<<

Symmetry of solutions of minimal gradient graph equations on punctured space

Zixiao Liu, Jiguang Bao*

April 23, 2021

Abstract In this paper, we study symmetry and existence of solutions of minimal gradient graph equations on punctured space Rn \{0}, which include the Monge-Ampere` equation, inverse harmonic Hessian equation and the special Lagrangian equation. This extends the classifica- tion results of Monge-Ampere` equations. Under some conditions, we also give the character- ization of the solvability on exterior Dirichlet problem in terms of their asymptotic behaviors. Keywords: minimal gradient graph equation, optimal symmetry, existence. MSC 2020: 35J60; 35B06.

1 Introduction

We consider the following fully nonlinear elliptic equations

2 n Fτ (λ(D u)) = C0 in R \{0}, (1.1)

2 2 where C0 is a constant, λ(D u) = (λ1, λ2, ··· , λn) are n eigenvalues of Hessian matrix D u, π τ ∈ [0, 2 ] and

 1 n  X  ln λi, τ = 0,  n  √ i=1  a2 + 1 n λ + a − b  X i π  ln , 0 < τ < 4 ,  2b λi + a + b arXiv:2104.10904v1 [math.AP] 22 Apr 2021  i=1  √ n  X 1 π Fτ (λ) := − 2 , τ = 4 , 1 + λi  √ i=1  2 n  a + 1 X λi + a − b  arctan , π < τ < π ,  b λ + a + b 4 2  i=1 i  n  X  arctan λ , τ = π ,  i 2 i=1

*Supported in part by National Natural Science Foundation of China (11871102 and 11631002).

1 a = cot τ, b = p|cot2 τ − 1|. Equations (1.1) origin from gradient graph (x, Du(x)) with zero n n mean curvature under various metrics equipped on R × R . In 2010, Warren [32] first proved 2 2 n that if u ∈ C (Ω) is a solution of Fτ (λ(D u)) = C0 in Ω ⊂ R , then the volume of (x, Du(x)) π π π 1 is a maximal for τ ∈ [0, 4 ) and minimal for τ ∈ ( 4 , 2 ] among all homologous, C , space-like n n n-surfaces in (R × R , gτ ), where h π i g = sin τδ + cos τg , τ ∈ 0, , τ 0 0 2 is the linearly combined metric of standard Euclidean metric

n n X X δ0 = dxi ⊗ dxi + dyj ⊗ dyj, i=1 j=1 and the pseudo-Euclidean metric

n n X X g0 = dxi ⊗ dyi + dyj ⊗ dxj. i=1 j=1

If τ = 0, then (1.1) becomes the Monge-Ampere` equation

det D2u = enC0 .

n For the Monge-Ampere` equation on entire R , the classical theorem by Jorgens¨ [22], Calabi [11] 2 n and Pogorelov [30] states that any convex classical solution of det D u = 1 in R must be a quadratic polynomial. See Cheng-Yau [13], Caffarelli [5] and Jost-Xin [23] for different proofs n and extensions. For the Monge-Ampere` equation in exterior domain of R , there are exterior Jorgens-Calabi-Pogorelov¨ type results by Ferrer-Mart´ınez-Milan´ [15] for n = 2 and Caffarelli-Li [7], which state that any locally convex solution must be asymptotic to quadratic polynomials (for n = 2 we need additional ln-term) near infinity. For the Monge-Ampere` equation on punctured n space R \{0}, there are classification results by Jorgens¨ [22] for n = 2 and Jin-Xiong [21] that shows every locally convex function of det D2u = 1, modulo an affine transform, has to be |x| 1 R n n u(x) = 0 (r + c) dr for some c ≥ 0. For further generalization and refined asymptotics on Jorgens-Calabi-Pogorelov¨ type results we refer to [3, 8, 18, 31] and the references therein. π If τ = 2 , then (1.1) becomes the special Lagrangian equation n X 2  arctan λi D u = C0. (1.2) i=1 n For special Lagrangian equation on entire R , there are Bernstein-type results by Yuan [33, 34], which state that any classical solution of (1.2) and ( −KI, n ≤ 4, n − 2 D2u ≥ or C > π, (1.3) −( √1 + (n))I, n ≥ 5, 0 3 2 must be a quadratic polynomial, where I denote the identity matrix, K is any constant and (n) n is a small dimensional constant. For special Lagrangian equation on exterior domain of R , there

2 is an exterior Bernstein-type result by Li-Li-Yuan [25], which states that any classical solution of (1.2) on exterior domain with (1.3) must be asymptotic to quadratic polynomial (for n = 2 we need additional ln-term) near infinity. Furthermore, Chen-Shankar-Yuan [12] proved that a convex viscosity solution of (1.2) must be smooth. π Pn 1 If τ = , then (1.1) is a translated inverse harmonic Hessian equation 2 = 1, 4 i=1 λi(D u) n which is a special form of Hessian quotient equation. For Hessian quotient equation on entire R , there is a Bernstein-type result by Bao-Chen-Guan-Ji [1]. π For general τ ∈ [0, 2 ], Warren [32] proved the Bernstein-type results under suitable semi- convex conditions by the results of Jorgens¨ [22]-Calabi [11]-Pogorelov [30], Flanders [16] and Yuan [33, 34]. In our earlier work [28], we generalized the results to equation (1.1) on exterior domain and provide finer asymptotic expansions. For further generalization with perturbed right hand side we refer to [29] etc. In this paper, we focus on n ≥ 3 and prove a type of classification results as in Jorgens¨ [22] and Jin-Xiong [21]. T n T Hereinafter, we let x denote the transpose of vector x ∈ R , O denote the transpose of n×n matrix O, Sym(n) denote the set of symmetric n × n matrix, σk(λ) denote the kth elementary form of λ and DFτ (λ(A)) denote the matrix with elements being value of partial derivative of Fτ (λ(M)) w.r.t Mij variable at matrix A. A function u(x) is called generalized symmetric with respect to A ∈ Sym(n) if it relies only on the value of xT Ax. As in [6], we say u ∈ C0(Ω) is n 2 a viscosity subsolution (supersolution) of (1.1) in Ω ⊂ R if for any function ψ ∈ C (Ω) with 2 2 2 2 π π D ψ > 0, D ψ > (−a + b)I, D ψ > −I, D ψ > −(a + b)I for τ = 0, τ ∈ (0, 4 ), τ = 4 , π π τ ∈ ( 4 , 2 ) respectively and point x ∈ Ω satisfying ψ(x) = u(x) and ψ ≥ (≤)u in Ω, we have 2 Fτ (λ(D ψ(x))) ≥ (≤)C0. If u is both viscosity subsolution and supersolution of (1.1), we say it is a viscosity solution. By a direct computation (see for instance [19, 32]), if λi > −a − b, for i = 1, 2, ··· , n, then n n   X λi + a − b X λi + a nπ arctan = arctan − . (1.4) λ + a + b b 4 i=1 i i=1

Hence replacing u(x) by  (a−b) 2 π  u(x) + 2 |x| , 0 < τ < 4 ,  1 2 π u(x) + 2 |x| , τ = 4 , (1.5)  u(x) a 2 π π  b + 2b |x| , 4 < τ < 2 , it is equivalent to consider 2 n Gτ (λ(D u)) = C0 in R \{0}, (1.6)

3 where  n  1 X  ln λi, τ = 0,  n  √ i=1  2 n  a + 1 X λi  ln , 0 < τ < π ,  2b λ + 2b 4  i=1 i  n  √ X 1 G (λ) := − 2 , τ = π , τ λ 4  i=1 i  √ n !  a2 + 1 X nπ  arctan λ − π < τ < π ,  b i 4 4 2  i=1  n  X  arctan λ , τ = π .  i 2  i=1 Our first main result provides a type of symmetry and relationship between asymptotic at infinity and value at origin.

2 n 0 n Theorem 1.1. Let u ∈ C (R \{0}) be a classical convex solution of (1.6). Then u ∈ C (R ),

2 n Gτ (λ(D u)) ≥ C0 in R (1.7)

n in viscosity sense and there exist c ∈ R, β ∈ R and A ∈ Sym(n) with Gτ (λ(A)) = C0 such that 1 u(x) ≤ xT Ax + βx + c in n. (1.8) 2 R Furthermore, u(x) − βx is symmetric in eigenvector directions of A, i.e.,

n T T u(xe) − βxe = u(x) − βx, ∀ {xe ∈ R :(O xe)i = ±(O x)i, ∀ i = 1, 2, ··· , n}, (1.9) where O is an orthogonal matrix such that OAOT is diagonal.

Remark 1.2. After a change of coordinate y := OT x, u(y) − βy is symmetric with respect to n coordinate planes. Especially, u(y) − βy is central symmetry i.e., u(y) − βy = u(−y) + βy.

Remark 1.3. The solutions of (1.1) and (1.6) have asymptotic behavior and higher order expan- sion at infinity, see for instance earlier results by the authors [28, 29].

π π  Remark 1.4. For τ ∈ 4 , 2 case, the convex assumption on u in Theorem 1.1 can be relaxed to either of the following

 D2u > −I, n ≤ 4,  n o D2u > − min 1, √1 + (n) I, n ≥ 5, 3 (1.10)  D2u > −I and √bC0 + nπ > n−2 π, n ≥ 5, a2+1 4 2

π where (n) is as in (1.3). For τ = 2 case, the convex assumption can be relaxed to (1.3). π Our second main result shows that the symmetry (1.9) is optimal for τ ∈ (0, 2 ] cases in the following sense.

4 Proposition 1.5. Let 0 < A ∈ Sym(n) satisfy Gτ (λ(A)) = C0. If there exists a generalized 1 T symmetric convex classical solution u(x) = U( 2 x Ax) of (1.6). Then for τ = 0 case,

1 (xT Ax) 2 Z 1 C0 n u = e (r + c1) n dr + c2 0 π π π π π for some c1 ≥ 0 and c2 ∈ R. For cases τ ∈ (0, 4 ), τ = 4 , τ ∈ ( 4 , 2 ) and τ = 2 , either A is a multiplication of identity matrix i.e., ! √ 2b 2C A = − 2b I,A = − 0 I, 1 − exp( √2b C ) 2n n a2+1 0     b π C0 A = tan √ C0 + I,A = tan I n a2 + 1 4 n 1 T respectively, or u is the quadratic function 2 x Ax up to some constant c.

Remark 1.6. Proposition 1.5 still holds if equation (1.6) is restricted to an annulus BR2 \ BR1 with any 0 < R1 < R2, where Br denotes the ball centered at origin with radius r. By (1.8) in Theorem 1.1, c ≥ u(0). Our final main result shows that c ≥ u(0) is also a sufficient condition for a classical solution exists. More explicitly, we prove the existence of π solution of (1.1), τ ∈ (0, 4 ) with prescribed value at origin and asymptotic behavior at infinity for all c ≥ u(0).

n Theorem 1.7. For any given c ≥ u0, β ∈ R and 0 < A ∈ Sym(n) with Gτ (λ(A)) = C0 satisfying n X min{λ1(A), ··· , λn(A)} + 2b δ := > 2, (1.11) 0 λ (A) + 2b i=1 i 0 n there exists a unique convex viscosity solution u ∈ C (R ) of  G (λ(D2u)) = C , in n \{0},  τ 0 R u(0) = u0, (1.12)  1 T 2−δ0 u(x) = 2 x Ax + βx + c + O(|x| ), as |x| → ∞,

π with τ ∈ (0, 4 ). Remark 1.8. At the end of Section 3, we prove that the solution found in Theorem 1.7 also satisfies (1.7) and (1.8). Condition (1.11) holds if A has two same minimal eigenvalues. For any C0 < 0, condition (1.11) is possible to fail. For instance there exists sequences such that Gτ (λ1, λ2, λ3) = C0 but 2b λ1() → − 2b and λ2(), λ3() → +∞ 1 − exp( √ 2b C ) a2+1 0 as  → 0+. By a direct computation, the left hand side of (1.11) tends to 1 as  → 0+. Whether (1.11) is optimal remains unknown.

5 π Remark 1.9. The study on exterior Dirichlet problem by Li-Li [24] for τ = 4 and Li [27] for π π τ ∈ ( 4 , 2 ] also works for punctured space after minor modification. From their proof, the results in Theorem 1.7 still holds with δ0 changed into

σn−1(λ(A)) · min{λ1(A), ··· , λn(A)} δ0 = σn(λ(A))

π for τ = 4 ,

Pn kc ( nπ + √ b C )σ (λ(A)) k=0 k 4 a2+1 0 k δ0 = Pn ξ ( nπ + √ b C , λ(A))c ( nπ + √ b C )σ (λ(A)) k=0 k 4 a2+1 0 k 4 a2+1 0 k

π π for τ ∈ ( 4 , 2 ) and Pn k=0 kck(C0)σk(λ(A)) δ0 = Pn k=0 ξk(C0, λ(A))ck(C0)σk(λ(A)) π for τ = 2 , where

 j+1 c2j(C) := (−1) sin C if k = 2j, j ∈ N, ck(C) := j c2j+1(C) := (−1) cos C if k = 2j + 1, j ∈ N,

Pn 2 2 i=1 σk−1;i(λ(A))λi (A)xi n Ξk(λ(A), x) := Pn 2 ∀ x ∈ R \{0}, σk(λ(A)) i=1 λi(A)xi ¯ ξk(λ(A)) := sup Ξk(λ(A), x), ξ (λ(A)) := inf Ξk(λ(A), x), k n x∈Rn\{0} x∈R \{0} and  ξ¯ (λ(A)) if c (C) > 0, ξ (C, λ(A)) := k k k ξ (λ(A)) if c (C) ≤ 0. k k For τ = 0, optimal existence of Dirichlet problem on exterior domain and punctured space have been studied in [3, 7, 21, 26] etc.

