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

Colorful Helly via induced matchings

Cosmin Pohoata Kevin Yang  and  Shengtong Zhang
Abstract.

We establish a theorem regarding the maximum size of an induced matching in the bipartite complement of the incidence graph of a set system (X,)(X,\mathscr{F}). We show that this quantity plus one provides an upper bound on the colorful Helly number of this set system, i.e. the minimum positive integer NN for which the following statement holds: if finite subfamilies 1,,N\mathscr{F}_{1},\ldots,\mathscr{F}_{N}\subset\mathscr{F} are such that FiF=0\cap_{F\in\mathscr{F}_{i}}F=0 for every i=1,,Ni=1,\ldots,N, then there exists FiiF_{i}\in\mathscr{F}_{i} such that F1FN=F_{1}\cap\ldots\cap F_{N}=\emptyset. We will also discuss some natural refinements of this result and applications.

Department of Mathematics, Emory University, Atlanta, GA. Email: cosmin.pohoata@emory.edu. Research supported by NSF Award DMS-2246659.
Skyline High School, Sammamish, WA. Email: yangkev27@issaquah.wednet.edu.
Department of Mathematics, Stanford University, Stanford, CA. Email: stzh1555@stanford.edu.

1. Introduction

In 1913, Helly famously proved that for every collection of convex sets C1,,CnC_{1},\ldots,C_{n} in d\mathbb{R}^{d} with an empty intersect, there exists a set S[n]S\subset[n] of size |S|d+1|S|\leqslant d+1 with iSCi=\bigcap_{i\in S}C_{i}=\emptyset. Moreover, d+1d+1 is the smallest number with this property. This result is known as Helly’s theorem and is fundamental result in combinatorial geometry, with a large number of generalizations and applications over the years. See for example [5], [9], [12], [24], and the references therein. We also recommend the surveys [3] and [6], focusing on more recent developments. To state some of these different variants of Helly’s theorem, it is often standard to introduce the following terminology. For a set XX and a family \mathscr{F} of subsets of XX, the Helly number h=h(X,)h=h(X,\mathscr{F})\in\mathbb{N} of \mathscr{F} is defined as the smallest integer hh such that for any finite subfamily KK of \mathscr{F}, if every subset of at most hh members of \mathscr{F} has a nonempty intersection then all sets in KK have a nonempty intersection. If no such hh exists, then h(X,):=h(X,\mathscr{F}):=\infty. Helly’s theorem above then asserts that the Helly number of the family of convex sets in d\mathbb{R}^{d} is d+1d+1. Another example of a known Helly number is the Helly number of spheres in d\mathbb{R}^{d}. In [23], Maehara proved that this equals d+2d+2. An additional example is the theorem of Doignon [11] which asserts that the Helly number of convex lattice sets in d\mathbb{R}^{d}, that is sets of the form CdC\cap\mathbb{Z}^{d} where CC is a convex set in d\mathbb{R}^{d} is 2d2^{d}. A combinatorial example is the fact that the Helly number of the collection of (sets of vertices of) subtrees of any given tree is 22.

An algebraic example that in some sense motivated the present paper is the fact that for any field 𝔽\mathbb{F} the Helly number of hypersurfaces of degree at most DD in 𝔽d\mathbb{F}^{d} is (D+dd){D+d\choose d}. This was proved in 1986 by Deza and Frankl [10] in the following elegant way: for any set of polynomials f1,,fnf_{1},\ldots,f_{n} of degree D\leqslant D in 𝔽[x1,,xd]\mathbb{F}[x_{1},\ldots,x_{d}] with the property that their zero sets Z(fi):={x𝔽d:fi(x)=0}Z(f_{i}):=\left\{x\in\mathbb{F}^{d}:f_{i}(x)=0\right\} satisfy i=1nZ(fi)=\bigcap_{i=1}^{n}Z(f_{i})=\emptyset, consider a minimal subset S[n]S\subset[n] such that iSZ(fi)=\bigcap_{i\in S}Z(f_{i})=\emptyset. By definition, this is a set SS which must also satisfy the property that for every jSj\in S, the intersection iS{j}Z(fi)\bigcap_{i\in S\setminus\{j\}}Z(f_{i}) is nonempty. In other words, for every jSj\in S there exists xj𝔽dx_{j}\in\mathbb{F}^{d} such that:

(1) fi(xj)0if and only ifij.f_{i}(x_{j})\neq 0\ \text{if and only if}\ i\neq j.

By a (nowadays) standard argument, it is not difficult to see that this condition implies that the set of polynomials {fi}iS\left\{f_{i}\right\}_{i\in S} is a set of linearly independent polynomials in 𝔽[x1,,xd]\mathbb{F}[x_{1},\ldots,x_{d}]. This yields the estimate |S|(D+dd)|S|\leqslant\binom{D+d}{d}, since (D+dd)\binom{D+d}{d} equals precisely the dimension of the vector space of polynomials of degree at most DD in 𝔽[x1,,xd]\mathbb{F}[x_{1},\ldots,x_{d}]. In [13], Frankl subsequently extended this argument to varieties of bounded degree. It is perhaps instructive to highlight at this point that a similar strategy cannot be used to prove Helly’s original theorem for convex sets. In particular, if d2d\geqslant 2 and C1,,CnC_{1},\ldots,C_{n} is a collection of convex sets in d\mathbb{R}^{d} such that i=1nCi=\bigcap_{i=1}^{n}C_{i}=\emptyset, then it is no longer true that the size of a minimal set S[n]S\subset[n] with the property that iSCi=\bigcap_{i\in S}C_{i}=\emptyset has to be bounded independently of nn (rather than a minimum set). Indeed, note that while for every jSj\in S there exists a point xjdx_{j}\in\mathbb{R}^{d} such that

(2) xjCiif and only ifij,x_{j}\not\in C_{i}\ \text{if and only if}\ i\neq j,

this condition no longer implies that |S|=Od(1)|S|=O_{d}(1) (let alone |S|d+1|S|\leqslant d+1).111Indeed, note that for any d2d\geqslant 2 and any arbitrarily large m1m\geqslant 1 one can construct mm convex sets C1,,CmdC_{1},\,\ldots,\,C_{m}\in\mathbb{R}^{d} such that for every i[m]i\in[m] there exists xix_{i} such that xijiCjCix_{i}\in\bigcap_{j\neq i}C_{j}\setminus C_{i}. Let PP be a convex polytope with at least mm facets. Pick mm facets f1,,fmf_{1},\,\ldots,\,f_{m} and let CiC_{i} be the half-space bounded by the hyperplane supporting fif_{i} and containing PP. Let xix_{i} be any point in the polytope determined by the hyperplane supporting fif_{i}, and the hyperplane supporting all the neighbor facets of fif_{i}. By construction, xiCjx_{i}\in C_{j} if and only if iji\neq j. Nevertheless, this type of argument turns out to be quite powerful in settings where conditions like (1) or (2) do generate a set of |S||S| linear independent objects in a vector space of bounded dimension. For example, in a recent nice paper [1], Alon, Jin, and Sudakov also used this strategy to determine that the Helly number of the set of Hamming balls of radius tt inside an nn-dimensional space XnX^{n} is 2t+12^{t+1} (where XX is an arbitrary finite or infinite set).

In this paper, we generalize this story by showing that in a set system exhibiting conditions like (1) or (2) one can also establish an optimal colorful Helly theorem. To state things more precisely going forward, it will be convenient to consider the following abstract setting. Let XX be a set and let \mathscr{F} be a collection of subsets of elements from XX. We define the colorful Helly number of the set system (X,)(X,\mathscr{F}) to be the minimum positive integer η=η(X,)\eta=\eta(X,\,\mathscr{F}) for which the following statement holds: if finite subfamilies 1,,η\mathscr{F}_{1},\,\ldots,\,\mathscr{F}_{\eta}\subset\mathscr{F} are such that FiF=\cap_{F\in\mathscr{F}_{i}}F=\emptyset for every i=1,,ηi=1,\,\ldots,\,\eta, then there exists FiiF_{i}\in\mathscr{F}_{i} such that F1Fη=F_{1}\cap\cdots\cap F_{\eta}=\emptyset.

Like before, if no such η\eta exists, then we say η(X,):=\eta(X,\,\mathscr{F}):=\infty. It is a celebrated theorem of Lovász (whose proof first appeared in a paper by Bárány [4]) that if X=dX=\mathbb{R}^{d} and \mathscr{F} is the family of convex sets in d\mathbb{R}^{d} then η(X,)=d+1\eta(X,\,\mathscr{F})=d+1. This generalization of Helly’s theorem has managed to inspire a lot of further interesting research directions on its own, due to its rather mysterious geometric nature. For example, while clearly h(X,)η(X,)h(X,\mathscr{F})\leqslant\eta(X,\mathscr{F}), it is not always the case that h(X,)=η(X,)h(X,\mathscr{F})=\eta(X,\mathscr{F}). Let X=dX=\mathbb{R}^{d} and let \mathscr{F} consist of a family of axis-parallel boxes: it is a folklore result that in this case h(X,)=2h(X,\mathscr{F})=2, while it is not difficult to see that η(X,)=d+1\eta(X,\mathscr{F})=d+1.222Lovász’s theorem implies η(X,)d+1\eta(X,\mathscr{F})\leqslant d+1, whereas η(X,)d+1\eta(X,\mathscr{F})\geqslant d+1 follows by considering the dd-coloring of the facets of a hypercube in d\mathbb{R}^{d} so that each opposite pair of facets carry the same color. This is a coloring with dd color classes 1,,d\mathscr{F}_{1},\ldots,\mathscr{F}_{d} where FiF=0\cap_{F\in\mathscr{F}_{i}}F=0 for every i=1,,ηi=1,\,\ldots,\,\eta, but such that F1FdF_{1}\cap\cdots\cap F_{d}\neq\emptyset for every FiiF_{i}\in\mathscr{F}_{i}. It is therefore natural to ask which Helly-type theorems admit appropriate colorful versions (see for example the survey of Amenta, De Loera, and Soberón [3, Section 2.2] for more context).

Our results below address this question by providing several natural sufficient conditions for the existence of a colorful Helly theorem. We will work with the incidence graph of XX and \mathscr{F}, which is the bipartite graph GX,G_{X,\,\mathscr{F}} on XX\cup\mathscr{F} where (x,F)X×(x,\,F)\in X\times\mathscr{F} is an edge if xFx\in F. The bipartite complement G¯X,\overline{G}_{X,\,\mathscr{F}} of GX,G_{X,\,\mathscr{F}} denotes the graph with vertex set XX\cup\mathscr{F} where (x,F)X×(x,\,F)\in X\times\mathscr{F} is an edge if xFx\not\in F. We will also be concerned with the size τ=τ(X,)\tau=\tau(X,\,\mathscr{F}) of the maximum induced matching inside G¯X,\overline{G}_{X,\,\mathscr{F}}. We shall sometimes refer to a matching in G¯X,\overline{G}_{X,\,\mathscr{F}} as a comatching of the set system (X,)(X,\mathscr{F}), and to τ(X,)\tau(X,\,\mathscr{F}) as the size of the largest comatching in (X,)(X,\mathscr{F}) (or simply, the comatching number of (X,)(X,\mathscr{F})). We are now ready to state our first result.

Theorem 1.1.

Let XX be a set and let \mathscr{F} be a collection of subsets of XX. Then we have

η(X,)1+τ(X,).\eta(X,\,\mathscr{F})\leqslant 1+\tau(X,\,\mathscr{F}).

As we will see in Section 2, this result is optimal and even the additive +1+1 term is necessary. Furthermore, in certain situations we can remove the assumption that the i\mathscr{F}_{i}’s are finite. Let η(X,)\eta^{\infty}(X,\mathscr{F}) be the colorful Helly number with the finiteness assumption on i\mathscr{F}_{i} removed. Clearly, we have η(X,)η(X,)\eta(X,\mathscr{F})\leqslant\eta^{\infty}(X,\mathscr{F}).

We say \mathscr{F} is Noetherian if for any sequence of sets F1,F2,F_{1},F_{2},\ldots in \mathscr{F}, the descending chain of sets

F1F1F2F1F2F3F_{1}\supset F_{1}\cap F_{2}\supset F_{1}\cap F_{2}\cap F_{3}\supset\cdots

stabilizes. For example, we have

  • If XX is finite, then \mathscr{F} is Noetherian.

  • If XX is n\mathbb{C}^{n} or n\mathbb{R}^{n}, and \mathscr{F} is any family of algebraic varieties in XX, then \mathscr{F} is Noetherian due to the Nullstellensatz and the fact that [x1,,xn]\mathbb{C}[x_{1},\ldots,x_{n}] is a Noetherian ring.

  • If XX is 2\mathbb{R}^{2}, and \mathscr{F} is the family of convex sets, then \mathscr{F} not Noetherian. We can check that η(X,)=3\eta(X,\mathscr{F})=3 but η(X,)=\eta^{\infty}(X,\mathscr{F})=\infty.

If \mathscr{F} is Noetherian, then the finiteness assumption on i\mathscr{F}_{i} is not necessary. While this observation is not hard to prove, we are unaware of any literature that discusses this subtlety.

Proposition 1.2.

If a set family \mathscr{F} on XX is Noetherian, then we have

η(X,)=η(X,).\eta^{\infty}(X,\mathscr{F})=\eta(X,\mathscr{F}).

In addition, we can consider a slightly stronger notion of τ(X,)\tau(X,\mathscr{F}). Let τ(X,)\tau^{\prime}(X,\mathscr{F}) be the largest τ\tau^{\prime} such that there exist x1,,xτ+1Xx_{1},\ldots,x_{\tau^{\prime}+1}\in X and F1,,FτF_{1},\ldots,F_{\tau^{\prime}} such that xiFjx_{i}\in F_{j} if and only if iji\neq j. We call this a comatching with intersection, and xτ+1x_{\tau^{\prime}+1} the common point. Clearly, we have

τ(X,){τ(X,)1,τ(X,)}.\tau^{\prime}(X,\mathscr{F})\in\{\tau(X,\mathscr{F})-1,\tau(X,\mathscr{F})\}.

Using this terminology, we will show the following refinement of Theorem 1.1, which is in some sense our main result.

Theorem 1.3.

For any set family \mathscr{F} on XX, we have

η(X,)1+τ(X,).\eta(X,\mathscr{F})\leqslant 1+\tau^{\prime}(X,\mathscr{F}).

In particular, if τ(X,)=τ(X,)1\tau^{\prime}(X,\mathscr{F})=\tau(X,\mathscr{F})-1, then η(X,)=1+τ(X,)=τ(X,)\eta(X,\mathscr{F})=1+\tau^{\prime}(X,\mathscr{F})=\tau(X,\mathscr{F}).

Theorem 1.3 has a few new consequences, which we will discuss in detail in Section 3. First, we will argue that Theorem 1.3 implies that the colorful Helly number for dd-dimensional spheres in d\mathbb{R}^{d} is d+2d+2, generalizing the work of Maehara [23]. Second, we will show that Theorem 1.3 establishes an optimal colorful Helly theorem for any set of hypersurfaces of bounded degree, generalizing the works of Motzkin [26] and Deza-Frankl [10]. Last but not least, we will show that the colorful Helly number of the set of Hamming balls of radius tt inside an nn-dimensional space XnX^{n} is 2t+12^{t+1}, generalizing the work of Alon, Jin, and Sudakov [1]. Turns out, it is also possible to prove the first two consequences of Theorem 1.3 using some more ad-hoc arguments. These arguments have some shortcomings, but we considered them to be quite instructive so we will discuss them in Section 2.

It is also natural to wonder whether set systems with bounded comatching number satisfy a so-called fractional Helly theorem. In the context of Helly’s theorem for convex sets, this theorem was first established by Katchalski and Liu in [19]: if \mathscr{F} is a family of nd+1n\geqslant d+1 convex sets in d\mathbb{R}^{d} such that the number of (d+1)(d+1)-tuples of \mathscr{F} with non-empty intersection is at least α(nd+1)\alpha{n\choose d+1} for some constant α>0\alpha>0, then there are at least βn\beta n members of \mathscr{F} whose intersection is non-empty, where β>0\beta>0 is a constant which depends only on α\alpha and dd. The optimal choice β=1(1α)1d+1\beta=1-(1-\alpha)^{\frac{1}{d+1}} was determined by Kalai [17]. In [15], Holmsen showed that the the fractional Helly theorem can be derived as a purely combinatorial consequence of the colorful Helly theorem, and then later Holmsen and Lee [16] used this result to prove that a fractional Helly theorem holds in all set systems with a bounded Radon number. We note that an abstract fractional Helly theorem for sets with bounded comatching number also holds.

Corollary 1.4.

Let (X,)(X,\,\mathscr{F}) be a system with finite comatching number τ\tau. For α(0, 1)\alpha\in(0,\,1) there exists β=β(α,τ)(0, 1)\beta=\beta(\alpha,\,\tau)\in(0,\,1) with the following property: Let FF be a subfamily of n>τ+1n>\tau+1 sets in \mathscr{F} and suppose at least α(nτ+1)\alpha\binom{n}{\tau+1} of the (τ+1)(\tau+1)-tuples in FF have nonempty intersection. Then there exists some βn\beta n members of FF whose intersection is non-empty.

While it is also possible to use the result [15] and derive a fractional Helly theorem using Theorem 1.3, we note that the statement from Corollary 1.4 (with the correct dependence on τ\tau) follows from the work of Matoušek [25]. Matoušek showed that the fractional Helly number of a set system is at most the dual VC dimension plus one. On the other hand, the dual VC-dimension of a set system is dominated by our parameter τ\tau^{\prime}: Suppose we are given a family AA\subset\mathscr{F} which is dual shattered by XX, meaning, for every BAB\subset A, there exists xbBBaAAx\in\bigcap_{b\in B}B\setminus\bigcap_{a\in A}A. Taking B:=A{a}B:=A\setminus\{a\} for aAa\in A and B:=AB:=A yields that AA is a comatching with intersection.

Ackowledgement

We thank Qiuyu Ren for numerous helpful discussions, and proving the d=2d=2 case of Proposition 4.2.


2. Warm up: Colorful Helly for hypersurfaces

In this section, we discuss some preliminary results that only require ad-hoc arguments (in hindsight). The first one concerns a colorful version of Helly’s theorem for spheres.

2.1. Spheres

In [23], Maehara proved that the Helly number for dd-dimensional spheres is d+2d+2. We show that the colorful Helly number is equal to the Helly number. We note that we also do not need the finiteness assumption on the families i\mathscr{F}_{i}.

Theorem 2.1 (Colorful Helly for spheres).

Let 1,,d+2\mathscr{F}_{1},\,\ldots,\,\mathscr{F}_{d+2} be families of spheres in d\mathbb{R}^{d}. If for every transversal {Sii:1id+2}\{S_{i}\in\mathscr{F}_{i}:1\leqslant i\leqslant d+2\}, the intersection i=1d+2Si\bigcap_{i=1}^{d+2}S_{i} is nonempty, then one of the i\mathscr{F}_{i} has nonempty intersection.

Proof.

We use the fact that given a dd-dimensional sphere AA and a (d1)(d-1)-dimensional sphere BB, they are either disjoint, tangent, AA contains BB, or ABA\cap B is a (d2)(d-2)-dimensional sphere.

Let S1S_{1} and S2S_{2} be spheres in 1\mathscr{F}_{1} and 2\mathscr{F}_{2}. Assume they are not tangent, as otherwise every S33S_{3}\in\mathscr{F}_{3} must contain this tangent point, so we are done. Then S1S_{1} and S2S_{2} intersect in a (d1)(d-1)-dimensional sphere Cd1C_{d-1}. If every S33S_{3}\in\mathscr{F}_{3} contains Cd1C_{d-1}, 3\mathscr{F}_{3} has a nonempty intersection and we are done. If there exists S33S_{3}\in\mathscr{F}_{3} that is tangent to Cd1C_{d-1}, every S44S_{4}\in\mathscr{F}_{4} must contain this tangent point, so we are done. Hence we may assume there exists S33S_{3}\in\mathscr{F}_{3} that intersects Cd1C_{d-1} in a (d2)(d-2)-dimensional sphere Cd2C_{d-2}. Continuing like this, we get spheres SiiS_{i}\in\mathscr{F}_{i} for 1id+11\leqslant i\leqslant d+1 that intersect in a point (0-dimensional sphere) C0C_{0}. Then all Sd+2d+2S_{d+2}\in\mathscr{F}_{d+2} must contain C0C_{0} so d+2\mathscr{F}_{d+2} has a nonempty intersection.  \Box

We would like to point out that Theorem 2.2 is an example where Theorem 1.1 by itself only yields a suboptimal result, as in this case η=τ=d+2\eta=\tau=d+2. On the other hand, we shall see in Section 3 that Theorem 1.3 does provide the right answer.

Proposition 2.2.

Let X:=dX:=\mathbb{R}^{d} and 𝒮dd\mathcal{S}_{d}\subset\mathbb{R}^{d} be the family of all dd-dimensional spheres. Then

τ(X,𝒮d)=d+2.\tau(X,\,\mathcal{S}_{d})=d+2.
Proof.

First, we show that τ(X,𝒮d)d+2\tau(X,\,\mathcal{S}_{d})\leqslant d+2 through a stronger statement: for positive integers dnd\leqslant n, X:=nX:=\mathbb{R}^{n}, and the family 𝒮dd\mathcal{S}_{d}\subset\mathbb{R}^{d} of all dd-dimensional spheres, we have τ(X,𝒮d)d+2\tau(X,\,\mathcal{S}_{d})\leqslant d+2.

We proceed with induction on dd. For d=0d=0, a sphere is a point. Clearly τ(X,𝒮0)=2\tau(X,\,\mathcal{S}_{0})=2.

Assume the statement for d1d-1. Assume for the sake of contradiction there exists d+3d+3 spheres S1,,Sd+3𝒮dS_{1},\,\ldots,\,S_{d+3}\in\mathcal{S}_{d} and d+3d+3 (distinct) points x1,,xd+3x_{1},\,\ldots,\,x_{d+3} such that xiSjx_{i}\in S_{j} if and only if iji\neq j. Consider the set Y={x1,,xd+2}Y=\{x_{1},\,\ldots,\,x_{d+2}\} of points on Sd+3S_{d+3}. Then SiS_{i} contains Y{xi}Y\setminus\{x_{i}\} for all ii. Let Si:=SiSd+3S_{i}^{\prime}:=S_{i}\cap S_{d+3} for 1id+21\leqslant i\leqslant d+2, which is a (d1)(d-1)-dimensional sphere. Then xiSjx_{i}\in S_{j}^{\prime} if and only if iji\neq j, so this forms an induced matching in G¯X,𝒮d1\overline{G}_{X,\,\mathcal{S}_{d-1}}. By the induction hypothesis, d+2(d1)+2d+2\leqslant(d-1)+2, a contradiction.

To see that τ(X,𝒮d)d+2\tau(X,\,\mathcal{S}_{d})\geqslant d+2 and thus that τ(X,𝒮d)=d+2\tau(X,\,\mathcal{S}_{d})=d+2, consider TT to be a regular dd-dimensional simplex. Let \mathscr{F} be the family of spheres consisting of each sphere centered at the vertices of TT and containing the center of TT, and the sphere circumscribing TT. It is easy to see that τ(X,)=d+2\tau(X,\,\mathscr{F})=d+2, and thus τ(X,𝒮d)d+2\tau(X,\,\mathcal{S}_{d})\geqslant d+2.

The construction for d=2d=2 is shown below.

Figure 1. Four spheres in 2\mathbb{R}^{2} forming a comatching.

\Box

We will rederive Theorem 2.2 as a consequence of Theorem 1.3 in Section 3.

2.2. Hypersurfaces

Another consequence of Theorem 1.3 is an optimal colorful Helly theorem for hypersurfaces of bounded degree over any field 𝔽\mathbb{F}, generalizing the works of Moktzin [26] and Deza-Frankl [10]. We will discuss this in Section 3, but for now we would like to include a separate argument inspired by the one of Lovász from [4], which only works in the case when 𝔽=\mathbb{F}=\mathbb{C} and which may be interesting for independent reasons. We would also like to highlight that in [14], De Loera et al also adapted this argument of Lovász in order to prove an optimal colorful version of Doignon’s theorem [11].

Theorem 2.3 (Colorful Helly for complex polynomials).

Let DD and dd be positive integers and m:=(D+dd)m:=\binom{D+d}{d}. If 1,,m[x1,,xd]\mathscr{F}_{1},\,\ldots,\,\mathscr{F}_{m}\subset\mathbb{C}[x_{1},\,\ldots,\,x_{d}] are finite sets of polynomials of degree at most DD such that

fiZ(f)=\bigcap_{f\in\mathscr{F}_{i}}Z(f)=\emptyset

for all i[m]i\in[m], then there exists {fii}i[m]\{f_{i}\in\mathscr{F}_{i}\}_{i\in[m]} such that iZ(fi)=\bigcap_{i}Z(f_{i})=\emptyset.

We will also require the following simple lemma.

Lemma 2.4.

Let VV be a \mathbb{C}-vector space and a,b,cVa,\,b,\,c\in V be distinct vectors. Let [b,c]:={tb+(1t)c:t}[b,\,c]:=\{tb+(1-t)c:t\in\mathbb{C}\}. If ac,bc0\langle a-c,\,b-c\rangle\neq 0, there exists d[b,c]d\in[b,\,c] such that ad<ac||a-d||<||a-c||.

Proof.

Let d(t):=tb+(1t)cd(t):=tb+(1-t)c. Then ad(t)=(ac)t(bc)a-d(t)=(a-c)-t(b-c) so

ad(t)2\displaystyle||a-d(t)||^{2} =(ad(t))(a¯d(t)¯)\displaystyle=(a-d(t))(\overline{a}-\overline{d(t)})
=((ac)t(bc))((ac¯)t¯(bc¯))\displaystyle=((a-c)-t(b-c))((\overline{a-c})-\overline{t}(\overline{b-c}))
=(ac)(ac¯)+tt¯(bc)(bc¯)t(ac¯)(bc)t¯(ac)(bc¯)\displaystyle=(a-c)(\overline{a-c})+t\overline{t}(b-c)(\overline{b-c})-t(\overline{a-c})(b-c)-\overline{t}(a-c)(\overline{b-c})
=ac2+t2bc22Re(t(ac¯)(bc)).\displaystyle=||a-c||^{2}+||t||^{2}||b-c||^{2}-2\mathrm{Re}(t(\overline{a-c})(b-c)).

Hence ad(t)<ac||a-d(t)||<||a-c|| if and only if t2bc2<2Re(t(ac¯)(bc))||t||^{2}||b-c||^{2}<2\mathrm{Re}(t(\overline{a-c})(b-c)). Note that (ac¯)(bc)(\overline{a-c})(b-c) is nonzero because ac,bc0\langle a-c,\,b-c\rangle\neq 0.

If (ac¯)(bc)(\overline{a-c})(b-c) has a real part r0r\neq 0, let tt be real. Then Re(t(ac¯)(bc))=rt\mathrm{Re}(t(\overline{a-c})(b-c))=rt so it suffices to prove that there exists tt such that t2bc2<2rtt^{2}||b-c||^{2}<2rt. Taking tt close enough to 0 works.

If (ac¯)(bc)(\overline{a-c})(b-c) can be written as riri for some real rr, let t=:sit=:si. Then Re(t(ac¯)(bc))=rs\mathrm{Re}(t(\overline{a-c})(b-c))=-rs so it suffices to prove that there exists ss such that s2bc2<2rss^{2}||b-c||^{2}<-2rs. Taking ss close enough to 0 works.  \Box

The next ingredient is an analogue of the so-called colorful Carathéodory theorem [4] for polynomials of bounded degree.

Proposition 2.5 (Colorful Carathéodory for polynomials).

Let DD and dd be positive integers and m:=(D+dd)m:=\binom{D+d}{d}. Let 1,,m[x1,,xd]\mathscr{F}_{1},\,\ldots,\,\mathscr{F}_{m}\subset\mathbb{C}[x_{1},\,\ldots,\,x_{d}] be finite sets of polynomials of degree at most DD. For any

pi=1mI(i),p\in\bigcap_{i=1}^{m}I(\mathscr{F}_{i}),

there exists {fii}i[m]\{f_{i}\in\mathscr{F}_{i}\}_{i\in[m]} such that pI(f1,,fm)p\in I(f_{1},\,\ldots,\,f_{m}).

Proof.

Let R:=[x1,,xd]R:=\mathbb{C}[x_{1},\,\ldots,\,x_{d}]. The ring RR is a \mathbb{C}-vector space so we have an inner product and norm. Define the distance between polynomials PP and QQ to be PQ||P-Q||. This is a metric. Let the distance between a polynomial PP and a subset SRS\subset R equal minQSPQ\min_{Q\in S}||P-Q||.

Suppose fiif_{i}\in\mathscr{F}_{i} for 1im1\leqslant i\leqslant m has the property that the ideal 𝔞:=I(f1,,fm)\mathfrak{a}:=I(f_{1},\,\ldots,\,f_{m}) is closest to pp, with h𝔞h\in\mathfrak{a} being the closest point. Assume for the sake of contradiction that ph>0||p-h||>0. Note that polynomials in RR of degree at most DD have at most mm terms. Thus the \mathbb{C}-vector space VV of polynomials of degree at most DD has dimension mm.

If f1,,fmf_{1},\,\ldots,\,f_{m} forms a basis of VV, there exists c1,,cmc_{1},\,\ldots,\,c_{m}\in\mathbb{C} not all 0 such that c1f1++cmfm=1c_{1}f_{1}+\cdots+c_{m}f_{m}=1. But then 1𝔞1\in\mathfrak{a} so pR=𝔞p\in R=\mathfrak{a}, a contradiction.

Hence there must exist c1,,cmc_{1},\,\ldots,\,c_{m}\in\mathbb{C} not all 0 for which c1f1++cmfm=0c_{1}f_{1}+\cdots+c_{m}f_{m}=0. Without loss of generality, suppose cm0c_{m}\neq 0; so 𝔞=I(f1,,fm1)\mathfrak{a}=I(f_{1},\,\ldots,\,f_{m-1}).

Suppose that for all fmf\in\mathscr{F}_{m} and g(f)g\in(f), it is true that gh,ph=0\langle g-h,\,p-h\rangle=0. Since 0(f)0\in(f) it follows that g,ph=gh,ph0h,ph=0\langle g,\,p-h\rangle=\langle g-h,\,p-h\rangle-\langle 0-h,\,p-h\rangle=0. Thus (m)(\mathscr{F}_{m}) is contained in (ph)(p-h)^{\perp}. However, p,ph=ph,ph0h,ph=ph20\langle p,\,p-h\rangle=\langle p-h,\,p-h\rangle-\langle 0-h,\,p-h\rangle=||p-h||^{2}\neq 0 gives p(ph)p\not\in(p-h)^{\perp} so p(m)p\not\in(\mathscr{F}_{m}), a contradiction.

Hence there exists fmf\in\mathscr{F}_{m} and g(f)g\in(f) such that gh,ph0\langle g-h,\,p-h\rangle\neq 0. By Lemma 2.4, there exists h1[g,h]I(f,h)I(f1,,fm1,f)h_{1}\in[g,\,h]\subset I(f,\,h)\subset I(f_{1},\,\ldots,\,f_{m-1},f) closer to pp than hh. Thus I(f1,,fm1,f)I(f_{1},\,\ldots,\,f_{m-1},f) is closer to pp than 𝔞\mathfrak{a}, a contradiction.  \Box

Proof of Theorem 2.3.

By Hilbert’s Nullstellensatz, 1(i)1\in(\mathscr{F}_{i}) for all i[m]i\in[m]. By Proposition 2.5, there exists a transversal {fii}i[m]\{f_{i}\in\mathscr{F}_{i}\}_{i\in[m]} such that 1(f1,,fm)1\in(f_{1},\,\ldots,\,f_{m}). Thus iZ(fi)=\bigcap_{i}Z(f_{i})=\emptyset by Hilbert’s Nullstellensatz again.  \Box

We will rederive Theorem 2.3 as a consequence of Theorem 1.3 in Section 3. In doing so, we will circumvent the need for the nontrivial direction in the Hilbert Nullstellensatz and thus get an optimal colorful Helly result for polynomials of bounded degree over any ground field 𝔽\mathbb{F}.


3. Proof of Theorem 1.1, Proposition 1.2 and Theorem 1.3

We begin with the proof of Theorem 1.3, which subsumes Theorem 1.1.

3.1. Proof of Theorem 1.3

The setup: let τ=τ(X,)\tau=\tau^{\prime}(X,\mathscr{F}), and 1,,τ+1\mathscr{F}_{1},\ldots,\mathscr{F}_{\tau+1} be finite subfamilies of \mathscr{F} with an empty intersection. We make the following key observation.

Claim. For each (τ+1)(\tau+1)-tuple C=(F1,,Fτ+1)1××τ+1C=(F_{1},\ldots,F_{\tau+1})\in\mathscr{F}_{1}\times\cdots\times\mathscr{F}_{\tau+1} with non-empty intersection, there exists some C=(F1,,Fτ+1)1××τ+1C^{\prime}=(F^{\prime}_{1},\ldots,F^{\prime}_{\tau+1})\in\mathscr{F}_{1}\times\cdots\times\mathscr{F}_{\tau+1} such that

F1Fτ+1F1Fτ+1.F^{\prime}_{1}\cap\cdots\cap F^{\prime}_{\tau+1}\subsetneq F_{1}\cap\cdots\cap F_{\tau+1}.

Proof of Claim. We first note that there must exist an i[τ+1]i\in[\tau+1] such that

FijiFj.F_{i}\supset\bigcap_{j\neq i}F_{j}.

If not, we can find xiXx_{i}\in X that lies in jiFj\bigcap_{j\neq i}F_{j} but not FiF_{i}. Setting xτ+2x_{\tau+2} to be any element in the intersection jFj\bigcap_{j}F_{j}, we obtain a comatching with intersection of size τ+1\tau+1, contradiction.

Consider any xjiFjx\in\bigcap_{j\neq i}F_{j}. As the sets in i\mathscr{F}_{i} have empty intersection, there exists some FiiF_{i}^{\prime}\in\mathscr{F}_{i} that does not contain xx. Letting Fj=FjF_{j}^{\prime}=F_{j} for each jij\neq i, we have

F1Fτ+1jiFj=jiFj=jFjF^{\prime}_{1}\cap\cdots\cap F^{\prime}_{\tau+1}\subsetneq\bigcap_{j\neq i}F^{\prime}_{j}=\bigcap_{j\neq i}F_{j}=\bigcap_{j}F_{j}

which proves the claim.

Let the (τ+1)(\tau+1)-tuple C=(F1,,Fτ+1)1××τ+1C=(F_{1},\ldots,F_{\tau+1})\in\mathscr{F}_{1}\times\cdots\times\mathscr{F}_{\tau+1} achieve the minimal non-empty intersection F1Fτ+1F_{1}\cap\cdots\cap F_{\tau+1} with respect to the inclusion ordering; such FiF_{i}’s exist because the i\mathscr{F}_{i}’s are finite. Applying the claim, we obtain C=(F1,,Fτ+1)1××τ+1C^{\prime}=(F^{\prime}_{1},\ldots,F^{\prime}_{\tau+1})\in\mathscr{F}_{1}\times\cdots\times\mathscr{F}_{\tau+1} such that

F1Fτ+1F1Fτ+1.F^{\prime}_{1}\cap\cdots\cap F^{\prime}_{\tau+1}\subsetneq F_{1}\cap\cdots\cap F_{\tau+1}.

By minimality, we must have F1Fτ+1=F^{\prime}_{1}\cap\cdots\cap F^{\prime}_{\tau+1}=\emptyset, as desired. This proves that

η(X,)1+τ(X,).\eta(X,\mathscr{F})\leqslant 1+\tau^{\prime}(X,\mathscr{F}).

In particular, if τ(X,)=τ(X,)1\tau^{\prime}(X,\mathscr{F})=\tau(X,\mathscr{F})-1, then η(X,)τ(X,)\eta(X,\mathscr{F})\leqslant\tau(X,\mathscr{F}). On the other hand, the reverse inequality η(X,)τ(X,)\eta(X,\mathscr{F})\geqslant\tau(X,\mathscr{F}) also holds. Write instead τ=τ(X,)\tau=\tau(X,\mathscr{F}). Let F1,,FτF_{1},\ldots,F_{\tau} be a maximum comatching in \mathscr{F}. Since τ(X,)=τ1\tau^{\prime}(X,\mathscr{F})=\tau-1, we have F1Fτ=F_{1}\cap\cdots\cap F_{\tau}=\emptyset; taking 1==τ1={F1,,Fτ}\mathscr{F}_{1}=\cdots=\mathscr{F}_{\tau-1}=\{F_{1},\ldots,F_{\tau}\} shows that \mathscr{F} has colorful Helly number at least τ\tau. All in all, we showed that

(3) ifτ(X,)=τ(X,)1,thenη(X,)=τ(X,).\text{if}\ \tau^{\prime}(X,\mathscr{F})=\tau(X,\mathscr{F})-1,\ \text{then}\ \eta(X,\mathscr{F})=\tau(X,\mathscr{F}).

3.2. Examples showing sharpness

We now show that Theorem 1.1 is tight by itself. For example, let M=2M=2, X={1, 2, 3, 4}X=\{1,\,2,\,3,\,4\}, and

1={{1, 2},{3, 4}},\mathscr{F}_{1}=\{\{1,\,2\},\,\{3,\,4\}\},
2={{2, 3},{4, 1}},\mathscr{F}_{2}=\{\{2,\,3\},\,\{4,\,1\}\},
=12.\mathscr{F}=\mathscr{F}_{1}\cup\mathscr{F}_{2}.

Then there is no comatching in \mathscr{F} of size 33: any three element subset of XX must be of the form {i,i+1,i+2}\{i,\,i+1,\,i+2\} modulo 44, and there is no subset of \mathscr{F} that contains {i,i+2}\{i,i+2\}. So τ(X,)=2\tau(X,\mathscr{F})=2, but η(X,)3\eta(X,\mathscr{F})\geqslant 3 because one can consider the aforementioned two families 1\mathscr{F}_{1} and 2\mathscr{F}_{2} (which clearly do not satisfy the colorful Helly property).

A more general example goes as follows. Take any M2M\geqslant 2. Let X={1,, 2M}X=\{1,\,\ldots,\,2M\}, and

1==M1={X\{1, 2},X\{3, 4},,X\{2M1, 2M}},\mathscr{F}_{1}=\cdots=\mathscr{F}_{M-1}=\{X\backslash\{1,\,2\},\,X\backslash\{3,\,4\},\,\ldots,\,X\backslash\{2M-1,\,2M\}\},
M={X\{2, 3},X\{4, 5},,X\{2M, 1}},\mathscr{F}_{M}=\{X\backslash\{2,\,3\},\,X\backslash\{4,\,5\},\,\ldots,\,X\backslash\{2M,\,1\}\},
=i=1Mi.\mathscr{F}=\bigcup_{i=1}^{M}\mathscr{F}_{i}.

Then there is no comatching of size (M+1)(M+1) in \mathscr{F}: any (M+1)(M+1) element subset MM of XX must contain three elements of the form {i,i+1,i+2}\{i,\,i+1,\,i+2\} modulo MM, and there is no set in \mathscr{F} that contains ii and i+2i+2 but not i+1i+1. So τ(X,)M\tau(X,\mathscr{F})\leqslant M. On the other hand, we have η(X,)M+1\eta(X,\mathscr{F})\geqslant M+1 because one can consider the aforementioned families 1,,M\mathscr{F}_{1},\cdots,\mathscr{F}_{M}, which do not satisfy the colorful Helly property.

3.3. Proof of Proposition 1.2

Let η=η(X,)\eta=\eta(X,\mathscr{F}). It is clear that η(X,)η\eta^{\infty}(X,\mathscr{F})\geqslant\eta. On the other hand, consider (possibly infinite) subfamilies 1,,η\mathscr{F}_{1},\,\ldots,\,\mathscr{F}_{\eta}\subset\mathscr{F} such that FiF=\cap_{F\in\mathscr{F}_{i}}F=\emptyset for every i=1,,ηi=1,\,\ldots,\,\eta. By the Noetherian hypothesis, there exists finite ii\mathscr{F}_{i}^{\prime}\subset\mathscr{F}_{i} such that FiF=\cap_{F\in\mathscr{F}_{i}^{\prime}}F=\emptyset. By the definition of η\eta, there exists FiiF_{i}\in\mathscr{F}_{i}^{\prime} such that F1Fη=F_{1}\cap\cdots\cap F_{\eta}=\emptyset. We conclude that η(X,)η.\eta^{\infty}(X,\mathscr{F})\leqslant\eta.

3.4. Applications

Finally, we would like to record three special cases of Theorem 1.3, in particular the criterion for colorful Helly from (3). The first one is the fact that the colorful Helly numbers of spheres in d\mathbb{R}^{d} equals the Helly number d+2d+2, already proved in Theorem 2.2.


Spheres. Let X=dX=\mathbb{R}^{d}, and let 𝒮d\mathcal{S}_{d} be the family of all dd-dimensional spheres. Then we have

τ(X,𝒮d)=d+2,τ(X,𝒮d)=d+1.\tau(X,\mathcal{S}_{d})=d+2,\tau^{\prime}(X,\mathcal{S}_{d})=d+1.

We have shown the first statement above, which implies that τ(X,𝒮d){d+1,d+2}\tau^{\prime}(X,\mathcal{S}_{d})\in\{d+1,d+2\}. To show that τ(X,𝒮d)=d+1\tau^{\prime}(X,\mathcal{S}_{d})=d+1, suppose for the sake of contradiction that points x1,,xd+3dx_{1},\ldots,x_{d+3}\in\mathbb{R}^{d} and spheres S1,,Sd+2𝒮dS_{1},\ldots,S_{d+2}\in\mathcal{S}_{d} form a comatching with intersection. We apply inversion ϕ\phi with center O=xd+3O=x_{d+3}. Then ϕ(S1),,ϕ(Sd+2)\phi(S_{1}),\,\ldots,\,\phi(S_{d+2}) are a set of (d+2)(d+2)-planes in d\mathbb{R}^{d} that form a comatching with ϕ(x1),,ϕ(xd+2)\phi(x_{1}),\ldots,\phi(x_{d+2}). This is impossible because d+2d+2 points in d\mathbb{R}^{d} are always affinely dependent. So we recover the fact that

η(X,𝒮d)=η(X,𝒮d)=d+2.\eta(X,\mathcal{S}_{d})=\eta^{\infty}(X,\mathcal{S}_{d})=d+2.

Hypersurfaces of bounded degree. Let 𝔽\mathbb{F} be an arbitrary field, let X:=𝔽dX:=\mathbb{F}^{d}, and let VD:=F[x1,,xd]DV_{D}:=F[x_{1},\ldots,x_{d}]_{\leqslant D} be the vector space of polynomials in x1,,xdx_{1},\ldots,x_{d} and of degree at most DD. Recall that τ(X,VD)(D+dd)\tau(X,\,V_{D})\leqslant\binom{D+d}{d}, since the elements of any comatching must correspond to linearly independent polynomials via (1). Suppose for the sake of contradiction that τ(X,VD)(D+dd)\tau^{\prime}(X,\,V_{D})\geqslant\binom{D+d}{d}. Set m=(D+dd)m=\binom{D+d}{d}. Then there exists polynomials f1,,fmF[x1,,xd]Df_{1},\,\ldots,\,f_{m}\in F[x_{1},\ldots,x_{d}]_{\leqslant D} and a0,a1,,amFa_{0},\,a_{1},\,\ldots,\,a_{m}\in F such that fi(aj)=0f_{i}(a_{j})=0 if and only if iji\neq j. Since the polynomials are linearly independent, they form a basis. Hence there exists c1,,cmFc_{1},\,\ldots,\,c_{m}\in F such that c1f1++cmfm=1c_{1}f_{1}+\cdots+c_{m}f_{m}=1. It follows that 0=c1f1(a0)++cmfm(a0)=10=c_{1}f_{1}(a_{0})+\cdots+c_{m}f_{m}(a_{0})=1, a contradiction. Thus τ(X,VD)(D+dd)1\tau^{\prime}(X,\,V_{D})\leqslant\binom{D+d}{d}-1. By (3), it follows that

η(X,VD)(D+dd).\eta(X,V_{D})\leqslant\binom{D+d}{d}.

In other words, this shows that Theorem 2.3 holds over any ground field 𝔽\mathbb{F}, not only over \mathbb{C}.


Hamming balls of a fixed radius. The Hamming distance between p,qnp,\,q\in\mathbb{R}^{n}, denoted dist(p,q)\mathrm{dist}(p,\,q), is the number of coordinates where pp and qq differ. The Hamming ball centered at xx with radius tt, denoted B(x,t)B(x,\,t), is the set of points yny\in\mathbb{R}^{n} with dist(x,y)t\mathrm{dist}(x,\,y)\leqslant t. In [1], Alon, Jin and Sudakov prove that the Helly number for the family \mathscr{F} of Hamming balls of radius tt in n\mathbb{R}^{n} is 2t+12^{t+1} by establishing τ(X,)=2t+1\tau(X,\mathscr{F})=2^{t+1}. We generalize this result by showing that the colorful Helly number of \mathscr{F} is also 2t+12^{t+1}.

Recall that τ=τ(X,)\tau^{\prime}=\tau^{\prime}(X,\mathscr{F}) is the largest τ\tau^{\prime} such that there exist x1,,xτ+1Xx_{1},\ldots,x_{\tau^{\prime}+1}\in X and F1,,FτF_{1},\ldots,F_{\tau^{\prime}} such that xiFjx_{i}\in F_{j} if and only if iji\neq j. We prove that τ\tau^{\prime} for Hamming balls of radius tt is 2t+112^{t+1}-1. Let m:=2t+1m:=2^{t+1}. Suppose for the sake of contradiction there exists a1,,am,b0,b1,,bmna_{1},\,\ldots,\,a_{m},\,b_{0},\,b_{1},\,\ldots,\,b_{m}\in\mathbb{R}^{n} such that dist(ai,bj)t\mathrm{dist}(a_{i},\,b_{j})\leqslant t if and only if iji\neq j. For each i[m]i\in[m], let Di:={k[n]:ai,kbi,k}D_{i}:=\{k\in[n]:a_{i,k}\neq b_{i,k}\} and did_{i} be the largest element of DiD_{i}. For all i[m]i\in[m], let si:=dist(ai,bi)ts_{i}:=\mathrm{dist}(a_{i},\,b_{i})-t. We call a pair (I1,I2)(I_{1},\,I_{2}) compatible with ii if I1DiI_{1}\subset D_{i}, |I1|t+s1+12|I_{1}|\geqslant t+\frac{s_{1}+1}{2}, I2[n]DiI_{2}\subset[n]\setminus D_{i}, or I1Di{di}I_{1}\subset D_{i}\setminus\{d_{i}\}, |I1|=t+si2|I_{1}|=t+\frac{s_{i}}{2}, I2[n]DiI_{2}\subset[n]\setminus D_{i}. Note that |I1|t+1|I_{1}|\geqslant t+1 in both cases. For every i[m]i\in[m] and every such pair (I1,I2)(I_{1},\,I_{2}), define a polynomial on xnx\in\mathbb{R}^{n} by

fi,I1,I2(x):=kI1I2(xkai,k)kDiI1(xkbi,k).f_{i,I_{1},I_{2}}(x):=\prod_{k\in I_{1}\cup I_{2}}(x_{k}-a_{i,k})\prod_{k\in D_{i}\setminus I_{1}}(x_{k}-b_{i,k}).

In [1, Theorem 2.3], Alon, Jin and Sudakov prove that these polynomials are linearly independent in the vector space of multilinear polynomials on nn variables Vn=[x1,,xn]/(x12x1,,xn2xn)V_{n}=\mathbb{R}[x_{1},\cdots,x_{n}]/(x_{1}^{2}-x_{1},\cdots,x_{n}^{2}-x_{n}). The vector space VnV_{n} has dimension 2n2^{n}. Let

Vn,d:={i=0(d1)/2(ni),if d is oddi=0d/21(ni)+(n1d/21),if d is evenV_{n,d}:=\begin{cases}\sum_{i=0}^{(d-1)/2}\binom{n}{i},&\text{if $d$ is odd}\\ \sum_{i=0}^{d/2-1}\binom{n}{i}+\binom{n-1}{d/2-1},&\text{if $d$ is even}\end{cases}

The number of pairs compatible with ii is 2n(t+si)Vt+si,si2^{n-(t+s_{i})}V_{t+s_{i},s_{i}} so

i=1m2n(t+si)Vt+si,si2n.\sum_{i=1}^{m}2^{n-(t+s_{i})}V_{t+s_{i},s_{i}}\leqslant 2^{n}.

By [1, Claim 2.2], Vt+si,si2si1V_{t+s_{i},s_{i}}\geqslant 2^{s_{i}-1} so

2ni=1m2n(t+si)Vt+si,sii=1m2nt1=2n.2^{n}\geqslant\sum_{i=1}^{m}2^{n-(t+s_{i})}V_{t+s_{i},s_{i}}\geqslant\sum_{i=1}^{m}2^{n-t-1}=2^{n}.

Hence the polynomials fi,I1,I2f_{i,I_{1},I_{2}} form a basis of VnV_{n}, and there exists coefficients ci,I1,I2c_{i,I_{1},I_{2}}\in\mathbb{R} such that

i,I1,I2ci,I1,I2fi,I1,I2(x)=1,for allx{0,1}[n].\sum_{i,I_{1},I_{2}}c_{i,I_{1},I_{2}}f_{i,I_{1},I_{2}}(x)=1,\text{for all}\ x\in\{0,1\}^{[n]}.

For all (i,I1,I2)(i,\,I_{1},\,I_{2}), as b0b_{0} has Hamming distance at most tt with aia_{i}, and |I1I2|t+1\lvert I_{1}\cup I_{2}\rvert\geqslant t+1, there must exists kI1I2k\in I_{1}\cup I_{2} such that b0,k=ai,kb_{0,k}=a_{i,k}. Therefore, the first product in the definition of fi,I1,I2f_{i,I_{1},I_{2}} vanishes at b0b_{0}, thus fi,I1,I2(b0)=0f_{i,I_{1},I_{2}}(b_{0})=0. Therefore, we have

0=i,I1,I2ci,I1,I2fi,I1,I2(b0)=1,0=\sum_{i,I_{1},I_{2}}c_{i,I_{1},I_{2}}f_{i,I_{1},I_{2}}(b_{0})=1,

a contradiction. Thus τ=2t+11\tau^{\prime}=2^{t+1}-1 and the conclusion follows by the criterion from (3).


4. Collapsibility

Motivated by our results on the colorful Helly theorem, we further explore the connection between the comatching number and the collapsible and Leray properties of the nerve complex. We first recall some relevant definitions (See e.g. [20]). All simplicial complexes in this section are abstract simplicial complexes. A dd-collapse of a simplicial complex removes a free face σ\sigma with |σ|=d\lvert\sigma\rvert=d and all faces containing σ\sigma. We say a simplicial complex KK is dd-collapsible if we can remove all simplices of size at least dd via a sequence of dd-collapses. We say KK is dd-Leray if for any induced subcomplex LL of KK, the reduced homology groups H~i(L)\tilde{H}_{i}(L) with coefficients in \mathbb{R} are trivial for all idi\geqslant d.

The nerve complex N()N(\mathscr{F}) of a set family \mathscr{F} is the simplicial complex on vertex set \mathscr{F} whose faces are the subfamilies of \mathscr{F} with non-empty intersection. We say a complex KK is dd-representable if it is isomorphic to the nerve complex of convex sets in d\mathbb{R}^{d}. The connections between dd-representable, dd-collapsible, dd-Leray and the Helly theorems are well-studied. Wegner [28] showed that dd-representable implies dd-collapsible, and dd-collapsible implies dd-Leray. In the opposite direction, Tancer [27] showed that for any dd there exists a simplicial complex that is 22-collapsible but not dd-representable. Alon and Kalai [2] showed that if N()N(\mathscr{F}) is dd-collapsible, then \mathscr{F} has fractional Helly number (d+1)(d+1) with optimal dependence between α\alpha and β\beta. Furthermore, Kalai and Meshulam [18] also showed that a dd-collapsible simplicial complex property has the colorful Helly property, thereby generalizing the colorful Helly theorem for convex set (or more precisely the stronger fact that the dd-Leray property implies that the colorful Helly number at most d+1d+1). For the sake of completeness, we include a version of their proof in the abstract setting for set systems.

Proposition 4.1.

If N()N(\mathscr{F}) is dd-collapsible, then \mathscr{F} has colorful Helly number at most (d+1)(d+1).

Proof.

Suppose 1,,d+1\mathscr{F}_{1},\ldots,\mathscr{F}_{d+1} are subfamilies of \mathscr{F} such that for every F11,,Fd+1d+1F_{1}\in\mathscr{F}_{1},\ldots,F_{d+1}\in\mathscr{F}_{d+1}, we have F1Fd+1F_{1}\cap\cdots\cap F_{d+1}\neq\emptyset. Then (F1,,Fd+1)(F_{1},\ldots,F_{d+1}) is a face in N()N(\mathscr{F}). Without loss of generality, assume (F1,,Fd+1)(F_{1},\ldots,F_{d+1}) is the first face in 1××d+1\mathscr{F}_{1}\times\cdots\times\mathscr{F}_{d+1} to be removed during the dd-collapsing process by collapsing σ=(F1,,Fd)\sigma=(F_{1},\ldots,F_{d}). Let τ\tau be the unique maximal face containing σ\sigma right before this dd-collapse. For any Fd+1F\in\mathscr{F}_{d+1}, σ{F}\sigma\cup\{F\} is a face right before the collapse, so we must have σ{F}τ\sigma\cup\{F\}\in\tau. We conclude that d+1τ\mathscr{F}_{d+1}\subset\tau, thus d+1\mathscr{F}_{d+1} is a face in N()N(\mathscr{F}) so

Fd+1F.\bigcap_{F\in\mathscr{F}_{d+1}}F\neq\emptyset.

\Box

A similar argument can also be found in [20]. In light of our Theorem 1.3, it is natural to wonder if τ(X,)d\tau(X,\mathscr{F})\leqslant d implies dd-collapsibility. If true, Proposition 4.1 would thus imply our Theorem 1.3 and would also yield Corollary 1.4 with optimal dependence between α\alpha and β\beta (via the work of Alon and Kalai [2]). Unfortunately, the following proposition shows that this is not the case.

Proposition 4.2.

For every d1d\geqslant 1, there exist a base set XX and a set family \mathscr{F} on XX with comatching number τ(X,)τ(X,)2d\tau^{\prime}(X,\mathscr{F})\leqslant\tau(X,\mathscr{F})\leqslant 2d, such that N()N(\mathscr{F}) is not (3d1)(3d-1)-Leray and thus not (3d1)(3d-1)-collapsible.

We now prove this proposition. First, we define an analogue of comatching number for abstract simplicial complexes.

Definition 4.3.

Let KK be an abstract simplicial complex on vertex set VV. A comatching in KK is a subset MVM\subset V such that the following holds: for any vMv\in M, there is a maximal face FF of KK with FM=M\{v}F\cap M=M\backslash\{v\}. Let τ(V,K)\tau(V,K) denote the maximum size of a comatching in KK.

We can easily relate the definition of comatching for set families and simplicial complexes.

Proposition 4.4.

1) For any set family \mathscr{F} on ground set XX, we have τ(,N())τ(X,)\tau(\mathscr{F},N(\mathscr{F}))\leqslant\tau(X,\mathscr{F}).

2) For any simplicial complex KK on vertex set VV without an isolated vertex, there is a ground set XX and a set family (K)\mathscr{F}(K) on XX such that KN((K))K\cong N(\mathscr{F}(K)) and τ(X,(K))max(2,τ(V,K))\tau(X,\mathscr{F}(K))\leqslant\max(2,\tau(V,K)).

Proof.

1) Suppose MM is a comatching in N()N(\mathscr{F}). We claim that MM is also a comatching in \mathscr{F}. Indeed, for any set fMf\in M, there is a maximal face FF of N()N(\mathscr{F}) such that FM=M\{f}F\cap M=M\backslash\{f\}. There must exist some aXa\in X such that

F={f:af}.F=\{f\in\mathscr{F}:a\in f\}.

Therefore, aa lies in every set in M\{f}M\backslash\{f\}, but not in ff. So MM forms a comatching in \mathscr{F}.

2) Let FF be the set of maximal faces in KK. Let X=VFX=V\sqcup F. For each vVv\in V, we associate the subset Fv={v}{fF:vf}F_{v}=\{v\}\cup\{f\in F:v\in f\} on XX. Let (K)\mathscr{F}(K) be the set family {Fv:vV}\{F_{v}:v\in V\} on XX.

We check that the map ϕ:sFs\phi:s\mapsto F_{s} gives an isomorphism of complexes KN((K))K\cong N(\mathscr{F}(K)). If f0f_{0} is a face in KK, then f0f_{0} is contained in some maximal face fFf\in F. Therefore, for any sF0s\in F_{0}, we have fϕ(s)f\in\phi(s), so fsF0ϕ(s)f\in\bigcap_{s\in F_{0}}\phi(s), thus ϕ(f0)\phi(f_{0}) is a face in the nerve complex by definition. On the other hand, suppose ϕ(f0)\phi(f_{0}) is a face in KK for some f0Vf_{0}\subset V. If |f0|=1\lvert f_{0}\rvert=1, then f0f_{0} is a face in KK as KK has no isolated vertex. If |f0|2\lvert f_{0}\rvert\geqslant 2, there exists some fVFf\in V\sqcup F such that fsf0ϕ(s)f\in\bigcap_{s\in f_{0}}\phi(s). We must have fTf\in T, thus sFs\in F for any sf0s\in f_{0}. So f0Ff_{0}\subset F, thus f0f_{0} is a face in VV, as desired.

Furthermore, suppose MM is a comatching of size at least 33 in (K)\mathscr{F}(K). We check that ϕ1(M)\phi^{-1}(M) is a comatching in KK. Consider any vϕ1(M)v\in\phi^{-1}(M). As MM is a comatching in (K)\mathscr{F}(K), there is some fVFf\in V\sqcup F such that for each FuMF_{u}\in M, FuF_{u} contains ff if and only if uvu\neq v. As |M\{v}|2\lvert M\backslash\{v\}\rvert\geqslant 2, at least two sets in (K)\mathscr{F}(K) contains FF, so we must have fFf\in F. Thus for uϕ1(M)u\in\phi^{-1}(M), we have uffFuuvu\in f\Leftrightarrow f\in F_{u}\Leftrightarrow u\neq v. We conclude that fϕ1(M)=ϕ1(M)\{v}f\cap\phi^{-1}(M)=\phi^{-1}(M)\backslash\{v\}, and ϕ1(M)\phi^{-1}(M) is a comatching in (K)\mathscr{F}(K), as desired.  \Box

In light of part 2) of this proposition, we can restate Proposition 4.2 in terms of the comatching number for simplicial complexes (note that isolated vertices don’t affect the collapsible or Leray property).

Proposition 4.5.

For every d2d\geqslant 2, there exists a vertex set VV and a simplicial complex KK on VV, such that τ(V,K)2d\tau(V,K)\leqslant 2d and KK is not (3d1)(3d-1)-Leray thus also not (3d1)(3d-1)-collapsible.

We first verify this proposition for d=1d=1 with an explicit construction. We make use of Leray’s celebrated nerve theorem, which allows us to easily compute the homology of certain nerve complexes.

Theorem 4.6 ([22]).

Let XX be a topological space, and let ={Ui}\mathscr{F}=\{U_{i}\} be an open cover of XX. Suppose all finite, non-empty intersections of sets in \mathscr{F} are contractible. Then the nerve complex N()N(\mathscr{F}) is homotopy-equivalent to XX. Therefore, H~i(N())H~i(X)\tilde{H}_{i}(N(\mathscr{F}))\cong\tilde{H}_{i}(X) for every i0i\geqslant 0.

Proposition 4.7.

There exists a vertex set VV and a simplicial complex KK on VV such that τ(V,K)=2\tau(V,K)=2, but KK is not 22-Leray.

Proof.
12345678910111213141516
Figure 2. The regions in red and blue are members of \mathscr{F}. To visualize K=N()K=N(\mathscr{F}), we identify the red region with grid square 11, and the blue region with grid square 1313.

Let XX be the surface of the torus (/)2(\mathbb{R}/\mathbb{Z})^{2}. We partition XX into a 4×44\times 4 square grid with the sides glued together. For the set of 2×22\times 2 subsquare SS of the grid, let SS^{\circ} be its interior. Let \mathscr{F} be the family of SS^{\circ}’s. Define KK to be the nerve complex N()N(\mathscr{F}) of \mathscr{F}. See Figure 2 above for an illustration.

We first check that τ(,K)=2\tau(\mathscr{F},K)=2. We identify each SS^{\circ}\in\mathscr{F} with its top-left grid square. Under this identification, the maximal faces of KK are the 2×22\times 2 subsquares of the grid. Suppose we have a comatching MM of size 33 in KK. Then each pair of corresponding grid squares in MM must lie in a common maximal face, which is possible only if the corresponding grid squares in MM form a right isoceles triangle. Without loss of generality, we may assume that MM consists of grid squares {1,2,5}\{1,2,5\}. Then there is no maximal face that contains {2,5}\{2,5\} but not 11, contradiction.

We now compute the homology of KK. As the non-empty intersections of subfamilies of \mathscr{F} are contractible, Leray’s nerve theorem shows that KK is homotopic-equivalent to XX. Therefore, we have H~2(K)H~2(X)H~2((/)2)\tilde{H}_{2}(K)\cong\tilde{H}_{2}(X)\cong\tilde{H}_{2}((\mathbb{R}/\mathbb{Z})^{2})\cong\mathbb{R}, so KK is not 22-Leray.  \Box

We generalize the construction to arbitrary comatching number by using the join [21, Definition 2.16] of two simplicial complexes.

Definition 4.8.

Let KK be a simplicial complex on VV and LL be a simplicial complex on WW. The join of KK and LL, denoted by KLK*L, is the simplicial complex on VWV\sqcup W with a face FGF\sqcup G for each face FF in KK and face GG in LL.

We now illustrate how the comatching number and the homology groups behave under joins.

Proposition 4.9.

We have τ(VW,KL)τ(V,K)+τ(W,L)\tau(V\sqcup W,K*L)\leqslant\tau(V,K)+\tau(W,L).

Proof.

If PP is a comatching in KLK*L, then PVP\cap V is a comatching in KK and PWP\cap W is a comatching in LL. So |P|=|PV|+|PW|τ(K)+τ(L)\lvert P\rvert=\lvert P\cap V\rvert+\lvert P\cap W\rvert\leqslant\tau(K)+\tau(L).  \Box

Definition 4.10.

We say KK is dd-good if H~d(K)0\tilde{H}_{d}(K)\neq 0 and H~i(K)=0\tilde{H}_{i}(K)=0 for every i>di>d.

Proposition 4.11.

If KK is dd-good and LL is dd^{\prime}-good, with d,d1d,d^{\prime}\geqslant 1, then KLK*L is (d+d+1)(d+d^{\prime}+1)-good.

Proof.

Fix any kd+d+1k\geqslant d+d^{\prime}+1. By an analog of Künneth’s theorem for join (see e.g. [18, Section 3], [29]), we have

H~k(KL)i+j=k1H~i(K)H~j(L).\tilde{H}_{k}(K*L)\cong\bigoplus_{i+j=k-1}\tilde{H}_{i}(K)\otimes\tilde{H}_{j}(L).

By the goodness assumption on KK and LL, all terms in the direct sum vanish except when k=d+d+1k=d+d^{\prime}+1, i=di=d and j=dj=d^{\prime}. So H~k(KL)\tilde{H}_{k}(K*L) is nonzero when k=d+d+1k=d+d^{\prime}+1 and zero when k>d+d+1k>d+d^{\prime}+1.  \Box

Proof of Proposition 4.5.

In Proposition 4.7, we constructed a simplicial complex K2K_{2} such that τ(K2)=2\tau(K_{2})=2 and K2K_{2} is 22-good. Now we take K2d=K2K2K_{2d}=K_{2}*\cdots*K_{2}, where there are dd copies of K2K_{2}. By Proposition 4.9, we have

τ(K2d)dτ(K2)2d.\tau(K_{2d})\leqslant d\tau(K_{2})\leqslant 2d.

By Proposition 4.11, K2dK_{2d} is (3d1)(3d-1)-good, thus not (3d1)(3d-1)-Leray. So K2dK_{2d} satisfies the desired properties.  \Box

We find it interesting to see whether bounded comatching number implies dd-collapsible or dd-Leray for some dd. We already don’t know the answer to the following question.

Question 4.12.

Is there an absolute constant d>0d>0 such that the following holds: for every set family \mathscr{F} with τ()2\tau(\mathscr{F})\leqslant 2, the nerve complex N()N(\mathscr{F}) is always dd-Leray?

References

  • [1] N. Alon, Z. Jin, B. Sudakov, The Helly number of Hamming balls and related problems, arXiv:2405.10275.
  • [2] N. Alon and G. Kalai, A simple proof of the upper bound theorem, European J. Combin. 6 (1985), no. 3, 211–214.
  • [3] N. Amenta, J. A. De Loera, and P. Soberón, Helly’s theorem: new variations and applications, Algebraic and geometric methods in discrete mathematics, Contemp. Math., vol. 685, Amer. Math. Soc., Providence, RI, 2017, pp. 55–95.
  • [4] I. Bárány, A generalization of Carathéodory’s theorem, Discrete Math. 40 (1982), no. 2-3, 141–152.
  • [5] I. Bárány, Combinatorial Convexity, AMS University Lecture Series 77, 2021.
  • [6] I. Bárány, G. Kalai, Helly-type problems, Bull. Amer. Math. Soc. (N.S.), 59(4):471–502, 2022.
  • [7] C. Carathéodory, Über den Variabilitätsbereich der Koeffizienten von Potenzreihen, Math. Annalen, 64 (1907), 95–115.
  • [8] G. R. Chelnokov and V. L. Dol’nikov, On transversals of quasialgebraic families of sets, J. Combin. Theory Ser. A 125 (2014), 194–213.
  • [9] L. Danzer, B. Grünbaum, V. Klee, Helly’s theorem and its relatives, Proc. Sympos. Pure Math., Vol. VII, Amer. Math. Soc., Providence, R.I., 1963, pp. 101–180.
  • [10] M. Deza and P. Frankl, A Helly type theorem for hypersurfaces, J. Combin. Th. Ser. A, 45 (1987), 27–30.
  • [11] J. P. Doignon, Convexity in cristallographical lattices, J. Geom 3 (1973), no. 1, 71–85.
  • [12] J. Eckhoff, Helly, Radon, and Carathéodory type theorems, Handbook of convex geometry, Vol. A, B, North-Holland, Amsterdam, 1993, pp. 389–448.
  • [13] P. Frankl, Helly-type theorems for varieties, Europ. J. Combinatorics, 10 (1989), 243–245.
  • [14] J. A. De Loera, R. N. La Haye, D. Oliveros, E. Roldán- Pensado, Helly numbers of Algebraic subsets of n\mathbb{R}^{n} and an extension of Doignon’s theorem, Advances in Geometry, Vol. 17 (4), (2017) 473–482.
  • [15] A. Holmsen, Large cliques in hypergraphs with forbidden substructures, Combinatorica 40 (2020), no. 4, 527–537.
  • [16] A. Holmsen, D. Lee, Radon numbers and the fractional Helly theorem, Israel J. Math. 241 (2021), no. 1, 433–447.
  • [17] G. Kalai. Intersection patterns of convex sets, Israel J. Math. 48(2-3), 161–174 (1984).
  • [18] G. Kalai, R. Meshulam, A topological colorful Helly theorem, Advances in Mathematics, 191 (2005), 305–311.
  • [19] M. Katchalski, A. Liu, A problem of geometry in n\mathbb{R}^{n}, Proc. Amer. Math. Soc. 75 (1979), 284–288.
  • [20] I. Khmelnitsky, dd-collapsibility and its applications, https://igorkhm.github.io/papers/thesis.pdf
  • [21] D. N. Kozlov, Combinatorial algebraic topology, Algorithms and Computation in Mathematics, 21, Springer, Berlin, 2008.
  • [22] J. Leray, Sur la forme des espaces topologiques et sur les points fixes des représentations, J. Math. Pures Appl. (9) 24 (1945), 95–167.
  • [23] H. Maehara, Helly-type theorems for spheres, Discrete Comput. Geom. 4 (1989), 279–285.
  • [24] J. Matousek, Lectures on Discrete Geometry, Springer-Verlag, New York, 2002.
  • [25] J. Matousek, Bounded VC-dimension implies a fractional Helly theorem, Discrete Comput. Geom. 31 (2004), 251–255.
  • [26] T. S. Motzkin, A proof of Hilbert’s Nullstellensatz, Math. Zeitschr. 63 (1955), 341–344.
  • [27] M. Tancer, Non-representability of finite projective planes by convex sets, Proc. Amer. Math. Soc. 138 (2010), no. 9, 3285–3291.
  • [28] G. Wegner, dd-collapsing and nerves of families of convex sets, Arch. Math. (Basel) 26 (1975), 317–321.
  • [29] G. W. Whitehead, Homotopy groups of joins and unions, Trans. Amer. Math. Soc. 83 (1956), 55–69.