Next Article in Journal
Numerical Analysis of the Effect of the Vortex Finder on the Hydrocyclone’s Split Water Ratio and Separation Performance
Next Article in Special Issue
Notes on the Transversality Method for Iterated Function Systems—A Survey
Previous Article in Journal
An Efficient Optimal Derivative-Free Fourth-Order Method and Its Memory Variant for Non-Linear Models and Their Dynamics
Previous Article in Special Issue
A Survey on Newhouse Thickness, Fractal Intersections and Patterns
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Survey on the Hausdorff Dimension of Intersections †

Department of Mathematics and Statistics, University of Helsinki, P.O. Box 68, FI-00014 Helsinki, Finland
This survey is based on the talk I gave in Karoly Simon’s 60+1 Birthday Conference in Budapest in June 2022.
Math. Comput. Appl. 2023, 28(2), 49; https://doi.org/10.3390/mca28020049
Submission received: 30 January 2023 / Revised: 9 March 2023 / Accepted: 19 March 2023 / Published: 22 March 2023
(This article belongs to the Special Issue Geometry of Deterministic and Random Fractals)

Abstract

:
Let A and B be Borel subsets of the Euclidean n-space with  dim A + dim B > n . This is a survey on the following question: what can we say about the Hausdorff dimension of the intersections  A ( g ( B ) + z )  for generic orthogonal transformations g and translations by z?

1. Introduction

The books [1,2] contain most of the required background information and the proofs of some of the results discussed below.
Let  L n  stand for the Lebesgue measure on the Euclidean n-space  R n  and let dim stand for the Hausdorff dimension and  H s  for the s-dimensional Hausdorff measure. For  A R n , denote by  M ( A )  the set of Borel measures  μ  with  0 < μ ( A ) <  and with the compact support  spt μ A .
We let  O ( n )  denote the orthogonal group of  R n  and  θ n  its Haar probability measure. The main fact needed about the measure  θ n  is the inequality
θ n ( { g O ( n ) : | x g ( z ) | < r } ) ( r / | z | ) n 1 for x , z R n , r > 0 .
This is quite easy, and is in fact trivial in the plane.
Let A and B be Borel subsets of  R n  with Hausdorff dimensions  s = dim A  and  t = dim B . What can we say about the Hausdorff dimensions of the intersections of A and typical rigid motions of B—more precisely, of  dim A ( g ( B ) + z )  for almost all  g O ( n )  and for  z R n  in a set of positive Lebesgue measure? Optimally, one could hope that this dimension is given by the larger of the numbers  s + t n  and 0, which happens when smooth surfaces meet in a general position.
The problem on the upper bound is much easier than on the lower bound. Let
V z = { ( x , y ) R n × R n : x = y + z } , z R n ,
be the z translate of the diagonal in  R n × R n , and let  π  be the projection  π ( x , y ) = x . Then,
A ( g ( B ) + z ) = π ( ( A × g ( B ) ) V z ) ,
and it follows from a Fubini-type inequality for the Hausdorff dimension [1] (Theorem 7.7) that for any  g O ( n ) ,
dim A ( g ( B ) + z ) dim ( A × B ) n for almost all z R n ,
provided  dim ( A × B ) n . We have always  dim ( A × B ) dim A + dim B  and the equation  dim ( A × B ) = dim A + dim B  holds if, for example,  0 < H s ( A ) < , 0 < H t ( B ) < , and one of the sets has positive lower density, say
θ * s ( A , x ) = lim inf r 0 r s H s ( A B ( x , r ) ) > 0 for H s almost all x A .
Even the weaker condition that the Hausdorff and packing dimensions of A agree suffices; see [1], pp. 115–116. Then, we have
dim A ( g ( B ) + z ) dim A + dim B n for almost all z R n ,
provided  dim A + dim B n . Without some extra condition, this inequality fails seriously: for any  0 s n , there exists a Borel set  A R n  of dimension s such that  dim A f ( A ) = s  for all similarity maps f of  R n . This was proven by Falconer in [3]; see also Example 13.19 in [1] and the further references given there.
We have the lower bound for the dimension of intersections if we use larger transformation groups—for example, similarities.
Theorem 1.
Let A and B be Borel subsets of  R n  with  dim A + dim B > n . Then, for every  ε > 0 ,
L n ( { z R n : dim A ( r g ( B ) + z ) dim A + dim B n ε } ) > 0 ,
for almost all  g O ( n )  and almost all  r > 0 .
If A and B have positive and finite Hausdorff measures,  ε  is not needed. This theorem was proven in the 1980s independently by Kahane [4] and in [5]. More generally, Kahane proved that the similarities can be replaced by any closed subgroup of the general linear group of  R n  that is transitive outside the origin. He gave applications to multiple points of stochastic processes.
There are many special cases where the equality  dim A ( g ( B ) + z ) = dim A + dim B n  holds for almost all g and for z in a set of positive measure. The case where one of the sets is a plane, initiated by Marstrand in [6], has been studied a lot; see discussions in [1] (Chapter 10) and [2] (Chapter 6), and [7] for a more recent result. More generally, one of the sets can be rectifiable; see [5].
The main open problem is as follows: what conditions on the Hausdorff dimensions or measures of A and B guarantee that for  θ n  almost all  g O ( n ) ,
L n ( { z R n : dim A ( g ( B ) + z ) dim A + dim B n } ) > 0 ,
or perhaps for all  ε > 0 ,
L n ( { z R n : dim A ( g ( B ) + z ) dim A + dim B n ε } ) > 0 ?
If one of the sets is a Salem set, i.e., it supports a measure with an optimal Fourier decay allowed by its Hausdorff dimension, then (8) holds without dimensional restrictions; see [8]. I expect (8) to be true for all Borel subsets A and B of  R n .
Below, I shall discuss some partial results on this question. I shall also say something about the exceptional sets of transformations.
In this survey, I shall concentrate on the Hausdorff dimension and general Borel sets. For remarks and references about related results on other dimensions, see [1] (Section 13.20) and [2] (Section 7.3). There is a rich body of literature on various questions about intersections of dynamically generated and related sets. For recent results and further references, see [9,10,11]. For probabilistic sets, see [12] and its references.
I would like to thank the referees for their useful comments.

