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

An affine isoperimetric inequality for log-concave functions

Zengle Zhang1 and Jiazu Zhou2,3,∗∗ zhangzengle128@163.com zhoujz@swu.edu.cn 1 Key Laboratory of Group and Graph Theories and Applications, Chongqing University of Arts and Sciences, Chongqing 402160, China. 2 School of Mathematics and Statistics, Southwest University, Chongqing 400715, China. 3 College of Science, Wuhan University of Science and Technology, Wuhan 430081, China.
Abstract.

The authors gave an affine isoperimetric inequality [38] that gives a lower bound for the volume of a polar body and the equality holds if and only if the body is a simplex. In this paper, we give a functional isoperimetric inequality for log-concave functions that contains the affine isoperimetric inequality of Lutwak, Yang and Zhang in [38].

Key words and phrases:
LYZ ellipsoid; LYZ polar inequality; log-concave function; affine isoperimetric inequality; reverse affine isoperitmetric inequality.
2010 Mathematics Subject Classification:
52A40, 52A41
*Supported in part by NSFC (No. 12071378, No. 12301071) and the Science and Technology Research Program of Chongqing Municipal Education Commission (No. KJQN202201339);
**The corresponding author

1. Introductions

The isoperimetric problem was known in Ancient Greece. However, the first mathematically rigorous proof was obtained only in the 19th century by Weierstrass based on works of Bernoulli, Euler, Lagrange and others. The isoperimetric problem is equivalent to the isoperimetric inequality that is bounding by the surface area and volume of the geometric domain KK in the Euclidian space n\mathbb{R}^{n} with equality if and only if KK is a standard ball. The natural generalizations of the isoperimetric inequality are Alexandrov-Fenchel inequalities that are bounding by mixed volumes of convex bodies in integral and convex geometry analysis. Blascheke-Santaló inequality, differential affine isoperimetric inequality, Busemann-Petty centroid inequality, Petty projection inequality and more affine isoperimetric inequalities are found with equalities for ellipsoids, polar bodies or simplices [8, 14, 15, 27, 34, 35, 36, 37, 39, 41, 46].

During past decades, the reverse affine isoperimetric inequalities, with simplices, cubes or polar bodies as extremal, received more attentions. Usually, reverse affine isoperimetric inequalities are much harder to establish. One of the known open problem about the reverse affine isoperimetric inequalities is the Mahler conjecture.

Let 𝒦on\mathcal{K}_{o}^{n} denote the set of compact convex sets that contain the origin in their interiors in n\mathbb{R}^{n}. Let K𝒦onK\in\mathcal{K}_{o}^{n}, the polar body KK^{\circ} of KK, is defined by

K={xn:xy1,for allyK},\displaystyle K^{\circ}=\{x\in\mathbb{R}^{n}:x\cdot y\leq 1,\quad\text{for all}\quad y\in K\}, (1.1)

and xyx\cdot y denotes the inner product of xx and yy.

If K𝒦onK\in\mathcal{K}_{o}^{n} is symmetric, Mahler[40] conjectured that

|K||K|4nn!,\displaystyle|K||K^{\circ}|\geq\frac{4^{n}}{n!}, (1.2)

where |K||K| denotes the volume of KK, with equality if and only if either KK or KK^{\circ} is a parallelepiped.

Inequality (1.2) was proved by Mahler [40] for n=2n=2, the 33-dimensional case was solved by Iriyeh and Shibata [28], and the case when n4n\geq 4 is still open. One can see [8, 9, 10, 14, 15, 25, 26, 27, 34, 35, 36, 37, 38, 39, 41, 46, 47, 48, 49] for more references on Mahler’s conjecture, affine isoperimetric inequalities and reverse affine isoperimetric inequalities.

Let K𝒦onK\in\mathcal{K}_{o}^{n}. The radial function of Γ2K\Gamma_{-2}K, usually called the LYZ ellipsoid, is defined by

ρΓ2K(u)2=1|K|K|unK(x)|2hK(nK(x))1𝑑n1(x),uSn1,\displaystyle\rho_{\Gamma_{-2}K}(u)^{-2}=\frac{1}{|K|}\int_{\partial K}|u\cdot n_{K}(x)|^{2}h_{K}(n_{K}(x))^{-1}d\mathcal{H}^{n-1}(x),\quad u\in S^{n-1}, (1.3)

where nK(x)n_{K}(x) denotes the outer unit normal vector of at xKx\in\partial K, hKh_{K} is the support function of KK and n1\mathcal{H}^{n-1} is the (n1)(n-1)-dimensional Hausdorff measure. Another known reverse affine isoperimetric inequality is the following LYZ polar inequality [38].

LYZ polar inequality. Let K𝒦onK\in\mathcal{K}_{o}^{n}, then

|K||Γ2K|(n+1)n+12ωnn!nn2\displaystyle|K^{\circ}||\Gamma_{-2}K|\geq\frac{(n+1)^{\frac{n+1}{2}}\omega_{n}}{n!n^{\frac{n}{2}}} (1.4)

with equality if and only if KK is a simplex whose centroid is at the origin.

Recently, Fang and Zhou [20] introduced the LYZ ellipsoid of log-concave functions by solving an extremum problem. A function f:nf:\mathbb{R}^{n}\to\mathbb{R} is called a log-concave function if logf\log f is concave. Log-concave functions would be considered as extension of convex bodies. Usually, the results of convex bodies can be covered by taking log-concave functions, such as the Gaussian function or the characteristic function of convex bodies. The researches on log-concave functions originated from Ball’s Phd thesis [7], in which the Blaschke-Santaló inequality of even log-concave function was established. In recent years, some considerable progress has been made in establishing the geometric inequalities of log-concave function. Colesanti and Fragalà[17] established the Minkowski inequality of log-concave function via the Prékopa-Leindler inequality. Fang, Xing and Ye[19] further gave LpL_{p} Minkowski inequalities and solved the existence of the even LpL_{p} Minkowski problem of log-concave functions for p>1p>1. John and Löwner ellipsoids for log-concave functions and their related inequalities were given in [3, 32]. Some affine isoperimetric inequalities of the affine surface areas of log-concave functions were obtained in [12, 13]. Other inequalities, such as Pólya-Szegö inequality, Rogers-Shephard inequality and its reverse and some stability results for log-concave functional inequalites can be found in [1, 2, 3, 4, 5, 6, 10, 11, 16, 22, 23, 29, 30, 31, 32, 33].

In this paper, for a convex function φ:n{}\varphi:\mathbb{R}^{n}\to\mathbb{R}\cup\{\infty\}, we use φ\varphi^{\ast} to denote the Legendre transform of φ\varphi, and f=eφf^{\circ}=e^{-\varphi^{\ast}} to denote the dual of a log-concave function ff. The integral of ff on n\mathbb{R}^{n} is denoted by J(f)J(f) and the domain of φ\varphi is denoted by dom(φ)={x:φ(x)<}{\rm dom}(\varphi)=\{x:\varphi(x)<\infty\}. For any function f:nf:\mathbb{R}^{n}\to\mathbb{R} and tt\in\mathbb{R}, the upper level set of ff is defined by

Kt(f)={xn:f(x)t}.K_{t}(f)=\{x\in\mathbb{R}^{n}:f(x)\geq t\}.

Let LCn{\rm LC}_{n} denote the class of all upper semi-continuous log-concave functions f=eφf=e^{-\varphi} with dom(φ)=n{\rm dom}(\varphi)=\mathbb{R}^{n} and 0<J(f)<0<J(f)<\infty.

Let fLCnf\in\ {\rm LC}_{n}, for the LYZ ellipsoid Γ2f\Gamma_{-2}f of the log-concave function ff, defined in Definition 4.1, we have the following affine isoperimetric inequality, the functional LYZ polar inequality of (1.4).

Theorem. Let f=eφLCnf=e^{-\varphi}\in{\rm LC}_{n} and μf\mu_{f} be the surface area measure of ff. Suppose that φ(x)>0\varphi^{\ast}(x)>0 on n\{o}\mathbb{R}^{n}\backslash\{o\} and logt=sup{φ(x/φ(x)):xsupp(μf)}-\log t=\sup\{\varphi^{\ast}(x/\varphi^{\ast}(x)):x\in{\rm supp}(\mu_{f})\}, then

|Kt(f)|J(Γ2f)8n2(n+1)n+12Γ(n2+1)ωnn!nn.\displaystyle|K_{t}(f^{\circ})|J(\Gamma_{-2}f)\geq\frac{8^{\frac{n}{2}}(n+1)^{\tfrac{n+1}{2}}\Gamma(\frac{n}{2}+1)\omega_{n}}{n!n^{n}}. (1.5)

Moreover, if φ(x)>φ(o)\varphi(x)>\varphi(o) for all xn\{o}x\in\mathbb{R}^{n}\backslash\{o\}, then equality holds if and only if f(x)=exK+1f(x)=e^{-{\|x\|_{K}}+1} and KK is a simplex whose centroid is at the origin.

