Loading [MathJax]/jax/element/mml/optable/MathOperators.js
Research article

A priori bounds and existence of smooth solutions to Minkowski problems for log-concave measures in warped product space forms

  • Received: 10 February 2023 Revised: 17 March 2023 Accepted: 17 March 2023 Published: 03 April 2023
  • MSC : 35B45, 35J96, 53C42

  • In the present paper, we prove the a priori bounds and existence of smooth solutions to a Minkowski type problem for the log-concave measure ef(|x|2)dx in warped product space forms with zero sectional curvature. Our proof is based on the method of continuity. The crucial factor of the analysis is the a priori bounds of an auxiliary Monge-Ampère equation on Sn. The main result of the present paper extends the Minkowski type problem of log-concave measures to the space forms and it may be an attempt to get some new analysis for the log-concave measures.

    Citation: Zhengmao Chen. A priori bounds and existence of smooth solutions to Minkowski problems for log-concave measures in warped product space forms[J]. AIMS Mathematics, 2023, 8(6): 13134-13153. doi: 10.3934/math.2023663

    Related Papers:

    [1] Amira A. Ishan, Meraj Ali Khan . Chen-Ricci inequality for biwarped product submanifolds in complex space forms. AIMS Mathematics, 2021, 6(5): 5256-5274. doi: 10.3934/math.2021311
    [2] Ali H. Alkhaldi, Akram Ali, Jae Won Lee . The Lawson-Simons' theorem on warped product submanifolds with geometric information. AIMS Mathematics, 2021, 6(6): 5886-5895. doi: 10.3934/math.2021348
    [3] Tong Wu, Yong Wang . Super warped products with a semi-symmetric non-metric connection. AIMS Mathematics, 2022, 7(6): 10534-10553. doi: 10.3934/math.2022587
    [4] Ibrahim Al-Dayel, Meraj Ali Khan . Ricci curvature of contact CR-warped product submanifolds in generalized Sasakian space forms admitting nearly Sasakian structure. AIMS Mathematics, 2021, 6(3): 2132-2151. doi: 10.3934/math.2021130
    [5] Ali H. Alkhaldi, Meraj Ali Khan, Shyamal Kumar Hui, Pradip Mandal . Ricci curvature of semi-slant warped product submanifolds in generalized complex space forms. AIMS Mathematics, 2022, 7(4): 7069-7092. doi: 10.3934/math.2022394
    [6] Jiale Gao, Kezheng Zuo, Qingwen Wang, Jiabao Wu . Further characterizations and representations of the Minkowski inverse in Minkowski space. AIMS Mathematics, 2023, 8(10): 23403-23426. doi: 10.3934/math.20231189
    [7] Noura Alhouiti, Fatemah Mofarreh, Akram Ali, Fatemah Abdullah Alghamdi . On gradient normalized Ricci-harmonic solitons in sequential warped products. AIMS Mathematics, 2024, 9(9): 23221-23233. doi: 10.3934/math.20241129
    [8] Pengfei Zhu, Kai Liu . Numerical investigation of convergence in the L norm for modified SGFEM applied to elliptic interface problems. AIMS Mathematics, 2024, 9(11): 31252-31273. doi: 10.3934/math.20241507
    [9] Fatemah Mofarreh, S. K. Srivastava, Anuj Kumar, Akram Ali . Geometric inequalities of PR-warped product submanifold in para-Kenmotsu manifold. AIMS Mathematics, 2022, 7(10): 19481-19509. doi: 10.3934/math.20221069
    [10] Biswabismita Bag, Meraj Ali Khan, Tanumoy Pal, Shyamal Kumar Hui . Geometric analysis on warped product semi-slant submanifolds of a locally metallic Riemannian space form. AIMS Mathematics, 2025, 10(4): 8131-8143. doi: 10.3934/math.2025373
  • In the present paper, we prove the a priori bounds and existence of smooth solutions to a Minkowski type problem for the log-concave measure ef(|x|2)dx in warped product space forms with zero sectional curvature. Our proof is based on the method of continuity. The crucial factor of the analysis is the a priori bounds of an auxiliary Monge-Ampère equation on Sn. The main result of the present paper extends the Minkowski type problem of log-concave measures to the space forms and it may be an attempt to get some new analysis for the log-concave measures.



    In the present paper, we focus on the geometry of log-concave measure which is defined as follows:

    Definition A.1 (Log-concave Measure(see [12,34,43])). A measure μ is called log-concave if its density dμ(x)dx is log-concave, that is, dμ(x)dx=ef(x) for some convex function f which means that

    μ(E)=Eef(x)dx (1.1)

    for every Borel set ERn+1 and some convex function f.

    Now, we provide some examples of log-concave measures.

    Examples A.2 (i) Gauss measure. The Gauss measure γn on Rn is defined as follows,

    dγn=1(2π)n+12e|x|22dx (1.2)

    which characterizes the Gaussian generalized random processes in stochastic analysis, see Bogachev [4].

    (ii) The weighted Bergman measure in Siegel domain. The domain

    Ω2={z=(z,zn+1)Cn+1:Im(zn+1)>|z|2}. (1.3)

    is a pseudo-convex domain in Cn+1. In order to analyze some potential theory on Ω2, such as the estimates of Cauchy-Szegö kernel on Ω2, the suitable Bergman space X(Ω2) may be chosen provided the Bergman norm X(Ω2) is well-defined where

    gX(Ω2)=Cn+1|g(z)|2e4πλ|z|2dV(z) (1.4)

    for any holomorphic function g in Cn+1, see pp. 45–66 of Chang and Tie [9]. We may call the measure

    dˉV(z)=e4πλ|z|2dV(z)

    be the weighted Bergman measure associated with the Siegel domain Ω2 and it is easy to see that dˉV is a log-concave measure for any fixed λ>0.

    (iii) Gibbs measure of some nonlinear Schrödinger equation. The Gibbs measure P(du) of some nonlinear Schrödinger equation is defined as follows:

    P(du)=eH(u)du (1.5)

    where

    H(u)=12Rn|u|2dx. (1.6)

    That is, H(u) is the Hamilton functions for the following Schrödinger equation with unit mass,

    itu=Δu, (1.7)

    (see a similar description of [17]). The Gibbs measures play an important role in quantum field theory and regularity and asymptotic behaviors of the Cauchy problem for some Schrödinger equations, see [17,19,35] and their references.

    It may be interesting to mention that some of the classical concepts and results in integral geometry have been generalized to the log-concave measures, such as the support function and Steiner type formulas. Moreover, the convexity of f can be used to deduce some interesting geometric inequalities for the measure ef(x)dx, such as Brunn-Minkowski inequality, Prékopa-Leindler inequalities or Blaschke-Santaló inequalities and so on, see [4,5,7,12,14,18,34,43]. Naturally, the prescribed log-concave measure problem has also been posed and studied which is called the Lp Minkowski problem of the log-concave measure in the present paper, see [12,15,30,38]. The works of [12,15,30,38] can be formulated in the following way:

    Problem A.3. For any fixed n1 and pR, given any Borel measures 1ψ(x)dx, find a convex function u such that

    (u)(up1ψ(x)dx)=ef(|y|2)dy. (1.8)

    In particular, if the measure dμ and ef(x)dx are both supported on the whole space Rn+1, Problem A.3 is the so-called Lp Minkowski problem for log-concave measure and has been analyzed, see [12,15,43].

    If N=Sn and the support set of the measure ef(|x|2)dx lies on the boundary of a hypersurface M, noting that in smooth case, the normal mapping ν and the support function u of a hypersurface M satisfies

    ν1=u, (1.9)

    Problem A.3 can be stated as follows,

    Problem A.4 (A Minkowski problem for log-concave measure). For any fixed n1 and pR, given a non-negative, finite Borel measure dμ=1ψ(ξ)dξ defined on the unit sphere Sn, find a convex hypersurface MRn+1 such that

    ν(u1pef(|x|2)dσ(x))=dμ(ξ), (1.10)

    where ν, u and dσ are the normal mapping, support function and surface measure of a convex hypersurface MRn+1 respectively, f is convex.

    In particular, if f0, Problem A.4 becomes the following classical Minkowski problem for p-curvature function which is also called Lp Minkowski problem.

    Problem A.5 (The classical Minkowski problem for p-curvature function). For any fixed n1 and pR, given a non-negative, finite Borel measure μ defined on the unit sphere Sn, find a convex hypersurface MRn+1 such that

    ν(u1pdσ(x))=dμ(ξ), (1.11)

    where ν, u and dσ are the normal mapping, support function and surface measure of a convex hypersurface MRn+1 respectively.

    In particular, if p=1, Problem A.5 was posed and analyzed by Minkowski for his wonderful construction of the Gaussian curvature (measure) via natural arguments in convex and integral geometry provided the measure dμ is the sum of the delta measure or the measure dμ is absolutely continuous with respect to the spherical Lebesgue measure whose density is continuous, see [44]. Later, Aleksandrov (see [44]) and Fenchel and Jensen (see [44]) generated the result of Minkowski for the general Borel measure on the unit sphere independently. Later, with the help of PDEs, Problem A.5 was resolved by Lewy, Nirenberg, Pogorelov, Cheng and Yau and so on, (see [44]).

