An oscillating dynamic model of collective cells in a ...

1 downloads 0 Views 4MB Size Report
Sep 29, 2017 - mesoderm invagination (Martin et al., 2009) and in amnioserosa cells during dorsal closure (Solon et al., 2009). Shape os- cillations are also ...
Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

Contents lists available at ScienceDirect

Journal of the Mechanics and Physics of Solids journal homepage: www.elsevier.com/locate/jmps

An oscillating dynamic model of collective cells in a monolayer Shao-Zhen Lin, Shi-Lei Xue, Bo Li∗, Xi-Qiao Feng∗ Institute of Biomechanics and Medical Engineering, AML and CNMM, Department of Engineering Mechanics, Tsinghua University, Beijing 100084, China

a r t i c l e

i n f o

Article history: Received 24 July 2017 Revised 9 September 2017 Accepted 23 September 2017 Available online 29 September 2017 Keywords: Cytodynamics Cellular vertex model Oscillation Chemomechanical coupling Hopf bifurcation

a b s t r a c t Periodic oscillations of collective cells occur in the morphogenesis and organogenesis of various tissues and organs. In this paper, an oscillating cytodynamic model is presented by integrating the chemomechanical interplay between the RhoA effector signaling pathway and cell deformation. We show that both an isolated cell and a cell aggregate can undergo spontaneous oscillations as a result of Hopf bifurcation, upon which the system evolves into a limit cycle of chemomechanical oscillations. The dynamic characteristics are tailored by the mechanical properties of cells (e.g., elasticity, contractility, and intercellular tension) and the chemical reactions involved in the RhoA effector signaling pathway. External forces are found to modulate the oscillation intensity of collective cells in the monolayer and to polarize their oscillations along the direction of external tension. The proposed cytodynamic model can recapitulate the prominent features of cell oscillations observed in a variety of experiments, including both isolated cells (e.g., spreading mouse embryonic fibroblasts, migrating amoeboid cells, and suspending 3T3 fibroblasts) and multicellular systems (e.g., Drosophila embryogenesis and oogenesis). © 2017 Elsevier Ltd. All rights reserved.

1. Introduction Oscillations occur in a wealth of biological processes such as circadian rhythms (Dunlap, 1999; Reppert and Weaver, 2002), calcium sparks (Cheng and Lederer, 2008; Jaggar et al., 20 0 0), DNA synthesis (Zhang et al., 20 07), cytokinesis (Grill et al., 2005; Sedzinski et al., 2011), embryogenesis (Blanchard et al., 2010; Solon et al., 2009), and collective cell migration (Cox and Snead, 2016; Serra-Picamal et al., 2012), with a period ranging from a few seconds to dozens of hours. These dynamic behaviors may stem from either an active programmed regulation for certain biological functions such as circadian rhythms (Dunlap, 1999; Reppert and Weaver, 2002), or a passive response to the circumstance such as the oscillation of cyclic adenosine monophosphate in starved Dictyostelium cells (Kamino et al., 2011). Biological oscillations are usually triggered by Hopf bifurcation, beyond which the system loses its stability at equilibrium and enters into a limit cycle (Goldbeter, 2002). Thus, they can be viewed as a non-equilibrated self-organization of the dissipative biological systems (Nicolis and Prigogine, 1977). As an important form of biological oscillations, cell shape or area oscillations play a pivotal role in morphogenesis including tissue elongation (Rauzi et al., 2010), embryo invagination (Martin et al., 2009), and tissue sealing (Blanchard et al., 2010). For example, during Drosophila embryogenesis, cell shape oscillations occur in ventral furrow cells at the onset of ∗

Corresponding authors. E-mail addresses: [email protected] (B. Li), [email protected] (X.-Q. Feng).

https://doi.org/10.1016/j.jmps.2017.09.013 0022-5096/© 2017 Elsevier Ltd. All rights reserved.

