-
the tissue is non-linearly biphasic. The stress rise in the termined at equilibrium, are used in an exact numerical
compressive phase depends on the sum of a linear time solution (solid line) to predict :he complete stress history. The
function and an exponential time function. The stress inter- agreement is excellent over the entire 8C00 s period. This
cept u0 is different from the total stress decay uR method provides a self-consistency check on the validity of
= c,, exp(e,M) (Holmes et al., in press). For a linearly the model as well as the asymptotic solutions (Holmes rt al.. in
biphasic material. CT”= CJ~ since .LI = 0. press).
Properties of articular cartilage: a review 19 I

durmg the compressive phase. However, during the iment, M is very sensitive to the observed stress peaks.
relaxation phase. the non-linear permeability causes a Thus we expect that this non-linear, flow-limiting
significant change in the rate ofstress relaxation. From parameter M could provide a sensitive quantitative
equation (73), it is seen that the rate of stress relaxation indicator of cartilage degeneration in diseases such as
is proportional to M,, which depends inversely on the osteoarthritis. Future work in this area must now focus
lower permeability value k. exp( -cc, M) of the com- on the correlation ofchanges of these intrinsic material
pressed tissue. The mechanism of stress relaxation is properties ofcartilage with data such as: (1) the natural
indeed fluid redistribution within the tissue, and variation of tissue structure and biochemistry; (2) the
interstitiai fluid flow in the compressed tissue must pathological. e.g. osteoarthritic, changes of collagen
pass through the matrix with this layer of lower fibrillar ultrastructure and proteoglycan conformation
permeability. Thus a higher stress-relaxation rate is found in diseased tissues; and (3) the variation of
seen relative to the peak stress op. It should be pointed monovalent (Na’), divalent (Ca’ ‘), and other ions in
out that the material parameters were determined the interstitium. These problems pose extraordinary
using the data from the compressive phase of the challenges not only to biomechanicians but to a whole
experiment; hence, the relaxation phase provides an host of allied biomedical scientists.
internal check on the self-consistency of the biphasic
theory, and, as can be seen in Fig. 14, the agreement is Arknowlrdgzmml-This material is based on research sup-
excellent. ported by National Science Foundation grant MEA
82-11968. National Institute of Arthritis, Diabetes, and
Digestive and Kidney Diseases grants AM 19093 and AM
26440. and the Clark and Crossan endowment 81 Rensselaer
Polytechnic Institute. Any opinions, findings and conclusions
This review has focused on three specific aspects of or recommendations expressed in this publication are those
of the authors and do not necessarily represent the wews of
our understanding of cartilage biomechanics. First,
the National Science Foundation. We wish to thank Mr.
relevant biochemical, physicochemical, and ultrastruc- Brendan McCormack and Ms. W. B. Zhu for their technical
tural knowledge is presented so that the reader can assistance in obtaining the stress-relaxation and creep data.
interpret the biomechanical properties of this tissue Mrs. Lynne Nagengast and Ms. Joyce A. Brock for their
assistance in typing this manuscript. and Ms. Rose .A. Boshoff
from a micromechanical point of view. Second, a
for her editorial assistance.
detailed historical account is presented of prior at-
tempts to model articular cartilage indentation be-
havior. An appreciation of the evolution of thought on
the development of cartilage deformational theories,
REFERENCES
leading up to the development of the biphasic model
for articular cartilage, is both instructive and en- Armstrong, C. G. and Mow, V. C. (1980) Friction. lubrication
lightening. Third, a summary of our most recent and wear of articular cartilage. Scienrijic F5undati5ns of

Orrhopaedics and Traumatalogy (Edited by Owen, R..


investigations on the influence of the non-linear strain-
Goodfellow, J. W. and Bullough, P. G.), pp. 223-232.
dependent permeability effect on the compressive William Heinemann, London.
creep and stress-relaxation behavior of the tissue is Armstrong, C. G. and Mow, V. C. (1982a) Variations in the
presented. We found, by two separate, independent intrinsic mechanical properties of human articular cartilage
with age degeneration and water content. J. Bone Jt Surg.
tests-the steady, direct permeability experiment
64A, 88-94.
and the transient compressive stress-relaxation Armstrong. C. G. and Mow, V. C. (1982b) Biomechanics of
experiment-that cartilage permeability must be de- normal and osteoarthrotic articular cartilage. Clinical
scribed by the function k = k, exp(EM) where k, and Trends in Orthopoedics, (Edited by Wilson, P. D. and
Mare the intrinsic permeability parametersand Eis the Straub, L. R.), pp. 189-197. Thieme-Stratton, New York.
