ArticlePDF Available

Focal Adhesion Kinase Stabilizes the Cytoskeleton

Authors:

Abstract and Figures

Focal adhesion kinase (FAK) is a central focal adhesion protein that promotes focal adhesion turnover, but the role of FAK for cell mechanical stability is unknown. We measured the mechanical properties of wild-type (FAKwt), FAK-deficient (FAK-/-), FAK-silenced (siFAK), and siControl mouse embryonic fibroblasts by magnetic tweezer, atomic force microscopy, traction microscopy, and nanoscale particle tracking microrheology. FAK-deficient cells showed lower cell stiffness, reduced adhesion strength, and increased cytoskeletal dynamics compared to wild-type cells. These observations imply a reduced stability of the cytoskeleton in FAK-deficient cells. We attribute the reduced cytoskeletal stability to rho-kinase activation in FAK-deficient cells that suppresses the formation of ordered stress fiber bundles, enhances cortical actin distribution, and reduces cell spreading. In agreement with this interpretation is that cell stiffness and cytoskeletal stability in FAK-/- cells is partially restored to wild-type level after rho-kinase inhibition with Y27632.
Content may be subject to copyright.
Focal Adhesion Kinase Stabilizes the Cytoskeleton
Ben Fabry, Anna H. Klemm, Sandra Kienle, Tilman E. Scha
¨ffer, and Wolfgang H. Goldmann*
Department of Physics, Friedrich-Alexander-University Erlangen-Nuremberg, Erlangen, Germany
ABSTRACT Focal adhesion kinase (FAK) is a central focal adhesion protein that promotes focal adhesion turnover, but the
role of FAK for cell mechanical stability is unknown. We measured the mechanical properties of wild-type (FAKwt), FAK-deficient
(FAK / ), FAK-silenced (siFAK), and siControl mouse embryonic fibroblasts by magnetic tweezer, atomic force microscopy,
traction microscopy, and nanoscale particle tracking microrheology. FAK-deficient cells showed lower cell stiffness, reduced
adhesion strength, and increased cytoskeletal dynamics compared to wild-type cells. These observations imply a reduced
stability of the cytoskeleton in FAK-deficient cells. We attribute the reduced cytoskeletal stability to rho-kinase activation in
FAK-deficient cells that suppresses the formation of ordered stress fiber bundles, enhances cortical actin distribution, and
reduces cell spreading. In agreement with this interpretation is that cell stiffness and cytoskeletal stability in FAK / cells is
partially restored to wild-type level after rho-kinase inhibition with Y27632.
INTRODUCTION
Cells that adhere to an extracellular matrix form an architec-
turally highly complex cytoskeleton. A major component of
the cytoskeleton is force-generating actomyosin stress fibers
that connect to focal adhesions (FAs). The interplay of
actomyosin stress fibers and FAs defines to a large part
the mechanical behavior of cells, such as their motility,
morphology, and contractility (1 6).
Focal adhesion kinase (FAK) is a central protein of FAs
and is known to regulate several cytoskeletal and other focal
adhesion proteins. FAK interacts with integrins (7), paxillin
(8), p130Cas (9), a-actinin (10), and other proteins that link
FAs to the actin cytoskeleton (4). The molecular details of
these interactions have not been fully characterized, but it
is generally agreed that FAK promotes a high FA turnover
through a rho-kinase (ROCK)-dependent pathway (11 14).
FAK knockout cells show high rhoA-kinase and ROCK
activity (12,13,15). Because ROCK inactivates myosin light
chain phosphatase, phosphorylates myosin light chain, and
therefore promotes actomyosin contractility, it has been
hypothesized that contractile tension in the cytoskeleton is
altered in FAK-deficient cells (10,14 18). Direct measure-
ments of traction forces, however, showed no difference
between wild-type and FAK-deficient fibroblasts (19).
With regard to the mechanical stability of cells, the role of
FAK is similarly unclear. Because FAK promotes high FA
turnover and high cell motility (12,13,19), it could be
expected that the cytoskeleton in FAK-deficient cells is
less dynamic and more rigid compared to FAK-expressing
cells. However, FAK-deficient cells show a rounded cell
morphology with a smaller spreading area, pronounced
cortical distribution of the actin cytoskeleton, and a loss of
actomyosin stress fibers (12,13), all of which are signs
of reduced cell stiffness. Treatment of FAK-deficient cells
with the ROCK-inhibitor Y27632 leads to a larger spreading
area and a reformation of stress fibers (13), which is
puzzling as this inhibitor induces the complete opposite
behavior in wild-type fibroblasts (20).
The aim of this work was first to directly measure the
impact of FAK on cell mechanics in mouse embryonic
fibroblasts (MEFs), and second to characterize how rho-
kinase contributes to the mechanical changes in wild-type
and FAK-deficient cells. Our results show that FAK is
important for maintaining cell rigidity (stiffness) through
promoting a static and highly aligned contractile cytoskel-
eton. FAK knockout leads to a pronounced speedup of cyto-
skeletal dynamics, which is independent of any decreased
FA turnover in these cells. We hypothesize that the effects
of FAK on cytoskeletal dynamics and organization are, to
a large extent, mediated through a compensatory activation
of ROCK in FAK knockout cells. In support of this hypoth-
esis, we find that treatment with the ROCK-inhibitor
Y27632 has opposite effects on cell rigidity and cytoskeletal
dynamics in FAK wild-type versus knockout cells.
MATERIALS AND METHODS
Cells and cell culture
FAK deficient (FAK / ) and FAK wild type (FAKwt) mouse embryonic
fibroblasts (MEFs) were purchased from American Type Culture Collection
(FAK / , cat. No. CRL 2644; FAKwt, cat. No. CRL 2645; ATCC,
Manassas, VA). FAKwt and FAK / cells from ATCC are reported to
carry a p53 knockout mutation (11). Moreover, FAK / cells overexpress
PYK2, a homologous protein of FAK (21). PYK2 overexpression has no
effect on cell mechanics (22), but p53 overexpression may have a small
effect (23). To rule out any of these secondary effects, we performed siRNA
downregulation experiments on differently derived MEF cells (obtained
from Dr. W. H. Ziegler, University of Leipzig, Leipzig, Germany) with
normal PYK2 and p53 expression levels. All cell lines were maintained
in low glucose (1 g/L) Dulbecco’s modified Eagle’s medium supplemented
with 10% fetal calf serum, 2 mM L glutamine, and 100 U/ml penicillin
streptomycin (i.e., Dulbecco’s modified Eagle’s medium complete
medium). This medium was also used during measurements except where
Submitted June 9, 2011, and accepted for publication September 23, 2011.
*Correspondence: wgoldmann@biomed.uni erlangen.de
Editor: Douglas Nyle Robinson.
Ó2011 by the Biophysical Society
0006 3495/11/11/2131/8 $2.00 doi: 10.1016/j.bpj.2011.09.043
Biophysical Journal Volume 101 November 2011 2131 2138 2131
stated otherwise. siRNA against FAK (siFAK) was targeted against both
splice variants of murine FAK (gene accession numbers NM 007982.2
and NM 001130409.1). The sequence was sense: R(GGG ACA UUG
CUG CUC GGA A)dTdT; antisense: R(UUC CGA GCA GCA AUG
UCC C)dTdG. siRNA was 30AlexaFluor546 labeled to assess transfection
efficiency. As a control (siControl), we used Alexa Fluor546 labeled
Allstar siRNA (Qiagen, Hilden, Germany), a nonsilencing siRNA with no
homology to any known mammalian gene. Transfection of 100,000
MEFs was performed in 35 mm wells using 6 ml HiPerFect transfection
reagent (Qiagen) with 20 nM siRNA.
Magnetic tweezer microrheology
Magnetic tweezer rheology exerts a mechanical shear stress to the cell by
applying lateral forces to magnetic beads that are connected to the cytoskel
eton through adhesion contacts on the apical cell surface (24). This
technique reports the passive mechanical properties of the cytoskeleton,
such as the elastic modulus and its time (or frequency) dependency. In brief,
superparamagnetic, epoxylated beads (4.5 mm, Dynabeads; Invitrogen,
Carlsbad, CA) were coated with fibronectin (5 mg per 1 10
7
beads; Roche,
Pleasanton, CA) at 4C for 24 h. Before measurements, the beads were
sonicated, 2 10
5
beads were added to ~10
5
subconfluent cells in a
35 mm dish, and cells were incubated with the beads for 30 min at 5%
CO
2
and 37C. Thereafter, the medium was exchanged with fresh, pre
warmed medium to remove unbound beads. Measurements were performed
on a heated inverted microscope stage at 40magnification (NA 0.6)
without CO
2
. The measuring time was limited to 30 min per dish. A sole
noid with a sharp tipped steel needle core was used to generate a defined
force on the bead.
