Loading [MathJax]/jax/output/SVG/jax.js
Research article

Some rigidity theorems on Finsler manifolds

  • Received: 30 October 2020 Accepted: 05 January 2021 Published: 11 January 2021
  • MSC : 53C24, 53C60

  • We prove that, for a Finsler manifold with the weighted Ricci curvature bounded below by a positive number, it is a Finsler sphere if and only if the diam attains its maximal value, if and only if the volume attains its maximal value, and if and only if the first closed eigenvalue of the Finsler-Laplacian attains its lower bound. These generalize some rigidity theorems in Riemannian geometry to the Finsler setting.

    Citation: Songting Yin. Some rigidity theorems on Finsler manifolds[J]. AIMS Mathematics, 2021, 6(3): 3025-3036. doi: 10.3934/math.2021184

    Related Papers:

    [1] Rattanasak Hama, Sorin V. Sabau . Randers metrics on two-spheres of revolution with simple cut locus. AIMS Mathematics, 2023, 8(11): 26213-26236. doi: 10.3934/math.20231337
    [2] Yanlin Li, Yuquan Xie, Manish Kumar Gupta, Suman Sharma . On projective Ricci curvature of cubic metrics. AIMS Mathematics, 2025, 10(5): 11305-11315. doi: 10.3934/math.2025513
    [3] Guangyue Huang, Botao Wang . Rigidity results for closed vacuum static spaces. AIMS Mathematics, 2023, 8(12): 28728-28737. doi: 10.3934/math.20231470
    [4] Xiaoling Zhang, Cuiling Ma, Lili Zhao . On some m-th root metrics. AIMS Mathematics, 2024, 9(9): 23971-23978. doi: 10.3934/math.20241165
    [5] 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
    [6] 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
    [7] Yanlin Li, Dipen Ganguly, Santu Dey, Arindam Bhattacharyya . Conformal η-Ricci solitons within the framework of indefinite Kenmotsu manifolds. AIMS Mathematics, 2022, 7(4): 5408-5430. doi: 10.3934/math.2022300
    [8] Rajesh Kumar, Sameh Shenawy, Lalnunenga Colney, Nasser Bin Turki . Certain results on tangent bundle endowed with generalized Tanaka Webster connection (GTWC) on Kenmotsu manifolds. AIMS Mathematics, 2024, 9(11): 30364-30383. doi: 10.3934/math.20241465
    [9] Aliya Naaz Siddiqui, Mohammad Hasan Shahid, Jae Won Lee . On Ricci curvature of submanifolds in statistical manifolds of constant (quasi-constant) curvature. AIMS Mathematics, 2020, 5(4): 3495-3509. doi: 10.3934/math.2020227
    [10] Noura Alhouiti, Fatemah Mofarreh, Fatemah Abdullah Alghamdi, Akram Ali, Piscoran-Ioan Laurian . Geometric topology of CR-warped products in six-dimensional sphere. AIMS Mathematics, 2024, 9(9): 25114-25126. doi: 10.3934/math.20241224
  • We prove that, for a Finsler manifold with the weighted Ricci curvature bounded below by a positive number, it is a Finsler sphere if and only if the diam attains its maximal value, if and only if the volume attains its maximal value, and if and only if the first closed eigenvalue of the Finsler-Laplacian attains its lower bound. These generalize some rigidity theorems in Riemannian geometry to the Finsler setting.



