You are on page 1of 16

Friction ISSN 2223-7690

https://doi.org/10.1007/s40544-023-0822-y CN 10-1237/TH
RESEARCH ARTICLE

Experimental study on boundary lubricity of superficial area of


articular cartilage and synovial fluid
Wenxiao LI1,*, Takehiro MORITA1,2, Yoshinori SAWAE1,2,3
1
Departure of Mechanical Engineering, Kyushu University, Fukuoka 8190395, Japan
2
International Institute for Carton-Neutral Energy Research, Kyushu University, Fukuoka 8190395, Japan
3
Advanced Research Center for Biomechanics, Kyushu University, Fukuoka 8190395, Japan
Received: 08 March 2023 / Revised: 09 August 2023 / Accepted: 30 August 2023
© The author(s) 2023.

Abstract: The boundary lubrication mechanism at the articulating surface of natural synovial joints has been
the subject of much discussion in tribology. In this study, to elucidate the lubricating function of the superficial
area of articular cartilage and synovial fluid (SF), cartilage specimens were processed with four different
treatments: gentle and severe washing with detergent, incubation in NaCl solution, and trypsin digestion to
selectively remove certain constituents from the cartilage surface. Subsequently, the frictional characteristics
were examined in phosphate-buffered saline (PBS) and SF against glass. Angularly reciprocating sliding tests
with a spherical glass probe and square articular cartilage specimens were performed at low contact loads in
the mN range to extract the frictional behavior in the superficial area of the cartilage specimens. Meanwhile,
the cartilage surface was observed to confirm the effects of treatments on the morphology of the cartilage
surface using a fluorescence microscope and water-immersion methods. The coefficient of friction (COF) of the
prepared cartilage specimens was varied from 0.05 to over 0.3 in PBS. However, a certain group of cartilage
specimens exhibited a low COF of less than 0.1 with limited variation. For the low COF group of specimens, all
four treatments increased the COF in PBS to different extents, and fluorescence microscopy revealed that the
integrity of the cartilage surface was deteriorated by treatments. This means that the intact cartilage surface had
lubricating constituents to maintain low friction, and the removal of such constituents resulted in the loss of the
intrinsic boundary lubricity of the cartilage surface. The variation in the COF of the cartilage specimens was
suppressed in SF because it had a clear boundary lubrication effect on the cartilage surface. The lubricating
effect of SF could be confirmed even after degenerative treatment.

Keywords: articular cartilage; synovial fluid; the uppermost superficial layer; boundary lubrication

1 Introduction covered with an acellular boundary layer, called


the uppermost superficial surface layer or lamina
The articular surfaces of natural synovial joints splendens [4, 5]. The lamina splendens is a continuous
are covered with compliant and highly hydrated biomembrane structure composed of a phospholipid
articular cartilage and lubricated with synovial fluid bilayer, proteins, glycoproteins, and proteoglycan
(SF) [1]. Since its structure and composition changes [4, 6, 7]. SF secreted from the synovial membrane
continuously from the bottom to the surface, articular contains many biological macromolecules, including
cartilage can be divided into four zones: the calcified hyaluronic acid (HA), phospholipids, albumin,
cartilage zone, the deep zone, the middle zone, and globulin, lubricin, and glycosaminoglycans [8–11].
the superficial zone [2, 3]. The superficial zone is The extremely smooth nature of synovial joint

* Corresponding author: Wenxiao LI, E-mail: li.wenxiao.345@s.kyushu-u.ac.jp


2 Friction

movements, with the Ra of the cartilage measured in After absorbing water, proteoglycan aggregates bind
the range of 1 μm, has attracted significant interest in with hyaluronic acid, which in turn adheres to the
tribology [12]. In the past several decades, numerous collagen network located on the surface of the
studies have been conducted to unravel the detailed cartilage. The contact load between natural synovial
mechanisms of synovial joint lubrication, such as joint surfaces is supported to some extent by water
boundary [13, 14], multimode [15], elastohydrodynamic molecules, which involves osmotic pressure. Lin et al.
[16, 17], hydrodynamic [18], weeping [19], boosted [27] introduced a complex paradigm of hydration
[20], and biphasic lubrication mechanisms [21, 22]. lubrication, where phospholipid tails adhere to one
In 1959, McCutchen [19] reported that the interstitial another while their hydrated phosphocholine heads
fluid contained in porous articular cartilage flows create a low-shear sliding interface on the cartilage
through the pores, and the pressurized fluid supports surface through hydration lubrication.
most of the applied load; this is called weeping Some studies have focused on the individual
lubrication. Ten years later, the lubrication effect influence of SF components and investigated the
of this model was demonstrated by Mow et al. [23] constituents that are responsible for the ultra-low
through a complicated mathematical analysis. Walker’s friction of articular cartilage. Murakami et al. [28],
boosted lubrication [20] hypothesis states that in using a reciprocating tester, showed the COF between
contrast to weeping lubrication, small molecular a lower reciprocating flat glass plate and a stationary
weight solutes, such as water, in the SF flow into the upper cartilage specimen under macroscale load
porous cartilage, and large macromolecular weight conditions and clarified the effect of synovial
solutes, such as HA, are retained on the cartilage constituents on the lubrication mechanism. Bonnevie
surface as boundary lubricants. Subsequently, the et al. [29], using a custom-built tribometer, reported
biphasic lubrication mechanism was elucidated. that the COF between cartilage and a polished glass
Ateshian et al. [21, 22, 24] proposed and validated a plate was significantly reduced by the addition of
boundary friction model based on biphasic lubrication lubricin to a HA solution. However, macroscopic
to predict the friction behavior of articular cartilage, friction involves the interaction of many microscale
including its time-, velocity-, and load-dependent factors over a large contact area. Therefore, the
responses. Mixed and multimode lubrication [4, 15] influence of some small factors on the friction behavior
have also been developed. will be ignored, such as the local damage of the
Owing to the low relative motion speed and large initial cartilage surface and molecular structure. It is
applied load between the articulating surfaces, some difficult to understand the effects of unique factors
scientists believe that the lubrication mechanism on friction behavior. These friction measurement
between cartilage surfaces is boundary lubrication, as results contain interstitial fluid effects owing to
supported by experimental evidence. In 1959, Charnley the biphasic nature of the cartilage. Therefore, the
[13] designed a pendulum machine with a cadaver effect of SF constituents cannot be isolated from these
joint as the fulcrum and found that the friction results.
resistance was independent of the speed. In 1968, Seror et al. [30] used atomic force microscopy to
Linn et al. [25] demonstrated that adding trypsin to demonstrate that the addition of lipids to HA-coated
SF significantly increased the coefficient of friction mica resulted in a low COF (0.001) at pressure >100 atm.
(COF), whereas the addition of hyaluronidase had a Chang et al. [31] found that friction between
limited effect on COF. hydrophobic surfaces significantly reduced because
Over the past few decades, several studies have of the addition of lubricin, whereas friction between
focused on the boundary lubrication mechanism of hydrophilic surfaces slightly increased. The friction
the molecular boundary layer coating the superficial between the model surfaces with a mixture of lubricin
area of the cartilage surface. Ishikawa et al. [26] and HA solutions did not change. It is believed that
showed that the proteoglycan gel layer on the cartilage the microscale contact force is too small to pressurize
surface contributes to low-friction characteristics. the cartilage. Therefore, microscale contact provides

