Download Energy barriers and cell migration in densely packed tissues†

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Endomembrane system wikipedia , lookup

Biochemical switches in the cell cycle wikipedia , lookup

Cell encapsulation wikipedia , lookup

Cell cycle wikipedia , lookup

Extracellular matrix wikipedia , lookup

Tissue engineering wikipedia , lookup

Cell growth wikipedia , lookup

Cytokinesis wikipedia , lookup

Cellular differentiation wikipedia , lookup

JADE1 wikipedia , lookup

Cell culture wikipedia , lookup

Mitosis wikipedia , lookup

Amitosis wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

List of types of proteins wikipedia , lookup

Transcript
Energy barriers and cell migration in densely packed tissues†
Dapeng Bi∗ , J. H. Lopez∗ , J. M. Schwarz∗ and M. Lisa Manning∗a
Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX
First published on the web Xth XXXXXXXXXX 200X
DOI: 10.1039/b000000x
Recent observations demonstrate that confluent tissues exhibit features of glassy dynamics, such as caging behavior and dynamical heterogeneities, although it has remained unclear how single-cell properties control this behavior. Here we develop numerical
and theoretical models to calculate energy barriers to cell rearrangements, which help govern cell migration in cell monolayers.
In contrast to work on sheared foams, we find that energy barrier heights are exponentially distributed and depend systematically
on the cell’s number of neighbors. Based on these results, we predict glassy two-time correlation functions for cell motion, with a
timescale that increases rapidly as cell activity decreases. These correlation functions are used to construct simple random walks
that reproduce the caging behavior observed for cell trajectories in experiments. This work provides a theoretical framework for
predicting collective motion of cells in wound-healing, embryogenesis and cancer tumorogenesis.
1
Introduction
Many important biological processes, including embryogenesis 1,2 , wound healing 3,4 , and tumorigenesis 5,6 , require cells to
move through tissues.
While numerous studies have quantified cell motility by
analyzing isolated cells in controlled environments 7,8 , recent
work has highlighted that cell motion in densely packed tissues is collective, and very different from isolated cell motion.
In densely packed or confluent tissues (no gaps between cells)
researchers have discovered signatures of collective motility
such as dynamical heterogeneities 9,10 and caging behavior 11 .
These signatures also occur in many glassy non-biological
materials, including polymers, granular materials, and foams
12 . They can be understood in terms of the potential energy
landscape, which specifies the total potential energy of a material as a function of the positions of all the degrees of freedom,
such as the particle positions. A glassy material spends most
of its time close to a mechanically stable minimum in the potential energy landscape, but rare fluctuations can overcome
the high energy barriers and allow the material to escape to
a new minimum. These collective, rare fluctuations typically
involve a particle escaping from a cage generated by its neighbors.
Inactive materials such as dry foams are jammed at confluence. Therefore, individual elements do not change neighbors
unless a sufficient external force is applied at the boundaries.
† Electronic Supplementary Information (ESI) available: [details of any
supplementary information available should be included here]. See DOI:
10.1039/b000000x/
∗ Department of Physics, Syracuse University, Syracuse, NY 13244, USA Fax:
+1 315 443 9103; Tel: +1 315 443 3920; E-mail: [email protected]
b Syracuse Biomaterials Institute, Syracuse, NY 13244, USA
Much effort has focused on understanding these rearrangements that occur when energy is injected globally; they tend
to occur at special weak regions or soft spots in the material 13
and the energy barriers to rearrangements are power-law distributed 14 .
Even in the absence of external forces, cells in confluent tissues regularly intercalate, or exchange neighbors 15 . They actively change their shapes and exert forces on contacts to overcome large mechanical energy barriers and transition from one
metastable state to another. Because energy is injected locally,
instead of globally at the boundaries, we hypothesize that the
statistics of energy barriers explored by cells might be very
different from those in inactive materials. The fact that glassy
dynamics are observed in confluent tissues suggests that cell
migration rates are governed by these energy barriers. In other
words, cell motility in tissues is set not by single-cell migration rates but instead by the rate at which cells can squeeze
past neighbors.
There is no existing theoretical framework for predicting
cell migration rates in confluent tissues. Although several
recent particle-based models for collective cell motion show
signatures of glassy dynamics 16,17 , these break down at confluence and do not capture changes to cell shapes that occur
during intercalation.
In this Communication, we develop a framework for predicting cell migration rates in tissues by first calculating energy barriers to cell rearrangements. We find that the distribution of energy barriers for local rearrangements is exponentially distributed, which is precisely the distribution required
for glassy dynamics in non-active matter 18 , and different from
that observed in foams. Our simulation and model also predict
that the height of the energy barriers depends systematically
on the topology of cell neighbors in the vicinity of the rear1–6 | 1
rangement. We utilize the ’trap’ model 18 and an extension of
the Soft Glassy Rheology (SGR) model 19 to convert our results for energy barrier distributions to testable predictions for
cell migration, including waiting times and two-time correlation functions. Finally, we carry out a minimal random walks
based on these two-time correlation functions which capture
caging and migration of cells and make qualitative comparisons to experiments.
Shape equilibrium or vertex models have been successfully used to predict the minimum energy shapes of 2D crosssections 3D cells in confluent tissues 1,20–22 . These models
develop an equation for the mechanical energy of a cell,
Ui = ξ Pi2 + γPi + β (Ai − A0 )2 ,
(1)
where Pi and Ai are the perimeter and area of the cell. Coarsegrained mechanical properties of single cells that influence
cell shape, which are discussed in 1,23 , include cortical elasticity, cortical surface tension, bulk incompressibility, and cellcell adhesion. The term quadratic in the perimeter accounts for
the elastic contractility of the actomyosin based cortex, with
modulus ξ . An effective ‘line tension’ γ couples linearly to the
perimeter. γ can be negative or positive and represents effects
due to cell-cell adhesion and cortical tension. The last term
quadratic in the area accounts for the bulk elasticity and additional cell-cell adhesion effects 20 . Quantities in Eq. (1) can
S1 E2
E1 S2
0.61
0.60
0.59
0.58
A
0.57
0.56
2.602
2.602
Etot /Ncells Etot /Ncells
Etot /Ncells Etot /Ncells
Etot /Ncells Etot /Ncells
utissue
S1 E2
E1 S2
S1 E2
E1 S2
B
2.602
2.600
2.600
2.598
2.598
2.598
2.596
2.596
E2.596
1 E
2.594
2.594
2.594
2.592
2.592‡
2.592
‡
2.590
2.590
2.590
2.590
2.590
2.585
2.585
2.585
2.580
2.580
2.580
2.575
2.575
2.575
2.570
2.570
2.570
2.600
uAB
‡‡‡ ‡
‡‡‡
‡‡‡
‡
‡
‡
‡
‡
‡
‡
‡
‡
‡
‡
‡‡ ‡‡‡‡ ‡‡ ‡‡
‡‡
‡
‡
‡
‡
‡
‡‡ ‡
‡
1
‡
E1
‡
‡
‡
E2
E2
‡
‡
‡
‡
‡
2.590
0.2 0.1
-0.2 -0.1-0.2 -0.1 0.0-0.1 0.0
0.0 0.1
−0.32.595-0.2−0.2
−0.1 0.0 0.10.1
0.2
2.595
2.595
`↵
‡
‡
‡‡
‡
‡
‡
‡
‡
E2
‡
‡
E2
C
‡
0.2
0.30.2
‡
‡E2
E2
‡
‡‡‡ ‡‡
‡
‡‡‡‡ ‡‡
‡
‡
‡‡‡‡
‡
‡
‡
‡
‡
‡‡
Fig. 1 A T1 transition and its typical
profile
from our
‡‡energy
‡‡
‡‡
E1 ‡E1‡ ‡ E‡1 ‡ E3‡ E3 E3
simulation. Cells E1 and E2 ‡share
an
edge
before
the
T1 and
‡
‡
become disjoint after the T1, while S1 and S2 are disjoint before the
T1 and share an edge after the T1. The energy increases as the edge
L L L
separating cells S1 and S2 decreases in length, and reaches a
maximum at length zero. A T1 swap takes place and then the energy
decreases as the edges separating E1 and E2 grows in length. The
energy difference ∆uAB marks the height of the energy barrier
associated with this transition.
2.565
2.565
-0.3
-0.2 2.565
-0.1
-0.3
-0.2
-0.3
0.0
-0.1
-0.2
0.1
0.0
-0.1
0.2
0.1
0.0
0.3
0.2
0.1
0.3
0.2
0.3
be non-dimensionalized
by an energy scale β A20 and a length
√
scale A0 :
utissue = ∑ ui ;
ui = κ p2i + 2κ p0 pi + (ai − 1)2 ,
i
2|
1–6
(2)
3/2
with κ = ξ /(β A0 ) and 2κ p0 = γ/(β A0 ).
This mechanical energy functional has been remarkably
successful in predicting cell shapes in embryonic tissues 1,21
and it allows for anisotropic interactions between cells. Although a few researchers have used these models to investigate
cell growth and division 1,20 , they have not been used to make
predictions about cell migration.
Standard methods 24 were used to generate a random 2D
pattern of N points, which was then mapped to a packing of N
polygons with periodic boundary conditions via voronoi tessellation. The program Surface Evolver 25 was used to find
the nearest local minimum of Eq. (2) via a steepest descent
algorithm.
Under confluent conditions, cells can only rearrange via T1
topological swaps, as illustrated in Fig. 1. Although cell division and death can lead to fluid-like behavior 26 , these are
not necessary for cell migration 2,11 and therefore we study
cell packings in the absence of these processes. To induce a
T1 transition at an edge, the total energy is minimized while
the length of the edge `α is actively decreased from Lα until the edge reaches zero length. Such processes are common
during planar junction remodeling in epithelial layers 15 . A
topological swap takes place at `α = 0. The new edge is actively increased to a length Lα and then allowed to relax to
its final unconstrained minimized state. Except for this T1
transition, the topology of the network of vertices and edges
remains fixed. We have also studied systems where passive
energy-minimizing T1 transitions are allowed in addition to
the active T1 transition, and this does not change any of the
results reported below 27 .
Fig. 1 shows the total energy of the system as a function
of the edge length during a typical T1 transition. The length
`α is displayed as a negative number before the T1 transition
and positive after the T1 transition. The energy barrier for
this process ∆uAB is defined as the minimum energy required
to escape state A towards another stable state C. Statistics of
∆u are collected by testing the T1 transition path on six randomly generated tissues each consisting of N = 64 cells. For
all cells in a tissue, we set the parameters such that the minimal
shape for each cell is a regular hexagon of area 1: κ = 1 and
p0 ≈ 3.722. The distribution of energy barriers ρ(∆u) of these
transitions is shown in Fig. 2(b). The tail obeys an exponential
distribution:
ρ(∆u) ∝ e−c ∆u/h∆ui = e−∆u/ε0 ,
(3)
where fitting has determined c = 1.18 and we define ε0 =
h∆ui/c. This exponential distribution is robust to changes
in model parameters κ, p0 , cell-to-cell variations (A0 → A0i )
and the method we use to initialize cell locations 27 . Our data
suggests that the exponential tails ultimately arise from an interplay between the statistics of edge lengths and the energy
∗ The dependence on the topological measure of E1 and E2 is not included
because the Aboav-Weaire law holds for our cellular packings (Fig. S1), and
therefore the topology of E1 and E2 are strongly constrained by QS .
1
0
10
(a)
P ( u/h ui)
u(L↵ )) / u
(u
0.5
0
−1
(b)
−1
10
−2
10
−3
−0.5
0
10
0
1
2
l↵ /L↵
3
4
5
6
u/h ui
f (Qs )
(c)
P ( u)
1.00
0.75
u
functional. Although the initial T1 edge lengths Lα are Gaussian distributed, we find that the change in energy due to a
reduction in cell perimeter is quadratic in Lα , resulting in an
exponential distribution for energy barriers.
Whereas simulations of sheared foams generically generate power-law distributed energy barriers with an exponential
cutoff 14 , exponential energy barriers appear to be a unique
feature of confluent tissues where energy is injected locally.
This is intriguing because it is precisely the distribution seen
in glassy systems with quenched disorder 18 .
In 28 it was shown that the the ground state of Eq. (2) forms
an ordered hexagonal lattice. However, cells in a biological tissue vary significantly in their number of neighbors or
contact topologies, giving rise to a highly degenerate set of
metastable states. The T1 transitions explore these metastable
states and we find an interesting dependence of the energy barrier heights on the local contact topology of cells involved. As
depicted in Fig. 1, cells S1 and S2 both gain one neighbor
while E1 and E2 lose one neighbor each after the transition.
To quantify the dependence of the energy barrier heights on
the local topology, we capture the local topology of four cells
with the measure QS = (6 − ZS1 ) + (6 − ZS2 ) where ZS1 and
ZS2 are the number of neighbors for cells S1 and S2 ∗ . Higher
values of QS correspond to S1, S2 pairs with fewer neighbors.
After a T1 transition, QS is always reduced by 2. In Fig. 2(c)
the energy barriers are categorized by their pre-T1 QS values.
∆u decreases monotonically with increasing QS and becomes
vanishingly small when QS = 2 (which becomes a QS = 0 state
after a T1 transition). This hints that the hexagonal configuration (all Z 0 s = 6) is not only the energetically preferred
state, but configurations further away from the ground state
also have higher energy barriers.
We observe that during a T1 transition most of the change
in energy is localized to the four cells S1, S2, E1 and E2 that
participate. Based on this observation, we develop a simple
mean-field model, which considers all four cells involved in
a T1 transition
√ to be initially regular polygons of equal edge
length ` = 2/33/4 ≈ 0.62. We allow only the coordination
of S1 and S2 to vary independently, and set ZE1 = ZE2 = 6,
the average value required by the Gauss-Bonnet theorem.
The total energy for the four cells can be calculated for
the transition path, yielding a generic profile for the energy
leading up to the T1 transition, shown by the black line in
Fig. 2(a), that is remarkably similar to simulation results. The
mean-field model also predicts the energy barrier height ∆um f
as a function of the topology of the cells involved, as shown
by red dotted line in Fig. 2(c). With no fitting parameters,
the mean-field model correctly predicts the magnitude of
the energy barrier and the observation that lower topological
0.50
0.25
0.00
−3
−2
−1
0
1
2
3
QS
Fig. 2 (a): The energy trace shows a universal behavior, as shown
by the collapse of numerical results(colored thick lines) onto one
curve which is predicted by the mean-field model. (b) Probability
density on a semi-log plot illustrates the exponential distribution of
energy barriers. The dashed line is an exponential fit with a slope of
−1.18. (c)The dependence of barrier heights on the contact
topology of the underlaying cells. A histogram (p(QS , ∆u)) of
energy barrier heights is shown at each value of the pre-T1
topological measure QS . Higher values of QS correspond to S1, S2
pairs with fewer neighbors. The average values are represented by
the black curve. p(QS , ∆u) exhibits exponential tail for the range of
QS shown here. The black solid line is the average value of ∆u and
the red dotted line is the meanfield theoretical prediction with no
fitting parameters. The overall distribution P(∆u) (black histogram
on right of figure) is obtained by convolving p(QS , ∆u) with the
distribution of topological measures f (QS ) (red histogram on top).
measures have higher energy barriers, although it does not fit
the shape of the simulation curve. This suggests the shape of
this curve is due to nontrival local correlations between cell
shapes.
To go from energy barrier distributions to cell migration
rates, we explore two of the simplest models to demonstrate
that the observed energy barrier distribution generically yields
glassy behavior, as measured by the time one has to wait to see
a cell change its neighbors. In confluent tissues, cell migration
rates are then proportional to neighbor exchange rates.
In traditional statistical mechanics, the rate at which a nearequilibrium system transitions from one metastable state to
another is described by an Arrhenius process,
R = ω0 e−∆uAB /ε ,
(4)
1–6 | 3
1010
0.75
!0 ⌧
Ctrap (0, t)
1014
decreasing "
106
0.50
102
0.25
0.00
10
2
102
3
10
106
!0 t
1010
10
1014
2
"/"0 = 1.1
4
2
0
2
6
4 2 0 2
log[b/(!0 "0 )]
0.9 1.00 1.10 1.20 1.30 1.40 1.50
1.25
(c)
(b)
Increasing b
(a)
1.00
log(!0 ⌧ )
where ∆uAB is the energy barrier separating two metastable
states A and C (Fig. 1)), ω0 is an inherent escape attempt frequency and ε = kB T is the scale of energy fluctuations.
While the assumptions on which Eq. (4) is based do not
necessarily hold in biological tissues, analogues to parameters ω0 , ∆uAB and ε exist in cells and likely govern cell motility. Several successful tissue models have characterized the
cell activity using an effective temperature ε estimated from
membrane ruffling 29 . Both ε and the rate at which cells attempt to cross barriers ω0 are correlated with cell protrusivity
and active shape fluctuations, which are determined in large
part by the cell’s individual biochemical makeup. For simplicity, we assume that ω0 and ε are single-cell properties that
are constant throughout the tissue, although other choices are
possible and would be interesting directions of future study. In
contrast, the distribution of energy barriers, ρ(∆u), is clearly
a collective property determined by cell-cell interactions and
the geometry of cell packing inside the tissue, as described in
the previous section.
1.00
"/"0
(d)
2
decreasing
↵2
h r2 i
10
1
0.75
"/"0
0.50
10
0
0.25
decreasing "/"0
10
10
1
!0 t
10
2
3
10
0.00 0
10
10
1
2
!0 t
10
3
10
Fig. 3 (a) Two-time correlation functions for
ε/ε0 = [2.00, 1.10, 1.32, 1.06, 1.02] in the trap model. As ε → ε0 ,
the correlations persist for increasingly long times, leading to glassy
behavior. (b) Colored lines are the caging time τ in the SGR model.
In the limit b/(ω0 ε0 ) → 0, the SGR model becomes the trap model
(thick black line). Inset: τ as a function of b/(ω0 ε0 ) at ε/ε0 = 1.1
(black dashed line in the main figure). (c) Mean squared
displacement for a random walk where the step sizes are determined
by the two-time correlation function Ctrap (0,t). Here we have used
b/(ω0 ε0 ) = 0.01 and ε/ε0 values ranging from 1.001 to 1.01. The
solid red line indicates slope 1. (d) Non-Gaussian parameter α2
(described in text) for random walk tracks shown in (c). α2 first rises
to a peak that coincides with the caging time τ(ε, b) and decays to
0 as the system becomes diffusive. α2 = 0 means diffusive behavior.
We first use a simple ‘trap’ model for glasses 18 to predict
waiting times for cell migration. In the trap model, a competition between ρ(∆u) and the Arrhenius rate (Eq. (4)) that sam4|
1–6
ples this distribution 18 determines the dynamics. For tissues
where ρ(∆u) has an exponential tail (Eq. (3)), the distribution
of the average time τ̃ spent in a metastable state is given by 18 :
f (τ̃) ∝ τ̃ −ε/ε0 ,
(5)
where τ̃ = R−1 is the inverse of the Arrhenius rate (Eq. (4)).
When ε < ε0 , Eq. (5) cannot be normalized, this means the
system cannot relax to an equilibrium state, resulting in solidlike glassy behavior.
For ε > ε0 , one can calculate the two-time correlation function Ctrap (0,t), which is the probability for a cell to rearrange
after spending time t in a state. In Fig. 3(a), Ctrap (0,t) exhibits
glassy or caging behavior at short times, but decays to zero at
longer times, indicating fluid-like behavior. The time scale of
this relaxation behavior depends on ε. We can define a caging
time as the value of τ such that Ctrap (0, τ) = e−1 . As a ε → ε0 ,
the system approaches a glass transition and τ(ε) diverges, as
shown by the black solid line in Fig. 3(b).
We next augment this simple model to account for an additional feature of single-cell motility: single cells on substrates tend to move along the same direction for long periods
of time due to polarization of the mechanical components that
generate traction forces 30,31 . This directed motion has been
shown to be important in other models for embryonic tissues 11
and occurs in addition to the random fluctuations induced by
changes to the cell shape that are modeled by ε. Therefore we
include directed cell motion in an SGR-like framework 19 .
We use the energy barrier height ∆u to label the state of a
T1 four-cell region (see Fig. S2). We model self-propelled,
directed motion by assuming the cell by assuming that the cell
actively increases the system’s potential energy at a constant
rate b. At time t, then the effective barrier height ∆u − bt.
There is also a finite probability for it to undergo a rearrangement due to non-directed fluctuations in its shape; we describe
this as an activated process controlled by a temperature-like
parameter ε 29 . Then the rate for overcoming a barrier at time
t can be written as:
R = ω0 e−(∆u−bt)/ε .
(6)
After escaping a trap with the rate given in Eq. (6), the T1 fourcell region enters into a new trap chosen from the distribution
ρ(∆u) as given by Eq. 3.
Simple extensions of the SGR analysis 19 can be used to
derive Ctrap (0,t), which is again the probability for a cell to
rearrange after spending time t in a state. Similar to the trap
model, a caging time τ can be defined. As shown by the colored lines in Fig. 3(b) adding a polarization energy b decreases
the caging time; in the limit of b → 0, the SGR model becomes
the trap model (a full contour plot of τ(ε, b) is also shown in
Fig. S3. In Fig. 3(b)(inset), we show that as a function of increasing b and constant ε, the caging time has a power-law
decay.
One possible way of implementing the trap model and comparing to direct experimental results of cell motility is to carry
out a random walk using the the two-time correlation function
Ctrap (0,t). First, at each time step, the state of a cell is determined by drawing a random state according to Ctrap (0,t):
it is either caged with probability Ctrap (0,t) and takes a small
step chosen from a χ 2 distribution or it migrates with probability 1 − Ctrap (0,t) and takes a larger step chosen from a
Gaussian distribution. In Fig. 3(c) we show the mean squared
displacements of these random walk tracks near the glass transition. Cells are caged at small time scales and diffusive behavior dominates at longer times; the transition between the
two regimes occurs at the time τ(b, ε) (Fig. 3(b)). To better
demonstrate cage breaking, we also analyzed the non-gaussian
parameter α2 11 for these random walks as shown in Fig. 3(d).
The peaks in α2 also coincide with the average time of cage
breaking events, directly set by τ(b, ε). As the glass transition is approached at ε → ε0 , the peak shifts further to larger
times, demonstrating a slowing down of dynamics in the system. Similar mean-squared displacements and non-gaussian
parameters have been seen in three-dimensional zebrafish embryos 11 and 2D epithelial sheets 32 , suggesting that our simple
model can explain those glassy features.
Both the trap and SGR-like models suggest that the energy barrier distribution we found in our simulations can lead
glassy cell dynamics, and that waiting times for cell migration
increase as the average barrier height (parameterized by ε0 )
decreases.
Discussion and Conclusion We have simulated confluent
tissue monolayers and numerically calculated the energy barriers required for cell rearrangements. We show that the distribution of energy barriers, ρ(∆u), is exponential and that ∆u
depends on a cell’s number of neighbors in a monolayer tissue.
Building on these results, we show that two minimal models 19
predict glassy dynamics, as measured by temporal correlation
functions and waiting times, and a simple random walk based
on these statistics reproduces features seen in experiments on
confluent tissues.
It should be possible to test these predictions in experiments
on confluent monolayers. Both the models predict that cell migration rates increase as the energy barriers decrease. Therefore, Fig. 2(c) predicts that cells are more likely to change
neighbors if they are in regions with high topological measure
(lower number of excess neighbors for S1 and S2). Although
it is difficult to track cell membranes in confluent tissues, one
could estimate cell topologies by taking a voronoi tessellation
of nuclei positions, and directly test this prediction.
Furthermore, both models make predictions about two-time
correlation functions, which could be studied experimentally
by looking at the decay in the overlap between a cell’s initial
and current voronoi areas as a function of time 33 . One could
decrease cell activity by adding drugs such as blebbistatin, and
compare directly to Eq. (6). In addition, there is a large-scale
cutoff for the exponential tail in our simulations which correlates with the largest edge length in the tissue. This suggests
that in real tissues we should always expect expect the twotime correlation function to decay to zero provided one waits
long enough.
Here we only model the simplest transition path leading to
a T1 transition by shortening (and subsequently growing) the
edges between cells. Realistically, the transition path can be
more complicated. For example, protrusions can be made as
the cell establishes new integrin bonds with the substrate, developing more complicated patterns such as Rosettes 15 . We
have studied a few such pathways using Surface Evolver and
find that they generically cost more energy, though a more systematic study is needed. In addition, we could analyze experimental cell shapes during T1 events to determine which
transition pathways the cells actually take, and estimate the
transition barrier across those pathways in silica.
For simplicity, our models and simulations make several assumptions about cell activity and dissipation, which should
be checked and modified if necessary. For example, we assume that dissipative processes, such as the actin network being remodeled by myosin, are not strongly dependent on cell
shapes/geometry and therefore we neglect them in our energy
functional. This could be checked using two point microrheology, and the model could be modified accordingly. Similarly, we have assumed that the rate at which cells attempt
to cross energy barriers ω0 , is also not geometry dependent.
However, since mechanosensing machinery influence cell polarization 34 it is possible that local cell shapes systematically
affect attempt frequencies, and this would be an interesting
avenue of future research. Furthermore, our model postulates
that the single-cell mechanical parameters κ, p0 are independent of the activities b and ε, but that is an assumption that we
intend to relax and study.
Finally, in writing down trap and SGR models, we have
implicitly assumed that the dynamics of cell monolayers are
dominated by the potential energy landscape (like a supercooled liquid or glass), in contrast to a higher temperature
normal liquid where rearrangements can happen anywhere
and are not strongly constrained by the potential energy landscape. This assumption is justified by the observations of
caging behavior and dynamical heterogeneities, but also by
the microscopic observation that cell structures are close to
that predicted by Eq. 2 23 , and transition between these nearequilibrium states quickly compared to the waiting times they
spend in each state 11 . Quantifying these transition times in
experiments (in addition to the waiting times) would therefore
be very useful.
Acknowledgements M.L.M. acknowledges the support
from NSF CMMI-1334611 and the Dean of A&S and the
Chancellor’s Fund at Syracuse University. J.H.L. and J.M.S.
1–6 | 5
acknowledge the support from NSF-DMR-0645373. The authors acknowledge useful discussions with Rastko Sknepnek
and Shiladitya Banerjee. The authors would also like to thank
the anonymous reviewers for their valuable comments and
suggestions to improve the quality of the paper.
References
1 R. Farhadifar, J.-C. Rper, B. Aigouy, S. Eaton and F. Jlicher, Current
Biology, 2007, 17, 2095 – 2104.
2 E.-M. Schoetz, R. D. Burdine, F. Julicher, M. S. Steinberg, C.-P. Heisenberg and R. A. Foty, HFSP journal, 2008, Vol.2 (1), 1–56.
3 M. Poujade, E. Grasland-Mongrain, A. Hertzog, J. Jouanneau,
P. Chavrier, B. Ladoux, A. Buguin and P. Silberzan, Proceedings of the
National Academy of Sciences, 2007, 104, 15988–15993.
4 L. Schneider, M. Cammer, J. Lehman, S. K. Nielsen, C. F. Guerra, I. R.
Veland, C. Stock, E. K. Hoffmann, B. K. Yoder, A. Schwab, P. Satir and
S. T. Christensen, Cellular Physiology and Biochemistry, 2010, 25, 279–
292.
5 P. Friedl and D. Gilmour, Nat Rev Mol Cell Biol, 2009, 10, 445–457.
6 D. Irimia and M. Toner, Integr. Biol., 2009, 1, 506–512.
7 K. Keren, Z. Pincus, G. M. Allen, E. L. Barnhart, G. Marriott,
A. Mogilner and J. A. Theriot, Nature, 2008, 453, 475–480.
8 R. J. Petrie, N. Gavara, R. S. Chadwick and K. M. Yamada, The Journal
of Cell Biology, 2012, 197, 439–455.
9 T. E. Angelini, E. Hannezo, X. Trepat, M. Marquez, J. J. Fredberg and
D. A. Weitz, Proceedings of the National Academy of Sciences, 2011,
108, 4714–4719.
10 K. D. Nnetu, M. Knorr, J. Käs and M. Zink, New Journal of Physics,
2012, 14, 115012.
11 E.-M. Schötz, M. Lanio, J. A. Talbot and M. L. Manning, Journal of The
Royal Society Interface, 2013, 10, 11.
12 M. D. Ediger, Annual Review of Physical Chemistry, 2000, 51, 99–128.
13 M. L. Manning and A. J. Liu, Phys. Rev. Lett., 2011, 107, 108302.
14 S. A. Langer and A. J. Liu, EPL (Europhysics Letters), 2000, 49, 68.
15 C. Guillot and T. Lecuit, Science, 2013, 340, 1185–1189.
6|
1–6
16 S. Henkes, Y. Fily and M. C. Marchetti, Phys. Rev. E, 2011, 84, 040301.
17 L. Berthier, Nonequilibrium glassy dynamics of self-propelled hard disks,
2013.
18 C. Monthus and J.-P. Bouchaud, Journal of Physics A: Mathematical and
General, 1996, 29, 3847.
19 P. Sollich, Phys. Rev. E, 1998, 58, 738–759.
20 L. Hufnagel, A. A. Teleman, H. Rouault, S. M. Cohen and B. I. Shraiman,
Proceedings of the National Academy of Sciences, 2007, 104, 3835–
3840.
21 S. Hilgenfeldt, S. Erisken and R. W. Carthew, Proceedings of the National
Academy of Sciences, 2008, 105, 907–911.
22 M. L. Manning, R. A. Foty, M. S. Steinberg and E.-M. Schoetz,
Proceedings of the National Academy of Sciences, 2010, 107, 12517–
12522.
23 K. K. Chiou, L. Hufnagel and B. I. Shraiman, PLoS Comput Biol, 2012,
8, e1002512.
24 S. Torquato, Author and H. Haslach, Jr, Applied Mechanics Reviews,
2002, 55, B62–B63.
25 K. A. Brakke, Experimental mathematics, 1992, 1, 141–165.
26 M. Basan, T. Risler, J.-F. Joanny, X. Sastre-Garau and J. Prost, HFSP J.,
2009, 3, 265.
27 D. Bi, J. H. Lopez, J. M. Schwarz and M. L. Manning, Unpublished, 2013.
28 D. B. Staple, R. Farhadifar, J. C. Roeper, B. Aigouy, S. Eaton and
F. Julicher, Eur. Phys. J. E, 2010, 33, 117–127.
29 J. Mombach, D. Robert, F. Graner, G. Gillet, G. Thomas, M. Idiart and
J. Rieu, Physica A: Statistical Mechanics and its Applications, 2005, 352,
525–534.
30 R. Baker, M. Brasch, M. L. Manning and J. H. Henderson, submitted,
2013.
31 V. Maruthamuthu, B. Sabass, U. Schwarz and M. Gardel, Proceedings of
the National Academy of Sciences, 2011, 108, 4708.
32 K. D. Nnetu, M. Knorr, S. Pawlizak, T. Fuhs and J. A. Kas, Soft Matter,
2013, 9, 9335–9341.
33 A. R. Abate and D. J. Durian, Phys. Rev. E, 2007, 76, 021306.
34 R. De, A. Zemel, S. A. Safran and G. V. Shivashankar, in Chapter 7 Theoretical Concepts and Models of Cellular Mechanosensing, Academic
Press, 2010, vol. Volume 98, pp. 143–175.
Supplementary Information for Energy barriers and cell migration in
densely packed tissues†
Dapeng Bi∗ , J. H. Lopez∗ , J. M. Schwarz∗ and M. Lisa Manning∗a
Received Xth XXXXXXXXXX 20XX, Accepted Xth XXXXXXXXX 20XX
First published on the web Xth XXXXXXXXXX 200X
DOI: 10.1039/b000000x
1
Aboov-Weaire law
In a cellular packing where all vertices are three-edge coordinated, the coordination number of nearest adjacent neighbors
are correlated and can be described by the empirical AboavWeaire law 1 :
zn.n. (z) ≈ A + B/z,
(1)
where z is the coordination number of a given cell and zn.n.
is the coordination number of its neighbor. The ... notation
means averaging over all cells in the packing. As shown in
Fig. 1, this also holds for the packings used in this work.
Zn.n. 8
2
Trap Model for glases
The trap model for glasses 2 models a system with a complex
potential energy landscape, that is captured by a quenched distribution ρ(Δu). The simple assumption is that the evolution
between different metastable states occur via activated processes. The probability for a state to be in trap labeled by Δu
at time t, P(Δu,t), evolves according to the master equation 2 :
d
P(Δu,t) = −ω0 e−Δu/ε P(Δu,t) + ω(t)ρ(Δu).
(2)
dt
In Eq. (2), the first term on the r.h.s. is the rate of escape
from trap labeled by Δu. This is modeled as an activated process where the fluctuation is given by ε. In thermal systems
ε = kB T . ω0 is an inherent attempt frequency which is assumed to be independent of ε. The second term on the r.h.s.
of Eq. (2) is the rate of entering a trap with Δu, where a time
dependent attempt frequency is given by
7
6
5
3
Therefore, for the four cells involved in a T1 transition labeled S1,S2,E1 and E2, their topologies are not independent
from each other. Statistically, given ZS1 and ZS2 it is possible
to determine ZE1 and ZE2 .
4
5
6
7
Z
8
9
ω(t) = ω0
10
Fig. 1 The topological correlations from cell packings used in our
simulations are represented by the blue squares. Z is the number of
neighbors for a cell, and Zn,n is the average number of neighbors
of the nearest neighbors. The solid blue line is fitting using
Aboav-weaire law, with A and B as fitting parameters
† Electronic Supplementary Information (ESI) available: [details of any
supplementary information available should be included here]. See DOI:
10.1039/b000000x/
∗ Department of Physics, Syracuse University, Syracuse, NY 13244, USA Fax:
+1 315 443 9103; Tel: +1 315 443 3920; E-mail: [email protected]
b Syracuse Biomaterials Institute, Syracuse, NY 13244, USA
∞
0
dΔuP(Δu,t) e−Δu/ε .
(3)
The rate of entering includes ρ(Δu) which is the underlaying
distribution of the trap depth Δu in the potential energy landscape.
Here we chose the form of ρ(Δu) from the simulation result
of the main text
1
(4)
ρ(Δu) = e−Δu/ε0 .
ε0
2.1
Steady state solution
For this work we are interested in the steady state solution to
Eq. (2) which is given by
ε − ε0 −Δu(1−1/x)
e
(5)
fss (Δu) =
ε ε0
1–3 | 1
where x = ε/ε0 . For x < 1 the system is glassy, since Eq. (5) is
not normalizable. For x > 1 which is normalizable only when
x ≥ 1. The steady state distribution (Eq. 5 in main text) can
also be written in terms of the trapping time τ̃, which is the
inverse of the escape rate τ̃ = eΔu/x
gss (τ̃) = fss (Δu)
2.2
dΔu
∝ τ −x .
d τ̃
(6)
Two-time correlation function
We also study dynamics of the trap model in terms of two-time
correlation for a particle starting in one trap at time t = 0 and
remaining in the same at t. The probability that a state is in
a trap labelled by Δu at t = 0 will stay in the same state Δu
at time t is given by the homogeneous part of the solution of
Eq. (2):
stay
(t) = fss (Δu) exp −ω0 e−Δu/xt .
(7)
FΔu
A two-time correlation function can be defined by including
the contribution from all initial traps Δu,
ctrap (0,t) =
∞
0
stay
dΔuFΔu
(t) =
x−1
(ω0t)x−1
ω0 t
0
dw wx−2 e−w .
(8)
A plot of ctrap (0,t) for different values of x = ε/ε0 is shown
in Fig. 3(a) of the main text.
3
closer to a T1 rearrangement, and making the effective barrier
height Δu − bt. There is also a finite probability for it to undergo a rearrangement due to non-directed fluctuations in its
shape; we describe this as an activated process controlled by a
temperature-like parameter ε. Then the rate for overcoming a
barrier at time t can be written as
Soft Glassy Rheology model for tissues
R = ω0 exp [−(Δu − bt)/ε] .
After escaping a trap with the rate given in Eq. (9), the T1
four-cell region enters into a new trap chosen from the distribution ρ(Δu). A master equation can be written 3 to describe
the evolution of the four-cell T1 location:
∂
P(Δu,t) = − ω0 e−[Δu−bt]/ε P(Δu,t)
∂t
+ρ(Δu)
dΔu ω0 e−[Δu −bt]/ε P(Δu ,t),
(10)
where P(Δu,t) is the probability that T1 location is in trap
labeled by Δu at time t. On the r.h.s., the first term describes
the rate of hopping out of a trap and second term is the total
rate of entering trap with Δu, which is proportional to ρ(Δu),
the distribution of choosing a new trap.
Equation 10 can be rewritten in terms of nondimensionalized time s = ω0t and energy u = Δu/ε0
∂
f (u, s) =
∂s
− e−(u−Bs)/x f (u, s) + e−u
du e−(u −Bs)/x f (u , s)
(11)
where we have also introduced the dimensionless parameters
B = b/(ω0 ε0 ) and x = ε/ε0 .
The steady-state solution of Eq. 11 is the same as the result
for the trap model Eq. (5):
Potential
Energy
fss (u) = (1 − 1/x)e−u(1−1/x) ,
bt
Δu
Reaction
Coordinate
Fig. 2 A schematic of the model for glassy dynamics in tissues. Δu
labels the trap (represented by the black dot with arrow). bt is the
amount of energy generated as the cell moves in a directed manner.
Following the ideas of Soft Glassy Rheology 3 , we model
self-propelled, directed motion in a simple way, by assuming the cell generates energy at a constant rate b that allows
it to traverse the Potential Energy Landscape. At time t, directed cell motion has consumed energy ∼ bt, bringing it
2|
(9)
1–3
(12)
which is not normalizable for 1 < 1/x or ε > ε0 . This indicates
a glass transition at ε = ε0 .
The probability that a T1 location with u at s = 0 will stay
in the same state u at time s is given by the homogeneous part
of the solution of Eq. (11):
x
(13)
Fustay (s) = fss (u) exp − e−u/x (eBs/x − 1) .
B
A two-time correlation function can be defined by including
the contribution from all traps depths u,
csgr (0, s) =
∞
0
du Fustay (t) =
x−1
y(s)x−1
y(s)
0
dw wx−2 e−w ,
(14)
where y(s) = Bx eBs/x − 1 . We can define a caging time as
the value of τ such that Csgr (0, τ) = e−1 . The dependence of
τ on x = ε/ε0 and B = b/(ω0 ε0 ) is plotted in Fig. 3.
ω0 τ
100
1020
1015
ω0 τ = ∞
b/(ω0 0 )
102
1010
10−2
−4
10
−6
10
105
100
1
1.05
/0
1.10
1.15
Fig. 3 Eq. (14) is used to define a caging time τ, such that
c(0, τ) = e−1 , as a function of the two dimensionless parameters of
the model.
References
1 D. L. Weaire and S. Hutzler, The physics of foams, Oxford University
Press, 1999.
2 C. Monthus and J.-P. Bouchaud, Journal of Physics A: Mathematical and
General, 1996, 29, 3847.
3 P. Sollich, Phys. Rev. E, 1998, 58, 738–759.
1–3 | 3