This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Null hypersurfaces as wave fronts in Lorentz-Minkowski space

S. Akamine, A. Honda, M. Umehara and K. Yamada Department of Liberal Arts, College of Bioresource Sciences, Nihon University,1866 Kameino, Fujisawa, Kanagawa, 252-0880, Japan [email protected] Department of Applied Mathematics, Faculty of Engineering, Yokohama National University, 79-5 Tokiwadai, Hodogaya, Yokohama 240-8501, Japan [email protected] Department of Mathematical and Computing Sciences, Tokyo Institute of Technology, Tokyo 152-8552, Japan [email protected] Department of Mathematics, Tokyo Institute of Technology, Tokyo 152-8551, Japan [email protected]
(Date: June 10, 2024)
Abstract.

In this paper, we show that “LL-complete null hypersurfaces” (i.e. ruled hypersurfaces foliated by entirety of light-like lines) as wave fronts in the (n+1)(n+1)-dimensional Lorentz-Minkowski space are canonically induced by hypersurfaces in the nn-dimensional Euclidean space. As an application, we show that most of null wave fronts can be realized as restrictions of certain LL-complete null wave fronts. Moreover, we determine LL-complete null wave fronts whose singular sets are compact.

Key words and phrases:
light-like hypersurface, null hypersurface, singular point, wave front
2010 Mathematics Subject Classification:
Primary 53C50; Secondary 53C42, 53B30.
The first author was partially supported by the Grant-in-Aid for Young Scientists No. 19K14527. The second author was partially supported by the Grant-in-Aid for Young Scientists No. 19K14526 and (B) No.  20H01801. The third and fourth authors were partially supported by (B) No.  21H00981 and (B) No.  17H02839, respectively, from Japan Society for the Promotion of Science.

Introduction

We denote by 1n+1{\mathbb{R}}^{n+1}_{1} the (n+1)(n+1)-dimensional Lorentz-Minkowski space. A null hypersurface in 1n+1{\mathbb{R}}^{n+1}_{1} is a CC^{\infty}-immersion whose induced metric degenerates everywhere. Such a hypersurface is also called a light-like hypersurface and is locally foliated by light-like lines (cf. [2, Fact 2.6]). Roughly speaking, a null hypersurface is said to be LL-complete if each light-like line is the entirety of a straight line in 1n+1{\mathbb{R}}^{n+1}_{1} (see Definition 2.1 for details).

In the authors’ previous work [2], it was shown that LL-complete null immersed hypersurfaces in 1n+1{\mathbb{R}}^{n+1}_{1} are totally geodesic. As is clear from this previous work, the study of global properties of such surfaces, considering only immersions is too restrictive. In this paper, we shall investigate the global behavior of null hypersurfaces with singular points in 1n+1{\mathbb{R}}^{n+1}_{1}.

More precisely, instead of immersions, we shall introduce null hypersurfaces as wave fronts (see Section 1). It can be easily observed that each hypersurface in the nn-dimensional Euclidean space 0n{\mathbb{R}}^{n}_{0} (n2n\geq 2) induces an associated parallel family, and this family can be considered as a section of a null hypersurface in 1n+1{\mathbb{R}}^{n+1}_{1}. For example, the light-cone

(0.1) Λn:={(t,x1,,xn)1n+1;(x1)2++(xn)2=t2}\Lambda^{n}:=\Big{\{}(t,x_{1},\ldots,x_{n})\in{\mathbb{R}}^{n+1}_{1}\,;\,(x_{1})^{2}+\cdots+(x_{n})^{2}=t^{2}\Big{\}}

is a typical example of an LL-complete null wave front, which corresponds to the family of parallel hypersurfaces of the unit sphere Sn1S^{n-1} in 0n{\mathbb{R}}^{n}_{0} centered at the origin. The origin as the cone-like singular point of Λn\Lambda^{n} corresponds to the parallel hypersurface of Sn1S^{n-1} just shrinking to a point. In general, a null hypersurface may have various singular points. For example, for each a(0,1]a\in(0,1], we consider an ellipse

(0.2) γ(θ):=(acosθ,sinθ)(0θ2π),\gamma(\theta):=(a\cos\theta,\,\sin\theta)\qquad(0\leq\theta\leq 2\pi),

in the Euclidean plane. Let 𝐧(θ){\mathbf{n}}(\theta) be the leftward unit normal vector field along γ\gamma. Then the map Fa:×(/2π)13F_{a}:{\mathbb{R}}\times({\mathbb{R}}/2\pi{\mathbb{Z}})\to{\mathbb{R}}^{3}_{1} defined by

(0.3) Fa(t,θ):=(0,γ(θ))+t(1,𝐧(θ))(a(0,1])F_{a}(t,\theta):=(0,\gamma(\theta))+t(1,{\mathbf{n}}(\theta))\qquad(a\in(0,1])

gives a null wave front having cuspidal edges and four swallowtails whenever 0<a<10<a<1 (the definitions of cuspidal edges and swallowtails are given in [3]). Each slice of the image of FaF_{a} by the horizontal plane t=t0t=t_{0} corresponds to the parallel curve of the ellipse γ\gamma of equi-distance t0t_{0}. When a=1a=1, the image of F1F_{1} just coincides with the light-cone Λ2\Lambda^{2} of 13{\mathbb{R}}^{3}_{1}.

Refer to caption
Figure 1. The complete null wave front associated with the family of parallel curves of the ellipse of a=1/2a=1/2.

In Section 2, we prove the converse assertion (cf. Theorem 2.6) as a fundamental theorem of null wave fronts, which states that any LL-complete null wave fronts in 1n+1{\mathbb{R}}^{n+1}_{1} are induced from wave fronts in 0n{\mathbb{R}}^{n}_{0}. As a consequence, we can say that an LL-complete null wave front is associated with the parallel family of a wave front in 0n{\mathbb{R}}^{n}_{0}.

We give two applications of this fundamental theorem. In Section 3, we show a “structure theorem of null hypersurfaces” (cf. Theorem 3.4) which asserts that most of null wave fronts (for example, real analytic null wave fronts and CC^{\infty} null wave fronts with a certain genericity) can be obtained as restrictions of LL-complete null wave fronts. There are several interesting geometric structures on null hypersurfaces in 1n+1{\mathbb{R}}^{n+1}_{1} without assuming LL-completeness (cf. [4, 7]). The authors hope that this theorem might play an important role in the further study of null hypersurfaces.

On the other hand, an LL-complete null wave front in 1n+1{\mathbb{R}}^{n+1}_{1} is called complete if its singular set is a non-empty compact subset in the domain of definition. For example, the light-cone Λn\Lambda^{n} in 1n+1{\mathbb{R}}^{n+1}_{1} and the null wave front FaF_{a} in 13{\mathbb{R}}^{3}_{1} given in (0.3) are complete. In Section 4, as the deepest application of the fundamental theorem, we show that each of complete null wave fronts corresponds to a parallel family of a closed convex hypersurface in 0n{\mathbb{R}}^{n}_{0} if n3n\geq 3 (cf. Theorem 4.4). When n=2n=2, we show that complete null wave fronts are induced by parallel families of locally convex closed regular curves in the Euclidean plane 02{\mathbb{R}}^{2}_{0}. If the curve is an ellipse, the null wave front corresponds to FaF_{a} as above (see Figure 1).

It should be remarked that the classical four vertex theorem for closed convex curves implies the existence of four non-cuspidal edge singular points on complete null wave fronts with embedded ends in 13{\mathbb{R}}^{3}_{1} (cf. Corollary 4.6).

There is one point to note when reading this paper: Readers who are interested only in the fundamental theorem for LL-complete null wave fronts or complete null wave fronts can skip Section 3. In fact, Section 4 is devoted to properties of complete null wave fronts and does not refer to Section 3. Appendix A of this paper is required for Section 2, but Appendix B is used for Section 3, so such readers also do not need to read Appendix B.

1. Properties of null wave fronts in 1n+1{\mathbb{R}}^{n+1}_{1}

We first recall the definition of wave fronts. Let (n+1)({\mathbb{R}}^{n+1})^{*} be the dual vector space of n+1{\mathbb{R}}^{n+1}, and we denote by P(n+1)P^{*}({\mathbb{R}}^{n+1}) the projective space associated with (n+1)({\mathbb{R}}^{n+1})^{*}. We let

π:(n+1){𝟎}P(n+1)\pi:({\mathbb{R}}^{n+1})^{*}\setminus\{{\mathbf{0}}\}\to P^{*}({\mathbb{R}}^{n+1})

be the canonical projection. Then, for each element ζP(n+1)\zeta\in P^{*}({\mathbb{R}}^{n+1}), there exists a linear function ω:n+1\omega:{\mathbb{R}}^{n+1}\to{\mathbb{R}} (which is an element of (n+1){𝟎}({\mathbb{R}}^{n+1})^{*}\setminus\{{\mathbf{0}}\}) such that π(ω)=ζ\pi(\omega)=\zeta. Since the kernel of ω\omega is invariant under non-zero scalar multiplications, the nn-dimensional subspace

Ker(ζ):=Ker(ω)\operatorname{Ker}(\zeta):=\operatorname{Ker}(\omega)

is well-defined.

In this paper, we set r=r=\infty or r=ωr=\omega and “CrC^{r}” means smoothness if r=r=\infty and real analyticity if r=ωr=\omega. We fix a CrC^{r}-differentiable nn-manifold MnM^{n}.

Definition 1.1.

Let F:Mnn+1F:M^{n}\to{\mathbb{R}}^{n+1} be a CrC^{r}-map. Then the map FF is called a CrC^{r}-wave front (or simply a CrC^{r}-front) if for each pMnp\in M^{n}, there exist a neighborhood UU of pp and a CrC^{r}-map α~:U(n+1){𝟎}\tilde{\alpha}:U\to({\mathbb{R}}^{n+1})^{*}\setminus\{{\mathbf{0}}\} such that

  1. (1)

    α:=πα~\alpha:=\pi\circ\tilde{\alpha} satisfies (dF)q(𝐯)Ker(αq)(qU,𝐯TqMn)(dF)_{q}({\mathbf{v}})\in\operatorname{Ker}(\alpha_{q})\,\,(q\in U,\,\,{\mathbf{v}}\in T_{q}M^{n}), and

  2. (2)

    L:=(F,α):Un+1×P(n+1)L:=(F,\alpha):U\to{\mathbb{R}}^{n+1}\times P^{*}({\mathbb{R}}^{n+1}) is an immersion.

Moreover, if we can take U=MnU=M^{n}, the map FF is said to be co-orientable. In this case, α:MnP(n+1)\alpha:M^{n}\to P^{*}({\mathbb{R}}^{n+1}) is called the Gauss map of FF, and the map α~:Mn(n+1){𝟎}\tilde{\alpha}:M^{n}\to({\mathbb{R}}^{n+1})^{*}\setminus\{{\mathbf{0}}\} is called the lift of α\alpha.

If a CrC^{r}-wave front FF is not co-orientable, taking the double covering π^:M^nMn\hat{\pi}:\hat{M}^{n}\to M^{n}, the composition Fπ^F\circ\hat{\pi} becomes a co-orientable wave front. So without loss of generality, we may assume that FF itself is co-orientable.

We let 1n+1{\mathbb{R}}^{n+1}_{1} be Lorentz-Minkowski (n+1)(n+1)-space of signature (++)(-+\cdots+), and denote by ,\left\langle{\,},{\,}\right\rangle the canonical Lorentzian inner product on 1n+1{\mathbb{R}}^{n+1}_{1}.

Proposition 1.2.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a ((co-orientable)) CrC^{r}-wave front. Then there exists a vector field ξ^\hat{\xi} without zeros ((which can be considered as a map ξ^:Mn1n+1{𝟎})\hat{\xi}:M^{n}\to{\mathbb{R}}^{n+1}_{1}\setminus\{{\mathbf{0}}\}) such that

ξ^p,dF(𝐯)=0(pMn,𝐯TpMn).\left\langle{\hat{\xi}_{p}},{dF({\mathbf{v}})}\right\rangle=0\qquad(p\in M^{n},\,\,{\mathbf{v}}\in T_{p}M^{n}).

A vector field ξ^\hat{\xi} along FF given in Proposition 1.2 is called a normal vector field along FF.

Proof.

Let α~\tilde{\alpha} be a lift of the Gauss map of FF. Since ,\left\langle{\,},{\,}\right\rangle is non-degenerate, there exists a vector field ξ^\hat{\xi} on MnM^{n} along FF such that

(1.1) ξ^p,𝐱=α~p(𝐱)(𝐱TF(p)1n+1).\left\langle{\hat{\xi}_{p}},{{\mathbf{x}}}\right\rangle=\tilde{\alpha}_{p}({\mathbf{x}})\qquad({\mathbf{x}}\in T_{F(p)}{\mathbb{R}}^{n+1}_{1}).

Since αp𝟎\alpha_{p}\neq{\mathbf{0}} for each pp, the vector field ξ^\hat{\xi} has no zeros on MnM^{n}. ∎

We then introduce ‘null wave fronts’ as follows:

Definition 1.3.

In the setting of Proposition 1.2, the CrC^{r}-wave front F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} is said to be null or light-like if ξ^p\hat{\xi}_{p} points in the light-like direction in 1n+1{\mathbb{R}}^{n+1}_{1} at each pMnp\in M^{n}.

We denote by (,)E(\,,\,)_{E} the canonical positive definite inner product on n+1(=1n+1){\mathbb{R}}^{n+1}(={\mathbb{R}}^{n+1}_{1}).

Definition 1.4 (EE-normalized normal vector field).

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a (co-orientable) null CrC^{r}-wave front. A normal vector field ξ^\hat{\xi} along FF is said to be EE-normalized if ξ^\hat{\xi} points in the future direction and satisfies |ξ^|E=2|\hat{\xi}|_{E}=\sqrt{2}, where

|𝐯|E:=(𝐯,𝐯)E(𝐯1n+1).|{\mathbf{v}}|_{E}:=\sqrt{({\mathbf{v}},{\mathbf{v}})_{E}}\qquad({\mathbf{v}}\in{\mathbb{R}}^{n+1}_{1}).

Since such a vector field ξ^\hat{\xi} is uniquely determined, we denote it by ξ^E\hat{\xi}_{E}.

Remark 1.5.

One can replace 2\sqrt{2} with any positive constant cc. However, the choice of c:=2c:=\sqrt{2} makes sense in the following reason: As we will show in Theorem 2.3, a hypersurface f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0} in Euclidean space 0n{\mathbb{R}}^{n}_{0} with unit normal vector field ν\nu induces null wave fronts F±:×Σn11n+1F_{\pm}:{\mathbb{R}}\times\Sigma^{n-1}\to{\mathbb{R}}^{n+1}_{1} whose EE-normalized normal vector fields are given by ξ^:=(1,±ν)\hat{\xi}:=(1,\pm\nu).

We can always take the EE-normalized normal vector field for a given co-orientable null wave front FF as follows: By Proposition 1.2, we can take a normal vector field ξ^\hat{\xi} along FF. Since FF is light-like, ξ^\hat{\xi} gives a light-like vector field along FF. By replacing ξ^\hat{\xi} by ξ^-\hat{\xi}, we may assume that ξ^\hat{\xi} points in the future direction, and

ξ^E:=2|ξ^|Eξ^\hat{\xi}_{E}:=\frac{\sqrt{2}}{|\hat{\xi}|_{E}}\hat{\xi}

gives the desired vector field.

Lemma 1.6.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a ((co-orientable)) null wave front. Then

(1.2) F:=(F,ξ^E):Mn1n+1×1n+1{\mathcal{L}}_{F}:=(F,\hat{\xi}_{E}):M^{n}\to{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1}

is an immersion into 1n+1×1n+1{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1}.

We call this F:Mn1n+1×1n+1{\mathcal{L}}_{F}:M^{n}\to{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1} the canonical lift of FF.

Proof.

We denote by Sn(2)S^{n}(\sqrt{2}) the sphere of radius 2\sqrt{2} centered at the origin in 1n+1{\mathbb{R}}^{n+1}_{1} with respect to the metric (,)E(\,,\,)_{E}. Since

Sn(2)𝐯π(𝐯,)P(n+1)S^{n}(\sqrt{2})\ni{\mathbf{v}}\mapsto\pi(\left\langle{{\mathbf{v}}},{*}\right\rangle)\in P^{*}({\mathbb{R}}^{n+1})

gives the double covering of P(n+1)P^{*}({\mathbb{R}}^{n+1}), it can be easily observed that F{\mathcal{L}}_{F} is an immersion into 1n+1×Sn(2){\mathbb{R}}^{n+1}_{1}\times S^{n}(\sqrt{2}) if and only if (F,πα~):Mnn+1×P(n+1)(F,\pi\circ\tilde{\alpha}):M^{n}\to{\mathbb{R}}^{n+1}\times P^{*}({\mathbb{R}}^{n+1}) is an immersion, where α~\tilde{\alpha} is the map α~:Mn(n+1){𝟎}\tilde{\alpha}:M^{n}\to({\mathbb{R}}^{n+1})^{*}\setminus\{{\mathbf{0}}\} induced by ξ^\hat{\xi} given in (1.1). ∎

The following assertion gives a characterization of null wave fronts:

Proposition 1.7.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a CrC^{r}-map. Then FF is a ((co-orientable)) null wave front if and only if there exists a vector field ξ^\hat{\xi} along FF defined on MnM^{n} such that

  1. (1)

    (ξ^,ξ^)E=2(\hat{\xi},\hat{\xi})_{E}=2,

  2. (2)

    ξ^\hat{\xi} is pointing in the future light-like direction, and

  3. (3)

    (F,ξ^)(F,\hat{\xi}) is an immersion into 1n+1×1n+1{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1} satisfying ξ^p,dFp(𝐯)=0\langle\hat{\xi}_{p},dF_{p}({\mathbf{v}})\rangle=0 for each 𝐯TpMn{\mathbf{v}}\in T_{p}M^{n} (pMn)(p\in M^{n}).

Proof.

We have already seen that any null wave front in 1n+1{\mathbb{R}}^{n+1}_{1} uniquely induces an EE-normalized normal vector field ξ^E\hat{\xi}_{E}. So it is sufficient to show the converse. Suppose that ξ^\hat{\xi} is a vector field along FF satisfying the above three conditions. We let α~:Mn(n+1){𝟎}\tilde{\alpha}:M^{n}\to({\mathbb{R}}^{n+1})^{*}\setminus\{{\mathbf{0}}\} be the map given in (1.1). By (1) and (3), (F,πα~):Mnn+1×P(n+1)(F,\pi\circ\tilde{\alpha}):M^{n}\to{\mathbb{R}}^{n+1}\times P^{*}({\mathbb{R}}^{n+1}) is an immersion by the same reason as in the proof of Lemma 1.6. ∎

Lemma 1.8.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a null CrC^{r}-immersion with a normal vector field ξ^\hat{\xi} on MnM^{n} as in Proposition 1.7. Then for each pMnp\in M^{n}, there exists a unique tangent vector ξpTpMn\xi_{p}\in T_{p}M^{n} such that dF(ξp)=ξ^pdF(\xi_{p})=\hat{\xi}_{p}. Moreover, the correspondence pξpp\mapsto\xi_{p} is CrC^{r}-differentiable.

Proof.

Since dFp(TpMn)dF_{p}(T_{p}M^{n}) is an nn-dimensional vector subspace of TF(p)1n+1T_{F(p)}{\mathbb{R}}^{n+1}_{1} at each pMnp\in M^{n}, it holds that

dFp(TpMn)={𝐯^1n+1;𝐯^,ξ^p=0}.dF_{p}(T_{p}M^{n})=\Big{\{}\hat{{\mathbf{v}}}\in{\mathbb{R}}^{n+1}_{1}\,;\,\left\langle{\hat{{\mathbf{v}}}},{\hat{\xi}_{p}}\right\rangle=0\Big{\}}.

Since ξ^pdFp(TpMn)\hat{\xi}_{p}\in dF_{p}(T_{p}M^{n}) and dFpdF_{p} is injective, the desired vector ξp\xi_{p} should be

ξp:=(dFp)1(ξ^p)(pMn).\xi_{p}:=(dF_{p})^{-1}(\hat{\xi}_{p})\qquad(p\in M^{n}).

Since (dFp)1(ξ^p)(dF_{p})^{-1}(\hat{\xi}_{p}) depends smoothly on pp, we obtain the assertion. ∎

Later, we will show that the assertion of Lemma 1.8 is extended for null CrC^{r}-wave fronts (cf. Theorem 1.14). The next assertion is a weak version of Lemma 1.8 for null wave fronts.

Proposition 1.9.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a ((co-orientable)) null CrC^{r}-wave front and ξ^\hat{\xi} a normal vector field along FF. Then, for each pMnp\in M^{n}, there exists a tangent vector 𝐯TpMn\boldsymbol{v}\in T_{p}M^{n} such that (dF)p(𝐯)=ξ^p(dF)_{p}(\boldsymbol{v})=\hat{\xi}_{p}.

Proof.

We may assume that ξ^=ξ^E\hat{\xi}=\hat{\xi}_{E}. If FF gives an immersion at pp, then the assertion follows from Lemma 1.8. So we may assume that pp is a singular point ((i.e. is a point where FF is not an immersion)) of FF. We can take a local coordinate system (u1,,un)(u_{1},\ldots,u_{n}) of MnM^{n} centered at pp such that the family of vectors

(u1)p,,(ur)p(r>0)(\partial_{u_{1}})_{p}\,\,,\ldots,\,\,(\partial_{u_{r}})_{p}\qquad(r>0)

spans the kernel of (dF)p(dF)_{p}, where uj:=/uj\partial_{u_{j}}:=\partial/\partial u_{j} (j=1,,nj=1,\ldots,n). By setting

Fui:=dF(ui)(i=1,,n),F_{u_{i}}:=dF(\partial_{u_{i}})\qquad(i=1,\ldots,n),

the vectors Fur+1(p),,Fun(p)F_{u_{r+1}}(p)\,\,,\ldots,\,\,F_{u_{n}}(p) are linearly independent. We denote by VV the subspace of 1n+1{\mathbb{R}}^{n+1}_{1} spanned by these vectors. We set

(1.3) η^i:=dξ^E(ui)(i=1,,n).\hat{\eta}_{i}:=d\hat{\xi}_{E}(\partial_{u_{i}})\qquad(i=1,\ldots,n).

If we think FF and ξ^E\hat{\xi}_{E} are column vector-valued functions, we can consider the (2n+2)×n(2n+2)\times n-matrix

(1.4) M0:=(Fu1FurFur+1Funη^1η^rη^r+1η^n)M_{0}:={\begin{pmatrix}F_{u_{1}}&\cdots&F_{u_{r}}&F_{u_{r+1}}&\cdots&F_{u_{n}}\\ \hat{\eta}_{1}&\cdots&\hat{\eta}_{r}&\hat{\eta}_{r+1}&\cdots&\hat{\eta}_{n}\end{pmatrix}}

at pp as the Jacobi matrix of the map (F,ξ^E)(F,\hat{\xi}_{E}) of MnM^{n} into 1n+1×1n+1{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1}, which can be computed as

M0=(𝟎𝟎Fur+1(p)Fun(p)η^1(p)η^r(p)η^r+1(p)η^n(p)).M_{0}={\begin{pmatrix}{\mathbf{0}}&\cdots&{\mathbf{0}}&F_{u_{r+1}}(p)&\cdots&F_{u_{n}}(p)\\ \hat{\eta}_{1}(p)&\cdots&\hat{\eta}_{r}(p)&\hat{\eta}_{r+1}(p)&\cdots&\hat{\eta}_{n}(p)\end{pmatrix}}.

Since FF is a wave front, the matrix M0M_{0} is of rank nn and so

η^1(p),,η^r(p)\hat{\eta}_{1}(p)\,\,,\ldots,\,\,\hat{\eta}_{r}(p)

are linearly independent at pp, which gives a basis of the vector space defined by

(1.5) W:={i=1raiη^i(p);a1,,ar}.W:=\left\{\sum_{i=1}^{r}a_{i}\hat{\eta}_{i}(p)\,;\,a_{1},\ldots,a_{r}\in{\mathbb{R}}\right\}.

Suppose (ξ^E)p(\hat{\xi}_{E})_{p} does not belong to VV. We set N:=(ξ^E)pN:={\mathbb{R}}(\hat{\xi}_{E})_{p}, that is, it is the 1-dimensional vector space generated by (ξ^E)p(\hat{\xi}_{E})_{p}. By the following Lemma 1.10, WN={𝟎}W\cap N=\{{\mathbf{0}}\} holds, and VV, WW and NN satisfy the assumption of Lemma A.3 in Appendix A. So we have WV={𝟎}W\cap V=\{{\mathbf{0}}\}. Since WW is of dimension rr and VV is of dimension nrn-r, the direct sum V+W+NV+W+N coincides with 1n+1{\mathbb{R}}^{n+1}_{1}. Then (ξ^E)p(\hat{\xi}_{E})_{p} vanishes because (ξ^E)p(\hat{\xi}_{E})_{p} is perpendicular to V,WV,W and NN, a contradiction. Thus, we can find a vector 𝒗TpMn\boldsymbol{v}\in T_{p}M^{n} such that (ξ^E)p=dFp(𝒗)V(\hat{\xi}_{E})_{p}=dF_{p}(\boldsymbol{v})\in V. ∎

Lemma 1.10.

The vector space WW given in (1.5) is perpendicular to dFp(TpMn)dF_{p}(T_{p}M^{n}) in 1n+1{\mathbb{R}}^{n+1}_{1}. Moreover, (ξ^E)p(\hat{\xi}_{E})_{p} does not belong to WW.

Proof.

As in the proof of Proposition 1.9, we denote by VV the subspace of 1n+1{\mathbb{R}}^{n+1}_{1} spanned by Fur+1(p),,Fun(p)F_{u_{r+1}}(p)\,\,,\ldots,\,\,F_{u_{n}}(p). Using the notation (1.3), for each i{1,,r}i\in\{1,\ldots,r\} and j{r+1,,n}j\in\{r+1,\ldots,n\}, we have

Fuj,η^i\displaystyle\left\langle{F_{u_{j}}},{\hat{\eta}_{i}}\right\rangle =Fuj,ξ^EuiFujui,ξ^E=Fujui,ξ^E\displaystyle={\left\langle{F_{u_{j}}},{\hat{\xi}_{E}}\right\rangle}_{u_{i}}-\left\langle{F_{u_{j}u_{i}}},{\hat{\xi}_{E}}\right\rangle=-\left\langle{F_{u_{j}u_{i}}},{\hat{\xi}_{E}}\right\rangle
=Fui,ξ^Euj+Fui,dξ^E(uj)=Fui,η^j=0\displaystyle=-\left\langle{F_{u_{i}}},{\hat{\xi}_{E}}\right\rangle_{u_{j}}+\left\langle{F_{u_{i}}},{d\hat{\xi}_{E}(\partial_{u_{j}})}\right\rangle=\left\langle{F_{u_{i}}},{\hat{\eta}_{j}}\right\rangle=0

at pp, where we used the fact Fui(p)=𝟎F_{u_{i}}(p)={\mathbf{0}}. By this computation, we have that

(1.6) v,w=0(vV,wW),\left\langle{v},{w}\right\rangle=0\qquad(v\in V,\,\,w\in W),

proving the first assertion.

We prove the second assertion. If (ξ^E)pW(\hat{\xi}_{E})_{p}\in W, then we can write ξ^E=i=1rai(ξ^E)ui\hat{\xi}_{E}=\sum_{i=1}^{r}a_{i}(\hat{\xi}_{E})_{u_{i}} at pp. Then it holds that

2=((ξ^E)p,(ξ^E)p)E=i=1rai(((ξ^E)ui)p,(ξ^E)p)E=0,2=\Big{(}(\hat{\xi}_{E})_{p},(\hat{\xi}_{E})_{p}\Big{)}_{E}=\sum_{i=1}^{r}a_{i}\Big{(}\big{(}(\hat{\xi}_{E})_{u_{i}}\big{)}_{p},\big{(}\hat{\xi}_{E}\big{)}_{p}\Big{)}_{E}=0,

a contradiction. ∎

We next prepare the following:

Proposition 1.11.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a null CrC^{r}-wave front and ξ^E\hat{\xi}_{E} its associated EE-normalized normal vector field. Suppose that pMnp\in M^{n} is a singular point of FF. Then

  1. (1)

    for each δ\delta\in{\mathbb{R}},

    (1.7) Fδ:=F+δξ^EF_{\delta}:=F+\delta\,\hat{\xi}_{E}

    is a null wave front defined on MnM^{n}, and

  2. (2)

    there exists a positive number ε0\varepsilon_{0} such that FδF_{\delta} (0<|δ|<ε0)(0<|\delta|<\varepsilon_{0}) is an immersion at pp.

Remark 1.12.

Later, we will show that the image of FδF_{\delta} coincides with that of FF under a suitable assumption of FF (see Corollary 2.8).

Proof.

Since ξ^E\hat{\xi}_{E} is a light-like vector field, it is a normal vector field of FδF_{\delta}. We set

Mδ:=((Fδ)u1(Fδ)ur(Fδ)ur+1(Fδ)un(ξ^E)u1(ξ^E)ur(ξ^E)ur+1(ξ^E)un)M_{\delta}:={\begin{pmatrix}(F_{\delta})_{u_{1}}&\cdots&(F_{\delta})_{u_{r}}&(F_{\delta})_{u_{r+1}}&\cdots&(F_{\delta})_{u_{n}}\\ (\hat{\xi}_{E})_{u_{1}}&\cdots&(\hat{\xi}_{E})_{u_{r}}&(\hat{\xi}_{E})_{u_{r+1}}&\cdots&(\hat{\xi}_{E})_{u_{n}}\end{pmatrix}}

for δ\delta\in{\mathbb{R}}. Since MδM_{\delta} (δ0\delta\neq 0) is obtained by adding the second row to the first row of M0M_{0}, the pair (Fδ,ξ^E)(F_{\delta},\hat{\xi}_{E}) gives an immersion of MnM^{n} into 1n+1×1n+1{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1} if and only if (F,ξ^E)(F,\hat{\xi}_{E}) is an immersion. So (1) is proved.

We next prove (2). By Proposition 1.9, there exists a vector 𝒗TpMn\boldsymbol{v}\in T_{p}M^{n} such that dFp(𝒗)=(ξ^E)pdF_{p}(\boldsymbol{v})=(\hat{\xi}_{E})_{p}. We then take a local coordinate system (u1,,un)(u_{1},\ldots,u_{n}) of MnM^{n} centered at pp such that

(u1)p,,(ur)p(r>0)(\partial_{u_{1}})_{p}\,\,,\ldots,\,\,(\partial_{u_{r}})_{p}\qquad(r>0)

belong to the kernel of (dF)p(dF)_{p} and (un)p=𝒗(\partial_{u_{n}})_{p}=\boldsymbol{v}. We let VV be the subspace of 1n+1{\mathbb{R}}^{n+1}_{1}, which is spanned by Fur+1(p),,Fun1(p)F_{u_{r+1}}(p)\,\,,\ldots,\,\,F_{u_{n-1}}(p). We consider the rr-dimensional vector space WW given by (1.5). By Lemma 1.10, it can be easily checked that the three vector subspaces V,WV,\,\,W and N:=ξ^EN:={\mathbb{R}}\hat{\xi}_{E} satisfy the assumption of Lemma A.3 in Appendix A. So we have

VW=VN=WN={𝟎}.V\cap W=V\cap N=W\cap N=\{{\mathbf{0}}\}.