Note that the inequality (1.5) is affine invariant, that is, the inequality (1.5) does not change under compositions of ff with affine transformations of n\mathbb{R}^{n}. Inequality (1.5) includes the LYZ polar inequality as a special case. In particular, when f(x)=exK+1f(x)=e^{-{\|x\|_{K}}+1} and KK is a convex body contains origin in its interior, the inequality (1.5) becomes the LYZ polar inequality (1.4). A detailed derivation can be found in Section 5. The supremum logt-\log t may achieve infinity and in this case the inequality (1.5) is strict. However, logt-\log t is often bounded and some examples of the function ff will be given in Section 5.

This paper is organized as follows. In Section 2, we provide some basics of convex bodies and log-concave functions. The Ball-Barth inequality for isotropic embeddings, which is a key tool to establish our main theorem, is introduced in Section 3. In Section 4, the definition of the LYZ ellipsoid of the log-concave function Γ2f\Gamma_{-2}f is given, and its properties is listed. In Section 5, we give the proof of the main theorem and show that LYZ’s polar inequality is a direct result of inequality (1.5) ((Theorem 5.1).

2. Preliminaries and notations

In this section, we list some preliminaries and notations about convex bodies and log-concave functions. Good general references for the theory of convex bodies and log-concave functions are provided by the books of Gardner [24], Schneider[45] and Rockfellar [43].

2.1. Basics regarding convex bodies

Let e1,,ene_{1},\cdots,e_{n} denote the standard Euclidean basis in the nn-dimensional Euclidean space n\mathbb{R}^{n}. For x,ynx,y\in\mathbb{R}^{n}, xyx\cdot y stands for the inner product of x,yx,y, and |x||x| for the Euclidean norm of xx. Let KK be a convex body (non-empty compact convex set of n\mathbb{R}^{n}) contains the origin in its interior. The Minkowski function of KK is defined by

xK=inf{λ0:xλK}.\displaystyle\|x\|_{K}=\inf\{\lambda\geq 0:x\in\lambda K\}. (2.1)

Clearly, if KK is the unit ball BnB^{n} in n\mathbb{R}^{n}, the Minkowski function K\|\cdot\|_{K} becomes the Euclidean norm.

For convex body KK, its support function hK:nh_{K}:\mathbb{R}^{n}\to\mathbb{R} is defined by

hK(x)=max{xy:yK},\displaystyle h_{K}(x)=\max\{x\cdot y:y\in K\},

and its radial function ρK:n\{o}\rho_{K}:\mathbb{R}^{n}\backslash\{o\}\to\mathbb{R} is defined by

ρK(x)=max{λ0:λxK}.\displaystyle\rho_{K}(x)=\max\{\lambda\geq 0:\lambda x\in K\}.

The support function and Minkowski function are homogeneous of degree 11 while the radial function is homogeneous of degree 1-1. From the definitions of the support, Minkowski and radial functions and the definition of the polar body, one can see that

xK=hK(x)=1ρK(x),\displaystyle\|x\|_{K}=h_{K^{\circ}}(x)=\frac{1}{\rho_{K}(x)}, (2.2)

for any xox\neq o. Let GL(n)GL(n) denote the general linear group of n\mathbb{R}^{n}. For TGL(n)T\in GL(n),

(TK)=TtK,\displaystyle(TK)^{\circ}=T^{-t}K^{\circ}, (2.3)

where TtT^{-t} is the inverse of transpose of TT.

2.2. Functional setting

Let φ:n{+}\varphi:\mathbb{R}^{n}\rightarrow\mathbb{R}\cup\{+\infty\}. If for every x,ynx,y\in\mathbb{R}^{n} and λ[0,1]\lambda\in[0,1],

φ((1λ)x+λy)(1λ)φ(x)+λφ(y),\varphi((1-\lambda)x+\lambda y)\leq(1-\lambda)\varphi(x)+\lambda\varphi(y),

then φ\varphi is a convex function. Let dom(φ)={xn:φ(x)<}.\text{dom}(\varphi)=\{x\in\mathbb{R}^{n}:\varphi(x)<\infty\}. Clearly, dom(φ)\text{dom}(\varphi) is a convex set for any convex function φ\varphi. The Legendre trasformation of φ\varphi is the convex function defined by

φ(y)=supxn{xyφ(x)}yn.\displaystyle\varphi^{*}(y)=\sup_{x\in\mathbb{R}^{n}}\left\{x\cdot y-\varphi(x)\right\}\quad\quad\forall y\in\mathbb{R}^{n}. (2.4)

Clearly, φ(x)+φ(y)xy\varphi(x)+\varphi^{*}(y)\geq x\cdot y for all x,ynx,y\in\mathbb{R}^{n}, there is an equality if and only if xdom(φ)x\in\text{dom}(\varphi) and yy is the gradient of φ\varphi at xx. Hence,

φ(φ(x))+φ(x)=xφ(x).\varphi^{*}(\nabla\varphi(x))+\varphi(x)=x\cdot\nabla\varphi(x).

The convex function φ:n{+}\varphi:\mathbb{R}^{n}\rightarrow\mathbb{R}\cup\{+\infty\} is lower semi-continuous, if the subset {xn:φ(x)t}\{x\in\mathbb{R}^{n}:\varphi(x)\leq t\} is a closed set for any t(,+]t\in(-\infty,+\infty]. If φ\varphi is a lower semi-continuous convex function, then also φ\varphi^{*} is a lower semi-continuous convex function, and φ=φ\varphi^{**}=\varphi.

Let ff be a log-concave function. The total mass of ff is defined by

J(f)=nf(x)𝑑x.\displaystyle J(f)=\int_{\mathbb{R}^{n}}f(x)dx.

For f=eφ,g=eψf=e^{-\varphi},g=e^{-\psi} and α,β>0\alpha,\beta>0, the Asplund sum of ff and gg is defined by

αfβg=e(αφ+βψ).\displaystyle\alpha\cdot f\oplus\beta\cdot g=e^{-(\alpha\varphi^{\ast}+\beta\psi^{\ast})^{\ast}}.

Let f=eφ,g=eψLCnf=e^{-\varphi},g=e^{-\psi}\in{\rm LC}_{n}. Rotem [44] gives the following first variational formula

δJ(f,g)=limt0+J(ftg)J(f)t=nψ(x)𝑑μf(x),\displaystyle\delta J(f,g)=\lim_{t\rightarrow 0^{+}}\frac{J(f\oplus t\cdot g)-J(f)}{t}=\int_{\mathbb{R}^{n}}\psi^{\ast}(x)d\mu_{f}(x), (2.5)

where μf\mu_{f} is the surface area measure of ff, which is defined by

μf(A)=φ(x)Af(x)𝑑x\mu_{f}(A)=\int_{\nabla\varphi(x)\in A}f(x)dx

for any AnA\subset\mathbb{R}^{n}.

3. The Ball-Barth inequality

Let μ\mu be a positive Borel measure μ\mu on SnS^{n}. A positive semi-definite n×nn\times n matrix [μ][\mu] can be generated by μ\mu, which is defined by

[μ]=Snuu𝑑μ(u)\displaystyle[\mu]=\int_{S^{n}}u\otimes ud\mu(u) (3.1)

where uuu\otimes u is the rank 11 matrix generated by uSnu\in S^{n}, or, equivalently by

v[μ]v=Sn|vu|2𝑑μ(u)\displaystyle v\cdot[\mu]v=\int_{S^{n}}|v\cdot u|^{2}d\mu(u) (3.2)

for all vSnv\in S^{n}. A positive Borel measure μ\mu is said to be an isotropic measure if

[μ]=In,\displaystyle[\mu]=I_{n}, (3.3)

where InI_{n} is the identity matrix. If μ\mu is isotropic, summing the equation (3.2) with v=ei,i=1,,n+1v=e_{i},i=1,\cdots,n+1, one has

μ(Sn)=n+1.\displaystyle\mu(S^{n})=n+1. (3.4)

The following Ball-Barth inequality was established in [36]. The Ball-Barth inequality. Let μ\mu an isotropic measure on SnS^{n} and l:Sn(0,)l:S^{n}\rightarrow(0,\infty) be a continuous function. Then

detSnl(u)uu𝑑μ(u)exp{Snlogl(u)𝑑μ(u)},\displaystyle\det\int_{S^{n}}l(u)u\otimes ud\mu(u)\geq\exp\left\{\int_{S^{n}}\log l(u)d\mu(u)\right\}, (3.5)

with equality if and only if l(u1)l(un+1)l(u_{1})\cdots l(u_{n+1}) is a constant for linearly independent u1,,un+1u_{1},\cdots,u_{n+1} in supp(ν){\rm supp}(\nu).

