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

[2]\fnmRyu \surUeno

[1,2]\orgdivDepartment of Mathematics, \orgnameHokkaido University, \orgaddress\citySapporo, \postcode060-0810, \countryJapan

A Variation Problem for Mappings between Statistical Manifolds

\fnmHitoshi \surFuruhata [email protected]    [email protected] *
Abstract

We present statistical biharmonic maps, a new class of mappings between statistical manifolds naturally derived from a variation problem. We give the Euler-Lagrange equation of this problem and prove that improper affine hyperspheres induce examples of such maps.

keywords:
statistical manifolds, harmonic maps, biharmonic maps, affine minimal hypersurfaces, improper affine hyperspheres
pacs:
[

Mathematical Subject Classification] 53B12, 53A15, 58E20, 53C43

1 Introduction

A statistical manifold is nothing but a Riemannian manifold equipped with a torsion-free affine connection satisfying the Codazzi equation. Although the origin of this curious terminology is in information geometry, it is a fact that various research fields have provided interesting examples of statistical manifolds. At the beginning of the 20th century, W. Blaschke developed the hypersurface theory in the equiaffine space. A pair of so-called affine fundamental form and the induced affine connection is a statistical structure on the hypersurface. Many types of research on affine hyperspheres, affine minimal hypersurfaces, and so on have a relation to the study of statistical structures (See Example 2.9).

In Riemannian geometry, harmonic maps might be understood as the most important research subject derived from the variation principle. Finding the corresponding subjects in the geometry of statistical manifolds is an interesting task for us. In fact, the Euler-Lagrange equation, the harmonic map equation, can be generalized in the statistical manifold setting (See [12, 13] for example). However, it is hard to find attempts to generalize them as the variation problem itself. In this paper, we will take notice of a certain class of mappings between statistical manifolds determined from a variation problem.

Let (M,g,M)(M,g,\nabla^{M}) be a statistical manifold. By definition, the affine connection M\nabla^{M} is not necessarily the Levi-Civita connection of gg. If so, we call the statistical manifold Riemannian. Assume temporarily that MM is compact for the sake of simplicity. Let (N,h,N)(N,h,\nabla^{N}) be a statistical manifold. For a smooth map u:MNu:M\to N, we define

τ(u)\displaystyle\tau(u) =\displaystyle= τ(g,M,N)(u)\displaystyle\tau^{(g,\nabla^{M},\nabla^{N})}(u)
=\displaystyle= trg{(X,Y)XNuYuXMY}Γ(u1TN),\displaystyle\mathrm{tr}_{g}\{(X,Y)\mapsto\nabla^{N}_{X}u_{*}Y-u_{*}\nabla^{M}_{X}Y\}\in\Gamma(u^{-1}TN),

and

E2(u)\displaystyle E_{2}(u) =\displaystyle= E2(g,M,h,N)(u)\displaystyle E_{2}^{(g,\nabla^{M},h,\nabla^{N})}(u)
=\displaystyle= 12Mτ(u)h2𝑑μg,\displaystyle\frac{1}{2}\int_{M}\|\tau(u)\|_{h}^{2}d\mu_{g},

where dμgd\mu_{g} is the standard measure derived from the Riemanian metric gg. We remark that the functional E2E_{2} is defined by using the statistical structures of both the source and the target. It seems that our E2E_{2} is the simplest one among such functionals. This functional has been known as the bi-energy for a smooth map in the case where both statistical manifolds are Riemannian. In this case, the first variation formula is given by Jiang Guoying, and its critical point is called a biharmonic map (See [4, 14] for example).

In the statistical manifold case, we obtain the first variation formula (Theorem 3.1), and τ2(u)=0\tau_{2}(u)=0 as the Euler-Lagrange equation, where

τ2(u)\displaystyle\tau_{2}(u) =\displaystyle= τ2(g,M,h,N)(u)\displaystyle\tau_{2}^{(g,\nabla^{M},h,\nabla^{N})}(u)
=\displaystyle= Δ¯uτ(u)+divg(trgKM)τ(u)\displaystyle\bar{\Delta}^{u}\tau(u)+\operatorname{div}^{g}(\mathrm{tr}_{g}K^{M})\tau(u)
i=1mLN(uei,τ(u))ueiKN(τ(u),τ(u))Γ(u1TN).\displaystyle\quad-\sum_{i=1}^{m}L^{N}(u_{*}e_{i},\tau(u))u_{*}e_{i}-K^{N}(\tau(u),\tau(u))\quad\in\Gamma(u^{-1}TN).

The symbols will be prepared in §2. In the case where both statistical manifolds are Riemannian, the above equation of course restores Jiang’s equation. We will call τ2(u)\tau_{2}(u) the statistical bi-tension field of uu, and say that uu is a statistical biharmonic map if τ2(u)=0\tau_{2}(u)=0. This class might be the first one derived from a variation problem for mappings between statistical manifolds.

We illustrate simple examples of statistical biharmonic maps in §4. We remark that a map uu with τ(u)=0\tau(u)=0 is statistical biharmonic by definition. We will introduce one of the interesting examples beforehand (Theorem 4.5). Let f:M(m+1,D,det)f:M\to(\mathbb{R}^{m+1},D,\det) be an improper affine hypersphere, and (g,M)(g,\nabla^{M}) the induced statistical structure on MM by ff. Although the setting is equiaffine geometry, the standard connection DD is considered as the Levi-Civita connection of the Euclidean metric g0g_{0}, and the volume form det\det is also naturally determined by g0g_{0}. An improper affine hypersphere ff is statistical biharmonic as a map from the statistical manifold (M,g,M)(M,g,\nabla^{M}) to the Riemannian statistical manifold (m+1,g0,D)(\mathbb{R}^{m+1},g_{0},D). In this case, τ(f)\tau(f) does not vanish, though τ2(f)\tau_{2}(f) vanishes. Improper affine hyperspheres are important research objects in affine geometry (See [1, 9] for example).

This paper is organized as follows. In §2, we prepare the basic notion of statistical manifolds, particularly, Laplacian, which is useful for our first variation formula. §3 devotes the proof of the first variation formula. In §5, we give some conditions which imply τ=0\tau=0 from τ2=0\tau_{2}=0. In the Riemannian setting, harmonic maps are naturally biharmonic maps, and in many settings, the converse is also true, as the Chen conjecture [2] has suggested. We give the corresponding properties in the statistical setting.

2 Preliminaries

2.1 Statistical manifolds

Throughout this paper, all the objects are assumed to be smooth. MM denotes a manifold of dimension m2m\geq 2, C(M)C^{\infty}(M) the set of functions on MM, and Γ(E)\Gamma(E) the set of sections of a vector bundle EE over MM.

In this section, we quickly fix the notation of the geometry of statistical manifolds. Let (g,)(g,\nabla) be a statistical structure on MM, that is, gg is a Riemannian metric and \nabla is a torsion-free affine connection satisfying (Xg)(Y,Z)=(Yg)(X,Z)(\nabla_{X}g)(Y,Z)=(\nabla_{Y}g)(X,Z) for X,Y,ZΓ(TM)X,Y,Z\in\Gamma(TM). We denote by g\nabla^{g} the Levi-Civita connection of gg.

Definition 2.1.

Let (M,g,)(M,g,\nabla) be a statistical manifold.

(1)  We set K=K(g,)Γ(TM(1,2))K=K^{(g,\nabla)}\in\Gamma(TM^{(1,2)}) by

K(X,Y)=KXY=XYXgY,X,YΓ(TM).K(X,Y)=K_{X}Y=\nabla_{X}Y-\nabla^{g}_{X}Y,\quad X,Y\in\Gamma(TM).

We also denote it by KMK^{M} if necessary.

(2)  We define an affine connection ¯\overline{\nabla} by

Xg(Y,Z)=g(XY,Z)+g(Y,¯XZ),X,Y,ZΓ(TM),Xg(Y,Z)=g(\nabla_{X}Y,Z)+g(Y,\overline{\nabla}_{X}Z),\quad X,Y,Z\in\Gamma(TM),

and call it the conjugate connection of \nabla with respect to gg.

(3)  We define the curvature tensor field RΓ(TM(1,3))R^{\nabla}\in\Gamma(TM^{(1,3)}) of \nabla by

R(X,Y)Z=XYZYXZ[X,Y]Z,X,Y,ZΓ(TM).R^{\nabla}(X,Y)Z=\nabla_{X}\nabla_{Y}Z-\nabla_{Y}\nabla_{X}Z-\nabla_{[X,Y]}Z,\quad X,Y,Z\in\Gamma(TM).

On a statistical manifold (M,g,)(M,g,\nabla), we often denote RR^{\nabla} by RR, RgR^{\nabla^{g}} by RgR^{g}, and R¯R^{\overline{\nabla}} by R¯\overline{R}, for short.

Remark 2.2.

The following formulas hold for X,Y,Z,WΓ(TM)X,Y,Z,W\in\Gamma(TM):

(g)(Y,Z;X)=(Xg)(Y,Z)=2g(K(X,Y),Z),\displaystyle(\nabla g)(Y,Z;X)=(\nabla_{X}g)(Y,Z)=-2g(K(X,Y),Z), (2.1)
¯XY=XgYK(X,Y),\displaystyle\overline{\nabla}_{X}Y=\nabla^{g}_{X}Y-K(X,Y), (2.2)
g(R¯(X,Y)Z,W)=g(Z,R(X,Y)W),\displaystyle g(\overline{R}(X,Y)Z,W)=-g(Z,R(X,Y)W), (2.3)
12(R+R¯)(X,Y)Z=Rg(X,Y)Z+[KX,KY]Z.\displaystyle\frac{1}{2}(R+\overline{R})(X,Y)Z=R^{g}(X,Y)Z+[K_{X},K_{Y}]Z. (2.4)
Definition 2.3.