2. Projections and Plane Intersections

This topic can be thought of as a study of the integral-geometric properties of fractal sets and the Hausdorff dimension. Let us briefly review some of the basic related results on projections and plane sections. This was started by Marstrand in [6] in the plane. His main results in general dimensions are the following. Let  G ( n , m )  be the Grassmannian of linear m-dimensional subspaces of  R n  and  P V : R n V  the orthogonal projection onto  V G ( n , m ) . Let also  γ n , m  be the orthogonally invariant Borel probability measure on  G ( n , m ) .
Theorem 2.
Let  A R n  be a Borel set. Then, for almost all  V G ( n , m ) ,
dim P V ( A ) = dim A if dim A m ,
and
H m ( P V ( A ) ) > 0 if dim A > m .
Theorem 3.
Let  n m s n  and let  A R n  be  H s  measurable with  0 < H s ( A ) < . Then, for almost all  V G ( n , m ) ,
H n m ( { u V : dim ( A ( V + u ) ) = s + m n } ) > 0 ,
and for almost all  V G ( n , n m )  and for  H s  almost all  x A ,
dim ( A ( V + x ) ) = s + m n .
One can sharpen these results by deriving estimates on the Hausdorff dimension of the exceptional sets of the planes V. For the first part of Theorem 2, this was first done by Kaufman in [13] in the plane, and then in [14,15] in higher dimensions. For the second part of Theorem 2, the exceptional set estimates were proven by Falconer in [16]. Thus we have, recall that  dim G ( n , m ) = m ( n m ) .
Theorem 4.
Let  A R n  be a Borel set with  s = dim A . Then,
dim { V G ( n , m ) : P V ( A ) < dim A } s m + m ( n m ) if s m ,
and
dim { V G ( n , m ) : H m ( P V ( A ) ) = 0 } ) m s + m ( n m ) if s > m ,
These inequalities are sharp by the examples in [14] (and their modifications), but the proof for (13) also gives the upper bound  t m + m ( n m )  if  dim A  on the left-hand side is replaced by  t , 0 t dim A . Then, for  t < dim A , this is not always sharp; see the discussion in [2] (Section 5.4).
For the plane sections, Orponen proved in [17]—see also [2] (Theorem 6.7)—the exceptional set estimate (which of course is sharp, as (14) is).
Theorem 5.
Let  n m s n  and let  A R n  be  H s  measurable with  0 < H s ( A ) < . Then, there is a Borel set  E G ( n , m )  such that
dim E n m s + m ( n m )
and for  V G ( n , m ) \ E ,
H n m ( { u V : dim ( A ( V + u ) ) = s + m n } ) > 0 .
We can also ask for exceptional set estimates corresponding to (12). We proved with Orponen [7] the following:
Theorem 6.
Let  n m s n  and let  A R n  be  H s  measurable with  0 < H s ( A ) < . Then, the set B of points  x R n  with
γ n , m ( { V G ( n , m ) : dim A ( V + x ) = s + m n } ) = 0
has dimension  dim B n m .
Very likely, the bound  n m  is not sharp. When  m = 1 , probably, the sharp bound should be  2 ( n 1 ) s  in accordance with Orponen’s sharp result for radial projections in [18].
Another open question is whether there could be some sort of non-trivial estimate for the dimension of the exceptional pairs  ( x , V ) .

