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

On the global behaviors for defocusing semilinear wave equations in 1+2\mathbb{R}^{1+2}

Dongyi Wei    Shiwu Yang
Abstract

In this paper, we study the asymptotic decay properties for defocusing semilinear wave equations in 1+2\mathbb{R}^{1+2} with pure power nonlinearity. By applying new vector fields to null hyperplane, we derive improved time decay of the potential energy, with a consequence that the solution scatters both in the critical Sobolev space and energy space for all p>1+8p>1+\sqrt{8}. Moreover combined with Brézis-Gallouet-Wainger type of logarithmic Sobolev embedding, we show that the solution decays pointwise with sharp rate t12t^{-\frac{1}{2}} when p>113p>\frac{11}{3} and with rate tp18+ϵt^{-\frac{p-1}{8}+\epsilon} for all 1<p1131<p\leq\frac{11}{3}. This in particular implies that the solution scatters in energy space when p>251p>2\sqrt{5}-1.

1 Introduction

Consider the Cauchy problem to the defocusing semilinear wave equation

ϕ=t2ϕ+Δϕ=|ϕ|p1ϕ,ϕ(0,x)=ϕ0(x),tϕ(0,x)=ϕ1(x)\Box\phi=-\partial_{t}^{2}\phi+\Delta\phi=|\phi|^{p-1}\phi,\quad\phi(0,x)=\phi_{0}(x),\quad\partial_{t}\phi(0,x)=\phi_{1}(x) (1)

in 1+2\mathbb{R}^{1+2} with 1<p<1<p<\infty. It is well known that this equation is energy subcritical and is local well-posed in energy space, see for example [7]. The energy conservation then immediately implies that the solution is indeed globally in time but without any asymptotic dynamics.

Due to the repulsive nature of the equation, it is expected that the solution should decay in some sense. It has been shown in the early work [9] of Glassey-Pecher that the solution decays as fast as linear solutions in the pointwise sense for the super-conformal case p>5p>5. The solution still decays as long as 1+8<p51+\sqrt{8}<p\leq 5 but with slower decay rate in time. As a consequence of these pointwise decay estimates, they also demonstrated that the solution scatters in energy space when p>4.15p>4.15.

While the above scattering result (asymptotic completeness) is measured in energy space and heavily relies on the smoothness of the data, Ginibre-Velo in [8] established a complete scattering theory (asymptotic completeness and existence of wave operator) with data in conformal energy space for the super-conformal case p>5p>5. However it is not clear whether such scattering theory can be extended beyond the conformal invariant power p=5p=5. For a more comprehensive discussion on the global dynamics for defocusing semilinear wave equation in general dimension, we refer the interested readers to [10], [16] and references therein.

Both of the above mentioned results (indeed all the previous related works) relied on the approximate conservation of conformal energy derived by using the conformal Killing vector field as multiplier. Since the conformal Killing vector field is order 22 with weights t2t^{2}, an effective estimate requires that the power pp can not be too small. A possible way to study the asymptotic behaviors for smaller pp is to use vector fields with less weights. This comes in the robust new vector field method introduced by Dafermos-Rodnianski in [3], which has been successfully applied to study the asymptotic decay properties for the energy subcritical defocusing semilinear wave equations in higher dimensions [16]. However this method fails for equation in dimension 22 (see detailed discussion in the end of the introduction).

With drawback of the aforementioned methods to investigate the global dynamics of equation (1) in 1+2\mathbb{R}^{1+2}, this paper is devoted to developing new approach to study linear and nonlinear wave equations in dimension 22. To derive weighted energy estimates with order less than 22, we apply new vector fields to regions bounded by null hyperplanes instead of null cones used in higher dimensions. The use of null hyperplane can be viewed as reduction of the equation to 1+1\mathbb{R}^{1+1}, on which the wave operator is invariant under a larger class of conformal symmetry (see [12]). This enables us to use vector fields of order p12\frac{p-1}{2} for all 1<p<51<p<5.

Another difficulty to study the decay of linear waves in 1+2\mathbb{R}^{1+2} is the failure of the embedding LH1L^{\infty}\hookrightarrow H^{1}. In space dimension 33, the work [17] made use of the weighted potential energy flux through backward light cone, which is sufficiently strong to control the nonlinearity. As pointed out above, such estimate is absent in dimension 22. This suggests us to explore the weighted energy estimate for the derivatives of the solution in addition to the potential energy decay. Since a priori the H2H^{2} norm of solution to (1) grows at most polynomially in time tt, we make use of Brézis-Gallouet-Wainger inequality (logarithmic Sobolev embedding) to show that the solution decays in time tt for all 1<p<51<p<5.

To state our main theorems, recall from [16] that the critical exponent

sp=p3p1s_{p}=\frac{p-3}{p-1}

for Sobolev space. In particular, equation (1) is energy subcritical for all 1<p<1<p<\infty, and is conformal invariant when p=5p=5 (corresponding to sp=12s_{p}=\frac{1}{2}). We also recall the linear operator 𝐋(t)\mathbf{L}(t)

𝐋(t)(ϕ0(x),ϕ1(x))=(ϕ(t,x),tϕ(t,x))\mathbf{L}(t)(\phi_{0}(x),\phi_{1}(x))=(\phi(t,x),\partial_{t}\phi(t,x))

with ϕ\phi solving the linear wave equation ϕ=0\Box\phi=0, ϕ(0,x)=ϕ0\phi(0,x)=\phi_{0}, tϕ(0,x)=ϕ1\partial_{t}\phi(0,x)=\phi_{1} on 1+2\mathbb{R}^{1+2}.

For constant 0γ20\leq\gamma\leq 2 define the weighted energy norm of the initial data

k,γ=lk2(1+|x|)γ+2l(|l+1ϕ0|2+|lϕ1|2)+(1+|x|)γ|ϕ0|p+1dx.\displaystyle\mathcal{E}_{k,\gamma}=\sum\limits_{l\leq k}\int_{\mathbb{R}^{2}}(1+|x|)^{\gamma+2l}(|\nabla^{l+1}\phi_{0}|^{2}+|\nabla^{l}\phi_{1}|^{2})+(1+|x|)^{\gamma}|\phi_{0}|^{p+1}dx.

In this paper we will only use the initial conformal energy 0,2\mathcal{E}_{0,2} and the first order energy 1,0\mathcal{E}_{1,0}.

Our first result is the time decay of the potential energy as well as the scattering results in critical Sobolev space.

Theorem 1.

For solution ϕ\phi to the defocusing semilinear wave equation (1) in 1+2\mathbb{R}^{1+2} with subconformal power 1<p51<p\leq 5, the potential energy decays as follows

2|ϕ(t,x)|p+1𝑑xC0,2(1+t)p12,t0\displaystyle\int_{\mathbb{R}^{2}}|\phi(t,x)|^{p+1}dx\leq C\mathcal{E}_{0,2}(1+t)^{-\frac{p-1}{2}},\quad\forall t\geq 0

for some constant CC depending only on pp.

As a consequence for all 1+8<p<51+\sqrt{8}<p<5, the solution ϕ\phi scatters to linear solutions in the critical Sobolev space and energy space, that is, there exist pairs ϕ0±H˙xspH˙x1\phi_{0}^{\pm}\in\dot{H}_{x}^{s_{p}}\cap\dot{H}_{x}^{1} and ϕ1±H˙xsp1Lx2\phi_{1}^{\pm}\in\dot{H}_{x}^{s_{p}-1}\cap L_{x}^{2} such that

limt±(ϕ(t,x),tϕ(t,x))𝐋(t)(ϕ0±(x),ϕ1±(x))H˙xs×H˙xs1=0\displaystyle\lim\limits_{t\rightarrow\pm\infty}\|(\phi(t,x),\partial_{t}\phi(t,x))-\mathbf{L}(t)(\phi_{0}^{\pm}(x),\phi_{1}^{\pm}(x))\|_{\dot{H}_{x}^{s}\times\dot{H}_{x}^{s-1}}=0

for all sps1s_{p}\leq s\leq 1.

Remark 1.1.

The superconformal case p5p\geq 5 has been well understood in [8], [9]. Since we are aiming at comparing the nonlinear solution with linear solutions, the larger power pp makes the problem easier due to the decay of the solution.

Remark 1.2.

The time decay of the potential energy obtained in [9] is

2|ϕ(t,x)|p+1𝑑xC0,2(1+t)3p,t0.\displaystyle\int_{\mathbb{R}^{2}}|\phi(t,x)|^{p+1}dx\leq C\mathcal{E}_{0,2}(1+t)^{3-p},\quad\forall t\geq 0.

For the subconformal case when p<5p<5 we see that p3<p12p-3<\frac{p-1}{2}. In particular we improved the time decay of the potential energy.

Remark 1.3.

The lower bound for the above scattering result is consistent with that in higher dimensions (p(d)=1+d2+4d4d1p(d)=\frac{1+\sqrt{d^{2}+4d-4}}{d-1}, see the main theorem in [16]). However we still need to require that the initial data belong to the conformal energy space.

Remark 1.4.

Scattering in the critical Sobolev space H˙sp\dot{H}^{s_{p}} is motivated by the open problem that whether equation (1) is globally well-posed in critical Sobolev space, see recent breakthrough [4], [5] by Dodson in 1+3\mathbb{R}^{1+3}.

Based on the above time decay of the potential energy as well as Brézis-Gallouet-Wainger inequality, we also obtain pointwise decay estimates for the solution.

Theorem 2.

Consider the Cauchy problem to the defocusing semilinear wave equation (1). For initial data (ϕ0,ϕ1)(\phi_{0},\phi_{1}) bounded in 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}, the solution ϕ\phi satisfies the following pointwise decay estimates for all t0t\geq 0:

  • For the case when 113<p<5\frac{11}{3}<p<5, the solution enjoys the sharp time decay estimates

    |ϕ(t,x)|C(1+t+|x|)12|\phi(t,x)|\leq C(1+t+|x|)^{-\frac{1}{2}}

    for some constant CC depending only on pp, 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}.

  • Otherwise for all 1<p1131<p\leq\frac{11}{3} and ϵ>0\epsilon>0, the solution still decays

    |ϕ(t,x)|Cϵ(1+t+|x|)p18+ϵ|\phi(t,x)|\leq C_{\epsilon}(1+t+|x|)^{-\frac{p-1}{8}+\epsilon}

    with CϵC_{\epsilon} depending on pp, ϵ\epsilon and 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}.

Remark 1.5.

The constants CC and CϵC_{\epsilon} depend polynomially on 0,2\mathcal{E}_{0,2} and linearly on 1,0ϵ\mathcal{E}_{1,0}^{\epsilon}. One can make such dependence explicitly in the course of the proof.

Remark 1.6.

In dimension d=3d=3 as shown in [17], the solution decays as quickly as linear waves (along the cone) for p>p(d)p>p(d). Since 113<p(2)=1+8\frac{11}{3}<p(2)=1+\sqrt{8}, the sharp pointwise time decay estimate even holds for pp less than p(2)p(2) in 1+2\mathbb{R}^{1+2}. In general, we conjecture that the solution to (1) in 1+d\mathbb{R}^{1+d} scatters in critical Sobolev space and decays sharply (only in dimension d=2,3d=2,3) for 1+4d<p<1+4d21+\frac{4}{d}<p<1+\frac{4}{d-2}.

Remark 1.7.

Pointwise decay estimates are only available in [9] when p>1+8p>1+\sqrt{8}. One novelty of our work is that the pointwise decay estimate holds for all p>1p>1. Moreover, our decay estimates are stronger than those obtained by Glassey-Pecher. Hence we also improved the scattering result in energy space.

As a corollary of the above pointwise decay estimate, we extend Glassey-Pecher’s scattering result [9] in energy space.

Corollary 1.

For p>251p>2\sqrt{5}-1 and initial data bounded in 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}, the solution ϕ\phi of the defocusing semilinear wave equation (1) in 1+2\mathbb{R}^{1+2} verifies

ϕLtpLx2p<.\displaystyle\|\phi\|_{L_{t}^{p}L_{x}^{2p}}<\infty.

Consequently the solution scatters in energy space, that is, there exists pairs (ϕ0±(x),ϕ1±(x))(\phi_{0}^{\pm}(x),\phi_{1}^{\pm}(x)) such that

limt±ϕ(t,x)𝐋(t)(ϕ0±(x),ϕ1±(x))H˙x1+tϕ(t,x)t𝐋(t)(ϕ0±(x),ϕ1±(x))Lx2=0.\displaystyle\lim\limits_{t\rightarrow\pm\infty}\|\phi(t,x)-\mathbf{L}(t)(\phi_{0}^{\pm}(x),\phi_{1}^{\pm}(x))\|_{\dot{H}_{x}^{1}}+\|\partial_{t}\phi(t,x)-\partial_{t}\mathbf{L}(t)(\phi_{0}^{\pm}(x),\phi_{1}^{\pm}(x))\|_{L_{x}^{2}}=0.

To better elaborate our idea for the proof, let’s first review the existing methods to study the asymptotic decay properties for energy subcritical defocusing semilinear wave equations. The pioneering work dates back to Strauss in [14] for the equation in 1+3\mathbb{R}^{1+3} with superconformal power 3<p<53<p<5. The key observation was the almost conservation of conformal energy obtained by using the conformal Killing vector field as multiplier. The superconformal power 1+4d1<p1+\frac{4}{d-1}<p plays the role that the associated conformal energy is uniformly bounded by the initial data, which leads to the sharp time decay of the potential energy

d|ϕ|p+1𝑑xC(1+t)2,d2\displaystyle\int_{\mathbb{R}^{d}}|\phi|^{p+1}dx\leq C(1+t)^{-2},\quad\forall d\geq 2

This time decay estimate is also the starting point to conclude the pointwise decay estimates (in dimension d=2d=2 or 33) and scattering results [15], [8]. To go beyond the conformal power, Pecher [13] observed that instead of uniform bound, the conformal energy grows at most polynomially by using Gronwall’s inequality. And for pp close to the conformal invariant power, the potential energy still decays

Rd|ϕ|p+1𝑑xC(1+t)1+dp(d1),d2,p>d+1d1.\displaystyle\int_{\mathbb{}{R}^{d}}|\phi|^{p+1}dx\leq C(1+t)^{1+d-p(d-1)},\quad\forall d\geq 2,\quad p>\frac{d+1}{d-1}.

This covers part of subconformal power [9], [11], [10].

Improvements both on the range of power pp (smaller pp makes the problem more difficult) and decay estimates of the solution in higher dimension d3d\geq 3 were achieved by the second author in [16], [17], [18], which rely on the new vector field method original developed by Dafermos-Rodnianski [3] for the study of decay estimates for linear waves on black hole background. The crucial new ingredient of this approach is a type of rr-weighted energy estimates obtained by using the vector fields rγ(t+r)r^{\gamma}(\partial_{t}+\partial_{r}) as multipliers for all 0γ20\leq\gamma\leq 2. This leads to the time decay of the potential energy (comparatively, the original estimate is an integral version in time tt)

d|ϕ(t,x)|p+1𝑑xC(1+t)(d1)(p1)2.\displaystyle\int_{\mathbb{R}^{d}}|\phi(t,x)|^{p+1}dx\leq C(1+t)^{-\frac{(d-1)(p-1)}{2}}.

One notices that for the subconformal case when p<d+3d1p<\frac{d+3}{d-1}

(d1)(p1)2>1+dp(d1).\displaystyle\frac{(d-1)(p-1)}{2}>1+d-p(d-1).

In particular the new approach greatly improved the time decay of the potential energy for the subconformal case.

However this new approach only works for the equation in higher dimensions 1+d\mathbb{R}^{1+d}, d3d\geq 3 and fails when d=2d=2 for the reason that the lower order term generated by the use of vector field rγ(t+r)r^{\gamma}(\partial_{t}+\partial_{r}) with 1<γ<21<\gamma<2 switches sign to be negative.

To overcome this difficulty, we first apply the vector field

X=(x22+(tx1)2+1)t+(x22(tx1)2)1+2(tx1)x22\displaystyle X=(x_{2}^{2}+(t-x_{1})^{2}+1)\partial_{t}+(x_{2}^{2}-(t-x_{1})^{2})\partial_{1}+2(t-x_{1})x_{2}\partial_{2}

to the region bounded by the null hyperplane {t=x1}\{t=x_{1}\} and the initial hypersurface, which leads to the weighted energy estimate

t0x22|(t+1)ϕ|2(t,t,x2)+|2ϕ(t,t,x2)|2+2p+1|ϕ(t,t,x2)|p+1dtdx2C0,2.\int_{t\geq 0}x_{2}^{2}|(\partial_{t}+\partial_{1})\phi|^{2}(t,t,x_{2})+|\partial_{2}\phi(t,t,x_{2})|^{2}+\frac{2}{p+1}|\phi(t,t,x_{2})|^{p+1}dtdx_{2}\leq C\mathcal{E}_{0,2}.

Then we consider the vector field

X=u1p12(t1)+u1p122x22(t+1)+2u1p121x22\displaystyle X=u_{1}^{\frac{p-1}{2}}(\partial_{t}-\partial_{1})+u_{1}^{\frac{p-1}{2}-2}x_{2}^{2}(\partial_{t}+\partial_{1})+2u_{1}^{\frac{p-1}{2}-1}x_{2}\partial_{2}

with u1=tx1+1u_{1}=t-x_{1}+1 and region bounded by the null hyperplane {t=x1}\{t=x_{1}\}, the constant time tt hypersurface {t=t}\{t=t\} as well as the initial hypersurface. This implies the weighted energy estimate

x1tu1p52(x22+u12)|ϕ(t,x)|p+1𝑑x\displaystyle\int_{x_{1}\leq t}u_{1}^{\frac{p-5}{2}}(x_{2}^{2}+u_{1}^{2})|\phi(t,x)|^{p+1}dx
C0,2+Ct0(|x2(t+1)ϕ|2+|2ϕ|2+2p+1|ϕ|p+1)(t,t,x2)𝑑t𝑑x2\displaystyle\leq C\mathcal{E}_{0,2}+C\int_{t\geq 0}(|x_{2}(\partial_{t}+\partial_{1})\phi|^{2}+|\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})(t,t,x_{2})dtdx_{2}
C0,2\displaystyle\leq C\mathcal{E}_{0,2}

in view of the previous weighted energy estimate through the null hyperplane {t=x1}\{t=x_{1}\}. Since we only study the equation in the future t0t\geq 0, the above estimate in particular shows that

x10(1+t)p12|ϕ(t,x)|p+1𝑑xC0,2\displaystyle\int_{x_{1}\leq 0}(1+t)^{\frac{p-1}{2}}|\phi(t,x)|^{p+1}dx\leq C\mathcal{E}_{0,2}

as u1=tx1+1u_{1}=t-x_{1}+1. By symmetry (x1x_{1} to x1-x_{1}) the above estimate also holds on the region x10x_{1}\geq 0 (the time tt is always fixed). This demonstrates the improved time decay of the potential energy at time tt.


