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

Boundedness of geometric invariants near a singularity which is a suspension of a singular curve

Luciana F. Martins Kentaro Saji Samuel P. dos Santos  and  Keisuke Teramoto Universidade Estadual Paulista (UNESP), Instituto de Biociências Letras e Ciências Exatas, R. Cristóvão Colombo, 2265, Jd Nazareth, 15054-000 São José do Rio Preto, São Paulo, Brazil luciana.martins@unesp.br samuel.paulino@unesp.br Department of Mathematics, Graduate School of Science, Kobe University, Rokkodai 1-1, Nada, Kobe, 657-8501, Japan saji@math.kobe-u.ac.jp Graduate School of Sciences and Technology for Innovation, Yamaguchi University, Yamaguchi, 753-8512, Japan kteramoto@yamaguchi-u.ac.jp
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 torsion
2020 Mathematics Subject Classification:
Primary 57R45; Secondary 58K05
Partly supported by the Japan Society for the Promotion of Science KAKENHI Grants numbered 18K03301, 19K14533, 22K03312, 22K13914, the Japan-Brazil bilateral project JPJSBP1 20190103 and the São Paulo Research Foundation grant 2018/17712-7.

1. 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 σ\sigma be an 𝒜\mathcal{A}-equivalence class of singular plane curve-germs. A σ\sigma-edge is a map-germ f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) such that it is 𝒜\mathcal{A}-equivalent to (u,v)(u,c1(v),c2(v))(u,v)\mapsto(u,c_{1}(v),c_{2}(v)), where c=(c1,c2)c=(c_{1},c_{2}) is a representative of σ\sigma, namely, a one-dimensional suspension of σ\sigma. Here, two map-germs h1,h2:(𝑹m,0)(𝑹n,0)h_{1},h_{2}:(\boldsymbol{R}^{m},0)\to(\boldsymbol{R}^{n},0) are 𝒜\mathcal{A}-equivalent if there exist diffeomorphisms Φs:(𝑹m,0)(𝑹m,0)\Phi_{s}:(\boldsymbol{R}^{m},0)\to(\boldsymbol{R}^{m},0) and Φt:(𝑹n,0)(𝑹n,0)\Phi_{t}:(\boldsymbol{R}^{n},0)\to(\boldsymbol{R}^{n},0) such that h2=Φth1Φs1h_{2}=\Phi_{t}\circ h_{1}\circ\Phi_{s}^{-1}. A cuspidal edge (𝒜\mathcal{A}-equivalent to the germ (u,v)(u,v2,v3)(u,v)\mapsto(u,v^{2},v^{3}) at the origin) and a 5/25/2-cuspidal edge (𝒜\mathcal{A}-equivalent to the germ (u,v)(u,v2,v5)(u,v)\mapsto(u,v^{2},v^{5})) are examples of σ\sigma-edges, and σ\sigma are a 3/23/2-cusp and 5/25/2-cusp respectively. If σ\sigma is of finite multiplicity, then the σ\sigma-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 𝑹3\boldsymbol{R}^{3}, then σ\sigma-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 33-space (see [8]).

In this paper, we study geometry of σ\sigma-edges. For this, we consider two classes of singular map-germs, that we shall call mm-type and (m,n)(m,n)-type edges, the first including (m,n)(m,n)-type edges and also σ\sigma-edges when σ\sigma has finite multiplicity (see Section 2). One observes that mm-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 mm-type edges, getting same names, except for the cuspidal curvature, which we called (m,m+i)(m,m+i)-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 mm-type edge is a front if and only if the (m,m+1)(m,m+1)-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 σ\sigma-edges

We give several classes similar to σ\sigma-edges. It includes σ\sigma-edges, and these classes will be useful to treat. We recall that a map-germ f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) is a frontal if there exists a unit vector field ν\nu along ff such that dfp(Xp),ν(p)=0\left\langle{df_{p}(X_{p})},{\nu(p)}\right\rangle=0 holds at any p(𝑹2,0)p\in(\boldsymbol{R}^{2},0) and any XpTp𝑹2X_{p}\in T_{p}\boldsymbol{R}^{2}, where ,\left\langle{\cdot},{\cdot}\right\rangle is the canonical inner product of 𝑹3\boldsymbol{R}^{3}. The vector field ν\nu is called a unit normal vector field of ff. A map-germ f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) is an mm-type edge if it is 𝒜\mathcal{A}-equivalent to (u,vm,vm+1a(u,v))(u,v^{m},v^{m+1}a(u,v)) for a function a(u,v)a(u,v). A map-germ f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) is a (m,n)(m,n)-type edge (m<nm<n) if it is 𝒜\mathcal{A}-equivalent to (u,vm,vnh(u,v))(u,v^{m},v^{n}h(u,v)), where h(0,0)=1h(0,0)=1. This is equivalent to being 𝒜n\mathcal{A}^{n}-equivalent to (u,vm,vn)(u,v^{m},v^{n}). Two map-germs are 𝒜n\mathcal{A}^{n}-equivalent if their nn-jets at the origin are 𝒜\mathcal{A}-equivalent. We show:

Lemma 2.1.

Let ff be a σ\sigma-edge ((respectively, an mm-type edge, an (m,n)(m,n)-type edge)). Then an intersection curve of ff with a surface TT which is transversal to f(S(f))f(S(f)) passing through pS(f)p\in S(f) near 0 is 𝒜\mathcal{A}-equivalent to σ\sigma ((respectively, 𝒜m\mathcal{A}^{m}-equivalent to (tm,0)(t^{m},0), 𝒜n\mathcal{A}^{n}-equivalent to (tm,tn))(t^{m},t^{n})).

Proof.

Since the assumption and the assertion do not depend on the choice of the coordinate systems, we can assume ff is given by (u,c1(v),c2(v))(u,c_{1}(v),c_{2}(v)), where c=(c1,c2)c=(c_{1},c_{2}) is 𝒜\mathcal{A}-equivalent to σ\sigma. Then TT can be represented by the graph {(x,y,z)|x=h(y,z)}\{(x,y,z)\,|\,x=h(y,z)\} in (𝑹3,0)(\boldsymbol{R}^{3},0) as the xyzxyz-space, and the intersection curve is (h(c1(v),c2(v)),c1(v),c2(v))(h(c_{1}(v),c_{2}(v)),c_{1}(v),c_{2}(v)). Since TT is transverse to the xx-axis, the orthogonal projection of TT to the yzyz-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 mm or (m,n)(m,n)-type edge under the assumption n<2mn<2m. We assume n<2mn<2m throughout this subsection. Let f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be a frontal-germ satisfying rankdf0=1\operatorname{rank}df_{0}=1. Then there exists a vector field η\eta such that ηp\eta_{p} generates kerdfp\ker df_{p} if pS(f)p\in S(f). We call η|S(f)\eta|_{S(f)} a null vector field, and η\eta 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 S(f)S(f) is a regular curve, and the tangent direction of S(f)S(f) is not in kerdf0\ker df_{0}. Let ξ\xi be a vector field such that ξp\xi_{p} is a non-zero tangent vector of S(f)S(f) for pS(f)p\in S(f). We consider the following conditions for (ξ,η)(\xi,\eta):

  1. [2.1]

    ηif=0\eta^{i}f=0 (1im1)(1\leq i\leq m-1) on S(f)S(f),

  2. [2.2]

    rank(ξf,ηmf)=2\operatorname{rank}(\xi f,\eta^{m}f)=2 on S(f)S(f),

  3. [2.3]

    rank(ξf,ηmf,ηif)=2\operatorname{rank}(\xi f,\eta^{m}f,\eta^{i}f)=2 (m<i<n)(m<i<n) on S(f)S(f),

  4. [2.4]

    rank(ξf,ηmf,ηnf)=3\operatorname{rank}(\xi{f},\eta^{m}f,\eta^{n}f)=3 at pp.

Here, for a vector field ζ\zeta and a map ff, the symbol ζif\zeta^{i}f stands for the ii-times directional derivative of ff by ζ\zeta. Moreover, for a coordinate system (u,v)(u,v) and a map ff, the symbol fvif_{v^{i}} stands for if/vi\partial^{i}f/\partial v^{i}.

Proposition 2.2.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be a frontal-germ satisfying rankdf0=1\operatorname{rank}df_{0}=1. Assume that the set of singular points S(f)S(f) is a regular curve, and the tangent direction of S(f)S(f) generated by ξ\xi is not in kerdf0\ker df_{0}. If there exists a null vector field η\eta satisfying [2.1], and (ξ,η)(\xi,\eta) satisfies [2.2], then ff is an mm-type edge. Moreover, if (ξ,η)(\xi,\eta) also satisfies [2.3][2.4], then ff is an (m,n)(m,n)-type edge.

As we will see, the conditions [2.2][2.4] does not depend on the choice of null vector field η\eta 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, ff is as in Proposition 2.2.

Lemma 2.3.

Let Φ\Phi be a diffeomorphism-germ on (𝐑3,0)(\boldsymbol{R}^{3},0), and set f^=Φ(f)\hat{f}=\Phi(f). If ff and (ξ,η)(\xi,\eta) satisfy the condition CC, then f^\hat{f} and (ξ,η)(\xi,\eta) satisfy CC, where C={[2.1]}C=\{\ref{itm:cri1}\}, C={[2.1],[2.2]}C=\{\ref{itm:cri1},\ref{itm:cri2}\}, C={[2.1]C=\{\ref{itm:cri1}[2.3]}\ref{itm:cri3}\} and C={[2.1]C=\{\ref{itm:cri1}[2.4]}\ref{itm:cri4}\}.

Proof.

Let us assume η\eta satisfies [2.1]. By a direct calculation, we have ηf^=dΦ(f)ηf\eta\hat{f}=d\Phi(f)\eta f, and

ηif^=j=0i1cijηj(dΦ(f))ηijf(cij𝑹{0}).\eta^{i}\hat{f}=\sum_{j=0}^{i-1}c_{ij}\eta^{j}(d\Phi(f))\eta^{i-j}f\quad(c_{ij}\in\boldsymbol{R}\setminus\{0\}). (2.1)

By [2.1], ηif^=0\eta^{i}\hat{f}=0 (2im1)(2\leq i\leq m-1) and ηmf^=dΦ(f)ηmf\eta^{m}\hat{f}=d\Phi(f)\eta^{m}f on S(f)S(f). Then we see the assertion for the cases C={[2.1]}C=\{\ref{itm:cri1}\} and C={[2.1],[2.2]}C=\{\ref{itm:cri1},\ref{itm:cri2}\}. We assume η\eta satisfies [2.1][2.3]. By (2.1) and c1n=1(0)c_{1n}=1(\neq 0) we see the assertion. ∎

It is clear that the conditions [2.2][2.4] do not depend on the choice of ξ\xi, i.e., non-zero functional multiple and extension other than S(f)S(f). Moreover, it does not depend on the non-zero functional multiple of η\eta:

Lemma 2.4.

Let hh be a non-zero function. If ff and (ξ,η)(\xi,\eta) satisfy the condition CC, then ff and (ξ,η^)(\xi,\hat{\eta}) satisfy CC, where η^=hη\hat{\eta}=h\eta and CC is the same as those in Lemma 2.3.

Proof.

Since (hη)if(h\eta)^{i}f is a linear combination of ηf,,ηif\eta f,\ldots,\eta^{i}f, and the coefficient of ηif\eta^{i}f is hih^{i}, we see the assertion. ∎

A coordinate system (u,v)(u,v) satisfying S(f)={v=0}S(f)=\{v=0\}, η|S(f)=v\eta|_{S(f)}=\partial_{v} is said to be adapted.

Lemma 2.5.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be a frontal-germ satisfying rankdf0=1\operatorname{rank}df_{0}=1. We assume that the set of singular points S(f)S(f) is a regular curve. For any null vector field η\eta, there exists an adapted coordinate system (u,v)(u,v) such that η=v\eta=\partial_{v} for any (u,v)(u,v).

In this lemma, we do not assume that ff is an mm-type edges.

Proof.

Since rankdf0=1\operatorname{rank}df_{0}=1, one can easily see that there exists a coordinate system (u,v)(u,v) such that η=v\eta=\partial_{v} for any (u,v)(u,v). Since S(f)S(f) is a regular curve, and the tangent direction of it is not in kerdf0\ker df_{0}, S(f)S(f) can be parametrized as (u,a(u))(u,a(u)). Define a new coordinate system (u~,v~)(\tilde{u},\tilde{v}) by u~=u\tilde{u}=u and v~=va(u)\tilde{v}=v-a(u). Then S(f)={v~=0}S(f)=\{\tilde{v}=0\} and /v~=/v\partial/\partial\tilde{v}=\partial/\partial v hold. This shows the assertion. ∎

Lemma 2.6.

If two null vector fields η,η~\eta,\tilde{\eta} satisfy [2.1], and (ξ,η)(\xi,\eta) satisfies CC, then (ξ,η~)(\xi,\tilde{\eta}) also satisfies CC. Here, CC is the collection of the conditions: C={[2.2]}C=\{\ref{itm:cri2}\}, C={[2.2],[2.3]}C=\{\ref{itm:cri2},\ref{itm:cri3}\}, C={[2.2]C=\{\ref{itm:cri2}[2.4]}\ref{itm:cri4}\}.

Proof.

Let us assume that η\eta and η~\tilde{\eta} 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 (u,v)(u,v) an adapted coordinate system with η=v\eta=\partial_{v} for any (u,v)(u,v) and ξ=u\xi=\partial_{u}. Since fv==fvm1=0f_{v}=\cdots=f_{v^{m-1}}=0 on the uu-axis, fvf_{v} has the form fv=vm1ψ(u,v)f_{v}=v^{m-1}\psi(u,v). If the pair (ξ,η)(\xi,\eta) satisfies [2.2], then rank(fu,ψ)=2\operatorname{rank}(f_{u},\psi)=2 on the uu-axis. On the other hand, any null vector field is written as a1(u,v)u+a2(u,v)va_{1}(u,v)\partial_{u}+a_{2}(u,v)\partial_{v}, (a1(u,0)=0,a2(u,v)0)(a_{1}(u,0)=0,\,a_{2}(u,v)\neq 0). By Lemma 2.4, dividing this by a2a_{2}, we may assume an extended null vector field η~\tilde{\eta} is

η~=va(u,v)u+v.\tilde{\eta}=va(u,v)\partial_{u}+\partial_{v}.

Since it holds that η~2f=0\tilde{\eta}^{2}f=0 on the uu-axis (when m>2m>2) and fu(u,0)0f_{u}(u,0)\neq 0, we have a(u,0)=0a(u,0)=0. Continuing this argument, we may assume

η~=vm1a(u,v)u+v.\tilde{\eta}=v^{m-1}a(u,v)\partial_{u}+\partial_{v}. (2.2)

Thus η~f=vm1(afu+ψ)\tilde{\eta}f=v^{m-1}(af_{u}+\psi) holds, and η~mf=(m1)!(afu+ψ)\tilde{\eta}^{m}f=(m-1)!(af_{u}+\psi) holds on the uu-axis. Therefore (ξ,η~)(\xi,\tilde{\eta}) satisfies [2.2]. We assume that the pair (ξ,η=v)(\xi,\eta=\partial_{v}) satisfies [2.1]-[2.3] and (ξ,η~)(\xi,\tilde{\eta}) does [2.1], [2.2]. By this assumption, rank(fu,ψ)=2\operatorname{rank}(f_{u},\psi)=2, rank(fu,ψ,ψvi)=2\operatorname{rank}(f_{u},\psi,\psi_{v^{i}})=2 (0<i<nm)(0<i<n-m). By the form of η~\tilde{\eta}, it holds that η~m+1f=(m1)!(avfu+afuv+ψv)\tilde{\eta}^{m+1}f=(m-1)!(a_{v}f_{u}+af_{uv}+\psi_{v}) on the uu-axis. Since fv=0f_{v}=0 on the uu-axis, fuv=0f_{uv}=0 on the uu-axis. Thus rank(ξf,η~mf,η~m+1f)=2\operatorname{rank}(\xi f,\tilde{\eta}^{m}f,\tilde{\eta}^{m+1}f)=2 on the uu-axis. Similarly, fvm1=0f_{v^{m-1}}=0 on the uu-axis, fuv2==fuvm1=0f_{uv^{2}}=\cdots=f_{uv^{m-1}}=0 on the uu-axis. Thus if im1i\leq m-1, then since n<2mn<2m, we have η~m+if=(m1)!(ψvi+avifu)\tilde{\eta}^{m+i}f=(m-1)!(\psi_{v^{i}}+a_{v^{i}}f_{u}) on the uu-axis. Thereby we have rank(ξf,η~mf,η~m+if)=2\operatorname{rank}(\xi f,\tilde{\eta}^{m}f,\tilde{\eta}^{m+i}f)=2 (im2i\leq m-2) on the uu-axis. The last assertion can be shown by the same calculation. ∎

Proof of Proposition 2.2.

We assume ff satisfies the condition of the proposition, and (ξ,η)(\xi,\eta) satisfies the conditions [2.1] and [2.2]. Then we take an adapted coordinate system (u,v)(u,v) such that η=v\eta=\partial_{v}. By the proof of Lemma 2.6, there exist p(u)p(u) and q(u,v)q(u,v) such that f(u,v)=p(u)+vmq(u,v)f(u,v)=p(u)+v^{m}q(u,v), and (p1)u(0,0)0(p_{1})_{u}(0,0)\neq 0, where p=(p1,p2,p3)p=(p_{1},p_{2},p_{3}). We set U=p1(u),V=vU=p_{1}(u),V=v. Then ff has the form (U,P2(U),P3(U))+VmQ(U,V)(U,P_{2}(U),P_{3}(U))+V^{m}Q(U,V). By a coordinate change on the target, ff has the form (U,0,0)+VmQ(U,V)(U,0,0)+V^{m}Q(U,V), where Q(U,V)=(0,Q2(U,V),Q3(U,V))Q(U,V)=(0,Q_{2}(U,V),Q_{3}(U,V)). Rewriting the notation, we may assume ff is written as

f(u,v)=(u,vmq2(u,v),vmq3(u,v)).f(u,v)=(u,v^{m}q_{2}(u,v),v^{m}q_{3}(u,v)).

