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

Width deviation of convex polygons

Shigeki Akiyama and Teturo Kamae Institute of Mathematics, University of Tsukuba (akiyama@math.tsukuba.ac.jp)Advanced Mathematical Institute, Osaka City University (kamae@apost.plala.or.jp)
Abstract

We consider the width XT(ω)X_{T}(\omega) of a convex nn-gon TT in the plane along the random direction ω/2π\omega\in\mathbb{R}/2\pi\mathbb{Z} and study its deviation rate:

δ(XT)=𝔼(XT2)𝔼(XT)2𝔼(XT).\delta(X_{T})=\frac{\sqrt{\mathbb{E}(X^{2}_{T})-\mathbb{E}(X_{T})^{2}}}{\mathbb{E}(X_{T})}.

We prove that the maximum is attained if and only if TT degenerates to a 22-gon. Let n2n\geq 2 be an integer which is not a power of 22. We show that

π4ntan(π2n)+π28n2sin2(π2n)1\sqrt{\frac{\pi}{4n\tan(\frac{\pi}{2n})}+\frac{\pi^{2}}{8n^{2}\sin^{2}(\frac{\pi}{2n})}-1}

is the minimum of δ(XT)\delta(X_{T}) among all nn-gons and determine completely the shapes of TT’s which attain this minimum. They are characterized as polygonal approximations of equi-Reuleaux bodies, found and studied by K. Reinhardt [9]. In particular, if nn is odd, then the regular nn-gon is one of the minimum shapes. When nn is even, we see that regular nn-gon is far from optimal. We also observe an unexpected property of the deviation rate on the truncation of the regular triangle.

1 Introduction

The width of an image along various directions is basic information in the Image Processing Technique. We are interested in the deviation of the widths of compact convex sets in the plane along a random direction. Although it is an important and useful quantity both practically and theoretically, as long as we know, there is no serious study on it.

Xα1(ω)X_{\alpha_{1}}(\omega)Xα2(ω)X_{\alpha_{2}}(\omega)α1\alpha_{1}α2\alpha_{2}TTXT(ω)X_{T}(\omega)ω\omega
Figure 1: XT(ω)X_{T}(\omega)

Let Ω=/2π=(π,π]\Omega=\mathbb{R}/2\pi\mathbb{Z}=(-\pi,\pi] and \mathbb{P} be the normalized Lebesgue measure on Ω\Omega. Consider the probability space (Ω,)(\Omega,\mathbb{P}); we write Ω\Omega for short. Let a compact convex set TT in the complex plane \mathbb{C} be given. We identify TT with its boundary. We discuss the length of the orthogonally projected shadow of TT by the light from a random direction ω\omega, say XT(ω)X_{T}(\omega). It is also interpreted as the width of TT along the orthogonal direction of ω\omega, that is, ω±π2\omega\pm\frac{\pi}{2}.

We are interested in the uniformity (or its contrary) of the deviation of XT(ω)X_{T}(\omega) with respect to ωΩ\omega\in\Omega. For this purpose, we consider the deviation rate of the random variable XTX_{T}, say δ(XT)\delta(X_{T}), defined by

δ(XT)=𝔼(XT2)𝔼(XT)2𝔼(XT),\delta(X_{T})=\frac{\sqrt{\mathbb{E}(X^{2}_{T})-\mathbb{E}(X_{T})^{2}}}{\mathbb{E}(X_{T})}, (1)

where 𝔼()\mathbb{E}(~{}) is the expectation of the random variables. Clearly, δ(XT)\delta(X_{T}) is invariant among the similarity images of TT. It is also clear that δ(XT)=0\delta(X_{T})=0 if and only if TT is a closed curve of constant width (c.f. [5]). Hence, δ(XT)\delta(X_{T}) measures the degree how far TT is from convex bodies of constant width. For the ellipse T={(x,y);(x2/a2)+(y2/b2)=1}T=\{(x,y);~{}(x^{2}/a^{2})+(y^{2}/b^{2})=1\} with the perimeter LL, it is not difficult to see that

δ(XT)=2π2(a2+b2)L2L.\delta(X_{T})=\frac{\sqrt{2\pi^{2}(a^{2}+b^{2})-L^{2}}}{L}.

In this paper, we study the deviation rate of the convex nn-gons TT with n2n\geq 2. We decompose XTX_{T} as the sum of random variables coming from its edges, say XT=j=1nXαjX_{T}=\sum_{j=1}^{n}X_{\alpha_{j}}. Throughout this paper, the branch of the argument of a complex number is chosen to be (π,π](-\pi,\pi]. For a complex number α\alpha\in\mathbb{C} with arg(α)=θ\arg(\alpha)=\theta, define a random variable XαX_{\alpha} on Ω\Omega by

Xα(ω)=|α|sin(θω)+,X_{\alpha}(\omega)=|\alpha|\sin(\theta-\omega)_{+},

where x+=max{x,0}x_{+}=\max\{x,0\}. In other words, Xα(ω)X_{\alpha}(\omega) is the length of the orthogonally projected shadow of the right side of the vector Oα\vec{O\alpha} by the light from the ω\omega direction (the left side of the vector is permeable and makes no shadow).

Let α1,α2,,αn(n2)\alpha_{1},\alpha_{2},\cdots,\alpha_{n}~{}(n\geq 2) be a sequence of distinct complex numbers. They are arranged in the counter clockwise order forming a convex nn-gon if and only if

()arg(α2α1)arg(α3α2)arg(αn+1αn)arg(α2α1)+2π(mod2π),(*)~{}\arg(\alpha_{2}-\alpha_{1})\leq\arg(\alpha_{3}-\alpha_{2})\leq\cdots\\ \leq\arg(\alpha_{n+1}-\alpha_{n})\leq\arg(\alpha_{2}-\alpha_{1})+2\pi~{}~{}({\rm mod}~{}2\pi),

where we always consider the suffix jj related to the nn-gon TT in modulo nn, so that αn+1=α1,α0=αn\alpha_{n+1}=\alpha_{1},~{}\alpha_{0}=\alpha_{n}, etc. In this case, the convex nn-gon with vertices α1,,αn\alpha_{1},\cdots,\alpha_{n} is denoted by T=T(α1,,αn)T=T(\alpha_{1},\cdots,\alpha_{n}).

The above TT is non-degenerate (i.e. all the vertices are extremal points) if and only if “<<” holds everywhere in the above inequalities. If this is not the case, then we identify T=T(α1,,αn)T=T(\alpha_{1},\cdots,\alpha_{n}) with T(α1,,αm)T(\alpha^{\prime}_{1},\cdots,\alpha^{\prime}_{m}), where {α1,,αm}\{\alpha^{\prime}_{1},\cdots,\alpha^{\prime}_{m}\} are the set of extremal points among {α1,,αn}\{\alpha_{1},\cdots,\alpha_{n}\} arranged in the counter clockwise order.

Define a random variable XTX_{T} by

XT(ω)=j=1nXαj+1αj(ω),X_{T}(\omega)=\sum_{j=1}^{n}X_{\alpha_{j+1}-\alpha_{j}}(\omega),

which agrees with the former explanation as to the length of the shadow of TT (see Figure 1).

The set of convex nn-gons T=T(α1,,αn)T=T(\alpha_{1},\cdots,\alpha_{n}) can be identified with the sequence of nn distinct complex numbers (α1,,αn)(\alpha_{1},\cdots,\alpha_{n}) satisfying ()(*). We denote the space consisting of these (α1,,αn)(\alpha_{1},\cdots,\alpha_{n}) by Θn\Theta_{n}. We consider the usual structures coming from n\mathbb{C}^{n} on Θn\Theta_{n}. Then, k=2n1Θk\cup_{k=2}^{n-1}\Theta_{k} is considered as the boundary of Θn\Theta_{n} in the above sense.

Denote the factor space of Θn\Theta_{n} divided by the similarity equivalence by Θ~n\tilde{\Theta}_{n}. Then it is a compact space. Most of our notions like the regular nn-gon are notions in Θ~n\tilde{\Theta}_{n} rather than Θn\Theta_{n}. The deviation rate δ(XT)\delta(X_{T}) can be considered as a continuous and piecewise smooth functional on Θ~n\tilde{\Theta}_{n}. Hence, it has the minimum and the maximum in Θn\Theta_{n}.

It turned out that the solution of the minimization problem of δ(XT)\delta(X_{T}) is related to the convex bodies of constant width. A Reuleaux body111In literature, it is often referred to Reuleaux ”polygon”, though its boundary is not linear. In this paper, we use the word ”body” to distinguish it from a genuine polygon. is a convex body of constant width whose boundary consists of a finite number of circular arcs with the center in it and the radius equal to the width.

The reader can consult [5, 6], but for the self-containedness, we review the Reuleaux body here briefly.

Let DD be a Reuleaux body with width rr. Then, C=DC=\partial D consists of a finite number of circular arcs, say C1,,CpC_{1},\cdots,C_{p}, of radius rr. The endpoints of the circular arcs are called essential if the 2 neighboring circular arcs have different centers, and hence cannot be joined into a single circular arc.

We assume that the circular arcs C1,,CpC_{1},\cdots,C_{p} are arranged in the counter clockwise order and all the endpoints are essential. Let A1,A2A_{1},A_{2} be the endpoints of C1C_{1}; A2,A3A_{2},A_{3} be the endpoints of C2C_{2}, \cdots; Ap,Ap+1A_{p},A_{p+1} be the endpoints of CpC_{p} (the suffices of CC or AA are considered modulo pp so that Ap+1=A1A_{p+1}=A_{1}). Note that a center of any one of the circular arcs is an essential endpoint of some other circular arcs on CC, since otherwise, we have 2 points in CC with a distance larger than rr. Conversely any endpoint is a center of some circular arc since there must be the same number of circular arcs and centers.

Hence, the center of C1C_{1} is one of A3,,ApA_{3},\cdots,A_{p}. Let it be AkA_{k}. Since DD is strictly convex, the tangent vector of CC at cCc\in C to the counter clockwise direction rotates counter clockwise as cc moves. The center of the circular arc containing cCc\in C is the other intersection of the normal line to the tangent vector at cc with CC. This intersection moves at the endpoints of the circular arcs to the counter clockwise direction since the normal lines from the same point rotate to the counter clockwise direction. Hence, the center of this circular arc is the next end point to that before. This means that the center of C1C_{1} is AkA_{k}, the center of C2C_{2} is Ak+1A_{k+1}, \cdots.

On the other hand, the center of the circular arc CkC_{k} must be A2A_{2} since otherwise, either it is A1A_{1} or there is l1,2l\neq 1,2 such that AkAl¯=r\overline{A_{k}A_{l}}=r. The latter is impossible since if so, by the convexity and the assumption on CC, the circular arc C1C_{1} can be extended to AlA_{l} contradicting the assumption that both A1A_{1} and A2A_{2} are essential end points. The former is impossible since if so, then we have a contradiction that A2Ak+1¯>r\overline{A_{2}A_{k+1}}>r. Thus, we have

A2=the center of Ck=A2k1,A_{2}=\mbox{the center of }C_{k}=A_{2k-1},

and hence, 22k1(modp)2\equiv 2k-1~{}(\mbox{mod}~{}p). This implies that pp is odd.

Consider the diagonals connecting 2 points in CC, one of the center and an interior point of the circular arc centered by it. Any of 2 diagonals with different centers always intersect just at one point. Let Λj\Lambda_{j} be the set of angles from AjA_{j} to the points in the circular arc centered by AjA_{j}. We always consider the angles in modulo 2π2\pi. Then, by the above argument, Λj\Lambda_{j} and Λl\Lambda_{l} with jlj\neq l are essentially disjoint. The same thing holds for Λj\Lambda_{j} and Λl+π\Lambda_{l}+\pi. Let Λ=j=1pΛj\Lambda=\cup_{j=1}^{p}\Lambda_{j}. Then, the union is essentially disjoint. Moreover, Λ\Lambda and Λ+π\Lambda+\pi is also essentially disjoint. It also holds that Λ(Λ+π)=(π,π]\Lambda\cup(\Lambda+\pi)=(-\pi,\pi]. This is because the Lebesgue measure of Λ\Lambda is π\pi, since the integration of the exterior angle of CC is 2π2\pi, which counts the angles in Λj(j=1,,p)\Lambda_{j}~{}(j=1,\cdots,p) twice, once at CjC_{j}, once at its center, where the exterior angle jumps just the amount of angles in Λj\Lambda_{j}.