3. Some Words about the Methods

The methods in all cases use Frostman measures. Suppose that the Hausdorff measures  H s ( A )  and  H t ( B )  are positive. Then, there are  μ M ( A )  and  ν M ( B )  such that  μ ( B ( x , r ) ) r s  and  ν ( B ( x , r ) ) r t  for  x R n , r > 0 . In particular, for  0 < s < dim A  and  0 < t < dim B , there are  μ M ( A )  and  ν M ( B )  such that  I s ( μ ) <  and  I t ( ν ) < , where the s energy  I s ( μ )  is defined by
I s ( μ ) = | x y | s d μ x d μ y .
Then, the goal is to find intersection measures  λ g , z M ( A ( g ( B ) + z ) )  such that
spt λ g , z spt μ ( g ( spt ν ) + z ) ,
λ g , z ( R n ) d L n z = μ ( R n ) ν ( R n ) for θ n almost all g O ( n ) ,
I s + t n ( λ g , z ) d L n z d θ n g I s ( μ ) I t ( ν ) .
This would give (8).
There are two closely related methods to produce these measures. The first, used in [5], is based on (3): the intersections  A ( g ( B ) + z )  can be realized as level sets of the projections  S g :
S g ( x , y ) = x g ( y ) , x , y R n ,
A ( g ( B ) + z ) = π ( ( A × g ( B ) ) S g 1 { z } ) , π ( x , y ) = x .
Notice that the map  S g  is essentially the orthogonal projection onto the n-plane  { ( x , g ( x ) ) : x R n } .
Thus, one slices (disintegrates)  μ × g # ν  ( g # ν  is the push-forward) with the planes  V z = { ( x , y ) : x = y + z } , z R n .  For this to work, one needs to know that
S g # ( μ × ν ) L n for θ n almost all g O ( n ) .
This is usually proven by establishing the  L 2  estimate
S g # ( μ × ν ) ( x ) 2 d x d θ n g 1 ,
which, by Plancherel’s formula, is equivalent to
F ( S g # ( μ × ν ) ) ( x ) 2 d x d θ n g 1 ,
where  F  stands for the Fourier transform.
The second method, used in [4], is based on convolution approximation. Letting  ψ ε , ε > 0 , be a standard approximate identity, set  ν ε = ψ ε ν  and
ν g , z , ε ( x ) = ν ε ( g 1 ( x z ) ) , x R n .
Then, the plan is to show that when  ε 0 , the measures  ν g , z , ε μ  converge weakly to the desired intersection measures.
No Fourier transform is needed to prove Theorem 1, but the proofs of all theorems discussed below, except Theorems 10 and 11, rely on the Fourier transform defined by
μ ^ ( x ) = e 2 π i x · y d μ y , x R n .
The basic reason for its usefulness in this connection is the formula
I s ( μ ) = | x y | s d μ x d μ y = c ( n , s ) | μ ^ ( x ) | 2 | x | s n d x ,
which is a consequence of Parseval’s formula and the fact that the distributional Fourier transform of the Riesz kernel  k s , k s ( x ) = | x | s , is a constant multiple of  k n s .
The decay of the spherical averages,
σ ( μ ) ( r ) = r 1 n S ( r ) | μ ^ ( x ) | 2 d σ r n 1 x , r > 0 ,
of  μ M ( R n ) , where  σ r n 1  is the surface measure on the sphere  S ( r ) = { x R n : | x | = r } , often plays an important role. By integration in polar coordinates, if  σ ( μ ) ( r ) r t  for  r > 0  and for some  t > 0 , then  I s ( μ ) <  for  0 < s < t . Hence, the best decay that we can hope for under the finite s energy assumption (or the Frostman assumption  μ ( B ( x , r ) ) r s ) ) is  r s . This is true when  s ( n 1 ) / 2 —see [2] (Lemma 3.5)—but false otherwise.
The decay estimates for  σ ( μ ) ( r )  have been studied by many people; a discussion can be found in [2] (Chapter 15). The best-known estimates, due to Wolff [19] when  n = 2  (the proof can also be found in [2] (Chapter 16)) and to Du and Zhang [20] in the general case, are the following. Let  μ M ( R n )  with  μ ( B ( x , r ) ) r s  for  x R n , r > 0 . Then, for all  ε > 0 , r > 1 ,
σ ( μ ) ( r ) r ( n 1 ) s / n + ε for all 0 < s < n , r ( n 1 ) / 2 + ε if ( n 1 ) / 2 s n / 2 , r s + ε if 0 < s ( n 1 ) / 2 .
The essential case for the first estimate is  s > n / 2 ; otherwise, the second and third are better. Up to  ε , these estimates are sharp when  n = 2 . When  n 3 , the sharp bounds are not known for all s; see [21] for a discussion and the most recent examples. As mentioned above, the last bound is always sharp.

