Nhels. 2019. Rigidez de MDCKII Por AFM

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Article

Stiffness of MDCK II Cells Depends on Confluency


and Cell Size
Stefan Nehls,1 Helen Nöding,1 Susanne Karsch,1 Franziska Ries,1 and Andreas Janshoff1,*
1
€ t Göttingen, Institute of Physical Chemistry, Göttingen, Germany
Georg-August-Universita

ABSTRACT Mechanical phenotyping of adherent cells has become a serious tool in cell biology to understand how cells
respond to their environment and eventually to identify disease patterns such as the malignancy of cancer cells. In the steady
state, homeostasis is of pivotal importance, and cells strive to maintain their internal stresses even in challenging environments
and in response to external chemical and mechanical stimuli. However, a major problem exists in determining mechanical prop-
erties because many techniques, such as atomic force microscopy, that assess these properties of adherent cells locally can
only address a limited number of cells and provide elastic moduli that vary substantially from cell to cell. The origin of this spread
in stiffness values is largely unknown and might limit the significance of measurements. Possible reasons for the disparity are
variations in cell shape and size, as well as biological reasons such as the cell cycle or polarization state of the cell. Here, we
show that stiffness of adherent epithelial cells rises with increasing projected apical cell area in a nonlinear fashion. This size
stiffening not only occurs as a consequence of varying cell-seeding densities, it can also be observed within a small area of
a particular cell culture. Experiments with single adherent cells attached to defined areas via microcontact printing show that
size stiffening is limited to cells of a confluent monolayer. This leads to the conclusion that cells possibly regulate their size
distribution through cortical stress, which is enhanced in larger cells and reduced in smaller cells.

INTRODUCTION
Cellular biophysics with emphasis on the mechanical prop- cortex, together with the plasma membrane of such cells,
erties of adherent cells has received increasing attention dominates the repulsive forces that are experienced by nano-
over the past decades, as multiple techniques have shown indenters (13,14). In epithelial cells, mechanical homeosta-
that biomechanics is involved in a variety of processes sis is of major importance and must continue even in
such as cell spreading (1,2), migration (3,4), proliferation, challenging environments, e.g., high or low osmolarity, to
and stem cell differentiation (5–7) and is altered depending ensure layer integrity (15). Disruption of this integrity by
on the malignancy of certain types of cancer (8,9). One way the creation of defects in the cell layer leads to altered me-
to assess the mechanical properties of cells is to externally chanical properties of the remaining cells surrounding the
deform them with a defined force or pressure and measure defect. This change holds for multiple spheres of influence
the response function. Viscoelastic responses to external up to several tens of micrometers away from the defect
mechanical stimuli are found to originate mainly from the (16–19).
actin cortex attached to the plasma membrane (10). Defor- Based on these findings, information on mechanical
mation of cells either globally or locally with a sharp changes within a layer may be naturally transmitted between
indenter is usually interpreted either in terms of a contact cells and poses an ideal way of establishing and maintaining
mechanics model, providing the Young’s modulus, or in mechanical homeostasis within the layer. Homeostasis of
terms of a tension model, in which cortical tension prevails cellular morphology within monolayers is a prerequisite
at low strain, whereas area dilatation dominates at large for their functionality. Although epithelial cells like the
strain (11,12). Here, we focus on the apical mechanics of MDCK II cell line show high proliferation rates under sub-
adherent epithelial cells in terms of contact mechanics, in confluent conditions, their growth is restricted upon reaching
which studies have already shown that the thin actomyosin confluency by contact inhibition (20,21). Although this
might seem like epithelial morphogenesis is finished in early
confluent monolayers of these cells, further development
Submitted November 27, 2018, and accepted for publication April 22, 2019. persists even for days after reaching confluency, during
*Correspondence: ajansho@gwdg.de which cell-cell contacts are strengthened and supercellular
Editor: Christopher Yip. structures like domes or tubules are being formed (22,23).
https://doi.org/10.1016/j.bpj.2019.04.028
Ó 2019 Biophysical Society.

2204 Biophysical Journal 116, 2204–2211, June 4, 2019


Cell Size Mechanics

Although downregulated, cell division still takes place and without Mg2þ or Ca2þ and a solution of bovine collagen I (0.2 mg/mL;
depends on the area occupied by the dividing cell, with larger Thermo Fisher Scientific) in PBS was added and incubated for another
2 h. Finally, the dish was again washed three times with PBS and twice
cells exhibiting higher division rates (24). with MEM, and 100,000 MDCK II cells suspended in 2.5 mL M10F40
In this study, we show that there exists a strong correla- (MEM containing 10% FCS (Biowest), 4 mM L-glutamine (Lonza), peni-
tion between cellular stiffness and the cells’ projected apical cillin/streptomycin (0.2 mg/mL; PAA, Pasching, Germany), 0.25 mg/mL
area, i.e., the occupied space within a cell monolayer. Our amphotericin B (Biochrom), and 40 mM HEPES (Biochrom)) were added
finding points toward a regulation of cell size in a confluent and incubated at 37 C. After 60 min, the sample was rinsed with M10F40 to
remove any nonadherent cells, and atomic force microscopy (AFM) mea-
layer driven by cortical tension. Knowledge of this correla- surements were started.
tion between cell stiffness and projected cell area permits us
to estimate and potentially also to eliminate the often-
observed variation in mechanical measurements due to a AFM
variation in cell size. We suggest following a precise proto- Experiments were performed using MLCT cantilevers (knorm ¼ 0.01
col to ensure that comparable data sets are gathered, in pN/nm; Bruker AFM Probes, Camarillo, CA) on an atomic force micro-
which variations in seeding density and incubation time scope (either MFP-3D Origin; Asylum Research, Santa Barbara, CA or
are kept at a minimum. NanoWizard III; JPK, Berlin, Germany). Samples were kept at 37 C in
pH-buffered media during all experiments. Force maps with a resolution
of 1–8 mm2 per force-distance curve (FDC) were acquired. To determine
the Young’s moduli, cells were indented up to a force of 0.5–1 nN at vertical
MATERIALS AND METHODS speeds of 2–5 mm/s. To receive the size of individual cells and correlate that
to their Young’s modulus, slope maps were calculated from the force-
Cell culture
distance data by baseline correcting and linear fitting of the lowest 20 nm
Madin Darby canine kidney cells (MDCK II; Health Protection Agency, of the data. In confluent cells, cell-cell contacts are much stiffer than central
Salisbury, UK) were cultivated in minimal essential medium (MEM with cell regions, and high slopes therefore mark contact lines (see Figs. S1 and
Earl’s salts, 2.2 g/L NaHCO3; Biochrom, Berlin, Germany) containing S2). These boundaries of cell-cell contacts were then chosen manually and
4 mM L-glutamine (Lonza, Basel, Switzerland) and 10% (v/v) fetal calf used for cell segmentation. In single-cell studies, cell bodies were chosen in
serum (FCS; BioWest, Nuaille, France). Stem and samples were kept at a similar fashion based on the substrate showing extremely steep slopes in
37 C and 7.5% CO2 (Heracell 150i; Thermo Fisher Scientific, Waltham, the FDCs.
MA). Subcultivation was performed using standard culture flasks (TPP, Tra- To determine viscoelastic properties, the cantilever, which has been in-
sadingen, Switzerland) by addition of 0.05% trypsin and 0.02% EDTA dented into the cellular cortex to a depth of about 1 mm, was actively oscil-
(Biochrom) and short incubation to remove the adherent cells from the cul- lated to examine the impact of the cytoskeleton (9,25,26). The cantilever
ture surface. Suspended cells were mixed with FCS and centrifuged, and the was exited to sinusoidal oscillations at different frequencies ranging from
pellet was dissolved in MEM and seeded into fresh flasks. 5 to 100 Hz with small amplitudes. The complex shear modulus G*(u)
Cells destined to serve in experiments of confluent cells were taken in the can then be obtained by the amplitude diminution and the phase shift be-
last step. 220 cells per square millimeter (c/mm2) were seeded into petri tween the in- and output signal. The real part of the resulting complex shear
dishes (Ibidi, Martinsried, Germany) and kept at 37 C and 7.5% CO2 for modulus, the storage modulus G0 (u), describes the frequency-dependent
48 h, unless stated otherwise. Before experiments, cell media were elastic properties, whereas the imaginary part, the loss modulus G00 (u), dis-
exchanged to MEM containing HEPES (15 mM on confluent cells, 40 mM plays the energy dissipation in the system and represents its viscous prop-
on single cells, 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid; Bio- erties. The fraction h(u) ¼ G00 (u)/G0 (u) is the so-called loss tangent, a
chrom) as well as 0.2 mg/mL penicillin/streptomycin (PAA, Pasching, Ger- model free parameter describing the material properties of the cell (9).
many) and 0.25 mg/mL amphotericin B (Biochrom). The complex shear modulus G* increases with frequency following a
weak power law with an exponent b usually in between 0.2 and 0.4. The
loss modulus G00 exhibits lower values than G0 in the low-frequency regime
(<50 Hz). Hence, in this regime, MDCK II cells behave more elastically
Cell patterning rather than viscous. The mild power law and the elastic nature of the cell
To create micrometer-scale patterns on culture dishes, plasma-induced justifies our approach using Young’s moduli from fitting a purely elastic
protein patterning was used. First, patterns were drawn in AutoCAD model to the data. In contrast, at larger frequencies, the cells display
(Autodesk, San Rafael, CA), and the corresponding mask was created more fluid-like properties. A successful attempt to explain this power law
(Compugraphics, Jena, Germany). The passivated, microstructured wafers behavior in G* in the sense of an active soft glassy material has been sug-
were then used to cast polydimethylsiloxane (SYLGARD 184; Dow gested by Fabry and co-workers (27,28). The idea is based on the soft glassy
Corning, Wiesbaden, Germany) molds by mixing polymer and curing agent rheology model, assuming that the cytoskeleton of the cell is held together
in a 10:1 ratio and curing it at 70 C for 4 h. The finished stamp was peeled by weak attractive forces between neighboring elements (28,29). Therefore,
off the wafer. the complex shear modulus G* of living cells is fitted with the power law
Glass-bottomed petri dishes (Ibidi) were cleaned by washing with ultra- structural damping (Eq. 1) using a complex, nonlinear, least-squares fitting
pure water and ethanol and dried in nitrogen stream. The freshly cured pol- routine:
ydimethylsiloxane stamp was cut into 5  5 mm pieces and placed onto the p    b
bp u
glass surface. While in contact, the dish was placed in a plasma cleaner and 
G ¼ G0 Gð1  bÞcos b 1 þ i tan þ ium;
exposed to oxygen plasma for 90 s. Immediately after plasma treatment, a 2 2 u0
solution of poly-L-lysine-graft-poly-ethyleneglycol (0.5 mL, PLL (20)-g
[3.5]-PEG (2)/ tetramethylrhodamine; SuSoS, D€ubendorf, Switzerland)
(1)
was added to each individual stamp, and the sample rested for 1 h at with G0 the scaling factor describing the stiffness of the sample; b,
23 C under exclusion of light. Within the given time period, the solvent the power-law coefficient; u0, the scaling factor of the frequency (set to
evaporated. After incubation, the stamps were removed, and the dish was 1 rad/s); and the viscosity m. G0 can be easily converted in the apparent
washed three times with phosphate-buffered saline (PBS; Biochrom) Young’s modulus E0 by E0 ¼ 2G0(1 þ n), where n is Poisson’s ratio

Biophysical Journal 116, 2204–2211, June 4, 2019 2205


Nehls et al.

(assumed to be 0.3 because cells consist predominately of water, which is necting proteins such as ZO-1, F-actin, or E-cadherin were
incompressible). Notably, the scaling factor of the frequency is set arbi- visually inspected (see Fig. S4). The mechanical properties
trarily, which prevents a direct comparison between the Young’s modulus
obtained from force indentation data using a purely elastic contact
of the cells were obtained from microrheology experiments
mechanics model. as described previously (9,32).
We investigated layers originating from three different
seeding densities (270, 540, and 815 c/mm2). Each sample
Theory was incubated for 48 h, with a clear tendency of smaller
cell sizes for higher seeding densities. On these samples,
Cellular mechanics measured by external probes is often described in terms
of contact mechanics such as the Hertz model, considering the deformation
AFM-based microrheology was performed to probe the
of a semi-infinite fully elastic space by a rigid indenter. Such deformations elastic and viscous responses as a function of cell size.
are usually characterized by a single parameter describing the elastic Essentially, two parameters were obtained from the micro-
response, the so-called Young’s modulus E. Employing a conical indenter rheological spectra: the apparent Young’s modulus, E0,
with half-opening angle q results in the following approximate relationship which describes the elastic stiffness (Fig. 1, top), and the
between the force f and the indentation depth d (30):
loss tangent h, which provides information about the rela-
2E tanðqÞ 2 tion of the viscous and elastic properties at a given fre-
f ¼ d: (2)
pð1  n2 Þ quency (Fig. 1, bottom). We have chosen 100 Hz because
it turned out in a previous study that at this frequency, differ-
In typical force spectroscopy experiments, a force setpoint fSP is given,
and indentation is continued until the specified force setpoint is reached; ences between different cell lines are most pronounced (9).
hence, all acquired data share roughly the same maximal force but different The loss tangent is a quantitative measure of to what extent
indentation depths. The equation above is only true during contact between
probe and samples and does not apply to regions of the FDC without con-
tact. To evaluate FDCs without knowledge of the contact point, which is
usually difficult to determine automatically, we use the integral of the whole
curve, which is invariant to the length of the noncontact region.
Z dSP
2E tanðqÞ 2
Welast ¼ d dd (3)
0 pð1  n2 Þ
We know that indentation depth at the force setpoint equals
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
fSP pð1  n2 Þ
dSP ¼ : (4)
2E tanðqÞ
Together with Eq. 3, the Young’s modulus can be obtained from the
elastic energy Welast via
3
fSP  
E ¼ 2
p 1  n2 : (5)
18 Welast tanðqÞ

This idea of obtaining the elastic modulus automatically from FDCs


without determining the contact point is similar to the force integration to
equal limits, which has been applied to extract relative values (31). Howev-
er, here, we actually obtain the absolute Young’s modulus. The procedure is
strictly correct for ideal data, but the addition of noise might lead to a biased
modulus even for large sampling. For typical FDCs, at least, as they occur
in our experiments, we can show that from simulated and noise-decorated
FDCs, the correct modulus is found with high precision (see Fig. S3).

RESULTS FIGURE 1 Young’s modulus E0 (top) and loss tangent at 100 Hz (bottom)
of MDCK II seeded at different cell densities. Box plots extend from the
Cell mechanics as a function of seeding density 25th to the 75th percentile, and whiskers from the 10th to the 90th. Individ-
ual data points are shown as dots; some outliers are not shown. Four
Our first approach was to seed different amounts of MDCK different cell-seeding densities were used, these being 190, 270, 540 and
II cells on equally sized petri dishes to obtain confluent cell 815 c/mm2. Cells were allowed to grow for 48 h before measurement, and
monolayers with different individual cell size. Seeding high in the case of 190 c/mm2, for 72 h. Number of viscoelastic spectra: n ¼
4092 (>20 cells), 7789 (>20 cells), 2027 (>20 cells), and 1622 (>12 cells),
numbers of cells leads to an early and dense confluent layer
respectively. A rank-sum test was performed to test the null hypothesis that
and consequently to a small average cell size. To verify that the data of the indicated data sets are from populations with equal medians.
a confluent monolayer was formed at each cell-seeding den- Asterisks indicate that the null hypothesis was rejected at the 0.05% (***)
sity, fluorescence stainings of important cell-cell intercon- significance level. To see this figure in color, go online.

2206 Biophysical Journal 116, 2204–2211, June 4, 2019


Cell Size Mechanics

the viscous properties (G00 ) prevail over the elastic properties decreasing the mean area that cells occupy within the layer
(G0 ). For a seeding density of 540 c/mm2, a median Young’s (21,23). To test whether the difference in area on its own is
modulus of E0(540 c/mm2) ¼ 850 Pa was found, which is in responsible for the variation in stiffness, we conducted ex-
good agreement with previously reported Young’s moduli of periments in which we acquired force maps across multiple
the same cell line under similar culture conditions (E0(330– cells within one particular sample and determined the area
540 c/mm2) ¼ 560 Pa (32)). For higher seeding densities, of each individual cell. By doing so, we can circumvent
the Young’s modulus decreased to E0(815 c/mm2) ¼ 460 effects that are caused by different lengths of postconfluent
Pa, whereas lower seeding densities resulted in a higher incubation periods. However, because this includes the
elastic modulus of E0(270 c/mm2) ¼ 1.0 kPa. Hence, higher acquisition and evaluation of large data sets, we used a fully
seeding densities generate smaller cells correlating with automated analysis in the following, yielding solely the
lower Young’s moduli. Notably, E0 obtained from rheolog- Young’s modulus as the description of the cellular elastic
ical data is not directly comparable with the Young’s behavior. This is further justified by the finding that viscous
modulus from fitting a pure elastic contact model to the contribution did not change much with seeding density.
data because E0 scales with the arbitrarily set time t0, here To determine the size distribution of cells under our
set to 1 s. experimental conditions as precisely as possible, we stained
The loss tangent computed at 100 Hz does not follow a for the tight junction protein ZO-1 to facilitate cell segmen-
systematic trend. For all samples, the loss tangent was found tation. Employing automated analysis of fluorescence
to be h > 1; thus, G00 > G0 . Comparing the lowest and highest micrographs over large areas of a culture dish, we receive
cell-seeding densities used in this study, we observed a slight reliable and well-resolved data on cell areas, with the major-
increase from h(270 c/mm2) ¼ 2.04 to h(815 c/mm2) ¼ 2.14. ity of cells taking up between 150 and 350 mm2 (see Fig. 2).
However, intermediate seeding densities showed an overall This distribution corresponds well to previous studies con-
lower loss tangent of h(540 c/mm2) ¼ 1.86. ducted by Puliafito et al. (21).
To further explore this relationship, we decided to The determination of Young’s moduli of individual cells
perform additional experiments with a seeding density of as a function of cell area was performed by AFM indenta-
only 190 c/mm2. Here, the cells grow so sparsely that an in- tion experiments. In these experiments, cell area and
cubation time of 3 days was necessary to obtain a confluent Young’s modulus were extracted concomitantly from the ac-
monolayer. However, after incubation, we found even larger quired force maps. Therein, cell-cell contacts are marked by
cells compared to the other seeding conditions, and the narrow bands of exceptionally stiff behavior, as is confirmed
Young’s modulus was indeed increased in these samples by overlay with optical phase-contrast images (see Fig. S1).
(Fig. 1). We also observed a significant drop of the loss We decided to use the force data for determination of cell
tangent to h(190 c/mm2) ¼ 1.68, which indicates a more borders instead of fluorescence images of fixed samples
solid-like behavior of these larger cells. This is expected because cell position and area change dynamically and
because more prestressed cells are exhibiting a smaller po- could easily vary between the force map acquisition and
wer-law factor, which indicates a less fluid behavior. The successful fixation.
elastic modulus follows the general trend that larger cells Binning the results of cells in bands of 50 mm2, Fig. 3
display a larger stiffness, i.e., E0(190 c/mm2) ¼ 1.3 kPa. shows that for increasing cell area within a monolayer, the
We rationalize this trend by considering that neighboring Young’s modulus indeed increases monotonically from
cells pull more strongly at larger cells, forcing them into a 1 kPa to 3 kPa within sizes ranging from about 100 mm2
highly prestressed state. up to about 600 mm2. Considering that the ratio between
These findings underline the importance of adhering circumference/area decreases for larger areas and because
rigidly to established protocols with respect to both cell- the circumference usually marks the stiffest region of such
seeding density and growth time when extracting mechani-
cal properties from cell culture samples. In fact, we suggest
that evaluation of cell density at the time of experiments
should be performed and the data supplied in biomechanical
studies. The question was now whether the observed effect
was due to variations in cell size or, for instance, polarity of
the cells.

Size-dependent elasticity of confluent cells


The variation in mechanical behavior of cells in monolayers
could be caused by several aspects that might change during FIGURE 2 Kernel density estimates of projected cell areas obtained from
layer maturation. Even after establishment of a confluent ZO-1 stainings of confluent MDCK II cells (n ¼ 506). An average cell area
monolayer, cell density still increases for days, continuously of 278 mm2 is found.

Biophysical Journal 116, 2204–2211, June 4, 2019 2207


Nehls et al.

of the patterned grids were separated far enough to allow for


only a single MDCK II cell per instance to adhere, in which
case cells successfully adapt to the matrix pattern.
Single-spread MDCK II cells show a different size distri-
bution compared to confluent ones. They are lower in height
but produce larger adhesion areas. Therefore, we investigated
cells with adhesion areas of 900, 1200, and 1500 mm2. This
approach not only enabled us to enforce a particular cell size
but also an exact adhesion geometry. Using highly symmetric
patterns—here, disks—we can average results by using the
four symmetry axes when analyzing parameter maps. By
doing so and averaging locally over all investigated cells,
we receive parameter maps with high statistical accuracy to
generate a distribution of stiffness values on different areas
of the cell surfaces. See the recent work of Garcia and Garcia
for a detailed discussion on this topic (33).
FIGURE 3 Plot of the cells’ Young’s modulus (median) as a function of As shown in Fig. 4, circular cells show a small ring
projected apical cell area. Cells within a range of 50 mm2 were collected and
around their perimeter of high Young’s moduli, surrounding
represented in points; error bars represent the standard error of the mean for
all condensed cells’ values. The dotted line (smoothed spline) serves as a a larger central region that is rather soft. Although the cen-
guide to the eye, showing the saturation effect for large areas. Data include tral region with a Young’s modulus of 1.2 kPa is quite
14,326 FDCs taken on 105 cells. similar to values of other studies on confluent cells, the
stiffer perimeter shows moduli that are up to two orders of
cells, larger areas should systematically decrease in median magnitude larger (11,34). The width of this ring differs
stiffness. However, the contrary was observed in our exper- slightly depending on the total adhesion area, but the general
iments, underlining the actual size-dependent stiffening. pattern is strikingly similar between all three tested sizes.
Because the increase in stiffness with occupied area of cells Notably, the modulus depends also on the distance to under-
within a confluent cell monolayer corresponds to our obser- lying substrate. The numbers are therefore rather apparent
vation of stiffness changes depending on initial seeding cell ones in the context of the corresponding experiment.
density (Fig. 1), we suggest that the cell area plays a pivotal The central region on the patterned cells, apart from the
role in cell-stiffness homeostasis. Interestingly, the Young’s circumference ring, represents the majority of the cell sur-
modulus increases in a nonlinear fashion and seems to satu- face. As seen in the one-dimensional kernel density distribu-
rate for very large cell sizes. tion (Fig. 4, bottom), the Young’s modulus is very similar for
all investigated cell sizes, with virtually no effect from the
Single-cell studies
Our experiments concerning size-stiffness correlation of
cells that are part of a confluent layer rely on the natural
size distribution, which is, among other factors, also a result
of seeding density and time. This brings up the question as
to what extent the size dependence of cell stiffness is a
feature of confluent cells and originates from neighboring
cells pulling more strongly to enlarge the area or is, instead,
a generic property. Moreover, the number of cells exhibiting
sizes at the lower or upper end of this distribution is very
low. To force a larger number of cells into extreme sizes
and thereby increase the significance of results, we used mi-
cropatterned culture substrates. These substrates consist of
well-defined areas coated with collagen I to foster adhesion
of cells surrounded by coatings with nonadhesive polymer, FIGURE 4 Elasticity of micropatterned cells. Parameter maps of the log-
forcing the cells to occupy and adapt their shape to that of arithmic Young’s modulus (top) at a resolution of 1.5  1.5 mm show softer
the extracellular matrix patterns. MDCK II cells have shown central regions and stiffer peripheries. For all adhesion areas (blue ¼
900 mm2, red ¼ 1200 mm2, green ¼ 1500 mm2), the Young’s modulus of
that if the patterns are packed densely enough for cells to
the central region is almost the same at 1.2 kPa. The stiffer perimeter has
form cell-cell contacts and form a confluent layer, the no dedicated peak in the one-dimensional distribution, with values up to
cell-cell adhesion dominates, and cells essentially disregard two orders of magnitude larger than the softer interior part. Nblue ¼ 2834,
the matrix patterns (unpublished data). Therefore, instances Nred ¼ 2924, Ngreen ¼ 2362. To see this figure in color, go online.

2208 Biophysical Journal 116, 2204–2211, June 4, 2019


Cell Size Mechanics

different adhesion areas. The stiffer perimeter of the cells, stiffness for higher amounts of cells seeded. Higher amounts
however, displays a size-dependent Young’s modulus, i.e., of seeded cells do also result in higher cell density after an
E900 ¼ 73.0 kPa, E1200 ¼ 18.6 kPa, and E1500 ¼ 29.9 kPa. equal amount of incubation time and therefore lead to smaller
In contrast to the confluent studies, this is not a simple stiff- cell areas (see Fig. S4). Apart from the decreasing Young’s
ening with increased size because the intermediate adhesion modulus, we also found that the loss tangent at 100 Hz drops
area results in the softest cell periphery, whereas the small- moderately with decreasing cell-seeding density. This means
est adhesion area leads to the softest cell periphery. These that cells behave more like solids if larger. This corresponds
results suggest that the mechanical behavior of MDCK II well to the general behavior of cells in the framework of soft
cells follow distinctively different rules depending on glassy materials in which stiffer cells also generate less en-
whether or not a confluent layer is present, which further ergy dissipation. However, care needs to be taken when cell
supports the idea of cell mechanics playing a role in layer densities are very low because polarization of cells might
organization and vice versa. Still, these results are surpris- not be completed yet (39). These findings once more under-
ing, given recent studies by Efremov et al. who compared line the importance of using equal incubation conditions for
the stiffness of vero cells either in single-spread or confluent all samples to provide correct comparability between
stages and observed stiffer behavior for single cells in a measurements.
setup similar to ours (35). Also, similar studies of single Further, given the correlation between seeding and incu-
cells on adhesive islands (human mesenchymal stem cells bation conditions with both elasticity and cell adhesion area,
and lung human microvascular endothelial cells) showed we wanted to elucidate whether we can also find a correla-
size stiffening in the respective studies, which we did not tion between cell size and elasticity in a single confluent
observe in our experiments on MDCK II cells (36,37). layer independent of seeding density or other factors. Fortu-
nately, even monoclonal cultures show cells with a broad
size distribution, and although the exact regulatory mecha-
DISCUSSION
nisms involved in size homeostasis are still under debate, in-
The mechanical properties of epithelial cells are of major fluence of cell size on processes like division rates has been
importance for their biological function as efficient barrier shown to exist (24,40,41). Alterations of cell mechanics
tissue. Therefore, epithelia have to ensure that their struc- depending on size may pose a way for cells to recognize
tural integrity is not compromised. One prime example for their current size and alter their biochemistry appropriately.
such cells is the MDCK II cell line, which forms a rather This is especially interesting because different stiffness and
flexible and fluid two-dimensional layer of cells where inter- thereby actomyosin network architectures can carry infor-
cellular gaps are tightly closed to provide control over the mation even from mother to daughter cells and should there-
transepithelial flow. In this study, we used MDCK II cells fore contribute to the ongoing discussion on how cells could
grown on standard culture surfaces and performed site-spe- be aware of their very own size (42).
cific nanoindentation experiments to probe the repulsive We have shown that for MDCK II cells, a decently broad
force in response to deformation. This permitted us to size distribution can be obtained, and by determining the
gain information about the apical mechanics of these cells. exact size and median Young’s modulus of individual cells,
We employ simple Hertzian mechanics here and refrain we were able to show that larger cells tend to have higher
from more complex viscoelastic models to keep the matter stiffness even if originating from the same cell layer. Inter-
simple. Essentially, more sophisticated models based on estingly, we found a trend including saturation effects for
shell mechanics also generate the same qualitative picture. large sizes. Such saturation has also been observed when
The goal of this work was to elaborate on the distribution investigating cells seeded on gels of different elasticity,
of mechanical properties between multiple cells within where cell stiffness correlates with substrate stiffness (37).
confluent layers and to search for correlations between Interestingly, we did not find a window of optimal size
cellular mechanics and the occupied area of cells within compared to Miettinen et al., who observed a peak metabolic
these layers. From a tensegrity point of view, we expect activity for an optimal cell size (43). This shows that the way
larger structures if neighboring cells exert stress through cellular properties correlate with cell size is diverse and com-
adherens junctions, leading to larger restoring forces upon plex in nature, and cell stiffness is one of these properties.
indentation. Lastly, we tested whether or not we can use single
Therefore, we initially performed rheological experiments patterned cells to find the same correlations we found in
on samples derived from cells seeded at different initial cell the confluent samples already. In other words, we wanted
densities. Force indentation curves were subjected to a to find out whether the size-dependent stiffness is a univer-
theoretical framework that has been well established sal feature or only occurs in confluent layers. Although we
before (9), and we found Young’s moduli that are comparable had to rely on much larger adhesion areas for single-cell
to those found in previous studies (11,15,38). However, studies, this approach provided better reproducibility of
by comparing the results of samples with different cell den- exact cell sizes and yielded laterally resolved parameter
sities, we found that there is a significant trend to decreasing maps of the apical mechanics of cells (44). The Young’s

Biophysical Journal 116, 2204–2211, June 4, 2019 2209


Nehls et al.

moduli of the stiffer, peripheral region of single patterned experiments on confluent cells related to individual cell size. S.N. per-
MDCK II cells do depend on the total adhesion area, formed analysis on these data sets. Single-cell studies were performed
and analyzed by S.N. S.N. and A.J. wrote the manuscript.
whereas the central region shows similar Young’s moduli
regardless of the adhesion area. Previous studies by Tee
et al. and Roca-Cusachs et al. did not consider this differ- ACKNOWLEDGMENTS
ence in behavior depending on the exact location on the
cell surface and could not report such bimodal distribution We gratefully acknowledge financial support from SFB 937 (A17) and SPP
because the authors did not acquire force maps across the 1782. H.N. additionally thanks the Deutsche Telekom Stiftung for financial
support.
complete cell surface. It is therefore even more surprising
that neither of the surface regions, which we managed to
distinguish by mechanical behavior, show the same changes REFERENCES
that have been observed on other cell lines in the respective
studies. This emphasizes how fundamentally different the 1. Raucher, D., and M. P. Sheetz. 2000. Cell spreading and lamellipodial
extension rate is regulated by membrane tension. J. Cell Biol. 148:127–
mechanical behavior of cells is regulated depending on 136.
cell type and how strongly it is influenced by the environ- 2. Pietuch, A., and A. Janshoff. 2013. Mechanics of spreading cells
mental conditions. probed by atomic force microscopy. Open Biol. 3:130084.
In larger cells, contractile forces and internal stresses are 3. Li, S., J. L. Guan, and S. Chien. 2005. Biochemistry and biomechanics
presumably enhanced. A similar trend was previously found of cell motility. Annu. Rev. Biomed. Eng. 7:105–150.
when cells were cultured on pores. Cells on larger pores 4. Plotnikov, S. V., A. M. Pasapera, ., C. M. Waterman. 2012. Force fluc-
tuations within focal adhesions mediate ECM-rigidity sensing to guide
were smaller and softer (32). directed cell migration. Cell. 151:1513–1527.
5. Lv, H., L. Li, ., Y. Li. 2015. Mechanism of regulation of stem cell dif-
ferentiation by matrix stiffness. Stem Cell Res. Ther. 6:103.
CONCLUSION
6. Olivares-Navarrete, R., E. M. Lee, ., Z. Schwartz. 2017. Substrate
We scrutinized the connection between the size of cells and stiffness controls osteoblastic and chondrocytic differentiation of
mesenchymal stem cells without exogenous stimuli. PLoS One.
their apical mechanics, showing that in confluent MDCK II 12:e0170312.
cells, stiffness is increased in samples with lower initial cell- 7. Narayanan, K., V. Y. Lim, ., J. Y. Ying. 2014. Extracellular matrix-
seeding density. Even in one cell monolayer, the stiffness of mediated differentiation of human embryonic stem cells: differentia-
individual cells is increased with increasing apical surface tion to insulin-secreting beta cells. Tissue Eng. Part A. 20:424–433.
area. This shows that special caution is required when 8. Suresh, S. 2007. Biomechanics and biophysics of cancer cells. Acta
Biomater. 3:413–438.
comparing mechanical studies of cells carried out under
9. Rother, J., H. Nöding, ., A. Janshoff. 2014. Atomic force micro-
different culture conditions. scopy-based microrheology reveals significant differences in the
Interestingly, this behavior could not be observed for sin- viscoelastic response between malign and benign cell lines.
gle patterned cells, where we found that mechanics differ be- Open Biol. 4:140046.
tween certain locations on a cell’s surface, but the stiffness 10. Br€uckner, B. R., H. Nöding, and A. Janshoff. 2017. Viscoelastic prop-
erties of confluent MDCK II cells obtained from force cycle experi-
obtained from the cell’s center does not depend on the cell’s ments. Biophys. J. 112:724–735.
size. This emphasizes the importance of cell-cell contacts for
11. Pietuch, A., B. R. Br€uckner, ., A. Janshoff. 2013. Elastic properties of
the elastic response of cells, as recently shown (38). Taken cells in the context of confluent cell monolayers: impact of tension and
together, these findings show that cellular mechanics react surface area regulation. Soft Matter. 9:11490–11502.
to environmental conditions in complex ways and that ten- 12. Krieg, M., G. Fl€aschner, ., D. J. M€uller. 2019. Atomic force micro-
sion homeostasis and cell size are inherently coupled. The scopy-based mechanobiology. Nat. Rev. Phys:1:41–57. Published
online November 1, 2018.
further understanding of the correlation between mechanical
13. Rouven Br€uckner, B., A. Pietuch, ., A. Janshoff. 2015. Ezrin is a
and biochemical properties of cells should ultimately lead to major regulator of membrane tension in epithelial cells. Sci. Rep.
a better understanding of how these cells are able to respond 5:14700.
appropriately to a large variety of different environmental 14. Nehls, S., and A. Janshoff. 2017. Elastic properties of pore-spanning
conditions and still maintain performance. apical cell membranes derived from MDCK II cells. Biophys. J.
113:1822–1830.
15. Pietuch, A., B. R. Br€uckner, and A. Janshoff. 2013. Membrane tension
homeostasis of epithelial cells through surface area regulation in
SUPPORTING MATERIAL response to osmotic stress. Biochim. Biophys. Acta. 1833:712–722.
Supporting Material can be found online at https://doi.org/10.1016/j.bpj. 16. Tambe, D. T., C. C. Hardin, ., X. Trepat. 2011. Collective cell guid-
2019.04.028. ance by cooperative intercellular forces. Nat. Mater. 10:469–475.
17. Das, T., K. Safferling, ., J. P. Spatz. 2015. A molecular mechanotrans-
duction pathway regulates collective migration of epithelial cells. Nat.
AUTHOR CONTRIBUTIONS Cell Biol. 17:276–287.
18. Karsch, S., D. Kong, ., A. Janshoff. 2017. Single-cell defects cause a
H.N. and F.R. performed experiments and analysis of the experiments long-range mechanical response in a confluent epithelial cell layer.
regarding different cell-seeding densities. S.N., H.N., and S.K. performed Biophys. J. 113:2601–2608.

2210 Biophysical Journal 116, 2204–2211, June 4, 2019


Cell Size Mechanics

19. Vishwakarma, M., J. Di Russo, ., J. P. Spatz. 2018. Mechanical inter- 32. Rother, J., M. B€uchsensch€utz-Göbeler, ., A. Janshoff. 2015. Cyto-
actions among followers determine the emergence of leaders in skeleton remodelling of confluent epithelial cells cultured on porous
migrating epithelial cell collectives. Nat. Commun. 9:3469. substrates. J. R. Soc. Interface. 12:20141057.
20. Li, S., E. R. Gerrard, Jr., and D. F. Balkovetz. 2004. Evidence for 33. Garcia, P. D., and R. Garcia. 2018. Determination of the elastic moduli
ERK1/2 phosphorylation controlling contact inhibition of proliferation of a single cell cultured on a rigid support by force microscopy.
in Madin-Darby canine kidney epithelial cells. Am. J. Physiol. Cell Biophys. J. 114:2923–2932.
Physiol. 287:C432–C439. 34. Br€uckner, B. R., and A. Janshoff. 2015. Elastic properties of epithelial
21. Puliafito, A., L. Hufnagel, ., B. I. Shraiman. 2012. Collective and sin- cells probed by atomic force microscopy. Biochim. Biophys. Acta.
gle cell behavior in epithelial contact inhibition. Proc. Natl. Acad. Sci. 1853:3075–3082.
USA. 109:739–744. 35. Efremov, Y. M., A. A. Dokrunova, ., K. V. Shaitan. 2013. The effects
22. Cao, F., and J. M. Burke. 1997. Protein insolubility and late-stage of confluency on cell mechanical properties. J. Biomech. 46:1081–
morphogenesis in long-term postconfluent cultures of MDCK epithelial 1087.
cells. Biochem. Biophys. Res. Commun. 234:719–728. 36. Roca-Cusachs, P., J. Alcaraz, ., D. Navajas. 2008. Micropatterning of
23. Rothen-Rutishauser, B., S. D. Kr€amer, ., H. Wunderli-Allenspach. single endothelial cell shape reveals a tight coupling between nuclear
1998. MDCK cell cultures as an epithelial in vitro model: cytoskeleton volume in G1 and proliferation. Biophys. J. 94:4984–4995.
and tight junctions as indicators for the definition of age-related stages 37. Tee, S. Y., J. Fu, ., P. A. Janmey. 2011. Cell shape and substrate rigid-
by confocal microscopy. Pharm. Res. 15:964–971. ity both regulate cell stiffness. Biophys. J. 100:L25–L27.
24. Puliafito, A., L. Primo, and A. Celani. 2017. Cell-size distribution in 38. Br€uckner, B. R., and A. Janshoff. 2018. Importance of integrity of
epithelial tissue formation and homeostasis. J. R. Soc. Interface. cell-cell junctions for the mechanics of confluent MDCK II cells.
14:20170032. Sci. Rep. 8:14117.
25. Mahaffy, R. E., S. Park, ., C. K. Shih. 2004. Quantitative analysis of 39. Schneider, D., T. Baronsky, ., A. Janshoff. 2013. Tension monitoring
the viscoelastic properties of thin regions of fibroblasts using atomic during epithelial-to-mesenchymal transition links the switch of pheno-
force microscopy. Biophys. J. 86:1777–1793. type to expression of moesin and cadherins in NMuMG cells. PLoS
26. Mahaffy, R. E., C. K. Shih, ., J. K€as. 2000. Scanning probe-based fre- One. 8:e80068.
quency-dependent microrheology of polymer gels and biological cells. 40. Vargas-Garcia, C. A., M. Soltani, and A. Singh. 2016. Conditions for
Phys. Rev. Lett. 85:880–883. cell size homeostasis: a stochastic hybrid system approach. IEEE
27. Kollmannsberger, P., and B. Fabry. 2009. Active soft glassy rheology of Life Sci. Lett. 2:47–50.
adherent cells. Soft Matter. 5:1771–1774. 41. Vargas-Garcia, C. A., K. R. Ghusinga, and A. Singh. 2018. Cell size
28. Kollmannsberger, P., and B. Fabry. 2011. Linear and nonlinear control and gene expression homeostasis in single-cells. Curr. Opin.
rheology of living cells. Annu. Rev. Mater. Res. 41:75–97. Syst. Biol. 8:109–116.
29. Sollich, P. 1998. Rheological constitutive equation for a model of soft 42. Ginzberg, M. B., R. Kafri, and M. Kirschner. 2015. Cell biology. On
glassy materials. Phys. Rev. E Stat. Phys. Plasmas Fluids Relat. Inter- being the right (cell) size. Science. 348:1245075.
discip. Topics. 58:738–759. 43. Miettinen, T. P., and M. Björklund. 2016. Cellular allometry of mito-
30. Sneddon, I. N. 1965. The relation between load and penetration in the chondrial functionality establishes the optimal cell size. Dev. Cell.
axisymmetric Boussinesq problem for a punch of arbitrary profile. Int. 39:370–382.
J. Eng. Sci. 3:47–57. 44. Rigato, A., F. Rico, ., S. Scheuring. 2015. Atomic force microscopy
31. A-Hassan, E., W. F. Heinz, ., J. H. Hoh. 1998. Relative microelastic mechanical mapping of micropatterned cells shows adhesion geome-
mapping of living cells by atomic force microscopy. Biophys. J. try-dependent mechanical response on local and global scales. ACS
74:1564–1578. Nano. 9:5846–5856.

Biophysical Journal 116, 2204–2211, June 4, 2019 2211


Biophysical Journal, Volume 116

Supplemental Information

Stiffness of MDCK II Cells Depends on Confluency and Cell Size


Stefan Nehls, Helen Nöding, Susanne Karsch, Franziska Ries, and Andreas Janshoff
Supporting Material

Stiffness of MDCK II Cells Depends on Confluency and


Cell Size
S. Nehlsa , H. Nödinga , S. Karscha , F. Riesa , A. Janshoffa∗
a
Georg-August-Universität Göttingen, Institute of Physical Chemistry,
37077 Göttingen, Germany

Figure S1: Stiffness map showing the slope of a linear fit of the last 20 nm of each FDC (top left, green).
Cell-cell borders are marked by bright areas. Comparing this to the phase-contrast image of the same
sample, nuclei and cell borders are visible (bottom left). The right image shows an overlap of both
images, confirming that cell-cell contacts are identified in stiffness maps. Small deviations between both
images can occur due to thermal drift since the acquisition of the force map takes several minutes. Box
sizes are 100 µm x 100 µm.

1
Figure S2: Representative height topography scan and associated force map analyzed with the Hertzian
contact model implemented in the JPK Data Processing Software (quadratic pyramid tip shape with half-
angle to edge of 17.5◦ and ν = 0.5) showing elevated Young’s moduli at cell-cell junctions.

2
Figure S3: Proof-of-principle of the fit routine to obtain the Young’s modulus from force distance curves.
Force distance data is simulated for a sample stiffness of 1 kPa and an indentation up to a force of 500 pN
(top, blue). The data was realistically sampled and Gaussian noise was added with typical amplitude.
The resulting fit is shown as dashed red line. The process was repeated for 10,000 times (center), each
time with random noise, and the distribution of the received modules from the noisy data are shown as
a kernel density plot (blue), with the red line marking the originally simulated Young’s modulus. The
real modulus is very well represented by the results. Fits can be filtered by the fit quality (bottom panel),
where interestingly higher quality factor, hence larger deviations of the fit, lead to a higher number of fits
that are accepted and a better overall estimation of the real modulus. This also shows that for our data,
quality factors larger than one can be just due to the random noise.

3
A1 A2 A3 A4

270 c/mm²

B1 B2 B3 B4 40 µm

540 c/mm²

Figure S4: Examplary immunostainings of MDCK II cells at either 270 c/mm2 (A) or 540 c/mm2 (B).
Shown are the basal (1) and apical (2) layers of f-actin stainings and E-cadherin (3) and ZO-1 (4) stainings.
All show a densely packed and closed cell monolayer despite large cell size differences.

You might also like