On this coordinate system, v\partial_{v} satisfies the condition [2.1], it satisfies [2.2] by Lemma 2.6. This implies (q2(0,0),q3(0,0))(0,0)(q_{2}(0,0),q_{3}(0,0))\neq(0,0). So we assume q2(0,0)0q_{2}(0,0)\neq 0. We set U=uU=u, V=vq2(u,v)1/mV=vq_{2}(u,v)^{1/m}. Rewriting the notation, we may assume ff is written as (u,vm,vmq3(u,v))(u,v^{m},v^{m}q_{3}(u,v)). By a coordinate change on the target, we may assume ff is written as (u,vm,vm+1q3(u,v))(u,v^{m},v^{m+1}q_{3}(u,v)). This proves the first assertion. We assume that η\eta also satisfies [2.3] and [2.4]. We may assume ff is written as (u,vm,vm+1q3(u,v))(u,v^{m},v^{m+1}q_{3}(u,v)). By Lemma 2.6, we may assume that v\partial_{v} satisfies [2.3] and [2.4]. By [2.3], the function q3(u,v)q_{3}(u,v) satisfies q3=(q3)v=(q3)vnm1=0q_{3}=(q_{3})_{v}=\cdots(q_{3})_{v^{n-m-1}}=0 on the uu-axis. Thus ff is written as (u,vm,vnq4(u,v))(u,v^{m},v^{n}q_{4}(u,v)). By [2.4], it holds that q40q_{4}\neq 0, and hence the assertion is proved. ∎

By the proof of Lemma 2.6, we have the following property:

Corollary 2.7.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be a frontal satisfying rankdf0=1\operatorname{rank}df_{0}=1 and the set of singular points S(f)S(f) is a regular curve. Furthermore, assume a vector field η\eta satisfying the condition [2.1] with v\partial_{v} satisfying [2.1]. Let (u,v)(u,v) be an adapted coordinate system. Then there exists ψ\psi such that ηf(u,v)=vm1ψ(u,v)\eta f(u,v)=v^{m-1}\psi(u,v).

2.2. Normal form of mm or (m,n)(m,n)-type edge

Given a curve-germ γ:(𝑹,0)(𝑹2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0), if there exists mm such that γ=tm1ρ\gamma^{\prime}=t^{m-1}\rho (ρ(0)0\rho(0)\neq 0), then γ\gamma at 0 is said to be of finite multiplicity, and such an mm is called the multiplicity or the order of γ\gamma at 0. Moreover, if there exists nn (n>mn>m and nkmn\neq km, k=2,3,k=2,3,\ldots) such that γ\gamma is 𝒜n\mathcal{A}^{n}-equivalent to (tm,tn)(t^{m},t^{n}), then γ\gamma is called of (m,n)(m,n)-type. This (m,n)(m,n) is well-defined since if γ\gamma is 𝒜r\mathcal{A}^{r}-equivalent to (tm,0)(t^{m},0) then it is not 𝒜r\mathcal{A}^{r}-equivalent to (tm,ti)(t^{m},t^{i}) for iri\leq r, ikmi\neq km (k=1,2,)(k=1,2,\ldots). We simplify a curve-germ of (m,n)(m,n)-type and an (m,n)(m,n)-type edge by coordinate changes on the source and by special orthonormal matrices on the target. Let (x,y)(x,y) be the ordinary coordinate system of (𝑹2,0)(\boldsymbol{R}^{2},0). A coordinate system (u,v)=(u(x,y),v(x,y))(u,v)=(u(x,y),v(x,y)) is positive if the determinant of the Jacobi matrix of (u(x,y),v(x,y))(u(x,y),v(x,y)) is positive. We have the following results.

Lemma 2.8.

Let γ:(𝐑,0)(𝐑2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a curve germ satisfying γ(i)(0)=0\gamma^{(i)}(0)=0 (i=1,,m1)(i=1,\ldots,m-1), and γ(m)(0)0\gamma^{(m)}(0)\neq 0. Then there exist a parameter tt and a special orthonormal matrix AA on 𝐑2\boldsymbol{R}^{2} such that

Aγ(t)=(tm,tm+1b(t)).A\gamma(t)=\left(t^{m},\ t^{m+1}b(t)\right). (2.3)

Let γ:(𝐑,0)(𝐑2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a curve germ of (m,n)(m,n)-type. Then there exist a parameter tt and a special orthonormal matrix AA on 𝐑2\boldsymbol{R}^{2} such that

Aγ(t)=(tm,i=2n/maitim+tnb(t))(b(0)0),A\gamma(t)=\left(t^{m},\ \sum_{i=2}^{\lfloor n/m\rfloor}a_{i}t^{im}+t^{n}b(t)\right)\quad(b(0)\neq 0), (2.4)

where k\lfloor k\rfloor is the greatest integer less than kk ((in our convention, n/mn/m is not an integer)).

Proof.

One can easily see the first assertion. We assume that γ\gamma is a curve germ of (m,n)(m,n)-type, then we may assume γ(t)=(tm,tm+1b(t))\gamma(t)=(t^{m},t^{m+1}b(t)). If tm+1b(t)t^{m+1}b(t) has a term tit^{i} (i<n,ikm)(i<n,i\neq km), then jnγ(0)j^{n}\gamma(0) is not 𝒜n\mathcal{A}^{n}-equivalent to (tm,tn)(t^{m},t^{n}). This proves the assertion. ∎

Proposition 2.9.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an mm-type edge. Then there exist a positive coordinate system (u,v)(u,v) and a special orthonormal matrix AA on 𝐑3\boldsymbol{R}^{3} such that

Af(u,v)=(u,u2a(u)2+vmm!,u2b0(u)2+vmm!bm(u,v))(bm(0,0)=0).Af(u,v)=\left(u,\dfrac{u^{2}a(u)}{2}+\dfrac{v^{m}}{m!},\dfrac{u^{2}b_{0}(u)}{2}+\dfrac{v^{m}}{m!}b_{m}(u,v)\right)\quad(b_{m}(0,0)=0). (2.5)

Moreover, if ff is an (m,n)(m,n)-type edge, then there exist a positive coordinate system (u,v)(u,v) and a special orthonormal matrix AA on 𝐑3\boldsymbol{R}^{3} such that

Af(u,v)=(u,u2a(u)2+vmm!,u2b0(u)2+i=2n/mvim(im)!bim(u)+vnbn(u,v)n!),Af(u,v)=\left(u,\dfrac{u^{2}a(u)}{2}+\dfrac{v^{m}}{m!},\dfrac{u^{2}b_{0}(u)}{2}+\sum_{i=2}^{\lfloor n/m\rfloor}\dfrac{v^{im}}{(im)!}b_{im}(u)+\dfrac{v^{n}b_{n}(u,v)}{n!}\right), (2.6)

bn(0,0)0b_{n}(0,0)\neq 0.

Proof.

By the proof of Proposition 2.2, we may assume

f(u,v)=(u,u2a2(u)+vma2m(u,v),u2a3(u)+vma3m(u,v)).f(u,v)=(u,u^{2}a_{2}(u)+v^{m}a_{2m}(u,v),u^{2}a_{3}(u)+v^{m}a_{3m}(u,v)).

By that proof again, (a2m(0,0),a3m(0,0))(0,0)(a_{2m}(0,0),a_{3m}(0,0))\neq(0,0). By a rotation on 𝑹3\boldsymbol{R}^{3}, we may assume a2m(0,0)>0a_{2m}(0,0)>0 and a3m(0,0)=0a_{3m}(0,0)=0. By a coordinate change vva2m(u,v)1/mv\mapsto va_{2m}(u,v)^{1/m}, we may assume f(u,v)=(u,u2a2(u)+vm/m!,u2a3(u)+vma3m(u,v))f(u,v)=(u,u^{2}a_{2}(u)+v^{m}/m!,u^{2}a_{3}(u)+v^{m}a_{3m}(u,v)), (a3m(0,0)=0)(a_{3m}(0,0)=0). This proves the first assertion. If ff is an (m,n)(m,n)-type edge, then the function a3m(u,v)a_{3m}(u,v) can be expanded by

i=0n1vibi(u)+vnbn(u,v).\sum_{i=0}^{n-1}v^{i}b_{i}(u)+v^{n}b_{n}(u,v).

Since ff is an (m,n)(m,n)-type edge, the curve vf(u,v)v\mapsto f(u,v) is of (m,n)(m,n)-type for any uu near 0. This implies that bi(u)=0b_{i}(u)=0 (ikm,k1)(i\neq km,k\geq 1). By a3m(0,0)=0a_{3m}(0,0)=0, b0(u)=0b_{0}(u)=0. This proves the assertion. ∎

Each form (2.5) and (2.6) is called the normal form of an mm-type edge and an (m,n)(m,n)-type edge, respectively. Looking the first and the second components in (2.5) and (2.6), we remark that the mm-jet of the coordinate system (u,v)(u,v) which gives the normal form is uniquely determined up to ±\pm when mm is even. Let f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an mm-type edge and η\eta a null vector field which satisfies the condition [2.1]. Then the subspace V1=df0(T0𝑹2)V_{1}=df_{0}(T_{0}\boldsymbol{R}^{2}) and the subspace V2V_{2} spanned by df0(T0𝑹2)df_{0}(T_{0}\boldsymbol{R}^{2}), ηmf(0)\eta^{m}f(0) do not depend on the choice of η\eta. We assume that the representation f=(f1,f2,f3)f=(f_{1},f_{2},f_{3}) of xyzxyz-space 𝑹3\boldsymbol{R}^{3} satisfies that V1V_{1} is the xx-axis and V2V_{2} is the xyxy-plane. Then the coordinate system (u,v)(u,v) gives the normal form (2.5) if and only if f1(u,v)=uf_{1}(u,v)=u and (f2)uv(f_{2})_{uv} is identically zero.

2.3. Geometric invariants

2.3.1. Cuspidal curvatures

Let ff be an mm-type edge. A pair of vector fields (ξ,η)(\xi,\eta) is said to be adapted if ξ\xi is tangent to S(f)S(f), and η\eta is a null vector field. We take an adapted pair of vector fields (ξ,η)(\xi,\eta) such that η\eta satisfies the condition [2.1], and (ξ,η)(\xi,\eta) is positively oriented. One can show the existence of such a pair by the definition of mm-type edge. We define

ωm,m+1(t)=|ξf|(m+1)/mdet(ξf,ηmf,ηm+1f)|ξf×ηmf|(2m+1)/m(μ(t))\omega_{m,m+1}(t)=\frac{|\xi{f}|^{(m+1)/m}\det(\xi{f},\eta^{m}{f},\eta^{m+1}f)}{|\xi{f}\times\eta^{m}{f}|^{(2m+1)/m}}(\mu(t))

where μ\mu is a parametrization of S(f)S(f). We call ωm,m+1\omega_{m,m+1} the (m,m+1)(m,m+1)-cuspidal curvature. We have the following lemma:

Proposition 2.10.

The function ωm,m+1\omega_{m,m+1} does not depend on the choice of (ξ,η)(\xi,\eta) satisfying the condition [2.1].

Proof.

Since it is not appeared in the formula, ωm,m+1\omega_{m,m+1} does not depend on the choice of the coordinate system. Let (ξ,η)(\xi,\eta) be an adapted pair of vector fields satisfying the condition [2.1]. It is clear that the function ωm,m+1\omega_{m,m+1} does not depend on the choice of ξ\xi. We take an adapted coordinate system (u,v)(u,v) satisfying v=η\partial_{v}=\eta. Then

ωm,m+1(u)=|fu|(m+1)/mdet(fu,fvm,fvm+1)|fu×fvm|(2m+1)/m.\omega_{m,m+1}(u)=|f_{u}|^{(m+1)/m}\det(f_{u},f_{v^{m}},f_{v^{m+1}})|f_{u}\times f_{v^{m}}|^{-(2m+1)/m}.

By Corollary 2.7, we have fv=vm1ψf_{v}=v^{m-1}\psi. Let η~\tilde{\eta} be another null vector field satisfying the condition [2.1]. We see that ωm,m+1\omega_{m,m+1} does not depend on the non-zero functional multiples of η\eta, we may assume η~=a(u,v)u+v\tilde{\eta}=a(u,v)\partial_{u}+\partial_{v}. By the proof of Lemma 2.6, we may assume that η~\tilde{\eta} is

η~=vm1a(u,v)u+v.\tilde{\eta}=v^{m-1}a(u,v)\partial_{u}+\partial_{v}. (2.7)

Then by fv=vm1ψf_{v}=v^{m-1}\psi,

η~f=vm1(afu+ψ).\tilde{\eta}f=v^{m-1}(af_{u}+\psi).

Thus

η~mf=(m1)!(afu+ψ)+(m1)(m1)!vη(afu+ψ)+v2g(u,v),\tilde{\eta}^{m}f=(m-1)!(af_{u}+\psi)+(m-1)(m-1)!v\eta(af_{u}+\psi)+v^{2}g(u,v), (2.8)

where gg is a function, and

η~m+1f=(m1)!η(afu+ψ)+(m1)(m1)!ηvη(afu+ψ)=m!(ηafu+aηfu+ηψ)\tilde{\eta}^{m+1}f=(m-1)!\eta(af_{u}+\psi)+(m-1)(m-1)!\eta v\eta(af_{u}+\psi)=m!(\eta af_{u}+a\eta f_{u}+\eta\psi)

hold on the uu-axis. Since ψ=((m1)!)1fvm\psi=((m-1)!)^{-1}f_{v^{m}} and ψv=(m!)1fvm+1\psi_{v}=(m!)^{-1}f_{v^{m+1}}, we have

|ξf|(m+1)/mdet(ξf,ηmf,ηm+1f)|ξf×ηmf|(2m+1)/m((m1)!)1/m(u,0)=\displaystyle\frac{|\xi{f}|^{(m+1)/m}\det(\xi{f},\eta^{m}{f},\eta^{m+1}f)}{|\xi{f}\times\eta^{m}{f}|^{(2m+1)/m}((m-1)!)^{1/m}}(u,0)= |fu|(m+1)/mdet(fu,ψ,ψv)|fu×ψ|(2m+1)/m(u,0)\displaystyle\frac{|f_{u}|^{(m+1)/m}\det(f_{u},\psi,\psi_{v})}{|f_{u}\times\psi|^{(2m+1)/m}}(u,0)
=\displaystyle= |fu|(m+1)/mdet(fu,fvm,fvm+1)|fu×fvm|(2m+1)/m(u,0).\displaystyle\frac{|f_{u}|^{(m+1)/m}\det(f_{u},f_{v^{m}},f_{v^{m+1}})}{|f_{u}\times f_{v^{m}}|^{(2m+1)/m}}(u,0).

This shows the assertion. ∎

We have the following proposition.

Proposition 2.11.

Let f:(𝐑2,0)(𝐑3,0)f\colon(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an mm-type edge. Then ff at 0 is an (m,m+1)(m,m+1)-type edge if and only if ωm,m+10\omega_{m,m+1}\neq 0 at 0.

Proof.

Since ff is an mm-type edge, by Proposition 2.9, we may assume that ff is given by the right-hand side of (2.5). Since bm(0,0)=0b_{m}(0,0)=0, there exist c1(u)c_{1}(u) and c2(u,v)c_{2}(u,v) such that bm(u,v)=c1(u)+vc2(u,v)b_{m}(u,v)=c_{1}(u)+vc_{2}(u,v). Since we can take η=v\eta=\partial_{v}, the function ωm,m+1\omega_{m,m+1} is a non-zero functional multiple of c2(u,0)c_{2}(u,0). Then we see the assertion. ∎

It is easy to show that (m,m+1)(m,m+1)-type edges are fronts and that an mm-type edge is a front if and only if ωm,m+10\omega_{m,m+1}\neq 0. In Appendix A, we define (m,n)(m,n)-cuspidal curvature for a curve germ of (m,n)(m,n)-type, denoting it by rm,nr_{m,n}. An intersection curve of (m,m+1)(m,m+1)-type edge ff with TT as in Lemma 2.1, is a curve-germ of (m,m+1)(m,m+1)-type. The following holds.

Corollary 2.12.

Let f:(𝐑2,0)(𝐑3,0)f\colon(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be a σ\sigma-edge, where σ\sigma is 𝒜\mathcal{A}-equivalent to v(vm,vm+1)v\mapsto(v^{m},v^{m+1}). Then the (m,m+1)(m,m+1)-cuspidal curvature ωm,m+1\omega_{m,m+1} at 0 coincides with the (m,m+1)(m,m+1)-cuspidal curvature rm,m+1r_{m,m+1} of the intersection curve ρ\rho of ff with a plane PP which is perpendicular to the tangent line to ff at 0.

Proof.

By the assumption, we may assume that ff is given by the normal form (2.5). Since fu(0,0)=(1,0,0)f_{u}(0,0)=(1,0,0) and fv(0,0)=(0,0,0)f_{v}(0,0)=(0,0,0), the plane PP is given by P={(0,y,z)𝑹3|y,z𝑹}P=\{(0,y,z)\in\boldsymbol{R}^{3}\ |\ y,z\in\boldsymbol{R}\}. Thus the intersection curve ρ\rho can be parametrized by

ρ(v)=f(0,v)=(0,vmm!,bm+1(0,v)vm+1(m+1)!).\rho(v)=f(0,v)=\left(0,\frac{v^{m}}{m!},\frac{b_{m+1}(0,v)v^{m+1}}{(m+1)!}\right).

This can be considered as a normal form of a curve which is 𝒜m+1\mathcal{A}^{m+1}-equivalent to v(vm,vm+1)v\to(v^{m},v^{m+1}). Hence we have the assertion by Example A.2. ∎

Let ff be an mm-type edge. We assume ωm,m+1\omega_{m,m+1} is identically zero on S(f)S(f). Let μ(t)\mu(t) be a parametrization of S(f)S(f). We define

ωm,m+2(t)=|ξf|(m+2)/mdet(ξf,ηmf,ηm+2f)|ξf×ηmf|(2m+2)/m(μ(t)).\omega_{m,m+2}(t)=\dfrac{|\xi f|^{(m+2)/m}\det(\xi f,\eta^{m}f,\eta^{m+2}f)}{|\xi f\times\eta^{m}f|^{(2m+2)/m}}(\mu(t)).

We will see this does not depend on the choice of (ξ,η)(\xi,\eta) which satisfies the conditions [2.1] and [2.2] in Proposition 2.2 and det(ξf,ηmf,ηjf)=0\det(\xi{f},\eta^{m}f,\eta^{j}f)=0 (j<m+2)(j<m+2). Inductively, we define ωm,m+i\omega_{m,m+i} when ωm,m+j=0\omega_{m,m+j}=0 (ji1)(j\leq i-1) by

ωm,m+i=|ξf|(m+i)/mdet(ξf,ηmf,ηm+if)|ξf×ηmf|(2m+i)/m(μ(t)).\omega_{m,m+i}=\dfrac{|\xi f|^{(m+i)/m}\det(\xi f,\eta^{m}f,\eta^{m+i}f)}{|\xi f\times\eta^{m}f|^{(2m+i)/m}}(\mu(t)).

We will also see this does not depend on the choice of (ξ,η)(\xi,\eta) satisfying the conditions [2.1] and [2.2] in Proposition 2.2 and det(ξf,ηmf,ηjf)=0\det(\xi{f},\eta^{m}f,\eta^{j}f)=0 (j<m+i)(j<m+i). If i=mi=m, we set βm,2m=ωm,2m\beta_{m,2m}=\omega_{m,2m}.

Proposition 2.13.

Under the assumption ωm,m+1==ωm,m+i1=0\omega_{m,m+1}=\cdots=\omega_{m,m+i-1}=0, the function ωm,m+i\omega_{m,m+i} (i=1,,m1)(i=1,\ldots,m-1) does not depend on the choice of the pair (ξ,η)(\xi,\eta) which satisfies the conditions [2.1] and [2.2] in Proposition 2.2 and det(ξf,ηmf,ηm+jf)=0\det(\xi{f},\eta^{m}f,\eta^{m+j}f)=0 (1j<i)(1\leq j<i).

Proof.

We already showed the case i=1i=1 in Proposition 2.10. Let (ξ,η)(\xi,\eta) be a pair of vector fields satisfying the assumption of lemma. We take an adapted coordinate system (u,v)(u,v) such that v=η\partial_{v}=\eta. By the proof of Lemma 2.6, we see fv=vm1ψf_{v}=v^{m-1}\psi.

Moreover, we have:

Lemma 2.14.

There exist functions α,β\alpha,\beta, and a vector valued function θ\theta such that

ψv=αfu+βψ+vi1θ.\psi_{v}=\alpha f_{u}+\beta\psi+v^{i-1}\theta. (2.9)
Proof.

Since fvm+1=(m1)!ψvf_{v^{m+1}}=(m-1)!\psi_{v} on the uu-axis, ωm,m+1=0\omega_{m,m+1}=0 implies that there exists α1,β1,θ1\alpha_{1},\beta_{1},\theta_{1} such that ψv=α1fu+β1ψ+vθ1\psi_{v}=\alpha_{1}f_{u}+\beta_{1}\psi+v\theta_{1}. We assume that there exist αk,βk,θk\alpha_{k},\beta_{k},\theta_{k} such that ψv=αkfu+βkψ+vkθk\psi_{v}=\alpha_{k}f_{u}+\beta_{k}\psi+v^{k}\theta_{k} (k=1,,i2)(k=1,\ldots,i-2). Differentiating this equation, we have

ψvk+1=l=0k(kl)((αk)vlfuvkl+(βk)vlψvkl+(vk)vl(θk)vkl)\psi_{v^{k+1}}=\sum_{l=0}^{k}{}{\begin{pmatrix}k\\ l\end{pmatrix}}\Big{(}(\alpha_{k})_{v^{l}}f_{uv^{k-l}}+(\beta_{k})_{v^{l}}\psi_{v^{k-l}}+(v^{k})_{v^{l}}(\theta_{k})_{v^{k-l}}\Big{)}

Thus by fuv==fuvm1=0f_{uv}=\cdots=f_{uv^{m-1}}=0 holds, and ψvjfu,ψ𝑹\psi_{v^{j}}\in\left\langle f_{u},\psi\right\rangle_{\boldsymbol{R}} (jk)(j\leq k) on the uu-axis, we have 2=rank(fu,ψ,ψvk+1)=rank(fu,ψ,θk)2=\operatorname{rank}(f_{u},\psi,\psi_{v^{k+1}})=\operatorname{rank}(f_{u},\psi,\theta_{k}) on the uu-axis. Hence there exist functions αk+1,βk+1\alpha_{k+1},\beta_{k+1}, and a vector valued function θk+1\theta_{k+1} such that θk=αk+1fu+βk+1ψ+vθk+1\theta_{k}=\alpha_{k+1}f_{u}+\beta_{k+1}\psi+v\theta_{k+1}. 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 η\eta as in the right-hand side of (2.7). By the same calculations in the proof of Proposition 2.10, we have ηf=vm1(afu+ψ)\eta f=v^{m-1}(af_{u}+\psi). Thus

ηm+if=k=0m+i1(m+i1k)ηkvm1ηm+i1k(afu+ψ).\eta^{m+i}f=\sum_{k=0}^{m+i-1}{\begin{pmatrix}m+i-1\\ k\end{pmatrix}}\eta^{k}v^{m-1}\eta^{m+i-1-k}(af_{u}+\psi).

Since ηkvm1=0\eta^{k}v^{m-1}=0 if km1k\neq m-1 and ηkvm1=(m1)!\eta^{k}v^{m-1}=(m-1)!,

ηm+if=(m+i1m1)(m1)!ηi(afu+ψ).\eta^{m+i}f={\begin{pmatrix}m+i-1\\ m-1\end{pmatrix}}(m-1)!\eta^{i}(af_{u}+\psi).

Thus ηm+if=vg(u,v)+(afu+ψ)vi\eta^{m+i}f=vg(u,v)+(af_{u}+\psi)_{v^{i}}, where gg is a function. By fuv==fuvm1=0f_{uv}=\ldots=f_{uv^{m-1}}=0 holds, and ψvjfu,ψ𝑹\psi_{v^{j}}\in\left\langle f_{u},\psi\right\rangle_{\boldsymbol{R}} (jk)(j\leq k) on the uu-axis by Lemma 2.14, we have

|ξf|(m+i)/mdet(ξf,ηmf,ηm+if)|ξf×ηmf|(2m+i)/m(u,0)=\displaystyle\dfrac{|\xi f|^{(m+i)/m}\det(\xi f,\eta^{m}f,\eta^{m+i}f)}{|\xi f\times\eta^{m}f|^{(2m+i)/m}}(u,0)= |fu|(m+i)/mdet(fu,fvm,fvm+i)|fu×fvm|(2m+i)/m(u,0),\displaystyle\dfrac{|f_{u}|^{(m+i)/m}\det(f_{u},f_{v^{m}},f_{v^{m+i}})}{|f_{u}\times f_{v^{m}}|^{(2m+i)/m}}(u,0),

and this shows the assertion. ∎

We call ωm,m+i\omega_{m,m+i} the (m,m+i)(m,m+i)-cuspidal curvature, and βm,2m\beta_{m,2m} the (m,2m)(m,2m)-bias. Note that βm,2m\beta_{m,2m} does not depend on the choice of (ξ,η)(\xi,\eta) satisfying [2.1], [2.2] and ξf,ηmf=0\left\langle{\xi{f}},{\eta^{m}{f}}\right\rangle=0 at pp by the same calculation. In this case, a(0,0)=0a(0,0)=0 by the additional assumption. If ff is an mm-type edge, and written as (2.5), then ωm,m+1(0)=(m+1)(bm)v(0,0)\omega_{m,m+1}(0)=(m+1)(b_{m})_{v}(0,0). If ff is an (m,n)(m,n)-edge (n<2m)(n<2m), and written as (2.6), then ωm,n(0)=bn(0,0)\omega_{m,n}(0)=b_{n}(0,0), and βm,2m(0,0)=b2m(0)\beta_{m,2m}(0,0)=b_{2m}(0). See Appendix A for geometric meanings of the terms bimb_{im} (i=2,,n/m)(i=2,\ldots,{\lfloor n/m\rfloor}).

2.3.2. Singular, normal curvatures and cuspidal torsion

Let ff be an mm-type edge, and μ(t)\mu(t) be a parametrization of the singular set. Let ν\nu be a unit normal vector field of ff, and we set λ=det(fu,fv,ν)\lambda=\det(f_{u},f_{v},\nu) for an oriented coordinate system (u,v)(u,v) on (𝑹2,0)(\boldsymbol{R}^{2},0). We set μ^=fμ\hat{\mu}=f\circ\mu. Then we define

κs(t)=sgn(δηm1λ(μ(t)))det(μ^,μ^′′,ν(μ))|μ^|3,κν(t)=μ^′′,ν(μ)|μ^|2\kappa_{s}(t)=\operatorname{sgn}\Big{(}\delta\ \eta^{m-1}\lambda(\mu(t))\Big{)}\dfrac{\det(\hat{\mu}^{\prime},\hat{\mu}^{\prime\prime},\nu(\mu))}{|\hat{\mu}^{\prime}|^{3}},\quad\kappa_{\nu}(t)=\dfrac{\left\langle{\hat{\mu}^{\prime\prime}},{\nu(\mu)}\right\rangle}{|\hat{\mu}^{\prime}|^{2}} (2.10)

and

κt(t)=det(ξf,ηmf,ξηmf)|ξf×ηmf|2(μ(t))det(ξf,ηmf,ξ2f)ξf,ηmf|ξf|2|ξf×ηmf|2(μ(t)),\kappa_{t}(t)=\dfrac{\det(\xi f,\eta^{m}f,\xi\eta^{m}f)}{|\xi f\times\eta^{m}f|^{2}}(\mu(t))-\dfrac{\det(\xi f,\eta^{m}f,\xi^{2}f)\left\langle{\xi f},{\eta^{m}f}\right\rangle}{|\xi f|^{2}|\xi f\times\eta^{m}f|^{2}}(\mu(t)), (2.11)

where δ=1\delta=1 if (μ,η)(\mu^{\prime},\eta) agrees the orientation of the coordinate system, and δ=1\delta=-1 if (μ,η)(\mu^{\prime},\eta) does not agree the orientation. We call κs\kappa_{s}, κν\kappa_{\nu} and κt\kappa_{t} 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, κs\kappa_{s} does not depend on the choice of ν\nu, nor the choice of η\eta when mm is even. To see the well-definedness of κt\kappa_{t}, we need:

Proposition 2.15.

The definition (2.11) does not depend on the choice of the adapted vector fields (ξ,η)(\xi,\eta) with η\eta which satisfies [2.1].

Proof.

One can easily to check it does not depend on the choice of functional multiplications of η\eta. Since the assertion does not depend on the choice of local coordinate system, one can choose an adapted coordinate system (u,v)(u,v) with v\partial_{v} satisfying [2.1]. Let η\eta be a null vector field which satisfies [2.1]. Then by the proof of Lemma 2.6, we may assume η\eta is given by (2.2). Then by (2.8), we see ηmf=(m1)!(afu+ψ)\eta^{m}f=(m-1)!(af_{u}+\psi) on the uu-axis, where ψ\psi is given in the proof of Lemma 2.6. Furthermore, by (2.8), we see and ξηmf=(m1)!(aufu+afu2+ψu)\xi\eta^{m}f=(m-1)!(a_{u}f_{u}+af_{u^{2}}+\psi_{u}) on the uu-axis. Substituting these formulas into the right-hand side of (2.11), we see it is

det(fu,ψ,ψu)|fu×ψ|2(u,0)det(fu,ψ,fu2)fu,ψ|fu|2|fu×ψ|2(u,0),\dfrac{\det(f_{u},\psi,\psi_{u})}{|f_{u}\times\psi|^{2}}(u,0)-\dfrac{\det(f_{u},\psi,f_{u^{2}})\left\langle{f_{u}},{\psi}\right\rangle}{|f_{u}|^{2}|f_{u}\times\psi|^{2}}(u,0),

and by fvm=(m1)!ψf_{v^{m}}=(m-1)!\psi, this shows the assertion. ∎

If an mm-type edge ff is given by the form (2.5), then κs(0)=a(0)\kappa_{s}(0)=a(0), κν(0)=b(0)\kappa_{\nu}(0)=b(0) and κt(0)=(bm)u(0,0)\kappa_{t}(0)=(b_{m})_{u}(0,0).

2.4. Boundedness of Gaussian curvature and mean curvature near an mm-type edge

Here we study the behavior of the Gaussian and mean curvatures.

Let g:(𝑹i,0)𝑹g:(\boldsymbol{R}^{i},0)\to\boldsymbol{R} be a function-germ (i=1,2)i=1,2). If there exists an integer nn (n1)(n\geq 1) such that ging\in\mathcal{M}_{i}^{n} and gin+1g\not\in\mathcal{M}_{i}^{n+1}, then gg is said to be of order nn, where i={g:(𝑹i,0)𝑹|g(0)=0}\mathcal{M}_{i}=\{g\colon(\boldsymbol{R}^{i},0)\to\boldsymbol{R}\ |\ g(0)=0\} is the unique maximal ideal of the local ring of function-germs and in\mathcal{M}_{i}^{n} denotes the nnth power of i\mathcal{M}_{i} (cf. [9, p. 46]). If gig\not\in\mathcal{M}_{i}, then the order of gg is 0. The order of gg is denoted by ord(g)\operatorname{ord}(g). If gg is of order nn (n0)(n\geq 0), then gg is said to be of finite order. Let g1,g2:(𝑹i,0)𝑹g_{1},g_{2}:(\boldsymbol{R}^{i},0)\to\boldsymbol{R} be two function-germs such that gig_{i} is of finite order. The rational order ord(f)\operatorname{ord}(f) of a function f=g1/g2:(𝑹iZ,0)𝑹f=g_{1}/g_{2}:(\boldsymbol{R}^{i}\setminus Z,0)\to\boldsymbol{R}, where Z=g21(0)Z=g_{2}^{-1}(0) is

ord(f)=ord(g1)ord(g2).\operatorname{ord}(f)=\operatorname{ord}(g_{1})-\operatorname{ord}(g_{2}).

For a function f=g1/(|g2|g3):(𝑹iZ,0)𝑹f=g_{1}/(|g_{2}|g_{3}):(\boldsymbol{R}^{i}\setminus Z,0)\to\boldsymbol{R}, we define ord(f)=ord(g1)ord(g2)ord(g3)\operatorname{ord}(f)=\operatorname{ord}(g_{1})-\operatorname{ord}(g_{2})-\operatorname{ord}(g_{3}), where Z=g21(0)g31(0)Z=g_{2}^{-1}(0)\cup g_{3}^{-1}(0). If g1ig_{1}\in\mathcal{M}_{i}^{\infty}, then we define ord(f)=\operatorname{ord}(f)=\infty. If ord(f)=0\operatorname{ord}(f)=0, then ff is called rationally bounded, and ord(f)=1\operatorname{ord}(f)=1, then ff is called rationally continuous ([12, Definition 3.4]). If i=1i=1, this is the usual one.

Since the property ging\in\mathcal{M}_{i}^{n} 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 f:(𝑹2,0)(𝑹3,0)f\colon(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an mm-type edge, and let (u,v)(u,v) be an adapted coordinate system with v\partial_{v} satisfying [2.1]. We take (m1)!ψ(m-1)!\psi in Corollary 2.7. Namely, here we set ψ\psi by fv=vm1ψ/(m1)!f_{v}=v^{m-1}\psi/(m-1)!. Since ff is an mm-type edge, fuf_{u} and ψ\psi is linearly independent (Proposition 2.2 and the independence of the condition [2.2]). Thus the unit normal vector ν\nu of ff can be taken as ν=ν^/|ν^|\nu=\hat{\nu}/|\hat{\nu}| (ν^=fu×ψ)(\hat{\nu}=f_{u}\times\psi). Using fuf_{u}, ψ\psi and ν\nu, we define the following functions:

E^\displaystyle\widehat{E} =fu,fu,\displaystyle=\left\langle{f_{u}},{f_{u}}\right\rangle, F^\displaystyle\widehat{F} =fu,ψ,\displaystyle=\left\langle{f_{u}},{\psi}\right\rangle, G^\displaystyle\widehat{G} =ψ,ψ,\displaystyle=\left\langle{\psi},{\psi}\right\rangle,
L^\displaystyle\widehat{L} =fu,ν^u,\displaystyle=-\left\langle{f_{u}},{\widehat{\nu}_{u}}\right\rangle, M^\displaystyle\widehat{M} =ψ,ν^u,\displaystyle=-\left\langle{\psi},{\widehat{\nu}_{u}}\right\rangle, N^\displaystyle\widehat{N} =ψ,ν^v.\displaystyle=-\left\langle{\psi},{\widehat{\nu}_{v}}\right\rangle.

We note that coefficients of the first and the second fundamental forms of σ\sigma-edges being of multiplicity mm can be written as

E\displaystyle E =E^,\displaystyle=\widehat{E}, F\displaystyle F =vm1(m1)!F^,\displaystyle=\dfrac{v^{m-1}}{(m-1)!}\widehat{F}, G\displaystyle G =(vm1(m1)!)2G^,\displaystyle=\left(\dfrac{v^{m-1}}{(m-1)!}\right)^{2}\widehat{G},
L\displaystyle L =L^|ν^|,\displaystyle=\dfrac{\widehat{L}}{|\widehat{\nu}|}, M\displaystyle M =vm1|ν^|(m1)!M^,\displaystyle=\dfrac{v^{m-1}}{|\widehat{\nu}|(m-1)!}\widehat{M}, N\displaystyle N =vm1(m1)!|ν^|N^.\displaystyle=\dfrac{v^{m-1}}{(m-1)!|\widehat{\nu}|}\widehat{N}.
Lemma 2.16.

The differentials νu\nu_{u} and νv\nu_{v} of ν\nu are written as

νu\displaystyle\nu_{u} =G^L^F^M^(E^G^F^2)|ν^|fuE^M^F^L^(E^G^F^2)|ν^|ψ,\displaystyle=-\dfrac{\widehat{G}\widehat{L}-\widehat{F}\widehat{M}}{(\widehat{E}\widehat{G}-\widehat{F}^{2})|\hat{\nu}|}f_{u}-\dfrac{\widehat{E}\widehat{M}-\widehat{F}\widehat{L}}{(\widehat{E}\widehat{G}-\widehat{F}^{2})|\hat{\nu}|}\psi,
νv\displaystyle\nu_{v} =vm1(m1)!G^M^F^N^(E^G^F^2)|ν^|fuE^N^vm1(m1)!F^M^(E^G^F^2)|ν^|ψ.\displaystyle=-\dfrac{\dfrac{v^{m-1}}{(m-1)!}\widehat{G}\widehat{M}-\widehat{F}\widehat{N}}{(\widehat{E}\widehat{G}-\widehat{F}^{2})|\hat{\nu}|}f_{u}-\dfrac{\widehat{E}\widehat{N}-\dfrac{v^{m-1}}{(m-1)!}\widehat{F}\widehat{M}}{(\widehat{E}\widehat{G}-\widehat{F}^{2})|\hat{\nu}|}\psi.
Proof.

Since νu,ν=νv,ν=0\left\langle{\nu_{u}},{\nu}\right\rangle=\left\langle{\nu_{v}},{\nu}\right\rangle=0, there exist functions A,B,C,DA,B,C,D on (𝑹2,0)(\boldsymbol{R}^{2},0) such that

νu=Afu+Bψ,νv=Cfu+Dψ.\nu_{u}=Af_{u}+B\psi,\quad\nu_{v}=Cf_{u}+D\psi.

Considering νu,fu,νu,ψ,νv,fu\left\langle{\nu_{u}},{f_{u}}\right\rangle,\left\langle{\nu_{u}},{\psi}\right\rangle,\left\langle{\nu_{v}},{f_{u}}\right\rangle and νv,ψ\left\langle{\nu_{v}},{\psi}\right\rangle, we have

1|ν^|(L^M^)=(E^F^F^G^)(AB),1|ν^|(vm1(m1)!M^N^)=(E^F^F^G^)(CD).-\frac{1}{|\hat{\nu}|}\begin{pmatrix}\widehat{L}\\ \widehat{M}\end{pmatrix}=\begin{pmatrix}\widehat{E}&\widehat{F}\\ \widehat{F}&\widehat{G}\end{pmatrix}\begin{pmatrix}A\\ B\end{pmatrix},\quad-\frac{1}{|\hat{\nu}|}\begin{pmatrix}\frac{v^{m-1}}{(m-1)!}\widehat{M}\\ \widehat{N}\end{pmatrix}=\begin{pmatrix}\widehat{E}&\widehat{F}\\ \widehat{F}&\widehat{G}\end{pmatrix}\begin{pmatrix}C\\ D\end{pmatrix}.

Solving these equations, we have the assertion. ∎

By this lemma, νv\nu_{v} can be written as

νv=N^(E^G^F^2)|ν^|(F^fuE^ψ)\nu_{v}=\frac{\widehat{N}}{(\widehat{E}\widehat{G}-\widehat{F}^{2})|\hat{\nu}|}(\widehat{F}f_{u}-\widehat{E}\psi)

along the uu-axis. Since fuf_{u} and ψ\psi are linearly independent and E^0\widehat{E}\neq 0, the condition νv(0)0\nu_{v}(0)\neq 0 is equivalent to N^(0)0\widehat{N}(0)\neq 0. To see this fact, we take the same setting in the proof of Proposition 2.10. Then we see

det(fu,fvm,fvm+1)=mdet(fu,ψ,ψv)=mν^,ψv=mN^\det(f_{u},f_{v^{m}},f_{v^{m+1}})=m\det(f_{u},\psi,\psi_{v})=m\left\langle{\hat{\nu}},{\psi_{v}}\right\rangle=m\widehat{N} (2.12)

along the uu-axis, where ν^=fu×ψ\hat{\nu}=f_{u}\times\psi and N^=ν^,ψv=ν^v,ψ\widehat{N}=\left\langle{\hat{\nu}},{\psi_{v}}\right\rangle=-\left\langle{\hat{\nu}_{v}},{\psi}\right\rangle. Since {fu,ψ,ν}\{f_{u},\psi,\nu\} is a frame of 𝑹3\boldsymbol{R}^{3}, and fu,νv=fv,νu=0\left\langle{f_{u}},{\nu_{v}}\right\rangle=\left\langle{f_{v}},{\nu_{u}}\right\rangle=0, ν,νv=0\left\langle{\nu},{\nu_{v}}\right\rangle=0, it holds that νv0\nu_{v}\neq 0 if and only if νv,ψ0\left\langle{\nu_{v}},{\psi}\right\rangle\neq 0. Moreover, since ν,ψ=0\left\langle{\nu},{\psi}\right\rangle=0, it holds that νv,ψ0\left\langle{\nu_{v}},{\psi}\right\rangle\neq 0 is equivalent to ν^v,ψ0\left\langle{\hat{\nu}_{v}},{\psi}\right\rangle\neq 0. Let f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an (m,n)(m,n)-type edge, and let us set

r=min({n}{im|bim(0)0 in the form (2.6),i=2,3,}).r=\min\Big{(}\{n\}\cup\{im\,|\,b_{im}(0)\neq 0\text{ in the form }\eqref{eq:edgenormal2},i=2,3,\ldots\}\Big{)}.
Theorem 2.17.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an (m,n)(m,n)-type edge. Then the rational order of the mean curvature HH is r2mr-2m. If the normal curvature does not vanish at 0, then the rational order of the Gaussian curvature KK is r2mr-2m.

Proof.

We take an adapted coordinate system (u,v)(u,v) such that v\partial_{v} satisfies [2.1]. Since L^=fuu,ν\hat{L}=\left\langle{f_{uu}},{\nu}\right\rangle, the normal curvature does not vanish if and only if L^(0)0\hat{L}(0)\neq 0. The Gaussian curvature KK and the mean curvature HH of ff are given by

K=(m1)!vm1L^N^vm1(m1)!M^2|ν^|2(E^G^F^2),H=(m1)!vm1E^N^2vm1(m1)!F^M^+vm1(m1)!G^L^2|ν^|(E^G^F^2).K=\dfrac{(m-1)!}{v^{m-1}}\dfrac{\widehat{L}\widehat{N}-\frac{v^{m-1}}{(m-1)!}\widehat{M}^{2}}{|\widehat{\nu}|^{2}(\widehat{E}\widehat{G}-\widehat{F}^{2})},\quad H=\dfrac{(m-1)!}{v^{m-1}}\dfrac{\widehat{E}\widehat{N}-2\frac{v^{m-1}}{(m-1)!}\widehat{F}\widehat{M}+\frac{v^{m-1}}{(m-1)!}\widehat{G}\widehat{L}}{2|\widehat{\nu}|(\widehat{E}\widehat{G}-\widehat{F}^{2})}.

This and E^0\widehat{E}\neq 0, E^G^F^20\widehat{E}\widehat{G}-\widehat{F}^{2}\neq 0 at 0, together with

N^=vrm1m(rm1)!(br(0)+vα(u,v)),\widehat{N}=\dfrac{v^{r-m-1}}{m(r-m-1)!}(b_{r}(0)+v\alpha(u,v)),

where α\alpha is a function, by using the form (2.6) and (2.12) give the assertion. ∎

By Theorem 2.17, the order of KK and HH coincide. Moreover since n<2mn<2m, they never bounded when the normal curvature does not vanish.

3. Curves passing through mm-type edges

In this section, we consider geometric invariants of a curve γ\gamma passing through an mm-type edge ff. If γ^=fγ\hat{\gamma}=f\circ\gamma is non-singular, then the usual invariants can be defined as well as the regular case. We consider the case when γ^\hat{\gamma} has a singular point, namely, γ\gamma passing through a singular point of ff in the direction of a null vector.

3.1. Normalized curvatures of singular curves

Following [15, 4], we introduce normalized curvature on curves in 𝑹2\boldsymbol{R}^{2}. Let γ^:(𝑹,0)(𝑹n,0)\hat{\gamma}:(\boldsymbol{R},0)\to(\boldsymbol{R}^{n},0) be a curve, and let 0 be a singular point. We assume that there exists kk such that γ^=tk1ρ\hat{\gamma}^{\prime}=t^{k-1}\rho (ρ(0)0\rho(0)\neq 0).

We set

s=|γ^|𝑑t,s=\int|\hat{\gamma}^{\prime}|\,dt, (3.1)

and

s~=sgn(s)|s|1/k,\tilde{s}=\operatorname{sgn}(s)|s|^{1/k}, (3.2)

we see s~\tilde{s} is a CC^{\infty} function and ds~/dt(0)>0d\tilde{s}/dt(0)>0. We call this parameter an 1/k1/k-arc-length.

Proposition 3.1.

The parameter tt is an 1/k1/k-arc-length parameter of γ^\hat{\gamma} if and only if |γ^(t)|=k|tk1||\hat{\gamma}^{\prime}(t)|=k|t^{k-1}|.

Proof.

If |γ^(t)|=k|tk1||\hat{\gamma}^{\prime}(t)|=k|t^{k-1}| and s(t)s(t) as in (3.1) it holds that

s(t)=0tk|ξk1|𝑑ξ=0tεkξk1𝑑ξ=εtk,(ε={sgn(t),if k is even1if k is odd).s(t)=\int_{0}^{t}k|\xi^{k-1}|d\xi=\int_{0}^{t}\varepsilon\,k\xi^{k-1}d\xi=\varepsilon\,t^{k},\ \left(\varepsilon=\begin{dcases}\operatorname{sgn}(t),\ &\text{if $k$ is even}\\ 1\ &\text{if $k$ is odd}\end{dcases}\right).

Since sgn(s)=sgn(t)\operatorname{sgn}(s)=\operatorname{sgn}(t), then |s|=|tk||s|=|t^{k}| and, therefore, t=sgn(s)|s|1/kt=\operatorname{sgn}(s)|s|^{1/k}.

Let us suppose now that tt is the 1/k1/k-arc-length, i.e., t=sgn(s)|s|1/kt=\operatorname{sgn}(s)|s|^{1/k}, with s(t)s(t) as in (3.1). Since sgn(s)=sgn(t)\operatorname{sgn}(s)=\operatorname{sgn}(t), thus tk=sgn(s)k|s|=sgn(s)k+1st^{k}=\operatorname{sgn}(s)^{k}|s|=\operatorname{sgn}(s)^{k+1}s and, consequently, s(t)=sgn(t)k+1ktk1=k|t|k1s^{\prime}(t)=\operatorname{sgn}(t)^{k+1}kt^{k-1}=k|t|^{k-1}. Therefore, it holds that |γ^(t)|=k|tk1||\hat{\gamma}^{\prime}(t)|=k|t^{k-1}|. ∎

Let us set n=2n=2. Then the curvature κ\kappa satisfies that

κ~=|s~k1|κ\tilde{\kappa}=|\tilde{s}^{k-1}|\kappa (3.3)

is a CC^{\infty} function. We call κ~\tilde{\kappa} the normalized curvature. This is originally introduced in [15] and generalized in [4]. Let f(t)f(t) be a given CC^{\infty}-function, and k2k\geq 2 be an integer. Then similarly to [15, Theorem 1.1], one can show that there exists a unique plane curve up to isometries in 𝑹2\boldsymbol{R}^{2} with normalized curvature given by κ~(t)=f(t)\tilde{\kappa}(t)=f(t), where tt is the 1/k1/k-arc-length parameter.

Using the frame {𝒆,𝒏}\{\boldsymbol{e},\boldsymbol{n}\} along γ^\hat{\gamma} defined by 𝒆=ρ/|ρ|\boldsymbol{e}=\rho/|\rho| and 𝒏\boldsymbol{n} the π/2\pi/2-rotation of 𝒆\boldsymbol{e}, the normalized curvature can be interpreted as follows: We set the function κ1\kappa_{1} by the equation

(𝒆𝒏)=(0κ1κ10)(𝒆𝒏),{\begin{pmatrix}\boldsymbol{e}^{\prime}\\ \boldsymbol{n}^{\prime}\end{pmatrix}}={\begin{pmatrix}0&\kappa_{1}\\ -\kappa_{1}&0\end{pmatrix}}{\begin{pmatrix}\boldsymbol{e}\\ \boldsymbol{n}\end{pmatrix}}, (3.4)

where is a differentiation by the 1/k1/k-arc-length. Then we have:

Proposition 3.2.

Let {𝐞,𝐧}\{\boldsymbol{e},\boldsymbol{n}\} be the above frame along γ^(t)\hat{\gamma}(t) in the Euclidean plane 𝐑2\boldsymbol{R}^{2} satisfying (3.4), where tt is the 1/k1/k-arc-length parameter. Then κ1=kκ~\kappa_{1}=k\tilde{\kappa} holds.

Proof.

Since γ^(t)=tk1ρ(t)\hat{\gamma}^{\prime}(t)=t^{k-1}\rho(t) where ρ(0)0\rho(0)\neq 0 and the 1/k1/k-arc-length parameter tt satisfies |γ^(t)|=k|t|k1|\hat{\gamma}^{\prime}(t)|=k|t|^{k-1}, so γ^′′(t)=(k1)tk2ρ(t)+tk1ρ(t)\hat{\gamma}^{\prime\prime}(t)=(k-1)t^{k-2}\rho(t)+t^{k-1}\rho^{\prime}(t) and |ρ(t)|=k|\rho(t)|=k. Then

κ(t)=1k3|t|k1det(ρ(t),ρ(t)).\kappa(t)=\dfrac{1}{k^{3}|t|^{k-1}}\det(\rho(t),\rho^{\prime}(t)).

Consequently,

κ~(t)=|t|k1κ(t)=1k3det(ρ(t),ρ(t)).\tilde{\kappa}(t)=|t|^{k-1}\kappa(t)=\dfrac{1}{k^{3}}\det(\rho(t),\rho^{\prime}(t)).

On the other hand, since κ1(t)=𝒆(t)𝒏(t)\kappa_{1}(t)=\boldsymbol{e}^{\prime}(t)\cdot\boldsymbol{n}(t), where 𝒆(t)=ρ(t)/|ρ(t)|=ρ(t)/k\boldsymbol{e}(t)=\rho(t)/|\rho(t)|=\rho(t)/k and 𝒏(t)\boldsymbol{n}(t) is the π/2\pi/2-counterclockwise rotation of 𝒆(t)\boldsymbol{e}(t), and the dot ‘\cdot’ denotes the canonical inner product of 𝑹2\boldsymbol{R}^{2}, it holds that:

κ1(t)=1kρ(t)𝒏(t)=1k2det(ρ(t),ρ(t)).\kappa_{1}(t)=\frac{1}{k}\rho^{\prime}(t)\cdot\boldsymbol{n}(t)=\frac{1}{k^{2}}\det(\rho(t),\rho^{\prime}(t)).

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 f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be a frontal and ν\nu a unit normal vector field of ff. Let γ:(𝑹,0)(𝑹2,0)\gamma\colon(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a curve. We set γ^=fγ\hat{\gamma}=f\circ\gamma. We assume there exists kk such that γ^=tk1ρ\hat{\gamma}^{\prime}=t^{k-1}\rho (ρ(0)0\rho(0)\neq 0). The geodesic curvature κg\kappa_{g}, the normal curvature κn\kappa_{n} and the geodesic torsion τg\tau_{g} are defined by

κg=det(γ^,γ^′′,ν)|γ^|3,κn=γ^′′,ν|γ^|2,τg=det(γ^,ν,ν)|γ^|2\kappa_{g}=\dfrac{\det(\hat{\gamma}^{\prime},\hat{\gamma}^{\prime\prime},\nu)}{|\hat{\gamma}^{\prime}|^{3}},\ \kappa_{n}=\dfrac{\left\langle{\hat{\gamma}^{\prime\prime}},{\nu}\right\rangle}{|\hat{\gamma}^{\prime}|^{2}},\ \tau_{g}=\dfrac{\det(\hat{\gamma}^{\prime},\nu,\nu^{\prime})}{|\hat{\gamma}^{\prime}|^{2}} (3.5)

on regular points (see [1, page 261]). These curvatures can be unbounded near singular points. Indeed, it holds that

κg=1|t|k1det(ρ,ρ,ν)|ρ|3,κn=1tk1ρ,ν|ρ|2,τg=1tk1det(ρ,ν,ν)|ρ|2.\kappa_{g}=\dfrac{1}{|t|^{k-1}}\dfrac{\det(\rho,\rho^{\prime},\nu)}{|\rho|^{3}},\ \kappa_{n}=\dfrac{1}{t^{k-1}}\dfrac{\left\langle{\rho^{\prime}},{\nu}\right\rangle}{|\rho|^{2}},\ \tau_{g}=\dfrac{1}{t^{k-1}}\dfrac{\det(\rho,\nu,\nu^{\prime})}{|\rho|^{2}}. (3.6)

One can easily see that

κ~g=|s~|k1κg,κ~n=s~k1κn,τ~g=s~k1τg\tilde{\kappa}_{g}=|\tilde{s}|^{k-1}\,\kappa_{g},\ \tilde{\kappa}_{n}=\tilde{s}^{k-1}\,\kappa_{n},\ \tilde{\tau}_{g}=\tilde{s}^{k-1}\,\tau_{g} (3.7)

are CC^{\infty} functions, where s~\tilde{s} is the function given by (3.2) for γ^\hat{\gamma}. We call κ~g,κ~n,τ~g\tilde{\kappa}_{g},\tilde{\kappa}_{n},\tilde{\tau}_{g} normalized geodesic curvature, normal curvature and geodesic torsion of γ~\tilde{\gamma}, respectively. These satisfy:

Lemma 3.3.

It holds that

κ~g\displaystyle\tilde{\kappa}_{g} =1k2k!1/kdet(γ^(k),γ^(k+1),ν)|γ^(k)|2+1/k,\displaystyle=\dfrac{1}{k^{2}\,k!^{-1/k}}\dfrac{\det\left(\hat{\gamma}^{(k)},\hat{\gamma}^{(k+1)},\nu\right)}{|\hat{\gamma}^{(k)}|^{2+1/k}}, (3.8)
κ~n\displaystyle\tilde{\kappa}_{n} =1k2k!1/kγ^(k+1),ν|γ^(k)|1+1/k,\displaystyle=\dfrac{1}{k^{2}\,k!^{-1/k}}\dfrac{\left\langle{\hat{\gamma}^{(k+1)}},{\nu}\right\rangle}{|\hat{\gamma}^{(k)}|^{1+1/k}}, (3.9)
τ~g\displaystyle\tilde{\tau}_{g} =1kk!1/kdet(γ^(k),ν,ν)|γ^(k)|1+1/k\displaystyle=\dfrac{1}{k\,k!^{-1/k}}\dfrac{\det(\hat{\gamma}^{(k)},\nu,\nu^{\prime})}{|\hat{\gamma}^{(k)}|^{1+1/k}} (3.10)

at t=0t=0.

Proof.

Since γ^(t)=tk1ρ(t)\hat{\gamma}^{\prime}(t)=t^{k-1}\rho(t), then ρ(0)=γ^(k)(0)(k1)!\rho(0)=\frac{\hat{\gamma}^{(k)}(0)}{(k-1)!}, ρ(0)=γ^(k+1)(0)k!\rho^{\prime}(0)=\frac{\hat{\gamma}^{(k+1)}(0)}{k!} and ρ0=|ρ(0)|=|γ^(k)(0)|(k1)!\rho_{0}=|\rho(0)|=\frac{|\hat{\gamma}^{(k)}(0)|}{(k-1)!}. Therefore, it holds that

s~k1=tk1(ρ0(k1)/kk(k1)/k+tO(t))\tilde{s}^{k-1}=t^{k-1}\left(\dfrac{\rho_{0}^{(k-1)/k}}{k^{(k-1)/k}}+tO(t)\right)

where O(t)O(t) is a smooth function of tt. Thus by (3.6) and (3.7), we get at t=0t=0:

κ~g\displaystyle\tilde{\kappa}_{g} =ρ0k1kk1/kkdet(ρ,ρ,ν)ρ03=k1/kkdet(ρ,ρ,ν)ρ02+1/k\displaystyle=\frac{\rho_{0}^{\frac{k-1}{k}}k^{1/k}}{k}\dfrac{\det(\rho,\rho^{\prime},\nu)}{\rho_{0}^{3}}=\dfrac{k^{1/k}}{k}\,\dfrac{\det(\rho,\rho^{\prime},\nu)}{\rho_{0}^{2+1/k}}
=k1/k(k1)!2+1/kk(k1)!k!det(γ^(k),γ^(k+1),ν)|γ^(k)|2+1/k=k!1/kk2det(γ^(k),γ^(k+1),ν)|γ^(k)|2+1/k,\displaystyle=\dfrac{k^{1/k}(k-1)!^{2+1/k}}{k\,(k-1)!\,k!}\dfrac{\det(\hat{\gamma}^{(k)},\hat{\gamma}^{(k+1)},\nu)}{|\hat{\gamma}^{(k)}|^{2+1/k}}=\dfrac{k!^{1/k}}{k^{2}}\dfrac{\det(\hat{\gamma}^{(k)},\hat{\gamma}^{(k+1)},\nu)}{|\hat{\gamma}^{(k)}|^{2+1/k}},
κ~n\displaystyle\tilde{\kappa}_{n} =ρ0(k1)/kk(k1)/kρ,νρ02=k1/kkρ,νρ01+1/k\displaystyle=\frac{\rho_{0}^{(k-1)/k}}{k^{(k-1)/k}}\,\frac{\left\langle{\rho^{\prime}},{\nu}\right\rangle}{\rho_{0}^{2}}=\frac{k^{1/k}}{k}\frac{\left\langle{\rho^{\prime}},{\nu}\right\rangle}{\rho_{0}^{1+1/k}}
=k1/k(k1)!1+1/kkk!γ^(k+1),ν|γ^(k)|1+1/k=k!1/kk2γ^(k+1),ν|γ^(k)|1+1/k\displaystyle=\frac{k^{1/k}(k-1)!^{1+1/k}}{k\,k!}\frac{\left\langle{\hat{\gamma}^{(k+1)}},{\nu}\right\rangle}{|\hat{\gamma}^{(k)}|^{1+1/k}}=\frac{k!^{1/k}}{k^{2}}\frac{\left\langle{\hat{\gamma}^{(k+1)}},{\nu}\right\rangle}{|\hat{\gamma}^{(k)}|^{1+1/k}}

and

τ~g\displaystyle\tilde{\tau}_{g} =ρ0(k1)/kk(k1)/kdet(ρ,ν,ν)ρ02=k1/kkdet(ρ,ν,ν)ρ01+1/k\displaystyle=\frac{\rho_{0}^{(k-1)/k}}{k^{(k-1)/k}}\dfrac{\det(\rho,\nu,\nu^{\prime})}{\rho_{0}^{2}}=\dfrac{k^{1/k}}{k}\dfrac{\det(\rho,\nu,\nu^{\prime})}{\rho_{0}^{1+1/k}}
=k1/k(k1)!1+1/kk(k1)!det(γ^(k),ν,ν)|γ^(k)|1+1/k=k!1/kkdet(γ^(k),ν,ν)|γ^(k)|1+1/k,\displaystyle=\dfrac{k^{1/k}(k-1)!^{1+1/k}}{k(k-1)!}\dfrac{\det(\hat{\gamma}^{(k)},\nu,\nu^{\prime})}{|\hat{\gamma}^{(k)}|^{1+1/k}}=\dfrac{k!^{1/k}}{k}\dfrac{\det(\hat{\gamma}^{(k)},\nu,\nu^{\prime})}{|\hat{\gamma}^{(k)}|^{1+1/k}},

which show the assertion. ∎

Similar with the case of plane curves, these invariants can be interpreted as follows. Under the same assumption above, we set e=ρ/|ρ|e=\rho/|\rho|, ν=ν(γ^)\nu=\nu(\hat{\gamma}) and b=e×νb=-e\times\nu. Then {e,ν,b}\{e,\nu,b\} is a frame along γ^\hat{\gamma}. We define κ1,κ2,κ3\kappa_{1},\kappa_{2},\kappa_{3} by

(ebν)=(0κ1κ2κ10κ3κ2κ30)(ebν),{\begin{pmatrix}e^{\prime}\\ b^{\prime}\\ \nu^{\prime}\end{pmatrix}}={\begin{pmatrix}0&\kappa_{1}&\kappa_{2}\\ -\kappa_{1}&0&\kappa_{3}\\ -\kappa_{2}&-\kappa_{3}&0\end{pmatrix}}{\begin{pmatrix}e\\ b\\ \nu\end{pmatrix}}, (3.11)

where =d/dt{}^{\prime}=d/dt is the differentiation by the 1/k1/k-arc-length parameter. With the above notation, we get the following:

Proposition 3.4.

If tt is the 1/k1/k-arc-length parameter, then

κ1=kκ~g,κ2=kκ~nandκ3=kτ~g\kappa_{1}=k\tilde{\kappa}_{g},\ \kappa_{2}=k\tilde{\kappa}_{n}\ \ \text{and}\ \ \kappa_{3}=k\tilde{\tau}_{g}

holds, for any tt.

Proof.

The 1/k1/k-arc-length parameter tt satisfies |γ^(t)|=k|tk1||\hat{\gamma}^{\prime}(t)|=k|t^{k-1}|. Then |ρ(t)|=k|\rho(t)|=k and e=ρ/ke=\rho/k. So, putting s~=t\tilde{s}=t at (3.7) and using (3.6) , it holds that

κ1\displaystyle\kappa_{1} =e,b=1k2det(ρ,ρ,ν)=kκ~g,\displaystyle=\left\langle{e^{\prime}},{b}\right\rangle=\dfrac{1}{k^{2}}\det(\rho,\rho^{\prime},\nu)=k\tilde{\kappa}_{g},
κ2\displaystyle\kappa_{2} =e,ν=1kρ,ν=kκ~n,\displaystyle=\left\langle{e^{\prime}},{\nu}\right\rangle=\dfrac{1}{k}\left\langle{\rho^{\prime}},{\nu}\right\rangle=k\tilde{\kappa}_{n}\,,
κ3\displaystyle\kappa_{3} =ν,b=1kdet(ρ,ν,ν)=kτ~g.\displaystyle=-\left\langle{\nu^{\prime}},{b}\right\rangle=\dfrac{1}{k}\det(\rho,\nu,\nu^{\prime})=k\tilde{\tau}_{g}.

Thus the assertion holds. ∎

3.3. Behaviors of κg,κn\kappa_{g},\kappa_{n} and τg\tau_{g} passing through an mm-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 mm-type edge, concluding on boundedness. Describing the condition, we use the curvature of such curve. Let f:(𝑹2,0)(𝑹3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an mm-type edge, m2m\geq 2, and γ:(𝑹,0)(𝑹2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a regular curve such that γ(0)\gamma^{\prime}(0) is a null vector of ff at 0. Let (u,v)(u,v) be a coordinate system which gives the form (2.5), and γ~(t)=(u(t),v(t))\tilde{\gamma}(t)=(u(t),v(t)) be a parametrization of γ\gamma where the coordinate system on the target space is (u,v)(u,v), and the orientation of γ~\tilde{\gamma} agrees the direction of vv at 0. Since such coordinate system is unique (unique up to (u,v)(u,v)(u,v)\mapsto(u,-v) if mm is even), the order of contact of γ~\tilde{\gamma} with the vv-axis at 0 and the curvature κ~\tilde{\kappa} of γ~\tilde{\gamma} is well-defined as a curve on ff. We call such order of contact the order of contact with the normalized null direction, and we call κ~\tilde{\kappa} the curvature written in the normal form. If γ~(t)=(tlc(t),t)\tilde{\gamma}(t)=(t^{l}c(t),t) (c(0)0c(0)\neq 0), then the order of contact with the normalized null direction is ll, and κ~(l2)(0)=l!c(0)\tilde{\kappa}^{(l-2)}(0)=-l!c(0) and κ~(l1)(0)=(l+1)!c(0)\tilde{\kappa}^{(l-1)}(0)=-(l+1)!c^{\prime}(0) hold.

Theorem 3.5.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an mm-type edge, m2m\geq 2, and γ:(𝐑,0)(𝐑2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a regular curve with order of contact l2l\geq 2 with the null direction of ff at 0 and κ~\tilde{\kappa} the curvature of γ\gamma written in the normal form of ff. Then, it holds that:

(1) The case lml\geq m.
For κg\kappa_{g},

  • if m<l2mm<l\leq 2m, then ordκg=l2m\operatorname{ord}\kappa_{g}=l-2m;

  • if l>2ml>2m, then ordκg1\operatorname{ord}\kappa_{g}\geq 1, and ordκg=1\operatorname{ord}\kappa_{g}=1 is equivalent to

    {(l1)!κt(0)ωm,m+1(0)m!(m+1)!κ~(l2)(0)0(if l=2m+1),κt(0)ωm,m+1(0)0(if l>2m+1);\begin{dcases}(l-1)!\kappa_{t}(0)\omega_{m,m+1}(0)-m!(m+1)!\tilde{\kappa}^{(l-2)}(0)\neq 0&(\text{if $l=2m+1$}),\\ \kappa_{t}(0)\omega_{m,m+1}(0)\neq 0&(\text{if $l>2m+1$});\end{dcases}
  • if l=ml=m, then ordκg1m\operatorname{ord}\kappa_{g}\geq 1-m, and ordκg=1m\operatorname{ord}\kappa_{g}=1-m if and only if κ~(l1)(0)0\tilde{\kappa}^{(l-1)}(0)\neq 0.

For κn\kappa_{n}, it holds that ordκn1m\operatorname{ord}\kappa_{n}\geq 1-m, and ordκn=1m\operatorname{ord}\kappa_{n}=1-m if and only if ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0. For τg\tau_{g},

  • if l<2ml<2m, then ordτgl2m+1\operatorname{ord}\tau_{g}\geq l-2m+1, and ordτg=l2m+1\operatorname{ord}\tau_{g}=l-2m+1 is equivalent to

    {ωm,m+1(0)0(if l<2m1),m(l1)!κt(0)+(m1)!2κ~(l2)(0)ωm,m+1(0)0(if l=2m1);\begin{dcases}\omega_{m,m+1}(0)\neq 0&(\text{if $l<2m-1$}),\\ m(l-1)!\kappa_{t}(0)+{(m-1)!^{2}}\,\tilde{\kappa}^{(l-2)}(0)\,\omega_{m,m+1}(0)\neq 0&(\text{if $l=2m-1$});\end{dcases}
  • if l2ml\geq 2m, then ordτg0\operatorname{ord}\tau_{g}\geq 0, and ordτg=0\operatorname{ord}\tau_{g}=0 if and only if κt(0)0\kappa_{t}(0)\neq 0.

(2) The case m/2<l<mm/2<l<m.
For this case, it holds that ordκg=m2l\operatorname{ord}\kappa_{g}=m-2l, ordκnm2l+1\operatorname{ord}\kappa_{n}\geq m-2l+1, and ordκn=m2l+1\operatorname{ord}\kappa_{n}=m-2l+1 is equivalent to

{ωm,m+1(0)0(if l>(m+1)/2),(m+1)!(ml+1)κν(0)(κ~(l2))2(0)+2l!2ωm,m+1(0)0(if l=(m+1)/2).\begin{dcases}\omega_{m,m+1}(0)\neq 0&(\text{if $l>(m+1)/2$}),\\ (m+1)!(m-l+1)\kappa_{\nu}(0)(\tilde{\kappa}^{(l-2)})^{2}(0)+2l!^{2}\omega_{m,m+1}(0)\neq 0&(\text{if $l=(m+1)/2$}).\end{dcases}

For τg\tau_{g}, it holds that ordτg1l\operatorname{ord}\tau_{g}\geq 1-l, and ordτg=1l\operatorname{ord}\tau_{g}=1-l is equivalent to ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0.

(3) The case lm/2l\leq m/2.
In this case, it holds that ordκg0\operatorname{ord}\kappa_{g}\geq 0, and ordκg=0\operatorname{ord}\kappa_{g}=0 is equivalent to

{κs(0)0(if l<m/2),m!κs(0)(κ~(l2))2(0)+2l!20(if l=m/2).\begin{dcases}\kappa_{s}(0)\neq 0&(\text{if $l<m/2$}),\\ m!\kappa_{s}(0)(\tilde{\kappa}^{(l-2)})^{2}(0)+2l!^{2}\neq 0&(\text{if $l=m/2$}).\end{dcases}

For κn\kappa_{n} it holds that ordκn0\operatorname{ord}\kappa_{n}\geq 0, and ordκn=0\operatorname{ord}\kappa_{n}=0 if and only if κn(0)0\kappa_{n}(0)\neq 0. For τg\tau_{g}, it holds that ordτg1l\operatorname{ord}\tau_{g}\geq 1-l, and ordτg=1l\operatorname{ord}\tau_{g}=1-l if and only if ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0.

If mm is even and (u,v)(u,v) be a coordinate system which gives the form (2.5), then (u,v)(u,-v) also gives (2.5). In this case, changing (u,v)(u,v) to (u,v)(u,-v), the signs of κ~\tilde{\kappa} and ωm,m+1\omega_{m,m+1} reverse, and them of κt\kappa_{t} and κs\kappa_{s} do not change. So, when mm is even, neither the condition

(l1)!κt(0)ωm,m+1(0)m!(m+1)!κ~(l2)(0)\displaystyle(l-1)!\kappa_{t}(0)\omega_{m,m+1}(0)-m!(m+1)!\tilde{\kappa}^{(l-2)}(0) 0,\displaystyle\neq 0,
m(l1)!κt(0)+(m+1)!2κ~(l2)(0)ωm,m+1(0)\displaystyle m(l-1)!\kappa_{t}(0)+(m+1)!^{2}\,\tilde{\kappa}^{(l-2)}(0)\,\omega_{m,m+1}(0) 0nor\displaystyle\neq 0\quad\text{nor}
m!κs(0)(κ~(l2))2(0)+2l!2\displaystyle m!\kappa_{s}(0)(\tilde{\kappa}^{(l-2)})^{2}(0)+2l!^{2} 0\displaystyle\neq 0

changes under the coordinate change (u,v)(u,v) to (u,v)(u,-v).

Proof.

Let γ^=fγ\hat{\gamma}=f\circ\gamma. One can assume that ff is given by the form (2.5) and, since v\partial v is a null vector of ff, one can take γ(t)=(x(t),t)\gamma(t)=(x(t),t), with x(0)=x(0)=0x(0)=x^{\prime}(0)=0. Then x(t)x(t) is of order ll and we set γ(t)=(tlc(t),t)\gamma(t)=(t^{l}c(t),t) (c(0)0c(0)\neq 0).

In the normal form (2.5), since bm(0,0)=0b_{m}(0,0)=0, further we may assume ff is given by f(u,v)=(u,u2a(u)/2+vm/m!,u2b0(u)/2+(vm/m!)(ubm1(u)+vbm2(u,v)))f(u,v)=(u,u^{2}a(u)/2+v^{m}/m!,u^{2}b_{0}(u)/2+(v^{m}/m!)(ub_{m1}(u)+vb_{m2}(u,v))). We recall that κs(0)=a(0),κν(0)=b(0)\kappa_{s}(0)=a(0),\kappa_{\nu}(0)=b(0) and κt(0)=(bm)u(0,0)\kappa_{t}(0)=(b_{m})_{u}(0,0). Furthermore, it holds that κt(0)=bm1(0)\kappa_{t}(0)=b_{m1}(0), ωm,m+1=(m+1)bm2(0,0)\omega_{m,m+1}=(m+1)b_{m2}(0,0), κ~(l2)(0)=l!c(0)\tilde{\kappa}^{(l-2)}(0)=-l!c(0) and κ~(l1)(0)=(l+1)!c(0)\tilde{\kappa}^{(l-1)}(0)=-(l+1)!c^{\prime}(0). We set φ\varphi by fv=vm1φ/(m1)!f_{v}=v^{m-1}\varphi/(m-1)!. Then ν~2=fu×φ\tilde{\nu}_{2}=f_{u}\times\varphi gives a non-zero normal vector field to ff.

(1). Assume lml\geq m. By (2.5) we get γ^=tmρ~\hat{\gamma}=t^{m}\tilde{\rho}, where

ρ~(t)\displaystyle\tilde{\rho}(t) =(tlmc(t),g2(t),tg3(t)),\displaystyle=(t^{l-m}c(t),g_{2}(t),tg_{3}(t)),
g2(t)\displaystyle g_{2}(t) =m!t2lma(t)c(t)2+22m!,\displaystyle=\dfrac{m!t^{2l-m}a(t)c(t)^{2}+2}{2m!},
g3(t)\displaystyle g_{3}(t) =m!t2lm1b0(t)c(t)2+2bm2(t)+2tl1bm1(t)c(t)2m!.\displaystyle=\dfrac{m!t^{2l-m-1}b_{0}(t)c(t)^{2}+2b_{m2}(t)+2t^{l-1}b_{m1}(t)c(t)}{2m!}.

Then γ^=tm1ρ\hat{\gamma}=t^{m-1}\rho, where ρ=mρ~+tρ~\rho=m\tilde{\rho}+t\tilde{\rho}^{\prime}. Note that ρ(0)0\rho(0)\neq 0. Setting ν2(t)=ν~2(γ(t))\nu_{2}(t)=\tilde{\nu}_{2}(\gamma(t)), we can show that ν2(t)=(tmd(t),te(t),1)\nu_{2}(t)=(t^{m}d(t),te(t),1), where

d(t)\displaystyle d(t) =12mm!(2mbm1+2mt2lmabm1c2m!2mtlmb0cm!\displaystyle=\dfrac{1}{2mm!}\bigg{(}-2mb_{m1}+2mt^{2l-m}ab_{m1}c^{2}m!-2mt^{l-m}b_{0}cm! (3.12)
+2t1+lmabm2cm!+2mt1+lmabm2cm!+mt3lmbm1c3m!a\displaystyle\hskip 56.9055pt+2t^{1+l-m}ab_{m2}cm!+2mt^{1+l-m}ab_{m2}cm!+mt^{3l-m}b_{m1}c^{3}m!a^{\prime}
+t1+2lmbm2c2m!a+mt1+2lmbm2c2m!amt2lmc2m!b0\displaystyle\hskip 56.9055pt+t^{1+2l-m}b_{m2}c^{2}m!a^{\prime}+mt^{1+2l-m}b_{m2}c^{2}m!a^{\prime}-mt^{2l-m}c^{2}m!b_{0}^{\prime}
+2t2+lmacm!bm2,v+t2+2lmc2m!abm2,v2mtlcbm12mtbm2,u),\displaystyle\hskip 56.9055pt+2t^{2+l-m}acm!b_{m2,v}+t^{2+2l-m}c^{2}m!a^{\prime}b_{m2,v}-2mt^{l}cb^{\prime}_{m1}-2mtb_{m2,u}\bigg{)},
e(t)\displaystyle e(t) =1m(mtl1bm1c+(1+m)bm2+tbm2,v).\displaystyle=\dfrac{-1}{m}\Big{(}mt^{l-1}b_{m1}c+(1+m)b_{m2}+tb_{m2,v}\Big{)}.

We abbreviate the variable, namely a=a(t)a=a(t), bm2=bm2(γ(t))b_{m2}=b_{m2}(\gamma(t)) for instance, and (bm2)v=bm2,v(b_{m2})_{v}=b_{m2,v}. Here, we see

g2(0)=1m!,g2(0)=0,g3(0)=bm2(0)m!,d(0)=bm1(0)m!(ifm<l),\displaystyle g_{2}(0)=\dfrac{1}{m!},\ g_{2}^{\prime}(0)=0,\ g_{3}(0)=\dfrac{b_{m2}(0)}{m!},\ d(0)=\dfrac{-b_{m1}(0)}{m!}\ (\text{if}\ m<l),
d(0)=m!b0(0)c(0)bm1(0)m!(ifm=l),e(0)=(m+1)bm2(0)m.\displaystyle d(0)=\dfrac{-m!b_{0}(0)c(0)-b_{m1}(0)}{m!}\ (\text{if}\ m=l),\ e(0)=-\dfrac{(m+1)b_{m2}(0)}{m}.

To see the rational order of the invariants κg,κn\kappa_{g},\kappa_{n}, τg\tau_{g} at 0, we may use ν2(t)\nu_{2}(t) instead of νγ(t)\nu\circ\gamma(t) in (3.6). Since g2(0)=0g_{2}^{\prime}(0)=0, we can write g2=tg~g_{2}^{\prime}=t\tilde{g}. We see

ρ\displaystyle\rho =(ltlmc+tlm+1c,mg2+t2g~2,(m+1)tg3+t2g3),\displaystyle=(lt^{l-m}c+t^{l-m+1}c^{\prime},mg_{2}+t^{2}\tilde{g}_{2},(m+1)tg_{3}+t^{2}g_{3}^{\prime}), (3.13)
ρ\displaystyle\rho^{\prime} ={(l(lm)tlm1c+tlmO(1),tO(1),(m+1)g3+tO(1))(l>m)((m+1)c+tO(1),tO(1),(m+1)g3+tO(1))(m=l),\displaystyle=\begin{dcases}\Big{(}l(l-m)t^{l-m-1}c+t^{l-m}O(1),\\ \hskip 56.9055pttO(1),(m+1)g_{3}+tO(1)\Big{)}\quad(l>m)\\ \Big{(}(m+1)c^{\prime}+tO(1),\\ \hskip 56.9055pttO(1),(m+1)g_{3}+tO(1)\Big{)}\ (m=l),\end{dcases} (3.14)

where O(1)O(1) means a smooth function depending on tt. Then we see |ρ,ρ,ν2||\rho,\rho^{\prime},\nu_{2}|, where ||=det()|\cdot|=\det(\cdot) is, for l>ml>m:

|ltlmc+tlm+1O(1)l(lm)tlm1c+tlmO(1)tmdmg2+tO(1)tO(1)te(m+1)tg3+t2O(1)(m+1)g3+tO(1)1|.{\begin{vmatrix}lt^{l-m}c+t^{l-m+1}O(1)&l(l-m)t^{l-m-1}c+t^{l-m}O(1)&t^{m}d\\ mg_{2}+tO(1)&tO(1)&te\\ (m+1)tg_{3}+t^{2}O(1)&(m+1)g_{3}+tO(1)&1\end{vmatrix}}. (3.15)

If 2ml+1>02m-l+1>0, then (3.15) is tlm1A1(t)t^{l-m-1}A_{1}(t), where

A1(0)=|0l(lm)c(0)0mg2(0)000(m+1)g3(0)1|=l(lm)(m1)!c(0).A_{1}(0)={\begin{vmatrix}0&l(l-m)c(0)&0\\ mg_{2}(0)&0&0\\ 0&(m+1)g_{3}(0)&1\end{vmatrix}}=-\dfrac{l(l-m)}{(m-1)!}c(0).

If 2ml+1=02m-l+1=0, then (3.15) is tmA2(t)t^{m}A_{2}(t), where

A2(0)\displaystyle A_{2}(0) =|0l(m+1)c(0)d(0)mg2(0)000(m+1)g3(0)1|=m(m+1)g2(0)(lcdg3)(0)\displaystyle={\begin{vmatrix}0&l(m+1)c(0)&d(0)\\ mg_{2}(0)&0&0\\ 0&(m+1)g_{3}(0)&1\end{vmatrix}}=-m(m+1)g_{2}(0)(lc-dg_{3})(0)
=m+1(m1)!(bm1bm2(m!)2+lc)(0).\displaystyle=-\dfrac{m+1}{(m-1)!}\left(\dfrac{b_{m1}b_{m2}}{(m!)^{2}}+lc\right)(0).

If 2ml+1<02m-l+1<0, then l>ml>m and (3.15) is tmA3(t)t^{m}A_{3}(t), where

A3(0)\displaystyle A_{3}(0) =|00d(0)mg2(0)000(m+1)g3(0)1|=m(m+1)d(0)g2(0)g3(0)\displaystyle={\begin{vmatrix}0&0&d(0)\\ mg_{2}(0)&0&0\\ 0&(m+1)g_{3}(0)&1\end{vmatrix}}=m(m+1)d(0)g_{2}(0)g_{3}(0)
=m(1+m)(m!)3bm1(0)bm2(0).\displaystyle=-\dfrac{m(1+m)}{(m!)^{3}}b_{m1}(0)b_{m2}(0).

If m=lm=l, by (3.13) and (3.14), we see the assertion, once ord|t|m1=m1\operatorname{ord}|t|^{m-1}=m-1 and ord|ρ|3=0\operatorname{ord}|\rho|^{3}=0. These shows the assertion for κg\kappa_{g}.

Since one can easily see that ρ,ν2=(m+1)bm2/m!\left\langle{\rho^{\prime}},{\nu_{2}}\right\rangle=(m+1)b_{m2}/m! at 0, the assertion for κn\kappa_{n} is proved. Next we see |ρ,ν2,ν2||\rho,\nu_{2},\nu_{2}^{\prime}| is

|ltlmc+tlm+1O(1)tmdmtm1d+tmO(1)mg2+tO(1)tee+tO(1)(m+1)tg3+t2O(1)10|.{\begin{vmatrix}lt^{l-m}c+t^{l-m+1}O(1)&t^{m}d&mt^{m-1}d+t^{m}O(1)\\ mg_{2}+tO(1)&te&e+tO(1)\\ (m+1)tg_{3}+t^{2}O(1)&1&0\end{vmatrix}}. (3.16)

If 2ml1>02m-l-1>0, then (3.16) is tlmB1(t)t^{l-m}B_{1}(t), where

B1(0)=|lc(0)00mg2(0)0e(0)010|=lc(0)e(0)=l(m+1)mbm2(0)c(0).B_{1}(0)={\begin{vmatrix}lc(0)&0&0\\ mg_{2}(0)&0&e(0)\\ 0&1&0\end{vmatrix}}=-lc(0)e(0)=\dfrac{l(m+1)}{m}b_{m2}(0)c(0).

If 2ml1=02m-l-1=0, then m<lm<l and (3.16) is tm1B2(t)t^{m-1}B_{2}(t), where

B2(0)\displaystyle B_{2}(0) =|lc(0)0md(0)mg2(0)0e(0)010|=(lce+m2dg2)(0)\displaystyle={\begin{vmatrix}lc(0)&0&md(0)\\ mg_{2}(0)&0&e(0)\\ 0&1&0\end{vmatrix}}=(-lce+m^{2}dg_{2})(0)
=bm1(0)(m1)!2+l(m+1)bm2(0)c(0)m.\displaystyle=-\dfrac{b_{m1}(0)}{(m-1)!^{2}}+\dfrac{l(m+1)b_{m2}(0)c(0)}{m}.

If 2ml1<02m-l-1<0, then m<lm<l, and (3.16) is tm1B3(t)t^{m-1}B_{3}(t), where

B3(0)\displaystyle B_{3}(0) =|00md(0)mg2(0)0e(0)010|=m2d(0)g2(0)=bm1(0)(m1)!2.\displaystyle={\begin{vmatrix}0&0&md(0)\\ mg_{2}(0)&0&e(0)\\ 0&1&0\end{vmatrix}}=m^{2}d(0)g_{2}(0)=-\dfrac{b_{m1}(0)}{(m-1)!^{2}}\,.

This show the assertion for τg\tau_{g}.

(2) and (3). We assume l<ml<m and we shall use the same notation of case (1). Setting ν2(t)=ν~2(γ(t))\nu_{2}(t)=\tilde{\nu}_{2}(\gamma(t)), we can show that ν2(t)=(tld(t),te(t),1)\nu_{2}(t)=(t^{l}d(t),te(t),1), where

d(t)\displaystyle d(t) =12mm!(2mtmlbm1+2mm!tlabm1c22mm!b0c+2m!tabm2c+2mm!tabm2c\displaystyle=\dfrac{1}{2mm!}\bigg{(}-2mt^{m-l}b_{m1}+2mm!t^{l}ab_{m1}c^{2}-2mm!b_{0}c+2m!tab_{m2}c+2mm!tab_{m2}c
+mm!t2lbm1c3a+m!tl+1bm2c2a+mm!tl+1bm2c2amtlc2m!b0\displaystyle\hskip 56.9055pt+mm!t^{2l}b_{m1}c^{3}a^{\prime}+m!t^{l+1}b_{m2}c^{2}a^{\prime}+mm!t^{l+1}b_{m2}c^{2}a^{\prime}-mt^{l}c^{2}m!b_{0}^{\prime}
+2m!t2acbm2,v+m!tl+2c2abm2,v2mtmcbm12mtml+1bm2,u)\displaystyle\hskip 56.9055pt+2m!t^{2}acb_{m2,v}+m!t^{l+2}c^{2}a^{\prime}b_{m2,v}-2mt^{m}cb^{\prime}_{m1}-2mt^{m-l+1}b_{m2,u}\bigg{)}

and ee is the same as in (3.12). We assume lm/2l\leq m/2. Then γ^=tlρ~\hat{\gamma}=t^{l}\tilde{\rho}, where ρ~(t)=(c(t),tlg2(t)\tilde{\rho}(t)=(c(t),t^{l}g_{2}(t), tlg3(t))t^{l}g_{3}(t)) and

g2(t)\displaystyle g_{2}(t) =2tm2l+m!a(t)c(t)22m!,\displaystyle=\dfrac{2t^{m-2l}+m!a(t)c(t)^{2}}{2m!},
g3(t)\displaystyle g_{3}(t) =2tmlbm1(t)c(t)+2tm2l+1bm2(γ(t))+m!b0(t)c(t)22m!.\displaystyle=\dfrac{2t^{m-l}b_{m1}(t)c(t)+2t^{m-2l+1}b_{m2}(\gamma(t))+m!b_{0}(t)c(t)^{2}}{2m!}\,.

Then γ^=tlρ\hat{\gamma}^{\prime}=t^{l}\rho, where ρ=lρ~+tρ~\rho=l\tilde{\rho}+t\tilde{\rho}^{\prime}, with ρ(0)0\rho(0)\neq 0. Since γ^\hat{\gamma} has multiplicity ll, we need replace m1m-1 in equations (3.6) by l1l-1. Here, we see

g2(0)=a(0)c(0)2/2(if l<m/2),g2(0)=a(0)c(0)2/2+1/m!(if m=2l),\displaystyle g_{2}(0)=a(0)c(0)^{2}/2\ (\text{if }l<m/2),\ g_{2}(0)=a(0)c(0)^{2}/2+1/m!\ (\text{if }m=2l),
g3(0)=b0(0)c(0)2/2,e(0)=(1+m)bm2(0)/m.\displaystyle g_{3}(0)=b_{0}(0)c(0)^{2}/2,\ e(0)=-(1+m)b_{m2}(0)/m\,.

To see the order, we may use ν2(t)\nu_{2}(t) instead of νγ(t)\nu\circ\gamma(t) in (3.6). We see

ρ\displaystyle\rho =(lc+tc,tl(2lg2+tg2),tl(2lg3+tg3)),\displaystyle=\Big{(}lc+tc^{\prime},t^{l}(2lg_{2}+tg_{2}^{\prime}),t^{l}(2lg_{3}+tg_{3}^{\prime})\Big{)}, (3.17)
ρ\displaystyle\rho^{\prime} =((l+1)c+tO(1),tl1(2l2g2+tO(1)),tl1(2l2g3+tO(1))).\displaystyle=\Big{(}(l+1)c^{\prime}+tO(1),t^{l-1}(2l^{2}g_{2}+tO(1)),t^{l-1}(2l^{2}g_{3}+tO(1))\Big{)}.

By applying the formula

|x11x12x13kx21x22x23kx31x32x33|=|x11kx12kx13x21x22x23x31x32x33|,{\begin{vmatrix}x_{11}&x_{12}&x_{13}\\ kx_{21}&x_{22}&x_{23}\\ kx_{31}&x_{32}&x_{33}\end{vmatrix}}={\begin{vmatrix}x_{11}&kx_{12}&kx_{13}\\ x_{21}&x_{22}&x_{23}\\ x_{31}&x_{32}&x_{33}\end{vmatrix}},

for k=tl1k=t^{l-1}, we see |ρ,ρ,ν2||\rho,\rho^{\prime},\nu_{2}| is

|lc+tO(1)(l+1)c+tO(1)tldtl(2lg2+tO(1))tl1(2l2g2+tO(1))tetl(2lg3+tO(1))tl1(2l2g3+tO(1))1|\displaystyle{\begin{vmatrix}lc+tO(1)&(l+1)c^{\prime}+tO(1)&t^{l}d\\ t^{l}(2lg_{2}+tO(1))&t^{l-1}(2l^{2}g_{2}+tO(1))&te\\ t^{l}(2lg_{3}+tO(1))&t^{l-1}(2l^{2}g_{3}+tO(1))&1\end{vmatrix}}
=\displaystyle= |lc+tO(1)tl1(l+1)c+tO(1)t2l1dt(2lg2+tO(1))tl1(2l2g2+tO(1))tet(2lg3+tO(1))tl1(2l2g3+tO(1))1|\displaystyle{\begin{vmatrix}lc+tO(1)&t^{l-1}(l+1)c^{\prime}+tO(1)&t^{2l-1}d\\ t(2lg_{2}+tO(1))&t^{l-1}(2l^{2}g_{2}+tO(1))&te\\ t(2lg_{3}+tO(1))&t^{l-1}(2l^{2}g_{3}+tO(1))&1\end{vmatrix}} (3.18)
=\displaystyle= tl1C1(t)(C1(t)=|lc+tO(1)(l+1)c+tO(1)t2l1dt(2lg2+tO(1))2l2g2+tO(1)tet(2lg3+tO(1))2l2g3+tO(1)1|).\displaystyle t^{l-1}C_{1}(t)\quad\left(C_{1}(t)={\begin{vmatrix}lc+tO(1)&(l+1)c^{\prime}+tO(1)&t^{2l-1}d\\ t(2lg_{2}+tO(1))&2l^{2}g_{2}+tO(1)&te\\ t(2lg_{3}+tO(1))&2l^{2}g_{3}+tO(1)&1\end{vmatrix}}\right).

Then C1(0)=2l3c(0)g2(0)C_{1}(0)=2l^{3}c(0)g_{2}(0). This shows the assertion for κg\kappa_{g}. By (3.17), we see ρ,ν2=tl1(2l2g3+tO(1))\left\langle{\rho^{\prime}},{\nu_{2}}\right\rangle=t^{l-1}(2l^{2}g_{3}+tO(1)) and |ρ,ν2,ν2|(0)=l(m+1)c(0)bm2/m|\rho,\nu_{2},\nu_{2}^{\prime}|(0)=l(m+1)c(0)b_{m2}/m. This shows the assertions for κn\kappa_{n} and τg\tau_{g}.

Next we assume l>m/2l>m/2. In this case, γ^=tl(c(t),tmlg2(t),tml+1g3(t))\hat{\gamma}=t^{l}(c(t),t^{m-l}g_{2}(t),t^{m-l+1}g_{3}(t)). We set ρ~(t)=(c(t),tmlg2(t),tml+1g3(t))\tilde{\rho}(t)=(c(t),t^{m-l}g_{2}(t),t^{m-l+1}g_{3}(t)) and

g2(t)\displaystyle g_{2}(t) =2+m!t2lma(t)c(t)22m!,\displaystyle=\dfrac{2+m!\,t^{2l-m}a(t)c(t)^{2}}{2m!},
g3(t)\displaystyle g_{3}(t) =2tl1bm1(t)c(t)+2bm2(γ(t))+m!t2lm1b0(t)c(t)22m!.\displaystyle=\dfrac{2t^{l-1}b_{m1}(t)c(t)+2b_{m2}(\gamma(t))+m!\,t^{2l-m-1}b_{0}(t)c(t)^{2}}{2m!}.

Here, it holds that

g2(0)=1/m!,e(0)=(1+m)bm2(0)/m,\displaystyle g_{2}(0)=1/m!,\ e(0)=-(1+m)b_{m2}(0)/m,
g3(0)=bm2(0)/m!(if 2lm1>0),\displaystyle g_{3}(0)=b_{m2}(0)/m!\ (\text{if }2l-m-1>0),
g3(0)=b0(0)c(0)2/2+bm2(0)/m!(if m=2l1).\displaystyle g_{3}(0)=b_{0}(0)c(0)^{2}/2+b_{m2}(0)/m!\ (\text{if }m=2l-1).

It holds that γ^=tl1ρ\hat{\gamma}^{\prime}=t^{l-1}\rho, with ρ=lρ~+tρ~\rho=l\tilde{\rho}+t\tilde{\rho}^{\prime} and ρ(0)0\rho(0)\neq 0. We see

ρ(t)\displaystyle\rho(t) =(lc+tc,tml(mg2+tg2),tml+1((m+1)g3+tg3)),\displaystyle=\Big{(}lc+tc^{\prime},t^{m-l}(mg_{2}+tg_{2}^{\prime}),t^{m-l+1}((m+1)g_{3}+tg_{3}^{\prime})\Big{)}, (3.19)
ρ(t)\displaystyle\rho^{\prime}(t) =((l+1)c+tO(1),tml1(m(ml)g2+tO(1)),\displaystyle=\Big{(}(l+1)c^{\prime}+tO(1),t^{m-l-1}(m(m-l)g_{2}+tO(1)),
tml((m+1)(ml+1)g3+tO(1))).\displaystyle\hskip 142.26378ptt^{m-l}((m+1)(m-l+1)g_{3}+tO(1))\Big{)}.

By the similar method to (3.18), we see |ρ,ρ,ν2||\rho,\rho^{\prime},\nu_{2}| is

|lc+tO(1)(1+l)c+tO(1)tldtml(mg2+tO(1))tml1(m(ml)g2+tO(1))tetml+1((m+1)g3+tO(1))tml((m+1)(ml+1)g3+tO(1))1|\displaystyle{\begin{vmatrix}lc+tO(1)&(1+l)c^{\prime}+tO(1)&t^{l}d\\ t^{m-l}(mg_{2}+tO(1))&t^{m-l-1}(m(m-l)g_{2}+tO(1))&te\\ t^{m-l+1}((m+1)g_{3}+tO(1))&t^{m-l}((m+1)(m-l+1)g_{3}+tO(1))&1\end{vmatrix}}
=\displaystyle= |lc+tO(1)tml1((1+l)c+tO(1))tm1dt(mg2+tO(1))tml1(m(ml)g2+tO(1))tet2((m+1)g3+tO(1))tml((m+1)(ml+1)g3+tO(1))1|\displaystyle{\begin{vmatrix}lc+tO(1)&t^{m-l-1}((1+l)c^{\prime}+tO(1))&t^{m-1}d\\ t(mg_{2}+tO(1))&t^{m-l-1}(m(m-l)g_{2}+tO(1))&te\\ t^{2}((m+1)g_{3}+tO(1))&t^{m-l}((m+1)(m-l+1)g_{3}+tO(1))&1\end{vmatrix}}
=\displaystyle= tml1C2(t),\displaystyle t^{m-l-1}C_{2}(t),
C2(t)=\displaystyle C_{2}(t)= |lc+tO(1)(1+l)c+tO(1)tm1dt(mg2+tO(1))m(ml)g2+tO(1)tet2((m+1)g3+tO(1))t((m+1)(ml+1)g3+tO(1))1|.\displaystyle{\begin{vmatrix}lc+tO(1)&(1+l)c^{\prime}+tO(1)&t^{m-1}d\\ t(mg_{2}+tO(1))&m(m-l)g_{2}+tO(1)&te\\ t^{2}((m+1)g_{3}+tO(1))&t((m+1)(m-l+1)g_{3}+tO(1))&1\end{vmatrix}}.

Then C2(0)=l(ml)mc(0)g2(0)=l(ml)c(0)/(m1)!C_{2}(0)=l(m-l)mc(0)g_{2}(0)=l(m-l)c(0)/(m-1)! and, replacing m1m-1 by l1l-1 in equations (3.6), this shows the assertion for κg\kappa_{g}. By (3.17), we see ρ,ν2=tmlC3(t)\left\langle{\rho^{\prime}},{\nu_{2}}\right\rangle=t^{m-l}C_{3}(t), where C3(t)=m(ml)g2(t)e2(t)+(m+1)(ml+1)g3(t)+tO(1)C_{3}(t)=m(m-l)g_{2}(t)e_{2}(t)+(m+1)(m-l+1)g_{3}(t)+tO(1). It holds that

C3(0)={(m+1)bm2(0)m!(m<2l1),(m+1)(lb0(0)c(0)22(l!)2+bm2(0)m!)(m=2l1).C_{3}(0)=\begin{dcases}\dfrac{(m+1)b_{m2}(0)}{m!}\ &(m<2l-1),\\ (m+1)\left(\dfrac{lb_{0}(0)c(0)^{2}}{2(l!)^{2}}+\dfrac{b_{m2}(0)}{m!}\right)&(m=2l-1).\end{dcases}

and |ρ,ν2,ν2|(0)=lc(0)e(0)=l(m+1)c(0)bm2(0)/m|\rho,\nu_{2},\nu_{2}^{\prime}|(0)=-lc(0)e(0)=l(m+1)c(0)b_{m2}(0)/m. This shows the assertions for κn\kappa_{n} and τg\tau_{g}. ∎

In particular, we have the following corollary on boundedness directly obtained from Theorem 3.5.

Corollary 3.6.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be an mm-type edge with m2m\geq 2, and γ:(𝐑,0)(𝐑2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a regular curve with order of contact l2l\geq 2 with the null direction of ff at 0.

  1. (1)

    The case lml\geq m. For κg\kappa_{g},

    • if l2ml\geq 2m, then κg\kappa_{g} is bounded at 0;

    • if m<l<2mm<l<2m, then κg\kappa_{g} is unbounded at 0;

    • if m=lm=l and κ~(l1)(0)0\tilde{\kappa}^{(l-1)}(0)\neq 0, then κg\kappa_{g} is unbounded at 0.

    For κn\kappa_{n}, if ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0, then κn\kappa_{n} is unbounded at 0. For τg\tau_{g},

    • if ml<2m1m\leq l<2m-1 and ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0, then τg\tau_{g} is unbounded at 0;

    • if l=2m1l=2m-1 and m(l1)!κt(0)+(m1)!2κ~(l2)(0)ωm,m+1(0)0m(l-1)!\kappa_{t}(0)+{(m-1)!^{2}}\,\tilde{\kappa}^{(l-2)}(0)\,\omega_{m,m+1}(0)\neq 0, then τg\tau_{g} is bounded at 0;

    • if l>2m1l>2m-1, then τg\tau_{g} is bounded at 0.

  2. (2)

    The case m/2<l<mm/2<l<m. In this case, κg\kappa_{g} is unbounded at 0. If l=(m+1)/2l=(m+1)/2, then κn\kappa_{n} is bounded at 0. If m>l>(m+1)/2m>l>(m+1)/2 and ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0, then κn\kappa_{n} is unbounded at 0. If ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0, then τg\tau_{g} is unbounded at 0.

  3. (3)

    The case lm/2l\leq m/2. In this case, κg\kappa_{g} and κn\kappa_{n} are bounded at 0. If ωm,m+1(0)0\omega_{m,m+1}(0)\neq 0, then τg\tau_{g} is unbounded at 0.

We consider the case that f:(𝑹2,0)(𝑹3,0)f\colon(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) is a cuspidal edge. By definition, it is a (2,3)(2,3)-edge, in particular, a 22-type edge. Then by Theorem 3.5, the following assertion holds.

Corollary 3.7.

Let f:(𝐑2,0)(𝐑3,0)f:(\boldsymbol{R}^{2},0)\to(\boldsymbol{R}^{3},0) be a cuspidal edge, and let γ:(𝐑,0)(𝐑2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a regular curve with order of contact l2l\geq 2 with the null direction of ff at 0 and κ~\tilde{\kappa} the curvature of γ\gamma written in the normal form of ff. Then, it holds that:

For κg\kappa_{g},

  • if l=2l=2, then ordκg1\operatorname{ord}\kappa_{g}\geq-1, and ordκg=1\operatorname{ord}\kappa_{g}=-1 if and only if κ~(l1)(0)0\tilde{\kappa}^{(l-1)}(0)\neq 0.

  • if l=3l=3 or 44, then ordκg=l4\operatorname{ord}\kappa_{g}=l-4;

  • if l5l\geq 5, then ordκg1\operatorname{ord}\kappa_{g}\geq 1, and ordκg=1\operatorname{ord}\kappa_{g}=1 is equivalent to

    {(l1)!κt(0)ω2,3(0)12κ~(l2)(0)0(if l=5),κt(0)ω2,3(0)0(if l>5);\begin{dcases}(l-1)!\kappa_{t}(0)\omega_{2,3}(0)-12\tilde{\kappa}^{(l-2)}(0)\neq 0&(\text{if $l=5$}),\\ \kappa_{t}(0)\omega_{2,3}(0)\neq 0&(\text{if $l>5$});\end{dcases}

For κn\kappa_{n}, it holds that ordκn=1\operatorname{ord}\kappa_{n}=-1.

For τg\tau_{g},

  • if l=2l=2 or 33, then ordτgl3\operatorname{ord}\tau_{g}\geq l-3, and ordτg=l3\operatorname{ord}\tau_{g}=l-3 is equivalent to

    {ω2,3(0)0(if l<3),2(l1)!κt(0)+κ~(l2)(0)ω2,3(0)0(if l=3);\begin{dcases}\omega_{2,3}(0)\neq 0&(\text{if $l<3$}),\\ 2(l-1)!\kappa_{t}(0)+\tilde{\kappa}^{(l-2)}(0)\,\omega_{2,3}(0)\neq 0&(\text{if $l=3$});\end{dcases}
  • if l4l\geq 4, then ordτg0\operatorname{ord}\tau_{g}\geq 0, and ordτg=0\operatorname{ord}\tau_{g}=0 if and only if κt(0)0\kappa_{t}(0)\neq 0.

Proof.

Since ω2,3\omega_{2,3} corresponds to the cuspidal curvature κc\kappa_{c} and it does not vanish at 0 ([12, Proposition 3.11]), we have the assertion by Theorem 3.5. ∎

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. (1)

    For the geodesic curvature κg\kappa_{g},

    • if l4l\geq 4, then κg\kappa_{g} is bounded at 0;

    • if l=3l=3, then κg\kappa_{g} is unbounded at 0;

    • if l=2l=2 and κ~(0)0\tilde{\kappa}^{\prime}(0)\neq 0, then κg\kappa_{g} is unbounded at 0.

  2. (2)

    The normal curvature κn\kappa_{n} is unbounded at 0.

  3. (3)

    For the geodesic torsion τg\tau_{g},

    • if l=2l=2, then τg\tau_{g} is unbounded at 0;

    • if l=3l=3 and 4κt(0)+κ~(0)κc(0)04\kappa_{t}(0)+\,\tilde{\kappa}^{\prime}(0)\,\kappa_{c}(0)\neq 0, then τg\tau_{g} is bounded at 0;

    • if l4l\geq 4, then τg\tau_{g} is bounded at 0,

    where κc\kappa_{c} is the cuspidal curvature (cf. [12]) corresponding to ω2,3\omega_{2,3}.

Note that ordκg1\operatorname{ord}\kappa_{g}\geq-1 for l2l\geq 2 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 f=(u,v22,v5)f=(u,\frac{v^{2}}{2},v^{5}) and γ(t)=(t4,t)\gamma(t)=(t^{4},t), it holds that m=2,l=4m=2,l=4, ordκg=0,ordκn=1,ordτg=3\operatorname{ord}\kappa_{g}=0,\operatorname{ord}\kappa_{n}=1,\operatorname{ord}\tau_{g}=3 (see Figure 1). Thus these three invariants are bounded at 0 (cf. Corollary 3.6). For the cuspidal edge f(u,v)=(u,v2,v3)f(u,v)=(u,v^{2},v^{3}) and the same γ\gamma, we see that κg\kappa_{g} and τg\tau_{g} are bounded, but κn\kappa_{n} is unbounded at 0 (cf. Corollary 3.8). Figure 2 shows the graphs of these invariants near 0.

Refer to caption
Refer to caption
Refer to caption
Figure 1. The graphs of κg\kappa_{g} (left), κn\kappa_{n} (middle) and τg\tau_{g} (right) of the curve γ^(t)=f(γ(t))\hat{\gamma}(t)=f(\gamma(t)), where f=(u,v22,v5)f=(u,\frac{v^{2}}{2},v^{5}) and γ(t)=(t4,t)\gamma(t)=(t^{4},t).
Refer to caption
Refer to caption
Refer to caption
Figure 2. The graphs of κg\kappa_{g} (left), κn\kappa_{n} (middle) and τg\tau_{g} (right) of the curve γ^(t)=f(γ(t))\hat{\gamma}(t)=f(\gamma(t)), where f(u,v)=(u,v2,v3)f(u,v)=(u,v^{2},v^{3}) and γ(t)=(t4,t)\gamma(t)=(t^{4},t).

Acknowledgements

The authors thank the referee, Yuki Hattori and Atsufumi Honda for valuable comments and suggestions.

Appendix A Generalized biases for plane curve

Let γ:(𝑹,0)(𝑹2,0)\gamma:(\boldsymbol{R},0)\to(\boldsymbol{R}^{2},0) be a curve-germ of (m,n)(m,n)-type which is given by the form (2.4) in the xyxy-plane (𝑹2,0)(\boldsymbol{R}^{2},0). The terms aia_{i} (i=2,,n/m)(i=2,\ldots,\lfloor n/m\rfloor) measures the bias of γ\gamma near singular point. We call ai+1a_{i+1} the (m,im)(m,im)-bias (i=2,,n/m)(i=2,\ldots,\lfloor n/m\rfloor) of γ\gamma at 0, and it is denoted by βm,im\beta_{m,im}. We call b(0)b(0) is called the (m,n)(m,n)-cuspidal curvature as in [7], and it is denoted by rm,nr_{m,n}.

If mm and nn are even, then it is a half part of a curve of (m/2,n/2)(m/2,n/2)-type, we consider the following cases: (1) both m,nm,n are odd, (2) mm is odd and nn is even, and (3) mm is even and nn is odd. Moreover, let aka_{k} denotes the first non-zero term of aia_{i} (i=2,,n/m)(i=2,\ldots,\lfloor n/m\rfloor). We consider the case (1)(1) and (2)(2). Then γ\gamma passes through the origin tangent to the xx-axis. In the case (1)(1), if kk is odd, it also passes across the xx-axis. If kk is even, it approaches to the origin from one side of xx-axis and goes away into the same side of xx-axis, and if there does not exist such kk (namely, the bias is zero), it passes through the xx-axis. In the case (2)(2), if the bias is zero, it approaches to the origin from one side of xx-axis and goes away into the same side of xx-axis. Figure 3 shows the images of the curves γ1:t(t3,a1t6+a2t9+t11)\gamma_{1}:t\mapsto(t^{3},a_{1}t^{6}+a_{2}t^{9}+t^{11}) with (a1,a2)=(1,0),(0,1),(0,0)(a_{1},a_{2})=(1,0),(0,1),(0,0) from left to right. Figure 4 shows the images of the curves γ2:t(t3,a1t6+a2t9+t14)\gamma_{2}:t\mapsto(t^{3},a_{1}t^{6}+a_{2}t^{9}+t^{14}) with (a1,a2)=(1,0),(0,1),(0,0)(a_{1},a_{2})=(1,0),(0,1),(0,0) from left to right.

Refer to caption
Refer to caption
Refer to caption
Figure 3. The images of the curves γ1\gamma_{1}
Refer to caption
Refer to caption
Refer to caption
Figure 4. The images of the curves γ2\gamma_{2}

We consider the case (3)(3). Then γ\gamma approaches the origin from a direction of the xx-axis and making a cusp, and back to the same direction. If kk is both odd and even, it approaches to the origin from one side of xx-axis and goes away into the same side of xx-axis. If the bias is zero, it passes through the xx-axis. Figure 5 shows the images of the curves γ3:t(t4,a1t8+a2t12+t13)\gamma_{3}:t\mapsto(t^{4},a_{1}t^{8}+a_{2}t^{12}+t^{13}) with (a1,a2)=(1,0),(0,1),(0,0)(a_{1},a_{2})=(1,0),(0,1),(0,0) from left to right.

Refer to caption
Refer to caption
Refer to caption
Figure 5. The images of the curves γ3\gamma_{3}
Example A.1.

Let γ\gamma be a curve-germ 𝒜3\mathcal{A}^{3}-equivalent to (t3,0)(t^{3},0). We set

a~i=γ(3)(0)γ(i)(0)i!|γ(3)(0)|,b~i=det(γ(3)(0),γ(i)(0))i!|γ(3)(0)|.\tilde{a}_{i}=\dfrac{\gamma^{(3)}(0)\cdot\gamma^{(i)}(0)}{i!|\gamma^{(3)}(0)|},\quad\tilde{b}_{i}=\dfrac{\det(\gamma^{(3)}(0),\gamma^{(i)}(0))}{i!|\gamma^{(3)}(0)|}.

One can calculate the invariants up to 1010 degrees as follows. The (3,4)(3,4)-cuspidal curvature r3,4r_{3,4} is

r3,4=b~4a~34/3.r_{3,4}=\dfrac{\tilde{b}_{4}}{\tilde{a}_{3}^{4/3}}. (A.1)

If r3,40r_{3,4}\neq 0, i.e., b~40\tilde{b}_{4}\neq 0, then γ\gamma is 𝒜\mathcal{A}-equivalent to (t3,t4)(t^{3},t^{4}). We assume b~4=0\tilde{b}_{4}=0. Then the (3,5)(3,5)-cuspidal curvature r3,5r_{3,5} is

r3,5=b~5a~35/3.r_{3,5}=\dfrac{\tilde{b}_{5}}{\tilde{a}_{3}^{5/3}}. (A.2)

If r3,50r_{3,5}\neq 0, i.e., b~50\tilde{b}_{5}\neq 0, then γ\gamma is 𝒜\mathcal{A}-equivalent to (t3,t5)(t^{3},t^{5}). We assume b~5=0\tilde{b}_{5}=0. Then the (3,6)(3,6)-bias β3,6\beta_{3,6} and the (3,7)(3,7)-cuspidal curvature r3,7r_{3,7} are

β3,6\displaystyle\beta_{3,6} =b~6a~32,\displaystyle=\dfrac{\tilde{b}_{6}}{\tilde{a}_{3}^{2}}, (A.3)
r3,7\displaystyle r_{3,7} =7a~4b~6+2a~3b~72a~310/3.\displaystyle=\dfrac{-7\tilde{a}_{4}\tilde{b}_{6}+2\tilde{a}_{3}\tilde{b}_{7}}{2\tilde{a}_{3}^{10/3}}. (A.4)

If r3,70r_{3,7}\neq 0, i.e., 7a~4b~6+2a~3b~70-7\tilde{a}_{4}\tilde{b}_{6}+2\tilde{a}_{3}\tilde{b}_{7}\neq 0, then γ\gamma is 𝒜7\mathcal{A}^{7}-equivalent to (t3,t7)(t^{3},t^{7}). We assume r3,7=0r_{3,7}=0, i.e., b~7=7a~4b~6/(2a~3)\tilde{b}_{7}=7\tilde{a}_{4}\tilde{b}_{6}/(2\tilde{a}_{3}). Then the (3,8)(3,8)-cuspidal curvature r3,8r_{3,8} is

r3,8=35a~42b~6+2a~3(28a~5b~6+5a~3b~8)10a~314/3.r_{3,8}=\dfrac{-35\tilde{a}_{4}^{2}\tilde{b}_{6}+2\tilde{a}_{3}(-28\tilde{a}_{5}\tilde{b}_{6}+5\tilde{a}_{3}\tilde{b}_{8})}{10\tilde{a}_{3}^{14/3}}. (A.5)

If r3,80r_{3,8}\neq 0, then γ\gamma is 𝒜8\mathcal{A}^{8}-equivalent to (t3,t8)(t^{3},t^{8}). We assume r3,8=0r_{3,8}=0, i.e., b~8=7(5a~42+8a~3a~5)b~6/(10a~32)\tilde{b}_{8}=7(5\tilde{a}_{4}^{2}+8\tilde{a}_{3}\tilde{a}_{5})\tilde{b}_{6}/(10\tilde{a}_{3}^{2}). Then the (3,9)(3,9)-bias β3,9\beta_{3,9} and the (3,10)(3,10)-cuspidal curvature r3,10r_{3,10} are

β3,9\displaystyle\beta_{3,9} =63a~4a~5b~6+42a~3a~6b~65a~32b~95a~35,\displaystyle=-\dfrac{63\tilde{a}_{4}\tilde{a}_{5}\tilde{b}_{6}+42\tilde{a}_{3}\tilde{a}_{6}\tilde{b}_{6}-5\tilde{a}_{3}^{2}\tilde{b}_{9}}{5\tilde{a}_{3}^{5}}, (A.6)
r3,10\displaystyle r_{3,10} =(10a~33b~10+945a~42a~5b~642a~3(3a~5210a~4a~6)b~615a~32(8a~7b~6+5a~4b~9))10a~319/3.\displaystyle=\dfrac{(10\tilde{a}_{3}^{3}\tilde{b}_{10}+945\tilde{a}_{4}^{2}\tilde{a}_{5}\tilde{b}_{6}-42\tilde{a}_{3}(3\tilde{a}_{5}^{2}-10\tilde{a}_{4}\tilde{a}_{6})\tilde{b}_{6}-15\tilde{a}_{3}^{2}(8\tilde{a}_{7}\tilde{b}_{6}+5\tilde{a}_{4}\tilde{b}_{9}))}{10\tilde{a}_{3}^{19/3}}. (A.7)
Proof of Example A.1.

By rotating γ\gamma in 𝑹3\boldsymbol{R}^{3}, we can write

γ(t)=(i=310a~ii!ti,i=410b~ii!ti)+O(10).\gamma(t)=\left(\sum_{i=3}^{10}\dfrac{\tilde{a}_{i}}{i!}t^{i},\ \sum_{i=4}^{10}\dfrac{\tilde{b}_{i}}{i!}t^{i}\right)+O(10). (A.8)

We set

φ(t)=t(6i=310a~ii!ti3)1/3,\varphi(t)=t\left(6\sum_{i=3}^{10}\dfrac{\tilde{a}_{i}}{i!}t^{i-3}\right)^{1/3},

and the inverse function of s=φ(t)s=\varphi(t) as t=ψ(s)t=\psi(s). We set ψ(s)=i=110ψisi/i!+O(10)\psi(s)=\sum_{i=1}^{10}\psi_{i}s^{i}/i!+O(10). Then we have

ψ1\displaystyle\psi_{1} =1/a~31/3,\displaystyle=1/\tilde{a}_{3}^{1/3},
ψ2\displaystyle\psi_{2} =a~4/(6a~35/3),\displaystyle=-\tilde{a}_{4}/(6\tilde{a}_{3}^{5/3}),
ψ3\displaystyle\psi_{3} =(5a~424a~3a~5)/(40a~33),\displaystyle=(5\tilde{a}_{4}^{2}-4\tilde{a}_{3}\tilde{a}_{5})/(40\tilde{a}_{3}^{3}),
ψ4\displaystyle\psi_{4} =(175a~43+252a~3a~4a~572a~32a~6)/(1080a~313/3),\displaystyle=(-175\tilde{a}_{4}^{3}+252\tilde{a}_{3}\tilde{a}_{4}\tilde{a}_{5}-72\tilde{a}_{3}^{2}\tilde{a}_{6})/(1080\tilde{a}_{3}^{13/3}),
ψ5\displaystyle\psi_{5} =(13475a~4427720a~3a~42a~5+10080a~32a~4a~6+432a~32(14a~525a~3a~7))/(45360a~317/3),\displaystyle=(13475\tilde{a}_{4}^{4}-27720\tilde{a}_{3}\tilde{a}_{4}^{2}\tilde{a}_{5}+10080\tilde{a}_{3}^{2}\tilde{a}_{4}\tilde{a}_{6}+432\tilde{a}_{3}^{2}(14\tilde{a}_{5}^{2}-5\tilde{a}_{3}\tilde{a}_{7}))/(45360\tilde{a}_{3}^{17/3}),
ψ6\displaystyle\psi_{6} =(1575a~45+4200a~3a~43a~51680a~32a~42a~6+96a~32a~4(21a~52+5a~3a~7)\displaystyle=(-1575\tilde{a}_{4}^{5}+4200\tilde{a}_{3}\tilde{a}_{4}^{3}\tilde{a}_{5}-1680\tilde{a}_{3}^{2}\tilde{a}_{4}^{2}\tilde{a}_{6}+96\tilde{a}_{3}^{2}\tilde{a}_{4}(-21\tilde{a}_{5}^{2}+5\tilde{a}_{3}\tilde{a}_{7})
+16a~33(42a~5a~65a~3a~8))/(2240a~37),\displaystyle\hskip 28.45274pt+16\tilde{a}_{3}^{3}(42\tilde{a}_{5}\tilde{a}_{6}-5\tilde{a}_{3}\tilde{a}_{8}))/(2240\tilde{a}_{3}^{7}),
ψ7\displaystyle\psi_{7} =(475475a~461556100a~3a~44a~5+655200a~32a~43a~642120a~32a~42(28a~52+5a~3a~7)\displaystyle=(475475\tilde{a}_{4}^{6}-1556100\tilde{a}_{3}\tilde{a}_{4}^{4}\tilde{a}_{5}+655200\tilde{a}_{3}^{2}\tilde{a}_{4}^{3}\tilde{a}_{6}-42120\tilde{a}_{3}^{2}\tilde{a}_{4}^{2}(-28\tilde{a}_{5}^{2}+5\tilde{a}_{3}\tilde{a}_{7})
+3240a~33a~4(182a~5a~6+15a~3a~8)\displaystyle\hskip 28.45274pt+3240\tilde{a}_{3}^{3}\tilde{a}_{4}(-182\tilde{a}_{5}\tilde{a}_{6}+15\tilde{a}_{3}\tilde{a}_{8})
1296a~33(91a~5360a~3a~5a~7+5a~3(7a~62+a~3a~9)))/(233280a~325/3),\displaystyle\hskip 28.45274pt-1296\tilde{a}_{3}^{3}(91\tilde{a}_{5}^{3}-60\tilde{a}_{3}\tilde{a}_{5}\tilde{a}_{7}+5\tilde{a}_{3}(-7\tilde{a}_{6}^{2}+\tilde{a}_{3}\tilde{a}_{9})))/(233280\tilde{a}_{3}^{25/3}),
ψ8\displaystyle\psi_{8} =(155520a~10a~36+11(4447625a~47+17243100a~3a~45a~57497000a~32a~44a~6\displaystyle=(-155520\tilde{a}_{10}\tilde{a}_{3}^{6}+11(-4447625\tilde{a}_{4}^{7}+17243100\tilde{a}_{3}\tilde{a}_{4}^{5}\tilde{a}_{5}-7497000\tilde{a}_{3}^{2}\tilde{a}_{4}^{4}\tilde{a}_{6}
+2570400a~32a~43(7a~52+a~3a~7)45360a~33a~42(238a~5a~6+15a~3a~8)\displaystyle\hskip 28.45274pt+2570400\tilde{a}_{3}^{2}\tilde{a}_{4}^{3}(-7\tilde{a}_{5}^{2}+\tilde{a}_{3}\tilde{a}_{7})-45360\tilde{a}_{3}^{3}\tilde{a}_{4}^{2}(-238\tilde{a}_{5}\tilde{a}_{6}+15\tilde{a}_{3}\tilde{a}_{8})
+15552a~34(98a~52a~6+20a~3a~6a~7+15a~3a~5a~8)\displaystyle\hskip 28.45274pt+15552\tilde{a}_{3}^{4}(-98\tilde{a}_{5}^{2}\tilde{a}_{6}+20\tilde{a}_{3}\tilde{a}_{6}\tilde{a}_{7}+15\tilde{a}_{3}\tilde{a}_{5}\tilde{a}_{8})
+5184a~33a~4(833a~53420a~3a~5a~7+5a~3(49a~62+5a~3a~9))))/(6998400a~329/3),\displaystyle\hskip 28.45274pt+5184\tilde{a}_{3}^{3}\tilde{a}_{4}(833\tilde{a}_{5}^{3}-420\tilde{a}_{3}\tilde{a}_{5}\tilde{a}_{7}+5\tilde{a}_{3}(-49\tilde{a}_{6}^{2}+5\tilde{a}_{3}\tilde{a}_{9}))))/(6998400\tilde{a}_{3}^{29/3}),
ψ9\displaystyle\psi_{9} =(17920a~10a~4a~36+2480625a~4811113200a~3a~46a~5+4939200a~32a~44(3a~52+a~4a~6)\displaystyle=(17920\tilde{a}_{10}\tilde{a}_{4}\tilde{a}_{3}^{6}+2480625\tilde{a}_{4}^{8}-11113200\tilde{a}_{3}\tilde{a}_{4}^{6}\tilde{a}_{5}+4939200\tilde{a}_{3}^{2}\tilde{a}_{4}^{4}(3\tilde{a}_{5}^{2}+\tilde{a}_{4}\tilde{a}_{6})
70560a~33a~42(84a~53+140a~4a~5a~6+25a~42a~7)+4032a~34(84a~54+840a~4a~52a~6\displaystyle\hskip 28.45274pt-70560\tilde{a}_{3}^{3}\tilde{a}_{4}^{2}(84\tilde{a}_{5}^{3}+140\tilde{a}_{4}\tilde{a}_{5}\tilde{a}_{6}+25\tilde{a}_{4}^{2}\tilde{a}_{7})+4032\tilde{a}_{3}^{4}(84\tilde{a}_{5}^{4}+840\tilde{a}_{4}\tilde{a}_{5}^{2}\tilde{a}_{6}
+600a~42a~5a~7+25a~42(14a~62+5a~4a~8))+2560a~36(12a~72+21a~6a~8+14a~5a~9)\displaystyle\hskip 28.45274pt+600\tilde{a}_{4}^{2}\tilde{a}_{5}\tilde{a}_{7}+25\tilde{a}_{4}^{2}(14\tilde{a}_{6}^{2}+5\tilde{a}_{4}\tilde{a}_{8}))+2560\tilde{a}_{3}^{6}(12\tilde{a}_{7}^{2}+21\tilde{a}_{6}\tilde{a}_{8}+14\tilde{a}_{5}\tilde{a}_{9})
4480a~35(72a~52a~7+a~5(84a~62+90a~4a~8)+5a~4(24a~6a~7+5a~4a~9)))/(89600a~311),\displaystyle\hskip 28.45274pt-4480\tilde{a}_{3}^{5}(72\tilde{a}_{5}^{2}\tilde{a}_{7}+\tilde{a}_{5}(84\tilde{a}_{6}^{2}+90\tilde{a}_{4}\tilde{a}_{8})+5\tilde{a}_{4}(24\tilde{a}_{6}\tilde{a}_{7}+5\tilde{a}_{4}\tilde{a}_{9})))/(89600\tilde{a}_{3}^{11}),
ψ10\displaystyle\psi_{10} =13(16865646875a~49+85717170000a~3a~47a~5+19595520a~10a~36(10a~42+3a~3a~5)\displaystyle=13(-16865646875\tilde{a}_{4}^{9}+85717170000\tilde{a}_{3}\tilde{a}_{4}^{7}\tilde{a}_{5}+19595520\tilde{a}_{10}\tilde{a}_{3}^{6}(-10\tilde{a}_{4}^{2}+3\tilde{a}_{3}\tilde{a}_{5})
38710980000a~32a~46a~6+2844072000a~32a~45(49a~52+5a~3a~7)\displaystyle\hskip 28.45274pt-38710980000\tilde{a}_{3}^{2}\tilde{a}_{4}^{6}\tilde{a}_{6}+2844072000\tilde{a}_{3}^{2}\tilde{a}_{4}^{5}(-49\tilde{a}_{5}^{2}+5\tilde{a}_{3}\tilde{a}_{7})
1422036000a~33a~44(70a~5a~6+3a~3a~8)\displaystyle\hskip 28.45274pt-1422036000\tilde{a}_{3}^{3}\tilde{a}_{4}^{4}(-70\tilde{a}_{5}\tilde{a}_{6}+3\tilde{a}_{3}\tilde{a}_{8})
+372314880a~34a~42(154a~52a~6+20a~3a~6a~7+15a~3a~5a~8)\displaystyle\hskip 28.45274pt+372314880\tilde{a}_{3}^{4}\tilde{a}_{4}^{2}(-154\tilde{a}_{5}^{2}\tilde{a}_{6}+20\tilde{a}_{3}\tilde{a}_{6}\tilde{a}_{7}+15\tilde{a}_{3}\tilde{a}_{5}\tilde{a}_{8})
+206841600a~33a~43(385a~53132a~3a~5a~7+a~3(77a~62+5a~3a~9))\displaystyle\hskip 28.45274pt+206841600\tilde{a}_{3}^{3}\tilde{a}_{4}^{3}(385\tilde{a}_{5}^{3}-132\tilde{a}_{3}\tilde{a}_{5}\tilde{a}_{7}+\tilde{a}_{3}(-77\tilde{a}_{6}^{2}+5\tilde{a}_{3}\tilde{a}_{9}))
1119744a~34a~4(10241a~547980a~3a~52a~7+150a~32(4a~72+7a~6a~8)\displaystyle\hskip 28.45274pt-1119744\tilde{a}_{3}^{4}\tilde{a}_{4}(10241\tilde{a}_{5}^{4}-7980\tilde{a}_{3}\tilde{a}_{5}^{2}\tilde{a}_{7}+150\tilde{a}_{3}^{2}(4\tilde{a}_{7}^{2}+7\tilde{a}_{6}\tilde{a}_{8})
+70a~3a~5(133a~62+10a~3a~9))+186624a~35(22344a~53a~610080a~3a~5a~6a~7\displaystyle\hskip 28.45274pt+70\tilde{a}_{3}\tilde{a}_{5}(-133\tilde{a}_{6}^{2}+10\tilde{a}_{3}\tilde{a}_{9}))+186624\tilde{a}_{3}^{5}(22344\tilde{a}_{5}^{3}\tilde{a}_{6}-10080\tilde{a}_{3}\tilde{a}_{5}\tilde{a}_{6}\tilde{a}_{7}
3780a~3a~52a~8+5a~3(392a~63+135a~3a~7a~8+105a~3a~6a~9)))/(1763596800a~337/3).\displaystyle\hskip 28.45274pt-3780\tilde{a}_{3}\tilde{a}_{5}^{2}\tilde{a}_{8}+5\tilde{a}_{3}(-392\tilde{a}_{6}^{3}+135\tilde{a}_{3}\tilde{a}_{7}\tilde{a}_{8}+105\tilde{a}_{3}\tilde{a}_{6}\tilde{a}_{9})))/(1763596800\tilde{a}_{3}^{37/3}).

Substituting t=ψ(s)t=\psi(s) into γ(t)\gamma(t), and by a straightforward calculation, we see γ(ψ(s))=(s3/6,r3,4s4/4!)+O(4)\gamma(\psi(s))=(s^{3}/6,r_{3,4}s^{4}/4!)+O(4), and we have (A.1). Under the condition r3,4=0r_{3,4}=0, we have γ(ψ(s))=(s3/6,r3,5s5/5!)+O(5)\gamma(\psi(s))=(s^{3}/6,r_{3,5}s^{5}/5!)+O(5), and we have (A.2). We assume r3,4=r3,5=0r_{3,4}=r_{3,5}=0, Then we see γ(ψ(s))=(s3/6,β3,6s6/6!+β3,7s7/7!)+O(7)\gamma(\psi(s))=(s^{3}/6,\beta_{3,6}s^{6}/6!+\beta_{3,7}s^{7}/7!)+O(7), and we have (A.3) and (A.4). We assume r3,7=0r_{3,7}=0. Then we see γ(ψ(s))=(s3/6,β3,6s6/6!+β3,8s8/8!)+O(8)\gamma(\psi(s))=(s^{3}/6,\beta_{3,6}s^{6}/6!+\beta_{3,8}s^{8}/8!)+O(8), and we have (A.5). We assume r3,8=0r_{3,8}=0. Then we see γ(ψ(s))=(s3/6,β3,6s6/6!+β3,9s9/9!+r3,10s10/10!)+O(10)\gamma(\psi(s))=(s^{3}/6,\beta_{3,6}s^{6}/6!+\beta_{3,9}s^{9}/9!+r_{3,10}s^{10}/10!)+O(10), and we have (A.6) and (A.7). ∎

Example A.2.

Let γ\gamma be a curve-germ 𝒜m+1\mathcal{A}^{m+1}-equivalent to (tm,tm+1)(t^{m},t^{m+1}). We set

γ(t)=(i=mm+1aii!ti,i=mm+1bii!ti)+O(m+1)((am,bm)(0,0)).\gamma(t)=\left(\sum_{i=m}^{m+1}\dfrac{a_{i}}{i!}t^{i},\ \sum_{i=m}^{m+1}\dfrac{b_{i}}{i!}t^{i}\right)+O(m+1)\quad((a_{m},b_{m})\neq(0,0)). (A.9)

Then by a standard rotation AA in 𝑹2\boldsymbol{R}^{2} and a parameter change

ta¯1/m(ta¯m+1m(m+1)a(m+1)/mt2),t\mapsto\bar{a}^{-1/m}\left(t-\dfrac{\bar{a}_{m+1}}{m(m+1)a^{(m+1)/m}}t^{2}\right),

we see

Aγ(t)=(tmm!,rm,m+1(m+1)!tm+1),(rm,m+1=b¯m+1a¯m(m+1)/m).A\gamma(t)=\left(\dfrac{t^{m}}{m!},\dfrac{r_{m,m+1}}{(m+1)!}t^{m+1}\right),\quad\left(r_{m,m+1}=\dfrac{\bar{b}_{m+1}}{\bar{a}_{m}^{(m+1)/m}}\right).

Thus the (m,m+1)(m,m+1)-cuspidal curvature is rm,m+1r_{m,m+1}. Here, a¯i\bar{a}_{i} and b¯i\bar{b}_{i} are

a¯i=γ(m)(0)γ(i)(0)i!|γ(m)(0)|,b¯i=det(γ(m)(0),γ(i)(0))i!|γ(m)(0)|.\bar{a}_{i}=\dfrac{\gamma^{(m)}(0)\cdot\gamma^{(i)}(0)}{i!|\gamma^{(m)}(0)|},\quad\bar{b}_{i}=\dfrac{\det(\gamma^{(m)}(0),\gamma^{(i)}(0))}{i!|\gamma^{(m)}(0)|}.

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 5/25/2-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