1 Introduction

The classical Brunn–Minkowski inequality for convex bodies (compact convex sets with nonempty interiors) states that for convex bodies K, L in Euclidean n-space, \(\mathbb {R}^n\), the volume of the bodies and of their Minkowski sum \(K+L=\{x+y: x\in \text {and}\; y\in L\}\), are related by

$$\begin{aligned} V\left(K+ L\right )^{\frac{1}{n}} \ge V(K)^{\frac{1}{n}}+ V(L)^{\frac{1}{n}}, \end{aligned}$$
(1)

with equality if and only if K and L are homothetic.

The Brunn–Minkowski inequality has an equivalent formulation as for all real \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )K+ \lambda L\right )\ge V(K)^{1-\lambda }V(L)^{\lambda }, \end{aligned}$$
(2)

and for \(\lambda \in (0,1)\), there is equality if and only if K and L are translates.

The excellent survey article of Gardner [3] gives a comprehensive account of various aspects and consequences of the Brunn–Minkowski inequality.

In the 1960s, Firey [2] introduced for \(p\ge 1\) the so-called Minkowski–Firey \(L_p\) sum of convex bodies that contain the origin in their interiors, and established the \(L_p\)-Brunn–Minkowski inequality, which states as follows:

$$\begin{aligned} V\left((1-\lambda )\cdot K+_p\lambda \cdot L\right )^{\frac{p}{n}} \ge (1-\lambda )V(K)^{\frac{p}{n}}+\lambda V(L)^{\frac{p}{n}}, \end{aligned}$$
(3)

with equality for \(\lambda \in (0,1)\) if and only if K and L are dilates.

In the mid-1990s, it was shown in Refs. [12, 13] that a study of the volume of \(L_p\)-Minkowski addition leads to an \(L_p\)-Brunn–Minkowski theory. This theory has expanded rapidly.

If K and L are convex bodies that contain the origin in their interiors and \(0\le \lambda \le 1\) then the Minkowski–Firey \(L_p\)-combination (\(p>0\)), \((1-\lambda )\cdot K+_p\lambda \cdot L\), is defined by

$$\begin{aligned} (1-\lambda )\cdot K+_p\lambda \cdot L=\bigcap _{u\in S^{n-1}}\{x\in \mathbb {R}^n: x\cdot u\le \left((1-\lambda )h_K(u)^p+\lambda h_L(u)^p\right )^{1/p}\}. \end{aligned}$$
(4)

It has been noticed that the \(L_p\)-Minkowski addition makes sense for all \(p>0\). The case \(p=0\) is known as the log-Minkowski addition, \((1-\lambda )\cdot K+_0 \lambda \cdot L\), of convex bodies K and L that contain the origin in their interior, defined by

$$\begin{aligned} (1-\lambda )\cdot K+_0\lambda \cdot L= \bigcap _{u\in S^{n-1}}\{x\in \mathbb {R}^n: x\cdot u \le h_K(u)^{1-\lambda }h_L(u)^{\lambda }\}. \end{aligned}$$
(5)

In Ref. [1], Böröczky, Lutwak, Yang and Zhang conjectured the log-Brunn–Minkowski inequality: If K and L are o-symmetric convex bodies in \(\mathbb {R}^n\), then for all \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )\cdot K+_0\lambda \cdot L\right ) \ge V(K)^{(1-\lambda )} V(L)^{\lambda }. \end{aligned}$$
(6)

The log-Brunn–Minkowski inequality is stronger than the \(L_p\)-Brunn–Minkowski inequality for \(p>0\). It was shown in Ref. [1] that the log-Brunn–Minkowski inequality is equivalent to the following log-Minkowski mixed volume inequality: If K and L are o-symmetric convex bodies in \(\mathbb {R}^n\), then

$$\begin{aligned} \int _{S^{n-1}}\log \frac{h_L}{h_K}d\bar{V}_K\ge \frac{1}{n} \log \frac{V(L)}{V(K)}. \end{aligned}$$
(7)

Here \(\bar{V}_K\) denotes the cone-volume probability measure of K.

The planar case of the log-Brunn–Minkowski inequality and the equivalent log-Brunn–Minkowski inequality were proved in Ref. [1].

Theorem 1.1

([1]) If K and L are o-symmetric convex bodies in \(\mathbb {R}^2\), then for all real \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )\cdot K+_0\lambda \cdot L\right ) \ge V(K)^{(1-\lambda )} V(L)^{\lambda }, \end{aligned}$$
(8)

with equality for \(\lambda \in (0,1)\) if and only if K and L are dilates or K and L are parallelograms with parallel sides.

Theorem 1.2

([1]) If K and L are o-symmetric convex bodies in \(\mathbb {R}^2\), then,

$$\begin{aligned} \int _{S^1}\log \frac{h_L}{h_K}d\bar{V}_K\ge \frac{1}{2} \log \frac{V(L)}{V(K)}, \end{aligned}$$
(9)

with equality if and only if K and L are dilates or K and L are parallelograms with parallel sides.

It is easily seen from definition (4) that for fixed convex bodies KL and fixed \(\lambda \in [0,1]\), the \(L_p\)-Minkowski–Firey combination \((1-\lambda )\cdot K+_p\lambda \cdot L\) is increasing with respect to set inclusion, as p increases, i.e., if \(0\le p\le q\),

$$\begin{aligned} (1-\lambda )\cdot K+_p\lambda \cdot L\subset (1-\lambda )\cdot K+_q\lambda \cdot L. \end{aligned}$$
(10)

From (9), the \(L_p\)-Brunn–Minkowski inequality and the \(L_p\)-Minkowski inequality were proved in Ref. [1] for \(p\in (0,1)\).

Theorem 1.3

([1]) Suppose \(0<p<1\). If K and L are o-symmetric convex bodies in \(\mathbb {R}^2\), then for all real \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )\cdot K+_p\lambda \cdot L\right ) \ge V(K)^{(1-\lambda )} V(L)^{\lambda }, \end{aligned}$$
(11)

with equality for \(\lambda \in (0,1)\) if and only if \(K=L\).

Theorem 1.4

([1]) Suppose \(0<p<1\). If K and L are o-symmetric convex bodies in \(\mathbb {R}^2\), then for all \(\lambda \in [0,1]\),

$$\begin{aligned} \left( \int _{S^1}\left( \frac{h_L}{h_K}\right) ^pd\bar{V}_K\right) ^{\frac{1}{p}}\ge \left( \frac{V(L)}{V(K)}\right) ^{\frac{1}{2}}, \end{aligned}$$
(12)

with equality for \(\lambda \in (0,1)\) if and only if K and L are dilates.

In Ref. [18], Ma gave an alternative proof of Theorem 1.2. Some results of the log-Brunn–Minkowski inequality for \(n\ge 3,\) see Refs. [19, 21, 25].

There is a counterexample, showing that, if K is an o-centered cube and L is a distinct translate of K, then (6) does not hold for general non-o-symmetric convex bodies. By introducing the notion of “dilation position”, Xi and Leng [23] proved the log-Brunn–Minkowski inequality and the equivalent log-Minkowski mixed volume inequality for general planar convex bodies.

Theorem 1.5

([23]) If K and L are convex bodies in \(\mathbb {R}^2\) with \(o\in K\cap L\), and KL are in dilation position, then for all real \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )\cdot K+_0\lambda \cdot L\right ) \ge V(K)^{(1-\lambda )} V(L)^{\lambda }, \end{aligned}$$
(13)

with equality for \(\lambda \in (0,1)\) if and only if K and L are dilates or K and L are parallelograms with parallel sides.

Theorem 1.6

([24]) If K and L are convex bodies in \(\mathbb {R}^2\) with \(o\in K\cap L\), and KL are in dilation position, then

$$\begin{aligned} \int _{S^1}\log \frac{h_L}{h_K}d\bar{V}_K\ge \frac{1}{2} \log \frac{V(L)}{V(K)}, \end{aligned}$$
(14)

with equality if and only if K and L are dilates or K and L are parallelograms with parallel sides.

The Orlicz–Brunn–Minkowski theory originated with the work of Lutwak et al. [15, 16]. By introducing the Orlicz–Minkowski addition, Gardner, Hug and Weil [4], and Xi et al. [24] proved the Orlicz–Brunn–Minkowski inequality and Orlicz–Minkowski inequality. It is a natural extension of the \(L_p\)-Brunn–Minkowski theory for \(p\ge 1\). For dual Orlicz–Brunn–Minkowski theory see [5, 26].

Let \(\Phi \) be the set of strictly increasing functions \(\phi : (0,\infty )\rightarrow I\subset \mathbb {R}\) which are continuously differentiable on \((0,\infty )\) with positive derivative, and satisfy that \(\lim _{t\rightarrow \infty }\phi (t)=\infty \) and that \(\log \circ \phi ^{-1}\) is concave. Observe that whenever \(\phi \in \Phi \) is convex, the composite function \(\log \circ \phi ^{-1}\) is concave. The collection of convex functions from \(\Phi \) shall be denoted by \(\mathcal {C}\).

Let \(\lambda \in [0,1]\) and \(\phi \in \Phi \). For \(u\in S^{n-1}\), we define a function \(h_{\lambda }(u)\) as

$$\begin{aligned} h_{\lambda }(u)=\inf \{\tau >0: (1-\lambda )\phi \left(\frac{h_K(u)}{\tau }\right)+\lambda \phi \left(\frac{h_L(u)}{\tau }\right)\le \phi (1)\}. \end{aligned}$$
(15)

By the strict monotonicity of \(\phi \), we have

$$\begin{aligned} \phi (1)=(1-\lambda )\phi \left(\frac{h_K(u)}{h_{\lambda }(u)}\right)+\lambda \phi \left(\frac{h_L(u)}{h_{\lambda }(u)}\right). \end{aligned}$$
(16)

The \(\phi \)-combination \((1-\lambda )\cdot K+_{\phi } \lambda \cdot L\) of \(K, L\in \mathcal {K}^n_o\) is defined in Ref. [17] by

$$\begin{aligned} (1-\lambda )\cdot K+_{\phi } \lambda \cdot L=\bigcap _{u\in S^{n-1}}\{x\in \mathbb {R}^n: x\cdot u\le h_{\lambda }(u)\}. \end{aligned}$$
(17)

Note that if \(\phi (t)=t^p\) with \(p>0\), then the \(\phi \)-combination reduces to the \(L_p\)-Minkowski combination. Further, if \(\phi (t)=\alpha \log (t) (\alpha >0)\), then we retrieve the log-Minkowski combination. In Ref. [17], Lv proved the \(\phi \)-Minkowski inequality and \(\phi \)-Brunn–Minkowski inequality for general functions \(\phi \) for o-symmetric planar convex bodies KL. If \(\phi (t)=t^p, p\in (0,1)\), then the \(\phi \)-Minkowski inequality reduces to the \(L_p\)-Minkowski inequality (12) and \(L_p\)-Brunn–Minkowski inequality (11).

In this paper, we first present a new proof Theorem 1.6, and extend Theorems 1.3 and 1.4 from \(p\in (0,1)\) and o-symmetric convex bodies KLto general case \(\phi \) and general convex bodies KL. More precisely, we have the following main results.

Theorem 1.7

Let \(\phi \in \Phi \) with \(\phi \ne \alpha \log (\alpha >0)\), and K and L are planar convex bodies containing the origin o in their interiors, and \(o\in K\cap L\). If K and L are at a dilation position, then

$$\begin{aligned} \int _{S^1}\phi \left( \frac{h_L}{h_K}\right) d\bar{V}_K\ge \phi \left( \frac{V(L)^{\frac{1}{2}}}{V(K)^{\frac{1}{2}}}\right), \end{aligned}$$
(18)

with equality if and only if K and L are dilates.

Theorem 1.8

Let \(\phi \in \Phi \), \(\phi \ne \alpha \log (\alpha >0)\) be concave on \((0,\infty )\), and K and L are planar convex bodies containing the origin o in their interiors, and \(o\in K\cap L\). If K and L are at a dilation position, then for all real \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )\cdot K+_{\phi }\lambda \cdot L\right ) \ge V(K)^{(1-\lambda )} V(L)^{\lambda }, \end{aligned}$$
(19)

with equality for \(\lambda \in (0,1)\) if and only if \(K=L\).

2 Preliminaries

Let \(\mathcal {K}^n\) be the class of convex bodies (compact convex sets with nonempty interiors) in \(\mathbb {R}^n\), and let \(\mathcal {K}^n_o\) be those sets in \(\mathcal {K}^n\) containing the origin in their interiors.

The support function \(h_K: \mathbb {R}^n\rightarrow \mathbb {R}\), of compact convex subset K of \(\mathbb {R}^n\) is defined by \(h_K(x)=\{x\cdot y: y\in K\}\), for \(x\in \mathbb {R}^n\), and uniquely determines the convex set.

A boundary point \(x\in \partial K\) of the convex body K is said to have \(u\in S^{n-1}\) as one of its outer unit normals provided \(x\cdot u=h_K(u)\). A boundary point is said to be singular if it has more than one unit normal vector. It is well known that the set of singular boundary points of a convex body has \((n-1)\)-dimensional Hausdorff measure \(\mathcal {H}^{n-1}\) equal to 0.

Let \(K\in \mathcal {K}^n\) and \(\nu _K:\partial K\rightarrow S^{n-1}\) the generalized Gauss map. For each Borel set \(\omega \subset S^{n-1}\), the inverse spherical image \(\nu _K^{-1}(\omega )\) of \(\omega \) is the set of all boundary points of K which have an outer unit normal belonging to the set \(\omega \). The surface area measure \(S_K\) of \(K\in \mathcal {K}^n\) is defined by

$$\begin{aligned} S_K(\omega )=\mathcal {H}^{n-1}(\nu _K^{-1}(\omega )), \end{aligned}$$
(20)

for each Borel set \(\omega \subset S^{n-1}\), i.e., \(S_K(\omega )\) is the \((n-1)\)-dimensional Hausdorff measure of the set of all points on \(\partial K\) that have a unit normal that lies in \(\omega \).

The Hausdorff distance \(d_H(K,L)\) of compact convex sets KL is defined by \(d_H(K,L)=\Vert h_K-h_L\Vert _{\infty }\). A sequence of convex bodies, \(K_i\), is said to converge to a body K, i.e., \(\lim _{i\rightarrow \infty }K_i=K\) if \(d_H(K_i,K)\rightarrow 0\). If K is a convex body and \(K_i\) is a sequence of convex bodies then

$$\begin{aligned} \lim _{i\rightarrow \infty }K_i=K \Rightarrow \lim _{i\rightarrow \infty }S_{K_i}=S_K, \;\text {weakly}. \end{aligned}$$
(21)

The cone-volume measure \(V_K\) of \(K\in \mathcal {K}^n\) is a Borel measure on the unit sphere \(S^{n-1}\) defined for a Borel set \(\omega \subset S^{n-1}\) by

$$\begin{aligned} V_K(\omega )=\frac{1}{n} \int _{x\in \nu _K^{-1}(\omega )}x\cdot \nu _K(x) \text{d}\mathcal {H}^{n-1}(x), \end{aligned}$$
(22)

and thus

$$\begin{aligned} \text{d}V_K=\frac{1}{n} h_KdS_K. \end{aligned}$$
(23)

Since,

$$\begin{aligned} V(K)=\frac{1}{n}\int _{u\in S^{n-1}}h_K(u)\text{d}S_K(u), \end{aligned}$$
(24)

we can define the cone-volume probability measure \(\bar{V}_K\) of K by

$$\begin{aligned} \bar{V}_K=\frac{1}{V(K)}V_K. \end{aligned}$$
(25)

Suppose \(K,L\in \mathcal {K}^n_o\). For \(p\ne 0\), the \(L_p\)-mixed volume \(V_p(K,L)\) can be defined as

$$\begin{aligned} V_p(K,L)=\int _{u\in S^{n-1}}\left( \frac{h_L}{h_K}\right) ^p\text{d}V_K. \end{aligned}$$
(26)

The normalized \(L_p\)-mixed volume \(\bar{V}_p(K,L)\) was first defined in Ref. [14],

$$\begin{aligned} \bar{V}_p(K,L)=\left(\int _{u\in S^{n-1}}\left( \frac{h_L}{h_K}\right) ^p\text{d}\bar{V}_K\right )^{\frac{1}{p}}. \end{aligned}$$
(27)

For \(p=\infty \), we define

$$\begin{aligned} \bar{V}_{\infty }(K,L)=\max \{h_L/h_K: u\in \text {supp}S_K \}, \end{aligned}$$
(28)

and we have

$$\begin{aligned} \lim _{p\rightarrow \infty } \bar{V}_p(K,L)=\bar{V}_{\infty }(K,L). \end{aligned}$$
(29)

Letting \(p\rightarrow 0\) gives

$$\begin{aligned} \bar{V}_0(K,L)=\exp \left( \int _{u\in S^{n-1}}\log \frac{h_L}{h_K}\text{d}\bar{V}_K\right) , \end{aligned}$$
(30)

which is the normalized log-mixed volume of K and L. From Jesen’s inequality we know that \(p\mapsto \bar{V}_p(K,L)\) is strictly monotone increasing, unless \(h_L/h_K\) is constant on \(\text {supp}S_K\).

Suppose \(K,L\in \mathcal {K}^n\). The inradius r(KL) and R(KL) of K with respect to L are defined by

$$\begin{aligned} r(K,L)=\sup \{t>0: x+tL \subset K \;\text {and}\; x\in \mathbb R^n\}, \end{aligned}$$
$$\begin{aligned} R(K,L)=\inf \{t>0: x+tL \supset K \;\text {and}\; x\in \mathbb R^n\}. \end{aligned}$$

From the definition, it follows that \(r(K,L)=1/R(L,K)\).

If KL happen to be o-symmetric convex bodies, then clearly

$$\begin{aligned} r(K,L)=\min _{u\in S^{n-1}}\frac{h_K(u)}{h_L(u)} \;\text {and}\; R(K,L)=\max _{u\in S^{n-1}}\frac{h_K(u)}{h_L(u)}. \end{aligned}$$
(31)

Let \(K,L \in \mathcal K^n\). K and L are said to be at a dilation position, if \(o\in K\cap L\), and

$$\begin{aligned} r(K,L)L \subset K \subset R(K,L) L. \end{aligned}$$
(32)

The definition and some properties of dilation position were first given by Xi and Leng [23]. It is easy to prove that if KL are o-symmetric convex bodies, then K and L are at a dilation position.

In general, we refer the reader to [20] for standard notation concerning convex bodies.

3 A new proof of Theorem 1.6

In Ref. [18], Ma gave a proof of Theorem 1.1. In the following, we demonstrate an alternate proof of Theorem 1.5 by employing Ma’s approach [18]. The following lemma is needed in our proof.

Lemma 3.1

([23]) Let \(K,L\in \mathcal K^2\) with \(o\in K\cap L\). If K and L are at a dilation position, then

$$\begin{aligned} \int _{S^1}\frac{h_K}{h_L}\text{d}\bar{V}_K \le \frac{V(L,K)}{V(L)}, \end{aligned}$$
(33)

with equality if and only if K and L are dilates, or K and L are parallelograms with parallel sides.

We repeat the statement of Theorem 1.6, and present our approach.

Theorem 3.2

([23]) If K and L are convex bodies in \(\mathbb {R}^2\) with \(o\in K\cap L\), and KL are at a dilation position, then

$$\begin{aligned} \int _{S^1}\log \frac{h_L}{h_K}\text{d}\bar{V}_K\ge \frac{1}{2} \log \frac{V(L)}{V(K)}, \end{aligned}$$
(34)

with equality if and only if K and L are dilates or K and L are parallelograms with parallel sides.