| https://mc03.manuscriptcentral.com/friction
Friction 3

a unique perspective for studying cartilage boundary was harvested from the proximal end of the middle
friction behavior without fluid pressurization. phalanx and shaped into an 8 × 8 mm square plate
In boundary lubrication, there is a growing realization specimen with a thickness of 2 mm, consisting of
that the synergistic effects of the constituents of the roughly 1-mm thick cartilage tissue and underlying
superficial area of the cartilage surface and SF play a subchondral bone. All the specimens were carefully
predominant role in the COF. Recently, Jahn et al. [32] prepared to prevent damage to the cartilage surface.
proposed the structure of the cartilage boundary The surfaces of the prepared plate specimens were
lubricant layer to explain the effects of the constituents observed by the naked eye, and those without
of the cartilage surface and SF on the friction behavior. fissures or chondral defects were selected and used
In this mechanism, the hydrated lubrication product for sliding tests.
near the phosphocholine head of the phospholipid In this study, four different degenerative treatments
absorbs the water molecule and supports the contact for articular cartilage were selected based on a literature
load. This indicates that lubricin plays a key role in survey and were employed to remove specific
the boundary lubrication of the articular cartilage lubricating elements from the cartilage surface. A
and in linking HA to the cartilage. However, this solution of 10% (v/v) detergent (Triton-X-100) was used
is simply an experiment on a mica surface, whose to remove adsorbed lipids and protein molecules
physical characteristics are different from those of from the cartilage surface [15, 33, 34]. The first group
natural articular cartilage. of cartilage specimens, called the gentle-wash group,
As mentioned above, the boundary lubrication
was gently washed with detergent solution for 30 min
mechanism of natural synovial joints, particularly the
using a tube mixer. The second group of cartilage
lubricating functions existing in the superficial area
specimens was also washed with the detergent solution
of the cartilage surface and SF, is still subject to active
for 30 min; however, in this case, it was washed
debate. Because the purpose of this study was to
intensively using an ultrasonic cleaner and was referred
experimentally identify the boundary lubricity of
to as the severe wash group. These cartilage specimens
molecular layers on the articular cartilage surface, the
were subsequently rinsed with phosphate-buffered
friction behavior of the surface was evaluated at
saline (PBS) three times and stored in PBS at room
extremely low contact loads to minimize contributions
temperature until the sliding experiment. The third
from the bulk of the cartilage tissue. Four different
group, called the NaCl group, was immersed in a 1.5 M
degenerative treatments were used to remove specific
NaCl solution for 20 min, rinsed three times, and
lubricating elements from the cartilage surface and
elucidate their lubricating function. We also examined re-equilibrated in PBS for 1 h to remove lubricin from
the friction behavior of treated and untreated cartilage the cartilage surface [35–37]. Trypsin digestion was
specimens under SF lubrication to elucidate the used to remove proteoglycans and lubricin from
synergistic effect of the lubricating elements existing the superficial area of the cartilage specimen [38, 39].
in the superficial area of the articular cartilage and SF For the last group, bovine cartilage specimens were
constituents. incubated in 30 mL of the reaction mixture (0.2 M
Tris-HCl buffer, pH = 7.3; 500 units mL–1 trypsin (Sigma))
at 37 °C for 20 min. Subsequently, 1 mg trypsin
2 Materials and methods inhibitor (Sigma) was added to the mixture. The final
group of cartilage specimens was called the trypsin
2.1 Cartilage specimen preparation and degenerative
group.
treatments
Fresh SF was extracted from healthy bovine
Cartilage specimens were prepared from fresh bovine metacarpal–phalangeal joints. The SF obtained from
metacarpal-phalangeal joints, from animals that were several joints were mixed to minimize sample-to-sample
approximately 20 months, obtained from a local variability. If potential blood contamination was
abattoir within 24 h of sacrifice. Using a micro cutter identified, a certain amount of SF was excluded from
(MC-201, MARUTO, Japan), the osteochondral block the sliding test.

www.Springer.com/journal/40544 | Friction
4 Friction

2.2 Friction tests speed. After finishing the friction measurement at


5 mN, the same procedure was repeated at increased
Rotational reciprocating sliding tests with an angular contact loads of 20 and 100 mN.
amplitude of 30° were performed using a ball-on- In the first series of experiments, the frictional
plate-type tribometer (Anton Paar NTR3, Austria) behavior of the untreated cartilage specimens was
with a smooth glass probe (Ra ≈ 0.045 μm) of 2 mm evaluated in this contact. Each experiment consisted
diameter to evaluate the frictional characteristics of three steps: the surface friction of the fresh specimen
of the cartilage–glass contact. Photographs of the was evaluated immediately after preparation in PBS
tribometer are shown in Fig. 1. The tribometer used (Step 1). Next, the specimen was kept in a liquid bath
in this study was developed to precisely investigate filled with PBS for 1.5 h, and its surface friction in
the surface friction of compliant materials under PBS was examined again to confirm the effect of
low-contact loads. The cartilage specimen was placed stabilization in PBS (Step 2). The COF obtained in
in a liquid bath and was fully immersed in the Step 2 was used as the control data in the subsequent
test lubricant. The glass probe was connected to second series of experiments. For selected specimens,
the cantilevers for force measurement through a PBS in the liquid bath was replaced with SF after
stainless-steel stem and mounted on a piezo-actuator Step 2, and the cartilage specimen was equilibrated
integrated in the tribometer. A dual quad-beam in SF for 1.5 h. Then, friction measurements were
cantilever with two separate high-resolution capacitive performed in SF (Step 3).
sensors was used to measure the normal and frictional The second series of experiments examined the
forces exerted between the glass probe and the effects of degenerative treatments on the surface
cartilage surface. A piezo-actuator reads the normal friction of articular cartilage. The specimen was first
force measured by the sensor in real time and instantly stabilized in PBS for 1.5 h after preparation. Then,
adjusts the vertical position of the glass probe to the surface friction before degenerative treatment
maintain a constant normal force during the sliding was characterized in PBS (Step 1). The specimen
test. Subsequently, the COF was recorded as the ratio was subsequently processed with one of the four
of the frictional force to the normal force measured degenerative treatments and equilibrated for 1 h in
by the sensors. PBS after the treatment before measuring the surface
To evaluate the surface friction of each specimen, friction again in PBS (Step 2). Finally, the treated
the glass probe was pressed against the cartilage specimens were equilibrated in SF for 1.5 h, and their
surface at a constant load, and the specimen in the surface friction in SF was examined (Step 3).
liquid bath was rotated by a servo motor at constant The process of the friction measurements is
speed. Initially, the contact load and sliding speed summarized schematically in Fig. 2. All experimental
were set to 5 mN and 0.05 mm/s, respectively. The procedures were conducted at room temperature.
sliding speed was subsequently increased to 0.1, 0.2, 2.3 Rheological test
0.5, 1.0, and 2.0 mm/s in a stepwise manner, and the
COF was recorded for five cycles at each sliding To determine the rheological characteristics of SF

Fig. 1 Photograph of the tribometer used in this study.

| https://mc03.manuscriptcentral.com/friction
Friction 5

Fig. 2 Schematic representation of the overall experimental design.

extracted from bovine metacarpal–phalangeal joints, using a fluorescence microscope equipped with a
a rotational rheometer (Physica MCR 301; Anton Paar, water-immersion objective.
Austria) with a cone-on-plate configuration was
employed. Sampled SF was cast between a flat plate 2.5 Statistical analysis
and shallow cone with a diameter of 50 mm and an Statistical analysis was conducted using one-way
apex angle of 0.5°, and the viscosity was measured at ANOVA for significance with Dunnett’s multiple
a constant shear rate of 1 × 104 s–1 at 25 °C. comparison test. Data and error bars are presented as
2.4 Staining and histology mean±the standard error of the mean (SEM). The
threshold for significance was set at P < 0.05.
Fluorescent staining was used to visualize the surface
morphology of the cartilage specimens and to confirm
3 Results
the damage to the surface caused by degenerative
treatments. A small section was obtained from the 3.1 Rheological test
surfaces of the treated and untreated specimens and
washed with 1% bovine serum albumin (BSA, Sigma) The viscosities of PBS and SF were plotted as functions
in PBS for 30 min before fluorescence staining with of shear rate on a log scale, as shown in Fig. 3. Bovine
keratan sulfate, which is a major side-chain component
of proteoglycan molecules. BSA was used as a
blocking agent to prevent non-specific binding of the
antibody to the cartilage tissue. The cartilage sections
were incubated with the primary antibody of
anti-keratan sulfate (MAB2022, Sigma; diluted 1:400
in PBS) at room temperature for 1.5 h in the dark. Then,
the sections were rinsed with PBS and subsequently
incubated with a secondary antibody solution consisting
of Alexa Flour 568 conjugated Goat anti-mouse IgG2b
(A-21144, ThermoFisher; diluted 1:400 in PBS) for 1 h
in the dark. The stained cartilage surface was imaged Fig. 3 Lubricant viscosity measurements of PBS and bovine SF.

www.Springer.com/journal/40544 | Friction
6 Friction

SF showed a characteristic shear-thinning behavior, dark spots might be defects caused in the superficial
reflected by a decrease in viscosity with increasing area of the cartilage surface by gentle washing in
shear rate. In contrast, PBS viscosity exhibited detergent solution. The severe wash group completely
Newtonian behavior. At the lowest shear rate, the lost its original smooth surface morphology, and rough
viscosity of SF was approximately 1,000 times higher and uneven surfaces with uniformly distributed small
than that of PBS. However, as the shear rate was dents were observed (Fig. 4(d)). The surface of the
increased to 104 s–1, the viscosity of SF decreased to NaCl group appeared to be slightly modified and
0.01 Pa·s, the difference in viscosity between PBS and had some similarities to that of the gentle-wash
SF became smaller, and the viscosity of SF became six group (Fig. 4(e)). Meanwhile, the collagen fiber network
times higher than that of PBS. The actual range of was exposed by trypsin digestion, and black areas
shear rates in the synovial joints extends from rest became dominant on the cartilage surface as a result
to 106 Pa·s [40]. This value is much higher than the of proteoglycan removal (Fig. 4(f)).
range of the shear rates in our rheological measurement.
The viscosity of SF was similar to that of PBS at this 3.3 Frictional characteristics of untreated cartilage
high shear rate. Therefore, the viscous effect of the SF specimens
may be much lower at a high shear rate in the actual
Figures 5(a)–5(c) shows the COF obtained from the
synovial joint.
untreated cartilage specimens in Steps 1, 2, and 3,
3.2 Fluorescence observation respectively, as a function of the sliding speed on a log
scale at different normal loads. In all steps, a visible
Fluorescence images of the cartilage specimen effect on the sliding speed was observed. The COF
surfaces are shown in Fig. 4. The surface of the fresh decreased with increasing sliding speed. The COF
specimen was consistently and uniformly smooth in Steps 1 and 3 remained almost constant with
(Fig. 4(a)). This represents the ideal appearance of the increasing normal load (Figs. 5(a) and 5(c)). Meanwhile,
intact uppermost superficial layer of the articular the COF in Step 2 exhibited a clear dependence on
cartilage. However, a slightly uneven surface was the contact load and decreased with an increase in
observed after stabilization in PBS for 1.5 h (Fig. 4(b)). the contact load (Fig. 5(b)). Consequently, the highest
The gently washed cartilage maintained a relatively COF was obtained at the lowest sliding speed and
smooth surface; however, randomly distributed dark contact load.
spots of different sizes were observed (Fig. 4(c)). These In Figs. 6(a) and 6(b), the highest COF of each

Fig. 4 Fluorescent images of the articular cartilage surfaces for treated and untreated cartilage specimens in PBS. (a) Fresh specimen;
(b) control specimen; (c) gentle wash group; (d) severe wash group; (e) NaCl group; (f) trypsin group. (Scale bars: 250 μm.)

| https://mc03.manuscriptcentral.com/friction
Friction 7

untreated specimen obtained in Steps 2 and 3 is plotted


against the highest COF in Step 1, respectively. There
was a certain variation in the highest COF of the
freshly prepared cartilage specimens (Step 1) obtained
at the lowest load of 5 mN and sliding speed of
0.05 mm/s. This might have been caused by variations
in the surface integrity of the prepared cartilage
specimens. The quality of the test specimens prepared
from animal tissues is inevitably affected by the
condition of the source animals, and it is difficult
to obtain uniform specimens. Such a difference in
specimen quality is responsible for the non-negligible
variation in the experimental data from biological
tissue specimens in most cases. After stabilization in
PBS (Step 2), the variation in the highest COF was
enhanced (Fig. 6(a)), particularly when the highest
COF in Step 1 exceeded 0.1. However, the effect of
PBS stabilization was limited for the untreated cartilage
specimens, which showed a relatively low COF of
<0.1. In this case, the cartilage specimen might have
had a preferable surface condition after the preparation
process and could preserve surface integrity during
the storage period in PBS. Subsequently, it showed
low COF, even under low-load and low-speed
conditions in Step 2. However, if the cartilage surface
was not perfect and had damage, its lubricating
function deteriorated and became vulnerable to storage
Fig. 5 Average values of the COF as a function of the sliding in PBS. Subsequently, some cartilage specimens
speed for the fresh specimen in (a) Step 1, (b) Step 2, and (c) Step 3 exhibited a notably high COF during Step 2.
under normal loads of 5 (blue), 20 (orange), and 100 (black) mN In contrast, the highest COF under SF lubrication
in PBS. (n = 15 for Step 1 and 2; n = 8 for Step 3.)
(Step 3) showed limited variation. It ranged from
0.07 to 0.20, and some specimens showed a reduced
COF compared to that in Step 1. These results clearly
indicate the lubricating function of the SF. Even if
the cartilage specimens had some damage to their
surface, the SF could compensate for it and reduce
the COF.