When a step force fwas applied to a cell bound bead, it moved with
a displacement d(t) toward the tweezer needle tip (Fig. 1). Following Kasza
et al. (25) and Kollmannsberger et al. (26), we estimate the typical strain
g(t)asd(t) divided by the bead radius r, and the typical stress sas the
applied force divided by the bead cross sectional area, r
2
p. The creep
compliance J(t) in units of Pa
1
is then given by g(t)/sand is fit to the equa
tion J(t)J
0
(t/t
0
)
b
with time normalized to t
0
1 s. The prefactor, J
0
, and
the power law exponent, b, were both force dependent. The value J
0
is the
creep compliance at t
0
1 s and corresponds, apart from a negligible
correction factor (the Gamma function, G,at1 b) to the inverse magnitude
of the cell’s dynamic shear modulus evaluated at a radian frequency u
0
1
rad/s (27,28),
G0ðu0ÞþiG00ðu0Þ
¼1
J0
Gð1bÞ:
The power law exponent breflects the dynamics of the force bearing
structures of the cell that are connected to the bead (27). A power law expo
nent of b0 is indicative of a purely elastic solid, and b1 is indicative
of a purely viscous fluid. In cells, the power law exponent usually falls in
the range between 0.1 and 0.5, whereby higher values have been linked
to a higher turnover rate of cytoskeletal structures (27).
Atomic force microscopy
Atomic force microscopy (AFM) is used as an alternative method to
measure cell stiffness. A nonfunctionalized sharp tip that is in contact
with the cell for <1 s serves as a probe. The measurements are therefore
not influenced by focal adhesion formation between the probe and the
cell. Cells were seeded on fibronectin coated (50 mg/ml) cell culture dishes
(Nalge Nunc, Rochester, NY) in CO
2
independent medium (Leibovitz L 15
medium, with L Glutamine; Gibco, Invitrogen, Carlsbad, CA) for 12 h
before and during measurements. Measurements were performed on a
MFP 3D Stand Alone AFM (Asylum Research, Goleta, CA) as described
previously (22). The spring constants of the cantilevers (Bio Lever;
Olympus, Melville, NY) were determined before cell measurements using
the thermal noise method (29) and were in the range of 5.9 7.7 pN/nm.
The force mapping mode with 260 pN maximum indentation force (inden
tation depth between 30 and 100 nm) was used to measure cell stiffness
and sample height (Fig. 2 A). Force versus z piezo extension curves
(Fig. 2 C) were acquired on different positions on the sample surface. The
local shear moduli (Fig. 2 B) were obtained by fitting the extended Hertz
model to each force versus z piezo extension curve in the force map (30).
In the extended Hertz model for a conical indenter, the relationship between
the applied force Fand the resulting sample indentation dis given by
F¼2
p
E
1n2d2tanðaÞ;
where Eis the Young’s modulus, nis the Poisson’s ratio, and ais the
opening half angle of the conical cantilever tip (30,31). We assumed
an incompressible sample with a Poisson’s ratio of n0.5. Therefore,
the shear modulus Gis related to the Young’s modulus Eby G
E/[2$(1þn)] E/3. A region of 80 80 mm was scanned to obtain
10 10 or 20 20 force curves. To reduce the influence of the underlying
substrate, only regions of the cell that were at least 60% of the maximum
cell height were analyzed.
Nanoscale particle tracking
This method is used to quantify internal remodeling processes of the
cytoskeleton. Fibronectin coated fluorescent beads (4.5 mm) are bound to
confluent cells via integrins (32,33). The spontaneous movement of the
beads was tracked for 5 min. The mean squared displacement (MSD) of
bead movements followed a power law with time (t) according to
MSD D$(t/t
0
)
b
(Fig. 3). The value t
0
was set to 1 s, Dreflects an apparent
diffusivity, equivalent to the square of the distance traveled during a 1 s
interval, and the power law exponent bis a measure of the persistence,
with b1 for randomly moving beads and b2 for directed, ballistic
motion along a straight path (32).
Immunofluorescence of focal adhesions and the actin
cytoskeleton
A quantity of 10,000 50,000 cells was seeded on 5 mg/ml fibronectin
coated glass slides and incubated overnight at 37C and 5% CO
2
. Adherent
FIGURE 1 Magnetic tweezer microrheology. Bead displacement
(geometric mean from >74 cells) versus time during force steps with
increasing force magnitude. Each step lasted 1 s. Numbers above the curve
indicate the lateral pulling force (in units of nN). After 10 s, the force was
reduced to zero. The differential cell stiffness (see Fig. 4 A) and the
exponent of the creep modulus (see Fig. 4 B) was computed as described
in Kollmannsberger et al. (26).
Biophysical Journal 101(9) 2131 2138
2132 Fabry et al.
cells were fixed for 20 min with 3% paraformaldehyde and lysed for 5 min
with 0.2% Triton X 100. Cells were then blocked for 1 h in 0.5% BSA and
PBS at room temperature. Primary antibodies (1:10
6
vinculin/hVIN 1;
Sigma, St. Louis, MO) and secondary antibodies (ml:200, FITC coupled
anti mouse IgG; Jackson ImmunoResearch, West Grove, PA) were both
diluted in 0.5% BSA and PBS at given ratios and incubated for 1 h at
room temperature each. Alexa Fluor546 Phalloidin (Molecular Probes,
Eugene, OR) was added simultaneously with the second antibody for actin
staining. Samples were mounted with Mowiol (Sigma) solution. Micros
copy was carried out on a DMI6000 microscope with a 63/1.3 NA
objective (Leica Microsystems, Wetzlar, Germany). Images were acquired
with a charge coupled device camera (ORCA ER; Hamamatsu, Hamamatsu
City, Japan).
Traction microscopy
This technique measures the forces that cells exert on their surroundings by
observing the displacements of beads embedded in a flexible gel substrate
on which the cells are cultured (32,34). Traction measurements were per
formed with 6.1% acrylamide/bisacrylamide (19:1) gels with 0.5 mm green
fluorescent beads. The Young’s modulus of the gels was 12.8 50.8 kPa
as measured from the linear extension of a cylinder of gel (16 mm diameter,
50 mm length) under force. Gels were coated with 50 mg bovine collagen G
(Biochrom AG, Berlin, Germany) diluted in 50 mM HEPES overnight at
4C. 10
4
cells were seeded on the gels and incubated under normal growth
conditions. During the measurements, the cells were maintained at 37C
and 5% CO
2
in a humidified atmosphere. Cell tractions were computed
from an unconstrained deconvolution of the gel surface displacement field
(32,35) measured before and after cells were detached from the substrate
with a cocktail of 80 mM cytochalasin D and 0.25% trypsin.
Statistical evaluation
FAK knockout and FAK silencing experiments were performed on MEF
cells derived from two different cell isolations. Therefore, for statistical
evaluation, we only compared FAKwt with FAK / cells, and siControl
cells with siFAK cells. Statistical significant differences were calculated
using a Student’s unpaired ttest, assuming unequal variances. Results
were considered to be significant and marked with an asterisk for p<
0.05. All data are expressed as arithmetic mean 5standard error of the
mean, except for bead detachment experiments (expressed as cumulative
probability) and for data that show a log normal distribution, namely cell
stiffness measured with magnetic tweezers and AFM, and apparent diffu
sivity (32,36). These data are expressed as geometric mean 5geometric
standard error of the mean. Number of measurements is between 46 and
184 cells for traction measurements, between 85 and 144 cells for adhesion
strength measurements, between 611 and 1031 beads for nanoscale particle
tracking experiments, between 10 and 30 cells for AFM measurements, and
between 74 and 129 cells for magnetic tweezer measurements.
RESULTS AND DISCUSSION
MEF cells, regardless of FAK expression levels, behave me-
chanically similar to other cells: they stiffen when probed
ABC
FIGURE 2 AFM force mapping. (A) Height.
Scan range: 80 mm80 mm. Color bar range
(black to white): 0 5 mm. (B) Corresponding shear
modulus. Grayscale range (black to white):
0 20 kPa. (C) Force versus z piezo extension
data at the position indicated (red cross) in panels
Aand B. A fit of the Hertz model to the data gives
the local shear modulus at this position.
Biophysical Journal 101(9) 2131 2138
FAK Stabilizes the Cytoskeleton 2133
at high forces (Fig. 4 A). Measurements with magnetic twee-
zers show that cell stiffness was approximately twofold
higher in FAKwt versus FAK/cells. Those differences
were significant (p<0.05) at all force levels. A similar trend
was observed in siControl versus siFAK knockdown cells,
although the differences were less pronounced and were
not consistently significant at all force levels.
To test whether these differences are attributable to the
mechanical behavior of the cytoskeleton, as opposed to
differences in FAK-mediated adhesion properties of the
fibronectin-coated magnetic beads, we performed stiffness
measurements using an atomic force microscope (AFM)
(Fig. 4 D). Because the cantilever tip of the AFM was
not functionalized and was in contact with the cell surface
for <1 s, these measurements are not influenced by specific
adhesion between the cell and the probe. The AFM
measurements confirm our results obtained with magnetic
tweezers, with a notable exception that the differences
between FAK-expressing and FAK-deficient cells were
even larger than those measured by magnetic tweezers.
The shear modulus magnitude from AFM measurements
approximately matched the values obtained with magnetic
tweezers (Fig. 4,Aand D). Differences are attributable to
the uncertainty in the geometric factor that is needed to
convert forces to stress, and displacements to strain
(27,37). Moreover, the shear modulus magnitude from
magnetic tweezer measurements reflects the combined
elastic and dissipative cell properties at a timescale of 1 s,
whereas the fit of the Hertz-model to the AFM force-inden-
tation data was performed under the assumptions of a purely
elastic response and a Poisson’s ratio of 0.5. These assump-
tions are reasonable because the cell’s elastic properties
dominate over dissipative properties, the timescale of the
AFM indentation measurements was also approximately
1 s, and the Poisson’s ratio has only a moderate effect on
the result.
In the magnetic tweezer experiments, the creep compli-
ance increased with time according to a power-law with
exponent bthat was significantly increased by ~50% in
FAK /cells compared to FAKwt cells (Fig. 4 B). A
similar trend was also seen in siControl versus siFAK cells,
although the differences were not always significant at all
force levels. The larger creep exponent in the FAK-deficient
cells indicates that the cytoskeletal structures that are
connected to the magnetic beads remodel more rapidly
and show a higher turnover compared to FAK-expressing
cells. This finding is somewhat unexpected, as the focal
adhesion complex, which is part of the structure probed
by the magnetic beads, has been previously reported to
be more stable in FAK-deficient cells (11,14). As a plausible
hypothesis, we suggest that not so much the focal adhesion
AD
B
C
E
F
FIGURE 4 FAK increases cell stiffness and adhesion strength, and
reduces cytoskeletal dynamics. Cell stiffness (magnitude of the shear
modulus) (A) and exponent of the creep modulus (B) measured with
magnetic tweezer microrheology at different forces. (C) Adhesion strength
of fibronectin coated beads expressed as percentage of detached beads
versus pulling force. (D) Cell stiffness (shear modulus) measured with an
uncoated AFM tip. Apparent diffusivity (E) and power law exponent of
the mean squared displacement (F) of spontaneous movements of
FN coated beads attached to the cell surface. (Bars) Standard error.
(Asterisk) Significant (p<0.05) differences.
FIGURE 3 Nanoscale particle tracking. Cytoskeletal remodeling
dynamics was estimated from the apparent diffusivity and the power law
exponent (see Fig. 4,Eand F) of the mean squared bead displacement
(geometric mean of >611 beads) versus time.
Biophysical Journal 101(9) 2131 2138
2134 Fabry et al.
complex, but the connected cytoskeletal structures are more
dynamic and unstable in FAK-deficient cells.
To test this hypothesis, we used the nanoscale particle
tracking technique. Fibronectin-coated fluorescent beads
are connected to the cytoskeleton via integrin-type cell
surface receptors, and their movements are followed over
time for several minutes. Although the type of probe is the
same as in the magnetic tweezer experiments, the properties
being measured are not. The magnetic beads are actively
forced and thus report the passive mechanical properties
of the cytoskeleton, such as elastic modulus and its time
(or frequency) dependency. In the particle tracking experi-
ments, the beads are passive in the sense that they are not
externally forced (except thermally) and report the active
material properties of the cytoskeleton. This is because the
beads act as fiducial markers of the cytoskeleton and
move only if the cytoskeletal structures to which they are
bound also move, for instance due to cytoskeletal remodel-
ing events (38), including those that lead to an overall cell
movement, and contractile force fluctuations (32). We find
an increased apparent diffusivity of the cytoskeleton-bound
beads in the FAK-deficient cells (Fig. 4 E), which is in
agreement with our hypothesis that the cytoskeleton of
FAK-deficient cells is less stable and more dynamic.
Additional support for this hypothesis comes from the
finding that nearly 50% of the magnetic beads attached to
FAK /cells detach at forces of 10 nN, as opposed to
only 10% of the beads that detach from FAKwt cells
(Fig. 4 C). Although we do not know whether the bead
detachment occurred due to a rupture of protein bonds in
the focal adhesion complex or due to a rupture of protein
bonds in the associated cytoskeleton, previous reports of
a more stable adhesion complex in FAK-deficient cells
(11,12,14) point to the cytoskeleton as the weakest link.
We noted that the fibronectin-coated fluorescent beads
moved more persistently and with a significantly larger
exponent bwhen attached to FAK-expressing cells as
compared to FAK-deficient cells (Fig. 4 F). The bcharacter-
izes the time evolution of the MSD of the beads and was
closer to a ballistic behavior in FAK-expressing cells, as
opposed to a more random behavior in FAK-deficient cells.
Because the beads generally follow and move along a path
that is determined by the cytoskeletal architecture, the
straighter and more persistent movement of the beads
attached to FAK-expressing cells suggests a more aligned
cytoskeleton, whereas the more random bead movement in
FAK-deficient cells in turn suggests a more isotropic
arrangement of the cytoskeleton.
Passive and active cytoskeletal properties are intricately
linked by universal scaling laws (39,40): a stiffer cytoskel-
eton is usually less frequency- or time-dependent (the
power-law exponent btends toward zero) (27), remodels
more slowly on short timescales (the apparent diffusivity
Dis smaller) (38) but more persistently (the MSD-exponent
bat longer timescales is higher) (32). A comparison of the
magnetic tweezer and particle tracking data (Fig. 4) shows
that FAK/and FAKwt cells obey these scaling laws.
Fluorescent images confirm, in agreement with previous
reports (12,15), that FAKwt cells have a highly ordered
and aligned actin cytoskeleton with prominent stress fibers,
whereas FAK/cells show a more pronounced cortical
actin cytoskeleton (Fig. 5 A). Such differences are not
visible in siControl versus siFAK cells, which is consistent
with the smaller differences that we found in all of the
mechanical parameters between these cells (Fig. 4). We
attribute these smaller mechanical differences and the lack
of any discernible differences in the cytoskeletal architec-
ture between siControl and siFAK cells to an incomplete
FAK knockdown, with expression levels of ~10% of base-
line FAK in the siRNA-silenced cells (data not shown). In
this regard, FAK may behave similarly to other focal
adhesion proteins. For example, a downregulation of the
focal adhesion protein vinculin to ~10% of control level is
sufficient for proper focal adhesion formation and mechan-
ical coupling (41,42), and only cells with full vinculin
knockout show substantial mechanical impairment.
It is known that cell stiffness scales linearly with cyto-
skeletal prestress, which in turn depends on the contractile
activation, cell area (both spreading area and cross-section
area), and cytoskeletal alignment (43). To test whether the
stiffness differences in cells with different FAK expression
levels are attributable to altered prestress, we measured
the traction forces of these cells (Fig. 5,Band C). As a scalar
value of cell tractions, we computed the elastic strain energy
that is stored in the matrix below the cell (35). Strain energy
was twofold decreased in FAK/cells (Fig. 5 D), but
this was solely due to a diminished spreading area of
the FAK/cells (Fig. 5 E). When strain energy was
normalized to spreading area (corresponding to an average
surface energy density) (Fig. 5 F), no difference could be
seen between FAKwt and FAK/cells. Consistent with
this finding, the maximum tractions and average traction
magnitude in all cells types were similar (Fig. 5 B), which
is also in line with previous reports (44,45). Moreover,
siControl and siFAK cells, which did not differ in their
spreading area (Fig. 5 E), showed no difference in strain
energy (Fig. 5 D). Taken together, these findings suggest
that the smaller stiffness of FAK-deficient cells was the
result of a reduced prestress (26,46), which in turn was
predominantly caused by a decreased spreading area and
therefore increased cross-sectional area. Note that this
interpretation rests on the assumption of a similar cell
volume in FAK-expressing and FAK-deficient cells.
Decreased stress fiber expression and a more pronounced
cortical organization in the actin cytoskeleton of FAK/
cells have been reported to result from rho-kinase, which
is known to be overactive in FAK/cells (12,13,15).
We confirmed, in line with previous reports (13), that
treatment of the FAK/cells with 10 mM of the ROCK-
inhibitor Y27632 for 30 min reestablishes a more wild-type
Biophysical Journal 101(9) 2131 2138
FAK Stabilizes the Cytoskeleton 2135
fibroblast cell morphology (data not shown). Y27632-treat-
ment not only affected the organization of the actomyosin
cytoskeleton, but it partially rescued the mechanical proper-
ties of the FAK/cells toward a FAK wild-type pheno-
type. The spontaneous motion of beads bound to FAK/
cells became significantly less diffusive and more directed
after Y27632-treatment, and the cell stiffness measured by
AFM increased (Fig. 6 A). Remarkably, FAKwt cells
responded to ROCK-inhibition with Y27632 with the exact
opposite behavior: lower cell stiffness and more diffusive,
less directed bead motion.
On a cell-by-cell basis, the stiffness change in FAK/
cells after Y27632-treatment was highly correlated (r
2
¼
0.51) with the stiffness before Y27632-treatment (Fig. 6 B,
and see Fig. S1 in the Supporting Material). Soft cells
stiffened after treatment with the ROCK-inhibitor Y27632,
whereas stiff cells softened. FAKwt cells, which were
generally stiffer than FAK/cells, tended to soften after
ROCK inhibition regardless of baseline stiffness (Fig. 6 B,
and see Fig. S1). This observation, together with the fact
that ROCK is more active in FAK/cells, suggests
that cytoskeletal architecture, mechanics, and dynamics
can change in response to altered ROCK activity only
along a specific trajectory, as conceptually outlined in
Fig. 6 C. Accordingly, cell stiffness, stress fiber formation,
cytoskeletal stability, and spreading area are highest at an
intermediate ROCK activity level as is present in normal
(FAKwt) cells. At lower or higher levels of ROCK activity,
all those parameters decrease. The question then arises
how FAK and ROCK are connected such that they alter cyto-
skeletal architecture and mechanics only along a specific
trajectory.
ABC
D
E
F
FIGURE 5 Traction force magnitude is not affected by FAK. (A) Fluorescence images (actin, red; vinculin, green) of cells plated on glass. (B) Traction
maps and corresponding (C) bright field images of cells on polyacrylamide gels. (Red dashed line) Cell outlines. Strain energy (D) and spreading area (E)
are reduced in FAK / cells, but the average traction force magnitude and strain energy normalized to the spreading area (F) are not dependent on FAK
expression levels.
Biophysical Journal 101(9) 2131 2138
2136 Fabry et al.
ROCK is activated by rhoA-kinase, which in turn is regu-
lated by FAK via rho-GEFs (activator) and rho-GAPs (inac-
tivator) (18). This gives FAK a dual role that is crucial for
the spatiotemporal regulation of rhoA-kinase and ROCK
activity (3,4,18). ROCK is known to regulate intermediate
filament assembly, actin polymerization via LIM-kinase,
cortical actin distribution via adducin (47), and actomyosin
contraction via myosin light chain (MLC) phosphatase inhi-
bition and direct MLC phosphorylation (48 50).
In our experiments, it was expected that inhibition of
rho-kinase by Y27632 decreases MLC-dependent contrac-
tility and, therefore cell stiffness. FAKwt cells behaved as
expected, but FAK/cells did not. The restructuring
of the cytoskeleton toward a more mesenchymal phenotype
in FAK/cells after ROCK inhibition seemed to dominate
over any decrease in actomyosin contractility. We speculate
that this may be because ROCK activates the trans-
membrane actin binding proteins ezrin, radixin, and moesin
(ERM) that are important for the cortical distribution of
actin (51 53). Because ROCK is overactive in the FAK/
cells, it could induce an overactivation of the ERM proteins
and thereby induce the binding of cortical actin bundles to
the plasma membrane. To validate the involvement of
ERM proteins, however, further work is needed.
Regardless of the molecular mechanisms, our data show
that FAK stabilizes the actin cytoskeleton through
a ROCK-mediated pathway.
SUPPORTING MATERIAL
One figure is available at http://www.biophysj.org/biophysj/supplemental/
S0006 3495(11)01134 9.
We thank Drs. Wolfgang H. Ziegler, Staffan Johansson, Bernd Hoffmann,
Gerold Diez, Carina Raupach, Philip Kollmannsberger, Jose Luis Alonso,
Daniel Paranhos Zitterbart, and Thorsten Koch for help with experiments
and stimulating discussions.
This work was supported by grants from Bayerische Forschungsallianz;
Deutscher Akademischer Austauschdienst; Bavaria California Technology
Center; National Institutes of Health (NIH HL65960); and Deutsche
Forschungsgemeinschaft. A.H.K. has been supported by a grant from the
University of Erlangen Nuremberg.
A.H.K. and W.G. designed the study, A.H.K. and S.K. performed and
analyzed the experiments, T.S. and B.F. developed the methods, and B.F.,
A.H.K. and W.G. wrote the manuscript.
REFERENCES
1. Parsons, J. T., K. H. Martin, ., S. A. Weed. 2000. Focal adhesion
kinase: a regulator of focal adhesion dynamics and cell movement.
Oncogene. 19:5606 5613.
2. Schlaepfer, D. D., S. K. Mitra, and D. Ilic. 2004. Control of motile and
invasive cell phenotypes by focal adhesion kinase. Biochim. Biophys.
Acta. 1692:77 102.
3. Tilghman, R. W., J. K. Slack Davis, ., J. T. Parsons. 2005. Focal adhe
sion kinase is required for the spatial organization of the leading edge
in migrating cells. J. Cell Sci. 118:2613 2623.
4. Mitra, S. K., D. A. Hanson, and D. D. Schlaepfer. 2005. Focal adhesion
kinase: in command and control of cell motility. Nat. Rev. Mol. Cell
Biol. 6:56 68.
5. Stamenovi
c, D. 2008. Rheological behavior of mammalian cells. Cell.
Mol. Life Sci. 65:3592 3605.
6. Geiger, B., J. P. Spatz, and A. D. Bershadsky. 2009. Environmental
sensing through focal adhesions. Nat. Rev. Mol. Cell Biol. 10:21 33.
7. Giancotti, F. G., and E. Ruoslahti. 1999. Integrin signaling. Science.
285:1028 1032.
8. Turner, C. E. 2000. Paxillin interactions. J. Cell Sci. 113:4139 4140.
9. Hanks, S. K., and T. R. Polte. 1997. Signaling through focal adhesion
kinase. Bioessays. 19:137 145.
10. Izaguirre, G., L. Aguirre, ., B. Haimovich. 2001. The cytoskeletal/
non muscle isoform of aactinin is phosphorylated on its actin binding
domain by the focal adhesion kinase. J. Biol. Chem. 276:28676 28685.
11. Ili
c, D., Y. Furuta, ., T. Yamamoto. 1995. Reduced cell motility and
enhanced focal adhesion contact formation in cells from FAK deficient
mice. Nature. 377:539 544.
12. Ren, X. D., W. B. Kiosses, ., M. A. Schwartz. 2000. Focal adhesion
kinase suppresses Rho activity to promote focal adhesion turnover.
J. Cell Sci. 113:3673 3678.
13. Chen, B. H., J. T. Tzen, ., H. C. Chen. 2002. Roles of Rho associated
kinase and myosin light chain kinase in morphological and migratory
defects of focal adhesion kinase null cells. J. Biol. Chem.
277:33857 33863.
14. Webb, D. J., K. Donais, ., A. F. Horwitz. 2004. FAK Src signaling
through paxillin, ERK and MLCK regulates adhesion disassembly.
Nat. Cell Biol. 6:154 161.
15. Schober, M., S. Raghavan, ., E. Fuchs. 2007. Focal adhesion kinase
modulates tension signaling to control actin and focal adhesion
dynamics. J. Cell Biol. 176:667 680.
16. Raftopoulou, M., and A. Hall. 2004. Cell migration: Rho GTPases lead
the way. Dev. Biol. 265:23 32.
A
BC
FIGURE 6 ROCK inhibitor has diverging effects on cell mechanical
behavior in FAKwt versus FAK / cells. (A) Change in cell stiffness,
bead diffusivity, and persistence of bead motion after ROCK inhibition
with 10 mM Y27632 for 30 min (relative to untreated cells, dashed line).
(B) Stiffness of individual cells measured with AFM before and after
ROCK inhibition with Y27632. (Shaded dashed line) Line of identity.
(C) Schematic representation of the proposed molecular pathway: FAK
exerts its influence on cytoskeletal stiffness, stability, and morphology
through modulation of the expression and activity level of ROCK.
(Arrowheads) Expected changes of cell stiffness and cytoskeletal properties
after ROCK inhibition with Y27632 along a specific trajectory.
Biophysical Journal 101(9) 2131 2138
FAK Stabilizes the Cytoskeleton 2137
17. Michael, K. E., D. W. Dumbauld, ., A. J. Garcı
´a. 2009. Focal
adhesion kinase modulates cell adhesion strengthening via integrin
activation. Mol. Biol. Cell. 20:2508 2519.
18. Tomar, A., and D. D. Schlaepfer. 2009. Focal adhesion kinase: switch
ing between GAPs and GEFs in the regulation of cell motility. Curr.
Opin. Cell Biol. 21:676 683.
19. Wang, H. B., M. Dembo, ., Y. Wang. 2001. Focal adhesion kinase
is involved in mechanosensing during fibroblast migration. Proc.
Natl. Acad. Sci. USA. 98:11295 11300.
20. Katoh, K., Y. Kano, ., K. Fujiwara. 2001. Stress fiber organization
regulated by MLCK and rho kinase in cultured human fibroblasts.
Am. J. Physiol. Cell Physiol. 280:C1669 C1679.
21. Sieg, D. J., D. Ili
c, ., D. D. Schlaepfer. 1998. Pyk2 and Src family
protein tyrosine kinases compensate for the loss of FAK in fibro
nectin stimulated signaling events but Pyk2 does not fully function
to enhance FAK cell migration. EMBO J. 17:5933 5947.
22. Klemm, A. H., S. Kienle, ., W. H. Goldmann. 2010. The influence
of Pyk2 on the mechanical properties in fibroblasts. Biochem. Biophys.
Res. Commun. 393:694 697.
23. Klemm, A. H., G. Diez, ., W. H. Goldmann. 2009. Comparing the
mechanical influence of vinculin, focal adhesion kinase and p53 in
mouse embryonic fibroblasts. Biochem. Biophys. Res. Commun.
379:799 801.
24. Kollmannsberger, P., and B. Fabry. 2007. High force magnetic twee
zers with force feedback for biological applications. Rev. Sci. Instrum.
78:114301 114306.
25. Kasza, K. E., F. Nakamura, ., D. A. Weitz. 2009. Filamin A is essen
tial for active cell stiffening but not passive stiffening under external
force. Biophys. J. 96:4326 4335.
26. Kollmannsberger, P., C. T. Mierke, and B. Fabry. 2011. Nonlinear
viscoelasticity of adherent cells is controlled by cytoskeletal tension.
Soft Matter. 7:3127 3132.
27. Fabry, B., G. N. Maksym, ., J. J. Fredberg. 2001. Scaling the micro
rheology of living cells. Phys. Rev. Lett. 87:148102.
28. Hildebrandt, J. 1969. Comparison of mathematical models for cat lung
and viscoelastic balloon derived by Laplace transform methods from
pressure volume data. Bull. Math. Biophys. 31:651 667.
29. Cook, S. M., T. E. Schaffer, ., K. M. Lang. 2006. Practical implemen
tation of dynamic methods for measuring atomic force microscope
cantilever spring constants. Nanotechnology. 17:2135 2145.
30. Jiao, Y., and T. E. Schaffer. 2004. Accurate height and volume
measurements on soft samples with the atomic force microscope.
Langmuir. 20:10038 10045.
31. Sneddon, I. N. 1965. The relation between load and penetration in the
axisymmetric Boussinesq problem for a punch of arbitrary profile. Int.
J. Eng. Sci. 3:47 57.
32. Raupach, C., D. P. Zitterbart, ., B. Fabry. 2007. Stress fluctuations and
motion of cytoskeletal bound markers. Phys. Rev. E. 76:011918.
33. Metzner, C., C. Raupach, ., B. Fabry. 2010. Fluctuations of cytoskel
eton bound microbeads the effect of bead receptor binding
dynamics. J. Phys. Condens. Matter. 22:194105.
34. Pelham, Jr., R. J., and Y. Wang. 1997. Cell locomotion and focal adhe
sions are regulated by substrate flexibility. Proc. Natl. Acad. Sci. USA.
94:13661 13665.
35. Butler, J. P., I. M. Toli
c Nørrelykke, ., J. J. Fredberg. 2002. Traction
fields, moments, and strain energy that cells exert on their surround
ings. Am. J. Physiol. Cell Physiol. 282:C595 C605.
36. Fabry, B., G. N. Maksym, ., J. J. Fredberg. 2001. Selected contribu
tion: time course and heterogeneity of contractile responses in cultured
human airway smooth muscle cells. J. Appl. Physiol. 91:986 994.
37. Mijailovich, S. M., M. Kojic, ., J. J. Fredberg. 2002. A finite element
model of cell deformation during magnetic bead twisting. J. Appl.
Physiol. 93:1429 1436.
38. Bursac, P., G. Lenormand, ., J. J. Fredberg. 2005. Cytoskeletal re
modeling and slow dynamics in the living cell. Nat. Mater. 4:557 561.
39. Trepat, X., G. Lenormand, and J. J. Fredberg. 2008. Universality in cell
mechanics. Soft Matter. 4:1750 1759.
40. Kollmannsberger, P., and B. Fabry. 2011. Linear and nonlinear
rheology of living cells. Annu. Rev. Mater. Res. 41:75 97.
41. Goldmann, W. H., R. Galneder, ., R. M. Ezzell. 1998. Differences in
elasticity of vinculin deficient F9 cells measured by magnetometry and
atomic force microscopy. Exp. Cell Res. 239:235 242.
42. Xu, W., J. L. Coll, and E. D. Adamson. 1998. Rescue of the mutant
phenotype by reexpression of full length vinculin in null F9 cells;
effects on cell locomotion by domain deleted vinculin. J. Cell Sci.
111:1535 1544.
43. Wang, N., K. Naruse, ., D. E. Ingber. 2001. Mechanical behavior in
living cells consistent with the tensegrity model. Proc. Natl. Acad.
Sci. USA. 98:7765 7770.
44. Wang, J. G., M. Miyazu, ., K. Naruse. 2001. Uniaxial cyclic stretch
induces focal adhesion kinase (FAK) tyrosine phosphorylation fol
lowed by mitogen activated protein kinase (MAPK) activation. Bio
chem. Biophys. Res. Commun. 288:356 361.
45. Pirone, D. M., W. F. Liu, ., C. S. Chen. 2006. An inhibitory role for
FAK in regulating proliferation: a link between limited adhesion and
RhoA ROCK signaling. J. Cell Biol. 174:277 288.
46. Wang, N., I. M. Toli
c Nørrelykke, ., D. Stamenovi
c. 2002. Cell
prestress. I. Stiffness and prestress are closely associated in adherent
contractile cells. Am. J. Physiol. Cell Physiol. 282:C606 C616.
47. Fukata, Y., M. Amano, and K. Kaibuchi. 2001. Rho Rho kinase
pathway in smooth muscle contraction and cytoskeletal reorganization
of non muscle cells. Trends Pharmacol. Sci. 22:32 39.
48. Kimura, K., M. Ito, ., K. Kaibuchi. 1996. Regulation of myosin phos
phatase by Rho and Rho associated kinase (rkinase). Science.
273:245 248.
49. Amano, M., H. Mukai, ., K. Kaibuchi. 1996. Identification of a puta
tive target for Rho as the serine threonine kinase protein kinase N.
Science. 271:648 650.
50. Geiger, B., and A. Bershadsky. 2001. Assembly and mechanosensory
function of focal contacts. Curr. Opin. Cell Biol. 13:584 592.
51. Arpin, M., M. Algrain, and D. Louvard. 1994. Membrane actin micro
filament connections: an increasing diversity of players related to band
4.1. Curr. Opin. Cell Biol. 6:136 141.
52. Oshiro, N., Y. Fukata, and K. Kaibuchi. 1998. Phosphorylation of moe
sin by Rho associated kinase (rkinase) plays a crucial role in the
formation of microvilli like structures. J. Biol. Chem. 273:34663
34666.
53. Niggli, V., and J. Rossy. 2008. Ezrin/radixin/moesin: versatile control
lers of signaling molecules and of the cortical cytoskeleton. Int. J. Bio
chem. Cell Biol. 40:344 349.
Biophysical Journal 101(9) 2131 2138
2138 Fabry et al.
... Hence, we propose the cytoskeletal rearrangement of TMCs as an important means for responding to the shear stress and regulating the aqueous humor outflow. Existing studies have proved that the shear stress can cause cytoskeletal arrangement for several different cell types (Galbraith et al., 1998;Kadi et al., 2007;Huang et al., 2010;Risca et al., 2012;Cheng et al., 2013;Molladavoodi et al., 2017;Son et al., 2020), and the changes in cytoskeleton and cellular functions after shear stress stimulation may be mediated by FAK (Girard and Nerem, 1995;Fabry et al., 2011;Cheng et al., 2013;Sun et al., 2018), ERK pathway (Fabry et al., 2011;Sun et al., 2018), Rho pathway (Tzima et al., 2002), and transient receptor potential melastatin type 7 (TRPM7) channel (Liu et al., 2015;Xiao et al., 2015). These uncovered mechanisms may also hold for the case of TMCs. ...
... Hence, we propose the cytoskeletal rearrangement of TMCs as an important means for responding to the shear stress and regulating the aqueous humor outflow. Existing studies have proved that the shear stress can cause cytoskeletal arrangement for several different cell types (Galbraith et al., 1998;Kadi et al., 2007;Huang et al., 2010;Risca et al., 2012;Cheng et al., 2013;Molladavoodi et al., 2017;Son et al., 2020), and the changes in cytoskeleton and cellular functions after shear stress stimulation may be mediated by FAK (Girard and Nerem, 1995;Fabry et al., 2011;Cheng et al., 2013;Sun et al., 2018), ERK pathway (Fabry et al., 2011;Sun et al., 2018), Rho pathway (Tzima et al., 2002), and transient receptor potential melastatin type 7 (TRPM7) channel (Liu et al., 2015;Xiao et al., 2015). These uncovered mechanisms may also hold for the case of TMCs. ...
Article
Full-text available
Mechanical microenvironment and cellular senescence of trabecular meshwork cells (TMCs) are suspected to play a vital role in primary open-angle glaucoma pathogenesis. However, central questions remain about the effect of shear stress on TMCs and how aging affects this process. We have investigated the effect of shear stress on the biomechanical properties and extracellular matrix regulation of normal and senescent TMCs. We found a more significant promotion of Fctin formation, a more obvious realignment of F-actin fibers, and a more remarkable increase in the stiffness of normal cells in response to the shear stress, in comparison with that of senescent cells. Further, as compared to normal cells, senescent cells show a reduced extracellular matrix turnover after shear stress stimulation, which might be attributed to the different phosphorylation levels of the extracellular signal-regulated kinase. Our results suggest that TMCs are able to sense and respond to the shear stress and cellular senescence undermines the mechanobiological response, which may lead to progressive failure of cellular TM function with age.
... It is known that the cell adhesion process can occur in substrates and components of the extracellular matrix (ECM) upon extracellular integrins domains, which modifies its intracellular portion and creates points of interaction with intracellular proteins, such as Focal Adhesion Kinases (FAK) and Src [22][23][24][25][26][27][28]. This supramolecular platform promotes an adaptation in the cytoskeleton structures and establishes important parameters regarding the survival of cells, such as adhesion, proliferation, migration, and differentiation [29,30]. Considering focal adhesion kinase (FAK) and Src, the phosphorylation cascade is modulated by the capacity of those cells to guide the ECM components' breakdown via specific matrix metalloproteinases and cathepsins activities, culminating immediately in the rearrangement of the cytoskeleton [9,11,31,32]. ...
Article
Full-text available
Since Branemark’s findings, titanium-based alloys have been widely used in implantology. However, their success in dental implants is not known when considering the heterogenicity of housing cells surrounding the peri-implant microenvironment. Additionally, they are expected to recapitulate the physiological coupling between endothelial cells and osteoblasts during appositional bone growth during osseointegration. To investigate whether this crosstalk was happening in this context, we considered the mechanotransduction-related endothelial cell signaling underlying laminar shear stress (up to 3 days), and this angiocrine factor-enriched medium was harvested further to use exposing pre-osteoblasts (pOb) for up to 7 days in vitro. Two titanium surfaces were considered, as follows: double acid etching treatment (w_DAE) and machined surfaces (wo_DAE). These surfaces were used to conditionate the cell culture medium as recommended by ISO10993-5:2016, and this titanium-enriched medium was later used to expose ECs. First, our data showed that there is a difference between the surfaces in releasing Ti molecules to the medium, providing very dynamic surfaces, where the w_DAE was around 25% higher (4 ng/mL) in comparison to the wo_DAE (3 ng/mL). Importantly, the ECs took up some of this titanium content for up to 3 days in culture. However, when this conditioned medium was used to expose pOb for up to 7 days, considering the angiocrine factors released from ECs, the concentration of Ti was lesser than previously reported, reaching around 1 ng/mL and 2 ng/mL, respectively. Thereafter, pOb exposed to this angiocrine factor-enriched medium presented a significant difference when considering the mechanosignaling subjected to the ECs. Shear-stressed ECs showed adequate crosstalk with osteoblasts, stimulating the higher expression of the Runx2 gene and driving higher expressions of Alkaline phosphatase (ALP), bone sialoprotein (BSP), and osteocalcin. Mechanotransduction-related endothelial cell signaling as a source of angiocrine molecules also stimulated the higher expression of the Col3A1 gene in osteoblasts, which suggests it is a relevant protagonist during trabecular bone growth. In fact, we investigated ECM remodeling by first evaluating the expression of genes related to it, and our data showed a higher expression of matrix metalloproteinase (MMP) 2 and MMP9 in response to mechanosignaling-based angiocrine molecules, independent of considering w_DAE or the wo_DAE, and this profile reflected on the MMP2 and MMP9 activities evaluated via gelatin-based zymography. Complimentarily, the ECM remodeling seemed to be a very regulated mechanism in mature osteoblasts during the mineralization process once both TIMP metallopeptidase inhibitor 1 and 2 (TIMP1 and TIMP2, respectively) genes were significantly higher in response to mechanotransduction-related endothelial cell signaling as a source of angiocrine molecules. Altogether, our data show the relevance of mechanosignaling in favoring ECs’ release of bioactive factors peri-implant, which is responsible for creating an osteogenic microenvironment able to drive osteoblast differentiation and modulate ECM remodeling. Taking this into account, it seems that mechanotransduction-based angiocrine molecules explain the successful use of titanium during osseointegration.
... Black triangles in (e) and (f) denote ON and OFF for each actuation cycle.response observed for a 4.5 μm-diameter fibronectin-coated Dynabead™ that has been attached to a murine embryonic fibroblast and actuated with an MT device[214]. In comparing Δ MT ( ) to Δ max ( ), we note that the observed difference between Δ MT ( OFF ) and Δ max ( OFF ) for each of the three actuation cycles varies from ~0.4 μm to ~0.5 μm (Figure 5.3(e)). ...
... Black triangles in (e) and (f) denote ON and OFF for each actuation cycle.response observed for a 4.5 μm-diameter fibronectin-coated Dynabead™ that has been attached to a murine embryonic fibroblast and actuated with an MT device[214]. In comparing Δ MT ( ) to Δ max ( ), we note that the observed difference between Δ MT ( OFF ) and Δ max ( OFF ) for each of the three actuation cycles varies from ~0.4 μm to ~0.5 μm (Figure 5.3(e)). ...
... Focal adhesion kinase (FAK), an intracellular kinase that binds to CD44 and integrins, also decreased, and signals for the phosphorylated forms of FAK (Tyr397, Tyr925) and total FAK protein reduced in a similar manner at 24 and 72 h after h4 # 147D administration (Figures 3(f )-3(h)). Based on the observed inhibition of CD44/integrin α3/integrin α6/FAK in tumors with h4 # 147D treatment, h4 # 147D was considered to have caused cytoskeletal stress-mediated cell death in the tumor because these molecules are crucial for the organization of cytoskeletal proteins [40][41][42]. ...
Article
Full-text available
CD147 is an immunoglobulin-like receptor that is highly expressed in various cancers and involved in the growth, metastasis, and activation of inflammatory pathways via interactions with various functional molecules, such as integrins, CD44, and monocarboxylate transporters. Through screening of CD147-targeting antibodies with antitumor efficacy, we discovered a novel rat monoclonal antibody #147D. This humanized IgG4-formatted antibody, h4#147D, showed potent antitumor efficacy in xenograft mouse models harboring the human PDAC cell line MIA PaCa-2, HCC cell line Hep G2, and CML cell line KU812, which featured low sensitivity to the corresponding standard-of-care drugs (gemcitabine, sorafenib, and imatinib, respectively). An analysis of tumor cells derived from MIA PaCa-2 xenograft mice treated with h4#147D revealed that cell surface expression of CD147 and its binding partners, including CD44 and integrin α3β1/α6β1, was significantly reduced by h4#147D. Inhibition of focal adhesion kinase (FAK), activation of multiple stress responsible signal proteins such as c-JunN-terminal kinase (JNK) and mitogen-activated protein kinase p38 (p38MAPK), and expression of SMAD4, as well as activation of caspase-3 were obviously observed in the tumor cells, suggesting that h4#147D induced tumor shrinkage by inducing multiple stress responsible signals. These results suggest that the anti-CD147 antibody h4#147D offers promise as a new antibody drug candidate.
Article
Full-text available
Intercellular communication is critical to the formation and homeostatic function of all tissues. Previous work has shown that cells can communicate mechanically via the transmission of cell‐generated forces through their surrounding extracellular matrix, but this process is not well understood. Here, mechanically defined, synthetic electrospun fibrous matrices are utilized in conjunction with a microfabrication‐based cell patterning approach to examine mechanical intercellular communication (MIC) between endothelial cells (ECs) during their assembly into interconnected multicellular networks. It is found that cell force‐mediated matrix displacements in deformable fibrous matrices underly directional extension and migration of neighboring ECs toward each other prior to the formation of stable cell‐cell connections enriched with vascular endothelial cadherin (VE‐cadherin). A critical role is also identified for calcium signaling mediated by focal adhesion kinase and mechanosensitive ion channels in MIC that extends to multicellular assembly of 3D vessel‐like networks when ECs are embedded within fibrin hydrogels. These results illustrate a role for cell‐generated forces and ECM mechanical properties in multicellular assembly of capillary‐like EC networks and motivates the design of biomaterials that promote MIC for vascular tissue engineering.
Preprint
Full-text available
Individuals with Junctional Epidermolysis Bullosa (JEB), a rare genetic skin disease characterised by loss of function mutations in the Laminin332 (Lam332), do not survive beyond their first birthday. Here we report that loss of Lam332 leads to absence of cholesterol lipid from the epidermis in vitro and in vivo. Stable knockdown of Lam332 chains (LAMA3, LAMB3 and LAMC2) was established using shRNA and were used to develop 3D skin equivalents. Changes in lipid synthesis were assessed by western blot analysis and immunohistochemistry. Findings were confirmed in an inducible mouse model of Lamα3 ( Lama3 flox/flox/K14CreERT ) and in anonymized, archival human tissue: JEB skin and normal age-matched controls. Further lipid analysis was explored using lipidomics in 3D skin equivalents and mouse tissue. Cholesterol biosynthesis genes were increased with loss of Lam332 in vitro, however a decrease in nile red lipid staining was observed in Lamα3 mouse (n = 6) and in JEB patient skin (n = 7). Further changes to the epidermal lipid profile with loss of Lam332 was confirmed with lipidomic analysis of Lamα3 mouse epidermis and Lam332 skin equivalents. Cholesterol transport within Lam332 KD keratinocytes was revealed to be disrupted, which in keratinocytes is dependent on the actomyosin network, which was reversed with recombinant human Lam332. In conclusion these findings suggest a role for Lam332 in lipid metabolism in the skin and a broader role in epidermal homeostasis and barrier formation. Restoration of cholesterol transport in JEB patients offers the potential to improve the skin barrier and survival.
Article
Full-text available
Skeletal stem and progenitor cells (SSPCs) are the multi-potent, self-renewing cell lineages that form the hematopoietic environment and adventitial structures of the skeletal tissues. Skeletal tissues are responsible for a diverse range of physiological functions because of the extensive differentiation potential of SSPCs. The differentiation fates of SSPCs are shaped by the physical properties of their surrounding microenvironment and the mechanical loading forces exerted on them within the skeletal system. In this context, the present review first highlights important biomolecules involved with the mechanobiology of how SSPCs sense and transduce these physical signals. The review then shifts focus towards how the static and dynamic physical properties of microenvironments direct the biological fates of SSPCs, specifically within biomaterial and tissue engineering systems. Biomaterial constructs possess designable, quantifiable physical properties that enable the growth of cells in controlled physical environments both in-vitro and in-vivo . The utilization of biomaterials in tissue engineering systems provides a valuable platform for controllably directing the fates of SSPCs with physical signals as a tool for mechanobiology investigations and as a template for guiding skeletal tissue regeneration. It is paramount to study this mechanobiology and account for these mechanics-mediated behaviors to develop next-generation tissue engineering therapies that synergistically combine physical and chemical signals to direct cell fate. Ultimately, taking advantage of the evolved mechanobiology of SSPCs with customizable biomaterial constructs presents a powerful method to predictably guide bone and skeletal organ regeneration.
Article
Lysophosphatidic acid (LPA) is a lysophospholipid that signals through six G-protein coupled receptors (LPARs), LPA1 to LPA6. LPA has been described as a potent modulator of fibrosis in different pathologies. In skeletal muscle, LPA increases fibrosis-related proteins and the number of fibro/adipogenic progenitors (FAPs). FAPs are the primary source of ECM-secreting myofibroblasts in acute and chronic damage. However, the effect of LPA on FAPs activation in vitro has not been explored. This study aimed to investigate FAPs' response to LPA and the downstream signaling mediators involved. Here, we demonstrated that LPA mediates FAPs activation by increasing their proliferation, expression of myofibroblasts markers, and upregulation of fibrosis-related proteins. Pretreatment with the LPA1/LPA3 antagonist Ki16425 or genetic deletion of LPA1 attenuated the LPA-induced FAPs activation, resulting in decreased expression of cyclin e1, α-SMA, and fibronectin. We also evaluated the activation of the focal adhesion kinase (FAK) in response to LPA. Our results showed that LPA induces FAK phosphorylation in FAPs. Treatment with the P-FAK inhibitor PF-228 partially prevented the induction of cell responses involved in FAPs activation, suggesting that this pathway mediates LPA signaling. FAK activation controls downstream cell signaling within the cytoplasm, such as the Hippo pathway. LPA induced the dephosphorylation of the transcriptional coactivator YAP (Yes-associated protein) and promoted direct expression of target pathway genes such as Ctgf/Ccn2 and Ccn1. The blockage of YAP transcriptional activity with Super-TDU further confirmed the role of YAP in LPA-induced FAPs activation. Finally, we demonstrated that FAK is required for LPA-dependent YAP dephosphorylation and the induction of Hippo pathway target genes. In conclusion, LPA signals through LPA1 to regulate FAPs activation by activating FAK to control the Hippo pathway.
Article
The tensegrity hypothesis holds that the cytoskeleton is a structure whose shape is stabilized predominantly by the tensile stresses borne by filamentous structures. Accordingly, cell stiffness must increase in proportion with the level of the tensile stress, which is called the prestress. Here we have tested that prediction in adherent human airway smooth muscle (HASM) cells. Traction microscopy was used to measure the distribution of contractile stresses arising at the interface between each cell and its substrate; this distribution is called the traction field. Because the traction field must be balanced by tensile stresses within the cell body, the prestress could be computed. Cell stiffness (G) was measured by oscillatory magnetic twisting cytometry. As the contractile state of the cell was modulated with graded concentrations of relaxing or contracting agonists (isoproterenol or histamine, respectively), the mean prestress (<(p)overbar>(t)) ranged from 350 to 1,900 Pa. Over that range, cell stiffness increased linearly with the prestress: G (Pa) = 0.18 <(p)overbar>(t) + 92. While this association does not necessarily preclude other interpretations, it is the hallmark of systems that secure shape stability mainly through the prestress. Regardless of mechanism, these data establish a strong association between stiffness of HASM cells and the level of tensile stress within the cytoskeleton.
Article
Responses of cells to mechanical properties of the adhesion substrate were examined by culturing normal rat kidney epithelial and 3T3 fibroblastic cells on a collagen-coated polyacrylamide substrate that allows the flexibility to be varied while maintaining a constant chemical environment. Compared with cells on rigid substrates, those on flexible substrates showed reduced spreading and increased rates of motility or lamellipodial activity. Microinjection of fluorescent vinculin indicated that focal adhesions on flexible substrates were irregularly shaped and highly dynamic whereas those on firm substrates had a normal morphology and were much more stable. Cells on flexible substrates also contained a reduced amount of phosphotyrosine at adhesion sites. Treatment of these cells with phenylarsine oxide, a tyrosine phosphatase inhibitor, induced the formation of normal, stable focal adhesions similar to those on firm substrates. Conversely, treatment of cells on firm substrates with myosin inhibitors 2,3-butanedione monoxime or KT5926 caused the reduction of both vinculin and phosphotyrosine at adhesion sites. These results demonstrate the ability of cells to survey the mechanical properties of their surrounding environment and suggest the possible involvement of both protein tyrosine phosphorylation and myosin-generated cortical forces in this process. Such response to physical parameters likely represents an important mechanism of cellular interaction with the surrounding environment within a complex organism.
Article
The viscoelastic response of living cells to small external forces and deformations is characterized by a weak power law in time. The elastic modulus of cells and the power law exponent with which the elastic stresses decay depend on the active contractile prestress in the cytoskeleton. It is unknown whether this also holds in the physiologically relevant regime of large external forces and deformations. We used magnetic tweezers to apply stepwise increasing forces to magnetic beads bound to the cytoskeleton of different cell lines, and recorded the resulting cell deformation (creep response). The creep response followed a weak power law at all force levels. Stiffness and power law exponent increased with force in all cells, indicating simultaneous stress stiffening and fluidization of the cytoskeleton. The amount of stress stiffening and fluidization differed greatly between cell types but scaled with the contractile prestress as the only free parameter. Our results demonstrate that by modulating the internal mechanical tension, cells can actively control their mechanical properties over an exceedingly large range. This behavior is of fundamental importance for protection against damage caused by large external forces, and allows the cells to adapt to the highly variable and nonlinear mechanical properties of the extracellular matrix.
Article
The cytoskeleton (CSK) of the adherent living cell is arguably the most complex form of soft matter that exists in nature. It is constituted by hundreds of different proteins that interact with each other in a highly specific manner and, as a requirement for life, exists out of thermodynamic equilibrium and in a constant state of remodeling. While such structural and dynamical complexity may have conferred the cell with diverse and unpredictable mechanical properties, recent evidence indicates that the behavior of the CSK conforms to a limited set of empirical laws that appear to be simple and universal. While mechanistic understanding of such laws is still lacking, their very existence suggests that rather than being addressed solely in terms of molecular details and specific interactions, cell mechanics need to be addressed also from an integrative point of view.
Article
Living cells are an active soft material with fascinating mechanical properties. Under mechanical loading, cells exhibit creep and stress relaxation behavior that follows a power-law response rather than a classical exponential response. Such a response puts cells in the context of soft colloidal glasses and other disordered metastable materials that share the same properties. In cells, however, both the power-law exponent and stiffness are related to the contractile prestress in the cytoskeleton. In addition, cells are made of a highly nonlinear material that stiffens and fluidizes under mechanical stress. They show active and adaptive mechanical behavior such as contraction and remodeling that sets them apart from any other nonliving material. Strikingly, all these observations can be linked by simple relationships with the power-law exponent as the only organizing parameter. Current theoretical models capture specific facets of cell mechanical behavior, but a comprehensive understanding is still emerging.
Article
Cell migration is a complex, highly regulated process that involves the continuous formation and disassembly of adhesions (adhesion turnover). Adhesion formation takes place at the leading edge of protrusions, whereas disassembly occurs both at the cell rear and at the base of protrusions. Despite the importance of these processes in migration, the mechanisms that regulate adhesion formation and disassembly remain largely unknown. Here we develop quantitative assays to measure the rate of incorporation of molecules into adhesions and the departure of these proteins from adhesions. Using these assays, we show that kinases and adaptor molecules, including focal adhesion kinase (FAK), Src, p130CAS, paxillin, extracellular signal-regulated kinase (ERK) and myosin light-chain kinase (MLCK) are critical for adhesion turnover at the cell front, a process central to migration.
Article
THE intracellular protein tyrosine kinase FAK (focal adhesion kinase) was originally identified by its high level of tyrosine phosphorylation in v-src-transformed cells1á¤-4. FAK is also highly phosphorylated during early development5,6. In cultured cells it is localized to focal adhesion contacts and becomes phosphorylated and activated in response to integrin-mediated binding of cells to the extracellular matrix, suggesting an important role in cell adhesion and/or migration. We have generated FAK-deficient mice by gene targeting to examine the role of FAK during development. Mutant embryos displayed a general defect of mesoderm development, and cells from these embryos had reduced mobilityin vitro. Surprisingly, the number of focal adhesions was increased in FAK-deficient cells, suggesting that FAK may be involved in the turnover of focal adhesion contacts during cell migration.
Article
Measurement of atomic force microscope cantilever spring constants (k) is essential for many of the applications of this versatile instrument. Numerous techniques to measure k have been proposed. Among these, we found the thermal noise and Sader methods to be commonly applicable and relatively user-friendly, providing an in situ, non-destructive, fast measurement of k for a cantilever independent of its material or coating. Such advantages recommend these methods for widespread use. An impediment thereto is the significant complication involved in the initial implementation of the methods. Some details of the implementation are discussed in publications, while others are left unsaid. Here we present a complete, cohesive, and practically oriented discussion of the implementation of both the thermal noise and Sader methods of measuring cantilever spring constants. We review the relevant theory and discuss practical experimental means for determining the required quantities. We then present results that compare measurements of k by these two methods over nearly two orders of magnitude, and we discuss the likely origins of both statistical and systematic errors for both methods. In conclusion, we find that the two methods agree to within an average of 4% over the wide range of cantilevers measured. Given that the methods derive from distinct physics we find the agreement a compelling argument in favour of the accuracy of both, suggesting them as practical standards for the field.
Article
The mechanical properties of some hollow organs are most conveniently described by a pressure-volume relationship. If the material exhibits hysteresis, thep-v relation must include provision for time-dependent or path-dependent properties. Provided the amplitude of deformation is fairly small and the hysteresis is primarily of the viscoelastic type, a linear description is possible. That this may take the form of a simple transfer function in which material properties are implicit is illustrated for the case of a rubber balloon. The transfer function was derived from the pressure transients which follow step changes in volume produced in a fluid-filled plethysmograph. The applicability of the transfer function in predicting responses to other forcing functions was tested by varying the balloon volume sinusoidally over a frequency range of 1000, at 4 different amplitudes. The good agreement between the linear model and all types of data justifies the use of Laplace transform methods and the assumption that superposition holds. When isolated cat lung is tested in the same manner, the transfer function quantitatively predicts the magnitude ratio of sinusoidal responses but only about two-thirds of the phase angle. The additional energy loss per cycle is interpreted as arising from static hysteresis. The analysis thus provides a simple means of estimating the relative contributions of viscoelastic (dynamic) and static hysteretic processes to the total damping in a material.
Article
A solution of the axisymmetric Boussinesq problem is derived from which are deduced simple formulae for the depth of penetration of the tip of a punch of arbitrary profile and for the total load which must be applied to the punch to achieve this penetration. Simple expressions are also derived for the distribution of pressure under the punch and for the shape of the deformed surface. The results are illustrated by the evaluation of the expressions for several simple punch shapes.RésuméL'auteur a établi une solution du problème axisymétrique de Boussinesq qui lui à permis de déduire des formules simples donnant la profondeur de pénétration d'un pénétrateur de profil arbitraire ainsi que la charge totale nécessaire pour assurer cette pénétration. Il donne également des expressions simples qui définissent la distribution de la pression sous le pénétrateur ainsi que la forme de la surface déformée.Les résultats sont illustrés par l'application de ces expressions aux cas de plusieurs pénétrateurs de forme simple.ZusammenfassungEine Lösung des Boussinesq'schen achsialsymmetrischen Problems wird abgeleitet. Von dieser werden dann weitere einfache Formeln abgeleitet, mit denen sich die Einschlagtiefe einer Stösselspitze mit willkürlichem Profil, und die zur Erzielung diesser Tiefe aufzubringende Gesamtlast bestimmen lässt Ausserdem werden einfache Ausdrücke für die Druckverteilung unter dem Stössel und die Form der deformierten Oberfläche angegeben. Die Ergebnisse werden durch die Auswertung der Ausdrücke für mehrere einfache Stösselformen erleutert.SumàrioSi deriva una soluzione del problema assisimmetrico del Boussinesq dalla quale si deducono semplici formule per la profondità di penetrazione della punta di un punzone di profilo arbitrario e per il carico totale ehe deve venire applicato al punzone per ottenere detta penetrazione. Si derivano inoltre semplici espressioni per la distribuzione della pressione sotto il punzone e per il profilo della superficie deformata. I risultati sono illustrati con la valutazione delle espressioni per vari profili semplici di punzone.РефератДaнo peшeниe aкcиcиммeTpичнoй пpoблeмы Бuccoнeк,a, из кoTopoгo вывeдeны пpocTыe фopмuлы дпя глuбины пoгpuжeния кoнцa пpoбoйникa пpoизвoльнoгo пpoфиля, и для пoлнoй нaгpuзки, кoTopaя дoлжнa быTь пpилoжeнa к пpoбoйиикu, чToбя дaTь эTu глuбинu пoгpuжeния.Taкжe вывeдeны пpocTыe выpaжeния для pacпpeдeлeния дaвлeния пoд чpoбoйникoм и для фopмы дeфopмиpoвaннoй пoвepxнocTи. peзuльTaTы иллюcTpнpoвaны вычиcлeниями Taкиx выpaжeний для нecкoлькиx пpocTыx фopм пpoбoйнкoв.