Soliton Solutions of the Mean Curvature Flow and Minimal Hypersurfaces

2 Equivalence of the soliton and minimal hypersurface equation

2.1 Monge-Ampère systems

Let \(N\) be a \((2n+1)\)-dimensional manifold carrying a contact structure, meaning a maximally nonintegrable codimension 1 subbundle \(D\subset TN\) which we assume to be given by the kernel of a globally defined contact form \(\theta\). Recall that a \(n\)-dimensional submanifold \(f: \Sigma \to N\) which satisfies \(f^*\theta=0\) is called a Legendre submanifold of \((N,D)\). A Monge-Ampère differential system on \((N,D)\) is a differential ideal \(\mathcal{M} \subset \mathcal{A}^*(N)\) in the exterior algebra of differential forms on \(N\) given by \[\mathcal{M}=\left\{\theta,\mathrm{d}\theta,\varphi\right\},\] where \(\varphi \in \mathcal{A}^n(N)\) is a \(n\)-form.3 The brackets \(\{\;\}\) denote the algebraic span of the elements within, i.e. the elements of \(\mathcal{M}\) may be written as \[\alpha\wedge\theta+\beta\wedge \mathrm{d}\theta+\gamma\wedge \varphi,\] where \(\alpha,\beta,\gamma\) are differential forms on \(N\). Note that \(\mathcal{M}\) is indeed a differential ideal since \(\mathrm{d}\varphi\) lies in the contact ideal \(\mathcal{C}=\left\{\theta,\mathrm{d}\theta\right\}\), cf. [5]. A Legendre submanifold of \((N,D)\) which pulls-back to \(0\) the \(n\)-form \(\varphi\) as well will be called a Legendre integral manifold of \(\mathcal{M}\). Two Monge-Ampère systems \((N,\mathcal{M})\) and \((\bar{N},\bar{\mathcal{M}})\) are called equivalent if there exists a diffeomorphism \(\psi : N \to \bar{N}\) identifying the two ideals. Note that this implies that \(\psi\) is a contact diffeomorphism.

2.2 Minimal hypersurfaces via frames

In order to fix notation we review the description of minimal hypersurfaces using moving frames. For \(n \geqslant 1\), let \((M,g)\) be an oriented Riemannian \((n+1)\)-manifold, \(\pi : F \to M\) its right principal \(SO(n+1)\)-bundle of positively oriented orthonormal frames and \(\tau : U \to M\) its (sphere) bundle of unit tangent vectors. Write the elements of \(F\) as \((p,e_0,\ldots,e_n)\) where \(p \in M\) and \(e_0,\ldots,e_n\) is a positively oriented \(g\)-orthonormal basis of \(T_pM\). The Lie group \(SO(n+1)\) acts smoothly from the right by \[(p,e_0,\ldots,e_n)\cdot r=\left(p,\sum_{i=0}^n e_i r_{i0},\dots, \sum_{i=0}^n e_i r_{in}\right),\] where \(r_{ik}\) for \(i,k=0,\ldots,n\) denote the entries of the matrix \(r\). The map \(\nu : F \to U\), given by \((p,e_0,\ldots,e_n) \mapsto (p,e_0)\) is a smooth surjection whose fibres are the \(SO(n)\)-orbits and thus makes \(F\) together with its right action into a \(SO(n)\)-bundle over \(U\). Here we embed \(SO(n)\) as the Lie subgroup of \(SO(n+1)\) given by \[\left\{\left(\begin{array}{cc} 1 & 0 \\ 0 & r \end{array}\right) \in SO(n+1), r \in SO(n) \right\}.\] Let \(\omega_i \in \mathcal{A}^1(F)\) denote the tautological forms of \(F\) given by \[\left(\omega_i\right)_{(p,e_0,\ldots,e_n)}(\xi)=g_p\left(e_i,\pi^{\prime}(\xi)\right),\] and \(\omega_{ik} \in \mathcal{A}^1(F)\) the Levi-Civita connection forms which satisfy \(\omega_{ik}+\omega_{ki}=0\). The dual vector fields to the coframing \(\left(\omega_i,\omega_{ik}\right), i < k\), will be denoted by \(\left(\mathbf{W}_i,\mathbf{W}_{ik}\right)\). Recall that we have the structure equations \[\tag{2.1} \begin{aligned} \mathrm{d}\omega_i+\sum_{k=0}^n\omega_{ik}\wedge\omega_k&=0,\\ \mathrm{d}\omega_{ik}+\sum_{l=0}^n\omega_{il}\wedge\omega_{lk}&=\Omega_{ik}, \end{aligned}\] where \(\Omega_{ik} \in \mathcal{A}^2(F)\) are the curvature forms. Denote by \(\hat{\omega}_i\) the wedge product of the forms \(\omega_1,\ldots \omega_n\), with the \(i\)-th form omitted \[\hat{\omega}_i=\omega_1\wedge\cdots\wedge\omega_{i-1}\wedge\omega_{i+1}\wedge \cdots \wedge \omega_n.\] For \(n=1\) set \(\hat{\omega}_1\equiv1.\) Note that the forms \[\begin{aligned} \theta&=\omega_0,\\ \omega&=\omega_1\wedge \cdots \wedge \omega_n,\\ \mu&=-\frac{1}{n}\sum_{i=1}^n (-1)^{i-1}\omega_{0i} \wedge \hat{\omega}_i,\\ \end{aligned}\] are \(\nu\)-basic, i.e. pullbacks of forms on \(U\) which, by abuse of language, will also be denoted by \(\theta,\omega,\mu\). Since \[\tag{2.2} \mathrm{d}\omega_0=-\sum_{k=1}^n\omega_{0k}\wedge \omega_k\] the \(1\)-form \(\theta\) is a contact form. Note also that \[\tag{2.3} \mathrm{d}\omega+(-1)^{n-1}\,n\,\mu\wedge\theta=0.\] The geometric significance of these forms is the following: Suppose \(f : \Sigma \to M\) is an oriented hypersurface and \(\mathcal{G}_f : \Sigma \to U\) its orientation compatible Gauss lift. In other words the value of \(\mathcal{G}_f\) at \(p \in \Sigma\) is the unique unit vector at \(f(p)\) which is \(g\)-orthogonal to \(f^{\prime}(T_p\Sigma)\) and together with a positively oriented basis of \(T_p\Sigma\) induces the positive orientation of \(T_{f(p)}M\). By construction we have \[\tag{2.4} \mathcal{G}_f^*\theta=0\] and simple computations show that \[\tag{2.5} \mathcal{G}_f^*\omega=\omega_{f^*g},\] where \(\omega_{f^*g}\) denotes the Riemannian volume form on \(\Sigma\) induced by \(f^*g\). Suppose \(\tilde{f} : V \subset \Sigma\to F\) is a local framing covering \(\mathcal{G}_f\) and \(f\). Then pulling back (2.4) and using (2.2) gives \[\sum_{k=1}^n\tilde{f}^*{\omega_{0k}}\wedge \tilde{f}^*{\omega_k}=0.\] The independence (2.5) implies that the forms \(\varepsilon_i=\tilde{f}^*\omega_i\) are linearly independent and thus Cartan’s lemma yields the existence of functions \(h_{ik} : V \to \mathbb{R}\), symmetric in the indices \(i,k\), such that \[\tilde{f}^*\omega_{0i}=\sum_{k=1}^n h_{ik} \varepsilon_k.\] In particular we have \[\tag{2.6} \mathcal{G}_f^*\mu=-H \varepsilon_1\wedge\cdots\wedge\varepsilon_n,\] where \(H=\frac{1}{n}\sum_{i=1}^n h_{ii}\) is the mean curvature of the hypersurface \(f : \Sigma \to M\). Conversely if \(\mathcal{G} : N \to U\) is an orientable \(n\)-submanifold with \(\mathcal{G}^*\theta=0\) and \(\mathcal{G}^*\omega \neq 0\), then \(\tau \circ \mathcal{G} : N \to M\) is an immersion. Shrinking \(N\) if necessary we can assume that \(f=\tau \circ \mathcal{G} : N \to M\) is a hypersurface which can be oriented in such a way that its Gauss lift agrees with \(\mathcal{G}\). Thus the Legendre integral manifolds \(\mathcal{G} : \Sigma \to U\) of the M.A.S. \(\mathcal{M}_g\) on \(U\) given by \[\mathcal{M}_g=\left\{\theta,\mathrm{d}\theta,\mu\right\}\] which satisfy the transversality conditions \(\mathcal{G}^*\omega\neq0\) locally correspond to minimal hypersurfaces on \((M,g)\).

2.3 \(\mathbf{X}\)-pseudosoliton hypersurfaces via frames

Given a vector field \(\mathbf{X}\) on \(M\) define the functions \(X_i : F \to \mathbb{R}\) by \[\tag{2.7} (p,e_0,\ldots,e_n) \mapsto g_p(\mathbf{X}(p),e_i).\] Of course \(X_0\) is the \(\nu\)-pullback of a function on \(U\) which will be denoted by \(X\). Using (1.1) and (2.6) it follows that an oriented hypersurface \(f : \Sigma \to M\) is a \(\mathbf{X}\)-pseudosoliton hypersurface if and only if \[\mathcal{G}_f^*\left(\mu-X\omega\right)=0.\] Thus the Legendre integral manifolds \(\mathcal{G} : \Sigma \to U\) of the M.A.S. \(\mathcal{M}_{g,\mathbf{X}}\) on \(U\) given by \[\mathcal{M}_{g,\mathbf{X}}=\left\{\theta,\mathrm{d}\theta,\mu-X\omega\right\}\] which satisfy the transversality conditions \(\mathcal{G}^*\omega\neq0\) locally correspond to \(\mathbf{X}\)-pseudosoliton hypersurfaces on \((M,g)\). Now suppose \(\mathbf{X}\) is a gradient vector field \(\mathbf{X}=\nabla_g u\) for some smooth function \(u : M \to \mathbb{R}\). Let \(\bar{g}=e^{-2u}g\), \(\bar{\pi} : \bar{F} \to M\) denote the bundle of positively oriented \(\bar{g}\)-orthonormal frames with canonical coframing \(\bar{\omega}_i, \bar{\omega}_{ik}\) and \(\tilde{\psi} : F \to \bar{F}\) the map which scales a \(\bar{g}\)-orthonormal frame by \(e^{u}\). Then by definition \[\tag{2.8} \tilde{\psi}^*\bar{\omega}_i=e^{-u}\omega_i,\] and the structure equations (2.1) yield \[\tag{2.9} \tilde{\psi}^*\bar{\omega}_{ik}=\omega_{ik}+u_k\omega_i-u_i\omega_k,\] where we expand \(\pi^*\mathrm{d}u=\sum_{k=0}^n u_k \omega_k\) for some smooth functions \(u_k : F \to \mathbb{R}\). Note that \(u_0\) is the \(\nu\)-pullback of the function \(X\). Let \(\bar{\tau} : \bar{U} \to M\) denote the \(\bar{g}\)-unit tangent bundle with canonical forms \(\bar{\mu},\bar{\omega}\) and \(\psi : U \to \bar{U}\) the map which scales a \(g\)-unit vector by \(e^u\). Then (2.8) implies \[\psi^*\bar{\omega}=e^{-nu}\omega,\] thus \(\psi\) is a contact diffeomorphism. Moreover (2.8) and (2.9) yield \[\tag{2.10} \begin{aligned} \psi^*\bar{\mu}&=-\frac{e^{-(n-1)u}}{n}\sum_{k=1}^n (-1)^{k-1}\left(\omega_{0k}+u_k\theta-u_0\omega_k\right)\wedge \hat{\omega}_k\\ &=-e^{-(n-1)u}\left(\mu-X\omega+\frac{1}{n}\theta\wedge\left(i_{\nabla_g u}\omega\right)\right)\\ \end{aligned}\] which can be written as \(\alpha\wedge\theta+\gamma\wedge(\mu-X\omega)\) for some \((n-1)\)-form \(\alpha\) and some smooth real-valued function \(\gamma\) on \(U\). This yields \[\psi^*\mathcal{M}_{e^{-2u}g}=\mathcal{M}_{g,\nabla_{g} u}.\] Summarising we have proved the

Proposition 2.1

Let \((M,g)\) be an oriented Riemannian manifold and \(\mathbf{X}=\nabla_g u\) a gradient vector field. Then the \(\mathbf{X}\)-pseudosoliton M.A.S. on \((M,g)\) is equivalent to the minimal hypersurface M.A.S. on \((M,e^{-2u}g)\).

2.4 The non-gradient case

Proposition 2.1 raises the question if there still exists a contact equivalence between minimal hypersurfaces and solitons if \(\mathbf{X}\) is not a gradient vector field. We will argue next that this is not possible for \(n \geqslant 2\), so assume in this subsection that \(n\geqslant 2\). Before providing the arguments we recall a result from symplectic linear algebra. Suppose \((V,\Theta)\) is a symplectic vector space of dimension \(2n\), i.e. \(\Theta \in \Lambda^2(V^*)\) is non-degenerate. If a form \(\beta\) of degree \(s \leqslant p\) satisfies \[\tag{2.11} \beta\wedge\Theta^{(n-p)}=0,\] then \(\beta=0\). This is a corollary of the Lepage decomposition theorem for \(p\)-forms on symplectic vector spaces. (cf. [10]). Of course in our setting the symplectic vector spaces are the fibres of the contact subbundle \(D\) and \(\Theta\) is obtained by restricting \(\mathrm{d}\theta\) to \(D\).

Lemma 2.2

A necessary condition for the \(\mathbf{X}\)-pseudosoliton M.A.S. to be equivalent to the minimal hypersurface M.A.S. is the existence of an exact \(1\)-form \(\rho\) such that \[\mathrm{d}\left(\left(\mu-X\omega\right)\wedge \theta\right)=\rho\wedge\left(\mu-X\omega\right)\wedge\theta\]

Proof. Write \(\varphi=\mu-X\omega\) and suppose there exists a Riemannian metric \(\bar{g}\) and a diffeomorphism \(\psi : U \to \bar{U}\) such that \(\psi^*\mathcal{M}_{\bar{g}}=\mathcal{M}_{g,\mathbf{X}}\). Then \[\tag{2.12} \psi^*\bar{\mu}=\alpha\wedge\theta+\beta\wedge \mathrm{d}\theta+\gamma\wedge\varphi,\] where \(\alpha\) is a \((n-1)\)-form, \(\beta\) a \((n-2)\)-form and \(\gamma\) a smooth real-valued function on \(U\). Note that we have \[\tag{2.13} \begin{aligned} 0&=\varphi \wedge \mathrm{d}\theta,\\ 0&=\bar{\mu}\wedge \mathrm{d}\bar{\theta}. \end{aligned}\] Wedging (2.12) with \(\psi^*\mathrm{d}\bar{\theta}\), using (2.13) and that \(\psi\) is a contact diffeomorphism gives \[\tag{2.14} \left(\beta\wedge\mathrm{d}\theta\wedge \mathrm{d}\theta\right)\vert_{D}=0,\] where \(\vert_{D}\) denotes the restriction to the contact subbundle \(D\subset TU\). For \(n=2\) equation (2.14) implies \(\beta=0\). For \(n \geqslant 3\) it follows with (2.11) and (2.14) that \(\beta\vert_{D}=0\), thus there exists a \((n-3)\)-form \(\beta^{\prime}\) such that \[\beta=\beta^{\prime}\wedge\theta.\] We can therefore assume that there exists a \((n-1)\)-form \(\alpha^{\prime}\) such that \[\tag{2.15} \psi^*\bar{\mu}=\alpha^{\prime}\wedge\theta+\gamma\wedge\varphi.\] Wedging both sides of (2.15) with \(\psi^*\bar{\theta}\) gives \[\psi^*\left(\bar{\mu}\wedge\bar{\theta}\right)=\left(\alpha^{\prime}\wedge\theta+\gamma\wedge\varphi\right)\wedge\psi^*\bar{\theta}.\] this is equivalent to \[\psi^*\left(\bar{\mu}\wedge\bar{\theta}\right)=\tilde{\gamma}\wedge\varphi\wedge\theta\] for some smooth non-vanishing real-valued function \(\tilde{\gamma}\). Since \(\bar{\mu}\wedge\bar{\theta}\) is an exact form, see (2.3), we must have \[\mathrm{d}\xi=\mathrm{d}f\wedge\xi,\] where we have written \(\xi=\varphi\wedge\theta\) and \(f=-\ln \vert \tilde{\gamma}\vert\).

Using this Lemma we can proof the

Theorem 2.3

The \(\mathbf{X}\)-pseudosoliton M.A.S. on an oriented Riemannian manifold \((M,g)\) of dimension \(n+1\geqslant 3\) is equivalent to a minimal hypersurface M.A.S. if and only if \(\mathbf{X}\) is a gradient vector field.

Remark 2.4

Before giving the proof we point out identities which hold for the functions \(X_i\) (recall (2.7) for their definition). Since the \(1\)-form \(O=(\omega_{ik}) \in \mathcal{A}^1(F,\mathfrak{so}(n+1))\) is a connection we have \(O(\mathbf{W}_v)=v\), where \(\mathbf{W}_v\) is the vector field obtained by differentiating the flow \[\left((p,e_0,\ldots,e_n),t\right) \mapsto (p,e_0,\ldots,e_n) \cdot \exp(tv)\] and \(v \in \mathfrak{so}(n+1)\), the Lie algebra of \(SO(n+1)\). In particular this implies that the time \(t\) flow of the vector field \(\mathbf{W}_{ik}\) for \(i<k\) maps the frame \[(p,e_0,\ldots,e_i,\ldots,e_k,\ldots,e_n)\] to the frame \[\left(p,e_0,\ldots,\cos (t) e_i-\sin (t)e_k,\ldots, \sin (t)e_i+\cos (t)e_k,\ldots, e_n\right)\] and thus \[\tag{2.16} \mathcal{L}_{\mathbf{W}_{ik}}X_j=\delta_{jk}X_i-\delta_{ij}X_k,\] where \(\mathcal{L}\) stands for the Lie-derivative.

