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

Can large inhomogeneities generate target patterns?

Gabriela Jaramillo [email protected] University of Houston, Department of Mathematics,
3551 Cullen Blvd., Houston, TX 77204-3008, USA
Abstract.

We study the existence of target patterns in oscillatory media with weak local coupling and in the presence of an impurity, or defect. We model these systems using a viscous eikonal equation posed on the plane, and represent the defect as a perturbation. In contrast to previous results we consider large defects, which we describe using a function with slow algebraic decay, i.e. gO(1/|x|m)g\sim\mathrm{O}(1/|x|^{m}) for m(1,2]m\in(1,2]. We prove that these defects are able to generate target patterns and that, just as in the case of strongly localized impurities, their frequency is small beyond all orders of the small parameter describing their strength. Our analysis consists of finding two approximations to target pattern solutions, one which is valid at intermediate scales and a second one which is valid in the far field. This is done using weighted Sobolev spaces, which allow us to recover Fredholm properties of the relevant linear operators, as well as the implicit function theorem, which is then used to prove existence. By matching the intermediate and far field approximations we then determine the frequency of the pattern that is selected by the system.

ORCID: 0000-0002-7724-3794
This work is supported by NSF DMS-1911742.

AMS subject classification: 35B36, 35B40, 35Q56, 35Q92

Keywords: Target pattern, spiral waves, bound states of Schrödinger equation.

1. Introduction

Target patterns are coherent structures that emerge in excitable and in oscillatory media. They are characterized by concentric waves that expand away from a center, or core region, creating a ‘bull’s-eye’ pattern. Although often associated with the Belousov-Zhabotinsky reaction [27], they also appear in colonies of slime mold [3, 5], in the oxidation of carbon monoxide on platinum [26], and in brain tissue [24].

In this paper we will focus on target patterns that arise in oscillatory media, where three key mechanisms, or processes, contribute to their formation. The first mechanism is associated with the intrinsic dynamics of the system, which must support a limit cycle that results in uniform time oscillations. The second is a transport process that allows for different spatial regions to interact, such as diffusion in chemical reactions, or coupling between neurons in brain tissue. While these two processes are enough to generate traveling and spiral waves, to obtain target patterns one needs a third ingredient, a defect. Indeed, it is believed that the role of defects, or impurities, is to alter the dynamics of the system in a localized area resulting in a change in the frequency of the time oscillations. As a consequence, these defects act as pacemakers entraining the rest of the medium and forming target patterns.

While experiments and previous analytical results confirm that small localized defects give rise to these patterns, [25, 17, 7, 18, 8, 22, 26, 15, 11, 13], in this paper we want to determine the exact level of localization that is needed to generate them. In particular, assuming the inhomogeneity is modeled as a function with algebraic decay of order O(1/|x|m)\mathrm{O}(1/|x|^{m}), we want to determine how small we can take m>0m>0 and still obtain a well defined target pattern.

To simplify the analysis we concentrate only on systems which involve weak local coupling. Because it is well known that under this assumption the amplitude of oscillations is tied, or enslaved, to the dynamics of the phase, this allows us to focus our analysis on this last variable. Indeed, the results presented in [4] show that coherent structures in these systems are well described by the following viscous eikonal equation

ϕt=Δϕb|ϕ|2εg(x),x2,\phi_{t}=\Delta\phi-b|\nabla\phi|^{2}-\varepsilon g(x),\quad x\in\mathbb{R}^{2}, (1)

where the perturbation, εg\varepsilon g, represents the defect. The above expression is derived using a multiple scale analysis and it therefore models phase changes that occur over long spatial and time scales. In this context, target patterns then correspond to solutions of the form ϕ(x,t)=ϕ~(x)Ωt\phi(x,t)=\tilde{\phi}(x)-\Omega t, satisfying the boundary condition ϕk\nabla\phi\to k as |x||x|\to\infty, where the constant kk then represents the pattern’s wavenumber.

Our motivation for considering large inhomogeneities is three fold. First, in all previous work the level of localization imposed on the inhomogeneity was tied to the tools used to prove the existence of these patterns. Yet, numerical simulation like the ones presented here in Section 6, show that these assumptions can be relaxed. For example, in [23] defects are modeled as functions with compact support and target pattern solutions are found using separation of variables. In contrast, in [15] the authors use spatial dynamics to prove the existence of these patterns. This then allows them to model the impurities as radially symmetric functions with exponential decay. In [13], thanks to the use of weighted Sobolev spaces, this assumption is relaxed and general (non-radially symmetric) inhomogeneities with decay of order O(1/|x|m)\mathrm{O}(1/|x|^{m}), m>2m>2, are considered.

Although using different approaches, the references mentioned above show that target patterns can only be generated by inhomogeneities with a postive and finite mass M=2gM=\int_{\mathbb{R}^{2}}g. This obviously restricts the level of decay of gg to be of order o(1/|x|2)\mathrm{o}(1/|x|^{2}). However, our numerical simulations show that one can obtain target patterns even in the case when the defect is assumed to decay only at order O(1/|x|m)\mathrm{O}(1/|x|^{m}), for m(1,2]m\in(1,2]. We are therefore interested in proving the existence of target patterns for these ‘large’ inhomogeneities of infinite mass.

Our second reason for considering this problem is tied to the existence of spiral waves in oscillatory media with nonlocal coupling. In [10] it was shown that the dynamics of these patterns are well described by the following amplitude equation

0=Kw+λw+α|w|2w+O(ε),w,0<ε<<10=K\ast w+\lambda w+\alpha|w|^{2}w+\mathrm{O}(\varepsilon),\quad w\in\mathbb{C},\qquad 0<\varepsilon<<1

where ww is a radially symmetric complex-valued function, and KK is a symmetric convolution kernel of diffusive type. Additional assumptions on KK imply that formally one can write this operator as (1εΔ1)1σΔ1(1-\varepsilon\Delta_{1})^{-1}\sigma\Delta_{1}, and suggest preconditioning the above equation with (1εΔ1)(1-\varepsilon\Delta_{1}), where Δ1=rr+1rr1r2\Delta_{1}=\partial_{rr}+\frac{1}{r}\partial_{r}-\frac{1}{r^{2}}. This then results in the following expression, which perhaps not surprisingly resembles the complex Ginzburg-Landau equation,

0=βΔ1w+λw+α|w|2w+O(ε),β=(σελ),r[0,).0=\beta\Delta_{1}w+\lambda w+\alpha|w|^{2}w+\mathrm{O}(\varepsilon),\quad\beta=(\sigma-\varepsilon\lambda),\quad r\in[0,\infty).

From there, a similar multiple-scale analysis as the one carried out in [4] and that we also summarize in Appendix, gives a hierarchy of equations at different powers of a small parameter δ=δ(ε)\delta=\delta(\varepsilon). In particular, at order δ2\delta^{2} one finds the steady state viscous eikonal equation,

Ω=Δ1ϕb|ϕ|2δ2g,-\Omega=\Delta_{1}\phi-b|\nabla\phi|^{2}-\delta^{2}g, (2)

as a description of the phase dynamics of spiral waves. However, in contrast to the case of target patterns, here the inhomogeneity does not represent a defect, but is instead related to the small variations, ρ\rho, of the amplitude of the pattern. More precisely, g(1ρ2)g\sim(1-\rho^{2}). Although not immediately obvious, one can check using numerical simulations that the perturbation (1ρ2)(1-\rho^{2}) decays at infinity at order O(1/|x|2)\mathrm{O}(1/|x|^{2}) (see the Appendix A). Therefore, the particular viscous eikonal equation that is connected to the phase dynamics of spiral waves in these systems is the same equation that we are trying to solve.

Finally, our third motivation comes from the following change of variables, ϕ=1blog(Ψ)\phi=-\frac{1}{b}\log(\Psi), which transforms the steady state viscous eikonal equation, (2), into a Schrödinger eigenvalue problem with potential εg\varepsilon g,

ΩΨ=ΔΨ+εg(x)Ψ.\Omega\Psi=\Delta\Psi+\varepsilon g(x)\Psi.

The transformation also shows that our target pattern solutions correspond to bound states of this operator. The only result solving the above eigenvalue problem that we are aware of is that of Simon [21], who proved that in the two dimensional case and under the assumption of localized potentials, i.e. 2g(x)(1+x2)𝑑x<\int_{\mathbb{R}^{2}}g(x)(1+x^{2})\;dx<\infty, bound states exists if and only if the mass 2g(x)𝑑x>0\int_{\mathbb{R}^{2}}g(x)\;dx>0.

Notice that in the context of the Schrödinger operator, our problem corresponds to the ‘supercritical’ case, in the sense that the potential, gg, no longer corresponds to a bounded perturbation of the Laplacian. To see this, fix g(x)=1/(1+|x|)mg(x)=1/(1+|x|)^{m} with 1<m21<m\leq 2, and consider the rescaling y=γxy=\gamma x. The Schrödinger operator then reads γ2ΔyΨ+εγm(γ+|y|)mΨ\gamma^{2}\Delta_{y}\Psi+\varepsilon\frac{\gamma^{m}}{(\gamma+|y|)^{m}}\Psi, and it is then clear that if we choose γ\gamma small, the potential is actually ‘large’ in the far field. Consequently, the results from [21] no longer apply for the case considered here.

In this paper we show that target pattern solutions to the viscous eikonal equation, or equivalently, bound states to the above Schrödinger operator, exists even for these large inhomogeneities. As with small defects, we prove that target pattern solutions have frequencies, Ω\Omega, that are small beyond all orders of the parameter ε\varepsilon. Consequently one cannot use a regular perturbation expansion to justify existence. To resolve this issue we first find two approximations to target patterns, one which is valid at intermediate scales and second one that accounts for the far field behavior of the solution. By matching these two approximations we are then able to determine the unique value of the frequency selected by the system.

It is in the course of this analysis that one sees that the slow decay rate of the inhomogeneity plays a major role in shaping the solution at intermediate scales. This is the main difference between the analysis presented here and that of [13], where inhomogeneities of finite mass are considered. It is also why we will split defects into a core region and a far field region, reflecting the fact that the defects we work with are still too small to alter the shape of the pattern at large scales, but do contribute to the form of the equation at intermediate scales. In particular, we write the impurity as the sum two functions, defined as

gc=(1χD)ggf=χDg,g_{c}=(1-\chi^{D})g\qquad g_{f}=\chi^{D}g, (3)

where χD\chi^{D} is a CC^{\infty} radial cut-off function, with χD(|x|)=0\chi^{D}(|x|)=0 for |x|<D|x|<D and χD(|x|)=1\chi^{D}(|x|)=1 for |x|>2D|x|>2D. To prove the existence of target patterns, the value of the parameter DD can remain arbitrary, so long as it is a finite number. This follows because even though in the above definition the function gcg_{c} has compact support, our results hold for more general ‘core’ functions. The only requirement being that this core defect has finite mass. We therefore make the following assumption.

Hypothesis 1.1.

The inhomogeneity, gg, lives in Hσk(2)H_{\sigma}^{k}(\mathbb{R}^{2}), with k2k\geq 2 and σ(0,1)\sigma\in(0,1), is radially symmetric, and positive. In addition, the defect can be split into the sum of two positive functions, gc,gfg_{c},g_{f}, satisfying

  • The function gfg_{f} is in Hσk(2)H^{k}_{\sigma}(\mathbb{R}^{2}) for 0<σ<10<\sigma<1. In particular, gfO(1/rm)g_{f}\sim\mathrm{O}(1/r^{m}) as rr\to\infty, with 1<m21<m\leq 2, while near the origin gf(|x|)=0g_{f}(|x|)=0 for |x|<1|x|<1.

  • The function gcg_{c} is in Hγk(2)H^{k}_{\gamma}(\mathbb{R}^{2}) for γ>1\gamma>1. In particular, gcO(1/rd)g_{c}\sim\mathrm{O}(1/r^{d}) with d>2d>2 as rr\to\infty.

Remark 1.2.

The spaces Hσk(2)H^{k}_{\sigma}(\mathbb{R}^{2}), with σ\sigma\in\mathbb{R}, are weighted Sobolev spaces with norm

uHσk(2)=|α|s(1+|x|2)σ/2Dαu(x)L2(2).\|u\|_{H^{k}_{\sigma}(\mathbb{R}^{2})}=\sum_{|\alpha|\leq s}\|(1+|x|^{2})^{\sigma/2}D^{\alpha}u(x)\|_{L^{2}(\mathbb{R}^{2})}.

Notice that for positive values of σ\sigma, they impose a level of decay on functions. For a precise definition of these spaces see Section 2.

With the above hypothesis and the approach just described, we prove the following result.

Theorem 1.

Let k2k\geq 2 and σ(0,1)\sigma\in(0,1) and consider a function gHr,σk(2)g\in H^{k}_{r,\sigma}(\mathbb{R}^{2}) satisfying Hypothesis 1.1. Then, there exists a constant ε0>0\varepsilon_{0}>0 and a C1([0,ε0))C^{1}([0,\varepsilon_{0})) family of eigenfunctions ϕ=ϕ(r;ε)\phi=\phi(r;\varepsilon) and eigenvalues Ω=Ω(ε)\Omega=\Omega(\varepsilon) that bifurcate from zero and solve the equation

Δ0ϕb(rϕ)2εg(r)+Ω=0r=|x|[0,).\Delta_{0}\phi-b(\partial_{r}\phi)^{2}-\varepsilon g(r)+\Omega=0\qquad r=|x|\in[0,\infty).

Moreover, this family has the form

ϕ(r;ε)=1bχ1(Λr)log(K0(Λr))+ϕ1(r;ε)+εc,Λ2=bΩ(ε)\phi(r;\varepsilon)=-\frac{1}{b}\chi^{1}(\Lambda r)\log(K_{0}(\Lambda r))+\phi_{1}(r;\varepsilon)+\varepsilon c,\qquad\Lambda^{2}=b\Omega(\varepsilon)

where

  1. i)

    cc is a constant that depends on the initial conditions of the problem,

  2. ii)

    K0(z)K_{0}(z) represents the zeroth-order Modified Bessel function of the second kind,

  3. iii)

    rϕ1Hr,σk(2)\partial_{r}\phi_{1}\in H^{k}_{r,\sigma}(\mathbb{R}^{2}), and

  4. iv)

    Ω=Ω(ε)=C(ε)4e2γεexp[2/a]\Omega=\Omega(\varepsilon)=C(\varepsilon)4e^{-2\gamma_{\varepsilon}}\exp[2/a], with

    a=εb0gc(r)r𝑑r,a=-\varepsilon b\int_{0}^{\infty}g_{c}(r)\;r\;dr,

    and C(ε)C(\varepsilon) a C1C^{1} constant that depends on ε\varepsilon.

Remark 1.3.

Notice that under Hypothesis 1.1 the viscous eikonal equation, (1), is invariant under rotations. As a result we can look for solutions that are radially symmetric. This assumption is made mainly for convenience, and one can follow the steps in [13] to tackle the more general case of non-symmetric inhomogeneties.

Remark 1.4.

If the inhomogeneity gg has strong algebraic decay, i.e. g(r)1/rmg(r)\sim 1/r^{m} with m>2m>2, then we are back in the regime considered in [13]. In this case, the impurity has finite mass and there is no need to split this function into the sum of its core and far field functions. In fact, one can set g=gc=gfg=g_{c}=g_{f} and the above theorem is equivalent to Theorem 1 in [13] with a=εb0g(r)r𝑑r<a=-\varepsilon b\int_{0}^{\infty}g(r)r\;dr<\infty.

Remark 1.5.

While the exact form of the cut-off function χD\chi^{D} appearing in the definition of gcg_{c} is not important for the proof of existence, it does play a role when approximating the pattern’s frequency, Ω\Omega. As our numerical simulations show, there is an optimal way of picking the parameter DD that allows one to obtain better estimates for the frequency, see Section 6. If a non-optimal choice is made, one can improve the estimates for Ω\Omega by using higher order approximations for the intermediate and far field solutions when carrying out the matched asymptotics, see Section 5.

We close this section with some comments regarding the mathematical tools used in this paper. As in reference [13], the proof of existence of solutions is based on the implicit function theorem. This requires that the linearization about our first order approximation, ϕ=ϕ0\phi=\phi_{0}, be an invertible, or at least Fredholm operator with closed range and finite dimensional kernel and cokernel. However, because the equations are posed on the plane, we obtain linear operators that are second order differential operator with essential spectrum near the origin. In addition, the translational symmetry of the system implies that these maps have a zero eigenvalue at the origin. Consequently, these operators are not invertible and they do not have a closed range when posed as maps between standard Sobolev spaces. To overcome this difficulty and recover Fredholm properties for these maps, we work instead with weighted Sobolev spaces. In particular, we make use of the results from [19], where Fredholm properties for the Laplace operator are derived. For other instances where this approach is used to prove existence of patterns see references [9, 11, 13, 12].