3.4 Effect of degenerative treatments

From the results of the second series of experiments,


the COF at the lowest sliding speed (0.05 mm/s)
Fig. 6 Variation of the COF in (a) PBS (Step 2) and (b) bovine
was chosen to evaluate the effect of degenerative
SF (Step 3) with the initial condition (Step 1) between the glass
probe and cartilage surface for the fresh and control specimens treatments on the boundary lubrication function of
at 5 mN and 0.05 mm/s in PBS. (n = 15 for the fresh and control the cartilage surface and was compared with the
specimens in PBS; n = 8 for the fresh and control specimens COF of untreated cartilage specimens under different
in SF.) loads in Figs. 7(a), 7(b), and 7(c). As indicated in the

www.Springer.com/journal/40544 | Friction
8 Friction

first series of experiments, the COF at the lowest sliding contact load. The highest average COF was observed
speed was sensitive to the initial condition of the in the severe wash group for all contact loads.
cartilage surface. Therefore, the effect of degenerative Meanwhile, the average COFs of the gentle-wash
treatments was examined using only selected cartilage and NaCl groups were very similar, regardless of the
specimens whose maximum COF in Step 1 was <0.1. load level.
At the lowest load of 5 mN, all treated specimens
exhibited a significant increase in the COF compared 3.5 Lubricating role of SF
with the untreated cartilage, except for the trypsin In Figs. 8(a)–8(e), the COFs at 5 mN and 0.05 mm/s in
group. Even when the load was increased to 20 and PBS and SF were compared for each cartilage specimen
100 mN, the treated cartilage exhibited a significantly regardless of the COF in Step 1 to demonstrate the
higher COF than the untreated cartilage, except lubrication effect of SF. The COF in PBS was highly
for the severe wash group at 100 mN. However, the dependent on the initial condition of the cartilage
difference in COF decreased with an increase in the surface and exhibited significant variation. However,
the COF variation in SF was comparatively smaller
(Fig. 6). This result indicates the lubricating effect
of the SF. The COF of most specimens in SF was
lower than that in PBS, particularly for the control
specimen and severe wash- and NaCl groups. The
average COF of the control specimen decreased
from 0.21 to 0.12 (Fig. 8(a)). This reduction in COF
was due to the SF constituents adsorbed on the
cartilage surface.
The average COF of the NaCl group (Fig. 8(d))

Fig. 7 Comparison of the COFs for all specimens lubricated


with PBS at the lowest sliding speed (0.05 mm/s) under different Fig. 8 Comparison of the COFs of the control and treated
normal loads. (a) 5, (b) 20, and (c) 100 mN. *P<0.05, **P<0.01, specimens lubricated with PBS and SF at 0.05 mm/s and 5 mN.
***
P<0.001, and ****P<0.0001 for differences from control specimen. (a) Control specimen; (b) gentle wash group; (c) severe wash group;
(n = 6 for the fresh and control specimens; n = 4 for the gentle (d) NaCl group; (e) trypsin group. (n = 8 for the control specimen;
wash, severe wash, and NaCl groups; n = 2 for the trypsin group n = 4 for the gentle wash and severe wash groups; n = 8 for the
in PBS.) NaCl and trypsin groups in PBS.)

| https://mc03.manuscriptcentral.com/friction
Friction 9

decreased from 0.28 to 0.18, whereas the average hydrogel layer. If the Péclet number exceeds a certain
COFs of the gentle-wash (Fig. 8(b)) and trypsin groups threshold, the contact size starts decreasing, and the
(Fig. 8(e)) only decreased from 0.20 to 0.18 and 0.23 COF decreases simultaneously. This phenomenon can
to 0.17, respectively. The average COF of the severe be attributed to the insufficient time for the interstitial
wash group in the SF was 0.11, which was lower water to be squeezed out from the contact area and
than that of the untreated control group. However, a activated biphasic lubrication mechanism under
limited number of severely washed specimens could high sliding speeds [21, 47]. Based on the results of
be tested in SF, and the sample number was only previous studies [42–46], we presumed that the COF
three. The gentle-wash, NaCl, and trypsin groups between the cartilage specimen and glass probe was
showed approximately similar average COFs that were evaluated under the boundary to mixed lubrication
larger than that of the control group in the SF. This condition between sliding surfaces in this series of
may be because the degenerative treatments removed experiments. The COF decrease with increasing sliding
certain molecules linked to the boundary lubrication speed can be attributed to hydrodynamic fluid film
mechanism at the cartilage surface. formation and enhanced fluid load support by the
water content of articular cartilage, even under the
lowest contact load of 5 mN.
4 Discussion
Furthermore, the COF of the control sample with a
Our research confirmed the influence of the sliding partially damaged uppermost superficial layer increased
speed and applied contact load on the friction behavior as the applied contact load decreased in PBS, as shown
of the cartilage surface in contact with a glass probe. in Fig. 5(b). The observed load dependency on the
The COF decreased with an increase in the sliding COF between the cartilage specimen and glass probe
speed. This is the typical frictional behavior of is consistent with the results of previous studies
lubricated bearings under the mixed-lubrication [48–50]. However, the fresh cartilage specimen with
condition suggested in the conventional Stribeck an intact surface did not show such load dependency
curve [41]. However, it is also widely known that on the COF (Fig. 5(a)). Consequently, the observed
compliant elastomers and hydrogels vary their friction load dependency is believed to be caused by the
significantly with sliding speed [42–44]. Therefore, partial depletion of the uppermost superficial layer
the COF between the compliant and highly hydrated on the cartilage surface.
articular cartilage and the glass probe may decrease As shown in Fig. 7, the specimens after degenerative
with increasing sliding speed, even under boundary treatments also showed a load dependency of their
lubrication conditions. Bonnevie et al. [45] reported COF, and the effects of degenerative treatments were
similar experimental results: the COF between articular significantly pronounced by reducing the contact
cartilage and a ridged spherical probe decreased with load to 5 mN. Based on this finding, the average
increasing sliding speed. They reported that the COF value and standard deviation of COFs shown in Fig. 7
was proportional to the contact area, which decreased were calculated for each specimen group and plotted
with increasing sliding speed because the stiffness against the applied contact load in Fig. 9 to elucidate
of the cartilage surface increased and this was the influence of degenerative treatments on the load
accompanied by a decrease in the ploughing (hysteresis) dependency of COFs. As shown in Fig. 9, each
component of the friction force. Delavoipière et al. degenerative treatment had a clear and intrinsic
[46] demonstrated a similar frictional behavior with impact on the load dependency of COF, and a distinct
thin hydrated hydrogel layers confined within contacts relationship between the applied load and the COF
between rigid glass substrates. They confirmed was demonstrated for each specimen group.
experimentally that frictional force and contact size As previously mentioned, all sliding tests in this
depended on the Péclet number defined by the ratio study were conducted under boundary to mixed
of the migration speed of the contact point on the gel lubrication conditions, and the COF decreased
surface to the flow rate of interstitial water within the with increasing sliding speed. The COF of the fresh

www.Springer.com/journal/40544 | Friction
10 Friction

articular cartilage is covered by the lamina splendens


[3, 6, 7, 52, 53]. The lamina splendens is an acellular
boundary layer that consists of an upper layer of an
amorphous substance and a lower layer of collagen
fibrils oriented parallel to the surface to provide the
highest possible tensile and shear strength [54–56].
Sawae et al. [57] confirmed that the amorphous
upper layer of lamina splendens mainly consists of
hydrated proteoglycan molecules. Furthermore, the
presence of adsorbed phospholipid bilayers on top
of the lamina splendens was confirmed by transmission
electron microscope (TEM) observations [7]. Jahn et al.
Fig. 9 Average values of COF as a function of the normal load [32, 58] suggested that hydrated phosphocholine
for cartilage specimens with sliding speed of 0.05 mm/s in PBS. heads aligned on phospholipid bilayers established
a low-shear sliding interface through the hydration
cartilage specimen increased with increasing contact lubrication mechanism and provided boundary
load. However, the control specimen showed an lubrication.
inconsistent load dependence: the COF was the highest In this study, we chose cartilage specimens whose
under 5 mN load and decreased as the load increased initial friction coefficient was < 0.1, even under the
to 20 mN. Subsequently, it remained almost constant, lowest load and speed conditions, to examine the
even when the load was increased to 100 mN. The effects of degenerative treatments. We assumed that
increase in COF under the lowest contact load was the selected specimens maintained their surface
further emphasized for the gentle-wash and NaCl integrity even after the specimen preparation process
groups. The two groups exhibited similar friction and retained intact lamina splendens and adsorbed
characteristics. The COF at 5 mN was three times higher phospholipid films [1, 4, 28] (Fig. 10(a)). The
than that of fresh cartilage and decreased considerably absorption of water by the hydrophilic phosphocholine
as the load increased to 20 mN. The significant heads in the phospholipid molecule results in hydration
difference in the load dependency mentioned above lubrication [30, 59, 60]. In this lubrication mechanism,
should be linked to the difference in the condition of the well-preserved adsorbed phospholipid film can
the superficial area of cartilage specimens demonstrated reduce attractive molecular interactions between
by fluorescence observation (Fig. 4). the surfaces, such as van der Waals forces, thereby
Fresh cartilage has a smooth surface, as shown in decreasing the adhesive force between the cartilage
Fig. 4(a), and the characteristic layered structure of and glass probe surfaces. Therefore, the COF of the
the cartilage superficial area has been reported in fresh specimens was the lowest among all the tested
numerous previous studies [5, 51]. As revealed by groups. On the other hand, fluorescence observation
several researchers, the uppermost surface of intact of the control specimen surface showed a slightly

Fig. 10 Schematic representation of articular cartilage surfaces of all treated and untreated specimens after the treatment process. (a) Fresh
specimen; (b) control specimen; (c) gentle wash group; (d) severe wash group; (e) NaCl group; (f) trypsin group. The dark blue: the
adsorbed phospholipids film; light blue: the lamina splendens; orange: denatured lamina splendens.

| https://mc03.manuscriptcentral.com/friction
Friction 11

uneven morphology (Fig. 4(b)). This may indicate that significantly altered (Fig. 10(d)). As a result, the cartilage
the absorbed phospholipid film was partially removed surface loses its intrinsic lubricating function and
from the cartilage surface (Fig. 10(b)). The imperfect exhibits high COF.
adsorbed film could not suppress the adhesive Trypsin group was digested by enzyme which
interaction between the surfaces and increased the removed proteoglycans from the cartilage surface and
COF under the lowest contact load. According to the exposed the collagen network (Fig. 4(f)) [39, 64–66].
Johnson–Kendall–Roberts (JKR) theory, the contact Therefore, the cartilage specimens in this group lost
area is influenced by the applied contact load and the their intrinsic lubrication function and exhibited
adhesive force between the rigid sphere and the higher COF. Some previous studies reported that a
compliant surface [61]. As the contact load decreases, trend of increased adhesion force can be observed in
the relative contribution of the adhesive force to trypsin-treated cartilage [66, 67]. However, removal
surface friction increased because of the significant of the compliant lamina splendens diminished the
deformation of the compliant cartilage surface by the marked increase in COF under the lowest contact
adhesive force [62]. load.
In the gentle-wash and NaCl groups, the cartilage Figure 11 shows the COF of the gently washed and
surface was washed out by the detergent or NaCl NaCl groups at 5 mN, 0.05 mm/s and 5 mN, 2 mm/s
solution and lost its integrity, as shown in Figs. 4(c) in SF. At 5 mN and 2 mm/s, the COF decreased
and 4(e). Jones et al. [37] demonstrated that NaCl significantly and showed a minimum COF compared
treatment could remove lubricin from the cartilage with that at 5 mN, 0.05 mm/s. The gentle-wash group
surface, while the overall tissue proteoglycan content exhibited the same COF as the NaCl group at the
was hardly changed. Jahn et al. [58] suggested lowest speed and contact load. However, the COF of
that lubricin play an important role in maintaining the NaCl group was larger than that of the gentle-wash
adsorbed phospholipids and HA complexes on group at the highest speed and contact load tested.
the cartilage surface. Consequently, the adsorbed Based on Figs. 4(a) and 4(e), it is clear that the
phospholipid film could be removed from the cartilage processing treatment by NaCl may have removed
surface by NaCl treatment. Therefore, the removal something that is necessary to activate the lubrication
of the adsorbed phospholipid film by NaCl treatment effect of the fluid film.
and gentle washing with the detergent might result We applied extremely low loads between the glass
in the exposure of the more compliant lamina and cartilage to evaluate the friction behavior in the
splendens with much lower stiffness, more than superficial layer on the cartilage surface, attempting
inactivating the hydration lubrication by phospholipid to minimize contributions from the bulk of the
films (Figs. 10(c) and 10(e)) [4, 63]. Consequently, the cartilage tissue. Higher contact loads lead to greater
effect of the adhesive interaction was further enhanced surface deformation in the cartilage and greater energy
by the larger deformation of the compliant lamina dissipation due to the viscoelastic deformation of the
splendens, which caused a marked increase in the cartilage tissue. In addition, cartilage deformation
COF under the lowest contact load. Based on the JKR can lead to increased interstitial fluid support. As
theory, the relative contribution of the adhesive force mentioned above, the Péclet number represents the
to the exerted friction should be emphasized at low
contact loads [61, 62].
As shown in Fig. 4(d), severe washing with the
ultrasonic cleaner completely changed the morphology
of the cartilage surface. It is likely that ultrasonication
caused a significant temperature increase and induced
the denaturation of proteins on the cartilage surface.
Therefore, the original layered structure on the cartilage
surface was severely damaged, and the chemical Fig. 11 Average COF values of gentle wash group and NaCl
and physical properties of the lamina splendens were group at (a) 5 mN, 0.05 mm/s and (b) 5 mN, 2 mm/s in SF.