Proof of Theorem 2.3. We have \[\mathrm{d}X_0=\sum_{i=0}^n P_i \omega_i-\sum_{k=1}^n X_k\omega_{0k}\] for some smooth functions \(P_i : F \to \mathbb{R}\). From this it follows with straightforward computations that the \(1\)-forms \(\rho\) on \(U\) which satisfy \(\mathrm{d}\xi=\rho \wedge \xi\) pull-back to \(F\) to become \[\tag{2.17} \nu^*\rho=\lambda \omega_0+n\sum_{k=1}^n X_k \omega_k\] for a smooth function \(\lambda : F \to \mathbb{R}\). Differentiating (2.17) gives \[\nu^*\mathrm{d}\rho=\mathrm{d}\lambda\wedge\omega_0-\lambda\sum_{k=1}^n\omega_{0k}\wedge\omega_k+n\sum_{k=1}^n\mathrm{d}X_k\wedge\omega_k-n\sum_{i=0}^n\sum_{k=1}^nX_k\omega_{ki}\wedge\omega_i.\] Wedging with \(\omega_0\wedge\hat{\omega}_1\) yields \[\nu^*\mathrm{d}\rho\wedge\omega_0\wedge\hat{\omega}_1=\left(\lambda\, \omega_{01}-n\,\mathrm{d}X_1-n\,\sum_{k=1}^n X_k\,\omega_{1k}\right)\wedge\omega_0\wedge\omega.\] Using (2.16) we can expand \[\begin{aligned} \mathrm{d}X_1\wedge\omega_0\wedge\omega&=\left(\left(\mathcal{L}_{\mathbf{W}_{01}}X_1\right)\omega_{01}+\sum_{k=1}^n\left(\mathcal{L}_{\mathbf{W}_{1k}}X_1\right)\omega_{1k}\right)\wedge\omega_0\wedge\omega\\ &=\left(X_0\,\omega_{01}-\sum_{k=1}^nX_k\,\omega_{1k}\right)\wedge\omega_0\wedge\omega.\\ \end{aligned}\] Concluding we get \[\nu^*\mathrm{d}\rho\wedge\omega_0\wedge\hat{\omega}_1=(\lambda-nX_0)\,\omega_{01}\wedge\omega_0\wedge\omega.\] Suppose the \(\mathbf{X}\)-pseudosoliton M.A.S. on \((M,g)\) is equivalent to a minimal hypersurface M.A.S. Then, by Lemma 2.2, \(\rho\) has to be exact, this implies \[\lambda=nX_0\] and thus \[\nu^*\rho=n\sum_{i=0}^n X_i \omega_i.\] Note that if \(\chi \in TF\) is a vector tangent to the frame \((p,e_0,\ldots,e_n)\) we have \[\sum_{i=0}^n\left(X_i\omega_i\right)(\chi)=\sum_{i=0}^ng_p\left(g_p(\mathbf{X}(p),e_i)e_i,\pi^{\prime}(\chi)\right)=g_p\left(\mathbf{X}(p),\pi^{\prime}(\chi)\right),\] hence \[\rho=n\,\tau^*\left(\mathbf{X}^{\flat}\right),\] where \(\mathbf{X}^{\flat}\) denotes the \(g\)-dual \(1\)-form to \(\mathbf{X}\). The \(1\)-form \(\rho\) is exact and thus \(\rho=\mathrm{d}f\) for some real-valued function \(f\) on \(U\) which is locally constant on the fibres of \(\tau : U \to M\). Since the \(\tau\)-fibres are connected, it follows that \(f\) is constant on the \(\tau\)-fibres and thus equals the pullback of a smooth function \(u\) on \(M\) for which \[du=n\,\mathbf{X}^{\flat}.\] In other words \(\mathbf{X}\) is a gradient vector field. Conversely if \(\mathbf{X}\) is a gradient vector field, then the \(\mathbf{X}\)-pseudosoliton M.A.S. on \((M,g)\) is equivalent to the minimal hypersurface M.A.S. on \((M,e^{-2u}g)\) by Proposition 2.1.

Remark 2.5

In [5], Bryant, Griffiths and Grossman study the calculus of variations on contact manifolds in the setting of differential systems. In particular they give necessary and sufficient conditions for a M.A.S. to be locally of Euler-Lagrange type, i.e. locally equivalent to a M.A.S. whose Legendre integral manifolds correspond to the solutions of a variational problem. In fact, if one replaces Lemma 2.2 with [5] a proof along the lines of Theorem 2.3 shows that for \(n \geqslant 2\) the \(\mathbf{X}\)-pseudosoliton M.A.S. is locally equivalent to a M.A.S. of Euler-Lagrange type if and only if \(\mathbf{X}\) is a gradient vector field.

2.5 The surface case

Recall that in the case \(n=1\) of a surface \((M,g)\), the solutions of the \(\mathbf{X}\)-pseudosoliton equation (1.1) are immersed curves on \(M\) which may be repa- rametrised to become geodesics of a Weyl connection. In his Ph.D. thesis [12], the second author has constructed a \(10\)-parameter family of Weyl connections on the \(2\)-sphere whose geodesics are the great circles, and thus in particular projectively equivalent to the Levi-Civita connection of the standard spherical metric. Inspection shows that there are Weyl connections in this \(10\)-parameter family whose vector field is not a gradient and thus they provide counterexamples to Theorem 2.3 in the surface case.

This raises the question what the necessary and sufficient conditions for the \(\mathbf{X}\)-pseudosolitons curves are, in order to be the geodesics of a Riemannian metric. In [12] it was also shown that on a surface locally every affine torsion-free connection is projectively equivalent to a Weyl connection. Finding the necessary and sufficient conditions thus comes down to finding the necessary and sufficient conditions for an affine torsion-free connection to be projectively equivalent to a Levi-Civita connection. Therefore the conditions follow by applying the results in [4] and we refer the reader to this source for further details.