Processing math: 56%

Sharp interface limit in a phase field model of cell motility

  • Received: 01 March 2017 Revised: 01 August 2017
  • Fund Project: The work of LB was supported by NSF grants DMS-1106666 and DMS-1405769. The work of VR and MP was partially supported by NSF grant DMS-1106666.
  • 35Q92, 35K51, 35B25

  • We consider a phase field model of cell motility introduced in [40] which consists of two coupled parabolic PDEs. We study the asymptotic behavior of solutions in the limit of a small parameter related to the width of the interface (sharp interface limit). We formally derive an equation of motion of the interface, which is mean curvature motion with an additional nonlinear term. In a 1D model parabolic problem we rigorously justify the sharp interface limit. To this end, a special representation of solutions is introduced, which reduces analysis of the system to a single nonlinear PDE that describes the interface velocity. Further stability analysis reveals a qualitative change in the behavior of the system for small and large values of the coupling parameter. Using numerical simulations we also show discontinuities of the interface velocity and hysteresis. Also, in the 1D case we establish nontrivial traveling waves when the coupling parameter is large enough.

    Citation: Leonid Berlyand, Mykhailo Potomkin, Volodymyr Rybalko. Sharp interface limit in a phase field model of cell motility[J]. Networks and Heterogeneous Media, 2017, 12(4): 551-590. doi: 10.3934/nhm.2017023

    Related Papers:

    [1] Leonid Berlyand, Mykhailo Potomkin, Volodymyr Rybalko . Sharp interface limit in a phase field model of cell motility. Networks and Heterogeneous Media, 2017, 12(4): 551-590. doi: 10.3934/nhm.2017023
    [2] Antonio DeSimone, Martin Kružík . Domain patterns and hysteresis in phase-transforming solids: Analysis and numerical simulations of a sharp interface dissipative model via phase-field approximation. Networks and Heterogeneous Media, 2013, 8(2): 481-499. doi: 10.3934/nhm.2013.8.481
    [3] Shi Jin, Min Tang, Houde Han . A uniformly second order numerical method for the one-dimensional discrete-ordinate transport equation and its diffusion limit with interface. Networks and Heterogeneous Media, 2009, 4(1): 35-65. doi: 10.3934/nhm.2009.4.35
    [4] Serge Nicaise, Cristina Pignotti . Asymptotic analysis of a simple model of fluid-structure interaction. Networks and Heterogeneous Media, 2008, 3(4): 787-813. doi: 10.3934/nhm.2008.3.787
    [5] Victor A. Eremeyev . Anti-plane interfacial waves in a square lattice. Networks and Heterogeneous Media, 2025, 20(1): 52-64. doi: 10.3934/nhm.2025004
    [6] Clément Cancès . On the effects of discontinuous capillarities for immiscible two-phase flows in porous media made of several rock-types. Networks and Heterogeneous Media, 2010, 5(3): 635-647. doi: 10.3934/nhm.2010.5.635
    [7] Patrick W. Dondl, Michael Scheutzow . Positive speed of propagation in a semilinear parabolic interface model with unbounded random coefficients. Networks and Heterogeneous Media, 2012, 7(1): 137-150. doi: 10.3934/nhm.2012.7.137
    [8] Mauro Garavello, Benedetto Piccoli . Coupling of microscopic and phase transition models at boundary. Networks and Heterogeneous Media, 2013, 8(3): 649-661. doi: 10.3934/nhm.2013.8.649
    [9] Markus Gahn, Maria Neuss-Radu, Peter Knabner . Effective interface conditions for processes through thin heterogeneous layers with nonlinear transmission at the microscopic bulk-layer interface. Networks and Heterogeneous Media, 2018, 13(4): 609-640. doi: 10.3934/nhm.2018028
    [10] Fadia Bekkal-Brikci, Giovanna Chiorino, Khalid Boushaba . G1/S transition and cell population dynamics. Networks and Heterogeneous Media, 2009, 4(1): 67-90. doi: 10.3934/nhm.2009.4.67
  • We consider a phase field model of cell motility introduced in [40] which consists of two coupled parabolic PDEs. We study the asymptotic behavior of solutions in the limit of a small parameter related to the width of the interface (sharp interface limit). We formally derive an equation of motion of the interface, which is mean curvature motion with an additional nonlinear term. In a 1D model parabolic problem we rigorously justify the sharp interface limit. To this end, a special representation of solutions is introduced, which reduces analysis of the system to a single nonlinear PDE that describes the interface velocity. Further stability analysis reveals a qualitative change in the behavior of the system for small and large values of the coupling parameter. Using numerical simulations we also show discontinuities of the interface velocity and hysteresis. Also, in the 1D case we establish nontrivial traveling waves when the coupling parameter is large enough.



    The problem of cell motility has been a classical subject in biology for several centuries. It dates back to the celebrated discovery by van Leeuwenhoek in the 17th century who drastically improved the microscope to the extent that he was able to observe motion of single celled organisms that moved due to contraction and extension. Three centuries later this problem continues to attract the attention of biologists, biophysicists and, more recently, applied mathematicians. A comprehensive review of the mathematical modeling of cell motility can be found in [27].

    This work is motivated by the problem of motility (crawling motion) of eukaryotic cells on substrates. The network of actin (protein) filaments (which is a part of the cytoskeleton in such cells) plays an important role in cell motility. We are concerned with cell shape dynamics, caused by extension of the front of the cell due to polymerization of the actin filaments and contraction of the back of the cell due to detachment of these filaments. Modeling of this process in full generality is at present a formidable challenge because several important biological ingredients (e.g., regulatory pathways [27]) are not yet well understood.

    In recent biophysical studies several simplified phase field models of cell motility have been proposed. Simulations performed for these models demonstrated good agreement with experiments (e.g., [40,38] and references therein). Recall that phase field models are typically used to describe the evolution of an interface between two phases (e.g., solidification or viscous fingering). The key ingredient of such models is an auxiliary scalar field, which takes two different values in domains describing the two phases (e.g., 1 and 0) with a diffuse interface of a small width. An alternative approach to cell motility involving free boundary problems is developed in [22,34,5,33,32].

    We consider the coupled system of parabolic PDEs, which is a modified version of the model from [40] in the diffusive scaling (tε2t, xεx):

    ρεt=Δρε1ε2W(ρε)Pερε+λε(t) in  Ω, (1)
    Pεt=εΔPε1εPεβρεin Ω, (2)

    where

    λε(t)=1|Ω|Ω(1ε2W(ρε)+Pερε)dx. (3)

    The unknowns here are the scalar phase field function ρε and the orientation vector Pε; Ω is a bounded domain in R2, λε is the Lagrange multiplier responsible for preservation of volume. We study solutions of system (1)-(3) in the sharp interface limit, when the parameter ε>0 (which is, loosely speaking, the width of the interface) tends to zero.

    While system (1)-(3) represents a modified version of the model from [40], the main features of the original model are preserved. First, the volume preservation constraint in [40] is imposed by introducing a penalization parameter into the double well potential. In (1)-(2) the volume preservation is enforced by the (dynamic) Lagrange multiplier λε given by (3). Both ways of introducing volume preservation are equivalent in the sharp interface limit, see [1,2,8]. Second, for technical simplicity we dropped two terms in the original equation for the orientation field P. One of them, responsible for a stronger damping in the phase ρε0, can be added to (2) without any qualitative changes, while the second one, the so-called γ-term, leads to an enormous technical complication, even existence is very hard to prove. Ref. [40] qualifies this term as a symmetry breaking mechanism, which is important for initiation of motion, however it is observed in [40] that self-sustained motion occurs even without γ-term (page 3 in [40]). Third, our study reveals another symmetry breaking mechanism in (1)-(2), emanated from asymmetry of the potential W(ρ) (see Subsection 1.2). That is, the effect of γ-term is replaced, to some extent, by asymmetry of the potential W. Note that symmetry of potential W(ρ)=W(1ρ) reflects an idealized view that the two phases ρ0 and ρ1 are equivalent resulting in the symmetric profile of the interface with respect to interchanging of the phases. In the model under consideration the phase ρ1 corresponds to the cell interior and ρ0 outside the cell, therefore the phases are not equivalent and it is natural to assume that the potential W is asymmetric.

    Heuristically, system (1)-(3) describes the motion of a interface caused by the competition between mean curvature motion (due to stiffness of interface) and the push of the orientation field on the interface curve. The main issue is to determine the influence of this competition on the qualitative behavior of the sharp interface solution. The parameter β>0 models this competition which is why it plays a key role in the analysis of system (1)-(3).

    Recall the Allen-Cahn equation which is at the core of system (1)-(3),

    ρεt=Δρε1ε2W(ρε), (4)

    where W(ρ) is the derivative of a double equal well potential W(ρ). We suppose that

    {W()C3(R), W(ρ)>0 when ρ{0,1},W(ρ)=W(ρ)=0 at {0,1}, W(0)>0, W(1)>0, (5)

    e.g. W(ρ)=14ρ2(ρ1)2. Equation (4) was introduced in [3] to model the motion of the phase-antiphase boundary (interface) between two grains in a solid material. Analysis of (4) as ε0 leads to the asymptotic solution that takes values ρε0 and ρε1 in the domains corresponding to two phases separated by an interface of width of order ε, the so-called sharp interface. Furthermore, it was shown that this sharp interface obeys mean curvature motion. Recall that in this motion the normal component of the velocity of each point of the surface is equal to the mean curvature of the surface at this point. This motion has been extensively studied in the geometrical community (e.g., [19,21,18,7] and references therein). It also received significant attention in PDE literature. Specifically [12] and [13] established existence of global viscosity solutions (weak solutions) for the mean curvature flow. Mean curvature motion of the interface in the limit ε0 was formally derived in [15], [36] and then justified in [14] by using the viscosity solutions techniques. The limit ε0 was also studied for a stochastically perturbed Allen-Cahn equation (4) in [23,29].

    Solutions of the stationary Allen-Cahn equation with the volume constraint were studied in [26] by Γ-convergence techniques applied to the stationary variational problem corresponding to (4). It was established that the Γ-limiting functional is the interface perimeter (curve length in 2D or surface area in higher dimensions). Subsequently in the work [35] an evolutionary reaction-diffusion equation with double-well potential and nonlocal term that describes the volume constraint was studied. The following asymptotic formula for evolution of the interface Γ in the form of volume preserving mean curvature flow was formally derived in [35]:

    V=κ1|Γ(t)|Γ(t)κds, (6)

    where V stands for the normal velocity of Γ(t) with respect to the inward normal, κ denotes the curvature of Γ(t), |Γ(t)| is the curve length. Formula (6) was rigorously justified in the radially symmetric case in [9] and in the general case in [11].

    Three main approaches to the study of asymptotic behavior (sharp interface limit) of solutions of phase field equations and systems have been developed.

    When a comparison principle for solutions applies, a PDE approach based on viscosity solutions techniques was successfully used in [14,4,17,24] and other works. This approach can not be applied to the system (1)-(3), because

    ● equations (1)-(2) are coupled through spatial gradients,

    ● equation (1) contains the nonlocal (volume preservation) term λε given by (3).

    Another technique used in such problems is Γ-convergence (see [37,23] and references therein). This technique also does not work for the system (1)-(3). The standard Allen-Cahn equation (4) is a gradient flow (in L2 metric) with Ginzburg-Landau energy functional, which is why one can use the Γ-convergence approach. However, there is no energy functional such that problem (1)-(3) can be written as a gradient flow.

    When none of the above elegant tools apply, one can use direct construction of an asymptotic expansion followed by its justification via energy bounds [28]. In Allen-Cahn type problems it typically requires a number of terms (e.g., at least five in [11]) in the expansion. In this work we use some ingredients of this technique. We construct an asymptotic formula (see e.g., (65)) with just two terms: the leading one with the only unknown location of the interface and the corrector vanishing in the limit ε0. This representation is supplemented with an additional condition that the unknown term in the corrector is orthogonal to the eigenfunction of the Allen-Cahn operator linearized around its standing wave (see (66)). This condition defines the interface location implicitly (in a "weak form", via an integral identity) but it allows us to apply a Poincaré type inequality (142) in derivation of necessary bounds for the corrector. This representation leads to a reduction of the coupled system to a single singularly perturbed non-linear PDE which for ε0 provides the sharp interface limit. This approach is rigorously justified in the 1D model problem, however we believe that this justification can be carried out in the 2D problem (1)-(3). For small β it is implemented via the contraction mapping principle; for large β it requires more subtle stability analysis of a semigroup generated by a nonlinear nonlocal operator.

    The main objectives of this work are: prove well-posedness of (1)-(3), reveal the effect of the coupling in (1)-(2) on the sharp interface limit, study qualitative behavior of system (1)-(2) versus values of the parameter β.

    The first main result, Theorem 1, demonstrates that there is no finite time blow up and that the sharp interface property of the initial data propagates in time. Theorem 1 establishes existence of solutions to problem (1)-(3) on the time-interval [0,T] for any T>0 and sufficiently small ε, ε<ε0(T). It also shows that a sharp (width ε) interface at t=0 remains sharp for t(0,T). This is proved by combining a maximum principle with energy type bounds.

    To study how coupling of equations (1)-(2) along with the nonlocal volume constraint (3) affect the sharp interface limit we use formal asymptotic expansions following the method of [15]. In this way we derive the equation of motion for the sharp interface,

    V=κ+1c0Φβ(V)1|Γ(t)|Γ(t)(κ+1c0Φβ(V))ds, (7)

    where c0 is a constant determined by the potential W(ρ) and the function Φ(V) is a given function (obtained by solving (31)).

    The parameter β in (2) can be thought of as the strength of coupling in system (1)-(3). If β is small, then (1) and (7) can be viewed as a perturbation of Allen-Cahn equation with volume preserving term and curvature driven motion (6), respectively. Results of the work [25], which addresses (7) for small (subcritical) values of β, show that curves evolving according to (7) behave similarly to those satisfying (6): they become close to circles quite fast exhibiting a little shift compared with curvature driven motion. On the other hand, if β is not small evolution of sharp interface changes dramatically. In this case the function c0VΦβ(V) is no longer invertible and one can expect quite complicated behavior of the interface curve. As the first step to study this case, it is natural to look for solutions for (1)-(3) with steady motion. We can predict existence of such solutions based on our results for a 1D analogue of (1)-(3). We prove that in the 1D case there exist traveling wave solutions with nonzero velocities, provided that β is large enough and the potential W(ρ) has certain asymmetry, e.g. W(ρ)=14ρ2(ρ1)2(ρ2+1). Existence of such traveling waves is consistent with experimental observations of motility on keratocyte cells which exhibit self-propagation along the straight line maintaining the same shape over many times of its length [22].

    Heuristically, for traveling waves with nonzero velocity, say Vε>0, the push of Pε on the front edge of the interface must be stronger than its pullback on the rear edge. This asymmetry in Pε comes forth with an asymmetry of W(ρ). We show that the velocity V=Vε solves simultaneously equations c0V=Φβ(V)λ and c0V=Φβ(V)λ, up to a small error. These equations are obtained in the sharp interface limit on the front and rear edges of the interface, respectively; Φβ(V) and Φβ(V) represent in these equations, loosely speaking, the push (and pullback) of Pε on the front and rear edges. Then eliminating λ one derives 2c0V=Φβ(V)Φβ(V), this yields the only solution V=0 unless the potential has certain asymmetry (for symmetric potentials, e.g., W(ρ)=14ρ2(ρ1)2, one has Φβ(V)=Φβ(V)). Theorem 2 justifies the equation 2c0V=Φβ(V)Φβ(V) for velocities of traveling waves in the sharp interface limit ε0. Its proof is based on Schauder's fixed point theorem.

    Finally, we study the 1D model parabolic problem without any restrictions on β, where the effects of curvature and volume preservation are mimicked by a given forcing term F(t). As already mentioned the main technical trick here is to introduce a special (two term) representation of solutions which allows us to reduce the study of the interface velocity to a single singularly perturbed nonlinear equation. Linearization of this equation and spectral analysis of the corresponding generator lead to a notion of stable and unstable velocities. The main result here, Theorem 6, can be informally stated as follows. If the interface velocity Vε belongs to the domain of stable velocities it keeps varying continuously obeying the law c0Vε(t)=Φβ(Vε(t))F(t)+o(1) until it becomes unstable (if so). This theoretical result is supplemented by numerical simulations which show that interface velocities exhibit jumps and reveal existence of a hysteresis loop. Note that velocity jumps were also observed in [40]. More precisely, the onset of cell motion was attributed to the subcritical instability (see Fig. 3 and discussion below this figure on page 5 in [40]) which is a typical example of hysteresis [39]. Our stability analysis also predicts that stationary solutions of (1)-(3) with circular shape of the phase field functions are unstable in the case of asymmetric potentials and large enough β. This conjecture is based on the fact that zero velocity is unstable in this case (see Remark 7).

    The paper is organized as follows. Section 2.1 is devoted to the well-posedness of the problem (1)-(3). In Section 2.2 the equation for the interface motion (7) is formally derived. Section 3 deals with traveling wave solutions. Section 4 contains the rigorous justification of the sharp interface limit in the context of the model 1D problem. Some results obtained in this manuscript were announced in [6] without proofs.

