Let $ \mathcal{K}^n $ denote the set of convex bodies (compact, convex subsets with nonempty interiors) in Euclidean $ n $ -space $ \mathbb{R}^n $; Let $ \mathcal{K}_{o}^{n} $ denote the set of convex bodies containing the origin in their interiors in $ \mathbb{R}^{n} $; $ \mathcal{S}_{o}^{n} $ denotes the set of star bodies (about the origin) in $ \mathbb{R}^n $. We use $ |K| $ to denote the $ n $ -dimensional volume of a body $ K $. Besides, write $ S^{n-1} $ for the unit sphere in $ \mathbb{R}^{n} $.
For a compact set $ K\subset\mathbb{R}^{n} $ which is star-shaped with respect to the origin, we will use $ \rho_{K}=\rho(K,\cdot) :\mathbb{R}^{n}\backslash\{o\}\rightarrow[0,\infty) $ to denote its radial function. That is,
If $ \rho_{K} $ is continuous and positive, then $ K $ is called star body. In [1], Lutwak introduced the intersection body $ IK $ of each $ K\in \mathcal{S}_{o}^{n} $, which is the star body with radial function
where $ v(\cdot) $ denotes $ (n-1) $ -dimensional volume and $ u^{\bot} $ is the hyperplane orthogonal to $ u $.
For $ 0<p<1 $, the $ L_{p} $ -intersection body $ I_{p}K $ of $ K\in\mathcal{S}_{o}^{n} $ is defined by (see, e.g., Cardner and Giannopoulos [2], Yuan [3])
where $ x\cdot u $ is the usual inner product of $ x,u\in\mathbb{R}^{n} $, and integration is with respect to Lebesgue measure.
Haberl and Ludwig [4] also gave the following definition of $ L_{p} $ -intersection body for convex polytopes and investigated its characterization. There is a different constant between the $ L_{p} $ -intersection body of (1.1) and the following $ L_{p} $ -intersection body in [5]. For $ u\in S^{n-1} $ and $ 0<p<1 $, the $ L_{p} $ -intersection body of $ K\in\mathcal{S}_{o}^{n} $ is defined by
where $ \Gamma $ denotes the Gamma function.
Intersection body [1] played an important role for the solution of the celebrated Busemann-Petty problem, and found applications in geometric tomography [6], affine isoperimetric inequalities [7, 8] and the geometry of Banach spaces [9]. For more details about intersection body we refer the reader to [10-14].
Progress towards an Orlicz Brunn-Minkowski theory was made by Lutwak, Yang and Zhang [15, 16]. This theory plays such a crucial role that it is undeniably applied to a number of areas of geometry. The more development of the Orlicz Brunn-Minkowski theory can be found in [17-31]. Recently, Ma et al. [32] considered convex function $ \varphi:\mathbb{R}\backslash\{o\}\rightarrow(0,\infty) $ such that $ \lim\limits_{t\rightarrow\infty}\varphi(t)=0 $, $ \lim\limits_{t\rightarrow0}\varphi(t)=\infty $ and $ \varphi(0) =\infty $. This means that $ \varphi $ must be increasing on $ (-\infty,0) $ and decreasing on $ (0,\infty) $. They assumed that $ \varphi $ is either strictly increasing on $ (-\infty,0) $ or strictly decreasing on $ (0,\infty) $, and denoted by $ \Phi $ the class of such $ \varphi $. Further, for $ K\in\mathcal{S}_{o}^{n} $, they gave the concept of Orlicz intersection body $ I_{\varphi}K $ as the star body whose radial function is given by
when $ \varphi(t)=|t|^{-p} $, $ 0<p<1 $, then $ I_{\varphi}K=\frac{1}{|K|}I_{p}K. $
In this paper, we also consider the above convex function $ \varphi $ and assume that $ K,L $ are two star bodies (about the origin) in $ \mathbb{R}^{n} $ with volume $ |K| $ and $ |L| $. If $ \varphi\in\Phi $, then the Orlicz mixed intersection bodies $ I_{\varphi}(K,L) $ of $ K $, $ L $ as the star body whose radial function at $ x\in\mathbb{R}^{n} $ is given by
when $ \varphi(t)=|t|^{-p} $ with $ 0<p<1 $, and $ K=L $ in (1.3), then $ I_{\varphi}(K,K)=I_{p}(K,K),$ where the radial function of $ I_{p}(K,K) $ is given by
Our main result deals with the affine isoperimetric inequality for Orlicz mixed intersection bodies. Here and in the following, for $ K\in \mathcal{K}^{n} $, we denote by $ B_{K} $ the $ n $ -ball with the same volume as $ K $ centered at the origin.
