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

Asymptotic expansion of 2-dimensional gradient graph with vanishing mean curvature at infinity

Zixiao Liu and Jiguang Bao111 Supported by the National Key Research and Development Program of China (No. 2020YFA0712904) and National Natural Science Foundation of China (No. 11871102 and No. 11631002).
Abstract

In this paper, we establish the asymptotic expansion at infinity of gradient graph in dimension 2 with vanishing mean curvature at infinity. This corresponds to our previous results in higher dimensions and generalizes the results for minimal gradient graph on exterior domain in dimension 2. Different from the strategies for higher dimensions, instead of the equivalence of Green’s function on unbounded domains, we apply a version of iteration methods from Bao–Li–Zhang [Calc.Var PDE, 52(2015), pp. 39-63] that is refined by spherical harmonic expansions to provide a more explicit asymptotic behavior than known results.
Keywords: Monge–Ampère equation, Mean curvature equation, Asymptotic behavior, Spherical harmonic expansion.
MSC 2020: 35J60, 35C20, 35B20.

1 Introduction

We consider the asymptotic expansion at infinity of solutions to a family of mean curvature equations of gradient graph in dimension 22.

Let (x,Du(x))(x,Du(x)) denote the gradient graph of uu in (n×n,gτ)(\mathbb{R}^{n}\times\mathbb{R}^{n},g_{\tau}), where DuDu denotes the gradient of scalar function uu and

gτ=sinτδ0+cosτg0,τ[0,π2]g_{\tau}=\sin\tau\delta_{0}+\cos\tau g_{0},\quad\tau\in\left[0,\frac{\pi}{2}\right]

is the linearly combined metric of standard Euclidean metric

δ0=i=1ndxidxi+j=1ndyjdyj,\delta_{0}=\sum_{i=1}^{n}dx_{i}\otimes dx_{i}+\sum_{j=1}^{n}dy_{j}\otimes dy_{j},

with the pseudo-Euclidean metric

g0=i=1ndxidyi+j=1ndyjdxj.g_{0}=\sum_{i=1}^{n}dx_{i}\otimes dy_{i}+\sum_{j=1}^{n}dy_{j}\otimes dx_{j}.

As proved in [37], if uC2(n)u\in C^{2}(\mathbb{R}^{n}) is a classical solution of

Fτ(λ(D2u))=f(x),F_{\tau}(\lambda(D^{2}u))=f(x), (1.1)

then Df(x)Df(x) is the mean curvature of gradient graph (x,Du(x))(x,Du(x)) in (n×n,gτ)\left(\mathbb{R}^{n}\times\mathbb{R}^{n},g_{\tau}\right). In (1.1), f(x)f(x) is a sufficiently regular function, λ(D2u)=(λ1,λ2,,λn)\lambda\left(D^{2}u\right)=\left(\lambda_{1},\lambda_{2},\cdots,\lambda_{n}\right) is the vector formed by nn eigenvalues of Hessian matrix D2uD^{2}u and

Fτ(λ):={1ni=1nlnλi,τ=0,a2+12bi=1nlnλi+abλi+a+b,0<τ<π4,2i=1n11+λi,τ=π4,a2+1bi=1narctanλi+abλi+a+b,π4<τ<π2,i=1narctanλi,τ=π2,F_{\tau}(\lambda):=\left\{\begin{array}[]{ccc}\displaystyle\frac{1}{n}\sum_{i=1}^{n}\ln\lambda_{i},&\tau=0,\\ \displaystyle\frac{\sqrt{a^{2}+1}}{2b}\sum_{i=1}^{n}\ln\frac{\lambda_{i}+a-b}{\lambda_{i}+a+b},&0<\tau<\frac{\pi}{4},\\ \displaystyle-\sqrt{2}\sum_{i=1}^{n}\frac{1}{1+\lambda_{i}},&\tau=\frac{\pi}{4},\\ \displaystyle\frac{\sqrt{a^{2}+1}}{b}\sum_{i=1}^{n}\arctan\displaystyle\frac{\lambda_{i}+a-b}{\lambda_{i}+a+b},&\frac{\pi}{4}<\tau<\frac{\pi}{2},\\ \displaystyle\sum_{i=1}^{n}\arctan\lambda_{i},&\tau=\frac{\pi}{2},\\ \end{array}\right.

a=cotτ,b=|cot2τ1|a=\cot\tau,b=\sqrt{\left|\cot^{2}\tau-1\right|}.

When τ=0\tau=0, Eq. (1.1) becomes the Monge–Ampère type equation

detD2u=enf(x).\det D^{2}u=e^{nf(x)}.

When τ=π4\tau=\frac{\pi}{4}, Eq. (1.1) can be translated into the inverse harmonic Hessian equation

2i=1n1λi=f(x),-\sqrt{2}\sum_{i=1}^{n}\dfrac{1}{\lambda_{i}}=f(x),

which is a special form of the quadratic Hessian equations σk(λ)σl(λ)=f(x)\frac{\sigma_{k}(\lambda)}{\sigma_{l}(\lambda)}=f(x), where σk(λ)\sigma_{k}(\lambda) with k=1,2,,nk=1,2,\cdots,n denotes the kk-th elementary symmetric function of λ(D2u)\lambda(D^{2}u).
When τ=π2\tau=\frac{\pi}{2}, Eq. (1.1) becomes the Lagrangian mean curvature type equation

i=1narctanλi(D2u)=f(x).\sum_{i=1}^{n}\arctan\lambda_{i}\left(D^{2}u\right)=f(x). (1.2)

Especially when f(x)f(x) is a constant, the Lagrangian mean curvature equation above is also known as the special Lagrangian equation.

For f(x)f(x) being a constant C0C_{0}, Warren [39] proved Bernstein-type results of (1.1) based on the results of Jörgens [26]–Calabi [11]–Pogorelov [35], Flanders [17] and Yuan [43, 44], which state that any classical solution with under suitable semi-convex conditions must be a quadratic. Especially when τ=0\tau=0, there are different proofs and extensions of the Bernstein-type results of Monge–Ampère equations by Cheng–Yau [15], Caffarelli [6], Jost–Xin [27], Fu [18], Li–Xu–Simon–Jia [29], etc. When τ=π4\tau=\frac{\pi}{4}, there are Bernstein-type results of Hessian quotient equations by Bao–Chen–Guan–Ji [1]. For generalizations of Bernstein-type results of Hessian and Hessian quotient equations, we refer to Chang–Yuan [12], Chen–Xiang [14], Li–Ren–Wang [31], Yuan [43], Du [16] etc. When τ=π2\tau=\frac{\pi}{2}, though the mean curvature relies only on DfDf, Yuan [44] reveals the importance of the value of phase C0C_{0}. The phase (also known as the Lagrangian angle) n22π\frac{n-2}{2}\pi is called critical since the level set

{λn|i=1narctanλi=C0}\left\{\lambda\in\mathbb{R}^{n}~{}|~{}\sum_{i=1}^{n}\arctan\lambda_{i}=C_{0}\right\}

is convex only when |C0|n22π|C_{0}|\geq\frac{n-2}{2}\pi. Another way to obtain a convexity/concaviety structure is to restrict in the range of D2u>0D^{2}u>0 as in Yuan [43]. For further relevant discussions, we refer to Warren–Yuan [40, 41], Wang–Yuan [38], Chen–Shankar–Yuan [13], Li–Li–Yuan [30], Bhattacharya–Shankar [4, 5], Bhattacharya [3] and the references therein.

For f(x)C0f(x)-C_{0} having compact support and n3n\geq 3, there are exterior Bernstein type results of (1.1) by the authors [34], which state that any classical solution with the same semi-convex conditions must be asymptotic to quadratic polynomial at infinity, together with higher order expansions that give the precise gap between exterior minimal gradient graph and the entire case. For f(x)C0f(x)-C_{0} having compact support and n=2n=2, the authors [32] proved similar exterior Bernstein type results of (1.1), which imply that exterior solutions are asymptotic to quadratic polynomial with additional ln\ln-term at infinity. When τ=0\tau=0, such results were partially proved earlier for Monge–Ampère equations by Caffarelli–Li [9] and Hong [23]. When τ=π2\tau=\frac{\pi}{2}, such results were partially proved earlier for special Lagrangian equations by Li–Li–Yuan [30]. The refined asymptotic expansions in our earlier results [34, 32] are new even for τ=0\tau=0 and π2\frac{\pi}{2} cases, it reveals that the gap between exterior minimal gradient graph and the entire case can be written into higher order errors.

For f(x)C0f(x)-C_{0} vanishing at infinity, n3n\geq 3 and τ[0,π4]\tau\in[0,\frac{\pi}{4}], there are exterior Bernstein-type results of (1.1) by the authors [33], which provide both the asymptotic behavior and finer expansions of error terms. Especially when τ=0\tau=0, the asymptotic behavior result of solutions to Monge–Ampère type equations were proved under stronger assumptions on f(x)f(x) by Bao–Li–Zhang [2].

Such asymptotic expansion for solutions of geometric curvature equations were earlier introduced in Han–Li–Li [21], which refines the previous study on the Yamabe equation and the σk\sigma_{k}-Yamabe equations by Caffarelli–Gidas–Spruck [8], Korevaar–Mazzeo–Pacard–Schoen [28], Han–Li–Teixeira [22], etc. We would also like to mention that for the Monge–Ampère type equations, there are also classification results and asymptotic behavior analysis for f(x)C0f(x)-C_{0} being a periodic function by Caffarelli–Li [10] or asymptotically periodic at infinity by Teixeira–Zhang [36], etc. Under additional assumptions on D2uD^{2}u at infinity, the asymptotic behavior results were obtained for general fully nonlinear elliptic equations by Jia [25].

In this paper, we consider the asymptotic behavior and further expansions or error terms at infinity of solutions to (1.1) with f(x)C0f(x)-C_{0} vanishing at infinity and n=2n=2. Some special structures in n=2n=2 case enable us to deal with all τ[0,π2]\tau\in[0,\frac{\pi}{2}], which is different from n3n\geq 3 case in [33]. However, there are also disadvantages caused by n=2n=2, especially the lack of equivalence on the Green’s function on unbounded domain. The asymptotic behavior obtained here is a refinement of known results of the Monge–Ampère equations by Bao–Li–Zhang [2].

Consider classical solutions of

Fτ(λ(D2u))=f(x)in 2,F_{\tau}\left(\lambda\left(D^{2}u\right)\right)=f(x)\quad\text{in }\mathbb{R}^{2}, (1.3)

where f(x)Cm(2)f(x)\in C^{m}(\mathbb{R}^{2}) converge to some constant f()f(\infty) in the sense of

lim sup|x||x|ζ+k|Dk(f(x)f())|<,k=0,1,2,,m\limsup_{|x|\rightarrow\infty}|x|^{\zeta+k}|D^{k}(f(x)-f(\infty))|<\infty,\quad\forall~{}k=0,1,2,\cdots,m (1.4)

for some ζ>2\zeta>2 and m3m\geq 3.

From the definition of FτF_{\tau} operator, λi\lambda_{i} must satisfy

{λi>0,for τ=0,λi+abλi+a+b>0,for τ(0,π4),λi1,for τ=π4,λi+a+b0,for τ(π4,π2),\left\{\begin{array}[]{llll}\lambda_{i}>0,&\text{for }\tau=0,\\ \frac{\lambda_{i}+a-b}{\lambda_{i}+a+b}>0,&\text{for }\tau\in(0,\frac{\pi}{4}),\\ \lambda_{i}\not=-1,&\text{for }\tau=\frac{\pi}{4},\\ \lambda_{i}+a+b\not=0,&\text{for }\tau\in(\frac{\pi}{4},\frac{\pi}{2}),\\ \end{array}\right.

for i=1,2i=1,2. Thus we separate the solution into semi-convex and semi-concave cases. For simplicity, we consider the semi-convex case

A>{0,τ=0,(ab)I,τ(0,π4),I,τ=π4,(a+b)I,τ(π4,π2),,τ=π2,A>\left\{\begin{array}[]{llll}0,&\tau=0,\\ -(a-b)I,&\tau\in\left(0,\frac{\pi}{4}\right),\\ -I,&\tau=\frac{\pi}{4},\\ -(a+b)I,&\tau\in\left(\frac{\pi}{4},\frac{\pi}{2}\right),\\ -\infty,&\tau=\frac{\pi}{2},\\ \end{array}\right. (1.5)

where II denotes the 2-by-2 identity matrix and the semi-concave case can be treated similarly.

Hereinafter, we assume D2uD^{2}u satisfy (1.5) in 2\mathbb{R}^{2},

f()0for τ=π2and f(){Fτ(λ(A))|A satisfies (1.5)},f(\infty)\not=0\quad\text{for }\tau=\frac{\pi}{2}\quad\text{and }f(\infty)\not\in\partial\left\{F_{\tau}(\lambda(A))~{}|~{}A\text{ satisfies }\eqref{equ:cond:semi-convex}\right\}, (1.6)

where the notation \partial denote the boundary of a set in \mathbb{R}. We may assume further without loss of generality that f()>0f(\infty)>0 for τ=π2\tau=\frac{\pi}{2} case, otherwise consider u-u instead.

Remark 1.1.

In condition (1.6), our major additional restriction is f()0f(\infty)\not=0 for τ=π2.\tau=\frac{\pi}{2}. It corresponds to the critical phase in 2\mathbb{R}^{2}, which leads to a different phenomenon than supercritical case.

If τ(0,π4]\tau\in(0,\frac{\pi}{4}] and f()=sup{Fτ(A)|A satisfies (1.5)}=0\displaystyle f(\infty)=\sup\{F_{\tau}(A)~{}|~{}A\text{ satisfies }\eqref{equ:cond:semi-convex}\}=0, then by the structure of Fτ(λ)F_{\tau}(\lambda), we have

λ1(D2u(x)),λ2(D2u(x))+as |x|.\lambda_{1}(D^{2}u(x)),\lambda_{2}(D^{2}u(x))\rightarrow+\infty\quad\text{as }|x|\rightarrow\infty.

These are not the asymptotic behavior under discussion and hence we rule out these situations by (1.6).

Let 𝚂𝚢𝚖(2)\mathtt{Sym}(2) denote the set of 2-by-2 symmetric matrix and xTx^{T} denote the transpose of a vector x2x\in\mathbb{R}^{2}. We say a scalar function φ=Ol(|x|k1(ln|x|)k2)\varphi=O_{l}\left(|x|^{-k_{1}}(\ln|x|)^{k_{2}}\right) with l,k1,k20l\in\mathbb{N},k_{1},k_{2}\geq 0 if it satisfies

|Dkφ|=O(|x|k1k(ln|x|)k2)as |x|\left|D^{k}\varphi\right|=O\left(|x|^{-k_{1}-k}(\ln|x|)^{k_{2}}\right)\quad\text{as }|x|\rightarrow\infty

for all k=0,1,2,,lk=0,1,2,\cdots,l.

Our main result shows the following asymptotic behavior and expansion result at infinity.

Theorem 1.2.

Let uC2(2)u\in C^{2}(\mathbb{R}^{2}) be a classical solution of (1.3) with D2uD^{2}u satisfying (1.5) and fCm(2)f\in C^{m}(\mathbb{R}^{2}) satisfy (1.4), (1.6) for some ζ>2\zeta>2 and m3m\geq 3. Assume further that

u(x)C(1+|x|2)in 2u(x)\leq C(1+|x|^{2})\quad\text{in }\mathbb{R}^{2} (1.7)

for some C>0C>0. Then there exist A𝚂𝚢𝚖(2)A\in\mathtt{Sym}(2) satisfying Fτ(λ(A))=f()F_{\tau}(\lambda(A))=f(\infty) and (1.5), b2,c,db\in\mathbb{R}^{2},c,d\in\mathbb{R} such that

u(x)(12xTAx+bx+c)dln(xTPx)={Om+1(|x|2min{ζ,3}),if ζ3,Om+1(|x|1(ln|x|)),if ζ=3,\begin{array}[]{lllll}&\displaystyle u(x)-\left(\frac{1}{2}x^{T}Ax+bx+c\right)-d\ln\left(x^{T}Px\right)\\ =&\displaystyle\left\{\begin{array}[]{lllll}O_{m+1}\left(|x|^{2-\min\{\zeta,3\}}\right),&\text{if }\zeta\not=3,\\ O_{m+1}\left(|x|^{-1}(\ln|x|)\right),&\text{if }\zeta=3,\\ \end{array}\right.\\ \end{array} (1.8)

as |x||x|\rightarrow\infty, where the matrix PP is given by

P=(DFτ(λ(A)))1=12(sinτA2+2cosτA+sinτI).P=(DF_{\tau}(\lambda(A)))^{-1}=\frac{1}{2}\left(\sin\tau A^{2}+2\cos\tau A+\sin\tau I\right). (1.9)

Furthermore, when ζ>3\zeta>3, there also exist d1,d2d_{1},d_{2}\in\mathbb{R} such that

u(x)(12xTAx+βx+c)dln(xTPx)=(xTPx)12(d1cosθ+d2sinθ)+{Om(|x|2ζ),if ζ<4,Om(|x|2(ln|x|)),if ζ4,\begin{array}[]{llll}&\displaystyle u(x)-\left(\frac{1}{2}x^{T}Ax+\beta x+c\right)-d\ln(x^{T}Px)\\ =&\displaystyle(x^{T}Px)^{-\frac{1}{2}}(d_{1}\cos\theta+d_{2}\sin\theta)+\left\{\begin{array}[]{llll}O_{m}(|x|^{2-\zeta}),&\text{if }\zeta<4,\\ O_{m}(|x|^{-2}(\ln|x|)),&\text{if }\zeta\geq 4,\\ \end{array}\right.\\ \end{array} (1.10)

as |x||x|\rightarrow\infty, where θ=P12x(xTPx)12\theta=\frac{P^{\frac{1}{2}}x}{(x^{T}Px)^{\frac{1}{2}}}.

Remark 1.3.

For τ=0\tau=0 case in Theorem 1.2, condition (1.7) is not necessary.

Remark 1.4.

Theorem 1.2 generalizes the asymptotic expansion results of previous work [32] by the authors, where f(x)f(x) being a constant since it corresponds to ζ=\zeta=\infty and m=m=\infty case in (1.8) and (1.10). But there are many differences in argumentation methods. By differentiating the equations we only obtain nonhomogeneous elliptic equations and inequalities on exterior domain. Furthermore, when ba2+1f(x)+π20\frac{b}{\sqrt{a^{2}+1}}f(x)+\frac{\pi}{2}\equiv 0 and τ(π4,π2)\tau\in(\frac{\pi}{4},\frac{\pi}{2}), the equation can be translated into harmonic equations. But only if ba2+1f()+π2=0\frac{b}{\sqrt{a^{2}+1}}f(\infty)+\frac{\pi}{2}=0, it yields an additional perturbation term involving the second order derivatives of uu. This leads to the difficult discussions as in (3.5). For a similar reason, τ=π2\tau=\frac{\pi}{2} case cannot be deduced into the Monge–Ampère equation detD2v=1\det D^{2}v=1 or harmonic equations by a simple change of variable as in [32]. For these perturbed cases, we turn to study the algebraic form as in (3.6) and apply iteration methods instead of using the asymptotic behavior of solutions to the Monge–Ampère equations directly.

Remark 1.5.

As in the discussions in [2, 34, 33] etc., ζ>2\zeta>2 in (1.4) is optimal in the sense that for ζ=2\zeta=2 we may construct radially symmetric solutions with u=12xTAx+O((ln|x|)2)u=\frac{1}{2}x^{T}Ax+O((\ln|x|)^{2}) as |x||x|\rightarrow\infty. Furthermore, the asymptotic expansion (1.10) is optimal in the sense that the next order term in (1.10) may contain error terms like |x|2ln|x||x|^{-2}\ln|x|, which cannot be represented into (xTPx)1(d3cos2θ+d4sin2θ)(x^{T}Px)^{-1}(d_{3}\cos 2\theta+d_{4}\sin 2\theta) for some d3,d4d_{3},d_{4}\in\mathbb{R}.

Remark 1.6.

By extension results as in Theorem 3.2 of [42], we may change the value of uu and ff on a dense subset without affecting the asymptotic behavior near infinity. Consequently by interior estimates as in Lemma 17.16 of [19], the regularity assumption on ff can be relaxed to fC0(2)f\in C^{0}(\mathbb{R}^{2}) with DmfD^{m}f exists outside a compact subset of 2\mathbb{R}^{2} for some m3m\geq 3. Especially since m3m\geq 3, we may assume without loss of generality that uC4(2)u\in C^{4}(\mathbb{R}^{2}).

The paper is organized as follows. In section 2 we prove existence results for Poisson equations on exterior domain of 2\mathbb{R}^{2}. In section 3 we prove that uu converge to a quadratic function 12xTAx\frac{1}{2}x^{T}Ax at infinity with a speed of O(|x|2ϵ)O(|x|^{2-\epsilon}) for some ϵ>0\epsilon>0, which is similar to the strategy used in [9, 2] etc. In section 4 we prove Theorem 1.2 by iteration and spherical harmonic decomposition, based on the results in sections 2 and 3.

2 Preliminary results on Poisson equations

In this section, we prove an existence result for Poisson equation on exterior domain.

Lemma 2.1.

Let gC(2)g\in C^{\infty}(\mathbb{R}^{2}) satisfy

g(r)Lp(𝕊1)c0rk1(lnr)k2r>1\|g(r\cdot)\|_{L^{p}\left(\mathbb{S}^{1}\right)}\leq c_{0}r^{-k_{1}}(\ln r)^{k_{2}}\quad\forall~{}r>1 (2.1)

for some c0>0,k1>0,k20c_{0}>0,k_{1}>0,k_{2}\geq 0 and p2p\geq 2. Then there exists a smooth solution vv of

Δv=gin 2B1¯\Delta v=g\quad\text{in }\mathbb{R}^{2}\setminus\overline{B_{1}} (2.2)

such that

|v(x)|{Cc0|x|2k1(ln|x|)k2,if k1,Cc0|x|2k1(ln|x|)k2+1,if k1{2},Cc0(ln|x|)k2+2,if k1=2,|v(x)|\leq\left\{\begin{array}[]{lllll}Cc_{0}|x|^{2-k_{1}}(\ln|x|)^{k_{2}},&\text{if }k_{1}\not\in\mathbb{N}_{*},\\ Cc_{0}|x|^{2-k_{1}}(\ln|x|)^{k_{2}+1},&\text{if }k_{1}\in\mathbb{N}_{*}\setminus\{2\},\\ Cc_{0}(\ln|x|)^{k_{2}+2},&\text{if }k_{1}=2,\\ \end{array}\right. (2.3)

for |x|>1|x|>1 for some C>0C>0.

For k1>2k_{1}>2 case, Lemma 2.1 is similar to the one proved earlier by the authors in [32]. The proof here is similar with minor modifications.

Proof of Lemma 2.1.

Here we only provide detail proof for 0<k1<10<k_{1}<1 case, the rest parts follow with minor modifications on the choice of ak,ma_{k,m} and ϵ>0\epsilon>0 below as in (2.7).

In polar coordinate we have

Δv=2vr2+1rvr+1r22vθ2,\Delta v=\dfrac{\partial^{2}v}{\partial r^{2}}+\frac{1}{r}\frac{\partial v}{\partial r}+\frac{1}{r^{2}}\frac{\partial^{2}v}{\partial\theta^{2}},

where r:=|x|r:=|x| represents the radial distance and θ:=x|x|\theta:=\frac{x}{|x|} the angle. Let

Y1(0)(θ)12π,Y1(k)(θ)=1πcoskθandY2(k)(θ)=1πsinkθ,Y_{1}^{(0)}(\theta)\equiv\dfrac{1}{\sqrt{2\pi}},\quad Y_{1}^{(k)}(\theta)=\dfrac{1}{\sqrt{\pi}}\cos k\theta\quad\text{and}\quad Y_{2}^{(k)}(\theta)=\dfrac{1}{\sqrt{\pi}}\sin k\theta,

which forms a complete standard orthogonal basis of L2(𝕊1)L^{2}(\mathbb{S}^{1}). Decompose gg and the wanted solution vv into

v(x)=a0,1(r)+k=1+m=12ak,m(r)Ym(k)(θ),g(x)=b0,1(r)+k=1+m=12bk,m(r)Ym(k)(θ),v(x)=a_{0,1}(r)+\sum_{k=1}^{+\infty}\sum_{m=1}^{2}a_{k,m}(r)Y_{m}^{(k)}(\theta),\quad g(x)=b_{0,1}(r)+\sum_{k=1}^{+\infty}\sum_{m=1}^{2}b_{k,m}(r)Y_{m}^{(k)}(\theta),

where

ak,m(r):=𝕊n1v(rθ)Ym(k)(θ)𝑑θ,bk,m(r):=𝕊n1g(rθ)Ym(k)(θ)𝑑θ.a_{k,m}(r):=\int_{\mathbb{S}^{n-1}}v(r\theta)\cdot Y_{m}^{(k)}(\theta)d\theta,\quad b_{k,m}(r):=\int_{\mathbb{S}^{n-1}}g(r\theta)\cdot Y_{m}^{(k)}(\theta)d\theta.

By the linear independence of Ym(k)(θ)Y_{m}^{(k)}(\theta), (2.2) implies that

a0,1′′(r)+1ra0,1(r)=b0,1(r)in r>1a_{0,1}^{\prime\prime}(r)+\frac{1}{r}a_{0,1}^{\prime}(r)=b_{0,1}(r)\quad\text{in }r>1

and for all kk\in\mathbb{N}_{*} with m=1,2,m=1,2,

ak,m′′(r)+1rak,m(r)k2r2ak,m(r)=bk,m(r)in r>1.a_{k,m}^{\prime\prime}(r)+\frac{1}{r}a_{k,m}^{\prime}(r)-\frac{k^{2}}{r^{2}}a_{k,m}(r)=b_{k,m}(r)\quad\text{in }r>1.

By solving the ODE, there exist constants Ck,m(1),Ck,m(2)C_{k,m}^{(1)},C_{k,m}^{(2)} such that for all r>1r>1,

ak,m(r)=Ck,m(1)rk+Ck,m(2)rk12krk2rτ1kbk,m(τ)𝑑τ+12krk2rτ1+kbk,m(τ)𝑑τ\begin{array}[]{lll}a_{k,m}(r)&=&C_{k,m}^{(1)}r^{k}+C_{k,m}^{(2)}r^{-k}\\ &&\displaystyle-\dfrac{1}{2k}r^{k}\int_{2}^{r}\tau^{1-k}b_{k,m}(\tau)d\tau+\dfrac{1}{2k}r^{-k}\int_{2}^{r}\tau^{1+k}b_{k,m}(\tau)d\tau\end{array} (2.4)

for k1k\geq 1 and

a0,1(r)=C0,1(1)+C0,1(2)lnr2rτlnτb0,1(τ)𝑑τ+lnr2rτb0,1(τ)𝑑τ.\begin{array}[]{lllll}a_{0,1}(r)&=&C_{0,1}^{(1)}+C_{0,1}^{(2)}\ln r\\ &&\displaystyle-\int_{2}^{r}\tau\ln\tau b_{0,1}(\tau)d\tau+\ln r\int_{2}^{r}\tau b_{0,1}(\tau)d\tau.\end{array}

By (2.1),

|b0,1(r)|2+k=1+m=12|bk,m(r)|2=||g(r)||L2(𝕊n1)2c02(2π)p2pr2k1(lnr)2k2|b_{0,1}(r)|^{2}+\sum_{k=1}^{+\infty}\sum_{m=1}^{2}|b_{k,m}(r)|^{2}=||g(r\cdot)||^{2}_{L^{2}(\mathbb{S}^{n-1})}\leq c_{0}^{2}(2\pi)^{\frac{p-2}{p}}r^{-2k_{1}}(\ln r)^{2k_{2}} (2.5)

for all r>1r>1. Then by 0<k1<10<k_{1}<1, we have r1kbk,m(r)L1(2,+)r^{1-k}b_{k,m}(r)\in L^{1}(2,+\infty) for all k2k\geq 2 and rk+1bk,m(r)L1(2,+)r^{k+1}b_{k,m}(r)\not\in L^{1}(2,+\infty) for all kk\in\mathbb{N}. We choose Ck,m(1)C_{k,m}^{(1)} and Ck,m(2)C_{k,m}^{(2)} in (2.4) such that

a0,1(r):=2rτlnτb0,1(τ)𝑑τ+lnr2rτb0,1(τ)𝑑τ,a_{0,1}(r):=-\int_{2}^{r}\tau\ln\tau b_{0,1}(\tau)d\tau+\ln r\int_{2}^{r}\tau b_{0,1}(\tau)d\tau,
ak,m(r):=12r2rbk,m(τ)𝑑τ+12r12rτ2bk,m(τ)𝑑τa_{k,m}(r):=-\dfrac{1}{2}r\int_{2}^{r}b_{k,m}(\tau)d\tau+\dfrac{1}{2}r^{-1}\int_{2}^{r}\tau^{2}b_{k,m}(\tau)d\tau

for k=1k=1 and

ak,m(r):=12krk+rτ1kbk,m(τ)𝑑τ+12krk2rτ1+kbk,m(τ)𝑑τa_{k,m}(r):=-\dfrac{1}{2k}r^{k}\int_{+\infty}^{r}\tau^{1-k}b_{k,m}(\tau)d\tau+\dfrac{1}{2k}r^{-k}\int_{2}^{r}\tau^{1+k}b_{k,m}(\tau)d\tau

for all k2k\geq 2.

