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

2D Thin obstacle problem with data at infinity

Runcao Lyu Department of Mathematical Sciences, University of Science and Technology of China, Hefei, China [email protected]  and  Zikai Ye Department of Mathematical Sciences, University of Science and Technology of China, Hefei, China [email protected]
Abstract.

In this paper, we consider the thin obstacle problem in 2\mathbb{R}^{2} with data at infinity. We first prove the existence and uniqueness of it. Then we show that its symmetric solutions are actually half-space solutions. Our results are needed when classifying the half-space (2k12)(2k-\frac{1}{2})-homogeneous solutions to the thin obstacle problems in 3\mathbb{R}^{3}. It is a generalization of one part of Savin-Yu’s work [9] on classifying the half-space 72\frac{7}{2}-homogeneous solutions.

1. Introduction

Thin obstacle problem, also called Signorini Problem [11], is a type of free boundary problem which studies minimizers of the Dirichlet energy

B1+|u|2\int_{B_{1}^{+}}|\nabla u|^{2}

in {uW1,2(B1+) — u=g on B1{xn+1>0}, u0 on B1{xn+1=0}},\{u\in W^{1,2}(B_{1}^{+})\text{ | }u=g\text{ on }\partial B_{1}\cap\{x_{n+1}>0\},\text{ }u\geq 0\text{ on }B_{1}\cap\{x_{n+1}=0\}\}, where B1+=B1{xn+1>0}.B_{1}^{+}=B_{1}\cap\{x_{n+1}>0\}. The conditions u=g on B1{xn+1>0}, u0 on B1{xn+1=0}u=g\text{ on }\partial B_{1}\cap\{x_{n+1}>0\},\text{ }u\geq 0\text{ on }B_{1}\cap\{x_{n+1}=0\} are understood in the sense of trace.

By standard variational approach and even reflection, these minimizers satisfy the following Euler-Lagrange equations

(1.1) {Δu0inB1u0inB1{xn+1=0}Δu=0inB1({u>0}{xn+10}).\begin{cases}\Delta u\leq 0\ \ &in\ B_{1}\\ u\geq 0\ \ &in\ B_{1}\cap\{x_{n+1}=0\}\\ \Delta u=0\ &in\ B_{1}\cap(\{u>0\}\cup\{x_{n+1}\neq 0\}).\end{cases}

in a Euclidean ball B1B_{1} in n+1\mathbb{R}^{n+1}. Since the odd part (with respect to xn+1x_{n+1}) of uu is harmonic and vanishes on {xn+1=0}\{x_{n+1}=0\}, it suffices to consider even solution. It is obvious that a rotation around xn+1x_{n+1}-axis or positive multiples (we call it normalization) of a solution uu to (1.1) is still a solution.

Early results about the regularity of uu were given by Richardson [8] and Ural’tseva [13]. Then the optimal regularity of uu was proved by Athanaspoulos and Caffarelli [4] that uu is locally Lipschitz in B1B_{1} and locally C1,1/2C^{1,1/2} in B1{xn+10}B_{1}\cap\{x_{n+1}\geq 0\}. With the help of optimal regularity, several studies were done to study the contact set Λ(u):={u=0}{xn+1=0}\Lambda(u):=\{u=0\}\cap\{x_{n+1}=0\} and the free boundary Γ(u):=nΛ(u)\Gamma(u):=\partial_{\mathbb{R}^{n}}\Lambda(u) as follows.

By making use of Almgren’s frequency and monotonicity formula [3], Athanasopoulos-Caffarelli-Salsa [5] showed that for any qΛ(u)q\in\Lambda(u), there is a constant λq\lambda_{q} such that

uL2(Br(q))rn2+λq\left\lVert u\right\rVert_{L^{2}(\partial B_{r}(q))}\sim r^{\frac{n}{2}+\lambda_{q}}

as r0r\to 0 and up to subsequence,

uq,r:=rn2u(r+q)uL2(Br(q))u0.u_{q,r}:=r^{\frac{n}{2}}\frac{u(r\cdot+q)}{\left\lVert u\right\rVert_{L^{2}(\partial B_{r}(q))}}\to u_{0}.

as r0r\to 0. The constant λq\lambda_{q} is called the frequency of uu at qq and u0u_{0} is a λq\lambda_{q}-homogeneous solution to (1.1), called a blow-up profile of uu at qq. A constant λ\lambda\in\mathbb{R} is called an admissible frequency if it is a frequency of some solution uu to (1.1) at some point qΛ(u)q\in\Lambda(u). There are many results on the classification of admissible frequencies

Φ:={λ — there is a non-trivial λ-homogeneous solution to (1.1)}\Phi:=\{\lambda\in\mathbb{R}\text{ | there is a non-trivial }\lambda\text{-homogeneous solution to \eqref{thin}}\}

and λ\lambda-homogeneous solutions

𝒫λ:={u — u is a solution to (1.1) with xu=λu}\mathcal{P}_{\lambda}:=\{u\text{ | }u\text{ is a solution to \eqref{thin} with }x\cdot\nabla u=\lambda u\}

for each admissible frequency λΦ\lambda\in\Phi.

The classification in the case of 2\mathbb{R}^{2} is completed. See Petrosyan-Shahgholian-Ural’tseva [2]. The set of all admissible frequencies is

Φ={2k12 — k}.\Phi=\mathbb{N}\cup\{2k-\frac{1}{2}\text{ | }k\in\mathbb{N}\}.

The homogeneous solutions with integer frequencies, up to normalization, are actually even reflections of polynomials. As for the (2k12)(2k-\frac{1}{2})-homogeneous solutions, up to normalization, one has

𝒫2k12={au2k12:=ar2k12cos((2k12)θ) — a0},\mathcal{P}_{2k-\frac{1}{2}}=\{au_{2k-\frac{1}{2}}:=ar^{2k-\frac{1}{2}}\cos((2k-\frac{1}{2})\theta)\text{ | }a\geq 0\},

where (r,θ)(r,\theta) is the polar coordinate of 2\mathbb{R}^{2}. They vanish on the half line {x10x2=0}\{x_{1}\geq 0\text{, }x_{2}=0\} and satisfy

spt(Δu)=Λ(u)={x10x2=0}\text{spt}(\Delta u)=\Lambda(u)=\{x_{1}\leq 0\text{, }x_{2}=0\}

In higher dimensions, the classifications of admissible frequencies and homogeneous solutions remain open. Athanasopoulos-Caffarelli-Salsa [5] showed that the set of admissible frequencies satisfies

{2k12 — k}Φ{1,32}[2,+),\mathbb{N}\cup\{2k-\frac{1}{2}\text{ | }k\in\mathbb{N}\}\subset\Phi\subset\{1,\frac{3}{2}\}\cup[2,+\infty),

while the left-hand side is obtained by extending the homogeneous solutions from 2\mathbb{R}^{2}. Moreover, Columbo-Spolaor-Velichkov [6] and Savin-Yu [10] showed that there is a frequency gap around each integer.

When it comes to the classification of homogeneous solutions, the case of integral frequencies are studied by Figalli-Ros-Oton-Serra [1] and Garofalo-Petrosyan [7]. It is shown by Athanasopoulos-Caffarelli-Salsa that

𝒫32={Normalizations of u32}.\mathcal{P}_{\frac{3}{2}}=\{\text{Normalizations of }u_{\frac{3}{2}}\}.

For the homogeneous solutions with frequency 2k122k-\frac{1}{2}, k=2,3,k=2,3,..., it is still tough now to classify them. However, for half-space solution uu of frequency 72\frac{7}{2} in 3\mathbb{R}^{3}, i.e. up to normalization,

either spt(Δu){x20x3=0}, or spt(Δu){x20x3=0},\text{either spt}(\Delta u)\subset\{x_{2}\leq 0\text{, }x_{3}=0\}\text{, or spt}(\Delta u)\supset\{x_{2}\leq 0\text{, }x_{3}=0\},

all the half-space solutions uu of frequency 72\frac{7}{2} in 3\mathbb{R}^{3} we currently know consist of (up to normalization)

1={u72+a1x1u52+a2(x1215r2)u32 — 0a25a12Γ(a2)},\mathcal{F}_{1}=\{u_{\frac{7}{2}}+a_{1}x_{1}u_{\frac{5}{2}}+a_{2}(x_{1}^{2}-\frac{1}{5}r^{2})u_{\frac{3}{2}}\text{ | }0\leq a_{2}\leq 5\text{, }a_{1}^{2}\leq\Gamma(a_{2})\},

where

Γ(a2):=min{4(a2(115a2)2425a2(72310a3))}\Gamma(a_{2}):=\text{min}\{4(a_{2}(1-\frac{1}{5}a_{2})\text{, }\frac{24}{25}a_{2}(\frac{7}{2}-\frac{3}{10}a_{3}))\}\text{, }
(r,θ) is the polar coordinate to (x2,x3) - plane.(r,\theta)\text{ is the polar coordinate to }(x_{2},x_{3})\text{ - plane}.

Savin-Yu [9] proved that all 72\frac{7}{2}-homogeneous solutions that are close to 1\mathcal{F}_{1} actually lies in 1\mathcal{F}_{1} (up to normalization). Then they characterized the rate of convergence of blow-up and proved the regularity of the contact set Λ72\Lambda_{\frac{7}{2}} which consists of all points in Λ\Lambda with frequency 72\frac{7}{2}. The main idea of their proof is to use the homogeneity of uu and reduce thin obstacle problem in B1B_{1} to 𝕊2\mathbb{S}^{2}. Since all the difficulties in the thin obstacle problem occur near the free boundary, we need to consider thin obstacle problem around 𝕊2{r=0}\mathbb{S}^{2}\cap\{r=0\}, that is, near east and west poles. Direct calculation implies that the corresponding solutions on these opposite spherical caps should satisfy a kind of (almost) symmetry. At the infinitesimal level, it reduces to the problem in 2D with data at infinity and the corresponding two solutions should be (almost) symmetric.

Similar result for general half-space (2k12)(2k-\frac{1}{2})-homogeneous solutions remains unknown. But motivated by Savin-Yu’s method [9], it is clear from above that if we want to study half-space (2k12)(2k-\frac{1}{2})-homogeneous solutions for k2k\geq 2, we need to consider the solution to the 2D thin obstacle problem with data at infinity that is close to a linear combination of ul12u_{l-\frac{1}{2}}, l=1,2,,2k,l=1,2,...,2k, which is the main purpose of this paper.

As a building block, we first prove the existence and uniqueness of 2D thin obstacle problem. The uniqueness is shown by maximum principle. To prove the existence, we construct a barrier function and meanwhile, we estimate finer expansion of the solution uu. With the building block, we find the Fourier coefficients vanish on sufficiently large circles. The corresponding statement on small spherical caps imply the boundedness of the solution. For the case of frequency 72\frac{7}{2}, see Appendix A in Savin-Yu [9].

Then we characterize a pair of symmetric solutions to the 2D thin obstacle problem. We first construct two auxiliary odd symmetric polynomials to indicate spt(Δu\Delta u), the support of Δu\Delta u as a measure. Using these polynomials, we argue by contradiction to prove that spt(Δu\Delta u) has only one connected component in \mathbb{R}, that is to say, uu is a half-space solution. In Savin-Yu [9], they also argued by contradiction but they carefully counted the number of each zero of these polynomials to get a contradiction with the degree of these polynomials. However, not all zeros can be found in the general case. As a consequence, their method does not work in general. Instead, we focus on the odevity of the multiplicities of certain zeros to obtain a contradiction, which does not have to take all zeros into account. This is the main improvement of our work. After that, we finish our characterization and extend it to the case of almost symmetric solutions, which is truly needed for the spherical case.

This paper is organized as follows:

In Section 2, we give detailed statement of 2D thin obstacle problem with data at infinity and main results. In Section 3, some preliminaries for the paper are introduced. In Section 4, we prove the existence and uniqueness of the problem. Moreover, we show that the Fourier coefficients of the solution on large circles vanish. In Section 5, we characterize a pair of symmetric solutions and extend it to almost symmetric case.

2. Main Results

To classify two blow-up solutions around opposite spherical caps, Savin-Yu [9] constructed two polynomials and counted the number of zeros of these polynomials. However, this method does not work if we study half-space (2k12)(2k-\frac{1}{2})-homogeneous solutions for k4k\geq 4.

In this paper, we extend the results of [9] on 2D thin obstacle problem with data at infinity to the general case. To be more specific, instead of counting the number of zeros of the polynomials, we check the odevity of the multiplicities of certain zeros of these polynomials to get a contradiction, which works for general cases.

Recall the definition from (2.1). For p=u2k12+l=12k1alu2k12lp=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a_{l}u_{2k-\frac{1}{2}-l}, we study the solutions to the thin obstacle problem in 2\mathbb{R}^{2} with data pp at infinity:

(2.1) {u solves (1.1) in 2,sup2|up|<+.\begin{cases}&u\text{ solves \eqref{thin} in }\mathbb{R}^{2},\\ &\sup\limits_{\mathbb{R}^{2}}|u-p|<+\infty.\end{cases}
Remark 2.1.

Similar to the definition of u2k12u_{2k-\frac{1}{2}} in n+1\mathbb{R}^{n+1}, we can define uk12(r,θ):=rk12cos((k12)θ)u_{k-\frac{1}{2}}(r,\theta):=r^{k-\frac{1}{2}}\cos((k-\frac{1}{2})\theta) for all kk\in\mathbb{Z}, where (r,θ)(r,\theta) is the polar coordinate in (xn,xn+1)(x_{n},x_{n+1}) We say uu solves (1.1) in 2\mathbb{R}^{2} if uu solves (1.1) in BRB_{R}, for any R>0R>0.

With similar approach as in Savin-Yu [9], we prove the existence and uniqueness of (2.1) as a building block. We also show that the Fourier coefficients of the solution vanishes along big circles, which is needed to bound the solution of thin obstacle problem on spherical caps.

Theorem 2.1.

For |al|1,l=1,,2k1|a_{l}|\leq 1,\ l=1,...,2k-1, there is a unique solution uu to (2.1). Moreover, for any N>0N>0, there are bjb_{j}\in\mathbb{R}. 1jN1\leq j\leq N such that |bj|M|b_{j}|\leq M,

|u(p+j=1Nbju12j)|M|x|Nu12for all x2\biggl{|}u-\biggl{(}p+\sum_{j=1}^{N}b_{j}u_{\frac{1}{2}-j}\biggr{)}\biggr{|}\leq M|x|^{-N}u_{-\frac{1}{2}}\ \text{for all }\ x\in\mathbb{R}^{2}

and

BR[u(p+j=1Nbju12j)]cos((l12))=0\int_{\partial B_{R}}\biggl{[}u-\biggl{(}p+\sum_{j=1}^{N}b_{j}u_{\frac{1}{2}-j}\biggr{)}\biggr{]}\cdot\cos((l-\frac{1}{2}))=0

for all 1lN and RM1\leq l\leq N\text{ and }R\geq M, where M>0M>0 is a universal constant.

Remark 2.2.

We will denote bj=bj[a1,,a2k1]b_{j}=b_{j}[a_{1},...,a_{2k-1}].

Before we state our second result, for simplicity, we shall introduce some definitions. We first define a transform operator UτU_{\tau} as

(2.2) Uτ(x1,x2)=(x1+τ,x2),Uτ(f)(x)=f(Uτx).U_{\tau}(x_{1},x_{2})=(x_{1}+\tau,x_{2}),\ U_{\tau}(f)(x)=f(U_{-\tau}x).

Then we give the definitions of conjugacy and symmetry

Definition 2.1.

(i) For each pp that can be written as the form p=u2k12+l=12k1alu2k12lp=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a_{l}u_{2k-\frac{1}{2}-l}. We say qq is conjugate to pp if

q=u2k12+l=12k1(1)lalu2k12l.q=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}(-1)^{l}a_{l}u_{2k-\frac{1}{2}-l.}

(ii) We say uu and vv are symmetric if there exists a constant τ\tau such that Uτ(u)U_{\tau}(u) is conjugate to Uτ(v)U_{-\tau}(v).

Remark 2.3.

From the definition, we see that in particular, symmetric solutions are half-space solutions.

The following theorem says that under anti-symmetric conditions, two solutions to 2D thin obstacle problem with conjugate data at infinity are half-space solutions and symmetric. As stated in the introduction, if we restrict the solutions of (1.1) to two opposite spherical caps in 𝕊2\mathbb{S}^{2}, we find that the conjugacy of data at infinity and symmetry of solutions are required. For instance, for the case of k=2k=2, we can consider

v12=(x146x12r2r4)u12, and v32:=(x15+10x13r215x1r4)u32,v_{-\frac{1}{2}}=(x_{1}^{4}-6x_{1}^{2}r^{2}-r^{4})u_{-\frac{1}{2}}\text{, and }v_{-\frac{3}{2}}:=(x_{1}^{5}+10x_{1}^{3}r^{2}-15x_{1}r^{4})u_{\frac{3}{2}},

where (r,θ)(r,\theta) is understood as the polar coordinate of (x2,x3)(x_{2},x_{3}). Both of them are 72\frac{7}{2}-homogeneous functions. Roughly speaking, if we restrict them to opposite spherical caps near (±1,0,0)(\pm 1,0,0) infinitesimally, we can view them as the counterparts of u12u_{-\frac{1}{2}} and u32u_{-\frac{3}{2}} in 2\mathbb{R}^{2} and obtain the anti-symmetry conditions of bib_{i} by their odd symmetry with respect to {x1=0}\{x_{1}=0\}. Similarly, we can obtain the anti-symmetry condition for ala_{l}.

Theorem 2.2.

Given the function p=u2k12+l=12k1alu2k12lp=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a_{l}u_{2k-\frac{1}{2}-l} and let qq be the conjugate function of pp with |al|1.|a_{l}|\leq 1. Suppose that uu and vv are solutions to (2.1) with the datum pp and qq at infinity respectively. Assume bi[a1,a2,,a2k2,a2k1]=(1)i+1bi[a1,a2,,a2k2,a2k1]b_{i}[a_{1},a_{2},\dots,a_{2k-2},a_{2k-1}]=(-1)^{i+1}b_{i}[-a_{1},a_{2},\dots,a_{2k-2},-a_{2k-1}], for all 1i2k21\leq i\leq 2k-2, then uu and vv are symmetric.

Moreover, we can find universally bounded constants αl, 1l2k2\alpha_{l},\ 1\leq l\leq 2k-2 and τ\tau such that

u=Uτ(u2k12+l=12k2αlu2k12l), v=Uτ(u2k12+l=12k2(1)lαlu2k12l),u=U_{\tau}\biggl{(}u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-2}\alpha_{l}u_{2k-\frac{1}{2}-l}\biggr{)},\text{ }v=U_{-\tau}\biggl{(}u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-2}(-1)^{l}\alpha_{l}u_{2k-\frac{1}{2}-l}\biggr{)},

3. Preliminaries

In this section, we first introduce some notations, which is similar to [9]. Unless there is a special explanation, we will consider the case of 2\mathbb{R}^{2} in this paper. Denote the polar coordinates of (x1,x2)(x_{1},x_{2}) by (r,θ)(r,\theta). Then the slit is defined as

(3.1) 𝒮:={θ=π}={x10,x2=0}.\mathcal{S}:=\{\theta=\pi\}=\{x_{1}\leq 0,x_{2}=0\}.

For a subset of 2\mathbb{R}^{2}, we can decompose it as E=E^E~E=\widehat{E}\cup\widetilde{E}, where E^=E𝒮,\widehat{E}=E\setminus\mathcal{S}, and E~=E𝒮\widetilde{E}=E\cap\mathcal{S}. Given a domain Ωn+1\Omega\subset\mathbb{R}^{n+1}, a harmonic function in the slit domain Ω^\widehat{\Omega} is a continuous function that is even with respect to {xn+1=0}\{x_{n+1}=0\} and satisfies

(3.2) {Δv=0 in Ω^,v=0 in Ω~.\begin{cases}\Delta v=0&\text{ in $\widehat{\Omega}$,}\\ v=0&\text{ in $\widetilde{\Omega}$.}\end{cases}

It is easy to check uk12u_{k-\frac{1}{2}} is harmonic in 2^\widehat{\mathbb{R}^{2}}. We denote the derivatives of u2k12u_{2k-\frac{1}{2}} by the following (for simplicity, we also denote w2k12:=u2k12w_{2k-\frac{1}{2}}:=u_{2k-\frac{1}{2}}):

w2k32:=x1u2k12,wl12:=x1wl+12l,l2k1.w_{2k-\frac{3}{2}}:=\frac{\partial}{\partial x_{1}}u_{2k-\frac{1}{2}},w_{l-\frac{1}{2}}:=\frac{\partial}{\partial x_{1}}w_{l+\frac{1}{2}}\text{, }l\in\mathbb{Z},\ l\leq 2k-1.

Given a non-negative integer mm, we define the following class of (m+12)(m+\frac{1}{2})-homogeneous functions

(3.3) m+12:={v:v is a harmonic function in 2^,xv=(m+12)v}.\mathcal{H}_{m+\frac{1}{2}}:=\{v:v\text{ is a harmonic function in }\widehat{\mathbb{R}^{2}},\quad x\cdot\nabla v=(m+\frac{1}{2})v\}.

The case in dimension 22 is rather simple. Its proof follows easily by writing vm+12v\in\mathcal{H}_{m+\frac{1}{2}} in the polar coordinate. Each element of m+12\mathcal{H}_{m+\frac{1}{2}} is a multiple of um+12u_{m+\frac{1}{2}}:

m+12={aum+12 — a}.\mathcal{H}_{m+\frac{1}{2}}=\{au_{m+\frac{1}{2}}\text{ | }a\in\mathbb{R}\}.

Then we can approximate the solution of (3.2) in B1B_{1} by functions in m+12\mathcal{H}_{m+\frac{1}{2}}.

Theorem 3.1 (Theorem 4.5 from [12]).

Let vv be a solution to (3.2) with Ω=B1n\Omega=B_{1}\subset\mathbb{R}^{n} and vL(B1)1\|v\|_{L^{\infty}(B_{1})}\leq 1. Given N0N\geq 0, we can find vm+12m+12v_{m+\frac{1}{2}}\in\mathcal{H}_{m+\frac{1}{2}} for m=0,1,,Nm=0,1,\dots,N, such that

vm+12L(B1)C\|v_{m+\frac{1}{2}}\|_{L^{\infty}(B_{1})}\leq C

and

|vm=0Nvm+12|(x)C|x|N+1u12 for xB12,\biggl{|}v-\sum_{m=0}^{N}v_{m+\frac{1}{2}}\biggr{|}(x)\leq C|x|^{N+1}u_{\frac{1}{2}}\text{ for $x\in B_{\frac{1}{2}}$,}

where CC depends only on mm.

Remark 3.1.

In the 2D case, this theorem can be directly derived by using a square-root transformation and the Taylor expansion near 0.

In the following two sections, we will generalize the results of Appendix B in [9] from data p=u72+a1u52+a2u32+a3u12p=u_{\frac{7}{2}}+a_{1}u_{\frac{5}{2}}+a_{2}u_{\frac{3}{2}}+a_{3}u_{\frac{1}{2}} to the general case p=u2k12+l=12k1alu2k12lp=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a_{l}u_{2k-\frac{1}{2}-l}.

4. Existence and Uniqueness

In this section, we prove Theorem 2.1. First, we prove the existence and uniqueness of the solution to the thin obstacle problem in 2\mathbb{R}^{2} with data pp at infinity (2.1). Then as a corollary, we deduce that Fourier coefficients of the solution vanish along big circles. Recall that p=u2k12+l=12k1alu2k12lp=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a_{l}u_{2k-\frac{1}{2}-l}.

Proposition 4.1 (Uniqueness).

Let |aj|1|a_{j}|\leq 1, if there are two solutions u1u_{1}, u2u_{2} to the system (2.1), then u1=u2u_{1}=u_{2}.

Proof.

Suppose that u1u_{1} and u2u_{2} are two solutions to (1.1) in 2\mathbb{R}^{2} with sup2|ujp|<+.\sup_{\mathbb{R}^{2}}|u_{j}-p|<+\infty. We first claim that there exists a constant M>0M>0 such that

Δuj=0 in {r>M}^, and uj=0 in {r>M}~.\Delta u_{j}=0\text{ in }\widehat{\{r>M\}},\text{ and }u_{j}=0\text{ in }\widetilde{\{r>M\}}.

To derive the first part of the claim, we point out that

sup2|ujp|C\sup_{\mathbb{R}^{2}}|u_{j}-p|\leq C

However, for MM big enough, one has p>Cp>C on {x1>M,x2=0}\{x_{1}>M,x_{2}=0\}, which implies that uj>0u_{j}>0 on {x1>M,x2=0}\{x_{1}>M,x_{2}=0\} and the first part of the claim follows.

For the second part, we first briefly summarize the idea of the proof. We will construct a function φ\varphi such that it is also a solution to the thin obstacle problem and φ=0\varphi=0 at (x0,0)(x_{0},0) with some fixed x0Mx_{0}\leq-M. Moreover, we show that φp+C\varphi\geq p+C along the boundary of a neighborhood Ω\Omega of (x0,0)(x_{0},0). Then, we have ujp+Cφu_{j}\leq p+C\leq\varphi on Ω\partial\Omega, and the maximum principle implies that uj0u_{j}\leq 0 at (x0,0)(x_{0},0), which together with the definition of uju_{j} yields uj=0u_{j}=0 at (x0,0)(x_{0},0).

Now, it suffices to show that for any (x0,0)(x_{0},0) where x0Mx_{0}\leq-M , we can construct φ\varphi. First, we set the neighborhood of (x0,0)(x_{0},0) by

Ω:={(x1,x2):|x1x0|d,|x2|d}\Omega:=\{(x_{1},x_{2}):|x_{1}-x_{0}|\leq d,|x_{2}|\leq d\}

with dd to be determined latter. We now define

φ(x1,x2):=(|x1x0|2|x2|2)/d3.\varphi(x_{1},x_{2}):=(|x_{1}-x_{0}|^{2}-|x_{2}|^{2})/d^{3}.

It follows directly that φ\varphi is a global solution to the thin obstacle problem. and now we would show

φp+ConΩ.\varphi\geq p+C\hskip 5.0pton\hskip 5.0pt\partial\Omega.

By even symmetry, it suffices to show it holds on Ω{x20}\partial\Omega\cap\{x_{2}\geq 0\}. For that, we point out that

x2pc|x0|2k12\partial_{x_{2}}p\leq-c|x_{0}|^{2k-\frac{1}{2}}

for MM big enough, where cc is a positive universal constant. Inside Ω+=Ω¯{x20}\Omega^{+}=\overline{\Omega}\cap\{x_{2}\geq 0\}, one has

φp\displaystyle\varphi-p (|x1x0|2|x2|2)/d3+c|x0|2k12x2\displaystyle\geq(|x_{1}-x_{0}|^{2}-|x_{2}|^{2})/d^{3}+c|x_{0}|^{2k-\frac{1}{2}}x_{2}
(|x1x0|2)/d3+c2|x0|2k12x2\displaystyle\geq(|x_{1}-x_{0}|^{2})/d^{3}+\frac{c}{2}|x_{0}|^{2k-\frac{1}{2}}x_{2}

When |x1x0|=d , 0x2d|x_{1}-x_{0}|=d\text{ , }0\leq x_{2}\leq d, we have

φp1d+c2|x0|2k12x21dC,\varphi-p\geq\frac{1}{d}+\frac{c}{2}|x_{0}|^{2k-\frac{1}{2}}x_{2}\geq\frac{1}{d}\geq C,

if we choose dd small enough. When |x1x0|d , |x2|=d|x_{1}-x_{0}|\leq d\text{ , }|x_{2}|=d, we have

φpc2|x0|2k12x2c2dM2k12C,\varphi-p\geq\frac{c}{2}|x_{0}|^{2k-\frac{1}{2}}x_{2}\geq\frac{c}{2}dM^{2k-\frac{1}{2}}\geq C,

if MM is sufficiently large. By even symmetry, we have φpC on Ω\varphi-p\geq C\text{ on }\partial\Omega and we finish the proof of the claim. Then, we define

w(x)=(u1u2)(M2x|x|2)w(x)=(u_{1}-u_{2})\biggl{(}\frac{M^{2}x}{|x|^{2}}\biggr{)}

to be the Kelvin transform of (u1u2)(u_{1}-u_{2}) with respect to BM\partial B_{M}. Then ww is a bounded harmonic function in BM^\widehat{B_{M}}. By Theorem 3.1 with N=0N=0, we have |w|Cu12|w|\leq Cu_{\frac{1}{2}}, which implies

|u1u2|Cu12in2|u_{1}-u_{2}|\leq Cu_{-\frac{1}{2}}\hskip 4.0ptin\hskip 4.0pt\mathbb{R}^{2}

by inverting Kelvin transform. By maximum principle, we conclude that u1=u2u_{1}=u_{2}. ∎

Next we construct a barrier function which will be used to prove the existence.

Lemma 4.1 (Barrier function).

Let |aj|1|a_{j}|\leq 1 for any 1j2k11\leq j\leq 2k-1. Then there is a solution QQ to (1.1) such that

Qp on {rM}Q\geq p\text{ on }\{r\geq M\}

for a universally large constant MM.

Proof.

Rewrite pp in the basis {u2k12,,w12}\{u_{2k-\frac{1}{2}},\dots,w_{\frac{1}{2}}\} as p=u2k12+l=12k1a~lu2k12lp=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}\tilde{a}_{l}u_{2k-\frac{1}{2}-l}. For τ>0\tau>0 to be chosen, we denote (α1,α2,,α2k2)(\alpha_{1},\alpha_{2},...,\alpha_{2k-2}) to be the solution to the system

j=0i1j!αijτj=a~i, 1i2k2\sum_{j=0}^{i}\frac{1}{j!}\alpha_{i-j}\tau^{j}=\tilde{a}_{i},\text{ }1\leq i\leq 2k-2

Here we take α0=1\alpha_{0}=1. For example, when k=3k=3, it’s

α1+τ=a~1,α2+α1τ+12τ2=a~2,α3+α2τ+12α1τ2+16τ3=a~3,\displaystyle\alpha_{1}+\tau=\tilde{a}_{1},\ \alpha_{2}+\alpha_{1}\tau+\frac{1}{2}\tau^{2}=\tilde{a}_{2},\ \alpha_{3}+\alpha_{2}\tau+\frac{1}{2}\alpha_{1}\tau^{2}+\frac{1}{6}\tau^{3}=\tilde{a}_{3},
α4+α3τ+12α2τ2+16α1τ3+124τ4=a~4.\displaystyle\alpha_{4}+\alpha_{3}\tau+\frac{1}{2}\alpha_{2}\tau^{2}+\frac{1}{6}\alpha_{1}\tau^{3}+\frac{1}{24}\tau^{4}=\tilde{a}_{4}.

We define

q=u2k12+l=12k2αlu2k12l.q=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-2}\alpha_{l}u_{2k-\frac{1}{2}-l}.

Then Taylor’s Theorem gives

Uτ(q)p(1(2k1)!τ2k1+l=12k2αl(2k1l)!τ2k1la~2k1)w12Cτ2kw12.U_{-\tau}(q)-p\geq\biggl{(}\frac{1}{(2k-1)!}\tau^{2k-1}+\sum_{l=1}^{2k-2}\frac{\alpha_{l}}{(2k-1-l)!}\tau^{2k-1-l}-\tilde{a}_{2k-1}\biggr{)}w_{-\frac{1}{2}}-C\tau^{2k}w_{-\frac{1}{2}}.

Choosing τ\tau large universally and invoking the definition of UτU_{\tau} from (2.2), then

Uτ(q)p0 on {rM}U_{-\tau}(q)-p\geq 0\text{ on }\{r\geq M\}

for a universally large MM. By choosing τ\tau larger, if necessary, it is straightforward to verify that (α1,,α2k2)(\alpha_{1},\dots,\alpha_{2k-2}) satisfies the condition which allows qq to solve the thin obstacle problem in 2\mathbb{R}^{2}, and consequently, Q:=UτqQ:=U_{-\tau}q solves the thin obstacle problem in 2.\mathbb{R}^{2}.

Then we can prove the existence of the 2D thin obstacle problem with data pp at infinity.

Proposition 4.2.

Let |aj|1|a_{j}|\leq 1, for any 1j2k11\leq j\leq 2k-1. Then there is a unique solution uu to the 2D thin obstacle problem with data pp at infinity. And there is a universal constant M>0M>0 such that

sup2uM,Δu=0 in {r>M}^, and u=0 on {r>M}~.\sup_{\mathbb{R}^{2}}u\leq M,\ \Delta u=0\text{ in }\widehat{\{r>M\}},\text{ and }u=0\text{ on }\widetilde{\{r>M\}}.

Moreover, for any NN, we can find b1,,bNb_{1},...,b_{N} satisfying |bj|M|b_{j}|\leq M such that

|u(p+j=1Nbju12j)|M|x|Nu12for all x2.\biggl{|}u-\biggl{(}p+\sum_{j=1}^{N}b_{j}u_{\frac{1}{2}-j}\biggr{)}\biggr{|}\leq M|x|^{-N}u_{-\frac{1}{2}}\ \text{for all }\ x\in\mathbb{R}^{2}.
Proof.

For large nn\in\mathbb{N}, let unu_{n} be the solution to the thin obstacle problem (1.1) in BnB_{n} with un=pu_{n}=p along Bn\partial B_{n}. By the maximum principle, we have unp in Bnu_{n}\geq p\text{ in }B_{n}. Also, since QQ and unu_{n} are solutions to the thin obstacle problem in BnB_{n}, we have

Δ(Qun)0 in Bn and (Qun)=0 on Bn,\Delta(Q-u_{n})^{-}\geq 0\text{ in }B_{n}\text{ and }(Q-u_{n})^{-}=0\text{ on }\partial B_{n},

which implies that (Qun)0 in Bn(Q-u_{n})^{-}\leq 0\text{ in }B_{n} i.e. unQ in Bnu_{n}\leq Q\text{ in }B_{n}. Thus we have

(4.1) punQ in Bnp\leq u_{n}\leq Q\text{ in }B_{n}

if nn is large. This gives us a locally uniformly bounded family {un}\{u_{n}\}. Then up to subsequence, it converges to some uu_{\infty} locally uniformly on 2\mathbb{R}^{2}, which is a solution to the thin obstacle problem in 2\mathbb{R}^{2}.

From (4.1), we have un=0 in Bn{x1M,x2=0}u_{n}=0\text{ in }B_{n}\cap\{x_{1}\leq-M,x_{2}=0\} and un1 in Bn{x1M,x2=0}u_{n}\geq 1\text{ in }B_{n}\cap\{x_{1}\geq M,x_{2}=0\} for a universal M>0M>0. Thus we have

Δu=0 in {r>M}^; and u=0 in {r>M}~.\Delta u_{\infty}=0\text{ in }\widehat{\{r>M\}};\text{ and }u_{\infty}=0\text{ in }\widetilde{\{r>M\}}.

On {r=M}\{r=M\}, we have 0upQpC0\leq u_{\infty}-p\leq Q-p\leq C. Thus the maximum principle, applied to the domain {r>M}\{r>M\}, gives

|up|C|u_{\infty}-p|\leq C

for a universal constant CC. In particular, uu_{\infty} is the unique solution to (2.1), according to Proposition 4.1. For simplicity, we then denote uu_{\infty} by uu.

Next we deduce a refined expansion of the solution uu. Let w(x):=(up)(M2x|x|2)w(x):=(u-p)(\frac{M^{2}x}{|x|^{2}}) be the Kelvin transform of upu-p with respect to BM\partial B_{M}. Results from previous steps imply that ww is a harmonic function in the slit domain BM^\widehat{B_{M}}. Applying Theorem 3.1 with N1N-1, we get universally bounded bj,j=1,,N,b_{j},\ j=1,...,N, such that

|wj=1Nbjuj12|C|x|Nu12 in BM.\biggl{|}w-\sum_{j=1}^{N}b_{j}u_{j-\frac{1}{2}}\biggr{|}\leq C|x|^{N}u_{\frac{1}{2}}\text{ in }B_{M}.

We immediately have

|wj=1Nbju12j|C|x|Nu12 in 2.\biggl{|}w-\sum_{j=1}^{N}b_{j}u_{\frac{1}{2}-j}\biggr{|}\leq C|x|^{-N}u_{-\frac{1}{2}}\text{ in }\mathbb{R}^{2}.

by inverting the Kelvin transform and the proof is completed. ∎

As a corollary, for the solution from the previous proposition, we show that its Fourier coefficients along big circles vanish:

Corollary 4.1.

For the solution uu obtained by Proposition 4.2, we have

BR[u(p+j=1Nbju12j)]cos((i12)θ)=0\int_{\partial B_{R}}\biggl{[}u-\biggl{(}p+\sum_{j=1}^{N}b_{j}u_{\frac{1}{2}-j}\biggr{)}\biggr{]}\cdot\cos((i-\frac{1}{2})\theta)=0

for all 1iN and RM.1\leq i\leq N\text{ and }R\geq M.

Proof.

For simplicity, we denote the extended data by

pext:=(p+j=1Nbju12j).p_{ext}:=\biggl{(}p+\sum_{j=1}^{N}b_{j}u_{\frac{1}{2}-j}\biggr{)}.

From Proposition 4.2, we have Δ(upext)=0 in {r>M}^, and upext=0 on {r>M}~.\Delta(u-p_{ext})=0\text{ in }\widehat{\{r>M\}},\text{ and }u-p_{ext}=0\text{ on }\widetilde{\{r>M\}}. For R>MR>M and each ii and 1li1\leq l\leq i fixed, define v:=(r2l12R2l1r2l12)cos(2l12θ).v:=(r^{\frac{2l-1}{2}}-R^{2l-1}r^{-\frac{2l-1}{2}})\cos(\frac{2l-1}{2}\theta). Then vv is a harmonic function in the slit domain 2\mathbb{R}^{2}. For any L>RL>R, by integration by parts, we have

0=BLBR(upext)ΔvΔ(upext)v=(BLBR)(upext)νv(upext)vν.0=\int_{B_{L}\setminus B_{R}}(u-p_{ext})\cdot\Delta v-\Delta(u-p_{ext})\cdot v=\int_{\partial(B_{L}\setminus B_{R})}(u-p_{ext})_{\nu}\cdot v-(u-p_{ext})\cdot v_{\nu}.

The asymptotic expansion of upextu-p_{ext} yields that |upext|=O(LN12)|u-p_{ext}|=O(L^{-N-\frac{1}{2}}), |(upext)ν|=O(LN32)|(u-p_{ext})_{\nu}|=O(L^{-N-\frac{3}{2}}), |v|=O(L2l12)|v|=O(L^{\frac{2l-1}{2}}) and |vν|=O(L2l32)|v_{\nu}|=O(L^{\frac{2l-3}{2}}) on BL\partial B_{L} as LL\to\infty. As a result,

BL(upext)νv(upext)vν=O(LN+l1).\int_{\partial B_{L}}(u-p_{ext})_{\nu}\cdot v-(u-p_{ext})\cdot v_{\nu}=O(L^{-N+l-1}).

On BR\partial B_{R}, we have v=0v=0 and vν=(2l1)R2l32cos(2l12θ)v_{\nu}=-(2l-1)R^{\frac{2l-3}{2}}\cos(\frac{2l-1}{2}\theta). Thus

BR(upext)cos(2l12θ)=O(R32l2LN+l1).\int_{\partial B_{R}}(u-p_{ext})\cdot\cos(\frac{2l-1}{2}\theta)=O(R^{\frac{3-2l}{2}}L^{-N+l-1}).

Sending LL\to\infty and then we finish our proof. ∎

Combining all of above, we complete the proof of Theorem 2.1

5. Characterization of Symmetric Solutions

In this section, we give the proof of Theorem 2.2 to characterize symmetric solutions and also show its perturbation version. We will first construct two important polynomials. Then we will use these auxiliary polynomials to show that uu and vv are actually half-space solutions to 2D thin obstacle problems with data pp and qq at infinity respectively. In the end, by applying Theorem 3.1, we can fully characterize uu and vv. We begin with the construction of two auxiliary polynomials as in Savin-Yu [9].

Lemma 5.1.

With the same assumption as in Theorem 2.2,

P(t):=Re(x1uix2u)2(t,0) , Q(t):=Re(x1vix2v)2(t,0)P(t):=\operatorname{Re}(\partial_{x_{1}}u-i\partial_{x_{2}}u)^{2}(t,0)\text{ , }Q(t):=\operatorname{Re}(\partial_{x_{1}}v-i\partial_{x_{2}}v)^{2}(t,0)

are two polynomials with degree 4k34k-3. Moreover, we have

(5.1) P(t)=Q(t).P(t)=-Q(-t).
Proof.

Recall that for simplicity, we define bj=bj[a1,a2,,a2k+1]b_{j}=b_{j}[a_{1},a_{2},...,a_{2k+1}] for j=1,,2k2j=1,...,2k-2, and

pext:=p+i=12k2bju12j, and qext:=q+i=12k2(1)jbju12j.p_{ext}:=p+\sum_{i=1}^{2k-2}b_{j}u_{\frac{1}{2}-j},\text{ and }q_{ext}:=q+\sum_{i=1}^{2k-2}(-1)^{j}b_{j}u_{\frac{1}{2}-j}.

Proposition 4.2 shows that

|upext|+|vqext|2M|x|2k12 in 2.|u-p_{ext}|+|v-q_{ext}|\leq 2M|x|^{-2k-\frac{1}{2}}\text{ in $\mathbb{R}^{2}$.}

Then it follows that

(5.2) |upext|+|vqext|C|x|2k32 for |x|1.|\nabla u-\nabla p_{ext}|+|\nabla v-\nabla q_{ext}|\leq C|x|^{-2k-\frac{3}{2}}\text{ for }|x|\geq 1.

Note that uu is an entire solution to the thin obstacle problem of order O(|x|2k12)O(|x|^{2k-\frac{1}{2}}) at infinity. Also note that the real part (x1u)2(x2u)2(\partial_{x_{1}}u)^{2}-(\partial_{x_{2}}u)^{2} and the imaginary part 2x1ux2u-2\partial_{x_{1}}u\partial_{x_{2}}u of (x1uix2u)2(\partial_{x_{1}}u-i\partial_{x_{2}}u)^{2} are both continuous in 2\mathbb{R}^{2} and harmonic outside {x2=0}\{x_{2}=0\}. Thus they are harmonic functions in 2\mathbb{R}^{2} and so is (x1uix2u)2(\partial_{x_{1}}u-i\partial_{x_{2}}u)^{2}. Then by Liouville theorem, (x1uix2u)2(\partial_{x_{1}}u-i\partial_{x_{2}}u)^{2} is a polynomial of degree 4k34k-3. By direct computation, we have

(x1pextix2pext)2=𝒫(x1+ix2)+j=14k3j(x1+ix2),(\partial_{x_{1}}p_{ext}-i\partial_{x_{2}}p_{ext})^{2}=\mathcal{P}(x_{1}+ix_{2})+\sum_{j=1}^{4k-3}\mathcal{R}_{j}(x_{1}+ix_{2}),

where 𝒫\mathcal{P} is a complex polynomial of degree 4k34k-3, and each j\mathcal{R}_{j} is a (j)(-j)-homogeneous rational function for j=1,2,,4k3j=1,2,\dots,4k-3.

With (5.2), it follows that

(x1uix2u)2=𝒫 in 2.(\partial_{x_{1}}u-i\partial_{x_{2}}u)^{2}=\mathcal{P}\text{ in }\mathbb{R}^{2}.

If we define

P(t):=Re(x1uix2u)2(t,0)=[(x1u)2(x2u)2](t,0)=Re𝒫(t),P(t):=\operatorname{Re}(\partial_{x_{1}}u-i\partial_{x_{2}}u)^{2}(t,0)=[(\partial_{x_{1}}u)^{2}-(\partial_{x_{2}}u)^{2}](t,0)=\operatorname{Re}\mathcal{P}(t),

then PP is a real polynomial. Similarly, corresponding to vv and qextq_{ext}, we have

(x1qextix2qext)2=𝒬(x1+ix2)+j=14k3𝒮j(x1+ix2),(\partial_{x_{1}}q_{ext}-i\partial_{x_{2}}q_{ext})^{2}=\mathcal{Q}(x_{1}+ix_{2})+\sum_{j=1}^{4k-3}\mathcal{S}_{j}(x_{1}+ix_{2}),

where 𝒬\mathcal{Q} is also a complex polynomial of degree 4k34k-3, and 𝒮j\mathcal{S}_{j} is a (j)(-j)-homogeneous rational function for j=1,2,,4k3j=1,2,\dots,4k-3. Moreover, we have

Q(t):=Re(x1vix2v)2(t,0)=[(x1v)2(x2v)2](t,0)=Re𝒬(t),Q(t):=\operatorname{Re}(\partial_{x_{1}}v-i\partial_{x_{2}}v)^{2}(t,0)=[(\partial_{x_{1}}v)^{2}-(\partial_{x_{2}}v)^{2}](t,0)=\operatorname{Re}\mathcal{Q}(t),

which is also a real polynomial. With the symmetry condition bi[a1,a2,,a2k+1]=(1)ibi[a1,a2,,a2k+1]b_{i}[a_{1},a_{2},...,a_{2k+1}]=(-1)^{i}b_{i}[-a_{1},a_{2},...,-a_{2k+1}], it just follows from direct computation that

(5.3) P(t)=Q(t).P(t)=-Q(-t).

With two auxiliary polynomials, we can show that uu and vv are in fact half-space solutions to 2.1. Its proof is the key difference between our method and Savin-Yu’s method in [9].

Proposition 5.1.

With the same assumption as Theorem 2.2, we have uu and vv are half-space solutions to (2.1).

Proof.

With (5.1), we show that up to a translation, uu must be a half-space solution. Since u=0u=0 on {r>M}~\widetilde{\{r>M\}} according to Proposition 2.1, it suffices to show that spt(Δu)\text{spt}(\Delta u) has only one component.

Suppose, on the contrary, that

(,a][b,+)spt(Δu)(,a][b,c] with b>a,(-\infty,a]\cup[b,+\infty)\supset\text{spt}(\Delta u)\supset(-\infty,a]\cup[b,c]\text{ with }b>a,

Here spt(Δu\Delta u) is the support of Δu\Delta u as a measure. For simplicity, we omit the second coordinate component since it is always 0 in the proof. Note that the second component has to terminate in finite length since Δu=0\Delta u=0 in {r>M}^\widehat{\{r>M\}}.

