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

Some results about equichordal convex bodies

Jesús Jerónimo-Castro1, Francisco G. Jimenez-Lopez2,
and Efrén Morales-Amaya3
1,2Facultad de Ingeniería
Universidad Autónoma de Querétaro, México
3Facultad de Matemáticas-Acapulco,
Universidad Autónoma de Guerrero, México
1[email protected],
2[email protected],
3[email protected]
Abstract

Let KK and LL be two convex bodies in n\mathbb{R}^{n}, n2n\geq 2, with LintKL\subset\text{int}\,K. We say that LL is an equichordal body for KK if every chord of KK tangent to LL has length equal to a given fixed value λ\lambda. In [1], J. Barker and D. Larman proved that if LL is a ball, then KK is a ball concentric with LL. In this paper we prove, derived from the proof of Theorem 1, that there exist an infinite number of closed curves, different from circles, which possess an equichordal convex body. If the dimension of the space is more than or equal to 3, then only Euclidean balls possess an equichordal convex body. We also prove some results about isoptic curves and give relations between isoptic curves and convex rotors in the plane.

1 Introduction

Let KK be a convex body in the plane, i.e., a compact and convex set with non-empty interior, and let 𝒫\mathcal{P} be a convex polygon. It is said that KK is a rotor in 𝒫\mathcal{P} if for every rotation ρ\rho, there is a translate of 𝒫\mathcal{P} that contains to ρ(K)\rho(K) and all sides of 𝒫\mathcal{P} are tangent to KK. There are many results about rotors in regular polygons, see for instance [6], and for the particular case of rotors in equilateral triangles see [21]. The case of rotors in squares is well-known, indeed, bodies of constant width are a very important topic in Convex Geometry and have many interesting properties and applications in mechanisms (see the quite nice book [13]).

Another topic, apparently not related to rotors, is the Equichordal Problem. Let xx be a point in the interior of a convex body KK, we say that xx is an equichordal point if every chord of KK through xx have the same length. The famous Equichordal Problem, due to W. Blaschke, W. Rothe, and R. Weitzenböck [2], asks about the existence of a convex body with two equichordal points. There are many false proofs about the non existence of such a body, however, M. Rychlik finally gave a complete proof about the non existence of a body with two equichordal points in [16]. It is worth mentioning that there are many convex bodies, different from the disc, which have exactly one equichordal point. Here we are interested in a generalization of the notion of equichordal point in the following way: Let KK and LL be two convex bodies in n\mathbb{R}^{n}, n2n\geq 2, with LintKL\subset\text{int}\,K. We say that LL is an equichordal body for KK if every chord of KK tangent to LL have length equal to a given fixed value λ\lambda. In [1], J. Barker and D. Larman proved that if KK is a convex body that possesses an equichordal ball then it is also a ball. However, we wonder if there exist convex bodies different from balls which possess an equichordal convex body in its interior. It seems that bodies which float in equilibrium in every position provide examples of such bodies in the plane (see for instance [20]), however, it is not clear if the considered bodies KK are convex or not.

In sections 3 and 4 of this paper we study convex bodies LL for which the chords of one of its isoptic curves (defined in the following section), that are tangent to LL have length equal to a constant number λ\lambda. Moreover, these bodies LL are examples of rotors in regular polygons and if we fix the convex body LL and the circumscribed polygon is rotated, while maintained circumscribed to KK, its vertices describe the isoptic curve of LL. In section 6 we also prove that in dimension 33 or higher, only Euclidean balls (or simply balls) possess and equichordal convex body in its interior.

2 Preliminary concepts

We give first some definitions and notation. Let KK be a given planar convex body; for every real number tt we denote by (t)\ell(t) the support line of KK with outward normal vector u(t)=(cost,sint)u(t)=(\cos t,\sin t). The function p:p:\mathbb{R}\longrightarrow\mathbb{R}, defined as p(t)=maxxKu(t),xp(t)=\max_{x\in K}\langle u(t),x\rangle, is known by the name of support function of KK. When the origin OO is contained in KK, p(t)p(t) is nothing else than the distance from OO to the support line (t).\ell(t). The distance between the support lines (t)\ell(t) and (t+π)\ell(t+\pi) is called the width of KK in direction u(t)u(t) and it is denoted by w(t)w(t), in other words, w(t)=p(t)+p(t+π)w(t)=p(t)+p(t+\pi). If w(t)w(t) is constant, independently of tt, we say that KK is a body of constant width. For any α(0,π)\alpha\in(0,\pi), the α\alpha-isoptic KαK_{\alpha} of KK is defined as the locus of points at which two tangent lines to KK intersect at an angle α\alpha. Using the support function, K\partial K is parameterized (see for instance [19]) by

γ(t)=p(t)u(t)+p(t)u(t),fort[0,2π].{}\gamma(t)=p(t)u(t)+p^{\prime}(t)u^{\prime}(t),\ \text{for}\ t\in[0,2\pi]. (1)

The isoptic curve KαK_{\alpha} can be parameterized by the same angle by the formula (see [4] or [10])

γα(t)=p(t)u(t)+[p(t)cotα+1sinαp(t+πα)]u(t).{}\gamma_{\alpha}(t)=p(t)u(t)+\left[p(t)\cot\alpha+\frac{1}{\sin\alpha}p(t+\pi-\alpha)\right]u^{\prime}(t). (2)

By Cauchy’s formula, the perimeter of KK can be obtained by (see [19])

L(K)=02πp(t)𝑑t.L(K)=\int_{0}^{2\pi}p(t)dt. (3)

For any tt\in\mathbb{R} we define (see Fig. 1)

a(t)\displaystyle a(t) =|γα(t)γ(t)|,\displaystyle=|\gamma_{\alpha}(t)-\gamma(t)|, (4)
b(t)\displaystyle b(t) =|γα(t)γ(t+πα)|,\displaystyle=|\gamma_{\alpha}(t)-\gamma(t+\pi-\alpha)|, (5)
c(t)\displaystyle c(t) =|γα(t)γα(t+πα)|=b(t)+a(t+πα),\displaystyle=|\gamma_{\alpha}(t)-\gamma_{\alpha}(t+\pi-\alpha)|=b(t)+a(t+\pi-\alpha), (6)
q(t)\displaystyle q(t) =|γ(t)γ(t+πα)|,\displaystyle=|\gamma(t)-\gamma(t+\pi-\alpha)|, (7)
λ(t)\displaystyle\lambda(t) =|γα(t+πα)γα(tπ+α)|.\displaystyle=|\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t-\pi+\alpha)|. (8)

We also define d(t)d(t) to be the distance between the points obtained by projecting the origin OO onto the support lines of KK at γ(t)\gamma(t) and γ(t+πα)\gamma(t+\pi-\alpha).