1.1. Outline:

The paper is organized as follows. In Section 2 we introduce a special class of weighted Sobolev spaces and summarize Fredholm properties for the Laplacian and related operators. In Section 3 we work with our model (1) and derive from it an equation that is valid at intermediate scales. We then prove existence of solutions to this equation that are bounded near the origin and that have appropriate growth conditions. Next, in Section 4 we work with the full model (1) and, treating the frequency as a parameter, find a first order approximation to target pattern solutions. Then, in Subsection 5.1 we use matched asymptotics to determine the value of the frequency, Ω\Omega, selected by the system. More precisely, we show that Ω\Omega is a C1C^{1} function of the parameter ε\varepsilon. This then allows us to prove existence of solutions using the implicit function theorem. The analysis is complemented by numerical simulations presented in Section 6, and a discussion in Section 7.

2. Preliminaries

In this section two different classes of Sobolev spaces are introduced, weighted Sobolev spaces and Kondratiev spaces. We also look at Fredholm properties for the specific operators that will appear in later sections. We will see how these properties depend on the weighted spaces used to define the domain and range of these operators. Throughout this section we use the symbol x=(1+|x|2)1/2\langle x\rangle=(1+|x|^{2})^{1/2}, which appears in the definition of the norms for the weighted Sobolev spaces introduced.

2.1. Weighted Sobolev Spaces

Let ss be a nonnegative integer, p(1,)p\in(1,\infty), and γ\gamma a real number. We denote by Wγs,p(d)W^{s,p}_{\gamma}(\mathbb{R}^{d}) the space of functions formed by taking the completion of C0(d,)C_{0}^{\infty}(\mathbb{R}^{d},\mathbb{C}) under the norm

uWγs,p(d)=|α|sxγDαu(x)Lp(d).\|u\|_{W^{s,p}_{\gamma}(\mathbb{R}^{d})}=\sum_{|\alpha|\leq s}\|\langle x\rangle^{\gamma}D^{\alpha}u(x)\|_{L^{p}(\mathbb{R}^{d})}.

When p=2p=2 we let Wγs,2(d)=Hγs(d)W^{s,2}_{\gamma}(\mathbb{R}^{d})=H^{s}_{\gamma}(\mathbb{R}^{d}). In this case these spaces are also Hilbert spaces, with inner product defined in the natural way by

f,g=|α|sdf(x)g¯(x)x2γ𝑑x\langle f,g\rangle=\sum_{|\alpha|\leq s}\int_{\mathbb{R}^{d}}f(x)\bar{g}(x)\;\langle x\rangle^{2\gamma}\;dx

where the overbar denotes the complex conjugate.

Notice in particular that depending on the sign of the weight γ\gamma, the functions in these spaces are either allowed to grow (γ<0\gamma<0 ), or forced to decay (γ>0)\gamma>0). We also have natural embeddings, with Wγs,p(d)Wσs,p(d)W^{s,p}_{\gamma}(\mathbb{R}^{d})\subset W^{s,p}_{\sigma}(\mathbb{R}^{d}) provided γ>σ\gamma>\sigma, and Wγs,p(d)Wγk,p(d)W^{s,p}_{\gamma}(\mathbb{R}^{d})\subset W^{k,p}_{\gamma}(\mathbb{R}^{d}) whenever s>ks>k. For 1<p<1<p<\infty we can also identify the dual, (Wγs,p(d))(W^{s,p}_{\gamma}(\mathbb{R}^{d}))^{*}, with the space Wγs,q(d)W^{-s,q}_{-\gamma}(\mathbb{R}^{d}), where pp and qq conjugate exponents.

Kondratiev Spaces: With s,ps,p and γ\gamma as in the previous section, we define Kondratiev spaces as the completion of C0(d,)C_{0}^{\infty}(\mathbb{R}^{d},\mathbb{C}) functions under the norm

uMγs,p(d)=|α|sxγ+|α|Dαu(x)Lp(d)\|u\|_{M^{s,p}_{\gamma}(\mathbb{R}^{d})}=\sum_{|\alpha|\leq s}\|\langle x\rangle^{\gamma+|\alpha|}D^{\alpha}u(x)\|_{L^{p}(\mathbb{R}^{d})}

and denote them by the symbol Mγs,p(d)M^{s,p}_{\gamma}(\mathbb{R}^{d}).

Again we see that these spaces are Hilbert spaces when p=2p=2, with inner product given by

f,g=|α|sdf(x)g¯(x)x2(γ+|α|)𝑑x.\langle f,g\rangle=\sum_{|\alpha|\leq s}\int_{\mathbb{R}^{d}}f(x)\bar{g}(x)\;\langle x\rangle^{2(\gamma+|\alpha|)}\;dx.

We also have the following natural embeddings. One can check that Mγs,p(d)Mγk,p(d)M^{s,p}_{\gamma}(\mathbb{R}^{d})\subset M^{k,p}_{\gamma}(\mathbb{R}^{d}) whenever s>ks>k, and Mγs,p(d)Mσs,p(d)M^{s,p}_{\gamma}(\mathbb{R}^{d})\subset M^{s,p}_{\sigma}(\mathbb{R}^{d}) provided γ>σ\gamma>\sigma. In addition, as in the case of standard Sobolev spaces, one can identify the dual space (Mγs,p(d))(M^{s,p}_{\gamma}(\mathbb{R}^{d}))^{*} with Mγs,q(d)M^{-s,q}_{-\gamma}(\mathbb{R}^{d}), where pp and qq are conjugate exponents.

As was the case with the weighted Sobolev spaces defined above, Kondratiev spaces encode growth or decay depending on the sign of γ\gamma. However, in contrast to Wγs,p(d)W^{s,p}_{\gamma}(\mathbb{R}^{d}), Kondratiev spaces enforce a specific algebraic growth or algebraic decay depending on the value of γ\gamma. In addition, we have the following result which summarizes decay properties for functions in Mγs,p(d)M^{s,p}_{\gamma}(\mathbb{R}^{d}) in terms of γ\gamma.

Lemma 2.1.

Let d,md,m\in\mathbb{N} with m>(d1)/2m>(d-1)/2. Then, for γ>d/2\gamma>-d/2, and for all fMγm+1,2(d)f\in M^{m+1,2}_{\gamma}(\mathbb{R}^{d}) there is a constant C>0C>0 such that

|f(x)|CfLγ+12(d)δfMγm+1,2(d)1δ|x|γd/2,|f(x)|\leq C\|\nabla f\|^{\delta}_{L^{2}_{\gamma+1}(\mathbb{R}^{d})}\|f\|^{1-\delta}_{M^{m+1,2}_{\gamma}(\mathbb{R}^{d})}\;|x|^{-\gamma-d/2},

whenever |x||x| is large and where δ=(d1)/2m\delta=(d-1)/2m.

Proof.

Let (θ,r)(\theta,r) represent spherical coordinates in d\mathbb{R}^{d}, with rr being the radial direction and θ\theta representing the coordinates in the unit sphere, Σ\Sigma. Then,

Σ|f(θ,R)|2𝑑θ\displaystyle\int_{\Sigma}|f(\theta,R)|^{2}\;d\theta Σ(R|rf(θ,s)|𝑑s)2𝑑θ\displaystyle\leq\int_{\Sigma}\left(\int_{\infty}^{R}|\partial_{r}f(\theta,s)|\;ds\right)^{2}\;d\theta
f(,R)L2(Σ)\displaystyle\|f(\cdot,R)\|_{L^{2}(\Sigma)} [Σ(Rsαsγ+1|rf(θ,s)|s(d1)/2𝑑s)2𝑑θ]1/2,\displaystyle\leq\left[\int_{\Sigma}\left(\int^{\infty}_{R}s^{\alpha}s^{\gamma+1}|\partial_{r}f(\theta,s)|s^{(d-1)/2}\;ds\right)^{2}\;d\theta\right]^{1/2},

where α=(γ+1)+(1d)/2\alpha=-(\gamma+1)+(1-d)/2 and RR is fixed. Since the inner integral is squared, we also switched the bounds of integration. Then, applying Minkowski’s inequality for integrals [6, Theorem 6.19] followed by Cauchy Schwartz,

f(,R)L2(Σ)\displaystyle\|f(\cdot,R)\|_{L^{2}(\Sigma)} Rsα[Σs2(γ+1)|rf(θ,s)|2s(d1)𝑑θ]1/2𝑑s\displaystyle\leq\int^{\infty}_{R}s^{\alpha}\left[\int_{\Sigma}s^{2(\gamma+1)}|\partial_{r}f(\theta,s)|^{2}s^{(d-1)}\;d\theta\right]^{1/2}\;ds
[Rs2α]1/2[RΣs2(γ+1)|rf(θ,s)|2s(d1)𝑑θ𝑑s]1/2\displaystyle\leq\left[\int^{\infty}_{R}s^{2\alpha}\right]^{1/2}\left[\int_{\infty}^{R}\int_{\Sigma}s^{2(\gamma+1)}|\partial_{r}f(\theta,s)|^{2}s^{(d-1)}\;d\theta\;ds\right]^{1/2}
C(γ,d)Rα+1/2fLγ+12(d)\displaystyle\leq C(\gamma,d)R^{\alpha+1/2}\|\nabla f\|_{L^{2}_{\gamma+1}(\mathbb{R}^{d})}

where the last line holds provided 2α+1<02\alpha+1<0. We therefore have the following inequality

f(,R)L2(Σ)C(γ,d)Rγd/2fLγ+12(d),γ>d/2.\|f(\cdot,R)\|_{L^{2}(\Sigma)}\leq C(\gamma,d)R^{-\gamma-d/2}\|\nabla f\|_{L^{2}_{\gamma+1}(\mathbb{R}^{d})},\qquad\gamma>-d/2.

Suppose now that fMγm+1,2(d)f\in M^{m+1,2}_{\gamma}(\mathbb{R}^{d}). Using the chain rule we find that |rDΣkf(,R)||Dk+1f|Rk|\partial_{r}D^{k}_{\Sigma}f(\cdot,R)|\leq|D^{k+1}f|R^{k}, where by DΣkD^{k}_{\Sigma} we mean derivatives with respect to unit sphere variables. As a result, similar calculations as the one above show that for integer values of k[1,m]k\in[1,m] we have that

DΣkf(,R)L2(Σ)Rγd/2Dk+1fLγ+k+12(d),γ>d/2.\|D_{\Sigma}^{k}f(\cdot,R)\|_{L^{2}(\Sigma)}\leq R^{-\gamma-d/2}\|D^{k+1}f\|_{L^{2}_{\gamma+k+1}(\mathbb{R}^{d})},\qquad\gamma>-d/2.

Next we use a result from Adams and Fournier [2, Theorem, 5.9], which states that given a domain Σ\Sigma of dimension nn, and conjugate exponents, p,qp,q, satisfying p>1p>1 and mp>nmp>n, 1qp1\leq q\leq p, there exist a constant CC such that for all uWm,p(Σ)u\in W^{m,p}(\Sigma)

uL(Σ)CuWm,p(Σ)δuLq(Σ)(1δ),\|u\|_{L^{\infty}(\Sigma)}\leq C\|u\|_{W^{m,p}(\Sigma)}^{\delta}\|u\|_{L^{q}(\Sigma)}^{(1-\delta)},

where δ=np/(np+(mpn)q)\delta=np/(np+(mp-n)q). Choosing p=q=2p=q=2, n=d1n=d-1, m>(d1)/2m>(d-1)/2, we obtain δ=(d1)/2m\delta=(d-1)/2m and

f(,R)C(m)Rγd/2fMγm+1,2(d)δfLγ+12(d)1δ,\|f(\cdot,R)\|_{\infty}\leq C(m)R^{-\gamma-d/2}\|f\|^{\delta}_{M^{m+1,2}_{\gamma}(\mathbb{R}^{d})}\|\nabla f\|^{1-\delta}_{L^{2}_{\gamma+1}(\mathbb{R}^{d})},

provided fMγm+1,2(d)f\in M^{m+1,2}_{\gamma}(\mathbb{R}^{d}).

Remark 2.2.

Although in the definitions presented above, the spaces Mγs,p(d)M^{s,p}_{\gamma}(\mathbb{R}^{d}) and Wγs,p(d)W^{s,p}_{\gamma}(\mathbb{R}^{d}) consist of complex-valued functions, in what follows we will assume that all functions are real-valued.

2.2. Fredholm operators

In this section and throughout the paper, we use the notation Mr,γs,p(d)M^{s,p}_{r,\gamma}(\mathbb{R}^{d}) and Hr,γs(d)H^{s}_{r,\gamma}(\mathbb{R}^{d}) to denote the subspaces of radially symmetric functions in Mγs,p(d)M^{s,p}_{\gamma}(\mathbb{R}^{d}) and Hγs(d)H^{s}_{\gamma}(\mathbb{R}^{d}), respectively.

The next Lemma shows that for λ>0\lambda>0, the operator r+1r+λ\partial_{r}+\frac{1}{r}+\lambda is invertible in appropriate spaces.

Lemma 2.3.

Let γ\gamma\in\mathbb{R}, λ>0\lambda>0, and kk\in\mathbb{N}. Then, the operator (r+1r+λ):Hr,γk(2)Hr,γk1(2)(\partial_{r}+\frac{1}{r}+\lambda):H^{k}_{r,\gamma}(\mathbb{R}^{2})\longrightarrow H^{k-1}_{r,\gamma}(\mathbb{R}^{2}) has a bounded inverse.

Proof.

A short calculation shows that the kernel of this operator is spanned by the function eλr/r\mathrm{e}^{-\lambda r}/r, which is singular near the origin and is therefore not in Lr,γ2(2)L^{2}_{r,\gamma}(\mathbb{R}^{2}). At the same time, the adjoint of this operator is given by r+λ:Lr,γ2(2)Hr,γ1(2)-\partial_{r}+\lambda:L^{2}_{r,-\gamma}(\mathbb{R}^{2})\longrightarrow H^{-1}_{r,-\gamma}(\mathbb{R}^{2}), and we find that the cokernel is spanned by the function eλr\mathrm{e}^{\lambda r}, which again is not in the space Lr,γ2(2)L^{2}_{r,-\gamma}(\mathbb{R}^{2}) no matter what the value of γ\gamma is. As a result, the kernel and co-kernel of this operator are trivial.

To prove the result, we are left with showing that that the inverse operator

Lr,γ2(2)Hr,γ1(2)f(r)u(r)=1r0reλ(sr)f(s)s𝑑s\begin{array}[]{c c c}L^{2}_{r,\gamma}(\mathbb{R}^{2})&\longrightarrow&H^{1}_{r,\gamma}(\mathbb{R}^{2})\\ f(r)&\mapsto&u(r)=\frac{1}{r}\int_{0}^{r}\mathrm{e}^{\lambda(s-r)}f(s)s\;ds\end{array}

is bounded. To show that uLr,γ2(2)fLr,γ2(2)\|u\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}\leq\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}, we use the inequality

uLr,γ2(2)uLr,γ2(B1)+uLr,γ2(2\B1),\|u\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}\leq\|u\|_{L^{2}_{r,\gamma}(B_{1})}+\|u\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2}\backslash B_{1})},

where B1B_{1} is the unit ball in 2\mathbb{R}^{2}. Then

uLr,γ2(2\B1)=\displaystyle\|u\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2}\backslash B_{1})}= [1|0rf(s)eλ(sr)sr𝑑s|2r2γr𝑑r]1/2\displaystyle\left[\int_{1}^{\infty}\left|\int_{0}^{r}f(s)\mathrm{e}^{\lambda(s-r)}\frac{s}{r}\;ds\right|^{2}\langle r\rangle^{2\gamma}r\;dr\right]^{1/2}
\displaystyle\leq 2|γ+1/2|[1|0rf(s)sγ+1/2eλ(rs)rs|γ+12|ds|2dr]1/2,\displaystyle 2^{|\gamma+1/2|}\left[\int_{1}^{\infty}\left|\int_{0}^{r}f(s)\langle s\rangle^{\gamma+1/2}\quad\mathrm{e}^{-\lambda(r-s)}\langle r-s\rangle^{|\gamma+\frac{1}{2}|}\;ds\right|^{2}\;dr\right]^{1/2},
\displaystyle\leq 2|γ+1/2|[0|0r|f(rz)|rzγ+1/2eλzz|γ+1/2|dz|2𝑑r]1/2\displaystyle 2^{|\gamma+1/2|}\left[\int_{0}^{\infty}\left|\int_{0}^{r}|f(r-z)|\langle r-z\rangle^{\gamma+1/2}\mathrm{e}^{-\lambda z}\langle z\rangle^{|\gamma+1/2|}\;dz\right|^{2}\;dr\right]^{1/2}
\displaystyle\leq 2|γ+1/2|0eλzz|γ+1/2|(z|f(rz)|2rz2γz𝑑z)1/2𝑑z\displaystyle 2^{|\gamma+1/2|\int_{0}^{\infty}e^{-\lambda z}\langle z\rangle^{|\gamma+1/2|}\left(\int_{z}^{\infty}|f(r-z)|^{2}\langle r-z\rangle^{2\gamma}z\;dz\right)^{1/2}\;dz}
\displaystyle\leq C1(γ,λ)fLr,γ2(2)\displaystyle C_{1}(\gamma,\lambda)\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}

where the second line follows from the fact that s/r<1s/r<1 and the relation sσrσ2|σ|rs|σ|\langle s\rangle^{-\sigma}\langle r\rangle^{\sigma}\leq 2^{|\sigma|}\langle r-s\rangle^{|\sigma|}. The inequality on the third line comes from using the change of coordinates z=rsz=r-s and extending the outer limits of integration to zero, while the fourth line follows from an application of Minkowski’s inequality for integrals, [6, Theorem 6.19]. In the final result we let C1(γ,λ)=2|γ+1/2|0eλzz|γ+1/2|𝑑zC_{1}(\gamma,\lambda)=2^{|\gamma+1/2|}\int_{0}^{\infty}e^{-\lambda z}\langle z\rangle^{|\gamma+1/2|}\;dz.

To prove the relation uLr,γ2(B1)CfLr,γ2(2)\|u\|_{L^{2}_{r,\gamma}(B_{1})}\leq C\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}, we bound

|u(r)|\displaystyle|u(r)| 1r0r|f(s)eλ(rs)s|𝑑s\displaystyle\leq\frac{1}{r}\int_{0}^{r}\left|f(s)\mathrm{e}^{-\lambda(r-s)}s\right|\;ds
1r0r|f(s)|s𝑑s\displaystyle\leq\frac{1}{r}\int_{0}^{r}\left|f(s)\right|s\;ds
1r(0r|f(s)|2s𝑑s)1/2(0rs𝑑s)1/2\displaystyle\leq\frac{1}{r}\left(\int_{0}^{r}|f(s)|^{2}s\;ds\right)^{1/2}\left(\int_{0}^{r}s\;ds\right)^{1/2}
1r(0|f(s)|2s2γs𝑑s)1/2(r2)\displaystyle\leq\frac{1}{r}\left(\int_{0}^{\infty}|f(s)|^{2}\langle s\rangle^{2\gamma}s\;ds\right)^{1/2}\left(\frac{r}{\sqrt{2}}\right)
12fLr,γ2(2)\displaystyle\leq\frac{1}{\sqrt{2}}\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}

where the third line follows from Hölder’s inequality. We then obtain that u(r)Lr,γ2(B1)C2(γ)fLr,γ2(2)\|u(r)\|_{L^{2}_{r,\gamma}(B_{1})}\leq C_{2}(\gamma)\;\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})} with C2(γ)=12(01r2γr𝑑r)1/2C_{2}(\gamma)=\frac{1}{\sqrt{2}}\left(\int_{0}^{1}\langle r\rangle^{2\gamma}r\;dr\right)^{1/2}, and consequently

u(r)Lr,γ2(2)(C1(γ,λ)+C2(γ))fLr,γ2(2).\|u(r)\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}\leq(C_{1}(\gamma,\lambda)+C_{2}(\gamma))\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}.

Next, to show that the derivative ruLr,γ2(2)\partial_{r}u\in L^{2}_{r,\gamma}(\mathbb{R}^{2}), we use the equation to write ru=fλuur\partial_{r}u=f-\lambda u-\frac{u}{r}, and thus obtain

ruLr,γ2(2B1)(1+(1+λ)C1(γ,λ))fLr,γ2(2).\|\partial_{r}u\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2}\setminus B_{1})}\leq(1+(1+\lambda)C_{1}(\gamma,\lambda))\;\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}.

To bound ruLr,γ2(B1)\|\partial_{r}u\|_{L^{2}_{r,\gamma}(B_{1})}, notice first that

|u(r)|r1r20r|f(s)|s𝑑s1r20r|f(s)f(y)|s𝑑s+12|f(y)|,\frac{|u(r)|}{r}\leq\frac{1}{r^{2}}\int_{0}^{r}|f(s)|s\;ds\leq\frac{1}{r^{2}}\int_{0}^{r}|f(s)-f(y)|s\;ds+\frac{1}{2}|f(y)|,

where we pick y[0,r]y\in[0,r]. Letting B~(y,2r)=B(y,2r)B(0,r)2\tilde{B}(y,2r)=B(y,2r)\cap B(0,r)\subset\mathbb{R}^{2} , where B(y,2r)B(y,2r) is the ball centered at yy of radius 2r2r, the inequality becomes

|u(r)|r|B~(y,2r)|2πr2(1|B~(y,2r)|B~(y,2r)|f(s)f(y)|s𝑑s𝑑θ)+12|f(y)|.\frac{|u(r)|}{r}\leq\frac{|\tilde{B}(y,2r)|}{2\pi r^{2}}\left(\frac{1}{|\tilde{B}(y,2r)|}\int_{\tilde{B}(y,2r)}|f(s)-f(y)|s\;ds\;d\theta\right)+\frac{1}{2}|f(y)|.

Here, |B~(y,2r)||\tilde{B}(y,2r)| denotes the measure of the set B~(y,2r)\tilde{B}(y,2r). Since L2(B(0,M))L1(B(0,M))L^{2}(B(0,M))\subset L^{1}(B(0,M)) for any ball B(0,M)2B(0,M)\subset\mathbb{R}^{2}, with finite radius MM, we have that ff is in Lloc1(B(0,M))L^{1}_{loc}(B(0,M)). By the Lebesgue Differentiation Theorem [6, Theorem 3.21], the expression in parenthesis approaches zero as r0r\to 0, while the fraction in front remains bounded since |B~(y,2r)|>2πr2|\tilde{B}(y,2r)|>2\pi r^{2}. Therefore, close to the origin, the function |u(r)|/r|u(r)|/r is bounded by f(r)f(r) and, using again the equation ru=fλuu/r\partial_{r}u=f-\lambda u-u/r, we find that

ruLr,γ2(B1)(2+(1+λ)C2(γ))fLr,γ2(2).\|\partial_{r}u\|_{L^{2}_{r,\gamma}(B_{1})}\leq(2+(1+\lambda)C_{2}(\gamma))\;\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}.

It then follows that

ruLr,γ2(2)(3+(1+λ)(C1(γ,λ)+C2(γ)))fLr,γ2(2).\|\partial_{r}u\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}\leq(3+(1+\lambda)(C_{1}(\gamma,\lambda)+C_{2}(\gamma)))\|f\|_{L^{2}_{r,\gamma}(\mathbb{R}^{2})}.

The above calculations then show that the map λ=r+1r+λ:Hr,γ1(2)Lr,γ2(2)\mathcal{L}_{\lambda}=\partial_{r}+\frac{1}{r}+\lambda:H^{1}_{r,\gamma}(\mathbb{R}^{2})\longrightarrow L^{2}_{r,\gamma}(\mathbb{R}^{2}) is invertible. Moreover, we also obtain that the operator norm of its inverse satisfies,

λ13+(2+λ)(C1(γ,λ)+C2(γ)).\|\mathcal{L}^{-1}_{\lambda}\|\leq 3+(2+\lambda)(C_{1}(\gamma,\lambda)+C_{2}(\gamma)).

To extend the result to the more general operator r+1r+λ:Hr,γk(2)Hr,γk1(2)\partial_{r}+\frac{1}{r}+\lambda:H^{k}_{r,\gamma}(\mathbb{R}^{2})\longrightarrow H^{k-1}_{r,\gamma}(\mathbb{R}^{2}) one can proceed by induction: Assuming that ff and uu are in Hr,γk1(2)H^{k-1}_{r,\gamma}(\mathbb{R}^{2}) one shows that rku\partial_{r}^{k}u is in Lr,γ2(2)L^{2}_{r,\gamma}(\mathbb{R}^{2}) using the relation rku=rk1(fuur)\partial^{k}_{r}u=\partial_{r}^{k-1}(f-u-\frac{u}{r}). The fact that rk1(ur)\partial_{r}^{k-1}\left(\dfrac{u}{r}\right) is in the correct space follows by a similar argument as the one done above to prove u/rLr,γ2(2)u/r\in L^{2}_{r,\gamma}(\mathbb{R}^{2}).

To simplify notation, we define (λ)=r+1r+λ\mathcal{L}(\lambda)=\partial_{r}+\frac{1}{r}+\lambda and prove in the next Lemma that its inverse, defined over appropriate weighted spaces, is continuously differentiable with respect to the parameter λ\lambda.

Lemma 2.4.

Let λ>0\lambda>0 and k{0}k\in\mathbb{N}\cup\{0\} and consider the operator defined by (λ)u=ru+1ru+λu\mathcal{L}(\lambda)u=\partial_{r}u+\frac{1}{r}u+\lambda u. Then, its inverse,

1(λ):Hr,γk(2)Hr,γk+1(2)\begin{array}[]{c c c c}\mathcal{L}^{-1}(\lambda):&H^{k}_{r,\gamma}(\mathbb{R}^{2})&\longrightarrow&H^{k+1}_{r,\gamma}(\mathbb{R}^{2})\\ \end{array}

is C1C^{1} with respect to the parameter λ\lambda.

Proof.

Lemma 2.3 shows that 1(λ)\mathcal{L}^{-1}(\lambda), with the specified domain and range, is a bounded operator for all λ(0,)\lambda\in(0,\infty). To prove the continuity of this operator with respect to λ\lambda we must show that given fHr,γk(2)f\in H^{k}_{r,\gamma}(\mathbb{R}^{2}),

supfHr,γk=1(1(λ+h)1(λ))fHr,γk+1Ch.\sup_{\|f\|_{H^{k}_{r,\gamma}}=1}\|(\mathcal{L}^{-1}(\lambda+h)-\mathcal{L}^{-1}(\lambda))f\|_{H^{k+1}_{r,\gamma}}\leq Ch.

Using the notation ϕ(λ)=1(λ)f\phi(\lambda)=\mathcal{L}^{-1}(\lambda)f, we notice that

(1(λ+h)1(λ))f=\displaystyle(\mathcal{L}^{-1}(\lambda+h)-\mathcal{L}^{-1}(\lambda))f= ϕ(λ+h)ϕ(λ)\displaystyle\phi(\lambda+h)-\phi(\lambda)
=\displaystyle= 1(λ)[((λ+h)(λ))]ϕ(λ+h)\displaystyle-\mathcal{L}^{-1}(\lambda)\left[(\mathcal{L}(\lambda+h)-\mathcal{L}(\lambda))\right]\phi(\lambda+h)
=\displaystyle= h1(λ)1(λ+h)f,\displaystyle-h\mathcal{L}^{-1}(\lambda)\mathcal{L}^{-1}(\lambda+h)f,

from which the desired result follows. This last expression also shows that for λ>0\lambda>0, the derivative of 1(λ)\mathcal{L}^{-1}(\lambda) with respect to this parameter is given by

λ1(λ):Hr,γk(2)Hr,γk+1(2)f1(λ)1(λ)f\begin{array}[]{c c c c}\partial_{\lambda}\mathcal{L}^{-1}(\lambda):&H^{k}_{r,\gamma}(\mathbb{R}^{2})&\longrightarrow&H^{k+1}_{r,\gamma}(\mathbb{R}^{2})\\ &f&\mapsto&-\mathcal{L}^{-1}(\lambda)\mathcal{L}^{-1}(\lambda)f\end{array}

Since the derivative λ1(λ)\partial_{\lambda}\mathcal{L}^{-1}(\lambda) is the composition of two continuous operators, it follows that it is itself continuous with respect to λ\lambda. ∎

Finally, the next proposition establishes Fredholm properties for the radial operators Δn:Mr,γ22,2(2)Lr,γ2(2)\Delta_{n}:M^{2,2}_{r,\gamma-2}(\mathbb{R}^{2})\rightarrow L^{2}_{r,\gamma}(\mathbb{R}^{2}) ,

Δn=rr+1rrn2r2,n{0}.\Delta_{n}=\partial_{rr}+\frac{1}{r}\partial_{r}-\frac{n^{2}}{r^{2}},\quad n\in\mathbb{N}\cup\{0\}.

The results follows from [19], where it is shown that the Laplace operator Δ:Mγ22,p(2)Lγp(2)\Delta:M^{2,p}_{\gamma-2}(\mathbb{R}^{2})\rightarrow L^{p}_{\gamma}(\mathbb{R}^{2}) is Fredholm, and the fact that when p=2p=2, one can decompose the space Mγ22,2(2)M^{2,2}_{\gamma-2}(\mathbb{R}^{2}) into a direct sum mn,γ22\oplus m^{2}_{n,\gamma-2} where

mn,γ22={uMγ22,2(2):u(r,θ)=vn(r)einθ,vnMr,γ22,2(2)},n.m^{2}_{n,\gamma-2}=\{u\in M^{2,2}_{\gamma-2}(\mathbb{R}^{2}):\;u(r,\theta)=v_{n}(r)\mathrm{e}^{\mathrm{i}n\theta},\quad v_{n}\in M^{2,2}_{r,\gamma-2}(\mathbb{R}^{2})\},\qquad n\in\mathbb{Z}.

Notice that because the functions in Mγ22,2(2)M^{2,2}_{\gamma-2}(\mathbb{R}^{2}) are real valued we have that vn=v¯nv_{-n}=\bar{v}_{n}. For a detailed proof of this next result, see [13, Lemma 3.1].

Proposition 2.5.

Let γ\\gamma\in\mathbb{R}\backslash\mathbb{Z}, and nn\in\mathbb{Z}. Then, the operator Δn:Mr,γ22,2(2)Lr,γ2(2)\Delta_{n}:M^{2,2}_{r,\gamma-2}(\mathbb{R}^{2})\rightarrow L^{2}_{r,\gamma}(\mathbb{R}^{2}) given by

Δnϕ=rrϕ+1rrϕn2r2ϕ\Delta_{n}\phi=\partial_{rr}\phi+\frac{1}{r}\partial_{r}\phi-\frac{n^{2}}{r^{2}}\phi

is a Fredholm operator and,

  1. (1)

    for 1|n|<γ<|n|+11-|n|<\gamma<|n|+1, the map is invertible;

  2. (2)

    for γ>|n|+1\gamma>|n|+1, the map is injective with cokernel spanned by r|n|r^{|n|};

  3. (3)

    for γ<1|n|\gamma<1-|n|, the map is surjective with kernel spanned by r|n|r^{|n|}.

On the other hand, the operator is not Fredholm for integer values of γ\gamma.

3. Intermediate Approximations to the Viscous Eikonal Equation

As mentioned in the introduction, our interest in the viscous eikonal equation

ϕ~t=Δϕ~b|ϕ~|2εg(x),x2,\tilde{\phi}_{t}=\Delta\tilde{\phi}-b|\nabla\tilde{\phi}|^{2}-\varepsilon g(x),\quad x\in\mathbb{R}^{2},

stems from its role as a model equation for the phase dynamics of target patterns and spiral waves in oscillatory systems. We are therefore interested in solutions of the form ϕ~(x,t)=ϕ(x)Ωt\tilde{\phi}(x,t)=\phi(x)-\Omega t, which then satisfy the steady state equation,

Δϕb|ϕ|2εg(x)+Ω=0x2.\Delta\phi-b|\nabla\phi|^{2}-\varepsilon g(x)+\Omega=0\qquad x\in\mathbb{R}^{2}. (4)

Because the gradient, ϕ\nabla\phi, then approximates the pattern’s wavenumber, target patterns then correspond to those ϕ\phi which in addition fulfill the boundary conditions, ϕk\nabla\phi\to k as |x||x|\to\infty. Consequently, we look for solutions to equation (4) that bifurcate from zero when ε>0\varepsilon>0, and whose gradients are bounded at infinity.

Notice that the condition on the gradient, ϕ\nabla\phi, provides enough information to derive an equation that is valid at intermediate scales. Indeed, assuming a regular perturbation for both ϕ\phi and Ω,\Omega, one obtains at order O(ε)\mathrm{O}(\varepsilon) the equation,

rrϕ1+1rrϕ1g=Ω1.\partial_{rr}\phi_{1}+\frac{1}{r}\partial_{r}\phi_{1}-g=-\Omega_{1}.

A short calculation then shows that in order to obtain solutions with bounded derivatives, the parameter Ω1\Omega_{1} must be zero. Continuing this perturbation analysis one checks that this condition must be satisfied at all orders of ε\varepsilon . In other words, the frequency, Ω\Omega, must be small beyond all orders of this parameter. Consequently, at intermediate scales the system is well approximated by the intermediate equation

Δ0ϕb(rϕ)2εgcεgf=0.\Delta_{0}\phi-b(\partial_{r}\phi)^{2}-\varepsilon g_{c}-\varepsilon g_{f}=0. (5)

Notice that we have explicitly written the inhomogeneity as the sum of two functions satisfying Hypothesis 1.1. This choice of notation will be used next in Subsection 3.1, where we construct a first order approximation for the above equation. We then use this information to prove existence of solutions to equation  (5) in Subsection 3.2.

Notation: Throughout this section, and in the rest of the paper, we use γϵ\gamma_{\epsilon} to denote the Euler Mascheroni constant, and the symbols χ,χMC(2)\chi,\chi_{M}\in C^{\infty}(\mathbb{R}^{2}) to denote smooth radial cut-off functions satisfying

χ(x)={0if|x|<11if|x|>2,χM(x)={0if|x|<11if2<|x|<M0if2M<|x|,\chi(x)=\left\{\begin{array}[]{c c c}0&\mbox{if}&|x|<1\\ 1&\mbox{if}&|x|>2\end{array},\right.\qquad\chi_{M}(x)=\left\{\begin{array}[]{c c c}0&\mbox{if}&|x|<1\\ 1&\mbox{if}&2<|x|<M\\ 0&\mbox{if}&2M<|x|\\ \end{array},\right.

where MM is a positive constant. Notice in particular, that χM\chi_{M} has compact support.

3.1. First Order Approximation

We construct a first order approximation, ϕ\phi, which is the sum of two functions, ϕ0\phi_{0} and ϕ1\phi_{1}. We take

ϕ0=1bχMlog(1+alogr+εK(r)),\phi_{0}=-\frac{1}{b}\chi_{M}\log(1+a\log r+\varepsilon K(r)), (6)

a choice that is motivated by the Hopf-Cole transform, ϕ=1blogΨ\phi=-\frac{1}{b}\log\Psi, which turns the eikonal equation (5) into the steady state Schrödinger equation with potential εg\varepsilon g. The value of the constant MM appearing in the definition of the cut-off function χM\chi_{M} is taken so that the expression log(1+alogr+εK(r)),\log(1+a\log r+\varepsilon K(r)), always remains bounded. The constant aa is a parameter that is determined when constructing the second part to the approximation, while the function K(r)K(r) satisfies

Δ0K+bgf=0.\Delta_{0}K+bg_{f}=0. (7)

The fact that we can solve this last equation follows from our assumptions on the inhomogeneity. Recall that gfHσk(2)g_{f}\in H^{k}_{\sigma}(\mathbb{R}^{2}) for 0<σ<10<\sigma<1. Proposition 2.5 then shows that the radial Laplacian, Δ0\Delta_{0}, is a surjective operator with a one dimensional kernel spanned by {1}\{1\}. We can therefore use Lyapunov-Schmidt reduction to solve this equation and find a family of solutions

K(r)=Kp(r)+c,c.K(r)=K_{p}(r)+c,\qquad c\in\mathbb{R}.

Since cc is arbitrary, without loss of generality we pick c=0c=0. We also find that the solution, KK, belongs to the space Mr,σ22,2(2)M^{2,2}_{r,\sigma-2}(\mathbb{R}^{2}). In fact, one can check that KK has more regularity and is in the space

Rσk={uMr,σ22,2(2):D2uHσk(2)}.R_{\sigma}^{k}=\{u\in M^{2,2}_{r,\sigma-2}(\mathbb{R}^{2}):D^{2}u\in H^{k}_{\sigma}(\mathbb{R}^{2})\}. (8)
Remark 3.1.

Notice that the function ϕ0\phi_{0} is as regular as the function KK, and that as a result the derivative rϕ0\partial_{r}\phi_{0} is in the space Hσk+1(2)H^{k+1}_{\sigma}(\mathbb{R}^{2}). In addition, this function is bounded and has compact support.

Remark 3.2.

Because gfg_{f} is in Hσk(2)H^{k}_{\sigma}(\mathbb{R}^{2}) with σ(0,1)\sigma\in(0,1) and k2k\geq 2, we then have the following decay properties for the solution to equation (7):

  • if gfg_{f} decays like 1/rm1/r^{m} in the far field, with 1<m<2,1<m<2, then KO(r2m)K\sim\mathrm{O}(r^{2-m}) at infinity, while

  • if gfg_{f} decays like 1/r21/r^{2} in the far field, then KO((logr)2)K\sim\mathrm{O}((\log r)^{2}) at infinity.

Next, we define the second function, ϕ1\phi_{1}, as the solution to the equation

Δ0ϕ1abΔ0(χlogr)εgc=0.\Delta_{0}\phi_{1}-\frac{a}{b}\Delta_{0}(\chi\log r)-\varepsilon g_{c}=0.

Here, the constant aa is the same as the one appearing in the definition of ϕ0\phi_{0}, and the function gcg_{c} is in the space Hγk(2)H^{k}_{\gamma}(\mathbb{R}^{2}) with 1<γ1<\gamma, by assumption.

To justify the existence of ϕ1\phi_{1}, we use again Proposition 2.5 which shows that for values of γ>1\gamma>1, the operator Δ0:Mr,γ22,2(2)Lr,γ2(2)\Delta_{0}:M^{2,2}_{r,\gamma-2}(\mathbb{R}^{2})\longrightarrow L^{2}_{r,\gamma}(\mathbb{R}^{2}) is Fredholm with index -1, and cokernel spanned by {1}\{1\}. Because the projection of Δ0(χlogr)\Delta_{0}(\chi\log r) onto the cokernel is non-trivial, i.e.

2Δ0(χlogr)𝑑x=2π,\int_{\mathbb{R}^{2}}\Delta_{0}(\chi\log r)\;dx=2\pi,

the Bordering Lemma stated at the end of this subsection then shows that the operator

Mr,γ22,2(2)×Lr,γ2(2)(ϕ,a)Δ0ϕ1abΔ0(χlogr)\begin{array}[]{c c c}M^{2,2}_{r,\gamma-2}(\mathbb{R}^{2})\times\mathbb{R}&\longrightarrow&L^{2}_{r,\gamma}(\mathbb{R}^{2})\\[8.61108pt] (\phi,a)&\longmapsto&\Delta_{0}\phi_{1}-\dfrac{a}{b}\Delta_{0}(\chi\log r)\end{array}

is invertible. Therefore, the equation for ϕ1\phi_{1} is indeed solvable. In addition, projecting onto the constant functions we also find that

a=εb0gc(r)r𝑑r.a=-\varepsilon b\int_{0}^{\infty}g_{c}(r)\;r\;dr. (9)

Finally, since gcHγk(2)g_{c}\in H^{k}_{\gamma}(\mathbb{R}^{2}), it follows that our solution ϕ1\phi_{1} is in the space RγkR_{\gamma}^{k}, defined as in (8).

Lemma 3.3.

[Bordering Lemma] Let XX and YY be Banach spaces, and consider the operator

S=[ABCD]:X×pY×q,S=\begin{bmatrix}A&B\\ C&D\end{bmatrix}:X\times\mathbb{R}^{p}\longrightarrow Y\times\mathbb{R}^{q},

with bounded linear operators A:XYA:X\longrightarrow Y, B:pYB:\mathbb{R}^{p}\longrightarrow Y, C:XqC:X\longrightarrow\mathbb{R}^{q}, D:pqD:\mathbb{R}^{p}\longrightarrow\mathbb{R}^{q}. If AA is Fredholm of index ii, then SS is Fredholm of index i+pqi+p-q.

Proof.

One can write SS as the sum of a block diagonal operator with the indicated index, i+pqi+p-q, and a compact operator consisting of the off-diagonal elements. Since compact perturbations do not alter the index of a Fredholm operator, the result then follows. ∎

3.2. Existence of Solutions to Intermediate Approximation

Using the first order approximation, ϕ0+ϕ1\phi_{0}+\phi_{1}, defined in the previous subsection we now prove the existence of solutions to equation (5) using the implicit function theorem.

Inserting the ansatz

ϕ=ϕ0+ϕ1+ϕ2\phi=\phi_{0}+\phi_{1}+\phi_{2}

into equation (5), one obtains the following expression for ϕ2\phi_{2},

Δ0ϕ2b(2rϕ0(rϕ1+rϕ2)+(rϕ1+rϕ2)2)+G1=0,\Delta_{0}\phi_{2}-b\Big{(}2\partial_{r}\phi_{0}(\partial_{r}\phi_{1}+\partial_{r}\phi_{2})+(\partial_{r}\phi_{1}+\partial_{r}\phi_{2})^{2}\Big{)}+G_{1}=0, (10)

where the term G1G_{1} is given by

G1=Δ0ϕ0b(rϕ0)2εgf+abΔ0(χlogr)G_{1}=\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}-\varepsilon g_{f}+\frac{a}{b}\Delta_{0}(\chi\log r)

To continue the analysis, we let ψ=rϕ2\psi=\partial_{r}\phi_{2} in equation (10), and add and subtract the term λψ\lambda\psi. We assume that the parameter λ\lambda is sufficiently small, positive, and fixed. The result is,

rψ+1rψ+λψb(2rϕ0(rϕ1+ψ)+(rϕ1+ψ)2)+G1λψ=0.\partial_{r}\psi+\frac{1}{r}\psi+\lambda\psi-b\Big{(}2\partial_{r}\phi_{0}(\partial_{r}\phi_{1}+\psi)+(\partial_{r}\phi_{1}+\psi)^{2}\Big{)}+G_{1}-\lambda\psi=0.

Letting λ=r+1r+λ\mathcal{L}_{\lambda}=\partial_{r}+\frac{1}{r}+\lambda, we may precondition this last equation with λ1\mathcal{L}^{-1}_{\lambda} and write

ψ+λ1[b(2rϕ0(rϕ1+ψ)+(rϕ1+ψ)2)+G1λψ]=0.\psi+\mathcal{L}_{\lambda}^{-1}\left[-b\Big{(}2\partial_{r}\phi_{0}(\partial_{r}\phi_{1}+\psi)+(\partial_{r}\phi_{1}+\psi)^{2}\Big{)}+G_{1}-\lambda\psi\right]=0. (11)

Because the above expression is equivalent to the intermediate equation (5), if we find a solution ψ\psi to (11), we immediately obtain a corresponding solution to (5) of the form

ϕ(r;ε)=ϕ0(r;ε)+ϕ1(r;ε)+ϕ2(r;ε)+εc,\phi(r;\varepsilon)=\phi_{0}(r;\varepsilon)+\phi_{1}(r;\varepsilon)+\phi_{2}(r;\varepsilon)+\varepsilon c,

where rϕ2=ψ\partial_{r}\phi_{2}=\psi and cc is a constant of integration.

To use the implicit function theorem, we view the left hand side of (11) as an operator F:Hδk(2)×+Hδk(2)F:H^{k}_{\delta}(\mathbb{R}^{2})\times\mathbb{R}_{+}\longrightarrow H^{k}_{\delta}(\mathbb{R}^{2}) for some appropriate δ\delta\in\mathbb{R}, and show that it is well defined, smooth with respect to ε\varepsilon, and that its Fréchet derivative DψF(0;0):Hδk(2)Hδk(2)D_{\psi}F(0;0):H^{k}_{\delta}(\mathbb{R}^{2})\longrightarrow H^{k}_{\delta}(\mathbb{R}^{2}) is invertible. The result is the following theorem.

Theorem 2.

Let k2k\geq 2, σ(0,1)\sigma\in(0,1) and consider functions gHr,σk(2)g\in H^{k}_{r,\sigma}(\mathbb{R}^{2}) satisfying Hypothesis 1.1. Then, the intermediate equation

Δ0ϕb(rϕ)2εg=0,\Delta_{0}\phi-b(\partial_{r}\phi)^{2}-\varepsilon g=0,

has a family of solutions

ϕ(r;ε)=1bχMlog(1+alogr+εK)+ϕ1(r;ε)+ϕ2(r;ε)+εc,c,\phi(r;\varepsilon)=-\frac{1}{b}\chi_{M}\log\Big{(}1+a\log r+\varepsilon K\Big{)}+\phi_{1}(r;\varepsilon)+\phi_{2}(r;\varepsilon)+\varepsilon c,\qquad c\in\mathbb{R},

that bifurcates from zero at ε=0\varepsilon=0 and is C1C^{1} in ε[0,)\varepsilon\in[0,\infty). Moreover, letting γ(1,)\gamma\in(1,\infty), and σ(0,1)\sigma\in(0,1) be defined as in Hypothesis 1.1, the family of solutions satisfies:

  • ϕ1{uMr,γ22,2(2):D2uHγk(2)}\phi_{1}\in\{u\in M^{2,2}_{r,\gamma-2}(\mathbb{R}^{2}):D^{2}u\in H^{k}_{\gamma}(\mathbb{R}^{2})\},

  • rϕ2Hδk(2)\partial_{r}\phi_{2}\in H^{k}_{\delta}(\mathbb{R}^{2}), where δ=min(γ1,σ)\delta=\min(\gamma-1,\sigma).

  • Δ0K=gf\Delta_{0}K=g_{f}, with KRγkK\in R^{k}_{\gamma}, and

  • a=εb0gc(r)r𝑑ra=-\varepsilon b\int_{0}^{\infty}g_{c}(r)\;r\;dr.

Proof.

As already mentioned, the result follows from finding solutions to equation (11) using the implicit function theorem. We therefore consider the left hand side of this equation as an operator F:Hδk(2)×+Hδk(2)F:H^{k}_{\delta}(\mathbb{R}^{2})\times\mathbb{R}_{+}\longrightarrow H^{k}_{\delta}(\mathbb{R}^{2}), with δ=min(γ1,σ)>0\delta=\min(\gamma-1,\sigma)>0.

Since the operator’s dependence on ε\varepsilon comes from the three functions rϕ0,rϕ1,\partial_{r}\phi_{0},\partial_{r}\phi_{1}, and G1G_{1}, and since these functions are all smooth with respect to ε\varepsilon on the interval [0,)[0,\infty), then the same result holds for the operator FF.

To show that the Fréchet derivative DψF(0;0)=Idλλ1D_{\psi}F(0;0)=Id-\lambda\mathcal{L}_{\lambda}^{-1} is invertible, we recall the results from Section 2. In particular, Lemma 2.3 shows that λ1:Hδk(2)Hδk+1(2)\mathcal{L}_{\lambda}^{-1}:H^{k}_{\delta}(\mathbb{R}^{2})\longrightarrow H^{k+1}_{\delta}(\mathbb{R}^{2}) is bounded. Since the embedding Hδk+1(2)Hδk(2)H^{k+1}_{\delta}(\mathbb{R}^{2})\subset H^{k}_{\delta}(\mathbb{R}^{2}) is continuous, it then follows that DψF(0;0):Hδk(2)Hδk(2)D_{\psi}F(0;0):H^{k}_{\delta}(\mathbb{R}^{2})\longrightarrow H^{k}_{\delta}(\mathbb{R}^{2}) is a small perturbation of the identity operator, and is thus invertible for a sufficiently small λ\lambda.

To complete the proof we need to show that the operator FF is well defined. Taking into account the results of Lemma 2.3, this is equivalent to showing that the expression

N(ψ,ε)=b(2rϕ0(rϕ1+ψ)+(rϕ1+ψ)2)+G1λψ,N(\psi,\varepsilon)=-b\Big{(}2\partial_{r}\phi_{0}(\partial_{r}\phi_{1}+\psi)+(\partial_{r}\phi_{1}+\psi)^{2}\Big{)}+G_{1}-\lambda\psi,

defines a bounded operator N:Hδk(2)×+Hδk1(2)N:H^{k}_{\delta}(\mathbb{R}^{2})\times\mathbb{R}_{+}\longrightarrow H^{k-1}_{\delta}(\mathbb{R}^{2}).

We start with the term rϕ0(rϕ1+ψ)\partial_{r}\phi_{0}(\partial_{r}\phi_{1}+\psi). From the definition of ϕ1\phi_{1} we know that this is a function in RγkR^{k}_{\gamma} with γ>1\gamma>1. In particular,

rϕ1{uMr,γ11,2(2):DuHγk(2)}Hγ1k+1(2).\partial_{r}\phi_{1}\in\{u\in M^{1,2}_{r,\gamma-1}(\mathbb{R}^{2}):Du\in H^{k}_{\gamma}(\mathbb{R}^{2})\}\subset H^{k+1}_{\gamma-1}(\mathbb{R}^{2}).

