Boundedness of geometric invariants near a singularity which is a suspension of a singular curve
Abstract.
Near a singular point of a surface or a curve, geometric invariants diverge in general, and the orders of diverge, in particular the boundedness about these invariants represent geometry of the surface and the curve. In this paper, we study boundedness and orders of several geometric invariants near a singular point of a surface which is a suspension of a singular curve in the plane and those of curves passing through the singular point. We evaluates the orders of Gaussian and mean curvatures and them of geodesic, normal curvatures and geodesic torsion for the curve.
Key words and phrases:
cuspidal edge; geodesic curvature; normal curvature; geodesic torsion2020 Mathematics Subject Classification:
Primary 57R45; Secondary 58K051. Introduction
In this paper, we study boundedness of several geometric invariants near a singular point of a surface which is a suspension of a singular curve in the plane. More precisely, let be an -equivalence class of singular plane curve-germs. A -edge is a map-germ such that it is -equivalent to , where is a representative of , namely, a one-dimensional suspension of . Here, two map-germs are -equivalent if there exist diffeomorphisms and such that . A cuspidal edge (-equivalent to the germ at the origin) and a -cuspidal edge (-equivalent to the germ ) are examples of -edges, and are a -cusp and -cusp respectively. If is of finite multiplicity, then the -edge is a frontal. A frontal is a class of surfaces with singular points, and it is well known that surfaces with constant curvature are frequently in this class. In these decades, there are several studies of frontals from the viewpoint of differential geometry and various geometric invariants at singular points are introduced (for instance [3, 5, 6, 7, 11, 12, 14]). If a surface is invariant under a group action on , then -edges will appear naturally. Singularities appearing on surfaces of revolution and a helicoidal surface are examples of such surfaces [13, 16]. Moreover, such singularities appear on the dual surface at cone like singular points of a constant mean curvature one surface in the de Sitter -space (see [8]).
In this paper, we study geometry of -edges. For this, we consider two classes of singular map-germs, that we shall call -type and -type edges, the first including -type edges and also -edges when has finite multiplicity (see Section 2). One observes that -type edges are frontals. In order to proceed our study we find a normal form for each one of these map-germs preserving the geometry of the initial map, since we only use isometries in the target (Proposition 2.9). In [11, 14] the authors define singular, normal and cuspidal curvatures, as well as cuspidal torsion for frontals. In an analogous way, we define similar geometric invariants for -type edges, getting same names, except for the cuspidal curvature, which we called -cuspidal curvature. These invariants are related with the coefficients of the normal form given in Proposition 2.9. It is worth mention that these cuspidal curvatures are similar. In fact, we know that a frontal-germ is a front if and only if the cuspidal curvature is not zero. We conclude from Proposition 2.11 that a -type edge is a front if and only if the -cuspidal curvature is non zero at 0. In particular, we study orders of geometric invariants and geometric invariants of curves passing through the singular point. We evaluates the orders of Gaussian and mean curvatures (Theorem 2.17) and the minimum orders of geodesic, normal curvatures and geodesic torsion for a singular curve passing through the singular point (Theorem 3.5). These minimum orders are written in terms of singular, cuspidal and normal curvatures and the cuspidal torsion. As a corollary, we give the boundedness of these curvatures under certain generic conditions (Corollary 3.6).
2. Geometry of -edges
We give several classes similar to -edges. It includes -edges, and these classes will be useful to treat. We recall that a map-germ is a frontal if there exists a unit vector field along such that holds at any and any , where is the canonical inner product of . The vector field is called a unit normal vector field of . A map-germ is an -type edge if it is -equivalent to for a function . A map-germ is a -type edge () if it is -equivalent to , where . This is equivalent to being -equivalent to . Two map-germs are -equivalent if their -jets at the origin are -equivalent. We show:
Lemma 2.1.
Let be a -edge respectively, an -type edge, an -type edge. Then an intersection curve of with a surface which is transversal to passing through near is -equivalent to respectively, -equivalent to , -equivalent to .
