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

Local fixed point results for centroid body operators

Chase Reuter Department of Mathematics 2750
North Dakota State University
PO BOX 6050
Fargo, ND 58108-6050
USA
[email protected]
Abstract.

We prove that, in a neighborhood of the Euclidean ball, there are no other fixed points of the pp-centroid body operator, using spherical harmonic techniques. We also show that the Euclidean ball is locally the only body whose centroid body is a dilate of its polar intersection body.

Key words and phrases:
Convex bodies, spherical harmonics, spherical Radon and Cosine transforms
2010 Mathematics Subject Classification:
52A20,33C55,44A12

1. Introduction

Characterizing the Euclidean space among all real normed spaces is one of the aims of the ten problems formulated in 1956 by Busemann and Petty [4]. These problems lead to the study of certain operators on convex bodies in n{\mathbb{R}^{n}}, such as the intersection body operator for the first Busemann-Petty problem. Proving a particular property of the intersection body operator in each dimension resulted in the resolution of the first Busemann-Petty problem in 1996. In the class of convex bodies, proving properties for operators globally is challenging. The local study of such operators on convex bodies appears to be a more approachable initial step. In [5], it was established that, other than linear transformations of the Euclidean ball, there are no additional fixed points of the intersection body operator in a neighborhood of the Euclidean ball. A similar result was obtained in [12] for the projection body operator. In [1], a local affirmative answer was obtained for the Busemann-Petty problems 5 and 8, by studying the local fixed points of the polar-intersection body operator.

In this paper we conduct a similar local study for several problems involving the centroid body operator. First, we consider its fixed points.
Problem 1. If for an origin-symmetric convex body KnK\subset{\mathbb{R}^{n}}, n3n\geq 3 we have

(1) cΓK=K,c\Gamma K=K,

where ΓK\Gamma K is the centroid body of KK and cc is some constant, need KK be an ellipsoid?

Below we will verify that the ball does indeed satisfy this condition, and that equation (1) is invariant under linear transformations. Consequently, this equation holds for all ellipsoids. We prove the following result.

Theorem 1.

Let n3n\geq 3. If an origin-symmetric convex body KnK\subset{\mathbb{R}^{n}} satisfies (1) and is sufficiently close to the Euclidean ball in the Banach-Mazur distance, then KK must be an ellipsoid.

A similar result is obtained for the pp-centroid body operators Γp\Gamma_{p}, p(1,+)p\in(1,+\infty).

Theorem 2.

Let n3n\geq 3. If an origin-symmetric convex body KnK\subset{\mathbb{R}^{n}} satisfies

(2) cΓpK=K,c\Gamma_{p}K=K,

and is sufficiently close to the Euclidean ball in the Banach-Mazur distance, then KK must be an ellipsoid.

In addition, we prove a local result involving the pp-centroid body and the polar intersection body of KK:

Theorem 3.

Let n3n\geq 3 and p1p\geq 1. If an origin-symmetric convex body KnK\subset{\mathbb{R}^{n}} satisfies

(3) cΓpK=K,c\Gamma_{p}K={\mathcal{I}}^{*}K,

and is sufficiently close to the Euclidean ball in the Banach-Mazur distance, then KK must be an ellipsoid.

Theorem 3 is similar to several open problems listed by Gardner in [6, pg 336-337], relating the intersection body of KK with the difference body ΔK\Delta K and the cross section body CKCK of KK:


Problem 8.1 Suppose that KK is a convex body such that ΔK\Delta K and K{\mathcal{I}}^{*}K are homothetic. Must KK be a centered ellipsoid?


Problem 8.9 (ii) If ΠK\Pi K is homothetic to either K{\mathcal{I}}K or CKCK, is KK an ellipsoid?


In two dimensions, the answer to Problem 1 is known to be affirmative: the only convex bodies which are dilates of their centroid bodies are ellipses. This was proven by Petty in [9, Theorem 5.6]. Indeed, the centroid body of KK has an important role in fluid statics. If KK is an origin-symmetric convex body with density 1/2 (with water having density 1), each point on the boundary of the centroid body of KK is the center of mass of the underwater part of KK in a given direction, and ΓK\Gamma K is the convex hull of the locus of all such centers of mass in all directions. The flotation properties of KK are governed by three Theorems of Dupin for ΓK\Gamma K (see for example [11]). In particular, in two dimensions, the Third Theorem of Dupin states that the radius of curvature R(ξ)R(\xi) of the centroid body ΓK\Gamma K at a boundary point perpendicular to a unit direction ξ\xi is proportional to the cube of the length of the chord (ξ)=Kξ\ell(\xi)=K\cap\xi^{\perp},

R(ξ)=(ξ)324,R(\xi)=\frac{\ell(\xi)^{3}}{24},

If we assume that ΓK\Gamma K is homothetic to KK, then the same relation is true between the curvature of KK at a point and the cube of the length of the corresponding central section. But in the two dimensional case this is a restatement of the Busemann-Petty Problem number 8 in dimension 2 [4, 9].

2. Analytic statement of the problems

Let 𝒮n1={xn:|x|=1}{\mathcal{S}^{n-1}}=\{x\in\mathbb{R}^{n}:|x|=1\} be the unit sphere in n\mathbb{R}^{n}, where |x||x| denotes the Euclidean norm. Let σn1\sigma_{n-1} be the spherical Lebesgue measure on 𝒮n1{\mathcal{S}^{n-1}} and λn\lambda_{n} be the nn-dimensional Lebesgue measure. A convex body KK is a convex compact subset of n\mathbb{R}^{n} with non-empty interior. We will assume that the origin is an interior point of convex bodies. KK is origin symmetric if for every xKx\in K, xK-x\in K.

Given a convex body KnK\subset{\mathbb{R}^{n}}, the support function hKh_{K} of KK is defined as

hK(θ)=max{xθ|xK},h_{K}(\theta)=\max\{x\cdot\theta\hskip 2.0pt\rvert\hskip 2.0ptx\in K\},

and its radial function is

ρK(θ)=max{t>0:tθK}.\rho_{K}(\theta)=\max\{t>0:\,t\theta\in K\}.

A body LL is called the centroid body of a convex body KK if

hL(ϕ)=1voln(K)K|ϕx|𝑑λn(x),ϕ𝒮n1.h_{L}(\phi)=\frac{1}{\text{vol}_{n}(K)}\int_{K}\left\lvert\phi\cdot x\right\rvert\hskip 2.0ptd\lambda_{n}(x),\hskip 10.0pt\forall\phi\in{\mathcal{S}^{n-1}}.

The body LL is denoted as ΓK\Gamma K. Switching this integral to polar coordinates and simplifying yields

hΓK(ϕ)\displaystyle h_{\Gamma K}(\phi) =1voln(K)𝒮n10ρK(θ)|ϕ(rθ)|rn1𝑑r𝑑σn1(θ)\displaystyle=\frac{1}{\text{vol}_{n}(K)}\int_{{\mathcal{S}^{n-1}}}\int_{0}^{\rho_{K}(\theta)}\left\lvert\phi\cdot(r\theta)\right\rvert r^{n-1}\hskip 2.0ptdr\hskip 2.0ptd\sigma_{n-1}(\theta)
=1(n+1)voln(K)𝒮n1|ϕθ|ρKn+1(θ)𝑑σn1(θ).\displaystyle=\frac{1}{(n+1)\text{vol}_{n}(K)}\int_{{\mathcal{S}^{n-1}}}\left\lvert\phi\cdot\theta\right\rvert\rho_{K}^{n+1}(\theta)\hskip 2.0ptd\sigma_{n-1}(\theta).

Examining the right hand side, we observe that it is a multiple of the cosine transform. The cosine transform 𝒞f{\mathcal{C}}f of an integrable function f:𝒮n1f:{\mathcal{S}^{n-1}}\rightarrow{\mathbb{R}} is defined as

𝒞f(ϕ):=𝒮n1|ϕθ|f(θ)𝑑σn1(θ).{\mathcal{C}}f(\phi):=\int_{{\mathcal{S}^{n-1}}}\left\lvert\phi\cdot\theta\right\rvert f(\theta)\hskip 2.0ptd\sigma_{n-1}(\theta).

Using this to rewrite the support function of the centroid body gives

(4) hΓK=1(n+1)voln(K)𝒞ρKn+1.h_{\Gamma K}=\frac{1}{(n+1)\text{vol}_{n}(K)}{\mathcal{C}}\rho_{K}^{n+1}.

Equation (1) can now be rewritten analytically as

(5) hK=c(n+1)voln(K)𝒞ρKn+1.h_{K}=\frac{c}{(n+1)\text{vol}_{n}(K)}{\mathcal{C}}\rho_{K}^{n+1}.

A body LL is called the intersection body of the convex body KK if

hL(θ)=voln(Kθ),θ𝒮n1.h_{L}(\theta)=\text{vol}_{n}(K\cap\theta^{\perp}),\hskip 10.0pt\forall\theta\in{\mathcal{S}^{n-1}}.

The body LL is denoted as K{\mathcal{I}}K. Expressing the volume as an integral and passing to polar coordinates, we obtain

ρK(ϕ)\displaystyle\rho_{{\mathcal{I}}K}(\phi) =Kϕ𝑑λn1(x)\displaystyle=\int_{K\cap\phi^{\perp}}\hskip 2.0ptd\lambda_{n-1}(x)
=𝒮n1ϕ0ρK(θ)rn2𝑑r𝑑σn2(θ)\displaystyle=\int_{{\mathcal{S}^{n-1}}\cap\phi^{\perp}}\int_{0}^{\rho_{K}(\theta)}r^{n-2}\hskip 2.0ptdr\hskip 2.0ptd\sigma_{n-2}(\theta)
=1n1𝒮n1ϕρKn1(θ)𝑑σn2(θ).\displaystyle=\frac{1}{n-1}\int_{{\mathcal{S}^{n-1}}\cap\phi^{\perp}}\rho_{K}^{n-1}(\theta)\hskip 2.0ptd\sigma_{n-2}(\theta).

The right hand side is almost the spherical Radon transform. The spherical Radon transform of an integrable function fL1(𝒮n1)f\in L^{1}({\mathcal{S}^{n-1}}) is denoted f{\mathcal{R}}f and is defined as

f(ϕ)=𝒮n1ϕf(θ)𝑑σn2(θ).{\mathcal{R}}f(\phi)=\int_{{\mathcal{S}^{n-1}}\cap\phi^{\perp}}f(\theta)\hskip 2.0ptd\sigma_{n-2}(\theta).

Thus, the radial function of the intersection body can be expressed in terms of the Radon transform of the radial function of the original body, by

(6) ρK=1n1ρKn1.\rho_{{\mathcal{I}}K}=\frac{1}{n-1}{\mathcal{R}}\rho_{K}^{n-1}.

A body LL is the polar body of a convex body KK if

hL(θ)=1ρK(θ),θ𝒮n1.h_{L}(\theta)=\frac{1}{\rho_{K}(\theta)},\forall\theta\in{\mathcal{S}^{n-1}}.

The body LL is denoted by KK^{*}. The polar intersection body of a body KK is K=(K){\mathcal{I}}^{*}K=({\mathcal{I}}K)^{*}. Hence,

hK=(n1)(ρKn1)1.h_{{\mathcal{I}}^{*}K}=(n-1)({\mathcal{R}}\rho_{K}^{n-1})^{-1}.

Combining this with equation (4) gives the reformulation of cΓK=Kc\Gamma K={\mathcal{I}}^{*}K as the following equation,

(7) c(n21)voln(K)𝒞ρKn+1=(ρKn1)1.\frac{c}{(n^{2}-1)\text{vol}_{n}(K)}{\mathcal{C}}\rho_{K}^{n+1}=({\mathcal{R}}\rho_{K}^{n-1})^{-1}.

To simplify the exposition, we will first consider the case p=1p=1. The analytic reformulation of equation (2) will be derived in Section 5, and the analytic reformulation of (3) will be considered in Section 6.


Remark: Even though Theorem 1 does not make such assumptions on the body KK, if KK satisfies (1), then it must be centrally symmetric, strictly convex and smooth, since ΓK\Gamma K has those properties (see [6, Sec. 9.1]).

2.1. Ellipsoids as solutions

2.1.1. cΓK=Kc\Gamma K=K

Let us denote by CKC_{K} the constant cc in (5) corresponding to a body KK that satisfies (1). If we evaluate (5) when KK is the unit Euclidean ball BB, we obtain

1=hB(θ)=CB(n+1)voln(B)𝒞1=2CBκn1(n+1)κn,1=h_{B}(\theta)=\frac{C_{B}}{(n+1)\text{vol}_{n}(B)}{\mathcal{C}}1=\frac{2C_{B}\kappa_{n-1}}{(n+1)\kappa_{n}},

since 𝒞1=2κn1{\mathcal{C}}1=2\kappa_{n-1}, where κn\kappa_{n} represents the volume of the unit Euclidean ball in n\mathbb{R}^{n}. We conclude that

CB=(n+1)κn2κn1.C_{B}=\frac{(n+1)\kappa_{n}}{2\kappa_{n-1}}.

Thus, BB satisfies equation (1) with this constant CBC_{B}.

Next, we will verify that (5) is invariant under linear transformations of the body KK. This will imply that ellipsoids are also solutions of (1). For TGLnT\in\text{GL}_{n}, we have

hΓTK(ϕ)\displaystyle h_{\Gamma TK}\left(\phi\right) =1voln(TK)TK|ϕx|𝑑x\displaystyle=\frac{1}{\text{vol}_{n}(TK)}\int_{TK}\left\lvert{\phi}\cdot{x}\right\rvert dx
=|detT1|voln(K)K|ϕTx||detT|𝑑x\displaystyle=\frac{\left\lvert\det T^{-1}\right\rvert}{\text{vol}_{n}(K)}\int_{K}\left\lvert{\phi}\cdot{Tx}\right\rvert\left\lvert\det T\right\rvert dx xTx\displaystyle x\rightarrow Tx
=1voln(K)K|(Tϕ)x|𝑑x\displaystyle=\frac{1}{\text{vol}_{n}(K)}\int_{K}\left\lvert(T^{*}\phi)\cdot x\right\rvert dx
=hΓK(Tϕ)=hTΓK(ϕ).\displaystyle=h_{\Gamma K}(T^{*}\phi)=h_{T\Gamma K}(\phi).

Since linear transformations commute with the centroid body operator, if KK satisfies K=CKΓKK=C_{K}\Gamma K, then so does TKTK, with the same constant CTK=CKC_{TK}=C_{K}.

2.1.2. cK=ΓKc{\mathcal{I}^{*}}K=\Gamma K

Let CKC_{K}^{\prime} be the constant associated with a body KK that satisfies the condition of Theorem 3. If we evaluate (7) when KK is the Euclidean ball, we see

(1)\displaystyle({\mathcal{R}}1) =CB(n1)(𝒞1(n+1)κn)1,\displaystyle=C_{B}^{\prime}(n-1)\left(\frac{{\mathcal{C}}1}{{(n+1)\kappa_{n}}}\right)^{-1},
CB\displaystyle C_{B}^{\prime} =(1n1)(𝒞1(n+1)κn).\displaystyle=\left(\frac{{\mathcal{R}}1}{n-1}\right)\left(\frac{{\mathcal{C}}1}{(n+1)\kappa_{n}}\right).

To show that all ellipsoids are fixed points, we need to check how linear transformations interact with (3). Using the calculations found in Gardner’s book [6, pgs 21 & 308], the equation below follows.

ΓTK\displaystyle\Gamma TK =TΓK=TCK(K)=CK(TK)\displaystyle=T\Gamma K=TC_{K}^{\prime}({\mathcal{I}}K)^{*}=C_{K}^{\prime}(T^{-*}{\mathcal{I}}K)^{*}
=CK(|det(T)|1TK)=CK|detT|TK.\displaystyle=C_{K}^{\prime}(\left\lvert\det(T)\right\rvert^{-1}{\mathcal{I}}TK)^{*}=C_{K}^{\prime}\left\lvert\det T\right\rvert{\mathcal{I}}^{*}TK.

Therefore if KK satisfies cΓK=Kc\Gamma K={\mathcal{I}}^{*}K, so does TKTK, albeit for a different constant, CTK=CK|detT|C_{TK}^{\prime}=C_{K}^{\prime}\left\lvert\det T\right\rvert.

2.2. From the Banach-Mazur distance to the Hausdorff distance

2.2.1. cΓK=Kc\Gamma K=K

The Banach-Mazur distance (see [13, pg 589]) between two origin symmetric convex bodies KK and LL, is defined as

dBM(K,L):=min{1+λ1λ1|(1λ)KTL(1+λ)K,TGLn}.d_{BM}(K,L):=\min\left\{\frac{1+\lambda}{1-\lambda}\geq 1\hskip 2.0pt\rvert\hskip 2.0pt(1-\lambda)K\subseteq TL\subseteq(1+\lambda)K,T\in\text{GL}_{n}\right\}.

Let ϵ>0\epsilon>0 and suppose that dBM(B,K)<1+ϵ1ϵd_{BM}(B,K)<\frac{1+\epsilon}{1-\epsilon}. Taking the linear transformation TT where the minimum is achieved yields

(1ϵ)BTK(1+ϵ)B.(1-\epsilon)B\subseteq TK\subseteq(1+\epsilon)B.

Then, for the radial and support functions of TKTK we have

(8) 1ϵρTK1+ϵ1ϵhTK1+ϵ,1-\epsilon\leq\rho_{TK}\leq 1+\epsilon\hskip 40.0pt1-\epsilon\leq h_{TK}\leq 1+\epsilon,

and taking volumes on each side gives

(9) (1ϵ)nvoln(TK)κn(1+ϵ)n.(1-\epsilon)^{n}\leq\frac{\text{vol}_{n}(TK)}{\kappa_{n}}\leq(1+\epsilon)^{n}.

While in similar problems (see, for example, [1, 5]), an appropriate dilation of the body can be chosen so that CK=CBC_{K}=C_{B}, here we have that the constant in (1) is invariant under linear transformations, making this choice impossible. Therefore, we will use equations (8) and (9) to obtain a bound on the constant CTK=CKC_{TK}=C_{K} in terms of CBC_{B} and ϵ\epsilon.

The cosine transform is a positive operator and as such it preserves inequalities. Applying the cosine transform to all sides of the radial function estimate (8), and adjusting the constants, we obtain

𝒞(1ϵ)n+1𝒞ρTKn+1𝒞(1+ϵ)n+1,{\mathcal{C}}(1-\epsilon)^{n+1}\leq{\mathcal{C}}\rho_{TK}^{n+1}\leq{\mathcal{C}}(1+\epsilon)^{n+1},
CB(n+1)κn𝒞(1ϵ)n+1CB(n+1)κn𝒞ρTKn+1CB(n+1)κn𝒞(1+ϵ)n+1.\frac{C_{B}}{(n+1)\kappa_{n}}{\mathcal{C}}(1-\epsilon)^{n+1}\leq\frac{C_{B}}{(n+1)\kappa_{n}}{\mathcal{C}}\rho_{TK}^{n+1}\leq\frac{C_{B}}{(n+1)\kappa_{n}}{\mathcal{C}}(1+\epsilon)^{n+1}.

Using equation (4), we can rewrite all three terms in the inequality as

(1ϵ)n+1CBκnvoln(TK)hΓ(TK)(1+ϵ)n+1,(1-\epsilon)^{n+1}\leq\frac{C_{B}}{\kappa_{n}}\text{vol}_{n}(TK)\,h_{\Gamma(TK)}\leq(1+\epsilon)^{n+1},
(1ϵ)n+1CBvoln(TK)CKκnhCKΓTK(1+ϵ)n+1.(1-\epsilon)^{n+1}\leq\frac{C_{B}\text{vol}_{n}(TK)}{C_{K}\kappa_{n}}h_{C_{K}\Gamma TK}\leq(1+\epsilon)^{n+1}.

Therefore, if TK=CKΓ(TK)TK=C_{K}\Gamma(TK), we have

(1ϵ)n+1CBCKvoln(TK)κnhTK(1+ϵ)n+1.(1-\epsilon)^{n+1}\leq\frac{C_{B}}{C_{K}}\frac{\text{vol}_{n}(TK)}{\kappa_{n}}h_{TK}\leq(1+\epsilon)^{n+1}.

Now we use the support function estimate in (8), together with (9), to obtain

(1ϵ)n+1(1+ϵ)n+1CBCK(1+ϵ)n+1(1ϵ)n+1.\frac{(1-\epsilon)^{n+1}}{(1+\epsilon)^{n+1}}\leq\frac{C_{B}}{C_{K}}\leq\frac{(1+\epsilon)^{n+1}}{(1-\epsilon)^{n+1}}.

Therefore, CK=CB+O(ϵ)C_{K}=C_{B}+O(\epsilon).


Among all linear transformations, we will choose one that places KK in the isotropic position, i.e., the position where

(10) K|xy|2𝑑y=c|x|2xn.\int_{K}\left\lvert x\cdot y\right\rvert^{2}dy=c\,|x|^{2}\qquad\qquad\forall x\in{\mathbb{R}^{n}}.

It should be noted that the isotropic position as stated above is not unique as any dilate of KK must also satisfy (10). We are interested in convex bodies close to the Euclidean ball and will accordingly select a dilate where voln(K)κn\text{vol}_{n}(K)\approx\kappa_{n} per (9). See, for example [3, Sec. 2.3.2] for more details about the derivation of the isotropic position. In [1, Sec. 5], it is shown that if ϵ>0\epsilon>0 and (1ϵ)BK(1+ϵ)B(1-\epsilon)B\subseteq K\subseteq(1+\epsilon)B, then for the isotropic position KK^{\prime} of KK with voln(K)=voln(K)\text{vol}_{n}(K)=\text{vol}_{n}(K^{\prime}), we have

(1ϵ)(1ϵ1+ϵ)n+22BK(1+ϵ)(1+ϵ1ϵ)n+22B,(1-\epsilon)\left(\frac{1-\epsilon}{1+\epsilon}\right)^{\frac{n+2}{2}}B\subseteq K^{\prime}\subset(1+\epsilon)\left(\frac{1+\epsilon}{1-\epsilon}\right)^{\frac{n+2}{2}}B,

and the above considerations guarantee that the constant still satisfies CK=CB+O(ϵ)C_{K^{\prime}}=C_{B}+O(\epsilon).

2.2.2. cK=ΓKc{\mathcal{I}^{*}}K=\Gamma K

Starting with (8), we perform similar estimates for the cosine and Radon transform, obtaining

1(1+ϵ)n+1(𝒞ρKn+1(n+1)voln(K))1(𝒞1(n+1)voln(K))1(1ϵ)n+1,\frac{1}{(1+\epsilon)^{n+1}}\leq\left(\frac{{\mathcal{C}}\rho_{K}^{n+1}}{(n+1)\text{vol}_{n}(K)}\right)^{-1}\left(\frac{{\mathcal{C}}1}{(n+1)\text{vol}_{n}(K)}\right)\leq\frac{1}{(1-\epsilon)^{n+1}},
(1ϵ)n1(ρKn1n1)(n11)(1+ϵ)n1.(1-\epsilon)^{n-1}\leq\left(\frac{{\mathcal{R}}\rho_{K}^{n-1}}{n-1}\right)\left(\frac{n-1}{{\mathcal{R}}1}\right)\leq(1+\epsilon)^{n-1}.

Taking the ratio of the two terms above, multiplying by (9) and using equation (7) yields

(1ϵ)n(1+ϵ)n+1CBCK(1+ϵ)n(1ϵ)n+1.\frac{(1-\epsilon)^{n}}{(1+\epsilon)^{n+1}}\leq\frac{C_{B}^{\prime}}{C_{K}^{\prime}}\leq\frac{(1+\epsilon)^{n}}{(1-\epsilon)^{n+1}}.

