Limiting shape of the -Minkowski problem
Abstract.
In [3], Andrews classified the limiting shape for isotropic curvature flow corresponding to the solutions of the -Minkowski problem as in the planar case. In this paper, we use the group-invariant method to study the asymptotic shape of solutions to the -Minkowski problem as in high dimensions. For any regular polytope , we establish the existence of a solution to the -Minkowski problem that converges to as , thereby revealing the intricate geometric structure underlying this limiting behavior. We also extend the result to the dual Minkowski problem.
Key words and phrases:
Minkowski problem, limiting shape, polytope.2010 Mathematics Subject Classification:
35J20; 35J60; 52A40; 53A151. Introduction
The classical Minkowski problem looks for a convex body such that its surface measure matches a given Radon measure on the unit sphere in . This problem dates back to the early works of Minkowski, Aleksandrov and Fenchel-Jessen (see [33]). For a convex body (the set of convex bodies containing the origin) and two parameters , Lutwak,Yang and Zhang [30] introduced the measure
for any measurable set , and proposed the following dual Minkowski problem. Given a finite Radon measure on , find a convex body such that . Here, is the Gauss map, is the support function of , and is the -dimensional Hausdorff measure. In the smooth category, the measure is given by for a positive function , where is the spherical Lebesgue measure, the dual Minkowski problem can then be formulated by the equation
(1.1) |
This is a fully nonlinear equation of Monge-Ampère type, which has received increasing attention in recent years (see e.g. [6, 11, 21, 22, 30, 35, 36]). When , equation (1.1) reduces to the -Minkowski problem, introduced in [28], given by
(1.2) |
Here, denotes the standard Levi-Civita connection on the sphere , and is the identity matrix.
The -Minkowski problem has been extensively studied in recent years (see e.g. [1, 2, 4, 7, 8, 16, 19, 20, 23, 24, 27, 29, 31, 32, 37, 38]). According to the Blaschke-Santaló inequality, the problem is divided into three cases, the sub-critical case , the critical case , and the super-critical case . There are rich phenomena on the existence and multiplicity of solutions. In the sub-critical case , the solution is usually not unique [25]. When , there may be infinitely many solutions [26]. In the critical case , due to a Kazdan-Warner type obstruction [13], there may be no solutions to the problem (1.2). In the super-critical case , surprisingly, Guang, Li and the second author [18] proved that for any and any positive, smooth function , there is a solution to (1.2). An amazing result was obtained by Andrews [3]. He proved that in the planar case , for any regular polygon (with -sides), there is a solution to (1.2) which converges to , as tends to negative infinity. This raises a natural question of whether such limiting behavior holds in high dimensions:
Problem 1.1.
(i) Does there exist a solution to the -Minkowski problem (1.2), such that the associated convex body converges to a polytope , which is tangential to , as ?
(ii) Moreover, for any regular polytope tangent to the unit sphere ,
does there exist a solution to (1.2) for negative large ,
such that the associated convex body converges to the polytope
as ?
In fact, Andrews classified the limiting shape of the solutions to (1.2) in the planar case by solving an isotropic curvature flow. In this paper, we explore a new method, called the group-invariant method, to study the limiting shape of the solutions to (1.2) in high dimensions. To do so, we assume that, throughout the paper, satisfies the following conditions:
-
(i)
is a measurable function satisfying
(1.3) for positive constants .
-
(ii)
is invariant with respect to a group , which is a discrete subgroup of satisfying the spanning property (introduced in Section 2).
A particular case is when is the discrete subgroup of generated by a regular polytope.
First, we obtain the asymptotic behavior of solutions to the -Minkowski problem (1.2) for negative large in high dimensions, which solves part (i) of Problem 1.1.
Theorem 1.1.
Let be a discrete subgroup of satisfies the spanning property. Then for , there exists an -invariant solution to the -Minkowski problem (1.2), such that the associated convex body sub-converges to a group invariant polytope , which is tangential to , as .
We say that a polytope is tangential to the unit sphere if every face of is tangential to . It is easy to see that if is tangential to , then is convex and contains the unit ball .
By the regularity theory of Caffarelli [9, 10], the solution is if is positive and bounded, and is if is also Hölder continuous. By the maximum principle, there is a unique solution to (1.2) when [13]. It is easy to see that the solution tends to constant one when tends to infinity. In fact, at the maximum point of , we have
by equation (1.2). Hence . Similarly, one can show that by evaluating (1.2) at the minimum point of . We conclude that the limiting shape of the solutions is the unit ball as .
Theorem 1.1 deals with the case . When , there is a trivial solution to (1.2). Theorem 1.1 tells that there is a non-trivial solution when is negative large. The solution is a maximizer of the functional
(1.4) |
among all -invariant convex bodies containing the origin, where is the support function of ,
Here the value is equal to times the volume . The spanning property of the group guarantees the uniform estimate of group invariant solutions, and one can obtain the existence of solutions for all . By property of the functional , we show that the maximizer is not a constant when is negative large, and converges to a limit as tends to negative infinity. Moreover, the limit is a polytope.
The next theorem shows that every regular polytope tangent to the unit sphere is a limit polytope, thereby resolving part (ii) of Problem 1.1.
Theorem 1.2.
For any regular polytope tangential to the unit sphere , there exists solution to (1.2) for negative large , such that the associated convex body converges to the polytope as .
A regular polytope has strong symmetry. We prove Theorem 1.2 by considering maximizers of the functional among convex bodies with the strong symmetry. Theorems 1.1 and 1.2 can be regarded as a higher dimensional extension of Andrews’s result in ([3], Theorem 1.5) in the planar case. Here we allow to be a measurable function.
Next we consider the dual Minkowski problem (1.1). Under the group invariant assumptions in Theorem 1.1, we have uniform estimate for the group invariant solution to (1.1), which is a maximizer of the functional
(1.5) |
among all -invariant convex bodies , where
and .
Here we are interested in the asymptotic behaviour of the solution as tends to negative infinity for fixed , and the asymptotic behaviour as tends to positive infinity for fixed .
Theorem 1.3.
Let be as in Theorem 1.1. Let be the maximizer of the functional and be the associated convex body.
-
(1)
For any given , sub-converges to a group invariant polytope , which contains the unit ball and is tangential to , as .
-
(2)
For any given , converges to the unit ball as .
-
(3)
For any given , sub-converges to a group invariant polytope , which contains the unit ball and is tangential to , as .
The next theorem shows that every regular polytope tangential to the unit sphere is a limit polytope.
Theorem 1.4.
Let be a regular polytope and let satisfy .
-
(1)
If is tangential to the unit sphere , there exists a local maximizer of the functional , such that when , converges to as .
-
(2)
When , there exists a local maximizer of , which converges to a regular polytope similar to as .
This paper is organized as follows. In Section 2, we introduce some notations and properties concerning subgroups and the corresponding group invariant polytopes. In Section 3, we prove the existence of non-trivial group invariant solutions to (1.2) by a variational method for negative large . By passing to the limit , we characterize the limiting convex body in Section 4. In Section 5, we show that a regular polytope is a strictly local maximizer of the functional and thus give a proof of Theorem 1.2. Finally, we consider the dual Minkowski problem (1.1) and prove Theorems 1.3-1.4 in Sections 6-7.
2. Discrete subgroup and group invariant polytope
2.1. Notations
We will work in Euclidean space, where for any . A compact convex set with non-empty interior in is called a convex body. Denote by the set of convex bodies in containing the origin. Given a convex body , the support function of is defined by
We denote by the Hessian matrix of with respect to an orthonormal frame on the unit sphere , where is the Kronecker delta symbol. When is uniformly convex, the matrix is positive definite, with inverse . Let the radial distance function of be given by
There holds , where is the unit outer normal of at the point , such that is the Gauss map. Thus, can be recovered from by
The surface measure of is given by
where is the surface measure of the sphere .
