
We consider a phase field model of cell motility introduced in [
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
[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 [
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 (
∂ρε∂t=Δρε−1ε2W′(ρε)−Pε⋅∇ρε+λε(t) in Ω, | (1) |
∂Pε∂t=εΔPε−1εPε−β∇ρεin Ω, | (2) |
where
λε(t)=1|Ω|∫Ω(1ε2W′(ρε)+Pε⋅∇ρε)dx. | (3) |
The unknowns here are the scalar phase field function
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
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
Recall the Allen-Cahn equation which is at the core of system (1)-(3),
∂ρε∂t=Δρε−1ε2W′(ρε), | (4) |
where
{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.
Solutions of the stationary Allen-Cahn equation with the volume constraint were studied in [26] by
V=κ−1|Γ(t)|∫Γ(t)κds, | (6) |
where
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
Another technique used in such problems is
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
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
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
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
The parameter
Heuristically, for traveling waves with nonzero velocity, say
Finally, we study the 1D model parabolic problem without any restrictions on
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
∂νρε=0 and Pε=0 on ∂Ω. | (8) |
Introduce the following energy-type functionals
Eε(t):=ε2∫Ω|∇ρε(x,t)|2dx+1ε∫ΩW(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
1 The exponent
Theorem 1. (Bounds on finite time intervals) If the initial data
ε∫T0∫Ω(∂ρε∂t)2dxdt≤C,Eε(t)+Fε(t)≤C∀t∈(0,T), | (12) |
where
Remark 1. This theorem implies that if the initial data are well-prepared in the sense of (10)-(11), then for
Remark 2. A solution of (1)-(2) is understood as follows
ρε∈C([0,T];H1(Ω)), ∂tρε∈L2((0,T)×Ω),Pε∈L2(0,T;H1(Ω)), ∂tPε∈L2((0,T)×Ω) |
and equations (1) and (2) hold in
Proof. First multiply (1) by
∫Ω|∂tρε|2dx+ddt∫Ω(12|∇ρε|2+1ε2W(ρε))dx=−∫ΩPε⋅∇ρε∂tρεdx≤12∫Ω|∂tρε|2dx+12∫Ω|Pε|2|∇ρε|2dx. | (13) |
Here we used the fact that, due to (3), the integral of
∫Ωλε(t)∂tρεdx=0. |
Next, using the maximum principle in (1) we get:
−2ε2supτ∈(0,t]|λε(τ)|≤ρε≤1+2ε2supτ∈(0,t]|λε(τ)|. | (14) |
Let
−ε1/4≤ρε≤1+ε1/4,when t≤Tε, | (15) |
and from now on assume that
Using (13), (15) and integrating by parts we obtain
ddtEε+ε4∫Ω|∂tρε|2dx≤ε∫(|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|)dx≤∫Ω|∂tρε||Pε|2dx+∫Ω|Pε⋅∇ρε||Pε|2dx+∫Ω|∇ρε||∇|Pε|2|dx+1ε2∫Ω|W′(ρε)||Pε|2dx+|λε|∫Ω|Pε|2dx=:5∑i=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,
3∑1Ii≤ε∫Ω(∂tρε)2dx+ε∫Ω|Pε|2|∇ρε|2dx+12ε∫Ω|Pε|4dx+∫Ω|∇|Pε|2|2dx+1εEε. |
Since
1ε2∫Ω|W′(ρ)||Pε|2dx≤Cε2∫ΩW(ρ)dx+12ε2∫Ω|Pε|4dx≤CεEε+12ε2∫Ω|Pε|4dx. |
Finally, in order to bound
|λε(t)|≤Cε2(∫ΩW(ρε)dx)1/2+(∫Ω|∇ρε|2dx)1/2(∫Ω|Pε|2dx)1/2≤Cε(Eεε)1/2+(2Eεε∫Ω|Pε|2dx)1/2, | (17) |
then
I5≤Cε(Eεε)1/2∫Ω|Pε|2dx+(2Eεε)1/2(∫Ω|Pε|2dx)3/2≤CεEε+12ε2∫Ω|Pε|4dx+E2ε+Cε2/3∫Ω|Pε|4dx. |
Thus,
5∑1Ii≤CεEε+E2ε+1+O(ε)ε2∫Ω|Pε|4dx+ε∫Ω(∂tρε)2dx+ε∫Ω|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
14∫Ω|∂tρε|2dx+ddtEεε≤CεEε+E2ε+1ε2∫Ω|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
ddt∫Ω(k|Pε|2+|Pε|4)dx+ε∫Ω(2k|∇Pε|2+4|∇Pε|2|Pε|2+2|∇|Pε|2|2)dx+2ε∫Ω(k|Pε|2+2|Pε|4)dx=−2βk∫ΩPε⋅∇ρεdx+4β∫Ωρεdiv(Pε|Pε|2)dx≤kCε∫Ω|∇ρε|2dx+kε∫Ω|Pε|2dx+ε∫Ω|∇Pε|2|Pε|2dx+C1ε∫Ω|Pε|2dx. |
We chose
ε∫Ω|∇|Pε|2|2dx+1ε∫Ω|Pε|4dx≤CEε−ddt∫Ω(k|Pε|2+|Pε|4)dx. | (19) |
Finally, introducing
dGεdt≤CGε+εG2ε, | (20) |
with a constant
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
Let
x=X0(s,t)+r(s,t)=X(r,s,t),where υ is an inward normal to Γ(t). |
The inverse mapping to
r=±dist(x,Γ(t)),s=S(x,t), |
where in the formula for
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
∂2θ0∂z2=W′(θ0), | (25) |
and
−∂2θ1∂z2+W″(θ0)θ1=V0∂θ0∂z−∂θ0∂zκ(s,t)−(Ψ0⋅υ)∂θ0∂z+λ0(t), | (26) |
−V0∂Ψ0∂z=∂2Ψ0∂z2−Ψ0−β∂θ0∂z, | (27) |
where
It is well-known that there exists a standing wave solution
c0V0(s,t)=c0κ(s,t)+∫(Ψ0⋅ υ)(∂θ0∂z)2dz−λ0(t), where c0=∫R(∂θ0∂z)2dz. | (28) |
Next we obtain the formula for
∂tρε=−θ′0(rε)V0(s,t)ε+O(1). |
Thus, in order to satisfy the condition
∫V(s,t)|∂∂sX0(s,t)|ds = 0. |
Using this fact and integrating (28) with respect to
λ0(t)=∫{c0κ(s,t)+∫(Ψ0⋅υ)(∂θ0∂z)2dz}|∂∂sX0(s,t)|ds. | (29) |
Finally, the unique solution of (27) is given by
∂2zψ+V∂zψ−ψ−βθ′0=0. | (30) |
The representation
∫Ψ0⋅υ(θ′0)2dz=Φβ(V):=∫ψ(z,V)(θ′0)2dz, | (31) |
where we have also taken into account the linearity of (30) in
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
We are interested in (localized in some sense) solutions of (1)-(2) with
0=∂2xρε+V∂xρε−W′(ρε)ε2−Pε∂xρε+λε, | (32) |
0=ε∂2xPε+V∂xPε−1εPε−β∂xρε. | (33) |
Let us postulate an ansatz for the phase field function
ρε=ϕε+εχε+εu, | (34) |
