You are on page 1of 26

BE18CH10-Klein ARI 15 June 2016 8:39

ANNUAL
REVIEWS Further
Click here to view this article's
online features:
• Download figures as PPT slides
• Navigate linked references
• Download citations
• Explore related articles
Lubrication of Articular
• Search keywords
Cartilage
Sabrina Jahn, Jasmine Seror, and Jacob Klein
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

Weizmann Institute of Science, Rehovot 76100, Israel; email: jacob.klein@weizmann.ac.il


Access provided by 196.190.84.122 on 02/18/24. For personal use only.

Annu. Rev. Biomed. Eng. 2016. 18:235–58 Keywords


The Annual Review of Biomedical Engineering is articular cartilage, osteoarthritis, friction and lubrication, boundary
online at bioeng.annualreviews.org
lubrication, cartilage lubrication, hydration lubrication, hyaluronan,
This article’s doi: lubricin, phospholipids, supramolecular synergy
10.1146/annurev-bioeng-081514-123305

Copyright  c 2016 by Annual Reviews. Abstract


All rights reserved
The major synovial joints such as hips and knees are uniquely efficient tri-
bological systems, able to articulate over a wide range of shear rates with
a friction coefficient between the sliding cartilage surfaces as low as 0.001
up to pressures of more than 100 atm. No human-made material can match
this. The means by which such surfaces maintain their very low friction has
been intensively studied for decades and has been attributed to fluid-film
and boundary lubrication. Here, we focus especially on the latter: the re-
duction of friction by molecular layers at the sliding cartilage surfaces. In
particular, we discuss such lubrication in the light of very recent advances
in our understanding of boundary effects in aqueous media based on the
paradigms of hydration lubrication and of the synergism between different
molecular components of the synovial joints (namely hyaluronan, lubricin,
and phospholipids) in enabling this lubrication.

235
BE18CH10-Klein ARI 15 June 2016 8:39

Contents
1. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
1.1. Some Basic Concepts of Frictional Dissipation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
1.2. Friction in the Major Joints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
1.3. Cartilage Friction and Osteoarthritis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
1.4. Fluid Films and Boundary Layers at Cartilage Surfaces . . . . . . . . . . . . . . . . . . . . . . . 239
2. PHYSIOLOGY OF ARTICULAR CARTILAGE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
2.1. Molecular Constituents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
2.2. Chondrocytes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
3. MODES OF CARTILAGE LUBRICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
4. BOUNDARY LUBRICATION OF CARTILAGE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

4.1. Hyaluronan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243


4.2. Lubricin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

4.3. Phospholipids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246


5. HYDRATION LUBRICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
6. SUPRAMOLECULAR SYNERGY IN CARTILAGE LUBRICATION . . . . . . . . . . 248
7. SUMMARY AND PERSPECTIVE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251

1. INTRODUCTION

1.1. Some Basic Concepts of Frictional Dissipation


Concepts of friction, lubrication, and wear are crucial for maintaining physiological functions
in living systems whenever surfaces in contact slide past each other, particularly in mechanically
stressed environments. The major human synovial joints such as hips and knees are prime examples.
They are remarkable biotribological systems, capable of withstanding maximal local pressures up
to 18 MPa (∼180 atm) or more during locomotion (1), while exhibiting very low friction. Indeed,
as discussed below, abnormally high friction at the joints may be closely correlated with joint
pathologies such as osteoarthritis (OA), and for this reason too it is important to understand the
origins of lubrication at the molecular level.
Friction and lubrication are most fruitfully considered in terms of energy dissipation as surfaces
slide past each other, traditionally thought of in terms of two main mechanisms (2). Viscous dis-
sipation within thin fluid films trapped between the sliding surfaces, where the energy dissipation
is Eviscous , occurs as the fluid is sheared. For (smooth, flat, parallel) surfaces separated by a fluid
film of thickness D and effective viscosity ηeff , sliding past each other at velocity dx/dt, this may

be (approximately) written as Eviscous = A [ηeff (dx/dt)/D]dx over the sliding distance, where
A is the surface area. An additional major mode of energy loss on sliding is boundary friction,
which arises when molecular contact is made between the sliding surfaces themselves, or between
molecular species attached to them, so-called boundary layers. Dissipation arises, typically, from
formation and hysteretic breakage of molecular bonds or passage of the opposing molecules over
local repulsive energy maxima, as phonons generated by these irreversible processes are converted
to heat (3). Boundary friction is more complicated to describe theoretically, but is characterized
by, for example, a much weaker velocity dependence of the friction force than the linear velocity
dependence of simple viscous dissipation of hydrodynamic lubrication (as Eviscous above). Wear
of surfaces, where molecular bonds are irreversibly broken in the formation of wear debris, is

236 Jahn · ·Seror Klein


BE18CH10-Klein ARI 15 June 2016 8:39

another clear pathway for energy dissipation by friction, as are viscoelastic losses arising when soft
surfaces undergo deformation on sliding under load. Lubrication may reduce friction between
the underlying surfaces either hydrodynamically by trapped fluid films or via suitable boundary
layers. For real surfaces in aqueous media, including biological surfaces, lubrication is often in
the so-called mixed regime, where the main mechanisms, fluid-film and boundary lubrication, are
active simultaneously, as when surfaces are in contact via asperities (4).

1.2. Friction in the Major Joints


Friction is often quantified in terms of a friction coefficient: μ = (force to slide the surfaces)/(load
compressing the surfaces). Although such a simple number does not by itself reveal the molecular
basis or the nature of the energy dissipation during the sliding—which is the underlying origin of
the friction—it provides a useful empirical comparative parameter. Measured friction coefficients
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

for whole joints in vivo have been reported to range from approximately 0.001 to 0.03 (see, e.g.,
citations in Reference 5), though it is important to emphasize that friction measurements on
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

intact, unperturbed, in vivo articular cartilage are, in fact, extremely challenging for two main
reasons, apart from the difficulty of manipulating the joints of live animals. First, the sliding of
cartilage surfaces is so well lubricated that any measured friction is likely to be dominated by other
energy dissipation pathways [such as viscoelastic losses arising from the distortion of adjacent
ligaments, membranes, and connective tissue (D. Dowson, personal communication)]. Indeed,
because of these additional dissipation modes, it is likely that the reported friction coefficient
values for living joints, low as they are (with μ down to ∼0.001), may be higher than the actual,
intrinsic, in vivo cartilage–cartilage friction coefficient. Second, studies on in vivo cartilage, as
well as on excised cartilage ex vivo, may be influenced by the known upregulation of cartilage-
degrading enzymes, which occurs in rapid response to “insult” of the living cartilage, including
incision and removal of the joint components, or undue strain on them (6–15). Such enzymes,
including matrix metalloproteases, phospholipases (13), and ADAMTS5 (an aggrecanase), are
likely to modify considerably the properties of the cartilage surface and near-surface regions
involved in friction processes relative to its native unperturbed state, as discussed in Section 1.3,
below. For these reasons, attempts to understand the extremely efficient lubrication of cartilage
have to date focused primarily on in vitro studies of the likely lubricants (both fluid-film and
boundary), using model substrates as well as excised cartilage, rather than on living joints (though
the surface properties of excised cartilage, too, are likely to differ significantly from the in vivo
surface). In particular, direct measurements in model systems have been made on the lubricating
role of synovial fluid (SF) and, separately or in combination, on molecules that are believed to
constitute the cartilage boundary lubricants, as we discuss below.

1.3. Cartilage Friction and Osteoarthritis


In vivo, joint surfaces slide past each other with modest gliding speeds, as when walking, in the
range from rest to ∼0.3 m/s (16). At the slip plane between the surfaces, where the sheared
interfacial region may be of the order of 10 nm thick, this may translate to local shear rates as high
as 106−7 /s (5, 17). When in health, the articular cartilage (see Figure 1 for a schematic) is a self-
lubricating, self-healing system, actively maintaining the lubrication between the opposing sliding
surfaces of the joints over their lifetime. Pathological conditions of the joint such as OA, which
can be described as a group of painful joint disorders with a complex etiology of both genetic and
acquired contributions (18), are manifested by damage and loss of cartilage tissue. OA is believed
to result from a cascade of events, beginning with damage to the cartilage, leading to a disruption

www.annualreviews.org • Lubrication of Articular Cartilage 237


BE18CH10-Klein ARI 15 June 2016 8:39

a b Articular surface Zones


Bone Superficial

Collagen II
Load
Synovial
membrane
Motion
Transitional
Joint cavity
Articular
cartilage

Load

Radial
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

c Aggrecan Lubricin Tidemark


Calcified
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

HA
Subchondral bone

Compressive modulus
collagen X and GAG

Figure 1
Schematic of a synovial joint, the structure of cartilage, and its molecular arrangement at the articular surface. (a) Synovial joints are
composed of opposing articulating bones, which are covered with a thin layer of cartilage embedded in synovial fluid filling the joint
cavity and enclosed by a synovial membrane. (b) Cartilage tissue is functionally and structurally organized into four main zones situated
above the subchondral bone. The superficial zone exposes the articulating, lubricated surface. Chondrocytes occupy less than 10% of
the volume of the cartilage tissue. The composition of matrix constituents of the tissue, as well as its modulus, varies with depth, as
indicated for two collagen types and glycosaminoglycans (GAGs). (c) The outer cartilage surface, showing charged macromolecules.
Hyaluronic acid (HA) (blue), lubricin ( green), and aggrecans (red ) are prominent macromolecules believed to participate in lubrication.
Phospholipids are present (see Figure 8) but are not shown here. Panel b modified with permission from Reference 23. Panel c
modified with permission from Reference 149.

of lubrication, and the disruption of lubrication in turn causes progressive structural and functional
damage of the cartilage tissue and the subchondral bone. A possible scenario correlates friction
with OA as follows: High friction [or some other initiating event, such as sport or accident traumas
(19)] leads to increased shear strain of the cartilage, which transmits itself to the chondrocyte cells
embedded within it (Figure 1b) and may cause upregulation of cartilage-degrading enzymes via
mechanotransduction (6–10, 14). Degraded tissue is then abraded and removed by the friction
upon articulation, with the resulting degraded cartilage surface exhibiting higher friction, and
thus higher shear strains on sliding, and so on in a self-reinforcing cycle. Figure 2 illustrates this
process schematically: For the sliding friction coefficient μ, pressure P, and shear modulus K, the
shear strain ε on the lower material as the top surface slides past it is given by ε = (μP/K). At typical
mean walking pressures [P ≈ 5 MPa, though this value can vary substantially over the cartilage
surface (1)] and a shear modulus of the cartilage K ≈ 5 × 105 N/m2 [though this value may vary
significantly with depth (20)], a sliding friction coefficient μ = 0.01 would lead to strains around
10% (Figure 2a), whereas μ = 0.001 would result in a 1% strain [though as the modulus increases
(20), the strain would fall with depth from the surface]. Assuming the strain is transmitted to and
sensed by the enclosed near-surface (chondrocyte) cells, the higher μ value would readily elicit
a mechanotransductive response (21, 22) and possible enzyme upregulation (9, 14), whereas the
lower μ value is less likely to do so (22). Although this is a simplified picture, its main elements

238 Jahn · ·Seror Klein


BE18CH10-Klein ARI 15 June 2016 8:39

Shear strain = (shear stress/K) = µP/K


(in steady sliding)
P = 5 × 106 Pa P = 5 × 106 Pa

μ = 0.001 μ = 0.01

1% strain 10% strain

(Roughly to scale)
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

Modulus Modulus
K = 5 × 105 N/m2 K = 5 × 105 N/m2
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

Figure 2
Schematic illustration (where the cartilage strain is indicated roughly to scale) showing how sliding friction,
with different friction coefficients μ, between upper and lower surfaces can lead to different shear strain ε of
the cartilage and of embedded entities within it. The orange disk represents a chondrocyte cell embedded in
the cartilage network and distorting affinely with it. Realistic values are given for the cartilage modulus K
(assumed constant over the region shown) and maximal normal cartilage pressures P during walking. The
shear strain in steady sliding is given by ε = μP/K.

are correct and illustrate how a friction coefficient that is significantly higher (say, μ ≈ 0.01 or
higher) than that associated with healthy cartilage (where μ ≈ 0.001 or lower) could be correlated
with the onset of OA.

