Uneingeschränkter Zugang

On the stabilizing effect of chemotaxis on bacterial aggregation patterns


Zitieren

Introduction

When grown on semi-solid agar surfaces, many species of bacteria exhibit a variety of complex spatial patterns which depend upon certain environmental conditions, such as the level of nutrients, the hardness of the agar substrate and the temperature. For instance, when concentrations of the bacterium Bacillus subtilis are inoculated on a moderately soft agar plate with poor nutrient levels, the bacteria form dense branching colonies with a smooth envelope propagating outward (cf. Ohgiwari et al. [18]). If the nutrient level is increased and the agar is softer, then the bacteria form simple circular patterns growing in an homogeneous fashion (cf. Wakita et al. [24]).

Continuous deterministic reaction-diffusion systems of equations for the bacterial density and the nutrient concentration appear to be good models for the transition from a dense branch morphology (DBM) regime to the homogeneous envelope front type of pattern (see, e.g., [3, 5, 9, 14, 21]). In the mid nineties, Kawasaki et al. [9] proposed a reaction diffusion model with a non-linear, degenerate, density-dependent, cross-diffusion coefficient, which captures many of the main features of the patterns of B. subtilis found experimentally. Since many aspects of the observed patterns can only be understood if one takes into account the effects of chemotaxis (that is, the tendency of some biological individuals to move preferably toward higher concentrations of chemicals [1, 23]), in a recent contribution Leyva et al. [14] proposed a chemotactic version of the original model by Kawasaki and collaborators, which incorporates a chemotactic toward nutrient concentration term into the equations. This new term is appropriately adapted to the cross diffusion coefficient under consideration, that is, it follows an empirical rule suggested by Ben-Jacob and his group [3, 8] (see section 2.1 below). In [14], the authors present the outcome of numerical simulations and observed that the colony envelope moves faster if the chemotactic signal is present. Moreover, they perform asymptotic estimations on the normal speed of the front as a function of the chemotactic sensitivity. In addition, simulations in the soft-agar, low-nutrient regime indicate a change of morphology from a DBM-like pattern to an homogeneous circular front when the chemotactic signal is increased.

In this paper we provide a quantitative estimation that explains, at first order approximation, this change in morphology. More precisely, we show that when the chemotactic sensitivity is increased, the eigenvalues of the linearized operator around the envelope front become more stable (that is, they move to the stable complex half plane), suggesting, in this fashion, that the homogeneous front is more difficult to destabilize. This mechanism seems to suppress the branching of the colony when the chemotactic signal toward nutrient is stronger. For that purpose, we make some approximations, which include the derivation of an effective scalar equation for the bacterial density (thanks to the balanced production of bacteria and consumption of nutrient), and the approximation of a smooth envelope front by a locally planar traveling front (cf. [7]). We employ energy estimates at the level of the resulting spectral equations which show that the front is spectrally stable under transversal perturbations. These estimates generalize previous calculations for systems with constant diffusion coefficients [2, 4], and make use of recently developed techniques to locate point spectra for stability problems with density-dependent, degenerate diffusivities [15].

The rest of the paper is organized as follows. In section 2 we present the model in [14] and show some particular numerical simulations that exhibit the change in morphology when chemotaxis is switched on. The central section 3 contains the energy estimates, performed on a generic model after proper approximations, which yield the transversal spectral stability of the front, as well as the energy bounds for the eigenvalues in terms of the chemotactic signal. The final section 4 contains a brief discussion on our (and related) results.

The effects of chemotaxis on bacterial aggregration dynamics

In this section we describe in detail a chemotaxis-reaction-diffusion model for bacterial growth, which includes a nonlinear, cross-diffusion term, as well as a chemotactic term toward nutrient. We also present the results of numerical simulations of this system in the DBM-regime, which keep the environmental conditions fixed whereas the chemotactic parameter is increased.

Chemotactic non-linear degenerate cross-diffusion model

Based on the model by Kawasaki et al. [9], Leyva and collaborators [14] have proposed the following chemotactic, non-linear degenerate cross-diffusion system of equations in order to model the dynamics of B. subtilis on thin agar plates:

ut=(σuvu)(ξ(u,v)χ(v)v)+θκuv,vt=DvΔvuv,$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {{u_t}} \hfill & = \hfill & {\nabla \cdot (\sigma uv\nabla u) - \nabla \cdot (\xi (u,v)\chi (v)\nabla v) + \theta \kappa uv,} \hfill \\ {{v_t}} \hfill & = \hfill & {{D_v}{\rm{\Delta }}v - uv,} \hfill \\ \end{array} \end{array}$$

for (x,y) ∊ Ω ⊂ ℝ2, t ∊ (0, +∞). Here Ω is an open, bounded spatial domain and t > 0 denotes time. Scalar functions v = v(x,y,t) and u = u(x,y,t) represent the concentration of nutrients and the density of bacterial cells, respectively. Moreover, k > 0 is a positive constant and θ > 0 is the constant conversion factor.

The constant Dv > 0 is the diffusion coefficient for the nutrient, whereas the diffusion coefficient of bacteria is a non-linear function of the concentrations, as proposed by Kawasaki et al. [9],

Du=Du(u,v)=σuv,$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {D_u} = {D_u}(u,v) = \sigma uv, \end{array} \end{array}$$

where σ > 0 measures the hardness of the agar substrate. This nonlinear diffusion coefficient was proposed to model the experimental observations by Oghiwari et al. [18], which suggest that bacteria are immotile when either the bacterial density u or the nutrient concentration v are low, and become active as u or v increase.

The chemotactic sensitivity function χ = χ(v) follows a Lapidus-Schiller receptor law [13]:

χ(v)=χ0Kd(Kd+v)2;$$\begin{array}{} \displaystyle \chi \left( v \right) = \frac{{{\chi _0}{K_d}}}{{{{\left( {{K_d} + v} \right)}^2}}}; \end{array}$$

here χ0 > 0 is a constant, measuring the strength of the chemotactic signal, and Kd > 0 is the receptor-ligand binding dissociation constant (measured in nutrient concentration units), representing the nutrient level needed for half receptor to be occupied. It depends on the bacterial species under consideration, alone. Finally, the bacterial response function ξ = ξ (u,v), which must be non-negative for attractive chemotaxis, measures the sensitivity of the bacteria to the effective sensed gradient, namely ξ(v)∇v, and it has the same units as a diffusion coefficient. In the DBM-regime of soft agar and nutrient poor medium, Leyva et al. [14] proposed the following bacterial response function:

ξ(u,v)=uDu(u,v)=σu2v.$$\begin{array}{} \displaystyle \xi \left( {u,v} \right) = u{D_u}\left( {u,v} \right) = \sigma {u^2}v. \end{array}$$

This choice obeys the empiric law proposed by E. Ben-Jacob and his group [3, 8] which states that the bacterial response function should be proportional to the product of the bacterial density and its diffusion coefficient:

|ξ|uDu.$$\begin{array}{} \displaystyle \left| \xi \right| \propto u{D_u}. \end{array}$$

The explanation by Ben-Jacob and collaborators goes as follows. Bacterial motion is, necessarily, cooperative at low bacterial densities. This can be roughly modeled by taking the diffusion coefficient to be dependent on bacterial density. Thus, the chemotactic flux should be modified accordingly as Jchem = −ξ(u,v)χ(v)∇v, where ξ follows the empirical rule (5). Notice that when the diffusion coefficient is constant, Du(u,v) = 1 (after normalizations), then the bacterial response function is ξ = u, recovering the standard Keller-Segel [10, 11] chemotactic flux: Jchem = −(v)∇v.

