Skip to main content
Log in

Hypersurfaces with Light-Like Points in a Lorentzian Manifold

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

Abstract

Consider a constant mean curvature immersion \(F:U(\subset \varvec{R}^n)\rightarrow M\) into an arbitrary Lorentzian \((n+1)\)-manifold M. A point \(o\in U\) is called a light-like point if the first fundamental form \(\mathrm{d}s^2\) of F degenerates at o. We denote by \(B_F\) the determinant function of the symmetric matrix associated to \(\mathrm{d}s^2\) with respect to a local coordinate system at o. A light-like point o is said to be degenerate if the exterior derivative of \(B_F\) vanishes at o. We show that if o is a degenerate light-like point, then the image of F contains a light-like geodesic segment of M passing through f(o) (cf. Theorem E). This explains why several known examples of constant mean curvature hypersurface in the Lorentz–Minkowski \((n+1)\)-space form \(\varvec{R}^{n+1}_1\) contain light-like lines on their sets of light-like points, under a suitable regularity condition of F. Several related results are also 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. Akamine, S.: Causal characters of zero mean curvature surfaces of Riemann type in Lorentz-Minkowski 3-space. Kyushu J. Math. 71, 211–249 (2017)

    Article  MathSciNet  Google Scholar 

  2. Akamine, S., Singh, R.K.: Wick rotations of solutions to the minimal surface equation, the zero mean curvature surface equation and the Born-Infeld equation. arXiv:1711.00299

  3. Estudillo, F.J.M., Romero, A.: Generalized maximal surfaces in Lorentz–Minkowski space \({\mathbb{L}}^3\). Math. Proc. Camb. Phil. Soc. 111, 515–524 (1992)

    Article  Google Scholar 

  4. Fujimori, S., Kim, Y.W., Koh, S.-E., Rossman, W., Shin, H., Takahashi, H., Umehara, M., Yamada, K., Yang, S.-D.: Zero mean curvature surfaces in \({\bf L}^3\) containing a light-like line. C.R. Acad. Sci. Paris. Ser. I(350), 975–978 (2012)

    Article  Google Scholar 

  5. Fujimori, S., Rossman, W., Umehara, M., Yamada, K., Yang, S.-D.: Embedded triply periodic zero mean curvature surfaces of mixed type in Lorentz–Minkowski 3-space. Mich. Math. J. 63, 189–207 (2014)

    Article  MathSciNet  Google Scholar 

  6. Fujimori, S., Kim, Y.W., Koh, S.-E., Rossman, W., Shin, H., Umehara, M., Yamada, K., Yang, S.-D.: Zero mean curvature surfaces in Lorentz–Minkowski \(3\)-space and \(2\)-dimensional fluid mechanics. Math. J. Okayama Univ. 57, 173–200 (2015)

    MathSciNet  MATH  Google Scholar 

  7. Fujimori, S., Kim, Y. W., Koh, S.-E., Rossman, W., Shin, H., Umehara, M., Yamada, K., Yang, S.-D.: Zero mean curvature surfaces in Lorentz-Minkowski \(3\)-space which change type across a light-like line, Osaka J. Math. 52, 285–297 (2015). Erratum to the article “Zero mean curvature surfaces in Lorentz-Minkowski \(3\)-space which change type across a light-like line”. Osaka J. Math. 53, 289–293 (2016)

  8. Fujimori, S., Kawakami, Y., Kokubu, M., Rossman, W., Umehara, M., Yamada, K.: Zero mean curvature entire graphs of mixed type in Lorentz-Minkowski \(3\)-space. Q. J. Math. 67, 801–837 (2016)

    MathSciNet  MATH  Google Scholar 

  9. Fujimori, S., Kawakami, Y., Kokubu, M., Rossman, W., Umehara, M., Yamada, K.: Analytic extension of Jorge-Meeks type maximal surfaces in Lorentz-Minkowski 3-space. Osaka J. Math. 54, 249–272 (2017)