Proof

Set

$$\begin{aligned} F(t)= \int _{S^1}\log \left( \frac{h_{L+tK}}{h_K}\right) \text{d}\bar{V}_K-{\frac{1}{2}}\log \left( \frac{V(L+tK)}{V(K)}\right) , \;\;t\in [0,\infty ). \end{aligned}$$
(35)

Since \(h_{L+tK}=h_L+th_K\) and \(V(L+tK)=V(L)+2V(L,K)t+V(K)t^2\), we have

$$\begin{aligned} F'(t)= & {} \int _{S^1}\frac{h_K}{h_L+th_K}\text{d}\bar{V}_K-\frac{(V(L,K)+V(K)t)}{V(L)+2V(L,K)t+V(K)t^2}\\= & {} \int _{S^1}\frac{h_K}{h_{L+tK}}\text{d}\bar{V}_K-\frac{V(L+tK,K)}{V(L+tK)}. \end{aligned}$$

By Lemma 5.2 of Ref. [23], we have K and \(L+tK\) are at a dilation position. Therefore, we get \(F'(t)\le 0\) from Lemma3.1, which implies that F(t) is decreasing on \([0,\infty )\).

By mean value theorem for integrals, there exists \(u_0\in S^1\) such that

$$\begin{aligned} \int _{S^1}\log \left(\frac{h_{L+tK}}{h_K}\right )\text{d}\bar{V}_K=\log \left(\frac{h_{L+tK}(u_0)}{h_K(u_0)}\right ). \end{aligned}$$
(36)

Let \(t\rightarrow \infty \), then

$$\begin{aligned} F(t)= & {} \log \left( \frac{h_{L+tK}(u_0)}{h_K(u_0)}\right) -{\frac{1}{2}}\log \left (\frac{V(L+tK)}{V(K)}\right )\\= & {} \log \left(\frac{h_L(u_0)+th_K(u_0)}{h_K(u_0)}\cdot \frac{V(K)^{\frac{1}{2}}}{V(L+tK)^{\frac{1}{2}}}\right )\\= & {} \log \left(\frac{h_L(u_0)+th_K(u_0)}{h_K(u_0)}\cdot \frac{V(K)^{\frac{1}{2}}}{(V(L)+2tV(L,K)+t^2V(K))^{\frac{1}{2}}}\right )\\\rightarrow & {} 0. \end{aligned}$$

Therefore, \(F(t)\ge 0\) for \(t\in [0,\infty )\). In particular, \(F(0)\ge 0\), which implies

$$\begin{aligned} \int _{S^1}\log \frac{h_L}{h_K}\text{d}\bar{V}_K \ge \frac{1}{2} \log \frac{V(L)}{V(K)}. \end{aligned}$$

If the equality holds in (34), then \(F(0)=0\), which implies \(F(t)\equiv 0\) for \(t\in [0,\infty )\). Therefore, \(F'(t)\equiv 0\) for all \(t\in [0,\infty )\). By Lemma 3.1, we have K and \(L+tK\) are dilates, or K and \(L+tK\) are parallelograms with parallel sides. So, K and L are dilates, or K and L are parallelograms with parallel sides. Conversely, if K and L are dilates, or K and L are parallelograms with parallel sides, the equality of (34) holds. \(\square \)

Remark 3.3

In Ref. [23], Xi and Leng proved that Theorems 1.5 and 1.6 are equivalent.

4 Proofs of Theorems 1.7 and 1.8

Suppose \(K,L\in \mathcal {K}^n_o\). For \(\phi \in \Phi \), the \(\phi \)-mixed volume \(V_{\phi }(K,L)\) was defined in Ref. [17] by

$$\begin{aligned} V_{\phi }(K,L)=\int _{S^{n-1}}\phi \left(\frac{h_L}{h_K}\right )\text{d}V_K. \end{aligned}$$
(37)

The normalized \(\phi \)-mixed volume \(\bar{V}_{\phi }(K,L)\) of \(K,L\in \mathcal {K}^n_o\) was defined in Ref. [17] by

$$\begin{aligned} \bar{V}_{\phi }(K,L)=\phi ^{-1}\left(\int _{S^{n-1}}\phi \left(\frac{h_L}{h_K}\right )d\bar{V}_K\right ). \end{aligned}$$
(38)

In particular, if \(\phi (t)=t^p\) with \(p>0\), the normalized \(\phi \)-mixed volume \(\bar{V}_{\phi }(K,L)\) reduces to the normalized \(L_p\)-mixed volume \(\bar{V}_p(K,L)\).

We repeat the statements of Theorems 1.7 and 1.8.

Theorem 4.1

Suppose that \(\phi \in \Phi \) with \(\phi \ne \alpha \log (\alpha >0)\), and \(K,L\in \mathcal K^2_o\) with \(o\in K\cap L\). If K and L are at a dilation position, then

$$\begin{aligned} \int _{S^1}\phi \left(\frac{h_L}{h_K}\right )\text{d}\bar{V}_K\ge \phi \left( \frac{V(L)^{\frac{1}{2}}}{V(K)^{\frac{1}{2}}}\right) , \end{aligned}$$
(39)

with equality if and only if K and L are dilates.

Proof

From the log-concavity of \(\phi ^{-1}\), we have

$$\begin{aligned} \int _{S^{n-1}}\log \frac{h_L}{h_K}\text{d}\bar{V}_K\le \log \circ \phi ^{-1}\left( \int _{S^{n-1}}\phi \left( \frac{h_L}{h_K}\right )\text{d}\bar{V}_K\right) , \end{aligned}$$
(40)

which is equivalent to

$$\begin{aligned} \exp \left( \int _{S^{n-1}}\log \frac{h_L}{h_K}\text{d}\bar{V}_K\right) \le \phi ^{-1}\left( \int _{S^{n-1}}\phi \left( \frac{h_L}{h_K}\right) \text{d}\bar{V}_K\right) . \end{aligned}$$
(41)

That is

$$\begin{aligned} \bar{V}_0(K,L)\le \bar{V}_{\phi }(K,L), \end{aligned}$$
(42)

with equality if and only if \(h_L/h_K\) is constant on \(\text {supp}S_K\). From (14), we have

$$\begin{aligned} \bar{V}_{\phi }(K,L)\ge \frac{V(L)^{\frac{1}{2}}}{V(K)^{\frac{1}{2}}}, \end{aligned}$$
(43)

which leads to (39). From the equality condition of (14) and (42), we have equality holds in (39) if and only if K and L are dilates. \(\square \)

Theorem 4.2

Suppose that \(\phi \in \Phi \), \(\phi \ne \alpha \log (\alpha >0)\) be concave on \((0,\infty )\), and \(K,L\in \mathcal K^2_o\) with \(o\in K\cap L\). If K and L are at a dilation position, then for all real \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )\cdot K+_{\phi }\lambda \cdot L\right ) \ge V(K)^{(1-\lambda )} V(L)^{\lambda }, \end{aligned}$$
(44)

with equality for \(\lambda \in (0,1)\) if and only if \(K=L\).

Proof

Set \(Q_{\lambda }=(1-\lambda )\cdot K+_{\phi } \lambda \cdot L\). From (16) and the concavity of \(\phi \), we have

$$\begin{aligned} \phi (1)=(1-\lambda )\phi \left(\frac{h_K(u)}{h_{\lambda }(u)}\right)+\lambda \phi \left(\frac{h_L(u)}{h_{\lambda }(u)}\right)\le \phi \left( \frac{(1-\lambda ) h_K+\lambda h_L}{h_{\lambda }}\right) . \end{aligned}$$
(45)

By the monotone property of \(\phi \), we have

$$\begin{aligned} h_{\lambda }\le (1-\lambda ) h_K+\lambda h_L. \end{aligned}$$
(46)

From (17), we have \(h_{\lambda }=h_{Q_{\lambda }}\) with respect to the surface area measure \(S_{Q_{\lambda }}\). Hence, we have

$$\begin{aligned} Q_{\lambda }\subset (1-\lambda )K+\lambda L. \end{aligned}$$
(47)

On the other hand, from (16), we have

$$\begin{aligned} 1=\phi ^{-1}\left( (1-\lambda )\phi \left(\frac{h_K(u)}{h_{\lambda }(u)}\right)+\lambda \phi \left(\frac{h_L(u)}{h_{\lambda }(u)}\right)\right) . \end{aligned}$$
(48)

From the log-concavity of \(\phi \), we have

$$\begin{aligned} 0&= (\log \circ \phi ^{-1})\left((1-\lambda )\phi \left(\frac{h_K(u)}{h_{\lambda }(u)}\right)+\lambda \phi \left(\frac{h_L(u)}{h_{\lambda }(u)}\right)\right )\\&\ge (1-\lambda )\log \frac{h_K(u)}{h_{\lambda }(u)}+\lambda \log \frac{h_L(u)}{h_{\lambda }(u)}\\&= \log \frac{h_K^{1-\lambda }h_L^{\lambda }}{h_{\lambda }}, \end{aligned}$$

which implies \(h_K^{1-\lambda }h_L^{\lambda }\le h_{\lambda }\). Hence,

$$\begin{aligned} (1-\lambda )\cdot K+_0\lambda \cdot L\subset Q_{\lambda }. \end{aligned}$$
(49)

From (13), we have

$$\begin{aligned} V(Q_{\lambda })\ge V((1-\lambda )\cdot K+_0\lambda \cdot L)\ge V(K)^{1-\lambda }V(L)^{\lambda }. \end{aligned}$$
(50)

If equality holds in (44), then \(V((1-\lambda )\cdot K+_0\lambda \cdot L)= V(K)^{1-\lambda }V(L)^{\lambda }\). By the equality condition of (13), we have K and L are dilates. In addition, from \(V(Q_{\lambda })= V((1-\lambda )\cdot K+_0\lambda \cdot L)\), we have \((1-\lambda )\cdot K+_0\lambda \cdot L= Q_{\lambda }\), which implies \(K=L\). \(\square \)

We can get the \(L_p\)-Minkowski inequality and \(L_p\)-Brunn–Minkowski inequality for general planar convex bodies by setting \(\phi (t)=t^p\) in Theorems 4.1 and 4.2.

Corollary 4.3

Suppose that \(0<p<1\), and \(K,L\in \mathcal K^2_o\) with \(o\in K\cap L\). If K and L are at a dilation position, then

$$\begin{aligned} \left( \int _{S^1}\left( \frac{h_L}{h_K}\right) ^p \text{d}\bar{V}_K \right) ^{\frac{1}{p}}\ge \left( \frac{V(L)}{V(K)}\right) ^{\frac{1}{2}}, \end{aligned}$$
(51)

with equality if and only if K and L are dilates.

Corollary 4.4

Suppose that \(0<p<1\), and \(K,L\in \mathcal K^2_o\) with \(o\in K\cap L\). If K and L are at a dilation position, then for all real \(\lambda \in [0,1]\),

$$\begin{aligned} V\left((1-\lambda )\cdot K+_{\phi }\lambda \cdot L\right ) \ge V(K)^{(1-\lambda )} V(L)^{\lambda }, \end{aligned}$$
(52)

with equality for \(\lambda \in (0,1)\) if and only if \(K=L\).

5 \(\phi \)-Minkowski measure of asymmetry

In the well-known paper [6], abstracting from some extremal problems arising from geometry or other mathematical branches and from the previous work of many mathematicians, Grünbaum formulated a concept of measures of asymmetry (or symmetry) for convex bodies which, among other applications, can be used to describe how far a convex set is from a (centrally) symmetric one. Since then, the properties and applications of these known asymmetry measures are studied by many mathematicians (see [7,8,9,10,11, 22] and references therein).

In Ref. [7], Guo introduced a family of measures of (central) asymmetry, the so-called p-measures of asymmetry, which have the well-known Minkowski measure of asymmetry as a special case, and showed some similar properties of the p-measures to the Minkowski one. In Ref. [11], Jin, Leng and Guo extended the p-Minkowski measure of asymmetry to an Orlicz version. In addition, Jin et al. [11] showed that \(p \;(1\le p\le \infty )\)-Minkowski measures of asymmetry are closely related to \(L_p\)-mixed volumes. More precisely, we can define \(p \;(1\le p\le \infty )\)-Minkowski measures of asymmetry by \(L_p\)-mixed volumes. In Ref. [9], Jin introduced a measure of asymmetry \(\text {as}_0(K)\) for planar convex bodies K in terms of the log-mixed volume, and extended the p-Minkowski measures of asymmetry to the case \(0\le p\le \infty \).

For \(K \in \mathcal {K}^n\), \(x\in \text {int}(K)\) and \(1 \le p\le \infty \), the p-Minkowski measure of asymmetry of K is defined by

$$\begin{aligned} \text {as}_p(C)=\inf _{x \in \text {int}(C)}\bar{V}_p(K_x, -K_x), \end{aligned}$$
(53)

where \(K_x\) denotes \(K+\{-x\}\). A point \(x \in \text {int}(K)\) satisfying \(\bar{V}_p(K_x, -K_x)=\text {as}_p(K)\) is called a p-critical point of K. The set of all p-critical points is denoted by \(\mathcal {C}_p(K)\). The well-known Minkowski measure of asymmetry is the special case that \(p=\infty \).

Theorem 5.1

([6, 7]) For \(1 \le p \le \infty \), if \(K \in \mathcal {K}^n\) then,

$$\begin{aligned} 1 \le \text {as}_p(K) \le n, \end{aligned}$$
(54)

equality holds on the left-hand side if and only if K is symmetric, and on the right-hand side if and only if K is a simplex.

For the p-critical set \(\mathcal {C}_p(K)\), we have the following theorem.

Theorem 5.2

([6, 7]) For \(1 \le p \le \infty \), and \(K \in \mathcal {K}^n\), we have the following statements:

(1) if \(p=1\), then \(\mathcal {C}_1(K)=\text {int}(K)\);

(2) if \(p=\infty \), then \(\mathcal {C}_{\infty }(K)\) is a convex set with \(\dim (\mathcal {C}_{\infty }(K))+\text {as}_{\infty }(K)\le n\);

(3) if \(p\in (1,\infty )\), then \(\mathcal {C}_{p}(K)\) is a singleton.

Note that if \(K\in \mathcal {K}^2\), then \(\mathcal {C}_{\infty }(K)\) is a singleton, i.e., each planar convex body has a unique critical \(\infty \)-critical point.

For fixed \(K\in \mathcal {K}^n\), we denotes the unique p-critical point of K by \(x_p\) for \(p\in (1,\infty )\). It is easy to see that \(x_p\) are coincide with the center of K if K is symmetric; if K is a simplex, then \(x_p\) are coincide with the centroid of K. There are some other convex bodies that have this property that all \(p(1<p<\infty )\)-critical points coincide.

Example 5.3

(1) If \(K:=a_1a_2a_3a_4\) with \(a_1(-3,0),a_2(0,-3),a_3(4,0)\) and \(a_4(0,3)\), then the quadrilateral K has centroid \(c(\frac{1}{4},0)\) and \(x_p(\frac{4}{15},0)\) for \(p\in (1,\infty ]\);

(2) If \(K:=a_1a_2a_3a_4\) with \(a_1(-5,0),a_2(0,-5),a_3(12,0)\) and \(a_4(0,5)\), then the quadrilateral K has centroid \(c(\frac{7}{3},0)\) and \(x_p(\frac{84}{41},0)\) for \(p\in (1,\infty ]\).

Therefore, we state the following problem.

Problem 5.4

Suppose that \(K \in \mathcal {K}^n\). Is it that \(\dim (\text {conv}\{x_p: p\in (1,\infty )\})=0\)?

The p-Minkowski measure of asymmetry for the case \(p\in [0,1)\) is introduced in Ref. [9].

Given \(K\in \mathcal {K}^2\), let \(s \in \mathcal {C}_{\infty }(K)\) be the unique \(\infty \)-critical point of K. The log-Minkowski measure \(\text {as}_0(K)\) of K is defined by

$$\begin{aligned} \text {as}_0(K)=\bar{V}_0(K_s, -K_s). \end{aligned}$$
(5.3)

Theorem 5.5

([9])

If \(K\in \mathcal K^2\), then,

$$\begin{aligned} 1 \le \text {as}_0(K) \le 2. \end{aligned}$$
(56)

Equality holds on the left-hand side if and only if K is symmetric, and equality holds on the right-hand side if and only if K is a triangle.

If we define \(\text {as}_0(K)=\inf _{x\in \text {int}(K)}\bar{V}_0(K_x, -K_x)\), then when K is a square, \(\text {as}_0(C)<1\). This result shows that \(\text {as}_0(K)\) is not a measure of asymmetry in the sense of Grünbaum [6].

In the following, we introduce a new measure of asymmetry in terms of the normalized \(\phi \)-mixed volume.

Definition 5.6

Suppose that \(\phi \in \Phi \) be concave on \((0,\infty )\), \(K\in \mathcal {K}^2\), and \(s \in \mathcal {C}_{\infty }(K)\) be the unique \(\infty \)-critical point of K. The \(\phi \)-Minkowski measure \(\text {as}_{\phi }(K)\) of K is defined by

$$\begin{aligned} \text {as}_{\phi }(K)=\bar{V}_{\phi }(K_s, -K_s). \end{aligned}$$
(57)

For the \(\phi \)-Minkowski measure, we have the following theorem.

Theorem 5.7

Suppose that \(\phi \in \Phi \) be concave on \((0,\infty )\). If \(K\in \mathcal K^2\), then,

$$\begin{aligned} 1 \le \text {as}_{\phi }(K) \le 2. \end{aligned}$$
(58)

Equality holds on the left-hand side if and only if K is symmetric, and equality holds on the right-hand side if and only if K is a triangle.

Proof

From (57), (42) and (56), we have

$$\begin{aligned} \text {as}_{\phi }(K)=\, & {} \bar{V}_{\phi }(K_s, -K_s)\\\ge \,& {} \bar{V}_0(K_s, -K_s)\\=\, & {} \text {as}_0(K)\\\ge\, & {} 1. \end{aligned}$$

On the other hand, from the concavity of \(\phi \), we have

$$\begin{aligned} \int _{S^{n-1}}\phi \left( \frac{h_{-K_s}}{h_{K_s}}\right) \text{d}\bar{V}_{K_s}\le \phi \left( \int _{S^{n-1}}\frac{h_{-K_s}}{h_{K_s}}\text{d}\bar{V}_{K_s}\right) . \end{aligned}$$
(59)

From (27), (38), (53), (54) and (59), we have

$$\begin{aligned} \text {as}_{\phi }(K)=\, & {} \bar{V}_{\phi }(K_s, -K_s)\\=\, & {} \phi ^{-1}\left( \int _{S^{n-1}}\phi \left( \frac{h_{-K_s}}{h_{K_s}}\right) \text{d}\bar{V}_{K_s}\right) \\\le\, & {} \int _{S^{n-1}}\frac{h_{-K_s}}{h_{K_s}}\text{d}\bar{V}_{K_s}\\=\, & {} \bar{V}_1(K_s, -K_s)\\=\, & {} \text {as}_1(K)\\\le\, & {} 2. \end{aligned}$$

Hence,

$$\begin{aligned} 1\le \text {as}_0(K)\le \text {as}_{\phi }(K)\le \text {as}_1(K)\le 2. \end{aligned}$$

If K is symmetric, then we have \(1= \text {as}_0(K)\le \text {as}_{\phi }(K)\le \text {as}_1(K)=1\), which implies \(\text {as}_{\phi }(K)=1\); Conversely, if \(\text {as}_{\phi }(K)=1\), then \(1\le \text {as}_0(K)\le \text {as}_{\phi }(K)=1\), which implies \(\text {as}_0(K)=1\), so K is symmetric.

If K is a triangle, then we have \(2= \text {as}_0(K)\le \text {as}_{\phi }(K)\le \text {as}_1(K)=2\), which implies \(\text {as}_{\phi }(K)=2\); Conversely, if \(\text {as}_{\phi }(K)=2\), then \(2= \text {as}_{\phi }(K)\le \text {as}_1(K)\le 2\), which implies \(\text {as}_1(K)=2\), so K is a triangle. \(\square \)