Local fixed point results for centroid body operators
Abstract.
We prove that, in a neighborhood of the Euclidean ball, there are no other fixed points of the -centroid body operator, using spherical harmonic techniques. We also show that the Euclidean ball is locally the only body whose centroid body is a dilate of its polar intersection body.
Key words and phrases:
Convex bodies, spherical harmonics, spherical Radon and Cosine transforms2010 Mathematics Subject Classification:
52A20,33C55,44A121. Introduction
Characterizing the Euclidean space among all real normed spaces is one of the aims of the ten problems formulated in 1956 by Busemann and Petty [4]. These problems lead to the study of certain operators on convex bodies in , such as the intersection body operator for the first Busemann-Petty problem. Proving a particular property of the intersection body operator in each dimension resulted in the resolution of the first Busemann-Petty problem in 1996. In the class of convex bodies, proving properties for operators globally is challenging. The local study of such operators on convex bodies appears to be a more approachable initial step. In [5], it was established that, other than linear transformations of the Euclidean ball, there are no additional fixed points of the intersection body operator in a neighborhood of the Euclidean ball. A similar result was obtained in [12] for the projection body operator. In [1], a local affirmative answer was obtained for the Busemann-Petty problems 5 and 8, by studying the local fixed points of the polar-intersection body operator.
In this paper we conduct a similar local study for several problems involving the centroid body operator. First, we consider its fixed points.
Problem 1.
If for an origin-symmetric convex body , we have
(1) |
where is the centroid body of and is some constant, need be an ellipsoid?
Below we will verify that the ball does indeed satisfy this condition, and that equation (1) is invariant under linear transformations. Consequently, this equation holds for all ellipsoids. We prove the following result.
Theorem 1.
Let . If an origin-symmetric convex body satisfies (1) and is sufficiently close to the Euclidean ball in the Banach-Mazur distance, then must be an ellipsoid.
A similar result is obtained for the -centroid body operators , .
Theorem 2.
Let . If an origin-symmetric convex body satisfies
(2) |
and is sufficiently close to the Euclidean ball in the Banach-Mazur distance, then must be an ellipsoid.
In addition, we prove a local result involving the -centroid body and the polar intersection body of :
Theorem 3.
Let and . If an origin-symmetric convex body satisfies
(3) |
and is sufficiently close to the Euclidean ball in the Banach-Mazur distance, then must be an ellipsoid.
Theorem 3 is similar to several open problems listed by Gardner in [6, pg 336-337], relating the intersection body of with the difference body and the cross section body of :
Problem 8.1 Suppose that is a convex body such that and are homothetic. Must be a centered ellipsoid?
Problem 8.9 (ii) If is homothetic to either or , is an ellipsoid?
In two dimensions, the answer to Problem 1 is known to be affirmative: the only convex bodies which are dilates of their centroid bodies are ellipses. This was proven by Petty in [9, Theorem 5.6]. Indeed, the centroid body of has an important role in fluid statics. If is an origin-symmetric convex body with density 1/2 (with water having density 1), each point on the boundary of the centroid body of is the center of mass of the underwater part of in a given direction, and is the convex hull of the locus of all such centers of mass in all directions. The flotation properties of are governed by three Theorems of Dupin for (see for example [11]). In particular, in two dimensions, the Third Theorem of Dupin states that the radius of curvature of the centroid body at a boundary point perpendicular to a unit direction is proportional to the cube of the length of the chord ,
If we assume that is homothetic to , then the same relation is true between the curvature of at a point and the cube of the length of the corresponding central section. But in the two dimensional case this is a restatement of the Busemann-Petty Problem number 8 in dimension 2 [4, 9].
2. Analytic statement of the problems
Let be the unit sphere in , where denotes the Euclidean norm. Let be the spherical Lebesgue measure on and be the -dimensional Lebesgue measure. A convex body is a convex compact subset of with non-empty interior. We will assume that the origin is an interior point of convex bodies. is origin symmetric if for every , .
