Skip to main content
Log in

Extrinsic Geometric Flows on Codimension-One Foliations

  • Published:
Journal of Geometric Analysis Aims and scope Submit manuscript

Abstract

We study the geometry of a codimension-one foliation with a time-dependent Riemannian metric. The work begins with formulae for deformations of geometric quantities as the Riemannian metric varies along the leaves of a foliation. Then the Extrinsic Geometric Flow depending on the second fundamental form of the foliation is introduced. Under suitable assumptions, this evolution yields the second-order parabolic PDEs, for which the existence/uniqueness and in some cases convergence of a solution are shown. Applications to the problem of prescribing the mean curvature function of a codimension-one foliation, and examples with harmonic and umbilical foliations (e.g., foliated surfaces) and with twisted product metrics are given.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Brendle, S.: Ricci Flow and the Sphere Theorem. Graduate Studies in Math., vol. 111. AMS, Providence (2010)

    MATH  Google Scholar 

  2. Candel, A., Conlon, L.: Foliations, I. AMS, Providence (2000)

    MATH  Google Scholar 

  3. Doyle, P.W., Vassiliou, P.J.: Separation of variables for the 1-dimensional non-linear diffusion equation. Int. J. Non-Linear Mech. 33(2), 315–326 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  4. Eells, J., Sampson, J.: Harmonic mappings of Riemannian manifolds. Am. J. Math. 86, 109–160 (1964)

    Article  MathSciNet  MATH  Google Scholar 

  5. Ilyin, A.M., Kalashnikov, A.S., Oleynik, O.A.: Linear second-order partial differential equations of the parabolic type. J. Math. Sci. 108(4), 435–542 (2002). Translated from Russian: Russ. Math. Surv. 17(3), 3–146 (1962)

    Article  MathSciNet  Google Scholar 

  6. Lovric, M., Min-Oo, M., Ruh, E.: Deforming transverse Riemannian metrics of foliations. Asian J. Math. 4(2), 303–314 (2000)

    MathSciNet  MATH  Google Scholar 

  7. Oshikiri, G.: A characterization of the mean curvature functions of codimension-one foliations. Tohoku Math. J., II. Ser. 49(4), 557–563 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  8. Prieto, J.I.R., Saralegi-Aranguren, M., Wolak, R.: Cohomological tautness for Riemannian foliations. Russ. J. Math. Phys. 16(3), 450–466 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  9. Rovenski, V., Walczak, P.: Topics in Extrinsic Geometry of Codimension-One Foliations. Springer, Berlin (2011). doi:10.1007/978-1-4419-9908-5

    Book  MATH  Google Scholar 

  10. Rovenski, V., Walczak, P.: Extrinsic geometric flows on foliated manifolds, I (2010), 34 pp. arXiv:1003.1607v3

  11. Rovenski, V., Walczak, P.: Extrinsic geometric flows on foliated manifolds, II (2010), 18 pp. arXiv:1009.6066v2

  12. Taylor, M.E.: Partial Differential Equations, III: Nonlinear Equations, 2nd edn. Applied Math. Sciences, vol. 117. Springer, Berlin (2011)

    MATH  Google Scholar 

  13. Tondeur, P.: Foliations on Riemannian manifolds. Springer, Berlin (1988)

    Book  MATH  Google Scholar 

  14. Topping, P.: Lectures on the Ricci flow. LMS Lecture Notes, vol. 325. London Math. Society and Cambridge University Press, London (2006)

    Book  MATH  Google Scholar 

  15. Walczak, P.: Mean curvature functions for codimension-one foliations with all leaves compact. Czechoslov. Math. J. 34(109), 146–162 (1984)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

The author would like to thank Igor Gaissinski (Technion, Haifa) for helpful discussion of Sect. A.1. The work was supported by the Marie Curie Actions grant EU-FP7-P-2010-RG, No. 276919.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vladimir Rovenski.

Additional information

Communicated by Jiaping Wang.

Appendix: PDEs

Appendix: PDEs

In Sect. A.1, following [12] we discuss parabolic PDEs and prove Proposition 12 about the quasi-linear heat equation. In Sect. A.2, using the generalized companion matrix, defined in [9], we prove Proposition 14 about an infinite system of parabolic PDEs.