Because ψHδk(2)\psi\in H^{k}_{\delta}(\mathbb{R}^{2}) and δ=min(γ1,σ)>0\delta=\min(\gamma-1,\sigma)>0 , it then follows that the sum (rϕ1+ψ)(\partial_{r}\phi_{1}+\psi) is also in this space. Since rϕ0\partial_{r}\phi_{0} has compact support and k+1k+1 bounded derivatives (see Remark 3.1), then the product rϕ0(rϕ1+ψ)\partial_{r}\phi_{0}(\partial_{r}\phi_{1}+\psi) is also well defined in Hδk(2)H^{k}_{\delta}(\mathbb{R}^{2}).

Next, since (rϕ1+ψ)(\partial_{r}\phi_{1}+\psi) is in Hδk(R2)H^{k}_{\delta}(R^{2}), with δ>0\delta>0 and k2k\geq 2, Lemma 3.4 below shows that (rϕ1+ψ)2(\partial_{r}\phi_{1}+\psi)^{2} is in Hδk1(2)H^{k-1}_{\delta}(\mathbb{R}^{2}). Finally, Lemma 3.5 at the end of this section shows that G1Hσk(2)G_{1}\in H^{k}_{\sigma}(\mathbb{R}^{2}), and because δ=min(γ1,σ)>0\delta=\min(\gamma-1,\sigma)>0, this term is also well defined.

Since the operator FF satisfies the assumptions of the implicit function theorem we obtain a family of solutions ψ(r;ε)\psi(r;\varepsilon) that bifurcates from zero and is smooth with respect to ε\varepsilon. Because ψ=rϕ2\psi=\partial_{r}\phi_{2}, we arrive at the family

ϕ(r;ε)=ϕ0(r;ε)+ϕ1(r;ε)+ϕ2(r;ε)+εc,c.\phi(r;\varepsilon)=\phi_{0}(r;\varepsilon)+\phi_{1}(r;\varepsilon)+\phi_{2}(r;\varepsilon)+\varepsilon c,\quad c\in\mathbb{R}.

This finishes the proof of the Theorem.

Lemma 3.4.

Let ψHγk(2)\psi\in H^{k}_{\gamma}(\mathbb{R}^{2}) with γ>0\gamma>0 and k2k\geq 2. Then, ψ2Hγk1(2)\psi^{2}\in H^{k-1}_{\gamma}(\mathbb{R}^{2}).

Proof.

To simplify the analysis we let DjD^{j} denote any jj-th order derivative, and we only prove that Dk1(ψ2)D^{k-1}(\psi^{2}) is in Lr,γ2(2)L^{2}_{r,\gamma}(\mathbb{R}^{2}), since a similar analysis shows that lower derivatives are in this same space. Because k2k\geq 2 and γ>0\gamma>0, it follows by Sobolev embeddings that DjψHr,γ2(2)H2(2)CB(2)D^{j}\psi\in H^{2}_{r,\gamma}(\mathbb{R}^{2})\subset H^{2}(\mathbb{R}^{2})\subset C_{B}(\mathbb{R}^{2}) for 0jk20\leq j\leq k-2. Then, writing

Dk1(ψ2)=j=0k1(k1j)Dk1jψDjψD^{k-1}(\psi^{2})=\sum_{j=0}^{k-1}{k-1\choose j}D^{k-1-j}\psi D^{j}\psi

we see that this derivate can be written as a product of a bounded function and a function that is in Lr,γ2(2)L^{2}_{r,\gamma}(\mathbb{R}^{2}). Hence ψ2Hr,γk1(2)\psi^{2}\in H^{k-1}_{r,\gamma}(\mathbb{R}^{2}). ∎

The next Lemma shows that G1G_{1} is in Hσk(2)H^{k}_{\sigma}(\mathbb{R}^{2}) with σ(0,1)\sigma\in(0,1).

Lemma 3.5.

Let k2k\geq 2, σ(0,1)\sigma\in(0,1) and take gfHσk(2)g_{f}\in H^{k}_{\sigma}(\mathbb{R}^{2}). Consider the function ϕ0\phi_{0} constructed from gfg_{f} and described above in (6). Then the expression

G1=Δ0ϕ0b(rϕ0)2εgf+abΔ0(χlogr)G_{1}=\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}-\varepsilon g_{f}+\frac{a}{b}\Delta_{0}(\chi\log r)

is also in Hσk(2)H^{k}_{\sigma}(\mathbb{R}^{2}).

Proof.

Using the notation ϕ0=χMϕ~0\phi_{0}=\chi_{M}\tilde{\phi}_{0}, we first expand G1G_{1}

G1=\displaystyle G_{1}= Δ0ϕ0b(rϕ0)2εgf+abΔ0(χlogr)\displaystyle\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}-\varepsilon g_{f}+\frac{a}{b}\Delta_{0}(\chi\log r)
G1=\displaystyle G_{1}= (ϕ~0Δ0χM+2χMϕ~0b(χMϕ~0)22bχMχMϕ~0rϕ~0+b(rϕ~0)2(χMχM2))\displaystyle\left(\tilde{\phi}_{0}\Delta_{0}\chi_{M}+2\chi_{M}^{\prime}\tilde{\phi}_{0}-b(\chi_{M}^{\prime}\tilde{\phi}_{0})^{2}-2b\chi_{M}^{\prime}\chi_{M}\tilde{\phi}_{0}\partial_{r}\tilde{\phi}_{0}+b(\partial_{r}\tilde{\phi}_{0})^{2}(\chi_{M}-\chi_{M}^{2})\right)
1bχM[Δ0(alogr+εK)1+alogr+εK]εgf+abΔ0(χlogr)\displaystyle-\frac{1}{b}\chi_{M}\left[\frac{\Delta_{0}(a\log r+\varepsilon K)}{1+a\log r+\varepsilon K}\right]-\varepsilon g_{f}+\frac{a}{b}\Delta_{0}(\chi\log r)
G1=\displaystyle G_{1}= (ϕ~0Δ0χM+2χMϕ~0b(χMϕ~0)22bχMχMϕ~0rϕ~0+b(rϕ~0)2(χMχM2))\displaystyle\left(\tilde{\phi}_{0}\Delta_{0}\chi_{M}+2\chi_{M}^{\prime}\tilde{\phi}_{0}-b(\chi_{M}^{\prime}\tilde{\phi}_{0})^{2}-2b\chi_{M}^{\prime}\chi_{M}\tilde{\phi}_{0}\partial_{r}\tilde{\phi}_{0}+b(\partial_{r}\tilde{\phi}_{0})^{2}(\chi_{M}-\chi_{M}^{2})\right)
1bχM[Δ0alogr1+alogr+εK]1bχM[εΔ0K1+alogr+εK]εgf+abΔ0(χlogr).\displaystyle-\frac{1}{b}\chi_{M}\left[\frac{\Delta_{0}a\log r}{1+a\log r+\varepsilon K}\right]-\frac{1}{b}\chi_{M}\left[\frac{\varepsilon\Delta_{0}K}{1+a\log r+\varepsilon K}\right]-\varepsilon g_{f}+\frac{a}{b}\Delta_{0}(\chi\log r).

Because the logr\log r is a fundamental solution of the Laplacian and since χM\chi_{M} is zero near the origin, then the term

χM[Δ0alogr1+alogr+εK]=0.\chi_{M}\left[\frac{\Delta_{0}a\log r}{1+a\log r+\varepsilon K}\right]=0.

Similarly, because the function KK is a solution to Δ0K+bgf=0\Delta_{0}K+bg_{f}=0, then we may write

1bχM[εΔ0K1+alogr+εK]εgf=\displaystyle-\frac{1}{b}\chi_{M}\left[\frac{\varepsilon\Delta_{0}K}{1+a\log r+\varepsilon K}\right]-\varepsilon g_{f}= εgf(1χM)εgfχM[(alogr+εK)1+alogr+εK].\displaystyle-\varepsilon g_{f}(1-\chi_{M})-\varepsilon g_{f}\chi_{M}\left[\frac{(a\log r+\varepsilon K)}{1+a\log r+\varepsilon K}\right].

Therefore,

G1=\displaystyle G_{1}= (ϕ~0Δ0χM+2χMϕ~0b(χMϕ~0)22bχMχMϕ~0rϕ~0+b(rϕ~0)2(χMχM2))\displaystyle\left(\tilde{\phi}_{0}\Delta_{0}\chi_{M}+2\chi_{M}^{\prime}\tilde{\phi}_{0}-b(\chi_{M}^{\prime}\tilde{\phi}_{0})^{2}-2b\chi_{M}^{\prime}\chi_{M}\tilde{\phi}_{0}\partial_{r}\tilde{\phi}_{0}+b(\partial_{r}\tilde{\phi}_{0})^{2}(\chi_{M}-\chi_{M}^{2})\right)
εgf(1χM)εgfχM[(alogr+εK)1+alogr+εK]+abΔ0(χlogr)\displaystyle-\varepsilon g_{f}(1-\chi_{M})-\varepsilon g_{f}\chi_{M}\left[\frac{(a\log r+\varepsilon K)}{1+a\log r+\varepsilon K}\right]+\frac{a}{b}\Delta_{0}(\chi\log r)

From the definition of χM\chi_{M} it is clear that all terms involving a derivative of this function are localized and have compact support. Because the value of MM in the definition of χM\chi_{M} was chosen to vanish whenever the expression 1+alogr+εK1+a\log r+\varepsilon K is 0\leq 0, we see that the term in brackets is also bounded and with compact support. In addition, this term is as regular as the function KRσkK\in R^{k}_{\sigma}, and is therefore in Hk+2(2)H^{k+2}(\mathbb{R}^{2}). On the other hand, the function (1χM)εgf(1-\chi_{M})\varepsilon g_{f} behaves like gfg_{f} at infinity and as a result it is in the same space as the inhomogeneity. Finally, because Δ0logr=0\Delta_{0}\log r=0 on 2{0}\mathbb{R}^{2}\setminus\{0\}, the function Δ0(χlogr)\Delta_{0}(\chi\log r) is localized and smooth. Taking all this into account, we may conclude that G1G_{1} is in the space Hσk(2)H^{k}_{\sigma}(\mathbb{R}^{2}).

4. Far Field Approximation to the Viscous Eikonal Equation

In this section we consider again the full equation

Δ0ϕb(rϕ)2εg(r)+Ω=0r=|x|[0,),\Delta_{0}\phi-b(\partial_{r}\phi)^{2}-\varepsilon g(r)+\Omega=0\qquad r=|x|\in[0,\infty), (12)

but assume that the value of Ω\Omega is fixed and different from zero. As in Section 3, we first find an appropriate expression for the far field behavior of the solution and a first order approximation for this new equation. We then use this result to prove existence of solutions using the implicit function theorem.

Because the inhomogeneity is algebraically decaying, for large values of rr the relevant terms in the equation are

Δ0ϕb(rϕ)2+Ω=0\Delta_{0}\phi-b(\partial_{r}\phi)^{2}+\Omega=0

To find a first order approximation, we can again use the Hopf-Cole transform, ϕ(r)=(1/b)log(K)\phi(r)=-(1/b)\log(K), rewriting the equation as

rrK+1rrKΛ2K=0Λ2=bΩ.\partial_{rr}K+\frac{1}{r}\partial_{r}K-\Lambda^{2}K=0\qquad\Lambda^{2}=b\Omega.

Notice that this is either a Bessel, or the Modified Bessel equation, depending on the sign of bΩb\Omega. Because we are interested in solutions, ϕ(r)\phi(r), that are real, we pick bΩ>0b\Omega>0 so that the solution to this last equation is K=K0K=K_{0}, the Modified Bessel function of the second kind. In particular, because K0(z)O(ez)K_{0}(z)\sim\mathrm{O}(e^{-z}) as zz\to\infty ( see Table 1 below), this implies that rϕ(r)\partial_{r}\phi(r) is bounded in the far field, as desired.

We therefore consider the ansatz

ϕ(r)=ϕ0(r)+ϕ1(r)\phi(r)=\phi_{0}(r)+\phi_{1}(r)

where ϕ0\phi_{0} is given by

ϕ0(r)=1bχ(Λr)log(K0(Λr)),Λ2=bΩ>0.\phi_{0}(r)=-\frac{1}{b}\chi(\Lambda r)\log(K_{0}(\Lambda r)),\qquad\Lambda^{2}=b\Omega>0.

Here, again χ\chi represents a cut-off function that removes the singular behavior of the log function near the origin. Inserting this expression into equation (12) gives

Δ0ϕ12brϕ0rϕ1b(rϕ1)2+(Δ0ϕ0b(rϕ0)2+Ω)εg=0.\Delta_{0}\phi_{1}-2b\partial_{r}\phi_{0}\partial_{r}\phi_{1}-b(\partial_{r}\phi_{1})^{2}+(\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}+\Omega)-\varepsilon g=0.

Since the terms appearing in the parenthesis represent a localized function with compact support, they do not contribute to the behavior of the solution for large values of rr. Thus, the far field behavior of the solution is determined by

Δ0ϕ12brϕ0rϕ1b(rϕ1)2εg=0.\Delta_{0}\phi_{1}-2b\partial_{r}\phi_{0}\partial_{r}\phi_{1}-b(\partial_{r}\phi_{1})^{2}-\varepsilon g=0. (13)

Letting ψ=rϕ1\psi=\partial_{r}\phi_{1} and adding and subtracting the term 2Λψ2\Lambda\psi gives us

rψ+1rψ+2Λψ+[2brϕ0ψbψ2εg2Λψ]=0.\partial_{r}\psi+\frac{1}{r}\psi+2\Lambda\psi+\left[-2b\partial_{r}\phi_{0}\psi-b\psi^{2}-\varepsilon g-2\Lambda\psi\right]=0.

We can then precondition this equation by 2Λ1\mathcal{L}_{2\Lambda}^{-1}, since by Lemma 2.3 we know that this operator is bounded for all values of Λ>0\Lambda>0, if its domain is Hr,σk(2)H^{k}_{r,\sigma}(\mathbb{R}^{2}). Thus, the equation can be written as

F(ψ;ε)=Id+2Λ1[2brϕ0ψbψ2εg2Λψ]=0.F(\psi;\varepsilon)=\mathrm{Id}+\mathcal{L}_{2\Lambda}^{-1}\Big{[}-2b\partial_{r}\phi_{0}\psi-b\psi^{2}-\varepsilon g-2\Lambda\psi\Big{]}=0. (14)

In what follows we will show that the operator F:Hr,σk(2)×Hr,σk(2)F:H^{k}_{r,\sigma}(\mathbb{R}^{2})\times\mathbb{R}\rightarrow H^{k}_{r,\sigma}(\mathbb{R}^{2}) satisfies the conditions of the implicit function theorem and prove the following theorem.

Theorem 3.

Take gHr,σkg\in H^{k}_{r,\sigma} with k2k\geq 2 and let σ(0,1)\sigma\in(0,1). Then there exist a positive constant Λ0\Lambda_{0} such that for any fixed Λ(0,Λ0)\Lambda\in(0,\Lambda_{0}), there is an ε0>0\varepsilon_{0}>0, and C1C^{1} family of solutions, ϕ=ϕ(r;ε)\phi=\phi(r;\varepsilon), to equation (13) that bifurcates from zero at ε=0\varepsilon=0 and is valid for ε(ε0,ε0)\varepsilon\in(-\varepsilon_{0},\varepsilon_{0}). Moreover, this family has the form

ϕ(r;ε)=1bχ(Λr)log(K0(Λr))+ϕ1(r;ε)+εc\phi(r;\varepsilon)=-\frac{1}{b}\chi(\Lambda r)\log(K_{0}(\Lambda r))+\phi_{1}(r;\varepsilon)+\varepsilon c

where

  1. i)

    K0(z)K_{0}(z) represents the zeroth-order Modified Bessel function of the second kind,

  2. ii)

    rϕ1Hr,σk+1(2)\partial_{r}\phi_{1}\in H^{k+1}_{r,\sigma}(\mathbb{R}^{2}),

  3. iii)

    and cc\in\mathbb{R} is an arbitrary constant.

Proof.

Because finding solutions to equation (13) is equivalent to finding the zeros of the operator F:Hr,σk(2)×Hr,σk(2)F:H^{k}_{r,\sigma}(\mathbb{R}^{2})\times\mathbb{R}\rightarrow H^{k}_{r,\sigma}(\mathbb{R}^{2}) defined in (14), we check that FF satisfies the assumptions of the implicit function theorem.

It is clear that F(0;0)=0F(0;0)=0 and that this operator is smooth with respect to the parameter ε\varepsilon. To check that the Fréchet derivative, DψF(0;0):Hr,σk(2)Hr,σk(2)D_{\psi}F(0;0):H^{k}_{r,\sigma}(\mathbb{R}^{2})\rightarrow H^{k}_{r,\sigma}(\mathbb{R}^{2}), given by

DψF(0;0)=Id+2Λ1[2brϕ02Λ],D_{\psi}F(0;0)=Id+\mathcal{L}_{2\Lambda}^{-1}[-2b\partial_{r}\phi_{0}-2\Lambda],

is invertible, notice that the term in the brackets can be written as the product of a bounded function times the constant 2Λ2\Lambda. Indeed, this can be checked by expanding this term,

2brϕ02Λ\displaystyle-2b\partial_{r}\phi_{0}-2\Lambda =2br[1bχ(Λr)log[K0(Λr)]]2Λ\displaystyle=-2b\partial_{r}[\;-\frac{1}{b}\chi(\Lambda r)\log[K_{0}(\Lambda r)]]-2\Lambda
=2Λ[χ(Λr)log[K0(Λr)]+χ(Λr)K0(Λr)K0(Λr)1]\displaystyle=2\Lambda\Big{[}\chi^{\prime}(\Lambda r)\log[K_{0}(\Lambda r)]+\chi(\Lambda r)\frac{K_{0}^{\prime}(\Lambda r)}{K_{0}(\Lambda r)}-1\Big{]}

and using the fact that the ratio K0(z)K0(z)=1\frac{K^{\prime}_{0}(z)}{K_{0}(z)}=-1 as rr\to\infty, and that χ\chi^{\prime} has compact support. Since the operator 2Λ1:Hr,σk(2)Hr,σk+1(2)\mathcal{L}_{2\Lambda}^{-1}:H^{k}_{r,\sigma}(\mathbb{R}^{2})\longrightarrow H^{k+1}_{r,\sigma}(\mathbb{R}^{2}) is bounded, it follows that there is a small number Λ0>0\Lambda_{0}>0 such that if Λ(0,Λ0)\Lambda\in(0,\Lambda_{0}), the derivative DψF(0;0)D_{\psi}F(0;0) is a small perturbation of the identity and is therefore invertible.

We are left with showing that the operator FF is well defined. Taking into account again that the map 2Λ1:Hr,σk1(2)Hr,σk(2)\mathcal{L}_{2\Lambda}^{-1}:H^{k-1}_{r,\sigma}(\mathbb{R}^{2})\longrightarrow H^{k}_{r,\sigma}(\mathbb{R}^{2}) is bounded, this is equivalent to showing that the terms

2brϕ0ψbψ2εg2Λψ-2b\partial_{r}\phi_{0}\psi-b\psi^{2}-\varepsilon g-2\Lambda\psi

define a bounded operator N:Hr,σk(2)×Hr,σk1(2)N:H^{k}_{r,\sigma}(\mathbb{R}^{2})\times\mathbb{R}\rightarrow H^{k-1}_{r,\sigma}(\mathbb{R}^{2}).

First, notice that by assumption, the impurity gg is in the desired space. As for the elements involving the variable ψHr,σk(2)\psi\in H^{k}_{r,\sigma}(\mathbb{R}^{2}), because the derivative rϕ0\partial_{r}\phi_{0} is a bounded function, we can easily check that they are both in the space Hr,σk1(2)H^{k-1}_{r,\sigma}(\mathbb{R}^{2}). Finally, since σ>0\sigma>0, Lemma 3.4 shows that the product ψ2\psi^{2} is in Hr,σk1(2)H^{k-1}_{r,\sigma}(\mathbb{R}^{2}).

This proves that the operator FF satisfies the conditions of the implicit function theorem and proves the existence of a family of solutions solving F(ψ;ε)=0F(\psi;\varepsilon)=0 . Going back to the definition of ψ=rϕ1\psi=\partial_{r}\phi_{1}, we see that the above result also gives us a family of solutions ϕ(r;ε)=ϕ0(r)+ϕ1(r;ε)+εc\phi(r;\varepsilon)=\phi_{0}(r)+\phi_{1}(r;\varepsilon)+\varepsilon c solving the far field equation (12), where ϕ1Hr,σk+1(2)\phi_{1}\in H^{k+1}_{r,\sigma}(\mathbb{R}^{2}) and cc is for now an arbitrary constant which is the result of integrating ψ\psi. This proves the result of the theorem. ∎

5. Existence of Target Patterns

5.1. Matching

To determine an expression for the eigenvalue Ω\Omega, we must match the intermediate and far field approximations of the wavenumber, rϕ\partial_{r}\phi. For convenience we recall their expressions,

ϕfar(r;ε,Λ)=\displaystyle\phi_{far}(r;\varepsilon,\Lambda)= 1bχ(Λr)log(K0(Λr))+ϕ1(r;ε)+εc\displaystyle-\frac{1}{b}\chi(\Lambda r)\log(K_{0}(\Lambda r))+\phi_{1}(r;\varepsilon)+\varepsilon c
ϕint(r;ε)=\displaystyle\phi_{int}(r;\varepsilon)= 1bχMlog(1+alogr+εK(r))+ϕ¯1(r;ε)+ϕ¯2(r;ε)+εc,\displaystyle-\frac{1}{b}\chi_{M}\log\Big{(}1+a\log r+\varepsilon K(r)\Big{)}+\bar{\phi}_{1}(r;\varepsilon)+\bar{\phi}_{2}(r;\varepsilon)+\varepsilon c,

As before, K0K_{0} denotes the Modified Bessel function of the first kind, while the function KK satisfies

Δ0K+bgf=0.\Delta_{0}K+bg_{f}=0.

Notice that the remaining terms, ϕ1,ϕ¯1,\phi_{1},\bar{\phi}_{1}, and ϕ¯2\bar{\phi}_{2}, all have derivatives that decay algebraically at infinity. In particular,

  1. 1.

    The function ϕ¯1\bar{\phi}_{1} defined in Subsection 3.1 is in the space RγkMr,γ22,2R^{k}_{\gamma}\subset M^{2,2}_{r,\gamma-2}, with γ>1\gamma>1. From Lemma 2.1 it follows that |ϕ¯1|<|x|γ+1|\bar{\phi}_{1}|<|x|^{-\gamma+1}. In particular, if the inhomogeneity gcO(r(d+2))g_{c}\sim\mathrm{O}(r^{-(d+2)}), with d>0d>0, we have that rϕ¯1O(r(d+1))\partial_{r}\bar{\phi}_{1}\sim\mathrm{O}(r^{-(d+1)}).

  2. 2.

    From Theorems 2 and 3 we know that the functions rϕ1\partial_{r}\phi_{1} and rϕ¯2\partial_{r}\bar{\phi}_{2} are in the space Hδk(2)H^{k}_{\delta}(\mathbb{R}^{2}), where δ(0,1)\delta\in(0,1). It then follows from Sobolev embeddings that these functions are bounded. In addition, because δ>0\delta>0, they must decay algebraically.

z0z\to 0 zz\to\infty
K0(z)K_{0}(z) log(z/2)γe+O(z2)-\log(z/2)-\gamma_{e}+\mathrm{O}(z^{2}) π2zez(1+O(1/z))\sqrt{\frac{\pi}{2z}}\mathrm{e}^{-z}\Big{(}1+\mathrm{O}(1/z)\Big{)}
K1(z)K_{1}(z) 1z+O(z)\frac{1}{z}+\mathrm{O}(z) π2zez(1+O(1/z))\sqrt{\frac{\pi}{2z}}\mathrm{e}^{-z}\Big{(}1+\mathrm{O}(1/z)\Big{)}
Table 1. Asymptotic behavior for the Modified Bessel functions of the second kind of zeroth and first-order, taken from [1, (9.6.8), (9.6.9), (9.7.2)]

To do the matching, recall from the analysis in Subsection 3 that the parameter Λ2=bΩ\Lambda^{2}=b\Omega is assumed to be small beyond all orders of ε\varepsilon. This justifies the scaling r=η(ε)rηr=\eta(\varepsilon)r_{\eta}, where rηr_{\eta} is a constant and η(ε)=ε/Λ\eta(\varepsilon)=\varepsilon/\Lambda. As a result, Λr0\Lambda r\to 0 as ε0\varepsilon\to 0, while rr\to\infty, and we find that for small value of η\eta we are in the region where both approximations are valid. Moreover, since εo(η(ε))\varepsilon\sim\mathrm{o}(\eta(\varepsilon)) there is always an open interval where the two approximations can be matched, even as ε0\varepsilon\to 0. Because in this region the functions χM=χ=1\chi_{M}=\chi=1, we obtain

rϕfar(r;ε,Λ)\displaystyle\partial_{r}\phi_{far}(r;\varepsilon,\Lambda) 1b[ΛK0(Λr)K0(Λr)]+rϕ1\displaystyle\sim-\frac{1}{b}\;\left[\Lambda\frac{K^{\prime}_{0}(\Lambda r)}{K_{0}(\Lambda r)}\right]+\partial_{r}\phi_{1}
rϕint(r;ε)\displaystyle\partial_{r}\phi_{int}(r;\varepsilon) 1b[a/r+εrK(r)1+alogr+εK(r)]+rϕ¯1+rϕ¯2.\displaystyle\sim-\frac{1}{b}\;\left[\frac{a/r+\varepsilon\partial_{r}K(r)}{1+a\log r+\varepsilon K(r)}\right]+\partial_{r}\bar{\phi}_{1}+\partial_{r}\bar{\phi}_{2}.

Setting the derivatives equal to each other, rϕfar=rϕint\partial_{r}\phi_{far}=\partial_{r}\phi_{int}, we find that

(ΛK0(Λr)bK0(Λr)rϕ1)\displaystyle\Big{(}\Lambda K^{\prime}_{0}(\Lambda r)-bK_{0}(\Lambda r)\partial_{r}\phi_{1}\Big{)} (1+alogr+εK(r))\displaystyle\left(1+a\log r+\varepsilon K(r)\right)
=K0(Λr)[(a/r+εrK(r))b(rϕ¯1+rϕ¯2)(1+alogr+εK(r))]\displaystyle=K_{0}(\Lambda r)\Big{[}(a/r+\varepsilon\partial_{r}K(r))-b(\partial_{r}\bar{\phi}_{1}+\partial_{r}\bar{\phi}_{2})(1+a\log r+\varepsilon K(r))\Big{]}
(Λ(1Λr+O(Λr))\displaystyle\Big{(}-\Lambda\left(\frac{1}{\Lambda r}+\mathrm{O}(\Lambda r)\right) bK0(Λr)rϕ1)(1+alogr+εK(r))\displaystyle-bK_{0}(\Lambda r)\partial_{r}\phi_{1}\Big{)}\left(1+a\log r+\varepsilon K(r)\right)
=(log(Λ/2)γelog(r)+O((Λr)2))\displaystyle=\Big{(}-\log(\Lambda/2)-\gamma_{e}-\log(r)+\mathrm{O}((\Lambda r)^{2})\Big{)}
×[(a/r+εrK(r))b(rϕ¯1+rϕ¯2)(1+alogr+εK(r))],\displaystyle\qquad\times\Big{[}(a/r+\varepsilon\partial_{r}K(r))-b(\partial_{r}\bar{\phi}_{1}+\partial_{r}\bar{\phi}_{2})(1+a\log r+\varepsilon K(r))\Big{]},

where in the second line we use the fact that K0(z)=K1(z)K^{\prime}_{0}(z)=-K_{1}(z) and the expansions from Table 1.

We now proceed with the matched asymptotic analysis to determine the value of Λ\Lambda. Notice that due to the relation Ωb=Λ2\Omega b=\Lambda^{2}, this will also allow us to obtain an expression for the frequency. The method is as follows: We first divide the above expression by different gage functions in order to select terms of similar order in ε\varepsilon. We then cancel any duplicate terms, let ε\varepsilon go to zero, and select the value of any undefined constant so that the remaining terms add up to zero.

Because we are interested only in finding the value of the constant Λ\Lambda, we can simplify these computations by noticing that terms of the form O(Λr)rϕ1\mathrm{O}(\Lambda r)\partial_{r}\phi_{1} will go to zero, as ε0\varepsilon\to 0, faster than any other term. Thus, they are not of the same order in ε\varepsilon as elements that involve Λ\Lambda. This follows from the scalings picked and the algebraic decay rate of the the function rϕ1\partial_{r}\phi_{1}. We may therefore consider instead the expression

1r(1+alogr+εK(r))=\displaystyle-\frac{1}{r}\left(1+a\log r+\varepsilon K(r)\right)= (log(Λ/2)γelog(r))\displaystyle(-\log(\Lambda/2)-\gamma_{e}-\log(r)) (15)
×[(a/r+εrK(r))b(rϕ¯1+rϕ¯2)(1+alogr+εK(r))].\displaystyle\times[(a/r+\varepsilon\partial_{r}K(r))-b(\partial_{r}\bar{\phi}_{1}+\partial_{r}\bar{\phi}_{2})(1+a\log r+\varepsilon K(r))].

It is worth pointing out here that, in contrast to more standard matched asymptotic analyses, the elements in equation (15) are not of order O(εn),n\mathrm{O}(\varepsilon^{n}),n\in\mathbb{N}. Moreover, we find that dominants terms depend on the yet to be determined approximations ϕ¯i,i=1,2\bar{\phi}_{i},i=1,2. Thus, we will not be able to match them exactly, but we can justify that the process can be done.

First, looking at the right hand side, one notices that the dominant term is εlog(r)rK-\varepsilon\log(r)\partial_{r}K. Because rr depends on ε\varepsilon, we may use this function as a gage function. Dividing by εlog(r)rK-\varepsilon\log(r)\partial_{r}K and letting ε0\varepsilon\to 0, or equivalently rr\to\infty, we are left with matching,

0=1brϕ¯1+rϕ¯2εrK(1+alogr+εK).0=1-b\frac{\partial_{r}\bar{\phi}_{1}+\partial_{r}\bar{\phi}_{2}}{\varepsilon\partial_{r}K}(1+a\log r+\varepsilon K).

By picking the value of δ(0,1)\delta\in(0,1) so that the higher order correction term, ϕ¯2Lδ2(2)\bar{\phi}_{2}\in L^{2}_{\delta}(\mathbb{R}^{2}), is in the same space as both, KK and rK\partial_{r}K, we see that it is possible to match these terms. Expression (15) then becomes

1r(1+alogr+εK(r))=\displaystyle-\frac{1}{r}\left(1+a\log r+\varepsilon K(r)\right)= (log(Λ/2)γe)\displaystyle(-\log(\Lambda/2)-\gamma_{e})
×[(a/r+εrK(r))b(rϕ¯1+rϕ¯2)(1+alogr+εK(r))]\displaystyle\times[(a/r+\varepsilon\partial_{r}K(r))-b(\partial_{r}\bar{\phi}_{1}+\partial_{r}\bar{\phi}_{2})(1+a\log r+\varepsilon K(r))]
arlogr.\displaystyle-\frac{a}{r}\log r.

Second, cancelling the term arlogr\frac{a}{r}\log r, using εK/r\varepsilon K/r as a gage function, and letting rr\to\infty, we obtain

1=(log(Λ/2)γe)[rKrKbrεK(rϕ¯1+rϕ¯2)(1+alogr+εK(r))].-1=(-\log(\Lambda/2)-\gamma_{e})\left[\frac{\partial_{r}K}{rK}-\frac{br}{\varepsilon K}(\partial_{r}\bar{\phi}_{1}+\partial_{r}\bar{\phi}_{2})(1+a\log r+\varepsilon K(r))\right].

Since ϕ¯i,i=1,2\bar{\phi}_{i},i=1,2 represent all higher order terms, and not just one function, one can again justify that these terms can be matched. As a result, equation (15) now reads

1r=(logΛ/2γ3)ar.-\frac{1}{r}=(-\log\Lambda/2-\gamma_{3})\frac{a}{r}.

Finally, solving for Λ\Lambda, we see that Λ=2eγeexp(1/a)\Lambda=2\mathrm{e}^{-\gamma_{e}}\exp(1/a). Using the relation Λ2=bΩ\Lambda^{2}=b\Omega, we also obtain that

Ω=4be2γeexp(1a).\Omega=\frac{4}{b}\mathrm{e}^{-2\gamma_{e}}\exp\left(\frac{1}{a}\right).

In particular, from the definition of aa, i.e. a=bε0gc(r)r𝑑ra=-b\varepsilon\int_{0}^{\infty}g_{c}(r)\;r\;dr, we may conclude that both Λ\Lambda and Ω\Omega are smooth functions of ε\varepsilon, for all ε(0,εM)\varepsilon\in(0,\varepsilon_{M}), with εM\varepsilon_{M} a positive constant. In addition, notice that as ε\varepsilon approaches zero, the value of Λ\Lambda and Ω\Omega also goes to zero.

Remark 5.1.

Notice that:

  1. 1.

    We need the constant a<0a<0 in order for Λ=bΩ\Lambda=b\Omega to satisfy our initial assumption of being small beyond all orders of ε\varepsilon. If ε>0\varepsilon>0, this condition is guaranteed from formula (9) and the assumption that gg is a positive function.

  2. 2.

    Notice also that if εg<0\varepsilon\int g<0, the gradient rϕint\partial_{r}\phi_{int} would also be negative and we would not be able to match the two approximations. This is in line with previous results which show that target pattern solutions (or thanks to the Hopf-Cole transform, ϕ=1blog(Ψ)\phi=-\frac{1}{b}\log(\Psi), ground states of the Schrödinger eigenvalue problem, ΔΨ+εgΨ\Delta\Psi+\varepsilon g\Psi) do not exist when the inhomogeneity (potential) satisfies εg<0\varepsilon\int g<0. See [21] for a proof of this result.

  3. 3.

    Because we rigorously proved the existence of solutions to the intermediate and far field approximations, we know that we can obtain approximations for ϕfar\phi_{far} and ϕinter\phi_{inter} to any desired order. Thus, by matching these higher order approximations, we can obtain better estimates for the parameter Λ\Lambda. In particular, if we consider a=εa1+ε2a2a=\varepsilon a_{1}+\varepsilon^{2}a_{2}, and find the corresponding expressions for ϕfar,ϕinter\phi_{far},\phi_{inter} and a2a_{2}, the above matching process leads to Λ=C(ε)2eγeexp(1/εa1)\Lambda=C(\varepsilon)2\mathrm{e}^{-\gamma_{e}}\exp(1/\varepsilon a_{1}), with C(ε)=exp(1/a1/εa1)C(\varepsilon)=\exp(1/a-1/\varepsilon a_{1}). In addition, by defining Λ(0)=εΛ(0)=0\Lambda(0)=\partial_{\varepsilon}\Lambda(0)=0, we obtain that this estimate is also C1C^{1} with respect to ε\varepsilon on [0,εM)[0,\varepsilon_{M}), for some εM>0\varepsilon_{M}>0.

5.2. Existence of Solutions

In this subsection we combine the results of the previous subsections and prove Theorem 1, which is stated in the introduction and reproduced below for convenience.

Theorem.

Let k2k\geq 2 and σ(0,1)\sigma\in(0,1) and consider a function gHr,σk(2)g\in H^{k}_{r,\sigma}(\mathbb{R}^{2}) satisfying Hypothesis 1.1. Then, there exists a constant ε0>0\varepsilon_{0}>0 and a C1([0,ε0))C^{1}([0,\varepsilon_{0})) family of eigenfunctions ϕ=ϕ(r;ε)\phi=\phi(r;\varepsilon) and eigenvalues Ω=Ω(ε)\Omega=\Omega(\varepsilon) that bifurcate from zero and solve the equation

Δ0ϕb(rϕ)2εg(r)+Ω=0r=|x|[0,).\Delta_{0}\phi-b(\partial_{r}\phi)^{2}-\varepsilon g(r)+\Omega=0\qquad r=|x|\in[0,\infty). (16)

Moreover, this family has the form

ϕ(r;ε)=1bχ(Λr)log(K0(Λr))+ϕ1(r;ε)+εc,Λ2=bΩ(ε)\phi(r;\varepsilon)=-\frac{1}{b}\chi(\Lambda r)\log(K_{0}(\Lambda r))+\phi_{1}(r;\varepsilon)+\varepsilon c,\qquad\Lambda^{2}=b\Omega(\varepsilon)

where

  1. i)

    cc is a constant that depends on the initial conditions of the problem,

  2. ii)

    K0(z)K_{0}(z) represents the zeroth-order Modified Bessel function of the second kind,

  3. iii)

    rϕ1Hr,σk(2)\partial_{r}\phi_{1}\in H^{k}_{r,\sigma}(\mathbb{R}^{2}), and

  4. iv)

    Ω=Ω(ε)=C(ε)4e2γεexp[2/a]\Omega=\Omega(\varepsilon)=C(\varepsilon)4e^{-2\gamma_{\varepsilon}}\exp[2/a], with

    a=εb0gc(r)r𝑑r,a=-\varepsilon b\int_{0}^{\infty}g_{c}(r)\;r\;dr,

    and C(ε)C(\varepsilon) a C1C^{1} constant that depends on ε\varepsilon.

Proof.

The proof mimics the analysis done for the far field approximation, except that now we consider the full equation (16). As above, we use the ansatz ϕ(r)=ϕ0(r)+ϕ1(r)\phi(r)=\phi_{0}(r)+\phi_{1}(r), with ϕ0\phi_{0} given by

ϕ0(r)=1bχ(Λr)log(K0(Λr)),Λ2=bΩ>0.\phi_{0}(r)=-\frac{1}{b}\chi(\Lambda r)\log(K_{0}(\Lambda r)),\qquad\Lambda^{2}=b\Omega>0.

In contrast to the analysis from Section 4, here we treat the parameter Λ\Lambda as a C1C^{1} function of ε\varepsilon, a result that follows from the matched asymptotic analysis of Subsection 5.1. Thus, given any ε>0\varepsilon>0 there is a corresponding value of Λ\Lambda that defines an approximation, ϕ0\phi_{0}, and a frequency, Ω=Λ2/b\Omega=\Lambda^{2}/b, both of which satisfy the equation (Δ0ϕ0b(rϕ0)2+Ω)=0(\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}+\Omega)=0 in the far field.

Inserting this ansatz into equation (16) gives

Δ0ϕ12brϕ0rϕ1b(rϕ1)2+(Δ0ϕ0b(rϕ0)2+Ω)εg=0.\Delta_{0}\phi_{1}-2b\partial_{r}\phi_{0}\partial_{r}\phi_{1}-b(\partial_{r}\phi_{1})^{2}+(\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}+\Omega)-\varepsilon g=0.

Letting ψ=rϕ1\psi=\partial_{r}\phi_{1}, adding and subtracting the term 2Λψ2\Lambda\psi, and precondition the result by 2Λ1\mathcal{L}_{2\Lambda}^{-1}, gives the following equivalent formulation of equation (16),

F(ψ;ε)=Id+2Λ1[2brϕ0ψbψ2+(Δ0ϕ0b(rϕ0)2+Ω)εg2Λψ]=0.F(\psi;\varepsilon)=\mathrm{Id}+\mathcal{L}_{2\Lambda}^{-1}\Big{[}-2b\partial_{r}\phi_{0}\psi-b\psi^{2}+(\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}+\Omega)-\varepsilon g-2\Lambda\psi\Big{]}=0. (17)

Our goal is to show that the operator F:Hr,σk(2)×Hr,σk(2)F:H^{k}_{r,\sigma}(\mathbb{R}^{2})\times\mathbb{R}\rightarrow H^{k}_{r,\sigma}(\mathbb{R}^{2}) satisfies the conditions of the implicit function theorem.

By Remark 5.1, Λ(0)=εΛ(0)=0\Lambda(0)=\partial_{\varepsilon}\Lambda(0)=0, so that the operator FF is C1([0,εM))C^{1}([0,\varepsilon_{M})) with respect to ε\varepsilon, for some εM>0\varepsilon_{M}>0. Moreover, thanks to the cut-off function in the definition of ϕ0\phi_{0}, i.e. χ=χ(Λr)\chi=\chi(\Lambda r), we find that the terms (Δ0ϕ0b(rϕ0)2+Ω)(\Delta_{0}\phi_{0}-b(\partial_{r}\phi_{0})^{2}+\Omega) tend to zero as ε\varepsilon goes to zero. Therefore, F(0;0)=0F(0;0)=0. In addition, because the elements in the parenthesis are smooth and have compact support, they belong to the space Hr,σk1(2)H^{k-1}_{r,\sigma}(\mathbb{R}^{2}), for any natural number kk and any real number σ\sigma. A similar analysis as in the proof of Theorem 3 then shows that the rest of the terms in FF belong to the space Hr,σk(2)H^{k}_{r,\sigma}(\mathbb{R}^{2}), with k2,k\geq 2, and σ(0,1)\sigma\in(0,1). As a result, the operator FF is also well defined. Since its Fréchet derivative, DψF(0;0)D_{\psi}F(0;0), is now the identity map on Hr,σk(2)H^{k}_{r,\sigma}(\mathbb{R}^{2}), we may apply the implicit function theorem to conclude the existence of solutions ψ=rϕHr,σk(2)\psi=\partial_{r}\phi\in H^{k}_{r,\sigma}(\mathbb{R}^{2}). The results of Theorem 1 then follow in a similar way as those done in Section 4. ∎

6. Simulations

In this section we numerically explore the effects of adding large inhomogeneities, εg\varepsilon g, as perturbations to the eikonal equation, i.e.

ϕt=Δϕ|ϕ|2εg.\phi_{t}=\Delta\phi-|\nabla\phi|^{2}-\varepsilon g. (18)

To run the simulations we model the equation on a square domain with periodic boundary conditions and employ a spectral RK4 method based on [14], using a mesh size h=100/512h=100/512 and a time step dt=0.5dt=0.5. The numerical scheme is continued until a steady state is reached. Different domain lengths were tested, (L={100,120,140,160,180,200}L=\{100,120,140,160,180,200\}), resulting in the same approximations for ϕ\phi. Thus, a domain of length L=100L=100 was chosen to run all numerical experiments for computational efficiency.

Refer to caption
(a)
Refer to caption
(b)
Figure 1. Numerical simulation of the time dependent eikonal equation (18) with gg as in (19), p=0.8p=0.8, initial condition ϕ=0\phi=0, and various values of εa=0.15(3:2:20)\varepsilon\sim a=0.15*(3:2:20), where a=ε03g(r)r𝑑ra=\varepsilon\int_{0}^{3}g(r)r\;dr. A) Plots of the gradient of the steady state solution. Top most curve corresponds to maximum value of ε\varepsilon used. B) Plot of 1log(k(a))1\frac{1}{\log(k(a))-1} vs. aa, where for large |x||x| the gradient ϕ\nabla\phi approximates the wavenumber, k(a)k(a).Circles represent data from the simulation while dotted line is the linear fit.

Simulations confirm our analytical results, finding that for inhomogeneities that take the form

g=(A(1+r2)p),p(1/2,1],A,g=\left(\frac{A}{(1+r^{2})^{p}}\right),\quad p\in(1/2,1],\quad A\in\mathbb{R}, (19)

the solutions to the eikonal equation grow linearly at infinity. This is depicted in Figure 1(a) where the gradient, ϕ\nabla\phi, is plotted for different values of the parameter ε\varepsilon. Notice that because we are using periodic boundary conditions, the value of ϕ\nabla\phi goes to zero at the boundary of the domain. As predicted by the analysis of the previous sections, we find that the wavenumber, k=lim|x|ϕΛ/bk=\lim_{|x|\to\infty}\nabla\phi\sim\Lambda/b, and as a result the frequency, Ω=Λ2/b\Omega=\Lambda^{2}/b, is small beyond all orders of ε\varepsilon. To confirm this result we approximate the wavenumber by evaluating the gradient ϕ\nabla\phi at large values of |x||x|. In Figure 1(b) we plot the relation 1log(k)1\frac{-1}{\log(k)-1} vs. aa, where aa represents the mass of gc=(1χ)gg_{c}=(1-\chi)g, which we take as a substitute for ε\varepsilon, since a=bεgc(r)r𝑑ra=-b\varepsilon\int g_{c}(r)r\;dr. Notice how in the figure the data points taken from the simulations follow a straight line, confirming that Λexp(1/a)\Lambda\sim\exp(1/a).

Finally, to determine how the the decay rate, pp, affects the wavenumber, we ran simulations for values of p(0.5,3]p\in(0.5,3]. Notice that using the notation from Hypothesis 1.1, where g1/rmg\sim 1/r^{m}, this is equivalent to considering values of m(1,6)m\in(1,6). These results are summarized in Figure 2(a). They show that the wavenumber decreases as the decay rate of the inhomogeneity, pp, increases. The figure also compares the numerical approximation to the wavenumber, kk, which we plot using stars, with the analytical result kexp(1/a)k\sim\exp(1/a). In particular, following Theorem 1 we use

a={εb03g(r)r𝑑rforp(1/2,1)εb0g(r)r𝑑rforp(1,3).a=\left\{\begin{array}[]{c c c }-\varepsilon b\int_{0}^{3}g(r)\;r\;dr&\mbox{for}&p\in(1/2,1)\\ -\varepsilon b\int_{0}^{\infty}g(r)\;r\;dr&\mbox{for}&p\in(1,3).\end{array}\right.

For values of p(0.5,1)m(1,2)p\in(0.5,1)\sim m\in(1,2), we are in the regime considered in this paper, where the impurity gg does not have finite mass and is thus a large inhomogeneity. In this case, we assume a value of D3D\sim 3 in the definition of gcg_{c} specified in the introduction, see equation (3). We then calculate the mass of this function by integrating from 0 to 3, since this provided the best fit to the data (see Remark 1.5). This approximation is plotted using a dashed line. On the other hand, when p(1,3)p\in(1,3) we are in the regime where the impurity, gg, has finite mass and the results from [13] apply (see also Theorem 1 and Remark 1.4). In this case, the value of a=εb0g(r)r𝑑ra=-\varepsilon b\int_{0}^{\infty}g(r)\;r\;dr. This approximation is plotted using a solid line.

Notice that both approximations for the wavenumberm, kk, do a good job of following the data in the respective regions of the pp-axis where they are valid, i.e. 0.5<p<10.5<p<1 for the dashed line, p>2p>2 for the solid line. However, the estimates for p(1,2)p\in(1,2) using the mass of gg (solid line) are not accurate, even though they follow the results from Theorem 1 in [13], or equivalently, Theorem 1 together with Remark 1.4 stated in this paper. This is not unreasonable given that the frequency of the pattern, Ω\Omega, and as a result its wavenumber, kk, are both small beyond all orders of the parameter ε\varepsilon. In particular, when p1p\to 1 we have that a=εbg(r)r𝑑ra=-\varepsilon b\int g(r)r\;dr\to-\infty. Because aO(ε)a\sim\mathrm{O}(\varepsilon), the estimates for Ωexp(1/a)\Omega\sim\exp(1/a) become worse and worse, and in this case one needs to approximate aa to higher orders in ε\varepsilon to obtain better estimates. Figure 2(a) then suggests that the interval 1<p<21<p<2 is a transitional regime, where one can numerically obtain a better fit to the data by using a cut-off function to better approximate the value of aa.

Finally, we also confirm numerically that for values of p0.5p\leq 0.5 the inhomogeneity no longer produces target patterns, but rather solutions with ϕO(r)\nabla\phi\sim\mathrm{O}(r) at infinity, see Figure 2(b). This is not a tight bound on the growth rate of ϕ\nabla\phi and is just a very rough estimate based on our numerical experiments.

Refer to caption
(a)
Refer to caption
(b)
Figure 2. A) Plot of wavenumber kk vs. pp for steady state solutions of the eikonal equation using g=A(1+r2)pg=\dfrac{A}{(1+r^{2})^{p}} with A=1.5A=1.5 and for p(0.5,3)p\in(0.5,3). Stars represents results from simulation, while solid and dashed lines represents approximation with kexp(1/a)k\sim\exp(-1/a), with a=A/(22p)a=-A/(2-2p) for solid line, a=A03g(r)r𝑑ra=A\int_{0}^{3}g(r)\;rdr for dashed line. B) Plot of ϕ\nabla\phi vs. rr, for values of p=0.3,0.8,1.5p=0.3,0.8,1.5.

7. Discussion

In this paper we showed that large defects can generate target patterns in oscillatory media. Under the assumption of weak coupling, we modeled such systems using a viscous eikonal equation, and represented the defect as a localized inhomogeneity. In contrast to previous results, which assume that the inhomogeneity is strongly localized, in this paper we relaxed this assumption and described impurities as functions with algebraic decay of order O(1/|x|m)\mathrm{O}(1/|x|^{m}), 1<m21<m\leq 2.

Our main motivation for studying this problem came from the universality of the viscous eikonal equation as a model for the phase dynamics of coherent structures in oscillatory media. In particular, our interest stems from the fact that this same equation can be used to describe the phase dynamics of spiral waves in oscillatory media with nonlocal coupling. In this context, the large inhomogeneity no longer represents a defect, but instead encodes information about variations in the amplitude of the pattern.

A second motivation came from the fact that the steady state viscous eikonal equation is conjugate to a Schrödinger eigenvalue problem. Indeed, it is well known that the Hopf-Cole transformation maps target pattern solutions to bound states of the corresponding Schrödinger operator, and that the frequency of target pattern solutions then corresponds to the energy of these states. In this context, the results presented here expand the conditions on the Schrödinger potential that allow for such bound states to exist. In particular, we show that Schrödinger operators with potentials that decay sufficiently fast at infinity can have bound states even when the mass of the potential 2g(r)r𝑑r\int_{\mathbb{R}^{2}}g(r)\;r\;dr is not finite.

In particular, our analysis provides a first order approximation for target pattern solutions and for their frequency. In agreement with simulations we show that, just as in the case of small defects, the frequency is small beyond all orders of the small parameter used to describe the strength of the impurity. As a result, solutions do not follow a regular expansion. Therefore, to obtain our results we first found intermediate and far field approximations to the steady state viscous eikonal equation. Then using a matched asymptotic analysis we were able to determine the value of the frequency selected by the system. This approach is similar in spirit to the one used to prove existence of target patterns and spiral waves in reaction-diffusion equations using spatial dynamics, [20, 15]. There, the modeling equations are viewed as a system of ordinary differential equation in the radial variable, and a center manifold reduction is used to obtain a vector field describing the amplitude of these patterns. Coherent structures then correspond to heteroclinic solutions, connecting a fixed point at infinity with solutions that are bounded near the origin. Our matching process is then equivalent to showing that the center-stable manifold of the fixed point intersects transversely the solution curve that lives in the center manifold.

Finally, the analysis presented in this paper is complemented by simulations of the viscous eikonal equation. Our numerical experiments are in good agreement with simulations. They confirm that the wavenumber, and therefore the frequency of target patterns, do not follow a regular expansion on the small parameter ε\varepsilon representing the strength of the impurity gg. They also confirm that when m1m\leq 1, the solutions to the viscous eikonal equation no longer represent target patterns, since in this case the gradient ϕ\nabla\phi does not approach a constant as |x||x|\to\infty.

8. Appendix

In [10] it was shown that the following amplitude equation governs the dynamics of one-armed spiral waves in nonlocal oscillatory media,

0=βΔ1w+λw+α|w|2w+N(w,ε),r[0,).0=\beta\Delta_{1}w+\lambda w+\alpha|w|^{2}w+N(w,\varepsilon),\quad r\in[0,\infty).

Here ww is a radial and complex-valued function, and

β=(σελ),λ,α,NO(|ε||w|4).\beta=(\sigma-\varepsilon\lambda),\quad\lambda,\alpha\in\mathbb{C},\quad N\sim\mathrm{O}(|\varepsilon||w|^{4}).

It was also established in [10] that the constant λI\lambda_{I} is an unknown parameter that needs to be determined when solving the equation.

In this section a multiple-scale analysis is used to derive a steady state viscous eikonal equation from the above expression. We will see that this eikonal equation is of the form considered in this paper and that it involves an inhomogeneity that decays at order O(1/|x|2)\mathrm{O}(1/|x|^{2}).

To accomplish this task we first let w=Aw~w=A\tilde{w}, with A2=λR/αRA^{2}=-\lambda_{R}/\alpha_{R}. This change of variables is done for convenience and leads to the following equation,

0=βΔ1w~+λw~+(λR+iα~I)|w~|2w~+N(w~,ε),α~I=αλR/αR.0=\beta\Delta_{1}\tilde{w}+\lambda\tilde{w}+(-\lambda_{R}+\mathrm{i}\tilde{\alpha}_{I})|\tilde{w}|^{2}\tilde{w}+N(\tilde{w},\varepsilon),\qquad\tilde{\alpha}_{I}=-\alpha\lambda_{R}/\alpha_{R}.

Letting w~=ρeiϕ\tilde{w}=\rho\mathrm{e}^{\mathrm{i}\phi} and separating the real and imaginary parts of the above expression, one finally obtains the system

0=\displaystyle 0= βR[Δ1ρ(rϕ)2ρ]βI[Δ0ϕρ+2rϕrρ]+λRρλRρ3+Re[N(w~;ε)eiϕ]\displaystyle\beta_{R}\left[\Delta_{1}\rho-(\partial_{r}\phi)^{2}\rho\right]-\beta_{I}\left[\Delta_{0}\phi\rho+2\partial_{r}\phi\partial_{r}\rho\right]+\lambda_{R}\rho-\lambda_{R}\rho^{3}+\mathrm{Re}\left[N(\tilde{w};\varepsilon)\mathrm{e}^{-\mathrm{i}\phi}\right] (20)
0=\displaystyle 0= βR[Δ0ϕρ+2rϕrρ]+βI[Δ1ρ(ϕ)2ρ]+λIρ+α~Iρ3+Im[N(w~;ε)eiϕ].\displaystyle\beta_{R}\left[\Delta_{0}\phi\rho+2\partial_{r}\phi\partial_{r}\rho\right]+\beta_{I}\left[\Delta_{1}\rho-(\partial\phi)^{2}\rho\right]+\lambda_{I}\rho+\tilde{\alpha}_{I}\rho^{3}+\mathrm{Im}\left[N(\tilde{w};\varepsilon)\mathrm{e}^{-\mathrm{i}\phi}\right]. (21)

Next, we proceed with a perturbation analysis following [4]. We rescale the variable rr by defining S=δrS=\delta r, where δ\delta is assumed to be a small positive parameter. We also use the following expressions for the unknown functions:

ρ=ρ0+δ2(R0+δR1),ρ0=ρ0(r),Ri=Ri(δr)i=0,1ϕ=ϕ0+δϕ1,ϕi=ϕi(δr)i=0,1.\begin{array}[]{r l c l }\rho=&\rho_{0}+\delta^{2}(R_{0}+\delta R_{1}),&\rho_{0}=\rho_{0}(r),&R_{i}=R_{i}(\delta r)\quad i=0,1\\ \phi=&\phi_{0}+\delta\phi_{1},&&\phi_{i}=\phi_{i}(\delta r)\hskip 14.22636pti=0,1.\end{array}

And for the parameter we choose λI=α~I+δ2λ~I,\lambda_{I}=-\tilde{\alpha}_{I}+\delta^{2}\tilde{\lambda}_{I}, with α~\tilde{\alpha} as above and λ~I\tilde{\lambda}_{I} a free parameter.

Inserting the above ansatz into the equations (20) and (21) we obtain a set of equations in powers of δ\delta. To write this equations more compactly, we use the subscript SS to distinguish operators that are applied to functions that depend on this variable, i.e. Δ0,S\Delta_{0,S}. The absence of this subscript indicates that the operator is applied to a function of the original variable rr.

At order O(1)\mathrm{O}(1) we find that ρ0\rho_{0} must satisfy,

0=\displaystyle 0= βRΔ1ρ0+λRρ0λRρ03,\displaystyle\beta_{R}\Delta_{1}\rho_{0}+\lambda_{R}\rho_{0}-\lambda_{R}\rho_{0}^{3},
0=\displaystyle 0= βIΔ1ρ0α~Iρ0+α~Iρ03.\displaystyle\beta_{I}\Delta_{1}\rho_{0}-\tilde{\alpha}_{I}\rho_{0}+\tilde{\alpha}_{I}\rho_{0}^{3}.

At the next order, O(δ2)\mathrm{O}(\delta^{2}), we find two equations involving R0R_{0} and ϕ0\phi_{0},

0=\displaystyle 0= βIρ0Δ0,Sϕ02βISϕ0Sρ0βRρ0(Sϕ0)2+λRR0(13ρ02),\displaystyle-\beta_{I}\rho_{0}\Delta_{0,S}\phi_{0}-2\beta_{I}\partial_{S}\phi_{0}\partial_{S}\rho_{0}-\beta_{R}\rho_{0}(\partial_{S}\phi_{0})^{2}+\lambda_{R}R_{0}(1-3\rho_{0}^{2}),
0=\displaystyle 0= βRρ0Δ0,Sϕ0+2βRSϕ0Sρ0βIρ0(Sϕ0)2+α~IR0(3ρ021)+λ~Iρ0.\displaystyle\beta_{R}\rho_{0}\Delta_{0,S}\phi_{0}+2\beta_{R}\partial_{S}\phi_{0}\partial_{S}\rho_{0}-\beta_{I}\rho_{0}(\partial_{S}\phi_{0})^{2}+\tilde{\alpha}_{I}R_{0}(3\rho_{0}^{2}-1)+\tilde{\lambda}_{I}\rho_{0}.

For our purposes, it is enough to stop at this stage and not list higher order terms.

We first focus on the order O(1)\mathrm{O}(1) system. The first equation can be solved, provided βR,λR>0\beta_{R},\lambda_{R}>0. This equation falls into a broader family of o.d.e. which were solved in [16]. In this reference, the authors showed that such equations posses a unique solution ρ\rho_{*} satisfying

ρ1asr,ρ(r)brwhenr0\rho_{*}\to 1\quad\mbox{as}\quad r\to\infty,\qquad\rho_{*}(r)\sim br\quad\mbox{when}\quad r\sim 0

Of course, the solution ρ\rho_{*} would not satisfy the second equation in the system. So we let

G=βIΔ1ρα~Iρ+α~Iρ3=(βIβRλR+α~I)ρ(ρ21),G=\beta_{I}\Delta_{1}\rho_{*}-\tilde{\alpha}_{I}\rho_{*}+\tilde{\alpha}_{I}\rho_{*}^{3}=\left(\frac{\beta_{I}}{\beta_{R}}\lambda_{R}+\tilde{\alpha}_{I}\right)\rho_{*}(\rho_{*}^{2}-1),

and add these terms to the order O(δ2)\mathrm{O}(\delta^{2}) system.

Refer to caption
Refer to caption
Figure 3. Solution to the boundary value problem (22)

Going back to the order O(δ2)\mathrm{O}(\delta^{2}) system, we first notice that because ρ0=ρbr=bS/δ\rho_{0}=\rho_{*}\sim br=bS/\delta near the origin, then the terms that involve this variable are in fact ’large’ when compared to the terms that do not. Concentrating only on these large terms, we find that in the first equation we can solve for R0R_{0} in terms of the variable ϕ0\phi_{0}. Inserting this result into the second equation gives us the viscous eikonal equation,

Δ0,Sϕ0b(Sϕ0)2+Ωcg=0\Delta_{0,S}\phi_{0}-b(\partial_{S}\phi_{0})^{2}+\Omega-cg=0

as expected, where

b=βIλRβRα~Iα~IβI+λRβR,Ω=λ~IλRα~IβI+λRβR,c=βIλR+α~IβRβR(α~IβI+λRβR),g=(1ρ2).b=\frac{\beta_{I}\lambda_{R}-\beta_{R}\tilde{\alpha}_{I}}{\tilde{\alpha}_{I}\beta_{I}+\lambda_{R}\beta_{R}},\qquad\Omega=\frac{\tilde{\lambda}_{I}\lambda_{R}}{\tilde{\alpha}_{I}\beta_{I}+\lambda_{R}\beta_{R}},\qquad c=-\frac{\beta_{I}\lambda_{R}+\tilde{\alpha}_{I}\beta_{R}}{\beta_{R}(\tilde{\alpha}_{I}\beta_{I}+\lambda_{R}\beta_{R})},\qquad g=(1-\rho_{*}^{2}).

Numerical simulations show that the perturbation gg decays at order O(1/r2)\mathrm{O}(1/r^{2}) as rr goes to infinity, see Figure 3. To obtain these results, we solved the boundary value problem

0=rrρ+1rrρ1r2ρ+ρρ3,ρ()=1,ρ(0)=0,0=\partial_{rr}\rho+\frac{1}{r}\partial_{r}\rho-\frac{1}{r^{2}}\rho+\rho-\rho^{3},\qquad\rho(\infty)=1,\quad\rho(0)=0, (22)

treating the equation as a system of o.d.e. and using a shooting method with condition

ρ(r)brwhenr0.\rho(r)\sim br\quad\mbox{when}\quad r\sim 0.

9. Declarations

Conflict of Interest: The author declares that she has no conflict of interest.

References

  • [1] Milton Abramowitz, Irene A Stegun, and Robert H Romer. Handbook of mathematical functions with formulas, graphs, and mathematical tables, 1988.
  • [2] Robert A Adams and John JF Fournier. Sobolev spaces. Elsevier, 2003.
  • [3] Fernanda Alcantara and Marilyn Monk. Signal propagation during aggregation in the slime mould dictyostelium discoideum. Microbiology, 85(2):321–334, 1974.
  • [4] Arjen Doelman, Björn Sandstede, Arnd Scheel, and Guido Schneider. The dynamics of modulated wave trains. American Mathematical Soc., 2009.
  • [5] A.J. Durston. Pacemaker activity during aggregation in dictyostelium discoideum. Developmental Biology, 37(2):225–235, 1974.
  • [6] Gerald B Folland. Real analysis: modern techniques and their applications, volume 40. John Wiley & Sons, 1999.
  • [7] Patrick S Hagan. Target patterns in reaction-diffusion systems. Advances in Applied Mathematics, 2(4):400–416, 1981.
  • [8] Matthew Hendrey, Keeyeol Nam, Parvez Guzdar, and Edward Ott. Target waves in the complex Ginzburg-Landau equation. Phys. Rev. E, 62:7627–7631, Dec 2000.
  • [9] Gabriela Jaramillo. Inhomogeneities in 3 dimensional oscillatory media. Networks and Heterogeneous Media, 10(2):387–399, 2015.
  • [10] Gabriela Jaramillo. Rotating spirals in oscillatory media with nonlocal interactions and their normal form. Discrete and Continuous Dynamical Systems - S, 0:–, 2022.
  • [11] Gabriela Jaramillo and Arnd Scheel. Pacemakers in large arrays of oscillators with nonlocal coupling. Journal of Differential Equations, 260(3):2060–2090, 2016.
  • [12] Gabriela Jaramillo, Arnd Scheel, and Qiliang Wu. The effect of impurities on striped phases. Proceedings of the Royal Society of Edinburgh: Section A Mathematics, 149(1):131– 168, 2019.
  • [13] Gabriela Jaramillo and Shankar C Venkataramani. Target patterns in a 2d array of oscillators with nonlocal coupling. Nonlinearity, 31(9):4162–4201, jul 2018.
  • [14] Aly-Khan Kassam and Lloyd N. Trefethen. Fourth-order time-stepping for stiff PDEs. SIAM Journal on Scientific Computing, 26(4):1214–1233, 2005.
  • [15] Richard Kollár and Arnd Scheel. Coherent structures generated by inhomogeneities in oscillatory media. SIAM Journal on Applied Dynamical Systems, 6(1):236–262, 2007.
  • [16] N Kopell and L.N Howard. Target pattern and spiral solutions to reaction-diffusion equations with more than one space dimension. Advances in Applied Mathematics, 2(4):417 – 449, 1981.
  • [17] Nancy Kopell. Target pattern solutions to reaction-diffusion equations in the presence of impurities. Advances in Applied Mathematics, 2(4):389–399, 1981.
  • [18] Yoshiki Kuramoto. Chemical oscillations, waves, and turbulence. Courier Corporation, 2003.
  • [19] Robert C. McOwen. The behavior of the Laplacian on weighted Sobolev spaces. Communications on Pure and Applied Mathematics, 32(6):783–795, 1979.
  • [20] Arnd Scheel. Bifurcation to spiral waves in reaction-diffusion systems. SIAM Journal on Mathematical Analysis, 29(6):1399–1418, 1998.
  • [21] Barry Simon. The bound state of weakly coupled Schrödinger operators in one and two dimensions. Annals of Physics, 97(2):279–288, 1976.
  • [22] Michael Stich and Alexander S. Mikhailov. Complex pacemakers and wave sinks in heterogeneous oscillatory chemical systems. 216(4):521–521, 2002.
  • [23] Michael Stich and Alexander S. Mikhailov. Target patterns in two-dimensional heterogeneous oscillatory reaction–diffusion systems. Physica D: Nonlinear Phenomena, 215(1):38–45, 2006.
  • [24] Rory G. Townsend, Selina S. Solomon, Spencer C. Chen, Alexander N.J. Pietersen, Paul R. Martin, Samuel G. Solomon, and Pulin Gong. Emergence of complex wave patterns in primate cerebral cortex. Journal of Neuroscience, 35(11):4657–4662, 2015.
  • [25] John J. Tyson and Paul C. Fife. Target patterns in a realistic model of the Belousov-Zhabotinskii reaction. The Journal of Chemical Physics, 73(5):2224–2237, 1980.
  • [26] Janpeter Wolff, Michael Stich, Carsten Beta, and Harm Hinrich Rotermund. Laser-induced target patterns in the oscillatory CO oxidation on Pt(110). The Journal of Physical Chemistry B, 108(38):14282–14291, 09 2004.
  • [27] A. N. Zaikin and A. M. Zhabotinsky. Concentration wave propagation in two-dimensional liquid-phase self-oscillating system. Nature, 225(5232):535–537, 1970.