System (1) is further endowed with initial conditions of the form

u(x,y,0)=u0(x,y),(x,y)Ω,$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {u\left( {x,y,0} \right) = {u_0}\left( {x,y} \right),}&{\left( {x,y} \right)} \end{array} \in {\rm{\Omega }}, \end{array}$$

v(x,y,0)=v0,constant,$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {v(x,y,0) = {v_0},} & {{\rm{constant}},} \\ \end{array} \end{array}$$

where v0 > 0 is the uniform initial distribution of nutrient, and the function u0 = u0(x,y) is the initial inoculation of bacteria in χ (for example, a localized Gaussian distribution). In addition, we impose no-flux or Neumann boundary conditions,

u.v=0andv.v=0,at Ω,$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {\begin{array}{*{20}{c}} {\begin{array}{*{20}{c}} {\nabla u.v = 0}&{{\rm{and}}} \end{array}}&{\nabla v.v = 0,} \end{array}}&{{\rm{at\ }}\partial {\rm{\Omega }}} \end{array}, \end{array}$$

where v ∊ ℝ2, |v| = 1, is the unit outer normal at each point of ∂Ω.

After an appropriate scaling of variables (see [14] for details), system (1) can be put in non-dimensional form,

ut=σ0(1+Δ^)(uvv)χ0σ0(1+Δ^)(u2v(1+v)2v)+uv,vt=Δvuv,$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {{u_t}} \hfill & = \hfill & {{\sigma _0}(1 + {\rm{\hat \Delta }})\nabla \cdot (uv\nabla v) - {\chi _0}{\sigma _0}(1 + {\rm{\hat \Delta }})\nabla \cdot (\frac{{{u^2}v}}{{{{(1 + v)}^2}}}\nabla v) + uv,} \hfill \\ {{v_t}} \hfill & = \hfill & {{\rm{\Delta }}v - uv,} \hfill \\ \end{array} \end{array}$$

after suitable modifications for the functions D, ξ and χ (with a slight abuse of notation, we use the same letters for all functions and independent variables). Following Kawasaki el al. [9] (see also [14]), and in order to avoid intrinsic symmetries in the bacterial patterns induced by numerical discretization, in (8) we have considered

D(u,v)=σuv=σ0(1+Δ^)uv,$$\begin{array}{} \displaystyle D\left( {u,v} \right) = \sigma uv = {\sigma _0}\left( {1 + {\rm{\hat \Delta }}} \right)uv, \end{array}$$

where σ0 > 0 is a non-dimensional constant, and Δ^$\begin{array}{} \displaystyle \hat \Delta \end{array}$ is a stochastic fluctuation. Whence, after the scaling, the only non-dimensional free parameters are σ0 > 0, which measures the hardness of the agar substrate (larger values of σ0 for lower agar concentrations); χ0 ≥ 0, which measures the strength of the chemotactic signal; and v0 > 0, the relative initial concentration of nutrient.

Numerical simulations

We now present the results of numerical simulations of system (8), subject to boundary conditions of Neumann type (7), and with initial conditions of form (6a) and (6b). The calculations were performed in a two-dimensional square domain

Ω=(L/2,L/2)×(L/2,L/2),$$\begin{array}{} \displaystyle \Omega = ( - L/2,L/2) \times ( - L/2,L/2), \end{array}$$

with center at the origin, and sides of length L = 680. We employed an explicit Runge-Kutta time stepper and finite differences scheme, on a spatial square grid of N = 4096 points. Thus, the discretization mesh width was taken as Δx = Δy = L/N ≈ 0.166. The time step size considered was Δt = 0.011, for which the method is stable.

For the sake of comparison with the numerical simulations in [9, 14], the initial distribution of bacterial cell density was of Gaussian type, centered at the origin,

u0(x,y)=uMe(x2+y2)/6.25,$$\begin{array}{} \displaystyle u_0(x,y) = u_M e^{-(x^2 + y^2)/6.25} \end{array}$$

where uM = 0.71 is the maximum density at the center. The uniform initial distribution of nutrient was taken as

v0(x,y)v0=0.71,$$\begin{array}{} \displaystyle v_0(x,y) \equiv v_0 = 0.71, \end{array}$$

as considered in Kawasaki et al. [9]. The coefficient σ > 0, that appears in the diffusion for bacteria, was perturbed from a mean value σ0 > 0 via a stochastic fluctuation Δ^$\begin{array}{} \displaystyle {\rm{\hat \Delta }} \end{array}$, following (9). The value for the non-dimensional parameter σ0 was taken as σ0 = 4.0, measuring the softness/hardness of the agar substrate. These parameter values (v0 = 0.71, σ0 = 4.0) correspond to an environment with semi-solid agar (soft medium) and low-to-medium concentration of nutrients, where DBM patterns arise, characterized by smooth envelope fronts moving outward. In the absence of chemotaxis (cf. [9,14]), these conditions produce dense branches (fingering) and ramified patterns.

While fixing these conditions for nutrient and agar substrate, we switched on the chemotactic parameter χ0 ≥ 0. We performed simulations for different parameter values of the chemotactic sensitivity 0. The results of the simulations can be observed in Figure 1. A sequence of three time steps (in rows) of the contour plots for the computed bacterial density are presented. The numerical values for v0 = 0.71 and σ0 = 4.0 are kept fixed in the simulations, whereas, in columns, the values for the chemotactic sensitivity are varied and taken as σ0 = 0, 2.5 and 5.0, respectively. Notice the appearance of a colony envelope front for u, moving outward as time goes by.

Fig. 1

A sequence of three time steps, in rows, of the contour plots of u computed numerically, for three values of χ0, in columns, showing the change of morphology in the bacterial patterns. Notice the stabilizing effect of chemotaxis on the bacteria envelope front.

The first apparent change with respect to absent or weak chemotaxis is that the colony envelope front moves faster outward, as the chemotaxis is increased. This expected behavior has been already quantified in [14], where estimates on the normal speed of the front in terms of the chemotactic signal were established. The other significant change is morphological. Notice that the envelope front goes from a branched (dendrite-like) pattern (σ0 = 0) to a more homogeneous bacterial pattern (σ0 = 5.0), in which the protuberances seem to disappear.

Apparently, the increase in the chemotactic signal produces a transition from the DBM-morphology to the homogeneous disk type morphology which is observed in soft-agar, high-nutrient regimes [3, 9]. The presence of the chemotactic signal seems to supress the onset of instability that causes the formation of branches. This behaviour was already noticed in the simulations performed by Leyva et al. [14]. However, the authors did not provide any justification of this sudden change in morphology of the patterns.