Thus, it holds that the circular arcs C1,,CpC_{1},\cdots,C_{p} can be embedded by parallel translations by v1,,vp\vec{v}_{1},\cdots,\vec{v}_{p}\in\mathbb{C} into a circle 𝕊\mathbb{S} of radius rr so that

Λ~:=j=1p(Cj+vj)𝕊\tilde{\Lambda}:=\cup_{j=1}^{p}(C_{j}+\vec{v}_{j})\subset\mathbb{S}

satisfies that

#((Cj+vj)(Ck+vk))<for anyjk,\displaystyle\#((C_{j}+\vec{v}_{j})\cap(C_{k}+\vec{v}_{k}))<\infty~{}\mbox{for any}~{}j\neq k,
#(Λ~RπΛ~)<andΛ~RπΛ~=𝕊,\displaystyle\#(\tilde{\Lambda}\cap R_{\pi}\tilde{\Lambda})<\infty~{}\mbox{and}~{}\tilde{\Lambda}\cup R_{\pi}\tilde{\Lambda}=\mathbb{S}, (2)

where RπR_{\pi} is the rotation on 𝕊\mathbb{S} by angle π\pi, and “#\#” implies the number of elements in a set.

We write a Reuleaux pp-body to indicate the number pp of circular arcs with different centers. In particular, a Reuleaux pp-body is called regular if all the circular arcs have the same length. A regular Reuleaux 3-body is well known as “Reuleaux triangle”.

A Reinhardt nn-gon ([3], [9]) is an equilateral convex nn-gon that can be inscribed in a Reuleaux body, containing all the essential endpoints of the circular arcs. It is referred to as a Reinhalt polygon if nn is not necessarily specified. If the lengths of the circular arcs of a Reuleaux pp-body have ratio n1:n2::npn_{1}:n_{2}:\dots:n_{p} with integer nin_{i}’s, we can divide the circular arcs into ni(i=1,,p)n_{i}\ (i=1,\dots,p) parts of equal length and take the convex-hull of all the division points to get a Reinhardt nn-gon, where n=n1++npn=n_{1}+\cdots+n_{p}, which is specially called a Reinhardt (n1,,np)(n_{1},\cdots,n_{p})-gon.

In particular, let p3p\geq 3 be an odd integer and take a regular Reuleaux pp-body DD. We divide all the circular arcs of DD into qq parts of equal length. Then, the convex hull of all the division points forms a Reinhardt pqpq-gon, which is specially a Reinhardt (q,,qp times)(\underbrace{q,\cdots,q}_{\mbox{$p$ times}})-gon, see Figure 2. Denote qp=(q,,qp times)q^{p}=(\underbrace{q,\cdots,q}_{\mbox{$p$ times}}) for short. Note that Reinhardt 1p1^{p}-gon is the regular pp-gon.

Refer to caption
Figure 2: Reinhardt i3i^{3}-gon for i=1,2,3i=1,2,3

A Reinhardt polygon gives the solution of three optimization problems on convex nn-gons when nn is not a power of 22:

Problem 1.

Maximize the perimeter for a fixed diameter [9],

Problem 2.

Maximize the width for a fixed diameter [2],

Problem 3.

Maximize the width for a fixed perimeter [1].

Given a cyclic integer vector (n1,n2,,np)(n_{1},n_{2},\dots,n_{p}), there exists a unique way to construct a Reinhardt (n1,n2,,np)(n_{1},n_{2},\dots,n_{p})-gon when this is possible. We reproduce a necessary and sufficient condition of Reinhardt [9] that a given integer cyclic vector (n1,n2,,np)(n_{1},n_{2},\dots,n_{p}) forms a Reinhardt (n1,n2,,np)(n_{1},n_{2},\dots,n_{p})-gon as a byproduct of our result (Theorem 6 and 7). In particular, when nn is not a power of 22, we have a Reinhardt (n/p)p(n/p)^{p}-gon for any odd divisor p>1p>1. Many interesting properties on Reinhardt polygons are discussed in [6, 7, 3, 4]. When nn is odd, the regular nn-gon is one of the Reinhardt polygons but it is not unique when nn is a composite. Reinhardt nn-gon is unique if and only if n=pn=p or 2p2p with a prime pp. A Reinhardt polygon may have no symmetry at all (see Figures in [3]).

In this paper, we prove the following results.
(1) The deviation rate for the compact convex sets TT attains the maximum if and only if TT is the 2-gon (Theorem 2).
(2) For the integer n2n\geq 2 which is not a power of 22, the minimum of δ(XT)\delta(X_{T}) for TΘnT\in\Theta_{n} is

νn:=π4ntan(π2n)+π28n2sin2(π2n)1.\nu_{n}:=\sqrt{\frac{\pi}{4n\tan(\frac{\pi}{2n})}+\frac{\pi^{2}}{8n^{2}\sin^{2}(\frac{\pi}{2n})}-1}. (3)

Theorem 6 gives a complete characterization of the shapes that attain the minimum value νn\nu_{n}. As a consequence, we show that the minimum shape is nothing but a Reinhardt nn-gon (Theorem 7).

The minimum values for odd n3n\geq 3 is strictly decreasing in nn.
(3) For even m4m\geq 4, regular mm-gon is far from the minimum shape. Let n(<m)n(<m) be the odd number such that either n=m2n=\frac{m}{2} or n=m2+1n=\frac{m}{2}+1. If n=m2n=\frac{m}{2}, then δ(XTm)=δ(XTn)\delta(X_{T_{m}})=\delta(X_{T_{n}}) holds, and if n=m2+1n=\frac{m}{2}+1, then δ(XTm)>δ(XTn)\delta(X_{T_{m}})>\delta(X_{T_{n}}) holds, where Tm,TnT_{m},~{}T_{n} are the regular mm-gon and nn-gon, respectively (Theorem 4).

Refer to caption
Figure 3: parallel truncation (left) and non-parallel truncation (right)

(4) Let TT be the regular triangle. We call a truncation of it a quadrangle obtained by cutting off a vertex by a line near it. It is called a parallel truncation if the line is parallel to the opposite side, otherwise a non-parallel truncation (Figure 3). We prove that a parallel truncation of the regular triangle increases the deviation rate, while a non-parallel truncation decreases it (Theorem 8).

2 Maximum of the deviation rate

Theorem 1.

For any α,β\alpha,\beta\in\mathbb{C} with η=arg(β/α)(π,π]\eta=\arg(\beta/\alpha)\in(-\pi,\pi], the following statements hold:
(i)(i) 𝔼(Xα)=|α|/π\mathbb{E}(X_{\alpha})=|\alpha|/\pi,
(ii)(ii) 𝔼(XαXβ)=14π|α||β|V(η)\mathbb{E}(X_{\alpha}X_{\beta})=\frac{1}{4\pi}|\alpha||\beta|V(\eta), where VV is a periodic function of period 2π2\pi such that

V(x)=(π|x|)cosx+sin|x|(π<xπ).V(x)=(\pi-|x|)\cos x+\sin|x|~{}~{}(-\pi<x\leq\pi).
Refer to caption
Figure 4: Function VV

Proof (i)(i) Let θ=arg(α)\theta=\arg(\alpha). Then, we have

𝔼(Xα)=12π02π|α|sin(θξ)+dξ=|α|2π0πsinξdξ=|α|π.\mathbb{E}(X_{\alpha})=\frac{1}{2\pi}\int_{0}^{2\pi}|\alpha|\sin(\theta-\xi)_{+}d\xi=\frac{|\alpha|}{2\pi}\int_{0}^{\pi}\sin\xi d\xi=\frac{|\alpha|}{\pi}.

(ii)(ii) We have

𝔼(XαXβ)=12π02π|α||β|sin(θξ)+sin(θ+ηξ)+dξ\displaystyle\mathbb{E}(X_{\alpha}X_{\beta})=\frac{1}{2\pi}\int_{0}^{2\pi}|\alpha||\beta|\sin(\theta-\xi)_{+}\sin(\theta+\eta-\xi)_{+}d\xi
=12π|α||β|[0,π][η,πη]sinξsin(ξ+η)𝑑ξ\displaystyle=\frac{1}{2\pi}|\alpha||\beta|\int_{[0,\pi]\cap[-\eta,\pi-\eta]}\sin\xi\sin(\xi+\eta)d\xi
=14π|α||β|[0,π][η,πη](cosηcos(2ξ+η))𝑑ξ.\displaystyle=\frac{1}{4\pi}|\alpha||\beta|\int_{[0,\pi]\cap[-\eta,\pi-\eta]}(\cos\eta-\cos(2\xi+\eta))d\xi.

Therefore, if η<0\eta<0, then

𝔼(XαXβ)=14π|α||β|ηπ(cosηcos(2ξ+η))𝑑ξ\displaystyle\mathbb{E}(X_{\alpha}X_{\beta})=\frac{1}{4\pi}|\alpha||\beta|\int_{-\eta}^{\pi}(\cos\eta-\cos(2\xi+\eta))d\xi
=14π|α||β|((π+η)cosηsinη)=14π|α||β|V(η).\displaystyle=\frac{1}{4\pi}|\alpha||\beta|((\pi+\eta)\cos\eta-\sin\eta)=\frac{1}{4\pi}|\alpha||\beta|V(\eta).

If η0\eta\geq 0, then

𝔼(XαXβ)=14π|α||β|0πη(cosηcos(2ξ+η))𝑑ξ\displaystyle\mathbb{E}(X_{\alpha}X_{\beta})=\frac{1}{4\pi}|\alpha||\beta|\int_{0}^{\pi-\eta}(\cos\eta-\cos(2\xi+\eta))d\xi
=14π|α||β|((πη)cosη+sinη)=14π|α||β|V(η).\displaystyle=\frac{1}{4\pi}|\alpha||\beta|((\pi-\eta)\cos\eta+\sin\eta)=\frac{1}{4\pi}|\alpha||\beta|V(\eta).

\Box

Theorem 2.

The maximum of δ(XT)\delta(X_{T}) for the compact convex sets TT is attained by a 22-gon.

Proof Since all 2-gons are similar, they have the same δ\delta-value. Let U=T(0,1)U=T(0,1) be a 2-gon. Then by Theorem 1, we have

𝔼(XU)=1π+1π=2π\displaystyle\mathbb{E}(X_{U})=\frac{1}{\pi}+\frac{1}{\pi}=\frac{2}{\pi}
𝔼(XU2)=14π(2V(0)+2V(π))=12.\displaystyle\mathbb{E}(X_{U}^{2})=\frac{1}{4\pi}(2V(0)+2V(\pi))=\frac{1}{2}.

Hence, δ(XU)=(π2/8)1\delta(X_{U})=\sqrt{(\pi^{2}/8)-1}.

Take any convex polygon T=T(α1,,αn)T=T(\alpha_{1},\cdots,\alpha_{n}). We prove that δ(XT)(π2/8)1\delta(X_{T})\leq\sqrt{(\pi^{2}/8)-1}, and that the equality holds only when TT is 2-gon. Let

βj=αj+1αj=rje𝐢θj(j=1,,n)\beta_{j}=\alpha_{j+1}-\alpha_{j}=r_{j}e^{\mathbf{i}\theta_{j}}~{}(j=1,\cdots,n)

and let θjk(π,π]\theta_{jk}\in(-\pi,\pi] satisfies that

θjkθjθk(mod2π)(j,k=1,,n).\theta_{jk}\equiv\theta_{j}-\theta_{k}~{}({\rm mod}~{}2\pi)~{}(j,k=1,\cdots,n).

It is sufficient to prove that 𝔼(XT2)𝔼(XT)2π28\frac{\mathbb{E}(X_{T}^{2})}{\mathbb{E}(X_{T})^{2}}\leq\frac{\pi^{2}}{8}. Hence, it is sufficient to prove that I:=π32𝔼(XT)24π𝔼(XT2)0I:=\frac{\pi^{3}}{2}\mathbb{E}(X_{T})^{2}-4\pi\mathbb{E}(X_{T}^{2})\geq 0. Then by Theorem 1, we have

I=π32𝔼(XT)24π𝔼(XT2)=π32(j=1n𝔼(Xj))24πj,k=1n𝔼(XjXk)\displaystyle I=\frac{\pi^{3}}{2}\mathbb{E}(X_{T})^{2}-4\pi\mathbb{E}(X_{T}^{2})=\frac{\pi^{3}}{2}\left(\sum_{j=1}^{n}\mathbb{E}(X_{j})\right)^{2}-4\pi\sum_{j,k=1}^{n}\mathbb{E}(X_{j}X_{k})
=π32(j=1nrjπ)2j,k=1nrjrkV(θjθk)=j,k=1nrjrk(π2V(θjθk))\displaystyle=\frac{\pi^{3}}{2}\left(\sum_{j=1}^{n}\frac{r_{j}}{\pi}\right)^{2}-\sum_{j,k=1}^{n}r_{j}r_{k}V(\theta_{j}-\theta_{k})=\sum_{j,k=1}^{n}r_{j}r_{k}\left(\frac{\pi}{2}-V(\theta_{j}-\theta_{k})\right)
=j,k=1nrjrk(π2(π|θjk|)cosθjk|sinθjk|).(#)\displaystyle=\sum_{j,k=1}^{n}r_{j}r_{k}\left(\frac{\pi}{2}-(\pi-|\theta_{jk}|)\cos\theta_{jk}-|\sin\theta_{jk}|\right).~{}~{}~{}~{}(\#)

Since rjrkcosθjkr_{j}r_{k}\cos\theta_{jk} is the inner product between Oβj\vec{O\beta_{j}} and Oβk\vec{O\beta_{k}}, we have

j,k=1nrjrkcosθjk=(j=1kOβj,j=1kOβj)=(0,0)=0.\sum_{j,k=1}^{n}r_{j}r_{k}\cos\theta_{jk}=\left(\sum_{j=1}^{k}\vec{O\beta_{j}},~{}\sum_{j=1}^{k}\vec{O\beta_{j}}\right)=(\vec{0},\vec{0})=0.

This implies that in the above equality (#)(\#), the constant in the coefficient of cosθjk\cos\theta_{jk} can be changed anyway keeping the equality. Hence, we have

I=j,k=1nrjrk(π2(π2|θjk|)cosθjk|sinθjk|).I=\sum_{j,k=1}^{n}r_{j}r_{k}\left(\frac{\pi}{2}-\left(\frac{\pi}{2}-|\theta_{jk}|\right)\cos\theta_{jk}-|\sin\theta_{jk}|\right).

Thus, to prove I0I\geq 0, it is sufficient to prove that

(π2|x|)cosx+|sinx|π2\left(\frac{\pi}{2}-|x|\right)\cos x+|\sin x|\leq\frac{\pi}{2}

for any x(π,π]x\in(-\pi,\pi], which can be verified easily. Moreover, the equality holds if and only if x=0x=0 or π\pi, which implies that TT is 2-gon.

For a general compact convex set SS with C=SC=\partial S, replacing the above II by

I=C×C(π2(π2θ(u,v))cosθ(u,v)|sinθ(u,v)|)𝑑u𝑑v,I=\int\int_{C\times C}\left(\frac{\pi}{2}-\left(\frac{\pi}{2}-\theta(u,v)\right)\cos\theta(u,v)-|\sin\theta(u,v)|\right)dudv,

where θ(u,v)\theta(u,v) is the angle between the tangent lines of CC at uu and vv, we have the same statement, which completes the proof. \Box

3 Symmetrization and asymmetrization of nn-gons

Let

={α;(α)>0,or,(α)=0and(α)0}\mathcal{H}=\{\alpha\in\mathbb{C};~{}\Im(\alpha)>0,~{}\mbox{or,}~{}\Im(\alpha)=0~{}\mbox{and}~{}\Re(\alpha)\geq 0\}

be the upper half plane endowed with the quotient topology of \mathbb{C} by identifying zz and z-z. For α\alpha\in\mathbb{C}, we define

ι(α)={α(α)α(α).\iota(\alpha)=\left\{\begin{array}[]{cc}\alpha&(\alpha\in\mathcal{H})\\ -\alpha&(\alpha\notin\mathcal{H})\end{array}\right..

For a finite set of nonzero complex numbers S={α1,,αn}S=\{\alpha_{1},\cdots,\alpha_{n}\}, define ι(S)\iota(S)\subset\mathcal{H} by

ι(S)=Abbreviation{ι(α1),,ι(αn)},\iota(S)=\mbox{Abbreviation}\{\iota(\alpha_{1}),\cdots,\iota(\alpha_{n})\},

where Abbreviation{α1,,αn}\mbox{Abbreviation}\{\alpha_{1}^{\prime},\cdots,\alpha_{n}^{\prime}\} is the set of complex numbers obtained by replacing any of αj,αk\alpha_{j}^{\prime},\alpha_{k}^{\prime} with arg(αj)=arg(αk)\arg(\alpha_{j}^{\prime})=\arg(\alpha_{k}^{\prime}) by αj+αk\alpha_{j}^{\prime}+\alpha_{k}^{\prime}.

A nonzero element in \mathcal{H} is sometimes called a pre-edge. A sequence of pre-edges β1,,βm\beta_{1},\cdots,\beta_{m}\in\mathcal{H}, say =(β1,,βm)\mathcal{B}=\mathcal{B}(\beta_{1},\cdots,\beta_{m}) is called a pre-edge bundle (of size mm) if

0arg(β1)<arg(β2)<<arg(βm)<π.0\leq\arg(\beta_{1})<\arg(\beta_{2})<\cdots<\arg(\beta_{m})<\pi.

Denote by Ξm\Xi_{m} the set of pre-edge bundles of size mm.

For a convex nn-gon T=T(α1,,αn)T=T(\alpha_{1},\cdots,\alpha_{n}), we define its asymmetrization ι(T)\iota(T) as the pre-edge bundle =(β1,,βm)\mathcal{B}=\mathcal{B}(\beta_{1},\cdots,\beta_{m}) such that

ι({α2α1,α3α2,,αn+1αn})={β1,β2,,βm}.\iota(\{\alpha_{2}-\alpha_{1},~{}\alpha_{3}-\alpha_{2},~{}\cdots,~{}\alpha_{n+1}-\alpha_{n}\})=\{\beta_{1},\beta_{2},\cdots,\beta_{m}\}.

In this case, TT is called a realization of \mathcal{B}. Let T=T(α1,,αn)T=T(\alpha_{1},\cdots,\alpha_{n}) be a convex nn-gon and =(β1,,βm)\mathcal{B}=\mathcal{B}(\beta_{1},\cdots,\beta_{m}) be its asymmetrization. Then,

U=\displaystyle U=
T(γ,γ+12β1,,γ+12(β1++βm),γ+12(β2++βm),,γ+12βm)\displaystyle T\left(\gamma,\gamma+\frac{1}{2}\beta_{1},\cdots,\gamma+\frac{1}{2}(\beta_{1}+\cdots+\beta_{m}),\gamma+\frac{1}{2}(\beta_{2}+\cdots+\beta_{m}),\cdots,\gamma+\frac{1}{2}\beta_{m}\right)

is another realization of \mathcal{B} than TT, where γ=14(β1++βm)\gamma=-\frac{1}{4}(\beta_{1}+\cdots+\beta_{m}). We call UU the symmetrization of TT (or \mathcal{B}).

In this case, UU is symmetric, that is, U=T(γ1,,γ2k)U=T(\gamma_{1},\cdots,\gamma_{2k}) with even size 2k2k and

γk+1=γ1,γk+2=γ2,,γ2k=γk\gamma_{k+1}=-\gamma_{1},~{}\gamma_{k+2}=-\gamma_{2},~{}\cdots,~{}\gamma_{2k}=-\gamma_{k}

holds. For a symmetric U=T(γ1,,γ2k)U=T(\gamma_{1},\cdots,\gamma_{2k}), it holds that

XU(ω)=2|γj|sin(arg(γj)ω)for anyωwitharg(γj1γj)ωarg(γjγj+1).X_{U}(\omega)=2|\gamma_{j}|\sin(\arg(\gamma_{j})-\omega)\\ \mbox{for any}~{}\omega~{}\mbox{with}~{}\arg(\gamma_{j-1}-\gamma_{j})\leq\omega\leq\arg(\gamma_{j}-\gamma_{j+1}). (4)

This representation of XUX_{U} is called the diagonal representation.

1-1𝐢-\mathbf{i}111+𝐢1+\mathbf{i}𝐢\mathbf{i}112+2𝐢2+2\mathbf{i}𝐢\mathbf{i}1+𝐢-1+\mathbf{i}12𝐢-\frac{1}{2}-\mathbf{i}𝐢-\mathbf{i}112𝐢-1-\frac{1}{2}\mathbf{i}1-1𝐢\mathbf{i}12+𝐢\frac{1}{2}+\mathbf{i}1+12𝐢1+\frac{1}{2}\mathbf{i}11
Figure 5: TT (left), \mathcal{B} (center) and UU (right)
Example 1.

Let T=T(𝐢,1,1+𝐢,𝐢,1)T=T(-\mathbf{i},1,1+\mathbf{i},\mathbf{i},-1). Then, its asymmetrization is

=(1,2+2𝐢,𝐢,1+𝐢)\mathcal{B}=\mathcal{B}(1,2+2\mathbf{i},\mathbf{i},-1+\mathbf{i})

and its symmetrization is

U=T(12𝐢,𝐢,1,1+12𝐢,12+𝐢,𝐢,1,112𝐢).U=T\left(-\frac{1}{2}-\mathbf{i},-\mathbf{i},1,1+\frac{1}{2}\mathbf{i},\frac{1}{2}+\mathbf{i},\mathbf{i},-1,-1-\frac{1}{2}\mathbf{i}\right).

See Figure 3.

Let β\beta be a pre-edge with arg(β)=θ\arg(\beta)=\theta. Define a random variable X~β\tilde{X}_{\beta} as

X~β(ω)=|β|2|sin(θω)|,\tilde{X}_{\beta}(\omega)=\frac{|\beta|}{2}|\sin(\theta-\omega)|,

and for a pre-edge bundle =(β1,,βm)\mathcal{B}=\mathcal{B}(\beta_{1},\cdots,\beta_{m}), let X~=j=1mX~βj\tilde{X}_{\mathcal{B}}=\sum_{j=1}^{m}\tilde{X}_{\beta_{j}}.

Theorem 3.

If TT is a realization of \mathcal{B}, then we have XT=X~X_{T}=\tilde{X}_{\mathcal{B}}.

Proof Since T=T(α1,,αn)T=T(\alpha_{1},\cdots,\alpha_{n}) is a convex polygon, we have XT(ω)=XT(ω+π)X_{T}(\omega)=X_{T}(\omega+\pi) for any ωΩ\omega\in\Omega. Hence,

XT(ω)=XT(ω)+XT(ω+π)2=j=1nXαj(ω)+Xαj(ω+π)2\displaystyle X_{T}(\omega)=\frac{X_{T}(\omega)+X_{T}(\omega+\pi)}{2}=\sum_{j=1}^{n}\frac{X_{\alpha_{j}}(\omega)+X_{\alpha_{j}}(\omega+\pi)}{2}
=j=1n|αj|2(sin(θjω)++sin(θjωπ)+)\displaystyle=\sum_{j=1}^{n}\frac{|\alpha_{j}|}{2}(\sin(\theta_{j}-\omega)_{+}+\sin(\theta_{j}-\omega-\pi)_{+})
=j=1n|αj|2(sin(θjω)++(sin(θjω))+)\displaystyle=\sum_{j=1}^{n}\frac{|\alpha_{j}|}{2}(\sin(\theta_{j}-\omega)_{+}+(-\sin(\theta_{j}-\omega))_{+})
=j=1n|αj|2|sin(θjω)|=j=1nX~ι(αj)=j=1mX~βj=X~\displaystyle=\sum_{j=1}^{n}\frac{|\alpha_{j}|}{2}|\sin(\theta_{j}-\omega)|=\sum_{j=1}^{n}\tilde{X}_{\iota(\alpha_{j})}=\sum_{j=1}^{m}\tilde{X}_{\beta_{j}}=\tilde{X}_{\mathcal{B}}

\Box

4 Minimum of deviation rate among the pre-edge bundles of fixed size

A regular nn-gon is denoted by Tn(n=2,3,)T_{n}~{}(n=2,3,\cdots). A pre-edge bundle Rm=(β1,,βm)R_{m}=\mathcal{B}(\beta_{1},\cdots,\beta_{m}) is said to be regular if

|β1|==|βm|andarg(βj+1)arg(βj)π/m(modπ)(j=1,,m).|\beta_{1}|=\cdots=|\beta_{m}|~{}\mbox{and}~{}\arg(\beta_{j+1})-\arg(\beta_{j})\equiv\pi/m~{}({\rm mod}~{}\pi)~{}(j=1,\cdots,m).
Theorem 4.

(i)(i) It holds that

δ(X~Rm)=νm(see(3))\delta(\tilde{X}_{R_{m}})=\nu_{m}~{}~{}(\mbox{see}~{}(3))

for m=1,2,m=1,2,\cdots, and hence, δ(X~R1)<δ(X~R2)<δ(X~R3)<\delta(\tilde{X}_{R_{1}})<\delta(\tilde{X}_{R_{2}})<\delta(\tilde{X}_{R_{3}})<\cdots.
(ii)(ii) If m2m\geq 2 is even, then let n(<m)n(<m) be the odd number such that either n=m2n=\frac{m}{2} or n=m2+1n=\frac{m}{2}+1. If n=m2n=\frac{m}{2}, then δ(XTm)=δ(XTn)=δ(X~Rn)\delta(X_{T_{m}})=\delta(X_{T_{n}})=\delta(\tilde{X}_{R_{n}}) holds, and if n=m2+1n=\frac{m}{2}+1, then δ(XTm)=δ(X~Rm/2)>δ(X~Rn)=δ(XTn)\delta(X_{T_{m}})=\delta(\tilde{X}_{R_{m/2}})>\delta(\tilde{X}_{R_{n}})=\delta(X_{T_{n}}) holds.

Proof 
(ii)(ii)
holds since both of TnT_{n} and T2nT_{2n} are realizations of RnR_{n} if n=m2n=\frac{m}{2} is odd, and δ(XTn)=δ(XT2n)=δ(XTm)=δ(X~Rn)\delta(X_{T_{n}})=\delta(X_{T_{2n}})=\delta(X_{T_{m}})=\delta(\tilde{X}_{R_{n}}) by Theorem 3. If n=m2+1n=\frac{m}{2}+1, then by the monotonicity in (i)(i), δ(XTn)=δ(X~Rn)<δ(X~Rm/2)\delta(X_{T_{n}})=\delta(\tilde{X}_{R_{n}})<\delta(\tilde{X}_{R_{m/2}}).

By Theorem 3, to prove (i)(i), it is sufficient to prove that

δ(XT2m)=νm(m=1,2,)\delta(X_{T_{2m}})=\nu_{m}~{}~{}(m=1,2,\cdots)

for T2m=T(e𝐢π/2m,e𝐢3π/2m,,e𝐢π(1+2(2m1))/2m)T_{2m}=T(e^{\mathbf{i}\pi/2m},e^{\mathbf{i}3\pi/2m},\cdots,e^{\mathbf{i}\pi(1+2(2m-1))/2m}). Using the diagonal representation (4), we have

𝔼(XT2m)=2m2ππ/2π/2+π/m2sin(π2mω)𝑑ω\displaystyle\mathbb{E}(X_{T_{2m}})=\frac{2m}{2\pi}\int_{-\pi/2}^{-\pi/2+\pi/m}2\sin(\frac{\pi}{2m}-\omega)d\omega
=2mππ/(2m)π/(2m)sin(π2ω)𝑑ω=2mππ/(2m)π/(2m)cosωdω=4mπsinπ2m\displaystyle=\frac{2m}{\pi}\int_{\pi/(2m)}^{-\pi/(2m)}\sin(\frac{\pi}{2}-\omega)d\omega=\frac{2m}{\pi}\int_{-\pi/(2m)}^{\pi/(2m)}\cos\omega d\omega=\frac{4m}{\pi}\sin\frac{\pi}{2m}
𝔼(XT2m2)=2m2ππ/2π/2+π/m4sin2(π2mω)𝑑ω\displaystyle\mathbb{E}(X^{2}_{T_{2m}})=\frac{2m}{2\pi}\int_{-\pi/2}^{-\pi/2+\pi/m}4\sin^{2}(\frac{\pi}{2m}-\omega)d\omega
=4mππ/(2m)π/(2m)sin2(π2ω)𝑑ω=4mππ/(2m)π/(2m)cos2ωdω\displaystyle=\frac{4m}{\pi}\int_{\pi/(2m)}^{-\pi/(2m)}\sin^{2}(\frac{\pi}{2}-\omega)d\omega=\frac{4m}{\pi}\int_{-\pi/(2m)}^{\pi/(2m)}\cos^{2}\omega d\omega
=2mπsinπm+2\displaystyle=\frac{2m}{\pi}\sin\frac{\pi}{m}+2

Hence,

δ(XT2m)=𝔼(XT2m2)𝔼(XT2m)21=πsinπm8msin2π2m+π28m2sin2π2m1\displaystyle\delta(X_{T_{2m}})=\sqrt{\frac{\mathbb{E}(X^{2}_{T_{2m}})}{\mathbb{E}(X_{T_{2m}})^{2}}-1}=\sqrt{\frac{\pi\sin\frac{\pi}{m}}{8m\sin^{2}\frac{\pi}{2m}}+\frac{\pi^{2}}{8m^{2}\sin^{2}\frac{\pi}{2m}}-1}
=π4mtan(π2m)+π28m2sin2(π2m)1.\displaystyle=\sqrt{\frac{\pi}{4m\tan(\frac{\pi}{2m})}+\frac{\pi^{2}}{8m^{2}\sin^{2}(\frac{\pi}{2m})}-1}.

Let x=π2mx=\frac{\pi}{2m} and II be the term inside the root in the above formula. Then, we have

I=x2tanx+x22sin2x1.I=\frac{x}{2\tan x}+\frac{x^{2}}{2\sin^{2}x}-1.

We show that II is an increasing function of x(0,π2]x\in(0,\frac{\pi}{2}]. We have

I(x)=cosxsinxx2sin2x+xsinxx2cosxsin3x\displaystyle I^{\prime}(x)=\frac{\cos x\sin x-x}{2\sin^{2}x}+\frac{x\sin x-x^{2}\cos x}{\sin^{3}x}
=cosxsin2x+xsinx2x2cosx2sin3x\displaystyle=\frac{\cos x\sin^{2}x+x\sin x-2x^{2}\cos x}{2\sin^{3}x}

Since

cosx112x2+124x4andsinxx16x3,\cos x\leq 1-\frac{1}{2}x^{2}+\frac{1}{24}x^{4}~{}~{}\mbox{and}~{}~{}\sin x\geq x-\frac{1}{6}x^{3},
(2sin3x)I(x)=cosxsin2x+xsinx2x2cosx\displaystyle(2\sin^{3}x)I^{\prime}(x)=\cos x\sin^{2}x+x\sin x-2x^{2}\cos x
x(x16x3)(2x2sin2x)(112x2+124x4)\displaystyle\geq x(x-\frac{1}{6}x^{3})-(2x^{2}-\sin^{2}x)(1-\frac{1}{2}x^{2}+\frac{1}{24}x^{4})
x(x16x3)2x2(112x2+124x4)+(x16x3)2(112x2+124x4)\displaystyle\geq x(x-\frac{1}{6}x^{3})-2x^{2}(1-\frac{1}{2}x^{2}+\frac{1}{24}x^{4})+(x-\frac{1}{6}x^{3})^{2}(1-\frac{1}{2}x^{2}+\frac{1}{24}x^{4})
=1172x6136x8+1864x10=172x6(12x22x+11)\displaystyle=\frac{11}{72}x^{6}-\frac{1}{36}x^{8}+\frac{1}{864}x^{10}=\frac{1}{72}x^{6}(12x^{2}-2x+11)

which is positive on x(0,π2]x\in(0,\frac{\pi}{2}]. Thus, I(x)I(x) is strictly increasing in xx, and hence, δ(XTm)\delta(X_{T_{m}}) is strictly decreasing in m=1,2,m=1,2,\cdots.

By a numerical calculation, we have δ(X~Rm)\delta(\tilde{X}_{R_{m}}) as follows.

m|1234567δ|0.483420.097720.041960.023330.014850.010280.00754\begin{array}[]{ccccccccc}m&|&1&2&3&4&5&6&7\\ \hline\cr\delta&|&0.48342&0.09772&0.04196&0.02333&0.01485&0.01028&0.00754\end{array}

\Box

Theorem 5.

The deviation rate δ(X~)\delta(\tilde{X}_{\mathcal{B}}) attains the minimum among Ξm\mathcal{B}\in\Xi_{m} if and only if =Rm\mathcal{B}=R_{m}.

Proof We use the induction on mm. If m=1m=1, the statement is clear since R1R_{1} is essentially the only element in Ξ1\Xi_{1}.

Let m2m\geq 2 and assume that the statement holds for Ξj(j=1,,m1)\Xi_{j}~{}(j=1,\cdots,m-1). The boundary of the closure of Ξm\Xi_{m} consists of j=1m1Ξj\cup_{j=1}^{m-1}\Xi_{j}, and the δ\delta-values there are larger than δ(X~Rm)\delta(\tilde{X}_{R_{m}}) since the assumption of the induction and Theorem 4. Hence, there is 0Ξm\mathcal{B}_{0}\in\Xi_{m} attaining the minimum of δ(X~)\delta(\tilde{X}_{\mathcal{B}}) in Ξm\Xi_{m}. We prove that 0=Rm\mathcal{B}_{0}=R_{m}.

For this purpose, we take the symmetrization T0Θ2mT_{0}\in\Theta_{2m} of 0\mathcal{B}_{0}. Then, δ(XT0)\delta(X_{T_{0}}) is minimum among δ(XT)\delta(X_{T}) for symmetric TΘ2mT\in\Theta_{2m}. This is equivalent to say that κ(XT0)\kappa(X_{T_{0}}) is minimum among κ(XT)\kappa(X_{T}) for symmetric TΘ2mT\in\Theta_{2m}, where κ(XT)=𝔼(XT2)𝔼(XT)2\kappa(X_{T})=\frac{\mathbb{E}(X_{T}^{2})}{\mathbb{E}(X_{T})^{2}}. We’ll conclude from this that T0T_{0} is the regular 2m2m-gon.

Let T0=T(α1,,α2m)T_{0}=T(\alpha_{1},\cdots,\alpha_{2m}). Consider the diagonal representation (4) of XT0X_{T_{0}}. Then, for any j=1,,2mj=1,\cdots,2m, we have

XT0(ω)=2|αj|sin(arg(αj)ω)ifωΩj,X_{T_{0}}(\omega)=2|\alpha_{j}|\sin(\arg(\alpha_{j})-\omega)~{}\mbox{if}~{}\omega\in\Omega_{j},

where

Ωj={ωΩ;arg(αj1αj)<ωarg(αjαj+1)}.\Omega_{j}=\{\omega\in\Omega;~{}\arg(\alpha_{j-1}-\alpha_{j})<\omega\leq\arg(\alpha_{j}-\alpha_{j+1})\}.

For a fixed j=1,,mj=1,\cdots,m and a real number λ\lambda near 0, let

T0λ=T(α1,,(1+λ)αj,αj+1,,(1+λ)αj+m,,α2m).T^{\lambda}_{0}=T(\alpha_{1},\cdots,(1+\lambda)\alpha_{j},\alpha_{j+1},\cdots,(1+\lambda)\alpha_{j+m},\cdots,\alpha_{2m}).

By the minimality, we must have ddλκ(XT0λ)|λ=0=0\frac{d}{d\lambda}\kappa(X_{T_{0}^{\lambda}})|_{\lambda=0}=0.

Let

A=𝔼(XT0),B=𝔼(XT02),aj=𝔼(XT01Ωj),bj=𝔼(XT021Ωj).A=\mathbb{E}(X_{T_{0}}),~{}B=\mathbb{E}(X^{2}_{T_{0}}),~{}a_{j}=\mathbb{E}(X_{T_{0}}1_{\Omega_{j}}),~{}b_{j}=\mathbb{E}(X^{2}_{T_{0}}1_{\Omega_{j}}).

Then, we have

𝔼(XT0λ2)𝔼(XT0λ)2=B2bj+2(1+λ)2bj(A2aj+2(1+λ)aj)2+o(λ).\frac{\mathbb{E}({X_{T_{0}^{\lambda}}}^{2})}{\mathbb{E}(X_{T_{0}^{\lambda}})^{2}}=\frac{B-2b_{j}+2(1+\lambda)^{2}b_{j}}{(A-2a_{j}+2(1+\lambda)a_{j})^{2}}+o(\lambda).

Therefore, we have

0=ddλκ(XT0λ)|λ=0=4bjA4ajBA3,0=\frac{d}{d\lambda}\kappa(X_{T_{0}^{\lambda}})|_{\lambda=0}=\frac{4b_{j}A-4a_{j}B}{A^{3}},

and hence,

bjaj=BAfor anyj=1,,2m.\frac{b_{j}}{a_{j}}=\frac{B}{A}~{}~{}\mbox{for any}~{}~{}j=1,\cdots,2m.

Denoting

uj=arg(αj)arg(αj1αj)π2,vj=arg(αj)arg(αjαj+1)π2,u_{j}=\arg(\alpha_{j})-\arg(\alpha_{j-1}-\alpha_{j})-\frac{\pi}{2},~{}v_{j}=\arg(\alpha_{j})-\arg(\alpha_{j}-\alpha_{j+1})-\frac{\pi}{2},

it holds that

aj=𝔼(XT01Ωj)=arg(αj1αj)arg(αjαj+1)2|αj|sin(arg(αj)ω)dω2π\displaystyle a_{j}=\mathbb{E}(X_{T_{0}}1_{\Omega_{j}})=\int_{\arg(\alpha_{j-1}-\alpha_{j})}^{\arg(\alpha_{j}-\alpha_{j+1})}2|\alpha_{j}|\sin(\arg(\alpha_{j})-\omega)\frac{d\omega}{2\pi}
=|αj|πuj+π/2vj+π/2sinω(1)𝑑ω=|αj|π(cos(vj+π/2)cos(uj+π/2))\displaystyle=\frac{|\alpha_{j}|}{\pi}\int_{u_{j}+\pi/2}^{v_{j}+\pi/2}\sin\omega~{}(-1)d\omega=\frac{|\alpha_{j}|}{\pi}(\cos(v_{j}+\pi/2)-\cos(u_{j}+\pi/2))
=|αj|π(sinujsinvj)\displaystyle=\frac{|\alpha_{j}|}{\pi}(\sin u_{j}-\sin v_{j})
bj=𝔼(XT021Ωj)=arg(αj1αj)arg(αjαj+1)2|αj|2sin2(arg(αj)ω)dω2π\displaystyle b_{j}=\mathbb{E}(X_{T_{0}}^{2}1_{\Omega_{j}})=\int_{\arg(\alpha_{j-1}-\alpha_{j})}^{\arg(\alpha_{j}-\alpha_{j+1})}2|\alpha_{j}|^{2}\sin^{2}(\arg(\alpha_{j})-\omega)\frac{d\omega}{2\pi}
=|αj|2πuj+π/2vj+π/2sin2(ω)(1)𝑑ω=|αj|2π(ujvj2+sin(2vj+π)sin(2uj+π)4)\displaystyle=\frac{|\alpha_{j}|^{2}}{\pi}\int_{u_{j}+\pi/2}^{v_{j}+\pi/2}\sin^{2}(\omega)~{}(-1)d\omega=\frac{|\alpha_{j}|^{2}}{\pi}\left(\frac{u_{j}-v_{j}}{2}+\frac{\sin(2v_{j}+\pi)-\sin(2u_{j}+\pi)}{4}\right)
=|αj|2π(ujvj2+sin2ujsin2vj4).\displaystyle=\frac{|\alpha_{j}|^{2}}{\pi}\left(\frac{u_{j}-v_{j}}{2}+\frac{\sin 2u_{j}-\sin 2v_{j}}{4}\right).
0αj\alpha_{j}αj1\alpha_{j-1}njn_{j}αj+1\alpha_{j+1}nj+1n_{j+1}vjv_{j}uju_{j}ujvju_{j}-v_{j}
Figure 6: αj1,αj,αj+1,nj,nj+1,uj,vj\alpha_{j-1},\alpha_{j},\alpha_{j+1},n_{j},n_{j+1},u_{j},v_{j}

For j=1,,2mj=1,\cdots,2m, let

nj=the perpendicular leg from 0 to the lineαj1αjn_{j}=\mbox{the perpendicular leg from 0 to the line}~{}\alpha_{j-1}\alpha_{j}

Then, the above uju_{j} and vjv_{j} have another representation (see Figure 6) that

uj=arg(αj)arg(nj),vj=arg(αj)arg(nj+1)(π,π]\displaystyle u_{j}=\arg(\alpha_{j})-\arg(n_{j}),~{}v_{j}=\arg(\alpha_{j})-\arg(n_{j+1})\in(-\pi,\pi] (5)
ujvj=the exterior angle atαj>0\displaystyle u_{j}-v_{j}=\mbox{the exterior angle at}~{}\alpha_{j}>0

and we have

BA=|αj|(2uj2vj+sin2ujsin2vj)4(sinujsinvj)forj=1,,2m.\frac{B}{A}=\frac{|\alpha_{j}|(2u_{j}-2v_{j}+\sin 2u_{j}-\sin 2v_{j})}{4(\sin u_{j}-\sin v_{j})}~{}\mbox{for}~{}j=1,\cdots,2m. (6)

For a fixed j=1,,mj=1,\cdots,m and a real number λ\lambda near 0, let

T0𝐢λ=T(α1,,(1+𝐢λ)αj,αj+1,,(1+𝐢λ)αj+m,,α2m).T^{\mathbf{i}\lambda}_{0}=T(\alpha_{1},\cdots,(1+\mathbf{i}\lambda)\alpha_{j},\alpha_{j+1},\cdots,(1+\mathbf{i}\lambda)\alpha_{j+m},\cdots,\alpha_{2m}).

Then, we have

𝔼(XT0λ2)𝔼(XT0λ)2=B+2dj2λ2π2dj+12λ2π(A+2djλ2π2dj+1λ2π)2+o(λ),\frac{\mathbb{E}({X_{T_{0}^{\lambda}}}^{2})}{\mathbb{E}(X_{T_{0}^{\lambda}})^{2}}=\frac{B+2d_{j}^{2}\frac{\lambda}{2\pi}-2d_{j+1}^{2}\frac{\lambda}{2\pi}}{(A+2d_{j}\frac{\lambda}{2\pi}-2d_{j+1}\frac{\lambda}{2\pi})^{2}}+o(\lambda),

where

dj=XT0(arg(αj1αj))\displaystyle d_{j}=X_{T_{0}}(\arg(\alpha_{j-1}-\alpha_{j}))
=2|αj|sin(arg(αj)arg(αj1αj))=2|αj|cosuj\displaystyle~{}~{}~{}=2|\alpha_{j}|\sin(\arg(\alpha_{j})-\arg(\alpha_{j-1}-\alpha_{j}))=2|\alpha_{j}|\cos u_{j}
dj+1=XT0(arg(αjαj+1))\displaystyle d_{j+1}=X_{T_{0}}(\arg(\alpha_{j}-\alpha_{j+1}))
=2|αj|sin(arg(αj)arg(αjαj+1))=2|αj|cosvj.\displaystyle~{}~{}~{}~{}~{}~{}=2|\alpha_{j}|\sin(\arg(\alpha_{j})-\arg(\alpha_{j}-\alpha_{j+1}))=2|\alpha_{j}|\cos v_{j}.

Therefore, we have

0=ddλκ(XT0𝐢λ)|λ=0=1π(dj2dj+12)A2π(djdj+1)BA3,0=\frac{d}{d\lambda}\kappa(X_{T_{0}^{\mathbf{i}\lambda}})|_{\lambda=0}=\frac{\frac{1}{\pi}(d_{j}^{2}-d_{j+1}^{2})A-\frac{2}{\pi}(d_{j}-d_{j+1})B}{A^{3}},

and hence, either dj=dj+1d_{j}=d_{j+1} or dj+dj+12=BA\frac{d_{j}+d_{j+1}}{2}=\frac{B}{A}.

If dj+dj+12=BA\frac{d_{j}+d_{j+1}}{2}=\frac{B}{A} holds, then by (6), we have

|αj|(2uj2vj+sin2ujsin2vj)4(sinujsinvj)=|αj|(cosuj+cosvj).\frac{|\alpha_{j}|(2u_{j}-2v_{j}+\sin 2u_{j}-\sin 2v_{j})}{4(\sin u_{j}-\sin v_{j})}=|\alpha_{j}|(\cos u_{j}+\cos v_{j}).

Hence,

2(ujvj)=sin2ujsin2vj+4sin(ujvj).2(u_{j}-v_{j})=\sin 2u_{j}-\sin 2v_{j}+4\sin(u_{j}-v_{j}).

Since ujvj>0u_{j}-v_{j}>0 and if ujvjπ2u_{j}-v_{j}\leq\frac{\pi}{2}, then sin(ujvj)2π(ujvj)\sin(u_{j}-v_{j})\geq\frac{2}{\pi}(u_{j}-v_{j}), we have a contradiction that sin2ujsin2vj+4sin(ujvj)>2(ujvj)\sin 2u_{j}-\sin 2v_{j}+4\sin(u_{j}-v_{j})>2(u_{j}-v_{j}). Therefore, ujvj>π2u_{j}-v_{j}>\frac{\pi}{2} and the exterior angle of αj\alpha_{j} is larger than π/2\pi/2. By the symmetry, the exterior angle of αk\alpha_{k} for kj,j+πk\neq j,j+\pi is smaller than π/2\pi/2. Therefore, for these kk, dj+dj+12=BA\frac{d_{j}+d_{j+1}}{2}=\frac{B}{A} is impossible, and hence, dk=dk+1d_{k}=d_{k+1} holds. Thus, uk=vk>0u_{k}=-v_{k}>0. This implies |nk|=|nk+1||n_{k}|=|n_{k+1}| for any kj,j+mk\neq j,j+m.

Any case, we have |nk|=|nk+1||n_{k}|=|n_{k+1}| except for k=j,j+mk=j,j+m. This implies that there are 2 classes

|nj+1|=|nj+2|==|nj+m|\displaystyle|n_{j+1}|=|n_{j+2}|=\cdots=|n_{j+m}|
|nj+m+1|=|nj+m+2|==|nj+2m|\displaystyle|n_{j+m+1}|=|n_{j+m+2}|=\cdots=|n_{j+2m}|
(suffices are considered modulo2m).\displaystyle(\mbox{suffices are considered modulo}~{}2m).

By symmetry, the values of these 2 classes coincide.

Thus, |n1|=|n2|==|n2m||n_{1}|=|n_{2}|=\cdots=|n_{2m}|, which implies that T0T_{0} has a inscribed circle with radius r:=|n1|r:=|n_{1}|. Hence, uj=vj(j=1,,2m)u_{j}=-v_{j}~{}(j=1,\cdots,2m). Then by (6),

BA=|αj|(2uj+sin2uj)4sinuj=|αj|cosuj(2uj+sin2uj)4sinujcosuj\displaystyle\frac{B}{A}=\frac{|\alpha_{j}|(2u_{j}+\sin 2u_{j})}{4\sin u_{j}}=\frac{|\alpha_{j}|\cos u_{j}(2u_{j}+\sin 2u_{j})}{4\sin u_{j}\cos u_{j}}
=r(2uj+sin2uj)2sin2uj=r22ujsin2uj+r2(j=1,,2m).\displaystyle=\frac{r(2u_{j}+\sin 2u_{j})}{2\sin 2u_{j}}=\frac{r}{2}~{}\frac{2u_{j}}{\sin 2u_{j}}+\frac{r}{2}~{}~{}(j=1,\cdots,2m).

It follows from this that 2ujsin2uj\frac{2u_{j}}{\sin 2u_{j}} is the same for j=1,,2mj=1,\cdots,2m. Since the correspondence xxsinxx\mapsto\frac{x}{\sin x} for x>0x>0 is one-to-one, we have u1==u2mu_{1}=\cdots=u_{2m}. Thus, T0T_{0} has the same exterior angle 2uj2u_{j} at vertex αj\alpha_{j} for j=1,,2mj=1,\cdots,2m. Together with that T0T_{0} has a inscribed circle, T0T_{0} is a regular 2m2m-gon, which completes the proof. \Box

5 Minimum of deviation rate among nn-gons

Lemma 2.

If n2n\geq 2, then the set

P(n):={(c0,,cn1){1,1}n|j=0n1cjexp(jπ𝐢n)=0}P(n):=\left\{(c_{0},\dots,c_{n-1})\in\{-1,1\}^{n}\ \left|\ \sum_{j=0}^{n-1}c_{j}\exp\left(\frac{j\pi\mathbf{i}}{n}\right)=0\right.\right\}

is empty if and only if nn is a power of 22.

Proof If nn is not a power of 22, then take an odd factor pp of nn. Denote j=0,1,,n1j=0,1,\cdots,n-1 as

j=knp+(=0,1,,np1;k=0,1,,p1).j=k\frac{n}{p}+\ell~{}~{}(\ell=0,1,\cdots,\frac{n}{p}-1;~{}k=0,1,\cdots,p-1).

Given c{1,1}(=0,1,,np1)c_{\ell}\in\{-1,1\}~{}(\ell=0,1,\cdots,\frac{n}{p}-1) arbitrary, define

cj=c(1)k(j=0,1,,n1).c_{j}=c_{\ell}(-1)^{k}~{}~{}(j=0,1,\cdots,n-1).

Then, we have

j=0n1cjejπ𝐢n=j=0n1c(1)kexp((knp+)π𝐢n)\displaystyle\sum_{j=0}^{n-1}c_{j}e^{\frac{j\pi\mathbf{i}}{n}}=\sum_{j=0}^{n-1}c_{\ell}(-1)^{k}\exp\left(\frac{(k\frac{n}{p}+\ell)\pi\mathbf{i}}{n}\right)
==0(n/p)1cexp(π𝐢n)k=0p1(1)kexp(kπ𝐢p)\displaystyle=\sum_{\ell=0}^{(n/p)-1}c_{\ell}\exp\left(\frac{\ell\pi\mathbf{i}}{n}\right)\sum_{k=0}^{p-1}(-1)^{k}\exp\left(\frac{k\pi\mathbf{i}}{p}\right)
==0(n/p)1cexp(π𝐢n)(k=0p12exp((2k𝐢)πp)k=1p12exp((2k1)π𝐢p))\displaystyle=\sum_{\ell=0}^{(n/p)-1}c_{\ell}\exp\left(\frac{\ell\pi\mathbf{i}}{n}\right)\left(\sum_{k=0}^{\frac{p-1}{2}}\exp\left(\frac{(2k\mathbf{i})\pi}{p}\right)-\sum_{k=1}^{\frac{p-1}{2}}\exp\left(\frac{(2k-1)\pi\mathbf{i}}{p}\right)\right)
==0(n/p)1cexp(2π𝐢n)(k=0p12exp(2kπ𝐢p)+k=1p12exp((2k1+p)π𝐢p))\displaystyle=\sum_{\ell=0}^{(n/p)-1}c_{\ell}\exp\left(\frac{2\pi\ell\mathbf{i}}{n}\right)\left(\sum_{k=0}^{\frac{p-1}{2}}\exp\left(\frac{2k\pi\mathbf{i}}{p}\right)+\sum_{k=1}^{\frac{p-1}{2}}\exp\left(\frac{(2k-1+p)\pi\mathbf{i}}{p}\right)\right)
==0(n/p)1cexp(2π𝐢n)(k=0(p1)/2exp(2kπ𝐢p)+k=(p+1)/2p1exp(2kπ𝐢p))\displaystyle=\sum_{\ell=0}^{(n/p)-1}c_{\ell}\exp\left(\frac{2\pi\ell\mathbf{i}}{n}\right)\left(\sum_{k=0}^{(p-1)/2}\exp\left(\frac{2k\pi\mathbf{i}}{p}\right)+\sum_{k=(p+1)/2}^{p-1}\exp\left(\frac{2k\pi\mathbf{i}}{p}\right)\right)
==0(n/p)1cexp(2π𝐢n)k=0p1exp(2kπ𝐢p)=0.\displaystyle=\sum_{\ell=0}^{(n/p)-1}c_{\ell}\exp\left(\frac{2\pi\ell\mathbf{i}}{n}\right)\sum_{k=0}^{p-1}\exp\left(\frac{2k\pi\mathbf{i}}{p}\right)=0.

Therefore P(n)P(n) contains a non-empty subset

Q(p):={(c0,,cn1){1,1}n;ck(n/p)+=c(1)k}Q(p):=\left\{(c_{0},\dots,c_{n-1})\in\{-1,1\}^{n};~{}c_{k(n/p)+\ell}=c_{\ell}(-1)^{k}\right\}

with Card(Q(p))=2n/p\mathrm{Card}(Q(p))=2^{n/p}. Thus we see that P(n)P(n) is non-empty.

Next assume that n=2sn=2^{s} with s1s\geq 1. Since xn+1[x]x^{n}+1\in\mathbb{Z}[x] is the minimum polynomial of w:=exp(𝐢πn)w:=\exp\left(\mathbf{i}\frac{\pi}{n}\right),

j=0n1cjexp(jπ𝐢n)=c0+c1w++cn1wn1\sum_{j=0}^{n-1}c_{j}\exp\left(\frac{j\pi\mathbf{i}}{n}\right)=c_{0}+c_{1}w+\cdots+c_{n-1}w^{n-1}

cannot be 0 for any (c0,c1,,cn1){1,1}n(c_{0},c_{1},\cdots,c_{n-1})\in\{-1,1\}^{n}. Hence, P(n)=P(n)=\emptyset. \Box

Theorem 6.

Let n2n\geq 2 be an integer which is not a power of 22. Then we have

minTΘnδ(XT)=δ(X~Rn)=νn(see(3)).\min_{T\in\Theta_{n}}\delta(X_{T})=\delta(\tilde{X}_{R_{n}})=\nu_{n}~{}~{}(\mbox{see}~{}(3)).

The minimum is attained if and only if the asymmetrization of TT is RnR_{n}, and hence, if and only if TT is similar to the polygon T(α1,,αn)T(\alpha_{1},\dots,\alpha_{n}) with

{αj+1αj;j=1,,n}={cjexp(jπ𝐢n);(c0,,cn1)P(n)}.\{\alpha_{j+1}-\alpha_{j};~{}j=1,\dots,n\}=\left\{c_{j}\exp\left(\frac{j\pi\mathbf{i}}{n}\right);~{}(c_{0},\dots,c_{n-1})\in P(n)\right\}.

Here P(n)P(n) is defined in Lemma 2.

Proof In light of Theorem 5, TΘnT^{*}\in\Theta_{n} attains the minimum of δ(XT)\delta(X_{T}) among TΘnT\in\Theta_{n} if TT^{*} is a realization of the pre-edge bundle RnR_{n}. A realization of RnR_{n}, say TT, may have more than nn edges but this number is reduced to nn if and only if there exists a one-to-one correspondence between the set of pre-edges {β1,,βn}\{\beta_{1},\cdots,\beta_{n}\} of RnR_{n} and the set of edges of TT, i.e., there exists (c0,c1,,cn1){1,1}n(c_{0},c_{1},\dots,c_{n-1})\in\{-1,1\}^{n} such that {c0β0,c1β1,,cn1βn1}\{c_{0}\beta_{0},c_{1}\beta_{1},\dots,c_{n-1}\beta_{n-1}\} is the set of edges of TT. Since we may assume βj=exp(jπ𝐢n)\beta_{j}=\exp\left(\frac{j\pi\mathbf{i}}{n}\right) for i=0,,n1i=0,\dots,n-1, this condition is satisfied if and only if

j=0n1cjexp(jπ𝐢n)=0\sum_{j=0}^{n-1}c_{j}\exp\left(\frac{j\pi\mathbf{i}}{n}\right)=0 (7)

is solvable in cj{1,1}c_{j}\in\{-1,1\}, i.e. P(n)P(n) is non-empty. \Box

Hereafter n>1n>1 is always assumed to be an integer which is not a power of 22. There is a natural map σ\sigma and τ\tau from P(n)P(n) to itself defined by

σ((c0,c1,,cn1))=(c1,cn1,c0)\sigma((c_{0},c_{1},\dots,c_{n-1}))=(c_{1},\dots c_{n-1},-c_{0})

and

τ((c0,c1,,cn1))=(cn1,cn2,,c0)\tau((c_{0},c_{1},\dots,c_{n-1}))=(c_{n-1},c_{n-2},\dots,c_{0})

which corresponds to the symmetry of dihedral group D2nD_{2n}: the rotation of angle π/n\pi/n and the reflexion. Two elements (c0,,cn1)(c_{0},\dots,c_{n-1}) and (c0,,cn1)(c^{\prime}_{0},\dots,c^{\prime}_{n-1}) of P(n)P(n) give congruent realizations if and only if

(c0,,cn1)=σj(c0,,cn1)(c_{0},\dots,c_{n-1})=\sigma^{j}(c^{\prime}_{0},\dots,c^{\prime}_{n-1})

or

(c0,,cn1)=σjτ(c0,,cn1)(c_{0},\dots,c_{n-1})=\sigma^{j}\tau(c^{\prime}_{0},\dots,c^{\prime}_{n-1})

for some j{0,,2n1}j\in\{0,\dots,2n-1\}. Lemma 2 and Theorem 6 can be restated in a geometric form.

Theorem 7.

If 1<n1<n\in\mathbb{N} is not a power of 22, then the minimum polygon in Θn\Theta_{n} is a Reinhardt nn-gon and vice versa. A Reinhardt nn-gon exists if and only if nn is not a power of 22.

Though it is not stated in this manner, the latter statement follows from the characterization of Reinhardt [9], see the discussion after the proof.

Proof Assume that 1<n1<n\in\mathbb{N} is not a power of 22. By Theorem 6, the minimum nn-polygon TT^{*} can be considered, without generality, to have the asymmetrization

(cexp(π𝐢2n),cexp(3π𝐢2n),,cexp((2n1)π𝐢2n))\mathcal{B}\left(c\exp\left(\frac{\pi\mathbf{i}}{2n}\right),c\exp\left(\frac{3\pi\mathbf{i}}{2n}\right),\cdots,c\exp\left(\frac{(2n-1)\pi\mathbf{i}}{2n}\right)\right)

for some appropriate c>0c>0, so that its symmetrization is

T2n=T(e(n)π𝐢2n,e(n+2)π𝐢2n,,e(3n2)π𝐢2n).T_{2n}=T\left(e^{\frac{(-n)\pi\mathbf{i}}{2n}},e^{\frac{(-n+2)\pi\mathbf{i}}{2n}},\cdots,e^{\frac{(3n-2)\pi\mathbf{i}}{2n}}\right).

Let T=T(P1,P2,,Pn)T^{*}=T(P_{1},P_{2},\cdots,P_{n}). Since XT=XT2nX_{T^{*}}=X_{T_{2n}} and by (4), XT(ω)X_{T^{*}}(\omega) is a periodic function of period π/n\pi/n such that if ωη(modπ/n)\omega\equiv\eta~{}(\mbox{mod}~{}\pi/n) with η(π2n,π2n]\eta\in(-\frac{\pi}{2n},\frac{\pi}{2n}],

XT(ω)=2cosη.X_{T^{*}}(\omega)=2\cos\eta.

Therefore, XT(ω)X_{T^{*}}(\omega) for ω(0,2π]\omega\in(0,2\pi] repeats its maximum 2 and its minimum 2cos(π2n)2\cos\left(\frac{\pi}{2n}\right), 2n2n times.

Since TT^{*} has the asymmetrization of the same size nn, there are no parallel edges in TT^{*}. Hence, the endpoints of the minimum shadow of TT^{*}, say at ω=jπn+π2n\omega=\frac{j\pi}{n}+\frac{\pi}{2n} come from a vertex, say PkP_{k}, and an edge, say PlPl+1P_{l}P_{l+1}. Those of the 2 neighboring maximum shadows come from the vertices Pk,PlP_{k},P_{l} and from the vertices Pk,Pl+1P_{k},P_{l+1}, respectively. Hence, PkPl¯=PkPl+1¯\overline{P_{k}P_{l}}=\overline{P_{k}P_{l+1}} holds. Replace the edge PlPl+1P_{l}P_{l+1} by the circular arc centered at PkP_{k} having the endpoints at PlP_{l} and Pl+1P_{l+1}. Repeating this replacement for j=1,2,,nj=1,2,\cdots,n, we get a convex body of constant width. It is easy to see that this convex body is a Reuleaux pp-body for some pp. Hence, TT^{*} is a Reinhardt nn-gon.

We prove the converse. Let TT be a Reinhardt nn-gon coming from a Reuleaux pp-body of width rr. By (2), we can rearrange the circular arc of SS by parallel translations into a circle 𝕊\mathbb{S} of radius rr so that the circular arcs are essentially disjoint and cover just a half part of 𝕊\mathbb{S}, and by the rotation of angle π\pi, they moved to the other half part of 𝕊\mathbb{S}. The edges of TT correspond to the chord of the circular arcs. By the above property, it is easy to see that the asymmetrization of TT is a regular pre-edge bundle of size nn. Hence, δ(XT)=νn\delta(X_{T})=\nu_{n} and TT is the minimum polygon.

The “if” part of the last statement follows from the first part. To prove the “only if” part, suppose that a Reinhardt nn-polygon TT exists for n=2kn=2^{k}. Then by the above argument, TT has the asymmetrization RnR_{n} which has a realization TT of the same size. This contradicts Lemma 2. \Box

Let us describe the correspondence between (c0,,cn1)P(n)(c_{0},\dots,c_{n-1})\in P(n) and cyclic integer vector. Choose i{0,1,,2n1}i\in\{0,1,\dots,2n-1\} that σi(c0,,cn1)=(d0,,dn1)\sigma^{i}(c_{0},\dots,c_{n-1})=(d_{0},\dots,d_{n-1}) with d0=dn1d_{0}=d_{n-1}. Count number of runs of 11 and 1-1 in (d0,,dn1)(d_{0},\dots,d_{n-1}), i.e., we write (d0,,dn1)(d_{0},\dots,d_{n-1}) like 1n1(1)n21^{n_{1}}(-1)^{n_{2}}\dots or (1)n11n2(-1)^{n_{1}}1^{n_{2}}\dots. Then (n1,n2,,np)(n_{1},n_{2},\dots,n_{p}) is the desired cyclic vector. For the converse direction, we choose either 1n1(1)n21^{n_{1}}(-1)^{n_{2}}\dots or (1)n11n2(-1)^{n_{1}}1^{n_{2}}\dots.

Reinhardt [9] gave an alternative characterization of the cyclic vector (n1,n2,,np)(n_{1},n_{2},\dots,n_{p}) with n=i=1pnin=\sum_{i=1}^{p}n_{i}: it is a cyclic vector if and only if pp is odd and the polynomial

1zn1+zn1+n2+zn1+n2++np11-z^{n_{1}}+z^{n_{1}+n_{2}}-\dots+z^{n_{1}+n_{2}+\dots+n_{p-1}}

is divisible by Φ2n(z)\Phi_{2n}(z), the 2n2n-th cyclotomic polynomial. For completeness, we show that this characterization is equivalent to ours. As above, we assume that d0=dn1d_{0}=d_{n-1}. From (7) we have

j=0n1djzj=0with z=exp(π𝐢n).\sum_{j=0}^{n-1}d_{j}z^{j}=0\quad\text{with }z=\exp\left(\frac{\pi\mathbf{i}}{n}\right).

We consider that zz is a variable and multiply z1z-1. Then we see

(z1)j=0n1djzj\displaystyle(z-1)\sum_{j=0}^{n-1}d_{j}z^{j}
=12zn1+2zn1+n2+2zn1++np1zn\displaystyle=1-2z^{n_{1}}+2z^{n_{1}+n_{2}}-\dots+2z^{n_{1}+\dots+n_{p-1}}-z^{n} (8)
22zn1+2zn1+n2+2zn1++np1(modΦ2n(z)).\displaystyle\equiv 2-2z^{n_{1}}+2z^{n_{1}+n_{2}}-\dots+2z^{n_{1}+\dots+n_{p-1}}\pmod{\Phi_{2n}(z)}.

Dividing by 22, we see the condition of Reinhardt. To get the converse, we just go backwards. Note that the polynomial (8) is divisible by z1z-1 as pp is odd. By Lemma 2, the second statement of Theorem 7 is derived from the Reinhardt criterion.

The subset Q(p)Q(p) in the proof of Lemma 2 corresponds to pp-fold rotational symmetry. We can find Reinhardt polygons without any symmetry [3, 4].

6 Truncation of the regular triangle

Let X,YX,Y be general \mathbb{R}-valued, square integrable random variables on the probability space Ω\Omega. Assume further that X0X\geq 0 everywhere and 𝔼(X)>0\mathbb{E}(X)>0. Recall that

κ(X)=𝔼(X2)𝔼(X)2=δ(X)2+1.\kappa(X)=\frac{\mathbb{E}(X^{2})}{\mathbb{E}(X)^{2}}=\delta(X)^{2}+1.

It holds for any tt\in\mathbb{R} with sufficiently small modulus that

κ(X+tY)=𝔼((X+tY)2)𝔼(X+tY)2=𝔼(X2)𝔼(X)21+2t𝔼(XY)𝔼(X2)+t2𝔼(Y2)𝔼(X2)1+2t𝔼(Y)𝔼(X)+t2𝔼(Y)2𝔼(X)2\displaystyle\kappa(X+tY)=\frac{\mathbb{E}((X+tY)^{2})}{\mathbb{E}(X+tY)^{2}}=\frac{\mathbb{E}(X^{2})}{\mathbb{E}(X)^{2}}\frac{1+2t\frac{\mathbb{E}(XY)}{\mathbb{E}(X^{2})}+t^{2}\frac{\mathbb{E}(Y^{2})}{\mathbb{E}(X^{2})}}{1+2t\frac{\mathbb{E}(Y)}{\mathbb{E}(X)}+t^{2}\frac{\mathbb{E}(Y)^{2}}{\mathbb{E}(X)^{2}}}
=𝔼(X2)𝔼(X)2(1+2t(𝔼(XY)𝔼(X2)𝔼(Y)𝔼(X))+O(t2)).\displaystyle=\frac{\mathbb{E}(X^{2})}{\mathbb{E}(X)^{2}}\left(1+2t\left(\frac{\mathbb{E}(XY)}{\mathbb{E}(X^{2})}-\frac{\mathbb{E}(Y)}{\mathbb{E}(X)}\right)+O(t^{2})\right).

If 𝔼(XY)𝔼(X2)𝔼(Y)𝔼(X)=0\frac{\mathbb{E}(XY)}{\mathbb{E}(X^{2})}-\frac{\mathbb{E}(Y)}{\mathbb{E}(X)}=0, then we have

κ(X+tY)=𝔼(X2)𝔼(X)2(1+t2(𝔼(Y2)𝔼(X2)𝔼(Y)2𝔼(X)2)+O(t3)).\kappa(X+tY)=\frac{\mathbb{E}(X^{2})}{\mathbb{E}(X)^{2}}\left(1+t^{2}\left(\frac{\mathbb{E}(Y^{2})}{\mathbb{E}(X^{2})}-\frac{\mathbb{E}(Y)^{2}}{\mathbb{E}(X)^{2}}\right)+O(t^{3})\right).

Hence, the following Lemma holds.

Lemma 3.

(i)(i) ddtδ(X+tY)|t=0>,=,<0\frac{d}{dt}\delta(X+tY)|_{t=0}>,=,<0 if and only if 𝔼(XY)𝔼(X2)>,=,<𝔼(Y)𝔼(X)\frac{\mathbb{E}(XY)}{\mathbb{E}(X^{2})}>,=,<\frac{\mathbb{E}(Y)}{\mathbb{E}(X)}, respectively.
(ii)(ii) Assume that “==” holds in (i)(i). Then, there exists ε>0\varepsilon>0 such that δ(X+tY)>,<δ(X)\delta(X+tY)>,<\delta(X) for any t(ε,ε){0}t\in(-\varepsilon,\varepsilon)\setminus\{0\} if 𝔼(Y2)𝔼(X2)>,<𝔼(Y)2𝔼(X)2\frac{\mathbb{E}(Y^{2})}{\mathbb{E}(X^{2})}>,<\frac{\mathbb{E}(Y)^{2}}{\mathbb{E}(X)^{2}}, respectively.

Proof (i)(i) follows since

ddtδ(X+tY)=(1/2)(κ(X+tY)1)1/2ddtκ(X+tY).\textstyle\frac{d}{dt}\delta(X+tY)=(1/2)(\kappa(X+tY)-1)^{-1/2}\frac{d}{dt}\kappa(X+tY).

(ii)(ii) follows since δ(X+tY)\delta(X+tY) is a monotone increasing function of κ(X+tY)\kappa(X+tY). \Box

Lemma 4.

Let TT and SS be triangles in \mathbb{C}. Then, XT=XSX_{T}=X_{S} (a.s.) holds if and only if there exists zz\in\mathbb{C} such that either S=T+zS=T+z or S=T+zS=-T+z.

Proof XT=XTX_{-T}=X_{T} holds since they have the same asymmetrization. Hence, the “if” part holds.

Let us prove the “only if” part. Let T=T(α,β,γ)T=T(\alpha,\beta,\gamma). Consider XT(ω)X_{T}(\omega) as a function of ω/\omega\in\mathbb{R}/\mathbb{Z}. Then, it is locally minimal if and only if ω\omega is equal to either of

±(arg(β)arg(α)),±(arg(γ)arg(β)),±(arg(α)arg(γ))\pm(\arg(\beta)-\arg(\alpha)),~{}\pm(\arg(\gamma)-\arg(\beta)),~{}\pm(\arg(\alpha)-\arg(\gamma))

modulo 2π2\pi. If XT=XSX_{T}=X_{S} (a.s.), then they should have the same set of ω\omega as this. Also, at any of these ω\omega, they should have the same height. This implies that either S=T+zS=T+z or S=T+zS=-T+z for some zz\in\mathbb{C}. \Box

Lemma 5.

Let TT and SS be triangles in \mathbb{C}. Then,

𝔼(XTXS)𝔼(XT2)1/2𝔼(XS2)1/2.\mathbb{E}(X_{T}X_{S})\leq\mathbb{E}(X_{T}^{2})^{1/2}\mathbb{E}(X_{S}^{2})^{1/2}.

The equality holds if and only if there exist λ0\lambda\geq 0 and zz\in\mathbb{C} such that either S=λT+zS=\lambda T+z or S=λT+zS=-\lambda T+z.

Proof This is the Cauchy-Schwartz inequality for the inner product X,Y=E(XY)\langle X,Y\rangle=E(XY). The equality holds if and only if there exists λ>0\lambda>0 such that XS=λXT=XλTX_{S}=\lambda X_{T}=X_{\lambda T} (a.s.). Hence, by Lemma 4, if and only if S=λT+zS=\lambda T+z or S=λT+zS=-\lambda T+z for some zz\in\mathbb{C}. \Box

Lemma 6.

Let T=T(α,β,γ)T=T(\alpha,\beta,\gamma) be a regular triangle. Let σ{0}\sigma\in\mathbb{C}\setminus\{0\}. Then,

𝔼(XTXσ)𝔼(XT2)𝔼(Xσ)𝔼(XT).\frac{\mathbb{E}(X_{T}X_{\sigma})}{\mathbb{E}(X_{T}^{2})}\leq\frac{\mathbb{E}(X_{\sigma})}{\mathbb{E}(X_{T})}.

The equality holds if and only if 0σ\vec{0\sigma} is parallel to one of the edges of TT, that is, arg(σ)\arg(\sigma) is equal to one of

±(arg(β)arg(α)),±(arg(γ)arg(β)),±(arg(α)arg(γ))\pm(\arg(\beta)-\arg(\alpha)),~{}\pm(\arg(\gamma)-\arg(\beta)),~{}\pm(\arg(\alpha)-\arg(\gamma))

modulo 2π2\pi.

Proof Without loss of generality, we assume that the length of the edges of TT is 1. It holds that 𝔼(Xμz1Xμz2)=𝔼(Xz1Xz2)\mathbb{E}(X_{\mu z_{1}}X_{\mu z_{2}})=\mathbb{E}(X_{z_{1}}X_{z_{2}}) for any μ,z1,z2\mu,z_{1},z_{2}\in\mathbb{C} with |μ|=1|\mu|=1. Moreover, since e2π𝐢/3T=Te^{2\pi\mathbf{i}/3}T=T, we have

𝔼(XTXσ)=𝔼(XTXe2π𝐢/3σ)=𝔼(XTXe4π𝐢/3σ).\mathbb{E}(X_{T}X_{\sigma})=\mathbb{E}(X_{T}X_{e^{2\pi\mathbf{i}/3}\sigma})=\mathbb{E}(X_{T}X_{e^{4\pi\mathbf{i}/3}\sigma}).

Hence, 𝔼(XTXσ)=(1/3)𝔼(XTXS)\mathbb{E}(X_{T}X_{\sigma})=(1/3)\mathbb{E}(X_{T}X_{S}) with the regular triangle S=T(0,σ,eπ𝐢/3σ)S=T(0,\sigma,e^{\pi\mathbf{i}/3}\sigma). Therefore by Lemma 5,

𝔼(XTXσ)=(1/3)𝔼(XTXS)(1/3)𝔼(XT2)1/2𝔼(XS2)1/2\displaystyle\mathbb{E}(X_{T}X_{\sigma})=(1/3)\mathbb{E}(X_{T}X_{S})\leq(1/3)\mathbb{E}(X_{T}^{2})^{1/2}\mathbb{E}(X_{S}^{2})^{1/2}
=(1/3)𝔼(XT2)1/2𝔼((|σ|XT)2)1/2=(|σ|/3)𝔼(XT2).\displaystyle=(1/3)\mathbb{E}(X_{T}^{2})^{1/2}\mathbb{E}((|\sigma|X_{T})^{2})^{1/2}=(|\sigma|/3)\mathbb{E}(X_{T}^{2}).

The equality holds if and only if 0σ\vec{0\sigma} is parallel to one of the edges of TT. Thus, we have

𝔼(XTXσ)𝔼(XT2)|σ|3=𝔼(Xσ)𝔼(XT)\frac{\mathbb{E}(X_{T}X_{\sigma})}{\mathbb{E}(X_{T}^{2})}\leq\frac{|\sigma|}{3}=\frac{\mathbb{E}(X_{\sigma})}{\mathbb{E}(X_{T})}

with the equality if and only if 0σ\vec{0\sigma} is parallel to one of the edges of TT. \Box

Theorem 8.