Armstronn. C. G.. Mow. V. C.. Lai. W. M. and Rosenberg. L.
dilatation given by E = trVu. This non-linear perme-
C. (198 1)Biorheological properties of proteoglycan macro-
ability function has profound effects on the compress- molecules. Transactions of the 27th Annttal Meeting of the
ive creep and stress-relaxation behavior of the tissue, Qrtkopaedic Research Sociery Vol. 6. p. 90. ORS, Chicago.
since these phenomenological viscoelastic effects are Aspden, R. M. and Hukins. D. W. L. (1981) Collagen
predominately governed by the rate of interstitial fluid organization in articular cartilage, determined by X-ray
diffraction and its relationship to tissue function. Proc. R.
flow. The physicochemical basis for this nonlinear Sot. 212B, 299-304.
flow-limiting effect derives from the significant direct Bar, E. (1926) Elastizitatsprufungen der gelenkknorpel. Arch.
correlation between cartilage permeability and tixed- EntwMech. Org. IO& 739-760.
charge density from steady-permeation experiments. Benninghoff, A. (1924) Experimentels untersuchungen uber
den einfluss verschiedenartiges mechanischer beanspruc-
Compaction of the tissue will correspondingly increase
hung auf den knorpel. Verh. anat. Ges., Jena 33, 194-200.
its fixed-charge density and hence decrease its perme- Benninghoff, A. (1925) Form und bau der gelenkknorpel in
ability. Thus the nonlinearly permeable biphasic ihren besiehungen zur funktion, 2. Der aufbau des gelenk-
model for cartilage should be used whenever detailed knorpel in seinen beziehungen zur funktion. Z. Zelljiirsch.
mikros. Anat. 2, 783-862.
deformational responses of the tissue in compression
Ballet. A. J. and Nance. J. L. (1966) Biochemical findings in
or detailed material-parameter characterizations of the normal and osteoarthritic articular cartilage. II.
tissue are desired. Preliminary experimental results Chondroitin sulfate concentration and chain length, water
show that, for example, in the stress-relaxation exper- and ash content. J. c/in. Incest. 45. 1170-l 177.
392 VAN C. Mow, MARK H. HOLMESand W. MICHAELLAI

Bowen, R. M. (1976) Theory of mixture. Conrinuum Physics, twhavior ofarticular cartilage in tension. J. biomech. Rngng
(Edited by Eringen, A. E.), Vol. III. pp. l-127. Academic 103,221~231.
Press, New York. Hardingham, T. E. and Muir. H. 11974) Hyaluronic acid in
Bowes, J. H. and Kenten. R. H. (1948) The amino-acid cartilage and proteoglycan aggregation. Eiochem. 1. 139,
composition and titration curve ofcollagen. Biochrm. 1.43, 565-581.
358-365. Hascall, V. C. (1977) Interaction of cartilage proteoglycans
Brandt. K. D. (1981) Pathogenesis ofosteoarthritis.Trxrbool. with hyaluronic acid. J. Suprumol. Strucr. 7. 101-120.
ofRheumnrology (Edited by Kelley, W. N., Harris, E. D., Jr.. Hayes, W. C. and Bodine, A. J. (1978) Flow-independent
Ruddy. S. and Sledge, C. B.), pp. 1457-1487. W. B. viscoelastic properties of articular cartilage mstrix. J.
Saunders. Philadelphia. Biomechanics 1I. 407420.
Brandt. K. D. and Palmoski. M. 1. (1976) Organimtion of Hayes, W. C.. Keer. L. M., Herrmann. G. and Mockros. L. F.
ground substance proteoglycans in normal and os- (1972) A mathematical analysis for indentation test of
teoarthritic knee cartilage. Arrhritis Rheum. 19. 209-215. articular cartilage. J. Biomechnnics 5, 541-551.
