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

On the sine polarity and the LpL_{p}-sine Blaschke-Santaló inequality 111Keywords: Blaschke-Santaló inequality, LpL_{p}-sine transform, LpL_{p}-sine centroid body, sine polar body.

Qingzhong Huang, Ai-Jun Li 222Corresponding author: Ai-Jun Li, Dongmeng Xi and Deping Ye
Abstract

This paper is dedicated to study the sine version of polar bodies and establish the LpL_{p}-sine Blaschke-Santaló inequality for the LpL_{p}-sine centroid body.

The LpL_{p}-sine centroid body ΛpK\Lambda_{p}K for a star body KnK\subset\mathbb{R}^{n} is a convex body based on the LpL_{p}-sine transform, and its associated Blaschke-Santaló inequality provides an upper bound for the volume of ΛpK\Lambda_{p}^{\circ}K, the polar body of ΛpK\Lambda_{p}K, in terms of the volume of KK. Thus, this inequality can be viewed as the “sine cousin” of the LpL_{p} Blaschke-Santaló inequality established by Lutwak and Zhang. As pp\rightarrow\infty, the limit of ΛpK\Lambda_{p}^{\circ}K becomes the sine polar body KK^{\diamond} and hence the LpL_{p}-sine Blaschke-Santaló inequality reduces to the sine Blaschke-Santaló inequality for the sine polar body. The sine polarity naturally leads to a new class of convex bodies 𝒞en\mathcal{C}_{e}^{n}, which consists of all origin-symmetric convex bodies generated by the intersection of origin-symmetric closed solid cylinders. Many notions in 𝒞en\mathcal{C}_{e}^{n} are developed, including the cylindrical support function, the supporting cylinder, the cylindrical Gauss image, and the cylindrical hull. Based on these newly introduced notions, the equality conditions of the sine Blaschke-Santaló inequality are settled.

2020 Mathematics Subject Classification: 52A20, 52A30, 52A40, 94A15.

1 Introduction

Geometric inequalities, such as isoperimetric or affine isoperimetric inequalities, are the central objects of interest in convex geometry. These inequalities aim to estimate geometric (or affine) invariants from above and/or below in terms of the volume. One of the most important affine isoperimetric inequalities is the celebrated Blaschke-Santaló inequality. It asserts that, in the nn-dimensional Euclidean space n\mathbb{R}^{n},

V(K)V(K)ωn2V(K)V(K^{\circ})\leq\omega_{n}^{2} (1.1)

holds for any origin-symmetric convex body KK or more generally for any convex body KK with its centroid at the origin, with equality if and only if KK is an origin-symmetric ellipsoid (see, e.g., [4, 57, 56, 46, 51, 48, 47]). Here V()V(\cdot) denotes the volume (Lebesgue measure) in n\mathbb{R}^{n} and the volume of the Euclidean unit ball BnB^{n} in n\mathbb{R}^{n} is solely written by ωn=πn/2/Γ(1+n/2)\omega_{n}=\pi^{n/2}/\Gamma(1+n/2). A convex body KK in n\mathbb{R}^{n} is a convex compact subset of n\mathbb{R}^{n} with nonempty interior. If the convex body KK contains the origin in its interior, then the polar body KK^{\circ} of KK is defined by

K={xn:xy1for allyK},K^{\circ}=\Big{\{}x\in\mathbb{R}^{n}:x\cdot y\leq 1\quad\textrm{for all}~{}y\in K\Big{\}},

where xyx\cdot y is the inner product of x,ynx,y\in\mathbb{R}^{n}. In 1990’s, a more general version of the Blaschke-Santaló inequality was established by Lutwak and Zhang [43]. It states that, for KnK\subset\mathbb{R}^{n} being a star body and 1p1\leq p\leq\infty, one has

V(K)V(ΓpK)ωn2,V(K)V(\Gamma_{p}^{\circ}K)\leq\omega_{n}^{2}, (1.2)

with equality if and only if KK is an origin-symmetric ellipsoid, where ΓpK\Gamma_{p}^{\circ}K denotes the polar body of the LpL_{p} centroid body ΓpK\Gamma_{p}K whose support function at xnx\in\mathbb{R}^{n} is given by

hΓpK(x)p=1cn,pV(K)K|xy|p𝑑y,h_{\Gamma_{p}K}(x)^{p}=\frac{1}{c_{n,p}V(K)}\int_{K}|x\cdot y|^{p}\,dy, (1.3)

where dy\,dy is the Lebesgue measure and

cn,p=ωn+pω2ωnωp1,c_{n,p}=\frac{\omega_{n+p}}{\omega_{2}\omega_{n}\omega_{p-1}}, (1.4)

with ωp=πn/2/Γ(1+p/2)\omega_{p}=\pi^{n/2}/\Gamma(1+p/2) for p>0p>0. The normalization above is chosen so that ΓpBn=Bn\Gamma_{p}B^{n}=B^{n}. The body ΓK\Gamma_{\infty}K is to be interpreted as the limit of ΓpK\Gamma_{p}K for pp\rightarrow\infty. In the case that KK is an origin-symmetric convex body and p=p=\infty, ΓK\Gamma_{\infty}^{\circ}K coincides with KK^{\circ}, and hence inequality (1.2) becomes the Blaschke-Santaló inequality (1.1). So inequality (1.2) is usually called the LpL_{p} Blaschke-Santaló inequality.

It can be observed that the definition of the LpL_{p} centroid body is based on the cosine function (i.e., the inner product). More specifically, integrating in polar coordinates, for xnx\in\mathbb{R}^{n},

hΓpK(x)p=nωn(n+p)cn,pV(K)Sn1|xu|pρK(u)n+p𝑑u=nωn(n+p)cn,pV(K)(𝒞pρK()n+p)(x),h_{\Gamma_{p}K}(x)^{p}=\!\frac{n\omega_{n}}{(n+p)c_{n,p}V(K)}\!\int_{S^{n-1}}|x\cdot u|^{p}\rho_{K}(u)^{n+p}\,du=\frac{n\omega_{n}}{(n+p)c_{n,p}V(K)}\big{(}\mathcal{C}_{p}\,\rho_{K}(\cdot)^{n+p}\big{)}(x),

where du\,du is the rotation invariant probability measure on the unit sphere Sn1S^{n-1} in n\mathbb{R}^{n} and 𝒞pμ\mathcal{C}_{p}\,\mu denotes the LpL_{p}-cosine transform of a Borel measure μ\mu defined on Sn1S^{n-1}. That is, for p>0p>0,

(𝒞pμ)(x)=Sn1|xu|p𝑑μ(u),xn.(\mathcal{C}_{p}\,\mu)(x)=\int_{S^{n-1}}|x\cdot u|^{p}\,d\mu(u),\ \ \ x\in\mathbb{R}^{n}. (1.5)

The LpL_{p}-cosine transform provides a very useful analytical operator for convex geometry and plays a dominating role in applications, see, e.g., the books [13, 26, 58, 24] and references [6, 18, 19, 29, 31, 35, 39, 40, 43, 55, 30, 59], among others.

Similar to (1.5), 𝒮pμ\mathcal{S}_{p}\,\mu, the LpL_{p}-sine transform of a Borel measure μ\mu defined on Sn1S^{n-1}, can be defined for p>0p>0 as follows:

(𝒮pμ)(x)=Sn1|Pux|p𝑑μ(u)=Sn1[x,u]p𝑑μ(u),xn,(\mathcal{S}_{p}\,\mu)(x)=\int_{S^{n-1}}|\textrm{P}_{u^{\perp}}x|^{p}\,d\mu(u)=\int_{S^{n-1}}[x,u]^{p}\,d\mu(u),\ \ \ x\in\mathbb{R}^{n}, (1.6)

where Pux\textrm{P}_{u^{\perp}}x is the orthogonal projection of xx onto the (n1)(n-1)-dimensional subspace uu^{\perp} perpendicular to uu, and [x,u]=|x|2|u|2|xu|2[x,u]=\sqrt{|x|^{2}|u|^{2}-|x\cdot u|^{2}} with |x||x| being the Euclidean norm of xnx\in\mathbb{R}^{n}. Note that [x,u][x,u] represents the sine function due to the Pythagoras theorem, and geometrically [x,u][x,u] is the 22-dimensional volume of the parallelepiped spanned by xx and uu. The sine, cosine, and Radon transforms are closely related and have been extensively studied, see, e.g., [16, 52, 54, 53, 45, 60, 14, 15, 20]. The LpL_{p}-sine transform (i.e., m=n1m=n-1) and the LpL_{p}-cosine transform (i.e., m=1m=1) on Sn1S^{n-1} are the special cases of the transform 𝒞m,pμ\mathcal{C}_{m,p}\,\mu on Grassmann manifolds 𝐆n,m\mathbf{G}_{n,m} (the set of mm-dimensional linear subspaces in n\mathbb{R}^{n}) [28]. Here, for a Borel measure μ\mu on 𝐆n,m\mathbf{G}_{n,m} and p>0p>0,

(𝒞m,pμ)(x)=𝐆n,m|Pξx|p𝑑μ(ξ),xn.(\mathcal{C}_{m,p}\,\mu)(x)=\int_{\mathbf{G}_{n,m}}|\textrm{P}_{\xi}x|^{p}d\mu(\xi),\ \ \ x\in\mathbb{R}^{n}.

It can be checked that, when p1p\geq 1, (𝒞m,pμ)1/p(\mathcal{C}_{m,p}\,\mu)^{1/p} defines a norm on n\mathbb{R}^{n} and hence its unit ball is an origin-symmetric convex body in n\mathbb{R}^{n}. Some isoperimetric and reverse isoperimetric inequalities for the unit balls induced by the norm (𝒞m,pμ)1/p(\mathcal{C}_{m,p}\,\mu)^{1/p} were established in [28]. Based on the L2L_{2}-sine transform, the first three authors of the present paper proposed two new sine ellipsoids, and obtained sharp volume inequalities and valuation properties for these sine ellipsoids in [27]. Note that these sine ellipsoids are closely related in the Pythagorean relation and duality to the Legendre ellipsoid and its dual ellipsoid (known as the LYZ ellipsoid) [36]. It is worth mentioning that, when n=2n=2, the LpL_{p}-sine transform coincides with the LpL_{p}-cosine transform up to a rotation of π/2\pi/2. Thus, properties for the LpL_{p}-cosine transform shall hold for the LpL_{p}-sine transform on 2\mathbb{R}^{2}. Unfortunately, this is no long true when n3n\geq 3, and the situation is totally different. The main difficulty for the LpL_{p}-sine transform is the lack of affine nature, which makes the analysis on the LpL_{p}-sine transform often more challenging.

The main goal of the present paper is to establish the LpL_{p}-sine Blaschke-Santaló inequality (1.9), which can be viewed as the “sine cousin” of the LpL_{p} Blaschke-Santaló inequality (1.2). Furthermore, a sine version of the Blaschke-Santaló inequality (1.10) with characterization of equality is also obtained. Thus, our results are complementary to the classical studies for the cosine transform.

In Section 3, we introduce a new convex body in n\mathbb{R}^{n} (i.e., the LpL_{p}-sine centroid body ΛpK\Lambda_{p}K for KK being an Ln+pL_{n+p}-star) in terms of the LpL_{p}-sine transform (1.6), whose support function at xnx\in\mathbb{R}^{n} is defined by, for p1p\geq 1,

hΛpK(x)p=nωn(n+p)c~n,pV(K)(𝒮pρK()n+p)(x)=1c~n,pV(K)K[x,y]p𝑑y,h_{\Lambda_{p}K}(x)^{p}=\frac{n\omega_{n}}{(n+p)\widetilde{c}_{n,p}V(K)}\big{(}\mathcal{S}_{p}\,\rho_{K}(\cdot)^{n+p}\big{)}(x)=\frac{1}{\widetilde{c}_{n,p}V(K)}\int_{K}[x,y]^{p}\,dy, (1.7)

where

c~n,p=(n1)ωn1ωn+p2(n+p)ωnωn+p3.\widetilde{c}_{n,p}=\frac{(n-1)\omega_{n-1}\omega_{n+p-2}}{(n+p)\omega_{n}\omega_{n+p-3}}. (1.8)

In particular, ΛpBn=Bn\Lambda_{p}B^{n}=B^{n} (see (4.1)). Note that ΛpK\Lambda_{p}K is well-defined since hΛpKh_{\Lambda_{p}K} is a convex function (see Section 3 for details). In the special case when KK is a star body in n\mathbb{R}^{n}, ΛpK\Lambda_{p}K was first proposed in [27]. Motivated by the limit of ΛpK\Lambda_{p}^{\circ}K as pp\rightarrow\infty, the sine polar body of KK, denoted by KK^{\diamond}, is introduced. Namely, for a subset KnK\subset\mathbb{R}^{n}, we let

K={xn:[x,y]1forallyK}.K^{\diamond}=\Big{\{}x\in\mathbb{R}^{n}:[x,y]\leq 1\ \ \mathrm{for\ all}\ y\in K\Big{\}}.

The sine polar body KK^{\diamond} differs from KK^{\circ} mainly with [x,y][x,y] replacing xyx\cdot y. Hence many properties for KK^{\diamond} are similar to those for KK^{\circ}, for instance, LKL^{\diamond}\subseteq K^{\diamond} if KLK\subseteq L, (Bn)=Bn(B^{n})^{\diamond}=B^{n}, and (cK)=c1K(cK)^{\diamond}=c^{-1}\cdot K^{\diamond} for any c>0c>0. However, some properties for KK^{\diamond} are completely different from those for KK^{\circ}. For instance, the volume product V(K)V(K)V(K)V(K^{\circ}) is GL(n)\textrm{GL}(n)-invariant (i.e., V(ϕK)V((ϕK))=V(K)V(K)V(\phi K)V((\phi K)^{\circ})=V(K)V(K^{\circ}) for any invertible linear map ϕ\phi on n\mathbb{R}^{n}), but the volume product V(K)V(K)V(K)V(K^{\diamond}) is in general not GL(n)\textrm{GL}(n)-invariant (except for n=2n=2). The non-GL(n)\textrm{GL}(n)-invariance for V(K)V(K)V(K)V(K^{\diamond}) does bring extra difficulty in characterizing the equality for the sine Blaschke-Santaló inequality (1.10) in Section 5. On the other hand, the sine bipolar property K=KK=K^{\diamond\diamond}, where K=(K)K^{\diamond\diamond}=(K^{\diamond})^{\diamond}, does not hold in general for all convex bodies. Indeed, the definition of KK^{\diamond} indicates that KK^{\diamond} is formed by the intersection of closed solid cylinders (see (3.7)). This brings our attention to a special class of convex bodies 𝒞en\mathcal{C}_{e}^{n} consisting of all origin-symmetric convex bodies generated by the intersection of origin-symmetric closed solid cylinders. The sine bipolar property K=KK=K^{\diamond\diamond} holds for all K𝒞enK\in\mathcal{C}_{e}^{n}, see Proposition 3.3 (v). Proposition 3.5 (ii) also shows that KK^{\diamond\diamond} is just the cylindrical hull of KK – the intersection of all origin-symmetric closed solid cylinders containing KK. The cylindrical hull of KK is a notion analogous to the convex hull, and any convex body K𝒞enK\in\mathcal{C}_{e}^{n} must be equal to the cylindrical hull of itself (see Proposition 3.5 (i)). Moreover, in Section 3, we develop many notions in 𝒞en\mathcal{C}_{e}^{n}, such as the cylindrical support function, the supporting cylinder, and the cylindrical Gauss image. These notions are completely parallel to their classical counterparts, e.g., the support function, the supporting hyperplane, and the Gauss image.

Last but not the least, we would like to mention that the notion 𝒞en\mathcal{C}_{e}^{n} is quite important. A basic example of 𝒞en\mathcal{C}_{e}^{n} is the intersection of two closed solid cylinders with equal radius at right angles. The volume of this bicylinder (also known as “Mou He Fang Gai” in China) in 3\mathbb{R}^{3} was first studied by Archimedes more than 2200 years ago and independently by the old Chinese mathematicians Hui Liu, Chongzhi Zu and Geng Zu more than 1500 years ago (see, e.g., [23, 25]). An amazing application of the bicylinder is the calculation of the volume of 33-dimensional balls by these old Chinese mathematicians [25]. The solids obtained as the intersection of two or three cylinders of equal radius at right angles are better known as the Steinmetz solids in Europe, see Figure 1. Applications of the Steinmetz solid are common in real life, such as the T-adapter and the circular pipes joining. The intersection of cylinders also naturally appears in the study of crystal (see, e.g., [1, 49]).

Refer to caption
Refer to caption
Figure 1: The left one is the bicylinder (“Mou He Fang Gai” in Chinese) and the right one is a tricylinder in 3\mathbb{R}^{3}. More figures can be found in e.g., [8, 22].