Transversal stability

In this section, we provide a quantitative argument that justifies the change in morphology experienced by the bacterial colony as the intensity of the chemotactic signal is increased, under the same conditions on motility and level of nutrient. For that purpose, let us consider the following generic chemotactic cross-diffusion model:

ut=(D¯(u,v)u)(ξ(u,v)χ(v)v)+uv,vt=Δvuv,$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {{u_t}} \hfill & = \hfill & {\nabla \cdot (\bar D(u,v)\nabla u) - \nabla \cdot (\xi (u,v)\chi (v)\nabla v) + uv,} \hfill \\ {{v_t}} \hfill & = \hfill & {{\rm{\Delta }}v - uv,} \hfill \\ \end{array} \end{array}$$

written in appropriate non-dimensional variables.

In the sequel, we make the following assumptions:

(H1) D¯C2(2)$\begin{array}{} \displaystyle \bar D \in {C^2}\left( {{\mathbb{R}^2}} \right) \end{array}$, D¯(u,v)0$\begin{array}{} \displaystyle \bar D\left( {u,v} \right) \ge 0 \end{array}$ for all u,v ∊ ℝ. Moreover, D¯=0$\begin{array}{} \displaystyle \bar D = 0 \end{array}$ if u = 0 or v = 0.

(H2) For all u,v ∊ ℝ under consideration:

ξ(u,v)=uD¯(u,v).$$\begin{array}{} \displaystyle \xi(u,v) = u \bar{D}(u,v). \end{array}$$

(H3) χC2(ℝ) and uniformly bounded: 0 ≤ χ(v) ≤ C for some constant C > 0 and all v ∊ ℝ under consideration.

Hypothesis (H1) generalizes the nonlinear degenerate diffusion function of Kawasaki et al. [9], given by (2). It is degenerate at u = 0 or v = 0, when either the nutrient or the bacterial density vanishes. Hypothesis (H2) simply expresses that the bacterial response function should satisfy (5), as proposed by Ben-Jacob in the DBM-regime [3,8]. Finally, hypothesis (H3) resembles a receptor-ligand chemotactic signal, which should be bounded thanks to the saturation of receptors, just like the Lapidus-Schiller receptor law (3). Under these assumptions, the chemotactic signal is attractive.

Notice that the kinetic terms in system (10) are of the form ±f (u,v) = ±uv, that is, there is a balance between the production term for bacterial density, and the loss of nutrient concentration. The product uv measures the probability of encounters of bacterial cells with the nutrient.

Approximations

In order to quantify the effects of chemotaxis on the branching patterns, we are going to make some approximations on the solutions to system (10). First, we employ the reduction suggested by Kawasaki et al. [9]. They observed that, due to the form of the reaction terms (production and loss terms are balanced), in the absence of diffusion the total mass is approximately conserved for small values of nutrient and bacterial density. In our model, if we neglect both diffusion and chemotaxis (the latter is also vanishing thanks to the bacterial response under consideration –hypothesis (H3)–), then we add the equations and integrate in time to obtain

u+vC,constant.$$\begin{array}{} \displaystyle u + v \approx C, \quad \text{constant.} \end{array}$$

This constant can be taken as C = v0, the initial nutrient reference value. This approximation is consistent with the numerical simulations of [9] (in the absence of chemotactic signal), and those of [14], where chemotaxis is incorporated. Such simulations show that, at first order, u is constant inside the envelope front and v is near zero (total consumption of resources), whereas away from the front u = 0 and the nutrient is not yet consumed. Thus, we make the following approximation

v=v0u.$$\begin{array}{} \displaystyle v = {v_0} - u. \end{array}$$

Upon substitution into system (10) one obtains a scalar equation for the bacterial density of the form

ut=(D(u)u)+g(u),$$\begin{array}{} \displaystyle {u_t} = \nabla \cdot \left( {D\left( u \right)\nabla u} \right) + g\left( u \right), \end{array}$$

where the effective nonlinear, degenerate, density-dependent diffusion coefficient is given by

D(u):=D(u,v0u)(1+uχ(v0u)),$$\begin{array}{} \displaystyle D\left( u \right){\rm{ : = }}D\left( {u,{v_0} - u} \right)\left( {1 + u\chi \left( {{v_0} - u} \right)} \right), \end{array}$$

and the effective reaction term is

g(u) := v0u(1uv0).$$\begin{array}{} \displaystyle g\left( u \right){\rm{ : = }}{v_0}u\left( {1 - \frac{u}{{{v_0}}}} \right). \end{array}$$

It is to be observed that, under the hypothesis of attractive chemotaxis towards nutrients (H3), the effective diffusion coefficient is a non-negative, density-dependent nonlinear function. Moreover, it is doubly-degenerate, in the sense that it vanishes at u = 0 and at u = v0 (equivalently, at v = 0). In addition, notice that the resulting reaction term (14) is of Fisher-KPP (or logistic) type [6, 12].

Finally, since χ(·) is a C2 function near v = 0, we approximate χ by a constant value,

χ(v)χ(0)=χ00,$$\begin{array}{} \displaystyle \chi(v) \equiv \chi(0) = \chi_0 \geq 0, \end{array}$$

that is, we assume that the chemotactic sensitivity does not vary for v near zero. This simplification allows us to work with the parameter χ0 ≥ 0 of the numerical simulations as a measure of the chemotactic signal, which can be switched on and off. Since for suitably normalized values of v ∊ [0,v0] the Lapidus-Schiller chemotactic law is uniformly bounded, this approximation represents no loss of generality.

Perturbations of the envelope front

Following Funaki et al. [7], we assume that, locally, the envelope front behaves like a planar front, and that its propagation happens in one preferred direction. Thus, let us consider an approximated one-dimensional planar front solution of the form

u(x,y,t)=ϕ(xct)=ϕ(z),$$\begin{array}{} \displaystyle u(x,y,t) = \phi(x-ct) = \phi(z), \end{array}$$

on a strip domain,

ΩL={(x,y)2:<x<,0<y<L},$$\begin{array}{} \displaystyle \Omega_L = \{ (x,y) \in \mathbb{R}^2 \, : \, -\infty < x < \infty, \; \; 0 < y < L \}, \end{array}$$

for some L > 0. Here z := xct denotes the (Galilean) variable of translation, and c ∊ ℝ is the constant speed of the front. Substituting the traveling front solution into (12) we obtain the following ordinary differential equation for the profile function φ,

cϕz=D(ϕ)ϕzz+D(ϕ)zϕz+g(ϕ).$$\begin{array}{} \displaystyle - c{\varphi _z} = D\left( \varphi \right){\varphi _{zz}} + D{\left( \varphi \right)_z}{\varphi _z} + g\left( \varphi \right). \end{array}$$

In pattern formation problems, it is natural to consider infinite domains and to neglect the influence of boundary conditions. Due to finite speed of propagation and the fact that we are interested in the local-in-time, local-in-space evolution near the interface of the envelope front, working on an infinite domain ΩL means no loss of generality. Consequently, we assume that the front has asymptotic limits of the form

u±=limz±ϕ(z).$$\begin{array}{} \displaystyle {u_ \pm } = \mathop {\lim }\limits_{z \to \pm \infty } \phi (z). \end{array}$$