    In Riemannian geometry, Myers's theorem proves that if (M,g) is a complete connected Riemannian n-manifold such that Ric(n1)k for some positive number k, then it is compact and Diam(M)πk. Further, Cheng's maximum diam theorem states that if its diameter attains its maximal value, then the manifold is isometric to the Euclidean sphere Sn(1k). Under the same curvature condition, Lichnerowicz estimate indicates that the first closed eigenvalue of the Laplacian is not less than nk, while Obata rigidity theorem shows if the first closed eigenvalue attains its lower bound, then it is isometric to Sn(1k). By the same token, Bishop-Gromov comparison theorem also demonstrates that if Vol(M)=Vol(Sn(1k)), then it is isometric to Sn(1k). Therefore, on a complete connected Riemannian n-manifold with Ric(n1)k for some positive k, the following conditions are equivalent:

    (M,g) is the Euclidean sphere Sn(1k);

    Diam(M)=πk;

    ● the first eigenvalue of Laplacian is λ1(M)=nk;

    Vol(M)=Vol(Sn(1k)).

    It is natural to ask:

    In the Finsler setting, can we also characterize and determine Finsler spheres by the diam, the first eigenvalue and the volume?

    In Finsler geometry, Cheng's maximum diam theorem is studied in [2] for reversible Finsler manifolds and in [10] for the general case. Lichnerowicz estimate and Obata rigidity theorem in Finsler situation are also considered in [8,9,10]. Along this line, the authors give a positive answer to the problem above. As to Bishop-Gromov comparison theorem, there are several generalized results established in [4,6,7,11], respectively. However, they (see [4,6,7]) do not study the rigidity phenomenon when the volume reaches its maximum value. In [11], Zhao-Shen study the rigidity problem and give some characterizations by using constant radial flag curvature and constant radial S curvature. Yet, there is still much to be desired.

    Let Sn(1k) denote a Finsler sphere which has Busemann Hausdorff volume form, constant flag curvature k and vanishing S curvature. When the Finsler metric is a Randers metric, the sphere is called a Randers sphere denoted by Sn(1k). There are infinitely many Randers spheres, and if the metric is reversible, the sphere is just the Euclidean sphere (see [10]). For more details, we refer to Section 2 below.

    In this paper, we characterize Finsler spheres and obtain some rigidity results in the following.

    Theorem 1.1. Let (M,F,dμ) be a complete connected Finsler n-manifold with Busemann-Hausdorff volume form. If the weighted Ricci curvature Ricn(n1)k>0, then the following conditions are equivalent:

    (1) (M,F,dμ) is a Finsler sphere Sn(1k);

    (2) Diam(M)=πk;

    (3) the first eigenvalue of Finsler-Laplacian is λ1=nk;

    (4) voldμF(M)=vol(Sn(1k)).

    Since the classification for Finsler metrics with constant flag curvature is not solved, we can not determine all Finsler sphere metrics. However, for a Randers sphere Sn(1k), the metric F can be expressed by navigation data (g,W), where g is the standard sphere metric and W is a Killing vector. Therefore, Randers spheres are of most importance among Finsler spheres. By narrowing the scope in Theorem 1.1, we have

    Theorem 1.2. Let (M,F,dμ) be a complete connected Randers n-manifold with Busemann-Hausdorff volume form. If the weighted Ricci curvature Ricn(n1)k>0, then the following conditions are equivalent:

    (1) (M,F,dμ) is a Randers sphere Sn(1k);

    (2) Diam(M)=πk;

    (3) the first eigenvalue of Finsler-Laplacian is λ1=nk;

    (4) voldμF(M)=vol(Sn(1k)).

    Remark 1.3. Theorems 1.1 and 1.2 show that, apart from the Euclid sphere, the maximum diameter, the maximum volume and the lower bound of the first eigenvalue can be attained on countless Finsler (especially Randers) spheres.

    The paper is organized as follows. In Section 2, some fundamental concepts and formulas which are necessary for the present paper are given, and some lemmas are contained. The volume comparison theorem and Theorem 1.1 are then proved in Sections 3 and 4, repectively.

    Let M be an n-dimensional smooth manifold and π:TMM be the natural projection from the tangent bundle TM. Let (x,y) be a point of TM with xM, yTxM, and let (xi,yi) be the local coordinates on TM with y=yi/xi. A Finsler metric on M is a function F:TM[0,+) satisfying the following properties:

    (ⅰ) Regularity: F(x,y) is smooth in TM0;

    (ⅱ) Positive homogeneity: F(x,λy)=λF(x,y) for λ>0;

    (ⅲ) Strong convexity: The fundamental quadratic form

    g:=gij(x,y)dxidxj,gij:=12[F2]yiyj

    is positively definite.