Theorem 1.1 If $ K,L\in \mathcal{K}_{o}^{n} $ and $ \varphi\in\Phi $, then
with equality if $ K $ and $ L $ are dilates of each other and have the same midpoints.
For a body $ K\in \mathcal{K}_{o}^{n} $, its support function, $ h_{K}=h(K,\cdot): S^{n-1}\rightarrow\mathbb{R} $ is defined by
For $ c > 0 $, the support function of the convex body $ cK=\{cy:y\in K\} $ is
Let $ {\rm GL}(n) $ denote the group of linear transformations. For $ T\in {\rm GL}(n) $, write $ T^{t} $ for the transpose of $ T $, $ T^{-1} $ for the inverse of $ T $ and $ T^{-t} $ for the inverse of the transpose of $ T $. The support function of the image $ T K=\{Ty:y\in K\} $ is given by
According to the definition of radial function $ \rho_{K} $, for $ K\in \mathcal{S}_{o}^{n} $ and $ a>0 $, we easily get
For $ T\in {\rm GL}(n) $, the radial function of the image $ TK=\{Ty:y\in K\} $ of $ K\in \mathcal{S}_{o}^{n} $ is given by
The radial distance between $ K,L\in \mathcal{S}_{o}^{n} $ is defined by
For $ K\in \mathcal{S}_{o}^{n} $, define the real number $ R_{K} $ and $ r_{K} $ by
For $ K\in \mathcal{K}_{o}^{n} $, then the polar body $ K^{*} $ of $ K $ is defined by
It is easily proved that $ (K^{*})^{*}=K $ if $ K\in\mathcal{K}_{o}^{n} $.
From the definitions of support and radial function, it follows obviously that for each $ K\in \mathcal{K}^{n}_{o} $, we easily get
In the following write $ x $, $ y $ for vectors in $ \mathbb{R}^{n} $ and $ x' $, $ y' $ for vectors in $ \mathbb{R}^{n-1} $. We will use $ (x',s) $, $ (y',s) $ for vectors in $ \mathbb{R}^{n}=\mathbb{R}^{n-1}×\mathbb{R} $. For a convex body $ K $ and a direction $ u\in S^{n-1} $, let $ K_{u} $ denote the image of the orthogonal projection of $ K $ onto $ u^{\bot} $, the subspace of $ \mathbb{R}^{n} $ orthogonal to $ u $. The undergraph and overgraph functions, $ \underline{l}_{u}(K,\cdot):K_{u}\rightarrow\mathbb{R} $ and $ \bar{l}_{u}(K,\cdot):K_{u}\rightarrow\mathbb{R} $, of $ K $ in the direction $ u $ are given by
Therefore the Steiner symmetral $ S_{u}K $ of $ K\in\mathcal{K}_{o}^{n} $ in the direction $ u $ is defined by the body whose orthogonal projection onto $ u^{\bot} $ is identical to that of $ K $ and whose undergraph and overgraph functions are
For more details about the Steiner symmetrization we can refer to[33].
The following lemma will be required.
Lemma 2.1 (see [16], Lemma 1.2) Suppose $ K \in\mathcal{K}_{o}^{n} $ and $ u \in S^{n-1} $. For $ y'\in {\rm relint}K_{u} $, the overgraph and undergraph functions of $ K $ in direction $ u $ are given by
and
From the definition of Orlicz mixed intersection bodies (1.3) and $ \varphi $ is strictly increasing on $ (-\infty,0) $ or strictly decreasing on $ (0,\infty) $, we have the function
is strictly increasing on $ (0,\infty) $ and it is also continuous.Thus we have
Lemma 2.2 If $ K,L\in \mathcal{S}_{o}^{n} $, then for $ u_{0}\in S^{n-1} $, \begin{eqnarray}\frac{1}{|K||L|}\int_{K}\int_{L}\varphi\bigg(\frac{|u_{0}\cdot(y-z)|}{λ_{0}}\bigg){\rm d}y{\rm d}z=1\nonumber\end{eqnarray}if and only if $ \rho_{I_{\varphi}(K,L)}^{-1}(u_{0})=λ_{0}. $
Lemma 2.3 (see [34]) Let $ x=(x_{1},x_{2},\cdots,x_{n})\in \mathbb{R}^{n} $ and $ y=(y_{1},y_{2},\cdots,y_{n})\in \mathbb{R}^{n} $. If $ 0 <m_{1}\leq x_{k}\leq M_{1} $, $ 0<m_{2}\leq y_{k}\leq M_{2} $, $ k=1,\cdots,n $, then
Lemma 2.3 implies that if $ x,y,z\in \mathbb{R}^{n} $, then there exists a constant $ c_{0}\in (0,1) $ such that
If $ K $ and $ L $ are bodies in $ \mathbb{R}^{n} $, their multiplicative $ d_{ab}(K,L) $ is defined by (see [35]) $ d_{ab}(K,L)=\inf\{ab:a,b>0,K\subseteq bL,L\subseteq aK\} $.Whenever we write
Denote by $ c_{\varphi}>0 $ the constant to meet $ c_{\varphi}=\min\{c>0: \max\{\varphi(c),\varphi(-c)\}\leq 1\} $ for each $ \varphi\in\Phi $. And denote distant $ d(K,L) $ of convex bodies $ K $, $ L $ by
Lemma 2.4 If $ K,L\in \mathcal{S}_{o}^{n} $ and $ \varphi\in\Phi $, then for all $ u\in S^{n-1} $, there exists a positive constant $ b $ and $ c_{0}\in (0,1) $ such that
Proof Let $ x_{0}\in S^{n-1} $ and $ \rho_{I_{\varphi}(K,L)}^{-1}(x_{0})=λ_{0} $. Then
We first obtain the upper estimate. From the definition of $ c_{\varphi} $, either $ \varphi(c_{\varphi})=1 $ or $ \varphi(-c_{\varphi})=1 $. If $ \varphi(c_{\varphi})=1 $, by (2.7), we have
With the definition of $ \varphi $, $ \varphi $ is strictly increasing on $ (-\infty,0) $ or strictly decreasing on $ (0,\infty) $. So we can immediately obtain $ c_{\varphi}\leq\frac{d(K,L)}{λ_{0}} $, i.e., $ λ_{0}\leq\frac{d(K,L)}{c_{\varphi}} $.
On the other hand, we obtain the lower estimate. Note that $ \varphi(c_{\varphi})=1 $, $ x_{0}\in S^{n-1} $, from (2.5), (2.6) and (2.3), it follows that
where $ b>0 $ and $ c_{0}\in (0,1) $. Since $ \varphi $ is strictly decreasing on $ (0,\infty) $, we immediately get $ c_{\varphi}\geq\frac{c_{0}}{λ_{0}}\big|\frac{1}{bR_{K}}-\frac{b}{r_{L}}\big| $, i.e., $ λ_{0}\geq\frac{c_{0}}{c_{\varphi}}\big|\frac{1}{bR_{K}}-\frac{b}{r_{L}}\big| $.We complete the proof.
The following shows that the Orlicz mixed intersection bodies operator $ I_{\varphi}: \mathcal{S}_{o}^{n}\rightarrow\mathcal{S}_{o}^{n} $ is continuous.
Lemma 2.5 Suppose $ \varphi\in\Phi $. If $ K_{i},L_{i}\in\mathcal{S}_{o}^{n} $ and $ K_{i}\rightarrow K\in\mathcal{S}_{o}^{n} $, $ L_{i}\rightarrow L\in \mathcal{S}_{o}^{n} $, then $ I_{\varphi}(K_{i},L_{i})\rightarrow I_{\varphi}(K,L) $.
Proof For $ \varphi\in\Phi $ and $ \varphi $ is convex, continuous, and either strictly increasing on $ (-\infty,0) $ or strictly decreasing on $ (0,\infty) $, then for $ u_{0}\in S^{n-1} $, we will show that
Suppose $ \rho_{I_{\varphi}(K_{i},L_{i})}^{-1}(u_{0})=λ_{i} $, by Lemma 2.2, we have
From Lemma 2.4, there exists a positive constant $ b $ and $ c_{0}\in(0,1) $ such that
Since $ K_{i}\rightarrow K\in \mathcal{S}_{o}^{n} $, $ L_{i}\rightarrow L\in \mathcal{S}_{o}^{n} $, we have $ d(K_{i},L_{i})\rightarrow d(K,L)>0 $, $ R_{K_{i}}\rightarrow R_{K}>0 $, $ r_{L_{i}}\rightarrow r_{L}>0 $ and there exist $ \alpha $, $ \beta $ such that $ 0< \alpha\leq λ_{i}\leq \beta< \infty $ for all $ i $. Let $ \{λ_{*}\} $ be a convergent subsequence of $ \{λ_{i}\} $, and suppose that $ λ_{*}\rightarrowλ_{0} $, together with the continuous of $ \varphi $ and (2.9), we have
which by Lemma 2.2 yields $ \rho^{-1}_{I_{\varphi}(K,L)}(u_{0})=λ_{0} $. This shows that
But for radial function on $ S^{n-1} $ pointwise and uniform convergence are equivalent (see Schneider [8], p.54). Thus the pointwise convergence $ \rho^{-1}_{I_{\varphi}(K_{i},L_{i})}\rightarrow \rho^{-1}_{I_{\varphi}(K,L)} $ on $ S^{n-1} $ completes the proof.
We next show that the Orlicz mixed intersection bodies operator is also continuous in $ \varphi $.
Lemma 2.6 For $ K,L\in \mathcal{S}_{o}^{n} $ and $ \varphi_{i}\rightarrow \varphi\in \Phi $, then $ I_{\varphi_{i}}(K,L)\rightarrow I_{\varphi}(K,L) $.
Proof Let $ K,L\in \mathcal{S}_{o}^{n} $ and $ u_{0}\in S^{n-1} $. We will show that $ \rho^{-1}_{I_{\varphi_{i}}(K,L)}\rightarrow \rho^{-1}_{I_{\varphi}(K,L)}. $ For $ \varphi\in\Phi $ with $ \varphi $ is convex, continuous, and either strictly increasing on $ (-\infty,0) $ or strictly decreasing on $ (0,\infty) $, and let $ \rho^{-1}_{I_{\varphi_{i}}(K,L)}(u_{0})=λ_{i} $, i.e.,
then together with Lemma 2.4 we have that there exists a positive constant $ b $ and $ c_{0}\in (0,1) $ such that
Since $ \varphi_{i}\rightarrow \varphi\in \Phi $, we have $ c_{\varphi_{i}}\rightarrow c_{\varphi}>0 $ and thus there exist $ \alpha $, $ \beta $ such that $ 0< \alpha\leq λ_{i}\leq \beta<\infty $ for all $ i $. Denote by $ \{λ_{*}\} $ the arbitrary convergent subsequence of $ \{λ_{i}\} $, and suppose that $ λ_{*}\rightarrow λ_{0} $, together with (2.10), we immediately have