2.2. Discrete subgroup and group invariant polytope
In this subsection, we introduce some notations and properties about the group invariant polytopes that will be used later. According to [34], a group is a set that is closed under multiplication, and satisfies the associative law, has a unit element, and has inverses. By definition, each discrete subgroup of is a finite group. For example,
is a discrete subgroup when is rational, and an infinite subgroup when is irrational.
Let be a discrete subgroup of . We say that satisfies the spanning property if for any , the set
forms a non-degenerate, -dimensional polytope in . Here we denote by the convex hull of the set . Moreover, we also define the bounded ratio condition by
(2.1) |
for the support function of . We have the following equivalence.
Proposition 2.1.
For a discrete subgroup of , the spanning property is equivalent to the bounded ratio condition (2.1).
Proof.
If the spanning property does not hold for some , then is a degenerate polytope in -dimension. In this case, the bounded ratio condition fails.
Conversely, if the bounded ratio condition does not hold for a sequence with tending to some , it is clear that is degenerate. ∎
The spanning property is important in the sense that it guarantees the a-priori estimates for the maximizing sequence of our variational problem in Section 3. Let us point out that in the spanning property, the set is generated by only one arbitrary point .
For a given discrete subgroup satisfying the spanning property, there may be more than one polytope corresponding to the subgroup . For example, let be the discrete subgroup associated with the regular -polygon for , then all regular -polygons with are also -invariant, for all integers .
Regular polytopes are of particular interest. According to [14, 15], we introduce the following definition for regular polytopes.
Definition 2.1.
A flag is a sequence of faces of a polytope , each contained in the next, with exactly one face from each dimension. More precisely, a flag of an -dimensional polytope is a set , where for is the -dimensional face of , such that for all .
Definition 2.2.
A regular polytope is a polytope whose symmetry group , as a maximal subgroup of , acts transitively on its flags in the sense for any flags , there exists such that .
Regular polytopes are the high dimensional analog of regular polygons () and regular polyhedra (). A concise symbolic representation for regular polytopes was developed by Schläfli in the 19th century, and a slightly modified form has become standard nowadays [14]. Regular polytopes can be classified according to the symmetry group . Regular polytopes with vertices in dimension are classified as follows,
(2.2) |
Example 2.1.
Let be a standard cube. Let be the associated symmetric group. The following statements hold.
-
(1)
When , is a standard cube in , which is a regular polytope.
-
(2)
When , is a cross-polytope in , which is also a regular polytope.
-
(3)
When , is a group invariant polytope formed by cutting eight solid angles from eight vertices of the cube, which is not a regular polytope.
3. Existence of group invariant solutions of (1.2)
Group invariant solutions have been studied in various geometric problems. Here we consider the discrete subgroup that satisfies the spanning property. By the bounded ratio condition in Section 2, we can prove the following existence result.
Theorem 3.1.
Let be a discrete subgroup of satisfying the spanning property. For , equation (1.2) admits a group invariant solution , which is a maximizer of the global variational problem
(3.1) |
over the group invariant family
where is the set of group invariant convex bodies containing the origin, and is the support function of .
To prove Theorem 3.1, it suffices to prove
Lemma 3.1.
To prove Lemma 3.1, we first prove the following lemmas.
Lemma 3.2.
For , , there exists a positive constant depending only on and , such that
(3.2) |
for each .
Proof.
Lemma 3.3.
For , , and a group invariant positive function on , there exists a positive constant depending only on such that
(3.5) |
for any support function with the associated convex body in .
Proof.
Denote and . Assume that for the south polar . By the convexity of , we have
where is the geodesic distance from to . Hence,
for a positive constant depending only on . By taking some large constant , noting that
holds for another positive constant depending only on , inequality (3.5) then follows from Lemma 3.2 when . ∎
As a corollary of Lemma 3.3, we have the following result.
Corollary 3.1.
Let be the solution obtained in Lemma 3.1 with the support function , we have
(3.6) |
Proof.