In Section 4, we will prove the following LpL_{p}-sine Blaschke-Santaló inequality: let p1p\geq 1 and KnK\subset\mathbb{R}^{n} be a star body. Then

V(K)V(ΛpK)ωn2,V(K)V(\Lambda_{p}^{\circ}K)\leq\omega_{n}^{2}, (1.9)

with equality if and only if KK is an origin-symmetric ellipsoid when n=2n=2, and is an origin-symmetric ball when n3n\geq 3. For p=2p=2, this inequality was also established by the first three authors [27, Theorem 4.1]. By letting pp\rightarrow\infty, the LpL_{p}-sine Blaschke-Santaló inequality will lead to the following sine Blaschke-Santaló inequality: let KnK\subset\mathbb{R}^{n} be a star body. Then

V(K)V(K)ωn2,V(K)V(K^{\diamond})\leq\omega_{n}^{2}, (1.10)

with equality if and only if KK is an origin-symmetric ellipsoid when n=2n=2, and is an origin-symmetric ball when n3n\geq 3.

Therefore, inequalities (1.9) and (1.10) can be viewed as the “sine cousin” of inequalities (1.2) and (1.1). However, the approximation as pp\rightarrow\infty does not yield the equality conditions of inequality (1.10). To characterize these equality conditions requires a lot of work mainly because of the extra difficulty arising from the non-GL(n)\textrm{GL}(n)-invariance of V(K)V(K)V(K)V(K^{\diamond}) for n3n\geq 3. In this sense, equality conditions in inequality (1.1) and its sine counterpart (1.10) are different for n3n\geq 3. However, inequalities (1.1) and (1.10) are in fact equivalent if KK is an origin-symmetric convex body in 2\mathbb{R}^{2}. Indeed, it will be shown in Section 4 that under this circumstance KK^{\diamond} is just a rotation of KK^{\circ} by an angle of π/2\pi/2. A similar situation also happens to (1.2) and (1.9).

We also mention that information theory (or probability theory) and convex geometry are closely related (see, e.g., [12, Section 14]). In recent years there was a growing body of works in this direction, see, e.g., [5, 9, 11, 17, 37, 38, 41, 42, 33, 34, 50, 44]. For example, an important application of the LpL_{p} Blaschke-Santaló inequality (1.2) to information theory leads to the celebrated LpL_{p} moment-entropy inequality (see (6.7)) due to Lutwak, Yang, and Zhang [38]. Similarly, the “sine cousin” of LpL_{p} moment-entropy inequality (see (6.16)) can be deduced by using the LpL_{p}-sine Blaschke-Santaló inequality. We believe that the applications of the LpL_{p}-sine transform to information theory (or probability theory) will likely produce more interesting and useful results for applications in many areas.

2 Preliminaries

Throughout the paper, n\mathbb{R}^{n} denotes nn-dimensional Euclidean space (n2n\geq 2) and the term subspace means a linear subspace. As usual, xyx\cdot y denotes the inner product of x,ynx,y\in\mathbb{R}^{n} and |x||x| the Euclidean norm of xx. The unit sphere {xn:|x|=1}\{x\in\mathbb{R}^{n}:|x|=1\} is denoted by Sn1S^{n-1} and the unit ball {xn:|x|1}\{x\in\mathbb{R}^{n}:|x|\leq 1\} by BnB^{n}. Furthermore, we denote by oo the origin in n\mathbb{R}^{n} and x¯=x/|x|Sn1\bar{x}=x/|x|\in S^{n-1} for all xn{o}x\in\mathbb{R}^{n}\setminus\{o\}. Note that the cosine of the angle between nonzero vectors xx and yy is indeed given by x¯y¯\bar{x}\cdot\bar{y}. Thus, for x,yn{o}x,y\in\mathbb{R}^{n}\setminus\{o\},

|xy|=|x||y||x¯y¯|=|x||Pxy|,|x\cdot y|=|x|\cdot|y|\cdot|\bar{x}\cdot\bar{y}|=|x|\cdot|\textrm{P}_{x}y|, (2.1)

where Pxy\textrm{P}_{x}y is the orthogonal projection of yy onto the 11-dimensional subspace containing xx. On the other hand, the sine of the angle between nonzero vectors xx and yy can be easily calculated through the formula 1|x¯y¯|2\sqrt{1-|\bar{x}\cdot\bar{y}|^{2}}. Corresponding to (2.1), for x,yn{o}x,y\in\mathbb{R}^{n}\setminus\{o\}, let

[x,y]:=|x||y|1|x¯y¯|2=|x|2|y|2|xy|2=|x||y|2|Pxy|2=|x||Pxy|,[x,y]:=|x|\cdot|y|\cdot\sqrt{1-|\bar{x}\cdot\bar{y}|^{2}}=\sqrt{|x|^{2}|y|^{2}-|x\cdot y|^{2}}=|x|\cdot\sqrt{|y|^{2}-|\textrm{P}_{x}y|^{2}}=|x|\cdot|\textrm{P}_{x^{\perp}}y|, (2.2)

where Pxy\textrm{P}_{x^{\perp}}y is the orthogonal projection of yy onto the (n1)(n-1)-dimensional subspace xx^{\perp} perpendicular to xx. In other words, [x,y][x,y] represents the 22-dimensional volume of the parallelepiped spanned by xx and yy. Of course, one can switch the roles of x,yx,y to get [x,y]=[y,x]=|y||Pyx|[x,y]=[y,x]=|y|\cdot|\textrm{P}_{y^{\perp}}x|. Moreover, if one or both of xx and yy are zero vectors, one can simply let [x,y]=0[x,y]=0.

A convex body in n\mathbb{R}^{n} is a convex compact subset of n\mathbb{R}^{n} with nonempty interior. By 𝒦on\mathscr{K}_{o}^{n} we mean the class of all convex bodies with the origin oo in their interiors. For each K𝒦onK\in\mathscr{K}_{o}^{n}, one can define its support function hK:n(0,)h_{K}:\mathbb{R}^{n}\rightarrow(0,\infty) by

hK(x)=max{xy:yK},xn.h_{K}(x)=\max\{x\cdot y:y\in K\},\ \ \ x\in\mathbb{R}^{n}. (2.3)

A convex body K𝒦onK\in\mathscr{K}_{o}^{n} is said to be origin-symmetric if xK-x\in K for all xKx\in K. Denote by 𝒦en\mathscr{K}_{e}^{n} the subclass of all origin-symmetric elements of 𝒦on\mathscr{K}_{o}^{n}.

Denote by GL(n)\textrm{GL}(n) the group of all invertible linear transforms on n\mathbb{R}^{n} and SL(n)\textrm{SL}(n) the subgroup of GL(n)\textrm{GL}(n) whose determinant is 11. By O(n)\textrm{O}(n) we mean the orthogonal group of n\mathbb{R}^{n}, that is, the group of linear transformations preserving the inner product. For each ϕGL(n)\phi\in\textrm{GL}(n), let ϕt\phi^{t} and ϕ1\phi^{-1} be the transpose and inverse of ϕ\phi, respectively. For ϕGL(n)\phi\in\textrm{GL}(n) and xnx\in\mathbb{R}^{n}, let ϕK={ϕy:yK}\phi K=\{\phi y:y\in K\} and then

hϕK(x)=hK(ϕtx).h_{\phi K}(x)=h_{K}(\phi^{t}x). (2.4)

Throughout the paper, du\,du denotes the rotation invariant probability measure on Sn1S^{n-1}. Let q>0q>0. We say that a function ρ:Sn1[0,)\rho:S^{n-1}\rightarrow[0,\infty) is qq-integrable on Sn1S^{n-1} if ρ\rho is measurable and

Sn1ρ(u)q𝑑u<.\int_{S^{n-1}}\rho(u)^{q}\,du<\infty.

Denote by Lq(Sn1)L^{q}(S^{n-1}) the set of all qq-integrable functions on Sn1S^{n-1}. As Sn1𝑑u=1\int_{S^{n-1}}\,du=1, one can easily obtain, by using the Hölder inequality, that if 0<q1<q0<q_{1}<q, then

Lq(Sn1)Lq1(Sn1).L^{q}(S^{n-1})\subseteq L^{q_{1}}(S^{n-1}). (2.5)

A set LnL\subset\mathbb{R}^{n} is said to be star-shaped about the origin if oLo\in L and

L={ru:0rρ(u)foruSn1},L=\big{\{}ru:0\leq r\leq\rho(u)\ \ \textrm{for}\ \ u\in S^{n-1}\big{\}},

where ρ:Sn1[0,)\rho:S^{n-1}\rightarrow[0,\infty) is a nonnegative function. Such a function ρ\rho is called the radial function of LL, which will be more often written as ρL.\rho_{L}. The radial function can be extended to ρL:n{o}[0,)\rho_{L}:\mathbb{R}^{n}\setminus\{o\}\rightarrow[0,\infty) by ρL(ru)=r1ρL(u)\rho_{L}(ru)=r^{-1}\rho_{L}(u) for r>0r>0 and uSn1u\in S^{n-1}, which gives

ρL(x)=max{λ0:λxL},xn{o}.\rho_{L}(x)=\max\big{\{}\lambda\geq 0:\lambda x\in L\big{\}},\ \ \ x\in\mathbb{R}^{n}\setminus\{o\}. (2.6)

Let p>0p>0. A star-shaped set LL in n\mathbb{R}^{n} is said to be an Ln+pL_{n+p}-star generated by ρLn+p(Sn1)\rho\in L^{n+p}(S^{n-1}) if ρL=ρ\rho_{L}=\rho. Clearly, it follows from (2.6) that for c>0c>0 and xn\{o}x\in\mathbb{R}^{n}\backslash\{o\},

ρcL(x)=cρL(x).\rho_{cL}(x)=c\rho_{L}(x). (2.7)

For ϕGL(n)\phi\in\textrm{GL}(n) and xn\{o}x\in\mathbb{R}^{n}\backslash\{o\}, one has ρϕL(x)=ρL(ϕ1x)\rho_{\phi L}(x)=\rho_{L}(\phi^{-1}x).

A set LnL\subset\mathbb{R}^{n} is called a star body if LL is star-shaped and ρL\rho_{L} restricted on Sn1S^{n-1} is positive and continuous. Denote by 𝒮n\mathcal{S}^{n} the class of all star bodies. Clearly, 𝒦on𝒮n\mathscr{K}_{o}^{n}\subset\mathcal{S}^{n}, and LL is an Ln+pL_{n+p}-star for all p>0p>0 if L𝒮nL\in\mathcal{S}^{n}. Moreover, the Minkowski functional L:n[0,)\|\cdot\|_{L}:\mathbb{R}^{n}\rightarrow[0,\infty) of L𝒮nL\in\mathcal{S}^{n} is defined by

xL=min{λ0:xλL},xn.\|x\|_{L}=\min\{\lambda\geq 0:x\in\lambda L\},\ \ \ x\in\mathbb{R}^{n}. (2.8)

Recall that the polar body KK^{\circ} of K𝒦onK\in\mathscr{K}_{o}^{n} is defined by

K={xn:xy1for allyK}.K^{\circ}=\Big{\{}x\in\mathbb{R}^{n}:x\cdot y\leq 1\quad\textrm{for all}~{}y\in K\Big{\}}. (2.9)

It can be easily verified that for K𝒦onK\in\mathscr{K}_{o}^{n},

(cK)\displaystyle(cK)^{\circ}\!\!\! =\displaystyle= c1K,foranyc>0,\displaystyle\!\!\!c^{-1}K^{\circ},\ \ \ \mathrm{for\ any}\ c>0, (2.10)
(ϕK)\displaystyle(\phi K)^{\circ}\!\!\! =\displaystyle= ϕtK,foranyϕGL(n).\displaystyle\!\!\!\phi^{-t}K^{\circ},\ \ \ \mathrm{for\ any}\ \phi\in\textrm{GL}(n). (2.11)

Moreover, it follows from (2.3), (2.6), (2.8), and (2.9) that if K𝒦onK\in\mathscr{K}_{o}^{n},

xK=1/ρK(x)=hK(x),foranyxn{o}.\|x\|_{K^{\circ}}=1/\rho_{K^{\circ}}(x)=h_{K}(x),\ \ \ \mathrm{for\ any}\ x\in\mathbb{R}^{n}\setminus\{o\}. (2.12)

In particular, (K)=K(K^{\circ})^{\circ}=K for any K𝒦onK\in\mathscr{K}_{o}^{n}. Note that if K𝒦enK\in\mathscr{K}_{e}^{n} is an origin-symmetric convex body in n\mathbb{R}^{n}, (2.9) can also be rewritten as

K={xn:|xy|1for allyK}.K^{\circ}=\{x\in\mathbb{R}^{n}:|x\cdot y|\leq 1\quad\textrm{for all}~{}y\in K\}. (2.13)

Let p>0p>0. For LL being an Ln+pL_{n+p}-star, its volume is given by

V(L)=ωnSn1ρL(u)n𝑑u.V(L)=\omega_{n}\int_{S^{n-1}}\rho_{L}(u)^{n}\,du. (2.14)

By (2.5), for an Ln+pL_{n+p}-star LL, it has a finite volume as ρLLn+p(Sn1).\rho_{L}\in L^{n+p}(S^{n-1}). Moreover, the LpL_{-p} dual mixed volume [38] of an Ln+pL_{n+p}-star LL and a star body L𝒮nL^{\prime}\in\mathcal{S}^{n} can be formulated by

V~p(L,L)=ωnSn1ρL(u)n+pρL(u)p𝑑u.\widetilde{V}_{-p}(L,L^{\prime})=\omega_{n}\int_{S^{n-1}}\rho_{L}(u)^{n+p}\rho_{L^{\prime}}(u)^{-p}\,du. (2.15)

In particular, for L𝒮nL\in\mathcal{S}^{n} and p>0p>0, one has

V~p(L,L)=V(L).\widetilde{V}_{-p}(L,L)=V(L). (2.16)

Following from the Hölder inequality, one can easily obtain the following dual LpL_{-p} Minkowski inequality [38]: for p>0p>0,

V~p(L,L)nV(L)n+pV(L)p\widetilde{V}_{-p}(L,L^{\prime})^{n}\geq V(L)^{n+p}V(L^{\prime})^{-p} (2.17)

holds for any Ln+pL_{n+p}-star LL and any L𝒮nL^{\prime}\in\mathcal{S}^{n}, with equality if and only if there exists a constant c>0c>0 such that ρL(u)=cρL(u)\rho_{L^{\prime}}(u)=c\cdot\rho_{L}(u) for almost all uSn1u\in S^{n-1} with respect to the spherical measure of Sn1S^{n-1}.

3 The LpL_{p}-sine centroid body, the sine polar body, and the cylindrical hull

Let p1p\geq 1 and n2n\geq 2. We first show that, under the assumption that KnK\subset\mathbb{R}^{n} is an Ln+pL_{n+p}-star with V(K)>0V(K)>0, the set ΛpK\Lambda_{p}K defined in (1.7) is well-defined. Indeed, we can get hΛpK(0,)h_{\Lambda_{p}K}\in(0,\infty) on n{o}\mathbb{R}^{n}\setminus\{o\}. To see this, for any xn{o}x\in\mathbb{R}^{n}\setminus\{o\} and any yKy\in K, (2.2) implies [x,y]|x||y|[x,y]\leq|x|\cdot|y| and hence

hΛpK(x)p\displaystyle h_{\Lambda_{p}K}(x)^{p} =1c~n,pV(K)K[x,y]p𝑑y\displaystyle=\frac{1}{\widetilde{c}_{n,p}V(K)}\int_{K}[x,y]^{p}\,dy
1c~n,pV(K)K|x|p|y|p𝑑y\displaystyle\leq\frac{1}{\widetilde{c}_{n,p}V(K)}\int_{K}|x|^{p}|y|^{p}\,dy
=nωn|x|p(n+p)c~n,pV(K)Sn1ρK(u)n+p𝑑u<.\displaystyle=\frac{n\omega_{n}|x|^{p}}{(n+p)\widetilde{c}_{n,p}V(K)}\int_{S^{n-1}}\rho_{K}(u)^{n+p}\,du<\infty.

On the other hand, as V(K)>0V(K)>0, then KK cannot be concentrated on a 11-dimensional space, and [x,y]>0[x,y]>0 for almost all yKy\in K (with respect the Lebesgue measure). Hence, hΛpK(x)>0h_{\Lambda_{p}K}(x)>0 for all xn{o}x\in\mathbb{R}^{n}\setminus\{o\}.

It can be easily verified that hΛpK(tx)=thΛpK(x)h_{\Lambda_{p}K}(tx)=t\cdot h_{\Lambda_{p}K}(x) for all t0t\geq 0 and xnx\in\mathbb{R}^{n}. Moreover, hΛpK(x)h_{\Lambda_{p}K}(x) is a convex function, namely for all τ[0,1]\tau\in[0,1] and x,xnx,x^{\prime}\in\mathbb{R}^{n}, one has

hΛpK(τx+(1τ)x)τhΛpK(x)+(1τ)hΛpK(x).h_{\Lambda_{p}K}(\tau x+(1-\tau)x^{\prime})\leq\tau\cdot h_{\Lambda_{p}K}(x)+(1-\tau)\cdot h_{\Lambda_{p}K}(x^{\prime}).