Hence, the ratio of the constants is of the order 1+O(ϵ)1+O(\epsilon). Incorporating the isotropic position we still have CK=CB+O(ϵ)C_{K}^{\prime}=C_{B}^{\prime}+O(\epsilon).

3. Spherical harmonics

Spherical harmonics are homogeneous harmonic polynomials restricted to the sphere. They form the eigenspaces for a collection of common linear operators, such as the spherical Radon transform, the cosine transform and the Laplace-Baltrami operator on the sphere. In this section, we will mostly follow the expositions from Groemer [7] and Atkison-Han [2].

The space of all harmonic, homogeneous polynomials of degree kk on nn variables whose domain is restricted to the sphere is denoted kn{{\mathcal{H}}_{k}^{n}}. Its elements are called the spherical harmonics of degree kk. The spherical harmonic spaces are orthogonal with respect to the L2(𝒮n1)L^{2}({\mathcal{S}^{n-1}}) inner product. The space of all finite sums of spherical harmonics on nn variables is denoted by n=k=0kn{{\mathcal{H}}^{n}}=\oplus_{k=0}^{\infty}{{\mathcal{H}}_{k}^{n}}, and called the space of spherical harmonics.

The space of spherical harmonics is dense in L2(𝒮n1)L^{2}({\mathcal{S}^{n-1}}) and by extension also in C(𝒮n1)C({\mathcal{S}^{n-1}}). The orthogonal projection onto the spherical harmonics of degree kk, will be denoted 𝒫k:C(𝒮n1)kn{\mathcal{P}_{k}}:C({\mathcal{S}^{n-1}})\rightarrow{{\mathcal{H}}_{k}^{n}} (see [2, Def. 2.11] for an explicit definition of the projection). The eigenvalues of the cosine transform on each of the spherical harmonic spaces are (see [7, Lemmas 3.4.1, 3.4.5)]), for k=0k=0, 𝒞𝒫0=2κn1𝒫0{\mathcal{C}}{\mathcal{P}_{0}}=2\kappa_{n-1}{\mathcal{P}_{0}} and for even k>0k>0

𝒞𝒫k\displaystyle{\mathcal{C}}{\mathcal{P}_{k}} =(1)(k2)/22κn1(k3)!!(n1)!!(k+n1)!!𝒫k=μn,k𝒫k.\displaystyle=(-1)^{(k-2)/2}2\kappa_{n-1}\frac{(k-3)!!(n-1)!!}{(k+n-1)!!}{\mathcal{P}_{k}}=\mu_{n,k}{\mathcal{P}_{k}}.

where (k+2)!!=(k+2)(k!!)(k+2)!!=(k+2)(k!!). Since the cosine transform is defined as an integral over a symmetric domain, 𝒞𝒫k=0{\mathcal{C}}{\mathcal{P}_{k}}=0 when kk is odd.

When the body KK is placed in the isotropic position, we have the following estimate of the second harmonic of ρK\rho_{K}, obtained in [1, Sec. 9]. Let r0=𝒫0ρK=𝒮n1ρK𝑑σr_{0}={\mathcal{P}_{0}}\rho_{K}=\int_{{\mathcal{S}^{n-1}}}\rho_{K}d\sigma be the mean value of ρK\rho_{K} on the sphere, and let ρKr0=𝒫2(ρK)+𝒫4(ρK)+\rho_{K}-r_{\scriptscriptstyle{0}}={\mathcal{P}_{2}}(\rho_{K})+{\mathcal{P}_{4}}(\rho_{K})+\dots be the spherical harmonic decomposition of ρKr0\rho_{K}-r_{\scriptscriptstyle{0}}. From the definition (10) of the isotropic position, we have that if p(x)p(x) is a quadratic harmonic polynomial, p(x)=i,jaijxixjp(x)=\sum\limits_{i,j}a_{ij}x_{i}x_{j} with i=1naii=0\sum\limits_{i=1}^{n}a_{ii}=0, then

0=Kp(x)𝑑x=cn𝒮n1ρKn+2(x)p(x)𝑑σ(x).0=\int\limits_{K}p(x)dx=c_{n}\int\limits_{{\mathcal{S}^{n-1}}}\rho_{K}^{n+2}(x)p(x)d\sigma(x).

This means ρKn+2\rho_{K}^{n+2} has no second order term in its spherical harmonic decomposition. If KK, additionally, was also close to the Euclidean ball in the sense |ρKr0|<ϵ\left\lvert\rho_{K}-r_{0}\right\rvert<\epsilon, then 𝒫2ρK{\mathcal{P}_{2}}\rho_{K} should also be small in the L2(𝒮n1)L^{2}({\mathcal{S}^{n-1}}) sense. Indeed, estimating the error of the first order Taylor polynomial of (x+r0)n+2(x+r_{0})^{n+2} at ρKr0\rho_{K}-r_{0}, yields

|ρKn+2(r0n+2+(n+2)r0n+1(ρKr0))|Cϵ|ρKr0|.|\rho_{K}^{n+2}-(r_{\scriptscriptstyle{0}}^{n+2}+(n+2)r_{\scriptscriptstyle{0}}^{n+1}(\rho_{K}-r_{\scriptscriptstyle{0}}))|\leq C\epsilon|\rho_{K}-r_{\scriptscriptstyle{0}}|.

The second harmonic 𝒫2(ρK){\mathcal{P}_{2}}(\rho_{K}) of ρK\rho_{K} can now be estimated from ρKn+2\rho_{K}^{n+2} using the previous equation. Applying the L2(𝒮n1)L^{2}({\mathcal{S}^{n-1}}) to each side and using the orthogonality of the spherical harmonic spaces, we obtain

(n+2)r0n+1𝒫2(ρK)L2(Sn1)CϵρKr0L2(Sn1),(n+2)r_{\scriptscriptstyle{0}}^{n+1}\|{\mathcal{P}_{2}}(\rho_{K})\|_{L^{2}(S^{n-1})}\leq C\epsilon\|\rho_{K}-r_{\scriptscriptstyle{0}}\|_{L^{2}(S^{n-1})},

which yields that

(11) 𝒫2(ρK)L2(Sn1)CϵρKr0L2(Sn1).\|{\mathcal{P}_{2}}(\rho_{K})\|_{L^{2}(S^{n-1})}\leq C^{\prime}\epsilon\|\rho_{K}-r_{\scriptscriptstyle{0}}\|_{L^{2}(S^{n-1})}.

3.1. Linear approximation of 𝒞ρKn+1{\mathcal{C}}\rho_{K}^{n+1}

In [1, Sec. 9], an estimate on the linear component of the operator ([ρK]α)β({\mathcal{R}}[\rho_{K}]^{\alpha})^{\beta} is obtained. In this subsection we derive a similar estimate for the operator [𝒞ρKα]β[{\mathcal{C}}\rho_{K}^{\alpha}]^{\beta} that appears in our problem.

Let c,γ{0}c,\gamma\in{\mathbb{R}}\setminus\{0\}, we know that the first order Taylor polynomial of xγx^{\gamma} at cc is

(12) xγcγ+γcγ1(xc).x^{\gamma}\approx c^{\gamma}+\gamma c^{\gamma-1}(x-c).

Take r0=𝒫0ρKr_{0}={\mathcal{P}_{0}}\rho_{K} be the constant cc, γ=α\gamma=\alpha, and ρK\rho_{K} be xx. We can then estimate the error of the first order approximation of ρKα\rho_{K}^{\alpha} when 1ϵ<ρK<1+ϵ1-\epsilon<\rho_{K}<1+\epsilon and ϵ>0\epsilon>0 is small. In particular, we see the Lipschitz constant CαC_{\alpha} appear in the error bound,

|ρKα(r0α+αr0α1(ρKr0))|Cαϵ|ρKr0|.\left\lvert\rho_{K}^{\alpha}-(r_{0}^{\alpha}+\alpha r_{0}^{\alpha-1}(\rho_{K}-r_{0}))\right\rvert\leq C_{\alpha}\epsilon\left\lvert\rho_{K}-r_{0}\right\rvert.

Observe that for any function f0f\geq 0, 𝒞f0{\mathcal{C}}f\geq 0. That is 𝒞{\mathcal{C}} is a positive operator in the vector space sense. Using the fact that 𝒞{\mathcal{C}} is a positive linear operator,

𝒞|ρKα(r0α+αr0α1(ρKr0))|Cαϵ𝒞|ρKr0|.{\mathcal{C}}\left\lvert\rho_{K}^{\alpha}-(r_{0}^{\alpha}+\alpha r_{0}^{\alpha-1}(\rho_{K}-r_{0}))\right\rvert\leq C_{\alpha}\epsilon\,{\mathcal{C}}\left\lvert\rho_{K}-r_{0}\right\rvert.

As positive linear operators admit a triangle inequality |𝒞f|𝒞|f|\left\lvert{\mathcal{C}}f\right\rvert\leq{\mathcal{C}}\left\lvert f\right\rvert, we get

|𝒞ρKα(𝒞r0α+αr0α1𝒞(ρKr0))|Cαϵ𝒞|ρKr0|.\left\lvert{\mathcal{C}}\rho_{K}^{\alpha}-({\mathcal{C}}r_{0}^{\alpha}+\alpha r_{0}^{\alpha-1}{\mathcal{C}}(\rho_{K}-r_{0}))\right\rvert\leq C_{\alpha}\epsilon\,{\mathcal{C}}\left\lvert\rho_{K}-r_{0}\right\rvert.

Finally, let us examine the first order Taylor polynomial (12) again but with x=𝒞ρKαx={\mathcal{C}}\rho_{K}^{\alpha}, c=𝒞r0αc={\mathcal{C}}r_{0}^{\alpha}, and γ=β\gamma=\beta.

(𝒞ρKα)β\displaystyle({\mathcal{C}}\rho_{K}^{\alpha})^{\beta} (𝒞r0α)β+β(𝒞r0α)β1[𝒞(ρKαr0α)]\displaystyle\approx({\mathcal{C}}r_{0}^{\alpha})^{\beta}+\beta({\mathcal{C}}r_{0}^{\alpha})^{\beta-1}\left[{\mathcal{C}}(\rho_{K}^{\alpha}-r_{0}^{\alpha})\right]
(𝒞r0α)β+β(𝒞r0α)β1[αr0α1𝒞(ρKr0)]\displaystyle\approx({\mathcal{C}}r_{0}^{\alpha})^{\beta}+\beta({\mathcal{C}}r_{0}^{\alpha})^{\beta-1}\left[\alpha r_{0}^{\alpha-1}{\mathcal{C}}(\rho_{K}-r_{0})\right]

Estimating the error of the approximation, we then see

(13) |(𝒞ρKα)β(r0αβ(𝒞1)β+αβ(𝒞1)β1r0αβ1𝒞(ρKr0))|Cα,βϵ𝒞|ρKr0|,\left\lvert({\mathcal{C}}\rho_{K}^{\alpha})^{\beta}-(r_{0}^{\alpha\beta}({\mathcal{C}}1)^{\beta}+\alpha\beta({\mathcal{C}}1)^{\beta-1}r_{0}^{\alpha\beta-1}{\mathcal{C}}(\rho_{K}-r_{0}))\right\rvert\leq C_{\alpha,\beta}\,\epsilon\,{\mathcal{C}}\left\lvert\rho_{K}-r_{0}\right\rvert,

where the Lipschitz constant Cα,βC_{\alpha,\beta} is dependent on the value of 𝒞1{\mathcal{C}}1, α\alpha, β\beta, and the maximum value of ϵ\epsilon allowed. Notably, this is bounded Note that in the estimate of (13) only the positivity of the linear operator 𝒞{\mathcal{C}} was used. Considering some other positive linear operator {\mathcal{M}} (positive in the f0{\mathcal{M}}f\geq 0 when f0f\geq 0 sense), the particular constant Cα,βC_{\alpha,\beta} will change but an identical estimate holds for [ρKα]β[{\mathcal{M}}\rho_{K}^{\alpha}]^{\beta}.

The next lemma requires a definition. Let {\mathcal{M}} be a bounded linear operator on L2(𝒮n1)L^{2}({\mathcal{S}^{n-1}}) whose eigenspaces are precisely kn{{\mathcal{H}}_{k}^{n}}, that is, there exist eigenvalues μk\mu_{k} such that for all fL2f\in L^{2},

f=k0μk𝒫kf.{\mathcal{M}}f=\sum_{k\geq 0}\mu_{k}{\mathcal{P}_{k}}f.

Then {\mathcal{M}} is said to be a strong contraction if

maxk0|μk|<1, and limkinfμk=0.\max_{k\geq 0}\left\lvert\mu_{k}\right\rvert<1,\hskip 20.0pt\text{ and }\hskip 20.0pt\lim_{k\rightarrow\inf}\mu_{k}=0.

It should be noted that small enough constant multiples of the cosine and Radon transforms are strong contractions, for example. The main result about the strong contraction is [1, Lemma 4], stated below.

Lemma 4. Assume {\mathcal{M}} is a strong contraction. Then there exists ϵ(0,1)\epsilon\in(0,1) such that for any symmetric convex body KK and any c(1ϵ,1+ϵ)c\in(1-\epsilon,1+\epsilon), the conditions

1ϵρK1+ϵ,(hKh0)c(ρKr0)L2ϵρKr0L2,1-\epsilon\leq\rho_{K}\leq 1+\epsilon,\hskip 20.0pt{\left\|(h_{K}-h_{0})-c{\mathcal{M}}(\rho_{K}-r_{0})\right\|}_{L^{2}}\leq\epsilon{\left\|\rho_{K}-r_{0}\right\|}_{L^{2}},

imply hK=ρK=h_{K}=\rho_{K}=const. Here h0h_{0} and r0r_{0} are the constant terms of the spherical harmonic decomposition.

4. Proof of Theorem 1

Theorem 1. There exists an ϵ2>0\epsilon_{2}>0, such that the only convex bodies KK satisfying (1) with dBM(B,K)<ϵ1d_{BM}(B,K)<\epsilon_{1} are ellipsoids.

Proof.

This will be proven in two steps, first under certain assumptions on KK and then in general. First, we will assume that KK satisfies (1), is close to the sphere (in the sense that 1ϵρK1+ϵ1-\epsilon\leq\rho_{K}\leq 1+\epsilon), and is placed in the isotropic position with voln(K)=κn\text{vol}_{n}(K)=\kappa_{n}.

Using the same approach as in the papers [1, 5, 12], we will separate the operator into linear and non-linear components and examine each separately. With our assumptions on KK, equation (5) reads as

hK=CK(n+1)κn𝒞ρKn+1.h_{K}=\frac{C_{K}}{(n+1)\kappa_{n}}{\mathcal{C}}\rho_{K}^{n+1}.

As we saw in Section 3, we write ρK=r0+γ\rho_{K}=r_{0}+\gamma, where r0r_{0} is the zero harmonic of ρK\rho_{K}. Since 𝒞{\mathcal{C}} is a linear operator and constants are eigenvectors of it, we have

(14) hK=CK(n+1)κn𝒞(r0+γ)n+1=CKCBr0n+1+CKr0nκn𝒞γ+G1,h_{K}=\frac{C_{K}}{(n+1)\kappa_{n}}{\mathcal{C}}(r_{0}+\gamma)^{n+1}=\frac{C_{K}}{C_{B}}r_{0}^{n+1}+\frac{C_{K}r_{0}^{n}}{\kappa_{n}}{\mathcal{C}}\gamma+G_{1},

where G1G_{1} contains all the higher order terms. Thus, the linear component may be rewritten as,

(15) r0nCKκn𝒞γ=(r0nCKCB)(CBκn𝒞)γ.r_{0}^{n}\frac{C_{K}}{\kappa_{n}}{\mathcal{C}}\gamma=\left(r_{0}^{n}\frac{C_{K}}{C_{B}}\right)\left(\frac{C_{B}}{\kappa_{n}}{\mathcal{C}}\right)\gamma.

Estimating the higher order terms, we solve for G1G_{1} in (14) and take L2L^{2} norms yielding

G1L2=[hKr0n+1CKCB](r0nCKCB)(CBκn𝒞)γL2.\displaystyle{\left\|G_{1}\right\|}_{L^{2}}={\left\|\left[h_{K}-r_{0}^{n+1}\frac{C_{K}}{C_{B}}\right]-\left(r_{0}^{n}\frac{C_{K}}{C_{B}}\right)\left(\frac{C_{B}}{\kappa_{n}}{\mathcal{C}}\right)\gamma\right\|}_{L^{2}}.

Using (13) with α=n+1,β=1\alpha=n+1,\beta=1 we obtain

G1L2\displaystyle{\left\|G_{1}\right\|}_{L^{2}} Cn+1,1CKCBϵCB(n+1)κn𝒞|γ|L2\displaystyle\leq C_{n+1,1}\frac{C_{K}}{C_{B}}\epsilon{\left\|\frac{C_{B}}{(n+1)\kappa_{n}}{\mathcal{C}}\left\lvert\gamma\right\rvert\right\|}_{L^{2}}
Cn+1,1CKCBϵγL2.\displaystyle\leq C_{n+1,1}\frac{C_{K}}{C_{B}}\epsilon{\left\|\gamma\right\|}_{L^{2}}.

In the notation of Lemma 4, we have h0=r0n+1CKCB1h_{0}=r_{0}^{n+1}C_{K}C_{B}^{-1}, c=r0nCKCB1c=r_{0}^{n}C_{K}C_{B}^{-1}, and =CBκn1(𝒞𝒞𝒫0𝒞𝒫2){\mathcal{M}}=C_{B}\kappa_{n}^{-1}({\mathcal{C}}-{\mathcal{C}}{\mathcal{P}_{0}}-{\mathcal{C}}{\mathcal{P}_{2}}). The degree zero harmonic of ρKr0\rho_{K}-r_{0} is zero, and the second harmonic of γ\gamma has already been estimated by (11). The eigenvalues of operator {\mathcal{M}} are scalar multiples of the eigenvalues of the cosine transform starting with degree 4. Since the even degree eigenvalues decay monotonically to zero in absolute value, we only need to check that the first non-zero eigenvalue is less than one in absolute value. Indeed, 𝒫4=(n+3)1𝒫4{\mathcal{M}}{\mathcal{P}_{4}}=-(n+3)^{-1}{\mathcal{P}_{4}}, and {\mathcal{M}} is a strong contraction as desired. By lemma 4, there is some 0<ϵ00<\epsilon_{0} where if KK satisfies,

  • 1ϵ0ρK1+ϵ01-\epsilon_{0}\leq\rho_{K}\leq 1+\epsilon_{0},

  • c(1ϵ0,1+ϵ0)c\in(1-\epsilon_{0},1+\epsilon_{0}),

  • (hKh0)c(ρKr0)L2ϵ0ρKr0L2{\left\|(h_{K}-h_{0})-c{\mathcal{M}}(\rho_{K}-r_{0})\right\|}_{L^{2}}\leq\epsilon_{0}{\left\|\rho_{K}-r_{0}\right\|}_{L^{2}},

KK must be the Euclidean ball.

First, we note that the constant cc is bounded by the product of the bounds on r0nr_{0}^{n} and the bounds on the ratio CB/CKC_{B}/C_{K}, giving c=1+O(ϵ)c=1+O(\epsilon). If ϵ\epsilon is sufficiently small 1ϵ0<c<1+ϵ01-\epsilon_{0}<c<1+\epsilon_{0} follows. The second criteria follows if ϵ<ϵ0\epsilon<\epsilon_{0}. Checking the third criteria,

[hKh0]c(ρKr0)L2{\left\|\left[h_{K}-h_{0}\right]-c{\mathcal{M}}\left(\rho_{K}-r_{0}\right)\right\|}_{L^{2}}
[hKh0]c(CBκn𝒞)γL2+cCBκn𝒞𝒫2γL2\leq{\left\|\left[h_{K}-h_{0}\right]-c\left(\frac{C_{B}}{\kappa_{n}}{\mathcal{C}}\right)\gamma\right\|}_{L^{2}}+c{\left\|\frac{C_{B}}{\kappa_{n}}{\mathcal{C}}{\mathcal{P}_{2}}\gamma\right\|}_{L^{2}}
(Cn+1,1CKCB+cCn+2r0n+1)ϵγL2.\leq\bigg{(}C_{n+1,1}\frac{C_{K}}{C_{B}}+c\frac{C_{n+2}}{r_{0}^{n+1}}\bigg{)}\epsilon{\left\|\gamma\right\|}_{L^{2}}.

Now we need to estimate the expression in parentheses. We have that

Cn+1,1CKCB+cCn+2r0n+1(n+1)(1+ϵ)2n+1(1ϵ)n+1+(1+ϵ0)(n+2)(1+ϵ)n+1(1ϵ)n+1C_{n+1,1}\frac{C_{K}}{C_{B}}+c\frac{C_{n+2}}{r_{0}^{n+1}}\leq(n+1)\frac{(1+\epsilon)^{2n+1}}{(1-\epsilon)^{n+1}}+(1+\epsilon_{0})(n+2)\frac{(1+\epsilon)^{n+1}}{(1-\epsilon)^{n+1}}

is bounded above. Hence, the product of this bound with ϵ\epsilon is bounded above by ϵ0\epsilon_{0}, provided ϵ\epsilon is small enough.

Now we prove Theorem 1 for all bodies close to the ball. Let 0<ϵ10<\epsilon_{1} be the maximal ϵ\epsilon for which the additional criteria for Lemma 4 hold in the isotropic case. Put

ϵ2=max{ϵ>0||1(1±ϵ)(1±ϵ1ϵ)n+22|ϵ1}.\epsilon_{2}=\max\left\{\epsilon>0\hskip 2.0pt\bigg{\rvert}\hskip 2.0pt\left\lvert 1-(1\pm\epsilon)\left(\frac{1\pm\epsilon}{1\mp\epsilon}\right)^{\frac{n+2}{2}}\right\rvert\leq\epsilon_{1}\right\}.

If KK is a convex body with dBM(B,K)ϵ2d_{BM}(B,K)\leq\epsilon_{2} and satisfying (1), then there exists a linear transformation TGLnT\in\text{GL}_{n} where TKTK satisfies the assumptions of the first part of the proof. Therefore, TKTK is a dilate of the Euclidean ball and KK is an ellipsoid. ∎

5. Generalization to pp-centroid bodies: Proof of Theorem 2

This section introduces pp-centroid body and extends the result of the previous section to p>1p>1. Subsection 5.1 gives the definition of the pp-centroid body and an analytic reformulation in terms of the pp-cosine transform. As the differences between the prior case and the p>1p>1 are purely technical, the subsequent subsections include relevant details for the extension of Theorem 1 to p>1p>1 for completeness.

5.1. Definition

The pp-centroid body of a star body, ΓpK\Gamma_{p}K is defined in a similar way to the centroid body by,

hΓpK(θ):=(K|θx|p𝑑x)1p=1(n+p)voln(K)p(𝒞pρKn+p(θ))1p,θ𝒮n1h_{\Gamma_{p}K}(\theta):=\left(\mathop{}\mkern-3.0mu\mathchoice{\hbox to0.0pt{$\displaystyle\vbox{\hbox to6.44583pt{\hss$\textstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\textstyle\vbox{\hbox to6.44583pt{\hss$\scriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptstyle\vbox{\hbox to4.60416pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptscriptstyle\vbox{\hbox to4.60416pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}\mkern-3.0mu\int_{K}\left\lvert\theta\cdot x\right\rvert^{p}dx\right)^{\frac{1}{p}}=\frac{1}{\sqrt[p]{(n+p)\text{vol}_{n}(K)}}\left({\mathcal{C}}^{p}\rho_{K}^{n+p}(\theta)\right)^{\frac{1}{p}},\hskip 20.0pt\forall\theta\in{\mathcal{S}^{n-1}}

where \mathop{}\mkern-3.0mu\mathchoice{\hbox to0.0pt{$\displaystyle\vbox{\hbox to5.83331pt{\hss$\textstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\textstyle\vbox{\hbox to5.83331pt{\hss$\scriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptstyle\vbox{\hbox to4.5833pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptscriptstyle\vbox{\hbox to3.74997pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}\mkern-3.0mu\int denotes the integral average K𝑑x=1voln(K)K𝑑x\mathop{}\mkern-3.0mu\mathchoice{\hbox to0.0pt{$\displaystyle\vbox{\hbox to6.44583pt{\hss$\textstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\textstyle\vbox{\hbox to6.44583pt{\hss$\scriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptstyle\vbox{\hbox to4.60416pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptscriptstyle\vbox{\hbox to4.60416pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}\mkern-3.0mu\int_{K}\,dx=\frac{1}{\text{vol}_{n}(K)}\int_{K}\,dx and 𝒞p{\mathcal{C}}^{p} is the pp-cosine transform,

𝒞p(f):=𝒮n1|ϕθ|pf(θ)𝑑σn1(θ).{\mathcal{C}}^{p}(f):=\int_{{\mathcal{S}^{n-1}}}\left\lvert\phi\cdot\theta\right\rvert^{p}f(\theta)\hskip 2.0ptd\sigma_{n-1}(\theta).

The pp-centroid bodies were introduced in [8]. For p1p\geq 1, the function hΓpKh_{\Gamma_{p}K} is the support function of a convex body, so ΓpK\Gamma_{p}K is well defined. In particular for p=1p=1, this is the centroid body studied in the previous sections, and for p=2p=2 this is the so called Legendre ellipsoid. For p=2p=2, the relation CKΓ2K=KC_{K}\Gamma_{2}K=K trivially implies that KK is an ellipsoid.

For p>1p>1, p2p\neq 2, we will study if

CKΓpK=KC_{K}\Gamma_{p}K=K

for some scalar CKC_{K} implies that KK is an ellipsoid, when KK is close enough to the Euclidean ball in the Banach-Mazur distance. The analytic reformulation of the above equation is

(16) CK(n+p)voln(K)p(𝒞pρKn+p(θ))1p=hK(θ).\frac{C_{K}}{\sqrt[p]{(n+p)\text{vol}_{n}(K)}}\left({\mathcal{C}}^{p}\rho_{K}^{n+p}(\theta)\right)^{\frac{1}{p}}=h_{K}(\theta).

5.2. Linear transformations and Stability

As in the p=1p=1 case, equation (16) is invariant under linear transformations. If TGLnT\in\text{GL}_{n}, then

hΓpTK(θ)=(TK|θx|p𝑑x)1p=(K|Tθx|p𝑑x)1p=hTΓpK(θ).h_{\Gamma_{p}TK}(\theta)=\left(\mathop{}\mkern-3.0mu\mathchoice{\hbox to0.0pt{$\displaystyle\vbox{\hbox to11.50868pt{\hss$\textstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\textstyle\vbox{\hbox to11.50868pt{\hss$\scriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptstyle\vbox{\hbox to8.22047pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptscriptstyle\vbox{\hbox to8.22047pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}\mkern-3.0mu\int_{TK}\left\lvert\theta\cdot x\right\rvert^{p}dx\right)^{\frac{1}{p}}=\left(\mathop{}\mkern-3.0mu\mathchoice{\hbox to0.0pt{$\displaystyle\vbox{\hbox to6.44583pt{\hss$\textstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\textstyle\vbox{\hbox to6.44583pt{\hss$\scriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptstyle\vbox{\hbox to4.60416pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}{\hbox to0.0pt{$\scriptscriptstyle\vbox{\hbox to4.60416pt{\hss$\scriptscriptstyle{-}$\hss}}$\hss}}\mkern-3.0mu\int_{K}\left\lvert T^{*}\theta\cdot x\right\rvert^{p}dx\right)^{\frac{1}{p}}=h_{T\Gamma_{p}K}(\theta).

Because the pp-centroid body operator commutes with linear transformations, if KK satisfies (16) then so does TKTK, and CTK=CKC_{TK}=C_{K}. Since the unit Euclidean ball BB satisfies (16), we have that ellipsoids also satisfy this equation with the constant

1=CB(n+p)κnp(𝒞p1)1p,or equivalently,CB=(n+p)κnp(𝒞p1)1/p.1=\frac{C_{B}}{\sqrt[p]{(n+p)\kappa_{n}}}({\mathcal{C}}^{p}1)^{\frac{1}{p}},\hskip 20.0pt\text{or equivalently,}\hskip 20.0ptC_{B}=\frac{\sqrt[p]{(n+p)\kappa_{n}}}{({\mathcal{C}}^{p}1)^{1/p}}.

Since CKC_{K} in (16) is invariant under linear transformations of the body, we cannot choose a particular value for this constant by picking an appropriate linear transformation. On the other hand, we can obtain stability bounds on CKC_{K} in same way as in the case p=1p=1. Let KK be close to the Euclidean ball, that is for some ϵ>0\epsilon>0,

(17) 1ϵρK1+ϵ,1ϵhK1+ϵ.1-\epsilon\leq\rho_{K}\leq 1+\epsilon,\hskip 40.0pt1-\epsilon\leq h_{K}\leq 1+\epsilon.

This implies

(18) 1(1+ϵ)npκnvoln(K)p1(1ϵ)np.\frac{1}{(1+\epsilon)^{\frac{n}{p}}}\leq\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}\leq\frac{1}{(1-\epsilon)^{\frac{n}{p}}}.

Applying the pp-cosine transform to the estimate on the radial function in (17), we get

CB(n+p)κnp(𝒞p(1ϵ)n+p)1pCB(n+p)κnp(𝒞pρKn+p)1p\frac{C_{B}}{\sqrt[p]{(n+p)\kappa_{n}}}\left({\mathcal{C}}^{p}(1-\epsilon)^{n+p}\right)^{\frac{1}{p}}\leq\frac{C_{B}}{\sqrt[p]{(n+p)\kappa_{n}}}\left({\mathcal{C}}^{p}\rho_{K}^{n+p}\right)^{\frac{1}{p}}
CB(n+p)κnp(𝒞p(1+ϵ)n+p)1p.\leq\frac{C_{B}}{\sqrt[p]{(n+p)\kappa_{n}}}\left({\mathcal{C}}^{p}(1+\epsilon)^{n+p}\right)^{\frac{1}{p}}.

Combining this estimate with (18) and with equation (16) we have

(1ϵ)n+ppCBCKvoln(K)κnphK(1+ϵ)n+pp,(1-\epsilon)^{\frac{n+p}{p}}\leq\frac{C_{B}}{C_{K}}\sqrt[p]{\frac{\text{vol}_{n}(K)}{\kappa_{n}}}\,h_{K}\leq(1+\epsilon)^{\frac{n+p}{p}},
(1ϵ)np+1(1+ϵ)npCBCKhK(1+ϵ)np+1(1ϵ)np,\frac{(1-\epsilon)^{\frac{n}{p}+1}}{(1+\epsilon)^{\frac{n}{p}}}\leq\frac{C_{B}}{C_{K}}\,h_{K}\leq\frac{(1+\epsilon)^{\frac{n}{p}+1}}{(1-\epsilon)^{\frac{n}{p}}},

and, since 1ϵhK1+ϵ1-\epsilon\leq h_{K}\leq 1+\epsilon,

(1ϵ1+ϵ)np+1CBCK(1+ϵ1ϵ)np+1.\left(\frac{1-\epsilon}{1+\epsilon}\right)^{\frac{n}{p}+1}\leq\frac{C_{B}}{C_{K}}\leq\left(\frac{1+\epsilon}{1-\epsilon}\right)^{\frac{n}{p}+1}.

5.3. The eigenvalues of 𝒞p{\mathcal{C}}^{p}

The eigenspaces of the pp-cosine transform are, as in the case of the cosine transform, the spaces of spherical harmonics of degree kk. From [10, Eq. (3.4)] we have that the eigenvalues 𝒞α𝒫k=mα,k𝒫k{\mathcal{C}}^{\alpha}{\mathcal{P}_{k}}=m_{\alpha,k}{\mathcal{P}_{k}} for p2,4,p\neq 2,4,\ldots are equal to

mp,k={2(1)k/2π(n1)/2σn1Γ((kp)/2)Γ(p/2)Γ((p+1)/2)Γ((k+n+p)/2):2k0:2k.m_{p,k}=\begin{cases}\frac{2(-1)^{k/2}\pi^{(n-1)/2}}{\sigma_{n-1}}\frac{\Gamma((k-p)/2)}{\Gamma(-p/2)}\frac{\Gamma((p+1)/2)}{\Gamma((k+n+p)/2)}&:2\mid k\\ 0&:2\nmid k\end{cases}.

When pp is an even integer and k>pk>p, the eigenvalues are 0. The above formula for the eigenvalues mp,km_{p,k} can be rewritten so that it works also when pp is an even integer and kpk\leq p. (For the reader who wants to consult reference [10], we note that in that paper the pp-cosine transform is denoted by Mα+1M^{\alpha+1}).

Observe that |mp,2k+2/mp,2k|=|2kp|/(2k+n+p)\left\lvert m_{p,2k+2}/m_{p,2k}\right\rvert=\left\lvert 2k-p\right\rvert/(2k+n+p). Expressing the higher order non-zero harmonics as a telescoping product of the lower order ones, we see that

|mp,2k|=mp,0=0k1|mp,2+2mp,2|=mp,0=0k1|2p2+n+p|,\left\lvert m_{p,2k}\right\rvert=m_{p,0}\prod_{\ell=0}^{k-1}\left\lvert\frac{m_{p,2\ell+2}}{m_{p,2\ell}}\right\rvert=m_{p,0}\prod_{\ell=0}^{k-1}\left\lvert\frac{2\ell-p}{2\ell+n+p}\right\rvert,

hence the non-zero eigenvalues are absolutely strictly decreasing as kk\rightarrow\infty, and their limit is 0. The asymptotic behaviour as pp varies is similar, for mp,00m_{p,0}\rightarrow 0 as pp\rightarrow\infty.

5.4. Linear approximation error

The pp-cosine tranansform is a positive linear operator, therefore if we set α=n+p\alpha=n+p, β=1/p\beta=1/p on the left hand side of (13), we have

|(𝒞pρKn+p)1pr0n+pp(𝒞p1)1pn+pp(𝒞p1)1ppr0np𝒞p(ρKr0))|.\displaystyle\left\lvert({\mathcal{C}}^{p}\rho_{K}^{n+p})^{\frac{1}{p}}-r_{0}^{\frac{n+p}{p}}({\mathcal{C}}^{p}1)^{\frac{1}{p}}-\frac{n+p}{p}({\mathcal{C}}^{p}1)^{\frac{1-p}{p}}r_{0}^{\frac{n}{p}}{\mathcal{C}}^{p}(\rho_{K}-r_{0}))\right\rvert.

Taking the L2L^{2} norm of this term and multiplying by the missing constants in equation (16), the left hand side can be rewritten as

(19) ||(hKr0n+ppCKCBCB(n+p)κnp(𝒞p1)1p)\bigg{\rvert}\bigg{\lvert}\left(h_{K}-r_{0}^{\frac{n+p}{p}}\frac{C_{K}}{C_{B}}\frac{C_{B}}{\sqrt[p]{(n+p)\kappa_{n}}}({\mathcal{C}}^{p}1)^{\frac{1}{p}}\right)
(r0npn+ppCK(n+p)κnp[(𝒞p1)1p]1p𝒞p)(ρKr0)||L2.-\left(r_{0}^{\frac{n}{p}}\,\frac{n+p}{p}\frac{C_{K}}{\sqrt[p]{(n+p)\kappa_{n}}}\left[({\mathcal{C}}^{p}1)^{\frac{1}{p}}\right]^{1-p}{\mathcal{C}}^{p}\right)(\rho_{K}-r_{0})\bigg{\rvert}\bigg{\lvert}_{L^{2}}.

Recalling that CB[(n+p)κn]1/p(𝒞p1)1/p=1C_{B}[(n+p)\kappa_{n}]^{-1/p}({\mathcal{C}}^{p}1)^{1/p}=1, we can simplify the last term, obtaining

(r0npn+ppCK(n+p)κnp(CB(n+p)κnp)p1𝒞p)(ρKr0)\left(r_{0}^{\frac{n}{p}}\frac{n+p}{p}\frac{C_{K}}{\sqrt[p]{(n+p)\kappa_{n}}}\left(\frac{C_{B}}{\sqrt[p]{(n+p)\kappa_{n}}}\right)^{p-1}{\mathcal{C}}^{p}\right)(\rho_{K}-r_{0})
=(r0npCKCB)(n+ppCBp(n+p)κn𝒞p)(ρKr0)=\left(r_{0}^{\frac{n}{p}}\frac{C_{K}}{C_{B}}\right)\left(\frac{n+p}{p}\frac{C_{B}^{p}}{(n+p)\kappa_{n}}{\mathcal{C}}^{p}\right)(\rho_{K}-r_{0})
=(r0npCKCB)(CBppκn𝒞p)(ρKr0).=\left(r_{0}^{\frac{n}{p}}\frac{C_{K}}{C_{B}}\right)\left(\frac{C_{B}^{p}}{p\,\kappa_{n}}{\mathcal{C}}^{p}\right)(\rho_{K}-r_{0}).

Thus, (19) simplifies to

(20) (hKr0n+ppCKCB)(r0npCKCB)(CBppκn𝒞p)(ρKr0)L2.{\left\|\left(h_{K}-r_{0}^{\frac{n+p}{p}}\frac{C_{K}}{C_{B}}\right)-\left(r_{0}^{\frac{n}{p}}\frac{C_{K}}{C_{B}}\right)\left(\frac{C_{B}^{p}}{p\kappa_{n}}{\mathcal{C}}^{p}\right)(\rho_{K}-r_{0})\right\|}_{L^{2}}.

Next, we will consider the right hand side of (13). After multiplying by CK[(n+p)κn]1/pC_{K}[(n+p)\kappa_{n}]^{-1/p} and taking L2L^{2} norms, it can be rewritten as

Cn+p,1p𝒞p1ϵ(CK(n+p)κnp)𝒞p|ρKr0|L2C_{n+p,\frac{1}{p}}\,{\mathcal{C}}^{p}1\,\epsilon\left(\frac{C_{K}}{\sqrt[p]{(n+p)\kappa_{n}}}\right){\left\|{\mathcal{C}}^{p}\left\lvert\rho_{K}-r_{0}\right\rvert\right\|}_{L^{2}}
=Cn+p,1p𝒞p1ϵCKCB(CB(n+p)κnp)1pCBp(n+p)κn𝒞p|ρKr0|L2=C_{n+p,\frac{1}{p}}{\mathcal{C}}^{p}1\,\epsilon\frac{C_{K}}{C_{B}}\left(\frac{C_{B}}{\sqrt[p]{(n+p)\kappa_{n}}}\right)^{1-p}{\left\|\frac{C_{B}^{p}}{(n+p)\kappa_{n}}{\mathcal{C}}^{p}\left\lvert\rho_{K}-r_{0}\right\rvert\right\|}_{L^{2}}
(21) =Cn+p,1pCKCB(𝒞p1)21pϵCBp(n+p)κn𝒞p|ρKr0|L2.=C_{n+p,\frac{1}{p}}\,\frac{C_{K}}{C_{B}}\left({\mathcal{C}}^{p}1\right)^{2-\frac{1}{p}}\epsilon{\left\|\frac{C_{B}^{p}}{(n+p)\kappa_{n}}{\mathcal{C}}^{p}\left\lvert\rho_{K}-r_{0}\right\rvert\right\|}_{L^{2}}.

Note that the exponent 21p2-\frac{1}{p} on 𝒞p1{\mathcal{C}}^{p}1 is non-negative for p1p\geq 1 and that

𝒞p1\displaystyle{\mathcal{C}}^{p}1 =2π(n1)/2σn1Γ(p+12)Γ(n+p2)=1πΓ(n2)Γ(p+12)Γ(n+p2).\displaystyle=\frac{2\pi^{(n-1)/2}}{\sigma_{n-1}}\frac{\Gamma(\frac{p+1}{2})}{\Gamma(\frac{n+p}{2})}=\frac{1}{\sqrt{\pi}}\frac{\Gamma(\frac{n}{2})\Gamma(\frac{p+1}{2})}{\Gamma(\frac{n+p}{2})}.

We claim that this expression is bounded above by 11 when the dimension nn is greater than 22. This is obtained by using the property zΓ(z)=Γ(z+1)z\Gamma(z)=\Gamma(z+1) several times to yield

𝒞p1=[Γ(n2n2+1)π][=0n21(n2n+p2)][Γ(p+12)Γ(p2+n2n2+1)].{\mathcal{C}}^{p}1=\left[\frac{\Gamma\left(\frac{n}{2}-\left\lceil\frac{n}{2}\right\rceil+1\right)}{\sqrt{\pi}}\right]\left[\prod_{\ell=0}^{\left\lceil\frac{n}{2}\right\rceil-1}\left(\frac{n-2\ell}{n+p-2\ell}\right)\right]\left[\frac{\Gamma\left(\frac{p+1}{2}\right)}{\Gamma\left(\frac{p}{2}+\frac{n}{2}-\left\lceil\frac{n}{2}\right\rceil+1\right)}\right].

Breaking up the three terms in the product, we note that the leftmost is bounded by 11, the middle term is maximal when p=1p=1 and n=3n=3 (which are the minimal values of p,np,n), and the last term is also maximal when p=1p=1 and nn odd (by the convexity of Γ\Gamma), yielding

𝒞p1134121<1.{\mathcal{C}}^{p}1\leq 1\cdot\frac{3}{4}\cdot\frac{1}{2}\cdot 1<1.

Using this estimate, and the fact that the L2L^{2} norm of the normalized pp-cosine transform is 1, (21) is bounded by

Cn+p,1pϵCKCBρKr0L2.C_{n+p,\frac{1}{p}}\epsilon\frac{C_{K}}{C_{B}}{\left\|\rho_{K}-r_{0}\right\|}_{L^{2}}.

Therefore, combining this with (20), we have the following estimate that we will need to apply Lemma 4.

(22) (hKr0n+ppCKCB)(r0npCKCB)(CBppκn𝒞p)(ρKr0)L2{\left\|\left(h_{K}-r_{0}^{\frac{n+p}{p}}\frac{C_{K}}{C_{B}}\right)-\left(r_{0}^{\frac{n}{p}}\frac{C_{K}}{C_{B}}\right)\left(\frac{C_{B}^{p}}{p\kappa_{n}}{\mathcal{C}}^{p}\right)(\rho_{K}-r_{0})\right\|}_{L^{2}}
Cn+p,1pϵCKCBρKr0L2.\leq C_{n+p,\frac{1}{p}}\epsilon\frac{C_{K}}{C_{B}}{\left\|\rho_{K}-r_{0}\right\|}_{L^{2}}.

5.5. The proof of Theorem 2.

The last thing that needs to be checked is that the operator CBppκn𝒞p\frac{C^{p}_{B}}{p\kappa_{n}}{\mathcal{C}}^{p} in equation (22) is a strong contraction (note that the constant multiplying this operator r0npCKCBr_{0}^{\frac{n}{p}}\frac{C_{K}}{C_{B}} is of the order 1+O(ϵ)1+O(\epsilon), as discussed in subsection 5.2). As mentioned in Section 5.3, the eigenvalues of the pp-cosine transform for the harmonic spaces of even degree greater than or equal to 4 are less than 1, strictly decreasing in absolute value, and their limit is 0. Thus, it is enough to check the eigenvalue of the degree four harmonic space of the scaled pp-cosine transform CBp[pκn]1𝒞pC_{B}^{p}[p\kappa_{n}]^{-1}{\mathcal{C}}^{p}. The computation yields

|n+ppCBp(n+p)κnmp,4|=|mp,4mp,2|=|2p|n+2+p<1\left\lvert\frac{n+p}{p}\frac{C_{B}^{p}}{(n+p)\kappa_{n}}m_{p,4}\right\rvert=\left\lvert\frac{m_{p,4}}{m_{p,2}}\right\rvert=\frac{\left\lvert 2-p\right\rvert}{n+2+p}<1

From this point, the proof of Theorem 2 is identical to that of Theorem 1.

6. On Theorem 3

6.1. Invariance under linear transformations

Let CKC_{K}^{\prime} be the constant associated with a body KK that satisfies the condition of Theorem 3. If we evaluate (7) when KK is the Euclidean ball, we see

(1)\displaystyle({\mathcal{R}}1) =CB(n1)(𝒞p1(n+1)κn)1p,\displaystyle=C_{B}^{\prime}(n-1)\left(\frac{{\mathcal{C}}^{p}1}{{(n+1)\kappa_{n}}}\right)^{-\frac{1}{p}},
CB\displaystyle C_{B}^{\prime} =(1n1)(𝒞p1(n+1)κn)1p.\displaystyle=\left(\frac{{\mathcal{R}}1}{n-1}\right)\left(\frac{{\mathcal{C}}^{p}1}{(n+1)\kappa_{n}}\right)^{\frac{1}{p}}.

We next show that if KK satisfies cΓK=Kc\Gamma K={\mathcal{I}}^{*}K, so does TKTK.

ΓpTK\displaystyle\Gamma_{p}TK =TΓpK=TCK(K)=CK(TK)\displaystyle=T\Gamma_{p}K=TC_{K}^{\prime}({\mathcal{I}}K)^{*}=C_{K}^{\prime}(T^{-*}{\mathcal{I}}K)^{*}
=CK(|det(T)|1TK)=CK|detT|TK.\displaystyle=C_{K}^{\prime}(\left\lvert\det(T)\right\rvert^{-1}{\mathcal{I}}TK)^{*}=C_{K}^{\prime}\left\lvert\det T\right\rvert{\mathcal{I}}^{*}TK.

Therefore if KK satisfies cΓK=Kc\Gamma K={\mathcal{I}}^{*}K, so does TKTK, albeit for a different constant, CTK=CK|detT|C_{TK}^{\prime}=C_{K}^{\prime}\left\lvert\det T\right\rvert.

6.2. Stability of the constants.

Applying the normalized pp-cosine transform and Radon transform to 1ϵρK1+ϵ1-\epsilon\leq\rho_{K}\leq 1+\epsilon, we have

1(1+ϵ)n+pp(𝒞pρKn+p(n+p)voln(K))1p(𝒞p1(n+p)voln(K))1p1(1ϵ)n+pp,\frac{1}{(1+\epsilon)^{\frac{n+p}{p}}}\leq\left(\frac{{\mathcal{C}}^{p}\rho_{K}^{n+p}}{(n+p)\text{vol}_{n}(K)}\right)^{\frac{-1}{p}}\left(\frac{{\mathcal{C}}^{p}1}{(n+p)\text{vol}_{n}(K)}\right)^{\frac{1}{p}}\leq\frac{1}{(1-\epsilon)^{\frac{n+p}{p}}},

and

(1ϵ)n1(ρKn1n1)(n11)(1+ϵ)n1.(1-\epsilon)^{n-1}\leq\left(\frac{{\mathcal{R}}\rho_{K}^{n-1}}{n-1}\right)\left(\frac{n-1}{{\mathcal{R}}1}\right)\leq(1+\epsilon)^{n-1}.

If KK satisfies cΓK=Kc\Gamma^{*}K={\mathcal{I}}K, taking the ratio of the two terms above and multiplying by (9) yields

(1ϵ)np(1+ϵ)n+ppCBCK(1+ϵ)np(1ϵ)n+pp.\frac{(1-\epsilon)^{\frac{n}{p}}}{(1+\epsilon)^{\frac{n+p}{p}}}\leq\frac{C_{B}^{\prime}}{C_{K}^{\prime}}\leq\frac{(1+\epsilon)^{\frac{n}{p}}}{(1-\epsilon)^{\frac{n+p}{p}}}.

Thus, the ratio of the constants is of the order 1+O(ϵ)1+O(\epsilon). Applying an appropriate linear transformation to put KK in the isotropic position, we still have CK=CB+O(ϵ)C_{K}^{\prime}=C_{B}^{\prime}+O(\epsilon).

6.3. Linearization estimate

From the calculation of the constant for the ball, we immediately have

(23) 1ρKn1=CKCBκnvoln(K)p(𝒞1pρKn+p)1p,{\mathcal{R}}_{1}\rho_{K}^{n-1}=\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}({\mathcal{C}}_{1}^{p}\rho_{K}^{n+p})^{\frac{-1}{p}},

where 11=𝒞1p1=1{\mathcal{R}}_{1}1={\mathcal{C}}_{1}^{p}1=1 are the normalized Radon and pp-Cosine transforms. As we assume 1ϵ<ρK<1+ϵ1-\epsilon<\rho_{K}<1+\epsilon, the ratio |ρK1r0|<1\left\lvert\frac{\rho_{K}-1}{r_{0}}\right\rvert<1 for sufficiently small ϵ\epsilon. Consequently, we have convergence of the binomial series expansion of (r0+(ρKr0))1p(r_{0}+(\rho_{K}-r_{0}))^{-\frac{1}{p}}. The binomial expansion of the right hand side yields,

CKCBκnvoln(K)pm=0(m+1p1m)(1)mr0(n+p)(1p+m)(𝒞1pρKn+pr0n+p)m.\displaystyle\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}\sum_{m=0}^{\infty}\binom{m+\frac{1}{p}-1}{m}(-1)^{m}r_{0}^{-(n+p)(\frac{1}{p}+m)}\left({\mathcal{C}}_{1}^{p}\rho_{K}^{n+p}-r_{0}^{n+p}\right)^{m}.

Taking the inverse (normalized) Radon transform of both sides of (23), and separating the constant (24), linear (25), and higher order terms (26)-(28) yields

(24) ρKr0+r0n1CKCBκnvoln(K)pr01nnpn1\displaystyle\rho_{K}-r_{0}+\frac{r_{0}}{n-1}-\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}\frac{r_{0}^{1-n-\frac{n}{p}}}{n-1}
(25) =(CKCBκnvoln(K)pr0n(1+1p))(n+pp(n1)11𝒞1p)(ρKr0)\displaystyle=-\left(\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}r_{0}^{-n(1+\frac{1}{p})}\right)\left(\frac{n+p}{p(n-1)}{\mathcal{R}}_{1}^{-1}\circ{\mathcal{C}}_{1}^{p}\right)\left(\rho_{K}-r_{0}\right)
(26) +CKCBκnvoln(K)pm=2(m+1p1m)(1)mr01mpn(m+1p+1)n111(𝒞1pρKn+pr0n+p)m\displaystyle+\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}\sum_{m=2}^{\infty}\binom{m+\frac{1}{p}-1}{m}\frac{(-1)^{m}r_{0}^{1-mp-n(m+\frac{1}{p}+1)}}{n-1}{\mathcal{R}}_{1}^{-1}\left({\mathcal{C}}_{1}^{p}\rho_{K}^{n+p}-r_{0}^{n+p}\right)^{m}
(27) +1n1m=2n1(n1m)(ρKr0)mr0n1m\displaystyle+\frac{1}{n-1}\sum_{m=2}^{n-1}\binom{n-1}{m}\left(\rho_{K}-r_{0}\right)^{m}r_{0}^{n-1-m}
(28) +CKCBκnvoln(K)p1p(n1)m=2n+p(n+pm)r0(1pm)(n+p)1(1𝒞1p)(ρkr0)m.\displaystyle+\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}\frac{1}{p(n-1)}\sum_{m=2}^{n+p}\binom{n+p}{m}r_{0}^{(\frac{1}{p}-m)(n+p)-1}({\mathcal{R}}^{-1}\circ{\mathcal{C}}_{1}^{p})(\rho_{k}-r_{0})^{m}.

Assuming that ρKr0<ϵ{\left\|\rho_{K}-r_{0}\right\|}_{\infty}<\epsilon, we observe a few facts for the above equation. The linear term (25) consists of a constant of the order 1+O(ϵ)1+O(\epsilon) multiplying an operator that will be shown to be a contraction. The higher order terms (26), (27) and (28) will be shown to be of the order of O(ϵ2)O(\epsilon^{2}).

6.4. Proof of Theorem 3

We will assume for now that the terms (26)-(28) are indeed of the order O(ϵ2)O(\epsilon^{2}), and focus on terms (24) and (25). Taking L2L^{2} norms, we obtain,

ρKr02ρKr0CKCBκnvoln(K)pr0npn1r0n1n12\displaystyle{\left\|\rho_{K}-r_{0}\right\|}_{2}\leq{\left\|\rho_{K}-r_{0}-\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}\frac{r_{0}^{\frac{n}{p}}}{n-1}-\frac{r_{0}^{n-1}}{n-1}\right\|}_{2}
(CKCBκnvoln(K)pr01pp+n+pn+pp(n1)11𝒞1pL2L2+O(ϵ))ρKr02,\displaystyle\leq\left(\frac{C_{K}^{\prime}}{C_{B}^{\prime}}\sqrt[p]{\frac{\kappa_{n}}{\text{vol}_{n}(K)}}r_{0}^{\frac{1-p}{p}+n+p}{\left\|\frac{n+p}{p(n-1)}{\mathcal{R}}_{1}^{-1}\circ{\mathcal{C}}_{1}^{p}\right\|}_{L^{2}\rightarrow L^{2}}+O(\epsilon)\right){\left\|\rho_{K}-r_{0}\right\|}_{2},

where L2L2{\left\|\cdot\right\|}_{L^{2}\rightarrow L^{2}} is the operator norm. The first inequality follows from the orthogonality of the spherical harmonic spaces and the definition of r0r_{0}. If we are able to show that the term in parenthesis on the right hand side is strictly less than one for any given p1p\geq 1 and ϵ\epsilon small enough, then there will exist some ϵ0\epsilon_{0} such that ϵ<ϵ0\epsilon<\epsilon_{0} implies ρKr02=0{\left\|\rho_{K}-r_{0}\right\|}_{2}=0 or, equivalently, ρK=r0\rho_{K}=r_{0}.

We will check the operator norm n+pp(n1)11𝒞1pL2L2{\left\|\frac{n+p}{p(n-1)}{\mathcal{R}}_{1}^{-1}\circ{\mathcal{C}}_{1}^{p}\right\|}_{L^{2}\rightarrow L^{2}}, as the terms multiplying it have already been shown to be 1+O(ϵ)1+O(\epsilon) in subsection 6.2. The eigenspaces of the spherical Radon transform are the spherical harmonic spaces and the eigenvalues of the normalized spherical Radon transform can be seen below

𝒫k=σn1(1)k2μk𝒫k={σn1𝒫k:k=0σn1(1)k2(k1)!!(n3)!!(n+k3)!!𝒫k:2k and k00:2k,{\mathcal{R}}{\mathcal{P}_{k}}=\sigma_{n-1}(-1)^{\frac{k}{2}}\mu_{k}{\mathcal{P}_{k}}=\begin{cases}\sigma_{n-1}{\mathcal{P}_{k}}&:k=0\\ \sigma_{n-1}(-1)^{\frac{k}{2}}\frac{(k-1)!!(n-3)!!}{(n+k-3)!!}{\mathcal{P}_{k}}&:2\mid k\text{ and }k\neq 0\\ 0&:2\nmid k,\end{cases}

see for example [7, Eq 3.4.16 pg 103]. Combining this with the eigenvalues for the pp-cosine transform, we have that

n+pp(n1)11𝒞1p𝒫2kL2L2=n+pp(n1)=0k1|2p2+12+n12+n+p|.{\left\|\frac{n+p}{p(n-1)}{\mathcal{R}}_{1}^{-1}\circ{\mathcal{C}}_{1}^{p}\circ{\mathcal{P}_{2k}}\right\|}_{L^{2}\rightarrow L^{2}}=\frac{n+p}{p(n-1)}\prod_{\ell=0}^{k-1}\left\lvert\frac{2\ell-p}{2\ell+1}\,\,\frac{2\ell+n-1}{2\ell+n+p}\right\rvert.

For k=1k=1, the right hand side equals 1. As kk increases, each additional term in the product is strictly less than 1. Therefore this operator norm is less than 11.

6.5. Higher order terms

The fact that (27) and (28) are of the order O(ϵ2)O(\epsilon^{2}) is immediate since m2m\geq 2. This leaves just (26) to be estimated.

We need to introduce a family of operators to help estimate the composition of the inverse Radon transform and the powers of the pp-cosine transform. Let 𝟙I:{0,1}\mathbbm{1}_{I}:{\mathbb{R}}\rightarrow\{0,1\} be the indicator function of the interval [0,1][0,1], and define Mk:L2L2M_{k}:L^{2}\rightarrow L^{2} by

Mk=j=0𝟙I(j2k)𝒫j,M_{k}=\sum_{j=0}^{\infty}\mathbbm{1}_{I}\left(\frac{j}{2^{k}}\right){\mathcal{P}_{j}},

and M~0:=M0\widetilde{M}_{0}:=M_{0} and M~k:=Mk+1Mk\widetilde{M}_{k}:=M_{k+1}-M_{k}. Observe that

MkfL2fL2,𝒫kfL(dim(kn)|𝒮n1|)1/2fL2,{\left\|M_{k}f\right\|}_{L^{2}}\leq{\left\|f\right\|}_{L^{2}},\hskip 40.0pt{\left\|{\mathcal{P}}_{k}f\right\|}_{L^{\infty}}\leq\left(\frac{\dim({\mathcal{H}}_{k}^{n})}{\left\lvert{\mathcal{S}^{n-1}}\right\rvert}\right)^{1/2}{\left\|f\right\|}_{L^{2}},

where the last estimates can be found in [2, Eq. 2.48, pg. 27]. For fC(𝒮n1)f\in C({\mathcal{S}^{n-1}}), MkffM_{k}f\rightarrow f in L2L^{2} as well. Thus, for any mm\in{\mathbb{N}}

(f)m\displaystyle(f)^{m} =k=0(Mk+1f)m(Mkf)m\displaystyle=\sum_{k=0}^{\infty}(M_{k+1}f)^{m}-(M_{k}f)^{m}
=k=0(Mk+1fMkf)(j=0m1(Mk+1f)j(Mkf)m1j)\displaystyle=\sum_{k=0}^{\infty}\left(M_{k+1}f-M_{k}f\right)\left(\sum_{j=0}^{m-1}(M_{k+1}f)^{j}(M_{k}f)^{m-1-j}\right)
=k=0(M~kf)(j=0m1(Mk+1f)j(Mkf)m1j),\displaystyle=\sum_{k=0}^{\infty}\left(\widetilde{M}_{k}f\right)\left(\sum_{j=0}^{m-1}(M_{k+1}f)^{j}(M_{k}f)^{m-1-j}\right),

where the last equality is uses the linearity of MkM_{k} and the definition of M~k\widetilde{M}_{k}. Breaking each of the summands into their dyadic harmonic projections yields

(f)m\displaystyle(f)^{m} =b=0k=0M~b[(M~kf)(j=0m1(Mk+1f)j(Mkf)m1j)].\displaystyle=\sum_{b=0}^{\infty}\sum_{k=0}^{\infty}\widetilde{M}_{b}\left[\left(\widetilde{M}_{k}f\right)\left(\sum_{j=0}^{m-1}(M_{k+1}f)^{j}(M_{k}f)^{m-1-j}\right)\right].

While the product of two harmonic polynomials need not be harmonic, any polynomial when restricted to the sphere may be written as the sum of homogeneous, harmonic polynomials of lesser or equal degree (see [7, Lemma 3.2.5, pg. 69] or [2, Theorem 2.18, pg 31]). Thus, if the degree of the polynomial in square brackets is smaller than 2b2^{b}, then the operator M~b\widetilde{M}_{b} will give 0 identically. For non-zero terms, the degree of the polynomial in brackets is

2b2k+1+(m1)2k+1=m2k+1bk+log2(2m).2^{b}\leq 2^{k+1}+(m-1)2^{k+1}=m2^{k+1}\implies b\leq k+\log_{2}(2m).

Now, taking the inverse Radon transform and L2L^{2} norms on both sides, we see that the right hand side is bounded by

b=011M~bk=blog2(2m)[(M~kf)(j=0m1(Mk+1f)j(Mkf)m1j)]L2\displaystyle\sum_{b=0}^{\infty}{\left\|{\mathcal{R}}^{-1}_{1}\widetilde{M}_{b}\sum_{k=b-\lfloor\log_{2}(2m)\rfloor}^{\infty}\left[\left(\widetilde{M}_{k}f\right)\left(\sum_{j=0}^{m-1}(M_{k+1}f)^{j}(M_{k}f)^{m-1-j}\right)\right]\right\|}_{L^{2}}
b=0μ2b+11k=blog2(2m)[(M~kf)(j=0m1(Mk+1f)j(Mkf)m1j)]L2\displaystyle\leq\sum_{b=0}^{\infty}\mu_{2^{b+1}}^{-1}{\left\|\sum_{k=b-\lfloor\log_{2}(2m)\rfloor}^{\infty}\left[\left(\widetilde{M}_{k}f\right)\left(\sum_{j=0}^{m-1}(M_{k+1}f)^{j}(M_{k}f)^{m-1-j}\right)\right]\right\|}_{L^{2}}
b=0μ2b+11k=blog2(2m)M~kfL2(j=0m1Mk+1fLjMkfLm1j).\displaystyle\leq\sum_{b=0}^{\infty}\mu_{2^{b+1}}^{-1}\sum_{k=b-\lfloor\log_{2}(2m)\rfloor}^{\infty}{\left\|\widetilde{M}_{k}f\right\|}_{L^{2}}\left(\sum_{j=0}^{m-1}{\left\|M_{k+1}f\right\|}_{L^{\infty}}^{j}{\left\|M_{k}f\right\|}_{L^{\infty}}^{m-1-j}\right).

Setting f=𝒞1pgf={\mathcal{C}}_{1}^{p}g, where g=ρKn+pr0n+pg=\rho_{K}^{n+p}-r_{0}^{n+p}, we obtain the following bound on the LL^{\infty} norm,

Mk𝒞1pgL\displaystyle{\left\|M_{k}{\mathcal{C}}_{1}^{p}g\right\|}_{L^{\infty}} a=02k𝒫a𝒞1pgL1|𝒮n1|a=02k|mp,amp,0|dim(an)𝒫agL2\displaystyle\leq\sum_{a=0}^{2^{k}}{\left\|{\mathcal{P}_{a}}{\mathcal{C}}_{1}^{p}g\right\|}_{L^{\infty}}\leq\frac{1}{\sqrt{\left\lvert{\mathcal{S}^{n-1}}\right\rvert}}\sum_{a=0}^{2^{k}}\left\lvert\frac{m_{p,a}}{m_{p,0}}\right\rvert\sqrt{\dim({{\mathcal{H}}_{a}^{n}})}{\left\|{\mathcal{P}_{a}}g\right\|}_{L^{2}}
supa{|mp,a|dim(an)}|mp,0||𝒮n1|gL2.\displaystyle\leq\frac{\sup_{a}\{\left\lvert m_{p,a}\right\rvert\sqrt{\dim({{\mathcal{H}}_{a}^{n}})}\}}{\left\lvert m_{p,0}\right\rvert\sqrt{\left\lvert{\mathcal{S}^{n-1}}\right\rvert}}{\left\|g\right\|}_{L^{2}}.

Observe that |mp,a|=O(a(n+2p)/2)\left\lvert m_{p,a}\right\rvert=O(a^{-(n+2p)/2}) and dim(an)=O(an)\dim({{\mathcal{H}}_{a}^{n}})=O(a^{n}), so this supremum is indeed finite though dependent on the dimension. Simplifying,

(29) 1(𝒞1pg)mL2\displaystyle{\left\|{\mathcal{R}}^{-1}({\mathcal{C}}_{1}^{p}g)^{m}\right\|}_{L^{2}}\leq
mCm1gL2m(b=0log2(m)μ2b+11+b=log2(2m)μ2p+11|mp,2blog2(2m)mp,0|).\leq mC^{m-1}{\left\|g\right\|}_{L^{2}}^{m}\left(\sum_{b=0}^{\lfloor\log_{2}(m)\rfloor}\mu_{2^{b+1}}^{-1}+\sum_{b=\lfloor\log_{2}(2m)\rfloor}^{\infty}\mu_{2^{p+1}}^{-1}\left\lvert\frac{m_{p,2^{b-\lfloor\log_{2}(2m)\rfloor}}}{m_{p,0}}\right\rvert\right).

where C=maxkMk𝒞1pLL2C=\max_{k}{\left\|M_{k}{\mathcal{C}}_{1}^{p}\right\|}_{L^{\infty}\rightarrow L^{2}}. Now we are in a position to estimate the higher order term (26). Including the binomial coefficient and power of r0r_{0} changing with mm, we see

(30) (m+1p1m)r0m(pn)n11(𝒞1pg)mL2(m+1)Γ(1p)(1ϵ)m|np|mCm1gL2m\displaystyle\binom{m+\frac{1}{p}-1}{m}\frac{r_{0}^{m(p-n)}}{n-1}{\left\|{\mathcal{R}}^{-1}({\mathcal{C}}_{1}^{p}g)^{m}\right\|}_{L^{2}}\leq\frac{(m+1)\Gamma(\frac{1}{p})}{(1-\epsilon)^{m\left\lvert n-p\right\rvert}}mC^{m-1}{\left\|g\right\|}_{L^{2}}^{m}\cdot
(b=0log2(m)μ2b+11+b=log2(2m)μ2p+11|mp,2blog2(2m)mp,0|).\displaystyle\cdot\left(\sum_{b=0}^{\lfloor\log_{2}(m)\rfloor}\mu_{2^{b+1}}^{-1}+\sum_{b=\lfloor\log_{2}(2m)\rfloor}^{\infty}\mu_{2^{p+1}}^{-1}\left\lvert\frac{m_{p,2^{b-\lfloor\log_{2}(2m)\rfloor}}}{m_{p,0}}\right\rvert\right).

The right hand side is of the order O(ϵm)O(\epsilon^{m}) times the sum in parentheses. Note that the sums are constant for m[2a,2a+1)m\in[2^{a},2^{a+1}) and at these values of mm we get a right hand side of

ϵ2a(b=0aμ2b+11+b=0μ2b+a+21|mp,2bmp,0|).\epsilon^{2^{a}}\left(\sum_{b=0}^{a}\mu_{2^{b+1}}^{-1}+\sum_{b=0}^{\infty}\mu_{2^{b+a+2}}^{-1}\left\lvert\frac{m_{p,2^{b}}}{m_{p,0}}\right\rvert\right).

Summing over m[2a,2a+1)m\in[2^{a},2^{a+1})\cap{\mathbb{N}}, we then obtain a right hand side of

[1ϵ2a+11ϵ]ϵ2a(b=0aμ2b+11+b=0μ2b+a+21|mp,2bmp,0|).\left[\frac{1-\epsilon^{2^{a}+1}}{1-\epsilon}\right]\epsilon^{2^{a}}\left(\sum_{b=0}^{a}\mu_{2^{b+1}}^{-1}+\sum_{b=0}^{\infty}\mu_{2^{b+a+2}}^{-1}\left\lvert\frac{m_{p,2^{b}}}{m_{p,0}}\right\rvert\right).

The term in square brackets is 1+O(ϵ)1+O(\epsilon) and will be omitted from now on. Now, summing over aa\in{\mathbb{N}} we get an upper bound for (26),

(26)a=1ϵ2a(b=0aμ2b+11+b=0μ2b+a+21|mp,2bmp,0|).\eqref{calip_gammma_Linearization_HOTC}\leq\sum_{a=1}^{\infty}\epsilon^{2^{a}}\left(\sum_{b=0}^{a}\mu_{2^{b+1}}^{-1}+\sum_{b=0}^{\infty}\mu_{2^{b+a+2}}^{-1}\left\lvert\frac{m_{p,2^{b}}}{m_{p,0}}\right\rvert\right).

Since μk1=O(kn2)\mu_{k}^{-1}=O(k^{\frac{n}{2}}) and mp,k=O(kn+2p2)m_{p,k}=O(k^{-\frac{n+2p}{2}}), we get the estimate

(26)\displaystyle\eqref{calip_gammma_Linearization_HOTC} a=1ϵ2a(b=0a2(b+1)n2+b=02(b+a+2)n22bn+22).\displaystyle\leq\sum_{a=1}^{\infty}\epsilon^{2^{a}}\left(\sum_{b=0}^{a}2^{(b+1)\frac{n}{2}}+\sum_{b=0}^{\infty}2^{(b+a+2)\frac{n}{2}}2^{-b\frac{n+2}{2}}\right).

Evaluating the partial sums of the geometric series, we have

(26)a=1ϵ2a(2(a+2)n212n21+2(a+2)n2+1)ϵ21ϵCn\displaystyle\eqref{calip_gammma_Linearization_HOTC}\leq\sum_{a=1}^{\infty}\epsilon^{2^{a}}\left(\frac{2^{(a+2)\frac{n}{2}}-1}{2^{\frac{n}{2}}-1}+2^{(a+2)\frac{n}{2}+1}\right)\leq\frac{\epsilon^{2}}{1-\epsilon}C_{n}

Finally, picking epsilon small enough then yields the desired estimate.


Acknowledgements: I would like to recognize my advisor María de los Ángeles Alfonseca for her guidance in this work. I would also like to thank Fedor Nazarov for his advice to use the p{\mathcal{M}}_{p} operators for the final estimate. I also thank Dmitry Ryabogin and Vladyslav Yaskin for several fruitful discussions.

References

  • [1] M. Angeles Alfonseca, Fedor Nazarov, Dmitry Ryabogin, and Vladyslav Yaskin. A solution to the fifth and the eighth Busemann-Petty problems in a small neighborhood of the Euclidean ball. Adv. Math., 390:Paper No. 107920, 28, 2021.
  • [2] Kendall Atkinson and Weimin Han. Spherical harmonics and approximations on the unit sphere: an introduction, volume 2044 of Lecture Notes in Mathematics. Springer, Heidelberg, 2012.
  • [3] Silouanos Brazitikos, Apostolos Giannopoulos, Petros Valettas, and Beatrice-Helen Vritsiou. Geometry of isotropic convex bodies, volume 196 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2014.
  • [4] H. Busemann and C. M. Petty. Problems on convex bodies. Math. Scand., 4:88–94, 1956.
  • [5] Alexander Fish, Fedor Nazarov, Dmitry Ryabogin, and Artem Zvavitch. The unit ball is an attractor of the intersection body operator. Adv. Math., 226(3):2629–2642, 2011.
  • [6] Richard J. Gardner. Geometric tomography, volume 58 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, New York, second edition, 2006.
  • [7] H. Groemer. Geometric applications of Fourier series and spherical harmonics, volume 61 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1996.
  • [8] Erwin Lutwak, Deane Yang, and Gaoyong Zhang. LpL_{p} affine isoperimetric inequalities. J. Differential Geom., 56(1):111–132, 2000.
  • [9] C. M. Petty. On the geometry of the Minkowski plane. Riv. Mat. Univ. Parma, 6:269–292, 1955.
  • [10] Boris Rubin. Intersection bodies and generalized cosine transforms. Adv. Math., 218(3):696–727, 2008.
  • [11] Dmitry Ryabogin. On bodies floating in equilibrium in every orientation. Geom. Dedicata, 217(4):Paper No. 70, 17, 2023.
  • [12] C. Saroglou and A. Zvavitch. Iterations of the projection body operator and a remark on Petty’s conjectured projection inequality. J. Funct. Anal., 272(2):613–630, 2017.
  • [13] Rolf Schneider. Convex bodies: the Brunn-Minkowski theory, volume 151 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, expanded edition, 2014.