Since the front spreads from the region occupied by the cells toward outer regions filled with nutrient, we consider

u+=0,u=v0.$$\begin{array}{} \displaystyle {u_ + } = 0,{u_ - } = {v_0}. \end{array}$$

Moreover, we shall assume that the front is monotone, that is,

ϕz(z)<0,for all zR.$$\begin{equation} \phi_z(z) \lt 0, \qquad \text{for all } \, z \in \mathbb{R}. \end{equation}$$

Remark 1. The existence of traveling fronts for reaction diffusion equations where the diffusion coefficient is density-dependent and degenerate, and the reaction term is of Fisher-KPP type, as well as its monotonicity properties, have been addressed in the literature by several authors; see, for example, [16, 19]. As a by-product of the existence analysis of fronts for degenerate diffusion coefficients, which decay to zero as they approach equilibria,

D(ϕ)0,as ϕ(z){0,z+v0,z,$$\begin{array}{} \displaystyle D(\phi ) \to 0,\;\;{\rm{as }}{\mkern 1mu} \phi (z) \to \left\{ {\begin{array}{*{20}{l}} {0,}&{z \to + \infty }\\ {{v_0},}&{z \to - \infty ,} \end{array}} \right. \end{array}$$

it turns out that the heteroclinic orbit belongs to a center manifold passing through the degenerate equilibrium points. For instance, when φ(z) → 0 as z → +∞, the center manifold on the phase plane looks like [19, 20],

ϕz=αϕ+O(ϕ2),$$\begin{array}{} \displaystyle {\phi _z} = - \alpha \phi + O({\phi ^2}), \end{array}$$

for some constant α > 0, from which we can deduce the exponential decay of φ and its derivatives,

|ϕ|,|ϕz|,|ϕzz|Ceαz,z+,$$\begin{array}{} \displaystyle |\phi |,|{\phi _z}|,|{\phi _{zz}}| \le C{e^{ - \alpha z}},\quad z \to + \infty , \end{array}$$

as well as the rate of decay of the diffusion coefficient

D(ϕ)=D(0)ϕ+12D(0)ϕ2+=O(ϕ)Ceαz,$$\begin{array}{} \displaystyle D(\phi ) = D'(0)\phi + {\textstyle{1 \over 2}}D''(0){\phi ^2} + \ldots = O(\phi ) \le C{e^{ - \alpha z}}, \end{array}$$

for φ ~ 0. Similar decay rates hold at the other degenerate point, as φv0 when z → -∞. See [15, 16, 19, 20] for further information.

Let us consider perturbations of the form w = w(z,y,t), satisfying the boundary conditions

w(±,y,t)=0,t>0,0<y<L,w(z,L,t)=w(z,0,t)=0,t>0,z.}$$\begin{array}{} \displaystyle \left.\begin{array}{*{20}{r}} {w\left( { \pm \infty ,y,t} \right) = 0,}&{t > 0,0 < y < L,}\\ {w\left( {z,L,t} \right) = w\left( {z,0,t} \right) = 0,}&{t > 0,z \in \mathbb{R}.} \end{array}\right\} \end{array}$$

Substituting u(x,y,t) = φ(z) + w(z,y,t) into (12) and expanding the nonlinear functions we obtain

cϕzcwz+wt=D(ϕ)(wzz+wyy)+(D(ϕ)ϕz)z+2D(ϕ)zwz+D(ϕ)zzw+g(ϕ)+g(ϕ)w+O(|w|2+|w||w|).$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} { - c{\phi _z} - c{w_z} + {w_t} = D(\phi )({w_{zz}} + {w_{yy}}) + {{(D(\phi ){\phi _z})}_z} + 2D{{(\phi )}_z}{w_z} + D{{(\phi )}_{zz}}w + }\\ {g(\phi ) + g'(\phi )w + O(|w{|^2} + \left| w \right|\left| {\nabla w} \right|).} \end{array} \end{array}$$

In view of the profile equation (15) and linearizing around the front, one arrives at the following linear equation for the perturbation

wt=D(ϕ)(wzz+wyy)+(2D(ϕ)z+c)wz+(D(ϕ)zz+g(ϕ))w.$$\begin{array}{} \displaystyle {w_t} = D(\phi )({w_{zz}} + {w_{yy}}) + (2D{(\phi )_z} + c){w_z} + (D{(\phi )_{zz}} + g'(\phi ))w. \end{array}$$

In order to establish the associated spectral problem, let us specialize (20) to solutions of the form

w(z,y,t)=eλtU(z,y),$$\begin{array}{} \displaystyle w(z,y,t) = {e^{\lambda t}}U(z,y), \end{array}$$

where λ ∊ C and UL2L), subject to the following boundary conditions

U(±,y)=0,y[0,L],Uy(z,0)=Uy(z,L)=0,z.}$$\begin{array}{} \displaystyle \left.\begin{array}{*{20}{r}} {U( \pm \infty ,y) = 0,}&{y \in [0,L],}\\ {{U_y}(z,0) = {U_y}(z,L) = 0,}&{z \in \mathbb{R}.} \end{array}\right\} \end{array}$$

Substituting into (20) we obtain the following spectral problem,

λU=D(ϕ)(Uzz+Uyy)+(2D(ϕ)z+c)Uz+(D(ϕ)zz+g(ϕ))U,$$\begin{array}{} \displaystyle \lambda U = D(\phi )({U_{zz}} + {U_{yy}}) + (2D{(\phi )_z} + c){U_z} + (D{(\phi )_{zz}} + g'(\phi ))U, \end{array}$$

where λ ∊ C plays the role of an eigenvalue, and UL2L) is the associated eigenfunction. The right hand side of (22) defines a closed, densely defined operator in L2L), namely

LU=D(ϕ)(Uzz+Uyy)+(2D(ϕ)z+c)Uz+(D(ϕ)zz+g(ϕ))U,L:D(L)=H2(ΩL)L2(ΩL)L2(ΩL),$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {{\cal L}U} \hfill & = \hfill & {D(\phi )({U_{zz}} + {U_{yy}}) + (2D{{(\phi )}_z} + c){U_z} + (D{{(\phi )}_{zz}} + {g^\prime }(\phi ))U,} \hfill \\ {{\cal L}} \hfill & : \hfill & {{\cal D}({\cal L}) = {H^2}({\Omega _L}) \subset {L^2}({\Omega _L}) \to {L^2}({\Omega _L}),} \hfill \\ \end{array} \end{array}$$

with domain 𝒟( ) = H2L). is the linearized operator around the front. For any UL2L) satisfying (21) and any m = 0,1,2,..., let us define

Um(z):=0LU(z,y)Ym(y)dy,$$\begin{array}{} \displaystyle {U_m}(z): = \int_0^L U (z,y){Y_m}(y){\mkern 1mu} dy, \end{array}$$

where

Ym(y)={1L,m=0,2Lcos(mπyL),m=1,2,$$\begin{array}{} \displaystyle {Y_m}(y) = \left\{ {\begin{array}{*{20}{l}} {\frac{1}{{\sqrt L }},}&{m = 0,}\\ {}&{}\\ {\sqrt {\frac{2}{L}} \cos\ (\frac{{m\pi y}}{L}),}&{m = 1,2, \ldots } \end{array}} \right. \end{array}$$