    Let X=Xixi be a differentiable vector field. Then the covariant derivative of X by vTxM with reference vector wTxM0 is defined by

    DwvX(x):={vjXixj(x)+Γijk(w)vjXk(x)}xi,

    where Γijk denote the coefficients of the Chern connection.

    Given two linearly independent vectors V,WTxM0, the flag curvature is defined by

    K(V,W):=gV(RV(V,W)W,V)gV(V,V)gV(W,W)gV(V,W)2,

    where RV is the Chern curvature:

    RV(X,Y)Z=DVXDVYZDVYDVXZDV[X,Y]Z.

    Then the Ricci curvature for (M,F) is defined as

    Ric(V)=n1α=1K(V,eα),

    where e1,,en1,VF(V) form an orthonormal basis of TxM with respect to gV.

    Let (M,F,dμ) be a Finsler n-manifold with dμ=σ(x)dx1dxn. The distortion is given by

    τ(x,V)=logdet(gij(x,V))σ(x).

    For the vector VTxM, let γ:(ε,ε)M be a geodesic with γ(0)=x,˙γ(0)=V. Then the S-curvature measures the rate of changes of the distortion along geodesics

    S(x,V):=ddt[τ(γ(t),˙γ(t))]t=0.

    Define

    ˙S(x,V):=F2(V)ddt[S(γ(t),˙γ(t))]t=0.

    Then the weighted Ricci curvature of (M,F,dμ) is defined by (see [3])

    {Ricn(V):={Ric(V)+˙S(V),forS(V)=0,,otherwise,RicN(V):=Ric(V)+˙S(V)S(V)2(Nn)F(V)2,N(n,),Ric(V):=Ric(V)+˙S(V),

    For a smooth function u, the gradient vector of u at x is defined by u(x):=L1(du), where L:TxMTxM is the Legendre transform. Let V=Vixi be a smooth vector field on M. The divergence of V with respect to an arbitrary volume form dμ is defined by

    divV:=ni=1(Vixi+Vilogσ(x)xi).

    Then the Finsler-Laplacian of u can be defined by

    Δu:=div(u).

    The equality is in the weak W1,2(M) sense. Namely, for any φC0(M), we have

    MφΔudμ=Mdφ(u)dμ.

    Recall that Bao-Shen [1] found a family Randers sphere metrics on S3, which shows that the maximum diam can be achieved in non-Riemannian case.

    Example 2.1. [1] View S3 as a compact Lie group. Let ζ1,ζ2,ζ3 be the standard right invariant 1-form on S3 satisfying

    dζ1=2ζ2ζ3,dζ2=2ζ3ζ1,dζ3=2ζ1ζ2.

    For k1, define

    αk(y)=(kζ1(y))2+k(ζ2(y))2+k(ζ3(y))2,βk(y)=k2kζ1(y).

    Then Fk=αk+βk is a Randers metric on S3 satisfying

    K1,S0,Diam(S3,Fk)=π.

    Inspired by Bao-Shen's example, we give the following definition.

    Definition 2.2. A Finsler manifold with Busemann-Hausdorff volume form is said to be a Finsler sphere if it has positively constant flag curvature and vanishing S-curvature. In particular, if the metric is a Randers metric, we call it a Randers sphere.

    Generally, we are not able to identify a Finsler sphere metric. However, for a Randers sphere metric, F can be expressed by (see [10])

    F=λg2+W20λW0λ,λ=1W2g,

    where g is the standard sphere metric, and W is a Killing vector field on Sn(1k).

    To prove our results, we further introduce the following lemmas.

    Lemma 2.3. [4] Let (M,F,dμ) be a Finsler n manifold. If the weighted Ricci curvature satisfies RicN(N1)k,N[n,), then the Laplacian of the distance function r(x)=dF(p,x) from any given point pM can be estimated as follows:

    Δr(N1)sk(r)sk(r),

    pointwise on M({p}Cut(p)) and in the sense of distributions on M{p}, where

    sk={1ksin(kt),k>0;t,k=0;1ksinh(kt),k<0.

    Lemma 2.4. [3] Let (M,F,dμ) be a Finsler n manifold. If the weighted Ricci curvature satisfies RicN(N1)k,N[n,), then for any 0<r<R (πk if k>0), it holds that

    max{voldμFB+x(R)voldμFB+x(r),voldμFBx(R)voldμFBx(r)}R0sN1kdtr0sN1kdt.

    In [6,7], the volume comparison theorems are established on Finsler manifolds satisfying Ric(n1)k and some S curvature condition. Later, Zhao-Shen [11] generalize them and further obtain the rigidity result. Using the weighted Ricci curvature condition, Ohta [3] obtain another version of the relative volume comparison (see Lemma 2.4 above). However, the problem about the rigidity result remains open. The main obstacle is that, for any volume form, the limit

    limr0voldμF(B+p(r))r0(sk(t))N1dt

    does not necessarily exists. Therefore, to obtain the rigidity result, it is suitable to give a restriction on the volume form.

    Let (M,F,dμ) be a Finsler n-manifold. Fix a point pM. Then on TpM the Finsler metric F(p,y) induces a Riemannian metric gp(y):=gij(p,y)dyidyj, which also induces a Riemannian metric ˙gp on SpM:={y|yTpM,F(y)=1}. Let (r,θ) be the polar coordinate around p and write the volume form by dμ=σp(r,θ)drdθ. Then we can give the volume comparison theorem as follows:

    Theorem 3.1. Let (M,F,dμ) be a forward complete Finsler n-manifold with arbitrary volume form. If the weighted Ricci curvature Ricn(n1)k, then for some positive number Cp:=limr0SpMσp(r,θ)rn1dθ,

    voldμF(B+p(r))Cpr0(sk(t))n1dt,0rip, (3.1)

    where B+p(r) denotes the forward geodesic ball centered at p of radius r, and ip is the cut value of p. Moreover, the equality holds for r0>0 if and only if for ySpM,

    K(˙γy(t);)=k,0tr0ip,

    where γy(t) is the geodesic satisfying γy(0)=p,˙γy(0)=y. In this case, under the polar coordinate (r,θ) of p, we have

    g(r|(r,θ))=drdr+s2k˙gp(θ),

    where ˙gp denotes the Riemannian metric on SpM.

    Proof. From Lemma 2.4, we have

    voldμF(B+p(r))r0(sk(t))n1dt

    is monotone decreasing on r. By

    limr0voldμF(B+p(r))r0(sk(t))n1dt=limr0r0SpMσp(r,θ)dθdrr0(sk(t))n1dt=limr0SpMσp(r,θ)rn1dθ=Cp,

    we obtain (3.1).

    Let r(x)=dF(p,x) be the distance function from p. Since Ricn(n1)k, by Laplacian comparison theorem (Lemma 2.3), we have

    Δr(n1)sk(r)sk(r), (3.2)

    which yields

    Δr=rlogσprlogsk(r)n1:=rlog˜σ,

    where ˜σ:=sk(r)n1. Define f(r)=σp(r,θ)˜σ(r). Then

    f(r)=σp˜σσp˜σ˜σ2=σp˜σr(logσplog˜σ)0.

    Hence, f(r) is monotone decreasing on r. As a result,

    σp(R,θ)˜σ(R)σp(r,θ)˜σ(r),rR.

    Assume that the equality holds in (3.1). That is,

    r0SpMσp(ρ,θ)dθdρ=Cpr0˜σ(ρ)dρ,rip.

    Differentiating it with respect to r on both sides gives

    SpMσp(r,θ)˜σ(r)dθ=Cp,rip.

    By the monotonicity of f(r), we deduce that

    Cp=SpMσp(R,θ)˜σ(R)dθSpMσp(r,θ)˜σ(r)dθ=Cp,rRip,

    which implies

    σp(R,θ)˜σ(R)=σp(r,θ)˜σ(r),rRip.

    Thus the equality holds in (3.2), which gives

    r(Δr)+(Δr)2n1=(n1)k. (3.3)

    Let Sp(r(x)) be the forward geodesic sphere of radius r(x) centered at p. Choosing the local gr-orthonormal frame E1,,En1 of Sp(r(x)) near x, we get local vector fields E1,,En1,En=r by parallel transport along geodesic rays. Thus, it follows from [7] that

    rtrrH(r)=Ric(r)i,j[H(r)(Ei,Ej)]2, (3.4)

    where H(r) is the Hessian of the distance function r. On the other hand, we also have (see [7])

    Δr=trrH(r)S(r)=trrH(r). (3.5)

    Therefore, by (3.4) and (3.5), we obtain

    (n1)k=r(Δr)+(Δr)2n1=r(trrH(r))+1n1(trrH(r))2rtrrH(r)+i,j[H(r)(Ei,Ej)]2=Ric(r)=Ricn(r)(n1)k. (3.6)

    It follows from (3.6) that

    i,j[H(r)(Ei,Ej)]2=1n1(trrH(r))2,

    which means

    2r(Ei,Ej):=H(r)(Ei,Ej)={trrH(r)n1=Δrn1=ctk(r),i=j<n,0,ij, (3.7)

    where ctk(r):=sk(r)sk(r). Next we shall compute the flag curvature. From (3.7), we know that {Ei}n1i=1 are (n1) eigenvectors of 2r. That is,

    DrEir=ctk(r)Ei,i=1,,n1.

    Noticed that r is a geodesic field of (M,F). Therefore, the flag curvature K(r;) equals to the sectional curvature of the weighted Riemannian manifold (M,gr). Note that {Ei}n1i=1 are (n1) eigenvectors of 2r and parallel along the geodesic ray. By a straightforward computation, we get, for 1in1,

    K(r;Ei)=Rr(Ei,r,Ei,r)=gr(Rr(Ei,r)r,Ei)=gr(DrEiDrrrDrrDrEirDr[Ei,r]r,Ei)=gr(Drr(ctk(r)Ei)+DrDrEirDrrEir,Ei)=gr(ctk(r)Ei+Drctk(r)Eir,Ei)=ctk(r)ctk(r)gr(DrEir,Ei)=ctk(r)ctk(r)2=k.

    We are now to prove that if K(˙γy(t);)=k, then the equality holds in (3.1). Under the polar coordinate of p, we have (r,θ)=(r(q),θ1(q),,θn1(q)) for qDp{p}, where

    r(q)=F(y),θα(q)=ˉθα(yF(y)),y=exp1p(q).

    Then

    θα|q=(dexpp)y(rˉθα).

    So, for ySpM, θα can be viewed as a Jacobi field on γy(t), and

    limr01rθα|q=ˉθα|y.

    Since K(r;)=k, Jα(t)=θα|(t,y)=sk(t)Eα(t) is the Jacobi field satisfying J(0)=0, where Eα(t) is a parallel vector field on γy(t), and Eα(0)=ˉθα|y. By Gauss lemma, gr(r,θα)=0. Therefore,

    g(r|(r,θ))=drdr+s2k˙gp(θ).

    Since S(˙γy(t))=0, we have

    0=ddtτ(t)=ddtlogdetgij(t)σp(t)=ddtlogsk(t)n1det(˙gp(θ)αβ)σp(t)=ddtlogsk(t)n1σp(t)=ddtlog˜σ(t)σp(t),

    which means that

    σp(R,θ)˜σ(R)=σp(r,θ)˜σ(r),0rRip.

    Thus, we obtain

    R0SpMσp(R,θ)dθdR=R0SpM˜σ(R)σp(r,θ)˜σ(r)dθdR=R0˜σ(R)dRSpMσp(r,θ)˜σ(r)dθ=R0˜σ(R)dRlimr0SpMσp(r,θ)˜σ(r)dθ=CpR0˜σ(R)dR,0Rip.

    That is,

    voldμF(B+(r))=Cpr0(sk(t))n1dt,0rip.

    From Theorem 3.1, it is easy to obtain the following:

    Corollary 3.2. Let (M,F,dμ) be a forward complete Finsler n-manifold with arbitrary volume form. If the weighted Ricci curvature Ricn(n1)k, and there exists some positive number C such that limr0SpMσp(r,θ)rn1dθ=C for pM, then

    voldμF(B+p(r))=Cr0(sk(t))n1dt,0rip,pM

    if and only if Kk.

    Recall that, for a Randers sphere (Sn(1k)), the metric F is expressed by (see [10])

    F=λg2+W20λW0λ,λ=1W2g,

    where g is the standard sphere metric, W is a Killing vector field on Sn(1k). Moreover, we have (see [10])

    voldμF(Sn(1k))=volg(Sn(1k));Diam(Sn(1k),F)=πk.

    In what follows, we show that the properties above still hold for a general Finsler sphere.

    Proposition 4.1. On a Finsler sphere Sn(1k), we have

    (1) voldμF(Sn(1k))=volg(Sn(1k));

    (2) Diam(Sn(1k))=πk.

    Proof. Since dμ is the Busemann-Hausdorff volume form, the constant Cp in (3.1) is

    Cp=limr0SpMσp(r,θ)rn1dθ=limr0voldμF(B+p(r))r0(sk(t))n1dt=vol(Sn1).

    Thus, from Corollary 3.2, we have

    voldμF(Sn(1k))=volg(Sn(1k))

    Now fix pSn(1k). Using K=k and Theorem 3.1 in [5], there exists qSn(1k) such that

    expp(πkξ)=q,ξSp(Sn(1k)),

    where Sp(Sn(1k)):={v|vTp(Sn(1k)),F(v)=1}. From the proof of the volume comparison theorem (Theorem 3.1),

    voldμF(B+p(r))σn(r),

    where σn(r) denotes the volume of the metric ball of radius r in Sn(1k). The equality holds if and only if B+p(r)Dp, i.e., ipr. By the Bonnet-Myers theorem, Diam(Sn(1k))πk, which means ¯B+p(πk)=Sn(1k). Therefore,

    voldμF(B+p(πk))=voldμF(Sn(1k))=volg(Sn(1k))=σn(πk).

    We deduce that ipπk, which yields dF(p,q)=πk.

    Theorem 4.2. Let (M,F,dμ) be a complete connected Finsler n-manifold with Busemann-Hausdorff volume form. If the weighted Ricci curvature Ricn(n1)k>0, and voldμF(M)=volg(Sn(1k)), then (M,F) is isometric to a Finsler sphere.

    Proof. Since dμ is the Busemann-Hausdorff volume form, the constant Cp in (3.1) is Cp=vol(Sn1). Then it follows Theorem 3.1 that, for any pM,

    voldμF(B+p(r))vol(Sn1)r0(sk(t))n1dt:=σn(r),

    where σn(r) denotes the volume of the metric ball of radius r in Sn(1k). The equality holds if and only if B+p(r)Dp. That is, ipr. Since

    voldμF(M)=volg(Sn(1k)),

    we deduce that, for pM and rip,

    voldμF(B+p(r))=σn(r).

    Then, from Corollary 3.2, we have

    Kk.

    This completes the proof.

    To prove Theorem 1.1, we further need the following theorems.

    Theorem 4.3. [10] Let (M,F,dμ) be a complete connected Finsler n-manifold with the Busemann-Hausdorff volume form. If the weighted Ricci curvature satisfies Ricn(n1)k>0, and Diam(M)=πk, then (M,F) is isometric to a Finsler sphere.

    Theorem 4.4. [10] Let (M,F,dμ) be a complete connected Finsler n-manifold with the Busemann-Hausdorff volume form. If the weighted Ricci curvature satisfies Ricn(n1)k>0, then the first eigenvalue of Finsler-Laplacian λ1=nk if and only if (M,F) is isometric to a Finsler sphere.

    Proof of Theorem 1.1.

    It follows from Proposition 4.1, Theorems 4.2–4.4 directly.

    From the proof of Theorem 3.1, we know that the key step is

    Δr=(n1)sk(r)sk(r).

    Therefore, we get another equivalent condition.

    Theorem 4.5. Let (M,F,dμ) be a complete connected Finsler n-manifold with Busemann-Hausdorff volume form. If the weighted Ricci curvature Ricn(n1)k>0, then the following conditions are equivalent:

    (1) (M,F,dμ) is a Finsler sphere Sn(1k);

    (2) Δr=(n1)cotr for any distance function r(x)=dF(p,x),pM.

    This work was supported by National Natural Science Foundation of China (Grant No. 11971253).

    The author declares that they have no conflicts of interest.



    [1] D. Bao, Z. Shen, Finsler metrics of constant cuevature on the Lie group S3, J. Lond. Math. Soc., 66 (2002), 453–467. doi: 10.1112/S0024610702003344
    [2] C. Kim, J. Yim, Finsler manifolds with positive constant flag curvature, Geometriae Dedicata, 98 (2003), 47–56. doi: 10.1023/A:1024034012734
    [3] S. Ohta, Finsler interpolation inequalities, Calc. Var. Partial Dif., 36 (2009), 211–249. doi: 10.1007/s00526-009-0227-4
    [4] S. Ohta, K. T. Sturm, Heat Flow on Finsler Manifolds, Commun. Pur. Appl. Math., 62 (2009), 1386–1433. doi: 10.1002/cpa.20273
    [5] Z. Shen, Lectures on Finsler geometry, World Scientific Publishing Company, 2001.
    [6] Z. Shen, Volume compasion and its applications in Riemann-Finaler geometry, Adv. Math., 128 (1997), 306–328. doi: 10.1006/aima.1997.1630
    [7] B. Wu, Y. Xin, Comparison theorems in Finsler geometry and their applications, Math. Ann., 337 (2007), 177–196.
    [8] S. Yin, Q. He, Y. Shen, On lower bounds of the first eigenvalue of Finsler-Laplacian, Publ. Math. Debrecen, 83 (2013), 385–405. doi: 10.5486/PMD.2013.5532
    [9] S. Yin, Q. He, The first eigenvalue of Finsler p-Laplacian, Differ. Geom. Appl., 35 (2014), 30–49. doi: 10.1016/j.difgeo.2014.04.009
    [10] S. Yin, Q. He, The maximum diam theorem on Finsler manifolds, arXiv: 1801.04527v1.
    [11] W. Zhao, Y. Shen, A universal volume comparison theorem for Finsler manifolds and related results, Can. J. Math., 65 (2013), 1401–1435. doi: 10.4153/CJM-2012-053-4
  • This article has been cited by:

    1. Songting Yin, Huarong Wang, Certain Conditions for a Finsler Manifold to Be Isometric with a Finsler Sphere, 2022, 10, 2299-3274, 290, 10.1515/agms-2022-0142
    2. Songting Yin, Qun He, The maximum diam theorem on Finsler manifolds, 2021, 31, 1050-6926, 12231, 10.1007/s12220-021-00715-z
  • Reader Comments
  • © 2021 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(1657) PDF downloads(33) Cited by(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog