
After Thomas James Willmore, many authors were looking for an immersion of a manifold in Euclidean space or Riemannian manifold, which is the critical point of functionals whose integrands depend on the mean curvature or the norm of the second fundamental form. We study a new Willmore-type variational problem for a foliated hypersurface in Euclidean space. Its general version is the Reilly-type functional, where the integrand depends on elementary symmetric functions of the eigenvalues of the restriction on the leaves of the second fundamental form. We find the 1st and 2nd variations of such functionals and show the conformal invariance of some of them. For a critical hypersurface with a transversally harmonic foliation, we derive the Euler-Lagrange equation and give examples with low-dimensional foliations. We present critical hypersurfaces of revolution and show that they are local minima for special variations of immersion.
Citation: Vladimir Rovenski. Willmore-type variational problem for foliated hypersurfaces[J]. Electronic Research Archive, 2024, 32(6): 4025-4042. doi: 10.3934/era.2024181
[1] | Fanqi Zeng . Some almost-Schur type inequalities and applications on sub-static manifolds. Electronic Research Archive, 2022, 30(8): 2860-2870. doi: 10.3934/era.2022145 |
[2] | Youngjin Hwang, Ildoo Kim, Soobin Kwak, Seokjun Ham, Sangkwon Kim, Junseok Kim . Unconditionally stable monte carlo simulation for solving the multi-dimensional Allen–Cahn equation. Electronic Research Archive, 2023, 31(8): 5104-5123. doi: 10.3934/era.2023261 |
[3] | Liu Yang, Yuehuan Zhu . Second main theorem for holomorphic curves on annuli with arbitrary families of hypersurfaces. Electronic Research Archive, 2024, 32(2): 1365-1379. doi: 10.3934/era.2024063 |
[4] | Víctor León, Bruno Scárdua . A geometric-analytic study of linear differential equations of order two. Electronic Research Archive, 2021, 29(2): 2101-2127. doi: 10.3934/era.2020107 |
[5] | Frédéric Campana . Algebraicity of foliations on complex projective manifolds, applications. Electronic Research Archive, 2022, 30(4): 1187-1208. doi: 10.3934/era.2022063 |
[6] | Huifu Xia, Yunfei Peng . Pointwise Jacobson type necessary conditions for optimal control problems governed by impulsive differential systems. Electronic Research Archive, 2024, 32(3): 2075-2098. doi: 10.3934/era.2024094 |
[7] | Marco G. Ghimenti, Anna Maria Micheletti . Compactness and blow up results for doubly perturbed Yamabe problems on manifolds with non umbilic boundary. Electronic Research Archive, 2022, 30(4): 1209-1235. doi: 10.3934/era.2022064 |
[8] | Li Du, Xiaoqin Yuan . The minimality of biharmonic hypersurfaces in pseudo-Euclidean spaces. Electronic Research Archive, 2023, 31(3): 1587-1595. doi: 10.3934/era.2023081 |
[9] | Nurlan A. Abiev . The Ricci curvature and the normalized Ricci flow on the Stiefel manifolds SO(n)/SO(n−2). Electronic Research Archive, 2025, 33(3): 1858-1874. doi: 10.3934/era.2025084 |
[10] | Juan Gerardo Alcázar, Carlos Hermoso, Hüsnü Anıl Çoban, Uğur Gözütok . Computation of symmetries of rational surfaces. Electronic Research Archive, 2024, 32(11): 6087-6108. doi: 10.3934/era.2024282 |
After Thomas James Willmore, many authors were looking for an immersion of a manifold in Euclidean space or Riemannian manifold, which is the critical point of functionals whose integrands depend on the mean curvature or the norm of the second fundamental form. We study a new Willmore-type variational problem for a foliated hypersurface in Euclidean space. Its general version is the Reilly-type functional, where the integrand depends on elementary symmetric functions of the eigenvalues of the restriction on the leaves of the second fundamental form. We find the 1st and 2nd variations of such functionals and show the conformal invariance of some of them. For a critical hypersurface with a transversally harmonic foliation, we derive the Euler-Lagrange equation and give examples with low-dimensional foliations. We present critical hypersurfaces of revolution and show that they are local minima for special variations of immersion.
Many authors, e.g., [1,2,3,4,5], were looking for an immersion ϕ:Mn→ˉMn+1 of a smooth manifold Mn (n≥2) into a Riemannian manifold (ˉM,ˉg) or Euclidean space Rn+1, which is a critical point of the following functionals for compactly supported variations of ϕ:
Wn,p=∫MHpdV,Jn,p=∫M‖h‖pdV. | (1.1) |
Here, dV is the volume form of the induced metric g=<⋅,⋅> on M, h is the scalar second fundamental form of ϕ(M), H=1ntracegh is the mean curvature, and p>0. These functionals measure how much ϕ(M) differs from a minimal hypersurface (H=0) or a totally geodesic hypersurface (h=0). The actions (1.1) are a particular case of functionals
WFn=∫MF(H)dV,JFn=∫MF(‖h‖)dV, |
where F is a C3-regular function of one variable, e.g., [6,7,8,9]. For a closed smooth hypersurface Mn in Rn+1, we get Wn,n≥Cn, where Cn=2π(n+1)/2Γ((n+1)/2) is the area of the unit n-sphere; the equality Wn,n=Cn holds if and only if Mn is embedded as a hypersphere; see [1].
Variational problems for (1.1) were first posed by Willmore in [10] for W2,2, which belongs to conformal geometry. The Euler-Lagrange equation for W2,2 is the well known elliptic PDE
ΔH+2H(H2−K)=0, | (1.2) |
where Δ is the Laplacian and K is the gaussian curvature of M2⊂R3. Solutions of (1.2) are called Willmore surfaces. An important class of Willmore surfaces in R3 arises from the stereographic projection of minimal surfaces in the 3-sphere. By Lawson's theorem, any compact, orientable surface can be minimally embedded in the 3-sphere. For a closed orientable surface M in R3, the inequality W2,2≥C2=4π holds with the equality for round spheres. If M2 is a torus in R3, then, according to the Willmore conjecture proven by Marques and Neves in [4], we have W2,2≥2π2; the equality holds if and only if the generating curve is a circle and the ratio of radii is 1√2. Willmore surfaces have applications in biophysics, computer graphics, materials science, architecture, etc., e.g., [11].