For a0,1(r)a_{0,1}(r), we notice that there are cancellation properties as below. By (2.5) we have

|a0,1(r)|=|2rlnrττb0,1(τ)𝑑τ|Cc02rlnrττ1k1(lnτ)k2𝑑τCc0(lnr2rτ1k1(lnτ)k2𝑑τ2rτ1k1(lnτ)k2+1𝑑τ).\begin{array}[]{llll}|a_{0,1}(r)|&=&\displaystyle\left|\int_{2}^{r}\ln\frac{r}{\tau}\tau b_{0,1}(\tau)d\tau\right|\\ &\leq&\displaystyle Cc_{0}\int_{2}^{r}\ln\frac{r}{\tau}\tau^{1-k_{1}}(\ln\tau)^{k_{2}}d\tau\\ &\leq&\displaystyle Cc_{0}\left(\ln r\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}}d\tau-\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}+1}d\tau\right).\\ \end{array} (2.6)

By a direct computation, for any 0<k1<20<k_{1}<2 we have

2rτ1k1(lnτ)k2+1𝑑τ=12k1(r2k1(lnr)k2+122k1(ln2)k2+1(k2+1)2rτ1k1(lnτ)k2𝑑τ).\begin{array}[]{llll}&\displaystyle\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}+1}d\tau\\ =&\displaystyle\dfrac{1}{2-k_{1}}\left(r^{2-k_{1}}(\ln r)^{k_{2}+1}-2^{2-k_{1}}(\ln 2)^{k_{2}+1}-(k_{2}+1)\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}}d\tau\right).\end{array}

Consequently

lnr2rτ1k1(lnτ)k2𝑑τ2rτ1k1(lnτ)k2+1𝑑τ=12k1(r2k1(lnr)k2+122k1(ln2)k2(lnr)k2lnr2rτ1k1(lnτ)k21𝑑τ)12k1(r2k1(lnr)k2+122k1(ln2)k2+1(k2+1)2rτ1k1(lnτ)k2𝑑τ)k2+12k12rτ1k1(lnτ)k2𝑑τk22k1lnr2rτ1k1(lnτ)k21𝑑τ+C(1+lnr)Cr2k1(lnr)k2\begin{array}[]{llll}&\displaystyle\ln r\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}}d\tau-\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}+1}d\tau\\ =&\displaystyle\dfrac{1}{2-k_{1}}\left(r^{2-k_{1}}(\ln r)^{k_{2}+1}-2^{2-k_{1}}(\ln 2)^{k_{2}}(\ln r)-k_{2}\ln r\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}-1}d\tau\right)\\ &-\displaystyle\dfrac{1}{2-k_{1}}\left(r^{2-k_{1}}(\ln r)^{k_{2}+1}-2^{2-k_{1}}(\ln 2)^{k_{2}+1}-(k_{2}+1)\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}}d\tau\right)\\ \leq&\displaystyle\frac{k_{2}+1}{2-k_{1}}\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}}d\tau-\frac{k_{2}}{2-k_{1}}\ln r\int_{2}^{r}\tau^{1-k_{1}}(\ln\tau)^{k_{2}-1}d\tau+C(1+\ln r)\\ \leq&Cr^{2-k_{1}}(\ln r)^{k_{2}}\\ \end{array}

for some C>0C>0 for all r>2r>2. This yields

|a0,1(r)|Cc0r2k1(lnr)k2,r>2.|a_{0,1}(r)|\leq Cc_{0}r^{2-k_{1}}(\ln r)^{k_{2}},\quad\forall~{}r>2.

Similar argument also holds for k1>2k_{1}>2 case with the integrate range changed from (2,r)(2,r) into (r,+)(r,+\infty). But for k1=2k_{1}=2, from (2.6) we have no cancellation and it yields

|a0,1(r)|Cc0r2k1(lnr)k2+2,r>2.|a_{0,1}(r)|\leq Cc_{0}r^{2-k_{1}}(\ln r)^{k_{2}+2},\quad\forall~{}r>2.

For 0<k1<10<k_{1}<1 case, we pick 0<ϵ:=12min{1,22k1,2k1}0<\epsilon:=\frac{1}{2}\min\{1,2-2k_{1},2k_{1}\} such that

{32k1ϵ>1,12k1ϵ>1,32k2k1+ϵ<1,k2.\left\{\begin{array}[]{llll}3-2k_{1}-\epsilon>-1,\\ 1-2k_{1}-\epsilon>-1,\\ 3-2k-2k_{1}+\epsilon<-1,&\forall~{}k\geq 2.\\ \end{array}\right.

Then by (2.5) and Hölder inequality, we have

a0,12(r)+k=1+m=12ak,m2(r)Cc02r42k1(lnr)2k2+2k=1+m=12r2k|2rτ1+kbk,m(τ)𝑑τ|2+2m=12r2|2rb1,m(τ)𝑑τ|2+2k=2+m=12r2k|+rτ1kbk,m(τ)𝑑τ|2Cc02r42k1(lnr)2k2+2k=1+m=12r2k2rτ3+2k2k1ϵ(lnτ)2k2𝑑τ2rτ2k1(lnτ)2k2bk,m2(τ)dττ1ϵ+2m=12r22rτ12k1ϵ(lnτ)2k2𝑑τ2rτ2k1(lnτ)2k2b1,m2(τ)dττ1ϵ+2k=2+m=12r2kr+τ32k2k1+ϵ(lnτ)2k2𝑑τr+τ2k1(lnτ)2k2bk,m2(τ)dττ1+ϵCc02r42k1(lnr)2k2+Cr42k1ϵ(lnr)2k22rk=1+m=12τ2k1(lnτ)2k2bk,m2(τ)dττ1ϵ+Cr42k1+ϵ(lnr)2k2r+k=2+m=12τ2k1(lnτ)2k2bk,m2(τ)dττ1+ϵCc02r42k1(lnr)2k2\begin{array}[]{llll}&\displaystyle a_{0,1}^{2}(r)+\sum_{k=1}^{+\infty}\sum_{m=1}^{2}a_{k,m}^{2}(r)\\ \leq&\displaystyle Cc_{0}^{2}r^{4-2k_{1}}(\ln r)^{2k_{2}}+2\sum_{k=1}^{+\infty}\sum_{m=1}^{2}r^{-2k}\left|\int_{2}^{r}\tau^{1+k}b_{k,m}(\tau)d\tau\right|^{2}\\ &\displaystyle+2\sum_{m=1}^{2}r^{2}\left|\int_{2}^{r}b_{1,m}(\tau)d\tau\right|^{2}+2\sum_{k=2}^{+\infty}\sum_{m=1}^{2}r^{2k}\left|\int_{+\infty}^{r}\tau^{1-k}b_{k,m}(\tau)d\tau\right|^{2}\\ \leq&\displaystyle Cc_{0}^{2}r^{4-2k_{1}}(\ln r)^{2k_{2}}\\ &+\displaystyle 2\sum_{k=1}^{+\infty}\sum_{m=1}^{2}r^{-2k}\int_{2}^{r}\tau^{3+2k-2k_{1}-\epsilon}(\ln\tau)^{2k_{2}}d\tau\cdot\int_{2}^{r}\tau^{2k_{1}}(\ln\tau)^{-2k_{2}}b_{k,m}^{2}(\tau)\frac{d\tau}{\tau^{1-\epsilon}}\\ &+\displaystyle 2\sum_{m=1}^{2}r^{2}\int_{2}^{r}\tau^{1-2k_{1}-\epsilon}(\ln\tau)^{2k_{2}}d\tau\cdot\int_{2}^{r}\tau^{2k_{1}}(\ln\tau)^{-2k_{2}}b_{1,m}^{2}(\tau)\frac{d\tau}{\tau^{1-\epsilon}}\\ &+\displaystyle 2\sum_{k=2}^{+\infty}\sum_{m=1}^{2}r^{2k}\int^{+\infty}_{r}\tau^{3-2k-2k_{1}+\epsilon}(\ln\tau)^{2k_{2}}d\tau\cdot\int^{+\infty}_{r}\tau^{2k_{1}}(\ln\tau)^{-2k_{2}}b_{k,m}^{2}(\tau)\frac{d\tau}{\tau^{1+\epsilon}}\\ \leq&\displaystyle Cc_{0}^{2}r^{4-2k_{1}}(\ln r)^{2k_{2}}+Cr^{4-2k_{1}-\epsilon}(\ln r)^{2k_{2}}\int_{2}^{r}\sum_{k=1}^{+\infty}\sum_{m=1}^{2}\tau^{2k_{1}}(\ln\tau)^{-2k_{2}}b_{k,m}^{2}(\tau)\frac{d\tau}{\tau^{1-\epsilon}}\\ &+\displaystyle Cr^{4-2k_{1}+\epsilon}(\ln r)^{2k_{2}}\int_{r}^{+\infty}\sum_{k=2}^{+\infty}\sum_{m=1}^{2}\tau^{2k_{1}}(\ln\tau)^{-2k_{2}}b_{k,m}^{2}(\tau)\frac{d\tau}{\tau^{1+\epsilon}}\\ \leq&\displaystyle Cc_{0}^{2}r^{4-2k_{1}}(\ln r)^{2k_{2}}\\ \end{array}

for some C>0C>0 relying only on k1,k2k_{1},k_{2} and pp.

For 1<k1<21<k_{1}<2 case, we only need to change a1,ma_{1,m} with m=1,2m=1,2 into

a1,m=12r+rτb1,m(τ)dτ+12r12rτ2b1,m(τ)dτ,a_{1,m}=-\dfrac{1}{2}r\int_{+\infty}^{r}\tau b_{1,m}(\tau)\mathrm{d}\tau+\frac{1}{2}r^{-1}\int_{2}^{r}\tau^{2}b_{1,m}(\tau)\mathrm{d}\tau, (2.7)

and 0<ϵ:=12min{1,42k1,2k12}0<\epsilon:=\frac{1}{2}\min\{1,4-2k_{1},2k_{1}-2\}. The estimates on a0,12(r)+k=1+m=12ak,m2(r)\displaystyle a_{0,1}^{2}(r)+\sum_{k=1}^{+\infty}\sum_{m=1}^{2}a_{k,m}^{2}(r) follow similarly.

For k1=1k_{1}=1 case, we choose ϵ:=12\epsilon:=\frac{1}{2} and we use the following estimates of a1,ma_{1,m}.

a1,m2(r)Cc02r22rτ1(lnτ)2k2𝑑τ2rτ2(lnτ)2k2b1,m2(τ)dττ+Cc02r22rτ3(lnτ)2k2𝑑τ2rτ2(lnτ)2k2b1,m2(τ)dττCc02r2(lnτ)2k2+2.\begin{array}[]{llll}a_{1,m}^{2}(r)&\leq&\displaystyle Cc_{0}^{2}r^{2}\int_{2}^{r}\tau^{-1}(\ln\tau)^{2k_{2}}d\tau\cdot\int_{2}^{r}\tau^{2}(\ln\tau)^{-2k_{2}}b_{1,m}^{2}(\tau)\frac{d\tau}{\tau}\\ &&+\displaystyle Cc_{0}^{2}r^{-2}\int_{2}^{r}\tau^{3}(\ln\tau)^{2k_{2}}d\tau\cdot\int_{2}^{r}\tau^{2}(\ln\tau)^{-2k_{2}}b_{1,m}^{2}(\tau)\frac{d\tau}{\tau}\\ &\leq&Cc_{0}^{2}r^{2}(\ln\tau)^{2k_{2}+2}.\end{array}

The rest parts of estimate follow similarly.

For k1=2k_{1}=2 case, we choose ϵ:=12\epsilon:=\frac{1}{2} and change a1,ma_{1,m} with m=1,2m=1,2 into (2.7). In this case, the estimates of a0,1a_{0,1} shall be

a0,12(r)Cc022rτ1(lnτ)2+2k2𝑑τ2rτ4(lnτ)2k2dττ+Cc02(lnr)22rτ1(lnτ)2k2𝑑τ2rτ4(lnτ)2k2b0,12(τ)dττCc02(lnr)2k2+4.\begin{array}[]{llll}a_{0,1}^{2}(r)&\leq&\displaystyle Cc_{0}^{2}\int_{2}^{r}\tau^{-1}(\ln\tau)^{2+2k_{2}}d\tau\cdot\int_{2}^{r}\tau^{4}(\ln\tau)^{-2k_{2}}\frac{d\tau}{\tau}\\ &&\displaystyle+Cc_{0}^{2}(\ln r)^{2}\int_{2}^{r}\tau^{-1}(\ln\tau)^{2k_{2}}d\tau\cdot\int_{2}^{r}\tau^{4}(\ln\tau)^{-2k_{2}}b_{0,1}^{2}(\tau)\frac{d\tau}{\tau}\\ &\leq&Cc_{0}^{2}(\ln r)^{2k_{2}+4}.\end{array}

The rest parts of estimate follow similarly.

This proves that v(r)v(r) is well-defined, is a solution of (2.2) in distribution sense [20] and satisfies

||v(r)||L2(𝕊1)2{Cc02r42k1(lnr)2k2,if k1,Cc02r42k1(lnr)2k2+2,if k1{2},Cc02(lnr)2k2+4,if k1=2,||v(r\cdot)||^{2}_{L^{2}(\mathbb{S}^{1})}\leq\left\{\begin{array}[]{lllll}Cc_{0}^{2}r^{4-2k_{1}}(\ln r)^{2k_{2}},&\text{if }k_{1}\not\in\mathbb{N}_{*},\\ Cc_{0}^{2}r^{4-2k_{1}}(\ln r)^{2k_{2}+2},&\text{if }k_{1}\in\mathbb{N}_{*}\setminus\{2\},\\ Cc_{0}^{2}(\ln r)^{2k_{2}+4},&\text{if }k_{1}=2,\\ \end{array}\right.

By interior regularity theory of elliptic differential equations, vv is smooth [19]. Then the pointwise decay rate at infinity follows from re-scaling method and weak Harnack inequality as Theorem 8.17 of [19] etc. (see also Lemma 3.1 of [32]). ∎

Remark 2.2.

By Hölder inequality, the constant CC relying on pp in (2.3) remains finite when p=p=\infty in (2.1). Furthermore, by approximation method, the results hold for more general right hand side term other than smooth functions.

Similar to Lemma 3.2 in [34], by interior estimate, we have the following

Lemma 2.3.

Let gC(2)g\in C^{\infty}(\mathbb{R}^{2}) satisfy

g=Ol(|x|k1(ln|x|)k2)as |x|g=O_{l}(|x|^{-k_{1}}(\ln|x|)^{k_{2}})\quad\text{as }|x|\rightarrow\infty

for some k1>0,k20k_{1}>0,k_{2}\geq 0 and l1l-1\in\mathbb{N}. Then

vg={Ol+1(|x|2k1(ln|x|)k2),if k1,Ol+1(|x|2k1(ln|x|)k2+1),if k1{2},Ol+1(|x|2k1(ln|x|)k2+2),if k1=2,as |x|,v_{g}=\left\{\begin{array}[]{lllll}O_{l+1}(|x|^{2-k_{1}}(\ln|x|)^{k_{2}}),&\text{if }k_{1}\not\in\mathbb{N}_{*},\\ O_{l+1}(|x|^{2-k_{1}}(\ln|x|)^{k_{2}+1}),&\text{if }k_{1}\in\mathbb{N}_{*}\setminus\{2\},\\ O_{l+1}(|x|^{2-k_{1}}(\ln|x|)^{k_{2}+2}),&\text{if }k_{1}=2,\\ \end{array}\right.\quad\text{as }|x|\rightarrow\infty,

where vgv_{g} denotes the solution found in Lemma 2.1.

3 Quadratic part of uu at infinity

In this section, we prove a weaker asymptotic behavior than (1.8), which concerns only on the quadratic part of uu at infinity.

Lemma 3.1.

Let u,fu,f be as in Theorem 1.2. Then there exist A𝚂𝚢𝚖(2)A\in\mathtt{Sym}(2) satisfying Fτ(λ(A))=f()F_{\tau}(\lambda(A))=f(\infty) and ϵ>0\epsilon>0 such that

|u(x)12xTAx|=O2(|x|2ϵ)\left|u(x)-\frac{1}{2}x^{T}Ax\right|=O_{2}(|x|^{2-\epsilon}) (3.1)

as |x|.|x|\rightarrow\infty.

For τ=0\tau=0, the results in Lemma 3.1 were proved earlier in Bao–Li–Zhang [2]. More rigorously, in the proof of Theorem 1.2 in [2] (see also Theorem 2.2 of [33]), we have the following result for Monge–Ampère type equations, which holds for general n2n\geq 2.

Theorem 3.2.

Let uC0(n)u\in C^{0}\left(\mathbb{R}^{n}\right) be a convex viscosity solution of

detD2u=ψ(x)in n\operatorname{det}D^{2}u=\psi(x)\quad\text{in }\mathbb{R}^{n}

with u(0)=minnu=0u(0)=\min_{\mathbb{R}^{n}}u=0, where 0<ψC0(n)0<\psi\in C^{0}\left(\mathbb{R}^{n}\right) and

ψ1n1Ln(n).\psi^{\frac{1}{n}}-1\in L^{n}\left(\mathbb{R}^{n}\right).

Then there exists a linear transform TT satisfying detT=1\operatorname{det}T=1 such that v:=uTv:=u\circ T satisfies

|v12|x|2|C|x|2ε,|x|1\left|v-\dfrac{1}{2}|x|^{2}\right|\leq C|x|^{2-\varepsilon},\quad\forall~{}|x|\geq 1

for some C>0C>0 and ε>0\varepsilon>0.

For τ(0,π4]\tau\in(0,\frac{\pi}{4}] cases, the results follow from Legendre transform and the asymptotic behavior for the Monge–Ampère type equations and Poisson equations. More explicitly, as proved for general n2n\geq 2 cases in Theorem 2.1 and Remark 2.6 of [33], we have the following result.

Theorem 3.3.

Let uC2(2)u\in C^{2}(\mathbb{R}^{2}) be a classical solution of (1.3) with τ[0,π4]\tau\in[0,\frac{\pi}{4}] and fC1(2)f\in C^{1}(\mathbb{R}^{2}) satisfy

|x|ζ|f(x)f()|+|x|1+ζ|Df(x)|C,|x|>1|x|^{\zeta}|f(x)-f(\infty)|+|x|^{1+\zeta^{\prime}}\left|Df(x)\right|\leq C,\quad\forall~{}|x|>1 (3.2)

for some C>0C>0 with ζ>1,ζ>0\zeta>1,\zeta^{\prime}>0 for τ[0,π4)\tau\in[0,\frac{\pi}{4}) and ζ,ζ>0\zeta,\zeta^{\prime}>0 for τ=π4\tau=\frac{\pi}{4}. Let D2uD^{2}u and f()f(\infty) satisfy (1.5) and (1.6) respectively. For τ(0,π4]\tau\in(0,\frac{\pi}{4}], we assume further that uu satisfies (1.7). Then there exist ϵ>0,A𝚂𝚢𝚖(2)\epsilon>0,A\in\mathtt{Sym}(2) satisfying Fτ(λ(A))=f()F_{\tau}(\lambda(A))=f(\infty) and (1.5) and C>0C>0 such that

D2uCα(2)Cand|D2u(x)A|C|x|ϵ,|x|>1.||D^{2}u||_{C^{\alpha}(\mathbb{R}^{2})}\leq C\quad\text{and}\quad\left|D^{2}u(x)-A\right|\leq\dfrac{C}{|x|^{\epsilon}},\quad\forall~{}|x|>1.
Proof of Lemma 3.1 with 0τπ40\leq\tau\leq\frac{\pi}{4}.

In Lemma 3.1, we have fCm(2)f\in C^{m}(\mathbb{R}^{2}) satisfies (1.4) for some ζ>2\zeta>2 and m3m\geq 3. Especially fC1(2)f\in C^{1}(\mathbb{R}^{2}) satisfies condition (3.2) with ζ=ζ>2\zeta^{\prime}=\zeta>2. By Theorem 3.3 we have D2uCα(2)||D^{2}u||_{C^{\alpha}(\mathbb{R}^{2})} is bounded for all 0<α<10<\alpha<1 and D2u(x)D^{2}u(x) converge to matrix AA at Hölder speed |x|ϵ|x|^{-\epsilon}. By Newton–Leibnitz formula, since Du,uDu,u are bounded on B1\partial B_{1}, for any |x|>1|x|>1 we let e:=x|x|B1e:=\frac{x}{|x|}\in\partial B_{1} to obtain

|Du(x)Du(e)0|x|1A𝑑se|=|0|x|1D2u((s+1)e)Adse|C|x|1ϵ.\begin{array}[]{llllll}\displaystyle\left|Du(x)-Du(e)-\int_{0}^{|x|-1}Ads\cdot e\right|&=&\displaystyle\left|\int_{0}^{|x|-1}D^{2}u((s+1)e)-Ads\cdot e\right|\\ &\leq&\displaystyle C|x|^{1-\epsilon}.\\ \end{array}

Consequently there exists C>0C>0 such that

|Du(x)Ax|C|x|1ϵ,|x|>1.\left|Du(x)-Ax\right|\leq C|x|^{1-\epsilon},\quad\forall~{}|x|>1.

By Newton–Leibnitz formula again we have

|u(x)u(e)0|x|1xTAe𝑑s|=|0|x1|(Du((s+1)e)Ax)e𝑑s|C|x|2ϵ,|x|>1\begin{array}[]{llll}\displaystyle\left|u(x)-u(e)-\int_{0}^{|x|-1}x^{T}A\cdot eds\right|&=&\displaystyle\left|\int_{0}^{|x-1|}\left(Du((s+1)e)-Ax\right)\cdot eds\right|\\ &\leq&C|x|^{2-\epsilon},\quad\forall~{}|x|>1\end{array}

for some C>0C>0. This finishes the proof of Lemma 3.1 with τ[0,π4]\tau\in[0,\frac{\pi}{4}]. ∎

Proof of Lemma 3.1 with π4<τ<π2\frac{\pi}{4}<\tau<\frac{\pi}{2}.

By semi-convex condition (1.5), we have λi>(a+b)\lambda_{i}>-(a+b) for i=1,2i=1,2 and consequently by a direct computation as in [24, 39],

arctanλi+abλi+a+b=arctanλi+abπ4.\arctan\frac{\lambda_{i}+a-b}{\lambda_{i}+a+b}=\arctan\frac{\lambda_{i}+a}{b}-\frac{\pi}{4}.

Consequently Eq. (1.3) with τ(π4,π2)\tau\in(\frac{\pi}{4},\frac{\pi}{2}) and semi-convex condition D2u>(a+b)ID^{2}u>-(a+b)I becomes

arctanλ1(D2u)+ab+arctanλ2(D2u)+ab=ba2+1f(x)+π2in 2.\arctan\dfrac{\lambda_{1}(D^{2}u)+a}{b}+\arctan\dfrac{\lambda_{2}(D^{2}u)+a}{b}=\dfrac{b}{\sqrt{a^{2}+1}}f(x)+\frac{\pi}{2}\quad\text{in }\mathbb{R}^{2}.

Let v:=1b(u+a2|x|2)v:=\frac{1}{b}(u+\frac{a}{2}|x|^{2}), then vv satisfies (1.7) for some new C>0C>0 and the Lagrangian mean curvature equation

arctanλ1(D2v)+arctanλ2(D2v)=ba2+1f(x)+π2in 2.\arctan\lambda_{1}(D^{2}v)+\arctan\lambda_{2}(D^{2}v)=\dfrac{b}{\sqrt{a^{2}+1}}f(x)+\frac{\pi}{2}\quad\text{in }\mathbb{R}^{2}. (3.3)

If ba2+1f()π2\frac{b}{\sqrt{a^{2}+1}}f(\infty)\not=-\frac{\pi}{2}, we may assume without loss of generality that ba2+1f()+π2>0\frac{b}{\sqrt{a^{2}+1}}f(\infty)+\frac{\pi}{2}>0, otherwise we consider the equation satisfied by v-v as replacement. Then the desired result (3.1) follows from τ=π2\tau=\frac{\pi}{2} case, which will be proved below.

It remains to prove for ba2+1f()+π2=0\frac{b}{\sqrt{a^{2}+1}}f(\infty)+\frac{\pi}{2}=0 case. For any sufficiently small δ>0\delta>0, there exists R0>0R_{0}>0 such that

|ba2+1f(x)+π2|<δ,|x|>R0.\left|\dfrac{b}{\sqrt{a^{2}+1}}f(x)+\frac{\pi}{2}\right|<\delta,\quad\forall~{}|x|>R_{0}.

Since D2v>ID^{2}v>-I, together with the continuity of D2uD^{2}u in 2\mathbb{R}^{2}, we have D2vD^{2}v bounded on entire 2\mathbb{R}^{2}. Consequently D2uD^{2}u is bounded on entire 2\mathbb{R}^{2}. For any sufficiently large |x|>1|x|>1, we set

R:=|x|anduR(y):=4R2u(x+R2y)in B1.R:=|x|\quad\text{and}\quad u_{R}(y):=\dfrac{4}{R^{2}}u(x+\frac{R}{2}y)\quad\text{in }B_{1}.

Then uRu_{R} is a classical solution of

Fτ(λ(D2uR(y)))=f(x+R2y)=:fR(y)in B1.F_{\tau}(\lambda(D^{2}u_{R}(y)))=f(x+\frac{R}{2}y)=:f_{R}(y)\quad\text{in }B_{1}.

Since D2uRD^{2}u_{R} are uniformly (to RR) bounded, the equations above are uniformly elliptic. Consequently, by the definition of uRu_{R}, we have uRC0(B1)||u_{R}||_{C^{0}(B_{1})} are also uniformly bounded. By interior Hölder estimates for second derivatives as in Theorem 17.11 of [19], we have

D2uRCα(B12)C||D^{2}u_{R}||_{C^{\alpha}(B_{\frac{1}{2}})}\leq C (3.4)

for some α,C>0\alpha,C>0 uniform to sufficiently large RR. Now we turn to the algebraic form of (3.3) i.e.,

Δv=(1detD2v)tan(ba2+1f(x)+π2)=:g(x)\Delta v=(1-\det D^{2}v)\cdot\tan\left(\frac{b}{\sqrt{a^{2}+1}}f(x)+\frac{\pi}{2}\right)=:g(x) (3.5)

in 2\mathbb{R}^{2}. By approximation and Lemma 2.1, there exists a solution vgv_{g} solving (3.5) on 2B1¯\mathbb{R}^{2}\setminus\overline{B_{1}} with estimate

|vg(x)|C|x|2ζ(ln|x|),|x|>1|v_{g}(x)|\leq C|x|^{2-\zeta}(\ln|x|),\quad\forall~{}|x|>1

for some C>0C>0. Let vR(y):=vg(x+R2y)v_{R}(y):=v_{g}(x+\frac{R}{2}y) in B1B_{1}, where R=|x|>2R=|x|>2, then

ΔvR=g(x+R2y)=:gR(y)in B1.\Delta v_{R}=g(x+\frac{R}{2}y)=:g_{R}(y)\quad\text{in }B_{1}.

By conditions (1.4) and (3.4), we have

vRC0(B1)CRζlnRandgRCα(B12)CRαζ||v_{R}||_{C^{0}(B_{1})}\leq CR^{-\zeta}\ln R\quad\text{and}\quad||g_{R}||_{C^{\alpha}(B_{\frac{1}{2}})}\leq CR^{\alpha-\zeta}

for some C>0C>0 uniform to R>2R>2. By interior Schauder estimates, we have

vRC2,α(B13)CRϵ||v_{R}||_{C^{2,\alpha}(B_{\frac{1}{3}})}\leq CR^{-\epsilon^{\prime}}

for some 0<α,ϵ<10<\alpha,\epsilon^{\prime}<1 and C>0C>0. Consequently there exists C>0C>0 such that

vg=O2(|x|2ϵ)as |x|.v_{g}=O_{2}(|x|^{2-\epsilon^{\prime}})\quad\text{as }|x|\rightarrow\infty.

Since vvgv-v_{g} is harmonic on 2B1¯\mathbb{R}^{2}\setminus\overline{B_{1}} with bounded Hessian matrix, by spherical harmonic expansion as in (2.4), there exists Av𝚂𝚢𝚖(2),βv2A_{v}\in\mathtt{Sym}(2),\beta_{v}\in\mathbb{R}^{2} and cv,dvc_{v},d_{v}\in\mathbb{R} such that

vvg=12xTAvx+βvx+dvln|x|+cv+Ol(|x|1)v-v_{g}=\frac{1}{2}x^{T}A_{v}x+\beta_{v}\cdot x+d_{v}\ln|x|+c_{v}+O_{l}(|x|^{-1})

for all ll\in\mathbb{N} as |x||x|\rightarrow\infty. Combining the two asymptotic behavior above, we have

u(x)=bv(x)a2|x|2=12xT(bAvaI)x+bβvx+bdvln|x|+bcv+bvg+O2(|x|1)=12xT(bAvaI)x+O2(|x|2ϵ)\begin{array}[]{llll}u(x)&=&bv(x)-\frac{a}{2}|x|^{2}\\ &=&\frac{1}{2}x^{T}(bA_{v}-aI)x+b\beta_{v}\cdot x+bd_{v}\ln|x|+bc_{v}+bv_{g}+O_{2}(|x|^{-1})\\ &=&\frac{1}{2}x^{T}(bA_{v}-aI)x+O_{2}(|x|^{2-\epsilon^{\prime}})\\ \end{array}

as |x||x|\rightarrow\infty. This finishes the proof of (3.1). ∎

Proof of Lemma 3.1 with τ=π2\tau=\frac{\pi}{2}.

Consider the algebraic form of Eq. (1.3) with τ=π2\tau=\frac{\pi}{2} i.e.,

cosf(x)Δu+sinf(x)detD2u=sinf(x)\cos f(x)\cdot\Delta u+\sin f(x)\det D^{2}u=\sin f(x)

in 2\mathbb{R}^{2}. By condition (1.6), cotf()0\cot f(\infty)\not=0 and consequently we have

detD2u+cotf()Δu=1+(cotf()cotf(x))Δu\det D^{2}u+\cot f(\infty)\cdot\Delta u=1+\left(\cot f(\infty)-\cot f(x)\right)\Delta u

in 2\mathbb{R}^{2}. Change of variable by setting

v(x):=u(x)+cotf()2|x|2,v(x):=u(x)+\frac{\cot f(\infty)}{2}|x|^{2},

which satisfies

detD2v=(λ1(D2u)+cotf())(λ2(D2u)+cotf())=1+cot2f()+(cotf()cotf(x))Δu=:g(x)\begin{array}[]{lll}\det D^{2}v&=&\displaystyle(\lambda_{1}(D^{2}u)+\cot f(\infty))\cdot(\lambda_{2}(D^{2}u)+\cot f(\infty))\\ &=&\displaystyle 1+\cot^{2}f(\infty)+\left(\cot f(\infty)-\cot f(x)\right)\Delta u\\ &=:&g(x)\\ \end{array} (3.6)

in 2\mathbb{R}^{2}. To obtain the desired results, we shall obtain the asymptotic behavior of g(x)g(x) at infinity and apply Theorem 3.2.

Step 1: We prove the boundedness of D2uD^{2}u by interior Hessian estimate.

Since f()(0,π)f(\infty)\in(0,\pi), for any sufficiently small 0<δ<f()0<\delta<f(\infty), there exists R0>0R_{0}>0 such that

|f(x)f()|<δ,|x|>R0.|f(x)-f(\infty)|<\delta,\quad\forall~{}|x|>R_{0}.

By Eq. (1.3) with τ=π2\tau=\frac{\pi}{2}, for all i=1,2i=1,2, we have

arctanλi+π2>arctanλ1+arctanλ2>f()δ>0,|x|>R0.\arctan\lambda_{i}+\frac{\pi}{2}>\arctan\lambda_{1}+\arctan\lambda_{2}>f(\infty)-\delta>0,\quad\forall~{}|x|>R_{0}.

Consequently by the monotonicity of arctan\arctan function, we have

D2u>cot(f()δ)I|x|>R0.D^{2}u>-\cot(f(\infty)-\delta)I\quad\forall~{}|x|>R_{0}.

By (1.7) and the quadratic growth condition from above, there exists C>0C>0 such that

|u(x)|C(1+|x|2),x2.|u(x)|\leq C(1+|x|^{2}),\quad\forall~{}x\in\mathbb{R}^{2}. (3.7)

For sufficiently large |x|>2R0|x|>2R_{0}, we set

R:=|x|>2R0anduR(y):=4R2u(x+R2y),yB1,R:=|x|>2R_{0}\quad\text{and}\quad u_{R}(y):=\dfrac{4}{R^{2}}u(x+\frac{R}{2}y),\quad y\in B_{1}, (3.8)

where Br(x)B_{r}(x) denote the ball centered at xx with radius rr and Br:=Br(0)B_{r}:=B_{r}(0). Then uRC4(B1)u_{R}\in C^{4}(B_{1}) satisfies

arctanλ1(D2uR)+arctanλ2(D2uR)=f(x+R2y)=:fR(y)in B1.\arctan\lambda_{1}(D^{2}u_{R})+\arctan\lambda_{2}(D^{2}u_{R})=f(x+\frac{R}{2}y)=:f_{R}(y)\quad\text{in }B_{1}. (3.9)

By a direct computation, (3.7) implies that there exists a constant CC uniform to R>2R0R>2R_{0} such that

supB1|uR|4R2supB|x|2(x)|u|C.\sup_{B_{1}}|u_{R}|\leq\dfrac{4}{R^{2}}\sup_{B_{\frac{|x|}{2}}(x)}|u|\leq C.

Together with condition (1.4) on ff with ζ>2\zeta>2 and m3m\geq 3, we have

fRC1,1(B1)=supB1(|fR|+|DfR|)+supy,zB1yz|DfR(y)DfR(z)||yz|supB1(|fR|+|DfR|+|D2fR|)CRζ,\begin{array}[]{lllll}||f_{R}||_{C^{1,1}(B_{1})}&=&\displaystyle\sup_{B_{1}}\left(|f_{R}|+|Df_{R}|\right)+\sup_{y,z\in B_{1}\atop y\not=z}\dfrac{|Df_{R}(y)-Df_{R}(z)|}{|y-z|}\\ &\leq&\displaystyle\sup_{B_{1}}\left(|f_{R}|+|Df_{R}|+|D^{2}f_{R}|\right)\\ &\leq&CR^{-\zeta},\\ \end{array}

for some constant C>0C>0 uniform to R>2R0R>2R_{0}. Furthermore, since

supB1|fR(y)f()|0as R.\sup_{B_{1}}\left|f_{R}(y)-f(\infty)\right|\rightarrow 0\quad\text{as }R\rightarrow\infty.

By f()>0f(\infty)>0, there exists δ>0\delta>0 uniform to sufficiently large RR such that |fR|δ>0|f_{R}|\geq\delta>0.

Now we introduce the following interior Hessian estimates for Lagrangian mean curvature equations as in Theorems 1.1 and 1.2 by Bhattacharya [3].

Theorem 3.4.

Let uu be a C4C^{4} solution of (1.2) in BrnB_{r}\subset\mathbb{R}^{n}, where fC1,1(Br)f\in C^{1,1}(B_{r}) and

|f|n22π+δ|f|\geq\frac{n-2}{2}\pi+\delta (3.10)

for some δ>0\delta>0. Then we have

|D2u(0)|C1exp(C2r2n2(oscBr(0)ur+1)2n2),\left|D^{2}u(0)\right|\leq C_{1}\exp\left(\frac{C_{2}}{r^{2n-2}}\left(\underset{B_{r}(0)}{osc}\frac{u}{r}+1\right)^{2n-2}\right),

where C1,C2C_{1},C_{2} are positive constants depending on fC1,1(Br),n\|f\|_{C^{1,1}\left(B_{r}\right)},n, and δ\delta.

Applying Theorem 3.4 to the equation satisfies by uRu_{R} in B1B_{1}, there exists C>0C>0 uniform to R>2R0R>2R_{0} such that

|D2uR(0)|C1exp(C2(oscB1(0)uR+1)2)C.|D^{2}u_{R}(0)|\leq C_{1}\exp\left(C_{2}\left(\underset{B_{1}(0)}{osc}u_{R}+1\right)^{2}\right)\leq C.

Consequently D2uD^{2}u is bounded on entire 2\mathbb{R}^{2}.

Step 2: Now we compute the asymptotic behavior of g(x)g(x) at infinity.

By the definition of g(x)g(x) in Eq. (3.6) and condition (1.4), since D2uD^{2}u is bounded on entire 2\mathbb{R}^{2}, we have

g(x)(1+cot2f())=(cotf()cotf(x))ΔuCsin2f()|f(x)f()|C|x|ζ\begin{array}[]{llll}g(x)-(1+\cot^{2}f(\infty))&=&(\cot f(\infty)-\cot f(x))\Delta u\\ &\leq&\dfrac{C}{\sin^{2}f(\infty)}|f(x)-f(\infty)|\\ &\leq&C|x|^{-\zeta}\\ \end{array}

for some constant C>0C>0. Consequently there exists C>0C>0 such that

2B1¯|g12(x)(1+cot2f())12|2𝑑xC2B1¯|g(x)(1+cot2f())|2𝑑xC2B1¯|x|2ζ𝑑x<.\begin{array}[]{lllll}&\displaystyle\int_{\mathbb{R}^{2}\setminus\overline{B_{1}}}\left|g^{\frac{1}{2}}(x)-(1+\cot^{2}f(\infty))^{\frac{1}{2}}\right|^{2}dx\\ \leq&\displaystyle C\int_{\mathbb{R}^{2}\setminus\overline{B_{1}}}|g(x)-(1+\cot^{2}f(\infty))|^{2}dx\\ \leq&\displaystyle C\int_{\mathbb{R}^{2}\setminus\overline{B_{1}}}|x|^{-2\zeta}dx\\ <&\infty.\\ \end{array}

Step 3: Obtain the asymptotic behavior of vv.

Since 1+cot2f()>01+\cot^{2}f(\infty)>0, by the continuity of D2vD^{2}v and n=2n=2, we have either D2v>0D^{2}v>0 or D2v<0D^{2}v<0 for sufficiently large |x|>2R0|x|>2R_{0}. We may assume without loss of generality that D2v>0D^{2}v>0, otherwise we consider v-v instead. Thus by extension result as Theorem 3.2 of [42], we can apply Theorem 3.2 (see also Corollary 2.3 in [33]) after re-scaling v~:=1(1+cot2f())12v\widetilde{v}:=\frac{1}{(1+\cot^{2}f(\infty))^{\frac{1}{2}}}v, there exist Av𝚂𝚢𝚖(2)A_{v}\in\mathtt{Sym}(2) satisfying detAv=1+cot2f()\det A_{v}=1+\cot^{2}f(\infty) such that

|v12xTAvx|C|x|2ϵ|x|>1\left|v-\dfrac{1}{2}x^{T}A_{v}x\right|\leq C|x|^{2-\epsilon}\quad\forall~{}|x|>1

for some C>0C>0 and ϵ>0\epsilon>0. By the definition of vv, we have

|u12xT(Avcotf()I)x|C|x|2ϵ|x|>1.\left|u-\frac{1}{2}x^{T}(A_{v}-\cot f(\infty)I)x\right|\leq C|x|^{2-\epsilon}\quad\forall~{}|x|>1.

By taking A:=Avcotf()IA:=A_{v}-\cot f(\infty)I, it is easy to verify that

cosf()trace(A)+sinf()detA=sinf(),\cos f(\infty)\operatorname{trace}(A)+\sin f(\infty)\det A=\sin f(\infty),

and hence arctanλ1(A)+arctanλ2(A)=f()\arctan\lambda_{1}(A)+\arctan\lambda_{2}(A)=f(\infty) and the first part of the desired result in Lemma 3.1 follows immediately.

Step 4: We prove interior gradient and Hessian estimates by scaling.

Let uR,fRu_{R},f_{R} be as in (3.8) and (3.9). From the results in Step 1, D2uD^{2}u is bounded on entire 2\mathbb{R}^{2}. Consequently equations (3.9) are uniformly (to R>2R0R>2R_{0}) elliptic. By interior Hölder estimates for second derivatives as in Theorem 17.11 of [19], we have

[D2uR]Cα(B12)C[D^{2}u_{R}]_{C^{\alpha}(B_{\frac{1}{2}})}\leq C

for some α,C>0\alpha,C>0, where α\alpha relies only on D2uRC0(B1)||D^{2}u_{R}||_{C^{0}(B_{1})} and CC relies only on uRC2(B1)||u_{R}||_{C^{2}(B_{1})} and fRC1(B1)||f_{R}||_{C^{1}(B_{1})}. Consequently

D2uRCα(B12)<C||D^{2}u_{R}||_{C^{\alpha}(B_{\frac{1}{2}})}<C

for some C>0C>0 for all R>2R0R>2R_{0}. We would like to mention that n=2n=2 is necessary to apply interior estimates as Theorem 17.11 in [19]. For higher dimensions, it is generally required that the operator has a concavity structure (see for instance Theorem 17.14 in [19] and Theorem 8.1 in [7]).

To obtain the desired result, it remains to prove the gradient and Hessian estimate on the difference between uu and 12xTAx\frac{1}{2}x^{T}Ax. Let

w(x):=u(x)12xTAxandwR(y):=4R2w(x+R2y)yB1.w(x):=u(x)-\frac{1}{2}x^{T}Ax\quad\text{and}\quad w_{R}(y):=\frac{4}{R^{2}}w(x+\frac{R}{2}y)\quad\forall~{}y\in B_{1}. (3.11)

From the results in Steps 1-4, there exists C>0C>0 uniform to all R>2R0R>2R_{0} such that

uRL(B1)CandwRL(B1)CRϵ.||u_{R}||_{L^{\infty}(B_{1})}\leq C\quad\text{and}\quad||w_{R}||_{L^{\infty}(B_{1})}\leq CR^{-\epsilon}.

