On the Hausdorff dimension of Weighted Singular Vectors
Abstract.
We prove a sharp upper bound on the Hausdorff dimension of weighted singular vectors in using dynamics on homogeneous spaces, specifically the method of integral inequalities. Together with the lower bound proved recently by Kim and Park [KimPark2024], this determines the Hausdorff dimension of weighted singular vectors, thereby generalizing to arbitrary dimension, the work of Liao, Shi, Solan, and Tamam [LSST], who determined the Hausdorff dimension of weighted singular vectors in two dimensions. We also provide the first known bounds for the Hausdorff dimension of weighted singular vectors restricted to fractal subsets.
Key words and phrases:
Diophantine approximation, ergodic theory, Hausdorff dimension, flows on homogeneous spaces2020 Mathematics Subject Classification:
11J13, 11J83, 37A171. Introduction
Singular vectors are a class of vectors for which Dirichlet’s theorem in Diophantine approximation can be ‘infinitely improved’. More precisely,
Definition 1.1.
Fix and such that and . A vector is called -singular if for any , there exists such that for all the following system of inequalities
has an integer solution . We denote by , the set of all -singular vectors in .
While the study of the classical ‘non-weighted’ situation, i.e. the study of is more ubiquitous, Diophantine approximation with weights has been extensively studied in recent times. We refer the reader to [CGGMS] for an introduction. The set of singular vectors is clearly non-empty and due to a classical result of Khintchine, it contains uncountably many vectors when the dimension is greater than . Moreover, a modification of a classical argument in [Casselsbook] shows that the set of singular vectors has zero Lebesgue measure. There is a closely related notion of Dirichlet improvable vectors, see [KleinbockRao] for the relation between singular vectors with Dirichlet improvable vectors with respect to different norms.
Two central questions about singular vectors have recently been the focus of several important works:
-
(1)
to estimate the measure of with respect to other natural measures. This problem encapsulates the themes of Diophantine approximation on manifolds as well as that on fractals,
-
(2)
to estimate the Hausdorff dimension of , again one could ask for the Hausdorff dimension of restricted to a manifold or a fractal.
Following work of several authors, an approach to question 1 was developed by Kleinbock and Weiss in [KWsingular] where it was shown that if belongs to a class of measures called friendly measures, then . Friendly measures form a reasonably large class of measures and include some IFS’s as well as pushforwards of Lebesgue measure by smooth maps. There has been recent progress in showing the existence of weighted singular vectors on manifolds, cf. [KMWW, DattaTamam].
The second question, i.e. of estimating the Hausdorff dimension of turns out to be a difficult problem and has received much attention in the last two decades. In a landmark work [Cheung], Y. Cheung showed that the dimensionof is . This was subsequently generalized to in an important work of Cheung and Chevallier [CheungChevallier]. Another important result was obtained by Kadyrov, Kleinbock, Lindenstrauss and Margulis in [KKLM] using methods from homogeneous dynamics. Namely, a sharp upper bound on the more general set of singular on average matrices was obtained using integral inequalities as introduced in the famous work [EMM] on the Oppenheim conjecture. The complementary lower bound was obtained by Das, Fishman, Simmons and Urbanski in [DFSU] using methods from the parametric geometry of numbers, see also the recent paper [Solan] for more results in this direction. In [Khalilsing], Khalil upgraded the integral inequality approach of [KKLM] using a beautiful argument to deal with singular vectors on fractals. That is, an upper bound on the Hausdorff dimension of singular vectors lying on self-similar fractals in that satisfied the open set condition was obtained. In [ShahYang], Shah and Yang obtained bounds for the dimension of certain singular vectors lying on affine subspaces.
All the impressive results towards question 2 discussed above are in the ‘unweighted’ setting. The literature in the more complicated weighted setting is sparser. In [LSST], the Hausdorff dimension of was shown to be equal to . As far as higher dimensions are concerned, the authors (page 836 in [LSST]) write that “it is likely” that the dimension of weighted singular vectors in is “where is the top Lyapunov exponent for the adjoint action of the corresponding one-parameter semigroup on the corresponding unipotent group.”
This expectation turns out to be accurate. The lower bound was recently obtained by Kim and Park in [KimPark2024] using an appropriate generalization of the technique of [LSST] that involves the construction of a fractal set contained in . A corollary of our main result provides the matching upper bound, and so together we have
Theorem 1.2.
Fix and such that and . Then the Hausdorff dimension of satisfies
(1) |
In fact, our methods also allow us to deal with Diophantine approximation on fractals; providing the very first dimension bounds for weighted singular vectors on fractals. Here is another corollary of our main theorem.
Theorem 1.3.
Fix . Fix such that and . For each , let be an IFS consisting of contractive similarities with equal contraction ratios on and satisfying the open set condition. Let be the limit set of and . Then, the dimension of satisfies
(2) |
where .
We will now state our main theorem, postponing some notation for later. Our approach is dynamical in nature and owes its existence to an influential work of Dani [Dani-Crelle] where a connection was established between Diophantine properties of vectors and dynamical and topological properties of certain orbits of an associated unimodular lattice. Let be the space of unimodular lattices in . For and , define
(3) |
For define the set
By Dani’s results mentioned above (the famous “Dani correspondence”), singular vectors correspond exactly to the above set for . The main result of the paper is the following theorem.
Theorem 1.4.
Fix . Fix such that and . For each , let be an IFS consisting of contractive similarities with equal contraction ratios on and satisfying the open set condition. Let be the limit set of and . Then for every ,
(4) |
1.1. Outline of proof
The core strategy of the proof closely mirrors the approach developed in [KKLM]. The key idea is to construct a family of height functions that satisfy the contraction hypothesis. These height functions quantify the depth of orbits as they approach the cusp. Once such a family is constructed, it is then used iteratively to generate an open covering of . Specifically, by applying the contraction hypothesis iteratively, we construct coverings of using balls of nearly uniform diameter, with the number of balls being bounded by a function of their common diameter. This estimate ensures the finiteness of the -dimensional Hausdorff measure of for specific values of smaller than , thereby providing an upper bound on . This bound corresponds to the one stated in Theorem 1.4.
Despite its apparent simplicity, the approach presents significant challenges. The biggest one is constructing a family of height functions that satisfy the optimal contraction conditions. The optimality is important because the bound for depends very sensitively on the contraction rate of the height functions. Existing techniques can create height functions, but they aren’t optimal unless all the weights are equal. As a result, the best possible bound for , without the new tools in this paper, is , which only matches the actual bound when , meaning all the weights are equal (where ).
To address this, we introduce some new concepts. First, we redefine the covolume of sublattices to depend on weights, as done in Section 3. Using this new definition of covolume , we as usual define the following proper functions on :
In the unweighted case, the linear combination of these acts as a height function, and one of the essential ingredients in proving that the linear combination is indeed a height function is a result from [EMM], Lemma 5.6, which states that for any two sublattices of and with usual definition of covolume , we have
This inequality is used to link the average rate of expansion of vectors in certain representations to recurrence results on , which leads to the height function. The most crucial step of the paper is to prove an analogue of this result using the redefined covolume. Although this may seem straightforward, the new covolume definition introduces complications. For instance, it is not even true that for any , one can find a constant such that
The analogous key inequality is referred to as the “weighted Mother inequality” and is discussed in Section 3. Section 4 proves the optimal expansion rate of vectors in the exterior powers of the standard representation of with respect to the quasinorm based on weights. Section 5 constructs the height functions using results from Section 4 and the weighted Mother inequality.
The second challenge lies in deriving dimension bounds from the contraction hypothesis. This is done iteratively. In the equal weight case, for , the process is as follows: Start with and divide it into parts, i.e., . Applying the contraction hypothesis with the base point , at most intervals are needed to cover .
Now, pick one of the -balls that intersect , say , and divide it into smaller parts. By applying the contraction hypothesis again with base point , and using the fact that
and that we know the point is near the cusp (since ), we conclude that at most intervals are needed to cover . Thus, -balls of diameter are needed to cover . Repeating this process iteratively, we find that -balls of diameter are needed to cover , which gives the bound .
In the equal weight case, this process is simpler because the action of (as defined in (3) for ) expands the interval evenly. This allows us to easily subdivide it into cubes like and conjugate them with , making the integral over these cubes behave like that over . However, in the unequal expansion case, such simple divisions of are no longer possible. Instead, the space must be divided into hypercuboids with unequal side lengths to ensure that they become under conjugation. We must also ensure that these hypercuboids are disjoint, since we rely on the small measure of their union to conclude that the number of hypercuboids needed for covering is small. These two restrictions together are difficult to satisfy. This issue becomes even more challenging when dealing with fractals. Overcoming this obstacle is the second main contribution of this paper, discussed in Section 6.
2. Notation
The following notation will be used throughout the paper.
2.1. Homogeneous spaces
Fix and set and . The space can be naturally identified with the space of unimodular lattices in , via the identification .
Fix such that and . Define and set
(5) |
for any subset of . For , define
(6) |
As in the Introduction, we define for , the matrix
(7) |
We define to be the following subgroup of ,
(8) |
2.2. Iterated Function Systems
A contracting similarity is a map of the form where , and . A finite similarity Iterated Function System with constant ratio (IFS) on is a collection of contracting similarities indexed by a finite set , called the alphabet, such that there exists a constant independent of so that
for all .
Let . The coding map of an IFS is the map defined by the formula
(9) |
It is well known that the limit in exists and that the coding map is continuous. The image of under the coding map called the limit set of , is a compact subset of , which we denote by . We define for ,
(10) |
We will say that satisfies an open set condition (OSC for short) if there exist a non-empty open subset such that the following holds
Let denote the space of probability measures on . For each we can consider the measure under the coding map. A measure of the form is called a Bernoulli measure.
The following proposition is well known (see for eg [[Hutchinson], Thm. 5.3(1)] for a proof).
Proposition 2.1.
Suppose is an IFS satisfying the open set condition with limit set . Let denote the common contraction ratio of and . Then the Hausdorff dimension of is . Moreover the -dimensional Hausdorff measure satisfies . If denotes the normalised restriction of to , then is a Bernoulli measure and equals , where is the uniform measure on , i.e, for all . Moreover, there exists a constant such that for all , we have
(11) |
For the rest of the paper, we fix for , the IFS with common contraction ratio and the limit set . Let , , and . Let denote the normalised restriction of to and define the measure on . Let us define the constant as
(12) |
An observations that will be needed later in the proof is that
(13) |
for all .
3. Weighted Mother Inequality
Let us define
Define action of on via the map . Suppose denote the standard basis of . For each index set , we define
(14) |
The collection of monomials gives a basis of for each . For and each index set , we denote by , the unique value so that , where the sum is taken over all index sets .
For each , we define quasi-norm on each of as
(15) |
where maximum is taken oven all index sets of cardinality and is defined as in (5). Note that for all and for all (), we have
(16) |
Definition 3.1.
For a discrete subgroup of of rank , we define as , where is a -basis of . Note that the definition of is independent of the choice of basis . We define as
(17) |
where on is defined as in (15). We also define .
The main result of this section is the following inequality which originated in [EMM]. Part of this proof is motivated from [[BQ12], Proof of Prop. 3.1] where it is termed the ‘Mother inequality’.
Theorem 3.2.
There exist a constant and a finite set such that the following holds. Fix a lattice in . Then for any sublattices of , we have
(18) |
where denotes the smallest discrete subgroup of containing and .
Proof.
The proof is divided into two cases:
Case : . In this case, (18) follows if the following holds: For any and , the following holds:
(19) |
To prove (19), note that
(20) | ||||
(21) | ||||
(22) | ||||
(23) | ||||
(24) | ||||
(25) |
Hence (19) follows.
Case : . In this case, (18) follows if the following holds: For any , and , the following holds:
(26) |
To prove this, we define the linear map such that for any and ,
where the sum is taken over all subsets of size . Let us explain each term in the sum.
-
•
the element is the exterior product , when one writes with .
-
•
the element is the exterior product , when one writes with .
-
•
the sign is the signature of the permutation of sending to for .
For , we define quasi-norms on as
(27) |
where . Let
Then it is easy to see that for every and , we have such that
(28) |
Let and . It is clear that is a finite set, not containing . Also note that .
Note that for , the action of (defined as in (6)) on via the map satisfies
(29) |
Also for all , , we have
(30) |
We define
(31) |
which is a finite value since is continuous and the set is a compact set.
Fix , and . Note that (26) hold trivially if or . Hence, we may assume and let . Then, we have using (29) that
(32) |
Note that
(33) |
Hence, we have
using (28) | ||||
using (33) | ||||
using (29) | ||||
using (30) | ||||
using (31), (32) | ||||
using (32) | ||||
using (29) | ||||
Thus theorem now follows by taking any larger than all the constants appearing above. ∎
4. Critical Exponent
For , we define
(34) |
Also, for any compact subset (defined as in (8)), define the set . It is easy to see that is still a compact set. Define
(35) |
Lemma 4.1.
Fix . Then for all , with , we have
Recall the following well-known fact.
Lemma 4.2.
There exists such that the following holds for all and all
The following lemma is a simple application of Fubini’s Theorem.
Lemma 4.3.
Let be a Borel measure and a non-negative Borel function on a separable metric space . Then,
The following Proposition studies the average rate of expansion of vectors in the exterior powers of , and is the main result in this section.
Proposition 4.4.
Fix . For all , there exists such that the following holds for all ,
Proof.
Fix , and . Without loss of generality, we may assume that . Consider the vector space spanned by . Also consider the vector space spanned by We will denote by the canonical projection. Similarly define . Note that for all , we have
Then we have
Thus, it is enough to prove that the quantity is bounded above by a constant independent of .
To do this, we define , where is defined as in (35). Since is compactly supported, we have that .
Case 1 . Then, since , we have . This means that
Thus, .
Case 2 . In this case, there exists a subset of cardinality such that , where . Let us define for all , the set
Let us define
It is clear that . Claim that for all , we have
(36) |
where is large enough so that (11) holds for all (). To see this, first of all, note that by explicit computation, we have for every , the following holds
for some . Thus for all , we have
Thus if we fix for , then for all and , we have that the condition is equivalent to the condition that belongs to an interval of size atmost , which has measure less than (using (11)). Hence, by Fubini’s Theorem, we get that
(37) |
From (13), we know that . Coming this with the fact that , we get (36) follows from (37).
5. Height Functions
We will need some notation before proceeding. For , let denote the set of all primitive subgroups of the lattice , i.e, the subgroups of the lattice satisfying , where is the smallest vector subspace containing . For every , we define as
(38) |
where is defined as in (17). For , we define . It is easy to see that .
Also, let and be defined as in Theorem 3.2.
Lemma 5.1.
For every and , we have .
Proof.
Fix and . It is easy to see that there exists a primitive sublattice of rank of co-volume less than or equal to . This follows from a simple application of induction and Minkowski’s Convex Body Theorem. This means that the vector (defined as in Def. 3.1) satisfies , where the sum is taken over all index sets of cardinality . Hence we have , which gives . Thus, .
∎
Lemma 5.2.
For any compact subset , the following holds for every , ,
(39) |
for all .
Proof.
The following Proposition introduces the method of ‘integral inequalities’ into the proof and is the place where weighted Mother inequality is used.
Proposition 5.3.
For every , there exists a constant , such that for every and ., there exists , so that for all , we have
(40) | ||||
(41) |
Proof.
Lemma 5.4.
There exists constants such that
(43) |
for all and .
Proof.
Let . We will define by induction. Let and suppose that is already defined for satisfying and (43) for satisfying . Then define inductively as
Then, it is clear that . Also for , such that , we have
using and the definition of | |||
Thus (43) holds for satisfying . Hence, by induction, the sequence exists. ∎
Fix as in Lemma 5.4. For and , we define the function
(44) |
Proposition 5.5.
For every , there exists a constant , depending only on such that for every , there exists and so that for all , the following holds
(45) |
6. The Contraction Hypothesis
Definition 6.1 (The Contraction Hypothesis).
Suppose is a metric space equipped with an action of . Given a collection of functions for some unbounded set and real numbers , we say that satisfies the -contraction hypothesis on if the following properties hold:
-
(1)
The set is independent of and is -invariant.
-
(2)
For every , is uniformly log Lipschitz with respect to the (defined in (8)) action. That is for every bounded neighborhood of identity in , there exists a constant such that for every , and ,
(48) -
(3)
There exists a constant such that the following holds: for every , there exists such that for all with ,
(49)
The functions will be referred to as height functions.
Remark 6.2.
The above definition of Contraction Hypothesis is motivated from the corresponding definition in [Khalilsing].
Theorem 6.3.
Let be a metric space equipped with an action by . Assume that satisfies the -contraction hypothesis on for a real numbers . Then for all ,
Proof.
Without loss of generality, we may assume that . Now fix large enough so that . Let us define as
(50) |
Let be the constant so that (48) holds for all , and with .
Fix . Fix large enough so that for all . For the sake of simplicity, let . For every and , we define as the unique integer such that . We define for all , the element as
(51) |
Let us quickly recall some notation. For , is the limit set of the IFS , with common contraction ratio and . The dimension of equals . The set has dimension . The measure denote the normalised restriction of to and the measure on is defined as . For and , we define as in (10) corresponding to . We also define as in (12). We define as . Let us define . Also define as restriction of . Note that since , so belongs to for all .
Let and be so that (49) holds for and with . Let .
Then we have that
(52) |
where . Fix .
Let us make some easy observations before proceeding:
Observation 1 For all , we have and belongs to (defined as in (50)).
Explanation: This follows from the definition of along with fact that
Observation 2 For , if is such that , then for all .
Explanation: Suppose . Then as and , we have . Now if , then the -th co-ordinate is less than . If define as the vector whose -th entry is times the -th entry of , then . Thus we have
Observation 3 For any measurable function , measurable set and , we have
Explanation: Using the fact that each satisfies open set condition, we get that for any , the following holds
The observation now follows immediately.
Observation 4 For any measurable function , and , , we have
Explanation: Let us define the matrix as
(53) |
Using the definition of and , it is easy to see that . Since and is a Bernoulli measure using the Proposition 2.1 corresponding to uniform measure on , we get that equals pushforward of under the map . Thus, we have
Observation 5 For any measurable function , and , , we have
Explanation: Note that for all , we have
Observation 6 For and such that , we have
Explanation: Since , we get from Observation 2 and fact that that . Combining this with Observation 1, we get that . Now above observation follows by using (49).
Note that for any , we have using Observation 2 that
(54) |
Also, we have for
(55) |
Using (54) and iteratively using (55) for , we get that for any the following holds
(56) |
where . This
For and , we define as the unique integer satisfying . Let us try to cover by balls of size less than or equal to . We do this by selecting sets in which intersect . Note that there are many elements in , each of which has equal -measure. Thus, we get that
where . Thus for every , we can cover by -many hypercuboid of diameter less than or equal to . Thus,
which is finite for all satisfying , i.e,
This gives that
(57) |
Using fact that for any countable collection of borel sets and (52), (57), we get that
Since is independent of , letting (which exists as is unbounded), we get that the theorem holds. Hence proved. ∎
7. Final Proof
Proof of Theorem 1.4.
Let be given. Using Proposition 5.5, for every , choose and define the collection of height functions
Now it is easy to see that the action of on satisfies -contraction hypothesis with respect to measure . Indeed the first properties of Definition 6.1 follow immediately from the definition of . The second condition follows from Lemma 5.2 and the fact that for all , is a linear combination of . The third property follows from Proposition 5.5 and for and corresponding to each (Note that the value of also depends on ).
Thus by Theorem 6.3, we get that for all the following holds
Since , we get that
Since is arbitrary, the theorem follows. ∎
8. Concluding Remarks
More generally, one could ask for the dimension of weighted singular matrices. Our approach is applicable in this setting. It is also possible to deal with the product of more general fractals rather than the same contraction ratios. The main changes needed in this case involve the contraction hypothesis. These generalizations are ongoing work in progress.