    MathSciNet  MATH  Google Scholar 

  10. Fujimori, S., Kawakami, Y., Kokubu, M., Rossman, W., Umehara, M., Yamada, K.: Analytic extension of exceptional constant mean curvature one catenoids in de Sitter 3-space. Math. J. Okayama Univ. arXiv:1507.06695v2 (to appear)

  11. Hashimoto, K., Kato, S.: Bicomplex extensions of zero mean curvature surfaces in \({\varvec {R}}^{2,1}\) and \({\varvec {R}}^{2,2}\)

  12. Honda, A., Koiso, M., Kokubu, M., Umehara, M., Yamada, K.: Mixed type surfaces with bounded mean curvature in \(3\)-dimensional space-times. Differ. Geom. Appl. 52, 64–77 (2017)

    Article  MathSciNet  Google Scholar 

  13. Klyachin, V.A.: Zero mean curvature surfaces of mixed type in Minkowski space. Izv. Math. 67, 209–224 (2003)

    Article  MathSciNet  Google Scholar 

  14. Kobayashi, O.: Maximal surfaces in the \(3\)-dimensional Minkowski space \(L^3\). Tokyo J. Math. 6, 297–309 (1983)

    Article  MathSciNet  Google Scholar 

  15. Thayer, J.: Notes on partial differential equations, Monografias de Mathemática 34., IMPA (1980)

  16. Umehara, M., Yamada, K.: Maximal surfaces with singularities in Minkowski space. Hokkaido Math. J. 35, 13–40 (2006)

    Article  MathSciNet  Google Scholar 

  17. Umehara, M., Yamada, K.: Surfaces with light-like points in Lorentz-Minkowski space with applications, in “Lorentzian Geometry and Related Topics”. Springer Proc. Math. Stat. 211, 253–273 (2017)

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank Toshizumi Fukui for fruitful discussions and the referees for valuable comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. Umehara.

Additional information

Umehara was partially supported by the Grant-in-Aid for Scientific Research (A) No. 26247005, and Yamada by (B) No. 17H02839 from Japan Society for the Promotion of Science.

Appendices

Appendix A: Fermi Coordinate Systems Along Light-Like Geodesics

Let (Mg) be a Lorentzian \((n+1)\)-manifold. In this appendix, we prove the following assertion:

Proposition A.1

Let \(I=(a,b)\) be a closed interval, and let \(\sigma :I\rightarrow M\) be a light-like geodesic. Then there exists a local diffeomorphism \((\epsilon >0\) is a constant)

$$\begin{aligned}&\Phi :\biggl \{(x_0,\dots ,x_n)\in \varvec{R}^{n+1}\,;\, x_0,x_n\in \left[ \frac{a+\epsilon }{\sqrt{2}},\frac{b-\epsilon }{\sqrt{2}}\right] , \\&\qquad \qquad \,\, |x_i|<\epsilon \,\, (i=1,\dots ,n-1) \biggr \} \rightarrow M \end{aligned}$$

such that

  1. (a1)

    \(\Phi (t,0,\dots ,t)=\sigma (t)\) holds for \(t\in \left[ {a}/{\sqrt{2}},{b}/{\sqrt{2}}\right] \),

  2. (a2)

    regarding \((x_0,\dots ,x_n)\) as a local coordinate system by \(\Phi \), \(g_{0,0}^{}=-1\),   \(g_{0,i}^{}=0\) along \(\sigma \) and \(g_{j,k}^{}=\delta _{j,k}\) hold for \(i,j,k=1,\dots ,n\) along \(\sigma \), where \(g_{j,k}^{}:=g(\partial _{x_j},\partial _{x_k})\), and

  3. (a3)

    all of the Christoffel symbols vanish along \(\sigma \). As a consequence, all derivatives \(\partial g_{\alpha ,\beta }^{}/\partial x_\gamma \)\((\alpha ,\beta ,\gamma =0,\dots ,n)\) vanish along \(\sigma \).