A statistical manifold (M,g,)(M,g,\nabla) is said to be conjugate symmetric if R=R¯R=\overline{R} holds.

The properties of the curvature tensor field RR^{\nabla} related to gg are similar to the ones of RgR^{g} if the considering statistical manifold is conjugate symmetric. In particular, we have

g(R(Z,W)X,Y)=g(R(X,Y)Z,W)g(R^{\nabla}(Z,W)X,Y)=g(R^{\nabla}(X,Y)Z,W) (2.5)

for X,Y,Z,WΓ(TM)X,Y,Z,W\in\Gamma(TM).

Definition 2.4.

A statistical manifold (M,g,)(M,g,\nabla) is said to be trace-free or to satisfy the apolarity condition if trgK=0\mathrm{tr}_{g}K=0.

The apolarity is in the terminology of equiaffine geometry. On the other hand, in centroaffine geometry, T=T(g,)=1mtrgKΓ(TM)T=T^{(g,\nabla)}=\frac{1}{m}\mathrm{tr}_{g}K\in\Gamma(TM) is often called the Tchebychev vector field, and 𝒯=𝒯(g,)=gTΓ(TM(1,1))\mathcal{T}=\mathcal{T}^{(g,\nabla)}=\nabla^{g}T\in\Gamma(TM^{(1,1)}) the Tchebychev operator. See [8] for example.

Definition 2.5.

Let (M,g,)(M,g,\nabla) be a statistical manifold. The curvature interchange tensor field L=L(g,)Γ(TM(1,3))L=L^{(g,\nabla)}\in\Gamma(TM^{(1,3)}) is defined by

g(L(Z,W)X,Y)=g(R(X,Y)Z,W),X,Y,Z,WΓ(TM).g(L(Z,W)X,Y)=g(R^{\nabla}(X,Y)Z,W),\quad X,Y,Z,W\in\Gamma(TM).

We denote L¯=L(g,¯)\bar{L}=L^{(g,\overline{\nabla})}.

Example 2.6.

(1)  If a statistical structure (g,)(g,\nabla) is conjugate symmetric, then L=RL=R holds.

(2)  Suppose that RR is written as

R(X,Y)Z=g(Y,Z)SXg(X,Z)SY,X,Y,ZΓ(TM)R(X,Y)Z=g(Y,Z)SX-g(X,Z)SY,\quad X,Y,Z\in\Gamma(TM) (2.6)

for some SΓ(TM(1,1))S\in\Gamma(TM^{(1,1)}) such that g(SX,Y)=g(X,SY)g(SX,Y)=g(X,SY). Then the curvature interchange tensor field of (g,)(g,\nabla) is given as

L(Z,W)X=g(X,Z)SW+g(SY,W)Z.L(Z,W)X=-g(X,Z)SW+g(SY,W)Z.
Remark 2.7.

The curvature interchange tensor field LL has the following properties for X,Y,Z,WΓ(TM)X,Y,Z,W\in\Gamma(TM):

L(Z,W)Y=L¯(W,Z)Y,\displaystyle L(Z,W)Y=-\bar{L}(W,Z)Y,
g(L(Z,W)Y,X)=g(Y,L(Z,W)X),\displaystyle g(L(Z,W)Y,X)=-g(Y,L(Z,W)X),
R¯(Y,Z)W=L(Y,W)ZL(Z,W)Y.\displaystyle\overline{R}(Y,Z)W=L(Y,W)Z-L(Z,W)Y.
Example 2.8.

On the Euclidean space (2,g0)(\mathbb{R}^{2},g_{0}), define a torsion-free affine connection \nabla as follows.

xx=x,xy=0,yy=y.\nabla_{\frac{\partial}{\partial x}}\frac{\partial}{\partial x}=\frac{\partial}{\partial x},\quad\nabla_{\frac{\partial}{\partial x}}\frac{\partial}{\partial y}=0,\quad\nabla_{\frac{\partial}{\partial y}}\frac{\partial}{\partial y}=\frac{\partial}{\partial y}.

The triplet (2,g0,)(\mathbb{R}^{2},g_{0},\nabla) is a statistical manifold. It is easy to see that \nabla is flat, and accordingly the statistical structure (g0,)(g_{0},\nabla) is conjugate symmetric. We have that g0trg0K=0\nabla^{g_{0}}\mathrm{tr}_{g_{0}}K=0 while trg0K0\mathrm{tr}_{g_{0}}K\neq 0 on some points.

Example 2.9.

Let (m+1,D,det)(\mathbb{R}^{m+1},D,\det) be a standard equiaffine space. By definition, DD denotes the standard flat affine connection of m+1\mathbb{R}^{m+1}, and det\det the standard volume form, which is given as the determinant with respect to the standard affine coordinate system with respect to DD. The group SL(m+1,)m+1SL(m+1,\mathbb{R})\ltimes\mathbb{R}^{m+1} acts on this space preserving these structures. Two hypersurfaces f1,f2:Mm+1f_{1},f_{2}:M\to\mathbb{R}^{m+1} are equiaffinely congruent if there exist ASL(m+1,)A\in SL(m+1,\mathbb{R}) and bm+1b\in\mathbb{R}^{m+1} such that f2(x)=Af1(x)+bf_{2}(x)=Af_{1}(x)+b for all xMx\in M.

Let f:Mm+1f:M\to\mathbb{R}^{m+1} be a locally strongly convex hypersurface with the Blaschke normal vector field ξΓ(f1Tm+1)\xi\in\Gamma(f^{-1}T\mathbb{R}^{m+1}). By definition, ff and ξ\xi satisfy the following properties: At each point xMx\in M, we have the decomposition Tf(x)m+1=(df)xTxMξxT_{f(x)}\mathbb{R}^{m+1}=(df)_{x}T_{x}M\oplus\mathbb{R}\xi_{x}. According to this decomposition, define ,h,S\nabla,h,S and τ\tau by

DXfY=fXY+h(X,Y)ξ,\displaystyle D_{X}f_{*}Y=f_{*}\nabla_{X}Y+h(X,Y)\xi,
DXξ=fSX+τ(X)ξ\displaystyle D_{X}\xi=-f_{*}SX+\tau(X)\xi

for X,YΓ(TM)X,Y\in\Gamma(TM). Then (1) τ\tau vanishes, and hh is a Riemannian metric on MM. (2) dμh(X1,,Xm)=det(fX1,,fXm,ξ)d\mu_{h}(X_{1},\ldots,X_{m})=\det(f_{*}X_{1},\ldots,f_{*}X_{m},\xi) holds for X1,,XmΓ(TM)X_{1},\ldots,X_{m}\in\Gamma(TM). We remark that ξ\xi is determined for ff, as a unit normal vector field in Euclidean geometry. We often call hh the Blaschke metric, and SS the affine shape operator of ff, respectively.

Such a hypersurface ff is called an improper affine hypersphere if S=0S=0, and an affine minimal hypersurface if trS=0\mathrm{tr}\,S=0.

By the Codazzi equation, a well-known integrability condition, the above (h,)(h,\nabla) is a statistical structure on MM, which is called the statistical structure equiaffinely induced by ff. Furthermore, the condition (2) implies that (h,)(h,\nabla) satisfies the apolarity condition: trhK=0\mathrm{tr}_{h}K=0. We have the same formula to (2.6) as the Gauss equation:

R(X,Y)Z=h(Y,Z)SXh(X,Z)SY,X,Y,ZΓ(TM).R^{\nabla}(X,Y)Z=h(Y,Z)SX-h(X,Z)SY,\quad X,Y,Z\in\Gamma(TM).

For more detail, see [9] for example.

Example 2.10.

Set Ω={1,,n+1}\Omega=\{1,\ldots,n+1\}, and 𝒮n\mathcal{S}_{n} as the set of all the positive probability functions on Ω\Omega, that is,

𝒮n={p:Ω(0,1)|ω=1n+1p(ω)=1}.\mathcal{S}_{n}=\left\{p:\Omega\to(0,1)\middle|\quad\sum_{\omega=1}^{n+1}p(\omega)=1\right\}.

The set 𝒮n\mathcal{S}_{n} is a smooth nn-dimensional manifold with an atlas consisted of a chart ϕ:𝒮np(p(1),,p(n))n\phi:\mathcal{S}_{n}\ni p\mapsto(p(1),\ldots,p(n))\in\mathbb{R}^{n}. Denote the coordinate system on 𝒮n\mathcal{S}_{n} derived from ϕ\phi by (η1,,ηn)(\eta_{1},\ldots,\eta_{n}). The Fisher information metric gFg^{F} on 𝒮n\mathcal{S}_{n} is defined as

gpF(ηi,ηj)=ω=1n+1p(ω)ηiln(p(ω))ηjln(p(ω))=δijp(i)+1p(n+1),p𝒮n.\begin{split}g_{p}^{F}\left(\frac{\partial}{\partial\eta^{i}},\frac{\partial}{\partial\eta^{j}}\right)&=\sum_{\omega=1}^{n+1}p(\omega)\frac{\partial}{\partial\eta^{i}}\mathrm{ln}(p(\omega))\frac{\partial}{\partial\eta^{j}}\mathrm{ln}(p(\omega))\\ &=\frac{\delta_{ij}}{p(i)}+\frac{1}{p(n+1)},\quad p\in\mathcal{S}_{n}.\end{split}

The Riemannian metric gFg^{F} has constant curvature 1/41/4. The dual metric of gFg^{F} can be determined by the following formula.

gpF(dηi,dηj)=p(i)p(j)+δijp(i).g_{p}^{F}(d\eta^{i},d\eta^{j})=-p(i)p(j)+\delta_{ij}p(i).