To see this, for yn{o}y\in\mathbb{R}^{n}\setminus\{o\}, one has [x,y]=|y||Pyx|[x,y]=|y|\cdot|\textrm{P}_{y^{\perp}}x| and hence

[τx+(1τ)x,y]\displaystyle[\tau x+(1-\tau)x^{\prime},y]\!\! =\displaystyle= |y||Py(τx+(1τ)x)|\displaystyle\!\!|y|\cdot\big{|}\textrm{P}_{y^{\perp}}\big{(}\tau x+(1-\tau)x^{\prime}\big{)}\big{|} (3.1)
=\displaystyle= |y||τPyx+(1τ)Pyx|\displaystyle\!\!|y|\cdot\big{|}\tau\textrm{P}_{y^{\perp}}x+(1-\tau)\textrm{P}_{y^{\perp}}x^{\prime}\big{|}
\displaystyle\leq τ|y||Pyx|+(1τ)|y||Pyx|\displaystyle\!\!\tau|y|\cdot\big{|}\textrm{P}_{y^{\perp}}x\big{|}+(1-\tau)|y|\cdot\big{|}\textrm{P}_{y^{\perp}}x^{\prime}\big{|}
=\displaystyle= τ[x,y]+(1τ)[x,y].\displaystyle\!\!\tau[x,y]+(1-\tau)[x^{\prime},y].

Clearly, (3.1) is also valid for y=oy=o. It then follows from the Minkowski inequality that

hΛpK(τx+(1τ)x)\displaystyle h_{\Lambda_{p}K}(\tau x+(1-\tau)x^{\prime})\!\! =\displaystyle= (1c~n,pV(K)K[τx+(1τ)x,y]p𝑑y)1/p\displaystyle\!\!\bigg{(}\frac{1}{\widetilde{c}_{n,p}V(K)}\int_{K}[\tau x+(1-\tau)x^{\prime},y]^{p}\,dy\bigg{)}^{1/p}
\displaystyle\leq (1c~n,pV(K)K(τ[x,y]+(1τ)[x,y])p𝑑y)1/p\displaystyle\!\!\bigg{(}\frac{1}{\widetilde{c}_{n,p}V(K)}\int_{K}\Big{(}\tau[x,y]+(1-\tau)[x^{\prime},y]\Big{)}^{p}\,dy\bigg{)}^{1/p}
\displaystyle\leq τ(1c~n,pV(K)K[x,y]p𝑑y)1/p+(1τ)(1c~n,pV(K)K[x,y]p𝑑y)1/p\displaystyle\!\!\tau\cdot\bigg{(}\frac{1}{\widetilde{c}_{n,p}V(K)}\int_{K}[x,y]^{p}\,dy\bigg{)}^{1/p}+(1-\tau)\cdot\bigg{(}\frac{1}{\widetilde{c}_{n,p}V(K)}\int_{K}[x^{\prime},y]^{p}\,dy\bigg{)}^{1/p}
=\displaystyle= τhΛpK(x)+(1τ)hΛpK(x).\displaystyle\!\!\tau\cdot h_{\Lambda_{p}K}(x)+(1-\tau)\cdot h_{\Lambda_{p}K}(x^{\prime}).

Furthermore, ΛpK\Lambda_{p}K is clearly origin-symmetric. We conclude that ΛpK𝒦en\Lambda_{p}K\in\mathscr{K}_{e}^{n} for p1p\geq 1.

Denote by ΛpK\Lambda_{p}^{\circ}K the polar body of ΛpK\Lambda_{p}K. It follows from (2.12), (1.7), and the polar coordinate that

ρΛpK(x)p=nωn(n+p)c~n,pV(K)Sn1[x,u]pρK(u)n+p𝑑u,xn{o}.\displaystyle\rho_{\Lambda_{p}^{\circ}K}(x)^{-p}=\frac{n\omega_{n}}{(n+p)\widetilde{c}_{n,p}V(K)}\int_{S^{n-1}}[x,u]^{p}\rho_{K}(u)^{n+p}\,du,\ \ \ x\in\mathbb{R}^{n}\setminus\{o\}. (3.2)

By (2.7) and (3.2), one has

Λp(cK)=c1ΛpK,foranyc>0.\Lambda_{p}^{\circ}(cK)=c^{-1}\Lambda_{p}^{\circ}K,\ \ \ \mathrm{for\ any}\ c>0. (3.3)

If K𝒮nK\in\mathcal{S}^{n} is a star body in n\mathbb{R}^{n}, then KK is an Ln+pL_{n+p}-star for all p1p\geq 1 with V(K)>0V(K)>0. Thus, by (1.7) and Stirling’s formula,

hΛK(x)=limphΛpK(x)=maxyK[x,y],foranyxn.h_{\Lambda_{\infty}K}(x)=\lim_{p\rightarrow\infty}h_{\Lambda_{p}K}(x)=\max_{y\in K}\,[x,y],\ \ \ \mathrm{for\ any}\ x\in\mathbb{R}^{n}.

Again it is easy to check that ΛK𝒦en\Lambda_{\infty}K\in\mathscr{K}_{e}^{n}. Denote by ΛK\Lambda_{\infty}^{\circ}K the polar body of ΛK\Lambda_{\infty}K. It follows from (2.12) that

xΛK=hΛK(x)=maxyK[x,y],foranyxn.\displaystyle\|x\|_{\Lambda_{\infty}^{\circ}K}=h_{\Lambda_{\infty}K}(x)=\max_{y\in K}\,[x,y],\ \ \ \mathrm{for\ any}\ x\in\mathbb{R}^{n}. (3.4)

Formula (2.8) for the Minkowski functional yields that, for any star body L𝒮nL\in\mathcal{S}^{n}, one must have L={xn:xL1}.L=\big{\{}x\in\mathbb{R}^{n}:\ \|x\|_{L}\leq 1\big{\}}. Together with (3.4), one gets, if K𝒮nK\in\mathcal{S}^{n},

ΛK={xn:xΛK1}={xn:maxyK[x,y]1}.\Lambda_{\infty}^{\circ}K=\Big{\{}x\in\mathbb{R}^{n}:\ \|x\|_{\Lambda_{\infty}^{\circ}K}\leq 1\Big{\}}=\Big{\{}x\in\mathbb{R}^{n}:\ \max_{y\in K}\,[x,y]\leq 1\Big{\}}.

This motivates our definition for the sine polar body.

Definition 3.1.

Let KK be a subset in n\mathbb{R}^{n} with n2n\geq 2. Define KK^{\diamond}, the sine polar body of KK, to be

K={xn:[x,y]1forallyK}.K^{\diamond}=\Big{\{}x\in\mathbb{R}^{n}:[x,y]\leq 1\ \ \mathrm{for\ all}\ y\in K\Big{\}}. (3.5)

Clearly, for K𝒮nK\in\mathcal{S}^{n}, one has

K=ΛK.K^{\diamond}=\Lambda_{\infty}^{\circ}K. (3.6)

If K𝒦enK\in\mathscr{K}_{e}^{n}, one sees that the major difference of (2.13) and (3.5) is in fact the replacement of |xy||x\cdot y| by [x,y][x,y], and hence the newly defined sine polar body KK^{\diamond} can be viewed as the sine counterpart of KK^{\circ}.

Throughout the paper, a closed solid cylinder C(u,r)nC^{-}(u,r)\subset\mathbb{R}^{n} with axis being the line {tu:t}\{tu:t\in\mathbb{R}\} for uSn1u\in S^{n-1} and base radius r>0r>0 is a subset of n\mathbb{R}^{n} of the following form:

C(u,r)={xn:[x,u]r}={xn:|Pux|r}.C^{-}(u,r)=\big{\{}x\in\mathbb{R}^{n}:[x,u]\leq r\big{\}}=\big{\{}x\in\mathbb{R}^{n}:|\textrm{P}_{u^{\perp}}x|\leq r\big{\}}.

Due to (2.13), one sees that if K𝒦enK\in\mathscr{K}_{e}^{n}, then KK^{\circ} is obtained by the intersection of slabs. However, it follows from (2.2) and (3.5) that

K\displaystyle K^{\diamond}\!\! =\displaystyle= {xn:|y||Pyx|1forallyK{o}}\displaystyle\!\!\Big{\{}x\in\mathbb{R}^{n}:|y|\cdot|\textrm{P}_{y^{\perp}}x|\leq 1\ \ \mathrm{for\ all}\ y\in K\setminus\{o\}\Big{\}} (3.7)
=\displaystyle= {xn:|Pyx|1|y|forallyK{o}}.\displaystyle\!\!\Big{\{}x\in\mathbb{R}^{n}:|\textrm{P}_{y^{\perp}}x|\leq\frac{1}{|y|}\ \ \mathrm{for\ all}\ y\in K\setminus\{o\}\Big{\}}.

This means that KK^{\diamond} is indeed formed by the intersection of closed solid cylinders (with axis being the line {ty:t}\{ty:t\in\mathbb{R}\} and base radius being |y|1|y|^{-1} for yK{o}y\in K\setminus\{o\}). As the polar body plays fundamental roles in convex geometry, such as in the celebrated Blaschke-Santaló inequality (1.1) and many other affine isoperimetric inequalities (see, e.g., [32]), we expect that the sine polar body will play roles similar to its “cosine cousin” in applications. In particular, as a start, the sine Blaschke-Santaló inequality for the sine polar body will be established in Section 5.

In view of Definition 3.1 and the nature of the sine polar body, it is worth to investigate 𝒞en𝒦en\mathcal{C}_{e}^{n}\subseteq\mathscr{K}_{e}^{n}, the family of origin-symmetric convex bodies generated by the intersection of origin-symmetric closed solid cylinders. Typical examples of 𝒞en\mathcal{C}_{e}^{n} are the Steinmetz solids, see Figure 1. As pointed out by one of the referees, our sine polarity turns out to be a special case of the notion of polarity with respect to a general function c(,)c(\cdot,\cdot) introduced by Artstein-Avidan, Sadovsky, and Wyczesany [2] (namely, by letting c(x,y)=[x,y]c(x,y)=-[x,y] and t=1t=-1). Consequently, 𝒞en\mathcal{C}_{e}^{n} is a type of the cc-class of sets in [2]. These concepts were further studied in [3].

It is indeed a surprise that to generate a convex body K𝒞enK\in\mathcal{C}_{e}^{n}, as few as only two closed solid cylinders whose axes are not parallel to each other would be enough. In other words, one needs the directions of the axes of the closed solid cylinders generating K𝒞enK\in\mathcal{C}_{e}^{n} being not concentrated on an 11-dimensional subspace. This result is summarized in the following proposition.

Proposition 3.2.

Let r1,r2>0r_{1},r_{2}>0 be two constants and directions u,vSn1u,v\in S^{n-1} such that uu is not parallel to vv. Then C(u,r1)C(v,r2)𝒞en.C^{-}(u,r_{1})\cap C^{-}(v,r_{2})\in\mathcal{C}_{e}^{n}.

Proof.

Let K=C(u,r1)C(v,r2)K=C^{-}(u,r_{1})\cap C^{-}(v,r_{2}). To prove K𝒞enK\in\mathcal{C}_{e}^{n}, it would be enough to prove K𝒦enK\in\mathscr{K}_{e}^{n}.

First of all, it is clear that KK is an origin-symmetric closed convex set with nonempty interior. Suppose that KK is unbounded. Then there exists a direction wSn1w\in S^{n-1} such that {τw:τ>0}K\{\tau w:\tau>0\}\subseteq K. In particular, [τw,u]r1[\tau w,u]\leq r_{1} for all τ>0\tau>0. This implies 0[w,u]r1/τ00\leq[w,u]\leq r_{1}/\tau\rightarrow 0 as τ\tau\rightarrow\infty. Thus, [w,u]=|Puw|=0[w,u]=|\textrm{P}_{u^{\perp}}w|=0, and uu is parallel to ww. Similarly, vv is parallel to ww, and thus uu is parallel to vv, a contradiction. So K=C(u,r1)C(v,r2)K=C^{-}(u,r_{1})\cap C^{-}(v,r_{2}) is bounded and then K𝒦enK\in\mathscr{K}_{e}^{n} as desired. ∎

Basic properties for the sine polar body can be listed as follows.

Proposition 3.3.

Let KK be a subset in n\mathbb{R}^{n} with n2n\geq 2. Then the following properties hold. (i) Suppose LnL\subseteq\mathbb{R}^{n} such that KLK\subseteq L. Then LKL^{\diamond}\subseteq K^{\diamond}. (ii) The sine polar body of BnB^{n} is again BnB^{n}, that is, (Bn)=Bn(B^{n})^{\diamond}=B^{n}. (iii) For any OO(n)O\in\textrm{O}(n) and c>0c>0, one has, (OK)=OK(OK)^{\diamond}=OK^{\diamond} and (cK)=c1K(cK)^{\diamond}=c^{-1}\cdot K^{\diamond}. (iv) If KnK\subset\mathbb{R}^{n} is bounded and is not concentrated on any 11-dimensional subspace, then K𝒞enK^{\diamond}\in\mathcal{C}_{e}^{n}. (v) Let K=(K).K^{\diamond\diamond}=(K^{\diamond})^{\diamond}. Then KKK\subseteq K^{\diamond\diamond}. Moreover, K=KK=K^{\diamond\diamond} for all K𝒞enK\in\mathcal{C}_{e}^{n}. (vi) For any K𝒞enK\in\mathcal{C}_{e}^{n}, one has

V(K)V(K).V(K^{\diamond})\leq V(K^{\circ}). (3.8)

We omit the proofs of properties (i)-(v), since they are easy to get. Properties (vi) is vital to prove the sine Blaschke-Santaló inequality, which will be proved in the end of this section. Note that properties (i) and (v) have also appeared in [2, Lemma 3.10] for cc-polarity.

Recall that KK^{\circ\circ} is the smallest convex body that contains a compact subset KnK\subset\mathbb{R}^{n} with the origin in its interior, namely, KK^{\circ\circ} is the convex hull of KK. Motivated by Proposition 3.3, the following cylindrical hull of a bounded set may be proposed.

Definition 3.4.

Let EnE\subset\mathbb{R}^{n} be bounded and not concentrated on any 11-dimensional subspace. The cylindrical hull of EE, denote by cyl(E)\mathrm{cyl}(E), is a convex body in n\mathbb{R}^{n} of the following form:

cyl(E)={C:CisaclosedsolidcylindersuchthatEC}.\mathrm{cyl}(E)=\bigcap\big{\{}C:\ C\ \mathrm{is\ a\ closed\ solid\ cylinder\ such\ that}\ E\subset C\big{\}}.

If EnE\subset\mathbb{R}^{n} is bounded and is not concentrated on any 11-dimensional subspace, then Ecyl(E)E\subseteq\mathrm{cyl}(E) and cyl(E)𝒞en.\mathrm{cyl}(E)\in\mathcal{C}_{e}^{n}. Indeed, as EE is not concentrated on any 11-dimensional subspace, there must be two nonparallel vectors x,yn{o}x,y\in\mathbb{R}^{n}\setminus\{o\} such that x,yEx,y\in E. It further yields that all the base radii of closed solid cylinders containing EE are positive. Therefore, cyl(E)𝒞en\textrm{cyl}(E)\in\mathcal{C}_{e}^{n}. Since cyl(E)\textrm{cyl}(E) is a convex body in n\mathbb{R}^{n} containing EE, one also has cyl(E)conv(E)\mathrm{cyl}(E)\supseteq\mathrm{conv}(E), as the convex hull conv(E)\mathrm{conv}(E) is the smallest convex body containing EE.

The following result asserts that KK^{\diamond\diamond} is exactly the cylindrical hull of KK. Again, the cylindrical hull turns out to be a special case of the cc-envelope defined in [2].

Proposition 3.5.

Let KnK\subset\mathbb{R}^{n} be bounded and not concentrated on any 11-dimensional subspace. Then the following results hold. (i) K𝒞enK\in\mathcal{C}_{e}^{n} if and only if K=cyl(K)K=\mathrm{cyl}(K). (ii) K=cyl(K).K^{\diamond\diamond}=\mathrm{cyl}(K).

Proof.

(i) Let K𝒞enK\in\mathcal{C}_{e}^{n}. Note that Kcyl(K)K\subseteq\mathrm{cyl}(K). Hence, K=cyl(K)K=\mathrm{cyl}(K) follows immediately if Kcyl(K)K\supseteq\mathrm{cyl}(K) is verified. This is an easy consequence of K𝒞enK\in\mathcal{C}_{e}^{n} as KK is an intersection of closed solid cylinders (of course) containing KK.

(ii) First of all, it follows from Proposition 3.3 (iv) that K𝒞enK^{\diamond}\in\mathcal{C}_{e}^{n} and K𝒞enK^{\diamond\diamond}\in\mathcal{C}_{e}^{n}. Together with KKK\subseteq K^{\diamond\diamond} due to Proposition 3.3 (v), one gets that KK^{\diamond\diamond} is an intersection of closed solid cylinders containing KK. This implies that cyl(K)K.\mathrm{cyl}(K)\subseteq K^{\diamond\diamond}.

Now let us prove cyl(K)K.\mathrm{cyl}(K)\supseteq K^{\diamond\diamond}. By Proposition 3.3 (i), the fact that Kcyl(K)K\subseteq\mathrm{cyl}(K) implies Kcyl(K)K^{\diamond}\supseteq\mathrm{cyl}(K)^{\diamond}. From Proposition 3.3 (i), (v) and the fact that cyl(K)𝒞en\mathrm{cyl}(K)\in\mathcal{C}_{e}^{n}, we further get Kcyl(K)=cyl(K)K^{\diamond\diamond}\subseteq\mathrm{cyl}(K)^{\diamond\diamond}=\mathrm{cyl}(K) as desired. ∎

Recall that any convex body K𝒦enK\in\mathscr{K}_{e}^{n} can be formed by

K=uSn1{xn:|xu|hK(u)}.K=\bigcap_{u\in S^{n-1}}\Big{\{}x\in\mathbb{R}^{n}:|x\cdot u|\leq h_{K}(u)\Big{\}}.

Note that hK(u)=maxyKuyh_{K}(u)=\max_{y\in K}u\cdot y is the distance from the origin to the tangent hyperplane of KK at the direction uSn1u\in S^{n-1}. Likewise, we can propose a definition for the cylindrical support function.

Definition 3.6.

The cylindrical support function cK:n[0,)c_{K}:\mathbb{R}^{n}\rightarrow[0,\infty) of K𝒞enK\in\mathcal{C}_{e}^{n} is defined by

cK(x)=maxyK[x,y],xn.c_{K}(x)=\max_{y\in K}\,[x,y],\ \ \ x\in\mathbb{R}^{n}. (3.9)

The function cKc_{K} is an even function and has positive homogeneity of degree 11 in the sense that cK(rx)=rcK(x)c_{K}(rx)=r\cdot c_{K}(x) for all r0r\geq 0 and xnx\in\mathbb{R}^{n}. Moreover, cKc_{K} is convex, that is

cK(τx+(1τ)x)τcK(x)+(1τ)cK(x)c_{K}(\tau x+(1-\tau)x^{\prime})\leq\tau c_{K}(x)+(1-\tau)c_{K}(x^{\prime})

for any x,xnx,x^{\prime}\in\mathbb{R}^{n} and τ[0,1]\tau\in[0,1]. Indeed, by (3.1), one has

cK(τx+(1τ)x)\displaystyle c_{K}(\tau x+(1-\tau)x^{\prime})\!\! =\displaystyle= maxyK[τx+(1τ)x,y]\displaystyle\!\!\max_{y\in K}\,[\tau x+(1-\tau)x^{\prime},y]
\displaystyle\leq maxyK(τ[x,y]+(1τ)[x,y])\displaystyle\!\!\max_{y\in K}\Big{(}\tau[x,y]+(1-\tau)[x^{\prime},y]\Big{)}
\displaystyle\leq τmaxyK[x,y]+(1τ)maxyK[x,y]\displaystyle\!\!\tau\max_{y\in K}\,[x,y]+(1-\tau)\max_{y\in K}\,[x^{\prime},y]
=\displaystyle= τcK(x)+(1τ)cK(x).\displaystyle\!\!\tau c_{K}(x)+(1-\tau)c_{K}(x^{\prime}).

Denote by K\partial K the boundary of KK. The cylindrical support function can be used to form K𝒞enK\in\mathcal{C}_{e}^{n}.

Proposition 3.7.

For any K𝒞enK\in\mathcal{C}_{e}^{n}, one has

K=uSn1C(u,cK(u)).K=\bigcap_{u\in S^{n-1}}C^{-}(u,c_{K}(u)). (3.10)

Moreover, for any xKx\in\partial K, there must exist u0Sn1u_{0}\in S^{n-1} such that

[x,u0]=cK(u0).[x,u_{0}]=c_{K}(u_{0}). (3.11)
Proof.

It follows from (3.9) that, for a given uSn1u\in S^{n-1}, [y,u]cK(u)[y,u]\leq c_{K}(u) holds for any yKy\in K. Thus, KC(u,cK(u))K\subseteq C^{-}(u,c_{K}(u)) and KuSn1C(u,cK(u)).K\subseteq\bigcap_{u\in S^{n-1}}C^{-}(u,c_{K}(u)).

Assume that KuSn1C(u,cK(u)).K\subsetneq\bigcap_{u\in S^{n-1}}C^{-}(u,c_{K}(u)). Then there exists xuSn1C(u,cK(u))x\in\bigcap_{u\in S^{n-1}}C^{-}(u,c_{K}(u)) but xKx\notin K. Note that K𝒞enK\in\mathcal{C}_{e}^{n} is a convex body obtained by the intersection of closed solid cylinders, say

K=uΩSn1C(u,r(u)),K=\bigcap_{u\in\Omega\subseteq S^{n-1}}C^{-}(u,r(u)),

where r(u)r(u) depends on uΩu\in\Omega. Thus, xKx\notin K means that there exists u0Ωu_{0}\in\Omega such that [x,u0]>r(u0)[x,u_{0}]>r(u_{0}). Note that [x,u0]cK(u0)[x,u_{0}]\leq c_{K}(u_{0}) and thus r(u0)<cK(u0)r(u_{0})<c_{K}(u_{0}). As KC(u0,r(u0))K\subset C^{-}(u_{0},r(u_{0})), then

maxyK[u0,y]r(u0)<cK(u0),\max_{y\in K}\,[u_{0},y]\leq r(u_{0})<c_{K}(u_{0}),

which contradicts with the maximality of cK(u0)c_{K}(u_{0}). Hence (3.10) follows.

Now let us prove (3.11) for xKx\in\partial K and some u0Sn1u_{0}\in S^{n-1}. As K𝒞enK\in\mathcal{C}_{e}^{n} is an origin-symmetric convex body, then the convex function cKc_{K} restricted on Sn1S^{n-1} is continuous and bounded. Moreover, xKx\in\partial K implies that λxK\lambda x\notin K for all λ>1\lambda>1. Thus, it follows from (3.10) that there exists uλSn1u_{\lambda}\in S^{n-1} such that [λx,uλ]>cK(uλ).[\lambda x,u_{\lambda}]>c_{K}(u_{\lambda}). As Sn1S^{n-1} is compact and cKc_{K} restricted on Sn1S^{n-1} is continuous, one can select a sequence λm=1+1m1\lambda_{m}=1+\frac{1}{m}\rightarrow 1 whose corresponding umu_{m} satisfies umu0u_{m}\rightarrow u_{0} for some u0Sn1u_{0}\in S^{n-1} and cK(um)cK(u0)c_{K}(u_{m})\rightarrow c_{K}(u_{0}) as mm\rightarrow\infty. It then follows that

[x,u0]=limm[λmx,um]limmcK(um)=cK(u0).[x,u_{0}]=\lim_{m\rightarrow\infty}[\lambda_{m}x,u_{m}]\geq\lim_{m\rightarrow\infty}c_{K}(u_{m})=c_{K}(u_{0}).

On the other hand, [x,u0]cK(u0)[x,u_{0}]\leq c_{K}(u_{0}) as xK=uSn1C(u,cK(u)).x\in K=\bigcap_{u\in S^{n-1}}C^{-}(u,c_{K}(u)). This concludes that if xKx\in\partial K, there must have u0Sn1u_{0}\in S^{n-1} such that [x,u0]=cK(u0)[x,u_{0}]=c_{K}(u_{0}). ∎

Definition 3.8.

Let K𝒞enK\in\mathcal{C}_{e}^{n} with n2n\geq 2. For uSn1u\in S^{n-1}, define the supporting cylinder of KK at direction uu by

C(u,cK(u))={zn:[z,u]=cK(u)}.C(u,c_{K}(u))=\big{\{}z\in\mathbb{R}^{n}:[z,u]=c_{K}(u)\big{\}}.

Note that cK(u)c_{K}(u) is just the distance from the origin to the supporting cylinder C(u,cK(u))C(u,c_{K}(u)). Moreover, the proof of Proposition 3.7 also gives cK(u)=min{r>0:KC(u,r)}.c_{K}(u)=\min\big{\{}r>0:K\subseteq C^{-}(u,r)\big{\}}. Hence, cK(u)c_{K}(u) is the minimal base radius of the cylinder with axis {tu:t}\{tu:t\in\mathbb{R}\}. From (3.10) and its proof, if K𝒞enK\in\mathcal{C}_{e}^{n} and xKx\in\partial K, then there must have (at least) one supporting cylinder, say C(u,cK(u))C(u,c_{K}(u)), containing xx. This observation can be used to define the so-called cylindrical Gauss image of KK and its reverse. More precisely, the cylindrical Gauss image of KK, denoted by ξK:KSn1\xi_{K}:\partial K\rightarrow S^{n-1}, is defined by, for xKx\in\partial K,

ξK(x)={uSn1:usatisfies[x,u]=cK(u)}.\xi_{K}(x)=\big{\{}u\in S^{n-1}:u\ \ \mathrm{satisfies}\ \ [x,u]=c_{K}(u)\big{\}}.

The reverse cylindrical Gauss image of KK, denoted by ξK1:Sn1K\xi_{K}^{-1}:S^{n-1}\rightarrow\partial K, is given by, for uSn1u\in S^{n-1},

ξK1(u)={xK:xsatisfies[u,x]=cK(u)}.\xi_{K}^{-1}(u)=\big{\{}x\in\partial K:x\ \ \mathrm{satisfies}\ \ [u,x]=c_{K}(u)\big{\}}.

Of course, both ξK\xi_{K} and ξK1\xi_{K}^{-1} may not be injective. Moreover, [u,ξK1(u)]=cK(u)[u,\xi_{K}^{-1}(u)]=c_{K}(u) and [x,ξK(x)]=cK(ξK(x))[x,\xi_{K}(x)]=c_{K}(\xi_{K}(x)), if ξK(x)\xi_{K}(x) and ξK1(u)\xi_{K}^{-1}(u) are both singleton sets.

By (2.12) and (2.14), the volume of KK^{\circ} for K𝒦onK\in\mathscr{K}_{o}^{n} can be calculated by

V(K)=ωnSn11hK(u)n𝑑u.V(K^{\circ})=\omega_{n}\int_{S^{n-1}}\frac{1}{h_{K}(u)^{n}}\,du. (3.12)

A similar result also holds for KK^{\diamond} if K𝒞enK\in\mathcal{C}_{e}^{n}.

Proposition 3.9.

Let K𝒞enK\in\mathcal{C}_{e}^{n} with n2n\geq 2. Then

ρK(u)cK(u)=ρK(u)cK(u)=1\rho_{K^{\diamond}}(u)\cdot c_{K}(u)=\rho_{K}(u)\cdot c_{K^{\diamond}}(u)=1

holds for all uSn1u\in S^{n-1}. Moreover,

V(K)=ωnSn11cK(u)n𝑑u.V(K^{\diamond})=\omega_{n}\int_{S^{n-1}}\frac{1}{c_{K}(u)^{n}}\,du. (3.13)
Proof.

By (3.9), one has, for a given uSn1u\in S^{n-1}, [y,u]cK(u)[y,u]\leq c_{K}(u) for any yKy\in K. Equivalently,

[y,ucK(u)]1forallyK.\Big{[}y,\frac{u}{c_{K}(u)}\Big{]}\leq 1\ \ \mathrm{for\ all}\ y\in K.

Hence ucK(u)K\frac{u}{c_{K}(u)}\in K^{\diamond} by Definition 3.1. It follows from (2.6) that ρK(u)1cK(u)\rho_{K^{\diamond}}(u)\geq\frac{1}{c_{K}(u)}. On the other hand, one can take y0ξK1(u)y_{0}\in\xi_{K}^{-1}(u). Therefore, [y0,ucK(u)]=1[y_{0},\frac{u}{c_{K}(u)}]=1 and [y0,tu]>1[y_{0},tu]>1 if t>1cK(u)t>\frac{1}{c_{K}(u)}. As y0Ky_{0}\in\partial K, it follows from Definition 3.1 that tuKtu\notin K^{\diamond} for any t>1cK(u)t>\frac{1}{c_{K}(u)}. The fact that ρK(u)uK\rho_{K^{\diamond}}(u)u\in\partial K^{\diamond} then yields ρK(u)1cK(u)\rho_{K^{\diamond}}(u)\leq\frac{1}{c_{K}(u)} and thus ρK(u)=1cK(u)\rho_{K^{\diamond}}(u)=\frac{1}{c_{K}(u)}. By Proposition 3.3 (v), one gets that, for uSn1u\in S^{n-1},

ρK(u)=ρK(u)=1cK(u).\rho_{K}(u)=\rho_{K^{\diamond\diamond}}(u)=\frac{1}{c_{K^{\diamond}}(u)}.

Finally, the volume of KK^{\diamond} can be calculated by

V(K)=ωnSn1ρK(u)n𝑑u=ωnSn11cK(u)n𝑑uV(K^{\diamond})=\omega_{n}\int_{S^{n-1}}\rho_{K^{\diamond}}(u)^{n}\,du=\omega_{n}\int_{S^{n-1}}\frac{1}{c_{K}(u)^{n}}\,du

due to (2.14).∎

After the above preparations, we now prove Proposition 3.3 (vi).

Proof of Proposition 3.3 (vi).

Recall that K=uSn1C(u,cK(u))K=\bigcap_{u\in S^{n-1}}C^{-}(u,c_{K}(u)) proved in (3.10). Then KC(u,cK(u))K\subset C^{-}(u,c_{K}(u)) for any fixed uSn1u\in S^{n-1} and [x,u]cK(u)[x,u]\leq c_{K}(u) for any xKx\in K. By (2.2), one must have

[x,u]=|Pux||Pvx|=|xv|[x,u]=|\textrm{P}_{u^{\perp}}x|\geq|\textrm{P}_{v}x|=|x\cdot v|

for any vSn1v\in S^{n-1} such that uvu\perp v. Taking the maximum over xKx\in K, one has

cK(u)hK(v),c_{K}(u)\geq h_{K}(v),

and thus, for any vSn1v\in S^{n-1},

Sn1v1cK(ζ)n𝑑ζ1hK(v)n,\int_{S^{n-1}\cap v^{\perp}}\frac{1}{c_{K}(\zeta)^{n}}\,d\zeta\leq\frac{1}{h_{K}(v)^{n}}, (3.14)

where dζd\zeta is the rotation invariant probability measure on Sn1vS^{n-1}\cap v^{\perp}. For any nonnegative continuous function f:Sn1f:S^{n-1}\rightarrow\mathbb{R}, [26, (2.22)] yields that

Sn1f(u)𝑑u=Sn1(Sn1vf(ζ)𝑑ζ)𝑑v.\int_{S^{n-1}}f(u)\,du=\int_{S^{n-1}}\bigg{(}\int_{S^{n-1}\cap v^{\perp}}f(\zeta)\,d\zeta\bigg{)}\,dv. (3.15)

By (3.13), (3.15), (3.14), and (3.12), one gets

V(K)\displaystyle V(K^{\diamond})\!\! =\displaystyle= ωnSn11cK(u)n𝑑u\displaystyle\!\!\omega_{n}\int_{S^{n-1}}\frac{1}{c_{K}(u)^{n}}\,du
=\displaystyle= ωnSn1(Sn1v1cK(ζ)n𝑑ζ)𝑑v\displaystyle\!\!\omega_{n}\int_{S^{n-1}}\bigg{(}\int_{S^{n-1}\cap v^{\perp}}\frac{1}{c_{K}(\zeta)^{n}}\,d\zeta\bigg{)}\,dv
\displaystyle\leq ωnSn11hK(v)n𝑑v,\displaystyle\!\!\omega_{n}\int_{S^{n-1}}\frac{1}{h_{K}(v)^{n}}\,dv,
=\displaystyle= V(K).\displaystyle\!\!V(K^{\circ}).

This concludes the proof of (3.8). ∎

4 The LpL_{p}-sine Blaschke-Santaló inequality

In this section, we will prove the LpL_{p}-sine Blaschke-Santaló inequality (1.9). The following result is needed.

Proposition 4.1.

Let p1p\geq 1 and n2n\geq 2. Then

ΛpBn=Bn.\Lambda_{p}B^{n}=B^{n}. (4.1)
Proof.

Lemma 7.5 in [28] states that for an mm-dimensional subspace VV of n\mathbb{R}^{n}, one has

Sn1|PVu|p𝑑u=mωmωn+p2nωnωm+p2forp1,\displaystyle\int_{S^{n-1}}\big{|}\textrm{P}_{V}u\big{|}^{p}\,du=\frac{m\omega_{m}\omega_{n+p-2}}{n\omega_{n}\omega_{m+p-2}}\ \ \mathrm{for}\ \ p\geq 1, (4.2)

where PVu\textrm{P}_{V}u is the orthogonal projection of uu onto the subspace VV. Recall that [x,u]=|x||Pxu|[x,u]=|x|\cdot|\textrm{P}_{x^{\perp}}u| for all xn{o}x\in\mathbb{R}^{n}\setminus\{o\} (see (2.2)). Hence, if xn{o}x\in\mathbb{R}^{n}\setminus\{o\}, one can let V=xV=x^{\perp} whose dimension is m=n1m=n-1 and then (4.2) yields that for all p1p\geq 1,

Sn1[x,u]p𝑑u=|x|pSn1|Pxu|p𝑑u=((n1)ωn1ωn+p2nωnωn+p3)|x|p.\int_{S^{n-1}}[x,u]^{p}\,du=|x|^{p}\cdot\int_{S^{n-1}}|\textrm{P}_{x^{\perp}}u|^{p}\,du=\bigg{(}\frac{(n-1)\omega_{n-1}\omega_{n+p-2}}{n\omega_{n}\omega_{n+p-3}}\bigg{)}|x|^{p}.

Together with (1.7) and (1.8), one gets

hΛpBn(x)\displaystyle h_{\Lambda_{p}B^{n}}(x)\!\! =\displaystyle= (nωn(n+p)c~n,pV(Bn)Sn1[x,u]pρBn(u)n+p𝑑u)1/p\displaystyle\!\!\Big{(}\frac{n\omega_{n}}{(n+p)\widetilde{c}_{n,p}V(B^{n})}\int_{S^{n-1}}[x,u]^{p}\rho_{B^{n}}(u)^{n+p}\,du\Big{)}^{1/p}
=\displaystyle= (n(n+p)c~n,pSn1[x,u]p𝑑u)1/p\displaystyle\!\!\Big{(}\frac{n}{(n+p)\widetilde{c}_{n,p}}\int_{S^{n-1}}[x,u]^{p}\,du\Big{)}^{1/p}
=\displaystyle= |x|=hBn(x).\displaystyle\!\!|x|=h_{B^{n}}(x).

This concludes (4.1). ∎

Moreover, it follows from (3.3) and (4.1) that

Λp(cBn)=c1Λp(Bn)=c1(Bn)=c1Bn,foranyc>0.\Lambda_{p}^{\circ}(cB^{n})=c^{-1}\Lambda_{p}^{\circ}(B^{n})=c^{-1}(B^{n})^{\circ}=c^{-1}B^{n},\ \ \ \mathrm{for\ any}\ c>0. (4.3)
Lemma 4.2.

Let p1p\geq 1 and n2n\geq 2. If KnK\subset\mathbb{R}^{n} is an Ln+pL_{n+p}-star with V(K)>0V(K)>0, then

V(ΛpΛpK)V(K),V(\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)\geq V(K), (4.4)

with equality if and only if ρK(u)=ρΛpΛpK(u)\rho_{K}(u)=\rho_{\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K}(u) for almost all uSn1u\in S^{n-1} with respect to the spherical measure on Sn1S^{n-1}.

Proof.

Let K,LnK,L\subset\mathbb{R}^{n} be two Ln+pL_{n+p}-stars with positive volumes. As discussed in Section 3, both ΛpK𝒦en\Lambda_{p}^{\circ}K\in\mathscr{K}_{e}^{n} and ΛpL𝒦en\Lambda_{p}^{\circ}L\in\mathscr{K}_{e}^{n} are origin-symmetric convex bodies. It follows from (2.15), (3.2), and Fubini’s theorem that

V~p(K,ΛpL)V(K)\displaystyle\frac{\widetilde{V}_{-p}(K,\Lambda_{p}^{\circ}L)}{V(K)}\!\! =\displaystyle= ωnV(K)Sn1ρK(u)n+pρΛpL(u)p𝑑u\displaystyle\!\!\frac{\omega_{n}}{V(K)}\int_{S^{n-1}}\rho_{K}(u)^{n+p}\rho_{\Lambda_{p}^{\circ}L}(u)^{-p}\,du (4.5)
=\displaystyle= nωn2(n+p)c~n,pV(K)V(L)Sn1Sn1[u,v]pρK(u)n+pρL(v)n+p𝑑v𝑑u\displaystyle\!\!\frac{n\omega_{n}^{2}}{(n+p)\widetilde{c}_{n,p}V(K)V(L)}\int_{S^{n-1}}\int_{S^{n-1}}[u,v]^{p}\rho_{K}(u)^{n+p}\rho_{L}(v)^{n+p}\,dv\,du
=\displaystyle= ωnV(L)Sn1ρL(v)n+p(nωn(n+p)c~n,pV(K)Sn1[u,v]pρK(u)n+p𝑑u)𝑑v\displaystyle\!\!\frac{\omega_{n}}{V(L)}\int_{S^{n-1}}\rho_{L}(v)^{n+p}\Big{(}\frac{n\omega_{n}}{(n+p)\widetilde{c}_{n,p}V(K)}\int_{S^{n-1}}[u,v]^{p}\rho_{K}(u)^{n+p}\,du\Big{)}\,dv
=\displaystyle= ωnV(L)Sn1ρL(v)n+pρΛpK(v)p𝑑v\displaystyle\!\!\frac{\omega_{n}}{V(L)}\int_{S^{n-1}}\rho_{L}(v)^{n+p}\rho_{\Lambda_{p}^{\circ}K}(v)^{-p}\,dv
=\displaystyle= V~p(L,ΛpK)V(L).\displaystyle\!\!\frac{\widetilde{V}_{-p}(L,\Lambda_{p}^{\circ}K)}{V(L)}.

By letting L=ΛpKL=\Lambda_{p}^{\circ}K, (2.16) and (4.5) imply that

V~p(K,ΛpΛpK)V(K)\displaystyle\frac{\widetilde{V}_{-p}(K,\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)}{V(K)} =\displaystyle= V~p(ΛpK,ΛpK)V(ΛpK)=1.\displaystyle\frac{\widetilde{V}_{-p}(\Lambda_{p}^{\circ}K,\Lambda_{p}^{\circ}K)}{V(\Lambda_{p}^{\circ}K)}=1. (4.6)

That is, V~p(K,ΛpΛpK)=V(K).\widetilde{V}_{-p}(K,\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)=V(K). Together with (2.17), one has, for all p1p\geq 1,

V(K)n=V~p(K,ΛpΛpK)nV(K)n+pV(ΛpΛpK)p,V(K)^{n}=\widetilde{V}_{-p}(K,\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)^{n}\geq V(K)^{n+p}V(\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)^{-p},

which is exactly (4.4) after rearrangement. Clearly if ρK(u)=ρΛpΛpK(u)\rho_{K}(u)=\rho_{\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K}(u) for almost all uSn1u\in S^{n-1} with respect to the spherical measure of Sn1S^{n-1}, then V(ΛpΛpK)=V(K)V(\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)=V(K). On the other hand, if V(ΛpΛpK)=V(K)V(\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)=V(K) is assumed, then the equality conditions of (2.17) yield that there exists a constant c>0c>0 such that ρK(u)=cρΛpΛpK(u)\rho_{K}(u)=c\cdot\rho_{\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K}(u) for almost all uSn1u\in S^{n-1} with respect to the spherical measure of Sn1S^{n-1}. This further yields that V(ΛpΛpK)=V(K)=cnV(ΛpΛpK)V(\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)=V(K)=c^{n}V(\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K) and hence c=1c=1, as desired. ∎

Denote by [x1,,xn][x_{1},\cdots,x_{n}] the nn-dimensional volume of the parallelotope spanned by vectors x1,,xnnx_{1},\cdots,x_{n}\in\mathbb{R}^{n}. Let p1p\geq 1 and n2n\geq 2. We shall need Tp(K2,,Kn)𝒦en\mathrm{T}_{p}(K_{2},\cdots,K_{n})\in\mathscr{K}_{e}^{n} for K2,,Kn𝒦onK_{2},\cdots,K_{n}\in\mathscr{K}_{o}^{n}, whose support function at xnx\in\mathbb{R}^{n} was given in [27] (up to a factor) by

hTp(K2,,Kn)(x)p=1V(K2)V(Kn)K2Kn[x,x2,,xn]p𝑑x2𝑑xn.h_{\mathrm{T}_{p}(K_{2},\cdots,K_{n})}(x)^{p}=\frac{1}{V(K_{2})\cdots V(K_{n})}\int_{K_{2}}\cdots\int_{K_{n}}[x,x_{2},\cdots,x_{n}]^{p}\,dx_{2}\cdots\,dx_{n}.

In particular, if we let K2=K𝒦onK_{2}=K\in\mathscr{K}_{o}^{n}, K3==Kn=BnK_{3}=\cdots=K_{n}=B^{n}, then [27, Theorem 4.3] asserts that

Tp(K,Bn,,Bnn2)=dn,pΛpK\mathrm{T}_{p}(K,\underbrace{B^{n},\cdots,B^{n}}_{n-2})=d_{n,p}\cdot\Lambda_{p}K

for all p1p\geq 1 with dn,pp=c~n,pd_{n,p}^{p}=\widetilde{c}_{n,p} for n=2n=2 and

(n+p)n2dn,ppnn2c~n,p\displaystyle\frac{(n+p)^{n-2}d_{n,p}^{p}}{n^{n-2}\widetilde{c}_{n,p}}\!\! =\displaystyle= Sn1Sn1([x,u2,,un][x,u2])p𝑑u2𝑑un\displaystyle\!\!\int_{S^{n-1}}\cdots\int_{S^{n-1}}\bigg{(}\frac{[x,u_{2},\cdots,u_{n}]}{[x,u_{2}]}\bigg{)}^{p}\,du_{2}\cdots\,du_{n} (4.7)
=\displaystyle= i=2n1Sn1|PViui+1|p𝑑ui+1=i=2n1(ni)ωniωn+p2nωnωni+p2\displaystyle\!\!\prod_{i=2}^{n-1}\int_{S^{n-1}}\big{|}\textrm{P}_{V_{i}^{\perp}}u_{i+1}\big{|}^{p}\,du_{i+1}=\prod_{i=2}^{n-1}\frac{(n-i)\omega_{n-i}\omega_{n+p-2}}{n\omega_{n}\omega_{n-i+p-2}}

for n3n\geq 3. The second equality in (4.7) follows from [27, (3.7)], where the dimensions of subspaces ViV_{i}^{\perp} are nin-i for i=2,,ni=2,\cdots,n. (We would like to point out that dvi\,dv_{i} in [27, Lemma 3.2] are equal to duidu_{i} in the present paper multiplying nωnn\omega_{n}.) The last equality in (4.7) follows from formula (4.2). Therefore, by (1.8), one has

dn,p=(i=1n1(ni)ωniωn+p2(n+p)ωnωni+p2)1/p.d_{n,p}=\bigg{(}\prod_{i=1}^{n-1}\frac{(n-i)\omega_{n-i}\omega_{n+p-2}}{(n+p)\omega_{n}\omega_{n-i+p-2}}\bigg{)}^{1/p}.

By (2.10), we further have, for p1p\geq 1 and n2n\geq 2,

Tp(K,Bn,,Bnn2)=dn,p1ΛpK.\displaystyle\mathrm{T}_{p}^{\circ}(K,\underbrace{B^{n},\cdots,B^{n}}_{n-2})=d_{n,p}^{-1}\cdot\Lambda_{p}^{\circ}K. (4.8)

With a different normalization, the following inequality was established by Haddad in [21, Theorem 1.3], which is equivalent to the LpL_{p}-Busemann random simplex inequality [7, Theorem 1.3]. By a different approach, inequality (4.9) for star bodies and for p=2p=2 was also established in [27, Corollary 4.5].

Lemma 4.3.

Let p1p\geq 1 and n2n\geq 2. If K2,,Kn𝒦onK_{2},\cdots,K_{n}\in\mathscr{K}_{o}^{n}, then

V(K2)V(Kn)V(Tp(K2,,Kn))(nωnn+p(n+p)ωn+pn(j=1n(p+j)ωp+jjωj))np,V(K_{2})\cdots V(K_{n})\cdot V(\mathrm{T}_{p}^{\circ}(K_{2},\cdots,K_{n}))\leq\bigg{(}\frac{n\omega_{n}^{n+p}}{(n+p)\omega_{n+p}^{n}}\Big{(}\prod_{j=1}^{n}\frac{(p+j)\omega_{p+j}}{j\omega_{j}}\Big{)}\bigg{)}^{\frac{n}{p}}, (4.9)

with equality if and only if K2,,KnK_{2},\cdots,K_{n} are origin-symmetric ellipsoids that are dilates.

Before the proof of Theorem 4.4, we shall mention the LpL_{p}-sine centroid body and the sine polar body for the case n=2n=2. Notice that for x,y2x,y\in\mathbb{R}^{2},

[x,y]=|x||Pxy|=|ψπ/2x||Pψπ/2xy|=|ψπ/2xy|,[x,y]=|x|\cdot|\textrm{P}_{x^{\perp}}y|=|\psi_{\pi/2}x|\cdot|\textrm{P}_{\psi_{\pi/2}x}y|=|\psi_{\pi/2}x\cdot y|, (4.10)

where

ψπ/2=(0110)\psi_{\pi/2}=\left(\begin{array}[]{ccc}0&-1\\ 1&0\\ \end{array}\right)

is a rotation by the angle of π/2\pi/2 in 2\mathbb{R}^{2}. For K𝒦o2K\in\mathscr{K}_{o}^{2}, it follows from (1.4) and (1.8) that c2,p=c~2,pc_{2,p}=\widetilde{c}_{2,p}. By (1.7), (4.10), (1.3), and (2.4), we have, for p1p\geq 1 and for x2x\in\mathbb{R}^{2},

hΛpK(x)p=1c~2,pV(K)K[x,y]p𝑑y=1c2,pV(K)K|ψπ/2xy|p𝑑y=hψπ/2tΓpK(x)p.h_{\Lambda_{p}K}(x)^{p}=\frac{1}{\widetilde{c}_{2,p}V(K)}\int_{K}[x,y]^{p}\,dy=\frac{1}{c_{2,p}V(K)}\int_{K}|\psi_{\pi/2}x\cdot y|^{p}\,dy=h_{\psi_{\pi/2}^{t}\Gamma_{p}K}(x)^{p}.

Together with (2.11) and the the facts that ψπ/21=ψπ/2\psi_{\pi/2}^{-1}=-\psi_{\pi/2} and ΓpK\Gamma_{p}^{\circ}K is origin-symmetric, one has

ΛpK=ψπ/2ΓpK.\Lambda_{p}^{\circ}K=\psi_{\pi/2}\Gamma_{p}^{\circ}K. (4.11)

Similar, for K𝒦e2K\in\mathscr{K}_{e}^{2}, it follows from (3.5), (2.13), and (4.10) that

K=ψπ/2K.K^{\diamond}=\psi_{\pi/2}K^{\circ}. (4.12)

Due to the affine natures of ΓpK\Gamma_{p}^{\circ}K and KK^{\circ}, it is easy to verify that in 2\mathbb{R}^{2}

Λp(ϕK)=|detϕ|1ϕΛpK,\Lambda_{p}^{\circ}(\phi K)=|\det\phi|^{-1}\phi\Lambda_{p}^{\circ}K, (4.13)

and

(ϕK)=|detϕ|1ϕK.(\phi K)^{\diamond}=|\det\phi|^{-1}\phi K^{\diamond}. (4.14)

Therefore, the volume product V(K)V(ΛpK)V(K)V(\Lambda_{p}^{\circ}K) is GL(2)\textrm{GL}(2)-invariant for K𝒦o2K\in\mathscr{K}_{o}^{2} and the volume product V(K)V(K)V(K)V(K^{\diamond}) is GL(2)\textrm{GL}(2)-invariant for K𝒦e2K\in\mathscr{K}_{e}^{2}. The volume products V(K)V(ΛpK)V(K)V(\Lambda_{p}^{\circ}K) and V(K)V(K)V(K)V(K^{\diamond}), however, is no longer GL(n)\textrm{GL}(n)-invariant for n3n\geq 3. This is a major difference between equality conditions in inequality (1.2), (1.1) and its sine counterpart (1.9), (1.10).

We are now in a position to prove the LpL_{p}-sine Blaschke-Santaló inequality (1.9) in the following theorem. For p=2p=2 and KK being a star body, one recovers [27, Theorem 4.1].

Theorem 4.4.

Let p1p\geq 1 and n2n\geq 2. If KnK\subset\mathbb{R}^{n} is an Ln+pL_{n+p}-star with V(K)>0V(K)>0, then

V(K)V(ΛpK)ωn2,V(K)V(\Lambda_{p}^{\circ}K)\leq\omega_{n}^{2}, (4.15)

with equality if and only if KK, up to sets of measure 0, is an origin-symmetric ellipsoid when n=2n=2, and is an origin-symmetric ball when n3n\geq 3.

Proof.

Note that when n=2n=2 and K𝒦o2K\in\mathscr{K}_{o}^{2}, it follows from (4.11) that inequality (4.15) is exactly the LpL_{p} Blaschke-Santaló inequality (1.2).

Now assume that n3n\geq 3 and K𝒦onK\in\mathscr{K}_{o}^{n}. Taking K2=KK_{2}=K and K3==Kn=BnK_{3}=\cdots=K_{n}=B^{n} in (4.9), one gets

V(K)V(Bn)V(Bn)n2V(Tp(K,Bn,Bnn2))(nωnn+p(n+p)ωn+pn(j=1n(p+j)ωp+jjωj))np,V(K)\cdot\underbrace{V(B^{n})\cdots V(B^{n})}_{n-2}\cdot V(\mathrm{T}_{p}^{\circ}(K,\underbrace{B^{n}\cdots,B^{n}}_{n-2}))\leq\bigg{(}\frac{n\omega_{n}^{n+p}}{(n+p)\omega_{n+p}^{n}}\Big{(}\prod_{j=1}^{n}\frac{(p+j)\omega_{p+j}}{j\omega_{j}}\Big{)}\bigg{)}^{\frac{n}{p}},

with equality if and only if K2=KK_{2}=K and K3==Kn=BnK_{3}=\cdots=K_{n}=B^{n} are origin-symmetric ellipsoids that are dilates (of course, this implies that KK is an origin-symmetric ball).

Together with (4.8), one has, for n3n\geq 3,

V(K)V(ΛpK)\displaystyle V(K)V(\Lambda_{p}^{\circ}K)\!\! \displaystyle\leq dn,pnωnn2(nωnn+p(n+p)ωn+pn(j=1n(p+j)ωp+jjωj))np\displaystyle\!\!\frac{d_{n,p}^{n}}{\omega_{n}^{n-2}}\bigg{(}\frac{n\omega_{n}^{n+p}}{(n+p)\omega_{n+p}^{n}}\Big{(}\prod_{j=1}^{n}\frac{(p+j)\omega_{p+j}}{j\omega_{j}}\Big{)}\bigg{)}^{\frac{n}{p}}
=\displaystyle= ωn2((p+1)(p+n)ωn+p1ωn+p2n(n+p)nωpωp1ωn+pn1)np\displaystyle\!\!\omega_{n}^{2}\bigg{(}\frac{(p+1)\cdots(p+n)\omega_{n+p-1}\omega_{n+p-2}^{n}}{(n+p)^{n}\omega_{p}\omega_{p-1}\omega_{n+p}^{n-1}}\bigg{)}^{\frac{n}{p}}
=\displaystyle= ωn2,\displaystyle\!\!\omega_{n}^{2},

with equality if and only if KK is an origin-symmetric ball.

In the above, we have established inequality (4.15) for all n2n\geq 2 and K𝒦onK\in\mathscr{K}_{o}^{n}. Now we consider KK being an Ln+pL_{n+p}-star with V(K)>0V(K)>0. Together with (4.4) and the fact that ΛpK𝒦on\Lambda_{p}^{\circ}K\in\mathscr{K}_{o}^{n}, we have

V(K)V(ΛpK)V(ΛpΛpK)V(ΛpK)ωn2.V(K)V(\Lambda_{p}^{\circ}K)\leq V(\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)V(\Lambda_{p}^{\circ}K)\leq\omega_{n}^{2}. (4.16)

Hence, the desired inequality (4.15) holds for n2n\geq 2 and for any Ln+pL_{n+p}-star KK with V(K)>0V(K)>0.

Now let us characterize the equality of (4.15). Equality holding in the second inequality of (4.16) yields that ΛpK\Lambda_{p}^{\circ}K is an origin-symmetric ellipsoid when n=2n=2 and is an origin-symmetric ball when n3n\geq 3. For n=2n=2, we may assume ΛpK=ϕB2\Lambda_{p}^{\circ}K=\phi B^{2} for some ϕGL(2)\phi\in\mathrm{GL}(2). By (4.13) and (4.3), one has

ΛpΛpK=Λp(ϕB2)=|detϕ|1ϕΛpB2=|detϕ|1ϕB2.\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K=\Lambda_{p}^{\circ}(\phi B^{2})=|\det\phi|^{-1}\phi\Lambda_{p}^{\circ}B^{2}=|\det\phi|^{-1}\phi B^{2}. (4.17)

For n3n\geq 3, we may assume ΛpK=cBn\Lambda_{p}^{\circ}K=cB^{n} for some constant c>0c>0. By (4.3), one has

ΛpΛpK=Λp(cBn)=c1Bn.\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K=\Lambda_{p}^{\circ}(cB^{n})=c^{-1}B^{n}.

From Lemma 4.2, equality holding in the first inequality of (4.16) implies ρK(u)=ρΛpΛpK(u)\rho_{K}(u)=\rho_{\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K}(u) for almost all uSn1u\in S^{n-1} with respect to the spherical measure on Sn1S^{n-1}. That is, to have equality in (4.16), KK, up to sets of measure 0, is an origin-symmetric ellipsoid when n=2n=2, and is an origin-symmetric ball when n3n\geq 3. Conversely, due to (4.11) and (4.3), equality holds in (4.15) if KK, up to sets of measure 0, is an origin-symmetric ellipsoid when n=2n=2, and is an origin-symmetric ball when n3n\geq 3. ∎

5 The sine Blaschke-Santaló inequality

If KK is a star body in n\mathbb{R}^{n}, then KK is an Ln+pL_{n+p}-star for all p1p\geq 1 with V(K)>0V(K)>0. By taking pp\rightarrow\infty on (4.15) and (3.6), one gets the sine Blaschke-Santaló inequality (1.10):

V(K)V(K)=V(K)V(ΛK)=limpV(K)V(ΛpK)ωn2.V(K)V(K^{\diamond})=V(K)V(\Lambda_{\infty}^{\circ}K)=\lim_{p\rightarrow\infty}V(K)V(\Lambda_{p}^{\circ}K)\leq\omega_{n}^{2}. (5.1)

However, this approximation argument does not yield equality conditions of (5.1). The full characterization of equality conditions of (5.1) will be presented in Theorem 5.2.

The following lemma will be used to extend the sine Blaschke-Santaló inequality (5.1) from star bodies to general bounded and measurable sets in n\mathbb{R}^{n}.

Lemma 5.1.

If KK is a bounded and measurable set in n\mathbb{R}^{n} with n2n\geq 2, then

V(K)V(K).V(K^{\diamond\diamond})\geq V(K). (5.2)

If in addition KK is a star body, equality holds in (5.2) if and only if K=KK=K^{\diamond\diamond}.

Proof.

As KK is bounded, KK^{\diamond\diamond} is also bounded by Proposition 3.3 (i)-(iii). Proposition 3.3 (v) implies KKK\subseteq K^{\diamond\diamond} and hence V(K)V(K)V(K^{\diamond\diamond})\geq V(K). Now suppose V(K)=V(K)V(K^{\diamond\diamond})=V(K) and KK is a star body in n\mathbb{R}^{n}. It follows from (2.14) that

0=V(K)V(K)=ωnSn1(ρK(u)nρK(u)n)𝑑u.0=V(K^{\diamond\diamond})-V(K)=\omega_{n}\int_{S^{n-1}}\Big{(}\rho_{K^{\diamond\diamond}}(u)^{n}-\rho_{K}(u)^{n}\Big{)}\,du.

Since ρKρK\rho_{K^{\diamond\diamond}}\geq\rho_{K} and ρK\rho_{K} is continuous on Sn1S^{n-1}, this further implies that ρK(u)=ρK(u)\rho_{K^{\diamond\diamond}}(u)=\rho_{K}(u) for all uSn1u\in S^{n-1}, and thus K=KK^{\diamond\diamond}=K. ∎

We are now in a position to prove the sine Blaschke-Santaló inequality (1.10).

Theorem 5.2.

Let KK be a bounded and measurable set in n\mathbb{R}^{n} with n2n\geq 2 such that KK is not concentrated on any 11-dimensional subspace. Then

V(K)V(K)ωn2.V(K)V(K^{\diamond})\leq\omega_{n}^{2}. (5.3)

If in addition KK is a star body, equality holds in (5.3) if and only if KK is an origin-symmetric ellipsoid when n=2n=2 and is an origin-symmetric ball when n3n\geq 3.

Proof.

Since KnK\subset\mathbb{R}^{n} is bounded and is not concentrated on any 11-dimensional subspace, it follows from Proposition 3.3 (iv) that K𝒞enK^{\diamond}\in\mathcal{C}_{e}^{n} and V(K)<V(K^{\diamond})<\infty. Clearly, inequality (5.3) holds trivially if V(K)=0V(K)=0.

Assume that V(K)>0V(K)>0. Lemma 5.1 yields

V(K)V(K)V(K)V(K).V(K)V(K^{\diamond})\leq V(K^{\diamond})V(K^{\diamond\diamond}). (5.4)

If K𝒮nK\in\mathcal{S}^{n}, then equality in (5.4) holds if and only if K=KK^{\diamond\diamond}=K. Applying (3.8) for L=K𝒞enL=K^{\diamond}\in\mathcal{C}_{e}^{n} and the Blaschke-Santaló inequality (1.1) for KK^{\diamond}, one gets

V(K)V(K)V(K)V((K))ωn2.V(K^{\diamond})V(K^{\diamond\diamond})\leq V(K^{\diamond})V((K^{\diamond})^{\circ})\leq\omega_{n}^{2}. (5.5)

Thus, the desired inequality (5.3) immediately follows from (5.4) and (5.5).

Now let us characterize the equality of (5.3) under the assumption that KK is a star body. Clearly, from (4.12), Proposition 3.3 (ii) and (iii), equality holds in (5.3) if KK is an origin-symmetric ellipsoid when n=2n=2 and is an origin-symmetric ball when n3n\geq 3. Conversely, the equality conditions of the Blaschke-Santaló inequality (1.1) imply KK^{\diamond} is an origin-symmetric ellipsoid. When n=2n=2, we may assume K=ϕB2K^{\diamond}=\phi B^{2} for some ϕGL(2)\phi\in\mathrm{GL}(2). It follows from the equality conditions of (5.4), (4.14) and Proposition 3.3 (ii) that

K=K=(ϕB2)=(detϕ)1ϕ(B2)=(detϕ)1ϕB2,K=K^{\diamond\diamond}=(\phi B^{2})^{\diamond}=(\det\phi)^{-1}\phi(B^{2})^{\diamond}=(\det\phi)^{-1}\phi B^{2},

which is an origin-symmetric ellipsoid. To get the desired equality conditions for n3n\geq 3, we only need to show that equality holding in (5.3) for n3n\geq 3 yields that KK^{\diamond} is an origin-symmetric ball. Suppose that KK^{\diamond} is an origin-symmetric ellipsoid but is not an origin-symmetric ball. Without loss of generality, one can assume that

K=={x=(x1,,xn)n:x12a12+x22a22++xn2an21},K^{\diamond}=\mathscr{E}=\Big{\{}x=(x_{1},\cdots,x_{n})\in\mathbb{R}^{n}:\ \ \frac{x_{1}^{2}}{a_{1}^{2}}+\frac{x_{2}^{2}}{a_{2}^{2}}+\cdots+\frac{x_{n}^{2}}{a_{n}^{2}}\leq 1\Big{\}},

where a1a2ana_{1}\leq a_{2}\leq\cdots\leq a_{n} such that a1<ana_{1}<a_{n}. We separate the proof into two cases.

Case 1: there exists 1<k<n1<k<n such that a1<akana_{1}<a_{k}\leq a_{n}. In this case, we shall prove that 𝒞en\mathscr{E}\notin\mathcal{C}_{e}^{n}, which is a contradiction with the fact =K𝒞en\mathscr{E}=K^{\diamond}\in\mathcal{C}_{e}^{n}.

We argue by contradiction by assuming that 𝒞en\mathscr{E}\in\mathcal{C}_{e}^{n}. Note that a1e1a_{1}e_{1}\in\partial\mathscr{E}. It follows from Proposition 3.7 that there is some u0Sn1u_{0}\in S^{n-1} such that

c(u0)=[a1e1,u0]=a1[e1,u0]a1.c_{\mathscr{E}}(u_{0})=[a_{1}e_{1},u_{0}]=a_{1}[e_{1},u_{0}]\leq a_{1}. (5.6)

Let z=(z1,,zn)u0z=(z_{1},\cdots,z_{n})\in u_{0}^{\perp}\cap\partial\mathscr{E} such that zspan{e1,,ek1}z\notin\textrm{span}\{e_{1},\ldots,e_{k-1}\}, where {e1,,en}\{e_{1},\cdots,e_{n}\} is the canonical basis of n\mathbb{R}^{n}. Since zspan{e1,,ek1}z\notin\textrm{span}\{e_{1},\ldots,e_{k-1}\}, then at least one of zk,,znz_{k},\cdots,z_{n} is not zero. Together with zz\in\partial\mathscr{E} and a1<akana_{1}<a_{k}\leq a_{n}, one can check that

1=z12a12++zk2ak2++zn2an2<|z|2a12,1=\frac{z_{1}^{2}}{a_{1}^{2}}+\cdots+\frac{z_{k}^{2}}{a_{k}^{2}}+\cdots+\frac{z_{n}^{2}}{a_{n}^{2}}<\frac{|z|^{2}}{a_{1}^{2}},

which implies |z|>a1|z|>a_{1}. On the other hand, due to zu0z\in u_{0}^{\perp}, one gets [z,u0]=|z|>a1.[z,u_{0}]=|z|>a_{1}. It follows from (3.9) that

c(u0)=maxy[y,u0][z,u0]>a1.c_{\mathscr{E}}(u_{0})=\max_{y\in\mathscr{E}}[y,u_{0}]\geq[z,u_{0}]>a_{1}.

This contradicts to (5.6) and hence 𝒞en\mathscr{E}\notin\mathcal{C}_{e}^{n}.

Case 2: let a:=a1==an1a:=a_{1}=\cdots=a_{n-1} and b:=anb:=a_{n} with a<ba<b. Then

={x=(x1,,xn)n:x12a2++xn12a2+xn2b21}.\mathscr{E}=\Big{\{}x=(x_{1},\cdots,x_{n})\in\mathbb{R}^{n}:\ \ \frac{x_{1}^{2}}{a^{2}}+\cdots+\frac{x_{n-1}^{2}}{a^{2}}+\frac{x_{n}^{2}}{b^{2}}\leq 1\Big{\}}. (5.7)

It is well-known that

={x=(x1,,xn)n:a2x12++a2xn12+b2xn21}.\mathscr{E}^{\circ}=\Big{\{}x=(x_{1},\cdots,x_{n})\in\mathbb{R}^{n}:\ \ a^{2}x_{1}^{2}+\cdots+a^{2}x_{n-1}^{2}+b^{2}x_{n}^{2}\leq 1\Big{\}}. (5.8)

We now claim that

{x=(x1,,xn)n:b2x12++b2xn12+a2xn21}:=0.\mathscr{E}^{\diamond}\subseteq\Big{\{}x=(x_{1},\cdots,x_{n})\in\mathbb{R}^{n}:\ \ b^{2}x_{1}^{2}+\cdots+b^{2}x_{n-1}^{2}+a^{2}x_{n}^{2}\leq 1\Big{\}}:=\mathscr{E}_{0}. (5.9)

Let U0=span{en1,en}U_{0}=\textrm{span}\{e_{n-1},e_{n}\}. Then it follows from Definition 3.1 that

U0\displaystyle\mathscr{E}^{\diamond}\cap U_{0}\!\! =\displaystyle= {xnU0:maxy[x,y]1}\displaystyle\!\!\big{\{}x\in\mathbb{R}^{n}\cap U_{0}:\max_{y\in\mathscr{E}}[x,y]\leq 1\big{\}} (5.10)
\displaystyle\subseteq {xnU0:maxyU0[x,y]1}.\displaystyle\!\!\big{\{}x\in\mathbb{R}^{n}\cap U_{0}:\max_{y\in\mathscr{E}\cap U_{0}}[x,y]\leq 1\big{\}}.

In the 22-dimensional subspace U0U_{0}, U0\mathscr{E}^{\circ}\cap U_{0} is an ellipse of the following form:

U0={z=(z1,,zn)n:a2zn12+b2zn21,z1==zn2=0}.\mathscr{E}^{\circ}\cap U_{0}=\Big{\{}z=(z_{1},\cdots,z_{n})\in\mathbb{R}^{n}:\ \ a^{2}z_{n-1}^{2}+b^{2}z_{n}^{2}\leq 1,\ \ z_{1}=\cdots=z_{n-2}=0\Big{\}}.

As shown in (4.12), restricted on the 22-dimensional subspace U0U_{0}, the set

={xnU0:maxyU0[x,y]1}\mathcal{E}=\big{\{}x\in\mathbb{R}^{n}\cap U_{0}:\max\limits_{y\in\mathscr{E}\cap U_{0}}[x,y]\leq 1\big{\}}

is just a rotation of U0\mathscr{E}^{\circ}\cap U_{0} by an angle of π/2\pi/2, namely

={z=(z1,,zn)n:b2zn12+a2zn21,z1==zn2=0}=0U0.\mathcal{E}=\Big{\{}z=(z_{1},\cdots,z_{n})\in\mathbb{R}^{n}:\ \ b^{2}z_{n-1}^{2}+a^{2}z_{n}^{2}\leq 1,\ \ z_{1}=\cdots=z_{n-2}=0\Big{\}}=\mathscr{E}_{0}\cap U_{0}.

Together with (5.10), one concludes

U0=0U0.\mathscr{E}^{\diamond}\cap U_{0}\subseteq\mathcal{E}=\mathscr{E}_{0}\cap U_{0}. (5.11)

Let OnO_{n} be a rotation of the form

On=(On1001)O_{n}=\left(\begin{array}[]{ccc}O_{n-1}&0\\ 0&1\\ \end{array}\right) (5.12)

where On1O_{n-1} is an arbitrary rotation on the (n1)(n-1)-dimensional subspace span{e1,,en1}.\textrm{span}\{e_{1},\cdots,e_{n-1}\}. It can be checked by Proposition 3.3 (iii) that

On(U0)=On()OnU0=(On)OnU0=OnU0,O_{n}(\mathscr{E}^{\diamond}\cap U_{0})=O_{n}(\mathscr{E}^{\diamond})\cap O_{n}U_{0}=(O_{n}\mathscr{E})^{\diamond}\cap O_{n}U_{0}=\mathscr{E}^{\diamond}\cap O_{n}U_{0},

where the last equality follows from the definition of \mathscr{E} defined by (5.7). Similarly,

On(0U0)=On0OnU0=0OnU0,O_{n}(\mathscr{E}_{0}\cap U_{0})=O_{n}\mathscr{E}_{0}\cap O_{n}U_{0}=\mathscr{E}_{0}\cap O_{n}U_{0},

where the the last equality follows from the definition of 0\mathscr{E}_{0} defined by (5.9). Applying OnO_{n} on both sides of (5.11), one gets

OnU0=On(U0)On(0U0)=0OnU0.\mathscr{E}^{\diamond}\cap O_{n}U_{0}=O_{n}(\mathscr{E}^{\diamond}\cap U_{0})\subseteq O_{n}(\mathscr{E}_{0}\cap U_{0})=\mathscr{E}_{0}\cap O_{n}U_{0}.

Taking the union over all OnO_{n} of the form (5.12), one gets

=n=(On(OnU0))=On(OnU0)On(0OnU0)=0,\mathscr{E}^{\diamond}=\mathscr{E}^{\diamond}\cap\mathbb{R}^{n}=\mathscr{E}^{\diamond}\bigcap\Big{(}\bigcup_{O_{n}}\big{(}O_{n}U_{0}\big{)}\Big{)}=\bigcup_{O_{n}}\Big{(}\mathscr{E}^{\diamond}\cap O_{n}U_{0}\Big{)}\subseteq\bigcup_{O_{n}}\Big{(}\mathscr{E}_{0}\cap O_{n}U_{0}\Big{)}=\mathscr{E}_{0},

which concludes the proof of (5.9). Thus, it follows from (5.8) and (5.9) that

V()V(0)=1bn1aωnandV()=1an1bωn.V(\mathscr{E}^{\diamond})\leq V(\mathscr{E}_{0})=\frac{1}{b^{n-1}a}\omega_{n}\quad\textrm{and}\quad V(\mathscr{E}^{\circ})=\frac{1}{a^{n-1}b}\omega_{n}.

Since n3n\geq 3 and b>ab>a, one clearly has V()<V()V(\mathscr{E}^{\diamond})<V(\mathscr{E}^{\circ}). It means V(K)<V((K))V(K^{\diamond\diamond})<V((K^{\diamond})^{\circ}) since K=K^{\diamond}=\mathscr{E}. Thus, from (5.4) and (5.5), equality cannot hold in inequality (5.3). ∎

6 Related functional inequalities

As applications of the LpL_{p}-sine Blaschke-Santaló inequality (4.15), we will provide some functional inequalities related to it. These functional inequalities can be proved along the routine approach in the literature, so we only give a detailed proof for Theorem 6.1 and omit the proofs for Theorems 6.2 and 6.3.

The following inequality is equivalent to the LpL_{p}-sine Blaschke-Santaló inequality (4.15).

Theorem 6.1.

Let p1p\geq 1 and n2n\geq 2. If KK and LL are Ln+pL_{n+p}-stars in n\mathbb{R}^{n}, then

KL[x,y]p𝑑x𝑑yn(n1)ωn1ωn+p2(n+p)2ωn+p3ωn1+2p/n[V(K)V(L)]n+pn.\int_{K}\int_{L}[x,y]^{p}\,dx\,dy\geq\frac{n(n-1)\omega_{n-1}\omega_{n+p-2}}{(n+p)^{2}\omega_{n+p-3}\omega_{n}^{1+2p/n}}\big{[}V(K)V(L)\big{]}^{\frac{n+p}{n}}. (6.1)

If V(K)V(L)>0V(K)V(L)>0, equality holds in (6.1) if and only if KK and LL, up to sets of measure 0, are dilates of an origin-symmetric ellipsoid when n=2n=2 and are origin-symmetric balls when n3n\geq 3.

Proof.

Without loss of generality, we assume that V(K)V(L)>0V(K)V(L)>0. From (4.5), one has

Sn1Sn1[u,v]pρK(u)n+pρL(v)n+p𝑑u𝑑v=(n+p)c~n,pV(L)V~p(K,ΛpL)nωn2.\int_{S^{n-1}}\int_{S^{n-1}}[u,v]^{p}\rho_{K}(u)^{n+p}\rho_{L}(v)^{n+p}\,du\,dv=\frac{(n+p)\widetilde{c}_{n,p}V(L)\widetilde{V}_{-p}(K,\Lambda_{p}^{\circ}L)}{n\omega_{n}^{2}}.

Together with the polar coordinate, the dual LpL_{-p} Minkowski inequality (2.17), and the LpL_{p}-sine Blaschke-Santaló inequality (4.15), one gets

KL[x,y]p𝑑x𝑑y\displaystyle\int_{K}\int_{L}[x,y]^{p}\,dx\,dy\!\! =\displaystyle= (nωnn+p)2Sn1Sn1[u,v]pρK(u)n+pρL(v)n+p𝑑u𝑑v\displaystyle\!\!\bigg{(}\frac{n\omega_{n}}{n+p}\bigg{)}^{2}\int_{S^{n-1}}\int_{S^{n-1}}[u,v]^{p}\rho_{K}(u)^{n+p}\rho_{L}(v)^{n+p}\,du\,dv
=\displaystyle= nc~n,pV(L)V~p(K,ΛpL)n+p\displaystyle\!\!\frac{n\widetilde{c}_{n,p}V(L)\widetilde{V}_{-p}(K,\Lambda_{p}^{\circ}L)}{n+p}
\displaystyle\geq nc~n,pV(L)V(K)n+pnV(ΛpL)pnn+p\displaystyle\!\!\frac{n\widetilde{c}_{n,p}V(L)V(K)^{\frac{n+p}{n}}V(\Lambda_{p}^{\circ}L)^{-\frac{p}{n}}}{n+p}
=\displaystyle= nc~n,p[V(L)V(K)]n+pnn+p[V(L)V(ΛpL)]pn\displaystyle\!\!\frac{n\widetilde{c}_{n,p}[V(L)V(K)]^{\frac{n+p}{n}}}{n+p}\cdot\big{[}V(L)V(\Lambda_{p}^{\circ}L)\big{]}^{-\frac{p}{n}}
\displaystyle\geq (nc~n,p[V(K)V(L)]n+pnn+p)ωn2pn\displaystyle\!\!\bigg{(}\frac{n\widetilde{c}_{n,p}[V(K)V(L)]^{\frac{n+p}{n}}}{n+p}\bigg{)}\cdot\omega_{n}^{-\frac{2p}{n}}
=\displaystyle= n(n1)ωn1ωn+p2(n+p)2ωn+p3ωn1+2p/n[V(K)V(L)]n+pn,\displaystyle\!\!\frac{n(n-1)\omega_{n-1}\omega_{n+p-2}}{(n+p)^{2}\omega_{n+p-3}\omega_{n}^{1+2p/n}}\big{[}V(K)V(L)\big{]}^{\frac{n+p}{n}},

which concludes the proof of inequality (6.1).

Now the equality condition of (6.1) immediately follows from the equality conditions of (4.15) and (2.17) together with the facts (4.3) and (4.17). ∎

Conversely, one can also deduce Theorem 4.4 by applying Theorem 6.1. To see this, for KnK\subset\mathbb{R}^{n} being an Ln+pL_{n+p}-star with positive volume, one has ΛpK𝒦en\Lambda_{p}^{\circ}K\in\mathscr{K}_{e}^{n}. Taking L=ΛpKL=\Lambda_{p}^{\circ}K in (6.1) yields

KΛpK[x,y]p𝑑x𝑑yn(n1)ωn1ωn+p2(n+p)2ωn+p3ωn1+2p/n[V(K)V(ΛpK)]n+pn.\int_{K}\int_{\Lambda_{p}^{\circ}K}[x,y]^{p}\,dx\,dy\geq\frac{n(n-1)\omega_{n-1}\omega_{n+p-2}}{(n+p)^{2}\omega_{n+p-3}\omega_{n}^{1+2p/n}}\big{[}V(K)V(\Lambda_{p}^{\circ}K)\big{]}^{\frac{n+p}{n}}. (6.3)

On the other side, it follows from (6) and (4.6) that

KΛpK[x,y]p𝑑x𝑑y=nc~n,pV(ΛpK)V~p(K,ΛpΛpK)n+p=nc~n,pV(ΛpK)V(K)n+p.\int_{K}\int_{\Lambda_{p}^{\circ}K}[x,y]^{p}\,dx\,dy=\frac{n\widetilde{c}_{n,p}V(\Lambda_{p}^{\circ}K)\widetilde{V}_{-p}(K,\Lambda_{p}^{\circ}\Lambda_{p}^{\circ}K)}{n+p}=\frac{n\widetilde{c}_{n,p}V(\Lambda_{p}^{\circ}K)V(K)}{n+p}. (6.4)

Now, inequality (4.15) immediately follows from (6.3) and (6.4).

Just like the proof of [43, Theorem A], for nonnegative functions f,gL1(Sn1)f,g\in L^{1}(S^{n-1}), by taking ρK=f1/(n+p)\rho_{K}=f^{1/(n+p)} and ρL=g1/(n+p)\rho_{L}=g^{1/(n+p)} in Theorem 6.1, one obtains the following result.

Theorem 6.2.

Let p1p\geq 1 and n2n\geq 2. For any nonnegative functions f,gL1(Sn1)f,g\in L^{1}(S^{n-1}), one has

Sn1Sn1[u,v]pf(u)g(v)𝑑u𝑑v((n1)ωn1ωn+p2nωnωn+p3)fnn+pgnn+p.\int_{S^{n-1}}\int_{S^{n-1}}[u,v]^{p}f(u)g(v)\,du\,dv\geq\bigg{(}\frac{(n-1)\omega_{n-1}\omega_{n+p-2}}{n\omega_{n}\omega_{n+p-3}}\bigg{)}\cdot\|f\|_{\frac{n}{n+p}}\|g\|_{\frac{n}{n+p}}. (6.5)

When n=2n=2 and f22+pg22+p>0\|f\|_{\frac{2}{2+p}}\|g\|_{\frac{2}{2+p}}>0, equality holds if and only if there exist ϕSL(2)\phi\in\mathrm{SL}(2) and real numbers d1,d2>0d_{1},d_{2}>0, such that, for almost all uS1u\in S^{1} (with respect to the spherical measure on S1S^{1}),

f(u)=d1|ϕu|2pandg(u)=d2|ϕu|2p.f(u)=d_{1}|\phi u|^{-2-p}\quad\mathrm{and}\quad g(u)=d_{2}|\phi u|^{-2-p}.

When n3n\geq 3 and fnn+pgnn+p>0\|f\|_{\frac{n}{n+p}}\|g\|_{\frac{n}{n+p}}>0, equality holds if and only if there exist real numbers d3,d4>0d_{3},d_{4}>0, such that, for almost all uSn1u\in S^{n-1} (with respect to the spherical measure on Sn1S^{n-1}),

f(u)=d3andg(u)=d4.f(u)=d_{3}\quad\mathrm{and}\quad g(u)=d_{4}.

As pointed out by one of the referees, Theorem 6.1 is a special case of [10, Corollary 4.2]. We also mention that Theorem 6.2 can be viewed as the “sine cousin” of the following inequality proved by Lutwak and Zhang [43]: for p1p\geq 1 and continuous functions f,g:Sn1(0,)f,g:S^{n-1}\rightarrow(0,\infty), one has

Sn1Sn1|uv|pf(u)g(v)𝑑u𝑑vωn+p2ω2ωn2ωp1fnn+pgnn+p,\int_{S^{n-1}}\int_{S^{n-1}}|u\cdot v|^{p}f(u)g(v)\,du\,dv\geq\frac{\omega_{n+p-2}}{\omega_{2}\omega_{n-2}\omega_{p-1}}\|f\|_{\frac{n}{n+p}}\|g\|_{\frac{n}{n+p}}, (6.6)

with equality if and only if there exist ϕSL(n)\phi\in\mathrm{SL}(n) and constants d1,d2>0d_{1},d_{2}>0, such that, for all uSn1u\in S^{n-1},

f(u)=d1|ϕu|(n+p)andg(u)=d2|ϕtu|(n+p).f(u)=d_{1}|\phi u|^{-(n+p)}\ \ \mathrm{and}\ \ g(u)=d_{2}|\phi^{-t}u|^{-(n+p)}.

Note that a stronger version of inequality (6.6) was proved by Nguyen [50], where the continuity of f,gf,g is replaced by integrability. Furthermore, Nguyen [50] showed that inequality (6.6) can be used to prove the following LpL_{p} moment-entropy inequality established by Lutwak, Yang, and Zhang [38]. Suppose p1p\geq 1, n2n\geq 2, and λ(nn+p,]\lambda\in(\frac{n}{n+p},\infty]. Let X,YX,Y be two independent random vectors in n\mathbb{R}^{n} with density functions f,g:n[0,)f,g:\mathbb{R}^{n}\rightarrow[0,\infty). If X,YX,Y have finite ppth moment (that is, n|x|pf(x)𝑑x<\int_{\mathbb{R}^{n}}|x|^{p}f(x)\,dx<\infty, and ff is assumed to be bounded if λ=\lambda=\infty), then

𝔼(|XY|p)c02(n+p)ωn+pnπωp1ωn1+2p/n[Nλ(X)Nλ(Y)]p/n.\mathbb{E}(|X\cdot Y|^{p})\geq\frac{c_{0}^{2}(n+p)\omega_{n+p}}{n\pi\omega_{p-1}\omega_{n}^{1+2p/n}}\big{[}N_{\lambda}(X)N_{\lambda}(Y)\big{]}^{p/n}. (6.7)

Here,

𝔼(|XY|p)=nn|xy|pf(x)g(y)𝑑x𝑑y,\mathbb{E}(|X\cdot Y|^{p})=\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}|x\cdot y|^{p}f(x)g(y)\,dx\,dy,

and the λ\lambda-Rényi entropy power Nλ(X)N_{\lambda}(X) of XX is defined by

Nλ(X)={(nf(x)λ𝑑x)11λ,ifλ1,exp(nf(x)logf(x)𝑑x),ifλ=1,limλNλ(X),ifλ=.\displaystyle N_{\lambda}(X)=\left\{\begin{array}[]{ll}\big{(}\int_{\mathbb{R}^{n}}f(x)^{\lambda}\,dx\big{)}^{\frac{1}{1-\lambda}},&\textrm{if}~{}\lambda\neq 1,\\ \exp(-\int_{\mathbb{R}^{n}}f(x)\log f(x)\,dx),&\textrm{if}~{}\lambda=1,\\ \lim_{\lambda\rightarrow\infty}N_{\lambda}(X),&\textrm{if}~{}\lambda=\infty.\end{array}\right. (6.11)

The precise value of c0c_{0} can be found in [38]. Equality in (6.7) holds if and only if there exists an origin-symmetric ellipsoid \mathscr{E}, such that, a.e.,

f(x)={b1pλ(a1ρ(x)),ifλ(nn+p,),𝟏a1(x)V(a1),ifλ=,andg(y)={b2pλ(a2ρ(y)),ifλ(nn+p,),𝟏a2(y)V(a2),ifλ=,f(x)\!=\!\left\{\begin{array}[]{ll}\!b_{1}\cdot p_{\lambda}\big{(}\frac{a_{1}}{\rho_{\mathscr{E}}(x)}\big{)},&\mathrm{if}\ \lambda\in(\frac{n}{n+p},\infty),\\ \!\frac{\mathbf{1}_{a_{1}\mathscr{E}}(x)}{V(a_{1}\mathscr{E})},&\mathrm{if}\ \lambda=\infty,\end{array}\right.\!\mathrm{and}\ \ g(y)\!=\!\left\{\begin{array}[]{ll}\!b_{2}\cdot p_{\lambda}\big{(}\frac{a_{2}}{\rho_{\mathscr{E}^{\circ}}(y)}\big{)},&\mathrm{if}\ \lambda\in(\frac{n}{n+p},\infty),\\ \!\frac{\mathbf{1}_{a_{2}\mathscr{E}^{\circ}}(y)}{V(a_{2}\mathscr{E}^{\circ})},&\mathrm{if}\ \lambda=\infty,\end{array}\right.

where a1,a2>0a_{1},a_{2}>0 are some constants, and b1,b2b_{1},b_{2} are constants chosen to make ff and gg density functions. Here, the function pλ:(0,)(0,)p_{\lambda}:(0,\infty)\rightarrow(0,\infty) is given by

pλ(s)={(1+sp)1/(λ1),ifλ<1,esp,ifλ=1,(1sp)+1/(λ1),ifλ>1,\displaystyle p_{\lambda}(s)=\left\{\begin{array}[]{ll}(1+s^{p})^{1/(\lambda-1)},&\textrm{if}\ \ \lambda<1,\\ e^{-s^{p}},&\textrm{if}\ \ \lambda=1,\\ (1-s^{p})_{+}^{1/(\lambda-1)},&\textrm{if}\ \ \lambda>1,\end{array}\right. (6.15)

where a+=max{a,0}a_{+}=\max\{a,0\} for aa\in\mathbb{R}.

Following from Theorem 6.2 and along the approach in [50] (with slight modifications), one can get the LpL_{p}-sine moment-entropy inequality in Theorem 6.3. Note that this theorem can also be proved by Theorem 4.4 and along the approach in [38] (with slight modifications).

Theorem 6.3.

Suppose p1p\geq 1, n2n\geq 2, and λ(nn+p,]\lambda\in(\frac{n}{n+p},\infty]. Let X,YX,Y be two independent random vectors in n\mathbb{R}^{n} with density functions f,g:n[0,)f,g:\mathbb{R}^{n}\rightarrow[0,\infty). If X,YX,Y have finite ppth moment, then

𝔼([X,Y]p)c02(n1)ωn1ωn+p2nωn+p3ωn1+2p/n[Nλ(X)Nλ(Y)]pn.\mathbb{E}([X,Y]^{p})\geq\frac{c_{0}^{2}(n-1)\omega_{n-1}\omega_{n+p-2}}{n\omega_{n+p-3}\omega_{n}^{1+2p/n}}\Big{[}N_{\lambda}(X)N_{\lambda}(Y)\Big{]}^{\frac{p}{n}}. (6.16)

When n=2n=2, equality in (6.16) holds if and only if there exists an origin-symmetric ellipsoid \mathscr{E}, such that, a.e.,

f(x)={b1pλ(a1ρ(x)),ifλ(22+p,),𝟏a1(x)V(a1),ifλ=,andg(y)={b2pλ(a2ρ(y)),ifλ(22+p,),𝟏a2(y)V(a2),ifλ=,f(x)\!=\!\left\{\!\!\begin{array}[]{ll}b_{1}\cdot p_{\lambda}\big{(}\frac{a_{1}}{\rho_{\mathscr{E}}(x)}\big{)},&\mathrm{if}\ \lambda\in(\frac{2}{2+p},\infty),\\ \frac{\mathbf{1}_{a_{1}\mathscr{E}}(x)}{V(a_{1}\mathscr{E})},&\mathrm{if}\ \lambda=\infty,\end{array}\right.\mathrm{and}\ g(y)\!=\!\left\{\!\!\begin{array}[]{ll}b_{2}\cdot p_{\lambda}\big{(}\frac{a_{2}}{\rho_{\mathscr{E}}(y)}\big{)},&\mathrm{if}\ \lambda\in(\frac{2}{2+p},\infty),\\ \frac{\mathbf{1}_{a_{2}\mathscr{E}}(y)}{V(a_{2}\mathscr{E})},&\mathrm{if}\ \lambda=\infty,\end{array}\right.

where a1,a2>0a_{1},a_{2}>0 are some constants, and b1,b2b_{1},b_{2} are constants chosen to make ff and gg density functions. When n3n\geq 3, equality in (6.16) holds if and only if, a.e.,

f(x)={b3pλ(a3ρBn(x)),ifλ(nn+p,),𝟏a3Bn(x)V(a3Bn),ifλ=,andg(y)={b4pλ(a4ρBn(y)),ifλ(nn+p,),𝟏a4Bn(y)V(a4Bn),ifλ=,f(x)\!=\!\!\left\{\!\!\begin{array}[]{ll}b_{3}\cdot p_{\lambda}(\frac{a_{3}}{\rho_{B^{n}}(x)}),&\mathrm{if}\ \lambda\in(\frac{n}{n+p},\infty),\\ \frac{\mathbf{1}_{a_{3}B^{n}}(x)}{V(a_{3}B^{n})},&\mathrm{if}\ \lambda=\infty,\end{array}\right.\mathrm{and}\ g(y)\!=\!\left\{\!\!\begin{array}[]{ll}b_{4}\cdot p_{\lambda}(\frac{a_{4}}{\rho_{B^{n}}(y)}),&\mathrm{if}\ \lambda\in(\frac{n}{n+p},\infty),\\ \frac{\mathbf{1}_{a_{4}B^{n}}(y)}{V(a_{4}B^{n})},&\mathrm{if}\ \lambda=\infty,\end{array}\right.

where a3,a4>0a_{3},a_{4}>0 are some constants, and b3,b4b_{3},b_{4} are constants chosen to make ff and gg density functions.

If XX and YY are random vectors with density functions 𝟏K/V(K)\mathbf{1}_{K}/V(K) and 𝟏L/V(L)\mathbf{1}_{L}/V(L), then inequality (6.16) reduces to inequality (6.1) and Theorem 6.1 can be extended to the case that KK and LL are bounded and measurable sets in n\mathbb{R}^{n}. In particular, taking pp\rightarrow\infty on both sides of inequality (6.1), Stirling’s formula yields that

supxK,yL[x,y]ωn2/n[V(K)V(L)]1/n.\sup_{x\in K,y\in L}[x,y]\geq\omega_{n}^{-2/n}[V(K)V(L)]^{1/n}. (6.17)

Note that the sine Blaschke-Santaló inequality (5.3) is a special case of inequality (6.17). Indeed, under the assumptions on KK in Theorem 5.2 and L=KL=K^{\diamond}, one has

1=supxK,yK[x,y]ωn2/n[V(K)V(K)]1/n,1=\sup_{x\in K,y\in K^{\diamond}}[x,y]\geq\omega_{n}^{-2/n}[V(K)V(K^{\diamond})]^{1/n},

as desired.

Finally, it is easy to see that inequalities (6.5) and (6.6), (6.16) and (6.7), together with their equality conditions are in fact equivalent when n=2n=2.

Acknowledgement. The authors would like to thank Professor Daniel Hug from Karlsruhe Institute of Technology for many helpful discussion on the proof of inequality (3.8). The authors are also indebted to the referees for many valuable suggestions and comments, which greatly improve the quality and presentation of the present paper. In particular, we are grateful to the referees for pointing out references [2, 10, 48] and providing the Mathematica code for the images in Figure 1.

The research of QH was supported by NSFC (No. 11701219) and AARMS postdoctoral fellowship (joint with Memorial University of Newfoundland, Canada). The research of AL was supported by Zhejiang Provincial Natural Science Foundation of China (LY22A010001). The research of DX was supported by NSFC (No. 12071277) and STCSM program (No. 20JC1412600). The research of DY was supported by a NSERC grant, Canada.

References

  • [1] I. Angell and M. Moore, Symmetrical intersections of cylinders, Acta Cryst. Sect. A 43 (1987), 244-250.
  • [2] S. Artstein-Avidan, S. Sadovsky, and K. Wyczesany, Optimal measure transportation with respect to non-traditional costs, arXiv:2104.04838.
  • [3] S. Artstein-Avidan, S. Sadovsky, and K. Wyczesany, A Zoo of Dualities, arXiv:2110.11308.
  • [4] W. Blaschke, Über affine Geometrie VII: Neue Extremeingenschaften von Ellipse und Ellipsoid, Ber. Verh. Sächs. Akad. Wiss., Math. Phys. Kl. 69 (1917), 412-420.
  • [5] S. Bobkov and M. Madiman, Reverse Brunn-Minkowski and reverse entropy power inequalities for convex measures, J. Funct. Anal. 262 (2012), 3309-3339.
  • [6] J. Bourgain and J. Lindenstrauss, Projection bodies, Geometric Aspects of Functional Analysis (1986-87) (J. Lindenstrauss and V.D. Milman, eds.), Lecture Notes in Math., Springer, Berlin, 1317 (1988), 250-270.
  • [7] J. Bourgain, M. Meyer, V. Milman, and A. Pajor, On a geometric inequality, Geometric Aspects of Functional Analysis (1986-87) (J. Lindenstrauss and V.D. Milman, eds.), Lecture Notes in Math., Springer, Berlin, 1317 (1988), 271-282.
  • [8] P. Bourke, Intersecting cylinders, http://paulbourke.net/geometry/cylinders/.
  • [9] A. Cianchi, E. Lutwak, D. Yang, and G. Zhang, A unified approach to Cramér-Rao inequalities, IEEE Trans. Inform. Theory 60 (2014), 643-650.
  • [10] S. Dann, G. Paouris, and P. Pivovarov, Bounding marginal densities via affine isoperimetry, Proc. Lond. Math. Soc. 113 (2016), 140-162.
  • [11] M. Fradelizi and A. Marsiglietti, On the analogue of the concavity of entropy power in the Brunn-Minkowski theory, Adv. in Appl. Math. 57 (2014), 1-20.
  • [12] R.J. Gardner, The Brunn-Minkowski inequality, Bull. Amer. Math. Soc. 39 (2002), 355-405.
  • [13] R.J. Gardner, Geometric tomography, second edition, Encyclopedia of Mathematics and its Applications, 58, Cambridge University Press, New York, 2006.
  • [14] P. Goodey and W. Weil, The determination of convex bodies from the mean of random sections, Math. Proc. Cambridge Philos. Soc. 112 (1992), 419-430.
  • [15] P. Goodey and W. Weil, A uniqueness result for mean section bodies, Adv. Math. 229 (2012), 596-601.
  • [16] E. Grinberg and B. Rubin, Radon inversion on Grassmannians via Gårding-Gindikin fractional integrals, Ann. Math. 159 (2004), 783-817.
  • [17] O. Guleryuz, E. Lutwak, D. Yang, and G. Zhang, Information-theoretic inequalities for contoured probability distributions, IEEE Trans. Inform. Theory 48 (2002), 2377-2383.
  • [18] C. Haberl, Minkowski valuations intertwining with the special linear group, J. Eur. Math. Soc. 14 (2012), 1565-1597.
  • [19] C. Haberl and F. Schuster, General LpL_{p} affine isoperimetric inequalities, J. Differential Geom. 83 (2009), 1-26.
  • [20] C. Haberl and F. Schuster, Affine vs. Euclidean isoperimetric inequalities, Adv. Math. 356 (2019), 106811, 26 pp.
  • [21] J. Haddad, A convex body associated to the Busemann random simplex inequality and the Petty conjecture, J. Funct. Anal. 281 (2021), Paper No. 109118, 28 pp.
  • [22] Heidelberg Institute for Theoretical Studies, Steinmetz Solids, https://www.h-its.org/projects/steinmetz-solids/.
  • [23] J.P. Hogendijk, The Surface Area of the Bicylinder and Archimedes’ Method, Historia Math. 29 (2002), 199-203.
  • [24] D. Hug and W. Weil, Lectures on convex geometry, Graduate Texts in Math. 286, Springer, Cham, 2020.
  • [25] T. Kiang, An old Chinese way of finding the volume of a sphere, Math. Gaz. 56 (1972), 88-91.
  • [26] A. Koldobsky, Fourier analysis in convex geometry, Mathematical Surveys and Monographs, Amer. Math. Soc., Providence, Rhode Island, 2005.
  • [27] A.-J. Li, Q. Huang, and D. Xi, New sine ellipsoids and related volume inequalities, Adv. Math. 353 (2019), 281-311.
  • [28] A.-J. Li, D. Xi, and G. Zhang, Volume inequalities of convex bodies from cosine transforms on Grassmann manifolds, Adv. Math. 304 (2017), 494-538.
  • [29] Y. Lonke, Derivatives of the LpL_{p}-cosine transform, Adv. Math. 176 (2003), 175-186.
  • [30] M. Ludwig, Projection bodies and valuations, Adv. Math. 172 (2002), 158-168.
  • [31] M. Ludwig, Minkowski valuations, Trans. Amer. Math. Soc. 357 (2005), 4191-4213.
  • [32] E. Lutwak, Selected affine isoperimetric inequalities, In Handbook of Convex Geometry (P.M. Gruber and J.M. Wills eds.), North-Holland, Amsterdam, 1993, pp. 151-176.
  • [33] E. Lutwak, S. Lv, D. Yang, and G. Zhang, Extensions of Fisher information and Stam’s inequality, IEEE Trans. Inform. Theory 3 (2012), 1319-1327.
  • [34] E. Lutwak, S. Lv, D. Yang, and G. Zhang, Affine moments of a random vector, IEEE Trans. Inform. Theory 59 (2013), 5592-5599.
  • [35] E. Lutwak, D. Yang, and G. Zhang, LpL_{p} affine isoperimetric inequalities, J. Differential Geom. 56 (2000), 111-132.
  • [36] E. Lutwak, D. Yang, and G. Zhang, A new ellipsoid associated with convex bodies, Duke Math. J. 104 (2000) 375-390.
  • [37] E. Lutwak, D. Yang, and G. Zhang, The Cramer-Rao inequality for star bodies, Duke Math. J. 112 (2002), 59-81.
  • [38] E. Lutwak, D. Yang, and G. Zhang, Moment-entropy inequalities, Ann. Probab. 32 (2004), 757-774.
  • [39] E. Lutwak, D. Yang, and G. Zhang, Volume inequalities for subspaces of LpL_{p}, J. Differential Geom. 68 (2004), 159-184.
  • [40] E. Lutwak, D. Yang, and G. Zhang, LpL_{p} John ellipsoids, Proc. London Math. Soc. 90 (2005), 497-520.
  • [41] E. Lutwak, D. Yang, and G. Zhang, Cramér-Rao and moment-entropy inequalities for Renyi entropy and generalized Fisher information, IEEE Trans. Inform. Theory 51 (2005), 473-478.
  • [42] E. Lutwak, D. Yang, and G. Zhang, Moment-Entropy Inequalities for a Random Vector, IEEE Trans. Inform. Theory 53 (2007), 1603-1607.
  • [43] E. Lutwak and G. Zhang, Blaschke-Santaló inequalities, J. Differential Geom. 47 (1997), 1-16.
  • [44] M. Madiman, J. Melbourne, and P. Xu, Forward and reverse entropy power inequalities in convex geometry, Convexity and Concentration (E. Carlen, M. Madiman, E. Werner, eds), IMA Vol. Math. Appl., Springer, New York, 161 (2017), 427-485.
  • [45] G. Maresch and F. Schuster, The sine transform of isotropic measures, Int. Math. Res. Not. 2012 (2012), 717-739.
  • [46] M. Meyer and A. Pajor, On the Blaschke-Santaló inequality, Arch. Math. (Basel) 55 (1990), 82-93.
  • [47] M. Meyer and S. Reisner, Shadow systems and volumes of polar convex bodies, Mathematika 53 (2006), 129-148.
  • [48] M. Meyer and E. Werner, The Santaló-regions of a convex body, Trans. Amer. Math. Soc. 350 (1998), 4569-4591.
  • [49] M. Moore, Symmetrical intersections of right circular cylinders, Math. Gaz. 58 (1974), 181-185.
  • [50] V.H. Nguyen, A simple proof of the moment-entropy inequalities, Adv. in Appl. Math. 108 (2019), 31-44.
  • [51] C.M. Petty, Affine isoperimetric problems, Discrete Geometry and Convexity (J. E. Goodman, E. Lutwak, J. Malkevitch, R. Pollack, eds), Ann. New York Acad. Sci. 440 (1985), 113-127.
  • [52] B. Rubin, Inversion formulas for the spherical Radon transform and the generalized cosine transform, Adv. in Appl. Math. 29 (2002), 471-497.
  • [53] B. Rubin, Radon, cosine and sine transforms on real hyperbolic space, Adv. Math. 170 (2002), 206-223.
  • [54] B. Rubin, Funk, cosine, and sine transforms on Stiefel and Grassmann manifolds, J. Geom. Anal. 23 (2013), 1441-1497.
  • [55] D. Ryabogin and A. Zvavitch, The Fourier transform and Firey projections of convex bodies, Indiana Univ. Math. J. 53 (2004), 667-682.
  • [56] J. Saint-Raymond, Sur le volume des corps convexes symétriques, Initiation Seminar on Analysis: G. Choquet-M. Rogalski-J. Saint-Raymond, 20th Year: 1980/1981, Publ. Math. Univ. Pierre et Marie Curie, Paris, 11 (1981), 25 pp.
  • [57] L.A. Santaló, Un invariante afin para los cuerpos convexos de espacio de nn dimensiones, Portugal. Math. 8 (1949), 155-161.
  • [58] R. Schneider, Convex bodies: the Brunn-Minkowski theory, second expanded edition, Encyclopedia of Mathematics and its Applications, Vol. 151, Cambridge University Press, Cambridge, 2014.
  • [59] V. Yaskin and M. Yaskina, Centroid bodies and comparison of volumes, Indiana Univ. Math. J. 55 (2006), 1175-1194.
  • [60] G. Zhang, Radon, cosine and sine transforms on Grassmannian manifolds, Int. Math. Res. Not. 2009 (2009), 1743-1772.

Qingzhong Huang,  [email protected]
College of Data Science, Jiaxing University, Jiaxing, 314001, China

Ai-Jun Li,   [email protected]
School of Science, Zhejiang University of Science and Technology, Hangzhou, Zhejiang, 310023, China

Dongmeng Xi,   [email protected]
Department of Mathematics, Shanghai University, Shanghai, 200444, China

Deping Ye,   [email protected]
Department of Mathematics and Statistics, Memorial University of Newfoundland, St. John’s, Newfoundland A1C 5S7, Canada