1 Introduction

The concept of John ellipsoid, introduced by Fritz John [20], is extremely useful in convex geometry and Banach space geometry. For each convex body (compact convex set with nonempty interior) K in the n-dimensional Euclidean \({\mathbb {R}}^{n}\), its John ellipsoid JK is defined as the unique ellipsoid of maximal volume contained in K.

Two fundamental results concerning the John ellipsoid are John’s inclusion and Ball’s volume-ratio inequality. Let K be an origin-symmetric convex body K in \({\mathbb {R}}^{n}\). John’s inclusion shows that

$$\begin{aligned} K\subseteq \sqrt{n}JK. \end{aligned}$$
(1.1)

As an application of John’s inclusion, the best upper bound of the Banach–Mazur distance is \(\sqrt{n}\), for an n-dimensional normed space to n-dimensional Euclidean space. Ball’s volume-ratio inequality states that

$$\begin{aligned} \frac{|K|}{|JK|}\le \frac{2^{n}}{\omega _{n}}, \end{aligned}$$
(1.2)

with equality if and only if K is a parallelotope. Here \(|\cdot |\) denotes n-dimensional volume and \(\omega _{n}=|B|=\pi ^{n/2}/\Gamma \left( 1+\frac{n}{2}\right) \) denotes the volume of the unit ball B in \({\mathbb {R}}^{n}\). The fact that there is equality in (1.2) only for parallelotopes was established by Barthe [3]. For more information about the John ellipsoid, one can refer to [1, 2, 12, 14, 15, 21, 22, 44] and the references within.

In 2005, Lutwak, Yang and Zhang [30] extend the John ellipsoid to \(L_{p}\) John ellipsoids, which is an important concept in the \(L_{p}\) Brunn–Minkowski theory initiated by Lutwak [27, 28]. During the last two decades, the \(L_{p}\) Brunn–Minkowski theory has achieved great developments and expanded rapidly, see, e.g., [4,5,6, 8, 9, 17,18,19, 24,25,26, 29, 31,32,34, 37, 38, 47,48,51]. Moreover, the Orlicz Brunn–Minkowski theory, as an extension of the \(L_{p}\) Brunn–Minkowski theory, emerged in [16, 35, 36]. In these papers, the fundamental notions of the \(L_{p}\) projection body and the \(L_{p}\) centroid body were extended to an Orlicz setting, see also [7, 53, 55]. For more information, please refer to the literature [11, 23, 39,40,41, 54, 56,57,58,59,60]. In particular, the classical John ellipsoid is extended to the \(L_{p}\) setting by Lutwak, Yang and Zhang [30] and to the Orlicz setting by Zou and Xiong [58].

Suppose \(p\in (0,\infty ]\) and K is a convex body in \({\mathbb {R}}^{n}\) with the origin in its interior. Among all origin-symmetric ellipsoids E, the unique ellipsoid that solves the constrained maximization problem

$$\begin{aligned} \max \limits _{E}\left( \frac{|E|}{\omega _{n}}\right) ^{\frac{1}{n}},\ \ \ \text{ subject } \text{ to }\ \ \ {\overline{V}}_{p}(K,E)\le 1, \end{aligned}$$
(1.3)

is called the \(L_{p}\) John ellipsoid [30] of K and denoted by \(E_{p}K\). Clearly, \(E_{p}B=B\). Here

$$\begin{aligned} {\overline{V}}_{p}(K,E)=\left( \frac{1}{n|K|}\int _{S^{n-1}} \left( \frac{h_{E}(u)}{h_{K}(u)}\right) ^{p}h_{K}(u)\mathrm{d}S(K,u)\right) ^{\frac{1}{p}},\ \ 0<p<\infty , \end{aligned}$$

is the normalized \(L_{p}\) mixed volume of K and E; \(S^{n-1}\) is the unit sphere in \({\mathbb {R}}^{n}\); \(h_{K}\) and \(h_{E}\) are the support functions (see Sect. 2) of K and E, respectively. In the case \(p=\infty \), we define

$$\begin{aligned} {\overline{V}}_{\infty }(K,E)=\sup \left\{ \frac{h_{E}(u)}{h_{K}(u)}: u\in {\mathrm{supp}}S(K,\cdot )\right\} . \end{aligned}$$

Therefore, when the John point of K, i.e., the center of JK, is at the origin, \(E_{\infty }K\) is precisely the classical John ellipsoid JK. In the case \(p=2\), the \(L_2\) John ellipsoid \(E_{2}K\) is the new ellipsoid \(\Gamma _{-2}K\) found by Lutwak, Yang and Zhang in [32], which is now called the LYZ ellipsoid and is in some sense dual to the Legendre ellipsoid of inertia in classical mechanics [42]. In the case \(p=1\), \(E_{1}K\) is the so-called Petty ellipsoid, see [13, 43]. The volume-normalized Petty ellipsoid is obtained by minimizing the surface area of K under \({\mathrm{SL}}(n)\)-transformations.

In general, the \(L_{p}\) John ellipsoid \(E_{p}K\) is not contained in K (except when \(p=\infty \)). However, when \(1\le p\le \infty \), it has \(|E_{p}K|\le |K|\). In reverse, for \(0<p\le \infty \), the \(L_{p}\) version of Ball’s volume-ratio inequality [30] states that

$$\begin{aligned} \frac{|K|}{|E_{p}K|}\le \frac{2^{n}}{\omega _{n}} \end{aligned}$$

with equality if and only if K is a parallelotope.

By \(L_p\) dual curvature measures, Lutwak, Yang and Zhang [31] introduced the notion of \(L_{p}\) dual mixed volumes which unifies \(L_{p}\) mixed volumes of convex bodies in the \(L_{p}\) Brunn–Minkowski theory and dual mixed volumes of star bodies in the dual Brunn–Minkowski theory. Therefore, \(L_{p}\) dual mixed volumes become to be a core concept in convex geometry with unifying some contents of the \(L_{p}\) Brunn–Minkowski theory and the dual Brunn–Minkowski theory.

Let \({\mathcal {K}}^{n}_{o}\) denote the class of convex bodies in \({\mathbb {R}}^{n}\) that contain the origin in their interiors. And let \({\mathcal {S}}^{n}_{o}\) denote the set of star bodies (compact star-shaped set about the origin) in \({\mathbb {R}}^{n}\).

Suppose K is a convex body in \({\mathbb {R}}^{n}\). For each \(v\in {\mathbb {R}}^{n}\backslash \{o\}\), the hyperplane

$$\begin{aligned} H_{K}(v)=\{x\in {\mathbb {R}}^{n}: x\cdot v=h_{K}(v)\} \end{aligned}$$

is called the supporting hyperplane to K with outer normal v.

The spherical image (Gauss image) of \(\sigma \subset \partial K\) is defined by

$$\begin{aligned} \varvec{\nu }_{K}(\sigma )=\{v\in S^{n-1}:x\in H_{K}(v)\ \text{ for } \text{ some }\ x\in \sigma \}\subset S^{n-1}. \end{aligned}$$

Let \(\sigma _{K}\subset \partial K\) be the set consisting of boundary points \(x\in \partial K\), for which the set \(\varvec{\nu }_{K}(\{x\})\) contains more than a single element. It is well known that the spherical Lebesgue measure of \(\sigma _{K} \) is \({\mathcal {H}}^{n-1}(\sigma _{K})=0\) (see, e.g., [46, p. 84]). On precisely the functions

$$\begin{aligned} \nu _{K}:\partial K\backslash \sigma _{K}\rightarrow S^{n-1}, \end{aligned}$$

is called the spherical image map (Gauss map) of K and is continuous (see, e.g., [46, Lemma 2.2.12]). The set \(\partial K\backslash \sigma _{K}\) is usually abbreviated by \(\partial 'K\). Since \({\mathcal {H}}^{n-1}(\sigma _{K})=0\), the integrals over subsets of \(\partial 'K\) and \(\partial K\) are equal with respect to \({\mathcal {H}}^{n-1}\).

For \(\omega \subset S^{n-1}\), the radial Gauss image of \(\omega \) is denoted by

$$\begin{aligned} \varvec{\alpha }_{K}(\omega )=\{v\in S^{n-1}: \rho _{K}(u)u\in H_{K}(v)\ \ \text{ for } \text{ some }\ u\in \omega \}. \end{aligned}$$

For a subset \(\eta \subset S^{n-1}\), the reverse radial Gauss image of \(\eta \) is denoted by

$$\begin{aligned} \varvec{\alpha }_{K}^{*}(\eta )=\{u\in S^{n-1}: \rho _{K}(u)u\in H_{K}(v)\ \ \text{ for } \text{ some }\ v\in \eta \}. \end{aligned}$$

For \(K\in {\mathcal {K}}^{n}_{o}\), the radial map of K, \(r_{K}: S^{n-1}\rightarrow \partial K,\) is defined by

$$\begin{aligned} r_{K}(u)=\rho _{K}(u)u\in \partial K, \end{aligned}$$

for \(u\in S^{n-1}\). Here, \(\rho _{K}(u)=\max \{\lambda >0: \lambda u\in K\}\) is the radial function of K for \(u\in S^{n-1}\). Note that \(r_{K}^{-1}: \partial K\rightarrow S^{n-1}\) is given by \(r_{K}^{-1}(x)=x/|x|\) for \(x\in \partial K\). Let \(\omega _{K}=\overline{\sigma _{K}}=r_{K}^{-1}(\sigma _{K})\). Observe that \(\omega _{K}\) has spherical Lebesgue measure 0, and the integrals over subsets of \(S^{n-1}\backslash \omega _{K}\) and \(S^{n-1}\) are equal with respect to the spherical Lebesgue measure.

The radial Gauss map of \(K\in {\mathcal {K}}^{n}_{o}\), \(\alpha _{K}: S^{n-1}\backslash \omega _{K}\rightarrow S^{n-1}\), is given by

$$\begin{aligned} \alpha _{K}=\nu _{K}\circ r_{K}. \end{aligned}$$

Obviously, for any \(\lambda >0\) and any \(u\in S^{n-1}\),

$$\begin{aligned} \alpha _{\lambda K}(u)=\alpha _{K}(u). \end{aligned}$$
(1.4)

For \(p,q\in {\mathbb {R}}\), \(K\in {\mathcal {K}}^{n}_{o}\), and \(Q\in {\mathcal {S}}^{n}_{o}\), the \(L_{p}\) dual curvature measures \({\widetilde{C}}_{p,q}(K,Q)\) are Borel measures on \(S^{n-1}\) given by

$$\begin{aligned} \int _{S^{n-1}}g(v){\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)=\frac{1}{n}\int _{S^{n-1}} g(\alpha _{K}(u))h_{K}(\alpha _{K}(u))^{-p}\rho _{K}(u)^{q}\rho _{Q}(u)^{n-q}{\mathrm{d}}u,\nonumber \\ \end{aligned}$$
(1.5)

for each continuous function \(g: S^{n-1}\rightarrow {\mathbb {R}}\). For each Borel set \(\eta \subseteq S^{n-1}\), we have

$$\begin{aligned} {\widetilde{C}}_{p,q}(K,Q,\eta )=\frac{1}{n}\int _{\varvec{\alpha }_{K}^{*} (\eta )}h_{K}(\alpha _{K}(u))^{-p}\rho _{K}^{q}(u)\rho _{Q}^{n-q}(u){\mathrm{d}}u. \end{aligned}$$
(1.6)

It has shown that [31, Proposition 5.4] that the \(L_{p}\) surface area measure, the dual curvature measure and the integral measure are all special cases of the \(L_{p}\) dual curvature measure. In particular, for \(p,q\in {\mathbb {R}}\), and \(K\in {\mathcal {K}}^{n}_{o}\),

$$\begin{aligned} {\widetilde{C}}_{p,q}(K,K,\cdot )= & {} \frac{1}{n}S_{p}(K,\cdot ),\end{aligned}$$
(1.7)
$$\begin{aligned} {\widetilde{C}}_{p,n}(K,B,\cdot )= & {} \frac{1}{n}S_{p}(K,\cdot ), \end{aligned}$$
(1.8)

where \(S_{p}(K,\cdot )\) is the \(L_{p}\)-surface area measure of K.

Using \(L_{p}\) dual curvature measures, Lutwak, Yang and Zhang [31] introduced the concept of (pq)-mixed volume volumes. For \(p,q\in {\mathbb {R}}\), and convex bodies \(K,L\in {\mathcal {K}}^{n}_{o}\), and a star body \(Q\in {\mathcal {S}}^{n}_{o}\), the (pq)-mixed volume \({\widetilde{V}}_{p,q}(K,L,Q)\) is defined by

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,L,Q)= & {} \int _{S^{n-1}}h_{L}^{p}(v){\mathrm{d}} {\widetilde{C}}_{p,q}(K,Q,v)\nonumber \\= & {} \frac{1}{n}\int _{S^{n-1}}h_{L}(\alpha _{K}(u))^{p}h_{K} (\alpha _{K}(u))^{-p}\rho _{K}(u)^{q}\rho _{Q}(u)^{n-q}{\mathrm{d}}u \end{aligned}$$
(1.9)
$$\begin{aligned}= & {} \frac{1}{n}\int _{S^{n-1}}\left( \frac{h_{L}(\alpha _{K}(u))}{h_{K} (\alpha _{K}(u))}\right) ^{p}\left( \frac{\rho _{K}(u)}{\rho _{Q}(u)} \right) ^{q}\rho _{Q}(u)^{n}{\mathrm{d}}u. \end{aligned}$$
(1.10)

The concept of the (pq)-mixed volume unifies the \(L_{p}\) mixed volume and the dual mixed volume in the sense that

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,L,K)=V_{p}(K,L),\ \ {\widetilde{V}}_{p,q}(K,K,Q)={\widetilde{V}}_{q}(K,Q). \end{aligned}$$
(1.11)

In this paper we will consider the problem of minimizing total \(L_{p}\) dual curvature measures under \(\mathrm{SL}(n)\)-transformations. Let K be a smooth convex body in \({\mathbb {R}}^{n}\) with the origin in its interior, and let Q be a smooth star body in \({\mathbb {R}}^{n}\). For real number pq, find

$$\begin{aligned} \min \limits _{\phi \in {\mathrm{SL}}(n)}\int _{S^{n-1}}{\mathrm{d}}{\widetilde{C}}_{p,q}(\phi K,\phi Q,u). \end{aligned}$$

From (1.9) and [31, Proposition 7.3] (see also Lemma 2.3 of our paper), it follows that the original problem of minimizing total \(L_{p}\) dual curvature under \({\mathrm{SL}}(n)\)-transformations can be rewritten as

$$\begin{aligned} \min \limits _{\phi \in {\mathrm{SL}}(n)}\int _{S^{n-1}}{\mathrm{d}}{\widetilde{C}}_{p,q} (\phi K,\phi Q,u)= & {} \min \limits _{\phi \in {\mathrm{SL}}(n)}{\widetilde{V}}_{p,q}(\phi K,B,\phi Q)\\= & {} \min \limits _{\phi \in {\mathrm{SL}}(n)}{\widetilde{V}}_{p,q}(K,\phi ^{-1}B,Q)\\= & {} \min \limits _{|E|=\omega _{n}}{\widetilde{V}}_{p,q}(K,E,Q), \end{aligned}$$

where the last minimum is taken over all origin-symmetric ellipsoids with volume \(\omega _n\). A \(\phi _{p,q}\in {\mathrm{SL}}(n)\) at which this minimum is attained defines an ellipsoid \({\bar{E}}_{p,q}(K,Q)\) which \(\phi _{p,q}\) maps into the unit ball B, i.e., \({\bar{E}}_{p,q}(K,Q)=\phi _{p,q}^{-1}B\). This ellipsoid is unique and will be called the volume-normalized (pq)-John ellipsoid of K and Q. For \(p=\infty \), define

$$\begin{aligned} {\bar{E}}_{\infty ,q}(K,Q)=\lim \limits _{p\rightarrow \infty }{\bar{E}}_{p,q}(K,Q). \end{aligned}$$

For \(r\in [0, +\infty )\), the normalized r-th dual area measure of \(K,Q\in {\mathcal {S}}^{n}_{o}\), \(\overline{{\widetilde{V}}}_{r}(K,Q; \cdot )\), is defined by

$$\begin{aligned} {\mathrm{d}}\overline{{\widetilde{V}}}_{r}(K,Q; u)=\frac{1}{n{\widetilde{V}}_{r}(K,Q)} \rho ^{r}_{K}(u)\rho _{Q}^{n-r}(u){\mathrm{d}}u,\ \ \ \text{ for }\ u\in S^{n-1}, \end{aligned}$$
(1.12)

where \({\widetilde{V}}_{r}(K,Q)\) is the r-th dual mixed volume of \(K,Q\in {\mathcal {S}}^{n}_{o}\). Clearly, \({\mathrm{d}}\overline{{\widetilde{V}}}_{r}(K,Q; \cdot )\) is a probability measure on \(S^{n-1}\). In the case \(Q=K\), \({\mathrm{d}}\overline{{\widetilde{V}}}_{r}(K,K;u)={\mathrm{d}}\overline{{\widetilde{V}}}_{K}(u) =\frac{1}{n|K|} \rho ^{n}_{K}{\mathrm{d}}u,\text{ for }\ u\in S^{n-1},\) is the normalized dual area measure of \(K\in {\mathcal {S}}^{n}_{o}\). And for the cases \(r=0,n\), we have \({\mathrm{d}}\overline{{\widetilde{V}}}_{0}(K,Q; \cdot )={\mathrm{d}}\overline{{\widetilde{V}}}_{Q}(\cdot )\) and \({\mathrm{d}}\overline{{\widetilde{V}}}_{n}(K,Q; \cdot )={\mathrm{d}}\overline{{\widetilde{V}}}_{K}(\cdot )\).