In (,a][b,c](-\infty,a]\cup[b,c], we have x1u=0\partial_{x_{1}}u=0. Thus P(t)=(x2u)20P(t)=-(\partial_{x_{2}}u)^{2}\leq 0 for t(,a][b,c]t\in(-\infty,a]\cup[b,c]. On the other hand, in (a,b)(a,b), x2u=0\partial_{x_{2}}u=0 and P(t)=(x1u)20P(t)=(\partial_{x_{1}}u)^{2}\geq 0. Moreover, since u(a)=u(b)=0u(a)=u(b)=0, u0u\geq 0 in (a,b)(a,b). we must have x1u(d)=0\partial_{x_{1}}u(d)=0 at some point d(a,b)d\in(a,b) and there exists ϵ>0\epsilon>0 such that x1u>0 on (dϵ,d)\partial_{x_{1}}u>0\text{ on }(d-\epsilon,d) and x1u<0 on (d,d+ϵ)\partial_{x_{1}}u<0\text{ on }(d,d+\epsilon). Note that uu only has at most finitely many zeros on (a,b)(a,b). Otherwise, there are infinitely many zeros of x1u\partial_{x_{1}}u and thus of P(t)P(t), which is impossible. So such dd and ϵ\epsilon exist. Thus dd is a zero of x1u\partial_{x_{1}}u of odd multiplicity λ\lambda. So dd is a root of PP of multiplicity 2λ2\lambda.

With the symmetry described in (5.1), we have Q<0Q<0 on (b,a)(-b,-a) except finitely many zeros. This implies v=0v=0 on (b,a)(-b,-a) except finitely many points. By the continuity of vv, we conclude that vv vanishes identically on (b,a)(-b,-a). So x1v=0 on (b,a)\partial_{x_{1}}v=0\text{ on }(-b,-a) and Q=(x2v)2Q=-(\partial_{x_{2}}v)^{2}.

Using the symmetry (5.1) again, we have d-d is a zero of QQ and x2v\partial_{x_{2}}v. Since x2v\partial_{x_{2}}v stays non-positive on (b,a)(-b,-a), d-d is a zero of x2v\partial_{x_{2}}v of even multiplicity μ\mu and thus a zero of QQ of multiplicity 2μ2\mu, which is a contradiction with the symmetry.

As a result, spt(Δu)\text{spt}(\Delta u) must be a half line. A similar result holds for spt(Δv).\text{spt}(\Delta v). With (5.1), we see that if spt(Δu)=(,a]\text{spt}(\Delta u)=(-\infty,a], then spt(Δv)=(,a].\text{spt}(\Delta v)=(-\infty,-a].

Remark 5.1.

We should point out that if we count the number of zeros of P(t)P(t) and Q(t)Q(t), we can only prove the case of pp with leading term u72u_{\frac{7}{2}} as in Savin-Yu [9] and u112u_{\frac{11}{2}} (with slight modifications). However, this method does not work for general case since we cannot find all zeros of P(t)P(t) and Q(t)Q(t). Instead, we check the multiplicity of the zero dd of P(t)P(t) to get a contradiction, which is the main advantage of our paper. It is independent of the leading term.

Finally, we can use the Lemma 5.1 and Proposition 5.1 above to prove Theorem 2.2.

Proof of Theorem 2.2.

With the help of Proposition 5.1, we can apply Theorem 3.1 to get

Uau=a0u2k12+l=12k2αlu2k12i+a2k+1u12.U_{-a}u=a_{0}^{\prime}u_{2k-\frac{1}{2}}+\sum_{l=1}^{2k-2}\alpha_{l}u_{2k-\frac{1}{2}-i}+a_{2k+1}^{\prime}u_{\frac{1}{2}}.

Since sup2|up|<+\sup\limits_{\mathbb{R}^{2}}|u-p|<+\infty, we have a5=0a_{5}^{\prime}=0 . The bound of |u(u2k12+l=12k2alu2k12l)|\biggl{|}u-\biggl{(}u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-2}a_{l}u_{2k-\frac{1}{2}-l}\biggr{)}\biggr{|} in 2\mathbb{R}^{2} yields a0=1.a_{0}^{\prime}=1. Therefore,

Uau=u2k12+l=12k2αlu2k12l.U_{-a}u=u_{2k-\frac{1}{2}}+\sum_{l=1}^{2k-2}\alpha_{l}u_{2k-\frac{1}{2}-l}.

Similarly, we have

Uav=u2k12+l=12k2βlu2k12l.U_{a}v=u_{2k-\frac{1}{2}}+\sum_{l=1}^{2k-2}\beta_{l}u_{2k-\frac{1}{2}-l}.

Thus uu and vv are symmetric. Finally, we can apply (5.1) and obtain that αl=(1)l+1βl\alpha_{l}=(-1)^{l+1}\beta_{l}. This finishes our proof of Theorem 2.2. ∎

With Theorem 2.2, we obtain the following corollary, which allows us to perturb the symmetry condition

bi[a1,a2,,a2k2,a2k1]=(1)i+1bi[a1,a2,,a2k2,a2k1],b_{i}[a_{1},a_{2},\dots,a_{2k-2},a_{2k-1}]=(-1)^{i+1}b_{i}[-a_{1},a_{2},\dots,a_{2k-2},-a_{2k-1}],

for all 1i2k21\leq i\leq 2k-2, to almost-symmetric case. It is what we truly need when studying half-space (2k12)(2k-\frac{1}{2})-homogeneous solutions. Its proof is almost the same as the Corollary B.2 of [9].

Corollary 5.1.

Given p=u2k12+l=12k1alu2k12l=u2k12+l=12k1a~lw2k12lp=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a_{l}u_{2k-\frac{1}{2}-l}=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}\tilde{a}_{l}w_{2k-\frac{1}{2}-l} with |aj|1|a_{j}|\leq 1, we set

bj+:=bj[a1,a2,,a2k2,a2k1]bj:=bj[a1,a2,,a2k2,a2k1]b_{j}^{+}:=b_{j}[a_{1},a_{2},\dots,a_{2k-2},a_{2k-1}]\text{, }b_{j}^{-}:=b_{j}[-a_{1},a_{2},\dots,a_{2k-2},-a_{2k-1}]

for j=1,2,,2k1,j=1,2,...,2k-1, and

pext:=p+j=12k2bju12j=p+j=12k2b~jw12j.p_{ext}:=p+\sum_{j=1}^{2k-2}b_{j}u_{\frac{1}{2}-j}=p+\sum_{j=1}^{2k-2}\tilde{b}_{j}w_{\frac{1}{2}-j}.

Then there is a universal modulus of continuity ω\omega such that

l=12k2|a~lj=0lαljj!τj|+|a~2k1j=12k1α2k1jj!τj|\displaystyle\sum_{l=1}^{2k-2}\biggl{|}\tilde{a}_{l}-\sum_{j=0}^{l}\frac{\alpha_{l-j}}{j!}\tau^{j}\biggr{|}+\biggl{|}\tilde{a}_{2k-1}-\sum_{j=1}^{2k-1}\frac{\alpha_{2k-1-j}}{j!}\tau^{j}\biggr{|}
+j=12k2|b~j+l=02k2α2k2l(l+j+1)!τl+j+1|ω(j=12k2|bj+bj|)\displaystyle+\sum_{j=1}^{2k-2}\biggl{|}\tilde{b}_{j}^{+}-\sum_{l=0}^{2k-2}\frac{\alpha_{2k-2-l}}{(l+j+1)!}\tau^{l+j+1}\biggr{|}\leq\omega(\sum_{j=1}^{2k-2}|b_{j}^{+}-b_{j}^{-}|)

for universally bounded αj\alpha_{j} (we naturally define α0=1\alpha_{0}=1) and τ\tau such that u2k12+l=12k2αlu2k12lu_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-2}\alpha_{l}u_{2k-\frac{1}{2}-l} is a solution to the thin obstacle problem in 2\mathbb{R}^{2}.

Proof.

Suppose there is no such ω\omega, then we can find a sequence (aln)(a_{l}^{n}) such that for any bounded αj\alpha_{j} and τ\tau such that u2k12+l=12k2αlu2k12lu_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-2}\alpha_{l}u_{2k-\frac{1}{2}-l} is a solution to the thin obstacle problem in 2\mathbb{R}^{2}, we have

l=12k2|a~lnj=0lαljj!τj|+|a~2k1nj=12k1α2k1jj!τj|\displaystyle\sum_{l=1}^{2k-2}\biggl{|}\tilde{a}^{n}_{l}-\sum_{j=0}^{l}\frac{\alpha_{l-j}}{j!}\tau^{j}\biggr{|}+\biggl{|}\tilde{a}^{n}_{2k-1}-\sum_{j=1}^{2k-1}\frac{\alpha_{2k-1-j}}{j!}\tau^{j}\biggr{|}
(5.4) +j=12k2|b~j+,nl=02k2α2k2l(l+j+1)!τl+j+1|ϵ>0\displaystyle+\sum_{j=1}^{2k-2}\biggl{|}\tilde{b}^{+,n}_{j}-\sum_{l=0}^{2k-2}\frac{\alpha_{2k-2-l}}{(l+j+1)!}\tau^{l+j+1}\biggr{|}\geq\epsilon>0

for some ϵ>0\epsilon>0. Meanwhile, the corresponding (bj±,n)(b_{j}^{\pm,n}) satisfy

(5.5) j=12k2|bj+,nbj,n|0,\sum_{j=1}^{2k-2}|b_{j}^{+,n}-b_{j}^{-,n}|\to 0,

as nn\to\infty. Then up to subsequence, we have

alnal and bj±,nbj±,.a_{l}^{n}\to a_{l}^{\infty}\text{ and }b_{j}^{\pm,n}\to b_{j}^{\pm,\infty}.

Set

pn+=u2k12+l=12k1alnu2k12l=u2k12+l=12k1a~lnw2k12lp_{n}^{+}=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a^{n}_{l}u_{2k-\frac{1}{2}-l}=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}\tilde{a}^{n}_{l}w_{2k-\frac{1}{2}-l}

and let un+u^{+}_{n} be the (unique) solution to (2.1) with data pn+p^{+}_{n} at infinity. Then by Proposition 4.2, we have

|un+pn+|M in 2.|u^{+}_{n}-p^{+}_{n}|\leq M\text{ in }\mathbb{R}^{2}.

Up to subsequence, we have un+u^{+}_{n} converge to u+u^{+}_{\infty} locally uniformly for some solution u+u^{+}_{\infty} to the thin obstacle problem in 2\mathbb{R}^{2}. Moreover, we have

|u+[u2k12+l=12k1alu2k12l]|M in 2.\biggl{|}u^{+}_{\infty}-\biggl{[}u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a^{\infty}_{l}u_{2k-\frac{1}{2}-l}\biggr{]}\biggr{|}\leq M\text{ in $\mathbb{R}^{2}$.}

Thus u+u^{+}_{\infty} is the solution to (2.1) with data p+=u2k12+l=12k1alnu2k12lp^{+}_{\infty}=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}a^{n}_{l}u_{2k-\frac{1}{2}-l} at infinity.

Corollary 4.1 implies that bj+,:=bj[al]=limnbj+,n.b_{j}^{+,\infty}:=b_{j}[a_{l}^{\infty}]=\lim\limits_{n\to\infty}b_{j}^{+,n}. Simliarly, for pn=u2k12+l=12k1(1)lalnu2k12lp_{n}^{-}=u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}(-1)^{l}a^{n}_{l}u_{2k-\frac{1}{2}-l}, we also have bj,:=bj[(1)lal]=limnbj,n.b_{j}^{-,\infty}:=b_{j}[(-1)^{l}a_{l}^{\infty}]=\lim\limits_{n\to\infty}b_{j}^{-,n}. Invoking (5.5), we conclude

bj[a1,a2,,a2k2,a2k1]=(1)j+1bj[a1,a2,,a2k2,a2k1]b_{j}[a_{1},a_{2},\dots,a_{2k-2},a_{2k-1}]=(-1)^{j+1}b_{j}[-a_{1},a_{2},\dots,a_{2k-2},-a_{2k-1}]

Then Proposition 5.1 yields that

u+=Uτ(u2k12+l=12k1αlu2k12l)u_{\infty}^{+}=U_{\tau}(u_{2k-\frac{1}{2}}+\sum\limits_{l=1}^{2k-1}\alpha_{l}u_{2k-\frac{1}{2}-l})

for αj\alpha_{j} such that u2k12+l=12k2αlu2k12lu_{2k-\frac{1}{2}}+\sum_{l=1}^{2k-2}\alpha_{l}u_{2k-\frac{1}{2}-l} is a solution to the thin obstacle problem in 2\mathbb{R}^{2}. As a consequence, we have

l=12k2|a~lj=0lαljj!τi|+|a~2k1j=12k1α2k1jj!τj|\displaystyle\sum_{l=1}^{2k-2}\biggl{|}\tilde{a}^{\infty}_{l}-\sum_{j=0}^{l}\frac{\alpha_{l-j}}{j!}\tau^{i}\biggr{|}+\biggl{|}\tilde{a}^{\infty}_{2k-1}-\sum_{j=1}^{2k-1}\frac{\alpha_{2k-1-j}}{j!}\tau^{j}\biggr{|}
+j=12k2|b~j+,l=02k2α2k2l(l+j+1)!τl+j+1|=0.\displaystyle+\sum_{j=1}^{2k-2}\biggl{|}\tilde{b}_{j}^{+,\infty}-\sum_{l=0}^{2k-2}\frac{\alpha_{2k-2-l}}{(l+j+1)!}\tau^{l+j+1}\biggr{|}=0.

Taking the limit of (5), we get a contradiction. ∎

Acknowledgment

We would like to thank Prof. Hui Yu for the helpful discussion.

References

  • [1] J. A. “Generic regularity of free boundaries for the obstacle problem” In Publications mathématiques de l’IHÉS 132.1 Springer, 2020, pp. 181–292
  • [2] Ural’tseva A. and N. Nikolaevna “Regularity of free boundaries in obstacle-type problems” American Mathematical Soc., 2012
  • [3] F. Almgren “Dirichlet’s problem for multiple valued functions and the regularity of mass minimizing integral currents” In Minimal Submanifolds and Geodesics: Proceedings of the Japan-United States Seminar on Minimal Submanifolds, Including Geodesics, Tokyo 1977 North-Holland, 1979
  • [4] L. I. “Optimal regularity of lower-dimensional obstacle problems” In Journal of Mathematical Sciences 132.3 Springer, 2006, pp. 274–284
  • [5] S. I. “The structure of the free boundary for lower dimensional obstacle problems” In American journal of mathematics 130.2 Johns Hopkins University Press, 2008, pp. 485–498
  • [6] B. M. “Direct epiperimetric inequalities for the thin obstacle problem and applications” In Communications on Pure and Applied Mathematics 73.2 Wiley Online Library, 2020, pp. 384–420
  • [7] A. N. “Some new monotonicity formulas and the singular set in the lower dimensional obstacle problem” In Inventiones mathematicae 177.2 Springer, 2009, pp. 415
  • [8] David Richardson “Variational problems with thin obstacles”, 1978
  • [9] O. Savin and H. Yu “Half-space solutions with 7/27/2 frequency in the thin obstacle problem”, 2021 arXiv:2109.06399 [math.AP]
  • [10] O. Savin and H. Yu “On the fine regularity of the singular set in the nonlinear obstacle problem”, 2021 arXiv:2101.11759 [math.AP]
  • [11] A. Signorini “Questioni di elasticità non linearizzata e semilinearizzata” In Rend. Mat. Appl 18.5, 1959, pp. 95–139
  • [12] D. Silva and O. Savin CC^{\infty} Regularity of Certain Thin Free Boundaries” In Indiana University Mathematics Journal 64.5 Indiana University Mathematics Department, 2015, pp. 1575–1608
  • [13] N. Ural’tseva “Regularity of solutions of variational inequalities” In Russian Mathematical Surveys 42.6 IOP Publishing, 1987, pp. 191