1 Introduction

In Ref. [2] it is proved that for every reduced polygon in the Euclidan plane there exists a boundary point of this polygon, such that the ball centered at this point and with radius equal to the thickness of the polygon covers it. In this paper we present a proof of an analogous theorem for sphere, but we use here a different approach.

At first recall some basic notions.

By \(S^2\) we denote the unit sphere in the three-dimensional Euclidean space \(E^{3}\). By a great circle of \(S^2\) we mean the intersection of \(S^2\) with any two-dimensional subspace of \(E^{3}\). The intersection of \(S^2\) with a one-dimensional subspace of \(E^{3}\) is a pair of points, which is called a pair of antipodes. Observe that if two different points ab are not antipodes, there is exactly one great circle containing them. This great circle consists of two parts laying between a and b. The shorter of these parts is called the arc connecting a and b, or shortly arc. We denote it by ab. By the spherical distance (or shortly distance) of a and b we mean the length of the arc ab and we denote it by |ab|. For ab being a pair of antipodes, we put \(|ab|=\pi \).

A subset of \(S^2\) is called convex if it does not contain any pair of antipodes of \(S^2\) and if together with every two points it contains the arc connecting them. A closed convex set with non-empty interior is called a spherical convex body.

Let \(\rho \in \left( 0, \frac{\pi }{2}\right] \). By the disk \(B_\rho (c)\) of radius \(\rho \) and center \(c\in S^2\) we mean the set of points of \(S^2\) at the distance at most \(\rho \) from c. The boundary of a disk is called a circle. We denote by \({\overline{B}}_\rho (c)\) the circle bounding \(B_\rho (c)\).

Let a and b be different points of \({\overline{B}}_\rho (c)\) such that abc do not lay on a great circle. Observe that a and b divide \({\overline{B}}_\rho (c)\) into two parts. The smaller of these parts is called a small arc of \({\overline{B}}_\rho (c)\). We denote it by \({\widehat{ab}}\). We say that \({\widehat{ab}}\) is of radius \(\rho \) and its center is c. Points a and b are called the endpoints of this small arc.

For any point p the disk \(B_{\frac{\pi }{2}}(p)\) is called a hemisphere and is denoted by H(p). If pq are antipodes, then H(p) and H(q) are called opposite hemispheres. By a semicircle we mean the intersection of a hemisphere with a great circle different from the boundary of this hemisphere. For any boundary point s of a convex body \(C \subset S^2\) we say that a hemisphere H supports C at s if \(C \subset H\) and s belongs to the great circle bounding H. If F is a subset of an open hemisphere, we define \(\text {conv} (F) \) as the smallest convex set containing F. We say that e is an extreme point of C if \(C\setminus \{ e\}\) is a convex set. We denote the set of all extreme points of C by E(C).

If hemispheres G and H are different and not opposite, then \(L = G \cap H\) is called a lune. The two semicircles bounding L and contained in G and H, respectively, are denoted by G/H and H/G. The thickness \(\Delta (L)\) of \(L \subset S^2\) is defined as the distance of the midpoints of G/H and H/G. The arc connecting these midpoints is called the string of the lune L.

For every hemisphere K supporting a convex body \(C\subset S^2\) we find the hemispheres \(K^*\) supporting C such that the lunes \(K\cap K^*\) are of the minimum thickness (by compactness arguments at least one such a hemisphere \(K^*\) exists). The thickness of the lune \(K \cap K^*\) is called the width of C determined by K and it is denoted by \({\textrm{width}}_K (C)\) (see [4]). If for all hemispheres K supporting C the numbers \({\textrm{width}}_K (C)\) are equal, we say that C is of constant width (see [4]). By the thickness \(\Delta (C)\) of C we understand the minimum width of C determined by K over all supporting hemispheres K of C (see [4]).

After [4] we call a spherical convex body \(R \subset S^2\) reduced if \(\Delta (Z) < \Delta (R)\) for every convex body \(Z \subset R\) different from R. Simple examples of reduced spherical convex bodies on \(S^2\) are spherical bodies of constant width and, in particular, the disks on \(S^2\). From the paper [5] we see which spherical polygons are reduced. Theorem 3.1 of this paper says that all reduced spherical polygons are of thickness at most \(\frac{\pi }{2}\).

Most of above notions are similar to ones in the Euclidan space \(E^d\), where we have a wide knowledge about reduced bodies, see for instance [9]. Since this subject is well explored in the Euclidan plane, it is natural to investigate reduced bodies on the sphere. There are many papers about reduced bodies on \(S^d\), see [3, 4, 7, 10, 11]. There are also some papers considering reduced bodies in finite-dimensional normal spaces, see for instance [8]. See also the monograph [12], which examines reduced bodies and bodies of constant width in various spaces, in particural on the sphere.

In Refs. [1] and [13] there are solutions of some of spherical reduced bodies problems, which earlier were solved for the plane. In this paper we give one more analogous property.

2 A few lemmas

For the proof of our theorem we need some lemmas.

Lemma 1

Let \(F\subset S^2\) and \(\rho \in \left( 0, \frac{\pi }{2}\right] \). If \(conv (F)\ne S^2\) and \(p \in conv (F)\), then \(\bigcap _{f\in F}B_\rho (f) \subset B_\rho (p)\).

Proof

In the first case assume that there are extreme points \(e_1,e_2\) of \(\text {conv} (F)\) such that \(p\in e_1e_2\). Clearly \(e_1,e_2 \in F\). Let \(x\in \bigcap _{f\in F}B_\rho (f)\). In particular \(x\in B_\rho (e_1)\) and \(x\in B_\rho (e_2)\), which means that \(|xe_1|\le \rho \) and \(|xe_2|\le \rho \). Let \(x'\) be such a point that \(p\in xx'\) and \(|px|=|px'|\). It is easy to observe that \(|x'e_1|=|xe_2|\). Therefore by the triangle inequality for the triangle \(xx'e_1\) we get

$$\begin{aligned} |px| = \frac{1}{2} |xx'|\le \frac{1}{2} \left( |xe_1| +|x'e_1|\right) =\frac{1}{2} \left( |xe_1| +|xe_2| \right) \le \max \{ |xe_1|, |xe_2|\} \le \rho . \end{aligned}$$

It means that \(x\in B_\rho (p)\) and so \(\bigcap _{f\in F}B_\rho (f) \subset B_\rho (p)\), which ends the proof in the first case.

Consider now the second case when p does not belong to any arc with endpoints at extreme points of \(\text {conv} (F)\). There exist an extreme point e and a boundary point b of \(\text {conv} (F)\) such that \(p\in eb\). There also exist extreme points \(e_1,e_2\) of \(\text {conv} (F)\) such that \(b\in e_1e_2\). By the same reason as in the first case we have \(B_\rho (e_1)\cap B_\rho (e_2)\subset B_\rho (b)\) and \(B_\rho (e)\cap B_\rho (b)\subset B_\rho (p)\). Since \(e,e_1,e_2\in F\), we conclude that

$$\begin{aligned} \bigcap _{f\in F}B_\rho (f)\subset B_\rho (e)\cap B_\rho (e_1)\cap B_\rho (e_2)\subset B_\rho (e)\cap B_\rho (b)\subset B_\rho (p). \end{aligned}$$

This ends the proof in the second case. \(\square \)

For any \(F\subset S^2\) we define the set \(F_\rho ^{\circ }\) as \(\left\{ p: F\subset B_\rho (p)\right\} \).

Lemma 2

If C is a spherical convex body and \(\rho \in \left( 0, \frac{\pi }{2}\right] \), then \(C_\rho ^{\circ }= \bigcap _{e\in E(C)}B_\rho (e)\).

Proof

We have:

$$\begin{aligned}{} & {} x\in C_\rho ^{\circ }\Leftrightarrow C \subset B_\rho (x) \Leftrightarrow \forall _{p\in C}\ p\in B_\rho (x) \Leftrightarrow \\{} & {} \quad \Leftrightarrow \forall _{p\in C}\ |px| \le \rho \Leftrightarrow \forall _{p\in C}\ x\in B_\rho (p) \Leftrightarrow x \in \bigcap _{p\in C} B_\rho (p). \end{aligned}$$

Thus in order to end the proof it is sufficient to show that \(\bigcap _{p\in C} B_\rho (p)=\bigcap _{e\in E(C)} B_\rho (e)\).

Since \(E(C) \subset C\), it is obvious that \(\bigcap _{p\in C} B_\rho (p)\subset \bigcap _{e\in E(C)} B_\rho (e)\). Let us show that \(\bigcap _{p\in C} B_\rho (p)\supset \bigcap _{e\in E(C)} B_\rho (e)\). Take any point \(p\in C\). From \(p \in \text {conv} (E(C))\) and Lemma 1 we conclude that \(\bigcap _{e\in E(C)} B_\rho (e) \subset B_\rho (p)\). This inclusion holds true for every \(p\in C\), therefore \(\bigcap _{e\in E(C)} B_\rho (e) \subset \bigcap _{p\in C}B_\rho (p)\) which ends the proof. \(\square \)

The following lemma is a special case of Lemma 3 from [6].

Lemma 3

If the endpoints of a small arc \({\widehat{ab}}\) of radius \(\rho \) are contained in \(B_\rho (c)\), then \({\widehat{ab}}\subset B_\rho (c)\).

Next lemma describes the shape of the boundary of a spherical convex body. We denote the boundary of C in the common way by \(\textrm{bd} (C)\) and its diameter by \(\textrm{diam}(C)\).

Lemma 4

Let F be the intersection of a finite number of different disks of radius \(\rho \) and F has the non-empty interior. Then \(bd (F)\) consists of small arcs of radius \(\rho \) with different centers.

Proof

We apply induction. For two disks the thesis of lemma is obvious. Assume that the intersection F of a finite family \({\mathcal {D}}\) of disks of radius \(\rho \) has non-empty interior and its boundary consists of small arcs of radius \(\rho \) with different centers. Intersect F with a disk B of radius \(\rho \) whose center is different from the centers of the disks from \({\mathcal {D}}\). Also assume that \(F\cap B\) has non-empty interior. If \(F\subset B\), then \(F\cap B= F\) and by the assumption we already know that the inductive thesis holds true. Otherwise the circle bounding B intersects the interior of F. Observe that if points ab belong to \(\textrm{bd}(B) \cap F\), then they belong to every disk from \({\mathcal {D}}\). Therefore by Lemma 3 it follows that \({\widehat{ab}}\subset \textrm{bd}(B)\). Hence \({\widehat{ab}}\) is a subset of every disk from \({\mathcal {D}}\) and thus it is a subset of F. As a consequence, \(\textrm{bd}(B) \cap F\) is a small arc. Without losing the generality we may assume that a and b are the endpoints of this arc.

We conclude that the boundary of \(B\cap F\) consists of the small arc \({\widehat{ab}}\) and small arcs being parts of boundaries of some disks from \({\mathcal {D}}\). Therefore also in this case the inductive thesis holds true. \(\square \)

We define \(E^*(V)\) as \(\left\{ e\in E(V): |\textrm{bd}(V_r^{\circ }) \cap {\overline{B}}_r(e)|>1\right\} \). From Lemmas 2 and 4 we immediately obtain the following corollary.

Corollary 1

If V is a spherical convex polygon, then \(V_\rho ^{\circ }= \bigcap _{e\in E^{*}(V)}B_\rho (e)\). For every point \(e\in E^*(V)\) the intersection of \(\textrm{bd}(V_\rho ^{\circ })\) and \({\overline{B}}_\rho (e)\) is a small arc of \({\overline{B}}_\rho (e)\).

Applying this tool we prove the following theorem.

3 The covering by a disk

Theorem

Every spherical reduced polygon V is contained in a disk of radius \(\Delta (V)\) centered at a boundary point of V.

Proof

Recall that \(V_{\Delta (V)}^{\circ }\) is the set of centers of disks of radius \(\Delta (V)\) that contain V. So it is sufficient to show that \(V_{\Delta (V)}^{\circ }\) has a common point with the boundary of V.

Assume the opposite that \(V_{\Delta (V)}^{\circ }\) is a subset of the interior of V.

By order we always mean the order according to the positive orientation.

Let \(e_1, e_2, \ldots , e_m\) be all points of \(E^*(V)\) laying on the boundary of V in that order. Put \(e_{m+1}=e_1\). For \(i=1,2, \ldots , m\) denote by \(p_{i+1}\) the point of \(V_{\Delta (V)}^{\circ }\) which belongs to \({\overline{B}}_{\Delta (V)}(e_i)\cap {\overline{B}}_{\Delta (V)}(e_{i+1})\). Put \(p_{m+1}=p_1\). According to Corollary 1 we see that the boundary of \(V_{\Delta (V)}^{\circ }\) consists only of small arcs \(\widehat{p_ip_{i+1}}\) of \({\overline{B}}_r(e_i)\).

Since for every \(i=1,2,\ldots , m\) the point \(e_i\) is an extreme point of V, by Theorem 4 of Ref. [4] there exists a hempishere \(L_i\) of thickness \(\Delta (V)\) containing V, whose string has an endpoint at \(e_i\). Denote the other end of this string by \(b_i\).

Since \(|e_1b_1| = \Delta (V)\), we see that \(b_1\) lays in the boundary of \(B_{\Delta (V)}(e_1)\). But from our assumption that \(V_{\Delta (V)}^{\circ }\) is a subset of the interior of V we conclude that \(b_1\notin \widehat{p_1p_2}\). Therefore \(b_1\) lays either on the same side of the great circle containing \(e_1p_2\) as \(p_3\), or on the same side of the great circle containing \(e_1p_1\) as \(e_2\). Without losing the generality we can assume that the first of these possibilites takes place.

figure a

Let vu be points of \(\textrm{bd}(V)\) such that |vu| is equal to \(\textrm{diam}(V)\). By Proposition 4.1 of Ref. [5] we know that v and u are vertices of V and \(\{ v,u\}\) = \(\{ v_i, v_j\}\) for some ij which differ by \(\frac{n-1}{2}\) or \(\frac{n+1}{2}\) modulo n. Let t and s be points of \(\textrm{bd}(V)\) such that vt and su are strings of lunes of thickness \(\Delta (V)\) containing V. Observe that exactly one of the following statements holds true: t, the midpoint of tu and u lay on the boundary of V in this order, or s, the midpoint of sv and v lay on the boundary of V in this order. We may assume without losing the generality, that the first of these statements holds true.

Clearly, there exists i such that points \(e_i\), v and \(e_{i+1}\) lay on the boundary of V in this order. We do not lose the generality assuming that \(i=1\), i.e. that \(e_1\), v and \(e_2\) are in this order. We do not exclude the possibility that \(e_1=v\), but we do not allow the possibility that \(e_2=v\).

By Corollary 3.8 from Ref. [5] the string of every lune of thickness \(\Delta (V)\) containing V halves perimeter of V. Thus strings of every pair of lunes of thickness \(\Delta (C)\) containing C intersect. Hence we conclude that \(b_1\), t and \(b_2\) lay on the boundary of V in this order. Observe that there exists \(k\ge 2\) such that t does not belong to \(B_{\Delta (V)}(e_k)\), otherwise \(t\in V_{\Delta (V)}^{\circ }\). Clearly, such \(e_k\) lays on the opossite side of the great circle containing \(e_1p_2\) than t.

Since \(|te_k|>\Delta (V)\), there exists a point x such that \(x\in te_k\) and \(|tx|=\Delta (V)\). The triangle vxt is isosceles with \(|tv|=|tx|\). Consider the great circle containing t and the midpoint m of the arc vx. Clearly it is the set of all points of \(S^2\) equally distanced from v and x. But since \(|\angle mtu| = |\angle mtv| + |\angle vtu|= |\angle mtv| + \frac{\pi }{2}< \pi \), we conclude that u lays on the same side of the great circle containing mt as v. Therefore \(|ux|>|uv| = \textrm{diam}(V)\). But this is impossible, because \(x\in V\) and there does not exist an arc inside V longer than \(\textrm{diam}(V)\). The obtained contradiction ends the proof. \(\square \)