Eigenfunctions are thus decomposed as

U(z,y)=m=0+Um(z)Ym(y).$$\begin{array}{} \displaystyle U(z,y) = \sum\limits_{m = 0}^{ + \infty } {{U_m}} (z){Y_m}(y). \end{array}$$

Substituting into (22) and collecting equal modes for each m ∊ ℤ, m ≥ 0, we establish the following hierarchy of spectral equations,

λU0=D(ϕ)z2U0+(2D(ϕ)z+c)zU0+(D(ϕ)zz+g(ϕ))U0,for m=0,$$\begin{array}{} \displaystyle \lambda {U_0} = D(\phi )\partial _z^2{U_0} + (2D{(\phi )_z} + c){\partial _z}{U_0} + (D{(\phi )_{zz}} + g'(\phi )){U_0},\quad {\rm{for }}{\mkern 1mu} m = 0, \end{array}$$

and,

λUm=D(ϕ)z2Um+(2D(ϕ)z+c)zUm+(D(ϕ)zz+g(ϕ)m2π2L2D(ϕ))Um,  for m,m1,$$\begin{array}{} \displaystyle \lambda {U_m} = D(\phi )\partial _z^2{U_m} + (2D{(\phi )_z} + c){\partial _z}{U_m} + (D{(\phi )_{zz}} + g'(\phi ) - \frac{{{m^2}{\pi ^2}}}{{{L^2}}}D(\phi )){U_m},\qquad {\rm{for }}{\mkern 1mu} m \in ,{\mkern 1mu} m \ge 1, \end{array}$$

for the same eigenvalue λ ∊ ℂ. We are interested in solutions UmL2(ℝ) to (23) and (24), for all m ≥ 0. The right hand sides of both equations are closed, densely defined differential operators of second order in L2(ℝ), with domain H2(ℝ).

Remark 2. Observe that, in view of the profile equation (15), the function Φ := φzH2(ℝ) is a solution to (23) with λ = 0 and m = 0. Indeed, if we differentiate (15) we obtain

0=D(ϕ)Φzz+(2D(ϕ)z+c)Φz+(D(ϕ)zz+g(ϕ))Φ.$$\begin{array}{} \displaystyle 0 = D(\phi) \Phi_{zz} + \big( 2D(\phi)_z + c \big) \Phi_z + \big( D(\phi)_{zz} + g'(\phi) \big) \Phi. \end{array}$$

This shows that λ = 0 is an eigenvalue (associated to translation) of (22) with eigenfunction U(z,y) = Φ(z) = φz(z). Clearly, its expansion is given by U0=Lϕz$\begin{array}{} \displaystyle {U_0} = \sqrt L {\phi _z} \end{array}$, and

Um(z)=0Lϕz(z)Ym(y)dy=0,$$\begin{array}{} \displaystyle {U_m}(z) = \int_0^L {{\phi _z}} (z){Y_m}(y){\mkern 1mu} dy = 0, \end{array}$$

for all m ≥ 1, in that case.

Energy estimates and transversal stability

Following [15], for each mode m ∊ ℤ, m ≥ 0, let us define the change of variables

Wm(z):=Um(z)exp(c2z0zdζD(ϕ(ζ))),$$\begin{array}{} \displaystyle {W_m}(z): = {U_m}(z)\exp (\frac{c}{2}\int_{{z_0}}^z {\frac{{d\zeta }}{{D(\phi (\zeta ))}}} ), \end{array}$$

where z0 ∊ ℝ is fixed but arbitrary. Let us suppose that the eigenfunction UH2L) ⊂ L2L) decays sufficiently fast to zero as z → ±∞ so that, for each m ≥ 1,

WmH2().$$\begin{array}{} \displaystyle W_m \in H^2(\mathbb{R}). \end{array}$$

Remark 3. This assumption simply implies that we restrict the analysis to the set of localized perturbations that decay fast enough. At the level of the spectral problem, by a standard cut-off argument one can always approximate any generic localized perturbation by rapidly decaying ones; at the level of the evolution equations, we are simply restricting the space of controlled perturbations to a smaller space with fast decay.

Hence, we can compute

zUm=exp(c2z0zdζD(ϕ(ζ)))(zWmc2D(ϕ)Wm),z2Um=exp(c2z0zdζD(ϕ(ζ)))(z2Wmc2D(ϕ)zWm+(cD(ϕ)z2D(ϕ)2+c24D(ϕ)2)Wm),$$\begin{array}{} \displaystyle \begin{array}{*{20}{c}} {{\partial _z}{U_m} = \exp (\frac{c}{2}\smallint _{{z_0}}^z\frac{{d\zeta }}{{D(\phi (\zeta ))}})\left( {{\partial _z}{W_m} - \frac{c}{{2D(\phi )}}{W_m}} \right),}\\ {\partial _z^2{U_m} = \exp (\frac{c}{2}\smallint _{{z_0}}^z\frac{{d\zeta }}{{D(\phi (\zeta ))}})\left( {\partial _z^2{W_m} - \frac{c}{{2D(\phi )}}{\partial _z}{W_m} + (\frac{{cD{{(\phi )}_z}}}{{2D{{(\phi )}^2}}} + \frac{{{c^2}}}{{4D{{(\phi )}^2}}}){W_m}} \right),} \end{array} \end{array}$$

and substitute into the spectral equation (24) to obtain

λWm=D(ϕ)z2Wm+2D(ϕ)zzWm+(H(z)m2π2L2D(ϕ))Wm,$$\begin{array}{} \displaystyle \lambda {W_m} = D(\phi )\partial _z^2{W_m} + 2D{(\phi )_z}{\partial _z}{W_m} + (H(z) - \frac{{{m^2}}}{{{\pi ^2}{L^2}}}D(\phi )){W_m}, \end{array}$$

for all m ∊ ℤ, m ≥ 0 and the eigenvalue λ ∊ ℂ associated to the perturbation. Here

H(z):=D(ϕ)zz+g(ϕ)c24D(ϕ)cD(ϕ)z2D(ϕ).$$\begin{array}{} \displaystyle H(z): = D{(\phi )_{zz}} + g'(\phi ) - \frac{{{c^2}}}{{4D(\phi )}} - \frac{{cD{{(\phi )}_z}}}{{2D(\phi )}}. \end{array}$$

The following lemma states that any eigenvalue is real and non-positive, that is, the front is spectrally stable under transversal perturbations.

Lemma 1

If λ ∊ ℂ is an eigenvalue of the associated spectral problem, such that (25) holds for each mode m ∊ ℤ, m ≥ 0 with WmH2(ℝ), then λ ∊ ℝ and λ ≤ 0.