Proof.
Since the assumption and the assertion do not depend on the choice of the coordinate systems, we can assume is given by , where is -equivalent to . Then can be represented by the graph in as the -space, and the intersection curve is . Since is transverse to the -axis, the orthogonal projection of to the -plane is a diffeomorphism, we see the assertion. One can show the other claims by the similar way. ∎
2.1. A sufficient condition
We give a sufficient condition of a frontal-germ being an or -type edge under the assumption . We assume throughout this subsection. Let be a frontal-germ satisfying . Then there exists a vector field such that generates if . We call a null vector field, and an extended null vector field. An extended null vector field is also called a null vector field if it does not induce a confusion. We assume that the set of singular points is a regular curve, and the tangent direction of is not in . Let be a vector field such that is a non-zero tangent vector of for . We consider the following conditions for :
-
[2.1]
on ,
-
[2.2]
on ,
-
[2.3]
on ,
-
[2.4]
at .
Here, for a vector field and a map , the symbol stands for the -times directional derivative of by . Moreover, for a coordinate system and a map , the symbol stands for .
Proposition 2.2.
Let be a frontal-germ satisfying . Assume that the set of singular points is a regular curve, and the tangent direction of generated by is not in . If there exists a null vector field satisfying [2.1], and satisfies [2.2], then is an -type edge. Moreover, if also satisfies [2.3]–[2.4], then is an -type edge.
As we will see, the conditions [2.2]–[2.4] does not depend on the choice of null vector field satisfying [2.1]. To show this fact, we show several lemmas which we shall need later. Firstly we show that the conditions does not depend on the choice of the diffeomorphism on the target. In what follows in this section, is as in Proposition 2.2.
Lemma 2.3.
Let be a diffeomorphism-germ on , and set . If and satisfy the condition , then and satisfy , where , , – and –.
Proof.
It is clear that the conditions [2.2]–[2.4] do not depend on the choice of , i.e., non-zero functional multiple and extension other than . Moreover, it does not depend on the non-zero functional multiple of :
Lemma 2.4.
Let be a non-zero function. If and satisfy the condition , then and satisfy , where and is the same as those in Lemma 2.3.
Proof.
Since is a linear combination of , and the coefficient of is , we see the assertion. ∎
A coordinate system satisfying , is said to be adapted.
Lemma 2.5.
Let be a frontal-germ satisfying . We assume that the set of singular points is a regular curve. For any null vector field , there exists an adapted coordinate system such that for any .
In this lemma, we do not assume that is an -type edges.
Proof.
Since , one can easily see that there exists a coordinate system such that for any . Since is a regular curve, and the tangent direction of it is not in , can be parametrized as . Define a new coordinate system by and . Then and hold. This shows the assertion. ∎
Lemma 2.6.
If two null vector fields satisfy [2.1], and satisfies , then also satisfies . Here, is the collection of the conditions: , , –.
Proof.
Let us assume that and satisfy [2.1]. Since the assumption [2.1] and the assertion do not depend on the choice of the coordinate system on the source by Lemma 2.5, we take an adapted coordinate system with for any and . Since on the -axis, has the form . If the pair satisfies [2.2], then on the -axis. On the other hand, any null vector field is written as , . By Lemma 2.4, dividing this by , we may assume an extended null vector field is
Since it holds that on the -axis (when ) and , we have . Continuing this argument, we may assume
(2.2) |
Thus holds, and holds on the -axis. Therefore satisfies [2.2]. We assume that the pair satisfies [2.1]-[2.3] and does [2.1], [2.2]. By this assumption, , . By the form of , it holds that on the -axis. Since on the -axis, on the -axis. Thus on the -axis. Similarly, on the -axis, on the -axis. Thus if , then since , we have on the -axis. Thereby we have () on the -axis. The last assertion can be shown by the same calculation. ∎
Proof of Proposition 2.2.