where
ϕε:=θ0((x+a)/ε)θ0((a−x)/ε),χε:=χ−ε+(χ+ε−χ−ε)ϕε, |
constants
Substitute representation (34) in (32)-(33) to find after rescaling the variable
∂2yu−W"(ϕε)u=−V∂yϕε+Pε(∂yϕε+ε∂yu)−λ+1ε(W′(ϕε+εχε)−∂2y(ϕε+εχε))+1ε(W′(ϕε+εχε+εu)−W′(ϕε+εχε)−εW"(ϕε)u)−εV∂y(u+χε), | (35) |
and
∂2yPε+V∂yPε−Pε=β∂yϕε+εβ∂y(χε+u). | (36) |
Note that the ansatz (34) yields the characteristic function of the interval
∂2zu−W′′(θ0(z))u=−Vθ′0(z)+Pεθ′0−λ and ∂2zPε+V∂zPε−Pε=βθ′0(z). |
Resolve the second equation to obtain
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
∂2yu−W"(ϕε)u−∂2yθ(1)ε−W′′(ϕε)θ(1)εθ(1)εu+Hε∫uθ(2)εdy=G(λ,V,Pε,u), | (37) |
where
Hε=1∫(θ(2)ε)2dy(W″(ϕε)θ(2)ε−∂2yθ(2)ε+∂2yθ(1)ε−W′′(ϕε)θ(1)εθ(1)εθ(2)ε) |
and
G(λ,V,Pε,u)=Hε∫uθ(2)εdy−∂2yθ(1)ε−W′′(ϕε)θ(1)εθ(1)εu−V∂yϕε+Pε(∂yϕε+ε∂yu)−εV∂y(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
Qεu:=∂2yu−W"(ϕε)u−∂2yθ(1)ε−W′′(ϕε)θ(1)εθ(1)εu+Hε∫uθ(2)εdy |
has two eigenfunctions
Lemma 1. Let
‖vε‖H1≤C‖fε‖L2 | (39) |
with
Proof. Multiplying
(Qεvε,vε)L2=∫(2∂yθ(1)ε∂ywε+θ(1)ε∂2ywε)θ(1)εwεdy=−∫(θ(1)ε)2(∂ywε)2dy, | (40) |
where the latter equality is obtained via integrating by parts, and the term with
∫(θ(1)ε)2(∂ywε)2dy≤‖fε‖L2‖vε‖L2. | (41) |
The statement of Lemma 1 immediately follows if we prove the following inequality
∫v2εdy≤C∫(θ(1)ε)2(∂ywε)2dy | (42) |
with
∫(θ(1)ε)2(∂ywε)2dy≤C‖fε‖2+Ce−r/ε, | (43) |
and combining (43) with
‖vε‖2H1≤∫(θ(1)ε)2(∂ywε)2dy+C∫v2εdy |
yields (39).
To prove (42) we use the Poincaré inequality (see Appendix A)
∫(θ′0(a/ε±y))2|wε−⟨wε⟩±|2dy≤C∫(θ(1)ε)2(∂ywε)2dy | (44) |
with a constant
⟨wε⟩±:=∫(θ′0(a/ε±y))2wεdy∫(θ′0)2dy. |
Due to orthogonality of
∫(θ′0(a/ε±y))2wεdy=−∫θ′0(y+a/ε)θ′0(a/ε−y)wεdy. |
Thanks to the exponential decay of
|∫(θ′0(a/ε±y))2wεdy|≤e−r/ε(∫(θ(1)ε)2w2εdy)1/2 | (45) |
for sufficiently small
Proposition 1. For sufficiently small
Proof. Given
∂2yu−W′′(0)u+(W′′(0)−W′′(ϕε))u−∂2yθ(1)ε−W′′(ϕε)θ(1)εθ(1)εu+Hε∫uθ(2)εdy=f. | (46) |
Since
Consider now the eigenfunction
Qεqε=Hε∫(θ(2))2dy−θ(2)∫Hε(θ(2)ε−qε)dy |
with Lemma 1 we obtain that
Let us consider now for a given
Pε(y)=ψ0(y+a/ε,V)−ψ0(a/ε−y,−V)+Bε | (47) |
and observe that
∫(∂yBε)2dy+∫(Bε)2dy≤εCM‖Bε‖L2 hence ‖Bε‖H1≤εC1M. |
Now consider
∫G(λ,V,Pε,u)θ(k)εdy=0, k=1,3. | (48) |
Calculate leading terms of (48) for small
W′(ϕε+εχε+εu)−W′(ϕε+εχε)−εW″(ϕε)u=O(ε2) | (49) |
and
W′(ϕε+εχε)−∂2y(ϕε+εχε)=ε(W′′(ϕε)χε−∂2yχε)+O(ε2), | (50) |
where
∫(W′′(ϕε)χε−∂2yχε)θ(k)εdy=∫(W′′(ϕε)θ(k)ε−∂2yθ(k)ε)χεdy, k=1,3 |
tend to zero, when
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
Theorem 2. (Existence of traveling waves) Assume that the equation
Proof. Consider the mapping
‖Q−1εG(λε,Vε,Pε,u)‖H1≤C4+εC5(M), |
with
‖Q−1εG(λε,Vε,Pε,u)‖H1≤M |
for sufficiently small
Also, the mapping
KM,r:={u:‖u‖H1≤M,|u|,|∂yu|≤Me−r(|y|−2aε) when |y|≥2aε}. | (52) |
We claim that for some
|Pε|≤C(1+εM)e−r1(|y|−2a/ε) for |y|≥2a/ε, | (53) |
with
∓∂2yPε∓V∂yPε±Pε≥β∂yϕε+εβ∂y(χε+u) for |y|≥2a/ε. |
This yields pointswise bounds
Thus the image of the convex closed set
u↦Q−1εG(λε,Vε,Pε,u) |
is contained in
Remark 3. Note that
Remark 4. As already mentioned, nontrivial traveling waves appear in the case when
Φ′β(0)=β∫10∫10e−|y(θ0)−y(˜θ0)|(1+|y(θ0)−y(˜θ0)|)W′(˜θ0)√W(θ0)W(˜θ0)dθ0d~θ0, | (54) |
where
In order to have more insight about this dependence assume that the diffusion coefficient in equation (2) for
Φβ(V)=∫χ(θ′0)2dy,−δ∂2yψ−V∂yψ+ψ=−βθ′0. |
One can write down an asymptotic expansion of
ψ=−βθ′0−δβθ′′′0+…,ψV=−βθ′′0−2δβθ(ⅳ)0+…. |
Then we have
Φ′β(0)=−2δβ∫θ(ⅳ)0(θ′0)2dy+O(βδ2), |
which yields, after integrating by parts and using the relations
Φ′β(0)=8√23δβ∫10W′′(ρ)dW3/2(ρ)+O(βδ2). | (55) |
The integral in (55) can be interpreted as a measure of asymmetry of the potential
β>βcritical=3c08√2δ∫10W′′(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
c0V−Φβ(V)=−F, | (56) |
may have multiple roots
F:=1|Γ|∫Γ(κ+Φβ(V))ds−κ. |
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
{∂ρε∂t=∂2xρε−W′(ρε)ε2−Pε∂xρε+F(t)ε,x∈R1, t>0,∂Pε∂t=ε∂2xPε−1εPε−β∂xρε. |
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
We study the asymptotic behavior of solutions to the system (57)-(58) as
ρε(x,0)=θ0(x/ε)+εvε(x/ε), | (59) |
where
ρε(x,t)=θ0(x−xε(t)ε)+εvε(x−xε(t)ε,t). | (60) |
The
The main goal of this Section is to prove that
c0V0(t)=Φβ(V0(t))−F(t), | (61) |
where
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(x,t)=θ0(y)+εχε(y,t)+εuε(y,t), Pε(x,t)=Qε(y,t),y=x−xε(t)ε, | (62) |
where
(ⅱ) Reduction of the system to a single equation. The unknown function
{(c0+o(1))Vε(t)=∫(θ0′)2Aεdy−F(t)+o(1),ε∂Aε∂t=∂2yAε+Vε(t)∂yAε−Aε−βθ′0. |
Taking the limit
(ⅲ) Analysis of reduced problem. For sufficiently small
In order to pass to the limit
ρε(x,t)=θ0(x−xε(t)ε)+εχε(x−xε(t)ε,t)+εuε(x−xε(t)ε,t), | (65) |
with the new unknown function
∫θ′0(y)uε(y,t)dy=0, | (66) |
and
χε(y,t)=χ−ε(t)+θ0(y)(χ+ε(t)−χ−ε(t)), |
where
ε2∂tχ+ε=−W′(1+εχ+ε)ε+F(t),ε2∂tχ−ε=−W′(εχ−ε)ε+F(t) | (67) |
with the initial data
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
|χε(y,t)|+|∂yχε(y,t)|+|∂2yχε"(y,t)|≤C∀t∈[0,T],y∈R, | (68) |
moreover, thanks to the continuity of
ε2‖∂tχε‖L∞→0uniformly on [0,T] as ε→0. | (69) |
Finally, we set
Remark 5. The choice of
∫(θ0(y+xε(0)/ε)−θ0(y))θ′0(y)dy=ε∫(χε(y,0)−vε(y+xε(0)/ε))θ′0(y)dy. |
If
The following theorem justifies the expansions (65) and will be used to obtain a reduced system for unknowns
Theorem 3. (Validation of representation (65)-(66)) Let
‖vε‖L2<C,‖vε‖L∞≤C/ε, | (70) |
and
‖pε‖L2(R)+‖∂ypε‖L2≤C. | (71) |
Then there exists
Proof. Step 1. (coupled system for
W′(θ0+ε(χε+uε))=W′(θ0+εχε)+εW′′(θ0)uε+ε2W′′′(ξε)χεuε+ε22W′′′(¯ξε)u2ε, |
where
ε2∂uε∂t=∂2yuε−W′′(θ0)uε+Vεθ′0−Qεθ′0+∂2yχε+W′(θ0)−W′(θ0+εχε)ε+F(t)−ε2∂χε∂t−εW′′′(ξε)χεuε−ε2W′′′(¯ξε)u2ε−εQε(∂yχε+∂yuε)+εVε(∂yχε+∂yuε). | (72) |
This equation is coupled with that for
ε∂Qε∂t=∂2yQε+Vε∂yQε−Qε−βθ′0−εβ(∂yχε+∂yuε). | (73) |
Finally, considering the solution
Vε(c0−ε∫(uε+χε)θ′′0dy)=ε2∫∂tχεθ′0dy−ε∫∂tρε(xε(t)+εy,t)θ′0dy. | (74) |
Note that if we obtain a uniform in
Step 2. (energy estimates for
∫(−∂2yuε+W′′(θ0)uε)uεdy=∫(θ′0)2(∂ywε)2dy, |
and
∫θ′0uεdy=0, ∫∂yχεuεdy=0, ∫∂yuεuεdy=0, |
we get
ε22ddt∫u2εdy+∫(θ′0)2(∂ywε)2dy≤∫(R1−Qεθ′0−εQε∂yχε)uεdy−ε∫Qε∂yuεuεdy+Cε∫(u2ε+|uε|3)dy, | (75) |
where
∫(θ′0)2(∂ywε)2dy≥Cθ0‖uε‖2H1 |
with
ε22ddt‖uε‖2L2+Cθ02‖uε‖2H1≤C+C1‖Qε‖2L2+ε2∫∂yQεu2εdy+Cε∫|uε|3dy+Cθ02(‖uε‖2L22−‖uε‖2H1)≤C+C1‖Qε‖2L2+ε‖∂yQε‖2L2+C2ε‖uε‖6L2 | (76) |
where we have also used the interpolation inequality
εddt‖Qε‖2L2+‖∂yQε‖2L2+‖Qε‖2L2≤C+Cε2‖uε‖2L2, | (77) |
εddt‖∂yQε‖2L2+‖∂2yQε‖2L2+‖∂yQε‖2L2≤C+Cε2‖uε‖2H1, | (78) |
by multiplying (73) by
Step 3. (uniform bound for
ddt‖uε‖2L2≥0at t=¯t. | (79) |
It follows from (77) that
∫¯t0‖uε‖2H1dt≤C(¯t+ε2‖uε(⋅,0)‖2L2+ε‖Qε(⋅,0)‖2L2+ε¯tM3)) | (80) |
with a constant
‖∂yQε‖2L2≤C(1ε+ε3‖uε(⋅,0)‖2L2+ε2‖Qε(⋅,0)‖2L2+ε2M3)+‖∂yQε(⋅,,0)‖2L2 |
for all
‖Qε‖2L2≤C(1+ε2M)+‖Qε(⋅,0)‖2L2∀t∈(0,¯t). |
We substitute the latter two bounds into (76) and consider the resulting inequality at
‖u(⋅,¯t)‖2H1≤C(1+‖Qε(⋅,0)‖2L2+ε‖∂yQε(⋅,,0)‖2L2+ε4‖uε(⋅,0)‖2L2+εM3), |
where
¯M=max{‖uε(⋅,0)‖2L2,C(1+‖Qε(⋅,0)‖2L2+ε‖∂yQε(⋅,,0)‖2L2+ε4‖uε(⋅,0)‖2L2)}, |
e.g.
Note that as a bi-product of the above proof we obtained the integral bound
∫T0‖uε‖2H1dt≤C, | (81) |
which plays an important role in the following derivation of a reduced system for
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
(c0+ε˜Oε(t))Vε(t)−∫(θ′0)2Aεdy+F(t)=εOε(t)+˜oε(t), | (82) |
where
ε∂Aε∂t=∂2yAε+Vε(t)∂yAε−Aε−βθ′0 | (83) |
with the initial condition
˜Oε(t):=−∫(χε+uε)θ′′0dy,Oε(t):=∫(12W′′′(˜ξε)χ2ε+W′′′(ξε)χεuε+12W′′′(ˉξε)u2ε)θ′0dy+1ε∫(Qε−Aε)(θ′0)2dy+∫Qε∂y(χε+uε)θ′0dy,˜oε(t):=ε2∫∂χε∂tθ′0dy | (84) |
with
Proposition 2. Let conditions of Theorem 3 be satisfied, then
Proof. By Theorem 3 the first term in (84) is bounded. To estimate the remaining terms represent
ε∂Bε∂t=∂2yBε+Vε∂yBε−Bε−εβ(∂yχε+∂yuε) | (85) |
with zero initial condition. Multiply this equation by
εddt‖Bε‖2L2+‖Bε‖2L2≤Cε2(1+‖uε‖2L2),ddt‖∂yBε‖2L2≤Cε(1+‖uε‖2H1). | (86) |
After integrating these inequalities from
εddt(‖Aε‖2L2+‖∂yAε‖2L2)+(‖Aε‖2L2+‖∂yAε‖2L2)≤C, |
then apply Gronwall's inequality to conclude that
1ε∫|Qε−Aε|(θ′0)2dy+|∫Qε∂y(χε+uε)θ′0dy|=1ε∫|Bε|(θ′0)2dy+|∫(χε+uε)∂y(Qεθ′0)dy|≤Cε‖Bε‖L2+C(1+‖u‖L2)(‖Aε‖H1+‖Bε‖H1)≤C1. |
From now on
The following Theorem establishes the sharp interface limit for sufficiently small
\|A_{\varepsilon}(\, \cdot\, , 0)\|_{L^2} < C. | (87) |
Theorem 4. (Sharp Interface Limit for subcritical
{{V}_{\varepsilon }}(t)\to {{V}_{0}}(t)\ in\ {{L}^{\infty }}(\delta ,T)\ as\ \varepsilon \to 0,\ \ \forall \delta > 0, | (88) |
where
Proof. Step 1. (Study of the boundary layer at
\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
\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
\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
\| \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
\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
\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
\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
\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
\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\|^2_{L^2} < c_0\beta^2, \;\;\;t\in [\sqrt{\varepsilon}, T]. | (99) |
Indeed, multiplying (93) by
\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
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
Step 3. Since
\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
\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
\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
\max\limits_{t\in[\sqrt{\varepsilon}, T]}\| D_\varepsilon\|_{L^2}=0 \quad \text{when}\ \varepsilon\to 0. | (105) |
Finally, since
|\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
In the case where
As a first step, we neglect terms
\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),
\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
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 \frac{dz_{\varepsilon}}{dt}=\mathcal{F}(z_{\varepsilon}, t), \;t\in[0, T]. | (110) |
Assume that there exists the unique root
Note that the equivalence of singular and large-time limits is straightforward for the singularly perturbed autonomous problems (
\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
Rescale the "fast" time
\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
V_0:=\frac{1}{c_0}\left(\int (\theta_0'(y))^2f_0 dy-F(t)\right), | (112) |
then
\partial_\tau f +\mathcal{T}(V_0) f=0, | (113) |
where
\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
Proposition 3. The spectrum
\sigma(\mathcal{A}(V))=\sigma_{\text{ess}}(\mathcal{A}(V))=\left\{k^2-iVk +1;\, k\in \mathbb R\right\}. |
The spectrum
Theorem 5. (On spectrum of the linearized operator) Consider the part
\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
Proof. We suppress dependence 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=(\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
Now suppose that
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
\left((\mathcal{A}-\lambda)^{-1}\partial_y\psi, (\theta_0')^2\right)_{L^2} =c_0, | (116) |
and
Thus, Theorem 5 reduces the study of the part of the spectrum
Proposition 4. If
Proof. Consider the function
-\partial_y^2\psi_{V}-V\partial_y\psi_{V}+\psi_V =\partial_y \psi, |
where
\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
Def 1. Define the set of stable velocities
\mathcal{S}:=\left\{V\in \mathbb R: \forall \lambda \in\sigma (\mathcal{T}(V))\ \text{has positive real part}\, \right\}, | (117) |
where
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
Proposition 4 implies that the inequality
\Phi'_\beta(V) < c_0 | (118) |
is a necessary condition for stability of
\mathcal{S}=\left\{V\in\mathbb R: \;\Phi'_\beta(V) < c_0\right\}. | (119) |
To support our hypothesis we consider
In this subsection we formulate the main result on the 1D sharp interface limit in the system (57)-(58) for arbitrary
(C1) Let
c_0V(t)-\Phi_\beta(V(t))=-F(t), \;\;t\in[0, T_\star], \; V(0)=V_0. | (120) |
(C2) Assume that
Theorem 6. (Sharp Interface Limit for all
Theorem 6 justifies the sharp interface equation (120) for any
Remark 8. Condition (C2) is crucial to determine the solution branch in the case if
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
Lemma 2. There exist
\|A_\varepsilon(\, \cdot\, , t)-\psi(\, \cdot\, , V(t))\|_{L^2}\leq \delta, | (122) |
then for any
\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
This Lemma shows that if the initial data are at distance at most
\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
\|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
Proof of Lemma 2. As in the first step of the proof of Theorem 4, consider the function
\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
\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
In order to proceed with the proof of Lemma 2 we first prove exponential stability of the semigroup
Lemma 3. There exists
(ⅰ) the following inequality holds
\label{exp_stable_semigroup} \|e^{-\mathcal{T}\tau}\|\leq Me^{-\omega \tau}, \;\;\tau\geq 0, | (127) |
where
(ⅱ) for every
\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
Moreover, constants
Proof of Lemma 3. Step 1. (proof of (ⅰ)). For every fixed
\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
Set
\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
\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
It follows from (130) that the family of holomorphic functions
Step 2. (proof of (ⅱ)). To prove (128) for
\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
\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
\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
\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
\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
\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
\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
q \in(0, \omega/{(2C_\ast)}), |
we see that the function
\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
\|\eta_\varepsilon(\, \cdot\, , \tau)\|_{L^2}^2\leq q e^{-\omega \tau/2}+2\delta_\varepsilon(\tau)/\omega, |
provided that
In view of the above analysis the equation(61) for large
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
Assume for simplicity of presentation that function
(Cr1) if
(Cr2) if
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
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
We also performed numerical simulations for the original PDE system (57)-(58) for
It is well known (see, e.g., [28]) that under conditions (5) on the potential
\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
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
\int (\theta_0^\prime)^2u^2dy\leq C_F\int(\theta_0^\prime)^2(u^\prime)^2dy, | (144) |
holds with
\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
\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
\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
\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
Corollary 1. Let
\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
Proof. Recall that standing waves
\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
In this appendix we study the set of stable of velocities
Theorem 5 implies that if
\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'}(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
\int_{\mathbb R} \frac{i \chi(k)}{(k^2-iVk+1)(k^2-iVk+(1-\lambda))}dk=1. | (150) |
Next, consider
\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
Proposition 5. (ⅰ) Assume
(ⅱ) There exists
Remark 9. Condition
Proof. Part (ⅰ). First, assume
Rewrite the imaginary part of
\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
Take
\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
\Psi_k(\mu)=\frac{2k^2+1+\mu}{((k^2+\mu)^2+V^2k^2)}. |
Compute
\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
Finally, assume by contradiction that
Consider
\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
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
{{\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
\delta = \frac{1}{2} +\mu\left(\int \rho dx - V_0\right) - \sigma |P|^2. | (156) |
Here
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. ![]() |
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 |