The isotropic embedding from n\mathbb{R}^{n} to SnS^{n} is given as follows.

Definition 3.1.

Let (n,μ)(\mathbb{R}^{n},\mu) be a Borel measure space. A continuous function h:nSnh:\mathbb{R}^{n}\rightarrow S^{n} is an isotropic embedding of (n,μ)(\mathbb{R}^{n},\mu) into SnS^{n} if

n|uh(x)|2𝑑μ(x)=1\displaystyle\int_{\mathbb{R}^{n}}|u\cdot h(x)|^{2}d\mu(x)=1 (3.6)

for all uSnu\in S^{n}.

Let μ\mu and hh have been given in Definition 3.1. Summing the equation (3.6) with u=e1,,en,en+1u=e_{1},\cdots,e_{n},e_{n+1}, one has

μ(n)=n+1.\displaystyle\mu(\mathbb{R}^{n})=n+1. (3.7)

Define a new Borel measure ν\nu on SnS^{n}, which is the push-forward of μ\mu by hh, by

Sng(u)𝑑ν(u)=ng(h(x))𝑑μ(x)\displaystyle\int_{S^{n}}g(u)d\nu(u)=\int_{\mathbb{R}^{n}}g(h(x))d\mu(x)

for any Borel function g:Sng:S^{n}\to\mathbb{R}. Definition 3.1 shows that ν\nu is an isotropic measure on SnS^{n}, by applying the Ball-Barthe inequality to ν\nu, we can obtain the following Ball-Barthe inequality for isotropic embeddings.

Proposition 3.1.

If h:nSnh:\mathbb{R}^{n}\rightarrow S^{n} is an isotropic embedding of the Borel measure space (n,μ)(\mathbb{R}^{n},\mu) into SnS^{n}, then for each continuous l:n(0,)l:\mathbb{R}^{n}\rightarrow(0,\infty)

detnl(x)h(x)h(x)𝑑μ(x)exp{nlogl(x)𝑑μ(x)},\displaystyle\det\int_{\mathbb{R}^{n}}l(x)h(x)\otimes h(x)d\mu(x)\geq\exp\left\{\int_{\mathbb{R}^{n}}\log l(x)d\mu(x)\right\}, (3.8)

with equality if and only if l(x1),,l(xn+1)l(x_{1}),\cdots,l(x_{n+1}) is constants for x1,,xn+1x_{1},\cdots,x_{n+1} in supp(ν){\rm supp}(\nu) such that h(x1),,h(xn+1)h(x_{1}),\cdots,h(x_{n+1}) are linearly independent.

4. The LYZ ellipsoid of log-concave functions

In this section, we will define the LYZ ellipsoid of log-concave functions, and introduce its properties.

Definition 4.1.

Let f=eφLCnf=e^{-\varphi}\in{\rm LC}_{n}. Suppose that φ(x)>0\varphi^{\ast}(x)>0 for any xn\{o}x\in{\mathbb{R}^{n}\backslash\{o\}}, then the LYZ ellipsoid of ff, denoted by Γ2f\Gamma_{-2}f, is defined as

logΓ2f(x)\displaystyle-\log\Gamma_{-2}f(x) =n28δJ(f,f)n|xy|2φ(y)1𝑑μf(y)\displaystyle=\frac{n^{2}}{8\delta J(f,f)}\int_{\mathbb{R}^{n}}|x\cdot y|^{2}\varphi^{*}(y)^{-1}d\mu_{f}(y)
=n28δJ(f,f)n|xφ(y)|2φ(φ(y))1eφ(y)𝑑y.\displaystyle=\frac{n^{2}}{8\delta J(f,f)}\int_{\mathbb{R}^{n}}|x\cdot\nabla\varphi(y)|^{2}\varphi^{*}(\nabla\varphi(y))^{-1}e^{-\varphi(y)}dy. (4.1)

Note that the function logΓ2f(x)\sqrt{-\log\Gamma_{-2}f(x)} is a Minkowski functional of an ellipsoid. Indeed, if EE is an ellipsoid generated by n×nn\times n real symmetric matrix AA, that is

E={xn:xAx1},\displaystyle E=\{x\in\mathbb{R}^{n}:x\cdot Ax\leq 1\},

one can check that the Minkowski function of EE is given by xE2=xAx.\|x\|_{E}^{2}=x\cdot Ax. From this, we see that the function logΓ2f(x)\sqrt{-\log\Gamma_{-2}f(x)} is a Minkowski functional of an ellipsoid, which is generated by n×nn\times n real symmetric matrix M=[mij(f)]M=[m_{ij}(f)], where

mij(f)=n28δJ(f,f)n(eiy)(ejy)φ(y)1𝑑μf(y).\displaystyle m_{ij}(f)={\frac{n^{2}}{8\delta J(f,f)}}\int_{\mathbb{R}^{n}}(e_{i}\cdot y)(e_{j}\cdot y)\varphi^{*}(y)^{-1}d\mu_{f}(y).
Property 4.1.

Let fLCnf\in{\rm LC}_{n} and TGL(n)T\in{\rm GL}(n). Then

Γ2(fT)=(Γ2f)T.\displaystyle\Gamma_{-2}(f\circ T)=(\Gamma_{-2}f)\circ T. (4.2)
Proof.

Let TtT^{t} be the transpose of TT. From the definition of Legendre transform (2.4) and the fact that Txy=xTtyTx\cdot y=x\cdot T^{t}y, one has

(φT)(y)\displaystyle(\varphi\circ T)^{*}(y) =\displaystyle= supxn{xyφ(Tx)}\displaystyle\sup_{x\in\mathbb{R}^{n}}\left\{x\cdot y-\varphi(Tx)\right\}
=\displaystyle= supxn{xTtyφ(x)}\displaystyle\sup_{x\in\mathbb{R}^{n}}\left\{x\cdot T^{-t}y-\varphi(x)\right\}
=\displaystyle= φ(Tty).\displaystyle\varphi^{*}(T^{-t}y). (4.3)

By (4.1) and the fact that (φT)(y)=Ttφ(Ty)\nabla(\varphi\circ T)(y)=T^{t}\nabla\varphi(Ty), one has

n|xy|2(φT)(y)1𝑑μfT(y)\displaystyle\int_{\mathbb{R}^{n}}|x\cdot y|^{2}(\varphi\circ T)^{*}(y)^{-1}d\mu_{f\circ T}(y)
=n|x(φT)(y)|2(φT)((φT)(y))1eφ(Ty)𝑑y\displaystyle\quad\quad=\int_{\mathbb{R}^{n}}|x\cdot\nabla(\varphi\circ T)(y)|^{2}(\varphi\circ T)^{*}(\nabla(\varphi\circ T)(y))^{-1}e^{-\varphi(Ty)}dy
=n|xTtφ(Ty)|2φ(φ(Ty))1eφ(Ty)𝑑y\displaystyle\quad\quad=\int_{\mathbb{R}^{n}}|x\cdot T^{t}\nabla\varphi(Ty)|^{2}\varphi^{*}(\nabla\varphi(Ty))^{-1}e^{-\varphi(Ty)}dy
=n|Txφ(Ty)|2φ(φ(Ty))1eφ(Ty)𝑑y\displaystyle\quad\quad=\int_{\mathbb{R}^{n}}|Tx\cdot\nabla\varphi(Ty)|^{2}\varphi^{*}(\nabla\varphi(Ty))^{-1}e^{-\varphi(Ty)}dy
=|detT|1n|Txy|2φ(y)1𝑑μf(y).\displaystyle\quad\quad=|\det T|^{-1}\int_{\mathbb{R}^{n}}|Tx\cdot y|^{2}\varphi^{*}(y)^{-1}d\mu_{f}(y).

This together with δJ(fT,fT)=|detT|1δJ(f,f)\delta J(f\circ T,f\circ T)=|\det T|^{-1}\delta J(f,f) gives (4.2). ∎

The following corollary shows that the quantity |Kt(f)|J(Γ2(f))|K_{t}(f^{\circ})|J(\Gamma_{-2}(f)) is invariant under the operation of GL(n)GL(n).

Corollary 4.1.

Let fLCnf\in{\rm LC}_{n} and TGL(n)T\in{\rm GL}(n). Then for any t>0t>0

|Kt((fT))|J(Γ2(fT))=|Kt(f)|J(Γ2(f)).|K_{t}((f\circ T)^{\circ})|J(\Gamma_{-2}(f\circ T))=|K_{t}(f^{\circ})|J(\Gamma_{-2}(f)).
Proof.

By (4.3), one has for any TGL(n)T\in GL(n),

Kt((fT))\displaystyle K_{t}((f\circ T)^{\circ}) ={xn:f(Tx)t}={xn:eφ(Ttx)t}\displaystyle=\{x\in\mathbb{R}^{n}:f^{\circ}(Tx)\geq t\}=\{x\in\mathbb{R}^{n}:e^{-\varphi^{*}{(T^{-t}x)}}\geq t\}
={Ttyn:eφ(y)t}=Tt(Kt(f)).\displaystyle=\{T^{t}y\in\mathbb{R}^{n}:e^{-\varphi^{*}{(y})}\geq t\}=T^{t}(K_{t}(f^{\circ})).

Property (4.1) implies that

J(Γ2(fT))=J((Γ2f)T)=n(Γ2f)(Tx)𝑑x=(detT)1J(Γ2f).J(\Gamma_{-2}(f\circ T))=J((\Gamma_{-2}f)\circ T)=\int_{\mathbb{R}^{n}}(\Gamma_{-2}f)(Tx)dx=(\det T)^{-1}J(\Gamma_{-2}f).

Hence

|Kt((fT))|J(Γ2(fT))=|(Kt(f))|J(Γ2f).|K_{t}((f\circ T)^{\circ})|J(\Gamma_{-2}(f\circ T))=|(K_{t}(f^{\circ}))|J(\Gamma_{-2}f).

This completes the proof. ∎

When f=exK22f=e^{-\frac{\|x\|_{K}^{2}}{2}} with K𝒦onK\in\mathcal{K}_{o}^{n}, Γ2f(x)\Gamma_{-2}f(x) can be calculated precisely.

Property 4.2.

Let K𝒦onK\in\mathcal{K}_{o}^{n}. If f=exK22f=e^{-\frac{\|x\|_{K}^{2}}{2}}, then

Γ2f(x)=exΓ2K22.\displaystyle\Gamma_{-2}f(x)=e^{-\frac{\|x\|_{\Gamma_{-2}K}^{2}}{2}}.
Proof.

The normalized cone measure σK\sigma_{K} of KK is defined by

dσK(z)=znK(z)n|K|dn1(z)forzK.d\sigma_{K}(z)=\frac{z\cdot n_{K}(z)}{n|K|}d\mathcal{H}^{n-1}(z)\quad\text{for}\quad z\in\partial K.

For yny\in\mathbb{R}^{n}, we write y=rzy=rz with zKz\in\partial K, then

dy=n|K|rn1drdσK(z).\displaystyle dy=n|K|r^{n-1}drd\sigma_{K}(z). (4.4)

Together with the fact that (xK22)=xK22\left(\frac{\|x\|_{K}^{2}}{2}\right)^{\ast}=\frac{\|x\|_{K^{\circ}}^{2}}{2} for any xnx\in\mathbb{R}^{n}, one has

n|xφ(y)|2φ(φ(y))1eφ(y)𝑑y\displaystyle\int_{\mathbb{R}^{n}}|x\cdot\nabla\varphi(y)|^{2}\varphi^{*}(\nabla\varphi(y))^{-1}e^{-\varphi(y)}dy
=n|xyKyK|2(yKyKK22)1eyK22𝑑y\displaystyle\quad\quad=\int_{\mathbb{R}^{n}}|x\cdot\|y\|_{K}\nabla\|y\|_{K}|^{2}\left(\tfrac{\|\|y\|_{K}\nabla\|y\|_{K}\|_{K^{\circ}}^{2}}{2}\right)^{-1}e^{-\tfrac{\|y\|_{K}^{2}}{2}}dy
=n|xyK|2(yKK22)1eyK22𝑑y\displaystyle\quad\quad=\int_{\mathbb{R}^{n}}|x\cdot\nabla\|y\|_{K}|^{2}\left(\tfrac{\|\nabla\|y\|_{K}\|_{K^{\circ}}^{2}}{2}\right)^{-1}e^{-\tfrac{\|y\|_{K}^{2}}{2}}dy
=2n|K|0Krn1|xzK|2(zKK2)1er22drdσK(z)\displaystyle\quad\quad=2n|K|\int_{0}^{\infty}\int_{\partial K}r^{n-1}|x\cdot\nabla\|z\|_{K}|^{2}\left(\|\nabla\|z\|_{K}\|_{K^{\circ}}^{2}\right)^{-1}e^{-\frac{r^{2}}{2}}drd\sigma_{K}(z)
=2n2Γ(n2)n|K|K|xzK|2(zKK2)1dσK(z).\displaystyle\quad\quad=2^{\frac{n}{2}}\Gamma\left(\tfrac{n}{2}\right)n|K|\int_{\partial K}|x\cdot\nabla\|z\|_{K}|^{2}\left(\|\nabla\|z\|_{K}\|_{K^{\circ}}^{2}\right)^{-1}d\sigma_{K}(z).

By the fact that zK=nK(z)nK(z)K\nabla\|z\|_{K}=\frac{n_{K}(z)}{\|n_{K}(z)\|_{K^{\circ}}} for zKz\in\partial K (see [45, Remark 1.7.14]), (2.2) and (1.3), we have

n|K|K|xzK|2(zKK2)1dσK(z)\displaystyle n|K|\int_{\partial K}|x\cdot\nabla\|z\|_{K}|^{2}\left(\|\nabla\|z\|_{K}\|_{K^{\circ}}^{2}\right)^{-1}d\sigma_{K}(z)
=n|K|K|xzK|2dσK(z)\displaystyle\quad\quad=n|K|\int_{\partial K}|x\cdot\nabla\|z\|_{K}|^{2}d\sigma_{K}(z)
=n|K|K|xnK(z)nK(z)K|2𝑑σK(z)\displaystyle\quad\quad=n|K|\int_{\partial K}\left|x\cdot\tfrac{n_{K}(z)}{\|n_{K}(z)\|_{K^{\circ}}}\right|^{2}d\sigma_{K}(z)
=K|xnK(z)|2(znK(z))1𝑑n1(z)\displaystyle\quad\quad=\int_{\partial K}\left|x\cdot n_{K}(z)\right|^{2}(z\cdot n_{K}(z))^{-1}d\mathcal{H}^{n-1}(z)
=|K|ρΓ2K(x)2\displaystyle\quad\quad=|K|\rho_{\Gamma_{-2}K}(x)^{-2}
=|K|xΓ2K2.\displaystyle\quad\quad=|K|\|x\|_{\Gamma_{-2}K}^{2}.

By using (2.5) and (4.4), we have

δJ(f,f)\displaystyle\delta J(f,f) =12nyKyKK2eyK22𝑑y\displaystyle=\frac{1}{2}\int_{\mathbb{R}^{n}}{\|\|y\|_{K}\nabla\|y\|_{K}\|_{K^{\circ}}^{2}}e^{-\tfrac{\|y\|_{K}^{2}}{2}}dy
=n|K|20rn+1er22𝑑rKzKK2dσK(z)\displaystyle=\frac{n|K|}{2}\int_{0}^{\infty}r^{n+1}e^{-\frac{r^{2}}{2}}dr\int_{\partial K}\|\nabla\|z\|_{K}\|_{K^{\circ}}^{2}d\sigma_{K}(z)
=2n22Γ(n2)n2|K|.\displaystyle=2^{\frac{n}{2}-2}\Gamma(\tfrac{n}{2})n^{2}|K|.

Hence

logΓ2f(x)=xΓ2K22.\displaystyle-\log\Gamma_{-2}f(x)=\frac{\|x\|_{\Gamma_{-2}K}^{2}}{2}.

This completes the proof. ∎

When ff is the Gaussian function e|x|22e^{-\frac{|x|^{2}}{2}}, the following result can be obtained directly from Property 4.2 and the fact that Γ2Bn=Bn\Gamma_{-2}B^{n}=B^{n}.

Corollary 4.1.

If f=γn=e|x|22f=\gamma_{n}=e^{-\frac{|x|^{2}}{2}}, then

Γ2γn=γn.\displaystyle\Gamma_{-2}\gamma_{n}=\gamma_{n}.

5. LYZ polar inequality for log-concave functions

In this section, we will establish the LYZ polar inequality for log-concave functions. The following lemmas are needed.

Lemma 5.1.

Let f=eφLCnf=e^{-\varphi}\in{\rm LC}_{n} and Kt(f)={xn:f(x)t}K_{t}(f)=\{x\in\mathbb{R}^{n}:f(x)\geq t\} be the upper level set of ff. If ν\nu is a finite, positive Borel measure on n\mathbb{R}^{n} and φ>0\varphi^{\ast}>0 on n\{o}\mathbb{R}^{n}\backslash\{o\}, then

nxφ(x)1𝑑ν(x)rKt(f),\displaystyle\int_{\mathbb{R}^{n}}x\varphi^{\ast}(x)^{-1}d\nu(x)\in rK_{t}(f^{\circ}),

where logt=sup{φ(x/φ(x)):xsupp(ν)}-\log t=\sup\{\varphi^{\ast}(x/\varphi^{\ast}(x)):{x\in{\rm supp}(\nu)}\} and r=|ν|r=|\nu|.

Proof.

Let

x0=nxφ(x)1𝑑ν(x).\displaystyle x_{0}=\int_{\mathbb{R}^{n}}x\varphi^{\ast}(x)^{-1}d\nu(x).

By the convexity of φ\varphi^{\ast}, we have

φ(x0r)nφ(xφ(x))dν(x)rlogt.\displaystyle\varphi^{\ast}\left(\frac{x_{0}}{r}\right)\leq\int_{\mathbb{R}^{n}}\varphi^{\ast}\left(\frac{x}{\varphi^{\ast}(x)}\right)\frac{d\nu(x)}{r}\leq-\log t.

Hence

x0r{xn:f(x)t}.\displaystyle x_{0}\in r\left\{x\in\mathbb{R}^{n}:f^{\circ}(x)\geq t\right\}.

Then x0rKt(f)x_{0}\in rK_{t}(f). ∎

Note that logt-\log t may reach infinity, but it is often bounded for many log-concave functions, such as f=eφLCnf=e^{-\varphi}\in{\rm LC}_{n} with φ(x)>0\varphi^{\ast}(x)>0 for xn\{o}x\in\mathbb{R}^{n}\backslash\{o\} and dom(φ)=n{\rm dom}(\varphi^{\ast})=\mathbb{R}^{n}. We first need to show J(f)<J(f^{\circ})<\infty. For the Legendre transformation of φ\varphi,

φ(y)=supxnxyφ(x)|y|φ(y|y|),\varphi^{\ast}(y)=\sup_{x\in\mathbb{R}^{n}}x\cdot y-\varphi(x)\geq|y|-\varphi\left(\frac{y}{|y|}\right),

we have φ(y)\varphi^{\ast}(y)\to\infty as |y||y|\to\infty and hence J(f)<J(f^{\circ})<\infty. Moreover, fLCnf^{\circ}\in{\rm LC}_{n}. [17, Lemma 2.5] shows that there exists two constants a>0a>0 and bb such that

φ(x)a|x|+b\displaystyle\varphi^{\ast}(x)\geq a|x|+b (5.1)

for any xnx\in\mathbb{R}^{n}. Hence

lim|x||x|φ(x)lim|x||x|a|x|+b=1a,\lim_{|x|\to\infty}\frac{|x|}{\varphi^{\ast}(x)}\leq\lim_{|x|\to\infty}\frac{|x|}{a|x|+b}=\frac{1}{a},

that is, for any ε>0\varepsilon>0, there exists a constant MM such that x/φ(x)(1a+ε)Bn{x}/{\varphi^{\ast}(x)}\in\left(\frac{1}{a}+\varepsilon\right)B^{n} for x(MBn)cx\in(MB^{n})^{c} and the unit ball BnB^{n} in n\mathbb{R}^{n}. Then φ(x/φ(x))\varphi^{\ast}\left({x}/{\varphi^{\ast}(x)}\right) is bounded for any x(M0Bn)cx\in(M_{0}B^{n})^{c} as the continuity of φ\varphi^{\ast}. Set m=min{φ(x):xMBn}m=\min\{\varphi^{\ast}(x):x\in MB^{n}\}, then x/φ(x)(M/m)Bnx/\varphi^{\ast}(x)\in(M/m)B^{n} for any xMBnx\in MB^{n}. Hence φ(x/φ(x))\varphi^{\ast}\left({x}/{\varphi^{\ast}(x)}\right) is bounded on MBnMB^{n} and logt-\log t is a finite constant.

The following lemma is the key for the equality condition of our main inequality.

Lemma 5.2.

Let φ\varphi be convex function and dom(φ)=n(\varphi)=\mathbb{R}^{n} and φ(x)>φ(o)\varphi(x)>\varphi(o) for all xn\{o}x\in\mathbb{R}^{n}\backslash\{o\}. Suppose that φ\varphi satisfies the equality

φ(x)=xφ(x)c\displaystyle\varphi(x)=x\cdot\nabla\varphi(x)-c (5.2)

a.e. on n\mathbb{R}^{n}, where cc is a constant. Then φ(x)=xKc\varphi(x)=\|x\|_{K}-c for some K𝒦onK\in\mathcal{K}_{o}^{n}.

Proof.

Firstly, we consider the case when c=0c=0, in which case we claim that

φ(λx)=λφ(x)\displaystyle\varphi(\lambda x)=\lambda\varphi(x) (5.3)

for all xnx\in\mathbb{R}^{n} and any λ0\lambda\geq 0. Since |φ(x)||\nabla\varphi(x)| is finite near the origin, hence |φ(o)|=lim|x|0|xφ(x)|=0|\varphi(o)|=\lim_{|x|\to 0}|x\cdot\nabla\varphi(x)|=0, which implies that (5.3) holds for λ=0\lambda=0. Otherwise, if φ(λx)λφ(x)\varphi(\lambda x)\neq\lambda\varphi(x) for some points xox\neq o, where φ\varphi is differentiable, then by the convexity of φ(x)\varphi(x), one has

φ(y)φ(x)φ(x)(yx)\displaystyle\varphi(y)-\varphi(x)\geq\nabla\varphi(x)\cdot(y-x)

for all yny\in\mathbb{R}^{n}. Note that the above inequality is strictly when y=oy=o due to the assumptions φ(x)>φ(o)\varphi(x)>\varphi(o) and φ(λx)λφ(x)\varphi(\lambda x)\neq\lambda\varphi(x). Then by taking y=oy=o, we have φ(x)<φ(x)x\varphi(x)<\nabla\varphi(x)\cdot x, which contradicts with (5.2). Thus φ(λx)=λφ(x)\varphi(\lambda x)=\lambda\varphi(x) at every differentiable points. Moreover, this also holds on whole n\mathbb{R}^{n} as φ\varphi is continuous and φ\varphi is differentiable almost everywhere. Combined with the convexity of φ\varphi, one sees φ\varphi is sublinear. By the fact that every sublinear function from n\mathbb{R}^{n} to \mathbb{R} is a support function of a convex body, we have φ(x)=hK(x)\varphi(x)=h_{K}(x) for some convex body KK. Since φ(x)>φ(o)=0\varphi(x)>\varphi(o)=0 for all oxno\neq x\in\mathbb{R}^{n}, one can see that KK is a convex body that contains the origin in its interior. In this case, one has φ(x)=hK(x)=xK\varphi(x)=h_{K}(x)=\|x\|_{K^{\circ}} and (5.2) follows. ∎

For a real function τ:(0,+)R\tau:(0,\ +\infty)\rightarrow R, let

0τ(t)el𝑑l=1πtel2𝑑l.\displaystyle\int_{0}^{\tau(t)}e^{-l}dl=\frac{1}{\sqrt{\pi}}\int_{-\infty}^{t}e^{-l^{2}}dl. (5.4)

The transformation T:n+1n+1T:\mathbb{R}^{n+1}\rightarrow\mathbb{R}^{n+1} is defined by

Ty=nh(x)τ(yh(x))hi(x)𝑑ν(x),\displaystyle Ty=\int_{\mathbb{R}^{n}}h(x)\frac{\tau(y\cdot h(x))}{h_{i}(x)}d\nu(x),

for yn+1y\in\mathbb{R}^{n+1}, with hi=heih_{i}=h\cdot e_{i} for i=1,,n+1i=1,\cdots,n+1.

By the Ball-Barth inequality, we obtain the following lemma.

Lemma 5.3.

Let (n,ν)(\mathbb{R}^{n},\nu) be a Borel measure space and h:nSnh:\mathbb{R}^{n}\to S^{n} be an isotropic embedding of the Borel measure space (n,ν)(\mathbb{R}^{n},\nu) into SnS^{n}. Then

(n+1)n+12T(n+1)exi𝑑x,\displaystyle(n+1)^{\frac{n+1}{2}}\leq\int_{T(\mathbb{R}^{n+1})}e^{-x_{i}}dx, (5.5)

with equality if and only if hih_{i} is constant on supp(ν){\rm supp}(\nu), and there exists a C>0C>0 with respect to yy such that

j=1n+1τ(yh(xi))=C\prod_{j=1}^{n+1}{\tau^{\prime}(y\cdot h(x_{i}))}=C

for x1,,xn+1supp(ν)x_{1},\cdots,x_{n+1}\in{\rm supp}(\nu) with h(x1),,h(xn+1)h(x_{1}),...,h(x_{n+1}) linearly independent.

Proof.

Deriving both sides of (5.4) with respect to tt, we have

τ(t)+logτ(t)=logπt2.\displaystyle-\tau(t)+\log\tau^{\prime}(t)=-\log\sqrt{\pi}-t^{2}.

Taking t=yh(x)t=y\cdot h(x), one has

|yh(x)|2=logπτ(yh(x))+logτ(yh(x))hi(x)+loghi(x).\displaystyle-|y\cdot h(x)|^{2}=\log\sqrt{\pi}-\tau(y\cdot h(x))+\log\frac{\tau^{\prime}(y\cdot h(x))}{h_{i}(x)}+\log h_{i}(x). (5.6)

Now we integrate (5.6) over all xnx\in\mathbb{R}^{n} with respect to the measure dνd\nu. From the definition of isotropic embedding (3.6), the integral of the left hand side on (5.6) is equal to

n|yh(x)|2|dν(x)=|y|2.\displaystyle-\int_{\mathbb{R}^{n}}|y\cdot h(x)|^{2}|d\nu(x)=-|y|^{2}. (5.7)

Next we deal with the integral of the last term on the right hand side of (5.6). From (3.7), one can see that the measure 1n+1dν\frac{1}{n+1}d\nu is a probability measure. By the fact that the L0L_{0}-mean of a function is dominated by its L2L_{2}-mean on a probability space and (3.6), one has

exp(1n+1nloghi(x)𝑑ν(x))(1n+1n|hi(x)|2𝑑ν(x))12=(1n+1)12,\displaystyle\exp\left(\frac{1}{n+1}\int_{\mathbb{R}^{n}}\log h_{i}(x)d\nu(x)\right)\leq\left(\frac{1}{n+1}\int_{\mathbb{R}^{n}}|h_{i}(x)|^{2}d\nu(x)\right)^{\frac{1}{2}}=\left(\frac{1}{n+1}\right)^{\frac{1}{2}}, (5.8)

with equality if and only if hi(x)h_{i}(x) is constant on supp(ν){\rm supp}(\nu). Hence

nloghi(x)dν(x)log(1n+1)n+12.\displaystyle\int_{\mathbb{R}^{n}}\log h_{i}(x)d\nu(x)\leq\log\left(\frac{1}{n+1}\right)^{\frac{n+1}{2}}.

Calculating the derivative of TyTy with respect to yy, we have

dTy=nh(x)h(x)τ(yh(x))hi(x)𝑑ν(x).\displaystyle dTy=\int_{\mathbb{R}^{n}}h(x)\otimes h(x)\frac{\tau^{\prime}(y\cdot h(x))}{h_{i}(x)}d\nu(x).

From Proposition 3.1, we infer that

det(dTy)\displaystyle\det(dTy) \displaystyle\geq exp{nlogτ(yh(x))hi(x)dν(x)},\displaystyle\exp\left\{\int_{\mathbb{R}^{n}}\log\frac{\tau^{\prime}(y\cdot h(x))}{h_{i}(x)}d\nu(x)\right\}, (5.9)

with equality if and only if

j=1n+1τ(yh(xj))hi(xj)\displaystyle\prod_{j=1}^{n+1}\frac{\tau^{\prime}(y\cdot h(x_{j}))}{h_{i}(x_{j})}

is constant for x1,,xn+1supp(μ)x_{1},\cdots,x_{n+1}\in{\rm supp}(\mu) such that h(x1),,h(xn+1)h(x_{1}),\cdots,h(x_{n+1}) are linearly independent.

Combining (5.6), (5.7), (5.8), (5.9) and the fact that ν(n)=n+1\nu(\mathbb{R}^{n})=n+1, we have

exp{|y|2}(πn+1)n+12det(d(Ty))exp{eiTy}.\displaystyle\exp\{-|y|^{2}\}\leq\left(\frac{\pi}{n+1}\right)^{\frac{n+1}{2}}\det(d(Ty))\exp\{-e_{i}\cdot Ty\}.

Integrating this inequality over all yn+1y\in\mathbb{R}^{n+1} gives the desired inequality (5.5).∎

Now we are ready to prove our functional isoperimetric inequality for log-concave functions.

Theorem 5.1.

Let f=eφLCnf=e^{-\varphi}\in{\rm LC}_{n} and μf\mu_{f} be the surface area measure of ff. Suppose that φ(x)>0\varphi^{\ast}(x)>0 on n\{o}\mathbb{R}^{n}\backslash\{o\} and logt=sup{φ(x/φ(x)):xsupp(μf)}-\log t=\sup\{\varphi^{\ast}(x/\varphi^{\ast}(x)):x\in{\rm supp}(\mu_{f})\}, then

|Kt(f)|J(Γ2f)8n2(n+1)n+12Γ(n2+1)ωnn!nn.\displaystyle|K_{t}(f^{\circ})|J(\Gamma_{-2}f)\geq\frac{8^{\frac{n}{2}}(n+1)^{\tfrac{n+1}{2}}\Gamma(\frac{n}{2}+1)\omega_{n}}{n!n^{n}}. (5.10)

Moreover, if φ(x)>φ(o)\varphi(x)>\varphi(o) for all xn\{o}x\in\mathbb{R}^{n}\backslash\{o\}, then equality holds if and only if f(x)=exK+1f(x)=e^{-{\|x\|_{K}}+1} and KK is a simplex whose centroid is at the origin.

Proof.

From Property 4.1, (2.3) and the fact that logΓ2f(x)\sqrt{-\log\Gamma_{-2}f(x)} is a Minkowski function of an ellipsoid, we only need to prove (5.10) holds for Γ2f(x)=e|x|2/2\Gamma_{-2}f(x)=e^{-{|x|^{2}}/{2}}. This means that the measure

ν()=n24δJ(f,f)(φ)1μf()\nu(\cdot)=\frac{n^{2}}{4\delta J(f,f)}(\varphi^{*})^{-1}\mu_{f}(\cdot)

is isotropic. Let cn=2/nc_{n}={2/n}. For any xnx\in\mathbb{R}^{n}, define a function h:nn+1h:\mathbb{R}^{n}\rightarrow\mathbb{R}^{n+1} by h(x)=(x,cnφ(x)),h(x)=(x,c_{n}\varphi^{\ast}(x)), and h¯:nSn\bar{h}:\mathbb{R}^{n}\rightarrow S^{n} by h¯=h/|h|.\bar{h}={h}/{|h|}. Assume that y=(z,s)n+1y=(z,s)\in\mathbb{R}^{n+1}. By the fact that the barycenter of μf\mu_{f} is at the origin, we have

n|yh¯(x)|2|h(x)|2𝑑ν(x)\displaystyle\int_{\mathbb{R}^{n}}|y\cdot\bar{h}(x)|^{2}|h(x)|^{2}d\nu(x)
=n|(z,s)(x,cnφ(x))|2𝑑ν(x)\displaystyle\quad\quad=\int_{\mathbb{R}^{n}}|(z,s)\cdot(x,c_{n}\varphi^{\ast}(x))|^{2}d\nu(x)
=n|zx|2𝑑ν(x)+2scnznx𝑑μf(x)+4s2n2n(φ)2𝑑ν(x)\displaystyle\quad\quad=\int_{\mathbb{R}^{n}}|z\cdot x|^{2}d\nu(x)+2sc_{n}z\cdot\int_{\mathbb{R}^{n}}xd\mu_{f}(x)+\frac{4s^{2}}{n^{2}}\int_{\mathbb{R}^{n}}(\varphi^{\ast})^{2}d\nu(x)
=|z|2+s2\displaystyle\quad\quad=|z|^{2}+s^{2}
=|y|2.\displaystyle\quad\quad=|y|^{2}.

Hence h¯:nSn\bar{h}:\mathbb{R}^{n}\rightarrow S^{n} is an isotropic embedding of the Borel measure space (n,|h|2dν)(\mathbb{R}^{n},|h|^{2}d\nu) into SnS^{n}. From Lemma 5.3, one can see that

(n+1)n+12n+1een+1Ty|dTy|𝑑y=T(n+1)een+1z𝑑z.\displaystyle(n+1)^{\frac{n+1}{2}}\leq\int_{\mathbb{R}^{n+1}}e^{-e_{n+1}\cdot Ty}|dTy|dy=\int_{T(\mathbb{R}^{n+1})}e^{-e_{n+1}\cdot z}dz. (5.11)

Here TyTy is given by

Ty\displaystyle Ty =nh¯(x)τ(yh¯(x))en+1h¯(x)|h(x)|2𝑑ν(x)\displaystyle=\int_{\mathbb{R}^{n}}\bar{h}(x)\frac{\tau(y\cdot\bar{h}(x))}{e_{n+1}\cdot\bar{h}(x)}|h(x)|^{2}d\nu(x)
=nh(x)τ(yh¯(x))en+1h(x)|h(x)|2𝑑ν(x)\displaystyle=\int_{\mathbb{R}^{n}}h(x)\frac{\tau(y\cdot\bar{h}(x))}{e_{n+1}\cdot h(x)}|h(x)|^{2}d\nu(x)
=n((cnφ(x))1x,1)τ(yh¯(x))|h(x)|2𝑑ν(x).\displaystyle=\int_{\mathbb{R}^{n}}((c_{n}\varphi^{\ast}(x))^{-1}x,1)\tau(y\cdot\bar{h}(x))|h(x)|^{2}d\nu(x).

Lemma 5.1 implies that

Tyr>0cn1rKt(f)×{r}=:D.\displaystyle Ty\in\bigcup_{r>0}c_{n}^{-1}rK_{{t}}(f^{\circ})\times\{r\}=:D.

Therefore,

T(n+1)een+1z𝑑z\displaystyle\int_{T(\mathbb{R}^{n+1})}e^{-e_{n+1}\cdot z}dz Deen+1z𝑑z=0cn1rKter𝑑x𝑑r\displaystyle\leq\int_{D}e^{-e_{n+1}\cdot z}dz=\int_{0}^{\infty}\int_{c_{n}^{-1}rK_{t}}e^{-r}dxdr
=nn2n0rner𝑑r|Kt(f)|=nn2nΓ(n+1)|Kt(f)|.\displaystyle=\frac{n^{n}}{2^{n}}\int_{0}^{\infty}r^{n}e^{-r}dr|K_{t}(f^{\circ})|=\frac{n^{n}}{2^{n}}\Gamma(n+1)|K_{t}(f^{\circ})|.

This together with the fact that J(e|x|22)=2n2Γ(n2+1)ωnJ(e^{-\frac{|x|^{2}}{2}})=2^{\frac{n}{2}}\Gamma(\frac{n}{2}+1)\omega_{n} gives (5.10).

From the equality condition of inequality (5.5), we have that φ\varphi^{\ast} is a positive constant on supp(τ(yh¯)|h|2ν){\rm supp}(\tau(y\cdot\overline{h})|h|^{2}\nu), that is, φ=c0>0\varphi^{\ast}=c_{0}>0 on supp(μf){\rm supp}(\mu_{f}) due to τ>0\tau>0 on \mathbb{R} and |h(x)|2>0|h(x)|^{2}>0 on n+1\mathbb{R}^{n+1}. Since dom(φ)=n{\rm dom}(\varphi)=\mathbb{R}^{n}, one can see that

φ(φ(x))=xφ(x)φ(x)=c0\displaystyle\varphi^{\ast}(\nabla\varphi(x))=x\cdot\nabla\varphi(x)-\varphi(x)=c_{0} (5.12)

almost everywhere on n\mathbb{R}^{n}. If φ(x)>φ(o)\varphi(x)>\varphi(o) for any xn\{o}x\in\mathbb{R}^{n}\backslash\{o\}, Lemma 5.2 implies that φ(x)=xKc0\varphi(x)=\|x\|_{K}-c_{0} for some convex body K𝒦onK\in\mathcal{K}^{n}_{o}. In this case, by (4.4), one can calculate that

δJ(f,f)=c0ec0nΓ(n)|K|\displaystyle\delta J(f,f)=c_{0}e^{c_{0}}n\Gamma(n)|K| (5.13)

and

n|xφ(y)|2φ(φ(y))1eφ(y)𝑑y\displaystyle\int_{\mathbb{R}^{n}}|x\cdot\nabla\varphi(y)|^{2}\varphi^{*}(\nabla\varphi(y))^{-1}e^{-\varphi(y)}dy
=c01ec0n|xyK|2eyK𝑑y\displaystyle\quad\quad=c_{0}^{-1}e^{c_{0}}\int_{\mathbb{R}^{n}}|x\cdot\nabla\|y\|_{K}|^{2}e^{-\|y\|_{K}}dy
=c01ec0n|K|0K|xzK|2rn1erdrdσK(z)\displaystyle\quad\quad=c_{0}^{-1}e^{c_{0}}n|K|\int_{0}^{\infty}\int_{\partial K}|x\cdot\nabla\|z\|_{K}|^{2}r^{n-1}e^{-r}drd\sigma_{K}(z)
=c01ec0Γ(n)n|K|K|xzK|2drdσK(z)\displaystyle\quad\quad=c_{0}^{-1}e^{c_{0}}\Gamma(n)n|K|\int_{\partial K}|x\cdot\nabla\|z\|_{K}|^{2}drd\sigma_{K}(z)
=c01ec0Γ(n)|K|xΓ2K2.\displaystyle\quad\quad=c_{0}^{-1}e^{c_{0}}\Gamma(n)|K|\|x\|_{\Gamma_{-2}K}^{2}.

This together with the definition of logΓ2(f)-\log\Gamma_{-2}(f) and (5.13) shows that

logΓ2(exK+c0)=n8c02xΓ2K2.\displaystyle-\log\Gamma_{-2}(e^{-\|x\|_{K}+c_{0}})=\frac{n}{8c_{0}^{2}}\|x\|_{\Gamma_{-2}K}^{2}.

For any c>0c>0 and convex body KK,

J(ecxK2)\displaystyle J(e^{-c\|x\|_{K}^{2}}) =necxK2𝑑x\displaystyle=\int_{\mathbb{R}^{n}}e^{-c\|x\|_{K}^{2}}dx
=n|K|0Krn1ecr2𝑑r𝑑σK(x)\displaystyle=n|K|\int_{0}^{\infty}\int_{\partial K}r^{n-1}e^{-cr^{2}}drd\sigma_{K}(x)
=cn2Γ(n2+1)|K|.\displaystyle=c^{-\tfrac{n}{2}}\Gamma(\tfrac{n}{2}+1)|K|.

Hence

J(Γ2(exK+c))=8n2c0nnn2Γ(n2+1)|Γ2K|.\displaystyle J(\Gamma_{-2}(e^{-\|x\|_{K}+c}))=\frac{8^{\frac{n}{2}}c_{0}^{n}}{n^{\frac{n}{2}}}\Gamma(\tfrac{n}{2}+1)|\Gamma_{-2}K|.

By the definition of Legendre transform and φ(x)=xKc0\varphi(x)=\|x\|_{K}-c_{0}, we have

φ(x)={c0,xK;+,xK.\varphi^{\ast}(x)=\left\{\begin{aligned} &\ c_{0},&x\in K^{\circ};\\ &+\infty,&x\notin K^{\circ}.\end{aligned}\right.

For any xsupp(μf)Kx\in{\rm supp}(\mu_{f})\subset K^{\circ} and c01c_{0}\geq 1, we obtain

logt=sup{φ(xφ(x)):xsupp(μf)}=c0.-\log t=\sup\left\{\varphi^{\ast}\left(\frac{x}{\varphi^{\ast}(x)}\right):x\in{\rm supp}{(\mu_{f})}\right\}=c_{0}.

Hence

Kt(f)={xn:φ(x)c0}=K\displaystyle K_{t}(f^{\circ})=\{x\in\mathbb{R}^{n}:\varphi^{\ast}(x)\leq c_{0}\}=K^{\circ}

and the inequality (5.10) becomes

|K||Γ2K|(n+1)n+12ωnc0nn!nn2.\displaystyle|K^{\circ}||\Gamma_{-2}K|\geq\frac{(n+1)^{\tfrac{n+1}{2}}\omega_{n}}{c_{0}^{n}n!n^{\frac{n}{2}}}. (5.14)

Let conv(A)conv(A) denote the convex hull of AnA\subset\mathbb{R}^{n} and A¯\overline{A} denote the closure of AA. For the surface area μf\mu_{f} of the log-concave function, [18, (16)] says that

conv(supp(μf))={φ<+}¯,{\rm conv}({\rm supp}(\mu_{f}))=\overline{\{\varphi^{\ast}<+\infty\}},

therefore conv(supp(μf))=K{\rm conv}({\rm supp}(\mu_{f}))=K^{\circ}.

For x(supp(μf))(K)x\in\partial({\rm supp}(\mu_{f}))\subset\partial(K^{\circ}) and c0<1c_{0}<1, we have x/c0Kx/c_{0}\notin K^{\circ} and logt=+-\log t=+\infty. Therefore |Kt(f)|=+|K_{t}(f^{\circ})|=+\infty and the inequality (5.10) is strict. Hence the best constant of (5.14) is c0=1c_{0}=1 and the inequality (5.14) becomes the LYZ polar inequality (1.4). Therefore the equality of (5.14) holds if and only if c0=1c_{0}=1 and KK is a simplex with centroid at the origin. ∎


References

  • [1] D. Alonso-Gutiérrez, J. Bernués and B.G. Merino, Zhang’s inequality for log-concave functions, In Geometric Aspects of Functional Analysis (pp. 29-48), Springer, Cham, 2020.
  • [2] D. Alonso-Gutiérrez, B.G. Merino, C.H. Jiménez and R. Villa, Rogers-Shephard inequality for log-concave functions, J. Funct. Anal., 271 (2016), 3269-3299.
  • [3] D. Alonso-Gutiérrez, B.G. Merino, C.H. Jiménez and R. Villa, John’s ellipsoid and the integral ratio of a log-concave function, J. Geom. Anal., 28 (2018), 1182-1201.
  • [4] S. Artstein-Avidan, D.I. Florentin and A. Segal, Functional Brunn-Minkowski inequalities induced by polarity, Adv. Math., 364 (2020), 107006.
  • [5] S. Artstein-Avidan, B. Klartag and V.D. Milman, The Santaló point of a function, and a functional form of Santaló inequality, Mathematika, 51 (2004), 33-48.
  • [6] S. Artstein-Avidan, B. Klartag, C. Schütt and E.M. Werner, Functional affine-isoperimetry and an inverse logarithmic Sobolev inequality, J. Funct. Anal., 262 (2012), 4181-4204.
  • [7] K. Ball, Isometric problems in LpL_{p} and sections of convex sets, PhD dissertation, Cambridge, 1986.
  • [8] K. Ball, Volume ratios and a reverse isoperimetric inequality, J. London Math. Soc., 44 (1991), 351-359.
  • [9] F. Barthe, On a reverse form of the Brascamp-Lieb inequality, Invent. Math., 134 (1998), 335-361.
  • [10] F. Barthe, K.J. Böröczky and M. Fradelizi, Stability of the functional forms of the Blaschke-Santaló inequality, Monatsh. Math., 173 (2014), 135-159.
  • [11] K. Böröczky and A. De, Stability of the Prékopa-Leindler inequality for log-concave functions, Adv. Math., 386 (2021), 107810.
  • [12] U. Caglar, M. Fradelizi, O. Guedon, J. Lehec, C. Schütt and E.M. Werner, Functional versions of LpL_{p}-affine surface area and entropy inequalities, Int. Math. Res. Not., 2016 (2016), 1223-1250.
  • [13] U. Caglar and D. Ye, Affine isoperimetric inequalities in the functional Orlicz-Brunn-Minkowski theory, Adv. Appl. Math., 81 (2016), 78-114.
  • [14] S. Campi and P. Gronchi, The LpL_{p}-Busemann-Petty centroid inequality, Adv. Math., 167 (2002), 128-141.
  • [15] S. Campi and P. Gronchi, On the reverse LpL_{p}-Busemann-Petty centroid inequality, Mathematika, 49 (2002), 1-11.
  • [16] A. Colesanti, Functional inequalities related to the Rogers-Shephard inequality, Mathematika, 53 (2006), 81-101.
  • [17] A. Colesanti and I. Fragalà, The first variation of the total mass of log-concave functions and related inequalities, Adv. Math., 244 (2013), 708-749.
  • [18] D. Cordero-Erausquin and B. Klartag, Moment measures, J. Funct. Anal., 268 (2015), 3834-3866.
  • [19] N. Fang, S. Xing and D. Ye, Geometry of log-concave functions: the LpL_{p} Asplund sum and the LpL_{p} Minkowski problem, Calc. Var. Partial Differ. Equ., 21 (2022), Art. 45.
  • [20] N. Fang and J. Zhou, LYZ ellipsoid and Petty projection body for log-concave functions, Adv. Math., 340 (2018), 914-959.
  • [21] N. Fang and J. Zhou, The Busemann-Petty problem on entropy of log-concave functions, Sci. China Math., 65 (2022), 2171-2182.
  • [22] M. Fradelizi and M. Meyer, Some functional forms of Blaschke-Santaló inequality, Math. Z., 256 (2007), 379-395.
  • [23] M. Fradelizi and M. Meyer, Some functional inverse Santaló inequality, Adv. Math., 218 (2008), 1430-1452.
  • [24] R. J. Gardner, Geometric tomography, Cambridge Univ. Press, Cambridge, 1995.
  • [25] A. A. Giannopoulos and V. D. Milman, Extremal problems and isotropic positions of convex bodies, Israel J. Math., 117 (2000), 29-60.
  • [26] A. A. Giannopoulos and M. Papadimitrakis, Isotropic surface area measures, Mathematika, 46 (1999), 1-13.
  • [27] C. Haberl and F. Schuster, General LpL_{p} affine isoperimetric inequalities, J. Differential Geom., 83 (2009), 1-26.
  • [28] H. Iriyeh and M. Shibata, Symmetric Mahler¡¯s conjecture for the volume product in the 33-dimensional case, Duke Math. J., 169 (2020), 1077-1134.
  • [29] G. Ivanov and M. Naszódi, Functional John ellipsoids, J. Funct. Anal., 282 (2022), 109441.
  • [30] B. Klartag and V.D. Milman, Geometry of log-concave functions and measures, Geom. Dedicata, 112 (2005), 169-182.
  • [31] B. Li, C. Schütt and E.M. Werner, Floating functions, Israel J. Math., 231 (2019), 181-210.
  • [32] B. Li, C. Schütt and E.M. Werner, The Löwner function of a log-concave function, J. Geom. Anal., 31 (2021), 423-456.
  • [33] Y. Lin, Affine Orlicz Pólya-Szegö principle for log-concave functions, J. Funct. Anal., 273 (2017), 3295-3326.
  • [34] E. Lutwak, D. Yang and G. Zhang, LpL_{p} affne isoperimetric inequalities, J. Differential Geom., 56 (2000), 111-132.
  • [35] E. Lutwak, D. Yang and G. Zhang, A new ellipsoid associated with convex bodies, Duke Math. J., 104 (2000), 375-390.
  • [36] E. Lutwak, D. Yang and G. Zhang, Volume inequalities for subspaces of LpL_{p}, J. Differential Geom., 68 (2004), 159-184.
  • [37] E. Lutwak, D. Yang and G. Zhang, Volume inequalities for isotropic measures, Amer. J. Math., 129 (2007), 1711-1723.
  • [38] E. Lutwak, D. Yang and G. Zhang, A volume inequality for polar bodies, J. Differential Geom., 84 (2010), 163-178.
  • [39] E. Lutwak and G. Zhang, Blaschke-Santaló inequalities, J. Differential Geom., 45 (1997), 1-16.
  • [40] K. Mahler, Ein Minimalproblem für konvexe Polygone, Mathematica (Zutphen), B7 (1939), 118-127.
  • [41] M. Meyer and E. Werner, On the pp-affine surface area, Adv. Math., 152 (2000), 288-313.
  • [42] V.D. Milman and L. Rotem, Mixed integrals and related inequalities, J. Funct. Anal., 264 (2013), 570-604.
  • [43] T. Rockafellar, Convex Analysis, Princeton University Press, Princeton, (1970).
  • [44] L. Rotem, The anisotropic total variation and surface area measures, arXiv preprint arXiv:2206.13146 (2022).
  • [45] R. Schneider, Convex bodies: the Brunn-Minkowski theory, Cambridge University Press, 2014.
  • [46] F. Schuster and M. Weberndorfer, Volume inequalities for asymmetric Wulff shapes, J. Differential Geom., 92 (2012): 263-283.
  • [47] B. Zhu, H. Hong and D. Ye, The Orlicz-Petty bodies, Interna. Math. Res. Notices, 14 (2018), 4356-4403.
  • [48] D. Zou and G. Xiong, Orlicz-John ellipsoids, Adv. Math., 265 (2014), 132-168.
  • [49] D. Zou and G. Xiong, Orlicz-Legendre ellipsoids, J. Geom. Anal., 26 (2016), 2474-2502.