Proof. When m = 0 and Φ = φz, the change of variables for this eigenfunction, namely,

Ψ(z):=Φ(z)exp(c2z0zdζD(ϕ(ζ))),$$\begin{array}{} \displaystyle \Psi (z): = \Phi (z)\exp (\frac{c}{2}\int_{{z_0}}^z {\frac{{d\zeta }}{{D(\phi (\zeta ))}}} ), \end{array}$$

yields

0=D(ϕ)z2Ψ+2D(ϕ)zzΨ+H(z)Ψ.$$\begin{array}{} \displaystyle 0 = D(\phi )\partial _z^2\Psi + 2D{(\phi )_z}{\partial _z}\Psi + H(z)\Psi . \end{array}$$

Multiply (26) and (25) by D(φ) ≥ 0 and rearrange the terms to obtain

λD(ϕ)Wm=z(D(ϕ)2zWm)+D(ϕ)H(z)Wmm2π2L2D(ϕ)2Wm,$$\begin{array}{} \displaystyle \lambda D(\phi ){W_m} = {\partial _z}(D{(\phi )^2}{\partial _z}{W_m}) + D(\phi )H(z){W_m} - \frac{{{m^2}}}{{{\pi ^2}{L^2}}}D{(\phi )^2}{W_m}, \end{array}$$

for all m ≥ 1, and

0=z(D(ϕ)2Ψz)+D(ϕ)H(z)Ψ.$$\begin{array}{} \displaystyle 0 = {\partial _z}(D{(\phi )^2}{\Psi _z}) + D(\phi )H(z)\Psi . \end{array}$$

Notice that the zero-eigenfunction Φ (and its corresponding transformed eigenfunction Ψ) are independent of the eigenfunction U associated to the eigenvalue λ under consideration. Nonetheless, equation (28) holds, with the same coefficient H(·) as in (27). Whence, from monotonicity of the front (see (16)), we have that Φ = φz ≠ 0 and, consequently, Ψ ≠ 0. Therefore, we can substitute

D(ϕ)H(z)=(D(ϕ)2Ψz)zΨ,$$\begin{array}{} \displaystyle D(\phi )H(z) = - \frac{{{{(D{{(\phi )}^2}{\Psi _z})}_z}}}{\Psi }, \end{array}$$

into (27) to arrive at

λD(ϕ)Wm=(D(ϕ)2zWm)z(D(ϕ)2ΨzΨ+m2π2L2D(ϕ)2)Wm.$$\begin{array}{} \displaystyle \lambda D(\phi ){W_m} = {(D{(\phi )^2}{\partial _z}{W_m})_z} - \left( {\frac{{D{{(\phi )}^2}{\Psi _z}}}{\Psi } + \frac{{{m^2}}}{{{\pi ^2}{L^2}}}D{{(\phi )}^2}} \right){W_m}. \end{array}$$

Take the L2 product of Wm with last equation and integrate by parts. The result is

λ+D(ϕ)|Wm|2dz=+W¯m(D(ϕ)2zWm)zdz+(D(ϕ)2Ψz)zΨ|Wm|2dz+m2π2L2+D(ϕ)2|Wm|2dz=+D(ϕ)2(Ψzz(|Wm|2Ψ)|zWm|2m2π2L2|Wm|2)dz=+D(ϕ)2Ψ2|z(WmΨ)|2dzm2π2L2+D(ϕ)2|Wm|2dz0,$$\begin{array}{} \displaystyle \begin{array}{*{20}{l}} {\lambda \int_{ - \infty }^{ + \infty } D (\phi )|{W_m}{|^2}{\mkern 1mu} dz}&{ = \int_{ - \infty }^{ + \infty } {{{\bar W}_m}} {{(D{{(\phi )}^2}{\partial _z}{W_m})}_z}{\mkern 1mu} dz - \int_{ - \infty }^{ + \infty } {\frac{{{{(D{{(\phi )}^2}{\Psi _z})}_z}}}{\Psi }} |{W_m}{|^2}{\mkern 1mu} dz + }\\ {}&{\; - \frac{{{m^2}}}{{{\pi ^2}{L^2}}}\int_{ - \infty }^{ + \infty } D {{(\phi )}^2}|{W_m}{|^2}{\mkern 1mu} dz}\\ {}&{ = \int_{ - \infty }^{ + \infty } D {{(\phi )}^2}\left( {{\Psi _z}{\partial _z}(\frac{{|{W_m}{|^2}}}{\Psi }) - |{\partial _z}{W_m}{|^2} - \frac{{{m^2}}}{{{\pi ^2}{L^2}}}|{W_m}{|^2}} \right){\mkern 1mu} dz}\\ {}&{ = - {\mkern 1mu} \int_{ - \infty }^{ + \infty } D {{(\phi )}^2}{\Psi ^2}{{\left| {{\partial _z}(\frac{{{W_m}}}{\Psi })} \right|}^2}{\mkern 1mu} dz\; - \frac{{{m^2}}}{{{\pi ^2}{L^2}}}\int_{ - \infty }^{ + \infty } D {{(\phi )}^2}|{W_m}{|^2}{\mkern 1mu} dz}\\ {}&{ \le 0,} \end{array} \end{array}$$

for any m ≥ 1. We conclude that λ is real and non-positive. □

Remark 4. Lemma 1 means that, at first order approximation (locally planar front for a scalar equation due to the balanced source and loss kinetic terms), the envelope front is stable under transversal small, local-in-space perturbations. This behaviour is verified by the numerical calculation of the actual (curved) envelope front.

Remark 5. The result of Lemma 1 can be extrapolated to a whole family of operators indexed by the transversal Fourier frequencies. Indeed, if we consider the eigenvalue problem (22) on L2(ℝ2) and eigenfunctions of the form U = e W (z) with WL2(ℝ) then we obtain the following family of linear operators

LξW=D(ϕ)Wzz+(2D(ϕ)z+c)Wz+(D(ϕ)zz+g(ϕ)ξ2D(ϕ))W,Lξ:L2L2,$$\begin{array}{} \displaystyle \begin{array}{*{20}{r,l}} {{{\cal L}_\xi }W} \hfill & = \hfill & {D(\phi ){W_{zz}} + (2D{{(\phi )}_z} + c){W_z} + (D{{(\phi )}_{zz}} + {g^\prime }(\phi ) - {\xi ^2}D(\phi ))W,} \hfill \\ {{{\cal L}_\xi }} \hfill & : \hfill & {{L^2} \to {L^2},} \hfill \\ \end{array} \end{array}$$

for each ξ ∊ ℝ. A similar analysis (under similar decay assumptions) leads to spectral stability of each operator ξ, with ξ ∊ ℝ fixed, generalizing in this fashion the observation by Butanda [4] in the constant diffusion case.

The stabilizing effect of chemotaxis

Let us define

η(z):=D(ϕ(z))0,  z,$$\begin{array}{} \displaystyle \eta (z): = \sqrt {D(\phi (z))} \ge 0,\qquad z \in \mathbb{R}, \end{array}$$

and the customary weighted energy function spaces

Hηk(;)={v:η(z)v(z)Hk(;)},$$\begin{array}{} \displaystyle H_\eta ^k(\mathbb{R};\mathbb{C}) = \{ v{\mkern 1mu} :{\mkern 1mu} \eta (z)v(z) \in {H^k}(\mathbb{R};\mathbb{C})\} , \end{array}$$

for k ∊ ℤ, k ≥ 0, which are Hilbert spaces endowed with the inner product (and norm),

u,vHηk:=ηu,ηvHk,  vHηk2:=ηvHk2=v,vHηk.$$\begin{array}{} \displaystyle {\langle u,v\rangle _{H_\eta ^k}}: = {\langle \eta u,\eta v\rangle _{{H^k}}},\quad \quad \left\| v \right\|_{H_\eta ^k}^2: = \left\| {\eta v} \right\|_{{H^k}}^2 = {\langle v,v\rangle _{H_\eta ^k}}. \end{array}$$

According to custom, we denote Hη0(;)=Lη2(;)$\begin{array}{} \displaystyle H_\eta ^0(\mathbb{R};\mathbb{C}) = L_\eta ^2(\mathbb{R};\mathbb{C}) \end{array}$.