4. The First Theorem

If one of the sets has a dimension greater than  ( n + 1 ) / 2 , we have the following theorem. It was proven in [22]; see also [1] (Theorem 13.11) or [2] (Theorem 7.4).
Theorem 7.
Let s and t be positive numbers with  s + t > n  and  s > ( n + 1 ) / 2 . Let A and B be Borel subsets of  R n  with  H s ( A ) > 0  and  H t ( B ) > 0 . Then,
L n ( { z R n : dim A ( g ( B ) + z ) dim A + dim B n } ) > 0
for almost all  g O ( n ) .
The proof is based on the slicing method. The key estimate is
μ × μ ( { ( x , y ) : r | x y | r + δ } ) I s ( μ ) δ r s 1
if  μ M ( R n ) , 0 < δ r  and  ( n + 1 ) / 2 s < n . This is combined with the inequality (1).
The inequality (28) is obtained with the help of the Fourier transform, and that is the only place in the proof of Theorem 7 where the Fourier transform is needed.
One problem of extending Theorem 7 below the dimension bound  ( n + 1 ) / 2  is that the estimate (28) then fails, at least in the plane by [2] (Example 4.9) and in  R 3  by [23].
In Section 7, we discuss estimates on the exceptional sets of orthogonal transformations. The proof of Theorem 13 gives another proof for Theorem 7 but under the stronger assumption  s + t > n + 1 . On the other hand, Theorem 12 below holds with the assumption  s + ( n 1 ) t / n > n  but under the additional condition of positive lower density. Of course,  s + ( n 1 ) t / n > n  is sometimes stronger and sometimes weaker than  s > ( n + 1 ) / 2 , s + t > n . For example, consider these in the plane. When  s = t , the first one says  s > 4 / 3  and the second one  s > 3 / 2 . On the other hand, when s is slightly larger than  3 / 2 , the first requires t to be at least 1, but the second allows  t = 1 / 2 .
Theorem 7 says nothing in  R 1 , and there is nothing to say: in [5], I constructed compact sets  A , B R  such that  dim A = dim B = 1  and  A ( B + z )  contains at most one point for any  z R . With  A , B R  as above, the n-fold Cartesian products  A × × A R n  and  B × × B R n  yield the corresponding examples in  R n —that is, simply, with translations, we obtain nothing in general.
Donoven and Falconer proved in [24] an analogue of Theorem 7 for the isometries of the Cantor space. They did not need any dimensional restrictions. They used martingales to construct the desired random measures with finite energy integrals on the intersections.

5. The Projections  S g

We now discuss further the projections  S g ; recall (19). They are particular cases of restricted projections, which recently have been studied extensively; see [2] (Section 5.4), and [25,26] and the references given there. Restricted means that we are considering a lower-dimensional subspace of the Grassmannian  G ( 2 n , n ) . For the full Grassmannian, we have Marstrand’s projection Theorem 2.
As mentioned above, to prove Theorem 7, one first needs to know (21) when  s + t > n  and  s > ( n + 1 ) / 2  and  μ  and  ν  have finite s and t energies. A simple proof using spherical averages is given in [2] (Lemma 7.1). This immediately yields the weaker result: with the assumptions of Theorem 7, for almost all  g O ( n ) ,
L n ( { z R n : A ( g ( B ) + z ) } ) > 0 ,
because (29) is equivalent to  L n ( S g ( A × B ) ) > 0 . Even for this, I do not know if the assumption  s > ( n + 1 ) / 2  is needed.
Let us first look at general Borel subsets of  R 2 n .
Theorem 8.
Let  A R 2 n  be a Borel set. If  dim A > n + 1 , then  L n ( S g ( A ) ) > 0  for  θ n  almost all  g O ( n ) .
This was proven in [26]. The paper also contains dimension estimates for  S g ( A )  when  dim A n + 1  and estimates on the dimension of exceptional sets of transformations g. In particular, if  n dim A n + 1 , then
dim S g ( A ) dim A 1 for θ n almost all g O ( n ) .
The bound  n + 1  in Theorem 8 is sharp. This was shown by Harris in [27]. First, (30) is sharp. The example for  dim A = n  is simply the diagonal  D = { ( x , x ) : x R n } . To see this, suppose that  g O ( n )  is such that  det g = ( 1 ) n + 1 , which is satisfied by half of the orthogonal transformations. Then, by some linear algebra, g has a fixed point, whence the kernel of  x S g ( x , x )  is non-trivial, so  dim S g ( D ) n 1 . Taking the Cartesian product of D with a one-dimensional set of zero  H 1  measure, we obtain A with  dim A = n + 1  and  L n ( S g ( A ) ) = 0 , which proves the sharpness.
However, this only gives an example A of dimension  n + 1  for which  L n ( S g ( A ) ) = 0  for  g O ( n )  with measure  1 / 2 . Is there a counter-example that works for almost all  g O ( n ) ?
Here are the basic ingredients of the proof of Theorem 8. They were inspired by Oberlin’s paper [28].
Let  0 < n + 1 < s < dim A  and  μ M ( A )  with  I s ( μ ) < , and let  μ g M ( S g ( A ) )  be the push-forward of  μ  under  S g . The Fourier transform of  μ g  is given by
μ g ^ ( ξ ) = μ ^ ( ξ , g 1 ( ξ ) ) .
By fairly standard arguments, using also the inequality (1), one can then show that for  R > 1 ,
R | ξ | 2 R | μ ^ ( ξ , g 1 ( ξ ) ) | 2 d ξ d θ n g R n + 1 s .
This is summed over the dyadic annuli,  R = 2 k , k = 1 , 2 , . The sum converges since  s > n + 1 . Hence, for  θ n  almost all  g O ( n ) μ g  is absolutely continuous with  L 2  density, and so  L n ( S g ( A ) ) > 0 .
For product sets, we can improve this, which is essential for the applications to intersections.
Theorem 9.
Let  A , B R n  be Borel sets. If  dim A + ( n 1 ) dim B / n > n  or  dim A + dim B > n  and  dim A > ( n + 1 ) / 2 , then  L n ( S g ( A × B ) ) > 0  for  θ n  almost all  g O ( n ) .
The case  dim A > ( n + 1 ) / 2  is a special case of Theorem 7; recall (29). The proof of the case  dim A + ( n 1 ) dim B / n > n  is based on the spherical averages and the first estimate of (26). Here is a sketch.
Let  0 < s < dim A , 0 < t < dim B  and  ε > 0  such that  s + ( n 1 ) t / n ε > n , and let  μ M ( A ) , ν M ( B )  with  μ ( B ( x , r ) ) r s , ν ( B ( x , r ) ) r t  for  x R n , r > 0 . Let  λ g = S g # ( μ × ν ) M ( S g ( A × B ) ) .  Then,  λ g ^ ( ξ ) = μ ^ ( ξ ) ν ^ ( g 1 ( ξ ) ) . By (26), we have
| λ g ^ ( ξ ) | 2 d ξ d θ g = | μ ^ ( ξ ) | 2 σ ( ν ) ( | ξ | ) d ξ | μ ^ ( ξ ) | 2 | ξ | ( n 1 ) t / n + ε d ξ = c I n ( n 1 ) t / n + ε ( μ ) I s ( μ ) < .
This gives Theorem 9.
In fact, for some results on the intersections below, we again need absolute continuity as in (21). In the case  s + ( n 1 ) t > n , we need the quantitative estimate: if  s + ( n 1 ) t > n , μ , ν M ( R n )  and  μ ( B ( x , r ) ) r s , ν ( B ( x , r ) ) r t  for  x R n , r > 0 , then
S g # ( μ × ν ) ( x ) 2 d x d θ n g 1 ,
with the implicit constant independent of  μ  and  ν . The arguments described above give this too.

6. Level Sets and Intersections

The estimate (33) can be used to derive information on the Hausdorff dimension of the level sets of  S g , and hence, by (20), of intersections. The following results were proven in [29]. We shall first discuss a more general version of this principle: a quantitative projection theorem leads to estimates of the Hausdorff dimension of level sets. This is also how, in [1] (Chapter 10), the proof for Marstrand’s section Theorem 3 runs.
We consider the following general setting. Let  P λ : R n R m , λ Λ ,  be orthogonal projections, where  Λ  is a compact metric space. Suppose that  λ P λ x  is continuous for every  x R n . Let also  ω  be a finite non-zero Borel measure on  Λ . We denote by  D ( μ , · )  the Radon–Nikodym derivative of a measure  μ  on  R m .
Theorem 10.
Let  s > m . Suppose that there exists a positive number C such that  P λ μ L m  for ω almost all  λ Λ  and
D ( P λ μ , u ) 2 d L m u d ω λ < C
whenever  μ M ( B n ( 0 , 1 ) )  is such that  μ ( B ( x , r ) ) r s  for  x R n , r > 0 .
If  A R n  is  H s  measurable,  0 < H s ( A ) <  and  θ * s ( A , x ) > 0  (recall (5)) for  H s  almost all  x A , then for ω almost all  λ Λ ,
L m ( { u R m : dim P λ 1 { u } A = s m } ) > 0 .
For an application to intersections, we shall need the following product set version of Theorem 10. There,  P λ : R n × R p R m , λ Λ , m < n + p ,  are orthogonal projections with the same assumptions as before.
Theorem 11.
Let  s , t > 0  with  s + t > m . Suppose that there exists a positive number C such that  P λ ( μ × ν ) L m  for ω almost all  λ Λ  and
D ( P λ ( μ × ν ) , u ) 2 d L m u d ω λ < C
whenever  μ M ( B n ( 0 , 1 ) ) , ν M ( B p ( 0 , 1 ) )  are such that  μ ( B ( x , r ) ) r s  for  x R n , r > 0 , and  ν ( B ( y , r ) ) r t  for  y R p , r > 0 .
If  A R n  is  H s  measurable,  0 < H s ( A ) < B R p  is  H t  measurable,  0 < H t ( B ) < θ * s ( A , x ) > 0  for  H s  almost all  x A , and  θ * t ( B , y ) > 0  for  H t  almost all  y B , then for ω almost all  λ Λ ,
L m ( { u R m : dim P λ 1 { u } ( A × B ) = s + t m } ) > 0 .
I do not know if the assumptions on positive lower density are needed.
I give a few words about the proof of Theorem 10. First, notice that  D ( P λ ( μ ) , u )  is given by
D ( P λ μ , u ) = lim δ 0 L m ( B ( 0 , 1 ) ) 1 δ m μ ( { y : | P λ ( y ) u | δ } ) .
Let  μ  be the restriction of  H s  to a subset of A so that  μ  satisfies the Frostman s condition. Then, (34) is applied to the measures
μ a , r , δ = r s T a , r ( μ δ B ( a , r ) ) M ( B ( 0 , 1 ) ) , a R n , r > 0 , δ > 0 ,
where  μ δ ( B ) = δ n B μ ( B ( x , r ) ) d L n x T a , r ( x ) = ( x a ) / r  is the blow-up map and  μ δ B ( a , r )  is the restriction of  μ δ  to  B ( a , r ) . This leads for almost all  x A , λ Λ , to
lim r 0 lim inf δ 0 r t δ m μ ( { y B ( x , r ) : | P λ ( y x ) | δ } ) = 0 ,
which is a Frostman-type condition along the level sets of the  P λ . With some further work, it leads to (35). The proof of Theorem 11 is similar.
Theorem 11, together with the quantitative version of Theorem 9 and with (20), can be applied to the projections  S g  to obtain the following result on the Hausdorff dimension of intersections.
Theorem 12.
Let  s , t > 0  with  s + ( n 1 ) t / n > n  and let  A R n  be  H s  measurable with  0 < H s ( A ) < , and let  B R n  be  H t  measurable with  0 < H t ( B ) < . Suppose that  θ * s ( A , x ) > 0  for  H s  almost all  x A  and  θ * t ( B , y ) > 0  for  H t  almost all  y B . Then, for  θ n  almost all  g O ( n ) ,
L n ( { z R n : dim A ( g ( B ) + z ) = s + t n } ) > 0 .
Again, I do not know if the positive lower density assumptions are needed for the lower bound  s + t n . As mentioned before, they are needed for the upper bound.

7. Exceptional Set Estimates

Recall the exceptional set estimates for orthogonal projections and for intersections with planes from Section 2. Now, we discuss some similar results from [30] for intersections.
First, we have an exceptional set estimate related to Theorem 7. However, we need a slightly stronger assumption: the sum of the dimensions is required to be larger than  n + 1 , rather than only one of the sets having a dimension larger than  ( n + 1 ) / 2 . Recall that the dimension of  O ( n )  is  n ( n 1 ) / 2 .
Theorem 13.
Let s and t be positive numbers with  s + t > n + 1 . Let A and B be Borel subsets of  R n  with  H s ( A ) > 0  and  H t ( B ) > 0 . Then, there is  E O ( n )  such that
dim E n ( n 1 ) / 2 ( s + t ( n + 1 ) )
and for  g O ( n ) \ E ,
L n ( { z R n : dim A ( g ( B ) + z ) s + t n } ) > 0 .
The proof is based on the Fourier transform and the convolution approximation mentioned in Section 3. Instead of  θ n , one uses a Frostman measure  θ  on the exceptional set E: if  α > ( n 1 ) ( n 2 ) / 2  is such that  θ ( B ( g , r ) ) r α  for all  g O ( n )  and  r > 0 , then for  x , z R n \ { 0 } , r > 0 ,
θ ( { g : | x g ( z ) | < r } ) ( r / | z | ) α ( n 1 ) ( n 2 ) / 2 .
This replaces the inequality (1).
In the case where one of the sets has a small dimension, we have the following improvement of Theorem 13.
Theorem 14.
Let A and B be Borel subsets of  R n  and suppose that  dim A ( n 1 ) / 2 . If  0 < u < dim A + dim B n , then there is  E O ( n )  with
dim E n ( n 1 ) / 2 u
such that for  g O ( n ) \ E ,
L n ( { z R n : dim A ( g ( B ) + z ) u } ) > 0 .
The last decay estimate in (26) of spherical averages is essential for the proof. The reason that the assumption  dim A ( n 1 ) / 2  leads to a better result is that this estimate in (26) is stronger than the others. For  dim A > ( n 1 ) / 2 , the inequalities (26) would only give weaker results with u replaced by a smaller number; see [30] (Section 4).
If one of the sets supports a measure with sufficiently fast decay of the averages  σ ( μ ) ( r ) , we can improve the estimate of Theorem 13. Then, the results even hold without any rotations, provided that the dimensions are large enough. In particular, we have the following result in the event that one of the sets is a Salem set. By definition, A is a Salem set if, for every  0 < s < dim A , there is  μ M ( A )  such that  | μ ^ ( x ) | 2 | x | s . A discussion on Salem sets can be found, for example, in [2], Section 3.6.
Theorem 15.
Let A and B be Borel subsets of  R n  and suppose that A is a Salem set. Suppose that  0 < u < dim A + dim B n .
(a)If  dim A + dim B > 2 n 1 , then
L n ( { z R n : dim A ( B + z ) u } ) > 0 .
(b)If  dim A + dim B 2 n 1 , then there is  E O ( n )  with
dim E n ( n 1 ) / 2 u
such that for  g O ( n ) \ E ,
L n ( { z R n : dim A ( g ( B ) + z ) u } ) > 0 .
Could this hold for general sets, perhaps in the form that  dim E = 0 , if  dim A + dim B > 2 n 1 ? It follows from Theorem 5 that this is true if one of the sets is a plane. In  R 2 , a slightly stronger question reads as follows: if  s + t > 2  and A and B are Borel subsets of  R 2  with  H s ( A ) > 0  and  H t ( B ) > 0 , is there  E O ( 2 )  with  dim E = 0 , if  s + t 3 , and  dim E 3 s t , if  s + t 3 , such that for  g O ( 2 ) \ E ,
L 2 ( { z R 2 : dim A ( g ( B ) + z ) s + t 2 } ) > 0 ?

