Research article

The geometry of geodesic invariant functions and applications to Landsberg surfaces

  • Received: 05 June 2024 Revised: 24 July 2024 Accepted: 01 August 2024 Published: 07 August 2024
  • MSC : 53B40, 53C60

  • In this paper, for a given spray S on an n-dimensional manifold M, we investigated the geometry of S-invariant functions. For an S-invariant function P, we associated a vertical subdistribution VP and found the relation between the holonomy distribution and VP by showing that the vertical part of the holonomy distribution is the intersection of all spaces VFS associated with FS where FS is the set of all Finsler functions that have the geodesic spray S. As an application, we studied the Landsberg Finsler surfaces. We proved that a Landsberg surface with S-invariant flag curvature is Riemannian or has a vanishing flag curvature. We showed that for Landsberg surfaces with non-vanishing flag curvature, the flag curvature is S-invariant if and only if it is constant; in this case, the surface is Riemannian. Finally, for a Berwald surface, we proved that the flag curvature is H-invariant if and only if it is constant.

    Citation: Salah G. Elgendi, Zoltán Muzsnay. The geometry of geodesic invariant functions and applications to Landsberg surfaces[J]. AIMS Mathematics, 2024, 9(9): 23617-23631. doi: 10.3934/math.20241148

    Related Papers:

    [1] Salah G. Elgendi . Parallel one forms on special Finsler manifolds. AIMS Mathematics, 2024, 9(12): 34356-34371. doi: 10.3934/math.20241636
    [2] Fangmin Dong, Benling Li . On a class of weakly Landsberg metrics composed by a Riemannian metric and a conformal 1-form. AIMS Mathematics, 2023, 8(11): 27328-27346. doi: 10.3934/math.20231398
    [3] Özgür Boyacıoğlu Kalkan, Süleyman Şenyurt, Mustafa Bilici, Davut Canlı . Sweeping surfaces generated by involutes of a spacelike curve with a timelike binormal in Minkowski 3-space. AIMS Mathematics, 2025, 10(1): 988-1007. doi: 10.3934/math.2025047
    [4] Ayman Elsharkawy, Clemente Cesarano, Abdelrhman Tawfiq, Abdul Aziz Ismail . The non-linear Schrödinger equation associated with the soliton surfaces in Minkowski 3-space. AIMS Mathematics, 2022, 7(10): 17879-17893. doi: 10.3934/math.2022985
    [5] M. Khalifa Saad, Nural Yüksel, Nurdan Oğraş, Fatemah Alghamdi, A. A. Abdel-Salam . Geometry of tubular surfaces and their focal surfaces in Euclidean 3-space. AIMS Mathematics, 2024, 9(5): 12479-12493. doi: 10.3934/math.2024610
    [6] Gülnur Şaffak Atalay . A new approach to special curved surface families according to modified orthogonal frame. AIMS Mathematics, 2024, 9(8): 20662-20676. doi: 10.3934/math.20241004
    [7] Kai Wang, Xiheng Wang, Xiaoping Wang . Curvature estimation for point cloud 2-manifolds based on the heat kernel. AIMS Mathematics, 2024, 9(11): 32491-32513. doi: 10.3934/math.20241557
    [8] Haiming Liu, Jiajing Miao . Geometric invariants and focal surfaces of spacelike curves in de Sitter space from a caustic viewpoint. AIMS Mathematics, 2021, 6(4): 3177-3204. doi: 10.3934/math.2021192
    [9] Haiming Liu, Xiawei Chen, Jianyun Guan, Peifu Zu . Lorentzian approximations for a Lorentzian α-Sasakian manifold and Gauss-Bonnet theorems. AIMS Mathematics, 2023, 8(1): 501-528. doi: 10.3934/math.2023024
    [10] Yanlin Li, Kemal Eren, Soley Ersoy . On simultaneous characterizations of partner-ruled surfaces in Minkowski 3-space. AIMS Mathematics, 2023, 8(9): 22256-22273. doi: 10.3934/math.20231135
  • In this paper, for a given spray S on an n-dimensional manifold M, we investigated the geometry of S-invariant functions. For an S-invariant function P, we associated a vertical subdistribution VP and found the relation between the holonomy distribution and VP by showing that the vertical part of the holonomy distribution is the intersection of all spaces VFS associated with FS where FS is the set of all Finsler functions that have the geodesic spray S. As an application, we studied the Landsberg Finsler surfaces. We proved that a Landsberg surface with S-invariant flag curvature is Riemannian or has a vanishing flag curvature. We showed that for Landsberg surfaces with non-vanishing flag curvature, the flag curvature is S-invariant if and only if it is constant; in this case, the surface is Riemannian. Finally, for a Berwald surface, we proved that the flag curvature is H-invariant if and only if it is constant.