    Noting that Gaussian curvature is the Jocabian of normal mapping,

    Problem A.6 (A prescribed Gaussian curvature problem). For any fixed n1 and pR, find a smooth convex hypersurface MRn+1 such that

    u1pef(ρ2)K=1ψ(ξ), (1.12)

    where u, ρ and K are the support function, radial function and Gaussian curvature function of a convex hypersurface MRn+1, f is convex.

    It may be worth mention that the arguments of Minkowski are based on some convex analysis of volume under the Minkowski sum, such as Brunn-Minkowski inequality and Hadamard variational formula, see [8,21,44]. A natural generalization of Minkowski sum is the so-called p-sum posed by Firey [16] for any fixed p1 due to the convexity of the function g(t)=tp for p1. Based on the concept of p-sum posed by Firey, Lutwak [39] introduced to p-Gaussian curvature function and posed and studied Problem A.5 which is called Lp Minkowski problem later. More results on Lp Minkowski problem can be referred to [6,11,22,26,27,28,31,32,40,41]

    On the one hand, recently, more and more researchers have been focusing on prescribed curvature problem in Riemannian manifolds, see [10,24,37] and so on. In the point of view of development of geometric analysis, it is interesting to focus on geometric problems for log-concave measures in Riemannian manifolds.

    On the other hand, recently, classical stochastic analysis has been developing in Riemannian manifolds [3,13,29,45]. It follows from Example A.2 that the log-concave measures have their origin in stochastic analysis, it is also interesting to focus on the theory of integral geometry of log-concave measures in Riemannian manifolds.

    The main focus of the present paper is on Problem A.3 when the support set of the log-concave measure ef(|x|2)dx enjoys a more interesting metric structure. Among them, one interesting object is the so-called warped product space forms.

    In the polar coordinate system, the metric of the hypersurface M satisfies

    ds2=dρ2(ξ)+ρ2(ξ)dξ2. (1.13)

    where ρ is the so-called radial function defined in (1.2). As a generalization to (1.13), one may consider a hypersurface M in Rn+1 whose metric satisfies

    ds2=dρ2(ξ)+φ2(ρ)(ξ)dξ2 (1.14)

    for any given function φ:(0,)R, see [2]. In particular, if φ(ρ)=ρ, M is a hypersurface in Euclidean Space. Motivated by the work of Aleksandrov's construction of integral Gaussian curvature, Oliker [42] focused on the existence of hypersurface which was prescribed the so-called integral Gaussian curvature and zero sectional curvature in a smooth frame. The works of Aleksandrov and Oliker provided new motivations on the geometric analysis of the warped product space forms, see [2,23,25,33,36,46] and so on.

    It is worth mentioning that the target hypersurfaces of [30,38] both lie in Rn+1, that is, φ(ρ)=ρ provided we suppose the metric of M with the form (1.14). It is natural to analyze Problem A.3 when the metric of the hypersurface M satisfies (1.14) for a given function φ.

    In a smooth frame, if the sectional curvature of the hypersurface M is zero, we know that the Gaussian curvature K and the support function u of M can be written as follows:

    K=det(ρij+2φ(ρ)φ(ρ)ρiρj+φ(ρ)φ(ρ)δij)φn2(ρ)(φ2(ρ)+|ρ|2)n2+1 (1.15)

    and

    u=φ2(ρ)φ2(ρ)+|ρ|2 (1.16)

    (see Lemma A in Appendix). Therefore, we focus on the existence of smooth solutions to the Eq (1.1) on the unit sphere Sn:

    det(ρij+2φ(ρ)φ(ρ)ρiρj+φ(ρ)φ(ρ)δij)(φ2(ρ)+|ρ|2)n+1+p2=ψ(ξ)ef(ρ2)φn2p(ρ). (1.17)

    Before stating the main result of the present paper, we assume the following conditions hold.

    (A.1.) f, φ and ψ are both C2 positive functions,

    fC2(R)+ψC2(R)+φC2(R)< (1.18)

    and

    {limρef(ρ2)φn+1p(ρ)(φ(ρ))n=0;limρ0ef(ρ2)φn+1p(ρ)(φ(ρ))n=. (1.19)

    (A.2.) The function ef(t2)φn+1+p(t) is non-increasing on (0,).

    (A.3.)inft>0φ(t)0 and there exists a positive number γ such that

    ntφ(t)γφ(t) (1.20)

    for any t>0.

    The main result of the present paper can be stated as follows,

    Theorem 1.1. For any fixed n1 and p>n1, suppose that the assumptions (A.1.)(A.3.) holds, then there exists a ρC2(Sn) to Eq (1.17) satisfying

    ρC2(Sn)c, (1.21)

    where c is independent of ρ.

    Remark 1.2. It follows from (1.1) that the equation is associated to the so-called prescribed Gauss curvature problem for the log-concave measure ef(|x|2)dx which may be an attempt on more differential geometric analysis for the log-concave measure ef(|x|2)dx. In particular, if f(t)=n+12ln(2π)+t2, some of these topics have been focused on, see [4,5,7,8,18].

    The rest of the paper is organized as follows: In Section 2, we get the a priori bounds of solutions. In Section 3, we prove Theorem 1.1. In the Appendix, we list some basic geometric quantity associated to the discussion.

    Section 2 devotes to the a priori bounds of solutions to the following equation on the unit sphere Sn:

    det(ρij+2φ(ρ)φ(ρ)ρiρj+φ(ρ)φ(ρ)δij)(φ2(ρ)+|ρ|2)n+1+p2=ψ(ξ)ef(ρ2)φn2p(ρ). (2.1)

    We let the set of the positive continuous function on the unit sphere Sn be C+(Sn) and

    C={ρC2,σ(Sn):(ρij+2φ(ρ)φ(ρ)ρiρj+φ(ρ)φ(ρ)δij) is positive definite}. (2.2)

    This main result of this section can be stated as follows,

    Theorem 2.0. For any fixed n1 and p>n1, let ρCC+(Sn) be a solution to (2.1) and f, φ and ψ satisfy the condition (A.1.). Then there exists a positive constant c, independent of ρ, such that

    0<c1ρC2,σ(Sn)c<, (2.3)

    where σ(0,1).

    Now, we divide the proof of Theorem 2.0 into following several of lemmas.

    Lemma 2.1. For any fixed n1, let ρCC+(Sn) be a solution to (2.1) and f, φ and ψ satisfy the condition (A.1.). Then there exists a positive constant c, independent of ρ, such that

    0<c1ρ(ξ)c<,ξSn. (2.4)

    Proof. We consider the following extremal problem,

    R=maxξSnρ(ξ). (2.5)

    It follows from the compactness of Sn and the continuity of ρ that there exists ξ1Sn such that

    R=ρ(ξ1). (2.6)

    It follows from (2.1) that at the point ξ=ξ1,

    (φ(R))nφ1+p(R)=(φ(R)φ(R))nφn+1+p(R)det(ρij+2φ(ρ)φ(ρ)ρiρj+φ(ρ)φ(ρ)δij)(φ2(ρ)+|ρ|2)n+1+p2=ψ(ξ1)ef(R2)φn2p(R), (2.7)

    that is,

    ef(R2)φn+1p(R)(φ(R))n1ψ(ξ1)1maxξSnψ(ξ)>0. (2.8)

    However, there exists a contradiction between (2.8) and (1.19) provided R is sufficiently large. This implies there exists a positive constant c>0 such that

    Rc<. (2.9)

    Adopting a similar argument, we also get

    r1c>0. (2.10)

    (2.9) and (2.10) yield the desired conclusion of lemma 2.1.

    The following lemma can be referred to [2]. For the sake of the completeness, we give the proof here.

    Lemma 2.2. We let ρ be a solution of (2.1) and v(ξ)=φ(ρ(ξ))φ(ρ(ξ)), then v solves the following equation,

    det(vij+vδij)(φ2(ρ)+|ρ|2)n+1+p2=ψ(ξ)ef(ρ2)φn2p(ρ). (2.11)

    Proof. By the definition of v, we have,

    vi=ρiφ2(ρ),vij=ρijφ2(ρ)+2φ(ρ)φ3(ρ)ρiρj (2.12)

    and

    vij+vδij=1φ2(ρ)(ρij+2φ(ρ)φ(ρ)ρiρj+φ(ρ)φ(ρ)δij). (2.13)

    Therefore,

    u1pef(ρ2)ψ(ξ)=K=det(ρij+2φ(ρ)φ(ρ)ρiρj+φ(ρ)φ(ρ)δij)φn2(ρ)(φ2(ρ)+|ρ|2)n+22=det(vij+vδij)φn2(ρ)(φ2(ρ)+|ρ|2)n+22. (2.14)

    Noting that

    u=φ2(ρ)φ2(ρ)+|ρ|2 (2.15)

    we have,

    det(vij+vδij)(φ2(ρ)+|ρ|2)n+1+p2=ψ(ξ)ef(ρ2)φn2p(ρ). (2.16)

    This is the desired conclusion of Lemma 2.2.

    In the rest of this section, we will consider the a priori bounds of solutions to Eq (2.11).

    It is easy to see that

    ρCv{vC2,σ(Sn):(vij+δijv) is positive definite} (2.17)

    Lemma 2.3. For any fixed n\ge 1 and p > -n-1 , we let v\in \bar{\mathcal{C}}\cap C_+(\mathbb{S}^n) be the solution of (2.11). Suppose that f , \varphi and \psi satisfy the condition (A.1.) . Then there exists a positive constant c , independent of \rho , such that

    \begin{equation} 0\le|\nabla v(\xi)|\le c, \forall \xi\in \mathbb{S}^n. \end{equation} (2.18)

    Thus,

    \begin{equation} 0\le|\nabla \rho(\xi)|\le c, \forall \xi\in \mathbb{S}^n. \end{equation} (2.19)

    Proof. The proof follows from the argument of Oliker [42]. Let \{\xi_i\}_{i = 1}^n be a system of smooth local orthogonal coordinates on \mathbb{S}^n , we therefore get

    \begin{equation} |d\xi|^2 = \delta_{ij}d\xi_id\xi_j, \end{equation} (2.20)

    where \delta_{ij} is the Dirac notation. It is easy to see that

    \begin{equation} |\nabla v|^2 = \delta_{ij}\frac{\partial v}{\partial \xi_i}\frac{\partial v}{\partial \xi_j} = \sum\limits_{i = 1}^n|\frac{\partial v}{\partial \xi_i}|^2, \end{equation} (2.21)

    (see pp. 812 of Oliker [42]). We let

    \begin{equation*} \label{2303} u = \frac{v^2+|\nabla v|^2}{2}. \end{equation*}

    Suppose that there exists \xi_2\in \mathbb{S}^n such that

    \begin{equation*} \label{2305} u(\xi_2) = \max\limits_{\xi \in \mathbb{S}^n}u(\xi). \end{equation*}

    Then, for any fixed i\in \{1, 2, \cdots, n\} , at the point \xi_2 ,

    \begin{equation} 0 = u_i = \sum\limits_{j = 1}^{n}(v_{ij}+v\delta_{ij})\frac{\partial v}{\partial \xi_j}. \end{equation} (2.22)

    It follows from Lemma 2.1 that there exists a positive constant c such that

    \begin{equation} \det(v_{ij}+v\delta_{ij}) = \psi(\xi)e^{f(\rho^2)}\varphi^{-n-2p}(\rho) (\varphi^2(\rho)+|\nabla\rho|^2)^{\frac{n+1+p}{2}}\ge c > 0 \end{equation} (2.23)

    at the point \xi_0 . This means that the matrix (v_{ij}+v\delta_{ij})_{n\times n} is nonsingular at the point \xi_2 . Therefore, combining with (2.22), we get

    \begin{equation*} \label{2308} v_k(\xi_2) = 0 \end{equation*}

    for any fixed k\in \{1, 2, \cdots, n\} . Therefore, it follows from Lemma 2.1 that there exists a positive constant c such that

    \begin{equation*} \label{2309} \frac{1}{2}|\nabla v|^2(\xi)\le u(\xi)\le u(\xi_2) = \frac{1}{2}\max\limits_{\xi}v(\xi)\le c, \forall \xi\in \mathbb{S}^n. \end{equation*}

    This completes the proof of Lemma 2.3.

    We first let W_{ij} = (v_{ij}+\delta_{ij}v) , \mathcal{G}(W_{ij}) = (\det W_{ij})^{\frac{1}{n}} and

    \Psi(\xi) = (\psi(\xi)e^{f(\rho^2)}\varphi^{-n-2p}(\rho) (\varphi^2(\rho)+|\nabla\rho|^2)^{\frac{n+1+p}{2}})^{\frac{1}{n}},

    then Eq (2.11) becomes

    \begin{equation} \mathcal{G}(W_{ij}) = \Psi. \end{equation} (2.24)

    Lemma 2.4. For any fixed n\ge 1 and p > -n-1 , let \rho\in \mathcal{C}\cap C_+(\mathbb{S}^n) be a solutions of (2.1). Suppose that f , \varphi and \psi satisfy the condition (A.1.) . Then there exists a positive constant c , independent of \rho , such that

    \begin{equation} -\Delta \rho\le c. \end{equation} (2.25)

    Proof. Let H = \sum_{i}W_{ii} = \Delta v+nv . By the commutator identity, we have,

    \begin{equation} H_{ii} = \Delta W_{ii}-nW_{ii}+H. \end{equation} (2.26)

    Suppose that H achieves it maximum at the point \xi = \xi_3 . Without loss of generality, we may (H_{ij})_{n\times n} is diagonal at the point \xi = \xi_3 . Therefore, at the point \xi = \xi_3 ,

    \begin{equation} 0\ge \mathcal{G}^{ij}H_{ij} = \mathcal{G}^{ii}(\Delta W_{ii})-n\mathcal{G}^{ii}+H\Sigma_{i}\mathcal{G}^{ii}. \end{equation} (2.27)

    It follows from (2.25) that

    \begin{equation} \mathcal{G}^{ij}W_{ij\alpha} = \Psi_\alpha, \mathcal{G}^{ij,rs}W_{ij\alpha}W_{rs\alpha}+\mathcal{G}^{ij}\Delta W_{ij} = \Delta\Psi \end{equation} (2.28)

    By the concavity of \mathcal{G} , we have

    \begin{equation} \mathcal{G}^{ij,rs}W_{ij\alpha}W_{rs\alpha}\le 0. \end{equation} (2.29)

    This implies that

    \begin{equation} \mathcal{G}^{ii}\Delta W_{ii}\ge \mathcal{G}^{ij,rs}W_{ij\alpha}W_{rs\alpha}+\mathcal{G}^{ij}\Delta W_{ij} = \Delta\Psi. \end{equation} (2.30)

    Putting (2.30) into (2.27), we have, at the point \xi = \xi_3 ,

    \begin{equation} 0\ge \Delta\Psi-n\Psi+H\Sigma_{i}\mathcal{G}^{ii}. \end{equation} (2.31)

    It follows from Newton-MacLaurin inequality that

    \begin{equation} \Sigma_{i}\mathcal{G}^{ii}\ge 1, \end{equation} (2.32)

    see [26].

    Now, we claim that at the point \xi = \xi_3 ,

    \begin{equation} \frac{\Delta\Psi}{\Psi}\ge \frac{n+1+p}{2n}\min\limits_{\xi\in \mathbb{S}^n}\frac{\varphi^2(\rho(\xi))} {\varphi^2(\rho(\xi))+|\nabla\rho(\xi)|^2}\sum\limits_{k\alpha}\rho^2_{k\alpha} -c. \end{equation} (2.33)

    Indeed, it follows from the definition of \Psi that

    \begin{equation} \begin{split} \log\Psi& = \frac{1}{n}\log \psi(\xi)-\frac{n+2p}{n}\log \varphi(\rho)+\frac{1}{n}f(\rho^2)\\ &\; \; \; +\frac{n+1+p}{2n}\log (\varphi^2(\rho)+|\nabla\rho|^2). \end{split} \end{equation} (2.34)

    For any fixed \alpha\in \{1, 2, \cdots, n\} , taking \alpha -th partial derivatives on both sides of (2.34) twice, we have

    \begin{equation} \begin{split} \frac{\Psi_\alpha}{\Psi}& = (\frac{1}{n}(\log \psi(\xi))' -\frac{n+2p}{n}(\log \varphi)')\rho_\alpha +\frac{2}{n}f'(\rho^2)\rho\rho_\alpha \\ &\; \; \; +\frac{n+1+p}{n}\frac{(\varphi(\rho)\varphi'(\rho)\rho_\alpha +\rho_k\rho_{k\alpha})}{\varphi^2(\rho)+|\nabla\rho|^2} \end{split} \end{equation} (2.35)

    and

    \begin{equation} \begin{split} \frac{\Delta\Psi}{\Psi}-\frac{|\nabla\Psi|^2}{\Psi^2} & = \Sigma_\alpha\frac{\Psi_{\alpha\alpha}}{\Psi}-\frac{\Psi^2_\alpha}{\Psi^2}\\ & = \Sigma_\alpha(\frac{1}{n}(\log\psi)'' -\frac{n+2p}{n}(\log \varphi)'\rho_{\alpha\alpha}\\ &\; \; \; -\frac{n+2p}{n}(\log \varphi)''\rho^2_{\alpha} +\frac{2f'(\rho^2)}{n}(\rho^2_\alpha+\rho\rho_{\alpha\alpha})\\ &\; \; \; +\frac{4f''(\rho^2)}{n}\rho^2\rho^2_\alpha +\frac{n+1+p}{n}\frac{(\varphi(\rho)\varphi''(\rho) +(\varphi'(\rho))^2)\rho^2_\alpha +\varphi(\rho)\varphi'(\rho)\rho_{\alpha\alpha}}{\varphi^2(\rho) +|\nabla\rho|^2}\\ &\; \; \; +\frac{n+1+p}{n}\frac{\sum\limits_k\rho^2_{k\alpha} +\sum\limits_k\rho_{k\alpha\alpha}\rho_k}{\varphi^2(\rho)+|\nabla\rho|^2} \\ &\; \; \; -\frac{n+1+p}{n}\frac{(\varphi(\rho)\varphi'(\rho)\rho_\alpha +\sum\limits_k\rho_{k}\rho_{k\alpha})^2}{(\varphi^2(\rho)+|\nabla\rho|^2)^2}) = \sum^4_{j = 1}I_j. \end{split} \end{equation} (2.36)

    where

    \begin{equation} I_1 = ( -\frac{n+2p}{n}(\log \varphi)'+\frac{n+1+p}{n} \frac{\varphi(\rho)\varphi'(\rho)}{\varphi^2(\rho)+|\nabla\rho|^2} +\frac{2}{n}f'(\rho^2)\rho)\Delta\rho, \end{equation} (2.37)
    \begin{equation} \begin{split} I_2& = (\log \psi)'' +(-\frac{n+2p}{n}(\log \varphi)'' +\frac{2}{n}f'(\rho^2)+\frac{4}{n}f''(\rho^2)\rho^2\\ &\; \; \; +\frac{n+1+p}{n}\frac{(\varphi(\rho)\varphi''(\rho) +(\varphi'(\rho))^2)}{\varphi^2(\rho)+|\nabla\rho|^2}\\ &\; \; \; -\frac{n+1+p}{n}\frac{(\varphi(\rho)\varphi'(\rho))^2} {(\varphi^2(\rho)+|\nabla\rho|^2)^2})|\nabla\rho|^2, \end{split} \end{equation} (2.38)
    \begin{equation} I_3 = -\frac{n+1+p}{n}(\frac{\sum_\alpha(\sum_k\rho_k\rho_{k\alpha})^2} {(\varphi^2(\rho)+|\nabla\rho|^2)^2} +\frac{2\sum_{k\alpha}\rho_k\rho_\alpha\rho_{k\alpha} \varphi(\rho)\varphi'(\rho)} {(\varphi^2(\rho)+|\nabla\rho|^2)^2}) \end{equation} (2.39)

    and

    \begin{equation} I_4 = \frac{n+1+p}{n}(\frac{\sum_{k\alpha}\rho^2_{k\alpha}} {\varphi^2(\rho)+|\nabla\rho|^2}+\frac{\nabla\rho\cdot\nabla\Delta\rho} {\varphi^2(\rho)+|\nabla\rho|^2}). \end{equation} (2.40)

    We first estimate the term I_1 . It follows from Hölder inequality that

    \begin{equation} \begin{split} |I_1|\le c\Delta \rho & = c\Sigma_i\rho_{ii}\\ &\le\frac{(n+1+p)\varepsilon}{n}\Sigma_i\rho^2_{ii}+\frac{c}{2\varepsilon}\\ &\le\frac{(n+1+p)\varepsilon}{n}\Sigma_{k\alpha}\rho^2_{k\alpha} +\frac{c}{2\varepsilon}. \end{split} \end{equation} (2.41)

    for some \varepsilon to be chosen later. Therefore,

    \begin{equation} I_1\ge-\frac{(n+1+p)\varepsilon}{n}\Sigma_{k\alpha}\rho^2_{k\alpha} +\frac{c}{2\varepsilon}. \end{equation} (2.42)

    Now, we turn to the estimate of the term I_2 . It follows from Lemma 2.2 that

    \begin{equation} I_2\ge-c. \end{equation} (2.43)

    Now, we estimate the term I_3 . It follows from Hölder inequality and Lemma 2.2 that

    \begin{equation} \begin{split} |\frac{2\sum\limits_{k\alpha}\rho_k\rho_\alpha \rho_{k\alpha}\varphi(\rho)\varphi'(\rho)} {(\varphi^2(\rho)+|\nabla\rho|^2)^2}| &\le c|\nabla\rho|^2(\sum\limits_{k\alpha}\rho_{k\alpha})^2)^{\frac{1}{2}}\\ &\le c(\sum\limits_{k\alpha}\rho_{k\alpha})^2)^{\frac{1}{2}}\\ &\le\frac{(n+1+p)\varepsilon}{n}\sum\limits_{k\alpha}\rho^2_{k\alpha}+\frac{c}{4\varepsilon} \end{split} \end{equation} (2.44)

    for the same \varepsilon as in (2.41) and

    \begin{equation} \frac{\sum\limits_\alpha(\sum\limits_k\rho_k\rho_{k\alpha})^2} {(\varphi^2(\rho)+|\nabla\rho|^2)^2}\le \frac{|\nabla\rho|^2}{(\varphi^2(\rho)+|\nabla\rho|^2)^2} \sum\limits_{k\alpha}\rho^2_{k\alpha} \end{equation} (2.45)

    Putting (2.44) and (2.45) into (2.36), we have

    \begin{equation} I_3\ge -\frac{n+1+p}{n}(\frac{|\nabla\rho|^2}{(\varphi^2(\rho)+|\nabla\rho|^2)^2} +\varepsilon)\sum\limits_{k\alpha}\rho^2_{k\alpha}-c. \end{equation} (2.46)

    Now, we estimate the term I_4 . Since

    \begin{equation} v_i = -\frac{\rho_i}{\varphi^2(\rho)}, \end{equation} (2.47)

    we have

    \begin{equation} v_{ii} = -\frac{\rho_{ii}}{\varphi^2(\rho)} +\frac{2\varphi'(\rho)}{\varphi^3(\rho)}\rho^2_i \end{equation} (2.48)

    and thus,

    \begin{equation} \Delta \rho = \frac{2\varphi'(\rho)}{\varphi(\rho)}|\nabla\rho|^2 -\varphi^2(\rho)\Delta v. \end{equation} (2.49)

    Moreover, since

    \begin{equation} H = \Delta v+nv, \end{equation} (2.50)

    we have,

    \begin{equation} \Delta \rho = (-H+nv)\varphi^2(\rho)+\frac{2\varphi'(\rho)} {\varphi(\rho)}|\nabla\rho|^2. \end{equation} (2.51)

    Therefore, for any i\in\{1, 2, \cdots, n\} , we have

    \begin{equation} \begin{split} (\Delta \rho)_i& = -H_i\varphi^2(\rho)+n\varphi^2(\rho)v_i -2\varphi(\rho)(H-nv)\varphi'(\rho)\rho_i\\ &\; \; \; +(\frac{2\varphi''(\rho)}{\varphi(\rho)} -\frac{2(\varphi'(\rho))^2}{\varphi^2(\rho)})\rho_i|\nabla\rho|^2 +\frac{2\varphi'(\rho)}{\varphi(\rho)}\sum\limits_l\rho_l\rho_{li}\\ & = -H_i\varphi^2(\rho)-n\rho_i -2\varphi(\rho)(H-nv)\varphi'(\rho)\rho_i\\ &\; \; \; +(\frac{2\varphi''(\rho)}{\varphi(\rho)} -\frac{2(\varphi'(\rho))^2}{\varphi^2(\rho)})\rho_i|\nabla\rho|^2 +\frac{2\varphi'(\rho)}{\varphi(\rho)}\sum\limits_l\rho_l\rho_{li}\\ \end{split} \end{equation} (2.52)

    and

    \begin{equation} \begin{split} \nabla\rho\cdot\nabla\Delta\rho& = -\nabla\rho\cdot\nabla H\varphi^2(\rho)-n|\nabla\rho|^2 -2\varphi(\rho)(H-nv)\varphi'(\rho)|\nabla\rho|^2\\ &\; \; \; +(\frac{2\varphi''(\rho)}{\varphi(\rho)} -\frac{2(\varphi'(\rho))^2}{\varphi^2(\rho)})|\nabla\rho|^4 +\frac{2\varphi'(\rho)}{\varphi(\rho)}\sum\limits_{li}\rho_l\rho_i\rho_{li}\\ & = -n|\nabla\rho|^2 -2\varphi(\rho)(H-nv)\varphi'(\rho)|\nabla\rho|^2\\ &\; \; \; +(\frac{2\varphi''(\rho)}{\varphi(\rho)} -\frac{2(\varphi'(\rho))^2}{\varphi^2(\rho)})|\nabla\rho|^4 +\frac{2\varphi'(\rho)}{\varphi(\rho)}\sum\limits_{li}\rho_l\rho_i\rho_{li} \end{split} \end{equation} (2.53)

    at the point \xi = \xi_3 since \xi_3 is a critical point of H . It follows from Lemma 2.2 that

    \begin{equation} -n|\nabla\rho|^2 -2\varphi(\rho)(H-nv)\varphi'(\rho)|\nabla\rho|^2 +(\frac{2\varphi''(\rho)}{\varphi(\rho)} -\frac{2(\varphi'(\rho))^2}{\varphi(\rho)^2})|\nabla\rho|^4\ge -c \end{equation} (2.54)

    at the point \xi = \xi_3 . It follows from Hölder inequality that

    \begin{equation} \frac{2\varphi'(\rho)}{\varphi(\rho)}\sum\limits_{li}\rho_l\rho_i\rho_{li}\le \frac{(n+1+p)\varepsilon}{n}\sum\limits_{li}\rho^2_{li}+\frac{c}{4\varepsilon} \end{equation} (2.55)

    for the same \varepsilon as in (2.41). Therefore, at the point x = x_3 , we have,

    \begin{equation} \frac{n+2-p}{n}\frac{\nabla\rho\cdot\nabla\Delta\rho} {\varphi^2(\rho)+|\nabla\rho|^2}\ge -\frac{(n+1+p)\varepsilon}{n}\sum\limits_{li}\rho^2_{li}-\frac{c}{4\varepsilon}-c. \end{equation} (2.56)

    Putting (2.56) into (2.40), we have,

    \begin{equation} I_4\ge\frac{n+1+p}{n}(\frac{1} {\varphi^2(\rho)+|\nabla\rho|^2} -\varepsilon)\sum\limits_{li}\rho^2_{li} +\frac{c}{2\varepsilon}-c. \end{equation} (2.57)

    Therefore,

    \begin{equation} \begin{split} \frac{\Delta\Psi}{\Psi} &\ge\sum^4_{j = 1}I_j\\ &\ge\frac{n+1+p}{n}(\frac{1} {\varphi^2(\rho)+|\nabla\rho|^2} -\frac{|\nabla\rho|^2}{(\varphi^2(\rho)+|\nabla\rho|^2)^2} -2\varepsilon)\sum\limits_{li}\rho^2_{li} +\frac{c}{2\varepsilon}-c\\ &\ge\frac{n+1+p}{n}(\frac{\varphi^2(\rho)} {\varphi^2(\rho)+|\nabla\rho|^2}-2\varepsilon) \sum\limits_{li}\rho^2_{li} -\frac{c}{2\varepsilon}-c. \end{split} \end{equation} (2.58)

    Let \varepsilon_0 = \min\limits_{\xi\in \mathbb{S}^n}\frac{\varphi^2(\rho(\xi))} {\varphi^2(\rho(\xi))+|\nabla\rho(\xi)|^2} , for any \varepsilon\in (0, 4\varepsilon_0) , we have,

    \begin{equation} \begin{split} \frac{\Delta\Psi}{\Psi}\ge\frac{n+1+p}{2n}\min\limits_{\xi\in \mathbb{S}^n}\frac{\varphi^2(\rho(\xi))} {\varphi^2(\rho(\xi))+|\nabla\rho(\xi)|^2}\sum\limits_{li}\rho^2_{li} -c. \end{split} \end{equation} (2.59)

    (2.31), (2.32) and (2.58) yields that there exists a positive constant c such that

    \begin{equation} \Sigma_{li}\rho^2_{li}\le c \end{equation} (2.60)

    at the point \xi = \xi_3 . Therefore, it follows from Hölder inequality that

    \begin{equation} -\Delta\rho = -\Sigma_l\rho_{ll}\le \sqrt{n}\sqrt{\Sigma_l\rho^2_{ll}}\le \sqrt{n}\sqrt{\Sigma_{li}\rho^2_{li}}\le c \end{equation} (2.61)

    at the point \xi = \xi_3 . This completes the proof of Lemma 2.4.

    Now, we are in a position the prove Theorem 2.0.

    Final proof of Theorem 2.0. It follows from (2.24) that Eq (2.1) becomes

    \begin{equation} \mathcal{F}(W_{ij}) = 0 \end{equation} (2.62)

    provided \mathcal{F}(W_{ij}) = \mathcal{G}(W_{ij})-\psi . We let \mathcal{F}_{ij} = \frac{\partial \mathcal{F}}{\partial W_{ij}} . It follows from {Lemmas 2.1–2.4} that there exist positive constants \lambda and \Lambda , independent of W_{ij} , such that

    \begin{equation} 0 < \lambda\zeta^2\le \mathcal{F}_{ij}\zeta_i\zeta_j\le \Lambda\zeta^2, \end{equation} (2.63)

    for any \zeta = (\zeta_1, \zeta_2, \cdots, \zeta_n)\in \mathbb{R}^n . That is,

    (i) (2.62) is elliptic uniformly.

    Moreover, it is easy to see that \mathcal{G} = \det^{\frac{1}{n}} is concave with respect to W_{ij} and therefore,

    (ii) \; \mathcal{F} is concave with respect to W_{ij} .

    Then, it follows from {Theorem 17.14} of Gilbarg and Trudinger [20] that there exist \tau_1\in (0, 1) and positive constant c , independent of W , such that

    \begin{equation} \|W\|_{C^{2,\tau_1}(\mathbb{S}^n)}\le c, \end{equation} (2.64)

    and therefore there exist \tau\in (0, 1) and positive constant c , independent of \rho , such that

    \begin{equation} \|\rho\|_{C^{2,\tau}(\mathbb{S}^n)}\le c, \end{equation} (2.65)

    (see pp. 457–461 of Gilbarg and Trudinger [20]). This is the desired conclusion of Theorem 2.0.

    This section devotes to the proof of Theorem 1.1.

    Motivated by [42], we consider the following auxiliary problem with a parameter t\in [0, 1] on the unit sphere \mathbb{S}^n ,

    \begin{equation} M(\rho) = \frac{\det (-\rho_{ij}+\frac{2\varphi'(\rho)}{\varphi(\rho)}\rho_i\rho_j +\varphi(\rho)\varphi'(\rho)\delta_{ij})}{ (\varphi^2(\rho)+|\nabla\rho|^2)^{\frac{n+1+p}{2}}} = t\psi(\xi)K(\rho)+(1-t)g(\rho)\triangleq K_t \end{equation} (3.1)

    where K(\rho) = e^{-f(\rho^2)}\varphi^{n-2p}(\rho) and g(\rho) = \frac{(\varphi'(\rho))^n} {\varphi^{1+p}(\rho)}\rho^{-\gamma} with \gamma > 0 .

    By (A.1.) and the definition of K_t , we have

    \begin{equation} \left\{\begin{array}{ll} \lim\limits_{\rho\to \infty}K_t\frac{\varphi^{1+p}(R)}{(\varphi'(R))^n} = 0;\\ \lim\limits_{\rho\to 0}K_t\frac{\varphi^{1+p}(R)}{(\varphi'(R))^n} = \infty. \end{array} \right. \end{equation} (3.2)

    for any t\in [0, 1] .

    We let the set of the positive continuous function on the unit sphere \mathbb{S}^n be C_+(\mathbb{S}^n) and

    \begin{equation} \mathcal{C} = \{\rho\in C^{2,\sigma}(\mathbb{S}^n): (-\rho_{ij}+\frac{2\varphi'(\rho)}{\varphi(\rho)}\rho_i\rho_j +\varphi(\rho)\varphi'(\rho)\delta_{ij}) \ \text{is positive definite}\} \end{equation} (3.3)

    and

    \begin{equation} \mathcal{I} = \{t\in [0,1]: \rho\in \mathcal{C}\cap C_+(\mathbb{S}^n), \ \text{(3.1) is solvable}\}. \end{equation} (3.4)

    Adopting a similar argument in Section 2, we get

    Lemma 3.1. For any fixed n\ge 1 , p > -n-1 and t\in [0, 1] , we let \rho_t\in \mathcal{C}\cap C_+(\mathbb{S}^n) be a solution of (3.1). Suppose that the condition (A.1.) holds, then there exists a constant c , independent on t , such that

    \begin{equation*} \label{3101} 0 < c^{-1}\le |\rho_t|_{C^{2,\sigma}(\mathbb{S}^n)}\le c, \end{equation*}

    for any t\in [0, 1] and some \sigma\in(0, 1) .

    As a corollary of Lemma 3.1, we have,

    Corollary 3.2. For any fixed n\ge 1 , p > -n-1 and t\in [0, 1] , we let \mathcal{I} is the set defined in (3.4) . Suppose that f , \varphi and \psi satisfy the condition (A.1) . Then \mathcal{I} is closed.

    Proof. It suffices to show that for any sequence \{t_j\}_{j = 1}^\infty\subseteq\mathcal{I} satisfying

    \begin{equation*} \label{3201} t_j\to t_0, \end{equation*}

    as j\to \infty for some t_0\in [0, 1] , we need to prove t_0\in\mathcal{I} .

    We let \rho_j be a solutions of problem (3.1) at t = t_j . It follows from the conclusion of Lemma 3.1 that there exists a positive constant c , independent of j such that

    \begin{equation*} \label{3202} \|\rho_j\|_{C^{2,\sigma}(\mathbb{S}^n)}\le c. \end{equation*}

    By Ascoli-Arzela Theorem, we see that, up to a subsequence, there exists a \rho_0\in C^2(\mathbb{S}^n)

    \begin{equation*} \label{3203} \|\rho_j-\rho_0\|_{C^2(\mathbb{S}^n)}\to 0 \end{equation*}

    as j\to \infty . It is easy to see that

    \begin{equation} M(\rho_j)\to M(\rho_0), K(\rho_j)\to K(\rho_0), g(\rho_j)\to g(\rho_0) \end{equation} (3.5)

    uniformly on \mathbb{S}^n as j\to \infty . Letting j\to \infty , we can see that (t_0, \rho_0) is a solution to the following problem:

    \begin{equation} \frac{\det (-\rho_{ij}+\frac{2\varphi'(\rho)}{\varphi(\rho)}\rho_i\rho_j +\varphi(\rho)\varphi'(\rho)\delta_{ij})}{ (\varphi^2(\rho)+|\nabla\rho|^2)^{\frac{n+1+p}{2}}} = t\psi(\xi)K(\rho)+(1-t)g(\rho) \end{equation} (3.6)

    This implies that t_0\in\mathcal{I} . This is the desired conclusion of Corollary 3.2.

    Lemma 3.3. For any fixed n\ge 1 , p > -n-1 and t\in [0, 1] , we let \mathcal{I} is the set defined in (3.4) . Suppose that f , \varphi and \psi satisfy the conditions (A.1.) , (A.1.) and (A.3.) . Then \mathcal{I} is open.

    Proof. Suppose that there exists a \bar{t}\in \mathcal{I} and a \delta > 0 , for any t_1\in B_\delta(\bar{t})\cap [0, 1] , we need to prove that t_1\in \mathcal{I} . To achieve this goal, joint with Implicit Function Theorem, we need to analyze the kernel of linearized equation associated to (3.1) . We assume that \bar{\rho} is a solution to equation (3.1) at t = \bar{t} . For any \zeta\in \mathbb{S}^n , we let M[\bar{\rho}](\zeta) = \frac{d}{d\varepsilon}M(\bar{\rho}+\varepsilon\zeta)|_{\varepsilon = 0} , K_t[\bar{\rho}](\zeta) = \frac{d}{d\varepsilon}K_t(\bar{\rho}+\varepsilon\zeta)|_{\varepsilon = 0} and

    \begin{equation} G_t(\bar{\rho}) = M(\bar{\rho})-K_t(\bar{\rho}). \end{equation} (3.7)

    It is easy to see that

    \begin{equation} \begin{split} G_t[\bar{\rho}](\zeta) = \frac{d}{d\varepsilon}G_t(\bar{\rho}+\varepsilon\zeta)|_{\varepsilon = 0}& = \frac{d}{d\varepsilon}M(\bar{\rho}+\varepsilon\zeta)|_{\varepsilon = 0} -\frac{d}{d\varepsilon}K_t(\bar{\rho}+\varepsilon\zeta)|_{\varepsilon = 0}\\ & = M[\bar{\rho}](\zeta)-K_t[\bar{\rho}](\zeta). \end{split} \end{equation} (3.8)

    We first calculate M[\rho](\zeta) . Taking logarithm on the left hand side of (3.1) , we get

    \begin{equation} \begin{split} \frac{M[\bar{\rho}](\zeta)}{M(\bar{\rho})}& = \bar{P}_{ij}B(\zeta) -(n+1+p)\frac{\varphi(\bar{\rho})\varphi'(\bar{\rho})\zeta +\nabla\bar{\rho}\cdot\nabla\zeta}{\varphi^2(\bar{\rho})+|\nabla \bar{\rho}|^2} \end{split} \end{equation} (3.9)

    where (\bar{P}_{ij})_{n\times n} is the inverse of the matrix (-\bar{\rho}_{ij} +\frac{2\varphi'(\bar{\rho})}{\varphi(\bar{\rho})} \bar{\rho}_i\bar{\rho}_j +\varphi(\bar{\rho})\varphi'(\bar{\rho})\delta_{ij})_{n\times n} and

    \begin{equation} B(\zeta) = -\zeta_{ij} +\frac{2\varphi'(\bar{\rho})}{\varphi(\bar{\rho})} (\bar{\rho}_i\zeta_j +\bar{\rho}_j\zeta_i) +((\frac{2\varphi'(\bar{\rho})}{\varphi(\bar{\rho})}) '\bar{\rho}_i\bar{\rho}_j +(\varphi(\bar{\rho})\varphi'(\bar{\rho}))'\delta_{ij})\zeta. \end{equation} (3.10)

    We first analyze the term -(n+1+p)\frac{\varphi(\bar{\rho})\varphi'(\bar{\rho})\zeta +\nabla\bar{\rho}\cdot\nabla\zeta}{\varphi^2(\bar{\rho})+|\nabla \bar{\rho}|^2} . We let \zeta = \varphi(\bar{\rho})\eta . Direct Calculation shows that

    \begin{equation} \zeta_i = \varphi(\bar{\rho})\eta_i+\varphi'(\bar{\rho})\bar{\rho}_i\eta \end{equation} (3.11)

    and

    \begin{equation} \zeta_{ij} = \varphi(\bar{\rho})\eta_{ij} +\varphi'(\bar{\rho})(\bar{\rho}_i\eta_j+\bar{\rho}_j\eta_i) +(\varphi''(\bar{\rho})\bar{\rho}_i\bar{\rho}_j +\varphi'(\bar{\rho})\bar{\rho}_{ij})\eta. \end{equation} (3.12)

    We can see that

    \begin{equation} (n+1+p)\frac{\varphi(\bar{\rho})\varphi'(\bar{\rho})\zeta +\nabla\bar{\rho}\cdot\nabla\zeta} {\varphi^2(\bar{\rho})+|\nabla\bar{\rho}|^2} = (n+1+p)\frac{\varphi'(\bar{\rho})}{\varphi(\bar{\rho})}\zeta +(n+1+p)\frac{\varphi(\bar{\rho})\nabla\bar{\rho}\cdot\nabla\eta} {\varphi^2(\bar{\rho})+|\nabla\bar{\rho}|^2}. \end{equation} (3.13)

    This implies that

    \begin{equation} -(n+1+p)\frac{\varphi(\bar{\rho})\varphi'(\bar{\rho})\zeta +\nabla\bar{\rho}\cdot\nabla\zeta} {\varphi^2(\bar{\rho})+|\nabla\bar{\rho}|^2} = -(n+1+p)\frac{\varphi(\bar{\rho})\nabla\bar{\rho}\cdot\nabla \eta}{\varphi^2(\bar{\rho})+|\nabla\bar{\rho}|^2}-(n+1+p)\varphi'(\bar{\rho})\eta. \end{equation} (3.14)

    Now, we move the term \bar{P}_{ij}B(\zeta) . It follows from (3.11) and (3.12) that

    \begin{equation} \begin{split} -\zeta_{ij}+\frac{2\varphi'(\bar{\rho})}{\varphi(\bar{\rho})} (\bar{\rho}_i\zeta_j+\bar{\rho}_j\zeta_i) & = -\varphi(\bar{\rho})\eta_{ij}+\varphi'(\bar{\rho})(\bar{\rho}_i\zeta_j +\bar{\rho}_j\zeta_i)\\ &\; \; \; +(-\varphi'(\bar{\rho})\bar{\rho}_{ij} +(\frac{4(\varphi'(\bar{\rho}))^2}{\varphi(\bar{\rho})} -\varphi''(\bar{\rho})) \bar{\rho}_i\bar{\rho}_j)\eta. \end{split} \end{equation} (3.15)

    Noting \frac{4(\varphi'(\bar{\rho}))^2}{\varphi(\bar{\rho})} -\varphi''(\bar{\rho})+2\varphi(\bar{\rho})(\frac{\varphi'(\bar{\rho})} {\varphi(\bar{\rho})})' = \frac{2(\varphi'(\bar{\rho}))^2}{\varphi(\bar{\rho})}+\varphi''(\bar{\rho}) and

    \varphi(\bar{\rho})(\varphi(\bar{\rho})\varphi'(\bar{\rho}))' = \varphi(\bar{\rho})(\varphi'(\bar{\rho}))^2 +\varphi^2(\bar{\rho})\varphi''(\bar{\rho}),

    we get

    \begin{equation} \begin{split} B(\zeta)& = -\varphi(\bar{\rho})\eta_{ij}+\varphi'(\bar{\rho}_iv_j +\bar{\rho}_j\eta_i)\\ &\; \; \; +(-\varphi'(\bar{\rho})\bar{\rho}_{ij} +(\frac{4(\varphi'(\bar{\rho}))^2}{\varphi(\bar{\rho})} -\varphi''(\bar{\rho}) +2\varphi(\bar{\rho})(\frac{\varphi'(\bar{\rho})}{\varphi(\bar{\rho})})') \bar{\rho}_i\bar{\rho}_j +\varphi(\bar{\rho})(\varphi(\bar{\rho})\varphi'(\bar{\rho}))' \delta_{ij})\eta\\ & = -\varphi(\bar{\rho})\eta_{ij} +\varphi'(\bar{\rho})(\bar{\rho}_i\eta_j+\bar{\rho}_j\eta_i)\\ &\; \; \; +\varphi'(\bar{\rho})(-\bar{\rho}_{ij} +\frac{2\varphi'(\bar{\rho})}{\varphi(\bar{\rho})} \bar{\rho}_i\bar{\rho}_j +\varphi(\bar{\rho})\varphi'(\bar{\rho})\delta_{ij})\eta +\varphi''(\bar{\rho})(\bar{\rho}_i\bar{\rho}_j +\varphi^2(\bar{\rho})\delta_{ij})\eta. \end{split} \end{equation} (3.16)

    Multiplying the matrix (\bar{P}_{ij})_{n\times n} on both sides of (3.16), we get

    \begin{equation} \begin{split} \bar{P}_{ij}B(\zeta) = -\varphi(\bar{\rho})\bar{P}_{ij}\eta_{ij} +2\varphi'(\bar{\rho})\bar{P}_{ij}\bar{\rho}_i\eta_j+n\varphi'(\bar{\rho}) +\varphi''(\bar{\rho})\bar{P}_{ij}(\bar{\rho}_i\bar{\rho}_j+\varphi^2(\bar{\rho}) \delta_{ij})\eta. \end{split} \end{equation} (3.17)

    Putting (3.17) and (3.14) into (3.9), we have,

    \begin{equation} \begin{split} M[\bar{\rho}](\zeta)& = -\varphi(\bar{\rho}) M(\bar{\rho})\bar{P}_{ij}\eta_{ij} -(n+2)\frac{\varphi(\bar{\rho})(\bar{\rho}) M(\bar{\rho})\nabla\bar{\rho}\cdot\nabla\eta}{\varphi^2(\bar{\rho}) +|\nabla \bar{\rho}|^2}+ 2M(\bar{\rho})\varphi'(\bar{\rho})\bar{P}_{ij}\bar{\rho}_i\eta_j\\ &\; \; \; +M(\bar{\rho}) \bar{P}_{ij}\varphi''(\bar{\rho})(\bar{\rho}_i\bar{\rho}_j +\varphi^2(\bar{\rho})\delta_{ij})\eta -(1+p)K_t(\bar{\rho})\varphi'(\bar{\rho})\eta. \end{split} \end{equation} (3.18)

    By the definition of K_t , we have

    \begin{equation} K_t[\bar{\rho}](\zeta) = K_t'(\bar{\rho})\zeta = t\psi(\xi)K'(\bar{\rho})\varphi(\bar{\rho})\eta +(1-t)g'(\bar{\rho})\varphi(\bar{\rho})\eta \end{equation} (3.19)

    Combining (3.19), (3.18) and (3.8), we have

    \begin{equation} \begin{split} \mathcal{L}(\eta) = G'[\bar{\rho}](\zeta) & = -\varphi(\bar{\rho})M(\bar{\rho})\bar{P}_{ij}\eta_{ij} -(n+1+p)\frac{\varphi(\bar{\rho})M(\bar{\rho})\nabla\bar{\rho}\cdot\nabla \eta}{\varphi^2(\bar{\rho})+|\nabla\bar{\rho}|^2} +2M(\bar{\rho})\varphi'(\bar{\rho})\bar{P}_{ij}\bar{\rho}_i\eta_j\\ &\; \; \; +M(\bar{\rho})(\varphi''(\bar{\rho})\bar{P}_{ij}(\bar{\rho}_i\bar{\rho}_j +\varphi^2(\bar{\rho}) \delta_{ij})\\ &\; \; \; -t\psi(\xi)(\varphi(\bar{\rho}) \frac{\partial K}{\partial \bar{\rho}}+(1+p)K\varphi(\bar{\rho})') -(1-t)(g'(\bar{\rho})\varphi(\bar{\rho}) +(1+p)g(\bar{\rho})\varphi'(\bar{\rho})) \eta\\ &\triangleq a_{ij}\eta_{ij}+b_i\eta_i+N\eta, \end{split} \end{equation} (3.20)

    where

    \begin{equation} a_{ij} = -\varphi(\bar{\rho})M(\bar{\rho})\bar{P}_{ij}, \end{equation} (3.21)
    \begin{equation} \begin{split} b_i = -(n+1+p)\frac{\varphi(\bar{\rho})M(\bar{\rho})\bar{\rho}_i} {\varphi^2(\bar{\rho})+|\nabla\bar{\rho}|^2} +2M(\bar{\rho})\varphi'(\bar{\rho})\bar{P}_{ji}\bar{\rho}_i \end{split} \end{equation} (3.22)

    and

    \begin{equation} \begin{split} N = &M(\bar{\rho})(\varphi''(\bar{\rho})\bar{P}_{ij}(\bar{\rho}_i\bar{\rho}_j +\varphi^2(\bar{\rho}) \delta_{ij})\\ &-t\psi(\xi)(\varphi(\bar{\rho}) \frac{\partial K}{\partial \bar{\rho}}+(1+p)K\varphi'(\bar{\rho})) -(1-t)(g'(\bar{\rho})\varphi(\bar{\rho})+(1+p)g(\bar{\rho})\varphi'(\bar{\rho})). \end{split} \end{equation} (3.23)

    Since \varphi(\bar{\rho}), K_t(\bar{\rho}) > 0 , (\bar{P}_{ij})_{n\times n} is positive, we see that a_{ij} is non-positive. It follows from Lemma 3.1 that b_i is bounded. Now, we claim that

    \begin{equation} N > 0. \end{equation} (3.24)

    Indeed, it is easy to see that the matrix (\bar{\rho}_i\bar{\rho}_j+\varphi^2(\bar{\rho})\delta_{ij}) is positive. Since (\bar{P}_{ij}) is positive, we have,

    \begin{equation} \varphi''(\bar{\rho})M(\bar{\rho})\bar{P}_{ij}(\bar{\rho}_i\bar{\rho}_j +\varphi^2(\bar{\rho})\delta_{ij})\ge 0 \end{equation} (3.25)

    provided \varphi''(\bar{\rho})\ge 0 . It follows from assumption (A.2.) and (A.3.) that

    \begin{equation} \varphi(\bar{\rho})K'(\bar{\rho}) +(1+p)K(\bar{\rho})\varphi'(\bar{\rho}) = \varphi^{-p}(\bar{\rho})(K(\bar{\rho})\varphi^{p+1}(\bar{\rho}))' < 0 \end{equation} (3.26)

    and

    \begin{equation} \begin{split} &\varphi(\bar{\rho})g'(\bar{\rho}) +(1+p)g(\bar{\rho})\varphi'(\bar{\rho})\\ & = \varphi^{-p}(\bar{\rho})(g(\bar{\rho}) \varphi^{p+1}(\bar{\rho}))'\\ & = (\varphi'(\bar{\rho}))^{n-1}\varphi^{-p}(\bar{\rho})\bar{\rho}^{-\gamma-1} (n\varphi''(\bar{\rho})\bar{\rho}-\gamma\varphi'(\bar{\rho}))\le 0. \end{split} \end{equation} (3.27)

    noting that \min\limits_{\xi\in \mathbb{S}^{n}}\psi(\xi)\ge 0 , (3.25), (3.26) and (3.28) imply (3.24). By Strong Maximum Principle for elliptic equations of second order, we see that

    \begin{equation} \eta\equiv 0 \end{equation} (3.28)

    (see pp. 35 of Gilbarg and Trudinger [20]) and thus,

    \begin{equation} \zeta\equiv 0 \end{equation} (3.29)

    since \varphi(\bar{\rho}) > 0 . Then by the standard Implicit Function Theorem, for any t\in B_\gamma(\bar{t})\cap [0, 1] , there exists a \rho\in C^{2, \sigma_1}(\mathbb{S}^n) such that G_t(\rho) = 0 for some \sigma_1\in (0, 1) . This means that t\in \mathcal{I} and completes the proof of Lemma 3.3.

    Now, we are in a position to prove Theorem 1.1.

    Final proof of Theorem 1.1. It is easy to see that \rho\equiv 1 is a solution to equation (3.1) at t = 0 , this means that 0\in \mathcal{I} and thus, \mathcal{I} is not empty. Combining this and Corollary 3.2 and Lemma 3.3, we see that \mathcal{I} = [0, 1] . Taking t = 1 , we get the desired conclusion of Theorem 1.1.

    The work was supported by China Postdoctoral Science Foundation (Grant: No.2021M690773) and the author of the present paper would like to thank Prof. Qiuyi Dai and Daomin Cao for their helpful guidance on the Theory of PDEs and useful comments on the first version of the present paper.

    The author declares no conflict of interest.

    In this section, we list some basic geometric quantity which has been used in the present paper and can be referred to [2].

    Lemma A. Suppose M is a hypersurface in \mathbb{R}^{n+1} with the metric ds^2 = d\rho^2+\varphi^2(\rho)d\xi^2 and with zero sectional curvature, then the following statements hold.

    (a) The components of the metric g and its inverse g^{-1} can be expressed as follows:

    \begin{equation} g_{ij} = \varphi^2(\rho)\delta_{ij}+\rho_i\rho_j, g^{ij} = \frac{1}{\varphi^2(\rho)}(\delta^{ij}-\frac{\rho^i\rho^j} {\varphi^2(\rho)+|\nabla \rho|^2}) \end{equation} (A.1)

    respectively and thus, \det(g_{ij}) = \varphi^{2n-2}(\rho)(\varphi^2(\rho)+|\nabla \rho|^2) .

    (b) The coefficients of the second fundamental form b_{ij} is given by:

    \begin{equation} b_{ij} = \frac{\varphi(\rho)}{\sqrt{\varphi^2(\rho)+|\nabla\rho|^2}}(-\rho_{ij} +\frac{2\varphi'(\rho)}{\varphi(\rho)}\rho_i\rho_j +\varphi(\rho)\varphi'(\rho)\delta_{ij}). \end{equation} (A.2)

    (c) The Gaussian curvature \mathcal{K} was given by:

    \begin{equation} \mathcal{K}(\xi) = \frac{\det b_{ij}}{\det g_{ij}} = \frac{\det (-\rho_{ij}+\frac{2\varphi'(\rho)}{\varphi(\rho)}\rho_i\rho_j +\varphi(\rho)\varphi'(\rho)\delta_{ij})} {\varphi^{n-2}(\rho)(\varphi^2(\rho)+|\nabla\rho|^2)^{\frac{n+2}{2}}}. \end{equation} (A.3)


    [1] C. Bär, F. Pfäffle, Wiener measures on Riemannian manifolds and the Feynman-Kac formula, Matematica Contemporanea, 40 (2011), 37–90.
    [2] M. S. Birman, S. Hildebrandt, V. A. Solonnikov, N. N. Uraltseva Nonlinear problems in mathematical physics and related topics I, New York: Springer, 2002. http://doi.org/10.1007/978-1-4615-0777-2
    [3] J. L. M. Barbosa, J. Lira, V. Oliker, J. H. S. de Lira, Uniqueness of starshaped compact hypersurfaces with prescribed m-th mean curvature in hyperbolic space, Illinois J. Math., 51 (2007), 571–582. http://doi.org/10.1215/ijm/1258138430 doi: 10.1215/ijm/1258138430
    [4] V. I. Bogachev, Gaussian measures, American Mathematical Society, 1998.
    [5] C. Borell, The Brunn-Minkowski inequality in Gauss space, Invent. Math., 30 (1975), 207–216. https://doi.org/10.1007/BF01425510 doi: 10.1007/BF01425510
    [6] K. J. Böröczky, E. Lutwak, D. Yang, G. Zhang, The logarithmic Minkowski problem, J. Amer. Math. Soc., 26 (2013), 831–852.
    [7] H. J. Brascamp, E. H. Lieb, On extensions of the Brunn-Minkowski and Prékopa-Leindler theorems, including inequalities for log concave functions, and with an application to the diffusion equation, J. Funct. Anal., 22 (1976), 366–389. http://doi.org/10.1016/0022-1236(76)90004-5 doi: 10.1016/0022-1236(76)90004-5
    [8] Y. D. Burago, V. A. Zalgaller, Geometric inequalities, Berlin: Springer, 1988. https://doi.org/10.1007/978-3-662-07441-1
    [9] D. C. Chang, J. Tie, The sub-Laplacian operators of some model domains, De Gruyter, 2022. https://doi.org/10.1515/9783110642995
    [10] D. J. Chen, H. Z. Li, Z. Z. Wang, Starshaped compact hypersurfaces with prescribed Weingarten curvature in warped product manifolds. Calc. Var., 57 (2018), 1–26. http://doi.org/10.1007/s00526-018-1314-1
    [11] K. S. Chou, X. J. Wang, The L_p-Minkowski problem and the Minkowski problem in centroaffine geometry. Adv. Math., 205 (2006), 33–83. http://doi.org/10.1016/j.aim.2005.07.004
    [12] A. Colesanti, I. Fragalà, The first variation of the total mass of log-concave functions and related inequalities, Adv. Math., 244 (2013), 708–749. http://doi.org/10.1016/j.aim.2013.05.015 doi: 10.1016/j.aim.2013.05.015
    [13] M. Émery, Stochastic calculus in manifolds, Berlin: Springer, 1989. http://doi.org/10.1007/978-3-642-75051-9
    [14] C. Eric, M. Mokshay, M. W. Elisabeth, Convexity and concentration, New York: Springer, 2017. http://doi.org/10.1007/978-1-4939-7005-6
    [15] N. F. Fang, S. D. Xing, D. P. Ye, Geometry of log-concave functions: the L_p Asplund sum and the L_p Minkowski problem, Calc. Var., 61 (2022), 1–37. http://doi.org/10.1007/s00526-021-02155-7 doi: 10.1007/s00526-021-02155-7
    [16] W. J. Firey, p-means of convex bodies, Math. Scand., 10 (1962), 17–24. http://doi.org/10.7146/math.scand.a-10510 doi: 10.7146/math.scand.a-10510
    [17] J. Fröhlich, A. Knowles, B. Schlein, V. Sohinger, Gibbs measures of nonlinear Schrödinger equations as limits of many-body quantum states in dimensions d\le 3, Commun. Math. Phys., 356 (2017), 883–980. http://doi.org/10.1007/s00220-017-2994-7 doi: 10.1007/s00220-017-2994-7
    [18] R. J. Gardner, A. Zvavitch, Gaussian Brunn-Minkowski inequalities, Trans. Amer. Math. Soc., 362 (2010), 5333–5353. http://doi.org/10.1090/S0002-9947-2010-04891-3 doi: 10.1090/S0002-9947-2010-04891-3
    [19] H. T. Georgii, Gibbs measures and phase transitions, De Gruyter, 2011. http://doi.org/10.1515/9783110250329
    [20] D. Gilbarg, N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Berlin: Springer, 2001. https://doi.org/10.1007/978-3-642-61798-0
    [21] P. M. Gruber, J. M. Wills, Convexity and its applications, Birkhäuser Basel: Springer, 1983. https://doi.org/10.1007/978-3-0348-5858-8
    [22] B. Guan, P. F. Guan, Convex hypersurfaces of prescribed curvatures. Ann. Math., 156 (2002), 655–673. http://doi.org/10.2307/3597202
    [23] P. F. Guan, J. f. Li, A mean curvature type flow in space forms. Int. Math. Res. Notices, 2015 (2015), 4716–4740. http://doi.org/10.1093/imrn/rnu081
    [24] P. F. Guan, C. S. Lin, X. N. Ma, The Christoffel-Minkowski problem II: weingarten curvature equations, Chinese Ann. Math. B, 27 (2006), 595–614. http://doi.org/10.1007/s11401-005-0575-0 doi: 10.1007/s11401-005-0575-0
    [25] P. F. Guan, J. F. Li, M. T. Wang, A volume preserving flow and the isoperimetric problem in warped product spaces, Trans. Amer. Math. Soc., 372 (2019), 2777–2798. http://doi.org/10.1090/tran/7661 doi: 10.1090/tran/7661
    [26] P. F. Guan, X. N. Ma, The Christoffel-Minkowski problem I: convexity of solutions of a Hessian equation, Invent. Math., 151 (2003), 553–577. http://doi.org/10.1007/s00222-002-0259-2 doi: 10.1007/s00222-002-0259-2
    [27] P. F. Guan, J. F. Li, Y. Y. Li, Hypersurfaces of prescribed curvature measure, Duke Math. J., 161 (2012), 1927–1942. http://doi.org/10.1215/00127094-1645550 doi: 10.1215/00127094-1645550
    [28] P. F. Guan, C. Y. Ren, Z. Z. Wang, Global C^2-estimates for convex solutions of curvature equations, Commun. Pure Appl. Math., 68 (2015), 1287–1325. http://doi.org/10.1002/cpa.21528 doi: 10.1002/cpa.21528
    [29] B. Güneysu, Covariant schrödinger semigroups on riemannian manifolds, Birkhäuser Cham: Springer, 2017. https://doi.org/10.1007/978-3-319-68903-6
    [30] Y. Huang, D. M. Xi, Y. M. Zhao, The Minkowski problem in Gaussian probability space, Adv. Math., 385 (2021), 107769. http://doi.org/10.1016/j.aim.2021.107769 doi: 10.1016/j.aim.2021.107769
    [31] Y. Huang, E. Lutwak, D. Yang, G. Y. Zhang, Geometric measures in the dual Brunn-Minkowski theory and their associated Minkowski problems. Acta Math., 216 (2016), 325–388. http://doi.org/10.1007/s11511-016-0140-6
    [32] Y. Huang, Y. M. Zhao, On the L_p dual Minkowski problem. Adv. Math., 332 (2018), 57–84. http://doi.org/10.1016/j.aim.2018.05.002
    [33] Q. N. Jin, Y. Y. Li, Starshaped compact hypersurfaces with prescribed k-th mean curvature in hyperbolic space, Discrete Cont. Dyn-A., 15 (2006), 367–377. http://doi.org/10.3934/dcds.2006.15.367 doi: 10.3934/dcds.2006.15.367
    [34] B. Klartag, V. D. Milman, Geometry of log-concave functions and measures, Geom. Dedicata, 112 (2005), 169–182. http://doi.org/10.1007/s10711-004-2462-3 doi: 10.1007/s10711-004-2462-3
    [35] M. Lewin, P. T. Nam, N. Rougerie, Gibbs measures based on 1d (an) harmonic oscillators as mean-field limits, J. Math. Phys., 59 (2018), 041901. http://doi.org/10.1063/1.5026963 doi: 10.1063/1.5026963
    [36] C. H. Li, Z. Z. Wang, The Weyl problem in warped product spaces, J. Differ. Geom., 114 (2020), 243–304. http://doi.org/10.4310/jdg/1580526016 doi: 10.4310/jdg/1580526016
    [37] Q. R. Li, W. M. Sheng, Closed hypersurfaces with prescribed Weingarten curvature in Riemannian manifolds, Calc. Var., 48 (2013), 41–66. http://doi.org/10.1007/s00526-012-0540-1 doi: 10.1007/s00526-012-0540-1
    [38] J. Q. Liu, The L_p-Gaussian Minkowski problem, Calc. Var., 61 (2022), 28. http://doi.org/10.1007/s00526-021-02141-z doi: 10.1007/s00526-021-02141-z
    [39] E. Lutwak, The Brunn-Minkowski-Firey theory. I. Mixed volumes and the minkowski problem, J. Differ. Geom., 38 (1993), 131–150. http://doi.org/10.4310/jdg/1214454097 doi: 10.4310/jdg/1214454097
    [40] E. Lutwak, V. Oliker, On the regularity of solutions to a generalization of the Minkowski problem, J. Differ. Geom., 41 (1995), 227–246.
    [41] E. Lutwak, D. Yang, G. Zhang, On the L_p-Minkowski problem, Trans. Amer. Math. Soc., 356 (2004), 4359–4370.
    [42] V. I. Oliker, Hypersurfaces in \mathbb{R}^{n+1} with prescribed Gaussian curvature and related equations of Monge-Ampère type. Commun. Part. Diff. Eq., 9 (1984), 807–838. http://doi.org/10.1080/03605308408820348
    [43] L. Rotem, Surface area measures of log-concave functions, Journal d'Analyse Mathématique, 147 (2022), 373–400. http://doi.org/10.1007/s11854-022-0227-2
    [44] R. Schneider, Convex bodies: The Brunn-Minkowski theory. Cambridge University Press, 2014. https://doi.org/10.1017/CBO9781139003858
    [45] D. W. Stroock, An introduction to the analysis of paths on a Riemannian manifold. American Mathematical Society, 2000.
    [46] Z. N. Sui, Strictly locally convex hypersurfaces with prescribed curvature and boundary in space forms, Commun. Part. Diff. Eq., 45 (2020), 253–283. http://doi.org/10.1080/03605302.2019.1670675 doi: 10.1080/03605302.2019.1670675
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1363) PDF downloads(63) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog