Null hypersurfaces as wave fronts in Lorentz-Minkowski space
Abstract.
In this paper, we show that “-complete null hypersurfaces” (i.e. ruled hypersurfaces foliated by entirety of light-like lines) as wave fronts in the -dimensional Lorentz-Minkowski space are canonically induced by hypersurfaces in the -dimensional Euclidean space. As an application, we show that most of null wave fronts can be realized as restrictions of certain -complete null wave fronts. Moreover, we determine -complete null wave fronts whose singular sets are compact.
Key words and phrases:
light-like hypersurface, null hypersurface, singular point, wave front2010 Mathematics Subject Classification:
Primary 53C50; Secondary 53C42, 53B30.Introduction
We denote by the -dimensional Lorentz-Minkowski space. A null hypersurface in is a -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 -complete if each light-like line is the entirety of a straight line in (see Definition 2.1 for details).
In the authors’ previous work [2], it was shown that -complete null immersed hypersurfaces in 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 .
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 -dimensional Euclidean space () induces an associated parallel family, and this family can be considered as a section of a null hypersurface in . For example, the light-cone
(0.1) |
is a typical example of an -complete null wave front, which corresponds to the family of parallel hypersurfaces of the unit sphere in centered at the origin. The origin as the cone-like singular point of corresponds to the parallel hypersurface of just shrinking to a point. In general, a null hypersurface may have various singular points. For example, for each , we consider an ellipse
(0.2) |
in the Euclidean plane. Let be the leftward unit normal vector field along . Then the map defined by
(0.3) |
gives a null wave front having cuspidal edges and four swallowtails whenever (the definitions of cuspidal edges and swallowtails are given in [3]). Each slice of the image of by the horizontal plane corresponds to the parallel curve of the ellipse of equi-distance . When , the image of just coincides with the light-cone of .

In Section 2, we prove the converse assertion (cf. Theorem 2.6) as a fundamental theorem of null wave fronts, which states that any -complete null wave fronts in are induced from wave fronts in . As a consequence, we can say that an -complete null wave front is associated with the parallel family of a wave front in .
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 null wave fronts with a certain genericity) can be obtained as restrictions of -complete null wave fronts. There are several interesting geometric structures on null hypersurfaces in without assuming -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 -complete null wave front in is called complete if its singular set is a non-empty compact subset in the domain of definition. For example, the light-cone in and the null wave front in 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 if (cf. Theorem 4.4). When , we show that complete null wave fronts are induced by parallel families of locally convex closed regular curves in the Euclidean plane . If the curve is an ellipse, the null wave front corresponds to 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 (cf. Corollary 4.6).
There is one point to note when reading this paper: Readers who are interested only in the fundamental theorem for -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
We first recall the definition of wave fronts. Let be the dual vector space of , and we denote by the projective space associated with . We let
be the canonical projection. Then, for each element , there exists a linear function (which is an element of ) such that . Since the kernel of is invariant under non-zero scalar multiplications, the -dimensional subspace
is well-defined.
In this paper, we set or and “” means smoothness if and real analyticity if . We fix a -differentiable -manifold .
Definition 1.1.
Let be a -map. Then the map is called a -wave front (or simply a -front) if for each , there exist a neighborhood of and a -map such that
-
(1)
satisfies , and
-
(2)
is an immersion.
Moreover, if we can take , the map is said to be co-orientable. In this case, is called the Gauss map of , and the map is called the lift of .
If a -wave front is not co-orientable, taking the double covering , the composition becomes a co-orientable wave front. So without loss of generality, we may assume that itself is co-orientable.
We let be Lorentz-Minkowski -space of signature , and denote by the canonical Lorentzian inner product on .
Proposition 1.2.
Let be a co-orientable -wave front. Then there exists a vector field without zeros which can be considered as a map such that
A vector field along given in Proposition 1.2 is called a normal vector field along .
Proof.
Let be a lift of the Gauss map of . Since is non-degenerate, there exists a vector field on along such that
(1.1) |
Since for each , the vector field has no zeros on . ∎
We then introduce ‘null wave fronts’ as follows:
Definition 1.3.
In the setting of Proposition 1.2, the -wave front is said to be null or light-like if points in the light-like direction in at each .
We denote by the canonical positive definite inner product on .
Definition 1.4 (-normalized normal vector field).
Let be a (co-orientable) null -wave front. A normal vector field along is said to be -normalized if points in the future direction and satisfies , where
Since such a vector field is uniquely determined, we denote it by .
Remark 1.5.
One can replace with any positive constant . However, the choice of makes sense in the following reason: As we will show in Theorem 2.3, a hypersurface in Euclidean space with unit normal vector field induces null wave fronts whose -normalized normal vector fields are given by .
We can always take the -normalized normal vector field for a given co-orientable null wave front as follows: By Proposition 1.2, we can take a normal vector field along . Since is light-like, gives a light-like vector field along . By replacing by , we may assume that points in the future direction, and
gives the desired vector field.
Lemma 1.6.
Let be a co-orientable null wave front. Then
(1.2) |
is an immersion into .
We call this the canonical lift of .
Proof.
We denote by the sphere of radius centered at the origin in with respect to the metric . Since
gives the double covering of , it can be easily observed that is an immersion into if and only if is an immersion, where is the map induced by given in (1.1). ∎
The following assertion gives a characterization of null wave fronts:
Proposition 1.7.
Let be a -map. Then is a co-orientable null wave front if and only if there exists a vector field along defined on such that
-
(1)
,
-
(2)
is pointing in the future light-like direction, and
-
(3)
is an immersion into satisfying for each .
Proof.
We have already seen that any null wave front in uniquely induces an -normalized normal vector field . So it is sufficient to show the converse. Suppose that is a vector field along satisfying the above three conditions. We let be the map given in (1.1). By (1) and (3), is an immersion by the same reason as in the proof of Lemma 1.6. ∎
Lemma 1.8.
Let be a null -immersion with a normal vector field on as in Proposition 1.7. Then for each , there exists a unique tangent vector such that . Moreover, the correspondence is -differentiable.
Proof.
Since is an -dimensional vector subspace of at each , it holds that
Since and is injective, the desired vector should be
Since depends smoothly on , we obtain the assertion. ∎
Later, we will show that the assertion of Lemma 1.8 is extended for null -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 be a co-orientable null -wave front and a normal vector field along . Then, for each , there exists a tangent vector such that .
Proof.
We may assume that . If gives an immersion at , then the assertion follows from Lemma 1.8. So we may assume that is a singular point i.e. is a point where is not an immersion of . We can take a local coordinate system of centered at such that the family of vectors
spans the kernel of , where (). By setting
the vectors are linearly independent. We denote by the subspace of spanned by these vectors. We set
(1.3) |
If we think and are column vector-valued functions, we can consider the -matrix
(1.4) |
at as the Jacobi matrix of the map of into , which can be computed as
Since is a wave front, the matrix is of rank and so
are linearly independent at , which gives a basis of the vector space defined by
(1.5) |
Suppose does not belong to . We set , that is, it is the 1-dimensional vector space generated by . By the following Lemma 1.10, holds, and , and satisfy the assumption of Lemma A.3 in Appendix A. So we have . Since is of dimension and is of dimension , the direct sum coincides with . Then vanishes because is perpendicular to and , a contradiction. Thus, we can find a vector such that . ∎
Lemma 1.10.
The vector space given in (1.5) is perpendicular to in . Moreover, does not belong to .
Proof.
As in the proof of Proposition 1.9, we denote by the subspace of spanned by . Using the notation (1.3), for each and , we have
at , where we used the fact . By this computation, we have that
(1.6) |
proving the first assertion.
We prove the second assertion. If , then we can write at . Then it holds that
a contradiction. ∎
We next prepare the following:
Proposition 1.11.
Let be a null -wave front and its associated -normalized normal vector field. Suppose that is a singular point of . Then
-
(1)
for each ,
(1.7) is a null wave front defined on , and
-
(2)
there exists a positive number such that is an immersion at .
Remark 1.12.
Later, we will show that the image of coincides with that of under a suitable assumption of (see Corollary 2.8).
Proof.
Since is a light-like vector field, it is a normal vector field of . We set
for . Since () is obtained by adding the second row to the first row of , the pair gives an immersion of into if and only if is an immersion. So (1) is proved.
We next prove (2). By Proposition 1.9, there exists a vector such that . We then take a local coordinate system of centered at such that
belong to the kernel of and . We let be the subspace of , which is spanned by . We consider the -dimensional vector space given by (1.5). By Lemma 1.10, it can be easily checked that the three vector subspaces and satisfy the assumption of Lemma A.3 in Appendix A. So we have
In particular, the vectors
are linearly independent at . We fix a non-zero real number . Then we have
If is sufficiently small, the right-hand side of the last equality is equal to at . So we can conclude that is an immersion on a sufficiently small neighborhood of for sufficiently small . ∎
Definition 1.13 (-normalized null vector field).
Let be a null -wave front and its associated -normalized normal vector field. If there exists a -differentiable vector field defined on such that , then is called the -normalized null vector field with respect to (in fact, is uniquely determined as follows).
The following is the deepest result in this section, which asserts the existence of -normalized null vector field:
Theorem 1.14.
Let be a null -wave front and its associated -normalized normal vector field. Then there exists a unique -differentiable vector field defined on such that
-
(1)
that is, is the -normalized null vector field, and
-
(2)
the image of each integral curve of is a part of a light-like line in .
Moreover, is a light-like vector space in .
Proof.
Roughly speaking, the key of the proof is that the image of the null wave front is foliated by light-like lines as mentioned in the introduction, and we will construct a vector field defined on so that each integral curve of corresponds to the leaf of the foliation as follows.
We fix a point arbitrarily. By Proposition 1.11, for a sufficiently small positive number , there exists a neighborhood of such that is an immersion on . Hence, by Lemma 1.8, there exists a unique -differentiable null vector field satisfying on . We let be the integral curve of such that for . Since is a null vector field along , [2, Fact 2.6] implies that parametrizes a segment of a light-like line. Thus is a geodesic in , and so we have where is the Levi-Civita connection of . So,
holds at . Since the light-like vector () lies in , the vector space is light-like. Since is arbitrary, the uniqueness of implies it can be defined on , proving the assertion. ∎
2. A fundamental theorem for -complete null wave fronts
We first define the following “-completeness” of null wave fronts in , which is analogous to the case of null immersions given in [2, Definition 2.7].
Definition 2.1.
Let be a (co-orientable) null -wave front. The map is called -complete if for each , there exists an integral curve of the -normalized null vector field passing through such that coincides with an entire light-like line in .
The following assertion holds:
Proposition 2.2.
Let be a co-orientable null -wave front. Then is -complete if and only if its -normalized null vector field is complete on that is, each integral curve of is defined on .
Proof.
Let be any integral curve of . Since , is parametrized by an affine parametrization of a light-like line. So is defined on if and only if the image of coincides with the entirety of a line. ∎
We consider the height function
(2.1) |
with respect to the time axis, where . The level set is a space-like hyperplane which is isometric to the Euclidean -space . So, we frequently use the identification
(2.2) |
Let be an -manifold. A -map is called a (co-orientable) wave front if there exists a unit normal vector field along defined on such that the map is an immersion, where is the unit sphere centered at the origin in . We now show that the following representation formula of -complete null wave fronts in from a given wave front in as follows:
Theorem 2.3.
Let be a co-orientable -wave front with unit normal vector field . Then for each choice of , the map defined by
(2.3) |
gives an -complete null wave front in , where Moreover, the regular set of is dense in .
When 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 by a hyperplane () is congruent to the image of a parallel hypersurface of .
Remark 2.4.
Since
the image of is congruent to that of in . So we call the normal form of the null wave front associated with .
Proof of Theorem 2.3.
Without loss of generality, we may assume that , and we set
Since () is orthogonal to with respect to the Lorentzian inner product , the vector field is a null vector field of defined on . We write
In the following discussions, we think that and take values in column vectors. Then we have
We fix arbitrarily, and take a local coordinate system of centered at . Then gives a local coordinate system of . To show that is an immersion, it is sufficient to show that matrix
is of rank . Since
(2.4) | ||||
holds, we have
(2.5) |
Since is a wave front, we have
which implies that is of rank , that is, is a null wave front by Proposition 1.7. Since we have already shown that points in the null direction, is a null wave front. By definition, it is obvious that is -complete.
We next show that is an immersion on an open dense subset of . By (2.4), the matrix is of rank at if
To prove that this holds for almost all , one can use the fact that the parallel hypersurface (for fixed ) has a singular point at if and only if coincides with one of the inverse of principal curvatures of at : If a given point is a singular point of , then is a regular point of for . In particular, is an accumulation point of the regular points of , and we can conclude that the regular set of is dense in . ∎
Lemma 2.5.
In the setting of Theorem 2.3, if
is an embedding, then by setting the map
is also an embedding, where .
Proof.
We remark that () is a map into . Since is a null wave front, is an immersion (cf. Proposition 1.7). So it is sufficient to prove that is a homeomorphism from to . For the sake of simplicity, we set .
We let be a point in . Then there exists a unique point at which the straight line meets the hyperplane in . We let
(2.6) |
be the canonical orthogonal projection. Since depends on and continuously, the map is a continuous map whose image coincides with . So
is well-defined, and gives the inverse map of . So is an embedding. ∎
Conversely, we can prove the following:
Theorem 2.6.
Let be a co-orientable -complete null -wave front, then there exists a co-orientable -wave front and a diffeomorphism such that coincides with the null wave front as in Theorem 2.3.
Proof.
Let be the -normalized null vector field of . By Theorem 1.14, the image of each integral curve of is a light-like geodesic. So the identity
(2.7) |
holds on , where and is the Levi-Civita connection of .
Since and points in the light-like direction, we have , in particular, by applying the implicit function theorem, the zero-level set
is an embedded hypersurface of . Let be the one parameter group of transformations on associated with . As in [2, Proposition A.3], the map defined by
(2.8) |
is an immersion. Since (cf. (2.7)) is a light-like geodesic in , we obtain the expression
(2.9) |
In particular, is an injection, and so,
is also an injection. We suppose that is not injective and holds. Since , the point lies on the integral curve passing through . Since is a straight line, meets exactly once. So we can conclude and . Thus is an injection. On the other hand, since is -complete, any integral curve of must meet . So is bijective and then it becomes a diffeomorphism. If we set (), then we have
(2.10) |
Since
(2.11) |
we have which implies that is a unit normal vector field of in the space-like hyperplane . We fix arbitrarily, then we can write where . Since , we have
(2.12) |
where the dot denotes the canonical Euclidean inner product on , where . So is a unit normal vector field of .
Finally, we show that is a wave front defined on : Since is a null wave front, the pair is an immersion on into . Then its restriction
is also an immersion. Moreover, since the map
is an immersion (where ), and so is a co-orientable wave front. Summarizing the above discussions, (2.9) can be written as
By setting , we obtain the assertion. ∎
Corollary 2.7.
The regular set of an -complete null -wave front is dense in .
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 be an -complete null -wave front. Then, for each the parallel hypersurface given in (1.7) is also an -complete null wave front and has the same image as .
Proof.
By Theorem 2.6, there exists a co-orientable -wave front and a diffeomorphism such that holds on . We let be the -normalized normal vector field along . Then is the -normalized normal vector field of . So we have that
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 -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 -complete null wave front.


Here is one example to illustrate the structure theorem: We consider the -tubular neighborhood of the logarithmic spiral (see Figure 2, left)
in the -plane, and consider the image with respect to the graph . Then is an open subset of the light-cone (see Figure 2, right). Since is a ruled surface, we can extend each of the ruling light-like lines to both sides, and obtain an -complete null wave front as an -completion of . However, this surface is different from the light-cone . In fact, each light-like line as a generator of the ruled surface meets several times and can be regarded as a kind of double covering the light-cone . To produce the actual from the extension of , we need to consider the quotient space with an appropriate equivalence relation in the image of . In this example, the resulting -completion is the light-cone , which is, of course, a Hausdorff space. However, in the general situation, in order for the domain of definition of the resulting -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 -wave front which may not be -complete. Let
(3.1) |
be the restriction of the height function to the hypersurface . he following lemma will play an important role:
Lemma 3.1.
Let be a co-orientable null -wave front, and let be a point in . Then, for each open neighborhood of , there exist
-
•
a neighborhood of ,
-
•
a positive number and a real number ,
-
•
a -differentiable -submanifold of ,
-
•
a surjective -submersion , and
-
•
a -wave front
such that
-
(1)
We set and denote by . Then gives a unit normal vector field of in by regarding is a map into .
-
(2)
We set
(3.2) Then the map given by
is a diffeomorphism between and satisfying
(3.3) where if and if .
-
(3)
The map given by
is an embedding satisfying
-
(4)
The map defined by
is an embedding.
In this setting, if itself is -complete and , then the above assertions hold by setting .
Definition 3.2.
In this setting, if is an embedding, is called an -adapted neighborhood of . Moreover, is called the fundamental -data at .
Proof of Lemma 3.1.
Let be the -normalized null vector field of . Since points in the light-like future direction, we have . In particular, by the implicit function theorem, the level set
is an embedded hypersurface of . We set
(3.4) |
Since is transversal to the hypersurface , there exist
-
•
an open interval of the form () or ,
-
•
a connected open submanifold of satisfying , and
-
•
an injective -immersion
such that (cf. [2, Proposition A.3]) the map () gives an integral curve of satisfying . In this setting, if is -complete and , then is a complete vector field of and we can set .
Since is an injective immersion between manifolds of the same dimension, by the inverse function theorem, is an open map and so
(3.5) |
is a neighborhood of , and gives a diffeomorphism from to . Moreover, if is -complete and , the map gives a diffeomorphism between and by the same reason why is a diffeomorphism in the proof of Theorem 2.6.
We define a map by
where is the canonical projection of onto . Since for (cf. (3.4)), is a surjective submersion satisfying
(3.6) |
By Theorem 1.14, (, ) lies on a light-like straight line, and so, the vector does not depend on the parameter . So we can write
where (since points in the future direction). We regard is a Euclidean -space. By the same argument as in the proof of Theorem 2.6 (cf. (2.10), (2.11) and (2.12)), we have and can write
(3.7) |
such that gives a unit normal vector field of the map
So is a frontal in . Moreover, since () parametrizes a light-like straight line, we have that
(3.8) |
We next show that is a wave front in the hyperplane . Since is a null wave front in , the pair is an immersion from into (cf. Lemma 1.6). By (3.8), we can write
By setting (), the rank of the matrix (2.5) for satisfies
Since the right-hand side is equal to , the matrix
is of rank at each point of . So is a wave front on .
We consider the map given by
which corresponds to the parallel hypersurface of having signed equi-distance in . Then is a wave front in whose unit normal vector field satisfies
Then, by Theorem 2.3,
(3.9) |
is an -complete null wave front. We set
and consider the following diffeomorphism
By (3.6), it holds that
(3.10) |
By this with (3.8) and (3.9), we have
for each . Thus, by setting
we obtain the desired fundamental data: In fact, if we set
then the two maps and are immersions, since we have already shown that and are wave fronts.
By Lemma 3.1, for each point , there exist fundamental -data
such that is an -adapted neighborhood of giving an -complete null wave front (cf. (3.2))
and an embedding
(3.11) |
where . Since satisfies the second axiom of the countability, there exist an at most countable set and a family of fundamental -data
such that
-
(1)
for each , there exists satisfying
- (2)
-
(3)
.
In particular, is a family of embeddings defined on -dimensional manifolds. To ensure that the domain of definition of the resulting is a Hausdorff space, we give the following definition:
Definition 3.3.
In the above setting, is said to be admissible if 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 be an admissible null -wave front if is real analytic, i.e. , it is admissible, see Lemma B.5. Then there exist
-
•
a -differentiable -manifold ,
-
•
a -immersion ,
-
•
an -complete null wave front written in the normal form and its canonical lift
such that
If is an embedding see Remark 3.5, then gives a diffeomorphism between and . Moreover, if is -complete, then is a surjection.
Remark 3.5.
Since is a null wave front, its canonical lift is an immersion. So, the assumption that is an embedding is not so restrictive. On the other hand, even if is an embedding, the admissibility of does not hold in general, see Example 3.7.
Proof.
We consider the disjoint union and give a relation for so that implies that there exists a pair of indices such that
-
•
is a neighborhood of ,
-
•
is a neighborhood of , and
-
•
is -related to in the sense of Definition B.1.
As seen in the proof of Proposition B.7, the symbol gives an equivalence relation, and is an -manifold. Then is the canonical projection as an open map. We set
Then is the differentiable structure of as shown in the proof of Proposition B.7. Moreover, an immersion is induced which satisfies
We can write
Then, by definition, we have
So, if we set
then it holds that
(3.12) | ||||
for and . So, if we consider the maps defined by
then we have
We set
Since is an embedding (see (3) of Lemma 3.1), then the restriction is also an embedding for each . In particular, is an -complete null wave front on .
We then fix arbitrarily, and assume that belongs to for some . By (3.3) and (3.12), we have
So it holds that
(3.13) |
Since and are immersions, the map
(3.14) |
is also an immersion.
If for some other index , then the immersion is also induced. Since and are embeddings, so is . Thus coincides with on , and the map is canonically induced. By (3.13) and (3.14),
holds. In particular, also holds on . We now consider the case that is an embedding. Suppose that (). Then we have
Since is an embedding, we have , proving the injectivity of . Since is an immersion between same dimensional manifolds, it is a diffeomorphism.
If is -complete, then and for each , which imply that the map is surjective (cf. (3.14)). So we can conclude that is a surjection. ∎
Example 3.6.
We set , and consider a null immersion
whose image lies on the light-cone passing through the origin in . In this case, the image of the -completion of is the light-cone and so the map covers twice on in . In particular, the induced map is not a diffeomorphism. This example shows that if we drop the condition that is injective, the injectivity of does not follow, in general.

Example 3.7.
Consider a plane curve (cf. Figure 3)
where is a -function into such that for and for , where is a sufficiently small positive number. Then generates the -complete null wave front (cf. Theorem 2.3). By the construction of , is not admissible, since the image of has non-transversal intersections. We then set
where is a unit normal vector field of . The image of is an embedded spiral-shaped curve on image of . Consider a sufficiently small tubular neighborhood of in the image of . Then is an embedded null surface, because has no self-intersections. If we think that is the inclusion map , then the -completion of is . So, this example shows that the admissibility of the -completion of in Theorem 3.4 is independent of the embeddedness of the canonical lift .
Moreover, if we apply our -completion procedure for , 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 -wave front is said to be complete if
-
•
is -complete, and
-
•
the singular set of is a non-empty compact subset of .
Remark 4.2.
As proved in [2], if is -complete and its singular set is empty, then the image is a subset of a light-like hyperplane of , which is a trivial case. So we only consider the case that the singular set of is non-empty, in the above definition.
Remark 4.3.
We prove the following:
Theorem 4.4.
Let be a null -wave front in . If is complete, then it can be reparametrized as a normal form , where is an immersion defined on a compact -manifold whose principal curvatures are non-zero and all same sign everywhere in . In particular, if , then is a compact convex hypersurface in . On the other hand, if , then is a closed locally convex regular curve in .
Proof.
We let be a complete null wave front. By Theorem 2.6, we may assume that there exist
-
•
an -manifold ,
-
•
a diffeomorphism , and
-
•
a co-orientable wave front
such that coincides with the normal form given in Theorem 2.3. We consider the height function given in (2.1). Since the singular set of is compact, for sufficiently large , the restriction
is an immersion such that . Without loss of generality, we may assume that and satisfies
(4.1) |
where can be considered as the unit normal vector field along the immersion . So, we may assume that is defined on . Since is co-orientable and is an immersion, is orientable. We let
be principal curvature functions of . Here, are eigenvalues of a symmetric matrix associated with the shape operator of at . 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 can be considered as a family of real-valued continuous functions defined on . We first show that each never changes sign: Suppose that there exists a sequence of points which converges to a point such that
Then
are singular points of which are unbounded on , contradicting the compactness of the singular set of . Thus, each as a continuous function on has no zeros unless it is identically zero. By Remark 4.2, is not a part of a hyperplane in . So, there exists an integer () such that are not identically zero, where is a subset of . By Hartman’s product theorem (cf. [8, Page 347]), is a product of a compact manifold and (). For each , we can define a continuous map by
Then coincides with the singular set of . Since is complete null wave front, is compact. Since the projection of via the continuous map
is compact, the hypersurface is also compact. Thus (i.e. ) and each () is either positive-valued or negative-valued on . Moreover, since is compact, there exists a point such that attains the farthest point of the image of from the origin in . Then are all positive or all negative at the same time.
Here, replacing by , we may assume that
Then, take the same sign on . Thus, Hadamard’s theorem [5] implies that is an embedded convex hypersurface in whenever . ∎
If , then is a locally strictly convex regular curve in . So we may express by defined as an -periodic curve (), parametrized by the arc-length. Then its curvature function can be taken so that is positive everywhere. Then we may assume that is expressed as
where is a parallel curve of , and the singular set of is
and
(4.2) |
just parametrizes the singular set image of . We can prove the following:
Proposition 4.5.
Let be a complete null wave front in which is generated by a closed locally convex regular curve in . Then non-cuspidal edge points of correspond to vertices on , where a vertex is a critical point of the curvature function of .
Proof.
The following assertion is equivalent to the classical four vertex theorem for convex plane curve:
Corollary 4.6.
Let be a complete null wave front in which has no self-intersections outside of a compact subset of . Then has at least four non-cuspidal edge singular points.
Proof.
Since has no self-intersections outside of a compact set, must be generated by a (closed) convex plane curve in the -plane. By the classical four vertex theorem, has at least four vertices, then such points corresponds to the non-cuspidal edge points of as seen in the proof of Proposition 4.5. So we obtain the assertion. ∎
Since a complete null wave front in can be considered as a complete flat front in the Euclidean -space, Corollary 4.6 is a special case of the four non-cuspidal edge point theorem given in [9, Theorem D].
Example 4.7.

Example 4.8.
Consider a locally convex plane curve with a self-intersection
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
We denote by the canonical Lorentzian inner product of . In this section, we prepare the following assertion for vector subspaces in .
Definition A.1.
A subspace of is said to be perpendicular to a subspace if holds.
For a subspace of , we set
(A.1) |
which is the maximal subspace perpendicular to in .
Definition A.2.
A subspace of is said to be degenerate if there exists a non-zero vector such that for all . We call such a vector a degenerate vector of . On the other hand, a subspace is said to be non-degenerate if it is not degenerate.
Lemma A.3.
Let and be three subspaces of such that
-
(1)
is a 1-dimensional degenerate subspace satisfying ,
-
(2)
and are perpendicular to , and
-
(3)
is perpendicular to .
Then .
Proof.
If is non-degenerate, then so is , and using (3), we have
which implies . So we may assume that is degenerate. By definition, there exists a degenerate vector belonging to . We let be a generator of the 1-dimensional vector . Then and are both degenerate vectors. In particular, they are light-like, that is, hold. So if we set , then by (1), contains two linearly independent light-like vectors and , and so, is a time-like vector space (cf. [10, Lemma 27, Page 141]), which is non-degenerate. So, is also non-degenerate. Since (cf. (2) and (3)) and , we have
proving . ∎
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 (). Let be two connected -dimensional manifolds. We let and be two -embeddings. A point is said to be -related to if there exist an open neighborhood of and an open neighborhood of such that
-
•
and
-
•
.
If there are no -related points on and , we say that “ is not -related to ”.
Remark B.2.
This “-relatedness” is an open condition. In fact, if is -related to , then there exist open neighborhoods and of and respectively such that each point of is -related to a certain point in .
In this setting, the following assertion holds.
Lemma B.3.
If we set
then resp. is an open subset of resp. . Moreover, if is non-empty, then there exists a unique -diffeomorphism satisfying on .
Proof.
Assume that (resp. ) is -related to (resp. ). Since and are embeddings, (resp. ) is uniquely determined, and so the map is also uniquely determined. The smoothness of is obvious. ∎
Definition B.4.
Let and be as in Definition B.1. Then the pair is said to be -admissible if, for each pair , it holds that
-
(i)
is -related to , or
-
(ii)
there exist a neighborhood of and a neighborhood of such that is not -related to .
By definition, if does not meet , then the pair is -admissible. The following assertion is immediately from the definition:
Lemma B.5.
In the setting of Definition B.4, the pair is -admissible if meets transversally or and are both real analytic.
Proof.
We fix a pair such that is not -related to . If meets transversally, then the assertion is obvious. So we may assume that and are both real analytic and does not meet transversally. Then and we can take -dimensional affine plane as a common tangential space of and at in . Since (resp. ) is locally connected, there exists a connected open neighborhood (resp. ) of (resp. ) such that (resp. ). By the implicit function theorem, we may assume that the images of the maps and are expressed as the graphs of certain functions defined on the same domain in .
It suffices to show that there exist a neighborhood of and a neighborhood of such that is not -related to . If not, there exists a pair such that is -related to , which implies that there exist open neighborhoods and of and , respectively, such that each point of is -related to a corresponding point in . Then the vector-valued function coincides with on some non-empty open subset of . By the connectedness of and the real analyticity of and , the two vector-valued functions coincide identically on , which implies that is -related to , a contradiction. ∎
Definition B.6.
Let be an at most countable set, and let be a family consisting of connected -dimensional manifolds and embeddings . Then is called admissible if is -admissible for any choice of .
We let be an admissible family as in Definition B.6. Consider the disjoint union and define a relation for so that implies that there exist a neighborhood () of and a neighborhood () of such that is -related to . Then it is easy to check that is an equivalence relation. Moreover, the following assertion holds:
Proposition B.7.
Let be an admissible family. Then there exist
-
•
a manifold ,
-
•
a -immersion , and
-
•
a diffeomorphism
such that is a differentiable structure of and coincides with on for each .
Proof.
Since is a disjoint union of open subsets, this is uniquely determined by . So we denote by . As we have already noted, is an equivalence relation, and the canonical projection is induced. Since is an at most countable set, satisfies the second axiom of countability. If we show that the quotient space is a Hausdorff space, then we can easy observe that has a structure of -manifold by using Lemma B.3, and we can construct the desired -immersion by setting
(B.1) |
So it is sufficient to prove that is a Hausdorff space: Consider the set defined by
By Remark B.2, the canonical projection is an open map. So we prove that is a closed subset of (cf. [6, Chapter 3, Theorem 11]). We consider a sequence in and suppose that and converge to and in , respectively. By definition, there exist and () and a positive integer such that
-
•
and ,
-
•
and for .
We remark that the admissibility of implies the admissibility between and . Since and , we have . We suppose . Then, by (ii) of Definition B.4, there exist a neighborhood of and a neighborhood of such that is not -related to . However, this contradicts the fact that . So 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, -singularities of wave fronts, Math. Proc. Camb. Phil. Soc. 146 (2009) 731–745.