www.Springer.com/journal/40544 | Friction
12 Friction

ratio of the sliding velocity to the velocity of the cartilage surface makes it challenging to achieve a
interstitial fluid flow relative to the solid matrix. perfect cartilage–cartilage contact pair.
It can be used to estimate the magnitude of the We applied a range of sliding speeds to study
interstitial fluid load support. The contact radius is the frictional behavior of the cartilage surface. As the
positively correlated with the Péclet number. Therefore, sliding velocity increases, the flow resistance of
reducing the contact load decreased the Péclet number water and the associated viscous drag force typically
and minimized the effect of biphasic lubrication by increase [70]. However, we observed a decrease in
interstitial water in the cartilage tissue. However, 5 mN the tangential force with an increase in the sliding
was the minimum contact load that could be applied velocity (Fig. 9). The viscosities of both PBS and
in this study because significant data instability was SF are very low. Consequently, the contribution of the
observed at contact loads below 5 mN. The Péclet drag force could be considered negligible compared
number was calculated from Pe  Va /H  A k where V is with that of the force detected in the current
sliding speed, a is the contact radius, H  A is aggregate experiments.
modulus, and k is permeability. According to the
Hertz contact theory [43], the contact radius between
the cartilage surface and glass probe can be estimated
5 Conclusions
as 0.236–0.641 mm. If we assume H  A = 13 MPa [68] The friction characteristics of treated and untreated
and k  6  10 16 m4/(N·s) [68], when the applied load cartilage specimens in contact with a smooth glass
varies between 5 and 100 mN and the sliding speed surface were evaluated using an angular reciprocating
changes between 0.05 and 2 mm/s, Pe ranges from sliding test with a range of low contact loads and
1.5 to 164.1, still greater than 1. This indicates that
sliding speeds in PBS and SF. To confirm the effects
biphasic lubrication still influenced the friction results
of the treatments on the morphology of the cartilage
obtained here. In addition, the elastic deformation
surface, a fluorescence microscope and water
of the cartilage tissue produced by contact with
immersion objectives were used.
the glass probe and viscoelastic energy dissipation
The notable findings of this work can be
should contribute to the friction result. These factors
summarized as follows:
represent the limitations of our experiment, leading
1) A certain variation was observed in the highest
to discrepancies between our results and those of
COF of the freshly prepared cartilage specimens
other studies. For example, Shoaib et al. [69] applied
obtained at the lowest load and sliding speed. The
a contact pressure of less than 1 kPa and a sliding
condition of the original source animals may be
velocity of 1 μm/s between lamina splendens and mica.
responsible for this variation. Such differences in
In this scenario, using a surface-force apparatus,
specimen quality in most biological tissue tests should
the COF in PBS was found to be in the range of 0.003.
However, the results showed that the COF of the fresh be considered.
samples in PBS reached 0.1 in our experiment, even 2) The COF of the treated specimens increased
for the samples with an intact uppermost superficial significantly after the degenerative treatments. This
layer. According to the Hertz contact theory [43], the may be due to the deterioration of the integrity of the
contact pressure in the center of the contact area cartilage surface caused by degenerative treatments.
between the cartilage surface and glass probe can 3) Compared with PBS, SF exhibited an obvious
be estimated as 42.8–116.2 kPa. This was a significant boundary lubrication effect on the cartilage surface,
increase compared to Shoaib’s experiment. Consequently, even for the treated specimens.
there should be contributions from cartilage bulk 4) The untreated and treated cartilage specimens
deformation, viscoelastic energy dissipation, or fluid exhibited different friction characteristics at the
flow inside the cartilage tissue. Another potential lowest sliding speed and normal load. The different
limitation of this study is that friction measurements adhesion forces acting between the untreated and
were performed between the cartilage and glass rather treated cartilage surfaces and glass probe surfaces
than against the cartilage. The unevenness of the may be responsible for these results.

| https://mc03.manuscriptcentral.com/friction
Friction 13

Acknowledgements layer of articular cartilage in the lubrication mechanism of


joints. J Anat 199(3): 241–250 (2001)
Financial support was given by the Grant-in Aid for [6] Orford C R, Gardner D L. Ultrastructural histochemistry of
Scientific Research (A) of Japan Society for the the surface lamina of normal articular cartilage. Histochem
Promotion of Science (21H04535). J 17(2): 223–233 (1985)
[7] Higaki H, Murakami T, Nakanishi Y, Miura H, Mawatari T,
Iwamoto Y. The lubricating ability of biomembrane models
Declaration of competing interest with dipalmitoyl phosphatidylcholine and gamma-globulin.
Proc Inst Mech Eng H 212(5): 337–346 (1998)
The authors have no competing interests to declare
[8] Cook S G, Guan Y, Pacifici N J, Brown C N, Czako E,
that are relevant to the content of this article.
Samak M S, Bonassar L J, Gourdon D. Dynamics of synovial
fluid aggregation under shear. Langmuir 35(48): 15887–15896
Open Access This article is licensed under a Creative
(2019)
Commons Attribution 4.0 International License, which [9] Schmidt T A, Gastelum N S, Nguyen Q T, Schumacher
permits use, sharing, adaptation, distribution and B L, Sah R L. Boundary lubrication of articular cartilage:
reproduction in any medium or format, as long as Role of synovial fluid constituents. Arthritis Rheum 56(3):
you give appropriate credit to the original author(s) 882–891 (2007)
and the source, provide a link to the Creative Commons [10] More S, Kotiya A, Kotia A, Ghosh S K, Spyrou L A, Sarris
licence, and indicate if changes were made. I E. Rheological properties of synovial fluid due to
The images or other third party material in this viscosupplements: A review for osteoarthritis remedy.
article are included in the article’s Creative Commons Comput Methods Programs Biomed 196: 105644 (2020)
licence, unless indicated otherwise in a credit line to [11] Jung S, Petelska A, Beldowski P, Augé W K, Casey T,
the material. If material is not included in the article’s Walczak D, Lemke K, Gadomski A. Hyaluronic acid and
Creative Commons licence and your intended use is phospholipid interactions useful for repaired articular cartilage
surfaces—A mini review toward tribological surgical
not permitted by statutory regulation or exceeds the
adjuvants. Colloid Polym Sci 295(3): 403–412 (2017)
permitted use, you will need to obtain permission
[12] Ghosh S, Bowen J, Jiang K, Espino D M, Shepherd D E T.
directly from the copyright holder.
Investigation of techniques for the measurement of articular
To view a copy of this licence, visit
cartilage surface roughness. Micron 44: 179–184 (2013)
http://creativecommons.org/licenses/by/4.0/. [13] Charnley J. The lubrication of animal joints in relation to
surgical reconstruction by arthroplasty. Ann Rheum Dis
References 19(1): 10–19 (1960)
[14] Schmidt T A, Sah R L. Effect of synovial fluid on boundary
[1] Murakami T, Nakashima K, Yarimitsu S, Sawae Y, Sakai N. lubrication of articular cartilage. Osteoarthritis Cartilage
Effectiveness of adsorbed film and gel layer in hydration 15(1): 35–47 (2007)
lubrication as adaptive multimode lubrication mechanism [15] Murakami T, Higaki H, Sawae Y, Ohtsuki N, Moriyama S,
for articular cartilage. Proc Inst Mech Eng Part J J Eng Nakanishi Y. Adaptive multimode lubrication in natural
Tribol 225(12): 1174–1185 (2011) synovial joints and artificial joints. Proc Inst Mech Eng H
[2] Eyre D. Articular cartilage and changes in Arthritis: Collagen 212(1): 23–35 (1998)
of articular cartilage. Arthritis Res Ther 4(1): 30 (2001) [16] Tanner R I. An alternative mechanism for the lubrication of
[3] Lane J M, Weiss C. Review of articular cartilage collagen synovial joints. Phys Med Biol 11(1): 119–127 (1966)
research. Arthritis Rheum 18(6): 553–562 (1975) [17] Dowson D. Paper 12: Modes of lubrication in human joints.
[4] Murakami T, Nakashima K, Sawae Y, Sakai N, Hosoda N. Proc Inst Mech Eng Conf Proc 181(10): 45–54 (1966)
Roles of adsorbed film and gel layer in hydration lubrication [18] Unsworth A. Tribology of human and artificial joints. Proc
for articular cartilage. Proc Inst Mech Eng Part J J Eng Inst Mech Eng H 205(3): 163–172 (1991)
Tribol 223(3): 287–295 (2009) [19] McCutchen C W. Mechanism of animal joint: Sponge-
[5] Kumar P, Oka M, Toguchida J, Kobayashi M, Uchida E, hydrostatic and Weeping Bearings. Nature 184: 1284–1285
Nakamura T, Tanaka K. Role of uppermost superficial surface (1959)