Let us suppose that λ ∊ (−∞,0] is an eigenvalue associated to an eigenfunction U = ΣUmYm Φ H2L) of problem (22). Let m ∊ ℝ, m ≥ 1 be a single mode for which Um ≠ 0. Consequently, Wm ≠ 0 and WmLη2>0$\begin{array}{} \displaystyle {\left\| {{W_m}} \right\|_{L_\eta ^2}} > 0 \end{array}$. (The modes m ≥ 1 for which Um ≡ 0 do not contribute to the energy estimate.) Then from equation (29) and Lemma 1 we have that

|λ|+D(ϕ)|Wm|2dz=+D(ϕ)2|zWm|2dz+m2π2L2+D(ϕ)2|Wm|2dz+J,$$\begin{aligned} |\lambda| \int_{-\infty}^{+\infty} D(\phi) |W_m|^2 \, dz &= \int_{-\infty}^{+\infty} D(\phi)^2 |\partial_z W_m|^2 \, dz + \frac{m^2}{\pi^2 L^2} \int_{-\infty}^{+\infty} D(\phi)^2 |W_m|^2 \, dz + J, \end{aligned}$$

where

J:=+D(ϕ)2Ψzz(|Wm|2Ψ)dz=+2D(ϕ)D(ϕ)ϕzΨzΨ|Wm|2dz++D(ϕ)2ΨzzΨ|Wm|2dz,$$\begin{array}{} \displaystyle J: = - \int_{ - \infty }^{ + \infty } D {(\phi )^2}{\Psi _z}{\partial _z}(\frac{{|{W_m}{|^2}}}{\Psi }){\mkern 1mu} dz = \int_{ - \infty }^{ + \infty } 2 D(\phi )D'(\phi ){\phi _z}\frac{{{\Psi _z}}}{\Psi }|{W_m}{|^2}{\mkern 1mu} dz{\mkern 1mu} + {\mkern 1mu} \int_{ - \infty }^{ + \infty } D {(\phi )^2}\frac{{{\Psi _{zz}}}}{\Psi }|{W_m}{|^2}{\mkern 1mu} dz, \end{array}$$

after integration by parts.

Lemma 2

The functions D′(φ)φzΨzand D(φzzare uniformly bounded in z ∊ ℝ.

Proof. Follows from the definition of Y, the rates of decay (17) and (18), and straightforward computations. □

In view of last lemma we have

|J|C+D(ϕ)|Wm|2dz=CWmLη22,$$\begin{array}{} \displaystyle \left| J \right| \le C\int_{ - \infty }^{ + \infty } {D(\phi )} {\left| {{W_m}} \right|^2}dz = C\left\| {{W_m}} \right\|_{L_\eta ^2}^2, \end{array}$$

for some uniform C > 0. Substituting back yields the estimate

|λ|WmLη22(supzD(ϕ))(zWmLη22+m2π2L2WmLη22)+CWmLη22.$$\begin{array}{} \displaystyle \left| \lambda \right|\left\| {{W_m}} \right\|_{L_\eta ^2}^2 \le (\mathop {\sup }\limits_{z \in \mathbb{R}} D(\phi ))(\left\| {{\partial _z}{W_m}} \right\|_{L_\eta ^2}^2 + \frac{{{m^2}}}{{{\pi ^2}{L^2}}}\left\| {{W_m}} \right\|_{L_\eta ^2}^2) + C\left\| {{W_m}} \right\|_{L_\eta ^2}^2. \end{array}$$

This estimate holds for each fixed eigenvalue λ and all values of m ≥ 1. It provides an upper bound for |λ| in terms of m (recall that |λ| = −λ > 0, in view of Lemma (1)). Notice that the effective diffusion coefficient D(φ) (that is, the weight of the Sobolev norms) depends on the intensity of the chemotactic signal χ0 as well. Therefore, we normalize with respect to the weighted L2-norm to obtain

|λ|ρmCm(supzD(ϕ))+C,$$\begin{array}{} \displaystyle |\lambda | \le {\rho _m}{C_m}(\mathop {\sup }\limits_{z \in \mathbb{R}} D(\phi )) + C, \end{array}$$

where

ρm:=WmHη12WmLη22>0,$$\begin{array}{} \displaystyle {\rho _m}: = \frac{{\left\| {{W_m}} \right\|_{H_\eta ^1}^2}}{{\left\| {{W_m}} \right\|_{L_\eta ^2}^2}} > 0, \end{array}$$

and Cm > 0 is a constant depending on m, satisfying

CmC1+m2C2,$$\begin{array}{} \displaystyle {C_m} \le {C_1} + {m^2}{C_2}, \end{array}$$

for some uniform constants C1,C2 > 0.

Lemma 3

For anyuHη1(;)$\begin{array}{} \displaystyle u \in H_\eta ^1(\mathbb{R};\mathbb{C}) \end{array}$withuLη2>0$\begin{array}{} \displaystyle {\left\| u \right\|_{L_\eta ^2}} > 0 \end{array}$, the ratio

ρ=uHη12uLη22$$\begin{array}{} \displaystyle \rho = \frac{{\left\| u \right\|_{H_\eta ^1}^2}}{{\left\| u \right\|_{L_\eta ^2}^2}} \end{array}$$

is a uniformly bounded function of χ0 > 0.

Proof. Since D(·) is uniformly bounded from above, it is clear that HηkHk$\begin{array}{} \displaystyle H_\eta ^k \subset {H^k} \end{array}$ for all k ≥ 0. The conclusion follows by noticing that both the numerator and the denominator are linear on χ0,

ρ=uHη112+χ0uHη222uLη122+χ0uLη222=O(1),$$\begin{array}{} \displaystyle \rho = \frac{{\left\| u \right\|_{H_{{\eta _1}}^1}^2 + {\chi _0}\left\| u \right\|_{H_{{\eta _2}}^2}^2}}{{\left\| u \right\|_{L_{{\eta _1}}^2}^2 + {\chi _0}\left\| u \right\|_{L_{{\eta _2}}^2}^2}} = O(1), \end{array}$$

for all χ0 > 0 and where the weight functions η1, η2 are given by

η1=σ0v0ϕ(1ϕ/v0),η2=σ0v0ϕ2(1ϕ/v0).$$\begin{array}{} \displaystyle {\eta _1} = \sqrt {{\sigma _0}{v_0}\phi (1 - \phi /{v_0})} ,\quad {\eta _2} = \sqrt {{\sigma _0}{v_0}{\phi ^2}(1 - \phi /{v_0})} . \end{array}$$