In particular, the vectors

(ξ^E)u1,,(ξ^E)ur,Fur+1,,Fun(=ξ^E)(\hat{\xi}_{E})_{u_{1}},\,\,\ldots,(\hat{\xi}_{E})_{u_{r}},\,\,F_{u_{r+1}}\,\,,\ldots,\,\,F_{u_{n}}(=\hat{\xi}_{E})

are linearly independent at pMnp\in M^{n}. We fix a non-zero real number δ\delta. Then we have

m(δ):=\displaystyle m(\delta):= rank(dFδ(u1),,dFδ(ur),dFδ(ur+1),,dFδ(un))\displaystyle\operatorname{rank}\Big{(}dF_{\delta}(\partial_{u_{1}}),\dots,dF_{\delta}(\partial_{u_{r}}),dF_{\delta}(\partial_{u_{r+1}}),\dots,dF_{\delta}(\partial_{u_{n}})\Big{)}
=rank(δdξ^E(u1),,δdξ^E(ur),dF(ur+1)+δdξ^E(ur+1),\displaystyle=\operatorname{rank}\Big{(}\delta d\hat{\xi}_{E}(\partial_{u_{1}}),\dots,\delta d\hat{\xi}_{E}(\partial_{u_{r}}),dF(\partial_{u_{r+1}})+\delta d\hat{\xi}_{E}(\partial_{u_{r+1}}),
,dF(un)+δdξ^E(un))\displaystyle\phantom{aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa}\dots,dF(\partial_{u_{n}})+\delta d\hat{\xi}_{E}(\partial_{u_{n}})\Big{)}
=rank(dξ^E(u1),,dξ^E(ur),dF(ur+1)+δdξ^E(ur+1),\displaystyle=\operatorname{rank}\Big{(}d\hat{\xi}_{E}(\partial_{u_{1}}),\dots,d\hat{\xi}_{E}(\partial_{u_{r}}),dF(\partial_{u_{r+1}})+\delta d\hat{\xi}_{E}(\partial_{u_{r+1}}),
,dF(un)+δdξ^E(un)).\displaystyle\phantom{aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa}\dots,dF(\partial_{u_{n}})+\delta d\hat{\xi}_{E}(\partial_{u_{n}})\Big{)}.

If |δ||\delta| is sufficiently small, the right-hand side of the last equality is equal to nn at pp. So we can conclude that FδF_{\delta} is an immersion on a sufficiently small neighborhood of pp for sufficiently small |δ|(>0)|\delta|(>0). ∎

Definition 1.13 (EE-normalized null vector field).

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a null CrC^{r}-wave front and ξ^E\hat{\xi}_{E} its associated EE-normalized normal vector field. If there exists a CrC^{r}-differentiable vector field ξE\xi_{E} defined on MnM^{n} such that dF(ξE)=ξ^EdF(\xi_{E})=\hat{\xi}_{E}, then ξE\xi_{E} is called the EE-normalized null vector field with respect to FF (in fact, ξE\xi_{E} is uniquely determined as follows).

The following is the deepest result in this section, which asserts the existence of EE-normalized null vector field:

Theorem 1.14.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a null CrC^{r}-wave front and ξ^E\hat{\xi}_{E} its associated EE-normalized normal vector field. Then there exists a unique CrC^{r}-differentiable vector field ξE\xi_{E} defined on MnM^{n} such that

  1. (1)

    dF(ξE)=ξ^EdF(\xi_{E})=\hat{\xi}_{E} ((that is, ξE\xi_{E} is the EE-normalized null vector field)), and

  2. (2)

    the image of each integral curve of ξE\xi_{E} is a part of a light-like line in 1n+1{\mathbb{R}}^{n+1}_{1}.

Moreover, dF(TpMn)dF(T_{p}M^{n}) is a light-like vector space in 1n+1{\mathbb{R}}^{n+1}_{1}.

Proof.

Roughly speaking, the key of the proof is that the image of the null wave front FF is foliated by light-like lines as mentioned in the introduction, and we will construct a vector field ξE\xi_{E} defined on MnM^{n} so that each integral curve of ξE\xi_{E} corresponds to the leaf of the foliation as follows.

We fix a point pMnp\in M^{n} arbitrarily. By Proposition 1.11, for a sufficiently small positive number δ\delta, there exists a neighborhood UU of pp such that FδF_{\delta} is an immersion on UU. Hence, by Lemma 1.8, there exists a unique CrC^{r}-differentiable null vector field ξE\xi_{E} satisfying dFδ(ξE)=ξ^EdF_{\delta}(\xi_{E})=\hat{\xi}_{E} on UU. We let γq(t)\gamma_{q}(t) be the integral curve of ξE\xi_{E} such that γq(0)=q\gamma_{q}(0)=q for qUq\in U. Since ξ^E\hat{\xi}_{E} is a null vector field along FδF_{\delta}, [2, Fact 2.6] implies that FδγqF_{\delta}\circ\gamma_{q} parametrizes a segment of a light-like line. Thus Fδγq(t)F_{\delta}\circ\gamma_{q}(t) is a geodesic in 1n+1{\mathbb{R}}^{n+1}_{1}, and so we have DξEdFδ(ξE)=𝟎,D_{\xi_{E}}dF_{\delta}(\xi_{E})={\mathbf{0}}, where DD is the Levi-Civita connection of 1n+1{\mathbb{R}}^{n+1}_{1}. So,

dF(ξE)=dFδ(ξE)δ(dξ^E)(ξE)=ξ^EδDξEdFδ(ξE)=ξ^EdF(\xi_{E})=dF_{\delta}(\xi_{E})-\delta(d\hat{\xi}_{E})(\xi_{E})=\hat{\xi}_{E}-\delta D_{\xi_{E}}dF_{\delta}(\xi_{E})=\hat{\xi}_{E}

holds at pp. Since the light-like vector (ξ^E)q(\hat{\xi}_{E})_{q} (qUq\in U) lies in dFq(TqMn)dF_{q}(T_{q}M^{n}), the vector space (dF)q(TqMn)(dF)_{q}(T_{q}M^{n}) is light-like. Since pp is arbitrary, the uniqueness of ξE\xi_{E} implies it can be defined on MnM^{n}, proving the assertion. ∎

2. A fundamental theorem for LL-complete null wave fronts

We first define the following “LL-completeness” of null wave fronts in 1n+1{\mathbb{R}}^{n+1}_{1}, which is analogous to the case of null immersions given in [2, Definition 2.7].

Definition 2.1.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a (co-orientable) null CrC^{r}-wave front. The map FF is called LL-complete if for each pMnp\in M^{n}, there exists an integral curve γ:Mn\gamma:{\mathbb{R}}\to M^{n} of the EE-normalized null vector field ξE\xi_{E} passing through pp such that Fγ()F\circ\gamma({\mathbb{R}}) coincides with an entire light-like line in 1n+1{\mathbb{R}}^{n+1}_{1}.

The following assertion holds:

Proposition 2.2.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a ((co-orientable)) null CrC^{r}-wave front. Then FF is LL-complete if and only if its EE-normalized null vector field ξE\xi_{E} is complete on MnM^{n} ((that is, each integral curve of ξE\xi_{E} is defined on ){\mathbb{R}}).

Proof.

Let γ(t)\gamma(t) be any integral curve of ξE\xi_{E}. Since |dF(ξE)|E=2|dF(\xi_{E})|_{E}=\sqrt{2}, γ(t)\gamma(t) is parametrized by an affine parametrization of a light-like line. So γ\gamma is defined on {\mathbb{R}} if and only if the image of FγF\circ\gamma coincides with the entirety of a line. ∎

We consider the height function

(2.1) τ^:1n+1𝐯𝐯,𝐞0\hat{\tau}:{\mathbb{R}}^{n+1}_{1}\ni{\mathbf{v}}\mapsto-\left\langle{{\mathbf{v}}},{{\mathbf{e}}_{0}}\right\rangle\in{\mathbb{R}}

with respect to the time axis, where 𝐞0:=(1,0,,0){\mathbf{e}}_{0}:=(1,0,\ldots,0). The level set τ^1(0)\hat{\tau}^{-1}(0) is a space-like hyperplane which is isometric to the Euclidean nn-space 0n{\mathbb{R}}^{n}_{0}. So, we frequently use the identification

(2.2) 0nx:=(x1,,xn)x~:=(0,x1,,xn)τ^1(0).{\mathbb{R}}^{n}_{0}\ni x:=(x_{1},\ldots,x_{n})\,\,\longleftrightarrow\,\,\tilde{x}:=(0,x_{1},\ldots,x_{n})\in\hat{\tau}^{-1}(0).

Let Σn1\Sigma^{n-1} be an (n1)(n-1)-manifold. A CrC^{r}-map f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0} is called a (co-orientable) wave front if there exists a unit normal vector field ν\nu along ff defined on Σn1\Sigma^{n-1} such that the map Σn1p(f(p),ν(p))0n×Sn1\Sigma^{n-1}\ni p\mapsto(f(p),\nu(p))\in{\mathbb{R}}^{n}_{0}\times S^{n-1} is an immersion, where Sn1S^{n-1} is the unit sphere centered at the origin in 0n{\mathbb{R}}^{n}_{0}. We now show that the following representation formula of LL-complete null wave fronts in 1n+1{\mathbb{R}}^{n+1}_{1} from a given wave front in 0n{\mathbb{R}}^{n}_{0} as follows:

Theorem 2.3.

Let f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0} be a ((co-orientable)) CrC^{r}-wave front with unit normal vector field ν\nu. Then for each choice of σ{+,}\sigma\in\{+,-\}, the map σf:×Σn11n+1{\mathcal{F}}^{f}_{\sigma}:{\mathbb{R}}\times\Sigma^{n-1}\to{\mathbb{R}}^{n+1}_{1} defined by

(2.3) σf(t,x):=f~(x)+tξ^σ(x),ξ^σ(x):=(1,σν(x))(t,xΣn1){\mathcal{F}}^{f}_{\sigma}(t,x):=\tilde{f}(x)+t\hat{\xi}_{\sigma}(x),\qquad\hat{\xi}_{\sigma}(x):=(1,\sigma\nu(x))\qquad(t\in{\mathbb{R}},\,\,x\in\Sigma^{n-1})

gives an LL-complete null wave front in 1n+1{\mathbb{R}}^{n+1}_{1}, where f~(x):=(0,f(x)).\tilde{f}(x):=(0,f(x)). Moreover, the regular set of σf{\mathcal{F}}^{f}_{\sigma} is dense in ×Σn1{\mathbb{R}}\times\Sigma^{n-1}.

When ff is an immersion, this formula is known (see Kossowski [7]). So this theorem can be considered as its generalization for wave fronts. The slice of the image of ±f{\mathcal{F}}^{f}_{\pm} by a hyperplane {t=c}\{t=c\} (cc\in{\mathbb{R}}) is congruent to the image of a parallel hypersurface fcf^{c} of ff.

Remark 2.4.

Since

f(t,x)=(100010001)+f(t,x),{\mathcal{F}}^{f}_{-}(t,x)={\begin{pmatrix}-1&0&\cdots&0\\ 0&1&\cdots&0\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\cdots&1\end{pmatrix}}\,{\mathcal{F}}^{f}_{+}(-t,x),

the image of f{\mathcal{F}}^{f}_{-} is congruent to that of +f{\mathcal{F}}^{f}_{+} in 1n+1{\mathbb{R}}^{n+1}_{1}. So we call +f{\mathcal{F}}^{f}_{+} the normal form of the null wave front associated with ff.

Proof of Theorem 2.3.

Without loss of generality, we may assume that σ=+\sigma=+, and we set

F:=+f,ξ^:=ξ^+(=(1,ν)).F:={\mathcal{F}}^{f}_{+},\qquad\hat{\xi}:=\hat{\xi}_{+}\Big{(}=\big{(}1,\nu\big{)}\Big{)}.

Since ξ^(x)\hat{\xi}(x) (xΣn1x\in\Sigma^{n-1}) is orthogonal to dF(Tp(×Σn1))dF(T_{p}({\mathbb{R}}\times\Sigma^{n-1})) with respect to the Lorentzian inner product ,\left\langle{\,},{\,}\right\rangle, the vector field t:=/t\partial_{t}:=\partial/\partial t is a null vector field of FF defined on ×Σn1{\mathbb{R}}\times\Sigma^{n-1}. We write

F=(F0,,Fn),ξ^=(ξ^0,,ξ^n),\displaystyle F=(F^{0},\ldots,F^{n}),\qquad\hat{\xi}=(\hat{\xi}^{0},\ldots,\hat{\xi}^{n}),
f=(f1,,fn), and ν=(ν1,,νn).\displaystyle f=(f^{1},\ldots,f^{n}),\,\,\text{ and }\,\,\,\nu=(\nu^{1},\ldots,\nu^{n}).

In the following discussions, we think that FF and ν\nu take values in column vectors. Then we have

F0=t,ξ^0=1,Fi=fi+tνi,ξ^i=νi(i=1,,n).F^{0}=t,\quad\hat{\xi}^{0}=1,\quad F^{i}=f^{i}+t\nu^{i},\quad\hat{\xi}^{i}=\nu^{i}\qquad(i=1,\ldots,n).

We fix xΣn1x\in\Sigma^{n-1} arbitrarily, and take a local coordinate system (u1,,un1)(u_{1},\ldots,u_{n-1}) of Σn1\Sigma^{n-1} centered at xx. Then (t,u1,,un1)(t,u_{1},\ldots,u_{n-1}) gives a local coordinate system of ×Σn1{\mathbb{R}}\times\Sigma^{n-1}. To show that F:=(F,ξ^)\mathcal{L}_{F}:=(F,\hat{\xi}) is an immersion, it is sufficient to show that (2n+2)×n(2n+2)\times n matrix

Mt:=(FtFu1Fun1ξ^tξ^u1ξ^un1)M_{t}:={\begin{pmatrix}F_{t}&F_{u_{1}}&\cdots&F_{u_{n-1}}\\ \hat{\xi}_{t}&\hat{\xi}_{u_{1}}&\ldots&\hat{\xi}_{u_{n-1}}\end{pmatrix}}

is of rank nn. Since

(2.4) (Ft(t,x),Fu1(t,x),,Fun1(t,x))\displaystyle(F_{t}(t,x),\,\,F_{u_{1}}(t,x),\,\,\ldots,\,\,F_{u_{n-1}}(t,x))
=(100ν(x)fu1(x)+tνu1(x)fun1(x)+tνun1(x))\displaystyle\phantom{aaaaaaaaaaaa}={\begin{pmatrix}1&0&\ldots&0\\ \nu(x)&f_{u_{1}}(x)+t\nu_{u_{1}}(x)&\ldots&f_{u_{n-1}}(x)+t\nu_{u_{n-1}}(x)\end{pmatrix}}

holds, we have

(2.5) Mt=(100νfu1+tνu1fun1+tνun1000𝟎νu1νun1).M_{t}={\begin{pmatrix}1&0&\ldots&0\\ \nu&f_{u_{1}}+t\nu_{u_{1}}&\ldots&f_{u_{n-1}}+t\nu_{u_{n-1}}\\ 0&0&\ldots&0\\ {\mathbf{0}}&\nu_{u_{1}}&\ldots&\nu_{u_{n-1}}\end{pmatrix}}.

Since ff is a wave front, we have

n1=rank(fu1fun1νu1νun1),n-1=\operatorname{rank}{\begin{pmatrix}f_{u_{1}}&\ldots&f_{u_{n-1}}\\ \nu_{u_{1}}&\ldots&\nu_{u_{n-1}}\end{pmatrix}},

which implies that MtM_{t} is of rank nn, that is, FF is a null wave front by Proposition 1.7. Since we have already shown that t\partial_{t} points in the null direction, FF is a null wave front. By definition, it is obvious that FF is LL-complete.

We next show that FF is an immersion on an open dense subset of ×Σn1{\mathbb{R}}\times\Sigma^{n-1}. By (2.4), the matrix (Ft,Fu1,,Fun1)(F_{t},F_{u_{1}},\ldots,F_{u_{n-1}}) is of rank nn at (t,x)×Σn1(t,x)\in{\mathbb{R}}\times\Sigma^{n-1} if

n1=rank(fu1(x)+tνu1(x),,fun1(x)+tνun1(x)).n-1=\operatorname{rank}\Big{(}f_{u_{1}}(x)+t\nu_{u_{1}}(x),\,\,\ldots,\,\,f_{u_{n-1}}(x)+t\nu_{u_{n-1}}(x)\Big{)}.

To prove that this holds for almost all (t,x)(t,x), one can use the fact that the parallel hypersurface ft:=f+tνf_{t}:=f+t\nu (for fixed tt) has a singular point at xx if and only if tt coincides with one of the inverse of principal curvatures of ff at xx: If a given point (t,x)×Σn1(t,x)\in{\mathbb{R}}\times\Sigma^{n-1} is a singular point of FF, then xx is a regular point of ftnf_{t_{n}} for tn:=t+1/nt_{n}:=t+1/n. In particular, (t,x)(t,x) is an accumulation point of the regular points {(tn,x)}n=1\{(t_{n},x)\}_{n=1}^{\infty} of FF, and we can conclude that the regular set of FF is dense in ×Σn1{\mathbb{R}}\times\Sigma^{n-1}. ∎

Lemma 2.5.