1.4. Fluid Films and Boundary Layers at Cartilage Surfaces


Earlier models (see also Section 3, below) have emphasized hydrodynamic (fluid-film) lubrication,
but there is a growing realization that molecular boundary layers must play a crucial role (4), and
over recent decades the nature of the surface boundary layer on articular cartilage, and the mecha-
nism by which it lubricates, has been increasingly investigated. Molecular components ubiquitous
in cartilage and in SF, which have been implicated as forming boundary lubrication layers in this
context (17), include both macromolecules [notably hyaluronan, also known as hyaluronic acid
(HA), a linear polysaccharide; the bottle-brush-like proteoglycan aggrecan; and the mucin-like
glycoprotein lubricin, also known as PRG4 or superficial zone protein (SZP) (24, 25); as well as
others] and phospholipids (PLs) (26–30). A crucial point, which we emphasize in this review, is that
any realistic model of cartilage boundary lubrication must, at the very least, be able to reproduce
its cardinal features: that is, the physiologically low friction coefficient of articular cartilage (down
to μ ≈ 10−3 or lower) up to the maximal joint pressures (100–200 atm or more), with known
components of the synovial joint, over a wide range of shear rates and shearing times.
This article provides a brief overview of the physiology of articular cartilage relevant to its
tribology, and notes briefly some of the fluid-film models of cartilage lubrication. The main
emphasis, however, concerns recent studies and new insights into boundary lubrication, because
that appears increasingly to be the major dissipation pathway that any understanding of the low
cartilage friction will need to account for.

www.annualreviews.org • Lubrication of Articular Cartilage 239


BE18CH10-Klein ARI 15 June 2016 8:39

2. PHYSIOLOGY OF ARTICULAR CARTILAGE


We present a brief account of the main features of articular cartilage with relevance to its lubrica-
tion; greater detail (and references to myriad investigations) is provided in more specialized recent
summaries (e.g., Reference 31). The articulating ends of the bones that make up synovial joints
in mammalian bodies (Figure 1a) are covered with a thin layer of hyaline (articular) cartilage,
which in the major human joints (hips and knees) is ∼1–3 mm in thickness. Such cartilage layers
are remarkable constructs, supporting a wide range of stresses and impacts; related primary func-
tions are the distribution and minimization of contact stresses on subchondral bone generated
during loading of the joint by providing a low-friction gliding surface. Hyaline articular carti-
lage is composed of a macromolecular fibrillar network matrix, primarily type II together with
type IX and XI collagen fibrils, which provide the superstructure tensile strength and stiffness.
This network is permeated with charged macromolecules, including, importantly, HA, a linear
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

polysaccharide, which is complexed with aggrecans (schematically shown in Figure 1c) and satu-
rated with water and mobile ions, whose osmotic pressure helps counter loads on the cartilage. An
elastin network also contributes to the matrix properties, particularly in the outer superficial zone
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

(Figure 1b) (32), as does the presence of fibronectin (33). Highly specialized cells, chondrocytes
(which synthesize the cartilage components), are embedded in the matrix (34). The cartilage is
avascular and alymphatic and lacks innervations (35). Although all synovial joints consist of the
same basic components, the thickness, matrix composition, and cell density of the articular carti-
lage, and thus its mechanical properties, can vary (as schematically depicted in Figure 1b) within
the same joint and among joints or species (18). Curiously, however, mean pressures at cartilage
surfaces, a crucial determinant in cartilage lubrication, appear rather uniform independently of
mammalian species, apparently with good reason, as discussed further in Section 2.2., below. Char-
acteristic values of the cartilage shear modulus (from a study on bovine joints) have been measured
in the range 0.1–1.5 MPa, where the modulus increases with depth from the outer superficial zone
(20). Figure 1 shows a schematic of a synovial joint, its composition, and the structure of articular
cartilage.

2.1. Molecular Constituents


Other than the ∼75% overall water content of healthy cartilage [increasing from 65% near the
subchondral bone to ∼80% at the outer superficial zone (36)], the main components of the car-
tilage network include different collagen types (see the preceding section), proteoglycans, and
noncollagenous proteins. The collagen fibrils form a highly specific ultrastructural arrangement
throughout the cartilage tissue. Close to the articular surface, the tightly woven collagen fibrils
align parallel to the joint surface to form the so-called superficial zone. Its network alignment has
been observed in X-ray diffraction, polarized light, and electron microscopy studies (37, 38). The
superficial zone is divided into a thin (of order micrometers), acellular boundary layer (outer super-
ficial zone or lamina splendens), composed of fine fibrils as well as an elastin (39) and fibronectin
(33) network, and an underlying zone of ellipsoid-shaped chondrocytes.
The main noncollagenous components of the network phase of articular cartilage are proteo-
glycans. In contrast to collagen and water, the concentration of proteoglycans is lowest close to
the surface and increases with tissue depth, as do the tissues’ compressive strength (40) and shear
modulus (20). Proteoglycans are composed of a protein core and one or more glycosaminoglycan
(GAG) chains. These GAGs are unbranched chains of repetitive disaccharide units of sulfated
hexoamines, chondroitin sulfate, and keratin sulfate. Morphologically, GAGs covalently attach to
a protein backbone, forming aggrecan molecules, which have a bottle-brush-like structure. These
aggrecans in turn are linked to HA via link proteins to form a higher-scale bottle-brush-like

240 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

structure. Figure 1c schematically illustrates the structure of typical proteoglycans present in


articular cartilage. GAGs carry a high concentration of closely spaced (5–14 Å) (41), negatively
charged carboxylate or sulfate groups. These strings of negative charges are highly hydrated,
repel other negatively charged molecules, and attract freely mobile counterions, resulting in
high osmolarity. The balance of fluids and electrolytes within the cartilage tissue is sustained
by proteoglycans. The length of the central backbone of HA can vary from ∼100 to more than
104 nm (42). In articular cartilage, two main classes of proteoglycans can be distinguished:
(a) large aggregating proteoglycan monomers (aggrecans) and (b) small proteoglycans such as
biglycan, decorin, and fibromodulin (18), which may help stabilize the network matrix. The
formation of large aggregates is essential to suppress the mobility of proteoglycans with respect
to other extracellular matrix (ECM) components in order to prevent their displacement during
tissue deformation. Other notable components that have been identified in the outer superficial
zone include lubricin (24), a mucin-like protein, and PLs (43). A summary of some of the physical
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

properties of the main macromolecular components is given in Reference 17.


Access provided by 196.190.84.122 on 02/18/24. For personal use only.

2.2. Chondrocytes
Articular chrondrocytes are responsible for synthesizing and maintaining the matrix infrastructure
by which they are surrounded. They form only 1–5% by volume of the cartilage without cell-to-
cell contacts (44). Chondrocytes are capable of sensing changes in the composition of the ECM
due to the degradation of macromolecules and demands at the interface. Thus, they respond by
synthesizing the required molecules in appropriate amounts, such as type II collagen, proteoglycan
aggregates, and other specific noncollagenous proteins (18).
The behavior of chondrocytes is highly regulated by load and pressure and by the way in which
these vary dynamically, although the precise mechanisms (and, for example, why they appear
to differ for shear as opposed to normal stressing of the cells) are not yet fully understood (9,
21, 39). Clearly, chondrocyte behavior relates closely to the effect of pressure and friction at
the cartilage surface, as considered above with respect to transmitted strains. Depending on the
distinct zones of the cartilage, chondrocytes differ in their size, shape, and metabolic activity.
Various studies (see Reference 18) have suggested that the primary cilium of chondrocytes plays a
crucial role in recognizing mechanical changes, such as those that might be caused by strain in the
cartilage induced in turn by friction during articulation. It might act as a mechanotransductory
sensory feedback loop translating extracellular information to the cell in order to facilitate the
secretion of ECM (18). We emphasize the issue of the mechanosensitivity of chondrocytes within
the cartilage in this review of friction at the articular cartilage surface, precisely because of the
relation, mentioned briefly above, between friction/lubrication and the development of OA. There
is an increasing realization that the degradation of cartilage in OA, for decades considered to be a
wear-and-tear disease, may be a consequence of the upregulation of specific cellular enzymes such
as metalloproteinases, collagenases, and phospholipases known to degrade the ECM, in response
to specific mechanical stimuli (8–15), including friction-induced ones.
Given that chondrocytes are similar in articular cartilage coatings of mammalian joints of
different species, it is interesting to inquire how their known regulation by pressure relates to the
fact that, across the animal kingdom, animals vary by many orders of magnitude in mass, so that
their major joints (e.g., legs or hind paws) support very different loads. For a mammal of size L, say,
one expects its mass M to vary as its volume, that is, roughly as M ∼ L3 , and the load-bearing area
A of its joints to vary roughly as A ∼ L2 . The pressure P on the cartilage coating the load-bearing
joints might then be expected to vary as P = (animal weight/animal joint area) = M/A ∼ L3 /L2
∼ L ∼ M1/3 . Because M can vary between, say, Mcow ≈ 600 kg to Mmouse ≈ 30 g, the ratio of the

www.annualreviews.org • Lubrication of Articular Cartilage 241


BE18CH10-Klein ARI 15 June 2016 8:39

load-bearing pressure in a cow’s leg joint to that in the joint of the hind leg of a mouse might
be expected to vary by a factor (Mcow /Mmouse )1/3 ≈ 25. In fact, in a comparative study (45) of the
pressures in the joints of different animals ranging from cows to sheep, dogs, rats, and mice, the
joint pressure varied at most by a factor of three or so. Indeed there was barely a 20% difference
in pressure between cow and mouse joints, rather than the ∼25-fold variation that simple physical
scaling might suggest. Physically, this arises because the load-bearing cartilage areas of the lighter
animals were much smaller than an M2 scaling would predict, so that the pressure was correspond-
ingly much higher. We believe this may be due to evolutionary selection, as the chondrocytes in
articular cartilage across species operate optimally under similarly high pressures (39).