Given a convex body , the support function of is defined as
and its radial function is
A body is called the centroid body of a convex body if
The body is denoted as . Switching this integral to polar coordinates and simplifying yields
Examining the right hand side, we observe that it is a multiple of the cosine transform. The cosine transform of an integrable function is defined as
Using this to rewrite the support function of the centroid body gives
(4) |
Equation (1) can now be rewritten analytically as
(5) |
A body is called the intersection body of the convex body if
The body is denoted as . Expressing the volume as an integral and passing to polar coordinates, we obtain
The right hand side is almost the spherical Radon transform. The spherical Radon transform of an integrable function is denoted and is defined as
Thus, the radial function of the intersection body can be expressed in terms of the Radon transform of the radial function of the original body, by
(6) |
A body is the polar body of a convex body if
The body is denoted by . The polar intersection body of a body is . Hence,
Combining this with equation (4) gives the reformulation of as the following equation,
(7) |
To simplify the exposition, we will first consider the case . The analytic reformulation of equation (2) will be derived in Section 5, and the analytic reformulation of (3) will be considered in Section 6.
Remark: Even though Theorem 1 does not make such assumptions on the body , if satisfies (1), then it must be centrally symmetric, strictly convex and smooth, since has those properties (see [6, Sec. 9.1]).
2.1. Ellipsoids as solutions
2.1.1.
2.1.2.
Let be the constant associated with a body that satisfies the condition of Theorem 3. If we evaluate (7) when is the Euclidean ball, we see
To show that all ellipsoids are fixed points, we need to check how linear transformations interact with (3). Using the calculations found in Gardner’s book [6, pgs 21 & 308], the equation below follows.
Therefore if satisfies , so does , albeit for a different constant, .
2.2. From the Banach-Mazur distance to the Hausdorff distance
2.2.1.
The Banach-Mazur distance (see [13, pg 589]) between two origin symmetric convex bodies and , is defined as
Let and suppose that . Taking the linear transformation where the minimum is achieved yields
Then, for the radial and support functions of we have
(8) |
and taking volumes on each side gives
(9) |
While in similar problems (see, for example, [1, 5]), an appropriate dilation of the body can be chosen so that , here we have that the constant in (1) is invariant under linear transformations, making this choice impossible. Therefore, we will use equations (8) and (9) to obtain a bound on the constant in terms of and .
The cosine transform is a positive operator and as such it preserves inequalities. Applying the cosine transform to all sides of the radial function estimate (8), and adjusting the constants, we obtain
Using equation (4), we can rewrite all three terms in the inequality as
Therefore, if , we have
Now we use the support function estimate in (8), together with (9), to obtain
Therefore, .
Among all linear transformations, we will choose one that places in the isotropic position, i.e., the position where
(10) |
It should be noted that the isotropic position as stated above is not unique as any dilate of must also satisfy (10). We are interested in convex bodies close to the Euclidean ball and will accordingly select a dilate where per (9). See, for example [3, Sec. 2.3.2] for more details about the derivation of the isotropic position. In [1, Sec. 5], it is shown that if and , then for the isotropic position of with , we have
and the above considerations guarantee that the constant still satisfies .
2.2.2.
3. Spherical harmonics
Spherical harmonics are homogeneous harmonic polynomials restricted to the sphere. They form the eigenspaces for a collection of common linear operators, such as the spherical Radon transform, the cosine transform and the Laplace-Baltrami operator on the sphere. In this section, we will mostly follow the expositions from Groemer [7] and Atkison-Han [2].
The space of all harmonic, homogeneous polynomials of degree on variables whose domain is restricted to the sphere is denoted . Its elements are called the spherical harmonics of degree . The spherical harmonic spaces are orthogonal with respect to the inner product. The space of all finite sums of spherical harmonics on variables is denoted by , and called the space of spherical harmonics.
The space of spherical harmonics is dense in and by extension also in . The orthogonal projection onto the spherical harmonics of degree , will be denoted (see [2, Def. 2.11] for an explicit definition of the projection). The eigenvalues of the cosine transform on each of the spherical harmonic spaces are (see [7, Lemmas 3.4.1, 3.4.5)]), for , and for even
where . Since the cosine transform is defined as an integral over a symmetric domain, when is odd.
When the body is placed in the isotropic position, we have the following estimate of the second harmonic of , obtained in [1, Sec. 9]. Let be the mean value of on the sphere, and let be the spherical harmonic decomposition of . From the definition (10) of the isotropic position, we have that if is a quadratic harmonic polynomial, with , then
This means has no second order term in its spherical harmonic decomposition. If , additionally, was also close to the Euclidean ball in the sense , then should also be small in the sense. Indeed, estimating the error of the first order Taylor polynomial of at , yields
The second harmonic of can now be estimated from using the previous equation. Applying the to each side and using the orthogonality of the spherical harmonic spaces, we obtain
which yields that
(11) |
3.1. Linear approximation of
In [1, Sec. 9], an estimate on the linear component of the operator is obtained. In this subsection we derive a similar estimate for the operator that appears in our problem.
Let , we know that the first order Taylor polynomial of at is
(12) |
Take be the constant , , and be . We can then estimate the error of the first order approximation of when and is small. In particular, we see the Lipschitz constant appear in the error bound,
Observe that for any function , . That is is a positive operator in the vector space sense. Using the fact that is a positive linear operator,
As positive linear operators admit a triangle inequality , we get
Finally, let us examine the first order Taylor polynomial (12) again but with , , and .
Estimating the error of the approximation, we then see
(13) |
where the Lipschitz constant is dependent on the value of , , , and the maximum value of allowed. Notably, this is bounded Note that in the estimate of (13) only the positivity of the linear operator was used. Considering some other positive linear operator (positive in the when sense), the particular constant will change but an identical estimate holds for .
The next lemma requires a definition. Let be a bounded linear operator on whose eigenspaces are precisely , that is, there exist eigenvalues such that for all ,
Then is said to be a strong contraction if
It should be noted that small enough constant multiples of the cosine and Radon transforms are strong contractions, for example. The main result about the strong contraction is [1, Lemma 4], stated below.
Lemma 4. Assume is a strong contraction. Then there exists such that for any symmetric convex body and any , the conditions
imply const. Here and are the constant terms of the spherical harmonic decomposition.
4. Proof of Theorem 1
Theorem 1. There exists an , such that the only convex bodies satisfying (1) with are ellipsoids.
Proof.
This will be proven in two steps, first under certain assumptions on and then in general. First, we will assume that satisfies (1), is close to the sphere (in the sense that ), and is placed in the isotropic position with .
Using the same approach as in the papers [1, 5, 12], we will separate the operator into linear and non-linear components and examine each separately. With our assumptions on , equation (5) reads as
As we saw in Section 3, we write , where is the zero harmonic of . Since is a linear operator and constants are eigenvectors of it, we have
(14) |
where contains all the higher order terms. Thus, the linear component may be rewritten as,
(15) |
Estimating the higher order terms, we solve for in (14) and take norms yielding
Using (13) with we obtain
In the notation of Lemma 4, we have , , and . The degree zero harmonic of is zero, and the second harmonic of has already been estimated by (11). The eigenvalues of operator are scalar multiples of the eigenvalues of the cosine transform starting with degree 4. Since the even degree eigenvalues decay monotonically to zero in absolute value, we only need to check that the first non-zero eigenvalue is less than one in absolute value. Indeed, , and is a strong contraction as desired. By lemma 4, there is some where if satisfies,
-
•
,
-
•
,
-
•
,
must be the Euclidean ball.
First, we note that the constant is bounded by the product of the bounds on and the bounds on the ratio , giving . If is sufficiently small follows. The second criteria follows if . Checking the third criteria,
Now we need to estimate the expression in parentheses. We have that
is bounded above. Hence, the product of this bound with is bounded above by , provided is small enough.
Now we prove Theorem 1 for all bodies close to the ball. Let be the maximal for which the additional criteria for Lemma 4 hold in the isotropic case. Put
If is a convex body with and satisfying (1), then there exists a linear transformation where satisfies the assumptions of the first part of the proof. Therefore, is a dilate of the Euclidean ball and is an ellipsoid. ∎
5. Generalization to -centroid bodies: Proof of Theorem 2
This section introduces -centroid body and extends the result of the previous section to . Subsection 5.1 gives the definition of the -centroid body and an analytic reformulation in terms of the -cosine transform. As the differences between the prior case and the are purely technical, the subsequent subsections include relevant details for the extension of Theorem 1 to for completeness.
5.1. Definition
The -centroid body of a star body, is defined in a similar way to the centroid body by,
where denotes the integral average and is the -cosine transform,
The -centroid bodies were introduced in [8]. For , the function is the support function of a convex body, so is well defined. In particular for , this is the centroid body studied in the previous sections, and for this is the so called Legendre ellipsoid. For , the relation trivially implies that is an ellipsoid.
For , , we will study if
for some scalar implies that is an ellipsoid, when is close enough to the Euclidean ball in the Banach-Mazur distance. The analytic reformulation of the above equation is
(16) |
5.2. Linear transformations and Stability
As in the case, equation (16) is invariant under linear transformations. If , then
Because the -centroid body operator commutes with linear transformations, if satisfies (16) then so does , and . Since the unit Euclidean ball satisfies (16), we have that ellipsoids also satisfy this equation with the constant
Since in (16) is invariant under linear transformations of the body, we cannot choose a particular value for this constant by picking an appropriate linear transformation. On the other hand, we can obtain stability bounds on in same way as in the case . Let be close to the Euclidean ball, that is for some ,
(17) |
This implies
(18) |
Applying the -cosine transform to the estimate on the radial function in (17), we get
Combining this estimate with (18) and with equation (16) we have
and, since ,
5.3. The eigenvalues of
The eigenspaces of the -cosine transform are, as in the case of the cosine transform, the spaces of spherical harmonics of degree . From [10, Eq. (3.4)] we have that the eigenvalues for are equal to
When is an even integer and , the eigenvalues are 0. The above formula for the eigenvalues can be rewritten so that it works also when is an even integer and . (For the reader who wants to consult reference [10], we note that in that paper the -cosine transform is denoted by ).
Observe that . Expressing the higher order non-zero harmonics as a telescoping product of the lower order ones, we see that
hence the non-zero eigenvalues are absolutely strictly decreasing as , and their limit is . The asymptotic behaviour as varies is similar, for as .
5.4. Linear approximation error
The -cosine tranansform is a positive linear operator, therefore if we set , on the left hand side of (13), we have
Taking the norm of this term and multiplying by the missing constants in equation (16), the left hand side can be rewritten as
(19) |
Recalling that , we can simplify the last term, obtaining
Thus, (19) simplifies to
(20) |
Next, we will consider the right hand side of (13). After multiplying by and taking norms, it can be rewritten as
(21) |
Note that the exponent on is non-negative for and that
We claim that this expression is bounded above by when the dimension is greater than . This is obtained by using the property several times to yield
Breaking up the three terms in the product, we note that the leftmost is bounded by , the middle term is maximal when and (which are the minimal values of ), and the last term is also maximal when and odd (by the convexity of ), yielding
Using this estimate, and the fact that the norm of the normalized -cosine transform is 1, (21) is bounded by
Therefore, combining this with (20), we have the following estimate that we will need to apply Lemma 4.
(22) |
5.5. The proof of Theorem 2.
The last thing that needs to be checked is that the operator in equation (22) is a strong contraction (note that the constant multiplying this operator is of the order , as discussed in subsection 5.2). As mentioned in Section 5.3, the eigenvalues of the -cosine transform for the harmonic spaces of even degree greater than or equal to 4 are less than 1, strictly decreasing in absolute value, and their limit is . Thus, it is enough to check the eigenvalue of the degree four harmonic space of the scaled -cosine transform . The computation yields
From this point, the proof of Theorem 2 is identical to that of Theorem 1.
6. On Theorem 3
6.1. Invariance under linear transformations
6.2. Stability of the constants.
Applying the normalized -cosine transform and Radon transform to , we have
and
If satisfies , taking the ratio of the two terms above and multiplying by (9) yields
Thus, the ratio of the constants is of the order . Applying an appropriate linear transformation to put in the isotropic position, we still have .
6.3. Linearization estimate
From the calculation of the constant for the ball, we immediately have
(23) |
where are the normalized Radon and -Cosine transforms. As we assume , the ratio for sufficiently small . Consequently, we have convergence of the binomial series expansion of . The binomial expansion of the right hand side yields,
Taking the inverse (normalized) Radon transform of both sides of (23), and separating the constant (24), linear (25), and higher order terms (26)-(28) yields
(24) | ||||
(25) | ||||
(26) | ||||
(27) | ||||
(28) |
Assuming that , we observe a few facts for the above equation. The linear term (25) consists of a constant of the order multiplying an operator that will be shown to be a contraction. The higher order terms (26), (27) and (28) will be shown to be of the order of .
6.4. Proof of Theorem 3
We will assume for now that the terms (26)-(28) are indeed of the order , and focus on terms (24) and (25). Taking norms, we obtain,
where is the operator norm. The first inequality follows from the orthogonality of the spherical harmonic spaces and the definition of . If we are able to show that the term in parenthesis on the right hand side is strictly less than one for any given and small enough, then there will exist some such that implies or, equivalently, .