www.Springer.com/journal/40544 | Friction
14 Friction

[20] Walker P S, Dowson D, Longfield M D, Wright V. recombinant human lubricin. J Orthop Res 27(6): 771–777
“Boosted lubrication” in synovial joints by fluid entrapment (2009)
and enrichment. Ann Rheum Dis 27(6): 512–520 (1968) [37] Jones A R C, Gleghorn J P, Hughes C E, Fitz L J, Zollner R,
[21] Ateshian G A. The role of interstitial fluid pressurization in Wainwright S D, Caterson B, Morris E A, Bonassar L J,
articular cartilage lubrication. J Biomech 42(9): 1163–1176 Flannery C R. Binding and localization of recombinant
(2009) lubricin to articular cartilage surfaces. J Orthop Res 25(3):
[22] Ateshian G A, Wang H Q, Lai W M. The role of interstitial 283–292 (2007)
fluid pressurization and surface porosities on the boundary [38] Pickard J E, Fisher J, Ingham E, Egan J. Investigation into
friction of articular cartilage. J Tribol 120(2): 241–248 (1998) the effects of proteins and lipids on the frictional properties of
[23] Mow M C, Ling F F. On weeping lubrication theory. Z Für articular cartilage. Biomaterials 19(19): 1807–1812 (1998)
Angew Math Und Phys ZAMP 20(2): 156–166 (1969) [39] Poole A R, Pidoux I, Reiner A, Tang L H, Choi H,
[24] Ateshian G A. A theoretical formulation for boundary friction Rosenberg L. Localization of proteoglycan monomer and
in articular cartilage. J Biomech Eng 119(1): 81–86 (1997) link protein in the matrix of bovine articular cartilage: An
[25] Linn F C, Radin E L. Lubrication of animal joints. iii. the immunohistochemical study. J Histochem Cytochem 28(7):
effect of certain chemical alterations of the cartilage and 621–635 (1980)
lubricant. Arthritis Rheum 11(5): 674–682 (1968) [40] Seror J, Merkher Y, Kampf N, Collinson L, Day A J,
[26] Ishikawa Y, Hiratsuka K I, Sasada T. Role of water in the Maroudas A, Klein J. Normal and shear interactions between
lubrication of hydrogel. Wear 261(5–6): 500–504 (2006) hyaluronan–aggrecan complexes mimicking possible boundary
[27] Lin W F, Klein J. Hydration lubrication in biomedical lubricants in articular cartilage in synovial joints.
applications: From cartilage to hydrogels. Acc Mater Res Biomacromolecules 13(11): 3823–3832 (2012)
3(2): 213–223 (2022) [41] Stachowiak G, Batchelor A. Engineering tribology, 4th edn.
[28] Murakami T, Yarimitsu S, Nakashima K, Sawae Y, Sakai N. Butterworth-Heinemann, 2013.
Influence of synovia constituents on tribological behaviors [42] Gong J P. Friction and lubrication of hydrogels—Its richness
of articular cartilage. Friction 1(2): 150–162 (2013) and complexity. Soft Matter 2(7): 544–552 (2006)
[29] Bonnevie E D, Galesso D, Secchieri C, Cohen I, Bonassar [43] Popov V L. Contact Mechanics and Friction: Physical
L J. Elastoviscous transitions of articular cartilage reveal a Principles and Applications. Springer, 2010.
mechanism of synergy between lubricin and hyaluronic [44] Persson B N J, Spencer N D. Sliding friction: Physical
acid. PLoS One 10(11): e0143415 (2015) principles and applications. Phys Today 52(1): 66–68 (1999)
[30] Seror J, Zhu L Y, Goldberg R, Day A J, Klein J. [45] Bonnevie E D, Baro V J, Wang L, Burris D L. In situ
Supramolecular synergy in the boundary lubrication of studies of cartilage microtribology: Roles of speed and
synovial joints. Nat Commun 6: 6497 (2015) contact area. Tribol Lett 41(1): 83–95 (2011)
[31] Chang D P, Abu-Lail N I, Coles J M, Guilak F, Jay G D, [46] Delavoipière J, Tran Y, Verneuil E, Heurtefeu B, Hui C Y,
Zauscher S. Friction force microscopy of lubricin and Chateauminois A. Friction of poroelastic contacts with thin
hyaluronic acid between hydrophobic and hydrophilic surfaces. hydrogel films. Langmuir 34(33): 9617–9626 (2018)
Soft Matter 5(18): 3438–3445 (2009) [47] Caligaris M, Ateshian G A. Effects of sustained interstitial
[32] Jahn S, Seror J, Klein J. Lubrication of articular cartilage. fluid pressurization under migrating contact area, and
Phys Today 71(4): 48–54 (2018) boundary lubrication by synovial fluid, on cartilage friction.
[33] Mendibil U, Ruiz-Hernandez R, Retegi-Carrion S, Garcia- Osteoarthritis Cartilage 16(10): 1220–1227 (2008)
Urquia N, Olalde-Graells B, Abarrategi A. Tissue-specific [48] Majd S E, Rizqy A I, Kaper H J, Schmidt T A, Kuijer R,
decellularization methods: Rationale and strategies to achieve Sharma P K. An in vitro study of cartilage–meniscus
regenerative compounds. Int J Mol Sci 21(15): 5447 (2020) tribology to understand the changes caused by a meniscus
[34] Gilbert T W, Sellaro T L, Badylak S F. Decellularization of implant. Colloids Surf B Biointerfaces 155: 294–303 (2017)
tissues and organs. Biomaterials 27(19): 3675–3683 (2006) [49] Katta J, Pawaskar S S, Jin Z M, Ingham E, Fisher J. Effect
[35] Gleghorn J P, Jones A, Flannery C R, Bonassar L J. of load variation on the friction properties of articular
Boundary mode frictional properties of engineered cartilage. Proc Inst Mech Eng Part J J Eng Tribol 221(3):
cartilaginous tissues. Eur Cells Mater 14: 20–29 (2007) 175–181 (2007)
[36] Gleghorn J P, Jones A R C, Flannery C R, Bonassar L J. [50] Furmann D, Nečas D, Rebenda D, Čípek P, Vrbka M,
Boundary mode lubrication of articular cartilage by Křupka I, Hartl M. The effect of synovial fluid composition,