Therefore, estimate (30) shows that the bounds for the eigenvalues and their dependence on the intensity of the chemotactic signal are controlled by the function

supzD(ϕ)=maxu[0,v0]D(u)=maxu[0,v0](σ0v0u(1uv0)(1+χ0u)).$$\begin{array}{} \displaystyle \mathop {\sup }\limits_{z \in \mathbb{R}} D(\phi ) = {\max _{u \in [0,{v_0}]}}D(u) = {\max _{u \in [0,{v_0}]}}({\sigma _0}{v_0}u(1 - \frac{u}{{{v_0}}})(1 + {\chi _0}u)). \end{array}$$

Let

G(u):=u(1uv0)(1+χ0u),  u[0,v0].$$\begin{array}{} \displaystyle G(u): = u(1 - \frac{u}{{{v_0}}})(1 + {\chi _0}u),\qquad u \in [0,{v_0}]. \end{array}$$

After straightforward calculations one finds that, for positive values of the chemotactic sensitivity χ0 > 0, the maximum of the cubic polynomial G on the interval [0,v0] occurs at

u*(v0,χ0)=13(v01χ0+(v01χ0)2+3v0χ0)(0,v0).$$\begin{array}{} \displaystyle {u_*}({v_0},{\chi _0}) = \frac{1}{3}\left( {{v_0} - \frac{1}{{{\chi _0}}} + \sqrt {{{({v_0} - \frac{1}{{{\chi _0}}})}^2} + \frac{{3{v_0}}}{{{\chi _0}}}} {\mkern 1mu} } \right) \in (0,{v_0}). \end{array}$$

Thus, the maximum of G (and hence the energy bound for |λ| in (30)) is controlled by

Θ(v0,χ0):=G(u*(v0,χ0))=u*(v0,χ0)(1u*(v0,χ0)v0)(1+χ0u*(v0,χ0)).$$\begin{array}{} \displaystyle \Theta ({v_0},{\chi _0}): = G({u_*}({v_0},{\chi _0})) = {u_*}({v_0},{\chi _0})(1 - \frac{{{u_*}({v_0},{\chi _0})}}{{{v_0}}})(1 + {\chi _0}{u_*}({v_0},{\chi _0})). \end{array}$$

By standard calculus tools, one easily verifies that, for fixed values of the initial nutrient concentration v0, the bound Θ is an increasing function of the chemotactic signal χ0 ≥ 0. Figure 2 shows this behaviour of Θ as a function of χ0, for different fixed values of v0 = 0.5,0.75,1.0,1.5.

Fig. 2

Plot of Θ(v0, χ0) as a function of χ0 ≥ 0 for different fixed values of v0 = 0.5,0.75,1.0,1.5 (color plot online).

Whence, in view of Lemma 1 and (30), we have

0>λ>(C+ρmCmσ0v0Θ(v0,χ0)).$$\begin{array}{} \displaystyle 0 > \lambda > - (C + {\rho _m}{C_m}{\sigma _0}{v_0}\Theta ({v_0},{\chi _0})). \end{array}$$

The right hand side of (31) is a negative, decreasing function of χ0 ∊ (0,+∞) for each fixed value of the nutrient level v0. In lay terms, this shows that for greater values of χ0 ≥ 0, the eigenvalues of the linearized operator around the front are allowed to be more negative. Since λ is actually the in-time growth rate of perturbations of the form w = eλtU, this suggests that λ is further pushed to the negative side as the chemotactic signal is increased, delaying the formation of unstable patterns (fingering) around the envelope front. To sum up, although the envelope front moves faster due to chemotaxis, the spectral bounds for the eigenvalues of the linearized problem actually decrease as functions of the chemotactic sensitivity χ0, demanding greater energy to make the patterns transversally unstable. Finally, it is to be observed that our estimation is elementary, in spite of working with a density-dependent, degenerate, nonlinear diffusion coefficient.

Discussion

In this paper we have shown that, for a generic nonlinear, degenerate cross-diffusion, chemotactic model of the form (10), underlying balanced kinetic terms, and under the assumptions of approximately constant chemotactic sensitivity, together with a scalar conservation law approximation that is suggested by the balanced source/loss terms, envelope traveling fronts (which are, at first order, locally planar in space) are spectrally stable under transversal perturbations. More precisely, the eigenvalues associated to the perturbations in both space dimensions are real and non-negative. This is shown using an appropriate change of variables and by performing energy estimates on the perturbation (spectral) equations. These energy estimates provide, in addition, bounds for the eigenvalues of the linearized operator around the front. Elementary calculations show that these bounds decrease as functions of the chemotactic sensitivity χ0 ≥ 0, suggesting that, as χ0 grows, the patterns are more stable.

Of course, these observations apply to a first order approximation of the general problem. The crucial approximation is to recast the stability problem of the front as an equivalent one for an effective scalar equation. This is possible thanks to the balanced terms of production of bacterial cells and loss of nutrient concentration, which induce a conservation law for the total density. As a consequence, the resulting effective nonlinear density-dependent diffusion coefficient for a scalar reaction-diffusion equation of Fisher-KPP type is also dependent on the chemotactic sensitivity parameter, pushing the interface to be deformable on shorter scales than the new diffusion length as the chemotactic signal is increased.

Our quantitative observations are consistent with the analysis of Arouh and Levine [2], who showed that chemotaxis suppresses the instability of fronts for a Kessler-Levine system with constant diffusivities and nutrient chemotactic terms. Their analysis applies to balanced source/loss kinetic terms of cut-off type. It is to be observed that the method employed here is elementary and applicable to density-dependent diffusion coefficients. In contrast, the method of Arouh and Levine applies (apparently) to constant diffusion models only, as they compute asymptotic expansions of the eigenvalues of the Laplace operator. Both theirs and our results show the supression of colony branching when the bacteria are assisted chemotactically toward nutrients, a fact that has been confirmed numerically in other systems, such as models for the aggregation of Paenibacillus dendritiformis (see, e.g., [22]) containing balanced kinetic terms as well.

This stabilizing effect of chemotaxis appears to contradict the results of Funaki et al. [7], which show that when the chemotactic signal is present traveling front solutions are destabilized. Indeed, the authors actually show that the transversal instability depends upon wave numbers which, as they increase, produce unstable eigenvalues of a linearized problem, provided that the chemotactic sensitivity function is uniformly convex. The model considered by Funaki and collaborators, however, is a reaction-diffusion-chemotaxis system for which the production and degradation terms are not balanced (unlike those in (10) and in the systems considered in [2,22]). These reaction terms modify the form of the linearized operator around the wave and hence, its stability. It is well known that balanced kinetic terms cause stable wave pinning, even in the case of reaction-diffusion systems without chemotaxis (see, for example, [17]). Therefore, we conjecture that this robustness is lost in the system studied by Funaki and collaborators, causing branching instabilities.

eISSN:
2444-8656
Sprache:
Englisch
Zeitrahmen der Veröffentlichung:
Volume Open
Fachgebiete der Zeitschrift:
Biologie, andere, Mathematik, Angewandte Mathematik, Allgemeines, Physik