We will check the operator norm , as the terms multiplying it have already been shown to be in subsection 6.2. The eigenspaces of the spherical Radon transform are the spherical harmonic spaces and the eigenvalues of the normalized spherical Radon transform can be seen below
see for example [7, Eq 3.4.16 pg 103]. Combining this with the eigenvalues for the -cosine transform, we have that
For , the right hand side equals 1. As increases, each additional term in the product is strictly less than 1. Therefore this operator norm is less than .
6.5. Higher order terms
The fact that (27) and (28) are of the order is immediate since . This leaves just (26) to be estimated.
We need to introduce a family of operators to help estimate the composition of the inverse Radon transform and the powers of the -cosine transform. Let be the indicator function of the interval , and define by
and and . Observe that
where the last estimates can be found in [2, Eq. 2.48, pg. 27]. For , in as well. Thus, for any
where the last equality is uses the linearity of and the definition of . Breaking each of the summands into their dyadic harmonic projections yields
While the product of two harmonic polynomials need not be harmonic, any polynomial when restricted to the sphere may be written as the sum of homogeneous, harmonic polynomials of lesser or equal degree (see [7, Lemma 3.2.5, pg. 69] or [2, Theorem 2.18, pg 31]). Thus, if the degree of the polynomial in square brackets is smaller than , then the operator will give identically. For non-zero terms, the degree of the polynomial in brackets is
Now, taking the inverse Radon transform and norms on both sides, we see that the right hand side is bounded by
Setting , where , we obtain the following bound on the norm,
Observe that and , so this supremum is indeed finite though dependent on the dimension. Simplifying,
(29) |
where . Now we are in a position to estimate the higher order term (26). Including the binomial coefficient and power of changing with , we see
(30) | |||
The right hand side is of the order times the sum in parentheses. Note that the sums are constant for and at these values of we get a right hand side of
Summing over , we then obtain a right hand side of
The term in square brackets is and will be omitted from now on. Now, summing over we get an upper bound for (26),
Since and , we get the estimate
Evaluating the partial sums of the geometric series, we have
Finally, picking epsilon small enough then yields the desired estimate.
Acknowledgements: I would like to recognize my advisor María de los Ángeles Alfonseca for her guidance in this work. I would also like to thank Fedor Nazarov for his advice to use the operators for the final estimate. I also thank Dmitry Ryabogin and Vladyslav Yaskin for several fruitful discussions.
References
- [1] M. Angeles Alfonseca, Fedor Nazarov, Dmitry Ryabogin, and Vladyslav Yaskin. A solution to the fifth and the eighth Busemann-Petty problems in a small neighborhood of the Euclidean ball. Adv. Math., 390:Paper No. 107920, 28, 2021.
- [2] Kendall Atkinson and Weimin Han. Spherical harmonics and approximations on the unit sphere: an introduction, volume 2044 of Lecture Notes in Mathematics. Springer, Heidelberg, 2012.
- [3] Silouanos Brazitikos, Apostolos Giannopoulos, Petros Valettas, and Beatrice-Helen Vritsiou. Geometry of isotropic convex bodies, volume 196 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2014.
- [4] H. Busemann and C. M. Petty. Problems on convex bodies. Math. Scand., 4:88–94, 1956.
- [5] Alexander Fish, Fedor Nazarov, Dmitry Ryabogin, and Artem Zvavitch. The unit ball is an attractor of the intersection body operator. Adv. Math., 226(3):2629–2642, 2011.
- [6] Richard J. Gardner. Geometric tomography, volume 58 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, New York, second edition, 2006.
- [7] H. Groemer. Geometric applications of Fourier series and spherical harmonics, volume 61 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1996.
- [8] Erwin Lutwak, Deane Yang, and Gaoyong Zhang. affine isoperimetric inequalities. J. Differential Geom., 56(1):111–132, 2000.
- [9] C. M. Petty. On the geometry of the Minkowski plane. Riv. Mat. Univ. Parma, 6:269–292, 1955.
- [10] Boris Rubin. Intersection bodies and generalized cosine transforms. Adv. Math., 218(3):696–727, 2008.
- [11] Dmitry Ryabogin. On bodies floating in equilibrium in every orientation. Geom. Dedicata, 217(4):Paper No. 70, 17, 2023.
- [12] C. Saroglou and A. Zvavitch. Iterations of the projection body operator and a remark on Petty’s conjectured projection inequality. J. Funct. Anal., 272(2):613–630, 2017.
- [13] Rolf Schneider. Convex bodies: the Brunn-Minkowski theory, volume 151 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, expanded edition, 2014.