When a-priori lower-upper bounds of functions have been derived in Lemmas 3.2 and 3.3, by Blaschke’s selection theorem, a maximizing sequence converges to a limiting convex body . Due to the uniform convergence of the support functions to , we see that is actually a maximizer of the variational problem (3.1). Consequently, we can prove Lemma 3.1.
Proof of Lemma 3.1.
To show Lemma 3.1, it remains to verify that satisfies the Euler-Lagrange equation (1.2). This can be achieved by adapting the argument in ([12], Section 3) (see also [13], Section 5) with slight modifications. First, by calculating the first variation along the Wulff shape, it follows that the Monge-Ampère measure associated to the solution is bounded from above. By the duality of the polar bodies, we have also that the Monge-Ampère measure is bounded from below. Then, using the work of Caffarelli in [9], we can demonstrate the strict convexity and regularity of the solution. Consequently, we can show that the maximizer satisfies that
(3.7) |
for all group invariant function , we thus reach the conclusion that satisfies the Euler-Lagrange equation (1.2). Actually, one needs only to consider group invariant Hilbert space
and use the fact that the orthonormal subspace of itself must be identical to . Then, (3.7) means that the group invariant function belongs to the orthonormal subspace of , and hence must be identical to zero.
4. Limiting extremal bodies
By Theorem 3.1 and its proof, we have
Theorem 4.1.
Proof.
For any convex body with support function , we normalize it to with support function
Then we have
Moreover, the equality holds at multiples of . This completes the proof. ∎
It is worth noting that when the Santaló center of is the origin and when , our functional is exactly Mahler’s volume [33]. By Lemmas 3.2 and 3.3, the sequence of maximizers of the functional is a-priori bounded in by convexity. Hence, along a subsequence, the convex body converges to a limiting convex body . We have the following characterization of .
Lemma 4.1.
Let be a discrete subgroup of satisfying the spanning property. The best constant of the functional converges to the best constant of the limiting functional
Moreover, is exactly a maximizer of the functional .
Proof.
For each , we have
(4.2) |
Conversely, for each , by the convergence property of -norm to -norm, we have
(4.3) |
Combining (4.3) with (4.2) yields
(4.4) |
On the other hand, by the uniform convergence of to , and the condition (1.3), we have
where the infinitesimal is small as long as is negative large. As a result,
by (4.4). This implies that is exactly a maxmizer of the functional . ∎
The next lemma rules out the trivial possibility of the limiting sphere .
Lemma 4.2.
Let be any group invariant polytope, then
(4.5) |
Consequently, the limiting shape of , when , cannot be a sphere.
Proof.
The next lemma implies that the maximizer of the limiting functional must be a polytope.
Lemma 4.3.
For any non-polytope group invariant convex body , there exists a group invariant polytope containing such that
(4.6) |
Proof.
Let and for the inner-radius of , we denote to be the tangent hyperplane of at , and let be the half space of containing . Define
Obviously, is a group invariant polytope containing , then
as long as is not a polytope. ∎
Combining with Theorems 3.1 and 4.1, along with Lemma 4.3, we obtain the following theorem, which characterizes the limiting extremal body as a maximal group-invariant polytope.
Theorem 4.2.
Let be a discrete subgroup of satisfying the spanning property. The maximizer of the functional must be a group invariant polytope, after scaling such that the volume .
5. Local maximizer of the functional and proof of Theorem 1.2
In this section, we use to denote a regular polytope that is tangential to the unit sphere , and to represent the symmetry group of . Denote
the -neighborhood of for a positive constant and the Hausdorff distance , and denote the boundary of . Denote also the normalized group invariant family
and the normalized -neighborhood
of . Then, we can prove that has strictly maximal volume in the normalized family for some as follows.
Lemma 5.1.
For any regular polytope tangential to the unit sphere , there exists a small constant such that
(5.1) |
Proof.
Step 1. Subdivision of a regular polytope.