We assume satisfies the condition of the proposition, and satisfies the conditions [2.1] and [2.2]. Then we take an adapted coordinate system such that . By the proof of Lemma 2.6, there exist and such that , and , where . We set . Then has the form . By a coordinate change on the target, has the form , where . Rewriting the notation, we may assume is written as
On this coordinate system, satisfies the condition [2.1], it satisfies [2.2] by Lemma 2.6. This implies . So we assume . We set , . Rewriting the notation, we may assume is written as . By a coordinate change on the target, we may assume is written as . This proves the first assertion. We assume that also satisfies [2.3] and [2.4]. We may assume is written as . By Lemma 2.6, we may assume that satisfies [2.3] and [2.4]. By [2.3], the function satisfies on the -axis. Thus is written as . By [2.4], it holds that , and hence the assertion is proved. ∎
By the proof of Lemma 2.6, we have the following property:
2.2. Normal form of or -type edge
Given a curve-germ , if there exists such that (), then at is said to be of finite multiplicity, and such an is called the multiplicity or the order of at . Moreover, if there exists ( and , ) such that is -equivalent to , then is called of -type. This is well-defined since if is -equivalent to then it is not -equivalent to for , . We simplify a curve-germ of -type and an -type edge by coordinate changes on the source and by special orthonormal matrices on the target. Let be the ordinary coordinate system of . A coordinate system is positive if the determinant of the Jacobi matrix of is positive. We have the following results.
Lemma 2.8.
Let be a curve germ satisfying , and . Then there exist a parameter and a special orthonormal matrix on such that
(2.3) |
Let be a curve germ of -type. Then there exist a parameter and a special orthonormal matrix on such that
(2.4) |
where is the greatest integer less than in our convention, is not an integer.
Proof.
One can easily see the first assertion. We assume that is a curve germ of -type, then we may assume . If has a term , then is not -equivalent to . This proves the assertion. ∎
Proposition 2.9.
Let be an -type edge. Then there exist a positive coordinate system and a special orthonormal matrix on such that
(2.5) |
Moreover, if is an -type edge, then there exist a positive coordinate system and a special orthonormal matrix on such that
(2.6) |
.
Proof.
By the proof of Proposition 2.2, we may assume
By that proof again, . By a rotation on , we may assume and . By a coordinate change , we may assume , . This proves the first assertion. If is an -type edge, then the function can be expanded by
Since is an -type edge, the curve is of -type for any near . This implies that . By , . This proves the assertion. ∎
Each form (2.5) and (2.6) is called the normal form of an -type edge and an -type edge, respectively. Looking the first and the second components in (2.5) and (2.6), we remark that the -jet of the coordinate system which gives the normal form is uniquely determined up to when is even. Let be an -type edge and a null vector field which satisfies the condition [2.1]. Then the subspace and the subspace spanned by , do not depend on the choice of . We assume that the representation of -space satisfies that is the -axis and is the -plane. Then the coordinate system gives the normal form (2.5) if and only if and is identically zero.
2.3. Geometric invariants
2.3.1. Cuspidal curvatures
Let be an -type edge. A pair of vector fields is said to be adapted if is tangent to , and is a null vector field. We take an adapted pair of vector fields such that satisfies the condition [2.1], and is positively oriented. One can show the existence of such a pair by the definition of -type edge. We define
where is a parametrization of . We call the -cuspidal curvature. We have the following lemma:
Proposition 2.10.
The function does not depend on the choice of satisfying the condition [2.1].
Proof.
Since it is not appeared in the formula, does not depend on the choice of the coordinate system. Let be an adapted pair of vector fields satisfying the condition [2.1]. It is clear that the function does not depend on the choice of . We take an adapted coordinate system satisfying . Then
By Corollary 2.7, we have . Let be another null vector field satisfying the condition [2.1]. We see that does not depend on the non-zero functional multiples of , we may assume . By the proof of Lemma 2.6, we may assume that is
(2.7) |
Then by ,
Thus
(2.8) |
where is a function, and
hold on the -axis. Since and , we have
This shows the assertion. ∎
We have the following proposition.
Proposition 2.11.
Let be an -type edge. Then at is an -type edge if and only if at .
Proof.
It is easy to show that -type edges are fronts and that an -type edge is a front if and only if . In Appendix A, we define -cuspidal curvature for a curve germ of -type, denoting it by . An intersection curve of -type edge with as in Lemma 2.1, is a curve-germ of -type. The following holds.
Corollary 2.12.
Let be a -edge, where is -equivalent to . Then the -cuspidal curvature at coincides with the -cuspidal curvature of the intersection curve of with a plane which is perpendicular to the tangent line to at .
Proof.
Let be an -type edge. We assume is identically zero on . Let be a parametrization of . We define
We will see this does not depend on the choice of which satisfies the conditions [2.1] and [2.2] in Proposition 2.2 and . Inductively, we define when by
We will also see this does not depend on the choice of satisfying the conditions [2.1] and [2.2] in Proposition 2.2 and . If , we set .
Proposition 2.13.
Proof.
We already showed the case in Proposition 2.10. Let be a pair of vector fields satisfying the assumption of lemma. We take an adapted coordinate system such that . By the proof of Lemma 2.6, we see .
Moreover, we have:
Lemma 2.14.
There exist functions , and a vector valued function such that
(2.9) |
Proof.
Since on the -axis, implies that there exists such that . We assume that there exist such that . Differentiating this equation, we have
Thus by holds, and on the -axis, we have on the -axis. Hence there exist functions , and a vector valued function such that . This shows the assertion. ∎
We continue the proof of Proposition 2.13. Since the assertion holds by multiplying the null vector field by a non-zero function, we take a null vector field as in the right-hand side of (2.7). By the same calculations in the proof of Proposition 2.10, we have . Thus
Since if and ,
Thus , where is a function. By holds, and on the -axis by Lemma 2.14, we have
and this shows the assertion. ∎
We call the -cuspidal curvature, and the -bias. Note that does not depend on the choice of satisfying [2.1], [2.2] and at by the same calculation. In this case, by the additional assumption. If is an -type edge, and written as (2.5), then . If is an -edge , and written as (2.6), then , and . See Appendix A for geometric meanings of the terms .
2.3.2. Singular, normal curvatures and cuspidal torsion
Let be an -type edge, and be a parametrization of the singular set. Let be a unit normal vector field of , and we set for an oriented coordinate system on . We set . Then we define
(2.10) |
and
(2.11) |
where if agrees the orientation of the coordinate system, and if does not agree the orientation. We call , and singular curvature, normal curvature and cuspidal torsion, respectively. These definitions are direct analogies of [14, 11]. It is easy to see that the definitions (2.10) do not depend on the choice of parametrization of the singular curve. Moreover, does not depend on the choice of , nor the choice of when is even. To see the well-definedness of , we need:
Proposition 2.15.
Proof.
One can easily to check it does not depend on the choice of functional multiplications of . Since the assertion does not depend on the choice of local coordinate system, one can choose an adapted coordinate system with satisfying [2.1]. Let be a null vector field which satisfies [2.1]. Then by the proof of Lemma 2.6, we may assume is given by (2.2). Then by (2.8), we see on the -axis, where is given in the proof of Lemma 2.6. Furthermore, by (2.8), we see and on the -axis. Substituting these formulas into the right-hand side of (2.11), we see it is
and by , this shows the assertion. ∎
If an -type edge is given by the form (2.5), then , and .
2.4. Boundedness of Gaussian curvature and mean curvature near an -type edge
Here we study the behavior of the Gaussian and mean curvatures.
Let be a function-germ (. If there exists an integer such that and , then is said to be of order , where is the unique maximal ideal of the local ring of function-germs and denotes the th power of (cf. [9, p. 46]). If , then the order of is . The order of is denoted by . If is of order , then is said to be of finite order. Let be two function-germs such that is of finite order. The rational order of a function , where is
For a function , we define , where . If , then we define . If , then is called rationally bounded, and , then is called rationally continuous ([12, Definition 3.4]). If , this is the usual one.
Since the property does not depend on the choice of coordinate system, the order and the rational order does not depend on the choice of coordinate system.
Let be an -type edge, and let be an adapted coordinate system with satisfying [2.1]. We take in Corollary 2.7. Namely, here we set by . Since is an -type edge, and is linearly independent (Proposition 2.2 and the independence of the condition [2.2]). Thus the unit normal vector of can be taken as . Using , and , we define the following functions:
We note that coefficients of the first and the second fundamental forms of -edges being of multiplicity can be written as
Lemma 2.16.
The differentials and of are written as
Proof.
Since , there exist functions on such that
Considering and , we have
Solving these equations, we have the assertion. ∎
By this lemma, can be written as
along the -axis. Since and are linearly independent and , the condition is equivalent to . To see this fact, we take the same setting in the proof of Proposition 2.10. Then we see
(2.12) |
along the -axis, where and . Since is a frame of , and , , it holds that if and only if . Moreover, since , it holds that is equivalent to . Let be an -type edge, and let us set
Theorem 2.17.
Let be an -type edge. Then the rational order of the mean curvature is . If the normal curvature does not vanish at , then the rational order of the Gaussian curvature is .
Proof.
By Theorem 2.17, the order of and coincide. Moreover since , they never bounded when the normal curvature does not vanish.
3. Curves passing through -type edges
In this section, we consider geometric invariants of a curve passing through an -type edge . If is non-singular, then the usual invariants can be defined as well as the regular case. We consider the case when has a singular point, namely, passing through a singular point of in the direction of a null vector.
3.1. Normalized curvatures of singular curves
Following [15, 4], we introduce normalized curvature on curves in . Let be a curve, and let be a singular point. We assume that there exists such that ().
We set
(3.1) |
and
(3.2) |
we see is a function and . We call this parameter an -arc-length.
Proposition 3.1.
The parameter is an -arc-length parameter of if and only if .
Proof.
Let us suppose now that is the -arc-length, i.e., , with as in (3.1). Since , thus and, consequently, . Therefore, it holds that . ∎
Let us set . Then the curvature satisfies that
(3.3) |
is a function. We call the normalized curvature. This is originally introduced in [15] and generalized in [4]. Let be a given -function, and be an integer. Then similarly to [15, Theorem 1.1], one can show that there exists a unique plane curve up to isometries in with normalized curvature given by , where is the -arc-length parameter.
Using the frame along defined by and the -rotation of , the normalized curvature can be interpreted as follows: We set the function by the equation
(3.4) |
where ′ is a differentiation by the -arc-length. Then we have:
Proposition 3.2.
Let be the above frame along in the Euclidean plane satisfying (3.4), where is the -arc-length parameter. Then holds.
Proof.
Since where and the -arc-length parameter satisfies , so and . Then
Consequently,
On the other hand, since , where and is the -counterclockwise rotation of , and the dot ‘’ denotes the canonical inner product of , it holds that:
Thus we have the assertion. ∎
3.2. Normalized curvatures on frontals
According to Section 3.1, we define the normalized curvatures for curves on a frontal. Let be a frontal and a unit normal vector field of . Let be a curve. We set . We assume there exists such that (). The geodesic curvature , the normal curvature and the geodesic torsion are defined by
(3.5) |
on regular points (see [1, page 261]). These curvatures can be unbounded near singular points. Indeed, it holds that
(3.6) |
One can easily see that
(3.7) |
are functions, where is the function given by (3.2) for . We call normalized geodesic curvature, normal curvature and geodesic torsion of , respectively. These satisfy:
Lemma 3.3.
It holds that
(3.8) | ||||
(3.9) | ||||
(3.10) |
at .
Proof.
Similar with the case of plane curves, these invariants can be interpreted as follows. Under the same assumption above, we set , and . Then is a frame along . We define by
(3.11) |
where is the differentiation by the -arc-length parameter. With the above notation, we get the following:
Proposition 3.4.
If is the -arc-length parameter, then
holds, for any .
3.3. Behaviors of and passing through an -type edge
In this section we shall study the orders of the geodesic and normal curvatures and the geodesic torsion of a curve passing through an -type edge, concluding on boundedness. Describing the condition, we use the curvature of such curve. Let be an -type edge, , and be a regular curve such that is a null vector of at . Let be a coordinate system which gives the form (2.5), and be a parametrization of where the coordinate system on the target space is , and the orientation of agrees the direction of at . Since such coordinate system is unique (unique up to if is even), the order of contact of with the -axis at and the curvature of is well-defined as a curve on . We call such order of contact the order of contact with the normalized null direction, and we call the curvature written in the normal form. If (), then the order of contact with the normalized null direction is , and and hold.
Theorem 3.5.
Let be an -type edge, , and be a regular curve with order of contact with the null direction of at and the curvature of written in the normal form of . Then, it holds that:
(1) The case .
For ,
-
•
if , then ;
-
•
if , then , and is equivalent to
-
•
if , then , and if and only if .
For , it holds that , and if and only if . For ,
-
•
if , then , and is equivalent to
-
•
if , then , and if and only if .
(2) The case .
For this case, it holds that , , and
is equivalent to
For , it holds that , and is equivalent to .
(3) The case .
In this case, it holds that , and
is equivalent to
For it holds that , and if and only if . For , it holds that , and if and only if .
If is even and be a coordinate system which gives the form (2.5), then also gives (2.5). In this case, changing to , the signs of and reverse, and them of and do not change. So, when is even, neither the condition
changes under the coordinate change to .
Proof.
Let . One can assume that is given by the form (2.5) and, since is a null vector of , one can take , with . Then is of order and we set ().
In the normal form (2.5), since , further we may assume is given by . We recall that and . Furthermore, it holds that , , and . We set by . Then gives a non-zero normal vector field to .
(1). Assume . By (2.5) we get , where
Then , where . Note that . Setting , we can show that , where
(3.12) | ||||
We abbreviate the variable, namely , for instance, and . Here, we see
To see the rational order of the invariants , at , we may use instead of in (3.6). Since , we can write . We see
(3.13) | ||||
(3.14) |
where means a smooth function depending on . Then we see , where is, for :
(3.15) |
If , then (3.15) is , where
If , then (3.15) is , where
If , then and (3.15) is , where
If , by (3.13) and (3.14), we see the assertion, once and . These shows the assertion for .
Since one can easily see that at , the assertion for is proved. Next we see is
(3.16) |
If , then (3.16) is , where
If , then and (3.16) is , where
If , then , and (3.16) is , where
This show the assertion for .
(2) and (3). We assume and we shall use the same notation of case (1). Setting , we can show that , where
and is the same as in (3.12). We assume . Then , where , and
Then , where , with . Since has multiplicity , we need replace in equations (3.6) by . Here, we see
To see the order, we may use instead of in (3.6). We see
(3.17) | ||||
By applying the formula
for , we see is
(3.18) | ||||
Then . This shows the assertion for . By (3.17), we see and . This shows the assertions for and .
Next we assume . In this case, . We set and
Here, it holds that
It holds that , with and . We see
(3.19) | ||||
By the similar method to (3.18), we see is
Then and, replacing by in equations (3.6), this shows the assertion for . By (3.17), we see , where . It holds that
and . This shows the assertions for and . ∎
In particular, we have the following corollary on boundedness directly obtained from Theorem 3.5.
Corollary 3.6.
Let be an -type edge with , and be a regular curve with order of contact with the null direction of at .
-
(1)
The case . For ,
-
•
if , then is bounded at ;
-
•
if , then is unbounded at ;
-
•
if and , then is unbounded at .
For , if , then is unbounded at . For ,
-
•
if and , then is unbounded at ;
-
•
if and , then is bounded at ;
-
•
if , then is bounded at .
-
•
-
(2)
The case . In this case, is unbounded at . If , then is bounded at . If and , then is unbounded at . If , then is unbounded at .
-
(3)
The case . In this case, and are bounded at . If , then is unbounded at .
We consider the case that is a cuspidal edge. By definition, it is a -edge, in particular, a -type edge. Then by Theorem 3.5, the following assertion holds.
Corollary 3.7.
Let be a cuspidal edge, and let be a regular curve with order of contact with the null direction of at and the curvature of written in the normal form of . Then, it holds that:
For ,
-
•
if , then , and if and only if .
-
•
if or , then ;
-
•
if , then , and is equivalent to
For , it holds that .
For ,
-
•
if or , then , and is equivalent to
-
•
if , then , and if and only if .
Proof.
About the boundedness, we have the following corollary immideately from Theorem 3.7.
Corollary 3.8.
Under the same assumption of Corollary 3.7, we have the following:
-
(1)
For the geodesic curvature ,
-
•
if , then is bounded at ;
-
•
if , then is unbounded at ;
-
•
if and , then is unbounded at .
-
•
-
(2)
The normal curvature is unbounded at .
-
(3)
For the geodesic torsion ,
-
•
if , then is unbounded at ;
-
•
if and , then is bounded at ;
-
•
if , then is bounded at ,
where is the cuspidal curvature (cf. [12]) corresponding to .
-
•
Note that for is pointed out in [2, Proposition 2.19].
We observe that although in the above results we could not guarantee that the three invariants are bounded at the same time near a singular point, it is easy to find an example where it happens: taking and , it holds that , (see Figure 1). Thus these three invariants are bounded at (cf. Corollary 3.6). For the cuspidal edge and the same , we see that and are bounded, but is unbounded at (cf. Corollary 3.8). Figure 2 shows the graphs of these invariants near .






Acknowledgements
The authors thank the referee, Yuki Hattori and Atsufumi Honda for valuable comments and suggestions.
Appendix A Generalized biases for plane curve
Let be a curve-germ of -type which is given by the form (2.4) in the -plane . The terms measures the bias of near singular point. We call the -bias of at , and it is denoted by . We call is called the -cuspidal curvature as in [7], and it is denoted by .
If and are even, then it is a half part of a curve of -type, we consider the following cases: (1) both are odd, (2) is odd and is even, and (3) is even and is odd. Moreover, let denotes the first non-zero term of . We consider the case and . Then passes through the origin tangent to the -axis. In the case , if is odd, it also passes across the -axis. If is even, it approaches to the origin from one side of -axis and goes away into the same side of -axis, and if there does not exist such (namely, the bias is zero), it passes through the -axis. In the case , if the bias is zero, it approaches to the origin from one side of -axis and goes away into the same side of -axis. Figure 3 shows the images of the curves with from left to right. Figure 4 shows the images of the curves with from left to right.






We consider the case . Then approaches the origin from a direction of the -axis and making a cusp, and back to the same direction. If is both odd and even, it approaches to the origin from one side of -axis and goes away into the same side of -axis. If the bias is zero, it passes through the -axis. Figure 5 shows the images of the curves with from left to right.



Example A.1.
Let be a curve-germ -equivalent to . We set
One can calculate the invariants up to degrees as follows. The -cuspidal curvature is
(A.1) |
If , i.e., , then is -equivalent to . We assume . Then the -cuspidal curvature is
(A.2) |
If , i.e., , then is -equivalent to . We assume . Then the -bias and the -cuspidal curvature are
(A.3) | ||||
(A.4) |
If , i.e., , then is -equivalent to . We assume , i.e., . Then the -cuspidal curvature is
(A.5) |
If , then is -equivalent to . We assume , i.e., . Then the -bias and the -cuspidal curvature are
(A.6) | ||||
(A.7) |
Proof of Example A.1.
By rotating in , we can write
(A.8) |
We set
and the inverse function of as . We set . Then we have
Substituting into , and by a straightforward calculation, we see , and we have (A.1). Under the condition , we have , and we have (A.2). We assume , Then we see , and we have (A.3) and (A.4). We assume . Then we see , and we have (A.5). We assume . Then we see , and we have (A.6) and (A.7). ∎
Example A.2.
Let be a curve-germ -equivalent to . We set
(A.9) |
Then by a standard rotation in and a parameter change
we see
Thus the -cuspidal curvature is . Here, and are
References
- [1] M. P. do Carmo, Differential geometry of curves and surfaces, Prentice-Hall, Inc., Englewood Cliffs, N.J., 1976. MR 0394451
- [2] W. Domitrz and M. Zwierzyński, The Gauss-Bonnet theorem for coherent tangent bundles over surfaces with boundary and its applications, J. Geom. Anal. 30 (2020), 3243–3274. MR 4105152
- [3] S. Fujimori, K. Saji, M. Umehara and K. Yamada, Singularities of maximal surfaces. Math. Z. 259 (2008), 827–848. MR 2403743
- [4] T. Fukui, Local differential geometry of singular curves with finite multiplicities, Saitama Math. J. 31 (2017), 79–88. MR 3663296
- [5] T. Fukunaga, and M. Takahashi, Framed surfaces in the Euclidean space. Bull. Braz. Math. Soc. (N.S.) 50 (2019), 37–65. MR 3935057
- [6] M. Hasegawa, A. Honda, K. Naokawa, K. Saji, M. Umehara and K. Yamada, Intrinsic properties of surfaces with singularities. Internat. J. Math. 26 (2015), 1540008, 34 pp. MR 3338072
- [7] A. Honda and K. Saji, Geometric invariants of -cuspidal edges. Kodai Math. J. 42 (2019), 496–525. MR 3815292
- [8] A. Honda and H. Sato, Singularities of spacelike mean curvature one surfaces in de Sitter space, prepint, arXiv:2103.13849.
- [9] S. Izumiya, M. C. Romero Fuster, M. A. S. Ruas and F. Tari, Differential geometry from a singularity theory viewpoint, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2016. MR 3409029
- [10] S. Izumiya, K. Saji and N. Takeuchi, Flat surfaces along cuspidal edges, J. Singul. 16 (2017), 73–100. MR 3655304
- [11] L. F. Martins and K. Saji, Geometric invariants of cuspidal edges. Canad. J. Math. 68 (2016), 445–462. MR 3484374
- [12] L. F. Martins, K. Saji, M. Umehara and K. Yamada, Behavior of Gaussian curvature and mean curvature near non-degenerate singular points on wave fronts, Geometry and Topology of Manifold, Springer Proc. Math. & Statistics, 2016, 247–282. MR 3555987
- [13] L. F. Martins, K. Saji, S. P. Santos and K. Teramoto, Singular surfaces of revolution with prescribed unbounded mean curvature, An. Acad. Brasil. Ciênc. 91 (2019), no. 3, e20170865, 10 pp. MR 4017299
- [14] K. Saji, M. Umehara, and K. Yamada, The geometry of fronts, Ann. of Math. 169 (2009), 491–529. MR 2480610
- [15] S. Shiba and M. Umehara, The behavior of curvature functions at cusps and inflection points, Differential Geom. Appl. 30 (2012), no. 3, 285–299. MR 2922645
- [16] M. Takahashi and K. Teramoto, Surfaces of revolution of frontals in the Euclidean space, Bull. Brazilian Math. Soc. 51 (2020), 887–914. MR 4167695