    A system of second-order homogeneous ordinary differential equations (SODE), whose coefficients do not depend explicitly on time, can be identified by a special vector field called spray. The solution of the SODE is called the geodesic of the spray. The spray corresponding to the geodesic equation of a Riemannian or Finslerian metric is called the geodesic spray of the corresponding metric.

    The concept of geodesic invariant functions (or, equivalently, S-invariant functions or first integrals of S) has various applications not only in Finsler and Riemann geometries, but also in physics. For example, the norm and energy functions are geodesic invariant functions on Finslerian or Riemannian manifolds; on Landsberg surfaces, the main scalar of the surface is S-invariant. Also, in physics, if a geodesic invariant function is given, then this function can be treated as a constant of motion; in other words, these functions are conserved along motion. Geodesic invariant functions can give important information on the geometric structure. See, for example, [3,14] and references therein.

    By [15], for a given spray S on a n-dimensional manifold M, we can associate the so-called holonomy distribution, which is generated by the horizontal vector fields and their successive Lie brackets. The functions on TM that are invariant with respect to the parallel translation are called holonomy invariant functions. These functions are constant along the holonomy distribution [8]. It is easy to see that the holonomy invariant functions are also S-invariant functions, that is, constant along the spray. However, the opposite is not true: not all functions constant along the spray are holonomy invariant. In the literature S-invariant functions are also known as first integrals of the spray S; for example, we refer to [3,14].

    In this paper, we investigate the geometry of distributions associated with homogeneous S-invariant functions of degree k0. A function P defined on TM is called k-homogeneous, if it satisfies the equation P(λv)=λkP(v) for any vTM. We show that, to any k-homogeneous S-invariant nontrivial function P, one can associate the decomposition of TTM

    TTM=HPSpan{S}VPSpan{C}, (1.1)

    where HP and VP are n1-dimensional sub-distribution of the horizontal (resp. the vertical) spaces associated with the spray. Moreover, if P is a holonomy invariant function, then

    KerdP=HVP, (1.2)

    where H is the horizontal distribution associated to S.

    As a special case, for a Finsler manifold (M,F), since F is constant along its geodesic spray S and also along the horizontal distribution H, we focus our attention on the distribution VF. In [8], the notion of metrizability freedom of sprays was introduced. For a given spray S, mS shows how many essentially different Finsler functions can be associated to it. The metrizability freedom of a spray can be determined with the help of its holonomy distribution Hol. We prove that VHol and VF coincide if and only if the metrizability freedom of S is one. In the case when mS1, then VHol is a sub-distribution of VF and we prove that

    VHol=FFSVF

    where FS denotes the set of Finsler functions associated with the spray S.

    As an application, we turn our attention to the Landsberg surfaces. We show that for a Landsberg surface, if the flag curvature is S-invariant, then the surface is Riemannian or has a vanishing flag curvature. Also, for a Landsberg surface with non-vanishing flag curvature K, we establish that K is S-invariant if and only if K is constant. In this case, the surface is Riemannian. Finally, we prove that, for a Berwald surface, the flag curvature is H-invariant if and only if K is constant.

    M is an n-dimensional smooth manifold, its tangent bundle (TM,πM,M), and its subbundle of non-zero tangent vectors (TM,π,M). On the base manifold M, we indicate local coordinates by (xi), while on TM, the induced coordinates are (xi,yi). The natural almost-tangent structure of TM is defined locally by J=yidxi, which is the vector 1-form J on TM. The canonical or Liouville vector field is the vertical vector field C=yiyi on TM.

    The geometry of sprays and Finsler manifolds has a vast literature. Here, we are using essentially the results and the terminology of [11,12].

    A vector field SX(TM) is called a spray if JS=C and [C,S]=S. Locally, a spray is expressed as follows

    S=yixi2Giyi, (2.1)

    where the spray coefficients Gi=Gi(x,y) are 2-homogeneous functions in the y=(y1,,yn) variable. A curve σ:IM is called regular if σ:ITM, where σ is the tangent lift of σ. A regular curve σ on M is called geodesic of a spray S if Sσ=σ. Locally, σ(t)=(xi(t)) is a geodesic of S if and only if it satisfies the equation

    d2xidt2+2Gi(x,dxdt)=0. (2.2)

    A nonlinear connection is described by a supplemental n-dimensional distribution to the vertical distribution, denoted as H:uTMHuTu(TM). For every uTM, we have

    Tu(TM)=HuVu. (2.3)

    Every spray S induces a canonical nonlinear connection [11] through the corresponding horizontal and vertical projectors,

    h=12(Id+[J,S]),v=12(Id[J,S]). (2.4)

