This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Environment-stored memory in active nematics and extra-cellular matrix remodeling

Ram M. Adar1 and Jean-François Joanny2,3,4 [email protected] 11 Department of Physics,Technion – Israel Institute of Technology, Haifa 32000,Israel
2 Collège de France, 11 place Marcelin Berthelot, 75005 Paris, France
3 Laboratoire Physico-Chimie Curie, Institut Curie, Centre de Recherche, Paris Sciences et Lettres Research University, Centre National de la Recherche Scientifique, 75005 Paris, France
4 Université Pierre et Marie Curie, Sorbonne Universités, 75248 Paris, France
Abstract

Many active systems display nematic order, while interacting with their environment. In this Letter, we show theoretically how environment-stored memory acts an effective external field that aligns active nematics. The coupling to the environment leads to substantial modifications of the known phase diagram and dynamics of active nematics, including nematic order at arbitrarily low densities and arrested domain coarsening. We are motivated mainly by cells that remodel fibers in their extra-cellular matrix (ECM), while being directed by the fibers during migration. Our predictions indicate that remodeling promotes cellular and ECM alignment, and possibly limits the range of ordered ECM domains, in accordance with recent experiments.

Active nematics are systems composed of self-propelling constituents capable of aligning along a shared axis with no preferred overall direction . The active isotropic-nematic transition has been studied extensively [1, 2, 3, 4]. Similar to passive liquid crystals, order is driven by strong aligning fields, obtained by a combination of strong interactions and high densities. Unlike passive systems, activity couples order with propulsion and allows for coexistence between a dilute isotropic phase and dense nematic phase.

Active nematics are ubiquitous in biological systems at different scales. Our main motivation is cells in extra-cellular matrix (ECM), which are both capable of displaying nematic order. Growing biological evidence suggests that the interplay between cellular and ECM order is essential for tissue patterning and multicellular migration [5, 6, 7, 8, 9]. In particular, aligned collagen structures have been shown to greatly promote metastasis [10, 11].

Cell-ECM coupling is especially evident in fibroblasts that deposit, degrade and re-arrange ECM fibers [12, 13]. This has been modeled in different contexts, including wound healing [14], fibroblast alignment [15], and ECM patterning [16, 8]. However, the macroscopic physical mechanisms underlying cell-environment interplay and their role in determining orientational order and dynamics are not well understood or quantified.

Our approach to understand cell-environment interplay is to consider them as a two-component active system. We recently applied such a description to explain mechanical feedback mechanisms between cells and ECM [17, 18]. Here we focus on chemical remodeling. We find that environment-stored memory acts as an external field that allows for steady-state nematic order at arbitrarily low densities and constrains angular dynamics. We relate our results to recent in-vitro experiments on fibroblasts [8, 9]. While we are motivated by cells in ECM, our findings are generic and imply that the understanding of standard active matter may not apply in a dynamic environment, highlighting the need for further investigation and adaptation of existing theories .

Theory. We consider active cells and passive environment (matrix) segments in two dimensions, each described by their position and orientation, 𝒓\boldsymbol{r} and 𝒏\boldsymbol{n} for the cells and 𝒓\boldsymbol{r}^{\prime} and 𝒏\boldsymbol{n}^{\prime} for the matrix. Cells self-propel with a velocity 𝒗=v𝒏\boldsymbol{v}=v\boldsymbol{n} and diffuse with a diffusion coefficient DD. They also align with neighboring cells and matrix segments. Matrix segments are considered to be apolar. They are enslaved to the cells that may deposit and degrade them (for more general choices, see SM in [19]) . These dynamics are described by the following equations:

tfc\displaystyle\partial_{t}f_{\mathrm{c}} =(fcv𝒏)+D2fckfc+kρceEc/Zc\displaystyle=-\nabla\cdot\left(f_{\mathrm{c}}v\boldsymbol{n}\right)+D\nabla^{2}f_{\mathrm{c}}-kf_{\mathrm{c}}+k\rho_{\mathrm{c}}\mathrm{e}^{-E_{\mathrm{c}}}/Z_{\mathrm{c}}
tfm\displaystyle\partial_{t}f_{\mathrm{m}} =k+2[fc(𝒓,𝒏)+fc(𝒓,𝒏)]kρcfm,\displaystyle=\frac{k_{+}}{2}\left[f_{\mathrm{c}}\left(\boldsymbol{r}^{\prime},\boldsymbol{n}^{\prime}\right)+f_{\mathrm{c}}\left(\boldsymbol{r}^{\prime},-\boldsymbol{n}^{\prime}\right)\right]-k_{-}\rho_{\mathrm{c}}f_{\mathrm{m}}, (1)

where t\partial_{t} denotes the partial time derivative. The function fcf_{\mathrm{c}} (fmf_{\mathrm{m}}) describes the distribution to find a cell (matrix segment) at position 𝒓\boldsymbol{r} (𝒓\boldsymbol{r}^{\prime}) with orientation 𝒏\boldsymbol{n} (𝒏\boldsymbol{n}^{\prime}). They are normalized such that d𝒏fc=ρc\int\mathrm{d}\boldsymbol{n}f_{\mathrm{c}}=\rho_{\mathrm{c}} is the cellular density and d𝒏fm=ρm\int\mathrm{d}\boldsymbol{n}^{\prime}f_{\mathrm{m}}=\rho_{\mathrm{m}} is the matrix density.

The cellular orientation dynamics are written in terms of a tumbling rate kk, and an orientation probability, given by the Boltzmann factor exp(Ec)/Zc\exp\left(-E_{\mathrm{c}}\right)/Z_{\mathrm{c}} with the effective alignment energy EcE_{\mathrm{c}} and partition function Zc=d𝒏exp(Ec)Z_{\mathrm{c}}=\int\mathrm{d}\boldsymbol{n}\exp\left(-E_{\mathrm{c}}\right) [20]. This is a convenient choice that allows for the recovery of passive systems in simple limits.

Matrix deposition and degradation are described by the rates k+ρck_{+}\rho_{\mathrm{c}} and kρck_{-}\rho_{\mathrm{c}}, respectively. Here we assume that cells locally deposit segments along their axis of motion and degrade segments in all orientations. Similar ingredients of cell and matrix dynamics were recently proposed as part of a two-layer Viscek model [8]. We note that Eq. (Environment-stored memory in active nematics and extra-cellular matrix remodeling ) is written within mean field.

Averaging the different moments of the orientation angles yield mesoscopic fields that are the focus of our theory. The active cellular current density is given by 𝒋=vd𝒏fc\boldsymbol{j}=v\int\mathrm{d}\boldsymbol{n}f_{\mathrm{c}}, the cellular nematic tensor density is 𝑸c=d𝒏(𝒏𝒏𝑰/2)fc\boldsymbol{Q}_{\mathrm{c}}=\int\mathrm{d}\boldsymbol{n}\,\left(\boldsymbol{n}\boldsymbol{n}-\boldsymbol{I}/2\right)f_{\mathrm{c}}, and the matrix nematic tensor density is 𝑸m=d𝒏(𝒏𝒏𝑰/2)fm.\boldsymbol{Q}_{\mathrm{m}}=\int\mathrm{d}\boldsymbol{n}^{\prime}\,\left(\boldsymbol{n}^{\prime}\boldsymbol{n}^{\prime}-\boldsymbol{I}/2\right)f_{\mathrm{m}}. These fields are all extensive in the number of cells/matrix segments.

We coarse-grain Eq. (Environment-stored memory in active nematics and extra-cellular matrix remodeling ) into equations in terms of the average fields, using an approximation that neglects higher moments of fcf_{\mathrm{c}} in 𝒏\boldsymbol{n} beyond the nematic tensor . In particular, we treat the orientation within mean field in terms of the interaction Ec(𝒏)=2Tr[(𝒏𝒏𝑰/2)𝑸t]E_{\mathrm{c}}\left(\boldsymbol{n}\right)=-2\text{Tr}\left[\left(\boldsymbol{n}\boldsymbol{n}-\boldsymbol{I}/2\right)\boldsymbol{Q}_{t}\right] with the total aligning field 𝑸t=βc𝑸c+βm𝑸m\boldsymbol{Q}_{t}=\beta_{\mathrm{c}}\boldsymbol{Q}_{\mathrm{c}}+\beta_{\mathrm{m}}\boldsymbol{Q}_{\mathrm{m}}. It includes cell-cell and cell-matrix alignment, with the interaction strengths βc\beta_{\mathrm{c}} and βm\beta_{\mathrm{m}}, respectively (more general choices are given in the SM [19]). In the absence of cell activity and cell-matrix interaction, our choice of EcE_{\mathrm{c}} leads to an equivalent of Maier-Saupe theory [21] for compressible two-dimensional systems.

The resulting field equations are [19]:

tρc\displaystyle\partial_{t}\rho_{\mathrm{c}} =D2ρc𝒋,\displaystyle=D\nabla^{2}\rho_{\mathrm{c}}-\nabla\cdot\boldsymbol{j},
t𝒋\displaystyle\partial_{t}\boldsymbol{j} =D2𝒋v2ρc/2v2𝑸ck𝒋,\displaystyle=D\nabla^{2}\boldsymbol{j}-v^{2}\nabla\rho_{\mathrm{c}}/2-v^{2}\nabla\cdot\boldsymbol{Q}_{\mathrm{c}}-k\boldsymbol{j},
t𝑸c\displaystyle\partial_{t}\boldsymbol{Q}_{\mathrm{c}} =D2𝑸c(𝒋+𝒋T𝒋𝑰)/4\displaystyle=D\nabla^{2}\boldsymbol{Q}_{\mathrm{c}}-\left(\nabla\boldsymbol{j}+\nabla\boldsymbol{j}^{\mathrm{T}}-\nabla\cdot\boldsymbol{j}\boldsymbol{I}\right)/4
k𝑸c+kρcg(Qt)𝑸t/Qt,\displaystyle-k\,\boldsymbol{Q}_{\mathrm{c}}+k\rho_{\mathrm{c}}\,g\left(Q_{t}\right)\boldsymbol{Q}_{t}/Q_{t},
tρm\displaystyle\partial_{t}\rho_{\mathrm{m}} =ρc(k+kρm),\displaystyle=\rho_{\mathrm{c}}\left(k_{+}-k_{-}\rho_{\mathrm{m}}\right),
t𝑸m\displaystyle\partial_{t}\boldsymbol{Q}_{\mathrm{m}} =k+𝑸ckρc𝑸m.\displaystyle=k_{+}\boldsymbol{Q}_{\mathrm{c}}-k_{-}\rho_{\mathrm{c}}\boldsymbol{Q}_{\mathrm{m}}. (2)

The first equation is the cellular continuity equation, given by the active cellular current 𝒋\boldsymbol{j} and passive diffusive current. The second equation is a polarization-rate equation for the active current, which we interpret below, at steady state, as a force balance equation.

The equation for 𝑸c\boldsymbol{Q}_{\mathrm{c}} includes diffusion and shear-alignment (first line), as well as non-linear alignment terms that dominate at large lengthscales (second line). They are written in terms of the function g(x)=I1(x)/I0(x),g(x)=I_{1}(x)/I_{0}(x), where In(x)I_{n}(x) is the modified Bessel function of the first kind [22], which results from an angular average of the Boltzmann factor exp(Ec)\exp\left(-E_{\mathrm{c}}\right). The cellular dynamics include the first and second moments of the angular distribution (𝒋\boldsymbol{j} and 𝑸c\boldsymbol{Q}_{\mathrm{c}}, respectively), similarly to “self-propelled rods” [23, 24, 25]. Finally, the matrix dynamics are governed by cellular deposition and degradation.

Refer to caption
Figure 1: (Color online) Heuristic description of cell-matrix feedback. Left panel: Cells (green) align with matrix segments (purple). Right panel: cells degrade exisitng segments (dashed black) and deposit new segments (bold black). The feedback between these processes drives the phenomena in our theory.

These equations define our framework for active nematics (cells) with environment-stored memory (matrix nematic order) , which we apply for the study of ECM remodeling. Cell-matrix interplay enters the theory in two ways: cellular alignment by the matrix as part of the nematic tensor 𝑸t\boldsymbol{Q}_{t} and matrix remodeling by the cells (see Fig. 1). Cellular activity enters our theory in the active current 𝒋\boldsymbol{j}, matrix deposition and degradation, and possibly in the alignment dynamics.

Next, we focus on the consequences of remodeling on the emergence of cellular and ECM orientational order at steady state as well as typical relaxation dynamics of the cell and matrix. For brevity, we rescale times with the run time 1/k1/k and lengths with the typical cellular persistence length v/kv/k, while keeping the same notation.

Results. The standard isotropic-nematic transition in active systems is similar to a gas-liquid transition [1, 26], where the alignment strength plays the role of inverse temperature. At low densities and high temperatures, the system forms a dilute isotropic gas, while at high densities and low temperatures - a nematic liquid. At intermediate densities and temperatures, the two phases co-exist and are generally linearly unstable. Here, we show how the matrix can break this behavior.

The key to understand the coexistence lies in the stress. In the hydrodynamic limit of large system size and long time, the total cellular current is proportional to a divergence of a tensor that we interpret as the stress  [19], 𝝈=[ρc𝑰+2𝑸c/(1+2D)]\boldsymbol{\sigma}=-\left[{\color[rgb]{0,0,0}\rho_{\mathrm{c}}}\boldsymbol{I}+2\boldsymbol{Q}_{\mathrm{c}}/\left(1+2D\right)\right]. The steady-state behavior of the cells is thus described by a constant stress tensor. We consider a possible density profile along the xx direction and focus on the xxxx component of the stress that we denote as σ\sigma for brevity,

σ=σxx=(ρc+Qc1+2D).\displaystyle\sigma=\sigma_{xx}=-\left(\rho_{\mathrm{c}}+\frac{Q_{\mathrm{c}}}{1+2D}\right). (3)

The first term is the ideal-gas contribution to the pressure, while the second term is an extensile active stress 𝑸c\sim\boldsymbol{Q}_{\mathrm{c}} [27]. Here we consider ordering either along the xx axis (Qc>0Q_{\mathrm{c}}>0) or the yy axis (Qc<0Q_{\mathrm{c}}<0).

Co-existence is possible when the active stress decreases with density, compensating for the increase in ideal-gas pressure. This is the case for alignment in the yy direction. The stress σ\sigma can be considered as a Lagrange multiplier that enforces the total number of cells. It is given by (minus) the density in the isotropic phase.

Next, we derive the isotropic-nematic phase diagram in the density-temperature plane, where βc,βm1/T\beta_{\mathrm{c}},\beta_{\mathrm{m}}\sim 1/T, and the ratio βc/βm\beta_{\mathrm{c}}/\beta_{\mathrm{m}} is kept fixed . Examples of such phase diagrams with and without a matrix (ECM) are given in Figs. 2a and 2b. The region of co-existence is delimited by the binodal line (solid blue line), within which lies a region of linear instability, delimited by the spinodal line (dashed red line).

Steady-state nematic order: matrix aligns cells at arbitrarily low densities. We solve Eq. (Environment-stored memory in active nematics and extra-cellular matrix remodeling ) at steady state. The matrix density is ρm=k+/k\rho_{\mathrm{m}}=k_{+}/k_{-}, independent of ρc\rho_{\mathrm{c}}. The matrix nematic tensor has the same direction as the cellular one, chosen here as the xx-axis. We define the intensive nematic order of the cells and matrix, qc=Qc/ρcq_{\mathrm{c}}=Q_{\mathrm{c}}/\rho_{\mathrm{c}} and qm=Qm/ρmq_{\mathrm{m}}=Q_{\mathrm{m}}/\rho_{\mathrm{m}}, and find that qm=qcq_{\mathrm{m}}=q_{\mathrm{c}} at steady state.

The matrix thus inherits the same intensive nematic order as the cells. Consequently Qt=(βcρc+βmρm)qcQ_{t}=\left(\beta_{\mathrm{c}}\rho_{\mathrm{c}}+\beta_{\mathrm{m}}\rho_{\mathrm{m}}\right)q_{\mathrm{c}} at steady state, and the cellular nematic tensor solves

qc=g[(βcρc+βmρm)qc].\displaystyle{\color[rgb]{0,0,0}q_{\mathrm{c}}=g\left[\left(\beta_{\mathrm{c}}\rho_{\mathrm{c}}+\beta_{\mathrm{m}}\rho_{\mathrm{m}}\right)q_{\mathrm{c}}\right].} (4)

This is one of our main results. By expanding the right-hand-side of Eq. (4), we find that nematic order is possible for βcρc+βmρm>2.\beta_{\mathrm{c}}\rho_{\mathrm{c}}+\beta_{\mathrm{m}}\rho_{\mathrm{m}}>2. The βmρm\beta_{\mathrm{m}}\rho_{\mathrm{m}} term quantifies the matrix contribution and allows for nematic order even for vanishing cellular densities ρc0\rho_{\mathrm{c}}\approx 0 (grey region in Fig. 2b). The mechanism is simple: even dilute cells deposit a finite-density matrix after sufficiently long time. The matrix then acts as an external field that aligns the cells. Alternatively, rather than being aligned by current neighbors, cells are aligned by the memory of past neighbors, recorded by the matrix.

Next, we analyze the effect of ECM remodeling on the spinodal and binodal lines, as is plotted in Fig. 2.

Spinodal: matrix stabilizes the nematic order. The spinodal is given by σ/ρc=0\partial\sigma/\partial\rho_{\mathrm{c}}=0 for fixed values of βc\beta_{\mathrm{c}} and βm\beta_{\mathrm{m}} [19]. This threshold of linear instability is due to a negative compressibility. As the cellular density increases, the active stress overcomes the osmotic pressure and pushes cells up their concentration gradient. Note that active nematics can also be unstable due to a combination of active stress and shear alignment [28, 29], but this is not the case here, where the cells are effectively extensile and align with the strain rate.

Negative compressibility occurs for Qc/ρc<(1+2D)\partial Q_{\mathrm{c}}/\partial\rho_{\mathrm{c}}<-\left(1+2D\right) [Eq. (3)]. In the isotropic state, Qc0Q_{\mathrm{c}}\equiv 0 and this is not possible. Deep in the ordered state, Qc=±ρcQ_{\mathrm{c}}=\pm\rho_{\mathrm{c}}, also ensuring stability. The instability is possible, therefore, only for intermediate QcQ_{\mathrm{c}} values. For such values we expand the nonlinear terms of Eq. (4) and find its possible roots. One solution is qc=0q_{\mathrm{c}}=0 and the other is qc=(βmρm+βcρc2)/(βmρm+βcρc)3.q_{\mathrm{c}}=-\sqrt{\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}-2\right)/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right)^{3}}.

First, we examine the case of βmρm<2.\beta_{\mathrm{m}}\rho_{\mathrm{m}}<2. The cells are isotropic at low densities and become ordered at ρ=(2βmρm)/βc\rho_{\ast}=\left(2-\beta_{\mathrm{m}}\rho_{\mathrm{m}}\right)/\beta_{\mathrm{c}}. As QcρρQ_{\mathrm{c}}\sim\sqrt{\rho-\rho_{\ast}} in this case, Qc/ρc1\partial Q_{\mathrm{c}}/\partial\rho_{\mathrm{c}}\ll-1 and the system is unstable. The cell density ρ\rho_{\ast} thus marks the gas spinodal line. Otherwise, for βmρm>2\beta_{\mathrm{m}}\rho_{\mathrm{m}}>2, the slope Qc/ρc\partial Q_{\mathrm{c}}/\partial\rho_{\mathrm{c}} at vanishing densities is given by (βmρm2)/(βmρm)3/2<1\sqrt{\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}-2\right)/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}\right)^{3/2}}<1. The matrix thus increases the compressibility and ensures stability. This is why the spinodal lies outside the grey region in Fig. 2b.

Binodal: matrix allows for co-existence between different orientations. The binodal describes, for a given temperature, the densities of the macroscopic phases at coexistence. We find it from the equation for Qc,Q_{\mathrm{c}}, while replacing ρc\rho_{\mathrm{c}} by its steady-state value, σQc/(1+2D)-\sigma-Q_{\mathrm{c}}/\left(1+2D\right). Upon proper rescaling of lengths [19], we find that

Qc′′=Qc+(σ+Qc1+2D)g(Qt)F(σ,Qc).\displaystyle Q_{\mathrm{c}}^{\prime\prime}=Q_{\mathrm{c}}+\left(\sigma+\frac{Q_{\mathrm{c}}}{1+2D}\right)g\left(Q_{t}\right)\equiv F\left(\sigma,Q_{\mathrm{c}}\right). (5)

This has the same structure as Newton’s equation, where QcQ_{\mathrm{c}} plays the role of position and the xx coordinate - the role of time, while FF is the force (see also [30]). The first integral (conservation of energy) yields E=Qc2/2+UE=Q_{\mathrm{c}}^{\prime 2}/2+U, where we have denoted the “potential energy” U=dQcF(σ,Qc).U=-\int\mathrm{d}Q_{\mathrm{c}}\,F\left(\sigma,Q_{\mathrm{c}}\right).

Co-existence requires two QcQ_{\mathrm{c}} values that have the same “potential energy” UU. The co-existing phases can be either finite-sized or macroscopic, depending on the value of FF. Macroscopic phases occur for F=0F=0, where it takes an infinite “time” for the Newtonian particle to switch between the phases. These two conditions set QlQ_{\mathrm{l}}, the nematic order in the dense liquid phase, as well as σ=ρg-\sigma=\rho_{\mathrm{g}}, the density in the isotropic gas phase. To summarize, we require that Qc=0,QlQ_{\mathrm{c}}=0,Q_{\mathrm{l}} are equally-valued maxima of UU at the binodal.

We highlight the effect of the environment by focusing on two limits: a cell-dominated interaction U(βm=0)=UcU\left(\beta_{\mathrm{m}}=0\right)=U_{\mathrm{c}} where there is no matrix, and a matrix-dominated one U(βc=0)=UmU\left(\beta_{\mathrm{c}}=0\right)=U_{\mathrm{m}}, where the cells are aligned only by the matrix. Explicitly,

Uc(Qc)\displaystyle U_{\mathrm{c}}\left(Q_{\mathrm{c}}\right) =0QcdQ[ρc(Q)g(βcQ)Q],\displaystyle=\int_{0}^{Q_{\mathrm{c}}}\mathrm{d}Q\left[\rho_{\mathrm{c}}(Q)g\left(\beta_{\mathrm{c}}Q\right)-Q\right],
Um(Qc)\displaystyle U_{\mathrm{m}}\left(Q_{\mathrm{c}}\right) =0QcdQ[ρc(Q)g(βm~qc(Q))Q],\displaystyle=\int_{0}^{Q_{\mathrm{c}}}\mathrm{d}Q\left[\rho_{\mathrm{c}}(Q)g\left(\tilde{\beta_{\mathrm{m}}}q_{\mathrm{c}}(Q)\right)-Q\right], (6)

where βm~=ρmβm.\tilde{\beta_{\mathrm{m}}}=\rho_{\mathrm{m}}\beta_{\mathrm{m}}. The difference between the two cases is the magnitude of the total nematic tensor (QtQ_{t}), which appears as the argument of the nonlinear gg function. In the cell-dominated case, the argument scales as the extensive QcQ_{\mathrm{c}} that vanishes at small densities, while the matrix-dominated cases - as the intensive qcq_{\mathrm{c}}. The two potentials are plotted in Fig. 2c.

The intensive nematic order qcq_{\mathrm{c}} in the cell-dominated case is a function of βcρc\beta_{\mathrm{c}}\rho_{\mathrm{c}} [Eq. (4)] and both the spinodal and binodal lines are given by βcρc=const,\beta_{\mathrm{c}}\rho_{\mathrm{c}}=\mathrm{const,} as is displayed on Fig. 2a. In particular, we find that the nematic order at the liquid binodal βcQl\beta_{\mathrm{c}}Q_{\mathrm{l}} is not necessarily small [19]. Therefore, we cannot find it from an expansion of UcU_{\mathrm{c}}, but rather from its full nonlinear form that we evaluate numerically (and see Fig. 2c) . We find that there is indeed a macroscopic coexistence between an isotropic gas and nematic liquid, obtained from the maxima of UcU_{\mathrm{c}} for a specific value of ρg\rho_{\mathrm{g}}. The value ρl\rho_{\mathrm{l}} is then found by requiring a fixed stress, i.e., ρg=ρl+Ql/(1+2D)\rho_{\mathrm{g}}=\rho_{\mathrm{l}}+Q_{\mathrm{l}}/\left(1+2D\right). Co-existence was validated by numerical solutions of Eq. (Environment-stored memory in active nematics and extra-cellular matrix remodeling ) in 1D [31] , plotted in Fig. 2d.

The situation is very different in the matrix-dominated case. The value of qcq_{\mathrm{c}} in this case depends only on βm~\tilde{\beta_{\mathrm{m}}} [Eq. (4)]. We expand for small QcQ_{\mathrm{c}} and find UmQc2[Qc216σ2βm~3(2+βm~)]U_{\mathrm{m}}\sim-Q_{\mathrm{c}}^{2}\left[Q_{\mathrm{c}}^{2}-16\sigma^{2}\tilde{\beta_{\mathrm{m}}}^{-3}\left(-2+\tilde{\beta_{\mathrm{m}}}\right)\right]. In this case, Qc=0Q_{\mathrm{c}}=0 is a local minimum and the global maxima are Qc=±2σβm~3(2+βm~)Q_{\mathrm{c}}=\pm 2\sigma\sqrt{\tilde{\beta_{\mathrm{m}}}^{-3}\left(-2+\tilde{\beta_{\mathrm{m}}}\right)}.