The paper is organized as follows. In the next three sections, we prove Theorem 1.1, Theorem 1.7 and Proposition 1.5 respectively.

2 Proof of Theorem 1.1

In this section, we prove Theorem 1.1 by maximum principle and removable singularity of viscos- ity solutions. The following knowledge on convex functions is necessary.

Theorem 2.1 (Proposition 2.1 of [21]). Let v be a locally convex function in B1 \{0} with n ≥ 2. Then u can be uniquely extended to be a convex function in B1.

n 0 Theorem 2.2 (Theorem 1.1 of [9]). Let Ω ⊂ R be a domain, x ∈ Ω and u ∈ C (Ω) be a viscosity solution of F (D2u) ≥ f(x) in Ω \{x},

6 where F ∈ C0(Sym(n)) is an elliptic operator and f ∈ C0(Ω). Suppose that u is upperconical at x, i.e., sup ((u + η)(x) − (u + η)(x) + |x − x|) > 0 (2.1) x∈Ω for any η ∈ C∞(Ω) and  > 0. Then

F (D2u) ≥ f(x) in Ω in viscosity sense. Remark 2.3. If u is a convex function in Ω, then for all x ∈ Ω, u is upperconical at x. By contradiction, we suppose there exist  ∈ (0, 1) and η ∈ C∞(Ω) such that

(u + η)(x) − (u + η)(x) + |x − x| ≤ 0 in Ω.

By Taylor expansion of η near x, there exists 0 < δ < 1 such that  u(x) ≤ u(x) − ∇η(x) · (x − x) − |x − x|, ∀ 0 < |x − x| < δ. 2

Since u is convex, take any e ∈ ∂B1 and we have δ δ δ 2u(x) ≤ u(x + e) + u(x − e) ≤ 2u(x) − < 2u(x), 2 2 4 which is a contradiction. For more examples and discussions on condition (2.1), we refer to Remark 1.2 of [9].

2 n Let u ∈ C (R \{0}) be a classical solution of (1.6) as in Theorem 1.1 or Remark 1.4. By n the asymptotic behavior result (see for instance Theorem 1.3 of [28]), there exists c ∈ R, β ∈ R and A ∈ Sym(n) with Gτ (λ(A)) = C0 such that    2+k−n k 1 T lim sup |x| D u(x) − x Ax + βx + c < ∞ (2.2) |x|→+∞ 2 for all k ∈ N. Step 1. we prove (1.7) and (1.8). The proof is separated into five cases according to τ. If u is a classical convex solution as in Theorem 1.1, (1.7) follows immediately from Theorems 2.1 and 2.2. By (2.2), we have 1 u(x) → xT Ax + βx + c =: V (x) as |x| → ∞. 2 2 n By a direct computation, V (x) is a quadratic solution of Gτ (λ(D u)) = C0 on entire R . By maximum principle (see for instance Theorem 17.1 of [17]), 1 u(x) ≤ V (x) = xT Ax + βx + c in n. 2 R π π π For τ ∈ ( 4 , 2 ) case in Remark 1.4, the desired result follows from τ = 2 case, which will be π proved immediately. For τ = 2 case of Remark 1.4, we prove that there exists K0 ≥ 0 such that 2 D u > −K0I.

7 n−2 For the first subcase of (1.3), the result is immediate. For C0 > 2 π, then for all i = 1, 2, ··· , n, n n − 1 X arctan λ (D2u) + π > arctan λ (D2u) = C . i 2 j 0 j=1

n−1 Thus we may take K0 := min{tan(C0 − 2 π), 0}. Let K v(x) := u(x) + 0 |x|2, 2 then 2 2 n D v = D u + K0I > 0 in R \{0} and v satisfies n X 2 n arctan(λi(D v) − K0) = C0 in R \{0}. i=1 By a direct computation, ∂ 1 arctan(λi − K0) = 2 > 0. ∂λi 1 + (λi − K0) Together with asymptotics (2.2), D2v is bounded from above. Hence the differential operator is elliptic (see for instance [10]) and the desired results follow from Theorems 2.1 and 2.2. Step 2. we prove the symmetric property (1.9) for diagonal case i.e., A = diag(a1, a2, ··· , an). We only need to prove for β = 0 case, otherwise replace u by u − βx. Similar to the proof of Theorem 1.1 in [21], we introduce the following maximum principle. Lemma 2.4. Let w be a classical solution of

n aijDijw = 0 in R \{0}, where aij(x) is a positive definite matrix. Suppose w(0) = 0 and lim w(x) = 0, |x|→∞ then w ≡ 0. n Proof. By contradiction, we may suppose there exists x0 ∈ R \{0} such that u(x0) > 0. Then by n continuity and boundary conditions, there exists x1 ∈ R \{0} such that u(x1) = maxRn\{0} u > n 1 0. Applying standard maximum principle in level set {x ∈ R \{0} : u > 2 u(x1)}, we have a 1 contradiction 2 u(x1) = u(x1). For any i = 1, 2, ··· , n, let

U(x) := u(xe), where xe = (x1, ··· , xi−1, −xi, xi+1, ··· , xn). By a direct computation, u, U satisfy

 2 2 n  Gτ (λ(D u)) = Gτ (λ(D U)) = C0, in R \{0}, u(x) = U(x), at x = 0,  1 T  u(x),U(x) → 2 x Ax + c , as |x| → ∞.

8 By Newton-Leibnitz formula between u, U and applying maximum principle as in Lemma 2.4, we have u(x) = U(x) = u(xe). Hence u(x) is hyperplane symmetry with respect to all n coordinate hyperplane. Step 3. we prove (1.9) for general A. Since A is a symmetric matrix, by eigen-decomposition there exists an orthogonal matrix O such that

T A = O ΛO, where Λ = diag(a1, ··· , an). (2.3)

Let v(x) := u(O−1x), then D2v(x) = OD2u(O−1x)O−1 has the same eigenvalues as D2u(O−1x). Hence 2 n Gτ (λ(D v)) = C0 in R \{0} and 1 1 v(0) = u(0), v(x) → xT OAO−1x + c = xT Λx + c 2 2 as |x| → ∞. By the previous step, we have

n v(xe) = v(x) ∀ {xe ∈ R : xei = ±xi, ∀ i = 1, 2, ··· , n}. Thus n T T u(xe) = u(x) ∀ {xe ∈ R :(O xe)i = ±(O x)i, ∀ i = 1, 2, ··· , n} and (1.9) follows immediately.

3 Proof of Theorem 1.7

π In this section, we consider the existence result of Dirichlet problem for τ ∈ (0, 4 ) with prescribed asymptotic behavior and value at origin. By similar translation as (1.5) and eigen-decomposition as (2.3), we may assume β = 0 and consider only the following Dirichlet problem.

 2 n  Gτ (λ(D u)) = C0, in R \{0}, u(0) = u0, (3.1)  1 T  u(x) → 2 x Ax + c , as |x| → ∞, where 0 < A = diag(a1, ··· , an) satisfies Gτ (a1, ··· , an) = C0 < 0.

We will write the vector (a1, ··· , an) simply as a when no confusion can arise. Similar to the strategy as in [2, 24, 27], we seek for generalized symmetric subsolution of form

n 1 X u(x) = U(s), where s(x) := a x2, (3.2) 2 i i i=1 and apply Perron’s method. By a direct computation,

0 0 00 Diu(x) = U (s)aixi,Diju(x) = U (s)aiδij + U (s)aiajxixj.

9 By Sylvester’s determinant theorem, we have

2 σn(λ(D u))  0     U a1 a1x1  ..  00  .   = det  .  + U (s)  .  · (a1x1, ··· , anxn) 0 U an anxn  0  U a1  ..  = det  .  0 U an   0 −1    (U a1) a1x1  00  ..   .  · 1 + U (a1x1, ··· , anxn)  .   .  0 −1 (U an) anxn n 0 n 00 0 n−1 X 2 = (U ) σn(a) + U (U ) σn(a) aixi , i=1 i.e., 2 0 n 00 0 n−1 σn(λ(D u)) = (U ) σn(a) + 2sU (U ) σn(a). (3.3) Similarly, 2 0 00 Dij(u + b|x| ) = (U ai + 2b)δij + U aiajxixj and

2 2 σn(λ(D (u + b|x| )))  0     U a1 + 2b a1x1  ..  00  .   = det  .  + U (s)  .  · (a1x1, ··· , anxn) 0 U an + 2b anxn  0  U a1 + 2b  ..  = det  .  0 U an + 2b   0 −1    (U a1 + 2b) a1x1  00  ..   .  · 1 + U (a1x1, ··· , anxn)  .   .  0 −1 (U an + 2b) anxn n n Y 0 00 X 2 2 Y 0 = (U ai + 2b) + U ai xi (U aj + 2b) i=1 i=1 j=1,··· ,n j6=i i.e.,

n n 2 2 Y 0 2b 00 X 2 Y 0 2b σn(λ(D (u + b|x| ))) = σn(a) (U + ) + U σn(a) aixi (U + ). (3.4) ai aj i=1 i=1 j=1,··· ,n j6=i

10 By (3.3) and (3.4), u(x) = U(s) is a subsolution of (1.6) as long as

0 n 00 0 n−1 (U ) σn(a) + 2sU (U ) σn(a)   ≥ c0, n n Y 0 00 X 2 Y 0 2b (U ai + 2b) + U σn(a) aixi (U + )  aj  i=1 i=1 j=1,··· ,n j6=i i.e., (U 0)n + 2sU 00(U 0)n−1   ≥ c0, (3.5) n n Y 0 2b 00 X 2 Y 0 2b (U + ) + U aixi (U + ) ai  aj  i=1 i=1 j=1,··· ,n j6=i where c = exp( √ 2b C ) ∈ (0, 1) satisfies 0 a2+1 0

n n Y ai Y 2b c = , i.e., 1 = c · (1 + ). (3.6) 0 a + 2b 0 a i=1 i i=1 i Hereinafter, we may assume without loss of generality that

0 < a1 ≤ a2 ≤ · · · ≤ an.

Lemma 3.1. Let U be a classical solution of

 n Y 2b  (U 0)n − c (U 0 + )  0 a  i=1 i  n ! 00 0 n−1 Y 0 2b (3.7)  +2(s + 1)U (U ) − c0 (U + ) = 0, in s > 0,  ai  i=2  U 0 ≥ 1,U 00 ≤ 0, in s > 0.

2 n Then u(x) = U(s) is convex and satisfies Gτ (λ(D u)) ≥ C0 in R \{0}. 0 Proof. By the assumptions on ai, for all U ≥ 1,   n n  n−1  n−1 X 2 Y 0 2b X 2 Y 0 2b Y 0 2b aixi (U + ) ≤ aixi (U + ) = 2s (U + ).  aj  aj aj i=1 j=1,··· ,n i=1 j=1 j=1 j6=i

We claim that the denominator part of (3.5) is positive. By U 0 ≥ 1,U 00 ≤ 0, we only need to prove U 0 + 2sU 00 > 0, (3.8)

11 then   n n Y 0 2b 00 X 2 Y 0 2b (U + ) + U aixi (U + ) ai  aj  i=1 i=1 j=1,··· ,n j6=i n n−1 Y 2b Y 2b ≥ (U 0 + ) + 2sU 00 (U 0 + ) a a i=1 i i=1 i n−1 ! Y 2b  2b  = (U 0 + ) · U 0 + + 2sU 00 a a i=1 i n > 0. By (3.7), we have

n 0 n Y 0 2b (U ) − c0 (U + ) ai 00 00 i=1 2sU ≥ 2(s + 1)U = − n . Y 2b (U 0)n−1 − c (U 0 + ) 0 a i=2 i Consequently, n n ! Y 0 2b 0 Y 0 2b c0 (U + ) − U (U + ) ai ai 0 00 i=1 i=2 U + 2sU ≥ n . Y 2b (U 0)n−1 − c (U 0 + ) 0 a i=2 i Notice that for all U 0 ≥ 1, in view of (3.6),

n n ! Y 2b Y 2b (U 0)n−1 − c (U 0 + ) = (U 0)n−1 1 − c (1 + ) > 0. 0 a 0 a U 0 i=2 i i=2 i Combining the calculus above, (3.8) is proved. By (3.3) and Proposition 1.2 of [2],

n 2 0 k 00 0 k−1 X 2 σk(λ(D u)) = (U ) σk(a) + U (U ) (aixi) σk−1;i(a). (3.9) i=1

Since U 0 ≥ 1 and U 00 ≤ 0, together with (3.8) and the fact that

σk(a) = σk;i(a) + aiσk−1;i(a), for all k = 1, 2, ··· , n, we have

n 2  0k 00 0k−1 X 2 σk λ D u = U σk(a) + U U aixi (σk(a) − σk;i(a)) i=1 0 k 00 0 k−1 ≥ (U ) σk(a) + 2sU (U ) σk(a) > 0.

12 2 n It follows that D u > 0 in R \{0}. By the calculus above, (3.5) holds as long as n 0 n Y 0 00 0 n−1 (U ) σn(a) − c0 (U ai + 2b) + 2sσn(a)U (U ) i=1  n 00 X 2 Y 0 2b −c0U σn(a) aixi (U + ) ≥ 0.  aj  i=1 j=1,··· ,n j6=i

Dividing both sides by σn(a) > 0, it is equivalent to   n n 0 n Y 0 2b 00 0 n−1 00 X 2 Y 0 2b (U ) − c0 (U + ) + 2sU (U ) − c0U aixi (U + ) ≥ 0. (3.10) ai  aj  i=1 i=1 j=1,··· ,n j6=i By the assumptions, we have   n n X 2 Y 0 2b Y 0 2b aixi (U + ) ≥ 2s (U + )  aj  ai i=1 j=1,··· ,n i=2 j6=i and   n 00 0 n−1 00 X 2 Y 0 2b 2sU (U ) − c0U aixi (U + )  aj  i=1 j=1,··· ,n j6=i n ! Y 2b ≥ 2sU 00 (U 0)n−1 − c (U 0 + ) . 0 a i=2 i By a direct computation, as long as U 0 ≥ 1, n n n ! Y 2b Y ai Y ai + 2b 2b (U 0)n−1−c (U 0+ ) ≥ (U 0)n−1· 1 − ( ) ( ) = (U 0)n−1· > 0. 0 a a + 2b a a + 2b i=2 i i=1 i i=2 i 1 Hence (3.10) holds if U 0 ≥ 1,U 00 ≤ 0 and n n ! Y 2b Y 2b (U 0)n − c (U 0 + ) + 2(s + 1)U 00 (U 0)n−1 − c (U 0 + ) = 0 0 a 0 a i=1 i i=2 i Especially when U satisfies (3.7), u(x) = U(s) becomes a subsolution of (3.1) and this finishes the proof.

Lemma 3.2. For any c ≥ 0, there exists a solution of (3.7) with U(0) = 0, U 0(0) ≥ 1 and U(s) = s + c + o(1) as s → ∞, (3.11) if a1, ··· , an satisfies n X a1 + 2b > 2. (3.12) a + 2b i=1 i

13 Proof. Step 1. existence of initial value problem  g(ψ(s))  ψ0(s) = , in s ∈ (0, ∞), s + 1 (3.13)  ψ(0) = α ≥ 1, where n Y 2b ψn − c (ψ + ) 0 a i=1 i g(ψ) := − n !. Y 2b 2 ψn−1 − c (ψ + ) 0 a i=2 i By a direct computation and (3.6), for any ψ > 1, Qn 2b n n c0 i=1(ψ + a ) Y 2b Y 2b i = c (1 + ) < c (1 + ) = 1, (3.14) ψn 0 a ψ 0 a i=1 i i=1 i for any ψ ≥ 1, Qn 2b n n c0 i=2(ψ + a ) Y 2b Y 2b i = c (1 + ) < c (1 + ) = 1 (3.15) ψn−1 0 a ψ 0 a i=2 i i=1 i and g(1) = 1. Thus for all ψ > 1, g(ψ) < 0 and ψ = 1 is the unique root of g(ψ) = 0 in ψ ≥ 1. If α = 1, then ψ ≡ 1 is the only solution by the uniqueness theorem of ordinary differential equations. If α > 1, by the local existence result of ODE, ψ exists in a neighbourhood of origin. Since ψ = 1 is the only root of g(ψ) = 0 in the range of ψ ≥ 1 and g(ψ) < 0 as long as ψ > 1, ψ(s) is monotone decreasing with respect to s. Thus ψ is bounded from above by α and hence by the extension theorem, ψ exists on entire s ∈ [0, ∞). Since g(1) = 0, by the uniqueness theorem of ODE, ψ(s) > 1 for all s ∈ [0, ∞). Hence ψ(s) admits a limit ψ∞ at infinity. We claim that the limit ψ∞ must be 1, i.e., ψ(s) → 1 as s → +∞. (3.16)

By contradiction, if ψ∞ > 1, then by equation in (3.13), there exists  > 0 and C = C(ψ∞, ) > 0 such that g(ψ(s)) ψ0(s) = < −, ∀ s > C. s + 1 Integral over s ∈ (C, ∞) and it contradicts to the fact that ψ converge to ψ∞ at infinity. This finishes the proof of (3.16). Step 2. refined asymptotics at infinity. By a direct computation and applying (3.14), (3.15),   n n !−1 0 1 X Y 2b Y 2b g (1) = − n − c0 (1 + ) · 1 − c0 (1 + ) . 2  aj  ai i=1 j=1,··· ,n i=2 j6=i Furthermore by (3.6),   n n n n X Y 2b X Y aj Y aj + 2b X ai c0 (1 + ) =  ( ) = aj  aj + 2b aj  ai + 2b i=1 j=1,··· ,n i=1 j=1 j=1,··· ,n i=1 j6=i j6=i

14 and n !−1 n n !−1 Y 2b Y ai Y ai + 2b a1 + 2b 1 − c (1 + ) = 1 − = . 0 a a + 2b a 2b i=2 i i=1 i i=2 i Thus n n 1 a1 + 2b X 2b 1 X a1 + 2b δ0 g0(1) = − = − = − < −1 2 2b a + 2b 2 a + 2b 2 i=1 i i=1 i and ψ − 1 satisfies g(ψ) − g(1) (ψ − 1)0 = and ψ − 1 = o(1) s + 1 as s → ∞. By standard ODE analysis, we prove

0 ψ(s) = 1 + O(sg (1)) as s → ∞. (3.17)

In fact, t := ln(s + 1) ∈ (0, +∞), ϕ(t) := ψ(s(t)) − 1 satisfies ϕ0(t) = ψ0(s(t)) · et = g(ϕ + 1) − g(1) =: h(ϕ) in (0, +∞). with h(0) = 0, h0(0) = g0(1) and ϕ(t) → 0 as t → +∞. Since g0(1) < 0 and h(ϕ) = g0(1)ϕ + O(ϕ2) as ϕ → 0, by the asymptotic stability of ODE (see for instance Theorem 1.1 of Chap.13 in [14] or Theorem 2.16 of [4] for more detailed analysis), we have

0 ϕ(t) = O(eg (1)t) as t → +∞, which is exactly (3.17). Step 3. construction of solution of second order ODE (3.7). By taking Z s U(s) := u0 + ψ(r)dr, s ≥ 0, 0 we have U 0(s) = ψ(s) ≥ 1 and thus U becomes a solution of (3.7). Step 4. computation on the asymptotics at infinity. By (3.17), for any α ≥ 1, there exist a solution of (3.13) (denoted by ψ(s, α)) and a constant µ(α) ≥ 0 such that

g0(1)+1 U(s) = u0 + s + µ(α) + O(s ) as s → ∞.

More explicitly, by Z s U(s) − s − u0 = (ψ(r, α) − 1)dr, 0 sending s → +∞ we have Z +∞ µ(α) = (ψ(s, α) − 1)ds. (3.18) 0

15 By the theorem of the differentiability of the solution with respect to the initial value, we can differentiate (3.13) with respect to α and obtain

( 0 ∂ψ(s, α) v0 = g (ψ(s,α)) · v, in s ∈ (0, ∞), v(s) := satisfies s+1 ∂α v(0) = 1, which is solved by Z s g0(ψ(r, α))  v(s) = exp dr > 0. 0 r + 1 Hence ψ(s, α) is monotone increasing w.r.t. α. Now we claim that for any fixed s > 0, lim ψ(s, α) = +∞. (3.19) α→+∞ ∞ By contradiction, we suppose there exist s1 > 0, M > 1 and {αk}k=1 such that

1 < αk → +∞, as k → ∞ and ψ(s1, αk) ≤ M for all k ∈ N. By a direct computation, there exists C > 0 independent of β such that 0 = g(1) < −g(ψ) < Cψ, ∀ ψ > 1. By (3.13), we have dψ dψ ds ≤ = − . Cψ −g(ψ) s + 1

Integrate in ψ ∈ (M, αk) for large k and using ψ(1, αk) = αk, we have

Z αk dψ Z ψ(1,αk) dψ Z 1 ds s + 1 ≤ = − = ln 1 < ∞. M Cψ ψ(s1,αk) −g(ψ) s1 s + 1 2 Sending k → +∞ and we have 1 α s + 1 ln k < ln 1 < +∞, C M 2 which becomes a contradiction. Consequently we prove that µ(α) → +∞ as α → +∞. (3.20) By contradiction, if µ(α) is uniformly bounded by some M for all α ≥ 1, then by (3.18) we have Z +∞ 2M ≥ (ψ(s, α) − 1)ds. 0

Fix s = 1 in (3.19), then there exists sufficiently large α0 such that ψ(1, α0) > 3M + 1. Since ψ(s, α) ≥ 1 is monotone decreasing with respect to s, we have ψ(s, α0) > 3M + 1 for all s ∈ (0, 1). This leads to a contradiction and proves (3.20). On the other hand, when α = 1, we have Z s U(s) = u0 + 1dr = s + u0, 0 which gives µ(1) = 0. By the continuity of solution w.r.t α, such a solution of (3.7) with U(0) = 0 and asymptotics (3.11) exists for all c ≥ 0.

16 By standard Perron’s method, we have the following existence result.

Theorem 3.3. For any given c ≥ u0 and 0 < A ∈ Sym(n) satisfying Gτ (λ(A)) = C0 and (1.11), 0 n there exists a unique convex viscosity solution u ∈ C (R ) of (3.1).

Proof. By Lemmas 3.1, 3.2 and the computation at the start part of this section, for any c ≥ u0 1 T we have a generalized symmetric subsolution u(x) := U( 2 x Ax) with 1 u(0) = u and u(x) = xT Ax + c + O(|x|2−δ0 ) 0 2

2 as |x| → ∞. By a direct computation, u is convex near origin and satisfies Gτ (λ(D u)) ≥ C0 in n R \{0}. By Theorem 2.2, 2 n Gτ (λ(D u)) ≥ C0 in R . On the other hand, 1 u(x) := xT Ax + c 2 2 n is a smooth convex solution of Gτ (λ(D u)) = C0 in R . From the proof of Lemma 3.2, we have n u ≤ u in R . In fact by (3.20),   Z s 1 T u(x) − x Ax + c = U − s − u0 − µα = (ψ(r, α) − 1)dr < 0, 2 +∞ where α ≥ 1 is the unique constant such that c = u0 + µα. Let

2 n u(x) := sup{v(x): Gτ (λ(D v)) ≥ C0, v(0) = u0 and u ≤ v ≤ u, in R \{0}}.

0 n 1 T By a direct computation, u ∈ C (R ) is convex, u(0) = u0 and u(x) → 2 x Ax + c at infinity. By Perron’s method (see for instance [2, 20]), u(x) defined above is a viscosity solution of (3.1). The uniqueness follows similarly by maximum principle as in Lemma 2.4.

Eventually, we finish this section by proving that the solution found satisfies (1.7) and (1.8). 2 n 2 In fact, by equation Gτ (λ(D u)) = C0 in R \{0} and convexity D u > 0, Theorem 2.2 implies n u is a subsolution on entire R . From the proof of Theorem 3.3, we have 1 u(x) ≤ u(x) = xT Ax + c in n. 2 R Hence (1.8) follows immediately.

4 Proof of Proposition 1.5

In this section, we prove Proposition 1.5 following the line of Bao-Li-Li [2]. Let U, s(x) be as in (3.2). For τ = 0 case, the result follows from the classification results by Jin-Xiong [21]. π For τ ∈ (0, 4 ) case, we only need to prove the following result for Gτ operator and desired result follows by eigen-decomposition (2.3).

17 Proposition 4.1. For any 0 < s1 < s2 < ∞, if there exists a generalized symmetric classical convex solution u(x) = U(s) of

 1  G (λ(D2u)) = C in x : xT Ax ∈ (s , s ) , (4.1) τ 0 2 1 2 where A = diag(a1, ··· , an) > 0 satisfies Gτ (λ(A)) = C0. Then either 1 a = a = ··· = a or u = xT Ax + c (4.2) 1 2 n 2 for some c ∈ R.

q 2s Proof. For any s ∈ (s1, s2) and i0 = 1, ··· , n, take x = (0, ··· , 0, , 0, ··· , 0). By (3.3) ai0 and (3.4), u(x) is a classical solution of (4.1) if

0 n 00 0 n−1 (U ) σn(a) + 2sU (U ) n = c0. Y 2b Y 2b (U 0 + ) + 2sU 00 (U 0 + ) ai ai i=1 i=1,··· ,n i6=i0 Consequently, n Y 2b (U 0)n + 2sU 00(U 0)n−1 − c (U 0 + ) 0 a i=1 i Y 2b (4.3) = c 2sU 00 (U 0 + ). 0 a i=1··· ,n i i6=i0 00 1 T If U ≡ 0 in (s1, s2), then u = 2 x Ax + c immediately. 00 If U (s) 6= 0 at a point s ∈ (s1, s2), since the left hand side of (4.3) is independent of choice of i0, we have Y 2b 2sU 00(s) (U 0(s) + ) = C(s) ∀ i = 1, 2, ··· , n. a 0 i=1··· ,n i0 i6=i0 Thus in this case, a1 = a2 = ··· = an.

π For τ = 4 case, similarly by eigen-decomposition, we only need to prove the following result.

Proposition 4.2. For any 0 < s1 < s2 < ∞, if there exists a generalized symmetric classical convex solution u(x) = U(s) of

 1  G (λ(D2u)) = C in x : xT Ax ∈ (s , s ) , (4.4) τ 0 2 1 2 where A = diag(a1, ··· , an) > 0 satisfies Gτ (λ(A)) = C0. Then (4.2) holds for some c ∈ R.

18 q 2s Proof. For any s ∈ (s1, s2) and i0 = 1, ··· , n, take x = (0, ··· , 0, , 0, ··· , 0). As a ai0 necessary condition for such solution exists, C0 < 0. Since u(x) is a classical solution of (4.4), we have 2 2 c0 · σn(λ(D u)) = σn−1(λ(D u)), √ 2 where c0 = − 2 C0 > 0. By (3.3) and Proposition 1.2 of [2],

n 0 n 00 0 n−1 X 2 c0 · (U ) σn(a) + c0 · U (U ) σn(a) aixi i=1 n 0 n−1 00 0 n−2 X 2 = (U ) σn−1(a) + U (U ) (aixi) σn−2;i(a), i=1 i.e.,

0 n 00 0 n−1 0 n−1 00 0 n−2 c0(U ) σn(a) + 2c0sU (U ) σn(a) − (U ) σn−1(a) = 2sU (U ) ai0 σn−2;i0 (a). (4.5)

00 0 σn−1(a) 1 T If U ≡ 0 in (s1, s2), then U = = 1 and u = x Ax + c is a quadratic solution. c0σn(a) 2 00 If U (s) 6≡ 0 at a point s ∈ (s1, s2), since the left hand side of (4.5) is independent of choice of i0, we have 00 0 n−2 2sU (U ) ai0 σn−2;i0 (a) = C(s) ∀ i0 = 1, 2, ··· , n. By the continuity of solution, we may assume without loss of generality that U 0(s) 6= 0, otherwise there exists s0 close to s such that U 00(s0) 6= 0 and U 0(s0) 6= 0. Together with the fact that

σn−1(a) = σn−1;i(a) + aiσn−2;i(a) ∀ i = 1, 2, ··· , n,

we have σn−1;i0 (a) is independent of the choice of i0. Thus in this case, all ai0 are the same. π π π For τ ∈ ( 4 , 2 ), the equations are very similar to τ = 2 case and only need to prove for the latter case for simplicity.

Proposition 4.3. For any 0 < s1 < s2 < ∞, if there exists a generalized symmetric classical convex solution u(x) = U(s) of

n X  1  arctan λ (D2u) = C in x : xT Ax ∈ (s , s ) , (4.6) i 0 2 1 2 i=1 Pn where A = diag(a1, ··· , an) > 0 satisfies i=1 arctan ai = C0. Then (4.2) holds up to some constant c.

Proof. By the algebraic form of (4.6) (see for instance [10]), λ(D2u) satisfies

X k 2  X k 2  cos C0 (−1) σ2k+1 λ D u − sin C0 (−1) σ2k λ D u = 0. 0≤2k+1≤n 0≤2k≤n

19 q 2s For any s ∈ (s1, s2) and i0 = 1, ··· , n, take x = (0, ··· , 0, , 0, ··· , 0). Since u(x) is a ai0 classical solution of (4.6), by (3.3) and (3.9) we have   X k  0 2k+1 00 0 2k  cos C0  (−1) (U ) σ2k+1(a) + 2sU (U ) ai0 σ2k;i0 (a)  0≤2k+1≤n   X k  0 2k 00 0 2k−1  = sin C0  (−1) (U ) σ2k(a) + 2sU (U ) ai0 σ2k−1;i0 (a)  , 0≤2k≤n i.e.,

 00 T 2sp1U   00 0 ai0 n−2  2sp2U U  X  00 0 2   ai0 σ1;i0 (a)  2sp U 00(U 0)ka σ (a) =  2sp3U (U )  ·  = G(s), (4.7) k i0 k;i0    .   .   .  k=0  .  00 0 n−2 ai0 σn−2;i0 (a) 2spn−2U (U ) where  k k   p = (−1)k cos C + − π k 0 2 2 for k = 0, 1, ··· , n − 1 and   X k 0 2k+1 G(s) = − cos C0  (−1) (U ) σ2k+1(a) 0≤2k+1≤n   X k 0 2k + sin C0  (−1) (U ) σ2k(a) 0≤2k≤n 00 0 n−1 −2spn−1U (U ) σn(a).

Since sin C0 = 0 and cos C0 = 0 cannot hold at the same time, we may only consider the non-zero part of (4.7) and the following proof still works. Hereinafter, we only prove for sin C0 ·cos C0 6= 0 case. 00 0 1 T If U ≡ 0 in (s1, s2), then U ≡ 1 and u = 2 x Ax + c is a quadratic polynomial. 00 If U (s) 6≡ 0 at a point s ∈ (s1, s2), then by the continuity of solution, we may assume 00 0 without loss of generality that U (s) 6= 0 and U (s) 6= 0 in (s3, s4) ⊂ (s1, s2). Pick n − 1 points such that s3 < r1 < r2 < ··· < rn−1 < s4, ri 6= 0 and we prove that the matrix formed by coefficients of (4.7) is invertible, i.e.,

 00 00 0 n−2  2r1p1U (r1) ··· 2r1pn−2U (r1)(U (r1)) 00 00 0 n−2  2r2p1U (r2) ··· 2r2pn−2U (r2)(U (r2))  det   6= 0. (4.8)  . .. .   . . .  00 00 0 n−2 2rnp1U (rn−1) ··· 2rn−1pn−2U (rn−1)(U (rn−1))

20 By (4.7) and (4.8), (ai0 , ··· , ai0 σn−1;i0 (a)) is a constant vector independent of i0. Thus in this case, all ai0 are the same. 00 0 To prove (4.8), since sin C0 · cos C0 6= 0, 0 < r1 < ··· < rn−1, U (s) 6= 0 and U (s) 6= 0, by the determinate of Vandermonde form matrix, we have

 0 0 n−2  1 U (r1) ··· (U (r1)) 0 0 n−2  1 U (r2) ··· (U (r2))  Y det   = (U 0(r ) − U 0(r )) 6= 0.  . . .. .  i j  . . . .  i>j 0 0 n−2 i,j=1,2,··· ,n−1 1 U (rn−1) ··· (U (rn−1))

Hence (4.8) is proved and the result follows immediately.

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 Yanyan Li. On the exterior Dirichlet problem for Hessian equations. Trans. Amer. Math. Soc., 366(12):6183–6200, 2014. [3] Jiguang Bao, Haigang Li, and Lei Zhang. Monge-Ampere` equation on exterior domains. Calc. Var. Partial Differential Equations, 52(1-2):39–63, 2015. [4] Sigrun Bodine and Donald Lutz. Asymptotic integration of differential and difference equations, volume 2129 of Lecture Notes in Mathematics. Springer, Cham, 2015. [5] Luis Caffarelli. Interior W 2,p estimates for solutions of the Monge-Ampere` equation. Ann. of Math. (2), 131(1):135–150, 1990. [6] Luis Caffarelli and Xavier Cabre.´ Fully nonlinear elliptic equations, volume 43 of American Mathe- matical Society Colloquium Publications. American Mathematical Society, Providence, RI, 1995. [7] Luis Caffarelli and Yan Yan Li. An extension to a theorem of Jorgens,¨ Calabi, and Pogorelov. Comm. Pure Appl. Math., 56(5):549–583, 2003. [8] Luis Caffarelli and Yan Yan Li. A Liouville theorem for solutions of the Monge-Ampere` equation with periodic data. Ann. Inst. H. Poincare´ Anal. Non Lineaire´ , 21(1):97–120, 2004. [9] Luis Caffarelli, Yanyan Li, and . Some remarks on singular solutions of nonlinear elliptic equations III: viscosity solutions including parabolic operators. Comm. Pure Appl. Math., 66(1):109–143, 2013. [10] Luis Caffarelli, Louis Nirenberg, and . The Dirichlet problem for nonlinear second-order elliptic equations. III. Functions of the eigenvalues of the Hessian. Acta Math., 155(3-4):261–301, 1985. [11] Eugenio Calabi. Improper affine hyperspheres of convex type and a generalization of a theorem by K. Jorgens.¨ Michigan Math. J., 5:105–126, 1958. [12] Jingyi Chen, Ravi Shankar, and Yu Yuan. Regularity for convex viscosity solutions of special La- grangian equation. arXiv:1911.05452, 2019. [13] 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. [14] Earl A. Coddington and Norman Levinson. Theory of ordinary differential equations. McGraw-Hill Book Company, Inc., New York-Toronto-London, 1955. [15] Leonor Ferrer, Antonio Mart´ınez, and Francisco Milan.´ An extension of a theorem by K. Jorgens¨ and a maximum principle at infinity for parabolic affine spheres. Math. Z., 230(3):471–486, 1999. [16] Harley Flanders. On certain functions with positive definite Hessian. Ann. of Math. (2), 71:153–156, 1960.

21 [17] 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. [18] Guanghao Hong. A remark on Monge-Ampere` equation over exterior domains. arXiv 2007.12479, 2020. [19] Rongli Huang and Zhizhang Wang. On the entire self-shrinking solutions to Lagrangian mean curva- ture flow. Calc. Var. Partial Differential Equations, 41(3-4):321–339, 2011. [20] Hitoshi Ishii. On uniqueness and existence of viscosity solutions of fully nonlinear second-order elliptic PDEs. Comm. Pure Appl. Math., 42(1):15–45, 1989. [21] Tianling Jin and Jingang Xiong. Solutions of some Monge-Ampere` equations with isolated and line singularities. Adv. Math., 289:114–141, 2016. [22] Konrad Jorgens.¨ Uber¨ die Losungen¨ der Differentialgleichung rt−s2 = 1. Math. Ann., 127:130–134, 1954. [23]J urgen¨ Jost and Yuan Long Xin. Some aspects of the global geometry of entire space-like submani- folds. volume 40, pages 233–245. 2001. Dedicated to Shiing-Shen Chern on his 90th birthday. [24] Dongsheng Li and Zhisu Li. On the exterior Dirichlet problem for Hessian quotient equations. J. Differential Equations, 264(11):6633–6662, 2018. [25] Dongsheng Li, Zhisu Li, and Yu Yuan. A Bernstein problem for special Lagrangian equations in exterior domains. Adv. Math., 361:106927, 29, 2020. [26] YanYan Li and Siyuan Lu. Existence and nonexistence to exterior Dirichlet problem for Monge- Ampere` equation. Calc. Var. Partial Differential Equations, 57(6):Paper No. 161, 17, 2018. [27] Zhisu Li. On the exterior Dirichlet problem for special Lagrangian equations. Trans. Amer. Math. Soc., 372(2):889–924, 2019. [28] Zixiao Liu and Jiguang Bao. Asymptotic expansion at infinity of solutions of special Lagrangian equations. arXiv. 2012.13509, 2020. [29] Zixiao Liu and Jiguang Bao. Asymptotic expansion at infinity of solutions of Monge-Ampere` type equations. arXiv. 2102.08723, 2021. [30] AlekseIˇ Vasil’evich Pogorelov. On the improper convex affine hyperspheres. Geometriae Dedicata, 1(1):33–46, 1972. [31] Eduardo V. Teixeira and Lei Zhang. Global Monge-Ampere` equation with asymptotically periodic data. Indiana Univ. Math. J., 65(2):399–422, 2016. [32] Micah Warren. Calibrations associated to Monge-Ampere` equations. Trans. Amer. Math. Soc., 362(8):3947–3962, 2010. [33] Yu Yuan. A Bernstein problem for special Lagrangian equations. Invent. Math., 150(1):117–125, 2002. [34] 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: [email protected] Email: [email protected]

22