For the pointwise decay estimate for the solution, we go back to the conformal energy identity to derive that

(1+t)2/ ϕL22+(1+|tr|)ϕL22+ϕL22C(1+t)5p2\displaystyle(1+t)^{2}\|\mbox{$\nabla\mkern-13.0mu/$\,}\phi\|_{L^{2}}^{2}+\|(1+|t-r|)\nabla\phi\|_{L^{2}}^{2}+\|\phi\|_{L^{2}}^{2}\leq C(1+t)^{\frac{5-p}{2}}

in view of the above time decay of the potential energy. Here / =r1xr\mbox{$\nabla\mkern-13.0mu/$\,}=\nabla-r^{-1}x\partial_{r} is the angular derivative. The rough pointwise decay estimates for the solution then follow by use of the following Brézis-Gallouet-Wainger inequality

uL(2)CuH1(2)(1+lnuH2(2)uH1(2))12\displaystyle\|u\|_{L^{\infty}(\mathbb{R}^{2})}\leq C\|u\|_{H^{1}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(\mathbb{R}^{2})}}{\|u\|_{H^{1}(\mathbb{R}^{2})}}\right)^{\frac{1}{2}}

adapted to regions away from the light cone {t=|x|}\{t=|x|\} and annulus near the light cone, see details in Lemma 4.1. This is sufficiently strong to conclude the pointwise decay estimate

|ϕ(t,x)|C(1+t)p18+ϵ|\phi(t,x)|\leq C(1+t)^{-\frac{p-1}{8}+\epsilon}

for all 1<p<51<p<5. The loss of decay rate (1+t)ϵ(1+t)^{\epsilon} arises due to the at most polynomial growth of the second order energy ϕ(t,x)H2\|\phi(t,x)\|_{H^{2}}.

Finally for the improved sharp decay estimate for larger pp, inspired by the idea from [6], we write the nonlinear solution as sum of linear solution and contribution from nonlinearity. By carefully estimating the nonlinearity, a type of Gronwall’s inequality then leads to sharp pointwise decay of the solution when p>113p>\frac{11}{3}.

The plan of the paper is as follows: In Section 2, we define some notations and recall the energy method for wave equations. In Section 3, we use vector field method applied to hyperplane instead of cones to derive time decay estimates for the potential energy. In section 4, we develop logarithmic Sobolev embedding based on Brézis-Gallouet-Wainger inequaltiy and conclude the pointwise decay estimates for the solution.

Acknowledgments. The authors would like to thank R. Frank for kindly pointing out the origin of Brézis-Gallouet-Wainger inequality. S. Yang is partially supported by NSFC-11701017.

2 Preliminaries and energy identities

Let’s first define some necessary notations. We use the standard Cartesian coordinates (t,x1,x2)(t,x_{1},x_{2}) and polar local coordinate system (t,r,θ)(t,r,\theta) of Minkowski space. The initial hypersurface {t=0}\{t=0\} will be denoted as Σ0\Sigma_{0} and Σt\Sigma_{t} stands for the constant time tt hypersurface. Since the wave equation is time reversible, without loss of generality we only prove estimates in the future, i.e., t0t\geq 0.

As discussed in the introduction, the main idea to study such large data problem is to start with some weighted energy estimate, which could be obtained by using vector field method in the following way: Define the associated energy momentum tensor for the scalar field ϕ\phi

T[ϕ]μν=μϕνϕ12mμν(γϕγϕ+2p+1|ϕ|p+1),\displaystyle T[\phi]_{\mu\nu}=\partial_{\mu}\phi\partial_{\nu}\phi-\frac{1}{2}m_{\mu\nu}(\partial^{\gamma}\phi\partial_{\gamma}\phi+\frac{2}{p+1}|\phi|^{p+1}),

where mμνm_{\mu\nu} is the flat Minkowski metric on 1+2\mathbb{R}^{1+2}.

For any vector field XX and any function χ\chi, define the current

JμX,χ[ϕ]=T[ϕ]μνXν12μχ|ϕ|2+12χμ|ϕ|2.J^{X,\chi}_{\mu}[\phi]=T[\phi]_{\mu\nu}X^{\nu}-\frac{1}{2}\partial_{\mu}\chi\cdot|\phi|^{2}+\frac{1}{2}\chi\partial_{\mu}|\phi|^{2}.

By using Stokes’ formula, we derive the energy identity

𝒟iJX,χ[ϕ]𝑑vol=𝒟T[ϕ]μνπμνX+χμϕμϕ12χ|ϕ|2+χϕϕ+X(ϕ)(ϕ|ϕ|p1ϕ)dvol\int_{\partial\mathcal{D}}i_{J^{X,\chi}[\phi]}d\textnormal{vol}=\iint_{\mathcal{D}}T[\phi]^{\mu\nu}\pi^{X}_{\mu\nu}+\chi\partial_{\mu}\phi\partial^{\mu}\phi-\frac{1}{2}\Box\chi\cdot|\phi|^{2}+\chi\phi\Box\phi+X(\phi)(\Box\phi-|\phi|^{p-1}\phi)d\textnormal{vol} (2)

for any domain 𝒟\mathcal{D} in 1+2\mathbb{R}^{1+2}. Here πX=12Xm\pi^{X}=\frac{1}{2}\mathcal{L}_{X}m is the deformation tensor of the metric mm along the vector field XX and iZdvoli_{Z}d\textnormal{vol} is the contraction of the vector field with the volume form dvold\textnormal{vol}. By suitably choosing vector field XX and function χ\chi, we derive weighted energy estimate for solution of (1). The ultimate goal is to show the time decay of the potential energy.

Once we have time decay estimate for the potential energy, we rely on the following representation formula for linear wave equation with initial data (ϕ0,ϕ1)(\phi_{0},\phi_{1})

2πϕ(t,x)=t|y|<1ϕ1(x+ty)1|y|2𝑑y+t(t|y|<1ϕ0(x+ty)1|y|2𝑑y)+0t(ts)|y|<1ϕ(s,x+(ts)y)1|y|2𝑑y𝑑s.\begin{split}2\pi\phi(t,x)=&t\int_{|y|<1}\frac{\phi_{1}(x+ty)}{\sqrt{1-|y|^{2}}}dy+\partial_{t}\left(t\int_{|y|<1}\frac{\phi_{0}(x+ty)}{\sqrt{1-|y|^{2}}}dy\right)\\ &+\int_{0}^{t}(t-s)\int_{|y|<1}\frac{\Box\phi(s,x+(t-s)y)}{\sqrt{1-|y|^{2}}}dyds.\end{split} (3)

Finally we make a convention through out this paper to avoid too many constants that ABA\lesssim B means that there exists a constant CC, depending possibly on pp, 0,2\mathcal{E}_{0,2}, 1,0\mathcal{E}_{1,0} such that ACBA\leq CB.

3 Improved time decay of the potential energy

For linear wave equations in space dimension greater than two, one can obtain the integrated local energy decay estimates as well as energy flux decay estimates through Dafermos-Rodnianski’s modified vector field method [3]. However such method fails in space dimension 22 due to the wrong sign of lower order terms. So far as the authors know, it is even not clear whether the associated version holds for linear waves in 1+2\mathbb{R}^{1+2}, at least through vector field method. Since our goal is to investigate the asymptotic decay properties for solutions of defocusing semilinear wave equation, we instead derive new estimates by introducing new vector fields to show the time decay of the potential energy, which substantially improves the existing time decay obtained by using the conformal Killing vector field as multiplier. As the superconformal case p5p\geq 5 is easier and has been well understood in [9], in the sequel we only study the solution with subconformal power

1<p<5.1<p<5.

We begin in this section with the improved time decay of the potential energy.

Proposition 3.1.

For solution ϕ\phi to the defocusing semilinear wave equation (1) in 1+2\mathbb{R}^{1+2} with subconformal power 1<p<51<p<5, the potential energy decays as follows

2|ϕ(t,x)|p+1(1+t+|x|)p12𝑑xC0,2,t0\displaystyle\int_{\mathbb{R}^{2}}|\phi(t,x)|^{p+1}(1+t+|x|)^{\frac{p-1}{2}}dx\leq C\mathcal{E}_{0,2},\quad\forall t\geq 0

for some constant CC depending only on pp.

As a consequence for all 1+8<p<51+\sqrt{8}<p<5, the solution ϕ\phi is bounded in the following spacetime norm

ϕL3(p1)2(1+2)C0,21p+1\displaystyle\|\phi\|_{L^{\frac{3(p-1)}{2}}(\mathbb{R}^{1+2})}\leq C\mathcal{E}_{0,2}^{\frac{1}{p+1}} (4)

for some constant CC relying only on pp.

Proof.

In the energy identity (2), choose the vector field XX and the function χ\chi as follows:

X=(r2t2)(t+1)+2(tx1)S\displaystyle X=(r^{2}-t^{2})(\partial_{t}+\partial_{1})+2(t-x_{1})S =(x22+(tx1)2)t+(x22(tx1)2)1+2(tx1)x22,\displaystyle=(x_{2}^{2}+(t-x_{1})^{2})\partial_{t}+(x_{2}^{2}-(t-x_{1})^{2})\partial_{1}+2(t-x_{1})x_{2}\partial_{2},
χ=tx1.\displaystyle\chi=t-x_{1}.

Here r=|x|r=|x| and S=tt+rrS=t\partial_{t}+r\partial_{r} is the spacetime scaling vector field. We then compute that

πX=2(tx1)m,tχ=1,(t+1)χ=0,χ=0.\pi^{X}=2(t-x_{1})m,\quad\partial_{t}\chi=1,\quad(\partial_{t}+\partial_{1})\chi=0,\quad\Box\chi=0.

where mm is the Minkowski metric on 1+2\mathbb{R}^{1+2}. Thus the bulk term on the right hand side of the energy identity (2) is

T[ϕ]μνπμνX+χμϕμϕ+χϕϕ12χ|ϕ|2\displaystyle T[\phi]^{\mu\nu}\pi^{X}_{\mu\nu}+\chi\partial_{\mu}\phi\partial^{\mu}\phi+\chi\phi\Box\phi-\frac{1}{2}\Box\chi|\phi|^{2}
=6(tx1)p+1|ϕ|p+1+(tx1)|ϕ|p+1=(5p)(tx1)p+1|ϕ|p+1.\displaystyle=-\frac{6(t-x_{1})}{p+1}|\phi|^{p+1}+(t-x_{1})|\phi|^{p+1}=-\frac{(5-p)(t-x_{1})}{p+1}|\phi|^{p+1}.

In particular, the bulk integral is nonnegative when p5p\leq 5 and tx1t\leq x_{1}.

Next we need to compute the left hand side of the energy identity (2) for the region 𝒟={0tmin(x1,T)}\mathcal{D}=\{0\leq t\leq\min(x_{1},T)\} with boundary 𝒟\partial\mathcal{D} consisting of the null hyperplane {tx1=0}\{t-x_{1}=0\} and the constant tt-slices. Recall the volume form

dvol=dtdx=dtdx1dx2.d\textnormal{vol}=dtdx=dtdx_{1}dx_{2}.

Hence for the null hypersurface {tx1=0}\{t-x_{1}=0\}, we can compute that

iJX,χ[ϕ]dvol\displaystyle i_{J^{X,\chi}[\phi]}d\textnormal{vol} =(JX,χ[ϕ])L1dtdx2=(T[ϕ]L1νXν12L1χ|ϕ|2+12χL1|ϕ|2)dtdx2\displaystyle=-(J^{X,\chi}[\phi])_{L_{1}}dtdx_{2}=-(T[\phi]_{L_{1}\nu}X^{\nu}-\frac{1}{2}L_{1}\chi|\phi|^{2}+\frac{1}{2}\chi\cdot L_{1}|\phi|^{2})dtdx_{2}
=T[ϕ]L1Xdtdx2=x22T[ϕ]L1L1dtdx2=x22|L1ϕ|2dtdx2.\displaystyle=-T[\phi]_{L_{1}X}dtdx_{2}=-x_{2}^{2}T[\phi]_{L_{1}L_{1}}dtdx_{2}=-x_{2}^{2}|L_{1}\phi|^{2}dtdx_{2}.

Here we denote L1=t+1L_{1}=\partial_{t}+\partial_{1} and have used the fact that on the hyperplane {tx1=0}\{t-x_{1}=0\}

L1χ=0,χ=0,X=x22L1.L_{1}\chi=0,\quad\chi=0,\quad X=x_{2}^{2}L_{1}.

Then on the constant tt-slice, we can show that

iJX,χ[ϕ]dvol=\displaystyle i_{J^{X,\chi}[\phi]}d\textnormal{vol}= (JX,χ[ϕ])0dx=(T[ϕ]0νXν12tχ|ϕ|2+12χt|ϕ|2)dx\displaystyle(J^{X,\chi}[\phi])^{0}dx=-(T[\phi]_{0\nu}X^{\nu}-\frac{1}{2}\partial_{t}\chi|\phi|^{2}+\frac{1}{2}\chi\cdot\partial_{t}|\phi|^{2})dx
=\displaystyle= 12((x22+(tx1)2)(|tϕ|2+|ϕ|2+2p+1|ϕ|p+1)|ϕ|2+(tx1)t|ϕ|2\displaystyle-\frac{1}{2}\big{(}(x_{2}^{2}+(t-x_{1})^{2})(|\partial_{t}\phi|^{2}+|\nabla\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})-|\phi|^{2}+(t-x_{1})\partial_{t}|\phi|^{2}
+2tϕ((x22(tx1)2)1ϕ+2(tx1)x22ϕ))dx\displaystyle\quad\qquad+2\partial_{t}\phi((x_{2}^{2}-(t-x_{1})^{2})\partial_{1}\phi+2(t-x_{1})x_{2}\partial_{2}\phi)\big{)}dx
=\displaystyle= 12(|x2(tϕ+1ϕ)+(tx1)2ϕ|2+|(tx1)(tϕ1ϕ)+x22ϕ+ϕ|2\displaystyle-\frac{1}{2}(|x_{2}(\partial_{t}\phi+\partial_{1}\phi)+(t-x_{1})\partial_{2}\phi|^{2}+|(t-x_{1})(\partial_{t}\phi-\partial_{1}\phi)+x_{2}\partial_{2}\phi+\phi|^{2}
+2(x22+(tx1)2)p+1|ϕ|p+11((x1t)ϕ2)2(x2ϕ2))dx.\displaystyle\qquad+\frac{2(x_{2}^{2}+(t-x_{1})^{2})}{p+1}|\phi|^{p+1}-\partial_{1}((x_{1}-t)\phi^{2})-\partial_{2}(x_{2}\phi^{2}))dx.

Observe that on the constant tt-slice, we also have

x1t(1((x1t)ϕ2)+2(x2ϕ2))(t,x)𝑑x=0.\displaystyle\int_{x_{1}\geq t}(\partial_{1}((x_{1}-t)\phi^{2})+\partial_{2}(x_{2}\phi^{2}))(t,x)dx=0.

This holds whenever ϕ\phi decays suitably at spatial infinity. Alternatively, one can choose bounded region and then take limit. The above computations together with the energy identity (2) then lead to the following weighted energy identity for 𝒟={0tmin(x1,T)}\mathcal{D}=\{0\leq t\leq\min(x_{1},T)\}

0tmin(x1,T)(5p)(x1t)p+1|ϕ|p+1+0tTx22|L1ϕ|2(t,t,x2)𝑑t𝑑x2\displaystyle\iint_{0\leq t\leq\min(x_{1},T)}\frac{(5-p)(x_{1}-t)}{p+1}|\phi|^{p+1}+\int_{0\leq t\leq T}x_{2}^{2}|L_{1}\phi|^{2}(t,t,x_{2})dtdx_{2}
+12x1t(|x2L1ϕ+(tx1)2ϕ|2+|(tx1)(tϕ1ϕ)+x22ϕ+ϕ|2\displaystyle+\frac{1}{2}\int_{x_{1}\geq t}(|x_{2}L_{1}\phi+(t-x_{1})\partial_{2}\phi|^{2}+|(t-x_{1})(\partial_{t}\phi-\partial_{1}\phi)+x_{2}\partial_{2}\phi+\phi|^{2}
+2(x22+(tx1)2)p+1|ϕ|p+1)dx|t=0t=T=0.\displaystyle\qquad+\frac{2(x_{2}^{2}+(t-x_{1})^{2})}{p+1}|\phi|^{p+1})dx\Big{|}_{t=0}^{t=T}=0.

Letting T+T\to+\infty we conclude that

t0x22|L1ϕ|2(t,t,x2)𝑑t𝑑x212x10(|x2(ϕ1+1ϕ0)x12ϕ0|2+|x1(ϕ11ϕ0)+x22ϕ0+ϕ0|2+2(x22+x12)p+1|ϕ0|p+1)dxC{t=0}(1+|x|2)(|ϕ1|2+|ϕ0|2+2p+1|ϕ0|p+1)𝑑x=C0,2.\begin{split}&\int_{t\geq 0}x_{2}^{2}|L_{1}\phi|^{2}(t,t,x_{2})dtdx_{2}\\ &\leq\frac{1}{2}\int_{x_{1}\geq 0}(|x_{2}(\phi_{1}+\partial_{1}\phi_{0})-x_{1}\partial_{2}\phi_{0}|^{2}+|-x_{1}(\phi_{1}-\partial_{1}\phi_{0})+x_{2}\partial_{2}\phi_{0}+\phi_{0}|^{2}\\ &\qquad\qquad+\frac{2(x_{2}^{2}+x_{1}^{2})}{p+1}|\phi_{0}|^{p+1})dx\\ &\leq C\int_{\{t=0\}}(1+|x|^{2})(|\phi_{1}|^{2}+|\nabla\phi_{0}|^{2}+\frac{2}{p+1}|\phi_{0}|^{p+1})dx\\ &=C\mathcal{E}_{0,2}.\end{split} (5)

Here and in the following the constant CC relies only on pp and some small positive constant ϵ\epsilon (in case such constant appears). In last step the integral of |ϕ0|2|\phi_{0}|^{2} can be controlled by the weighted energy by using Hardy’s inequality.


Next in the energy identity (2), choose the vector field X=t,χ=0,X=\partial_{t},\ \chi=0, then πX=0,\pi^{X}=0, and

T[ϕ]μνπμνX+χμϕμϕ+χϕϕ12χ|ϕ|2=0.\displaystyle T[\phi]^{\mu\nu}\pi^{X}_{\mu\nu}+\chi\partial_{\mu}\phi\partial^{\mu}\phi+\chi\phi\Box\phi-\frac{1}{2}\Box\chi|\phi|^{2}=0.