In the setting of Theorem 2.3, if

lf:Σn1x(f(x),ν(x))0n×Sn1l_{f}:\Sigma^{n-1}\ni x\mapsto(f(x),\nu(x))\in{\mathbb{R}}^{n}_{0}\times S^{n-1}

is an embedding, then by setting ξ^+(x):=(1,ν(x)),\hat{\xi}_{+}(x):=(1,\nu(x)), the map

F:×Σn1(t,x)(+f(t,x),ξ^+(x))1n+1×1n+1{\mathcal{L}}_{F}:{\mathbb{R}}\times\Sigma^{n-1}\ni(t,x)\mapsto\big{(}\mathcal{F}^{f}_{+}(t,x),\hat{\xi}_{+}(x)\big{)}\in{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1}

is also an embedding, where F:=+fF:=\mathcal{F}^{f}_{+}.

Proof.

We remark that F{\mathcal{L}}_{F} (F:=+fF:={\mathcal{F}}^{f}_{+}) is a map into 1n+1×(1n+1{𝟎}){\mathbb{R}}^{n+1}_{1}\times({\mathbb{R}}^{n+1}_{1}\setminus\{{\mathbf{0}}\}). Since FF is a null wave front, F{\mathcal{L}}_{F} is an immersion (cf. Proposition 1.7). So it is sufficient to prove that F{\mathcal{L}}_{F} is a homeomorphism from MnM^{n} to F(Mn){\mathcal{L}}_{F}(M^{n}). For the sake of simplicity, we set ξ^:=ξ^+\hat{\xi}:=\hat{\xi}_{+}.

We let (P,ξ^)(P,\hat{\xi}) be a point in F(×Σn1){\mathcal{L}}_{F}({\mathbb{R}}\times\Sigma^{n-1}). Then there exists a unique point Φ(P,ξ^)\Phi(P,\hat{\xi}) at which the straight line {P+tξ^;t}\{P+t\hat{\xi}\,;\,t\in{\mathbb{R}}\} meets the hyperplane τ^1(0)\hat{\tau}^{-1}(0) in 1n+1{\mathbb{R}}^{n+1}_{1}. We let

(2.6) π0:1n+1τ^1(0)\pi_{0}:{\mathbb{R}}^{n+1}_{1}\to\hat{\tau}^{-1}(0)

be the canonical orthogonal projection. Since Φ(P,ξ^)\Phi(P,\hat{\xi}) depends on PP and ξ^\hat{\xi} continuously, the map (P,ξ^)(Φ(P,ξ^),π0(ξ^))(P,\hat{\xi})\mapsto(\Phi(P,\hat{\xi}),\pi_{0}(\hat{\xi})) is a continuous map whose image coincides with lf(Σn1)l_{f}(\Sigma^{n-1}). So

F(×Σn1)(P,ξ^)(τ^(P),lf1Φ(P,ξ^))×Σn1{\mathcal{L}}_{F}({\mathbb{R}}\times\Sigma^{n-1})\ni(P,\hat{\xi})\mapsto(\hat{\tau}(P),l_{f}^{-1}\circ\Phi(P,\hat{\xi}))\in{\mathbb{R}}\times\Sigma^{n-1}

is well-defined, and gives the inverse map of F{\mathcal{L}}_{F}. So F{\mathcal{L}}_{F} is an embedding. ∎

Conversely, we can prove the following:

Theorem 2.6.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a ((co-orientable)) LL-complete null CrC^{r}-wave front, then there exists a co-orientable CrC^{r}-wave front f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0} and a diffeomorphism Φ:Mn×Σn1\Phi:M^{n}\to{\mathbb{R}}\times\Sigma^{n-1} such that FΦ1F\circ\Phi^{-1} coincides with the null wave front F+fF^{f}_{+} as in Theorem 2.3.

Proof.

Let ξE\xi_{E} be the EE-normalized null vector field of FF. By Theorem 1.14, the image of each integral curve γ(t)\gamma(t) of ξE\xi_{E} is a light-like geodesic. So the identity

(2.7) DξEξ^E=Dγ(t)dF(γ(t))=𝟎D_{\xi_{E}}\hat{\xi}_{E}=D_{\gamma^{\prime}(t)}dF(\gamma^{\prime}(t))={\mathbf{0}}

holds on MnM^{n}, where ξ^E:=dF(ξE)\hat{\xi}_{E}:=dF(\xi_{E}) and DD is the Levi-Civita connection of 1n+1{\mathbb{R}}^{n+1}_{1}.

Since |ξ^E|E=2|\hat{\xi}_{E}|_{E}=\sqrt{2} and ξ^E\hat{\xi}_{E} points in the light-like direction, we have dτ^(dF(ξE))𝟎d\hat{\tau}(dF(\xi_{E}))\neq{\mathbf{0}}, in particular, by applying the implicit function theorem, the zero-level set

Σn1:=(τ^F)1(0)(Mn)\Sigma^{n-1}:=(\hat{\tau}\circ F)^{-1}(0)\,(\subset M^{n})

is an embedded hypersurface of MnM^{n}. Let {φt}t\{\varphi_{t}\}_{t\in{\mathbb{R}}} be the one parameter group of transformations on MnM^{n} associated with ξE\xi_{E}. As in [2, Proposition A.3], the map defined by

(2.8) Ψ:×Σn1(t,p)φt(p)Mn\Psi:{\mathbb{R}}\times\Sigma^{n-1}\ni(t,p)\mapsto\varphi_{t}(p)\in M^{n}

is an immersion. Since (cf. (2.7)) γ^(t):=F(φt(p))(t)\hat{\gamma}(t):=F(\varphi_{t}(p))\,\,(t\in{\mathbb{R}}) is a light-like geodesic in 1n+1{\mathbb{R}}^{n+1}_{1}, we obtain the expression

(2.9) FΨ(t,p)=Fφt(p)=F(p)+tξ^E(p)(t,pΣn1).F\circ\Psi(t,p)=F\circ\varphi_{t}(p)=F(p)+t\hat{\xi}_{E}(p)\qquad(t\in{\mathbb{R}},\,p\in\Sigma^{n-1}).

In particular, tFΨ(t,p)1n+1{\mathbb{R}}\ni t\mapsto F\circ\Psi(t,p)\in{\mathbb{R}}^{n+1}_{1} is an injection, and so,

tφt(p)Mn{\mathbb{R}}\ni t\mapsto\varphi_{t}(p)\in M^{n}

is also an injection. We suppose that Ψ\Psi is not injective and Ψ(t,p)=Ψ(s,q)\Psi(t,p)=\Psi(s,q) holds. Since φst(q)=p\varphi_{s-t}(q)=p, the point qq lies on the integral curve passing through pp. Since γ^(=Fγ)\hat{\gamma}(=F\circ\gamma) is a straight line, γ\gamma meets Σn1\Sigma^{n-1} exactly once. So we can conclude p=qp=q and s=ts=t. Thus Ψ\Psi is an injection. On the other hand, since FF is LL-complete, any integral curve of ξE\xi_{E} must meet Σn1\Sigma^{n-1}. So Ψ\Psi is bijective and then it becomes a diffeomorphism. If we set ξ^E(p)=(a(p),ν(p))\hat{\xi}_{E}(p)=(a(p),\nu(p)) (pMnp\in M^{n}), then we have

(2.10) 2=(ξ^E(p),ξ^E(p))E=a(p)2+|ν(p)|2.2=(\hat{\xi}_{E}(p),\hat{\xi}_{E}(p))_{E}=a(p)^{2}+|\nu(p)|^{2}.

Since

(2.11) 0=ξ^E(p),ξ^E(p)=a(p)2+|ν(p)|2,0=\langle\hat{\xi}_{E}(p),\hat{\xi}_{E}(p)\rangle=-a(p)^{2}+|\nu(p)|^{2},

we have a(p)2=|ν(p)|2=1,a(p)^{2}=|\nu(p)|^{2}=1, which implies that ν\nu is a unit normal vector field of f~:=F|Σn1\tilde{f}:=F|_{\Sigma^{n-1}} in the space-like hyperplane τ^1(0)\hat{\tau}^{-1}(0). We fix 𝐯TpΣn1{\mathbf{v}}\in T_{p}\Sigma^{n-1} arbitrarily, then we can write (df~)p(𝐯)=(0,𝐚)(d\tilde{f})_{p}({\mathbf{v}})=(0,{\mathbf{a}}) where 𝐚0n{\mathbf{a}}\in{\mathbb{R}}^{n}_{0}. Since ξ^E=(1,ν)\hat{\xi}_{E}=(1,\nu), we have

(2.12) 0=(df~)p,ξ^=𝐚ν,0=\left\langle{(d\tilde{f})_{p}},{\hat{\xi}}\right\rangle={\mathbf{a}}\cdot\nu,

where the dot denotes the canonical Euclidean inner product on 0n{\mathbb{R}}^{n}_{0}, where τ^1(0)={0}×0n\hat{\tau}^{-1}(0)=\{0\}\times{\mathbb{R}}^{n}_{0}. So ν\nu is a unit normal vector field of ff.

Finally, we show that f~\tilde{f} is a wave front defined on Σn1\Sigma^{n-1}: Since FF is a null wave front, the pair F:=(F,ξ^E){\mathcal{L}}_{F}:=(F,\hat{\xi}_{E}) is an immersion on MnM^{n} into 2n+2{\mathbb{R}}^{2n+2}. Then its restriction

(f~,ξ^E|Σn1):Σn11n+1×1n+1(\tilde{f},\hat{\xi}_{E}|_{\Sigma^{n-1}}):\Sigma^{n-1}\to{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1}

is also an immersion. Moreover, since ξE|Σn1=(1,ν),\xi_{E}|_{\Sigma^{n-1}}=(1,\nu), the map

(f,ν):Σn10n×0n(f,\nu):\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0}\times{\mathbb{R}}^{n}_{0}

is an immersion (where f~=(0,f)\tilde{f}=(0,f)), and so f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0} is a co-orientable wave front. Summarizing the above discussions, (2.9) can be written as

FΨ(t,x)\displaystyle F\circ\Psi(t,x) =Fφt(x)=F(x)+tξ^E(x)\displaystyle=F\circ\varphi_{t}(x)=F(x)+t\hat{\xi}_{E}(x)
=f~(x)+t(1,ν(x))=F+f(t,x)((t,x)×Σn1).\displaystyle=\tilde{f}(x)+t(1,\nu(x))=F^{f}_{+}(t,x)\qquad((t,x)\in{\mathbb{R}}\times\Sigma^{n-1}).

By setting Φ:=Ψ1\Phi:=\Psi^{-1}, we obtain the assertion. ∎

Corollary 2.7.

The regular set of an LL-complete null CrC^{r}-wave front F:Mn1m+1F:M^{n}\to{\mathbb{R}}^{m+1}_{1} is dense in MnM^{n}.

Proof.

We have just shown that such a null wave front can be reparametrized as a normal form. So, the assertion follows from the last statement of Theorem 2.3. ∎

Corollary 2.8.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be an LL-complete null CrC^{r}-wave front. Then, for each δ\delta\in{\mathbb{R}} the parallel hypersurface FδF_{\delta} given in (1.7) is also an LL-complete null wave front and has the same image as FF.

Proof.

By Theorem 2.6, there exists a co-orientable CrC^{r}-wave front f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0} and a diffeomorphism Φ:Mn×Σn1\Phi:M^{n}\to{\mathbb{R}}\times\Sigma^{n-1} such that FΦ1=F+fF\circ\Phi^{-1}=F^{f}_{+} holds on MnM^{n}. We let ξ^E\hat{\xi}_{E} be the EE-normalized normal vector field along FF. Then ξ^EΦ1\hat{\xi}_{E}\circ\Phi^{-1} is the EE-normalized normal vector field of F+fF^{f}_{+}. So we have that

FδΦ1(t,x)\displaystyle F_{\delta}\circ\Phi^{-1}(t,x) =FΦ1(t,x)+δξ^EΦ1(t,x)\displaystyle=F\circ\Phi^{-1}(t,x)+\delta\hat{\xi}_{E}\circ\Phi^{-1}(t,x)
=(f~(x)+tξ^EΦ1(t,x))+δξ^EΦ1(t,x)\displaystyle=\Big{(}\tilde{f}(x)+t\hat{\xi}_{E}\circ\Phi^{-1}(t,x)\Big{)}+\delta\hat{\xi}_{E}\circ\Phi^{-1}(t,x)
=f~(x)+(t+δ)ξ^EΦ1(t,x)=FΦ1(t+δ,x),\displaystyle=\tilde{f}(x)+(t+\delta)\hat{\xi}_{E}\circ\Phi^{-1}(t,x)=F\circ\Phi^{-1}(t+\delta,x),

which proves the assertion. ∎

3. A structure theorem of null wave fronts

In this section, we give a structure theorem of null wave fronts without assuming LL-completeness. Roughly speaking, a null wave front is foliated by line segments, so by extending each line segment to a whole line and connecting them appropriately, we will obtain an LL-complete null wave front.

Refer to caption
Refer to caption
Figure 2. The tubular neighborhood 𝒰{\mathcal{U}} of the logarithmic spiral γ\gamma and its image F(𝒰)F({\mathcal{U}}) in the light-cone Λ2\Lambda^{2}.

Here is one example to illustrate the structure theorem: We consider the 1/21/2-tubular neighborhood 𝒰{\mathcal{U}} of the logarithmic spiral (see Figure 2, left)

γ(t)=et/15(cost,sint)(π/2<t<4π+π/2)\gamma(t)=e^{t/15}(\cos t,\sin t)\qquad(\pi/2<t<4\pi+\pi/2)

in the xyxy-plane, and consider the image F(𝒰)F({\mathcal{U}}) with respect to the graph F(x,y):=(x,y,x2+y2)F(x,y):=(x,y,\sqrt{x^{2}+y^{2}}). Then F(𝒰)F({\mathcal{U}}) is an open subset of the light-cone Λ2\Lambda^{2} (see Figure 2, right). Since F(𝒰)F({\mathcal{U}}) is a ruled surface, we can extend each of the ruling light-like lines to both sides, and obtain an LL-complete null wave front Fˇ\check{F} as an LL-completion of F(𝒰)F({\mathcal{U}}). However, this surface Fˇ\check{F} is different from the light-cone Λ2\Lambda^{2}. In fact, each light-like line as a generator of the ruled surface Fˇ\check{F} meets F(𝒰)F({\mathcal{U}}) several times and Fˇ\check{F} can be regarded as a kind of double covering the light-cone Λ2\Lambda^{2}. To produce the actual Λ2\Lambda^{2} from the extension of F(𝒰)F({\mathcal{U}}), we need to consider the quotient space with an appropriate equivalence relation in the image of Fˇ\check{F}. In this example, the resulting LL-completion is the light-cone Λ2\Lambda^{2}, which is, of course, a Hausdorff space. However, in the general situation, in order for the domain of definition of the resulting LL-complete null wave front to be a Hausdorff space, it is necessary to assume appropriate conditions on the original null wave front.

From now on, we fix a (co-orientable) null CrC^{r}-wave front F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} which may not be LL-complete. Let

(3.1) τ:Mnpτ^F(p),\tau:M^{n}\ni p\mapsto\hat{\tau}\circ F(p)\in{\mathbb{R}},

be the restriction of the height function τ^\hat{\tau} to the hypersurface FF. he following lemma will play an important role:

Lemma 3.1.

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a ((co-orientable)) null CrC^{r}-wave front, and let pp be a point in MnM^{n}. Then, for each open neighborhood U~\tilde{U} of pp, there exist

  • a neighborhood U(U~)U(\subset\tilde{U}) of pp,

  • a positive number εU(0,]\varepsilon_{U}\in(0,\infty] and a real number tUt_{U},

  • a CrC^{r}-differentiable (n1)(n-1)-submanifold ΣU\Sigma_{U} of UU,

  • a surjective CrC^{r}-submersion ρU:UΣU\rho_{U}:U\to\Sigma_{U}, and

  • a CrC^{r}-wave front g~U:ΣU{0}×0n\tilde{g}_{U}:\Sigma_{U}\to\{0\}\times{\mathbb{R}}^{n}_{0}

such that

  1. (1)

    We set ξ^U(x):=ξ^E|ΣU(x)\hat{\xi}_{U}(x):=\hat{\xi}_{E}|_{\Sigma_{U}}(x) (xΣU)(x\in\Sigma_{U}) and denote by ξ^U=(1,νU)\hat{\xi}_{U}=(1,\nu_{U}). Then νU(x)\nu_{U}(x) gives a unit normal vector field of g~U\tilde{g}_{U} in 0n{\mathbb{R}}^{n}_{0} ((by regarding g~U\tilde{g}_{U} is a map into 0n){\mathbb{R}}^{n}_{0}).

  2. (2)

    We set

    (3.2) GU:×ΣU(t,x)g~U(x)+tξ^U(x)1n+1(xΣU,t),\displaystyle G_{U}:{\mathbb{R}}\times\Sigma_{U}\ni(t,x)\mapsto\tilde{g}_{U}(x)+t\hat{\xi}_{U}(x)\in{\mathbb{R}}^{n+1}_{1}\qquad(x\in\Sigma_{U},\,\,t\in{\mathbb{R}}),
    τU(q):=τ(q)(qU).\displaystyle\tau_{U}(q):=\tau(q)\qquad(q\in U).

    Then the map given by

    ΦU(q):=(τU(q),ρU(q))(qU)\Phi_{U}(q):=(\tau_{U}(q),\rho_{U}(q))\qquad(q\in U)

    is a diffeomorphism between UU and IU×ΣUI_{U}\times\Sigma_{U} satisfying

    (3.3) F(q)=GUΦU(q)=g~UρU(q)+τU(q)ξ^UρU(q)(qU),F(q)=G_{U}\circ\Phi_{U}(q)=\tilde{g}_{U}\circ\rho_{U}(q)+\tau_{U}(q)\hat{\xi}_{U}\circ\rho_{U}(q)\qquad(q\in U),

    where IU:=(εU+tU,εU+tU)I_{U}:=(-\varepsilon_{U}+t_{U},\varepsilon_{U}+t_{U}) if εU\varepsilon_{U}\neq\infty and IU=I_{U}={\mathbb{R}} if εU=\varepsilon_{U}=\infty.

  3. (3)

    The map given by

    U:×ΣU(t,x)(GU(t,x),ξ^U(x))1n+1×1n+1,{\mathcal{L}}_{U}:{\mathbb{R}}\times\Sigma_{U}\ni(t,x)\mapsto(G_{U}(t,x),\hat{\xi}_{U}(x))\in{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1},

    is an embedding satisfying

    F(q)=UΦU(q)(qU).{\mathcal{L}}_{F}(q)={\mathcal{L}}_{U}\circ\Phi_{U}(q)\qquad(q\in U).
  4. (4)

    The map defined by

    l~U:ΣUx(g~U(x),ξ^U(x))({0}×0n)×1n+1\tilde{l}_{U}:\Sigma_{U}\ni x\mapsto(\tilde{g}_{U}(x),\hat{\xi}_{U}(x))\in(\{0\}\times{\mathbb{R}}^{n}_{0})\times{\mathbb{R}}^{n+1}_{1}

    is an embedding.

In this setting, if FF itself is LL-complete and U~:=Mn\tilde{U}:=M^{n}, then the above assertions hold by setting IU:=I_{U}:={\mathbb{R}}.

Definition 3.2.

In this setting, if U:×ΣU1n+1×1n+1{\mathcal{L}}_{U}:{\mathbb{R}}\times\Sigma_{U}\to{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1} is an embedding, UU is called an FF-adapted neighborhood of pp. Moreover, (U,ΣU,τU,ρU,g~U,ξ^U,IU)(U,\Sigma_{U},\tau_{U},\rho_{U},\tilde{g}_{U},\hat{\xi}_{U},I_{U}) is called the fundamental FF-data at pp.

Proof of Lemma 3.1.

Let ξE\xi_{E} be the EE-normalized null vector field of FF. Since ξ^E(𝟎)\hat{\xi}_{E}(\neq{\mathbf{0}}) points in the light-like future direction, we have dτ(ξE)>0d\tau(\xi_{E})>0. In particular, by the implicit function theorem, the level set

Σ~n1:=τ1(τ(p))U~(Mn)\tilde{\Sigma}^{n-1}:=\tau^{-1}(\tau(p))\cap\tilde{U}\,(\subset M^{n})

is an embedded hypersurface of U~\tilde{U}. We set

(3.4) t0:=τ(p).t_{0}:=\tau(p).

Since ξE\xi_{E} is transversal to the hypersurface Σ~n1\tilde{\Sigma}^{n-1}, there exist

  • an open interval of the form J:=(ε,ε)J:=(-\varepsilon,\varepsilon) (ε>0\varepsilon>0) or J:=J:={\mathbb{R}},

  • a connected open submanifold Σn1\Sigma^{n-1} of Σ~n1\tilde{\Sigma}^{n-1} satisfying pΣn1p\in\Sigma^{n-1}, and

  • an injective CrC^{r}-immersion

    Ψ~:J×Σn1(t,x)φt(x)U~\tilde{\Psi}:J\times\Sigma^{n-1}\ni(t,x)\mapsto\varphi_{t}(x)\in\tilde{U}

such that (cf. [2, Proposition A.3]) the map tφt(x)t\mapsto\varphi_{t}(x) (xΣn1x\in\Sigma^{n-1}) gives an integral curve of ξE\xi_{E} satisfying φ0(x)=x\varphi_{0}(x)=x. In this setting, if FF is LL-complete and U~=Mn\tilde{U}=M^{n}, then ξE\xi_{E} is a complete vector field of MnM^{n} and we can set J:=J:={\mathbb{R}}.

Since Ψ~\tilde{\Psi} is an injective immersion between manifolds of the same dimension, by the inverse function theorem, Ψ~\tilde{\Psi} is an open map and so

(3.5) U:=Ψ~(J×Σn1)(U~)U:=\tilde{\Psi}(J\times\Sigma^{n-1})(\subset\tilde{U})

is a neighborhood of pp, and Ψ~\tilde{\Psi} gives a diffeomorphism from J×Σn1J\times\Sigma^{n-1} to UU. Moreover, if FF is LL-complete and U~:=Mn\tilde{U}:=M^{n}, the map Ψ~\tilde{\Psi} gives a diffeomorphism between ×Σn1{\mathbb{R}}\times\Sigma^{n-1} and MnM^{n} by the same reason why Ψ\Psi is a diffeomorphism in the proof of Theorem 2.6.

We define a map by

ρ:Uqπ2Ψ~1(q)Σn1,\rho:U\ni q\mapsto\pi_{2}\circ\tilde{\Psi}^{-1}(q)\in\Sigma^{n-1},

where π2\pi_{2} is the canonical projection of J×Σn1J\times\Sigma^{n-1} onto Σn1\Sigma^{n-1}. Since τφt(x)=t+t0\tau\circ\varphi_{t}(x)=t+t_{0} for xΣn1x\in\Sigma^{n-1} (cf. (3.4)), ρ\rho is a surjective submersion satisfying

(3.6) (τ(q)t0,ρ(q))=Ψ~1(q)(qU).\big{(}\tau(q)-t_{0},\rho(q)\big{)}=\tilde{\Psi}^{-1}(q)\qquad(q\in U).

By Theorem 1.14, Fφt(x)F\circ\varphi_{t}(x) (tJt\in J,  xΣn1x\in\Sigma^{n-1}) lies on a light-like straight line, and so, the vector ξ^Eφt(x)1n+1\hat{\xi}_{E}\circ\varphi_{t}(x)\in{\mathbb{R}}^{n+1}_{1} does not depend on the parameter tt. So we can write

(ξ^E(x):=)ξ^Eφt(x)=(a(x),ν(x))(xΣn1),(\hat{\xi}_{E}(x):=)\hat{\xi}_{E}\circ\varphi_{t}(x)=(a(x),\nu(x))\qquad(x\in\Sigma^{n-1}),

where a(x)>0a(x)>0 (since ξ^E\hat{\xi}_{E} points in the future direction). We regard τ^1(t0)\hat{\tau}^{-1}(t_{0}) is a Euclidean nn-space. By the same argument as in the proof of Theorem 2.6 (cf. (2.10), (2.11) and (2.12)), we have a=±1a=\pm 1 and can write

(3.7) ξ^E(x)=(1,ν(x)),|ν(x)|=1(xΣn1)\hat{\xi}_{E}(x)=(1,\nu(x)),\quad|\nu(x)|=1\qquad(x\in\Sigma^{n-1})

such that (0,ν)(0,\nu) gives a unit normal vector field of the map

f~:=F|Σn1:Σn1τ^1(t0)(1n+1).\tilde{f}:=F|_{\Sigma^{n-1}}:\Sigma^{n-1}\to\hat{\tau}^{-1}(t_{0})(\subset{\mathbb{R}}^{n+1}_{1}).

So ff is a frontal in τ^1(t0)\hat{\tau}^{-1}(t_{0}). Moreover, since F(φt(x))F(\varphi_{t}(x)) (xΣn1x\in\Sigma^{n-1}) parametrizes a light-like straight line, we have that

(3.8) FΨ~(t,x)=f~(x)+tξ^E(x)(tJ,xΣn1).F\circ\tilde{\Psi}(t,x)=\tilde{f}(x)+t\hat{\xi}_{E}(x)\qquad(t\in J,\,\,x\in\Sigma^{n-1}).

We next show that f~\tilde{f} is a wave front in the hyperplane τ^1(t0)\hat{\tau}^{-1}(t_{0}). Since FF is a null wave front in 1n+1{\mathbb{R}}^{n+1}_{1}, the pair F:=(F,ξ^E){\mathcal{L}}_{F}:=(F,\hat{\xi}_{E}) is an immersion from MnM^{n} into 1n+1×1n+1{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1} (cf. Lemma 1.6). By (3.8), we can write

FΨ~(t,x)=(f~(x)+tξ^E(x),ξ^E(x))((t,x)J×Σn1).{\mathcal{L}}_{F}\circ\tilde{\Psi}(t,x)=\big{(}\tilde{f}(x)+t\hat{\xi}_{E}(x),\hat{\xi}_{E}(x)\big{)}\qquad((t,x)\in J\times\Sigma^{n-1}).

By setting f~(x)=(t0,f(x))\tilde{f}(x)=(t_{0},f(x)) (xΣn1x\in\Sigma^{n-1}), the rank of the matrix (2.5) for t=t0t=t_{0} satisfies

rank(100νfu1+tνu1fun1+tνun1000𝟎νu1νun1)=rank(100νfu1fun1000𝟎νu1νun1).\operatorname{rank}{\begin{pmatrix}1&0&\ldots&0\\ \nu&f_{u_{1}}+t\nu_{u_{1}}&\ldots&f_{u_{n-1}}+t\nu_{u_{n-1}}\\ 0&0&\ldots&0\\ {\mathbf{0}}&\nu_{u_{1}}&\ldots&\nu_{u_{n-1}}\end{pmatrix}}=\operatorname{rank}{\begin{pmatrix}1&0&\ldots&0\\ \nu&f_{u_{1}}&\ldots&f_{u_{n-1}}\\ 0&0&\ldots&0\\ \bf 0&\nu_{u_{1}}&\ldots&\nu_{u_{n-1}}\end{pmatrix}}.

Since the right-hand side is equal to nn, the matrix

(fu1fun1νu1νun1){\begin{pmatrix}f_{u_{1}}&\ldots&f_{u_{n-1}}\\ \nu_{u_{1}}&\ldots&\nu_{u_{n-1}}\end{pmatrix}}

is of rank n1n-1 at each point of Σn1\Sigma^{n-1}. So f~\tilde{f} is a wave front on Σn1\Sigma^{n-1}.

We consider the map g~:Σn1τ^1(0)\tilde{g}:\Sigma^{n-1}\to\hat{\tau}^{-1}(0) given by

g~(x):=f~(x)t0ξ^E(x)(xΣn1),\tilde{g}(x):=\tilde{f}(x)-t_{0}\hat{\xi}_{E}(x)\qquad(x\in\Sigma^{n-1}),

which corresponds to the parallel hypersurface of ff having signed equi-distance t0-t_{0} in 0n{\mathbb{R}}^{n}_{0}. Then g~\tilde{g} is a wave front in τ^1(0)\hat{\tau}^{-1}(0) whose unit normal vector field ν\nu satisfies

(1,ν(x))=ξ^E|Σn1(x)(xΣn1).(1,\nu(x))=\hat{\xi}_{E}|_{\Sigma^{n-1}}(x)\qquad(x\in\Sigma^{n-1}).

Then, by Theorem 2.3,

(3.9) G(t,x):=g~(x)+t(1,ν(x))(xΣn1,t)G(t,x):=\tilde{g}(x)+t(1,\nu(x))\qquad(x\in\Sigma^{n-1},\,\,t\in{\mathbb{R}})

is an LL-complete null wave front. We set

IU:={(ε+t0,ε+t0)if ε,if ε=,I_{U}:=\begin{cases}(-\varepsilon+t_{0},\varepsilon+t_{0})&\text{if $\varepsilon\neq\infty$},\\ {\mathbb{R}}&\text{if $\varepsilon=\infty$},\end{cases}

and consider the following diffeomorphism

Υ:IU×Σn1(t,x)(tt0,x)J×Σn1.\Upsilon:I_{U}\times\Sigma^{n-1}\ni(t,x)\mapsto(t-t_{0},x)\in J\times\Sigma^{n-1}.

By (3.6), it holds that

(3.10) q=Ψ~(τ(q)t0,ρ(q))=Ψ~Υ(τ(q),ρ(q))(qU).q=\tilde{\Psi}(\tau(q)-t_{0},\rho(q))=\tilde{\Psi}\circ\Upsilon(\tau(q),\rho(q))\qquad(q\in U).

By this with (3.8) and (3.9), we have

F(q)\displaystyle F(q) =FΨ~Υ(τ(q),ρ(q))=FΨ~(τ(q)t0,ρ(q))\displaystyle=F\circ\tilde{\Psi}\circ\Upsilon(\tau(q),\rho(q))=F\circ\tilde{\Psi}(\tau(q)-t_{0},\rho(q))
=f~ρ(q)+(τ(q)t0)ξ^Eρ(q)\displaystyle=\tilde{f}\circ\rho(q)+(\tau(q)-t_{0})\hat{\xi}_{E}\circ\rho(q)
=g~ρ(q)+τ(q)ξ^Eρ(q)=G(τ(q),ρ(q))\displaystyle=\tilde{g}\circ\rho(q)+\tau(q)\hat{\xi}_{E}\circ\rho(q)=G(\tau(q),\rho(q))

for each qUq\in U. Thus, by setting

ΦU:=(Ψ~Υ)1,εU:=ε,tU:=t0,ρU:=ρ,\displaystyle\Phi_{U}:=(\tilde{\Psi}\circ\Upsilon)^{-1},\quad\varepsilon_{U}:=\varepsilon,\quad t_{U}:=t_{0},\quad\rho_{U}:=\rho,
νU:=ν,ξ^U:=(1,νU),ΣU:=Σn1,GU:=G,g~U:=g~,\displaystyle\nu_{U}:=\nu,\quad\hat{\xi}_{U}:=(1,\nu_{U}),\quad\Sigma_{U}:=\Sigma^{n-1},\quad G_{U}:=G,\quad\tilde{g}_{U}:=\tilde{g},

we obtain the desired fundamental data: In fact, if we set

U:×Σn1(t,x)(GU(t,x),ξ^U(x))1n+1×1n+1,\displaystyle{\mathcal{L}}_{U}:{\mathbb{R}}\times\Sigma^{n-1}\ni(t,x)\mapsto\big{(}G_{U}(t,x),\hat{\xi}_{U}(x)\big{)}\in{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1},
l~U:Σn1x(g~U(x),ξ^U(x))({0}×0n)×1n+1,\displaystyle\tilde{l}_{U}:\Sigma^{n-1}\ni x\mapsto\big{(}\tilde{g}_{U}(x),\hat{\xi}_{U}(x)\big{)}\in(\{0\}\times{\mathbb{R}}^{n}_{0})\times{\mathbb{R}}^{n+1}_{1},

then the two maps l~U\tilde{l}_{U} and U{\mathcal{L}}_{U} are immersions, since we have already shown that g~U\tilde{g}_{U} and GUG_{U} are wave fronts.

Since every immersion is locally an embedding, if we choose the open neighborhood U(U~)U(\subset\tilde{U}) of pp to be sufficiently small (that is, we choose a sufficiently small Σn1\Sigma^{n-1}, see (3.5)), then l~U\tilde{l}_{U} gives an embedding. Then, by Lemma 2.5, U{\mathcal{L}}_{U} is also an embedding. ∎

By Lemma 3.1, for each point pMnp\in M^{n}, there exist fundamental FF-data

(Up,Σp,τp,ρp,g~p,ξ^p,Ip)(U_{p},\Sigma_{p},\tau_{p},\rho_{p},\tilde{g}_{p},\hat{\xi}_{p},I_{p})

such that UpU_{p} is an FF-adapted neighborhood of pp giving an LL-complete null wave front (cf. (3.2))

Gp:×Σp(t,x)g~p(x)+tξ^p(x)1n+1G_{p}:{\mathbb{R}}\times\Sigma_{p}\ni(t,x)\mapsto\tilde{g}_{p}(x)+t\hat{\xi}_{p}(x)\in{\mathbb{R}}^{n+1}_{1}

and an embedding

(3.11) l~p:Σpx(g~p(x),ξ^p(x))({0}×0n)×1n+11n+1×1n+1,\tilde{l}_{p}:\Sigma_{p}\ni x\mapsto(\tilde{g}_{p}(x),\hat{\xi}_{p}(x))\in(\{0\}\times{\mathbb{R}}^{n}_{0})\times{\mathbb{R}}^{n+1}_{1}\subset{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1},

where ξ^p(x)=(1,νp(x))\hat{\xi}_{p}(x)=(1,\nu_{p}(x)). Since MnM^{n} satisfies the second axiom of the countability, there exist an at most countable set 𝒜{\mathcal{A}} and a family of fundamental FF-data

{(Uλ,Σλ,τλ,ρλ,g~λ,ξ^λ,Iλ)}λ𝒜\{(U_{\lambda},\Sigma_{\lambda},\tau_{\lambda},\rho_{\lambda},\tilde{g}_{\lambda},\hat{\xi}_{\lambda},I_{\lambda})\}_{\lambda\in{\mathcal{A}}}

such that

  1. (1)

    for each λ𝒜\lambda\in{\mathcal{A}}, there exists pλUλp_{\lambda}\in U_{\lambda} satisfying

    (Uλ,Σλ,τλ,ρλ,g~λ,ξ^λ,Iλ):=(Up,Σpλ,τpλ,ρpλ,g~pλ,ξ^pλ,Ipλ),\displaystyle(U_{\lambda},\Sigma_{\lambda},\tau_{\lambda},\rho_{\lambda},\tilde{g}_{\lambda},\hat{\xi}_{\lambda},I_{\lambda}):=(U_{p},\Sigma_{p_{\lambda}},\tau_{p_{\lambda}},\rho_{p_{\lambda}},\tilde{g}_{p_{\lambda}},\hat{\xi}_{p_{\lambda}},I_{p_{\lambda}}),
    g~λ:=g~pλ,\displaystyle\tilde{g}_{\lambda}:=\tilde{g}_{p_{\lambda}},
  2. (2)

    the map l~λ:Σλ1n+1×1n+1\tilde{l}_{\lambda}:\Sigma_{\lambda}\to{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1} defined by

    l~λ(x):=(g~λ(x),ξ^λ(x))(λ𝒜)\tilde{l}_{\lambda}(x):=(\tilde{g}_{\lambda}(x),\hat{\xi}_{\lambda}(x))\qquad(\lambda\in{\mathcal{A}})

    is an embedding (cf. (4) of Lemma 3.1),

  3. (3)

    Mn=λ𝒜UλM^{n}=\bigcup_{\lambda\in{\mathcal{A}}}U_{\lambda}.

In particular, {(Σλ,l~λ)}λ𝒜\{(\Sigma_{\lambda},\tilde{l}_{\lambda})\}_{\lambda\in{\mathcal{A}}} is a family of embeddings defined on (n1)(n-1)-dimensional manifolds. To ensure that the domain of definition of the resulting Fˇ\check{F} is a Hausdorff space, we give the following definition:

Definition 3.3.

In the above setting, FF is said to be admissible if {(Σλ,l~λ)}λ𝒜\{(\Sigma_{\lambda},\tilde{l}_{\lambda})\}_{\lambda\in{\mathcal{A}}} is an admissible family as in Definition B.6 in Appendix B.

We prove the following:

Theorem 3.4 (A structure theorem of null wave fronts).

Let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be an admissible null CrC^{r}-wave front ((if FF is real analytic, i.e. r=ωr=\omega, it is admissible, see Lemma B.5)). Then there exist

  • a CrC^{r}-differentiable (n1)(n-1)-manifold Σˇn1\check{\Sigma}^{n-1},

  • a CrC^{r}-immersion :Mn×Σˇn1{\mathcal{I}}:M^{n}\to{\mathbb{R}}\times\check{\Sigma}^{n-1},

  • an LL-complete null wave front Fˇ:×Σˇn11n+1\check{F}:{\mathbb{R}}\times\check{\Sigma}^{n-1}\to{\mathbb{R}}^{n+1}_{1} written in the normal form and its canonical lift ˇ:×Σˇn11n+1×1n+1\check{{\mathcal{L}}}:{\mathbb{R}}\times\check{\Sigma}^{n-1}\to{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1}

such that

Fˇ(q)=F(q),ˇ(q)=F(q)(qMn).\check{F}\circ{\mathcal{I}}(q)=F(q),\qquad\check{{\mathcal{L}}}\circ{\mathcal{I}}(q)={\mathcal{L}}_{F}(q)\qquad(q\in M^{n}).

If F{\mathcal{L}}_{F} is an embedding ((see Remark 3.5)), then {\mathcal{I}} gives a diffeomorphism between MnM^{n} and (Mn){\mathcal{I}}(M^{n}). Moreover, if FF is LL-complete, then {\mathcal{I}} is a surjection.

Remark 3.5.

Since FF is a null wave front, its canonical lift F{\mathcal{L}}_{F} is an immersion. So, the assumption that F{\mathcal{L}}_{F} is an embedding is not so restrictive. On the other hand, even if F{\mathcal{L}}_{F} is an embedding, the admissibility of FF does not hold in general, see Example 3.7.

Proof.

We consider the disjoint union 𝒮:=λ𝒜Σλ,\mathcal{S}:=\coprod_{\lambda\in{\mathcal{A}}}\Sigma_{\lambda}, and give a relation xyx\sim y for x,y𝒮x,y\in\mathcal{S} so that xyx\sim y implies that there exists a pair (λ,μ)𝒜×𝒜(\lambda,\mu)\in{\mathcal{A}}\times{\mathcal{A}} of indices such that

  • Σλ\Sigma_{\lambda} is a neighborhood of xx,

  • Σμ\Sigma_{\mu} is a neighborhood of yy, and

  • xx is (l~λ,l~μ)(\tilde{l}_{\lambda},\tilde{l}_{\mu})-related to yy in the sense of Definition B.1.

As seen in the proof of Proposition B.7, the symbol \sim gives an equivalence relation, and Σˇn1:=𝒮/\check{\Sigma}^{n-1}:=\mathcal{S}/{\sim} is an (n1)(n-1)-manifold. Then π:𝒮Σˇn1\pi:\mathcal{S}\to\check{\Sigma}^{n-1} is the canonical projection as an open map. We set

φλ:=π|Σλ.\varphi_{\lambda}:=\pi|_{\Sigma_{\lambda}}.

Then {(Σλ,φλ)}λ𝒜\{(\Sigma_{\lambda},\varphi_{\lambda})\}_{\lambda\in{\mathcal{A}}} is the differentiable structure of Σˇn1\check{\Sigma}^{n-1} as shown in the proof of Proposition B.7. Moreover, an immersion lˇ:Σˇn11n+1×1n+1\check{l}:\check{\Sigma}^{n-1}\to{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1} is induced which satisfies

lˇφλ(x)=(g~λ,ξ^λ)(=l~λ)(xΣλ,λ𝒜).\check{l}\circ\varphi_{\lambda}(x)=(\tilde{g}_{\lambda},\hat{\xi}_{\lambda})(=\tilde{l}_{\lambda})\qquad(x\in\Sigma_{\lambda},\,\,\lambda\in{\mathcal{A}}).

We can write

lˇ(xˇ)=(g~(xˇ),ξ^(xˇ))(xˇΣˇn1).\check{l}(\check{x})=(\tilde{g}(\check{x}),\hat{\xi}(\check{x}))\qquad(\check{x}\in\check{\Sigma}^{n-1}).

Then, by definition, we have

g~φλ(x)=g~λ(x),ξ^φλ(x)=ξ^λ(x)(xΣλ)\tilde{g}\circ\varphi_{\lambda}(x)=\tilde{g}_{\lambda}(x),\quad\hat{\xi}\circ\varphi_{\lambda}(x)=\hat{\xi}_{\lambda}(x)\qquad(x\in\Sigma_{\lambda})

So, if we set

Fˇ(t,xˇ)\displaystyle\check{F}(t,\check{x}) :=g~(xˇ)+tξ^(xˇ)((t,xˇ)×Σˇn1),\displaystyle:=\tilde{g}(\check{x})+t\hat{\xi}(\check{x})\qquad((t,\check{x})\in{\mathbb{R}}\times\check{\Sigma}^{n-1}),
Gλ(t,x)\displaystyle G_{\lambda}(t,x) :=g~λ(x)+tξ^λ(x)(t,xΣλ),\displaystyle:=\tilde{g}_{\lambda}(x)+t\hat{\xi}_{\lambda}(x)\qquad(t\in{\mathbb{R}},\,x\in\Sigma_{\lambda}),

then it holds that

(3.12) Fˇ(t,φλ(x))\displaystyle\check{F}(t,\varphi_{\lambda}(x)) =g~φλ(x)+tξ^φλ(x)\displaystyle=\tilde{g}\circ\varphi_{\lambda}(x)+t\hat{\xi}\circ\varphi_{\lambda}(x)
=g~λ(x)+tξ^λ(x)=Gλ(t,x)\displaystyle=\tilde{g}_{\lambda}(x)+t\hat{\xi}_{\lambda}(x)=G_{\lambda}(t,x)

for tt\in{\mathbb{R}} and xΣλx\in\Sigma_{\lambda}. So, if we consider the maps defined by

ˇ:×Σˇn1(t,xˇ)(g~(xˇ)+tξ^(xˇ),ξ^(xˇ))1n+1×1n+1,\displaystyle\check{{\mathcal{L}}}:{\mathbb{R}}\times\check{\Sigma}^{n-1}\ni(t,\check{x})\mapsto\Big{(}\tilde{g}(\check{x})+t\hat{\xi}(\check{x}),\hat{\xi}(\check{x})\Big{)}\in{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1},
λ:×Σλ(t,x)(Gλ(t,x),ξ^λ(x))1n+1×1n+1,\displaystyle{\mathcal{L}}_{\lambda}:{\mathbb{R}}\times\Sigma_{\lambda}\ni(t,x)\mapsto\big{(}G_{\lambda}(t,x),\hat{\xi}_{\lambda}(x)\big{)}\in{\mathbb{R}}^{n+1}_{1}\times{\mathbb{R}}^{n+1}_{1},

then we have

ˇ(t,φλ(x))=λ(t,x)(t,xΣλ).\check{{\mathcal{L}}}(t,\varphi_{\lambda}(x))={\mathcal{L}}_{\lambda}(t,x)\qquad(t\in{\mathbb{R}},\,\,x\in\Sigma_{\lambda}).

We set

Σˇλ:=φλ(Σλ).\check{\Sigma}_{\lambda}:=\varphi_{\lambda}(\Sigma_{\lambda}).

Since λ{\mathcal{L}}_{\lambda} is an embedding (see (3) of Lemma 3.1), then the restriction ˇ|×Σˇλ\check{{\mathcal{L}}}|_{{\mathbb{R}}\times\check{\Sigma}_{\lambda}} is also an embedding for each λ𝒜\lambda\in\mathcal{A}. In particular, Fˇ\check{F} is an LL-complete null wave front on ×Σˇn1{\mathbb{R}}\times\check{\Sigma}^{n-1}.

We then fix pMnp\in M^{n} arbitrarily, and assume that pp belongs to UλU_{\lambda} for some λ𝒜\lambda\in{\mathcal{A}}. By (3.3) and (3.12), we have

F(q)\displaystyle F(q) =g~λρλ(q)+τλ(q)ξ^λρλ(q)\displaystyle=\tilde{g}_{\lambda}\circ\rho_{\lambda}(q)+\tau_{\lambda}(q)\,\hat{\xi}_{\lambda}\circ\rho_{\lambda}(q)
=Gλ(ρλ(q),τλ(q))=Fˇ(τλ(q),φλρλ(q))(qUλ).\displaystyle=G_{\lambda}(\rho_{\lambda}(q),\tau_{\lambda}(q))=\check{F}(\tau_{\lambda}(q),\varphi_{\lambda}\circ\rho_{\lambda}(q))\qquad(q\in U_{\lambda}).

So it holds that

(3.13) F(q)=ˇ(τλ(q),φλρλ(q))(qUλ).{\mathcal{L}}_{F}(q)=\check{{\mathcal{L}}}(\tau_{\lambda}(q),\varphi_{\lambda}\circ\rho_{\lambda}(q))\qquad(q\in U_{\lambda}).

Since F{\mathcal{L}}_{F} and ˇ\check{{\mathcal{L}}} are immersions, the map

(3.14) λ:Uλq(τλ(q),φλρλ(q))×Σˇλ{\mathcal{I}}_{\lambda}:U_{\lambda}\ni q\mapsto(\tau_{\lambda}(q),\varphi_{\lambda}\circ\rho_{\lambda}(q))\in{\mathbb{R}}\times\check{\Sigma}_{\lambda}

is also an immersion.

If pUμp\in U_{\mu} for some other index μ𝒜\mu\in{\mathcal{A}}, then the immersion μ:Uμ×Σˇμ{\mathcal{I}}_{\mu}:U_{\mu}\to{\mathbb{R}}\times\check{\Sigma}_{\mu} is also induced. Since ˇ|×Σˇλ\check{{\mathcal{L}}}|_{{\mathbb{R}}\times\check{\Sigma}_{\lambda}} and ˇ|×Σˇμ\check{{\mathcal{L}}}|_{{\mathbb{R}}\times\check{\Sigma}_{\mu}} are embeddings, so is ˇ|×(ΣˇλΣˇμ)\check{{\mathcal{L}}}|_{{\mathbb{R}}\times(\check{\Sigma}_{\lambda}\cap\check{\Sigma}_{\mu})}. Thus λ{\mathcal{I}}_{\lambda} coincides with μ{\mathcal{I}}_{\mu} on UλUμU_{\lambda}\cap U_{\mu}, and the map :Mn×Σˇn1{\mathcal{I}}:M^{n}\to{\mathbb{R}}\times\check{\Sigma}^{n-1} is canonically induced. By (3.13) and (3.14),

F(q)=ˇ(q)(qMn){\mathcal{L}}_{F}(q)=\check{{\mathcal{L}}}\circ{\mathcal{I}}(q)\qquad(q\in M^{n})

holds. In particular, Fˇ=F\check{F}\circ{\mathcal{I}}=F also holds on MnM^{n}. We now consider the case that F{\mathcal{L}}_{F} is an embedding. Suppose that (p)=(q){\mathcal{I}}(p)={\mathcal{I}}(q) (p,qMnp,q\in M^{n}). Then we have

F(p)=ˇ(p)=ˇ(q)=F(q).{\mathcal{L}}_{F}(p)=\check{{\mathcal{L}}}\circ{\mathcal{I}}(p)=\check{{\mathcal{L}}}\circ{\mathcal{I}}(q)={\mathcal{L}}_{F}(q).

Since F{\mathcal{L}}_{F} is an embedding, we have p=qp=q, proving the injectivity of {\mathcal{I}}. Since {\mathcal{I}} is an immersion between same dimensional manifolds, it is a diffeomorphism.

If FF is LL-complete, then τλ(Uλ)=\tau_{\lambda}(U_{\lambda})={\mathbb{R}} and ρλ(Uλ)=Σλ\rho_{\lambda}(U_{\lambda})=\Sigma_{\lambda} for each λ𝒜\lambda\in{\mathcal{A}}, which imply that the map λ{\mathcal{I}}_{\lambda} is surjective (cf. (3.14)). So we can conclude that {\mathcal{I}} is a surjection. ∎

Example 3.6.

We set 2:=2{(0,0)}{\mathbb{R}}^{2}_{*}:={\mathbb{R}}^{2}\setminus\{(0,0)\}, and consider a null immersion

F:2(u,v)(u2+v2,2uv,u2v2)(13;t,x,y),F:{\mathbb{R}}^{2}_{*}\ni(u,v)\mapsto(u^{2}+v^{2},2uv,u^{2}-v^{2})\in({\mathbb{R}}^{3}_{1};t,x,y),

whose image lies on the light-cone Λ2\Lambda^{2} passing through the origin in 13{\mathbb{R}}^{3}_{1}. In this case, the image of the LL-completion of FF is the light-cone Λ2\Lambda^{2} and so the map FF covers twice on Λ2{t>0}\Lambda^{2}\cap\{t>0\} in 13{\mathbb{R}}^{3}_{1}. In particular, the induced map {\mathcal{I}} is not a diffeomorphism. This example shows that if we drop the condition that F{\mathcal{L}}_{F} is injective, the injectivity of Φ\Phi does not follow, in general.

Refer to caption
Figure 3. The image of the curve γ\gamma in Example 3.7.
Example 3.7.

Consider a plane curve (cf. Figure 3)

γ(t):=eω(t)(cost,sint)(0t4π),\gamma(t):=e^{\omega(t)}(\cos t,\sin t)\qquad(0\leq t\leq 4\pi),

where ω(t)\omega(t) is a CC^{\infty}-function into [0,1][0,1] such that ω(t)=0\omega(t)=0 for t(3πε,3π+ε)t\not\in(3\pi-\varepsilon,3\pi+\varepsilon) and ω(t)>0\omega(t)>0 for t(3πε,3π+ε)t\in(3\pi-\varepsilon,3\pi+\varepsilon), where ε\varepsilon is a sufficiently small positive number. Then γ\gamma generates the LL-complete null wave front +γ\mathcal{F}^{\gamma}_{+} (cf. Theorem 2.3). By the construction of γ\gamma, +γ\mathcal{F}^{\gamma}_{+} is not admissible, since the image of +γ\mathcal{F}^{\gamma}_{+} has non-transversal intersections. We then set

Γ(t):=(0,γ(t))+t(1,ν(t))(0t4π),\Gamma(t):=(0,\gamma(t))+t(1,\nu(t))\qquad(0\leq t\leq 4\pi),

where ν(t)\nu(t) is a unit normal vector field of γ(t)\gamma(t). The image C:=Γ([0,4π])C:=\Gamma([0,4\pi]) of Γ\Gamma is an embedded spiral-shaped curve on image of +γ\mathcal{F}^{\gamma}_{+}. Consider a sufficiently small tubular neighborhood 𝒰C{\mathcal{U}}_{C} of CC in the image of +γ\mathcal{F}^{\gamma}_{+}. Then 𝒰C{\mathcal{U}}_{C} is an embedded null surface, because Γ\Gamma has no self-intersections. If we think that FF is the inclusion map 𝒰C13{\mathcal{U}}_{C}\hookrightarrow{\mathbb{R}}^{3}_{1}, then the LL-completion of FF is +γ\mathcal{F}^{\gamma}_{+}. So, this example shows that the admissibility of the LL-completion of FF in Theorem 3.4 is independent of the embeddedness of the canonical lift F{\mathcal{L}}_{F}.

Moreover, if we apply our LL-completion procedure for 𝒰C{\mathcal{U}}_{C}, then the resulting quotient space is not a Hausdorff space, which implies the necessity of the admissibility condition as in Definition 3.3.

4. Classification of complete null wave fronts

In this section, we define “completeness” of null wave fronts and prove a structure theorem of them. The content of this section is completely independent of Section 3.

Definition 4.1.

A null CrC^{r}-wave front F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} is said to be complete if

  • FF is LL-complete, and

  • the singular set of FF is a non-empty compact subset of MnM^{n}.

Remark 4.2.

As proved in [2], if FF is LL-complete and its singular set is empty, then the image F(Mn)F(M^{n}) is a subset of a light-like hyperplane of 1n+1{\mathbb{R}}^{n+1}_{1}, which is a trivial case. So we only consider the case that the singular set of FF is non-empty, in the above definition.

Remark 4.3.

When n=2n=2, a null wave front FF in 13{\mathbb{R}}^{3}_{1} is a surface with vanishing Gaussian curvature (cf. [1, Appendix]). If FF is complete, then it is complete as a flat front in the Euclidean 3-space in the sense of Murata-Umehara [9].

We prove the following:

Theorem 4.4.

Let FF be a null CrC^{r}-wave front in 1n+1{\mathbb{R}}^{n+1}_{1}. If FF is complete, then it can be reparametrized as a normal form +f{\mathcal{F}}^{f}_{+}, where f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0} is an immersion defined on a compact (n1)(n-1)-manifold Σn1\Sigma^{n-1} whose principal curvatures are non-zero and all same sign everywhere in 0n{\mathbb{R}}^{n}_{0}. In particular, if n3n\geq 3, then ff is a compact convex hypersurface in 0n{\mathbb{R}}^{n}_{0}. On the other hand, if n=2n=2, then ff is a closed locally convex regular curve in 02{\mathbb{R}}^{2}_{0}.

Proof.

We let F:Mn1n+1F:M^{n}\to{\mathbb{R}}^{n+1}_{1} be a complete null wave front. By Theorem 2.6, we may assume that there exist

  • an (n1)(n-1)-manifold Σn1\Sigma^{n-1},

  • a diffeomorphism Φ:Mn×Σn1\Phi:M^{n}\to{\mathbb{R}}\times\Sigma^{n-1}, and

  • a co-orientable wave front f:Σn10nf:\Sigma^{n-1}\to{\mathbb{R}}^{n}_{0}

such that FΦ1F\circ\Phi^{-1} coincides with the normal form +f{\mathcal{F}}^{f}_{+} given in Theorem 2.3. We consider the height function τ^:1n+1\hat{\tau}:{\mathbb{R}}^{n+1}_{1}\to{\mathbb{R}} given in (2.1). Since the singular set of FF is compact, for sufficiently large t0t_{0}, the restriction

f~:=FΦ1|Φ(τ1(t0)):Σn1τ^1(t0)\tilde{f}:=F\circ\Phi^{-1}|_{\Phi(\tau^{-1}(t_{0}))}:\Sigma^{n-1}\to\hat{\tau}^{-1}(t_{0})

is an immersion such that f~(x)=(t0,f(x))\tilde{f}(x)=(t_{0},f(x)). Without loss of generality, we may assume that t0=0t_{0}=0 and FF satisfies

(4.1) FΦ1(t,x)=f~(x)+tξ^E(x),ξ^E(x):=(1,ν(x))(t,xΣn1),F\circ\Phi^{-1}(t,x)=\tilde{f}(x)+t\hat{\xi}_{E}(x),\quad\hat{\xi}_{E}(x):=(1,\nu(x))\qquad(t\in{\mathbb{R}},\,\,x\in\Sigma^{n-1}),

where ν\nu can be considered as the unit normal vector field along the immersion ff. So, we may assume that FF is defined on ×Σn1{\mathbb{R}}\times\Sigma^{n-1}. Since ff is co-orientable and is an immersion, Σn1\Sigma^{n-1} is orientable. We let

λ1λn1\lambda_{1}\leq\cdots\leq\lambda_{n-1}

be principal curvature functions of ff. Here, {λi(x)}i=1n1\{\lambda_{i}(x)\}_{i=1}^{n-1} are eigenvalues of a symmetric matrix associated with the shape operator of ff at xx. Since the characteristic polynomial of a real symmetric matrix consists only of real roots, the well-known fact that the roots of a polynomial depend continuously on its coefficients implies that {λi}i=1n1\{\lambda_{i}\}_{i=1}^{n-1} can be considered as a family of real-valued continuous functions defined on Σn1\Sigma^{n-1}. We first show that each λi\lambda_{i} never changes sign: Suppose that there exists a sequence of points {xk}k=1\{x_{k}\}_{k=1}^{\infty} which converges to a point xx_{\infty} such that

λi(xk)0 and λi(x)=0.\lambda_{i}(x_{k})\neq 0\,\,\text{ and }\,\,\lambda_{i}(x_{\infty})=0.

Then

(1λi(xk),xk)×Σn1(k=1,2,3,)\left(\frac{1}{\lambda_{i}(x_{k})},x_{k}\right)\in{\mathbb{R}}\times\Sigma^{n-1}\qquad(k=1,2,3,\ldots)

are singular points of FF which are unbounded on ×Σn1{\mathbb{R}}\times\Sigma^{n-1}, contradicting the compactness of the singular set of FF. Thus, each λi\lambda_{i} as a continuous function on Σn1\Sigma^{n-1} has no zeros unless it is identically zero. By Remark 4.2, ff is not a part of a hyperplane in 0n{\mathbb{R}}^{n}_{0}. So, there exists an integer rr (1rn11\leq r\leq n-1) such that λis(s=1,,r)\lambda_{i_{s}}\,\,(s=1,\ldots,r) are not identically zero, where {i1,,ir}\{i_{1},\ldots,i_{r}\} is a subset of {1,,n1}\{1,\ldots,n-1\}. By Hartman’s product theorem (cf. [8, Page 347]), Σn1\Sigma^{n-1} is a product of a compact manifold and l{\mathbb{R}}^{l} (l:=n1rl:=n-1-r). For each s{1,,r}s\in\{1,\ldots,r\}, we can define a continuous map ψs:Σn1×Σn1\psi_{s}:\Sigma^{n-1}\to{\mathbb{R}}\times\Sigma^{n-1} by

ψs(x):=(1λis(x),x)×Σn1.\psi_{s}(x):=\left(\frac{1}{\lambda_{i_{s}}(x)},x\right)\in{\mathbb{R}}\times\Sigma^{n-1}.

Then S:=s=1rψs(Σn1)S:=\bigcup_{s=1}^{r}\psi_{s}(\Sigma^{n-1}) coincides with the singular set of FF. Since FF is complete null wave front, SS is compact. Since the projection of SS via the continuous map

×Σn1(t,x)xΣn1{\mathbb{R}}\times\Sigma^{n-1}\ni(t,x)\mapsto x\in\Sigma^{n-1}

is compact, the hypersurface Σn1\Sigma^{n-1} is also compact. Thus r=n1r=n-1 (i.e. l=0l=0) and each λi\lambda_{i} (i=1,,n1i=1,\ldots,n-1) is either positive-valued or negative-valued on Σn1\Sigma^{n-1}. Moreover, since Σn1\Sigma^{n-1} is compact, there exists a point y0Σn1y_{0}\in\Sigma^{n-1} such that f(y0)f(y_{0}) attains the farthest point of the image of ff from the origin in 0n{\mathbb{R}}^{n}_{0}. Then λ1(y0),,λn1(y0)\lambda_{1}(y_{0}),\,\,\dots,\,\,\lambda_{n-1}(y_{0}) are all positive or all negative at the same time.

Here, replacing ν\nu by ν-\nu, we may assume that

λn1(x)>0(xΣn1).\lambda_{n-1}(x)>0\qquad(x\in\Sigma^{n-1}).

Then, λ1,,λn1\lambda_{1},\dots,\lambda_{n-1} take the same sign on Σn1\Sigma^{n-1}. Thus, Hadamard’s theorem [5] implies that ff is an embedded convex hypersurface in 0n{\mathbb{R}}^{n}_{0} whenever n3n\geq 3. ∎

If n=2n=2, then ff is a locally strictly convex regular curve in 02{\mathbb{R}}^{2}_{0}. So we may express ff by γ(s)\gamma(s) (s)(s\in{\mathbb{R}}) defined as an ll-periodic curve (l>0l>0), parametrized by the arc-length. Then its curvature function can be taken so that κ(s)\kappa(s) is positive everywhere. Then we may assume that FF is expressed as

F(t,s)=(t,γt(s))(s,t),F(t,s)=(t,\gamma_{t}(s))\qquad(s,t\in{\mathbb{R}}),

where γt(s):=γ(s)+tν(s)\gamma_{t}(s):=\gamma(s)+t\nu(s) is a parallel curve of γ\gamma, and the singular set of FF is

S:={(1/κ(s),s);s},S:=\{(1/\kappa(s),s)\,;\,s\in{\mathbb{R}}\},

and

(4.2) Cγ(s):=F(1κ(s),s)C_{\gamma}(s):=F\left(\frac{1}{\kappa(s)},s\right)

just parametrizes the singular set image of FF. We can prove the following:

Proposition 4.5.

Let FF be a complete null wave front in 13{\mathbb{R}}^{3}_{1} which is generated by a closed locally convex regular curve γ\gamma in 02{\mathbb{R}}^{2}_{0}. Then non-cuspidal edge points of FF correspond to vertices on γ\gamma, where a vertex is a critical point of the curvature function of γ\gamma.

Proof.

In the setting of (4.2), Cγ(s)𝟎C^{\prime}_{\gamma}(s)\neq{\mathbf{0}} if and only if κ(s0)0\kappa^{\prime}(s_{0})\neq 0, and so FF has a cuspidal edge singular point at (1/κ(s0),s0)(1/\kappa(s_{0}),s_{0}) only when κ(s0)0\kappa^{\prime}(s_{0})\neq 0 (see the criterion in [3] for cuspidal edge). ∎

The following assertion is equivalent to the classical four vertex theorem for convex plane curve:

Corollary 4.6.

Let FF be a complete null wave front in 13{\mathbb{R}}^{3}_{1} which has no self-intersections outside of a compact subset of 13{\mathbb{R}}^{3}_{1}. Then FF has at least four non-cuspidal edge singular points.

Proof.

Since FF has no self-intersections outside of a compact set, FF must be generated by a (closed) convex plane curve γ\gamma in the xyxy-plane. By the classical four vertex theorem, γ\gamma has at least four vertices, then such points corresponds to the non-cuspidal edge points of FF as seen in the proof of Proposition 4.5. So we obtain the assertion. ∎

Since a complete null wave front in 13{\mathbb{R}}^{3}_{1} can be considered as a complete flat front in the Euclidean 33-space, Corollary 4.6 is a special case of the four non-cuspidal edge point theorem given in [9, Theorem D].

Example 4.7.

We consider the complete null wave front FaF_{a} associated with the ellipse as in the introduction which has four vertices when 0<a<10<a<1. They correspond to the swallowtail singular points of the complete null wave front in Figure 1 in the introduction, which satisfies the assumption of Corollary 4.6.

Refer to caption
Figure 4. The null wave front associated with a locally convex curve with exactly one crossing.
Example 4.8.

Consider a locally convex plane curve with a self-intersection

γ(θ):=(12sinθ)(cosθ,sinθ)(0θ2π),\gamma(\theta):=(1-2\sin\theta)(\cos\theta,\sin\theta)\qquad(0\leq\theta\leq 2\pi),

which admits only two vertices. These vertices correspond to two swallowtail singular points of the corresponding null wave front as in Figure 4.

Appendix A A lemma of subspaces in 1n+1{\mathbb{R}}^{n+1}_{1}

We denote by ,\left\langle{},{}\right\rangle the canonical Lorentzian inner product of 1n+1{\mathbb{R}}^{n+1}_{1}. In this section, we prepare the following assertion for vector subspaces in 1n+1{\mathbb{R}}^{n+1}_{1}.

Definition A.1.

A subspace VV of 1n+1{\mathbb{R}}^{n+1}_{1} is said to be perpendicular to a subspace WW if 𝐯,𝐰=0(𝐯V,𝐰W)\left\langle{{\mathbf{v}}},{{\mathbf{w}}}\right\rangle=0\,\,({\mathbf{v}}\in V,\,\,{\mathbf{w}}\in W) holds.

For a subspace VV of 1n+1{\mathbb{R}}^{n+1}_{1}, we set

(A.1) V:={𝐱1n+1;𝐱,𝐯=0for all 𝐯V},V^{\perp}:=\{{\mathbf{x}}\in{\mathbb{R}}^{n+1}_{1}\,;\,\left\langle{{\mathbf{x}}},{{\mathbf{v}}}\right\rangle=0\,\,\text{for all ${\mathbf{v}}\in V$}\},

which is the maximal subspace perpendicular to VV in 1n+1{\mathbb{R}}^{n+1}_{1}.