    Equivalently, the canonical nonlinear connection defined by a spray is expressed as an almost product structure Γ=[J,S]=hv. A spray S is horizontal with regard to the induced nonlinear connection; this means that S=hS. Moreover, the two projectors, h and v, have the following local expressions

    h=δδxidxi,v=yiδyi,

    and the distributions are generated by the vector fields

    δδxi=xiGji(x,y)yj,δyi=dyi+Gji(x,y)dxi,

    where Gji(x,y)=Gjyi. If XX(M), then LX and iX stand for the Lie derivative with respect to X and the interior product by X, respectively. df represents the differential of fC(M). A skew-symmetric C(M)-linear map L:(X(M))X(M) is a vector -form on M. Each vector -form L defines two graded derivations of the Grassmann algebra of M, namely iL and dL, as follows

    iLf=0,iLdf=dfL(fC(M)),
    dL:=[iL,d]=iLd(1)1diL.

    The curvature tensor R of the nonlinear connection is

    R=12[h,h], (2.5)

    and the Jacobi endomorphism [12] is defined by

    Φ=v[S,h]=Rijyidxj=(2GixjS(Gij)GikGkj)yidxj.

    The two curvature tensors are related by

    3R=[J,Φ],Φ=iSR.

    For simplicity, we use the notations

    δi:=δδxi,i:=xi,˙i:=yi.

    Definition 2.1. A Finsler manifold of dimension n is a pair (M,F), where M is a smooth manifold of dimension n, and F is a continuous function F:TMR such that:

    a) F is smooth and strictly positive on TM.

    b) F is positively homogenous of degree 1 in the directional argument y: LCF=F.

    c) The metric tensor gij=˙i˙jE has rank n on TM, where E:=12F2 is the energy function.

    Since the 2-form ddJE is non-degenerate, the Euler-Lagrange equation

    ωE:=iSddJEd(ELCE)=0 (2.6)

    uniquely determines a spray S on TM. This spray is called the geodesic spray of the Finsler function. The ωE is called the Euler-Lagrange form associated with S and E.

    Definition 2.2. [15] The holonomy distribution Hol of a spray S is the distribution on TM generated by the horizontal vector fields and their successive Lie-brackets, that is

    Hol:=Xh(TM)Lie={[X1,[[Xm1,Xm]...]] | XiXh(TM)} (2.7)

    where Xh(TM) is the modules of horizontal vector fields.

    The parallel translation along curves with respect to the canonical nonlinear connection associated with a spray S can be introduced through horizontal lifts. Let c:[0,1]M be a piecewise smooth curve such that c(0)=p and c(1)=q, and let ch be a horizontal lift of the curve c (that is, πch=c and ˙ch(t)Hch(t)). The parallel translation τ:TpMTqM along c is defined as follows: If ch(0)=v and ch(1)=w, then τ(v)=w.

    Definition 2.3. Let S be a spray. A function EC(TM) is called a holonomy invariant function if it is invariant with respect to the parallel translation induced by the associated canonical nonlinear connection to S. That is, we have E(τ(v))=E(v), where vTM and τ is any parallel translation. The set of holonomy invariant functions is denoted by CHol.

    Since the parallel translations can be interpreted as travelling along the horizontal lift of curves [8], one can characterize the element of CHol as functions with vanishing horizontal derivatives. It follows that

    CHol={EC(TM) | LXE=0, XHol}. (2.8)

    Definition 2.4. Suppose S is a spray on a manifold M. If there is a Finsler function F such that its geodesic spray is S, then S is called Finsler metrizable.

    Let us denote by FS the set of Finsler function F generating S as a geodesic spray. Then, we have

    FFSE=12F2CHol (2.9)

    meaning that F is a Finsler function of S if and only if the energy function associated is a 2-homogenous regular element of CHol.

    The questions of how many essentially different Finsler metrics can be associated with a spray, and how to determine this number in terms of geometric quantities were considered in [8]. In the case when the holonomy distribution (2.7) of a spray S is regular, then the metrizability freedom mS(N) can be calculated by the following

    Theorem. ([8, Theorem 4.4]) Let S be a metrizable spray with regular holonomy distribution Hol. Then, the metrizability freedom can be calculated as mS=codim(Hol).

    In the case when the metrizability freedom of S is mS1, then for every v0TM there exists a neighborhood UTM and functionally independent element E1,,Ev of CHol on U such that any ECHol can be expressed as

     vU.E(v)=φ(E1(v),,Ev(v)), vU,

    with some function φ:RvR. We also remark that in that case, since Hol is generated by horizontal vector fields and their Lie brackets, it contains H, therefore

    Hol=HVHol, (2.10)

    where VHol denotes the vertical part of Hol. Since dim(H)=n, we get

    dimVHol=nmS. (2.11)

    Definition 3.1. Let S be a spray on M. Then, PC(TM) is called a geodesic invariant function, if for any geodesics c(t) of S it satisfies P(c(t))const.