| https://mc03.manuscriptcentral.com/friction
Friction 15

speed and load on frictional behaviour of articular cartilage. [62] He B, Chen W, Wang Q J. Surface texture effect on friction
Materials 13(6): 1334 (2020) of a microtextured poly(dimethylsiloxane) (PDMS). Tribol
[51] Jurvelin J S, Müller D J, Wong M, Studer D, Engel A, Lett 31(3): 187–197 (2008)
Hunziker E B. Surface and subsurface morphology of bovine [63] Singh S, Afara I O, Tehrani A H, Oloyede A. Effect
humeral articular cartilage as assessed by atomic force and of decellularization on the load-bearing characteristics of
transmission electron microscopy. J Struct Biol 117(1): articular cartilage matrix. Tissue Eng Regen Med 12(5):
45–54 (1996) 294–305 (2015)
[52] Murakami T, Sawae Y, Ihara M. Protective mechanism of [64] Chan S M T, Neu C P, Duraine G, Komvopoulos K, Reddi
articular cartilage to severe loading: Roles of lubricants, A H. Atomic force microscope investigation of the boundary-
cartilage surface layer, extracellular matrix and chondrocyte. lubricant layer in articular cartilage. Osteoarthritis Cartilage
JSME Int J, Ser C 46(2): 594–603 (2003) 18(7): 956–963 (2010)
[53] Saito T. The superficial zone of articular cartilage. Inflamm
[65] Wahlquist J A, DelRio F W, Randolph M A, Aziz A H,
Regen 42(1): 14 (2022)
Heveran C M, Bryant S J, Neu C P, Ferguson V L. Indentation
[54] Thielen N G M, van der Kraan P M, van Caam A P M.
mapping revealed poroelastic, but not viscoelastic, properties
TGFβ/BMP signaling pathway in cartilage homeostasis. Cells
spanning native zonal articular cartilage. Acta Biomater 64:
8(9): 969 (2019)
41–49 (2017)
[55] Bhosale A M, Richardson J B. Articular cartilage: Structure,
[66] Rojas F P, Batista M A, Lindburg C A, Dean D, Grodzinsky
injuries and review of management. Br Med Bull 87: 77–95
A J, Ortiz C, Han L. Molecular adhesion between cartilage
(2008)
extracellular matrix macromolecules. Biomacromolecules
[56] Fujioka R, Aoyama T, Takakuwa T. The layered structure
15(3): 772–780 (2014)
of the articular surface. Osteoarthritis Cartilage 21(8):
[67] Lawrence A, Xu X, Bible M D, Calve S, Neu C P, Panitch
1092–1098 (2013)
A. Synthesis and characterization of a lubricin mimic
[57] Sawae Y, Murakami T, Matsumoto K, Horimoto M. Study
on morphology and lubrication of articular cartilage surface (mLub) to reduce friction and adhesion on the articular
with atomic force microscopy. Jpn J Tribol 45(1): 51–62 cartilage surface. Biomaterials 73: 42–50 (2015)
(2000) [68] Soltz M A, Ateshian G A. A conewise linear elasticity
[58] Jahn S, Klein J. Hydration lubrication: The macromolecular mixture model for the analysis of tension-compression
domain. Macromolecules 48(15): 5059–5075 (2015) nonlinearity in articular cartilage. J Biomech Eng 122(6):
[59] Gaisinskaya-Kipnis A, Klein J. Normal and frictional 576–586 (2000)
interactions between liposome-bearing biomacromolecular [69] Shoaib T, Yuh C, Wimmer M A, Schmid T M, Espinosa-
bilayers. Biomacromolecules 17(8): 2591–2602 (2016) Marzal R M. Nanoscale insight into the degradation
[60] Jahn S, Seror J, Klein J. Lubrication of articular cartilage. mechanisms of the cartilage articulating surface preceding
Annu Rev Biomed Eng 18: 235–258 (2016) OA. Biomater Sci 8(14): 3944–3955 (2020)
[61] Johnson K L, Kendall K, Roberts A D. Surface energy and [70] Bahrami M, Le Houérou V, Rühe J. Lubrication of surfaces
contact of elastic solids. Proc R Soc A 324(1558): 301–313 covered by surface-attached hydrogel layers. Tribol Int 149:
(1971) 105637 (2020)

Wenxiao LI. She is currently a University in Japan, under the supervision of Prof.
doctoral student majoring in Yoshinori SAWAE. Her current research focuses on
mechanical engineering at Kyushu biomechanics of articular cartilage.

www.Springer.com/journal/40544 | Friction
16 Friction

Takehiro MORITA. He received Departure of Mechanical Engineering, Faculty of


his master’s degree in mechanical Engineering, Kyushu university, Japan. His major
engineering from Kumamoto research areas include friction and wear of prosthetic
University, Japan, in 1982. He is joint materials and polymer tribology in hydrogen
currently an associate professor at atmosphere.

Yoshinori SAWAE. He received Engineering, Faculty of Engineering, Kyushu university,


his Ph.D. degree in mechanical Japan. His major research areas include friction and
engineering from Kyushu University, wear of prosthetic joint materials, biomechanics of
Japan, in 1996. He is currently a articular cartilage and chondrocytes, and polymer
professor at Departure of Mechanical tribology in hydrogen atmosphere.

| https://mc03.manuscriptcentral.com/friction

You might also like