Let be a regular polytope that is tangential to the unit sphere . By the high symmetry of , each -dimensional face is also a regular polytope, where and . We will use this high symmetry to decompose into as many congruent pieces as possible. Let , the origin of , be the centroid of . For an -dimensional face of , we denote by the centroid of . Let be one of the -dimensional face of , we denote its centroid by . The bootstrap procedure will produce a sequence of faces
with lower dimensions, and yield a sequence of corresponding centroids
Note that
(5.2) |
Since , one may assume that is parallell to the -axis by rotation and set . Define , we have the recursive formula
(5.3) |
Then, can be decomposed into small pieces that are congruent to 111Sometimes, we may regard as a subdomain of for by simply omitting the coordinates for . For example, for each .. Let be the point of the piece defined above with projection
satisfying for support function of another normalized group invariant convex body .
Figure: subdivision of a polyhedron tangential to the unit sphere in .
Step 2. Claim : The regular polytope has local strictly maximal volume on the normalized group invariant family if the following integral function
(5.4) |
is positive for all closing to . (See footnote 1 for .)
Proof of claim .
At first, the hyperplane tangential to the unit sphere at is given by the radial function for . While, the hyperplane tangential to the at is given by the radial function . Therefore, we obtain that
(5.5) |
where represents the volume of the solid enclosed by the hyperplane and the infinite cone formed by connecting the origin to . Note that for each normalized group invariant convex body , we also have
(5.6) |
Regarding the coordinates in the hyperplane as the local coordinates of by projection , we have the metric
since , and thus the volume element is
Therefore, the volume difference function
holds for . (See footnote 1 in page 14 for .) Letting be the -th coordinates of and setting , there holds
(5.7) |
due to . This, together with (5.6), completes the proof of claim . ∎
Step 3. Claim : The regular polytope has locally strictly maximal volume on the normalized group invariant family for some . More precisely, there exists a positive constant such that for any , there holds
(5.8) |
Proof of claim .
The proof of the claim is equivalent to verifying that: If approaches without equaling , then the volume difference function in (5.4) is positive. Along the direction segment 222, the volume difference function can be simplified by for ,
Taking the first differentiation in and then evaluating at , we get
(5.9) |
using the fact that (See footnote 1 in page 14 for ). This, together with Claim , yields , and hence the desired inequality (5.8) holds. ∎
Then Lemma 5.1 follows directly from claim . ∎
For the given regular polytope tangential to , define the group invariant family
As shown in Lemma 5.1, is a local strict maximizer of the functional on the normalized family for some . Hence, there exists a positive constant such that
(5.10) |
We now consider the local variational scheme
(5.11) |
where and . Note that , i.e., as . In fact, the convergence of to the limiting family
(5.12) |
under the Hausdorff distance. To clarify this convergence, we mean that for each , there exists a sequence of converges to under Hausdorff distance. Vice versa, for any sequence of , we have sub-converges to some .
We now show the existence of maximizer of (5.11) for negative large , whose support function satisfies the Euler-Lagrange equation up to a constant multiplication. The main ingredient of the proof is to show that the maximizer of (5.11) stays away from the boundary of , that is, and .
Lemma 5.2.
Let be a regular polytope tangential to the unit sphere . For each , the variational problem (5.11) admits a maximizer staying away from the boundary of for a small .
Proof.
For simplicity, let denote the union of for , that is, . By a-priori Lipschitz bound of convex bodies , the boundary is a sequential compact set under the Hausdorff distance. Noting that the limiting functional is continuous under the Hausdorff distance, it follows from (5.10) that for some ,
(5.13) |
Noting that for fixed , we have the convergence
(5.14) |
We claim that the functionals are equi-continuous for negative large . Actually, by Minkowski’s inequality for the -weighted -space defined by the norm
and a positive constant , we have for some constant ,
Thus, the equi-continuity of the functional follows from the definition (1.4).
Combining with the sequential compactness of and the equi-continuity of the functionals for negative large , we have
(5.15) |
in particular,
For short, we write
By the compactness of and formula (5.12), we have the convergence
(5.16) |
Noting that by Lemma 5.1, it follows from (5.13)-(5.16) that
hold for negative large and some infinitesimal , where in the second inequality we have also used (5.13) and the convergence of to under Hausdorff distance. Using again the relation
(5.17) |
for some infinitesimal and negative large , we conclude from (5) that
(5.18) |
holds for small and negative large . We thus arrive at the existence result of local maximizer of the functional for the negative large . ∎
Lemma 5.3.
Proof.
The existence of maximizer follows from Lemma 5.2. Owing to the property that stays away from the boundary of , we can prove that satisfies the Euler-Lagrange equation (1.2) as in the proof of global variational problem. Noting that converges to as , the maximizer converges to a maximizer of the limiting functional. However, by Lemma 5.1, is the unique maximizer of the limiting functional on the family . This, together with the observation that for all , implies that . This completes the proof of the lemma. ∎
From our proof, a regular polytope is a local maximizer of the functional . We would like to point out that it may not be a global maximizer of .
6. The dual Minkowski problem
In this section, we turn to study the dual Minkowski problem (1.1) and prove Theorem 1.3. Parallel to Section 3, we consider a variational problem
(6.1) |
on the group invariant family
A similar argument of Lemma 3.1 gives the following solvability result.
Proposition 6.1.
Recall the functional
defined on -invariant convex bodies . Similarly to Section 4, the maximizer in Proposition 6.1 also serves as the extremal body of the following inequality.
Proposition 6.2.
Remark 6.1.
Let be the polar body of . Then there exists a duality
(6.3) |
Thus, the critical point of the functional corresponds to the critical point of the functional . Consequently, an analogous solvability result to Proposition 6.1 also holds for and .
By the group invariance of the solution , one has the uniform a-priori bound for . Passing to the limit , we obtain the limit which is the extremal body of the functional .
Lemma 6.1.
Let be a discrete subgroup of satisfying the spanning property. The best constant of the functional converges to the best constant of the limiting functional
Moreover, is a maximizer of the functional .
Lemma 6.2.
Let be a discrete subgroup of satisfying the spanning property.
-
(1)
The maximizer is a group invariant polytope if .
-
(2)
The maximizer is a ball for some , if .
Proof.
Part (1) can be proven as that in Theorem 4.2. For part (2), let be the polar body of . Then,
Noting that the functional is invariant under scaling, we may assume that . In this case, we have
with equality if and only if . This implies that the extremal body maximizing the functional must be a ball. Hence, we conclude that is also a ball by duality. The proof is completed. ∎
Lemma 6.3.
Let be a discrete subgroup of satisfying the spanning property. For each , the limiting shape of as is a group invariant polytope.
Proof.
When and , the maximizer of is also the maximizer of the functional
As , we conclude that is the maximizer of the limiting functional
Noting that the function is invariant under scaling, we may assume that for some . Let be a group invariant polytope. It is clear that is contained inside of by convexity. Thus,
as long as is not a polytope. Therefore, the limiting shape can only be a polytope for .
When and , the maximizer of is also the minimizer of the functional . Passing to the limit as , we conclude that is the minimizer of the limiting functional . Without loss of generality, we may also assume that for some . Then, it is clear that the minimizer of the functional is attained at for . ∎
7. The proof of Theorem 1.4
This section aims to prove Theorem 1.4. Firstly, let us consider . In this case, we have the following result, analogous to Lemma 5.1.
Lemma 7.1.
For , a regular polytope is a strict local maximal of the functional in the family for some . Namely, there exists a positive constant such that for any , there holds
(7.1) |
Proof.
Similar to the proof of Claim in Lemma 5.1, we can check that for some ,
if and only if
(7.2) |
for all closing to .
To complete the proof of Theorem 1.4 for the case , we shall consider the local variational scheme
(7.3) |
Then we will show the existence of maximizer of (7.3) for negative large , whose support function satisfies the Euler-Lagrange equation (1.1) up to a constant.
Lemma 7.2.
Proof.
The existence of maximizer to the variational scheme (7.3) follows from a-priori estimates for group invariant polytopes.
Next, we show that the property that . In fact, by Lemma 7.1, we know that is a local strictly maximizer of the functional on the family for some . Hence, there exists a positive constant such that
(7.4) |
Thus we have that for some ,
(7.5) |
and
(7.6) |
Similar to the proof of Lemma 5.2, we have
holds for negative large and some infinitesimal , and hence
(7.7) |
holds for small and negative large . This implies that the property that stays away from the boundary of . Then we can prove that satisfies the equation (1.1) as in the proof of the variational problem (6.1). Since
by the compactness of . Thus, the maximizer converges to a maximizer of the limiting functional. However, by Lemma 7.1, is the unique maximizer of the limiting functional on the family . We reach the conclusion that by the observation that for . This completes the proof. ∎
Lemma 7.3.
For any regular polytope with outer radius one and positive group invariant function on , there exists a small constant such that
(7.8) |
Proof.
Firstly, let and assume is positive group invariant function. For any , there exists at least one such that
Set , it is clearly that
(7.9) |
since . We remain to show that
(7.10) |
if approaches without being equal to . Noting that , if one takes their dual bodies (for some ), respectively, we have are closing from each other under the Hausdorff distance. So, to prove (7.10) is equivalent to proving
(7.11) |
where
The proof of (7.11) is similar to the proof of (5.8). Actually, after subdivision of into congruent pieces similar to one polytope , the comparison of with can be reduced to the verification of the negativity of the function
for approaching . Along the direction segment , we have
Therefore, it follows from
(7.12) |
for that (7.11) holds true.
After modifying the proof of (7.11) slightly, we can also show that: If , for approaches without being equal to , there holds
which is equivalent to
This, together with the fact that
(7.13) |
gives the desired result for case . ∎
With the help of Lemma 7.3, we know that the regular polytope is of local strictly extremal of for and . Considering the same local variational scheme (7.3) and modifying the argument in Section 5 slightly, we reach the following result.
Lemma 7.4.
Acknowledgement. The authors would like to thank Xudong Wang for many helpful comments.
References
- [1] B. Andrews, Gauss curvature flow: the fate of the rolling stones, Invent. Math., 138 (1999), 151-161.
- [2] B. Andrews, Motion of hypersurfaces by gauss curvature, Pacific J. Math., 195 (2000), 1-34.
- [3] B. Andrews, Classification of limiting shapes for isotropic curve flows, J. Amer. Math. Soc., 16 (2003), 443-459.
- [4] G. Bianchi, K. Böröczky, A. Colesanti and D. Yang, The -Minkowski problem for , Adv. Math., 341 (2019), 493-535.
- [5] K. Böröczky, The logarithmic Minkowski conjecture and the -Minkowski problem, Adv. Anal. Geom., 9 (2023), 83-118.
- [6] K. Böröczky and F. Fodor, The dual Minkowski problem for and , J. Diff. Equa., 266 (2019), 7980-8033.
- [7] K. Böröczky, E. Lutwak, D. Yang and G. Zhang, The logarithmic Minkowski problem, J. Amer. Math. Soc., 26 (2013), 831-852.
- [8] P. Bryan, M. Ivaki and J. Scheuer, A unified flow approach to smooth, even -Minkowski problems, Analysis & PDE, 12 (2019), 259-280.
- [9] L.A. Caffarelli, A localization property of viscosity solutions to the Monge-Ampère equation and their strict convexity, Ann. of Math., 131 (1990), 129-134.
- [10] L.A. Caffarelli, Interior estimates for solutions of the Monge-Ampère equation, Ann. of Math., 131 (1990), 135-150.
- [11] C. Chen, Y. Huang and Y. Zhao, Smooth solutions to the dual Minkowski problem, Math. Ann., 373 (2019), 953-976.
- [12] H. Chen, S. Chen and Q. Li, Variations of a class of Monge-Ampère-type functionals and their applications, Analysis & PDE, 14 (2021), 689-716.
- [13] K.S. Chou and X.-J. Wang, The -Minkowski problem and the Minkowski problem in centroaffine geometry, Adv. Math., 205 (2006), 33-83.
- [14] H.S.M. Coxeter, Regular polytopes, Courier Corporation, 1973.
- [15] P.R. Cromwell, Polyhedra, Cambridge University Press, 1997.
- [16] S. Du, On the planar -Minkowski problem, J. Diff. Equa., 287 (2021), 37-77.
- [17] P.M. Gruber, Convex and discrete geometry, Springer-Verlag, Berlin Heidelberg, 2007.
- [18] Q. Guang, Q. Li and X.-J. Wang, The -Minkowski problem with super-critical exponents, preprint.
- [19] Y. He, Q. Li and X.-J. Wang, Multiple solutions of the -Minkowski problem, Calc. Var. PDEs, 55 (2016), 13 pp.
- [20] Y. Huang, J. Liu and L. Xu, On the uniqueness of -Minkowski problems: the constant -curvature case in , Adv. Math., 281 (2015), 906-927.
- [21] Y. Huang, E. Lutwak, D. Yang and G. Zhang, Geometric measures in the dual Brunn-Minkowski theory and their associated Minkowski problems, Acta Math., 216 (2016), 325-388.
- [22] Y. Huang and Y. Zhao, On the dual Minkowski problem, Adv. Math., 332 (2018), 57-84.
- [23] D. Hug, E. Lutwak, D. Yang and G. Zhang, On the -Minkowski problem for polytopes, Discrete Comput. Geom., 33 (2005), 699-715.
- [24] M. Ivaki and E. Milman, -Minkowski problem under curvature pinching, Int. Math. Res. Not., IMRN (2024), 8638-8652.
- [25] H. Jian, J. Lu and X.-J. Wang, Nonuniqueness of solutions to the -Minkowski problem, Adv. Math., 281 (2015), 845-856.
- [26] Q. Li, Infinitely many solutions for centro-affine Minkowski problem, Int. Math. Res. Not., 18 (2019), 5577-5596.
- [27] J. Lu and X.-J. Wang, Rotationally symmetric solutions to the -Minkowski problem, J. Diff. Equa., 254 (2013), 983-1005.
- [28] E. Lutwak, The Brunn-Minkowski-Firey theory I. Mixed volumes and the Minkowski problem, J. Diff. Geom., 38 (1993), 131-150.
- [29] E. Lutwak, D. Yang and G. Zhang, On the -Minkowski problem, Trans. Am. Math. Soc., 356 (2004), 4359-4370.
- [30] E. Lutwak, D. Yang and G. Zhang, dual curvature measures, Adv. Math., 329 (2018), 85-132.
- [31] E. Milman, Centro-affine differential geometry and the log-Minkowski problem, J. Eur. Math. Soc., to appear.
- [32] E. Milman, A sharp centro-affine isospectral inequality of Szegö-Weinberger type and the -Minkowski problem, J. Diff. Geom., 127 (2024), 373-408.
- [33] R. Schneider, Convex bodies: the Brunn-Minkowski theory, Encyclopedia of Mathematics and its Applications, 151, Cambridge University Press, Cambridge, 2014.
- [34] L. Serge, Algebra, Grad. Texts in Math., 211, Springer-Verlag, New York, 2002, xvi+914 pp.
- [35] Y. Zhao, The dual Minkowski problem for negative indices, Calc. Var. PDEs, 56 (2017), 16 pp.
- [36] Y. Zhao, Existence of solutions to the even dual Minkowski problem, J. Diff. Geom., 110 (2018), 543-572.
- [37] G. Zhu, The -Minkowski problem for polytopes for , J. Funct. Anal., 269 (2015), 1070-1094.
-
[38]
G. Zhu, The -Minkowski problem for polytopes for negative ,
Indiana Univ. Math. J., 66 (2017), 1333-1350.