Proof

We can take vectors \(\varvec{e}_0,\dots ,\varvec{e}_n\in T_{\sigma (a)}M\) such that

  1. (1)

    \(\varvec{e}_0,\varvec{e}_1,\dots ,\varvec{e}_n\) (\(\varvec{e}_0:=\sigma '(a)\)) gives a basis of \(T_{\sigma (a)} M\),

  2. (2)

    \(g(\varvec{e}_0,\varvec{e}_0)= g(\varvec{e}_1,\varvec{e}_1)=0\) and \(g(\varvec{e}_0,\varvec{e}_1)=-1\),

  3. (3)

    \(g(\varvec{e}_0,\varvec{e}_i)=g(\varvec{e}_1,\varvec{e}_i)=0\) for \(i=2,\dots ,n\),

  4. (4)

    \(g(\varvec{e}_j,\varvec{e}_k)=\delta _{j,k}\) for \(j,k=2,\dots ,n\).

We let \(E_\alpha (t)\) (\(a\le t\le b\),   \(\alpha =0,\dots ,n\)) be the parallel vector field along \(\sigma (t)\) such that

$$\begin{aligned} E_\alpha (a)=\varvec{e}_\alpha \qquad (\alpha =0,\dots ,n). \end{aligned}$$

Since \(\sigma \) is a geodesic, \(\sigma '(t)=E_0(t)\) holds for \(t\in I\). We then set

$$\begin{aligned} \Phi :\varvec{R}\times \varvec{R}^n\ni (t,y_1,\dots ,y_n)\mapsto {{\text {Exp}}}_{\sigma (t)}\left( y_1E_1(t)+\cdots +y_n E_n(t)\right) \in M, \end{aligned}$$

where \({{\text {Exp}}}_p\) is the exponential map at \(p\in M\) with respect to the Lorentzian metric g. Then there exists \(\epsilon >0\) such that the restriction of \(\Phi \) to the open domain

$$\begin{aligned} \left\{ (y_0,y_1\dots ,y_n)\in \varvec{R}\times \varvec{R}^n;\, a<y_0<b,\quad |y_i|<\epsilon \,\,\, (i=1,\dots ,n)\right\} \end{aligned}$$

gives a diffeomorphism, where \(y_0:=t\). By definition, we have

$$\begin{aligned} E_{\alpha }(t) =\left( \frac{\partial }{\partial y_{\alpha }}\right) _{\sigma (t)}\qquad (\alpha =0,\dots ,n), \end{aligned}$$

and

  • \(g(\partial _{y_0},\partial _{y_0})= g(\partial _{y_1},\partial _{y_1})=0\) and \(g(\partial _{y_0},\partial _{y_1})=-1\),

  • \(g(\partial _{y_0},\partial _{y_i})=g(\partial _{y_1}, \partial _{y_i})=0\) for \(i=1,\dots ,n\) holds along \(\sigma \),

  • \(g(\partial _{y_j},\partial _{y_k})=\delta _{j,k}\) for \(j,k=2,\dots ,n\) along \(\sigma \).

Let D be the Levi-Civita connection of M. Then the Christoffel symbols with respect to this local coordinate system \((y_0,\dots ,y_n)\) are defined by

$$\begin{aligned} D_{\partial _{y_\alpha }}\partial _{y_\beta }= \sum _{\gamma =0}^n\Gamma _{\alpha ,\beta }^\gamma \partial _{y_\gamma } \qquad (\alpha ,\beta =0,\dots ,n). \end{aligned}$$

We would like to show that all of the Christoffel symbols \(\Gamma _{\alpha ,\beta }^\gamma \) (\(\alpha ,\beta ,\gamma =0,\dots ,n\)) with respect to the local coordinates \((y_0,\dots ,y_n)\) vanish along the \(y_0\)-axis (i.e., along the curve \(\sigma \)). We fix \((a_1,\dots ,a_n)\in \varvec{R}^n\setminus \{\varvec{0}\}\), and consider a curve \( c(t):=(0,a_1t,\dots ,a_n t) \) in M. Then, by our definition of the local coordinate system \((y_0,\dots ,y_n)\), this curve c(t) gives a geodesic on M. So \(c(t)=(c_0(t),\dots ,c_n(t))\) satisfies

$$\begin{aligned} c''_\gamma (t)+\sum _{\alpha ,\beta =0}^n \Gamma _{\alpha ,\beta }^\gamma (c(t)) c'_\alpha (t) c'_\beta (t)=0 \qquad (\gamma =0,\dots ,n). \end{aligned}$$

Since \(c'_0(t)=c''_0(t)=\cdots =c''_n(t)=0\), and \(c'_i(t)=a_i\) for \(i=1,\dots ,n\), this reduces to

$$\begin{aligned} \sum _{i,j=1}^n \Gamma _{i,j}^\gamma (c(t))a_i a_j=0 \qquad (\gamma =0,\dots ,n). \end{aligned}$$

If we set \(a_i=a_j=1\) and other \(a_k=0\) for \(k\ne i,j\), then we have

$$\begin{aligned} \Gamma _{i,j}^\gamma (c(t))+\Gamma _{j,i}^\gamma (c(t))=0 \qquad (i,j=1,\dots ,n,\,\,\gamma =0,\dots ,n). \end{aligned}$$

Since the connection is torsion free, \(\Gamma _{\alpha ,\beta }^\gamma =\Gamma _{\beta ,\alpha }^\gamma \) holds (\(0\le \alpha ,\beta ,\gamma \le n\)), and we get

$$\begin{aligned} \Gamma _{i,j}^\gamma (c(t))=0 \qquad (i,j=1,\dots ,n,\,\,\gamma =0,\dots ,n). \end{aligned}$$
(A.1)

On the other hand, since \(\sigma \) is a geodesic and \(E_0(t),\dots , E_n(t)\) are parallel vector fields along \(\sigma \), we have

$$\begin{aligned} \Gamma _{0,\beta }^\gamma (c(t))=0 \qquad (\beta ,\gamma =0,\dots ,n). \end{aligned}$$
(A.2)

By (A.1) and (A.2) together with the property \(\Gamma _{\alpha ,\beta }^\gamma =\Gamma _{\beta ,\alpha }^\gamma \), we can conclude that all of the Christoffel symbols along the curve \(\sigma \) vanish.

We now set

$$\begin{aligned} x_0:=\frac{y_0+y_1}{\sqrt{2}},\quad x_n:=\frac{y_0-y_1}{\sqrt{2}}, \quad x_i:=y_{i+1} \qquad (i=1,\dots ,n-1). \end{aligned}$$
(A.3)

Then the properties (a1) and (a2) for this new coordinate system \((x_0,\dots ,x_n)\) are obvious. Since the property that all of the Christoffel symbols vanish along \(\sigma \) is preserved under linear coordinate changes, (a3) is also obtained. \(\square \)

Appendix B: Computations in \(\varvec{R}^{n+1}_1\)

We denote by the dot ‘\(\cdot \)’ the canonical Lorentzian inner product of \(\varvec{R}^{n+1}_1\) with signature \((-+\cdots +)\). In this appendix, we compute \(B:=B_F\) and \(A:=A_F\) with respect to the canonical coordinate system \((x_0,x_1,\dots ,x_n)\) of \(\varvec{R}^{n+1}_1\). Let \(f(x_1,\dots ,x_n)\) be a \(C^2\)-function of n variables defined on a neighborhood of the origin \(o\in \varvec{R}^n\). We set

$$\begin{aligned} F=(f(x_1,\dots ,x_n),x_1,\dots ,x_n) \end{aligned}$$

and

$$\begin{aligned} \varvec{e}_0:=(1,0,\dots ,0),\quad \varvec{e}_1:=(0,1,\dots ,0),\quad \dots ,\quad \varvec{e}_n:=(0,0,\dots ,1). \end{aligned}$$

They give a canonical frame in \(\varvec{R}^{n+1}_1\) satisfying \(\varvec{e}_0\cdot \varvec{e}_0=-1\). Then \((x_1,\dots ,x_n)\) gives a local coordinate system of the domain of F, and we have

$$\begin{aligned} F_{x_i}=f_{x_i}\varvec{e}_0+\varvec{e}_i \qquad (i=1,\dots ,n). \end{aligned}$$

