This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Vertical projections in the Heisenberg group
via cinematic functions and point-plate incidences

Katrin Fässler and Tuomas Orponen Department of Mathematics and Statistics
University of Jyväskylä, P.O. Box 35 (MaD)
FI-40014 University of Jyväskylä
Finland
[email protected] [email protected]
Abstract.

Let {πe:𝕎e:eS1}\{\pi_{e}\colon\mathbb{H}\to\mathbb{W}_{e}:e\in S^{1}\} be the family of vertical projections in the first Heisenberg group \mathbb{H}. We prove that if KK\subset\mathbb{H} is a Borel set with Hausdorff dimension dimK[0,2]{3}\dim_{\mathbb{H}}K\in[0,2]\cup\{3\}, then

dimπe(K)dimK\dim_{\mathbb{H}}\pi_{e}(K)\geq\dim_{\mathbb{H}}K

for 1\mathcal{H}^{1} almost every eS1e\in S^{1}. This was known earlier if dimK[0,1]\dim_{\mathbb{H}}K\in[0,1].

The proofs for dimK[0,2]\dim_{\mathbb{H}}K\in[0,2] and dimK=3\dim_{\mathbb{H}}K=3 are based on different techniques. For dimK[0,2]\dim_{\mathbb{H}}K\in[0,2], we reduce matters to a Euclidean problem, and apply the method of cinematic functions due to Pramanik, Yang, and Zahl.

To handle the case dimK=3\dim_{\mathbb{H}}K=3, we introduce a point-line duality between horizontal lines and conical lines in 3\mathbb{R}^{3}. This allows us to transform the Heisenberg problem into a point-plate incidence question in 3\mathbb{R}^{3}. To solve the latter, we apply a Kakeya inequality for plates in 3\mathbb{R}^{3}, due to Guth, Wang, and Zhang. This method also yields partial results for Borel sets KK\subset\mathbb{H} with dimK(5/2,3)\dim_{\mathbb{H}}K\in(5/2,3).

Key words and phrases:
Vertical projections, Heisenberg group, Hausdorff dimension, Incidences
2010 Mathematics Subject Classification:
28A80 (primary) 28A78 (secondary)
K.F. is supported by the Academy of Finland via the project Singular integrals, harmonic functions, and boundary regularity in Heisenberg groups, grant No. 321696. T.O. is supported by the Academy of Finland via the project Incidences on Fractals, grant No. 321896.

1. Introduction

Fix eS1×{0}e\in S^{1}\times\{0\}\subset\mathbb{H}, and consider the vertical plane 𝕎e:=e\mathbb{W}_{e}:=e^{\perp} in the first Heisenberg group \mathbb{H}, see Section 2 for the definitions. Every point pp\in\mathbb{H} can be uniquely decomposed as p=wvp=w\cdot v, where

w𝕎eandv𝕃e:=span(e).w\in\mathbb{W}_{e}\quad\text{and}\quad v\in\mathbb{L}_{e}:=\operatorname{span}(e).

This decomposition gives rise to the vertical projection πe:=π𝕎e:𝕎e\pi_{e}:=\pi_{\mathbb{W}_{e}}\colon\mathbb{H}\to\mathbb{W}_{e}, defined by πe(p):=w\pi_{e}(p):=w. A good way to visualise πe\pi_{e} is to note that the fibres πe1{w}\pi_{e}^{-1}\{w\}, w𝕎ew\in\mathbb{W}_{e}, coincide with the horizontal lines w𝕃ew\cdot\mathbb{L}_{e}. These lines foliate \mathbb{H}, as ww ranges in 𝕎e\mathbb{W}_{e}, but are not parallel. Thus, the projections πe\pi_{e} are non-linear maps with linear fibres. For example, in the special cases e1=(1,0,0)e_{1}=(1,0,0) and e2=(0,1,0)e_{2}=(0,1,0) we have the concrete formulae

πe1(x,y,t)=(0,y,t+xy2)andπe2(x,y,t)=(x,0,txy2).\pi_{e_{1}}(x,y,t)=\left(0,y,t+\tfrac{xy}{2}\right)\quad\text{and}\quad\pi_{e_{2}}(x,y,t)=\left(x,0,t-\tfrac{xy}{2}\right). (1.1)

From the point of view of geometric measure theory in the Heisenberg group, the vertical projections are the Heisenberg analogues of orthogonal projections to (d1)(d-1)-planes in d\mathbb{R}^{d}. One of the fundamental theorems concerning orthogonal projections in d\mathbb{R}^{d} is the Marstrand-Mattila projection theorem [19, 20]: if KdK\subset\mathbb{R}^{d} is a Borel set, then

dimEπV(K)=min{dimEK,d1}\dim_{\mathrm{E}}\pi_{V}(K)=\min\{\dim_{\mathrm{E}}K,d-1\} (1.2)

for almost all (d1)(d-1)-planes VdV\subset\mathbb{R}^{d}. Here dimE\dim_{\mathrm{E}} refers to Hausdorff dimension in Euclidean space – in contrast to the notation "dim\dim_{\mathrm{\mathbb{H}}}" which will refer to Hausdorff dimension in the Heisenberg group. In d\mathbb{R}^{d}, orthogonal projections are Lipschitz maps, so the upper bound in (1.2) is trivial, and the main interest in (1.2) is the lower bound.

The vertical projections πe\pi_{e} are not Lipschitz maps 𝕎e\mathbb{H}\to\mathbb{W}_{e} relative to the natural metric dd_{\mathbb{H}} in \mathbb{H} and 𝕎e\mathbb{W}_{e}. Indeed, they can increase Hausdorff dimension: an easy example is a horizontal line, which is 11-dimensional to begin with, but gets projected to a 22-dimensional set – a parabola – in almost all directions. For general (sharp) results on how much πe\pi_{e} can increase Hausdorff dimension, see [1, Theorem 1.3]. We note that the vertical planes 𝕎e\mathbb{W}_{e} themselves are 33-dimensional, and \mathbb{H} is 44-dimensional.

Can the vertical projections lower Hausdorff dimension? In some directions they can, and the general (sharp) universal lower bound was already found in [1, Theorem 1.3]:

dimπe(K)max{0,12(dimK1),2dimK5},eS1.\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq\max\{0,\tfrac{1}{2}(\dim_{\mathrm{\mathbb{H}}}K-1),2\dim_{\mathrm{\mathbb{H}}}K-5\},\qquad e\in S^{1}.

Our main result states that the dimension drop cannot occur in a set of directions of positive measure for sets of dimension in [0,2]{3}[0,2]\cup\{3\}:

Theorem 1.3.

Let KK\subset\mathbb{H} be a Borel set with dimK[0,2]{3}\dim_{\mathrm{\mathbb{H}}}K\in[0,2]\cup\{3\}. Then dimπe(K)dimK\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq\dim_{\mathrm{\mathbb{H}}}K for 1\mathcal{H}^{1} almost every eS1e\in S^{1}.

The result is sharp for all values dimK[0,2]{3}\dim_{\mathrm{\mathbb{H}}}K\in[0,2]\cup\{3\}, and new for dimK(1,2]{3}\dim_{\mathrm{\mathbb{H}}}K\in(1,2]\cup\{3\}. It makes progress in [1, Conjecture 1.5] which proposes that

dimπe(K)min{dimK,3}\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq\min\{\dim_{\mathrm{\mathbb{H}}}K,3\} (1.4)

for 1\mathcal{H}^{1} almost every eS1e\in S^{1}. The cases dimK[0,1]\dim_{\mathrm{\mathbb{H}}}K\in[0,1] were established around a decade ago by Balogh, Durand-Cartagena, the first author, Mattila, and Tyson [1, Theorem 1.4]. For dimK>1\dim_{\mathrm{\mathbb{H}}}K>1, the strongest previous partial result is due to Harris [14] who in 2022 proved that

dimπe(K)min{1+dimK2,2}for 1 a.e. eS1.\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq\min\left\{\frac{1+\dim_{\mathrm{\mathbb{H}}}K}{2},2\right\}\quad\text{for $\mathcal{H}^{1}$ a.e. $e\in S^{1}$}.

Other partial results, also higher dimensions, are contained in [2, 4, 13, 15].

The "disconnected" assumption dimK[0,2]{3}\dim_{\mathrm{\mathbb{H}}}K\in[0,2]\cup\{3\} is due to the fact that Theorem 1.3 is a combination of two separate results, with different proofs. Perhaps surprisingly, the cases dimK[0,2]\dim_{\mathrm{\mathbb{H}}}K\in[0,2] are a consequence of a "11-dimensional" projection theorem. Namely, consider the (nonlinear) projections ρe:3\rho_{e}\colon\mathbb{R}^{3}\to\mathbb{R} obtained as the tt-coordinates of the projections πe\pi_{e}:

ρe=πTπe,πT(x,y,t)=(0,0,t).\rho_{e}=\pi_{T}\circ\pi_{e},\qquad\pi_{T}(x,y,t)=(0,0,t). (1.5)

Since the tt-axis in \mathbb{H} is 22-dimensional, it is conceivable that the maps ρe\rho_{e} do not a.e. lower the Hausdorff dimension of Borel sets of dimension at most 22. This is what we prove:

Theorem 1.6.

Let K3K\subset\mathbb{R}^{3} be a Borel set. Then

dimEρe(K)=min{dimEK,1}anddimρe(K)min{dimK,2}\dim_{\mathrm{E}}\rho_{e}(K)=\min\{\dim_{\mathrm{E}}K,1\}\quad\text{and}\quad\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K)\geq\min\{\dim_{\mathrm{\mathbb{H}}}K,2\}

for 1\mathcal{H}^{1} almost every eS1e\in S^{1}. In fact, the following shaper conclusion holds: for 0s<min{dimK,2}0\leq s<\min\{\dim_{\mathrm{\mathbb{H}}}K,2\}, we have dimE{eS1:dimρe(K)s}s2\dim_{\mathrm{E}}\{e\in S^{1}:\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K)\leq s\}\leq\tfrac{s}{2}.

Theorem 1.6 implies the cases dimK[0,2]\dim_{\mathrm{\mathbb{H}}}K\in[0,2] of Theorem 1.3, because the map πT\pi_{T} is Lipschitz when restricted to any plane 𝕎e\mathbb{W}_{e}, thus dimπe(K)dimρe(K)\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K) for all eS1e\in S^{1}.

The proof of Theorem 1.6 is a fairly straightforward application of recently developed technology to study the restricted projections problem in 3\mathbb{R}^{3} (see [8, 9, 11, 17, 22]). Even though the maps ρe\rho_{e} are nonlinear, Theorem 1.6 falls within the scope of the cinematic function framework introduced by Pramanik, Yang, and Zahl [22]. In Theorem 3.2, we apply this framework to record a more general version of Theorem 1.6 which simultaneously generalises [22, Theorem 1.3] and Theorem 1.6. The details can be found in Section 3.

The case dimK=3\dim_{\mathrm{\mathbb{H}}}K=3 of Theorem 1.3 is the harder result. This time we do not know how to deduce it from a purely Euclidean statement. Instead, it is deduced from the following "mixed" result between Heisenberg and Euclidean metrics:

Theorem 1.7.

Let KK\subset\mathbb{H} be a Borel set with dimK2\dim_{\mathrm{\mathbb{H}}}K\geq 2. Then,

dimEπe(K)min{dimK1,2}\dim_{\mathrm{E}}\pi_{e}(K)\geq\min\{\dim_{\mathrm{\mathbb{H}}}K-1,2\} (1.8)

for 1\mathcal{H}^{1} almost every eS1e\in S^{1}, and consequently

dimπe(K)min{2dimK3,3}\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq\min\{2\dim_{\mathrm{\mathbb{H}}}K-3,3\} (1.9)

for 1\mathcal{H}^{1} almost every eS1e\in S^{1}.

Theorem 1.7 will further be deduced from a δ\delta-discretised result which may have independent interest. We state here a simplified version (the full version is Theorem 5.11):

Theorem 1.10.

Let 0t30\leq t\leq 3 and η>0\eta>0. Then, the following holds for δ,ϵ>0\delta,\epsilon>0 small enough, depending only on η\eta. Let \mathcal{B} be a non-empty (δ,t,δϵ)(\delta,t,\delta^{-\epsilon})-set of Heisenberg balls of radius δ\delta, all contained in B(1)B_{\mathbb{H}}(1). Then, there exists eS1e\in S^{1} such that

Leb(πe())δ3t+η.\mathrm{Leb}(\pi_{e}(\cup\mathcal{B}))\geq\delta^{3-t+\eta}. (1.11)

Here Leb\mathrm{Leb} denotes Lebesgue measure on 𝕎e\mathbb{W}_{e}, identified with 2\mathbb{R}^{2}. For the definition of (δ,t)(\delta,t)-sets of δ\delta-balls, see Definition 5.1. Theorems 1.7 and 1.10 are proved in Sections 5-7.

Remark 1.12.

It seems likely that the lower bound (1.11) remains valid under the alternative assumptions that ||=δt|\mathcal{B}|=\delta^{-t} and

|{B:BB(p,r)}|δϵ(rδ)3,p,rδ.|\{B\in\mathcal{B}:B\subset B_{\mathbb{H}}(p,r)\}|\leq\delta^{-\epsilon}\cdot\left(\frac{r}{\delta}\right)^{3},\qquad p\in\mathbb{H},\,r\geq\delta. (1.13)

This is because the estimate (1.11) ultimately follows from Proposition 6.7 which works under the non-concentration condition (1.13). We will not need this version of Theorem 1.10, so we omit the details.

1.1. Sharpness of the results

Theorem 1.3 is sharp for all values dimK[0,2]{3}\dim_{\mathrm{\mathbb{H}}}K\in[0,2]\cup\{3\}. The "mixed" inequality (1.8) in Theorem 1.7 is sharp for all values dimK2\dim_{\mathrm{\mathbb{H}}}K\geq 2, even though the Heisenberg corollary (1.9) is unlikely to be sharp for any value dimK<3\dim_{\mathrm{\mathbb{H}}}K<3 (in fact, Theorem 1.3 shows that (1.9) is not sharp for dimK<5/2)\dim_{\mathrm{\mathbb{H}}}K<5/2).

The sharpness examples are as follows: if s:=dimK2s:=\dim_{\mathrm{\mathbb{H}}}K\leq 2, take an ss-dimensional subset of the tt-axis, and note that the tt-axis is preserved by the projections ρe\rho_{e} and πe\pi_{e}. If s>2s>2, take KK to be a union of translates of the tt-axis, thus K:=K0×K:=K_{0}\times\mathbb{R}. The πe\pi_{e}-projections send vertical lines to vertical lines, so πe(K)\pi_{e}(K) is a union of vertical lines on 𝕎e\mathbb{W}_{e}; more precisely πe(K)=π¯e(K0)×\pi_{e}(K)=\bar{\pi}_{e}(K_{0})\times\mathbb{R}, where π¯e\bar{\pi}_{e} is an orthogonal projection in 2\mathbb{R}^{2}. These observations lead to the sharpness of (1.8), and the sharpness of conjecture (1.4).

Theorem 1.10 is sharp for all values of t[0,3]t\in[0,3]. Indeed, it is possible that ||=δt|\mathcal{B}|=\delta^{-t}, and then Leb(πe())δ3t\mathrm{Leb}(\pi_{e}(\cup\mathcal{B}))\lesssim\delta^{3-t} for every eS1e\in S^{1}. It also follows from (1.11) that the smallest number of dd_{\mathbb{H}}-balls of radius δ\delta needed to cover πe()\pi_{e}(\cup\mathcal{B}) is δt+η\gtrsim\delta^{-t+\eta}. One might think that this solves Conjecture 1.4 for all dimK[0,3]\dim_{\mathrm{\mathbb{H}}}K\in[0,3], but we were not able to make this deduction rigorous: the difficulty appears when attempting to δ\delta-discretise Conjecture 1.4, and is caused by the non-Lipschitz behaviour of πe:(,d)(𝕎e,d)\pi_{e}\colon(\mathbb{H},d_{\mathbb{H}})\to(\mathbb{W}_{e},d_{\mathbb{H}}). This problem will be apparent in the proof of Theorem 1.7 in Section 7. Another, more heuristic, way of understanding the difference between Theorem 1.10 and Conjecture 1.4 is this: Leb(πe(K))\mathrm{Leb}(\pi_{e}(K)) is invariant under left-translating KK, but dimπe(K)\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K) is generally not.

As we already explained, the proof of Theorem 1.6, therefore the cases dimK[0,2]\dim_{\mathrm{\mathbb{H}}}K\in[0,2] of Theorem 1.3, follow from recent developments in the theory of restricted projections in 3\mathbb{R}^{3}, notably the cinematic function framework in [22]. The proof of Theorem 1.7 does not directly overlap with these results (see Section 1.2 for more details), and for example does not use the 2\ell^{2}-decoupling theorem, in contrast with [8, 9, 11]. That said, the argument was certainly inspired by the recent developments in the restricted projection problem.

1.2. Proof outline for Theorem 1.7

The proof of Theorem 1.7 is mainly based on two ingredients. The first one is a point-line duality principle between horizontal lines in \mathbb{H}, and 3\mathbb{R}^{3}. To describe this principle, let \mathcal{L}_{\mathbb{H}} be the family of all horizontal lines in \mathbb{H}, and let 𝒞\mathcal{L}_{\mathcal{C}} be the family of all lines in 3\mathbb{R}^{3} which are parallel to some line contained in a conical surface 𝒞\mathcal{C}. In Section 4, we show that there exist maps :3\ell\colon\mathbb{R}^{3}\to\mathcal{L}_{\mathbb{H}} and :𝒞\ell^{\ast}\colon\mathbb{H}\to\mathcal{L}_{\mathcal{C}} (whose ranges cover almost all of \mathcal{L}_{\mathbb{H}} and 𝒞\mathcal{L}_{\mathcal{C}}) which preserve incidence relations in the following way:

q(p)p(q),p3,q.q\in\ell(p)\quad\Longleftrightarrow\quad p\in\ell^{\ast}(q),\qquad p\in\mathbb{R}^{3},\,q\in\mathbb{H}.

Thus, informally speaking, incidence-geometric questions between points in \mathbb{H} and lines in \mathcal{L}_{\mathbb{H}} can always be transformed into incidence-geometric questions between points in 3\mathbb{R}^{3} and lines in 𝒞\mathcal{L}_{\mathcal{C}}. The point-line duality principle described here was used implicitly by Liu [18] to study Kakeya sets (formed by horizontal lines) in \mathbb{H}. However, making the principle explicit has already proved very useful since the first version of this paper appeared: we used it in [5] to study Kakeya sets associated with SL(2)SL(2)-lines in 3\mathbb{R}^{3}, and Harris [12] used it to treat the case dimK>3\dim_{\mathrm{\mathbb{H}}}K>3 of Theorem 1.3 (in this case the projections πe(K)\pi_{e}(K) turn out to have positive measure almost surely).

The question about vertical projections in \mathbb{H} can – after suitable discretisation – be interpreted as an incidence geometric problem between points in \mathbb{H} and lines in \mathcal{L}_{\mathbb{H}}. It can therefore be transformed into an incidence-geometric problem between points in 3\mathbb{R}^{3} and lines in 𝒞\mathcal{L}_{\mathcal{C}}. Which problem is this? It turns out that while the dual (p)\ell^{\ast}(p) of a point pp\in\mathbb{H} is a line in 𝒞\mathcal{L}_{\mathcal{C}}, the dual (B)\ell^{\ast}(B_{\mathbb{H}}) of a Heisenberg δ\delta-ball resembles an δ\delta-plate in 3\mathbb{R}^{3} – a rectangle of dimensions 1×δ×δ21\times\delta\times\delta^{2} tangent to 𝒞\mathcal{C}. So, the task of proving Theorem 1.10 (hence Theorem 1.7) is (roughly) equivalent to the task of solving an incidence-geometric problem between points in 3\mathbb{R}^{3}, and family of δ\delta-plates.

Moreover: the plates in our problem appear as duals of certain Heisenberg δ\delta-balls, approximating a tt-dimensional set KK\subset\mathbb{H}, with 0t30\leq t\leq 3. Consequently, the plates can be assumed to satisfy a tt-dimensional "non-concentration condition" relative to the metric dd_{\mathbb{H}}. In common jargon, the plate family is a (δ,t)(\delta,t)-set relative to dd_{\mathbb{H}}.

In [10], Guth, Wang, and Zhang proved the sharp (reverse) square function estimate for the cone in 3\mathbb{R}^{3}. A key component in their proof was a new incidence-geometric ("Kakeya") estimate [10, Lemma 1.4] for points and δ\delta-plates in 3\mathbb{R}^{3} (see Section 6 for the details). While this was not relevant in [10], it turns out that the incidence estimate in [10, Lemma 1.4] interacts perfectly with a (δ,3)(\delta,3)-set condition relative to dd_{\mathbb{H}}. This allows us to prove, roughly speaking, that the vertical projections of 33-Frostman measures on \mathbb{H} have L2L^{2}-densities. See Corollary 5.6 for a more precise statement.

For 0t<30\leq t<3, the (δ,t)(\delta,t)-set condition relative to dd_{\mathbb{H}} no longer interacts so well with [10, Lemma 1.4]. However, we were able to (roughly speaking) reduce Theorem 1.10 for (δ,t)(\delta,t)-sets, 0t30\leq t\leq 3, to the special case t=3t=3. This argument is explained in Section 5, so we omit the discussion here.

Acknowledgements

We thank the reviewer for a careful reading of the manuscript, and for providing us with helpful comments.

2. Preliminaries on the Heisenberg group

We briefly introduce the Heisenberg group and relevant related concepts. A more thorough introduction to the geometry of the Heisenberg group can be found in many places, for instance in the monograph [3].