For the null hypersurface {tx1=0}\{t-x_{1}=0\}, we have

iJX,χ[ϕ]dvol\displaystyle i_{J^{X,\chi}[\phi]}d\textnormal{vol} =(JX,χ[ϕ])L1dtdx2=T[ϕ]L1Xdtdx2\displaystyle=-(J^{X,\chi}[\phi])_{L_{1}}dtdx_{2}=-T[\phi]_{L_{1}X}dtdx_{2}
=(tϕ(t+1)ϕ+12(|tϕ|2+|ϕ|2+2p+1|ϕ|p+1))dtdx2\displaystyle=-(\partial_{t}\phi(\partial_{t}+\partial_{1})\phi+\frac{1}{2}(-|\partial_{t}\phi|^{2}+|\nabla\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1}))dtdx_{2}
=12(|L1ϕ|2+|2ϕ|2+2p+1|ϕ|p+1)dtdx2.\displaystyle=-\frac{1}{2}(|L_{1}\phi|^{2}+|\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})dtdx_{2}.

On the constant tt-slice, we can show that

iJX,χ[ϕ]dvol=\displaystyle i_{J^{X,\chi}[\phi]}d\textnormal{vol}= (JX,χ[ϕ])0dx=(T[ϕ]0νXν)dx\displaystyle(J^{X,\chi}[\phi])^{0}dx=-(T[\phi]_{0\nu}X^{\nu})dx
=\displaystyle= 12(|tϕ|2+|ϕ|2+2p+1|ϕ|p+1)dx.\displaystyle-\frac{1}{2}(|\partial_{t}\phi|^{2}+|\nabla\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})dx.

We therefore derive that

0tT(|L1ϕ|2+|2ϕ|2+2p+1|ϕ|p+1)(t,t,x2)𝑑t𝑑x2\displaystyle\int_{0\leq t\leq T}(|L_{1}\phi|^{2}+|\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})(t,t,x_{2})dtdx_{2}
+x1t(|tϕ|2+|ϕ|2+2p+1|ϕ|p+1)𝑑x|t=0t=T=0.\displaystyle+\int_{x_{1}\geq t}(|\partial_{t}\phi|^{2}+|\nabla\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})dx\Big{|}_{t=0}^{t=T}=0.

Similarly by letting T+T\to+\infty, we conclude that

t0(|L1ϕ|2+|2ϕ|2+2p+1|ϕ|p+1)(t,t,x2)𝑑t𝑑x2\displaystyle\int_{t\geq 0}(|L_{1}\phi|^{2}+|\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})(t,t,x_{2})dtdx_{2} (6)
x10(|ϕ1|2+|ϕ0|2+2p+1|ϕ0|p+1)𝑑x0,0.\displaystyle\leq\int_{x_{1}\geq 0}(|\phi_{1}|^{2}+|\nabla\phi_{0}|^{2}+\frac{2}{p+1}|\phi_{0}|^{p+1})dx\leq\mathcal{E}_{0,0}.

Finally in the energy identity (2), choose the vector field XX and the function χ\chi as follows:

X\displaystyle X =u1q(t1)+u1q2x22(t+1)+2u1q1x22,\displaystyle=u_{1}^{q}(\partial_{t}-\partial_{1})+u_{1}^{q-2}x_{2}^{2}(\partial_{t}+\partial_{1})+2u_{1}^{q-1}x_{2}\partial_{2},
χ\displaystyle\chi =u1q1,u1=tx1+1,q=12(p1).\displaystyle=u_{1}^{q-1},\quad u_{1}=t-x_{1}+1,\quad q=\frac{1}{2}(p-1).

Here we assume tx1t\geq x_{1}. Recall that L1=(t+1),L¯1=(t1)L_{1}=(\partial_{t}+\partial_{1}),\ \underline{L}_{1}=(\partial_{t}-\partial_{1}). We then compute that

L1X=0,\displaystyle\nabla_{L_{1}}X=0,\quad L¯1X=2qu1q1L¯1+2(q2)u1q3x22L1+4(q1)u1q2x22,\displaystyle\nabla_{\underline{L}_{1}}X=2qu_{1}^{q-1}\underline{L}_{1}+2(q-2)u_{1}^{q-3}x_{2}^{2}L_{1}+4(q-1)u_{1}^{q-2}x_{2}\partial_{2},
2X=2u1q2x2L1+2u1q12.\displaystyle\nabla_{2}X=2u_{1}^{q-2}x_{2}L_{1}+2u_{1}^{q-1}\partial_{2}.

In particular the non-vanishing components of the deformation tensor πμνX\pi_{\mu\nu}^{X} are

πL1L¯1X\displaystyle\pi^{X}_{L_{1}\underline{L}_{1}} =2qu1q1,πL¯1L¯1X=4(q2)u1q3x22,\displaystyle=-2qu_{1}^{q-1},\quad\pi^{X}_{\underline{L}_{1}\underline{L}_{1}}=-4(q-2)u_{1}^{q-3}x_{2}^{2},
πL¯12X\displaystyle\pi^{X}_{\underline{L}_{1}\partial_{2}} =2(q2)u1q2x2,π22X=2u1q1.\displaystyle=2(q-2)u_{1}^{q-2}x_{2},\quad\pi^{X}_{\partial_{2}\partial_{2}}=2u_{1}^{q-1}.

Since χ=0\Box\chi=0, we therefore can compute that

T[ϕ]μνπμνX+χμϕμϕ+χϕϕ12χ|ϕ|2\displaystyle T[\phi]^{\mu\nu}\pi^{X}_{\mu\nu}+\chi\partial_{\mu}\phi\partial^{\mu}\phi+\chi\phi\Box\phi-\frac{1}{2}\Box\chi|\phi|^{2}
=qu+q1(|2(ϕ)|2+2p+1|ϕ|p+1)+2u1q1(|2(ϕ)|212μϕμϕ1p+1|ϕ|p+1)\displaystyle=-qu_{+}^{q-1}(|\partial_{2}(\phi)|^{2}+\frac{2}{p+1}|\phi|^{p+1})+2u_{1}^{q-1}(|\partial_{2}(\phi)|^{2}-\frac{1}{2}\partial^{\mu}\phi\partial_{\mu}\phi-\frac{1}{p+1}|\phi|^{p+1})
(q2)u1q3x22|L1ϕ|22(q2)u1q2x22ϕL1ϕ+u1q1μϕμϕ+u1q1|ϕ|p+1\displaystyle\quad-(q-2)u_{1}^{q-3}x_{2}^{2}|L_{1}\phi|^{2}-2(q-2)u_{1}^{q-2}x_{2}\partial_{2}\phi\cdot L_{1}\phi+u_{1}^{q-1}\partial_{\mu}\phi\partial^{\mu}\phi+u_{1}^{q-1}|\phi|^{p+1}
=(2q)u1q3|x2L1ϕ+u12ϕ|2.\displaystyle=(2-q)u_{1}^{q-3}|x_{2}L_{1}\phi+u_{1}\partial_{2}\phi|^{2}.

Here we used 2q+2=p+1.2q+2=p+1. In particular when p5p\leq 5, we have q2q\leq 2 and the bulk integral is nonnegative.

Now we compute the boundary integrals for the region 𝒟={max(x1,0)tT}\mathcal{D}=\{\max(x_{1},0)\leq t\leq T\}. On the null hyperplane {tx1=0}\{t-x_{1}=0\}, we have

χ=u1=1,X=L¯1+x22L1+2x22.\chi=u_{1}=1,\quad X=\underline{L}_{1}+x_{2}^{2}L_{1}+2x_{2}\partial_{2}.

Therefore we have

iJX,χ[ϕ]dvol\displaystyle i_{J^{X,\chi}[\phi]}d\textnormal{vol} =(JX,χ[ϕ])L1dtdx2=(T[ϕ]L1νXν12L1χ|ϕ|2+12χL1|ϕ|2)dtdx2\displaystyle=-(J^{X,\chi}[\phi])_{L_{1}}dtdx_{2}=-(T[\phi]_{L_{1}\nu}X^{\nu}-\frac{1}{2}L_{1}\chi|\phi|^{2}+\frac{1}{2}\chi\cdot L_{1}|\phi|^{2})dtdx_{2}
=(x22T[ϕ]L1L1+T[ϕ]L1L¯1+2x2T[ϕ]L12+12L1|ϕ|2)dtdx2\displaystyle=-(x_{2}^{2}T[\phi]_{L_{1}L_{1}}+T[\phi]_{L_{1}\underline{L}_{1}}+2x_{2}T[\phi]_{L_{1}\partial_{2}}+\frac{1}{2}L_{1}|\phi|^{2})dtdx_{2}
=(x22|L1ϕ|2+|2(ϕ)|2+2p+1|ϕ|p+1+2x22ϕL1ϕ+12L1|ϕ|2)dtdx2\displaystyle=-(x_{2}^{2}|L_{1}\phi|^{2}+|\partial_{2}(\phi)|^{2}+\frac{2}{p+1}|\phi|^{p+1}+2x_{2}\partial_{2}\phi\cdot L_{1}\phi+\frac{1}{2}L_{1}|\phi|^{2})dtdx_{2}
=(|x2L1ϕ+2ϕ|2+2p+1|ϕ|p+1+12L1|ϕ|2)dtdx2.\displaystyle=-(|x_{2}L_{1}\phi+\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1}+\frac{1}{2}L_{1}|\phi|^{2})dtdx_{2}.

On the constant tt-slice, we can show that

iJX,χ[ϕ]dvol=\displaystyle i_{J^{X,\chi}[\phi]}d\textnormal{vol}= (JX,χ[ϕ])0dx=(T[ϕ]0νXν12tχ|ϕ|2+12χt|ϕ|2)dx\displaystyle(J^{X,\chi}[\phi])^{0}dx=-(T[\phi]_{0\nu}X^{\nu}-\frac{1}{2}\partial_{t}\chi|\phi|^{2}+\frac{1}{2}\chi\cdot\partial_{t}|\phi|^{2})dx
=\displaystyle= 12u1q2((x22+u12)(|tϕ|2+|ϕ|2+2p+1|ϕ|p+1)(q1)|ϕ|2+u1t|ϕ|2\displaystyle-\frac{1}{2}u_{1}^{q-2}\big{(}(x_{2}^{2}+u_{1}^{2})(|\partial_{t}\phi|^{2}+|\nabla\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})-(q-1)|\phi|^{2}+u_{1}\partial_{t}|\phi|^{2}
+2(x22u12)tϕ1ϕ+4u1x2tϕ2ϕ)dx\displaystyle\qquad\qquad+2(x_{2}^{2}-u_{1}^{2})\partial_{t}\phi\cdot\partial_{1}\phi+4u_{1}x_{2}\partial_{t}\phi\cdot\partial_{2}\phi\big{)}dx
=\displaystyle= 12u1q2(|x2(tϕ+1ϕ)+u12ϕ|2+|u1(tϕ1ϕ)+x22ϕ+ϕ|2\displaystyle-\frac{1}{2}u_{1}^{q-2}\big{(}|x_{2}(\partial_{t}\phi+\partial_{1}\phi)+u_{1}\partial_{2}\phi|^{2}+|u_{1}(\partial_{t}\phi-\partial_{1}\phi)+x_{2}\partial_{2}\phi+\phi|^{2}
+2(x22+u12)p+1|ϕ|p+1)dx12(1(u1q1ϕ2)2(x2u1q2ϕ2))dx.\displaystyle\qquad\qquad+\frac{2(x_{2}^{2}+u_{1}^{2})}{p+1}|\phi|^{p+1}\big{)}dx-\frac{1}{2}(\partial_{1}(u_{1}^{q-1}\phi^{2})-\partial_{2}(x_{2}u_{1}^{q-2}\phi^{2}))dx.

On the boundary 𝒟\partial\mathcal{D} (for 𝒟={max(x1,0)tT}\mathcal{D}=\{\max(x_{1},0)\leq t\leq T\}) with suitable decay of the solution ϕ\phi at spatial infinity, note that

x1t(1(u1q1ϕ2)2(x2u1q2ϕ2))𝑑x|t=0t=T=0tT(L1|ϕ|2)(t,t,x2)𝑑t𝑑x2.\displaystyle\int_{x_{1}\leq t}(\partial_{1}(u_{1}^{q-1}\phi^{2})-\partial_{2}(x_{2}u_{1}^{q-2}\phi^{2}))dx\Big{|}_{t=0}^{t=T}=\int_{0\leq t\leq T}(L_{1}|\phi|^{2})(t,t,x_{2})dtdx_{2}.

Denote

E[ϕ]q(t)=x1tu1q2(\displaystyle E[\phi]_{q}(t)=\int_{x_{1}\leq t}u_{1}^{q-2}( |x2(tϕ+1ϕ)+u12ϕ|2+|u1(tϕ1ϕ)+x22ϕ+ϕ|2\displaystyle|x_{2}(\partial_{t}\phi+\partial_{1}\phi)+u_{1}\partial_{2}\phi|^{2}+|u_{1}(\partial_{t}\phi-\partial_{1}\phi)+x_{2}\partial_{2}\phi+\phi|^{2}
+2(x22+u12)p+1|ϕ|p+1)(t,x)dx.\displaystyle+\frac{2(x_{2}^{2}+u_{1}^{2})}{p+1}|\phi|^{p+1})(t,x)dx.

Then the above computations together with the energy identity (2) lead to the following weighted energy identity

max(x1,0)tT(2q)u1q3|x2Lϕ+u12ϕ|2+12(E[ϕ]q(T)E[ϕ]q(0))\displaystyle\iint_{\max(x_{1},0)\leq t\leq T}(2-q)u_{1}^{q-3}|x_{2}L\phi+u_{1}\partial_{2}\phi|^{2}+\frac{1}{2}(E[\phi]_{q}(T)-E[\phi]_{q}(0))
0tT(|x2L1ϕ+2ϕ|2+2p+1|ϕ|p+1)(t,t,x2)𝑑t𝑑x2=0,\displaystyle-\int_{0\leq t\leq T}(|x_{2}L_{1}\phi+\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})(t,t,x_{2})dtdx_{2}=0,

from which we conclude that

E[ϕ]q(T)E[ϕ]q(0)+20tT(|x2L1ϕ+2ϕ|2+2p+1|ϕ|p+1)(t,t,x2)𝑑t𝑑x2.\displaystyle E[\phi]_{q}(T)\leq E[\phi]_{q}(0)+2\int_{0\leq t\leq T}(|x_{2}L_{1}\phi+\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})(t,t,x_{2})dtdx_{2}.

Note that on the initial hypersurface {t=0}\{t=0\},

E[ϕ]q(0)\displaystyle E[\phi]_{q}(0) =x10(1x1)q2(|x2(ϕ1+1ϕ0)+(1x1)2ϕ0|2+2(x22+(1x1)2)p+1|ϕ0|p+1\displaystyle=\int_{x_{1}\leq 0}(1-x_{1})^{q-2}(|x_{2}(\phi_{1}+\partial_{1}\phi_{0})+(1-x_{1})\partial_{2}\phi_{0}|^{2}+\frac{2(x_{2}^{2}+(1-x_{1})^{2})}{p+1}|\phi_{0}|^{p+1}
+|(1x1)(ϕ11ϕ0)+x22ϕ0+ϕ0|2)dx\displaystyle\qquad\qquad+|(1-x_{1})(\phi_{1}-\partial_{1}\phi_{0})+x_{2}\partial_{2}\phi_{0}+\phi_{0}|^{2})dx
C0,2.\displaystyle\leq C\mathcal{E}_{0,2}.

Now by using estimates (5) and (6), we can bound the integral on the null hyperplane

20tT(|x2L1ϕ+2ϕ|2+2p+1|ϕ|p+1)(t,t,x2)𝑑t𝑑x2\displaystyle 2\int_{0\leq t\leq T}(|x_{2}L_{1}\phi+\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})(t,t,x_{2})dtdx_{2}
\displaystyle\leq 40tT(x22|L1ϕ|2+|2ϕ|2+2p+1|ϕ|p+1)(t,t,x2)𝑑t𝑑x2\displaystyle 4\int_{0\leq t\leq T}(x_{2}^{2}|L_{1}\phi|^{2}+|\partial_{2}\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1})(t,t,x_{2})dtdx_{2}
\displaystyle\leq C0,2.\displaystyle C\mathcal{E}_{0,2}.

We therefore conclude that

x1T2(u1q+u1q2x22)|ϕ|p+1p+1𝑑xE[ϕ]q(T)C0,2,T0.\displaystyle\int_{x_{1}\leq T}\frac{2(u_{1}^{q}+u_{1}^{q-2}x_{2}^{2})|\phi|^{p+1}}{p+1}dx\leq E[\phi]_{q}(T)\leq C\mathcal{E}_{0,2},\quad\forall T\geq 0.

Recall that u1=tx1+1u_{1}=t-x_{1}+1. By restricting the integral to the region {x10}\{x_{1}\leq 0\}, the above estimate in particular implies that

x102(t+1)qp+1|ϕ|p+1(t,x)𝑑x\displaystyle\int_{x_{1}\leq 0}\frac{2(t+1)^{q}}{p+1}|\phi|^{p+1}(t,x)dx x1t2u1qp+1|ϕ|p+1(t,x)𝑑xC0,2,\displaystyle\leq\int_{x_{1}\leq t}\frac{2u_{1}^{q}}{p+1}|\phi|^{p+1}(t,x)dx\leq C\mathcal{E}_{0,2},
x1tr2(1+2r)q2x22|ϕ|p+1p+1𝑑x\displaystyle\int_{x_{1}\leq t\leq r}\frac{2(1+2r)^{q-2}x_{2}^{2}|\phi|^{p+1}}{p+1}dx x1t2u1q2x22|ϕ|p+1p+1𝑑xCE[ϕ]q(T)C0,2.\displaystyle\leq\int_{x_{1}\leq t}\frac{2u_{1}^{q-2}x_{2}^{2}|\phi|^{p+1}}{p+1}dx\leq CE[\phi]_{q}(T)\leq C\mathcal{E}_{0,2}.

Here r=|x|=x12+x22r=|x|=\sqrt{x_{1}^{2}+x_{2}^{2}} and 0<q=p12<20<q=\frac{p-1}{2}<2. This in particular shows that

x10(1+t)p12|ϕ|p+1(t,x)𝑑x+x10,rt(1+r)p122x22|ϕ|p+1(t,x)𝑑x0,2.\displaystyle\int_{x_{1}\leq 0}(1+t)^{\frac{p-1}{2}}|\phi|^{p+1}(t,x)dx+\int_{x_{1}\leq 0,r\geq t}(1+r)^{\frac{p-1}{2}-2}x_{2}^{2}|\phi|^{p+1}(t,x)dx\leq\mathcal{E}_{0,2}.

