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

Hyperplane families creating envelopes

Takashi Nishimura Research Institute of Environment and Information Sciences, Yokohama National University, 240-8501 Yokohama, JAPAN nishimura-takashi-yx@ynu.ac.jp
Abstract.

A simple geometric mechanism: “the locus of intersections of perpendicular bisectors and normal lines”,  often arises in many guises in Nonlinear Sciences. In this paper, a new application of this simple geometric mechanism is given. Namely, we show that this mechanism gives answers to all four basic problems on envelopes created by hyperplane families (existence problem, representation problem, equivalence problem of definitions, uniqueness problem) at once.

Key words and phrases:
Hyperplane family, Frontal, Envelope, Creative, Creator, Mirror-image mapping, Anti-orthotomic, Orthotomic, Cahn-Hoffman vector formula.
2010 Mathematics Subject Classification:
57R45, 58C25

1. Introduction

Throughout this paper, let nn be a positive integer. Moreover, all manifolds, functions and mappings are of class CC^{\infty} unless otherwise stated.

A simple geometric mechanism: “the locus of intersections of perpendicular bisectors and normal lines”, often arises in many guises in Physical Sciences. For example, as Richard Feynman elegantly explained in [9], the orbit of a planet around the sun can be understood as a consequence of this mechanism under the assumption of the inverse-square law (see Figure 2 where the circle is the hodograph of the velocity vectors of a planet ​​, that is to say, the circle is a curve drawn by the end points of the vectors that are parallel to the velocity vectors and start at a fixed point PP. The orbit of the planet is similar to the locus of intersections BtB_{t} of the perpendicular bisectors of velocity vectors PAt\overrightarrow{PA_{t}} and the normal lines to the circle at AtA_{t}). This is an example in Celestial Mechanics. In the same book [9], one can find that even the historical discovery of atomic nucleus due to Ernest Rutherford can be explained as a consequence of this simple geometric mechanism (see Figure 2 where the center of circle OO is an atomic nucleus. The orbit of an α\alpha particle is the locus of intersections BtB_{t} of the perpendicular bisectors of the segment PAt¯\overline{PA_{t}} and the normal lines to the circle at AtA_{t} ). This is an example in Nuclear Physics.

Refer to caption
Figure 1. Locus similar to the orbit of a planet.
Refer to caption
Figure 2. Locus of an α\alpha particle.

In Crystallography, one can find such the mechanism in the so-called Wulff construction for the equilibrium shape of a crystal. A brief explanation of the Wulff construction is as follows. Given an equilibrium crystal, take an arbitrary point PP inside the crystal and fix it. Georg Wulff discovered in [20] the so-called Gibbs-Wulff theorem which asserts that the length from the fixed point PP to the foot of the perpendicular to the tangent space to the face of the crystal is proportional to its surface energy density of the face. Let γ:S2\gamma:S^{2}\to\mathbb{R} be the surface energy density function of the equilibrium crystal. The graph of γ\gamma with respect to the polar coordinates about the point PP defines the mapping g:S23g:S^{2}\to\mathbb{R}^{3}. The mapping gg is often called the polar plot of γ\gamma or the γ\gamma-plot or the Wulff plot. Set f=2gf=2g and suppose that the image f(S2)f(S^{2}) has the well-defined normal vectors at any point f(x)f(x). Then, by the Gibbs-Wulff theorem, the accurate shape of the crystal surface is proportional to the shape obtained by our simple geometric mechanism: “the locus of the intersection of the perpendicular bisector of the vector Pf(x)\overrightarrow{Pf(x)} and the normal line to f(S2)f(S^{2}) at f(x)f(x)”. This is the Wulff construction and the constructed shape is called the Wulff shape. Notice that in general ff is a continuous mapping and thus from the viewpoint of rigorous mathematics, the Wulff construction is not a well-defined construction method in general. Nevertheless, Hoffman and Cahn showed in [11] that if γ:S2\gamma:S^{2}\to\mathbb{R} is differentiable, then the image f(S2)f(S^{2}) has a well-defined normal vector at each point f(x)f(x) and the set {γ(x)+γ(x)x|xS2}\{\nabla\gamma(x)+\gamma(x)x\;|\;x\in S^{2}\} is exactly the shape obtained by our simple geometric mechanism for the point PP and the surface f(S2)f(S^{2}). The Wulff construction and the Cahn-Hoffman formula in the plane is depicted in Figure 3. For details on the Wulff construction and Wulff shapes, see for instance [8, 10].

Refer to caption
Figure 3. The Wulff construction and the Cahn-Hoffman vector formula in the plane.

Moreover, it is a surprising fact that our simple geometric mechanism: “the locus of intersections of perpendicular bisectors and normal lines”can be applied even to Seismic Survey (see 7.14 (9) of [6]).

In Mathematics, our simple geometric mechanism: “the locus of intersections of perpendicular bisectors and normal lines”is called the anti-orthotomic of a mapping ff having a well-defined normal vector to its image at each point (for details on anti-orthotomics, see 7.14 of [6]. See also [15] where anti-orthotomics are generalized to frontals and [16] where more elementary explanation on anti-orthotomics can be found). In Mathematics as well, there are examples where anti-orthotomics are effectively applied (see [6]).

In order to understand better the powerfulness of the simple geometric mechanism, we would like to have more striking examples in Mathematics where anti-orthotomics are effectively applied. Namely, we want to seek mathematical problems which can be geometrically solved by our simple geometric mechanism though it seems difficult to solve them by other methods. This is the primitive motivation of this paper. In this paper, we show that the existence and uniqueness problem of envelopes for a given hyperplane family is one of such problems. Namely, we give a necessary and sufficient condition (see Definition 2) for a given hyperplane family to create an envelope. And then, we give a necessary and sufficient condition for the uniqueness of created envelopes if the given hyperplane family creates an envelope. It seems difficult to prove that the condition given in Definition 2 is actually a sufficient condition to create envelopes by other methods. In order to apply our simple geometric mechanism, we need some geometric objects to which the normal line can be reasonably well-defined at each point. Hyperplane families themselves are far from the reasonable geometric objects for our purpose. The reasonable geometric objects are frontals (the definition of frontal is given in the next paragraph). In order to obtain a frontal from a given hyperplane family, the mirror-image mapping will be locally introduced. Then, it turns out that if the given hyperplane family is creative (see Definition 2 below), then the mirror-image mapping is actually a frontal such that the normal line at each point intersects the corresponding hyperplane. Thus, we can apply the anti-orthotomic method developed in [15] to obtain Theorem 1 and Theorem 2. The existence and uniqueness problem of envelopes for a given hyperplane family can be easily interpreted as the existence and uniqueness problem of solutions for a certain type of system of first order differential equations with one constraint condition. In the author’s opinion, one of the most attractive features of our simple geometric mechanism is that it can make all solutions and their precise expressions clear in one shot by geometry without the need to solve the corresponding system of differential equations with one constraint condition.

Let SnS^{n} be the nn-dimensional unit sphere in the (n+1)(n+1)-dimensional vector space n+1\mathbb{R}^{n+1}. Given a point PP of n+1\mathbb{R}^{n+1} and an (n+1)(n+1)-dimensional unit vector 𝐧Snn+1{\bf n}\in S^{n}\subset\mathbb{R}^{n+1}, the hyperplane H(P,𝐧)H_{\left(P,{\bf n}\right)} relative to PP and 𝐧{\bf n} is naturally defined as follows, where the dot in the center stands for the standard scalar product of two vectors (XP)(X-P) and 𝐧{\bf n} in the vector space n+1\mathbb{R}^{n+1}.

H(P,𝐧)={Xn+1|(XP)𝐧=0}.H_{\left(P,{\bf n}\right)}=\{X\in\mathbb{R}^{n+1}\;|\;(X-P)\cdot{\bf n}=0\}.

Let NN be an nn-dimensional manifold without boundary. Given two mappings φ~:Nn+1\widetilde{\varphi}:N\to\mathbb{R}^{n+1} and ν~:NSn\widetilde{\nu}:N\to S^{n}, the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} relative to φ~\widetilde{\varphi} and ν~\widetilde{\nu} is naturally defined as follows.

(φ~,ν~)={H(φ~(x),ν~(x))}xN.\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\left\{H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)}\right\}_{x\in N}}.

A mapping f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} is called a frontal if there exists a mapping ν~:NSn\widetilde{\nu}:N\to S^{n} such that df~x(𝐯)ν~(x)=0d\widetilde{f}_{x}({\bf v})\cdot\widetilde{\nu}(x)=0 for any xNx\in N and any 𝐯TxN{\bf v}\in T_{x}N, where two vector spaces Tf~(x)n+1T_{\widetilde{f}(x)}\mathbb{R}^{n+1} and n+1\mathbb{R}^{n+1} are identified. By definition, it is natural to call ν~:NSn\widetilde{\nu}:N\to S^{n} a Gauss mapping of the frontal f~\widetilde{f}. The notion of frontal has been recently investigated (for instance, see [13]). In this paper, as the definition of envelope created by a hyperplane family, the following is adopted.

Definition 1.

Let (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} be a hyperplane family. A mapping f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} is called an envelope created by (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} if the following two conditions are satisfied.

  1. (a)

    f~(x)H(φ~(x),ν~(x))\widetilde{f}(x)\in H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)} for any xNx\in N.

  2. (b)

    df~x(𝐯)ν~(x)=0d\widetilde{f}_{x}({\bf v})\cdot\widetilde{\nu}(x)=0 for any xNx\in N and any 𝐯TxN{\bf v}\in T_{x}N.

In other words, an envelope created by (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is a mapping f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} giving a solution of the following system of first order differential equations with one constraint condition, where (U,(x1,,xn))\left(U,\left(x_{1},\ldots,x_{n}\right)\right) is an arbitrary coordinate neighborhood of NN such that xUx\in U.

{f~x1(x)ν~(x)=0,f~xn(x)ν~(x)=0,(f~(x)φ~(x))ν~(x)=0.\left\{\begin{array}[]{ccc}\frac{\partial\widetilde{f}}{\partial x_{1}}(x)\cdot\widetilde{\nu}(x)&=&0,\\ \vdots&{}\hfil&{}\hfil\\ \frac{\partial\widetilde{f}}{\partial x_{n}}(x)\cdot\widetilde{\nu}(x)&=&0,\\ \left(\widetilde{f}(x)-\widetilde{\varphi}(x)\right)\cdot\widetilde{\nu}(x)&=&0.\end{array}\right.

By definition, any envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} created by a hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} must be a frontal with Gauss mapping ν~:NSn\widetilde{\nu}:N\to S^{n}. For details on envelopes created by families of plane regular curves, refer to [6]. In Chapter 5 of [6], several definitions for envelope are given. For a hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}, Definition 1 is a generalization of their definition E2E_{2} from a viewpoint of parametrization (E2E_{2} envelope is a variety tangent to all lines of the given line family. Thus, in the case of plane, an envelope defined by Definition 1 is the same notion of E2E_{2} envelope. For details on the definition E2E_{2}, see 5.12 of [6]). The following definition, which may be regarded as a higher dimensional generalization of E1E_{1} from a viewpoint of parametrization (E1E_{1} envelope is the set of the limits of intersections with nearby members of the given line family. For details on the definition E1E_{1}, see 5.8 of [6] and for the relation between Definition 2 in the plane case and E1E_{1}, see Subsection 2.3), is the key notion for this paper.

Definition 2.

Let NN be an nn-dimensional manifold without boundary and let φ~:Nn+1\widetilde{\varphi}:N\to\mathbb{R}^{n+1}, ν~:NSn\widetilde{\nu}:N\to S^{n} be mappings. Let γ~:N\widetilde{\gamma}:N\to\mathbb{R} be the function defined by γ~(x)=φ~(x)ν~(x)\widetilde{\gamma}(x)=\widetilde{\varphi}(x)\cdot\widetilde{\nu}(x). Let TSnT^{*}S^{n} be the cotangent bundle of SnS^{n}. A hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is said to be creative if there exists a mapping Ω~:NTSn\widetilde{\Omega}:N\to T^{*}S^{n} with the form Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right) such that for any x0Nx_{0}\in N the equality dγ~=ω~d\widetilde{\gamma}=\widetilde{\omega} holds as germs of 11-form at x0x_{0}.

TSn{T^{*}S^{n}}N{N}Sn{S^{n}}Ω~\scriptstyle{\widetilde{\Omega}}ν~\scriptstyle{\widetilde{\nu}}

Namely, (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative if there exists a 11-form Ω~\widetilde{\Omega} along ν~\widetilde{\nu} such that for any x0Nx_{0}\in N by using of a coordinate neighborhood (U,(x1,,xn))\left(U,\left(x_{1},\ldots,x_{n}\right)\right) of NN at x0x_{0} and a normal coordinate neighborhood (V,(Θ1,,Θn))\left(V,\left(\Theta_{1},\ldots,\Theta_{n}\right)\right) of SnS^{n} at ν~(x0)\widetilde{\nu}\left(x_{0}\right), the 11-form germ dγ~d\widetilde{\gamma} at x0x_{0} is expressed as follows.

dγ~=i=1n(ω~(x)(Π(ν~(x),ν~(x0))(Θi)))d(Θiν~),d\widetilde{\gamma}=\sum_{i=1}^{n}\left(\widetilde{\omega}(x)\left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Pi}_{\left(\widetilde{\nu}(x),\widetilde{\nu}(x_{0})\right)}\left(\frac{\partial}{\partial\Theta_{i}}\right)\right)\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right),

where a normal coordinate neighborhood (V,(Θ1,,Θn))\left(V,\left(\Theta_{1},\ldots,\Theta_{n}\right)\right) is a local coordinate neighborhood at ν~(x0)\widetilde{\nu}\left(x_{0}\right) obtained by the inverse mappping of the exponential mapping at ν~(x0)\widetilde{\nu}\left(x_{0}\right), SnS^{n} inherits its metric from the ambient space n+1\mathbb{R}^{n+1} and Π(ν~(x),ν~(x0)):Tν~(x0)SnTν~(x)Sn{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Pi}_{\left(\widetilde{\nu}(x),\widetilde{\nu}(x_{0})\right)}:T_{\widetilde{\nu}\left(x_{0}\right)}S^{n}\to T_{\widetilde{\nu}\left(x\right)}S^{n} is the Levi-Civita translation. Notice that our objective manifold is the unit sphere SnS^{n} with metric inherited from n+1\mathbb{R}^{n+1}. Therefore, the Levi-Civita translation Π(ν~(x),ν~(x0))\Pi_{\left(\widetilde{\nu}(x),\widetilde{\nu}(x_{0})\right)} is the restriction of the rotation R:n+1n+1R:\mathbb{R}^{n+1}\to\mathbb{R}^{n+1} satisfying R(ν~(x0))=ν~(x)R(\widetilde{\nu}(x_{0}))=\widetilde{\nu}(x) to the tangent space Tν~(x0)SnT_{\widetilde{\nu}\left(x_{0}\right)}S^{n}. In particular, in the case n=1n=1, a normal coordinate Θ\Theta at ν~(x)\widetilde{\nu}\left(x\right) is nothing but the radian (or, its negative) between two unit vectors ν~(x0)\widetilde{\nu}\left(x_{0}\right) and ν~(x)\widetilde{\nu}\left(x\right) and the Levi-Civita translation Π(ν~(x),ν~(x0))\Pi_{\left(\widetilde{\nu}(x),\widetilde{\nu}(x_{0})\right)} is just the restriction of the plane rotation through Θ\Theta to the tangent space Tν~(x0)S1T_{\widetilde{\nu}\left(x_{0}\right)}S^{1}.

Remark 1.1.
  1. (1)

    It is reasonable to say that γ~\widetilde{\gamma} is totally differentiable with respect to ν~\widetilde{\nu} if (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative.

  2. (2)

    For a creative hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}, the map-germ (ν~,γ~):(N,x0)Sn×\left(\widetilde{\nu},\widetilde{\gamma}\right):(N,x_{0})\to S^{n}\times\mathbb{R} at any x0Nx_{0}\in N is called an opening of ν~:(N,x0)Sn\widetilde{\nu}:(N,x_{0})\to S^{n} (for opening germs, see for example [12]). Thus, Definition 2 may be regarded as a globalization of the notion of opening.

  3. (3)

    Definition 2 may be interpreted as follows. Let θ\theta be a canonical contact 11-form on J1(Sn,)J^{1}(S^{n},\mathbb{R}), namely at any (X0,Y0,P0)J1(Sn,)(X_{0},Y_{0},P_{0})\in J^{1}\left(S^{n},\mathbb{R}\right) the 11-form germ θ\theta is expressed as θ=dYi=1nCidΘi\theta=dY-\sum_{i=1}^{n}C_{i}d\Theta_{i}, where (Θ1,,Θn)\left(\Theta_{1},\ldots,\Theta_{n}\right) is a normal coordinate system at X0X_{0} and (Θ1,,Θn,Y,C1,,Cn)\left(\Theta_{1},\ldots,\Theta_{n},Y,C_{1},\ldots,C_{n}\right) is a canonical coordinate system of J1(Sn,)J^{1}\left(S^{n},\mathbb{R}\right) at (X0,Y0,P0)(X_{0},Y_{0},P_{0}). Then, a hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative if there exists a mapping Ω:NJ1(Sn,){\Omega}:N\to J^{1}\left(S^{n},\mathbb{R}\right) with the form Ω(x)=(ν~(x),γ~(x),c~1(x),,c~n(x)){\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\gamma}(x),\widetilde{c}_{1}(x),\ldots,\widetilde{c}_{n}(x)\right) such that Ωθ=0{\Omega}^{*}\theta=0, where c~1,,c~n:N\widetilde{c}_{1},\ldots,\widetilde{c}_{n}:N\to\mathbb{R} are some functions.

    J1(Sn,){J^{1}\left(S^{n},\mathbb{R}\right)}N{N}Sn{S^{n}}Ω\scriptstyle{\Omega}ν~\scriptstyle{\widetilde{\nu}}

    Notice that in Legendrian Singularity Theory, at any point x0Nx_{0}\in N, the map-germ Ω:(N,x0)J1(Sn,){\Omega}:\left(N,x_{0}\right)\to J^{1}\left(S^{n},\mathbb{R}\right) is assumed to be immersive and it is called a Legendrian immersion; and for Legendrian immersion Ω{\Omega}, the mapping Nx(ν~(x),γ~(x))N\ni x\mapsto\left(\widetilde{\nu}(x),\widetilde{\gamma}(x)\right) is called a wavefront or front (for details on Legendrian Singularity Theory and fronts, see for instance [1, 2, 17]). On the other hand, in Definition 2, Ω{\Omega} is not assumed to be immersive in general and the mapping Ω{\Omega} is called a Legendrian mapping (for details on Legendrian mappings, see for instance [12, 13, 18]). Thus, in Definition 2, in general, the set-germ (Ω(N),Ω(x0))\left({\Omega}(N),{\Omega}\left(x_{0}\right)\right) may be singular at some point x0Nx_{0}\in N (for example, see Example 4.1(4)).

  4. (4)

    Notice that the 11-form Ω~\widetilde{\Omega} along ν~\widetilde{\nu} in Definition 2 is not necessarily the pullback of a 11-form over SnS^{n} by ν~\widetilde{\nu} (for example, see Example 4.1(3), (4)) and the “creativeness”  does not depend on the particular choice of φ~,ν~\widetilde{\varphi},\widetilde{\nu} and depends only on the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. In the case that N=SnN=S^{n} and ν~:SnSn\widetilde{\nu}:S^{n}\to S^{n} is the identity mapping, for any φ~:Snn+1\widetilde{\varphi}:S^{n}\to\mathbb{R}^{n+1} the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is always creative by the following equality.

    dγ~=i=1nγ~ΘidΘi.d\widetilde{\gamma}=\sum_{i=1}^{n}\frac{\partial\widetilde{\gamma}}{\partial\Theta_{i}}d\Theta_{i}.