    In this section we consider the system (1)-(3) supplemented with the Neumann and the Dirichlet boundary conditions on Ω for ρε and Pε, respectively,

    νρε=0 and Pε=0 on Ω. (8)

    Introduce the following energy-type functionals

    Eε(t):=ε2Ω|ρε(x,t)|2dx+1εΩW(ρε(x,t))dx,Fε(t):=Ω(|Pε(x,t)|2+|Pε(x,t)|4)dx. (9)

    Assyme that system (1)-(2) is supplied with initial data that satisfy:

    ε1/4<ρε(x,0)<1+ε1/4, (10)

    and

    Eε(0)+Fε(0)C. (11)

    The first condition (10) is a weakened form of a standard condition 0ρε(x,0)1 for the phase field variable. If λε0, then the maximum principle implies 0ρε(x,t)1 for t>0. The presence of nontrivial λε leads to an "extended interval" for ρε.1 The second condition (11) means that at t=0 the function ρε has the structure of an "ε-transition layer", that is, the domain Ω consists of three subdomains: one where ρε1 (inside the cell), another where ρε0 (outside the cell), and they are separated by a transition layer of width ε (a diffusive interface). Furthermore, it can be shown that the magnitude of the orientation field Pε is small everywhere except the ε-transition layer (see (19)).

    1 The exponent 1/4 in (10) can be replaced by any positive number less than 1/2 as will be seen in the proof the next theorem.

    Theorem 1. (Bounds on finite time intervals) If the initial data ρε(x,0), Pε(x,0) satisfy (10) and (11), then for any T>0 the solution of (1)-(2) ρε, Pε with boundary conditions (8) exists on the time interval (0,T) for sufficiently small ε>0, ε<ε0(T). Moreover, ε1/4ρε(x,t)1+ε1/4 and

    εT0Ω(ρεt)2dxdtC,Eε(t)+Fε(t)Ct(0,T), (12)

    where C is independent of t and ε.

    Remark 1. This theorem implies that if the initial data are well-prepared in the sense of (10)-(11), then for 0<t<T the solution exists and has the structure of an ε-transition layer. Moreover, the bound on initial data (10) remains true for t>0. While it relies on a maximum principle argument, it also requires additional estimates on λε as seen from (14) below. Note also that in the proof below the interval of existence of the solution extends to [0,Tε], where Tε can be estimated from below by C0|logε|. However, energy bounds (12) are proved only for Tε=O(1).

    Remark 2. A solution of (1)-(2) is understood as follows

    ρεC([0,T];H1(Ω)),  tρεL2((0,T)×Ω),PεL2(0,T;H1(Ω)),  tPεL2((0,T)×Ω)

    and equations (1) and (2) hold in H1(Ω) for almost all t[0,T].

    Proof. First multiply (1) by tρε and integrate over Ω:

    Ω|tρε|2dx+ddtΩ(12|ρε|2+1ε2W(ρε))dx=ΩPερεtρεdx12Ω|tρε|2dx+12Ω|Pε|2|ρε|2dx. (13)

    Here we used the fact that, due to (3), the integral of tρε over Ω is zero and thus

    Ωλε(t)tρεdx=0.

    Next, using the maximum principle in (1) we get:

    2ε2supτ(0,t]|λε(τ)|ρε1+2ε2supτ(0,t]|λε(τ)|. (14)

    Let Tε>0 be the maximal time such that

    ε1/4ρε1+ε1/4,when tTε, (15)

    and from now on assume that tTε.

    Using (13), (15) and integrating by parts we obtain

    ddtEε+ε4Ω|tρε|2dxε(|Pε|2|Δρε|+||Pε|2||ρε|)dx (16)

    We proceed by deriving an upper bound for the integral in the right hand side of (16). By (1) we have

    Ω(|Δρε||Pε|2+||Pε|2||ρε|)dxΩ|tρε||Pε|2dx+Ω|Pερε||Pε|2dx+Ω|ρε|||Pε|2|dx+1ε2Ω|W(ρε)||Pε|2dx+|λε|Ω|Pε|2dx=:5i=1Ii.

    The following bounds are obtained by routine application of the Cauchy-Schwarz and Young's inequalities. For the sum of the first three terms in (17) we get,

    31IiεΩ(tρε)2dx+εΩ|Pε|2|ρε|2dx+12εΩ|Pε|4dx+Ω||Pε|2|2dx+1εEε.

    Since (W(ρε))2CW(ρε) we also have

    1ε2Ω|W(ρ)||Pε|2dxCε2ΩW(ρ)dx+12ε2Ω|Pε|4dxCεEε+12ε2Ω|Pε|4dx.

    Finally, in order to bound I5 we first derive,

    |λε(t)|Cε2(ΩW(ρε)dx)1/2+(Ω|ρε|2dx)1/2(Ω|Pε|2dx)1/2Cε(Eεε)1/2+(2EεεΩ|Pε|2dx)1/2, (17)

    then

    I5Cε(Eεε)1/2Ω|Pε|2dx+(2Eεε)1/2(Ω|Pε|2dx)3/2CεEε+12ε2Ω|Pε|4dx+E2ε+Cε2/3Ω|Pε|4dx.

    Thus,

    51IiCεEε+E2ε+1+O(ε)ε2Ω|Pε|4dx+εΩ(tρε)2dx+εΩ|Pε|2|ρε|2dx+Ω||Pε|2|2dx,

    and using this inequality, (17) and (15) in (16), then substituting the resulting bound in (13) we obtain, for sufficiently small ε,

    14Ω|tρε|2dx+ddtEεεCεEε+E2ε+1ε2Ω|Pε|4dx+Ω||Pε|2|2dx. (18)

    Now we obtain a bound for the last two terms in (18). Taking the scalar product of (2) with 2kPε+4|Pε|2Pε, k>0, integrating over Ω and using (15) we get

    ddtΩ(k|Pε|2+|Pε|4)dx+εΩ(2k|Pε|2+4|Pε|2|Pε|2+2||Pε|2|2)dx+2εΩ(k|Pε|2+2|Pε|4)dx=2βkΩPερεdx+4βΩρεdiv(Pε|Pε|2)dxkCεΩ|ρε|2dx+kεΩ|Pε|2dx+εΩ|Pε|2|Pε|2dx+C1εΩ|Pε|2dx.

    We chose k:=C1+1 to obtain

    εΩ||Pε|2|2dx+1εΩ|Pε|4dxCEεddtΩ(k|Pε|2+|Pε|4)dx. (19)

    Finally, introducing Gε=Eε+Ω(4k|Pε|2+|Pε|4)dx, by (18) and (19) we have the differential inequality,

    dGεdtCGε+εG2ε, (20)

    with a constant C>0 independent of ε. Considering the bounds on the initial data and assuming that ε is sufficiently small, one can easily construct a bounded supersolution ˜G of (20) on [0,T] such that ˜G(0)Gε. We now have, GεC on [0,Tε] for sufficiently small ε. By (14) and (17) we then conclude that Tε in (15) actually coincides with T when ε is small. The theorem is proved.

    In this section we formally derive equation (7) for the 2D system (1)-(2). While the derivation is analogous to the single Allen-Cahn equation (e.g., [36], [11]), the gradient coupling in (1)-(2) results in a nonlinear term that modifies the mean curvature motion.

    Assume that that initial data ρε(x,0) converge to the characteristic function of a smooth subdomain ω0Ω as ε0. Next we want to describe the evolution of the interface Γ(t)=ωt with t, where ωt is the support of limε0ρε(x,t). We will assume that the initial data coincide with initial values of asymptotic expansions for ρε and Pε to be constructed below.

    Let X0(s,t) be a parametrization of Γ(t). In a vicinity of Γ(t) the parameters s and the signed distance r to Γ(t) will be used as local coordinates, so that

    x=X0(s,t)+r(s,t)=X(r,s,t),where υ is an inward normal to Γ(t).

    The inverse mapping to x=X(r,s,t) is given by

    r=±dist(x,Γ(t)),s=S(x,t),

    where in the formula for r we choose + if xωt and , if xωt. Recall that Γ(t) is the limiting location of interface as ε0. Next we seek ρε and Pε in the following forms in local coordinates (r,s):

    ρε(x,t)=˜ρε(r(x,t)ε,S(x,t),t) and Pε(x,t)=˜Pε(r(x,t)ε,S(x,t),t). (21)

    Introduce asymptotic expansions in local coordinates:

    ˜ρε(z,s,t)=θ0(z,s,t)+εθ1(z,s,t)+... (22)
    ˜Pε(z,s,t)=Ψ0(z,s,t)+ (23)
    λε(t)=λ0(t)ε+λ1(t)+ελ2(t)+... (24)

    Now, substitute (22)-(24) into (1) and (2). Collecting terms with likewise powers of ε (ε2 and ε1) and equating them to zero we successively get,

    2θ0z2=W(θ0), (25)

    and

    2θ1z2+W(θ0)θ1=V0θ0zθ0zκ(s,t)(Ψ0υ)θ0z+λ0(t), (26)
    V0Ψ0z=2Ψ0z2Ψ0βθ0z, (27)

    where κ(s,t) is the curvature of Γ0(t) and V0(t):=tr is the limiting velocity. The curvature κ appears in the equation when one rewrites the Laplace operator in (1) in local coordinates (r,s).

    It is well-known that there exists a standing wave solution θ0(z) of (25) which tends to 1 as z and to 0 as z, respectively. Moreover, all derivatives of the function θ0(z) exponentially decay to 0 as |z| and θ0(z) is an eigenfunction of the linearized Allen-Cahn operator Lu:=u"+W"(θ0)u corresponding to the eigenvalue 0. Then multiplying (26) by θ0(z) and integrating over z we are lead to the solvability condition for (26)

    c0V0(s,t)=c0κ(s,t)+(Ψ0 υ)(θ0z)2dzλ0(t), where c0=R(θ0z)2dz. (28)

    Next we obtain the formula for λ0(t). It follows from (3) that Ωtρε=0. Substitute expansion (22) for ρε into Ωtρε=0 and take into account the fact that

    tρε=θ0(rε)V0(s,t)ε+O(1).

    Thus, in order to satisfy the condition Ωtρε=0 to the leading order, V0(s,t) must have

    V(s,t)|sX0(s,t)|ds = 0.

    Using this fact and integrating (28) with respect to s with the weight |sX0(s,t)|, we get

    λ0(t)={c0κ(s,t)+(Ψ0υ)(θ0z)2dz}|sX0(s,t)|ds. (29)

    Finally, the unique solution of (27) is given by Ψ0(z,s,t)=ψ(z,V0(s,t))υ(s,t) where ψ=ψ(z,V) is the unique (bounded) solution of

    2zψ+Vzψψβθ0=0. (30)

    The representation Ψ0(z,s,t)=ψ(z;V0(s,t))υ(s,t) yields

    Ψ0υ(θ0)2dz=Φβ(V):=ψ(z,V)(θ0)2dz, (31)

    where we have also taken into account the linearity of (30) in β. Now substitute (31) and (29) into equation (28) to conclude the derivation of sharp interface equation (7).

    In this section we study special solutions of system (1)-(2) in the 1D case. Specifically, we look for traveling waves (traveling pulses). Therefore it is natural to switch to the entire space R1 setting. We show that, not surprisingly, there are nonconstant stationary solutions, standing waves. However, we prove that apart from standing waves there are true traveling waves when the parameter β is large enough and the potential W(ρ) has certain asymmetry, e.g. W(ρ)=14(ρ2+ρ4)(ρ1)2, see also the discussion in Remark 4.

    We are interested in (localized in some sense) solutions of (1)-(2) with ρε=ρε(xVt), Pε=Pε(xVt). They satisfy the following stationary equations with unknown constant velocity V and constant λ:

    0=2xρε+VxρεW(ρε)ε2Pεxρε+λε, (32)
    0=ε2xPε+VxPε1εPεβxρε. (33)

    Let us postulate an ansatz for the phase field function ρε. Given a>0, we look for solutions of (32)-(33) for sufficiently small ε>0 with ρε having the form

    ρε=ϕε+εχε+εu, (34)

    where

    ϕε:=θ0((x+a)/ε)θ0((ax)/ε),χε:=χε+(χ+εχε)ϕε,

    constants χε and χ+ε are the smallest (in absolute value) solutions of W(εχε)=ελ and W(1+εχ+ε)=ελ, respectively, and u is the new unknown function vanishing at ±. The role of the constant χε in (34) is to amend the first term of the representation so that u decays at ±. Similarly, χ+ε is introduced to end up with u which is exponentially close to one in (a,a) away from points ±a (see also Fig. 1).

    Figure 1.  Illustration of the ansatz (34). Function ρε decays to a non-zero constant of the order ε for x± and to a constant slightly different from 1 for axa (solid). Dashed line represents the limiting profile, which is the characteristic function of (a,a).

    Substitute representation (34) in (32)-(33) to find after rescaling the variable y:=x/ε and rearranging terms,

    2yuW"(ϕε)u=Vyϕε+Pε(yϕε+εyu)λ+1ε(W(ϕε+εχε)2y(ϕε+εχε))+1ε(W(ϕε+εχε+εu)W(ϕε+εχε)εW"(ϕε)u)εVy(u+χε), (35)

    and

    2yPε+VyPεPε=βyϕε+εβy(χε+u). (36)

    Note that the ansatz (34) yields the characteristic function of the interval (a,a) in the limit ε0, provided that u=uε remains bounded. In this sense we seek solutions with localized profiles of the phase field function ρε. The idea of the construction of traveling wave solutions is based on the observation that solvability of the above equations (35) and (36) can be handled by local analysis near the points ±a. Indeed, setting z=y+a (35)-(36) and keeping only leading order terms we (formally) obtain

    2zuW(θ0(z))u=Vθ0(z)+Pεθ0λ  and  2zPε+VzPεPε=βθ0(z).

    Resolve the second equation to obtain Pε(z)=ψ(z,V), then solvability of the first equation (recall that 2zθ0(z)W(θ0(z))θ0(z)=0) requires that c0V=Φβ(V)λ, where we have used (31). Similarly, local analysis near the point a leads to the equation c0V=Φβ(V)λ. Thus, we have reduced the infinite dimensional system (35)-(36) to a two dimensional one.

    In order to transform the above heuristics into a rigorous analysis we reset (35)-(36) as a fixed point problem. To this end rewrite (35) in the following form, introducing auxiliary functions θ(1)ε(y)=θ0(y+a/ε)+θ0(a/εy) and θ(2)ε(y)=θ0(y+a/ε)θ0(a/εy),

    2yuW"(ϕε)u2yθ(1)εW(ϕε)θ(1)εθ(1)εu+Hεuθ(2)εdy=G(λ,V,Pε,u), (37)

    where

    Hε=1(θ(2)ε)2dy(W(ϕε)θ(2)ε2yθ(2)ε+2yθ(1)εW(ϕε)θ(1)εθ(1)εθ(2)ε)

    and

    G(λ,V,Pε,u)=Hεuθ(2)εdy2yθ(1)εW(ϕε)θ(1)εθ(1)εuVyϕε+Pε(yϕε+εyu)εVy(u+χε)λ+1ε(W(ϕε+εχε)2y(ϕε+εχε))+1ε(W(ϕε+εχε+εu)W(ϕε+εχε)εW"(ϕε)u). (38)

    Note that the operator Qε in the left hand side of (37),

    Qεu:=2yuW"(ϕε)u2yθ(1)εW(ϕε)θ(1)εθ(1)εu+Hεuθ(2)εdy

    has two eigenfunctions θ(1)ε and θ(2)ε corresponding to the zero eigenvalue.

    Lemma 1. Let vε be any function from H1(R) orthogonal to both θ(1)ε and θ(2)ε in L2(R). Assume also that fε:=Qεvε belongs to L2(R). Then for sufficiently small ε

    vεH1CfεL2 (39)

    with C independent of ε and vε.

    Proof. Multiplying Qεvε by vε in L2(R) and representing vε as vε=θ(1)εwε (note that θ(1)ε>0) we derive

    (Qεvε,vε)L2=(2yθ(1)εywε+θ(1)ε2ywε)θ(1)εwεdy=(θ(1)ε)2(ywε)2dy, (40)

    where the latter equality is obtained via integrating by parts, and the term with Hε vanishes thanks to orthogonality of vε to θ(2)ε. Thus

    (θ(1)ε)2(ywε)2dyfεL2vεL2. (41)

    The statement of Lemma 1 immediately follows if we prove the following inequality

    v2εdyC(θ(1)ε)2(ywε)2dy (42)

    with C independent of ε and vε. Indeed, using (42) in (41) we get

    (θ(1)ε)2(ywε)2dyCfε2+Cer/ε, (43)

    and combining (43) with

    vε2H1(θ(1)ε)2(ywε)2dy+Cv2εdy

    yields (39).

    To prove (42) we use the Poincaré inequality (see Appendix A)

    (θ0(a/ε±y))2|wεwε±|2dyC(θ(1)ε)2(ywε)2dy (44)

    with a constant C independent of ε and

    wε±:=(θ0(a/ε±y))2wεdy(θ0)2dy.

    Due to orthogonality of θ(1)εwε to θ(1)ε and θ(1)ε, we have

    (θ0(a/ε±y))2wεdy=θ0(y+a/ε)θ0(a/εy)wεdy.