By symmetry (or changing variable x1x1x_{1}\rightarrow-x_{1}), we also have

x10(1+t)p12|ϕ|p+1(t,x)𝑑x+x10,rt(1+r)p122x22|ϕ|p+1(t,x)𝑑xC0,2.\displaystyle\int_{x_{1}\geq 0}(1+t)^{\frac{p-1}{2}}|\phi|^{p+1}(t,x)dx+\int_{x_{1}\geq 0,r\geq t}(1+r)^{\frac{p-1}{2}-2}x_{2}^{2}|\phi|^{p+1}(t,x)dx\leq C\mathcal{E}_{0,2}.

Therefore we derive that

2(1+t)p12|ϕ|p+1(t,x)𝑑x+rt(1+r)p122x22|ϕ|p+1(t,x)𝑑xC0,2.\displaystyle\int_{\mathbb{R}^{2}}(1+t)^{\frac{p-1}{2}}|\phi|^{p+1}(t,x)dx+\int_{r\geq t}(1+r)^{\frac{p-1}{2}-2}x_{2}^{2}|\phi|^{p+1}(t,x)dx\leq C\mathcal{E}_{0,2}.

By symmetry again (switching the variable x1x_{1} and x2x_{2}), we also have

rt(1+r)p122x12|ϕ|p+1(t,x)𝑑xC0,2.\displaystyle\int_{r\geq t}(1+r)^{\frac{p-1}{2}-2}x_{1}^{2}|\phi|^{p+1}(t,x)dx\leq C\mathcal{E}_{0,2}.

Adding the above two estimate leads to the time decay of the potential energy

2(1+t+r)p12|ϕ|p+1(t,x)𝑑xC0,2.\displaystyle\int_{\mathbb{R}^{2}}(1+t+r)^{\frac{p-1}{2}}|\phi|^{p+1}(t,x)dx\leq C\mathcal{E}_{0,2}.

By our convention, the constant relies only on pp. We hence have shown the time decay of the potential energy of the Proposition.


Now based on the time decay estimate of the solution, we demonstrate that the solution is bounded in the spacetime norm L3(p1)2(1+2)L^{\frac{3(p-1)}{2}}(\mathbb{R}^{1+2}) for all 1+8<p<51+\sqrt{8}<p<5. Without loss of generality, it suffices to prove the estimate in the future t0t\geq 0. By using Hölder’s inequality, we can show that

ϕL3(p1)2(1+2{t0})\displaystyle\|\phi\|_{L^{\frac{3(p-1)}{2}}(\mathbb{R}^{1+2}\cap\{t\geq 0\})} (1+t+r)p12(p+1)ϕLxp+1(1+t+r)p12(p+1)Lxq1Lt3(p1)2({t0})\displaystyle\leq\|\|(1+t+r)^{\frac{p-1}{2(p+1)}}\phi\|_{L_{x}^{p+1}}\|(1+t+r)^{-\frac{p-1}{2(p+1)}}\|_{L_{x}^{q_{1}}}\|_{L_{t}^{\frac{3(p-1)}{2}}(\{t\geq 0\})}
C0,21p+1(1+t+r)p12(p+1)Lxq1Lt3(p1)2({t0}),\displaystyle\leq C\mathcal{E}_{0,2}^{\frac{1}{p+1}}\|\|(1+t+r)^{-\frac{p-1}{2(p+1)}}\|_{L_{x}^{q_{1}}}\|_{L_{t}^{\frac{3(p-1)}{2}}(\{t\geq 0\})},

where

1q1=23(p1)1p+1=5p3(p1)(p+1).\displaystyle\frac{1}{q_{1}}=\frac{2}{3(p-1)}-\frac{1}{p+1}=\frac{5-p}{3(p-1)(p+1)}.

Since p>1+8p>1+\sqrt{8}, we in particular have that

2<3(p1)22(5p).\displaystyle 2<\frac{3(p-1)^{2}}{2(5-p)}.

We therefore can compute that

(1+t+r)p12(p+1)Lxq1\displaystyle\|(1+t+r)^{-\frac{p-1}{2(p+1)}}\|_{L_{x}^{q_{1}}} =(2(1+t+r)p12(p+1)3(p1)(p+1)5p𝑑x)1q1\displaystyle=\left(\int_{\mathbb{R}^{2}}(1+t+r)^{-\frac{p-1}{2(p+1)}\frac{3(p-1)(p+1)}{5-p}}dx\right)^{\frac{1}{q_{1}}}
C(1+t)17+2p3p26(p1)(p+1).\displaystyle\leq C(1+t)^{\frac{17+2p-3p^{2}}{6(p-1)(p+1)}}.

This leads to

ϕL3(p1)2(1+2{t0})\displaystyle\|\phi\|_{L^{\frac{3(p-1)}{2}}(\mathbb{R}^{1+2}\cap\{t\geq 0\})} C0,2[ϕ]1p+1(0(1+t)17+2p3p26(p1)(p+1)3(p1)2𝑑t)23(p1)\displaystyle\leq C\mathcal{E}_{0,2}[\phi]^{\frac{1}{p+1}}\left(\int_{0}^{\infty}(1+t)^{\frac{17+2p-3p^{2}}{6(p-1)(p+1)}\cdot\frac{3(p-1)}{2}}dt\right)^{\frac{2}{3(p-1)}}
C0,21p+1\displaystyle\leq C\mathcal{E}_{0,2}^{\frac{1}{p+1}}

in view of the relation

17+2p3p2<4(p+1)\displaystyle 17+2p-3p^{2}<-4(p+1)

by the assumption p>1+8p>1+\sqrt{8}. We hence finished the proof for Proposition 3.1. ∎

The time decay estimate of main Theorem 1 follows from Proposition 4.2 immediately. The scattering results of Theorem 1 is a consequence of the spacetime bound (4). For details regarding this, we refer to Proposition 5.1 of [16]. We thus have shown Theorem 1.

4 Pointwise decay estimate of the solution

In this section, we rely on the potential energy decay obtained in the previous section to show the pointwise decay estimate for the solution. As we have discussed in the introduction, the large power pp makes the problem easier due to the better decay of the nonlinearity. And it is expected that for sufficiently large pp, the solution behaves like linear wave as shown in [9], [8] for the superconformal case p5p\geq 5. For the subconformal case studied in this paper, the solution decays faster for larger pp. The treatment varies for different range of pp. We first demonstrate a rough decay estimate for the solution for all 1<p<51<p<5.

4.1 Rough decay estimate for all pp

The obvious challenge in 2\mathbb{R}^{2} is the failure of traditional Sobolev inequality for LL^{\infty} estimate by the energy. This can be compensated by Brézis-Gallouet-Wainger inequality.

In 2\mathbb{R}^{2}, for 0s<R0\leq s<R, denote

BR={x:|x|R},AR,s={x:Rs|x|R+s}.\displaystyle B_{R}=\{x:|x|\leq R\},\quad A_{R,s}=\{x:R-s\leq|x|\leq R+s\}.
Lemma 4.1 (Brézis-Gallouet-Wainger inequality).

In 2\mathbb{R}^{2}, we have the following logarithmic Sobolev inequalities

uL(2)CuH1(2)(1+lnuH2(2)uH1(2))12,\displaystyle\|u\|_{L^{\infty}(\mathbb{R}^{2})}\leq C\|u\|_{H^{1}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(\mathbb{R}^{2})}}{\|u\|_{H^{1}(\mathbb{R}^{2})}}\right)^{\frac{1}{2}},
uL(B1/2)CuH1(B3/4)(1+lnuH2(B3/4)uH1(B3/4))12,\displaystyle\|u\|_{L^{\infty}(B_{1/2})}\leq C\|u\|_{H^{1}(B_{3/4})}\left(1+\ln\frac{\|u\|_{H^{2}(B_{3/4})}}{\|u\|_{H^{1}(B_{3/4})}}\right)^{\frac{1}{2}},
uL(B1c)CuH1(B5/6c)(1+lnuH2(B5/6c)uH1(B5/6c))12,\displaystyle\|u\|_{L^{\infty}(B_{1}^{c})}\leq C\|u\|_{H^{1}(B_{5/6}^{c})}\left(1+\ln\frac{\|u\|_{H^{2}(B_{5/6}^{c})}}{\|u\|_{H^{1}(B_{5/6}^{c})}}\right)^{\frac{1}{2}},
uL(A1,1/2)2CuH1(2)(u,/ u)L2(2)(1+lnuH2(2)(u,/ u)L2(2)).\displaystyle\|u\|_{L^{\infty}(A_{1,1/2})}^{2}\leq C\|u\|_{H^{1}(\mathbb{R}^{2})}\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(\mathbb{R}^{2})}}{\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}}\right).

The first inequality originally appeared in [1] with generalizations in [2]. For our needs, we will use the second inequality for the region near the axis |x|=0|x|=0 (for fixed time tt), while the third inequality will be applied on the region far away from the light cone. Near the light cone, the better decay of the angular derivative plays a crucial role. And we rely on the fourth inequality to prove the decay estimate for the solution. For readers’ interests, we will reprove the first inequality, based on which we show the rest ones needed in this paper.

Proof.

We define the Fourier transform as

u^(ξ)=12π2u(x)eixξ𝑑x,ξ2.\displaystyle\hat{u}(\xi)=\frac{1}{2\pi}\int_{\mathbb{R}^{2}}u(x)e^{-ix\cdot\xi}dx,\quad\xi\in\mathbb{R}^{2}.

In particular we have reverse transform

u(x)=12π2u^(ξ)eixξ𝑑ξ.\displaystyle u(x)=\frac{1}{2\pi}\int_{\mathbb{R}^{2}}\hat{u}(\xi)e^{ix\cdot\xi}d\xi.

By definition, we recall that

uHk(2)=(1+|ξ|2)k2u^L2(2),k=0,1,2.\displaystyle\|{u}\|_{H^{k}(\mathbb{R}^{2})}=\|(1+|\xi|^{2})^{\frac{k}{2}}\hat{u}\|_{L^{2}(\mathbb{R}^{2})},\quad k=0,1,2.

For a fixed constant R>1,R>1, we have

uL(2)\displaystyle\|{u}\|_{L^{\infty}(\mathbb{R}^{2})} 12πu^L1(2)\displaystyle\leq\frac{1}{2\pi}\|\hat{u}\|_{L^{1}(\mathbb{R}^{2})}
(1+|ξ|2)12(R+|ξ|2)12u^L2(2)(1+|ξ|2)12(R+|ξ|2)12L2(2).\displaystyle\leq\|(1+|\xi|^{2})^{\frac{1}{2}}(R+|\xi|^{2})^{\frac{1}{2}}\hat{u}\|_{L^{2}(\mathbb{R}^{2})}\|(1+|\xi|^{2})^{-\frac{1}{2}}(R+|\xi|^{2})^{-\frac{1}{2}}\|_{L^{2}(\mathbb{R}^{2})}.

We now compute that

(1+|ξ|2)12(R+|ξ|2)12L2(2)2\displaystyle\|(1+|\xi|^{2})^{-\frac{1}{2}}(R+|\xi|^{2})^{-\frac{1}{2}}\|_{L^{2}(\mathbb{R}^{2})}^{2} =2(1+|ξ|2)1(R+|ξ|2)1𝑑ξ\displaystyle=\int_{\mathbb{R}^{2}}(1+|\xi|^{2})^{-1}(R+|\xi|^{2})^{-1}d\xi
=0+2π(1+r2)1(R+r2)1r𝑑r\displaystyle=\int_{0}^{+\infty}2\pi(1+r^{2})^{-1}(R+r^{2})^{-1}rdr
=πR1ln1+r2R+r2|r=0+=πlnRR1.\displaystyle=\frac{\pi}{R-1}\ln\frac{1+r^{2}}{R+r^{2}}\Big{|}_{r=0}^{+\infty}=\frac{\pi\ln R}{R-1}.

Moreover we estimate that

(1+|ξ|2)12(R+|ξ|2)12u^L2(2)2\displaystyle\|(1+|\xi|^{2})^{\frac{1}{2}}(R+|\xi|^{2})^{\frac{1}{2}}\hat{u}\|_{L^{2}(\mathbb{R}^{2})}^{2} =(1+|ξ|2)u^L2(2)2+(R1)(1+|ξ|2)12u^L2(2)2\displaystyle=\|(1+|\xi|^{2})\hat{u}\|_{L^{2}(\mathbb{R}^{2})}^{2}+(R-1)\|(1+|\xi|^{2})^{\frac{1}{2}}\hat{u}\|_{L^{2}(\mathbb{R}^{2})}^{2}
=uH2(2)2+(R1)uH1(2)2.\displaystyle=\|{u}\|_{H^{2}(\mathbb{R}^{2})}^{2}+(R-1)\|{u}\|_{H^{1}(\mathbb{R}^{2})}^{2}.

Therefore

uL(2)2(uH2(2)2+(R1)uH1(2)2)πlnRR1.\displaystyle\|{u}\|_{L^{\infty}(\mathbb{R}^{2})}^{2}\leq(\|{u}\|_{H^{2}(\mathbb{R}^{2})}^{2}+(R-1)\|{u}\|_{H^{1}(\mathbb{R}^{2})}^{2})\frac{\pi\ln R}{R-1}.

Now we take R=1+uH2(2)/uH1(2)R=1+\|{u}\|_{H^{2}(\mathbb{R}^{2})}/\|{u}\|_{H^{1}(\mathbb{R}^{2})}. Then

uL(2)22πuH1(2)2ln(1+uH2(2)/uH1(2)),\displaystyle\|{u}\|_{L^{\infty}(\mathbb{R}^{2})}^{2}\leq 2\pi\|{u}\|_{H^{1}(\mathbb{R}^{2})}^{2}\ln(1+\|{u}\|_{H^{2}(\mathbb{R}^{2})}/\|{u}\|_{H^{1}(\mathbb{R}^{2})}),

which implies the first inequality on the whole space 2\mathbb{R}^{2}.

Now for the second inequality on a finite region and the third inequality on the complement of finite region, fix an even smooth function ψ:[0,1]{\psi}:\mathbb{R}\to[0,1], which is supported on [1,1][-1,1] and equals to one in [56,56][-\frac{5}{6},\frac{5}{6}]. Then define smooth functions

η1(x):=ψ(|x|),η2(x):=ψ(43|x|)\eta_{1}(x):={\psi}(|x|),\quad\eta_{2}(x):={\psi}(\frac{4}{3}|x|)

such that η2\eta_{2} is supported in B34B_{\frac{3}{4}} and 1η11-\eta_{1} is supported in B56c¯\overline{B_{\frac{5}{6}}^{c}}. Therefore

η2uH1(2)CuH1(B3/4),η2uH2(2)CuH2(B3/4),\displaystyle\|\eta_{2}u\|_{H^{1}(\mathbb{R}^{2})}\leq C\|u\|_{H^{1}(B_{3/4})},\ \|\eta_{2}u\|_{H^{2}(\mathbb{R}^{2})}\leq C\|u\|_{H^{2}(B_{3/4})}, (7)
(1η1)uH1(2)CuH1(B5/6c),(1η1)uH2(2)CuH2(B5/6c).\displaystyle\|(1-\eta_{1})u\|_{H^{1}(\mathbb{R}^{2})}\leq C\|u\|_{H^{1}(B_{5/6}^{c})},\ \|(1-\eta_{1})u\|_{H^{2}(\mathbb{R}^{2})}\leq C\|u\|_{H^{2}(B_{5/6}^{c})}. (8)

Here C>1C>1 is a universal constant. As η2=1\eta_{2}=1 in B1/2B_{1/2}, by (7) and the above logarithmic Sobolev inequality, we have

uL(B1/2)\displaystyle\|u\|_{L^{\infty}(B_{1/2})} η2uL(2)\displaystyle\leq\|\eta_{2}u\|_{L^{\infty}(\mathbb{R}^{2})}
Cη2uH1(2)(1+lnη2uH2(2)η2uH1(2))12\displaystyle\leq C\|\eta_{2}u\|_{H^{1}(\mathbb{R}^{2})}\left(1+\ln\frac{\|\eta_{2}u\|_{H^{2}(\mathbb{R}^{2})}}{\|\eta_{2}u\|_{H^{1}(\mathbb{R}^{2})}}\right)^{\frac{1}{2}}
Cη2uH1(2)(1+lnuH2(B3/4)η2uH1(2))12\displaystyle\leq C\|\eta_{2}u\|_{H^{1}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(B_{3/4})}}{\|\eta_{2}u\|_{H^{1}(\mathbb{R}^{2})}}\right)^{\frac{1}{2}}
CuH1(B3/4)(1+lnuH2(B3/4)uH1(B3/4))12,\displaystyle\leq C\|u\|_{H^{1}(B_{3/4})}\left(1+\ln\frac{\|u\|_{H^{2}(B_{3/4})}}{\|u\|_{H^{1}(B_{3/4})}}\right)^{\frac{1}{2}},

which implies the second logarithmic Sobolev inequality. Here the last step follows from the fact that δ(1+ln(1/δ))12\delta\cdot(1+\ln(1/\delta))^{\frac{1}{2}} is increasing for δ(0,1].\delta\in(0,1]. The third Sobolev inequality on the complement of finite region follows in a similar way in view of uL(B1c)(1η1)uL(2)\|u\|_{L^{\infty}(B_{1}^{c})}\leq\|(1-\eta_{1})u\|_{L^{\infty}(\mathbb{R}^{2})}.

Finally for the improved Sobolev inequality on annulus, let

u~(r,θ)=u(rcosθ,rsinθ).\widetilde{u}(r,\theta)=u(r\cos\theta,r\sin\theta).

Then

u~Hk([1/4,3]×[π/2,π/2])CuHk(2),k=0,1,2,\displaystyle\|\widetilde{u}\|_{H^{k}([1/4,3]\times[-\pi/2,\pi/2])}\leq C\|u\|_{H^{k}(\mathbb{R}^{2})},\quad k=0,1,2,
θu~L2([1/4,3]×[π/2,π/2])C/ uL2(2).\displaystyle\|\partial_{\theta}\widetilde{u}\|_{L^{2}([1/4,3]\times[-\pi/2,\pi/2])}\leq C\|\mbox{$\nabla\mkern-13.0mu/$\,}u\|_{L^{2}(\mathbb{R}^{2})}.

Now we define functions

η3(r,θ):=ψ(r5/4)ψ(2θ/π),u1(r,θ)=η3(r,θ)u~(r,θ)\eta_{3}(r,\theta):={\psi}(r-5/4){\psi}(2\theta/\pi),\quad u_{1}(r,\theta)=\eta_{3}(r,\theta)\widetilde{u}(r,\theta)

on 2\mathbb{R}^{2} with Cartesian coordinates r,θr,\theta\in\mathbb{R} (not polar coordinates for u1u_{1}). In particular η3\eta_{3} is smooth and supported in [1/4,3]×[π/2,π/2][1/4,3]\times[-\pi/2,\pi/2]. Moreover η3\eta_{3} is constant 11 in [1/2,3/2]×[π/3,π/3][1/2,3/2]\times[-\pi/3,\pi/3]. These properties lead to the following relations

u1Hk(2)Cu~Hk([1/4,3]×[π/2,π/2])CuHk(2),k=0,1,2,\displaystyle\|{u}_{1}\|_{H^{k}(\mathbb{R}^{2})}\leq C\|\widetilde{u}\|_{H^{k}([1/4,3]\times[-\pi/2,\pi/2])}\leq C\|u\|_{H^{k}(\mathbb{R}^{2})},\ k=0,1,2, (9)
(u1,θu1)L2(2)C(u~,θu~)L2([1/4,3]×[π/2,π/2])C(u,/ u)L2(2).\displaystyle\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})}\leq C\|(\widetilde{u},\partial_{\theta}\widetilde{u})\|_{L^{2}([1/4,3]\times[-\pi/2,\pi/2])}\leq C\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}. (10)

For fixed constant R1,R\geq 1, let uR(r,θ)=u1(r/R,θR)u_{R}(r,\theta)=u_{1}(r/R,\theta R). We compute that

ruRL2(2)\displaystyle\|\partial_{r}{u}_{R}\|_{L^{2}(\mathbb{R}^{2})} =R1ru1L2(2),θuRL2(2)=Rθu1L2(2),\displaystyle=R^{-1}\|\partial_{r}{u}_{1}\|_{L^{2}(\mathbb{R}^{2})},\quad\|\partial_{\theta}{u}_{R}\|_{L^{2}(\mathbb{R}^{2})}=R\|\partial_{\theta}{u}_{1}\|_{L^{2}(\mathbb{R}^{2})},
R1u1H1(2)\displaystyle R^{-1}\|{u}_{1}\|_{H^{1}(\mathbb{R}^{2})} uRH1(2)R1u1H1(2)+R(u1,θu1)L2(2),\displaystyle\leq\|{u}_{R}\|_{H^{1}(\mathbb{R}^{2})}\leq R^{-1}\|{u}_{1}\|_{H^{1}(\mathbb{R}^{2})}+R\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})},
uRH2(2)\displaystyle\|{u}_{R}\|_{H^{2}(\mathbb{R}^{2})} R2u1H2(2),uRL2(2)=u1L2(2).\displaystyle\leq R^{2}\|{u}_{1}\|_{H^{2}(\mathbb{R}^{2})},\quad\|{u}_{R}\|_{L^{2}(\mathbb{R}^{2})}=\|{u}_{1}\|_{L^{2}(\mathbb{R}^{2})}.

As a compactly supported function on 2\mathbb{R}^{2}, by using the logarithmic Sobolev embedding on the whole space, we have

u1L(2)\displaystyle\|u_{1}\|_{L^{\infty}(\mathbb{R}^{2})} =uRL(2)\displaystyle=\|u_{R}\|_{L^{\infty}(\mathbb{R}^{2})}
CuRH1(2)(1+lnuRH2(2)uRH1(2))12\displaystyle\leq C\|u_{R}\|_{H^{1}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u_{R}\|_{H^{2}(\mathbb{R}^{2})}}{\|u_{R}\|_{H^{1}(\mathbb{R}^{2})}}\right)^{\frac{1}{2}}
C(R1u1H1(2)+R(u1,θu1)L2(2))(1+lnR2u1H2(2)R1u1H1(2))12,\displaystyle\leq C(R^{-1}\|{u}_{1}\|_{H^{1}(\mathbb{R}^{2})}+R\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})})\left(1+\ln\frac{R^{2}\|u_{1}\|_{H^{2}(\mathbb{R}^{2})}}{R^{-1}\|u_{1}\|_{H^{1}(\mathbb{R}^{2})}}\right)^{\frac{1}{2}},

Now by taking R=u1H1(2)12(u1,θu1)L2(2)12R=\|{u}_{1}\|_{H^{1}(\mathbb{R}^{2})}^{\frac{1}{2}}\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})}^{-\frac{1}{2}}, we obtain that

R1u1H1(2)=R(u1,θu1)L2(2)=u1H1(2)12(u1,θu1)L2(2)12,\displaystyle R^{-1}\|{u}_{1}\|_{H^{1}(\mathbb{R}^{2})}=R\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})}=\|{u}_{1}\|_{H^{1}(\mathbb{R}^{2})}^{\frac{1}{2}}\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})}^{\frac{1}{2}},
lnR2u1H2(2)R1u1H1(2)32lnR2u1H2(2)u1H1(2)=32lnu1H2(2)(u1,θu1)L2(2).\displaystyle\ln\frac{R^{2}\|u_{1}\|_{H^{2}(\mathbb{R}^{2})}}{R^{-1}\|u_{1}\|_{H^{1}(\mathbb{R}^{2})}}\leq\frac{3}{2}\ln\frac{R^{2}\|u_{1}\|_{H^{2}(\mathbb{R}^{2})}}{\|u_{1}\|_{H^{1}(\mathbb{R}^{2})}}=\frac{3}{2}\ln\frac{\|u_{1}\|_{H^{2}(\mathbb{R}^{2})}}{\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})}}.

Therefore from the previous inequality and estimates (9), (10), we conclude that

u1L(2)2\displaystyle\|u_{1}\|_{L^{\infty}(\mathbb{R}^{2})}^{2} Cu1H1(2)(u1,θu1)L2(2)(1+lnu1H2(2)(u1,θu1)L2(2))\displaystyle\leq C\|{u}_{1}\|_{H^{1}(\mathbb{R}^{2})}\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u_{1}\|_{H^{2}(\mathbb{R}^{2})}}{\|({u}_{1},\partial_{\theta}u_{1})\|_{L^{2}(\mathbb{R}^{2})}}\right)
CuH1(2)(u,/ u)L2(2)(1+lnuH2(2)(u,/ u)L2(2)).\displaystyle\leq C\|{u}\|_{H^{1}(\mathbb{R}^{2})}\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(\mathbb{R}^{2})}}{\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}}\right).

As η3=1\eta_{3}=1 in [12,32]×[π3,π3][\frac{1}{2},\frac{3}{2}]\times[-\frac{\pi}{3},\frac{\pi}{3}], u1=η3u~u_{1}=\eta_{3}\widetilde{u}, we therefore have that

u~L([12,32]×[π3,π3])2\displaystyle\|\widetilde{u}\|_{L^{\infty}([\frac{1}{2},\frac{3}{2}]\times[-\frac{\pi}{3},\frac{\pi}{3}])}^{2} =u1L(2)2\displaystyle=\|u_{1}\|_{L^{\infty}(\mathbb{R}^{2})}^{2}
CuH1(2)(u,/ u)L2(2)(1+lnuH2(2)(u,/ u)L2(2)).\displaystyle\leq C\|{u}\|_{H^{1}(\mathbb{R}^{2})}\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(\mathbb{R}^{2})}}{\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}}\right).

By symmetry (or considering u(a)(rcosθ,rsinθ)=u(rcos(θ+a),rsin(θ+a))u_{(a)}(r\cos\theta,r\sin\theta)=u(r\cos(\theta+a),r\sin(\theta+a))), we also have

u~L([12,32]×[aπ3,a+π3])2\displaystyle\|\widetilde{u}\|_{L^{\infty}([\frac{1}{2},\frac{3}{2}]\times[a-\frac{\pi}{3},a+\frac{\pi}{3}])}^{2} CuH1(2)(u,/ u)L2(2)(1+lnuH2(2)(u,/ u)L2(2)),a.\displaystyle\leq C\|{u}\|_{H^{1}(\mathbb{R}^{2})}\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(\mathbb{R}^{2})}}{\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}}\right),\ \forall\ a\in\mathbb{R}.

In particular we derive that

u~L([12,32]×)2\displaystyle\|\widetilde{u}\|_{L^{\infty}([\frac{1}{2},\frac{3}{2}]\times\mathbb{R})}^{2} CuH1(2)(u,/ u)L2(2)(1+lnuH2(2)(u,/ u)L2(2)),\displaystyle\leq C\|{u}\|_{H^{1}(\mathbb{R}^{2})}\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}\left(1+\ln\frac{\|u\|_{H^{2}(\mathbb{R}^{2})}}{\|(u,\mbox{$\nabla\mkern-13.0mu/$\,}u)\|_{L^{2}(\mathbb{R}^{2})}}\right),

which implies the fourth logarithmic Sobolev inequality as x=(rcosθ,rsinθ)A1,1/2x=(r\cos\theta,r\sin\theta)\in A_{1,1/2} is equivalent to (r,θ)[12,32]×[0,2π](r,\theta)\in[\frac{1}{2},\frac{3}{2}]\times[0,2\pi] with u~(r,θ)=u(rcosθ,rsinθ)\widetilde{u}(r,\theta)=u(r\cos\theta,r\sin\theta). ∎

To apply the above logarithmic Sobolev inequality, we use the potential energy decay of the previous section to derive the following necessary bounds on fixed time tt by using the conformal energy identity.

Proposition 4.1.

The solution ϕ\phi to (1) verifies the following estimates for all ϵ>0\epsilon>0

(1+t)2/ ϕL22+(1+|tr|)ϕL22+ϕL22\displaystyle(1+t)^{2}\|\mbox{$\nabla\mkern-13.0mu/$\,}\phi\|_{L^{2}}^{2}+\|(1+|t-r|)\nabla\phi\|_{L^{2}}^{2}+\|\phi\|_{L^{2}}^{2} Cϵ(1+t)5p2,\displaystyle\leq C_{\epsilon}(1+t)^{\frac{5-p}{2}}, (11)
ϕH2\displaystyle\|\phi\|_{H^{2}} Cϵ(1+t)1+ϵ\displaystyle\leq C_{\epsilon}(1+t)^{1+\epsilon} (12)

for some constant CϵC_{\epsilon} depending on ϵ\epsilon, the initial conformal energy 0,2\mathcal{E}_{0,2} and the first order energy 1,0\mathcal{E}_{1,0}.

Proof.

The proof relies on the conformal energy identity

Q0(t)+2p+12(t2+r2)|ϕ(t,x)|p+1𝑑x+p5p+1st2τ2|ϕ(τ,x)|p+1𝑑x𝑑τ\displaystyle Q_{0}(t)+\frac{2}{p+1}\int_{\mathbb{R}^{2}}(t^{2}+r^{2})|\phi(t,x)|^{p+1}dx+\frac{p-5}{p+1}\int_{s}^{t}2\tau\int_{\mathbb{R}^{2}}|\phi(\tau,x)|^{p+1}dxd\tau
=Q0(s)+2p+12(s2+|x|2)|ϕ(s,x)|p+1𝑑x,\displaystyle=Q_{0}(s)+\frac{2}{p+1}\int_{\mathbb{R}^{2}}(s^{2}+|x|^{2})|\phi(s,x)|^{p+1}dx,

which can be obtained by using the conformal Killing vector field as multiplier (see for example Lemma 1 in [9]). Here

Q0(t)=2|x12ϕx21ϕ|2+|Sϕ+ϕ|2+j=12|xjtϕ+tjϕ|2dx\displaystyle Q_{0}(t)=\int_{\mathbb{R}^{2}}|x_{1}\partial_{2}\phi-x_{2}\partial_{1}\phi|^{2}+|S\phi+\phi|^{2}+\sum\limits_{j=1}^{2}|x_{j}\partial_{t}\phi+t\partial_{j}\phi|^{2}dx

with S=tt+x11+x22S=t\partial_{t}+x_{1}\partial_{1}+x_{2}\partial_{2} the scaling vector field.

Now by setting s=0s=0 and using the time decay estimate of Proposition 3.1, we derive that

Q0(t)\displaystyle Q_{0}(t) C0,2+C0,20ts(1+s)p12𝑑s\displaystyle\leq C\mathcal{E}_{0,2}+C\mathcal{E}_{0,2}\int_{0}^{t}s(1+s)^{-\frac{p-1}{2}}ds
C0,2(1+t)5p2.\displaystyle\leq C\mathcal{E}_{0,2}(1+t)^{\frac{5-p}{2}}.

Here and in the following CC denotes a universal constant. On the other hand, we can estimate that

2(t2+r2)|/ ϕ|2+|trϕ+rtϕ|2+|Sϕ+ϕ|2dxCQ0(t)C0,2(1+t)5p2.\displaystyle\int_{\mathbb{R}^{2}}(t^{2}+r^{2})|\mbox{$\nabla\mkern-13.0mu/$\,}\phi|^{2}+|t\partial_{r}\phi+r\partial_{t}\phi|^{2}+|S\phi+\phi|^{2}dx\leq CQ_{0}(t)\leq C\mathcal{E}_{0,2}(1+t)^{\frac{5-p}{2}}. (13)

To control the L2L^{2} norm of ϕ\phi, we note that

2(ttϕ+rrϕ+ϕ)ϕ𝑑x=2ttϕϕdx=t2t2|ϕ|2𝑑x.\displaystyle\int_{\mathbb{R}^{2}}(t\partial_{t}\phi+r\partial_{r}\phi+\phi)\phi dx=\int_{\mathbb{R}^{2}}t\partial_{t}\phi\cdot\phi dx=\frac{t}{2}\partial_{t}\int_{\mathbb{R}^{2}}|\phi|^{2}dx.

Therefore

t2ddtϕ(t)L22=2(ttϕ+rrϕ+ϕ)ϕ𝑑xϕ(t)L2(ttϕ+rrϕ+rϕ)(t)L2.\displaystyle\frac{t}{2}\frac{d}{dt}\|\phi(t)\|_{L^{2}}^{2}=\int_{\mathbb{R}^{2}}(t\partial_{t}\phi+r\partial_{r}\phi+\phi)\phi dx\leq\|\phi(t)\|_{L^{2}}\|(t\partial_{t}\phi+r\partial_{r}\phi+r\phi)(t)\|_{L^{2}}.

By using the bound

2|Sϕ+ϕ|2𝑑xQ0(t)C0,2(1+t)5p2,\displaystyle\int_{\mathbb{R}^{2}}|S\phi+\phi|^{2}dx\leq Q_{0}(t)\leq C\mathcal{E}_{0,2}(1+t)^{\frac{5-p}{2}},

we conclude that

ϕ(t)L2ϕ(1)L2+C0,2(t+1)5p4,t1.\displaystyle\|\phi(t)\|_{L^{2}}\leq\|\phi(1)\|_{L^{2}}+C\sqrt{\mathcal{E}_{0,2}}(t+1)^{\frac{5-p}{4}},\quad\forall\ t\geq 1.

We also have

ϕ(s)L2ϕ(0)L2+01tϕ(t)L2𝑑t,s[0,1]\displaystyle\|\phi(s)\|_{L^{2}}\leq\|\phi(0)\|_{L^{2}}+\int_{0}^{1}\|\partial_{t}\phi(t)\|_{L^{2}}dt,\quad\forall\ s\in[0,1]

with

ϕ(0)L2=ϕ0L2C0,2\|\phi(0)\|_{L^{2}}=\|\phi_{0}\|_{L^{2}}\leq C\sqrt{\mathcal{E}_{0,2}}

by using Hardy’s inequality. This together with the standard energy conservation

2|tϕ|2+|ϕ|2+2p+1|ϕ|p+1dx=0,0,\displaystyle\int_{\mathbb{R}^{2}}|\partial_{t}\phi|^{2}+|\nabla\phi|^{2}+\frac{2}{p+1}|\phi|^{p+1}dx=\mathcal{E}_{0,0}, (14)

leads to

ϕ(t)L2C0,2(t+1)5p4,t0.\displaystyle\|\phi(t)\|_{L^{2}}\leq C\sqrt{\mathcal{E}_{0,2}}(t+1)^{\frac{5-p}{4}},\ \forall\ t\geq 0. (15)

Therefore we can show that

2(tr)2|ϕ|2𝑑x\displaystyle\int_{\mathbb{R}^{2}}(t-r)^{2}|\nabla\phi|^{2}dx
C2|(t+r)(t+r)ϕ+ϕ|2+|(tr)(tr)ϕ+ϕ|2+(t2+r2)|/ ϕ|2+|ϕ|2dx\displaystyle\leq C\int_{\mathbb{R}^{2}}|(t+r)(\partial_{t}+\partial_{r})\phi+\phi|^{2}+|(t-r)(\partial_{t}-\partial_{r})\phi+\phi|^{2}+(t^{2}+r^{2})|\mbox{$\nabla\mkern-13.0mu/$\,}\phi|^{2}+|\phi|^{2}dx
C2(t2+r2)|/ ϕ|2+|trϕ+rtϕ|2+|Sϕ+ϕ|2+|ϕ|2dx\displaystyle\leq C\int_{\mathbb{R}^{2}}(t^{2}+r^{2})|\mbox{$\nabla\mkern-13.0mu/$\,}\phi|^{2}+|t\partial_{r}\phi+r\partial_{t}\phi|^{2}+|S\phi+\phi|^{2}+|\phi|^{2}dx
C0,2(1+t)5p2.\displaystyle\leq C\mathcal{E}_{0,2}(1+t)^{\frac{5-p}{2}}.

Estimate (11) then follows from the above bounds (13), (14), (15).


Finally for the H2H^{2} estimate (12) for ϕ\phi, note that

ϕH1C0,2(1+t)5p4+C0,0C0,2(1+t)\displaystyle\|\phi\|_{H^{1}}\leq C\sqrt{\mathcal{E}_{0,2}}(1+t)^{\frac{5-p}{4}}+C\sqrt{\mathcal{E}_{0,0}}\leq C\sqrt{\mathcal{E}_{0,2}}(1+t)

as p>1p>1. It remains to estimate 2ϕL2\|\nabla^{2}\phi\|_{L^{2}}. For any 0<ϵ<p1p+10<\epsilon<\frac{p-1}{p+1}, by using Gagliardo-Nirenberg inequality, we can show that

|ϕ||ϕ|p1L2\displaystyle\||\nabla\phi|\cdot|\phi|^{p-1}\|_{L^{2}} ϕL2+ϵϕL2(2+ϵ)(p1)ϵp1\displaystyle\leq\|\nabla\phi\|_{L^{2+\epsilon}}\|\phi\|_{L^{\frac{2(2+\epsilon)(p-1)}{\epsilon}}}^{p-1}
CϵϕL222+ϵ2ϕL2ϵ2+ϵϕLp+1(p+1)ϵ2(2+ϵ)ϕL2p1(p+1)ϵ2(2+ϵ)\displaystyle\leq C_{\epsilon}\|\nabla\phi\|_{L^{2}}^{\frac{2}{2+\epsilon}}\|\nabla^{2}\phi\|_{L^{2}}^{\frac{\epsilon}{2+\epsilon}}\|\phi\|_{L^{p+1}}^{\frac{(p+1)\epsilon}{2(2+\epsilon)}}\|\nabla\phi\|_{L^{2}}^{p-1-\frac{(p+1)\epsilon}{2(2+\epsilon)}}
Cϵ0,0p2(p+1)ϵ4(2+ϵ)2ϕL2ϵ2+ϵ.\displaystyle\leq C_{\epsilon}\mathcal{E}_{0,0}^{\frac{p}{2}-\frac{(p+1)\epsilon}{4(2+\epsilon)}}\|\nabla^{2}\phi\|_{L^{2}}^{\frac{\epsilon}{2+\epsilon}}.

Here we have used the energy conservation to control ϕL2\|\nabla\phi\|_{L^{2}} and ϕLp+1\|\phi\|_{L^{p+1}}. Then in view of energy estimate we derive that

2ϕL2\displaystyle\|\nabla^{2}\phi\|_{L^{2}} 1,0+p0t|ϕ||ϕ|p1L2𝑑s\displaystyle\leq\sqrt{\mathcal{E}_{1,0}}+p\int_{0}^{t}\||\nabla\phi|\cdot|\phi|^{p-1}\|_{L^{2}}ds
1,0+Cϵ0,0p2(p+1)ϵ4(2+ϵ)0t2ϕL2ϵ2+ϵ𝑑s,\displaystyle\leq\sqrt{\mathcal{E}_{1,0}}+C_{\epsilon}\mathcal{E}_{0,0}^{\frac{p}{2}-\frac{(p+1)\epsilon}{4(2+\epsilon)}}\int_{0}^{t}\|\nabla^{2}\phi\|_{L^{2}}^{\frac{\epsilon}{2+\epsilon}}ds,

from which we derive that

2ϕL2C1,0+Cϵ0,0p2+(p1)ϵ8(1+t)1+ϵ2.\displaystyle\|\nabla^{2}\phi\|_{L^{2}}\leq C\sqrt{\mathcal{E}_{1,0}}+C_{\epsilon}\mathcal{E}_{0,0}^{\frac{p}{2}+\frac{(p-1)\epsilon}{8}}(1+t)^{1+\frac{\epsilon}{2}}.

Replacing ϵ\epsilon by 2ϵ2\epsilon, we conclude estimate (12).

As a consequence of the above Proposition and the logarithmic Sobolev inequality, we derive the pointwise decay estimate for the solution for all 1<p<51<p<5.

Proposition 4.2.

For any ϵ>0\epsilon>0, the solution ϕ\phi of (1) verifies the following pointwise decay estimate

|ϕ|Cϵ(1+t+|x|)p18+ϵ\displaystyle|\phi|\leq C_{\epsilon}(1+t+|x|)^{-\frac{p-1}{8}+\epsilon}

for some constant CϵC_{\epsilon} depending on ϵ\epsilon, the initial conformal energy 0,2\mathcal{E}_{0,2} and the initial first order energy 1,0\mathcal{E}_{1,0}.

The proof also implies that the solution decays better away from the light cone, that is,

|ϕ|Cϵ(1+t)p14+ϵ,|tr|t2.\displaystyle|\phi|\leq C_{\epsilon}(1+t)^{-\frac{p-1}{4}+\epsilon},\quad|t-r|\geq\frac{t}{2}.
Proof.

The estimate trivially holds when t+|x|10t+|x|\leq 10. In the sequel, let’s assume without loss of generality that t+|x|>10t+|x|>10. Divide 2\mathbb{R}^{2} into three regions: Bt2B_{\frac{t}{2}}, B3t2cB_{\frac{3t}{2}}^{c}, {t2|x|3t2}\{\frac{t}{2}\leq|x|\leq\frac{3t}{2}\}. Here recall that BRB_{R} denotes the spatial ball with radius RR.

On region Bt2B_{\frac{t}{2}}, for fixed t>1t>1, define

ϕ~(x)=ϕ(t,tx),forx2.\displaystyle\tilde{\phi}(x)=\phi(t,tx),\ \text{for}\ x\in\mathbb{R}^{2}. (16)

In view of the decay estimate (11), we conclude that

|x|3t/4|ϕ|2𝑑x42t2|x|t(tr)2|ϕ|2𝑑x(1+t)1p2.\displaystyle\int_{|x|\leq 3t/4}|\nabla\phi|^{2}dx\leq 4^{2}t^{-2}\int_{|x|\leq t}(t-r)^{2}|\nabla\phi|^{2}dx\lesssim(1+t)^{\frac{1-p}{2}}.

Thus we can compute that

|x|3/4|ϕ~|2+|ϕ~|2dxC|x|3t/4(t2|ϕ|2+|ϕ|2)𝑑x(1+t)1p2.\displaystyle\int_{|x|\leq 3/4}|\tilde{\phi}|^{2}+|\nabla\tilde{\phi}|^{2}dx\leq C\int_{|x|\leq 3t/4}(t^{-2}|\phi|^{2}+|\nabla\phi|^{2})dx\lesssim(1+t)^{\frac{1-p}{2}}.

Similarly for the second order derivative

|x|3/4|2ϕ~|2𝑑x=|x|3t/4t2|2ϕ|2𝑑xt2+(1+t)4+ϵ(1+t)4+ϵ.\displaystyle\int_{|x|\leq 3/4}|\nabla^{2}\tilde{\phi}|^{2}dx=\int_{|x|\leq 3t/4}t^{2}|\nabla^{2}\phi|^{2}dx\lesssim t^{2}+(1+t)^{4+\epsilon}\lesssim(1+t)^{4+\epsilon}.

Hence from the logarithmic Sobolev embedding of Lemma 4.1, we derive that on the region {|x|t2}\{|x|\leq\frac{t}{2}\}, the solution satisfies

|ϕ||ϕ~|(1+t)1p4+ϵ2ln(Cϵt)(1+t)1p4+ϵ.\displaystyle|\phi|\leq|\tilde{\phi}|\lesssim(1+t)^{\frac{1-p}{4}+\frac{\epsilon}{2}}\sqrt{\ln(C_{\epsilon}t)}\lesssim(1+t)^{\frac{1-p}{4}+\epsilon}.

Here by using our notation the implicit constant CϵC_{\epsilon} relies on ϵ\epsilon, the initial data 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}.

On the region B3t2B_{\frac{3t}{2}}, fix a point (t,x0)(t,x_{0}) with r0=|x0|3t2r_{0}=|x_{0}|\geq\frac{3t}{2}. Define

ψ(x)=ϕ(t,r0x),|x|1.\displaystyle\psi(x)=\phi(t,r_{0}x),\quad|x|\geq 1.

Similarly we can show that

|x|5/6|ψ|2+|ψ|2\displaystyle\int_{|x|\geq 5/6}|\psi|^{2}+|\nabla\psi|^{2} 62r022|ϕ|2+|(rt)ϕ|2dx\displaystyle\leq 6^{2}r_{0}^{-2}\int_{\mathbb{R}^{2}}|\phi|^{2}+|(r-t)\nabla\phi|^{2}dx
r02(1+t)5p2\displaystyle\lesssim r_{0}^{-2}(1+t)^{\frac{5-p}{2}}
(1+r0)1p2.\displaystyle\lesssim(1+r_{0})^{\frac{1-p}{2}}.

Here as we have assumed that t+|x0|10t+|x_{0}|\geq 10 and |x0|32t|x_{0}|\geq\frac{3}{2}t. For the second order derivative of ψ\psi, we have

|x|5/6|2ψ|2𝑑x=r02|x|5r0/6|2ϕ|2𝑑xr02+(1+r0)4+ϵ(1+r0)4+ϵ.\displaystyle\int_{|x|\geq 5/6}|\nabla^{2}\psi|^{2}dx=r_{0}^{2}\int_{|x|\geq 5r_{0}/6}|\nabla^{2}\phi|^{2}dx\lesssim r_{0}^{2}+(1+r_{0})^{4+\epsilon}\lesssim(1+r_{0})^{4+\epsilon}.

Therefore according the logarithmic Sobolev embedding, we derive that

|ϕ(t,x0)|(1+r0)1p4+ϵ,r0=|x0|32t.\displaystyle|\phi(t,x_{0})|\lesssim(1+r_{0})^{\frac{1-p}{4}+\epsilon},\quad r_{0}=|x_{0}|\geq\frac{3}{2}t.

Finally on the region near the light cone, consider the annulus At,t/2A_{t,t/2}. For fixed t>2t>2, we still define ϕ~\widetilde{\phi} as in (16). Therefore in view of the estimates in Proposition 4.1, we can show that

ϕ~L2(2)=t1ϕ(t)L2(2)\displaystyle\|\widetilde{\phi}\|_{L^{2}(\mathbb{R}^{2})}=t^{-1}\|{\phi}(t)\|_{L^{2}(\mathbb{R}^{2})} (1+t)1p4,\displaystyle\lesssim(1+t)^{\frac{1-p}{4}},
ϕ~L2(2)=ϕ(t)L2(2)\displaystyle\|\nabla\widetilde{\phi}\|_{L^{2}(\mathbb{R}^{2})}=\|\nabla{\phi}(t)\|_{L^{2}(\mathbb{R}^{2})} 1,\displaystyle\lesssim 1,
/ ϕ~L2(2)=/ ϕ(t)L2(2)\displaystyle\|\mbox{$\nabla\mkern-13.0mu/$\,}\widetilde{\phi}\|_{L^{2}(\mathbb{R}^{2})}=\|\mbox{$\nabla\mkern-13.0mu/$\,}{\phi}(t)\|_{L^{2}(\mathbb{R}^{2})} (1+t)1p4,\displaystyle\lesssim(1+t)^{\frac{1-p}{4}},
ϕ~H2(2)tϕ(t)H2(2)\displaystyle\|\widetilde{\phi}\|_{H^{2}(\mathbb{R}^{2})}\leq t\|{\phi}(t)\|_{H^{2}(\mathbb{R}^{2})} t+(1+t)2+ϵ2(1+t)2+ϵ2.\displaystyle\lesssim t+(1+t)^{2+\frac{\epsilon}{2}}\lesssim(1+t)^{2+\frac{\epsilon}{2}}.

Therefore by using the logarithmic Sobolev embedding adapted to the annulus A1,1/2A_{1,1/2} for ϕ~\widetilde{\phi}, we derive that

ϕ(t)L(At,t/2)2\displaystyle\|\phi(t)\|_{L^{\infty}(A_{t,t/2})}^{2} =ϕ~L(A1,1/2)2\displaystyle=\|\widetilde{\phi}\|_{L^{\infty}(A_{1,1/2})}^{2}
Cϕ~H1(2)(ϕ~,/ ϕ~)L2(2)(1+lnϕ~H2(2)(ϕ~,/ ϕ~)L2(2))\displaystyle\leq C\|\widetilde{\phi}\|_{H^{1}(\mathbb{R}^{2})}\|(\widetilde{\phi},\mbox{$\nabla\mkern-13.0mu/$\,}\widetilde{\phi})\|_{L^{2}(\mathbb{R}^{2})}\left(1+\ln\frac{\|\widetilde{\phi}\|_{H^{2}(\mathbb{R}^{2})}}{\|(\widetilde{\phi},\mbox{$\nabla\mkern-13.0mu/$\,}\widetilde{\phi})\|_{L^{2}(\mathbb{R}^{2})}}\right)
(1+t)1p4(1+lnt)\displaystyle\lesssim(1+t)^{\frac{1-p}{4}}\left(1+\ln t\right)
Cϵ(1+t)1p4+2ϵ.\displaystyle\leq C_{\epsilon}(1+t)^{\frac{1-p}{4}+2\epsilon}.

By our notation the implicit constant as well as CϵC_{\epsilon} depends on ϵ>0\epsilon>0, the initial data 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}. We thus finished the proof for Proposition 4.2. ∎

As a corollary of the above pointwise decay estimate, the solution scatters in energy space when p>251p>2\sqrt{5}-1.

Corollary 4.1.

When p>251p>2\sqrt{5}-1, the solution ϕ\phi to (1) verifies the following spacetime bound

ϕLtpLx2pC\displaystyle\|\phi\|_{L_{t}^{p}L_{x}^{2p}}\leq C

for some constant CC depending on pp, the initial data 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}. Consequently the solution scatters in energy space, that is, there exists pairs (ϕ0±(x),ϕ1±(x))(\phi_{0}^{\pm}(x),\phi_{1}^{\pm}(x)) such that

limt±ϕ(t,x)𝐋(t)(ϕ0±(x),ϕ1±(x))Lx2=0.\displaystyle\lim\limits_{t\rightarrow\pm\infty}\|\partial\phi(t,x)-\partial\mathbf{L}(t)(\phi_{0}^{\pm}(x),\phi_{1}^{\pm}(x))\|_{L_{x}^{2}}=0.

Here 𝐋(t)\mathbf{L}(t) is the linear evolution operator for wave equation.

Proof.

Without loss of generality we only consider the spacetime norm in the future t0t\geq 0. In view of the potential energy decay estimate of Proposition 3.1 as well as the pointwise decay estimate of the above proposition, we can show that

ϕLtpLx2pp\displaystyle\|\phi\|_{L_{t}^{p}L_{x}^{2p}}^{p} =0(2|ϕ|p+1|ϕ|p1𝑑x)12𝑑t\displaystyle=\int_{0}^{\infty}\left(\int_{\mathbb{R}^{2}}|\phi|^{p+1}|\phi|^{p-1}dx\right)^{\frac{1}{2}}dt
0((1+t)p12((1+t)p18+ϵ)p1)12𝑑t\displaystyle\lesssim\int_{0}^{\infty}\left((1+t)^{-\frac{p-1}{2}}((1+t)^{-\frac{p-1}{8}+\epsilon})^{p-1}\right)^{\frac{1}{2}}dt
0(1+t)(p1)(p+3)16+p12ϵ𝑑t.\displaystyle\lesssim\int_{0}^{\infty}(1+t)^{-\frac{(p-1)(p+3)}{16}+\frac{p-1}{2}\epsilon}dt.

Since p>251p>2\sqrt{5}-1, we in particular have that

(p1)(p+3)>16.\displaystyle(p-1)(p+3)>16.

Choose ϵ\epsilon sufficiently small such that

(p1)(p+3)16+p12ϵ<1.\displaystyle-\frac{(p-1)(p+3)}{16}+\frac{p-1}{2}\epsilon<-1.

We then conclude that

ϕLtpLx2pCϵ\displaystyle\|\phi\|_{L_{t}^{p}L_{x}^{2p}}\leq C_{\epsilon}

for some constant CϵC_{\epsilon} depending only on ϵ>0\epsilon>0 (which could be fixed), pp and the initial data 0,2\mathcal{E}_{0,2}, 1,0\mathcal{E}_{1,0}. The scattering result follows by a standard argument, see for example [9]. ∎

4.2 Improved sharp decay estimates for larger pp

Since the solution decays in time, the nonlinearity decays faster for larger pp. It is believed that for sufficiently large pp, the solution to the nonlinear wave equation (1) decays as fast as linear waves (for example the superconformal case p5p\geq 5 addressed in [9]). We conjecture that for equation (1) on 1+2\mathbb{R}^{1+2}, this holds when p>3p>3 (that is for sufficiently smooth and localized data the solution decays like t12t^{-\frac{1}{2}}, which is stronger than that obtained in the previous section). In this section, we show that this conjecture holds when p>113p>\frac{11}{3}.

For fixed t0,r>0t\geq 0,\ r>0 define

ϕ(t,r)=sup{|ϕ(t,x)|:|x|=r}.\displaystyle\phi_{*}(t,r)=\sup\{|\phi(t,x)|:|x|=r\}. (17)
Lemma 4.2.

For the solution ϕ\phi of (1), let ϕ\phi_{*} be defined as above in (17). Then for 1<p<51<p<5 and t0t\geq 0

0+(1+t+r)p12|ϕ(t,r)|p+32𝑑rC\displaystyle\int_{0}^{+\infty}(1+t+r)^{\frac{p-1}{2}}|\phi_{*}(t,r)|^{\frac{p+3}{2}}dr\leq C

for some constant CC depending only on 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}.

Proof.

For fixed t0,r>0t\geq 0,\ r>0 define

ϕ~(t,r,θ)\displaystyle\widetilde{\phi}(t,r,\theta) =ϕ(t,rcosθ,rsinθ),\displaystyle=\phi(t,r\cos\theta,r\sin\theta),
ϕ(t,r)\displaystyle{\phi}_{-}(t,r) =inf{|ϕ(t,x)|:|x|=r}=inf{|ϕ~(t,r,θ)|:θ},\displaystyle=\inf\{|\phi(t,x)|:|x|=r\}=\inf\{|\widetilde{\phi}(t,r,\theta)|:\theta\in\mathbb{R}\},
A1(t,r)\displaystyle A_{1}(t,r) =12πππϕ~(t,r,θ)𝑑θ,\displaystyle=\frac{1}{2\pi}\int_{-\pi}^{\pi}\widetilde{\phi}(t,r,\theta)d\theta,
A2(t,r)\displaystyle A_{2}(t,r) =ππ|θϕ~(t,r,θ)|2𝑑θ,\displaystyle=\int_{-\pi}^{\pi}|\partial_{\theta}\widetilde{\phi}(t,r,\theta)|^{2}d\theta,
A3(t,r)\displaystyle A_{3}(t,r) =ππ|ϕ~(t,r,θ)|p+1𝑑θ.\displaystyle=\int_{-\pi}^{\pi}|\widetilde{\phi}(t,r,\theta)|^{p+1}d\theta.

Then we have 0ϕ(t,r)|A1(t,r)|0\leq{\phi}_{-}(t,r)\leq|A_{1}(t,r)| (if ϕ\phi is real valued). By the definition of ϕ,ϕ,ϕ~,A1,A2,A3\phi_{*},\ {\phi}_{-},\ \widetilde{\phi},\ A_{1},\ A_{2},\\ A_{3} and Hölder inequality, we have

ϕp+32(t,r)ϕp+32(t,r)\displaystyle\phi_{*}^{\frac{p+3}{2}}(t,r)-\phi_{-}^{\frac{p+3}{2}}(t,r) =supθ[π,π)|ϕ~(t,r,θ)|p+32infθ[π,π)|ϕ~(t,r,θ)|p+32\displaystyle=\sup_{\theta\in[-\pi,\pi)}|\widetilde{\phi}(t,r,\theta)|^{\frac{p+3}{2}}-\inf_{\theta\in[-\pi,\pi)}|\widetilde{\phi}(t,r,\theta)|^{\frac{p+3}{2}}
ππ|θ(|ϕ~|p+32)(t,r,θ)|𝑑θ\displaystyle\leq\int_{-\pi}^{\pi}|\partial_{\theta}(|\widetilde{\phi}|^{\frac{p+3}{2}})(t,r,\theta)|d\theta
=p+32ππ|θϕ~||ϕ~|p+12(t,r,θ)𝑑θ\displaystyle=\frac{p+3}{2}\int_{-\pi}^{\pi}|\partial_{\theta}\widetilde{\phi}||\widetilde{\phi}|^{\frac{p+1}{2}}(t,r,\theta)d\theta
p+32|A2(t,r)A3(t,r)|12.\displaystyle\leq\frac{p+3}{2}|A_{2}(t,r)A_{3}(t,r)|^{\frac{1}{2}}.

In particular we obtain that

ϕp+32(t,r)ϕp+32(t,r)+C|A2(t,r)A3(t,r)|12|A1(t,r)|p+32+C|A2(t,r)A3(t,r)|12.\displaystyle\phi_{*}^{\frac{p+3}{2}}(t,r)\leq\phi_{-}^{\frac{p+3}{2}}(t,r)+C|A_{2}(t,r)A_{3}(t,r)|^{\frac{1}{2}}\leq|A_{1}(t,r)|^{\frac{p+3}{2}}+C|A_{2}(t,r)A_{3}(t,r)|^{\frac{1}{2}}. (18)

Now using polar coordinate x=r(cosθ,sinθ)x=r(\cos\theta,\sin\theta) and the definition of A1,A2,A3A_{1},A_{2},A_{3}, we write that

2|ϕ(t,x)|p+1(1+t+r)p12𝑑x\displaystyle\int_{\mathbb{R}^{2}}|\phi(t,x)|^{p+1}(1+t+r)^{\frac{p-1}{2}}dx =0+A3(t,r)(1+t+r)p12r𝑑r,\displaystyle=\int_{0}^{+\infty}A_{3}(t,r)(1+t+r)^{\frac{p-1}{2}}rdr,
2(1+t2+r2)|/ ϕ|2𝑑x\displaystyle\int_{\mathbb{R}^{2}}(1+t^{2}+r^{2})|\mbox{$\nabla\mkern-13.0mu/$\,}\phi|^{2}dx =0+ππ(1+t2+r2)|r1θϕ~(t,r,θ)|2r𝑑θ𝑑r\displaystyle=\int_{0}^{+\infty}\int_{-\pi}^{\pi}(1+t^{2}+r^{2})|r^{-1}\partial_{\theta}\widetilde{\phi}(t,r,\theta)|^{2}rd\theta dr
=0+A2(t,r)(1+t2+r2)r1𝑑r,\displaystyle=\int_{0}^{+\infty}A_{2}(t,r)(1+t^{2}+r^{2})r^{-1}dr,
2|ϕ|2𝑑x\displaystyle\int_{\mathbb{R}^{2}}|\phi|^{2}dx =0+ππ|ϕ~(t,r,θ)|2r𝑑θ𝑑r2π0+|A1(t,r)|2r𝑑r,\displaystyle=\int_{0}^{+\infty}\int_{-\pi}^{\pi}|\widetilde{\phi}(t,r,\theta)|^{2}rd\theta dr\geq 2\pi\int_{0}^{+\infty}|A_{1}(t,r)|^{2}rdr,
(1+|tr|)ϕL22\displaystyle\|(1+|t-r|)\nabla\phi\|_{L^{2}}^{2} (1+|tr|)rϕL22\displaystyle\geq\|(1+|t-r|)\partial_{r}\phi\|_{L^{2}}^{2}
=0+ππ(1+|tr|)2|rϕ~(t,r,θ)|2r𝑑θ\displaystyle=\int_{0}^{+\infty}\int_{-\pi}^{\pi}(1+|t-r|)^{2}|\partial_{r}\widetilde{\phi}(t,r,\theta)|^{2}rd\theta
2π0+(1+|tr|)2|rA1(t,r)|2r𝑑r.\displaystyle\geq 2\pi\int_{0}^{+\infty}(1+|t-r|)^{2}|\partial_{r}A_{1}(t,r)|^{2}rdr.

In view of the time decay of Proposition 3.1, we conclude that

0+A3(t,r)(1+t+r)p12r𝑑rC0,2.\displaystyle\int_{0}^{+\infty}A_{3}(t,r)(1+t+r)^{\frac{p-1}{2}}rdr\leq C\mathcal{E}_{0,2}. (19)

By using estimates (13) and (14), we derive that

2(1+t2+r2)|/ ϕ|2𝑑x\displaystyle\int_{\mathbb{R}^{2}}(1+t^{2}+r^{2})|\mbox{$\nabla\mkern-13.0mu/$\,}\phi|^{2}dx C0,2(1+t)5p2,\displaystyle\leq C\mathcal{E}_{0,2}(1+t)^{\frac{5-p}{2}},
0+A2(t,r)(1+t2+r2)r1𝑑r\displaystyle\int_{0}^{+\infty}A_{2}(t,r)(1+t^{2}+r^{2})r^{-1}dr C0,2(1+t)5p2,\displaystyle\leq C\mathcal{E}_{0,2}(1+t)^{\frac{5-p}{2}},
0+A2(t,r)(1+t+r)p12r1𝑑r\displaystyle\int_{0}^{+\infty}A_{2}(t,r)(1+t+r)^{\frac{p-1}{2}}r^{-1}dr C(1+t)p1220+A2(t,r)(1+t2+r2)r1𝑑r\displaystyle\leq C(1+t)^{\frac{p-1}{2}-2}\int_{0}^{+\infty}A_{2}(t,r)(1+t^{2}+r^{2})r^{-1}dr (20)
C0,2.\displaystyle\leq C\mathcal{E}_{0,2}.

Similarly estimate (11) of Proposition 4.1 leads to

0+(|A1(t,r)|2+(1+|tr|)2|rA1(t,r)|2)r𝑑r\displaystyle\int_{0}^{+\infty}(|A_{1}(t,r)|^{2}+(1+|t-r|)^{2}|\partial_{r}A_{1}(t,r)|^{2})rdr (1+t)5p2,\displaystyle\lesssim(1+t)^{\frac{5-p}{2}}, (21)
t+14+rp121|A1(t,r)|2𝑑r\displaystyle\int_{\frac{t+1}{4}}^{+\infty}r^{\frac{p-1}{2}-1}|A_{1}(t,r)|^{2}dr C(1+t)p1220+|A1(t,r)|2r𝑑r\displaystyle\leq C(1+t)^{\frac{p-1}{2}-2}\int_{0}^{+\infty}|A_{1}(t,r)|^{2}rdr (22)
1.\displaystyle\lesssim 1.

Therefore from the above estimates (20) and (19), we show that

0+|A2(t,r)A3(t,r)|12(1+t+r)p12𝑑r\displaystyle\int_{0}^{+\infty}|A_{2}(t,r)A_{3}(t,r)|^{\frac{1}{2}}(1+t+r)^{\frac{p-1}{2}}dr 0+(r1A2(t,r)+rA3(t,r))(1+t+r)p12𝑑r\displaystyle\leq\int_{0}^{+\infty}(r^{-1}A_{2}(t,r)+rA_{3}(t,r))(1+t+r)^{\frac{p-1}{2}}dr (23)
1.\displaystyle\leq 1.

By using Hölder’s inequality and the definition of A1A_{1} and A3A_{3}, we have

2π|A1(t,r)|p+1A3(t,r).\displaystyle 2\pi|A_{1}(t,r)|^{p+1}\leq A_{3}(t,r).

We thus can show that

|A1(t,r)|p+32=|A1(t,r)|p+12|A1(t,r)||A3(t,r)|12|A1(t,r)|r|A3(t,r)|+r1|A1(t,r)|2,\displaystyle|A_{1}(t,r)|^{\frac{p+3}{2}}=|A_{1}(t,r)|^{\frac{p+1}{2}}|A_{1}(t,r)|\leq|A_{3}(t,r)|^{\frac{1}{2}}|A_{1}(t,r)|\leq r|A_{3}(t,r)|+r^{-1}|A_{1}(t,r)|^{2}, (24)
|r(|A1(t,r)|p+32)|=p+32|A1(t,r)|p+12|rA1(t,r)|C|A3(t,r)|12|rA1(t,r)|.\displaystyle|\partial_{r}(|A_{1}(t,r)|^{\frac{p+3}{2}})|=\frac{p+3}{2}|A_{1}(t,r)|^{\frac{p+1}{2}}|\partial_{r}A_{1}(t,r)|\leq C|A_{3}(t,r)|^{\frac{1}{2}}|\partial_{r}A_{1}(t,r)|. (25)

From estimates (24), (19) and (22), we conclude that

t+14+rp12|A1(t,r)|p+32𝑑rt+14+rp12(r|A3(t,r)|+r1|A1(t,r)|2)𝑑r1.\displaystyle\int_{\frac{t+1}{4}}^{+\infty}r^{\frac{p-1}{2}}|A_{1}(t,r)|^{\frac{p+3}{2}}dr\leq\int_{\frac{t+1}{4}}^{+\infty}r^{\frac{p-1}{2}}(r|A_{3}(t,r)|+r^{-1}|A_{1}(t,r)|^{2})dr\lesssim 1. (26)

In view of bounds (25), (19) and (21), we show that

0t+1|A1(t,r)|p+32(t+12r)𝑑r\displaystyle\int_{0}^{t+1}|A_{1}(t,r)|^{\frac{p+3}{2}}(t+1-2r)dr
=0t+1r(t+1r)r(|A1(t,r)|p+32)dr\displaystyle=-\int_{0}^{t+1}r(t+1-r)\partial_{r}(|A_{1}(t,r)|^{\frac{p+3}{2}})dr
C0t+1r(t+1r)|A3(t,r)|12|rA1(t,r)|𝑑r\displaystyle\leq C\int_{0}^{t+1}r(t+1-r)|A_{3}(t,r)|^{\frac{1}{2}}|\partial_{r}A_{1}(t,r)|dr
C0t+1(t+1)|A3(t,r)|r𝑑r+C(t+1)10t+1(t+1r)2|rA1(t,r)|r𝑑r\displaystyle\leq C\int_{0}^{t+1}(t+1)|A_{3}(t,r)|rdr+C(t+1)^{-1}\int_{0}^{t+1}(t+1-r)^{2}|\partial_{r}A_{1}(t,r)|rdr
C(t+1)1p120+A3(t,r)(1+t+r)p12r𝑑r+C(t+1)10,2(1+t)5p2\displaystyle\leq C(t+1)^{1-\frac{p-1}{2}}\int_{0}^{+\infty}A_{3}(t,r)(1+t+r)^{\frac{p-1}{2}}rdr+C(t+1)^{-1}\mathcal{E}_{0,2}(1+t)^{\frac{5-p}{2}}
(1+t)3p2,\displaystyle\lesssim(1+t)^{\frac{3-p}{2}},

which together with (26) implies that

0t+12|A1(t,r)|p+32(t+12r)𝑑r\displaystyle\int_{0}^{\frac{t+1}{2}}|A_{1}(t,r)|^{\frac{p+3}{2}}(t+1-2r)dr
=0t+1|A1(t,r)|p+32(t+12r)𝑑r+t+12t+1|A1(t,r)|p+32(2rt1)𝑑r\displaystyle=\int_{0}^{t+1}|A_{1}(t,r)|^{\frac{p+3}{2}}(t+1-2r)dr+\int_{\frac{t+1}{2}}^{t+1}|A_{1}(t,r)|^{\frac{p+3}{2}}(2r-t-1)dr
(1+t)3p2+(1+t)3p2t+12t+1|A1(t,r)|p+32rp12𝑑r\displaystyle\lesssim(1+t)^{\frac{3-p}{2}}+(1+t)^{\frac{3-p}{2}}\int_{\frac{t+1}{2}}^{t+1}|A_{1}(t,r)|^{\frac{p+3}{2}}r^{\frac{p-1}{2}}dr
(1+t)3p2.\displaystyle\lesssim(1+t)^{\frac{3-p}{2}}.

Moreover for small rr, we bound that

0t+14|A1(t,r)|p+32𝑑r2t+10t+12|A1(t,r)|p+32(t+12r)𝑑r(1+t)1p2.\displaystyle\int_{0}^{\frac{t+1}{4}}|A_{1}(t,r)|^{\frac{p+3}{2}}dr\leq\frac{2}{t+1}\int_{0}^{\frac{t+1}{2}}|A_{1}(t,r)|^{\frac{p+3}{2}}(t+1-2r)dr\lesssim(1+t)^{\frac{1-p}{2}}. (27)

Combining estimates (26) and (27), we deduce that

0+|A1(t,r)|p+32(1+t+r)p12𝑑r1,\displaystyle\int_{0}^{+\infty}|A_{1}(t,r)|^{\frac{p+3}{2}}(1+t+r)^{\frac{p-1}{2}}dr\lesssim 1,

which together with estimates (18) and (23) implies the Lemma. Here recall that by our notation the implicit constant relies only on the initial data 0,2\mathcal{E}_{0,2} and 1,0\mathcal{E}_{1,0}. ∎

We also need the following integration lemma.

Lemma 4.3.

For real constant A1A\neq 1, let

F(A)=θ[π,π),A+cosθ>01A+cosθ𝑑θ.\displaystyle F(A)=\int_{\theta\in[-\pi,\pi),A+\cos\theta>0}\frac{1}{\sqrt{A+\cos\theta}}d\theta.

Then we can bound that

F(A)C(1+ln(1+1/|A1|))F(A)\leq C(1+\ln(1+1/|A-1|))

for some constant CC independent of AA.

Proof.

If A1A\leq-1 then F(A)=0F(A)=0. For the case when 1<A<1-1<A<1, we first can write that

F(A)\displaystyle F(A) =2θ[0,π),A+cosθ>01A+cosθ𝑑θ\displaystyle=2\int_{\theta\in[0,\pi),A+\cos\theta>0}\frac{1}{\sqrt{A+\cos\theta}}d\theta
=2A11A+z1z2𝑑z\displaystyle=2\int_{-A}^{1}\frac{1}{\sqrt{A+z}\sqrt{1-z^{2}}}dz
=40π/21(1+A)sin2s+(1A)𝑑s.\displaystyle=4\int_{0}^{\pi/2}\frac{1}{\sqrt{(1+A)\sin^{2}s+(1-A)}}ds.

Now for the case when 1<A0-1<A\leq 0, it trivially has that

(1+A)sin2s+(1A)(1A)1.(1+A)\sin^{2}s+(1-A)\geq(1-A)\geq 1.

In particular

F(A)=\displaystyle F(A)= 40π/21(1+A)sin2s+(1A)𝑑s2π.\displaystyle 4\int_{0}^{\pi/2}\frac{1}{\sqrt{(1+A)\sin^{2}s+(1-A)}}ds\leq 2\pi.

For the case when 0A<10\leq A<1, we estimate that

(1+A)sin2s+(1A)sin2s+(1A)12(2sπ+1A)2,0sπ2.\displaystyle(1+A)\sin^{2}s+(1-A)\geq\sin^{2}s+(1-A)\geq\frac{1}{2}(\frac{2s}{\pi}+\sqrt{1-A})^{2},\quad\forall 0\leq s\leq\frac{\pi}{2}.

Therefore we can bound that

F(A)\displaystyle F(A) 420π/212sπ+1A𝑑s=22πln1+1A1A22π(1+ln(1+1/|A1|)).\displaystyle\leq 4\sqrt{2}\int_{0}^{\pi/2}\frac{1}{\frac{2s}{\pi}+\sqrt{1-A}}ds=2\sqrt{2}\pi\ln\frac{1+\sqrt{1-A}}{\sqrt{1-A}}\leq 2\sqrt{2}\pi(1+\ln(1+1/|A-1|)).

Similarly if A>1A>1, we change variable as follows

F(A)\displaystyle F(A) =20π1A+cosθ𝑑θ\displaystyle=2\int_{0}^{\pi}\frac{1}{\sqrt{A+\cos\theta}}d\theta
=40π/212sin2s+(A1)𝑑s\displaystyle=4\int_{0}^{\pi/2}\frac{1}{\sqrt{2\sin^{2}s+(A-1)}}ds
4π20π/2122s+πA1𝑑s\displaystyle\leq 4\pi\sqrt{2}\int_{0}^{\pi/2}\frac{1}{2\sqrt{2}s+\pi\sqrt{A-1}}ds
=2πln2+1A1A\displaystyle=2\pi\ln\frac{\sqrt{2}+\sqrt{1-A}}{\sqrt{1-A}}
22π(1+ln(1+1/|A1|)).\displaystyle\leq 2\sqrt{2}\pi(1+\ln(1+1/|A-1|)).

This completes the proof. ∎

We will also use the following lemma.

Lemma 4.4.

Let gg be nonnegative function defined on r>0r>0. Then for any s>0s>0 and xx\in\mathbb{R}, it holds

r>0g(r)(1+ln(1+2r|xr|1))𝑑r\displaystyle\int_{r>0}g(r)(1+\ln(1+2r|x-r|^{-1}))dr Cr>0g(r)(1+ln(1+s+r))𝑑r+CgL(1+s)1\displaystyle\leq C\int_{r>0}g(r)(1+\ln(1+s+r))dr+C\|g\|_{L^{\infty}}(1+s)^{-1}

for some constant CC which is independent of ss and xx.

Proof.

Denote

δ=(1+|x|+s)1.\delta=(1+|x|+s)^{-1}.

When |xr|1,r>0|x-r|\geq 1,\ r>0, we bound that

1+2rmin(δ1,|xr|1)1+2r(1+r)22(1+s+r)2.\displaystyle 1+2r\min(\delta^{-1},|x-r|^{-1})\leq 1+2r\leq(1+r)^{2}\leq 2(1+s+r)^{2}.

Otherwise if |xr|1,r>0|x-r|\leq 1,\ r>0 (in particular |x|1+r|x|\leq 1+r), we have

1+2rmin(δ1,|xr|1)1+2r(1+|x|+s)1+2r(2+r+s)2(1+s+r)2.\displaystyle 1+2r\min(\delta^{-1},|x-r|^{-1})\leq 1+2r(1+|x|+s)\leq 1+2r(2+r+s)\leq 2(1+s+r)^{2}.

Thus for r>0r>0, we always have

1+2rmin(δ1,|xr|1)2(1+s+r)2.\displaystyle 1+2r\min(\delta^{-1},|x-r|^{-1})\leq 2(1+s+r)^{2}.

Note that

1+2r|xr|1(1+2rmin(δ1,|xr|1))max(δ|xr|1,1).\displaystyle 1+2r|x-r|^{-1}\leq(1+2r\min(\delta^{-1},|x-r|^{-1}))\cdot\max(\delta|x-r|^{-1},1).

We therefore can show that

r>0g(r)(1+ln(1+2r|xr|1))𝑑r\displaystyle\int_{r>0}g(r)(1+\ln(1+2r|x-r|^{-1}))dr
r>0g(r)(1+ln(1+2rmin(δ1,|xr|1)))𝑑r+r>0g(r)lnmax(δ|xr|1,1)dr\displaystyle\leq\int_{r>0}g(r)(1+\ln(1+2r\min(\delta^{-1},|x-r|^{-1})))dr+\int_{r>0}g(r)\ln\max(\delta|x-r|^{-1},1)dr
r>0g(r)(1+ln(2(1+s+r)2))𝑑r+gLlnmax(δ|xr|1,1)dr\displaystyle\leq\int_{r>0}g(r)(1+\ln(2(1+s+r)^{2}))dr+\|g\|_{L^{\infty}}\int_{\mathbb{R}}\ln\max(\delta|x-r|^{-1},1)dr
2r>0g(r)(1+ln(1+s+r))𝑑r+2gL(1+s)1.\displaystyle\leq 2\int_{r>0}g(r)(1+\ln(1+s+r))dr+2\|g\|_{L^{\infty}}(1+s)^{-1}.