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

651

mesoderm invagination (Martin et al., 2009) and in amnioserosa cells during dorsal closure (Solon et al., 2009). Shape oscillations are also found in the follicle cells during stage 9 in the development of Drosophila egg chambers (He et al., 2010; Koride et al., 2014). Besides, in Xenopus embryogenesis, ectodermal and mesodermal cells can exhibit fluctuating contractions during convergence and extension (Kim and Davidson, 2011). In addition to the aforementioned collective oscillations, isolated cells may also display oscillatory features, as observed in some biological systems, e.g., spreading mouse embryonic fibroblast cells (Giannone et al., 2004), locomotive amoeboid cells (Satoh et al., 1985), suspended 3T3 fibroblast cells (Salbreux et al., 2007), and cytokinesis of Hela cells and L929 fibroblasts (Sedzinski et al., 2011). Among them, an isolated amoeboid cell exhibits distinct shape oscillation with a period of 2–10 min when migrating on an adherent glass surface (Satoh et al., 1985); suspended 3T3 fibroblasts can undergo spontaneous shape oscillations with the period ranging from 1 to 5 min, depending on the culture condition (Salbreux et al., 2007). It has been recognized that the dynamics of cytoskeletal systems is of vital importance for both physiology of individual cells (Jülicher and Prost, 1997; Kruse and Jülicher, 2005; Viatchenko-Karpinski et al., 1999) and morphogenesis of multicellular epithelia (He et al., 2015), organs or organisms (Angelini et al., 2010; Gorfinkiel and Blanchard, 2011; Paluch and Heisenberg, 2009; Vedula et al., 2014). Cell oscillations involve biological, chemical, and mechanical mechanisms that couple at the molecular, subcellular, and supracellular scales. Along with shape oscillations, myosin activity within the oscillating cells has been revealed to fluctuate in nearly an antiphase behavior (Blanchard et al., 2010; Dierkes et al., 2014; He et al., 2010; Koride et al., 2014; Sedzinski et al., 2011). Besides, experiments also evidenced that cell area oscillations disappear when actomyosin contraction is inhibited (He et al., 2010; Sedzinski et al., 2011), suggesting that the oscillations are intimately associated with cell contractility generated by the actomyosin cytoskeleton network. The actomyosin cytoskeleton network is a thin layer consisting of crosslinked actin filaments beneath the plasma membrane and links the neighboring cells through adherens junctions (Lecuit and Lenne, 2007; Lecuit et al., 2011; Murrell et al., 2015). The contractility of the actomyosin network is attributed to the activity of myosin II, which is an ATP-dependent motor protein and “walks” along the actin filaments to generate contractile forces (Murrell et al., 2015). Recently, considerable attentions have been paid to the dynamics of cell oscillations. It was found that amnioserosa cells undergo remarkable collective area oscillations with a period of ∼4 min after germband retraction and these oscillations can last for over an hour until the mid-stage of dorsal closure during Drosophila embryogenesis (Blanchard et al., 2010; Solon et al., 2009). A theoretical model involving sinusoidal contractions or area-related contraction forces has been proposed to reproduce the amnioserosa oscillations (Solon et al., 2009). Representing the amnioserosa by 81 connected regular hexagons, Wang et al. (2012) investigated the biomechanical mechanism behind cell oscillations. The contractility of cells was described by springs connecting neighboring vertices and an interactive regulation between the myosin activity and the contraction force was involved. Dierkes et al. (2014) proposed a one-dimensional contractile rod model which involved myosin turnover to capture the characteristics of cell oscillations. Recently, Lin et al. (2017b) established a time-delayed biomechanical model to address collective oscillations in Drosophila amnioserosa. These studies provide mechanistic insights into the collective cell oscillations. To date, however, there is still a lack of a generic model that rationalizes real biochemical signaling pathways and is capable of dissecting cell oscillations in systems with wide varied cell populations ranging from single to multiple cells. In fact, actomyosin-regulated cell contractility may be mediated by such complex signaling pathways as the RhoA effector signaling pathway and calcium pathway (Ahmadzadeh et al., 2017; Shenoy et al., 2016). Insights into the coupling between detailed biochemical signaling pathway and mechanical cues would deepen our understanding of cell shape oscillations. Taking RhoA effector signaling pathway as a representative one, we here develop a chemomechanical model that integrates a clear interactive regulation between myosin activity and cell deformation to address cell oscillations. We aim to providing a generic theoretical framework to delineate oscillatory dynamics observed in vastly diverse cellular systems ranging from isolated cells to cell aggregations. The paper is organized as follows. In Section 2, the chemomechanical cytodynamic model is provided by considering the mechanical and biochemical coupling mechanisms. In Section 3, we investigate the dynamic stability for both isolated cells and collective cells in an epithelial-like monolayer. In Section 4, we examine the characteristics of collective cell oscillations via spectral analysis. In Section 5, we explore the role of mechanical stretches in modulating collective cell oscillations. Finally, the main conclusions are drawn in Section 6.

2. Model 2.1. Collective cells in a confluent monolayer We apply the dynamic vertex model to investigate the collective cell oscillations in a confluent cell monolayer. The vertex model has been extensively used to study epithelial morphodynamics, e.g., wound closure (Nagai and Honda, 2009), tumor invasion (Lin et al., 2017c), and embryo morphogenesis (Osterfield et al., 2013). In the cellular vertex model, cells are approximated by connected polygons with intercellular boundaries linking to neighboring cells (Farhadifar et al., 2007; Lin et al., 2017a), as shown in Fig. 1(a) and (b). Different from the traditional vertex model, we will involve chemomechanical feedback. Within the vertex description, the dynamics of an epithelial-like monolayer is determined by the evolution of the cell vertices, at which three neighboring cells joint together. The total potential energy in the cell monolayer can be expressed

652

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

Fig. 1. Cellular vertex model. (a) Phase contrast image of a confluent Madin–Darby canine kidney (MDCK) cell monolayer, in which cells manifest polygonal shape (see the marked cells). (b) Geometric description of epithelial cell monolayer. Refer to the coordinate system (x, y), where the x axis is along the horizontal direction and the y axis is along the vertical direction. (c) Forces acting at a vertex in a cell monolayer. Fci , Fai , Fsi , and Fri denote forces arising from cell contractility, cell area elasticity, intercellular interfacial tension, and friction, respectively. (d) Schematic diagram of the interactive regulation between RhoA effector signaling pathway and cell deformation.

as (Farhadifar et al., 2007; Fletcher et al., 2014; Nagai and Honda, 2001; Xu et al., 2016b)

U=

1 1  2  (J ) 2 km fMLC LJ + Ka AJ − A0 + li j , 2 2 J

J

(1)

i, j

where the first term on the right-hand side accounts for cell contraction due to active contractility of the actomyosin cortex layer (Farhadifar et al., 2007; Salbreux et al., 2012); the second term corresponds to cell area elasticity resulting from the resistance to monolayer’s height fluctuations and cell apical area changes (Zehnder et al., 2015); the third term refers to the interfacial tension set by a competition between the cortical tension and the intercellular adhesion (Manning et al., 2010). LJ and AJ represent the perimeter and area of the Jth cell, respectively, and A0 is the preferred area. Ka denotes the area stiffness of a cell.  represents the interfacial tension acting on the intercellular interface, and can be either positive (when the cortical tension dominates over the intercellular adhesive energy) or negative (when the intercellular adhesive energy (J ) dominates); lij is the length of intercellular interface ij connecting vertices i and j. f MLC refers to fraction of activated myosin (J )

light chain (MLC), i.e. the phosphorylation level of MLC (pMLC) which dictates the activity of myosin II in cell J. km fMLC

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

653

characterizes the contractility of the actin cortex of the Jth cell, where we have assumed a linear relationship between cell contractility and pMLC. Consequently, km will be referred to as the cell contractility constant hereafter. It is noted that the intercellular adhesion  may also be related to myosin II activity through the complex cytoskeleton network. For example, the intercellular cohesion of epithelium is attributed to the adhesive contacts constituted mainly by E-cadherin molecules, which concentrate in the adherens junctions and form an adhesive belt, stitching cells together (Guillot and Lecuit, 2013). The adhesive belt is mechanically coupled to the cortical actomyosin network via a complex long chain, which includes α -catenin, β -catenin, vinculin, and so on (Lecuit et al., 2011). In this paper, we take  as a constant for simplicity. Since cellular motion is mainly damped by the energy dissipation due to viscosity, the inertia of cells is ignored in the dynamic vertex model (Fletcher et al., 2014). Therefore, the mechanical equilibrium of each vertex requires the balance p between the dissipative frictional force Fri = −ηdri /dt and the potential force Fi = −∂ U /∂ ri arising from cell contraction, cell elasticity, and interfacial tension (Fig. 1(c)), with ri the position of vertex i and η the effective coefficient of friction. Accordingly, the temporal evolution of vertex i is determined by

η

dri ∂U =− . dt ∂ ri

(2)

In the present model, we do not consider the detailed cell–substrate interactions, which may play a key role in cell differentiation, migration, adhesion, polarity, and reorientation (Chen et al., 2015; Engler et al., 2006; He et al., 2014; Qian et al., 2013; Sunyer et al., 2016; Xu et al., 2016a). The friction between the cells and their substrate is taken into account by the dissipative frictional force Fri . Substituting Eq. (1) into (2) leads to

 dr (J ) η i = − km fMLC LJ dt



J∈Ci



1      ri − r j  +  − , Ka AJ − A0 k × r j2 − r j1 −  ri − r j1  ri − r j2  r i − r j  2 J∈Ci j∈Vi ri − r j1

ri − r j2

where k denotes a unit vector normal to the cell monolayer; the summation

vertex i, while the summation

J∈Ci

(3)

computes over all neighboring cells Ci of

is made over all neighboring vertices Vi of vertex i; j1 and j2 are the neighboring vertices

j∈Vi

of vertex i in cell J, as shown in Fig. 1(c). Besides, we here focus on the dynamics of cell shape oscillations, and thus do not consider topological transitions in the polygon network for brevity. To reveal the biochemical mechanism of cell oscillations, we invoke the RhoA effector signaling pathway, which is responsible for cell contractility (Ahmadzadeh et al., 2017; Lessey et al., 2012; Miller and Davidson, 2013; Shenoy et al., 2016). RhoA is a small GTPase protein of Rho family, which is the key regulator of cytoskeleton and thus crucial for morphogenesis (Lecuit and Lenne, 2007; Lecuit et al., 2011). It has been illustrated that RhoA can be activated by tensile forces (Smith et al., 2003; Zhao et al., 2007). The active RhoA (GTP-binding form) acts upon Rho kinase (ROCK) to promote pMLC, which facilitates the assembly of myosin II and thus enhances its ATPase activity. Consequently, the contractile force of actin filaments generated by myosin II is enhanced. Taken together, there exists a feedback mechanism between cell deformation and myosin II activity through the RhoA effector signaling pathway, as shown in Fig. 1(d). Different upstream regulators for myosin II activity may also exist, e.g., Rho1 (Munjal et al., 2015) and Ca2+ (Ahmadzadeh et al., 2017; Shenoy et al., 2016). However, the corresponding regulating processes and outcomes of these upstream regulators for myosin II activity are similar (Ahmadzadeh et al., 2017; Munjal et al., 2015; Shenoy et al., 2016). Therefore, replacing the RhoA effector signaling pathway with another similar signaling pathway (e.g., Rho1 and Ca2+ ) would not interfere with our main conclusions. We employ the Hill function (Ferrell et al., 2011; Mackey and Glass, 1977) to describe the chemomechanical coupling of myosin activation and cell deformation. The kinetics of Rho, ROCK, and MLC activation and deactivation in the Jth cell is quantified as () d fRho



J

dt

= αRho

λL(J ) 

KRho +



() d fROCK = αROCK dt K J

nRho

λL

() fRho J

ROCK

() d fMLC = αMLC dt K J



MLC

 (J ) (J ) 1 − fRho − βRho fRho , n Rho (J ) 

+

() fROCK J



nROCK

(J ) nROCK fRho



(J ) (J ) 1 − fROCK − βROCK fROCK ,

(4)

(5)

nMLC

(J )

+ fROCK

 (J ) (J ) nMLC 1 − fMLC − βMLC fMLC ,

(6)

where f represents the fraction of activated biochemical factor (i.e., activity), α the activation coefficient, β the deactivation rate, K the apparent dissociation constant, and n the Hill coefficient. The subscripts Rho, ROCK, and MLC denote the quantities pertaining to Rho, ROCK, and MLC, respectively. λL = L/Ls characterizes the cell deformation, where Ls is the perimeter of the cell at rest. Because the cells in a monolayer statistically prefer a hexagonal shape (Nagai and Honda, 2001), Ls can be obtained by theoretically solving the equilibrium state of a regular hexagonal cell, which will be shown in Section 2.2. We

654

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

notice the chemical communication among cells is not a necessity for an oscillatory behavior in some systems (FernandezGonzalez and Zallen, 2011; Jayasinghe et al., 2013). For simplicity, therefore, we ignore the transmembrane transport of myosin II between cells. So far, we have developed a chemomechanical model that integrates a feedback mechanism between myosin activity and cell deformation to address cellular dynamics in a monolayer. This model will be used to investigate the cellular morphody namics of both isolated cells and collective interacting cells. We initially normalize Eqs. (3)–(6) by using A0 as the length scale and τ = η/(Ka A0 ) as the time scale. The normalized equations can be written as

 d ri (J ) ˜ =− k˜ m fMLC LJ dt˜



J∈Ci



 1     ri − r j  +  − , A˜ J − 1 k × r j2 − r j1 −   ri − r j1   ri − r j2   ri − r j  2 J∈Ci j∈Vi ri − r j1

ri − r j2

n (J ) L˜J Rho  d fRho (J ) (J ) = α˜ Rho 1 − fRho − β˜Rho fRho , n n dt˜ KRho L˜s Rho + L˜ Rho

(7)

(8)

J



() d fROCK = α˜ ROCK dt˜ K J

() fRho J



ROCK

() d fMLC = α˜ MLC dt˜ K J

nROCK



+ fRho

() fROCK J



MLC

(J ) nROCK



(J ) (J ) 1 − fROCK − β˜ROCK fROCK ,

(9)

nMLC

(J )

+ fROCK

 (J ) (J ) nMLC 1 − fMLC − β˜MLC fMLC ,

(10)

where the dimensionless parameters are defined as

˜ = 3/2 , L˜J = √LJ , r˜i = √ri , t˜ = τt , k˜ m = KkamA0 ,  Ka A0 A0 A0 A L˜s = √Ls , A˜ J = J , α˜ p = αp τ , β˜p = βp τ , A0

(11)

A0

with the subscript p stands for Rho, ROCK, and MLC. Eqs. (7)–(10) establish a dimensionless system controlling the dynamic evolution of cells, and they will be numerically solved by the finite difference method. 2.2. Isolated cell The above model can analyze cell oscillations in a monolayer system consisting of an arbitrary number of cells. To clearly reveal the physical mechanisms for the occurrence of cell oscillation, we first consider an isolated cell, simplified as a regular hexagonal unit. In this case, Eqs. (3)–(6) reduce to

η

√ 1 dR 3 = −6km fMLC R − Ka (A − A0 )R − , dt 2 2

(12)

λL Rho d fRho = αRho (1 − fRho ) − βRho fRho , n dt KRho + λL Rho

(13)

nROCK fRho d fROCK = αROCK nROCK (1 − f ROCK ) − βROCK f ROCK , dt KROCK + fRho

(14)

nMLC fROCK d fMLC = αMLC nMLC (1 − f MLC ) − βMLC f MLC , dt KMLC + fROCK

(15)

n

where R is the radius (i.e. edge length) of the cell. Note that /2 in the interfacial tension term in Eq. (12) refers to the surface tension between the cell and environment for the isolated cell, but rather intercellular surface tension for collective √ interacting cells. Using R0 = 2A0 /(3 3 ) (i.e., the cell radius corresponding to the preferred area A0 ) as the length scale and τ = η/(Ka A0 ) as the time scale, we have the following normalized equations:

√  1 ¯ dR¯ 3 2 = −6k¯ m fMLC R¯ − , R¯ − 1 R¯ −  2 2 dt¯

(16)

d fRho R¯ nRho = α¯ Rho (1 − fRho ) − β¯ Rho fRho , n dt¯ KRho R¯ s Rho + R¯ nRho

(17)

nROCK fRho d fROCK = α¯ ROCK (1 − fROCK ) − β¯ ROCK fROCK , KROCK + f nROCK dt¯

(18)

Rho

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

655

nMLC fROCK d fMLC ¯ = α¯ MLC nMLC (1 − f MLC ) − βMLC f MLC , KMLC + fROCK dt¯

(19)

where the dimensionless parameters are defined as



¯ = 3 3 3/2 = t¯ = τt , R¯ = RR0 , k¯ m = KkamA0 = k˜ m ,  2 Ka A 0 R R¯ s = R0s , α¯ p = αp τ = α˜ p , β¯ p = βp τ = β˜p .

√ 3 3 2

˜, 

(20)

The equilibrium state of the isolated cell satisfies dR¯ /dt¯ = 0 and d fp /dt¯ = 0, which give

 √ √  3 3 3 1 ¯ (s ) − = 0, R¯ s + 6k¯ m fMLC R¯ s +  2 2 2 ( ) fRho = s

(21)

α¯ Rho , ¯ α¯ Rho + βRho (KRho + 1 )

(22)

 (s) nROCK α¯ ROCK fRho  fROCK =  (s) nROCK  (s) nROCK  , α¯ ROCK fRho + β¯ ROCK KROCK + fRho

(23)

 (s) nMLC α¯ MLC fROCK  fMLC =  (s) nMLC  (s) nMLC  . α¯ MLC fROCK + β¯ MLC KMLC + fROCK

(24)

(s )

(s )





Substituting Eqs. (22)–(24) into Eq. (21), we can obtain the cell radius R¯ s at rest, or equivalently, R˜s = Rs / A0 = √ 4 2R¯ s / 27. Thus, the cell perimeter at equilibrium is Ls = 6Rs = 6Rs R0 = 6R˜s A0 , with the dimensionless form L˜s = 6R˜s .

2.3. Estimation of the dimensionless parameters According to experimental measurements, the viscosity is in the order of 104 − 105 N s m−2 for embryonic tissues (Forgacs et al., 1998); and the cell contraction force is in the range of 1 − 10 nN for epithelial cell sheets (Hannezo et al., 2014). Hence, we can estimate the friction coefficient η ∼ 0.01 − 0.1 N s m−1 and the cell contractility constant km ∼ 10−4 − 10−3 N m−1 . Taking the cell areal stiffness Ka ∼ 105 − 107 N m−3 for epithelial tissues (Girard et al., 2007; Solon et al., 2009), and the preferred area A0 = 400 μm2 , we have the scale of the dimensionless cell contractility k˜ m ∼ 0.1, the length scale = A0 = 20 μm and the time scale τ = η/(Ka A0 ) ∼ 100 s. The activation coefficients and deactivation rates of Rho, ROCK, and MLC have the scale ∼ 0.01 s−1 (Koride et al., 2014). Therefore, the corresponding dimensionless activation coefficients and deactivation rates are α˜ p ∼ 1 and β˜p ∼ 1, respectively. Furthermore, we take the apparent dissociation constants Kp ∼ 1. ˜ = 0, α˜ Rho = 2.0, Specifically, the parameter values used in our calculation are set, unless stated otherwise, as: k˜ m = 0.2,  β˜Rho = 0.3, α˜ ROCK = 1.0, β˜ROCK = 0.1, α˜ MLC = 100, β˜MLC = 1.0, KRho = 1.0, nRho = 8, KROCK = 0.8, nROCK = 8, KMLC = 0.8, and nMLC = 8. It should be noted that here we have set a large value of α˜ MLC (= 100) for clear illustration of cell oscillation. In fact, only if α˜ MLC and β˜MLC are in a proper range, can cell oscillations take place, as we will show in Section 3. 3. Hopf bifurcation Experiments have evidenced that the cell shape oscillations that occur during Drosophila germband extension and dorsal closure are autonomous for individual cells, which do not require stable contact with neighboring cells (Fernandez-Gonzalez and Zallen, 2011; Jayasinghe et al., 2013). In this section, we examine the cellular dynamics for both isolated cells and collective interacting cells dictated by chemomechanical coupling. To gain insights into the dynamic mechanisms of cell oscillations, we will first derive the analytical solution for the oscillation of a single cell and then analyze the dynamic behavior of collective cells in a monolayer. 3.1. Oscillation of an isolated cell Introducing a small perturbation to the equilibrium state of an isolated cell, the incremental equations of the dynamic system Eqs. (16)–(19) become

d

  δ R¯ dt¯



(s ) = − 6k¯ m fMLC +

 √  3 2 ¯ 3Rs − 1 δ R¯ − 6k¯ m R¯ s δ fMLC , 2

  d(δ fRho ) n K α¯ (s ) = Rho Rho Rho 1 − fRho δ R¯ − 2 ¯ dt¯ K + 1 R ( Rho ) s



α¯ Rho KRho + 1

+ β¯ Rho

(25)

 δ fRho ,

(26)

656

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666





nROCK −1

( )  nROCK KROCK fRho d(δ fROCK ) (s ) = α¯ ROCK  1 − fROCK δ fRho −  2  ¯ dt (s ) nROCK KROCK + fRho



s

nMLC −1

( )  nMLC KMLC fROCK d(δ fMLC ) (s ) = α¯ MLC  1 − fMLC δ fROCK −  2  ¯ nMLC dt (s ) KMLC + fROCK s



α¯ ROCK

 α¯ MLC

( ) fRho s



( ) KROCK + fRho



( ) fROCK s



nROCK s

nROCK + β¯ ROCK δ fROCK , 

nMLC



( ) KMLC + fROCK s

nMLC

(27)

+ β¯ MLC

δ fMLC ,

(28)

where δ R¯ , δ fRho , δ fROCK , and δ fMLC are the increments of cell radius, Rho activity, ROCK activity, and MLC activity from the equilibrium state, respectively. From the above perturbed Eqs. (25)–(28), the Jacobian (or characteristic) matrix for the system at the equilibrium state is





−bS 0 0 − aS ⎜ cS − dS 0 0 ⎟ J=⎝ , 0 eS − f S 0 ⎠ 0 0 gS − hS where

(29)

√  3 2 (s ) aS = 6k¯ m R¯ s , bS = 6k¯ m fMLC + 3R¯ s − 1 ,   2 n K α¯ α¯ Rho (s ) cS = Rho Rho Rho 1 − f , dS = + β¯ Rho , Rho 2 ¯ K Rho + 1 (KRho + 1 ) Rs



eS = α¯ ROCK

( ) nROCK KROCK fRho





nROCK −1

 (s ) 2 1 − fROCK , n (s ) ROCK s

KROCK + fRho



fS = α¯ ROCK

( ) fRho s

nROCK

nROCK + β¯ ROCK , nMLC −1 (s )  nMLC KMLC fROCK (s ) gS = α¯ MLC   (s) nMLC 2 1 − fMLC ,  

(30)

( ) KROCK + fRho s

KMLC + fROCK



hS = α¯ MLC

( ) fROCK s



nMLC

( ) KMLC + fROCK s

nMLC + β¯ MLC .

The dynamics of the isolated cell is characterized by the eigenvalue of the corresponding Jacobian matrix, ζ , which satisfies

det (ζ I − J ) = 0,

(31)

where I is the identity matrix. Substituting Eq. (29) into (31), we obtain the characteristic equation for the chemomechanical coupling dynamics of an isolated cell

( ζ + b S ) ( ζ + dS ) ( ζ + f S ) ( ζ + hS ) + aS cS e S gS = 0 .

(32)

Setting ζ = r + iω, Eq. (32) becomes

( r + bS ) ( r + dS ) ( r + f S ) ( r + hS ) − [ ( r + bS ) ( r + dS ) + ( r + bS ) ( r + f S ) + ( r + bS ) ( r + hS ) + (r + dS )(r + fS ) + (r + dS )(r + hS ) + (r + fS )(r + hS )]ω2 + ω4 + aS cS eS gS = 0,

(33)

[ ( r + bS ) ( r + dS ) ( r + f S ) + ( r + bS ) ( r + dS ) ( r + hS ) + ( r + bS ) ( r + f S ) ( r + hS ) + (r + dS )(r + fS )(r + hS )]ω − (4r + bS + dS + fS + hS )ω3 = 0,

(34)

where r and ω are real. Solving Eqs. (33) and (34) numerically, we can obtain the eigenvalue ζ that determines the stability of the isolated cell. Depending on ζ , three states can be distinguished: (i) If Re(ζ ) < 0, the equilibrium of the cell is stable and cannot be broken up by perturbations, which will be referred to as a stable state. (ii) If Re(ζ ) > 0 and Im(ζ ) = 0, the equilibrium of the cell is unstable. In this case, the equilibrium state may be broken due to infinitesimal perturbations, and then the system will evolve into a limit cycle of oscillation around the equilibrium state in the phase space. Such a state will be referred to as an oscillatory one hereafter. (iii) If Re(ζ ) > 0 and Im(ζ ) = 0, the cell will collapse, referred to as the collapsed state.

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

657

Fig. 2. Oscillation of an isolated cell. (a) Temporal evolution of cell radius R, Rho activity fRho , ROCK activity fROCK and MLC activity fMLC in an oscillating cell. (b) Projection on the R − f MLC plane of trajectory of an oscillating cell, with the red arrows indicating the moving direction. The green dash line corresponds to the intersection of the nullclines of fRho , fROCK , and fMLC ; whereas the blue dash line refers to the intersection of the nullclines of R, fRho , and fROCK . The black arrows indicate the direction of motion of trajectories as they cross the nullclines. The gray spot denotes the unstable equilibrium state. Here, the nullcline is the locus of points where the corresponding rate is zero. For instance, the nullcline of R is controlled by the equation dR/dt = 0. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

There exists a critical state, upon which a Hopf bifurcation happens and the system could undergo a transition from the stable state to the oscillatory state. In particular, a Hopf bifurcation occurs when the eigenvalue ζ becomes pure imaginary, i.e. Re(ζ ) = 0 and Im(ζ ) = 0. Thus, the critical parameters for such a stability transition satisfy

ω4 + [bS dS + bS fS + bS hS + dS fS + dS hS + fS hS ]ω2 + (aS cS eS gS + bS dS fS hS ) = 0,

(35)

(bS + dS + fS + hS )ω2 − (bS dS fS + bS dS hS + bS fS hS + dS fS hS ) = 0,

(36)

where ω > 0 is the dimensionless angular frequency at the occurrence of Hopf bifurcation in the dynamic system. From Eq. (36), the corresponding angular frequency ωH is given as



ωH =

bS dS f S + bS dS hS + bS f S hS + dS f S hS . bS + dS + f S + hS

(37)

Thus, the corresponding oscillation period is

TH =



ωH



= 2π

bS + dS + f S + hS . bS dS f S + bS dS hS + bS f S hS + dS f S hS

(38)

It is noted that the critical angular frequency ωH in Eq. (37) is deduced from Eq. (36), and thus the parameters bS , dS , fS , and hS should satisfy relations in Eqs. (35) and (36). We also perform direct simulations of Eqs. (16)–(19) to investigate the dynamic behavior of an isolated cell. The numerical results demonstrate that the proposed chemomechanical model can produce spontaneous cell oscillation. Fig. 2(a) gives the temporal evolution of the cell radius R, Rho activity fRho , ROCK activity fROCK and MLC activity fMLC of the cell, displaying apparent oscillations. The trajectory of (R, fRho , fROCK , fMLC ) is a limit cycle, which is a closed curve in a four-dimensional space. The projection of the trajectory on the R − fMLC plane is shown in Fig. 2(b). Thus, the cell oscillation reported here is autonomous, in agreement with experimental observations on, for example, germband extension (Fernandez-Gonzalez and Zallen, 2011), dorsal closure (Jayasinghe et al., 2013), and follicle cell oscillation (He et al., 2010) in Drosophila embryogenesis or oogenesis. Note that this projection moves clockwise (indicated by red arrows). The encapsulated area by the projection corresponds to the energy dissipation per cycle, replenished by the chemical reactions of RhoA pathway. Besides, the three different states (i–iii) can also be identified through numerical simulations, as shown in Fig. 3. Here, the collapsed state also includes the case where there exists no equilibrium state for the isolated cell, i.e., the equilibrium Eq. (21) has no positive solution of R¯ s . Now we perform a parametric study to capture the predominant features of cell oscillation. The characteristic equation (32) is solved to obtain the eigenvalue ζ . Comparing the theoretical and numerical results, we find that the above linear stability analysis well delineates the dynamic features of an isolated cell. In what follows, we will discuss the influence of cellular mechanical properties and the chemical reactions of the RhoA pathway on the cell dynamics. Specifically, we concentrate on the roles of cell contraction, surface tension, and the activation coefficient and deactivation rate of MLC. Firstly, we scrutinize how cell contraction and surface tension dictate the dynamic behavior of a single cell. A phase diagram is constructed by using both the linear stability analysis and numerical simulations, as shown in Fig. 4(a). It is seen (cr ) that when the cell contractility constant k˜ m is small (i.e., k˜ m < k˜ m ), the eigenvalue given by the characteristic equation

658

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

Fig. 3. Dynamics of an isolated cell. (a) Evolution of cell radius R corresponding to three different states, i.e. stable, oscillatory, and collapsed. (b) Projection on the R − f MLC plane of trajectories of the three states shown in (a), with the black arrows indicating the moving directions.

Fig. 4. Dynamics of an isolated cell regulated by cell’s mechanical properties. (a) The phase diagram regulated by the cell contractility constant k˜ m and ˜ . In the oscillatory region, the heat map corresponds to the oscillatory amplitude of cell radius RD , obtained from spectral analysis. The interfacial tension  white solid line indicates the critical parameters for a Hopf bifurcation, whereas the white dashed line corresponds to a transition from oscillatory state to collapsed state. (b) The eigenvalue ζ of an isolated cell system versus the cell contractility constant k˜ m . Here, the four eigenvalues are distinguished by different colors, and Re(ζ ) and Im(ζ ) are plotted as solid and dashed lines, respectively. The magenta point marks the Hopf bifurcation.

(32) has negative real parts (Fig. 4(b)) and thus the cell is arrested in its stable equilibrium state (Fig. 4(a)). When the (cr ) cell contractility constant k˜ m is large enough (i.e., k˜ m > k˜ m ), the eigenvalue of the dynamic system could have positive real parts (Fig. 4(b)) and spontaneous oscillation may occur (Fig. 4(a)). It suggests that sufficiently strong cell contractility is necessitated to trigger sustained oscillation. However, when the cell contractility constant k˜ m is too large, the cell will ˜ is below a critical value collapse. Besides, Fig. 4(a) reveals that surface tension influences the cellular dynamics. When  ˜ exceeds (the white solid line in Fig. 4(a)), the originally oscillatory cell may transit to a stable state. In addition, when  a second critical value (the white dashed line in Fig. 4(a)), the cell will collapse because of strong contraction. Therefore, ˜ ) tends to stabilize the cell. enhancing cell–substrate adhesion (i.e., decreasing  We next examine how the chemical reactions of the RhoA effector signaling pathway regulate the cell dynamics by identifying the roles of the activation coefficient α˜ MLC and the deactivation rate β˜MLC of MLC. Fig. 5(a) provides a phase diagram of (α˜ MLC , β˜MLC ). The coefficient of MLC activation, α˜ MLC , influences the cellular dynamics in the following manner. (cr ) When α˜ MLC is too small (α˜ MLC < α˜ MLC ), the eigenvalue of the chemomechanical coupling system has negative real parts, (cr ) and thus the cell can stay in a stable equilibrium state (Fig. 5(b)). When α˜ MLC is large enough (α˜ MLC > α˜ MLC ), the real parts of the eigenvalue could become positive, and the cell will undergo spontaneous oscillation (Fig. 5(a)). When α˜ MLC is too large, however, the cell will collapse. Moreover, Fig. 5(a) also reveals that an enhanced MLC deactivation rate β˜MLC tends to stabilize the cell dynamics, since cell contractility is reduced due to down-regulated MLC activity. A small β˜MLC favors spontaneous oscillations, whereas the cell will collapse when β˜MLC is too small (Fig. 5(a)). In addition, we find that the physical parameters involved in the chemomechanical coupling can also tailor the oscillatory period within the range from 6τ to 30τ (Fig. 6), i.e. 10 min −50 min, comparable to experimental observations in a rich variety of biological systems (Giannone et al., 2004; Salbreux et al., 2007; Satoh et al., 1985; Sedzinski et al., 2011).

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

659

Fig. 5. Dynamics of an isolated cell regulated by the chemical reaction of MLC. (a) The phase diagram regulated by the MLC activation coefficient α˜ MLC and MLC deactivation rate β˜MLC . (b) The eigenvalues ζ of an isolated cell system versus the MLC activation coefficient α˜ MLC .

˜ ) and (b) biochemical reactions (α˜ MLC , β˜MLC ). In the oscilFig. 6. Oscillation period of an isolated cell. Regulations of (a) cell mechanical properties (k˜ m ,  latory region, the heat map corresponds to the oscillatory period T, obtained from spectral analysis. The white solid line indicates the critical parameters for a Hopf bifurcation, whereas the white dashed line corresponds to a transition from oscillatory state to collapsed state.

3.2. Collective cell oscillations in a monolayer Now we examine the collective dynamics of a cell monolayer through numerical computations of Eqs. (7)–(10). Consider a cell monolayer consisting of plenty of cells (∼10,0 0 0). Periodic boundary conditions are adopted to eliminate the influence of boundary constraints. In order to eliminate the influence of stresses or stretches, we perform simulations based on the stress-free configuration of the cell monolayer. We now analyze how cell contraction and intercellular surface tension influence the collective dynamics of a cell mono˜ ) is established in Fig. 7(a). Similar to the dynamics of an isolated cell, there exist three diflayer. A phase diagram of (k˜ m ,  ferent dynamic states for collective cells, i.e., the stable state, the oscillatory state, and the collapsed state. For the collapsed state of a cell monolayer, there exists no stress-free configuration due to strong cell contraction or intercellular tension. ˜ are within a proper range. Collective cell oscillations occur when the cell contractility constant k˜ m and surface tension  Specifically, a small k˜ m renders a stable state, whereas sufficiently large k˜ m may trigger chemomechanical Hopf bifurcation, as shown in Fig. 7(b). Above the Hopf bifurcation, the cell monolayer undergoes sustained collective shape oscillations. When k˜ m is too large, however, the cells in the monolayer may collapse. Besides, the intercellular adhesion is found to stabilize the ˜ < 0, i.e., cell–cell adhesion dominates over cell cortex tension, strong enough intercellular adhesion cell dynamics: when  ˜ > 0, a may bring collective cell oscillations to a stable state that all cells in the monolayer are arrested; whereas when  large intercellular surface tension may result in collapse (Fig. 7(a)). The roles of the activation coefficient α˜ MLC and the deactivation rate β˜MLC of MLC in the collective cell dynamics are illustrated in Fig. 8(a). It demonstrates that small α˜ MLC corresponds to a stable state, and only when α˜ MLC is large enough that collective cell oscillations could take place, as a consequence of Hopf bifurcation (Fig. 8(b)). When α˜ MLC is too large, however, cell elasticity cannot resist so high cell contraction and the cells will collapse.

660

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

Fig. 7. Collective dynamics of a cell monolayer regulated by cell’s mechanical properties. (a) A phase diagram regulated by the cell contractility constant ˜ . AD = A(J ) J is the averaged areal oscillatory amplitude of the cell monolayer, with A(J ) being the areal oscillatory amplitude k˜ m and interfacial tension  D D of the Jth cell, obtained from spectral analysis. (b) The bifurcation diagram of the cell dynamics in a monolayer dictated by the cell contractility constant ˜km . Here, Amax = maxt [AJ (t )]J and Amin = mint [AJ (t )]J .

Fig. 8. Collective dynamics of a cell monolayer regulated by the chemical reaction of MLC. (a) The phase diagram regulated by the MLC activation coefficient

α˜ MLC and MLC deactivation rate β˜MLC . (b) The bifurcation diagram of the cell dynamics in a monolayer dictated by the MLC activation coefficient α˜ MLC .

4. Dynamic features of collective cell oscillations 4.1. Collective oscillatory morphodynamics In this section, we investigate the collective oscillatory dynamics in a cell monolayer. Fig. 9(a) illustrates the morphological evolution of an arbitrarily selected cell (denoted by C0) and its six neighbors (denoted by C1–C6). It shows the shape changes of the connected cells with time. The temporal evolution of the area, RhoA activity, ROCK activity and MLC activity of the selected cell C0 are plotted in Fig. 9(b), displaying apparent oscillations. During this dynamic process, oscillations are transmitted and synchronized through physical linkages between cells. Fig. 9(c) shows the collective areal oscillations of a small ensemble consisting of those seven cells (C0–C6). Spectral analysis is performed to examine the collective dynamic behavior of the whole monolayer. It can be seen from Fig. 10(a) that almost all cells oscillate with the same frequency of f ∼ 0.12τ −1 . Thus, the oscillation period is T = 8τ ∼ 13 min, comparable to experimental measurements (Blanchard et al., 2010; Sokolow et al., 2012; Solon et al., 2009). We calculate the oscillation phases of cell area ϕ A , RhoA activity ϕ Rho , ROCK activity ϕ ROCK , and MLC activity ϕ MLC for all cells by using the discrete Fourier transform method. Fig. 10(b) shows that almost all values of ϕ A and ϕ MLC are close to the two lines ϕA − ϕMLC = ±180o , indicating that the dynamics of cell area and pMLC (i.e., myosin activity) exhibits approximately an antiphase feature in the whole domain. The distribution of ϕA − ϕMLC shown in Fig. 10(c) further verifies the antiphase feature between ϕ A and ϕ MLC in the monolayer. As stated in the oscillation of an isolated cell, the antiphase feature arises (J ) from the negative feedback between myosin activity and cell deformation (Fig. 1(d)): cell expansion (λL > 1) promotes myosin activation, while elevated MLC activity (i.e., myosin activity) hinders cell expansion, as described in Eqs. (3)–(6). Such antiphase oscillatory dynamic behaviors of cell area and myosin activity have been observed in experiments (Blanchard et al., 2010; Dierkes et al., 2014; Martin et al., 2009; Solon et al., 2009). In addition, we also examine the oscillatory phases of RhoA activity ϕ Rho and ROCK activity ϕ ROCK . It demonstrates that the dynamics of cell area and RhoA activity are nearly

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

661

Fig. 9. Oscillating dynamics of collective cells. (a) Temporal evolution of cellular morphology: a local perspective, focusing on a typical cell (C0) and its neighbors (C1–C6). (b) Evolution of the area A, Rho activity fRho , ROCK activity fROCK , and MLC activity fMLC in an typical oscillating cell, i.e., the cell marked as C0 in (a). (c) Evolution of cell area of a typical oscillating cell C0 and its neighbors C1–C6, as marked in (a).

Fig. 10. Spectral analysis of collective cell oscillations. (a) Spectrograms of area and pMLC of the cells. (b) Scatter diagram of the oscillation phases ϕ MLC and ϕ A of the cells. (c) Distribution of phase differences ϕA − ϕRho , ϕA − ϕROCK , and ϕA − ϕMLC .

inphase due to the direct activation of RhoA from cell deformation, whereas ϕ A and ϕ ROCK have an phase difference of about 100o (Fig. 10(c)). 4.2. Supracellular oscillatory pattern To elucidate the spatial features of collective cell oscillations, we make a heat map of oscillation phases of cell area

ϕ A (Fig. 11(a)). It shows that cells with parallel contraction and expansion phases are orchestrated into random distributed

662

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

Fig. 11. Supracellular pattern of collective cell oscillations. (a) Spatial distribution of the areal oscillation phases ϕ A of the cells. (b) Histogram of the (J ) (J ) . (c) Spatial correlation of the oscillatory pattern shown in (c), with dc being the mean and the oscillatory orientation θaniso alignment orientation θalign distance between neighboring cells.

supracellular strings. Similar features are also observed in the spatial distributions of ϕ Rho , ϕ ROCK , and ϕ MLC . The oscillatory pattern appears to be labyrinthine, reminiscent of the patterns generated in growing soft tissues (Li et al., 2011). To examine (J ) the cell alignment during this phase pattern formation, we define the alignment orientation θalign of cell J as

     rJ − rI |J · ex    θalign = arccos , rJ − rI|J  (J )

(39)

where rJ is the geometric center of cell J; rI|J is the center of the neighboring cell I whose oscillation phase is the closest (J )

to cell J. Fig. 11(b) shows that the alignment orientation θalign distributes uniformly within the range (0o , 90o ), which is indicative of random oscillations. We further define the oscillatory orientation of cell J as



(J )

θaniso = arctan (J )

(J )

y(J ) x(J )

 ,

(40)

where x and y are the projective components of the area oscillation amplitude of cell J on the x and y axes, respectively. It is found that the distribution of oscillatory orientations is peaked and symmetric around 45o , suggesting that the oscillations in the whole tissue are isotropic (Fig. 11(b)). The supracellular oscillatory pattern shown in Fig. 11(a) manifests an intrinsic spatial correlation of collective cell oscillations. To identify the spatial correlation (i.e., wavelength) in the supracellular string-like patterns, we define the spatial

S.-Z. Lin et al. / Journal of the Mechanics and Physics of Solids 112 (2018) 650–666

663

Fig. 12. Mechanical stretches mediate the collective cellular dynamics in an oscillating cell monolayer. (a) The averaged areal oscillation amplitude AD versus mechanical stretches. Here, four kinds of stretches are examined, including uniaxial stretch (λy = 1), equiaxial stretch (λx = λy ), isometric stretch (λx λy = 1), and non-equiaxial stretch (λy = (λx − 1 )/2 + 1). (b) Dynamic evolution of area, Rho activity, ROCK activity and MLC activity of an arbitrary cell (J ) for different stretches. (d) Comparison in the monolayer sustained an increasing stretch. (c) Comparison of the distribution of oscillatory orientation θaniso (J ) of the distribution of oscillatory alignment θalign for different stretches. (e) The mean oscillatory orientation θ aniso and the mean oscillatory alignment θ align (J ) (J ) J and θalign = θalign J . versus stretch λx , with λy = 1/λx . Here, θaniso = θaniso

correlation function of oscillatory phases as



Cϕ (r ) = cos



ϕA(I ) − ϕA(J )



|

|

r −r < rI −rJ