8. Some Relations to the Distance Set Problem

There are some connections of this topic to Falconer’s distance set problem. For a general discussion and references, see, for example, [2]. Falconer showed in [31] that, for a Borel set  A R n , the distance set  { | x y | : x , y A }  has a positive Lebesgue measure if  dim A > ( n + 1 ) / 2 . We had the same condition in Theorem 7. Moreover, for distance sets, it is expected that  dim A > n / 2  should be enough.
When  n = 2 , Wolff [19] improved  3 / 2  to  4 / 3  using (26). Observe that when  dim A = dim B , the assumption  dim A + dim B / 2 > 2  in Theorem 12 becomes  dim A > 4 / 3  and it is the same as Wolff’s. This is no coincidence: both results use Wolff’s estimate (26).
The proofs of distance set results often involve the distance measure  δ ( μ )  of a measure  μ  defined by
δ ( μ ) ( B ) = μ × μ ( { ( x , y ) : | x y | B } ) , B R .
The crucial estimate (28) means that  δ ( μ )  is absolutely continuous with bounded density if  I ( n + 1 ) / 2 ( μ ) < . Hence, it yields Falconer’s result. As mentioned before, we cannot hope to obtain bounded density when  s < ( n + 1 ) / 2 , at least when  n = 2  or 3. In many of the later improvements, one verifies absolute continuity with  L 2  density. For example, Wolff showed that  δ ( μ ) L 2 ( R ) , if  I s ( μ ) <  for some  s > 4 / 3 . To do this, he used decay estimates for the spherical averages  σ ( μ ) ( r )  and proved (26) for  n = 2 . The proofs of the most recent, and so far the best known, distance set results in [20,32,33,34] involve using deep harmonic analysis techniques; restriction and decoupling. In the plane, the result of [33] says that the distance set of A has a positive Lebesgue measure if  dim A > 5 / 4 . See Shmerkin’s survey [35] for the distance set and related problems.
Distance measures are related to the projections  S g  by the following:
D ( S g # ( μ × ν ) ) ( z ) 2 d L n z d θ n g = c δ ( μ ) ( t ) δ ( ν ) ( t ) t 1 n d t ,
at least if  μ  and  ν  are smooth functions with compact support; see [26] (Section 5.2).
Since, by an example in [33], when  n = 2 , for any  s < 4 / 3 I s ( μ ) <  is not enough for  δ ( μ )  to be in  L 2 , probably, because of (45), it is not enough for  S g # ( μ × μ )  to be in  L 2 . However, in [33], it was shown that if  I s ( μ ) <  for some  s > 5 / 4 , there is a complex valued modification of  μ  with good  L 2  behavior. In even higher dimensions, similar results were proven in [34] with  n / 2 + 1 / 4  in place of  5 / 4 . Could these methods be used to show, for instance, that if  n = 2  and  dim A = dim B > 5 / 4 , then  L 2 ( S g ( A × B ) ) > 0  for almost all  g O ( 2 ) ?

