In this paper, the geominimal integral curvature on the convex body is introduced. The existence and uniqueness of the geominimal integral curvature are proved. Some other properties for the geominimal integral curvature, such as continuity, are investigated.
Citation: Shuang Mou. The geominimal integral curvature[J]. AIMS Mathematics, 2022, 7(8): 14338-14353. doi: 10.3934/math.2022790
[1] | Tian Gao, Dan Ma . The $ L_p $-Petty projection inequality for $ s $-concave functions. AIMS Mathematics, 2025, 10(4): 7706-7716. doi: 10.3934/math.2025353 |
[2] | Haiming Liu, Xiawei Chen, Jianyun Guan, Peifu Zu . Lorentzian approximations for a Lorentzian $ \alpha $-Sasakian manifold and Gauss-Bonnet theorems. AIMS Mathematics, 2023, 8(1): 501-528. doi: 10.3934/math.2023024 |
[3] | Kai Wang, Xiheng Wang, Xiaoping Wang . Curvature estimation for point cloud 2-manifolds based on the heat kernel. AIMS Mathematics, 2024, 9(11): 32491-32513. doi: 10.3934/math.20241557 |
[4] | Gülnur Şaffak Atalay . A new approach to special curved surface families according to modified orthogonal frame. AIMS Mathematics, 2024, 9(8): 20662-20676. doi: 10.3934/math.20241004 |
[5] | Youngjin Hwang, Jyoti, Soobin Kwak, Hyundong Kim, Junseok Kim . An explicit numerical method for the conservative Allen–Cahn equation on a cubic surface. AIMS Mathematics, 2024, 9(12): 34447-34465. doi: 10.3934/math.20241641 |
[6] | Salah G. Elgendi, Zoltán Muzsnay . The geometry of geodesic invariant functions and applications to Landsberg surfaces. AIMS Mathematics, 2024, 9(9): 23617-23631. doi: 10.3934/math.20241148 |
[7] | Hyun Geun Lee, Youngjin Hwang, Yunjae Nam, Sangkwon Kim, Junseok Kim . Benchmark problems for physics-informed neural networks: The Allen–Cahn equation. AIMS Mathematics, 2025, 10(3): 7319-7338. doi: 10.3934/math.2025335 |
[8] | Samah Gaber, Asmahan Essa Alajyan, Adel H. Sorour . Construction and analysis of the quasi-ruled surfaces based on the quasi-focal curves in $ \mathbb{R}^{3} $. AIMS Mathematics, 2025, 10(3): 5583-5611. doi: 10.3934/math.2025258 |
[9] | Özgür Boyacıoğlu Kalkan, Süleyman Şenyurt, Mustafa Bilici, Davut Canlı . Sweeping surfaces generated by involutes of a spacelike curve with a timelike binormal in Minkowski 3-space. AIMS Mathematics, 2025, 10(1): 988-1007. doi: 10.3934/math.2025047 |
[10] | Xudong Wang, Tingting Xiang . Dual Brunn-Minkowski inequality for $ C $-star bodies. AIMS Mathematics, 2024, 9(4): 7834-7847. doi: 10.3934/math.2024381 |
In this paper, the geominimal integral curvature on the convex body is introduced. The existence and uniqueness of the geominimal integral curvature are proved. Some other properties for the geominimal integral curvature, such as continuity, are investigated.
The geominimal surface area belongs to the research category of convex geometric analysis. The classical geominimal surface area, which can be dated back to 1974, was firstly introduced by Petty [21], in his paper, the classical Petty body was obtained and some classical affine isoperimetric inequalities were established. With the development of the Lp-Brunn-Minkowski theory, the classical geominimal surface area has been extended to Lp cases by Lutwak [14] (for p>1) and Ye [26] (for −n≠p<1) (also see [28,30]). The dual Lp-Brunn-Minkowski theory was introduced by Lutwak [13,15]. Wang and Chen [25] introduced the dual Lp geominimal surface area. For more results of the geominimal surface area, one can refer to [9,10,11,18,27,29,31] and so on.
We shall work in the n-dimensional Euclidean space Rn with the standard inner product x⋅y of x and y in Rn. For x∈Rn, write |x|=√x⋅x for the Euclidean norm of x. We call a set K⊂Rn is convex if provided that for any two points x,y∈K and λ∈[0,1], one has λx+(1−λ)y∈K. A convex subset K⊂Rn is called a convex body if K is compact with nonempty interior, and the interior point of a convex body K can be written as intK. Moreover, if for any two points x,y∈K (x≠y) and λ∈(0,1), λx+(1−λ)y∈intK holds, we call the convex body K is a strictly convex.
Throughout this paper, let Kn be the class of all convex bodies in Rn and Kn0 be the class of all convex bodies in Rn that contain the origin in their interiors and Kne be the class of all origin-symmetric convex bodies in Rn. The standard Lebesgue measure of a set K in Rn will be denoted by |K|, we call |K| is the n-dimensional volume of a set K∈Kn. The volume radius of the convex body K is defined as vrad(K)=(|K|/ωn)1/n. The origin-symmetric unit ball in n-dimensional Euclidean space is denote by Bn2, i.e., Bn2={x∈Rn:|x|≤1}, and we use ωn to denote the volume of the unit ball Bn2. We use Sn−1 to denote the unit sphere in Rn, and we let κn be the surface area measure of Bn2. Let K|u⊥ be the orthogonal projection of K onto the subspace orthogonal to u and Vn−1(K|u⊥) be the (n−1)-dimensional volume of K|u⊥. Let C(Sn−1) be the set of continuous functions defined on the unit sphere Sn−1.
As we all know, the notion of geominimal surface area was introduced by Petty [21]. For K∈Kn0, the geominimal surface area of K, is defined by the following optimal problem:
inf{∫Sn−1hQ(u)dS(K,u):Q∈Knsand |Q∗|=ωn}. | (1.1) |
Here Q∗ denotes polar body of Q, hQ(⋅) is the support function of convex body Q (see Section 2), Kns is the class of convex bodies in Rn whose Santalˊo points are at origin point o (see [23]), and S(K,⋅) is the surface area measure of convex body K, i.e., for any K∈Kn and any measurable subset Ω⊂Sn−1, the surface area measure S(K,Ω), is defined by
S(K,Ω)=∫ν−1K(Ω)dHn−1, |
where ν−1K:Sn−1→∂K is the inverse Gauss map and Hn−1 is the (n−1)-dimensional Hausdorff measure on ∂K.
In [16], Lutwak extended over body Q restricted to Kn0. It follows that
inf{∫Sn−1hQ(u)dS(K,u):Q∈Kn0and |Q∗|=ωn}. | (1.2) |
For any two convex bodies K,Q∈Kn, the mixed volume V1(K,Q) of K,Q is defined by
V1(K,Q)=1n∫Sn−1hQ(u)dS(K,u). |
Moreover, for any K∈Kn, the volume of K is defined by
|K|=1n∫Sn−1hK(u)dS(K,u). | (1.3) |
Together with (1.2), one has,
inf{nV1(K,Q):Q∈Kn0and |Q∗|=ωn}. | (1.4) |
Lutwak [16] proved the uniqueness of the solution to (1.4). This shows that the mixed volume V1(⋅,⋅) is a necessary premise of classical geominimal surface area. Hence, utilizing the relationship between the minimum surface area and the corresponding mixed volume, one can investigate the geominimal integral curvature with the help of the mixed entropy.
We define a new concept of the entropy functional of the convex body, i.e., the entropy functional E(K) of the convex body K∈Kn0 can be defined by
E(K)=−∫Sn−1loghK(u)du, |
where the integration is with respect to spherical Lebesgue measure (see, Huang et al.[7]).
For two convex bodies K,L∈Kn0, the Lp mixed entropy is defined by
Ep(K,L)=limε→0E(Kˆ+pε⋄L)−E(K)ε, |
where Kˆ+pε⋄L is the harmonic Lp combination of K,L for p≥1 and Kˆ+pε⋄L∈Kn0 is a new convex body. In Section 2, we will show that the convex body Kˆ+pε⋄L∈Kn0 can be defined for all p∈R and even for negative ϵ of sufficiently small absolute value. In this paper, we mainly study the mixed entropy functional when p=0, i.e., for K,L∈Kn0
E0(K,L)=limε→0E(Kˆ+0ε⋄L)−E(K)ε=−∫Sn−1logρL(u)dJ(K,u), | (1.5) |
where ρL is the radial function of convex body L and J(K,⋅) is the Aleksandrov integral curvature of the convex body K (see Section 2). For some of the the Aleksandrov integral recent resilts, see [1,3,8,19,20,22].
There exists a natural problem: whether there is a convex body L∈Kn0 with |L∗|=ωn such that L is a solution to the following problem:
inf{E0(K,Q):Q∈Kn0 with |Q∗|=ωn}. | (1.6) |
In this paper, let GE(K) be the geominimal integral curvature of the convex body K∈Kn0, it can be defined by optimal problem in (1.6). Together with (1.5), it is equivalent to solve the optimal problem as follows:
GE(K)=sup{∫Sn−1logρQ(u)dJ(K,u):Q∈Kn0,|Q∗|=ωn}. |
Based on the concept of the Petty body, we will study the entropy form of the Petty body. The following is our main result.
Theorem 1.1. Let K∈Kno, then there exists a convex body M∈Kne with |M∗|=ωn such that
∫Sn−1logρM(u)dJ(K,u)=sup{∫Sn−1logρL(u)dJ(K,u):L∈Kne,|L∗|=ωn}. |
In addition, in the plane R2, these bodies are unique or their polar bodies are parallelograms with parallel sides.
This paper is organized as follows. In Section 2, we collect some basic concepts and various facts that will be used in the proofs of our results. In Section 3, we study the properties of the L0 mixed entropy. In Section 4, we obtain the existence and uniqueness of the geominimal integral curvature.
We now introduce some basic facts and standard notations needed in this paper. For more details and concepts in convex geometry, please see [4,5,24].
The Minkowski sum of two convex sets K and L is denoted by
K+L={x+y:x∈K, y∈L}. |
The scalar product of λ∈R and K∈Kn is defined by λK={λx:x∈K}. For any convex body K in Rn, the support function of K, hK:Sn−1→R, is defined by
hK(u)=max{x⋅u:x∈K}, for all u∈Sn−1. |
For the convex body K and u∈Sn−1, the hyperplane
HK(u)={x∈Rn:x⋅u=hK(u)} |
is called the supporting hyperplane of K with unit normal u. For x∈∂K, if there is only one supporting hyperplane of K passing through point x, we call x∈∂K is a smooth point. If there exist more than a supporting hyperplane of K passing through point x, we call x∈∂K is a singlar point. For each x∈∂K is a smooth point, we call the convex body K is smooth. If a convex body K is strictly convex and smooth, we say that K is a regular convex body. For each K∈Kn, there exists a regular convex body sequence Ki such that Ki converges to K as i→∞ (see [24]).
Let lu={tu:t≥0} for u∈Sn−1. If L∩lu is a closed line segment for all u∈Sn−1, we say L⊂Rn is star-shaped with respect to the origin. Let L be compact and star-shaped with respect to the origin, the radial function ρL:Sn−1→[0,∞) is defined by
ρL(u)=max{λ≥0:λu∈L}, for all u∈Sn−1. |
A compact star-shaped set with respect to the origin is uniquely determined by its radial function. If ρL is positive and continuous on Sn−1, then the star-shaped L is called a star body about the origin. Let In0 be the set of all star bodies about the origin. Clearly, the radial function of a convex body in Kn0 is continuous and positive, i.e., Kn0⊂In0. If K∈Kn0, then
∂K={ρK(u)u:u∈Sn−1}. |
And the volume of K∈Kn0 can be rewritten by
|K|=1n∫Sn−1ρnK(u)du. | (2.1) |
For Borel set η⊂Sn−1, let ρ:η→(0,∞) be a continuous function, then the set {ρ(u)u:u∈η} is a Borel set. For a set K⊂Rn, the convex hull of K, write convK, is the intersection of all convex sets containing K. Hence, the convex hull ⟨ρ⟩ generated by ρ,
⟨ρ⟩=conv{ρ(u)u:u∈η} | (2.2) |
is a compact set (see [24]). Let η⊂Sn−1 always be a closed set and not contained in any great hemisphere of Sn−1, then we have ⟨ρ⟩∈Kn0. By the definition of convex hull of the function ρ, we have
ρ⟨ρ⟩(u)≥ρ(u), for all u∈Sn−1. | (2.3) |
If the function ρ is an even function on Sn−1, then the convex hull conv{ρ(u)u:u∈Sn−1} is an origin-symmetric body. Thus a convex body K is origin-symmetric if and only if the radial function ρK of the convex body K is an even function.
The definitions of radial function and support function immediately give that for λ>0 and K,L∈Kn0, one has hλK=λhK, ρλK=λρK and hK+L=hK+hL. Moreover
K⊂L ⇔ hK(u)≤hL(u) and K⊂L⇔ρK(u)≤ρL(u), for all u∈Sn−1. |
If K∈Kn0, the following formulas hold for all u∈Sn−1,
hK∗(u)=1ρK(u)andρK∗(u)=1hK(u), | (2.4) |
where K∗ is the polar body of K, and it is given by
K∗={x∈Rn:x⋅y≤1, for all y∈K}. |
On the set Kn, we consider the topology generated by the Hausdorff metric dH(⋅,⋅). For K,K′∈Kn, the Hausdorff metric dH(K,K′) is defined by
dH(K,K′)=‖hK−hK′‖∞=supu∈Sn−1|hK(u)−hK′(u)|. |
A sequence {Ki}i≥1⊂Kn converges to a convex body K0∈Kn if dH(Ki,K0)→0 as i→∞, i.e.,
dH(Ki,K0)→0 if and only if hKi→hK0 uniformly as i→∞. |
The radial metric is defined by
dρ(K,L)=‖ρK−ρL‖∞=supu∈Sn−1|ρK(u)−ρL(u)| |
for K,L∈In0. We use the fact that on Kn0, the Hausdorrf metric and the radial metric are topologically equivalent, i.e.,
dH(Ki,K0)→0 if and only if dρ(Ki,K0)→0 |
for {Ki}i≥1⊂Kn0 and K0∈Kn0.
If the function f:η→(0,∞) is continuous, the Wulff shape [f]∈Kn0 determined by f is a convex body defined by
[f]=⋂u∈η{x∈Rn:x⋅u≤f(u)}. |
Note that, if f=hK is a support function of convex body K∈Kn0, one has
[f]=K. |
Furthermore, if the function f:η→(0,∞) is continuous, we have (see e.g., [17] p.95)
[f]∗=⟨1/f⟩. | (2.5) |
The Lp Minkowski combination is a basic concept in the Lp-Brunn-Minkowski theory. For each p≥1, the Minkowski-Firey Lp-combination K+pL introduced by Firey (see, e.g., [24]) can be defined by the support function as follows, i.e., for K,L∈Kn0 and a,b>0
hpa⋅K+pb⋅L=ahpK+bhpL. |
Now we fix p≠0. For K,L∈Kn0 and a,b>0, we define the general Lp Minkowski combination, a⋅K+pb⋅L∈Kn0, via the Wulff shape,
a⋅K+pb⋅L=[(ahpK+bhpL)1/p]. |
When p=0, we define a⋅K+0b⋅L∈Kn0 via the Wulff shape,
a⋅K+0b⋅L=[haKhbL]. | (2.6) |
Note that "⋅" is written without its subscript p.
For any p∈R, Huang et al.[7] gave the definition of the Lp-harmonic combination (1−λ)⋄Kˆ+pλ⋄L∈Kn0, i.e.,
(1−λ)⋄Kˆ+pλ⋄L=((1−λ)⋅K∗+pλ⋅L∗)∗. | (2.7) |
Hence, together with (2.4)–(2.6), we obtain that
[haK∗hbL∗]∗=⟨ρaKρbL⟩. | (2.8) |
Let {Ki}∞i=1⊂Kn0 and K∈Kn0, combined with (2.4), this implies that
Ki→K if and only if K∗i→K∗. | (2.9) |
For a convex body K in Rn, the Gauss image of σ⊂∂K is defined by
ννK(σ)={v∈Sn−1:x∈HK(v) for some x∈σ}⊂Sn−1. |
The reverse Gauss image of η⊂Sn−1 is defined by
νν∗K(η)={x∈∂K:x∈HK(v) for some v∈η}⊂∂K. |
Let σK={x:x∈∂K is a singlar point}⊂∂K. It is known that Hn−1(σK)=0 (see, p.84 of Schneider [24]). The Gauss map of the convex body K is defined by
νK:∂K∖σK→Sn−1. |
From Lemma 2.2.12 of Schneider [24] we know that the Gauss map νK is continuous. The set ηK⊂Sn−1 consisting of all v∈Sn−1, for which the set νν∗K({v}), abbreviated as νν∗K(v), contains more than a single element, is of Hn−1-measure 0. The inverse Gauss map of the convex body K is defined by
ν−1K:Sn−1∖ηK→∂K, |
From Lemma 2.2.12 of Schneider [24] we also know that the function ν−1K is continuous.