Definition A.2.

A subspace VV of 1n+1{\mathbb{R}}^{n+1}_{1} is said to be degenerate if there exists a non-zero vector 𝐯V{\mathbf{v}}\in V such that 𝐯,𝐰=0\left\langle{{\mathbf{v}}},{{\mathbf{w}}}\right\rangle=0 for all 𝐰V{\mathbf{w}}\in V. We call such a vector 𝐯{\mathbf{v}} a degenerate vector of VV. On the other hand, a subspace V(1n+1)V(\subset{\mathbb{R}}^{n+1}_{1}) is said to be non-degenerate if it is not degenerate.

Lemma A.3.

Let V,WV,W and NN be three subspaces of 1n+1{\mathbb{R}}^{n+1}_{1} such that

  1. (1)

    NN is a 1-dimensional degenerate subspace satisfying NW={𝟎}N\cap W=\{{\mathbf{0}}\},

  2. (2)

    VV and WW are perpendicular to NN, and

  3. (3)

    WW is perpendicular to VV.

Then VW={𝟎}V\cap W=\{{\mathbf{0}}\}.

Proof.

If WW is non-degenerate, then so is WW^{\perp}, and using (3), we have

{𝟎}=WWVW,\{{\mathbf{0}}\}=W^{\perp}\cap W\supset V\cap W,

which implies VW={𝟎}V\cap W=\{{\mathbf{0}}\}. So we may assume that WW is degenerate. By definition, there exists a degenerate vector 𝐯0{\mathbf{v}}_{0} belonging to WW. We let ξN{𝟎}\xi\in N\setminus\{{\mathbf{0}}\} be a generator of the 1-dimensional vector NN. Then 𝐯0{\mathbf{v}}_{0} and ξ\xi are both degenerate vectors. In particular, they are light-like, that is, 𝐯0,𝐯0=ξ,ξ=0\left\langle{{\mathbf{v}}_{0}},{{\mathbf{v}}_{0}}\right\rangle=\left\langle{\xi},{\xi}\right\rangle=0 hold. So if we set W~:=W+N\tilde{W}:=W+N, then by (1), W~\tilde{W} contains two linearly independent light-like vectors 𝐯0{\mathbf{v}}_{0} and ξ\xi, and so, W~\tilde{W} is a time-like vector space (cf. [10, Lemma 27, Page 141]), which is non-degenerate. So, W~\tilde{W}^{\perp} is also non-degenerate. Since VW~V\subset\tilde{W}^{\perp} (cf. (2) and (3)) and WW~W\subset\tilde{W}, we have

{𝟎}=W~W~VW,\{{\mathbf{0}}\}=\tilde{W}^{\perp}\cap\tilde{W}\supset V\cap W,

proving VW={𝟎}V\cap W=\{{\mathbf{0}}\}. ∎

Appendix B A method to make an immersion defined on a connected manifold from a family of embeddings.

Definition B.1.

We fix positive integers n,mn,m (n<mn<m). Let U,VU,V be two connected nn-dimensional manifolds. We let f:Umf:U\to{\mathbb{R}}^{m} and g:Vmg:V\to{\mathbb{R}}^{m} be two CrC^{r}-embeddings. A point xUx\in U is said to be (f,g)(f,g)-related to yVy\in V if there exist an open neighborhood Ox(U)O_{x}(\subset U) of xx and an open neighborhood Oy(V)O_{y}(\subset V) of yy such that

  • f(x)=g(y)f(x)=g(y) and

  • f(Ox)=g(Oy)f(O_{x})=g(O_{y}).

If there are no (f,g)(f,g)-related points on UU and VV, we say that “UU is not (f,g)(f,g)-related to VV”.

Remark B.2.

This “(f,g)(f,g)-relatedness” is an open condition. In fact, if xUx\in U is (f,g)(f,g)-related to yVy\in V, then there exist open neighborhoods Ox(U)O_{x}(\subset U) and Oy(V)O_{y}(\subset V) of xx and yy respectively such that each point of OxO_{x} is (f,g)(f,g)-related to a certain point in OyO_{y}.

In this setting, the following assertion holds.

Lemma B.3.

If we set

OU,V\displaystyle O_{U,V} :={xU;x is (f,g)-related to some yV},\displaystyle:=\{x\in U\,;\,\text{$x$ is $(f,g)$-related to some $y\in V$}\},
OV,U\displaystyle O_{V,U} :={yV;y is (g,f)-related to some xU},\displaystyle:=\{y\in V\,;\,\text{$y$ is $(g,f)$-related to some $x\in U$}\},

then OU,VO_{U,V} ((resp. OV,U)O_{V,U}) is an open subset of UU ((resp. V)V). Moreover, if OU,VO_{U,V} is non-empty, then there exists a unique CrC^{r}-diffeomorphism φ:OU,VOV,U\varphi:O_{U,V}\to O_{V,U} satisfying gφ=fg\circ\varphi=f on OU,VO_{U,V}.

Proof.

Assume that xUx\in U (resp. yVy\in V) is (f,g)(f,g)-related to yVy\in V (resp. xUx\in U). Since ff and gg are embeddings, yVy\in V (resp. xUx\in U) is uniquely determined, and so the map φ\varphi is also uniquely determined. The smoothness of φ\varphi is obvious. ∎

Definition B.4.

Let f:Umf:U\to{\mathbb{R}}^{m} and g:Vmg:V\to{\mathbb{R}}^{m} be as in Definition B.1. Then the pair (U,V)(U,V) is said to be (f,g)(f,g)-admissible if, for each pair (x,y)U×V(x,y)\in U\times V, it holds that

  • (i)

    xx is (f,g)(f,g)-related to yy, or

  • (ii)

    there exist a neighborhood Ox(U)O_{x}(\subset U) of xx and a neighborhood Oy(V)O_{y}(\subset V) of yy such that OxO_{x} is not (f,g)(f,g)-related to OyO_{y}.

By definition, if f(U)f(U) does not meet g(V)g(V), then the pair (U,V)(U,V) is (f,g)(f,g)-admissible. The following assertion is immediately from the definition:

Lemma B.5.

In the setting of Definition B.4, the pair (U,V)(U,V) is (f,g)(f,g)-admissible if f(U)f(U) meets g(V)g(V) transversally or ff and gg are both real analytic.

Proof.

We fix a pair (x,y)U×V(x,y)\in U\times V such that xx is not (f,g)(f,g)-related to yy. If f(U)f(U) meets g(V)g(V) transversally, then the assertion is obvious. So we may assume that ff and gg are both real analytic and f(U)f(U) does not meet g(V)g(V) transversally. Then P:=f(x)=f(y)P:=f(x)=f(y) and we can take nn-dimensional affine plane TnT^{n} as a common tangential space of f(U)f(U) and g(V)g(V) at PP in m{\mathbb{R}}^{m}. Since UU (resp. VV) is locally connected, there exists a connected open neighborhood U1U_{1} (resp. V1V_{1}) of xx (resp. yy) such that U1UU_{1}\subset U (resp. V1VV_{1}\subset V). By the implicit function theorem, we may assume that the images of the maps f|U1f|_{U_{1}} and g|V1g|_{V_{1}} are expressed as the graphs of certain functions F,G:ΩmnF,G:\Omega\to{\mathbb{R}}^{m-n} defined on the same domain Ω\Omega in TnT^{n}.

It suffices to show that there exist a neighborhood Ox(U1)O_{x}(\subset U_{1}) of xx and a neighborhood Oy(V1)O_{y}(\subset V_{1}) of yy such that OxO_{x} is not (f,g)(f,g)-related to OyO_{y}. If not, there exists a pair (x1,y1)Ox×Oy(x_{1},y_{1})\in O_{x}\times O_{y} such that x1x_{1} is (f,g)(f,g)-related to y1y_{1}, which implies that there exist open neighborhoods O1(Ox)O_{1}(\subset O_{x}) and O2(Oy)O_{2}(\subset O_{y}) of x1x_{1} and y1y_{1}, respectively, such that each point of O1O_{1} is (f,g)(f,g)-related to a corresponding point in O2O_{2}. Then the vector-valued function FF coincides with GG on some non-empty open subset of Ω\Omega. By the connectedness of Ω\Omega and the real analyticity of FF and GG, the two vector-valued functions coincide identically on Ω\Omega, which implies that xx is (f,g)(f,g)-related to yy, a contradiction. ∎

Definition B.6.

Let 𝒜{\mathcal{A}} be an at most countable set, and let {(Uλ,fλ)}λ𝒜\{(U_{\lambda},f_{\lambda})\}_{\lambda\in{\mathcal{A}}} be a family consisting of connected nn-dimensional manifolds UλU_{\lambda} and embeddings fλ:Uλmf_{\lambda}:U_{\lambda}\to{\mathbb{R}}^{m}. Then {(Uλ,fλ)}λ𝒜\{(U_{\lambda},f_{\lambda})\}_{\lambda\in{\mathcal{A}}} is called admissible if (Uλ,Uμ)(U_{\lambda},U_{\mu}) is (fλ,fμ)(f_{\lambda},f_{\mu})-admissible for any choice of (λ,μ)𝒜×𝒜(\lambda,\mu)\in{\mathcal{A}}\times{\mathcal{A}}.

We let {(Uλ,fλ)}λ𝒜\{(U_{\lambda},f_{\lambda})\}_{\lambda\in{\mathcal{A}}} be an admissible family as in Definition B.6. Consider the disjoint union 𝒮:=λ𝒜Uλ,\mathcal{S}:=\coprod_{\lambda\in{\mathcal{A}}}U_{\lambda}, and define a relation xyx\sim y for x,y𝒮x,y\in\mathcal{S} so that xyx\sim y implies that there exist a neighborhood UλU_{\lambda} (λ𝒜\lambda\in{\mathcal{A}}) of xx and a neighborhood UμU_{\mu} (μ𝒜\mu\in{\mathcal{A}}) of yy such that xx is (fλ,fμ)(f_{\lambda},f_{\mu})-related to yy. Then it is easy to check that \sim is an equivalence relation. Moreover, the following assertion holds:

Proposition B.7.

Let {(Uλ,fλ)}λ𝒜\{(U_{\lambda},f_{\lambda})\}_{\lambda\in{\mathcal{A}}} be an admissible family. Then there exist

  • a manifold Mn:=𝒮/M^{n}:=\mathcal{S}/{\sim},

  • a CrC^{r}-immersion g:Mnmg:M^{n}\to{\mathbb{R}}^{m}, and

  • a diffeomorphism Ψλ:UλΨλ(Uλ)Mn\Psi_{\lambda}:U_{\lambda}\to\Psi_{\lambda}(U_{\lambda})\subset M^{n}

such that {(Uλ,Ψλ)}λ𝒜\{(U_{\lambda},\Psi_{\lambda})\}_{\lambda\in{\mathcal{A}}} is a differentiable structure of MnM^{n} and gΨλg\circ\Psi_{\lambda} coincides with fλf_{\lambda} on UλU_{\lambda} for each λ𝒜\lambda\in{\mathcal{A}}.

Proof.

Since 𝒮\mathcal{S} is a disjoint union of open subsets, this UλU_{\lambda} is uniquely determined by xx. So we denote UλU_{\lambda} by VxV_{x}. As we have already noted, \sim is an equivalence relation, and the canonical projection π:𝒮Mn:=𝒮/\pi:\mathcal{S}\to M^{n}:=\mathcal{S}/{\sim} is induced. Since 𝒜{\mathcal{A}} is an at most countable set, MnM^{n} satisfies the second axiom of countability. If we show that the quotient space MnM^{n} is a Hausdorff space, then we can easy observe that MnM^{n} has a structure of CrC^{r}-manifold by using Lemma B.3, and we can construct the desired CrC^{r}-immersion g:Mnmg:M^{n}\to{\mathbb{R}}^{m} by setting

(B.1) Ψλ:=π|Uλ.\Psi_{\lambda}:=\pi|_{U_{\lambda}}.

So it is sufficient to prove that MnM^{n} is a Hausdorff space: Consider the set defined by

:={(p,q)𝒮×𝒮;pq}.\mathcal{R}:=\{(p,q)\in\mathcal{S}\times\mathcal{S}\,;\,p\,\,\sim\,\,q\}.

By Remark B.2, the canonical projection π:𝒮Mn\pi:\mathcal{S}\to M^{n} is an open map. So we prove that \mathcal{R} is a closed subset of 𝒮×𝒮\mathcal{S}\times\mathcal{S} (cf. [6, Chapter 3, Theorem 11]). We consider a sequence {(pk,qk)}k=1\{(p_{k},q_{k})\}_{k=1}^{\infty} in \mathcal{R} and suppose that {pk}k=1\{p_{k}\}_{k=1}^{\infty} and {qk}k=1\{q_{k}\}_{k=1}^{\infty} converge to pp and qq in 𝒮\mathcal{S}, respectively. By definition, there exist (Uλ,fλ)(U_{\lambda},f_{\lambda}) and (Uμ,fμ)(U_{\mu},f_{\mu}) (λ,μ𝒜\lambda,\mu\in{\mathcal{A}}) and a positive integer ll such that

  • Vp=UλV_{p}=U_{\lambda} and Vq=UμV_{q}=U_{\mu},

  • Vpk=UλV_{p_{k}}=U_{\lambda} and Vqk=UμV_{q_{k}}=U_{\mu} for klk\geq l.

We remark that the admissibility of {(Uλ,fλ)}λ𝒜\{(U_{\lambda},f_{\lambda})\}_{\lambda\in{\mathcal{A}}} implies the admissibility between UλU_{\lambda} and UμU_{\mu}. Since pkpp_{k}\to p and qkqq_{k}\to q, we have fλ(p)=fμ(q)f_{\lambda}(p)=f_{\mu}(q). We suppose p≁qp\not\sim q. Then, by (ii) of Definition B.4, there exist a neighborhood Op(Uλ)O_{p}(\subset U_{\lambda}) of pp and a neighborhood Oq(Uμ)O_{q}(\subset U_{\mu}) of qq such that OpO_{p} is not (fλ,fμ)(f_{\lambda},f_{\mu})-related to OqO_{q}. However, this contradicts the fact that pkqkp_{k}\sim q_{k}. So MnM^{n} is a Hausdorff space. ∎

Acknowledgements.

The authors thank Riku Kishida and the reviewer for valuable comments.

References

  • [1] S. Akamine, M. Umehara and K. Yamada, Improvement of the Bernstein-type theorem for space-like zero mean curvature graphs in Lorentz-Minkowski space using fluid mechanical duality, Proc. Amer. Math. Soc. Ser. B, 7 (2020), 17–27.
  • [2] S. Akamine, A. Honda, M. Umehara and K. Yamada, Null hypersurfaces in Lorentzian manifolds with the null energy condition, J. Geom. Phys. 155 (2020), 103751, 6pp.
  • [3] M. Kokubu, W. Rossman, K. Saji, M. Umehara and K. Yamada, Singularities of flat fronts in hyperbolic 3-space, Pacific J. Math. 221 (2005), 303–351.
  • [4] K. L. Duggal and A. Gimenez, Lightlike hypersurfaces of Lorentzian manifolds with distinguished screen, Journal of Geometry and Physics 55 (2005), 107–122.
  • [5] J. Hadamard, Sur certaines propriétés des trajectories en dynamique, J. Math. Pures Appl. 3, (1839), 331-387.
  • [6] J. L. Kelley, General Topology, Dover Books on Mathematics, 2017.
  • [7] M. Kossowski, The intrinsic conformal structure and Gauss map of a light-like hypersurface in Minkowski space Trans. Amer. Math. Soc. 316 (1989), 369–383.
  • [8] S. Kobayashi and K. Nomizu, Foundations of differential geometry. Vol. II, Reprint of the 1969 original. Wiley Classics Library. A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1996.
  • [9] S. Murata and M. Umehara, Flat surfaces with singularities in Euclidean 3-space, Journal of Differential Geometry 82 (2009), 279–316.
  • [10] B. O’Neill, Semi-Riemannian Geometry, Academic Press 1983, USA.
  • [11] K. Saji, M. Umehara and K. Yamada, AkA_{k}-singularities of wave fronts, Math. Proc. Camb. Phil. Soc. 146 (2009) 731–745.