    Obviously, for a given spray S, the function PC(TM) is a geodesic invariant function if and only if

    LSP=0, (3.1)

    that is, P is a first integral of S [3]. In that spirit, we can call such a function an S-invariant function, referring also to the spray determining the geodesic structure. We remark that P is constant along S if and only if the dynamical covariant derivative of P vanishes; see for example [4].

    As the results of [4,9] show, certain geometric distributions associated with sprays and their deformation can play a central role in the investigation of their metrizability property. This is why, motivated by [9], for further computation and analysis, we introduce a decomposition of the horizontal (resp. the vertical) distributions adapted to an S-invariant function P, homogeneous of degree k0; we introduce the endomorphisms

    hP=hdJPkPS,vP=vdvPkPC, (3.2)

    and we set

    HP:=ImhP,VP:=ImvP. (3.3)

    We have the following

    Lemma 3.2.

    1. Properties of vP and VP:

    i) ker(vP)=HSpan{C},

    ii) Im(vP)=VP is an (n1)-dimensional involutive subdistribution of V,

    iii) any XVP is an infinitesimal symmetry of P that is LXP=0,

    iv) the vertical distribution has the decomposition V=VPSpan{C}.

    2. Properties of hP and HP:

    i) ker(hP)=VSpan{S},

    ii) Im(hP)=HP is an (n1)-dimensional subdistribution of H,

    iii) the horizontal distribution has the decomposition H=HPSpan{S},

    3. J(HP)=VP.

    Proof. We prove (1) in detail. The computations for (2) are similar.

    ad i) We note that H=Kerv, therefore HKervP. Moreover, if VkervP is vertical, then using v(V)=V we get

    vP(V)=0V=V(P)kPC,

    that is VSpan{C} and we get i).

    ad ii) We introduce the simplified notation Pi:=˙iP and the vector fields

    hi:=hP(δi)=δiPikPS, (3.4a)
    vi:=vP(˙i)=˙iPikPC (3.4b)

    for i=1,,n. We get

    HP=Span{h1,,hn}, (3.5a)
    VP=Span{v1,,vn}. (3.5b)

    We note that the vector fields in (3.5a) (resp., in (3.5b)) are not independent since yihi=0 (resp., yivi=0). Because the k-homogeneity property of P (and the (k1)-homogeneity property of Pi) for any vi,vjVP, their Lie bracket is

    [vi,vj]=[˙iPikPyk˙k, ˙jPjkPy˙]=PikP˙jPjkP˙i=PikPvjPjkPvi

    and hence, from (3.5b), we get that [vi,vj]VP hence VP is involutive.

    ad iii) One can check that the generators (3.5b) of the distribution are infinitesimal symmetry of P. Indeed, using Euler's theorem of the homogeneous functions, we get for the k-homogeneous P:

    LCP=kP, (3.6)

    and therefore

    LviP=˙i(P)PikPC(P)=PiPikPkP=0. (3.7)

    ad iv) Supposing CVP we get from (3.5b) that C=Civi with some coefficients Ci. Solving this equation, since C(P)=kP and vi(P)=0, we find that C(P)=Civi(P)=0, which is a contradiction.

    For 3), we note that for the generators (3.4a) of (3.5a) and (3.4b) of (3.5b), we get

    Jhi=JδiPikPJS=˙iPikPC=vi, (3.8)

    i=1,,n, and this proves 3).

    From Lemma 3.2 we get the following

    Corollary 3.3. For a given spray S on TM, then any non-trivial S-invariant function PC(TM) and homogeneous of degree k0 gives rise to the direct sum decomposition (1.1). Moreover, if P is constant along HP, then we have also (1.2).

    We have the following

    Proposition 3.4. Let (M,F) be a Finsler manifold with geodesic spray S. If P is a k-homogeneous holonomy invariant function with k0, then

    VHolVP. (3.9)

    Proof. Assume that P is a k-homogeneous holonomy invariant function with k0, then PCHol, and according to (2.8), we have VHolHolKerdP. It follows that

    VHolVKerdP=VP,

    where we use the notation (3.3).

    Remark 3.5. Let (M,F) be a Finsler manifold with geodesic spray S. If P is a k-homogeneous S-invariant (but not necessarily holonomy invariant) function with k0 and VHolVP, then dhdhP=0.

    Proof. We note that, since P is not necessarily a holonomy invariant function, we do not have dhP=0. However, the image of the curvature tensor R is in the holonomy distribution. If VHolVP, then dRP=0. On the other hand, using (2.5) and the properties d[h,h]=[dh,dh] and

    [dh,dh]=dhdh(1)dhdh=2dhdh,

    we have

    dhdhP=12d[h,h]P=dRP=0,

    which shows the statement of the remark.