In order to rewrite the formulation of our problem for the case \(p=\infty \), we next introduce a normalized version of (pq)-dual mixed volumes. If \(K,L\in {\mathcal {K}}^{n}_{o}\), \(Q\in {\mathcal {S}}^{n}_{o}\) and \(q\ge p>0\) with \(r=q-p\ge 0\), then we define the normalized (pq)-dual mixed volume by

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(K,L,Q)= & {} \left( \frac{{\widetilde{V}}_{p,q} (K,L,Q)}{{\widetilde{V}}_{r}(K,Q)}\right) ^{\frac{1}{p}}\nonumber \\= & {} \left( \int _{S^{n-1}}\left( \frac{h_{L}(\alpha _{K}(u)) \rho _{K}(u)}{h_{K}(\alpha _{K}(u))\rho _{Q}(u)}\right) ^{p} {\mathrm{d}}\overline{{\widetilde{V}}}_{r}(K,Q; u)\right) ^{\frac{1}{p}}. \end{aligned}$$
(1.13)

In the case \(p=\infty \) (then \(q=\infty \)), define

$$\begin{aligned} \overline{{\widetilde{V}}}_{\infty ,\infty }(K,L,Q)=\max \left\{ \frac{h_{L} (\alpha _{K}(u))\rho _{K}(u)}{h_{K}(\alpha _{K}(u))\rho _{Q}(u)}: u\in {\mathrm{supp}}{\widetilde{V}}_{r}(K,Q; \cdot )\right\} . \end{aligned}$$
(1.14)

Unless \(\frac{h_{L}(\alpha _{K}(u))\rho _{K}(u)}{h_{K}(\alpha _{K}(u))\rho _{Q}(u)}\) is constant on \({\mathrm{supp}}{\widetilde{V}}_{r}(K,Q; \cdot )\), it follows from (1.13) and Jensen’s inequality that

$$\begin{aligned} \overline{{\widetilde{V}}}_{p_{1},q_{1}}(K,L,Q)<\overline{{\widetilde{V}}}_{p_{2},q_{2}}(K,L,Q), \end{aligned}$$
(1.15)

for \(0<p_{1}<p_{2}\le \infty , 0<q_{1}=p_{1}+r\le p_{2}+r=q_{2}\le \infty \), and

$$\begin{aligned} \lim \limits _{p\rightarrow \infty }\overline{{\widetilde{V}}}_{p,q}(K,L,Q)=\overline{{\widetilde{V}}}_{\infty ,\infty }(K,L,Q). \end{aligned}$$

We shall require the fact that, for \(p_{0}\in (0,\infty ], q_{0}=p_{0}+r\in (0,\infty ]\) and \(r\in [0,\infty )\),

$$\begin{aligned} \lim \limits _{p\rightarrow p_{0}}\overline{{\widetilde{V}}}_{p,q}(K,L,Q) =\overline{{\widetilde{V}}}_{p_{0},q_{0}}(K,L,Q). \end{aligned}$$
(1.16)

In fact, we have already proved a more general conclusion, see Theorem 3.1 in subsequent. By (1.14), we have

$$\begin{aligned} \overline{{\widetilde{V}}}_{\infty ,\infty }(K,L,Q)\le 1\ \ \ \text{ if } \text{ and } \text{ only } \text{ if }\ \ \ L\subseteq \left( \frac{\rho _{Q}}{\rho _{K}}\right) K. \end{aligned}$$
(1.17)

In the sequel, we use \({\mathcal {E}}^{n}\) to denote the class of origin-symmetric ellipsoids in \({\mathbb {R}}^{n}\).

Inspired by the constrained maximization problem (1.3) posed by Lutwak, Yang and Zhang [30], this paper will consider a (pq)-version of the problem:

Optimization Problems 1.1

Let \(0<p\le q\) with \(q=p+r\), \(r\ge 0\). For \(K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\), find an ellipsoid, among all origin-symmetric ellipsoids, which solves the following constrained maximization problem:

$$\begin{aligned} \max \limits _{E\in {\mathcal {E}}^{n}}\left( \frac{|E|}{\omega _{n}}\right) ^{\frac{1}{n}}\ \ \ \text{ subject } \text{ to }\ \ \ \overline{{\widetilde{V}}}_{p,q}(K,E,Q)\le 1. \end{aligned}$$
($S_{p,q}$)

An ellipsoid that solves the constrained maximization problem will be called a \(S_{p,q}\) solution for K and Q. The dual problem is

$$\begin{aligned} \min \limits _{E\in {\mathcal {E}}^{n}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\ \ \ \text{ subject } \text{ to }\ \ \ \left( \frac{|E|}{\omega _{n}}\right) ^{\frac{1}{n}}\ge 1. \end{aligned}$$
(${\mathaccentV {bar}24E{S)

An ellipsoid that solves the dual problem will be called a \({\bar{S}}_{p.q}\) solution for K and Q.

We will prove in Sect. 4 there is a unique solution to the constrained maximization problem, which will be called the (pq)-John ellipsoid \(E_{p,q}(K,Q)\) in Definition 4.6. The dual problem is equivalent to the problem of minimizing total \(L_{p}\) dual curvature measures under \(\mathrm{SL}(n)\)-transformations. The dual problem has a unique solution with volume \(\omega _n\), which differs by only a scale factor to the \(S_{p,q}\) solution. Therefore, it is called the normalized (pq)-John ellipsoid \({\overline{E}}_{p,q}(K,Q)\).

In the case of \(Q=K\), \(E_{p,q}(K,Q)=E_{p}(K)\) is the \(L_{p}\) John ellipsoid studied by Lutwak, Yang and Zhang [30]. In the case that \(Q=B\) and \(p=n\), one also has \(E_{p,q}(K,Q)=E_{p}(K)\).

This paper is organized as follows. In Sect. 2 we recall some basic results in convex geometry. Section 3 proves the continuity of \({\widetilde{V}}_{p,q}\) and \(\overline{{\widetilde{V}}}_{p,q}\). We prove in Sect. 4 the existence, uniqueness and geometric characterization of the (pq)-John ellipsoid which solves Problem 1.1. Using the continuity of \({\widetilde{V}}_{p,q}\) and \(\overline{{\widetilde{V}}}_{p,q}\), we study continuity of (pq)-John ellipsoids in Sect. 5 In Sect. 6, we discuss generalizations of John’s inclusion for (pq)-John ellipsoids. In the last section, the inequality for the volume ratio is established.

2 Preliminaries

For quick reference we recall some basic results of convex geometry. We refer the reader to [10, 46] for details.

The setting will be the n-dimensional Euclidean space \({\mathbb {R}}^{n}\). As usual \(x\cdot y\) denotes the standard inner product of x and y in \({\mathbb {R}}^{n}\). For \(x\in {\mathbb {R}}^{n}\), let \(|x|=\sqrt{x\cdot x}\) be the Euclidean norm of x. For \(x\in {\mathbb {R}}^{n}\backslash \{o\}\), we use both \({\bar{x}}\) and \(\langle x\rangle \) to denote x/|x|.

In addition to its denoting absolute value, without confusion we will use \(|\cdot |\) to denote the standard Euclidean norm on \({\mathbb {R}}^{n}\), often to denote n-dimensional volume, and on occasion to denote the absolute value of the determinant of an \(n\times n\) matrix.

For \(K\in {\mathcal {K}}^{n}_{o}\), its support function, \(h_{K}: {\mathbb {R}}^{n}\rightarrow {\mathbb {R}}\) is defined by \(h_{K}(x)=\max \{x\cdot y: y\in K\}\), for \(x\in {\mathbb {R}}^{n}\). Obviously, for real \(\lambda >0\),

$$\begin{aligned} h_{\lambda K}(x)=\lambda h_{K}(x), \text{ for } x\in {\mathbb {R}}^{n}. \end{aligned}$$
(2.1)

More generally, for \(\phi \in {\mathrm{GL}}(n)\) the image \(\phi K=\{\phi x: x\in K\}\) have that

$$\begin{aligned} h_{\phi K}(x)= h_{K}(\phi ^{t}x), \end{aligned}$$
(2.2)

where \(\phi ^{t}\) denotes the transpose of \(\phi \).

The Hausdorff distance between convex bodies K and L is given by

$$\begin{aligned} \delta _{H}(K,L):=|h_{K}-h_{L}|_{\infty }=\max \limits _{u\in S^{n-1}}|h_{K}(u)-h_{L}(u)|. \end{aligned}$$

If \(K,L\in {\mathcal {K}}^{n}_{o}\), then for real \(p>0\), the \(L_{p}\)-mixed volume of K and L is defined by

$$\begin{aligned} V_{p}(K,L)=\frac{1}{n}\int _{S^{n-1}}h_{L}^{p}(u){\mathrm{d}}S_{p}(K,u). \end{aligned}$$
(2.3)

If K contains the origin in its interior, then its polar body \(K^{*}\) is given by \(K^{*}=\{x\in {\mathbb {R}}^{n}: x\cdot y\le 1\ \ \text{ for } \text{ all }\ y\in K\}\). Obviously, for \(\phi \in {\mathrm{GL}}(n)\),

$$\begin{aligned} (\phi K)^{*}=\phi ^{-t}K^{*}, \end{aligned}$$
(2.4)

where \(\phi ^{-t}\) denotes the inverse of the transpose of \(\phi \).

A star body \(K\subset {\mathbb {R}}^{n}\) is a compact star-shaped set about the origin whose radial function \(\rho _{K}: {\mathbb {R}}^{n}\backslash \{o\}\rightarrow {\mathbb {R}}\), defined for \(x\in {\mathbb {R}}^{n}\backslash \{o\}\) by \(\rho _{Q}(x)=\max \{\lambda >0: \lambda x\in Q\}\), is continuous. We call two star bodies K and L in \({\mathcal {S}}^{n}_{o}\) are dilates (of one another) if \(\rho _{K}(u)/ \rho _{L}(u)\) is independent of \(u\in S^{n-1}\). If \(\lambda >0\), we have

$$\begin{aligned} \rho _{\lambda K}(x)=\lambda \rho _{K}(x),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}\backslash \{o\}. \end{aligned}$$
(2.5)

More generally, for \(\phi \in {\mathrm{GL}}(n)\), the image \(\phi K=\{\phi x: x\in K\}\) of K have the property

$$\begin{aligned} \rho _{\phi K}(x)=\rho _{K}(\phi ^{-1}x), \end{aligned}$$
(2.6)

for all \(x\in {\mathbb {R}}^{n}\backslash \{o\}\).

The radial distance between \(K,L\in {\mathcal {S}}^{n}_{o}\) is

$$\begin{aligned} {\widetilde{\delta }}_{H}(K,L):=|\rho _{K}-\rho _{L}|_{\infty }=\max \limits _{u\in S^{n-1}}|\rho _{K}(u)-\rho _{L}(u)|. \end{aligned}$$

The dual Brunn–Minkowski theory is a theory of dual mixed volumes of star bodies. For \(q\in {\mathbb {R}}\), the q-th dual mixed volume of \(K,Q\in {\mathcal {S}}^{n}_{o}\), is defined by (see [31])

$$\begin{aligned} {\widetilde{V}}_{q}(K,Q)=\frac{1}{n}\int _{S^{n-1}}\rho _{K}^{q}(u)\rho _{Q}^{n-q}(u){\mathrm{d}}u, \end{aligned}$$
(2.7)

where the integral is with respect to spherical Lebesgue measure. It is well know that for \(\phi \in {\mathrm{GL}}(n)\),

$$\begin{aligned} {\widetilde{V}}_{q}(\phi K, \phi Q)=|\phi |{\widetilde{V}}_{q}(K,Q),\ \ q\in {\mathbb {R}}\backslash \{0\}. \end{aligned}$$
(2.8)

Dual Minkowski inequality can be expressed as follows: If \(0\le q\le n\) and \(K,Q\in {\mathcal {S}}^{n}_{o}\), then

$$\begin{aligned} {\widetilde{V}}_{q}(K,Q)^{n}\le |K|^{q}|Q|^{n-q}, \end{aligned}$$
(2.9)

with equality if and only if K and Q are dilates when \(0<q<n\).

If \(K\in {\mathcal {K}}^{n}_{o}\), then it is easy to see that the radial function and the support function of K are related by

$$\begin{aligned} h_{K}(v)= & {} \max \limits _{u\in S^{n-1}}(u\cdot v)\rho _{K}(u),\ \ \text{ for }\ v\in S^{n-1}, \end{aligned}$$
(2.10)
$$\begin{aligned} \frac{1}{\rho _{K}(u)}= & {} \max \limits _{v\in S^{n-1}}\frac{u\cdot v}{h_{K}(v)},\ \ \text{ for }\ u\in S^{n-1}. \end{aligned}$$
(2.11)

From definitions of \({\widetilde{V}}_{p,q}\) and the radial Gauss map, the support function and the radial function imply that

Lemma 2.1

Let \(\lambda >0\), then

$$\begin{aligned} {\widetilde{V}}_{p,q}(\lambda K,L,Q)= & {} \lambda ^{q-p}{\widetilde{V}}_{p,q}(K,L,Q), \end{aligned}$$
(2.12)
$$\begin{aligned} {\widetilde{V}}_{p,q}(K,\lambda L,Q)= & {} \lambda ^{p}{\widetilde{V}}_{p,q}(K,L,Q), \end{aligned}$$
(2.13)
$$\begin{aligned} {\widetilde{V}}_{p,q}(K,L,\lambda Q)= & {} \lambda ^{n-q}{\widetilde{V}}_{p,q}(K,L,Q). \end{aligned}$$
(2.14)

For \(\lambda >0\) and \(p\in (0,\infty ], q=p+r, r\in [0,\infty )\), based on the (1.13), (2.1) and (2.5), we can immediately obtain the results,

Lemma 2.2

Let \(\lambda >0\), then

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(\lambda K,L,Q)= & {} \overline{{\widetilde{V}}}_{p,q}(K,L,Q), \end{aligned}$$
(2.15)
$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(K,\lambda L,Q)= & {} \lambda \overline{{\widetilde{V}}}_{p,q}(K,L,Q),\end{aligned}$$
(2.16)
$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(K,L,\lambda Q)= & {} \lambda ^{-1}\overline{{\widetilde{V}}}_{p,q}(K,L,Q). \end{aligned}$$
(2.17)

We shall need the following fact.

Lemma 2.3

(cf. [31]) The (pq)-mixed volume is \({\mathrm{SL}}(n)\)-invariant, in that for \(p,q\in {\mathbb {R}}\), and \(K,L\in {\mathcal {K}}^{n}_{o}\), with \(Q\in {\mathcal {S}}^{n}_{o}\),

$$\begin{aligned} {\widetilde{V}}_{p,q}(\phi K, \phi L, \phi Q)={\widetilde{V}}_{p,q}(K,L,Q), \end{aligned}$$
(2.18)

for each \(\phi \in {\mathrm{SL}}(n)\).

Lemma 2.1, together with Lemma 2.3, shows that for \(\phi \in {\mathrm{GL}}(n)\),

$$\begin{aligned} {\widetilde{V}}_{p,q}(\phi K, \phi L, \phi Q)=|\phi |{\widetilde{V}}_{p,q}(K,L,Q). \end{aligned}$$
(2.19)

We will also need the fact that for \(\phi \in {\mathrm{GL}}(n)\) and \(p\in (0,\infty ], q=p+r, r\in [0,\infty )\),

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(\phi K, \phi L, \phi Q)=\overline{{\widetilde{V}}}_{p,q}(K,L,Q). \end{aligned}$$
(2.20)

This follows immediately from (2.8) and (2.19) for all \(p\in (0,\infty ], q=p+r\) and \(r\in [0,\infty )\).

The following inequality for (pq)-mixed volume is a generalization of the \(L_{p}\) Minkowski inequality for mixed volume (see [31]).

Lemma 2.4

Suppose pq are such that \(1\le \frac{q}{n}\le p\). If \(K,L\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\), then

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,L,Q)^{n}\ge |K|^{q-p}|L|^{p}|Q|^{n-q}, \end{aligned}$$
(2.21)

with equality if and only if KLQ are dilates when \(1<\frac{q}{n}<p\), while only K and L need be dilates when \(q=n\) and \(p >1\), and K and L are homothets when \(q=n\) and \(p =1\).

We shall require the following definition.

Definition 2.5

(cf. [31]) Suppose \(p\in {\mathbb {R}}\). If \(\mu \) is a Borel measure on \(S^{n-1}\) and \(\phi \in {\mathrm{SL}}(n)\) then, \(\phi _{p}\dashv \mu \), the \(L_{p}\) image of \(\mu \) under \(\phi \), is a Borel measure such that

$$\begin{aligned} \int _{S^{n-1}}f(u){\mathrm{d}}\phi _{p}\dashv \mu (u)=\int _{S^{n-1}}|\phi ^{-1}u|^{p}f(\langle \phi ^{-1}u\rangle ){\mathrm{d}}\mu (u) \end{aligned}$$

for each Borel \(f: S^{n-1}\rightarrow {\mathbb {R}}\).

Lemma 2.6

(cf. [31]) Suppose \(p\ne 0\) and \(q\ne 0\). Then for all \(Q\in {\mathcal {S}}^{n}_{o}\) and \(K,L\in {\mathcal {K}}^{n}_{o}\), and \(\phi \in {\mathrm{SL}}(n)\),

$$\begin{aligned} {\widetilde{C}}_{p,q}(\phi K,\phi Q, \cdot )=\phi ^{t}_{p}\dashv {\widetilde{C}}_{p,q}(K,L,\cdot ). \end{aligned}$$
(2.22)

We also need the following lemma:

Lemma 2.7

(cf. [19]) Suppose \(K_{i}\in {\mathcal {K}}^{n}_{o}\) with \(\lim \limits _{i\rightarrow \infty }K_{i}=K_{0}\). Let \(\omega =\cup _{i=0}^{\infty }\omega _{K_{i}}\), be the set (of \({\mathcal {H}}^{n-1}\)-measure 0) off of which all of the \(\alpha _{K_{i}}\) are defined. Then if \(u_{i}\in S^{n-1}\backslash \omega \) are such that \(\lim \limits _{i\rightarrow \infty }u_{i}=u_{0}\in S^{n-1}\backslash \omega \), then \(\lim \limits _{i\rightarrow \infty }\alpha _{K_{i}}(u_{i})=\alpha _{K_{0}}(u_{0})\).

Let \(K\in {\mathcal {K}}^{n}_{o}\). The classical projection body \(\Pi K\) of K is given by (see [10])

$$\begin{aligned} h_{\Pi K}(u)={\mathrm{vol}}_{n-1}(K|u^{\bot })=\frac{1}{2}\int _{S^{n-1}}|u\cdot v|{\mathrm{d}}S(K,v),\ \ \forall u\in S^{n-1}. \end{aligned}$$

We will use the concept of a \(L_{p}\)-projection body (see [28, 29, 45, 52]). For \(p\ge 1\), the \(L_{p}\)-projection body \(\Pi _{p}K\) is given by

$$\begin{aligned} h_{\Pi _{p}K}(u)=\left( \frac{1}{2n}\int _{S^{n-1}}|u\cdot v|^{p}{\mathrm{d}}S_{p}(K,v)\right) ^{\frac{1}{p}},\ \ u\in S^{n-1}, \end{aligned}$$

where \(S_{p}(K,\cdot )\) is the \(L_{p}\)-surface area measure. Clearly, \(\Pi _1 K=\frac{1}{n}\Pi K\).

We shall use the concepts of (pq)-mixed projection body and (pq)-mixed polar projection body. For each \(K\in {\mathcal {K}}^{n}_{o}\) with a star body \(Q\in {\mathcal {S}}^{n}_{o}\), and \(p>0, q>0\), the (pq)-mixed projection body, \(\Pi _{p,q}(K,Q)\), of K and Q is the origin-symmetric convex body whose support function is defined by

$$\begin{aligned} h_{\Pi _{p,q}(K,Q)}(u)=\left( \frac{1}{2}\int _{S^{n-1}}|u \cdot v|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\right) ^{\frac{1}{p}}, \ \ \text{ for } \text{ all }\ u\in S^{n-1}.\qquad \end{aligned}$$
(2.23)

In particular, we have \(\Pi _{p,n}(K,B)=\Pi _{p,q}(K,K)=\Pi _{p}K\) for \(p>1\), and \(\Pi _{1,n}(K,B)=\Pi _{1,q}(K,K)=\Pi _{1}(K)=\frac{1}{n}\Pi K\).

If \(K\in {\mathcal {K}}^{n}_{o}\) and real \(p>0\), the star body \(\Gamma _{-p}K\) (called by \(L_{p}\)-polar projection body, see [30]) is defined as, for \(u\in S^{n-1}\) :

$$\begin{aligned} \rho _{\Gamma _{-p}K}(u)^{-1}=\left( \frac{1}{|K|}\int _{S^{n-1}}|u\cdot v|^{p}{\mathrm{d}}S_{p}(K,v)\right) ^{\frac{1}{p}}. \end{aligned}$$

If \(K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\), and real \(p>0,q>0\) and \(q=p+r, r\in [0, +\infty )\), the star body \(\Gamma _{-p,-q}(K,Q)\) is defined by, for \(x\in {\mathbb {R}}^{n}\),

$$\begin{aligned} \rho _{\Gamma _{-p,-q}(K,Q)}^{-1}(x)=\left( \frac{n}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}}|x \cdot v|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\right) ^{\frac{1}{p}}. \end{aligned}$$
(2.24)

The star body \(\Gamma _{-p,-q}(K,Q)\) is called the (pq)-mixed polar projection body of K and Q. It is easy to know that \(\Gamma _{-p,-q}(K,K)=\Gamma _{-p}K\).

Note that for \(q\ge p\ge 1\), the body \(\Gamma _{-p,-q}(K,Q)\) is a convex body. Define \(\Gamma _{-\infty ,-\infty }(K,Q)\) by

$$\begin{aligned} \Gamma _{-\infty ,-\infty }(K,Q)=\lim \limits _{p\rightarrow \infty }\Gamma _{-p,-q}(K,Q). \end{aligned}$$
(2.25)

For real \(p >0,q=p+r, r\in [0, +\infty )\), and using(1.5), we can rewrite (2.25) as

$$\begin{aligned} n^{-\frac{1}{p}}\rho _{\Gamma _{-p,-q}(K,Q)}(u)^{-1} =\left( \int _{S^{n-1}}\left( \frac{|u \cdot v|\rho _{K}(v)}{h_{K}(\alpha _{K}(v))\rho _{Q}(v)}\right) ^{p}{\mathrm{d}}{\widetilde{V}}_{r}(K,Q; v)\right) ^{\frac{1}{p}}, \end{aligned}$$
(2.26)

for \(u\in S^{n-1}\). Thus, from (2.25) and (2.26),

$$\begin{aligned}&\rho _{\Gamma _{-\infty ,-\infty }(K,Q)}(u)^{-1}=\max \left\{ \frac{|u \cdot v|\rho _{K}(v)}{h_{K}(\alpha _{K}(v))\rho _{Q}(v)}: v\in {\mathrm{supp}}{\widetilde{V}}_{r}(K,Q; \cdot )\right\} ,\nonumber \\&\quad \ u\in S^{n-1}. \end{aligned}$$
(2.27)

3 The Continuity of \({\widetilde{V}}_{p,q}\) and \(\overline{{\widetilde{V}}}_{p,q}\)

In this section, we consider the continuity of \({\widetilde{V}}_{p,q}\) and \(\overline{{\widetilde{V}}}_{p,q}\).

Theorem 3.1

Suppose \(K,K_{i},L,L_{j}\in {\mathcal {K}}^{n}_{o}, Q,Q_{k}\in {\mathcal {S}}^{n}_{o}\) and \(p_{l},p, q_{m},q\in (0,\infty ]\), where \(i,j,k,l,m\in {\mathbb {N}}\). Let \(r\in [0,+\infty )\). If \(K_{i}\rightarrow K,L_{j}\rightarrow L,Q_{k}\rightarrow Q, p_{l}\rightarrow p\), and \(q_{m}\rightarrow q\) as \(i,j,k,l,m\rightarrow \infty \), then

$$\begin{aligned} \lim \limits _{i,j,k,l,m\rightarrow \infty }{\widetilde{V}}_{p_{l},q_{m}}(K_{i},L_{j},Q_{k})={\widetilde{V}}_{p,q}(K,L,Q), \end{aligned}$$
(3.1)

and

$$\begin{aligned} \lim \limits _{i,j,k,l\rightarrow \infty }\overline{{\widetilde{V}}}_{p_{l},p_{l}+r}(K_{i},L_{j},Q_{k}) =\overline{{\widetilde{V}}}_{p,p+r}(K,L,Q). \end{aligned}$$
(3.2)

Proof

Let

$$\begin{aligned} c_{m}=\min \{c_1,c_2\},~ c_{M}=\max \{c_3,c_4\}, \end{aligned}$$

where

$$\begin{aligned} c_{1}= & {} \frac{\inf \left( \left\{ \min \limits _{S^{n-1}}h_{L}\right\} \cup \left\{ \min \limits _{S^{n-1}}h_{L_{j}}: j\in {\mathbb {N}}\right\} \right) }{\sup \left( \left\{ \max \limits _{S^{n-1}}h_{K}\right\} \cup \left\{ \max \limits _{S^{n-1}}h_{K_{i}}: i\in {\mathbb {N}}\right\} \right) },\\ c_2= & {} \frac{\inf \left( \left\{ \min \limits _{S^{n-1}}\rho _{K}\right\} \cup \left\{ \min \limits _{S^{n-1}}\rho _{K_{i}}: i\in {\mathbb {N}}\right\} \right) }{\sup \left( \left\{ \max \limits _{S^{n-1}}\rho _{Q}\right\} \cup \left\{ \max \limits _{S^{n-1}}\rho _{Q_{k}}: k\in {\mathbb {N}}\right\} \right) },\\ c_3= & {} \frac{\sup \left( \left\{ \max \limits _{S^{n-1}}h_{L}\right\} \cup \left\{ \max \limits _{S^{n-1}}h_{L_{j}}: j\in {\mathbb {N}}\right\} \right) }{\inf \left( \left\{ \min \limits _{S^{n-1}}h_{K}\right\} \cup \left\{ \min \limits _{S^{n-1}}h_{K_{i}}: i\in {\mathbb {N}}\right\} \right) }, \end{aligned}$$

and

$$\begin{aligned} c_4=\frac{\sup \left( \left\{ \max \limits _{S^{n-1}}\rho _{K}\right\} \cup \left\{ \max \limits _{S^{n-1}}\rho _{K_{i}}: i\in {\mathbb {N}}\right\} \right) }{\inf \left( \left\{ \min \limits _{S^{n-1}}\rho _{Q}\right\} \cup \left\{ \min \limits _{S^{n-1}}\rho _{Q_{k}}: k\in {\mathbb {N}}\right\} \right) }. \end{aligned}$$

We first claim \(0<c_{m}\le c_{M}<\infty \). Since \(K_{i}\rightarrow K,L_{j}\rightarrow L\) and \(Q_{k}\rightarrow Q, p_{l}\rightarrow p\) as \(i,j,k\rightarrow \infty \), we have \(h_{K_{i}}\rightarrow h_{K}\), \(h_{L_{j}}\rightarrow h_{L}\) and \(h_{L_{k}}\rightarrow h_{L}\) uniformly on \(S^{n-1}\), respectively. From \(K,K_{i},L,L_{j}\in {\mathcal {K}}^{n}_{o}, Q,Q_{k}\in {\mathcal {S}}^{n}_{o}\), it follows that there exists an \(N_{0}\in N\), such that for all \(i,j,k>N_{0}\) and \(u\in S^{n-1}\),

$$\begin{aligned} \min _{S^{n-1}}h_{\frac{1}{2}K}\le & {} h_{K_{i}}(u)\le \max _{S^{n-1}}h_{2K}\ \ \text{ and }\ \ \min _{S^{n-1}}h_{\frac{1}{2}L}\le h_{L_{j}}(u)\le \max _{S^{n-1}}h_{2L},\\ \min _{S^{n-1}}\rho _{\frac{1}{2}K}\le & {} \rho _{K_{i}}(u)\le \max _{S^{n-1}}\rho _{2K}\ \ \text{ and }\ \ \min _{S^{n-1}}\rho _{\frac{1}{2}Q}\le \rho _{Q_{k}}(u)\le \max _{S^{n-1}}\rho _{2Q}. \end{aligned}$$

For brevity, we write

$$\begin{aligned} a_{m}=\min \{a:a\in A_{1}\cup A_{2}\},\ \ a_{M}=\max \{a:a\in A_{3}\cup A_{4}\}, \end{aligned}$$

where

$$\begin{aligned} A_{1}= & {} \bigcup \limits _{u\in S^{n-1}}\left\{ h_{\frac{1}{2}K}(u),h_{\frac{1}{2}L}(u), \rho _{\frac{1}{2}K}(u),\rho _{\frac{1}{2}Q}(u)\right\} ,\\ A_{2}= & {} \bigcup \limits _{1\le i\le N_{0}}\bigcup \limits _{u\in S^{n-1}}\left\{ h_{K_{i}}(u), h_{L_{j}}(u), \rho _{K_{i}}(u), \rho _{Q_{k}}(u)\right\} ,\\ A_{3}= & {} \bigcup \limits _{u\in S^{n-1}}\left\{ h_{2K}(u),h_{2L}(u),\rho _{2K}(u),\rho _{2Q}(u)\right\} , \end{aligned}$$

and

$$\begin{aligned} A_{4}=\bigcup \limits _{1\le i\le N_{0}}\bigcup \limits _{u\in S^{n-1}}\left\{ h_{K_{i}}(u),h_{L_{j}}(u),\rho _{K_{i}}(u), \rho _{Q_{k}}(u)\right\} . \end{aligned}$$