The Heisenberg group =(3,)\mathbb{H}=(\mathbb{R}^{3},\cdot) is the set 3\mathbb{R}^{3} equipped with the non-commutative group product defined by

(x,y,t)(x,y,t)=(x+x,y+y,t+t+12(xyyx)).(x,y,t)\cdot(x^{\prime},y^{\prime},t^{\prime})=\left(x+x^{\prime},y+y^{\prime},t+t^{\prime}+\tfrac{1}{2}(xy^{\prime}-yx^{\prime})\right).

The Heisenberg dilations are the group automorphisms δλ\delta_{\lambda}, λ>0\lambda>0, defined by

δλ(x,y,t)=(λx,λy,λ2t).\delta_{\lambda}(x,y,t)=(\lambda x,\lambda y,\lambda^{2}t).

The group product gives rise to projection-type mappings onto subgroups that are invariant under Heisenberg dilations. For eS1e\in S^{1}, we define the horizontal subgroup

𝕃e:={(se,0):s}.\mathbb{L}_{e}:=\{(se,0):\,s\in\mathbb{R}\}.

The vertical subgroup 𝕎e\mathbb{W}_{e} is the Euclidean orthogonal complement of 𝕃e\mathbb{L}_{e} in 3\mathbb{R}^{3}; in particular it is a plane containing the vertical axis. Every point pp\in\mathbb{H} can be written in a unique way as a product p=p𝕎ep𝕃ep=p_{\mathbb{W}_{e}}\cdot p_{\mathbb{L}_{e}} with p𝕎e𝕎ep_{\mathbb{W}_{e}}\in\mathbb{W}_{e} and p𝕃e𝕃ep_{\mathbb{L}_{e}}\in\mathbb{L}_{e}. The vertical Heisenberg projection onto the vertical plane 𝕎e\mathbb{W}_{e} is the map

πe:𝕎e,p=p𝕎ep𝕃ep𝕎e.\pi_{e}:\mathbb{H}\to\mathbb{W}_{e},\quad p=p_{\mathbb{W}_{e}}\cdot p_{\mathbb{L}_{e}}\mapsto p_{\mathbb{W}_{e}}.

The vertical projection to the xtxt-plane {(x,0,t):x,t}\{(x,0,t):x,t\in\mathbb{R}\} will play a special role; this projection will be denoted πxt\pi_{xt}, and it has the explicit formula stated in (1.1). Preliminaries about Heisenberg projections can be found for instance in [21, 2, 1]. These mappings have turned out to play an important role in geometric measure theory of the Heisenberg group endowed with a left-invariant non-Euclidean metric. The Korányi metric dd_{\mathbb{H}} is defined by

d(p,q):=q1p,d_{\mathbb{H}}(p,q):=\|q^{-1}\cdot p\|,

where \|\cdot\| is the Korányi norm given by

(x,y,t)=(x2+y2)2+16t24.\|(x,y,t)\|=\sqrt[4]{(x^{2}+y^{2})^{2}+16t^{2}}.

We will use the symbol B(p,r)B_{\mathbb{H}}(p,r) to denote the ball centered at pp with radius rr with respect to the Korányi metric. Balls centred at the origin are denoted B(r)B_{\mathbb{H}}(r). All vertical planes 𝕎e\mathbb{W}_{e}, eS1e\in S^{1}, equipped with dd_{\mathbb{H}} are isometric to each other via rotations of 3\mathbb{R}^{3} about the vertical axis. The Heisenberg dilations are similarities with respect to dd_{\mathbb{H}}, and it is easy to see that (,d)(\mathbb{H},d_{\mathbb{H}}) is a 44-regular space, while the vertical subgroups 𝕎e\mathbb{W}_{e} are 33-regular with respect to dd_{\mathbb{H}}. Moreover, there exists a constant 0<c<0<c<\infty, independent of ee, such that under the obvious identification of 𝕎e\mathbb{W}_{e} with 2\mathbb{R}^{2}, the restriction of the 33-dimensional Hausdorff measure 3\mathcal{H}^{3} to 𝕎e\mathbb{W}_{e} agrees with the 22-dimensional Lebesgue measure Leb\mathrm{Leb} on 2\mathbb{R}^{2} up to the multiplicative constant cc.

Vertical projections are neither group homomorphisms nor Lipschitz mappings with respect dd_{\mathbb{H}}. However, they behave well with respect to the Lebesgue measure on vertical planes. Namely, for every Borel set EE\subset\mathbb{H}, we have that

Leb(πe(pE))=Leb(πe(E)),p3,eS1,\mathrm{Leb}\left(\pi_{e}(p\cdot E)\right)=\mathrm{Leb}\left(\pi_{e}(E)\right),\qquad p\in\mathbb{R}^{3},\,e\in S^{1}, (2.1)

see the formula at the bottom of page 1970 in the proof of [7, Lemma 2.20].

3. Proof of Theorem 1.6

In this section, we prove Theorem 1.6, and therefore the cases dimK[0,2]\dim_{\mathrm{\mathbb{H}}}K\in[0,2] of Theorem 1.3. Further, Theorem 1.6 will be inferred from a more general statement, Theorem 3.2, modelled after [22, Theorem 1.3]. We first discuss Theorem 3.2, and then explain in Section 3.2 how it can be applied to deduce Theorem 1.6.

3.1. Projections induced by cinematic functions

We start by introducing terminology from [22, Definition 1.6] which will be needed for the formulation of Theorem 3.2.

Definition 3.1 (Cinematic family).

Let II\subset\mathbb{R} be a compact interval, and let C2(I)\mathcal{F}\subset C^{2}(I) be a family of functions satisfying the following conditions:

  1. (1)

    II is a compact interval, and \mathcal{F} has finite diameter in (C2(I),C2(I))(C^{2}(I),\|\cdot\|_{C^{2}(I)}).

  2. (2)

    (,C2(I))(\mathcal{F},\|\cdot\|_{C^{2}(I)}) is a doubling metric space.

  3. (3)

    For all f,gf,g\in\mathcal{F}, we have

    infθI|f(θ)g(θ)|+|f(θ)g(θ)|+|f′′(θ)g′′(θ)|fgC2(I).\inf_{\theta\in I}|f(\theta)-g(\theta)|+|f^{\prime}(\theta)-g^{\prime}(\theta)|+|f^{\prime\prime}(\theta)-g^{\prime\prime}(\theta)|\gtrsim\|f-g\|_{C^{2}(I)}.

Then, \mathcal{F} is called a cinematic family.

The following projection theorem is modelled after [22, Theorem 1.3]:

Theorem 3.2.

Let L>0L>0, let II\subset\mathbb{R} be a compact interval, and let {ρθ}θI\{\rho_{\theta}\}_{\theta\in I} be a family of LL-Lipschitz maps ρθ:B\rho_{\theta}\colon B\to\mathbb{R}, where B3B\subset\mathbb{R}^{3} is a ball. For pBp\in B, define the function fp:If_{p}\colon I\to\mathbb{R} by fp(θ):=ρθ(p)f_{p}(\theta):=\rho_{\theta}(p). Assume that pfpp\mapsto f_{p} is a bilipschitz embedding BC2(I)B\to C^{2}(I), and assume that ={fp:pB}\mathcal{F}=\{f_{p}:p\in B\} is a cinematic family.

Then, the projections {ρθ}θI\{\rho_{\theta}\}_{\theta\in I} satisfy (3.8): if K3K\subset\mathbb{R}^{3} is a Borel set, then

dimE{θI:dimEρθ(K)s}s,0s<min{dimEK,1}.\dim_{\mathrm{E}}\{\theta\in I:\dim_{\mathrm{E}}\rho_{\theta}(K)\leq s\}\leq s,\qquad 0\leq s<\min\{\dim_{\mathrm{E}}K,1\}.

We only sketch the proof of Theorem 3.2 since it is virtually the same as the proof of [22, Theorem 1.3]: this is the special case of Theorem 3.2, where

fp(θ)=ρθ(p):=γ(θ)p,p3,f_{p}(\theta)=\rho_{\theta}(p):=\gamma(\theta)\cdot p,\qquad p\in\mathbb{R}^{3}, (3.3)

and γ:IS2\gamma\colon I\to S^{2} parametrises a curve on S2S^{2} satisfying span{γ,γ˙,γ¨}=3\operatorname{span}\{\gamma,\dot{\gamma},\ddot{\gamma}\}=\mathbb{R}^{3} (this condition is needed to guarantee that the family {fp:pB}\{f_{p}:p\in B\} is cinematic for every ball B3B\subset\mathbb{R}^{3}, see the proof of [22, Proposition 2.1]).

The proof of [22, Theorem 1.3] is based on a reduction to [22, Theorem 1.7]. This is a "Kakeya-type" estimate concerning δ\delta-neighbourhoods of graphs of cinematic functions. More precisely, [22, Theorem 1.7] is only used via [22, Proposition 2.1], a special case of [22, Theorem 1.7] concerning the cinematic family {θγ(θ)p}pB\{\theta\mapsto\gamma(\theta)\cdot p\}_{p\in B}. We formulate a more general version of this proposition below: the only difference is that the cinematic family {θγ(θ)p}pB\{\theta\mapsto\gamma(\theta)\cdot p\}_{p\in B} is replaced by the family {θρθ(p)}pB\{\theta\mapsto\rho_{\theta}(p)\}_{p\in B} relevant for Theorem 3.2:

Proposition 3.4.

Fix ϵ>0\epsilon>0 and 0<αζ10<\alpha\leq\zeta\leq 1. Let II\subset\mathbb{R} be a compact interval, let B3B\subset\mathbb{R}^{3} be a ball, and let ρθ:B\rho_{\theta}\colon B\to\mathbb{R} be a family of uniformly Lipschitz functions with the properties assumed in Theorem 3.2: thus, ={fp:pB}\mathcal{F}=\{f_{p}:p\in B\} is a cinematic family, and the map pfpp\mapsto f_{p} is a bilipschitz embedding BC2(I)B\to C^{2}(I), where fp(θ):=ρθ(p)f_{p}(\theta):=\rho_{\theta}(p). Then there exists δ0>0\delta_{0}>0 such that the following holds for all δ(0,δ0]\delta\in(0,\delta_{0}]:

Let E2E\subset\mathbb{R}^{2} be a (δ,α;δϵ)1×(δ,α;δϵ)1(\delta,\alpha;\delta^{-\epsilon})_{1}\times(\delta,\alpha;\delta^{-\epsilon})_{1} quasi-product. Let ZδBZ_{\delta}\subset B be a δ\delta-separated set that satisfies

|ZδB(p,r)|δϵ(r/δ)ζ,p3,rδ.|Z_{\delta}\cap B(p,r)|\leq\delta^{-\epsilon}(r/\delta)^{\zeta},\qquad p\in\mathbb{R}^{3},\,r\geq\delta. (3.5)

Then

E(pZδ𝟏Γpδ)3/2δ2α/2ζ/2Cϵ|Zδ|,\int_{E}\Big{(}\sum_{p\in Z_{\delta}}\mathbf{1}_{\Gamma_{p}^{\delta}}\Big{)}^{3/2}\leq\delta^{2-\alpha/2-\zeta/2-C\epsilon}|Z_{\delta}|,

where C>0C>0 is absolute, and Γpδ\Gamma_{p}^{\delta} is the δ\delta-neighbourhood of the graph of fpf_{p}.

Proof.

The proof of [22, Proposition 2.1] is easy (given [22, Theorem 1.7]), but the proof of Proposition 3.4 is almost trivial. Indeed, the first part in the proof of [22, Proposition 2.1] is to verify that the family {θρθ(p)}pB\{\theta\mapsto\rho_{\theta}(p)\}_{p\in B} is cinematic in the case ρθ(p)=γ(θ)p\rho_{\theta}(p)=\gamma(\theta)\cdot p, but this is already a part of our hypothesis. The second part in the proof of [22, Proposition 2.1] is to verify that pfpp\mapsto f_{p} is a bilipschitz embedding BC2(I)B\to C^{2}(I), and this is – again – part of our hypothesis. In other words, all the work in the proof of [22, Proposition 2.1] has been made part of the hypotheses of Proposition 3.4. ∎

The reduction from [22, Theorem 1.3] to [22, Proposition 2.1] (in our case from Theorem 3.2 to Proposition 3.4) is presented in [22, Sections 2.1-2.4], and does not use the special form (3.3) (for example the linearity) of the maps ρθ:3\rho_{\theta}\colon\mathbb{R}^{3}\to\mathbb{R}: it is only needed that

  1. (1)

    the maps ρθ\rho_{\theta} are uniformly Lipschitz, for θI\theta\in I,

  2. (2)

    suppBsupθI|θρθ(p)|<\sup_{p\in B}\sup_{\theta\in I}|\partial_{\theta}\rho_{\theta}(p)|<\infty.

Property (1) is assumed in Theorem 3.2, whereas property (2) follows from the assumption that the family \mathcal{F} is cinematic (and in particular a bounded subset of C2(I)C^{2}(I)).

The argument in [22, Sections 2.1-2.4] is extremely well-written, and our notation is deliberately the same, so we will not copy the whole proof. We only make a few remarks, below. If the reader is unfamiliar with the ideas involved, we warmly recommend reading first the heuristic section [22, Section 1.2].

Proof sketch of Theorem 3.2.

The argument in [22, Section 2.1] can be copied verbatim; nothing changes. The most substantial change occurs in [22, Section 2.2]. Namely, [22, (2.10)] uses the fact (true in [22]) that the ρθ\rho_{\theta}-image of a δ\delta-cube Q3Q\subset\mathbb{R}^{3} has length |ρθ(Q)|δ|\rho_{\theta}(Q)|\gtrsim\delta. For the general Lipschitz maps ρθ\rho_{\theta} in Theorem 3.2 this may not be the case; it would be true for the special maps ρθ\rho_{\theta} needed in Theorem 3.7, so also this part of [22] would work verbatim for these maps. However, even in the generality of Theorem 3.2 the problem can be completely removed: one only needs to replace every occurrence of ρθ(Q)\rho_{\theta}(Q) in [22, Section 2.2] by an interval

Iθ(Q):=[ρθ(zQ)δ,ρθ(zQ)+δ]I_{\theta}(Q):=[\rho_{\theta}(z_{Q})-\delta,\rho_{\theta}(z_{Q})+\delta]

of length δ\sim\delta centred at ρθ(zQ)\rho_{\theta}(z_{Q}), where zQQz_{Q}\in Q is the centre of QQ. Since ρθ(Q)\rho_{\theta}(Q) only appears as a "tool" in [22, Section 2.2], the rest of the argument will remain unchanged. Let us, however, discuss what changes in [22, Section 2.2] when ρθ(Q)\rho_{\theta}(Q) is replaced by Iθ(Q)I_{\theta}(Q). We assume familiarity with the notation in [22].

First and foremost, [22, (2.9)] remains valid: whenever Q𝒬Q\in\mathcal{Q} is a cube that intersects ρθ1(Gθ)\rho_{\theta}^{-1}(G_{\theta}), then dist(ρθ(zQ),Gθ)Lδ\operatorname{dist}(\rho_{\theta}(z_{Q}),G_{\theta})\lesssim L\delta by our assumption that the maps ρθ\rho_{\theta} are LL-Lipschitz. Therefore,

Iθ(Q)Gθ:=NLδ(Gθ).I_{\theta}(Q)\subset G_{\theta}^{\prime}:=N_{L\delta}(G_{\theta}).

This gives [22, (2.9)] with the slightly modified definition of GθG_{\theta}^{\prime}, stated above. Consequently, also the version of [22, (2.10)] is true where ρθ(Q)\rho_{\theta}(Q) is replaced by Iθ(Q)I_{\theta}(Q): here the length bound |Iδ(Q)|δ|I_{\delta}(Q)|\gtrsim\delta is used. Finally, to deduce [22, (2.13)] from [22, (2.10)], we need to know that [22, (2.12)] remains valid when ρθ(Q)\rho_{\theta}(Q) is replaced with Iθ(Q)I_{\theta}(Q). This is clear: if yIθ(Q)y\in I_{\theta}(Q), then |yρθ(zQ)|δ|y-\rho_{\theta}(z_{Q})|\leq\delta by definition, and therefore (θ,y)ΓzQδ(\theta,y)\in\Gamma_{z_{Q}}^{\delta}, where

Γz={(θ,ρθ(z)):θI},zB,\Gamma_{z}=\{(\theta,\rho_{\theta}(z)):\theta\in I\},\qquad z\in B,

is the analogue of [22, (1.12)], and Γzδ\Gamma_{z}^{\delta} is the δ\delta-neighbourhood of Γz\Gamma_{z}. We have now verified [22, (2.13)]. The intervals ρθ(Q)\rho_{\theta}(Q) or Iθ(Q)I_{\theta}(Q) play no further role in the proof. The rest of [22, Section 2.2] works verbatim.

The same is also true for [22, Section 2.3]: the argument is fairly abstract down to [22, (2.19)], where it is needed that suppBsupθI|θρθ(p)|<\sup_{p\in B}\sup_{\theta\in I}|\partial_{\theta}\rho_{\theta}(p)|<\infty. The maps ρθ\rho_{\theta} in Theorem 3.2 satisfy this property automatically, as noted in (2) above.

Finally, we arrive at the short [22, Section 2.4]. The only difference is that we need to apply Proposition 3.4 in place of [22, Proposition 2.1]. This completes the proof of Theorem 3.2. ∎

3.2. From vertical projections to cinematic functions

We explain how the general projection result, Theorem 3.2, can be applied to prove Theorem 3.7, which concerns the special projections ρe=πTπe\rho_{e}=\pi_{T}\circ\pi_{e}. Recall that πe\pi_{e} is the vertical projection to the plane 𝕎e=e\mathbb{W}_{e}=e^{\perp}. For e=(e1,e2)S1e=(e_{1},e_{2})\in S^{1}, we write J(e):=(e2,e1)S1eJ(e):=(-e_{2},e_{1})\in S^{1}\cap e^{\perp} is the counterclockwise rotation of ee by π/2\pi/2. With this notation, the map πe\pi_{e} has the explicit formula

πe(z,t)=(z,Je,t+12z,ez,Je),\pi_{e}(z,t)=(\langle z,Je\rangle,t+\tfrac{1}{2}\langle z,e\rangle\langle z,Je\rangle), (3.6)

where ,\langle\cdot,\cdot\rangle is the Euclidean dot product in 2\mathbb{R}^{2}. In the formula (3.6), we have also identified each plane 𝕎e\mathbb{W}_{e} with 2\mathbb{R}^{2} via the map (yJe,t)(y,t)(yJe,t)\cong(y,t). It is worth noting that the distance dd_{\mathbb{H}} restricted to the plane 𝕎e\mathbb{W}_{e} (for eS1e\in S^{1} fixed) is bilipschitz equivalent to the parabolic distance on 2\mathbb{R}^{2}, namely dpar((x,s),(y,t))=|xy|+|st|d_{\mathrm{par}}((x,s),(y,t))=|x-y|+\sqrt{|s-t|}.

With the explicit expression (3.6) in hand, the nonlinear projections ρe=πTπe\rho_{e}=\pi_{T}\circ\pi_{e} introduced in (1.5) have the following formula:

ρe(z,t)=t+12z,ez,Je,(z,t)2×,eS1,\rho_{e}(z,t)=t+\tfrac{1}{2}\langle z,e\rangle\langle z,Je\rangle,\qquad(z,t)\in\mathbb{R}^{2}\times\mathbb{R},\,e\in S^{1},

By a slight abuse of notation, we write "dd_{\mathbb{H}}" for the square root metric on \mathbb{R}: thus d(s,t):=std_{\mathbb{H}}(s,t):=\sqrt{s-t}. The projection πT\pi_{T} restricted to any fixed plane 𝕎e\mathbb{W}_{e} is a Lipschitz map (𝕎e,d)(,d)(\mathbb{W}_{e},d_{\mathbb{H}})\to(\mathbb{R},d_{\mathbb{H}}), even though πT\pi_{T} is not "globally" a Lipschitz map (,d)(,d)(\mathbb{H},d_{\mathbb{H}})\to(\mathbb{R},d_{\mathbb{H}}). Therefore dimπe(K)dimρe(K)\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K) for all eS1e\in S^{1}, and the cases dimK[0,2]\dim_{\mathrm{\mathbb{H}}}K\in[0,2] of Theorem 1.3 follow from Theorem 1.6, whose contents are repeated here:

Theorem 3.7.

Let K3K\subset\mathbb{R}^{3} be Borel, and let 0s<min{dimEK,1}0\leq s<\min\{\dim_{\mathrm{E}}K,1\}. Then,

dimE{eS1:dimEρe(K)s}s.\dim_{\mathrm{E}}\{e\in S^{1}:\dim_{\mathrm{E}}\rho_{e}(K)\leq s\}\leq s. (3.8)

As a consequence, for every 0s<min{dimK,2}0\leq s<\min\{\dim_{\mathrm{\mathbb{H}}}K,2\},

dimE{eS1:dimρe(K)s}s2.\dim_{\mathrm{E}}\{e\in S^{1}:\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K)\leq s\}\leq\tfrac{s}{2}. (3.9)

In particular, dimρe(K)min{dimK,2}\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K)\geq\min\{\dim_{\mathrm{\mathbb{H}}}K,2\} for 1\mathcal{H}^{1} almost every eS1e\in S^{1}.

Remark 3.10.

We explain why (3.8) implies (3.9). It is well-known that

dimK2dimEK.\dim_{\mathrm{\mathbb{H}}}K\leq 2\dim_{\mathrm{E}}K.

for all sets KK\subset\mathbb{H}. This simply follows from the fact that the identity map (,dEuc)(,d)(\mathbb{H},d_{\mathrm{Euc}})\to(\mathbb{H},d_{\mathbb{H}}) is locally 12\tfrac{1}{2}-Hölder continuous. Therefore, if 0s<min{dimK,2}0\leq s<\min\{\dim_{\mathrm{\mathbb{H}}}K,2\}, as in (3.9), we have 0s2<min{dimEK,1}0\leq\tfrac{s}{2}<\min\{\dim_{\mathrm{E}}K,1\}, and (3.8) is applicable. Since

{eS1:dimρe(K)s}={eS1:dimEρe(K)s2}\{e\in S^{1}:\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K)\leq s\}=\{e\in S^{1}:\dim_{\mathrm{E}}\rho_{e}(K)\leq\tfrac{s}{2}\}

(the square root metric on \mathbb{R} doubles Euclidean dimension), we have

dimE{eS1:dimρe(K)s}=dimE{eS1:dimEρe(K)s2}(3.8)s2.\dim_{\mathrm{E}}\{e\in S^{1}:\dim_{\mathrm{\mathbb{H}}}\rho_{e}(K)\leq s\}=\dim_{\mathrm{E}}\{e\in S^{1}:\dim_{\mathrm{E}}\rho_{e}(K)\leq\tfrac{s}{2}\}\stackrel{{\scriptstyle\eqref{form46}}}{{\leq}}\tfrac{s}{2}.

This is what we claimed in (3.9).

For the remainder of this section, we focus on proving the Euclidean statement (3.8). This is chiefly based on verifying that the projections ρe:3\rho_{e}\colon\mathbb{R}^{3}\to\mathbb{R} give rise to a cinematic family of functions, as in Definition 3.1. Let us introduce the relevant cinematic family. We re-parametrise the projections ρe\rho_{e}, eS1e\in S^{1}, as ρθ\rho_{\theta}, θ\theta\in\mathbb{R}, where

ρθ:=ρe(θ),e(θ):=(cosθ,sinθ).\rho_{\theta}:=\rho_{e(\theta)},\qquad e(\theta):=(\cos\theta,\sin\theta).

With this notation, we define the following functions fp:f_{p}\colon\mathbb{R}\to\mathbb{R}, p3p\in\mathbb{R}^{3}:

fp(θ):=ρθ(p):=t+12z,e(θ)z,Je(θ),p=(z,t)3.f_{p}(\theta):=\rho_{\theta}(p):=t+\tfrac{1}{2}\langle z,e(\theta)\rangle\langle z,Je(\theta)\rangle,\qquad p=(z,t)\in\mathbb{R}^{3}. (3.11)
Proposition 3.12.

Let p03{(0,0,t):t}p_{0}\in\mathbb{R}^{3}\,\setminus\,\{(0,0,t):t\in\mathbb{R}\}. Then, there exists a radius r=r(p0)>0r=r(p_{0})>0 such that (B(p0,r)):={fp:pB(p0,r)}\mathcal{F}(B(p_{0},r)):=\{f_{p}:p\in B(p_{0},r)\} is a cinematic family.

The compact interval appearing in conditions (1)-(3) of Definition 3.1 can be taken to be [0,2π][0,2\pi] – this makes no difference, since the functions fpf_{p} are 2π2\pi-periodic. It turns out that the conditions (1)-(2) are satisfied for the family (B)\mathcal{F}(B), whenever B3B\subset\mathbb{R}^{3} is an arbitrary ball. To verify condition (3), we will need to assume that BB lies outside the tt-axis; we will return to this a little later. We first compute the derivatives of the functions in \mathcal{F}. For fpf_{p}\in\mathcal{F}, we have

fp(θ)=12z,e(θ)z,Je(θ)+12z,e(θ)z,Je(θ).f_{p}^{\prime}(\theta)=\tfrac{1}{2}\langle z,e^{\prime}(\theta)\rangle\langle z,Je(\theta)\rangle+\tfrac{1}{2}\langle z,e(\theta)\rangle\langle z,Je^{\prime}(\theta)\rangle.

This expression can be further simplified by noting that e(θ)=Je(θ)e^{\prime}(\theta)=Je(\theta), and Je(θ)=e(θ)Je^{\prime}(\theta)=-e(\theta). Therefore,

fp(θ)=12z,Je(θ)212z,e(θ)2.f_{p}^{\prime}(\theta)=\tfrac{1}{2}\langle z,Je(\theta)\rangle^{2}-\tfrac{1}{2}\langle z,e(\theta)\rangle^{2}. (3.13)

From this expression, we may compute the second derivative:

fp′′(θ)=z,Je(θ)z,Je(θ)z,e(θ)z,e(θ)=2z,e(θ)z,Je(θ).f_{p}^{\prime\prime}(\theta)=\langle z,Je(\theta)\rangle\langle z,Je^{\prime}(\theta)\rangle-\langle z,e(\theta)\rangle\langle z,e^{\prime}(\theta)\rangle=-2\langle z,e(\theta)\rangle\langle z,Je(\theta)\rangle. (3.14)

The formulae (3.11)-(3.14) immediately show that the map pfpp\mapsto f_{p} is locally Lipschitz:

supθ|fp(θ)fq(θ)|+|fp(θ)fq(θ)|+|fp′′(θ)fq′′(θ)|B|pq|,p,qB.\sup_{\theta\in\mathbb{R}}|f_{p}(\theta)-f_{q}(\theta)|+|f_{p}^{\prime}(\theta)-f_{q}^{\prime}(\theta)|+|f_{p}^{\prime\prime}(\theta)-f_{q}^{\prime\prime}(\theta)|\lesssim_{B}|p-q|,\qquad p,q\in B. (3.15)

This implies conditions (1)-(2) in Definition 3.1 for the family (B)\mathcal{F}(B). Regarding condition (3) in Definition 3.1, we claim the following:

Proposition 3.16.

If p03{(0,0,t):t}p_{0}\in\mathbb{R}^{3}\,\setminus\,\{(0,0,t):t\in\mathbb{R}\}, there exists a radius r=r(p0)>0r=r(p_{0})>0 and a constant c=c(p0)>0c=c(p_{0})>0 such that

|fp(θ)fq(θ)|+|fp(θ)fq(θ)|+|fp′′(θ)fq′′(θ)|c|pq||f_{p}(\theta)-f_{q}(\theta)|+|f_{p}^{\prime}(\theta)-f_{q}^{\prime}(\theta)|+|f_{p}^{\prime\prime}(\theta)-f_{q}^{\prime\prime}(\theta)|\geq c|p-q| (3.17)

for all p,qB(p0,r)p,q\in B(p_{0},r) and θ\theta\in\mathbb{R}.

We start with the following lemma:

Lemma 3.18.

For every p03{(0,0,t):t}p_{0}\in\mathbb{R}^{3}\,\setminus\,\{(0,0,t):t\in\mathbb{R}\} there exists a constant c>0c>0 and a radius r>0r>0 such that the following holds:

|fp(0)fq(0)|+|fp(0)fq(0)|+|fp′′(0)fq′′(0)|c|pq|,p,qB(p0,r).|f_{p}(0)-f_{q}(0)|+|f_{p}^{\prime}(0)-f_{q}^{\prime}(0)|+|f_{p}^{\prime\prime}(0)-f_{q}^{\prime\prime}(0)|\geq c|p-q|,\qquad p,q\in B(p_{0},r). (3.19)
Proof.

Recall that e(0)=(1,0)e(0)=(1,0) and Je(0)=(0,1)Je(0)=(0,1). We then define F:33F\colon\mathbb{R}^{3}\to\mathbb{R}^{3} by

F(p):=(fp(0),fp(0),fp′′(0))=(t+12z1z2,12(z22z12),2z1z2),p=(z,t)3.F(p):=(f_{p}(0),f_{p}^{\prime}(0),f_{p}^{\prime\prime}(0))=(t+\tfrac{1}{2}z_{1}z_{2},\tfrac{1}{2}(z_{2}^{2}-z_{1}^{2}),-2z_{1}z_{2}),\qquad p=(z,t)\in\mathbb{R}^{3}.

Then, we note that |detDF(p)|=2|z|2|\mathrm{det}DF(p)|=2|z|^{2}, so in particular the Jacobian of FF is non-vanishing outside the tt-axis. Now (3.19) follows from the inverse function theorem. ∎

We then prove Proposition 3.16:

Proof of Proposition 3.16.

To deduce (3.17) from (3.19), we record the following rotation invariance:

fRφ(p)(k)(θ+φ)=fp(k)(θ),p3,θ,φ.f^{(k)}_{R_{\varphi}(p)}(\theta+\varphi)=f^{(k)}_{p}(\theta),\qquad p\in\mathbb{R}^{3},\,\theta,\varphi\in\mathbb{R}. (3.20)

Here Rφ(z,t):=(eiφz,t)R_{\varphi}(z,t):=(e^{i\varphi}z,t) is a counterclockwise rotation around the tt-axis. The proof is evident from the formulae (3.11)-(3.14), and noting that

eiφz,e(θ+φ)=z,e(θ)andeiφz,Je(θ+φ)=z,Je(θ).\langle e^{i\varphi}z,e(\theta+\varphi)\rangle=\langle z,e(\theta)\rangle\quad\text{and}\quad\langle e^{i\varphi}z,Je(\theta+\varphi)\rangle=\langle z,Je(\theta)\rangle.

Now we are in a position to conclude the proof of (3.17). Fix p03{(0,0,t):t}p_{0}\in\mathbb{R}^{3}\,\,\setminus\,\{(0,0,t):t\in\mathbb{R}\} and θ0\theta_{0}\in\mathbb{R}. Then, apply Lemma 3.18 to the point

Rθ0(p0)3{(0,0,t):t}.R_{-\theta_{0}}(p_{0})\in\mathbb{R}^{3}\,\setminus\,\{(0,0,t):t\in\mathbb{R}\}.

This yields a constant c=c(p0,θ0)>0c=c(p_{0},\theta_{0})>0 and a radius r0=r0(p0,θ0)>0r_{0}=r_{0}(p_{0},\theta_{0})>0 such that

|fp(0)fq(0)|+|fp(0)fq(0)|+|fp′′(0)fq′′(0)|c|pq||f_{p}(0)-f_{q}(0)|+|f_{p}^{\prime}(0)-f_{q}^{\prime}(0)|+|f_{p}^{\prime\prime}(0)-f_{q}^{\prime\prime}(0)|\geq c|p-q| (3.21)

for all p,qB(Rθ0(p0),2r0)p,q\in B(R_{-\theta_{0}}(p_{0}),2r_{0}). Next, we choose I(θ0)=[θ0r1,θ0+r1]I(\theta_{0})=[\theta_{0}-r_{1},\theta_{0}+r_{1}] to be a sufficiently short interval around θ0\theta_{0} such that the following holds:

Rθ(p),Rθ(q)B(Rθ0(p0),2r0),p,qB(p0,r0),θI(θ0).R_{-\theta}(p),R_{-\theta}(q)\in B(R_{-\theta_{0}}(p_{0}),2r_{0}),\qquad p,q\in B(p_{0},r_{0}),\,\theta\in I(\theta_{0}).

Then, it follows from a combination of (3.20) and (3.21) that

k=02|fp(k)(θ)fq(k)(θ)|=(3.20)k=02|fRθ(p)(k)(0)fRθ(q)(k)(0)|(3.21)c|Rθ(p)Rθ(q)|=c|pq|\sum_{k=0}^{2}|f_{p}^{(k)}(\theta)-f_{q}^{(k)}(\theta)|\stackrel{{\scriptstyle\eqref{form56}}}{{=}}\sum_{k=0}^{2}|f_{R_{-\theta}(p)}^{(k)}(0)-f_{R_{-\theta}(q)}^{(k)}(0)|\stackrel{{\scriptstyle\eqref{form57}}}{{\geq}}c|R_{-\theta}(p)-R_{-\theta}(q)|=c|p-q|

for all p,qB(p0,r0)p,q\in B(p_{0},r_{0}) and all θI(θ0)\theta\in I(\theta_{0}). This completes the proof of (3.17) of all θI(θ0)\theta\in I(\theta_{0}). To extend the argument of all θ\theta\in\mathbb{R}, note that the functions fpf_{p}, and all of their derivatives, are 2π2\pi-periodic. So, it suffices to show that (3.17) holds for θ[0,2π]\theta\in[0,2\pi]. This follows by compactness from what we have already proven, by covering [0,2π][0,2\pi] by finitely many intervals of the form I(θ0)I(\theta_{0}), and finally defining "rr" and "cc" to be the minima of the constants r(p0,θ0)r(p_{0},\theta_{0}) and c(p0,θ0)c(p_{0},\theta_{0}) obtained in the process. ∎

Proposition 3.12 now follows from Proposition 3.16, and the discussion above it (where we verified Definition 3.1(1)-(2)). We then conclude the proof of Theorem 3.7:

Proof of Theorem 3.7.

Given Remark 3.10, it suffices to prove (3.8), which will be a consequence of Theorem 3.2. Indeed, since the projections ρe\rho_{e} are isometries on the tt-axis, we may assume that

dimE(K{(0,0,t):t})=dimEK.\dim_{E}(K\,\setminus\,\{(0,0,t):t\in\mathbb{R}\})=\dim_{E}K.

Consequently, for ϵ>0\epsilon>0, we may fix a point p0Kp_{0}\in K outside the tt-axis such that

dimE(KB(p0,r))>dimEKϵ,r>0.\dim_{\mathrm{E}}(K\cap B(p_{0},r))>\dim_{\mathrm{E}}K-\epsilon,\qquad r>0. (3.22)

Apply Proposition 3.12 to find a radius r>0r>0 such that the family of functions :=(B(p0,r))\mathcal{F}:=\mathcal{F}(B(p_{0},r)) is cinematic. It follows from a combination of (3.15) and Proposition 3.16 that pfpp\mapsto f_{p} is a bilipschitz embedding BC2()B\to C^{2}(\mathbb{R}). Therefore Theorem 3.2 is applicable: for every 0s<min{dimE(KB(p0,r)),1}0\leq s<\min\{\dim_{\mathrm{E}}(K\cap B(p_{0},r)),1\} we have

dimE{θ[0,2π]:dimEρθ(KB(p0,r))s}s.\dim_{\mathrm{E}}\{\theta\in[0,2\pi]:\dim_{\mathrm{E}}\rho_{\theta}(K\cap B(p_{0},r))\leq s\}\leq s.

Now (3.8) follows from (3.22) by letting ϵ0\epsilon\to 0. ∎

4. Duality between horizontal lines and 3\mathbb{R}^{3}

This section contains preliminaries to prove Theorem 1.7. Most importantly, we introduce a notion of duality that associates to points and horizontal lines in \mathbb{H} certain lines and points in 3\mathbb{R}^{3}. The lines in 3\mathbb{R}^{3} will be light rays – translates of lines on a fixed conical surface. To define these, we let 𝒞0\mathcal{C}_{0} be the vertical cone

𝒞0={(z1,z2,z3)3:z12+z22=z32},\mathcal{C}_{0}=\{(z_{1},z_{2},z_{3})\in\mathbb{R}^{3}:\,z_{1}^{2}+z_{2}^{2}=z_{3}^{2}\},

and we denote by 𝒞\mathcal{C} the (4545^{\circ}) rotated cone

𝒞=R(𝒞0)={(z1,z2,z3)3:z22=2z1z3},\mathcal{C}=R(\mathcal{C}_{0})=\{(z_{1},z_{2},z_{3})\in\mathbb{R}^{3}:\,z_{2}^{2}=2z_{1}z_{3}\},

where R(z1,z2,z3)=((z1+z3)/2,z2,(z1+z3)/2)R(z_{1},z_{2},z_{3})=\left((z_{1}+z_{3})/\sqrt{2},z_{2},(-z_{1}+z_{3})/\sqrt{2}\right). The cone 𝒞\mathcal{C} is foliated by lines

Ly=span(1,y,y2/2),y,L_{y}=\mathrm{span}_{\mathbb{R}}(1,-y,y^{2}/2),\quad y\in\mathbb{R}, (4.1)

cf. the proof of [18, Theorem 1.2], where a similar parametrization is used. To be accurate, the lines LyL_{y} only foliate 𝒞{(0,0,z):z}\mathcal{C}\,\setminus\,\{(0,0,z):z\in\mathbb{R}\}. We will abuse notation by writing Ly(s)=(s,sy,sy2/2)L_{y}(s)=(s,-sy,sy^{2}/2) for the parametrisation of the line LyL_{y}.

Definition 4.2 (Light rays).

We say that a line LL in 3\mathbb{R}^{3} is a light ray if L=z+LyL=z+L_{y} for some z3z\in\mathbb{R}^{3} and yy\in\mathbb{R}. In other words, LL is a (Euclidean) translate of a line contained in 𝒞\mathcal{C} (excluding the tt-axis).

Remark 4.3.

Every light ray can be written as (0,u,v)+Ly(0,u,v)+L_{y} for a unique (u,v)2(u,v)\in\mathbb{R}^{2}.

Definition 4.4 (Horizontal lines).

A line \ell in 3\mathbb{R}^{3} is horizontal if it is a Heisenberg left translate of a horizontal subgroup, that is, there exists pp\in\mathbb{H} and eS1e\in S^{1} such that =p𝕃e\ell=p\cdot\mathbb{L}_{e}.

Remark 4.5.

Every horizontal line, apart from left translates of the xx-axis, can be written as ={(as+b,s,(b/2)s+c):s}\ell=\{(as+b,s,(b/2)s+c):\,s\in\mathbb{R}\} for a uniquely determined point (a,b,c)3(a,b,c)\in\mathbb{R}^{3}.

Definition 4.6.

We define the following correspondence between points and lines:

  • To a point p=(x,y,t)p=(x,y,t)\in\mathbb{H}, we associate the light ray

    (p)=(0,x,txy/2)+Ly(0,x,txy/2)+𝒞3.\ell^{\ast}(p)=(0,x,t-xy/2)+L_{y}\subset(0,x,t-xy/2)+\mathcal{C}\subset\mathbb{R}^{3}. (4.7)

    (This formula will be motivated by Lemma 4.11 below.)

  • To a point p=(a,b,c)3p^{\ast}=(a,b,c)\in\mathbb{R}^{3}, we associate the horizontal line

    (p)={(as+b,s,b2s+c):s}.\ell(p^{\ast})=\{(as+b,s,\tfrac{b}{2}s+c):\,s\in\mathbb{R}\}.

Given a set 𝒫\mathcal{P} of points in \mathbb{H}, we define the family of light rays

(𝒫)=p𝒫(p).\ell^{\ast}(\mathcal{P})=\bigcup_{p\in\mathcal{P}}\ell^{\ast}(p). (4.8)
Remark 4.9.

It is worth observing that the point (0,x,txy/2)(0,x,t-xy/2) appearing in formula (4.7) is nearly the vertical projection of (x,y,t)(x,y,t) to the xtxt-plane; the actual formula for this projection would be πxt(x,y,t)=(x,0,txy/2)\pi_{xt}(x,y,t)=(x,0,t-xy/2). It follows from this observation that

((u,0,v)(0,y,0))=(0,u,v)+Ly,u,v,y,\ell^{\ast}((u,0,v)\cdot(0,y,0))=(0,u,v)+L_{y},\qquad u,v,y\in\mathbb{R}, (4.10)

because πxt((u,0,v)(0,y,0))=(u,0,v)\pi_{xt}((u,0,v)\cdot(0,y,0))=(u,0,v).

Under the point-line correspondence in Definition 4.6, incidences between points and horizontal lines in \mathbb{H} are in one-to-one correspondence with incidences between light rays and points in 3\mathbb{R}^{3}.

Lemma 4.11 (Incidences are preserved under duality).

For pp\in\mathbb{H} and p3p^{\ast}\in\mathbb{R}^{3}, we have

p(p)p(p).p\in\ell(p^{\ast})\quad\Longleftrightarrow\quad p^{\ast}\in\ell^{\ast}(p).
Proof.

Let p=(x,y,t)p=(x,y,t)\in\mathbb{H} and p=(a,b,c)3p^{\ast}=(a,b,c)\in\mathbb{R}^{3}. The condition p(p)p\in\ell(p^{\ast}) is equivalent to

{ay+b=xb2y+c=t.\left\{\begin{array}[]{l}ay+b=x\\ \tfrac{b}{2}y+c=t.\end{array}\right.

Recalling the notation Ly(s)=(s,sy,sy2/2)L_{y}(s)=(s,-sy,sy^{2}/2), this is further equivalent to

p=(a,b,c)=(0,x,txy/2)+Ly(a).p^{\ast}=(a,b,c)=(0,x,t-xy/2)+L_{y}(a). (4.12)

Finally, (4.12) is equivalent to p(p)p^{\ast}\in\ell^{\ast}(p). ∎

4.1. Measures on the space of horizontal lines

The duality p(p)p\mapsto\ell(p) between points in p3p\in\mathbb{R}^{3} and horizontal lines (p)\ell(p) in Definition 4.6 allows one to push-forward Lebesgue measure "Leb\mathrm{Leb}" on 3\mathbb{R}^{3} to construct a measure "𝔪\mathfrak{m}" on the set of horizontal lines:

𝔪():=(Leb)()=Leb({p3:(p)}).\mathfrak{m}(\mathcal{L}):=(\ell_{\sharp}\mathrm{Leb})(\mathcal{L})=\mathrm{Leb}(\{p\in\mathbb{R}^{3}:\ell(p)\in\mathcal{L}\}).

There is, however, a more commonly used measure on the space of horizontal lines. This measure "𝔥\mathfrak{h}" is discussed extensively for example in [6, Section 2.3]. The measure 𝔥\mathfrak{h} is the unique (up to a multiplicative constant) non-zero left invariant measure on the set of horizontal lines. One possible formula for it is the following:

𝔥()=S13({w𝕎e:πe1{w}})𝑑1(e).\mathfrak{h}(\mathcal{L})=\int_{S^{1}}\mathcal{H}^{3}(\{w\in\mathbb{W}_{e}:\pi_{e}^{-1}\{w\}\in\mathcal{L}\})\,d\mathcal{H}^{1}(e). (4.13)

Let fL1()f\in L^{1}(\mathbb{H}), and consider the weighted measure μf:=fdLeb\mu_{f}:=f\,d\mathrm{Leb}. Then, starting from the definition (4.13), it is easy to check that

S1πeμfL22𝑑1(e)=Xf()2𝑑𝔥(),\int_{S^{1}}\|\pi_{e}\mu_{f}\|_{L^{2}}^{2}\,d\mathcal{H}^{1}(e)=\int Xf(\ell)^{2}\,d\mathfrak{h}(\ell), (4.14)

where Xf():=f𝑑1Xf(\ell):=\int_{\ell}f\,d\mathcal{H}^{1}.

While the measure 𝔥\mathfrak{h} is mutually absolutely continuous with respect to 𝔪\mathfrak{m}, the Radon-Nikodym derivative is not bounded (from above and below): with our current notational conventions, the lines (p)\ell(p) are never parallel to the xx-axis, and the 𝔪\mathfrak{m}-density of lines making a small angle with the xx-axis is smaller than their 𝔥\mathfrak{h}-density. The problem can be removed by restricting our considerations to lines which make a substantial angle with the xx-axis. For example, let \mathcal{L}_{\angle} be the set of horizontal lines which have slope at most 11 relative to the yy-axis; thus

=({(a,b,c)3:|a|1}).\mathcal{L}_{\angle}=\ell(\{(a,b,c)\in\mathbb{R}^{3}:|a|\leq 1\}).

Then, 𝔪()𝔥()\mathfrak{m}(\mathcal{L})\sim\mathfrak{h}(\mathcal{L}) for all Borel sets \mathcal{L}\subset\mathcal{L}_{\angle}. The lines in \mathcal{L}_{\angle} coincide with pre-images of the form πe1{w}\pi_{e}^{-1}\{w\}, eSS1e\in S\subset S^{1}, where SS consists of those vectors making an angle at most 4545^{\circ} with the yy-axis. Now, (4.14) also holds in the following restricted form:

SπeμfL22𝑑1(e)=Xf()2𝑑𝔥()Xf()2𝑑𝔪().\int_{S}\|\pi_{e}\mu_{f}\|_{L^{2}}^{2}d\mathcal{H}^{1}(e)=\int_{\mathcal{L}_{\angle}}Xf(\ell)^{2}\,d\mathfrak{h}(\ell)\sim\int_{\mathcal{L}_{\angle}}Xf(\ell)^{2}\,d\mathfrak{m}(\ell). (4.15)

This equation will be useful in establishing Theorem 5.2. This will, formally, only prove Theorem 5.2 with "SS" in place of "S1S^{1}", but the original version is easy to deduce from this apparently weaker version.

4.2. Ball-plate duality

Recall from (4.8) the definition of the (dual) line set (P)\ell^{\ast}(P) for PP\subset\mathbb{H}. What does (B(p,r))\ell^{\ast}(B_{\mathbb{H}}(p,r)) look like? The answer is: a plate tangent to the cone 𝒞\mathcal{C}. Informally speaking, for r(0,12]r\in(0,\tfrac{1}{2}], an rr-plate tangent to 𝒞\mathcal{C} is a rectangle of dimensions (1×r×r2)\sim(1\times r\times r^{2}) whose long side is parallel to a light ray, and whose orientation is such that the plate is roughly tangent to 𝒞\mathcal{C}, see Figure 1. To prove rigorously that (B(p,r))\ell^{\ast}(B_{\mathbb{H}}(p,r)) looks like such a plate (inside B(1)B(1)), we need to be more precise with the definitions.

Recall that the cone 𝒞\mathcal{C} is a rotation of the "standard" cone 𝒞0={(x,y,z):z2=x2+y2}\mathcal{C}_{0}=\{(x,y,z):z^{2}=x^{2}+y^{2}\}.

\begin{overpic}[scale={1}]{Image1.pdf} \put(90.0,40.0){$x$} \put(-1.0,20.0){$y$} \put(45.0,88.0){$z$} \put(86.0,70.0){$\mathbb{P}=\{(1,-y,\tfrac{y^{2}}{2}):y\in\mathbb{R}\}$} \put(51.0,72.0){$\mathcal{C}$} \end{overpic}
Figure 1. The cone 𝒞\mathcal{C}, the parabola \mathbb{P}, and three rr-plates.

The intersection of 𝒞\mathcal{C} with the plane {x=1}\{x=1\} is the parabola

={(1,y,y2/2):y}.\mathbb{P}=\{(1,-y,y^{2}/2):y\in\mathbb{R}\}.

For every r(0,12]r\in(0,\tfrac{1}{2}] and pp\in\mathbb{P}, choose a rectangle =r(p)\mathcal{R}=\mathcal{R}_{r}(p) of dimensions r×r2r\times r^{2} in the plane {x=1}\{x=1\}, centred at pp, such that the longer rr-side is parallel to the tangent line of \mathbb{P} at pp. Then B(0,cr)\mathbb{P}\cap B(0,cr)\subset\mathcal{R} for an absolute constant c>0c>0. Now, the rr-plate centred at pp is the set obtained by sliding the rectangle \mathcal{R} along the light ray containing pp inside {|x|1}\{|x|\leq 1\}, see Figure 1. We make this even more formal in the next definition.

Definition 4.16 (rr-plate).

Let r(0,12]r\in(0,\tfrac{1}{2}], and let p=(1,y,y2/2)𝒞p=(1,-y,y^{2}/2)\in\mathbb{P}\subset\mathcal{C} with y[1,1]y\in[-1,1]. Let r(0):=[r,r]×[r2,r2]\mathcal{R}_{r}(0):=[-r,r]\times[-r^{2},r^{2}], and define r(y):=My(r(0))2\mathcal{R}_{r}(y):=M_{y}(\mathcal{R}_{r}(0))\subset\mathbb{R}^{2}, where

My=(10y1)M_{y}=\begin{pmatrix}1&0\\ -y&1\end{pmatrix}

(The rectangle r(y)\mathcal{R}_{r}(y) is the intersection of an rr-plate with the plane {x=0}\{x=0\}.) Define

𝒫r(p):={(0,r)+Ly([1,1]):rr(y)},\mathcal{P}_{r}(p):=\{(0,\vec{r})+L_{y}([-1,1]):\vec{r}\in\mathcal{R}_{r}(y)\},

The set 𝒫r(p)\mathcal{P}_{r}(p) is called the rr-plate centred at pp\in\mathbb{P}. In general, an rr-plate is any translate of one of the sets 𝒫r(p)\mathcal{P}_{r}(p), for p=(1,y,y2/2)p=(1,-y,y^{2}/2) with y[1,1]y\in[-1,1], and r(0,12]r\in(0,\tfrac{1}{2}].

For the rr-plate 𝒫r(p)\mathcal{P}_{r}(p), we also commonly use the notation 𝒫r(y)\mathcal{P}_{r}(y), where p=(1,y,y2/2)p=(1,-y,y^{2}/2).

Remark 4.17.

Since we require y[1,1]y\in[-1,1] in Definition 4.16, it is clear that an rr-plate contains, and is contained in, a rectangle of dimensions (1×r×r2)\sim(1\times r\times r^{2}). It is instructive to note that the number of "essentially distinct" rr-plates intersecting B(0,1)B(0,1) is roughly r4r^{-4}: to see this, take a maximal rr-separated subset of r\mathbb{P}_{r}\subset\mathbb{P}, and note that for each prp\in\mathbb{P}_{r}, the plate 𝒫r(p)\mathcal{P}_{r}(p) has volume r3r^{3}. Therefore it takes r3\sim r^{-3} translates of 𝒫r(p)\mathcal{P}_{r}(p) to cover B(0,1)B(0,1). This r4r^{-4}-numerology already suggests that the various rr-plates might correspond to Heisenberg rr-balls via duality.

To relate the plates 𝒫r\mathcal{P}_{r} to Heisenberg balls, we define a slight modification of the plates 𝒫r\mathcal{P}_{r}. Whereas 𝒫r\mathcal{P}_{r} is a union of (truncated) light rays in one fixed direction, the following "modified" plates contain full light rays in an rr-arc of directions. These "modified" plates will finally match the duals of Heisenberg balls, see Proposition 4.22.

Definition 4.18 (Modified rr-plate).

Let r(0,12]r\in(0,\tfrac{1}{2}] and y[1,1]y\in[-1,1]. Let r(y0)2\mathcal{R}_{r}(y_{0})\subset\mathbb{R}^{2} be the rectangle from Definition 4.16. For (u,v)2(u,v)\in\mathbb{R}^{2}, define the modified rr-plate

Πr(u,v,y):=(0,u,v)+{(0,r)+Ly:rr(y) and |yy|r}.\Pi_{r}(u,v,y):=(0,u,v)+\{(0,\vec{r})+L_{y^{\prime}}:\vec{r}\in\mathcal{R}_{r}(y)\text{ and }|y^{\prime}-y|\leq r\}. (4.19)
Remark 4.20.

The relation between the sets 𝒫r\mathcal{P}_{r} and Πr\Pi_{r} is that the following holds for some absolute constant c>0c>0: if r(0,12]r\in(0,\tfrac{1}{2}], y[1,1]y\in[-1,1], and u,vu,v\in\mathbb{R}, then

Πcr(u,v,y){(s,y,z):|s|2}(0,u,v)+𝒫r(y)Πr(u,v,y).\Pi_{cr}(u,v,y)\cap\{(s,y,z):|s|\leq 2\}\subset(0,u,v)+\mathcal{P}_{r}(y)\subset\Pi_{r}(u,v,y). (4.21)

(The constant "22" is arbitrary, but happens to be the one we need.) To see this, it suffices to check the case u=0=vu=0=v. Consider the "slices" of Πr(0,0,y)\Pi_{r}(0,0,y) and 𝒫r(y)\mathcal{P}_{r}(y) with a fixed plane {x=s}\{x=s\} for |s|1|s|\leq 1. If s=0s=0, both slices coincide with the rectangle y(y)\mathcal{R}_{y}(y). If 0<|s|10<|s|\leq 1, the slice Πr(0,0,y){x=s}\Pi_{r}(0,0,y)\cap\{x=s\} can be written as a sum

Πr(0,0,y){x=s}=r(y)+{Ly(s):|yy|r},\Pi_{r}(0,0,y)\cap\{x=s\}=\mathcal{R}_{r}(y)+\{L_{y^{\prime}}(s):|y-y^{\prime}|\leq r\},

whereas 𝒫r(y){x=s}=r(y)+{Ly(s)}\mathcal{P}_{r}(y)\cap\{x=s\}=\mathcal{R}_{r}(y)+\{L_{y}(s)\}. The relationship between these two slices is depicted in Figure 2.

\begin{overpic}[scale={0.9}]{Image2.pdf} \end{overpic}
Figure 2. The red box is the slice 𝒫r(y){x=s}\mathcal{P}_{r}(y)\cap\{x=s\}. The slice Πr(0,0,y){x=s}\Pi_{r}(0,0,y)\cap\{x=s\} is a union of the yellow boxes centred along the black curve {Ly(s):|yy|r}\{L_{y^{\prime}}(s):|y^{\prime}-y|\leq r\}. All the boxes individually are translates of r(y)\mathcal{R}_{r}(y).

After this, we leave it to the reader to verify that Πcr(0,0,y){x=s}𝒫r(y){x=s}\Pi_{cr}(0,0,y)\cap\{x=s\}\subset\mathcal{P}_{r}(y)\cap\{x=s\} if c>0c>0 is sufficiently small, and for |s|2|s|\leq 2.

We record the following consequence of (4.21): Πr(u,v,y){(s,y,z):|s|1}\Pi_{r}(u,v,y)\cap\{(s,y,z):|s|\leq 1\} is contained in a tube of width rr around the line (0,u,v)+Ly(0,u,v)+L_{y}. This is because 𝒫r(y)\mathcal{P}_{r}(y) is obviously contained in a tube of width r\sim r around LyL_{y} (this is a very non-sharp statement, using only that the longer side of y(r)\mathcal{R}_{y}(r) has length rr.)

We then show that the \ell^{\ast}-duals of Heisenberg balls are essentially modified plates:

Proposition 4.22.

Let p=(u0,0,v0)(0,y0,0)p=(u_{0},0,v_{0})\cdot(0,y_{0},0), r(0,12]r\in(0,\tfrac{1}{2}], and B:=B(p,r)B:=B_{\mathbb{H}}(p,r). Then,

(B)Π2r(u0,v0,y0)(CB),\ell^{\ast}(B)\subset\Pi_{2r}(u_{0},v_{0},y_{0})\subset\ell^{\ast}(CB), (4.23)

where C>0C>0 is an absolute constant, and CB=B(p,Cr)CB=B_{\mathbb{H}}(p,Cr).

Remark 4.24.

To build a geometric intuition, it will be helpful to notice the following. The yy-coordinate of the point p=(u0,0,v0)(0,y0,0)=(u0,y0,v0+12u0y0)p=(u_{0},0,v_{0})\cdot(0,y_{0},0)=(u_{0},y_{0},v_{0}+\tfrac{1}{2}u_{0}y_{0}) is "y0y_{0}". On the other hand, while the modified plate Π2r(u0,v0,y0)\Pi_{2r}(u_{0},v_{0},y_{0}) contains many lines, they are all "close" to the "central" line (0,u0,v0)+Ly0(0,u_{0},v_{0})+L_{y_{0}} (see Definition 4.19). According to the inclusions in (4.23), this means that the "direction" Ly0L_{y_{0}} of the modified plate containing the dual (B(p,r))\ell^{\ast}(B(p,r)) is determined by the yy-coordinate of pp. Even less formally: Heisenberg balls whose centres have the same yy-coordinate are dual to parallel plates.

Proof of Proposition 4.22.

To prove the inclusion (B)Π2r(u0,v0,y0)\ell^{\ast}(B)\subset\Pi_{2r}(u_{0},v_{0},y_{0}), let qB(p,r)q\in B_{\mathbb{H}}(p,r), and write q:=(u,0,v)(0,y,0)q:=(u,0,v)\cdot(0,y,0) with (u,v)2(u,v)\in\mathbb{R}^{2} and yy\in\mathbb{R}. First, we note that

|yy0|d(p,q)r.|y-y_{0}|\leq d_{\mathbb{H}}(p,q)\leq r. (4.25)

Let πxt\pi_{xt} be the vertical projection to the xtxt-plane {(u,0,v):u,v}\{(u^{\prime},0,v^{\prime}):u^{\prime},v^{\prime}\in\mathbb{R}\}. Then (u,0,v)=πxt(q)πxt(B)(u,0,v)=\pi_{xt}(q)\in\pi_{xt}(B) by the definition of πxt\pi_{xt}. We now observe that B=(u0,0,v0)B((0,y0,0),r)B=(u_{0},0,v_{0})\cdot B_{\mathbb{H}}((0,y_{0},0),r), so

πxt(B)=(u0,0,v0)+πxt(B((0,y0,0),r)).\pi_{xt}(B)=(u_{0},0,v_{0})+\pi_{xt}(B_{\mathbb{H}}((0,y_{0},0),r)).

We claim that

πxt(B((0,y0,0),r)){(u,0,v):(u,v)2r(y0)}.\pi_{xt}(B_{\mathbb{H}}((0,y_{0},0),r))\subset\{(u^{\prime},0,v^{\prime}):(u^{\prime},v^{\prime})\in\mathcal{R}_{2r}(y_{0})\}. (4.26)

This will prove that

(u,0,v)(u0,0,v0)+{(u,0,v):(u,v)2r(y0)}.(u,0,v)\in(u_{0},0,v_{0})+\{(u^{\prime},0,v^{\prime}):(u^{\prime},v^{\prime})\in\mathcal{R}_{2r}(y_{0})\}. (4.27)

Recalling the definition (4.19), a combination of (4.25) and (4.27) now shows that

(q)=((u,0,v)(0,y,0))=(4.10)(0,u,v)+LyΠ2r(u0,v0,y0).\ell^{\ast}(q)=\ell^{\ast}((u,0,v)\cdot(0,y,0))\stackrel{{\scriptstyle\eqref{form28}}}{{=}}(0,u,v)+L_{y}\subset\Pi_{2r}(u_{0},v_{0},y_{0}).

This will complete the proof of the inclusion (B)Π2r(u0,v0,y0)\ell^{\ast}(B)\subset\Pi_{2r}(u_{0},v_{0},y_{0}).

Let us then prove (4.26). Pick (x,y,t)B((0,y0,0),r)(x,y,t)\in B_{\mathbb{H}}((0,y_{0},0),r). Then,

(x,yy0,t+12xy0)=d((x,y,t),(0,y0,0))r,\|(x,y-y_{0},t+\tfrac{1}{2}xy_{0})\|=d_{\mathbb{H}}((x,y,t),(0,y_{0},0))\leq r,

so

|x|r,|yy0|r,and|t+12xy0|r2.|x|\leq r,\quad|y-y_{0}|\leq r,\quad\text{and}\quad|t+\tfrac{1}{2}xy_{0}|\leq r^{2}. (4.28)

Now, to prove (4.26), recall that πxt(x,y,t)=(x,0,t12xy)\pi_{xt}(x,y,t)=(x,0,t-\tfrac{1}{2}xy). Thus, we need to show that (x,t12xy)2r(y0)=My0(2r(0))(x,t-\tfrac{1}{2}xy)\in\mathcal{R}_{2r}(y_{0})=M_{y_{0}}(\mathcal{R}_{2r}(0)). Equivalently, My01(x,t12xy)2r(0)M_{y_{0}}^{-1}(x,t-\tfrac{1}{2}xy)\in\mathcal{R}_{2r}(0). Recalling the definition of MyM_{y}, one checks that

My01(x,t12xy)\displaystyle M_{y_{0}}^{-1}(x,t-\tfrac{1}{2}xy) =(10y01)(x,t12xy)\displaystyle=\begin{pmatrix}1&0\\ y_{0}&1\end{pmatrix}(x,t-\tfrac{1}{2}xy)
=(x,xy0+t12xy)\displaystyle=(x,xy_{0}+t-\tfrac{1}{2}xy)
=(x,t+12xy0+12x(y0y)).\displaystyle=(x,t+\tfrac{1}{2}xy_{0}+\tfrac{1}{2}x(y_{0}-y)).

Using (4.28), we finally note that the point on the right lies in the parabolic rectangle 2r(0)\mathcal{R}_{2r}(0). This concludes the proof of (4.26).

Let us then prove the inclusion Πr(u0,v0,y0)(CB)\Pi_{r}(u_{0},v_{0},y_{0})\subset\ell^{\ast}(CB). The set Πr(u0,v0,y0)\Pi_{r}(u_{0},v_{0},y_{0}) is a union of the lines (0,u0,v0)+(0,r)+Ly(0,u_{0},v_{0})+(0,\vec{r})+L_{y}, where rr(y0)\vec{r}\in\mathcal{R}_{r}(y_{0}) and |yy0|r|y-y_{0}|\leq r. We need to show that every such line can be realised as (q)\ell^{\ast}(q) for some qB(p,Cr)q\in B_{\mathbb{H}}(p,Cr). In this task, we are aided by the formula

((u,0,v)(0,y,0))=(0,u,v)+Ly\ell^{\ast}((u,0,v)\cdot(0,y,0))=(0,u,v)+L_{y}

observed in (4.7). This formula shows that we need to define q:=(u,0,v)(0,y,0)q:=(u,0,v)\cdot(0,y,0), where (u,v):=(u0,v0)+r(u,v):=(u_{0},v_{0})+\vec{r}, and yy is as in "LyL_{y}". Then we just have to hope that qB(p,Cr)q\in B_{\mathbb{H}}(p,Cr).

Recalling that p=(u0,0,v0)(0,y0,0)p=(u_{0},0,v_{0})\cdot(0,y_{0},0), one can check by direct computation that

d(p,q)=(u0u,y0y,v0v+y0(u0u)+12(uu0)(y0y).d_{\mathbb{H}}(p,q)=\|(u_{0}-u,y_{0}-y,v_{0}-v+y_{0}(u_{0}-u)+\tfrac{1}{2}(u-u_{0})(y_{0}-y)\|. (4.29)

On the other hand, one may easily check that (u,v)(u0,v0)+r(y0)(u,v)\in(u_{0},v_{0})+\mathcal{R}_{r}(y_{0}) is equivalent to

(uu0,vv0+y0(uu0))r(0),(u-u_{0},v-v_{0}+y_{0}(u-u_{0}))\in\mathcal{R}_{r}(0),

which implies |uu0|r|u-u_{0}|\leq r and |vv0+y0(uu0)|r2|v-v_{0}+y_{0}(u-u_{0})|\leq r^{2}. Since moreover |yy0|r|y-y_{0}|\leq r by assumption, it follows from (4.29) and the definition of the norm \|\cdot\| that d(p,q)rd_{\mathbb{H}}(p,q)\lesssim r. This completes the proof. ∎

We close the section with two additional auxiliary results:

Proposition 4.30.

Let p,qp,q\in\mathbb{H} and r(0,12]r\in(0,\tfrac{1}{2}], and assume that p1/10\|p\|\leq 1/10. Assume moreover that (p)B(1)(B(q,r))\ell^{\ast}(p)\cap B(1)\subset\ell^{\ast}(B_{\mathbb{H}}(q,r)). Then pB(q,Cr)p\in B_{\mathbb{H}}(q,Cr) for some absolute constant C>0C>0.

Proof.

Write p=(u,0,v)(0,y,0)p=(u,0,v)\cdot(0,y,0), so that (p)=(0,u,v)+Ly\ell^{\ast}(p)=(0,u,v)+L_{y}. Since p1/10\|p\|\leq 1/10, in particular |u|+|v|1/5|u|+|v|\leq 1/5. By the previous proposition, we already know that

[(0,u,v)+Ly]B(1)=(p)B(1)Π2r(u0,v0,y0),[(0,u,v)+L_{y}]\cap B(1)=\ell^{\ast}(p)\cap B(1)\subset\Pi_{2r}(u_{0},v_{0},y_{0}),

where we have written q=(u0,0,v0)(0,y0,0)q=(u_{0},0,v_{0})\cdot(0,y_{0},0). Since (0,u,v)B(1)(0,u,v)\in B(1), we know that (0,u,v)(p)Π2r(u0,v0,y0)(0,u,v)\in\ell^{\ast}(p)\cap\Pi_{2r}(u_{0},v_{0},y_{0}). But

Π2r(u0,v0,y0){x=0}={(0,u,v):(u,v)(u0,v0)+y0(r)},\Pi_{2r}(u_{0},v_{0},y_{0})\cap\{x=0\}=\{(0,u^{\prime},v^{\prime}):(u^{\prime},v^{\prime})\in(u_{0},v_{0})+\mathcal{R}_{y_{0}}(r)\},

so we may deduce that

(u,v)(u0,v0)+y0(r).(u,v)\in(u_{0},v_{0})+\mathcal{R}_{y_{0}}(r). (4.31)

Moreover, in Remark 4.20 we noted that Π2r(u0,v0,y0)B(1)\Pi_{2r}(u_{0},v_{0},y_{0})\cap B(1) is contained in the r\sim r-neighbourhood TT of the line (0,u0,v0)+Ly0(0,u_{0},v_{0})+L_{y_{0}}. Therefore also (0,u,v)+LyB(1)T(0,u,v)+L_{y}\cap B(1)\subset T. This implies that (Ly,Ly0)r\angle(L_{y},L_{y_{0}})\lesssim r, and hence |yy0|r|y-y_{0}|\lesssim r.

Now, we want to use (4.31) and |yy0|r|y-y_{0}|\lesssim r to deduce that d(p,q)rd_{\mathbb{H}}(p,q)\lesssim r. We first expand

d(p,q)=(u0u,y0y,v0v+y0(u0u)+12(uu0)(y0y).d_{\mathbb{H}}(p,q)=\|(u_{0}-u,y_{0}-y,v_{0}-v+y_{0}(u_{0}-u)+\tfrac{1}{2}(u-u_{0})(y_{0}-y)\|. (4.32)

Then, using the definition of y0(r)=My(0(r))\mathcal{R}_{y_{0}}(r)=M_{y}(\mathcal{R}_{0}(r)), we note that (4.31) is equivalent to

(uu0,vv0+y0(uu0))r(0).(u-u_{0},v-v_{0}+y_{0}(u-u_{0}))\in\mathcal{R}_{r}(0).

Combined with |yy0|r|y-y_{0}|\lesssim r, and recalling the definition of \|\cdot\|, this shows that the right hand side of (4.32) is bounded by r\lesssim r, as claimed. ∎

We already noted in Remark 4.24 that the (modified) 2r2r-plates containing (B(p1,r))\ell^{\ast}(B(p_{1},r)) and (B(p2,r))\ell^{\ast}(B(p_{2},r)) have (almost) the same direction if the points p1,p2p_{1},p_{2} have (almost) the same yy-coordinate. In this case, if d(p1,p2)Crd_{\mathbb{H}}(p_{1},p_{2})\geq Cr, it is natural to expect that (B(p1,r))\ell^{\ast}(B(p_{1},r)) and (B(p2,r))\ell^{\ast}(B(p_{2},r)) are disjoint, at least inside B(1)B(1). The next lemma verifies this intuition.

Lemma 4.33.

Let p1=(u1,0,v1)(0,y1,0)B(1)p_{1}=(u_{1},0,v_{1})\cdot(0,y_{1},0)\in B_{\mathbb{H}}(1) and p2=(u2,0,v2)(0,y2,0)B(1)p_{2}=(u_{2},0,v_{2})\cdot(0,y_{2},0)\in B_{\mathbb{H}}(1) be points with the properties

|y1y2|rand(B(p1,r))(B(p2,r))B(1).|y_{1}-y_{2}|\leq r\quad\text{and}\quad\ell^{\ast}(B_{\mathbb{H}}(p_{1},r))\cap\ell^{\ast}(B_{\mathbb{H}}(p_{2},r))\cap B(1)\neq\emptyset. (4.34)

Then, d(p1,p2)rd_{\mathbb{H}}(p_{1},p_{2})\lesssim r.

Proof.

We may reduce to the case y1=y2y_{1}=y_{2} by the following argument. Start by choosing a point p2B(p2,r)p_{2}^{\prime}\in B_{\mathbb{H}}(p_{2},r) such that the yy-coordinate of p2p_{2}^{\prime} equals y1y_{1}. This is possible, because |y1y2|r|y_{1}-y_{2}|\leq r, and the projection of B(p2,r)B_{\mathbb{H}}(p_{2},r) to the xyxy-plane is a Euclidean disc of radius rr. Then, notice that B(p2,r)B(p2,2r)B_{\mathbb{H}}(p_{2},r)\subset B_{\mathbb{H}}(p_{2}^{\prime},2r), so

(B(p1,2r))(B(p2,2r))B(1).\ell^{\ast}(B_{\mathbb{H}}(p_{1},2r))\cap\ell^{\ast}(B_{\mathbb{H}}(p_{2}^{\prime},2r))\cap B(1)\neq\emptyset.

Now, if we have already proven the lemma in the case y1=y2y_{1}=y_{2} (and for "2r2r" in place of "rr"), it follows that d(p1,p2)rd_{\mathbb{H}}(p_{1},p_{2}^{\prime})\lesssim r, and finally d(p1,p2)d(p1,p2)+d(p2,p2)rd_{\mathbb{H}}(p_{1},p_{2})\leq d_{\mathbb{H}}(p_{1},p_{2}^{\prime})+d_{\mathbb{H}}(p_{2}^{\prime},p_{2})\lesssim r.

Let us then assume that y1=y2=yy_{1}=y_{2}=y. It follows from (4.34) and the first inclusion in Proposition 4.22 combined with the first inclusion in (4.21) that

((0,u1,v1)+𝒫Cr(y))((0,u2,v2)+𝒫Cr(y))((0,u_{1},v_{1})+\mathcal{P}_{Cr}(y))\cap((0,u_{2},v_{2})+\mathcal{P}_{Cr}(y))\neq\emptyset

for some absolute constant C>0C>0. Let "xx" be a point in the intersection, and (using the definition of 𝒫Cr(y)\mathcal{P}_{Cr}(y)), express xx in the two following ways:

(0,u1,v1)+(0,r1)+Ly(s)=x=(0,u2,v2)+(0,r2)+Ly(s),(0,u_{1},v_{1})+(0,\vec{r}_{1})+L_{y}(s)=x=(0,u_{2},v_{2})+(0,\vec{r}_{2})+L_{y}(s),

where r1Cr(y)=My(Cr(0))\vec{r_{1}}\in\mathcal{R}_{Cr}(y)=M_{y}(\mathcal{R}_{Cr}(0)) and r2My(r(0))\vec{r_{2}}\in M_{y}(\mathcal{R}_{r}(0)), and s[1,1]s\in[-1,1]. The terms Ly(s)L_{y}(s) conveniently cancel out, and we find that

(u1,v1)(u2,v2)=r2r1My(2Cr(0)),(u_{1},v_{1})-(u_{2},v_{2})=\vec{r_{2}}-\vec{r_{1}}\in M_{y}(\mathcal{R}_{2Cr}(0)),

or equivalently

(u1u2,v1v2+y(u1u2))=My1(u1u2,v1v2)2Cr(0).(u_{1}-u_{2},v_{1}-v_{2}+y(u_{1}-u_{2}))=M_{y}^{-1}(u_{1}-u_{2},v_{1}-v_{2})\in\mathcal{R}_{2Cr}(0). (4.35)

We have already computed in (4.32) that

d(p1,p2)=(u1u2,0,v1v2+y(u1u2),d_{\mathbb{H}}(p_{1},p_{2})=\|(u_{1}-u_{2},0,v_{1}-v_{2}+y(u_{1}-u_{2})\|,

and now it follows immediately from (4.35) that d(p1,p2)rd_{\mathbb{H}}(p_{1},p_{2})\lesssim r. ∎

5. Discretising Theorem 1.7

The purpose of this section is to reduce the proof of Theorem 1.7 to Theorem 5.2 which concerns (δ,3)(\delta,3)-sets. We start by defining these precisely:

Definition 5.1 ((δ,t,C)\delta,t,C)-set).

Let (X,d)(X,d) be a metric space, and let t0t\geq 0 and C,δ>0C,\delta>0. A non-empty bounded set PXP\subset X is called a (δ,t,C)(\delta,t,C)-set if

|PB(x,r)|δCrt|P|δ,xX,rδ.|P\cap B(x,r)|_{\delta}\leq Cr^{t}\cdot|P|_{\delta},\qquad x\in X,\,r\geq\delta.

Here |A|δ|A|_{\delta} is the smallest number of balls of radius δ\delta needed to cover AA. A family of sets \mathcal{B} (typically: disjoint δ\delta-balls) is called a (δ,t,C)(\delta,t,C)-set if P:=P:=\cup\mathcal{B} is a (δ,t,C)(\delta,t,C)-set.

If PP\subset\mathbb{H}, or 𝒫()\mathcal{B}\subset\mathcal{P}(\mathbb{H}), the (δ,t,C)(\delta,t,C)-set condition is always tested relative to the metric dd_{\mathbb{H}}. We then state a δ\delta-discretised version of Theorem 1.7 for sets of dimension 33:

Theorem 5.2.

For every η>0\eta>0, there exists ϵ>0\epsilon>0 and δ0>0\delta_{0}>0 such that the following holds for all δ(0,δ0]\delta\in(0,\delta_{0}]. Let \mathcal{B} be a non-empty (δ,3,δϵ)(\delta,3,\delta^{-\epsilon})-set of δ\delta-balls contained in B(1)B_{\mathbb{H}}(1), with δ\delta-separated centres. Let μ=μf\mu=\mu_{f} be the measure on \mathbb{H} with density

f:=(δ4||)1B𝟏B.f:=(\delta^{4}|\mathcal{B}|)^{-1}\sum_{B\in\mathcal{B}}\mathbf{1}_{B}. (5.3)

Then,

S1πeμL22𝑑1(e)δη.\int_{S^{1}}\|\pi_{e}\mu\|_{L^{2}}^{2}\,d\mathcal{H}^{1}(e)\leq\delta^{-\eta}.

The proof of Theorem 5.2 will be given in Section 6. Deducing Theorem 1.7 from Theorem 5.2 involves two steps. The first one, carried out in Section 7, is to reduce Theorem 1.7 to a δ\delta-discretised version, which concerns (δ,t)(\delta,t)-sets with all possible values t[0,3]t\in[0,3]. This statement is Theorem 5.11 below, a simplified version of which was stated as Theorem 1.10 in the introduction.

The second – and less standard – step, carried out in this section, is to deduce Theorem 5.11 from Theorem 5.2. Heuristically, Theorem 5.2 is nothing but the 33-dimensional case of Theorem 5.11 – although in this case the statement looks more quantitative. We therefore need to argue that if we already have Theorem 5.11 for sets of dimension 33, then we also have it for sets of dimension t[0,3]t\in[0,3]. The heuristic is simple: given a set KK\subset\mathbb{H} of dimension t[0,3]t\in[0,3], we start by "adding" (from the left) to KK another – random – set HH\subset\mathbb{H} of dimension 3t3-t. Then, we apply the 33-dimensional version of Theorem 5.11 to HKH\cdot K, and this gives the correct conclusion for KK. A crucial point is that Theorem 5.11 concerns the Lebesgue measure (not the dimension) of πe(K)\pi_{e}(K). This quantity is invariant under left translating KK. This allows us to control Leb(πe(HK))\mathrm{Leb}(\pi_{e}(H\cdot K)) in a useful way.

We turn to the details. To deduce Theorem 1.7 from Theorem 5.2, we need a corollary of Theorem 5.2, stated in Corollary 5.6, which concerns slightly more general measures than ones of the form μ=μf\mu=\mu_{f} (as in (5.3)):

Definition 5.4 (δ\delta-measure).

Let δ(0,1]\delta\in(0,1] and C>0C>0. A Borel measure μ\mu on \mathbb{H} is called a (δ,C)(\delta,C)-measure if μ\mu has a density with respect to Lebesgue measure, also denoted μ\mu, and the density satisfies

μ(x)Cμ(B(x,δ))Leb(B(x,δ)),x.\mu(x)\leq C\cdot\frac{\mu(B_{\mathbb{H}}(x,\delta))}{\mathrm{Leb}(B_{\mathbb{H}}(x,\delta))},\qquad x\in\mathbb{H}.

If the constant C>0C>0 irrelevant, a (δ,C)(\delta,C)-measure may also be called a δ\delta-measure.

We will use the following notion of δ\delta-truncated Riesz energy:

Isδ(μ):=dμ(x)dμ(y)d,δ(x,y)s,I_{s}^{\delta}(\mu):=\iint\frac{d\mu(x)\,d\mu(y)}{d_{\mathbb{H},\delta}(x,y)^{s}}, (5.5)

where μ\mu is a Radon measure, 0s40\leq s\leq 4, and d,δ(x,y):=max{d(x,y),δ}d_{\mathbb{H},\delta}(x,y):=\max\{d_{\mathbb{H}}(x,y),\delta\}.

Corollary 5.6.

For every η>0\eta>0, there exists δ0,ϵ0>0\delta_{0},\epsilon_{0}>0 such that the following holds for all δ(0,δ0]\delta\in(0,\delta_{0}] and ϵ(0,ϵ0]\epsilon\in(0,\epsilon_{0}]. Let μ\mu be a (δ,δϵ)(\delta,\delta^{-\epsilon})-probability measure on B(1)B_{\mathbb{H}}(1) with I3δ(μ)δϵI_{3}^{\delta}(\mu)\leq\delta^{-\epsilon}. Then, there exists a Borel set GG\subset\mathbb{H} such that μ(G)1δϵ0\mu(G)\geq 1-\delta^{\epsilon_{0}}, and

S1πe(μ|G)L22d1(e)δη.\int_{S^{1}}\|\pi_{e}(\mu|_{G})\|_{L^{2}}^{2}\,d\mathcal{H}^{1}(e)\leq\delta^{-\eta}. (5.7)
Proof.

Fix η>0\eta>0, ϵ(0,ϵ0]\epsilon\in(0,\epsilon_{0}], and δ(0,δ0]\delta\in(0,\delta_{0}]. The dependence of δ0,ϵ0\delta_{0},\epsilon_{0} on η\eta will eventually be determined by an application of Theorem 5.2, but we will require at least that ϵ0η\epsilon_{0}\leq\eta.

It follows from I3δ(μ)δϵI^{\delta}_{3}(\mu)\leq\delta^{-\epsilon} and Chebychev’s inequality that there exists a set G0G_{0}\subset\mathbb{H} of measure μ(G0)13δϵ0\mu(G_{0})\geq 1-3\delta^{\epsilon_{0}} such that μ(B(x,r))δϵϵ0r3δ2ϵ0r3\mu(B_{\mathbb{H}}(x,r))\lesssim\delta^{-\epsilon-\epsilon_{0}}r^{3}\leq\delta^{-2\epsilon_{0}}r^{3} for all xG0x\in G_{0} and rδr\geq\delta. Now, for dyadic rationals 0<αδ32ϵ0δ20<\alpha\lesssim\delta^{3-2\epsilon_{0}}\leq\delta^{2}, let

G0,α:={xG0:α2μ(B(x,δ))α}.G_{0,\alpha}:=\{x\in G_{0}:\tfrac{\alpha}{2}\leq\mu(B_{\mathbb{H}}(x,\delta))\leq\alpha\}.

We discard immediately the sets G0,αG_{0,\alpha} with αδ10\alpha\leq\delta^{10}: the union of these sets has measure δ5δϵ0\leq\delta^{5}\leq\delta^{\epsilon_{0}} for δ>0\delta>0 small enough, so μ(G1)12δϵ0\mu(G_{1})\geq 1-2\delta^{\epsilon_{0}}, where

G1:=G0αδ10G0,α.G_{1}:=G_{0}\,\setminus\,\bigcup_{\alpha\leq\delta^{10}}G_{0,\alpha}.

Now, G1G_{1} is covered by the sets G0,αG_{0,\alpha} with δ10αδ2\delta^{10}\leq\alpha\lesssim\delta^{2}, and the number of such sets is mlog(1/δ)m\lesssim\log(1/\delta). We let {α1,,αm}\{\alpha_{1},\ldots,\alpha_{m}\} be an enumeration of these values of "α\alpha", and we abbreviate Gj:=G0,αjG^{j}:=G_{0,\alpha_{j}}. We note that the union of the sets GjG^{j} with μ(Gj)δ2ϵ0\mu(G^{j})\leq\delta^{2\epsilon_{0}} has measure at most mδ2ϵ0δϵ0m\cdot\delta^{2\epsilon_{0}}\leq\delta^{\epsilon_{0}} (for δ>0\delta>0 small), so finally

G:=G1{Gj:1jm and μ(Gj)δ2ϵ0}G:=G_{1}\,\setminus\,\bigcup\{G^{j}:1\leq j\leq m\text{ and }\mu(G^{j})\leq\delta^{2\epsilon_{0}}\}

has measure μ(G)12δϵ0δϵ01δϵ0\mu(G)\geq 1-2\delta^{\epsilon_{0}}-\delta^{\epsilon_{0}}\geq 1-\delta^{\epsilon_{0}}. Moreover, GG is covered by the sets GjG^{j} with μ(Gj)δ2ϵ0\mu(G^{j})\geq\delta^{2\epsilon_{0}}. Re-indexing if necessary, we now assume that μ(Gj)δ2ϵ0\mu(G^{j})\geq\delta^{2\epsilon_{0}} for all 1jm1\leq j\leq m.

For 1jm1\leq j\leq m fixed, let j\mathcal{B}_{j} be a finitely overlapping (Vitali) cover of GjG^{j} by balls of radius δ\delta, centred at GjG^{j}. Using the facts GjG0G^{j}\subset G_{0} and μ(Gj)δ2ϵ0\mu(G^{j})\geq\delta^{2\epsilon_{0}}, and the uniform lower bound μ(B(x,δ))αj/2\mu(B_{\mathbb{H}}(x,\delta))\geq\alpha_{j}/2 for xGjx\in G^{j}, it is easy to check that each j\mathcal{B}_{j} is a (δ,3,δCϵ0)(\delta,3,\delta^{-C\epsilon_{0}})-set with

|j|αj1.|\mathcal{B}_{j}|\lesssim\alpha_{j}^{-1}. (5.8)

Thus, writing

fj:=(δ4|j|1)Bj𝟏Bandμj:=μfj,f_{j}:=(\delta^{4}|\mathcal{B}_{j}|^{-1})\sum_{B\in\mathcal{B}_{j}}\mathbf{1}_{B}\quad\text{and}\quad\mu_{j}:=\mu_{f_{j}},

and assuming that δ0,ϵ0>0\delta_{0},\epsilon_{0}>0 are sufficiently small in terms of η\eta, we may deduce from Theorem 5.2 that

S1πe(μj)L22𝑑1(e)δη,1jm.\int_{S^{1}}\|\pi_{e}(\mu_{j})\|_{L^{2}}^{2}d\mathcal{H}^{1}(e)\leq\delta^{-\eta},\qquad 1\leq j\leq m.

Finally, it follows from the (δ,δϵ)(\delta,\delta^{-\epsilon})-property of μ\mu that

μ(x)δϵμ(B(x,δ))δ4δϵαjδ4(5.8)δϵδ4|j|δϵμj(x),xGj.\mu(x)\lesssim\delta^{-\epsilon}\cdot\frac{\mu(B_{\mathbb{H}}(x,\delta))}{\delta^{4}}\leq\delta^{-\epsilon}\cdot\frac{\alpha_{j}}{\delta^{4}}\stackrel{{\scriptstyle\eqref{form32}}}{{\lesssim}}\frac{\delta^{-\epsilon}}{\delta^{4}|\mathcal{B}_{j}|}\leq\delta^{-\epsilon}\cdot\mu_{j}(x),\qquad x\in G^{j}.

Thus, also the density of πe(μ|Gj)\pi_{e}(\mu|_{G^{j}}) is bounded from above by the density of πe(μj)\pi_{e}(\mu_{j}):

S1πe(μ|G)L22d1(e)δϵj=1mS1πe(μj)L22d1(e)log(1/δ)δηϵδ3η.\int_{S^{1}}\|\pi_{e}(\mu|_{G})\|_{L^{2}}^{2}\,d\mathcal{H}^{1}(e)\lesssim\delta^{-\epsilon}\sum_{j=1}^{m}\int_{S^{1}}\|\pi_{e}(\mu_{j})\|_{L^{2}}^{2}\,d\mathcal{H}^{1}(e)\lesssim\log(1/\delta)\cdot\delta^{-\eta-\epsilon}\leq\delta^{-3\eta}.

This completes the proof of (5.7) (with "3η3\eta" in place of "η\eta"). ∎

The concrete δ\delta-measures we will consider have the form ημ\eta\ast_{\mathbb{H}}\mu, where μ=μf\mu=\mu_{f} has a density of the form (5.3) (these are almost trivially δ\delta-measures), and η\eta is a (discrete) probability measure. The notation ημ\eta\ast_{\mathbb{H}}\mu refers to the (non-commutative!) Heisenberg convolution of η\eta and μ\mu, that is, the push-forward of η×μ\eta\times\mu under the group product (p,q)pq(p,q)\mapsto p\cdot q. Let us verify that such measures ημ\eta\ast_{\mathbb{H}}\mu are also δ\delta-measures:

Lemma 5.9.

Let μ\mu be (δ,C)(\delta,C) measure, and let η\eta be an arbitrary Borel probability measure on \mathbb{H}. Then ημ\eta\ast_{\mathbb{H}}\mu is again a (δ,C)(\delta,C)-measure.

Proof.

Recall that a (δ,C)(\delta,C) measure is absolutely continuous by definition, so the notation "μ(p)\mu(p)" is well-defined for Lebesgue almost every pp\in\mathbb{H}. The following formulae are valid, and easy to check, for Lebesgue almost every pp\in\mathbb{H}:

(ημ)(p)=μ(q1p)𝑑η(q)(\eta\ast_{\mathbb{H}}\mu)(p)=\int\mu(q^{-1}\cdot p)\,d\eta(q)

and

(ημ)(B(p,r))Leb(B(p,r))=μ(B(q1p,r))Leb(B(p,r))𝑑η(q).\frac{(\eta\ast_{\mathbb{H}}\mu)(B_{\mathbb{H}}(p,r))}{\mathrm{Leb}(B_{\mathbb{H}}(p,r))}=\int\frac{\mu(B_{\mathbb{H}}(q^{-1}\cdot p,r))}{\mathrm{Leb}(B_{\mathbb{H}}(p,r))}\,d\eta(q). (5.10)

Now, if one applies the δ\delta-measure assumption to the formula on the left hand side, one obtains

(ημ)(p)Cμ(B(q1p,δ))Leb(B(q1p,δ))𝑑η(q).(\eta\ast_{\mathbb{H}}\mu)(p)\leq C\int\frac{\mu(B_{\mathbb{H}}(q^{-1}\cdot p,\delta))}{\mathrm{Leb}(B_{\mathbb{H}}(q^{-1}\cdot p,\delta))}\,d\eta(q).

Lebesgue measure is invariant under left translations, so

Leb(B(q1p,δ))=Leb(B(p,δ)).\mathrm{Leb}(B_{\mathbb{H}}(q^{-1}\cdot p,\delta))=\mathrm{Leb}(B_{\mathbb{H}}(p,\delta)).

Therefore, it follows from equation (5.10) that

(ημ)(p)C(ημ)(B(p,δ))Leb(B(p,δ))(\eta\ast_{\mathbb{H}}\mu)(p)\leq C\cdot\frac{(\eta\ast_{\mathbb{H}}\mu)(B_{\mathbb{H}}(p,\delta))}{\mathrm{Leb}(B_{\mathbb{H}}(p,\delta))}

for Lebesgue almost every pp\in\mathbb{H}. This is what we claimed. ∎

We are then ready to state and prove the δ\delta-discretised counterpart of Theorem 1.7.

Theorem 5.11.

Let 0s<t30\leq s<t\leq 3. Then, there exist ϵ,δ0>0\epsilon,\delta_{0}>0, depending only on s,ts,t, such that the following holds for all δ(0,δ0]\delta\in(0,\delta_{0}]. Let \mathcal{B}\neq\emptyset be a (δ,t,δϵ)(\delta,t,\delta^{-\epsilon}) set of δ\delta-balls with δ\delta-separated centres, all contained in B(1)B_{\mathbb{H}}(1), and let SS1S\subset S^{1} be a Borel set of length 1(S)δϵ\mathcal{H}^{1}(S)\geq\delta^{\epsilon}. Then, there exists eSe\in S such that the following holds: if \mathcal{B}^{\prime}\subset\mathcal{B} is any sub-family with ||δϵ|||\mathcal{B}^{\prime}|\geq\delta^{\epsilon}|\mathcal{B}|, then

Leb(πe())δ3s.\mathrm{Leb}(\pi_{e}(\cup\mathcal{B}^{\prime}))\geq\delta^{3-s}.

In particular, πe()\pi_{e}(\cup\mathcal{B}^{\prime}) cannot be covered by fewer than δs\delta^{-s} parabolic balls of radius δ\delta.

Proof.

To reach a contradiction, assume that there exists a (δ,t,δϵ)(\delta,t,\delta^{-\epsilon})-set \mathcal{B} of δ\delta-balls with δ\delta-separated centres, contained in B(1)B_{\mathbb{H}}(1), and violating the conclusion of Theorem 5.11: there exists s<ts<t, and for every eSe\in S (Borel subset of S1S^{1} of length 1(S)δϵ\mathcal{H}^{1}(S)\geq\delta^{\epsilon}), there exists a subset e\mathcal{B}_{e}\subset\mathcal{B} with |e|δϵ|||\mathcal{B}_{e}|\geq\delta^{\epsilon}|\mathcal{B}| with the property

Leb(πe(e))δ3s.\mathrm{Leb}(\pi_{e}(\cup\mathcal{B}_{e}))\leq\delta^{3-s}. (5.12)

We aim for a contradiction if ϵ,δ\epsilon,\delta are sufficiently small. We fix an auxiliary parameter 0<η<(ts)/20<\eta<(t-s)/2. Then, we apply Corollary 5.6 to find the constant ϵ0>0\epsilon_{0}>0 which depends only on η\eta. Finally, we will assume, presently, that ϵ<ϵ0/2\epsilon<\epsilon_{0}/2, and η+3ϵ<ts\eta+3\epsilon<t-s.

Let μ\mu be the uniformly distributed probability measure on \cup\mathcal{B}; in particular μ\mu is a δ\delta-measure (with absolute constant), and Itδ(μ)δϵI_{t}^{\delta}(\mu)\lessapprox\delta^{-\epsilon}. Apply Proposition A.1 to find a set HB(1)H\subset B_{\mathbb{H}}(1) of cardinality |H|δt3|H|\leq\delta^{t-3} such that I3δ(τμ)δϵI^{\delta}_{3}(\tau\ast_{\mathbb{H}}\mu)\lessapprox\delta^{-\epsilon}, where τ\tau is the uniformly distributed probability measure on HH. Write ν:=τμ\nu:=\tau\ast_{\mathbb{H}}\mu, so ν\nu is a δ\delta-probability measure by Lemma 5.9. Since ϵ<ϵ0/2\epsilon<\epsilon_{0}/2 and I3δ(ν)δϵI_{3}^{\delta}(\nu)\lessapprox\delta^{-\epsilon}, it follows form Corollary 5.6 that there exists a set GG\subset\mathbb{H} of measure ν(G)1δϵ0\nu(G)\geq 1-\delta^{\epsilon_{0}} such that

11(S)Sπe(ν|G)L22d1(e)11(S)S1πe(ν|G)L22d1(e)δηϵ.\frac{1}{\mathcal{H}^{1}(S)}\int_{S}\|\pi_{e}(\nu|_{G})\|_{L^{2}}^{2}\,d\mathcal{H}^{1}(e)\leq\frac{1}{\mathcal{H}^{1}(S)}\int_{S^{1}}\|\pi_{e}(\nu|_{G})\|_{L^{2}}^{2}\,d\mathcal{H}^{1}(e)\leq\delta^{-\eta-\epsilon}. (5.13)

Finally, write Be:=H(e)B_{e}:=H\cdot(\cup\mathcal{B}_{e}) for all eS1e\in S^{1}, and note that ν(Be)δϵ\nu(B_{e})\geq\delta^{\epsilon} for all eSe\in S (this is a consequence of the general inequality (μ1μ2)(AB)(μ1×μ2)(A×B)(\mu_{1}\ast_{\mathbb{H}}\mu_{2})(A\cdot B)\geq(\mu_{1}\times\mu_{2})(A\times B)). Consequently, also ν(GBe)ν(G)+ν(Be)1δϵδϵ0δϵ/2\nu(G\cap B_{e})\geq\nu(G)+\nu(B_{e})-1\geq\delta^{\epsilon}-\delta^{\epsilon_{0}}\geq\delta^{\epsilon}/2, using ϵ<ϵ0/2\epsilon<\epsilon_{0}/2. Therefore,

δ2ϵ/4πe(ν|GBe)L12Leb(πe(Be))πe(ν|G)L22,eS1,\delta^{2\epsilon}/4\leq\|\pi_{e}(\nu|_{G\cap B_{e}})\|_{L^{1}}^{2}\leq\mathrm{Leb}(\pi_{e}(B_{e}))\cdot\|\pi_{e}(\nu|_{G})\|_{L^{2}}^{2},\quad e\in S^{1},

using Cauchy-Schwarz, and it follows from (5.13) that Leb(πe(Be))δη+3ϵ\mathrm{Leb}(\pi_{e}(B_{e}))\gtrsim\delta^{\eta+3\epsilon} for at least one vector eSe\in S. On the other hand, note that Be=H(e)B_{e}=H\cdot(\cup\mathcal{B}_{e}) is a union of δt3\leq\delta^{t-3} left translates of e\cup\mathcal{B}_{e}, and recall from (2.1) that

Leb(πe(pB))=Leb(πe(B)),p,B.\mathrm{Leb}(\pi_{e}(p\cdot B))=\mathrm{Leb}(\pi_{e}(B)),\qquad p\in\mathbb{H},\,B\subset\mathbb{H}.

Therefore, we have the upper bound

Leb(πe(Be))=Leb(πe(H(e)))(5.12)δt3δ3s=δts,eS1.\mathrm{Leb}(\pi_{e}(B_{e}))=\mathrm{Leb}(\pi_{e}(H\cdot(\cup\mathcal{B}_{e})))\stackrel{{\scriptstyle\eqref{form36}}}{{\leq}}\delta^{t-3}\cdot\delta^{3-s}=\delta^{t-s},\qquad e\in S^{1}.

Since η+3ϵ<ts\eta+3\epsilon<t-s by assumption, the previous lower and upper bounds for Leb(πe(Be))\mathrm{Leb}(\pi_{e}(B_{e})) are not compatible for δ>0\delta>0 small enough. A contradiction has been reached. ∎

6. Kakeya estimate of Guth, Wang, and Zhang

The purpose of this section is to prove Theorem 5.2. This will be based on the duality between horizontal lines and light rays developed in Section 4, and an application of a (reverse) square function inequality for the cone, due to Guth, Wang, and Zhang [10]. To be precise, we will not need the full power of this "oscillatory" statement, but rather only a Kakeya inequality for plates in [10, Lemma 1.4]. To introduce the statement, we need to recap some of the terminology and notation in [10]. This discussion follows [10, Section 1], but we prefer a different scaling: more precisely, in our discussion the geometric objects (plates and rectangles) of [10] are dilated by "RR" on the frequency side and (consequently) by R1R^{-1} on the spatial side.

Fix R1R\geq 1, and let

Γ:=ΓR:=𝒞{R/2|ξ|R}.\Gamma:=\Gamma_{R}:=\mathcal{C}\cap\{R/2\leq|\xi|\leq R\}. (6.1)

Let Γ(1)\Gamma(1) be the 11-neighbourhood of Γ\Gamma, and let Θ:=ΘR\Theta:=\Theta_{R} be a finitely overlapping cover of Γ(1)\Gamma(1) by rectangles of dimensions R×R1/2×1R\times R^{1/2}\times 1, whose longest side is parallel to a light ray. The statements in [10] are not affected by the particular construction of Θ\Theta, but in our application, the relevant rectangles are translates of dual rectangles of the δ\delta-plates in Definition 4.16, with δ=R1/2\delta=R^{-1/2}. Indeed, δ\delta-plates are rectangles of dimensions δ2×δ×1\sim\delta^{2}\times\delta\times 1 tangent to 𝒞\mathcal{C}, so their dual rectangles are plates of dimensions R×R1/2×1\sim R\times R^{1/2}\times 1, also tangent to 𝒞\mathcal{C} (this is because 𝒞\mathcal{C} has opening angle π/2\pi/2, see Figure 3). For concreteness, we will use translated duals of R1/2R^{-1/2}-plates (as in Definition 4.16) to form the collection Θ\Theta.

For each θΘ\theta\in\Theta, let fθL2(3)f_{\theta}\in L^{2}(\mathbb{R}^{3}) be a function with sptf^θθ\operatorname{spt}\hat{f}_{\theta}\subset\theta, and consider the square function

Sf:=(θΘ|fθ|2)1/2.Sf:=\Big{(}\sum_{\theta\in\Theta}|f_{\theta}|^{2}\Big{)}^{1/2}.

Then, [10, Lemma 1.4] contains an inequality of the following form:

3|Sf|4R1/2s1d(τ)=sUUτLeb(U)1SUfL24.\int_{\mathbb{R}^{3}}|Sf|^{4}\lesssim\sum_{R^{-1/2}\leq s\leq 1}\sum_{d(\tau)=s}\sum_{U\|U_{\tau}}\mathrm{Leb}(U)^{-1}\|S_{U}f\|_{L^{2}}^{4}. (6.2)

To understand the meaning of the "partial" square functions SUS_{U} we need to introduce more terminology from [10]. Fix a dyadic number s[R1/2,1]s\in[R^{-1/2},1] (an "angular" parameter), and write R:=s2R[1,R]R^{\prime}:=s^{2}R\in[1,R]. The 11-neighbourhood of the truncated cone ΓR=𝒞{|ξ|R}\Gamma_{R^{\prime}}=\mathcal{C}\cap\{|\xi|\sim R^{\prime}\} can be covered by a finitely overlapping family ΘR\Theta_{R^{\prime}} of rectangles of dimensions

R×(R)1/2×1=s2R×sR1/2×1.R^{\prime}\times(R^{\prime})^{1/2}\times 1=s^{2}R\times sR^{1/2}\times 1.

(Here ΘR\Theta_{R} agrees with Θ\Theta, as defined above.) Consequently, the (R)1(R^{\prime})^{-1}-neighbourhood of ΓR\Gamma_{R} is covered by the rescaled rectangles

𝒯s:={s2θ:θΘR}\mathcal{T}_{s}:=\{s^{-2}\theta:\theta\in\Theta_{R^{\prime}}\}

of dimensions R×s1R1/2×s2R\times s^{-1}R^{1/2}\times s^{-2}. Note that the family 𝒯1\mathcal{T}_{1} coincides with ΘR\Theta_{R} (at least if it is defined appropriately), whereas 𝒯R1/2\mathcal{T}_{R^{-1/2}} consists of 1\sim 1 balls of radius RR. For every s[R1/2,1]s\in[R^{-1/2},1], the rectangles in 𝒯s\mathcal{T}_{s} are at least as large as those in ΘR\Theta_{R}, so we may assume that every θΘR\theta\in\Theta_{R} is contained in at least one rectangle τ𝒯s\tau\in\mathcal{T}_{s}.

For θΘR\theta\in\Theta_{R} and τ𝒯s\tau\in\mathcal{T}_{s}, let θ\theta^{\ast} and τ\tau^{\ast} be the dual rectangles of θ\theta and τ\tau (here the word "dual" refers to the common notion in Euclidean Fourier analysis, and not the duality in the sense of Proposition 4.22). Then both θ\theta^{\ast} and τ\tau^{\ast} are rectangles centred at the origin, with dimensions

R1×R1/2×1andR1×sR1/2×s2,R^{-1}\times R^{-1/2}\times 1\quad\text{and}\quad R^{-1}\times sR^{-1/2}\times s^{2},

respectively. The longest sides of both θ\theta^{\ast} and τ\tau^{\ast} remain parallel to a light ray on 𝒞\mathcal{C}: this is again the convenient property of the "standard" cone 𝒞\mathcal{C} with opening angle π/2\pi/2, see Figure 3. Of course, θ\theta^{\ast} is an R1/2R^{-1/2}-plate in the sense of Definition 4.16, since the elements θΘ\theta\in\Theta were defined as (translates of) duals or R1/2R^{-1/2}-plates.

\begin{overpic}[scale={0.7}]{Image3.pdf} \put(5.5,37.5){$\theta$} \put(-7.0,27.0){$\sim R$} \put(-7.0,41.0){$\sim R^{1/2}$} \put(6.5,45.0){$1$} \put(6.0,8.0){$\tfrac{R}{2}$} \put(50.0,3.0){$\theta^{\ast}$} \put(70.0,-2.0){$1$} \put(99.0,2.0){$\sim R^{-1/2}$} \end{overpic}
Figure 3. On the left: the truncated cone Γ\Gamma and one of the plates θ\theta. On the right: the cone 𝒞\mathcal{C} and the dual plate θ\theta^{\ast}.

The set τ\tau^{\ast} turns out to be (essentially) a dilate of an (s2R)1/2(s^{2}R)^{-1/2}-plate. For every τ𝒯s\tau\in\mathcal{T}_{s}, consider Uτ:=s2τU_{\tau}:=s^{-2}\tau^{\ast}, which is a rectangle of dimensions

s2R1×s1R1/2×1=(s2R)1×(s2R)1/2×1.s^{-2}R^{-1}\times s^{-1}R^{-1/2}\times 1=(s^{2}R)^{-1}\times(s^{2}R)^{-1/2}\times 1.

In particular, UτU_{\tau} is an (s2R)1/2(s^{2}R)^{-1/2}-plate, and hence larger than (or at least as large as) θ\theta^{\ast}: if θτ\theta\subset\tau, then every translate of θ\theta^{\ast} is contained in some translate of 10Uτ10U_{\tau}. We let 𝒰τ\mathcal{U}_{\tau} be a tiling of 3\mathbb{R}^{3} by rectangles parallel to UτU_{\tau}. Now we may finally define the "partial" square function SUfS_{U}f:

SUf:=(θτ|fθ|2)1/2𝟏U,U𝒰τ.S_{U}f:=\Big{(}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\Big{)}^{1/2}\cdot\mathbf{1}_{U},\qquad U\in\mathcal{U}_{\tau}. (6.3)

We have now explained the meaning of (6.2), except the sum over "d(τ)=sd(\tau)=s". In our notation, this means the same as summing over τ𝒯s\tau\in\mathcal{T}_{s}.

We are then prepared to prove Theorem 5.2.

Proof of Theorem 5.2.

Let δ(0,12]\delta\in(0,\tfrac{1}{2}], and let \mathcal{B} be a (δ,3,δϵ)(\delta,3,\delta^{-\epsilon})-set of δ\delta-balls with δ\delta-separated centres. In the statement of Theorem 5.2, it was assumed that B(1)\cup\mathcal{B}\subset B_{\mathbb{H}}(1), but for slight technical convenience we strengthen this (with no loss of generality) to B(c)\cup\mathcal{B}\subset B_{\mathbb{H}}(c) for a small absolute constant c>0c>0. As in the statement of Theorem 5.2, let μ\mu be the measure on \mathbb{H} with density

f:=(δ4||)1B𝟏B.f:=(\delta^{4}|\mathcal{B}|)^{-1}\sum_{B\in\mathcal{B}}\mathbf{1}_{B}.

Following the discussion Section 4.1, and in particular recalling equation (4.15), Theorem 5.2 will be proven if we manage to establish that

Xf()2𝑑𝔪()δη,\int_{\mathcal{L}_{\angle}}Xf(\ell)^{2}\,d\mathfrak{m}(\ell)\leq\delta^{-\eta}, (6.4)

assuming that ϵ,δ>0\epsilon,\delta>0 are small enough, depending on η\eta. Recall that =({(a,b,c):|a|1})\mathcal{L}_{\angle}=\ell(\{(a,b,c):|a|\leq 1\}). To estimate the quantity in (6.4), notice first that

Xf()=f𝑑1(δ3||)1|{B:B}|,,Xf(\ell)=\int_{\ell}f\,d\mathcal{H}^{1}\lesssim(\delta^{3}|\mathcal{B}|)^{-1}\cdot|\{B\in\mathcal{B}:\ell\cap B\neq\emptyset\}|,\qquad\ell\in\mathcal{L}_{\angle}, (6.5)

because 1(B)δ\mathcal{H}^{1}(B\cap\ell)\lesssim\delta for all BB\in\mathcal{B}. Write N():=|{B:B}|N(\ell):=|\{B\in\mathcal{B}:\ell\cap B\neq\emptyset\}|. Then, as we just saw,

Xf()2𝑑𝔪()\displaystyle\int_{\mathcal{L}_{\angle}}Xf(\ell)^{2}\,d\mathfrak{m}(\ell) (δ3||)2N()2𝑑𝔪()\displaystyle\lesssim(\delta^{3}|\mathcal{B}|)^{-2}\int_{\mathcal{L}_{\angle}}N(\ell)^{2}\,d\mathfrak{m}(\ell)
(δ3||)2B(2)N((p))2𝑑Leb(p).\displaystyle\leq(\delta^{3}|\mathcal{B}|)^{-2}\int_{B(2)}N(\ell(p))^{2}\,d\mathrm{Leb}(p).

The second inequality is based on (a) the definition of the measure 𝔪=Leb\mathfrak{m}=\ell_{\sharp}\mathrm{Leb}, and (b) the observation that if (p)\ell(p)\in\mathcal{L}_{\angle} and N((p))0N(\ell(p))\neq 0, then (p)B(c)\ell(p)\cap B_{\mathbb{H}}(c)\neq\emptyset, and this forces pB(2)p\in B(2) (if c>0c>0 was taken small enough). Finally, by Lemma 4.11, we have

N((p))|{B:p(B)}|=B𝟏(B)(p).N(\ell(p))\leq|\{B\in\mathcal{B}:p\in\ell^{\ast}(B)\}|=\sum_{B\in\mathcal{B}}\mathbf{1}_{\ell^{\ast}(B)}(p).

Indeed, whenever (p)B\ell(p)\cap B\neq\emptyset for some BB\in\mathcal{B}, there exists a point q(p)Bq\in\ell(p)\cap B, and then Lemma 4.11 implies that p(q)(B)p\in\ell^{\ast}(q)\subset\ell^{\ast}(B). Therefore, combining (6.4)-(6.5), it will suffice to show that for η>0\eta>0 fixed, the inequality

B(2)(B𝟏(B))2δη(δ3||)2\int_{B(2)}\Big{(}\sum_{B\in\mathcal{B}}\mathbf{1}_{\ell^{\ast}(B)}\Big{)}^{2}\leq\delta^{-\eta}\cdot(\delta^{3}|\mathcal{B}|)^{2} (6.6)

holds assuming that we have picked ϵ>0\epsilon>0 (in the (δ,3,δϵ)(\delta,3,\delta^{-\epsilon})-set hypothesis for \mathcal{B}) sufficiently small, depending on η\eta. We formulate a slightly more general version of this inequality in Proposition 6.7 below, and then explain in the remark afterwards why (6.6) is a consequence. This completes the proof of Theorem 5.2. ∎

Proposition 6.7.

For every ϵ>0\epsilon>0, there exists δ0>0\delta_{0}>0 such that the following holds for all δ(0,δ0]\delta\in(0,\delta_{0}]. Let \mathcal{B} be a family of δ\delta-balls contained in B(1)B_{\mathbb{H}}(1) with δ\delta-separated centres, and satisfying the following non-concentration condition for some 𝐂>0\mathbf{C}>0:

|{B:BB(p,r)}|𝐂(rδ)3,p,rδ.|\{B\in\mathcal{B}:B\subset B_{\mathbb{H}}(p,r)\}|\leq\mathbf{C}\cdot\left(\frac{r}{\delta}\right)^{3},\qquad p\in\mathbb{H},\,r\geq\delta. (6.8)

Then,

B(2)(B𝟏(B))2𝐂δ3ϵ||.\int_{B(2)}\Big{(}\sum_{B\in\mathcal{B}}\mathbf{1}_{\ell^{\ast}(B)}\Big{)}^{2}\leq\mathbf{C}\cdot\delta^{3-\epsilon}|\mathcal{B}|. (6.9)
Remark 6.10.

Why is (6.6) a consequence of (6.9)? In (6.6), we assumed that \mathcal{B} is a (δ,3,δϵ)(\delta,3,\delta^{-\epsilon})-set. This implies

|{B:BB(p,r)}|δϵr3||,p,rδ.|\{B\in\mathcal{B}:B\subset B_{\mathbb{H}}(p,r)\}|\lesssim\delta^{-\epsilon}\cdot r^{3}|\mathcal{B}|,\qquad p\in\mathbb{H},\,r\geq\delta.

Therefore, (6.8) is satisfied with constant 𝐂δ3ϵ||\mathbf{C}\sim\delta^{3-\epsilon}|\mathcal{B}|. Hence (6.9) implies (6.6) if we choose ϵ<η/2\epsilon<\eta/2 and then δ>0\delta>0 sufficiently small.

We chose to formulate Proposition 6.7 separately because the "meaning" of (6.9) is easier to appreciate than that of (6.6): namely, if all the sets (B)\ell^{\ast}(B) had a disjoint intersection inside B(1)B(1), then the left hand side of (6.9) would be roughly δ3||\delta^{3}|\mathcal{B}|. Thus, (6.9) tells us that under the non-concentration condition (6.8), the sets (B)\ell^{\ast}(B) are nearly disjoint inside B(1)B(1), at least at the level of L2L^{2}-norms.

Proof of Proposition 6.7.

By the discussion in Section 4.2, the intersections (B)B(2)\ell^{\ast}(B)\cap B(2) are essentially δ\delta-plates – rectangles of dimensions 1×δ×δ21\times\delta\times\delta^{2} tangent to 𝒞\mathcal{C}. More precisely, for every BB\in\mathcal{B}, let 𝒫B3\mathcal{P}_{B}\subset\mathbb{R}^{3} be a CδC\delta-plate (as in Definition 4.16) with the property

(B)B(2)𝒫B.\ell^{\ast}(B)\cap B(2)\subset\mathcal{P}_{B}.

This is possible by first applying Proposition 4.22 (which yields a modified 2δ2\delta-plate containing (B)\ell^{\ast}(B)), and then the first inclusion in (4.21), which shows that the intersection of the modified 2δ2\delta-plate with B(2)B(2) is contained in a CδC\delta-plate 𝒫B\mathcal{P}_{B}. Now, we will prove (6.9) by establishing that

(B𝟏𝒫B)2𝐂δ3ϵ||.\int\Big{(}\sum_{B\in\mathcal{B}}\mathbf{1}_{\mathcal{P}_{B}}\Big{)}^{2}\leq\mathbf{C}\cdot\delta^{3-\epsilon}|\mathcal{B}|. (6.11)

Every plate 𝒫B\mathcal{P}_{B} has a direction, denoted θ(𝒫B)\theta(\mathcal{P}_{B}): this is the direction of the longest axis of 𝒫B\mathcal{P}_{B}, or more formally the real number "y[1,1]y\in[-1,1]" associated to the line "LyL_{y}" in Definitions 4.16. By enlarging the plates 𝒫B\mathcal{P}_{B} slightly (if necessary), we may assume that their directions lie in the set Θ:=(δ)[1,1]\Theta:=(\delta\mathbb{Z})\cap[-1,1]: this is because if two plates coincide in all other parameters, and differ in direction by δ\leq\delta, both are contained in constant enlargements of the other (this is not hard to check). The reason why we may restrict attention to [1,1][-1,1] is that all the plates 𝒫B\mathcal{P}_{B} were associated to the balls BB(1)B\subset B_{\mathbb{H}}(1), and in fact the yy-coordinate of the centre of BB determines the direction of 𝒫B\mathcal{P}_{B} (see (4.10)).

We next sort the family {𝒫B}B\{\mathcal{P}_{B}\}_{B\in\mathcal{B}} according to their directions:

{𝒫B:B}=:θΘ𝒫(θ),\{\mathcal{P}_{B}:B\in\mathcal{B}\}=:\bigcup_{\theta\in\Theta}\mathcal{P}(\theta),

where 𝒫(θ):={𝒫B:θ(𝒫B)=θ}\mathcal{P}(\theta):=\{\mathcal{P}_{B}:\theta(\mathcal{P}_{B})=\theta\}. Thus, for θΘ\theta\in\Theta fixed, the plates in 𝒫(θ)\mathcal{P}(\theta) are all translates of each other. Also, the plates in 𝒫(θ)\mathcal{P}(\theta) for a fixed θ\theta have bounded overlap: this follows from the assumption that the balls in \mathcal{B} have δ\delta-separated centres, and uses Lemma 4.33 (the plates with a fixed direction correspond precisely to Heisenberg balls whose yy-coordinates are, all, within "δ\delta" of each other).

Write R:=δ2R:=\delta^{-2}, thus δ=R1/2\delta=R^{-1/2}, and recall the truncated cone Γ=ΓR\Gamma=\Gamma_{R} from (6.1). Since the plates 𝒫𝒫(θ)\mathcal{P}\in\mathcal{P}(\theta) are translates of each other, they all have a common dual rectangle 𝒫θ\mathcal{P}^{\ast}_{\theta} of dimensions R×R1/2×1\sim R\times R^{1/2}\times 1. The rectangle 𝒫θ\mathcal{P}^{\ast}_{\theta} is centred at 0, but we may translate it by R\sim R in the direction of its longest RR-side (a light ray depending on θ\theta) so that the translate lies in the O(1)O(1)-neighbourhood of ΓR\Gamma_{R}. Committing a serious abuse of notation, we will denote this translated dual rectangle again by "θ\theta", and the collection of all these sets is denoted Θ\Theta. This notation coincides with the discussion below (6.1). There is a 11-to-11 correspondence between the directions θΘ=δ[1,1]\theta\in\Theta=\delta\mathbb{Z}\cap[-1,1] and the rectangles θΘ\theta\in\Theta defined just above, so the notational inconsistency should not cause confusion.

We gradually move towards applying the inequality (6.2) of Guth, Wang, and Zhang. The next task is to define the functions fθf_{\theta} and f=θΘfθf=\sum_{\theta\in\Theta}f_{\theta}. Fix θΘ\theta\in\Theta, 𝒫𝒫(θ)\mathcal{P}\in\mathcal{P}(\theta), and let φ𝒫𝒮(3)\varphi_{\mathcal{P}}\in\mathcal{S}(\mathbb{R}^{3}) be a non-negative Schwartz function with the properties

  1. (1)

    𝟏𝒫φ𝒫1\mathbf{1}_{\mathcal{P}}\leq\varphi_{\mathcal{P}}\lesssim 1,

  2. (2)

    φ𝒫\varphi_{\mathcal{P}} has rapid decay outside 𝒫\mathcal{P},

  3. (3)

    φ^𝒫𝒫θ\widehat{\varphi}_{\mathcal{P}}\subset\mathcal{P}^{\ast}_{\theta}.

Here "rapid decay outside 𝒫\mathcal{P} has" the usual meaning: if λ𝒫\lambda\mathcal{P} denotes a λ\lambda-times dilated, concentric, version of 𝒫\mathcal{P}, then φ(x)NλN\varphi(x)\lesssim_{N}\lambda^{-N} for all x3λ𝒫x\in\mathbb{R}^{3}\,\setminus\,\lambda\mathcal{P} (and for any NN\in\mathbb{N}). Then, define the function

fθ:=𝒫𝒫(θ)eθφ𝒫.f_{\theta}:=\sum_{\mathcal{P}\in\mathcal{P}(\theta)}e_{\theta}\cdot\varphi_{\mathcal{P}}.

Here eθe_{\theta} is a modulation, depending only on θ\theta, such that

eθφ𝒫^θ.\widehat{e_{\theta}\cdot\varphi_{\mathcal{P}}}\subset\theta.

Now the function f=θΘfθf=\sum_{\theta\in\Theta}f_{\theta} satisfies all the assumptions of the inequality (6.2), so

3(B𝟏𝒫B)2\displaystyle\int_{\mathbb{R}^{3}}\Big{(}\sum_{B\in\mathcal{B}}\mathbf{1}_{\mathcal{P}_{B}}\Big{)}^{2} =3(θΘ𝒫𝒫(θ)𝟏𝒫B)2\displaystyle=\int_{\mathbb{R}^{3}}\Big{(}\sum_{\theta\in\Theta}\sum_{\mathcal{P}\in\mathcal{P}(\theta)}\mathbf{1}_{\mathcal{P}_{B}}\Big{)}^{2}
3(θΘ|𝒫𝒫(θ)eθφ𝒫|2)2\displaystyle\leq\int_{\mathbb{R}^{3}}\Big{(}\sum_{\theta\in\Theta}\Big{|}\sum_{\mathcal{P}\in\mathcal{P}(\theta)}e_{\theta}\cdot\varphi_{\mathcal{P}}\Big{|}^{2}\Big{)}^{2}
=3|Sf|4R1/2s1d(τ)=sUUτLeb(U)1SUfL24.\displaystyle=\int_{\mathbb{R}^{3}}|Sf|^{4}\lesssim\sum_{R^{-1/2}\leq s\leq 1}\sum_{d(\tau)=s}\sum_{U\|U_{\tau}}\mathrm{Leb}(U)^{-1}\|S_{U}f\|_{L^{2}}^{4}. (6.12)

Recall the notation on the right hand side, in particular that δ=R1/2s1\delta=R^{-1/2}\leq s\leq 1 only runs over dyadic rationals, and the definition of the "partial" square function SUfS_{U}f from (6.3). The rectangles UU are Δ\Delta-plates with Δ=(s2R)1/2=s1δ\Delta=(s^{2}R)^{-1/2}=s^{-1}\delta. In particular, every UU is essentially the \ell^{\ast}-dual of a Heisenberg Δ\Delta-ball: this will allow us to control SUfL2\|S_{U}f\|_{L^{2}} by applying the non-concentration condition (6.8) between scales δ\delta and 11.

By definition,

SUf22=Uθτ|fθ|2=Uθτ(𝒫𝒫(θ)φ𝒫)2Uθτ𝒫𝒫(θ)φ𝒫.\|S_{U}f\|_{2}^{2}=\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}=\int_{U}\sum_{\theta\subset\tau}\Big{(}\sum_{\mathcal{P}\in\mathcal{P}(\theta)}\varphi_{\mathcal{P}}\big{)}^{2}\lessapprox\int_{U}\sum_{\theta\subset\mathcal{\tau}}\sum_{\mathcal{P}\in\mathcal{P}(\theta)}\varphi_{\mathcal{P}}. (6.13)

Above, and in the sequel, the notation ABA\lessapprox B means that for every ρ>0\rho>0, there exists a constant Cρ>0C_{\rho}>0 such that ACρδρBA\leq C_{\rho}\delta^{-\rho}B. In (6.13), the final "\lessapprox" inequality follows easily from the rapid decay of the functions φ𝒫\varphi_{\mathcal{P}}, and the bounded overlap of the plates 𝒫𝒫(θ)\mathcal{P}\in\mathcal{P}(\theta) for θΘ\theta\in\Theta fixed.

For θτ\theta\subset\tau, each plate 𝒫𝒫(θ)\mathcal{P}\in\mathcal{P}(\theta) is contained in some translate of 10Uτ10U_{\tau} (this was discussed above (6.3)), but this translate may not be UU. Let 𝐔U\mathbf{U}\supset U be an (RϵΔ)(R^{\epsilon}\Delta)-plate which is concentric with UU. We then decompose the right hand side of (6.13) as

Uθτ𝒫𝒫(θ)φ𝒫θτ𝒫𝒫(θ)𝒫𝐔φ𝒫+Uθτ𝒫𝒫(θ)𝒫𝐔φ𝒫.\int_{U}\sum_{\theta\subset\mathcal{\tau}}\sum_{\mathcal{P}\in\mathcal{P}(\theta)}\varphi_{\mathcal{P}}\leq\int\sum_{\theta\subset\tau}\mathop{\sum_{\mathcal{P}\in\mathcal{P}(\theta)}}_{\mathcal{P}\subset\mathbf{U}}\varphi_{\mathcal{P}}+\int_{U}\sum_{\theta\subset\tau}\mathop{\sum_{\mathcal{P}\in\mathcal{P}(\theta)}}_{\mathcal{P}\not\subset\mathbf{U}}\varphi_{\mathcal{P}}. (6.14)

Since each 𝒫𝒫(θ)\mathcal{P}\in\mathcal{P}(\theta) is contained in element of the tiling 𝒰τ\mathcal{U}_{\tau} (consisting of translates of UU) every plate 𝒫(θ)\mathcal{P}(\theta) with 𝒫𝐔\mathcal{P}\not\subset\mathbf{U} is far away from UU: more precisely, Rϵ/2𝒫U=R^{\epsilon/2}\mathcal{P}\cap U=\emptyset. By the rapid decay of φ𝒫\varphi_{\mathcal{P}} outside 𝒫\mathcal{P}, this implies that φ𝒫ϵδ100\varphi_{\mathcal{P}}\lesssim_{\epsilon}\delta^{100} on UU, and therefore the second term of (6.14) is bounded by, say, ϵδ50\lesssim_{\epsilon}\delta^{50}.

We then focus on the first term of (6.14), and we first note that

θτ𝒫𝒫(θ)𝒫𝐔φ𝒫δ3|{𝒫:𝒫𝐔}|,\int\sum_{\theta\subset\tau}\mathop{\sum_{\mathcal{P}\in\mathcal{P}(\theta)}}_{\mathcal{P}\subset\mathbf{U}}\varphi_{\mathcal{P}}\lesssim\delta^{3}\cdot|\{\mathcal{P}:\mathcal{P}\subset\mathbf{U}\}|, (6.15)

since φ𝒫L1Leb(𝒫)δ3\|\varphi_{\mathcal{P}}\|_{L^{1}}\sim\mathrm{Leb}(\mathcal{P})\sim\delta^{3}. So, we need to find out how many δ\delta-plates 𝒫\mathcal{P} are contained in 𝐔\mathbf{U}. Since 𝐔\mathbf{U} is an (RϵΔ)(R^{\epsilon}\Delta)-plate, it follows from the second inclusion (4.21), combined with the second inclusion in Proposition 4.22, that

𝐔(B(pU,CRϵΔ))=:(BU).\mathbf{U}\subset\ell^{\ast}(B_{\mathbb{H}}(p_{U},CR^{\epsilon}\Delta))=:\ell^{\ast}(B_{U}).

for some pUp_{U}\in\mathbb{H}, and for some absolute constant C>0C>0. On the other hand, the plates 𝒫=𝒫B\mathcal{P}=\mathcal{P}_{B}, BB\in\mathcal{B}, were initially chosen in such a way that (B){(s,y,z):|s|1}𝒫B\ell^{\ast}(B)\cap\{(s,y,z):|s|\leq 1\}\subset\mathcal{P}_{B}. Thus, whenever 𝒫B𝐔\mathcal{P}_{B}\subset\mathbf{U}, we have

(B){(s,y,z):|s|1}𝒫B𝐔(BU).\ell^{\ast}(B)\cap\{(s,y,z):|s|\leq 1\}\subset\mathcal{P}_{B}\subset\mathbf{U}\subset\ell^{\ast}(B_{U}).

This implies by Proposition 4.30 that BBUB\subset B_{U}, where possibly BUB_{U} was inflated by another constant factor. Thus,

|{𝒫:𝒫𝐔}||{B:BBU}|.|\{\mathcal{P}:\mathcal{P}\subset\mathbf{U}\}|\lesssim|\{B\in\mathcal{B}:B\subset B_{U}\}|.

Using (6.8), this will easily yield useful upper bounds for |{𝒫:𝒫𝐔}||\{\mathcal{P}:\mathcal{P}\subset\mathbf{U}\}|.

To make this precise, we sort the sets "UU" appearing in (6.12) according to the "richness"

ρ(U):=|{B:BBU}|(6.8)𝐂(CRϵΔδ)3.\rho(U):=|\{B\in\mathcal{B}:B\subset B_{U}\}|\stackrel{{\scriptstyle\eqref{form38a}}}{{\leq}}\mathbf{C}\cdot\left(\frac{CR^{\epsilon}\Delta}{\delta}\right)^{3}. (6.16)

For s[R1/2,1]s\in[R^{-1/2},1] fixed, we choose a (dyadic) value ρ=ρs\rho=\rho_{s} such that

d(τ)=sUUτLeb(U)1SUfL24d(τ)=sUUτρ(U)ρLeb(U)1SUfL24.\sum_{d(\tau)=s}\sum_{U\|U_{\tau}}\mathrm{Leb}(U)^{-1}\|S_{U}f\|_{L^{2}}^{4}\lessapprox\sum_{d(\tau)=s}\mathop{\sum_{U\|U_{\tau}}}_{\rho(U)\sim\rho}\mathrm{Leb}(U)^{-1}\|S_{U}f\|_{L^{2}}^{4}. (6.17)

Here "\lessapprox" hides a constant of the form Clog(1/δ)C\log(1/\delta). Let 𝒰(ρ)\mathcal{U}(\rho) be the collection of sets "UU" appearing on the right hand side, and let \mathcal{B}^{\prime}\subset\mathcal{B} be the subset of the original δ\delta-balls which are contained in some ball BUB_{U}, U𝒰(ρ)U\in\mathcal{U}(\rho). Then, evidently,

||ρ|𝒰(ρ)|RCϵ||.|\mathcal{B}^{\prime}|\lesssim\rho\cdot|\mathcal{U}(\rho)|\lesssim R^{C\epsilon}|\mathcal{B}^{\prime}|. (6.18)

The factor "RCϵR^{C\epsilon}" arises from the fact that while distinct sets "UU" are the duals of essentially disjoint Heisenberg Δ\Delta-balls, the inflated balls BUB_{U} only have bounded overlap, depending on the inflation factor RϵR^{\epsilon}.

Now, for U𝒰(ρ)U\in\mathcal{U}(\rho), we may estimate (6.15) as follows:

SUfL22ϵθτ𝒫𝒫(θ)𝒫𝐔φ𝒫δ3ρδ3RCϵ|||𝒰(ρ)|.\|S_{U}f\|_{L^{2}}^{2}\lessapprox_{\epsilon}\int\sum_{\theta\subset\tau}\mathop{\sum_{\mathcal{P}\in\mathcal{P}(\theta)}}_{\mathcal{P}\subset\mathbf{U}}\varphi_{\mathcal{P}}\lesssim\delta^{3}\cdot\rho\lesssim\delta^{3}\cdot R^{C\epsilon}\cdot\frac{|\mathcal{B}^{\prime}|}{|\mathcal{U}(\rho)|}.

(In this estimate, we have omitted the term "δ50\delta^{50}" from the second part of (6.14), because this term will soon turn out to be much smaller than the best bounds for what remains.) Plugging this estimate into (6.17), and observing that Leb(U)=Δ3\mathrm{Leb}(U)=\Delta^{3}, we obtain

d(τ)=sUUτLeb(U)1SUfL24\displaystyle\sum_{d(\tau)=s}\sum_{U\|U_{\tau}}\mathrm{Leb}(U)^{-1}\|S_{U}f\|_{L^{2}}^{4} ϵ|𝒰(ρ)|Δ3(δ3RCϵ|||𝒰(ρ)|)2\displaystyle\lessapprox_{\epsilon}|\mathcal{U}(\rho)|\cdot\Delta^{-3}\cdot\Big{(}\delta^{3}\cdot R^{C\epsilon}\cdot\frac{|\mathcal{B^{\prime}}|}{|\mathcal{U}(\rho)|}\Big{)}^{2}
=Δ3δ6R2Cϵ||2|𝒰(ρ)|\displaystyle=\Delta^{-3}\cdot\delta^{6}\cdot R^{2C\epsilon}\cdot\frac{|\mathcal{B}^{\prime}|^{2}}{|\mathcal{U}(\rho)|}
(6.16)&(6.18)𝐂R3Cϵδ3||.\displaystyle\stackrel{{\scriptstyle\eqref{form19}\&\eqref{form17}}}{{\lesssim}}\mathbf{C}\cdot R^{3C\epsilon}\cdot\delta^{3}|\mathcal{B}|.

Notably, this estimate is independent of "Δ\Delta" and the parameter "ss", so we may finally deduce from (6.12) that

3(B𝟏𝒫B)2ϵ𝐂R3Cϵδ3||.\int_{\mathbb{R}^{3}}\Big{(}\sum_{B\in\mathcal{B}}\mathbf{1}_{\mathcal{P}_{B}}\Big{)}^{2}\lessapprox_{\epsilon}\mathbf{C}\cdot R^{3C\epsilon}\cdot\delta^{3}|\mathcal{B}|.

Since R=δ2R=\delta^{-2} and ϵ>0\epsilon>0 was arbitrary, this implies (6.9) by renaming variables, and the proof of Proposition 6.7 is complete. ∎

7. Proof of Theorem 1.7

We recall the statement:

Theorem 7.1.

Let KK\subset\mathbb{H} be a Borel set with dimK=t[2,3]\dim_{\mathrm{\mathbb{H}}}K=t\in[2,3]. Then, dimEπe(K)t1\dim_{\mathrm{E}}\pi_{e}(K)\geq t-1 for 1\mathcal{H}^{1} almost every eS1e\in S^{1}. Consequently, dimπe(K)2t3\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K)\geq 2t-3 for 1\mathcal{H}^{1} almost every eS1e\in S^{1}.

Proof.

The lower bound for dimπe(K)\dim_{\mathrm{\mathbb{H}}}\pi_{e}(K) follows immediately from the lower bound for dimE(K)\dim_{\mathrm{E}}(K), combined with a general inequality between Hausdorff dimensions relative to Euclidean and Heisenberg metrics of subsets of 𝕎e\mathbb{W}_{e}, see [1, Theorem 2.8]. So, we focus on proving that dimE(K)t1\dim_{\mathrm{E}}(K)\geq t-1 for 1\mathcal{H}^{1} almost every eS1e\in S^{1}.

The first steps of the proof are standard; similar arguments have appeared, for example the deduction of [16, Theorem 2] from [16, Theorem 1]. So we only sketch the first part of the proof, and provide full details where they are non-standard. First, we may assume that KB(1)K\subset B_{\mathbb{H}}(1), and we may assume, applying Frostman’s lemma, that K=spt(μ)K=\operatorname{spt}(\mu) for some Borel probability measure μ\mu satisfying μ(B(p,r))rt\mu(B_{\mathbb{H}}(p,r))\lesssim r^{t} for all pp\in\mathbb{H} and r>0r>0.

We make the counter assumption that there exists s(1,t)s\in(1,t) such that

1({eS1:dimEπe(K)s1})>0.\mathcal{H}^{1}(\{e\in S^{1}:\dim_{\mathrm{E}}\pi_{e}(K)\leq s-1\})>0.

By several applications of the pigeonhole principle, this assumption can be applied to find the following objects for any ϵ>0\epsilon>0, and for arbitrarily small δ>0\delta>0:

  1. (1)

    A Borel subset SS1S^{\prime}\subset S^{1} of length 1(S)δϵ/2\mathcal{H}^{1}(S^{\prime})\geq\delta^{\epsilon/2}.

  2. (2)

    For every eSe\in S^{\prime} a collection of δ1s\leq\delta^{1-s} Euclidean δ\delta-discs 𝒲e\mathcal{W}_{e}, contained in 𝕎e\mathbb{W}_{e}.

  3. (3)

    If We:=𝒲eW_{e}:=\cup\mathcal{W}_{e} and eSe\in S^{\prime}, then

    μ(πe1(We))δϵ/2.\mu(\pi_{e}^{-1}(W_{e}))\geq\delta^{\epsilon/2}. (7.2)

We claim that (1)-(3) violate Theorem 5.11 if δ,ϵ>0\delta,\epsilon>0 are small enough. To this end, we first need to construct a relevant (δ,t,δϵ)(\delta,t,\delta^{-\epsilon})-set of (Heisenberg) δ\delta-balls \mathcal{B} contained in B(1)B_{\mathbb{H}}(1). Morally, this collection is a δ\delta-approximation of K=spt(μ)K=\operatorname{spt}(\mu). More precisely, we need to decompose KK to the following subsets:

Kα:={pK:α2μ(B(p,δ))α},K_{\alpha}:=\{p\in K:\tfrac{\alpha}{2}\leq\mu(B_{\mathbb{H}}(p,\delta))\leq\alpha\},

where α>0\alpha>0 runs over dyadic rationals with αδt\alpha\lesssim\delta^{t}. By one final application of the pigeonhole principle, and recalling (7.2), one can find a fixed index α2\alpha\in 2^{-\mathbb{N}} such that

μ(πe1(We)Kα)δϵ\mu(\pi_{e}^{-1}(W_{e})\cap K_{\alpha})\geq\delta^{\epsilon} (7.3)

for all eSSe\in S\subset S^{\prime}, where 1(S)δϵ\mathcal{H}^{1}(S)\geq\delta^{\epsilon}. In particular, μ(Kα)δϵ\mu(K_{\alpha})\geq\delta^{\epsilon}. Then, we let \mathcal{B} be a (Vitali) cover of KαK_{\alpha} by finitely overlapping Heisenberg δ\delta-balls with (δ/5)(\delta/5)-separated centres. Note that δϵα1||α1\delta^{\epsilon}\alpha^{-1}\lesssim|\mathcal{B}|\lesssim\alpha^{-1}. Using the definition of KαK_{\alpha}, and the Frostman condition for μ\mu, it is now easy to check that \mathcal{B} is a (δ,t,Cδϵ)(\delta,t,C\delta^{-\epsilon})-set of δ\delta-balls, where CC is roughly the Frostman constant of μ\mu.

Finally, from (7.3) and α||1\alpha\lesssim|\mathcal{B}|^{-1}, we deduce that if eSe\in S, then πe1(We)\pi_{e}^{-1}(W_{e}) intersects δϵ||\gtrsim\delta^{\epsilon}|\mathcal{B}| elements of \mathcal{B}, since

δϵμ(πe1(We)Kα)α|{B:πe1(We)B}|,eS.\delta^{\epsilon}\leq\mu(\pi_{e}^{-1}(W_{e})\cap K_{\alpha})\leq\alpha\cdot|\{B\in\mathcal{B}:\pi_{e}^{-1}(W_{e})\cap B\neq\emptyset\}|,\qquad e\in S.

Write e:={B:πe1(We)B}\mathcal{B}_{e}:=\{B\in\mathcal{B}:\pi_{e}^{-1}(W_{e})\cap B\neq\emptyset\}, thus |e|δϵ|||\mathcal{B}_{e}|\gtrsim\delta^{\epsilon}|\mathcal{B}|. We now arrive at the point where it is crucial that the elements of 𝒲e\mathcal{W}_{e} are Euclidean δ\delta-discs. Namely, if BeB\in\mathcal{B}_{e}, then πe1(D)B\pi_{e}^{-1}(D)\cap B\neq\emptyset for some D𝒲eD\in\mathcal{W}_{e}. Then, because DD is a Euclidean δ\delta-disc, and the Euclidean diameter of πe(B)\pi_{e}(B) is δ\lesssim\delta, we may conclude that πe(B)2D\pi_{e}(B)\subset 2D. This could seriously fail if DD were a disc in the metric dd_{\mathbb{H}}. Now, however, we see that

πe(e){2D:D𝒲e},\pi_{e}(\cup\mathcal{B}_{e})\subset\cup\{2D:D\in\mathcal{W}_{e}\},

and in particular Leb(πe(e))δ2|𝒲e|δ3s\mathrm{Leb}(\pi_{e}(\cup\mathcal{B}_{e}))\lesssim\delta^{2}\cdot|\mathcal{W}_{e}|\leq\delta^{3-s} for all eSe\in S. This violates the conclusion of Theorem 5.11, and the proof of Theorem 7.1 is complete. ∎

Appendix A Completing (δ,t)(\delta,t)-sets to (δ,3)(\delta,3)-sets

In this section, we use the following notation for the δ\delta-truncated ss-dimensional Riesz energy of a Radon measure ν\nu on \mathbb{H}:

Isδ(ν):=dν(x)dν(y)d,δ(x,y)s+t,I_{s}^{\delta}(\nu):=\iint\frac{d\nu(x)\,d\nu(y)}{d_{\mathbb{H},\delta}(x,y)^{s+t}},

where d,δ(x,y):=max{d(x,y),δ}d_{\mathbb{H},\delta}(x,y):=\max\{d_{\mathbb{H}}(x,y),\delta\}. We also recall that μν\mu\ast_{\mathbb{H}}\nu is the Heisenberg convolution of μ\mu and ν\nu, that is, the push-forward of μ×ν\mu\times\nu under the group operation (p,q)pq(p,q)\to p\cdot q.

Proposition A.1.

Let 0s,t30\leq s,t\leq 3 with s+t3s+t\leq 3, and let δ(0,12]\delta\in(0,\tfrac{1}{2}]. Let μ\mu be a Borel probability measure on B(1)B_{\mathbb{H}}(1) with Itδ(μ)𝐂I^{\delta}_{t}(\mu)\leq\mathbf{C}. Then, there exists a set HB(1)H\subset B_{\mathbb{H}}(1) with |H|δs|H|\leq\delta^{-s} such that the uniformly distributed (discrete) measure η\eta on HH satisfies

Is+tδ(ημ)𝐂,I_{s+t}^{\delta}(\eta\ast\mu)\leq\mathbf{C}^{\prime},

where 𝐂Clog(1/δ)C𝐂\mathbf{C}^{\prime}\leq C\log(1/\delta)^{C}\cdot\mathbf{C} for some absolute constant C>0C>0.

Proof.

Let Z:=δ3B(1)Z:=\delta\cdot\mathbb{Z}^{3}\cap B_{\mathbb{H}}(1) be a grid of Euclidean δ\delta-separated lattice points in B(1)B_{\mathbb{H}}(1). Then |Z|δ3|Z|\sim\delta^{-3}. Let HωZH_{\omega}\subset Z be a random set, where each point of ZZ is included independently with probability δs/(2|Z|)\delta^{-s}/(2|Z|). In particular, 𝔼ω|Hω|=δs/2\mathbb{E}_{\omega}|H_{\omega}|=\delta^{-s}/2. While we use the symbol "ω\omega" to index the elements in the underlying probability space, no explicit reference to this space will be needed. Let ηω\eta_{\omega} be the random measure

ηω:=δspHωδp=δspZ𝟏Hω(p)δp.\eta_{\omega}:=\delta^{s}\sum_{p\in H_{\omega}}\delta_{p}=\delta^{s}\sum_{p\in Z}\mathbf{1}_{H_{\omega}}(p)\cdot\delta_{p}.

We claim that

𝔼ω(Is+tδ(ηωμ))=𝔼ωdηω(p)dηω(q)d,δ(px,qy)s+t𝑑μ(x)𝑑μ(y)𝐂.\mathbb{E}_{\omega}\left(I^{\delta}_{s+t}(\eta_{\omega}\ast_{\mathbb{H}}\mu)\right)=\iint\mathbb{E}_{\omega}\iint\frac{d\eta_{\omega}(p)d\eta_{\omega}(q)}{d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)^{s+t}}\,d\mu(x)\,d\mu(y)\leq\mathbf{C}^{\prime}. (A.2)

for some 𝐂𝐂\mathbf{C}^{\prime}\lessapprox\mathbf{C}. In this argument, the notation "\lessapprox" hides a constant of the form Clog(1/δ)CC\log(1/\delta)^{C}. The inequality (A.2) will complete the proof of the proposition, because |Hω|δs|H_{\omega}|\leq\delta^{-s} with probability 12\geq\tfrac{1}{2} (for δ>0\delta>0 small enough), and therefore, by Chebychev’s inequality, Is+tδ(ηωμ)𝐂I_{s+t}^{\delta}(\eta_{\omega}\ast_{\mathbb{H}}\mu)\lesssim\mathbf{C}^{\prime} for some "ω\omega" with |Hω|δs|H_{\omega}|\leq\delta^{-s}.

To prove (A.2), it clearly suffices to establish that

𝔼ωdηω(p)dηω(q)d,δ(px,qy)s+t1d,δ(xy)t,x,yspt(μ)B(1).\mathbb{E}_{\omega}\iint\frac{d\eta_{\omega}(p)d\eta_{\omega}(q)}{d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)^{s+t}}\lessapprox\frac{1}{d_{\mathbb{H},\delta}(x\cdot y)^{t}},\qquad x,y\in\operatorname{spt}(\mu)\subset B_{\mathbb{H}}(1). (A.3)

By definition of ηω\eta_{\omega}, we have

dηω(p)dηω(q)d,δ(px,qy)s+t\displaystyle\iint\frac{d\eta_{\omega}(p)d\eta_{\omega}(q)}{d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)^{s+t}} =δ2sp,qZ𝟏Hω(p)𝟏Hω(q)d,δ(px,qy)s+t\displaystyle=\delta^{2s}\sum_{p,q\in Z}\frac{\mathbf{1}_{H_{\omega}}(p)\mathbf{1}_{H_{\omega}}(q)}{d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)^{s+t}}
=δ2spZ𝟏Hω(p)d,δ(x,y)s+t+δ2sp,qZpq𝟏Hω(p)𝟏Hω(q)d,δ(px,qy)s+t=:Σ1(ω)+Σ2(ω).\displaystyle=\delta^{2s}\sum_{p\in Z}\frac{\mathbf{1}_{H_{\omega}}(p)}{d_{\mathbb{H},\delta}(x,y)^{s+t}}+\delta^{2s}\sum_{\begin{subarray}{c}p,q\in Z\\ p\neq q\end{subarray}}\frac{\mathbf{1}_{H_{\omega}(p)}\mathbf{1}_{H_{\omega}(q)}}{d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)^{s+t}}=:\Sigma_{1}(\omega)+\Sigma_{2}(\omega).

We consider the expectations of Σ1(ω)\Sigma_{1}(\omega) and Σ2(ω)\Sigma_{2}(\omega) separately. The former one is simple, using that 𝔼ω(𝟏Hω(p))=ω{pHω}=δs/(2|Z|)δ3s\mathbb{E}_{\omega}(\mathbf{1}_{H_{\omega}}(p))=\mathbb{P}_{\omega}\{p\in H_{\omega}\}=\delta^{-s}/(2|Z|)\sim\delta^{3-s}:

𝔼ωΣ1(ω)δ2spZδ3sd,δ(x,y)s+t=|Z|δ3+sd,δ(x,y)s+tδsd,δ(x,y)s+t1d,δ(x,y)t,\mathbb{E}_{\omega}\Sigma_{1}(\omega)\sim\delta^{2s}\sum_{p\in Z}\frac{\delta^{3-s}}{d_{\mathbb{H},\delta}(x,y)^{s+t}}=\frac{|Z|\cdot\delta^{3+s}}{d_{\mathbb{H},\delta}(x,y)^{s+t}}\lesssim\frac{\delta^{s}}{d_{\mathbb{H},\delta}(x,y)^{s+t}}\leq\frac{1}{d_{\mathbb{H},\delta}(x,y)^{t}},

recalling that |Z|δ3|Z|\lesssim\delta^{-3}. To handle the expectation of Σ2(ω)\Sigma_{2}(\omega), we note that {pHω}\{p\in H_{\omega}\} and {qHω}\{q\in H_{\omega}\} are independent events for pqp\neq q, hence

𝔼ωΣ2(ω)δ2sp,qZpqδ62sd,δ(px,qy)s+tδ6pZδr1rst|{qZ:d,δ(px,qy)r}|,\mathbb{E}_{\omega}\Sigma_{2}(\omega)\sim\delta^{2s}\sum_{\begin{subarray}{c}p,q\in Z\\ p\neq q\end{subarray}}\frac{\delta^{6-2s}}{d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)^{s+t}}\sim\delta^{6}\sum_{p\in Z}\sum_{\delta\leq r\leq 1}r^{-s-t}|\{q\in Z:d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)\sim r\}|,

where "rr" runs over dyadic rationals. Since the product "\cdot" is not commutative, in general d,δ(px,qy)d,δ(pxy1,q)d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)\neq d_{\mathbb{H},\delta}(p\cdot x\cdot y^{-1},q), so the set {qZ:d,δ(px,qy)r}\{q\in Z:d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)\sim r\} is not contained in a \mathbb{H}-ball of radius r\sim r around pxy1p\cdot x\cdot y^{-1}. This is the key inefficiency in the argument, and causes the restriction s+t3s+t\leq 3: under this restriction, it actually suffices to note that {qZ:d,δ(px,qy)r}\{q\in Z:d_{\mathbb{H},\delta}(p\cdot x,q\cdot y)\sim r\} is contained in a Euclidean CrCr-ball. To see this, note that if qZq\in Z satisfies d,δ(px,qy)rd_{\mathbb{H},\delta}(p\cdot x,q\cdot y)\lesssim r with rδr\geq\delta, then

qB(px,Cr)y1.q\in B_{\mathbb{H}}(p\cdot x,Cr)\cdot y^{-1}.

Here B(px,Cr)B_{\mathbb{H}}(p\cdot x,Cr) is contained in a Euclidean ball of radius r\lesssim r (using r1r\leq 1). The same remains true after the right translation by y1y^{-1}, because |y|1|y|\lesssim 1 (by assumption), and the right translation zzy1z\mapsto z\cdot y^{-1} is Euclidean Lipschitz with constant depending only on |y||y|.

Now, since a Euclidean rr-ball contains (r/δ)3\lesssim(r/\delta)^{3} points of ZZ, we see that

𝔼ωΣ2(ω)δ3pZδr1r3st11d,δ(x,y)s+t,\mathbb{E}_{\omega}\Sigma_{2}(\omega)\lesssim\delta^{3}\sum_{p\in Z}\sum_{\delta\leq r\leq 1}r^{3-s-t}\lessapprox 1\leq\frac{1}{d_{\mathbb{H},\delta}(x,y)^{s+t}},

where in the final inequality we used again that x,yspt(μ)B(1)x,y\in\mathrm{spt}(\mu)\subset B_{\mathbb{H}}(1). This completes the proof of (A.3), and therefore the proof of the proposition. ∎

References

  • [1] Zoltán M. Balogh, Estibalitz Durand-Cartagena, Katrin Fässler, Pertti Mattila, and Jeremy T. Tyson. The effect of projections on dimension in the Heisenberg group. Rev. Mat. Iberoam., 29(2):381–432, 2013.
  • [2] Zoltán M. Balogh, Katrin Fässler, Pertti Mattila, and Jeremy T. Tyson. Projection and slicing theorems in Heisenberg groups. Adv. Math., 231(2):569–604, 2012.
  • [3] Luca Capogna, Donatella Danielli, Scott D. Pauls, and Jeremy T. Tyson. An introduction to the Heisenberg group and the sub-Riemannian isoperimetric problem, volume 259 of Progress in Mathematics. Birkhäuser Verlag, Basel, 2007.
  • [4] Katrin Fässler and Risto Hovila. Improved Hausdorff dimension estimate for vertical projections in the Heisenberg group. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 15:459–483, 2016.
  • [5] Katrin Fässler and Tuomas Orponen. A note on Kakeya sets of horizontal and SL(2)SL(2) lines. Bull. London Math. Soc. (to appear).
  • [6] Katrin Fässler, Tuomas Orponen, and Séverine Rigot. Semmes surfaces and intrinsic Lipschitz graphs in the Heisenberg group. Trans. Amer. Math. Soc., 373(8):5957–5996, 2020.
  • [7] Bruno Franchi and Raul Paolo Serapioni. Intrinsic Lipschitz graphs within Carnot groups. J. Geom. Anal., 26(3):1946–1994, 2016.
  • [8] Shengwen Gan, Shaoming Guo, Larry Guth, Terence L. J. Harris, Dominique Maldague, and Hong Wang. On restricted projections to planes in 3\mathbb{R}^{3}. arXiv e-prints, arXiv:2207.13844, July 2022.
  • [9] Shengwen Gan, Larry Guth, and Dominique Maldague. An exceptional set estimate for restricted projections to lines in 3\mathbb{R}^{3}. arXiv e-prints, page arXiv:2209.15152, September 2022.
  • [10] Larry Guth, Hong Wang, and Ruixiang Zhang. A sharp square function estimate for the cone in 3\mathbb{R}^{3}. Ann. of Math. (2), 192(2):551–581, 2020.
  • [11] Terence L. J. Harris. Length of sets under restricted families of projections onto lines. to appear in Contemp. Math. (AMS Conference proceedings).
  • [12] Terence L. J. Harris. Vertical projections in the Heisenberg group for sets of dimension greater than 3. Indiana Univ. Math. J. (to appear).
  • [13] Terence L. J. Harris. An a.e. lower bound for Hausdorff dimension under vertical projections in the Heisenberg group. Ann. Acad. Sci. Fenn. Math., 45(2):723–737, 2020.
  • [14] Terence L. J. Harris. A Euclidean Fourier-analytic approach to vertical projections in the Heisenberg group. Bull. Lond. Math. Soc., 55(2):961–977, 2023.
  • [15] Terence L. J. Harris, Chi N. Y. Huynh, and Fernando Román-García. Dimension distortion by right coset projections in the Heisenberg group. J. Fractal Geom., 8(4):305–346, 2021.
  • [16] Weikun He. Orthogonal projections of discretized sets. J. Fractal Geom., 7(3):271–317, 2020.
  • [17] Antti Käenmäki, Tuomas Orponen, and Laura Venieri. A Marstrand-type restricted projection theorem in 3\mathbb{R}^{3}. Amer. J. Math. (to appear).
  • [18] Jiayin Liu. On the dimension of Kakeya sets in the first Heisenberg group. Proc. Amer. Math. Soc., 150(8):3445–3455, 2022.
  • [19] J. M. Marstrand. Some fundamental geometrical properties of plane sets of fractional dimensions. Proc. London Math. Soc. (3), 4:257–302, 1954.
  • [20] Pertti Mattila. Hausdorff dimension, orthogonal projections and intersections with planes. Ann. Acad. Sci. Fenn. Ser. A I Math., 1(2):227–244, 1975.
  • [21] Pertti Mattila, Raul Serapioni, and Francesco Serra Cassano. Characterizations of intrinsic rectifiability in Heisenberg groups. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 9(4):687–723, 2010.
  • [22] Malabika Pramanik, Tongou Yang, and Joshua Zahl. A Furstenberg-type problem for circles, and a Kaufman-type restricted projection theorem in 3\mathbb{R}^{3}. arXiv e-prints, arXiv:2207.02259, July 2022.