For K∈Kn0, define the radial map of the convex body K
rK(⋅):Sn−1→∂K by rK(u)=ρK(u)u∈∂K, |
for u∈Sn−1. Note that r−1K(⋅):∂K→Sn−1 is the map r−1K(x)=ˉx=x/|x|. For ω⊂Sn−1, define the radial Gauss image of ω by
ααK(ω)=ννK(rK(ω))⊂Sn−1. |
Thus, for u∈Sn−1
ααK(u)={v∈Sn−1:rK(u)∈HK(v)}. |
Define the radial Gauss map of the convex body K∈Kn0
αK:Sn−1∖ωK→Sn−1 by αK=νK∘rK, |
where ωK={x/|x|:x∈σK}.
Define the reverse radial Gauss image of η⊂Sn−1 by
αα∗K:Sn−1→Sn−1 by αα−1K(η)=r−1K(νν∗K(η)). |
The inverse radial Gauss map of the convex body K∈Kn0 is defined by
α−1K:Sn−1∖ηK→Sn−1 by α−1K=r−1K∘ν−1K. |
Note that since both r−1K and ν−1K are continuous, α−1K is continuous.
The integral curvature J(K,⋅) of convex body K∈Kn0 is defined by,
J(K,ω)=Hn−1(ααK(ω)), | (2.10) |
for each Borel set ω⊂Sn−1. The total integral curvature of convex body K, is the surface area of the unit sphere Sn−1, thus J(K,Sn−1)=κn. The concept of integral curvature was introduced by Aleksandrov.
Following formula (2.10), and characteristic function I on Sn−1, then
∫Sn−1Iω(u)dJ(K,u)=∫Sn−1IααK(ω)(u)du=∫Sn−1Iω(α−1K(u))du, | (2.11) |
the last identity holds from the fact that v∈ααK(ω) if and only if α−1K∈ω for almost all u with respect to the spherical Lebesgue measure (see (2.20) in [6]). Furthermore, from formula (2.11), we have that
∫Sn−1f(u)dJ(K,u)=∫Sn−1f(α∗K(u))du | (2.12) |
for each continuous function f on Sn−1.
Lemma 2.1. ([6] Lemma 2.2) Let Ki∈Kn0 be such that limi→∞Ki=K0∈Kn0. Let ω=⋃∞i=0ωKi be the set (of Hn−1-measure zero) off of which all of the αKi are defined. If ui∈Sn−1∖ω are such that limi→∞ui=u0∈Sn−1∖ω, then limi→∞αKi(ui)=αK0(u0).
Lemma 2.2. ([6] Lemma 2.5) Let K∈Kn0, then
αα∗K(η)=ααK∗(η) |
for each η⊂Sn−1.
Note that if Ki→K0 in the Hausdorff metric, then for all f∈C(Sn−1), by formulas (2.9), (2.12), Lemmas 2.1 and 2.2, one has
limi→∞∫Sn−1f(u)dJ(Ki,u)=∫Sn−1f(u)dJ(K0,u). |
This proves that the integral curvature J(K,⋅) is weakly convergence measure.
Lemma 2.3. ([12] Lemma 2.1) If a sequence of measures {μi}∞i=1 on Sn−1 converges weakly to a finite measure μ on Sn−1 and a sequence of functions {fi}i≥1⊂C(Sn−1) converges uniformly to a function f0∈C(Sn−1), then
limi→∞∫Sn−1fi(u) dμi=∫Sn−1f0(u) dμ. |
Thus, by Lemma 2.3, if {fi}i≥1⊂C(Sn−1) is uniformly convergent to f0∈C(Sn−1) and {Ki}i≥1⊂Kn0 converges to K0∈Kn0 in the Hausdorff metric, together with the weak convergence of J(K,⋅), we have
limi→∞∫Sn−1fi(u) dJ(Ki,u)=∫Sn−1f0(u) dJ(K0,u). | (2.13) |
The Blaschke selection theorem is a powerful tool in convex geometry (see [5,24]) and will be often used in this paper. It reads: Every bounded sequence of convex bodies has a subsequence that converges to a convex set.
We will also use the following lemmas in the proofs of our main results.
Lemma 2.4. (see [12]) If {Ki}i≥1⊂Kn0 is a bounded sequence and {|K∗i|}i≥1 is also a bounded sequence, there is a subsequence {Kij}j≥1 of the sequence {Ki}i≥1 and a body K∈Kn0 such that Kij→K. In addition, if |K∗i|=ωn, then |K∗|=ωn.
Lemma 2.5. (see [16]) Let {Ki}i≥1⊂Kn0 be a convergent sequence with limit K0, i.e., Ki→K0 in the Hausdorff distance. If the sequence{|K∗i|}i≥1 is bounded, then K0∈Kn0.
In this section, we mainly prove some properties for the L0 mixed entropy. We now prove the continuity of the L0 mixed entropy as follows.
Proposition 3.1. Let {Ki}∞i=0⊂Kn0 and {Li}∞i=0⊂Kn0 be two sequences of convex bodies such that Ki→K0∈Kn0 and Li→L0∈Kn0 as i→∞ in the Hausdorff metric, then
E0(Ki,Li)→E0(K0,L0) as i→∞. |
Proof. Since Ki→K0∈Kn0 and Li→L0∈Kn0 as i→∞ in the Hausdorff metric, then J(Ki,⋅) converges weakly to J(K0,⋅) and the radial functions ρLi converges uniformly to ρL0, as i→∞. And there are two constants r, R>0 such that for all i≥1,
rBn2⊂ Li⊂RBn2. |
We have r≤ρLi≤R, for all i≥1. Furthermore, together with the continuity of the logarithmic function on [r,R], we get
logρLi(u)→logρL0(u) uniformly on Sn−1. |
By formula (2.13), one has
limi→∞E0(Ki,Li)=limi→∞−∫Sn−1logρLidJ(Ki,u)limi→∞E0(Ki,Li)=−∫Sn−1logρL0dJ(K0,u)limi→∞E0(Ki,Li)=E0(K0,L0). |
Proposition 3.2. Let {Ki}∞i=1⊂Kn0 and K∈Kn0 be regular convex bodies such that Ki→K as i→∞ in the Hausdorff metric. For {Li}∞i=1⊂Kne, then {E0(Ki,Li)}∞i=1 is bounded and {Li}∞i=1 is uniformly bounded if and only if there exist α,r>0 such that for all i≥1
rBn2⊂Li and |L∗i|≥α. |
Proof. The boundedness of {E0(Ki,Li)}∞i=1 is equivalent to the boundedness of {−E0(Ki,Li)}∞i=1, which shows that there are constants c and C such that c≤−E0(Ki,Li)≤C for all i≥1.
We first show that the sequence {Li}∞i=1 contains a small ball. For ui∈Sn−1, let
Ri(ui)=max{ρLi(u):u∈Sn−1}. |
Since the sequence {Li}∞i=1 is bounded, there is a constant β>0, such that Ri(ui)≤β for all i, we have β−1Bn2⊂L∗i for all i. Hence
|L∗i|≥ωnβn |
for all i.
On the other hand, by the Blaschke selection theorem, there is a subsequence of {Li}∞i=1, for convenience, we still record it as {Li}∞i=1, and a compact convex set L0, such that Li→L0 as i→∞ in the Hausdorff metric, the radial function sequence ρLi is uniformly continuous, we have that ρL0 is continuous, now we prove that L0 contains a small ball rBn2, if not, then there is a nonzero set ω and a sufficiently small real ϵ>0 such that ω={u:ρL0(u)<ϵ}, according to the regularity of the convex body K, we have J(K,ω)>0 and J(K,Sn−1∖ω)logR<∞. Thus, by Proposition 3.1,
c≤−limi→∞E0(Ki,Li)=−E0(K,L0)=∫Sn−1logρL0(u)dJ(K,u)limi→∞E(Ki,Li)E(K,L0)≤∫ωlogϵdJ(K,u)+∫Sn−1∖ωlogρL0(u)dJ(K,u)limi→∞E(Ki,Li)E(K,L0)≤J(K,ω)logϵ+J(K,Sn−1∖ω)logR. |
Let ϵ→0+, hence J(K,ω)logϵ→−∞, This is a contradiction to the boundedness of the mixed entropy E0(Ki,Li).
Now we prove that the sequence {Li}∞i=1 is bounded. We let Ri(ui)=max{ρLi(u):u∈Sn−1} for some ui∈Sn−1. Since the sequence {Li}∞i=1 contains a small ball, there is a constant r>0 such that rBn2⊂Li for all i≥1, let Qi be the convex hull of point Ri(ui)ui and rBn2|u⊥i, i.e.,
Qi=conv{Ri(ui)ui,rBn2|u⊥i}. |
Obviously, Qi⊂Li, together with the monotonicity of volume and (1.3),
|Li|≥|Qi|=1nRi(ui)Vn−1(rBn2|u⊥i). |
By the Blaschke-Stantalˊo inequality, i.e., for Li∈Kne,
|Li||L∗i|≤ω2n. |
Combined with |L∗i|≥α, this implies that
Ri(ui)=nVn−1(rBn2|u⊥i)|Qi|≤nVn−1(rBn2|u⊥i)|Li|≤nω2nαrn−1ωn−1, | (3.1) |
for all i. Thus the sequence {Li}∞i=1 is bounded. There are two constants r,R>0 such that rBn2⊂Li⊂RBn2 for all i, together with (2.10), we know that J(Ki,Sn−1)=κn, then for all i,
−κnlogR≤E0(Ki,Li)=−∫Sn−1logρLidJ(Ki,u)≤−κnlogr. |
This shows that the sequence {E0(Ki,Li)}∞i=1 is bounded.
Remark 1. According to the above proof of Proposition 3.2, if {Li}∞i=1⊂Kne and some α>0 such that rBn2⊂Li and |L∗i|≥α, then we can remove the condition that K∈Kn0 is a regular convex body, we also obtain the results that {E0(Ki,Li)}∞i=1 is bounded and {Li}∞i=1 is uniformly bounded.
Throughout this section, we suppose that K∈Kn0, we mainly prove the existence and uniqueness of the Entropy-Petty body. For further discussion, we introduce the continuity of the geominimal integral curvature GE(K). We first define the geominimal integral curvature GE(K) as follows:
Definition 4.1. Suppose K∈Kn0 is a convex body, the geominimal integral curvature of K is defined by
GE(K)=supL∈ Kne{∫Sn−1logρ(vrad(L∗)L,u)dJ(K,u)}˜GE(K)=sup{∫Sn−1logρ(L,u)dJ(K,u):L∈ Kne with |L∗|=ωn}. |
Remark 2. We show that the above definition is well defined. In fact, since |L∗|=ωn and L∈Kne, by the Blaschke-Stantalˊo inequality, we have |L|≤ωn. Hence, by (2.1) and the Jensen inequality
ωn≥1n∫Sn−1ρnL(u)du≥κnn(1κn∫Sn−1ρL(u)du)n. |
Hence ∫Sn−1ρL(u)du is uniformly bounded. In the next, we assume K∈Kn0 is a regular convex body, by the concavity of Logarithmic function, we obtain that
∫Sn−1logρ(L,u)dJ(K,u)≤κnlog(1κn∫Sn−1ρ(L,u)dJ(K,u))=κnlog(1κn∫Sn−1ρ(L,u)du)<∞. |
For any K∈Kn0, we choose a regular convex body sequence Ki such that Ki→K as i→∞, combined with Proposition 3.1, this implies that
∫Sn−1logρ(L,u)dJ(K,u)<∞. |
Hence Definition 4.1 is well defined.
In the next, we prove our mainly result Theorem 1.1.
proof of Theorem 1.1. Firstly, we prove the existence. By Definition 4.1, there is a sequence {Mi}∞i=1⊂Kne with |M∗i|=ωn such that
0=∫Sn−1logρBn2(u)dJ(K,u)≤∫Sn−1logρMi(u)dJ(K,u)<∞,for all i≥1. |
Let K∈Kn0 be a regular convex body, we have
0≤∫Sn−1logρMi(u)dJ(K,u)=∫Sn−1logρMi(u)du<∞. |
By formula (2.4), we get
−∞<∫Sn−1loghM∗i(u)du≤0. |
Let Ri(ui)=max{ρM∗i(u):u∈Sn−1}, and since Mi∈Kne, we have [−Ri(ui)ui,Ri(ui)ui]⊂M∗i. Hence h(M∗i,u)≥Ri(ui)|u⋅ui| for all u∈Sn−1. Therefore
κnlogRi(ui)+∫Sn−1log|u⋅ui|du≤∫Sn−1loghM∗i(u)du≤0. |
Now, assume K∈Kn0 is not a regular convex body, we can choose a regular convex body sequence Ki such that Ki→K as i→∞, combined with Proposition 3.1, this implies that
κnlogRi(ui)+∫Sn−1log|u⋅ui|du≤∫Sn−1loghM∗i(u)dJ(K,u)≤0. |
Since the integral on the left is independent of ui, this implies that Ri(ui) is uniformly bounded. Hence, there exists r>0 such that rBn2⊂Mi for all i≥1. By Proposition 3.2, the sequence {Mi}∞i=1 is bounded. By the Blaschke selection theorem, there is a subsequence, for convenience, it is still recorded it as {Mi}∞i=1, which converges to a compact convex set M. Since |M∗i|=ωn, by Lemma 2.5, we have M∈Kn0. Therefore Mi∈Kne, gives M∈Kne. By Proposition 3.1, we obtain that
GE(K)=limi→∞∫Sn−1logρMi(u)dJ(K,u)=∫Sn−1logρM(u)dJ(K,u) with |M∗|=ωn. |
Next, we prove the uniqueness of theorem in the plane R2. Assume that there are two convex bodies M1,M2∈K2e with |M∗1|=|M∗2|=π such that
GE(K)=∫Sn−1logρM1(u)dJ(K,u)=∫Sn−1logρM2(u)dJ(K,u). |
Now we define a new set M⊂Rn about M1,M2 and together with (2.7), we have
M=12⋄M1ˆ+012⋄M2=(12⋅M∗1+012⋅M∗2)∗. |
Combining (2.3) and (2.8), we obtain M=⟨ρ1/2M1ρ1/2M2⟩, this together with M1,M2∈K2e implies that the function ρ(u)=ρ1/2M1(u)ρ1/2M2(u) is even function on Sn−1. Hence M∈K2e and
ρ(M,u)≥ρ(M1,u)12ρ(M2,u)12, for all u∈Sn−1. | (4.1) |
Furthermore, by the log Brunn-Minkowski inequality in the plane (see [2]),
|M∗|=|12⋅M∗1+012⋅M∗2|≥√|M∗1|⋅|M∗2|=π, |
with equality if and only if M∗1 and M∗2 are dilates or they are parallelograms with parallel sides. If M∗1 and M∗2 are dilates, we let M∗1=sM∗2 for real number s>0 and together with |M∗1|=s2|M∗2|, we have s=1, thus we see M1=M2, which can be checked that vrad(M∗)≥1, with equality if and only if M1=M2 or their polar bodies are parallelograms with parallel sides. By (4.1) and Definition 4.1, we have
GE(K)≥∫S1logρ(vrad(M∗)M,u)dJ(K,u)GE(K)≥∫S1logρ(M,u)dJ(K,u)GE(K)≥∫S1log[ρ12(M1,u)ρ12(M2,u)]dJ(K,u)GE(K)=∫S112[logρM1(u)+logρM2(u)]dJ(K,u)GE(K)=GE(K). |
Hence, this forces vrad(M∗)=1 and then M1=M2 or their polar bodies are parallelograms with parallel sides.
We will prove the continuity of GE(K) as follows:
Theorem 4.1. Let {Ki}∞i=1⊂Kn0 and K∈Kn0 be such that Ki→K as i→∞ in the Hausdorff metric, then limi→∞GE(Ki)=GE(K).
Proof. Let {Ki}∞i=1⊂Kn0 be such that Ki→K∈Kn0 as i→∞. For any fixed small ε>0, by Definition 4.1 and Proposition 3.1, there is a convex body Mε with |M∗ε|=ωn, we have
GE(K)−ε≤−E0(K,Mε)=−limi→∞E0(Ki,Mε)=−lim infi→∞E0(Ki,Mε)≤lim infi→∞GE(Ki). |
Since ε>0 is arbitrary small, one has
GE(K)≤lim infi→∞GE(Ki). | (4.2) |
We now assume that Mi∈Kne with |M∗i|=ωn such that GE(Ki)=−E0(Ki,Mi),
0=∫Sn−1logρBn2(u)dJ(Ki,u)≤∫Sn−1logρMi(u)dJ(Ki,u)<∞,for all i≥1. |
Since both the sequence E0(Ki,Mi) and {Mi}∞i=1 are bounded, the Blaschke selection theorem now yields that a subsequence {Mij}∞j=1 of {Mi}∞i=1 converges to some compact convex set M′. But |M∗ij|=ωn, by LemmaS 2.4, 2.5, and Mi∈Kne, the set M′∈Kne is an origin-symmetric convex body and |(M′)∗|=ωn. Together with the Definition 4.1, Proposition 3.1 and Theorem 1.1, we obtain
GE(K)≥−E0(K,M′)=−limi→∞E0(Ki,Mi)=limi→∞GE(Ki)=lim supi→∞GE(Ki). | (4.3) |
Combining (4.2) and (4.3), we complete the proof, i.e.,
limi→∞GE(Ki)=GE(K). |
In the following corollary, we show that if the convex body K is an origin-symmetric polytope, then the optimal problem has an origin-symmetric polytope solution.
Corollary 4.1. Let K∈Kne be a polytope with vertices u1,u2,⋅⋅⋅,um. If M∈Kne such that
GE(K)=−E0(K,M) with |M∗|=ωn, |
then M is a polytope with vertices v1,v2,⋅⋅⋅,vm. Moreover, vi=λiui for λi>0, i∈{1,2,⋅⋅⋅,m}.
Proof. Let K∈Kne be a polytope with vertK={u1,u2,⋅⋅⋅,um} (m=2N≥n+1). Obviously, {u1|u1|,⋅⋅⋅,um|um|}⊂Sn−1 are not concentrated in any closed hemisphere of Sn−1. Then the integral curvature measure J(K,⋅) about convex body K is the discrete measure concentrated on {u1|u1|,⋅⋅⋅,um|um|}⊂Sn−1. Let P be a polytope
P=conv{ρ(M,ˉu1)ˉu1,ρ(M,ˉu2)ˉu2,⋅⋅⋅,ρ(M,ˉum)ˉum}. | (4.4) |
where ˉui=ui|ui|∈Sn−1 for i=1,...m. Let uP,i=ρ(M,ˉui)ˉui∈∂P be vertices of polytope P, then these are λi>0 such that uP,i=λiui for i∈{1,2,⋅⋅⋅,m}.
In the next, we only need prove P=M. By (4.4), we have ρ(P,ˉui)=ρ(M,ˉui) (1≤i≤m) and P⊂M. Thus
vrad(P∗)≥vrad(M∗)=1. |
We obtain
GE(K)=supL∈Kne{∫Sn−1log[vrad(L∗)ρ(L,u)]dJ(K,u)}GE(K)≥∫Sn−1log[vrad(P∗)ρ(P,u)]dJ(K,u)GE(K)≥∫Sn−1logρ(P,u)dJ(K,u)GE(K)=m∑i=1logρ(P,ˉui)⋅J(K,{ˉui})GE(K)=m∑i=1logρ(M,ˉui)⋅J(K,{ˉui})GE(K)=∫Sn−1logρ(M,u)dJ(K,u)GE(K)=supL∈Kne{∫Sn−1log[vrad(L∗)ρ(L,u)]dJ(K,u)}GE(K)=GE(K). |
This shows that vrad(P∗)=vrad(M∗)=1. Hence we know that |M|=|P|. Thus P=M.
In this paper, the geominimal integral curvature on the convex body is introduced. The existence and uniqueness of the geominimal integral curvature are proved. Some other properties for the geominimal integral curvature, such as continuity, are investigated.
The author declares that they have no competing interests.
This research is supported by Youth Thousand Talents Plan, Grant (No. GK202101008) and Postgraduate Innovation Team Project of Shaanxi Normal University (No. TD2020008Z).
[1] | A. D. Aleksandrov, Existence and uniqueness of a convex surface with a given integral curvature, Acad. Sci. USSR, 35 (1942), 131–134. |
[2] |
K. J. B¨or¨oczky, E. Lutwak, D. Yang, G. Zhang, The log-Brunn-Minkowski inequality, Adv. Math., 231 (2012), 1974–1997. https://doi.org/ 10.1016/j.aim.2012.07.015 doi: 10.1016/j.aim.2012.07.015
![]() |
[3] |
Y. Feng, B. He, The Orlicz Aleksandrov problem for Orlicz integral curvature, Int. Math. Res. Not., 2021 (2021), 5492–5519. https://doi.org/10.1093/imrn/rnz384 doi: 10.1093/imrn/rnz384
![]() |
[4] | R. J. Gardner, Geometric tomography, Cambridge Univ. Press, Cambridge, 1995. |
[5] | P. M. Gruber, Convex and discrete geometry, Springer-Verlag, Berlin Heidelberg, 2007. |
[6] |
Y. Huang, E. Lutwak, D. Yang, G. Zhang, Geometric measures in the dual Brunn-Minkowski theory and their associated Minkowski problems, Acta Math., 216 (2016), 325–388. https://doi.org/10.1007/s11511-016-0140-6 doi: 10.1007/s11511-016-0140-6
![]() |
[7] |
Y. Huang, E. Lutwak, D. Yang, G. Zhang, The Lp-Aleksandrov problem for Lp-integral curvature, J. Differ. Geom., 110 (2018), 1–29. https://doi.org/10.4310/jdg/1536285625 doi: 10.4310/jdg/1536285625
![]() |
[8] |
Q. Li, W. Sheng, X. Wang, Flow by Gauss curvature to the Aleksandrov and dual Minkowski problems, J. Eur. Math. Soc., 22 (2020), 893–923. https://doi.org/10.4171/JEMS/936 doi: 10.4171/JEMS/936
![]() |
[9] |
N. Li, S. Mou, The general dual orlicz geominimal surface area, J. Funct. Space., 2020 (2020), 1–6. https://doi.org/10.1155/2020/1387269 doi: 10.1155/2020/1387269
![]() |
[10] | M. Ludwig, General affine surface areas, Adv. Math., 224 (2010), 2346–2360. https://doi.org/10.1016/j.aim.2010.02.004 |
[11] |
M. Ludwig, M. Reitzner, A characterization of affine surface area, Adv. Math., 147 (1999), 138–172. https://doi.org/10.1006/aima.1999.1832 doi: 10.1006/aima.1999.1832
![]() |
[12] |
X. Luo, D. Ye, B. Zhu, On the polar Orlicz-Minkowski problems and the p-capacitary Orlicz-Petty bodies, Indiana U. Math. J., 69 (2020), 385–420. https://doi.org/10.1512/iumj.2020.69.7777 doi: 10.1512/iumj.2020.69.7777
![]() |
[13] | E. Lutwak, Dual mixed volume, Pac. J. Math., 58 (1975), 531–538. https://doi.org/10.2140/pjm.1975.58.531 |
[14] | E. Lutwak, Mixed affine surface area, J. Math. Anal. Appl., 125 (1987), 351–360. https://doi.org/10.1016/0022-247X(87)90097-7 |
[15] |
E. Lutwak, Centroid bodies and dual mixed volumes, P. Lond. Math. Soc., 2 (1990), 365–391. https://doi.org/10.1112/plms/s3-60.2.365 doi: 10.1112/plms/s3-60.2.365
![]() |
[16] |
E. Lutwak, The Brunn-Minkowski-Firey theory II. Affine and geominimal surface areas, Adv. Math., 118 (1996), 244–294. https://doi.org/10.1006/aima.1996.0022 doi: 10.1006/aima.1996.0022
![]() |
[17] |
E. Lutwak, D. Yang, G. Zhang, Lp dual curvature measures, Adv. Math., 329 (2018), 85–132. https://doi.org/10.1016/j.aim.2018.02.011 doi: 10.1016/j.aim.2018.02.011
![]() |
[18] |
S. Mou, B. Zhu, The orlicz-minkowski problem for measure in Rn and Orlicz geominimal measures, Int. J. Math., 30 (2019), 1950052. https://doi.org/10.1142/S0129167X19500526 doi: 10.1142/S0129167X19500526
![]() |
[19] |
V. Oliker, Hypersurfaces in Rn+1 with prescribed Gaussian curvature and related equations of Monge-Ampˊere type, Commun. Part. Diff. Eq., 9 (1984), 807–838. https://doi.org/10.1080/03605308408820348 doi: 10.1080/03605308408820348
![]() |
[20] |
V. Oliker, Embedding Sn−1 into Rn+1 with given integral Gauss curvature and optimal mass transport on Sn−1, Adv. Math., 213 (2007), 600–620. https://doi.org/10.1016/j.aim.2007.01.005 doi: 10.1016/j.aim.2007.01.005
![]() |
[21] | C. M. Petty, Geominimal surface area, Geometriae Dedicata, 3 (1974), 77–97. https://doi.org/10.1007/BF00181363 |
[22] |
P. Guan, Y. Li, C1,1 estimates for solutions of a problem of Alexandrov, Commun. Pure Appl. Math., 50 (1997), 189–811. https://doi.org/10.1002/(SICI)1097-0312(199708)50:8<789::AID-CPA4>3.0.CO;2-2 doi: 10.1002/(SICI)1097-0312(199708)50:8<789::AID-CPA4>3.0.CO;2-2
![]() |
[23] | L. A. Santalˊo, Un invariante afin para los cuerpos convexos del espacio de n-dimensiones, Port. Math., 8 (1949), 155–161. |
[24] | R. Schneider, Convex Bodies: The Brunn-Minkowski theory, second edition, Cambridge Univ. Press, 2014. |
[25] | W. Wang, Q. Chen, Lp Dual geominimal surface area, J. Ineq. Appl., 6 (2011), 264–275. |
[26] |
D. Ye, Lp geominimal surface areas and their inequalities, Int. Math. Res. Not., 2015 (2015), 2465–2498. https://doi.org/10.1093/imrn/rnu009 doi: 10.1093/imrn/rnu009
![]() |
[27] |
D. Ye, Dual Orlicz-Brunn-Minkowski theory: Dual Orlicz Lϕ affine and geominimal surface areas, J. Math. Anal. Appl., 443 (2016), 352–371. https://doi.org/10.1016/j.jmaa.2016.05.027 doi: 10.1016/j.jmaa.2016.05.027
![]() |
[28] |
D. Ye, B. Zhu, J. Zhou, The mixed Lp geominimal surface area for multiple convex bodies, Indiana U. Math. J., 64 (2015), 1513–1552. https://doi.org/10.1512/iumj.2015.64.5623 doi: 10.1512/iumj.2015.64.5623
![]() |
[29] |
S. Yuan, H. Jin, G. Leng, Orlicz geominimal surface areas, Math. Ineq. Appl., 18 (2015), 353–362. https://doi.org/10.7153/mia-18-25 doi: 10.7153/mia-18-25
![]() |
[30] |
B. Zhu, J. Zhou, W. Xu, Lp mixed geominimal surface area, J. Math. Anal. Appl., 422 (2015), 1247–1263. https://doi.org/10.1016/j.jmaa.2014.09.035 doi: 10.1016/j.jmaa.2014.09.035
![]() |
[31] | B. Zhu, N. Li, J. Zhou, Isoperimetric inequalities for Lp geominimal surface area, Glasg. Math. J. 53 (2011), 717–726. https://doi.org/10.1017/S0017089511000292 |
1. | Shuang Mou, Lp geominimal Gaussian surface area, 2024, 38, 0354-5180, 4991, 10.2298/FIL2414991M |