3. MODES OF CARTILAGE LUBRICATION


Observations relating to lubrication of natural joints have been documented at least since medieval
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

times. One of the first publications, by the sixteenth-century physician Paracelsus, who coined the
word synovia for the fluid, described the oily character of the liquid in the joint space (46). In 1691,
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

Havers (47) reported the chemical, physical, and lubrication properties of mammalian SF. More
systematically, over the past 80 years or so, many concepts and theories of how the synovial joint
is lubricated have been proposed (see, e.g., References 28, 41, 43, and 48–57) based on fluid-film
or boundary lubrication or on a mixed regime of the two (see References 4 and 17 for some recent
reviews).
Much of the early work on cartilage lubrication concerned fluid-film models, stimulated by re-
search on hydrodynamic lubrication of journal bearings in machine tribology by Osborne Reynolds
(1842–1912) and others (4, 58). In 1932, the anatomist MacConaill (59) made the first attempt
to develop a coherent theory about joint lubrication based on hydrodynamic theory, and shortly
afterward Jones (60) performed the first detailed experimental study on joint friction by mounting
a horse knee joint as a fulcrum in a pendulum machine. Pioneering work in the 1960s on hydrody-
namic (61, 62), elastohydrodynamic (61, 63), and microelastohydrodynamic (49) joint lubrication
took into account the deformability of articular cartilage—soft relative to the rigid surfaces for
which the original hydrodynamic models were developed. Further important developments in
fluid-film lubrication included the squeeze-film mechanism (64, 65) and the idea that the lubri-
cant is retained between the loaded surfaces due to some specific properties of the SF and the
articular cartilage. The latter idea was explored through Walker’s mechanism of boosted lubrica-
tion (66), Maroudas’s ultrafiltration theory (51, 67), and particularly by McCutchen’s concept of
weeping lubrication (52, 68). This model invoked the porosity of the cartilage tissue in a mech-
anism maintaining interstitial fluid pressure supporting joint load and lubrication (69), although
the nature of fluid flow/transport through cartilage tissue has been debated (48, 70, 71). Other
models included biphasic/triphasic (41) and mixed and multimode lubrication (72, 73). A more
detailed description of these and other fluid-film lubrication ideas is beyond the scope of this
article; a review is given in Reference 4.
The fluid-film theories noted above represent mixed modes of lubrication characterized by the
coexistence of interfacial regions operating under fluid-film and boundary lubrication conditions,
implying that multiple lubrication regimes might occur simultaneously. Experimental evidence
has generally supported mixed or boundary lubrication regimes rather than full hydrodynamic
lubrication (74). Thus, friction in joints has typically been measured as either independent of
velocity (indicative of boundary lubrication) (53) or decreasing with increasing velocity (indicative
of mixed lubrication) (75). A direct experimental indication that boundary contact occurs between
opposing, compressed cartilage layers is provided by measurements of the pressure distribution
across the cartilage in human hip joints. These have been carried out either in vivo, using a hip

242 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

implant incorporating pressure sensors in the femoral head, or on cadavers, with pressure-sensitive
film (1, 76). Such measurements reveal that the pressure distribution over the cartilage surface is
very nonuniform and that adjacent areas of the cartilage support highly unequal pressures, strongly
suggesting that the high-pressure regions are in boundary contact (as the fluid between them is
rapidly squeezed out to adjacent lower-pressure regions). The associated boundary lubrication on
sliding in contrast to fluid-film models must clearly take into account the detailed biochemical
and macromolecular composition and structure at the articular cartilage surface itself (17). Indeed,
much of the effort of the past decades to understand articular cartilage lubrication has been devoted
to identifying the nature of the molecular boundary layers. The following sections provide an
overview of some of this work and, in particular, describe our understanding of cartilage boundary
lubrication in the light of the recently emergent hydration lubrication paradigm.
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

4. BOUNDARY LUBRICATION OF CARTILAGE


The several lines of evidence noted above suggest that boundary friction plays a central role in
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

cartilage articulation. For this reason, the focus of most recent studies has shifted from using fluid-
film-based models (Section 3.1) to identifying the nature of the molecular boundary layer coating
the articular cartilage surface, and the mechanism that might lead to the very low friction between
two such surfaces as they slide under load (77, 78). As indicated in the introduction, the main
molecules that have been implicated in forming these lubricating boundary layers, either alone or
in combination, are HA (55, 57, 79–86) (though in earlier studies there was sometimes ambiguity
between HA as a component of SF and its role as a boundary lubricant); aggrecans (87–90); lubricin
(25, 56, 91–96); and PLs (27, 29, 30, 43, 86, 97–103). The role of other molecules present in the
synovial joint has also been examined (78, 104–109). It is appropriate to inquire concerning the
relation of these studies—most of which involve measuring friction between boundary layers of
the molecules on smooth, hard substrates, very different to articular cartilage—to their boundary
lubrication in living joints. The point is that in boundary friction the energy dissipation on sliding,
which manifests itself as the measured friction force, arises from the dynamic interaction between
the contacting boundary layers themselves, and is independent of the substrate. For rough, soft,
or viscoelastic substrates (such as cartilage) there may be additional dissipation pathways that
will manifest as additional friction, but the boundary friction itself will be a function only of the
interaction between the opposed molecular boundary layers, and thus well represented by the
studies cited above. In the following subsections, we briefly survey some direct studies of these
molecules as boundary lubricants.

4.1. Hyaluronan
HA is a high-molecular-weight, highly charged polysaccharide ubiquitous both in SF (where it
accounts for its high viscosity at low shear rates) and in articular cartilage (where, together with ag-
grecan, it contributes to the osmotic pressure giving the collagen network its mechanical resilience)
(17, 33). It was originally considered the molecule responsible for the lubricity of SF (79, 80, 110),
until it was demonstrated that HA enzymatic digestion made little difference to SF lubrication
ability (91, 111). Nonetheless, HA has often been, and is still, used as an intra-articularly injected
viscosupplement to alleviate OA symptoms.1 This is curious in view of the fact that HA solutions

1
As of 2013, the American Academy of Orthopedic Surgeons “cannot recommend” the use of HA viscosupplements for OA
symptoms (see http://www.aaos.org/news/aaosnow/jun13/cover1.asp).

www.annualreviews.org • Lubrication of Articular Cartilage 243


BE18CH10-Klein ARI 15 June 2016 8:39

P (atm)
0 9 11 12 14 17 20
0.3
a b P = 8.2 atm
40 μ = 0.3
Shear (friction) FII (mN)

Damage at point A: Fs
0.2 P ~ 3.0 atm, 30 μ = 0.1
D ~ 400 Å Fs

Fs (µN)
20
A
0.1 μ = FII /F = 0.19
10 μ = 0.01
P ~ 1.5 atm,
D ~ 500 Å
0
0
0 0.2 0.4 0.6 0.8 0 50 100 150 200 400 600
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

Normal force, F (mN) Fn (µN)


Access provided by 196.190.84.122 on 02/18/24. For personal use only.

Figure 3
Sliding friction mediated by hyaluronic acid (HA) boundary layers: friction forces as a function of normal force measured with surface
force balance. (a) Friction between a bare mica surface and a mica surface bearing HA layers on top of a lipid bilayer. The friction
coefficient (μ) is 0.19, increasing sharply when damage occurs at point A. When HA faces HA, the behavior is similar, or with a slightly
higher μ (∼0.3) (84). (b) Friction between two mica surfaces bearing attached HA layers (open triangles, stars). The friction, which may
also be influenced by bridging effects, corresponds to μ ≈ 0.3. Friction forces between surface layers consisting of complexes of HA and
aggrecan are also shown ( full and crossed symbols). Complexation of the HA with aggrecan reduces the friction coefficient substantially at
lower pressures (10 atm) relative to HA alone. Panel a modified with permission from Reference 84. Panel b modified with permission
from Reference 90.

at the concentrations in SF (and in viscosupplements) are highly shear-thinning, and at the typical
shear rates across the sliding articular cartilage surfaces (which can reach 105−6 /s) their viscosity
becomes comparable to that of water (112). In recent years HA has been examined as a boundary
lubricant, as it is known to be at the outer cartilage surface (Figure 1c) (see Section 6, below). Direct
measurements have been carried out using a variety of approaches, including surface force balance
(SFB) methods (84, 89, 90, 113) and friction force (colloid-tipped) microscopy (FFM) techniques
(82) for HA attached to model substrates, as well as torque-based methods using articular cartilage
samples (55, 57), although these were at rather low normal stresses [some 100-fold lower than
maximal pressures in joints (1)]. Friction coefficients measured directly using an SFB method (84,
89, 90, 114) reveal that HA is a rather poor boundary lubricant at high pressures (above several at-
mospheres), with μ ≈ 0.2–0.5 (Figure 3). Linking aggrecan molecules to the surface-attached HA
(Figure 3b) improves the lubrication, with μ ≈ 0.01, but again only up to several atmospheres of
pressure, beyond which the value of μ increases sharply. These values are far above the low friction
(μ down to 0.001 or less) at the high maximal pressures (up to 100 atm or more) in the major joints.

4.2. Lubricin
In 1968, Linn & Radin (115) showed that digesting HA from the cartilage surfaces of canine
joints does not affect the lubrication. In contrast, following trypsin digestion, which decomposes
proteins present on the articular surface, an increase in the friction coefficient was observed (115).
On the basis of this observation, the authors claimed that a protein is a necessary component of
the lubricant (or the mucin, as they called it), and interestingly, they conjectured that this protein
might work together with HA, possibly linking HA to the cartilage (115). [Several later studies,
using a number of approaches, directly indicated that such an interaction indeed exists (82, 85,

244 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

94–96).] Subsequently, Radin et al. (91) suggested that an HA-free fraction of SF contains the
actual lubricant, which was later identified as a glycoprotein and named lubricin by Swann et al.
(92) in 1981. Later still, this glycoprotein was shown to be identical to SZP or PRG4 (24, 25) and
present both in the outer superficial zone of cartilage and in SF. It is worth remarking that the
original study (91), based on which lubricin was claimed to be the boundary lubricant, showed
a rather moderate reduction in friction—relative to saline or HA-containing solution without
lubricin—of only a factor of two or so. Since then, many investigations (see, e.g., References 24,
93, 94, 111, and 116–118), including studies showing somewhat increased cartilage friction and
wear in lubricin-lacking mice (111, 116), have indicated that lubricin plays a role in the lubrication
of articular cartilage, and that its absence may be related to joint diseases. Direct measurements of
friction between lubricin boundary layers, however, reveal that lubricin per se is not an especially
efficient boundary lubricant, particularly in the context of the known low friction at high pressures
of articular cartilage, and its precise mechanistic role in articular cartilage lubrication remains far
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

from clear (119).