Then we have \(0<a_{m}\le a_{M}<\infty \), and

$$\begin{aligned}&a_{m}B\subseteq K\subseteq a_{M}B,\ \ \ a_{m}B\subseteq K_{i}\subseteq a_{M}B\ \ \ \text{ for }\ i\in {\mathbb {N}},\\&a_{m}B\subseteq L\subseteq a_{M}B,\ \ \ a_{m}B\subseteq L_{j}\subseteq a_{M}B\ \ \ \text{ for }\ j\in {\mathbb {N}},\\&a_{m}B\subseteq Q\subseteq a_{M}B,\ \ \ a_{m}B\subseteq Q_{k}\subseteq a_{M}B\ \ \ \text{ for }\ k\in {\mathbb {N}}. \end{aligned}$$

Thus, by the definitions of \(c_{m}\) and \(c_{M}\), it yields

$$\begin{aligned} 0<\frac{a_{m}}{a_{M}}\le c_{m}\le c_{M}\le \frac{a_{M}}{a_{m}}<\infty . \end{aligned}$$

Next, we prove

$$\begin{aligned} \lim \limits _{i,j,k,l,m\rightarrow \infty }{\widetilde{V}}_{p_{l},q_{m}}(K_{i},L_{j},Q_{k})={\widetilde{V}}_{p,q}(K,L,Q). \end{aligned}$$

For any \(\varepsilon >0\), three observations are in order. Firstly, let \(f(t)=t^p,f_{l}(t)=t^{p_{l}},l=1,2,\cdots ,\) defined on \([c_{m},c_{M}]\), then the sequence of \(\{f_{l}\}\) converges uniformly to f on \([c_{m},c_{M}]\). And let \(g(t)=t^p,g_{m}(t)=t^{p_{m}},m=1,2,\cdots ,\) defined on \([c_{m},c_{M}]\), then the sequence of \(\{g_{m}\}\) converges uniformly to g on \([c_{m},c_{M}]\). For all \(u\in S^{n-1}\),

$$\begin{aligned} c_{m}\le \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))}\le c_{M},\ \ c_{m}\le \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\le c_{M}, \end{aligned}$$

there exists an \(N_{1}\in {\mathbb {N}}\), such that for all \(l,m\ge N_{1}\),

$$\begin{aligned} \left| \left( \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))} \right) ^{p_{l}}\left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q_{m}}- \left( \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))} \right) ^{p}\left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q}\right| <\frac{\varepsilon }{3}, \end{aligned}$$
(3.3)

independently of i and j and uniformly on \(u\in S^{n-1}\).

Secondly, since \(K_{i}\rightarrow K,L_{j}\rightarrow L\) and \(Q_{k}\rightarrow Q, p_{l}\rightarrow p\) as \(i,j,k\rightarrow \infty \), and Lemma 2.7, there exists an \(N_{2}\in {\mathbb {N}}\) such that for all \(i,j,k>N_{2}\) and for all \(u\in S^{n-1}\),

$$\begin{aligned} \left| \left( \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))} \right) ^{p}\left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q}- \left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q}\right| <\frac{\varepsilon }{3}. \end{aligned}$$
(3.4)

Indeed, since functions f and g are all Lipschitzian on \([c_{m},c_{M}]\), there exist constants \(C_{1}, C_{2}>0\), such that for all \(u\in S^{n-1}\),

$$\begin{aligned}&\left| \left( \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))} \right) ^{p}\left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q}- \left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q}\right| \\&\quad \le \left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q} \left| \left( \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))} \right) ^{p}-\left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p}\right| \\&\qquad +\left( \frac{h_{L}(\alpha _{K})(u)}{h_{K}(\alpha _{K})(u)}\right) ^{p} \left| \left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q} -\left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q}\right| \\&\quad \le C_{1}\left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q} \left| \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))}- \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right| \\&\qquad +C_{2}\left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))} \right) ^{p}\left| \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}- \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right| \\&\quad \le C_{M}^qC_{1}\cdot \frac{\delta _{H}(L_{j},L)\max _{S^{n-1}}h_{K} +\delta _{H}(K_{i},K)\max _{S^{n-1}}h_{L}}{\min _{S^{n-1}}h_{K_{i}} \min _{S^{n-1}}h_{K}}\\&\qquad +C_{M}^PC_{2}\cdot \frac{{\widetilde{\delta }}_{H}(K_{i},K )\max _{S^{n-1}}\rho _{Q}+{\widetilde{\delta }}_{H}(Q_{k},Q)\max _{S^{n-1}}\rho _{K}}{\min _{S^{n-1}}\rho _{Q_{k}}\min _{S^{n-1}}\rho _{Q}}. \end{aligned}$$

Thirdly, since the measure sequence \(\{\overline{{\widetilde{V}}}_{Q_{k}}\}\) weakly converges to \(\overline{{\widetilde{V}}}_{Q}\), there exists an \(N_{3}\in {\mathbb {N}}\), such that for all \(k\ge N_{3}\),

$$\begin{aligned}&\left| \int _{S^{n-1}}\left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))} \right) ^{p}\left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q}{\mathrm{d}}\overline{{\widetilde{V}}}_{Q_{k}}(u)\right. \nonumber \\&\quad \left. -\int _{S^{n-1}} \left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q}{\mathrm{d}}\overline{{\widetilde{V}}}_{Q}(u)\right| <\frac{\varepsilon }{3}. \end{aligned}$$
(3.5)

From (3.3), (3.4) and (3.5), it follows that for all \(i,j,k,l,m\ge \max \{N_{1},N_{2},N_{3}\}\),

$$\begin{aligned}&\left| \int _{S^{n-1}}\left( \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))}\right) ^{p_{l}}\left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q_{m}}{\mathrm{d}}\overline{{\widetilde{V}}}_{Q_{k}}(u)- \int _{S^{n-1}}\left( \frac{h_{L}(\alpha _{K}(u))}{h_{K} (\alpha _{K}(u))}\right) ^{p}\left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q} {\mathrm{d}}\overline{{\widetilde{V}}}_{Q}(u)\right| \\&\quad \le \int _{S^{n-1}}\left| \left( \frac{h_{L_{j}} (\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))}\right) ^{p_{l}} \left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q_{m}}- \left( \frac{h_{L_{j}}(\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))} \right) ^{p}\left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q} \right| {\mathrm{d}}\overline{{\widetilde{V}}}_{Q_{k}}(u)\\&\qquad +\int _{S^{n-1}}\left| \left( \frac{h_{L_{j}} (\alpha _{K_{i}}(u))}{h_{K_{i}}(\alpha _{K_{i}}(u))}\right) ^{p} \left( \frac{\rho _{K_{i}}(u)}{\rho _{Q_{k}}(u)}\right) ^{q}- \left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q}\right| {\mathrm{d}} \overline{{\widetilde{V}}}_{Q_{k}}(u)\\&\qquad +\left| \int _{S^{n-1}}\left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p}\left( \frac{\rho _{K}(u)}{\rho _{Q}(u)} \right) ^{q}{\mathrm{d}}\overline{{\widetilde{V}}}_{Q_{k}}(u)- \left( \frac{h_{L}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q}{\mathrm{d}} \overline{{\widetilde{V}}}_{Q}(u)\right| \\&\quad <\varepsilon . \end{aligned}$$

Namely,

$$\begin{aligned} \lim \limits _{i,j,k,l,m\rightarrow \infty }\frac{{\widetilde{V}}_{p_{l},q_{m}}(K_{i},L_{j},Q_{k})}{|Q_{k}| }=\frac{{\widetilde{V}}_{p,q}(K,L,Q)}{|Q|}. \end{aligned}$$

The first conclusion follows from the fact \(|Q_{k}|\rightarrow |Q|\) by sending k to infinity.

Finally, we proceed to prove

$$\begin{aligned} \lim \limits _{i,j,k,l\rightarrow \infty }\overline{{\widetilde{V}}}_{p_{l},p_{l}+r}(K_{i},L_{j},Q_{k}) =\overline{{\widetilde{V}}}_{p,p+r}(K,L,Q). \end{aligned}$$

Fix \(\delta >0\). For \(0\le r<\infty \), we note that

$$\begin{aligned} \frac{{\widetilde{V}}_{p_{l},p_{l}+r}(K_{i},L_{j},Q_{k})}{{\widetilde{V}}_{r}(K_{i},Q_{k})}, \frac{{\widetilde{V}}_{p,p+r}(K,L,Q)}{{\widetilde{V}}_{r}(K,Q)}\in [c_1,c_3],\ \ \text{ for } \text{ each }\ i,j,k,l\in {\mathbb {N}}. \end{aligned}$$

The continuity of \(t^{\frac{1}{p}}\) on \([c_1,c_3]\) implies there exists an \(N_4>0\) such that for all \(l\ge N_4\),

$$\begin{aligned} \left| \left( \frac{{\widetilde{V}}_{p_{l},p_{l}+r}(K_{i},L_{j},Q_{k})}{{\widetilde{V}}_{r}(K_{i},Q_{k})}\right) ^{\frac{1}{p_{l}}} -\left( \frac{{\widetilde{V}}_{p_{l},p_{l}+r}(K_{i},L_{j},Q_{k})}{{\widetilde{V}}_{r}(K_{i},Q_{k})}\right) ^{\frac{1}{p}}\right| <\frac{\delta }{2} \end{aligned}$$
(3.6)

holds independently of ij and k.

From (1.11) and (3.1), it follows \(\lim \limits _{i,k\rightarrow \infty }{\widetilde{V}}_{r}(K_{i},Q_{k})={\widetilde{V}}_{r}(K,Q)\). Combining this with (3.1), the continuity of \(t^{\frac{1}{p}}\) on \([c_1,c_3]\) shows there exists an \(N_{5}>0\), such that for all \(i,j,k,l>N_{5}\),

$$\begin{aligned} \left| \left( \frac{{\widetilde{V}}_{p_{l},p_{l}+r}(K_{i},L_{j},Q_{k})}{{\widetilde{V}}_{r}(K_{i},Q_{k})}\right) ^{\frac{1}{p}}-\left( \frac{{\widetilde{V}}_{p,p+r}(K,L,Q)}{{\widetilde{V}}_{r}(K,Q)}\right) ^{\frac{1}{p}}\right| <\frac{\delta }{2}. \end{aligned}$$
(3.7)

In terms of (3.6) and (3.7), it follows that for \(i,j,k,l\ge \max \{N_{4},N_{5}\}\),

$$\begin{aligned} \left| \left( \frac{{\widetilde{V}}_{p_{l},q_{l}}(K_{i},L_{j},Q_{k})}{{\widetilde{V}}_{r}(K_{i},Q_{k})}\right) ^{\frac{1}{p_{l}}} -\left( \frac{{\widetilde{V}}_{p,q}(K,L,Q)}{{\widetilde{V}}_{r}(K,Q)} \right) ^{\frac{1}{p}}\right| <\delta . \end{aligned}$$

That is,

$$\begin{aligned} \lim \limits _{i,j,k,l\rightarrow \infty }\overline{{\widetilde{V}}}_{p_{l},p_{l}+r}(K_{i},L_{j},Q_{k}) =\overline{{\widetilde{V}}}_{p,p+r}(K,L,Q). \end{aligned}$$

\(\square \)

4 (pq)-John Ellipsoids

In this section, we focus on the main Problem 1.1 proposed in Sect. 1.

Optimization Problems. Let \(0<p\le q\) with \(q=p+r\), \(r\ge 0\). For \(K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\), find an ellipsoid, among all origin-symmetric ellipsoids, which solves the following constrained maximization problem:

$$\begin{aligned} \max \limits _{E\in {\mathcal {E}}^{n}}\left( \frac{|E|}{\omega _{n}}\right) ^{\frac{1}{n}}\ \ \ \text{ subject } \text{ to }\ \ \ \overline{{\widetilde{V}}}_{p,q}(K,E,Q)\le 1. \end{aligned}$$
($S_{p,q}$)

An ellipsoid that solves the constrained maximization problem will be called a \(S_{p,q}\) solution for K and Q. The dual problem is

$$\begin{aligned} \min \limits _{E\in {\mathcal {E}}^{n}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\ \ \ \text{ subject } \text{ to }\ \ \ \left( \frac{|E|}{\omega _{n}}\right) ^{\frac{1}{n}}\ge 1. \end{aligned}$$
(${\mathaccentV {bar}24E{S)

An ellipsoid that solves the dual problem will be called a \({\bar{S}}_{p.q}\) solution for K and Q.

The following theorem gives the existence of Problem \(S_{p,q}\) when \(0<p\le q\), and proves its uniqueness when \(1\le p\le q\).

Theorem 4.1

For any \(0<p\le q\), there exists an ellipsoid which solves Problem \(S_{p,q}\). The solution is unique for \(1\le p\le q\).

Proof

For an ellipsoid \(E\in {\mathcal {E}}^{n}\) (the class of origin-symmetric ellipsoids in \({\mathbb {R}}^n\)), we use \(d_{E}\) to denote its maximal principal radius. There exists a \(v_{E}\in S^{n-1}\) such that \(d_{E}|v_{E}\cdot u|\le h_{E}(u),\ \text{ for } \text{ all }\ u\in S^{n-1}\). From definitions of the (pq)-mixed projection body and the \(L_{p}\)-dual mixed volume, it yields

$$\begin{aligned}&\left( \frac{2}{{\widetilde{V}}_{r}(K,Q)}\right) ^{\frac{1}{p}}d_{E} \min \limits _{S^{n-1}}h_{\Pi _{p,q}(K,Q)}(v_{E})\nonumber \\&\quad \le \left( \frac{2}{{\widetilde{V}}_{r}(K,Q)}\right) ^{\frac{1}{p}} d_{E}h_{\Pi _{p,q}(K,Q)}(v_{E})\nonumber \\&\quad =\left( \frac{1}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} (d_{E}|u\cdot v_{E}|)^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u) \right) ^{\frac{1}{p}}\nonumber \\&\quad \le \left( \frac{1}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} h_{E}^{p}(u){\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)\right) ^{\frac{1}{p}}\nonumber \\&\quad =\overline{{\widetilde{V}}}_{p,q}(K,E,Q). \end{aligned}$$
(4.1)

Let \({\mathcal {E}}_{p,q}=\left\{ E\in {\mathcal {E}}^{n}:\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\le 1\right\} \). Then, the above inequality yields that

$$\begin{aligned} d_{E}\le & {} \left( \frac{{\widetilde{V}}_{r}(K,Q)}{2}\right) ^{\frac{1}{p}} \frac{\overline{{\widetilde{V}}}_{p,q}(K,E,Q)}{\min \limits _{S^{n-1}}h_{\Pi _{p,q}(K,Q)}}\nonumber \\\le & {} \left( \frac{{\widetilde{V}}_{r}(K,Q)}{2}\right) ^{\frac{1}{p}} \frac{1}{\min \limits _{S^{n-1}}h_{\Pi _{p,q}(K,Q)}},\ \ \text{ for } \text{ all }\ E\in {\mathcal {E}}_{p,q}. \end{aligned}$$
(4.2)

Thus, the set \({\mathcal {E}}_{p,q}\) is bounded in the metric space \(({\mathcal {E}}^{n},\delta _{H})\). Using Theorem 3.1, the functional \(\overline{{\widetilde{V}}}_{p,q}(K,\cdot ,Q)\) is continuous, then \({\mathcal {E}}_{p,q}\) is also closed. According to the Blaschke selection theorem, each maximizing sequence of ellipsoids for Problem \(S_{p,q}\) has a convergent subsequence whose limit is still in \({\mathcal {E}}_{p,q}\). Therefore, a solution to Problem \(S_{p,q}\) exists.