1.1 A.1 Parabolic PDEs in One Space Variable

We recall some facts about parabolic systems of quasi-linear PDEs.

Let A=(a ij (t,x,u)) be an n×n-matrix, \(a=(a_{i}(t,x,u,\partial_{x}\/u))\)n-vector, and t,x∈ℝ. Consider the quasilinear system of PDEs, n equations in n unknown functions u=(u 1,…,u n )

$$ \partial_tu= A(t,x,u)\,\partial^2_{xx}\/u+ a(t,x,u,\partial_x\/u).$$
(53)

PDEs (53) are homogeneous if a≡0. When A depends on t and x only, the system is semilinear. When A and a are functions of t and x, the system is linear. The initial value problem for (53) with given smooth data A, a, and u 0,

$$ u(0,x)=u_0(x),$$
(54)

consists of finding a smooth function u(t,x) satisfying (53)–(54).

The parabolicity condition for (53) says that there is c=const>0 such that

$$ \langle A\,v,v\rangle \ge c\langle v,v\rangle ,\quad\forall\,v\in \mathbb{R}^n.$$
(55)

Theorem A

(12, Ch. 15, Proposition 8.3)

Suppose that A of class C satisfies the parabolicity condition (55). Then the system (53)(54) with u 0H s(ℝ) (the Sobolev space) for some s>1, admits a unique solution uC([0,T), H s(ℝ))∩C ((0,T]×ℝ), which persists as long as \(\|u\|_{C^{r}}\) is bounded, given r>0.

One may apply this and the maximum principle to obtain the global existence and uniqueness result for a scalar parabolic equation (i.e., n=1 and A 11=k(u)>0) on S 1.

The following proposition is standard; for the convenience of the reader, we give its proof.

Proposition 12

Let a function k(u)∈C (ℝ) (the thermal diffusivity) satisfy

$$0<c_1\le k(u) \le c_2<\infty $$

for some real c 1c 2. Then the quasi-linear heat equation on a unit circle S 1

$$ \partial_tu = \partial_x\/\bigl(k(u)\,\partial_x\/u\bigr), \qquad u(0,x)=u_0(x) \in C^\infty\bigl(S^1\bigr)$$
(56)

admits a unique solution uC ([0,∞)×S 1). Moreover, there exists lim t→∞ u(t,x)=u ∈ℝ, and for some real α,K>0 the following inequalities are satisfied:

$$ \|u(t,\cdot) -u_\infty\|_{S^1}\le K\,\frac{c_2}{c_1}\,e^{-\alpha\,t}\| u_0 -u_\infty \|_{S^1}.$$
(57)

Proof

By Proposition 9.11 in [12, Ch. 15], there is a unique solution uC ([0,∞)×S 1). By the maximum principle, \(\|u(t,\cdot)\|_{S^{1}}\le\|u_{0}\|_{S^{1}}\) for all t>0. Define the function v=φ(u) of variables (t,x), where φ′(u)=k(u). The monotone function φ (of one variable) satisfies inequalities

$$ c_1\le\varphi'\le c_2.$$
(58)

The PDE (56)1 reads as \(\partial_{t}u= \partial^{2}_{xx}\/v\), hence (in view of derivation t v=φ ′(u)  t u) is equivalent to

$$\partial_tv = k(u)\,\partial^2_{xx}\/v,\qquad v(0,x)=\varphi\bigl(u_0(x)\bigr).$$

Let us compare it with the linear PDE on S 1,

$$ \partial_t\widetilde{v} =\tilde{k}(t,x)\,\partial^2_{xx}\/\widetilde{v},\qquad\widetilde{v}(0,x)=\varphi\bigl(u_0(x)\bigr),$$
(59)

where \(\tilde{k}(t,x)=k(u(t,x))\) is given. Indeed, \(c_{1}\le\tilde{k}(t,x)\le c_{2}\) for all t>0 and xS 1. From the existence and uniqueness of a solution to (59) we conclude that \(\widetilde{v}=v\). Denote by u ∈ℝ the average of u 0 over S 1, and set v =φ(u ). The function w=v(t,x)−v solves the linear PDE on S 1

$$ \partial_t\widetilde{w} =\tilde{k}(t,x)\,\partial^2_{xx}\/\widetilde{w},\qquad \widetilde{w}(0,x)=\varphi\bigl(u_0(x)\bigr)-v_\infty.$$
(60)

By the theory of linear PDEs, (60) possesses a fundamental solution \(\widetilde{G}(t,x,y)\) which can be built by the classical parametrix method, and satisfies the inequalities \(0\le\widetilde{G}\le K G\) for some real K>0. Here G is the fundamental solution of the heat equation \(\partial_{t}u=\alpha\,\partial^{2}_{xx}\/u\) for some constant α>0; see survey [5]. Recall that \(G(t,x,y)=\sum_{j\ge0} e^{-\alpha\, j^{2}\,t}\phi_{j}(x)\,\phi_{j}(y)\), where ϕ j denotes the eigenfunction (of operator \(-\alpha\,\partial^{2}_{xx}\/\) on S 1 with eigenvalue λ j =αj 2) satisfying \(\int_{S^{1}}\phi^{2}_{j}(x)\,dx=1\). Indeed, \(\widetilde{w}(t,x)=\int_{S^{1}} \widetilde{G}(t,x,y)\,(\varphi (u_{0}(y))-v_{\infty})\,dy\). In particular,

$$\bigl\|\widetilde{v}(t,\cdot)-v_\infty\bigr\|_{S^1}\le e^{-\alpha\, t}\bigl\|\widetilde{v}(0,\cdot)-v_\infty\bigr\|_{S^1}.$$

Thus, for all t>0 and xS 1 we have the a priori estimate

$$\bigl\|\varphi\bigl(u(t,\cdot)\bigr)-\varphi(u_\infty)\bigr\|_{S^1}\le K\,e^{-\alpha\,t}\bigl\| \varphi(u_0)-\varphi(u_\infty)\bigr\|_{S^1}.$$

From this, using inequalities

see (58), we deduce (57). □

Example 8

(i) Consider the heat equation over infinite (or semi-infinite) spatial interval ℝ,

$$ \partial_tu = \partial^2_{xx}\/u+ a(t,x),\qquad u(0,x)=u_0(x)\in L^2(\mathbb{R}).$$
(61)

Here we assume that u 0(x) is a bounded function. All the solutions of (61),

$$u(t,x)=\int_\mathbb{R}u_0(\xi) G(t,x,y)\,d\,y +\int _0^t\int_\mathbb{R}a(y,s)G(t-s,x,y)\,d\,y\,ds,$$

include the heat kernel \(G(t,x,y)={(4\pi t)^{-1/2}}\,e^{-(x-y)^{2}/(4t)}\) (the fundamental solution of homogeneous (61) for u 0(x)=δ x (ξ) – the Dirac delta function). For any function u 0L 2(ℝ), a unique solution of the homogeneous heat equation, u(t,x)=∫ u 0(y)G(t,x,y) dy, converges uniformly to a linear function, as t→∞. Indeed, if u 0 is bounded then the linear function is constant.

(ii) The rough Laplacian for tensors is defined by \(\Delta=\operatorname{div}\nabla\). For a unit circle S 1, the eigenvalues of −Δ are the numbers λ j =j 2 with eigenfunctions \(\phi_{j}(x)=\frac{1}{\sqrt{2\pi}}e^{-i j x}\), where j∈ℤ∖{0}. The heat equation on S 1,

$$ \partial_tu = \partial^2_{xx}\/u,\qquad u(0,\cdot)=u_0\in H^2\bigl(S^1\bigr),$$
(62)

has a unique solution for functions uC([0,∞), H 2(S 1))∩C 1((0,∞], L 2(S 1)). By the Sobolev embedding theorem, H 2(S 1)⊂C 1(S 1). The solution of (62) has the property that u(t,⋅)∈C (S 1) for all t>0. Moreover, \(u(t,\cdot)\to\bar{u}_{0}\) as t→∞, where \(\bar{u}_{0}=\frac{1}{2\pi}\int_{S^{1}} u_{0}(x)\,dx\) and

$$\bigl\|u(t,\cdot)-\bar{u}_0\bigr\|\le e^{-t}\|u_0-\bar{u}_0\|.$$

Consider the Jacobi theta function

$$\theta(z,\tau)=\sum_{n=1}^\infty e^{\pi i n^2\tau+2\pi i n z} =1+2\sum_{n=1}^\infty \bigl(e^{\pi i\tau}\bigr)^{n^2}\cos(2\pi n z),$$

where i 2=−1, z is a complex number, and τ is confined to the upper half-plane. Taking z=x∈ℝ and τ=4πti with real t>0, we write \(\theta(x, 4\pi t\,i)=1+2\sum_{n=1}^{\infty}e^{-4 \pi^{2} n^{2} t}\cos(2\pi n x)\), which satisfies the heat equation (62)1. Since lim t→0 θ(x,4πti)=∑ n∈ℤ δ(xn), the solution of (62) can be specified by convolving the periodic boundary condition at t=0 with θ(x,4πti).

(iii) Following [3], denote \(U(t,x)=\frac{\sin x}{\sqrt{\cos^{2}x+e^{2t}}}\ (t\ge0)\) and \(k(u)=\frac{1}{1+u^{2}}\). We have a 3-parameter family u(t,x)=U(ω 2(t+β 1),ω(x+β 2)) of exact solutions to the PDE (56), \(\partial_{t}u=\partial_{x}\/(k(u)\,\partial_{x}\/u))\) on S 1. Set ω=1 and β 1=β 2=0, hence u(t,x)=U(t,x) and \(u_{0}(x)=\frac{\sin x}{\sqrt{\cos^{2} x+1}}\). Indeed, lim t→∞ u(t,x)=u =0 for all xS 1. Since 0≤u 2≤1, we conclude that \(\frac{1}{2}\le k(u)\le1\). Finally, we have \(\|u(t,\cdot)\|_{S^{1}}\le e^{-t}\) and \(\|u_{0}\|_{S^{1}}=1\), which is consistent with (57).

1.2 A.2 The Generalized Companion Matrix

Let P n =k np 1 k n−1−⋯−p n−1 kp n be a polynomial over ℝ and k 1k 2≤⋯≤k n be the roots of P n for n>0. Hence, p i =(−1)i−1 σ i , where σ i are elementary symmetric functions of the roots k i . The following generalized companion matrix (see [9]) plays a key role in this work:

(63)

Lemma 6

(see [9])

The matrix (63) has the following properties:

  1. (a)

    The characteristic polynomial of B n,1 is P n .

  2. (b)

    \(v_{j}=(1,\,2\,k_{j},\,3\,k_{j}^{2},\ldots, n\,k_{j}^{n-1})\) is the eigenvector of B n,1 for the eigenvalue k j .

  3. (c)

    B n,1 V=VD, where \(V=\{\frac{n}{i} k_{j}^{i-1}\}_{1\le i,j\le n}\) is the Vandermonde type matrix, and D=diag(k 1,…,k n ) is a diagonal matrix. (If all k i ’s are distinct, then V −1 B n,1V=D.)

Proposition 13

(see [9])

Let τ i (t,x) (i∈ℕ) be the power sums of smooth functions k i (t,x) (1≤in). Given m>0, consider the infinite system of linear PDEs

$$ \partial_t\tau_i=-\frac{i\,m}{2(i+m-1)}\,\partial_x\/\tau_{i+m-1},\quad i\in \mathbb{N}.$$
(64)

Then the n-truncated (64), i.e., τ n+i ’s are eliminated using suitable polynomials of τ 1,…,τ n , is

$$\partial_t\overrightarrow{\!\tau} =- \frac{m}{2}\,(B_{n,1})^{m-1}\partial_x\/\overrightarrow{\!\tau}.$$

Remark that for m=1 the system (64) has diagonal form: \(\partial_{t}\tau_{i} =-\frac{1}{2}\,\partial_{x}\/\tau_{i}\), and for m=2 the n-truncated system (64) reads as \(\partial_{t}\tau_{i} =-B_{n,1}\,\partial_{x}\/\tau_{i}\).

Proposition 14

Let τ i (t,x) (i∈ℕ) be as in Proposition 13. Given m>0, consider the infinite system of linear PDEs

$$ \partial_t\tau_i=\frac{i\,m}{2(i+m-1)}\,\partial^2_{xx}\/\tau_{i+m-1},\quad i\in \mathbb{N}.$$
(65)

Then the n-truncated system (65) has the form

$$\partial_t\overrightarrow{\!\tau} = \frac{m}{2}\,(B_{n,1})^{m-1}\partial^2_{xx}\/\overrightarrow{\!\tau }+a_m(\overrightarrow{\!\tau},\partial_x\/\overrightarrow{\!\tau}).$$

Proof

By Proposition 13, we have \(\partial_{x}\/\tau_{n+i}=\sum_{j=1}^{n}\tilde{b}_{ij}\,\partial_{x}\/\tau_{j}\), where \(\tilde{b}_{ij}\) are the coefficients of the matrix B n,m−1. Derivation of this leads to \(\partial^{2}_{xx}\/\tau_{n+i}=\sum_{j=1}^{n}\tilde{b}_{ij}\,\partial^{2}_{xx}\/\tau_{j}+\partial_{x}\/\tilde{b}_{ij}\,\partial_{x}\/\tau_{j}\). □

Remark 5

For m=1 the system (65) has diagonal form: \(\partial_{t}\tau_{i}=\frac{1}{2}\,\partial^{2}_{xx}\/\tau_{i}\), and for m=2 the n-truncated system (65) reads as \(\partial_{t}\overrightarrow{\!\tau} =B_{n,1}\,\partial^{2}_{xx}\/\overrightarrow{\!\tau}\).

For small values of m, m=1,2, the terms \(\partial^{2}_{xx}\/\tau_{n+m}\) are given by

(66)
(67)

By Proposition 14, the last row of the matrix B n,1 or \(\frac{3}{2}(B_{n,1})^{2}\) consists of the coefficients at \(\partial^{2}_{xx}\/\tau_{i}\)’s on the right-hand side of (66), respectively, of (67), and so on.

Example 9

For f j =−2 δ j1, (65) reduces to the (system of) heat equations

$$\partial_t\tau_i =\partial^2_{xx}\/\tau_{i},\quad i\in \mathbb{N},$$

whose solution is known. Consider more complicated cases.

1. For f j =−δ j2, (65) reduces to the system

$$ \partial_t\tau_i =\frac{i}{i+1}\,\partial^2_{xx}\/\tau_{i+1},\quad i\in \mathbb{N},$$
(68)

whose n-truncated version reads as: \(\partial_{t}\overrightarrow{\!\!\tau}=B_{n,1}\,\partial^{2}_{xx}\/\overrightarrow{\!\!\tau}+a_{2}(\overrightarrow{\!\tau},\partial_{x}\/\overrightarrow{\!\tau})\). For n=2, we have two PDEs

and the matrix . If the roots k 1k 2, the eigenvectors of B 2,1 are v j =(1, 2 k j ), j=1,2. For n=3, (68) reduces to the quasilinear system of three PDEs with the matrix

$$B_{3,1}=\left ( \begin{array}{c@{\quad }c@{\quad }c}0 & \frac{1}{2} & 0 \\0 & 0 & \frac{2}{3} \\3\,\sigma_{3} & -\frac{3}{2}\,\sigma_{2} & \sigma_{1} \\\end{array} \right ),$$

whose eigenvalues are k j , and the eigenvectors are \(v_{j}=(1, 2\,k_{j}, 3\,k_{j}^{2})\).

2. For f j =−δ j3, (65) reduces to the system

For n=3, the system has the following matrix:

$$\frac{3}{2} (B_{3,1})^2 =\left ( \begin{array}{c@{\quad }c@{\quad }c}0 & 0 & \frac{1}{2} \\3\,\sigma_3 &-\frac{3}{2}\sigma_2 & \sigma_1 \\\frac{9}{2}\sigma_1\sigma_3 &\frac{9}{4}(\sigma_3-\sigma_1\sigma_2) &\frac{3}{2}(\sigma_1^2-\sigma_2) \\\end{array}\right ),$$

whose eigenvalues are \(\frac{3}{2}\,k_{j}^{2}\). This series of examples can be continued as long as one desires.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Rovenski, V. Extrinsic Geometric Flows on Codimension-One Foliations. J Geom Anal 23, 1530–1558 (2013). https://doi.org/10.1007/s12220-012-9297-1

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12220-012-9297-1

Keywords

Mathematics Subject Classification (2000)

Navigation