    More generally, if γ~:U\widetilde{\gamma}:U\to\mathbb{R} may be expressed as the composition of ν~:USn\widetilde{\nu}:U\to S^{n} and a certain function ξ:Sn\xi:S^{n}\to\mathbb{R} over an open set UNU\subset N, then the hyperplane family (φ~|U,ν~|U)\mathcal{H}_{\left(\widetilde{\varphi}|_{U},\widetilde{\nu}|_{U}\right)} is creative. However, there are examples showing that there does not exist a function α~:Sn\widetilde{\alpha}:S^{n}\to\mathbb{R} such that γ~=α~ν~\widetilde{\gamma}=\widetilde{\alpha}\circ\widetilde{\nu} although (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative (for example, see Example 4.1(3), (4)). Moreover, there are many examples such that (φ~|U,ν~|U)\mathcal{H}_{\left(\widetilde{\varphi}|_{U},\widetilde{\nu}|_{U}\right)} is not creative. For instance, for any constant mapping ν~:S1\widetilde{\nu}:\mathbb{R}\to S^{1}, the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is not creative where φ~:2\widetilde{\varphi}:\mathbb{R}\to\mathbb{R}^{2} is defined by φ~(t)=t2ν~(t)\widetilde{\varphi}(t)=t^{2}\widetilde{\nu}(t). And, it is clear in this case that (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} does not create an envelope in the sense of Definition 1. However, it is easily seen that

    𝒟\displaystyle\qquad\qquad\quad\mathcal{D} =\displaystyle= {(X1,X2)2|t s.t. F(X1,X2,t)=Ft(X1,X2,t)=0}\displaystyle\left\{\left(X_{1},X_{2}\right)\in\mathbb{R}^{2}\,|\,\exists t\mbox{ s.t. }F\left(X_{1},X_{2},t\right)=\frac{\partial F}{\partial t}\left(X_{1},X_{2},t\right)=0\right\}
    =\displaystyle= {(X1,X2)2|(X1,X2)ν~(0)=0},\displaystyle\left\{(X_{1},X_{2})\in\mathbb{R}^{2}\,|\,\left(X_{1},X_{2}\right)\cdot\widetilde{\nu}(0)=0\right\}\neq\emptyset,

    where F(X1,X2,t)=((X1,X2)φ~(t))ν~(t)F\left(X_{1},X_{2},t\right)=\left(\left(X_{1},X_{2}\right)-\widetilde{\varphi}(t)\right)\cdot\widetilde{\nu}(t). Thus, for this example, the envelope defined by Definition 1 is different from the envelope in the sense of classical definition (see 5.3 of [6]), For more examples on creative/non-creative hyperplane families and on comparison of Definition 2 with the classical envelope 𝒟\mathcal{D}, see Section 4. Therefore, it seems that the current situation on both the definitions of envelope and the relation of the creative condition (Defnition 2) with an envelope seems to be wrapped in mystery.

By definition, any frontal f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} with Gauss mapping ν~:NSn\widetilde{\nu}:N\to S^{n} is an envelope created by (f~,ν~)\mathcal{H}_{(\widetilde{f},\widetilde{\nu})}. Therefore, the notion of envelope created by a hyperplane family is the same as the notion of frontal. Moreover, it is clear that for any mapping ν~:NSn\widetilde{\nu}:N\to S^{n}, a constant mapping f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} is an envelope created by (f~,ν~)\mathcal{H}_{\left(\widetilde{f},\widetilde{\nu}\right)}. On the other hand, for a constant mapping ν~:S1\widetilde{\nu}:\mathbb{R}\to S^{1}, if the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} does not create an envelope then φ~:2\widetilde{\varphi}:\mathbb{R}\to\mathbb{R}^{2} must be not constant. From these elementary observations, it is natural to ask to obtain a necessary and sufficient condition for a given hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} to create an envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} in terms of γ~:N\widetilde{\gamma}:N\to\mathbb{R} and ν~:NSn\widetilde{\nu}:N\to S^{n}. Moreover, it is also desirable to solve the following two incidentally. “Suppose that a given hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1}. Then, obtain a representation formula of f~\widetilde{f}.”  “Suppose that n=1n=1. Then, find the precise relation between E1E_{1} envelope and E2E_{2} envelope.” In this paper, as an application of our simple geometric mechanism, all of these problems are solved as follows.

Theorem 1.

Let NN be an nn-dimensional manifold without boundary and let φ~:Nn+1\widetilde{\varphi}:N\to\mathbb{R}^{n+1}, ν~:NSn\widetilde{\nu}:N\to S^{n} be mappings. Then, the following three hold.

  1. (1)

    The hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope if and only if it is creative.

  2. (2)

    Suppose that the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1}. Then, for any xNx\in N, under the canonical identifications Tν~(x)SnTν~(x)SnTν~(x)n+1n+1T^{*}_{\widetilde{\nu}(x)}S^{n}\cong T_{\widetilde{\nu}(x)}S^{n}\subset T_{\widetilde{\nu}(x)}\mathbb{R}^{n+1}\cong\mathbb{R}^{n+1}, the (n+1)(n+1)-dimensional vector f~(x)\widetilde{f}(x) is represented as follows.

    f~(x)=ω~(x)+γ~(x)ν~(x),\widetilde{f}(x)=\widetilde{\omega}(x)+\widetilde{\gamma}(x)\widetilde{\nu}(x),

    where the (n+1)(n+1)-dimensional vector ω~(x)\widetilde{\omega}(x) is identified with the corresponding nn-dimensional cotangent vector ω~(x)\widetilde{\omega}(x) under these identifications.

  3. (3)

    Suppose that n=1n=1. Then, the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope (E2E_{2}-envelope) if and only if it creates an E1E_{1} envelope. Moreover, these two envelopes are exactly the same.

By Theorem 1, it is natural to call ω~\widetilde{\omega} the creator for an envelope f~\widetilde{f} created by (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. Recall that E1E_{1} envelope (resp., E2E_{2} envelope) is the set of the limit of intersections with nearby lines (resp., a parametrization tangent to all members of the given family). Thus, even in the case of plane, E2E_{2} envelope is exactly the same as the envelope in Definition 1.

The key idea for the proof of Theorem 1 is to regard the given hyperplane family as a moving mirror parametrized by xNx\in N. Then, for any parameter x0Nx_{0}\in N, by taking a point Pn+1P\in\mathbb{R}^{n+1} outside the mirror H(φ~(x0),ν~(x0))H_{\left(\widetilde{\varphi}(x_{0}),\widetilde{\nu}\left(x_{0}\right)\right)}, the mirror-image

fP(x)=2((φ~(x)P)ν~(x))ν~(x)+Pf_{{}_{P}}(x)=2\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\right)\widetilde{\nu}(x)+P

of PP by the mirror H(φ~(x),ν~(x))H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)} must have the same information as the mirror since the mirror is reconstructed as the perpendicular bisectors of the segment PfP(x)¯\overline{Pf_{{}_{P}}(x)}, where xx is a point in a sufficiently small neighborhood UPU_{{}_{P}} of x0x_{0}.

Refer to caption
Figure 4. The mirror-image mapping fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1}.

Hence, investigation of the given hyperplane family (φ~|UP,ν~|UP)\mathcal{H}_{\left(\widetilde{\varphi}|_{{}U_{{}P}},\widetilde{\nu}|_{{}U_{{}P}}\right)} may be replaced with analyzing the associated mirror-image mapping fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} (see Figure 4). This suggests applicability of results in [15] to the problem of this paper.

A sketch of the proof of Theorem 1 (1) may be given as follows. Suppose that the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative. Then, by definition, there exists a mapping Ω~:NTSn\widetilde{\Omega}:N\to T^{*}S^{n} having the form Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right) such that the equality dγ~=ω~d\widetilde{\gamma}=\widetilde{\omega} holds as germs of 11-form at x0x_{0}. Then, by investigating the Jacobian matrix of the mirror-image mapping fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} at xUPx\in U_{{}_{P}} directly, it turns out that for any xUPx\in U_{{}_{P}} the non-zero vector

𝐯P(x)=i=1n((ω~(x)P)(Θ(i,ν~(x))))Θ(i,ν~(x))((φ~(x)P)ν~(x))ν~(x){\bf v}_{{}_{P}}(x)=\sum_{i=1}^{n}\left(\left(\widetilde{\omega}\left(x\right)-P\right)\left(\frac{\partial}{\partial\Theta_{\left(i,\widetilde{\nu}(x)\right)}}\right)\right)\frac{\partial}{\partial\Theta_{\left(i,\widetilde{\nu}(x)\right)}}-\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}\left(x\right)\right)\widetilde{\nu}\left(x\right)

is perpendicular to the vector d(fP)x(𝐯)d\left(f_{{}_{P}}\right)_{x}\left({\bf v}\right) for any 𝐯TxN{\bf v}\in T_{x}N, where n+1\mathbb{R}^{n+1}, Tν~(x)n+1T_{\widetilde{\nu}\left(x\right)}\mathbb{R}^{n+1} and Tν~(x)n+1T^{*}_{\widetilde{\nu}\left(x\right)}\mathbb{R}^{n+1} are identified and Θ(i,ν~(x))=P(ν~(x),ν~(x0))(Θi)\frac{\partial}{\partial\Theta_{\left(i,\widetilde{\nu}(x)\right)}}=P_{\left(\widetilde{\nu}(x),\widetilde{\nu}\left(x_{0}\right)\right)}\left(\frac{\partial}{\partial\Theta_{i}}\right). Thus, fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} is a frontal. From the construction, the mapping f~P=𝐯P+fP:UPn+1\widetilde{f}_{{}_{P}}={\bf v}_{{}_{P}}+f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} must be exactly the same as the mapping f~P\widetilde{f}_{{}_{P}} given in Theorem 1 of [15]. Therefore, by Theorem 1 of [15] asserting that f~P\widetilde{f}_{{}_{P}} satisfies both conditions (a), (b) of Definition 1, f~P\widetilde{f}_{{}_{P}} is an envelope created by the hyperplane family (φ~|UP,ν~|UP)\mathcal{H}_{\left(\widetilde{\varphi}|_{U_{{}_{P}}},\widetilde{\nu}|_{U_{{}_{P}}}\right)}. The mapping f~P:UPn+1\widetilde{f}_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} is called the anti-orthotomic of fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} relative to PP. Calculation shows

() f~P(x0)=ω~(x0)+γ~(x0)ν~(x0).\widetilde{f}_{{}_{P}}(x_{0})=\widetilde{\omega}\left(x_{0}\right)+\widetilde{\gamma}\left(x_{0}\right)\widetilde{\nu}\left(x_{0}\right).

Thus, unlike fP(x0)f_{{}_{P}}(x_{0}), the location f~P(x0)\widetilde{f}_{{}_{P}}(x_{0}) does not depend on the particular choice of PP. In other words, in order to discover the formula ()(*), the role of PP is merely an auxiliary point just like an auxiliary line in elementary geometry (see Figure 5).

Refer to caption
Figure 5. The location f~P(x0)\widetilde{f}_{{}_{P}}(x_{0}) does not depend on the particular choice of PP.

Since x0x_{0} is an arbitrary point of NN, the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1}.

Conversely, suppose that the given hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1}. Then, the mirror-image mapping fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} (resp., the mapping gP:UPn+1g_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} defined by gP(x)=(f~(x)P)ν~(x)+Pg_{{}_{P}}(x)=\left(\widetilde{f}(x)-P\right)\cdot\widetilde{\nu}(x)+P) is called the orthotomic (resp., pedal) of f~|UP\widetilde{f}|_{U_{{}_{P}}} relative to the point PP. It is known that both the orthotomic fPf_{{}_{P}} and the pedal gPg_{{}_{P}} are frontals (see Proposition 1 and Corollary 1 of [15]). We prefer to investigate the orthotomic fPf_{{}_{P}} rather than the pedal gPg_{{}_{P}} because its Gauss mapping νP:UPSn\nu_{{}_{P}}:U_{{}_{P}}\to S^{n} has characteristic properties: νP(x)=f~(x)fP(x)f~(x)fP(x)\nu_{{}_{P}}(x)=\frac{\widetilde{f}(x)-f_{{}_{P}}(x)}{||\widetilde{f}(x)-f_{{}_{P}}(x)||} and ν~(x)νP(x)0\widetilde{\nu}(x)\cdot\nu_{{}_{P}}(x)\neq 0 for any xUPx\in U_{{}_{P}}, and thus we can take a bird’s eye view of f~(x)\widetilde{f}(x). Set ω~(x)=f~(x)γ~(x)ν~(x)\widetilde{\omega}(x)=\widetilde{f}(x)-\widetilde{\gamma}(x)\widetilde{\nu}(x) and Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right) for any xUPx\in U_{{}_{P}}. Then, under the identification of n+1\mathbb{R}^{n+1} and Tν~(x)n+1T^{*}_{\widetilde{\nu}(x)}\mathbb{R}^{n+1}, Ω~\widetilde{\Omega} having the form Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right) is a well-defined mapping UPTSnU_{{}_{P}}\to T^{*}S^{n}. By investigating the Jacobian matrix of the mirror image mapping fPf_{{}_{P}} at xUPx\in U_{{}_{P}} directly again, it turns out that ω~\widetilde{\omega} is actually the creator for the envelope f~|UP\widetilde{f}|_{U_{{}_{P}}}. Since the vector ω~(x0)\widetilde{\omega}(x_{0}) does not depend on the particular choice of PP and the point x0x_{0} is an arbitrary point of NN, (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative.

Theorem 1 (2) is a direct by-product of the proof of Theorem 1 (1) (see Figure 5). Theorem 1 (3) seems to be not a direct by-product of the proof of Theorem 1 (1) although it can be proved relatively easily by using the above argument (see Subsection 2.3).

When N=SnN=S^{n} and ν~:SnSn\widetilde{\nu}:S^{n}\to S^{n} is the identity mapping, it is easily seen ω~(x)=γ~(x)\widetilde{\omega}(x)=\nabla\widetilde{\gamma}(x). Therefore, in the case that N=SnN=S^{n} and ν~:SnSn\widetilde{\nu}:S^{n}\to S^{n} is the identity mapping, Theorem 1 (2) has been known as the Cahn-Hoffman vector formula ([11]). Theorem 1 (2) is a comprehensive generalization of their formula. Any Wulff shape is clearly a convex body and conversely it is known that any convex body can be constructed by the Wulff construction (for instance, see [19]). There are many Wulff shapes such that the surface energy density functions γ:Sn\gamma:S^{n}\to\mathbb{R} are not differentiable (convex polytopes are typical examples). Thus, for studing Wulff shapes having non-smooth surface energy functions, it is very significant to answer the following two problems: “(a) Generalize Cahn-Hoffman vector formula to the corresponding formula for any ν~:NSn\widetilde{\nu}:N\to S^{n}”ȧnd “(b) Resolution of singularities of the boundary of a convex body having non-smooth boundary by a frontal f~:Snn+1\widetilde{f}:S^{n}\to\mathbb{R}^{n+1}”. By Theorem 1 (2), the problem (a) is completely solved. As for the problem (b), to the best of author’s knowledge, only the boundary of a square has been realized as a frontal f~:S12\widetilde{f}:S^{1}\to\mathbb{R}^{2} so far (see [15]). Although there are apparently no published proofs at present, it is a comparatively straightforward generalization of this result to show that the boundary of a convex polygon is realized as a frontal f~:S12\widetilde{f}:S^{1}\to\mathbb{R}^{2}. However, even in the plane case, the problem (b) for the boundary of a convex body in general seems to be wrapped in mystery.

Moreover, Theorem 1 (2) might be useful even for the study of force problems in higher dimensional vector spaces. In [4], Petr Blaschke discovered that pedal coordinates are more suitable settings to study force problems in 2\mathbb{R}^{2}. Readers who want to confirm their usefulness are recommeded to refer to [4] (see also 7.24 (6) of [6] though this is not a force problem but a very suitable problem for understanding how useful pedal coordinates are). Theorem 1 (2) may be regarded as a higher dimensional generalization of pedal coordinates. Hence, it is expected that Theorem 1 (2) is a very suitable expression to study force problems etc. in all finite-dimensional vector spaces over \mathbb{R}. Example 4.2 (2) might be regarded as examples in which higher dimensional version of pedal coordinates are effectively used.

As an application of Theorem 1, a characterization for a hyperplane family to create a unique envelope is given as follows.

Theorem 2.

Let φ~:Nn+1\widetilde{\varphi}:N\to\mathbb{R}^{n+1}, ν~:NSn\widetilde{\nu}:N\to S^{n} be mappings. Then, the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates a unique envelope if and only if it is creative and the set consisting of regular points of ν~\widetilde{\nu} is dense in NN.

Under the assumption that Ω\Omega in Remark 1.1 (2) is immersive and some conditions are satisfied, a unique existence result of envelopes for hyperplane families has been obtained in [7]. Since their assumptions clearly imply that the creative condition defined in Definition 2 is satisfied and the set consisting of regular points of ν~\widetilde{\nu} is dense, their result follows from Theorem 1 and Theorem 2.

Notice that non-unique existence cases, too, are intriguing cases since Theorem 1 may be effectively applied even in such cases (see Example 4.2 (1), (2)).


This paper is organized as follows. Theorem 1 and Theorem 2 are proved in Section 2 and Section 3 respectively. In Section 4, examples are given. Section 5 is an appendix where an alternative proof of Theorem 1 except for Theorem 1 (3) is given. The alternative proof is a proof by a gauge theoretic approach. In order to avoid unnecessary complication, the alternative proof is given only in the case n=1n=1. The author has no idea on how to prove Theorem 1 (3) by using the alternative proof.

2. Proof of Theorem 1

2.1. Proof of Theorem 1 (1)

2.1.1. Proof of “if”  part

Let x0x_{0} be an arbitrary point of NN. Take one point PP of n+1H(φ~(x0),ν~(x0))\mathbb{R}^{n+1}-H_{\left(\widetilde{\varphi}\left(x_{0}\right),\widetilde{\nu}\left(x_{0}\right)\right)} and fix it. It follows (φ~(x0)P)ν~(x0)0\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\neq 0. Let U~P\widetilde{U}_{{}_{P}} be the set of points xNx\in N satisfying

(2.1) (φ~(x)P)ν~(x)0.\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\neq 0.

Then, it is clear that U~P\widetilde{U}_{{}_{P}} is an open neighborhood of x0x_{0} and the mirror image of the fixed point PP by the mirror H(φ~(x),ν~(x))H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)} is given by

2((φ~(x)P)ν~(x))ν~(x)+P2\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\right)\widetilde{\nu}(x)+P

for any xU~Px\in\widetilde{U}_{{}_{P}}.

Refer to caption
Figure 6. The mirror-image mapping fP:UPn+1f_{{}_{P}}:U_{P}\to\mathbb{R}^{n+1}.

Since the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is assumed to be creative, there exists a mapping Ω~:NTSn\widetilde{\Omega}:N\to T^{*}S^{n} with the form Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right) such that for any xNx\in N the following equality holds as 11-form germs at xx.

dγ~=ω~.d\widetilde{\gamma}=\widetilde{\omega}.

Let (V,(Θ1,,Θn))\left(V,\left(\Theta_{1},\ldots,\Theta_{n}\right)\right) be a normal coordinate neighborhood of SnS^{n} at ν~(x0)\widetilde{\nu}\left(x_{0}\right). Set UP=U~Pν~1(V)U_{{}_{P}}=\widetilde{U}_{{}_{P}}\cap\widetilde{\nu}^{-1}(V). Consider the mirror-image mapping fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} defined by

fP(x)=2((φ~(x)P)ν~(x))ν~(x)+Pf_{{}_{P}}(x)=2\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\right)\widetilde{\nu}(x)+P

for any xUPx\in U_{{}_{P}}. In order to show that fPf_{{}_{P}} is a frontal, it is sufficient to construct a Gauss mapping with respect to fPf_{{}_{P}}. By using the mapping Ω~|UP\widetilde{\Omega}|_{U_{{}_{P}}}, a Gauss mapping for fPf_{{}_{P}} is constructed as follows. For any xUPx\in U_{{}_{P}} set X=ν~(x)X=\widetilde{\nu}(x). Let Π(X,X0):TX0SnTXSn{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Pi}_{\left(X,X_{0}\right)}:T_{X_{0}}S^{n}\to T_{X}S^{n} be the Levi-Civita translation. For any ii (1in)(1\leq i\leq n), set Θ(i,X)=Π(X,X0)(Θi)\frac{\partial}{\partial\Theta_{\left(i,X\right)}}={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Pi}_{\left(X,X_{0}\right)}\left(\frac{\partial}{\partial\Theta_{i}}\right). Then notice that for any xUPx\in U_{{}_{P}}, under the identification of n+1\mathbb{R}^{n+1} and TfP(x)n+1T_{f_{{}_{P}}(x)}\mathbb{R}^{n+1},

Θ(1,X),,Θ(n,X),ν~(x)\left\langle\frac{\partial}{\partial\Theta_{\left(1,X\right)}},\ldots,\frac{\partial}{\partial\Theta_{\left(n,X\right)}},\widetilde{\nu}(x)\right\rangle

is an orthonormal basis of the tangent vector space TfP(x)n+1T_{f_{{}_{P}}(x)}\mathbb{R}^{n+1}.

Refer to caption
Figure 7. Figure for Proof of “if”  part.
Lemma 2.1.

For any xUPx\in U_{{}_{P}}, the following equality holds.

d(Pν~)=i=1n(PΘ(i,X))d(Θiν~).d\left(P\cdot\widetilde{\nu}\right)=\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right).

Proof of Lemma 2.1.

d(Pν~)\displaystyle d\left(P\cdot\widetilde{\nu}\right) =\displaystyle= j=1n(Pν~)xj(x)dxj\displaystyle\sum_{j=1}^{n}\frac{\partial\left(P\cdot\widetilde{\nu}\right)}{\partial x_{j}}(x)dx_{j}
=\displaystyle= j=1n(P(i=1n(Θiν~)xj(x)Θ(i,X)))dxj\displaystyle\sum_{j=1}^{n}\left(P\cdot\left(\sum_{i=1}^{n}\frac{\partial\left(\Theta_{i}\circ\widetilde{\nu}\right)}{\partial x_{j}}(x)\frac{\partial}{\partial\Theta_{(i,X)}}\right)\right)dx_{j}
=\displaystyle= i=1n(PΘ(i,X))(j=1n(Θiν~)xj(x)dxj)\displaystyle\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{(i,X)}}\right)\left(\sum_{j=1}^{n}\frac{\partial\left(\Theta_{i}\circ\widetilde{\nu}\right)}{\partial x_{j}}(x)dx_{j}\right)
=\displaystyle= i=1n(PΘ(i,X))d(Θiν~).\displaystyle\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right).

\Box

By Lemma 2.1, under the identification of Tν~(x)SnT_{\widetilde{\nu}(x)}S^{n} and Tν~(x)SnT^{*}_{\widetilde{\nu}(x)}S^{n}, it follows

d((φ~P)ν~)\displaystyle d\left(\left(\widetilde{\varphi}-P\right)\cdot\widetilde{\nu}\right) =\displaystyle= d(φ~ν~)d(Pν~)\displaystyle d\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right)-d\left(P\cdot\widetilde{\nu}\right)
=\displaystyle= dγ~d(Pν~)\displaystyle d\widetilde{\gamma}-d\left(P\cdot\widetilde{\nu}\right)
=\displaystyle= ω~d(Pν~)\displaystyle\widetilde{\omega}-d\left(P\cdot\widetilde{\nu}\right)
=\displaystyle= i=1n(ω~(x)Θ(i,X))d(Θiν~)i=1n(PΘ(i,X))d(Θiν~)\displaystyle\sum_{i=1}^{n}\left(\widetilde{\omega}(x)\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)-\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)
=\displaystyle= i=1n((ω~(x)P)Θ(i,X))d(Θiν~)\displaystyle\sum_{i=1}^{n}\left(\left(\widetilde{\omega}(x)-P\right)\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)

for any xUPx\in U_{{}_{P}}. Set

𝐯P(x)=i=1n((ω~(x)P)Θ(i,X))Θ(i,X)((φ~(x)P)ν~(x))ν~(x){\bf v}_{{}_{P}}(x)=\sum_{i=1}^{n}\left(\left(\widetilde{\omega}(x)-P\right)\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)\frac{\partial}{\partial\Theta_{(i,X)}}-\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\right)\widetilde{\nu}(x)

for any xUPx\in U_{{}_{P}} where n+1\mathbb{R}^{n+1} and TfP(x)n+1T_{f_{{}_{P}}\left(x\right)}\mathbb{R}^{n+1} are identified and TfP(x)SnT_{f_{{}_{P}}\left(x\right)}S^{n} and TfP(x)SnT_{f_{{}_{P}}\left(x\right)}^{*}S^{n} are identified. By ((2.1)), 𝐯P(x){\bf v}_{{}_{P}}(x) is not the zero vector. Moreover, the following holds.

Lemma 2.2.

For any 𝐯Tx0N{\bf v}\in T_{x_{0}}N, 𝐯P(x0){\bf v}_{{}_{P}}(x_{0}) is perpendicular to d(fP)x0(𝐯)d\left(f_{{}_{P}}\right)_{{}x_{0}}({\bf v}).

Proof of Lemma 2.2. Calculation of the product of the vector 𝐯P(x0){\bf v}_{{}_{P}}\left(x_{0}\right) and the Jacobian matrix of fPf_{{}_{P}} at x0x_{0} (denoted by J(fP)x0J\left(f_{{}_{P}}\right)_{x_{0}}) is carried out as follows, where n+1\mathbb{R}^{n+1} and TfP(x0)n+1T_{f_{{}_{P}}\left(x_{0}\right)}\mathbb{R}^{n+1} are identified and TfP(x0)SnT_{f_{{}_{P}}\left(x_{0}\right)}S^{n} and TfP(x0)SnT_{f_{{}_{P}}\left(x_{0}\right)}^{*}S^{n} are identified.

𝐯P(x0)J(fP)x0\displaystyle{\bf v}_{{}_{P}}\left(x_{0}\right)J\left(f_{{}_{P}}\right)_{x_{0}}
=\displaystyle= 2i=1n((ω~(x0)P)Θi)((φ~(x0)P)ν~(x0))d(Θiν~)\displaystyle 2\sum_{i=1}^{n}\left(\left(\widetilde{\omega}\left(x_{0}\right)-P\right)\cdot\frac{\partial}{\partial\Theta_{i}}\right)\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)
2((φ~(x0)P)ν~(x0))d((φ~P)ν~)at x0\displaystyle\qquad-2\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)d\left(\left(\widetilde{\varphi}-P\right)\cdot\widetilde{\nu}\right)_{\mbox{at }x_{0}}
=\displaystyle= 2((φ~(x0)P)ν~(x0))i=1n((ω~(x0)P)Θi)d(Θiν~)\displaystyle 2\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\sum_{i=1}^{n}\left(\left(\widetilde{\omega}\left(x_{0}\right)-P\right)\cdot\frac{\partial}{\partial\Theta_{i}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)
2((φ~(x0)P)ν~(x0))i=1n((ω~(x0)P)Θi)d(Θiν~)\displaystyle\qquad-2\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\sum_{i=1}^{n}\left(\left(\widetilde{\omega}\left(x_{0}\right)-P\right)\cdot\frac{\partial}{\partial\Theta_{i}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)
=\displaystyle= 0.\displaystyle 0.

\Box

We may consider that the point x0x_{0} is an arbitrary point of UPU_{{}_{P}}. Thus we have the following.

Lemma 2.3.

The mapping fP:UPn+1f_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} is a frontal with Gauss mapping νP:UPSn\nu_{{}_{P}}:U_{{}_{P}}\to S^{n} such that νP(x)ν~(x)0\nu_{{}_{P}}(x)\cdot\widetilde{\nu}(x)\neq 0, where νP(x)=𝐯P(x)𝐯P(x)\nu_{{}_{P}}(x)=\frac{{\bf v}_{{}_{P}}(x)}{\|{\bf v}_{{}_{P}}(x)\|}.

By Lemma 2.3, the hyperplane H(φ~(x),ν~(x))H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)} and the line x={fP(x)+tνP(x)|t}\ell_{x}=\left\{f_{{}_{P}}(x)+t\nu_{{}_{P}}(x)\left|\right.t\in\mathbb{R}\right\} must intersect only at one point for any xUPx\in U_{{}_{P}}. Define the mapping f~P:UPn+1\widetilde{f}_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} by

{f~P(x)}=H(φ~(x),ν~(x))x.\left\{\widetilde{f}_{{}_{P}}(x)\right\}=H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)}\cap\ell_{x}.

Then, from the construction, f~P\widetilde{f}_{{}_{P}} must have the following form (see p.7 of [15]).

f~P(x)=fP(x)fP(x)P22(fP(x)P)νP(x)νP(x).\widetilde{f}_{{}_{P}}(x)=f_{{}_{P}}(x)-\frac{||f_{{}_{P}}(x)-P||^{2}}{2\left(f_{{}_{P}}(x)-P\right)\cdot\nu_{{}_{P}}(x)}\nu_{{}_{P}}(x).

By Theorem 1 of [15] (more precisely, by 3.1 in p.9 of [15]) and Lemma 2.3, we have the following.

Lemma 2.4.

The mapping f~P\widetilde{f}_{{}_{P}} is a frontal with Gauss mapping ν~|UP:UPSn\widetilde{\nu}|_{U_{{}_{P}}}:U_{{}_{P}}\to S^{n}. In other words, f~P:UPn+1\widetilde{f}_{{}_{P}}:U_{{}_{P}}\to\mathbb{R}^{n+1} is an envelope created by the hyperplane family (φ~|UP,ν~|UP)\mathcal{H}_{\left(\widetilde{\varphi}|_{U_{{}_{P}}},\widetilde{\nu}|_{U_{{}_{P}}}\right)}.

On the other hand, it is easily seen that (fP(x0)+𝐯P(x0)φ~(x0))ν~(x0)=0\left(f_{{}_{P}}\left(x_{0}\right)+{\bf v}_{{}_{P}}\left(x_{0}\right)-\widetilde{\varphi}\left(x_{0}\right)\right)\cdot\widetilde{\nu}\left(x_{0}\right)=0 (see Figure 7). Thus, the vector fP(x0)+𝐯P(x0)f_{{}_{P}}\left(x_{0}\right)+{\bf v}_{{}_{P}}\left(x_{0}\right) must belong to H(φ~(x0),ν~(x0))H_{\left(\widetilde{\varphi}\left(x_{0}\right),\widetilde{\nu}\left(x_{0}\right)\right)}. From the construction and by using the equality P=i=1n(PΘi)Θi+(Pν~(x0))ν~(x0),P=\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{i}}\right)\frac{\partial}{\partial\Theta_{i}}+\left(P\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right), we have the following.

f~P(x0)\displaystyle\widetilde{f}_{{}_{P}}\left(x_{0}\right) =\displaystyle= fP(x)+𝐯P(x0)\displaystyle f_{{}_{P}}(x)+{\bf v}_{{}_{P}}\left(x_{0}\right)
=\displaystyle= 2((φ~(x0)P)ν~(x0))ν~(x0)+P\displaystyle 2\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)+P
+i=1n((ω~(x0)P)Θi)Θi((φ~(x0)P)ν~(x0))ν~(x0)\displaystyle\quad\quad+\sum_{i=1}^{n}\left(\left(\widetilde{\omega}\left(x_{0}\right)-P\right)\cdot\frac{\partial}{\partial\Theta_{i}}\right)\frac{\partial}{\partial\Theta_{i}}-\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)
=\displaystyle= ((φ~(x0)P)ν~(x0))ν~(x0)+P+i=1n((ω~(x0)P)Θi)Θi\displaystyle\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)+P+\sum_{i=1}^{n}\left(\left(\widetilde{\omega}\left(x_{0}\right)-P\right)\cdot\frac{\partial}{\partial\Theta_{i}}\right)\frac{\partial}{\partial\Theta_{i}}
=\displaystyle= (φ~(x0)ν~(x0))ν~(x0)+i=1n(ω~(x0)Θi)Θi\displaystyle\left(\widetilde{\varphi}\left(x_{0}\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)+\sum_{i=1}^{n}\left(\widetilde{\omega}\left(x_{0}\right)\cdot\frac{\partial}{\partial\Theta_{i}}\right)\frac{\partial}{\partial\Theta_{i}}
=\displaystyle= γ~(x0)ν~(x0)+ω~(x0).\displaystyle\widetilde{\gamma}\left(x_{0}\right)\widetilde{\nu}\left(x_{0}\right)+\widetilde{\omega}\left(x_{0}\right).

This proves the following lemma.

Lemma 2.5.

The following equality holds.

f~P(x0)=γ~(x0)ν~(x0)+ω~(x0).\widetilde{f}_{{}_{P}}\left(x_{0}\right)=\widetilde{\gamma}\left(x_{0}\right)\widetilde{\nu}\left(x_{0}\right)+\widetilde{\omega}\left(x_{0}\right).

Lemma 2.5 shows that f~P(x0)\widetilde{f}_{{}_{P}}\left(x_{0}\right) does not depend on the particular choice of Pn+1H(φ~(x0),ν~(x0))P\in\mathbb{R}^{n+1}-H_{\left(\widetilde{\varphi}\left(x_{0}\right),\widetilde{\nu}\left(x_{0}\right)\right)}. Define the mapping f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} by f~(x)=γ~(x)ν~(x)+ω~(x)\widetilde{f}(x)=\widetilde{\gamma}(x)\widetilde{\nu}(x)+\widetilde{\omega}(x). Since x0x_{0} is an arbitrary point of NN, by Lemma 2.4 and Lemma 2.5, it follows that the mapping f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1} is an envelope created by (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. This completes the proof of “if”  part. \Box

2.1.2. Proof of “only if”  part

Suppose that the hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1}. Then, by definition, f~\widetilde{f} is a frontal such that the inclusion f~(x)+df~x(TxN)H(φ~(x),ν~(x))\widetilde{f}(x)+d\widetilde{f}_{x}(T_{x}N)\subset H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)} holds for any xNx\in N. Let ω~:Nn+1\widetilde{\omega}:N\to\mathbb{R}^{n+1} be the mapping defined by ω~(x)=f~(x)γ~(x)ν~(x)\widetilde{\omega}(x)=\widetilde{f}(x)-\widetilde{\gamma}(x)\widetilde{\nu}(x) (see Figure 8).

Refer to caption
Figure 8. Figure for Proof of “only if”  part.

It is sufficient to show that under some identifications, ω~\widetilde{\omega} is actually a creator for the envelope f~\widetilde{f}.

It is easily seen that ω~(x)ν~(x)=0\widetilde{\omega}(x)\cdot\widetilde{\nu}(x)=0 for any xNx\in N. Thus, under the identification of n+1\mathbb{R}^{n+1} and Tν~(x)n+1T^{*}_{\widetilde{\nu}(x)}\mathbb{R}^{n+1}, we have

Lemma 2.6.

For any xNx\in N, ω~(x)Tν~(x)Sn\widetilde{\omega}(x)\in T^{*}_{\widetilde{\nu}(x)}S^{n} holds.

Let Ω~:NTSn\widetilde{\Omega}:N\to T^{*}S^{n} be the mapping defined by Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right). Let x0x_{0} be an arbitrary point of NN and let PP be a point of n+1H(φ~(x0),ν~(x0))\mathbb{R}^{n+1}-H_{\left(\widetilde{\varphi}(x_{0}),\widetilde{\nu}(x_{0})\right)}. Again, we consider the mirror-image mapping fP:U~Pn+1f_{{}_{P}}:\widetilde{U}_{{}_{P}}\to\mathbb{R}^{n+1} defined by

fP(x)=2((φ~(x)P)ν~(x))ν~(x)+P,f_{{}_{P}}(x)=2\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\right)\widetilde{\nu}(x)+P,

where U~P={xN|(φ~(x)P)ν~(x)0}\widetilde{U}_{{}_{P}}=\left\{x\in N\,\left|\,\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\neq 0\right\}\right.. The mapping fPf_{{}_{P}} is exactly the orthotomic of f~|U~P\widetilde{f}|_{\widetilde{U}_{{}_{P}}} relative to the point PP. Thus, by Proposition 1 of [15] (more precisely, by 2.1 in pp. 7–8 of [15]) , fPf_{{}_{P}} is a frontal and the mapping νP:U~PSn\nu_{{}_{P}}:\widetilde{U}_{{}_{P}}\to S^{n} define by

νP(x)=f~(x)fP(x)f~(x)fP(x)\nu_{{}_{P}}(x)=\frac{\widetilde{f}(x)-f_{{}_{P}}(x)}{\|\widetilde{f}(x)-f_{{}_{P}}(x)\|}

is its Gauss mapping. In particular, we have the following.

Lemma 2.7.

For any xU~Px\in\widetilde{U}_{{}_{P}} and any 𝐯TxN{\bf v}\in T_{x}N, the following holds.

(f~(x)fP(x))d(fP)x(𝐯)=0.\left(\widetilde{f}(x)-f_{{}_{P}}(x)\right)\cdot d\left(f_{{}_{P}}\right)_{x}({\bf v})=0.

For any xU~Px\in\widetilde{U}_{{}_{P}}, set

gP(x)=12(fP(x)P)+P=((φ~(x)P)ν~(x))ν~(x)+P.g_{{}_{P}}(x)=\frac{1}{2}\left(f_{{}_{P}}(x)-P\right)+P=\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\right)\widetilde{\nu}(x)+P.

Then, since fP(x)f_{{}_{P}}(x) is the mirror-image of PP with respect to the mirror H(φ~(x),ν~(x))H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)}, the following clearly holds.

Lemma 2.8.

The vector f~(x)gP(x)\widetilde{f}(x)-g_{{}_{P}}(x) is perpendicular to the vector gP(x)fP(x)=((φ~(x)P)ν~(x))ν~(x)g_{{}_{P}}(x)-f_{{}_{P}}(x)=-\left(\left(\widetilde{\varphi}(x)-P\right)\cdot\widetilde{\nu}(x)\right)\widetilde{\nu}(x) for any xU~Px\in\widetilde{U}_{{}_{P}}.

Thus,

f~(x)fP(x)=(f~(x)gP(x))+(gP(x)fP(x))\widetilde{f}(x)-f_{{}_{P}}(x)=\left(\widetilde{f}(x)-g_{{}_{P}}(x)\right)+\left(g_{{}_{P}}(x)-f_{{}_{P}}(x)\right)

is an orthogonal decomposition of f~(x)fP(x)\widetilde{f}(x)-f_{{}_{P}}(x) for any xU~Px\in\widetilde{U}_{{}_{P}} (see Figure 8).

In order to decompose the vector f~(x)gP(x)\widetilde{f}(x)-g_{{}_{P}}(x) reasonably, the open neighborhood U~P\widetilde{U}_{{}_{P}} of x0x_{0} is reduced as follows. Let (V,(Θ1,,Θn))\left(V,\left(\Theta_{1},\ldots,\Theta_{n}\right)\right) be a normal coordinate neighborhood of SnS^{n} at ν~(x0)\widetilde{\nu}\left(x_{0}\right). Set again UP=U~Pν~1(V)U_{{}_{P}}=\widetilde{U}_{{}_{P}}\cap\widetilde{\nu}^{-1}(V). Notice that dΘ1,,dΘn\left\langle d\Theta_{1},\ldots,d\Theta_{n}\right\rangle is an orthonormal basis of the cotangent space Tν~(x0)SnT^{*}_{\widetilde{\nu}\left(x_{0}\right)}S^{n}.

Lemma 2.9.

The equality

f~(x0)gP(x0)=ω~(x0)i=1n(PΘi)Θi\widetilde{f}\left(x_{0}\right)-g_{{}_{P}}\left(x_{0}\right)=\widetilde{\omega}\left(x_{0}\right)-\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{i}}\right)\frac{\partial}{\partial\Theta_{i}}

holds where three vector spaces n+1\mathbb{R}^{n+1}, Tν~(x0)n+1T_{\widetilde{\nu}\left(x_{0}\right)}\mathbb{R}^{n+1} and Tν~(x0)n+1T^{*}_{\widetilde{\nu}\left(x_{0}\right)}\mathbb{R}^{n+1} are identified.

Proof of Lemma 2.9. 

f~(x0)gP(x0)\displaystyle\widetilde{f}\left(x_{0}\right)-g_{{}_{P}}\left(x_{0}\right) =\displaystyle= f~(x0)(((φ~(x0)P)ν~(x0))ν~(x0)+P)\displaystyle\widetilde{f}\left(x_{0}\right)-\left(\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)+P\right)
=\displaystyle= (f~(x0)(φ~(x0)ν~(x0))ν~(x0))+((Pν~(x0))ν~(x0)P)\displaystyle\left(\widetilde{f}\left(x_{0}\right)-\left(\widetilde{\varphi}\left(x_{0}\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)\right)+\left(\left(P\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)-P\right)
=\displaystyle= (f~(x0)γ~(x0)ν~(x0))+((Pν~(x0))ν~(x0)P)\displaystyle\left(\widetilde{f}\left(x_{0}\right)-\widetilde{\gamma}\left(x_{0}\right)\widetilde{\nu}\left(x_{0}\right)\right)+\left(\left(P\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right)-P\right)
=\displaystyle= ω~(x0)i=1n(PΘi)Θi.\displaystyle\widetilde{\omega}\left(x_{0}\right)-\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{i}}\right)\frac{\partial}{\partial\Theta_{i}}.

\Box

By Lemma 2.9, the following holds.

f~(x0)fP(x0)\displaystyle\widetilde{f}\left(x_{0}\right)-f_{{}_{P}}\left(x_{0}\right) =\displaystyle= (f~(x0)gP(x0))+(gP(x0)fP(x0))\displaystyle\left(\widetilde{f}\left(x_{0}\right)-g_{{}_{P}}\left(x_{0}\right)\right)+\left(g_{{}_{P}}\left(x_{0}\right)-f_{{}_{P}}\left(x_{0}\right)\right)
=\displaystyle= ω~(x0)i=1n(PΘi)Θi((φ~(x0)P)ν~(x0))ν~(x0).\displaystyle\widetilde{\omega}\left(x_{0}\right)-\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{i}}\right)\frac{\partial}{\partial\Theta_{i}}-\left(\left(\widetilde{\varphi}\left(x_{0}\right)-P\right)\cdot\widetilde{\nu}\left(x_{0}\right)\right)\widetilde{\nu}\left(x_{0}\right).

Hence, by Lemma 2.1 and Lemma 2.7, the germ of 11-form dγ~d\widetilde{\gamma} at x0x_{0} is calculated as follows, where X=ν~(x)X=\widetilde{\nu}(x), Θ(i,X)=P(X,X0)(Θi).\frac{\partial}{\partial\Theta_{\left(i,X\right)}}=P_{\left(X,X_{0}\right)}\left(\frac{\partial}{\partial\Theta_{i}}\right). and P(X,X0):TX0SnTXSnP_{\left(X,X_{0}\right)}:T_{X_{0}}S^{n}\to T_{X}S^{n} is the Levi-Civita translation.

dγ~\displaystyle d\widetilde{\gamma} =\displaystyle= dγ~d(Pν~)+d(Pν~)\displaystyle d\widetilde{\gamma}-d\left(P\cdot\widetilde{\nu}\right)+d\left(P\cdot\widetilde{\nu}\right)
=\displaystyle= d((φ~P)ν~)+d(Pν~)\displaystyle d\left(\left(\widetilde{\varphi}-P\right)\cdot\widetilde{\nu}\right)+d\left(P\cdot\widetilde{\nu}\right)
=\displaystyle= i=1n((ω~P)Θ(i,X))d(Θiν~)+i=1n(PΘ(i,X))d(Θiν~)\displaystyle\sum_{i=1}^{n}\left(\left(\widetilde{\omega}-P\right)\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)+\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)
=\displaystyle= (ω~i=1n(PΘ(i,X))d(Θiν~))+i=1n(PΘ(i,X))d(Θiν~)\displaystyle\left(\widetilde{\omega}-\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)\right)+\sum_{i=1}^{n}\left(P\cdot\frac{\partial}{\partial\Theta_{\left(i,X\right)}}\right)d\left(\Theta_{i}\circ\widetilde{\nu}\right)
=\displaystyle= ω~.\displaystyle\widetilde{\omega}.

This calculation proves the following lemma.

Lemma 2.10.

The equality

dγ~=ω~d\widetilde{\gamma}=\widetilde{\omega}

holds as germs of 11-form at x0x_{0}.

Since x0x_{0} is an arbitrary point of NN, by Lemma 2.10, ω~\widetilde{\omega} is actually the creator for the given envelope f~:Nn+1\widetilde{f}:N\to\mathbb{R}^{n+1}. This completes the proof of “only if”  part. \Box

2.2. Proof of Theorem 1 (2)

Theorem 1 (2) is a direct by-product of the proof of Theorem 1 (1). ∎

2.3. Proof of Theorem 1 (3)

Recall that the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is said to create an E1E_{1} envelope (denoted by (E1)(E_{1}) in this subsection) if for any fixed t0Nt_{0}\in N and any tNt\in N near t0t_{0} the limit limtt0H(φ~(t),ν~(t))H(φ~(t0),ν~(t0))\lim_{t\to t_{0}}H_{\left(\widetilde{\varphi}(t),\widetilde{\nu}(t)\right)}\cap H_{\left(\widetilde{\varphi}(t_{0}),\widetilde{\nu}(t_{0})\right)} exists. On the other hand, the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is said to create an E2E_{2} envelope (denoted by (E2)(E_{2}) in this subsection) if it creates an envelope in the sense of Definition 1.

(E1)(E2)(E_{1})\Rightarrow(E_{2}) Let t0t_{0} be a point of NN and let tiNt_{i}\in N (i=1,2,)(i=1,2,\ldots) be a sequence conversing to t0t_{0}. Since (E1)(E_{1}) is assumed, we can assume that a point XtiX_{t_{i}} can be taken from the intersection H(φ~(t),ν~(t))H(φ~(t0),ν~(t0))H_{\left(\widetilde{\varphi}(t),\widetilde{\nu}(t)\right)}\cap H_{\left(\widetilde{\varphi}(t_{0}),\widetilde{\nu}(t_{0})\right)} such that limtit0Xti\lim_{t_{i}\to t_{0}}X_{t_{i}} exists. Denote the limit by Xt0X_{t_{0}}. Then, we have the following.

(Xtiφ~(ti))ν~(ti)\displaystyle\left(X_{t_{i}}-\widetilde{\varphi}({t_{i}})\right)\cdot\widetilde{\nu}(t_{i}) =\displaystyle= 0,\displaystyle 0,
(Xtiφ~(t0))ν~(t0)\displaystyle\left(X_{t_{i}}-\widetilde{\varphi}(t_{0})\right)\cdot\widetilde{\nu}(t_{0}) =\displaystyle= 0.\displaystyle 0.

This implies

Xti(ν~(ti)ν~(t0))=γ~(ti)γ~(t0).X_{t_{i}}\cdot\left(\widetilde{\nu}(t_{i})-\widetilde{\nu}(t_{0})\right)=\widetilde{\gamma}(t_{i})-\widetilde{\gamma}(t_{0}).

Thus we have

Xt0ν~t(t0)=γ~t(t0).X_{t_{0}}\cdot\frac{\partial\widetilde{\nu}}{\partial t}(t_{0})=\frac{\partial\widetilde{\gamma}}{\partial t}(t_{0}).

This implies that there exists a real number α(t0)\alpha(t_{0}) such that the following identity holds where d(Θν~)d\left(\Theta\circ\widetilde{\nu}\right) and dγ~d\widetilde{\gamma} stand for the 11-dimensional cotangent vectors in Tt0NT_{t_{0}}^{*}N, namely the following identity is nothing but the identity of two real numbers.

α(t0)d(Θν~)=dγ~.\alpha(t_{0})d\left(\Theta\circ\widetilde{\nu}\right)=d\widetilde{\gamma}.

It is not difficult to see that the function α:N\alpha:N\to\mathbb{R} is of class CC^{\infty}. This means that the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative. Therefore, by Theorem 1 (1), the line family creates an E2E_{2} envelope. ∎

(E2)(E1)(E_{2})\Rightarrow(E_{1}) For the proof of this implication, it is used the notions and notations introduced in the proof of Theorem 1 (1). The assumption (E2)(E_{2}) implies that γ~\widetilde{\gamma} is totally differentiable with respect to ν~\widetilde{\nu}. Take an arbitrary point t0Nt_{0}\in N and fixed it. Since γ~\widetilde{\gamma} is totally differentiable with respect to ν~\widetilde{\nu} at t0t_{0}, for any tt near t0t_{0} if the length of the vector fP(t0)fP(t)\overrightarrow{f_{P}(t_{0})f_{P}(t)} is positive, then the horizontal vector of fP(t0)fP(t)\overrightarrow{f_{P}(t_{0})f_{P}(t)} must be non-zero, where PP is a point taken outside the line H(φ~(t0),ν~(t0))H_{\left(\widetilde{\varphi}(t_{0}),\widetilde{\nu}(t_{0})\right)} and fPf_{P} is a mirror-image mapping introduced in the proof of Theorem 1 (1). Denote the intersection of the perpendicular bisector of fP(t0)fP(t)\overrightarrow{f_{P}(t_{0})f_{P}(t)} and the line H(φ~(t0),ν~(t0))H_{\left(\widetilde{\varphi}(t_{0}),\widetilde{\nu}(t_{0})\right)} by JtJ_{t}. Then, from the construction, it follows that the triangre JtfP(t0)fP(t)\triangle J_{t}f_{P}(t_{0})f_{P}(t) is an isosceles triangle with legs JtfP(t0)J_{t}f_{P}(t_{0}) and JtfP(t)J_{t}f_{P}(t). This implies the following (see Figure 9).

JtH(φ~(t),ν~(t))H(φ~(t0),ν~(t0)).J_{t}\in H_{\left(\widetilde{\varphi}(t),\widetilde{\nu}(t)\right)}\cap H_{\left(\widetilde{\varphi}(t_{0}),\widetilde{\nu}(t_{0})\right)}.
Refer to caption
Figure 9. Figure for (E2)(E1)(E_{2})\Rightarrow(E_{1}).

Notice that limtt0JtfP(t0)\lim_{t\to t_{0}}||J_{t}f_{P}(t_{0})|| is positive. Thus, we have

limtt0JtfP(t0)fP(t)=limtt0JtfP(t)fP(t0)=π2.\lim_{t\to t_{0}}\angle J_{t}f_{P}(t_{0})f_{P}(t)=\lim_{t\to t_{0}}\angle J_{t}f_{P}(t)f_{P}(t_{0})=\frac{\pi}{2}.

By Proposition 1 of [15] asserting that fPf_{P} is a frontal with its Gauss mapping fP(t0)f~P(t0)fP(t0)f~P(t0)\frac{f_{P}(t_{0})-\widetilde{f}_{P}(t_{0})}{||f_{P}(t_{0})-\widetilde{f}_{P}(t_{0})||}, it follows

limtt0Jt=f~P(t0),\lim_{t\to t_{0}}J_{t}=\widetilde{f}_{P}(t_{0}),

where f~P\widetilde{f}_{P} is the anti-orthotomic of fPf_{P} relative to the point PP introduced in the proof of Theorem 1 (1). Since t0t_{0} is an arbitrary point of NN, the given E2E_{2} envelope must be an E1E_{1} envelope by Theorem 1 (1). ∎

3. Proof of Theorem 2

Proof of “if”  part.  Since the hyperplane (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative, by Theorem 1, it creates an envelope. Let f~1,f~2:Nn+1\widetilde{f}_{1},\widetilde{f}_{2}:N\to\mathbb{R}^{n+1} be envelopes created by (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}.

Let x0Nx_{0}\in N be a regular point of ν~\widetilde{\nu}. Then, there exists an open coordinate neighborhood (U,(x1,,xn))\left(U,\left(x_{1},\ldots,x_{n}\right)\right) such that x0Ux_{0}\in U and ν~|U:Uν~(U)\widetilde{\nu}|_{U}:U\to\widetilde{\nu}(U) is a diffeomorphism. Then, the germ of 11-form d(φ~ν~)d\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right) at x0Ux_{0}\in U is

d(φ~ν~)\displaystyle d\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right) =\displaystyle= j=1n(φ~ν~)xj(x)dxj\displaystyle\sum_{j=1}^{n}\frac{\partial\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right)}{\partial x_{j}}(x)dx_{j}
=\displaystyle= j=1n(φ~ν~)xj(x)(i=1n(xjν~1)Θ(i,ν~(x))(ν~(x))dΘi)\displaystyle\sum_{j=1}^{n}\frac{\partial\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right)}{\partial x_{j}}(x)\left(\sum_{i=1}^{n}\frac{\partial\left(x_{j}\circ\widetilde{\nu}^{-1}\right)}{\partial\Theta_{\left(i,\widetilde{\nu}(x)\right)}}\left(\widetilde{\nu}(x)\right)d\Theta_{i}\right)
=\displaystyle= i=1n(j=1n(φ~ν~)xj(x)(xjν~1)Θ(i,ν~(x))(ν~(x)))dΘi.\displaystyle\sum_{i=1}^{n}\left(\sum_{j=1}^{n}\frac{\partial\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right)}{\partial x_{j}}(x)\frac{\partial\left(x_{j}\circ\widetilde{\nu}^{-1}\right)}{\partial\Theta_{\left(i,\widetilde{\nu}(x)\right)}}\left(\widetilde{\nu}(x)\right)\right)d\Theta_{i}.

Let Ω~:NTSn\widetilde{\Omega}:N\to T^{*}S^{n} be the mapping with the form Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right) such that ω~\widetilde{\omega} is the creator for f~\widetilde{f}. Then, by the above calculation, ω~|U\widetilde{\omega}|_{U} must have the following form.

ω~|U(x)=i=1n(j=1n(φ~ν~)xj(x)(xjν~1)Θ(i,ν~(x))(ν~(x)))dΘi.\widetilde{\omega}|_{U}(x)=\sum_{i=1}^{n}\left(\sum_{j=1}^{n}\frac{\partial\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right)}{\partial x_{j}}(x)\frac{\partial\left(x_{j}\circ\widetilde{\nu}^{-1}\right)}{\partial\Theta_{\left(i,\widetilde{\nu}(x)\right)}}\left(\widetilde{\nu}(x)\right)\right)d\Theta_{i}.

Hence, by Theorem 1 (2), we have the following.

Lemma 3.1.

At a regular point x0Nx_{0}\in N of ν~\widetilde{\nu}, the equality f~1(x0)=f~2(x0)\widetilde{f}_{1}(x_{0})=\widetilde{f}_{2}(x_{0}) holds.

Let x0Nx_{0}\in N be a singular point of ν~\widetilde{\nu}. Then, since we have assumed that the set of regular points of ν~\widetilde{\nu} is dense, there exists a point-sequence {yi}i=1,2,N\left\{y_{i}\right\}_{i=1,2,\ldots}\subset N such that yiy_{i} is a regular point of ν~\widetilde{\nu} for any ii\in\mathbb{N} and limiyi=x0\lim_{i\to\infty}y_{i}=x_{0}. Then, by Lemma 3.1, we have

f~1(x0)=f~1(limiyi)=limif~1(yi)=limif~2(yi)=f~2(limiyi)=f~2(x0).\widetilde{f}_{1}(x_{0})=\widetilde{f}_{1}\left(\lim_{i\to\infty}y_{i}\right)=\lim_{i\to\infty}\widetilde{f}_{1}(y_{i})=\lim_{i\to\infty}\widetilde{f}_{2}(y_{i})=\widetilde{f}_{2}\left(\lim_{i\to\infty}y_{i}\right)=\widetilde{f}_{2}(x_{0}).

Thus, we have the following.

Lemma 3.2.

Even at a singular point x0Nx_{0}\in N of ν~\widetilde{\nu}, the equality f~1(x0)=f~2(x0)\widetilde{f}_{1}(x_{0})=\widetilde{f}_{2}(x_{0}) holds.

\Box

Proof of “only if”  part.  Suppose that the hyperplane (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative and the set of regular points of ν~\widetilde{\nu} is not dense in NN. Then, there exists an open set UU of NN such that any point xUx\in U is a singular point of ν~\widetilde{\nu}. Then, there exist an integer kk (0k<n)(0\leq k<n) and an open set UkU_{k} such that UkUU_{k}\subset U and the rank of ν~\widetilde{\nu} at xx is kk for any xUkx\in U_{k}. Let x0x_{0} be a point of UkU_{k}. We may assume that UkU_{k} is sufficiently small open neighborhood of x0x_{0}. Then, by the rank theorem (for the rank theorem, see for example [5]), we have the following.

Lemma 3.3.

There exist functions η1,,ηk:N\eta_{1},\ldots,\eta_{k}:N\to\mathbb{R} such that the following three hold.

  1. (1)

    For any ii (1in)(1\leq i\leq n), ηi(x)=0\eta_{i}(x)=0 if xUkx\not\in U_{k}.

  2. (2)

    There exists an ii (1in)(1\leq i\leq n) such that ηi(x0)0\eta_{i}\left(x_{0}\right)\neq 0.

  3. (3)

    The following equality holds for any xNx\in N.

    i=1nηi(x)d(Θiν~)=0.\sum_{i=1}^{n}\eta_{i}(x)d\left(\Theta_{i}\circ\widetilde{\nu}\right)=0.

Since we have assumed that (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative, there exists a mapping Ω~:NTSn\widetilde{\Omega}:N\to T^{*}S^{n} with he form Ω~(x)=(ν~(x),ω~(x))\widetilde{\Omega}(x)=\left(\widetilde{\nu}(x),\widetilde{\omega}(x)\right) such that d(φ~ν~)=ω~d\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right)=\widetilde{\omega}. By Lemma 3.3, the following holds.

Lemma 3.4.

For any function α:N\alpha:N\to\mathbb{R} and any xNx\in N, the following equality holds as germs of 11-form at xx.

d(φ~ν~)=ω~(x)+α(x)i=1nηi(x)d(Θiν~).d\left(\widetilde{\varphi}\cdot\widetilde{\nu}\right)=\widetilde{\omega}(x)+\alpha(x)\sum_{i=1}^{n}\eta_{i}(x)d\left(\Theta_{i}\circ\widetilde{\nu}\right).

Therefore, by Theorem 1 (2), uncountably many distinct envelopes f~\widetilde{f} are created by the same hyperplane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. \Box

4. Examples

Example 4.1 (Uniform spin of affine tangent lines).
  1. (1)

    Let α:\alpha:\mathbb{R}\to\mathbb{R} be a non-constant function. Notice that α\alpha is of class CC^{\infty} as stated at the top of Section 1. Let φ~:2\widetilde{\varphi}:\mathbb{R}\to\mathbb{R}^{2} be the mapping defined by φ~(t)=(α(t),0)\widetilde{\varphi}(t)=(\alpha(t),0). Let ν~:S1\widetilde{\nu}:\mathbb{R}\to S^{1} be the constant mapping ν~(t)=(0,1)\widetilde{\nu}(t)=(0,1). For any fixed θ0\theta_{0}\in\mathbb{R}, let Rθ0:22R_{\theta_{0}}:\mathbb{R}^{2}\to\mathbb{R}^{2} be the linear mapping representing the rotation through angle θ0\theta_{0}. Set ν~θ0(t)=Rθν~(t)=(sinθ0,cosθ0)\widetilde{\nu}_{\theta_{0}}(t)=R_{\theta}\circ\widetilde{\nu}(t)=\left(-\sin\theta_{0},\cos\theta_{0}\right) and γ~θ0(t)=φ~(t)ν~θ0(t)=α(t)sinθ0\widetilde{\gamma}_{\theta_{0}}(t)=\widetilde{\varphi}(t)\cdot\widetilde{\nu}_{\theta_{0}}(t)=-\alpha(t)\sin\theta_{0}. Figure is depicted in Figure 10.

    Refer to caption
    Figure 10. Figure for Example 4.1 (1).

    It follows d(Θν~θ0)0d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)\equiv 0 and dγθ0=sinθ0dαd\gamma_{\theta_{0}}=-\sin\theta_{0}d\alpha. Since α\alpha is non-constant, there exists a regular point of α\alpha, that is to say, there exists a tt\in\mathbb{R} such that α(t)0\alpha^{\prime}(t)\neq 0. Therefore, by Theorem 1, the line family (φ~,ν~θ0)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}_{\theta_{0}}\right)} creates an envelope if and only if θ0π\theta_{0}\in\pi\mathbb{Z}. Suppose that θ0π\theta_{0}\in\pi\mathbb{Z}. In this case, by Theorem 2, uncountably many distinct envelope f~:2\widetilde{f}:\mathbb{R}\to\mathbb{R}^{2} can be created by the given line family (φ~,ν~θ0)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}_{\theta_{0}}\right)}. Let β:\beta:\mathbb{R}\to\mathbb{R} be a function. Since d(Θν~θ0)0d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)\equiv 0 and dγθ00d\gamma_{\theta_{0}}\equiv 0 in this case, the 11-form tβ(t)d(Θν~θ0)t\mapsto\beta(t)d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right) along ν~θ0\widetilde{\nu}_{\theta_{0}} may be a creator ω~\widetilde{\omega} for the line family. By Theorem 1 (2), the envelope f~\widetilde{f} has the following form.

    f~(t)=ω~(t)+(γ~θ0(t)ν~θ0(t))ν~θ0(t)=(±β(t),0)+(0,0)=(±β(t),0),\widetilde{f}(t)=\widetilde{\omega}(t)+{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\widetilde{\gamma}_{\theta_{0}}(t)\cdot\widetilde{\nu}_{\theta_{0}}(t)}}\right)\widetilde{\nu}_{\theta_{0}}(t)}}=({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\pm}}}\beta(t),0)+(0,0)=({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\pm}}}\beta(t),0),

    where double sign should be read in the same order and β(t)d(Θν~θ0)\beta(t)d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right), β(t)Rπ2ν~θ0(t)\beta(t)R_{\frac{\pi}{2}}\circ\widetilde{\nu}_{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}{\theta_{0}}}(t) are identified (both are denoted by the same symbol ω~(t)\widetilde{\omega}(t)).

    Set Fθ0(X1,X2,t)=(X1α(t),X2)ν~θ0(t)F_{\theta_{0}}\left(X_{1},X_{2},t\right)=\left(X_{1}-\alpha(t),X_{2}\right)\cdot\widetilde{\nu}_{\theta_{0}}(t). Suppose that θ0π\theta_{0}\not\in\pi\mathbb{Z}. In this case, the classical common definition of envelope 𝒟\mathcal{D} relative to Fθ0F_{\theta_{0}} is as follows.

    𝒟={(X1,X2)|t s.t. α(t)=0,X1=cotθ0X2+α(t)}.\mathcal{D}=\left\{\left(X_{1},X_{2}\right)\,|\,\exists t\mbox{ s.t. }\alpha^{\prime}(t)=0,X_{1}=\cot\theta_{0}X_{2}+\alpha(t)\right\}.

    Therefore, in this case, 𝒟=E1=E2=\mathcal{D}=E_{1}=E_{2}=\emptyset if and only if α\alpha is non-singular. Suppose that θ0π\theta_{0}\in\pi\mathbb{Z}. Then,

    𝒟={(X1,X2)|X2=0}.\mathcal{D}=\left\{\left(X_{1},X_{2}\right)\,|\,X_{2}=0\right\}.

    Therefore, in this case, E1=E2=𝒟E_{1}=E_{2}=\mathcal{D} if and only if β\beta is surjective.

  2. (2)

    Let ν~:S1\widetilde{\nu}:\mathbb{R}\to S^{1} be the mapping given by ν~(t)=(cost,sint)\widetilde{\nu}(t)=(\cos t,\sin t). Set ν~θ0=Rθ0ν~\widetilde{\nu}_{\theta_{0}}=R_{\theta_{0}}\circ\widetilde{\nu}, where Rθ0R_{\theta_{0}} is the rotation defined in the above example. Then, since d(Θν~θ0)dt(t)=1\frac{d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)}{dt}(t)=1, it follows d(Θν~θ0)=dtd\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)=dt. Thus, by Theorem 1(1) and Theorem 2, for any φ~:2\widetilde{\varphi}:\mathbb{R}\to\mathbb{R}^{2} the line family (φ~,ν~θ0)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}_{\theta_{0}}\right)} creates a unique envelope f~θ0\widetilde{f}_{\theta_{0}}. For any φ~:2\widetilde{\varphi}:\mathbb{R}\to\mathbb{R}^{2}, set γ~θ0(t)=φ~(t)ν~θ0(t)\widetilde{\gamma}_{\theta_{0}}(t)=\widetilde{\varphi}(t)\cdot\widetilde{\nu}_{\theta_{0}}(t). Since dγ~θ0=dγ~θ0dt(t)d(Θν~θ0)d\widetilde{\gamma}_{\theta_{0}}=\frac{d\widetilde{\gamma}_{\theta_{0}}}{dt}(t)d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right), by Theorem 1 (2), it follows

    f~(t)\displaystyle\widetilde{f}(t) =\displaystyle= dγ~θ0dt(t)Rπ/2ν~θ0(t)+γ~θ0(t)ν~θ0(t)\displaystyle\frac{d\widetilde{\gamma}_{\theta_{0}}}{dt}(t)R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}\left(t\right)+\widetilde{\gamma}_{\theta_{0}}(t)\widetilde{\nu}_{\theta_{0}}(t)
    =\displaystyle= dγ~θ0dt(t)Rπ/2ν~θ0(t)+γ~θ0(t)(cos(t+θ0),sin(t+θ0)),\displaystyle\frac{d\widetilde{\gamma}_{\theta_{0}}}{dt}(t)R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}\left(t\right)+\widetilde{\gamma}_{\theta_{0}}(t)\left(\cos\left(t+\theta_{0}\right),\sin\left(t+\theta_{0}\right)\right),

    where the 11-form d(Θν~)d\left(\Theta\circ\widetilde{\nu}\right) and the vector field Rπ/2ν~θ0(t)R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}\left(t\right) are identified. Let α:\alpha:\mathbb{R}\to\mathbb{R} be a function and set φ~(t)=ν~(t)+α(t)Rπ/2ν~θ0(t)\widetilde{\varphi}(t)=\widetilde{\nu}(t)+\alpha(t)R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}(t).

    Refer to caption
    Figure 11. Figure for Example 4.1 (2).

    Then, it follows dγ~θ0dt(t)0\frac{d\widetilde{\gamma}_{\theta_{0}}}{dt}(t)\equiv 0. Thus, as expected, the envelope created by the line family (φ~,ν~θ0)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}_{\theta_{0}}\right)} in this case is actually the circle with radius |c||c| centered at the origin, where c=γ~θ0(t)=cosθ0c=\widetilde{\gamma}_{\theta_{0}}(t)=\cos\theta_{0} (see Figure 11).

  3. (3)

    Let ν~:S1\widetilde{\nu}:\mathbb{R}\to S^{1} be the mapping defined by ν~(t)=11+9t4(3t2,1)\widetilde{\nu}(t)=\frac{1}{\sqrt{1+9t^{4}}}\left(-3t^{2},1\right). Set ν~θ0=Rθ0ν~\widetilde{\nu}_{\theta_{0}}=R_{\theta_{0}}\circ\widetilde{\nu} where Rθ0R_{\theta_{0}} is as above. Let α:\alpha:\mathbb{R}\to\mathbb{R} be a function and set φ~θ0(t)=(t,t3)+α(t)Rπ/2ν~θ0(t)\widetilde{\varphi}_{\theta_{0}}(t)=(t,t^{3})+\alpha(t)R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}(t). Set γ~θ0(t)=φ~θ0(t)ν~θ0(t)\widetilde{\gamma}_{\theta_{0}}(t)=\widetilde{\varphi}_{\theta_{0}}(t)\cdot\widetilde{\nu}_{\theta_{0}}(t). It is easily seen that 0 is a singular point of γ~θ0\widetilde{\gamma}_{\theta_{0}} if and only if θ0π\theta_{0}\in\pi\mathbb{Z}. On the other hand, by calculation, we have d(Θν~θ0)dt(t)=6t1+9t4\frac{d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)}{dt}(t)=\frac{6t}{1+9t^{4}} and thus 0 is a unique singular point of ν~θ0\widetilde{\nu}_{\theta_{0}} for any θ0\theta_{0}. Therefore, by Theorem 1, the hyperplane family (φ~,ν~θ)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}_{{}_{\theta}}\right)} does not create an envelope if θ0π\theta_{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}0}\not\in\pi\mathbb{Z}.

    Next, suppose that θ0π\theta_{0}\in\pi\mathbb{Z}. Then, calculations show

    d(γ~θ0)=(6t2+18t6)(1+9t4)32dt=(t+3t5)1+9t4d(Θν~θ0)dt(t)dt=(t+3t5)1+9t4d(Θν~θ0),\quad\quad d\left(\widetilde{\gamma}_{\theta_{0}}\right)=\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(6t^{2}+18t^{6})}{(1+9t^{4})^{\frac{3}{2}}}dt=\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(t+3t^{5})}{\sqrt{1+9t^{4}}}\frac{d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)}{dt}(t)dt=\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(t+3t^{5})}{\sqrt{1+9t^{4}}}d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0},}

    where double sign should be read in the same order. Set ω~(t)=(t+3t5)1+9t4d(Θν~θ0)\widetilde{\omega}(t)=\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(t+3t^{5})}{\sqrt{1+9t^{4}}}d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right). By Theorem 1 and Theorem 2, the hyperplane family (φ~,ν~θ0)\mathcal{H}_{(\widetilde{\varphi},\widetilde{\nu}_{\theta_{0}})} creates a unique envelope with the desired form

    f~(t)\displaystyle\widetilde{f}(t) =\displaystyle= ω~(t)+γ~θ0(t)ν~θ0(t)\displaystyle\widetilde{\omega}(t)+\widetilde{\gamma}_{\theta_{0}}(t)\widetilde{\nu}_{\theta_{0}}(t)
    =\displaystyle= (t+3t5)1+9t4(1,3t2)2t31+9t4(3t2,±1)\displaystyle\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(t+3t^{5})}{1+9t^{4}}\left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}1,{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}3t^{2}\right){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}\frac{2t^{3}}{1+9t^{4}}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}3t^{2},{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\pm}1)
    =\displaystyle= 11+9t4(t+3t5+6t5,3t3+9t72t3)\displaystyle\frac{1}{1+9t^{4}}\left(t+3t^{5}+6t^{5},3t^{3}+9t^{7}-2t^{3}\right)
    =\displaystyle= (t,t3),\displaystyle\left(t,t^{3}\right),

    where for each tt\in\mathbb{R} the cotangent vector (t+3t5)1+9t4d(Θν~θ0)\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(t+3t^{5})}{\sqrt{1+9t^{4}}}d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right) and the vector (t+3t5)1+9t4Rπ/2ν~θ0(t)\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(t+3t^{5})}{\sqrt{1+9t^{4}}}R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}(t) in the vector space 2\mathbb{R}^{2} are identified (see Figure 12).

    Refer to caption
    Figure 12. Figure for Example 4.1 (3) in the case θ02π\theta_{0}\in 2\pi\mathbb{Z}.

    Set U={0}U=\mathbb{R}-\{0\}. It is easily seen that ν~θ0|U\widetilde{\nu}_{\theta_{0}}|_{U} is non-singular even in the case θ0π\theta_{0}\not\in\pi\mathbb{Z}. Hence, by Theorem 1 and Theorem 2, the hyperplane family (φ~|U,ν~θ0|U)\mathcal{H}_{\left(\widetilde{\varphi}|_{U},\widetilde{\nu}_{\theta_{0}}|_{U}\right)} creates a unique envelope f~θ0:U2\widetilde{f}_{\theta_{0}}:U\to\mathbb{R}^{2} even when θ0π\theta_{0}\not\in\pi\mathbb{Z} and limt0f~θ0(t)=\lim_{t\to 0}\|\widetilde{f}_{\theta_{0}}(t)\|=\infty when θ0π\theta_{0}\not\in\pi\mathbb{Z}.

  4. (4)

    Let ν~:S1\widetilde{\nu}:\mathbb{R}\to S^{1} be the mapping defined by ν~(t)=14+25t6(5t3,2)\widetilde{\nu}(t)=\frac{1}{\sqrt{4+25t^{6}}}\left(-5t^{3},2\right). Set ν~θ0=Rθ0ν~\widetilde{\nu}_{\theta_{0}}=R_{\theta_{0}}\circ\widetilde{\nu} where Rθ0R_{\theta_{0}} is as above. Let α:\alpha:\mathbb{R}\to\mathbb{R} be a function and set φ~θ0(t)=(t2,t5)+α(t)Rπ/2ν~θ0(t)\widetilde{\varphi}_{\theta_{0}}(t)=(t^{2},t^{5})+\alpha(t)R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}(t). Set γ~θ0(t)=φ~θ0(t)ν~θ0(t)=3t5cosθ02t2sinθ05t8sinθ04+25t6\widetilde{\gamma}_{\theta_{0}}(t)=\widetilde{\varphi}_{\theta_{0}}(t)\cdot\widetilde{\nu}_{\theta_{0}}(t)=\frac{-3t^{5}\cos\theta_{0}-2t^{2}\sin\theta_{0}-5t^{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}8}\sin\theta_{0}}{\sqrt{4+25t^{6}}}. By calculation, we have d(Θν~θ0)dt(t)=30t24+25t6\frac{d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)}{dt}(t)=\frac{30t^{2}}{4+25t^{6}}. Therefore, the hyperplane family (φ~,ν~θ)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}_{{}_{\theta}}\right)} is not creative if θπ\theta\not\in\pi\mathbb{Z} and it creates no envelope in this case by Theorem 1.

    Next, suppose that θ0π\theta_{0}\in\pi\mathbb{Z}. Then, calculation shows

    d(γ~θ0)\displaystyle d\left(\widetilde{\gamma}_{\theta_{0}}\right) =\displaystyle= 30t2(2t2+5t8)(4+25t6)4+25t6dt\displaystyle\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp 30t^{2}\left(2t^{2}+5t^{8}\right)}{(4+25t^{6})\sqrt{4+25t^{6}}}dt
    =\displaystyle= (2t2+5t8)4+25t6d(Θν~θ0)dt(t)dt=(2t2+5t8)4+25t6d(Θν~θ0),\displaystyle\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(2t^{2}+5t^{8})}{\sqrt{4+25t^{6}}}\frac{d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right)}{dt}(t)dt=\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(2t^{2}+5t^{8})}{\sqrt{4+25t^{6}}}d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0},}

    where double sign should be read in the same order. Therefore, the hyperplane family (φ~,ν~θ)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}_{{}_{\theta}}\right)} is creative. Set ω~(t)=(2t2+5t8)4+25t6.d(Θν~θ0)\widetilde{\omega}(t)=\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(2t^{2}+5t^{8})}{\sqrt{4+25t^{6}}}.d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right). By Theorem 1 and Theorem 2, (φ~,ν~θ0)\mathcal{H}_{(\widetilde{\varphi},\widetilde{\nu}_{\theta_{0}})} creates a unique envelope with the desired form

    f~(t)\displaystyle\widetilde{f}(t) =\displaystyle= ω~(t)+γ~θ0(t)ν~θ0(t)\displaystyle\widetilde{\omega}(t)+\widetilde{\gamma}_{\theta_{0}}(t)\widetilde{\nu}_{\theta_{0}}(t)
    =\displaystyle= (2t2+5t8)4+25t6(2,5t3)+3t54+25t6(5t3,±2)\displaystyle\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(2t^{2}+5t^{8})}{4+25t^{6}}\left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}2,{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}5t^{3}\right)+\frac{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}3t^{5}}{4+25t^{6}}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}5t^{3},{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\pm}2)
    =\displaystyle= 14+25t6(4t2+10t8+15t8,10t5+25t116t5)\displaystyle\frac{1}{4+25t^{6}}\left(4t^{2}+10t^{8}+15t^{8},10t^{5}+25t^{11}-6t^{5}\right)
    =\displaystyle= (t2,t5),\displaystyle\left(t^{2},t^{5}\right),

    where for each tt\in\mathbb{R} the cotangent vector (2t2+5t8)4+25t6d(Θν~θ0)\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(2t^{2}+5t^{8})}{\sqrt{4+25t^{6}}}d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right) and the vector (2t2+5t8)4+25t6Rπ/2ν~θ0(t)\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(2t^{2}+5t^{8})}{\sqrt{4+25t^{6}}}R_{\pi/2}\circ\widetilde{\nu}_{\theta_{0}}(t) in the vector space 2\mathbb{R}^{2} are identified (see Figure 13).

    Refer to caption
    Figure 13. Figure for Example 4.1 (4) in the case θ02π\theta_{0}\in 2\pi\mathbb{Z}.

    In the case θ0=0\theta_{0}=0, consider the mapping Ω~:TS1\widetilde{\Omega}:\mathbb{R}\to T^{*}S^{1} given in Definition 2 and Ω:J1(S1,)\Omega:\mathbb{R}\to J^{1}\left(S^{1},\mathbb{R}\right) given in Remark 1.1(1). Namely, consider the following two mappings.

    Ω~(t)\displaystyle\widetilde{\Omega}(t) =\displaystyle= (14+25t6(5t3,±2),30t2(2t2+5t8)(4+25t6)32),\displaystyle\left(\frac{1}{\sqrt{4+25t^{6}}}\left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}5t^{3},{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\pm}2\right),\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp 30t^{2}(2t^{2}+5t^{8})}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\left(4+25t^{6}\right)^{\frac{3}{2}}}\right),
    Ω(t)\displaystyle\Omega(t) =\displaystyle= (14+25t6(5t3,±2),3t54+25t6,30t2(2t2+5t8)(4+25t6)32).\displaystyle\left(\frac{1}{\sqrt{4+25t^{6}}}\left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}5t^{3},{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\pm}2\right),\frac{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp}3t^{5}}{\sqrt{4+25t^{6}}},\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp 30t^{2}(2t^{2}+5t^{8})}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\left(4+25t^{6}\right)^{\frac{3}{2}}}\right).

    Since d(γ~θ0)=(2t2+5t8)4+25t6d(Θν~θ0)d\left(\widetilde{\gamma}_{\theta_{0}}\right)=\frac{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mp(2t^{2}+5t^{8})}{\sqrt{4+25t^{6}}}d\left(\Theta\circ\widetilde{\nu}_{\theta_{0}}\right), the map-germ of Ω\Omega at any tt is nothing but an opening of the map-germ Ω~:(,t)TS1\widetilde{\Omega}:(\mathbb{R},t)\to T^{*}S^{1}. At t=0t=0, the map-germ of each of them is not immersive and has singular images.

    Set U={0}U=\mathbb{R}-\{0\}. It is easily seen that ν~θ0|U\widetilde{\nu}_{\theta_{0}}|_{U} is non-singular even in the case θ0π\theta_{0}\not\in\pi\mathbb{Z}. Hence, by Theorem 1 and Theorem 2, the hyperplane family (φ~|U,ν~θ0|U)\mathcal{H}_{\left(\widetilde{\varphi}|_{U},\widetilde{\nu}_{\theta_{0}}|_{U}\right)} creates a unique envelope f~θ0:U2\widetilde{f}_{\theta_{0}}:U\to\mathbb{R}^{2} even when θ0π\theta_{0}\not\in\pi\mathbb{Z} and limt0f~θ0(t)=\lim_{t\to 0}\|\widetilde{f}_{\theta_{0}}(t)\|=\infty when θ0π\theta_{0}\not\in\pi\mathbb{Z}.

Example 4.2 (Unit speed curves).
  1. (1)

    Let 𝐫:2{\bf r}:\mathbb{R}\to\mathbb{R}^{2} be a unit speed curve. As usual, set 𝐭(s)=𝐫(s){\bf t}(s)={\bf r}^{\prime}(s) and 𝐧(s){\bf n}(s) is defined from t(s) by rotating anticlockwise through π2\frac{\pi}{2}. The Serret-Frenet formulas for the plane curve 𝐫{\bf r} is as follows.

    {𝐭(s)=κ(s)𝐧(s)𝐧(s)=κ(s)𝐭(s).\left\{\begin{array}[]{cccc}{\bf t}^{\prime}(s)&=&{}\hfil&\kappa(s){\bf n}(s)\\ {\bf n}^{\prime}(s)&=&-\kappa(s){\bf t}(s).&{}\hfil\end{array}\right.

    Set φ~=𝐫\widetilde{\varphi}={\bf r} and ν~=𝐧\widetilde{\nu}={\bf n}. Then, the line family (φ~,ν~)=(𝐫,𝐧)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}=\mathcal{H}_{\left({\bf r},{\bf n}\right)} is the affine tangent line family of the curve 𝐫{\bf r}. In this case, the correspondence 𝐫(𝐫,𝐧){\bf r}\mapsto\mathcal{H}_{\left({\bf r},{\bf n}\right)} may be regarded as the Legendre transformation of the given curve 𝐫{\bf r}. Set γ~(s)=φ~(s)ν~(s)\widetilde{\gamma}(s)=\widetilde{\varphi}(s)\cdot\widetilde{\nu}(s). Then,

    γ~(s)=𝐫(s)(κ(s)𝐭(s))=(𝐫(s)𝐭(s))(Θ𝐭ν~)(s),\widetilde{\gamma}^{\prime}(s)={\bf r}(s)\cdot\left(-\kappa(s){\bf t}(s)\right)={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}-}\left({\bf r}(s)\cdot{\bf t}(s)\right)\left(\Theta_{{\bf t}}\circ\widetilde{\nu}\right)^{\prime}(s){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0},}

    where ν~(s)=(cosΘ𝐭ν~(s),sinΘ𝐭ν~(s))\widetilde{\nu}(s)=\left(\cos\Theta_{{\bf t}}\circ\widetilde{\nu}(s),\sin\Theta_{{\bf t}}\circ\widetilde{\nu}(s)\right). Therefore, by Theorem 1, the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope.

    Suppose that the set of regular points of ν~\widetilde{\nu} is dense, that is to say, the set {s|κ(s)0}\{s\in\mathbb{R}\;|\;\kappa(s)\neq 0\} is dense. Then, by Theorem 2, the created envelopes are unique. By Theorem 1, the unique envelope is as follows (see Figure 14).

    f~(s)\displaystyle\widetilde{f}(s) =\displaystyle= ω~(s)+γ~(s)ν~(s)\displaystyle\widetilde{\omega}(s)+\widetilde{\gamma}(s)\cdot\widetilde{\nu}(s)
    =\displaystyle= (𝐫(s)𝐭(s))𝐭(s)+(𝐫(s)𝐧(s))𝐧(s)\displaystyle\left({\bf r}(s)\cdot{\bf t}(s)\right){\bf t}(s)+\left({\bf r}(s)\cdot{\bf n}(s)\right){\bf n}(s)
    =\displaystyle= 𝐫(s).\displaystyle{\bf r}(s).

    Notice that if there is a point ss\in\mathbb{R} such that κ(s)=0\kappa(s)=0, then the full discriminant of the line family is different from the unique desired envelope since the full discriminant includes the affine tangent line at ss. This is one of advantages of our method. The correspondence

    (𝐫,𝐧)𝐫\mathcal{H}_{\left({\bf r},{\bf n}\right)}\mapsto{\bf r}

    may be regarded as the inverse Legendre transformation for plane curves.

    Next, suppose that the set of regular points of ν~\widetilde{\nu} is not dense. Then, there exists an open interval (a,b)(a,b) such that κ(s)=0\kappa(s)=0 for any s(a,b)s\in(a,b). Then, for any s(a,b)s\in(a,b) and any function α:\alpha:\mathbb{R}\to\mathbb{R} such that α((a,b))={0}\alpha(\mathbb{R}-(a,b))=\{0\}, it follows

    γ~(s)=α(s)(Θ𝐭ν~)(s).\widetilde{\gamma}^{\prime}(s)=\alpha(s)\left(\Theta_{{\bf t}}\circ\widetilde{\nu}\right)^{\prime}(s).

    By Theorem 1,

    f~(s)\displaystyle\widetilde{f}(s) =\displaystyle= ω~(s)+γ~(s)ν~(s)\displaystyle\widetilde{\omega}(s)+\widetilde{\gamma}(s)\cdot\widetilde{\nu}(s)
    =\displaystyle= α(s)𝐭(s)+(𝐫(s)𝐧(s))𝐧(s)\displaystyle\alpha(s){\bf t}(s)+\left({\bf r}(s)\cdot{\bf n}(s)\right){\bf n}(s)
    =\displaystyle= ((α(s)(𝐫(s)𝐭(s)))+(𝐫(s)𝐭(s)))𝐭(s)+(𝐫(s)𝐧(s))𝐧(s)\displaystyle\left(\left(\alpha(s)-\left({\bf r}(s)\cdot{\bf t}(s)\right)\right)+\left({\bf r}(s)\cdot{\bf t}(s)\right)\right){\bf t}(s)+\left({\bf r}(s)\cdot{\bf n}(s)\right){\bf n}(s)
    =\displaystyle= 𝐫(s)+β(s)𝐭(s),\displaystyle{\bf r}(s)+\beta(s){\bf t}(s),

    where β(s)=α(s)(𝐫(s)𝐭(s))\beta(s)=\alpha(s)-\left({\bf r}(s)\cdot{\bf t}(s)\right). Hence, in this case, the inverse Legendre transformation does not work well.

Refer to caption
Figure 14. Example 4.2 (1).
  1. (2)

    Let 𝐫:3{\bf r}:\mathbb{R}\to\mathbb{R}^{3} be a unit speed space curve. As usual, set 𝐭(s)=𝐫(s){\bf t}(s)={\bf r}^{\prime}(s) and assume 𝐭(s)>0||{\bf t}^{\prime}(s)||>0 for any ss\in\mathbb{R} so that the principal normal vector 𝐧(s){\bf n}(s) can be defined by 𝐭(s)=𝐭(s)𝐧(s){\bf t}^{\prime}(s)=||{\bf t}^{\prime}(s)||{\bf n}(s). As usual, the binormal vector 𝐛(s){\bf b}(s) is defined by det(𝐭(s),𝐧(s),𝐛(s))=1\det\left({\bf t}(s),{\bf n}(s),{\bf b}(s)\right)=1. The Serret-Frenet formulas for the space curve 𝐫{\bf r} is as follows.

    {𝐭(s)=κ(s)𝐧(s)𝐧(s)=κ(s)𝐭(s)+τ(s)𝐛(s)𝐛(s)=τ(s)𝐧(s).\left\{\begin{array}[]{ccccc}{\bf t}^{\prime}(s)&=&{}\hfil&\kappa(s){\bf n}(s)&{}\hfil\\ {\bf n}^{\prime}(s)&=&-\kappa(s){\bf t}(s)&{}\hfil&+\tau(s){\bf b}(s)\\ {\bf b}^{\prime}(s)&=&{}\hfil&-\tau(s){\bf n}(s)&{}.\end{array}\right.

    Define φ~:23\widetilde{\varphi}:\mathbb{R}^{2}\to\mathbb{R}^{3} and ν~:2S2\widetilde{\nu}:\mathbb{R}^{2}\to S^{2} by φ~(s,u)=𝐫(s)\widetilde{\varphi}(s,u)={\bf r}(s) and ν~(s,u)=𝐛(s)\widetilde{\nu}(s,u)={\bf b}(s) respectively. Then, the plane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is the family of osculating planes of the space curve 𝐫{\bf r}. Set γ~(s,u)=𝐫(s)𝐛(s)\widetilde{\gamma}(s,u)={\bf r}(s)\cdot{\bf b}(s). Then, all of the following six identities are clear.

    γ~s(s,u)=𝐫(s)(τ(s)𝐧(s)),γ~u(s,u)=0,(Θ𝐭ν~)s(s,u)=0,\frac{\partial\widetilde{\gamma}}{\partial s}(s,u)={\bf r}(s)\cdot\left(-\tau(s){\bf n}(s)\right),\quad\frac{\partial\widetilde{\gamma}}{\partial u}(s,u)=0,\quad\frac{\partial\left(\Theta_{{\bf t}}\circ\widetilde{\nu}\right)}{\partial s}(s,u)=0,\quad
    (Θ𝐭ν~)u(s,u)=0,(Θ𝐧ν~)s(s,u)=τ(s),(Θ𝐧ν~)u(s,u)=0.\frac{\partial\left(\Theta_{{\bf t}}\circ\widetilde{\nu}\right)}{\partial u}(s,u)=0,\quad\frac{\partial\left(\Theta_{{\bf n}}\circ\widetilde{\nu}\right)}{\partial s}(s,u)=-\tau(s),\quad\frac{\partial\left(\Theta_{{\bf n}}\circ\widetilde{\nu}\right)}{\partial u}(s,u)=0.

    Therefore, we have the following.

    γ~s(s,u)\displaystyle\frac{\partial\widetilde{\gamma}}{\partial s}(s,u) =\displaystyle= α1(s,u)(Θ𝐭ν~)s(s,u)+(𝐫(s)𝐧(s))(Θ𝐧ν~)s(s,u),\displaystyle\alpha_{1}(s,u)\frac{\partial\left(\Theta_{{\bf t}}\circ\widetilde{\nu}\right)}{\partial s}(s,u)+\left({\bf r}(s)\cdot{\bf n}(s)\right)\frac{\partial\left(\Theta_{{\bf n}}\circ\widetilde{\nu}\right)}{\partial s}(s,u),
    γ~u(s,u)\displaystyle\frac{\partial\widetilde{\gamma}}{\partial u}(s,u) =\displaystyle= α2(s,u)(Θ𝐭ν~)u(s,u)+α3(s,u)(Θ𝐧ν~)u(s,u),\displaystyle\alpha_{2}(s,u)\frac{\partial\left(\Theta_{{\bf t}}\circ\widetilde{\nu}\right)}{\partial u}(s,u)+\alpha_{3}(s,u)\frac{\partial\left(\Theta_{{\bf n}}\circ\widetilde{\nu}\right)}{\partial u}(s,u),

    where α1,α2,α3:2\alpha_{1},\alpha_{2},\alpha_{3}:\mathbb{R}^{2}\to\mathbb{R} are arbitrary functions. Thus, by Theorem 1, the plane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} creates an envelope if and only if (𝐫(s)𝐧(s))=α3(s,u)\left({\bf r}(s)\cdot{\bf n}(s)\right)=\alpha_{3}(s,u) and α1(s,u)=α2(s,u)\alpha_{1}(s,u)=\alpha_{2}(s,u). Therefore, again by Theorem 1, we have the following concrete expression of the created envelopes.

    f~(s,u)\displaystyle\widetilde{f}(s,u)
    =\displaystyle= ω~(s,u)+γ~(s)ν~(s)\displaystyle\widetilde{\omega}(s,u)+\widetilde{\gamma}(s)\widetilde{\nu}(s)
    =\displaystyle= (𝐫(s)𝐧(s))𝐧(s)+α(s,u)𝐭(s)+(𝐫(s)𝐛(s))𝐛(s)\displaystyle\left({\bf r}(s)\cdot{\bf n}(s)\right){\bf n}(s)+\alpha(s,u){\bf t}(s)+\left({\bf r}(s)\cdot{\bf b}(s)\right){\bf b}(s)
    =\displaystyle= (𝐫(s)𝐧(s))𝐧(s)+(𝐫(s)𝐭(s))𝐭(s)+(α(s,u)(𝐫(s)𝐭(s)))𝐭(s)+(𝐫(s)𝐛(s))𝐛(s)\displaystyle\left({\bf r}(s)\cdot{\bf n}(s)\right){\bf n}(s)+\left({\bf r}(s)\cdot{\bf t}(s)\right){\bf t}(s)+\left(\alpha(s,u)-\left({\bf r}(s)\cdot{\bf t}(s)\right)\right){\bf t}(s)+\left({\bf r}(s)\cdot{\bf b}(s)\right){\bf b}(s)
    =\displaystyle= 𝐫(s)+β(s,u)𝐭(s),\displaystyle{\bf r}(s)+\beta(s,u){\bf t}(s),

    where α(s,u)=α1(s,u)=α2(s,u)\alpha(s,u)=\alpha_{1}(s,u)=\alpha_{2}(s,u) and β(s,u)=α(s,u)(𝐫(s)𝐭(s))\beta(s,u)=\alpha(s,u)-\left({\bf r}(s)\cdot{\bf t}(s)\right). All envelopes created by the osculating family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} can be exactly expressed as above. Hence, for example, both the tangent developable of 𝐫{\bf r} (in the case β(s,u)=u\beta(s,u)=u) and the space curve 𝐫{\bf r} (in the case β(s,u)=0\beta(s,u)=0) are envelopes of (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. Not only these two, there are uncountably many envelopes created by (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. All envelopes for the osculating plane family are created only by the given curve 𝐫{\bf r} and its unit tangent curve 𝐭{\bf t}.

    Next, we consider envelopes created by (𝐫,𝐛)\mathcal{H}_{\left({\bf r},{\bf b}\right)} and (f~,𝐧)\mathcal{H}_{\left(\widetilde{f},{\bf n}\right)}. Namely, we obtain all solutions g~(s,u){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}(s,u) for the following system of PDEs with one constraint condition.

    {g~s(s,u)𝐛(s)=0,g~u(s,u)𝐛(s)=0,g~s(s,u)𝐧(s)=0,g~u(s,u)𝐧(s)=0,(g~(s,u)𝐫(s))𝐛(s)=0.\left\{\begin{array}[]{ccc}\frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial s}(s,u)\cdot{\bf b}(s)&=&0,\\ \frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial u}(s,u)\cdot{\bf b}(s)&=&0,\\ \frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial s}(s,u)\cdot{\bf n}(s)&=&0,\\ \frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial u}(s,u)\cdot{\bf n}(s)&=&0,\\ \left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}(s,u)-{\bf r}(s)\right)\cdot{\bf b}(s)&=&0.\end{array}\right.

    Since κ(s)>0\kappa(s)>0 for any ss\in\mathbb{R} and

    f~s(s,u)\displaystyle\frac{\partial\widetilde{f}}{\partial s}(s,u) =\displaystyle= 𝐭(s)+βs(s,u)𝐭(s)+β(s,u)(κ(s)𝐧(s)),\displaystyle{\bf t}(s)+\frac{\partial\beta}{\partial s}(s,u){\bf t}(s)+\beta(s,u)\left(\kappa(s){\bf n}(s)\right),
    f~u(s,u)\displaystyle\frac{\partial\widetilde{f}}{\partial u}(s,u) =\displaystyle= βu(s,u)𝐭(s),\displaystyle\frac{\partial\beta}{\partial u}(s,u){\bf t}(s),

    if f~\widetilde{f} itself is a solution of the above system of PDEs, then β(s,u)\beta(s,u) must be constant 0. Conversely, it is clear that 𝐫{\bf r} itself is a solution of the above system of PDEs with one constraint condition. Therefore, for the above system of PDEs with one constraint condition, there are no solutions except for the trivial solution 𝐫{\bf r}. This implies that even for a space curve 𝐫:3{\bf r}:\mathbb{R}\to\mathbb{R}^{3}, the inverse Legendre transformation

    (𝐫,{𝐛,𝐧})𝐫\mathcal{H}_{\left({\bf r},\{{\bf b},{\bf n}\}\right)}\mapsto{\bf r}

    works well.

    Finally, we consider envelopes created by (𝐫,𝐛)\mathcal{H}_{\left({\bf r},{\bf b}\right)} and (f~,𝐭)\mathcal{H}_{\left(\widetilde{f},{\bf t}\right)}. Namely, we obtain all solutions g~(s,u){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}(s,u) for the following system of PDEs with one constraint condition.

    {g~s(s,u)𝐛(s)=0,g~u(s,u)𝐛(s)=0,g~s(s,u)𝐭(s)=0,g~u(s,u)𝐭(s)=0,(g~(s,u)𝐫(s))𝐛(s)=0.\left\{\begin{array}[]{ccc}\frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial s}(s,u)\cdot{\bf b}(s)&=&0,\\ \frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial u}(s,u)\cdot{\bf b}(s)&=&0,\\ \frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial s}(s,u)\cdot{\bf t}(s)&=&0,\\ \frac{\partial{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}}{\partial u}(s,u)\cdot{\bf t}(s)&=&0,\\ \left({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{g}}(s,u)-{\bf r}(s)\right)\cdot{\bf b}(s)&=&0.\end{array}\right.

    By the above calculations, if f~\widetilde{f} is a solution of the above system of PDEs, then both 1+βs(s,u)=01+\frac{\partial\beta}{\partial s}(s,u)=0 and βu(s,u)=0\frac{\partial\beta}{\partial u}(s,u)=0 must be satisfied. It follows β(s,u)=s+c\beta(s,u)=-s+c (c)(c\in\mathbb{R}). It is easily seen that for any cc\in\mathbb{R}, the space curve s𝐫(s)+(s+c)𝐭(s)s\mapsto{\bf r}(s)+(-s+c){\bf t}(s) is a solution of the above system of PDEs with one constraint condition. . Thus, in this case, the system of PDEs with one constraint condition has uncountably many solutions.

Example 4.3.
  1. (1)

    (The shoe surface : Example 1 of [3]) In this example, along the general theory developed in this paper, we start from making several general formulas for the envelope created by the affine tangent plane family of the surface having the form φ~:23\widetilde{\varphi}:\mathbb{R}^{2}\to\mathbb{R}^{3},  φ~(x,y)=(x,y,φ~1(x,y))\widetilde{\varphi}(x,y)=(x,\,y,\,\widetilde{\varphi}_{1}(x,y)) such that the origin (0,0)(0,0) is a singular point of the function φ~1:2\widetilde{\varphi}_{1}:\mathbb{R}^{2}\to\mathbb{R} and there are no other singular points of φ~1\widetilde{\varphi}_{1}. Then, by calculating the obtained general formulas in the case of the shoe surface φ~(x,y)=(x,y,13x312y2)\widetilde{\varphi}(x,y)=\left(x,\,y,\,\frac{1}{3}x^{3}-\frac{1}{2}y^{2}\right), just by calculations, we confirm that the concrete representation form of the envelope created by the affine tangent plane family of the shoe surface φ~\widetilde{\varphi} is actually the shoe surface itself.

    Let φ~:23\widetilde{\varphi}:\mathbb{R}^{2}\to\mathbb{R}^{3} be the mapping having the form φ~(x,y)=(x,y,φ~1(x,y))\widetilde{\varphi}(x,y)=(x,y,\widetilde{\varphi}_{1}(x,y)), where the function φ~1:2\widetilde{\varphi}_{1}:\mathbb{R}^{2}\to\mathbb{R} has a unique singularity at the origin, namely φ~1x(0,0)=φ~1y(0,0)=0\frac{\partial\widetilde{\varphi}_{1}}{\partial x}(0,0)=\frac{\partial\widetilde{\varphi}_{1}}{\partial y}(0,0)=0 and (φ~1x(x,y),φ~1y(x,y))(0,0)\left(\frac{\partial\widetilde{\varphi}_{1}}{\partial x}(x,y),\,\frac{\partial\widetilde{\varphi}_{1}}{\partial y}(x,y)\right)\neq(0,0) for any (x,y)2{(0,0)}(x,y)\in\mathbb{R}^{2}-\{(0,0)\}. Then, the mapping ν~:2S2\widetilde{\nu}:\mathbb{R}^{2}\to S^{2} defined by

    ν~(x,y)=φ~1x(x,y)×φ~1y(x,y)φ~1x(x,y)×φ~1y(x,y)=(φ~1x,φ~1y, 1)(φ~1x)2+(φ~1y)2+1\widetilde{\nu}(x,y)=\frac{\frac{\partial\widetilde{\varphi}_{1}}{\partial x}(x,y)\times\frac{\partial\widetilde{\varphi}_{1}}{\partial y}(x,y)}{\parallel\frac{\partial\widetilde{\varphi}_{1}}{\partial x}(x,y)\times\frac{\partial\widetilde{\varphi}_{1}}{\partial y}(x,y)\parallel}=\frac{\left(-\frac{\partial\widetilde{\varphi}_{1}}{\partial x},\,-\frac{\partial\widetilde{\varphi}_{1}}{\partial y},\,1\right)}{\sqrt{\left(\frac{\partial\widetilde{\varphi}_{1}}{\partial x}\right)^{2}+\left(\frac{\partial\widetilde{\varphi}_{1}}{\partial y}\right)^{2}+1}}

    is a Gauss mapping of the tangent plane family of φ~\widetilde{\varphi}. Here, the tangent plane family of φ~\widetilde{\varphi} is (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. Let (x0,y0)(x_{0},y_{0}) be an arbitrary point of 2{(0,0)}\mathbb{R}^{2}-\{(0,0)\}. Then, by the assumption on the function φ~1\widetilde{\varphi}_{1}, it follows that ν~(x0,y0)(0, 0, 1)\widetilde{\nu}(x_{0},y_{0})\neq(0,\,0,\,1). Set

    𝐯0(x0,y0)\displaystyle{\bf v}_{0}(x_{0},y_{0}) =\displaystyle= ν~(x0,y0),\displaystyle\widetilde{\nu}(x_{0},y_{0}),
    𝐯1(x0,y0)\displaystyle{\bf v}_{1}(x_{0},y_{0}) =\displaystyle= (0, 0, 1)((0, 0, 1)𝐯0(x0,y0))𝐯0(x0,y0)(0, 0, 1)((0, 0, 1)𝐯0(x0,y0))𝐯0(x0,y0),\displaystyle\frac{(0,\,0,\,1)-\left((0,\,0,\,1)\cdot{\bf v}_{0}(x_{0},y_{0})\right){\bf v}_{0}(x_{0},y_{0})}{\parallel(0,\,0,\,1)-\left((0,\,0,\,1)\cdot{\bf v}_{0}(x_{0},y_{0})\right){\bf v}_{0}(x_{0},y_{0})\parallel},
    𝐯2(x0,y0)\displaystyle{\bf v}_{2}(x_{0},y_{0}) =\displaystyle= 𝐯0(x0,y0)×𝐯1(x0,y0).\displaystyle{\bf v}_{0}(x_{0},y_{0})\times{\bf v}_{1}(x_{0},y_{0}).

    Then, 𝐯0(x0,y0),𝐯1(x0,y0),𝐯2(x0,y0)\langle{\bf v}_{0}(x_{0},y_{0}),{\bf v}_{1}(x_{0},y_{0}),{\bf v}_{2}(x_{0},y_{0})\rangle is an orthonormal basis of 3\mathbb{R}^{3}, and under the identification of two vector spaces 3\mathbb{R}^{3} and Tν~(x0,y0)3T_{\widetilde{\nu}(x_{0},y_{0})}\mathbb{R}^{3}, 𝐯1(x0,y0),𝐯2(x0,y0)\langle{\bf v}_{1}(x_{0},y_{0}),{\bf v}_{2}(x_{0},y_{0})\rangle is an orthonormal basis of the tangent vector space Tν~(x0,y0)S2T_{\widetilde{\nu}(x_{0},y_{0})}S^{2}. Let ε\varepsilon be a sufficiently small positive number and denote the set {Θ1𝐯1(x0,y0)+Θ2𝐯2(x0,y0)|ε<Θ1,Θ2<ε}\{\Theta_{1}{\bf v}_{1}(x_{0},y_{0})+\Theta_{2}{\bf v}_{2}(x_{0},y_{0})\,|\,-\varepsilon<\Theta_{1},\Theta_{2}<\varepsilon\} by VV^{\prime}. Let exp:VS2\exp:V^{\prime}\to S^{2} be the restriction of the exponential mapping at ν~(x0,y0)\widetilde{\nu}(x_{0},y_{0}) to VV^{\prime} and set V=exp(V)V=\exp(V^{\prime}). Let (V,(Θ1,Θ2))\left(V,\left(\Theta_{1},\Theta_{2}\right)\right) be the normal coordinate neighborhood at ν~(x0,y0)\widetilde{\nu}(x_{0},y_{0}) defined by exp1:VV\exp^{-1}:V\to V^{\prime}. Set

    γ~(x,y)=φ~(x,y)ν~(x,y)=xφ~1xyφ~1y+φ~1(x,y)(φ~1x)2+(φ~1y)2+1.\widetilde{\gamma}(x,y)=\widetilde{\varphi}(x,y)\cdot\widetilde{\nu}(x,y)=\frac{-x\frac{\partial\widetilde{\varphi}_{1}}{\partial x}-y\frac{\partial\widetilde{\varphi}_{1}}{\partial y}+\widetilde{\varphi}_{1}(x,y)}{\sqrt{\left(\frac{\partial\widetilde{\varphi}_{1}}{\partial x}\right)^{2}+\left(\frac{\partial\widetilde{\varphi}_{1}}{\partial y}\right)^{2}+1}}.

    Since ν~:2S2\widetilde{\nu}:\mathbb{R}^{2}\to S^{2} is a Gauss mapping of φ~:23\widetilde{\varphi}:\mathbb{R}^{2}\to\mathbb{R}^{3}, we have

    γ~x(x0,y0)\displaystyle\frac{\partial\widetilde{\gamma}}{\partial x}(x_{0},y_{0})
    =\displaystyle= φ~(x0,y0)ν~x(x0,y0)\displaystyle\widetilde{\varphi}(x_{0},y_{0})\cdot\frac{\partial\widetilde{\nu}}{\partial x}(x_{0},y_{0})
    =\displaystyle= (φ~(x0,y0)𝐯1(x0,y0))(Θ1ν~)x(x0,y0)+(φ~(x0,y0)𝐯2(x0,y0))(Θ2ν~)x(x0,y0)\displaystyle\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0})+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0})

    and

    γ~y(x0,y0)\displaystyle\frac{\partial\widetilde{\gamma}}{\partial y}(x_{0},y_{0})
    =\displaystyle= φ~(x0,y0)ν~y(x0,y0)\displaystyle\widetilde{\varphi}(x_{0},y_{0})\cdot\frac{\partial\widetilde{\nu}}{\partial y}(x_{0},y_{0})
    =\displaystyle= (φ~(x0,y0)𝐯1(x0,y0))(Θ1ν~)y(x0,y0)+(φ~(x0,y0)𝐯2(x0,y0))(Θ2ν~)y(x0,y0).\displaystyle\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0})+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0}).

    Thus, as the equality of 22-dimensional cotangent vectors of T(x0,y0)2T^{*}_{(x_{0},y_{0})}\mathbb{R}^{2}, we have the following equality.

    dγ~\displaystyle d\widetilde{\gamma} =\displaystyle= γ~x(x0,y0)dx+γ~y(x0,y0)dy\displaystyle\frac{\partial\widetilde{\gamma}}{\partial x}(x_{0},y_{0})dx+\frac{\partial\widetilde{\gamma}}{\partial y}(x_{0},y_{0})dy
    =\displaystyle= (φ~(x0,y0)𝐯1(x0,y0))d(Θ1ν~)+(φ~(x0,y0)𝐯2(x0,y0))d(Θ2ν~).\displaystyle\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)d\left(\Theta_{1}\circ\widetilde{\nu}\right)+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)d\left(\Theta_{2}\circ\widetilde{\nu}\right).

    Set U=2{(0,0)}U=\mathbb{R}^{2}-\{(0,0)\} and assume that the singular set of ν~\widetilde{\nu} is of Lebesgue measure zero. Then, since (x0,y0)(x_{0},y_{0}) is an arbitrary point of UU, by Theorem 1 (1) and Theorem 2, it follows that (φ~|U,ν~|U)\mathcal{H}_{(\widetilde{\varphi}|_{U},\widetilde{\nu}|_{U})} creates a unique envelope. Set

    ω~(x0,y0)=(φ~(x0,y0)𝐯1(x0,y0))d(Θ1ν~)+(φ~(x0,y0)𝐯2(x0,y0))d(Θ2ν~).\widetilde{\omega}(x_{0},y_{0})=\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)d\left(\Theta_{1}\circ\widetilde{\nu}\right)+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)d\left(\Theta_{2}\circ\widetilde{\nu}\right).

    Then, under the canonical identifications

    Tν~(x0,y0)S2Tν~(x0,y0)S2Tν~(x0,y0)33,T^{*}_{\widetilde{\nu}(x_{0},y_{0})}S^{2}\cong T_{\widetilde{\nu}(x_{0},y_{0})}S^{2}\subset T_{\widetilde{\nu}(x_{0},y_{0})}\mathbb{R}^{3}\cong\mathbb{R}^{3},

    the 22-dimensional cotangent vector

    ω~(x0,y0)=(φ~(x0,y0)𝐯1(x0,y0))dΘ1+(φ~(x0,y0)𝐯2(x0,y0))dΘ2\widetilde{\omega}(x_{0},y_{0})=\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)d\Theta_{1}+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)d\Theta_{2}

    may be regarded as the following 33-dimensional vector (denoted by the same symbol ω~(x0,y0)\widetilde{\omega}(x_{0},y_{0})).

    ω~(x0,y0)=(φ~(x0,y0)𝐯1(x0,y0))𝐯1(x0,y0)+(φ~(x0,y0)𝐯2(x0,y0))𝐯2(x0,y0).\widetilde{\omega}(x_{0},y_{0})=\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right){\bf v}_{1}(x_{0},y_{0})+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right){\bf v}_{2}(x_{0},y_{0}).

    Therefore, by Theorem 1 (2), the envelope vector at (x0,y0)(x_{0},y_{0}) must have the following form:

    f~(x0,y0)\displaystyle\widetilde{f}(x_{0},y_{0}) =\displaystyle= ω~(x0,y0)+γ~(x0,y0)ν~(x0,y0)\displaystyle\widetilde{\omega}(x_{0},y_{0})+\widetilde{\gamma}(x_{0},y_{0})\widetilde{\nu}(x_{0},y_{0})
    =\displaystyle= (φ~(x0,y0)𝐯1(x0,y0))𝐯1(x0,y0)+(φ~(x0,y0)𝐯2(x0,y0))𝐯2(x0,y0)\displaystyle\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right){\bf v}_{1}(x_{0},y_{0})+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right){\bf v}_{2}(x_{0},y_{0})
    +(φ~(x0,y0)𝐯0(x0,y0))𝐯0(x0,y0)\displaystyle\qquad\qquad+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{0}(x_{0},y_{0})\right){\bf v}_{0}(x_{0},y_{0})
    =\displaystyle= φ~(x0,y0).\displaystyle\widetilde{\varphi}(x_{0},y_{0}).

    By continuity, it follows that f~=φ~\widetilde{f}=\widetilde{\varphi} is the unique envelope created by the given plane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}.

    Next, we apply the above formulas to the shoe surface. The shoe surface is the image of φ~:23\widetilde{\varphi}:\mathbb{R}^{2}\to\mathbb{R}^{3} defined by φ~(x,y)=(x,y,13x312y2)\widetilde{\varphi}(x,y)=\left(x,y,\frac{1}{3}x^{3}-\frac{1}{2}y^{2}\right). Set φ~1(x,y)=13x312y2\widetilde{\varphi}_{1}(x,y)=\frac{1}{3}x^{3}-\frac{1}{2}y^{2}. Then, the origin (0,0)(0,0) is a unique singular point of φ~1\widetilde{\varphi}_{1}. For the given φ~\widetilde{\varphi}, we have ν~(x,y)=φ~x(x,y)×φ~y(x,y)φ~x(x,y)×φ~y(x,y)=(x2,y, 1)x4+y2+1\widetilde{\nu}(x,y)=\frac{\frac{\partial\widetilde{\varphi}}{\partial x}(x,y)\times\frac{\partial\widetilde{\varphi}}{\partial y}(x,y)}{||\frac{\partial\widetilde{\varphi}}{\partial x}(x,y)\times\frac{\partial\widetilde{\varphi}}{\partial y}(x,y)||}=\frac{\left(-x^{2},\,y,\,1\right)}{\sqrt{x^{4}+y^{2}+1}}. It is easily confirmed that the set consisting of regular points of ν~\widetilde{\nu} is dense. In fact, it is known that any singularity of ν~\widetilde{\nu} is a fold singularity (see [3]). Set U=2{(0,0)}U=\mathbb{R}^{2}-\{(0,0)\} and take an arbitrary point (x0,y0)(x_{0},y_{0}) of UU. For the shoe surface φ~\widetilde{\varphi}, we set

    𝐯0(x0,y0)\displaystyle{\bf v}_{0}(x_{0},y_{0}) =\displaystyle= ν~(x0,y0)=(x02,y0, 1)x04+y02+1,\displaystyle\widetilde{\nu}(x_{0},y_{0})=\frac{\left(-x_{0}^{2},\,y_{0},\,1\right)}{\sqrt{x_{0}^{4}+y_{0}^{2}+1}},
    𝐯1(x0,y0)\displaystyle{\bf v}_{1}(x_{0},y_{0}) =\displaystyle= (0, 0, 1)((0, 0, 1)𝐯0(x0,y0))𝐯0(x0,y0)(0, 0, 1)((0, 0, 1)𝐯0(x0,y0))𝐯0(x0,y0)=(x02,y0,x04+y02)(x04+y02)(x04+y02+1),\displaystyle\frac{(0,\,0,\,1)-\left((0,\,0,\,1)\cdot{\bf v}_{0}(x_{0},y_{0})\right){\bf v}_{0}(x_{0},y_{0})}{\parallel(0,\,0,\,1)-\left((0,\,0,\,1)\cdot{\bf v}_{0}(x_{0},y_{0})\right){\bf v}_{0}(x_{0},y_{0})\parallel}=\frac{\left(x_{0}^{2},\,-y_{0},\,x_{0}^{4}+y_{0}^{2}\right)}{\sqrt{\left(x_{0}^{4}+y_{0}^{2}\right)\left(x_{0}^{4}+y_{0}^{2}+1\right)}},
    𝐯2(x0,y0)\displaystyle{\bf v}_{2}(x_{0},y_{0}) =\displaystyle= 𝐯0(x0,y0)×𝐯1(x0,y0)=(y0,x02, 0)x04+y02.\displaystyle{\bf v}_{0}(x_{0},y_{0})\times{\bf v}_{1}(x_{0},y_{0})=\frac{\left(y_{0},\,x_{0}^{2},\,0\right)}{\sqrt{x_{0}^{4}+y_{0}^{2}}}.

    By calculation, we have

    φ~(x0,y0)𝐯1(x0,y0)\displaystyle\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0}) =\displaystyle= x03y02+(13x0312y02)(x04+y02)(x04+y02)12(x04+y02+1)12,\displaystyle\frac{x_{0}^{3}-y_{0}^{2}+\left(\frac{1}{3}x_{0}^{3}-\frac{1}{2}y_{0}^{2}\right)\left(x_{0}^{4}+y_{0}^{2}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}},
    φ~(x0,y0)𝐯2(x0,y0)\displaystyle\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0}) =\displaystyle= x0y0+x02y0(x04+y02)12.\displaystyle\frac{x_{0}y_{0}+x_{0}^{2}y_{0}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}}.

    Let (V,(Θ1,Θ2))(V,\left(\Theta_{1},\Theta_{2}\right)) be the normal cordinate neighborhood of S2S^{2} defined above. By calculations using the following two identities

    ν~x(x0,y0)\displaystyle\frac{\partial\widetilde{\nu}}{\partial x}(x_{0},y_{0}) =\displaystyle= 𝐯1(x0,y0)(Θ1ν~)x(x0,y0)+𝐯2(x0,y0)(Θ2ν~)x(x0,y0),\displaystyle{\bf v}_{1}(x_{0},y_{0})\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0})+{\bf v}_{2}(x_{0},y_{0})\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0}),
    ν~y(x0,y0)\displaystyle\frac{\partial\widetilde{\nu}}{\partial y}(x_{0},y_{0}) =\displaystyle= 𝐯1(x0,y0)(Θ1ν~)y(x0,y0)+𝐯2(x0,y0)(Θ2ν~)y(x0,y0),\displaystyle{\bf v}_{1}(x_{0},y_{0})\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0})+{\bf v}_{2}(x_{0},y_{0})\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0}),

    we have the following.

    (Θ1ν~)x(x0,y0)\displaystyle\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0}) =\displaystyle= 2x03(x04+y02)12(x04+y02+1),\displaystyle\frac{-2x_{0}^{3}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)},
    (Θ2ν~)x(x0,y0)\displaystyle\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0}) =\displaystyle= 2x0y02x0y032x05y0(x04+y02)12(x04+y02+1)32,\displaystyle\frac{-2x_{0}y_{0}-2x_{0}y_{0}^{3}-2x_{0}^{5}y_{0}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{3}{2}}},
    (Θ1ν~)y(x0,y0)\displaystyle\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0}) =\displaystyle= y0(x04+y02)12(x04+y02+1),\displaystyle\frac{-y_{0}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)},
    (Θ2ν~)y(x0,y0)\displaystyle\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0}) =\displaystyle= x02(x04+y02)12(x04+y02+1)12.\displaystyle\frac{x_{0}^{2}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}.

    On the other hand, from the form γ~(x,y)=φ~(x,y)ν~(x,y)=23x3+12y2x4+y2+1\widetilde{\gamma}(x,y)=\widetilde{\varphi}(x,y)\cdot\widetilde{\nu}(x,y)=\frac{-\frac{2}{3}x^{3}+\frac{1}{2}y^{2}}{\sqrt{x^{4}+y^{2}+1}}, we have

    γx(x0,y0)\displaystyle\frac{\partial\gamma}{\partial x}\left(x_{0},y_{0}\right) =\displaystyle= 2x022x02y02x03y0223x06(x04+y02+1)32,\displaystyle\frac{-2x_{0}^{2}-2x_{0}^{2}y_{0}^{2}-x_{0}^{3}y_{0}^{2}-\frac{2}{3}x_{0}^{6}}{\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{3}{2}}},
    γy(x0,y0)\displaystyle\frac{\partial\gamma}{\partial y}\left(x_{0},y_{0}\right) =\displaystyle= y0+12y03+23x03y0+x04y0(x04+y02+1)32.\displaystyle\frac{y_{0}+\frac{1}{2}y_{0}^{3}+\frac{2}{3}x_{0}^{3}y_{0}+x_{0}^{4}y_{0}}{\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{3}{2}}}.

    Thus, we have the following desired identity at (x0,y0)(x_{0},y_{0}).

    dγ~\displaystyle d\widetilde{\gamma}
    =\displaystyle= γ~x(x0,y0)dx+γ~y(x0,y0)dy\displaystyle\frac{\partial\widetilde{\gamma}}{\partial x}(x_{0},y_{0})dx+\frac{\partial\widetilde{\gamma}}{\partial y}(x_{0},y_{0})dy
    =\displaystyle= 2x022x02y02x03y0223x06(x04+y02+1)32dx+y0+12y03+23x03y0+x04y0(x04+y02+1)32dy\displaystyle\frac{-2x_{0}^{2}-2x_{0}^{2}y_{0}^{2}-x_{0}^{3}y_{0}^{2}-\frac{2}{3}x_{0}^{6}}{\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{3}{2}}}dx+\frac{y_{0}+\frac{1}{2}y_{0}^{3}+\frac{2}{3}x_{0}^{3}y_{0}+x_{0}^{4}y_{0}}{\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{3}{2}}}dy
    =\displaystyle= (x03y02+(13x0312y02)(x04+y02)(x04+y02)12(x04+y02+1)122x03(x04+y02)12(x04+y02+1)+(x0y0+x02y0)(x04+y02)12(2x0y02x0y032x05y0)(x04+y02)12(x04+y02+1)32)dx\displaystyle\left(\frac{x_{0}^{3}-y_{0}^{2}+\left(\frac{1}{3}x_{0}^{3}-\frac{1}{2}y_{0}^{2}\right)\left(x_{0}^{4}+y_{0}^{2}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}\frac{-2x_{0}^{3}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)}+\frac{\left(x_{0}y_{0}+x_{0}^{2}y_{0}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}}\frac{\left(-2x_{0}y_{0}-2x_{0}y_{0}^{3}-2x_{0}^{5}y_{0}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{3}{2}}}\right)dx
    +(x03y02+(13x0312y02)(x04+y02)(x04+y02)12(x04+y02+1)12y0(x04+y02)12(x04+y02+1)+(x0y0+x02y0)(x04+y02)12x02(x04+y02)12(x04+y02+1)12)dy\displaystyle\quad+\left(\frac{x_{0}^{3}-y_{0}^{2}+\left(\frac{1}{3}x_{0}^{3}-\frac{1}{2}y_{0}^{2}\right)\left(x_{0}^{4}+y_{0}^{2}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}\frac{-y_{0}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)}+\frac{\left(x_{0}y_{0}+x_{0}^{2}y_{0}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}}\frac{x_{0}^{2}}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}\right)dy
    =\displaystyle= ((φ~(x0,y0)𝐯1(x0,y0))(Θ1ν~)x(x0,y0)+(φ~(x0,y0)𝐯2(x0,y0))(Θ2ν~)x(x0,y0))dx\displaystyle\left(\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0})+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial x}(x_{0},y_{0})\right)dx
    +((φ~(x0,y0)𝐯1(x0,y0))(Θ1ν~)y(x0,y0)+(φ~(x0,y0)𝐯2(x0,y0))(Θ2ν~)y(x0,y0))dy\displaystyle\quad+\left(\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{1}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0})+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)\frac{\partial\left(\Theta_{2}\circ\widetilde{\nu}\right)}{\partial y}(x_{0},y_{0})\right)dy
    =\displaystyle= (φ~(x0,y0)𝐯1(x0,y0))d(Θ1ν~)+(φ~(x0,y0)𝐯2(x0,y0))d(Θ2ν~).\displaystyle\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)d\left(\Theta_{1}\circ\widetilde{\nu}\right)+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)d\left(\Theta_{2}\circ\widetilde{\nu}\right).

    Hence, by Theorem 1 (1) and Theorem 2, the plane family (φ~|U,ν~|U)\mathcal{H}_{\left(\widetilde{\varphi}|_{U},\widetilde{\nu}|_{U}\right)} for the shoe surface φ~(x,y)=(x,y,13x312y2)\widetilde{\varphi}(x,y)=\left(x,y,\frac{1}{3}x^{3}-\frac{1}{2}y^{2}\right) has a unique envelope f~:U3\widetilde{f}:{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}U}\to\mathbb{R}^{3}, where U=2{(0,0)}U=\mathbb{R}^{2}-\{(0,0)\}. Then, under the canonical identifications

    Tν~(x0,y0)S2Tν~(x0,y0)S2Tν~(x0,y0)33,T^{*}_{\widetilde{\nu}(x_{0},y_{0})}S^{2}\cong T_{\widetilde{\nu}(x_{0},y_{0})}S^{2}\subset T_{\widetilde{\nu}(x_{0},y_{0})}\mathbb{R}^{3}\cong\mathbb{R}^{3},

    the 22-dimensional cotangent vector

    ω~(x0,y0)=(φ~(x0,y0)𝐯1(x0,y0))dΘ1+(φ~(x0,y0)𝐯2(x0,y0))dΘ2\widetilde{\omega}(x_{0},y_{0})=\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right)d\Theta_{1}+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right)d\Theta_{2}

    is identified with the following 33-dimensional vector (denoted by the same symbol ω~(x0,y0)\widetilde{\omega}(x_{0},y_{0})).

    ω~(x0,y0)\displaystyle\widetilde{\omega}(x_{0},y_{0}) =\displaystyle= (φ~(x0,y0)𝐯1(x0,y0))𝐯1(x0,y0)+(φ~(x0,y0)𝐯2(x0,y0))𝐯2(x0,y0)\displaystyle\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{1}(x_{0},y_{0})\right){\bf v}_{1}(x_{0},y_{0})+\left(\widetilde{\varphi}(x_{0},y_{0})\cdot{\bf v}_{2}(x_{0},y_{0})\right){\bf v}_{2}(x_{0},y_{0})
    =\displaystyle= (x03y02+(13x0312y02)(x04+y02))(x04+y02)12(x04+y02+1)12(x02,y0,x04+y02)(x04+y02)12(x04+y02+1)12\displaystyle\frac{\left(x_{0}^{3}-y_{0}^{2}+\left(\frac{1}{3}x_{0}^{3}-\frac{1}{2}y_{0}^{2}\right)\left(x_{0}^{4}+y_{0}^{2}\right)\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}\frac{\left(x_{0}^{2},\,-y_{0},\,x_{0}^{4}+y_{0}^{2}\right)}{{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}}
    +(x0y0+x02y0)(x04+y02)12(y0,x02, 0)(x04+y02)12.\displaystyle\qquad+\frac{\left(x_{0}y_{0}+x_{0}^{2}y_{0}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}}\frac{\left(y_{0},\,x_{0}^{2},\,0\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}}.

    Therefore, by Theorem 1 (2), the unique envelope f~\widetilde{f} must have the following desired parametric representation on U=2{(0,0)}U=\mathbb{R}^{2}-\{(0,0)\}.

    f~(x0,y0)\displaystyle\widetilde{f}(x_{0},y_{0})
    =\displaystyle= ω~(x0,y0)+γ~(x0,y0)ν~(x0,y0)\displaystyle\widetilde{\omega}(x_{0},y_{0})+\widetilde{\gamma}(x_{0},y_{0})\widetilde{\nu}(x_{0},y_{0})
    =\displaystyle= (x03y02+(13x0312y02)(x04+y02))(x04+y02)12(x04+y02+1)12(x02,y0,x04+y02)(x04+y02)12(x04+y02+1)12\displaystyle\frac{\left(x_{0}^{3}-y_{0}^{2}+\left(\frac{1}{3}x_{0}^{3}-\frac{1}{2}y_{0}^{2}\right)\left(x_{0}^{4}+y_{0}^{2}\right)\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}\frac{\left(x_{0}^{2},\,-y_{0},\,x_{0}^{4}+y_{0}^{2}\right)}{{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}}
    +(x0y0+x02y0)(x04+y02)12(y0,x02, 0)(x04+y02)12+(23x03+12y02)(x04+y02+1)12(x02,y0, 1)(x04+y02+1)12\displaystyle\qquad+\frac{\left(x_{0}y_{0}+x_{0}^{2}y_{0}\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}}\frac{\left(y_{0},\,x_{0}^{2},\,0\right)}{\left(x_{0}^{4}+y_{0}^{2}\right)^{\frac{1}{2}}}+\frac{\left(-\frac{2}{3}x_{0}^{3}+\frac{1}{2}y_{0}^{2}\right)}{\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}\frac{\left(-x_{0}^{2},\,y_{0},\,1\right)}{\left(x_{0}^{4}+y_{0}^{2}+1\right)^{\frac{1}{2}}}
    =\displaystyle= (x0,y0,13x0312y02)\displaystyle\left(x_{0},\,y_{0},\,\frac{1}{3}x_{0}^{3}-\frac{1}{2}y_{0}^{2}\right)
    =\displaystyle= φ~(x0,y0).\displaystyle\widetilde{\varphi}(x_{0},y_{0}).

    By continuity, it follows that the given shoe surface φ~\widetilde{\varphi} itself is the unique envelope created by the tangent plane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}.

    The set called the parabolic line of φ~:23\widetilde{\varphi}:\mathbb{R}^{2}\to\mathbb{R}^{3} consists of points (x,y)2(x,y)\in\mathbb{R}^{2} at which ν~\widetilde{\nu} is singular. For the shoe surface, the parabolic line is the yy-axis {(0,y)|y}\{(0,y)\,|\,y\in\mathbb{R}\}. Thus, as similar as the case of unit speed plane curves 𝐫:2{\bf r}:\mathbb{R}\to\mathbb{R}^{2} with inflection points, the full discriminant of the tangent plane family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} for the shoe surface φ~:23\widetilde{\varphi}:\mathbb{R}^{2}\to\mathbb{R}^{3} is different from the unique desired envelope φ~\widetilde{\varphi} itself, since the full discriminant includes an affine tangent line {(λ,y,12y2)|λ}\left\{\left.\left(\lambda,y,-\frac{1}{2}y^{2}\right)\;\right|\;\lambda\in\mathbb{R}\right\} at any point (0,y)(0,y). Therefore, even in the case of surfaces in 3\mathbb{R}^{3}, by our method, one can distinguish the envelope in the sense of Definition 1 and the full discriminant. This means that, in the case of surfaces in 3\mathbb{R}^{3} as well, our method has an advantage.

  1. (2)

    (Example 4.1 of [14]) Let ν~:nSnn+1\widetilde{\nu}:\mathbb{R}^{n}\to S^{n}\subset\mathbb{R}^{n+1} be the mapping defined by ν~(p1,,pn)=1i=1npi2+1(p1,,pn,1)\widetilde{\nu}\left(p_{1},\ldots,p_{n}\right)=\frac{1}{\sqrt{\sum_{i=1}^{n}p_{i}^{2}+1}}\left(p_{1},\ldots,p_{n},-1\right). Then, ν~\widetilde{\nu} is non-singular and its inverse mapping ν~1:ν~(n+1)n+1\widetilde{\nu}^{-1}:\widetilde{\nu}\left(\mathbb{R}^{n+1}\right)\to\mathbb{R}^{n+1} is the central projection relative to the south pole (0,,0,1)(0,\ldots,0,-1) of SnS^{n}. Let φ~:nn+1\widetilde{\varphi}:\mathbb{R}^{n}\to\mathbb{R}^{n+1} be an arbitrary mapping. Set γ~(p)=φ~(p)ν~(p)\widetilde{\gamma}(p)=\widetilde{\varphi}(p)\cdot\widetilde{\nu}(p) where p=(p1,,pn)p=\left(p_{1},\ldots,p_{n}\right) be a point of n+1\mathbb{R}^{n+1}. Let (X=(X1,,Xn),Y)\left(X=\left(X_{1},\ldots,X_{n}\right),Y\right) be a point of n×\mathbb{R}^{n}\times\mathbb{R}. Since J1(n,)J^{1}(\mathbb{R}^{n},\mathbb{R}) and n××n\mathbb{R}^{n}\times\mathbb{R}\times\mathbb{R}^{n} are identified, (X,Y,p)(X,Y,p) may be regarded as the canonical coordinate system of J1(n,)J^{1}\left(\mathbb{R}^{n},\mathbb{R}\right). Since Xiν~(p)Yν~(p)=pi\frac{X_{i}\circ\widetilde{\nu}(p)}{Y\circ\widetilde{\nu}(p)}=-p_{i} for any ii (1in)(1\leq i\leq n) and any pn+1p\in\mathbb{R}^{n+1}, considering the first order differential equation

    ((X,Y)φ~(p))ν~(p)=0\left(\left(X,Y\right)-\widetilde{\varphi}(p)\right)\cdot\widetilde{\nu}(p)=0

    is exactly the same as considering the following Clairaut equation

    Y=i=1nXipi+φ~(p)ν~(p)Yν~(p).Y=\sum_{i=1}^{n}X_{i}p_{i}+\frac{\widetilde{\varphi}(p)\cdot\widetilde{\nu}(p)}{Y\circ\widetilde{\nu}(p)}.

    Thus, for each xnx\in\mathbb{R}^{n} the hyperplane H(φ~(x),ν~(x))H_{\left(\widetilde{\varphi}(x),\widetilde{\nu}(x)\right)} is a complete solution of the above Clairaut equation. Since ν~\widetilde{\nu} is non-singular, by Theorem 1 and Theorem 2, the above Clairaut equation has a unique singular solution f~:nn+1\widetilde{f}:\mathbb{R}^{n}\to\mathbb{R}^{n+1}. By Theorem 1 again, the unique singular solution f~\widetilde{f} has the following expression where xx is an arbitrary point of n\mathbb{R}^{n} and (V,(Θ1,,Θn))\left(V,\left(\Theta_{1},\ldots,\Theta_{n}\right)\right) is a sufficiently small normal coordinate neighborhood of SnS^{n} at ν~(x)\widetilde{\nu}(x).

    f~(x)=i=1(γ~ν~1)Θi(ν~(x))Θi+γ~(x)ν~(x).\widetilde{f}(x)=\sum_{i=1}\frac{\partial\left(\widetilde{\gamma}\circ\widetilde{\nu}^{-1}\right)}{\partial\Theta_{i}}\left(\widetilde{\nu}(x)\right)\frac{\partial}{\partial\Theta_{i}}+\widetilde{\gamma}(x)\widetilde{\nu}(x).

    By this expression, for instance, it is easily seen that when γ~(x)c(0)\widetilde{\gamma}(x)\equiv c(\neq 0) for any xn+1x\in\mathbb{R}^{n+1}, then the unique singular solution Y:UcY:U_{c}\to\mathbb{R} must be an explicit solution with the following expression where Uc={X|X<|c|}U_{c}=\{X\;|\;\|X\|<|c|\}.

    Y(X)={|c|2i=1nXi2( if c>0)|c|2i=1nXi2( if c<0).Y(X)=\left\{\begin{array}[]{rr}-\sqrt{|c|^{2}-\sum_{i=1}^{n}X_{i}^{2}}&(\mbox{ if }c>0)\\ \sqrt{|c|^{2}-\sum_{i=1}^{n}X_{i}^{2}}&(\mbox{ if }c<0).\end{array}\right.

5. Appendix: Alternative proof of Theorem 1 except for the assertion (3)
in the case n=1n=1

Let NN be a 11-dimensional manfold and let φ~:N2\widetilde{\varphi}:N\to\mathbb{R}^{2}, ν~:NS1\widetilde{\nu}:N\to S^{1} be mappings. Define the function Θ~:N\widetilde{\Theta}:N\to\mathbb{R} by ν~(t)=(cosΘ~(t),sinΘ~(t))\widetilde{\nu}(t)=\left(\cos\widetilde{\Theta}(t),\sin\widetilde{\Theta}(t)\right). Define also τ~(t):=(sinΘ~(t),cosΘ~(t))\widetilde{\tau}(t):=\left(\sin\widetilde{\Theta}(t),-\cos\widetilde{\Theta}(t)\right). Then, the following trivially holds.

Fact 5.1.

For any h:N2h:N\to\mathbb{R}^{2},

h(t)=(h(t)τ~(t))τ~(t)+(h(t)ν~(t))ν~(t).h(t)=\left(h(t)\cdot\widetilde{\tau}(t)\right)\widetilde{\tau}(t)+\left(h(t)\cdot\widetilde{\nu}(t)\right)\widetilde{\nu}(t).

We first show that the creative condition can be naturally obtained from an envelope by introducing a gauge theoretic approach. Suppose that f~:N2\widetilde{f}:N\to\mathbb{R}^{2} is an envelope created by the line family (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}. Then, we have the following.

γ~(t)=(f~(t)ν~(t))=f~(t)ν~(t)+f~(t)ν~(t)=0(f~(t)τ~(t))Θ~(t).\widetilde{\gamma}^{\prime}(t)=\left(\widetilde{f}(t)\cdot\widetilde{\nu}(t)\right)^{\prime}=\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)+\widetilde{f}(t)\cdot\widetilde{\nu}^{\prime}(t)=0-\left(\widetilde{f}(t)\cdot\widetilde{\tau}(t)\right)\widetilde{\Theta}^{\prime}(t).

Let h:NNh:N\to N be a bijective mapping. Then, notice that

(φ~,ν~)=(φ~h,ν~h)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)}=\mathcal{H}_{\left(\widetilde{\varphi}\circ h,\widetilde{\nu}\circ h\right)}

and

(γ~h)(t)=(f~(h(t))τ~(h(t)))Θ~(h(t))h(t).\left(\widetilde{\gamma}\circ h\right)^{\prime}(t)=-\left(\widetilde{f}(h(t))\cdot\widetilde{\tau}(h(t))\right)\widetilde{\Theta}^{\prime}(h(t))h^{\prime}(t).

From these simple observations, we see that it is important to extract a significant quantity which does not depend on the particular choice of hh. Then, we naturally reach the following setting.

ω~(t):=(f~(t)τ~(t))dΘ~.\widetilde{\omega}(t):=-\left(\widetilde{f}(t)\cdot\widetilde{\tau}(t)\right)d\widetilde{\Theta}.

and we trivially have dγ~=ω~d\widetilde{\gamma}=\widetilde{\omega}. Take an arbitrary point t0t_{0} of NN and fix it. Let (V,Θ)(V,\Theta) be a normal coordinate neighborhood of S1S^{1} at ν~(t0)\widetilde{\nu}(t_{0}) such that Θ(ν~(t0))=0\Theta\left(\widetilde{\nu}(t_{0})\right)=0 and Θ~(t)=(Θν~)(t)\widetilde{\Theta}(t)=\left(\Theta\circ\widetilde{\nu}\right)(t) for any tν~1(V)t\in\widetilde{\nu}^{-1}(V). In other words, (Θν~)(t)\left(\Theta\circ\widetilde{\nu}\right)(t) (tν~1(V))\left(t\in\widetilde{\nu}^{-1}(V)\right) is just the radian (or its negative) between two unit vectors ν~(t0)\widetilde{\nu}(t_{0}) and ν~(t)\widetilde{\nu}(t). By using the function Θ:V\Theta:V\to\mathbb{R}, the 11-form ω~(t)\widetilde{\omega}(t) may be written as follows.

ω~(t)=(f~(t)τ~(t))ν~dΘ,\widetilde{\omega}(t)=-\left(\widetilde{f}(t)\cdot\widetilde{\tau}(t)\right)\widetilde{\nu}^{*}d\Theta,

where ν~dΘ\widetilde{\nu}^{*}d{\Theta} stands for the pullback of the 11-form dΘd\Theta by ν~\widetilde{\nu}. Hence, we naturally reach the following 11-form which is denoted by the same symbol ω~\widetilde{\omega}.

ω~(t)=(f~(t)τ~(t))dΘ.\widetilde{\omega}(t)=-\left(\widetilde{f}(t)\cdot\widetilde{\tau}(t)\right)d\Theta.

It is easily seen that for any tν~1(V)t\in\widetilde{\nu}^{-1}(V), under the canonical identifications

Tν~(t)S1Tν~(t)S1Tν~(t)22,T^{*}_{\widetilde{\nu}(t)}S^{1}\cong T_{\widetilde{\nu}(t)}S^{1}\subset T_{\widetilde{\nu}(t)}\mathbb{R}^{2}\cong\mathbb{R}^{2},

the 11-dimensional cotangent vector

ω~(t)=(f~(t)τ~(t))dΘTν~(t)S1\widetilde{\omega}(t)=-\left(\widetilde{f}(t)\cdot\widetilde{\tau}(t)\right)d\Theta\in T^{*}_{\widetilde{\nu}(t)}S^{1}

is identified with the 22-dimensional vector

ω~(t)=(f~(t)τ~(t))τ~(t)2.\widetilde{\omega}(t)=\left(\widetilde{f}(t)\cdot\widetilde{\tau}(t)\right)\widetilde{\tau}(t)\in\mathbb{R}^{2}.

Since t0t_{0} is an arbitrary point of NN, we naturally see that the creative condition is satisfied for (φ,ν~~)\mathcal{H}_{\left(\widetilde{\varphi,\widetilde{\nu}}\right)} and the following horizontal-vertical decomposition formula holds for any tNt\in N.

Fact 5.2.
f~(t)=(f~(t)τ~(t))τ~(t)+(f~(t)ν~(t))ν~(t)=ω~(t)+γ~(t)ν~(t).\widetilde{f}(t)=\left(\widetilde{f}(t)\cdot\widetilde{\tau}(t)\right)\widetilde{\tau}(t)+\left(\widetilde{f}(t)\cdot\widetilde{\nu}(t)\right)\widetilde{\nu}(t)=\widetilde{\omega}(t)+\widetilde{\gamma}(t)\widetilde{\nu}(t).

Conversely, suppose that (φ~,ν~)\mathcal{H}_{\left(\widetilde{\varphi},\widetilde{\nu}\right)} is creative. Then, there exists a function α:N\alpha:N\to\mathbb{R} such that dγ~=αdΘ~d\widetilde{\gamma}=\alpha d\widetilde{\Theta}. Set ω~=αdΘ~\widetilde{\omega}=\alpha d\widetilde{\Theta}. Let t0Nt_{0}\in N be an arbitrary point. Then, under the canonical identifications

Tν~(t)S1Tν~(t)S1Tν~(t)22,T^{*}_{\widetilde{\nu}(t)}S^{1}\cong T_{\widetilde{\nu}(t)}S^{1}\subset T_{\widetilde{\nu}(t)}\mathbb{R}^{2}\cong\mathbb{R}^{2},

the 11-dimensional cotangent vector

ω~(t)=α(t)dΘTν~(t)S1\widetilde{\omega}(t)=\alpha(t)d\Theta\in T^{*}_{\widetilde{\nu}(t)}S^{1}

is identified with the 22-dimensional vector

ω~(t)=α(t)τ~(t)2,\widetilde{\omega}(t)=-\alpha(t)\widetilde{\tau}(t)\in\mathbb{R}^{2},

where (V,Θ)(V,\Theta) is a normal coordinate system of S1S^{1} at ν~(t0)\widetilde{\nu}(t_{0}) such that Θ(ν~(t0))=0\Theta(\widetilde{\nu}(t_{0}))=0 and tν~1(V)t\in\widetilde{\nu}^{-1}(V). Set

f~(t)=ω~(t)+γ~(t)ν~(t)=α(t)τ~(t)+γ~(t)ν~(t).\widetilde{f}(t)=\widetilde{\omega}(t)+\widetilde{\gamma}(t)\widetilde{\nu}(t)=-\alpha(t)\widetilde{\tau}(t)+\widetilde{\gamma}(t)\widetilde{\nu}(t).

Then, f~\widetilde{f} clearly satisfies the condition (b) of Definition 1 for any tν~1(V)t\in\widetilde{\nu}^{-1}(V). Moreover we have the following.

Lemma 5.1.

For any tν~1(V)t\in\widetilde{\nu}^{-1}(V), f~(t)ν~(t)=0\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)=0 holds.

Proof of Lemma 5.1 We have

γ~(t)=(f~(t)ν~(t))=f~(t)ν~(t)(f~(t)τ~(t))Θ~(t)=f~(t)ν~(t)+α(t)Θ~(t).\widetilde{\gamma}^{\prime}(t)=\left(\widetilde{f}(t)\cdot\widetilde{\nu}(t)\right)^{\prime}=\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)-\left(\widetilde{f}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}(t)}\cdot\widetilde{\tau}(t)\right)\widetilde{\Theta}^{\prime}(t)=\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)+\alpha(t)\widetilde{\Theta}^{\prime}(t).

Thus, we have the following.

ω~(t)=dγ~=γ~(t)dt\displaystyle\widetilde{\omega}(t)=d\widetilde{\gamma}=\widetilde{\gamma}^{\prime}(t)dt =\displaystyle= (f~(t)ν~(t))dt+α(t)Θ~(t)dt\displaystyle\left(\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)\right)dt+\alpha(t)\widetilde{\Theta}^{\prime}(t)dt
=\displaystyle= (f~(t)ν~(t))dt+α(t)dΘ~\displaystyle\left(\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)\right)dt+\alpha(t)d\widetilde{\Theta}
=\displaystyle= (f~(t)ν~(t))dt+ω~(t).\displaystyle\left(\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)\right)dt+\widetilde{\omega}(t).

It follows (f~(t)ν~(t))dt=0\left(\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)\right)dt=0. Since tt is a coordinate function on an open set ν~1(V)\widetilde{\nu}^{-1}(V) of NN, for any fixed tν~1(V)t\in\widetilde{\nu}^{-1}(V), the 11-dimensional cotangent vector dtdt at tt is not zero. Therefore, the number (f~(t)ν~(t))\left(\widetilde{f}^{\prime}(t)\cdot\widetilde{\nu}(t)\right) is always zero for any tν~1(V)t\in\widetilde{\nu}^{-1}(V). Since t0t_{0} is an arbitrary point of NN, Theorem 1 (1) holds. By the above decomposition of f~(t)\widetilde{f}(t), Theorem 1 (2) holds as well. ∎

Acknowledgements

The author is most grateful to two anonymous reviewers. Their comments/suggestions are hard to replace. The author would like to thank Richard Montgomery for appropriate comments. His suggestions improved this paper.

References

  • [1] V. I. Arnol’d, Singularities of Caustics and Wavefronts, Mathematics and its Applications, 62, Springer Netherland, Dordrecht, 1990. https://doi.org/10.1007/978-94-011-3330-2
  • [2] V. I. Arnol’d, S. M. Gusein-Zade, and A. N. Varchenko, Singularities of Differentiable Maps I, Monographs in Mathematics 82, Birkhäuser, Boston Basel Stuttgart, 1985. https://doi.org/10.1007/978-1-4612-3940-6
  • [3] T. Banchoff, T. Gaffney and C. MacCrory, Cusps of Gauss Mappings, Pitman Advanced Pub. Progman, 1982.
  • [4] P. Blaschke, Pedal coordinates, dark Kepler and other force problems, J. Math. Phys., 58 (2017), 063505. https://doi.org/10.1063/1.4984905
  • [5] T. Bröcker, Differentiable Germs and Catastrophes, Cambridge University Press, Cambridge, 1975. https://doi.org/10.1017/CBO9781107325418
  • [6] J. W. Bruce and P. J. Giblin, Curves and Singularities (second edition), Cambridge University Press, Cambridge, 1992. https://doi.org/10.1017/CBO9781139172615
  • [7] R. J. Fisher and H. Turner Laquer, Hyperplane envelopes and the Clairaut equation, J. Geom. Anal., 20 (2010), 609–650. https://doi.org/10.1007/s12220-010-9129-0
  • [8] Y. Giga, Surface Evolution Equations, Monographs of Mathematics, 99, Springer, 2006.
    https://doi.org/10.1007/3-7643-7391-1
  • [9] D. Goodstein and J. R. Goodstein, Feynman’s Lost Lecture: the Motion of Planets Around the Sun (first edition), W.W. Norton & Company, New York, 1996.
  • [10] H. Han and T. Nishimura, Spherical method for studying Wulff shapes and related topics, Adv. Stud. Pure Math., 78, 1–53, Math. Soc. Japan. Tokyo, 2018. https://doi.org/10.2969/aspm/07810001
  • [11] D. W. Hoffman and J. W. Cahn, A vector thermodynamics for anisotropic surfaces, Surface Science, 31 (1972), 368–388. https://doi.org/10.1016/0039-6028(72)90268-3
  • [12] G. Ishikawa, Opening of differentiable map-germs and unfoldings, Topics on real and complex singularities, 87-–113, World Sci. Publ., Hackensack, NJ, 2014. https://doi.org/10.1142/9789814596046_0007
  • [13] G. Ishikawa, Singularities of frontals, Adv. Stud. Pure Math., 78, 55–106, Math. Soc. Japan, Tokyo, 2018. https://doi.org/10.2969/aspm/07810055
  • [14] S. Izumiya, Singular solutions of first order differential equations, Bull. London Math. Soc., 26 (1994), 69–74. https://doi.org/10.1112/blms/26.1.69
  • [15] S. Janeczko and T. Nishimura, Anti-orthotomics of frontals and their applications, J. Math. Anal. Appl., 487 (2020), 124019. https://doi.org/10.1016/j.jmaa.2020.124019
  • [16] T. Nishimura, Kato’s chaos created by quadratic mappings associated with spherical orthotomic curves, J. Singul., 21 (2020), 205–211. https://doi.org/10.5427/jsing.2020.21l.
  • [17] K. Saji, M. Umehara and K. Yamada, The geometry of fronts, Ann. Math., 169-2 (2009), 491–529.
    https://doi.org/10.4007/annals.2009.169.491
  • [18] M. Takahashi, Envelopes of families of Legendre mappings in the unit tangent bundle over the Euclidean space, J. Math. Anal. Appl., 473 (2019), 408–420. https://doi.org/10.1016/j.jmaa.2018.12.057.
  • [19] J. E. Taylor, Crystalline variational problems, Bull. Amer. Math. Soc., 84 (1978), 568–588.
  • [20] G. Wulff, Zur frage der geschwindindigkeit des wachstrums und der auflösung der krystallflachen, Z. Kristallographine und Mineralogie, 34 (1901), 449–530.