We next prove the uniqueness by contradiction. We assume that the ellipsoids \(E_{1}\) and \(E_{2}\) are two different solutions to Problem \(S_{p,q}\). Let \(E_{1}=T_{1}B\) and \(E_{2}=T_{2}B\), where \(T_{1},T_{2}\in {\mathrm{GL}}(n)\). Then \(\det (T_{1})=\det (T_{2})\) and \(\overline{{\widetilde{V}}}_{p,q}(K,E_{i},Q)\le 1\), for \(i=1,2\).

Since each symmetric matrices \(T\in {\mathrm{GL}}(n)\) could be represented in the form \(T=PQ\), where P is symmetric, positive definite and Q is orthogonal. Then we may assume that \(T_{1}\) and \(T_{2}\) are symmetric and positive definite. Then \(T_{1}\ne \lambda T_{2}\), for all \(\lambda >0\). The Minkowski inequality for positive definite matrices implies

$$\begin{aligned} \det \left( \frac{1}{2}T_{1}+\frac{1}{2}T_{2}\right) ^{\frac{1}{n}} >\frac{1}{2}\det (T_{1})^{\frac{1}{n}}+\frac{1}{2}\det (T_{2})^{\frac{1}{n}}. \end{aligned}$$

Let \(E_{3}=\frac{1}{2}(T_{1}+T_{2})B\). Then we have

$$\begin{aligned} |E_{3}|>|E_{1}|=|E_{2}|. \end{aligned}$$
(4.3)

From (2.2) and the triangle inequality, one has for all \(u\in S^{n-1}\),

$$\begin{aligned} h_{E_{3}}(u)=\left| \frac{T^{t}_{1}+T^{t}_{2}}{2}u\right| \le \frac{|T^{t}_{1}u|+|T^{t}_{2}u|}{2}=\frac{h_{E_{1}}(u)+h_{E_{2}}(u)}{2}. \end{aligned}$$
(4.4)

Now, from Definition (1.13), the monotonicity of \(f(t)=t^{p},~p\ge 1\), (4.4), and the convexity of \(f(t)=t^{p}\), it follows that

$$\begin{aligned}&\overline{{\widetilde{V}}}_{p,q}(K,E_{3},Q)^{p} \\&\quad =\int _{S^{n-1}} \left( \frac{h_{E_{3}}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{p}{\mathrm{d}}{\widetilde{V}}_{r}(K,Q; u)\\&\quad \le \int _{S^{n-1}}\left( \frac{h_{E_{1}}(\alpha _{K}(u)) +h_{E_{2}}(\alpha _{K}(u))}{2h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{p}{\mathrm{d}}{\widetilde{V}}_{r}(K,Q; u)\\&\quad \le \int _{S^{n-1}}\left[ \frac{1}{2} \left( \frac{h_{E_{1}}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))}\right) ^{p} \left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{p} +\frac{1}{2}\left( \frac{h_{E_{2}}(\alpha _{K}(u))}{h_{K}(\alpha _{K}(u))} \right) ^{p}\left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{p}\right] \\&\qquad {\mathrm{d}}{\widetilde{V}}_{r}(K,Q; u)\\&\quad =\frac{1}{2}\overline{{\widetilde{V}}}_{p,q}(K,E_{1},Q)^{p} +\frac{1}{2}\overline{{\widetilde{V}}}_{p,q}(K,E_{2},Q)^{p} \le 1. \end{aligned}$$

Then \(E_{3}\in {\mathcal {E}}_{p,q}\). That is, \(E_{3}\) satisfies the constraint \(\overline{{\widetilde{V}}}_{p,q}(K,E_{3},Q)\le 1\). Then, it will result in \(|E_{3}|\le |E_{1}|=|E_{2}|\), which contradicts (4.3). \(\square \)

Our main problems \(S_{p,q}\) and \({\overline{S}}_{p,q}\) are two equivalent description. The solutions to \(S_{p,q}\) and \({\bar{S}}_{p,q}\) differ by only a scale factor. To prove this conclusion, we need the next lemma.

Lemma 4.2

Let \(p,q>0, K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\). Then

$$\begin{aligned} \max \limits _{\{E\in {\mathcal {E}}^{n}:\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\le 1\}}|E| =\max \limits _{\{E\in {\mathcal {E}}^{n}:\overline{{\widetilde{V}}}_{p,q}(K,E,Q)= 1\}}|E|; \end{aligned}$$
(4.5)

and

$$\begin{aligned} \min \limits _{\{E\in {\mathcal {E}}^{n}:|E|\ge \omega _n\}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q) =\min \limits _{\{E\in {\mathcal {E}}^{n}:|E|= \omega _n\}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q). \end{aligned}$$
(4.6)

Proof

We first prove that the ellipsoid \(E_1\) with \(\overline{{\widetilde{V}}}_{p,q}(K,E_{1},Q)<1\) cannot be the maximizer of \(\max \nolimits _{\{E\in {\mathcal {E}}^{n}:\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\le 1\}}|E|\). In fact, for the ellipsoid \(\overline{{\widetilde{V}}}_{p,q}(K,E_{1},Q)^{-1}E_{1}\), its volume is greater than the volume of \(E_1\), i.e.,

$$\begin{aligned} \left| \overline{{\widetilde{V}}}_{p,q}(K,E_{1},Q)^{-1}E_{1}\right| >|E_{1}|. \end{aligned}$$

And one has from (2.16),

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}\left( K,\overline{{\widetilde{V}}}_{p,q} (K,E_{1},Q)^{-1}E_{1},Q\right) =1, \end{aligned}$$

as required.

We next prove (4.6). For any ellipsoid \(E_2\) with \(|E_2|>\omega _n\), the ellipsoid \(\left( \frac{\omega _n}{|E_2|}\right) ^{\frac{1}{n}}E_2\) satisfies \(\left| \left( \frac{\omega _n}{|E_2|}\right) ^{\frac{1}{n}}E_2\right| =\omega _n\). And from (2.16), it follows that

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}\left( K,\left( \frac{\omega _n}{|E_2|}\right) ^{\frac{1}{n}}E_2,Q\right) =\left( \frac{\omega _n}{|E_2|}\right) ^{\frac{1}{n}}\overline{{\widetilde{V}}}_{p,q}(K,E_2,Q) <\overline{{\widetilde{V}}}_{p,q}(K,E_2,Q). \end{aligned}$$

\(\square \)

Theorem 4.3

Suppose \(p,q>0\) and K is an origin-symmetric convex body in \({\mathbb {R}}^{n}\), and Q is a star body in \({\mathbb {R}}^{n}\) about the origin.

\({\mathrm{(1)}}\) If \(E_{M}\) is an origin-symmetric ellipsoid that is a \(S_{p,q}\) solution for K and Q, then

$$\begin{aligned} \left( \frac{\omega _{n}}{|E_{M}|}\right) ^{\frac{1}{n}}E_{M} \end{aligned}$$
(4.7)

is a solution to Problem \({\bar{S}}_{p,q}\).

\({\mathrm{(2)}}\) If \(E_{m}\) is an origin-symmetric ellipsoid that is a \({\bar{S}}_{p,q}\) solution for K and Q, then

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(K,E_{m},Q)^{-1}E_{m} \end{aligned}$$
(4.8)

is a solution to Problem \(S_{p,q}\).

Proof

\({\mathrm{(1)}}\) Let \(E\in \{E\in {\mathcal {E}}^{n}: |E|\ge \omega _{n}\}\). It follows from (2.16) that

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}\left( K,\overline{{\widetilde{V}}}_{p,q}(K,E,Q)^{-1}E,Q\right) =1. \end{aligned}$$

Then, from the assumption that \(E_M\) is a \(S_{p,q}\) solution, it follows

$$\begin{aligned} |E_{M}|\ge \left| \overline{{\widetilde{V}}}_{p,q}(K,E,Q)^{-1}E\right| =\overline{{\widetilde{V}}}_{p,q}(K,E,Q)^{-n}\left| E\right| . \end{aligned}$$

Therefore,

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(K,E,Q)\ge \left( \frac{|E|}{|E_{M}|}\right) ^{\frac{1}{n}}\ge \left( \frac{\omega _{n}}{|E_{M}|}\right) ^{\frac{1}{n}} =\overline{{\widetilde{V}}}_{p,q}\left( K,\left( \frac{\omega _{n}}{|E_{M}|}\right) ^{\frac{1}{n}}E_{M},Q\right) , \end{aligned}$$

where the last equality uses the fact \(\overline{{\widetilde{V}}}_{p,q}(K,E_{M},Q)=1\) by (4.5). Added that \(\left( \frac{\omega _{n}}{|E_{M}|}\right) ^{\frac{1}{n}}E_{M}\in \{E\in {\mathcal {E}}^{n}: |E|\ge \omega _{n}\}\), it implies that the ellipsoid \(\left( \frac{\omega _{n}}{|E_{M}|}\right) ^{\frac{1}{n}}E_{M}\) is a solution to Problem \({\bar{S}}_{p,q}\).

\({\mathrm{(2)}}\) Let \(E\in \left\{ E\in {\mathcal {E}}^{n}:\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\le 1\right\} \). Since \(E_m\) is an \({\overline{S}}_{p,q}\) solution, and \(\left( \frac{\omega _{n}}{|E|}\right) ^{\frac{1}{n}}E\in \{E\in {\mathcal {E}}^{n}: |E|=\omega _{n}\},\) it follows from (2.16) that

$$\begin{aligned} \left( \frac{\omega _{n}}{|E|}\right) ^{\frac{1}{n}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q) =\overline{{\widetilde{V}}}_{p,q}\left( K,\left( \frac{\omega _{n}}{|E|}\right) ^{\frac{1}{n}}E,Q\right) \ge \overline{{\widetilde{V}}}_{p,q}(K,E_{m},Q). \end{aligned}$$

Using (4.6), we have \(|E_m|=\omega _n\). Then \(\overline{{\widetilde{V}}}_{p,q}(K,E_{m},Q)^{-1}|E_m|^{\frac{1}{n}} \ge \overline{{\widetilde{V}}}_{p,q}(K,E,Q)^{-1}|E|^{\frac{1}{n}}.\) Thus, it results in

$$\begin{aligned} \left( \frac{\left| \overline{{\widetilde{V}}}_{p,q}(K,E_{m},Q)^{-1}E_{m}\right| }{\omega _{n}}\right) ^{\frac{1}{n}}\ge \left( \frac{\left| \overline{{\widetilde{V}}}_{p,q}(K,E,Q)^{-1}E\right| }{\omega _{n}}\right) ^{\frac{1}{n}}\ge \left( \frac{|E|}{\omega _{n}}\right) ^{\frac{1}{n}}. \end{aligned}$$

Then the proof is completed by observing \(\overline{{\widetilde{V}}}_{p,q}\left( K,\overline{{\widetilde{V}}}_{p,q}(K,E_{m},Q)^{-1}E_{m},Q\right) =1\). \(\square \)

In Theorem 4.1, we proved the existence for all cases of \(0<p\le q\), and the uniqueness for the cases of \(1<p\le q\). In order to show the uniqueness of for all cases of \(0<p\le q\), we need the next lemma that shows that, without loss of generality, we may assume that the ellipsoid E is the unit ball B in \({\mathbb {R}}^{n}\).

Lemma 4.4

Suppose real \(p, q\ne 0, K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\). If \(\phi \in {\mathrm{GL}}(n)\), then

$$\begin{aligned}&{\widetilde{V}}_{p,q}(\phi ^{-1}K, B, \phi ^{-1}Q)|x|^{2}\nonumber \\&\quad =n\int _{S^{n-1}}|x \cdot v|^{2}{\mathrm{d}}{\widetilde{C}}_{p,q}(\phi ^{-1}K, \phi ^{-1}Q,v),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}, \end{aligned}$$
(4.9)

if and only if

$$\begin{aligned}&{\widetilde{V}}_{p,q}(K,\phi B, Q)h_{(\phi B)^{*}}^{2}(x)\nonumber \\&\quad =n\int _{S^{n-1}}|x \cdot v|^{2}h_{\phi B}^{p-2}(v){\mathrm{d}}{\widetilde{C}}_{p,q}(K, Q,v),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$
(4.10)

Proof

In light of Lemma 2.1, it suffices to prove the statement for \({\mathrm{SL}}(n)\). In terms of (2.2), (2.4) and Lemma 2.3, we have, for all \(x\in {\mathbb {R}}^{n}\),

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,\phi B, Q)h_{(\phi B)^{*}}^{2}(x)= & {} {\widetilde{V}}_{p,q}(K,\phi B, Q)h_{\phi ^{-t}B^{*}}^{2}(x)\\= & {} {\widetilde{V}}_{p,q}(\phi ^{-1}K,B, \phi ^{-1}Q)h_{B^{*}}^{2}(\phi ^{-1}x). \end{aligned}$$

Then, using Definition 2.5, (4.10) is equivalent to, for all \(x\in {\mathbb {R}}^{n}\),

$$\begin{aligned} {\widetilde{V}}_{p,q}(\phi ^{-1}K,B,\phi ^{-1}Q)h_{B^{*}}^{2}(x)= & {} n\int _{S^{n-1}}|\phi x\cdot v|^{2}h_{B}^{p-2}(\phi ^{t}v) {\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\\= & {} n\int _{S^{n-1}}|x\cdot \phi ^{t}v|^{2}|\phi ^{t}v|^{p-2} {\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\\= & {} n\int _{S^{n-1}}|x\cdot \langle \phi ^{t}v\rangle |^{2} |\phi ^{t}v|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\\= & {} n\int _{S^{n-1}}|x\cdot v|^{2}{\mathrm{d}}\phi ^{-t}_{p} \dashv {\widetilde{C}}_{p,q}(K,Q,v), \end{aligned}$$

which by Lemma 2.6 is in turn equivalent to

$$\begin{aligned}&{\widetilde{V}}_{p,q}(\phi ^{-1}K, B, \phi ^{-1}Q)|x|^{2}\\&\quad =n\int _{S^{n-1}}|x \cdot v|^{2}{\mathrm{d}}{\widetilde{C}}_{p,q}(\phi ^{-1}K, \phi ^{-1}Q,v),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$

\(\square \)

Now we show the existence and uniqueness of solution \(S_{p,q}\) and \({\bar{S}}_{p,q}\) for all cases \(0<p\le q\).

Theorem 4.5

Suppose that \(0<p\le q=p+r, r\in [0,\infty ), K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\). Then \(S_{p,q}\) as well as \({\bar{S}}_{p,q}\) has a unique solution. Moreover, an ellipsoid \(E\in {\mathcal {E}}^{n}\) solves \({\bar{S}}_{p,q}\) if and only if it satisfies

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,E,Q)h^{2}_{E^{*}}(x)=n\int _{S^{n-1}}|x \cdot u|^{2}h^{p-2}_{E}(u){\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n},\nonumber \\ \end{aligned}$$
(4.11)

and an ellipsoid \(E\in {\mathcal {E}}^{n}\) solves \(S_{p,q}\) if and only if it satisfies

$$\begin{aligned} {\widetilde{V}}_{r}(K,Q)h^{2}_{E^{*}}(x)=n\int _{S^{n-1}}|x \cdot u|^{2}h^{p-2}_{E}(u){\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$
(4.12)

Proof

We first show that an ellipsoid \(E\in {\mathcal {E}}^{n}\) solves \({\bar{S}}_{p,q}\) if and only if it satisfies (4.11). Without loss of generality, we may assume \(E=B\) by using Lemma 4.4. Namely, we will show that B is a \({\bar{S}}_{p,q}\) solution for K and Q if and only if

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,B, Q)|x|^{2}=n\int _{S^{n-1}}|x\cdot u|^{2}{\mathrm{d}}{\widetilde{C}}_{p,q}(K, Q,u),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$
(4.13)

Firstly, we show if \(B\in {\mathcal {E}}^{n}\) solves \({\bar{S}}_{p,q}\), then (4.13) holds. Indeed, suppose that \(T\in {\mathrm{SL}}(n)\). Choose \(\varepsilon _{0}>0\) sufficiently small so that for all \(\varepsilon \in (-\varepsilon _{0},\varepsilon _{0})\), \(I_n+\epsilon T\) is invertible, where \(I_n\) is identity matrix. For \(\varepsilon \in (-\varepsilon _{0},\varepsilon _{0})\), define \(T_{\varepsilon }\in {\mathrm{SL}}(n)\) by

$$\begin{aligned} T_{\varepsilon }=|I_{n}+\varepsilon T|^{-\frac{1}{n}}(I_{n}+\varepsilon T). \end{aligned}$$

Since \(|T_{\varepsilon }|=1\), the ellipsoid \(E_{\varepsilon }=T^{t}_{\varepsilon }B\) clearly has volume \(\omega _{n}\). The support function of \(E_{\varepsilon }\) is given by

$$\begin{aligned} h_{E_{\varepsilon }}(u)=h_{T^{t}_{\varepsilon }B}(u)=|T_{\varepsilon }u|. \end{aligned}$$

Since \(E_0=B\) is a \({\bar{S}}_{p,q}\) solution, we have

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,E_{0},Q)\le {\widetilde{V}}_{p,q}(K,E_{\varepsilon },Q),\ \ \text{ for } \text{ all }\ \varepsilon , \end{aligned}$$

and hence using (1.9), it is equivalent to

$$\begin{aligned} \frac{\mathrm{d}}{{\mathrm{d}}\varepsilon }\Big |_{\varepsilon =0}\int _{S^{n-1}}|T_{\varepsilon }u|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)=0. \end{aligned}$$
(4.14)

Note that

$$\begin{aligned} |I_{n}+\varepsilon T|^{\frac{1}{n}}=1+\frac{\varepsilon }{n}{\mathrm{tr}}T+O(\varepsilon ^{2}) \end{aligned}$$

and

$$\begin{aligned} |u+\varepsilon Tu|=[1+2\epsilon \cdot Tu+\epsilon ^2(Tu\cdot Tu)]^{\frac{1}{2}} =1+\varepsilon (u\cdot T u)+O(\varepsilon ^{2}), \end{aligned}$$

then (4.14) implies

$$\begin{aligned}&\frac{\mathrm{d}}{{\mathrm{d}}\varepsilon }\Big |_{\varepsilon =0}\int _{S^{n-1}}\left( \frac{1+\varepsilon ( u\cdot T u)+O(\varepsilon ^{2})}{1+\frac{\varepsilon }{n}{\mathrm{tr}}T+O(\varepsilon ^{2})}\right) ^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)\nonumber \\&\quad =p\int _{S^{n-1}}\left( u \cdot T u-\frac{1}{n}{\mathrm{tr}}T\right) {\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)\nonumber \\&\quad =0. \end{aligned}$$
(4.15)

Let \(T=x\otimes x\) for nonzero \(x\in {\mathbb {R}}^{n}\), where the notation \(x\otimes x\) represents the rank 1 linear operator on \({\mathbb {R}}^{n}\) that takes y to \((x\cdot y)x\). It immediately gives that \({\mathrm{tr}}(x\otimes x)=|x|^{2}\). Using the facts \({\mathrm{tr}}(x\otimes x)=|x|^{2}\) and \(u\cdot (x\otimes x)u=(u\cdot x)^{2}\), (4.15) is

$$\begin{aligned} \int _{S^{n-1}}|u\cdot x|^{2}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)=\frac{{\widetilde{V}}_{p,q}(K,B,Q)}{n}|x|^{2},\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$

Secondly, we show if

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,B, Q)|x|^{2}=n\int _{S^{n-1}}|x\cdot u|^{2} {\mathrm{d}}{\widetilde{C}}_{p,q}( K,Q,u),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}, \end{aligned}$$
(4.16)

then B is a solution to Problem \({\bar{S}}_{p,q}\). Moreover, B is a unique \({\bar{S}}_{p,q}\) solution.

To prove that B is a \({\bar{S}}_{p,q}\) solution for KQ, we show that for any ellipsoid E with \(|E|=\omega _n\), one has

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,E,Q)\ge {\widetilde{V}}_{p,q}(K,B,Q), \end{aligned}$$
(4.17)

with equality if and only if \(E=B\). It is equivalent to show that for any ellipsoid E with \(E=P^t B\), \(P\in {\mathrm{SL}}(n)\), one has

$$\begin{aligned} \left( \frac{1}{{\widetilde{V}}_{p,q}(K,B,Q)} \int _{S^{n-1}}|Pu|^p{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u) \right) ^{\frac{1}{p}}\ge 1, \end{aligned}$$
(4.18)

with equality if and only if \(Pu=1\) for all \(u\in S^{n-1}\). From Jensen’s inequality,

$$\begin{aligned}&\left( \frac{1}{{\widetilde{V}}_{p,q}(K,B,Q)} \int _{S^{n-1}}|Pu|^p{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u) \right) ^{\frac{1}{p}}\\&\quad \ge \exp \left( \frac{1}{{\widetilde{V}}_{p,q}(K,B,Q)} \int _{S^{n-1}}\log |Pu|{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u) \right) , \end{aligned}$$

with equality if and only if there exists \(c>0\) such that \(|Pu|=c\) for all \(u\in {\mathrm{supp}} {\widetilde{C}}_{p,q}(K,Q,\cdot )\). Hence, we need show

$$\begin{aligned} \int _{S^{n-1}}\log |Pu|{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)\ge 0, \end{aligned}$$
(4.19)

We write P as \(P=O^tDO\), where \(D={\mathrm{diag}}(\lambda _1,\lambda _2,\cdots ,\lambda _n)\) is a diagonal matrix with eigenvalues \(\lambda _1,\lambda _2,\cdots ,\lambda _n\), and O is orthogonal.

From Definition 2.5 and Lemma 2.6, it follows that

$$\begin{aligned} \int _{S^{n-1}}\log |Pu|{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)&= \int _{S^{n-1}}|Ou|^p\log |O^tDOu|{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)\\&=\int _{S^{n-1}}\log |O^tDv|{\mathrm{d}}O^t_p\dashv {\widetilde{C}}_{p,q}(K,Q,v)\\&=\int _{S^{n-1}}\log |Dv|{\mathrm{d}}{\widetilde{C}}_{p,q}(OK,OQ,v). \end{aligned}$$

Then by the concavity of the \(\log \) function and (4.16),

$$\begin{aligned} \int _{S^{n-1}}\log |Pu|{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)&= \frac{1}{2}\int _{S^{n-1}}\log \left( \sum \limits _{i=1}^n\lambda _i^2u_i^2\right) {\mathrm{d}}{\widetilde{C}}_{p,q}(OK,OQ,v)\\&\ge \sum \limits _{i=1}^n\int _{S^{n-1}}u_i^2\log (\lambda _i) {\mathrm{d}}{\widetilde{C}}_{p,q}(O K,O Q,v)\\&= \sum \limits _{i=1}^n\log (\lambda _i) \int _{S^{n-1}}|u\cdot e_i|^2{\mathrm{d}}{\widetilde{C}}_{p,q}(O K,O Q,v)\\&= \frac{1}{n}{\widetilde{V}}_{p,q}(K,B, Q)\sum \limits _{i=1}^n\log (\lambda _i), \end{aligned}$$

where \(u_i\) denotes \(u\cdot e_i\) for \(i=1,\cdots ,n\). Since \(|D|=1\), we have \(\sum \limits _{i=1}^n\log (\lambda _i)=\log (\prod \limits _{i=1}^n\lambda _i)=0\) Thus (4.19) holds. And then we have (4.16), namely B is a solution to Problem \({\bar{S}}_{p,q}\).

For the uniqueness of Problem \({\bar{S}}_{p,q}\), we only need consider the equality condition. Note that the strict concavity of \(\log \) function implies that equality in (4.16) holds only if \(u_{i_1},\cdots ,u_{i_N}\ne 0\) implies \(\lambda _{i_1}=\cdots =\lambda _{i_N}\), for \(u\in {\mathrm{supp}}{\widetilde{C}}_{p,q}(O K,O Q,\cdot )\). Thus \(|Du|=\lambda _i\) when \(u_i\ne 0\) for \(u\in {\mathrm{supp}}{\widetilde{C}}_{p,q}(O K,O Q,\cdot )\). Equality in (4.18) forces \(|Pu|=c\) for all \(u\in {\mathrm{supp}}{\widetilde{C}}_{p,q}(O K,O Q,\cdot )\). Since \({\mathrm{supp}}{\widetilde{C}}_{p,q}(O K,O Q,\cdot )\) is not contained in an \((n-1)\)-dimensional subspace of \({\mathbb {R}}^n\), we have \(\lambda _i=c\) for all i. Combining with \(|D|=\lambda _1\cdots \lambda _n=1\), we have \(\lambda _i=1\) for all i. Thus \(D=I_n\), and \(P=I_n\).

Note that Theorems 4.1 and 4.3 get the existence of the solution to Problems \(S_{p,q}\) and \({\bar{S}}_{p,q}\). And their uniqueness is proved from the above proof and Theorem 4.3.

Finally, we let the ellipsoid \(E\in {\mathcal {E}}^n\) solve Problem \(S_{p,q}\). Using Theorem 4.3, it is equivalent to that \(c_0E\) is a solution to Problem \({\bar{S}}_{p,q}\), where \(c_0=(\frac{\omega _n}{|E|})^{\frac{1}{n}}\). It holds if and only if (4.11) holds, i.e.,

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,E,Q)h^{2}_{E^{*}}(x)=n\int _{S^{n-1}}|x \cdot u|^{2}h^{p-2}_{E}(u){\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u),\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$

This completes the result by noticing that \(\overline{{\widetilde{V}}}_{p,q}(K,E,Q)= \left( \frac{{\widetilde{V}}_{p,q}(K,E,Q)}{{\widetilde{V}}_{r}(K,Q)}\right) ^{\frac{1}{p}} =1\) from Lemma 4.2. \(\square \)

Let \(0<p\le q\le \infty \). Theorem 4.5 shows that problem (\(S_{p,q}\)) has a unique solution. In the case \(Q=K\), the \(S_{p,q}\) problem had been considered by Lutwak, Yang and Zhang in [30].

In the case \(p=\infty \), with the aid of (1.16), we may rephrase \((S_{\infty ,q})\) as: Among all origin-symmetric ellipsoids, find an ellipsoid which solves the following constrained maximization problem:

$$\begin{aligned} \max \left( \frac{|E|}{\omega _{n}}\right) ^{\frac{1}{n}}\ \ \text{ subject } \text{ to }\ \ E\subseteq \left( \frac{\rho _{Q}}{\rho _{K}}\right) K. \end{aligned}$$
($S_{\infty ,\infty }$)

When \(Q=K\), the problem is the classical John-ellipsoid problem (see, e.g., Giannopoulos and Milman [12]).

In light of Theorem 4.1, Theorem 4.3 and Theorem 4.5, we introduce a family of ellipsoids, which is an extension of LYZ’s \(L_{p}\) John ellipsoids.

Definition 4.6

Let \(0<p\le q=p+r\le \infty , r\in [0,\infty )\). Suppose K is a convex body in \({\mathbb {R}}^{n}\) that contains the origin in its interior and Q is a star body (about the origin) in \({\mathbb {R}}^{n}\). Among all origin-symmetric ellipsoids, the unique ellipsoid that solves the constrained maximization problem

$$\begin{aligned} \max \limits _{E\in {\mathcal {E}}^{n}}|E|\ \ \ \ \text{ subject } \text{ to }\ \ \ \overline{{\widetilde{V}}}_{p,q}(K,E, Q)\le 1 \end{aligned}$$

will be called the (pq)-John ellipsoid of K and Q, and will be denoted by \(E_{p,q}(K,Q)\).

Among all origin-symmetric ellipsoids, the unique ellipsoid that solves the constrained minimization problem

$$\begin{aligned} \min \limits _{E\in {\mathcal {E}}^{n}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\ \ \ \ \text{ subject } \text{ to }\ \ \ |E|=\omega _{n} \end{aligned}$$

will be called the normalized (pq)-John ellipsoid of K and Q, and will be denoted by \({\bar{E}}_{p,q}(K,L)\).

Note that in the case \(Q=K\), \(E_{p,q}(K,K)=E_{p}(K)\) is the \(L_p\)-John ellipsoid. In the case that \(q=n\) and \(Q=B\), \(E_{p,n}(K,B)=E_{p}(K)\) is also the \(L_p\)-John ellipsoid. In the case that \(p=\infty \) and \(Q=K\), \(E_{\infty ,\infty }(K,K)=J(K)\) is also the classic John ellipsoid.

From Definition 4.6 and (2.20), we immediately obtain

Lemma 4.7

Suppose \(K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\), and \(0<p\le q\le \infty \). Then for \(\phi \in {\mathrm{GL}}(n)\),

$$\begin{aligned} E_{p,q}(\phi K,\phi Q)=\phi E_{p,q}(K,Q). \end{aligned}$$

From \(E_{p,q}(B,B)=E_pB=B\) and Lemma 4.7, we see that if \(E\in {\mathcal {E}}^{n}\), then

$$\begin{aligned} E_{p,q}(E,E)=E. \end{aligned}$$
(4.20)

Note that if the John point of K is at the origin (e.g., if K is origin-symmetric), then

$$\begin{aligned} E_{\infty ,\infty }(K,Q)\subseteq \left( \frac{\rho _{Q}}{\rho _{K}}\right) K. \end{aligned}$$

From (2.24), (4.12) of Theorem 4.5, we immediately obtain

Lemma 4.8

Suppose \(K\in {\mathcal {K}}^{n}_{o}, Q\in {\mathcal {S}}^{n}_{o}\) and \(2\le q\le \infty \). Then

$$\begin{aligned} E_{2,q}(K,Q)=\Gamma _{-2,-q}(K,Q). \end{aligned}$$

A finite positive Borel measure \(\mu \) on \(S^{n-1}\) is said to be isotropic if (see [12])

$$\begin{aligned} \int _{S^{n-1}}|u\cdot v|^{2}{\mathrm{d}}\mu (u)=\frac{|\mu |}{n}, \end{aligned}$$

for all \(v\in S^{n-1}\), where \(|\mu |\) denotes the total mass of \(\mu \). For nonzero \(x\in {\mathbb {R}}^{n}\), the notation \(x\otimes x\) represents the rank 1 linear operator on \({\mathbb {R}}^{n}\) that takes y to \((x\cdot y)x\). It immediately gives that \({\mathrm{tr}}x\otimes x=|x|^{2}\). Equivalently, \(\mu \) is isotropic if

$$\begin{aligned} \int _{S^{n-1}}u\otimes u{\mathrm{d}}\mu (u)=\frac{|\mu |}{n}I_{n}. \end{aligned}$$

From definition (1.6) and (1.9), we see that

$$\begin{aligned} {\widetilde{V}}_{p,q}(K,B,Q)= & {} \int _{S^{n-1}}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)\\= & {} \frac{1}{n}\int _{\varvec{\alpha }_{K}^{*}(S^{n-1})}h_{K} (\alpha _{K}(u))^{-p}\rho _{K}^{q}(u)\rho _{Q}^{n-q}(u){\mathrm{d}}u ={\widetilde{C}}_{p,q}(K,Q,S^{n-1}). \end{aligned}$$

Therefore, the condition (4.11) is equivalent to

$$\begin{aligned} \int _{S^{n-1}}|x\cdot u|^{2}{\mathrm{d}}{\widetilde{C}}_{p,q}(K, Q,u)=\frac{{\widetilde{C}}_{p,q}(K,Q,S^{n-1})}{n}|x|^{2},\ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$

Then an immediate consequence of Theorem 4.5 is

Corollary 4.9

Suppose \(K\in {\mathcal {K}}^{n}_{o}\) with \(Q\in {\mathcal {S}}^{n}_{o}\), and \(0<p\le q\in (0,\infty ]\). Then there exists a unique solution to the following constrained minimization problem:

$$\begin{aligned} \min \{{\widetilde{V}}_{p,q}(K,TB,Q): T\in {\mathrm{SL}}(n)\}. \end{aligned}$$

Moreover, the identity operator \(I_{n}\) is the solution if and only if \(L_{p}\) dual curvature measures \({\widetilde{C}}_{p,q}(K,Q,\cdot )\) are isotropic on \(S^{n-1}\).

Corollary 4.10

Suppose \(K\in {\mathcal {K}}^{n}_{o}\) with \(Q\in {\mathcal {S}}^{n}_{o}\), and \(0<p\le q\in (0,\infty ]\).

  1. (1)

    There exists an \({\mathrm{SL}}(n)\) transformation T, such that \({\widetilde{C}}_{p,q}(TK,TQ,\cdot )\) is isotropic on \(S^{n-1}\).

  2. (2)

    If \(T_{1},T_{2}\in {\mathrm{SL}}(n)\) such that \({\widetilde{C}}_{p,q}(T_{1}K,T_{1}Q,\cdot )\), \({\widetilde{C}}_{p,q}(T_{2}K,T_{2}Q,\cdot )\) are both isotropic on \(S^{n-1}\), then there exists an orthogonal \(O\in {\mathrm{O}}(n)\) such that \(T_{2}=O T_{1}\).

5 Continuity of (pq)-John Ellipsoids

In this section, we show that the family of (pq)-John ellipsoids associated with a convex body and a star body in \({\mathbb {R}}^{n}\) is continuous in \(p\in (0,\infty ]\).

We assume that \(K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\) are two fixed bodies in this section.

Lemma 5.1

Suppose \(0<p\le q\le \infty \). If \(aB\subseteq K\subseteq bB\) and \(aB\subseteq Q\subseteq bB\) for \(a,b>0\), then

$$\begin{aligned} {\bar{E}}_{p,q}(K,Q)\subseteq \left( \frac{b}{a}\right) ^{\frac{p+2q+n}{p}}(c_{n-2,p})^{-\frac{1}{p}}B, \end{aligned}$$

where

$$\begin{aligned} c_{n-2,p}=\frac{(n+p)\omega _{n+p}}{n\omega _{2}\omega _{n}\omega _{p-1}},\ \ \ \omega _{n}=\frac{\pi ^{\frac{n}{2}}}{\Gamma \left( 1+\frac{n}{2}\right) }. \end{aligned}$$

Proof

From (4.1) and the definition of \({\bar{E}}_{p,q}(K,Q)\), we have

$$\begin{aligned} d_{E}\le \left( \frac{{\widetilde{V}}_{r}(K,Q)}{2}\right) ^{\frac{1}{p}} \frac{\overline{{\widetilde{V}}}_{p,q}(K,B,Q)}{h_{\Pi _{p,q}}(K,Q)}, \end{aligned}$$
(5.1)

Now, we estimate the value of \(\overline{{\widetilde{V}}}_{p,q}(K,B,Q)\). By the definition of \(\overline{{\widetilde{V}}}_{p,q}(K,L,Q)\), we have

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}(K,B,Q)= & {} \left( \frac{1}{{\widetilde{V}}_{r}(K,Q)} \int _{S^{n-1}}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\right) ^{\frac{1}{p}}\nonumber \\= & {} \left( \frac{1}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} \left( \frac{\rho _{K}(v)}{h_{K}(\alpha _{K}(v))\rho _{Q}(v)}\right) ^{p}{\mathrm{d}} {\widetilde{V}}_{r}(K,Q; v)\right) ^{\frac{1}{p}}\nonumber \\\le & {} \frac{b}{a^{2}}\left( \frac{1}{{\widetilde{V}}_{r}(K,Q)} \int _{S^{n-1}}{\mathrm{d}}{\widetilde{V}}_{r}(K,Q; v)\right) ^{\frac{1}{p}} =\frac{b}{a^{2}}. \end{aligned}$$
(5.2)

Note that

$$\begin{aligned} \int _{S^{n-1}}|u\cdot v|^{p}{\mathrm{d}}u=\frac{(n+p)\omega _{n+p}}{\omega _{2}\omega _{p-1}}. \end{aligned}$$
(5.3)

By the definition of (pq)-mixed projection body and (5.3), we have

$$\begin{aligned} h_{\Pi _{p,q}(K,Q)}(v_{E})= & {} \left( \frac{1}{2}\int _{S^{n-1}} |u\cdot v_{E}|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,u)\right) ^{\frac{1}{p}}\nonumber \\= & {} \left( \frac{1}{2n}\int _{S^{n-1}}|u\cdot v_{E}|^{p}h_{K} (\alpha _{K}(u))^{-p}\left( \frac{\rho _{K}(u)}{\rho _{Q}(u)}\right) ^{q} \rho _{Q}^{n}(u){\mathrm{d}}u\right) ^{\frac{1}{p}}\nonumber \\\ge & {} \left( \frac{a^{q+n}}{2nb^{p+q}}\int _{S^{n-1}}|u \cdot v_{E}|^{p}{\mathrm{d}}u\right) ^{\frac{1}{p}}\nonumber \\= & {} \left( \frac{(n+p)\omega _{n+p}a^{q+n}}{2n\omega _{2}\omega _{p-1}b^{p+q}}\right) ^{\frac{1}{p}}. \end{aligned}$$
(5.4)

Together with (5.1), (5.2) and (5.4), and note that \({\widetilde{V}}_{r}(K,Q)\le \frac{\omega _{n}b^{n+r}}{a^{r}}\), we have

$$\begin{aligned} d_{{\bar{E}}_{p,q}(K,Q)}\le \left( \frac{b}{a}\right) ^{\frac{p+2q+n}{p}}(c_{n-2,p})^{-\frac{1}{p}}. \end{aligned}$$

Therefore,

$$\begin{aligned} {\bar{E}}_{p,q}(K,Q)\subseteq \left( \frac{b}{a}\right) ^{\frac{p+2q+n}{p}}(c_{n-2,p})^{-\frac{1}{p}}B. \end{aligned}$$

Note that \(\lim \limits _{p\rightarrow \infty }(c_{n-2,p})^{\frac{1}{p}}=1\), then \({\bar{E}}_{\infty ,\infty }(K,L)\subseteq \frac{b}{a}B\). \(\square \)

From Definition 4.6, we recall that for each \(p\in (0,\infty ]\) and \(q=p+r, r\in (0,\infty )\), the ellipsoid \({\bar{E}}_{p,q}(K,Q)\) is the unique ellipsoid that satisfies

$$\begin{aligned} \overline{{\widetilde{V}}}_{p,q}\left( K,{\bar{E}}_{p,q}(K,Q),Q\right) =\min \limits _{|E|=\omega _{n}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q). \end{aligned}$$
(5.5)

Lemma 5.2

If \(p,p_{0}\in (0,\infty ], q=p+r, r\in (0,\infty ),p\rightarrow p_{0}, q\rightarrow p_{0}+r=q_{0}, K\in {\mathcal {K}}^{n}_{o}\), and \(Q\in {\mathcal {S}}^{n}_{o}\), then

$$\begin{aligned} \lim \limits _{p\rightarrow p_{0}}\overline{{\widetilde{V}}}_{p,q}(K,{\bar{E}}_{p,q}(K,Q),Q) =\overline{{\widetilde{V}}}_{p_{0},q_{0}}(K,{\bar{E}}_{p_{0},q_{0}}(K,Q),Q). \end{aligned}$$

Proof

Using the Definition \({\bar{E}}_{p,q}(K,Q)\), Theorem 3.1, (5.5), and again the definition of \({\bar{E}}_{p,q}(K,Q)\), we have

$$\begin{aligned} \lim \limits _{p\rightarrow p_{0}}\overline{{\widetilde{V}}}_{p,q} (K,{\bar{E}}_{p,q}(K,Q),Q)= & {} \lim \limits _{p\rightarrow p_{0}}\min \limits _{|E| =\omega _{n}}\overline{{\widetilde{V}}}_{p,q}(K,E,Q)\\= & {} \min \limits _{|E|=\omega _{n}}\overline{{\widetilde{V}}}_{p_{0},q_{0}}(K,E,Q)\\= & {} \overline{{\widetilde{V}}}_{p_{0},q_{0}}(K,{\bar{E}}_{p_{0},q_{0}}(K,Q),Q). \end{aligned}$$

\(\square \)

Lemma 5.3

Suppose that \(p,p_{0}\in (0,\infty ], q=p+r, r\in (0,\infty ),p\rightarrow p_{0}, q\rightarrow p_{0}+r=q_{0}\), and \(K\in {\mathcal {K}}^{n}_{o}, Q\in {\mathcal {S}}^{n}_{o}\). If \(aB\subseteq Q\subseteq K\subseteq bB\) or \(aB\subseteq K\subseteq Q\subseteq bB\), for \(a,b>0\), then

$$\begin{aligned} \lim \limits _{p\rightarrow p_{0}}{\bar{E}}_{p,q}(K,Q)={\bar{E}}_{p_{0},q_{0}}(K,Q). \end{aligned}$$

Proof

We argue by contradiction and assume the conclusion to be false. Lemma 5.1, the Blaschke selection theorem, and our assumption, give a sequence \(p_{i}\rightarrow p_{0}\), as \(i\rightarrow \infty \), such that \(\lim \limits _{i\rightarrow \infty }{\bar{E}}_{p_{i},q_{i}}(K,Q)=E'\ne {\bar{E}}_{p_{0},q_{0}}(K,Q)\). Since the solution to Problem \(({\bar{S}}_{p,q})\) is unique, and by the uniform convergence established in Theorem 3.1, we get

$$\begin{aligned} \overline{{\widetilde{V}}}_{p_{0},q_{0}}\left( K,{\bar{E}}_{p_{0},q_{0}}(K,Q),Q \right)< & {} \overline{{\widetilde{V}}}_{p_{0},q_{0}}(K,E',Q)\\= & {} \lim _{i\rightarrow \infty }\overline{{\widetilde{V}}}_{p_{0},q_{0}} \left( K,{\bar{E}}_{p_{i},q_{i}}(K,Q),Q\right) \\= & {} \lim _{i\rightarrow \infty }\overline{{\widetilde{V}}}_{p_{i},q_{i}} \left( K,{\bar{E}}_{p_{i},q_{i}}(K,Q),Q\right) . \end{aligned}$$

This contradicts to Lemma 5.2. \(\square \)

Since, by Theorem 4.3, \(E_{p,q}(K,Q)=\overline{{\widetilde{V}}}_{p,q}\left( K,{\bar{E}}_{p,q}(K,Q),Q\right) ^{-1}{\bar{E}}_{p,q}(K,Q)\), the above gives

Theorem 5.4

If \(p,p_{0}\in (0,\infty ], q=p+r, r\in (0,\infty ),p\rightarrow p_{0}, q\rightarrow p_{0}+r=q_{0}, K\in {\mathcal {K}}^{n}_{o}\) and \(Q\in {\mathcal {S}}^{n}_{o}\), then

$$\begin{aligned} \lim \limits _{p\rightarrow p_{0}}E_{p,q}(K,Q)=E_{p_{0},q_{0}}(K,Q). \end{aligned}$$

6 Generalizations of John’s Inclusion

John’s inclusion states that if K is an origin-symmetric convex body in \({\mathbb {R}}^{n}\), then

$$\begin{aligned} E_{\infty }K\subseteq K\subseteq \sqrt{n}E_{\infty }K. \end{aligned}$$
(6.1)

\(L_{p}\) version of John’s inclusion is (see [30]): If K is a convex body in \({\mathbb {R}}^{n}\) that contains the origin in its interior, then

$$\begin{aligned}&E_{p}K\supseteq \Gamma _{-p}K\supseteq n^{\frac{1}{2}-\frac{1}{p}}\ \ \ \text{ when }\ \ \ 0<p\le 2,\\&E_{p}K\subseteq \Gamma _{-p}K\subseteq n^{\frac{1}{2}-\frac{1}{p}}\ \ \ \text{ when }\ \ \ 2\le p\le \infty . \end{aligned}$$

In this section, we shall prove a (pq)-version of John’s inclusion.

From (1.4), (2.1), (2.5) and Definition (2.26), we see immediately that if \(\lambda >0\), then

$$\begin{aligned} \Gamma _{-p,-q}(\lambda K,\lambda Q)=\lambda \Gamma _{-p,-q}(K,Q). \end{aligned}$$
(6.2)

Lemma 6.1

If \(p\in (0,\infty ], q=p+r, r\in [0,\infty )\) and \(K\in {\mathcal {K}}^{n}_{o}\), as well as \(Q\in {\mathcal {S}}^{n}_{o}\), then for \(\phi \in {\mathrm{GL}}(n)\)

$$\begin{aligned} \Gamma _{-p,-q}(\phi K,\phi Q)=\phi \Gamma _{-p,-q}(K,Q). \end{aligned}$$

Proof

From (6.2) it is sufficient to prove the formula when \(\phi \in {\mathrm{SL}}(n)\). For real \(p>0\), it follows from Definition (2.24), Lemma 2.6, Definition 2.5, Definition (2.24) again, and (2.6) that for \(u\in S^{n-1}\),

$$\begin{aligned} \rho _{\Gamma _{-p,-q}(\phi K,\phi Q)}(u)^{-p}= & {} \frac{n}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} |u\cdot v|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(\phi K,\phi Q,v)\\= & {} \frac{n}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} |u\cdot v|^{p}{\mathrm{d}}\phi ^{t}_{p}\dashv {\widetilde{C}}_{p,q}(K,Q,v)\\= & {} \frac{n}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} |u\cdot \langle \phi ^{-t}v\rangle |^{p}|\phi ^{-t}v|^{p}{\mathrm{d}} {\widetilde{C}}_{p,q}(K,Q,v)\\= & {} \frac{n}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} |u\cdot \phi ^{-t}v|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\\= & {} \frac{n}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} |\phi ^{-1}u\cdot v|^{p}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v)\\= & {} \rho _{\Gamma _{-p,-q}(K,Q)}(\phi ^{-1}u)^{-p}. \end{aligned}$$

The \(p=\infty \) case is now a direct consequence of the real case and Definition (2.25). \(\square \)

Lemma 6.2

If \(K\in {\mathcal {K}}^{n}_{o}, Q\in {\mathcal {S}}^{n}_{o}, p\in (0,\infty ]\) and \(q=p+r, r\in [0,\infty )\), then

$$\begin{aligned}&E_{p,q}(K,Q)\supseteq \Gamma _{-p,-q}(K,Q)\ \ \ \text{ when }\ \ \ 0<p<2,\\&E_{p,q}(K,Q)\subseteq \Gamma _{-p,-q}(K,Q)\ \ \ \text{ when }\ \ \ 2\le p\le \infty . \end{aligned}$$

Proof

Lemmas 4.7 and 6.1 show that it suffices to prove the inclusions when \(E_{p,q}(K,Q)=B\). For \(0<p<2\), Definition (2.24) and Theorem 4.5 show that for each \(u\in S^{n-1}\),

$$\begin{aligned} \rho _{\Gamma _{-p,-q}(K,Q)}(u)^{-p}= & {} \frac{n}{{\widetilde{V}}_{r}(K,Q)} \int _{S^{n-1}}|u\cdot v|^{p}{\mathrm{d}}{\tilde{C}}_{p,q}(K,Q,v)\\\ge & {} \frac{n}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}}|u \cdot v|^{2}{\mathrm{d}}{\tilde{C}}_{p,q}(K,Q,v)\\= & {} 1. \end{aligned}$$

This gives \(\Gamma _{-p,-q}(K,Q)\subseteq B=E_{p,q}(K,Q)\) when \(0<p<2\).

When \(2\le p<\infty \), the inequality is reversed. Thus \(E_{p,q}(K,Q)\subseteq \Gamma _{-p,-q}(K,Q)\) for \(2\le p<\infty \). The case \(p=\infty \) follows from the real case together with Theorem 5.4 and Definition (2.25). \(\square \)

Of course the case \(p =2\) of Lemma 6.2 is known from Lemma 4.8: \(E_{2,q}(K,Q)=\Gamma _{-2,-q}(K,Q)\).

Our general \(L_{p}\) version of John’s inclusion will be a corollary of

Theorem 6.3

If \(K\in {\mathcal {K}}^{n}_{o}, Q\in {\mathcal {S}}^{n}_{o}, p_i\in (0,\infty ], q_i=p_i+r, r\in [0,\infty )\), \(i=1,2\), then

$$\begin{aligned}&\Gamma _{-p_{1},-q_{1}}(K,Q)\supseteq n^{\frac{1}{2}-\frac{1}{p_{1}}}E_{p_{2},q_{2}}(K,Q)\ \ \ \text{ when }\ \ \ 0<p_{1}\le p_{2}\le 2,\\&\Gamma _{-p_{1},-q_{1}}(K,Q)\subseteq n^{\frac{1}{2}-\frac{1}{p_{1}}}E_{p_{2},q_{2}}(K,Q)\ \ \ \text{ when }\ \ \ 2\le p_{2}\le p_{1}\le \infty . \end{aligned}$$

Proof

Note that \(q_{i}=p_{i}+r, i=1,2\) and \(0\le r<\infty \). Lemmas 4.7 and 6.1 show that it suffices to prove the inclusions when \(E_{p_{2},q_{2}}(K,Q)\) is the unit ball B. Since \(E_{p_{2},q_{2}}(K,Q)=B\), Definition 4.6 gives

$$\begin{aligned} {\widetilde{V}}_{p_{2},q_{2}}(K,B,Q)={\widetilde{V}}_{r}(K,Q). \end{aligned}$$
(6.3)

Suppose \(0<p_{2}\le 2\). Now Definition (2.26), Definition (1.6), Jensen’s inequality, Definition (1.6) again, (6.3), Jensen’s inequality again, (6.3) again, and finally Theorem 4.5 show that for each \(u\in S^{n-1}\),

$$\begin{aligned} \rho _{\Gamma _{-p_{1},-q_{1}}(K,Q)}(u)^{-1}= & {} n^{\frac{1}{p_{1}}} \left[ \int _{S^{n-1}}\left( \frac{|u\cdot v|\rho _{K}(v)}{h_{K} (\alpha _{K}(v))\rho _{Q}(v)}\right) ^{p_{1}}{\mathrm{d}} \overline{{\widetilde{V}}}_{r}(K,Q; v)\right] ^{\frac{1}{p_{1}}}\\\le & {} n^{\frac{1}{p_{1}}}\left[ \int _{S^{n-1}}\left( \frac{|u\cdot v|\rho _{K}(v)}{h_{K}(\alpha _{K}(v))\rho _{Q}(v)} \right) ^{p_{2}}{\mathrm{d}}\overline{{\widetilde{V}}}_{r}(K,Q; v)\right] ^{\frac{1}{p_{2}}}\\= & {} n^{\frac{1}{p_{1}}}\left[ \frac{1}{{\widetilde{V}}_{r}(K,Q)} |u\cdot v|^{p_{2}}{\mathrm{d}}{\widetilde{C}}_{p_{2},q_{2}}(K,Q,v) \right] ^{\frac{1}{p_{2}}}\\= & {} n^{\frac{1}{p_{1}}}\left[ \frac{1}{{\widetilde{V}}_{p_{2},q_{2} }(K,B,Q)}\int _{S^{n-1}}|u\cdot v|^{p_{2}}{\mathrm{d}} {\widetilde{C}}_{p_{2},q_{2}}(K,Q,v)\right] ^{\frac{1}{p_{2}}}\\\le & {} n^{\frac{1}{p_{1}}}\left[ \frac{1}{{\widetilde{V}}_{p_{2},q_{2}} (K,B,Q)}\int _{S^{n-1}}|u\cdot v|^{2}{\mathrm{d}}{ \widetilde{C}}_{p_{2},q_{2}}(K,Q,v)\right] ^{\frac{1}{2}}\\= & {} n^{\frac{1}{p_{1}}}\left[ \frac{1}{{\widetilde{V}}_{r}(K,Q)} \int _{S^{n-1}}|u\cdot v|^{2}{\mathrm{d}}{\widetilde{C}}_{p_{2},q}(K,Q,v )\right] ^{\frac{1}{2}}\\= & {} n^{\frac{1}{p_{1}}-\frac{1}{2}}. \end{aligned}$$

Thus, \(n^{\frac{1}{2}-\frac{1}{p_{1}}}E_{p_{2},q_{2}}(K,Q)\subseteq \Gamma _{-p_{1},q_{1}}(K,Q)\).

When \(2\le p_{1}\le p_{2}<\infty \), the inequality above is reversed. Thus,

$$\begin{aligned} \Gamma _{-p_{1},q_{1}}(K,Q)\subseteq n^{\frac{1}{2}-\frac{1}{p_{1}}}E_{p_{2},q_{2}}(K,Q). \end{aligned}$$

The case \(p=\infty \) follows from the real case together with Theorem 5.4 and Definition (2.25). \(\square \)

By taking \(p_{1} =p_{2}=p\) in Theorem 6.3 and combining the inclusions with those of Lemma 6.2 we get the general \(L_{p}\) version of John’s inclusion:

Corollary 6.4

If \(K\in {\mathcal {K}}^{n}_{o}, Q\in {\mathcal {S}}^{n}_{o}, p,q\in (0,\infty ]\) with \(p\le q\), then

$$\begin{aligned}&E_{p,q}(K,Q)\supseteq \Gamma _{-p-q}(K,Q)\supseteq n^{\frac{1}{2}-\frac{1}{p}}E_{p,q}(K,Q)\ \ \ \text{ when }\ \ \ 0<p\le 2,\\&E_{p,q}(K,Q)\subseteq \Gamma _{-p,-q}(K,Q)\subseteq n^{\frac{1}{2}-\frac{1}{p}}E_{p,q}(K,Q)\ \ \ \text{ when }\ \ \ 2\le p \le \infty . \end{aligned}$$

7 Volume-Ratio Inequalities

We first established the following inequality.

Theorem 7.1

If \(K\in {\mathcal {K}}^{n}_{o}, Q\in {\mathcal {S}}^{n}_{o}\), \(r\in [0,\infty )\) and \(p_{1},p_{2},p_{1},p_{2}\in (0,+\infty ]\) with satisfying that \(p_1<p_2\), \(q_{1}=p_1+r\) and \(q_2=p_{2}+r\), then

$$\begin{aligned} |E_{p_{1},q_{1}}(K,Q)|\le |E_{p_{2},q_{2}}(K,Q)|. \end{aligned}$$

Proof

From Definitions (1.10), together with Jensen’s inequality, it follows that for \(0<p_{1}\le p_{2}\le \infty \),

$$\begin{aligned} \left( \frac{{\widetilde{V}}_{p_{1},q_{1}}(K,L,Q)}{{\widetilde{V}}_{r}(K,Q)} \right) ^{\frac{1}{p_{1}}}= & {} \left( \int _{S^{n-1}}\left( \frac{h_{L} (\alpha _{K}(u))\rho _{K}(u)}{h_{K}(\alpha _{K}(u))\rho _{Q}(u)}\right) ^{p_{1}} {\mathrm{d}}\overline{{\widetilde{V}}}_{r}(K,Q; u)\right) ^{\frac{1}{p_{1}}}\\\le & {} \left( \int _{S^{n-1}}\left( \frac{h_{L}(\alpha _{K}(u)) \rho _{K}(u)}{h_{K}(\alpha _{K}(u))\rho _{Q}(u)}\right) ^{p_{2}} {\mathrm{d}}\overline{{\widetilde{V}}}_{r}(K,Q; u)\right) ^{\frac{1}{p_{2}}}\\= & {} \left( \frac{{\widetilde{V}}_{p_{2},q_{2}}(K,L,Q)}{{\widetilde{V}}_{r}(K,Q)}\right) ^{\frac{1}{p_{2}}}. \end{aligned}$$

This together with Definition 4.6 immediately gives the desired result for real \(p_{2}\) and \(q_{2}\). For the case \(p_{2}=\infty , q_{2}=\infty \), the result follows from the real case and Theorem 5.4. \(\square \)

In general, the (pq)-John ellipsoid \(E_{p,q}(K,Q)\) is not contained in K or Q. However when \(1\le \frac{q}{n}\le p\le q \le n+p\le \infty \), the volume of \(E_{p,q}(K,Q)\) can be dominated by volume of Q.

Theorem 7.2

If \(K\in {\mathcal {K}}^{n}_{o}, Q\in {\mathcal {S}}^{n}_{o}\) and \(1\le \frac{q}{n}\le p\le q \le n+p\le \infty \), then

$$\begin{aligned} |E_{p,q}(K,Q)|\le |Q|, \end{aligned}$$
(7.1)

with equality if and only if KQ are origin-symmetric ellipsoids with dilates of each other when \(1\le \frac{q}{n}< p\), while KQ are an ellipsoid with dilates of each other when \(p=1,q=n\).

Proof

First suppose \(p<\infty \). From Definition (1.9), Definition 4.6 and the \(L_{p}\)-Minkowski inequality (see Lemma 2.4), we have

$$\begin{aligned} {\widetilde{V}}_{r}(K,Q)= & {} {\widetilde{V}}_{p,q}(K,E_{p,q}(K,Q),Q)\nonumber \\\ge & {} |K|^{\frac{q-p}{n}}|E_{p,q}(K,Q)|^{\frac{p}{n}}|Q|^{\frac{n-q}{n}}, r=q-p>0, \end{aligned}$$
(7.2)

with equality if and only if KQ and \(E_{p,q}(K,Q)\) are dilates when \(1<\frac{q}{n}<p\), while K and \(E_{p,q}(K,Q)\) are dilates when \(q=n\) and \(p >1\), and K and \(E_{p,q}(K,Q)\) are homothetic when \(q=n, p=1\).

From the dual \(L_{p}\)-Minkowski inequality (2.9), we have

$$\begin{aligned} {\widetilde{V}}_{r}(K,Q)^{n}\le |K|^{\frac{r}{n}}|Q|^{\frac{n-r}{n}}, \end{aligned}$$
(7.3)

with equality if and only if K and Q are dilates for \(0<r=q-p<n\).

Together with (7.2) and (7.3), we immediately get

$$\begin{aligned} |E_{p,q}(K,Q)|\le |Q|. \end{aligned}$$

The condition of equality follows from ones in (7.2) and (7.3).

For \(p=\infty \) the results follows from the argument for the real case and Theorem 7.1. \(\square \)

When \(Q=K\), an immediate consequence of Theorem 7.2 is

Corollary 7.3

If \(K\in {\mathcal {K}}^{n}_{o}\) and \(1\le p\le \infty \), then

$$\begin{aligned} |E_{p}(K)|\le |K|, \end{aligned}$$
(7.4)

with equality for \(p >1\), if and only if K is an origin-symmetric ellipsoid, and equality for \(p=1\) if and only if K is an ellipsoid.

Note that this inequality is about \(L_{p}\) John ellipsoid proved by Lutwak, Yang and Zhang [30].

If \(p,q\in (0,\infty ]\), K is an origin-symmetric convex body in \({\mathbb {R}}^{n}\), and Q is a star body (about the origin) in \({\mathbb {R}}^{n}\), then K is said to be (pq)-isotropic with respect to Q, if there exists a \(c >0\), such that

$$\begin{aligned} c|x|^{2}=n\int _{S^{n-1}}|x\cdot v|^{2}{\mathrm{d}}{\widetilde{C}}_{p,q}(K,Q,v),\ \ \ \text{ for } \text{ all }\ x\in {\mathbb {R}}^{n}. \end{aligned}$$

For \(Q=K\), then K is said to be \(L_{p}\) isotropic (see [30]).

Theorem 4.5 shows that K is (pq)-isotropic with respect to Q if and only if there exists a \(\lambda >0\), such that

$$\begin{aligned} E_{p,q}(K,Q)=\lambda B. \end{aligned}$$

Theorem 7.4

If \(0\le r\le n\), K and Q are origin-symmetric convex body in \({\mathbb {R}}^{n}\), and K is \((1,1+r)\)-isotropic with respect to Q, then for \(u\in S^{n-1}\),

$$\begin{aligned} h_{\Pi _{1,1+r}(K,Q)}(u)\le \frac{1}{2\sqrt{n}}|K|^{\frac{r}{n}}|Q|^{\frac{n-r}{n}} \left( \frac{\omega _n}{|E_{1,1+r}(K,Q)|}\right) ^{\frac{1}{n^2}}. \end{aligned}$$
(7.5)

Proof

If inequality (7.5) holds for bodies K and Q, then it obviously holds for all \(\lambda K\) and \(\lambda Q\) with \(\lambda >0\). Thus for K that is \((1,1+r)\)-isotropic with respect to Q we may assume that \(E_{1,1+r}(K,Q)=B\). It is necessary to show that

$$\begin{aligned} h_{\Pi _{1,1+r}(K,Q)}(u)\le \frac{1}{2\sqrt{n}}|K|^{\frac{r}{n}}|Q|^{\frac{n-r}{n}}. \end{aligned}$$

Definition 4.6 combined with Definition (1.13) gives

$$\begin{aligned} {\widetilde{V}}_{1,1+r}(K,B,Q)={\widetilde{V}}_{r}(K,Q). \end{aligned}$$
(7.6)

From Definition (2.23), (7.6), Jensen’s inequality, (7.6) again, and finally Theorem 4.5, it follows

$$\begin{aligned}&\frac{2}{{\widetilde{V}}_{r}(K,Q)}h_{\Pi _{1,r+1}(K,Q)}(u)\nonumber \\&\quad =\frac{1}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}}|u\cdot v |{\mathrm{d}}{\widetilde{C}}_{1,r+1}(K,Q,v)\\&\quad =\frac{1}{{\widetilde{V}}_{1,1+r}(K,B,Q)}\int _{S^{n-1}}| u\cdot v|{\mathrm{d}}{\widetilde{C}}_{1,1+r}(K,Q,v)\nonumber \\&\quad \le \left[ \frac{1}{\int _{S^{n-1}}{\mathrm{d}}{\widetilde{C}}_{1,1 +r}(K,Q,v)}\int _{S^{n-1}}|u\cdot v|^{2}{\mathrm{d}} {\widetilde{C}}_{1,1+r}(K,Q,v)\right] ^{\frac{1}{2}}\nonumber \\&\quad =\left[ \frac{1}{{\widetilde{V}}_{r}(K,Q)}\int _{S^{n-1}} |u\cdot v|^{2}{\mathrm{d}}{\widetilde{C}}_{1,1+r}(K,Q,v)\right] ^{\frac{1}{2}} =\frac{1}{\sqrt{n}}.\nonumber \end{aligned}$$
(7.7)

Then we have,

$$\begin{aligned} h_{\Pi _{1,r+1}(K,Q)}(u)\le \frac{1}{2\sqrt{n}}{\widetilde{V}}_{r}(K,Q),\ \ \text{ for }\ u\in S^{n-1}. \end{aligned}$$
(7.8)

Note that \(0\le r\le n\), by using dual Minkowski inequality (2.9), we have

$$\begin{aligned} h_{\Pi _{1,1+r}(K,Q)}(u)\le \frac{1}{2\sqrt{n}}|K|^{\frac{r}{n}}|Q|^{\frac{n-r}{n}}. \end{aligned}$$

\(\square \)

In particular, by taking \(Q=K\) in (7.5), and \(h_{\Pi _{1,1+r}(K,Q)}(u)=\frac{1}{n}h_{\Pi (K)}(u)=\frac{1}{n}{\mathrm{vol}}_{n-1}(K|u^{\bot })\), we have (see [30])

$$\begin{aligned} {\mathrm{vol}}_{n-1}(K|u^{\bot })\le \frac{\sqrt{n}}{2}|K|\left( \frac{\omega _n}{|J(K)|}\right) ^{\frac{1}{n^2}}. \end{aligned}$$
(7.9)