Abstract
These are lecture notes for a mini-course given in Banff in June 2024. They discuss the problem of bounding the number of -incidences under various hypotheses on and . The main focus will be on hypotheses relevant for the Furstenberg set problem.
Introduction
These lectures are about approximate incidence geometry, but they are far from a complete introduction to this topic. A more accurate name for the contents would be introduction to upper bounds in linear approximate incidence geometry in the plane. For further reading, I mention a few important topics not covered in these notes, and some of the most recent references (as of summer 2025). I will not discuss the high-low method pioneered in [28] and applied, e.g. in [6, 8, 9, 23, 24, 53]. I will not discuss Falconer’s distance set problem, or other “curvilinear” incidence problems [20, 25, 26, 36, 55, 64]. I will not discuss incidence lower bound problems [8, 9, 14]. Finally, I will not discuss the Kakeya problem [32, 59–61].
A prototypical problem in approximate incidence geometry is the following. Let be a finite set, and let be a finite set of lines in the plane. Fix a (small) scale parameter , and consider the -incidences
Here and below, stands for the closed Euclidean -neighbourhood of .
Question 1
(Incidences) How large can be in terms of P and ?
The following formulation is equivalent, but often gives a useful, slightly different perspective to the problem:
Question 2
(Rich lines) Given a finite set of points , and a parameter , how large can be the set of lines in whose -neighbourhood contains at least r points from P?
The answers to Questions 1–2 heavily depend on various hypotheses we place on P and . If we make no hypotheses, the best possible answer is . This is attained if P is contained in a single -disc , and all the lines also intersect B. To avoid this situation, one typically asks—at least—that both P and are -separated; in the formulation of Question 2, one only counts -separated elements in . The -separation of is defined with respect to some natural metric on , the space of all (affine) lines in . A common choice (also used in these notes) is
whenever , is a 1-dimensional subspace, and . The notation refers to orthogonal projection to L, and is the operator norm.
If both P and are -separated, the sharp estimate for is
| 1.1 |
see [19] (or [11, Exercise 7.5] or [21, Theorem 1.5] for alternative arguments which yield (1.1) with a -loss). This bound is (e.g.) attained whenever the points form a -net inside a -disc for some , and a -net of lines intersecting B.
Despite its sharpness, (1.1) is not very useful for solving “interesting” problems. These problems tend to involve more hypotheses on P and , which make the sharp bound for more difficult to prove. We will return to this in later sections.
Bounds for -incidences under -separation
When , we abbreviate
and . The “exact incidence” makes the problem of estimating easier than the problem in Question 1, although still non-trivial. Under no additional hypotheses on P and , Szemerédi and Trotter [57] in 1983 proved that
| 2.1 |
Equivalently, if is finite, then
The Szemerédi-Trotter bound is quite sharp. Perhaps, the simplest example is where , and : then . This shows that the implicit constant in front of the |P| (similarly ) term cannot be lower than 1. This leaves open the sharpness of the term . The exponent 2/3 is sharp, indeed the simple Example 2.2 below shows that the implicit constant in front of the term cannot be lower than :
Example 2.2
Consider and
Note that every line is n-rich, because if and , then for all . Thus, . On the other hand, if (2.1) holds with constant , then
This can only hold for all if .
Remark 2.3
The sharpest currently known form of (2.1) is with , see [1]. It is also known [52, Remark 4.2] that this inequality cannot hold with constant (see also [3, Sect. 1.3]). In that example, , and consists of -rich lines for any .
Crossing number proof
We start with a proof of (2.1) using the crossing number inequality discovered by Székely [56]. This proof also yields some information about -incidences, provided that both the points and lines are -separated:
Theorem 2.4
Let , and let be a -separated set. Let , and let be a family of -separated lines such that for all . Then,
| 2.5 |
Equivalently, .
Here is the crossing number inequality (proved independently in [37] and [2]):
Lemma 2.6
(Crossing number inequality) Let be a simple graph with . Then, the crossing number of G has the lower bound
The crossing number is the minimal number of edge crossings in any planar drawing of G. In particular, if and only if G is a planar graph.
Proof of Theorem 2.4
We may assume that , since otherwise , and Theorem 2.4 is trivial. Fix , so . Since the set is -separated, and is far larger than the width of the tube , the points in lie on a -Lipschitz graph over . In particular, they have a natural ordering (say, given by their projections to ). We form a set of unordered edges , in a graph with vertex set , by placing an edge between any pair of consecutive points p, q in this ordering. Then, we define
The (Euclidean) length of any edge is . It may happen that [p, q] lies in multiple families , but then the associated lines all lie at distance from each other in ,1 Since was assumed -separated, it follows that , i.e. the families have bounded overlap. Therefore,
since . Now, if , we have , and therefore, . This bound corresponds to the second term in (2.5).
Assume then that . In this case, the crossing number inequality is applicable in the graph and yields
We claim that , which will complete the proof of (2.5). The idea is roughly: every crossing pair of edges determines (hopefully) a unique pair of lines such that and . Therefore, . The uniqueness is not entirely trivial, however. We know that the edges in lie on an -Lipschitz graph over , but this alone is not helpful (unless ): It is well possible that two -Lipschitz graphs intersect each other multiple times, see Fig. 1.
Fig. 1.

The danger of many crossings
What saves the day is the fact that if , then all the crossings between must occur in a disc of radius . Let us do this carefully. We write
It suffices to show that individually .
If , there are no crossing pairs with . So, consider . Let be pairs of crossing edges with for all . Then, every edge and meets the intersection , which is contained in a disc of radius . So, in fact for all . Thus,
But for by the -separation of P, so , as claimed.
Cell decomposition proof
In this section, we give a second proof for (2.1), based on cell decompositions. This method was pioneered by Szemerédi and Trotter, but a simpler implementation can be achieved by the use of polynomials (see Lemma 2.11). The cell decomposition proof of (2.1) could also be modified to yield the -separated version stated in Theorem 2.4. We leave this as a (reasonably challenging) exercise for the reader.
We repeat below exactly what we are planning to prove:
Theorem 2.7
Let be a finite set. Then,
| 2.8 |
Besides the idea of cell decompositions, a second ingredient is the following simple incidence estimate:
Proposition 2.9
For finite and ,
| 2.10 |
Proof
For , write . Using the definition of and Cauchy–Schwarz:
Splitting the sum into diagonal and off-diagonal parts (where and , respectively), we obtain . This yields (2.10).
Here is the lemma from [27] on polynomial cell decompositions, we will need:
Lemma 2.11
Let be a degree. If is a finite set, there exists a polynomial surface with of degree such that every component contains points of P.
Remark 2.12
The components will be the “cells” in the “cell decomposition” proof of Theorem 2.7. Szemerédi and Trotter did not have access to Lemma 2.11, so they crafted their “cells” using straight lines and line segments. In many interesting, non-trivial cases (for example when P is a product set to begin with) crafting the cells by hand (using vertical and horizontal line) is easy.
Remark 2.13
The cases of Theorem 2.7 are easy and can be obtained from the following 2-ends argument. Let be a set of lines such that for all . In particular, every contains pairs of distinct points . As varies, these pairs are distinct, so we obtain
Rearranging yields . For , this coincides with the bound from Theorem 2.7. A key idea underlying the cell decomposition proof of Theorem 2.7 is to decompose P into smaller pieces , where we “expect” (although this inequality does not literally appear in the proof).
We are then equipped to prove Theorem 2.7.
Proof
We may assume that , since the opposite case follows from Remark 2.13. Write . We may also assume that . Namely when , we may use the elementary bound (2.10) to deduce that
If the second term dominates, then , and if the first term dominates, then , and therefore .
Assume , and apply Lemma 2.11 with degree to obtain a non-trivial polynomial surface with such that for all connected components . These components will be denoted .
Let
Since Z can contain distinct lines,2 We infer that , and therefore , using the assumption . Thus , and it suffices to verify the bound (2.8) for in place of .
Fix . By assumption . In particular, contains segments of the form [x, y] with distinct and consecutive (i.e. there are no points of P between x and y). Let be the collection of all such segments. A segment is cellular if for a common component , otherwise I is non-cellular. Note that every line contains non-cellular segments. Indeed, intersects Z somewhere on each non-cellular segment. So, if the number of non-cellular segments were , then , and hence (contrary to the definition of ).
The cellular segments are denoted , and their number is
Every line contains segments, but only non-cellular segments. Therefore, contains cellular segments, denoted . When varies, the collections are clearly disjoint. Therefore,
Consequently , as claimed.
Bounds for -incidences under s-dimensional separation
In Theorem 2.4, we saw that the Szemerédi-Trotter bound for -incidences is valid under the hypothesis that both P and are -separated. For applications in “continuum” incidence problems in fractal geometry, this hypothesis on is not so natural, see Remark 3.3. To keep the discussion concrete, we now introduce one distinguished “continuum” incidence question, which will follow us for the rest of these notes.
Problem 1
(Furstenberg set problem) Let and . Let be an (s, t)-Furstenberg set. This means that there exists a t-dimensional family such that for all . What is the best lower bound for ?
The Furstenberg set problem (for ) was proposed by Wolff [62, 63] in the late 90 s. After plenty of partial progress [4, 13, 15, 22, 28–30, 34, 38, 48, 49], it was finally solved in 2023 by Ren and Wang [53]. The sharp lower bound for is
| 3.1 |
Here is a simple but illustrative example of Furstenberg sets:
Example 3.2
For , let , where C is an s-dimensional Cantor set. Then , and F is an (s, 2)-Furstenberg set: Every non-vertical line satisfies In particular, the term “” in (3.1) cannot be omitted.
Remark 3.3
The estimate (3.1) can be viewed as a continuum analogue of the Szemerédi-Trotter bound (2.1). We clarify this somewhat informally. Instead of aiming for (3.1), let us attempt to prove (the box dimension estimate) for all small enough. Here is the -covering number.
Pick a -separated subset with , and for each a -separated set with . Let be a maximal -separated set in . Then,
| 3.4 |
(We used here that every line in is -incident to points in , although these points may not be the ones in .) On the other hand, if the Szemerédi-Trotter bound was valid under the -separation hypotheses on and , we could estimate
Combining these estimates would lead to , and eventually .
First, this is too good to be true: the term “” in (3.1) is necessary, as we saw in Example 3.2. Second, the Szemerédi-Trotter bound is not valid under -separation alone; the sharp bound under this hypothesis was mentioned in (1.1). The argument above using (1.1) would produce the unsharp estimate .
One might try to use Theorem 2.4, namely the version of Szemerédi-Trotter valid under -separation. This approach can—at best—yield an unsharp estimate. The argument above also runs into a difficulty: after extracting suitable -separated sets and , it seems hard to maintain any lower bound on the -incidences .
Discretising incidence problems involving Hausdorff dimension
We have seen that sharp upper bounds on under -separation or -separation hypotheses on P and fail to produce sharp estimates in the (s, t)-Furstenberg set problem. This is not surprising: The hypotheses and involve Hausdorff dimension, and that information is lost when passing to (merely) -separated or -separated subsets. We need to consider a separation condition that keeps the “Hausdorff dimension information”. Such a condition was introduced by Katz and Tao [34] in 2000:
Definition 3.5
(-set) Let (X, d) be a metric space, and . A -separated set is called a -set if
If the value of the constant is irrelevant, a -set may be called a -set.
We also extend the definition of -sets to families of dyadic cubes. For , a family of dyadic cubes is a -set if
where .
Remark 3.6
The -sets of these notes are sometimes called Katz-Tao -sets in the literature, while “-set” refers to sets satisfying .
Example 3.7
The property of being a -set gets weaker as s increases. In fact, every finite -separated set in is a -set.
Example 3.8
Every -separated set is a -set, because for . On the other hand, for , any r-disc can contain at most discs of radius . In particular for .
Jump ahead to Fig. 2 for a rather “opposite” example of a -set.
Fig. 2.

The train tracks example
We now record two standard tools for discretising continuum incidence problems involving Hausdorff dimension. Proposition 3.9 allows to find large -subsets, while Proposition 3.12 allows to find covers by -sets. We first recall that, for , the s-dimensional Hausdorff content of is defined by
where the runs over all countable set families such that .
Proposition 3.9
(Large -subsets) Let be a bounded set with . Then, for every , there exists a -set , and .
The proof is taken from [18, Proposition A.1]. The idea is the same as in the proof of Frostman’s lemma (see, e.g. [40, Theorem 8.8]).
Proof
Without loss of generality, assume that for some (if is arbitrary, first prove the proposition for the largest element in ). We may also assume that K is contained in some dyadic cube with .
Let be a finite set containing exactly one point from for all . Then, as desired, but may fail the -set condition. We modify as follows. Fix . If
define to be a maximal set with . Then automatically
Repeat this for all cubes to obtain . Then, repeat the procedure at all dyadic scales between and : if has already been defined, and there is a cube such that
choose a subset satisfying
| 3.10 |
The process stops in finitely many steps, because .
We claim that for every (and not only !) there exists a dyadic cube such that and
| 3.11 |
If , then (3.11) holds for . If , the point x was deleted from at some stage. Let be the largest cube containing x, where point deletion occurred. If this happened while defining , then (3.10) holds with . But since was the largest cube containing x where point deletion occurred, . This gives (3.11).
Let be the maximal elements in . Then, is a disjoint cover of K, because every element of is contained in some element of , as we just showed. This combined with (3.11) yields
We claim that P is a -set. For with , (3.10) yields . This implies for all and .
Proposition 3.12
(Covering by -sets) Let , and let be a compact set with . Then, for every , there exists a sequence of -sets such that , and
A slightly different version of this proposition was first proved by Katz and Tao in [34], but for more recent incarnations, see [23, 45].
Proof of Proposition 3.12
Let be a compact set with . Consider the following dyadic variant of the Hausdorff content:
where the “” only runs over families of dyadic cubes with side-length . It will be crucial that the “” exists thanks to the restriction to “large” cubes.
Fix . Using the compactness of K, and , one may find such that .3 This means that there exists a (finite) cover of K by dyadic cubes such that for all , and
| 3.13 |
It turns out that the sub-families , , are automatically -sets, that is,
If this inequality failed for some , a new competitor for is obtained by
Moreover, the analogue of the sum (3.13) for is strictly smaller than the one for , because
This violates the minimality of the sum (3.13).
To summarise, we have now found a cover of K by -sets, and it follows from (3.13) that for all (non-empty) families .
Incidences between -sets of points and lines
We return to the problem of estimating . After the introduction of -sets, a natural question is the following:
Problem 2
Let . Assume that is a -set, and is a -set. How large can be in terms of
and ?
|P| and ?
We use here the notation “” and “” to avoid confusion with the parameters “s” and “t” appearing in the Furstenberg set problem.
A sharp answer to part (a) was found by Fu and Ren in [22], for all values of . They provide an explicit continuous function such that
| 3.14 |
whenever are as in Problem 2. In particular for . They also give examples showing that is the minimal function satisfying (3.14). The full formula for the (piecewise linear) function f is a bit complicated, see [22, Theorem 1.4]. The method of [22] also gives an answer to part (b), but I’m not sure if that is sharp for all cardinalities |P| and . We will prove the bound in Theorem 3.22. For now, let us compare Problem 2 with the Furstenberg set problem.
First some bad news: Even a sharp answer to Problem 2 does not solve all the cases of the (s, t)-Furstenberg set problem. The reason is, roughly, that the hypothesis available for every (s, t)-Furstenberg set is very strong, and the extremal pairs for Problem 2.7 do not satisfy (a -discretised version) of this lower bound, e.g. in the form that , , would contain large -sets for some .
Then, some good news: A sharp answer to Problem 2.7 still solves the cases of the Furstenberg set problem. In fact, we have the following proposition:
Proposition 3.15
Assume that is a continuous function with the following property. If is a -set and is a -set, then
for all small enough. Then, every compact (s, t)-Furstenberg set satisfies
| 3.16 |
Remark 3.17
The compactness hypothesis in Proposition 3.15 could be relaxed to boundedness, but verify this, we would need a version of Proposition 3.12 for bounded sets. This is true, see [23, Lemma 2].
By (3.14), the explicit Fu-Ren function f satisfies the hypothesis (3.16) up to the harmless “”, so one can derive lower bounds for from the inequality (3.16).
The bound for , proven in Theorem 3.22, leads to the lower bound for (s, t)-Furstenberg sets with with . Here’s why. Assume that (without loss of generality), and assume that there exists an (s, t)-Furstenberg set with . Then , so
and this can be rearranged to . This is the sharp bound (predicted by (3.1)) in the range . The reader is encouraged to check that a similar application of Proposition 3.15 in the range would only yield unsharp results.
Proof of Proposition 3.15
By definition, there exists a line set with such that for all . It is easy to reduce matters to the case where our hypotheses are slightly strengthened as follows:
.
for all .
Fix . According to the Katz-Tao covering lemma, Proposition 3.12, we may find a sequence of families of dyadic cubes such that is a -set, and
In particular, for fixed, the intersection is covered by the sets for every . By our assumption (b), and the sub-additivity of Hausdorff content, there exists an index such that .
For , let . By assumption (a), there exists such that . We now set for this index , and we abbreviate . Then, is a -set. Define by selecting one point in each square in .
Since , by the subset finding lemma, Proposition 3.9, there exists a -set of cardinality . Here, the “” notation hides factors of order . Moreover, since for all , we know that
| 3.18 |
This implies that for , and consequently,
| 3.19 |
On the other hand, since P is a -set and is a -set, we have by the definition of the function f:
These inequalities are only compatible (for all small ) if . Since was arbitrary, we infer by the continuity of f.
The train tracks example
We will shortly prove the bound in Problem 2, in particular . Let us first see why this is sharp.
Example 3.20
Consider the points P and lines drawn in the train tracks example in Fig. 2. There are points organised in “columns”. Both the separation and height of the columns are . There are also lines, with lines incident to each column. In fact, every point is contained in for lines . Therefore,
Both P and are -sets. Therefore, under the hypothesis that both P and are -sets, the sharp upper bound for cannot beat . This is weaker than the Szemerédi-Trotter bound in Theorem 2.4, which would predict .
Remark 3.21
Recall from Remark 3.8 that -separated sets are -sets. The train tracks example demonstrates that the -separation of P and in Theorem 2.4 cannot be relaxed to the hypothesis that both P and are -sets.
One can next ask (at least) the following two questions:
-
(i)
Would it suffice to assume that only one of the sets P or is -separated—say P for concreteness—while is merely -separated, or perhaps a -set?
-
(ii)
Are the train tracks the only sharpness example if both P, are -sets?
The answer to (i) is negative, as far as I know, but turns positive if one additionally assumes that |P| is “maximal”, that is . Guth, Solomon, and Wang [28, Theorem 1.1] proved that if is -separated with , and is -separated, then
In fact, this is only a special case of their result. The article [28] pioneered the powerful high-low method for studying -incidences. We will not cover the high-low method in these notes, but the only reason is the lack of space and time. The high-low method is a central tool in -discretised incidence geometry and was also used in the resolution of the (s, t)-Furstenberg set conjecture.
Regarding question (ii), the answer is roughly positive, except that the individual “tracks” (the columns of red balls in Fig. 2) may be rotated and translated arbitrarily, as long as the -set conditions are preserved. The precise statement is that if is a pair of -sets with , then contains “cliques” such that . This is a recent result of the author and Yi [51]. In Fig. 2, the “cliques” are the pairs corresponding to each “track”.
Cases of the Furstenberg set problem
In this section, we prove a special case of Fu and Ren’s theorem [22, Theorem 1.4], which yields (in the notation of (3.14)), for . As discussed in Remark 3.17, this solves the (s, t)-Furstenberg set problem for .
Theorem 3.22
Let , and . Let be a -set, and let be a -set. Then,
| 3.23 |
Remark 3.24
The bound (3.23) with an additional -factor follows from [22, Theorem 1.5] and was originally proved using the high-low method. A finite field precedent is due to Vinh [58]. The -free proof below is from [46], and it is based on classical Sobolev smoothing properties of the X-ray transform. Let us note that the idea of using X-ray transforms and related operators to study incidences has been applied several times prior to [46], for example in [17, 33].
We then gather some preliminaries for the proof of Theorem 3.22. The arguments are a little sketchy, for full details see [46].
Definition 3.25
(X-ray transform) For , we define the X-ray transform by the formula
The X-ray transform is -smoothing of order , see [42, Theorem 5.3]. The following (special case) is [46, Theorem 2.16]:
Proposition 3.26
For every , there exists a constant such that
The norm on the right hand side is the standard homogeneous Sobolev norm
On the left hand side, the Sobolev norm is defined for functions in : This can be accomplished rigorously by identifying with , and then defining the Sobolev norm on instead of . We do not need the details here, except for the inequality
| 3.27 |
The proof of this inequality can be summarised as Plancherel + Cauchy–Schwarz.
Sobolev norms of negative order can be expressed in terms of the Riesz energy as follows: If is a Radon measure on , then
| 3.28 |
For a proof, see [40, Lemma 12.12]. Regarding Riesz energies, we need the following elementary computation concerning measures supported on -sets:
Lemma 3.29
Let be a -set with and . Consider the measure , and let be a standard mollification of . Then,
Proof sketch
The mollification causes small technicalities omitted here. Essentially the proof boils down to the computation
using .
We are then ready to prove Theorem 3.22.
Proof of Theorem 3.22
Recall that . Choose and such that . Note that either or , and we assume with no loss of generality that .
Let be a -set, and let be a -set. We associate the following measures and to P and , respectively:
One may now check that
Here is a mollification of (to make sense of this, identify with ). The inner integral is closely related to the X-ray transform of evaluated at : in fact
where is a mollification of . Therefore,
| 3.30 |
Next, we use the “duality” inequality (3.27) with , and eventually the boundedness of X between homogeneous Sobolev spaces (Proposition 3.26):
Since and , we may use Lemma 3.29 to deduce
and similarly . Plugging these estimates back into (3.30),
recalling that .
Cases of the Furstenberg set problem
Every (s, t)-Furstenberg set with satisfies . This follows from [38] or [31, Theorem A.1] (the special case was contained in [30]). In contrast with the cases , the sharp result in any of the cases cannot be derived by combining Proposition 3.15 with Fu and Ren’s (sharp) bound for for -sets P and -sets .
Exercise 1
What is the best lower bound you can obtain on the Hausdorff dimension of (s, t)-Furstenberg sets via Proposition 3.15, when ?
Remark 3.31
It may appear puzzling that Fu and Ren’s sharp upper bound on -incidences between does not always yield a sharp lower bound on the dimension of (s, t)-Furstenberg sets. To illustrate the reason, fix , , and consider the following two questions concerning a -set and a -set or lines :
Is it possible that for all ?
Is it possible that contains a -set with cardinality for all ?
The Fu-Ren bound always gives the correct answer to (Q1), but fails to see the difference between (Q1) and (Q2)—because the non-concentration conditions on , and the number of -incidences are the same in both questions. So, whenever is a triple relevant for the (s, t)-Furstenberg set problem such that (Q1)-(Q2) have opposite answers, the Fu-Ren bound fails to yield a sharp estimate on Furstenberg sets.
For example, take and . Now the answer to (Q1) is positive thanks to the “train tracks” Example 3.20. The answer to (Q2) is negative: A positive answer would essentially show that P is a -Furstenberg set with “dimension” 1. However, it has been known since the early 2000 s [4, 34] that this is not possible. Indeed, now we know that -Furstenberg sets have dimension by (3.1).
Remark 3.31 gives hint for how to proceed dealing with the cases : We need variants of incidence bounds which take into account the extra information present in (Q2). Consider the following generalised notion of -incidences:
Definition 3.32
Let , and let be an arbitrary collection of subsets of . We define .
Our familiar is a special case of this definition. We will apply the definition in a case where each is a -subset of a line. In this context have the following incidence bound, which can be viewed as a -incidence counterpart of Proposition 2.9:
Proposition 3.33
Let be -separated. Let . Let be a family of sets, where is a -set, and each is individually a -set. Then,
Remark 3.34
This proposition is “folklore” in the sense that the proof strategy has appeared in many places; I’m not sure where first, but it is certainly implicit, e.g. in the proof of [31, Theorem A.1]. A more explicit appearance is [48, Proposition 2.13].
Proof of Proposition 3.33
Write
The value of the “diagonal” sum is precisely . If the “diagonal” sum dominates, we obtain after rearranging .
For the “off-diagonal” sum with , observe that is contained in a ball of radius . Using the -separation of P, and the -set property of the sets , we deduce
Therefore, using the -set property of (which is implied by the -set property since ),
Since , this completes the proof.
Corollary 3.35
Every (s, t)-Furstenberg set with satisfies .
Remark 3.36
Corollary 3.35 is evidently sharp: for example any product set with and is an (s, t)-Furstenberg set, and it is often (if not always) the case that .
Proof of Corollary 3.35
The proof is similar to the proof of Proposition 3.15, we only point out the differences. We make a counter assumption: for some .
Extract the -set and the -set exactly the same way as in the proof of Proposition 3.15. Recall from (3.18) that
In the proof of Proposition 3.15, this information was used in the (wasteful) way to deduce that . We now do something more sophisticated: applying the subset finding lemma, Proposition 3.9, we locate, for each , a -set such that . Writing , we then have
Compare this lower bound with (3.19)!
The family satisfies the hypotheses of Proposition 3.33, so
Recalling that with , and , we arrive at a contradiction. It is worth noting that the -property of P was not used here in any other ways than to ensure that P is -separated, and (this actually means that the Katz–Tao covering lemma, Proposition 3.12, is not really needed in the proof).
Formal -discretisation of the Furstenberg set problem
We discussed above (Remark 3.31) that the full solution of the (s, t)-Furstenberg set problem cannot be obtained from Problem 2 (incidences between -sets of points and -sets of lines). There nonetheless exists a -discretised incidence problem which implies lower bounds for the (s, t)-Furstenberg set problem. This is formalised by the next proposition:
Proposition 3.37
Fix , , and . Consider the following statement (P):
-
(P)For every , there exists such that the following holds. Let be a -set with . For every , let be a -set with . Then, the union satisfies
If (P) holds for some , then every (s, t)-Furstenberg set satisfies .
In fact, Ren and Wang [53] showed precisely that (P) in the previous proposition holds with the choice as in (3.1).
Proposition 3.37 is proven by a pigeonholing argument similar to the ones we have seen in Proposition 3.15 and Corollary 3.35, see [31, Lemma 3.3] for the details.
Cases (s, 1) under maximal separation
We have previously covered the cases of the (s, t)-Furstenberg set problem. The remaining cases are more complicated, and beyond the scope of these notes.
In this section, we establish property (P) in Proposition 3.15 with (the sharp value) for all pairs (s, 1) with , but under the additional hypothesis that the -sets are maximally separated. This is a special case of [21, Theorem 4], but the proof we present here is different, and avoids using the crossing number lemma.
For technical convenience, we work under the hypothesis that contains a -net in . A modification of the argument would, however, work under the hypotheses that and is -separated.
As a lemma we need the following fact about (generalised) Kakeya sets:
Lemma 3.38
Let , and let be a -set such that for all . Then, the set satisfies .
Proof
For every , let be a maximal -separated set (in particular is a -set). Let be a maximal -separated set. Write . Then, . Combining this with Proposition 3.33 with , we obtain
which can be rearranged to . Thus .
Remark 3.39
In the previous proof, we could have also used Theorem 3.22.
Theorem 3.40
Let , and let be such that for all . Set . Let be a -dense set for all . Then, the set satisfies
In particular, if is a -set with , then (by Lemma 3.38).
Proof
Write , so . For fixed, we say that two tubes with are Q-distinct if
see Fig. 3 for intuition. Let be the maximal cardinality of a Q-distinct subset of .
Fig. 3.

Here , even though 9 tubes in total intersect Q
Claim 3.41
We have
Proof
Let be a maximal Q-distinct collection. Then, every tube satisfies for some . In particular, is contained in the union of the sets with . The Lebesgue measure of each of these sets is , so , as claimed.
The proof of the next claim is a variant of the 2-ends argument in Remark 2.13.
Claim 3.42
We have
Proof
Let be a maximal -separated set. Recall that the points in form a -net in . Therefore, if , and , the set contains two distinct points with separation , see Fig. 3.
Therefore, we obtain a map which associates to every tube with a pair with separation . It now remains to observe that this map is C-to-1 on every collection of Q-distinct tubes. Therefore , as claimed.
Finally, observe that for , so
This completes the proof.
Applications and connections of Furstenberg sets
In 2000, Wolff [62, 63] had recently posed the (s, 1)-Furstenberg set problem, and it was, e.g. known that every -Furstenberg set has . The conjecture (now a theorem) states that . In their influential 2001 paper, Katz and Tao [34] showed that an -improvement to the bound is logically equivalent (at a -discretised level) to an -improvement in a version of Falconer’s distance set conjecture, and also an -improvement in the -discretised version of the Erdős-Szemerédi sum-product problem. All of these equivalent -improvements were shortly afterwards obtained by Bourgain [4].
During the past 20 years, more connections have been discovered, for example to orthogonal and radial projections. The connection to orthogonal projections is straightforward: Non-trivial results on the -discretised Furstenberg set problem formulated in Proposition 3.37 imply non-trivial bounds on the dimension of exceptional sets of orthogonal projections. The mechanism is explained in [48, Sect. 3.2]. Notably, the full solution, due to Ren and Wang [53], implies the following sharp bound stated in [53, Theorem 1.2]:
Theorem 4.1
Let be analytic. Then, writing for and ,
A weaker estimate was earlier obtained by Kaufman [35].
The following sections contain brief accounts on the implications of Furstenberg set estimates to the (continuous) sum-product problem, to radial projections, and to arithmetic sums of fractal sets on the parabola.
The sum-product problem
The Erdős-Szemerédi sum-product conjecture [7] asks to prove that if (original formulation) or (plausible extension) is a finite set, then for all . This problem remains open and very actively studied. Elekes in the late 90 s [16] connected the problem to the Szemerédi-Trotter incidence bound and allowed him to establish the partial result . This is nowadays far below the state-of-the-art in the discrete version of Erdős and Szemerédi’s problem (see [54]), but his argument has the benefit of extending easily to a “continuum” setting. The main idea is the following observation:
Lemma 4.2
Let be sets. Let . Then, the family of lines
has the property that contains the affine copy of A given by .
Proof
Note that for all , since . Therefore for all .
To use Lemma 4.2, one checks that the line set is diffeomorphic to , and in particular . Therefore, Lemma 4.2 tells us that is an (s, t)-Furstenberg set with
One can now deduce various inequalities about using the Furstenberg set theorem. We only consider the case . Write , and note that . Thus, is an -Furstenberg set, and
Since (where is the packing dimension), it follows that either
Here “” is Elekes’ exponent. This argument also works well at a -discretised level, but we leave the details to the reader (or see [49, Corollary 6.6]). For further recent work on the continuum and -discretised sum-product problems, see [39, 43].
Radial projections and the dimension of quotients
In 2020, the best general lower bounds for the dimension of (s, t)-Furstenberg sets were the following:
| 4.3 |
The 2s-bound is a special case of Corollary 3.35, whereas the -bound is due to Héra [29] (see also [41] for a partial result). In 2021, Shmerkin and myself [48] proved the following “-improvement” over the 2s-bound:
Theorem 4.4
For and , the Hausdorff dimension of every (s, t)-Furstenberg set satisfies for some .
Theorem 4.4 was applied in [50] to prove a new radial projection theorem, which, in turn, played a role in the solution of the full Furstenberg set problem (see [48, 53]). I will now explain (very informally) the connection between Theorem 4.4 and radial projections.
For , the radial projection to x is the map defined by . A good way to think about radial projections is the following: if and , then , where
In [50], we proved the following:
Theorem 4.5
Let be disjoint Borel sets such that E is not contained on any line. Then, .
The following argument is a very rough indication of how the -improvement in the Furstenberg set problem, Theorem 4.4, appears in the proof of Theorem 4.5.
“Proof” of Theorem 4.5
Write . If necessary replacing F by an s-dimensional subset, we may assume that . To reach a contradiction, assume that for all , where . In other words for all .
The set is a “dual” -Furstenberg set: It contains an -dimensional line family incident to every point in the s-dimensional set E. Since , Theorem 4.4 implies
for some . Or does it? A potential problem is that in Theorem 4.4, so if (i.e. for ), we are in trouble. However, we may assume here that . This part of the argument uses the hypothesis that E is not contained on a line, and is based on a much earlier result [44].
Now we pose a rather unrealistic simplifying assumption: since , but , it is somewhat reasonable to expect that for . If this were the case, as we now assume, then F is an -Furstenberg set with . Consequently, by the second “classical” bound in (4.3),
This contradiction completes the “proof” of Theorem 4.5.
As a corollary, we obtain the following sum-quotient estimate:
Corollary 4.6
Let be Borel sets. Then
Proof
We may assume that both A, B contain at least two points. We apply Theorem 4.5 to the sets and . Since E is a Borel set not contained on a line, for every , there exists a point such that
Now, it remains to observe that agrees with the dimension of “slopes” spanned between the point and the set , namely
Since the quotient set contains all such slopes, the corollary follows.
Arithmetic sums of subsets of the parabola
The Furstenberg set problem has some hidden “curvature”, which can be brought to light by the following observation. The map sends every non-vertical line to some translate of the standard up-ward pointing parabola . More precisely, if is the line with slope passing through , then
The map is locally bi-Lipschitz, so s-dimensional subsets of lines are carried to s-dimensional subsets of the corresponding parabolas (for an alternative argument, observe that leaves the first coordinate fixed). The map also sends every t-dimensional set of lines to a family of parabolas of the form , where .
Using the map , and its properties described above, one obtains (with the argument of [47, Sect. 1.1]) the following corollary of the Furstenberg set theorem:
Corollary 4.7
Let be a set with , and let be a set with . Then,
| 4.8 |
In particular, .
Proof
The set is an (s, t)-Furstenberg set, so (4.8) is implied by the Furstenberg set theorem. The bound follows by applying (4.8) to and F (and noting first that ).
Remark 4.9
The bound is a continuum analogue of an observation of Bourgain and Demeter [5, Proposition 2.15]. They showed that if is a finite set, then . They also ask in [5, Question 2.13] whether , and deduce a positive answer from the -decoupling theorem if is -separated. Analogously, it seems unlikely that the bound for in Corollary 4.7 is sharp. For further literature on this topic, see [10, 12, 47].
Acknowledgements
I would like to thank the reviewers for making numerous helpful suggestions to improve the presentation.
Funding
Open Access funding provided by University of Jyväskylä (JYU).
Data availability
No datasets were generated or analysed during this study.
Declarations
Conflict of interest
The author is supported by the Research Council of Finland via the project Approximate incidence geometry, Grant No. 355453, and by the European Research Council (ERC) under the European Union’s Horizon Europe research and innovation programme (Grant Agreement No. 101087499). There are no other conflicts of interest.
Footnotes
If , then the segment [p, q] is contained in , which forces .
If Z contained more lines than its degree, then almost every lines of would be contained in Z. This would lead to , and therefore contrary to Lemma 2.11.
First find a sequence of arbitrary dyadic cubes covering K with . Enlarge the cubes slightly to make them open, but still retaining . Finally, pick a finite sub-cover, and let be the smallest side-length in that sub-cover.
T.O. is supported by the Research Council of Finland via the project Approximate incidence geometry, Grant No. 355453, and by the European Research Council (ERC) under the European Union’s Horizon Europe research and innovation programme (Grant Agreement No. 101087499).
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
References
- 1.Ackerman, E.: On topological graphs with at most four crossings per edge. Comput. Geom. 85, 101574 (2019) [Google Scholar]
- 2.Ajtai, M., Chvátal, V., Newborn, M.M., Szemerédi, E..: Crossing-free subgraphs. In: Theory and Practice of Combinatorics, Volume 60 of North-Holland Mathematics Studies, pp. 9–12. North-Holland, Amsterdam (1982)
- 3.Balko, M., Cibulka, J., Valtr, P.: Covering lattice points by subspaces and counting point-hyperplane incidences. Discrete Comput. Geom. 61(2), 325–354 (2019) [Google Scholar]
- 4.Bourgain, J.: On the Erdős–Volkmann and Katz–Tao ring conjectures. Geom. Funct. Anal. 13(2), 334–365 (2003) [Google Scholar]
- 5.Bourgain, J., Demeter, C.: The proof of the decoupling conjecture. Ann. Math. (2) 182(1), 351–389 (2015) [Google Scholar]
- 6.Bright, P., Gan, S.: Exceptional set estimates for radial projections in . Ann. Fenn. Math. 49(2), 631–661 (2024) [Google Scholar]
- 7.Choi, S.L.G., Erdős, P., Szemerédi, E.: Some additive and multiplicative problems in number theory. Acta Arith 27, 37–50 (1975) [Google Scholar]
- 8.Cohen, A., Pohoata, C., Zakharov, D.: A new upper bound for the Heilbronn triangle problem. arXiv e-prints, arXiv:2305.18253 (May 2023)
- 9.Cohen, A., Pohoata, C., Zakharov, D.: Lower bounds for incidences. Invent. Math. 240(3), 1045–1118 (2025) [Google Scholar]
- 10.Dasu, S., Demeter, C.: Fourier decay for curved Frostman measures. Proc. Am. Math. Soc. 152(1), 267–280 (2024) [Google Scholar]
- 11.Demeter, C.: Fourier Restriction, Decoupling, and Applications. Cambridge Studies in Advanced Mathematics, vol. 184. Cambridge University Press, Cambridge (2020) [Google Scholar]
- 12.Demeter, C., Wang, H.: Szemerédi-Trotter bounds for tubes and applications. Ars Inveniendi Analytica, Paper No. 1, p. 46 (2025)
- 13.Di Benedetto, D., Zahl, J.: New estimates on the size of -Furstenberg sets. arXiv e-prints, arXiv:2112.08249 (December 2021)
- 14.Dąbrowski, D., Goering, M., Orponen, T.: On the dimension of -Nikodým sets. arXiv e-prints, arXiv:2407.00306 (June 2024)
- 15.Dąbrowski, D., Orponen, T., Villa, M.: Integrability of orthogonal projections, and applications to Furstenberg sets. Adv. Math. 407, 108567 (2022) [Google Scholar]
- 16.Elekes, G.: On the number of sums and products. Acta Arith 81(4), 365–367 (1997) [Google Scholar]
- 17.Eswarathasan, S., Iosevich, A., Taylor, K.: Fourier integral operators, fractal sets, and the regular value theorem. Adv. Math. 228(4), 2385–2402 (2011) [Google Scholar]
- 18.Fässler, K., Orponen, T.: On restricted families of projections in . Proc. Lond. Math. Soc. (3) 109(2), 353–381 (2014) [Google Scholar]
- 19.Fässler, K., Orponen, T., Pinamonti, A.: Planar incidences and geometric inequalities in the Heisenberg group. arXiv e-prints, arXiv:2003.05862 (March 2020)
- 20.Fraser, J.M.: Distance sets, orthogonal projections and passing to weak tangents. Israel J. Math. 226(2), 851–875 (2018) [Google Scholar]
- 21.Fu, Y., Gan, S., Ren, K.: An incidence estimate and a Furstenberg type estimate for tubes in . J. Fourier Anal. Appl. 28(4), Paper No. 59, 28 (2022) [Google Scholar]
- 22.Yuqiu, F., Ren, K.: Incidence estimates for -dimensional tubes and -dimensional balls in . J. Fractal Geom. 11(1–2), 1–30 (2024) [Google Scholar]
- 23.Gan, S., Guo, S., Guth, L., Harris, T.L.J., Maldague, D., Wang, H.: On restricted projections to planes in . Am. J. Math. (2025) (to appear)
- 24.Gan, S., Guth, L., Maldague, D.: An exceptional set estimate for restricted projections to lines in . J. Geom. Anal. 34(1), Paper No. 15, 13 (2024) [Google Scholar]
- 25.Green, J., Harris, T.L.J., Ou, Y., Ren, K., Tammen, S.: Incidence bounds related to circular Furstenberg sets. arXiv e-prints, arXiv:2502.10686 (February 2025)
- 26.Guth, L., Iosevich, A., Yumeng, O., Wang, H.: On Falconer’s distance set problem in the plane. Invent. Math. 219, 779–830 (2020) [Google Scholar]
- 27.Guth, L., Katz, N.H.: On the Erdős distinct distances problem in the plane. Ann. Math. (2) 181(1), 155–190 (2015) [Google Scholar]
- 28.Guth, L., Solomon, N., Wang, H.: Incidence estimates for well spaced tubes. Geom. Funct. Anal. 29(6), 1844–1863 (2019) [Google Scholar]
- 29.Héra, K.: Hausdorff dimension of Furstenberg-type sets associated to families of affine subspaces. Ann. Acad. Sci. Fenn. Math. 44(2), 903–923 (2019) [Google Scholar]
- 30.Héra, K., Keleti, T., Máthé, A.: Hausdorff dimension of unions of affine subspaces and of Furstenberg-type sets. J. Fractal Geom. 6(3), 263–284 (2019) [Google Scholar]
- 31.Héra, K., Shmerkin, P., Yavicoli, A.: An improved bound for the dimension of -Furstenberg sets. Rev. Mat. Iberoam. 38(1), 295–322 (2022) [Google Scholar]
- 32.Hickman, J., Rogers, K.M., Zhang, R.: Improved bounds for the Kakeya maximal conjecture in higher dimensions. Am. J. Math. 144(6), 1511–1560 (2022) [Google Scholar]
- 33.Iosevich, A., Jorati, H., Łaba, I.: Geometric incidence theorems via Fourier analysis. Trans. Am. Math. Soc. 361(12), 6595–6611 (2009) [Google Scholar]
- 34.Katz, N.H., Tao, T.: Some connections between Falconer’s distance set conjecture and sets of Furstenburg type. N. Y. J. Math. 7, 149–187 (2001) [Google Scholar]
- 35.Kaufman, R.: On Hausdorff dimension of projections. Mathematika 15, 153–155 (1968) [Google Scholar]
- 36.Keleti, T., Shmerkin, P.: New bounds on the dimensions of planar distance sets. Geom. Funct. Anal. 29(6), 1886–1948 (2019) [Google Scholar]
- 37.Leighton, T.: Complexity Issues in VLSI. MIT Press, Cambridge, MA (1983) [Google Scholar]
- 38.Lutz, N., Stull, D.M.: Bounding the dimension of points on a line. In: Theory and Applications of Models of Computation, volume 10185 of Lecture Notes in Computer Science, pp. 425–439. Springer, Cham (2017)
- 39.Máthé, A., O’Regan, W.L.: Discretised sum-product theorems by Shannon-type inequalities. arXiv e-prints, arXiv:2306.02943 (June 2023)
- 40.Mattila, P.: Geometry of Sets and Measures in Euclidean Spaces. Fractals and Rectifiability. 1st Paperback Edition. Cambridge University Press, Cambridge (1999) [Google Scholar]
- 41.Molter, U., Rela, E.: Furstenberg sets for a fractal set of directions. Proc. Am. Math. Soc. 140(8), 2753–2765 (2012) [Google Scholar]
- 42.Natterer, F.: The Mathematics of Computerized Tomography. B. G. Teubner, Stuttgart. Wiley, Chichester (1986) [Google Scholar]
- 43.O’Regan, W.: Sum-product phenomena for Ahlfors-regular sets. arXiv e-prints, arXiv:2501.02131 (January 2025)
- 44.Orponen, T.: On the dimension and smoothness of radial projections. Anal. PDE 12(5), 1273–1294 (2019) [Google Scholar]
- 45.Orponen, T.: Combinatorial proofs of two theorems of Lutz and Stull. Math. Proc. Camb. Philos. Soc. 171(3), 503–514 (2021) [Google Scholar]
- 46.Orponen, T.: On the Hausdorff dimension of radial slices. Ann. Fenn. Math. 49(1), 183–209 (2024) [Google Scholar]
- 47.Orponen, T., Puliatti, C., Pyörälä, A.: On Fourier transforms of fractal measures on the parabola. Trans. Am. Math. Soc. (2025) (to appear)
- 48.Orponen, T., Shmerkin, P.: On the Hausdorff dimension of Furstenberg sets and orthogonal projections in the plane. Duke Math. J. 172(18), 3559–3632 (2023) [Google Scholar]
- 49.Orponen, T., Shmerkin, P.: Projections, Furstenberg sets, and the sum-product problem. arXiv e-prints, arXiv:2301.10199 (January 2023)
- 50.Orponen, T., Shmerkin, P., Wang, H.: Kaufman and Falconer estimates for radial projections and a continuum version of Beck’s theorem. Geom. Funct. Anal. 34(1), 164–201 (2024) [Google Scholar]
- 51.Orponen, T., Yi, G.: Large cliques in extremal incidence configurations. Rev. Mat. Iberoam. 41(4), 1431–1460 (2025) [Google Scholar]
- 52.Pach, J., Tóth, G.: Graphs drawn with few crossings per edge. Combinatorica 17(3), 427–439 (1997) [Google Scholar]
- 53.Ren, K., Wang, H.: Furstenberg sets estimate in the plane. arXiv e-prints, arXiv:2308.08819 (August 2023)
- 54.Rudnev, M., Stevens, S.: An update on the sum-product problem. Math. Proc. Camb. Philos. Soc. 173(2), 411–430 (2022) [Google Scholar]
- 55.Shmerkin, P., Wang, H.: On the distance sets spanned by sets of dimension in . Geom. Funct. Anal. 35(1), 283–358 (2025) [Google Scholar]
- 56.Székely, L.A.: Crossing numbers and hard Erdős problems in discrete geometry. Combin. Probab. Comput. 6(3), 353–358 (1997) [Google Scholar]
- 57.Szemerédi, E., Trotter, W.T., Jr.: Extremal problems in discrete geometry. Combinatorica 3(3–4), 381–392 (1983) [Google Scholar]
- 58.Vinh, L.A.: The Szemerédi–Trotter type theorem and the sum-product estimate in finite fields. Eur. J. Combin. 32(8), 1177–1181 (2011) [Google Scholar]
- 59.Wang, H., Zahl, J.: Sticky Kakeya sets and the sticky Kakeya conjecture. arXiv e-prints, arXiv:2210.09581 (October 2022)
- 60.Wang, H., Zahl, J.: The Assouad dimension of Kakeya sets in . Invent. Math. 241(1), 153–206 (2025) [Google Scholar]
- 61.Wang, H., Zahl, J.: Volume estimates for unions of convex sets, and the Kakeya set conjecture in three dimensions. arXiv e-prints, arXiv:2502.17655 (February 2025)
- 62.Wolff, T.: Decay of circular means of Fourier transforms of measures. Int. Math. Res. Notices 10, 547–567 (1999) [Google Scholar]
- 63.Wolff, T.: Recent work connected with the Kakeya problem. In: Prospects in Mathematics (Princeton, NJ, 1996), pp. 129–162. American Mathematical Society, Providence, RI (1999)
- 64.Zahl, J.: On maximal functions associated to families of curves in the plane. Duke Math. J. (2025) (to appear)
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
No datasets were generated or analysed during this study.