We set

$$\begin{aligned} S:=(s_{i,j}^{})_{i,j=1,\dots ,n},\qquad s_{i,j}^{}:=F_{x_i}\cdot F_{x_j} \end{aligned}$$

for \(i,j=1,\dots ,n\), where the dot ‘\(\cdot \) ’ is the canonical inner product of \(\varvec{R}^{n+1}_1\). Then we have

$$\begin{aligned} s_{i,j}^{}=\delta _{i,j}-f_{x_i}f_{x_j}\qquad \qquad (i,j=1,\dots ,n), \end{aligned}$$
(B.1)

and

$$\begin{aligned} S=I_n-(\nabla f)^T(\nabla f), \qquad \nabla f:=(f_{x_1},\dots ,f_{x_n}) \end{aligned}$$

hold, where \(I_n\) is the identity matrix and \((\nabla f)^T\) is the transpose of \((\nabla f)\). Then we have

$$\begin{aligned} B={{\text {det}}}(S)=(1-\lambda _1)\cdots (1-\lambda _n), \end{aligned}$$

where \(\lambda _1,\dots ,\lambda _n\) are eigenvalues of the matrix \((\nabla f)^T(\nabla f)\). Since \((\nabla f)^T(\nabla f)\) is of rank 1, we may assume that \(\lambda _2=\cdots =\lambda _n=0\). Then we have

$$\begin{aligned} B=1-\lambda _1=1-{{\text {trace}}}((\nabla f)^T(\nabla f)) =1-(f_{x_1})^2-\cdots -(f_{x_n})^2. \end{aligned}$$
(B.2)

Using this, it can be easily checked that the inverse matrix of S is given by

$$\begin{aligned} S^{-1}=I_n+\frac{1}{B}(\nabla f)^T(\nabla f). \end{aligned}$$

In particular, the cofactor matrix \(\tilde{S}=({\tilde{s}}^{i,j})_{i,j=1,\dots ,n}\) of S satisfies

$$\begin{aligned} \tilde{S}=B I_n+(\nabla f)^T(\nabla f), \end{aligned}$$
(B.3)

that is

$$\begin{aligned} {\tilde{s}}^{i,j}=B\delta _{i,j}+f_{x_i}f_{x_j}\qquad \qquad (i,j=1,\dots ,n), \end{aligned}$$
(B.4)

where \(\delta _{i,j}\) denotes Kronecker’s delta.

On the other hand,

$$\begin{aligned} {\tilde{\nu }}=-(1,f_{x_1},\dots ,f_{x_n}) \end{aligned}$$
(B.5)

gives a normal vector field defined by (2.5). Then the coefficients \({\tilde{h}}_{i,j}\) of the normalized second fundamental form given in (2.6) are written as

$$\begin{aligned} {\tilde{h}}_{i,j}=F_{x_i,x_j}\cdot {\tilde{\nu }}=f_{x_i,x_j}. \end{aligned}$$
(B.6)

Thus the matrix \({\tilde{h}}:=({\tilde{h}}_{i,j})_{i,j=1,\dots ,n}\) is just the Hessian matrix of f. By using the identity (B.2), the function A given in (2.6) can be computed as follows:

$$\begin{aligned} A&={{\text {trace}}}(\tilde{S} {\tilde{h}})= \sum _{i,j=1}^n (B\delta _{i,j}+f_{x_i}f_{x_j})f_{x_i,x_j}\\&=B \sum _{i=1}^n \left( f_{x_i,x_i} -\frac{1}{2} B_{x_i}f_{x_i}\right) =B \triangle f -\frac{1}{2} \nabla B\star \nabla f, \end{aligned}$$

where ‘\(\star \)’ is the canonical inner product of \(\varvec{R}^n\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Umehara, M., Yamada, K. Hypersurfaces with Light-Like Points in a Lorentzian Manifold. J Geom Anal 29, 3405–3437 (2019). https://doi.org/10.1007/s12220-018-00118-7

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12220-018-00118-7

Keywords

Mathematics Subject Classification

Navigation