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Mattila, P. Geometry of Sets and Measures in Euclidean Spaces; Cambridge University Press: Cambridge, UK, 1995. [Google Scholar]
  2. Mattila, P. Fourier Analysis and Hausdorff Dimension; Cambridge University Press: Cambridge, UK, 2015. [Google Scholar]
  3. Falconer, K.J. Classes of sets with large intersection. Mathematika 1985, 32, 191–205. [Google Scholar] [CrossRef]
  4. Kahane, J.-P. Sur la dimension des intersections. Asp. Math. Appl. North-Holl. Math. Libr. 1986, 34, 419–430. [Google Scholar]
  5. Mattila, P. Hausdorff dimension and capacities of intersections of sets in n-space. Acta Math. 1984, 152, 77–105. [Google Scholar] [CrossRef]
  6. Marstrand, J.M. Some fundamental geometrical properties of plane sets of fractional dimensions. Proc. Lond. Math. Soc. 1954, 4, 257–302. [Google Scholar] [CrossRef]
  7. Mattila, P.; Orponen, T. Hausdorff dimension, intersection of projections and exceptional plane sections. Proc. Am. Math. Soc. 2016, 144, 3419–3430. [Google Scholar] [CrossRef] [Green Version]
  8. Mattila, P. Spherical averages of Fourier transforms of measures with finite energy; dimension of intersections and distance sets. Mathematika 1987, 34, 207–228. [Google Scholar] [CrossRef]
  9. Shmerkin, P. On Furstenberg’s intersection conjecture, self-similar measures, and the Lq norms of convolutions. Ann. Math. 2019, 189, 319–391. [Google Scholar]
  10. Wu, M. A proof of Furstenberg’s conjecture on the intersections of ×p- and ×q-invariant sets. Ann. Math. 2019, 189, 707–751. [Google Scholar] [CrossRef] [Green Version]
  11. Yavicoli, A. A Survey on Newhouse Thickness, Fractal Intersections and Patterns. Math. Comput. Appl. 2022, 27, 111. [Google Scholar] [CrossRef]
  12. Shmerkin, P.; Suomala, V. Spatially Independent Martingales, Intersections, and Applications; Memoirs of the American Mathematical Society: Providence, RI, USA, 2018; Volume 251, 1195p. [Google Scholar]
  13. Kaufman, R. On Hausdorff dimension of projections. Mathematika 1968, 15, 153–155. [Google Scholar] [CrossRef]
  14. Kaufman, R.; Mattila, P. Hausdorff dimension and exceptional sets of linear transformations. Ann. Acad. Sci. Fenn. A Math. 1975, 1, 387–392. [Google Scholar] [CrossRef]
  15. Mattila, P. Hausdorff dimension, orthogonal projections and intersections with planes. Ann. Acad. Sci. Fenn. A Math. 1975, 1, 227–244. [Google Scholar] [CrossRef]
  16. Falconer, K.J. Hausdorff dimension and the exceptional set of projections. Mathematika 1982, 29, 109–115. [Google Scholar] [CrossRef]
  17. Orponen, T. Slicing sets and measures, and the dimension of exceptional parameters. J. Geom. Anal. 2014, 24, 47–80. [Google Scholar] [CrossRef] [Green Version]
  18. Orponen, T. A sharp exceptional set estimate for visibility. Bull. Lond. Math. Soc. 2018, 50, 1–6. [Google Scholar] [CrossRef] [Green Version]
  19. Wolff, T.W. Decay of circular means of Fourier transforms of measures. Int. Math. Res. Not. 1999, 10, 547–567. [Google Scholar] [CrossRef]
  20. Du, X.; Zhang, R. Sharp L2 estimates of the Schrödinger maximal function in higher dimensions. Ann. Math. 2019, 189, 837–861. [Google Scholar] [CrossRef]
  21. Du, X. Upper bounds of Fourier decay rate of fractal measures. J. Lond. Math. Soc. 2020, 102, 1318–1336. [Google Scholar] [CrossRef]
  22. Mattila, P. On the Hausdorff dimension and capacities of intersections. Mathematika 1985, 32, 213–217. [Google Scholar] [CrossRef]
  23. Iosevich, A.; Senger, S. Sharpness of Falconer’s d+12 estimate. Ann. Sci. Fenn. Math. 2016, 41, 713–720. [Google Scholar] [CrossRef]
  24. Donoven, C.; Falconer, K.J. Codimension formulae for the intersection of fractal subsets of Cantor spaces. Proc. Amer. Math. Soc. 2016, 144, 651–663. [Google Scholar] [CrossRef] [Green Version]
  25. Gan, S.; Guo, S.; Guth, L.; Harris, T.J.; Maldague, D.; Wang, H. On restricted projections to planes in R3. arXiv 2022, arXiv:2207.13844. [Google Scholar]
  26. Mattila, P. Hausdorff dimension and projections related to intersections. Publ. Mat. 2022, 66, 305–323. [Google Scholar] [CrossRef]
  27. Harris, T.L.J. Restricted families of projections onto planes: The general case of nonvanishing geodesic curvature. Anal. PDE 2022, 15, 1655–1701. [Google Scholar] [CrossRef]
  28. Oberlin, D.M. Exceptional sets of projections, unions of k-planes, and associated transforms. Israel J. Math. 2014, 202, 331–342. [Google Scholar] [CrossRef] [Green Version]
  29. Mattila, P. Hausdorff dimension of intersections with planes and general sets. J. Fractal Geom. 2021, 66, 389–401. [Google Scholar] [CrossRef]
  30. Mattila, P. Exceptional set estimates for the Hausdorff dimension of intersections. Ann. Acad. Sci. Fenn. A Math. 2017, 42, 611–620. [Google Scholar] [CrossRef]
  31. Falconer, K.J. On the Hausdorff dimension of distance sets. Mathematika 1982, 29, 206–212. [Google Scholar] [CrossRef]
  32. Du, X.; Guth, L.; Ou, Y.; Wang, H.; Wilson, B.; Zhang, R. Weighted restriction estimates and application to Falconer distance set problem. Amer. J. Math. 2021, 143, 175–211. [Google Scholar] [CrossRef]
  33. Guth, L.; Iosevich, A.; Ou, Y.; Wang, H. On Falconer’s distance set problem in the plane. Invent. Math. 2020, 219, 779–830. [Google Scholar] [CrossRef] [Green Version]
  34. Du, X.; Iosevich, I.; Ou, Y.; Wang, H.; Zhang, R. An improved result for Falconer’s distance set problem in even dimensions. Math. Ann. 2021, 380, 1215–1231. [Google Scholar] [CrossRef]
  35. Shmerkin, P. Slices and distances: On two problems of Furstenberg and Falconer. arXiv 2021, arXiv:2109.12157. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mattila, P. A Survey on the Hausdorff Dimension of Intersections. Math. Comput. Appl. 2023, 28, 49. https://doi.org/10.3390/mca28020049

AMA Style

Mattila P. A Survey on the Hausdorff Dimension of Intersections. Mathematical and Computational Applications. 2023; 28(2):49. https://doi.org/10.3390/mca28020049

Chicago/Turabian Style

Mattila, Pertti. 2023. "A Survey on the Hausdorff Dimension of Intersections" Mathematical and Computational Applications 28, no. 2: 49. https://doi.org/10.3390/mca28020049

Article Metrics

Back to TopTop