Reilly [12] and some mathematicians studied variations of more general functionals than (1.1):
WFn=∫MF(σ1,…,σn)dV,JFn=∫MF(τ1,…,τn)dV, | (1.3) |
where F∈C3(Rn). The elementary symmetric functions σr=∑i1<…<irki1…kir (0≤r≤n) of the principal curvatures ki satisfy the equality ∑nr=0σrtr=det(idTM+tA), where A is the Weingarten operator, i.e., <AX,Y>=h(X,Y). The power sums of the principal curvatures, τi=ki1+…+kin=traceAi, can be expressed as polynomials of σr using the Newton formulas, e.g., [13]. For example, σ0=1, τ1=σ1=nH, σn=detA, τ2=‖h‖2, and 2σ2=τ21−τ2. The r-th (r≤n) order Willmore functional, introduced by Guo in [9],
Wconfn,r=∫MQn/rrdV,Qr=∑rj=0(−1)j+1CjrSr−j1Sj, | (1.4) |
is a special case of (1.3), invariant under the conformal group of (ˉM,ˉg) and vanishing on totally umbilical hypersurfaces. Here, Sr=σr/Crn (where Crn=n!r!(n−r)! is a binomial coefficient) is the r-th mean curvature function of a hypersurface. In particular, Q2=S21−S2=1n2(n−1)((n−1)σ21−2nσ2). Examples of hypersurfaces in Rn+1 that are critical for (1.4) are given in [7,8].
An interesting problem is the generalization of the Willmore functional to submanifolds with additional structures, such as foliations or almost products. Let Mn (n≥2) equipped with an s-dimensional (1≤s≤n) foliation F be immersed into a Riemannian manifold (ˉM,ˉg). All leaves of the foliation under consideration have the same dimension. Let hF be the restriction of the second fundamental form of M on the leaves of F. Denote by τFi (1≤i≤s) the power sums, σFr (1≤r≤s) elementary symmetric functions of the eigenvalues kF1≤…≤kFs of hF, and set SFr=σFr/Crs. We have τF1=σF1=sHF=traceghF, τF2=‖hF‖2, (τF1)2−τF2=2σF2, etc. For foliation theory, we refer to [14]. The extrinsic geometry of foliations was developed in [13]. We study Reilly-type functionals for compactly supported variations of (Mn,F) immersed in Rn+1:
WFn,s=∫MF(σF1,…,σFs)dV,JFn,s=∫MF(τF1,…,τFs)dV, | (1.5) |
which, for s=n, reduces to (1.3). For F=(σF1/s)p and F=(τF2)p/2, the actions (1.5) read as
Wn,p,s=∫MHpFdV,Jn,p,s=∫M‖hF‖pdV, | (1.6) |
which reduces to (1.1) for s=n.
Remark 1. A foliated hypersurface in Rn+1, whose leaves {L} are minimal submanifolds in Rn+1 is an example of a minimizer for Wn,p,s in (1.6) with even p. A foliated hypersurface in Rn+1 with an asymptotic distribution TF (e.g., a ruled hypersurface) is a minimizer for Jn,p,s in (1.6). It is interesting to find critical hypersurfaces for actions (1.6) with HF≠0 or hF≠0 on an open dense set of M.
The following special case of (1.5) is invariant under the conformal group of (ˉM,ˉg), see Theorem 1:
Wconfn,s,r=∫M(QFr)n/rdV,QFr=∑rj=0(−1)j+1Cjr(SF1)r−jSFj,r≤s, | (1.7) |
and reduces to (1.4) for s=n. Note that QF2=(SF1)2−SF2=1s2(s−1)((s−1)(σF1)2−2sσF2) is true and QFr=0 if kF1=…=kFs, e.g., for hypersurfaces of revolution in Rn+1 foliated by parallels.
We hope that foliated hypersurfaces, which are local minima for (1.5), will be useful for natural sciences and technology related to layered (laminated) or non-isotropic materials.
The paper is organized as follows: Section 2 contains some lemmas that help us calculate variations of Reilly-type functionals on foliated hypersurfaces in Rn+1. In Section 3, conformal invariance of (1.7) is shown, the first variations of the functionals (1.5)–(1.7) are found, and the corresponding Euler-Lagrange equations for the case of transversally harmonic (for example, Riemannian) foliation are obtained. Then the second variation on critical hypersurfaces of some Willmore-type functionals is calculated. In Section 4, applications to hypersurfaces with low-dimensional foliations are given, the critical hypersurfaces of revolution for the actions (1.6) are presented, and it is shown that they are local minima for special variations of immersion.
Let r:Mn→Rn+1 be an immersion of a manifold M into Rn+1 with Euclidean metric ˉg and the Levi–Civita connection ˉ∇. We identify M with its image r(M) and restrict our calculations to a relatively compact neighborhood U⊂M with induced metric g=<⋅,⋅> and normal coordinates (x1,…,xn) centered at a point x∈M. Thus, gij=δij (the Kronecker symbol) and Γkij=0 at x. We will denote differentiation of a function f (or a tensor) with respect to the variable xi by fi.
Let ∂i=∂/∂xi be the coordinate vector fields on U. So, the vectors ri=ˉ∇∂ir form a local coordinate basis for the tangent bundle TM along U, and we get g=gijdxidxj, where gij=ˉg(ri,rj)=<∂i,∂j> and the Einstein summation rule is used. Let N be a unit normal vector field to M on U. The vectors Ni=ˉ∇∂iN belong to the tangent space TxM, i.e., <Ni,N>=0.
Let h be the scalar second fundamental form of M with respect to unit normal N, A=−ˉ∇N the Weingarten operator, and H=1ntracegh the mean curvature. Denote by hj the symmetric tensor dual to Aj, i.e., hj(X,Y)=<AjX,Y>. For example, h2=gklhlihkjdxidxj=hkihkjdxidxj.
Consider a one-parameter family of hypersurfaces rt=r+tuN (|t|<1). We get a variation δr=uN, where δ=(d/dt)|t=0 is the variational derivative operator, and u:U→R is a smooth function supported on a relatively compact neighborhood U. Obviously, (δr)i=uiN+uNi. The Hessian of a function u is a (0, 2)-tensor (Hessu)(X,Y)=X(Y(u))−(∇XY)u=(uij−Γkijuk)dxidxj; see [15], where Γkij are the Christoffel symbols. The Laplacian is Δu=tracegHessu=gijuij. Note that <g,Hessu>=Δu. The divergence of a vector field X=Xi∂i on M is divX=∇iXi.
Lemma 1 (see [3] for n=2). The following evolution equations are true:
δgij=−2uhij, | (2.1) |
δgij=2uhij, | (2.2) |
δhij=uij−uhlihjl ⇔ δh=Hessu−uh2, | (2.3) |
δ‖h‖2=2<h, uh2+Hessu>, | (2.4) |
δ(nH)=Δu+u‖h‖2, | (2.5) |
δdV=−nuHdV. | (2.6) |
Proof. Using δri=uiN+uNi, we calculate
<δri,rj>=<uiN+uNi,rj>=u<Ni,rj>=−u<N,rij>=−uhij. |
Thus, since the symmetry hij=hji we get the equality (2.1): δgij=<δri,rj>+<ri,δrj>=−2uhij. From gilglj=δij, it follows that (δgil)glj=−gil(δglj)=2ugilhlj; hence, (2.2) is true.
We will compute the variation of h. Using <N,Ni>=0, we find
<N, δrij>=<N, (uN)ij>=uij−u<Ni, Nj>=uij−u<hlirl, hkjrk>=uij−uhlihjl. |
Note that δN=ciri for some functions ci. Using <N,rj>=0, we get
gijci=<δN, rj>=−<N, δrj>=−<N, ujN+uNj>=−uj. |
It follows that ci=−gijuj and δN=−gijujri=−uiri. Using the Gauss equation for a hypersurface in Rn+1, we get at x: <δN, rij>=<−uiri, hjlN+Γkjlrk>=0. Thus, (2.3) is true:
δhij=δ<rij, N>=<δN, rij>+<N, δrij>=uij−uhlihjl. |
Calculating the variation of the mean curvature, we get (2.5):
δ(nH)=δ(gijhij)=2uhijhij+gij(uij−uhlihjl)=uhijhij+gijuij=Δu+u‖h‖2. |
The formula δ(dV)=12(tracegδg)dV for variation of dV is valid for any variation δg of a metric, for example, [13]. Applying (2.1) to the above gives (2.6). Next, we calculate the variation of ‖h‖2:
δ‖h‖2=δ(gikgjlhklhij)=2u(hikgjl+hjlgik)hklhij+gikgjl((ukl−uhqkhlq)hij+(uij−uhqihjq)hkl)=2(uhijhikhjk+uklgikgjl)=2u<h, h2>+2<h,Hessu>, |
that proves (2.4).
We carry out further calculations for a foliated hypersurface (M,F) and a foliated neighborhood U⊂M with normal coordinates (x1,…,xn) adapted to F, i.e., (x1,…,xs) are variables along the leaves; see [14]. Let ∇F:TM×TF→TF be the induced Levi–Civita connection on the leaves of F. The leafwise Laplacian on functions is ΔF=tracegHessF=divF∘∇F, where HessF is the Hessian on the leaves of F. Let P:TM→TF be the orthoprojector, thus, P2=P and P is self-adjoint. For hF and its dual self-adjoint operator AF, we can write hF(X,Y)=h(PX,PY) (X,Y∈XM) and AF=PAP. Let hjF be the symmetric tensor dual to AjF, i.e., hjF(X,Y)=<AjFX,Y>. The symmetric tensor hmix is given by
hmix(X,Y)=12(h(PX,Y)+h(X,PY))−h(PX,PY),X,Y∈XM. |
We have <hF,hmix>=0. The equality hmix=0 means that PA=AP, i.e., TF is an invariant subbundle for A. Let hF⊥ be the restriction of h on the normal distribution to F in M; then h2F⊥=gγϵhαγhαϵdxαdxβ, where s<α,β,γ,ϵ≤n. Define symmetric tensors h2mix=gαβhαihβjdxidxj+gijhαihβjdxαdxβ and Hessmixu=gijgαβuiαdxjdxβ, where 1≤i,j≤s and s<α,β,γ,ϵ≤n. Let Amix be the (1, 1)-tensor dual to hmix; then A2mix is dual to h2mix.
The Newton transformations Tr(AF) of AF are defined inductively or explicitly by, e.g., [13],
T0(AF)=idTF,Tr(AF)=σFridTF−AFTr−1(AF)(0<r≤s),Tr(AF)=∑rj=0(−1)jσFr−jAjF=σFridTF−σFr−1AF+…+(−1)rArF. |
For example, T1(AF)=σF1idTF−AF and Ts(AF)=0, and the following equalities are true:
traceTr(AF)=(s−r)σFr,trace(AFTr(AF))=(r+1)σFr+1,trace(A2FTr(AF))=σF1σFr+1−(r+2)σFr+2. | (2.7) |
The "musical" isomorphism ♯:T∗M→TM is used for tensors, e.g., h♯=A, and for (0,2)-tensors B and C, we have <B,C>=trace(B♯C♯)=<B♯,C♯>.
Lemma 2. The variations of τFi and σFr are the following:
1iδτFi=<hi−1F,HessFu>+u(τFi+1+<hi−1F,h2mix>), | (2.8a) |
δσFr=<Tr−1(AF),HessF♯u>+u(σF1σFr−1−(r+1)σFr+1+<Tr−1(AF),A2mix>). | (2.8b) |
Proof. By (2.3), we obtain δAF=HessF♯u+u(A2F+A2mix). Using this, (2.7) and the following variations of τFi and σFr; see [13]:
δτFi=itrace(Ai−1FδAF),δσFr=trace(Tr−1(AF)δAF), |
we get (2.8a,b).
Lemma 3. The following evolution equations are true:
δ(sHF)=ΔFu+u(‖hF‖2−‖hmix‖2), | (2.9) |
δ‖hF‖2=2<hF, u(h2F+h2mix)+HessFu>, | (2.10) |
δ‖hmix‖2=u<hF+hF⊥,h2mix>+2<hmix,Hessmixu>. | (2.11) |
For any smooth function f:M×R→R, the following evolution equation is true:
δ(ΔFf)=ΔF˙f+2u<hF,HessFf>+su<∇FHF,∇Ff>+2h(∇Fu,∇Ff)−sHF<∇Fu,∇Ff>. | (2.12) |
Proof. The Eqs (2.9) and (2.10) can be deduced from Lemma 2, but we will prove them directly. First, using (2.1) and (2.3), we get for 1≤i,j≤s, and 1≤q≤n,
δ(sHF)=δ(gijhij)=2uhijhij+gij(uij−uhqihjq)=ΔFu+u(‖hF‖2−‖hmix‖2), |
that proves (2.9). Also for 1≤i,j,k,l≤s and 1≤q≤n, we obtain (2.10):
δ‖hF‖2=δ(gikgjlhklhij)=4u<hF,h2F>+2<hF,HessFu>−2u<hF,h2F>−2u<hF,h2mix>. |
For 1≤i,j,k,l≤s, s<α,β,γ≤n, and 1≤q≤n, we obtain (2.11):
δ‖hmix‖2=δ(gijgαβhiαhjβ)=2u(hijgαβ+gijhαβ)hiαhjβ+gijgαβ(uiαhjβ+ujβhiα)−ugijgαβ(hlihαlhjβ+hljhαihlβ)=<hF+hF⊥,uh2mix>+2<hmix,Hessmixu>. |
The proof of (2.12) is similar to the proof of (19) in [3]: instead of M2, we consider s-dimensional leaves of F. The variation of the Christoffel symbols is the following tensor, e.g., [3]:
δΓkij=−ugkl(hjl,i+hil,j−hij,l)−gkl(uihjl+ujhil−ulhij). | (2.13) |
For the Laplacian ΔFf=gij(fij−Γkijfk) with 1≤i,j,k≤s it follows that
δ(ΔFf)=δ(gijfij)−δ(gijΓkijfk). | (2.14) |
For the first term, we get (for 1≤i,j≤s)
δ(gijfij)=2uhijfij+gij˙fij=2u<hF,HessFf>+ΔF˙f. | (2.15) |
For the second term, using (2.13) and Γkij=0 at x, we get for 1≤i,j,k≤s, and 1≤q≤n,
δ(gijΓkijfk)=gijδ(Γkij)fk=−gijgkq{u(hjq,i+hiq,j−hij,q)fk−(uihjq+ujhiq−uqhij)fk}=−2ugijgkqhjq,ifk+ugijgkqhij,qfk−2gijgkquihjqfk+gijgkquqhijfk. | (2.16) |
Using the Codazzi–Mainardi equation ∇khij−∇jhik=0, e.g., [15], we get for 1≤i,j,k,l≤s,
gijgkl(∇jhil)fk=gkl(gij∇lhij)fk=s(∇kHF)fk=s<∇FHF,∇Ff>. |
Thus, using normal coordinates and −2ugijgklhjl,ifk+ugijgklhij,lfk=−ugkl(gijhij,l)fk, we get s(HF)l=gij∇lhij for 1≤i,j,l≤s. Therefore, (2.16) becomes
δ(gijΓkijfk)=−ugkl(gijhij,l)fk−2gijgkluihjlfk+gklul(gijhij)fk=−su<∇FHF,∇Ff>−2h(∇Fu,∇Ff)+sHF<∇Fu,∇Ff>. | (2.17) |
Applying (2.15) and (2.17) to (2.14) completes the proof of (2.12).
Remark 2. To find the second variation of the functionals (1.5), we also need the variation δ<h,HessFu>, but we omit this calculation and consider the second variation of the functionals (1.6) only.
The following property helps to find the Euler-Lagrange equations using the first variation of the functionals (1.3)–(1.6):
(divP)∘P=0. | (2.18) |
Here, (divP)X=∑ni=1<(∇eiP)X,ei>, where e1,…,en is a local orthonormal basis on M. Note that Riemannian foliations (and the leaves of twisted products, e.g., [13]) satisfy (2.18).
Lemma 4. A foliated Riemannian manifold (M,g,F) satisfies (2.18) if and only if F is transversally harmonic, i.e., the normal distribution has zero mean curvature.
Proof. Using a local orthonormal frame on M such that ei∈TF (1≤i≤s), we calculate:
(divP)(PX)=∑ni=1<(∇eiP)(PX),ei>=∑ni=1{<∇ei(P2X),ei>−<P∇ei(PX),ei>}=∑ni=1{<∇ei(PX),ei>−<∇ei(PX),Pei>}=∑i>s<∇ei(PX),ei>=−<X,(n−s)H⊥>, |
where (n−s)H⊥=P∑i>s∇eiei is the mean curvature vector of (TF)⊥ and X∈XM.
For any 2-tensor B on M, define the adjoint of the covariant derivative ∇∗B=−∑i(∇iB)(ei,⋅); see [15]. We have the formula ∫M<B,∇B′>dV=∫M<∇∗B,B′>dV; see [15]; thus,
∫M<B,Hessu>dV=∫M<B,∇(∇u)>dV=∫M<∇∗B,∇u>dV=∫Mu(∇∗)2(B)dV. | (2.19) |
The next lemma generalizes (2.19) and the well-known Green's formula, e.g., [15].
Lemma 5. If a foliated Riemannian manifold (M,g,F) satisfies (2.18), then the following formulas are valid for any compactly supported functions u,f, and 2-tensor B:
∫Mf(ΔFu)dV=∫Mu(ΔFf)dV, | (2.20) |
∫M<B,HessFu>dV=∫Mu(∇F∗)2(B)dV. | (2.21) |
Proof. We have ΔFf2=divF(∇Ff2). One can show that divF(PX)=div(PX)−(divP)(PX) for all X∈XM. Hence, using ∇Ff=P∇f and (2.18), we get
f1ΔFf2=f1divF(∇Ff2)=f1{div(P∇f2)−(divP)(P∇f2)}=div(f1P∇f2)−<P∇f1,P∇f2>=div(f1P∇f2)−<∇Ff1,∇Ff2>. |
Using the divergence theorem, gives
∫Mf1(ΔFf2)dV=−∫M⟨∇Ff1,∇Ff2⟩dV. | (2.22) |
By this, the formula (2.20) is true. Next, using (2.18), we will prove
∫M<φ1,∇Fφ2>dV=∫M<∇F∗φ1,φ2>dV | (2.23) |
for any compactly supported (s,t)-tensor φ1 and (s,t+1)-tensor φ2. Define a compactly supported 1-form ω by ω(Y)=<ιYφ2,φ1> for Y∈XM. Take an orthonormal frame (ei) such that ∇Yei=0 for all Y∈TxM and some x∈M. To simplify calculations, assume that s=t=1, and then at x∈M,
−∇∗Fω=∑j(∇Fejω)(ej)=∑i,j<Pej,ei>(∇eiω)(ej)=∑i,j,c<Pej,ei>(<∇eiφ2(ej,ec),φ1(ec)>+<φ2(ej,ec),∇eiφ1(ec)>)=∑i,j,c[<<Pej,ei>∇eiφ2(ej,ec),φ1(ec)>+<φ2(ej,ec),<Pej,ei>∇eiφ1(ec)>]=∑j,c[<∇Fejφ2(ej,ec),φ1(ec)>+<φ2(ej,ec),∇Fejφ1(ec)>]=<φ2,∇Fφ1>−<∇∗Fφ2,φ1>. |
The ∇F∗ is related to the F-divergence of a vector field ω♯ by divFω♯=−∇F∗ω. By the above and ∫M(divF ω♯)dV=∫Mdiv(Pω♯)dV=0, we obtain (2.23). Applying this twice, we get (2.21).
In Section 3.1, we find the Euler-Lagrange equations (and first variations) for (1.5)–(1.7), and in Section 3.2, we find the second variations of (1.5) and (1.6). First, we check the conformity of (1.7).
Theorem 1. The functional Wconfn,s,r is a conformal invariant of a foliated hypersurface (M,F) in a Riemannian manifold (ˉM,ˉg).
Proof. Define a new Riemannian metric on ˉM by ˉgc=μ2ˉg for some positive function μ∈C3(ˉM). Then gc=μ2g is the new induced metric on M; thus, the new volume form of M is dVc=μndV. If X is a ˉg-unit vector, then Xc=X/μ is a gc-unit vector. By the well known formula for the Levi–Civita connection, e.g., [13], we get 2ˉ∇cXY=2ˉ∇XY+μ−2(X(μ2)Y+Y(μ2)X−<X,Y>ˉ∇μ2). By this, with X∈TF and Y=Nc, the operators A and Ac are related by Ac=1μ(A−1μ⟨ˉ∇μ,N⟩idTF), see also [13]. By the above and AF=PAP, AcF=PAcP, we get
AcF=1μ(AF−1μ⟨ˉ∇μ,N⟩idTF),HcF=1straceAcF=1μ(HF−1μ⟨ˉ∇μ,N⟩). |
Set BF=HFidTF−AF. Let λBi be the eigenvalues of BF on F and σBr be the elementary symmetric functions of BF. Obviously, BcF=1μBF holds; hence, λB,ci=1μλBi. One can show that QFr=−σBr/Csr is true; see [9]. By the above, μrQF,cr=QFr holds. Hence, (QFr)n/rdV is a conformal invariant of (M,F) in (ˉM,ˉg): (QF,cr)n/rdVc=(QFr)n/rdV. Note that if AF is a conformal operator on TF (i.e., proportional to idTF), then BF=0, hence, QFr=0.
We can state our main theorem.
Theorem 2. If (2.18) is valid, then Euler-Lagrange equations for the functionals (1.5) are
∑sr=1{(∇F∗)2(F′r⋅Tr−1(AF))+F′r(σF1σFr−1−(r+1)σFr+1+<Tr−1(AF),A2mix>)}−nFH=0, | (3.1a) |
∑si=11i{(∇F∗)2(F′i⋅Ai−1F)+F′i(τFi+1+<hi−1F,h2mix>)}−nFH=0. | (3.1b) |
Proof. Using (2.6), we get the following:
δWFn,s=∫Mδ(F(σF1,…,σFs)dV)=∫M{∑sr=1F′r⋅δσFr−nuFH}dV,δJFn,s=∫Mδ(F(τF1,…,τFs)dV)=∫M{∑si=1F′i⋅δτFi−nuFH}dV. | (3.2) |
From (3.2) and (2.8a,b), we find the first variations of functionals (1.5):
δWFn,s=∫M{∑sr=1<F′r⋅Tr−1(AF),HessF♯u> | (3.3a) |
+u∑sr=1F′r(σF1σFr−1−(r+1)σFr+1+<Tr−1(AF),A2mix>)−nuFH}dV, | (3.3b) |
δJFn,s=∫M{∑si=11iF′i(<hi−1F,HessFu>+u(τFi+1+<hi−1F,h2mix>))−nuFH}dV. | (3.3c) |
From (3.3a,b), using (2.18) and (2.20), we obtain (3.1a,b).
Equations (3.4) and (3.5) of the next statement follow from Theorem 2, but we will prove them.
Corollary 1. If (2.18) is valid, then the Euler-Lagrange equations for the functionals Wn,p,s,Jn,p,s, see (1.6), and Wconfn,s,2, see (1.7), are, respectively, the following:
ΔF(Hp−1F)+Hp−1F(‖hF‖2−‖hmix‖2−nspHHF)=0, | (3.4) |
(∇F∗)2(‖hF‖p−2hF)+‖hF‖p−2(<hF, h2F+h2mix>−np‖hF‖2H)=0, | (3.5) |
ΔF((QF2)n/2−1σF1)−ss−1(∇F∗)2((QF2)n/2−1T1(AF))+{σF1(σF1−2σF2+‖Amix‖2)−ss−1(σF1σF2−3σF3+<T1(AF),A2mix>)−s2QF2H}(QF2)n/2−1=0. | (3.6) |
Proof. Using (2.6), (2.9), and (2.10), we calculate the variation
δ(HpFdV)=Hp−1F{ps(ΔFu+u(‖hF‖2−‖hmix‖2))−nuHHF}dV,δ(‖hF‖pdV)=‖hF‖p−2{p<hF, u(h2F+h2mix)+HessFu>−nu‖hF‖2H}dV. |
Hence, using (2.18), (2.20), (2.21), and (2.6), we find the first variation of the actions (1.6):
δWn,p,s=∫Mδ(HpFdV)=∫MHp−1F{ps(ΔFu+u(‖hF‖2−‖hmix‖2))−nuHFH}dV=ps∫Mu{ΔF(Hp−1F)+Hp−1F(‖hF‖2−‖hmix‖2−nspHHF)}dV, | (3.7) |
δJn,p,s=∫M‖hF‖p−2{p<hF,u(h2F+h2mix)+HessFu>−nu‖hF‖2H}dV=∫Mu{p(∇F∗)2(‖hF‖p−2hF)+p‖hF‖p−2(<hF,h2F+h2mix>−n‖hF‖2H)}dV. | (3.8) |
From (3.7) and (3.8), the Euler-Lagrange equations (3.4) and (3.5) follow. By (2.8b) we get
δσF1=ΔFu+u(σF1−2σF2+‖Amix‖2,δσF2=<T1(AF),HessF♯u>+u(σF1σF2−3σF3+<T1(AF),A2mix>). |
Using QF2=1s2(s−1)((s−1)(σF1)2−2sσF2) and (2.8b) for r=1,2, we get
s2(s−1)δQF2=2(s−1)σF1δσF1−2sδσF2=2(s−1)σF1(ΔFu+u(σF1−2σF2+‖Amix‖2))−2s(<T1(AF),HessF♯u>+u(σF1σF2−3σF3+<T1(AF),A2mix>)). |
Hence
δWconfn,s,2=n2∫M(QF2)n/2−1{1s2(s−1)[2(s−1)σF1(ΔFu+u(σF1−2σF2+‖Amix‖2))−2s(<T1(AF),HessF♯u>+u(σF1σF2−3σF3+<T1(AF),A2mix>))]−2QF2uH}dV. | (3.9) |
Using (2.20) with f=(QF2)n2−1σF1 and (2.21) with B=(QF2)n2−1T1(AF) in (3.9), we get (3.6).
Remark 3. (ⅰ) For a hypersurface M⊂Rn+1 equipped with a line field (i.e., s=1) of the normal curvature κ, the functionals (1.5) and (1.6) coincide with WFn,1=∫MF(κ)dV and Wn,p,1=∫MκpdV. For Wn,2,1 and Jn,2,1, from (3.4) and (3.5) with p=2 and s=1, using (∇F∗)2hF=ΔFκ, we get the following leaf-wise elliptic PDE: ΔFκ+(κ2−‖hmix‖2−n2Hκ)κ=0.
(ⅱ) The first variation of the functional Wconfn.s,r and the Euler-Lagrange equation can be obtained from (3.1a) and (3.3a), similarly to the corresponding equations in [9] for Wconfn.r.
(ⅲ) By (3.7) and (3.8) with s=n, the first variations of functionals (1.1) are given by
δWn,p=∫MHp−1{pn(Δu+u‖h‖2)−nuH2}dV,δJn,p=∫M‖h‖p−2{p<h, uh2+Hessu>−nu‖h‖2H}dV. |
The corresponding Euler-Lagrange equations are well known:
ΔHp−1+Hp−1(‖h‖2−n2pH2)=0, | (3.10) |
(∇∗)2(‖h‖p−2h)+‖h‖p−2(<h,h2>−np‖h‖2H)=0, | (3.11) |
for example, [3], where n=2 and M2⊂R3. For p=n=2, we can use the identity ‖h‖2−2H2=12(k1−k2)2=2(H2−K), where k1 and k2 are the principal curvatures, H=(k1+k2)/2, and K=k1k2 is the gaussian curvature of a surface M2⊂R3. In this case, the Euler-Lagrange equation (3.10) reduces to (1.2). Using the identity <h,h2>=8H3−6HK, the Euler-Lagrange equation (3.11) for p=n=2 reads as (∇∗)2h+4H(H2−K)=0.
The following statement generalizes Corollary 1 in [3] when M2⊂R3.
Theorem 3. If (2.18) is valid, then the Euler-Lagrange equation for WFn,s of (1.5) with F=F(HF) is
ΔFF′+F′(‖hF‖2−‖hmix‖2)−snFH=0. | (3.12) |
At a critical hypersurface satisfying (2.18), the second variation of WFn,s with F=F(HF) is
δ2WFn,s=−∫Mns{F′ΔFu−uΔFF′}uHdV+∫M{F′s(2u<hF,HessFu>+su<∇FHF,∇Fu>+2h(∇Fu,∇Fu)−sHF‖∇Fu‖2)+F″s2ΔFu(ΔFu+u(‖hF‖2−‖hmix‖2))}dV+∫Mu{(F″s2(‖hF‖2−‖hmix‖2)−nsHF′)(ΔFu+u(‖hF‖2−‖hmix‖2))−F(Δu+u‖h‖2)+F′s(2<hF, u(h2F+h2mix)+HessFu>−u<hF+hF⊥,h2mix>−2<hmix,Hessmixu>)}dV. | (3.13) |
Proof. By (3.3a) with F=F(σF1/s), using <idTM,h2mix>=‖hmix‖2 and <idTF,HessFu>=ΔFu, we find the first variation of the functional WFn,s with F=F(HF), see (1.5):
δWFn,s=∫M{F′sΔFu+(F′s(‖hF‖2−‖hmix‖2)−nFH)u}dV=0. | (3.14) |
If (2.18) is valid, then using (3.14) and (2.20), we obtain (3.12). Our next aim is to calculate
δ2WFn,s=δ∫M{F′sΔFu+(F′s(‖hF‖2−‖hmix‖2)−nFH)u}dV=−∫M{F′sΔFu+(F′s(‖hF‖2−‖hmix‖2)−nFH)u}nuHdV+∫Mδ(F′sΔFu)dV+∫Mδ{(F′s(‖hF‖2−‖hmix‖2)−nFH)u}dV. | (3.15) |
For the first integral in the last line of (3.15), using (2.9), (2.12), and δu=0, we get
∫Mδ(F′sΔFu)dV=∫M{F′s(2u<hF,HessFu>+su<∇FHF,∇Fu>+2h(∇Fu,∇Fu)−sHF<∇Fu,∇Fu>)+F″s2ΔFu(ΔFu+u(‖hF‖2−‖hmix‖2))}dV. | (3.16) |
For the second integral in the last line of (3.15), using (2.11), we get
∫Mδ{(F′s(‖hF‖2−‖hmix‖2)−nFH)u}dV=∫Mu{(F″s2(‖hF‖2−‖hmix‖2)−nF′sH)(ΔFu+u(‖hF‖2−‖hmix‖2))−F(Δu+u‖h‖2)+F′s(2<hF, u(h2F+h2mix)+HessFu>−u<hF+hF⊥,h2mix>−2<hmix,Hessmixu>)}dV. | (3.17) |
By (3.15), (3.16), and (3.17), noting that the first variation vanishes at a critical immersion, we get
δ2WFn,s=−∫Mns{F′ΔFu+(F′(‖hF‖2−‖hmix‖2)−snFH)u}uHdV+∫M{F′s(2u<hF,HessFu>+su<∇FHF,∇Fu>+2h(∇Fu,∇Fu)−sHF‖∇Fu‖2)+1s2F″ΔFu(ΔFu+u(‖hF‖2−‖hmix‖2))}dV+∫Mu{(F″s2(‖hF‖2−‖hmix‖2)−nF′sH)(ΔFu+u(‖hF‖2−‖hmix‖2))−F(Δu+u‖h‖2)+F′s(2<hF, u(h2F+h2mix)+HessFu>−u<hF+hF⊥,h2mix>−2<hmix,Hessmixu>)}dV. | (3.18) |
From (3.18) and (3.12), at a critical hypersurface, we get (3.13).
Similarly, one can obtain the Euler-Lagrange equation for the functional with , see (1.5), but we do not present it here. From Theorem 3, with , we obtain the following.
Corollary 2. At a critical hypersurface satisfying (2.18), the second variation of the action in (1.6) is
(3.19) |
Proof. Substituting and in (3.13), we obtain (3.19).
Remark 4. By (3.18) with , the second variation of the action is
(3.20) |
This is compatible with a special case of Eq (7) in [3] for . As a special case of (3.20), the second variation of the functional in (1.1) with has the following form compatible with [3]:
We consider critical hypersurfaces equipped with two-dimensional foliations (i.e., ) in Section 4.1, and discuss critical hypersurfaces of revolution and their stability in Section 4.2.
For , it is natural to present the functionals (1.5) in the following form:
(4.1) |
where is the Gaussian curvature of the leaves. For , (4.1) reduces to the functional seen in [3]. The following equalities are true:
From (2.9) and (2.10) with , we obtain the following evolution equations:
(4.2) |
(4.3) |
Using (4.2) and (4.3) in , we get the evolution equation
(4.4) |
For , (4.2)–(4.4) reduce to the equations in [3].
The next statement for immersed in generalizes Theorem 1 in [3] with .
Theorem 4. If (2.18) is valid, then the Euler–Lagrange equation for the action (4.1) with is
(4.5) |
where denote partial derivatives of with respect to and . At a critical hypersurface foliated by surfaces and satisfying (2.18), the second variation of the functional (4.1) with is given by
(4.6) |
Proof. Using (4.2) and (4.4) in , and applying (2.6), we calculate the first variation of the functional (4.1) with :
(4.7) |
From (4.7), using (2.21), we get (4.5). From Theorem 3 with we get (4.6).
Remark 5. Let , then from (4.5) we obtain
From this, with , or from (3.4), we get the Euler-Lagrange equation for , see (1.6):
The following particular case of (4.6), or Corollary 2 with , is true.
Corollary 3. At a critical hypersurface satisfying (2.18), the second variation of is
The following consequence of Theorem 4 was proven for in [3].
Corollary 4. The Euler-Lagrange equation for the functional with is given by
Hypersurfaces of revolution in Euclidean space are naturally foliated into -spheres (parallels) and equipped with rotationally symmetric metrics – a special case of a warped product metric; see [15]. Such a hypersurface can be represented as a graph , where the function is monotonic, and are Cartesian coordinates in . We obtain the parametrization , of , where are cylindrical coordinates in . The principal curvatures of (functions of are for parallels and for profile curves (geodesics on ). If the profile curve is a straight line (), then and is a cone, a cylinder, or a hyperplane. To exclude these cases, we will assume . We get
(4.8) |
Recall that corresponds to the solutions with multiplicities of the eigenvalue problem on a unit -sphere. Any constant function on the round sphere spans the space of -eigenfunctions of the Laplacian. Let denote the orthogonality of functions with respect to the inner product. The sphere of radius in is not a local minimum of under volume-preserving deformations for . For , is a local minimum of under volume-preserving, nonconstant deformations provided , see Propositions 2 and 3 in [3]. According to (3.12), a hypersurface of revolution in foliated by -spheres-parallels is a critical point of the action with , see (1.5), if and only if
In this case, and are functionally related; hence, is a Weingarten hypersurface.
The following theorem studies the stability of hypersurfaces of revolution critical for (1.6).
Theorem 5. A hypersurface of revolution in foliated by -spheres-parallels is a critical point of the action or , see (1.6), if and only if
(4.9) |
A critical hypersurface is not a local minimum of for with respect to general variations, but it is a local minimum for variations satisfying .
Proof. 1. Let be critical for the action under general deformations. Since all principal curvatures are constant on parallels, from (3.4) and (3.5), we get, respectively,
(4.10) |
Using (4.8) in (4.10), yields , which is the differential equation for ,
(4.11) |
The solution of (4.11) is given by (4.9).
2. Let be the eigenfunction of on with the eigenvalue , then . Since our hypersurface of revolution is a warped product, its volume form is decomposed as , see [16]. For any function we have, see (2.22),
Using (3.19) with , Example 1, the equalities and
we find the second variation of :
(4.12) |
If the variation satisfies , then by (4.12) we get
which is negative for ; hence, our critical hypersurface is not a local minimum of .
Let satisfy . Using the inequalities and , see, for example, [3], we get
By these estimates, (4.12) and the inequality , we find
Hence, for all .
Example 1. (ⅰ) Let , be a hypersurface of revolution in foliated by parallels (2-spheres). We get and . Let be critical for the functional or with , see (1.6), under general deformations. From (4.10) we get . The solution to (4.11) is , where .
(ⅱ) Let , be a surface of revolution in foliated by parallels (circles). The principal curvatures are for parallels and for profile curves. Let be critical for the action or with under general deformations. Then ; hence, the equality is true. The solution to (4.11) is , where , it is illustrated on Figure 1 for .
This paper explores a generalized (for foliated hypersurfaces in a Riemannian manifold) form of the classical Willmore functionals, which is the Reilly-type functional. The 1st and 2nd variations of such functionals in the Euclidean space are computed, and the conformal properties of some of them are shown. Examples of critical hypersurfaces with low-dimensional transversally harmonic foliations and critical hypersurfaces of revolution, which are local minima for a specialized family of variations, are given. The results obtained are important for researchers working in the field of geometric variational problems and for scientists involved in the design of layered (laminated) or non-isotropic materials.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
Vladimir Rovenski is a guest editor for Electronic Research Archive and was not involved in the editorial review or the decision to publish this article. The author declares there is no conflict of interest.
[1] |
B. Chen, On a theorem of Fenchel-Borsuk-Willmore-Chern-Lashof, Math. Ann., 194 (1971), 19–26. https://doi.org/10.1007/BF01351818 doi: 10.1007/BF01351818
![]() |
[2] |
B. Chen, On the total curvature of immersed manifolds, II. Mean curvature and length of second fundamental form, Am. J. Math., 94 (1972), 799–809. https://doi.org/10.2307/2373759 doi: 10.2307/2373759
![]() |
[3] |
A. Gruber, M. Toda, H. Tran, On the variation of curvature functionals in a space form with application to a generalized Willmore energy, Ann. Glob. Anal. Geom., 56 (2019), 147–165. https://doi.org/10.1007/s10455-019-09661-0 doi: 10.1007/s10455-019-09661-0
![]() |
[4] |
F. C. Marques, A. Neves, Min-max theory and the Willmore conjecture, Ann. Math., 179 (2014), 683–782. https://doi.org/10.4007/annals.2014.179.2.6 doi: 10.4007/annals.2014.179.2.6
![]() |
[5] | Y. Zhu, J. Liu, G. Wu, Gap phenomenon of an abstract Willmore type functional of hypersurface in unit sphere, Sci. World J., (2014), 697132. https://doi.org/10.1155/2014/697132 |
[6] |
Y. Chang, Willmore surfaces and -Willmore surfaces in space forms, Taiwanese J. Math., 17 (2013), 109–131. http://doi.org/10.11650/tjm.17.2013.1840 doi: 10.11650/tjm.17.2013.1840
![]() |
[7] |
Z. Guo, Higher order Willmore hypersurfaces in Euclidean space, Acta. Math. Sin., Engl. Ser., 25 (2009), 77–84. https://doi.org/10.1007/s10114-008-6422-y doi: 10.1007/s10114-008-6422-y
![]() |
[8] | J. Li, Z. Guo, Higher order Willmore revolution hypersurfaes in , Lifelong Educ., 10 (2021) 75–89. |
[9] |
Z. Guo, Generalized Willmore functionals and related variational problems, Differ. Geom. Appl., 25 (2007), 543–551. https://doi.org/10.1016/j.difgeo.2007.06.004 doi: 10.1016/j.difgeo.2007.06.004
![]() |
[10] | T. J. Willmore, Note on embedded surfaces, An. Sti. Univ. Al. I. Cuza Iasi, N. Ser., 11 (1965), 493–496. |
[11] |
A. Song, Generation of tubular and membranous shape textures with curvature functionals, J. Math. Imaging Vis., 64 (2022), 17–40. https://doi.org/10.1007/s10851-021-01049-9 doi: 10.1007/s10851-021-01049-9
![]() |
[12] |
R. C. Reilly, Variational properties of functions of the mean curvatures for hypersurfaces in space forms, J. Differ. Geom., 8 (1973), 465–477. https://doi.org/10.4310/jdg/1214431802 doi: 10.4310/jdg/1214431802
![]() |
[13] | V. Rovenski, P. Walczak, Extrinsic Geometry of Foliations, Springer, Cham, 2021. https://doi.org/10.1007/978-3-030-70067-6 |
[14] | A. Candel, L. Conlon, Foliations, I, American Mathematical Society, Rhode Island, 2000. |
[15] | P. Petersen, Riemannian Geometry, edition, Springer, Cham, 2016. |
[16] | M. Berger, A Panoramic View of Riemannian Geometry, Springer-Verlag, Heidelberg, 2003. https://doi.org/10.1007/978-3-642-18245-7 |