Buckwalter, J. A. and Rosenberg. L. C. (1982) Electron Hayes, W. C. and Mockros, L. F. (1971) Viscoelastic pro-
microscopic studies of cartilage proteoglycans. Direct perties of human articular cartilage. J. appl. Ph,rsiol. 31.
evidence for the variable length of the chondroitin sulfate- 562-568.
rich region of proteoglycan subunit core protein. J. hi&. Higginson. G. R. (1978) Squeeze films between compliant
Chem. 257, 893&8939. solids. Wear 46. 387-395.
Bullough, P. and Goodfellow. J. (1968)The significance of the Higginson, G. R.. Litchfield M. R. and Snaith J. (1976) Load
tine structure of articular cartilage. J. BOW Jt Surg. SOB, deformation/time characteristics of articular cartilage. fnr.
852-857. J. me& Sci. IS, 481488.
Cameron, H., Pillar, R. M. and MacNab, 1. (1975) The Hildebrand, 0. (1921) Ueber neuropathische gelenkerkran-
microhardness of articular cartilage. Chn. Orthop. Rel. Top. kungen. Lnng. Arch. Klin. Chir. Ver. Dem. 2. Chir. 1,
108, 275-278. 443-493.
Camosso, M. E. and Marotti, G. (1962) The mechanical Hirsch, C. (1944) A contribution to the pathogenesis of
behavior of articular cartilage under compressive stress. 1. chondromalacia of the patella. Aclcr ckir. stand. 90 (83j.
Bone Jt Surg. 44A, 699-709. l-106.
Clarke, I. C. (1971) Surface characteristics of human articular Holmes, M. H. (1983) A nonlinear ditfusion equation arising
cartilage -a scanning electron microscope study. J. Anat. in the study of soft tissue. Q. J. uppl. Mush. 61, 209-
108. 23m30. 220.
Clarke, I. C. (1974) Articular cartilage: a review and scanning Holmes, M. H. (in press) Comparison theorems and similarity
electron microscope study. J. Anat. 118. 261-280. solution approximations for a nonlinear diffusion equation
Coletti. J. M., Akeson, W. H. and Woo, S. L-Y. (1972) A arising in the study of soft tissue. S/A&f Appl. Math. To
comparison of the physical behavior of normal articular appear.
cartilageand thearthroplasty surface. J. Bone Jr Surg. 54A. Holmes, M. H., Lai. W. M. and Mow, V. C. (in press)
147-160. Compression etfects on cartilage permeability. Tissue
Collins, R. (1982) A model of lubricant gelling in synovial Nutrition and Vidility (Edited by Hargens. A.). Springer.
joints. J. nppl. moth. Phpr. 33, 93-123. New York.
Craine. R. E.. Green, A. E. and Naghdi. P. M. (1970)A mixture Honner, R. and Thompson, R. C. (1971) The nutritional
of viscous elastic materials with ditTerent constituent pathways of articular cartilage. J. Bone Jr Sury. 53A.
temperatures. Q. JI Mech. appl. Math. 23, 171-184. 742-748.
Dowson, D., Unsworth, A.,Cooke,A. F. andGvozdanovic, D. Hori. R. Y. and Mockros, L. F. (1976) Indentation tests of
(1981) Lubrication of joints. An Introduction IO the human articular cartilage. J. Biomechonics 9, 259-268.
Biomechanics of Joints and Joint Replacement, (Edited by Hultkrantz, J. W. (1898) Ueber die spaltrichtungen der
Dowson, D. and Wright, V.). pp. 12&145. Mechanical gelenkknorpel, Verh. anat. Ges. 12, 248-256.
Engineering, London. Kempson, G. E. (1980)The mechanical properties of articular
Edwards, J. (1967) Physical characteristics of articular car- cartilage. 7%e Joints and Synociul Fluid, (Edited by
tilage. Proceedings ojthe Insritute o/Mechanical Engineers. Sokololf, L.), Vol. 2. pp. 177-238. Academic Press. New-
Vol. 181, pp. 16-24. London. York.
Ekholm. R. and Norback. B. (1951) On the relationship Kempson, G. E., Freeman, M. A. R. and Swanson, S. A. V.
between cartilage change and function. Acta orthop. stand. (1971) The determination of the creep modulus for ar-
21, 81-89. titular cartilage from indentation test on a human femoral
Elmore, S. M., Sokololf, L., Norris, G.and Carmeci, P. (1963) head. j. Biomechanics 4, 239-250.