which by Lemma 2.2 yields $ \rho^{-1}_{I_{\varphi}(K,L)}(u_{0})=λ_{0} $. This shows that $ \rho^{-1}_{I_{\varphi_{i}}(K,L)}(u_{0})\rightarrow\rho^{-1}_{I_{\varphi}(K,L)}(u_{0}) $. Since the radial function $ \rho_{I_{\varphi_{i}}(K,L)}\rightarrow \rho_{I_{\varphi}(K,L)}(u_{0}) $ pointwise on $ S^{n-1} $ they converge uniformly and hence $ I_{\varphi_{i}}(K,L)\rightarrow I_{\varphi}(K,L). $
Lemma 2.7 Suppose $ \varphi\in \Phi $. For $ K,L\in\mathcal{S}_{o}^{n} $ and a linear transformation $ T\in {\rm GL}(n) $, then $ I_{\varphi}(TK,TL)= T^{-t}(I_{\varphi}(K,L)). $
Proof Suppose $ x_{0}\in\mathbb{R}^{n} $ and
Let $ s=Ty $, $ t=Tz $, then $ |TK|=|{\rm det}T||K| $, $ |TL|=|{\rm det}T||L| $, $ {\rm d}s=|{\rm det}T|{\rm d}y $, $ {\rm d}t=|{\rm det}T|{\rm d}z $, where $ |{\rm det}T| $ is the absolute value of the determinant of $ T $. From Lemma 2.2, we have
i.e., $ \rho^{-1}_{I_{\varphi}(K,L)}(T^{t}x_{0})=λ_{0}. $ By (2.2), we have $ λ_{0}=\rho^{-1}_{I_{\varphi}(K,L)}(T^{t}x_{0})=\rho^{-1}_{T^{-t}(I_{\varphi}(K,L))}(x_{0}). $ Combining with equality (2.11), we immediately obtain
That is, $ I_{\varphi}(TK,TL)= T^{-t}(I_{\varphi}(K,L)). $
Lemma 3.1 Suppose $ \varphi\in \Phi $ is strictly convex and $ K,L\in \mathcal{K}_{o}^{n} $. If $ u\in S^{n-1} $ and $ x_{1}',x_{2}'\in u^{\bot} $, then
Equality in either inequality holds if $ \rho_{I_{\varphi}(K,L)}(x_{1}',1) =\rho_{I_{\varphi}(K,L)}(x_{2}',-1) $ with $ K $ and $ L $ are dilates having the same midpoints.
Proof We only prove (3.1). Inequality (3.2) can be established in the same way.
For each $ z'\in K_{u} $, $ y'\in L_{u} $, let $ \sigma_{z'}=\sigma_{z'}(u)=|K\cap (z'+\mathbb{R}u)| $ and $ \sigma_{y'}=\sigma_{y'}(u)=|L\cap (y'+\mathbb{R}u)| $ be the lengths of the chords $ K\cap (z'+\mathbb{R}u) $, $ L\cap(y'+\mathbb{R}u) $.
Define $ m_{z'}=m_{z'}(u) $ by $ m_{z'}(u)=\frac{1}{2}\underline{l}_{u}(K,z')+\frac{1}{2}\bar{l}_{u}(K,z') $ such that $ z'+m_{z'}u $ is the midpoint of the chord $ K\cap (z'+\mathbb{R}u) $. And define $ m_{y'}=m_{y'}(u) $ by $ m_{y'}(u)=\frac{1}{2}\underline{l}_{u}(L,y')+\frac{1}{2}\bar{l}_{u}(L,y') $ such that $ y'+m_{y'}u $ is the midpoint of the chord $ L\cap (y'+\mathbb{R}u) $. Note that the midpoints of the chords of $ K $ in the direction $ u $ lie in a subspace if and only if there exists a $ \mu\in K_{u} $ such that $ \mu\cdot z'=m_{z'} \ \mbox{for all} \ z'\in K_{u}. $ Similarly, the midpoints of the chords of $ L $ in the direction $ u $ lie in a subspace if and only if there exists a $ \mu\in L_{u} $ such that $ \mu\cdot y'=m_{y'} \ \mbox{for all} \ y'\in L_{u}. $ If $ λ_{1}> 0 $, then we have
by making the change of variables $ s=m_{y'}+s_{1} $ and $ t=m_{z'}+t_{1} $.
On the other hand, for $ λ_{2}>0 $, we have
by making the change of variables $ s=m_{y'}-s_{1} $ and $ t=m_{z'}-t_{1} $.
Let
From the convexity of $ \varphi $, it follows that
By (3.3), (3.4) and (3.5), it yields that
From Lemma 2.2, we get
Combining with the fact that $ |K|=|S_{u}K| $, $ |L|=|S_{u}L| $ and (3.6), we get
In light of the continuity of $ \varphi $ and (1.3) yields
with the equality requiring equality in (3.5) for all $ z'\in K_{u} $, $ y'\in L_{u} $, and $ s_{1} \in [-\frac{\sigma_{y'}}{2},\frac{\sigma_{y'}}{2}] $, $ t_{1} \in [-\frac{\sigma_{z'}}{2},\frac{\sigma_{z'}}{2}] $.
Since $ \varphi $ is strictly convex, this means that we must have $ \varphi $ can not be linear in a neighborhood of the origin given by
for all $ s_{1} \in (-\frac{\sigma_{y'}}{2},\frac{\sigma_{y'}}{2}) $, $ t_{1} \in (-\frac{\sigma_{z'}}{2},\frac{\sigma_{z'}}{2}) $. Choosing $ λ_{1}=\rho^{-1}_{I_{\varphi}(K,L)}(x_{1}',1) $ and $ λ_{2}=\rho^{-1}_{I_{\varphi}(K,L)}(x_{2}',-1) $, equation (3.7) immediately yields
for all $ y'\in L_{u} $ and $ z'\in K_{u} $.
But this means that the midpoints $ \{(y',m_{y'}): y'\in L_{u} \} $ and $ \{(z',m_{z'}): z'\in K_{u} \} $ of the chords of $ L $, $ K $ parallel to $ u $ lie in the subspaces
of $ \mathbb{R}^{n} $, respectively.As we can observe the equality holds if $ \rho_{I_{\varphi}(K,L)}(x_{1}',1) =\rho_{I_{\varphi}(K,L)}(x_{2}',-1) $ with $ K $ and $ L $ are dilates having the same midpoints.
Lemma 3.2 Suppose $ K,L\in \mathcal{K}_{o}^{n} $, $ \varphi\in\Phi $. If $ u\in S^{n-1} $, then
Proof Let $ y'\in {\rm relint}(I^{*}_{\varphi}(K,L)_{u}) $. According to Lemma 2.1, there exist $ x_{1}'=x_{1}'(y') $ and $ x_{2}'=x_{2}'(y') $ in $ u^{\bot} $ such that
From (2.4), (3.8), (3.9) and Lemma 3.1, it follows that
(3.10) and (3.11) give the inclusion.
Proof of Theorem 1.1 Combining with the Steiner symmetrization argument, there is a sequence of direction $ \{u_{i}\} $, such that the sequences $ \{K_{i}\} $ and $ \{L_{i}\} $ converge to $ B_{K} $ and $ B_{L} $, respectively, where the sequences $ \{K_{i}\} $ and $ \{L_{i}\} $ are defined by
with $ |K|=|K_{i}| $ and $ |L|=|L_{i}| $. Thus $ |K|=|B_{K}| $ and $ |L|=|B_{L}| $.
Since the Steiner symmetrization keeps the volume, by Lemma 3.2, we have
when $ i\rightarrow\infty $, we have $ |I^{*}_{\varphi}(B_{K},B_{L})|\leq |I^{*}_{\varphi}(K,L)|. $
According to the equality condition of Lemma 3.1, above equality holds if $ K $ and $ L $ are dilates of each other and have the same midpoints.