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

More unit distances in arbitrary norms

Josef Greilhuber Carl Schildkraut  and  Jonathan Tidor Department of Mathematics, Stanford University, Stanford, CA 94305, USA {jgreil, carlsch, jtidor}@stanford.edu
Abstract.

For d2d\geq 2 and any norm on d\mathbb{R}^{d}, we prove that there exists a set of nn points that spans at least (d2o(1))nlog2n(\tfrac{d}{2}-o(1))n\log_{2}n unit distances under this norm for every nn. This matches the upper bound recently proved by Alon, Bucić, and Sauermann for typical norms (i.e., norms lying in a comeagre set). We also show that for d3d\geq 3 and a typical norm on d\mathbb{R}^{d}, the unit distance graph of this norm contains a copy of Kd,mK_{d,m} for all mm.

2020 Mathematics Subject Classification:
Primary: 52C10; Secondary: 52A20, 05C62
Schildkraut was supported by NSF Graduate Research Fellowship Program DGE-2146755. Tidor was supported by a Stanford Science Fellowship.

1. Introduction

One of the most well-known problems in discrete geometry is the Erdős unit distance problem. This asks for the maximum number of pairs of points at distance 1 among a set of nn points in 2\mathbb{R}^{2}. In 1946, Erdős conjectured that the answer is given by an appropriately scaled n×n\sqrt{n}\times\sqrt{n} section of the integer lattice, which determines n1+c/loglognn^{1+c/\log\log n} unit distances [3]. Despite considerable effort, the best known upper bound on this problem is O(n4/3)O(n^{4/3}), proved in 1984 by Spencer, Szemerédi, and Trotter [8].

One reason that explains the difficulty of improving this bound comes from studying the problem in other norms. A later proof of Székely [9] shows that the Spencer–Szemerédi–Trotter bound of O(n4/3)O(n^{4/3}) unit distances holds for any strictly convex norm on 2\mathbb{R}^{2}. Furthermore, Valtr observed that the norm whose unit ball is given by |y|+x21|y|+x^{2}\leq 1 achieves this bound [11]. Thus, to improve the O(n4/3)O(n^{4/3}) bound, one needs to use a property of the Euclidean norm which is not true of all strictly convex norms.

Matoušek first studied the unit distance problem for typical norms. He showed that most norms on 2\mathbb{R}^{2} span at most O(nlognloglogn)O(n\log n\log\log n) unit distances. Here, “most” means for a comeagre set of norms in the sense of the Baire category theorem. We will define this formally in Section 3. Given a norm {\lVert\cdot\rVert} on d\mathbb{R}^{d}, we define U(n)U_{{\lVert\cdot\rVert}}(n) to be the maximum number of unit distances spanned by a set of nn points in d\mathbb{R}^{d} under this norm. In this notation, Matoušek’s result can be stated as follows:

Theorem 1.1 ([6, Theorem 1.1]).

For most norms {\lVert\cdot\rVert} on 2\mathbb{R}^{2},

U(n)=O(nlognloglogn).U_{{\lVert\cdot\rVert}}(n)=O(n\log n\log\log n).

In addition to greatly improving the bound of O(n4/3)O(n^{4/3}), Matoušek’s bound is even smaller than n1+c/loglognn^{1+c/\log\log n}. Thus, the usual Euclidean norm has some special property that allows it to span somewhat more unit distances than typical norms. Furthermore, a simple construction shows that U(n)=Ω(nlogn)U_{{\lVert\cdot\rVert}}(n)=\Omega(n\log n) for every norm, so Matoušek’s result is tight up to a loglogn\log\log n factor.

Recently, Alon, Bucić, and Sauermann removed the loglogn\log\log n factor, proving a bound which is tight up to a constant multiplicative factor. They also generalized the result to all dimensions.

Theorem 1.2 ([1, Theorem 1.1]).

For most norms {\lVert\cdot\rVert} on d\mathbb{R}^{d},

U(n)d2nlog2n.U_{{\lVert\cdot\rVert}}(n)\leq\frac{d}{2}n\log_{2}n.

This upper bound holds for any norm such that the set of unit vectors do not satisfy unusually many short rational linear dependencies. Alon, Bucić, and Sauermann then showed that for most norms, their unit vectors do not satisfy these short rational linear dependencies.

For each norm, they also gave a family of constructions with almost-matching leading constant.

Theorem 1.3 ([1, Theorem 1.2]).

For every norm {\lVert\cdot\rVert} on d\mathbb{R}^{d},

U(n)(d212o(1))nlog2n.U_{{\lVert\cdot\rVert}}(n)\geq\left(\frac{d}{2}-\frac{1}{2}-o(1)\right)n\log_{2}n.

Here the o(1)o(1) term goes to 0 as nn\to\infty for each fixed dd.

Our main result removes the 1/21/2 in this bound, matching the upper bound provided by Theorem 1.2.

Theorem 1.4.

For d2d\geq 2 and every norm {\lVert\cdot\rVert} on d\mathbb{R}^{d},

U(n)(d2o(1))nlog2n.U_{{\lVert\cdot\rVert}}(n)\geq\left(\frac{d}{2}-o(1)\right)n\log_{2}n.

Here the o(1)o(1) term goes to 0 as nn\to\infty for each fixed dd.

Our second result is about large complete bipartite graphs in the unit distance graph of typical norms. For a norm {\lVert\cdot\rVert} on d\mathbb{R}^{d}, its unit distance graph is the graph with vertex set d\mathbb{R}^{d} where two vertices are adjacent if they are at distance 1 under {\lVert\cdot\rVert}. We show that, for d3d\geq 3, one can find a copy of Kd,mK_{d,m} for arbitrarily large mm in the unit distance graph of a typical norm on d\mathbb{R}^{d}.

This means that for a typical norm {\lVert\cdot\rVert} on d\mathbb{R}^{d}, one can find dd translates of the unit sphere in this norm whose intersection has arbitrarily large finite size. Heuristically, the intersection of d1d-1 translates of the unit sphere should be a 1-dimensional manifold, so one should expect to be able to find a copy of Kd1,K_{d-1,\infty} in the unit distance graph of any norm. (Indeed, Alon, Bucić, and Sauermann confirm this intuition [1, Lemma 7.4].) However the intersection of dd translates of a unit sphere is usually 0-dimensional and one might expect its size to be typically bounded. For example, the unit distance graph of any strictly convex norm on 2\mathbb{R}^{2} is K2,3K_{2,3}-free. For d3d\geq 3, we disprove this intuition for most norms, though the proof exploits the peculiarities of the definition of “most” quite strongly.

Theorem 1.5.

For d3d\geq 3 and most norms {\lVert\cdot\rVert} on d\mathbb{R}^{d}, the unit distance graph of {\lVert\cdot\rVert} contains a copy of Kd,mK_{d,m} for every mm.

This result can be used to give an alternative proof of Theorem 1.4 for most norms in dimension d3d\geq 3. We will discuss this more at the end of Section 5.

Notation

We write 2{\lVert\cdot\rVert}_{2} for the standard Euclidean norm on d\mathbb{R}^{d} and e1,,edde_{1},\ldots,e_{d}\in\mathbb{R}^{d} for the standard orthonormal basis of d\mathbb{R}^{d}. We write 𝕊d1\mathbb{S}^{d-1} for the standard unit sphere in d\mathbb{R}^{d}.

We use standard additive combinatorics notation. Given two sets X,YX,Y, write X+Y={x+y:xX,yY}X+Y=\{x+y:x\in X,y\in Y\} for their sumset. For a scalar xx and a set of vectors YY, write xY={xy:yY}x\cdot Y=\{xy:y\in Y\}. Similarly, for a vector yy and a set of scalars XX, write Xy={xy:xX}X\cdot y=\{xy:x\in X\}. We write [n]={1,,n}[n]=\{1,\ldots,n\}.

Acknowledgments

We thank the anonymous referees for a very careful reading of this paper.

2. Warm-up construction: d=2d=2

In this section we give a sketch of the proof of Theorem 1.4 in dimension d=2d=2. Let {\lVert\cdot\rVert} be any norm on 2\mathbb{R}^{2} and let B={x2:x1}B=\{x\in\mathbb{R}^{2}:\|x\|\leq 1\} be its unit ball. One can easily check that BB is a compact, convex subset of 2\mathbb{R}^{2} which is symmetric about 0 and contains a neighborhood of 0.

Let h=sup(x,y)By>0h=\sup_{(x,y)\in B}y>0 be the height of BB above the xx-axis. Then for each t[0,h]t\in[0,h], the horizontal line t:={(x,t):x}\ell_{t}:=\{(x,t):x\in\mathbb{R}\} intersects BB in a line segment. Define λ:[0,h]0\lambda\colon[0,h]\to\mathbb{R}_{\geqslant 0} by setting λ(t)\lambda(t) to be the length of tB\ell_{t}\cap B. It is not hard to check that λ\lambda is continuous and takes every value in the interval [0,w][0,w] where w=λ(0)w=\lambda(0).

Define t1,t2,,tm(0,h)t_{1},t_{2},\ldots,t_{m}\in(0,h) so that λ(ti)=im+1w\lambda(t_{i})=\tfrac{i}{m+1}w. Let pi,qiBp_{i},q_{i}\in\partial B be the left- and right-endpoint of tiB\ell_{t_{i}}\cap B, respectively. Then, defining v=we1/(m+1)v=we_{1}/(m+1), we have qi=pi+ivq_{i}=p_{i}+iv.

p1p_{1}q1q_{1}p2p_{2}q2q_{2}p3p_{3}q3q_{3}p4p_{4}q4q_{4}p5p_{5}q5q_{5}p6p_{6}q6q_{6}p7p_{7}q7q_{7}B\partial Bvv
Figure 1. Selection of the points pip_{i} and qiq_{i}.

Now define the set

S={a0v+a1p1++ampm:a0{0,1,,m21} and ai{0,1} for all i[m]}.S=\{a_{0}v+a_{1}p_{1}+\cdots+a_{m}p_{m}:a_{0}\in\{0,1,\ldots,m^{2}-1\}\text{ and }a_{i}\in\{0,1\}\text{ for all }i\in[m]\}.

This is a set of at most |S|2mm2|S|\leq 2^{m}m^{2} points. For now, suppose that |S|=2mm2|S|=2^{m}m^{2}.

Note that (q,q+pi)S2(q,q+p_{i})\in S^{2} is a pair of points at distance 1 for each qq that comes from a tuple (a0,,am)(a_{0},\ldots,a_{m}) with ai=0a_{i}=0. The same is true of (q,q+pi+iv)S2(q,q+p_{i}+iv)\in S^{2} for each qq with ai=0a_{i}=0 and a0<m2ia_{0}<m^{2}-i. Under the assumption that |S|=2mm2|S|=2^{m}m^{2}, there are at least 2m1m2=|S|/22^{m-1}m^{2}=|S|/2 pairs of points separated by the vector pip_{i} for each i[m]i\in[m] and at least 2m1(m2i)(11/m)|S|/22^{m-1}(m^{2}-i)\geq(1-1/m)|S|/2 pairs separated by pi+ivp_{i}+iv for each i[m]i\in[m]. This sums up to at least (m1/2)|S||S|log2|S|(m-1/2)|S|\approx|S|\log_{2}|S| unit distances.

We will show later (see Lemma 4.5) that collisions among elements of SS only help us; in other words, even if |S|<2mm2|S|<2^{m}m^{2}, the set still spans at least (m1/2)|S|(m-1/2)|S| unit distances. This construction works for each m1m\geq 1. Taking the union of these constructions for various values of mm allows one to produce a set of nn points for any nn with at least (1o(1))nlog2n(1-o(1))n\log_{2}n unit distances.

In the rest of the paper, we will fill in the details of this sketch and generalize it to all dimensions. In dimensions d3d\geq 3, it will take more work to find the points p1,,pmp_{1},\ldots,p_{m}; we will need to use some topological dimension theory to perform this step.

Our argument shares several ideas with Alon, Bucić, and Sauermann’s proof of Theorem 1.3. Both proofs use the Hurewicz dimension lowering theorem as part of the argument to find the points p1,,pmp_{1},\ldots,p_{m}, though the additional properties of our point set require a more involved argument. Once these points are found, both proofs use them to construct generalized arithmetic progressions (GAPs) that span many unit distances. In the Alon–Bucić–Sauermann argument, they are able to guarantee that the GAP is proper, while we cannot do this and instead show how to deal with non-proper GAPs. The main innovation in this paper is that the specific structure of our point set p1,,pmp_{1},\ldots,p_{m} produces a GAP which is even denser in the unit distance graph.

3. Preliminaries

We call a norm on d\mathbb{R}^{d} a dd-norm. There is a one-to-one correspondence between dd-norms {\lVert\cdot\rVert} and their unit balls {xd:x1}\{x\in\mathbb{R}^{d}:\|x\|\leq 1\}.

Definition 3.1.

A set BdB\subset\mathbb{R}^{d} is a unit ball if BB is compact, convex, symmetric about 0, and contains a neighborhood of 0. Given a unit ball BB, define the norm B{\lVert\cdot\rVert}_{B} by xB=r\|x\|_{B}=r where r0r\geq 0 is the smallest non-negative real such that xrBx\in r\cdot B. This is the norm whose unit ball is BB. A unit ball BB is strictly convex if B\partial B does not contain a line segment of positive length.

We record the property that the boundary B\partial B of a unit ball BdB\subset\mathbb{R}^{d} is homeomorphic to 𝕊d1\mathbb{S}^{d-1}. Indeed, one such homeomorphism B𝕊d1\partial B\to\mathbb{S}^{d-1} is given explicitly by xx/x2x\mapsto x/\|x\|_{2}.

Write d\mathcal{B}_{d} for the set of unit balls in d\mathbb{R}^{d}. We consider d\mathcal{B}_{d} as a metric space under the Hausdorff distance

dH(A,B):=max{supaAinfbBab2,supbBinfaAab2}.d_{H}(A,B):=\max\left\{\sup_{a\in A}\inf_{b\in B}\|a-b\|_{2},\sup_{b\in B}\inf_{a\in A}\|a-b\|_{2}\right\}.
Definition 3.2.

A set 𝒜d\mathcal{A}\subseteq\mathcal{B}_{d} is comeagre if it can be written as a countable intersection of sets, each of which has dense interior. We say that a property is true of most norms if there exists a comeagre set 𝒜d\mathcal{A}\subseteq\mathcal{B}_{d} such that the property holds for all B{\lVert\cdot\rVert}_{B} with B𝒜B\in\mathcal{A}.

By the Baire category theorem, d\mathcal{B}_{d} is a Baire space (this follows from, e.g., [4, Theorem 6.4]) meaning that every comeagre set is dense. To prove Theorem 1.5, we exploit some counterintuitive properties of the definition of comeagre and the Hausdorff distance. We will prove the following.

Proposition 3.3.

For each d3d\geq 3 and m1m\geq 1, there exists a dense open set 𝒜md\mathcal{A}_{m}\subseteq\mathcal{B}_{d} of unit balls which contain a Kd,mK_{d,m} in their unit distance graph.

This result implies Theorem 1.5, since the set m1𝒜m\bigcap_{m\geq 1}\mathcal{A}_{m} is a comeagre set of unit balls which contain a Kd,mK_{d,m} in their unit distance graph for all m1m\geq 1 simultaneously.

4. More unit distances in all dimensions

For d2d\geq 2, let BdB\in\mathcal{B}_{d} be a strictly convex unit ball. For a nonzero vector wdw\in\mathbb{R}^{d} and xBx\in\partial B, we say that ww is tangent to BB at xx if the line x+span{w}x+\operatorname{span}\{w\} intersects BB only at xx. Define φw:B\varphi_{w}\colon\partial B\to\mathbb{R} so that φw(x)=0\varphi_{w}(x)=0 if ww is tangent to BB at xx; otherwise φw(x)\varphi_{w}(x) is the unique nonzero scalar for which xφw(x)wBx-\varphi_{w}(x)w\in\partial B. We know that B(x+span{w})B\cap(x+\operatorname{span}\{w\}) is an interval; by the strict convexity of BB, the points inside this interval do not lie in B\partial B, so φw\varphi_{w} is well-defined.

We will need the following properties of the function φw\varphi_{w}.

Lemma 4.1.

For any strictly convex unit ball BdB\in\mathcal{B}_{d} and any non-zero vector ww,

  1. (1)

    the map φw:B\varphi_{w}\colon\partial B\to\mathbb{R} is continuous; and

  2. (2)

    the set Sw:=φw1(0)BS_{w}:=\varphi_{w}^{-1}(0)\subseteq\partial B of points xBx\in\partial B such that ww is tangent to BB at xx is homeomorphic to 𝕊d2\mathbb{S}^{d-2}.

The set SwS_{w} is called a shadow boundary of BB. Portions of this lemma appear in the literature, for example in [5]. For completeness, we give a proof here.

Proof.

Write W=span{w}W=\operatorname{span}\{w\} and let π:dW\pi\colon\mathbb{R}^{d}\to W^{\bot} denote the orthogonal projection. The image K:=π(B)K:=\pi(B) is then a unit ball in Wd1W^{\bot}\cong\mathbb{R}^{d-1}.

For each yKy\in K, let ψ(y)=min{t:y+twB}\psi_{-}(y)=\min\{t\in\mathbb{R}:y+tw\in B\} and ψ+(y)=max{t:y+twB}\psi_{+}(y)=\max\{t\in\mathbb{R}:y+tw\in B\}. We claim that ψ+\psi_{+} is continuous. Indeed, assume yKy\in K is such that there exist (yk)k=1(y_{k})_{k=1}^{\infty} in KK with limkyk=y\lim_{k\to\infty}y_{k}=y, but ψ+(yk)\psi_{+}(y_{k}) does not converge to ψ+(y)\psi^{+}(y). By the boundedness of ψ+\psi_{+}, we can assume limkψ+(yk)=ψ~ψ+(y)\lim_{k\to\infty}\psi_{+}(y_{k})=\tilde{\psi}\neq\psi^{+}(y), after passing to a subsequence. Both y+ψ+(y)wy+\psi_{+}(y)w and y+ψ~wy+\tilde{\psi}w lie in B\partial B, the latter since B\partial B is closed and yk+ψ+(yk)By_{k}+\psi_{+}(y_{k})\in\partial B. Hence ψ~<ψ+(y)\tilde{\psi}<\psi_{+}(y), and y+12(ψ+(y)+ψ~)wBy+\frac{1}{2}(\psi_{+}(y)+\tilde{\psi})w\in B. By the strict convexity of BB, this point does not lie in B\partial B (otherwise B\partial B would contain the three collinear points) so there exists a small Euclidean ball JJ around y+12(ψ+(y)+ψ~)wy+\frac{1}{2}(\psi_{+}(y)+\tilde{\psi})w that is contained in BB. This, however, implies that ψ+(y)12(ψ+(y)+ψ~)\psi_{+}(y^{\prime})\geq\frac{1}{2}(\psi_{+}(y)+\tilde{\psi}) for all yπ(J)y^{\prime}\in\pi(J), contradicting the assumption that limkψ+(yk)=ψ~<12(ψ+(y)+ψ~)\lim_{k\to\infty}\psi_{+}(y_{k})=\tilde{\psi}<\frac{1}{2}(\psi_{+}(y)+\tilde{\psi}). The continuity of ψ\psi_{-} follows analogously.

Observe that |φw(x)|=ψ+(π(x))ψ(π(x))|\varphi_{w}(x)|=\psi_{+}(\pi(x))-\psi_{-}(\pi(x)). Thus, |φw|=(ψ+ψ)π|\varphi_{w}|=(\psi_{+}-\psi_{-})\circ\pi is continuous. At xBx\in\partial B with φw(x)=0\varphi_{w}(x)=0, this implies φw\varphi_{w} is continuous as well. Now consider xBx\in\partial B with φw(x)0\varphi_{w}(x)\neq 0. Assume for the sake of contradiction that there exists a sequence (xk)k=1(x_{k})_{k=1}^{\infty} with limkxk=x\lim_{k\to\infty}x_{k}=x such that limkφw(xk)=φw(x)\lim_{k\to\infty}\varphi_{w}(x_{k})=-\varphi_{w}(x). (Any such sequence must have a subsequence with limit in {φw(x),φw(x)}\{-\varphi_{w}(x),\varphi_{w}(x)\} by continuity of |φw||\varphi_{w}|). Since B\partial B is closed, xφw(x)w=limk(xk+φk(xk)w)x-\varphi_{w}(x)w=\lim_{k\to\infty}(x_{k}+\varphi_{k}(x_{k})w) lies in B\partial B. By construction of φw\varphi_{w}, so does x+φw(x)wx+\varphi_{w}(x)w. But now B\partial B contains three collinear points, contradicting the assumption of strict convexity. Hence, φw\varphi_{w} is continuous at any point xBx\in\partial B with φw(x)0\varphi_{w}(x)\neq 0 as well.

For part (2), we claim that π\pi is a bijection between SwS_{w} and K\partial K. Consider xSwx\in S_{w}. By the separating hyperplane theorem (applied to intB\operatorname{int}B and x+Wx+W), there exists a supporting hyperplane HH to BB that contains x+Wx+W. Then π(H)\pi(H) is a supporting hyperplane to KK that contains π(x)\pi(x). Thus, π(x)K\pi(x)\in\partial K. On the other hand, for each pKp\in\partial K, pick some xBπ1(p)x\in\partial B\cap\pi^{-1}(p). Then the preimage Bπ1(p)B\cap\pi^{-1}(p) is the closed interval (x+W)B(x+W)\cap B and also lies in B\partial B. By the strict convexity of BB we conclude that (x+W)B={x}(x+W)\cap B=\{x\}. This means that xSwx\in S_{w}, showing that π\pi is a bijection between SwS_{w} and K\partial K.

Both SwS_{w} and K\partial K are closed and bounded, hence compact, and as subsets of Euclidean spaces they are Hausdorff topological spaces. As continuous bijections between compact Hausdorff spaces are homeomorphisms (see, e.g., [7, Theorem 26.6]), we conclude that π|Sw:SwK\pi|_{S_{w}}\colon S_{w}\to\partial K is a homeomorphism. ∎

In the next lemma we will use some dimension theory. Throughout the proof, dimension will be the Lebesgue covering dimension, defined in [2, Definition 1.6.7]. By Urysohn’s theorem, this coincides with the notion of small and large inductive dimension (defined in [2, Definitions 1.1.1 and 1.6.1]) for separable metric spaces [2, Theorem 1.7.7]. All that we will use are the following facts.

Proposition 4.2.

For nonempty sets XmX\subseteq\mathbb{R}^{m} and YnY\subseteq\mathbb{R}^{n},

  1. (1)

    dimX0\dim X\in\mathbb{Z}_{\geqslant 0};

  2. (2)

    if XX is homeomorphic to YY, then dimX=dimY\dim X=\dim Y;

  3. (3)

    if dimX1\dim X\geq 1, then XX is infinite;

  4. (4)

    if XYX\subseteq Y, then dimXdimY\dim X\leq\dim Y;

  5. (5)

    if XX is compact and f:XYf\colon X\to Y is a continuous map such that dimf1(y)k\dim f^{-1}(y)\leq k for all yYy\in Y, then dimXdimY+k\dim X\leq\dim Y+k; and

  6. (6)

    dimXm\dim X\leq m with equality if and only if XX has nonempty interior.

The first three follow from the definitions while the fourth is [2, Theorem 1.1.2]. The fifth is the Hurewicz dimension lowering theorem, given as [2, Theorem 1.12.4], and the sixth is [2, Theorems 1.8.2 and 1.8.10].

Lemma 4.3.

Let BdB\in\mathcal{B}_{d} be a strictly convex unit ball and let w1,,wd1w_{1},\ldots,w_{d-1} be linearly independent vectors in d\mathbb{R}^{d}. Define the map Φ:Bd1\Phi\colon\partial B\to\mathbb{R}^{d-1} by Φ=(φw1,,φwd1)\Phi=(\varphi_{w_{1}},\ldots,\varphi_{w_{d-1}}). Then, for each positive integer mm, there exist distinct vectors p1,,pmBp_{1},\ldots,p_{m}\in\partial B, a vector td1t\in\mathbb{R}^{d-1}, and a scalar λ0\lambda\geq 0 satisfying the following conditions:

  • the points p1,,pmp_{1},\ldots,p_{m} lie (strictly) on the same side of the hyperplane span{w1,,wd1}\operatorname{span}\{w_{1},\ldots,w_{d-1}\};

  • no coordinate of tt is zero; and

  • Φ(pi)=(1+iλ)t\Phi(p_{i})=(1+i\lambda)t for each 1im1\leq i\leq m.

Proof.

By Lemma 4.1(2), Φ\Phi is continuous.

Let wdw_{d} be a unit vector orthogonal to w1,,wd1w_{1},\ldots,w_{d-1}. Pick a connected open set UBU\subset\partial B such that its closure, U¯\overline{U}, is disjoint from the union of shadow boundaries Sw1Swd1S_{w_{1}}\cup\cdots\cup S_{w_{d-1}} as well as from the hyperplane wdw_{d}^{\perp}. To see such a UU exists, note that by Lemma 4.1(1), each of

Sw1,Sw2,,Swd1,wdBS_{w_{1}},S_{w_{2}},\ldots,S_{w_{d-1}},w_{d}^{\perp}\cap\partial B

are subsets of B\partial B homeomorphic to 𝕊d2\mathbb{S}^{d-2}, while B\partial B is homeomorphic to 𝕊d1\mathbb{S}^{d-1}. Clearly, this suffices for such a UU to exist.

By definition, no coordinate of any point in Φ(U¯)\Phi(\overline{U}) is zero. By the continuity of Φ\Phi and the compactness of U¯\overline{U}, there exists ε>0\varepsilon>0 such that all coordinates of Φ(x)\Phi(x) have magnitude at least ε\varepsilon for all xUx\in U. Pick η>0\eta>0 such that |xwd|>η|x\cdot w_{d}|>\eta for all xUx\in U. By the central symmetry of BB, we may assume that xwd>ηx\cdot w_{d}>\eta for all xUx\in U.

Define

Y={(t1,,td1)d1:min(|t1|,,|td1|)<ε},Y=\{(t_{1},\ldots,t_{d-1})\in\mathbb{R}^{d-1}:\min(|t_{1}|,\ldots,|t_{d-1}|)<\varepsilon\},

an open neighborhood of the union of the coordinate hyperplanes. Also, define the closed half-space

Z={xd:xwdη}.Z=\{x\in\mathbb{R}^{d}:x\cdot w_{d}\geq\eta\}.

Write V=(BΦ1(Y))ZV=(\partial B\setminus\Phi^{-1}(Y))\cap Z. Since Φ\Phi is continuous, VV is compact. We also have UVU\subset V. Since B\partial B is homeomorphic to 𝕊d1\mathbb{S}^{d-1}, some subset of UU is homeomorphic to a non-empty open set in d1\mathbb{R}^{d-1}. Therefore, by 4.2(2)(4) (6), VV has dimension d1d-1.

By 4.2(5) applied to Φ|V:Vd1Y\Phi|_{V}\colon V\to\mathbb{R}^{d-1}\setminus Y, one of the following must hold:

  1. (a)

    some fiber Φ1(t)\Phi^{-1}(t) for tYt\not\in Y has positive dimension when intersected with ZZ, or

  2. (b)

    the image Φ(V)Φ(B)Y\Phi(V)\subset\Phi(\partial B)\setminus Y has dimension d1d-1.

In case (a), such a vector tt is not on any coordinate hyperplane, since it is not in YY. Since Φ1(t)Z\Phi^{-1}(t)\cap Z has positive dimension, by 4.2(3) it contains infinitely many points, and so we can simply take arbitrary p1,,pmp_{1},\ldots,p_{m} among them, with λ=0\lambda=0.

We now treat case (b). By 4.2(6), Φ(V)\Phi(V) contains some open ball TT in d1\mathbb{R}^{d-1}. Let tt be the center of such a ball; note that tt is not on any coordinate hyperplane. Let λ\lambda be small enough that (1+mλ)tT(1+m\lambda)t\in T and define ti=(1+iλ)tt_{i}=(1+i\lambda)t for each 1im1\leq i\leq m. Since tiTt_{i}\in T for each ii, we can find some piΦ1(ti)BZp_{i}\in\Phi^{-1}(t_{i})\cap\partial B\cap Z. These points lie in ZZ, and so they all lie on the same side of the hyperplane span{w1,,wd1}\operatorname{span}\{w_{1},\ldots,w_{d-1}\}, as desired. ∎

To prove Theorem 1.4 we will need a lemma about the number of unit distances spanned by a generalized arithmetic progression (GAP) whose increments are unit vectors. This is easy to compute for proper GAPs but we will show that similar bounds hold in general in terms of the size of the GAP. In the next two lemmas, we write [a,b]={a,a+1,,b}[a,b]=\{a,a+1,\ldots,b\}. We say that a set of vectors u1,,umu_{1},\ldots,u_{m} is non-overlapping if the 2m2m vectors ±u1,,±um\pm u_{1},\ldots,\pm u_{m} are distinct.

Lemma 4.4.

For integers kc2k\geq c\geq 2, a vector xdx\in\mathbb{R}^{d}, and a finite set XdX\subset\mathbb{R}^{d} we have the inequality

|X+[0,kc1]x|(1ck)|X+[0,k1]x|.|X+[0,k-c-1]\cdot x|\geq\left(1-\frac{c}{k}\right)|X+[0,k-1]\cdot x|.
Proof.

Define

Δi=|X+[0,i]x||X+[0,i1]x|\Delta_{i}=|X+[0,i]\cdot x|-|X+[0,i-1]\cdot x|

for i0i\geq 0. (Note that Δ0=|X|\Delta_{0}=|X|.)

We claim that Δ0Δ1Δ2Δ3\Delta_{0}\geq\Delta_{1}\geq\Delta_{2}\geq\Delta_{3}\geq\cdots. This is because

Δi\displaystyle\Delta_{i} =|(X+[0,i]x)(X+[0,i1]x)|\displaystyle=|(X+[0,i]\cdot x)\setminus(X+[0,i-1]\cdot x)|
=|(X+{ix})(X+[0,i1]x)|\displaystyle=|(X+\{ix\})\setminus(X+[0,i-1]\cdot x)|
=|X(X+[i,1]x)|.\displaystyle=|X\setminus(X+[-i,-1]\cdot x)|.

Clearly the sets on the final line decrease in size as ii increases. Finally we conclude that

|X+[0,kc1]x|\displaystyle|X+[0,k-c-1]\cdot x| =Δ0+Δ1++Δkc1\displaystyle=\Delta_{0}+\Delta_{1}+\cdots+\Delta_{k-c-1}
(1ck)(Δ0+Δ1++Δk1)\displaystyle\geq\left(1-\frac{c}{k}\right)(\Delta_{0}+\Delta_{1}+\cdots+\Delta_{k-1})
=(1ck)|X+[0,k1]x|.\displaystyle=\left(1-\frac{c}{k}\right)|X+[0,k-1]\cdot x|.\qed
Lemma 4.5.

Let v1,,vmdv_{1},\ldots,v_{m}\in\mathbb{R}^{d} be vectors and let k1,,km2k_{1},\ldots,k_{m}\geq 2 be integers. Suppose U[0,k1]××[0,km]U\subseteq[0,k_{1}]\times\cdots\times[0,k_{m}] is such that

{c1v1++cmvm:(c1,,cm)U}\{c_{1}v_{1}+\cdots+c_{m}v_{m}\colon(c_{1},\ldots,c_{m})\in U\}

is a non-overlapping set of |U||U| unit vectors. Define

S={a1v1++amvm:ai[0,ki1] for i[m]}.S=\{a_{1}v_{1}+\cdots+a_{m}v_{m}:a_{i}\in[0,k_{i}-1]\text{ for }i\in[m]\}.

Then SS spans at least

|S|cUi=1m(1ciki)|S|\cdot\sum_{c\in U}\prod_{i=1}^{m}\left(1-\frac{c_{i}}{k_{i}}\right)

unit distances.

Proof.

Set A=i=1m[0,ki1]mA=\prod_{i=1}^{m}[0,k_{i}-1]\subset\mathbb{Z}^{m}. Then define Ψ:md\Psi\colon\mathbb{Z}^{m}\to\mathbb{R}^{d} by Ψ(a)=a1v1++amvm\Psi(a)=a_{1}v_{1}+\cdots+a_{m}v_{m} where aia_{i} denotes the iith component of aa. For each cUc\in U, define Ac=i=1m[0,kici1]A_{c}=\prod_{i=1}^{m}[0,k_{i}-c_{i}-1].

Note that |S|=|Ψ(A)||S|=|\Psi(A)| and SS spans at least |Ψ(Ac)||\Psi(A_{c})| unit distances in the direction Ψ(c)\Psi(c). The latter is because for each xΨ(Ac)x\in\Psi(A_{c}), there exists aAca\in A_{c} such that Ψ(a)=x\Psi(a)=x. Then (Ψ(a),Ψ(a+c))=(x,x+Ψ(c))(\Psi(a),\Psi(a+c))=(x,x+\Psi(c)) spans a unit distance in the direction Ψ(c)\Psi(c). Since {Ψ(c):cU}\{\Psi(c):c\in U\} is non-overlapping, these unit distances are distinct, so SS spans at least cU|Ψ(Ac)|\sum_{c\in U}|\Psi(A_{c})| unit distances. Now by mm applications of Lemma 4.4,

|Ψ(Ac)||Ψ(A)|i=1m(1ciki),\frac{|\Psi(A_{c})|}{|\Psi(A)|}\geq\prod_{i=1}^{m}\left(1-\frac{c_{i}}{k_{i}}\right),

implying the desired result. ∎

Combining the previous results in this section, we find GAPs that span many unit distances. For technical reasons, we need to construct a nested sequence of GAPs S1S2SmS_{1}\subseteq S_{2}\subseteq\cdots\subseteq S_{m}.

Proposition 4.6.

Let BdB\in\mathcal{B}_{d} be a strictly convex unit ball. For each m1m\geq 1, there exist vectors v1,,vm+2d2v_{1},\ldots,v_{m+2d-2} with the following property. For [m]\ell\in[m], define the sets SdS_{\ell}\subset\mathbb{R}^{d} by

S={a1v1++am+2d2vm+2d2:ai[0,ki1] for i[m+2d2]}S_{\ell}=\left\{a_{1}v_{1}+\cdots+a_{m+2d-2}v_{m+2d-2}:a_{i}\in[0,k_{i}-1]\text{ for }i\in[m+2d-2]\right\}

where k1==k=2k_{1}=\cdots=k_{\ell}=2 and k+1==km=1k_{\ell+1}=\cdots=k_{m}=1 and km+1==km+d1=k_{m+1}=\cdots=k_{m+d-1}=\ell and km+d==km+2d2=2k_{m+d}=\cdots=k_{m+2d-2}=\ell^{2}. Then |S|23(d1)|S_{\ell}|\leq 2^{\ell}\ell^{3(d-1)} and SS_{\ell} spans at least d(2)|S|/2d(\ell-2)|S_{\ell}|/2 unit distances under B{\lVert\cdot\rVert}_{B}.

Proof.

Let w1,,wd1w_{1},\ldots,w_{d-1} be arbitrary linearly independent vectors and let H=wdH=w_{d}^{\perp} be the hyperplane they span. Now, use Lemma 4.3 to find some t({0})dt\in(\mathbb{R}\setminus\{0\})^{d}, some scalar λ0\lambda\geq 0, and some p1,,pmBp_{1},\ldots,p_{m}\in\partial B on the same side of HH for which

Φ(pj)=(1+jλ)t for all j[m].\Phi(p_{j})=(1+j\lambda)t\qquad\text{ for all }j\in[m].

By swapping the sign of wdw_{d} if necessary, we may assume that p1wd,,pmwd>0p_{1}\cdot w_{d},\ldots,p_{m}\cdot w_{d}>0.

Now, by the definition of Φ\Phi, we have

qij:=pjφwi(pj)wi=pj(1+jλ)tiwiBfor all j[m],i[d1].q_{ij}:=p_{j}-\varphi_{w_{i}}(p_{j})w_{i}=p_{j}-(1+j\lambda)t_{i}w_{i}\in\partial B\qquad\text{for all }j\in[m],i\in[d-1].

We claim that the set

𝒰:={pj:j[m]}{qij:i[d1],j[m]}B\mathcal{U}:=\{p_{j}:j\in[m]\}\cup\left\{q_{ij}:i\in[d-1],j\in[m]\right\}\subset\partial B

is non-overlapping of size dmdm. Indeed,

  1. (1)

    No two elements of 𝒰\mathcal{U} are antipodes: since qijwd=pjwd>0q_{ij}\cdot w_{d}=p_{j}\cdot w_{d}>0, the set 𝒰\mathcal{U} is contained within the half-space {x:wdx>0}\{x:w_{d}\cdot x>0\}.

  2. (2)

    The pjp_{j} are distinct by definition.

  3. (3)

    We do not have qij=pjq_{ij}=p_{j^{\prime}} for any i,j,ji,j,j^{\prime}, since the quantities

    φwi(pj)=(1+jλ)tiandφwi(qij)=φwi(pj)=(1+jλ)ti\varphi_{w_{i}}(p_{j^{\prime}})=(1+j^{\prime}\lambda)t_{i}\quad\text{and}\quad\varphi_{w_{i}}(q_{ij})=-\varphi_{w_{i}}(p_{j})=-(1+j\lambda)t_{i}

    differ in sign (as ti0t_{i}\neq 0).

  4. (4)

    We do not have have qij=qijq_{ij}=q_{ij^{\prime}} for jjj\neq j^{\prime} since

    qijφwi(qij)wi=pj,q_{ij}-\varphi_{w_{i}}(q_{ij})w_{i}=p_{j},

    so if qij=qijq_{ij}=q_{ij^{\prime}} then pj=pjp_{j}=p_{j^{\prime}}.

  5. (5)

    We do not have qij=qijq_{ij}=q_{i^{\prime}j^{\prime}} for iii\neq i^{\prime}: if these were equal, then

    qij\displaystyle q_{i^{\prime}j} =pjφwi(pj)wi=qij+φwi(pj)wiφwi(pj)wi\displaystyle=p_{j}-\varphi_{w_{i^{\prime}}}(p_{j})w_{i^{\prime}}=q_{ij}+\varphi_{w_{i}}(p_{j})w_{i}-\varphi_{w_{i^{\prime}}}(p_{j})w_{i^{\prime}}
    =qij+(1+jλ)(tiwitiwi)\displaystyle=q_{ij}+(1+j\lambda)\left(t_{i}w_{i}-t_{i^{\prime}}w_{i^{\prime}}\right)
    qij\displaystyle q_{ij^{\prime}} =pjφwi(pj)wi=qij+φwi(pj)wiφwi(pj)wi\displaystyle=p_{j^{\prime}}-\varphi_{w_{i}}(p_{j^{\prime}})w_{i}=q_{i^{\prime}j^{\prime}}+\varphi_{w_{i^{\prime}}}(p_{j^{\prime}})w_{i^{\prime}}-\varphi_{w_{i}}(p_{j^{\prime}})w_{i}
    =qij(1+jλ)(tiwitiwi).\displaystyle=q_{ij}-(1+j^{\prime}\lambda)\left(t_{i}w_{i}-t_{i^{\prime}}w_{i^{\prime}}\right).

    Since tiwitiwi0t_{i}w_{i}-t_{i^{\prime}}w_{i^{\prime}}\neq 0, the three points qij,qij,qijq_{i^{\prime}j},q_{ij^{\prime}},q_{ij} are distinct and collinear. Therefore, some line intersects B\partial B three times, contradicting the strict convexity of BB.

pjp_{j^{\prime}}pjp_{j}qijq_{ij^{\prime}}qij=qijq_{ij}=q_{i^{\prime}j^{\prime}}qijq_{i^{\prime}j}(1+jλ)tiwi(1+j^{\prime}\lambda)t_{i}w_{i}(1+jλ)tiwi(1+j^{\prime}\lambda)t_{i^{\prime}}w_{i^{\prime}}(1+jλ)tiwi(1+j\lambda)t_{i}w_{i}(1+jλ)tiwi(1+j\lambda)t_{i^{\prime}}w_{i^{\prime}}B\partial B
Figure 2. Case (5) of the non-overlapping property of 𝒰\mathcal{U}.

Now for [m]\ell\in[m], define

(v1,,vm+2d2)=(p1,,pm,t1w1,,td1wd1,λt1w1,,λtd1wd1),(v_{1},\ldots,v_{m+2d-2})=\left(p_{1},\ldots,p_{m},-t_{1}w_{1},\ldots,-t_{d-1}w_{d-1},-\lambda t_{1}w_{1},\ldots,-\lambda t_{d-1}w_{d-1}\right),

and select k1==k=2k_{1}=\cdots=k_{\ell}=2 and k+1==km=1k_{\ell+1}=\cdots=k_{m}=1 and km+1==km+d1=k_{m+1}=\cdots=k_{m+d-1}=\ell and km+d==km+2d2=2k_{m+d}=\cdots=k_{m+2d-2}=\ell^{2}. Set

U={ej:j[m]}{ej+em+i+jem+d1+i:j[m],i[d1]}m+2d2U=\{e_{j}:j\in[m]\}\cup\left\{e_{j}+e_{m+i}+je_{m+d-1+i}:j\in[m],i\in[d-1]\right\}\subset\mathbb{Z}^{m+2d-2}

so that

𝒰={c1v1++cm+2d2vm+2d2:cU}.\mathcal{U}=\{c_{1}v_{1}+\cdots+c_{m+2d-2}v_{m+2d-2}:c\in U\}.

We just proved that 𝒰\mathcal{U} is a non-overlapping set of size dmdm. Clearly the elements of 𝒰\mathcal{U} are unit vectors under B{\lVert\cdot\rVert}_{B}. Thus, by Lemma 4.5, the set

S={a1v1++am+2d2vm+2d2:ai{0,1,,ki1} for i[m+2d2]}S_{\ell}=\{a_{1}v_{1}+\cdots+a_{m+2d-2}v_{m+2d-2}:a_{i}\in\{0,1,\ldots,k_{i}-1\}\text{ for }i\in[m+2d-2]\}

with |S|iki=23(d1)|S_{\ell}|\leq\prod_{i}k_{i}=2^{\ell}\ell^{3(d-1)} spans at least

|S|cUi=1m+2d2(1ciki)|S_{\ell}|\cdot\sum_{c\in U}\prod_{i=1}^{m+2d-2}\left(1-\frac{c_{i}}{k_{i}}\right)

unit distances under B{\lVert\cdot\rVert}_{B}. This quantity is

|S|j=1(12+(d1)1212j2)|S|(12)j=1d2=d2(2)|S|.|S_{\ell}|\cdot\sum_{j=1}^{\ell}\left(\frac{1}{2}+(d-1)\frac{1}{2}\frac{\ell-1}{\ell}\frac{\ell^{2}-j}{\ell^{2}}\right)\geq|S_{\ell}|\left(1-\frac{2}{\ell}\right)\sum_{j=1}^{\ell}\frac{d}{2}=\frac{d}{2}(\ell-2)|S_{\ell}|.\qed

Now all that remains to prove the main theorem is to construct a set of exactly nn points by taking a union of translates of the GAPs provided by the previous proposition.

Proof of Theorem 1.4.

Let BdB\in\mathcal{B}_{d} be a unit ball. If BB is not strictly convex, it is well-known that UB(n)=Θ(n2)U_{{\lVert\cdot\rVert}_{B}}(n)=\Theta(n^{2}). Indeed, suppose B\partial B contains the segment connecting xyx-y and x+yx+y for some x,ydx,y\in\mathbb{R}^{d}. Then the subgraph of the unit distance graph of B{\lVert\cdot\rVert}_{B} induced by the segments (0,y)(0,y) and (x,x+y)(x,x+y) contains a copy of K,K_{\infty,\infty}. So, we may henceforth assume that BB is strictly convex.

We apply 4.6 with m=nm=n, to find a nested sequence of sets S1S2SnS_{1}\subseteq S_{2}\subseteq\cdots\subseteq S_{n} with the following properties: sm:=|Sm|2mm3(d1)s_{m}:=|S_{m}|\leq 2^{m}m^{3(d-1)} and SmS_{m} spans tmd(m2)sm/2t_{m}\geq d(m-2)s_{m}/2 unit distances. Note that since the SmS_{m} are nested, we have 1s1s2sn1\leq s_{1}\leq s_{2}\leq\cdots\leq s_{n}. We also have the easy bound snns_{n}\geq n.

Define S0S_{0} to be a single point and set s0=1s_{0}=1 and t0=0t_{0}=0. For each nn, we define a set SS of nn points that determines many unit distances as follows. Write n=i=1rsmin=\sum_{i=1}^{r}s_{m_{i}} where nm1m2mr0n\geq m_{1}\geq m_{2}\geq\cdots\geq m_{r}\geq 0 is the lexicographically largest sequence with this property. Since s0=1s_{0}=1, there exists at least one sequence with this property. We will define S=i=1r(xi+Smi)S=\bigcup_{i=1}^{r}(x_{i}+S_{m_{i}}) where x1,,xrx_{1},\ldots,x_{r} are generically chosen vectors. In particular, the xi+Smix_{i}+S_{m_{i}} are disjoint, giving |S|=n|S|=n. Write tt for the number of unit distances spanned by SS. We know that ti=1rtmit\geq\sum_{i=1}^{r}t_{m_{i}} which we will now show is large.

From the definition of the SmS_{m}, we have that Sm1SmS_{m-1}\subseteq S_{m} for each mm. Applying Lemma 4.4, we see that for m2m\geq 2

smsm12(mm1)d1(m2(m1)2)d12e3(d1)/(m1).\frac{s_{m}}{s_{m-1}}\leq 2\left(\frac{m}{m-1}\right)^{d-1}\left(\frac{m^{2}}{(m-1)^{2}}\right)^{d-1}\leq 2e^{3(d-1)/(m-1)}.

In particular, sm4sm1s_{m}\leq 4s_{m-1} for m5dm\geq 5d. Let MM be the smallest integer such that sMn/log2ns_{M}\geq n/\log_{2}n. (Since snns_{n}\geq n, this is well-defined.) From the bound sm2mm3(d1)s_{m}\leq 2^{m}m^{3(d-1)}, we see that (for each fixed dd) MM goes to infinity as nn goes to infinity.

If nn is sufficiently large, then M5dM\geq 5d, implying that sM4sM1<4n/log2ns_{M}\leq 4s_{M-1}<4n/\log_{2}n. Then, since we chose (m1,m2,)(m_{1},m_{2},\ldots) to be lexicographically largest, we have the property

i:miMsmi>nsM>n4nlog2n.\sum_{i:m_{i}\geq M}s_{m_{i}}>n-s_{M}>n-\frac{4n}{\log_{2}n}.

Furthermore, for miMm_{i}\geq M, the conclusion of 4.6 gave us the bound

tmismid(mi2)2(d2o(1))log2smi(d2o(1))(log2nlog2log2n)\frac{t_{m_{i}}}{s_{m_{i}}}\geq\frac{d(m_{i}-2)}{2}\geq\left(\frac{d}{2}-o(1)\right)\log_{2}s_{m_{i}}\geq\left(\frac{d}{2}-o(1)\right)(\log_{2}n-\log_{2}\log_{2}n)

where o(1)0o(1)\to 0 as nn\to\infty for each fixed dd. (The last inequality follows since for miMm_{i}\geq M we have smisMn/log2ns_{m_{i}}\geq s_{M}\geq n/\log_{2}n.)

Therefore we see that

ti:miMtmi\displaystyle t\geq\sum_{i:m_{i}\geq M}t_{m_{i}} (i:miMsmi)(d2o(1))(log2nlog2log2n)\displaystyle\geq\left(\sum_{i:m_{i}\geq M}s_{m_{i}}\right)\left(\frac{d}{2}-o(1)\right)\left(\log_{2}n-\log_{2}\log_{2}n\right)
(n4nlog2n)(d2o(1))log2n\displaystyle\geq\left(n-\frac{4n}{\log_{2}n}\right)\left(\frac{d}{2}-o(1)\right)\log_{2}n
(d2o(1))nlog2n.\displaystyle\geq\left(\frac{d}{2}-o(1)\right)n\log_{2}n.\qed

5. Kd,mK_{d,m}’s in the unit distance graph

In this section we prove 3.3, finding for each positive integer mm a copy of Kd,mK_{d,m} in the unit distance graph for an open dense subset of norms.

We begin by introducing the machinery we will use. To find a copy of Kd,mK_{d,m} in the unit distance graph of B{\lVert\cdot\rVert}_{B}, we must find dd translates of B\partial B which intersect in mm points. To show that the set 𝒜m\mathcal{A}_{m} of norms we construct is open, we want to show that these intersections persist under small perturbations of the unit ball. The principal tool to ensure this kind of stability is the Brouwer mapping degree.

The Brouwer mapping degree is an invariant of continuous maps which should be thought of as a robust “signed count” of preimages. We refer the interested reader to the textbook [10, Chapter 10] for a treatment of the mapping degree requiring only elementary analysis and measure theory.

Consider a bounded open set UdU\subset\mathbb{R}^{d}, a continuous map f:U¯df\colon\overline{U}\to\mathbb{R}^{d}, and a point ydf(U)y\in\mathbb{R}^{d}\setminus f(\partial U). The degree of ff with respect to UU and yy, denoted by deg(f,U,y)\deg(f,U,y), is an integer that satisfies the following properties:

  1. (1)

    if deg(f,U,y)0\deg(f,U,y)\neq 0, then yf(U)y\in f(U);

  2. (2)

    if f,g:U¯df,g\colon\overline{U}\to\mathbb{R}^{d} are continuous maps such that f(x)g(x)2<f(x)y2\|f(x)-g(x)\|_{2}<\|f(x)-y\|_{2} for all xUx\in\partial U, then deg(f,U,y)=deg(g,U,y)\deg(f,U,y)=\deg(g,U,y);

  3. (3)

    if f:U¯df\colon\overline{U}\to\mathbb{R}^{d} is continuously differentiable and the Jacobian Jf(x)=det(ifj(x))i,j=1dJ_{f}(x)=\det(\partial_{i}f_{j}(x))_{i,j=1}^{d} is nonzero at all xf1(y)x\in f^{-1}(y), then deg(f,U,y)=xf1(y)sgnJf(x)\deg(f,U,y)=\sum_{x\in f^{-1}(y)}\operatorname{sgn}\,J_{f}(x).

The existence of such a notion of degree follows from [10, Theorems 10.1 and 10.4].

In the first part of the proof, we construct a local model of a unit ball which contains a Kd,mK_{d,m} in its unit distance graph and show that this property is stable under small perturbations, ensuring openness. In the second part of the proof, we show that, near any unit ball in d\mathcal{B}_{d}, we can find one which looks like our local model, establishing density.

Fix n1n\geq 1. The local model will be the graph of a convex function of d1d-1 real variables. Let χ:d1[0,1]\chi\colon\mathbb{R}^{d-1}\to[0,1] be a smooth compactly-supported bump function with the properties that χ(x)=1\chi(x)=1 for x21\|x\|_{2}\leq 1 and χ(x)=0\chi(x)=0 for x22\|x\|_{2}\geq 2 (as well as 0χ(x)10\leq\chi(x)\leq 1 for all xx). Choose a constant h>0h>0 small enough such that the function

ρ(x1,x2,,xd1)=x12+x22++xd12+hχ(x)cos(πnx1)\displaystyle\rho(x_{1},x_{2},\ldots,x_{d-1})=x_{1}^{2}+x_{2}^{2}+\ldots+x_{d-1}^{2}+h\chi(x)\cos(\pi nx_{1})

is convex. This is possible since the Hessian of x12++xd12x_{1}^{2}+\ldots+x_{d-1}^{2} is twice the identity matrix, 2Id12I_{d-1}, while the Hessian of χ(x)cos(πnx1)\chi(x)\cos(\pi nx_{1}) is entry-wise bounded. Choosing hh small enough, the Hessian can be made to be positive definite everywhere. Let us denote the graph of ρ\rho over the ball of radius 4 by Σ0=Σ0(n)\Sigma_{0}=\Sigma_{0}(n), i.e.,

Σ0(n)={(x;ρ(x))d:xd1 with x24}.\Sigma_{0}(n)=\{(x;\rho(x))\in\mathbb{R}^{d}:x\in\mathbb{R}^{d-1}\text{ with }\|x\|_{2}\leq 4\}.
Lemma 5.1.

For d3d\geq 3 and n1n\geq 1, let B0dB_{0}\in\mathcal{B}_{d} be a unit ball such that B0\partial B_{0} contains the image of Σ0(n)\Sigma_{0}(n) under an invertible affine transformation. Then there exists ε>0\varepsilon>0 such that, for every BdB\in\mathcal{B}_{d} with dH(B0,B)<εd_{H}(B_{0},B)<\varepsilon, the unit distance graph of B{\lVert\cdot\rVert}_{B} contains a copy of Kd,2nK_{d,2n}.

Proof.

The conclusion is equivalent to the existence of dd translates of B\partial B which intersect in 2n2n points.

Let p1,,pd1{0}×d2d1p_{1},\ldots,p_{d-1}\in\{0\}\times\mathbb{R}^{d-2}\subset\mathbb{R}^{d-1} be affinely independent points which satisfy 2<pj2<32<\|p_{j}\|_{2}<3 for all j=1,,d1j=1,\ldots,d-1. Now consider the d1d-1 translates of Σ0\Sigma_{0} defined by

Σj=Σ0(pj;ρ(pj))\Sigma_{j}=\Sigma_{0}-(p_{j};\rho(p_{j}))

for j=1,,d1j=1,\ldots,d-1. We can compute that the intersection of these dd surfaces contains the following 2n2n points:

0jd1Σj{(2k+12n,0,,0,(2k+12n)2):nk<n}.\bigcap_{0\leq j\leq d-1}\Sigma_{j}\supset\left\{\left(\frac{2k+1}{2n},0,\ldots,0,\left(\frac{2k+1}{2n}\right)^{2}\right):-n\leq k<n\right\}.

We now summarize the remainder of the proof. The normal vectors to Σ0,,Σd1\Sigma_{0},\ldots,\Sigma_{d-1} at each of these intersection points can be computed explicitly; they are linearly independent. In other words, each of these 2n2n intersections is transversal. We can conclude by the well-known fact that transversal intersections persist under small perturbations. We will give an elementary deduction of this fact in our setting from the degree theory described above.

Refer to caption
Figure 3. The construction from Lemma 5.1 in three dimensions. Recall that Σ1\Sigma_{1} and Σ2\Sigma_{2} are the translates of Σ0\Sigma_{0} which send (p1;ρ(p1))(p_{1};\rho(p_{1})) resp. (p2;ρ(p2))(p_{2};\rho(p_{2})) to (0,0)(0,0). The two parabolas through (p1;ρ(p1))(p_{1};\rho(p_{1})) and (p2;ρ(p2))(p_{2};\rho(p_{2})) lie on Σ0\Sigma_{0}, hence Σ1Σ2\Sigma_{1}\cap\Sigma_{2} is the parabola through (0;0)(0;0). The surface Σ0\Sigma_{0} is “crinkled” near the origin so that this parabola cuts it transversally, creating transversal intersections of Σ0\Sigma_{0}, Σ1\Sigma_{1} and Σ2\Sigma_{2}.

For simplicity, we will perform the computation with a specific choice of B0B_{0}, namely

B0={(x;y)d:|y|16ρ(x)}.B_{0}=\{(x;y)\in\mathbb{R}^{d}:|y|\leq 16-\rho(x)\}.

The only section of B0B_{0} we will make reference to is the portion near the translated copy of Σ0\Sigma_{0} in B0\partial B_{0}. So, this specification has the benefit that we may work in a simple coordinate system relative to the coordinates in which we have defined Σ0\Sigma_{0}. All of the following arguments go through identically for any ball B0B_{0} whose boundary contains an affine image of Σ0\Sigma_{0} once an appropriate coordinate transformation is applied.

Set q0=(0;0)q_{0}=(0;0) and qj=(pj;ρ(pj))q_{j}=(p_{j};\rho(p_{j})) for 1jd11\leq j\leq d-1. Since B0\partial B_{0} contains the translated surface Σ0(0;16)\Sigma_{0}-(0;16), we see that B0qj\partial B_{0}-q_{j} contains the translated surfaces Σj(0;16)\Sigma_{j}-(0;16). Thus the computation above shows that the dd translates of B0q0,,B0qd1\partial B_{0}-q_{0},\ldots,\partial B_{0}-q_{d-1} intersect in 2n2n points. We must show that there exists ε>0\varepsilon>0 such that the same is true for any other unit ball which is ε\varepsilon-close to B0B_{0} in Hausdorff distance. To begin, for each BdB\in\mathcal{B}_{d}, define the continuous function ΦB:dd\Phi_{B}\colon\mathbb{R}^{d}\to\mathbb{R}^{d} by

ΦB(x)=(x+qjB1)j=0d1.\Phi_{B}(x)=\left(\|x+q_{j}\|_{B}-1\right)_{j=0}^{d-1}.

Fix nk<n-n\leq k<n. Then set x0=(2k+12n,0,,0,(2k+12n)216)x_{0}=\left(\frac{2k+1}{2n},0,\ldots,0,(\frac{2k+1}{2n})^{2}-16\right). By construction, ΦB0(x0)=0\Phi_{B_{0}}(x_{0})=0. In a neighborhood of x0x_{0}, the function ΦB0\Phi_{B_{0}} inherits smoothness from ρ\rho. We now are in a position to apply property (3) to compute the degree of ΦB0\Phi_{B_{0}}; to do this, we need to compute the Jacobian matrix of ΦB0\Phi_{B_{0}} at x0x_{0}.

For any point xΣj(0;16)x\in\Sigma_{j}-(0;16), one can see that the gradient of the function x+qjB01\|x+q_{j}\|_{B_{0}}-1 is a non-zero multiple of the normal vector to Σj(0;16)\Sigma_{j}-(0;16) at that point. Furthermore, since Σ0(0;16)\Sigma_{0}-(0;16) is the graph of the function ρ16\rho-16, we see that the normal vector to Σ0(0;16)\Sigma_{0}-(0;16) at the point (x,ρ(x)16)(x,\rho(x)-16) is a non-zero multiple of (ρ(x);1)(\nabla\rho(x);-1). In particular, at x0x_{0} we can compute that this is

ν0:=((2k+1n(1)khπn)e1;1).\nu_{0}:=\left(\left(\frac{2k+1}{n}-(-1)^{k}h\pi n\right)e_{1};-1\right).

For 1jd11\leq j\leq d-1, using the fact that Σj\Sigma_{j} is a translate of Σ0\Sigma_{0}, we compute that the normal to Σj(0;16)\Sigma_{j}-(0;16) at x0x_{0} is νj:=((2k+1)/ne1+2pj;1)\nu_{j}:=((2k+1)/n\,e_{1}+2p_{j};-1). Thus we see that the Jacobian matrix of ΦB0\Phi_{B_{0}} at x0x_{0} has columns which are non-zero multiples of ν0,,νd1\nu_{0},\ldots,\nu_{d-1}. The vectors ν0,,νd1\nu_{0},\ldots,\nu_{d-1} are linearly independent: the affine span of ν1,,νd1\nu_{1},\ldots,\nu_{d-1} is the (d2)(d-2)-flat defined by x1=(2k+1)/nx_{1}=(2k+1)/n and xd=1x_{d}=-1, so the linear span of these vectors is the hyperplane x1+(2k+1)/nxd=0x_{1}+(2k+1)/n\,x_{d}=0, and ν0\nu_{0} is not in this hyperplane. Therefore, the Jacobian of ΦB0\Phi_{B_{0}} does not vanish at x0x_{0}.

We now apply the inverse function theorem to ΦB0\Phi_{B_{0}}. Since ΦB0(x)\Phi_{B_{0}}(x) is differentiable with continuous derivative and its Jacobian does not vanish at x0x_{0}, there exists an open neighborhood UU of x0x_{0} so that ΦB0:U¯d\Phi_{B_{0}}\colon\overline{U}\to\mathbb{R}^{d} is injective. We can pick UU sufficiently small such that U¯\overline{U} does not contain (2k+12n,0,,0,(2k+12n)216)\left(\frac{2k^{\prime}+1}{2n},0,\ldots,0,(\frac{2k^{\prime}+1}{2n})^{2}-16\right) for any kkk^{\prime}\neq k. By property (3) of degree, we see that deg(ΦB0,U,0)=±1\deg(\Phi_{B_{0}},U,0)=\pm 1, since 0 has precisely one preimage in UU.

Now, choose ε>0\varepsilon>0 small enough that ΦB(y)ΦB0(y)2<ΦB0(y)2\|\Phi_{B}(y)-\Phi_{B_{0}}(y)\|_{2}<\|\Phi_{B_{0}}(y)\|_{2} for all yUy\in\partial U and all BdB\in\mathcal{B}_{d} with dH(B,B0)<εd_{H}(B,B_{0})<\varepsilon. This is possible since U\partial U is compact, 0ΦB0(U)0\not\in\Phi_{B_{0}}(\partial U), and xB\|x\|_{B} is a continuous function of (x,B)d×d(x,B)\in\mathbb{R}^{d}\times\mathcal{B}_{d}. By property (2) of degree, we see that deg(ΦB,U,0)=deg(ΦB0,U,0)\deg(\Phi_{B},U,0)=\deg(\Phi_{B_{0}},U,0) for all BdB\in\mathcal{B}_{d} with dH(B,B0)<εd_{H}(B,B_{0})<\varepsilon. In particular, this degree is nonzero, so by property (1) of degree, we see that there exists xUx\in U with ΦB(x)=0\Phi_{B}(x)=0. By definition, this point xx is in the intersection of the dd translates of B\partial B centered at q0,,qd1q_{0},\ldots,q_{d-1}. Repeating this argument for each nk<n-n\leq k<n, we find 2n2n distinct points in the intersection of these dd translates of B\partial B. ∎

Proof of 3.3.

Set n=m/2n=\left\lceil m/2\right\rceil. Let 𝒳nd\mathcal{X}_{n}\subseteq\mathcal{B}_{d} be the set of unit balls whose boundary contains, for some δ>0\delta>0, a translated copy of the scaled surface

Σδ:={(x;δ3ρ(x/δ))d:xd1 with x24δ}.\Sigma_{\delta}:=\{(x;\delta^{3}\rho(x/\delta))\in\mathbb{R}^{d}:x\in\mathbb{R}^{d-1}\text{ with }\|x\|_{2}\leq 4\delta\}.

By Lemma 5.1, for each B𝒳nB\in\mathcal{X}_{n}, there exists some εB>0\varepsilon_{B}>0 such that every BdB^{\prime}\in\mathcal{B}_{d} with dH(B,B)<εBd_{H}(B,B^{\prime})<\varepsilon_{B} contains a copy of Kd,2nK_{d,2n} in its unit distance graph. We define the union of these open neighborhoods

𝒜m={Bd:there exists B𝒳n such that dH(B,B)<εB}.\mathcal{A}_{m}=\{B^{\prime}\in\mathcal{B}_{d}:\text{there exists }B\in\mathcal{X}_{n}\text{ such that }d_{H}(B,B^{\prime})<\varepsilon_{B}\}.

Clearly 𝒜md\mathcal{A}_{m}\subseteq\mathcal{B}_{d} is open. To complete the proof, we must show that it is dense.

The prefactor δ3\delta^{3} is chosen so that δ3ρ(x/δ)\delta^{3}\rho(x/\delta) converges uniformly to the zero function on compact sets as δ0\delta\to 0. More precisely, we have the bounds |δ3ρ(x/δ)δx22|hδ3|\delta^{3}\rho(x/\delta)-\delta\|x\|_{2}^{2}|\leq h\delta^{3}. Recall that we defined ρ\rho so that infρ=h\inf\rho=h where h>0h>0 is taken very small. Then, for δ,R>0\delta,R>0, define the set

Xδ,R={(x;y)d:x2R and yδ3(ρ(x/δ)h).}X_{\delta,R}=\{(x;y)\in\mathbb{R}^{d}:\|x\|_{2}\leq R\text{ and }y\geq\delta^{3}(\rho(x/\delta)-h).\}

The above calculation shows that as δ0\delta\to 0, the sets Xδ,RX_{\delta,R} all live in the upper half-space and converge to the cylinder {(x;y)d:x2R and y0}\{(x;y)\in\mathbb{R}^{d}:\|x\|_{2}\leq R\text{ and }y\geq 0\} in Hausdorff distance.

Let BdB\in\mathcal{B}_{d} be an arbitrary unit ball. For any ε>0\varepsilon>0, we will find an ε\varepsilon-close element of 𝒳n\mathcal{X}_{n}. Suppose that BB has height 2y0>02y_{0}>0 in the xdx_{d}-direction. In other words, Bd1×[y0,y0]B\subset\mathbb{R}^{d-1}\times[-y_{0},y_{0}] and there exists some x0d1x_{0}\in\mathbb{R}^{d-1} such that (x0;y0),(x0;y0)B(x_{0};-y_{0}),(-x_{0};y_{0})\in\partial B. First we chop a thin slice off the bottom and top of BB. In particular, we can pick 0<y1<y00<y_{1}<y_{0} such that B:=B(d1×[y1,y1])B^{\prime}:=B\cap(\mathbb{R}^{d-1}\times[-y_{1},y_{1}]) is ε/2\varepsilon/2-close to BB. Since BB is convex and contains a neighborhood of 0, it contains a cone with apex (x0;y0)(x_{0};-y_{0}) and base centered at the origin. Thus BB^{\prime} contains a frustum with bases on the hyperplanes d1×{y1}\mathbb{R}^{d-1}\times\{-y_{1}\} and d1×{0}\mathbb{R}^{d-1}\times\{0\}. Next, we can find a small right cylinder CC in BB^{\prime} with one base on the hyperplane d1×{y1}\mathbb{R}^{d-1}\times\{-y_{1}\}. Say its bases have radius 4δ04\delta_{0}, the lower base is centered at (x1;y1)(x_{1};-y_{1}), and the height is (16h)δ03(16-h)\delta_{0}^{3}, i.e., C={(x;y):xx124δ0 and y[y1,y1+(16h)δ03]}C=\{(x;y):\|x-x_{1}\|_{2}\leq 4\delta_{0}\text{ and }y\in[-y_{1},-y_{1}+(16-h)\delta_{0}^{3}]\}. See Fig. 4 for an illustration of this in two dimensions.

CC(x0;y0)(x_{0};-y_{0})(x1;y1)(x_{1};-y_{1})(x0;y0)(-x_{0};y_{0})BBBB^{\prime}
Figure 4. Construction of CC inside of BB^{\prime}.

Pick RR so that BB^{\prime} is contained within the Euclidean ball of radius R/2R/2. Then for each 0<δ<δ00<\delta<\delta_{0}, we modify BB^{\prime} to place a copy of Σδ\Sigma_{\delta} inside the cylinder CC. More precisely, define the set

Bδ=B(Xδ,R+(x1;y1))(Xδ,R(x1;y1)).B_{\delta}=B^{\prime}\cap(X_{\delta,R}+(x_{1};-y_{1}))\cap(-X_{\delta,R}-(x_{1};-y_{1})).

Note that BδB_{\delta} is clearly still a unit ball for all δ>0\delta>0. Furthermore, Bδ𝒳nB_{\delta}\in\mathcal{X}_{n} for all δ(0,δ0)\delta\in(0,\delta_{0}) since Bδ\partial B_{\delta} contains a translate of Σδ\Sigma_{\delta} in the cylinder CC. Finally, we claim that BδB_{\delta} converges to BB^{\prime} in Hausdorff distance as δ0\delta\to 0. This is because Xδ,R+(x1;y1)X_{\delta,R}+(x_{1};-y_{1}) converges to the cylinder {(x;y):xx12R and yy1}\{(x;y):\|x-x_{1}\|_{2}\leq R\text{ and }y\geq-y_{1}\} and Xδ,R(x1;y1)-X_{\delta,R}-(x_{1};-y_{1}) converges to the cylinder {(x;y):x+x12R and yy1}\{(x;y):\|x+x_{1}\|_{2}\leq R\text{ and }y\leq y_{1}\}. We chose R,x1,y1R,x_{1},y_{1} so that BB^{\prime} is contained in the intersection of these two cylinders. Thus BδB_{\delta} converges to BB^{\prime} as δ0\delta\to 0, so there exists some choice of δ(0,δ0)\delta\in(0,\delta_{0}) so that dH(Bδ,B)<ε/2d_{H}(B_{\delta},B^{\prime})<\varepsilon/2. For this δ\delta we have dH(B,Bδ)<εd_{H}(B,B_{\delta})<\varepsilon and Bδ𝒳nB_{\delta}\in\mathcal{X}_{n}. This proves that 𝒜m\mathcal{A}_{m} is open and dense, as desired. ∎

Remark.

It is possible to use Theorem 1.5 to provide a different proof of Theorem 1.4 for d3d\geq 3 and a comeagre set of dd-norms. Indeed, consider a Kd,2d2mK_{d,2d^{2}m} in the unit distance graph of B{\lVert\cdot\rVert}_{B}. Let u1,,udu_{1},\ldots,u_{d} be the vertices on the left. By greedily selecting vertices, one can find v1,,vmv_{1},\ldots,v_{m} on the right so that {viuj:i[m],j[d]}\{v_{i}-u_{j}:i\in[m],j\in[d]\} is a non-overlapping set of dmdm unit vectors. Then a similar construction to 4.6 produces a set of points spanning many unit distances.

References

  • [1] Noga Alon, Matija Bucić, and Lisa Sauermann, Unit and distinct distances in typical norms, Geom. Funct. Anal. 35 (2025), 1–42.
  • [2] Ryszard Engelking, Dimension theory, North-Holland Mathematical Library, vol. 19, North-Holland Publishing Co., Amsterdam-Oxford-New York; PWN—Polish Scientific Publishers, Warsaw, 1978, Translated from the Polish and revised by the author.
  • [3] P. Erdős, On sets of distances of nn points, Amer. Math. Monthly 53 (1946), 248–250.
  • [4] Peter M. Gruber, Convex and discrete geometry, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 336, Springer, Berlin, 2007.
  • [5] Á. G. Horváth, On shadow boundaries of centrally symmetric convex bodies, Beiträge Algebra Geom. 50 (2009), 219–233.
  • [6] Jiří Matoušek, The number of unit distances is almost linear for most norms, Adv. Math. 226 (2011), 2618–2628.
  • [7] James R. Munkres, Topology, second ed., Prentice Hall, Inc., Upper Saddle River, NJ, 2000.
  • [8] J. Spencer, E. Szemerédi, and W. Trotter, Jr., Unit distances in the Euclidean plane, Graph theory and combinatorics (Cambridge, 1983), Academic Press, London, 1984, pp. 293–303.
  • [9] László A. Székely, Crossing numbers and hard Erdős problems in discrete geometry, Combin. Probab. Comput. 6 (1997), 353–358.
  • [10] Gerald Teschl, Topics in Linear and Nonlinear Functional Analysis, Graduate Studies in Mathematics, Amer. Math. Soc., Providence, to appear, available at https://www.mat.univie.ac.at/gerald/ftp/book-fa/.
  • [11] Pavel Valtr, Strictly convex norms allowing many unit distances and related touching questions, 2005, manuscript.