A parallel truncation of the regular triangle increases the deviation rate, while a non-parallel truncation decreases it.

Parallel truncation:

Let T=T(α,β,γ)T=T(\alpha,\beta,\gamma) be a regular triangle of the edge length 1. We also assume that it is of the counter clockwise order. Let t>0t>0 be sufficiently small. Let

βt=(1t)α+tβ,γt=(1t)α+tγ.\beta_{t}=(1-t)\alpha+t\beta,~{}\gamma_{t}=(1-t)\alpha+t\gamma.

Let

Vt=T(βt,β,γ,γt)V_{t}=T(\beta_{t},\beta,\gamma,\gamma_{t})

be a parallel truncation of TT at α\alpha. Then, we have

XVt=Xββt+Xγβ+Xγtγ+Xβtγt\displaystyle X_{V_{t}}=X_{\beta-\beta_{t}}+X_{\gamma-\beta}+X_{\gamma_{t}-\gamma}+X_{\beta_{t}-\gamma_{t}}
=XTtXβαtXαγ+tXβγ\displaystyle=X_{T}-tX_{\beta-\alpha}-tX_{\alpha-\gamma}+tX_{\beta-\gamma}
=XTtXY\displaystyle=X_{T}-tX_{Y}

with Y=Xc+XbXaY=X_{\vec{c}}+X_{\vec{b}}-X_{-\vec{a}}, where we denote a=γβ,b=αγ,c=βα\vec{a}=\gamma-\beta,~{}\vec{b}=\alpha-\gamma,~{}\vec{c}=\beta-\alpha. Since 𝔼(Xc)=𝔼(Xb)=𝔼(Xa)=1/π\mathbb{E}(X_{\vec{c}})=\mathbb{E}(X_{\vec{b}})=\mathbb{E}(X_{-\vec{a}})=1/\pi, we have by Lemma 6,

𝔼(XTXc)𝔼(XT2)=𝔼(XTXb)𝔼(XT2)=𝔼(XTXa)𝔼(XT2)=1/π𝔼(XT),\frac{\mathbb{E}(X_{T}X_{\vec{c}})}{\mathbb{E}(X_{T}^{2})}=\frac{\mathbb{E}(X_{T}X_{\vec{b}})}{\mathbb{E}(X_{T}^{2})}=\frac{\mathbb{E}(X_{T}X_{-\vec{a}})}{\mathbb{E}(X_{T}^{2})}=\frac{1/\pi}{\mathbb{E}(X_{T})},

and hence,

𝔼(XTY)𝔼(XT2)=𝔼(XTXc)+𝔼(XTXb)𝔼(XTXa)𝔼(XT2)+𝔼(XT2)𝔼(XT2)=1/π𝔼(XT)=𝔼(Y)𝔼(XT).\frac{\mathbb{E}(X_{T}Y)}{\mathbb{E}(X_{T}^{2})}=\frac{\mathbb{E}(X_{T}X_{\vec{c}})+\mathbb{E}(X_{T}X_{\vec{b}})-\mathbb{E}(X_{T}X_{-\vec{a}})}{\mathbb{E}(X_{T}^{2})+\mathbb{E}(X_{T}^{2})-\mathbb{E}(X_{T}^{2})}=\frac{1/\pi}{\mathbb{E}(X_{T})}=\frac{\mathbb{E}(Y)}{\mathbb{E}(X_{T})}.

Now, we prove that

𝔼(Y2)𝔼(XT2)>𝔼(Y)2𝔼(XT)2\frac{\mathbb{E}(Y^{2})}{\mathbb{E}(X_{T}^{2})}>\frac{\mathbb{E}(Y)^{2}}{\mathbb{E}(X_{T})^{2}}

so that δ(Vt)>δ(T)\delta(V_{t})>\delta(T) for sufficiently small |t|0|t|\neq 0 by Lemma 3. We have

𝔼(Y)=𝔼(Xb+XcXa)=1π\displaystyle\mathbb{E}(Y)=\mathbb{E}(X_{\vec{b}}+X_{\vec{c}}-X_{-\vec{a}})=\frac{1}{\pi}
𝔼(XT)=3π\displaystyle\mathbb{E}(X_{T})=\frac{3}{\pi}
𝔼(Y2)=𝔼((Xb+XcXa)2)\displaystyle\mathbb{E}(Y^{2})=\mathbb{E}((X_{\vec{b}}+X_{\vec{c}}-X_{-\vec{a}})^{2})
=34+(2/4π)(V(2π/3)2V(π/3))=1334π\displaystyle=\frac{3}{4}+(2/4\pi)(V(2\pi/3)-2V(\pi/3))=\frac{1}{3}-\frac{\sqrt{3}}{4\pi}
𝔼(XT2)=34+(6/4π)V(2π/3)=12+334π\displaystyle\mathbb{E}(X_{T}^{2})=\frac{3}{4}+(6/4\pi)V(2\pi/3)=\frac{1}{2}+\frac{3\sqrt{3}}{4\pi}
𝔼(Y2)𝔼(XT2)=1334π12+334π=0.214>19=𝔼(Y)2𝔼(XT)2,\frac{\mathbb{E}(Y^{2})}{\mathbb{E}(X_{T}^{2})}=\frac{\frac{1}{3}-\frac{\sqrt{3}}{4\pi}}{\frac{1}{2}+\frac{3\sqrt{3}}{4\pi}}=0.214\cdots>\frac{1}{9}=\frac{\mathbb{E}(Y)^{2}}{\mathbb{E}(X_{T})^{2}},

and complete the proof that δ(Vt)>δ(T)\delta(V_{t})>\delta(T) for sufficiently small |t|0|t|\neq 0 in the case of parallel truncation.

Non-parallel truncation:

Let T=T(α,β,γ)T=T(\alpha,\beta,\gamma) be a regular triangle of the edge length 1. We also assume that it is of the counter clockwise order. Let 0<λ10<\lambda\neq 1 and t>0t>0 be sufficiently small. Let

βt=(1t)α+tβ,γt=(1λt)α+λtγ.\beta_{t}=(1-t)\alpha+t\beta,~{}\gamma_{t}=(1-\lambda t)\alpha+\lambda t\gamma.

Let

Vt=T(βt,β,γ,γt)V_{t}=T(\beta_{t},\beta,\gamma,\gamma_{t})

be a non-parallel truncation of TT at α\alpha. Then, we have

XVt=Xββt+Xγβ+Xγtγ+Xβtγt\displaystyle X_{V_{t}}=X_{\beta-\beta_{t}}+X_{\gamma-\beta}+X_{\gamma_{t}-\gamma}+X_{\beta_{t}-\gamma_{t}}
=XTtXctXλb+tXe\displaystyle=X_{T}-tX_{\vec{c}}-tX_{\lambda\vec{b}}+tX_{\vec{e}}
=XTtY,\displaystyle=X_{T}-tY,

where we denote b=αγ,c=βα,e=c+λb\vec{b}=\alpha-\gamma,~{}\vec{c}=\beta-\alpha,~{}\vec{e}=\vec{c}+\lambda\vec{b} and Y=Xc+XλbXeY=X_{\vec{c}}+X_{\lambda\vec{b}}-X_{\vec{e}}.

By Lemma 6, we have

𝔼(XTXc)𝔼(XT2)=𝔼(XTXb)𝔼(XT2)=1/π𝔼(XT)and𝔼(XTXe)𝔼(XT2)<|e|/π𝔼(XT).\frac{\mathbb{E}(X_{T}X_{\vec{c}})}{\mathbb{E}(X_{T}^{2})}=\frac{\mathbb{E}(X_{T}X_{\vec{b}})}{\mathbb{E}(X_{T}^{2})}=\frac{1/\pi}{\mathbb{E}(X_{T})}~{}\mbox{and}~{}\frac{\mathbb{E}(X_{T}X_{\vec{e}})}{\mathbb{E}(X_{T}^{2})}<\frac{|\vec{e}|/\pi}{\mathbb{E}(X_{T})}.

Hence,

𝔼(XTXc)+λ𝔼(XTXb)𝔼(XTXe))𝔼(XT2)+λ𝔼(XT2)|e|𝔼(XT2)>1/π𝔼(XT),\frac{\mathbb{E}(X_{T}X_{\vec{c}})+\lambda\mathbb{E}(X_{T}X_{\vec{b}})-\mathbb{E}(X_{T}X_{\vec{e}}))}{\mathbb{E}(X_{T}^{2})+\lambda\mathbb{E}(X_{T}^{2})-|\vec{e}|\mathbb{E}(X_{T}^{2})}>\frac{1/\pi}{\mathbb{E}(X_{T})},

and we have

𝔼(XTY)𝔼(XT2)>(1+λ|e|)/π𝔼(XT).\frac{\mathbb{E}(X_{T}Y)}{\mathbb{E}(X_{T}^{2})}>\frac{(1+\lambda-|\vec{e}|)/\pi}{\mathbb{E}(X_{T})}.

Thus, by Lemma 3, ddtδ(XT+tY)|t=0>0\frac{d}{dt}\delta(X_{T}+tY)|_{t=0}>0, which implies that

δ(XVt)=δ(XTtY)<δ(XT)\delta(X_{V_{t}})=\delta(X_{T}-tY)<\delta(X_{T})

if t>0t>0 is small, which completes the proof.

7 Remaining problems

When nn is a power of 22, the method in this paper does not apply because by Lemma 2, there is no nn-gon realization of the regular pre-edge bundle RnR_{n}. The case n=4n=4 may be of special interest.

Refer to caption
Figure 7: Left: δ\delta-minimum, Center: Problem 1 & 2, Right: Problem 3

By numerical calculation, a possible minimum of δ\delta in Θ4\Theta_{4} is attained by the kite-shape with vertices

(A,B,C,D)((0,0.24213332485),(1,0),(0,1.67502597318),(1,0))(A,B,C,D)\approx((0,-0.24213332485),(1,0),(0,1.67502597318),(-1,0))

having the deviation rate 0.0353064253050.035306425305. This is not a solution to any of three optimization problems for n=4n=4 in the introduction. See Figure 7 for a comparison of solutions having the same horizontal width. Indeed from AB<AC<BC<BDAB<AC<BC<BD, we see that this shape is not optimal for Problems 1 and 2. The optimizers for these problems are the same, which is based on the regular triangle with an additional vertex similar to the construction of Reinhardt polygon, see [6, 2]. The kite ABCDABCD is not optimal for Problem 3 either. Indeed [1] showed that the maximum width of quadrangles with the unit perimeter is close to 9+6340.295\frac{\sqrt{-9+6\sqrt{3}}}{4}\approx 0.295, possibly attained by

((0,3+233),(1,0),(0,1+23),(1,0)),\left(\left(0,-\frac{\sqrt{-3+2\sqrt{3}}}{3}\right),(1,0),\left(0,\sqrt{1+\frac{2}{\sqrt{3}}}\right),(-1,0)\right),

while the kite shape gives the value 0.288. For Problem 1,2 and 3, the solutions have algebraic expressions. We do not know if our kite has such an algebraic expression.

In the next paper, we will discuss the minimality of δ\delta-values and the minimal shapes under all infinitesimal deformations including the cases when nn is a power of 22.

References

  • [1] C. Audet, P. Hansen, F. Messine, Isoperimetric polygons of maximum width. Discrete Comput. Geom. 41(1), 45–60 (2009)
  • [2] A. Bezdek, F. Fodor, On convex polygons of maximal width. Arch. Math. (Basel) 74(1), 75–80 (2000)
  • [3] K. G. Hare, M. J. Mossinghoff, Sporadic Reinhardt Polygons, Discrete Comp. Geom. 49, 540-557 (2013)
  • [4] K. G. Hare, M. J. Mossinghoff, Most Reinhardt polygons are sporadic. Geom. Dedicata 198 (2019), 1–18.
  • [5] H. Martini, L. Montejano, D. Oliveros, Bodies of constant width. An introduction to convex geometry with applications. Birkhäuser/Springer, Cham, 2019. xi+486 pp.
  • [6] M. J. Mossinghoff, A $1 problem, Amer. Math. Monthly 113 (2006), no. 5, 385–402.
  • [7] M. J. Mossinghoff, Enumerating isodiametric and isoperimetric polygons, Journal of Combinatorial Theory, Series A 118 (2011) 1801–1815.
  • [8] H. Rademacher, O. Toeplitz, The enjoyment of math, Princeton Science Library. Princeton University Press, Princeton, NJ, 1994. iv+205 pp. ISBN: 0-691-02351-4
  • [9] Von K. Reinhardt, Extremale Polygone gegebenen Durchmessers, Jahresber. Deutsh. Math. Verein. 31, 251-270 (1922)