ArticlePDF Available

Active viscoelastic models for cell and tissue mechanics

The Royal Society
Royal Society Open Science
Authors:

Abstract and Figures

Living cells are out of equilibrium active materials. Cell-generated forces are transmitted across the cytoskeleton network and to the extracellular environment. These active force interactions shape cellular mechanical behaviour, trigger mechano-sensing, regulate cell adaptation to the microenvironment and can affect disease outcomes. In recent years, the mechanobiology community has witnessed the emergence of many experimental and theoretical approaches to study cells as mechanically active materials. In this review, we highlight recent advancements in incorporating active characteristics of cellular behaviour at different length scales into classic viscoelastic models by either adding an active tension-generating element or adjusting the resting length of an elastic element in the model. Summarizing the two groups of approaches, we will review the formulation and application of these models to understand cellular adaptation mechanisms in response to various types of mechanical stimuli, such as the effect of extracellular matrix properties and external loadings or deformations.
Models of active contractility relating force/tension generation and the rate of contraction. (a) Model proposed by Hill [29] to describe muscle contraction [86]. CE serves as a force-generating contractile element and SE is an elastic spring that affects the dynamics of contraction. (i) Force–velocity curve obtained in Hill’s experiments. (ii) Normalized speed of shortening ( v = V / V m a x ) and normalized force generation ( f = F / F m a x ) in single cells in response to various substrate stiffnesses can be represented by the reduced Hill equation (3.2) [37]. (b) Active fluid model proposed to study mechanotransduction in cells [38]. (i) This model accounts for the transiently cross-linked nature of the actomyosin network and force generation via myosin motor activity. (ii) Equivalent of the model presented in (i), which has the same constitutive equation. The net effect of two dashpots is similar to one dashpot. (iii) Evolution of traction force generation over time for low (grey) and high (black) substrate stiffness. The two main characteristics of force generation–time curves, that is, the plateau force F p and the rate of force generation ( d F / d t ) are dependent on the substrate stiffness [37]. (iv) The plateau force has a linear relationship with substrate stiffnesses when the stiffness is below 60 nN μm⁻¹ and for higher values of stiffness, F p saturates [37]. (c) Active Maxwell model that assumes a time-dependent increase in cellular contractile force with the same profile as actin and myosin accumulation in the RhoA-activated areas [87]. (i) Changes in the local intensity of actin and myosin in the regions of RhoA activation. (ii) The plateauing exponential function used in the model to represent the stress profile during activation of RhoA. (iii) Experimental (black) and theoretical (red) evolution of strain energy during intervals of RhoA activation and relaxation.
… 
Active models for solid-like materials. (a). Active rheological model that includes a constant CE, σ a , in parallel to a standard solid model to account for the prestress in cell monolayers [18]. Temporal evolution of strain (i) and tissue stress (ii) as a function of device strain presented for epithelial monolayers undergoing compression at a low strain rate (0.5% s − 1 ). (b) The active element proposed by Muñoz et al., which adapts its resting length in response to deformations [108]. The evolution of stress over time in stress relaxation tests predicted by the Maxwell model and active models with two definitions of elastic strain is presented in a table. Stress in the models is defined using the elastic strain, that is, σ ( t ) = E ε e . Changing the definition of elastic strain can result in an equivalent evolution of stress in the active element and the Maxwell model for γ ≈ τ − 1 . When the strain is very small, the difference between the evolution of stresses in all three models will be negligible. However, at larger strains, the stress in the system would be dependent on the way that the strain is defined. (c) The modified version of the active element proposed by Muñoz et al., which also considers the effect of cellular contractility [109]. (i) The average curves representing the evolution of stress over time for MDCK monolayers stretched at 75% s − 1 to various amplitudes of strain plotted on a semi-log scale. (ii) Demonstration of the correlation between the τ model calculated from the active rheological model and the characteristic time τ calculated from fitting the stress–time curves with an empirical function, that is, σ = A e − t / τ t − α + B , for different loading conditions and actomyosin treatments. (d) An example of a hybrid vertex model employed to study wound healing. The top view of the epithelial tissue with ablated cells is shown along with the rheological models of nodal segments (cell-centre connections) in black and vertex segments (cell boundaries) in red [110].
… 
This content is subject to copyright.
Active viscoelastic models for
cell and tissuemechanics
Bahareh Tajvidi Safa1, Changjin Huang2, Alexandre
Kabla3 and Ruiguo Yang1,4,5
1Department of Mechanical and Materials Engineering, University of Nebraska-Lincoln,
Lincoln, NE 68588, USA
2School of Mechanical & Aerospace Engineering, Nanyang Technological University, Singapore
639798, Singapore
3Department of Engineering, University of Cambridge, Cambridge CB2 1PZ, UK
4Department of Biomedical Engineering, and 5Institute for Quantitative Health Science and
Engineering (IQ), Michigan State University, East Lansing, MI 48824, USA
BTS,0000-0003-1280-5319; AK,0000-0002-0280-3531;
RY,0000-0002-1361-4277
Living cells are out of equilibrium active materials. Cell-
generated forces are transmied across the cytoskeleton
network and to the extracellular environment. These active
force interactions shape cellular mechanical behaviour,
trigger mechano-sensing, regulate cell adaptation to the
microenvironment and can aect disease outcomes. In recent
years, the mechanobiology community has witnessed the
emergence of many experimental and theoretical approaches
to study cells as mechanically active materials. In this review,
we highlight recent advancements in incorporating active
characteristics of cellular behaviour at dierent length scales
into classic viscoelastic models by either adding an active
tension-generating element or adjusting the resting length of
an elastic element in the model. Summarizing the two groups
of approaches, we will review the formulation and application
of these models to understand cellular adaptation mechanisms
in response to various types of mechanical stimuli, such as the
eect of extracellular matrix properties and external loadings
or deformations.
1. Introduction
Living cells, their surrounding extracellular matrices (ECM) and
tissues as a whole exhibit viscoelastic properties, that is, having
both an elastic and a viscous behaviour. Mechanical tests are
used to characterize the mechanical properties of cells and tissues
and help us understand and predict cellular behaviour in healthy
or pathological conditions at dierent time scales [1–4]. For
example, experimental studies on cell monolayers devoid of
substrate have reported a viscoelastic solid-like behaviour under
© 2024 The Authors. Published by the Royal Society under the terms of the Creative
Commons Attribution License http://creativecommons.org/licenses/by/4.0/, which permits
unrestricted use, provided the original author and source are credited.
Review
Cite this article: Tajvidi Safa B, Huang C, Kabla
A, Yang R. 2024 Active viscoelastic models for cell
and tissue mechanics. R. Soc. Open Sci. 11:
231074.
https://doi.org/10.1098/rsos.231074
Received: 24 July 2023
Accepted: 25 February 2024
Subject Category:
Physics and biophysics
Subject Areas:
biomechanics, biomaterials
Keywords:
active model, viscoelasticity, cell modeling, cell
mechanics, tissue mechanics
Authors for correspondence:
Alexandre Kabla
e-mail: ajk61@cam.ac.uk
Ruiguo Yang
e-mail: ryang6@unl.edu
constant strains [5]. In other words, on the time scale of seconds, cell monolayers behave like a
viscous uid and dissipate stress, then reach a plateau in stress on the time scale of minutes, which
is a characteristic of elastic solid materials [5]. On the other hand, cell aggregates exhibit solid-like
behaviour on short time scales, for example, in a few seconds, and uid-like behaviour at time scales
of the order of minutes to hours [4,6,7]. In this case, the behaviour of cell aggregates is similar to
a viscoelastic uid material. Additionally, the mechanisms governing the viscous-like behaviour can
also dier depending on the time scale of load application [8]. For example, on the time scale of tens
of seconds to minutes, stress dissipation occurs owing to the turnover of actin laments and reorgan-
ization of the actomyosin network [4,8,9]. On the time scale of minutes to hours, stress dissipation
mechanisms at the cellular scale, such as oriented cell division and cell rearrangements, can start to
inuence the response to mechanical loadings [8,10–12]. Transitions between uid-like and solid-like
behaviours in living cells can also occur in response to mechanical stimuli. For instance, short-term
uidization has been reported immediately upon strain application, which is usually followed by
stiening [13–16].
The experimental results are often analysed by mathematical models to capture the important
features of the material response, presented in terms of model parameters. These parameters can
then be used for classication, comparison and prediction of the mechanical behaviour of cells and
tissues subjected to other loading conditions [17]. By relating the model parameters to the underlying
biological processes on the molecular and cellular scale, the physical meaning of the parameters can
sometimes be assigned. Since tissues exhibit time-dependent mechanical behaviours [5,18], a common
modelling approach is to consider the cells and tissues as a viscoelastic continuum and to describe their
mechanical response from quantitative mechanical interrogations, often stress–strain relationships, in
terms of a combination of stiness and viscosity, or elastic and loss moduli under dynamic loadings.
These extracted mechanical properties have long been regarded as disease biomarkers [19,20]. For
instance, metastatic cancerous cells exhibit lower stiness than benign cells [21].
Perhaps more importantly, cell mechanics is not only a by-product of the underlying molecular
structure but also a means for cells to actively adapt to environmental cues, in service to a preferred
cellular function, such as cell migration in wound healing [22] and tissue morphogenesis in develop-
ment [23]. Being viscoelastic in nature, cells can dissipate the imposed stress owing to external strains.
However, sometimes this passive response is insucient to maintain the mechanical integrity of the
cell. For instance, cells need to use eective stress relaxation mechanisms such as actin polymerization
to prevent tissue fracture. In other instances, under rapid strain applications, cells may need to stien
or actively pull back to make further deformation dicult, thus preventing further damage to their
cytoskeleton. From the adaptation perspective, the evolving elastic and viscous properties owing to
the active adjustments of cytoskeleton tension can be considered as a way to facilitate the response
and adaptation to external stimuli. Probing active cellular behaviour through the lens of mechanics
is particularly intriguing because it oers a window from which the adaptation can be quantitatively
examined up close with dened mechanical stimuli.
In the cell cortex as well as at the cell population scale, living maer not only responds to exter-
nal forces or deformation as any traditional material would but also often exhibits force-generating
mechanisms emerging from actin polymerization, adhesion dynamics and actomyosin contractility.
This process is referred to as being ‘active’ in this review. Active maer theory is another continuum-
based model used to describe the dynamics of cell cytoskeleton and cell monolayers (reviewed in
[9,24–26]). This type of model is based on the theory of liquid crystals and can be employed to describe
the mechanics of the actomyosin cortex for time scales longer than the turnover time of actin laments
[27]. However, these models rarely capture the way such materials respond to external deformations or
stresses; it remains challenging to capture, both experimentally and theoretically, the impact of active
processes on the mechanical state of living maer.
Experimentally, a common approach is to subject the cellular materials to tension or compression
and then assess their response using stress–strain curves. At the tissue scale, mechanical testing
machines can be used to investigate the macroscopic behaviour of tissue samples (gure 1a) [28,29].
A similar technique has been developed to study the response of cell monolayers when subjected to
compression or tension (gure 1b) [30]. In addition, to examine the microscale characteristics of living
tissues, atomic force microscopy (AFM) and nanoindentation techniques can be employed, where a
probe tip is used to apply forces in the pico-nano Newton range [31,32].
The AFM technique can also be used to study the active and passive behaviour of individual
cells [33–35] and can be adapted for cell pair studies in single-cell force spectroscopy (SCFS), where
a single cell adhered to the cantilever beam serves as the probe tip (gure 1c) [36]. Furthermore,
2
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
various methods are developed to study the response of single cells and cell doublets in tension and
compression tests. For instance, in microplate assays, cells are compressed or stretched between a xed
and a deformable plate [37,38], microbead assays use microbeads to apply force via magnetic [39,40]
or optical tweezers [41–43] (gure 1d), and the micropipee aspiration method involves subjecting
cells to negative pressures and measuring their deformation [44–46] (gure 1d). This method can
also be used for cell pair studies by bringing two cells into contact via micropipees, that is, dual
pipee aspiration (DPA) (gure 1c) [47], or it can be combined with other techniques such as optical
tweezers [48]. Recently, a new micromanipulator device has been introduced that can directly measure
forces in cell pairs under controlled loading conditions, leading to advancements in the precision of
the interrogation of cell pair mechanics (gure 1c) [49]. Dierent techniques have varied ranges of
resolution and loading rates; thus, one has to consider the application requirements when choosing an
experimental technique. For instance, AFM techniques benet from higher spatial resolution and force
sensitivity compared with methods employing micropipee aspiration [50,51].
These experimental techniques are employed to probe cellular response at various temporal and
spatial scales. In addition to measuring the global mechanical behaviour of cells and tissues, these
techniques can also be employed to examine the mechanical properties of specic components of
(a)
(c)
(d)
(b)Tissue
Cell pair
Single Cell
Cell layer
Figure 1. Experimental techniques used to probe cell mechanics in various spatial scales. (a) Mechanical testing machine used to
study the response of tissue samples in compression and tension tests. (b) The recent technique developed for investigating the
mechanical response of cell layers to uniaxial stretch and compression. (c) In cell pair studies, mechanical stimulus is applied through
various methods such as a single cell adhered to a tipless atomic force microscopy (AFM) cantilever in single-cell force spectroscopy
(SCFS), two micropipettes in dual pipette aspiration (DPA) and a microstructure fabricated using two-photon polymerization technique
in single-cell adhesion micro tensile tester (SCATT). (d) Microplate assay is a commonly used method to monitor the force generated
by individual cells when exposed to variations in substrate stiness. Other techniques, such as micropipette aspiration, AFM and optical
and magnetic tweezers, involve applying controlled displacement as the mechanical stimulus and monitoring the forces within cells or
investigating changes in displacement in response to controlled forces.
3
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
cells. For example, micropipee aspiration techniques have been used to investigate the microrheology
of the cell nuclei [52–54]. The monitored stress/strain response of the cellular materials will be the
outcome of both active force-generating and passive mechanisms operating at those scales. From a
modelling perspective, capturing active behaviour requires a mechanism of introducing change to the
otherwise xed viscoelastic models. Researchers over the years have incorporated active adaptation
mechanisms by introducing active empirical mathematical models to link mechanical characteristics to
the underlying biological processes without simulating the details of the underlying chemical signals.
Rheological models for describing the mechanical behaviour of cellularized materials across various
length scales have previously been reviewed [4,15,17,55]. In this article, our primary goal is to focus
on dierent forms of integrating activity into rheological models, emphasizing variations in their
denitions. We here summarized these reported mechanisms into two broad categories. The rst group
integrates a force/stress (force divided by cross-sectional area)-generating element to classic viscoelas-
tic models. This active element can be constant or time-dependent. The second group introduced
mathematical methods that adjust the resting length of an elastic spring or the reference stress-free
shape in two dimensions/three dimensions in classic viscoelastic models. Mathematical models in both
categories are developed in conjunction with specic techniques that probe the active cell response to
various types of mechanical stimuli. In some cases, the active models can be mathematically equiva-
lent, but dierent parameters may oer a beer link with the underlying biology. In this article, we
present a selection of such models, introducing rst the experimental ndings and then focusing on
the formulation and application of each type of active viscoelastic model. We will begin with a short
introduction to the biology of active cell behaviour.
2. Biology of cellular active behaviour
Cell cytoskeleton mainly comprises lamentous proteins that preserve the cell structure, arrange
organelles, and resist, transmit and generate forces [56,57]. These proteins can be categorized into
three groups: microtubules, intermediate laments and actin laments [58]. Active force generation
is realized by actin laments. They are constructed by assembling monomeric actin. Actin-binding
proteins bind to actin laments and form dierent structures, such as the lamellipodium network,
contractile bundles of stress bres and the contractile network of the cell cortex [57,59]. Actin la-
ments are engaged in active processes such as actin treadmilling and force generation by consuming
the energy provided by adenosine triphosphate (ATP) hydrolysis [60–62]. ATP molecules aach to
ATP binding sites on actin monomers, and the ATP-bound actin monomers will be assembled at
the plus/barbed end of the actin laments leading to lament growth [60,63]. ATP molecules will
slowly hydrolyse to adenosine diphosphate (ADP) and the ADP-bound actin monomers will start to
disassemble from the minus/pointed end of the actin laments [60,64]. The process of assembling and
disassembling actin monomers is referred to as actin treadmilling [65]. Actin laments use the energy
from ATP hydrolysis to generate protrusion forces to help cells in spreading and migration [66,67].
Actin treadmilling is also crucial in endocytosis, exocytosis and phagocytosis to engulf large particles
[57].
In addition, ATP hydrolysis provides the energy for contractile stress generation in the actomyosin
network. Myosin motors convert chemical energy from ATP hydrolysis to mechanical energy and slide
actin laments past one another to produce force [68–70]. This process is similar to the shortening of
sarcomeres in muscle cells [37,38,64,70]. The contractile forces generated from this process will then be
transmied to neighbouring cells and the ECM via cell–cell junctions and focal adhesions, respectively
[66,71,72]. At the cell and tissue level, contractile force production controls cellular activities, such as
cell migration [67,73–75], proliferation [76,77], stem cell lineage determination [78,79], tissue regener-
ation [80] and morphogenesis [81,82]. Important for cellular mechanical characterization, contractile
forces also regulate cellular response to substrate stiness and mechanosensing [83]. The active
ATP-dependent processes that are at play in subcellular scales can also regulate cellular response
under various loading conditions at cell and tissue scales [5,59,84,85].
3. Active contractile elements
Rheological models are useful tools to quantitatively analyse the results of mechanical tests. However,
associating the molecular scale origin with the model parameters is challenging, in particular, where
4
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
cell behaviour deviates from the response of traditional passive materials [17]. Thus, adding elements
that represent the active behaviour of cells can equip classic models with the ability to empirically link
the mechanical response to the underlying biological processes.
To represent active cellular dynamics from intracellular contractility, contractile/force-generating
elements (CEs) can be added to standard viscoelastic models. In this section, we will discuss
active CEs that can mimic cell behaviour when they are exposed to dierent types of mechanical
stimuli. We begin with the rst active model, which was presented by Hill to elucidate muscle
contraction and its application in studying mechanosensing in single cells. In the following section,
we will explore an active uid model that considers the transient nature of the actin network to
describe characteristics of the single-cell response to the substrate stiness. In this model frame-
work, tension generation via myosin activity will be denoted by , and regulating mechanisms
of the dynamics of tension build-up will be discussed. Additionally, we will also provide some
examples for both models.
3.1. Muscle contraction dened by the Hill model
Muscles function as intricate biochemical mechanisms that transform chemical energy into mechanical
energy through actomyosin interactions to provide movement in our bodies. Hill proposed the rst
theory to describe muscle contraction back in 1938 [29]. He conducted experiments using the sartorius
muscle of a frog and explained the macro properties of muscles. In these experiments, Hill quantied
both force generation and velocity of length changes in muscles under various loading conditions. The
two endpoints of the curve representing the force–velocity relationship in muscles (gure 2a(i)) were
generated in two extreme cases: zero force (isotonic condition) and zero velocity (isometric condition).
Specically, when force is kept at 0, the muscle can reach its maximum shortening velocity, max.
Similarly, when the muscle is restricted from changing its length, the maximum level of force, that is,
the stall force max, can be generated at the steady-state condition.
Hill proposed an empirical function to describe the correlation between the active force generation,
, and the shortening velocity, , as follows [29]:
(3.1)(+)(+) = (max +),
where a, b and max += are constants specic to each muscle type. The Hill model was rst
discovered on frog skeletal muscles. However, later studies have shown that the dimensionless form of
this model with a shape factor =/max =/max 0.25 can be used to describe the behaviour of other
muscle types,
(3.2)(+)(+) = (1 + ),
where =
max and =
max [88]. Hill also proposed a phenomenological model to describe muscle
mechanics. The original model consists of a contractile element (CE) that generates force and an elastic
element (SE), as shown in gure 2a. CE is governed by equation (3.2) and SE aects the length and rate
of change in the length of CE during contractions [29,86]. Other forms of this model have also been
introduced by integrating more elastic and viscous elements to account for the viscoelastic properties
of the muscle and its interactions with the connective tissue surrounding muscle bres [89]. The Hill
model has enabled researchers to explore the mechanics of muscles using only a few rheological
parameters. However, the Hill model falls short of elucidating the underlying biological mechanisms
of force generation in muscles [89–91]. Another shortcoming of this model is that it fails to consider
variations in the contractile characteristics of various bre types within muscles and the dependence of
muscle tension on the movement history [89,91]. Modications aimed at enhancing the accuracy of the
Hill-type model predictions are reviewed in [91].
This model is extremely versatile and not only can it be used to describe the behaviour of various
types of muscles, but also it can be adapted and employed to describe force generation in non-muscle
cells and their response to the physical and mechanical properties of their microenvironment [92–94].
Mitrossilis et al. [37] studied the response of a single C2.7 myoblast cell and a 3T3 broblast mounted
between two microplates and showed a time-dependent force generation in single cells in response
to the microplate stiness (gure 2b(iii)). The force–velocity curves for single cells have a similar
shape as the force–velocity curves reported for muscle bres. Moreover, they showed that the reduced
5
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
Hill equation (3.2), with the same shape factor = 0.25 as for muscles, can describe the normalized
shortening speed, =/max versus the normalized force, =/max, as shown in gure 2a(ii). In
these experiments, the maximum force and velocity are measured at innite microplate stiness and
very low stinesses, respectively. Therefore, showing that the coupling between force generation in
cells in contact with their surroundings still follows Hill’s model is quite remarkable, as it implies that
structured muscle actomyosin and cytoskeletal actomyosin exhibit similar behaviours independent of
the network architecture.
In addition, since the same function could explain the experimental results of dierent muscle
types, Huxley suggested that the force–velocity relationship is generic in muscles and proposed a
molecular explanation [95]. This molecular model incorporates the dynamics of the interaction between
actin laments and myosin motors, that is, the number of myosin heads connected to actin laments
and the formation of temporary connections between actin and myosin heads (i.e. cross bridges).
Understanding these dynamics brings valuable insight into the biological processes regulating the
predictions of the Hill model. For instance, cross-bridge-type models have shown that the maximum
Schematic Diagram
of the Model
(a)
(b)
(i)
(i)
(iii)
(ii)
(ii) (iii)
(i) (ii)
(iv)
(c)
Constitutive Equations
The Hill model
Contractile element including the dynamics of cytoskeleton turnover
Time-dependent contractile element
Experimental Results
Figure 2. Models of active contractility relating force/tension generation and the rate of contraction. (a) Model proposed by Hill [29]
to describe muscle contraction [86]. CE serves as a force-generating contractile element and SE is an elastic spring that aects the
dynamics of contraction. (i) Force–velocity curve obtained in Hill’s experiments. (ii) Normalized speed of shortening =/max
and normalized force generation =/max in single cells in response to various substrate stinesses can be represented by the
reduced Hill equation (3.2) [37]. (b) Active uid model proposed to study mechanotransduction in cells [38]. (i) This model accounts
for the transiently cross-linked nature of the actomyosin network and force generation via myosin motor activity. (ii) Equivalent of
the model presented in (i), which has the same constitutive equation. The net eect of two dashpots is similar to one dashpot. (iii)
Evolution of traction force generation over time for low (grey) and high (black) substrate stiness. The two main characteristics of force
generation–time curves, that is, the plateau force and the rate of force generation / are dependent on the substrate stiness
[37]. (iv) The plateau force has a linear relationship with substrate stinesses when the stiness is below 60 nN μm−1 and for higher
values of stiness, saturates [37]. (c) Active Maxwell model that assumes a time-dependent increase in cellular contractile force
with the same prole as actin and myosin accumulation in the RhoA-activated areas [87]. (i) Changes in the local intensity of actin
and myosin in the regions of RhoA activation. (ii) The plateauing exponential function used in the model to represent the stress prole
during activation of RhoA. (iii) Experimental (black) and theoretical (red) evolution of strain energy during intervals of RhoA activation
and relaxation.
6
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
speed of muscle shortening happens owing to the rate of myosin aachment and detachment [89,95]
rather than the extent of lament overlap [96,97]. This model has its limitations as well. For example,
this model does not include the eect of power stroke (i.e. a crucial step in the force generation cycle in
muscles where myosin heads pull the actin laments and generate force) [89,91].
Further aempts have been made to incorporate various aspects of molecular biology with the Hill
model. An example involves a model that integrated the dynamics of actin lament slippage during
the process of force generation, allowing the prediction of the relationship between force generation
and stiness of micropillars [98]. This model was further rened to account for the temporal evolution
of the force over time by integrating an internal variable, specically representing the progression of
myosin motor stalling over time [99]. Despite these improvements, this model falls short in explaining
the limiting factors for the maximum force generation and maximum speed of shortening of the cells.
A model that addresses these shortcomings is discussed in the next section.
3.2. Active element that includes the dynamic turnover of the actomyosin network
Étienne et al. developed an active uid model to link the fundamental features of cellular molecular
mechanics with a passive phenomenological model [38]. These features encompass the transient nature
of the actomyosin network, force generation via myosin motor activity and actin polymerization. The
transiently cross-linked actomyosin network is modelled as a Maxwell uid in series with an active
element , mimicking force generation via myosin motors (gure 2b). The constitutive equation of the
network is shown as follows:
(3.3)
˙+2
˙=.
Here, is the characteristic time scale of cross-linker unbinding (i.e. elastic-like in short time scales
(<) and viscous over longer periods of time (>)), and is the elastic modulus of the cell.
represents the maximum value of contractility or stall force that could be generated in cells,
determined by two factors. First, it is inuenced by the rate at which myosin motors can generate
stress (1/) and contract the cell with an elastic modulus of . Second, it is aected by the rate of
cross-linker unbinding (1/), which counteracts the increase of stress in the system. Consequently,
is proportional to (/). This model eectively captures the evolution of force generation
in cells leading to the establishment of tension in the steady state. Using a single dashpot (gure
2b(i)), instead of the proposed two (gure 2b(ii)), results in a similar constitutive equation. However,
using two dashpots highlights the loss of force generation in the steady-state condition when the net
displacement of the microplates is 0, indicating internal creep.
This model can predict the critical stiness over which the plateau force () remains constant,
the response of cells to step changes in substrate stiness, and the rate at which force is generated
across various substrate stinesses. Interestingly, the constitutive equation (3.3) can also be wrien in a
similar form as the Hill model of muscle contraction,
(3.4)
+ +=,
where =,= 2+, and =+ (
˙)/(2) . Here, =
˙ is the speed of shortening,
represents the rate of actin polymerization and =/2 is the internal creep. This equation allows the
examination of non-muscle cell behaviour in two extreme cases: = 0 (stiness of the substrate equal
to 0) and = 0 (stiness of the substrate equal to innity). These extreme scenarios illustrate the role
of the molecular mechanisms that both govern and limit cell responses. Basically, cells initiate force
generation upon aachment to the substrate. If the resistance of the external environment is lower
than the force generated via myosin motors (e.g. when = 0), cells will start to contract the microplate,
which would, in turn, increase its resistance against the cell. This increased external resistance leads to
a reduction in the rate of retrograde ow. Moreover, cross-linker unbinding and actin polymerization
are two internal mechanisms that antagonize the rate of retrograde ow. This explanation can also be
shown according to equation (3.4); when is 0, the maximum shortening length is max =
22.
These two internal factors also determine the maximum force generated in the system in the case
of very high stinesses. As demonstrated by max =1+
2
+ 2, the maximum force in the
7
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
cell is not equal to . In this scenario, actin polymerization requires extra work by myosin motors,
which will be lost as a boundary creep. Furthermore, cross-linker unbinding will also result in force
dissipation, leading to internal creep.
An analogous active Maxwell uid model was employed to understand the mechanical character-
istics underlying the observed cellular response to local RhoA activation [87]. In their experiments,
Oakes et al. used optogenetic probes to recruit a cytosolic photo-recruitable protein RhoA-specic
guanine exchange factor (prGEF) to the plasma membrane and activated RhoA over periods of
15 minutes. During local activation of RhoA, the uorescent intensity of both actin and myosin II
increased exponentially, plateaued and then decreased during the relaxation period (gure 2c(i)). A
similar response has been observed while measuring traction forces and strain energy in cells (gure
2c(iii)). Besides, RhoA activation resulted in a sudden enhancement of traction force generation in the
cell borders, whereas in the activation region traction forces did not change.
In the model used to describe this behaviour, contractile stress is introduced as an internal
boundary condition, that is, 0<, = ± 0,=0, to represent RhoA activation in the cell area.
The prole of 0 follows a similar paern as the actin and myosin accumulation in the activation area
(gure 2c(ii)). In addition, a two-dimensional model of stress bres embedded in a passive viscoelastic
environment was used to estimate the direction and magnitude of the actomyosin ow towards the
activation region. Finally, through the application of this model, they showcased the role of Zyxin in
regulating the time scale of the initial elastic behaviour.
Active CEs are employed in chemomechanical models as well. These models integrate the eect of
the interplay between the mechanical characteristics of cells, external and internal mechanical stimuli
and the cascade of biochemical signals to simulate cell behaviour in various scenarios. For example,
a chemomechanical model illustrated the growth dynamics of cell–ECM adhesion structures and
highlighted the regulatory inuence of the stiness of the nucleus and ECM [100]. Another model
demonstrated how the interaction between cells and ECM aects both gene expression and nuclear
architecture [101]. This group of models is reviewed in [102,103].
4. Integrating activity in viscoelastic solid models
The previous section demonstrated how active tension originating from actomyosin dynamics can be
integrated within constitutive equations that relate force and rate of contraction. Étienne et al.’s model
provides, in particular, a detailed description of the transient regimes leading to the establishment
of steady active stress [38]. These descriptions treat active materials as uid without a reference
to an intrinsic shape. It is, however, common for tissues to exhibit solid-like characteristics, with a
well-dened reference shape, and the role of active stresses may then be interpreted as an apparent
tension or a change in reference shape. In this section, we review these dierent approaches, highlight-
ing their similarities and dierences.
4.1. Constant active element
A simple way to introduce active contraction across a material is to include a stress-generating unit of
constant value. Such an approach was used by Wya et al., when they investigated the short-time-scale
response of Madin–Darby canine kidney (MDCK) monolayers to in-plane compressions [18], a process
observed during morphogenetic processes [104,105] and the normal physiological function of many
epithelial tissues [106,107]. The MDCK monolayers were placed between two rods and compressed
at dierent rates. Quickly compressing the monolayer to strains below a threshold level (ε ~ 33%)
resulted in transient folds that disappeared in time scales of the order of seconds, while the folds
created owing to higher strains were permanent. The same buckling threshold of ε ~ 33% was observed
when the monolayers were compressed both rapidly and at a low rate. In addition, using actomyosin
inhibitors, they have demonstrated that actomyosin activity regulates the rate of tissue aening, the
buckling threshold, pretension and the long-time-scale stiness of the monolayers.
A simple active rheological model could reproduce the results of their studies conducted under
dierent loading conditions (gure 3a). The model consists of a constant active element , which
brings the system to a tensile state even at 0 external load, in parallel to a standard linear solid
model. The MDCK monolayer buckles when the stress in the monolayer approaches the compression
range. Therefore, under compressive strains that would normally cause compressive stress, the model
8
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
assumes that stress in the monolayer remains 0. The constitutive equation under tensile and compres-
sive stresses is dened as
(4.1)
= ++1and =under tensile stress,
= 0 and =
1under compressive stress,
where is the pretension in the monolayer, is the device strain, characterizes the long-term
stiness of the tissue, and describe the short-time-scale response, 1 is the strain in the spring
with stiness satisfying
˙1+1
=
˙, and =/ is the characteristic time scale. Based on the model
predictions for the steady state, when the applied strain is larger than the buckling threshold , >,
(a) Constant contractile element
(b)
(c)
(i) (ii)
(i)(ii)
Maxwell model
Considering cellular contractility
(d) Considering local actin concentration at the wound edge
The Munoz model
Figure 3. Active models for solid-like materials. (a). Active rheological model that includes a constant CE, , in parallel to a standard
solid model to account for the prestress in cell monolayers [18]. Temporal evolution of strain (i) and tissue stress (ii) as a function
of device strain presented for epithelial monolayers undergoing compression at a low strain rate (0.5% s1). (b) The active element
proposed by Muñoz et al., which adapts its resting length in response to deformations [108]. The evolution of stress over time in stress
relaxation tests predicted by the Maxwell model and active models with two denitions of elastic strain is presented in a table. Stress
in the models is dened using the elastic strain, that is, =. Changing the denition of elastic strain can result in an equivalent
evolution of stress in the active element and the Maxwell model for 1. When the strain is very small, the dierence between
the evolution of stresses in all three models will be negligible. However, at larger strains, the stress in the system would be dependent
on the way that the strain is dened. (c) The modied version of the active element proposed by Muñoz et al., which also considers
the eect of cellular contractility [109]. (i) The average curves representing the evolution of stress over time for MDCK monolayers
stretched at 75% s1 to various amplitudes of strain plotted on a semi-log scale. (ii) Demonstration of the correlation between the
model calculated from the active rheological model and the characteristic time calculated from tting the stress–time curves with
an empirical function, that is, =/+, for dierent loading conditions and actomyosin treatments. (d) An example of a
hybrid vertex model employed to study wound healing. The top view of the epithelial tissue with ablated cells is shown along with the
rheological models of nodal segments (cell-centre connections) in black and vertex segments (cell boundaries) in red [110].
9
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
tissue strain will be equal to the device strain, = and the stress in the tissue will follow =+ .
However, when the monolayer is compressed past the model buckling threshold, stress levels fall to 0
and tissue strain plateaus at =/. These model predictions are consistent with the experimental
results shown in gure 3a(i) and (ii). The model can also predict the response of the monolayer to
a step of compressive strain of dierent magnitudes. Besides, the model provides a simple way to
capture how treatments aecting actomyosin control the amount of active tension revealing that the
buckling threshold observed in the experiments agrees with model predictions / in all conditions.
4.2. Active behaviour modelled as change in the resting length of a spring
The previous approach focused on how active stress, combined with external perturbations, would set
the mechanical evolution of a viscoelastic tissue. The Maxwell branch and its dashpot, in particular,
account for the remodelling and plasticity of the material. Another way to represent this is through the
explicit evolution of the rest shape, or stress-free state, of the material. Both remodelling and tensioning
can be accounted for through the evolution of the resting length of cells and tissues over time, as
proposed by Muñoz et al. [108]. We rst describe how a Maxwell-like behaviour can emerge from this
strategy and then how activity can be added to this class of models.
4.2.1. Model proposed by Muñoz et al.
Muñoz et al. proposed a model to account for the inuence of cytoskeleton activity on cell shape by
representing plasticity as a change in the rest length of a spring [108]. A material’s constitutive equation
therefore takes the form of a relationship between the rate of change of the rest length and the stress (or
equivalently some metric of elastic strain) in the material. In this model, the current resting length of
the material (lament, cell or tissue) , that is, the total length of the material when no external load is
applied to it, is proportional to the elastic strain . The rate of changing the resting length under strain
is dened as
(4.2)
˙
=,
where is the remodelling rate of the network dened as the network resistance to adjusting its
conguration to the applied deformation, is the current elastic strain, =
and is the current total
length of the network. According to Muñoz’s model, the denition of current elastic strain is dierent
from the apparent strain =0
0 , where 0 is the initial length and the resting length of the network.
For small deformations, Muñoz’s approach is mathematically equivalent to a linear Maxwell model
with a characteristic time of =
[108]. However, at large deformation, the models dier, leading to
rather complex relaxation dynamics for Muñoz’s model (see gure 3b). This dierence results from
the particular denitions of strains and the resulting nonlinearities emerging from them. For instance,
using a logarithmic form for the elastic strain = ln
(i.e. using the true strain denition rather than
the engineering strain), Muñoz’s model would match the Maxwell model up to large deformations,
with a relaxation time scale independent of the strain amplitude, as shown in the table in gure 3b. At
a mathematical level, controlling the rest length of a spring or having a dashpot in series is therefore
largely equivalent.
However, introducing a dynamic rest length enables a slightly dierent interpretation of the
physiological mechanisms [108]. The dashpot element in the Maxwell model is often associated with
remodelling the system but could also account for the resistance of the cytoplasmic uid to the
applied strain rate and dissipate power. On the other hand, for Muñoz’s active element, the inelastic
part of the external power will be used to overcome the resistance of the cytoskeleton laments
to adapt to the new conguration imposed by the external strain. The active model proposed by
Muñoz has been generalized to two-dimensional/three-dimensional continuum models [111] and also
integrated into discrete models such as cell-centred [112], vertex [113] and cell-centred/vertex hybrid
[109,110] approaches. By incorporating a porosity parameter representing the density of polymers
in the cell cytoskeleton, the continuum model proposed by Asadipour et al. can also replicate the
immediate uidization in cells in response to transient strains and the subsequent gradual stiening
10
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
[111]. These adaptations have facilitated the study of epithelial tissue behaviour in both two and three
dimensions. A modelling approach similar to the Muñoz model proposed by Esfahani et al. could also
demonstrate stiening in response to high strain rates applied to epithelial cell pairs [49]. A few of the
modications made to the Muñoz model are presented in the following sections.
4.2.2. Active element considering cellular contractility
Mosaa et al. [109] modied the evolution law of the resting length of the material proposed by Muñoz
et al. [108] by introducing a contractility parameter to account for the inherent contractility of the
cells,
(4.3)
˙
= .
In this model, when the elastic strain reaches to , the resting length will not change any more,
and as previously stated [108], when is zero, the model behaves similarly to the Maxwell model.
Mosaa et al. implemented the modied active element in a hybrid cell-centred/vertex model where
cells interact through both cell centrs, presented by nodes and cell–cell junctions, presented by the
connection between vertices [109]. This model could successfully simulate tissue extension and wound
healing.
Khalilgharibi et al. have used a similar approach for ing the results of stress relaxation tests
conducted on MDCK monolayers [5]. Their studies have shown that stress in the MDCK monolayer
increases promptly after strain application. Then, the stress will gradually relax along with an increase
in the monolayer length, which is regulated by actomyosin activity. Moreover, they have noticed a
strain-dependent characteristic time for monolayers stretched at 75% 1 strain rate (gure 3c(i)),
which cannot be explained by standard linear viscoelastic models. Therefore, they proposed a model
that consists of an elastic spring in parallel to an active element that sustains a constant pre-strain
and changes its resting length to relax the imposed stress and return its strain to . The changing
of the resting length of the monolayer in response to an applied strain 0 is dened as
(4.4)
˙
=
= 0.
Here, is the eective strain = / with representing the actual length of the
monolayer after applying the deformation and is the rate of changing the resting length. The
characteristic time predicted by this model,  =0/1+0, increases with the applied strain,
which is consistent with their experimental observations, as shown in (gure 3c(ii)).
Another example of the active element that accounts for cellular contractility involves integrating
this model into a hybrid two-dimensional cell-centred/vertex model to analyse the wound healing
process [110]. In this model, the vertex segments (cell boundaries) and nodal segments (cell-centre
connections) connect the apical and basal sides. Nodal segments are characterized using an elastic
spring, and the behaviour of vertex segments is described by an active model consisting of two
branches in parallel (gure 3d). The rst branch is an active element that accounts for the changes
in the resting length of the vertex following equation (4.3). The second branch is an elastic spring
with an additional time-varying contractility parameter Υ
^, which accounts for the eect of high actin
concentration at the wound edge (i.e. purse string contractility) and increases the stress in the elastic
spring. The outcomes of their simulations have demonstrated the regulatory mechanism of both purse
string contractility and tissue contractility on wound healing speed.
4.3. Active element considering a time delay in the active rest length changes
One of the potential reasons for the oscillatory response observed in tissues during various processes,
such as morphogenesis, could be the delay between the signal and the response, as stated by Muñoz
et al. [113]. These delays in the responses can be owing to the distance between the sender and
receiver of the biochemical signals or the time necessary for signal processing [113,114]. To study
this phenomenon, they modied the active element that was previously proposed by Muñoz et al. by
11
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
considering the eect of a time delay between the mechanical signals and the active rest length changes
[108],
(4.5)
˙= .
Analysing the stability of the delay dierential equation resulted in the limits of oscillation and
stability as follows:
(4.6)oscil =1
 and stab =π
2.
oscil is the time beyond which the rest length of the element oscillates, is the exponential constant,
and for time scales above stab the value of the rest length is unstable, and its oscillation amplitude will
rise over time. These values might trigger oscillations during embryogenesis.
In addition, delays can also be dependent on the apparent size of the element . Muñoz et al.
implemented the eect of the size-dependent delays, = , into a vertex model to analyse the
oscillation in the cellular area in biological tissues [113]. In the vertex model, the rest length of the
nodal elements was maintained constant, and changes in the resting length of the vertex elements
were dened using equation (4.5). For constant delays, oscillations in the cell area were periodic and
synchronous. However, for size-dependent delays, oscillations started to get increasingly out of sync.
It is argued that this model demonstrated the role of delay in the mechanical response in inducing
oscillations even in the absence of external sources.
5. Summary and future perspectives
The active viscoelastic models outlined here are capable of capturing numerous aspects of cell
behaviour at multiple spatial and temporal scales with a small number of model parameters without
considering the details of the structural components and dynamics of cell–cell and cell–ECM adhesion
sites. These active models provided insight into the results of experiments and predicted the system
behaviour in other arbitrary conditions. Additionally, despite the diculty in establishing a clear
connection between biological processes and model parameters, researchers have used drug treatments
or targeted mutations to demonstrate correlations between model parameters like Young’s modulus,
viscosity or active pretension and biological processes such as actomyosin activity, even molecules that
regulate these processes.
Each active viscoelastic model is described by a constitutive equation that represents a particular
cellular behaviour, such as sensing changes in the substrate stiness, and response to strains at
dierent magnitudes and rates in cell doublets and cell monolayers. Therefore, the existence and
use of a generalized model that can be employed to describe and predict the response of cells in
dierent scenarios is still an open question. Integrating and bridging the gap between phenomenologi-
cal and biophysical models is an important step to improve our understanding of these systems. For
example, a molecular model of the actomyosin cortex inspired a phenomenological model for cell-scale
mechanosensing [38], and a phenomenological model that included the role of actin polymerization
in changing the resting length of the material was incorporated in vertex models to study dier-
ent aspects of tissue dynamics [113]. Consequently, a comprehensive phenomenological model that
captures all the signicant facets of rheological data might also enrich the ndings of biophysical
models and allow us to improve the precision of simulations of cell activity.
Numerical models are an excellent research tool to complement, analyse and interpret experimental
data in the eld of cell mechanics. Numerous models with varying degrees of complexity and details
of the structural elements involved in the observed phenomena have been presented over the years.
Power law [15], fractional (reviewed in [17]), viscoelastic and active viscoelastic models are included
in this category, where the eects of subcellular microstructures on cell rheology are represented by
model parameters. For example, virtual cell (VCell) is a powerful model that includes details of the
nucleus, cytoskeleton, cytoplasm and chromatin bres [115]. This level of detail is computationally
expensive and might not always be necessary. In other words, based on the research question, the
phenomenon of interest, length and time scales and characteristics of the relevant microstructures
can be incorporated into biophysical models. For example, at the molecular scale, chemomechanical
[103,116] and molecular clutch models [117–119] are used to study cell–cell and cell–ECM adhesions,
at the cell-scale statistical approaches can be employed to study cell mechanics [120–122], and cellular
Pos (CPM) [123–125], vertex [126,127] and self-propelled Voronoi [128–130] models are introduced to
12
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
study the mechanical behaviour of epithelial monolayers in two dimensions and three dimensions. The
level of detail can be further reduced by describing the outcomes of experiments using phenomenolog-
ical models before examining the underlying mechanisms.
Ethics. This work did not require ethical approval from a human subject or animal welfare commiee.
Data accessibility. This article has no additional data.
Declaration of AI use. We have not used AI-assisted technologies in creating this article.
Authors’ contributions. B.T.S.: writing—original draft, writing—review and editing; C.H.: writing—review and editing;
A.K.: writing—review and editing; R.Y.: writing—review and editing.
All authors gave nal approval for publication and agreed to be held accountable for the work performed
therein.
Conict of interest declaration. We declare we have no competing interests.
Funding. This study was supported by the NSF (1826135, 2143997) and the NIH (R35GM150623, P20GM113126).
Acknowledgements. R.Y. acknowledges fundings from the NSF (1826135, 2143997) and the NIH (R35GM150623,
P20GM113126).
References
1. Fritsch A, Höckel M, Kiessling T, Nnetu KD, Wetzel F, Zink M, Käs JA. 2010 Are biomechanical changes necessary for tumour progression? Nat.
Phys. 6, 730–732. (doi:10.1038/nphys1800)
2. Phipps S, Yang THJ, Habib FK, Reuben RL, McNeill SA. 2005 Measurement of tissue mechanical characteristics to distinguish between benign
and malignant prostatic disease. Urology 66, 447–450. (doi:10.1016/j.urology.2005.03.017)
3. Palacio-Torralba J, Hammer S, Good DW, Alan McNeill S, Stewart GD, Reuben RL, Chen Y. 2015 Quantitative diagnostics of soft tissue through
viscoelastic characterization using time-based instrumented palpation. J. Mech. Behav. Biomed. Mater. 41, 149–160. (doi:10.1016/j.jmbbm.
2014.09.027)
4. Khalilgharibi N, Fouchard J, Recho P, Charras G, Kabla A. 2016 The dynamic mechanical properties of cellularised aggregates. Curr. Opin. Cell
Biol. 42, 113–120. (doi:10.1016/j.ceb.2016.06.003)
5. Khalilgharibi N et al. 2019 Stress relaxation in epithelial monolayers is controlled by the actomyosin cortex. Nat. Phys. 15, 839–847. (doi:10.
1038/s41567-019-0516-6)
6. Forgacs G, Foty RA, Shafrir Y, Steinberg MS. 1998 Viscoelastic properties of living embryonic tissues: a quantitative study. Biophys. J. 74, 2227–
2234. (doi:10.1016/S0006-3495(98)77932-9)
7. Guevorkian K, Colbert MJ, Durth M, Dufour S, Brochard-Wyart F. 2010 Aspiration of biological viscoelastic drops. Phys. Rev. Lett. 104, 218101.
(doi:10.1103/PhysRevLett.104.218101)
8. Charras G, Yap AS. 2018 Tensile forces and mechanotransduction at cell-cell junctions. Curr. Biol. 28, R445–R457. (doi:10.1016/j.cub.2018.02.
003)
9. Prost J, Jülicher F, Joanny JF. 2015 Active gel physics. Nature Phys. 11, 111–117. (doi:10.1038/nphys3224)
10. Aigouy B, Farhadifar R, Staple DB, Sagner A, Röper JC, Jülicher F, Eaton S. 2010 Cell ow reorients the axis of planar polarity in the wing
epithelium of Drosophila. Cell 142, 773–786. (doi:10.1016/j.cell.2010.07.042)
11. Wyatt TPJ, Harris AR, Lam M, Cheng Q, Bellis J, Dimitracopoulos A, Kabla AJ, Charras GT, Baum B. 2015 Emergence of homeostatic epithelial
packing and stress dissipation through divisions oriented along the long cell axis. Proc. Natl Acad. Sci. USA 112, 5726–5731. (doi:10.1073/pnas.
1420585112)
12. Hart KC, Tan J, Siemers KA , Sim JY, Pruitt BL, Nelson WJ, G loerich M. 2017 E-cadherin and LGN align epithelial cell divisions with tissue tension
independently of cell shape. Proc. Natl Acad. Sci. 114, E5845–E5853. (doi:10.1073/pnas.1701703114)
13. Trepat X, Deng L, An SS, Navajas D, Tschumperlin DJ, Gerthoer WT, Butler JP, Fredberg JJ. 2007 Universal physical responses to stretch in the
living cell. Nature 447, 592–595. (doi:10.1038/nature05824)
14. Krishnan R etal. 2009 Reinforcement versus uidization in cytoskeletal mechanoresponsiveness. PLoS one 4, e5486. (doi:10.1371/journal.pone.
0005486)
15. Kollmannsberger P, Fabry B. 2011 Linear and nonlinear rheology of living cells. Annu. Rev. Mater. Res. 41, 75–97. (doi:10.1146/annurev-
matsci-062910-100351)
16. Rosowski KA, Boltyanskiy R, Xiang Y, Van den Dries K, Schwartz MA, Dufresne ER. 2018 Vinculin and the mechanical response of adherent
broblasts to matrix deformation. Sci. Rep. 8, 17967. (doi:10.1038/s41598-018-36272-9)
17. Bonfanti A, Kaplan JL, Charras G, Kabla A. 2020 Fractional viscoelastic models for power-law materials. Soft Matter 16, 6002–6020. (doi:10.
1039/d0sm00354a)
18. Wyatt TPJ, Fouchard J, Lisica A, Khalilgharibi N, Baum B, Recho P, Kabla AJ, Charras GT. 2020 Actomyosin controls planarity and folding of
epithelia in response to compression. Nat. Mater. 19, 109–117. (doi:10.1038/s41563-019-0461-x)
19. Darling EM, Di Carlo D. 2015 High-throughput assessment of cellular mechanical properties. Annu. Rev. Biomed. Eng. 17, 35–62. (doi:10.1146/
annurev-bioeng-071114-040545)
13
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
20. Quan FS, Kim KS. 2016 Medical applications of the intrinsic mechanical properties of single cells. Acta Biochim. Biophys. Sin. 48, 865–871. (doi:
10.1093/abbs/gmw081)
21. Cross SE, Jin YS, Rao J , Gimzewski JK. 2007 Nanomechanical analysis of cells from cancer patients. Nat. Nanotechnol. 2, 780–783. (doi:10.1038/
nnano.2007.388)
22. Wong VW, Akaishi S, Longaker MT, Gurtner GC. 2011 Pushing back: wound mechanotransduction in repair and regeneration. J. Invest. Dermatol.
131, 2186–2196. (doi:10.1038/jid.2011.212)
23. Wozniak MA, Chen CS. 2009 Mechanotransduction in development: a growing role for contractility. Nat. Rev. Mol. Cell Biol. 10, 34–43. (doi:10.
1038/nrm2592)
24. Balasubramaniam L, Mège RM, Ladoux B. 2022 Active nematics across scales from cytoskeleton organization to tissue morphogenesis. Curr.
Opin. Genet. Dev. 73, 101897. (doi:10.1016/j.gde.2021.101897)
25. Doostmohammadi A, Ignés-Mullol J, Yeomans JM, Sagués F. 2018 Active nematics. Nat. Commun. 9, 3246. (doi:10.1038/s41467-018-05666-8)
26. Needleman D, Dogic Z . 2017 Active matter at the interface between materials science and cell biology. Nat. Rev. Mater. 2, 1–14. (doi:10.1038/
natrevmats.2017.48)
27. Kumar KV. 2021 The actomyosin cortex of cells: a thin lm of active matter. J. Indian Inst. Sci. 101, 97–112. (doi:10.1007/s41745-020-00220-2)
28. Brown AL, Farhat W, Merguerian PA, Wilson GJ, Khour y AE, Woodhouse KA. 2002 22 week assessment of bladder acellular matrix as a bladder
augmentation material in a porcine model. Biomaterials 23, 2179–2190. (doi:10.1016/s0142-9612(01)00350-7)
29. Hill AV. 1938 The heat of shortening and the dynamic constants of muscle. Proc. R. Soc. Lond. B 126, 136–195. (doi:10.1098/rspb.1938.0050)
30. Harris AR, Bellis J, Khalilgharibi N, Wyatt T, Baum B, Kabla AJ, Charras GT. 2013 Generating suspended cell monolayers for mechanobiological
studies. Nat. Protoc. 8, 2516–2530. (doi:10.1038/nprot.2013.151)
31. Crichton ML, Donose BC, Chen X, Raphael AP, Huang H, Kendall MAF. 2011 The viscoelastic, hyperelastic and scale dependent behaviour of
freshly excised individual skin layers. Biomaterials 32, 4670–4681. (doi:10.1016/j.biomaterials.2011.03.012)
32. Uriarte JJ, Meirelles T, Gorbenko Del Blanco D, Nonaka PN, Campillo N, Sarri E, Navajas D, Egea G, Farré R. 2016 Early impairment of lung
mechanics in a murine model of Marfan syndrome. PLoS one 11, e0152124. (doi:10.1371/journal.pone.0152124)
33. Okajima T. 2012 Atomic force microscopy for the examination of single cell rheology. Curr. Pharm. Biotechnol. 13, 2623–2631. (doi:10.2174/
138920101314151120122846)
34. Weafer PP, Reynolds NH, Jarvis SP, McG arry JP. 2015 Single cell active force generation under dynamic loading – Part I: AFM experiments. Acta
Biomater. 27, 236–250. (doi:10.1016/j.actbio.2015.09.006)
35. Fischer-Friedrich E, Toyoda Y, Cattin CJ, Müller DJ, Hyman AA, Jülicher F. 2016 Rheology of the active cell cortex in mitosis. Biophys. J. 111,
589–600. (doi:10.1016/j.bpj.2016.06.008)
36. Helenius J, Heisenberg CP, Gaub HE, Muller DJ. 2008 Single-cell force spectroscopy. J. Cell Sci. 121, 1785–1791. (doi:10.1242/jcs.030999)
37. Mitrossilis D, Fouchard J, Guiroy A, Desprat N, Rodriguez N, Fabry B, Asnacios A. 2009 Single-cell response to stiness exhibits muscle-like
behavior. Proc. Natl Acad. Sci. USA 106, 18 243–18 248. (doi:10.1073/pnas.0903994106)
38. Étienne J, Fouchard J, Mitrossilis D, Bu N, Durand-Smet P , Asnacios A. 2015 Cells as liquid motors: mechanosensitivity emerges from collective
dynamics of actomyosin cortex. Proc. Natl Acad. Sci. USA 112, 2740–2745. (doi:10.1073/pnas.1417113112)
39. Bonakdar N, Gerum R, Kuhn M, Spörrer M, Lippert A, Schneider W, Aifantis KE, Fabry B. 2016 Mechanical plasticity of cells. Nat. Mater. 15,
1090–1094. (doi:10.1038/nmat4689)
40. Kilinc D, Lee GU. 2014 Advances in magnetic tweezers for single molecule and cell biophysics. Integr. Biol. 6, 27–34. (doi:10.1039/c3ib40185e)
41. Ayala YA etal. 2016 Rheological properties of cells measured by optical tweezers. BMC Biophys. 9, 5. (doi:10.1186/s13628-016-0031-4)
42. Arbore C, Perego L, Sergides M, Capitanio M. 2019 Probing force in living cells with optical tweezers: from single-molecule mechanics to cell
mechanotransduction. Biophys. Rev. 11, 765–782. (doi:10.1007/s12551-019-00599-y)
43. Nussenzveig HM. 2018 Cell membrane biophysics with optical tweezers. Eur. Biophys. J. 47, 499–514. (doi:10.1007/s00249-017-1268-9)
44. Hochmuth RM. 2000 Micropipette aspiration of living cells. J. Biomech. 33, 15–22. (doi:10.1016/s0021-9290(99)00175-x)
45. González-Bermúdez B, Guinea GV, Plaza GR. 2019 Advances in micropipette aspiration: applications in cell biomechanics, models, and
extended studies. Biophys. J. 116, 587–594. (doi:10.1016/j.bpj.2019.01.004)
46. Wang H, Zhou F, Guo Y, Ju LA. 2022 Micropipette-based biomechanical nanotools on living cells. Eur. Biophys. J. 51, 119–133. (doi:10.1007/
s00249-021-01587-5)
47. Vedula SRK, Lim TS, Kausalya PJ, Lane EB, Rajagopal G, Hunziker W, Lim CT. 2009 Quantifying forces mediated by integral tight junction
proteins in cell–cell adhesion. Exp. Mech. 49, 3–9. (doi:10.1007/s11340-007-9113-1)
48. Tashiro H, Uchida M, Sato-Maeda M . 1993 Three-dimensional cell manipulator by means of optical trapping for the specication of cell-to-cell
adhesion. Opt. Eng. 32, 2812. (doi:10.1117/12.147714)
49. Esfahani AM etal. 2021 Characterization of the strain-rate-dependent mechanical response of single cell-cell junctions. Proc. Natl Acad. Sci. USA
118, e2019347118. (doi:10.1073/pnas.2019347118)
50. Yang R, Broussard JA, Green KJ, Espinosa HD. 2018 Techniques to stimulate and interrogate cell-cell adhesion mechanics. Extreme Mech. Lett.
20, 125–139. (doi:10.1016/j.eml.2017.12.002)
51. Monemian Esfahani A, Rosenbohm J, Reddy K, Jin X, Bouzid T, Riehl B, Kim E, Lim JY, Yang R. 2019 Tissue regeneration from mechanical
stretching of cell-cell adhesion. Tissue Eng. Part C. Methods 25, 631–640. (doi:10.1089/ten.TEC.2019.0098)
52. Dahl KN, Engler AJ , Pajerowski JD, Discher DE. 2005 Power-law rheology of isolated nuclei with deformation mapping of nuclear substructures.
Biophys. J. 89, 2855–2864. (doi:10.1529/biophysj.105.062554)
14
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
53. Davidson PM, Fedorchak GR, Mondésert-Deveraux S, Bell ES, Isermann P, Aubry D, Allena R, Lammerding J. 2019 High-throughput microuidic
micropipette aspiration device to probe time-scale dependent nuclear mechanics in intact cells. Lab Chip 19, 3652–3663. (doi:10.1039/
c9lc00444k)
54. Hobson CM, Falvo MR, Superne R. 2021 A survey of physical methods for studying nuclear mechanics and mechanobiology. APL Bioeng. 5,
041508. (doi:10.1063/5.0068126)
55. Tlili S, Gay C, Graner F, Marcq P, Molino F, Saramito P. 2015 Colloquium: mechanical formalisms for tissue dynamics. Eur. Phys. J. E. 38, 1–31.
(doi:10.1140/epje/i2015-15033-4)
56. Bernheim-Groswasser A, Gov NS, Safran SA, Tzlil S. 2018 Living matter: mesoscopic active materials. Adv. Mater. 30, e1707028. (doi:10.1002/
adma.201707028)
57. Banerjee S, Gardel ML, Schwarz US. 2020 The actin cytoskeleton as an active adaptive material. Annu. Rev. Condens. Matter Phys. 11, 421–439.
(doi:10.1146/annurev-conmatphys-031218-013231)
58. Fletcher DA, Mullins RD. 2010 Cell mechanics and the cytoskeleton. Nature 463, 485–492. (doi:10.1038/nature08908)
59. Xi W, Saw TB, Delacour D, Lim CT, Ladoux B. 2019 Material approaches to active tissue mechanics. Nat. Rev. Mater. 4, 23–44. (doi:10.1038/
s41578-018-0066-z)
60. Carlier MF, Shekhar S. 2017 Global treadmilling coordinates actin turnover and controls the size of actin networks. Nat. Rev. Mol. Cell Biol. 18,
389–401. (doi:10.1038/nrm.2016.172)
61. Pollard TD, Berro J. 2009 Mathematical models and simulations of cellular processes based on actin laments. J. Biol. Chem. 284, 5433–5437.
(doi:10.1074/jbc.R800043200)
62. Pollard TD. 2016 Theory from the Oster laboratory leaps ahead of experiment in understanding actin-based cellular motility. Biophys. J. 111,
1589–1592. (doi:10.1016/j.bpj.2016.08.044)
63. Wegner A. 1976 Head to tail polymerization of actin. J. Mol. Biol. 108, 139–150. (doi:10.1016/s0022-2836(76)80100-3)
64. Banerjee S, Gardel ML, Schwarz US. 2020 The actin cytoskeleton as an active adaptive material. Annu. Rev. Condens. Matter Phys. 11, 421–439.
(doi:10.1146/annurev-conmatphys-031218-013231)
65. Neuhaus JM, Wanger M, Keiser T, Wegner A. 1983 Treadmilling of actin. J. Muscle Res. Cell Motil. 4, 507–527. (doi:10.1007/BF00712112)
66. Chaudhuri O, Cooper-White J, Janmey PA, Mooney DJ, Shenoy VB. 2020 Eects of extracellular matrix viscoelasticity on cellular behaviour.
Nature 584, 535–546. (doi:10.1038/s41586-020-2612-2)
67. Alert R, Trepat X. 2020 Physical models of collective cell migration. Annu. Rev. Condens. Matter Phys 11, 77–101. (doi:10.1146/annurev-
conmatphys-031218-013516)
68. Robert-Paganin J, Pylypenko O, Kikuti C, Sweeney HL, Houdusse A. 2020 Force generation by myosin motors: a structural perspective. Chem.
Rev. 120, 5–35. (doi:10.1021/acs.chemrev.9b00264)
69. Houdusse A, Sweeney HL. 2016 How myosin generates force on actin laments. Trends Biochem. Sci. 41, 989–997. (doi:10.1016/j.tibs.2016.09.
006)
70. Murrell M, Oakes PW, Lenz M, Gardel ML. 2015 Forcing cells into shape: the mechanics of actomyosin contractility. Nat. Rev. Mol. Cell Biol. 16,
486–498. (doi:10.1038/nrm4012)
71. Livne A, Geiger B. 2016 The inner workings of stress bers from contractile machinery to focal adhesions and back. J. Cell Sci. 129, 1293–
1304. (doi:10.1242/jcs.180927)
72. Yap AS, Duszyc K, Viasno V. 2018 Mechanosensing and mechanotransduction at cell-cell junctions. Cold Spring Harb. Perspect. Biol. 10,
a028761. (doi:10.1101/cshperspect.a028761)
73. Shukla VC, Higuita-Castro N, Nana-Sinkam P, Ghadiali SN. 2016 Substrate stiness modulates lung cancer cell migration but not epithelial to
mesenchymal transition. J. Biomed. Mater. Res. A. 104, 1182–1193. (doi:10.1002/jbm.a.35655)
74. Hadden WJ et al. 2017 Stem cell migration and mechanotransduction on linear stiness gradient hydrogels. Proc. Natl Acad. Sci. USA 114,
5647–5652. (doi:10.1073/pnas.1618239114)
75. Alert R, Trepat X. 2019 Physical models of collective cell migration. arXiv. See https://arxiv.org/abs/1905.07675
76. Sun M et al. 2018 Eects of matrix stiness on the morphology, adhesion, proliferation and osteogenic dierentiation of mesenchymal stem
cells. Int. J. Med. Sci. 15, 257–268. (doi:10.7150/ijms.21620)
77. Liu N etal. 2018 Eect of substrate stiness on proliferation and dierentiation of periodontal ligament stem cells. Cell Prolif. 51, e12478. (doi:
10.1111/cpr.12478)
78. Engler AJ, Sen S, Sweeney HL, Discher DE. 2006 Matrix elasticity directs stem cell lineage specication. Cell 126, 677–689. (doi:10.1016/j.cell.
2006.06.044)
79. Mao AS, Shin JW, Mooney DJ. 2016 Eects of substrate stiness and cell-cell contact on mesenchymal stem cell dierentiation. Biomaterials
98, 184–191. (doi:10.1016/j.biomaterials.2016.05.004)
80. Li B, Wang JHC. 2011 Fibroblasts and myobroblasts in wound healing: force generation and measurement. J. Tissue Viability 20, 108–120.
(doi:10.1016/j.jtv.2009.11.004)
81. Bosveld F et al. 2012 Mechanical control of morphogenesis by Fat/Dachsous/Four-jointed planar cell polarity pathway. Science 336, 724–727.
(doi:10.1126/science.1221071)
82. Yamada KM et al. 2019 Extracellular matrix dynamics in cell migration, invasion and tissue morphogenesis. Int. J. Exp. Pathol. 100, 144–152.
(doi:10.1111/iep.12329)
15
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
83. Chowdhury F, Huang B, Wang N. 2021 Cytoskeletal prestress: the cellular hallmark in mechanobiology and mechanomedicine. Cytoskeleton
(Hoboken) 78, 249–276. (doi:10.1002/cm.21658)
84. Rodriguez ML, McGarr y PJ , Sniadecki NJ. 2013 Review on cell mechanics: experimental and modeling approaches. Appl. Mech. Rev. 65. (doi:10.
1115/1.4025355)
85. Doss BL, Pan M, G upta M, Grenci G, Mège RM, Lim CT, Sheetz MP , Voituriez R, Ladoux B. 2020 Cell response to substrate rigidity is regulated by
active and passive cytoskeletal stress. Proc. Natl Acad. Sci. USA 117, 12 817–12 825. (doi:10.1073/pnas.1917555117)
86. Herzog W. 2019 The problem with skeletal muscle series elasticity. BMC Biomed. Eng. 1, 28. (doi:10.1186/s42490-019-0031-y)
87. Oakes PW, Wagner E, Brand CA, Probst D, Linke M, Schwarz US, Glotzer M, Gardel ML. 2017 Optogenetic control of RhoA reveals zyxin-mediated
elasticity of stress bres. Nat. Commun. 8, 15817. (doi:10.1038/ncomms15817)
88. McMahon TA. 1984 Muscles, reexes, and locomotion. Princeton University Press. (doi:10.1515/9780691221540)
89. Arslan YZ, Karabulut D, Ortes F, Popovic MB. 2019 Exoskeletons, exomusculatures, exosuits: dynamic modeling and simulation. In
Biomechatronics, (ed. MB Popovic), pp. 305–331. London, UK/San Diego, CA: Elsevier.
90. Regazzoni F, Dedè L, Quarteroni A. 2021 Active force generation in cardiac muscle cells: mathematical modeling and numerical simulation of
the actin-myosin interaction. Vietnam J. Math. 49, 87–118. (doi:10.1007/s10013-020-00433-z)
91. Wakeling JM, Febrer-Nafría M, De Groote F. 2023 A review of the eorts to develop muscle and musculoskeletal models for biomechanics in the
last 50 years. J. Biomech. 155, 111657. (doi:10.1016/j.jbiomech.2023.111657)
92. Discher DE, Janmey P, Wang YL. 2005 Tissue cells feel and respond to the stiness of their substrate. Science 310, 1139–1143. (doi:10.1126/
science.1116995)
93. Janmey PA, Fletcher DA, Reinhart-King CA. 2020 Stiness sensing by cells. Physiol. Rev. 100, 695–724. (doi:10.1152/physrev.00013.2019)
94. Chakraborty M et al. 2021 Mechanical stiness controls dendritic cell metabolism and function. Cell Rep. 34, 108609. (doi:10.1016/j.celrep.
2020.108609)
95. Huxley AF. 1957 Muscle structure and theories of contraction. Prog. Biophys. Biophys. Chem 7, 255–318. (doi:10.1016/s0096-4174(18)30128-8)
96. Huxley AF, Simmons RM. 1971 Proposed mechanism of force generation in striated muscle. Nature 233, 533–538. (doi:10.1038/233533a0)
97. Powers JD, Malingen SA, Regnier M , Daniel TL. 2021 The sliding lament theory since Andrew Huxley: multiscale and multidisciplinary muscle
research. Annu. Rev. Biophys. 50, 373–400. (doi:10.1146/annurev-biophys-110320-062613)
98. Moreo P, García-Aznar JM, Doblaré M. 2008 Modeling mechanosensing and its eect on the migration and proliferation of adherent cells. Acta
Biomater. 4, 613–621. (doi:10.1016/j.actbio.2007.10.014)
99. Borau C, Kamm RD, García-Aznar JM. 2014 A time-dependent phenomenological model for cell mechano-sensing. Biomech. Model.
Mechanobiol. 13, 451–462. (doi:10.1007/s10237-013-0508-x)
100. Cao X, Lin Y, Driscoll TP, Franco-Barraza J, Cukierman E, Mauck RL, Shenoy VB. 2015 A chemomechanical model of matrix and nuclear rigidity
regulation of focal adhesion size. Biophys. J. 109, 1807–1817. (doi:10.1016/j.bpj.2015.08.048)
101. Alisafaei F, Jokhun DS, Shivashankar GV, Shenoy VB. 2019 Regulation of nuclear architecture, mechanics, and nucleocytoplasmic shuttling of
epigenetic factors by cell geometric constraints. Proc. Natl Acad. Sci. USA 116, 13 200–13 209. (doi:10.1073/pnas.1902035116)
102. Sun SY Zhang H Fang W Chen X Li B Feng XQ. 2022 Bio-chemo-mechanical coupling models of soft biological materials: a review. Adv. Appl.
Mech 55, 309–392. (doi:10.1016/bs.aams.2022.05.004)
103. Senthilkumar I, Howley E, McEvoy E. 2022 Thermodynamically-motivated chemo-mechanical models and multicellular simulation to provide
new insight into active cell and tumour remodelling. Exp. Cell Res. 419, 113317. (doi:10.1016/j.yexcr.2022.113317)
104. Sidhaye J, Norden C. 2017 Concerted action of neuroepithelial basal shrinkage and active epithelial migration ensures ecient optic cup
morphogenesis. Elife 6, e22689. (doi:10.7554/eLife.22689)
105. Tallinen T, Chung JY, Rousseau F , Girard N, Lefèvre J, Mahadevan L. 2016 On the growth and form of cortical convolutions. Nat. Phys. 12, 588–
593. (doi:10.1038/nphys3632)
106. Grainge CL, Lau LCK, Ward JA, Dulay V, Lahi G, Wilson S, Holgate S, Davies DE, Howarth PH. 2011 Eect of bronchoconstriction on airway
remodeling in asthma. N. Engl. J. Med. 364, 2006–2015. (doi:10.1056/NEJMoa1014350)
107. Park JA etal. 2015 Unjamming and cell shape in the asthmatic airway epithelium. Nat. Mater. 14, 1040–1048. (doi:10.1038/nmat4357)
108. Muñoz JJ, Albo S. 2013 Physiology-based model of cell viscoelasticity. Phys. Rev. E. Stat. Nonlin. Soft Matter Phys. 88, 012708. (doi:10.1103/
PhysRevE.88.012708)
109. Mosaa P, Rodríguez-Ferran A, Muñoz JJ. 2018 Hybrid cell-centred/vertex model for multicellular systems with equilibrium-preserving
remodelling. Int. J. Numer. Method. Biomed. Eng. 34, e2928. (doi:10.1002/cnm.2928)
110. Ioannou F, Dawi MA, Tetley RJ, Mao Y, Muñoz JJ. 2020 Development of a new 3D hybrid model for epithelia morphogenesis. Front. Bioeng.
Biotechnol. 8, 405. (doi:10.3389/fbioe.2020.00405)
111. Asadipour N, Trepat X, Muñoz JJ. 2016 Porous-based rheological model for tissue uidisation. J. Mech. Phys. Solids 96, 535–549. (doi:10.1016/j.
jmps.2016.07.002)
112. Mosaa P, Asadipour N, Millán D, Rodríguez-Ferran A, J Muñoz J. 2015 Cell-centred model for the simulation of curved cellular monolayers.
Comp. Part. Mech. 2, 359–370. (doi:10.1007/s40571-015-0043-x)
113. Muñoz JJ, Dingle M, Wenzel M. 2018 Mechanical oscillations in biological tissues as a result of delayed rest-length changes. Phys. Rev. E 98,
052409. (doi:10.1103/PhysRevE.98.052409)
114. Solon J, Kaya-Copur A, Colombelli J, Brunner D. 2009 Pulsed forces timed by a ratchet-like mechanism drive directed tissue movement during
dorsal closure. Cell 137, 1331–1342. (doi:10.1016/j.cell.2009.03.050)
16
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
115. Heydari T, Heidari M, Mashinchian O, Wojcik M, Xu K, Dalby MJ, Mahmoudi M, Ejtehadi MR. 2017 Development of a virtual cell model to predict
cell response to substrate topography. ACS Nano 11, 9084–9092. (doi:10.1021/acsnano.7b03732)
116. McEvoy E etal. 2022 Feedback between mechanosensitive signaling and active forces governs endothelial junction integrity. Nat. Commun. 13,
7089. (doi:10.1038/s41467-022-34701-y)
117. Elosegui-Artola A, Oria R, Chen Y, Kosmalska A, Pérez-González C, Castro N, Zhu C, Trepat X, Roca-Cusachs P. 2016 Mechanical regulation of a
molecular clutch denes force transmission and transduction in response to matrix rigidity. Nat. Cell Biol. 18, 540–548. (doi:10.1038/ncb3336)
118. Bennett M, Cantini M, Reboud J, Cooper JM, Roca-Cusachs P, Salmeron-Sanchez M. 2018 Molecular clutch drives cell response to surface
viscosity. Proc. Natl Acad. Sci. USA 115, 1192–1197. (doi:10.1073/pnas.1710653115)
119. Andreu I et al. 2021 The force loading rate drives cell mechanosensing through both reinforcement and cytoskeletal softening. Nat. Commun.
12, 4229. (doi:10.1038/s41467-021-24383-3)
120. McEvoy E, Shishvan SS, Deshpande VS, McGarry JP. 2018 Thermodynamic modeling of the statistics of cell spreading on ligand-coated elastic
substrates. Biophys. J. 115, 2451–2460. (doi:10.1016/j.bpj.2018.11.007)
121. Shishvan SS, Vigliotti A, Deshpande VS. 2018 The homeostatic ensemble for cells. Biomech. Model. Mechanobiol. 17, 1631–1662. (doi:10.1007/
s10237-018-1048-1)
122. McEvoy E, Deshpande VS, McGarry P. 2019 Transient active force generation and stress bre remodelling in cells under cyclic loading. Biomech.
Model. Mechanobiol. 18, 921–937. (doi:10.1007/s10237-019-01121-9)
123. Wortel IMN, Niculescu I, Kolijn PM, Gov NS, de Boer RJ, Textor J. 2021 Local actin dynamics couple speed and persistence in a cellular Potts
model of cell migration. Biophys. J. 120, 2609–2622. (doi:10.1016/j.bpj.2021.04.036)
124. Hirway SU, Lemmon CA, Weinberg SH. 2021 Multicellular mechanochemical hybrid cellular Potts model of tissue formation during epithelial‐
mesenchymal transition. Comp. Sys. Onco. 1, e1031. (doi:10.1002/cso2.1031)
125. Odagiri K, Fujisaki H, Takada H, Ogawa R. 2022 Numerical simulation using cellular Potts model for wound closure with ATP release and the
Mechanobioligical eects. arXiv See https://arxiv.org/pdf/2209.01354.pdf.
126. Fletcher AG, Ostereld M, Baker RE, Shvartsman SY. 2014 Vertex models of epithelial morphogenesis. Biophys. J. 106, 2291–2304. (doi:10.
1016/j.bpj.2013.11.4498)
127. Alt S, Ganguly P, Salbreux G. 2017 Vertex models: from cell mechanics to tissue morphogenesis. Phil. Trans. R. Soc. Lond. B. 372, 20150520. (doi:
10.1098/rstb.2015.0520)
128. Li B, Sun SX. 2014 Coherent motions in conuent cell monolayer sheets. Biophys. J. 107, 1532–1541. (doi:10.1016/j.bpj.2014.08.006)
129. Bi D, Yang X, Marchetti MC, Manning ML . 2016 Motility-driven glass and jamming transitions in biological tissues. Phys. Rev. X 6, 021011. (doi:
10.1103/PhysRevX.6.021011)
130. Giavazzi F, Paoluzzi M, Macchi M, Bi D, Scita G, Manning ML, Cerbino R, Marchetti MC. 2018 Flocking transitions in conuent tissues. Soft Matter
14, 3471–3477. (doi:10.1039/c8sm00126j)
17
royalsocietypublishing.org/journal/rsos R. Soc. Open Sci. 11: 231074
ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
The formation and recovery of gaps in the vascular endothelium governs a wide range of physiological and pathological phenomena, from angiogenesis to tumor cell extravasation. However, the interplay between the mechanical and signaling processes that drive dynamic behavior in vascular endothelial cells is not well understood. In this study, we propose a chemo-mechanical model to investigate the regulation of endothelial junctions as dependent on the feedback between actomyosin contractility, VE-cadherin bond turnover, and actin polymerization, which mediate the forces exerted on the cell-cell interface. Simulations reveal that active cell tension can stabilize cadherin bonds, but excessive RhoA signaling can drive bond dissociation and junction failure. While actin polymerization aids gap closure, high levels of Rac1 can induce junction weakening. Combining the modeling framework with experiments, our model predicts the influence of pharmacological treatments on the junction state and identifies that a critical balance between RhoA and Rac1 expression is required to maintain junction stability. Our proposed framework can help guide the development of therapeutics that target the Rho family of GTPases and downstream active mechanical processes. Gap formation in the vasculature underpins immune and tumour cell infiltration. Here the authors propose a chemo-mechanical model to analyse how feedback between mechanosensitive signalling, active cellular forces and adhesion governs the breakdown, recovery, and integrity of endothelial junctions.
Article
Full-text available
Mechanobiology is an emerging field at the interface of biology and mechanics, investigating the roles of mechanical forces within biomolecules, organelles, cells, and tissues. As a highlight, the recent advances of micropipette-based aspiration assays and dynamic force spectroscopies such as biomembrane force probe (BFP) provide unprecedented mechanobiological insights with excellent live-cell compatibility. In their classic applications, these assays measure force-dependent ligand–receptor-binding kinetics, protein conformational changes, and cellular mechanical properties such as cortical tension and stiffness. In recent years, when combined with advanced microscopies in high spatial and temporal resolutions, these biomechanical nanotools enable characterization of receptor-mediated cell mechanosensing and subsequent organelle behaviors at single-cellular and molecular level. In this review, we summarize the latest developments of these assays for live-cell mechanobiology studies. We also provide perspectives on their future upgrades with multimodal integration and high-throughput capability.
Article
Full-text available
Biological tissues are composed of various cell types working cooperatively to perform their respective function within organs and the whole body. During development, embryogenesis followed by histogenesis relies on orchestrated division, death, differentiation and collective movements of cellular constituents. These cells are anchored to each other and/or the underlying substrate through adhesion complexes and they regulate force generation by active cytoskeleton remodelling. The resulting contractility related changes at the level of each single cell impact tissue architecture by triggering changes in cell shape, cell movement and remodelling of the surrounding environment. These out of equilibrium processes occur through the consumption of energy, allowing biological systems to be described by active matter physics. ‘Active nematics’ a subclass of active matter encompasses cytoskeleton filaments, bacterial and eukaryotic cells allowing them to be modelled as rod-like elements to which nematic liquid crystal theories can be applied. In this review, we will discuss the concept of active nematics to understand biological processes across subcellular and multicellular scales, from single cell organization to cell extrusion, collective cell movements, differentiation and morphogenesis.
Article
Full-text available
Epithelial‐mesenchymal transition (EMT) is the transdifferentiation of epithelial cells to a mesenchymal phenotype, in which cells lose epithelial‐like cell–cell adhesions and gain mesenchymal‐like enhanced contractility and mobility. EMT is crucial for tissue regeneration and is also implicated in pathological conditions, such as cancer metastasis. Prior work has shown that transforming growth factor‐β1 (TGF‐β1) is a potent inducer of this biological process. In this study, we develop a computational model coupling mechanical and biochemical signaling in a multicellular tissue undergoing EMT. Specifically, we utilize a recently developed formulation that integrates a multicellular cellular Potts model (CPM), a lattice‐based stochastic model governing cell movement; a first moment of area model, governing cellular traction and junctional forces; a finite element model, which defines extracellular matrix (ECM) substrate strains; an intracellular signaling TGF‐β1‐mediated EMT model that governs cellular phenotype; and an extracellular signaling component governing ECM and TGF‐β1 signaling. In this study, we modeled the spatial cellular patterns that occur in tissue and the ECM during EMT. Our model predicts that EMT often initially occurs at a tissue boundary due to mechanochemical coupling, which results in transdifferentiation to progress inwards toward the center. Variation in model parameters demonstrated conditions enhancing and suppressing EMT, especially to drive EMT in the absence of TGF‐β1 and inhibit EMT in the presence of TGF‐β1. Specifically, enhancing the mechanochemical feedback typically promoted EMT, whereas greater assembled ECM degradation suppressed EMT. Simulated scratch test experiments illustrate that ECM composition can impact closure directly through EMT signaling. In conclusion, we integrated mechanical, biochemical, and extracellular signaling networks in a novel hybrid computational model to reproduce tissue formation dynamics of EMT.
Article
Full-text available
It is increasingly appreciated that the cell nucleus is not only a home for DNA but also a complex material that resists physical deformations and dynamically responds to external mechanical cues. The molecules that confer mechanical properties to nuclei certainly contribute to laminopathies and possibly contribute to cellular mechanotransduction and physical processes in cancer such as metastasis. Studying nuclear mechanics and the downstream biochemical consequences or their modulation requires a suite of complex assays for applying, measuring, and visualizing mechanical forces across diverse length, time, and force scales. Here, we review the current methods in nuclear mechanics and mechanobiology, placing specific emphasis on each of their unique advantages and limitations. Furthermore, we explore important considerations in selecting a new methodology as are demonstrated by recent examples from the literature. We conclude by providing an outlook on the development of new methods and the judicious use of the current techniques for continued exploration into the role of nuclear mechanobiology.
Article
Full-text available
Cell response to force regulates essential processes in health and disease. However, the fundamental mechanical variables that cells sense and respond to remain unclear. Here we show that the rate of force application (loading rate) drives mechanosensing, as predicted by a molecular clutch model. By applying dynamic force regimes to cells through substrate stretching, optical tweezers, and atomic force microscopy, we find that increasing loading rates trigger talin-dependent mechanosensing, leading to adhesion growth and reinforcement, and YAP nuclear localization. However, above a given threshold the actin cytoskeleton softens, decreasing loading rates and preventing reinforcement. By stretching rat lungs in vivo, we show that a similar phenomenon may occur. Our results show that cell sensing of external forces and of passive mechanical parameters (like tissue stiffness) can be understood through the same mechanisms, driven by the properties under force of the mechanosensing molecules involved. Cells sense mechanical forces from their environment, but the precise mechanical variable sensed by cells is unclear. Here, the authors show that cells can sense the rate of force application, known as the loading rate, with effects on YAP nuclear localization and cytoskeletal stiffness remodelling.
Article
Full-text available
Cell migration is astoundingly diverse. Molecular signatures, cell-cell interactions, and environmental structures each play their part in shaping cell motion, yielding numerous morphologies and migration modes. Nevertheless, in recent years, a simple unifying law was found to describe cell migration across many different cell types and contexts: faster cells turn less frequently. This universal coupling between speed and persistence (UCSP) was explained by retrograde actin flow from front to back, but it remains unclear how this mechanism generalizes to cells with complex shapes and cells migrating in structured environments, which may not have a well-defined front-to-back orientation. Here, we present an in-depth characterisation of an existing Cellular Potts Model (CPM), in which cells polarize dynamically from a combination of local actin dynamics (stimulating protrusions) and global membrane tension along the perimeter (inhibiting protrusions). We first show that the UCSP emerges spontaneously in this model through a crosstalk of intracellular mechanisms, cell shape, and environmental constraints, resembling the dynamic nature of cell migration in vivo. Importantly, we find that local protrusion dynamics suffice to reproduce the UCSP — even in cases where no clear global, front-to-back polarity exists. We then harness the spatial nature of the CPM to show how cell shape dynamics limit both the speed and persistence a cell can reach, and how a rigid environment such as the skin can restrict cell motility even further. Our results broaden the range of potential mechanisms underlying the speed-persistence coupling that has emerged as a fundamental property of migrating cells.
Article
Both the Hill and the Huxley muscle models had already been described by the time the International Society of Biomechanics was founded 50 years ago, but had seen little use before the 1970s due to the lack of computing. As computers and computational methods became available in the 1970s, the field of musculoskeletal modeling developed and Hill type muscle models were adopted by biomechanists due to their relative computational simplicity as compared to Huxley type muscle models. Muscle forces computed by Hill type muscle models provide good agreement in conditions similar to the initial studies, i.e. for small muscles contracting under steady and controlled conditions. However, more recent validation studies have identified that Hill type muscle models are least accurate for natural in vivo locomotor behaviours at submaximal activations, fast speeds and for larger muscles, and thus need to be improved for their use in understanding human movements. Developments in muscle modelling have tackled these shortcomings. However, over the last 50 years musculoskeletal simulations have been largely based on traditional Hill type muscle models or even simplifications of this model that neglected the interaction of the muscle with a compliant tendon. The introduction of direct collocation in musculoskeletal simulations about 15 years ago along with further improvements in computational power and numerical methods enabled the use of more complex muscle models in simulations of whole-body movement. Whereas Hill type models are still the norm, we may finally be ready to adopt more complex muscle models into musculoskeletal simulations of human movement.
Chapter
Physiological and pathological processes of living soft tissues always involve complicated interplays of biological, chemical, mechanical, and physical cues. Understanding the coupling mechanisms of these factors and their biological implications has attracted extensive interest from the fields of biomechanics, biophysics, biomaterials, and medical engineering. In this paper, we review the state-of-the-art of bio-chemo-mechanical coupling theories for soft biological tissues (e.g., tumors, blood vessels, brains, and nerves), synthetic materials (e.g., polymers and hydrogels), and organoids. Particular attention is paid to the modeling methods of bio-chemo-mechanical coupling mechanisms spanning from the molecular level, the cellular level, to the tissular level. Representative examples are provided to reveal the essential roles of bio-chemo-mechanical interplay in biological dynamic processes, including those associated with diseases. This review may not only help to further develop theoretical models and experimental methods of cells, tissues, and biological materials, but also gain insights into the dependence of the macroscopic mechanical properties and biological functions on the microscopic bio-chemo-mechanical coupling mechanisms.
Article
Computational models can shape our understanding of cell and tissue remodelling, from cell spreading, to active force generation, adhesion, and growth. In this mini-review, we discuss recent progress in modelling of chemo-mechanical cell behaviour and the evolution of multicellular systems. In particular, we highlight recent advances in (i) free-energy based single cell models that can provide new fundamental insight into cell spreading, cancer cell invasion, stem cell differentiation, and remodelling in disease, and (ii) mechanical agent-based models to simulate large numbers of discrete interacting cells in proliferative tumours. We describe how new biological understanding has emerged from such theoretical models, and the trade-offs and constraints associated with current approaches. Ultimately, we aim to make a case for why theory should be integrated with an experimental workflow to optimise new in-vitro studies, to predict feedback between cells and their microenvironment, and to deepen understanding of active cell behaviour.