Applying Newton–Leibnitz formula between (3.9) and arctanλ1(A)+arctanλ2(A)=f()\arctan\lambda_{1}(A)+\arctan\lambda_{2}(A)=f(\infty), we have

aijRDijwR=fR(y)f()in B1,a_{ij}^{R}D_{ij}w_{R}=f_{R}(y)-f(\infty)\quad\text{in }B_{1}, (3.12)

where

aijR(y)=01DMijFτ(A+tD2wR(y))𝑑t,a_{ij}^{R}(y)=\int_{0}^{1}D_{M_{ij}}F_{\tau}(A+tD^{2}w_{R}(y))dt,

are uniformly elliptic and having uniformly bounded (to R>2R0R>2R_{0}) CαC^{\alpha} norm. Hereinafter, we let [aij][a_{ij}] denote the 2-by-2 matrix with the i,ji,j-position being aija_{ij} and DMijFτ(M)D_{M_{ij}}F_{\tau}(M) denote the value of partial derivative of Fτ(λ(M))F_{\tau}(\lambda(M)) with respect to MijM_{ij} variable at M=[Mij]M=[M_{ij}]. By condition (1.4), there exists uniform C>0C>0 such that

fR1L(B1)+k=1m1DkfRCα(B1)CRζ.\left\|f_{R}-1\right\|_{L^{\infty}\left(B_{1}\right)}+\sum_{k=1}^{m-1}\left\|D^{k}f_{R}\right\|_{C^{\alpha}\left(B_{1}\right)}\leq CR^{-\zeta}. (3.13)

By interior Schauder estimates as Theorem 6.2 in [19],

wRC2,α(B12)C(wRL(B1)+fRf()Cα(B1¯))CRmin{ϵ,ζ}.||w_{R}||_{C^{2,\alpha}(B_{\frac{1}{2}})}\leq C\left(||w_{R}||_{L^{\infty}(B_{1})}+||f_{R}-f(\infty)||_{C^{\alpha}(\overline{B_{1}})}\right)\leq CR^{-\min\{\epsilon,\zeta\}}.

This finishes the proof of Lemma 3.1 by choosing ϵ\epsilon as the minimum of ϵ\epsilon and ζ\zeta. ∎

4 Asymptotic behavior and expansions of uu at infinity

In this section, we prove the asymptotic behavior and expansions of solution uu at infinity following the line of iteration method as by Bao–Li–Zhang [2] with an improvement from spherical harmonic expansion.

Lemma 4.1.

Let u,fu,f be as in Theorem 1.2 and

w(x):=u(x)12xTAx,w(x):=u(x)-\frac{1}{2}x^{T}Ax,

where A𝚂𝚢𝚖(2)A\in\mathtt{Sym}(2) is from Lemma 3.1. Then there exist C,α,ϵ>0C,\alpha,\epsilon^{\prime}>0 such that

{|Dkw(x)|C|x|2kϵ,|Dm+1w(x1)Dm+1w(x2)||x1x2|αC|x1|1mϵα,\left\{\begin{array}[]{ l }{\left|D^{k}w(x)\right|\leq C|x|^{2-k-\epsilon^{\prime}},}\\ {\frac{\left|D^{m+1}w\left(x_{1}\right)-D^{m+1}w\left(x_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\leq C\left|x_{1}\right|^{1-m-\epsilon^{\prime}-\alpha},}\end{array}\right. (4.1)

for all |x|>2,k=0,,m+1|x|>2,k=0,\ldots,m+1 and |x1|>2,x2B|x1|/2(x1)\left|x_{1}\right|>2,x_{2}\in B_{\left|x_{1}\right|/2}\left(x_{1}\right).

Proof.

For sufficiently large |x|>1|x|>1, we set R,uR,fR,wRR,u_{R},f_{R},w_{R} as in the proof of Lemma 3.1 i.e., (3.8), (3.9) and (3.11). As proved in Lemma 3.1, together with uC2(2)u\in C^{2}(\mathbb{R}^{2}) we have

|Dkw(x)|C|x|2kϵ,|x|>1|D^{k}w(x)|\leq C|x|^{2-k-\epsilon},\quad\forall~{}|x|>1

for some C>0C>0 and ϵ>0\epsilon>0 from Lemma 3.1 for k=0,1,2k=0,1,2. It remains to prove the higher order derivatives following the Step 4 in the proof of Lemma 3.1.

In fact, we consider the scaled equation

Fτ(λ(D2uR))=fR(y)in B1.F_{\tau}(\lambda(D^{2}u_{R}))=f_{R}(y)\quad\text{in }B_{1}. (4.2)

Since D2uRD^{2}u_{R} are uniformly (to RR) bounded, Fτ(λ(D2uR))F_{\tau}(\lambda(D^{2}u_{R})) are uniformly elliptic. By Theorem 17.11 in [19] there exist α,C,R0>0\alpha,C,R_{0}>0 such that

D2uRCα(B23)C,R>2.||D^{2}u_{R}||_{C^{\alpha}(B_{\frac{2}{3}})}\leq C,\quad\forall~{}R>2. (4.3)

Apply Newton–Leibnitz formula between (4.2) and Fτ(λ(A))=f()F_{\tau}(\lambda(A))=f(\infty) to obtain linearized equation (3.12). By the boundedness of D2uRD^{2}u_{R} and estimate (4.3), the coefficients aijR(y)a_{ij}^{R}(y) are uniformly elliptic and having finite CαC^{\alpha}-norm for some α>0\alpha>0 uniform to R>2R>2. Together with (3.13), by interior Schauder estimates we have

wRC2,α(B12)C(wRL(B23)+fRf()Cα(B23¯))CRmin{ϵ,ζ}||w_{R}||_{C^{2,\alpha}(B_{\frac{1}{2}})}\leq C\left(||w_{R}||_{L^{\infty}(B_{\frac{2}{3}})}+||f_{R}-f(\infty)||_{C^{\alpha}(\overline{B_{\frac{2}{3}}})}\right)\leq CR^{-\min\{\epsilon,\zeta\}}

for some C>0C>0 uniform to all R>2R>2.

It remains to obtain higher order derivative estimates. For any eB1e\in\partial B_{1}, we act partial derivative DeD_{e} to both sides of (4.2) and obtain

a~ijRDij(DewR)=DefR(y)in B1,wherea~ijR(y):=DMijFτ(D2uR(y)).\widetilde{a}_{ij}^{R}D_{ij}(D_{e}w_{R})=D_{e}f_{R}(y)\quad\text{in }B_{1},\quad\text{where}\quad\widetilde{a}_{ij}^{R}(y):=D_{M_{ij}}F_{\tau}(D^{2}u_{R}(y)).

By the boundedness of D2uRD^{2}u_{R} and estimate (4.3), the coefficients a~ijR\widetilde{a}_{ij}^{R} are uniformly elliptic with uniformly bounded CαC^{\alpha}-norm to all R>2R>2. By interior Schauder estimate again and the arbitrariness of eB1e\in\partial B_{1}, we have

wRC3,α(B¯1/3)C(wRL(B¯2/3)+|Df1,RCα(B¯2/3))CRmin{ϵ,ζ}.\left\|w_{R}\right\|_{C^{3,\alpha}\left(\overline{B}_{1/3}\right)}\leq C\left(\left\|w_{R}\right\|_{L^{\infty}\left(\overline{B}_{2/3}\right)}+|Df_{1,R}\|_{C^{\alpha}\left(\overline{B}_{2/3}\right)}\right)\leq CR^{-\min\{\epsilon,\zeta\}}.

By taking further derivatives of Eq. (4.2), higher order derivatives follow from Schauder estimate and this finishes the proof of (4.1). ∎

Next we prove an iteration type lemma that improves the estimates in Lemma 4.1. The result follows as in Lemma 2.2 in Bao–Li-Zhang [2] with minor modifications.

Lemma 4.2.

Let uC2(2)u\in C^{2}(\mathbb{R}^{2}) be a classical solution of

F(D2u)=f(x)in 2.F(D^{2}u)=f(x)\quad\text{in }\mathbb{R}^{2}. (4.4)

Suppose FF is smooth up to the boundary of the range of D2uD^{2}u and is uniformly elliptic. Suppose further that fCm(2)f\in C^{m}(\mathbb{R}^{2}) satisfies (1.4) for some ζ>2,m3\zeta>2,m\geq 3. If there exists a quadratic polynomial vv satisfying F(D2v)=f()F(D^{2}v)=f(\infty) and w:=uvw:=u-v satisfies

{|Dkw(x)|C|x|2ϵk(ln|x|)p0,|Dm+1w(x1)Dm+1w(x2)||x1x2|αC|x1|1mϵα(ln|x|)p0\left\{\begin{array}[]{ l }{\left|D^{k}w(x)\right|\leq C|x|^{2-\epsilon-k}(\ln|x|)^{p_{0}},}\\ {\frac{\left|D^{m+1}w\left(x_{1}\right)-D^{m+1}w\left(x_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\leq C\left|x_{1}\right|^{1-m-\epsilon-\alpha}(\ln|x|)^{p_{0}}}\end{array}\right. (4.5)

for some 0<ϵ<120<\epsilon<\frac{1}{2} and p0p_{0}\in\mathbb{N} for all |x|>2,k=0,,m+1|x|>2,k=0,\ldots,m+1 and |x1|>2,x2B|x1|/2(x1)\left|x_{1}\right|>2,x_{2}\in B_{\left|x_{1}\right|/2}\left(x_{1}\right). Then

{|Dkw(x)|C|x|22ϵk(ln|x|)2p0,|Dm+1w(x1)Dm+1w(x2)||x1x2|αC|x1|1m2ϵα(ln|x|)2p0,\left\{\begin{array}[]{ l }{\left|D^{k}w(x)\right|\leq C|x|^{2-2\epsilon-k}(\ln|x|)^{2p_{0}},}\\ {\frac{\left|D^{m+1}w\left(x_{1}\right)-D^{m+1}w\left(x_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\leq C\left|x_{1}\right|^{1-m-2\epsilon-\alpha}(\ln|x|)^{2p_{0}},}\end{array}\right. (4.6)

for all |x|>2,k=0,,m+1|x|>2,k=0,\ldots,m+1 and |x1|>2,x2B|x1|/2(x1)\left|x_{1}\right|>2,x_{2}\in B_{\left|x_{1}\right|/2}\left(x_{1}\right).

Proof.

Acting partial derivative DkD_{k} to both sides of Eq.  (4.4), we have

aij(x)Dij(Dku(x))=Dkf(x),whereaij(x)=DMijF(D2u(x)).a_{ij}(x)D_{ij}(D_{k}u(x))=D_{k}f(x),\quad\text{where}\quad a_{ij}(x)=D_{M_{ij}}F(D^{2}u(x)). (4.7)

By the assumptions on FF and uu, aij(x)a_{ij}(x) are uniformly elliptic coefficients. By the first formula of (4.5), since m3m\geq 3, there exists C>0C>0 such that

|aij(x)DMijF(A)|C|x|ϵ(ln|x|)p0,|Daij(x)|C|x|1ϵ(ln|x|)p0|a_{ij}(x)-D_{M_{ij}}F(A)|\leq C|x|^{-\epsilon}(\ln|x|)^{p_{0}},\quad|Da_{ij}(x)|\leq C|x|^{-1-\epsilon}(\ln|x|)^{p_{0}}

for all |x|>2|x|>2. For the given α(0,1)\alpha\in(0,1), together with the second formula in (4.5) we have

|Daij(x1)Daij(x2)||x1x2|αC|x1|1ϵα(ln|x1|)p0,|x1|>2,x2B|x1|/2(x1).\frac{\left|Da_{ij}\left(x_{1}\right)-Da_{ij}\left(x_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\leq C\left|x_{1}\right|^{-1-\epsilon-\alpha}(\ln|x_{1}|)^{p_{0}},\quad\left|x_{1}\right|>2,\quad x_{2}\in B_{\left|x_{1}\right|/2}\left(x_{1}\right).

For any l=1,2l=1,2, we act partial derivative DlD_{l} to both sides of (4.7). Let h1:=Dkluh_{1}:=D_{kl}u, then we have

DMij,MqrF(D2u)DijkuDqrlu+DMijF(D2u)Dijh1=Dklf(x),D_{M_{ij},M_{qr}}F(D^{2}u)D_{ijk}uD_{qrl}u+D_{M_{ij}}F(D^{2}u)D_{ij}h_{1}=D_{kl}f(x),

i.e.,

aij()Dijh1=f2(x)in 2,a_{ij}(\infty)D_{ij}h_{1}=f_{2}(x)\quad\text{in }\mathbb{R}^{2},

where

f2(x):=DklfDMij,MqrF(D2u)DijkuDqrlu+(aij()aij(x))Dijh1.f_{2}(x):=D_{kl}f-D_{M_{ij},M_{qr}}F(D^{2}u)D_{ijk}uD_{qrl}u+\left(a_{ij}(\infty)-a_{ij}(x)\right)D_{ij}h_{1}.

Notice that [aij()]=[DMijF(A)][a_{ij}(\infty)]=[D_{M_{ij}}F(A)] is a positive symmetric matrix, we set Q:=[aij()]12Q:=[a_{ij}(\infty)]^{\frac{1}{2}} and h~1(x):=h1(Qx)\widetilde{h}_{1}(x):=h_{1}(Qx). Since trace is invariant under cyclic permutations,

Δh~1(x)=f2(Qx)=:f~2(x)\Delta\widetilde{h}_{1}(x)=f_{2}(Qx)=:\widetilde{f}_{2}(x) (4.8)

in Q1(R2B1¯)Q^{-1}(R^{2}\setminus\overline{B_{1}}). Since QQ is invertible, by ζ>2>2ϵ\zeta>2>2\epsilon and (4.5), we have

|f~2(x)||Dklf(Qx)|+C|Dijku(Qx)||Dqklu(Qx)|+C|aij()aij(Qx)|C|x|2ζ+C|x|22ϵ(ln|x|)2p0C|x|22ϵ(ln|x|)2p0\begin{array}[]{lllll}|\widetilde{f}_{2}(x)|&\leq&|D_{kl}f(Qx)|+C|D_{ijk}u(Qx)|\cdot|D_{qkl}u(Qx)|+C|a_{ij}(\infty)-a_{ij}(Qx)|\\ &\leq&C|x|^{-2-\zeta}+C|x|^{-2-2\epsilon}(\ln|x|)^{2p_{0}}\\ &\leq&C|x|^{-2-2\epsilon}(\ln|x|)^{2p_{0}}\\ \end{array}

in xQ1(2B1¯)x\in Q^{-1}(\mathbb{R}^{2}\setminus\overline{B_{1}}). By Lemmas 2.1 and 2.3 and 0<2ϵ<10<2\epsilon<1, there exists a function h~2\widetilde{h}_{2} satisfying (4.8) on an exterior domain with estimate

|Dkh~2(x)|C|x|2ϵk(ln|x|)2p0,k=0,1.|D^{k}\widetilde{h}_{2}(x)|\leq C|x|^{-2\epsilon-k}(\ln|x|)^{2p_{0}},\quad\forall~{}k=0,1.

By the definition of h~1\tilde{h}_{1} and the first line of (4.5), h~1h~2\widetilde{h}_{1}-\widetilde{h}_{2} is harmonic on exterior domain of 2\mathbb{R}^{2} with vanishing speed

h~1h~2δkl=O(|x|ϵ(ln|x|)p0)as |x|.\widetilde{h}_{1}-\widetilde{h}_{2}-\delta_{kl}=O(|x|^{-\epsilon}(\ln|x|)^{p_{0}})\quad\text{as }|x|\rightarrow\infty.

By spherical harmonic expansion as in (2.4), see also formula (2.23) in [2], there exists C>0C>0 such that

|h~1h~2δkl|C|x|1|\widetilde{h}_{1}-\widetilde{h}_{2}-\delta_{kl}|\leq C|x|^{-1}

for all |x|>2|x|>2. By taking h2(x):=h~2(Q1x)h_{2}(x):=\widetilde{h}_{2}(Q^{-1}x), there exists C>0C>0 such that

|h1(x)h2(x)Akl|C|x|1,|x|>2\left|h_{1}(x)-h_{2}(x)-A_{kl}\right|\leq C|x|^{-1},\quad\forall~{}|x|>2

and hence

|Dkw(x)|C|x|22ϵk(ln|x|)2p0,|x|>2.|D^{k}w(x)|\leq C|x|^{2-2\epsilon-k}(\ln|x|)^{2p_{0}},\quad\forall~{}|x|>2.

By taking higher order derivatives as in the proof of Lemma 4.1, this finishes the proof of the first formula in (4.6).

It remains to prove the Hölder semi-norm part. In fact for sufficiently large |x||x|, we set R:=|x|R:=|x| and

h2,R(y):=h~2(x+R4y),f2,R(y)=R216f~2(x+R4y),|y|2.h_{2,R}(y):=\widetilde{h}_{2}\left(x+\frac{R}{4}y\right),\quad f_{2,R}(y)=\frac{R^{2}}{16}\widetilde{f}_{2}\left(x+\frac{R}{4}y\right),\quad|y|\leq 2.

By condition (1.4) on f(x)f(x) and (4.5), for any large |x1||x_{1}| and x2B|x1|/2(x1)x_{2}\in B_{|x_{1}|/2}(x_{1}) with x1x2x_{1}\not=x_{2}, we have

|f~2(x1)f2~(x2)||x1x2|α|D~klf(Qx1)Dklf(Qx2)||x1x2|α+|(aij()aij(Qx1))Dijh1(Qx1)(aij()aij(Qx2))Dijh1(Qx2)||x1x2|α+|FMij,Mqr(D2u)DijkuDqrlu(Qx1)FMij,Mqr(D2u)DijkuDqrlu(Qx2)||x1x2|αC|x1|ζ2α+C|x1|1ϵ(ln|x1|)p0|x1|1ϵα(ln|x1|)p0C|x1|22ϵα(ln|x1|)2p0.\begin{array}[]{lllll}&\dfrac{\left|\widetilde{f}_{2}\left(x_{1}\right)-\widetilde{f_{2}}\left(x_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\\ \leq&\dfrac{\left|\widetilde{D}_{kl}f\left(Qx_{1}\right)-D_{kl}f\left(Qx_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\\ &+\dfrac{\left|\left(a_{ij}(\infty)-a_{ij}(Qx_{1})\right)D_{ij}h_{1}\left(Qx_{1}\right)-\left(a_{ij}(\infty)-a_{ij}(Qx_{2})\right)D_{ij}h_{1}\left(Qx_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\\ &+\dfrac{\left|F_{M_{ij},M_{qr}}\left(D^{2}u\right)D_{ijk}uD_{qrl}u\left(Qx_{1}\right)-F_{M_{ij},M_{qr}}\left(D^{2}u\right)D_{ijk}uD_{qrl}u\left(Qx_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\\ \leq&C|x_{1}|^{-\zeta-2-\alpha}+C|x_{1}|^{-1-\epsilon}(\ln|x_{1}|)^{p_{0}}\cdot|x_{1}|^{-1-\epsilon-\alpha}(\ln|x_{1}|)^{p_{0}}\\ \leq&C|x_{1}|^{-2-2\epsilon-\alpha}(\ln|x_{1}|)^{2p_{0}}.\\ \end{array}

Thus by a direct computation, there exists C>0C>0 uniform to all R>1R>1 such that

f2,RCα(B1¯)CR2ϵ(ln|x|)2p0.||f_{2,R}||_{C^{\alpha}(\overline{B_{1}})}\leq CR^{-2\epsilon}(\ln|x|)^{2p_{0}}.

By the interior Schauder estimates of Poisson equation, we have

h2,RC2,α(B1)C(h2,RL(B2)+f2,RCα(B2))CR2ϵ(lnR)2p0\left\|h_{2,R}\right\|_{C^{2,\alpha}\left(B_{1}\right)}\leq C\left(\left\|h_{2,R}\right\|_{L^{\infty}\left(B_{2}\right)}+\left\|f_{2,R}\right\|_{C^{\alpha}\left(B_{2}\right)}\right)\leq CR^{-2\epsilon}(\ln R)^{2p_{0}}

By the definition of h2,Rh_{2,R} and the non-degeneracy of QQ we have the second formula in (4.6).

For higher order derivatives, the results follow from taking further derivatives to both sides of the equation and apply interior Schauder estimates as in the proof of Lemma 4.1. ∎

Now we are able to prove the asymptotic behavior at infinity by iteration. More explicitly, we have the following results.

Proposition 4.3.

Let u,fu,f be as in Theorem 1.2, then there exists A𝚂𝚢𝚖(2)A\in\mathtt{Sym}(2) satisfying Fτ(λ(A))=f()F_{\tau}(\lambda(A))=f(\infty) and (1.5), b2b\in\mathbb{R}^{2} and d,c,R1d,c,R_{1}\in\mathbb{R} such that (1.8) holds as |x||x|\rightarrow\infty, where QQ is given by (1.9). Furthermore, when ζ>3\zeta>3, then there exist d1,d2d_{1},d_{2} such that (1.10) holds.

Proof.

By Lemmas 3.1 and 4.1, there exist α,ϵ>0\alpha,\epsilon^{\prime}>0 such that (4.1) holds. Let p1p_{1}\in\mathbb{N} be the positive integer such that

2p1ϵ<1and1<2p1+1ϵ<2.2^{p_{1}}\epsilon^{\prime}<1\quad\text{and}\quad 1<2^{p_{1}+1}\epsilon^{\prime}<2.

(If necessary, we may choose ϵ\epsilon^{\prime} smaller to make both inequalities hold.) Let ϵ1:=2p1ϵ\epsilon_{1}:=2^{p_{1}}\epsilon^{\prime}. Applying Lemma 4.2 p1p_{1} times, we have

{|Dkw(x)|C|x|2ϵ1k,|Dm+1w(x1)Dm+1w(x2)||x1x2|αC|x1|1mϵ1α,\left\{\begin{array}[]{ l }{\left|D^{k}w(x)\right|\leq C|x|^{2-\epsilon_{1}-k},}\\ {\frac{\left|D^{m+1}w\left(x_{1}\right)-D^{m+1}w\left(x_{2}\right)\right|}{\left|x_{1}-x_{2}\right|^{\alpha}}\leq C\left|x_{1}\right|^{1-m-\epsilon_{1}-\alpha},}\end{array}\right. (4.9)

for all |x|>2,k=0,,m+1|x|>2,k=0,\ldots,m+1 and |x1|>2,x2B|x1|/2(x1)\left|x_{1}\right|>2,x_{2}\in B_{\left|x_{1}\right|/2}\left(x_{1}\right).

Now we consider the linearized equation again. Applying Newton–Leibnitz formula between Eq. (1.3) and Fτ(λ(A))=f()F_{\tau}(\lambda(A))=f(\infty), we have a~ijDijw=f(x)f()\widetilde{a}_{ij}D_{ij}w=f(x)-f(\infty) i.e.,

a~ij()Dijw=f(x)f()+(a~ij()a~ij(x))Dijw=:f3(x)\widetilde{a}_{ij}(\infty)D_{ij}w=f(x)-f(\infty)+(\widetilde{a}_{ij}(\infty)-\widetilde{a}_{ij}(x))D_{ij}w=:f_{3}(x)

in 2\mathbb{R}^{2}, where the coefficients are uniformly elliptic and

a~ij(x)=01DMijFτ(A+tD2w(x))𝑑ta~ij()=DMijFτ(A)\widetilde{a}_{ij}(x)=\int_{0}^{1}D_{M_{ij}}F_{\tau}(A+tD^{2}w(x))dt\rightarrow\widetilde{a}_{ij}(\infty)=D_{M_{ij}}F_{\tau}(A)

as |x||x|\rightarrow\infty. Let Q:=[DMijF(A)]12Q:=[D_{M_{ij}}F(A)]^{\frac{1}{2}} and w~(x):=w(Qx)\widetilde{w}(x):=w(Qx). By the invariance of trace under cyclic permutations again, we have

Δw~=f3(Qx)=:f~3(x).\Delta\widetilde{w}=f_{3}(Qx)=:\widetilde{f}_{3}(x). (4.10)

By the definition of a~ij(x)\widetilde{a}_{ij}(x), condition (1.4) on ff and (4.9) we have

|f~3(x)|C|x|2ϵ1|\widetilde{f}_{3}(x)|\leq C|x|^{-2\epsilon_{1}}

for some C,R1>0C,R_{1}>0 for all |x|>2R1|x|>2R_{1}. By Lemmas 2.1 and 2.3, there exist a function h~3\widetilde{h}_{3} solving (4.10) in 2BR1¯\mathbb{R}^{2}\setminus\overline{B_{R_{1}}} with estimate

|h~3(x)|C|x|22ϵ1|\widetilde{h}_{3}(x)|\leq C|x|^{2-2\epsilon_{1}}

for some C>0C>0 for all |x|>2R1|x|>2R_{1}. Thus w~h~3\widetilde{w}-\widetilde{h}_{3} is harmonic on an exterior domain of 2\mathbb{R}^{2} with w~h~3=O(|x|2ϵ1)\widetilde{w}-\widetilde{h}_{3}=O(|x|^{2-\epsilon_{1}}) as |x||x|\rightarrow\infty. By spherical harmonic expansion as in (2.4) or the proof of (2.31) in [2], there exist b~2\widetilde{b}\in\mathbb{R}^{2} and d~1,d~2\widetilde{d}_{1},\widetilde{d}_{2}\in\mathbb{R} such that

w~(x)h~3(x)=b~x+d~1ln|x|+d~2+O(|x|1)\widetilde{w}(x)-\widetilde{h}_{3}(x)=\widetilde{b}\cdot x+\widetilde{d}_{1}\ln|x|+\widetilde{d}_{2}+O(|x|^{-1})

as |x||x|\rightarrow\infty and consequently

|w~(x)b~x||h~3(x)|+|d~1ln|x|+d~2+O(|x|1)|=O(|x|22ϵ1)+|d~1ln|x|+d~2+O(|x|1)|=O(|x|22ϵ1)=o(|x|)\begin{array}[]{lllll}|\widetilde{w}(x)-\widetilde{b}\cdot x|&\leq&|\widetilde{h}_{3}(x)|+|\widetilde{d}_{1}\ln|x|+\widetilde{d}_{2}+O(|x|^{-1})|\\ &=&O(|x|^{2-2\epsilon_{1}})+|\widetilde{d}_{1}\ln|x|+\widetilde{d}_{2}+O(|x|^{-1})|\\ &=&O(|x|^{2-2\epsilon_{1}})=o(|x|)\\ \end{array}

as |x||x|\rightarrow\infty.

Let

w~1(x):=w~(x)b~x.\widetilde{w}_{1}(x):=\widetilde{w}(x)-\widetilde{b}\cdot x.

By interior estimates as used in Lemma 3.1, we have

|Dkw~1(x)|C|x|22ϵ1k,|D^{k}\widetilde{w}_{1}(x)|\leq C|x|^{2-2\epsilon_{1}-k},

for some C>0C>0 for all k=0,,m+1k=0,\cdots,m+1 and |x|>2R1|x|>2R_{1}. As in the process in obtaining (4.10), w~1\widetilde{w}_{1} satisfies

Δw~1=f~4(x)=O(|x|ζ)+O(|x|4ϵ1).\Delta\widetilde{w}_{1}=\widetilde{f}_{4}(x)=O(|x|^{-\zeta})+O(|x|^{-4\epsilon_{1}}). (4.11)

Since ζ>2\zeta>2 and 4ϵ1(2,4)4\epsilon_{1}\in(2,4), Lemmas 2.1 and 2.3, there exists a function h~4\widetilde{h}_{4} solving (4.11) in 2BR1¯\mathbb{R}^{2}\setminus\overline{B_{R_{1}}} with estimate

|h~4(x)|{C|x|2ζ(ln|x|)+C|x|24ϵ1,ζ,C|x|2ζ+C|x|24ϵ1,ζ,|\widetilde{h}_{4}(x)|\leq\left\{\begin{array}[]{llll}C|x|^{2-\zeta}(\ln|x|)+C|x|^{2-4\epsilon_{1}},&\zeta\not\in\mathbb{N}_{*},\\ C|x|^{2-\zeta}+C|x|^{2-4\epsilon_{1}},&\zeta\in\mathbb{N}_{*},\\ \end{array}\right.

for some C>0C>0 for all |x|>2R1|x|>2R_{1}. Thus w~1h~4\widetilde{w}_{1}-\widetilde{h}_{4} is harmonic in |x|>2R1|x|>2R_{1} with |w~1h~4|=O(|x|22ϵ1)|\widetilde{w}_{1}-\widetilde{h}_{4}|=O(|x|^{2-2\epsilon_{1}}). Since 22ϵ1<12-2\epsilon_{1}<1, by spherical harmonic expansion, there exist d~,d~3\widetilde{d},\widetilde{d}_{3}\in\mathbb{R} such that

w~1(x)h~4(x)=d~ln|x|+d~3+O(|x|1)\widetilde{w}_{1}(x)-\widetilde{h}_{4}(x)=\widetilde{d}\ln|x|+\widetilde{d}_{3}+O(|x|^{-1})

as |x||x|\rightarrow\infty and consequently,

|w~1(x)d~ln|x|||h~4(x)|+d~3+O(|x|1)={O(|x|2ζ),if ζ<4ϵ1,and ζ3O(|x|2ζ(ln|x|)),if ζ<4ϵ1,and ζ=3O(|x|24ϵ1),if ζ4ϵ1.\begin{array}[]{llllll}|\widetilde{w}_{1}(x)-\widetilde{d}\ln|x||&\leq&|\widetilde{h}_{4}(x)|+\widetilde{d}_{3}+O(|x|^{-1})\\ &=&\left\{\begin{array}[]{lll}O(|x|^{2-\zeta}),&\text{if }\zeta<4\epsilon_{1},&\text{and }\zeta\not=3\\ O(|x|^{2-\zeta}(\ln|x|)),&\text{if }\zeta<4\epsilon_{1},&\text{and }\zeta=3\\ O(|x|^{2-4\epsilon_{1}}),&\text{if }\zeta\geq 4\epsilon_{1}.\\ \end{array}\right.\\ \end{array} (4.12)

Again, we follow the process in obtaining (4.10).