Nature of’imperfect’elastirityofarticular cartilage. J. oppl. Kempson. G. E.. Muir, H., Pollard. C. and Tuke. M. (1973)
Physiol. JR, 393-396. The tensile properties of cartilage of human femoral
Favenesi, J. A., Shaffer, J. C. and Mow, V. C. (1983) Biphasic condyles related to the content ofcollagen and glycosamin-
mechanical properties of knee meniscus. Transactions ofrhe oglycans. Biochim. biophp. Acru 297, 456-472.
29th Annual Meeting of rhe Orthopuedic Research Society, Lai, W. M., Kuei. S. C. and Mow, V. C. (1978) Rheological
Vol. 8, p. 57. ORS, Chicago. equations for synovial fluids. J. biomech. Engng lOOK,
Fessler. J. H. (1960) A structural function of mucopolysac- 169-186.
charide in connective tissue. Biochem. J. 76, 124-132. Lai, W. M. and Mow, V. C. (1978) Ultrafiltration of synovial
Fung, Y. C. (198 I) Biomechanics. Springer-Verlag, New York. fluid by cartilage. J. Enyng Me&. Dir. Am. Sot. cit. Engrs
Ghadially, F. N.. Ghadially, J. A., Oryschak, A. F. and Yang, 104, EMl. 79-96.
N. K. (1976) Experimental production of ridges on rabbit Lai, W. M. and Mow, V. C. (1980) Drag induced compression
articular cartilage: a scanning electron microscope study. J. of articular cartilage during a permeation experiment.
Anal. 121. 119-132. Biorheology 17, 111-123.
Gildemeister. M. (1914) Uber dieelastizitat von leimgallerten. Lai, W. M., Mow. V. C. and Roth, V. (1981) Effects of a
2. Biol. 63, 175-190. nonlinear strain-dependent permeability and rate of com-
Gocke. E. (1927) Elastizitatsstudien am jungen und alten pression on the stress behavior of articular cartilage. J.
gelenkknorpel. Verh. dr. orthop. Ges. 22, 130-147. biomech. Engng 103, 221-231.
Grodzinsky. A. J., Roth, V.. Myers, E. R., Grossman, W. D. Lanir, Y. (1983)Constitutiveequations for fibrous connective
and Mow, V. C. (1981) The significance of electromecha- tissues. J. Biontecknnics 16, l-12.
nical and osmotic forces in the nonequilibrium swelling Lebedev. N. N. and Ufliand, I. A. (1958) Axisymmetriccontact
Properties of articular cartilage: a review

problem for an elasttc layer. J. app1. &fur/t. .2freh. 22, osteoarthrosis in the dog J. Bone J! Surq. 588. 9410 1.
4.4210. McKibbin 8. and hlaroudas. A. (1979) Nutrition and mcta-
Lt. J. T.. Armstrong. C. G. and Mow. V. C. (19S3) EtTect of holism. Ad& Articukrr Cnrrilaye (Edited by Freeman. !vl.
strain rate on mechanical properties of articular cartilage in A. R.). pp. 461--186. Pitman Medical. Kent, C.K.
tension. -IS.\f& Biomechonics Symposrum (Edited by Woo, Meachim. G. and Stockwell. R. A. (1979) The matrix. .&:,i(
S. L-Y. and Mates, R.). pp. 117-120. Arricutur Cnrrihge (Edited by Freeman. \I. A. R. I. pp. I-50.
Linn. F. C. and Sokoloff, L. (1965) Movement and com- Pitman Medical. Kent, U.K.
position of interstitial fluid ofcartilage. .4rtkritis Rheum. 8, Moskowitz, R. W.. Howell. D. S.. Goldberg. V. M.. \luniz. 0
38 1~~9.4. and Pita, J. C. I 1979) Cartilage proteoglycan altcra:ions in
Lipshi:r, H.. Etheridge. R. and Glimcher. M. J. (1976a) an experimentally induced model of rabbit osteoarthritis.
Changes in hexosamine content and swelling ratio of Arthritis Rhenm 22. 15s 163.
articular cartilage as a function of depth from the surface. J. Mow, V. C., Kuei. S. C., Lai. W. M. and Armstrong, C. G.
&me Jr Sury. 58A. 114991153. (1980) Biphasic creep and stress relaxation of articular
Lipshitz. H., Movv. V. C., Torzilli, P. A., Eisenfeld, J. and cartilage: theory and cxpcriments. J. hiumech. EU~UJ 102.
Glimcher, M. J. (1976b) On the stress relaxation ofarticular 73-84.
cartilage during confined compression. Procerdiriys of’the Mow, V. C. and Lai, W. M. (19791 Mechamcs ofanimal joints.
?rh In~er~~~onal Congrers on Rheoloyy (Edited by Klason. Ann. Rec. Fluid Mech.. (Edited by Van Dyke. x1.1. Vol. 11.
C. and Kubat. J.) p. 19X. Swedish Society of Rheology. pp. 247-288.
Gottenburg. Mow, V. C. and Lai. W. M. (1980) Recent development in
Livingston. D. I.. Yeh. G. S.. Rohall. P. and Gohman. S. D. synovial joint biomechanics. SfA.bf Rer. 22. 275.317.
(196 1)Viscoelastic factors in the strength of the elastomers Mow. V. C.: Lai. W. M. and Holmes. M. H. (1982) Advanced
under complex stress by a puncture method. J. appl. Pol~m. theoretical and experimental techniques in cartilage re-
Sci. 5. Ml-45 1. search. Biomechankx Principks unil Appliccrriuns, (Edited
Mak. A. F.. Lai, W. M. and Mow. V. C. (1982) Indentation of by Huiskes, R.. van Campcn. D. and DeWiJn. J i. pp. .l7 T.i.
articular cartilage: a biphasic analysis. Ailrarlces in Martinus Nijhotf. The Hague.
Biarnyinceriny. (Edited by Thibault. L.). pp. 71-74. Mow. V. C.. Lai. W. ,M. and Redler. 1. (1974) Some surface
American Society of Mechanical Engineers, New York. characteristics of articular cartilage, I. J. Birrmtyh~mics 7.
Malcom. L. L. (1976) An Experimenfal Inresrigorion of the 449356.
Frii-riond and Dej;vmurionol Responses u/ Ar~icular Mow, V. C. and Mak, A. (in press) Lubrication of diathrodial
Cortil~lge Inter_aces to Static and Dynamic Lwding. Ph.D. joints. H~~ndhook of Bioenyinc,erin~. (Edited by Skalak. R
Thesis. University of California. San Diego. and Chien, S.). McGraw Hill, New York.
Mankin, H. J. (1974) The reaction of articular cartilage to Mow, V. C., Mak, A. F.. Lai. W. M., Rosenberg. L. C. and
injury and osteoarthritis. NTH. Engl. J. Med. 291, Tang, L-H. (in press) Viscoelastic properties of pro-
128s 1292. teoglycan subunits and aggregates in varying solution
Mansour, J. and Mow, V. C. (1976) The permeability of concentrations. J. Biomrcknnics To appear.
articular cartilage under compressive strain and at high Mow, V. C., Myers. E. R.. Roth, V. and Lahh. P. (1981 J
pressures. J. Bone Jt Surg. %A. 509-516. Implications for collagen-proteoglycan interactions from
Maroudas. A. (1970) Distribution and diffusion of solutes in cartilage stress relaxation behavior in isometric tension.
articular cartilage Biophs. J. 10. 365-379. Srmin. Arthrilis Rheum. 1 I. Supp I. 41-43.
Maroudas. A. (1973) Physicochemical properties ofarticular Mow, V. C. and Schoonbcck. J. M. (19X2) Comparison of
cartilage. Adult Arriculur Corriluye. (Edited by Freeman, compressive properties and water. uranic acid and ash
M. A. R.). pp. 131-170. Grune & Stratton, New York. contents of bovine nasal, bovine articular and human
Maroudas. A. (1975a) Fluid transport in cartilage. Ann. articular cartilage. Trrrnscrcrions q[!/rhe l&h Anr~lcoi .1feerin‘l
rhewn. Dis. 36, Suppl. 2, 77. u/ [he Orfhopcrcdk Rcsecrrch Socier,v. Vol. 7. p. 209. ORS.
Maroudas, A. I1975b) Biophysical chemistry of cartilaginous Chicago.
tissue with special reference to solute and fluid transport. Mow, V. C.. Schoonbeck. J. M.. Koob, T. J. and E>re. D R.
Biorheology 12, 233-248. (1983) Correlative studies on viscoelastic shear propertie>
Maroudas. A. (1976) Balance between swelling pressure and and chemical composition of articuhr cartikigr.
collagen tension in normal and degenerate cartilage. Trunsacrions oj‘rhe 2901 Anmcd Meet&y ojtlie O.r!&~pr~r~/r<
.vatirre 260, 808-809. Research Society. Vol. 8, p, 201. ORS. Chiwgo
Maroudas. A. (1979) Physicochemical properties of articular Muir, H. (1977) A molecular approach to the undertt.indiiig
cartilage. Atlulr Arricubrr Cart&aye (Edited by Freeman. M. of ostroarthrosis. ._(nn. rhwm. Di.s. 36. 199.-70s.
A. R.). pp. 215-290. Pitman Medical, Kent, U.K. Muir, H.. Bullough. P. and tMaroudas. A. (19701 The dtstri-
Maroudas. A.. Bullough. P.. Swanson, S. A. V. and Freeman, bution of collagen in human articular cartilage with some
M. A. R. (1968) The permeability of articular cartilage. J. of its physiological implications. J. Bow Jr Sxy. 52B.
Bone Jr Sury. 5OB. 166177. 554563.
Maroudas, A.. Muir, H. and Wingham, 1. (1969) The corre- Muir, I. H. M. (1980) The chemistry of the ground substance
lation of hxed negative charge with glycosaminoglycan ofjoint cartilage. The Joims cd Sytwicrl F/d. ~Edited by
content of human articular cartilage. Biochim. hiophys. Sokoloff, L.), Vol. 2. pp. 27-94. Academic Prcsi. %ew York.
.4ctu 177, 492-500. Myers, E. R., Lai, W. M. and Mon. V. C. (in pressa) Isometric
McCall, J. G. (1968) Scanning electron microscopy of ar- swelling studies of articular cartilage: experiments and
ticu!ar surfaces. Lnncel ii, 1194. theory. Proceedings oi ~/IL’ I tf Cllino-Juyuft-L’.S..1.
McCormack. B. A. 0. (1983) Determination of intrinsic Biomechnnics Conference. Chinese Academy of Science.
permeability coefficient of human articular cartilage from Beijing.
rate-controlled compressive stress-relaxation experiments. Myers, E. R.. Lai. W. M. and Mow, V. C. (in press b) ,A
MS Thesis. Rensselaer Polytechnic Institute, Trov. NY. continuum theoryand experimental verification ofthe ion-
McCutchen. C. W. (1962) The frictional properties ofanimal induced swelling behavior ofarticular cartilage. J. hbmwch.
joints. Wear 5, I-17. Enyny.
McCutchen. C. W. (197X) Lubrication ofjoints.The Joinrsund Myers, E. R. and Mow, V. C. (19X3) Biomechanics ofcartilage
Synord F‘luid. (Edited by Sokoloff, L.). Vol. 1, pp. 437483. and its response to biomechanical stimuli. Currilrtyi.
Academic Press. New York. (Edited by Hall. 8. I(.). Vol. 1. pp. 3 13-342. Academic Press.
McDevitt. C. A. and Muir. H. (1976) Biochemical changes in New York.
the wrtilage of the knee in experimental and natural Parsons. J. R. and Black, J. (1977) The viscoelastic shear
394 VAN C. Mow. MARK H. HOLMES and W. MICHAEL L~I

behavior of normal rabbit articular cartilage. J. Torzilli. P. A. and Mow, V. C. I 19761) On the fundamental
Biomechanics 10.21-30. fluid transport mechanisms through normal and patholo-
Parsons. J. R. and Black, J. (1979) Mechanical behavior of gical articular cartilage during function, I. J. Biomechunics
articular cartilage; quantitative changes with alteration of 9. 541-552.
ionic environment. J. Biomechanics 12, 765-773. Tortilli, P. A. and Mow, V. C. 11976b) On the fundamental
Pasternack, S. G.. Veis, A. and Breen, M. (1974) Solvent- fluid transport mechanisms through normal and patholo-
depndent changes in proteoglycan subunit conformation gical articular cartilage during function. Il. J. Biomrchanics
in aqueous guanidine hydrochloride solution. J. bid. Chem. 9, 587606.