The flat affine connection (m)\nabla^{(m)} defined by (m)ηi=0\nabla^{(m)}\frac{\partial}{\partial\eta^{i}}=0 on 𝒮n\mathcal{S}_{n}, is often called the mixture connection. The pair (gF,(m))(g^{F},\nabla^{(m)}) is a statistical structure on 𝒮n\mathcal{S}_{n}, and the conjugate connection of (m)\nabla^{(m)} with respect to gFg^{F} is denoted by (e)\nabla^{(e)}, often called the exponential connection.

On the statistical manifold (𝒮n,gF,(e))(\mathcal{S}_{n},g^{F},\nabla^{(e)}), K=(e)gFK=\nabla^{(e)}-\nabla^{g^{F}} satisfies the following formulas for each p𝒮np\in\mathcal{S}_{n}:

gpF(K(ηi,ηj),ηk)=12(δijδjkp(i)21p(n+1)2),\displaystyle g_{p}^{F}\left(K\left(\frac{\partial}{\partial\eta^{i}},\frac{\partial}{\partial\eta^{j}}\right),\frac{\partial}{\partial\eta^{k}}\right)=-\frac{1}{2}\left(\frac{\delta_{ij}\delta_{jk}}{p(i)^{2}}-\frac{1}{p(n+1)^{2}}\right),
trgKp=12i=1n((n+1)p(i)1)ηip,\displaystyle\mathrm{tr}_{g}K_{p}=\frac{1}{2}\sum_{i=1}^{n}\left((n+1)p(i)-1\right)\frac{\partial}{\partial\eta^{i}}_{p},
divg(trgK)(p)=14(n21+i=1n+11p(i)).\displaystyle\operatorname{div}^{g}(\mathrm{tr}_{g}K)(p)=\frac{1}{4}\left(n^{2}-1+\sum_{i=1}^{n+1}\frac{1}{p(i)}\right).

2.2 Connection Laplacians of statistical manifolds

Definition 2.11.

Let (M,g,)(M,g,\nabla) be a statistical manifold, EME\to M a vector bundle with connection E:Γ(E)Γ(ETM)\nabla^{E}:\Gamma(E)\to\Gamma(E\otimes T^{*}M). We define the operator ΔE=Δ(g,,E):Γ(E)Γ(E)\Delta^{E}=\Delta^{(g,\nabla,\nabla^{E})}:\Gamma(E)\to\Gamma(E) as

ΔEξ=trg{(X,Y)XEYEξXYEξ},\Delta^{E}\xi=\mathrm{tr}_{g}\{(X,Y)\mapsto\nabla^{E}_{X}\nabla^{E}_{Y}\xi-\nabla^{E}_{\nabla_{X}Y}\xi\},

and call it the statistical connection Laplacian with respect to (g,)(g,\nabla) and E\nabla^{E}.

In the case where (g,)(g,\nabla) is Riemannian, it is known as the connection Laplacian or the rough Laplacian with respect to gg as well.

Let ,\langle\cdot,\cdot\rangle be a fiber metric on EE. We have a connection ¯iE\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}} on EE such that

Xξ,η=XEξ,η+ξ,¯XiEηX\langle\xi,\eta\rangle=\langle\nabla^{E}_{X}\xi,\eta\rangle+\langle\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{X}\eta\rangle

for XΓ(TM)X\in\Gamma(TM) and ξ,ηΓ(E)\xi,\eta\in\Gamma(E). The symbol Δ¯E\bar{\Delta}^{E} denotes the statistical connection Laplacian with respect to (g,¯)(g,\overline{\nabla}) and ¯iE\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}.

Example 2.12.

Let (M,g,)(M,g,\nabla) be a statistical manifold of dimension mm. Take an orthonormal basis {ei}i=1,,m\{e_{i}\}_{i=1,\ldots,m} with respect to gg.

(1)  For a trivial bundle E=M×E=M\times\mathbb{R}, the statistical connection Laplacian Δ=ΔE:C(M)C(M)\Delta=\Delta^{E}:C^{\infty}(M)\to C^{\infty}(M) is given as

Δf\displaystyle\Delta f =\displaystyle= i=1m{ei(eif)(eiei)f}\displaystyle\sum_{i=1}^{m}\{e_{i}(e_{i}f)-(\nabla_{e_{i}}e_{i})f\} (2.7)
=\displaystyle= i=1m{ei(eif)(eigei)f}i=1mK(ei,ei)f\displaystyle\sum_{i=1}^{m}\{e_{i}(e_{i}f)-(\nabla^{g}_{e_{i}}e_{i})f\}-\sum_{i=1}^{m}K(e_{i},e_{i})f
=\displaystyle= Δgf(trgK)f\displaystyle\Delta_{g}f-(\mathrm{tr}_{g}K)f

for fC(M)f\in C^{\infty}(M), where Δg\Delta_{g} is the Laplacian with respect to gg, which is the standard tool in Riemannian geometry. Furthermore, we have Δ¯f=Δgf+(trgK)f\bar{\Delta}f=\Delta_{g}f+(\mathrm{tr}_{g}K)f.

(2)  Let NN be a manifold with an affine connection N\nabla^{N}. For a map u:MNu:M\to N, we denote by u\nabla^{u} the connection on u1TNu^{-1}TN induced from N\nabla^{N} by uu: XuU=uXNUΓ(u1TN)\nabla^{u}_{X}U=\nabla^{N}_{u_{*}X}U\in\Gamma(u^{-1}TN) for XΓ(TM)X\in\Gamma(TM) and UΓ(u1TN)U\in\Gamma(u^{-1}TN). The symbol Δu=Δu1TN:Γ(u1TN)Γ(u1TN)\Delta^{u}=\Delta^{u^{-1}TN}:\Gamma(u^{-1}TN)\to\Gamma(u^{-1}TN) denotes the statistical connection Laplacian with respect to (g,)(g,{\nabla}) and u\nabla^{u}.

We now review divergences and Green’s formula for later use.

For a volume form θ\theta on MM, the θ\theta-divergence divθ:Γ(TM)C(M)\operatorname{div}^{\theta}:\Gamma(TM)\to C^{\infty}(M) is defined by

divθXθ=Xθ,XΓ(TM),\operatorname{div}^{\theta}X\theta=\mathcal{L}_{X}\theta,\quad X\in\Gamma(TM),

where \mathcal{L} is the Lie derivative on MM. On the other hand, for an affine connection \nabla on MM, the \nabla-divergence div\operatorname{div}^{\nabla} is defined by

divX=trX,XΓ(TM).\operatorname{div}^{\nabla}X=\mathrm{tr}\nabla X,\quad X\in\Gamma(TM).

If θ\theta is parallel with respect to \nabla, that is, θ=0\nabla\theta=0, then

divθ=div.\operatorname{div}^{\theta}=\operatorname{div}^{\nabla}.

For a Riemannian metric gg, we will denote the g\nabla^{g}-divergence by divg\operatorname{div}^{g}.

The statistical Laplacian has the formula of

Δf=div¯(gradgf),\Delta f=\operatorname{div}^{\overline{\nabla}}(\mathrm{grad}_{g}f),

where the gradient vector field gradgf\mathrm{grad}_{g}f of ff with respect to gg is defined by g(gradgf,X)=df(X)g(\mathrm{grad}_{g}f,X)=df(X) for XΓ(TM)X\in\Gamma(TM).

The following facts are well known. See [7] for example.

Proposition 2.13.

(1) For any fC(M)f\in C^{\infty}(M) and XΓ(TM)X\in\Gamma(TM), we have

div(fX)=Xf+fdivX.\operatorname{div}^{\nabla}(fX)=Xf+f\operatorname{div}^{\nabla}X. (2.8)

(2) (Green’s formula)  If MM is compact, for any XΓ(TM)X\in\Gamma(TM) we have

MdivθXθ=0.\int_{M}\operatorname{div}^{\theta}X\theta=0.

Besides, if XΓ(TM)X\in\Gamma(TM) satisfies divθX=0\operatorname{div}^{\theta}X=0, then for any fC(M)f\in C^{\infty}(M) we have

MXfθ=0.\int_{M}\,Xf\,\theta=0.

Using these facts, we have the following formula which is well-known in the case where a statistical manifold is Riemannian.

Proposition 2.14.

Let (M,g,)(M,g,\nabla) be a compact statistical manifold. Let EE be a vector bundle over MM with a connection E\nabla^{E} and a fiber metric ,\langle\cdot,\cdot\rangle. The formula

MΔEξ,η𝑑μg=Mξ,Δ¯Eη𝑑μg+Mdivg(trgK)ξ,η𝑑μg\int_{M}\langle\Delta^{E}\xi,\eta\rangle d\mu_{g}=\int_{M}\langle\xi,\bar{\Delta}^{E}\eta\rangle d\mu_{g}+\int_{M}\operatorname{div}^{g}(\mathrm{tr}_{g}K)\langle\xi,\eta\rangle d\mu_{g} (2.9)

holds for any ξ,ηΓ(E)\xi,\eta\in\Gamma(E).

Proof.

We set vector fields as

X=i=1meiEξ,ηei,Y=i=1mξ,¯eiiEηei,Z=ξ,ηtrgK,X=\sum_{i=1}^{m}\langle\nabla_{e_{i}}^{E}\xi,\eta\rangle e_{i},\quad Y=\sum_{i=1}^{m}\langle\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{e_{i}}\eta\rangle e_{i},\quad Z=\langle\xi,\eta\rangle\mathrm{tr}_{g}K,

where {ei}i=1,,m\{e_{i}\}_{i=1,\ldots,m} is an orthonormal basis with respect to gg. Then, we have

divg(XYZ)=ΔEξ,ηξ,Δ¯Eηdivg(trgK)ξ,η,\operatorname{div}^{g}(X-Y-Z)=\langle\Delta^{E}\xi,\eta\rangle-\langle\xi,\bar{\Delta}^{E}\eta\rangle-\operatorname{div}^{g}(\mathrm{tr}_{g}K)\langle\xi,\eta\rangle, (2.10)

and Green’s formula implies (2.9). The calculation is as follows.

divgX\displaystyle\operatorname{div}^{g}X =\displaystyle= i=1mg(eigX,ei)\displaystyle\sum_{i=1}^{m}g(\nabla^{g}_{e_{i}}X,e_{i}) (2.11)
=\displaystyle= i=1meieiEξ,η+i,j=1mejEξ,ηg(eigej,ei)\displaystyle\sum_{i=1}^{m}e_{i}\langle\nabla_{e_{i}}^{E}\xi,\eta\rangle+\sum_{i,j=1}^{m}\langle\nabla_{e_{j}}^{E}\xi,\eta\rangle g(\nabla^{g}_{e_{i}}e_{j},e_{i})
=\displaystyle= i=1m{eiEeiEξ,η+eiEξ,¯eiiEη}\displaystyle\sum_{i=1}^{m}\{\langle\nabla^{E}_{e_{i}}\nabla_{e_{i}}^{E}\xi,\eta\rangle+\langle\nabla_{e_{i}}^{E}\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{e_{i}}\eta\rangle\}
i,j=1mi,j=1mejEξ,ηg(eigei,ej)\displaystyle\quad-\sum_{i,j=1}^{m}\sum_{i,j=1}^{m}\langle\nabla_{e_{j}}^{E}\xi,\eta\rangle g(\nabla^{g}_{e_{i}}e_{i},e_{j})
=\displaystyle= i=1m{eiEeiEξ,η+eiEξ,¯eiiEη}i=1meigeiEξ,η\displaystyle\sum_{i=1}^{m}\{\langle\nabla^{E}_{e_{i}}\nabla_{e_{i}}^{E}\xi,\eta\rangle+\langle\nabla_{e_{i}}^{E}\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{e_{i}}\eta\rangle\}-\sum_{i=1}^{m}\langle\nabla_{\nabla^{g}_{e_{i}}e_{i}}^{E}\xi,\eta\rangle
=\displaystyle= i=1m{eiEeiEξ,η+eiEξ,¯eiiEη}\displaystyle\sum_{i=1}^{m}\{\langle\nabla^{E}_{e_{i}}\nabla_{e_{i}}^{E}\xi,\eta\rangle+\langle\nabla_{e_{i}}^{E}\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{e_{i}}\eta\rangle\}
i=1meieiEξ,η+trgKEξ,η\displaystyle\quad-\sum_{i=1}^{m}\langle\nabla_{\nabla_{e_{i}}e_{i}}^{E}\xi,\eta\rangle+\langle\nabla_{\mathrm{tr}_{g}K}^{E}\xi,\eta\rangle
=\displaystyle= ΔEξ,η+trgKEξ,η+i=1meiEξ,¯eiiEη.\displaystyle\langle\Delta^{E}\xi,\eta\rangle+\langle\nabla^{E}_{\mathrm{tr}_{g}K}\xi,\eta\rangle+\sum_{i=1}^{m}\langle\nabla_{e_{i}}^{E}\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{e_{i}}\eta\rangle.

Similarly, we have

divgY=ξ,Δ¯Eηξ,¯trgKiEη+i=1meiEξ,¯eiiEη.\operatorname{div}^{g}Y=\langle\xi,\bar{\Delta}^{E}\eta\rangle-\langle\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{\mathrm{tr}_{g}K}\eta\rangle+\sum_{i=1}^{m}\langle\nabla^{E}_{e_{i}}\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{e_{i}}\eta\rangle. (2.12)

Lastly, by (2.8) we have

divgZ=trgKξ,η+ξ,ηdivg(trgK)=trgKEξ,η+ξ,¯trgKiEη+ξ,ηdivg(trgK).\begin{split}\operatorname{div}^{g}Z&=\mathrm{tr}_{g}K\langle\xi,\eta\rangle+\langle\xi,\eta\rangle\operatorname{div}^{g}(\mathrm{tr}_{g}K)\\ &=\langle\nabla^{E}_{\mathrm{tr}_{g}K}\xi,\eta\rangle+\langle\xi,\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{\mathrm{tr}_{g}K}\eta\rangle+\langle\xi,\eta\rangle\operatorname{div}^{g}(\mathrm{tr}_{g}K).\end{split} (2.13)

By combining (2.11)(\ref{divX}), (2.12)(\ref{divY}), and (2.13)(\ref{divZ}), we obtain the desired (2.10)(\ref{lapfor0}). ∎

Let EE be a vector bundle over a statistical manifold (M,g,)(M,g,\nabla) with a fiber metric ,\langle\cdot,\cdot\rangle. For a connection E:Γ(E)Γ(ETM)\nabla^{E}:\Gamma(E)\to\Gamma(E\otimes T^{*}M), the adjoint (E):Γ(ETM)Γ(E)(\nabla^{E})^{*}:\Gamma(E\otimes T^{*}M)\to\Gamma(E) is defined by

(E)=i=1m{¯eiiE((ei))(¯eiei)}(\nabla^{E})^{*}\mathcal{E}=-\sum_{i=1}^{m}\left\{\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}}_{e_{i}}(\mathcal{E}(e_{i}))-\mathcal{E}(\overline{\nabla}_{e_{i}}{e_{i}})\right\}

for Γ(ETM)\mathcal{E}\in\Gamma(E\otimes T^{*}M).

We remark that the fiber metric of ETME\otimes T^{*}M is naturally derived from gg and ,\langle\cdot,\cdot\rangle, and we denote it by ,\langle\cdot,\cdot\rangle again. Using it, we write the adjointness of E\nabla^{E} and (E)(\nabla^{E})^{*} as follows.

Proposition 2.15 ([11]).

Let (M,g,),E,,,E(M,g,\nabla),E,\langle\cdot,\cdot\rangle,\nabla^{E} be as above.

(1)  Suppose that MM is compact and there is a \nabla-parallel volume form θ\theta on MM. The formula

M,Eηθ=M(E),ηθ\int_{M}\langle\mathcal{E},\nabla^{E}\eta\rangle\theta=\int_{M}\langle(\nabla^{E})^{*}\mathcal{E},\eta\rangle\theta

holds for any ηΓ(E)\eta\in\Gamma(E) and Γ(ETM)\mathcal{E}\in\Gamma(E\otimes T^{*}M).

(2)  We have

ΔEξ=(¯iE)Eξ,ξΓ(E).\Delta^{E}\xi=-(\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ E\end{subarray}})^{*}\nabla^{E}\xi,\quad\xi\in\Gamma(E).

We remark that our sign convention is different from [11] for example. Moreover, the statistical connection Laplacian ΔE\Delta^{E} is apart from \triangle in [5] as well.

3 The first variation formula of the statistical bi-energy

Let u:MNu:M\to N be a smooth map from a compact statistical manifold (M,g,M)(M,g,\nabla^{M}) to a statistical manifold (N,h,N)(N,h,\nabla^{N}), and let τ2(u)Γ(u1TN)\tau_{2}(u)\in\Gamma(u^{-1}TN) be the statistical bi-tension field for uu given in (1). A smooth variation F={ut}t(ϵ,ϵ)F=\{u_{t}\}_{t\in(-\epsilon,\epsilon)} of uu is a smooth map F:M×(ϵ,ϵ)NF:M\times(-\epsilon,\epsilon)\to N such that F(,0)=u0=uF(\cdot,0)=u_{0}=u. The smooth map FF yields a vector field VΓ(u1TN)V\in\Gamma(u^{-1}TN) by

V(x)=ddt|t=0ut(x),xM,V(x)=\left.\frac{d}{dt}\right|_{t=0}u_{t}(x),\quad x\in M,

called the variation vector field of FF. Conversely, it is known that if we take any VΓ(u1TN)V\in\Gamma(u^{-1}TN), there exists a smooth variation of uu that generates VV. In this section, we give the first variation formula of the statistical bi-energy E2E_{2} for smooth maps between statistical manifolds, defined in (1).

Theorem 3.1.

Let u:MNu:M\to N be a smooth map from a compact statistical manifold (M,g,M)(M,g,\nabla^{M}) to a statistical manifold (N,h,N)(N,h,\nabla^{N}). For an arbitrary smooth variation F={ut}t(ϵ,ϵ)F=\{u_{t}\}_{t\in(-\epsilon,\epsilon)} generating VV, the first variation formula is

ddt|t=0E2(ut)=MV,τ2(u)𝑑μg.\left.\displaystyle\frac{d}{dt}\right|_{t=0}E_{2}(u_{t})=\int_{M}\left\langle V,\tau_{2}(u)\right\rangle d\mu_{g}. (3.1)
Proof.

We take the standard Euclidean metric and its Levi-Civita connection on (ϵ,ϵ)(-\epsilon,\epsilon)\subset\mathbb{R}. On the product statistical manifold B=M×(ϵ,ϵ)B=M\times(-\epsilon,\epsilon), we denote the induced statistical connection by B\nabla^{B}. Identifying XΓ(TM)X\in\Gamma(TM) with (X,0)Γ(TB)Γ(TMT(ϵ,ϵ))(X,0)\in\Gamma(TB)\simeq\Gamma(TM\oplus T(-\epsilon,\epsilon)), we have

XBY=XMY,XBt=tBX=0,tBt=0\nabla^{B}_{X}Y=\nabla^{M}_{X}Y,\quad\nabla^{B}_{X}\frac{\partial}{\partial t}=\nabla^{B}_{\frac{\partial}{\partial t}}X=0,\quad\nabla^{B}_{\frac{\partial}{\partial t}}\frac{\partial}{\partial t}=0 (3.2)

for any X,YΓ(TM)X,Y\in\Gamma(TM).

Let F\nabla^{F} be the connection on Γ(F1TN)\Gamma(F^{-1}TN) induced from N\nabla^{N}, and set E=F1TNTBE=F^{-1}TN\otimes T^{*}B. We remark that E\nabla^{E} is given by

(XEΦ)(Y)=XF(ΦY)ΦXBYΓ(FTN)(\nabla^{E}_{X}\Phi)(Y)=\nabla^{F}_{X}(\Phi Y)-\Phi\nabla^{B}_{X}Y\in\Gamma(F^{*}TN)

for X,YΓ(TB)X,Y\in\Gamma(TB) and ΦΓ(E)\Phi\in\Gamma(E). The curvature tensor field RER^{E} of E\nabla^{E} is written by

RE(X,Y)Φ(Z)=((E)2Φ)(Z;Y;X)((E)2Φ)(Z;X;Y)R^{E}(X,Y)\Phi(Z)=\left((\nabla^{E})^{2}\Phi\right)(Z;Y;X)-\left((\nabla^{E})^{2}\Phi\right)(Z;X;Y) (3.3)

for X,Y,ZΓ(TB)X,Y,Z\in\Gamma(TB) and ΦΓ(E)\Phi\in\Gamma(E). Besides, it holds that

RE(X,Y)dF(Z)=RN(FX,FY)FZF(RB(X,Y)Z)R^{E}(X,Y)dF(Z)=R^{N}\left(F_{*}X,F_{*}Y\right)F_{*}Z-F_{*}\left(R^{B}(X,Y)Z\right) (3.4)

for X,Y,ZΓ(TB)X,Y,Z\in\Gamma(TB), where RBR^{B} and RNR^{N} are the curvature tensor fields of B\nabla^{B} and N\nabla^{N}, respectively.

Using (2.1), we have

ddtE2(ut)=12Mtτ(ut)h2𝑑μg=Mh(tFτ(ut),τ(ut))𝑑μg+12M(FtNh)(τ(ut),τ(ut))𝑑μg=Mh(tFτ(ut),τ(ut))𝑑μgMh(KN(τ(ut),τ(ut)),Ft)𝑑μg.\begin{split}\displaystyle\frac{d}{dt}E_{2}(u_{t})&=\displaystyle\frac{1}{2}\int_{M}\frac{\partial}{\partial t}||\tau(u_{t})||_{h}^{2}d\mu_{g}\\ &=\int_{M}h\left(\nabla^{F}_{\frac{\partial}{\partial t}}\tau(u_{t}),\tau(u_{t})\right)d\mu_{g}+\displaystyle\frac{1}{2}\int_{M}(\nabla^{N}_{F_{*}\frac{\partial}{\partial t}}h)\Bigl{(}\tau(u_{t}),\tau(u_{t})\Bigr{)}d\mu_{g}\\ &=\int_{M}h\left(\nabla^{F}_{\frac{\partial}{\partial t}}\tau(u_{t}),\tau(u_{t})\right)d\mu_{g}-\int_{M}h\left(K^{N}\bigl{(}\tau(u_{t}),\tau(u_{t})\bigr{)},F_{*}\frac{\partial}{\partial t}\right)d\mu_{g}.\end{split}

Taking an orthonormal basis {ei}i=1,,m\{e_{i}\}_{i=1,\ldots,m} with respect to gg, we have, by (3.3) and by (3.4)(\ref{curvtens}) and (3.2)(\ref{conF}),

tFτ(ut)=i=1mtF((eiEdF)(ei))=i=1m((E)2dF)(ei;ei;t)=i=1m((E)2dF)(ei;t;ei)+i=1m(RE(t,ei)dF)(ei)=i=1m((E)2dF)(ei;t;ei)+i=1mRN(Ft,F(ei))F(ei)=Δut(Ft)+i=1mRN(Ft,F(ei))F(ei).\begin{split}\nabla^{F}_{\frac{\partial}{\partial t}}\tau(u_{t})&=\sum_{i=1}^{m}\nabla^{F}_{\frac{\partial}{\partial t}}\left((\nabla^{E}_{e_{i}}dF)(e_{i})\right)\\ &=\sum_{i=1}^{m}\left((\nabla^{E})^{2}dF\right)\left(e_{i};e_{i};\frac{\partial}{\partial t}\right)\\ &=\sum_{i=1}^{m}\left((\nabla^{E}\right)^{2}dF)\left(e_{i};\frac{\partial}{\partial t};e_{i}\right)+\sum_{i=1}^{m}\left(R^{E}\left(\frac{\partial}{\partial t},e_{i}\right)dF\right)(e_{i})\\ &=\sum_{i=1}^{m}\left((\nabla^{E}\right)^{2}dF)\left(e_{i};\frac{\partial}{\partial t};e_{i}\right)+\sum_{i=1}^{m}{R^{N}}\left(F_{*}\frac{\partial}{\partial t},F_{*}(e_{i})\right)F_{*}(e_{i})\\ &=\Delta^{u_{t}}\left(F_{*}\frac{\partial}{\partial t}\right)+\sum_{i=1}^{m}{R^{N}}\left(F_{*}\frac{\partial}{\partial t},F_{*}(e_{i})\right)F_{*}(e_{i}).\end{split} (3.5)

Indeed, the last equality is obtained by (3.2)(\ref{conF}) as follows:

i=1m((E)2dF)(ei;t;ei)=i=1m(eiF((tEdF)(ei))(tEdF)(eiBei))=i=1m(eiuteiut(Ft)eiMeiut(Ft))=Δut(Ft).\begin{split}\sum_{i=1}^{m}\left((\nabla^{E})^{2}dF\right)\left(e_{i};\frac{\partial}{\partial t};e_{i}\right)&=\sum_{i=1}^{m}\left(\nabla^{F}_{e_{i}}\left((\nabla^{E}_{\frac{\partial}{\partial t}}dF)(e_{i})\right)-(\nabla^{E}_{\frac{\partial}{\partial t}}dF)(\nabla^{B}_{e_{i}}{e_{i}})\right)\\ &=\sum_{i=1}^{m}\left(\nabla^{u_{t}}_{e_{i}}\nabla^{u_{t}}_{e_{i}}\left(F_{*}\frac{\partial}{\partial t}\right)-\nabla^{u_{t}}_{\nabla^{M}_{e_{i}}{e_{i}}}\left(F_{*}\frac{\partial}{\partial t}\right)\right)\\ &=\Delta^{u_{t}}\left(F_{*}\frac{\partial}{\partial t}\right).\end{split}

By Proposition 2.14, we have

Mh(Δut(Ft),τ(ut))𝑑μg=Mh(Ft,Δ¯utτ(ut)+divg(trgK)τ(ut))𝑑μg.\begin{split}&\int_{M}h\left(\Delta^{u_{t}}\left(F_{*}\frac{\partial}{\partial t}\right),\tau(u_{t})\right)d\mu_{g}\\ &=\int_{M}h\left(F_{*}\frac{\partial}{\partial t},\bar{\Delta}^{u_{t}}\tau(u_{t})+\operatorname{div}^{g}(\mathrm{tr}_{g}K)\tau(u_{t})\right)d\mu_{g}.\end{split}

By combining them and remembering V=Ft|t=0V=\left.F_{*}\frac{\partial}{\partial t}\right|_{t=0}, we have the first variation formula:

ddt|t=0E2(ut)\displaystyle\left.\displaystyle\frac{d}{dt}\right|_{t=0}E_{2}(u_{t}) =\displaystyle= MV,Δ¯uτ(u)+divg(trgKM)τ(u)\displaystyle\int_{M}\langle\,V,\,\bar{\Delta}^{u}\tau(u)+\operatorname{div}^{g}(\mathrm{tr}_{g}K^{M})\tau(u)
i=1mLN(uei,τ(u))ueiKN(τ(u),τ(u))dμg.\displaystyle\qquad\qquad-\sum_{i=1}^{m}L^{N}(u_{*}e_{i},\tau(u))u_{*}e_{i}-K^{N}(\tau(u),\tau(u))\,\rangle d\mu_{g}.

4 Examples of statistical biharmonic maps

As stated in §1, we call a map uu statistical biharmonic if the statistical bi-tension field of uu vanishes identically. Under suitable assumptions, we will rewrite the statistical bi-tension field.

Let u:(M,g,M)(N,h,N)u:(M,g,\nabla^{M})\to(N,h,\nabla^{N}) be a smooth map between statistical manifolds. Suppose that MM is trace-free and NN is conjugate symmetric. Then the statistical bi-tension field of uu is given as

τ2(u)=Δ¯uτ(u)i=1mRN(uei,τ(u))ueiKN(τ(u),τ(u)),\tau_{2}(u)=\bar{\Delta}^{u}\tau(u)-\sum_{i=1}^{m}R^{N}(u_{*}e_{i},\tau(u))u_{*}e_{i}-K^{N}(\tau(u),\tau(u)), (4.1)

where {ej}j=1,,m\{e_{j}\}_{j=1,\ldots,m} is an orthonormal basis with respect to gg.

Example 4.1.

(1)  Let c:(M,g,M)(N,h,N)c:(M,g,\nabla^{M})\to(N,h,\nabla^{N}) be a smooth map between statistical manifolds. Suppose that (M,g,M)=((a,b),g0,g0)(M,g,\nabla^{M})=\left((a,b),g_{0},\nabla^{g_{0}}\right), where g0g_{0} is the Euclidean metric on (a,b)(a,b)\subset\mathbb{R}. Since τ(c)=c˙Nc˙\tau(c)=\nabla^{N}_{\dot{c}}\dot{c}, where c˙=cddt\dot{c}=c_{*}\dfrac{d}{dt}, the statistical bi-tension field of cc is given as

τ2(c)=¯c˙iN¯c˙iNc˙Nc˙LN(c˙,c˙Nc˙)c˙KN(c˙Nc˙,c˙Nc˙).\tau_{2}(c)=\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ N\end{subarray}}_{\dot{c}}\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ N\end{subarray}}_{\dot{c}}\nabla^{N}_{\dot{c}}\dot{c}-L^{N}(\dot{c},\nabla^{N}_{\dot{c}}\dot{c})\dot{c}-K^{N}\left(\nabla^{N}_{\dot{c}}\dot{c},\nabla^{N}_{\dot{c}}\dot{c}\right).

Accordingly, any N\nabla^{N}-geodesic cc is a statistical biharmonic map.

(2)  Take the statistical manifold (2,g0,)(\mathbb{R}^{2},g_{0},\nabla) in Example 2.8 as (N,h,N)(N,h,\nabla^{N}) in (1). For any real numbers a,b,ca,b,c, and dd, the curve c(t)=(at+c,bt+d)2c(t)=(at+c,bt+d)\in\mathbb{R}^{2}, tt\in\mathbb{R}, satisfies τ2(c)=0\tau_{2}(c)=0. Here, if a0a\neq 0 or b0b\neq 0, then τ(c)0\tau(c)\neq 0.

Example 4.2.

(1)  Let f:(M,g,M)(N,h,N)f:(M,g,\nabla^{M})\to(N,h,\nabla^{N}) be a smooth map between statistical manifolds. Suppose that (N,h,N)=(,g0,g0)(N,h,\nabla^{N})=(\mathbb{R},g_{0},\nabla^{g_{0}}). Since τ(f)=Δf\tau(f)=\Delta f, the statistical bi-tension field of ff is given as

τ2(f)=Δ¯Δf+divg(trgK)Δf,\tau_{2}(f)=\bar{\Delta}\Delta f+\operatorname{div}^{g}(\mathrm{tr}_{g}K)\Delta f,

where Δ=Δ(g,,g0)\Delta=\Delta^{(g,\nabla,\nabla^{g_{0}})} and Δ¯=Δ(g,¯,g0)\bar{\Delta}=\Delta^{(g,\overline{\nabla},\nabla^{g_{0}})}.

(2)  If we take the statistical manifold (2,g0,)(\mathbb{R}^{2},g_{0},\nabla) in Example 2.8, we have

τ2(f)=Δ¯Δf={(2x2+2y2)2(x+y)2}f.\tau_{2}(f)=\bar{\Delta}\Delta f=\left\{\left(\frac{\partial^{2}}{\partial x^{2}}+\frac{\partial^{2}}{\partial y^{2}}\right)^{2}-\left(\frac{\partial}{\partial x}+\frac{\partial}{\partial y}\right)^{2}\right\}f.

The function f(x,y)=sinhx+coshy,(x,y)2f(x,y)=\sinh{x}+\cosh{y},~{}(x,y)\in\mathbb{R}^{2}, satisfies τ2(f)=0\tau_{2}(f)=0, while τ(f)\tau(f) is not globally equal to 0.

Lemma 4.3.

Suppose that (M,g,)(M,g,\nabla) is a compact statistical manifold with divg(trgK)=0\operatorname{div}^{g}(\mathrm{tr}_{g}K)=0. For fC(M)f\in C^{\infty}(M), if Δf\Delta f is constant, then so is ff.

Proof.

We put XΓ(TM)X\in\Gamma(TM) as X=ΔfgradgfX=\Delta f\,\mathrm{grad}_{g}f. By (2.7) and (2.8), we have

divgX=ΔfΔgf=(Δf)2+ΔftrgKf=(Δf)2+divg(fΔftrgK).\begin{split}\operatorname{div}^{g}X&=\Delta f\Delta_{g}f\\ &=(\Delta f)^{2}+\Delta f\mathrm{tr}_{g}Kf\\ &=(\Delta f)^{2}+\operatorname{div}^{g}(f\Delta f\mathrm{tr}_{g}K).\end{split}

By Green’s formula, we have

0=M(Δf)2𝑑μg,0=\int_{M}(\Delta f)^{2}d\mu_{g},

thus we obtain Δf=0\Delta f=0.

We remark that Δgf2=2fΔgf+2dfg2\Delta_{g}f^{2}=2f\Delta_{g}f+2\|df\|_{g}^{2}. By (2.7) and Green’s formula again, we have

0=MfΔf𝑑μg=MfΔgf𝑑μgMftrgKf𝑑μg=Mdfg2𝑑μg12MtrgKf2𝑑μg=Mdfg2𝑑μg,\begin{split}0&=\int_{M}f\Delta fd\mu_{g}\\ &=\int_{M}f\Delta_{g}fd\mu_{g}-\int_{M}f\mathrm{tr}_{g}Kfd\mu_{g}\\ &=-\int_{M}\|df\|_{g}^{2}d\mu_{g}-\frac{1}{2}\int_{M}\mathrm{tr}_{g}Kf^{2}d\mu_{g}\\ &=-\int_{M}\|df\|_{g}^{2}d\mu_{g},\end{split}

therefore df=0df=0 on MM. ∎

Proposition 4.4.

Suppose that (M,g,)(M,g,\nabla) is a compact statistical manifold with divg(trgK)=0\operatorname{div}^{g}(\mathrm{tr}_{g}K)=0. A function f:(M,g,)(,g0,g0)f:(M,g,\nabla)\to(\mathbb{R},g_{0},\nabla^{g_{0}}) is statistical biharmonic if and only if ff is constant.

Proof.

Since divg(trgK)=0\operatorname{div}^{g}(\mathrm{tr}_{g}K)=0, we have

τ2(f)=Δ¯Δf=ΔgΔf+trgKΔf.\tau_{2}(f)=\bar{\Delta}\Delta f=\Delta_{g}\Delta f+\mathrm{tr}_{g}K\Delta f.

By integration, we obtain

0=M(ΔgΔf+trgKΔf)Δf𝑑μg=M{dΔfg2+12Δg(Δf)2+12trgK(Δf)2}𝑑μg=MdΔfg2𝑑μg,\begin{split}0&=\int_{M}(\Delta_{g}\Delta f+\mathrm{tr}_{g}K\Delta f)\Delta fd\mu_{g}\\ &=\int_{M}\left\{-\|d\Delta f\|_{g}^{2}+\frac{1}{2}\Delta_{g}(\Delta f)^{2}+\frac{1}{2}\mathrm{tr}_{g}K(\Delta f)^{2}\right\}d\mu_{g}\\ &=-\int_{M}\|d\Delta f\|_{g}^{2}d\mu_{g},\end{split}

hence, Δf\Delta f is constant. From Lemma 4.3, we conclude that ff is a constant function. ∎

Theorem 4.5.

Let f:Mm+1f:M\to\mathbb{R}^{m+1} be a locally strongly convex hypersurface with affine shape operator SS, and (h,)(h,\nabla) the statistical structure equiaffinely induced by ff as in Example 2.9. Consider ff as a map from a statistical manifold (M,h,)(M,h,\nabla) to a statistical manifold (m+1,g0,g0)(\mathbb{R}^{m+1},g_{0},\nabla^{g_{0}}) for the Euclidean metric g0g_{0} compatible with the equiaffine structure of m+1\mathbb{R}^{m+1}. Then, τ(f)\tau(f) does not vanish. Moreover, ff is statistical biharmonic if and only if (1) trS=0\mathrm{tr}\,S=0 and (2) trhS=0\mathrm{tr}_{h}\nabla S=0 hold.

In particular, an improper affine hypersphere is a statistical biharmonic map.

Proof.

Let denote g0\nabla^{g_{0}} by DD for short. For f:(M,h,)(m+1,g0,D)f:(M,h,\nabla)\to(\mathbb{R}^{m+1},g_{0},D),

τ(f)=trh{(X,Y)DXfYfXY}=trh{(X,Y)h(X,Y)ξ}=mξ,\tau(f)=\mathrm{tr}_{h}\{(X,Y)\mapsto D_{X}f_{*}Y-f_{*}\nabla_{X}Y\}=\mathrm{tr}_{h}\{(X,Y)\mapsto h(X,Y)\xi\}=m\xi,

which does not vanish. On the other hand, since trhK\mathrm{tr}_{h}K vanishes, by (4.1) the statistical bi-tension field τ2(f)\tau_{2}(f) of ff is obtained as

1mτ2(f)=i=1m(DeiDeiξD¯eieiξ)=i=1m(DeifSei+fS¯eiei)=i=1m(feiSeih(ei,Sei)ξ+fSeiei)=f(trhS)trSξ.\begin{split}\frac{1}{m}\tau_{2}(f)&=\sum_{i=1}^{m}\left(D_{e_{i}}D_{e_{i}}\xi-D_{\overline{\nabla}_{e_{i}}e_{i}}\xi\right)\\ &=\sum_{i=1}^{m}\left(-D_{e_{i}}f_{*}Se_{i}+f_{*}S\overline{\nabla}_{e_{i}}e_{i}\right)\\ &=\sum_{i=1}^{m}\left(-f_{*}\nabla_{e_{i}}Se_{i}-h(e_{i},Se_{i})\xi+f_{*}S\nabla_{e_{i}}e_{i}\right)\\ &=-f_{*}\left(\mathrm{tr}_{h}\nabla S\right)-\mathrm{tr}S\,\xi.\end{split}

Therefore, τ2(f)\tau_{2}(f) vanishes if and only if (1) and (2) hold. ∎

5 Reduction properties of statistical biharmonic maps

Let (M,g,M)(M,g,\nabla^{M}), (N,h,N)(N,h,\nabla^{N}) be statistical manifolds, and u:MNu:M\to N a smooth map with τ(u)Γ(u1TN)\tau(u)\in\Gamma(u^{-1}TN), which is defined for uu by using (g,M)(g,\nabla^{M}) and N\nabla^{N} as in (1).

Remark that if (g,M)(g,\nabla^{M}) and (h,N)(h,\nabla^{N}) are Riemannian, τ(u)\tau(u) is called the tension field of uu, and a map u:MNu:M\to N satisfying τ(u)=0\tau(u)=0 is classically called a harmonic map. In the case where M\nabla^{M} is flat, and (h,N)(h,\nabla^{N}) is Riemannian, a map u:MNu:M\to N satisfying τ(u)=0\tau(u)=0 is studied as an affine harmonic map in [6].

Remark 5.1.

Let (M,g,M)(M,g,\nabla^{M}) a compact statistical manifold equipped with a ¯iM\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ M\end{subarray}}-parallel volume form θ\theta, and (N,h,N)(N,h,\nabla^{N}) be a Riemannian one. By Proposition 2.15, the equation τ(u)=0\tau(u)=0 holds if and only if the map u:MNu:M\to N is a critical point of the following functional:

E(u)=12Mdug,h2θ.E(u)=\frac{1}{2}\int_{M}\|du\|_{g,h}^{2}\,\theta.

In the following, we observe some settings in which τ2(u)=0\tau_{2}(u)=0 implies τ(u)=0\tau(u)=0.

Lemma 5.2.

Let (M,g,M)(M,g,\nabla^{M}), (N,h,N)(N,h,\nabla^{N}) be statistical manifolds. Assume that MM is compact, and that there exists a ¯iM\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ M\end{subarray}}-parallel volume form θ\theta on MM. For a map u:MNu:M\to N, if ¯iNτ(u)\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ N\end{subarray}}\tau(u) vanishes, then so does τ(u)\tau(u).

Proof.

We put XΓ(TM)X\in\Gamma(TM) as

X=i=1mh(uei,τ(u))ei,X=\sum_{i=1}^{m}h\bigl{(}u_{*}e_{i},\tau(u)\bigr{)}e_{i},

and obtain

divθX=i=1mg(¯eiiMX,ei)=i=1meih(uei,τ(u))+i,j=1mh(uej,τ(u))g(¯eiiMej,ei)=i=1mh(eiNuei,τ(u))i=1mh(u(eiMei),τ(u))=h(τ(u),τ(u)).\begin{split}\operatorname{div}^{\theta}X&=\sum_{i=1}^{m}g(\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ M\end{subarray}}_{e_{i}}X,e_{i})\\ &=\sum_{i=1}^{m}e_{i}h\bigl{(}u_{*}e_{i},\tau(u)\bigr{)}+\sum_{i,j=1}^{m}h\left(u_{*}e_{j},\tau(u)\right)g\left(\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ M\end{subarray}}_{e_{i}}{e_{j}},e_{i}\right)\\ &=\sum_{i=1}^{m}h\bigl{(}\nabla^{N}_{e_{i}}u_{*}e_{i},\tau(u)\bigr{)}-\sum_{i=1}^{m}h\left(u_{*}(\nabla^{M}_{e_{i}}{e_{i}}),\tau(u)\right)\\ &=h(\tau(u),\tau(u)).\end{split}

Since θ\theta is ¯iM\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ M\end{subarray}}-parallel, by Green’s formula for the integration by θ\theta, we have

0=Mτ(u)h2θ,0=\int_{M}\|\tau(u)\|_{h}^{2}\,\theta,

which concludes τ(u)=0\tau(u)=0. ∎

Proposition 5.3.

Let (M,g,M)(M,g,\nabla^{M}), (N,h,N)(N,h,\nabla^{N}) be statistical manifolds. For (M,g,M)(M,g,\nabla^{M}), assume that MM is compact, there exists a ¯iM\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ M\end{subarray}}-parallel volume form θ\theta on MM, and that divg(trgKM)=0\operatorname{div}^{g}(\mathrm{tr}_{g}K^{M})=0 on MM. For (N,h,N)(N,h,\nabla^{N}), assume that (N,h,N)(N,h,\nabla^{N}) is conjugate symmetric, and that the UNU^{N}-sectional curvature is non-positive on (N,h)(N,h), that is, for unit tangent vectors X,YX,Y on (N,h)(N,h),

h(UN(X,Y)Y,X)0,h\left(U^{N}(X,Y)Y,X\right)\leq 0,

where UN=2RhRNU^{N}=2R^{h}-R^{N}. If a statistical biharmonic map u:(M,g,M)(N,h,N)u:(M,g,\nabla^{M})\to(N,h,\nabla^{N}) satisfies KN(τ(u),τ(u))=0K^{N}\bigl{(}\tau(u),\tau(u)\bigr{)}=0, then τ(u)=0\tau(u)=0.

The above UNU^{N} is found as an UN=2Rh(RN+R¯iN)/2Γ(TN(1,3))U^{N}=2R^{h}-(R^{N}+\overline{R}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ N\end{subarray}})/2\in\Gamma(TN^{(1,3)}) in [3], in which the sectional curvature is defined by using UNU^{N} even if (h,N)(h,\nabla^{N}) is not conjugate symmetric. This proposition is proved in [4] in the case where (g,M)(g,\nabla^{M}) and (h,N)(h,\nabla^{N}) are Riemannian.

Proof.

Using the condition KN(τ(u),τ(u))=0K^{N}\bigl{(}\tau(u),\tau(u)\bigr{)}=0 and the conjugate symmetricity of (h,N)(h,\nabla^{N}), we have

Δgτ(u)h2\displaystyle\Delta_{g}\,\|\tau(u)\|_{h}^{2}
=h(Δuτ(u),τ(u))+h(τ(u),Δ¯uτ(u))\displaystyle\quad=h\left(\Delta^{u}\tau(u),\tau(u)\right)+h\left(\tau(u),\bar{\Delta}^{u}\tau(u)\right)
+i=1m{4h((RNRh)(uei,τ(u))τ(u),uei)\displaystyle\qquad+\sum_{i=1}^{m}\left\{4h\left(\left(R^{N}-R^{h}\right)\bigl{(}u_{*}e_{i},\tau(u)\bigr{)}\tau(u),u_{*}e_{i}\right)\right.
+h(eiuτ(u),eiuτ(u))+h(¯eiiuτ(u),¯eiiuτ(u))}.\displaystyle\qquad\qquad\left.+h(\nabla^{u}_{e_{i}}\tau(u),\nabla^{u}_{e_{i}}\tau(u))+h(\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ u\end{subarray}}_{e_{i}}\tau(u),\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ u\end{subarray}}_{e_{i}}\tau(u))\right\}.

By Green’s theorem, Proposition 2.14, and by the condition divg(trgKM)=0\operatorname{div}^{g}(\mathrm{tr}_{g}K^{M})=0, the conjugate symmetricity of (h,N)(h,\nabla^{N}) and τ2(u)=0\tau_{2}(u)=0, we have

0\displaystyle 0 =\displaystyle= MΔgτ(u)h2𝑑μg\displaystyle\int_{M}\Delta_{g}\,\|\tau(u)\|_{h}^{2}\,d\mu_{g}
=\displaystyle= 2Mh(Δuτ(u),τ(u))𝑑μg\displaystyle 2\int_{M}h\left(\Delta^{u}\tau(u),\tau(u)\right)d\mu_{g}
+4i=1mMh((RNRh)(uei,τ(u))τ(u),uei)𝑑μg\displaystyle\quad+4\sum_{i=1}^{m}\int_{M}h\left(\left(R^{N}-R^{h}\right)\bigl{(}u_{*}e_{i},\tau(u)\bigr{)}\tau(u),u_{*}e_{i}\right)d\mu_{g}
+M{uτ(u)g,h2+¯iuτ(u)g,h2}𝑑μg\displaystyle\qquad+\int_{M}\{\|\nabla^{u}\tau(u)\|^{2}_{g,h}+\|\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ u\end{subarray}}\tau(u)\|^{2}_{g,h}\}d\mu_{g}
=\displaystyle= 2i=1mMh(UN(uei,τ(u))τ(u),uei)𝑑μg\displaystyle-2\sum_{i=1}^{m}\int_{M}h\left(U^{N}(u_{*}e_{i},\tau(u))\tau(u),u_{*}e_{i}\right)d\mu_{g}
+M{uτ(u)g,h2+¯iuτ(u)g,h2}𝑑μg.\displaystyle\quad+\int_{M}\{\|\nabla^{u}\tau(u)\|^{2}_{g,h}+\|\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ u\end{subarray}}\tau(u)\|^{2}_{g,h}\}d\mu_{g}.

Since the sectional curvature of UNU^{N} is non-positive we have

uτ(u)=¯iuτ(u)=0.\nabla^{u}\tau(u)=\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ u\end{subarray}}\tau(u)=0.

Therefore, Lemma 5.2 completes the proof. ∎

The following proposition states a local property in contrast to the previous ones. In the Riemannian setting it is obtained by [10].

Proposition 5.4.

Let u:(M,g,M)(N,h,N)u:(M,g,\nabla^{M})\to(N,h,\nabla^{N}) be a statistical hypersurface, that is, (g,M)(g,\nabla^{M}) is the statistical structure on MM induced by uu from (h,N)(h,\nabla^{N}), and m=n1m=n-1. Suppose that (N,h,N)(N,h,\nabla^{N}) is a trace-free and satisfies

RicNRic¯iN2Rich0.\operatorname{Ric}^{N}-\overline{\operatorname{Ric}}^{\begin{subarray}{c}\scalebox{0.2}{\phantom{i}}\\ N\end{subarray}}-2\operatorname{Ric}^{h}\geq 0. (5.1)

If uu is a biharmonic map such that KN(τ(u),τ(u))=0K^{N}\bigl{(}\tau(u),\tau(u)\bigr{)}=0, and τ(u)h\|\tau(u)\|_{h} is constant on MM, then τ(u)=0\tau(u)=0.

Proof.

Assume that τ(u)0\tau(u)\neq 0.

We take an orthonormal basis {ue1,,uem,τ(u)h1τ(u)}\{u_{*}e_{1},\ldots,u_{*}e_{m},\|\tau(u)\|_{h}^{-1}\tau(u)\} for hh, remarking that uu is an isometric immersion. Since (h,N)(h,\nabla^{N}) is trace-free,

trgKM\displaystyle\mathrm{tr}_{g}K^{M} =\displaystyle= i=1mh(uKM(ej,ej),uei)ei\displaystyle\sum_{i=1}^{m}h(u_{*}K^{M}(e_{j},e_{j}),u_{*}e_{i})e_{i}
=\displaystyle= i=1mτ(u)h2h(KN(τ(u),τ(u)),uei)ei=0.\displaystyle-\sum_{i=1}^{m}\|\tau(u)\|^{-2}_{h}h\left(K^{N}\bigl{(}\tau(u),\tau(u)\bigr{)},u_{*}e_{i}\right)e_{i}=0.

By using trgKM=0\mathrm{tr}_{g}K^{M}=0 and KN(τ(u),τ(u))=0K^{N}\bigl{(}\tau(u),\tau(u)\bigr{)}=0 again, we calculate the statistical Laplacian of τ(u)h2\|\tau(u)\|^{2}_{h} as

12Δτ(u)h2\displaystyle\frac{1}{2}\Delta\|\tau(u)\|^{2}_{h} =\displaystyle= h(Δ¯uτ(u),τ(u))+uτ(u),¯iuτ(u)g,h\displaystyle h\left(\bar{\Delta}^{u}\tau(u),\tau(u)\right)+\left\langle\nabla^{u}\tau(u),\overline{\nabla}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ u\end{subarray}}\tau(u)\right\rangle_{g,h} (5.2)
=\displaystyle= h(Δ¯uτ(u),τ(u))+^uτ(u)g,h2i=1mKN(uei,τ(u))h2\displaystyle h\left(\bar{\Delta}^{u}\tau(u),\tau(u)\right)+\|\widehat{\nabla}^{u}\tau(u)\|^{2}_{g,h}-\sum_{i=1}^{m}\|K^{N}\bigl{(}u_{*}e_{i},\tau(u)\bigr{)}\|^{2}_{h}
=\displaystyle= 12{RicN(τ(u),τ(u))Ric¯iN(τ(u),τ(u))2Rich(τ(u),τ(u))}\displaystyle\dfrac{1}{2}\left\{\operatorname{Ric}^{N}\bigl{(}\tau(u),\tau(u)\bigr{)}-\overline{\operatorname{Ric}}^{\begin{subarray}{c}\scalebox{0.2}{\phantom{i}}\\ N\end{subarray}}\bigl{(}\tau(u),\tau(u)\bigr{)}-2\operatorname{Ric}^{h}\bigl{(}\tau(u),\tau(u)\bigr{)}\right\}
+^uτ(u)g,h2,\displaystyle\qquad+\|\widehat{\nabla}^{u}\tau(u)\|^{2}_{g,h},

where ^u\widehat{\nabla}^{u} is the connection on u1TNu^{-1}TN induced by uu from ^=h\widehat{\nabla}=\nabla^{h}. Indeed, the last equality is obtained as follows. For the third term, by (2.4), we have

i=1mKN(uei,τ(u))h2=i=1mh(Kτ(u)NKueiNτ(u),uei)=i=1mh([Kτ(u)N,KueiN]τ(u),uei)=12RicN(τ(u),τ(u))12Ric¯iN(τ(u),τ(u))+Rich(τ(u),τ(u)),\begin{split}\sum_{i=1}^{m}\|K^{N}\bigl{(}u_{*}e_{i},\tau(u)\bigr{)}\|^{2}_{h}&=\sum_{i=1}^{m}h\left(K^{N}_{\tau(u)}K^{N}_{u_{*}e_{i}}\tau(u),u_{*}e_{i}\right)\\ &=\sum_{i=1}^{m}h\left([K^{N}_{\tau(u)},K^{N}_{u_{*}e_{i}}]\tau(u),u_{*}e_{i}\right)\\ &=-\frac{1}{2}\operatorname{Ric}^{N}\left(\tau(u),\tau(u)\right)-\frac{1}{2}\overline{\operatorname{Ric}}^{\begin{subarray}{c}\scalebox{0.2}{\phantom{i}}\\ N\end{subarray}}\left(\tau(u),\tau(u)\right)\\ &\quad+\operatorname{Ric}^{h}\left(\tau(u),\tau(u)\right),\end{split}

and for the first term, since τ2(u)=0\tau_{2}(u)=0,

h(Δ¯uτ(u),τ(u))=i=1mh(LN(uei,τ(u))uei,τ(u))=i=1mh(RN(τ(u),uei)uei,τ(u))=i=1mh(R¯iN(uei,τ(u))τ(u),uei)=Ric¯iN(τ(u),τ(u)),\begin{split}h\left(\bar{\Delta}^{u}\tau(u),\tau(u)\right)&=\sum_{i=1}^{m}h\left(L^{N}(u_{*}e_{i},\tau(u))u_{*}e_{i},\tau(u)\right)\\ &=-\sum_{i=1}^{m}h\left(R^{N}(\tau(u),u_{*}e_{i})u_{*}e_{i},\tau(u)\right)\\ &=-\sum_{i=1}^{m}h\left(\overline{R}^{\begin{subarray}{c}\scalebox{0.3}{\phantom{i}}\\ N\end{subarray}}(u_{*}e_{i},\tau(u))\tau(u),u_{*}e_{i}\right)\\ &=-\overline{\operatorname{Ric}}^{\begin{subarray}{c}\scalebox{0.2}{\phantom{i}}\\ N\end{subarray}}\bigl{(}\tau(u),\tau(u)\bigr{)},\end{split}

from which we have (5.2).

By (5.2) and (5.1)(\ref{ricciinq}) we obtain ^uτ(u)=0\widehat{\nabla}^{u}\tau(u)=0 since τ(u)\|\tau(u)\| is constant.

We will show that τ(u)h=0\|\tau(u)\|_{h}=0, which contradicts that hh is a positive definite metric, which completes the proof.

0=i=1mh(uei,^eiuτ(u))=i=1mh(^eiuuei,τ(u))=h(τ(u),τ(u))+i=1mh(KN(uei,uei),τ(u))=τ(u)h2τ(u)h2h(KN(τ(u),τ(u)),τ(u))=τ(u)h2.\begin{split}0&=\sum_{i=1}^{m}h\left(u_{*}e_{i},\widehat{\nabla}^{u}_{e_{i}}\tau(u)\right)\\ &=-\sum_{i=1}^{m}h\left(\widehat{\nabla}^{u}_{e_{i}}u_{*}e_{i},\tau(u)\right)\\ &=-h\bigl{(}\tau(u),\tau(u)\bigr{)}+\sum_{i=1}^{m}h\left(K^{N}\bigl{(}u_{*}e_{i},u_{*}e_{i}\bigr{)},\tau(u)\right)\\ &=-\|\tau(u)\|_{h}^{2}-\|\tau(u)\|^{-2}_{h}h(K^{N}\bigl{(}\tau(u),\tau(u)\bigr{)},\tau(u))\\ &=-\|\tau(u)\|_{h}^{2}.\end{split}

Statements and Declarations

Acknowledgments

\bmhead

Funding This work was supported by the Japan Society for the Promotion of Science KAKENHI, Grant Number JP22K03279, and by the Japan Science and Technology agency, the establishment of university fellowships towards the creation of science technology innovation, Grant Number JPMJFS2101.

References

  • [1] Baues, O., Cortés, V.: Realisation of special Kähler manifolds as parabolic spheres. Proc. Amer. Math. Soc. 129(8), 2403-2407 (2001)
  • [2] Cheng, B.-Y.: Some open problems and conjectures on submanifolds of finite type. Soochow J. Math. 17(2), 169-188 (1991)
  • [3] Furuhata, H., Hasegawa I., Satoh N.: Chen invariants and statistical submanifolds. Commun. Korean Math. Soc. 37(3), 851-864 (2022)
  • [4] Jiang, G.Y.: 2-Harmonic maps and their first and second variational formulas. Translated from the Chinese by Hajime Urakawa. Note Mat. 28, 209-232 (2009)
  • [5] Jiang, R., Tavakoli, J., Zhao, Y.: Laplacian operator on statistical manifold. Inf. Geom. 6(1), 69-79 (2023)
  • [6] Jost, J., Şimşir, F.M.: Affine harmonic maps. Analysis (Munich). 29(2), 185-197 (2009)
  • [7] Kobayashi, S., Nomizu, K.: Foundations of differential geometry. Vol. I John Wiley & Sons, Inc., New York (1963)
  • [8] Liu, H.L., Wang, C.P.: The centroaffine Tchebychev operator. Results Math. 27(1-2), 77-92 (1995)
  • [9] Nomizu, K., Sasaki, T.: Affine differential geometry. Cambridge University Press, Cambridge (1994)
  • [10] Oniciuc, C.: Biharmonic maps between Riemannian manifolds. An. Ştiinţ. Univ. Al. I. Cuza Iaşi. Mat. (N.S.) 48(2), 237-248 (2002)
  • [11] Opozda, B.: Bochner’s technique for statistical structures. Ann. Global Anal. Geom. 48(4), 357-395 (2015)
  • [12] Şimşir, F.M.: A note on harmonic maps of statistical manifolds. Commun. Fac. Sci. Univ. Ank. Ser. A1. Math. Stat. 68(2), 1370-1376 (2019)
  • [13] Uohashi, K.: Harmonic maps relative to α\alpha-connections. In: Nielsen, F. (ed.) Geometric theory of information, pp. 81-96. Springer Cham.(2014)
  • [14] Urakawa, H.: Geometry of biharmonic mappings. World Scientific Publishing Co. Pte. Ltd., Hackensack, N.J. (2019)