Since

Fτ(λ(D2(12xTAx+dln|x|)))=f()+O(|x|4)F_{\tau}(\lambda(D^{2}(\frac{1}{2}x^{T}Ax+d\ln|x|)))=f(\infty)+O(|x|^{-4})

as |x||x|\rightarrow\infty, we set

w~2(x):=w~1(x)d~ln|x|,\widetilde{w}_{2}(x):=\widetilde{w}_{1}(x)-\widetilde{d}\ln|x|,

which satisfies

Δw~2=O(|x|ζ)+O(|x|4)+{O(|x|2ζ),if ζ<4ϵ1,and ζ3O(|x|2ζ(ln|x|)2),if ζ<4ϵ1,and ζ=3O(|x|8ϵ1),if ζ4ϵ1.\Delta\widetilde{w}_{2}=O(|x|^{-\zeta})+O(|x|^{-4})+\left\{\begin{array}[]{lll}O(|x|^{-2\zeta}),&\text{if }\zeta<4\epsilon_{1},&\text{and }\zeta\not=3\\ O(|x|^{-2\zeta}(\ln|x|)^{2}),&\text{if }\zeta<4\epsilon_{1},&\text{and }\zeta=3\\ O(|x|^{-8\epsilon_{1}}),&\text{if }\zeta\geq 4\epsilon_{1}.\\ \end{array}\right.

Since 8ϵ1(4,8)8\epsilon_{1}\in(4,8), we have

Δw~2=O(|x|ζ)+O(|x|4).\Delta\widetilde{w}_{2}=O(|x|^{-\zeta})+O(|x|^{-4}).

By Lemmas 2.1 and 2.3, we have a solution h~5\widetilde{h}_{5} on exterior domain with estimate

|h~5(x)|{C|x|2ζ+C|x|2(ln|x|),if ζ,C|x|2ζ(ln|x|)+C|x|2(ln|x|),if ζ,|\widetilde{h}_{5}(x)|\leq\left\{\begin{array}[]{llll}C|x|^{2-\zeta}+C|x|^{-2}(\ln|x|),&\text{if }\zeta\not\in\mathbb{N}_{*},\\ C|x|^{2-\zeta}(\ln|x|)+C|x|^{-2}(\ln|x|),&\text{if }\zeta\in\mathbb{N}_{*},\\ \end{array}\right.

for some C>0C>0 for all |x|>2R1|x|>2R_{1}. Together with (4.12), by spherical harmonic expansion we have c~\widetilde{c}\in\mathbb{R} such that

w~2(x)h~5(x)=c~+O(|x|1).\widetilde{w}_{2}(x)-\widetilde{h}_{5}(x)=\widetilde{c}+O(|x|^{-1}). (4.13)

Rotating back by Q1Q^{-1} matrix, since P=Q2P=Q^{-2}, we have β2,c,d\beta\in\mathbb{R}^{2},c,d\in\mathbb{R} such that

|u(x)(12xTAx+βx+dln(xTPx)+c)|C|w~2c~|C|h~5(Q1x)|+C|x|1{C|x|2ζ+C|x|1,if ζ3,C|x|2ζ(ln|x|)+C|x|1,if ζ=3,\begin{array}[]{llll}&\displaystyle\left|u(x)-\left(\frac{1}{2}x^{T}Ax+\beta x+d\ln(x^{T}Px)+c\right)\right|\\ \leq&C|\widetilde{w}_{2}-\widetilde{c}|\\ \leq&C|\widetilde{h}_{5}(Q^{-1}x)|+C|x|^{-1}\\ \leq&\left\{\begin{array}[]{llll}C|x|^{2-\zeta}+C|x|^{-1},&\text{if }\zeta\not=3,\\ C|x|^{2-\zeta}(\ln|x|)+C|x|^{-1},&\text{if }\zeta=3,\\ \end{array}\right.\\ \end{array}

for some C>0C>0 for sufficiently large |x||x|. Estimates for higher order derivatives follow similarly as in Lemma 4.1. The second equality in (1.9) can be obtained in (1.4) of [32]. This finishes the proof of (1.8).

It remains to prove that when ζ>3\zeta>3, we have (1.10) at infinity. In fact from (4.13), we iterate once more by setting w~3:=w~2(x)c~\widetilde{w}_{3}:=\widetilde{w}_{2}(x)-\widetilde{c}, which satisfies

Δw~3=O(|x|ζ)+O(|x|4)as |x|.\Delta\widetilde{w}_{3}=O(|x|^{-\zeta})+O(|x|^{-4})\quad\text{as }|x|\rightarrow\infty.

By Lemmas 2.1 and 2.3, we have a solution h~6\widetilde{h}_{6} on exterior domain with estimate

|h~6(x)|{C|x|2ζ+C|x|2(ln|x|),if ζ,C|x|2ζ(ln|x|)+C|x|2(ln|x|),if ζ,|\widetilde{h}_{6}(x)|\leq\left\{\begin{array}[]{lllll}C|x|^{2-\zeta}+C|x|^{-2}(\ln|x|),&\text{if }\zeta\not\in\mathbb{N}_{*},\\ C|x|^{2-\zeta}(\ln|x|)+C|x|^{-2}(\ln|x|),&\text{if }\zeta\in\mathbb{N}_{*},\\ \end{array}\right.

for some C>0C>0 in |x|>2R1|x|>2R_{1}. Since w~3h~6\widetilde{w}_{3}-\widetilde{h}_{6} is harmonic on an exterior domain of 2\mathbb{R}^{2} and satisfies |w~3h~6|=O(|x|1)|\widetilde{w}_{3}-\widetilde{h}_{6}|=O(|x|^{-1}) as |x||x|\rightarrow\infty, by spherical harmonic expansion there exists d~4,d~5\widetilde{d}_{4},\widetilde{d}_{5}\in\mathbb{R} such that

w~3h~6=d~4cosθ|x|1+d~5sinθ|x|1+O(|x|2)\widetilde{w}_{3}-\widetilde{h}_{6}=\widetilde{d}_{4}\cos\theta|x|^{-1}+\widetilde{d}_{5}\sin\theta|x|^{-1}+O(|x|^{-2})

as |x||x|\rightarrow\infty, where θ=x|x|\theta=\frac{x}{|x|} here. By rotating back through Q1Q^{-1}, we have the 0-order estimates in (1.10) since P=Q2P=Q^{-2}. For higher order derivatives, the result follows from interior estimate as in [33]. ∎

References

  • [1] Jiguang Bao, Jingyi Chen, Bo Guan, and Min Ji. Liouville property and regularity of a Hessian quotient equation. Amer. J. Math., 125(2):301–316, 2003.
  • [2] Jiguang Bao, Haigang Li, and Lei Zhang. Monge-Ampère equation on exterior domains. Calc. Var. Partial Differential Equations, 52(1-2):39–63, 2015.
  • [3] Arunima Bhattacharya. Hessian estimates for Lagrangian mean curvature equation. Calc. Var. Partial Differential Equations, 60(6):Paper No. 224, 2021.
  • [4] Arunima Bhattacharya and Ravi Shankar. Optimal regularity for Lagrangian mean curvature type equations. arXiv. 2009.04613, 2020.
  • [5] Arunima Bhattacharya and Ravi Shankar. Regularity for convex viscosity solutions of Lagrangian mean curvature equation. arXiv. 2006.02030, 2020.
  • [6] Luis Caffarelli. Interior W2,pW^{2,p} estimates for solutions of the Monge-Ampère equation. Ann. of Math. (2), 131(1):135–150, 1990.
  • [7] Luis Caffarelli and Xavier Cabré. Fully nonlinear elliptic equations, volume 43 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 1995.
  • [8] Luis Caffarelli, Basilis Gidas, and Joel Spruck. Asymptotic symmetry and local behavior of semilinear elliptic equations with critical Sobolev growth. Comm. Pure Appl. Math., 42(3):271–297, 1989.
  • [9] Luis Caffarelli and Yanyan Li. An extension to a theorem of Jörgens, Calabi, and Pogorelov. Comm. Pure Appl. Math., 56(5):549–583, 2003.
  • [10] Luis Caffarelli and Yanyan Li. A Liouville theorem for solutions of the Monge-Ampère equation with periodic data. Ann. Inst. H. Poincaré Anal. Non Linéaire, 21(1):97–120, 2004.
  • [11] Eugenio Calabi. Improper affine hyperspheres of convex type and a generalization of a theorem by K. Jörgens. Michigan Math. J., 5:105–126, 1958.
  • [12] Sun-Yung Alice Chang and Yu Yuan. A Liouville problem for the sigma-2 equation. Discrete Contin. Dyn. Syst., 28(2):659–664, 2010.
  • [13] Jingyi Chen, Ravi Shankar, and Yu Yuan. Regularity for convex viscosity solutions of special Lagrangian equation. arXiv. 1911.05452, 2019.
  • [14] Li Chen and Ni Xiang. Rigidity theorems for the entire solutions of 2-Hessian equation. J. Differential Equations, 267(9):5202–5219, 2019.
  • [15] Shiu Yuen Cheng and Shing-Tung Yau. Complete affine hypersurfaces. I. The completeness of affine metrics. Comm. Pure Appl. Math., 39(6):839–866, 1986.
  • [16] Shi-Zhong Du. Necessary and sufficient conditions to Bernstein theorem of a Hessian equation. arXiv. 2106.06211, 2021.
  • [17] Harley Flanders. On certain functions with positive definite Hessian. Ann. of Math. (2), 71:153–156, 1960.
  • [18] Lei Fu. An analogue of Bernstein’s theorem. Houston J. Math., 24(3):415–419, 1998.
  • [19] David Gilbarg and Neil S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [20] Matthias Günther. Conformal normal coordinates. Ann. Global Anal. Geom., 11(2):173–184, 1993.
  • [21] Qing Han, Xiaoxiao Li, and Yichao Li. Asymptotic expansions of solutions of the Yamabe equation and the σk\sigma_{k}-Yamabe equation near isolated singular points. Comm. Pure Appl. Math., 74(9):1915–1970, 2021.
  • [22] Zheng-Chao Han, Yanyan Li, and Eduardo V. Teixeira. Asymptotic behavior of solutions to the σk\sigma_{k}-Yamabe equation near isolated singularities. Invent. Math., 182(3):635–684, 2010.
  • [23] Guanghao Hong. A Remark on Monge-Ampère equation over exterior domains. arXiv. 2007.12479, 2020.
  • [24] Rongli Huang and Zhizhang Wang. On the entire self-shrinking solutions to Lagrangian mean curvature flow. Calc. Var. Partial Differential Equations, 41(3-4):321–339, 2011.
  • [25] Xiaobiao Jia. Asymptotic behavior of solutions of fully nonlinear equations over exterior domains. C. R. Math. Acad. Sci. Paris, 358(11-12):1187–1197, 2020.
  • [26] Konrad Jörgens. Über die Lösungen der Differentialgleichung rts2=1rt-s^{2}=1. Math. Ann., 127:130–134, 1954.
  • [27] Jürgen Jost and Yuan Long Xin. Some aspects of the global geometry of entire space-like submanifolds. volume 40, pages 233–245. 2001. Dedicated to Shiing-Shen Chern on his 90th birthday.
  • [28] Nick Korevaar, Rafe Mazzeo, Frank Pacard, and Richard Schoen. Refined asymptotics for constant scalar curvature metrics with isolated singularities. Invent. Math., 135(2):233–272, 1999.
  • [29] An-Min Li, Ruiwei Xu, Udo Simon, and Fang Jia. Affine Bernstein problems and Monge-Ampère equations. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2010.
  • [30] Dongsheng Li, Zhisu Li, and Yu Yuan. A Bernstein problem for special Lagrangian equations in exterior domains. Adv. Math., 361:106927, 29, 2020.
  • [31] Ming Li, Changyu Ren, and Zhizhang Wang. An interior estimate for convex solutions and a rigidity theorem. J. Funct. Anal., 270(7):2691–2714, 2016.
  • [32] Zixiao Liu and Jiguang Bao. Asymptotic expansion and optimal symmetry of minimal gradient graph equations in dimension 2. Communications in Contemporary Mathematics, 0(0):2150110, 0.
  • [33] Zixiao Liu and Jiguang Bao. Asymptotic expansion at infinity of solutions of Monge-Ampère type equations. Nonlinear Anal., 212:Paper No. 112450, 17, 2021.
  • [34] Zixiao Liu and Jiguang Bao. Asymptotic Expansion at Infinity of Solutions of Special Lagrangian Equations. J. Geom. Anal., 32(3):90, 2022.
  • [35] Aleksei Vasil’evich Pogorelov. On the improper convex affine hyperspheres. Geometriae Dedicata, 1(1):33–46, 1972.
  • [36] Eduardo V. Teixeira and Lei Zhang. Global Monge-Ampére equation with asymptotically periodic data. Indiana Univ. Math. J., 65(2):399–422, 2016.
  • [37] Chong Wang, Rongli Huang, and Jiguang Bao. On the second boundary value problem for Lagrangian mean curvature equation. arXiv:1808.01139, 2018.
  • [38] Dake Wang and Yu Yuan. Hessian estimates for special Lagrangian equations with critical and supercritical phases in general dimensions. Amer. J. Math., 136(2):481–499, 2014.
  • [39] Micah Warren. Calibrations associated to Monge-Ampère equations. Trans. Amer. Math. Soc., 362(8):3947–3962, 2010.
  • [40] Micah Warren and Yu Yuan. Hessian estimates for the sigma-2 equation in dimension 3. Communications on Pure and Applied Mathematics, 62(3):305–321, 2009.
  • [41] Micah Warren and Yu Yuan. Hessian and gradient estimates for three dimensional special Lagrangian equations with large phase. Amer. J. Math., 132(3):751–770, 2010.
  • [42] Min Yan. Extension of convex function. J. Convex Anal., 21(4):965–987, 2014.
  • [43] Yu Yuan. A Bernstein problem for special Lagrangian equations. Invent. Math., 150(1):117–125, 2002.
  • [44] Yu Yuan. Global solutions to special Lagrangian equations. Proc. Amer. Math. Soc., 134(5):1355–1358, 2006.

Z. Liu & J. Bao

School of Mathematical Sciences, Beijing Normal University
Laboratory of Mathematics and Complex Systems, Ministry of Education
Beijing 100875, China
Email: liuzixiao@mail.bnu.edu.cn, jgbao@bnu.edu.cn