Here for the last step, we used

lnmax(δ/|xr|,1)dr=20δln(δr1)𝑑r=2δ2(1+s)1.\displaystyle\int_{\mathbb{R}}\ln\max(\delta/|x-r|,1)dr=2\int_{0}^{\delta}\ln(\delta r^{-1})dr=2\delta\leq 2(1+s)^{-1}.

We thus finished the proof for the Lemma. ∎

With the above preparations, we are now ready to establish the following improved decay estimates for the solution for p>113p>\frac{11}{3}, which together with Proposition 4.2 implies Theorem 2.

Proposition 4.3.

When 113<p<5\frac{11}{3}<p<5, the solution to (1) verifies the following sharp time decay estimate

|ϕ(t,x)|C(1+t+|x|)12\displaystyle|\phi(t,x)|\leq C(1+t+|x|)^{-\frac{1}{2}}

for some constant CC depending only on p,0,2,1,0p,\ \mathcal{E}_{0,2},\ \mathcal{E}_{1,0}.

Proof.

Let’s fix time t>0t>0. Define

f(t)=supx2|ϕ(t,x)(1+t)12|,f(t)=supx2|ϕ(t,x)(1+t+|x|)12|.\displaystyle f(t)=\sup\limits_{x\in\mathbb{R}^{2}}|\phi(t,x)(1+t)^{\frac{1}{2}}|,\quad f_{*}(t)=\sup\limits_{x\in\mathbb{R}^{2}}|\phi(t,x)(1+t+|x|)^{\frac{1}{2}}|.

In view of the rough decay estimate of Proposition 4.2, it is obvious that f(t)f(t) is finite for all t>0t>0. By the definition of ϕ\phi_{*} (in (17)) and ff, we have

ϕ(t,r)f(t)(1+t)12,\displaystyle\phi_{*}(t,r)\leq f(t)(1+t)^{-\frac{1}{2}},

which together with Lemma 4.2 implies that

0+(1+t+r)p12|ϕ(t,r)|p𝑑r\displaystyle\int_{0}^{+\infty}(1+t+r)^{\frac{p-1}{2}}|\phi_{*}(t,r)|^{p}dr |supr>0ϕ(t,r)|p320+(1+t+r)p12|ϕ(t,r)|p+32𝑑r\displaystyle\leq\left|\sup_{r>0}\phi_{*}(t,r)\right|^{\frac{p-3}{2}}\int_{0}^{+\infty}(1+t+r)^{\frac{p-1}{2}}|\phi_{*}(t,r)|^{\frac{p+3}{2}}dr (28)
|f(t)(1+t)12|p32.\displaystyle\lesssim|f(t)(1+t)^{-\frac{1}{2}}|^{\frac{p-3}{2}}.

Using the rough decay estimate of Proposition 4.2 again and in view of the definition of ϕ\phi_{*}, we have

ϕ(t,r)\displaystyle\phi_{*}(t,r) (1+t+r)p18+ϵ,\displaystyle\lesssim(1+t+r)^{-\frac{p-1}{8}+\epsilon},
r12ϕp(t,r)\displaystyle r^{\frac{1}{2}}\phi_{*}^{p}(t,r) (1+t+r)p(p1)8+12+pϵ.\displaystyle\lesssim(1+t+r)^{-\frac{p(p-1)}{8}+\frac{1}{2}+p\epsilon}.

For the case when 113<p<5\frac{11}{3}<p<5, we can choose ϵ\epsilon such that

0<ϵ<p181p.0<\epsilon<\frac{p-1}{8}-\frac{1}{p}.

In particular we conclude that

r12ϕp(t,r)(1+t+r)12.\displaystyle r^{\frac{1}{2}}\phi_{*}^{p}(t,r)\lesssim(1+t+r)^{-\frac{1}{2}}. (29)

Now for any point (t0,x0)1+2(t_{0},x_{0})\in\mathbb{R}^{1+2}, t00,t_{0}\geq 0, denote r0=|x0|r_{0}=|x_{0}|. In view of the representation formula for linear wave equation, the solution ϕ\phi to the semilinear wave equation (1) verifies

|ϕ(t0,x0)|(1+t0+r0)12+0t0|x0y|<t0s|ϕ|p(s,y)(t0s)2|x0y|2𝑑y𝑑s.|\phi(t_{0},x_{0})|\lesssim(1+t_{0}+r_{0})^{-\frac{1}{2}}+\int_{0}^{t_{0}}\int_{|x_{0}-y|<t_{0}-s}\frac{|\phi|^{p}(s,y)}{\sqrt{(t_{0}-s)^{2}-|x_{0}-y|^{2}}}dyds. (30)

Using polar coordinate y=r(cosθ,sinθ)y=r(\cos\theta,\sin\theta) and the definition of ϕ\phi_{*} we have

|x0y|<t0s|ϕ|p(s,y)(t0s)2|x0y|2𝑑y\displaystyle\int_{|x_{0}-y|<t_{0}-s}\frac{|\phi|^{p}(s,y)}{\sqrt{(t_{0}-s)^{2}-|x_{0}-y|^{2}}}dy
\displaystyle\leq r>0θI(r)ϕp(s,r)(t0s)2|x0r(cosθ,sinθ)|2r𝑑θ𝑑r,\displaystyle\int_{r>0}\int_{\theta\in I(r)}\frac{\phi_{*}^{p}(s,r)}{\sqrt{(t_{0}-s)^{2}-|x_{0}-r(\cos\theta,\sin\theta)|^{2}}}rd\theta dr,

where (for fixed x0,t0,sx_{0},t_{0},s)

I(r)={θ[π,π):(t0s)>|x0r(cosθ,sinθ)|}.\displaystyle I(r)=\{\theta\in[-\pi,\pi):(t_{0}-s)>|x_{0}-r(\cos\theta,\sin\theta)|\}.

Note that

|x0r(cosθ,sinθ)|2=r02+r22rr0cos(θ0θ),x0=r0(cosθ0,sinθ0).\displaystyle|x_{0}-r(\cos\theta,\sin\theta)|^{2}=r_{0}^{2}+r^{2}-2rr_{0}\cos(\theta_{0}-\theta),\quad x_{0}=r_{0}(\cos\theta_{0},\sin\theta_{0}).

Thus we can estimate that

θI(r)1(t0s)2|x0r(cosθ,sinθ)|2𝑑θ\displaystyle\int_{\theta\in I(r)}\frac{1}{\sqrt{(t_{0}-s)^{2}-|x_{0}-r(\cos\theta,\sin\theta)|^{2}}}d\theta
=\displaystyle= θI(r)1(t0s)2r02r2+2rr0cos(θ0θ)𝑑θ\displaystyle\int_{\theta\in I(r)}\frac{1}{\sqrt{(t_{0}-s)^{2}-r_{0}^{2}-r^{2}+2rr_{0}\cos(\theta_{0}-\theta)}}d\theta
=\displaystyle= θI~(r)1(t0s)2r02r2+2rr0cosθ𝑑θ\displaystyle\int_{\theta\in\widetilde{I}(r)}\frac{1}{\sqrt{(t_{0}-s)^{2}-r_{0}^{2}-r^{2}+2rr_{0}\cos\theta}}d\theta
=\displaystyle= F(21r1r01((t0s)2r02r2))2rr0.\displaystyle\frac{F(2^{-1}r^{-1}r_{0}^{-1}((t_{0}-s)^{2}-r_{0}^{2}-r^{2}))}{\sqrt{2rr_{0}}}.

Here (for fixed x0,t0,sx_{0},t_{0},s) the function FF is defined in Lemma 4.3 and

I~(r)={θ[π,π):(t0s)2r02r2+2rr0cosθ>0}.\displaystyle\widetilde{I}(r)=\{\theta\in[-\pi,\pi):(t_{0}-s)^{2}-r_{0}^{2}-r^{2}+2rr_{0}\cos\theta>0\}.

By using Lemma 4.3, we then can show that

|x0y|<t0s|ϕ|p(s,y)(t0s)2|x0y|2𝑑y\displaystyle\int_{|x_{0}-y|<t_{0}-s}\frac{|\phi|^{p}(s,y)}{\sqrt{(t_{0}-s)^{2}-|x_{0}-y|^{2}}}dy
\displaystyle\leq r>0ϕp(s,r)2rr0F(21r1r01((t0s)2r02r2))r𝑑r\displaystyle\int_{r>0}\frac{\phi_{*}^{p}(s,r)}{\sqrt{2rr_{0}}}F(2^{-1}r^{-1}r_{0}^{-1}((t_{0}-s)^{2}-r_{0}^{2}-r^{2}))rdr
\displaystyle\leq Cr>0ϕp(s,r)2rr0(1+ln(1+2rr0|(t0s)2(r0+r)2|1)rdr\displaystyle C\int_{r>0}\frac{\phi_{*}^{p}(s,r)}{\sqrt{2rr_{0}}}(1+\ln(1+2rr_{0}|(t_{0}-s)^{2}-(r_{0}+r)^{2}|^{-1})rdr
\displaystyle\leq Cr>0ϕp(s,r)2rr0(1+ln(1+2r/|t0sr0r|))r𝑑r.\displaystyle C\int_{r>0}\frac{\phi_{*}^{p}(s,r)}{\sqrt{2rr_{0}}}(1+\ln(1+2r/|t_{0}-s-r_{0}-r|))rdr.

Here in the last step we used (for t0>s>0t_{0}>s>0)

|(t0s)2(r0+r)2|=|t0sr0r||t0s+r0+r||t0sr0r|r0.\displaystyle|(t_{0}-s)^{2}-(r_{0}+r)^{2}|=|t_{0}-s-r_{0}-r||t_{0}-s+r_{0}+r|\geq|t_{0}-s-r_{0}-r|r_{0}.

Now using Lemma 4.4 with g(r)=r12ϕp(s,r)g(r)=r^{\frac{1}{2}}\phi_{*}^{p}(s,r), x=t0sr0x=t_{0}-s-r_{0} as well as estimate (29) , we conclude that

|x0y|<t0s|ϕ|p(s,y)(t0s)2|x0y|2𝑑yr>0ϕp(s,r)2rr0(1+ln(1+s+r))r𝑑r+r012(1+s)32.\begin{split}&\int_{|x_{0}-y|<t_{0}-s}\frac{|\phi|^{p}(s,y)}{\sqrt{(t_{0}-s)^{2}-|x_{0}-y|^{2}}}dy\\ &\lesssim\int_{r>0}\frac{\phi_{*}^{p}(s,r)}{\sqrt{2rr_{0}}}(1+\ln(1+s+r))rdr+r_{0}^{-\frac{1}{2}}(1+s)^{-\frac{3}{2}}.\end{split} (31)

Note that for positive constant γ>0\gamma>0, the function zγ(1+γlnz)z^{-\gamma}(1+\gamma\ln z) of zz is decreasing for z>1z>1. Hence for s>0,r>0s>0,\quad r>0 and γ=p/21\gamma=p/2-1, we can show that

r12(1+s)p22(1+γln(1+s+r))\displaystyle r^{\frac{1}{2}}(1+s)^{\frac{p-2}{2}}(1+\gamma\ln(1+s+r)) r12(1+s+r)p22(1+γln(1+s))\displaystyle\leq r^{\frac{1}{2}}(1+s+r)^{\frac{p-2}{2}}(1+\gamma\ln(1+s))
(1+s+r)p12(1+γln(1+s)),\displaystyle\leq(1+s+r)^{\frac{p-1}{2}}(1+\gamma\ln(1+s)),

which together with estimates (31), (28) leads to

|x0y|<t0s|ϕ|p(s,y)(t0s)2|x0y|2𝑑y\displaystyle\int_{|x_{0}-y|<t_{0}-s}\frac{|\phi|^{p}(s,y)}{\sqrt{(t_{0}-s)^{2}-|x_{0}-y|^{2}}}dy
(1+s)p22r012(ln(1+s)+1)r>0(1+s+r)p12ϕp(s,r)𝑑r+r012(1+s)32\displaystyle\lesssim(1+s)^{-\frac{p-2}{2}}r_{0}^{-\frac{1}{2}}(\ln(1+s)+1)\int_{r>0}(1+s+r)^{\frac{p-1}{2}}\phi_{*}^{p}(s,r)dr+r_{0}^{-\frac{1}{2}}(1+s)^{-\frac{3}{2}}
(1+s)p22r012(ln(1+s)+1)(1+s)p34|f(s)|p32+r012(1+s)32\displaystyle\lesssim(1+s)^{-\frac{p-2}{2}}r_{0}^{-\frac{1}{2}}(\ln(1+s)+1)(1+s)^{-\frac{p-3}{4}}|f(s)|^{\frac{p-3}{2}}+r_{0}^{-\frac{1}{2}}(1+s)^{-\frac{3}{2}}
r012(ln(1+s)+1)(1+s)73p4|f(s)|p32+r012(1+s)32.\displaystyle\lesssim r_{0}^{-\frac{1}{2}}(\ln(1+s)+1)(1+s)^{\frac{7-3p}{4}}|f(s)|^{\frac{p-3}{2}}+r_{0}^{-\frac{1}{2}}(1+s)^{-\frac{3}{2}}.

Multiply both side of (30) with r012r_{0}^{\frac{1}{2}}, we have

|ϕ(t0,x0)|r0121+0t0((ln(1+s)+1)(1+s)73p4|f(s)|p32+(1+s)32)𝑑s.\displaystyle|\phi(t_{0},x_{0})|r_{0}^{\frac{1}{2}}\lesssim 1+\int_{0}^{t_{0}}((\ln(1+s)+1)(1+s)^{\frac{7-3p}{4}}|f(s)|^{\frac{p-3}{2}}+(1+s)^{-\frac{3}{2}})ds.

On the other hand if r0=|x0|12max(1,t0)r_{0}=|x_{0}|\leq\frac{1}{2}\max(1,t_{0}), then by Proposition 4.2 we have the improved decay estimates

|ϕ(t0,x0)|(1+t0)p14+ϵ(1+t0)12.\displaystyle|\phi(t_{0},x_{0})|\lesssim(1+t_{0})^{-\frac{p-1}{4}+\epsilon}\lesssim(1+t_{0})^{-\frac{1}{2}}.

Hence taking supreme in terms of x0x_{0} and in view of the definition for f(s),f(s)f(s),\ f_{*}(s), we derive that

f(t0)f(t0)1+0t0(ln(1+s)+1)(1+s)73p4|f(s)|p32𝑑s.\displaystyle f(t_{0})\leq f_{*}(t_{0})\lesssim 1+\int_{0}^{t_{0}}(\ln(1+s)+1)(1+s)^{\frac{7-3p}{4}}|f(s)|^{\frac{p-3}{2}}ds.

For the case when 113<p<5\frac{11}{3}<p<5, we in particular have that

3p74>1>p32>0,\frac{3p-7}{4}>1>\frac{p-3}{2}>0,

which together with the previous estimate leads to the conclusion that

f(t0)1.f_{*}(t_{0})\lesssim 1.

The Proposition then follows by our convention that the implicit constant relies only on pp and the initial data 0,2\mathcal{E}_{0,2}, 1,0\mathcal{E}_{1,0}. ∎

References

  • [1] H. Brézis and T. Gallouet. Nonlinear Schrödinger evolution equations. Nonlinear Anal., 4(4):677–681, 1980.
  • [2] H. Brézis and S. Wainger. A note on limiting cases of Sobolev embeddings and convolution inequalities. Comm. Partial Differential Equations, 5(7):773–789, 1980.
  • [3] M. Dafermos and I. Rodnianski. A new physical-space approach to decay for the wave equation with applications to black hole spacetimes. In XVIth International Congress on Mathematical Physics, pages 421–432. World Sci. Publ., Hackensack, NJ, 2010.
  • [4] B. Dodson. Global well-posedness and scattering for the radial, defocusing, cubic nonlinear wave equation. 2018. arXiv:1809.08284.
  • [5] B. Dodson. Global well-posedness for the radial, defocusing, nonlinear wave equation for 3<p<53<p<5. 2018. arXiv:1810.02879.
  • [6] D. Eardley and V. Moncrief. The global existence of Yang-Mills-Higgs fields in 44-dimensional Minkowski space. I. Local existence and smoothness properties. Comm. Math. Phys., 83(2):171–191, 1982.
  • [7] J. Ginibre and G. Velo. The global Cauchy problem for the nonlinear Klein-Gordon equation. Math. Z., 189(4):487–505, 1985.
  • [8] J. Ginibre and G. Velo. Conformal invariance and time decay for nonlinear wave equations. I, II. Ann. Inst. H. Poincaré Phys. Théor., 47(3):221–261, 263–276, 1987.
  • [9] R. Glassey and H. Pecher. Time decay for nonlinear wave equations in two space dimensions. Manuscripta Math., 38(3):387–400, 1982.
  • [10] K. Hidano. Scattering problem for the nonlinear wave equation in the finite energy and conformal charge space. J. Funct. Anal., 187(2):274–307, 2001.
  • [11] K. Hidano. Conformal conservation law, time decay and scattering for nonlinear wave equations. J. Anal. Math., 91:269–295, 2003.
  • [12] G. Luli, S. Yang, and P. Yu. On one-dimension semi-linear wave equations with null conditions. Adv. Math., 329:174–188, 2018.
  • [13] H. Pecher. Decay of solutions of nonlinear wave equations in three space dimensions. J. Funct. Anal., 46(2):221–229, 1982.
  • [14] W. Strauss. Decay and asymptotics for cmu=F(u)cmu=F(u). J. Functional Analysis, 2:409–457, 1968.
  • [15] W. von Wahl. Some decay-estimates for nonlinear wave equations. J. Functional Analysis, 9:490–495, 1972.
  • [16] S. Yang. Global behaviors for defocusing semilinear wave equations. 2019. arXiv:1908.00606.
  • [17] S. Yang. Pointwise decay for semilinear wave equations in 3+1\mathbb{R}^{3+1}. 2019. arXiv:1908.00607.
  • [18] S. Yang. Uniform bound for solutions of semilinear wave equations in 3+1\mathbb{R}^{3+1}. 2019. arXiv:1910.02230.

School of Mathematical Sciences, Peking University, Beijing, China

E-mail address: [email protected]

Beijing International Center for Mathematical Research, Peking University, Beijing, China

E-mail address: [email protected]