Thus, Zappone et al. (56, 120), using an SFB method, showed that the friction coefficient
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

between two hydrophilic surfaces bearing lubricin is relatively low (μ = 0.02–0.04) only under
low pressures (up to ∼4 atm) and increases to μ ≈ 0.2 under the higher pressures typical of
synovial joints (Figure 4a). Using colloidal-probe friction force microscopy, Chang et al. (82,
96) showed friction coefficient values of ∼0.1 between lubricin boundary layers, much higher
than what is expected for articular cartilage. Furthermore, these authors studied the interaction

30
a Mica: friction at high loads b
28 Hydrophobized mica
1.2 26 1 mg/mL mucin
Rinsed, 0.1 M Na+
24
Rinsed, pure water
1.0 22 Filled: compression
20 Open: μ ≈ 0.2–0.3
Friction force, f (mN)

μ = 0.6 decompression
0.8 18
Fs* (µN)

30
0.6
14
μ = 0.2 12
0.4
10

0.2 8 μ ≈ 0.01–0.02
6
0.0 4
0 1 2 3 4 5 2
Load, F (mN) 0
0 50 100 150 200
Fn (µN)

Figure 4
Sliding friction mediated by boundary layers of different mucins. (a) Friction force f as a function of the load F between sliding
lubricin-bearing mica surfaces immersed in lubricin solutions of different concentrations or following surface attachment or rinsing
(different symbols refer to the different conditions). Up to ∼4 atm the friction is low (μ ≈ 0.02–0.04), but at higher pressures the friction
is high and falls within the shaded area within the range μ = 0.2–0.6. (b) Friction force Fs∗ as a function of load Fn (both compression
and decompression) between mica surfaces bearing porcine gastric mucin, both in mucin solution and rinsed, in both water and salt
solution, as indicated. Up to pressures between the surfaces of ∼7–10 atm the friction is low (μ ≈ 0.01–0.02), increasing at higher
pressures to μ ≈ 0.2–0.3. Panel a modified with permission from Reference 56. Panel b modified with permission from Reference 106.

www.annualreviews.org • Lubrication of Articular Cartilage 245


BE18CH10-Klein ARI 15 June 2016 8:39

between lubricin and HA (82, 96) and found that, once mixed together, the lubrication between
the surfaces does not improve (82). Other studies showed that boundary layers consisting of
lubricin together with other macromolecules [including HA (95, 121) and fibronectin (109)] led
to friction coefficients at physiological pressures that were some one to two orders of magnitude
higher than in living joints. A recent SFB study (106) on the boundary-lubricating properties of a
model mucin is also of interest, as lubricin is a glycoprotein with its dominant linear central region
possessing the characteristic mucin structure (122). This model mucin (Figure 4b) has boundary
lubrication properties very similar to those of lubricin: At low pressures (up to 7–10 atm) it leads
to relatively low friction, μ ≈ 0.01–0.02, whereas at higher pressures the friction increases sharply
(μ ≈ 0.1–0.2), much as for lubricin (Figure 4a).

4.3. Phospholipids
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

A different concept for boundary lubrication of articular cartilage was proposed in 1984 by Hills &
Butler (27) on the basis of what they called surface-active phospholipids (SAPLs). Such SAPLs are
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

generally phosphatidylcholines (PCs), with diacyl tails and a phosphocholine head group; they form
the major component of the lipids present in SF and at the surface of articular cartilage (43), and
are present on other tissues (such as lungs). Hills and Butler’s basic idea (Figure 5a) assumes that
the PC lipids behave much as synthetic surfactant-based boundary lubricants do in classic machine
tribology (2). That is, the zwitterionic phosphocholine head groups are supposed to attach to the
underlying, negatively charged cartilage substrate, with the acyl tails extending out and rubbing
past each other as the surfaces slide under compression (Figure 5a). Such a mechanism—where
frictional dissipation occurs via shearing of relatively weak van der Waals bonds at the interface

a b
– – – –
– –
+ + – +
– – + – – + –
– – –
+– – –
+

O O O O

– + – – +

O
P
O

O
P
O

O
P
O

O
P
O
– + + – O O O O
– –
Cohesive Synovial
bonds fluid – Hydrophobic N+ N+ N+ N+
– – palmitoyl chains
– + + + – N+ N+ N+ N+
Slip
+ – – – + plane
– – –
– O O O O
– + – + – O O O O
– P P P P
+ – – – O– O– O– O–
+ + + O O O O
– –
– – Articular surface – –

Positive groups Negative groups


for adsorption on surface

Figure 5
Proposed mechanisms for cartilage lubrication by phosphatidylcholine (PC) lipids. (a) Lipid monolayers adsorb to the negatively
charged cartilage surface via their dipolar head groups, exposing the diacyl (fatty acid) chains to the interface, while frictional
dissipation on sliding occurs as the chains slide past each other at the slip plane, much as in the classic boundary lubrication of metal
surfaces by surfactants. Such a mechanism is known to lead to friction coefficients (μ) of approximately 0.05–0.1 (2). (b) In the hydration
lubrication mechanism, assemblies of PCs (whether monolayers, bilayers, or vesicles attached or complexed at the cartilage surface)
exposing their phosphocholine head groups slide past each other. Frictional dissipation takes place through shear of the hydration
layers at the slip plane. This mechanism leads to low friction coefficients (μ ≈ 0.001 or less) up to high pressures (∼100 atm or more)
(29, 100, 126–128). Panel a modified with permission from Reference 27.

246 Jahn · · Seror Klein


BE18CH10-Klein ARI 15 June 2016 8:39

between the tails—is known to result in friction coefficients of ∼0.05–0.1 (2, 123), that is, some
50–100-fold larger than in joints. At the same time, lubrication by such surfactant-like boundary
layers in aqueous environment—as in joints—is known to occur via a very different mechanism
involving hydration lubrication (124), as discussed further in the following sections. Hills (98)
also found that a mixture of HA and dipalmitoylphosphatidylcholine (DPPC) lipids, probably in
a multilamellar vesicle form, when added to saline, led to rather low friction (μ ≈ 0.004) between
sliding quartz surfaces, although at pressures of 3 atm, some 50-fold lower than maximal joint
pressures (1). The same study (98) found that adding DPPC alone to quartz surfaces in saline
resulted in only modest lubrication (μ ≈ 0.01–0.1). In a different (SFB) study, Yu et al. (114)
found that adding dioleoylphosphatidylcholine (DOPC) lipids to surfaces bearing HA in aqueous
media resulted in rather high friction coefficients (μ > 0.5). Direct measurements (125) between
sliding surfaces bearing lipid bilayers showed that DPPC bilayers (but not DOPC bilayers) could
reduce friction to low values (μ ≈ 0.002), but only up to rather low pressures (3 atm). Mirea et al.
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

(103) used similarly low pressures to examine friction at other liquid-phase bilayers. In a separate
study, Wang et al. (86) found relatively low friction (μ ≈ 0.02) between surfaces bearing layers
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

of DPPC and HA, at pressures of ∼600 atm, although this study was not able (because of large
scatter) to measure friction in the range of physiological pressures of interest (up to, say, 150 atm).
Clearly, an understanding of the boundary lubrication of articular cartilage must, as noted
above, be able to reproduce its main features. These include especially the physiologically low
friction coefficient of articular cartilage (down to μ ≈ 10−3 or lower) up to the maximal joint
pressures (100–200 atm or higher), with known components of the synovial joint, over a wide range
of shear rates and shearing times. From these facts we may conclude that few, if any, of the above
studies, which directly measured friction between surfaces bearing the main molecular candidates
that have been proposed as cartilage lubricants, alone or in combination, appear to provide a
realistic model for cartilage boundary lubrication. Indeed, HA and lubricin by themselves (or in
combination) are rather poor boundary lubricants. A key to further understanding lies instead in
the concept of hydration lubrication, a paradigm for reduction of friction in aqueous media that
has emerged over the past decade or so.

5. HYDRATION LUBRICATION
Most of the studies described above attempted to identify the nature of the lubricating boundary
layer on the articular cartilage, rather than the detailed frictional mechanism—that is, the mecha-
nism by which energy is dissipated as two such layers slide past each other in the aqueous environ-
ment of the joint. Understanding this mechanism, as we discuss further below, can lead to deep
insight into the boundary lubrication of cartilage. In 2002, on the basis of SFB measurements, it was
proposed (126) that hydrated charges in an aqueous medium may provide extremely efficient lubri-
cating elements. This is because water in the hydration shell surrounding a charge, on the one hand,
is very tenaciously held (due essentially to the interaction of the large water dipole with the en-
closed charge) and, on the other hand, may be very rapidly relaxing (due to ultrafast exchange rates
of up to 109 /s, depending on the nature of the enclosed charge, with surrounding water molecules).
These twin attributes mean that it is difficult to squeeze the hydration water out under compression
between two surfaces but that, when sheared, the hydration shells respond in a fluidlike manner
and the surfaces may slide easily past each other. This combination of supporting a large compres-
sive load together with easy sliding enables hydration layers to act as excellent lubrication elements
in aqueous surroundings (resembling molecular ball bearings, as it were), especially in biological
systems (127, 128). This mechanism, termed hydration lubrication, was extended to several bound-
ary lubrication systems. These include different hydrated ions (126, 129–132), charged polymer

www.annualreviews.org • Lubrication of Articular Cartilage 247


BE18CH10-Klein ARI 15 June 2016 8:39

brushes (133–135), boundary layers of surfactants under water (124), polyzwitterionic brushes
(136–138), and in particular, liposomes or bilayers of PC lipids, exposing phosphocholine groups
at their outer surfaces (26, 29, 97, 100, 125). The energy dissipation mechanism as hydration layers
slide past each other has also been examined in detail, and can be accounted for by the viscous losses
within the subnanometer hydration shells themselves as they are compressed and sheared (139).
In the context of biological lubrication, the most interesting finding within the hydration
lubrication paradigm is that phosphocholine groups—the head groups of PC lipids, which are
known to be highly hydrated with up to ∼15 water molecules in the primary hydration shell (136)—
constitute extremely efficient lubrication elements. The high hydration arises from the particular
zwitterionic structure of the phosphocholine groups (140). Indeed, boundary layers based on such
phosphocholine elements, whether in the form of polymer brushes (136, 137), bilayers (26), or
surface layers of PC liposomes (26, 29, 97, 100), provided remarkably low friction (μ values down
to 10−4 or less) at pressures up to 200 atm or more, over extended sliding periods, and over orders
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

of magnitude in sliding velocity. A subtle interplay between the extent of hydration, the lipid phase
nature (i.e., gel or liquid phase), and the robustness of the lipid bilayer structures, along with the
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

possibility of their self-healing, plays an important role in the comparative lubricating ability of
different PC lipids (26, 29, 97, 100, 141); this topic, however, is beyond the scope of this review. The
lubrication itself—for example, between PC bilayer assemblies or between PC vesicles adsorbed
on surfaces—takes place at the interface between the hydrated phosphocholine head groups as they
slide past each other (Figure 5b). The coefficient-of-friction values—μ values down to 10−4 or
less at pressures up to 200 atm or more—are very much consistent with the lubrication properties
of living joints. Because PC lipids are ubiquitous in joints, this finding strongly suggests that
such lipids must play an important role in cartilage lubrication. Indeed, very recent SFB studies
of friction between boundary layers of PLs extracted from human and bovine SF and cartilage
showed that PC lipids provide similarly good lubrication (μ values down to 10−3 ) up to high
pressures (up to 100 atm or more) ( J. Seror, Y. Merkher, A. Maroudas & J. Klein, manuscript
submitted).

6. SUPRAMOLECULAR SYNERGY IN CARTILAGE LUBRICATION


Given that PC lipids are capable of providing, via the hydration lubrication mechanism, the known
low friction of joints at their physiologically high pressures, a central question arises: How do such
lipids, ubiquitous in joints (43, 142), arrange themselves to form suitable boundary layers at the
cartilage surface? A recent SFB investigation by Seror et al. (30) has made progress in resolving
this question. In this study, HA molecules were attached to a mica surface in order to mimic their
presence at the surface of articular cartilage (Figure 1c), and DPPC (or hydrogenated soy PC)
lipids were introduced in the form of small unilamellar vesicles (SUVs). Many investigations have
demonstrated that PC lipids interact with HA (82, 85, 86, 94–96, 98, 143–145). This interaction has
been suggested to be either hydrophobic [between the acyl chains in the PLs and the hydrophobic
patches in the HA polymer chain (143)] or hydrophilic [between the dipolar phosphocholine head
groups and the negatively charged HA chain (30, 144, 145)], or a combination of the two (30, 145).
In their study, Seror et al. (30) used atomic force microscopy (AFM) to image the surface
structures resulting from addition of the DPPC SUVs to the HA-coated mica (Figure 6). On the
basis of these micrographs and the known interactions of HA and PC lipids (145), together with
the normal interaction profiles between the surfaces, which revealed the absolute thickness of the
surface layers, the authors proposed that the DPPC SUVs interacted with the surface-attached HA
to form complexes exposing phosphocholine groups (Figure 6b). When two surfaces bearing such
HA/PC–lipid complexes were compressed and made to slide past each other, they revealed very

248 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

b
DPPC
HA
Biotin
Avidin

a
450
1.5

1.0

0.5
400
nm

nm
0.0
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

–0.5
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

–1.0
350
–1.5

300 350 400 450


nm
Figure 6
Structure of surface-attached hyaluronic acid (HA) with phosphatidylcholine (PC) lipids. (a) An atomic force
microscopy image of an avidin-coated mica surface bearing biotinylated HA (bHA) layers to which a
dispersion of dipalmitoylphosphatidylcholine small unilamellar vesicles (DPPC SUVs) had been added.
(Inset) Image of one such vesicle on the same scale as the main image. Both micrographs indicate (30)
complexation of the DPPC SUVs with the surface-bound HA (two such linear complexes are colored red).
(b) A schematic of such HA–DPPC complexes attached to the avidin layer, showing the lipids attached to the
HA either as a monolayer ( green) via hydrophobic interactions or as a bilayer (blue) via electrostatic
interactions. In both cases, the PC head groups are exposed to the aqueous interface. Figure modified with
permission from Reference 30.

low friction coefficients (μ ≈ 10−3 ) up to pressures well over 100 atm (Figure 7a). This observation
is in line with measurements of PC vesicles [in their gel phase (29, 100)], although with somewhat
higher friction, probably due to the rougher nature of the HA–lipid surface complexes. Crucially, it
exhibits the low friction at the high pressures characteristic of articular cartilage. The contrast with
the high frictional forces (μ ≈ 0.3) between the HA layers alone prior to the introduction of the
lipids (e.g., Figure 3) is particularly striking. Moreover, the low friction was essentially constant
over some three orders of magnitude in sliding velocity (Figure 7b), indicative of boundary rather
than fluid-film lubrication, and the surface complexes were robust and lubricious over at least 1 h
of back-and-forth sliding at high pressure (Figure 7c).
These results (30) have clear implications for the mechanism of cartilage boundary lubrication.
HA, which, as described above, is ubiquitous in both cartilage and SF, is exposed at the cartilage
outer surface and complexes with PCs, which are likewise ubiquitous, to form robust boundary lay-
ers that are capable of providing the low friction (μ ≈ 10−3 ) at the physiologically high pressures (up
to 100 atm or more) of the major joints. The attachment of the HA at the cartilage surface may be
due to either physical entanglements with the near-surface network or its interactions with lubricin,
as noted elsewhere (82, 85, 94–96), or both acting together. In this scenario, therefore (Figure 8)

www.annualreviews.org • Lubrication of Articular Cartilage 249


BE18CH10-Klein ARI 15 June 2016 8:39

a μ = 3 × 10 –3
70
P ≈ 122 atm
μ = 7 × 10 –3
60 P ≈ 160 atm
–3
10
50 1.5 ×
μ=

Fs (µN)
40

30
μ = 6 × 10 –4
P ≈ 220 atm
20
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

10
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

0
0 5 10 15 20 25 30 35
Fn (mN)

30 60
b c
20
Fs (µN)

Fs (µN)
50

10 40

0 30
0.01 0.1 1.0 0 20 40 60
Vs (µm/s) Time (min)

Figure 7
Friction up to physiologically high pressures between sliding surfaces bearing boundary layers of
surface-attached hyaluronic acid (HA) to which dipalmitoylphosphatidylcholine small unilamellar vesicles
(DPPC SUVs) had been added, as indicated in Figure 6. (a) Friction forces Fs as a function of load Fn ,
measured with the surface force balance method. The shaded area includes all the Fs versus Fn profiles under
water. The black symbols represent the limiting profiles across water (highest and lowest μ values at the high
pressure indicated; the broken line indicates the mean value over all measurements). The red symbols
represent Fs versus Fn profiles across a 0.15 M salt solution. (b) Friction force Fs as function of sliding
velocity v s under pressure P ≈ 160 atm, showing little change over three orders of magnitude in v s .
(c) Friction force as a function of sliding time, showing little change following 1 h of sliding at v ≈ 0.4 μm/s
and P ≈ 60 atm. These results show that such HA–lipid complexes achieve the low friction coefficients
(down to μ ≈ 0.001) of articular cartilage up to physiologically high pressures of more than 100 atm. Figure
modified with permission from Reference 30.

(30), all three of the main synovial joint components, which have been widely conjectured to act
independently as the cartilage boundary lubricant, are observed to act together, each with a very
different role, to provide the lubrication characteristic of healthy joints. Lubricin in the outer super-
ficial zone (25), or possibly adsorbed from SF at the articular surface (122), interacts with and immo-
bilizes HA at the cartilage surface, which in turn complexes with PC lipids to provide the boundary
lubrication at the exposed phosphocholine groups via the hydration lubrication mechanism. This
picture can naturally account for the healing of the boundary layers as they wear, given that all three
of the components, as described above, are widely present both in cartilage and in SF, and are thus
readily available to replenish and heal any lubricin–HA–PC surface-attached complexes as they are

250 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

PC

Bilayer Monolayer
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

Collagen network Lubricin HA


Access provided by 196.190.84.122 on 02/18/24. For personal use only.

Figure 8
Structure of the cartilage boundary lubricant layer, proposed by Seror et al. (30), based on results shown in
Figure 6 and 7 and described in detail in Reference 30. Lubricin ( purple) in the outer superficial zone and at
the cartilage surface interacts with and immobilizes hyaluronic acid (HA) at the surface ( gray); the surface-
attached HA in turn complexes with phosphatidylcholine (PC) lipids to form monolayer ( green) or bilayer
(blue) assemblies exposing phosphocholine groups. These exposed groups slide past similar groups from the
opposing surface with low friction (μ values down to 10−3 ) up to high pressures (100 atm or more) via the
hydration lubrication mechanism (Figure 5b).

naturally removed by friction. Moreover, this scenario emphasizes the importance of all three com-
ponents working together in the lubrication and may account, for example, for the fact that lubricin
deficiency can affect joint friction (116), even if lubricin itself is not the actual lubricating molecule.

7. SUMMARY AND PERSPECTIVE


We review developments in the understanding of articular cartilage tribology, noting the various
modes of lubrication and the possible connection between high friction and OA, and focusing on
model studies of boundary lubrication. Direct measurements reveal that the three main candidates
for forming such boundary layers, HA, lubricin, and PLs, do not, by themselves, appear to provide
the low friction (μ values down to 10−3 ) at the high physiological pressures (of order 100 atm)
that are characteristic of the major joints. A recent study has shown that a surface layer of HA
complexed with PLs can provide boundary lubrication with such characteristic high-pressure/low-
friction behavior, via the recently discovered hydration lubrication mechanism. This evidence in
turn suggests a scenario where these three components (HA, lubricin, and PLs) may act together
at the articular surface, each with a different role, to provide the remarkable boundary lubrication
of the major joints.
Future work should refine the scenario for articular cartilage boundary lubrication described in
this review, and should aim to exploit the new insights revealed. In particular, better understanding
of the nature and mechanism of the process underlying the low friction at articular cartilage has
implications, through biomimetic engineering of their surfaces, for improved prosthetics (135,
146) and other biomedical devices [such as catheters or contact lenses (147)], as well as artificial
cartilage (148) and tissue engineering scaffolds where lubrication is at a premium (149). Beyond
that, such an understanding, based on the emerging scenarios described above, can have important

www.annualreviews.org • Lubrication of Articular Cartilage 251


BE18CH10-Klein ARI 15 June 2016 8:39

treatment and clinical implications for OA. As an example, such treatment might involve replacing
the often-used HA-based viscosupplement intra-articular injections (see footnote 1 on p. 243,
above) by injections introducing suitable PC/polymer-based lubricants to augment or repair the
body’s natural biolubrication mechanisms.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
J.K. thanks Tonia Vincent for useful discussions and Duncan Dowson for correspondence. We
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

thank the European Research Council (Advanced Grant HydrationLube), the McCutchen Foun-
dation, the Petroleum Research Fund of the American Chemical Society (grant 55089-ND10),
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

the Israel Science Foundation, and the Israel Science Foundation–Natural Science Foundation
of China joint research program (grant 875/13) for support of the work described in this review.
J.S.’s current affiliation is CollPlant Ltd., Weizmann Science Park, Ness-Ziona 74140, Israel.

LITERATURE CITED
1. Hodge WA, Fuan RS, Carlson KL, Burgess RG, Harris WH, Mann RW. 1986. Contact pressures in
the human hip joint measured in vivo. PNAS 83:2879–83
2. Bowden FP, Tabor D. 1973. Friction: An Introduction to Tribology. New York: Anchor/Doubleday
3. Tabor D. 1992. Friction as a dissipative process. In Fundamentals of Friction: Macroscopic and Microscopic
Processes, ed. IL Singer, HM Pollock, pp. 3–20. Dordrecht, Neth.: Kluwer
4. Dowson D. 2012. Bio-tribology. Faraday Discuss. Pap. 156:9–30
5. Forster H, Fisher J. 1996. The influence of loading time and lubricant on the friction of articular cartilage.
Proc. Inst. Mech. Eng. H 210:109–19
6. Blain EJ, Gilbert SJ, Wardale RJ, Capper SJ, Mason DJ, Duance VC. 2001. Up-regulation of matrix
metalloproteinase expression and activation following cyclical compressive loading of articular cartilage
in vitro. Arch. Biochem. Biophys. 396:49–55
7. Lee JH, Fitzgerald JB, DiMicco MA, Grodzinsky AJ. 2005. Mechanical injury of cartilage explants causes
specific time-dependent changes in chondrocyte gene expression. Arthritis Rheum. 52:2386–95
8. Vincent T, Hermansson M, Bolton M, Wait R, Saklatvala J. 2002. Basic FGF mediates an immediate
response of articular cartilage to mechanical injury. PNAS 99:8259–64
9. Vincent TL. 2013. Targeting mechanotransduction pathways in osteoarthritis: a focus on the pericellular
matrix. Curr. Opin. Pharmacol. 13:449–54
10. Watt FE, Ismail HM, Didangelos A, Peirce M, Vincent TL, et al. 2013. Src and fibroblast growth factor
2 independently regulate signaling and gene expression induced by experimental injury to intact articular
cartilage. Arthritis Rheum. 65:397–407
11. Vignon E, Balblanc JC, Mathieu P, Louisot P, Richard M. 1993. Metalloprotease activity, phospholipase
A2 activity and cytokine concentration in osteoarthritis synovial fluids. Osteoarthr. Cartil. 1:115–20
12. Goldring MB. 2012. Chondrogenesis, chondrocyte differentiation, and articular cartilage metabolism in
health and osteoarthritis. Ther. Adv. Musculoskelet. Dis. 4:269–85
13. Lindström T, Gullichsen E, Heinonen O, Grönroos J, Nevalainen T, Niinikoski J. 1997. Group II
phospholipase A2 in serum after knee surgery and intramedullary nailing of tibial shaft fracture. Injury
28:169–71
14. Burleigh A, Chanalaris A, Boruc O, Saklatvala J, Vincent TL. 2012. Fgf2 is a mechanotransducer in
destabilised joints in vivo. Int. J. Exp. Pathol. A 93:23–24

252 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

15. Burleigh A, Chanalaris A, Saklatvala J, Vincent T. 2012. Pathogenic protease expression in murine OA
is critically dependent upon mechanical joint loading. Arthritis Res. Ther. 14(Suppl. 1):P73
16. Morrell KC, Hodge WA, Krebs DE, Mann RW. 2005. Corroboration of in vivo cartilage pressures with
implications for synovial joint tribology and osteoarthritis causation. PNAS 102:14819–24
17. Klein J. 2006. Molecular mechanisms of synovial joint lubrication. Proc. Inst. Mech. Eng. H 220:691–710
18. Buckwalter JA, Mankin HJ, Grodzinky AJ. 2005. Articular cartilage and osteoarthritis. In AAOS Instruc-
tional Course Lectures, ed. VD Pellegrini Jr., 54:465–80. Rosemont, IL: Am. Acad. Orthop. Surg.
19. Buckwalter JA. 2003. Sports, joint injury, and posttraumatic osteoarthritis. J. Orthop. Sports Phys. Ther.
33:578–88
20. Silverberg JL, Barrett AR, Das M, Petersen PB, Bonassar LJ, Cohen I. 2014. Structure–function relations
and rigidity percolation in the shear properties of articular cartilage. Biophys. J. 107:1721–30
21. O’Conor CJ, Leddy HA, Benefield HC, Liedtke WB, Guilak F. 2014. TRPV4-mediated mechanotrans-
duction regulates the metabolic response of chondrocytes to dynamic loading. PNAS 111:1316–21
22. Jin M, Frank EH, Quinn TM, Hunziker EB, Grodzinsky AJ. 2001. Tissue shear deformation stimulates
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

proteoglycan and protein biosynthesis in bovine cartilage explants. Arch. Biochem. Biophys. 395:41–48
23. Lyons TJ, McClure SF, Stoddard TW, McClure J. 2006. The normal human chondro-osseus junctional
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

region: evidence for contact of uncalcified cartilage with subchondral bone and marrow spaces. BMC
Musculoskelet. Disord. 7:52
24. Flannery CR, Hughes CE, Schumacher BL, Tudor D, Aydelotte MB, et al. 1999. Articular cartilage
superficial zone protein (SZP) is homologous to megakaryocyte stimulating factor precursor and is a
multifunctional proteoglycan with potential growth-promoting, cytoprotective, and lubricating proper-
ties in cartilage metabolism. Biochem. Biophys. Res. Commun. 254:535–41
25. Jay GD, Tantravahi U, Britt DE, Barrach HJ, Cha C. 2001. Homology of lubricin and superficial zone
protein (SZP): products of megakaryocyte stimulating factor (MSF) gene expression by human synovial
fibroblasts and articular chondrocytes localized to chromosome 1q25. J. Orthop. Res. 19:677–87
26. Sorkin R, Dror Y, Kampf N, Klein J. 2014. Mechanical stability and lubrication by phosphatidylcholine
boundary layers in the vesicular and in the extended lamellar phases. Langmuir 30:5005–14
27. Hills BA, Butler BD. 1984. Surfactants identified in synovial fluid and their ability to act as boundary
lubricants. Ann. Rheum. Dis. 43:641–48
28. Hills BA. 2000. Boundary lubrication in vivo. Proc. Inst. Mech. Eng. H 214:83–94
29. Goldberg R, Schroeder A, Silbert G, Turjeman K, Barenholz Y, Klein J. 2011. Boundary lubricants with
exceptionally low friction coefficients based on 2D close-packed phosphatidylcholine liposomes. Adv.
Mater. 23:3517–21
30. Seror J, Zhu L, Goldberg R, Day AJ, Klein J. 2015. Supramolecular synergy in the boundary lubrication
of synovial joints. Nat. Commun. 6:6497
31. Athanasiou KA, Darling EM, Hu JC, DuRaine GD, Reddi HA. 2013. Articular Cartilage. Boca Raton,
FL: CRC. 439 pp.
32. Yu J, Urban JPG. 2010. The elastic network of articular cartilage: an immunohistochemical study of
elastin fibres and microfibrils. J. Anat. 216:533–41
33. Balazs E. 2009. The role of hyaluronan in the structure and function of the biomatrix of connective
tissues. Struct. Chem. 20:233–43
34. Eyre DR. 2004. Collagens and cartilage matrix homeostasis. Clin. Orthop. Relat. Res. 427:S118–22
35. Grande DA, Sgaglione NA. 2010. Regenerative medicine. Self-directed articular resurfacing: a new
paradigm? Nat. Rev. Rheumatol. 6:677–78
36. Klein TJ, Schumacher BL, Schmidt TA, Li KW, Voegtline MS, et al. 2003. Tissue engineering of
stratified articular cartilage from chondrocyte subpopulations. Osteoarthr. Cartil. 11:595–602
37. Minns RJ, Steven FS. 1977. The collagen fibril organization in human articular cartilage. J. Anat.
123:437–57
38. Aspden RM, Hukins DW. 1981. Collagen organization in articular cartilage, determined by X-ray
diffraction, and its relationship to tissue function. Proc. R. Soc. B 212:299–304
39. Urban JPG. 1994. The chondrocyte: a cell under pressure. Br. J. Rheumatol. 33:901–8

www.annualreviews.org • Lubrication of Articular Cartilage 253


BE18CH10-Klein ARI 15 June 2016 8:39

40. Chen SS, Falcovitz YH, Schneiderman R, Maroudas A, Sah RL. 2001. Depth-dependent compressive
properties of normal aged human femoral head articular cartilage: relationship to fixed charge density.
Osteoarthr. Cartil. 9:561–69
41. Mow VC, Lai WM. 1980. Recent developments in synovial joint biomechanics. SIAM Rev. 22:275–317
42. Buckwalter JA, Rosenberg LC. 1982. Electron-microscopic studies of cartilage proteoglycans—direct
evidence for the variable length of the chondroitin sulfate–rich region of proteoglycan subunit core
protein. J. Biol. Chem. 257:9830–39
43. Sarma AV, Powell GL, LaBerge M. 2001. Phospholipid composition of articular cartilage boundary
lubricant. J. Orthop. Res. 19:671–76
44. Muir H. 1995. The chondrocyte, architect of cartilage—biomechanics, structure, function and molecular
biology of cartilage matrix macromolecules. BioEssays 17:1039–48
45. Simon WH. 1970. Scale effects in animal joints. 1. Articular cartilage thickness and compressive stress.
Arthritis Rheum. 13:244–55
46. Pagel W. 1982. Paracelsus: An Introduction to Philosophical Medicine in the Era of Renaissance. New York:
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

S. Karger
47. Havers C. 1691. Osteologia Nova. London: Smith
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

48. Ateshian GA. 2009. The role of interstitial fluid pressurization in articular cartilage lubrication. J. Biomech.
42:1163–76
49. Dowson D, Jin ZM. 1987. An analysis of micro-elastohydrodynamic lubrication in synovial joints consid-
ering cyclic loading and entraining velocities. In Proceedings of the 23rd Leeds–Lyon Symposium on Tribology,
ed. V Štěpina, V Veselý, pp. 375–86. Amsterdam: Elsevier
50. Jones ES. 1936. Joint lubrication. Lancet 227:1043–45
51. Maroudas A. 1968. Physicochemical properties of cartilage in the light of ion exchange theory. Biophys.
J. 8:575–95
52. McCutchen CW. 1959. Mechanism of animal joints: sponge-hydrostatic and weeping bearings. Nature
184:1284–85
53. Charnley J. 1960. The lubrication of animal joints in relation to surgical reconstruction by arthroplasty.
Ann. Rheum. Dis. 19:10–19
54. Jay GD. 1992. Characterization of a bovine synovial-fluid lubricating factor. 1. Chemical, surface-activity
and lubricating properties. Connect. Tissue Res. 28:71–88
55. Schmidt TA, Gastelum NS, Nguyen QT, Schumacher BL, Sah RL. 2007. Boundary lubrication of
articular cartilage—role of synovial fluid constituents. Arthritis Rheum. 56:882–91
56. Zappone B, Ruths M, Greene GW, Jay GD, Israelachvili JN. 2007. Adsorption, lubrication, and wear
of lubricin on model surfaces: polymer brush–like behavior of a glycoprotein. Biophys. J. 92:1693–708
57. Singh A, Corvelli M, Unterman SA, Wepasnick KA, McDonnell P, Elisseeff JH. 2014. Enhanced lu-
brication on tissue and biomaterial surfaces through peptide-mediated binding of hyaluronic acid. Nat.
Mater. 13:988–95
58. Reynolds O. 1886. On the theory of lubrication and its application to Mr. Beauchamp Tower’s experi-
ments, including an experimental determination of the viscosity of olive oil. Philos. Trans. R. Soc. Lond.
177:157–234
59. MacConaill MA. 1932. The function of intra-articular fibrocartilages, with special reference to the knee
and inferior radio-ulnar joints. J. Anat. 66:210–27
60. Jones ES. 1934. Joint lubrication. Lancet 1:1426–27
61. Dowson D. 1967. Modes of lubrication in human joints. Proc. Inst. Mech. Eng. J 181:45–54
62. Unsworth A. 1991. Tribology of human and artifical joints. Proc. Inst. Mech. Eng. H 205:163–72
63. Tanner RI. 1966. An alternative mechanism for the lubrication of synovial joints. Phys. Med. Biol. 11:119–
27
64. Fein RS. 1966. Are synovial joints squeeze-film lubricated? Proc. Inst. Mech. Eng. J 181:125–28
65. Higginson GR, Unsworth A. 1981. The lubrication of natural joints. In Tribology of Natural and Artificial
Joints, ed. JH Dumbleton, pp. 47–72. Amsterdam: North-Holland
66. Walker PS, Dowson D, Longfield MD, Wright V. 1968. “Boosted lubrication” in synovial joints by fluid
entrapment and enrichment. Ann. Rheum. Dis. 27:512–20

254 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

67. Maroudas A. 1975. Biophysical chemistry of cartilaginous tissues with special reference to solute and
fluid transport. Biorheology 12:233–48
68. McCutchen CW. 1962. The frictional properties of animal joints. Wear 5:1–17
69. McCutchen CW. 1964. Lubrication of joints. BMJ 1:384–85
70. Soltz MA, Basalo IM, Ateshian GA. 2003. Hydrostatic pressurization and depletion of trapped lubricant
pool during creep contact of a rippled indenter against a biphasic articular cartilage layer. J. Biomech.
Eng. 125:585–93
71. McCutchen CW. 2004. Comment on “Hydrostatic pressurization and depletion of trapped lubricant
pool during creep contact of a rippled indenter against a biphasic articular cartilage layer”. J. Biomech.
Eng. 126:536–37
72. Murakami T, Higaki H, Sawae Y, Ohtsuki N, Moriyama S, Nakanishi Y. 1998. Adaptive multimode
lubrication in natural synovial joints and artificial joints. Proc. Inst. Mech. Eng. H 212:23–35
73. Murakami T, Nakashima K, Sawae Y, Sakai N, Hosoda N. 2009. Roles of adsorbed film and gel layer in
hydration lubrication for articular cartilage. Proc. Inst. Mech. Eng. J 223:287–95
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

74. Jin ZM, Dowson D. 2005. Elastohydrodynamic lubrication in biological systems. Proc. Inst. Mech. Eng.
J 219:367–80
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

75. O’Kelly J, Unsworth A, Dowson D, Hall DA, Wright V. 1978. A study of the role of synovial fluid and
its constituents in the friction and lubrication of human hip joints. Eng. Med. 7:73–83
76. Afoke NY, Byers PD, Hutton WC. 1987. Contact pressures in the human hip joint. J. Bone Joint Surg.
Br. 69:536–41
77. Ateshian GA. 1997. A theoretical formulation for boundary friction in articular cartilage. J. Biomech.
Eng. 119:81–86
78. Sotres J, Arenebrant T. 2013. Experimental investigations of biological lubrication at the nanoscale: the
cases of synovial joints and the oral cavity. Lubricants 1:102–31
79. Laurent TC, Laurent UBG, Fraser JRE. 1995. Functions of hyaluronan. Ann. Rheum. Dis. 54:429–32
80. Ogston AG, Stanier JE. 1953. The physiological function of hyaluronic acid in synovial fluid; viscous,
elastic and lubricant properties. J. Physiol. 119:244–52
81. Maroudas A. 1969. Studies on the formation of hyaluronic acid films. In Lubrication and Wear in Joints,
ed. V Wright, pp. 124–30. London: Sector
82. Chang DP, Abu-Lail NI, Coles JM, Guilak F, Jay GD, Zauscher S. 2009. Friction force microscopy of
lubricin and hyaluronic acid between hydrophobic and hydrophilic surfaces. Soft Matter 5:3438–45
83. Swann DA, Radin EL, Nazimiec M, Weisser PA, Curran N, Lewinnek G. 1974. Role of hyaluronic acid
in joint lubrication. Ann. Rheum. Dis. 33:318–26
84. Benz M, Chen NH, Israelachvili J. 2004. Lubrication and wear properties of grafted polyelectrolytes,
hyaluronan and hylan, measured in the surface forces apparatus. J. Biomed. Mater. Res. A 71:6–15
85. Jay GD, Lane BP, Sokoloff L. 1992. Characterization of a bovine synovial fluid lubricating factor. III.
The interaction with hyaluronic acid. Connect. Tissue Res. 28:245–55
86. Wang M, Liu C, Thormann E, Dedinaite A. 2013. Hyaluronan and phospholipid association in biolu-
brication. Biomacromolecules 14:4198–206
87. Han L, Dean D, Ortiz C, Grodzinsky AJ. 2007. Lateral nanomechanics of cartilage aggrecan macro-
molecules. Biophys. J. 92:1384–98
88. Ng L, Grodzinsky AJ, Patwari P, Sandy J, Plaas A, Ortiz C. 2003. Individual cartilage aggrecan macro-
molecules and their constituent glycosaminoglycans visualized via atomic force microscopy. J. Struct.
Biol. 143:242–57
89. Seror J, Merkher Y, Kampf N, Collinson L, Day AJ, et al. 2011. Articular cartilage proteoglycans as
boundary lubricants: structure and frictional interaction of surface-attached hyaluronan and hyaluronan–
aggrecan complexes. Biomacromolecules 12:3432–43
90. Seror J, Merkher Y, Kampf N, Collinson L, Day AJ, et al. 2012. Normal and shear interactions between
hyaluronan–aggrecan complexes mimicking possible boundary lubricants in articular cartilage in synovial
joints. Biomacromolecules 13:3823–32
91. Radin EL, Swann DA, Weisser PA. 1970. Separation of a hyaluronate-free lubricating fraction from
synovial fluid. Nature 228:377–78

www.annualreviews.org • Lubrication of Articular Cartilage 255


BE18CH10-Klein ARI 15 June 2016 8:39

92. Swann DA, Slayter HS, Silver FH. 1981. The molecular structure of lubricating glycoprotein-I, the
boundary lubricant for articular cartilage. J. Biol. Chem. 256:5921–25
93. Flannery CR, Zollner R, Corcoran C, Jones AR, Root A, et al. 2009. Prevention of cartilage degeneration
in a rat model of osteoarthritis by intraarticular treatment with recombinant lubricin. Arthritis Rheum.
60:840–47
94. Jay GD, Torres JR, Warman ML, Laderer MC, Breuer KS. 2007. The role of lubricin in the mechanical
behavior of synovial fluid. PNAS 104:6194–99
95. Das S, Banquy X, Zappone B, Greene GW, Jay GD, Israelachvili JN. 2013. Synergistic interac-
tions between grafted hyaluronic acid and lubricin provide enhanced wear protection and lubrication.
Biomacromolecules 14:1669–77
96. Chang DP, Abu-Lail NI, Guilak F, Jay GD, Zauscher S. 2008. Conformational mechanics, adsorption,
and normal force interactions of lubricin and hyaluronic acid on model surfaces. Langmuir 24:1183–93
97. Goldberg R, Schroeder A, Barenholz Y, Klein J. 2011. Interactions between adsorbed hydrogenated soy
phosphatidylcholine (HSPC) vesicles at physiologically high pressures and salt concentrations. Biophys.
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

J. 100:2403–11
98. Hills BA. 1989. Oligolamellar lubrication of joints by surface-active phospholipid. J. Rheumatol. 16:82–91
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

99. Hills BA. 2002. Surface-active phospholipid: a Pandora’s box of clinical applications. Part I. The lung
and air spaces. Intern. Med. J. 32:170–78
100. Sorkin R, Kampf N, Dror Y, Shimoni E, Klein J. 2013. Origins of extreme boundary lubrication by
phosphatidylcholine liposomes. Biomaterials 34:5465–75
101. Sivan S, Schroeder A, Verberne G, Merkher Y, Diminsky D, et al. 2010. Liposomes act as effective
biolubricants for friction reduction in human synovial joints. Langmuir 26:1107–16
102. Williams PF 3rd, Powell GL, LaBerge M. 1993. Sliding friction analysis of phosphatidylcholine as a
boundary lubricant for articular cartilage. Proc. Inst. Mech. Eng. H 207:59–66
103. Mirea DA, Trunfio-Sfarghiu AM, Matei CI, Munteanu B, Piednoir A, et al. 2013. Role of the biomolec-
ular interactions in the structure and tribological properties of synovial fluid. Tribol. Int. 59:302–11
104. Crockett R. 2009. Boundary lubrication in natural articular joints. Tribol. Lett. 35:77–84
105. Crockett R, Grubelnik A, Roos S, Dora C, Born W, Troxler H. 2007. Biochemical composition of the
superficial layer of articular cartilage. J. Biomed. Mater. Res. A 82:958–64
106. Harvey NM, Yakubov GE, Stokes JR, Klein J. 2011. Normal and shear forces between surfaces bearing
porcine gastric mucin, a high-molecular-weight glycoprotein. Biomacromolecules 12:1041–50
107. Murakami T, Yarimitsu S, Nakashima K, Sawae Y, Sakai N. 2013. Influence of synovia constituents on
tribological behaviors of articular cartilage. Friction 1:150–62
108. Lee S, Spencer ND. 2008. Materials science—sweet, hairy, soft, and slippery. Science 319:575–76
109. Andresen Eguiluz RC, Cook SG, Brown CN, Wu F, Pacifici NJ, et al. 2015. Fibronectin mediates
enhanced wear protection of lubricin during shear. Biomacromolecules 16:2884–94
110. Maroudas A. 1967. Hyaluronic acid films. Proc. Inst. Mech. Eng. 181:122
111. Jay GD, Harris DA, Cha CJ. 2001. Boundary lubrication by lubricin is mediated by O-linked β(1–3)Gal-
GalNAc oligosaccharides. Glycoconj. J. 18:807–15
112. Krause WE, Bellomo EG, Colby RH. 2001. Rheology of sodium hyaluronate under physiological con-
ditions. Biomacromolecules 2:65–69
113. Tadmor R, Chen NH, Israelachvili JN. 2002. Thin film rheology and lubricity of hyaluronic acid solu-
tions at a normal physiological concentration. J. Biomed. Mater. Res. 61:514–23
114. Yu J, Banquy X, Greene GW, Lowrey DD, Israelachvili JN. 2012. The boundary lubrication of chemi-
cally grafted and cross-linked hyaluronic acid in phosphate buffered saline and lipid solutions measured
by the surface forces apparatus. Langmuir 28:2244–50
115. Linn FC, Radin EL. 1968. Lubrication of animal joints. 3. Effect of certain chemical alterations of
cartilage and lubricant. Arthritis Rheum. 11:674–78
116. Jay GD, Torres JR, Rhee DK, Helminen HJ, Hytinnen MM, et al. 2007. Association between friction
and wear in diarthrodial joints lacking lubricin. Arthritis Rheum. 56:3662–69
117. Marcelino J, Carpten JD, Suwairi WM, Gutierrez OM, Schwartz S, et al. 1999. CACP, encoding a
secreted proteoglycan, is mutated in camptodactyly–arthropathy–coxa vara–pericarditis syndrome. Nat.
Genet. 23:319–22

256 Jahn · ·
Seror Klein
BE18CH10-Klein ARI 15 June 2016 8:39

118. Abubacker S, Dorosz SG, Ponjevic D, Jay GD, Matyas J, Schmidt TA. 2015. Full-length recombinant
human proteoglycan 4 interacts with hyaluronan to provide cartilage boundary lubrication. Ann. Biomed.
Eng. In press. doi:10.1007/s10439-015-1390-8
119. Hills BA, Jay GD. 2002. Identity of the joint lubricant. J. Rheumatol. 29:200–1
120. Zappone B, Greene GW, Oroudjev E, Jay GD, Israelachvili JN. 2007. Molecular aspects of boundary
lubrication by human lubricin: effect of disulfide bonds and enzymatic digestion. Langmuir 24:1495–508
121. Greene GW, Banquy X, Lee DW, Lowrey DD, Yu J, Israelachvili JN. 2011. Adaptive mechanically
controlled lubrication mechanism found in articular joints. PNAS 108:5255–59
122. Jones ARC, Gleghorn JP, Hughes CE, Fitz LJ, Zollner R, et al. 2007. Binding and localization of
recombinant lubricin to articular cartilage surfaces. J. Orthop. Res. 25:283–92
123. Briscoe BJ, Evans DCB. 1982. The shear properties of Langmuir–Blodgett layers. Proc. R. Soc. A 380:389–
407
124. Briscoe WH, Titmuss S, Tiberg F, Thomas RK, McGillivray DJ, Klein J. 2006. Boundary lubrication
under water. Nature 444:191–94
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

125. Trunfio-Sfarghiu AM, Berthier Y, Meurisse MH, Rieu JP. 2008. Role of nanomechanical properties in
the tribological performance of phospholipid biomimetic surfaces. Langmuir 24:8765–71
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

126. Raviv U, Klein J. 2002. Fluidity of bound hydration layers. Science 297:1540–43
127. Jahn S, Klein J. 2015. Hydration lubrication: the macromolecular domain. Macromolecules 48:5059–75
128. Klein J. 2013. Hydration lubrication. Friction 1:1–23
129. Chai L, Goldberg R, Kampf N, Klein J. 2008. Selective adsorption of poly(ethylene oxide) onto a charged
surface mediated by alkali metal ions. Langmuir 24:1570–76
130. Gaisinskaya A, Ma LR, Silbert G, Sorkin R, Tairy O, et al. 2012. Hydration lubrication: exploring a new
paradigm. Faraday Discuss. Pap. 156:217–33
131. Perkin S, Goldberg R, Chai L, Kampf N, Klein J. 2009. Dynamics of confined hydration layers. Faraday
Discuss. Pap. 141:399–413
132. Donose BC, Vakarelski IU, Higashitani K. 2005. Silica surface lubrication by hydrated cations adsorption
from electrolyte solutions. Langmuir 21:1834–39
133. Kampf N, Gohy JF, Jerome R, Klein J. 2005. Normal and shear forces between a polyelectrolyte brush
and a solid surface. J. Polym. Sci. B 43:193–204
134. Raviv U, Giasson S, Gohy J-F, Jerome R, Klein J. 2008. Normal and frictional forces between surfaces
bearing polyelectrolyte brushes. Langmuir 24:8678–87
135. Raviv U, Giasson S, Kampf N, Gohy JF, Jerome R, Klein J. 2003. Lubrication by charged polymers.
Nature 425:163–65
136. Chen M, Briscoe WH, Armes SP, Klein J. 2009. Lubrication at physiological pressures by polyzwitteri-
onic brushes. Science 323:1698–701
137. Tairy O, Kampf N, Driver MJ, Armes SP, Klein J. 2015. Dense, highly hydrated polymer brushes via
modified atom-transfer-radical-polymerization: structure, surface interactions, and frictional dissipation.
Macromolecules 48:140–51
138. Banquy X, Burdyńska J, Lee DW, Matyjaszewski K, Israelachvili JN. 2014. Bioinspired bottle-brush
polymer exhibits low friction and amontons-like behavior. J. Am. Chem. Soc. 136:6199–202
139. Ma L, Gaisinskaya-Kipnis A, Kampf N, Klein J. 2015. Origins of hydration lubrication. Nat. Commun.
6:6060
140. Foglia F, Lawrence MJ, Lorenz CD, McLain SE. 2010. On the hydration of the phosphocholine head-
group in aqueous solution. J. Chem. Phys. 133:145103
141. Seror J, Sorkin R, Klein J. 2014. Boundary lubrication by macromolecular layers and its relevance to
synovial joints. Polym. Adv. Technol. 25:468–77
142. Kosinska MK, Liebisch G, Lochnit G, Wilhelm J, Klein H, et al. 2013. A lipidomic study of phospholipid
classes and species in human synovial fluid. Arthritis Rheum. 65:2323–33
143. Ghosh P, Hutadilok N, Adam N, Lentini A. 1994. Interactions of hyaluronan (hyaluronic-acid) with
phospholipids as determined by gel-permeation chromatography, multi-angle laser-light-scattering pho-
tometry and 1 H-NMR spectroscopy. Int. J. Biol. Macromol. 16:237–44
144. Nitzan DW, Nitzan U, Dan P, Yedgar S. 2001. The role of hyaluronic acid in protecting surface-active
phospholipids from lysis by exogenous phospholipase A2. Rheumatology 40:336–40

www.annualreviews.org • Lubrication of Articular Cartilage 257


BE18CH10-Klein ARI 15 June 2016 8:39

145. Pasquali-Ronchetti I, Quaglino D, Mori G, Bacchelli B. 1997. Hyaluronan–phospholipid interactions.


J. Struct. Biol. 120:1–10
146. Moro T, Takatori Y, Ishihara K, Konno T, Takigawa Y, et al. 2004. Surface grafting of artificial joints
with a biocompatible polymer for preventing periprosthetic osteolysis. Nat. Mater. 3:829–36
147. Nichols JJ, Sinnott LT. 2006. Tear film, contact lens, and patient-related factors associated with contact
lens–related dry eye. Investig. Ophthalmol. Vis. Sci. 47:1319–28
148. Murakami T, Sawae Y, Nakashima K, Fisher J. 2000. Tribological behaviour of artificial cartilage in thin
film lubrication. In Tribology Series, vol. 38: Thinning Films and Tribological Interfaces, ed. D Dowson, M
Priest, CM Taylor, P Ehret, THC Childs, et al., pp. 317–27. Amsterdam: Elsevier
149. Klein J. 2009. Repair or replacement—a joint perspective. Science 323:47–48
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

258 Jahn · ·
Seror Klein
BE18-FrontMatter ARI 24 May 2016 19:41

Annual Review of
Biomedical
Engineering
Contents Volume 18, 2016

Tissue Patterning: Translating Design Principles from In Vivo


Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

to In Vitro
Alison P. McGuigan and Sahar Javaherian p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

Rational Design of Targeted Next-Generation Carriers for Drug


and Vaccine Delivery
Balaji Narasimhan, Jonathan T. Goodman, and Julia E. Vela Ramirez p p p p p p p p p p p p p p p p p p25
Drugging Membrane Protein Interactions
Hang Yin and Aaron D. Flynn p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p51
Lensless Imaging and Sensing
Aydogan Ozcan and Euan McLeod p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p77
The Virtual Physiological Human: Ten Years After
Marco Viceconti and Peter Hunter p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 103
The Lymphatic System in Disease Processes and Cancer Progression
Timothy P. Padera, Eelco F.J. Meijer, and Lance L. Munn p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 125
Engineered Models of Confined Cell Migration
Colin D. Paul, Wei-Chien Hung, Denis Wirtz, and Konstantinos Konstantopoulos p p p p 159
Immune Tolerance for Autoimmune Disease and Cell Transplantation
Xunrong Luo, Stephen D. Miller, and Lonnie D. Shea p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 181
Implications of Lymphatic Transport to Lymph Nodes in Immunity
and Immunotherapy
Susan N. Thomas, Nathan A. Rohner, and Erin E. Edwards p p p p p p p p p p p p p p p p p p p p p p p p p p p p 207
Lubrication of Articular Cartilage
Sabrina Jahn, Jasmine Seror, and Jacob Klein p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 235
Innovative Tools and Technology for Analysis of Single Cells
and Cell–Cell Interaction
Tania Konry, Saheli Sarkar, Pooja Sabhachandani, and Noa Cohen p p p p p p p p p p p p p p p p p p p p 259

v
BE18-FrontMatter ARI 24 May 2016 19:41

Microfluidics for High-Throughput Quantitative Studies of Early


Development
Thomas J. Levario, Bomyi Lim, Stanislav Y. Shvartsman, and Hang Lu p p p p p p p p p p p p p p 285
Design of Catalytic Peptides and Proteins Through Rational
and Combinatorial Approaches
Yoshiaki Maeda, Olga V. Makhlynets, Hiroshi Matsui, and Ivan V. Korendovych p p p p p 311
Electrical Chips for Biological Point-of-Care Detection
Bobby Reddy, Eric Salm, and Rashid Bashir p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 329
Optical-Based Analysis of Soft Tissue Structures
Will Goth, John Lesicko, Michael S. Sacks, and James W. Tunnell p p p p p p p p p p p p p p p p p p p p p 357
Annu. Rev. Biomed. Eng. 2016.18:235-258. Downloaded from www.annualreviews.org

Emerging Themes in Image Informatics and Molecular Analysis for


Digital Pathology
Access provided by 196.190.84.122 on 02/18/24. For personal use only.

Rohit Bhargava and Anant Madabhushi p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 387

Indexes

Cumulative Index of Contributing Authors, Volumes 9–18 p p p p p p p p p p p p p p p p p p p p p p p p p p p p 413


Cumulative Index of Article Titles, Volumes 9–18 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 417

Errata

An online log of corrections to Annual Review of Biomedical Engineering articles may be


found at http://www.annualreviews.org/errata/bioeng

vi Contents

You might also like