249, 2X6-22 I I. Torzilli, P. A.. Rose. D. E. and Dethemers. D. A. (1982)
Policard. A. ( 1936) Phgsiologie Generule des Arriculurions (I Equilibrium water partition in articular cartilage.
1’Etnt Sumal ef Pafhotogique. Paris. Biorheolnyy 19, 519-537.
Rosenberg, L. C., Hellmann, W. and Kleinschmidt, A. K. Urban. J. P. G.. Maroudas. A.. Bayliss, M. T. and Dillon, J.
(1975) Electron microscopic studies of proteoglycan aggre- (1979) Swelling pressures of proteoglyc;ins at concentra-
gates from bovine articular cartilage. J. biol. Chem. 250. tions found in cartilaginous tissues. Biorhrulogy 16,
1877-1883. 447-464.
Roth, V. and Mow. V. C. (1980)The intrinsic tensile behavior Venn. M. F. (1978) Variation of chemical composition with
of the matrix of bovine articular cartilage and its variation age in human femoral head cartilage. rlnn. rheum. Dis. 37,
with age. J. Bone Jt Surg. 62A, 1102-l 117. 168-174.
Salter, R. B., Simmonds. D. F., Malcolm, B. W.. Rumble, E. J., Venn, M. and Maroudas. A. (1977) Chemical composition of
MacMichaei, D. and Clements. N. D. (1980) The biological normal and osteoarthrotic femoral head cartilage. Ann.
effect of continuous passive motion on the healing of full- rheum. Dis. 36. 121 -129.
thickness defects in articular cartilage. J. Bone Jf Surg. Warwick, R. and Williams. P. L. (1973) Gray’s .tnafomy. W. B.
62A. 1232-1251. Saunders, Philadslphia.
Schofield, J. D. and Weightman, B. (1978) New knowledge of Waters, N. E. (1965)The indentation of thin rubber sheets by
connective ageing. JI c/in. Puth. 31, Suppl. (Roy. ?oll. cylindrical indenters. Br. J. appl. Php. 16. 1387-1392. _
Path.) 12. 1711190. Woo. S. L.-Y.. Akeson. W. H. and Jemmott. G. (1976)
Schubert, G. and Hamerman. D. (1968) A Primer on Mrasuremenis of nonhomogeneous. directional mecha:
Connecrice Tissue Biochemistry. Lea and Febiger, nical properties of articular artilage in tension. J.
Philadelphia. Biomechanics 9. 785-791.
Simon. W. H. (1971) Scale effect in animal joints. Arthritis Woo, S. L.-Y.. Lubock, P., Gomez, M. A., Jemmott, G. F.,
Rheum. 14, 493-502. Kuei, S. C. and Akeson. W. H. (1979) Large deformation
Sokoloff, L. (1963) Elasticity of articular cartilage: effect of nonhomogeneous and directional properties of articular
ions and viscous solutions. Science 141, 1055-1056. cartilage in uniaxial tension. J. Biomechanics 12,437-446.
Sokoloff, L. (1966) Elasticity of aging cartilage. F&n Proc. Woo, S. L.-Y., Simon, B. R.. Kuei. S. C. and Akeson, W. H.
Fedn Am. Sots exp. Bid. 25, 1089-1095. (1980) Quasi-linsar viscorlastic properties of normal ar-
Sokoloff, L. (1969) Tire Biology o/Drgenerotioe Joint Disease. titular artilage. J. biomerh. Engng 102. 85-90.
The University of Chicago Press. Chicago. Woo, S. L.-Y., Gomez, M. A. and Akeson, W. H. (1981) The
Sokoloff, L. (Ed.) (1978)7%e Joints und Synooial Fluid, Vol. I. time and history-dependent viscoclastic properties of the
Academic Press, New York. canine medial collateral ligament. J. biumech. Engrlg 103,
Sokoloff, L. (Ed.) (1980) The Joints and Synod Fluid, Vol. 2. 293-298.
Academic Press. New York. Zarek, J. M. and Edwards, J. (1964) Dynamic considerations
Speer, D. P. and Dahners. L. (1979) The collagenous architec- of the human skeletal system. Biomechanics d Relared
ture of articular cartilage. C/in. Orthop. Rel. Res. 139, Bio-Engineering Topics, (Edited by Kenedi, R. M.), pp.
267-276. 187-293. Pergamon Press, Oxford.

You might also like