    Thanks to the exponential decay of θ0, θ0(y)α0eκ|y| (see, e.g., [28]), it follows that

    |(θ0(a/ε±y))2wεdy|er/ε((θ(1)ε)2w2εdy)1/2 (45)

    for sufficiently small ε and r>0 independent of ε. Combining (45) and (44) we obtain (42), the Lemma is proved.

    Proposition 1. For sufficiently small ε the operator Qε adjoint to Qε (with respect to the scalar product in L2(R)) has two eigenfunctions θ(1)ε and θ(3)ε=θ(2)ε+qε corresponding to the zero eigenvalue, with qεH1=o(ε). Moreover the equation Qεu=f has a solution if and only if fL2(R) is orthogonal to the eigenfunctions θ(1)ε and θ(3)ε of Qε.

    Proof. Given fL2(R), consider the equation Qεu=f rewriting it in the form

    2yuW(0)u+(W(0)W(ϕε))u2yθ(1)εW(ϕε)θ(1)εθ(1)εu+Hεuθ(2)εdy=f. (46)

    Since W(0)>0, the equation 2yuW(0)u=˜f has the unique solution u=G˜f for every ˜fL2(R) with a bounded resolving operator G:L2(R)L2(R). Moreover, by applying the operator G to (46) we reduce this equation to u+Ku=Gf, where K is a compact operator (this can be easily shown using the properties of the function θ0). Thus we can apply the Fredholm theorem to study the solvability of (46). Note that Qε does not have other eigenfunctions corresponding to the zero eigenvalue besides θ(1)ε and θ(2)ε. Indeed, existence of such an eigenfunction vε orthogonal to θ(1)ε, θ(2)ε in L2(R) and normalized by v2εdy=1 would contradict (41) derived in the proof of Lemma 1.

    Consider now the eigenfunction θ(3)ε of Qε orthogonal to θ(1)ε, and represent it as θ(3)ε=θ(2)ε+qε with qε orthogonal to both θ(1)ε and θ(2)ε. Then combining the equality

    Qεqε=Hε(θ(2))2dyθ(2)Hε(θ(2)εqε)dy

    with Lemma 1 we obtain that qεH1=o(ε) as ε0.

    Let us consider now for a given uH1(R), V and λ a solution Pε of (35), assuming that ε is sufficiently small and uH1M, |λ|M, |V|M for some finite M. We represent Pε in the form

    Pε(y)=ψ0(y+a/ε,V)ψ0(a/εy,V)+Bε (47)

    and observe that Bε can be estimated as follows,

    (yBε)2dy+(Bε)2dyεCMBεL2  hence  BεH1εC1M.

    Now consider u in the left hand side of (37) as an unknown function to write down the solvability condition

    G(λ,V,Pε,u)θ(k)εdy=0, k=1,3. (48)

    Calculate leading terms of (48) for small ε taking into account the fact that

    W(ϕε+εχε+εu)W(ϕε+εχε)εW(ϕε)u=O(ε2) (49)

    and

    W(ϕε+εχε)2y(ϕε+εχε)=ε(W(ϕε)χε2yχε)+O(ε2), (50)

    where O(ε2) in (49) and (50) stand for functions whose L-norm is bounded by Cε2. Note also that integrals

    (W(ϕε)χε2yχε)θ(k)εdy=(W(ϕε)θ(k)ε2yθ(k)ε)χεdy, k=1,3

    tend to zero, when ε0. Thus (48) can be rewritten as

    0=Φβ(V)+Φβ(V)2λ+ε˜Φ1(V,λ,u) and 2c0V=Φβ(V)Φβ(V)+ε˜Φ2(V,λ,u), (51)

    where functions ˜Φ1, ˜Φ2 and their first partial derivatives in V and λ are uniformly bounded by some constant depending on M only. Note that if V0 is a nondegenerate root of the equation 2c0V=Φβ(V)Φβ(V) then for sufficiently small ε, in a neighborhood of V0 and λ0=12(Φβ(V0)+Φβ(V0)) there exists a unique pair Vε(u) and λε(u) solving (51) and depending continuously on u.

    Theorem 2. (Existence of traveling waves) Assume that the equation 2c0V=Φβ(V)Φβ(V) has a nondegenerate root V0. Then for sufficiently small ε there exists a function uε, with uεH1C and C being independent of ε, a function Pε and constants V=Vε, λ=λε such that ρε given by (34) and Pε are solutions of (32)-(33). Moreover, the velocity Vε and the constant λε converge to V0 and λ0:=12(Φβ(V0)+Φβ(V0)) as ε0.

    Proof. Consider the mapping uQ1εG(λε,Vε,Pε,u), where λε=λε(u) and Vε=Vε(u) solve (51), and Pε is the solution of (36) with V=Vε and λ=λε. The operator Qε has two eigenfunctions θ(1)ε and θ(2)ε corresponding to the zero eigenvalue, we choose v:=Q1εG(λε,Vε,Pε,u) to be orthogonal to θ(1)ε and θ(2)ε in L2(R). Then the following holds. If uH1M, then Q1εG(λε,Vε,Pε,u)H1M for large M and sufficiently small ε. Indeed, by virtue of Lemma 1 it suffices to find an appropriate bound for the norm G(λε,Vε,Pε,u)L2. Considering formula (38) for G(λε,Vε,Pε,u) observe that the only third line and terms Vεyϕε and Pεyϕε in the second line have non vanishing norm in L2(R) as ε0. Moreover, these norms can be bounded by C+εC1(M) with C independent of M, while the norm of other therms can be estimated by εC2(M). Thus G(λε,Vε,Pε,u)L2C+εC3(M) and using Lemma 1 we obtain

    Q1εG(λε,Vε,Pε,u)H1C4+εC5(M),

    with C4 independent of M. It remains to choose M>C4 to conclude that

    Q1εG(λε,Vε,Pε,u)H1M

    for sufficiently small ε.

    Also, the mapping uQ1εG(λε,Vε,Pε,u) is continuous in H1. Thus, we can apply the Schauder fixed point theorem provided we establish the compactness of the mapping under consideration. To this end we consider a subset of functions u which decay exponentially with their first derivatives:

    KM,r:={u:uH1M,|u|,|yu|Mer(|y|2aε) when |y|2aε}. (52)

    We claim that for some M>0 and r>0 the solution v of the equation Qεv=G(λε,Vε,Pε,u) (orthogonal to θ(1)ε and θ(2)ε) belongs to KM,r for every uKM,r, when ε is sufficiently small. Indeed, the required bound for the norm of v in H1(R) is already established. It remains to prove that v and yv decay exponentially when |y|2a/ε. To this end we observe first that

    |Pε|C(1+εM)er1(|y|2a/ε) for |y|2a/ε, (53)

    with C>0 independent of M and ε, and r1>0 depending on M only. The proof of (53) is carried out in two steps. First, we multiply (36) by Pε, integrate on R and apply the Cauchy-Schwarz inequality. As a result we get PεLCPεH1C1(1+εM). Second, observe that the function θ0(y) decays exponentially when y±. Therefore there exists C2C1(1+εM) and r1>0 such that the functions P±(y):=±C2er1(|y|2a/ε) satisfy

    2yPεVyPε±Pεβyϕε+εβy(χε+u) for |y|2a/ε.

    This yields pointswise bounds C2er1(|y|2a/ε)Pε(y)C2er1(|y|2a/ε) for all y2a/ε and y2a/ε. Next using (53) in the equation Qεv=G(λ,V,Pε,u) and arguing similarly one can establish that |v|C(1+εC1(M))er2(|y|2a/ε) for |y|2a/ε. Finally, taking an integral from to y (or from y to +) of the equation Qεv=G(λ,V,Pε,u) we get the required bound for yv on (,2a/ε] (or [2a/ε,+)).

    Thus the image of the convex closed set KM,r under the mapping

    uQ1εG(λε,Vε,Pε,u)

    is contained in KM,r. Also, the equation Qεv=G(λε,Vε,Pε,u) provides the following bound |2yv|C(|u|,|yu|,|v|,|yv|) while |u|, |v|, |yu| and |yv| have pointwise bounds with decay estimates for large y, both u and v being elements of KM,r. This implies that the mapping uQ1εG(λε,Vε,Pε,u) is compact on KM,r (in the topology of H1(R)). Thus there exists a fixed point of this mapping in KM,r. Since the principal part 0=Φβ(V)+Φβ(V)2λ and 2c0V=Φβ(V)Φβ(V) of the system (51) is nondegenerate in the neighborhood of V0 and λ0, we have VεV0 and λελ0 as ε0.

    Remark 3. Note that V0=0 and λ0=Φβ(0) are always solutions of the principal part of the system (51). Moreover one can establish a traveling wave (in fact standing wave) solution with Vε equal to zero exactly, by following the line of Theorem 2 but considering subspace of even functions uH1(R). The existence of nontrivial traveling waves (with nonzero velocities) is granted by Theorem 2 in the case when the equation 2c0V=Φβ(V)Φβ(V) has a nonzero (nondegenerate) root. Such a solution does not exist for the standard potential W(ρ)=14ρ2(ρ1)2, in this case Φβ(V)=Φβ(V) for all V due to the fact that θ0 is an odd function. However, if the potential has two equally deep wells but possesses certain asymmetry, e.g. W(ρ)=14(ρ2+ρ4)(ρ1)2, we have Φβ(V)>Φβ(V) for V>0, so that nontrivial solutions of 2c0V=Φβ(V)Φβ(V) do exist for sufficiently large β, β>βcritical. The plot of the function Φβ(V) for β=1 and W(ρ)=14(ρ2+ρ4)(ρ1)2 is depicted in Fig. 2, as well as the corresponding standing wave.

    Figure 2.  Left: Φβ(V) for β=150 and Was(ρ)=14ρ2(1+ρ2)(ρ1)2, positive slope illustrates Φβ(0)>0; Right: θ0, standing wave for the Allen-Cahn equation for Wsym(ρ)=14ρ2(ρ1)2 (dashed) and Was(ρ)=14ρ2(1+ρ2)(ρ1)2 (solid).

    Remark 4. As already mentioned, nontrivial traveling waves appear in the case when W(ρ) has certain asymmetry, that, in particular, makes the derivative Φβ(V) of Φβ(V) positive at V=0. The function Φβ(V) depends on the potential W(ρ) in a complex way. Its derivative at V=0 is given by Φβ(0)=β(θ0)2ψV(y)dy with ψV(y) solving (2y+I)2ψV=θ0". The following representation ψV(y)=e|yz|(1+|yz|)θ0(z)dz can be obtained in a standard way by using fundamental solution, so that

    Φβ(0)=β1010e|y(θ0)y(˜θ0)|(1+|y(θ0)y(˜θ0)|)W(˜θ0)W(θ0)W(˜θ0)dθ0d~θ0, (54)

    where y(θ0)=θ01/2dθ02W(θ0) (this relation between y and θ0 follows from the equation θ0=2W(θ0) obtained by multiplication of θ0"=W(θ0) by θ0 and integration with respect to y). On the other hand, Φβ is a bounded function, therefore if the right hand side of (54) is positive, then for sufficiently large β there are non-zero solutions of equation 2c0V=Φβ(V)Φβ(V).

    In order to have more insight about this dependence assume that the diffusion coefficient in equation (2) for Pε is given by δε, where δ is a positive parameter independent of ε. This leads to redefining Φβ(V) as follows,

    Φβ(V)=χ(θ0)2dy,δ2yψVyψ+ψ=βθ0.

    One can write down an asymptotic expansion of ψ and its derivative ψV with respect to V at V=0 for sufficiently small δ>0

    ψ=βθ0δβθ0+,ψV=βθ02δβθ()0+.

    Then we have

    Φβ(0)=2δβθ()0(θ0)2dy+O(βδ2),

    which yields, after integrating by parts and using the relations (θ)2=2W(θ), θ=W(θ),

    Φβ(0)=823δβ10W(ρ)dW3/2(ρ)+O(βδ2). (55)

    The integral in (55) can be interpreted as a measure of asymmetry of the potential W(ρ), and nontrivial traveling waves emerge if this integral is positive and

    β>βcritical=3c082δ10W(ρ)dW3/2(ρ)+O(δ2).

    The equation of motion (7) formally derived in Subsection 2.2 exhibits qualitative changes for large values of the parameter β. This is indicated, in particular, by the fact that the equation

    c0VΦβ(V)=F, (56)

    may have multiple roots V. Note that combining the curvature and integral (constant) terms in (7) yields the equation of the form (56) with

    F:=1|Γ|Γ(κ+Φβ(V))dsκ.

    In this Section we analyze a 1D analogue of the original model and rigorously derive a law of motion in the sharp interface limit. For given F(t)C[0,T] we consider bounded solutions of the system

    {ρεt=2xρεW(ρε)ε2Pεxρε+F(t)ε,xR1, t>0,Pεt=ε2xPε1εPεβxρε.

    Analysis of the 1D problem (57)-(58) is a necessary step for understanding the original problem (1)-(2). Observe that motion of the interface in the 2D system (1)-(2) occurs in the normal direction, and therefore it is essentially one-dimensional. Thus, the 1D model (57)-(58) is anticipated to capture the main features of (1)-(2). The effects of curvature and mass conservation in (7) are modeled by a given function F(t). We believe that qualitative conclusions obtained for the 1D problem (57)-(58) apply for the 2D model (1)-(2).

    We study the asymptotic behavior of solutions to the system (57)-(58) as ε0 with "well-prepared" initial data for ρε,

    ρε(x,0)=θ0(x/ε)+εvε(x/ε), (59)

    where θ0 is a standing wave solution of the Allen-Cahn equation (25) such that θ0(z)0 as z and θ0(z)1 as z+. We seek ρε in the form

    ρε(x,t)=θ0(xxε(t)ε)+εvε(xxε(t)ε,t). (60)

    The xε(t) in (60) can be viewed as a location of the interface. Remark 5 explains that a choice of xε is not unique, however it is well defined in the limit ε0.

    The main goal of this Section is to prove that xε(t) converges as ε0 to x0(t), whose velocity V0(t)=˙x0(t) solves the sharp interface equation

    c0V0(t)=Φβ(V0(t))F(t), (61)

    where Φ(V) is the known nonlinear function given by (31). This equation can be formally obtained in the limit ε0 as in the Section 2.2.

    Next for reader's convenience we summarize key steps of the asymptotic analysis of (57)-(58):

    (ⅰ) Choice of a special representation. The function ρε is represented in the form

    ρε(x,t)=θ0(y)+εχε(y,t)+εuε(y,t), Pε(x,t)=Qε(y,t),y=xxε(t)ε, (62)

    where θ0 and χε are known, and uε, Qε are the new unknown functions. Existence of xε(t) with estimates on uε uniform in ε and t are established in Section 4.2.

    (ⅱ) Reduction of the system to a single equation. The unknown function uε is eliminated by showing that the third term in representation (62) is small. Next, we split Qε into two parts, Qε=Aε+Bε, where Bε depends on uε but is small, and Aε does not depend on uε. Thus, the original system (57)-(58) is reduced to

    {(c0+o(1))Vε(t)=(θ0)2AεdyF(t)+o(1),εAεt=2yAε+Vε(t)yAεAεβθ0.

    Taking the limit ε0 in the system (63)-(64) is non-trivial because of the product term Vε(t)yAε.

    (ⅲ) Analysis of reduced problem. For sufficiently small β we prove that xε(t)x0(t) as ε0 by the contraction mapping principle. For larger β, system (63)-(64) further reduces to a singularly perturbed non-linear non-local equation. The limiting transition in this equation is based on the stability analysis of the semigroup generated by the linearized operator.

    In order to pass to the limit ε0 in (57)-(58) we further specify vε in (60). Namely, we introduce the representation

    ρε(x,t)=θ0(xxε(t)ε)+εχε(xxε(t)ε,t)+εuε(xxε(t)ε,t), (65)

    with the new unknown function uε satisfying

    θ0(y)uε(y,t)dy=0, (66)

    and χε(y,t) defined by

    χε(y,t)=χε(t)+θ0(y)(χ+ε(t)χε(t)),

    where χ+ and χ are solutions of the following ODEs

    ε2tχ+ε=W(1+εχ+ε)ε+F(t),ε2tχε=W(εχε)ε+F(t) (67)

    with the initial data χ+ε(0)=F(0)/W(1) and χε(0)=F(0)/W(0).

    The idea of the decomposition of the lower order term in (60) into two parts is suggested by the observation that it is the most important to control behavior of ρε in the vicinity of the interface. So, ideally we would like to localize the analysis by considering functions that are negligibly small outside the interface. However, the right hand side F(t) prevents ρε from being localized. The function χε absorbs this nonlocal part of ρε: the new unknown function uε decays at infinity and, therefore, it allows one to work in Sobolev spaces on R. Note that the standard ODE methods yield the following bounds

    |χε(y,t)|+|yχε(y,t)|+|2yχε"(y,t)|Ct[0,T],yR, (68)

    moreover, thanks to the continuity of F(t) and a particular choice of the initial values χ±ε(0) we have

    ε2tχεL0uniformly on [0,T] as ε0. (69)

    Finally, we set Qε(y,t):=Pε(xε+εy,t).

    Remark 5. The choice of xε in the representation (60) is not unique, e.g. its perturbation with a term of order ε2 still leads to an expansion of the form (60). We introduced the additional orthogonality condition (66) which implicitly specifies xε(t). This condition allows us to use Poincaré type inequalities (see Appendix A) when deriving various bounds for uε. If the initial value of uε in the expansion (65) does not satisfy (66), it can be fixed by perturbing the initial value xε(0)=0 with a higher order term. Indeed, this amounts to solving the equation

    (θ0(y+xε(0)/ε)θ0(y))θ0(y)dy=ε(χε(y,0)vε(y+xε(0)/ε))θ0(y)dy.

    If vεL2C then the latter equation has a solution xε(0) and |xε(0)|=o(ε).

    The following theorem justifies the expansions (65) and will be used to obtain a reduced system for unknowns xε(t) and Qε(y,t) by eliminating uε.

    Theorem 3. (Validation of representation (65)-(66)) Let ρε and Pε be solutions of problem (57)-(58) with initial data ρε(x,0)=θ0(x/ε)+εvε(x/ε) and Pε(x,0)=pε(xε), where

    vεL2<C,vεLC/ε, (70)

    and

    pεL2(R)+ypεL2C. (71)

    Then there exists xε(t) such that the expansion (65)-(66) holds with uε(,t)L2 C for t[0,T].

    Proof. Step 1. (coupled system for uε, Qε and Vε:=˙xε) Note that the maximum principle applied to (57) yields ρεLC. This bound in conjunction with (68) allow one to write down the expansion

    W(θ0+ε(χε+uε))=W(θ0+εχε)+εW(θ0)uε+ε2W(ξε)χεuε+ε22W(¯ξε)u2ε,

    where ξε and ¯ξε are some bounded functions (while ξε and ¯ξε depend on θ0, χε and uε, this dependence is omitted for brevity). Then substituting the expansion (65) into equation (57) leads to

    ε2uεt=2yuεW(θ0)uε+Vεθ0Qεθ0+2yχε+W(θ0)W(θ0+εχε)ε+F(t)ε2χεtεW(ξε)χεuεε2W(¯ξε)u2εεQε(yχε+yuε)+εVε(yχε+yuε). (72)

    This equation is coupled with that for Qε

    εQεt=2yQε+VεyQεQεβθ0εβ(yχε+yuε). (73)

    Finally, considering the solution ρε as a given function we differentiate (65) in time, multiply by θ0(y) and integrate in y over R to obtain the equation for Vε. Thanks to (66) we get

    Vε(c0ε(uε+χε)θ0dy)=ε2tχεθ0dyεtρε(xε(t)+εy,t)θ0dy. (74)

    Note that if we obtain a uniform in t a priori bound of the form uεL2C with C independent of ε, (74) can be resolved with respect to ˙xε=Vε to come up with a well posed system (72)-(74).

    Step 2. (energy estimates for uε and Qε) Represent uε as uε=θ0wε, then multiply the equation (72) by uε and integrate in y over R. Since

    (2yuε+W(θ0)uε)uεdy=(θ0)2(ywε)2dy,

    and

    θ0uεdy=0, yχεuεdy=0, yuεuεdy=0,

    we get

    ε22ddtu2εdy+(θ0)2(ywε)2dy(R1Qεθ0εQεyχε)uεdyεQεyuεuεdy+Cε(u2ε+|uε|3)dy, (75)

    where R1=2yχε+W(θ0)W(θ0+εχε)εε2χεt. Due to the construction of χε we have, R1L2C with C independent of ε and t. Also, by a Poincaré type inequality (see Appendix A)

    (θ0)2(ywε)2dyCθ0uε2H1

    with Cθ0>0 independent of uε. Thus (75) implies that

    ε22ddtuε2L2+Cθ02uε2H1C+C1Qε2L2+ε2yQεu2εdy+Cε|uε|3dy+Cθ02(uε2L22uε2H1)C+C1Qε2L2+εyQε2L2+C2εuε6L2 (76)

    where we have also used the interpolation inequality |u|4dyCuH1u3L2 which yields |u|4dyC(u2H1+u6L2). Next we derive differential inequalities

    εddtQε2L2+yQε2L2+Qε2L2C+Cε2uε2L2, (77)
    εddtyQε2L2+2yQε2L2+yQε2L2C+Cε2uε2H1, (78)

    by multiplying (73) by Qε and 2yQε, and integrating on R.

    Step 3. (uniform bound for uεL2) We show that differential inequalities (76)-(78) imply that uε2L2 remains uniformly bounded on [0,T] when ε>0 is small. To this end fix M>max{1,uε(,0)2L2}, to be specified later, and consider the first time t=¯t(0,T) when uε(,t)2L2 reaches M (if any). We have, uε(,t)2L2<M on (0,¯t) and

    ddtuε2L20at t=¯t. (79)

    It follows from (77) that Qε2L2C+Cε2MεddtQε2L2; the same bound also holds for yQε2L2. Substitute these bounds in (76) and integrate from 0 to ¯t to conclude that

    ¯t0uε2H1dtC(¯t+ε2uε(,0)2L2+εQε(,0)2L2+ε¯tM3)) (80)

    with a constant C independent of ¯t, M and ε. Now integrate (78) from 0 to t, in view of (80) this results in the following pointwise inequality

    yQε2L2C(1ε+ε3uε(,0)2L2+ε2Qε(,0)2L2+ε2M3)+yQε(,,0)2L2

    for all t(0,¯t). Also, Gronwall's inequality applied to (77) yields

    Qε2L2C(1+ε2M)+Qε(,0)2L2t(0,¯t).

    We substitute the latter two bounds into (76) and consider the resulting inequality at t=¯t. In view of (79) we have that u(,¯t)L2u(,¯t)H1 and

    u(,¯t)2H1C(1+Qε(,0)2L2+εyQε(,,0)2L2+ε4uε(,0)2L2+εM3),

    where C is independent of ¯t, M and ε. Thus, taking M bigger than

    ¯M=max{uε(,0)2L2,C(1+Qε(,0)2L2+εyQε(,,0)2L2+ε4uε(,0)2L2)},

    e.g. M:=2¯M, and considering sufficiently small ε>0 we see that u(,¯t)2L2<M. This shows that u(,¯t)2L2<M on [0,T], and the Theorem is proved.

    Note that as a bi-product of the above proof we obtained the integral bound

    T0uε2H1dtC, (81)

    which plays an important role in the following derivation of a reduced system for Vε and Qε.

    The special form of the representation (65) (cf. (66)) together with estimates of Theorem 3 and (81) allow us to derive a system of the form (63)-(64) for Vε and Qε. To this end multiply (72) by θ0(y) and integrate in y over R, this results in

    (c0+ε˜Oε(t))Vε(t)(θ0)2Aεdy+F(t)=εOε(t)+˜oε(t), (82)

    where Aε is the solution of

    εAεt=2yAε+Vε(t)yAεAεβθ0 (83)

    with the initial condition Aε(y,0)=pε(y)(=Qε(y,0)) and

    ˜Oε(t):=(χε+uε)θ0dy,Oε(t):=(12W(˜ξε)χ2ε+W(ξε)χεuε+12W(ˉξε)u2ε)θ0dy+1ε(QεAε)(θ0)2dy+Qεy(χε+uε)θ0dy,˜oε(t):=ε2χεtθ0dy (84)

    with ˜ξε being a bounded function (as well as ξε and ¯ξε). It follows from (69) that ˜oε uniformly converges to 0 as ε0 (|˜oε|Cε if F is Lipschitz or W(0)=W(1)). Next we show that Oε(t) is bounded in L(0,T) uniformly in ε.

    Proposition 2. Let conditions of Theorem 3 be satisfied, then Oε(t) introduced in (84) is bounded uniformly in t[0,T] and ε.

    Proof. By Theorem 3 the first term in (84) is bounded. To estimate the remaining terms represent Qε as Qε=Aε+Bε, where Bε solves

    εBεt=2yBε+VεyBεBεεβ(yχε+yuε) (85)

    with zero initial condition. Multiply this equation by Bε and integrate on R, then multiply (85) by 2yBε and integrate on R to obtain

    εddtBε2L2+Bε2L2Cε2(1+uε2L2),ddtyBε2L2Cε(1+uε2H1). (86)

    After integrating these inequalities from 0 to t we make use of (81) to derive Bε2H1Cε. Also, Gronwall's inequality applied to (86) yields Bε2L2Cε2. Similarly, in order to bound AεL2 and yAεL2 we first get

    εddt(Aε2L2+yAε2L2)+(Aε2L2+yAε2L2)C,

    then apply Gronwall's inequality to conclude that Aε2H1C. Thus,

    1ε|QεAε|(θ0)2dy+|Qεy(χε+uε)θ0dy|=1ε|Bε|(θ0)2dy+|(χε+uε)y(Qεθ0)dy|CεBεL2+C(1+uL2)(AεH1+BεH1)C1.

    From now on ˜Oε, ˜oε and Oε are regarded as given functions in the reduced system (82)-(83), and their influence on the behavior of the system is small. Observe that taking the formal limit as ε0 in the system (82)-(83) leads to (61). Indeed, the formal limit as ε0 in (83) is nothing but (30) whose unique solution is ψ(y;V(t)). Then substituting this function into the limit of (82) yields (61).

    The following Theorem establishes the sharp interface limit for sufficiently small β. We assume that initial data P_\varepsilon(\varepsilon y, 0)=A_\varepsilon(y, 0) are bounded in L^2(\mathbb{R}) by a constant C independent of \varepsilon:

    \|A_{\varepsilon}(\, \cdot\, , 0)\|_{L^2} < C. (87)

    Theorem 4. (Sharp Interface Limit for subcritical \beta) Let A_{\varepsilon}, V_{\varepsilon} be solution of the reduced system (82)-(83) with \tilde{\mathcal{O}_{\varepsilon}}, \mathcal{O}_{\varepsilon}\in L^{\infty}(0, T) and \tilde o_\varepsilon converging to 0 in L^\infty(0, T) as \varepsilon\to 0. Assume also that (87) holds. Then there exists \beta_0>0 (e.g., \forall 0< \beta_0<2/\max\{\|(\theta_0^\prime)^2\|_{L^2}, \sqrt{c_0}\}) such that for 0\leq\beta<\beta_0

    {{V}_{\varepsilon }}(t)\to {{V}_{0}}(t)\ in\ {{L}^{\infty }}(\delta ,T)\ as\ \varepsilon \to 0,\ \ \forall \delta > 0, (88)

    where V_0 is the unique solution of (61).

    Proof. Step 1. (Study of the boundary layer at t=0). We show that the function \eta_{\varepsilon}(y, t)=A_\varepsilon(y, t)-\psi(y, V_0(0) ) behaves as a boundary layer at t=0. Since \psi satisfies \partial_y^2\psi+V_0(0)\partial_y\psi-\psi=\beta\theta_0^\prime, c_0V_0(0)=\int (\theta_0')^2 \psi dy-F(0) and A_\varepsilon, V_\varepsilon solve (82)-83, we have

    \begin{aligned} \varepsilon \partial_t \eta_\varepsilon =&\partial_y^2\eta_{\varepsilon}+V_\varepsilon\partial_y\eta_{\varepsilon}-\eta_\varepsilon+\frac{1}{c_0} \partial_y\psi\int (\theta_0')^2 \eta_\varepsilon dy\\ &+\frac{\partial_y \psi}{c_0+\varepsilon\tilde{\mathcal{O}}_{\varepsilon}} \left( F(0)(1+\varepsilon\tilde{\mathcal{O}}_{\varepsilon}/c_0)-F(t)-\varepsilon\frac{\tilde{\mathcal{O}}_{\varepsilon}} {c_0} \int (\theta_0')^2A_\varepsilon+\varepsilon\mathcal{O}_{\varepsilon}+\tilde {o}_\varepsilon \right). \end{aligned} \label{pde_for_eta} (89)

    Multiply (89) by \eta_\varepsilon and integrate on \mathbb{R},

    \begin{aligned} \frac{\varepsilon}{2}\frac{d}{dt}\|\eta\|_{L^2}^2 +\|\partial_y\eta\|_{L^2}^2 + \|\eta\|_{L^2}^2 &\leq \frac{1}{c_0} \|(\theta^\prime)^2\|_{L^2} \|\partial_y\psi\|_{L^2} \|\eta\|_{L^2} ^2\\&+C \left(|F(0)-F(t)|+|\tilde o_\varepsilon|+\varepsilon\right)(1+\|\eta\|_{L^2}^2 ). \end{aligned}

    Note that \|\partial_y \psi\|_{L^2}^2+\|\psi\|_{L^2}^2=-\beta\int\theta_0^\prime\psi dy, therefore \|\partial_y \psi\|_{L^2}^2\leq \beta^2\| \theta_0^\prime\|_{L^2}^2/4=\beta^2 c_0/4. Thus, if \beta \|(\theta^\prime)^2\|_{L^2}<2, then for sufficiently small \varepsilon and 0<t<\sqrt{\varepsilon} we have

    \label{energy_eq_for_eta} \frac{\varepsilon}{2}\frac{d}{dt}\|\eta\|_{L^2}^2 +\omega \|\eta\|_{L^2} ^2\leq C\left(|F(0)-F(t)|+|\tilde o_\varepsilon|+\varepsilon\right) (90)

    with some \omega>0 independent of \varepsilon. Now apply Gronwall's inequality to (90) to obtain that

    \| \eta_\varepsilon\|_{L^2}^2 \leq Ce^{-2\omega t/\varepsilon}+C\max\limits_{\tau\in(0, t)}\left(|F(0)-F(t)|+|\tilde o_\varepsilon|\right)+C\varepsilon \forall t\in[0, \sqrt{\varepsilon}],

    in particular,

    \label{bound_on_bl} \|A_\varepsilon(\, \cdot\, , \sqrt{\varepsilon})-\psi(\, \cdot\, , V_0(0) )\|_{L^2}\to 0\ \ \text{as}\ \varepsilon\to 0. (91)

    Step 2. (Resetting of (82)-(83) as a fixed point problem). Consider an arbitrary V\in L^\infty(\sqrt{\varepsilon}, T) and define \mathcal{F}_{\varepsilon}:L^{\infty}(\sqrt{\varepsilon}, T)\mapsto L^{\infty}(\sqrt{\varepsilon}, T) by

    \label{def_of_big_F} \mathcal{F}_{\varepsilon}(V):=\frac{1}{c_0+\varepsilon \tilde{\mathcal{O}}_\varepsilon}\left[\int (\theta_0')^2 (A+\tilde{\eta}_\varepsilon) dy-F(t)+\varepsilon \mathcal{O}_\varepsilon+\tilde o_\varepsilon\right], (92)

    where A is the unique solution of

    \left\{ \begin{align} & \quad \varepsilon {{\partial }_{t}}A=\partial _{y}^{2}A+V{{\partial }_{y}}A-A-\beta \theta _{0}^{\prime }, \\ & A(y,\sqrt{\varepsilon })=\psi (y,{{V}_{0}}(0)) \\ \end{align} \right.

    on \mathbb{R}\times (\sqrt{\varepsilon}, T] and \tilde{\eta}_\varepsilon solves

    \left\{ \begin{align} &\quad \varepsilon {{\partial }_{t}}{{{\tilde{\eta }}}_{\varepsilon }}=\partial _{y}^{2}{{{\tilde{\eta }}}_{\varepsilon }}+V{{\partial }_{y}}{{{\tilde{\eta }}}_{\varepsilon }}-{{{\tilde{\eta }}}_{\varepsilon }}, \\ &{{{\tilde{\eta }}}_{\varepsilon }}(y, \sqrt{\varepsilon })={{A}_{\varepsilon }}(y, \sqrt{\varepsilon })-\psi (y, {{V}_{0}}(0)). \\ \end{align} \right.

    Note that thanks to (91),

    \mathop {\max }\limits_{t \in [\sqrt \varepsilon, T]} \|{\tilde \eta _\varepsilon }\|{_{{L^2}}} \to 0, \quad {\rm{as}}\;\varepsilon \to 0. (95)

    It follows from the construction of \mathcal{F}_\varepsilon that V_\varepsilon is a fixed point of this mapping. Next we prove that, for sufficiently small \beta, \mathcal{F}_\varepsilon is a contraction mapping. Consider V_1, V_2\in L^{\infty}(\sqrt{\varepsilon}, T) and let A_1, A_2 be solutions of (93)-(94) with V=V_1 and V=V_2, respectively. The function \bar{A}:=A_1-A_2 solves the following problem

    \left\{ \begin{align} & \quad \varepsilon {{\partial }_{t}}\bar{A}=\partial _{y}^{2}\bar{A}+{{V}_{1}}{{\partial }_{y}}\bar{A}-\bar{A}+({{V}_{1}}-{{V}_{2}}){{\partial }_{y}}{{A}_{2}}, \\ & \bar{A}(y,\sqrt{\varepsilon })=0. \\ \end{align} \right.

    Multiplying equation (96) by \bar{A} and integrating in y we get

    \begin{aligned} \frac{\varepsilon}{2}\frac{\text{d}}{\text{d}t}\| \bar{A} \|^2_{L^2} +\| \bar{A} \|^2_{L^2} +\| \partial_y\bar{A} \|^2_{L^2} &=(V_1-V_2)\int \bar{A}\partial_y A_2 dy =(V_2-V_1)\int A_2 \partial_y \bar{A} dy \\ &\leq |V_1-V_2|^2\frac{ \|A_2\|_{L^2}^2}{4}+ \|\partial_y \bar{A} \|_{L^2}^2.\label{contr_map_energy_eq_for_A_bar} \end{aligned} (98)

    On the other hand every solution A of (93)-(94), in particular A_2, satisfies

    \|A\|^2_{L^2} < c_0\beta^2, \;\;\;t\in [\sqrt{\varepsilon}, T]. (99)

    Indeed, multiplying (93) by A and integrating in y we get

    \varepsilon\frac{d}{dt}\|A\|^2_{L^2}+2\|A\|^2_{L^2}+2\|\partial_y A\|^2_{L^2} =-2\beta \int \theta_0'A dy\leq c_0\beta^2+\|A\|^2_{L^2},

    which yields \varepsilon\frac{d}{dt}\|A\|^2_{L^2}+\|A\|^2_{L^2}\leq c_0\beta^2, the latter inequality in turn implies that \|A\|^2_{L^2}\leq \|\psi\|^2_{L^2}e^{-t/\varepsilon}+\beta^2c_0(1-e^{-t/\varepsilon}) for t\in[\sqrt{\varepsilon}, T]. Observing that \|\psi\|_{L^2}^2\leq\beta^2c_0, we are led to (99).

    Substitute now (99) in (98) to conclude that

    \|\mathcal{F}_{\varepsilon}(V_1)-\mathcal{F}_\varepsilon(V_2)\|^2_{L^{\infty}(\sqrt{\varepsilon}, T)} \leq \frac{c_0}{4}\beta^2 \|V_1-V_2\|^2_{L^{\infty}(\sqrt{\varepsilon}, T)}. (100)

    Thus, for \beta<2/\sqrt{c_0}, \mathcal{F}_{\varepsilon} is a contraction mapping.

    Step 3. Since V_\varepsilon is a fixed point of the mapping \mathcal{F}_{\varepsilon}, we have

    \begin{aligned} \|V_\varepsilon -V_0\|_{L^\infty(\sqrt{\varepsilon}, T)}&= \|\mathcal{F}_{\varepsilon}(V_{\varepsilon})-\mathcal{F}_\varepsilon (V_0)\|_{L^\infty(\sqrt{\varepsilon}, T)} + \|\mathcal{F}_{\varepsilon}(V_0)-V_0\|_{L^\infty(\sqrt{\varepsilon}, T)} \\ & \leq \frac{\sqrt{c_0}}{2}\beta \|V_\varepsilon -V_0\|_{L^\infty(\sqrt{\varepsilon}, T)}+ \|\mathcal{F}_{\varepsilon}(V_0)-V_0\|_{L^\infty(\sqrt{\varepsilon}, T)}. \end{aligned}

    Thus,

    \label{contr_map_one_minus_q} \|V_\varepsilon -V_0\|_{L^\infty(\sqrt{\varepsilon}, T)}\leq \frac{1}{1-\sqrt{c_0}\beta/2} \|\mathcal{F}_{\varepsilon}(V_0)-V_0\|_{L^\infty(\sqrt{\varepsilon}, T)}. (101)

    It remains to prove that

    \|{{\mathcal{F}}_{\varepsilon }}({{\text{V}}_{\text{0}}})\text{-}{{\text{V}}_{\text{0}}}{{\|}_{{{\text{L}}^{\infty }}(\sqrt{\varepsilon }, \text{T})}}\to \text{0}\ \text{as}\ \varepsilon \to \text{0}. (102)

    Step 4. (Proof of (102)). First, we approximate V_0(t), which can be a non-differentiable function, by a smooth function. Namely, construct V_{0\varepsilon}(t)\in C^1[0, T], e.g., as a mollification of V_0(t), such that

    \lim\limits_{\varepsilon\to 0}V_{0\varepsilon}= V_0\text{ in }C[0, T]\text{ and }\left|\frac{d}{dt}V_{0\varepsilon}\right| < \frac{C}{\sqrt{\varepsilon}}, \;\forall t\in [0, T]. (103)

    Let A be the solution of (93)-(94) with V=V_0(t). Consider D_\varepsilon(y, t):=A(y, t)-\psi(y, V_{0\varepsilon}(t)), it satisfies the following equality

    \varepsilon\partial_t D_\varepsilon- \partial^2_y D_\varepsilon- V_0 \partial_y D_\varepsilon +D_\varepsilon=-\varepsilon \frac{\partial \psi}{\partial V}(y;V_{0\varepsilon}(t))\frac{d}{dt}V_{0\varepsilon}+(V_0-V_{0\varepsilon})\partial_y \psi(y, V_{0\varepsilon}(t)) (104)

    on \mathbb{R}\times (\sqrt{\varepsilon}, \infty). Since the right hand side of (104) converges to 0 in L^\infty([0, T], L^2(\mathbb{R})) and the norm of initial values \|D_\varepsilon(y, \sqrt{\varepsilon})\|_{L^2}= \|\psi(y, V_0(0))-\psi(y, V_{0\varepsilon}(\sqrt{\varepsilon}))\|_{L^2}\to 0 as \varepsilon\to 0, we have

    \max\limits_{t\in[\sqrt{\varepsilon}, T]}\| D_\varepsilon\|_{L^2}=0 \quad \text{when}\ \varepsilon\to 0. (105)

    Finally, since \int (\theta_0^\prime)^2\psi(y, V_{0\varepsilon})dy=c_0V_0+F(t)+O(|V_{0\varepsilon}-V_0|) we see that

    |\mathcal{F}_{\varepsilon}(V_0)-V_0|\leq C(|V_{0\varepsilon}-V_0|+\|D_\varepsilon\|_{L^2}+\|\tilde \eta_\varepsilon\|_{L^2}+|\tilde o_\varepsilon|+\varepsilon)

    Then combining (95), (103) and (105) we establish (102), and the Theorem is proved.

    For larger \beta the contraction principle no longer applies and both analysis and the results become more complex. Here the stability analysis of the semigroup generated by a non-local non self-adjoint operator is used in place of the contraction mapping principle.

    In the case where \beta is not small, solutions of (61) are no longer unique, see Fig. 3. However, the original PDE problem (57)-(58) (as well as the reduced system (82)-(83) has the unique solution. This indicates that analysis for large \beta must be complemented by a criterion of how to select the limiting solution of equation (61) among all solutions of this equation.

    Figure 3.  Left: Plot of function \Phi_\beta(V) for \beta =150>\beta_{\text{cr}}; {\it Right}: Plot c_0V- \Phi_\beta(V) for \beta=150 vs F. For -F=1.5 there is one intersection ((61) has one root). For each -F=1.762 and -F=2.264 there are two intersections ((61) has two roots). For -F=2 there are three intersections ((61) has three roots).

    As a first step, we neglect terms \varepsilon\tilde{\mathcal{O}}_{\varepsilon}(t), \varepsilon\mathcal{O}_{\varepsilon}(t) and \tilde o_\varepsilon(t) in the reduced system (82)-(83) and study the system

    \left\{ \begin{align} & {{c}_{0}}{{V}_{\varepsilon }}(t)=\int{{{({{\theta }_{{{0}'}}}(y))}^{2}}}{{f}_{\varepsilon }}(y,t)dy-F(t), \\ & \varepsilon {{\partial }_{t}}{{f}_{\varepsilon }}=\partial _{y}^{2}{{f}_{\varepsilon }}+{{V}_{\varepsilon }}(t){{\partial }_{y}}{{f}_{\varepsilon }}-{{f}_{\varepsilon }}-\beta \theta _{0}^{'} \\ \end{align} \right.

    (in (106)-(107), f_\varepsilon replaces A_\varepsilon from (82)-(83)). Substitute (106) into (107) to rewrite the (106)-(107) as a single equation

    \varepsilon \partial_t f_{\varepsilon}=\partial_y^2 f_{\varepsilon}+\frac{1}{c_0}\left(\int (\theta_0')^2f_{\varepsilon} dy-F(t)\right) \partial_yf_{\varepsilon}-f_{\varepsilon}-\beta\theta_0'. (108)

    In the limit \varepsilon\to 0 this equation (formally) leads to the PDE

    0=\partial_y^2 f_{\varepsilon}+\frac{1}{c_0}\left(\int (\theta_0')^2f_{\varepsilon} dy-F(t)\right) \partial_yf_{\varepsilon}-f_{\varepsilon}-\beta\theta_0'. (109)

    Taking the formal limit is justified below for passing from (108) to (109).

    Remark 6. Equation (108) is a singular perturbation of (109) and both equations are non-autonomous. It is well-known that singular limit problems, including non-autonomous equations, can be reduced to the analysis of large time behavior of autonomous equations. To illustrate this, recall a standard example of an ODE with a small parameter \varepsilon from [20],

    \varepsilon \frac{dz_{\varepsilon}}{dt}=\mathcal{F}(z_{\varepsilon}, t), \;t\in[0, T]. (110)

    Assume that there exists the unique root \phi(t) of \mathcal{F}, i.e., 0=\mathcal{F}(\phi(t), t), \;t\in[0, T]. Then the singular limit \phi(t)=\lim\limits_{\varepsilon\to 0}z_{\varepsilon}(t) holds provided that \phi(t) is a stable root, i.e., all solutions u(\tau) of an autonomous problem \frac{du(\tau)}{d\tau}=\mathcal{F}(u(\tau), t) (t is fixed) converge to the large time limit \phi(t): \mathop {\lim }\limits_{\tau \to \infty } u(\tau ) = \phi (t). Note that the problem (110) has two time scales: a slow time t and a fast time \tau. Also the large-time limit corresponds to \tau \to \infty for a fixed parameter t.

    Note that the equivalence of singular and large-time limits is straightforward for the singularly perturbed autonomous problems (\mathcal{F} does not depend on t in (110)). In this case, the simple rescaling

    \tau :{\rm{ = }}t{\rm{/}}\varepsilon \quad u(\tau ):{\rm{ = }}{z_\varepsilon }(\varepsilon \tau )

    reduces the singular limit problem to a problem of stability of steady state.

    To justify the transition from (108) to (109) we introduce three time scales: slow, fast, and intermediate. More precisely, we employ the following three step procedure: (Ⅰ) partition the interval [0, T] by segments of length \sqrt{\varepsilon} on which the equation (108) is "almost" autonomous (F(t) is "almost" constant on each of these intervals); (Ⅱ) on the first interval (0, \sqrt{\varepsilon}), by appropriate scaling \tau=t/\varepsilon and stability analysis find large-time asymptotics \tau\to\infty (here we used equivalence of singular and large-time limits for autonomous equations); (Ⅲ) use the asymptotics found in (Ⅱ) as initial conditions for the next interval (\sqrt{\varepsilon}, 2\sqrt{\varepsilon}), repeat step (Ⅱ) on this interval, and continue to obtain global asymptotics on [0, T]. A crucial ingredient here is an exponential stability of the linearized problem which prevents accumulating of errors (see bound (124) in Lemma 2).

    Rescale the "fast" time \tau=t/\varepsilon in the unknown f_\varepsilon in (108) and "freeze" time t in F(t) (as described in step (Ⅱ) above)

    \label{eq_for_f} \partial_{\tau} f=\partial_y^2f+\frac{1}{c_0}\left(\int (\theta_0'(y))^2f(y, \tau) dy-F(t)\right)\partial_y f-f-\beta\theta_0', (111)

    here t\in[0, T] is considered to be a fixed parameter. Steady states of (111) are solutions of (109). Let f_0 be such a solution, we define its velocity by

    V_0:=\frac{1}{c_0}\left(\int (\theta_0'(y))^2f_0 dy-F(t)\right), (112)

    then f_0(y)=\psi(y, V_0), where \psi(y;V) is defined in (30). Linearizing equation (111) around f_{0} we obtain

    \partial_\tau f +\mathcal{T}(V_0) f=0, (113)

    where \mathcal{T}(V):L^2(\mathbb R)\mapsto L^2(\mathbb R) is a linear operator parameterized by V\in \mathbb{R} and given by

    \label{def_of_T} \mathcal{T}(V) f:= -\partial_y^2 f -V \partial_y f + f -\frac{1}{c_0}\left(\int (\theta_0')^2 f dy \right)\partial_y\psi(y, V). (114)

    Operator \mathcal{T}(V) is a perturbation of a local operator \mathcal{A}(V) f:= -\partial_y^2 f -V \partial_y f + f by a non-local rank one operator \mathcal{P}(V)f= -\partial_y\psi(y, V)\frac{1}{c_0}(f, (\theta_0^\prime)^2)_{L^2}, where (\, \cdot\, , \, \cdot\, )_{L^2} stands for the standard inner product in L^2(\mathbb{R}). The spectrum \sigma(\mathcal{A}(V)) of operator \mathcal{A}(V) is described by the following straightforward proposition.

    Proposition 3. The spectrum \sigma(\mathcal{A}(V)) consists only of its essential part:

    \sigma(\mathcal{A}(V))=\sigma_{\text{ess}}(\mathcal{A}(V))=\left\{k^2-iVk +1;\, k\in \mathbb R\right\}.

    The spectrum \sigma(\mathcal{T}(V)) of \mathcal{T}(V) is described in

    Theorem 5. (On spectrum of the linearized operator) Consider the part \sigma_p(\mathcal{T}(V)) of the spectrum \sigma(\mathcal{T}(V)) laying in \mathbb C\backslash \sigma_{\text{ess}}(\mathcal{A}). Then \sigma_p(\mathcal{T}(V)) is given by

    \sigma_p(\mathcal{T}(V))=\left\{\lambda\in \mathbb C\backslash \sigma_{\text{ess}}(\mathcal{A});\, \left(\, (\mathcal{A}(V)-\lambda)^{-1}\partial_y \psi, (\theta_0^\prime)^2\right)_{L^2}=c_0\right\}.

    Moreover, all \lambda from \sigma_{\text{p}}(\mathcal{T}(V)) are eigenvalues with finite algebraic multiplicities, and geometric multiplicity one.

    Proof. We suppress dependence of \mathcal{A}, \mathcal{T}, and \eta on V for brevity. Consider \lambda \notin \sigma_{\text{ess}}(\mathcal{A}) \cup\sigma_p(\mathcal{T}) and g\in L^2(\mathbb R). There exists the solution f of

    (\mathcal{A}-\lambda)f-\frac{1}{c_0}\partial_y \psi (f, (\theta_0^\prime )^2)_{L^2} = g, \ \ \text{or}\ f=\frac{1}{c_0}(\mathcal{A}-\lambda)^{-1} \partial_y\psi (f, (\theta_0^\prime )^2)_{L^2}+(\mathcal{A}-\lambda)^{-1}g

    which can be represented as

    \label{Resolvent_applied} f=\frac{1}{c_0}(\mathcal{A}-\lambda)^{-1} \partial_y\psi (f, (\theta_0^\prime )^2)_{L^2}+(\mathcal{A}-\lambda)^{-1}g. (115)

    Eliminate (f, (\theta_0^\prime )^2)_{L^2} from the latter equation to find that

    f=(\mathcal{A}-\lambda)^{-1}\partial_y\psi \frac{\left((\mathcal{A}-\lambda)^{-1}g, (\theta_0^\prime )^2 \right)_{L^2}}{c_0-\left((\mathcal{A}-\lambda)^{-1} \partial_y\psi, (\theta_0^\prime )^2 \right)_{L^2}} +(\mathcal{A}-\lambda)^{-1}g.

    Thus, if \lambda\notin\left\{\lambda\in \mathbb{C}; ((\mathcal{A}-\lambda)^{-1}\partial_y\psi, (\theta_0^\prime )^2)_{L^2}= c_0 \right\}\cup \sigma_{\text{ess}}(\mathcal{A}), then \lambda belongs to the resolvent set of \mathcal{T}.

    Now suppose that \lambda\in \sigma_p(\mathcal{T}(V)). Then ((\mathcal{A}-\lambda)^{-1}\partial_y\psi, (\theta_0^\prime )^2)_{L^2}= c_0 and by Fredholm's theorem applied to (115), \lambda is an eigenvalue of finite multiplicity. Let f be a corresponding eigenfunction, then by (115) we have

    f=\frac{1}{c_0}(\mathcal{A}-\lambda)^{-1} \partial_y\psi (f, (\theta_0^\prime )^2)_{L^2}.

    Take the scalar product of this equality with (\theta_0^\prime)^2 to conclude that \lambda \in \sigma_p(\mathcal{T}) if and only if

    \left((\mathcal{A}-\lambda)^{-1}\partial_y\psi, (\theta_0')^2\right)_{L^2} =c_0, (116)

    and (\mathcal{A}-\lambda)^{-1} \partial_y\psi is the unique (up to multiplication by a constant) eigenfunction.

    Thus, Theorem 5 reduces the study of the part of the spectrum \sigma_p(\mathcal{T})=\sigma(\mathcal{T})\setminus \sigma_{\text{ess}}(\mathcal{A}) of operator \mathcal{T}(V) to the equation (116). Next, using the obtained characterization of the \sigma_p(\mathcal{T}) we study the stability of \mathcal{T}.

    Proposition 4. If \Phi^\prime_\beta(V)\geq c_0, then there exists a real non positive eigenvalue \lambda\in \sigma_p(\mathcal{T}(V)).

    Proof. Consider the function \zeta(\lambda):= \left((\mathcal{A}(V)-\lambda)^{-1}\partial_y\psi, (\theta_0')^2\right)_{L^2} for real \lambda\in (-\infty, 0]. We claim that \zeta(0)=\Phi^\prime(V). Indeed, differentiate (30) to find that

    -\partial_y^2\psi_{V}-V\partial_y\psi_{V}+\psi_V =\partial_y \psi,

    where \psi_{V} denotes the partial derivative of \psi in V. Thus

    \zeta(0)=\left([\mathcal{A}(V)]^{-1}\partial_y\psi, (\theta_0')^2\right)_{L^2} =\left(\psi_V, (\theta_0')^2\right)_{L^2}=\Phi^\prime_\beta(V).

    On the other hand it is easy to see that \zeta(\lambda)\to 0 as \lambda\to-\infty, consequently \zeta(\lambda)=c_0 for some \lambda\in (-\infty, 0]. By Theorem 5 this \lambda is a non positive eigenvalue of \mathcal{T}(V).

    Def 1. Define the set of stable velocities \mathcal{S} by

    \mathcal{S}:=\left\{V\in \mathbb R: \forall \lambda \in\sigma (\mathcal{T}(V))\ \text{has positive real part}\, \right\}, (117)

    where \mathcal{T}(V) is the linearized operator given by (114).

    Remark 7. In the case of 2D sytem (1)-(3) one can expect (yet to be proved) that there exist standing wave solutions with circular symmetry when \Omega is a disk. However our preliminary reasonings show that these solutions are not stable if \Phi_\beta^\prime(0)>c_0 (this latter inequality holds for asymmetric potentials W(\rho) and sufficiently large \beta). This conjecture originates from the fact that zero velocity and its small perturbations does not belong to the set of stable velocities as shown in Proposition 4.

    Proposition 4 implies that the inequality

    \Phi'_\beta(V) < c_0 (118)

    is a necessary condition for stability of V. We hypothesize that (118) is also a sufficient condition, and therefore (118) describes the set \mathcal{S}, that is,

    \mathcal{S}=\left\{V\in\mathbb R: \;\Phi'_\beta(V) < c_0\right\}. (119)

    To support our hypothesis we consider W(\rho)=\frac14\rho^2(\rho-1)^2. In this case, the set \left\{V\in\mathbb R: \;\Phi'_\beta(V)< c_0\right\} is the complement to the open interval (V_{\text{min}}, V_{\text{max}}), where V_{\text{min}} and V_{\text{max}} are the local maximum and minimum, respectively (see Fig. 3 and the sketch of c_0V-\Phi_\beta(V) in Fig. 4). Numerical simulations clearly show that (119) holds. We can also rigorously prove that there exist such V_1 and V_2 that the set of stable velocities \mathcal{S} is non-empty and, moreover, contains the compliment to the open interval (V_1, V_2). This is done by means of Fourier analysis which allows us to rewrite (116) as an integral equation for a complex number \lambda. Details are relegated to Appendix B.

    Figure 4.  Sketch of the function F(V)=-c_0V+\Phi_\beta (V); F(V) has one local minimum, F_{\text{min}}=F(V_{\text{min}}), and one local maximum, F_{\text{max}}=F(V_{\text{max}}). Left: Until F<F_{\max} we stay on the left branch. When F exceeds F_{\max} we jump on the right branch; Right: Until F>F_{\min} we stay on the right branch; When F becomes less than F_{\min} we jump on the left branch. Red arrows on both figures illustrate jumps in velocities..

    In this subsection we formulate the main result on the 1D sharp interface limit in the system (57)-(58) for arbitrary \beta>0. Introduce the following conditions:

    (C1) Let V_0 \in \mathcal{S} solve c_0V_0- \Phi_\beta(V_0)=F(0) and let [0, T_\star] be a time interval such that there exists V(t)\in \mathcal{S} a continuous solution of

    c_0V(t)-\Phi_\beta(V(t))=-F(t), \;\;t\in[0, T_\star], \; V(0)=V_0. (120)

    (C2) Assume that P_{\varepsilon}(x, 0)=p_{\varepsilon}(x/\varepsilon) and \|p_\varepsilon(\, \cdot\, )-\psi(\, \cdot\, , V_0)\|_{L^2}< \delta with a small constant \delta>0 independent of \varepsilon (the function \psi=\psi(y;V_0) is defined by (30)).

    Theorem 6. (Sharp Interface Limit for all \beta) Let x_{\varepsilon} be as in Theorem 3 and assume that conditions (C1) and (C2) hold along with the conditions of Theorem 3. Then x_{\varepsilon}(t) converges to x_0(t) in C^1[0, T_\star], where V(t):= \dot{x}_{0}(t) is the solution of (120) as defined in (C1).

    Theorem 6 justifies the sharp interface equation (120) for any \beta. Its proof consists of two steps: (ⅰ) reduction to a single equation (nonlinear, singularly perturbed) which is done in Section 4.2 and (ⅱ) passage to the limit in this equation based on stability analysis presented below, which is the main ingredient of the proof.

    Remark 8. Condition (C2) is crucial to determine the solution branch in the case if V_0 is not a unique solution of (120).

    Proof of Theorem 6. Rewrite (82)-(83) in the form of the single PDE

    \begin{align} & \varepsilon \frac{\partial {{A}_{\varepsilon }}}{\partial t}=\partial _{y}^{2}{{A}_{\varepsilon }}+\frac{1}{{{c}_{0}}+\varepsilon {{\widetilde{\mathcal{O}}}_{\varepsilon }}(t)}\left( \int{{{(\theta _{0}^{'})}^{2}}}{{A}_{\varepsilon }}dy-F(t)+\varepsilon {{\mathcal{O}}_{\varepsilon }}(t)+{{{\tilde{o}}}_{\varepsilon }}(t) \right){{\partial }_{y}}{{A}_{\varepsilon }} \\ & \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad -{{A}_{\varepsilon }}-\beta \theta _{0}^{'}(y), \\ \end{align} (121)

    Recall that \tilde{\mathcal{O}}_{\varepsilon}(t) and {\mathcal{O}}_{\varepsilon}(t) are uniformly bounded functions, \tilde {o}_\varepsilon(t) tends to 0 uniformly on [0, T_{\ast}] as \varepsilon\to 0. We next pass to the limit in equation (121) using exponential stability (established in (127)) of the semigroup corresponding to the linearized operator. The following local stability result plays the crucial role in the proof.

    Lemma 2. There exist \omega>0 and \delta>0 such that if

    \|A_\varepsilon(\, \cdot\, , t)-\psi(\, \cdot\, , V(t))\|_{L^2}\leq \delta, (122)

    then for any 0<r<1 and sufficiently small \varepsilon, \varepsilon<\varepsilon_0 (T_\ast), the function \eta_\varepsilon(y, t, \tau)=A_{\varepsilon}(y, t+\varepsilon \tau)-\psi(y, V(t)) satisfies

    \label{stab_eq_2} \|\eta_\varepsilon(\, \cdot\, t, \tau)\|^2_{L^2}\leq C\left(e^{-\frac{\omega}{2}\tau}\|\eta_\varepsilon(\, \cdot\, t, 0)\|^2_{L^2}+ \max\limits_{s\in [t, t+\varepsilon \tau]}\left(|F(t)-F(s)|^2+\tilde o_\varepsilon^2(s)\right) +\varepsilon^2\right) (123)

    for 0\leq \tau \leq \frac{1}{\varepsilon^r}. The constants \omega, \delta and C in are independent of t, \tau and \varepsilon.

    This Lemma shows that if the initial data are at distance at most \delta from \psi (in the L^2-norm), then the solution A_\varepsilon(y, t+\varepsilon \tau) approaches \psi exponentially fast in \tau (first term in the RHS of (123)) with a deviation that is bounded from above independently of t (described by the second and the third terms in the RHS of (123)). The conclusion of Theorem 6 immediately follows from this Lemma. Indeed, consider the time interval (0, t_1), t_1:=\sqrt{\varepsilon}. Then by Lemma 2 we obtain

    \label{stab_eq_3} \| A_{\varepsilon}(\, \cdot\, , t_1)-\psi(\, \cdot\, , V(t_1))\|_{L^2}^2\leq C \left(e^{-\frac{\omega}{2\sqrt{\varepsilon}}}\delta+m^2(\sqrt{\varepsilon}) +\max\limits_{s\in [0, T_{\ast}]}\tilde o_\varepsilon^2(s)+\varepsilon^2 \right)+C_1\varepsilon, (124)

    where m denotes the modulus of continuity of F on [0, T_\star]. Choose \varepsilon small enough so that \log\frac{1}{\varepsilon}\leq \frac{\omega}{2}\sqrt{\varepsilon} and the right hand side of (124) is bounded by \delta. Similarly, for intervals (t_1, t_2), where t_2:=2\sqrt{\varepsilon}, (t_2, t_3, where t_3:=3\sqrt{\varepsilon}, etc., we obtain

    \|A_{\varepsilon}(\, \cdot\, , t_i)-\psi(\, \cdot\, , V(t_i))\|^2_{L^2}\leq C\left(\varepsilon +m^2(\sqrt{\varepsilon})+\max\limits_{s\in [0, T_{\ast}]}\tilde o_\varepsilon^2(s)+\varepsilon^2\right)+C_1\varepsilon < \delta.

    To complete the proof of Theorem 6 we again use Lemma 2 to bound \|A_{\varepsilon}(\, \cdot\, , t)-\psi(\, \cdot\, , V(t))\|^2_{L^2} for t\in (t_i, t_{i+1}), i=1, 2\dots.

    Proof of Lemma 2. As in the first step of the proof of Theorem 4, consider the function \eta_{\varepsilon}(y, \tau):=A_{\varepsilon}(y, t+\varepsilon\tau)-\psi(y, V(t)), hereafter t is considered as a fixed parameter. It follows from (121) and (30) that \eta_{\varepsilon} satisfies the following PDE

    \begin{aligned} \frac{\partial \eta_{\varepsilon}}{\partial \tau}+\mathcal{T}\eta_{\varepsilon}=&\frac{\partial_y\eta_\varepsilon}{c_0+\tilde O_\varepsilon} \int(\theta_0^\prime)^2\eta_\varepsilon dy + \frac{\Lambda_\varepsilon}{c_0+\tilde O_\varepsilon}\partial_y\eta_\varepsilon\\&-\frac{\varepsilon \tilde O_\varepsilon}{c_0(c_0+\tilde O_\varepsilon)}\partial_y\psi \int(\theta_0^\prime)^2\eta_\varepsilon dy +\frac{\Lambda_\varepsilon}{c_0+\tilde O_\varepsilon}\partial_y\psi, \end{aligned} \label{DLINNOE_URAVNENIE} (125)

    where

    \begin{aligned} \Lambda_\varepsilon(t, \tau):=&F(t)-F(t+\varepsilon\tau)\\&+\varepsilon O_\varepsilon(t+\varepsilon\tau)+\tilde o_\varepsilon(t+\varepsilon\tau) +\varepsilon\frac{\tilde O_\varepsilon(t+\varepsilon\tau)}{c_0}\left(\int(\theta_0^\prime)^2\psi(y, V(t)) dy\right). \end{aligned}

    Introduce the semigroup operator e^{-\mathcal{T}\tau}, \tau>0 in L^2(\mathbb R), then by Duhamel's principle

    \eta_\varepsilon (\, \cdot\, , \tau )=e^{-\mathcal{T}\tau}\eta_\varepsilon(\, \cdot\, , 0) +\int_0^{\tau} e^{-\mathcal{T}(\tau - \tau')} R_\varepsilon (\, \cdot\, , \tau')d\tau', (126)

    where R_\varepsilon (y, \tau) denotes the right hand side of (125).

    In order to proceed with the proof of Lemma 2 we first prove exponential stability of the semigroup e^{-\mathcal{T}t} and establish its consequences in the following

    Lemma 3. There exists \omega>0 such that

    (ⅰ) the following inequality holds

    \label{exp_stable_semigroup} \|e^{-\mathcal{T}\tau}\|\leq Me^{-\omega \tau}, \;\;\tau\geq 0, (127)

    where \|e^{-\mathcal{T}\tau}\| stands for the operator norm of e^{-\mathcal{T}\tau} in L^2(\mathbb R);

    (ⅱ) for every g(y, t),

    \label{stabb_ineq} \Bigl\|\int\limits_0^{\tau} e^{-\mathcal{T}(\tau-\tau')}\frac{\partial^k g}{\partial y^k}(\, \cdot\, , \tau')d\tau'\Bigr\|^2_{L^2} \leq C \int_0^{\tau} e^{-\omega(\tau-\tau')}\|g(\, \cdot\, , \tau')\|^2_{L^2}d\tau', \quad k=0, 1 (128)

    with a constant C independent of g.

    Moreover, constants \omega, M and C can be chosen independently of t (recall that \mathcal{T}=\mathcal{T}(V(t)) depends on t).

    Proof of Lemma 3. Step 1. (proof of (ⅰ)). For every fixed V\in \mathcal{S}, it follows from Gerhardt-Prúss theorem (see, e.g., [16,31]) that (127) holds with some constants M and \omega>0. However, for later use we need a stronger result, we prove that these constants can be chosen independently of V=V(t) for t\in[0, T_\ast]. To this end we establish the following bound

    \label{sector_bound} \|(\mathcal{T}(V(t))-\lambda-\omega)^{-1}\|\leq \frac{C}{|\lambda|} \ \text{for}\ \lambda\in \Pi_{\varphi_0}:= \{-re^{i\varphi}\, ;\, |\varphi|\leq \pi/2+\varphi_0, r > 0 \}, (129)

    with constants \omega>0, \varphi_0>0 and C all independent of t\in [0, T_\ast]. Then Theorem I.7.7 from [30] yields the inequality \|e^{-(\mathcal{T}(V(t))-\omega)\tau}\|\leq M for \tau>0 with constants \omega>0 and M independent of t, and this latter inequality is equivalent to (127).

    Set \mathcal{T}^\prime(t, \omega):=\mathcal{T}(V(t))-\omega and \mathcal{A}^\prime(t, \omega):=\mathcal{A}(V(t))-\omega. To prove (129) we first derive by Fourier analysis,

    \label{sector_bound_for_A} \|(\mathcal{A}^\prime(t, \omega)-\lambda)^{-1}\| \leq \max\limits_{k\in\mathbb{R}} \frac{1}{|k^2-iV(t)k+1- \lambda-\omega|}\leq \frac{C}{|\lambda|+1} \ \text{for}\ \lambda\in \Pi_{\overline{\varphi}}, (130)

    where \overline{\varphi}=\frac{1}{2}\arctan \frac{1}{\max_t |V(t)|}, constant C is independent of both t\in [0, T_\ast] and 0\leq\omega<1/2. Next we make use of the representation (cf. Theorem 5)

    \begin{aligned} (\mathcal{T}^\prime(t, \omega)-\lambda)^{-1}v =&\frac{((\mathcal{A}^\prime(t, \omega)-\lambda)^{-1}v, (\theta_0^\prime)^2)_{L^2}} {\mu(\lambda;t, \omega)} (\mathcal{A}^\prime(t, \omega)-\lambda)^{-1}\partial_y \psi \\ & + (\mathcal{A}^\prime(t, \omega)-\lambda)^{-1}v, \end{aligned} \label{repr_resolv_mu} (131)

    where \mu(\lambda;t, \omega)=c_0- \left((\mathcal{A}^\prime(t, \omega)-\lambda)^{-1}\partial_y \psi, (\theta_0^\prime)^2\right)_{L^2}.

    It follows from (130) that the family of holomorphic functions \mu(\cdot;t, \omega):\Pi_{\overline{\varphi}}\to \mathbb C satisfies |\mu|>1/2 everywhere but on a fixed bounded subset K of \Pi_{\overline{\varphi}} which is independent of 0\leq \omega\leq 1/2 and t\in [0, T_{\ast}]. On the other hand the functions \mu(\lambda;t, \omega) are uniformly bounded in \{\lambda\in\mathbb{C};\, {\rm Re} \lambda<1/4\} and they depend continuously on t and \omega. Now taking into account the fact that V(t)\in \mathcal{S} for all t\in [0, T_{\ast}] we show that |\mu(\lambda;t, 0)|\geq \mu_0 when \lambda \in K and |\text{Re}\lambda|\leq 2\omega for some 1/2\geq\mu_0>0 and 1/2\geq \omega>0. Indeed, otherwise there is a sequence t_k\to t_0, \lambda_k\to \lambda_0 such that \text{Re}\lambda_0=0 and \mu(\lambda_k;t_k, 0)\to 0. Then, by Montel's theorem, up to extracting a subsequence \mu(\lambda_k;t_k, 0)\to\mu(\lambda_0;t_0, 0), but \mu(\lambda_0;t_0, 0)\neq 0 as V(t_0)\in \mathcal{S} (cf. proof of Theorem 5). Thus there are \varphi_0>0 (\varphi_0\leq\overline{\varphi}) such that |\mu(\lambda;t, \omega)|\geq \mu_0 for \lambda\in \Pi_{\varphi_0}. Using this fact and inequality (130) to bound terms in in (131) we get (129), and therefore (127) holds for some \omega>0 and M, both being independent of t. This result immediately yields (128) for k=0.

    Step 2. (proof of (ⅱ)). To prove (128) for k=1 consider first \tau\geq 1 and show that

    \label{parabolic_regularity} \|e^{-\mathcal{T}\tau}\partial_y g\|_{L^2}\leq Ce^{-\omega\tau}\|g\|^2_{L^2}. (132)

    The idea here is to establish a short time parabolic regularization property. Consider f:=e^{-\mathcal{T}s}\partial_y g, it can be represented as f=\partial_y v with v solving

    \left\{ \begin{align} & \quad {{\partial }_{s}}v=\partial _{y}^{2}v+V{{\partial }_{y}}v-v-\frac{2\psi }{{{c}_{0}}}\int{\theta _{0}^{''}}{{\theta }_{{{0}'}}}vdy, \\ & v(y,0)=g(y). \\ \end{align} \right.

    In a standard way, multiplying (133) by v and integrating in y we get

    \label{soviet_star} \frac{1}{2}\frac{d}{ds}\|v\|^2_{L^2}+\|\partial_y v\|_{L^2}^2\leq C \|v\|^2_{L^2}. (135)

    Then an application of Gronwall's inequality yields the uniform bound

    \|v\|_{L^2}\leq C \|g\|_{L^2}\;\text{ for }\;0\leq s \leq 1.

    Using this bound in (135) we derive

    \int_0^1\|\partial_y v(\, \cdot\, , s)\|^2_{L^2}ds \leq C\|g\|^2_{L^2}.

    It follows that \|\partial_y v(\, \cdot\, , s_0)\|_{L^2}\leq C_1\|g\|_{L^2} for some 0<s_0\leq 1. Then by the semigroup property we have

    \begin{align} & \|{{e}^{-\mathcal{T}\tau }}{{\partial }_{y}}g{{\|}_{{{L}^{2}}}}=\|{{e}^{-\mathcal{T}(\tau -{{s}_{0}})}}{{\partial }_{y}}v(\cdot ,{{s}_{0}}){{\|}_{{{L}^{2}}}}\le M{{e}^{-\omega (\tau -{{s}_{0}})}}{{C}_{1}}\|g{{\|}_{{{L}^{2}}}} \\ & \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \le {{C}_{2}}{{e}^{-\omega \tau }}\|g{{\|}_{{{L}^{2}}}}\quad \text{for }\tau \ge 1, \\ \end{align}

    where we have used (127). The bound (132) being established, we conclude with the estimate

    \label{first_bound} \begin{aligned} \Bigl\|\int_0^{\tau-1}e^{-\mathcal{T}(\tau-\tau')}\partial_y g(\, \cdot\, , \tau')d\tau'\Bigr\|^2_{L^2}&\leq \Bigl(C \int_0^{\tau-1} e^{-\omega(\tau-\tau')}\|g(\, \cdot\, \tau')\|_{L^2}d\tau'\Bigr)^2 \\ &\leq C_1 \int_0^{\tau-1} e^{-\omega(\tau-\tau')}\|g(\, \cdot\, \tau')\|_{L^2}^2d\tau'. \end{aligned} (136)

    To complete the proof of (128) consider

    \tilde{f}(\, \cdot\, ):=\int_{\tau-1}^{\tau}e^{-\mathcal{T}(\tau-\tau')}\partial_y g(\, \cdot\, , \tau')d\tau'=\int_{0}^{1}e^{-\mathcal{T}(1-s)}\partial_y g(\, \cdot\, , \tau-1+s)ds

    (if \tau<1, we set g(y, \tau^\prime)\equiv 0 for \tau^\prime<0).It follows from the definition of \tilde{f} that \tilde{f}(y)=\tilde{v}(1, y), where v solves

    \left\{ \begin{align} & {{\partial }_{s}}\tilde{v}+\mathcal{T}\tilde{v}={{\partial }_{y}}g(y,\tau -1+s), \\ & \quad \quad \tilde{v}(0,y)=0. \\ \end{align} \right.

    Multiply equation (137) by v and integrate in y to obtain

    \begin{align} & \frac{1}{2}\frac{d}{ds}\|\tilde{v}\|_{{{L}^{2}}}^{2}+\|{{\partial }_{y}}\tilde{v}\|_{{{L}^{2}}}^{2}\le -\int{g}(y,\tau -1+s){{\partial }_{y}}\tilde{v}(y,s)dy+C\|\tilde{v}\|_{{{L}^{2}}}^{2} \\ & \quad \quad \quad \quad \quad \quad \quad \quad \quad \le \|{{\partial }_{y}}\tilde{v}\|_{{{L}^{2}}}^{2}+\frac{1}{4}\|g(\cdot ,\tau -1+s)\|_{{{L}^{2}}}^{2}+C\|\tilde{v}\|_{{{L}^{2}}}^{2}. \\ \end{align}

    Now apply Gonwall's inequality. As a result we get

    \|\tilde{v}(\, \cdot\, , 1)\|^2_{L^2}\leq C \int_0^1\|g(\, \cdot\, , \tau-1+s)\|_{L^2}^2 ds.

    Thus

    \begin{aligned} \Bigl\|\int_{\tau-1}^{\tau}e^{-\mathcal{T}(\tau-\tau')}\partial_y g(\, \cdot\, , \tau')d\tau'\Bigr\|^2_{L^2} &\leq C \int_0^1\|g(\, \cdot\, , \tau-1+s)\|^2_{L^2}ds \\&\leq C_1 \int_{\tau-1}^{\tau}e^{-\omega(\tau-\tau')}\|g(\tau')\|^2_{L^2}d\tau'. \end{aligned} \label{second_bound} (139)

    Combining (139) with (136) completes the proof of Lemma 3.

    Now we apply Lemma 3 to (126) to obtain the bound

    \begin{aligned} \|\eta_\varepsilon(\, \cdot\, , \tau)\|_{L^2}^2 \leq& 2M^2 \|\eta_\varepsilon(\, \cdot\, , 0)\|^2_{L^2}e^{-\omega \tau} \\&+\int_0^{\tau}e^{-\omega(\tau-\tau')}\left( C_\ast\|\eta_\varepsilon(\, \cdot\, , 0)\|^4_{L^2}+\delta_\varepsilon\|\eta_\varepsilon(\, \cdot\, , 0)\|^2_{L^2}+\delta_\varepsilon\right) d\tau', \end{aligned}\label{stabbb_ineq} (140)

    for 0\leq\tau\leq 1/\varepsilon^r (0<r<1), where C_\ast depends only on F(t) and T_{\ast} and \delta_\varepsilon=C(\max_{s\in [t, t+\varepsilon \tau]}\left(|F(t)-F(s)|^2+\tilde o_\varepsilon^2(s)\right) +\varepsilon^2). Let \alpha(\tau) be the right hand side of (140). Consider s\in [0, \tau], by (140) we have

    \begin{aligned} \dot{\alpha}(s)= &-\omega \alpha(s) +C_{\ast}\|\eta_\varepsilon(\, \cdot\, , 0)\|^4_{L^2} +\delta_\varepsilon(s)\|\eta_\varepsilon(\, \cdot\, , 0)\|^2_{L^2}+\delta_\varepsilon\\& \leq -\omega \alpha(s) +C_{\ast}\alpha^2(s) +\delta_\varepsilon(\tau)\alpha(s)+\delta_\varepsilon(\tau) \end{aligned}

    and \alpha(0)=2M^2\|\eta_\varepsilon(\, \cdot\, , 0)\|_L^2. Choosing an arbitrary q from the interval

    q \in(0, \omega/{(2C_\ast)}),

    we see that the function \overline{\alpha}(s):=q e^{-\omega s/2}+2\delta_\varepsilon(\tau)/\omega satisfies for sufficiently small \varepsilon the differential inequality

    \dot{\overline{\alpha}}(s)+\omega \overline{\alpha}(s) - C_{\ast}\overline{\alpha}^2(s) -\delta_\varepsilon(\tau)\overline{\alpha}(s)-\delta_\varepsilon(\tau) > 0.

    Therefore, if \alpha(0)\leq \overline{\alpha}(0)=q+2\delta_\varepsilon(\tau)/\omega, then \alpha(s)\leq \overline{\alpha}(s) \forall 0\leq s \leq \tau. Thus we have proved that

    \|\eta_\varepsilon(\, \cdot\, , \tau)\|_{L^2}^2\leq q e^{-\omega \tau/2}+2\delta_\varepsilon(\tau)/\omega,

    provided that \|\eta_\varepsilon(\, \cdot\, , 0)\|_{L^2}\leq \delta with 0<\delta<\sqrt{q}/(\sqrt{2}M). This concludes the proof of Lemma 2 and Theorem 6.

    In view of the above analysis the equation(61) for large \beta may have many solutions of quite complicated structure (e.g., discontinuous). Therefore, we need to introduce a criterion for selection of the "correct" solutions that are limiting solutions to the problem with \varepsilon>0. This is analogous, e.g. to viscosity solutions of Allen-Cahn when physical solutions are obtained (by regularization) in the sharp interface limit \varepsilon\to 0, [14].

    We now introduce such a criterion based on numerical observations and suggested by the stability analysis depicted in Fig. 4. Define the left velocity interval \mathcal{B}_{\text{L}}:=(-\infty, V_{\min}] and the right velocity interval \mathcal{B}_{\text{R}}:=[V_{\max}, \infty) for stable velocities V.

    Assume for simplicity of presentation that function F(t)\in C[0, T] is strictly increasing. Then the solution of (61) is chosen based on the following two criteria

    (Cr1) if V(0)\in \mathcal{B}_L, there is a unique V(t)\in\mathcal{B}_L satisfying (61) for all t\in[0, T]. Note that this V(t) is the only solution which is continuous and never enters the "forbidden" interval [V_{\min}, V_{\max}]

    (Cr2) if V(0)\in \mathcal{B}_R, then for any t\in[0, T] the solution V(t) of (61) is chosen in the right velocity interval \mathcal{B}_R, unless it is impossible (F(t)>F_{\min}, where F_{\min} is defined in Fig. 4). In the latter case V(t) is chosen from the left velocity interval \mathcal{B}_L.

    Intuitively, evolution of the sharp interface velocity can be described as follows. Consider for example the left part of Figure 4 left. As time evolves, the velocity increases along the right green branch until it reaches V_{\max}, then it jumps (along the horizontal red dashed line) to the solution of (61) on the left green branch, and continues increasing along this branch.

    Finally, numerical simulations show that the criterion (Cr2) predicts hysteresis in the system (61). Consider two forcing terms corresponding to the right and the left parts of Fig. 4:

    F_{\downarrow}(t)=-1.0 +(-2.25+1.0)t, \;\;F_{\uparrow}(t)=-2.25+(-1.0+2.25)t

    and \beta=150. For t\in [0,1] both F_{\downarrow}(t) and F_{\uparrow} (t) have the same values but in the opposite order in time t. Fig. 5 (left) depicts the solution of equation (61) according to to the criteria (Cr1) and (Cr2). The red and blue branches coincide when F~\notin~[F_{\min}, F_{\max}]. Moreover, a surprising hysteresis loop is observed when F\in [F_{\min}, F_{\max}].

    Figure 5.  Hysteresis loop in the problem of cell motility. Simulations of V=V(F) Left: (61) Jumping from the left to the right branches and back; Right: PDE system (57)-(58). On both figures arrows show in what direction the system (V(t), F(t)) evolves as time t grows; red curve is for F_{\uparrow}(t), blue curve is for F_{\downarrow}(t).

    We also performed numerical simulations for the original PDE system (57)-(58) for \rho_{\varepsilon}(x, 0)=\theta_0(x/\varepsilon), P_{\varepsilon}(x, 0)=\theta_0'(x/\varepsilon), \varepsilon=0.01, and defining \varepsilon-interface x_{\varepsilon}(t) as a number such that \rho_{\varepsilon}(x_{\varepsilon}(t), t)\approx 0.5 (\rho_{\varepsilon}(+\infty, t)+\rho_{\varepsilon}(-\infty, t)). The branches corresponding to F_\downarrow and F_{\uparrow} are depicted in Fig. 5 (right). The same hysteresis is observed which justifies numerically the above criteria.

    It is well known (see, e.g., [28]) that under conditions (5) on the potential W(\rho) the corresponding standing wave \theta_0 satisfies, for some \alpha_0>1,

    \begin{matrix} \alpha _{0}^{-1}{{e}^{-{{\kappa }_{-}}y}} < {{(\theta _{0}^{'}(y))}^{2}}\le {{\alpha }_{0}}{{e}^{-{{\kappa }_{-}}y}},\ y\le 0 \\ \alpha _{0}^{-1}{{e}^{-{{\kappa }_{+}}y}} < {{(\theta _{0}^{'}(y))}^{2}}\le {{\alpha }_{0}}{{e}^{-{{\kappa }_{+}}y}},\ y\ge 0, \\ \end{matrix} (141)

    where {{\kappa }_{\pm }}=2\sqrt{{W}''((1\pm 1)/2)}. In the case of the symmetric potential W(\rho)=\frac14 \rho^2(\rho-1)^2, \kappa_-=\kappa_+ and the standing wave \theta_0 is explicitly given by \theta_0(y)=\frac{1}{2}(1+\tanh\frac{y}{2\sqrt{2}}).

    Theorem 7. (Poincaré inequality) The following inequality holds

    \int{{{(\theta _{0}^{\prime })}^{2}}}{{(v-\langle v\rangle )}^{2}}dy\le {{C}_{P}}\int{{{(\theta _{0}^{\prime })}^{2}}}{{({{v}^{\prime }})}^{2}}dy,\ \ \forall \ v\in {{C}^{1}}(\mathbb{R}), (142)

    where

    \langle v \rangle=\frac{1}{\int(\theta_0^\prime)^2dy}\int(\theta_0^\prime)^2vdy. (143)

    Proof. Step 1. (Friedrich's inequality). Let u\in C^1(\mathbb R) satisfy u(0)=0. Then we show that the inequality

    \int (\theta_0^\prime)^2u^2dy\leq C_F\int(\theta_0^\prime)^2(u^\prime)^2dy, (144)

    holds with C_F independent of u. Indeed,

    \begin{aligned} &\int_0^{\infty}e^{-\kappa_+ y} u^2 dy =2 \int_0^{\infty} \left(\int_{y}^{\infty}e^{-\kappa_+ t}dt\right) u^\prime\, u dy \leq 2\int_{0}^{\infty}\left(\int_{y}^{\infty}e^{-\kappa_+ t}dt\right) |u^\prime||u|dy\nonumber\\ &=\frac{2}{\kappa_+}\int_0^{\infty}e^{-\kappa_+ y}|u^\prime||u|dy\leq \frac{2}{\kappa_+}\left(\int_{0}^{\infty} e^{-\kappa_+ y}(u^\prime)^2 dy\right)^{1/2}\left(\int_{0}^{\infty} e^{-\kappa_+ y}u^2 dy\right)^{1/2}. \end{aligned}

    Thus,

    \int_0^{\infty} (\theta_0^\prime)^2 u^2 dy\leq \frac{2\alpha_0^2}{\kappa_+}\left(\int_{0}^{\infty} (\theta_0^\prime)^2(u^\prime)^2 dy\right)^{1/2}\left(\int_{0}^{\infty} (\theta_0^\prime)^2u^2 dy\right)^{1/2}.\nonumber

    Dividing this inequality by \left(\int_{0}^{\infty} (\theta_0^\prime)^2u^2 dy\right)^{1/2}, and than taking square of both sides we get

    \int_{0}^{\infty}(\theta_0^\prime)^2u^2dy \leq \frac{4\alpha_0^4}{\kappa_+^2}\int_0^{\infty}(\theta_0^\prime)^2(u^\prime)^2 dy (145)

    Similarly we obtain

    \int_{-\infty}^{0}(\theta_0^\prime)^2u^2dy \leq \frac{c_0^4}{\kappa_-^2}\int_{-\infty}^{0}(\theta_0^\prime)^2(u^\prime)^2 dy, (146)

    Then adding (145) to (145) yields (144).

    Step 2. We prove the Poincaré inequality (142) by contradiction. Namely, assume that there exists a sequence v_n\in C^1(\mathbb R)\cap L^{\infty}(\mathbb R) such that

    \int{{{(\theta _{0}^{\prime })}^{2}}}v_{n}^{2}dy=1,\ \ \int{{{(\theta _{0}^{\prime })}^{2}}}{{v}_{n}}dy=0\text{ and }\int{{{(\theta _{0}^{\prime })}^{2}}}{{(v_{n}^{\prime })}^{2}}dy\to 0.

    Apply Friedrich's inequality (144) to functions v_n(y)-v_n(0):

    \nonumber \int (\theta_0^\prime)^2(v_n(y)-v_n(0))^2 dy \leq C_F \int (\theta_0^\prime)^2 (v_n^\prime)^2 dy\rightarrow 0.

    On the other hand,

    \nonumber \int (\theta_0^\prime)^2(v_n(y)-v_n(0))^2 dy=\int (\theta_0^\prime)^2 v_n^2dy +v_n^2(0)\int(\theta_0^\prime)^2 dy\geq \int (\theta_0^\prime)^2 v_n^2dy.

    Hence,

    \nonumber \int (\theta_0^\prime)^2 v_n^2 dy \rightarrow 0

    which contradicts the normalization \int (\theta_0^\prime)^2 v_n^2 dy=1. The Theorem is proved.

    Corollary 1. Let u\in H^1(\mathbb{R}), then

    \begin{align} &\|u-{{\langle u\rangle }_{\theta _{0}^{\prime }}}\theta _{0}^{\prime }\|_{{{H}^{1}}}^{2}\le C\int{{{(\theta _{0}^{\prime })}^{2}}}{{({{v}^{\prime }})}^{2}}dy, \\ &where\ {{\langle u\rangle }_{\theta _{0}^{\prime }}}=\frac{1}{\int{{{(\theta _{0}^{\prime })}^{2}}}dy}\int{u}\theta _{0}^{\prime }dy\ and\ v=u/\theta _{0}^{\prime }, \\ \end{align} (147)

    with a constant C independent of u.

    Proof. Recall that standing waves \theta_0 of the Allen-Cahn equation along with (141) satisfy

    \begin{align} & \alpha _{1}^{-1}{{e}^{-{{\kappa }_{-}}y}} < {{(\theta _{0}^{\prime \prime }(y))}^{2}}\le {{\alpha }_{1}}{{e}^{-{{\kappa }_{-}}y}},y\le 0 \\ & \alpha _{1}^{-1}{{e}^{-{{\kappa }_{+}}y}} < {{(\theta _{0}^{\prime \prime }(y))}^{2}}\le {{\alpha }_{1}}{{e}^{-{{\kappa }_{+}}y}},\ y\ge 0 \\ \end{align}

    for some \alpha_1>0. Then applying Theorem 7 to v=u/\theta_0^\prime and using density of C^1(\mathbb{R}) in H^1(\mathbb{R}) one derives (147).

    In this appendix we study the set of stable of velocities \mathcal{S}, i.e., the set of such V\in \mathbb R that the point spectrum of the linearized operator \mathcal{T}(V) defined by (114) lies in the right half of the complex plane. We restrict ourselves here to the case W(\rho)=\frac14 \rho^2(\rho-1)^2.

    Theorem 5 implies that if \text{Re}\lambda\leq 0, then \lambda solves the equation (116). Though (116) is a scalar equation with respect to \lambda\in \mathbb C, the evaluation of its left hand side requires solution of the PDE (30). By means of Fourier analysis, we can avoid solving the PDE and rewrite (116) in the form

    \int_{\mathbb R} \frac{-i\beta k \tilde{\theta_0'}\overline{\widetilde{(\theta_0')^2}} }{(k^2-iVk+1)(k^2-iVk+(1-\lambda))}dk=1, (148)

    where \tilde{\theta_0'} and {\widetilde{(\theta_0')^2}} are Fourier transforms of \theta_0' and (\theta_0')^2, respectively. In the case W(\rho)=\frac14 \rho^2(1-\rho)^2:

    \tilde{\theta_0'}(k):=\sqrt{\pi}\text{csch} (\sqrt{2}\pi k), \;\;\;\widetilde{(\theta_0')^2}(k)=\frac{\sqrt{2\pi}}{12}k(2k^2+1)\text{csch}(\sqrt{2}\pi k). (149)

    Introduce \chi(k):=-\frac{\beta\pi \sqrt{2}}{12} k^2 (2k^2+1)\text{csch}^2 (\sqrt{2}\pi k), then equation (148) becomes

    \int_{\mathbb R} \frac{i \chi(k)}{(k^2-iVk+1)(k^2-iVk+(1-\lambda))}dk=1. (150)

    Next, consider \lambda=\lambda_r+i\lambda_i. Denote by \mathcal{H}_\lambda(k) the integrand in (150) and rewrite it in the form

    \begin{align} & {{\mathcal{H}}_{{{\lambda }_{r}}+i{{\lambda }_{i}}}}(k)=-\chi (k)\frac{\left[ Vk({{k}^{2}}+\mu )+({{k}^{2}}+1)(Vk+{{\lambda }_{i}}) \right]}{\left( {{({{k}^{2}}+1)}^{2}}+{{V}^{2}}{{k}^{2}} \right)\left( {{({{k}^{2}}+\mu )}^{2}}+{{(Vk+{{\lambda }_{i}})}^{2}}) \right)} \\ & \quad \quad \quad \quad \quad +i\chi (k)\frac{\left[ ({{k}^{2}}+1)({{k}^{2}}+\mu )-Vk(Vk+{{\lambda }_{i}}) \right]}{\left( {{({{k}^{2}}+1)}^{2}}+{{V}^{2}}{{k}^{2}} \right)\left( {{({{k}^{2}}+\mu )}^{2}}+{{(Vk+{{\lambda }_{i}})}^{2}}) \right)}, \\ \end{align}

    where \mu=1-\lambda_r.

    Proposition 5. (ⅰ) Assume V<\sqrt{2}. If \Phi'_\beta(V)<c_0, then all eigenvalues \lambda \in \sigma_p(\mathcal{T}(V)) have positive real part, \text{Re}\lambda>0.

    (ⅱ) There exists \bar{V}>0 such that for all V>\bar{V} all eigenvalues of \mathcal{T}(V) have positive real part.

    Remark 9. Condition V<\sqrt{2} is a technical assumption in the proof which guarantees that integral (153) is negative. However, numerical simulations show that integral (153) is negative for all V.

    Proof. Part (ⅰ). First, assume 0<|V|<\sqrt{2}. We prove that if \lambda=\lambda_r+i\lambda_i with \lambda_r<1 (\mu>0) is a root of equation \zeta(\lambda)=1, then \lambda_i=0. In particular, the condition \lambda_r<1 guarantees that \lambda\notin \sigma_{\text{ess}}(\mathcal{A}(V)).

    Rewrite the imaginary part of \zeta(\lambda_r+i\lambda_i):

    \begin{align} & \text{Im}\zeta (\lambda )=\int\limits_{-\infty }^{\infty }{{{\mathcal{H}}_{\lambda }}}(k)dk \\ & \quad =\int\limits_{0}^{\infty }{\frac{{{\lambda }_{i}}V\chi (k)(-2({{k}^{2}}+1)({{k}^{2}}+\mu )+{{V}^{2}}{{k}^{2}}-{{({{k}^{2}}+\mu )}^{2}}-\lambda _{i}^{2})}{({{({{k}^{2}}+1)}^{2}}+{{V}^{2}}{{k}^{2}})({{({{k}^{2}}+\mu )}^{2}}+{{(Vk+{{\lambda }_{i}})}^{2}})({{({{k}^{2}}+\mu )}^{2}}+{{(Vk-{{\lambda }_{i}})}^{2}})}}dk. \\ \end{align}

    Since the numerator is the difference between (V^2-2)k^2 and a positive expression, we obtain \text{Im}\zeta(\lambda)\neq 0 for \lambda_i \neq 0.

    Take \lambda_i=0 and rewrite the real part of \zeta(1-\mu):

    \text{Re}\zeta(1-\mu)= - V \int\limits_{-\infty}^{\infty} k\chi(k)\frac{2k^2+1+\mu}{((k^2+1)^2+V^2k^2)((k^2+\mu)^2+V^2k^2)}dk. (151)

    The function \text{Re}\zeta(1-\mu) is obviously monotone for \mu>0. Indeed, denote by \Psi_k(\mu) the term of integrand in (151) which depends on \mu:

    \Psi_k(\mu)=\frac{2k^2+1+\mu}{((k^2+\mu)^2+V^2k^2)}.

    Compute \Psi'_k(\mu):

    \Psi'_k(\mu)=\frac{ (V^2-2-4\mu)k^2-k^4-2\mu-\mu^2}{((k^2+\mu)^2+V^2k^2)^2}. (152)

    If |V|<\sqrt{2}, then \Psi'_{k}(\mu)<0, which proves the monotonicity of \text{Re}\zeta(1-\mu).

    Finally, assume by contradiction that \beta c_0^{-1}\Phi(V)<1, but there exists an eigenvalue \lambda_0 with zero or negative real part, \text{Re}\lambda_0\leq 0. Then \zeta(\text{Re}\lambda_0)=\zeta(\lambda_0)\leq\zeta(0)<1 that contradicts \zeta(\lambda)=1.

    Consider V\leq 0. Then \text{Re}\zeta(\lambda_0)\leq 0. Indeed, observe that \text{Re}\zeta(\lambda_0) equals to

    \int\limits_{0}^{\infty} \frac{-4Vk\chi(k)\left[(2k^2+1+\mu)((k^2+\mu)^2+V^2k^2)\right]dk}{((k^2+1)^2+V^2k^2)\left((k^2+1)^2+(Vk-\lambda_i)^2\right)\left((k^2+\mu)^2+(Vk+\lambda_i)^2)\right)}. (153)

    The integral in (153) is negative or zero and, thus, cannot be equal to 1, so equality (116) does not hold and, in particular, there does not exist eigenvalues with negative real part. Thus, part (ⅰ) is proved.

    Part (ⅱ) follows immediately from (153).

    In this appendix we present the original phase-field model for the motion of a keratocyte cell on a substrait introduced in [40]. It consists ofequations for the phase-field \rho and the orientation vector P:

    {{\partial }_{t}}\rho ={{D}_{\rho }}\Delta \rho -{W}'(\rho )-\alpha \nabla \rho \cdot P, (154)
    {{\partial }_{t}}P={{D}_{P}}\Delta P-\frac{1}{{{\tau }_{1}}}P-\frac{1}{{{\tau }_{2}}}(1-{{\rho }^{2}})P-\beta \nabla \rho -\gamma (\nabla \rho \cdot P)P. (155)

    Coefficients D_{\rho} and D_{P} describe diffusion of \rho and P; \alpha and \beta are the actin protrusion and polymerization strengths; \tau_1 and \tau_2 are decay rates for P (depolymerization) inside and outside the cell; \gamma is the strength of myosin motors. The second term in the right hand side of (154) is defined as follows W'(\rho) = \rho(\delta-\rho)(1-\rho) where

    \delta = \frac{1}{2} +\mu\left(\int \rho dx - V_0\right) - \sigma |P|^2. (156)

    Here \mu is stiffness of the volume constraint, V_0 is the initial area of cell, and \sigma describes contraction due to actin bundles. In this model, the area penalization is introduced via parameter \delta in the double well potential W(\rho) as in the well-known Belousov-Zhabotinskii model [10,24]. A dimensionless parameter \sigma describes contractility due to bundles; in [40] this parameter ranges from 0 to 0.7 (see Table 1 in [40]). For the sake of simplicity in this work we considered the case \sigma=0 only.

    The authors are grateful to I. Aronson and F. Ziebert for useful discussions on the phase field model of cell motility introduced in their paper. The authors wish to thank M. Mizuhara who assisted in the proof-reading of the manuscript.

    [1] Generation, motion and thickness of transition layers for a nonlocal Allen-Cahn equation. Nonlinear Analysis (2010) 72: 3324-3336.
    [2] Convergence of a mass conserving Allen-Cahn equation whose Lagrange multiplier is nonlocal and local. Interfaces Free Bound. (2014) 16: 243-268.
    [3] A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening. Acta Metallurgica (1979) 27: 1085-1095.
    [4] A geometrical approach to front propagation problems in bounded domains with Neumann-type boundary conditions. Interfaces Free Bound. (2016) 5: 239-274.
    [5] Balance between cellsubstrate adhesion and myosin contraction determines the frequency of motility initiation in fish keratocytes. Proceedings of the National Academy of Sciences (2015) 112: 5045-5050.
    [6] Phase-field model of cell motility: Traveling waves and sharp interface limit. Comptes Rendus Mathematique (2016) 354: 986-992.
    [7] K. A. Brakke, The Motion of a Surface by Its Mean Curvature, Mathematical Notes, 20. Princeton University Press, Princeton, N. J. , 1978. i+252 pp.
    [8] A modified phase field approximation for mean curvature flow with conservation of the volume. Mathematical Methods in the Applied Sciences (2011) 34: 1157-1180.
    [9] Volume-preserving mean curvature flow as a limit of a nonlocal Ginzburg-Landau equation. SIAM J. Math. Anal. (1997) 28: 769-807.
    [10] Asymptotic behavior of solutions of an Allen-Cahn equation with a nonlocal term. Nonlinear Analysis: Theory, Methods & Applications (1997) 28: 1283-1298.
    [11] Mass conserving Allen-Cahn equation and volume preserving mean curvature flow. Interfaces Free Bound. (2010) 12: 527-549.
    [12] Uniqueness and existence of viscosity solutions of generalized mean curvature flow equations. J. Differential Geom. (1991) 33: 749-786.
    [13] Motion of level sets by mean curvature. Ⅰ. J. Differential Geom. (1991) 33: 635-681.
    [14] Phase transitions and generalized motion by mean curvature. Comm. Pure Appl. Math. (1992) 45: 1097-1123.
    [15] P. C. Fife, Dynamics of Internal Layers and Diffusive Interfaces, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 1988. vi+93 pp. doi: 10.1137/1.9781611970180
    [16] Spectral theory for contraction semigroups on Hilbert space. Trans. Amer. Math. Soc. (1978) 236: 385-394.
    [17] The volume preserving motion by mean curvature as an asymptotic limit of reaction-diffusion equations. Q. of Appl. Math. (1997) 55: 243-298.
    [18] The heat equation shrinks embedded plane curves to points. J. Differential Geom. (1987) 26: 285-314.
    [19] Three-manifolds with positive Ricci curvature. J. Differential Geom. (1982) 17: 255-306.
    [20] M. H. Holmes, Introduction to Perturbation Methods, 2nd edition. Texts in Applied Mathematics, 20. Springer, New York, 2013. xviii+436 pp. doi: 10.1007/978-1-4614-5477-9
    [21] Flow by mean curvature of convex surfaces into spheres. J. Differential Geom. (1984) 20: 237-266.
    [22] Mechanism of shape determination in motile cells. Nature (2008) 453: 475-480.
    [23] Action minimization and sharpinterface limits for the stochastic Allen-Cahn equation. Comm. Pure Appl. Math. (2007) 60: 393-438.
    [24] Nonlocal front propagation problems in bounded domains with Neumann-type boundary conditions and applications. Asymptot. Anal. (2004) 37: 257-292.
    [25] On an evolution equation in a cell motility model. Phys. D (2016) 318/319: 12-25.
    [26] The gradient theory of phase transitions and the minimal interface criterion. Arch. Rational Mech. Anal. (1987) 98: 123-142.
    [27] Mathematics of cell motility: Have we got its number?. J. Math. Biol. (2009) 58: 105-134.
    [28] Geometrical evolution of developed interfaces. Trans. Amer. Math. Soc. (1995) 347: 1533-1589.
    [29] Invariant measure of the stochastic Allen-Cahn equation: the regime of small noise and large system size. Electron. J. Probab. (2014) 19: 1-76.
    [30] A. Pazy, Semigroups of Linear Operators and Applications to Partial Differential Equations, Applied Mathematical Sciences, 44. Springer-Verlag, New York, 1983. viii+279 pp. doi: 10.1007/978-1-4612-5561-1
    [31] On the spectrum of C0-semigroups. Trans. Amer. Math. Soc. (1984) 284: 847-857.
    [32] Mechanics of motility initiation and motility arrest in crawling cells. J. Mech. Phys. Solids (2015) 84: 469-505.
    [33] P. Recho and L. Truskinovsky, Asymmetry between pushing and pulling for crawling cells, Phys. Rev. E, 87 (2013), 022720. doi: 10.1103/PhysRevE.87.022720
    [34] Multiscale two-dimensional modeling of a motile simple-shaped cell. Multiscale Model. Simul. (2005) 3: 413-439.
    [35] Nonlocal reaction-diffusion equations and nucleation. IMA J. Appl. Math. (1992) 48: 249-264.
    [36] Fast reaction, slow diffusion, and curve shortening. SIAM J. Appl. Math. (1989) 49: 116-133.
    [37] Gamma-convergence of gradient flows on Hilbert and metric spaces and applications. Discrete Contin. Dyn. Syst. A (2011) 31: 1427-1451.
    [38] D. Shao, W. Rappel and H. Levine, Computational model for cell morphodynamics, Physical Review Letters, 105 (2010), 108104. doi: 10.1103/PhysRevLett.105.108104
    [39] S. H. Strogatz, Nonlinear Dynamics and Chaos: With Applications to Physics, Biology, Chemistry, and Engineering, Westview press, 2014, xiii+513 pp.
    [40] Model for self-polarization and motility of keratocyte fragments. J. R. Soc. Interface (2011) 9: 1084-1092.
  • This article has been cited by:

    1. Alessandro Cucchi, Antoine Mellet, Nicolas Meunier, A Cahn--Hilliard Model for Cell Motility, 2020, 52, 0036-1410, 3843, 10.1137/19M1267969
    2. Mohammad Abu Hamed, Alexander A. Nepomnyashchy, A simple model of Keratocyte membrane dynamics: The case of motionless living cell, 2020, 408, 01672789, 132465, 10.1016/j.physd.2020.132465
    3. Alessandro Cucchi, Antoine Mellet, Nicolas Meunier, Self polarization and traveling wave in a model for cell crawling migration, 2022, 42, 1078-0947, 2381, 10.3934/dcds.2021194
    4. Nicolas Bolle, Matthew S. Mizuhara, Dynamics of a cell motility model near the sharp interface limit, 2020, 505, 00225193, 110420, 10.1016/j.jtbi.2020.110420
  • Reader Comments
  • © 2017 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(5199) PDF downloads(123) Cited by(4)

Figures and Tables

Figures(5)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog