ebook img

Deforming metrics of foliations PDF

0.22 MB·
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Deforming metrics of foliations

Deforming metrics of foliations Vladimir Rovenski and Robert Wolak ∗ † Abstract We study geometry of a manifold endowed with two complementary orthogonal distributi- 2 ons (plane fields) and a time-dependent Riemannian metric. The work begins with formulae 1 concerningdeformationsofgeometricquantitiesasthemetricvariesconformallyalongoneofthe 0 distributions. Then we introduce the geometric flow depending on the mean curvature vector 2 of the distribution, and show existence/uniquenes and convergence of a solution as t → ∞, n when the complementary distribution is integrable with compact leaves. We apply the method a to the problem of prescribing mean curvature vector field of a foliation, and give examples for J harmonic and umbilical foliations and for the double-twisted product metrics, including the 5 codimension-one case. ] Keywords: Riemannian metric; foliation; distribution; geometric flow; second fundamental G tensor; mean curvature; harmonic; umbilical; heat equation; laplacian; double-twisted product D Mathematics Subject Classifications (2010) Primary 53C12;Secondary 53C44 . h t a 1 Introduction m [ Geometric Flows (GFs) are important in many fields of mathematics and physics. A GF is an 3 evolutionofageometricstructureunderadifferentialequationrelatedtoafunctionalonamanifold, v usually associated with some curvature. The most popular GFs in mathematics are the heat flow 8 ([5], [7] etc), the Ricci flow ([2], [19] etc) and the mean curvature flow. They all correspond to 6 8 dynamical systems in the infinite-dimensional space of all appropriate geometric structures on a 1 given manifold. GF equations are quite difficult to solve in all generality, because of nonlinearity. . 9 Although the short time existence of solutions is guaranteed by the parabolic or hyperbolic nature 0 of the equations, their (long time) convergence to canonical geometric structures is analyzed under 1 1 various conditions. : Extrinsic geometry of foliated Riemannian manifolds describes properties which can be ex- v i pressed in terms of the second fundamental form of the leaves and its invariants (mean curvature X vector, higher mean curvatures and so on). One of the principal problems of extrinsic geometry r a of foliations reads as follows: Given a foliation F on a manifold M and an extrinsic geometric property (P), does there exist a Riemannian metric g on M such that F enjoys (P) w.r.t. g? Such problems (first posed by H.Gluck in 1979 for geodesic foliations) were studied already in the 1970’s when D.Sullivan provided a topological condition (called topological tautness) for a foliation, equivalent totheexistence ofaRiemannianmetricmakingalltheleaves minimal(see[3]). In recent decades, several tools providing results of this sort have been developed. Among them, one may find foliated cycles [18] and new integral formulae, [14, Part 1], [15], [16], [20], the very first of which is Reeb’s vanishing of the integral of the mean curvature. Few works consider GFs on foliated manifolds, see [1], [6], [10]). A GF on a foliated manifold is extrinsic, if the evolution dependson thesecond fundamentaltensor of the(leaves of the) foliation. Recently, thefirstauthor and P.Walczak introduced and studied extrinsic GFs on codimension-one foliations, see [14, Parts 2 and 3]. The paper extends this field of research for foliations of arbitrary codimension. ∗Mathematical Department,University of Haifa, E-mail: [email protected] †FacultyofMathematicsandComputerScience,TheInstituteofMathematicsofJagiellonianUniversity,Krakow, E-mail: [email protected] 1 Let(M,g)beaconnectedclosedRiemannianmanifoldwithapairofcomplementaryorthogonal distributionsDandD (planefields)ofdimensionsnandp,respectively. IfD(orD )isintegrable, ⊥ ⊥ itistangenttoafoliation F (respectively, F ). Denotebybandb thesecondfundamentaltensors ⊥ ⊥ of D and D (w.r.t. to g), H = Tr b ∈ Γ(D ) and H = Tr (b ) ∈ Γ(D) their mean curvature ⊥ g ⊥ ⊥ g ⊥ vectors. We call D (or D ) umbilical, harmonic or totally geodesic if ⊥ nb = H ·g , H = 0 or b = 0 (respectively, pb = H ·g , H = 0 or b = 0). D ⊥ ⊥ D⊥ ⊥ ⊥ | | By means of the natural representation of the structure group O(p)×O(n) on TM, Naveira [11] obtained thirty-six distinguished classes of Riemannian almost-product manifolds. Following this line of research, several geometers, [4], [8] and [9], completed the geometric interpretation, gave nontrivial examples for each class, and studied the behavior of the different conditions (and hence of the different classes of almost-product structures) under a conformal change of the metric. The notion of the D-truncated (r,k)-tensor field Sˆ (where r = 0,1, and denotes the D- component) will be helpful: Sˆ(X ,...,X ) = S(Xˆ ,...,Xˆ ) (X ∈ TM). We introduce the Ex- 1 k 1 k i b trinsic Geometric Flow (EGF) as a family g of Riemannian metrics on M satisfying the PDE t 2 ∂ g = − (div H )gˆ, (1) t t ⊥ t t n where g = g is given. Here gˆ is the D-truncated metric tensor g on M, i.e., gˆ(X,Y) = g(X,Y) 0 and gˆ(ξ,·) = 0 for all X,Y ∈ D and ξ ∈ D . Some operations w.r.t. EGF metrics g restricted ⊥ t on D are t-independent (e.g., ∇ and div ). Given a Riemannian metric g, the mean curvature ⊥ ⊥ ⊥ vector H can be expressed in terms of the first partial derivatives of g. Therefore g 7→ div H is a ⊥ second-order partial differential operator. We show that EGFs serve as a tool for studying the following: Question: Under what conditions on (M,D) the EGF metrics g converge to one for which D t enjoys a given extrinsic geometric property (P), e.g., is harmonic, umbilical, or totally geodesic? The structure of the paper is as follows. Section 2 collects main results. Section 3 develops formulae for the deformation of extrinsic geometric quantities of a foliation as the Riemannian metric varies conformally along D. Section 4 introduces EGF and shows the existence/uniquenes and converging of a global solution g (t ≥ 0). In particular, is is verified that t (i) the 1-form θ (dual to H) satisfies the heat equation along D . H ⊥ (ii) the metrics g preserve the ”umbilical” (”totally geodesic”, etc.) property of D. t (iii) for appropriate g the metrics g converge to a metric g with “harmonic distribution D”. 0 t ∞ In Theorem 1, the orthogonal distribution D is integrable and the method of proof is based on ⊥ solvingtheheatequationfor1-formsontheleavesofD . InTheorem2themethodisappliedtothe ⊥ problem of prescribing mean curvature vector field of D. Section 5 contains results and examples for codimension-one foliations (Proposition 6), and double-twisted product metrics. Appendix (Section 6) collects the necessary facts about the heat equation and the heat flow for 1-forms. 2 Main Results We define the connection ∇ induced on D by ∇ ξ = (∇ ξ) (i.e., ∇ ξ is projected onto D ), ⊥ ⊥ ⊥X X ⊥ X ⊥ where ξ ∈ D and X ∈ TM. In this section we suppose that D is integrable with all leaves ⊥ ⊥ compact and orientable. (If both D and D are integrable and the foliation tangent to D is ⊥ ⊥ totally geodesic then the foliation tangent to D is Riemannian). Denote θ the 1-form on D dual to the vector field ξ ∈ Γ(D ). The 1-form θ is D -harmonic ξ ⊥ ⊥ ξ ⊥ (see Section 6.2) if and only if δ θ = 0 (i.e., div ξ =0) and d θ = 0 (i.e., ∇ ξ is symmetric). ⊥ ξ ⊥ ⊥ ξ ⊥ i π Theorem 1. Let the leaves of D compose a fibration L ֒→ M → B. If dimD = p > 1, ⊥ ⊥ ⊥ suppose that d θ = 0. Then (1) admits a unique smooth solution g for all t ≥ 0 that converges ⊥ H t in C -topology as t → ∞ to a Riemannian metric g for which D is harmonic. ∞ ∞ 2 Example 1. Let the leaves of integrable distribution D be flat tori Tp. Any differential 1-form ⊥ on Tp can be written as ω = ω dxi. The form ω is harmonic if and only if the functions ω are i i i harmonic, and therefore constPant. In this case, the vector field dual to ω is constant. Indeed, the space of constant vector fields on a flat torus Tp is isomorphic to H1(Tp,R) ≃ Rp. A foliation, whose normal plane field is umbilical, is locally conformally equivalent to a Rie- mannian foliation, see [9]. The next corollary completes this fact. Corollary 1. Under the assumptions of Theorem 1, let D be g -umbilical. Then D is g -totally 0 ∞ geodesic (g is the limit of metrics (1) as t → ∞). ∞ A foliation with vanishing mean curvature is called harmonic. Every leaf of such foliation is a minimal submanifold of M. A foliation F is taut if there is at least one metric on M for which F is harmonic. In particular, if there is an immersed closed transversal manifold that intersects each leaf, then F is taut (see [13] and survey in [3]). For example, a Reeb foliation on S3 is not taut. The known proofs of existence of “taut” metrics use the Hahn–Banach Theorem and are not constructive. Corollary 2 (of Theorem 1) shows how to produce in some cases a family of metrics converging to the metric for which F is harmonic (i.e., H = 0). Certain results for ∞ codimension-one foliations are given in Section 5. Corollary 2. Let F be a foliation on (M,g) tangent to D of codimension p > 1. Suppose that the i π leaves of D compose a fibration L ֒→ M → B, and the equality d θ = 0 is satisfied. Then (1) ⊥ ⊥ ⊥ H0 admits a unique solution g (t ≥ 0), converging in C -topology as t → ∞ to a Riemannian metric t ∞ g , for which F is harmonic. ∞ Let F be a foliation of any codimension of a closed manifold M and X be a vector field on M. Recently, P. Schweitzer and P. Walczak [17] provided some necessary and sufficient conditions for X to become the mean curvature vector of F with respect to some Riemannian metric on M. Extending the definition of EGF and method of Theorem 1, we show how to produce in some cases (e.g., Theorem 2) a one-parameter family of metrics converging to the metric with prescribed mean curvature vector field of F. i π Theorem 2. Let p > 1 and the leaves of D compose a fibration L ֒→ M → B. Then for any ⊥ ⊥ smooth vector field X on M orthogonal to F and satisfying d θ = 0, the PDE ⊥ H X − 2 ∂ g = − div (H −X)gˆ, g = g (2) t t ⊥ t t 0 n admits a unique solution g (t ≥ 0), converging in C -topology as t → ∞ to a Riemannian metric t ∞ g with D -harmonic 1-form θ . If H1(L ,R) = 0 for the leaves of D , then H = X. ⊥ H∞ X ⊥ ⊥ ∞ − ∞ Example 2. (a) By Bochner theorem (see [5]), if the Ricci curvature of any leaf L of D is ⊥ ⊥ non-negative everywhere and positive for some point then H1(L ,R) = 0 (see Theorem 2). ⊥ (b)Thefollowingexample(communicatedtoauthorsbyP.Walczak)showsusthatthecondition d θ = 0 (see Theorem 2) and the assumption d θ = 0 (see Theorem 1 and Corollary 2) are ⊥ H X ⊥ H neede−d. Let X be a divergence free (e.g., a Killing) vector field on the leaves Sp of the product M = M ×Sp of a unit p-sphere and a Riemannian manifold (M ,g ). Let the distribution D on 1 1 1 M corresponds to TM . Then the product metric g has the mean curvature H = 0, and g is a 1 fixed point of the dynamical system (2). Consequently, H = 0 for all t ≥ 0 and H = 0 6= X. t ∞ 3 D-conformal variations of geometric quantities In this section we develop formulae for deformations of geometric quantities as the Riemannian metric varies conformally along one of the distributions. 3 3.1 Preliminaries Denote by M the space of smooth Riemannian metrics of finite volume on M such that D is ⊥ orthogonal to D. Elements of M are called (D,D )-adapted metrics. Let M ⊂ M be the ⊥ 1 subspace of (D,D )-adapted metrics of unit volume, and π : M → M , where π(g) = g¯ = ⊥ 1 vol(M,g) 2/ngˆ ⊕g be the D-conformal projection. Let g ∈ M (with 0 ≤ t < ε) be a family of − ⊥ t m(cid:0) etrics with g o(cid:1)f unit volume. Consider the D-truncated tensor field 0 S = ∂ g . t t t Recall that the musical isomorphism ♯ : T M → TM sends a covector ω = ω dxi to ω♯ = ωi∂ = ∗ i i gijω ∂ , and ♭ : TM → T M sends a vector X = Xi∂ to X♭ = X dxi = g Xjdxi. We denote by j i ∗ i i ij S♯ the (1,1)-tensor field on M which is g-dual to a symmetric (0,2)-tensor S, S(X,Y) = g(S♯(X),Y) for all vectors X,Y. Thevolume form vol of g evolves as d vol = 1 (Tr S♯)vol , see [19]. If S = s gˆ with s :M → R t t dt t 2 t t t t t (i.e., g are conformally equivalent along D and gˆ is the D-truncated metric g ) then t t t d n vol = s vol . (3) t t t dt 2 Hence, the metrics g˜ = (φ gˆ)⊕g with dilating factors φ = vol(M,g ) 2/n, belong to M . t t t t⊥ t t − 1 Recall that the Levi-Civita connection ∇t of a metric g on M is given by t 2g (∇t Y,Z) = X(g (Y,Z))+Y(g (X,Z))−Z(g (X,Y)) t X t t t + g ([X,Y],Z)−g ([X,Z],Y)−g ([Y,Z],X) (4) t t t for all vector fields X,Y and Z on M. Since the difference of two connections is always a tensor, Π := ∂ ∇t is a(1,2)-tensor fieldon(M,g ). Differentiation (4)w.r.t. tyieldstheformula, see[19], t t t 1 g (Π (X,Y),Z) = (∇t S )(Y,Z)+(∇t S )(X,Z)−(∇t S )(X,Y) (5) t t 2 X t Y t Z t (cid:2) (cid:3) for all X,Y,Z ∈ TM. Indeed, if the vector fields X = X(t), Y = Y(t) are t-dependent, then ∂ ∇t Y = Π (X,Y)+∇ (∂ Y)+∇ Y. (6) t X t X t ∂tX Notice the symmetry Π (X,Y)= Π (Y,X) of the tensor Π . t t t We will use the following condition for convergence of evolving metrics (see [2, Appendix A]). Proposition 1. Let ∂ g = s gˆ (t ≥ 0) be a one-parameter family of Riemannian metrics on t t t t a closed manifold M with complementary distributions D and D . Define functions u (t) = ⊥ m supM |(∇t)mst|g(t) and assume that 0∞um(t)dt < ∞ for all m ≥ 0. Then, as t → ∞, the metrics g converge in C -topology to a smoRoth Riemannian metric g . t ∞ ∞ Proof. Our assumptions ensure that g converge in C -topology to a symmetric (0,2)-tensor g . t ∞ The metrics are uniformly equivalent: c 1gˆ ≤ gˆ ≤cgˆ for some c> 0 and all t ≥ 0. Hence, g ∞is − 0 t 0 ∞ positive definite. Let ∇ φ be the D -component of the gradient of a function φ ∈ C1(M). The second funda- ⊥ ⊥ mental tensor of D (similarly of D ) is defined by ⊥ 1 b(X,Y) = (∇ Y +∇ X) , X,Y ∈ D. (7) X Y ⊥ 2 The second fundamental forms of D with respect to metrics g and g˜= (e2φgˆ)⊕g are related ⊥ by the following lemma. 4 Lemma 1 (see[14]forcodimension-onefoliations). Let (M, g = gˆ⊕g ) be a Riemannian manifold ⊥ with complementary orthogonal distributions D and D . Given φ ∈ C1(M), define a metric g˜ = ⊥ (e2φgˆ)⊕g . Then the second fundamental forms and the mean curvature vectors of D w.r.t. g˜ ⊥ and g are related by ˜b = e2φ b−(∇ φ)gˆ , H˜ = H −(dimD)∇ φ. (8) ⊥ ⊥ (cid:0) (cid:1) So, if φ is constant then, by (8), we have: ˜b = e2φb and H˜ = H. Proof. By (4), for any X,Y ∈ D and ξ ∈ D we have ⊥ 1 g(∇˜ Y,ξ)= e2φg(∇ Y,ξ)−e2φg(X,Y)ξ(φ)− (e2φ −1)g([X,Y],ξ). X X 2 From this and definition (7), formula (8) follows. Since H = Tr b, we have (8) . 1 g 2 3.2 The integral formula The divergence of a vector field X on (M,g) is given by divX = g(∇ X,e ), where (e ) is a s es s s local orthonormal frame on (M,g). Recall the identity for a smootPh function f : M → R, div(f ·X) = f ·divX +X(f). For a compact manifold M with boundary and inner normal n, the Divergence Theorem reads as divXdvol = g(X,n)dω. (9) Z Z M ∂M For a closed manifold M, we have divXdvol =0. The D -divergence, div ξ, of a vector field M ⊥ ⊥ ξ ∈ Γ(D ) is defined similarly to dRivξ, using a local orthonormal frame (ε ) of D . ⊥ α ⊥ Lemma 2. For a vector field ξ ∈ Γ(D ) on a closed manifold M, we have the identity ⊥ (div ξ)dvol = g(H,ξ)dvol. (10) ⊥ Z Z M M So, (div ξ)dvol = 0 for any vector field ξ ∈ Γ(D ) if and only if H = 0. M ⊥ ⊥ R Proof. Using the definition H = b(e ,e ), we have i p i i P≤ divξ−div ξ = g(∇ ξ,e )= − g(b(e ,e ),ξ) = −g(H,ξ). ⊥ Xi p ei i Xi p i i ≤ ≤ By the Divergence Theorem, div ξdvol = 0, we obtain (10). M R Remark 1. (i) By Lemma 2 with ξ = H, we have (div H)dvol = g(H,H)dvol ≥ 0. (11) ⊥ Z Z M M (ii) By Lemma 2 with ξ = ∇ f, for a function f ∈ C2(M), we have ⊥ (∆ f)dvol = g(∇f,H)dvol. ⊥ Z Z M M Here ∆ f = div (∇ f) is the D -Laplacian of f. ⊥ ⊥ ⊥ ⊥ (iii) In analogy with the fact that on a closed connected Riemannian manifold, every harmonic function (i.e., ∆f = 0) is constant, we claim: If ∆ f = g(∇f,H) then ∇ f = 0. Indeed, ⊥ ⊥ div(f∇ f)+f(H(f)−∆ f)= g(∇ f,∇ f). ⊥ ⊥ ⊥ ⊥ Using the Divergence Theorem, we obtain g(∇ f,∇ f)dvol = 0, and then ∇ f = 0. M ⊥ ⊥ ⊥ R 5 3.3 D-related geometric quantities Let {e , ε } (i ≤ n, α≤ p) be a local g -orthonormal frame on TM adapted to D and D . i α 0 ⊥ Lemma 3. Let {e } be a local g -orthonormal frame of D (on a set U ⊂ M), and ∂ g = sgˆ. i 0 q t t t Suppose that {e (t)} evolves according to i 1 ∂ e (t) = − se (t). (12) t i i 2 Then {e (t)} is a g -orthonormal frame of D on U for all t. i t q Proof. We have ∂ (g (e ,e )) = g (∂ e (t),e (t))+g (e (t),∂ e (t))+(∂ g )(e (t),e (t)) t t i j t t i j t i t j t t i j 1 1 =sgˆ(e (t),e (t))− g (se (t),e (t))− g (e (t),se (t)) = 0. t i j t i j t i j 2 2 The following lemma is compatible with Lemma 1. Lemma 4. Let g ∈ M and ∂ g = s gˆ for some s ∈ C1(M). Then the second fundamental t t t t t t tensor b, its mean curvature vector H and the D -divergence div H are evolved by ⊥ ⊥ 1 1 ∂ b(X,Y) = sb(X,Y)− gˆ(X,Y)∇ s+ (∇ ∂ Y +∇ Y +∇ ∂ X +∇ X) , (13) t 2 ⊥ 2 X t ∂tX Y t ∂tY ⊥ n n n ∂ H = − ∇ s, ∂ (div H) = − ∆ s, ∂ θ = − d s. (14) t ⊥ t ⊥ ⊥ t H ⊥ 2 2 2 Proof. Let S = ∂ g be D-truncated. By (5), (7), symmetry of S and S(·,D )= 0, we have t t ⊥ 1 g (∂ b(X,Y),ξ) = g ∂ (∇t Y)+∂ (∇t X), ξ t t 2 t t X t Y (cid:0) (cid:1) 1 = (∇t S)(Y,ξ)+(∇t S)(X,ξ)−(∇tS)(X,Y) +Q 2 X Y ξ (cid:2) (cid:3) for all ξ ∈ D and t-dependent X,Y ∈ D. Here, Q := 1 g (∇t ∂ Y +∇t Y +∇t ∂ X+∇t X, ξ) ⊥ 2 t X t ∂tX Y t ∂tY due to (6). Substituting S = sgˆ, we obtain the required (13): 1 g(∂ b(X,Y), ξ) = − sgˆ(Y,∇t ξ)+sgˆ(X,∇t ξ)+ξ(s)gˆ(X,Y) +Q t 2 X Y (cid:2) (cid:3) 1 = sg(b(X,Y), ξ)− gˆ(X,Y)ξ(s)+Q. 2 Let {e (t)} be a local g -orthonormal frame of D on U for all t, hence (12) holds, see Lemma 3. i t q By the above we obtain (14) (see also alternative proof in Remark 2): 1 ∂ H = ∂ b(e (t),e (t)) t t i i Xi 1 n = sb(e (t),e (t))− g (e (t),e (t))∇ s − sb(e (t),e (t)) = − ∇ s. i i t i i ⊥ i i ⊥ Xi(cid:2) 2 (cid:3) Xi 2 To show (14) , let S = ∂ g be D-truncated (i.e., S(D ,·) = 0). Using (5), we have 2 t t ⊥ ∂ (div H) = ∂ (g(∇ H,ε )) = (∂ g)(∇ H,ε )+g (∂ (∇ H),ε ) t ⊥ Xα≤p t εα α Xα≤p(cid:2) t εα α t t εα α (cid:3) 1 = S(∇ H,ε )+ (∇ S)(ε ,ε ) +div (∂ H) = div (∂ H). Xα≤p(cid:2) εα α 2 H α α (cid:3) ⊥ t ⊥ t Here, weusedS(ε ,·) = 0and(∇ S)(ε ,ε ) = H(S(ε ,ε ))−2S(∇ ε ,ε )= 0. Now, assuming α H α α α α H α α S = sgˆ, and using (14) , we obtain (14) : ∂ (div H)= div (∂ H)= −n div (∇ s)= −n ∆ s. 1 2 t ⊥ ⊥ t 2 ⊥ ⊥ 2 ⊥ To show (14) , we use (14) to calculate for any X ∈ D : 3 1 ⊥ n ∂ θ (X) = ∂ (g (H ,X)) = sgˆ(H ,X)+g (∂ H ,X)+g (H ,∂ X) = − g (∇ s,X)+g (H ,∂ X). t H t t t t t t t t t t t ⊥ t t t 2 Hence, (∂ θ )(X) = −n (∇ s)♭(X)+g (H ,∂ X)−θ (∂ X) = −nd s(X). t H 2 ⊥ t t t H t 2 ⊥ 6 Remark 2. The alternative proof of (14) is based on the identity (see [14, Lemma 2.4] for k = 0) 1 ∂ (Tr B)= Tr (∂ B)−hB,Si , t gt gt t gt where S = ∂ g, B – a t-dependent symmetric (k,2)-tensor on (M,g), and hB,Si= BijS . t ij In our case, k = 1, B = b, S = sgˆ and Tr B = H . Thus, using (13), we have t gt t n ∂ (Tr B) = ∂ H, Tr (∂ B)= Tr (∂ b) = sH − ∇ s, t gt t gt t gt t 2 ⊥ hB,Si = hb,sgˆi = shb ,gˆi = sTr b = sH . gt t gt t t gt gt t t Next we show that D-conformal variations of metrics preserve the umbilicity of D. Proposition 2. Let ∂ g = s gˆ (s : M → R), be a D-conformal family of Riemannian metrics t t t t t on a manifold (M,D,D ). If D is umbilical for g , then D is umbilical for any g . ⊥ 0 t Proof. Since D is g -umbilical, we have b = 1 Hgˆ at t = 0, where H is the mean curvature 0 n vector field of D. Applying to (13) the theorem on existence/uniqueness of a solution of ODEs, we conclude that b = 1 H˜ gˆ for all t, for some H˜ ∈ Γ(D ). Tracing this, we see that H˜ is the mean t n t t t ⊥ t curvature vector of b , hence D is umbilical for any g . t t 3.4 D -related geometric quantities ⊥ Lemma 5. Let g ∈ M and ∂ g = s gˆ for some s ∈ C1(M). Then the second fundamental t t t t t t tensor b and its mean curvature vector H are evolved as ⊥ ⊥ ∂ b = −sb , ∂ H =−sH . (15) t ⊥ ⊥ t ⊥ ⊥ Proof. We shall show for the more general setting S = ∂ g that t t ∂ b = −S♯◦b , ∂ H = −S♯(H ). (16) t ⊥ ⊥ t ⊥ ⊥ Using (5), we compute for any X ∈ D and ε ,ε ∈ D , α β ⊥ 1 g (∂ b (ε ,ε ),X) = g (∂ (∇t ε )+∂ (∇t ε ), X) t t ⊥ α β 2 t t εα β t εβ α 1 = (∇t S)(X,ε )+(∇t S)(X,ε )−(∇t S)(ε ,ε ) 2 εα β εβ α X α β (cid:2) (cid:3) 1 = − S(∇t ε ,X)+S(∇t ε ,X) = −S(b (ε ,ε ),X). 2 εα β εβ α ⊥ α β (cid:2) (cid:3) From this (16) follows when S = sgˆ. Next, for any X ∈ Γ(D), we have 1 1 g (∂ H ,X) = g (∂ (∇ ε ),X) = 2(∇ S)(ε ,X)−(∇ S)(ε ,ε ) t t ⊥ Xα t t εα α 2(cid:2) εα α X α α (cid:3) = − S(∇ ε ,X) = −S(H ,X), Xα εα α ⊥ which confirms (16) . From (16) for S = sgˆ we obtain (15). 2 Next we show that D-conformalvariations of metrics preserve“umbilical” (i.e., b = H ·g ) ⊥ ⊥ D⊥ | and “harmonic” (i.e., H = 0) properties of D . ⊥ ⊥ Proposition 3. Let ∂ g = s gˆ (s :M → R), be a D-conformal family of Riemannian metrics on t t t t t a manifold (M,D,D ). If D is either umbilical (e.g., totally geodesic) or harmonic for g , then ⊥ ⊥ 0 D is the same for any g . ⊥ t Proof. If D is g -umbilical, then b = 1 H g at t = 0, where H is the mean curvature vector ⊥ 0 ⊥ p ⊥ ⊥ ⊥ field of D . Applying to (15) the theorem on existence and uniqueness of a solution of ODEs, we ⊥ 1 conclude that b = 1H˜ g for all t, where H˜ ∈ Γ(D). Tracing this, we show that H˜ is the mean ⊥t p t t⊥ t t curvature vector of b , hence D is umbilical for any g . Indeed, H˜ ≡ 0 when D is g -totally ⊥t ⊥ t t ⊥ 0 geodesic. The remaining property, i.e., D is harmonic, can be proved similarly. ⊥ 7 4 Proofs of main results 4.1 Introducing the Extrinsic Geometric Flow Definition 1. Given g = g, a family of (D,D )-adapted Riemannian metrics g , t ∈ [0,ε), on M 0 ⊥ t will be called (a) an Extrinsic Geometric Flow (EGF) if (1) holds; (b) a normalized EGF if 2 2 ∂ g = − div H +r(t) gˆ, where r(t) = − (div H )dvol /vol(M,g ). (17) t t n ⊥ t t nZ ⊥ t t t (cid:0) (cid:1) M TheEGF(1)andits normalized companion (17)providesomemethodsof evolving Riemannian metrics on foliated manifolds. Obviously, both EGFs preserve harmonic (and totally geodesic) foliations. If g ∈ M , then all metrics g (t ≥ 0) of (17) belong to M , because, see (3), 0 1 t 1 d n vol(M,g )= − div H + r(t) dvol = r(t)vol(M,g )− r(t)dvol = 0. dt t ZM (cid:16) ⊥ t 2 (cid:17) t t ZM t Substituting (11) in the definition (17) of r(t), we have 2 r(t)= − g(H ,H )dvol /vol(M,g )≤ 0. n Z t t t t M ByProposition2, EGFspreservethefollowing propertiesofD : (i)umbilical(b = H ·g ), ⊥ ⊥ ⊥ D⊥ | (ii) totally geodesic (b = 0), (iii) harmonic (H = 0); which in case of integrable D mean that ⊥ ⊥ the foliation tangent to D is (i) conformal, (ii) Riemannian, (iii) transversally harmonic. Let g be a family of Riemannian metrics of finite volume on (M,D,D ). Metrics g˜ = (φ gˆ)⊕ t ⊥ t t t g with φ = vol(M,g ) 2/n have unit volume: dvol = 1. The next proposition shows that t⊥ t t − M t unnormalized and normalized EGFs differ only byRrescaling along the distribution D. f Proposition 4. Let g be a solution (of finite volume) to (1) on (M,D,D ). Then the metrics t ⊥ g˜ = (φ gˆ)⊕g , where φ = vol(M,g ) 2/n, t t t ⊥ t t − evolve according to the normalized EGF 2 2 ∂ g˜ = − div H˜ +ρ gˆ˜, where ρ = − (div H)dvol /vol(M,g ). (18) t t n ⊥ t t t t n Z ⊥ t t (cid:0) (cid:1) M Proof. Since φ depends only on t, by Lemma 1, H˜ = H for metrics g˜ and g . Hence, div H˜ = t t t t ⊥ t div H . From (3) with s = −2(div H ) we get the derivative of the volume function ⊥ t n ⊥ t d d vol(M,g ) = dvol = − (div H )dvol . dt t dt Z t Z ⊥ t t M M Thus φ = vol(M,g ) 2/n is a smooth function of variable t. By Lemma 1, we have ˜b = φ ·b . t t − t t t Therefore 2 φ ∂ g˜ = φ ∂ g +φ gˆ = − div H − ′t gˆ˜. t t t t t ′t t n ⊥ t φ t (cid:0) t(cid:1) e n/2 Notice that dvol = φ dvol . Using this and (3), we obtain t t t fddt volt = ddt (φtn2 volt) = (cid:0)n2 φtn2−1φ′t+ 21φtn2ns(cid:1)volt = n2(cid:16)φφt′t +s(cid:17)volt. f f Let ρ be the average of s, see (18). From the above we get t d φ φ 0 = 2 dvol = n ′t +s dvol = n ′t +ρ . dt ZM t ZM (cid:16)φt (cid:17) t (cid:0)φt t(cid:1) f f This shows that ρ = −φ /φ . Hence, g˜ evolves according to (18). t ′t t 8 4.2 Proof of Theorems 1–2 and Corollaries 1–2 Proposition 5. Let D be integrable. The mean curvature vector H of D and its D -divergence ⊥ t ⊥ with respect to g of EGF (1) or (17) satisfy the following PDEs on any leaf of D : t ⊥ ∂ H = ∇ (div H), ∂ (div H) = ∆ (div H). (19) t ⊥ ⊥ t ⊥ ⊥ ⊥ Proof. By Lemma 4 with s = −2 (div H) or s = −2 div H −r(t), we obtain (19) . Similarly, n ⊥ n ⊥ 1 from (14) we deduce (19) . 2 2 The eigenvalue problem, −∆ u = λu, on a leaf L (of D when it is integrable) has solution ⊥ ⊥ ⊥ with a sequence of eigenvalues with repetition (each one as many times as the dimension of its finite-dimensional eigenspace) 0 = λ < λ ≤ λ ≤ ··· ↑ ∞. Let φ be an eigenfunction with 0 1 2 j eigenvalue λj satisfying L⊥φ2j(x)dx = 1. Then G⊥(t,x,y) = jeλjtφj(x)φj(y) is a fundamental solution of the heat equaRtion on L , see details in Section 6.1. APsolution satisfying u(x,0) =u (x) ⊥ 0 is given by u(x,t) = G (t,x,y)u (y)dy. Moreover, if u ∈ L2(L ) then the solution converges L⊥ ⊥ 0 0 ⊥ uniformly, as t → ∞,Rto a D -harmonic function (constant λ when L is closed). ⊥ 1 ⊥ Proof of Theorem 1. This is based on the heat flow for 1-forms, see Section 6.2. Let θt be the H dual 1-form on D to the mean curvature vector field H (with respect to g for t ≥ 0). Using ⊥ t t div H = −δ θt and applying (14) with s = −2 div H, we show similarly to (19) that ⊥ t ⊥ H 3 n ⊥ 1 ∂ θt = −d δ θt , (20) t H ⊥ ⊥ H where θ0 = θ is known. We obtain (for L -product along the leaves of D ) H H 2 ⊥ 1 ∂ (kθt k2)= ∂ hθt , θt i = −hd δ θt , θt i= −kδ θt k2 ≤ 0. 2 t H t H H ⊥ ⊥ H H ⊥ H By the above, we have uniqueness of a solution of the linear PDE (20). By Theorem B in Section 6.2, the heat equation (considered on the leaves of D ) ⊥ ∂ ω = ∆ ω, where ω = θ , (21) t ⊥d 0 H admits a unique solution ω (t ≥ 0). As t → ∞, the 1-form ω converges exponentially to a t t D -harmonic 1-form ω , i.e., kω − ω k ≤ ce λt for some constants c,λ > 0. Since ω = θ ⊥ t − 0 H ∞ ∞ is D -closed, again by Theorem B, we have d ω = 0 for all t ≥ 0. (Notice that for p = 1 the ⊥ ⊥ t 1-form θ is always D -closed). Comparing (20) with (21), we conclude that θt = ω is a unique H ⊥ H t solution of (20). By the above, div H = 0 (here H ∈ Γ(D ) is dual to θ = ω ). Applying ⊥ ⊥ H∞ the integral formula (11), we conclude t∞hat lim kH k2∞=0, hence H = 0. ∞ t t ∞ With known H , the PDE (1) also has a →un∞ique smooth global solution, and t 2 t gˆ = gˆ exp − (div H )ds . t 0 n Z ⊥ s (cid:0) 0 (cid:1) By Lemma 5, we also have b = b exp(2 t(div H )ds) for t ≥ 0. Since the leaves of D are ⊥t ⊥0 n 0 ⊥ s ⊥ compact, and div H satisfies (19) (on anyR leaf of D ), we have |div H |≤ e λ1t|div H | and ⊥ t 2 ⊥ ⊥ t − ⊥ 0 t t t 1−e λ1t c˜ (div H )ds ≤ div H ds < c˜ e λ1tds = c˜ − < (cid:12)Z0 ⊥ s (cid:12) Z0 (cid:12) ⊥ s(cid:12) Z0 − λ1 λ1 (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) for some c˜>0 and all t ≥ 0. Now, for some c≥ 1 and all t ≥ 0, we have the uniform bounds c 1gˆ ≤ gˆ ≤ cgˆ , c 1b ≤ b ≤ cb . − 0 t 0 − ⊥0 ⊥t ⊥0 From the above and Proposition 1, g → g and b → b as t → ∞ in C -topology. t ⊥t ⊥ ∞ ∞ ∞ 9 Proof of Corollary 1. By Theorem 1, the EGF metrics g converge to a smooth metric g with t ∞ H = 0. By Proposition 2, EGFs preserve the umbilicity of D, hence D is g -umbilical. Note ∞ ∞ that an umbilical foliation with vanishing mean curvature is totally geodesic. Proof of Corollary 2. The leaves of orthogonal foliation are immersed compact cross-sections, hence F is taut. On the other hand, by Theorem 1, the EGF metrics g converge as t → ∞ to a t smooth Riemannian metric g with H = 0. ∞ ∞ Proof of Theorem 2. The vector field H = H −X satisfies PDEs of Proposition 5, t t e ∂ H = ∇ div H , ∂ (div H )= ∆ (div H ). (22) t t ⊥ ⊥ t t ⊥ t ⊥ ⊥ t e e e e Since ∂ X = 0, the 1-form θt := θ −θ satisfies the PDE, see (20), t He Ht X ∂tθHe = −d⊥δ⊥θHe (23) withtheinitial valueθ0 = θ −θ . As in theproofof Theorem1, wehave uniquenessof a solution e H X H of (23) for t ≥ 0. By Theorem B in Section 6.2, the heat equation (considered on the leaves of D ) ⊥ ∂ ω = ∆ ω (24) t ⊥d admits a unique smooth solution ω (t ≥ 0) with the initial value ω = θ −θ . As t → ∞, the t 0 H X 1-form ω converges exponentially to a D -harmonic 1-form ω , and kω −ω k ≤ ce λt for some t ⊥ t − ∞ ∞ constants c,λ > 0. Since ω is D -closed, again by Theorem B, we have d ω = 0 for all t ≥ 0. 0 ⊥ ⊥ t Comparing (23) with (24), as in the proof of Theorem 1, we conclude that θt = ω is a unique e t H solution of (23). By the above, div (H −X) = 0 (here H −X is dual to ω ). With known H , ⊥ t the PDE (2) also has a unique smooth∞global solution, and∞gˆ = gˆ exp − 2 ∞tdiv (H −X)ds . t 0 n 0 ⊥ s ByLemma5,wealsohaveb = b exp(2 tdiv (H −X)ds)forallt ≥(cid:0)0. UsRing|div (H −X)| ≤(cid:1) ⊥t ⊥0 n 0 ⊥ s ⊥ t e λ1t|div (H −X)|, we obtain R − ⊥ 0 t t t 1−e λ1t c˜ (div (H −X))ds ≤ div (H −X) ds < c˜ e λ1tds = c˜ − < (cid:12)Z0 ⊥ s (cid:12) Z0 (cid:12) ⊥ s (cid:12) Z0 − λ1 λ1 (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) for some c˜ > 0 and all t ≥ 0. As in the proof of Theorem 1, we conclude that g converges in t C -topology as t → ∞ to a Riemannian metric g with D -harmonic 1-form θ . ∞ ⊥ H∞ X Certainly, if H1(L ,R) = 0 for the leaves of D∞, then H = X. − ⊥ ⊥ ∞ 5 More examples 5.1 The codimension-one case Let(M,g)beaclosedRiemannianmanifoldwithacodimension-onedistributionD(i.e.,p = 1). Let N be the unit vector field (orthogonal to D), and b the scalar second fundamental form of D with respecttoN. Indeed,2b(X,Y)= g(∇ Y +∇ X, N). Hence, H = τ N andτ = g(N,H) = TrA, X Y 1 1 where A= b♯ is the shape operator. By Lemma 4, we find the variations (see also [14]) 1 n ∂ A= − N(s)iˆd, ∂ τ = − N(s), (25) t t 1 2 2 where ∂ g = s gˆ. For a codimension-one case, the EGF definitions (1) and (17) read as: t t t t 2 ∂ g = − N(τ )gˆ, (26) t t 1 t n 2 2 ∂ g = − N(τ )+r(t) gˆ, r(t) = − N(τ )dvol /vol(M,g ), (27) t t n 1 t nZ 1 t t (cid:0) (cid:1) M 10

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.