Equation (4) ensures that for any solution qc=qq_{\mathrm{c}}=q of F=0,F=0, qc=qq_{\mathrm{c}}=-q is also a solution. It can be shown analytically [19] that qc<0q_{\mathrm{c}}<0 is the global maximum, while qc>0q_{\mathrm{c}}>0 is a local one, as is demonstrated by a numerical plot of UmU_{\mathrm{m}} in Fig. 2c. This form of UmU_{\mathrm{m}} allows for co-existence between finite domains with nematic order in the xx and yy directions. For example, a nematic order qc=q>0q_{\mathrm{c}}=q>0, forced by surface anchoring, will transition to qc=qq_{\mathrm{c}}=-q in the bulk, along a thickness that diverges logarithmically with βm~2\tilde{\beta_{\mathrm{m}}}-2 [19].

The co-existence between differently-oriented domains is verified by numerical solutions of Eq. (Environment-stored memory in active nematics and extra-cellular matrix remodeling ) in 1D [31], plotted in Fig. 2e. This new type of coexistence is possible because cells order at arbitrarily low densities. Then, cells aligned along the xx direction at very low densities can exert a positive active stress that matches σ\sigma. The exact form of co-existence profiles depends on angular dynamics, as explained next.

Refer to caption
Figure 2: (Color online) (a+b) Phase diagrams in the density and temperature plane (a) without a matrix and (b) with a matrix. We consider ρ=ρc\rho=\rho_{\mathrm{c}} and T=1/βcT=1/\beta_{\mathrm{c}}. Solid blue lines are the binodal and dahsed red ones are the spinodal. The values used are: D=0.5D=0.5, βm,ρm=0\beta_{\mathrm{m}},\rho_{\mathrm{m}}=0 (a) and D=0.5D=0.5, βm=βc,ρm=5\beta_{\mathrm{m}}=\beta_{\mathrm{c}},\rho_{\mathrm{m}}=5 (b).(c) Comparison between cell-dominated and matrix-dominated potentials. (d+e) Snapshots of co-existence curves from a numerical solution to the hydrodynamic equations [Eq. (Environment-stored memory in active nematics and extra-cellular matrix remodeling )] in the cell-dominated (d) and matrix-dominated cases (e). The green ellipses are a heuristic description of cellular orientational order. The values used are: D=0.5D=0.5, σ=1\sigma=-1, and β=2.05\beta=2.05 (βc\beta_{\mathrm{c}} in cell-dominated case and βm~\tilde{\beta_{\mathrm{m}}} in matrix-dominated case).

Angular dynamics: matrix possibly arrests domain coarsening. Finally, we focus on angular dynamics. While the system is invariant under global rotations of the cells and matrix together, their preferred mutual alignment results in a finite relaxation rate of their relative angle that is independent of system size. We define the angle between the preferred axis of the cells and the xx axis as ϕc\phi_{\mathrm{c}} such that the two independent terms in 𝑸c\boldsymbol{Q}_{\mathrm{c}} are Qccos(2ϕc)/2Q_{\mathrm{c}}\cos(2\phi_{\mathrm{c}})/2 and Qcsin(2ϕc)/2Q_{\mathrm{c}}\sin(2\phi_{\mathrm{c}})/2. We similarly define ϕm\phi_{\mathrm{m}} for the matrix . The relative angle between them is α/2=ϕcϕm\alpha/2=\phi_{\mathrm{c}}-\phi_{\mathrm{m}}. We rewrite Eq. (Environment-stored memory in active nematics and extra-cellular matrix remodeling ) in terms of QcQ_{\mathrm{c}}, QmQ_{\mathrm{m}}, ϕc\phi_{\mathrm{c}} and ϕm\phi_{\mathrm{m}}, and find that  [19].

tα\displaystyle\partial_{t}\alpha =[kβmρmqmqcg(Qt)Qt+k+ρcρmqcqm]sinα,\displaystyle=-\left[k\beta_{\mathrm{m}}\rho_{\mathrm{m}}\frac{q_{\mathrm{m}}}{q_{\mathrm{c}}}\frac{g\left(Q_{t}\right)}{Q_{t}}+k_{+}\frac{\rho_{\mathrm{c}}}{\rho_{\mathrm{m}}}\frac{q_{\mathrm{c}}}{q_{\mathrm{m}}}\right]\sin\alpha, (7)

where we have included the time scale 1/k1/k explicitly. Note that all the densities and nematic orders also evolve in time and are coupled with α\alpha, e.g., via shear alignment.

The two terms in the parenthesis on the right-hand side of Eq. (7) describe the dynamics of the cells and matrix, respectively. In the ordered state, their characteristic rates scale as kβmρm/(βmρm+βcρc)k\beta_{\mathrm{m}}\rho_{\mathrm{m}}/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right) and kρc,k_{-}\rho_{\mathrm{c}}, respectively [19]. The cellular rate depends on the typical cellular re-orientation rate and the strength of its alignment to the matrix field, while the matrix rate is defined by the degradation rate. The interplay between these two rates determines whether the cells are free to rotate with the matrix constantly remodeling according to the cells [kρckβmρm/(βmρm+βcρc)][k_{-}\rho_{\mathrm{c}}\gg k\beta_{\mathrm{m}}\rho_{\mathrm{m}}/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right)] or the cells are pinned to the matrix [kρckβmρm/(βmρm+βcρc)][k_{-}\rho_{\mathrm{c}}\ll k\beta_{\mathrm{m}}\rho_{\mathrm{m}}/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right)].

The latter implies that suppression of cellular relaxation dynamics. For example, consider ordered cellular domains of typical size ll with different orientations (different ϕc\phi_{\mathrm{c}} values), such as alternating bands of width ll . As long as kβmρm/(βmρm+βcρc)Dt/l2,kρck\beta_{\mathrm{m}}\rho_{\mathrm{m}}/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right)\gg D_{t}/l^{2},\,k_{-}\rho_{\mathrm{c}}, we expect these domains to remain frozen, rather than relax into a common orientation, as is the usual case (see Supplemental Figure in [19]) . Here we have denoted DtD_{t} as the total translational diffusion coefficient. In our model, it is given by Dt=D+v2/(4k)D_{t}=D+v^{2}/(4k).

Discussion. This work demonstrates how environment-stored memory qualitatively changes the known behavior of active nematics. The underlying mechanism is generic: active particles generate a finite external field even for vanishing densities. Our findings open an avenue for novel behavior of active systems. Arrested domain coarsening, for example, suggests that the steady state may contain a signature of the initial conditions . Environment-induced relaxation dynamics should also slow down defect dynamics (as was shown very recently in  [32]) and possibly arrest typical instabilities, such as nematic bands at coexistence [1] and flow transitions [28]. Finally, this may also decrease the role of fluctuations beyond mean field .

Our finding are useful in understanding ECM remodeling by cells and its consequences on cellular and tissue dynamics. We focus on quasi-2D, in-vitro studies of fibroblasts and their derived matrices (see, e.g., Ref. [9].) The cells exchange momentum with the underlying substrate, as is the case in “dry” active systems. The rigid substrate also suppresses elastic matrix deformations. Nevertheless, ECM displays orientational order for cellular densities of the order 104μm210^{-4}\mu\mathrm{m}^{-2}, which correspond to ρc102\rho_{\mathrm{c}}\approx 10^{-2}, as can be understood from the memory effect in our theory (see also [15]). In setups where elasticity is important, it is expected to serve as another mechanism for alignment [33, 34]. Generally, ECM rheology is complex, including visco-plastic contributions [35].

It was recently reported [8] that fibroblast-ECM interaction promotes alignment in non-aligned ECMs, but may also decrease the range of alignment. This is explained by our theory in a simple way: increasing the interaction means a larger cellular aligning field QtQ_{t}, leading to alignment. At the same time, increasing βm\beta_{\mathrm{m}} also increases the rate of cellular relaxation to the matrix, and may thus suppress domain coarsening. For dilute cells and assuming that the translational diffusion is mainly active (Dtv2/kD_{t}\sim v^{2}/k), we predict a domain size of the order of the cellular persistence length, i.e., of the order 10μ10\mum. This is consistent with experimental findings [9, 8]. In a future work, we will further apply our framework to predict ECM patterns observed in vivo.

In conclusion, our work demonstrates the profound effect of environment-stored memory on the steady-state and dyanmics of active nematics, especially in the biological context of ECM remodeling. It is generic in nature and is expected to play a similar role in additional active systems, including polar and synthetic.

Acknowledgements. We thank Erik Sahai and Raphaël Voituriez for useful discussions.

References

  • Chaté [2020] H. Chaté, Dry aligning dilute active matter, Annual Review of Condensed Matter Physics 11, 189 (2020).
  • Peshkov et al. [2012] A. Peshkov, I. S. Aranson, E. Bertin, H. Chaté, and F. Ginelli, Nonlinear field equations for aligning self-propelled rods, Physical review letters 109, 268701 (2012).
  • Ngo et al. [2014] S. Ngo, A. Peshkov, I. S. Aranson, E. Bertin, F. Ginelli, and H. Chaté, Large-scale chaos and fluctuations in active nematics, Physical review letters 113, 038302 (2014).
  • Großmann et al. [2016] R. Großmann, F. Peruani, and M. Bär, Mesoscale pattern formation of self-propelled rods with velocity reversal, Physical Review E 94, 050602 (2016).
  • Clark and Vignjevic [2015] A. G. Clark and D. M. Vignjevic, Modes of cancer cell invasion and the role of the microenvironment, Current Opinion in Cell Biology 36, 13 (2015).
  • Alexander and Cukierman [2016] J. Alexander and E. Cukierman, Stromal dynamic reciprocity in cancer: Intricacies of fibroblastic-ecm interactions, Current Opinion in Cell Biology 42, 80 (2016).
  • Helvert et al. [2018] S. V. Helvert, C. Storm, and P. Friedl, Mechanoreciprocity in cell migration, Nature Cell Biology 20, 8 (2018).
  • Wershof et al. [2019] E. Wershof, D. Park, R. P. Jenkins, D. J. Barry, E. Sahai, and P. A. Bates, Matrix feedback enables diverse higher-order patterning of the extracellular matrix, PLoS Computational Biology 1510.1371/journal.pcbi.1007251 (2019).
  • Park et al. [2020] D. Park, E. Wershof, S. Boeing, A. Labernadie, R. P. Jenkins, S. George, X. Trepat, P. A. Bates, and E. Sahai, Extracellular matrix anisotropy is determined by tfap2c-dependent regulation of cell collisions, Nature Materials 19, 227 (2020).
  • Conklin et al. [2011] M. W. Conklin, J. C. Eickhoff, K. M. Riching, C. A. Pehlke, K. W. Eliceiri, P. P. Provenzano, A. Friedl, and P. J. Keely, Aligned collagen is a prognostic signature for survival in human breast carcinoma, The American journal of pathology 178, 1221 (2011).
  • Kopanska et al. [2016] K. S. Kopanska, Y. Alcheikh, R. Staneva, D. Vignjevic, and T. Betz, Tensile forces originating from cancer spheroids facilitate tumor invasion, PLoS One 11, e0156442 (2016).
  • Phillips et al. [2009] R. Phillips, J. Kondev, J. Theriot, et al.Physical biology of the cell (Garland Science, 2009).
  • Grinnell [2003] F. Grinnell, Fibroblast biology in three-dimensional collagen matrices, Trends in cell biology 13, 264 (2003).
  • [14] L. Olsen, P. K. Mainia, J. A. Sherratt’, and B. Marchant, Simple modelling of extracellular matrix alignment in dermal wound healing i. cell flux induced alignment.
  • Li et al. [2017] X. Li, R. Balagam, T.-F. He, P. P. Lee, O. A. Igoshin, and H. Levine, On the mechanism of long-range orientational order of fibroblasts, Proceedings of the National Academy of Sciences 114, 8974 (2017).
  • Dallon et al. [1999] J. C. Dallon, J. A. Sherratt, and P. K. Maini, Mathematical modelling of extracellular matrix dynamics using discrete cells: Fiber orientation and tissue regeneration (1999).
  • Adar and Joanny [2021] R. M. Adar and J. F. Joanny, Permeation instabilities in active polar gels, Physical Review Letters 12710.1103/PhysRevLett.127.188001 (2021).
  • Adar and Joanny [2022] R. M. Adar and J.-F. Joanny, Active-gel theory for multicellular migration of polar cells in the extra-cellular matrix, New Journal of Physics 24, 073001 (2022).
  • [19] See Supplemental Material at [URL will be inserted by publisher] for the derivation of Eqs. (Environment-stored memory in active nematics and extra-cellular matrix remodeling )-(7) and the details of the calculations presented in the Letter .
  • [20] As cells are deformable and interact among themselves and with ECM via adhesion bonds, we describe the alignment via an interaction, EcE_{\mathrm{c}}. Generally, alignment can also be driven by excluded volume between hard rods.
  • De Gennes and Prost [1993] P.-G. De Gennes and J. Prost, The physics of liquid crystals, 83 (Oxford university press, 1993).
  • Abramowitz et al. [1988] M. Abramowitz, I. A. Stegun, and R. H. Romer, Handbook of mathematical functions with formulas, graphs, and mathematical tables (American Association of Physics Teachers, 1988).
  • Marchetti et al. [2013] M. C. Marchetti, J. F. Joanny, S. Ramaswamy, T. B. Liverpool, J. Prost, M. Rao, and R. A. Simha, Hydrodynamics of soft active matter, Reviews of Modern Physics 85, 1143 (2013).
  • Peruani et al. [2006] F. Peruani, A. Deutsch, and M. Bär, Nonequilibrium clustering of self-propelled rods, Physical Review E 74, 030904 (2006).
  • Baskaran and Marchetti [2008] A. Baskaran and M. C. Marchetti, Enhanced diffusion and ordering of self-propelled rods, Physical Review Letters 101, 268101 (2008).
  • Solon and Tailleur [2013] A. P. Solon and J. Tailleur, Revisiting the flocking transition using active spins, Physical review letters 111, 078101 (2013).
  • [27] While our theory corresponds to an extensile active stress, cells such as fibroblasts are often contractile. The same qualitative results apply in the contractile case. The main difference is that the nematic order will form along the xx direction, rather than the yy direction.
  • Voituriez et al. [2005] R. Voituriez, J.-F. Joanny, and J. Prost, Spontaneous flow transition in active polar gels, Europhysics Letters 70, 404 (2005).
  • Duclos et al. [2018] G. Duclos, C. Blanch-Mercader, V. Yashunsky, G. Salbreux, J. F. Joanny, J. Prost, and P. Silberzan, Spontaneous shear flow in confined cellular nematics, Nature Physics 14, 728 (2018).
  • Solon et al. [2015] A. P. Solon, J.-B. Caussin, D. Bartolo, H. Chaté, and J. Tailleur, Pattern formation in flocking models: A hydrodynamic description, Physical Review E 92, 062111 (2015).
  • [31] Equation (2) was solved numerically in 1D using the “pdepe” function in Matlab with zero-flux boundary conditions. As an initial condition, we consider a homogeneous cellular density ρc\rho_{\mathrm{c}} with j,Qc=0j,Q_{\mathrm{c}}=0 and no matrix segments (ρm,Qm=0\rho_{\mathrm{m}},Q_{\mathrm{m}}=0), which are supplemented by small normally distributed fluctuations in ρc,\rho_{\mathrm{c}}, j,j, and QcQ_{\mathrm{c}}.
  • Jacques et al. [2023] C. Jacques, J. Ackermann, S. Bell, C. Hallopeau, C. Perez-Gonzalez, L. Balasubramaniam, X. Trepat, B. Ladoux, A. Maitra, R. Voituriez, et al., Aging and freezing of active nematic dynamics of cancer-associated fibroblasts by fibronectin matrix remodeling, bioRxiv , 2023 (2023).
  • De et al. [2007] R. De, A. Zemel, and S. A. Safran, Dynamics of cell orientation, Nature Physics 3, 655 (2007).
  • Livne et al. [2014] A. Livne, E. Bouchbinder, and B. Geiger, Cell reorientation under cyclic stretching, Biophysical Journal 106, 42a (2014).
  • Elosegui-Artola [2021] A. Elosegui-Artola, The extracellular matrix viscoelasticity as a regulator of cell and tissue dynamics, Current Opinion in Cell Biology 72, 10 (2021).
  • You et al. [2020] Z. You, A. Baskaran, and M. C. Marchetti, Nonreciprocity as a generic route to traveling states, Proceedings of the National Academy of Sciences 117, 19767 (2020).
  • Fruchart et al. [2021] M. Fruchart, R. Hanai, P. B. Littlewood, and V. Vitelli, Non-reciprocal phase transitions, Nature 592, 363 (2021).