Refer to caption
Figure 1: Parameters of the isoptic curve

By some tedious but simple calculations we can express the lengths of all these chords in terms of the support function of KK:

a(t)\displaystyle a(t) =1sinα[p(t+πα)+p(t)cosαp(t)sinα],\displaystyle=\frac{1}{\sin\alpha}[p(t+\pi-\alpha)+p(t)\cos\alpha-p^{\prime}(t)\sin\alpha], (9)
b(t)\displaystyle b(t) =1sinα[p(t+πα)cosα+p(t+πα)sinα+p(t)],\displaystyle=\frac{1}{\sin\alpha}[p(t+\pi-\alpha)\cos\alpha+p^{\prime}(t+\pi-\alpha)\sin\alpha+p(t)], (10)
c(t)\displaystyle c(t) =1sinα[2p(t+πα)cosα+p(t)+p(t2α)],\displaystyle=\frac{1}{\sin\alpha}[2p(t+\pi-\alpha)\cos\alpha+p(t)+p(t-2\alpha)], (11)
d(t)\displaystyle d(t) =p2(t)+p2(t+πα)+2p(t)p(t+πα)cosα.\displaystyle=\sqrt{p^{2}(t)+p^{2}(t+\pi-\alpha)+2p(t)p(t+\pi-\alpha)\cos\alpha}. (12)

Finally, the support function of a convex body is a periodic function, with period 2π2\pi, and it is also absolutely continuous, so we can consider its expansion in terms of the Fourier series (see [6]), i.e.,

p(t)=a0+n=1(ancosnt+bnsinnt).p(t)=a_{0}+\sum_{n=1}^{\infty}\left(a_{n}\cos nt+b_{n}\sin nt\right). (13)

The first and second derivatives of pp are expressed as

p(t)\displaystyle p^{\prime}(t) =n=1(nansinntnbncosnt),\displaystyle=-\sum_{n=1}^{\infty}\left(na_{n}\sin nt-nb_{n}\cos nt\right), (14)
p′′(t)\displaystyle p^{\prime\prime}(t) =n2n=1(ancosnt+bnsinnt).\displaystyle=-n^{2}\sum_{n=1}^{\infty}\left(a_{n}\cos nt+b_{n}\sin nt\right). (15)

3 Some results about isoptic curves in the plane

Our first result about isoptic chords is the following.

Theorem 1.

Let KK be a strictly convex body in the plane with differentiable boundary and let α(0,π)\alpha\in(0,\pi) be a fixed angle such that απ\frac{\alpha}{\pi} is an irrational number. Suppose c(t)=c0c(t)=c_{0}, for every t[0,2π]t\in[0,2\pi], for a positive number c0.c_{0}. Then KK is a disc.

Proof. Since c(t)=c0c(t)=c_{0} we have by (11) that

c(t)=1sinα[2p(t+πα)cosα+p(t)+p(t2α)]=c0,c(t)=\frac{1}{\sin\alpha}[2p(t+\pi-\alpha)\cos\alpha+p(t)+p(t-2\alpha)]=c_{0},

it follows that

2p(t+πα)cosα+p(t)+p(t2α)=0.2p^{\prime}(t+\pi-\alpha)\cos\alpha+p^{\prime}(t)+p^{\prime}(t-2\alpha)=0. (16)

If we substitute the Fourier coefficients of p(t)p(t) in the differential equation (16), by (14) we have

2cosα\displaystyle 2\cos\alpha n=1(nbncosn(t+πα)nansinn(t+πα))\displaystyle\sum_{n=1}^{\infty}(nb_{n}\cos n(t+\pi-\alpha)-na_{n}\sin n(t+\pi-\alpha))
+\displaystyle+ n=1(nbncosntnansinnt)\displaystyle\sum_{n=1}^{\infty}(nb_{n}\cos nt-na_{n}\sin nt)
+\displaystyle+ n=1(nbncosn(t2α)nansinn(t2α))=0.\displaystyle\sum_{n=1}^{\infty}(nb_{n}\cos n(t-2\alpha)-na_{n}\sin n(t-2\alpha))=0.

Since this holds for every real number tt, we must have that for every nn

cosnt[2nansinn(πα)cosα+nansin2nα+2nbncosn(πα)cosα\displaystyle\cos nt[-2na_{n}\sin n(\pi-\alpha)\cos\alpha+na_{n}\sin 2n\alpha+2nb_{n}\cos n(\pi-\alpha)\cos\alpha
+nbn+nbncos2nα]+sinnt[2nancosn(πα)cosαnannancos2nα\displaystyle+nb_{n}+nb_{n}\cos 2n\alpha]+\sin nt[-2na_{n}\cos n(\pi-\alpha)\cos\alpha-na_{n}-na_{n}\cos 2n\alpha
2nbnsinn(πα)cosα+nbnsin2nα]=0.\displaystyle-2nb_{n}\sin n(\pi-\alpha)\cos\alpha+nb_{n}\sin 2n\alpha]=0.

The coefficients of cosnt\cos nt and sinnt\sin nt must be both equal to 0, hence we have that

[2sinn(πα)cosα+sin2nα2cosn(πα)cosα+1+cos2nα2cosn(πα)cosα1cos2nα2sinn(πα)cosα+sin2nα][anbn]=[00].\begin{bmatrix}-2\sin n(\pi-\alpha)\cos\alpha+\sin 2n\alpha&2\cos n(\pi-\alpha)\cos\alpha+1+\cos 2n\alpha\\ -2\cos n(\pi-\alpha)\cos\alpha-1-\cos 2n\alpha&-2\sin n(\pi-\alpha)\cos\alpha+\sin 2n\alpha\end{bmatrix}\begin{bmatrix}a_{n}\\ b_{n}\end{bmatrix}=\begin{bmatrix}0\\ 0\end{bmatrix}.

The determinant of the matrix above is given by

(2sinn(πα)cosα+sin2nα)2+(2cosn(πα)cosα+1+cos2nα)2.(-2\sin n(\pi-\alpha)\cos\alpha+\sin 2n\alpha)^{2}+(2\cos n(\pi-\alpha)\cos\alpha+1+\cos 2n\alpha)^{2}.

This determinant is zero only if

2sinn(πα)cosα+sin2nα=0-2\sin n(\pi-\alpha)\cos\alpha+\sin 2n\alpha=0 (17)

and

2cosn(πα)cosα+1+cos2nα=0.2\cos n(\pi-\alpha)\cos\alpha+1+\cos 2n\alpha=0. (18)

Since α(0,π)\alpha\in(0,\pi), for every n2n\geq 2 we have that none of (17) and (18) are satisfied if απ\frac{\alpha}{\pi} is an irrational number. Hence, we have that the determinant is non zero and then an=bn=0a_{n}=b_{n}=0 for every n2n\geq 2. It follows that p(t)=a0+a1cost+b1sintp(t)=a_{0}+a_{1}\cos t+b_{1}\sin t, i.e., pp is the support function of a disc (see for instance [19]). \Box

Remark 1.

If απ\frac{\alpha}{\pi} is a rational number, then there exist convex bodies KK different from discs for which c(t)c(t) is constant. For instance: for the angles α1=π5\alpha_{1}=\frac{\pi}{5} and α2=3π5\alpha_{2}=\frac{3\pi}{5} let KK be the convex body whose support function is given by p(t)=60+cos5t+sin5tp(t)=60+\cos 5t+\sin 5t (see Fig. 2). In this case the isoptic curves Kα1K_{\alpha_{1}} and Kα2K_{\alpha_{2}} have the property that its chords tangent to KK have constant values c1c_{1} and c2c_{2}, respectively, with c1c2=12τ\frac{c_{1}}{c_{2}}=\frac{1}{2-\tau}, where τ=1+52\tau=\frac{1+\sqrt{5}}{2}. Moreover, Kα1K_{\alpha_{1}} is homothetic to Kα2K_{\alpha_{2}} with ratio of homothety equal to 12τ-\frac{1}{2-\tau}.

Refer to caption
Figure 2: A convex body with two isoptics with corresponding values of c(t)c(t) both of constant value for the angles α1=π5\alpha_{1}=\frac{\pi}{5} and α2=3π5\alpha_{2}=\frac{3\pi}{5}

However, if we impose the additional condition that q(t)q(t) is also constant, then KK must be a disc.

Theorem 2.

Let KK be a strictly convex body in the plane with differentiable boundary K\partial K, and let α(0,π)\alpha\in(0,\pi) be a fixed angle. Suppose c(t)=c0c(t)=c_{0}, and q(t)=q0q(t)=q_{0}, for every t[0,2π]t\in[0,2\pi], and for two positive numbers c0c_{0} and q0q_{0}. Then KK is a disc.

Proof. By some simple calculations we have that |γα(t)|=q(t)sinα|\gamma^{\prime}_{\alpha}(t)|=\frac{q(t)}{\sin\alpha} (see for instance [4]). Since c(t)=c0c(t)=c_{0} is also constant, we have

ddt(|γα(t+πα)γα(t)|2)=ddt(γα(t+πα)γα(t),γα(t+πα)γα(t))=0,\frac{d}{dt}(|\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t)|^{2})=\frac{d}{dt}(\langle\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t),\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t)\rangle)=0,

hence

γα(t+πα)γα(t),γα(t+πα)=γα(t+πα)γα(t),γα(t).\langle\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t),\gamma^{\prime}_{\alpha}(t+\pi-\alpha)\rangle=\langle\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t),\gamma^{\prime}_{\alpha}(t)\rangle.

It follows that

c0|γα(t+πα)|cosβ1=c0|γα(t)|cosβ2,c_{0}\cdot|\gamma^{\prime}_{\alpha}(t+\pi-\alpha)|\cos\beta_{1}=c_{0}\cdot|\gamma^{\prime}_{\alpha}(t)|\cos\beta_{2},

which implies that β1=β2\beta_{1}=\beta_{2}, where β1\beta_{1} is the angle between the vectors γα(t+πα)γα(t)\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t) and γα(t+πα)\gamma^{\prime}_{\alpha}(t+\pi-\alpha), and β2\beta_{2} is the angle between the vectors γα(t+πα)γα(t)\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t) and γα(t)\gamma^{\prime}_{\alpha}(t). It follows that the angle between the chord [γα(t+πα),γα(t)][\gamma_{\alpha}(t+\pi-\alpha),\gamma_{\alpha}(t)] and the tangent lines at γα(t)\gamma_{\alpha}(t) and γα(t+πα)\gamma_{\alpha}(t+\pi-\alpha), are equal (see Fig. 3). Similarly, we obtain that the angles between the chord [γα(tπ+α),γα(t)][\gamma_{\alpha}(t-\pi+\alpha),\gamma_{\alpha}(t)] and the tangent lines at γα(t)\gamma_{\alpha}(t) and γα(tπ+α)\gamma_{\alpha}(t-\pi+\alpha), are equal. Since the length of the tangent vector γα(t)\gamma^{\prime}_{\alpha}(t) is constant for every tt, and all the triangles γα(tπ+α)γα(t)γα(t+πα)\triangle\gamma_{\alpha}(t-\pi+\alpha)\gamma_{\alpha}(t)\gamma_{\alpha}(t+\pi-\alpha) are congruent, we also have that the angles between the chord [γα(tπ+α),γα(t+πα)][\gamma_{\alpha}(t-\pi+\alpha),\gamma_{\alpha}(t+\pi-\alpha)] and the tangent lines at γα(t+πα)\gamma_{\alpha}(t+\pi-\alpha) and γα(tπ+α)\gamma_{\alpha}(t-\pi+\alpha), are equal. By elementary geometry we have that the circle circumscribed to γα(tπ+α)γα(t)γα(t+πα)\triangle\gamma_{\alpha}(t-\pi+\alpha)\gamma_{\alpha}(t)\gamma_{\alpha}(t+\pi-\alpha) and the body KαK_{\alpha}, share the tangent lines at the points γα(tπ+α)\gamma_{\alpha}(t-\pi+\alpha), γα(t)\gamma_{\alpha}(t), and γα(t+πα)\gamma_{\alpha}(t+\pi-\alpha); under this condition it was proved in Lemma 3.3 in [8] that KαK_{\alpha} must be a disc. Now we use Theorem 2 (b) in [9] and conclude that KK is a disc. \Box

Refer to caption
Figure 3: The angle between the chord [γα(t+πα),γα(t)][\gamma_{\alpha}(t+\pi-\alpha),\gamma_{\alpha}(t)] and the tangent lines are equal

Let h(t)h(t) denote the length of the segment from γα(t)\gamma_{\alpha}(t) to the projection of OO into the support line of KK at γ(t)\gamma(t). It is easy to show that

h(t)=1sinα[p(t)cosα+p(t+πα)].h(t)=\frac{1}{\sin\alpha}[p(t)\cos\alpha+p(t+\pi-\alpha)]. (19)
Refer to caption
Figure 4: The value of h(t)h(t) is constant

We have the following result.

Theorem 3.

Let KK be a strictly convex body in the plane with differentiable boundary and let α(0,π)\alpha\in(0,\pi) be a fixed angle. Suppose h(t)=h0h(t)=h_{0}, for every t[0,2π]t\in[0,2\pi], for a positive number h0.h_{0}. Then KK is a disc.

Proof. By (19) we have that

p(t)cosα+p(t+πα)=0.p^{\prime}(t)\cos\alpha+p^{\prime}(t+\pi-\alpha)=0. (20)

Using the expansion in Fourier series for p(t)p^{\prime}(t) (see equation (14)) we have that

n=1(nbncosntcosαnansinntcosα+nbncosn(t+πα)nansinn(t+πα))=0.\sum_{n=1}^{\infty}(nb_{n}\cos nt\cos\alpha-na_{n}\sin nt\cos\alpha+nb_{n}\cos n(t+\pi-\alpha)-na_{n}\sin n(t+\pi-\alpha))=0.

Using trigonometric identities and simplifying we conclude that for all nn and for all t[0,2π]t\in[0,2\pi] we must have

0\displaystyle 0 =[nsinn(πα)an+(ncosα+ncosn(πα))bn]cosnt\displaystyle=[-n\sin n(\pi-\alpha)a_{n}+(n\cos\alpha+n\cos n(\pi-\alpha))b_{n}]\cos nt
+[(ncosαncosn(πα))annsinn(πα)bn]sinnt.\displaystyle+[(-n\cos\alpha-n\cos n(\pi-\alpha))a_{n}-n\sin n(\pi-\alpha)b_{n}]\sin nt.

This yields to the system of equations

[nsinn(πα)ncosα+ncosn(πα)ncosαncosn(πα)nsinn(πα)][anbn]=[00].\begin{bmatrix}-n\sin n(\pi-\alpha)&n\cos\alpha+n\cos n(\pi-\alpha)\\ -n\cos\alpha-n\cos n(\pi-\alpha)&-n\sin n(\pi-\alpha)\end{bmatrix}\begin{bmatrix}a_{n}\\ b_{n}\end{bmatrix}=\begin{bmatrix}0\\ 0\end{bmatrix}.

Notice that the determinant of the matrix above is given by

n2sin2n(πα)+n2(cosα+cosn(πα))2.n^{2}\sin^{2}n(\pi-\alpha)+n^{2}(\cos\alpha+\cos n(\pi-\alpha))^{2}.

This determinant is never zero, for if sinn(πα)=0\sin n(\pi-\alpha)=0, then n(πα)=kπn(\pi-\alpha)=k\pi, for some integer kk. Nonetheless, in this case cosn(πα)=(1)k\cos n(\pi-\alpha)=(-1)^{k} and since α(0,π)\alpha\in(0,\pi), it is impossible to have cosα+cosn(πα)=0\cos\alpha+\cos n(\pi-\alpha)=0. It follows that the only solutions for the system of equations is an=bn=0a_{n}=b_{n}=0 for all nn. Thus, the solutions of the differential equation (20) are constant functions and KK must be a disc centred at OO. \Box

Theorem 4.

Let KK be a strictly convex body in the plane and let α(0,π)\alpha\in(0,\pi) be a fixed angle. Suppose λ(t)=λ0\lambda(t)=\lambda_{0}, for every t[0,2π]t\in[0,2\pi], for a positive number λ0,\lambda_{0}, and KK has rotational symmetry of angle πα\pi-\alpha or 2α2\alpha. Then KαK_{\alpha} is a circle.

Proof. The vector ν(t)=γα(t+πα)γα(tπ+α)\nu(t)=\gamma_{\alpha}(t+\pi-\alpha)-\gamma_{\alpha}(t-\pi+\alpha) can be expressed using the parametrization given in (2) by

ν(t)\displaystyle\nu(t) =[2p(t+πα)cosα+p(t2α)+p(t)]u(t)\displaystyle=[2p(t+\pi-\alpha)\cos\alpha+p(t-2\alpha)+p(t)]u(t)
+1sinα[p(t+πα)cos2αp(tπ+α)+p(t2α)cosαp(t)cosα]u(t).\displaystyle+\frac{1}{\sin\alpha}[p(t+\pi-\alpha)\cos 2\alpha-p(t-\pi+\alpha)+p(t-2\alpha)\cos\alpha-p(t)\cos\alpha]u^{\prime}(t).

If KK has rotational symmetry of angle πα\pi-\alpha or 2α2\alpha we have that

ν(t)\displaystyle\nu(t) =[2p(t+πα)cosα+2p(t)]u(t)\displaystyle=[2p(t+\pi-\alpha)\cos\alpha+2p(t)]u(t)
+1sinα[p(t+πα)cos2αp(tπ+α)]u(t),\displaystyle+\frac{1}{\sin\alpha}[p(t+\pi-\alpha)\cos 2\alpha-p(t-\pi+\alpha)]u^{\prime}(t),

or equivalently

ν(t)=[2p(t+πα)cosα+2p(t)]u(t)+[p(t+πα)(2sinα)]u(t).\nu(t)=[2p(t+\pi-\alpha)\cos\alpha+2p(t)]u(t)+[p(t+\pi-\alpha)(-2\sin\alpha)]u^{\prime}(t).

In this case |ν(t)|2=λ2(t)|\nu(t)|^{2}=\lambda^{2}(t) can be easily calculated as

λ02=4[p2(t+πα)+p2(t)+2p(t+πα)p(t)cosα],\lambda_{0}^{2}=4[p^{2}(t+\pi-\alpha)+p^{2}(t)+2p(t+\pi-\alpha)p(t)\cos\alpha],

hence

λ024=p2(t+πα)+p2(t)2p(t+πα)p(t)cos(πα)=d(t)2,\frac{\lambda_{0}^{2}}{4}=p^{2}(t+\pi-\alpha)+p^{2}(t)-2p(t+\pi-\alpha)p(t)\cos(\pi-\alpha)=d(t)^{2},

i.e., the value of d(t)=λ02.d(t)=\frac{\lambda_{0}}{2}.

Refer to caption
Figure 5: The value of d(t)d(t) is constant

Let x(t)x(t) and y(t)y(t) be the projections of OO into the support lines of KK at γ(t)\gamma(t) and γ(t+πα)\gamma(t+\pi-\alpha), respectively (see Fig. 5). The quadrilateral Ox(t)γα(t)y(t)Ox(t)\gamma_{\alpha}(t)y(t) is cyclic, i.e., there exist a circle which passes through its four vertices, hence |γα(t)O|=d(t)sinα=λ02sinα.|\gamma_{\alpha}(t)-O|=\frac{d(t)}{\sin\alpha}=\frac{\lambda_{0}}{2\sin\alpha}. It follows that KαK_{\alpha} is a circle centred at OO. \Box

4 Some comments about rotors in the plane

In this section we give some words about how the results obtained in this work are related to rotors in polygons. Moreover, in all the examples shown below, if we fix the convex body and the circumscribed polygon is rotated, while maintained circumscribed to KK, the vertices describe a isoptic of KK.

When c(t)c(t) has constant value, using the Fourier series for the support function of pp in the proof of theorem 1 we arrived to the equations

2sinn(πα)cosα+sin2nα=0-2\sin n(\pi-\alpha)\cos\alpha+\sin 2n\alpha=0

and

2cosn(πα)cosα+1+cos2nα=0.2\cos n(\pi-\alpha)\cos\alpha+1+\cos 2n\alpha=0.

If nn is even, then

sinnα(cosnα+cosα)=0\sin n\alpha(\cos n\alpha+\cos\alpha)=0

and

cosnα(cosnα+cosα)=0.\cos n\alpha(\cos n\alpha+\cos\alpha)=0.

Both of them are zero if cosnα+cosα=0,\cos n\alpha+\cos\alpha=0, or after some trigonometric transformations

cos(nα+α2)cos(nαα2)=0.\cos\left(\frac{n\alpha+\alpha}{2}\right)\cos\left(\frac{n\alpha-\alpha}{2}\right)=0.

It follows that α=(2r+1n+1)π\alpha=\left(\frac{2r+1}{n+1}\right)\pi or α=(2r+1n1)π\alpha=\left(\frac{2r+1}{n-1}\right)\pi, where rr is any integer number. In this case the determinant of the associated matrix is zero and we can choose the coefficients ana_{n}, bnb_{n}, arbitrarily. For instance, for n=4n=4 we select α=π3\alpha=\frac{\pi}{3}, a0=30a_{0}=30, a4=0a_{4}=0, b4=1b_{4}=1, and for any other natural number nn we have that an=bn=0a_{n}=b_{n}=0, i.e., the support function of KK is p(t)=30+sin4tp(t)=30+\sin 4t (see Fig. 6). The body KK in this example is centrally symmetric.

Refer to caption
Figure 6: A centrally symmetric rotor in the equilateral triangle

If nn is odd, then

sinnα(cosnαcosα)=0\sin n\alpha(\cos n\alpha-\cos\alpha)=0

and

cosnα(cosnαcosα)=0.\cos n\alpha(\cos n\alpha-\cos\alpha)=0.

Both of them are zero if cosnαcosα=0,\cos n\alpha-\cos\alpha=0, or after some trigonometric transformations

sin(nα+α2)sin(nαα2)=0.\sin\left(\frac{n\alpha+\alpha}{2}\right)\sin\left(\frac{n\alpha-\alpha}{2}\right)=0.

It follows that α=(2rn+1)π\alpha=\left(\frac{2r}{n+1}\right)\pi or α=(2rn1)π\alpha=\left(\frac{2r}{n-1}\right)\pi, where rr is any integer number. Notice that we can obtain an example for any angle of the form α=sqπ\alpha=\frac{s}{q}\pi, where ss and qq are integers such that 0<s<q0<s<q. We just use this case, since α=2s2qπ\alpha=\frac{2s}{2q}\pi, we select r=sr=s and n=2q+1n=2q+1 or n=2q1n=2q-1. For instance, for α=2π3π=4π6π\alpha=\frac{2\pi}{3}\pi=\frac{4\pi}{6}\pi, and then we select n=7n=7. In Fig. 7 we show the body KK with its isoptic K2π/3,K_{2\pi/3}, with the property that c(t)c(t) has a constant value. The support function for this example is p(t)=80+cos7tp(t)=80+\cos 7t. The body in this example has constant width.

Refer to caption
Figure 7: A rotor with constant width in the regular hexagon

The example shown in Fig. 8 has coefficients different from zero for n=4n=4 and n=5n=5, consequently, the body KK obtained is neither of constant width or centrally symmetric. The support function of KK is p(t)=70+sin4t+cos5t.p(t)=70+\sin 4t+\cos 5t.

Refer to caption
Figure 8: A rotor in the equilateral triangle, which is neither of constant width nor centrally symmetric

Finally, we give an example of a rotor in the square. The support function is p(t)=60+cos5tp(t)=60+\cos 5t.

Refer to caption
Figure 9: A rotor for the square

5 An inequality about the length of some special chords

The following inequality and characterization of the disc, in terms of the length q(t)q(t), was proved in [7].

Theorem JY. Let KK be a strictly convex body in the plane with minimal width ω0\omega_{0}. For any fixed α(0,π)\alpha\in(0,\pi) there exists at least one value of the parameter tt, t(α)[0,2π]t(\alpha)\in[0,2\pi], such that q(t(α))ω0cosα2q(t(\alpha))\geq\omega_{0}\cos\frac{\alpha}{2}. Moreover, if there is not such a chord with length exceeding ω0cosα2\omega_{0}\cos\frac{\alpha}{2}, then KK is a disc.

Here we prove the following similar result.

Theorem 5.

Let KK be a convex body in the plane with minimal width ω0\omega_{0}. For any fixed α(0,π)\alpha\in(0,\pi) there exists at least one value of the parameter tt, t(α)[0,2π]t(\alpha)\in[0,2\pi], such that λ(t(α))2ω0cosα2\lambda(t(\alpha))\geq 2\omega_{0}\cos\frac{\alpha}{2}. Moreover, if there is not such a chord with length exceeding 2ω0cosα22\omega_{0}\cos\frac{\alpha}{2}, then KK is a convex body of constant width ω0\omega_{0}.

Proof. The mean value of the chords of KαK_{\alpha} that are tangent to KK is

c(t)¯\displaystyle\overline{c(t)} =12πsinα02π[2p(t+πα)cosα+p(t)+p(t2α)]𝑑t,\displaystyle=\frac{1}{2\pi\sin\alpha}\int_{0}^{2\pi}[2p(t+\pi-\alpha)\cos\alpha+p(t)+p(t-2\alpha)]dt,
=12πsinα[2L(K)cosα+2L(K)],(using Cauchy’s formula (3))\displaystyle=\frac{1}{2\pi\sin\alpha}[2L(K)\cos\alpha+2L(K)],\hskip 17.07182pt\text{(using Cauchy's formula (\ref{Cauchy}))}
=cotα2πL(K).\displaystyle=\frac{\cot\frac{\alpha}{2}}{\pi}L(K).

Indeed, by the same argument, the mean value of c(t)+c(tπ+α)c(t)+c(t-\pi+\alpha) is

c(t)+c(tπ+α)¯=2cotα2πL(K).\overline{c(t)+c(t-\pi+\alpha)}=\frac{2\cot\frac{\alpha}{2}}{\pi}L(K).

Let t[0,2π]t\in[0,2\pi] be any number and consider the triangle with sides of length c(t)c(t), c(tπ+α)c(t-\pi+\alpha), λ(t)\lambda(t), and angle α\alpha, as shown in Fig. 10. By Problem C4 in the book by I. Niven [15], we have that the minimum of λ(t)\lambda(t), among all triangles with a fixed angle α\alpha and the sum c(t)+c(tπ+α)=c0c(t)+c(t-\pi+\alpha)=c_{0}, for a constant number c0c_{0}, is obtained if and only if c(t)=c(tπ+α)=c0/2c(t)=c(t-\pi+\alpha)=c_{0}/2. It follows that for every t[0,2π]t\in[0,2\pi] it holds that

λ(t)[c(t)+c(tπ+α)]sinα2,\lambda(t)\geq[c(t)+c(t-\pi+\alpha)]\sin\frac{\alpha}{2},

with equality if and only if c(t)=c(tπ+α).c(t)=c(t-\pi+\alpha). Let t0[0,2π]t_{0}\in[0,2\pi] be such that c(t0)+c(t0π+α)=c(t)+c(tπ+α)¯,c(t_{0})+c(t_{0}-\pi+\alpha)=\overline{c(t)+c(t-\pi+\alpha)}, we have that

λ(t0)[c(t0)+c(t0π+α)]sinα2=2cosα2πL(K).\lambda(t_{0})\geq[c(t_{0})+c(t_{0}-\pi+\alpha)]\sin\frac{\alpha}{2}=\frac{2\cos\frac{\alpha}{2}}{\pi}L(K). (21)
Refer to caption
Figure 10: The minimum of λ(t)\lambda(t) is when c(t)=c(tπ+α)c(t)=c(t-\pi+\alpha)

Using Cauchy’s formula for the perimeter of KK we can also prove that L(K)πω0L(K)\geq\pi\omega_{0}, with equality if and only KK is a body of constant width ω0.\omega_{0}. Hence λ(t0)2ω0cosα2.\lambda(t_{0})\geq 2\omega_{0}\cos\frac{\alpha}{2}.

Now, if there is no chord with λ(t)>2ω0cosα2\lambda(t)>2\omega_{0}\cos\frac{\alpha}{2}, then λ(t0)=2ω0cosα2\lambda(t_{0})=2\omega_{0}\cos\frac{\alpha}{2}, which by (21) implies that L(K)=πω0,L(K)=\pi\omega_{0}, i.e., KK must be a body of constant width ω0\omega_{0} (see Fig. 11). \Box

Refer to caption
Figure 11: A convex body with λ(t)=2ω0cosα2\lambda(t)=2\omega_{0}\cos\frac{\alpha}{2} for every tt

6 Some results in higher dimensions

The following lemma is needed for some of the subsequent results.

Lemma 1.

Let K,LnK,L\subset\mathbb{R}^{n}, n3n\geq 3, be convex bodies with LintKL\subset\emph{int}\,K, such that every (n1)(n-1)-dimensional section of KK tangent to LL is an (n1)(n-1)-dimensional ball, then KK is a ball. If additionally, the centre of every tangent ball is at the point of contact with LL, then KK and LL are concentric balls.

Proof. There are several ways to prove this lemma. Here we give the following proof. First note that if all the 33-dimensional sections of a convex body through a given point in its interior are 33-dimensional balls, then it is an nn-dimensional ball. We consider any point in the interior of LL for such a point and since the hypothesis of the theorem is inheritated to every 33-dimensional section, we have that it is sufficient to prove the theorem in the 33-dimensional case.

Let xx be any point in K\partial K and let \ell be any line supporting KK at xx. Consider the two support planes of LL which share the line \ell, to say H1H_{1} and H2H_{2}. Let HH be any other support plane of LL through the point xx. By hypothesis HKH\cap K is a disc which intersects each one of the circles H1KH_{1}\cap\partial K and H2KH_{2}\cap\partial K at two points. The circle HKH\cap\partial K passes through three points of the set (H1K)(H2K)(H_{1}\cap\partial K)\cup(H_{2}\cap\partial K), and since there is a unique sphere which contains the two circles H1KH_{1}\cap\partial K and H2KH_{2}\cap\partial K, it holds that this sphere contains HKH\cap\partial K. Since HH is any support plane of LL through xx, we have that x={HK:H is a support plane of L through x}\mathcal{R}_{x}=\{H\cap\partial K:H\text{ is a support plane of }L\text{ through }x\} is a closed subset of a sphere. Let yKy\in\partial K be any point such that yx\mathcal{R}_{y}\cap\mathcal{R}_{x}\neq\emptyset, then yx\mathcal{R}_{y}\cup\mathcal{R}_{x} is contained in the same sphere. Continuing in this way, since KK is a compact set, we can prove that K\partial K is a sphere.

Now, let HH be any hyperplane supporting LL at a point zz and suppose the centre of the (n1)(n-1)-dimensional ball HKH\cap K is zz. The line orthogonal to HH through zz passes through the centre of the ball KK, this implies that indeed HH is the tangent plane of LL through zz, i.e., K\partial K is a differentiable surface. We have that all the normal lines of L\partial L passes through the centre of KK, this implies (see for instance [18]) that L\partial L is a sphere with centre at the centre of KK. We conclude that KK and LL are concentric balls. \Box

Let β=πα\beta=\pi-\alpha. In [5], J. W. Green proved that KK is a Euclidean disc if KαK_{\alpha} is a circle and any of the following conditions hold: β\beta is an irrational multiple of π\pi, or β=2mnπ\beta=\frac{2m}{n}\pi, with 2m2m and nn relatively prime integer numbers. On the other side, M.S. Klamkin conjectured [11] and J. Nitsche proved [14] that two different values α1\alpha_{1}, α2\alpha_{2}, such that Kα1K_{\alpha_{1}} and Kα2K_{\alpha_{2}} are circles, are enough to prove that KK is a Euclidean disc. A similar result was proved by Á. Kurusa and T. Ódor in [12] in n\mathbb{R}^{n}, n3n\geq 3, where the isoptic surface is defined using the solid angle under which is seen a convex body KK. They proved that if two isoptic surfaces of a given convex body are concentric spheres, then it is a ball. The following is a variant of this kind of results.

Theorem 6.

Let K3K\subset\mathbb{R}^{3} be a strictly convex body contained in the interior of a sphere 𝒮\mathcal{S}. Suppose that for every 22-dimensional plane HH, with HKH\cap K\neq\emptyset, it holds that H𝒮H\cap\mathcal{S} is a isoptic curve of HK.H\cap K. Then KK is a ball concentric with 𝒮\mathcal{S}.

Proof. We first observe the following: If for some angle α(0,π)\alpha\in(0,\pi) the α\alpha-isoptic curve of a planar body MM is a circle SS, then its centre xx belongs to MM. Suppose this is not the case. Let yMy\in M be the point which is closest to xx. Let z𝒮z\in\mathcal{S} be the point where the ray xy\overrightarrow{xy} intersects to 𝒮\mathcal{S} and let zz^{\prime} be the point in 𝒮\mathcal{S} such that [z,z][z,z^{\prime}] is a diameter. Let \ell be the line through xx which is orthogonal to [z,z][z,z^{\prime}]. The central projection of MM into the line \ell is a segment [a,b][a,b] that contains the centre xx. The angle formed by the lines zaz^{\prime}a and zbz^{\prime}b contains MM, however, at least one of the lines zaz^{\prime}a or zbz^{\prime}b does not intersect MM. Since the angles azb\measuredangle azb and azb\measuredangle az^{\prime}b are equal, we obtain that the angles under which MM is seen from the points zz and zz^{\prime} are not equal. This contradicts that SS is the isoptic curve of angle α\alpha of MM. Hence, we have that xx belongs to MM.

Refer to caption
Figure 12: The centre xx belong to MM

Now let qKq\in\partial K be any arbitrary point and let HqH_{q} be a plane supporting KK at qq. By hypothesis we have that Hq𝒮H_{q}\cap\mathcal{S} is a circle; we are going to prove that its centre is qq. Consider a sequence of planes {Hr}\{H_{r}\} parallel to HqH_{q} in such a way that the sequence converges to HqH_{q}. Let xrx_{r} be the centre of the circle Hr𝒮H_{r}\cap\mathcal{S}, for every natural number rr. By the comment above, we have that xrHrKx_{r}\in H_{r}\cap K, for every rr\in\mathbb{N}, and since HrKHqK=qH_{r}\cap K\longrightarrow H_{q}\cap K=q we have that xrqx_{r}\longrightarrow q when rr\longrightarrow\infty. In other words, the centre of Hq𝒮H_{q}\cap\mathcal{S} is qq. Now we apply Lemma 1 and conclude that KK is a ball concentric with 𝒮\mathcal{S}. \Box

The following result shows that in dimension 3 or higher, only Euclidean balls have an equichordal convex body.

Theorem 7.

Let KnK\subset\mathbb{R}^{n}, n3n\geq 3, be a strictly convex body which possesses an equichordal convex body LL in its interior. Then KK and LL are concentric balls.

Proof. Suppose the length of the chords of KK tangent to LL is λ\lambda. First we prove that LL is strictly convex, i.e., LL has not segments in its boundary. Suppose to the contrary that there is a segment [a,b]L[a,b]\subset\partial L and consider any 22-dimensional plane HH supporting LL at [a,b][a,b]. By hypothesis, every chord of HKH\cap K through aa has length λ\lambda, and the same happens for every chord through bb. This implies that HKH\cap K has two equichordal points, but as were proved by M. R. Rychlik [16] this is not possible.

Now, let HH be any 22-dimensional plane supporting LL at a point xx. We will prove that the diameter of HKH\cap K is λ\lambda. Suppose this is not the case and there is a chord [a,b][a,b] of HKH\cap K with |ab|>λ.|a-b|>\lambda. Clearly, [a,b]L=.[a,b]\cap L=\emptyset. Consider a 22-dimensional plane Π\Pi that contains to [a,b][a,b] and such that ΠintL.\Pi\cap\text{int}\,L\neq\emptyset. Let c,d(ΠL)c,d\in\partial(\Pi\cap L) such that the lines through cc and dd, respectively, that are parallel to [a,b][a,b] are support lines of ΠL\Pi\cap L. Since |ab|>λ|a-b|>\lambda, we have that there exist a chord [e,f][e,f] of ΠK\Pi\cap K that is separated from ΠL\Pi\cap L by the chord [a,b][a,b] and |ef|=λ|e-f|=\lambda. We have three chords of ΠK\Pi\cap K parallel to [a,b][a,b] and with length λ\lambda. This contradicts that KK is strictly convex, then the diameter of HKH\cap K must be λ\lambda.

Every chord of HKH\cap K through xx has length equal to λ\lambda, then every chord of HKH\cap K through xx is a binormal of HKH\cap K; it follows that HKH\cap K is a disc centred at xx (see for instance [3]). Since this is true for every 22-dimensional plane HH through xx, we have that any (n1)(n-1)-dimensional section of KK tangent to LL at xx is an (n1)(n-1)-dimensional sphere centred at xx. This is also true for every xLx\in\partial L, so we apply Lemma 1 and conclude that KK and LL are concentric balls. \Box

Corollary 1.

Let K,LnK,L\subset\mathbb{R}^{n}, n3n\geq 3, be two convex bodies. Suppose that for every hyperplane HH that intersects to LL it holds that HLH\cap L is an equichordal body of HKH\cap K. Then KK and LL are concentric balls.

Proof. It is quite simple to prove that all the chords of KK tangent to LL have the same length. The conclusion follows applying Theorem 7. \Box

Suppose now that KK is a strictly convex body in 3\mathbb{R}^{3}. For z1,z2Kz_{1},z_{2}\in\partial K, we will say that the segment [z1,z2][z_{1},z_{2}] is an α\alpha-chord of KK, if there exist tangent planes at z1z_{1} and z2z_{2} meeting at an angle α\alpha. We present the following analogue to Theorem JY.

Theorem 8.

Let KK be a strictly convex body in 3\mathbb{R}^{3} with boundary of class C1C^{1} and with minimal width ω0\omega_{0}. For any fixed α(0,π)\alpha\in(0,\pi) there exists an α\alpha-chord of KK with length at least ω0cosα2\omega_{0}\cos\frac{\alpha}{2}. Moreover, if there is not an α\alpha-chord with length exceeding ω0cosα2\omega_{0}\cos\frac{\alpha}{2}, then KK is a ball.

Proof. Let u0𝕊2u_{0}\in\mathbb{S}^{2} be a direction such that the width of KK in this direction has minimal value ω0\omega_{0}. Let υ𝕊2\upsilon\in\mathbb{S}^{2} be a vector orthogonal to u0u_{0} and let’s denote the orthogonal projection on the plane υ\upsilon^{\bot} by πυ\pi_{\upsilon}. Clearly, the minimum width of K=πυ(K)K^{\prime}=\pi_{\upsilon}(K) is also ω0\omega_{0} and every α\alpha-chord of KK^{\prime} is obtained as the projection of some α\alpha-chord of KK. By Theorem JY there is an α\alpha-chord of KK^{\prime} with length at least ω0cosα2\omega_{0}\cos\frac{\alpha}{2} and so there is an α\alpha-chord of KK with length larger than or equal to ω0cosα2\omega_{0}\cos\frac{\alpha}{2}.

Suppose now that all the α\alpha-chords of KK have length less than or equal to ω0cosα2\omega_{0}\cos\frac{\alpha}{2}. Then all the α\alpha-chords of πυ(K)\pi_{\upsilon}(K), for every υ𝕊2\upsilon\in\mathbb{S}^{2}, have length less than or equal to ω0cosα2\omega_{0}\cos\frac{\alpha}{2}. It follows from Theorem JY that πυ(K)\pi_{\upsilon}(K) is a disc. Since υ\upsilon is an arbitrary direction, we have that every 22-dimensional projection of KK is a disc and therefore KK is a ball. \Box

Theorem 9.

Let KK be a strictly convex body in 3\mathbb{R}^{3} with boundary of class C1C^{1} and with π2\frac{\pi}{2}-chords of constant length λ\lambda, then KK is a ball.

Proof. For any point zKz\in\partial K, let HzH_{z} denote the tangent plane of KK at zz and let Az={wK:[z,w] is an π2-chord of K}A_{z}=\{w\in\partial K:[z,w]\text{ is an }\frac{\pi}{2}\text{-chord of }K\}. For each point ww in AzA_{z}, we have that HwH_{w} intersects HzH_{z} with angle π2\frac{\pi}{2}. Let z,zKz,z^{\prime}\in\partial K be two points such that [z,z][z,z^{\prime}] is normal to HzH_{z} and HzH_{z^{\prime}}. For every wAzw\in A_{z}, the triangle Δzzw\Delta zz^{\prime}w is isosceles with sides [z,w][z,w] and [z,w][z^{\prime},w] of equal length. Then AzA_{z} is a circle centred at (z+z)/2(z+z^{\prime})/2. Moreover, given any two antipodal points z1z_{1} and z1z_{1}^{\prime} of AzA_{z}, the segment [z1,z1][z_{1},z_{1}^{\prime}] must be orthogonal to the planes Hz1H_{z_{1}} and Hz1H_{z_{1}^{\prime}}. By repeating the previous argument to each pair of antipodal points of AzA_{z}, we have that any section of KK containing [z,z][z,z^{\prime}] is a disc. This implies that KK is a ball. \Box

References

  • [1] J. A. Barker, and D. G. Larman, Determination of convex bodies by certain sets of sectional volumes. Discrete Math. 241 (2001), 799679-96.
  • [2] W. Blaschke, W. Rothe, and R. Weitzenböck, Aufgabe 552, Arch. Math. Phys. 27 (1917), 8282.
  • [3] G. D. Chakerian, and H. Groemer. Convex bodies of constant width, in Convexity and its Applications, Eds. P.M. Gruber and J.M. Wills, Birkhäuser, Basel, (1983).
  • [4] W. Cieślak, A. Miernowski, and W. Mozgawa. Isoptics of a closed strictly convex curve, Lecture Notes in Mathematics 1481 (1991), 283528-35.
  • [5] J. W. Green. Sets subtending a constant angle on a circle, Duke Math. J. 17 (1950), 263267263-267.
  • [6] H. Groemer. Geometric applications of Fourier series and spherical harmonics, Cambridge Univ. Press, Cambridge (1996).
  • [7] J. Jerónimo-Castro, and C. Yee-Romero, An inequality for the length of isoptic chords of convex bodies. Aequat. Math. 93 (3) (2019), 619628619-628.
  • [8] J. Jerónimo-Castro, and F. G. Jimenez-Lopez. Symmetries of convex sets in the hyperbolic plane, J. Conv. Analysis 26 (2019), 107710881077-1088.
  • [9] J. Jerónimo-Castro, F. G. Jimenez-Lopez, and R. Jiménez-Sánchez. On convex bodies with isoptic chords of constant length, Aequationes Math. 94 (2020), no. 6, 118911991189-1199.
  • [10] J. Jerónimo-Castro, M. A. Rojas-Tapia, U. Velasco-García, and C. Yee-Romero. An isoperimetric type inequality for isoptic curves of convex bodies, Results Math. 75 (2020).
  • [11] M. S. Klamkin. Conjectured isoptic characterization of a circle, Amer. Math. Monthly 95 (1988), 845845.
  • [12] Á. Kurusa, and T. Ódor. Isoptic characterization of spheres, J. Geom. 106 (2015), 637363-73.
  • [13] H. Martini, L. Montejano, D. Oliveros. Bodies of Constant Width: An Introduction to Convex Geometry with Applications, Birkhäuser, (2019).
  • [14] J. Nitsche. Isoptic Characterization of a Circle (Proof of a conjecture of M. S. Klamkin), Amer. Math. Monthly 97 (1990), 454745-47.
  • [15] I. Niven. Maxima and Minima without Calculus, Dolciani Mathematical Expositions No. 6, (1981).
  • [16] M. R. Rychlik. A complete solution to the equichordal point problem of Fujiwara, Blaschke, Rothe and Weitzenböck, Invent. Math. 129 (1997), no. 1, 141212141-212.
  • [17] C. Schütt, and E. Werner. The convex floating body, Math. Scand. 66 (1990), 275290.275-290.
  • [18] V. A. Toponogov. Differential geometry of curves and surfaces, a concise guide, Birkhäuser, Boston-Basel-Berlin (2006).
  • [19] F. A. Valentine. Convex sets. McGraw-Hill Series in Higher Mathematics, McGraw-Hill, New York, (1964).
  • [20] F. Wegner. Floating bodies of equilibrium, Stud. Appl. Math. 111 (2003), no. 2, 167183167-183.
  • [21] I. M. Yaglom, and V. G. Boltyanski. Convex Figures, Holt, Rinehart and Winston (1961)

This manuscript has not any associated data.