    It should be noted that in the generic case, the holonomy distribution of a spray is the 2n-dimensional distribution TTM and the metrizability freedom is mS=0. For mS=1 we get the following

    Theorem 3.6. Let S be a given spray metrizability freedom mS=1, that is (essentially) uniquely metrizable by a Finsler function F. Then, for any 1-homogeneous S-invariant function P, we have VHol=VP if and only if F=cP where cR{0}.

    Proof. Since the metrizability freedom of S is 1, then by [8] the codimension of Hol is one. That is, the dimension of Hol is 2n1 and by the fact that the dimension of HHol is n, we can conclude that the dimension of VHol=n1.

    Assume that F=cP, then P is holonomy invariant 1-homogenous function. From {Proposition 3.4}, we have VHolVP. Since the dimension of both spaces is n1, we get their equality.

    Conversely, assume that VHol=VP, then

    dvPF=0dvFdvPPdCF=0.

    Since dCF=F, then we have

    dvFdvPPF=0dvFF=dvPP.

    Then, there exists a function a(x) on M such that F=ea(x)P. Now, since P is S-invariant, then LSP=0 and also LSF=0; therefore, LSa(x)=0. Locally, we obtain that

    yiia(x)2Gi˙ia(x)=0yiia(x)=0.

    By differentiation with respect to yj, we get ja(x)=0, that is a(x) is constant function. Hence, we get F=cP.

    Corollary 3.7. Let (M,F) be a Finsler manifold with isotropic non-vanishing curvature. Then, for any 1-homogeneous S-invariant function P, we have VHol=VP if and only if F=cP, where c is a non-zero constant.

    Proof. In the case where the Finsler manifold has a non-vanishing isotropic curvature, then by [8], the metrizability freedom of its geodesic spray is 1. Therefore, the result follows by Theorem 3.6.

    The next theorem characterizes VHol and therefore Hol as the intersection of distributions associated with geodesic invariant functions:

    Theorem 3.8. Let S be a metrizable spray with regular holonomy distribution. Then, we have

    VHol=FFSVF. (3.10)

    Proof. Let us assume that S is a metrizable spray with regular holonomy distribution on an n-dimensional manifold M, and its metric freedom is mS (1). According to [8, Theorem 4.4], we have codim(Hol)=mS, or equivalently,

    dim(Hol)=2nmS, (3.11)

    and at the neighborhood of any (x,y)TM, there exists a set {E1,EmS} of energy functions associated with S such that any energy function of S can be locally written as a functional combination of E1,EmS. It follows that the corresponding Finsler functions {F1,FmS} are functionally independent, and locally generating the set of Finsler functions of S, that is, every Finsler function F of S can be written as a functional combination

    F=ϕ(F1,,Fv)

    with some 1-homogeneous function ϕ. It follows that

    FFSKer(dF)=vμ=1Ker(dFμ). (3.12)

    Since {F1,,Fv} are functionally independent, their derivatives are linearly independent, therefore vμ=1Ker(dFμ) is characterized by mS linearly independent equations in TTM. It follows that

    dim(vμ=1Ker(dFμ))=dim(TTM)mS=2nmS. (3.13)

    Moreover, the functions Fμ are all holonomy invariant functions; therefore, Ker(dFμ) contains the holonomy distribution for μ=1,,mS, and as a consequence, their intersection vμ=1Ker(dFμ) also contains Hol. Since the dimension of the intersection (3.13) and the dimension of the holonomy distribution (3.11) are equal, we get

    Hol=vμ=1Ker(dFμ). (3.14)

    Using the vertical projection for (3.14) we get

    VHol=v(Hol)(3.14)=v(vμ=1Ker(dFμ))(3.12)==v(FFSKer(dF))=FFSv(Ker(dF))=FFSVF

    showing the statement of the theorem.

    Corollary 3.9. Let S be a metrizable spray by a Finsler function F. Then, VHol=VF if and only if the metrizability freedom of S is mS=1.

    Theorem 3.10. Let F be a Finsler function and S its geodesic spray. Then, if P is a 1-homogeneous nontrivial VF-invariant function, then it is regular. Moreover, if P is S-invariant, then P=cF with some constant cR.

    We remark that the theorem shows that the S-invariant and VF-invariant properties are essentially characterizing the Finsler function associated with S.

    Proof. Let P be a 1-homogeneous VF-invariant function. It follows that it satisfies the the system

    dXP=0,XVF.

    Then, we have

    dvFP=dvPdvFFP=0dvFF=dvPP.

    Then, there exists a function a(x) on M such that F=ea(x)P. Then P=ea(x)F, and hence P inherits its regularity from the Finsler function F.