Supplemental material for “Environment-stored memory in active nematics and extra-cellular matrix remodeling”

This Supplemental Material (SM) provides, in greater detail, the derivation of the hydrodynamic equations and phase diagram. The outline of the SM is as follows. In Sec. I, we coarse grain the microscopic dynamics [Eq. (1) of the Letter] into hydrodynamic equations [Eq. (2) of the Letter]. In Sec II, we solve the equations at steady-state, analyze the linear stability of the steady-state (spinodal) and derive the conditions for phase coexistence (binodal). Finally, in Sec. III, we derive the nonlinear equations for the angular dynamics.

I Derivation of hydrodynamic equations

As our starting point, we consider the following microscopic dynamics of the cells and matrix segments [Eq. (1) of the Letter],

fct\displaystyle\frac{\partial f_{\mathrm{c}}}{\partial t} =(fcv𝒏)+D2fckfc+kρceEcZc\displaystyle=-\nabla\cdot\left(f_{\mathrm{c}}v\boldsymbol{n}\right)+D\nabla^{2}f_{\mathrm{c}}-kf_{\mathrm{c}}+k\rho_{\mathrm{c}}\frac{\mathrm{e}^{-E_{\mathrm{c}}}}{Z_{\mathrm{c}}}
fmt\displaystyle\frac{\partial f_{\mathrm{m}}}{\partial t} =k+2[fc(𝒓,𝒏)+fc(𝒓,𝒏)]kρcfm\displaystyle=\frac{k_{+}}{2}\left[f_{\mathrm{c}}\left(\boldsymbol{r},\boldsymbol{n}\right)+f_{\mathrm{c}}\left(\boldsymbol{r},-\boldsymbol{n}\right)\right]-k_{-}\rho_{\mathrm{c}}f_{\mathrm{m}}
k0fm+k0ρmeEmZm.\displaystyle-k_{0}f_{\mathrm{m}}+k_{0}\rho_{\mathrm{m}}\frac{\mathrm{e}^{-E_{\mathrm{m}}}}{Z_{\mathrm{m}}}. (8)

The function fcf_{\mathrm{c}} (fmf_{\mathrm{m}}) describe the distribution to find a cell (matrix segment) at position 𝒓\boldsymbol{r} with orientation 𝒏\boldsymbol{n} (𝒓\boldsymbol{r}^{\prime} with 𝒏\boldsymbol{n}^{\prime}). The microscopic equations describe how it changes due to diffusion, advection, and re-orientation. Here, in addition to the terms in Eq. (1) of the main text, we account for possible re-arrangment of matrix segments by the cells. This is descirbed similarly to cellular alignment, with a typical rate k0k_{0} and a Boltzmann factor, defined by the effective matrix interaction energy EmE_{\mathrm{m}} and partition function Zm=D𝒏exp(Em)Z_{\mathrm{m}}=\int D\boldsymbol{n}^{\prime}\exp\left(-E_{\mathrm{m}}\right). Generally, the rates may depend on the matrix and cellular densities.

Averaging the different moments of the orientation angles yield mesoscopic fields that are the focus of the hydrodynamic equations:

ρc(𝒓)\displaystyle\rho_{\mathrm{c}}\left(\boldsymbol{r}\right) =d𝒏fc(𝒓,𝒏),\displaystyle=\int\mathrm{d}\boldsymbol{n}\,f_{\mathrm{c}}\left(\boldsymbol{r},\boldsymbol{n}\right),
𝒋(𝒓)\displaystyle\boldsymbol{j}\left(\boldsymbol{r}\right) =d𝒏v𝒏fc(𝒓,𝒏),\displaystyle=\int\mathrm{d}\boldsymbol{n}\,v\boldsymbol{n}f_{\mathrm{c}}\left(\boldsymbol{r},\boldsymbol{n}\right),
𝑸c(𝒓)\displaystyle\boldsymbol{Q}_{\mathrm{c}}\left(\boldsymbol{r}\right) =d𝒏(𝒏𝒏𝑰/d)fc(𝒓,𝒏),\displaystyle=\int\mathrm{d}\boldsymbol{n}\,\left(\boldsymbol{n}\boldsymbol{n}-\boldsymbol{I}/d\right)f_{\mathrm{c}}\left(\boldsymbol{r},\boldsymbol{n}\right),
ρm(𝒓)\displaystyle\rho_{\mathrm{m}}\left(\boldsymbol{r}^{\prime}\right) =d𝒏fm(𝒓,𝒏),\displaystyle=\int\mathrm{d}\boldsymbol{n}^{\prime}\,f_{\mathrm{m}}\left(\boldsymbol{r}^{\prime},\boldsymbol{n}^{\prime}\right),
𝑸m(𝒓)\displaystyle\boldsymbol{Q}_{\mathrm{m}}\left(\boldsymbol{r}^{\prime}\right) =d𝒏(𝒏𝒏𝑰/d)fm(𝒓,𝒏),\displaystyle=\int\mathrm{d}\boldsymbol{n}^{\prime}\,\left(\boldsymbol{n}^{\prime}\boldsymbol{n}^{\prime}-\boldsymbol{I}/d\right)f_{\mathrm{m}}\left(\boldsymbol{r}^{\prime},\boldsymbol{n}^{\prime}\right),

where dd is the space dimension (d=2d=2 hereafter). We emphasize that these fields are extensive in the number of cells/matrix segments. In particular, the 𝑸\boldsymbol{Q} tensors are proportional to the density, as compared to the standard definition of nematic tensors. We define also the intensive nematic order parameter of the cells qc=Qc/ρcq_{\mathrm{c}}=Q_{\mathrm{c}}/\rho_{\mathrm{c}} and of the matrix qm=Qm/ρmq_{\mathrm{m}}=Q_{\mathrm{m}}/\rho_{\mathrm{m}}. Note that the first moment of the matrix angle (polarization) vanishes due to our choice of nematic interaction (see next) and absence of matrix advection.

According to Eq. (I), the cells and matrix segments reorient due to effective interactions EcE_{\mathrm{c}} and EmE_{\mathrm{m}}, respectively. This is a convenient choice that allows for the recovery of passive systems in simple limits. We focus on nematic interactions, which prefer a certain axis but not a direction. Furthermore, we treat the interactions within a mean-field (MF) approximation, such that the energies can be written in terms of the average fields introduced in Eq. (I) as

Ec(𝒏)\displaystyle E_{\mathrm{c}}\left(\boldsymbol{n}\right) =2Tr[(𝒏𝒏𝑰/2)(βcc𝑸c+βmc𝑸m)],\displaystyle=-2\text{Tr}\left[\left(\boldsymbol{n}\boldsymbol{n}-\boldsymbol{I}/2\right)\left(\beta_{\mathrm{c}\mathrm{c}}\boldsymbol{Q}_{\mathrm{c}}+\beta_{\mathrm{m}\mathrm{c}}\boldsymbol{Q}_{\mathrm{m}}\right)\right],
Em(𝒏)\displaystyle E_{\mathrm{m}}\left(\boldsymbol{n}^{\prime}\right) =2Tr[(𝒏𝒏𝑰/2)(βmm𝑸m+βcm𝑸c)\displaystyle=-2\text{Tr}\left[\left(\boldsymbol{n}^{\prime}\boldsymbol{n}^{\prime}-\boldsymbol{I}/2\right)\left(\beta_{\mathrm{m}\mathrm{m}}\boldsymbol{Q}_{\mathrm{m}}+\beta_{\mathrm{c}\mathrm{m}}\boldsymbol{Q}_{\mathrm{c}}\right)\right. (10)

Here the coefficients βcc\beta_{\mathrm{c}\mathrm{c}} and βmm\beta_{\mathrm{m}\mathrm{m}} describe the strength of the cell-cell and matrix-matrix aligning interactions, respectively, while βmc\beta_{\mathrm{m}\mathrm{c}} and βcm\beta_{\mathrm{c}\mathrm{m}} describe how the cells are aligned by the matrix and how the matrix is aligned by the cells, respectively. The two are not necessarily equal in active, non-equilibrium systems (non-reciprocal interaction [36, 37]).

Hereafter, we focus on the reciprocal case, such that βmc=βcm\beta_{\mathrm{m}\mathrm{c}}=\beta_{\mathrm{c}\mathrm{m}}. Furthermore, we assume a preferred parallel alignment and that the matrix is enslaved to the cells. We simplify the notations and write

Ec(𝒏)\displaystyle E_{\mathrm{c}}\left(\boldsymbol{n}\right) =2Tr[(𝒏𝒏𝑰/2)𝑸t],\displaystyle=-2\text{Tr}\left[\left(\boldsymbol{n}\boldsymbol{n}-\boldsymbol{I}/2\right)\boldsymbol{Q}_{t}\right],
Em(𝒏)\displaystyle E_{\mathrm{m}}\left(\boldsymbol{n}^{\prime}\right) =2Tr[(𝒏𝒏𝑰/2)βm𝑸c],\displaystyle=-2\text{Tr}\left[\left(\boldsymbol{n}^{\prime}\boldsymbol{n}^{\prime}-\boldsymbol{I}/2\right)\beta_{\mathrm{m}}\boldsymbol{Q}_{\mathrm{c}}\right], (11)

where we have defined the ”total” nematic tensor that aligns the cells 𝑸t=βc𝑸c+βm𝑸m\boldsymbol{Q}_{t}=\beta_{\mathrm{c}}\boldsymbol{Q}_{\mathrm{c}}+\beta_{\mathrm{m}}\boldsymbol{Q}_{\mathrm{m}}.

Next, we describe the coarse-graining procedure. It is similarly possible in the more general case of Eq. (I) and d=3d=3. We reserve this calculation for a future work.

Multiplying Eq. (I) by the appropriate powers of 𝒏\boldsymbol{n} and 𝒏\boldsymbol{n}^{\prime} and carrying out the integration leads to

ρct\displaystyle\frac{\partial\rho_{\mathrm{c}}}{\partial t} =𝒋+D2ρc,\displaystyle=-\nabla\cdot\boldsymbol{j}+D\nabla^{2}\rho_{\mathrm{c}},
𝒋t\displaystyle\frac{\partial\boldsymbol{j}}{\partial t} =v2𝑸cv22ρc+D2𝒋k𝒋,\displaystyle=-v^{2}\nabla\cdot\boldsymbol{Q}_{\mathrm{c}}-\frac{v^{2}}{2}\nabla\rho_{\mathrm{c}}+D\nabla^{2}\boldsymbol{j}-k\boldsymbol{j},
𝑸ct\displaystyle\frac{\partial\boldsymbol{Q}_{\mathrm{c}}}{\partial t} =v𝒏(𝒏𝒏𝑰/2)+D2𝑸ck𝑸c+kρcg(Qt)𝑸tQt,\displaystyle=-v\nabla\cdot\langle\boldsymbol{n}\,\left(\boldsymbol{n}\boldsymbol{n}-\boldsymbol{I}/2\right)\rangle+D\nabla^{2}\boldsymbol{Q}_{\mathrm{c}}-k\boldsymbol{Q}_{\mathrm{c}}+k\rho_{\mathrm{c}}\,g\left(Q_{t}\right)\frac{\boldsymbol{Q}_{t}}{Q_{t}},
ρmt\displaystyle\frac{\partial\rho_{\mathrm{m}}}{\partial t} =ρc(k+kρm),\displaystyle=\rho_{\mathrm{c}}\left(k_{+}-k_{-}\rho_{\mathrm{m}}\right),
𝑸mt\displaystyle\frac{\partial\boldsymbol{Q}_{\mathrm{m}}}{\partial t} =k+𝑸ckρc𝑸mk0𝑸m+k0ρmg(βmQc)𝑸cQc,\displaystyle=k_{+}\boldsymbol{Q}_{\mathrm{c}}-k_{-}\rho_{\mathrm{c}}\boldsymbol{Q}_{\mathrm{m}}-k_{0}\boldsymbol{Q}_{\mathrm{m}}+k_{0}\rho_{\mathrm{m}}\,g\left(\beta_{\mathrm{m}}Q_{\mathrm{c}}\right)\frac{\boldsymbol{Q}_{\mathrm{c}}}{Q_{\mathrm{c}}}, (12)

where \left<...\right> denotes and angular average with the probability density fcf_{\mathrm{c}}. Here we have defined the function g(x)=I1(x)/I0(x),g(x)=I_{1}(x)/I_{0}(x), where In(x)I_{n}(x) is the modified Bessel function of the nnth kind [22].

The nonlinear terms were obtained by carrying out integrals of the form

d𝒏e2Tr(𝒏𝒏𝑸)(𝒏𝒏𝑰/2)d𝒏e2Tr(𝒏𝒏𝑸)\displaystyle\frac{\int\mathrm{d}\boldsymbol{n}\,\mathrm{e}^{2\text{Tr}\left(\boldsymbol{n}\boldsymbol{n}\cdot\boldsymbol{Q}\right)}\left(\boldsymbol{n}\boldsymbol{n}-\boldsymbol{I}/2\right)}{\int\mathrm{d}\boldsymbol{n}\,\mathrm{e}^{2\text{Tr}\left(\boldsymbol{n}\boldsymbol{n}\cdot\boldsymbol{Q}\right)}} =02πdθeQcos2θ12(cos2θsin2θsin2θcos2θ)02πdθeQcos2θ\displaystyle=\frac{\int_{0}^{2\pi}\mathrm{d}\theta\mathrm{e}^{Q\cos 2\theta}\frac{1}{2}\left(\begin{matrix}\cos 2\theta&\sin 2\theta\\ \sin 2\theta&-\cos 2\theta\end{matrix}\right)}{\int_{0}^{2\pi}\mathrm{d}\theta\mathrm{e}^{Q\cos 2\theta}}
=I1(Q)I0(Q)𝑸Q,\displaystyle=\frac{I_{1}\left(Q\right)}{I_{0}\left(Q\right)}\frac{\boldsymbol{Q}}{Q}, (13)

where 𝑸=12Q(1001)\boldsymbol{Q}=\frac{1}{2}Q\left(\begin{matrix}1&0\\ 0&-1\end{matrix}\right) is a generic nematic tensor ( 𝑸=𝑸t\boldsymbol{Q}=\boldsymbol{Q}_{t} for the cell dynamics and 𝑸=βm𝑸c\boldsymbol{Q}=\beta_{\mathrm{m}}\boldsymbol{Q}_{\mathrm{c}} for the matrix dynamics).

The only term for which we do not have an exact expression is the advective term for 𝑸c\boldsymbol{Q}_{\mathrm{c}}, 𝒏𝒏𝒏\sim\langle\boldsymbol{n}\boldsymbol{n}\boldsymbol{n}\rangle. It is given by a higher moment of 𝒏\boldsymbol{n}, whose dynamics are determined by an even higher moment and so forth. We close our equations by considering moments only up to second order and by inserting the ansatz

fc\displaystyle f_{\mathrm{c}} =f0ρ+f1𝒋𝒏+f2Tr(𝑸c𝒏𝒏).\displaystyle=f_{0}\rho+f_{1}\boldsymbol{j}\cdot\boldsymbol{n}+f_{2}\text{Tr}\left(\boldsymbol{Q}_{\mathrm{c}}\cdot\boldsymbol{n}\boldsymbol{n}\right). (14)

Inserting this expression in Eq. (I) and enforcing the equalities leads to the final expression

fc\displaystyle f_{\mathrm{c}} =ρ2π+1πv𝒋𝒏+2πTr(𝑸c𝒏𝒏).\displaystyle=\frac{\rho}{2\pi}+\frac{1}{\pi v}\boldsymbol{j}\cdot\boldsymbol{n}+\frac{2}{\pi}\text{Tr}\left(\boldsymbol{Q}_{\mathrm{c}}\cdot\boldsymbol{n}\boldsymbol{n}\right). (15)

The above form of the probability density function allows for the calculation of the average

v𝒏(𝒏𝒏𝑰2)\displaystyle v\nabla\cdot\langle\boldsymbol{n}\left(\boldsymbol{n}\boldsymbol{n}-\frac{\boldsymbol{I}}{2}\right)\rangle =12π02πdθ(cos2θsin2θsin2θcos2θ)(cosθx+sinθy)(cosθjx+sinθjy)\displaystyle=\frac{1}{2\pi}\int_{0}^{2\pi}\mathrm{d}\theta\,\left(\begin{matrix}\cos 2\theta&\sin 2\theta\\ \sin 2\theta&-\cos 2\theta\end{matrix}\right)\left(\cos\theta\partial_{x}+\sin\theta\partial_{y}\right)\left(\cos\theta j_{x}+\sin\theta j_{y}\right)
=14π02πdθ(cos22θ(xjxyjy)sin22θ(xjy+yjx)sin22θ(xjy+yjx)cos22θ(xjxyjy))\displaystyle=\frac{1}{4\pi}\int_{0}^{2\pi}\mathrm{d}\theta\,\left(\begin{matrix}\cos^{2}2\theta\left(\partial_{x}j_{x}-\partial_{y}j_{y}\right)&\sin^{2}2\theta\left(\partial_{x}j_{y}+\partial_{y}j_{x}\right)\\ \sin^{2}2\theta\left(\partial_{x}j_{y}+\partial_{y}j_{x}\right)&-\cos^{2}2\theta\left(\partial_{x}j_{x}-\partial_{y}j_{y}\right)\end{matrix}\right)
=14(𝒋+𝒋T𝒋𝑰).\displaystyle=\frac{1}{4}\left(\nabla\boldsymbol{j}+\nabla\boldsymbol{j}^{\mathrm{T}}-\nabla\cdot\boldsymbol{j}\boldsymbol{I}\right). (16)

Substituting this result in Eq. (I) yields Eq. (2) of the main text.

Finally, we rewrite the hydrodynamic equations in dimensionless form. For the sake of brevity, we retain the same notations as before. Times are scaled with the average time between cellular tumbles 1/k1/k, and lengths are scaled with the cellular persistence length, l=v/kl=v/k, the typical distance a cell covers between tumbles. We find that

ρct\displaystyle\frac{\partial\rho_{\mathrm{c}}}{\partial t} =D2ρc𝒋,\displaystyle=D\nabla^{2}\rho_{\mathrm{c}}-\nabla\cdot\boldsymbol{j},
𝒋t\displaystyle\frac{\partial\boldsymbol{j}}{\partial t} =D2𝒋𝑸c12ρc𝒋,\displaystyle=D\nabla^{2}\boldsymbol{j}-\nabla\cdot\boldsymbol{Q}_{\mathrm{c}}-\frac{1}{2}\nabla\rho_{\mathrm{c}}-\boldsymbol{j},
𝑸ct\displaystyle\frac{\partial\boldsymbol{Q}_{\mathrm{c}}}{\partial t} =D2𝑸c14(𝒋+𝒋T𝒋𝑰)𝑸c+ρcg(Qt)𝑸tQt,\displaystyle=D\nabla^{2}\boldsymbol{Q}_{\mathrm{c}}-\frac{1}{4}\left(\nabla\boldsymbol{j}+\nabla\boldsymbol{j}^{\mathrm{T}}-\nabla\cdot\boldsymbol{j}\boldsymbol{I}\right)-\boldsymbol{Q}_{\mathrm{c}}+\rho_{\mathrm{c}}\,g\left(Q_{t}\right)\frac{\boldsymbol{Q}_{t}}{Q_{t}},
ρmt\displaystyle\frac{\partial\rho_{\mathrm{m}}}{\partial t} =ρc(k+kρm),\displaystyle=\rho_{\mathrm{c}}\left(k_{+}-k_{-}\rho_{\mathrm{m}}\right),
𝑸mt\displaystyle\frac{\partial\boldsymbol{Q}_{\mathrm{m}}}{\partial t} =k+𝑸ckρc𝑸mk0𝑸m+k0ρmg(βmQc)𝑸cQc.\displaystyle=k_{+}\boldsymbol{Q}_{\mathrm{c}}-k_{-}\rho_{\mathrm{c}}\boldsymbol{Q}_{\mathrm{m}}-k_{0}\boldsymbol{Q}_{\mathrm{m}}+k_{0}\rho_{\mathrm{m}}\,g\left(\beta_{\mathrm{m}}Q_{\mathrm{c}}\right)\frac{\boldsymbol{Q}_{\mathrm{c}}}{Q_{\mathrm{c}}}. (17)

II Steady-state phase diagram

The phase diagram in the standard active nematic case (no matrix) is analyzed similarly to a liquid-gas transition in the density-temperature plane, where the alignment strength plays the role of inverse temperature [1]. The system forms an isotropic ”gaseous” phase at low densities and high temperatures, and a nematic ”liquid” phase at high densities at low temperatures. In between, there is a co-existence region that is generally unstable. Here, we explain how the matrix modifies this picture and derive the new phase diagram from Eq. (I).

II.1 Steady state

We solve the hydrodynamic equations [Eq. (I)] at steady-state. First, we focus on the active cellular current density. For large lengthscales, we neglect the diffusion term and retain

𝒋=12(ρc𝑰+2𝑸c).\displaystyle\boldsymbol{j}=-\frac{1}{2}\nabla\cdot\left(\rho_{\mathrm{c}}\boldsymbol{I}+2\boldsymbol{Q}_{\mathrm{c}}\right). (18)

Substituting in the cellular continuity equation, we find

ρct\displaystyle\frac{\partial\rho_{\mathrm{c}}}{\partial t} =(D+12)[(ρc𝑰+2𝑸c1+2D)].\displaystyle=-\left(D+\frac{1}{2}\right)\nabla\cdot\left[\nabla\cdot\left(\rho_{\mathrm{c}}\boldsymbol{I}+\frac{2\boldsymbol{Q}_{\mathrm{c}}}{1+2D}\right)\right]. (19)

The right-hand side of this equation is the divergence of the total cellular current density. The expression for it can be interpreted as force balance. The friction due the current is balanced by the divergence of the stress tensor, 𝝈=ρc𝑰2𝑸c/(1+2D)\boldsymbol{\sigma}=-\rho_{\mathrm{c}}\boldsymbol{I}-2\boldsymbol{Q}_{\mathrm{c}}/\left(1+2D\right). It includes an ideal-gas-like term and an extensile active stress 𝑸c\sim\boldsymbol{Q}_{\mathrm{c}}. Within this picture, the cells can be considered as “pushers”. Similar results can be obtained for ”pullers” (contractile active stress), by rotating the nematic tensor by π/2\pi/2 (setting 𝑸c𝑸c\boldsymbol{Q}_{\mathrm{c}}\to-\boldsymbol{Q}_{\mathrm{c}}).

At steady-state, the effective stress tensor is fixed, introducing a constraint between the density and nematic order parameter. We consider variations of these fields along the xx axis. The above constraint reduces to

ρc+Qc1+2D=σxx=σ,\displaystyle\rho_{\mathrm{c}}+\frac{Q_{\mathrm{c}}}{1+2D}=-\sigma_{xx}=-\sigma, (20)

where we have denoted σxx=σ\sigma_{xx}=\sigma for brevity. For now, we consider a homogeneous steady-state, where the stress is fixed by definition. Below, we analyze also possible profiles along the xx direction to find the binodal of the phase diagram.

In the homogeneous case, the cellular density is determined by the initial condition. Cellular nematic order may form in any direction and we define its direction as the xx direction. Its magnitude is determined by the nonlinear terms in Eq. (I). In order to find it, we must first determine the steady state of the matrix.

The matrix density is given by ρm=k+/k\rho_{\mathrm{m}}=k_{+}/k_{-}. Importantly, this expression is independent of ρc\rho_{\mathrm{c}}. Even rather dilute cells can deposit a finite-density matrix after sufficient time. The matrix nematic order 𝑸m\boldsymbol{Q}_{\mathrm{m}} is enslaved to the cellular one and effectively renormalizes cell-cell interactions. It forms along the same axis as the cellular tensor 𝑸c\boldsymbol{Q}_{\mathrm{c}}. It is given by

𝑸m\displaystyle\boldsymbol{Q}_{\mathrm{m}} =ρmkρc+k0[k+k0g(βmQc)Qc]𝑸c.\displaystyle=\frac{\rho_{\mathrm{m}}}{k_{-}\rho_{\mathrm{c}}+k_{0}}\left[k_{-}+k_{0}\frac{g\left(\beta_{\mathrm{m}}Q_{\mathrm{c}}\right)}{Q_{\mathrm{c}}}\right]\boldsymbol{Q}_{\mathrm{c}}. (21)

The magnitude QmQ_{\mathrm{m}} is proportional to the matrix density, as expected. The first term in the parenthesis of Eq. (21) originates from active matrix deposition and degradation (note that ρm\rho_{\mathrm{m}} itself includes k+k_{+}) , while the second term - from re-arrangement due to alignment. The weighted contribution of each term is determined by the rates kk_{-} and k0k_{0}, respectively. We focus hereafter on matrix deposition and set k0=0,k_{0}=0, restricting ourselves to the limit presented in the main text. In this case, qm=qcq_{\mathrm{m}}=q_{\mathrm{c}}. In any case, the matrix dominates QtQ_{t} at low cellular densities.

The magnitude of the cellular nematic order QcQ_{\mathrm{c}} is determined by

FQcρcg(Qt)=0,\displaystyle F\equiv Q_{\mathrm{c}}-\rho_{\mathrm{c}}g\left(Q_{t}\right)=0, (22)

While Qc=0Q_{\mathrm{c}}=0 is always a solution, a non-vanishing solution also appears for FQ(0)0,F_{Q}(0)\leq 0, where FQ=F/QcF_{Q}=\partial F/\partial Q_{\mathrm{c}}. We similarly define the partial derivative with respect to the density as Fρ=F/ρcF_{\rho}=\partial F/\partial\rho_{\mathrm{c}}.

We expand FF for small qcq_{\mathrm{c}} and find that

Fρc[qc(1βcρc+βmρm2)+qc316(βcρc+βmρm)3].\displaystyle F\approx\rho_{\mathrm{c}}\left[q_{\mathrm{c}}\left(1-\frac{\beta_{\mathrm{c}}\rho_{\mathrm{c}}+\beta_{\mathrm{m}}\rho_{\mathrm{m}}}{2}\right)+\frac{q_{\mathrm{c}}^{3}}{16}\left(\beta_{\mathrm{c}}\rho_{\mathrm{c}}+\beta_{\mathrm{m}}\rho_{\mathrm{m}}\right)^{3}\right]. (23)

The roots of FF are either Qc=0Q_{\mathrm{c}}=0 or

Qc=±ρcβmρm+βcρc2(βmρm+βcρc)3.\displaystyle Q_{\mathrm{c}}=\pm\rho_{\mathrm{c}}\sqrt{\frac{\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}-2}{\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right)^{3}}}. (24)

This means that qcq_{\mathrm{c}} is only a function of the mean field βmρm+βcρc\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}.

First, we examine the case βmρm<2.\beta_{\mathrm{m}}\rho_{\mathrm{m}}<2. The cells are isotropic at low densities and become ordered at ρ=(2βmρm)/βc\rho_{\ast}=\left(2-\beta_{\mathrm{m}}\rho_{\mathrm{m}}\right)/\beta_{\mathrm{c}}. We see that QcρρQ_{\mathrm{c}}\sim\sqrt{\rho-\rho_{\ast}} in this case. Otherwise, for βmρm>2,\beta_{\mathrm{m}}\rho_{\mathrm{m}}>2, the cells can be ordered even at vanishing densities. This is the main effect of the matrix: the cells are aligned by the cell-matrix interaction, because the matrix has a finite density and it acts as an external field even at vanishing cellular densities.

II.2 Linear stability analysis: spinodal

Next, we analyze the linear stability of the steady-state solution in the limit of infinite wavelength. The onset of instability defines the spinodal line. In this hydrodynamic limit, we can treat the cellular current, as well as the density and nematic order of the matrix as fast variables. The stability analysis, therefore, is restricted to the cellular density and nematic tensor. It is convenient to write the nematic tensor in terms of its magnitude QcQ_{\mathrm{c}} and angle ϕc\phi_{\mathrm{c}}

𝑸c=Qc2(cos2ϕcsin2ϕcsin2ϕccos2ϕc).\displaystyle\boldsymbol{Q}_{\mathrm{c}}=\frac{Q_{\mathrm{c}}}{2}\left(\begin{matrix}\cos 2\phi_{\mathrm{c}}&\sin 2\phi_{\mathrm{c}}\\ \sin 2\phi_{\mathrm{c}}&-\cos 2\phi_{\mathrm{c}}\end{matrix}\right). (25)

We analyze the linear stability of the steady state with respect to perturbations with a growth rate ss and wave vector 𝒑=p(cosθ,sinθ)\boldsymbol{p}=p\left(\cos\theta,\sin\theta\right) of the form 𝒙=𝒙0+𝒙1exp(st+i𝒑𝒓),\boldsymbol{x}=\boldsymbol{x}^{0}+\boldsymbol{x}^{1}\exp\left(st+i\boldsymbol{p}\cdot\boldsymbol{r}\right), where 𝒙=(ρc,Qc,ϕc)\boldsymbol{x}=(\rho_{\mathrm{c}},Q_{\mathrm{c}},\phi_{\mathrm{c}}).

Linearizing Eq. (I) yields

sρc1\displaystyle s\rho_{\mathrm{c}}^{1} =p2[D+12+12cos2θQc1+sin2θQc0ϕc1],\displaystyle=-p^{2}\left[D+\frac{1}{2}+\frac{1}{2}\cos 2\theta Q_{\mathrm{c}}^{1}+\sin 2\theta Q_{\mathrm{c}}^{0}\phi_{\mathrm{c}}^{1}\right],
sQc1\displaystyle sQ_{\mathrm{c}}^{1} =p2[(D+14)Qc1+14cos2θρc1]Fρρc1FQQc1,\displaystyle=-p^{2}\left[\left(D+\frac{1}{4}\right)Q_{\mathrm{c}}^{1}+\frac{1}{4}\cos 2\theta\rho_{\mathrm{c}}^{1}\right]-F_{\rho}\rho_{\mathrm{c}}^{1}-F_{Q}Q_{\mathrm{c}}^{1},
2Qc0sϕc1\displaystyle 2Q_{\mathrm{c}}^{0}s\phi_{\mathrm{c}}^{1} =p2[(D+14)2Qc0ϕc1+14sin2θρc1],\displaystyle=-p^{2}\left[\left(D+\frac{1}{4}\right)2Q_{\mathrm{c}}^{0}\phi_{\mathrm{c}}^{1}+\frac{1}{4}\sin 2\theta\rho_{\mathrm{c}}^{1}\right], (26)

where, as before, FQ=F/QcF_{Q}=\partial F/\partial Q_{\mathrm{c}} and Fρ=F/ρcF_{\rho}=\partial F/\partial\rho_{\mathrm{c}}.

The equation on ρc1\rho_{\mathrm{c}}^{1} infers that the scaling of the growth rate is O(p2)O\left(p^{2}\right). In the hydrodynamic limit p1p\ll 1 we retain only such terms of order p2p^{2}. This means that the contribution of rotations (ϕc1\phi_{\mathrm{c}}^{1}) to density changes is negligible and that we can write Qc1Fρρc1/FQQ_{\mathrm{c}}^{1}\approx-F_{\rho}\rho_{\mathrm{c}}^{1}/F_{Q}. Inserting back in the equation for the density and focusing on θ=0\theta=0, where the destabilizing term is largest, leads to

s\displaystyle s =p2(D+1212FρFQ).\displaystyle=-p^{2}\left(D+\frac{1}{2}-\frac{1}{2}\frac{F_{\rho}}{F_{Q}}\right). (27)

The system is thus linearly unstable for FρFQ1+2D,\frac{F_{\rho}}{F_{Q}}\geq 1+2D, where the equality defines the spinodal.

The interpretation of this instability becomes straightforward when we notice that Fρ/FQ=Qc/ρc-F_{\rho}/F_{Q}=-\partial Q_{\mathrm{c}}/\partial\rho_{\mathrm{c}}. Eq. (20) then yields

σρc0.\displaystyle\frac{\partial\sigma}{\partial\rho_{\mathrm{c}}}\geq 0. (28)

This is, therefore, a mechanical instability. It occurs when the cells become sufficiently ordered upon a density increase, such that the active stress overcomes the pressure and pushes cells up their density gradient.

II.3 Analysis of the spinodal criterion

The instability criterion is by Qc/ρc1+2D-\partial Q_{\mathrm{c}}/\partial\rho_{\mathrm{c}}\geq 1+2D. It can be understood from the functional dependence of Qc(ρc)Q_{\mathrm{c}}\left(\rho_{\mathrm{c}}\right). In the absence of a matrix, at a given temperature, the cells are isotropic up to a finite density ρ\rho_{\ast}. Around this density, the nematic order scales as QcρcρQ_{\mathrm{c}}\sim\sqrt{\rho_{\mathrm{c}}-\rho_{\ast}}. Therefore, the derivative diverges at this point and the criterion for instability is fulfilled for Qc<0.Q_{\mathrm{c}}<0. This defines the gas spinodal. At large densities, all the cells are ordered such that Qc=ρcQ_{\mathrm{c}}=-\rho_{\mathrm{c}} and the system is stable. The density where stability sets in defines the liquid spinodal.

The matrix may break this behavior for sufficiently strong interactions. As we have found in Eq. (24), the matrix allows for the cells to be aligned at zero density for βmρm>2\beta_{\mathrm{m}}\rho_{\mathrm{m}}>2. In this case Qc/ρc\partial Q_{\mathrm{c}}/\partial\rho_{\mathrm{c}} is sufficiently small, such that the system is always stable. Otherwise, for βmρm<2\beta_{\mathrm{m}}\rho_{\mathrm{m}}<2, the gas spinodal is given by βcρ=2βmρm\beta_{\mathrm{c}}\rho_{\ast}=2-\beta_{\mathrm{m}}\rho_{\mathrm{m}}.

II.4 Coexistence criteria: binodal

Consider an isotropic dilute (gas) phase with density ρg\rho_{\mathrm{g}} and an ordered dense (liquid) phase with density ρl\rho_{\mathrm{l}} and nematic order parameter QlQ_{\mathrm{l}}. Co-existence requires an equal stress σ\sigma across the system. This sets

ρg\displaystyle\rho_{\mathrm{g}} =σ=ρl+Ql1+2D.\displaystyle=-\sigma=\rho_{\mathrm{l}}\,+\frac{Q_{\mathrm{l}}}{1+2D}. (29)

The ideal-gas pressure in the gaseous phase is balanced by a combination of an ideal-gas pressure and active stress in the liquid phase. As the liquid phase is denser, this requires that that the active stress will act in the opposite direction. For pusher cells, this is possible for Qł<0Q_{\l}<0, i.e., the cells are oriented in the yy direction, normal to the varying density profile.

The values of ρg\rho_{\mathrm{g}} and ρl\rho_{\mathrm{l}} at co-existence define the binodal. We continue its derivation by inserting 𝒋\boldsymbol{j} in terms of 𝑸c\boldsymbol{Q}_{\mathrm{c}} in the equation for the nematic order to find

D(1+14(1+2D))Qc′′=Qc+(σ+Qc1+2D)g(Qt),\displaystyle D\left(1+\frac{1}{4\left(1+2D\right)}\right)Q_{\mathrm{c}}^{\prime\prime}=Q_{\mathrm{c}}+\left(\sigma+\frac{Q_{\mathrm{c}}}{1+2D}\right)\,g\left(Q_{t}\right), (30)

where the total nematic order is Qt=βcQc+βmQmQ_{t}=\beta_{\mathrm{c}}Q_{\mathrm{c}}+\beta_{m}Q_{\mathrm{m}}. To simplify the equations, we further rescale lengths with D(1+14(1+2D)),\sqrt{D\left(1+\frac{1}{4\left(1+2D\right)}\right)}, such that the left-hand side of Eq. (30) simplifies to Qc′′Q_{\mathrm{c}}^{\prime\prime}. Furthermore, we define the right-hand side of the equation as F(σ,Qc)F\left(\sigma,Q_{\mathrm{c}}\right), such that

Qc′′=F(σ,Qc).\displaystyle Q_{\mathrm{c}}^{\prime\prime}=F\left(\sigma,Q_{\mathrm{c}}\right). (31)

Equation (31) has the same structure as Newton’s equation, where the xx coordinate plays the role of time and FF plays the role of the force (see also [30]). The first integral (conservation of energy) yields

E=12Qc2+U,\displaystyle E=\frac{1}{2}Q_{\mathrm{c}}^{\prime 2}+U, (32)

where we have denoted the “potential energy” U=dQcF(σ,Qc).U=-\int\mathrm{d}Q_{\mathrm{c}}\,F\left(\sigma,Q_{\mathrm{c}}\right). The binodal line describe the density and nematic order at macroscopically phase-separated states. Continuing the analogy to a Newtonian particle, this requires two QcQ_{\mathrm{c}} values that have the same energy (such that there is co- existence) and where the force vanishes, F=0F=0, such that it takes an infinite time to escape them. Without loss of generality, we take the energy value to be E=0E=0. These two conditions set QlQ_{\mathrm{l}}, the nematic order in the dense phase, as well as σ=ρg-\sigma=\rho_{\mathrm{g}}, the density in the isotropic gas phase. To summarize, we require that Qc=0,QłQ_{\mathrm{c}}=0,Q_{\l} are degenerate roots of UU.

Generally, co-existence between finite-sized domains with different QcQ_{\mathrm{c}} values is possible as long as they share the same value of UU. The width of the domains is then determined by the values of FF.

II.5 Cell-dominated and matrix-dominated limits

It instructive to focus on two limiting cases: a cell-dominated case, where βm=0\beta_{\mathrm{m}}=0, and a matrix-dominated case, where βc=0\beta_{\mathrm{c}}=0.

Cell-dominated case. In the absence of a matrix, we denote U(βm=0)=UcU\left(\beta_{\mathrm{m}}=0\right)=U_{\mathrm{c}} and find that

Uc(Qc)\displaystyle U_{\mathrm{c}}\left(Q_{\mathrm{c}}\right) =0QcdQ[ρc(Q)g(βcQ)Q].\displaystyle=\int_{0}^{Q_{\mathrm{c}}}\mathrm{d}Q\left[\rho_{\mathrm{c}}(Q)g\left(\beta_{\mathrm{c}}Q\right)-Q\right]. (33)

Multiplying UcU_{\mathrm{c}} by βc2,\beta_{\mathrm{c}}^{2}, we see that it is a function of βcρc\beta_{\mathrm{c}}\rho_{\mathrm{c}} and βcQc\beta_{\mathrm{c}}Q_{\mathrm{c}}. This means that the phase diagram will be written in terms of lines of the form βcρc=const.\beta_{\mathrm{c}}\rho_{\mathrm{c}}=\mathrm{const.}

The value of the nematic order at the binodal is found by differentiating UcU_{\mathrm{c}} and requiring F=0F=0. Expanding for small QcQ_{\mathrm{c}} we find that βcQł=16/[3(1+2D)βcρg]\beta_{\mathrm{c}}Q_{\l}=-16/\left[3\left(1+2D\right)\beta_{\mathrm{c}}\rho_{\mathrm{g}}\right]. In particular, having found the spinodal at βcρ=2\beta_{\mathrm{c}}\rho_{\ast}=2, we know that βcρg<2\beta_{\mathrm{c}}\rho_{\mathrm{g}}<2. This means that βcQl<8/[3(1+2D)]\beta_{\mathrm{c}}Q_{\mathrm{l}}<-8/\left[3\left(1+2D\right)\right]. For small DD values, as we expect for living cells, we find that βcQł<2\beta_{\mathrm{c}}Q_{\l}<-2. For such values, the linear expansion that we have used is not valid. It indicates that, in order to derive the phase diagram, UU should not be expanded around Qc=0Q_{\mathrm{c}}=0, but rather its full nonlinear form should be used and QcQ_{\mathrm{c}} shoud be found numerically.

The value of ρg=σ\rho_{\mathrm{g}}=-\sigma at the binodal is found by requiring that Qc=0,QlQ_{\mathrm{c}}=0,Q_{\mathrm{l}} have equal values of the “potential energy” UcU_{\mathrm{c}} with a vanishing “force” F=0F=0. Then, the liquid branch of the binodal is found by requiring that the stress is fixed,

βcρł=βcρg+163(1+2D)21βcρg.\displaystyle\beta_{\mathrm{c}}\rho_{\l}=\beta_{\mathrm{c}}\rho_{\mathrm{g}}+\frac{16}{3\left(1+2D\right)^{2}}\frac{1}{\beta_{\mathrm{c}}\rho_{\mathrm{g}}}. (34)

Matrix-dominated case. In the absence of cell-cell interactions, we denote U(βc=0)=UmU\left(\beta_{\mathrm{c}}=0\right)=U_{\mathrm{m}} and find that

Um(Qc)\displaystyle U_{\mathrm{m}}\left(Q_{\mathrm{c}}\right) =0QcdQ[ρc(Q)g(βm~qc(Q))Q]\displaystyle=\int_{0}^{Q_{\mathrm{c}}}\mathrm{d}Q\left[\rho_{\mathrm{c}}(Q)g\left(\tilde{\beta_{\mathrm{m}}}q_{\mathrm{c}}(Q)\right)-Q\right]
0QcdQ[(1+βm~2)Qβm~316σ2Q3]\displaystyle\approx\int_{0}^{Q_{\mathrm{c}}}\mathrm{d}Q\left[\left(-1+\frac{\tilde{\beta_{\mathrm{m}}}}{2}\right)Q-\frac{\tilde{\beta_{\mathrm{m}}}^{3}}{16\sigma^{2}}Q^{3}\right]
=βm~364σ2Qc2(Qc216σ22+βm~βm~3).\displaystyle=-\frac{\tilde{\beta_{\mathrm{m}}}^{3}}{64\sigma^{2}}Q_{\mathrm{c}}^{2}\left(Q_{\mathrm{c}}^{2}-16\sigma^{2}\frac{-2+\tilde{\beta_{\mathrm{m}}}}{\tilde{\beta_{\mathrm{m}}}^{3}}\right). (35)

The potential is qualitatively different in this case. For βm~<2,\tilde{\beta_{\mathrm{m}}}<2, it has a local maximum at Qc=0Q_{\mathrm{c}}=0 and no order forms. For βm~>2,\tilde{\beta_{\mathrm{m}}}>2, however, Qc=0Q_{\mathrm{c}}=0 is a local minimum and there are two maxima at Qc=±2σ(2+βm~)βm~3Q_{\mathrm{c}}=\pm 2\sigma\sqrt{\left(-2+\tilde{\beta_{\mathrm{m}}}\right)\tilde{\beta_{\mathrm{m}}}^{-3}}. Here we assumed that we are close to the transition βm~2.\tilde{\beta_{\mathrm{m}}}\approx 2.

The value of the intensive, nematic order parameter, qc=Qc/ρcq_{\mathrm{c}}=Q_{\mathrm{c}}/\rho_{\mathrm{c}} in this case, depends only on βm~\tilde{\beta_{\mathrm{m}}}. This is obtained from the steady-state condition qc=g(βm~qc).q_{\mathrm{c}}=g\left(\tilde{\beta_{\mathrm{m}}}q_{\mathrm{c}}\right). The explicit dependence on βm~\tilde{\beta_{\mathrm{m}}} also infers that the phase diagram is not given by straight lines, as was in the cell-dominated case.

The potential Um(βm~2)U_{\mathrm{m}}\left(\tilde{\beta_{\mathrm{m}}}\gtrapprox 2\right) infers co-existence of nematic order along the xx and yy directions at any value of the density. This is possible because cells order at arbitrarily low densities. Then, cells aligned along the xx direction at very low densities can exert a positive active stress that matches σ\sigma. Note that σ\sigma here does not correspond to the density of the gas phase, but is simply the average cell density.

The co-existence in this case is between finite-sized domains. To demonstrate this fact, we analytically calculate the energy difference between two solutions qc=±qq_{\mathrm{c}}=\pm q of F=0F=0 [Eq. (4) of the main text]. We make use of the identity Qc=σqc/[1+q/(1+2D)]Q_{\mathrm{c}}=-\sigma q_{\mathrm{c}}/\left[1+q/\left(1+2D\right)\right] to rewrite UmU_{\mathrm{m}} as an integral over the intensive order parameter. We find that for q>0,q>0, the energy difference ΔUm=Um(q)Um(q)\Delta U_{\mathrm{m}}=U_{\mathrm{m}}\left(q\right)-U_{\mathrm{m}}\left(-q\right) is given by

ΔUm\displaystyle\Delta U_{\mathrm{m}} =σ2qqdq[q+g(βm~q)][(1+q1+2D)3]\displaystyle=\sigma^{2}\int_{-q}^{q}\mathrm{d}q^{\prime}\left[-q^{\prime}+g\left(\tilde{\beta_{\mathrm{m}}}q^{\prime}\right)\right]\left[\left(1+\frac{q^{\prime}}{1+2D}\right)^{-3}\right]
=σ20qdq[qg(βm~q)][(1q1+2D)3(1+q1+2D)3]\displaystyle=\sigma^{2}\int_{0}^{q}\mathrm{d}q^{\prime}\left[q^{\prime}-g\left(\tilde{\beta_{\mathrm{m}}}q^{\prime}\right)\right]\left[\left(1-\frac{q^{\prime}}{1+2D}\right)^{-3}-\left(1+\frac{q^{\prime}}{1+2D}\right)^{-3}\right]
<0.\displaystyle<0. (36)

Here we have used the fact that qg(βm~q)<0q^{\prime}-g\left(\tilde{\beta_{\mathrm{m}}}q^{\prime}\right)<0 for 0<q<q0<q^{\prime}<q. The above calculation means that finite-sized bands of q<qc<0-q<q_{\mathrm{c}}<0 (in yy-direction) can co-exist with positive qc>0q_{\mathrm{c}}>0 (in xx-direction).

In the presence of a surface at x=0x=0, the form of UmU_{\mathrm{m}} allows for nematic order that is anchored along the xx direction to gradually change into its steady-state configuration along the yy direction in the bulk. This transition is characterized by a coherence length ll that, close to the ordering transition, diverges logarithmically with βm~2\tilde{\beta_{\mathrm{m}}}-2, as we show next. The length ll can also be regarded as the thickness of a pre-wetting layer.

We consider a surface at x=0x=0 that enforces a nematic order qc(x=0)=q>0q_{\mathrm{c}}\left(x=0\right)=q>0, where F(q)=0F(q)=0. Far away from the surface, the nematic order reaches the value qc=qq_{\mathrm{c}}=-q, where the “potential energy” is at its global maximum, Um(q)=EU_{\mathrm{m}}(-q)=E. We define ll as the distance over which the nematic order vanishes, q(x=l)=0q\left(x=l\right)=0. Note that the nematic order here is treated as a scalar. It does not rotate, but rather changes sign, inferring a disordered intermediate region.

The thickness ll is given by

l\displaystyle l =Qc(Qc=Q)Qc(Qc=0)dQcQc′′=2ΔUm2EdQcF,\displaystyle=-\int_{Q_{\mathrm{c}}^{\prime}\left(Q_{\mathrm{c}}=Q\right)}^{Q_{\mathrm{c}}^{\prime}\left(Q_{\mathrm{c}}=0\right)}\frac{\mathrm{d}Q_{\mathrm{c}}^{\prime}}{Q_{\mathrm{c}}^{\prime\prime}}=-\int_{\sqrt{2\Delta U_{\mathrm{m}}}}^{\sqrt{2E}}\frac{\mathrm{d}Q_{\mathrm{c}}^{\prime}}{F}, (37)

where Q=ρcqQ=\rho_{\mathrm{c}}q. Here we have used the equation (“Newton’s law”) Qc′′=FQ_{\mathrm{c}}^{\prime\prime}=F for the integrand and the first integral (“conservation of energy”) Qc2/2+Um=EQ_{\mathrm{c}}^{\prime 2}/2+U_{\mathrm{m}}=E for the integration limits.

We focus on βm~2\tilde{\beta_{\mathrm{m}}}\gtrapprox 2 and expand UmU0Qc2(Qc22Q2),U_{\mathrm{m}}\approx-U_{0}Q_{\mathrm{c}}^{2}\left(Q_{\mathrm{c}}^{2}-2Q^{2}\right), where U0>0U_{0}>0 is a prefactor. Conservation of energy then allows to relate between QcQ_{\mathrm{c}} and QcQ_{\mathrm{c}}^{\prime} according to Qc2U0(Qc2Q2)Q_{\mathrm{c}}^{\prime}\approx\sqrt{2U_{0}}\left(Q_{\mathrm{c}}^{2}-Q^{2}\right). We similarly expand F4U0Qc(Qc2Q2)F\approx 4U_{0}Q_{\mathrm{c}}\left(Q_{\mathrm{c}}^{2}-Q^{2}\right) and find that, to leading order, FQcF\sim Q_{\mathrm{c}}^{\prime}. Inserting back in Eq. (38) yields

l2ΔUm2EdQcQc=12lnEΔUm.\displaystyle l\sim\int_{\sqrt{2\Delta U_{\mathrm{m}}}}^{\sqrt{2E}}\frac{\mathrm{d}Q_{\mathrm{c}}^{\prime}}{Q_{\mathrm{c}}^{\prime}}=\frac{1}{2}\ln\frac{E}{\Delta U_{\mathrm{m}}}. (38)

While E/ΔUmE/\Delta U_{\mathrm{m}} is always larger than unity, both EE and ΔUm\Delta U_{\mathrm{m}} vanish at βm~=2.\tilde{\beta_{\mathrm{m}}}=2. For completeness, we find their scaling with βm~2\tilde{\beta_{\mathrm{m}}}-2 close to the transition. First, E=Um(Qc=Q)=U0Q4(βm~2)2E=U_{\mathrm{m}}\left(Q_{\mathrm{c}}=Q\right)=U_{0}Q^{4}\sim\left(\tilde{\beta_{\mathrm{m}}}-2\right)^{2}. Second, expanding Eq. (II.5) yields ΔUm(βm~2)q3(βm~2)5/2.\Delta U_{\mathrm{m}}\sim\left(\tilde{\beta_{\mathrm{m}}}-2\right)q^{3}\sim\left(\tilde{\beta_{\mathrm{m}}}-2\right)^{5/2}. Overall, we find that lln(βm~2)l\sim-\ln\left(\tilde{\beta_{\mathrm{m}}}-2\right).

III Non-linear angular dynamics

Next, we focus on the angular dynamics of both the cells and matrix in the limits of large wavelengths. In the absence of a matrix, the cellular angle is a soft mode and its global rotation does not decay. The matrix, however, introduces a preferred axis, and any relative angle between the cells and matrix is expected to decay over time with a rate that is independent of system size. The relative angle closes via both cellular and matrix dynamics, and their interplay depends on the typical cellular and matrix rates.

We analyze the angular dynamics by writing the nematic tensors in terms of their modulus and angle

𝑸c=Qc2(cos2ϕcsin2ϕcsin2ϕccos2ϕc),𝑸m=Qm2(cos2ϕmsin2ϕmsin2ϕmcos2ϕm),𝑸t=Qt2(cos2ϕtsin2ϕtsin2ϕtcos2ϕt).\boldsymbol{Q}_{\mathrm{c}}=\frac{Q_{\mathrm{c}}}{2}\left(\begin{matrix}\cos 2\phi_{\mathrm{c}}&\sin 2\phi_{\mathrm{c}}\\ \sin 2\phi_{\mathrm{c}}&-\cos 2\phi_{\mathrm{c}}\end{matrix}\right),\,\boldsymbol{Q}_{\mathrm{m}}=\frac{Q_{\mathrm{m}}}{2}\left(\begin{matrix}\cos 2\phi_{\mathrm{m}}&\sin 2\phi_{\mathrm{m}}\\ \sin 2\phi_{\mathrm{m}}&-\cos 2\phi_{\mathrm{m}}\end{matrix}\right),\,\boldsymbol{Q}_{t}=\frac{Q_{t}}{2}\left(\begin{matrix}\cos 2\phi_{t}&\sin 2\phi_{t}\\ \sin 2\phi_{t}&-\cos 2\phi_{t}\end{matrix}\right). (39)

The equations on 𝑸c\boldsymbol{Q}_{\mathrm{c}} and 𝑸m\boldsymbol{Q}_{\mathrm{m}} can be thought of as equations on vectors that can be represented in polar coordinates. This results in equations on the moduli QcQ_{\mathrm{c}} and QmQ_{\mathrm{m}} and the angles ϕc\phi_{\mathrm{c}} and ϕm,\phi_{\mathrm{m}}, respectively,

tQc\displaystyle\partial_{t}Q_{\mathrm{c}} =Qc+ρcg(Qt)cos2(ϕcϕt),\displaystyle=-Q_{\mathrm{c}}+\rho_{\mathrm{c}}g\left(Q_{t}\right)\cos 2\left(\phi_{\mathrm{c}}-\phi_{t}\right),
2Qctϕc\displaystyle 2Q_{\mathrm{c}}\partial_{t}\phi_{\mathrm{c}} =ρcg(Qt)sin2(ϕcϕt),\displaystyle=\rho_{\mathrm{c}}g\left(Q_{t}\right)\sin 2\left(\phi_{\mathrm{c}}-\phi_{t}\right),
tQm\displaystyle\partial_{t}Q_{\mathrm{m}} =ρckQm+k+Qccos2(ϕcϕm),\displaystyle=-\rho_{\mathrm{c}}k_{-}Q_{\mathrm{m}}+k_{+}Q_{\mathrm{c}}\cos 2\left(\phi_{\mathrm{c}}-\phi_{\mathrm{m}}\right),
2Qmtϕm\displaystyle 2Q_{\mathrm{m}}\partial_{t}\phi_{\mathrm{m}} =k+Qcsin2(ϕcϕm).\displaystyle=k_{+}Q_{\mathrm{c}}\sin 2\left(\phi_{\mathrm{c}}-\phi_{\mathrm{m}}\right). (40)

As the dynamics are invariant to global rotations of the systems, the angle ϕtϕc\phi_{t}-\phi_{\mathrm{c}} should depend only on ϕcϕm\phi_{\mathrm{c}}-\phi_{\mathrm{m}}. We find it by taking ϕc=0\phi_{\mathrm{c}}=0, such that Qtsin2ϕt=βmQmsin2ϕmQ_{t}\sin 2\phi_{t}=\beta_{\mathrm{m}}Q_{\mathrm{m}}\sin 2\phi_{\mathrm{m}} and Qtcos2ϕt=βcQc+βmQmcos2ϕm.Q_{t}\cos 2\phi_{t}=\beta_{\mathrm{c}}Q_{\mathrm{c}}+\beta_{\mathrm{m}}Q_{\mathrm{m}}\cos 2\phi_{\mathrm{m}}. This yields

tQc\displaystyle\partial_{t}Q_{\mathrm{c}} =Qc+ρcQtg(Qt)[βcQc+βmQmcos2(ϕcϕm)],\displaystyle=-Q_{\mathrm{c}}+\frac{\rho_{\mathrm{c}}}{Q_{t}}g\left(Q_{t}\right)\left[\beta_{\mathrm{c}}Q_{\mathrm{c}}+\beta_{\mathrm{m}}Q_{\mathrm{m}}\cos 2\left(\phi_{\mathrm{c}}-\phi_{\mathrm{m}}\right)\right],
2Qctϕc\displaystyle 2Q_{\mathrm{c}}\partial_{t}\phi_{\mathrm{c}} =ρcQtg(Qt)βmQmsin2(ϕcϕm),\displaystyle=-\frac{\rho_{\mathrm{c}}}{Q_{t}}g\left(Q_{t}\right)\beta_{\mathrm{m}}Q_{\mathrm{m}}\sin 2\left(\phi_{\mathrm{c}}-\phi_{\mathrm{m}}\right),
tQm\displaystyle\partial_{t}Q_{\mathrm{m}} =ρckQm+k+Qccos2(ϕcϕm),\displaystyle=-\rho_{\mathrm{c}}k_{-}Q_{\mathrm{m}}+k_{+}Q_{\mathrm{c}}\cos 2\left(\phi_{\mathrm{c}}-\phi_{\mathrm{m}}\right),
2Qmtϕm\displaystyle 2Q_{\mathrm{m}}\partial_{t}\phi_{\mathrm{m}} =k+Qcsin2(ϕcϕm).\displaystyle=k_{+}Q_{\mathrm{c}}\sin 2\left(\phi_{\mathrm{c}}-\phi_{\mathrm{m}}\right). (41)

In particular, we find that the relative angle decays according to

2t(ϕcϕm)\displaystyle 2\partial_{t}\left(\phi_{\mathrm{c}}-\phi_{m}\right) =[βmρmqmqcg(Qt)Qt+k+ρcρmqcqm]sin2(ϕcϕm).\displaystyle=-\left[\beta_{\mathrm{m}}\rho_{m}\frac{q_{\mathrm{m}}}{q_{\mathrm{c}}}\frac{g\left(Q_{t}\right)}{Q_{t}}+k_{+}\frac{\rho_{\mathrm{c}}}{\rho_{\mathrm{m}}}\frac{q_{\mathrm{c}}}{q_{\mathrm{m}}}\right]\sin 2\left(\phi_{\mathrm{c}}-\phi_{m}\right). (42)

The first term on the right-hand side is the cellular contribution, while the second is the matrix contribution.
Analysis of the relaxation rates. We focus on sufficiently ordered systems, such that we can neglect the dynamics in the densities and nematic order parameters, and focus only on angular dynamics. In this case, qc=g(Qt)q_{\mathrm{c}}=g(Q_{t}), qc=qmq_{\mathrm{c}}=q_{\mathrm{m}} and ρm=k+/k\rho_{\mathrm{m}}=k_{+}/k_{-}. Eq. (42) then reduces to

2t(ϕcϕm)\displaystyle 2\partial_{t}\left(\phi_{\mathrm{c}}-\phi_{m}\right) =(βmρmβmρm+βcρc+kρc)sin2(ϕcϕm).\displaystyle=-\left(\frac{\beta_{\mathrm{m}}\rho_{\mathrm{m}}}{\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}}+k_{-}\rho_{\mathrm{c}}\right)\sin 2\left(\phi_{\mathrm{c}}-\phi_{m}\right). (43)

The matrix rotation rate is kρck_{-}\rho_{\mathrm{c}}, because it is completely determined by the degradation rate (recall that we have omitted the k0k_{0} terms from our analysis), while the cellular rate is of order kβmρm/(βmρm+βcρc)k\beta_{\mathrm{m}}\rho_{\mathrm{m}}/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right), where we have reintroduced the timescale 1/k1/k.

Arrested domain coarsening. As is discussed in the main text, cellular alignment with the matrix is expected to arrest the coarsening of differently-oriented cellular domains. Explicitly, we expect domains of typical size ll to remain frozen as long as kβmρm/(βmρm+βcρc)Dt/l2,kρck\beta_{\mathrm{m}}\rho_{\mathrm{m}}/\left(\beta_{\mathrm{m}}\rho_{\mathrm{m}}+\beta_{\mathrm{c}}\rho_{\mathrm{c}}\right)\gg D_{t}/l^{2},\,k_{-}\rho_{\mathrm{c}}. Here DtD_{t} is the total translational diffusion coefficient. In our model, it is given by Dt=D+v2/(4k)D_{t}=D+v^{2}/(4k).

We test this prediction by numerically solving the hydrodynamic equations [Eq. (2) of the main text]. We consider initial conditions of fully aligned cells and matrix (qc=qm=1q_{\mathrm{c}}=q_{\mathrm{m}}=1), whose direction alternates along the yy-direction, according to

ϕc=ϕm\displaystyle\phi_{\mathrm{c}}=\phi_{\mathrm{m}} =π2cos2πly.\displaystyle=\frac{\pi}{2}\cos\frac{2\pi}{l}y. (44)

Here, ll is the band width. We integrate the equations for 10410^{4} time steps in the two limits of cell-dominated interaction (βm=0\beta_{\mathrm{m}}=0) and matrix dominated interaction (βc=0\beta_{\mathrm{c}}=0). In accordance with our predictions, the domains coarsen in the cell-dominated case and display negligible dynamics in the matrix-dominated case, as is evident from Fig. 3.

Refer to caption
Figure 3: (Color online) Cellular angles ϕc\phi_{\mathrm{c}} obtained from numerical solutions of Eq. (2) of the main text, 10410^{4} time steps after common initial state (left panel) in the form of alternating bands, according to Eq. (44). We compare between a matrix-dominated case (βc=0,βm=10,\beta_{\mathrm{c}}=0,\beta_{\mathrm{m}}=10, center panel) and cell-dominated case (βc=10,βm=0,\beta_{\mathrm{c}}=10,\beta_{\mathrm{m}}=0, right panel). The values used are: L=200L=200 (system size), D=0.5,D=0.5, ρc0=1\rho_{\mathrm{c}}^{0}=1 (initial cellular density), k+=1,k_{+}=1, and k=0.1.k_{-}=0.1.. Periodic boundary conditions are used. The matrix-dominated snapshot is barely distinguishable from the initial state, while the cell-dominated snapshot displays coarsening.