ArticlePDF Available

Abstract and Figures

Endothelial cells are constantly exposed to fluid shear stresses that regulate vascular morphogenesis, homeostasis, and disease. The mechanical responses of endothelial cells to relatively high shear flow such as that characteristic of arterial circulation has been extensively studied. Much less is known about the responses of endothelial cells to slow shear flow such as that characteristic of venous circulation, early angiogenesis, atherosclerosis, intracranial aneurysm, or interstitial flow. Here we used a novel, to our knowledge, microfluidic technique to measure traction forces exerted by confluent vascular endothelial cell monolayers under slow shear flow. We found that cells respond to flow with rapid and pronounced increases in traction forces and cell-cell stresses. These responses are reversible in time and do not involve reorientation of the cell body. Traction maps reveal that local cell responses to slow shear flow are highly heterogeneous in magnitude and sign. Our findings unveil a low-flow regime in which endothelial cell mechanics is acutely responsive to shear stress.
Content may be subject to copyright.
Biophysical Letter
Traction Forces of Endothelial Cells under Slow Shear Flow
Cecile M. Perrault,
1,2,3,
*Agusti Brugues,
3
Elsa Bazellieres,
3
Pierre Ricco,
2
Damien Lacroix,
1,2,3
and Xavier Trepat
3,4,5,6
1
Institute for In Silico Medicine and
2
Department of Mechanical Engineering, University of Sheffield, Sheffield, United Kingdom;
3
Institute for
Bioengineering of Catalonia, Barcelona, Spain;
4
Facultat de Medicina, Universitat de Barcelona, Barcelona, Spain;
5
Centro de Investigacio
´n
Biome
´dica en Red en Bioingenierı
´a, Biomateriales y Nanomedicina, Barcelona, Spain; and
6
Institucio
´Catalana de Recerca i Estudis Avanc¸ats,
Barcelona, Spain
ABSTRACT Endothelial cells are constantly exposed to fluid shear stresses that regulate vascular morphogenesis, homeosta-
sis, and disease. The mechanical responses of endothelial cells to relatively high shear flow such as that characteristic of arterial
circulation has been extensively studied. Much less is known about the responses of endothelial cells to slow shear flow such as
that characteristic of venous circulation, early angiogenesis, atherosclerosis, intracranial aneurysm, or interstitial flow. Here we
used a novel, to our knowledge, microfluidic technique to measure traction forces exerted by confluent vascular endothelial cell
monolayers under slow shear flow. We found that cells respond to flow with rapid and pronounced increases in traction forces
and cell-cell stresses. These responses are reversible in time and do not involve reorientation of the cell body. Traction maps
reveal that local cell responses to slow shear flow are highly heterogeneous in magnitude and sign. Our findings unveil a low-flow
regime in which endothelial cell mechanics is acutely responsive to shear stress.
Received for publication 4 March 2015 and in final form 11 August 2015.
*Correspondence: c.perrault@sheffield.ac.uk
The ability of endothelial cells (ECs) to sense and adapt to
shear flow is one of the best-studied phenomena in all me-
chanobiology. In response to flow, ECs are known to change
their orientation, remodel cell-cell and cell-matrix adhe-
sions, modify patterns of gene expression, and alter protein
localization at the cell membrane (1). Because these re-
sponses are downstream of a mechanical stimulus, a number
of studies have analyzed the time evolution of cell-matrix
tractions during the application of shear flow. Some of these
studies reported increases in traction forces with shear flow
(2–4), whereas others reached the opposite conclusion (5,6).
Previous studies of traction forces exerted by ECs in the
presence of constant shear flow focused on the application
of shear stresses >1 Pa (10 dyn/cm
2
). Shear stresses in this
range are characteristic of arterial flow during physiological
function. In many other physiological and pathological
conditions, however, shear stresses are much weaker. This
is the case of shear stresses during venous (7) and interstitial
flow (8), as well as during atherosclerosis (9) and intracranial
aneurysm (10). The biochemical and structural responses of
ECs to high versus low shear stress have been extensively
shown to differ in terms of cell morphology, orientation,
and expression of vasoactive agents, antioxidant enzymes,
growth regulators, inflammatory mediators, and adhesion
molecules (reviewed by Malek et al. (9)). Moreover, in the
presence of ultraslow flow such as interstitial flow (11),
ECs are capable of forming numerous capillary-like struc-
tures and have a greater rate of invasion (12). Many of the
phenomena described above are likely to involve a synergy
between flow sensing and force generation (13), but the
link between slow flow and cell contractility is unknown.
To address this question, we combined traction micro-
scopy (TM) and monolayer stress microscopy (MSM)
with microfluidic techniques and explored cellular traction
forces in reaction to slow shear flow (Fig. 1). TM maps
the magnitude, location, and direction of the forces exerted
by cells against their underlying soft substrate (14). Sub-
strate displacements caused by cell tractions are mapped
using fiduciary markers embedded in the soft substrate.
The displacement fields are then used to compute tractions
by inverting the elasticity equations in Fourier space.
TM was integrated into a microfluidic chamber, created in
PDMS (polydimethylsiloxane) by soft lithography. A mold
was machined from Plexiglas to create rectangular flow
channel of 2 mm in width and 2 cm in length. The channel
was designed to have two different heights over its length,
thus creating a channel with two different shear stress values
(15). The circulating media entered a chamber with an
initial height of 300 mm, and moved into a chamber with
a height of 600 mm. Corresponding shear stress values can
be found in Table S1 in the Supporting Material. Cell trac-
tions were monitored in both chambers, away from the tran-
sition zone between the two.
Editor: Gijsje Koenderink.
Ó2015 by the Biophysical Society
http://dx.doi.org/10.1016/j.bpj.2015.08.036
Biophysical Journal Volume 109 October 2015 1–4 1
Please cite this article in press as: Perrault et al., Traction Forces of Endothelial Cells under Slow Shear Flow, Biophysical Journal (2015), http://dx.doi.org/
10.1016/j.bpj.2015.08.036
BPJ 6761
Monolayers of human umbilical vein endothelial cells
were exposed to a time-varying protocol alternating no-
flow and applied flow in the range 0.014–0.133 Pa (see
Fig. 2 and the Supporting Material). The temporal stress
pattern consisted of two consecutive flow steps of 30-min
duration and increasing magnitude (5 and 10 mL/h), fol-
lowed by a 30-min period of no flow. After this period, we
applied a second pulse of flow (10 mL/h) lasting 30 min.
Upon exposure to flow, we observed an acute increase in
strain energy (the total energy transmitted by the cells on
the substrate) with no significant differences between the
two flow levels (Fig. 2). Subsequent doubling of the shear
flow did not trigger a second increase in traction forces.
Instead, cells tended to plateau at values that were 50–
100% higher than baseline levels. Quickly after stopping
the flow, the strain energy relaxed toward baseline levels,
thus indicating reversibility of responses to flow. Applica-
tion of an additional flow pulse triggered a second acute in-
crease in tractions, with pronounced differences between
the two flow levels, which is suggestive of a memory effect.
Finally, stopping the flow led to a relaxation toward base-
line levels.
The responses shown in Fig. 2 are spatial averages of trac-
tion maps. As previously shown in a diversity of cell types
(6,16), these maps exhibited a punctate distribution with
large spatial heterogeneities (Fig. 3,Aand B). The response
to shear flow was also heterogeneous; although the overall
traction of the monolayer increased, several cells displayed
significant traction drops (Fig. 3 C). The magnitude of local
changes in traction in response to shear was similar to the
global traction average. Upon flow application, tractions
showed a weak but significant tendency to orient perpendic-
ular to the direction of flow (p<0.001, Rayleigh test, Fig. 3,
Dand E).
Web 3C
FIGURE 2 ECs display acute responses to
slow shear flow. Strain energy was normalized
to its baseline (t¼0). (Green line) Flow values.
(Red and black lines) High shear (HS) stress
and low shear (LS) stress; both shear stresses
were at least one order-of-magnitude lower
than previously reported shear stresses in
TM experiments. n¼6 monolayers per condi-
tion. Differences between HS and LS are only
significant during the second shear pulse
(t¼105 min and t¼125 min, p%0.05). The
slopes of the strain energy between the first
and second flow periods were not significant.
To see this figure in color, go online.
Web 3C
FIGURE 1 Microfluidic traction assay. A
PDMS flow chamber is assembled over a
strip of polyacrylamide (E¼1.25 KPa) poly-
merized on a glass coverslip. The chamber
is divided into two sections of varying
heights. The flow is controlled by a syringe
pump. To see this figure in color, go online.
Biophysical Journal 109(8) 1–4
2Biophysical Letters
Please cite this article in press as: Perrault et al., Traction Forces of Endothelial Cells under Slow Shear Flow, Biophysical Journal (2015), http://dx.doi.org/
10.1016/j.bpj.2015.08.036
BPJ 6761
Finally, we used MSM to measure cell-cell stresses (14).
As reported previously (6,14), cell-cell stresses showed
supracellular spatial fluctuations (see Fig. S1). Upon flow
application, these fluctuations increased in magnitude but
cell-cell stresses did not change in orientation. Unlike
cell-substrate tractions, cell-cell stresses did not show sig-
nificant differences depending on flow magnitude.
Traction forces in the presence of constant shear stresses
of relatively high magnitude (>1 Pa) have been extensively
characterized in previous studies, with conflicting results
(2–6). Here we used shear stresses between one and two
orders-of-magnitude smaller than those applied in previous
studies based on TM. In response to these low stresses, ECs
exhibited acute but reversible increases in traction. These re-
sponses were fast and more pronounced than those reported
in previous studies using higher shear flows (2–6). The low
shear stresses applied here fall within the range of physio-
logical interstitial flow (8). Because interstitial flow induces
angiogenesis, we speculate that increased traction forces
observed here might recapitulate those required to initiate
the formation of new blood vessels in vivo (17). Low flows
are also characteristic of pathological conditions such as
intracranial aneurysm (10) and atherosclerosis (9); our find-
ings raise the question of whether increases in traction
forces might be protective or disruptive in these conditions.
A remarkable feature of our experiments was the hetero-
geneity of the responses. Heterogeneous responses of ECs to
flow have been previously reported (18) in terms of protein
and mRNA levels, calcium signaling, and organelle locali-
zation. Heterogeneity has been attributed to the topography
of the monolayer (19), the heterogeneous location, proper-
ties of cytoskeleton elements (20), and flow sensors. Our
A
B
C
D
E
Web 3C
FIGURE 3 Instantaneous maps of traction
forces of an endothelial monolayer at (A)t¼
96 min and (B)t¼126 min. (C) Difference
between (A) and (B). Distribution of the angle
between traction vectors and the direction of
flow at (D)t¼96 min and (E)t¼126 min. To
see this figure in color, go online.
Biophysical Journal 109(8) 1–4
Biophysical Letters 3
Please cite this article in press as: Perrault et al., Traction Forces of Endothelial Cells under Slow Shear Flow, Biophysical Journal (2015), http://dx.doi.org/
10.1016/j.bpj.2015.08.036
BPJ 6761
findings of heterogeneous force distributions might underlie
heterogenous responses in signaling and molecular localiza-
tion through mechanotransduction activity. Whether such
activity is dominated by cell-matrix or cell-cell stresses
could be elicited based on the differences in the character-
istic lengths of the corresponding fluctuations.
In conclusion, we used a new device, to our knowledge, to
combine microfluidics, TM, and MSM to assess the me-
chanical response of ECs to slow shear flow. The observed
acute increases in traction generation provide fresh insights
into the synergy between flow and the biomechanical reac-
tion of cells, with potential implications in morphogenesis
and disease.
SUPPORTING MATERIAL
Supporting Materials and Methods, one table, and one figure, are available at
http://www.biophysj.org/biophysj/supplemental/S0006-3495(15)00871-1.
AUTHOR CONTRIBUTIONS
C.M.P., D.L., and X.T. designed the study; C.M.P. implemented the flow
chamber and carried out experiments; C.M.P., A.B., and E.B. processed
data; C.M.P. and X.T. wrote the article; P.R. calculated shear and gel
displacement in the chamber; and all authors discussed and interpreted re-
sults and commented on the article.
ACKNOWLEDGMENTS
We thank Daniel Navajas, and members of the Integrative Cell and Tissue
Dynamics Lab and the MechanoBio Lab at the Institute for Bioengineering
of Catalonia, for fruitful discussions.
C.M.P. received funding from the Institute for Bioengineering of Catalonia.
This work was funded by the Spanish Ministry of Economy and Competi-
tiveness (under grant No. BFU2012-38146), the Generalitat de Catalunya
(under grant No. 2014-SGR-927), and the European Research Council (un-
der grant No. CoG-616480).
SUPPORTING CITATIONS
References (21–24) appear in the Supporting Material.
REFERENCES
1. Davies, P. F. 1995. Flow-mediated endothelial mechanotransduction.
Physiol. Rev. 75:519–560.
2. Hur, S. S., J. C. del A
´lamo, ., S. Chien. 2012. Roles of cell confluency
and fluid shear in 3-dimensional intracellular forces in endothelial
cells. Proc. Natl. Acad. Sci. USA. 109:11110–11115.
3. Shiu, Y.-T., S. Li, ., S. Chien. 2004. Rho mediates the shear-enhance-
ment of endothelial cell migration and traction force generation. Bio-
phys. J. 86:2558–2565.
4. Ting, L. H., J. R. Jahn, ., N. J. Sniadecki. 2012. Flow mechanotrans-
duction regulates traction forces, intercellular forces, and adherens
junctions. Am. J. Physiol. Heart Circ. Physiol. 302:H2220–H2229.
5. Conway, D. E., M. T. Breckenridge, ., M. A. Schwartz. 2013. Fluid
shear stress on endothelial cells modulates mechanical tension across
VE-cadherin and PECAM-1. Curr. Biol. 23:1024–1030.
6. Steward, R., D. Tambe, ., J. J. Fredberg. 2015. Fluid shear, intercel-
lular stress, and endothelial cell alignment. Am. J. Physiol. Cell Phys-
iol. Published online February 4, 2015. http://dx.doi.org/10.1152/
ajpcell.00363.2014.
7. Wragg, J. W., S. Durant, ., R. Bicknell. 2014. Shear stress regulated
gene expression and angiogenesis in vascular endothelium. Microcircu-
lation. 21:290–300.
8. Swartz, M. A., and M. E. Fleury. 2007. Interstitial flow and its effects in
soft tissues. Annu. Rev. Biomed. Eng. 9:229–256.
9. Malek, A. M., S. L. Alper, and S. Izumo. 1999. Hemodynamic shear
stress and its role in atherosclerosis. J. Amer. Med. Assoc. 282:2035–
2042.
10. Meng, H., V. M. Tutino, ., A. Siddiqui. 2014. High WSS or low WSS?
Complex interactions of hemodynamics with intracranial aneurysm
initiation, growth, and rupture: toward a unifying hypothesis. Am. J.
Neuroradiol. 35:1254–1262.
11. Park, J. Y., J. B. White, ., S. Takayama. 2011. Responses of endothe-
lial cells to extremely slow flows. Biomicrofluidics. 5:22211.
12. Song, J. W., and L. L. Munn. 2011. Fluid forces control endothelial
sprouting. Proc. Natl. Acad. Sci. USA. 108:15342–15347.
13. Kniazeva, E., and A. J. Putnam. 2009. Endothelial cell traction and
ECM density influence both capillary morphogenesis and maintenance
in 3-D. Am. J. Physiol. Cell Physiol. 297:C179–C187.
14. Tambe, D. T., C. C. Hardin, ., X. Trepat. 2011. Collective cell guid-
ance by cooperative intercellular forces. Nat. Mater. 10:469–475.
15. Galie, P. A., A. van Oosten, ., P. A. Janmey. 2015. Application of
multiple levels of fluid shear stress to endothelial cells plated on poly-
acrylamide gels. Lab Chip. 15:1205–1212.
16. Trepat, X., M. R. Wasserman, ., J. J. Fredberg. 2009. Physical forces
during collective cell migration. Nat. Phys. 5:426–430.
17. Herna
´ndez Vera, R., E. Genove
´,., C. E. Semino. 2009. Interstitial
fluid flow intensity modulates endothelial sprouting in restricted
Src-activated cell clusters during capillary morphogenesis. Tissue
Eng. A. 15:175–185.
18. Davies, P. F., T. Mundel, and K. A. Barbee. 1995. A mechanism
for heterogeneous endothelial responses to flow in vivo and in vitro.
J. Biomech. 28:1553–1560.
19. Barbee, K. A. 2002. Role of subcellular shear-stress distributions in
endothelial cell mechanotransduction. Ann. Biomed. Eng. 30:472–482.
20. Davies, P. F., J. Zilberberg, and B. P. Helmke. 2003. Spatial microsti-
muli in endothelial mechanosignaling. Circ. Res. 92:359–370.
21. Kandow, C. E., P. C. Georges, ., K. A. Beningo. 2007. Polyacryl-
amide hydrogels for cell mechanics: steps toward optimization and
alternative uses. Methods Cell Biol. 83:29–46.
22. White, F. M. 2003. Fluid Mechanics. McGraw-Hill, New York.
23. Yeung, T., P. C. Georges, ., P. A. Janmey. 2005. Effects of substrate
stiffness on cell morphology, cytoskeletal structure, and adhesion.
Cell Motil. Cytoskeleton. 60:24–34.
24. Tambe, D. T., U. Croutelle, ., J. J. Fredberg. 2013. Monolayer stress
microscopy: limitations, artifacts, and accuracy of recovered intercel-
lular stresses. PLoS One. 8:e55172.
Biophysical Journal 109(8) 1–4
4Biophysical Letters
Please cite this article in press as: Perrault et al., Traction Forces of Endothelial Cells under Slow Shear Flow, Biophysical Journal (2015), http://dx.doi.org/
10.1016/j.bpj.2015.08.036
BPJ 6761
TRACTION FORCES OF ENDOTHELIAL CELLS UNDER SLOW
SHEAR FLOW
C.M. Perrault1,2,3, A. Brugués3, E. Bazellieres3 , P. Ricco2 , D. Lacroix1,2,3 and X. Trepat3,4,5
1 INSIGNEO Institute for in-silico Medicine, University of Sheffield, Sheffield, UK; 2 Department of Mechanical Engineering,
University of Sheffield, Sheffield, UK; 3 Institute for Bioengineering of Catalonia, Barcelona, Spain; 4 Facultat de Medicina,
Universitat de Barcelona, and Ciber-BBN, Barcelona, Spain; 5 Institució Catalana de Recerca i Estudis Avançats, Barcelona,
Spain.
Supplementary material
1. Materials and Methods
1.1. Preparation of polyacrylamide gels and cell culture
Large rectangular coverslips were activated by using a 1:1:14 solution of acetic acid/bind-
silane/ethanol. The dishes were washed twice with ethanol and air-dried for 10 min.
Polyacrylamide gels (E= 1.25 kPa) were prepared as described in Kandow et al (1) and Yeung et al
(2). Briefly, a solution containing 5% acrylamide, 0.1% bis acrylamide, 0.5% ammonium
persulphate, 0.05% tetramethylethylenediamine, 0.4% of 200-nm-diameter red fluorescent
carboxylate-modified beads and 2mgml1 NH-acrylate was prepared. A molding channel was
created on the coverslip using double sided-tape and transparency paper to create polyacrylamide gels
with dimensions of 1mm in width and 2cm in length. A drop of 10 μl was placed on one end of the
channel, which then fills by capillary action.
After polymerization, tape and paper were carefully removed and the gels were washed with PBS and
incubated with 100μl of a collagen I solution (0.1 mg/ml, Millipore) overnight at 4°C. The gels were
washed afterwards with PBS and cells were seeded and incubated with cell culture media with 10%
FBS for 6h.
HUVEC cells were cultured in EGM™ BulletKit™ (Lonza, MA).
1.2. Flow experiments
A small volume (8 μl) containing 150,000 cells was placed on the polyacrylamide gel. Once the cells
were attached to the polyacrylamide gel (20 min), the unattached cells were washed away and 2 ml of
medium were added. Twelve hours after seeding the cells, the coverslip was attached to the PDMS
flow chamber and held together by a custom-made holder. The ensemble was then placed on the
microscope and connected on one side to a syringe pump (World Precision Instruments Aladdin 1000,
WPI, FL) and to the other side to a reservoir of degassed medium.
1.3. Time-lapse microscopy
Multidimensional acquisition routines were performed on an automated inverted microscope (Nikon
Eclipse Ti) equipped with thermal, CO2 and humidity control, using MetaMorph (Universal Imaging)
software. Time-lapse recording started approximately 10 min after assembly. The interval between
image acquisition was 1 min.
1.4. Traction microscopy (TM)
Cell tractions were evaluated using monolayer Fourier-transform traction microscopy (3). Briefly, the
displacement field was calculated by comparing fluorescent microbead images obtained during the
experiment with a reference image taken at the end of the experiment after the trypsinization and the
consequent detachment of the cells from the underlying substrate. A particle imaging velocimetry
algorithm (3) was used to determine the deformation of the substrate caused by the traction forces.
1.5 Monolayer Stress Microscopy (MSM)
In a 2D approximation, monolayer stress is fully captured by a tensor possessing two independent
normal components ( and ) and two identical shear components ( and ). At every pixel
of the monolayer, these four components of the stress tensor define two particular directions of the
plane, one in which the normal stress is maximum and one in which it is minimum. These directions,
which are mutually orthogonal, are called principal stress orientations, and the stress values in each
principal orientation are called maximum () and minimum () stress components. The average
normal stress is defined as =(+)/2, while the maximum shear stress is defined as =(-
)/2. The spatial resolution and force precision of MSM are formally set by those in the original
traction maps. How the reconstructed stress field is affected by the choice of boundary conditions and
by the assumptions of continuity, incompressibility, and homogeneity was extensively studied
elsewhere (3,6).
1.6 Statistical analysis
Summary data are expressed as mean ± SEM (standard error of the mean). Unless noted otherwise,
statistical comparisons were computed by Student’s t-test. Traction angles were compared to an
isotropic distribution using a Raleigh test. A value of p 0.05 was considered statistically significant.
2. Flow analysis in the microfluidic chamber
The microfluidic chamber was created to limit contact of the cells with PDMS prior to the experiment.
As a result, the system was designed with two components: 1) a strip of PAA gels on a glass slide and
2) the PDMS flow chamber. To maintain both sides together during operation, a customized
aluminium holder was created to apply mechanical forces over the assembly. To facilitate alignment
and limit leakage, the PAA gels were created slightly less wide than the flow channel, but the gap was
small enough to avoid fluctuations in the fluid distribution, as observed by flow of fluorescent beads.
To ensure that the displacement of the beads in the polyacrylamide gels was only caused by the cells,
the displacement of the upper layer of the gels due to shear stress was calculated. As the Reynolds
number of the flow (  , where u is the velocity, d is the hydraulic diameter and v is the
kinematic viscosity) is low, the flow can be assumed to be laminar. Assuming Poiseuille flow, the
ow rate per unit width (m2/s) is related to the streamwise mean pressure gradient as follows (4):

 (1)
Where Px represents the pressure gradient along the streamwise direction and μ is the viscosity of the
fluid.
As the ow rate per unit width is related to the volumetric ow rate  , the pressure
gradient is readily found. The laminar ow is given by

  (2)
where the y coordinate is taken from the middle of the gap so that the wall-shear stress τw is found

 
(3)
The angle γ of deformation of the surface is

(4)
where G is the shear modulus of the flexible surface.
The streamwise displacement Δs of the exible surface is thus
∆  tan (5)
The results for flow rates of 5 and 10 mL/hr in the small and large region of the chambers are
summarized in the table 1.
Small chamber (high shear) Large chamber (low shear)
Wall shear stress Gel displacement Wall shear stress Gel displacement
Q= 5mL/hr 0.066 Pa 2.7x10-9 m 0.014 Pa 5.5x10-10 m
Q= 10 mL/hr 0.133 Pa 5.3x10-9 m 0.028 Pa 1.1x10-9 m
Table1.Wallshearstressandgeldisplacementvaluesfortheflowratesappliedduringtheexperimentsinboth
chambersoftheflowdevice.Gelstiffnessis1.25kPa.
The displacement of the gel due to the flow is thus minimal and the shear stress values correspond
to a low flow regime (5).
3. Supplementary Figure
Supplementary Fig S1. Cell-cell stresses display acute and reversible responses to low shear flow.
A-B) Maps of the average normal stress at t= 96 min and t=126 min (same time points as in Fig. 3
in the main text). C) Time evolution of the average cell-cell stresses (norm of the average normal
stress) in response to the flow pattern depicted in green (n=6 per condition). D) Angular change of
the maximum principal stress between t=96 min and t=126min. These data show that changes in
cell-cell stress orientation in response to shear flow were minimal.
SUPPORTING REFERENCES
1. Kandow, C.E., P.C. Georges, P.A. Janmey, and K.A. Beningo. 2007. Polyacrylamide
hydrogels for cell mechanics: steps toward optimization and alternative uses. Methods Cell
Biol. 83: 29–46.
2. Yeung, T., P.C. Georges, L.A. Flanagan, B. Marg, M. Ortiz, M. Funaki, N. Zahir, W. Ming, V.
Weaver, and P.A. Janmey. 2005. Effects of substrate stiffness on cell morphology, cytoskeletal
structure, and adhesion. Cell Motil. Cytoskeleton. 60: 24–34.
3. Tambe, D.T., C.C. Hardin, T.E. Angelini, K. Rajendran, C.Y. Park, X. Serra-Picamal, E.H.
Zhou, M.H. Zaman, J.P. Butler, D.A. Weitz, J.J. Fredberg, and X. Trepat. 2011. Collective cell
guidance by cooperative intercellular forces. Nat. Mater. 10: 469–75.
4. White, F.M. 2003. Fluid Mechanics. McGraw-Hill.
5. Swartz, M.A., and M.E. Fleury. 2007. Interstitial flow and its effects in soft tissues. Annu.
Rev. Biomed. Eng. 9: 229–56.
6. Tambe DT, Croutelle U, Trepat X, Park CY, Kim JH, Millet E, Butler JP, and Fredberg JJ.
Monolayer stress microscopy: limitations, artifacts, and accuracy of recovered intercellular
stresses. PloS one 8, e55172 (2013).
... [326][327][328][329] Flow-induced traction or intercellular stress is another crucial mechanical cue regulating endothelial cell behaviors in vivo. Perrault et al. observed that endothelial cells have an acute but reversible response to slow flow (similar to interstitial flow) due to the heightened traction and intercellular stress, [330] which is also supported by Steward et al. [331] As interstitial flow can induce the formation of new vessels in vivo, [332] it is speculated that enhanced traction and cell-cell stress are required to initiate the process. In addition to flow-induced traction force, hydrostatic pressure from fluid flow may disrupt the vascular endothelialcadherin junction, leading to a multilayer endothelium structure with increased intercellular gap and permeability. ...
... 2023, 2300670 www.advancedsciencenews.com www.advancedscience.com 10%-strain stretching Epithelial crack and healing [259] 0-15% cyclic strain, 0.2 Hz Increased cell-cell interaction [260] Compression Loaded with 150 g weights Improve skin reconstitution [341] Endothelial cell Flow-induced shear stress >10 dyn cm −2 Sprouting, matrix invasion [342] 20 dynes cm −2 steady shear Reduced endothelial-to-mesenchymal transition [325] 1.2 Pa laminar fluid shear Retarded elongation, alignment [330] Mesenchymal stem cell Compression 10% peak compressive sinusoidal strain, 1 Hz, 5% compressive tare strain Enhanced chondrogenic differentiation, survival [196] 0-2.5% strain, 1 Hz, 1-h ON +23-h OFF Distinct collagen expression, mineral deposits [197] 10 kPa, 0.25 Hz, 1 h day −1 , start day 1 or 21 ...
Article
Full-text available
Cells interact with their surrounding environment through a combination of static and dynamic mechanical signals that vary over stimulus types, intensity, space, and time. Compared to static mechanical signals such as stiffness, porosity, and topography, the current understanding on the effects of dynamic mechanical stimulations on cells remains limited, attributing to a lack of access to devices, the complexity of experimental set‐up, and data interpretation. Yet, in the pursuit of emerging translational applications (e.g., cell manufacturing for clinical treatment), it is crucial to understand how cells respond to a variety of dynamic forces that are omnipresent in vivo so that they can be exploited to enhance manufacturing and therapeutic outcomes. With a rising appreciation of the extracellular matrix (ECM) as a key regulator of biofunctions, researchers have bioengineered a suite of ECM‐mimicking hydrogels, which can be fine‐tuned with spatiotemporal mechanical cues to model complex static and dynamic mechanical profiles. This review first discusses how mechanical stimuli may impact different cellular components and the various mechanobiology pathways involved. Then, how hydrogels can be designed to incorporate static and dynamic mechanical parameters to influence cell behaviors are described. The Scopus database is also used to analyze the relative strength in evidence, ranging from strong to weak, based on number of published literatures, associated citations, and treatment significance. Additionally, the impacts of static and dynamic mechanical stimulations on clinically relevant cell types including mesenchymal stem cells, fibroblasts, and immune cells, are evaluated. The aim is to draw attention to the paucity of studies on the effects of dynamic mechanical stimuli on cells, as well as to highlight the potential of using a cocktail of various types and intensities of mechanical stimulations to influence cell fates (similar to the concept of biochemical cocktail to direct cell fate). It is envisioned that this progress report will inspire more exciting translational development of mechanoresponsive hydrogels for biomedical applications.
... Cellular reactions to external mechanical cues play a crucial role in cellular processes such as stem cell differentiation, adhesion, migration, and proliferation (Paluch 2015;Lv 2015;Engler et al. 2006;Cui 2015;Brugués 2014). Furthermore, focal adhesion clusters grow in response to external shearing (Riveline 2001;Paul et al. 2008) which might help cells to withstand shear forces, e.g., forces exerted by the blood flow on endothelial cells (Davies 1995;Perrault 2015). ...
... Techniques to exert mechanical stimuli to cells include atomic force microscopy (AFM), which can be employed to measure forces necessary to rupture cellular adhesions (Kadem 2016;Selhuber-Unkel 2010) or forces exerted by cells (Huth et al. 2017;Brunner 2006), hydrodynamic shear stress (Davies 1995;Perrault 2015;Hanke et al. 2019), optical or magnetic tweezers (Rief et al. 1997;Neuman and Nagy 2008;Jiang et al. 2003;Roca-Cusachs et al. 2017), microneedle assays (Fedorchak and Lammerding 2016;Riveline 2001;Paul et al. 2008) and optical stretchers (Chan 2015;Micoulet et al. 2005). Despite the fact that such a large variety of physical cell manipulation techniques has been established and cellular forces exerted to surfaces can be measured via TFM or elastic resonator interference stress microscopy (Kronenberg 2017), a quantification of cellular force adaptation as a response to well-defined mechanical stimuli applied to cells has not yet been realized. ...
Article
Full-text available
Mammalian cells have evolved complex mechanical connections to their microenvironment, including focal adhesion clusters that physically connect the cytoskeleton and the extracellular matrix. This mechanical link is also part of the cellular machinery to transduce, sense and respond to external forces. Although methods to measure cell attachment and cellular traction forces are well established, these are not capable of quantifying force transmission through the cell body to adhesion sites. We here present a novel approach to quantify intracellular force transmission by combining microneedle shearing at the apical cell surface with traction force microscopy at the basal cell surface. The change of traction forces exerted by fibroblasts to underlying polyacrylamide substrates as a response to a known shear force exerted with a calibrated microneedle reveals that cells redistribute forces dynamically under external shearing and during sequential rupture of their adhesion sites. Our quantitative results demonstrate a transition from dipolar to monopolar traction patterns, an inhomogeneous distribution of the external shear force to the adhesion sites as well as dynamical changes in force loading prior to and after the rupture of single adhesion sites. Our strategy of combining traction force microscopy with external force application opens new perspectives for future studies of force transmission and mechanotransduction in cells.
... Increased membrane tension limits endocytic cellular trafficking by inducing rapid disassembly of caveolae (211), which are involved in intercellular bacterial spread (213). Variations in the magnitude of shear stresses and disturbances in shear flow significantly alter the cytoskeletal organization of cells, including the way they transduce forces to each other and to their environment (243,244). After long-term exposure to shear stress, cells exhibit increased cortical stiffness, and their surface is rougher and decorated with more "ridges" (245). ...
Article
To combat infectious diseases, it is important to understand how host cells interact with bacterial pathogens. Signals conveyed from pathogen to host, and vice versa, may be either chemical or mechanical. While the molecular and biochemical basis of host-pathogen interactions has been extensively explored, relatively less is known about mechanical signals and responses in the context of those interactions. Nevertheless, a wide variety of bacterial pathogens appear to have developed mechanisms to alter the cellular biomechanics of their hosts in order to promote their survival and dissemination, and in turn many host responses to infection rely on mechanical alterations in host cells and tissues to limit the spread of infection. In this review, we present recent findings on how mechanical forces generated by host cells can promote or obstruct the dissemination of intracellular bacterial pathogens. In addition, we discuss how in vivo extracellular mechanical signals influence interactions between host cells and intracellular bacterial pathogens. Examples of such signals include shear stresses caused by fluid flow over the surface of cells and variable stiffness of the extracellular matrix on which cells are anchored. We highlight bioengineering-inspired tools and techniques that can be used to measure host cell mechanics during infection. These allow for the interrogation of how mechanical signals can modulate infection alongside biochemical signals. We hope that this review will inspire the microbiology community to embrace those tools in future studies so that host cell biomechanics can be more readily explored in the context of infection studies.
... Stretch-induced reorientation of cells involves both passive mechanical response to cyclic substrate deformation, and dynamic changes to the cytoskeleton (Livne et al. 2014). Contractile stress fibers (SF's), comprised primarily of actin and myosin II, are essential in the development of intracellular stresses which are transmitted through focal adhesion (FA) complexes, and depend on the activity of RhoA, active Rac1, or Cdc42 (Balaban et al. 2001;Cirka et al. 2016;Perrault et al. 2015). Cyclic uniaxial stretching of fibroblasts, seeded on isotropic elastomeric substrates, leads to cell reorientation from random to a near-perpendicular angle with respect to the loading direction (Buck 1980;Wang et al. 2001). ...
Article
Full-text available
Stress fibers in the cytoskeleton are essential in maintaining cellular shape and influence cellular adhesion and migration. Cyclic uniaxial stretching results in cellular reorientation orthogonal to the applied stretch direction. The mechanistic cues underlying changes to cellular form and function to stretch stimuli are currently underexplored. We show stretch-induced stress fiber lengthening, their realignment, and increased cortical actin in NIH 3T3 fibroblasts stretched over varied amplitudes and durations. Higher amounts of actin and stress fiber alignment were accompanied with an increase in the effective elastic modulus of cells. Microtubules did not contribute to the measured stiffness or reorientation response but were essential to the nuclear reorientation. We used a phenomenological growth and remodeling law, based on the experimental data, to model stress fiber elongation and reorientation dynamics based on a nonlinear, orthotropic, fiber-reinforced continuum representation of the cell. The model predicts the changes observed fibroblast morphology and increased cellular stiffness under uniaxial cyclic stretch which agrees with experimental results. Such studies are important in exploring the differences underlying mechanotransduction and cellular contractility under stretch.
... Moreover, traction forces of breast cancer cells in the interaction with such surface were quantified (Toyjanova et al., 2014b). Among advances, combining TFM with a microfluidic approach has to be underlined (Perrault et al., 2015). A comparison between static and flow conditions revealed the distinct level of traction forces generated by cells that affect our understanding of their interaction with the surrounding microenvironment in normal and pathological conditions. ...
Article
Cells generate mechanical forces (traction forces, TFs) while interacting with the extracellular matrix or neighbouring cells. Forces are generated by both cells and extracellular matrix (ECM) and transmitted within the cell-ECM or cell-cell contacts involving focal adhesions or adherens junctions. Within more than two decades, substantial progress has been achieved in techniques that measure TFs. One of the techniques is traction force microscopy (TFM). This review discusses the TFM and its advances in measuring TFs exerted by cells (single cells and multicellular systems) at cell-ECM and cell-cell junctional intracellular interfaces. The answers to how cells sense, adapt and respond to mechanical forces unravel their role in controlling and regulating cell behaviour in normal and pathological conditions.
... Finally, it is important to note that, although shear stress is an important regulator in inducing leukocyte TEM (Cinamon et al., 2001), most existing force microscopy studies of leukocyte TEM have not considered shear flow conditions so far. While it requires a more complicated experimental setting, it is not unfeasible to consider shear flow in force microscopy assays and, in fact, there are established TFM and FRET imaging assays that include shear flow (Hur et al., 2012;Perrault et al., 2015;Heemskerk et al., 2016). Future efforts shall exploit these tools to study how shear affects the mechanics of leukocyte TEM by directly measure the forces involved in the process. ...
Article
Full-text available
Leukocyte transendothelial migration is crucial for innate immunity and inflammation. Upon tissue damage or infection, leukocytes exit blood vessels by adhering to and probing vascular endothelial cells (VECs), breaching endothelial cell-cell junctions, and transmigrating across the endothelium. Transendothelial migration is a critical rate-limiting step in this process. Thus, leukocytes must quickly identify the most efficient route through VEC monolayers to facilitate a prompt innate immune response. Biomechanics play a decisive role in transendothelial migration, which involves intimate physical contact and force transmission between the leukocytes and the VECs. While quantifying these forces is still challenging, recent advances in imaging, microfabrication, and computation now make it possible to study how cellular forces regulate VEC monolayer integrity, enable efficient pathfinding, and drive leukocyte transmigration. Here we review these recent advances, paying particular attention to leukocyte adhesion to the VEC monolayer, leukocyte probing of endothelial barrier gaps, and transmigration itself. To offer a practical perspective, we will discuss the current views on how biomechanics govern these processes and the force microscopy technologies that have enabled their quantitative analysis, thus contributing to an improved understanding of leukocyte migration in inflammatory diseases.
Article
Endothelial cells (ECs) within the vascular system encounter fluid shear stress (FSS). High, laminar FSS promotes vasodilation and anti-inflammatory responses, while low or disturbed FSS induces dysfunction and inflammation. However, the adaptation of ECs to dynamically changing FSS patterns remains underexplored. Here, by combining traction force microscopy with a custom flow chamber, we examined human umbilical vein endothelial cells adapting their traction during transitions from short-term low shear to long-term high shear stress. We discovered that the initial low FSS elevates the traction by only half of the amount by direct high FSS even after flow changes to high FSS. However, in the long term under high FSS, the flow started with low FSS triggers a substantial second rise in traction, for over 10 hours. In contrast, direct high FSS exposure results in a quick traction surge followed by a huge reduction below the baseline traction. Importantly, we find that the orientation of traction vectors is steered by initial shear exposure. Using Granger Causality analysis, we show that the traction that aligns in the flow direction under direct high FSS functionally causes cell alignment. Similarly, EC traction that orients perpendicular to the flow that starts with temporary low FSS functionally causes cell orientation perpendicular to the flow. Taken together, our findings elucidate the significant influence of initial short-term low FSS on lasting changes in endothelial traction that induces EC alignment.
Preprint
Full-text available
Fluid shear stress (FSS) plays a vital role in endothelial cell (EC) function, impacting vascular health. Low FSS triggers an inflammatory, atheroprone EC phenotype, while high FSS promotes atheroprotective responses. Yet, the adaptation of ECs to high FSS following low FSS exposure remains poorly understood. We explored this process's influence on vascular tension by measuring traction in human umbilical vein endothelial cells (HUVECs) subjected to high FSS (1 Pa) after prior low FSS exposure (0.1 Pa). We discovered that HUVECs experienced an immediate traction increase after low FSS exposure. Intriguingly, the transition to high FSS did not further enhance traction; instead, a secondary traction rise emerged after 2 hours under high FSS, lasting over 10 hours before gradually diminishing. Conversely, HUVECs exposed directly to high FSS displayed an initial traction surge within 30 minutes, followed by a rapid decline within an hour, falling below initial levels. Notably, even after more than 20 hours of exposure to high FSS, HUVECs previously preconditioned with 1-hour of low FSS exhibited traction alignment perpendicular to the flow direction, whereas those directly exposed to high FSS displayed alignment with the flow direction, both in the short and long terms. Collectively, these results suggest that even brief exposure to low FSS can initiate a low-shear-sensitive response in ECs, which requires a much-extended period to recover towards tension relaxation under laminar high shear flow conditions.
Article
The inherently dynamic and anisotropic microenvironment of cells imposes not only global and slow physical stimulations on cells but also acute and local perturbations. However, cell mechanical responses to transient subcellular physical signals remain unclear. In this study, acoustically activated targeted microbubbles were used to exert mechanical perturbations to single cells. The cellular contractile force was sensed by elastic micropillar arrays, while the pillar deformations were imaged using brightfield high-speed video microscopy at a frame rate of 1k frames per second for the first 10s and then confocal fluorescence microscopy. Cell mechanical responses are accompanied by cell membrane integrity changes. Both processes are determined by the perturbation strength generated by microbubble volumetric oscillations. The instantaneous cellular traction force relaxation exhibits two distinct patterns, correlated with two cell fates (survival or permanent damage). The mathematical modeling unveils that force-induced actomyosin disassembly leads to gradual traction force relaxation in the first few seconds. The perturbation may also influence the far end subcellular regions from the microbubbles and may propagate into connected cells with attenuations and delays. This study carefully characterizes the cell mechanical responses to local perturbations induced by ultrasound and microbubbles, advancing our understanding of the fundamentals of cell mechano-sensing, -responsiveness, and -transduction. Statement of Significance Subcellular physical perturbations commonly exist but haven't been fully explored yet. The subcellular perturbation generated by ultrasound and targeted microbubbles covers a wide range of strength, from mild, intermediate to intense, providing a broad biomedical relevance. With µm² spatial sensing ability and up to 1ms temporal resolution, we present spatiotemporal details of the instantaneous cellular contractile force changes followed by attenuated and delayed global responses. The correlation between the cell mechanical responses and cell fates highlights the important role of the instantaneous mechanical responses in the entire cellular reactive processes. Supported by mathematical modeling, our work provides new insights into the dynamics and mechanisms of cell mechanics.
Article
Full-text available
The vascular endothelial cells constitute the innermost layer. The cells are exposed to mechanical stress by the flow, causing them to express their functions. To elucidate the functions, methods involving seeding endothelial cells as a layer in a chamber were studied. The chambers are known as parallel plate, T-chamber, step, cone plate, and stretch. The stimulated functions or signals from endothelial cells by flows are extensively connected to other outer layers of arteries or organs. The coculture layer was developed in a chamber to investigate the interaction between smooth muscle cells in the middle layer of the blood vessel wall in vascular physiology and pathology. Additionally, the microfabrication technology used to create a chamber for a microfluidic device involves both mechanical and chemical stimulation of cells to show their dynamics in in vivo microenvironments. The purpose of this study is to summarize the blood flow (flow inducing) for the functions connecting to endothelial cells and blood vessels, and to find directions for future chamber and device developments for further understanding and application of vascular functions. The relationship between chamber design flow, cell layers, and microfluidics was studied.
Article
Full-text available
Atherosclerosis, the leading cause of death in the developed world and nearly the leading cause in the developing world, is associated with systemic risk factors including hypertension, smoking, hyperlipidemia, and diabetes mellitus, among others. Nonetheless, atherosclerosis remains a geometrically focal disease, preferentially affecting the outer edges of vessel bifurcations. In these predisposed areas, hemodynamic shear stress, the frictional force acting on the endothelial cell surface as a result of blood flow, is weaker than in protected regions. Studies have identified hemodynamic shear stress as an important determinant of endothelial function and phenotype. Arterial-level shear stress (>15 dyne/cm2) induces endothelial quiescence and an atheroprotective gene expression profile, while low shear stress (<4 dyne/cm2), which is prevalent at atherosclerosis-prone sites, stimulates an atherogenic phenotype. The functional regulation of the endothelium by local hemodynamic shear stress provides a model for understanding the focal propensity of atherosclerosis in the setting of systemic factors and may help guide future therapeutic strategies.
Article
Full-text available
Endothelial cell alignment along the direction of laminar fluid flow is widely understood to be a defining morphological feature of vascular homeostasis. While the role of associated signaling and structural events have been well studied, associated intercellular stresses under laminar fluid shear have remained ill-defined and the role of these stresses in the alignment process has remained obscure. To fill this gap, we report here the tractions and the complete in-plane intercellular stress fields measured within the human umbilical vein endothelial cell (HUVEC) monolayer subjected to a steady laminar fluid shear of 1.2 Pa. Tractions, intercellular stresses, as well as their time course, heterogeneity and anisotropy, were measured using monolayer traction force microscopy and monolayer stress microscopy. Prior to application of laminar fluid flow, intercellular stresses were largely tensile but fluctuated dramatically in space and in time (317+122 Pa). Within 12 hours of the onset of laminar fluid flow, the intercellular stresses decreased substantially, but continued to fluctuate dramatically (142+84 Pa). Moreover, tractions and intercellular stresses aligned strongly and promptly (within 1 hour) along the direction of fluid flow, whereas the endothelial cell body aligned less strongly and substantially more slowly (12 hours). Taken together, these results reveal that steady laminar fluid flow induces prompt reduction in magnitude and alignment of tractions and intercellular stress tensor components followed by the retarded elongation and alignment of the endothelial cell body. Appreciably smaller intercellular stresses supported by cell-cell junctions logically favor smaller incidence of gap formation and thus improved barrier integrity. Copyright © 2015, American Journal of Physiology - Cell Physiology.
Article
Full-text available
Measurements of endothelial cell response to fluid shear stress have previously been performed on unphysiologically rigid substrates. We describe the design and implementation of a microfluidic device that applies discrete levels of shear stress to cells plated on hydrogel-based substrates of physiologically-relevant stiffness. The setup allows for measurements of cell morphology and inflammatory response to the combined stimuli, and identifies mechanisms by which vascular stiffening leads to pathological responses to blood flow. We found that the magnitude of shear stress required to affect endothelial cell morphology and inflammatory response depended on substrate stiffness. Endothelial cells on 100 Pa substrates demonstrate a greater increase in cell area and cortical stiffness and decrease in NF-κB nuclear translocation in response to TNF-α treatment compared to controls than cells plated on 10 kPa substrates. The response of endothelial cells on soft substrates to shear stress depends on the presence of hyaluronan (HA). These results emphasize the importance of substrate stiffness on endothelial function, and elucidate a means by which vascular stiffening in aging and disease can impact the endothelium.
Article
Full-text available
Fundamental biological processes including morphogenesis, tissue repair and tumour metastasis require collective cell motions, and to drive these motions cells exert traction forces on their surroundings. Current understanding emphasizes that these traction forces arise mainly in `leader cells' at the front edge of the advancing cell sheet. Our data are contrary to that assumption and show for the first time by direct measurement that traction forces driving collective cell migration arise predominately many cell rows behind the leading front edge and extend across enormous distances. Traction fluctuations are anomalous, moreover, exhibiting broad non-Gaussian distributions characterized by exponential tails. Taken together, these unexpected findings demonstrate that although the leader cell may have a pivotal role in local cell guidance, physical forces that it generates are but a small part of a global tug-of-war involving cells well back from the leading edge.
Article
Full-text available
Fluid shear stress (FSS) from blood flow acting on the endothelium critically regulates vascular morphogenesis, blood pressure, and atherosclerosis [1]. FSS applied to endothelial cells (ECs) triggers signaling events including opening of ion channels, activation of signaling pathways, and changes in gene expression. Elucidating how ECs sense flow is important for understanding both normal vascular function and disease. EC responses to FSS are mediated in part by a junctional mechanosensory complex consisting of VE-cadherin, PECAM-1, and VEGFR2 [2]. Previous work suggested that flow increases force on PECAM-1, which initiates signaling [2-4]. Deletion of PECAM-1 blocks responses to flow in vitro and flow-dependent vascular remodeling in vivo [2, 5]. To understand this process, we developed and validated FRET-based tension sensors for VE-cadherin and PECAM-1 using our previously developed FRET tension biosensor [6]. FRET measurements showed that in static culture, VE-cadherin in cell-cell junctions bears significant myosin-dependent tension, whereas there was no detectable tension on VE-cadherin outside of junctions. Onset of shear stress triggered a rapid (<30 s) decrease in tension across VE-cadherin, which paralleled a decrease in total cell-cell junctional tension. Flow triggered a simultaneous increase in tension across junctional PECAM-1, while nonjunctional PECAM-1 was unaffected. Tension on PECAM-1 was mediated by flow-stimulated association with vimentin. These data confirm the prediction that shear increases force on PECAM-1. However, they also argue against the current model of passive transfer of force through the cytoskeleton to the junctions [7], showing instead that flow triggers cytoskeletal remodeling, which alters forces across the junctional receptors.
Article
Mechanical forces associated with blood flow play important roles in the acute control of vascular tone, the regulation of arterial structure and remodeling, and the localization of atherosclerotic lesions. Major regulation of the blood vessel responses occurs by the action of hemodynamic shear stresses on the endothelium. The transmission of hemodynamic forces throughout the endothelium and the mechanotransduction mechanisms that lead to biophysical, biochemical, and gene regulatory responses of endothelial cells to hemodynamic shear stresses are reviewed.
Article
The behavior of vascular EC is greatly altered in sites of pathological angiogenesis, such as a developing tumor or atherosclerotic plaque. Until recently it was thought that this was largely due to abnormal chemical signaling, i.e., endothelial cell chemo transduction, at these sites. However, we now demonstrate that the shear stress intensity encountered by EC can have a profound impact on their gene expression and behavior. We review the growing body of evidence suggesting that mechanotransduction, too, is a major regulator of pathological angiogenesis. This fits with the evolving story of physiological angiogenesis, where a combination of metabolic and mechanical signaling is emerging as the probable mechanism by which tight feedback regulation of angiogenesis is achieved in vivo.
Article
The behavior of vascular endothelial cells is greatly altered in sites of pathological angiogenesis, such as a developing tumour or atherosclerotic plaque. Until recently it was thought that this was largely due to abnormal chemical signalling, i.e. endothelial cell chemotransduction, at these sites. However, we now demonstrate that the shear stress intensity encountered by endothelial cells can have a profound impact on their gene expression and behaviour. We review the growing body of evidence suggesting that mechanotransduction, too, is a major regulator of pathological angiogenesis. This fits with the evolving story of physiological angiogenesis, where a combination of metabolic and mechanical signalling is emerging as the probable mechanism by which tight feedback regulation of angiogenesis is achieved in vivo. This article is protected by copyright. All rights reserved.
Article
Endothelial cells react to shear stresses with mechanotransduction responses that modify the cytoskeleton and cell-cell contacts. Cultures of endothelial cells were patterned as monolayers on micropost arrays and different shear flow profiles were applied to investigate the interplay between traction forces, intercellular forces, and adherens junctions (A–C). Cells exposed to laminar flow had elevated traction forces compared to static conditions while cells experiencing unsteady or disturbed flow exhibited lower traction forces (F). Similarly, the size of cell adherens junctions increased after laminar flow and decreased after disturbed flow. Decreasing cytoskeletal tension with Y-27632 decreased the size of adherens junctions (D), while increasing tension through Calyculin-A increased their size (E). A novel approach to measure intercellular forces between cells in the monolayers was developed (G) and these forces were found to be significantly higher for laminar flow than for static or disturbed conditions (H) with adherens junction size reflecting these tension changes.View Large Image | View Hi-Res Image | Download PowerPoint Slide These results indicate that laminar flow can increase cytoskeletal tension while disturbed flow decreases cytoskeletal tension. The corresponding change in cytoskeletal tension under shear can produce intercellular forces that can potentially affect the assembly of adherens junction.