    Now, assume that P is S-invariant; then, we have LSP=0 and using the fact that LSF=0, we have

    LSF=LSea(x)P=ea(x)PLSa(x)=0.

    Then, we obtain that yiia(x)=0. But by differentiating with respect to the yj variable, we get ja(x)=0. That is a(x)=const. Consequently, we get F=cP.

    Definition 4.1. A Finsler metric F on a manifold M is called a Berwald metric, if in any standard local coordinate system in TM the connection coefficients Gij(x,y) are linear. A Finsler metric F is called Landsberg metric if Landsberg tensor with the components Lijk=12FGhijkFyh is identically zero.

    The Berwald- and Landsberg-type Finsler metrics are the most important particular cases in Finler geometry. For Berwald metrics, the associated canonical connection is linear; for Landsberg metrics the parallel transport with respect to the canonical connection preserves the metric [1]. It is well known that all Berwald-type Finsler metrics are also Landsbergian, but there is the long-open, so-called unicorn problem: Is there a Landsberg metric that is not Berwald? In higher dimensions (n3), there are non-regular Landsberg metrics that are not Berwladian; for more details, we refer to [7,17]. In dimension two, L. Zhou [19] investigated a class of Landsberg surfaces and claimed that this class is not Bewaldian. Later, in [10], it was shown that the class is, in fact, Berwaldian. Up to the best of our knowledge, there is no example of non-Berwaldian Landsberg surfaces.

    A Finsler function F with the geodesic spray S is said to be of scalar flag curvature if there exists a function KC(TM) such that the Jacobi endomorphism Φ of the geodesic spray S is given by

    Φ=K(F2JFdJFC). (4.1)

    Since the Jacobi endomorphism Φ of any Finsler surface is in the above form, then it is clear that all Finsler surfaces are of scalar flag curvature K(x,y). Also, since the curvature R of a spray vanishes if and only if the Jacobi endomorphism vanishes, then the curvature of any Finsler surface vanishes if and only if K vanishes.

    Whenever the scalar curvature K of the Finsler surface is non-vanishing, we will use the so-called Berwald frame, introduced by Berwald in [6]: It is a frame on TM canonically associated with a 2-dimensional Finsler manifold and used to investigate projectively flat 2-dimensional Finsler manifolds. We note that when the scalar curvature vanishes, the Berwald frame is not defined. For more details, we refer, for instance, to [18].

    Lemma 4.2. [2] Let (M,F) be a Finslerian surface with the geodesic spray S and of flag curvature K0. Then, the Berwald frame {S,H,C,V} satisfies JH=V,

    [S,H]=KV, (4.2a)
    [S,V]=H, (4.2b)
    [H,V]=S+IH+S(I)V, (4.2c)

    and

    H(F)=V(F)=0. (4.3)

    Moreover, the Bianchi's identity is given by [14, Proposition 1.4]

    S2(I)+V(K)+IK=0, (4.4)

    where K is the flag curvature and I is the main scalar of (M,F).

    One can characterize the Berwald- and Landsberg-type Finler metrics in terms of the main scalar:

    Lemma 4.3. [5] A Finsler surface (M,F) is

    1. Landsberg if and only if S(I)=0.

    2. Berwald if and only if S(I)=0 and H(I)=0.

    Proposition 4.4. All Landsberg surfaces with basic flag curvature are either Riemannian or have vanishing flag curvature.

    Proof. Let (M,F) be a Landsberg surface with basic flag curvature, that is, K=K(x) is a function on the manifold M. Then, V(K)=0, and by using the fact that S(I)=0 together with (4.4), we have

    KI=0.

    Then, we have either K=0 or I=0 and this completes the proof.

    Proposition 4.5. For any Landsberg surface (M,F) with non-vanishing curvature, we have

    β+IV(β)+H(I)+V2(β)=0, (4.5)

    where β:=S(K0)K0S(t0I(t)dt), K0C(TM), V(K0)=0, I is the main scalar of (M,F) and the integration here is taken with respect to V.

    Proof. Assume that (M,F) is a Landsberg surface with non-vanishing K. We work on a neighborhood of a point (x0,y0)TM where F is regular. Then, from Lemma 4.3, we get that S(I)=0 and hence S2(I)=S(S(I))=0. Then, (4.4) has the form

    V(K)=IK. (4.6)

    Since K0, then we can write

    V(K)K=I.

    Using integration as in [16] we obtain

    K=K0exp(t0I(t)dt), (4.7)

    where K0C(TM) and V(K0)=0. But since K is homogeneous of degree 0 and by the fact that [C,V]=0, then K0 must be homogeneous of degree 0, that is, C(K0)=0. That is, V(K0)=0 and C(K0)=0, hence K0=K0(x).

    Taking the fact that S(I)=0, (4.7) implies

    S(K)=S(K0)exp(t0I(t)dt)+KS(t0I(t)dt)=S(K0)KK0+KS(t0I(t)dt).

    From which we can write

    S(K)K=S(K0)K0+S(t0I(t)dt). (4.8)

    Then, (4.8) can be written in the form

    S(K)=βK, (4.9)

    where β=S(K0)K0+S(t0I(t)dt). Applying S on (4.6) and using (4.9), we have

    S(V(K))=IS(K)=βIK. (4.10)

    Applying V on (4.9) and using (4.6), we have

    V(S(K))=V(β)K+βV(K)=V(β)KβIK. (4.11)

    Now, by the property that [V,S]=H (4.2b), (4.10), and (4.11) we have

    H(K)=V(β)K. (4.12)

    From which, together with (4.6), we get

    V(H(K))=V2(β)K+V(β)V(K)=V2(β)KIK V(β). (4.13)
    H(V(K))=H(I)KIH(K)=H(I)KIK V(β). (4.14)

    Since [H,V]K=H(V(K))V(H(K)) then by (4.2c), (4.13), and (4.14), we have

    S(K)+IH(K)=KH(I)KV2(β)

    from which, together with the fact that K0, and by (4.9), (4.12), we have

    β+IV(β)+H(I)+V2(β)=0.

    This completes the proof.

    As a consequence of the above proposition, we have the following result, which is obtained by [13] and [18], proved in a different way.

    Theorem 4.6. Let (M,F) be a Landsberg surface with non-zero flag curvature. If the flag curvature is S-invariant, then the surface is Riemannian.

    Proof. Let (M,F) be a Landsberg surface with non-vanishing flag curvature K and the property that S(K)=0. Then, by (4.9), we get that β=0 and therefore V(β)=V2(β)=0. Now, by (4.5), we obtain that H(I)=0 and the surface is Berwaldian. Moreover, by (4.12), we have H(K)=0 and using the fact that S(K)=0, (4.2a) implies

    KV(K)=0,

    from which, together with Proposition 4.4, the result follows.

    Theorem 4.7. Let (M,F) be a Landsberg surface with non-vanishing flag curvature K; then, K is S-invariant if and only if K is constant. In this case, F is Riemannian.

    Proof. Let (M,F) be a surface with non-vanishing flag curvature K. It is obvious that if K is constant, then S(K)=0 and hence K is S-invariant. Now, assume that K is S-invariant, that is, S(K)=0. By (4.9), β=0 and then by (4.12) we get that H(K)=0. Since [S,H]=KV, then KV(K)=S(H(K))H(S(K))=0, and hence V(K)=0 since K0. Moreover, K is zero homogeneous in y, then C(K)=0. Therefore, we have

    S(K)=0,H(K)=0,V(K)=0,C(K)=0

    which implies that K is constant. Then, F is Riemnnian by Theorem 4.6.

    A smooth function f on TM is said to be H-invariant if H(f)=0. Let's end this work by the following result.

    Theorem 4.8. Let (M,F) be a Berwald surface with non-vanishing flag curvature. Then, the flag curvature K is H-invariant if and only if K is constant.

    Proof. Let (M,F) be a Berwald surface. If K is constant, then it is clear that H(K)=0 and hence it is H-invariant. Now, assume that H(K)=0. If K=0, then the proof is done. If K0, then by (4.12), V(β)=0. Since the surface is Berwaldian, then H(I)=0. Therefore, by (4.5), β=0 and by (4.9), we have S(K)=0. Using (4.2a), we get that V(K)=0 since K0. Since C(K)=0, we have

    S(K)=0,H(K)=0,V(K)=0,C(K)=0

    which means that K is constant.

    In this work, we have investigated the concept of geodesically invariant functions (or, equivalently, S-invariant functions for a given spray S) and some of its geometric consequences. For a given S-invariant function P and homogeneous of degree k0, we manged to express the horizontal and vertical subbundles as a direct sum of associated distributions depending on the function P. Moreover, we study the relationship between the holonomy distribution and the kernel distribution of P. Also, we pay some attentions to the role of the metrizability freedom and its effect on the geometry of an S-invariant function. Finally, as an application, we focus on the Berwald- and Landsberg-type surfaces.

    Salah G. Elgendi and Zoltán Muzsnay: Conceptualization, methodology, validation, writing-original draft, writing-review & editing. All authors of this article have been contributed equally. All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors would like to thank the referee for the constructive comments and recommendations which contributed to improving the paper.

    The authors declare no conflict of interest.



    [1] D. Bao, On two curvature-driven problems in Riemann-Finsler geometry, Adv. Stud. Pure Math., 48 (2007), 19–71.
    [2] I. Bucataru, G. Cretu, E. H. Taha, Frobenius integrability and Finsler metrizability for 2-dimensional sprays, Differ. Geom. Appl., 56 (2018), 308–324. https://doi.org/10.1016/j.difgeo.2017.10.002 doi: 10.1016/j.difgeo.2017.10.002
    [3] I. Bucataru, O. Constantinescu, G. Cretu, A class of Finsler metrics admitting first integrals, J. Geom. Phys., 166 (2021), 104254. https://doi.org/10.1016/j.geomphys.2021.104254 doi: 10.1016/j.geomphys.2021.104254
    [4] I. Bucataru, Z. Muzsnay, Projective and Finsler metrizability: Parameterization-rigidity of the geodesics, Int. J. Math., 23 (2012). https://doi.org/10.1142/S0129167X12500991
    [5] S. Bacsó, M. Matsumoto, Reduction theorems of certain Landsberg spaces to Berwald spaces, Publ. Math.-Debrecen, 48 (1996), 357–366. https://doi.org/10.5486/pmd.1996.1733 doi: 10.5486/pmd.1996.1733
    [6] L. Berwald, On Finsler and Cartan geometries. Ⅲ: Two-dimensional Finsler spaces with rectilinear extremals, Ann. Math., 42 (1941), 84–112. https://doi.org/10.2307/1968989 doi: 10.2307/1968989
    [7] S. G. Elgendi, Solutions for the Landsberg unicorn problem in Finsler geometry, J. Geom. Phys., 159 (2021), 103918. https://doi.org/10.1016/j.geomphys.2020.103918 doi: 10.1016/j.geomphys.2020.103918
    [8] S. G. Elgendi, Z. Muzsnay, Freedom of h(2)-variationality and metrizability of sprays, Differ. Geom. Appl., 54 (2017), 194–207. https://doi.org/10.1016/j.difgeo.2017.03.020 doi: 10.1016/j.difgeo.2017.03.020
    [9] S. G. Elgendi, Z. Muzsnay, Metrizability of holonomy invariant projective deformation of sprays, Can. Math. Bull., 66 (2020), 701–714. https://doi.org/10.4153/s0008439520000016 doi: 10.4153/s0008439520000016
    [10] S. G. Elgendi, N. L. Youssef, A note on a result of L. Zhou's on Landsberg surfaces with K=0 and J=0, Differ. Geom. Appl., 77 (2021), 101779. https://doi.org/10.1016/j.difgeo.2021.101779 doi: 10.1016/j.difgeo.2021.101779
    [11] J. Grifone, Structure presque-tangente et connexions Ⅰ, Ann. Inst. Fourier, 22 (1972), 287–334. https://doi.org/10.5802/aif.407 doi: 10.5802/aif.407
    [12] J. Grifone, Z. Muzsnay, Variational principles for second-order differential equations, World Scientific Publishing, 2000. https://doi.org/10.1142/9789812813596
    [13] F. Ikeda, On two-dimensional Landsberg spaces, Tensor, 33 (1979), 43–48.
    [14] P. Foulon, R. Ruggiero, A first integral for C,k-basic Finsler surfaces and application to rigidity, Proc. Am. Math. Soc., 144 (2016), 3847–3858. https://doi.org/10.1090/proc/13079 doi: 10.1090/proc/13079
    [15] Z. Muzsnay, The Euler-Lagrange PDE and Finsler metrizability, Houston J. Math., 32 (2006), 79–98.
    [16] S. V. Sabau, H. Shimada, Riemann-Finsler surfaces, Math. Soc. Japan, 48 (2007), 125–163. https://doi.org/10.2969/aspm/04810125
    [17] Z. Shen, On a class of Landsberg metrics in Finsler geometry, Canad. J. Math., 61 (2009), 1357–1374. https://doi.org/10.4153/cjm-2009-064-9 doi: 10.4153/cjm-2009-064-9
    [18] E. H. Taha, On a class of Landsberg metrics in Finsler geometry, Int. J. Geom. Meth. Mod. Phys., 20 (2023), 2350002. https://doi.org/10.1142/s0219887823500020 doi: 10.1142/s0219887823500020
    [19] L. Zhou, The Finsler surface with K=0 and J=0, Differ. Geom. Appl., 35 (2014), 370–380. https://doi.org/10.1016/j.difgeo.2014.02.003 doi: 10.1016/j.difgeo.2014.02.003
  • This article has been cited by:

    1. Salah G. Elgendi, Parallel one forms on special Finsler manifolds, 2024, 9, 2473-6988, 34356, 10.3934/math.20241636
  • Reader Comments
  • © 2024 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(814) PDF downloads(45) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog