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

On the Subspace Choosability in Graphs111An extended abstract of this work appeared in Proc. of the European Conference on Combinatorics, Graph Theory and Applications (EuroComb), 2021 [7].

Dror Chawin School of Computer Science, The Academic College of Tel Aviv-Yaffo, Tel Aviv 61083, Israel. Research supported by the Israel Science Foundation (grant No. 1218/20).    Ishay Haviv22footnotemark: 2
Abstract

A graph GG is said to be kk-subspace choosable over a field 𝔽\mathbb{F} if for every assignment of kk-dimensional subspaces of some finite-dimensional vector space over 𝔽\mathbb{F} to the vertices of GG, it is possible to choose for each vertex a nonzero vector from its subspace so that adjacent vertices receive orthogonal vectors over 𝔽\mathbb{F}. The subspace choice number of GG over 𝔽\mathbb{F} is the smallest integer kk for which GG is kk-subspace choosable over 𝔽\mathbb{F}. This graph parameter, introduced by Haynes, Park, Schaeffer, Webster, and Mitchell (Electron. J. Comb., 2010), is inspired by well-studied variants of the chromatic number of graphs, such as the (color) choice number and the orthogonality dimension.

We study the subspace choice number of graphs over various fields. We first prove that the subspace choice number of every graph with average degree dd is at least Ω(d/lnd)\Omega(\sqrt{d/\ln d}) over any field. We then focus on bipartite graphs and consider the problem of estimating, for a given integer kk, the smallest integer mm for which the subspace choice number of the complete bipartite graph Kk,mK_{k,m} over a field 𝔽\mathbb{F} exceeds kk. We prove upper and lower bounds on this quantity as well as for several extensions of this problem. Our results imply a substantial difference between the behavior of the choice number and that of the subspace choice number. We also consider the computational aspect of the subspace choice number, and show that for every k3k\geq 3 it is 𝖭𝖯\mathsf{NP}-hard to decide whether the subspace choice number of a given bipartite graph over 𝔽\mathbb{F} is at most kk, provided that 𝔽\mathbb{F} is either the real field or any finite field.

1 Introduction

Graph coloring is the problem of minimizing the number of colors in a vertex coloring of a graph GG where adjacent vertices receive distinct colors. This minimum is known as the chromatic number of GG and is denoted by χ(G)\chi(G). Being one of the most popular topics in graph theory, the graph coloring problem was extended and generalized over the years in various ways. One classical variant, initiated independently by Vizing in 1976 [19] and by Erdös, Rubin, and Taylor in 1979 [8], is that of choosability, also known as list coloring, which deals with vertex colorings with some restrictions on the colors available to each vertex. A graph G=(V,E)G=(V,E) is said to be kk-choosable if for every assignment of a set SvS_{v} of kk colors to each vertex vVv\in V, there exists a choice of colors cvSvc_{v}\in S_{v} that form a proper coloring of GG (that is, cvcvc_{v}\neq c_{v^{\prime}} whenever vv and vv^{\prime} are adjacent in GG). The choice number of a graph GG, denoted ch(G)\mathop{\mathrm{ch}}(G), is the smallest integer kk for which GG is kk-choosable. It is well known that the choice number ch(G)\mathop{\mathrm{ch}}(G) behaves quite differently from the standard chromatic number χ(G)\chi(G). In particular, it can be arbitrarily large even for bipartite graphs (see, e.g., [8]). The choice number of graphs enjoys an intensive study in graph theory involving combinatorial, algebraic, and probabilistic tools (see, e.g., [1]). The computational decision problem associated with the choice number is unlikely to be tractable, because it is known to be complete for the complexity class Π2\Pi_{2} of the second level of the polynomial-time hierarchy even for bipartite planar graphs [8, 10, 11].

Another interesting variant of graph coloring, introduced by Lovász [14] in the study of Shannon capacity of graphs, is that of orthogonal representations, where the vertices of the graph do not receive colors but vectors from some given vector space. A tt-dimensional orthogonal representation of a graph G=(V,E)G=(V,E) over \mathbb{R} is an assignment of a nonzero vector xvtx_{v}\in\mathbb{R}^{t} to every vertex vVv\in V, such that xv,xv=0\langle x_{v},x_{v^{\prime}}\rangle=0 whenever vv and vv^{\prime} are adjacent in GG.222Orthogonal representations of graphs are sometimes defined in the literature as orthogonal representations of the complement, namely, the definition requires vectors associated with non-adjacent vertices to be orthogonal. The orthogonality dimension of a graph GG over \mathbb{R} is the smallest integer tt for which there exists a tt-dimensional orthogonal representation of GG over \mathbb{R}. The orthogonality dimension parameter is closely related to several other well-studied graph parameters, and in particular, for every graph GG it is bounded from above by the chromatic number χ(G)\chi(G). The orthogonality dimension of graphs and its extensions to fields other than the reals have found a variety of applications in combinatorics, information theory, and theoretical computer science (see, e.g., [15, Chapter 10] and [12]). As for the computational aspect, the decision problem associated with the orthogonality dimension of graphs is known to be 𝖭𝖯\mathsf{NP}-hard over every field [16] (see also [9]).

In 2010, Haynes, Park, Schaeffer, Webster, and Mitchell [13] introduced another variant of the chromatic number of graphs that captures both the choice number and the orthogonality dimension. In this setting, which we refer to as subspace choosability, each vertex of a graph GG is assigned a kk-dimensional subspace of some finite-dimensional vector space, and the goal is to choose for each vertex a nonzero vector from its subspace so that adjacent vertices receive orthogonal vectors. The smallest integer kk for which such a choice is guaranteed to exist for all possible subspace assignments is called the subspace choice number of the graph GG, formally defined as follows.

Definition 1.1.

For a graph G=(V,E)G=(V,E) and a function f:Vf:V\rightarrow\mathbb{N}, GG is ff-subspace choosable over a field 𝔽\mathbb{F} if for every integer tt and for every assignment of subspaces Wv𝔽tW_{v}\subseteq\mathbb{F}^{t} with dim(Wv)=f(v)\dim(W_{v})=f(v) to the vertices vVv\in V (which we refer to as an ff-subspace assignment), there exists a choice of a nonzero vector xvWvx_{v}\in W_{v} for each vertex vVv\in V, such that xv,xv=0\langle x_{v},x_{v^{\prime}}\rangle=0 whenever vv and vv^{\prime} are adjacent in GG. For an integer kk, the graph GG is kk-subspace choosable over 𝔽\mathbb{F} if it is ff-subspace choosable over 𝔽\mathbb{F} for the constant function ff defined by f(v)=kf(v)=k. The subspace choice number of GG over 𝔽\mathbb{F}, denoted ch-s(G,𝔽)\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F}), is the smallest kk for which GG is kk-subspace choosable over 𝔽\mathbb{F}.

Here and throughout the paper, we associate with the real field \mathbb{R} and with every finite field 𝔽\mathbb{F} the inner product defined by x,y=xiyi\langle x,y\rangle=\sum{x_{i}y_{i}}, whereas for the complex field \mathbb{C} we consider, as usual, the one defined by x,y=xiyi¯\langle x,y\rangle=\sum{x_{i}\overline{y_{i}}}.

The work [13] has initiated the study of the subspace choice number of graphs over the real and complex fields. Among other things, it was shown there that a graph is 22-subspace choosable over \mathbb{R} if and only if it contains no cycles. We note that this is in contrast to the characterization given in [8] for the (chromatic) 22-choosable graphs, which include additional graphs such as even cycles. This implies that the choice number and the subspace choice number do not coincide even on the 44-cycle graph. Over the complex field \mathbb{C}, however, it was shown in [13] that a graph is 22-subspace choosable if and only if it either contains no cycles or contains only one cycle and that cycle is even. This demonstrates the possible effect of the field on the subspace choice number. It further follows from [13] that for every graph GG and every field 𝔽\mathbb{F}, it holds that ch-s(G,𝔽)Δ(G)+1\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F})\leq\Delta(G)+1 where Δ(G)\Delta(G) stands for the maximum degree in GG. In fact, a similar argument shows that Δ(G)\Delta(G) can be replaced in this bound by the degeneracy of GG (i.e., the smallest integer kk for which every subgraph of GG contains a vertex of degree at most kk).

1.1 Our Contribution

The current work studies the subspace choice number of graphs over various fields. Our first result provides a lower bound on the subspace choice number of a general graph over any field in terms of its average degree.

Theorem 1.2.

There exists a constant c>0c>0 such that for every graph GG with average degree d>1d>1 and for every field 𝔽\mathbb{F},

ch-s(G,𝔽)>cdlnd.\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F})>c\cdot\sqrt{\frac{d}{\ln d}}.

The proof of Theorem 1.2 is based on a probabilistic argument. For certain graphs, we provide an improved lower bound on the subspace choice number, avoiding the logarithmic term (see Theorem 2.7). This improvement relies on an explicit construction of finite projective planes.

We note that a result of Saxton and Thomason [17], improving on a result of Alon [2], asserts that for every graph GG with average degree dd, it holds that ch(G)(1+o(1))log2d\mathop{\mathrm{ch}}(G)\geq(1+o(1))\cdot\log_{2}d, where the o(1)o(1) term tends to 0 when dd tends to infinity. Erdös et al. [8] proved that the choice number of the complete bipartite graph Kn,nK_{n,n} satisfies ch(Kn,n)=(1+o(1))log2n\mathop{\mathrm{ch}}(K_{n,n})=(1+o(1))\cdot\log_{2}n, hence the lower bound of [17] is tight on these graphs. Theorem 1.2 thus shows a substantial difference between the behavior of the subspace choice number and that of the standard choice number in terms of the average degree.

For the complete graph KnK_{n}, it is easy to see that ch-s(Kn,𝔽)=n\mathop{\mathrm{ch\text{-}s}}(K_{n},\mathbb{F})=n whenever 𝔽\mathbb{F} is a field over which no nonzero vector is self-orthogonal, such as \mathbb{R} and \mathbb{C}. For finite fields, however, we show that the subspace choice number of KnK_{n} is strictly smaller than nn for every sufficiently large nn. This in particular shows that the subspace choice number over finite fields can be smaller than the choice number.

Theorem 1.3.

There exists a constant c>0c>0 such that for every sufficiently large integer nn and for every finite field 𝔽\mathbb{F},

ch-s(Kn,𝔽)ncn.\mathop{\mathrm{ch\text{-}s}}(K_{n},\mathbb{F})\leq n-c\cdot\sqrt{n}.

We next put our focus on complete bipartite graphs. For the color choosability problem, it was observed in [8] that the graph Kk,mK_{k,m} is kk-choosable for every m<kkm<k^{k} whereas ch(Kk,m)=k+1\mathop{\mathrm{ch}}(K_{k,m})=k+1 for every mkkm\geq k^{k}. Considering the subspace choice number of these graphs, for every field 𝔽\mathbb{F} it holds that ch-s(Kk,m,𝔽)k+1\mathop{\mathrm{ch\text{-}s}}(K_{k,m},\mathbb{F})\leq k+1, because Kk,mK_{k,m} is kk-degenerate. We consider here the problem of identifying the values of mm for which this k+1k+1 upper bound is tight. Namely, for an integer kk and a field 𝔽\mathbb{F}, let m(k,𝔽)m(k,\mathbb{F}) denote the smallest integer mm for which it holds that ch-s(Kk,m,𝔽)=k+1\mathop{\mathrm{ch\text{-}s}}(K_{k,m},\mathbb{F})=k+1. We provide the following lower bound.

Theorem 1.4.

For every integer kk and for every field 𝔽\mathbb{F},

m(k,𝔽)>i=1k1k1i.m(k,\mathbb{F})>\sum_{i=1}^{k-1}{\Big{\lfloor}\frac{k-1}{i}\Big{\rfloor}}.

In particular, for every field 𝔽\mathbb{F} it holds that m(k,𝔽)=Ω(klogk)m(k,\mathbb{F})=\Omega(k\cdot\log k).

We next provide a general approach for proving upper bounds on m(k,𝔽)m(k,\mathbb{F}). The following theorem reduces this challenge to constructing families of vectors with certain linear independence constraints.

Theorem 1.5.

If there exists a collection of m=k(t1)+1m=k\cdot(t-1)+1 nonzero vectors in 𝔽k\mathbb{F}^{k} satisfying that every tt of them span the entire space 𝔽k\mathbb{F}^{k}, then m(k,𝔽)mm(k,\mathbb{F})\leq m.

The above theorem allows us to derive upper bounds on m(k,𝔽)m(k,\mathbb{F}) for various fields 𝔽\mathbb{F}.

Corollary 1.6.

Let kk be an integer and let 𝔽\mathbb{F} be a field.

  1. 1.

    If |𝔽|k2k+1|\mathbb{F}|\geq k^{2}-k+1 then m(k,𝔽)k2k+1m(k,\mathbb{F})\leq k^{2}-k+1.

  2. 2.

    If 𝔽\mathbb{F} is a finite field of size qkq\geq k then m(k,𝔽)kqk11q1+1m(k,\mathbb{F})\leq k\cdot\frac{q^{k-1}-1}{q-1}+1.

We remark that the first item of Corollary 1.6 is obtained by applying Theorem 1.5 with collections of vectors that form the columns of Vandermonde matrices. It implies that m(k,𝔽)=O(k2)m(k,\mathbb{F})=O(k^{2}) whenever the field 𝔽\mathbb{F} is infinite or sufficiently large as a function of kk, leaving us with a nearly quadratic gap from the lower bound given in Theorem 1.4. This again demonstrates a significant difference between the behavior of the choice number and that of the subspace choice number.

In fact, Theorems 1.4 and 1.5 are proved in a more general form with respect to asymmetric subspace assignments, where the left and right vertices of the complete bipartite graphs might be assigned subspaces of different dimensions. For the precise generalized statements, see Theorems 5.4 and 5.7. We note that this is analogous to the asymmetric setting of color choosability that was recently studied by Alon, Cambie, and Kang [3].

We particularly consider the bipartite graph K2,mK_{2,m} whose left side consists of only two vertices. For an integer nn, we say that K2,mK_{2,m} is (n;2,2)(n;2,2)-subspace choosable over a field 𝔽\mathbb{F} if it is ff-subspace choosable over 𝔽\mathbb{F} for the function ff that assigns the integer nn to one vertex of the left side and the integer 22 to each of the other vertices. We consider the problem of determining, for a given integer nn, the smallest mm for which K2,mK_{2,m} is (n;2,2)(n;2,2)-subspace choosable over a given field 𝔽\mathbb{F}, and prove the following.

Theorem 1.7.

For every integer n1n\geq 1 the following holds.

  1. 1.

    The graph K2,n1K_{2,n-1} is (n;2,2)(n;2,2)-subspace choosable over every field 𝔽\mathbb{F}.

  2. 2.

    The graph K2,nK_{2,n} is (n;2,2)(n;2,2)-subspace choosable over \mathbb{C}.

  3. 3.

    The graph K2,nK_{2,n} is (n;2,2)(n;2,2)-subspace choosable over \mathbb{R} if and only if nn is odd.

We finally consider the computational aspect of the subspace choice number and prove the following hardness result.

Theorem 1.8.

Let k3k\geq 3 be an integer and let 𝔽\mathbb{F} be either \mathbb{R} or some finite field. Then, the problem of deciding whether a given bipartite graph GG satisfies ch-s(G,𝔽)k\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F})\leq k is 𝖭𝖯\mathsf{NP}-hard.

The proof of Theorem 1.8 is inspired by the approach taken in a proof due to Rubin [8] for the Π2\Pi_{2}-hardness of the decision problem associated with the (color) choice number. His proof involves a delicate construction of several gadget graphs used to efficiently map an instance of the Π2\Pi_{2}-variant of the satisfiability problem to an instance of the color choosability problem. These gadgets, however, do not fit the setting of subspace choosability. In fact, the characterization of 22-subspace choosable graphs over the reals, given in [13], implies that the instances produced by the reduction of [8] are never subspace choosable over this field. To overcome this difficulty, we construct and analyze a different gadget graph that allows us, combined with ideas of Gutner and Tarsi [10, 11], to obtain the 𝖭𝖯\mathsf{NP}-hardness result stated in Theorem 1.8. Our analysis involves a characterization, stated below, of the 22-subspace choosable graphs over finite fields, extending the characterizations given in [13] for the real and complex fields.

Proposition 1.9.

For every finite field 𝔽\mathbb{F}, a graph is 22-subspace choosable over 𝔽\mathbb{F} if and only if it contains no cycles.

While Theorem 1.8 indicates the hardness of efficiently determining the subspace choice number of bipartite graphs, it would be natural to expect the stronger notion of Π2\Pi_{2}-hardness to hold for this problem.

1.2 Outline

The rest of the paper is organized as follows. In Section 2, we prove Theorem 1.2, relating the subspace choice number of a graph over a general field to its average degree. We also prove there an improved bound for certain graphs and discuss a limitation of our approach. In section 3, we prove the upper bound on the subspace choice number of complete graphs over finite fields given in Theorem 1.3. In Section 4, we prove the characterization of 22-subspace choosable graphs over finite fields given in Proposition 1.9, which will be used in the following sections. In Section 5, we prove several upper and lower bounds on the subspace choosability of complete bipartite graphs in the asymmetric setting, and in particular confirm Theorems 1.41.5, and 1.7. Finally, in Section 6, we prove our hardness result given in Theorem 1.8.

2 Subspace Choosability and Average Degree

In this section we relate the subspace choice number of a graph over a general field to its average degree and prove Theorem 1.2. We start with the following definition of kk-partitioned graphs (for an example, see Lemma 2.6).

Definition 2.1.

Let G=(V,E)G=(V,E) be a graph. For every vertex vVv\in V, let EvEE_{v}\subseteq E denote the set of edges of GG that are incident with vv. We say that the graph GG is kk-partitioned if it is possible to partition every set EvE_{v}, vVv\in V, into kk sets Ev(1),,Ev(k)E_{v}^{(1)},\ldots,E_{v}^{(k)} (some of which may be empty), such that for every function g:V[k]g:V\rightarrow[k] there exist two adjacent vertices v1,v2Vv_{1},v_{2}\in V such that {v1,v2}Ev1(g(v1))Ev2(g(v2))\{v_{1},v_{2}\}\in E_{v_{1}}^{(g(v_{1}))}\cap E_{v_{2}}^{(g(v_{2}))}.

The following theorem shows that the subspace choice number of a kk-partitioned graph exceeds kk over any field.

Theorem 2.2.

For every kk-partitioned graph GG and for every field 𝔽\mathbb{F}, ch-s(G,𝔽)>k\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F})>k.

  • Proof:

    Fix an arbitrary field 𝔽\mathbb{F}. Let G=(V,E)G=(V,E) be a kk-partitioned graph, and for every vertex vVv\in V, let Ev=Ev(1)Ev(k)E_{v}=E_{v}^{(1)}\cup\cdots\cup E_{v}^{(k)} be the corresponding partition of the edges incident with vv, as in Definition 2.1. We use these partitions to define a kk-subspace assignment over 𝔽\mathbb{F} to the vertices of GG involving vectors from the space 𝔽|E|\mathbb{F}^{|E|}, where each entry corresponds to an edge eEe\in E. To a vertex vVv\in V we assign the subspace WvW_{v} spanned by the kk vectors wv(1),,wv(k)w^{(1)}_{v},\ldots,w^{(k)}_{v}, where wv(i)w^{(i)}_{v} is the 0,10,1 indicator vector of the subset Ev(i)E_{v}^{(i)} of EE. In fact, some of the sets Ev(i)E^{(i)}_{v} might be empty, and thus some of the vectors wv(i)w^{(i)}_{v} might be zeros, resulting in subspaces WvW_{v} of dimension smaller than kk. To fix it, one can increase the length of the vectors from |E||E| to |E|+k|V||E|+k\cdot|V| and to add to each of the k|V|k\cdot|V| vectors wv(i)w^{(i)}_{v} a nonzero entry in a coordinate on which all the others have zeros. These entries ensure that the dimension of every subspace WvW_{v} is precisely kk. For simplicity of notation, we refer from now on to these modified vectors as wv(i)w^{(i)}_{v}.

    We show now that no choice of nonzero vectors from these subspaces satisfies that every two adjacent vertices receive orthogonal vectors over 𝔽\mathbb{F}. To see this, consider some choice of a nonzero vector xvWvx_{v}\in W_{v} for each vertex vVv\in V. We define a function g:V[k]g:V\rightarrow[k] as follows. For every vVv\in V, xvx_{v} is a nonzero linear combination of the vectors wv(1),,wv(k)w^{(1)}_{v},\ldots,w^{(k)}_{v}, hence there exists some jv[k]j_{v}\in[k] for which the coefficient of wv(jv)w^{(j_{v})}_{v} in this linear combination is nonzero. We define g(v)g(v) to be such an index jvj_{v}. By assumption, there exist two adjacent vertices v1,v2Vv_{1},v_{2}\in V such that {v1,v2}Ev1(g(v1))Ev2(g(v2))\{v_{1},v_{2}\}\in E_{v_{1}}^{(g(v_{1}))}\cap E_{v_{2}}^{(g(v_{2}))}. This implies that the entry that corresponds to the edge {v1,v2}\{v_{1},v_{2}\} of GG is nonzero in both xv1x_{v_{1}} and xv2x_{v_{2}}. However, the supports of the subspaces Wv1W_{v_{1}} and Wv2W_{v_{2}} intersect at this single entry, implying that the vectors xv1x_{v_{1}} and xv2x_{v_{2}} are not orthogonal over 𝔽\mathbb{F}. This implies that there exists a kk-subspace assignment over 𝔽\mathbb{F} to the vertices of GG with no appropriate choice of nonzero vectors, yielding that ch-s(G,𝔽)>k\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F})>k, as required.  

Theorem 2.2 motivates the problem of determining the largest integer kk for which a given graph is kk-partitioned. The following lemma uses a probabilistic argument to prove a lower bound on this quantity in terms of the average degree.

Lemma 2.3.

There exists a constant c>0c>0 such that every graph with average degree d>1d>1 is kk-partitioned for some kcdlndk\geq c\cdot\sqrt{\frac{d}{\ln d}}.

  • Proof:

    Let G=(V,E)G=(V,E) be a graph with average degree d>1d>1. Note that 2|E|=|V|d2\cdot|E|=|V|\cdot d. Let kk be the largest integer satisfying

    d>2k2lnk.\displaystyle d>2k^{2}\cdot\ln k. (1)

    Observe that for an appropriate choice of the constant cc, it holds that kcdlndk\geq c\cdot\sqrt{\frac{d}{\ln d}}. We prove that GG is kk-partitioned by a probabilistic argument. For every vertex vVv\in V, we define a random partition of the set EvE_{v} of the edges incident with vv, into kk sets Ev(1),,Ev(k)E_{v}^{(1)},\ldots,E_{v}^{(k)} (some of which may be empty) as follows. For each edge eEve\in E_{v}, we pick at random, uniformly and independently, some j[k]j\in[k], and put ee in Ev(j)E_{v}^{(j)}. We claim that the obtained partitions satisfy with positive probability the condition given in Definition 2.1, namely, that for every function g:V[k]g:V\rightarrow[k] there exist two adjacent vertices v1,v2Vv_{1},v_{2}\in V such that

    {v1,v2}Ev1(g(v1))Ev2(g(v2)).\displaystyle\{v_{1},v_{2}\}\in E_{v_{1}}^{(g(v_{1}))}\cap E_{v_{2}}^{(g(v_{2}))}. (2)

    Indeed, for every fixed function g:V[k]g:V\rightarrow[k] and for every edge {v1,v2}E\{v_{1},v_{2}\}\in E, the probability that the event (2) occurs is 1/k21/k^{2}. Hence, the probability that for all edges of EE this event does not occur is (11/k2)|E|(1-1/k^{2})^{|E|}. By the union bound, the probability that there exists a function g:V[k]g:V\rightarrow[k] such that for all edges of EE the event (2) does not occur is at most

    k|V|(11k2)|E|k|V|e|E|/k2=(elnkd/(2k2))|V|.k^{|V|}\cdot\bigg{(}1-\frac{1}{k^{2}}\bigg{)}^{|E|}\leq k^{|V|}\cdot e^{-|E|/k^{2}}=\big{(}e^{\ln k-d/(2k^{2})}\big{)}^{|V|}.

    By (1), the above is smaller than 11, hence with positive probability the random partition satisfies the required condition, and thus GG is kk-partitioned  

Combining Theorem 2.2 and Lemma 2.3 completes the proof of Theorem 1.2.

It is natural to ask whether Theorem 2.2 can be used to obtain better lower bounds on the subspace choice number of graphs than the one achieved by Theorem 1.2. The following lemma shows that for graphs with similar average and maximum degrees, Lemma 2.3 is tight up to the logarithmic term. Hence, the approach suggested by Theorem 2.2 cannot yield significantly better bounds for such graphs.

Lemma 2.4.

There exists a constant c>0c>0 such that every graph with maximum degree DD is not kk-partitioned whenever kcDk\geq c\cdot\sqrt{D}.

The proof of Lemma 2.4 uses the Lovász local lemma stated below (see, e.g., [4, Chapter 5]).

Lemma 2.5 (Lovász Local Lemma).

Let {\cal E} be a collection of events such that for each AA\in{\cal E}, it holds that Pr[A]p<1{\Pr\left[{A}\right]}\leq p<1 and that AA is mutually independent of a set of all but at most dd of the other events of {\cal E}. If ep(d+1)1e\cdot p\cdot(d+1)\leq 1, then with positive probability none of the events of {\cal E} occurs.

  • Proof of Lemma 2.4:

    Let G=(V,E)G=(V,E) be a graph with maximum degree DD, and let kcDk\geq c\cdot\sqrt{D} be an integer for some constant cc to be determined. We prove that GG is not kk-partitioned by a probabilistic argument. For every vertex vVv\in V, consider a partition Ev=Ev(1)Ev(k)E_{v}=E_{v}^{(1)}\cup\cdots\cup E_{v}^{(k)} of the set of the edges incident with vv into kk sets (some of the sets may be empty). We claim that there exists a function g:V[k]g:V\rightarrow[k] such that no edge {v1,v2}\{v_{1},v_{2}\} of GG satisfies {v1,v2}Ev1(g(v1))Ev2(g(v2))\{v_{1},v_{2}\}\in E_{v_{1}}^{(g(v_{1}))}\cap E_{v_{2}}^{(g(v_{2}))}. To prove it, consider a random function g:V[k]g:V\rightarrow[k] such that each value g(v)g(v) for vVv\in V is chosen uniformly and independently at random from [k][k]. For every edge e={v1,v2}Ee=\{v_{1},v_{2}\}\in E, let AeA_{e} denote the event that eEv1(g(v1))Ev2(g(v2))e\in E_{v_{1}}^{(g(v_{1}))}\cap E_{v_{2}}^{(g(v_{2}))}. The probability of each event AeA_{e} is clearly 1/k21/k^{2}. In addition, every event AeA_{e} is mutually independent of the set of all the other events AeA_{e^{\prime}} but those satisfying eee\cap e^{\prime}\neq\emptyset, whose number is at most 2(D1)2\cdot(D-1). By the Lovász local lemma (Lemma 2.5), it follows that if

    e1k2(2D1)1e\cdot\tfrac{1}{k^{2}}\cdot(2D-1)\leq 1

    then with positive probability no event AeA_{e} occurs. This implies that for an appropriate choice of the constant cc, there exists a function gg with the required property. Since this holds for all possible partitions of the sets EvE_{v} into kk sets, it follows that GG is not kk-partitioned, and we are done.  

We end this section by proving that for certain graphs, the logarithmic term in Lemma 2.3 can be avoided. Here, the proof does not use a probabilistic construction of partitions, but an explicit one, based on finite projective planes.

Lemma 2.6.

For a prime power qq, let HH be the (q+1)(q+1)-partite graph with qq vertices in every part. Let GG be a graph obtained from HH by removing at most q1q-1 of its vertices. Then, GG is qq-partitioned.

  • Proof:

    The proof is based on a well-known construction of projective planes, some of whose properties are described next (see, e.g., [6, Chapter 9]). For every prime power qq, there exists a collection of n=q2+q+1n=q^{2}+q+1 elements called points, and nn sets of points, called lines, satisfying that every two lines intersect at a single point, every two points belong together to a single line, every point belongs to precisely q+1q+1 of the lines, and every line includes precisely q+1q+1 of the points. Fix some point pp, let L1,,Lq+1L_{1},\ldots,L_{q+1} be the q+1q+1 lines that include pp, and put Li=Li{p}L^{\prime}_{i}=L_{i}\setminus\{p\} for every i[q+1]i\in[q+1]. Note that the sets LiL^{\prime}_{i} are pairwise disjoint. We view the graph HH as the graph on the vertex set i[q+1]Li\cup_{i\in[q+1]}{L^{\prime}_{i}} in which two vertices are adjacent if they belong to distinct sets LiL^{\prime}_{i}. Observe that every two vertices of HH are adjacent if and only if the line that includes their points does not include pp.

    Let G=(V,E)G=(V,E) be some subgraph of HH obtained by removing at most q1q-1 of its vertices, and observe that the number of its vertices satisfies

    |V|(q+1)q(q1)=q2+1.|V|\geq(q+1)\cdot q-(q-1)=q^{2}+1.

    We show that GG is qq-partitioned. To do so, we assign to every edge of the graph GG the line that includes the points represented by its vertices. This assignment induces for every vertex vVv\in V a partition of the set EvE_{v} of the edges incident with vv in GG, where the sets of the partition correspond to the lines associated with the edges. Observe that this partition of EvE_{v} consists of at most qq sets. Indeed, the vertex vv represents a point that belongs to q+1q+1 lines, but no edge of EvE_{v} is assigned the line that includes the point pp and the point of vv.

    In order to show that these partitions satisfy the condition of Definition 2.1, we shall verify that if one chooses for every point represented by a vertex in GG a line that corresponds to an edge incident with it, then there exist two adjacent vertices in GG for which the same line was chosen. This indeed follows from the fact that no edge of GG corresponds to a line that includes pp, hence the total number of lines associated with the edges of GG is at most q2q^{2}. Since the number of vertices in GG exceeds q2q^{2}, it follows that two vertices are assigned the same line. Since this line does not include pp, the two vertices must be adjacent in GG, completing the proof.  

Note that the graph HH from Lemma 2.6 is regular with degree q2q^{2}, hence the minimum degree of its subgraph GG is at least q2q+1q^{2}-q+1, and yet GG is qq-partitioned. This shows that the logarithmic term from Lemma 2.3 is not needed for GG. By combining Lemma 2.6 with Theorem 2.2, we derive the following.

Theorem 2.7.

For a prime power qq, let HH be the (q+1)(q+1)-partite graph with qq vertices in every part. Let GG be a graph obtained from HH by removing at most q1q-1 of its vertices. Then, for every field 𝔽\mathbb{F}, ch-s(G,𝔽)>q\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F})>q.

3 Subspace Choosability in Complete Graphs over Finite Fields

In this section we prove Theorem 1.3, which provides an upper bound on the subspace choice number of complete graphs over finite fields. We start with two useful lemmas.

Lemma 3.1.

For a finite field 𝔽\mathbb{F} and an integer tt, let w1,w2,w3w_{1},w_{2},w_{3} and z1,z2,z3z_{1},z_{2},z_{3} be two triples of vectors in 𝔽t\mathbb{F}^{t}. Then, there exist α1,α2,α3𝔽\alpha_{1},\alpha_{2},\alpha_{3}\in\mathbb{F}, not all zeros, such that

i[3]αiwi,i[3]αizi=0.\big{\langle}\sum_{i\in[3]}{\alpha_{i}\cdot w_{i}},\sum_{i\in[3]}{\alpha_{i}\cdot z_{i}}\big{\rangle}=0.
  • Proof:

    Consider the function f:𝔽3𝔽f:\mathbb{F}^{3}\rightarrow\mathbb{F} defined by f(α1,α2,α3)=i[3]αiwi,i[3]αizif(\alpha_{1},\alpha_{2},\alpha_{3})=\big{\langle}\sum_{i\in[3]}{\alpha_{i}\cdot w_{i}},\sum_{i\in[3]}{\alpha_{i}\cdot z_{i}}\big{\rangle}. The function ff is a degree 22 polynomial on 33 variables over 𝔽\mathbb{F}, and (0,0,0)(0,0,0) forms a root of ff. The Chevalley theorem (see, e.g., [18, Chapter IV, Theorem 1D]) implies that ff has another root, as required.  

Lemma 3.2.

For a finite field 𝔽\mathbb{F} and an integer tt, let U1,U2,U3U_{1},U_{2},U_{3} be three subspaces of 𝔽t\mathbb{F}^{t} whose dimensions satisfy

dim(U1)2,dim(U2)2,anddim(U1)+dim(U2)dim(U3)5.\dim(U_{1})\geq 2,~{}~{}\dim(U_{2})\geq 2,~{}~{}\mbox{and}~{}~{}\dim(U_{1})+\dim(U_{2})-\dim(U_{3})\geq 5.

Then, there exist nonzero vectors x1U1x_{1}\in U_{1} and x2U2x_{2}\in U_{2} such that x1,x2=0\langle x_{1},x_{2}\rangle=0 and

dim(U3(x1)(x2))dim(U3)1.\dim(U_{3}\cap(x_{1})^{\perp}\cap(x_{2})^{\perp})\geq\dim(U_{3})-1.
  • Proof:

    Let U1,U2,U3𝔽tU_{1},U_{2},U_{3}\subseteq\mathbb{F}^{t} be three subspaces as in the statement of the lemma.

    Assume first that dim(U1U2)3\dim(U_{1}\cap U_{2})\geq 3. In this case, there exist three linearly independent vectors in U1U2U_{1}\cap U_{2}. By Lemma 3.1, there exists a nonzero self-orthogonal linear combination of them. By choosing x1x_{1} and x2x_{2} to be this vector, it obviously holds that x1,x2=0\langle x_{1},x_{2}\rangle=0 and that

    dim(U3(x1)(x2))=dim(U3(x1))dim(U3)1,\dim(U_{3}\cap(x_{1})^{\perp}\cap(x_{2})^{\perp})=\dim(U_{3}\cap(x_{1})^{\perp})\geq\dim(U_{3})-1,

    as required.

    Assume next that dim(U1U3)1\dim(U_{1}\cap U_{3}^{\perp})\geq 1. Here, x1x_{1} can be chosen as an arbitrary nonzero vector of U1U3U_{1}\cap U_{3}^{\perp}, and x2x_{2} as an arbitrary nonzero vector of U2U_{2} satisfying x1,x2=0\langle x_{1},x_{2}\rangle=0. Such a vector exists because dim(U2)2\dim(U_{2})\geq 2. By x1U3x_{1}\in U_{3}^{\perp}, it follows that

    dim(U3(x1)(x2))=dim(U3(x2))dim(U3)1.\dim(U_{3}\cap(x_{1})^{\perp}\cap(x_{2})^{\perp})=\dim(U_{3}\cap(x_{2})^{\perp})\geq\dim(U_{3})-1.

    The case dim(U2U3)1\dim(U_{2}\cap U_{3}^{\perp})\geq 1 is handled similarly.

    Otherwise, we have dim(U1U2)2\dim(U_{1}\cap U_{2})\leq 2 and dim(U1U3)=dim(U2U3)=0\dim(U_{1}\cap U_{3}^{\perp})=\dim(U_{2}\cap U_{3}^{\perp})=0. This implies that

    dim(U1+U2)\displaystyle\dim(U_{1}+U_{2}) =\displaystyle= dim(U1)+dim(U2)dim(U1U2)\displaystyle\dim(U_{1})+\dim(U_{2})-\dim(U_{1}\cap U_{2})
    \displaystyle\geq dim(U1)+dim(U2)2dim(U3)+3,\displaystyle\dim(U_{1})+\dim(U_{2})-2\geq\dim(U_{3})+3,

    where for the last inequality we have used the assumption dim(U1)+dim(U2)dim(U3)5\dim(U_{1})+\dim(U_{2})-\dim(U_{3})\geq 5. It thus follows that

    dim((U1+U2)U3)\displaystyle\dim((U_{1}+U_{2})\cap U_{3}^{\perp}) =\displaystyle= dim(U1+U2)+dim(U3)dim(U1+U2+U3)\displaystyle\dim(U_{1}+U_{2})+\dim(U_{3}^{\perp})-\dim(U_{1}+U_{2}+U_{3}^{\perp})
    \displaystyle\geq dim(U1+U2)+(tdim(U3))t3.\displaystyle\dim(U_{1}+U_{2})+(t-\dim(U_{3}))-t\geq 3.

    Hence, there exist vectors w1,w2,w3U1w_{1},w_{2},w_{3}\in U_{1} and z1,z2,z3U2z_{1},z_{2},z_{3}\in U_{2} for which the three sums

    w1+z1,w2+z2,w3+z3w_{1}+z_{1},~{}w_{2}+z_{2},~{}w_{3}+z_{3}

    are linearly independent vectors that belong to U3U_{3}^{\perp}. By Lemma 3.1, there exist α1,α2,α3𝔽\alpha_{1},\alpha_{2},\alpha_{3}\in\mathbb{F}, not all zeros, such that the vectors x1=i[3]αiwix_{1}=\sum_{i\in[3]}{\alpha_{i}\cdot w_{i}} and x2=i[3]αizix_{2}=\sum_{i\in[3]}{\alpha_{i}\cdot z_{i}} satisfy x1,x2=0\langle x_{1},x_{2}\rangle=0. These vectors further satisfy that

    x1+x2=i[3]αi(wi+zi)U3.x_{1}+x_{2}=\sum_{i\in[3]}{\alpha_{i}\cdot(w_{i}+z_{i})}\in U_{3}^{\perp}.

    It follows that x1+x2x_{1}+x_{2} is nonzero, because the vectors wi+ziw_{i}+z_{i} are linearly independent. Observe that x1x_{1} belongs to U1U_{1} and that it is nonzero, because otherwise the vector x2x_{2} would be a nonzero vector that belongs to U2U3U_{2}\cap U_{3}^{\perp}, in contradiction to dim(U2U3)=0\dim(U_{2}\cap U_{3}^{\perp})=0. By the same reasoning, x2x_{2} is a nonzero vector of U2U_{2}. Finally, notice that for every vector uU3u\in U_{3} such that u,x1=0\langle u,x_{1}\rangle=0, it also holds that u,x2=0\langle u,x_{2}\rangle=0, and thus

    dim(U3(x1)(x2))=dim(U3(x1))dim(U3)1,\dim(U_{3}\cap(x_{1})^{\perp}\cap(x_{2})^{\perp})=\dim(U_{3}\cap(x_{1})^{\perp})\geq\dim(U_{3})-1,

    so we are done.  

Remark 3.3.

It can be shown that for the binary field 𝔽2\mathbb{F}_{2}, the third condition of Lemma 3.2 can be slightly weakened to dim(U1)+dim(U2)dim(U3)4\dim(U_{1})+\dim(U_{2})-\dim(U_{3})\geq 4.

We are ready to prove the following result, which implies Theorem 1.3.

Theorem 3.4.

For an integer k1k\geq 1, put n=k2+2k+3n=k^{2}+2k+3. Then, for every finite field 𝔽\mathbb{F},

ch-s(Kn,𝔽)nk.\mathop{\mathrm{ch\text{-}s}}(K_{n},\mathbb{F})\leq n-k.
  • Proof:

    For an integer k1k\geq 1 and a finite field 𝔽\mathbb{F}, consider the complete graph KnK_{n} on n=k2+2k+3n=k^{2}+2k+3 vertices. Let V=ABCV=A\cup B\cup C be the vertex set of the graph, where A={v1,,vk}A=\{v_{1},\ldots,v_{k}\} is a set of kk vertices, BB is a set of k2+kk^{2}+k vertices, and CC consists of the three remaining vertices. To prove that ch-s(Kn,𝔽)nk\mathop{\mathrm{ch\text{-}s}}(K_{n},\mathbb{F})\leq n-k, suppose that for some integer tt, we are given a subspace Uv𝔽tU_{v}\subseteq\mathbb{F}^{t} with dim(Uv)=nk\dim(U_{v})=n-k for every vVv\in V. Our goal is to show that there exist pairwise orthogonal nonzero vectors xvUvx_{v}\in U_{v} for vVv\in V. We describe now a process with several steps for choosing the vectors. Throughout the process we maintain for every vertex vVv\in V a subspace UvU^{\prime}_{v} defined as the subspace of the vectors currently available to the vertex vv. Namely, for every partial choice of vectors, UvU^{\prime}_{v} is the subspace of UvU_{v} that consists of all the vectors of UvU_{v} that are orthogonal to all the previously chosen vectors. Initially, we have Uv=UvU^{\prime}_{v}=U_{v} for every vVv\in V.

    Consider some partition of the set BB into kk sets, B=B1BkB=B_{1}\cup\cdots\cup B_{k}, where |Bi|=2(ki+1)|B_{i}|=2\cdot(k-i+1) for every i[k]i\in[k]. Note that this is possible, because |B|=k2+k=2i=1k(ki+1)|B|=k^{2}+k=2\cdot\sum_{i=1}^{k}({k-i+1)}. Our process starts with kk initial steps, where the role of the iith step (i[k]i\in[k]) is to choose vectors for the vertices of BiB_{i} in a way that poses only ki+1k-i+1 linear constraints on the choice of the vector for viv_{i}. Note that for the other vertices, the choice of the vectors for the vertices of BiB_{i} might pose twice this number of linear constraints.

    For i[k]i\in[k], the iith step is performed as follows. Consider an arbitrary partition of the set BiB_{i} into ki+1k-i+1 pairs, denoted by (a1,b1),,(aki+1,bki+1)(a_{1},b_{1}),\ldots,(a_{k-i+1},b_{k-i+1}). For every j[ki+1]j\in[k-i+1], we choose two nonzero vectors uajUaju_{a_{j}}\in U^{\prime}_{a_{j}} and ubjUbju_{b_{j}}\in U^{\prime}_{b_{j}} such that uaj,ubj=0\langle u_{a_{j}},u_{b_{j}}\rangle=0 and

    dim(Uvi(uaj)(ubj))dim(Uvi)1.\dim(U^{\prime}_{v_{i}}\cap(u_{a_{j}})^{\perp}\cap(u_{b_{j}})^{\perp})\geq\dim(U^{\prime}_{v_{i}})-1.

    Observe that such a choice, if it exists, satisfies that uaju_{a_{j}} and ubju_{b_{j}} are nonzero vectors that belong to the subspaces of the vertices aja_{j} and bjb_{j} respectively, they are orthogonal to all the previously chosen vectors and to one another, and in addition, their choice reduces the dimension of UviU^{\prime}_{v_{i}} by at most 11. To prove the existence of such a choice we apply Lemma 3.2. The number of vectors chosen before the (i,j)(i,j) iteration is l=1i1|Bl|+2(j1)\sum_{l=1}^{i-1}{|B_{l}|}+2(j-1), hence each of dim(Uaj)\dim(U^{\prime}_{a_{j}}) and dim(Ubj)\dim(U^{\prime}_{b_{j}}) is at least

    (nk)l=1i1|Bl|2(j1).(n-k)-\sum_{l=1}^{i-1}{|B_{l}|}-2(j-1).

    Additionally, since the 2(j1)2(j-1) already chosen vectors of the iith step reduce the dimension of UviU^{\prime}_{v_{i}} by at most j1j-1, it can be assumed that

    dim(Uvi)=(nk)l=1i1|Bl|(j1).\dim(U^{\prime}_{v_{i}})=(n-k)-\sum_{l=1}^{i-1}{|B_{l}|}-(j-1).

    It thus follows that in the (i,j)(i,j) iteration, it holds that

    dim(Uaj)+dim(Ubj)dim(Uvi)\displaystyle\dim(U^{\prime}_{a_{j}})+\dim(U^{\prime}_{b_{j}})-\dim(U^{\prime}_{v_{i}}) \displaystyle\geq (nk)l=1i1|Bl|3(j1)\displaystyle(n-k)-\sum_{l=1}^{i-1}{|B_{l}|}-3(j-1)
    =\displaystyle= l=ik|Bl|+33(j1)\displaystyle\sum_{l=i}^{k}{|B_{l}|}+3-3(j-1)
    =\displaystyle= (ki+1)(ki+2)3j+6\displaystyle(k-i+1)(k-i+2)-3j+6
    \displaystyle\geq (ki+1)(ki+2)3(ki+1)+6\displaystyle(k-i+1)(k-i+2)-3(k-i+1)+6
    \displaystyle\geq (ki+1)(ki1)+6=(ki)2+55,\displaystyle(k-i+1)(k-i-1)+6=(k-i)^{2}+5\geq 5,

    where for the first equality we use the fact that n=|B|+k+3n=|B|+k+3, and for the second inequality we use the fact jki+1j\leq k-i+1. The above bound, which also implies that dim(Uaj)2\dim(U_{a_{j}})\geq 2 and that dim(Ubj)2\dim(U_{b_{j}})\geq 2, allows us to apply Lemma 3.2 and to obtain the required vectors uaju_{a_{j}} and ubju_{b_{j}}.

    We next show that given the above choice for the vertices of BB, one can choose vectors for the vertices of ACA\cup C to obtain the required pairwise orthogonal vectors. First, for the three vertices of CC, choose arbitrary pairwise orthogonal nonzero vectors from the currently available subspaces. This is indeed possible, because so far we chose n(k+3)n-(k+3) vectors, so the dimension of the subspace available to each of them is at least 33. The choice for the first one leaves the available subspaces of the other two with dimension at least 22, and the choice of the second one leaves the available subspace of the third with dimension at least 11, allowing us to choose its nonzero vector.

    Finally, we choose the vectors for the vertices of AA. For each i[k]i\in[k], among the nkn-k vectors chosen so far, there are ki+1k-i+1 pairs of vectors whose choice reduced the dimension of UviU^{\prime}_{v_{i}} by at most 11. This implies that we currently have

    dim(Uvi)(nk)((nk)(ki+1))=ki+1.\dim(U^{\prime}_{v_{i}})\geq(n-k)-((n-k)-(k-i+1))=k-i+1.

    This allows us to go over the vertices vk,vk1,,v1v_{k},v_{k-1},\ldots,v_{1}, in this order, and to choose a nonzero vector from the subspace currently available to each of them, completing the proof.  

Remark 3.5.

For the binary field 𝔽2\mathbb{F}_{2}, it can be shown that ch-s(Kn,𝔽2)nk\mathop{\mathrm{ch\text{-}s}}(K_{n},\mathbb{F}_{2})\leq n-k for n=k2+2k+2n=k^{2}+2k+2. This follows by applying the above proof with the version of Lemma 3.2 mentioned in Remark 3.3.

4 Characterization of 22-Subspace Choosable Graphs

In this section we prove Proposition 1.9, which asserts that for every finite field 𝔽\mathbb{F} and for every graph GG, ch-s(G,𝔽)2\mathop{\mathrm{ch\text{-}s}}(G,\mathbb{F})\leq 2 if and only if GG contains no cycles.

  • Proof of Proposition 1.9:

    If GG contains no cycles then it is 11-degenerate, implying that it is 22-subspace choosable over every field 𝔽\mathbb{F}. To complete the proof, we fix some finite field 𝔽\mathbb{F} and turn to show that for every 3\ell\geq 3, the \ell-cycle CC_{\ell} satisfies ch-s(C,𝔽)>2\mathop{\mathrm{ch\text{-}s}}(C_{\ell},\mathbb{F})>2. We first prove it for =3\ell=3 and for =4\ell=4.

    • For =3\ell=3, assign to the vertices of the cycle C3C_{3} the subspaces of 𝔽3\mathbb{F}^{3} defined by

      U1=span(e1,e2),U2=span(e1,e2+e3), and U3=span(e1+αe3,e2),U_{1}=\mathop{\mathrm{span}}(e_{1},e_{2}),~{}~{}~{}U_{2}=\mathop{\mathrm{span}}(e_{1},e_{2}+e_{3}),\mbox{~{}~{}~{}and~{}~{}~{}}U_{3}=\mathop{\mathrm{span}}(e_{1}+\alpha\cdot e_{3},e_{2}),

      where α𝔽\alpha\in\mathbb{F} is some field element to be determined. We claim that for some α𝔽\alpha\in\mathbb{F} it is impossible to choose three pairwise orthogonal nonzero vectors xiUix_{i}\in U_{i} (i[3]i\in[3]). Indeed, it is easy to verify that x1x_{1} cannot be chosen as a scalar multiple of e1e_{1} nor of e2e_{2}. So assume without loss of generality that x1x_{1} is proportional to e1+ze2e_{1}+z\cdot e_{2} for some z0z\neq 0. If x1x_{1} is orthogonal to x2x_{2} and to x3x_{3}, then x2x_{2} is proportional to ze1e2e3z\cdot e_{1}-e_{2}-e_{3} and x3x_{3} is proportional to ze1+αze3e2z\cdot e_{1}+\alpha z\cdot e_{3}-e_{2}. However, the inner product of the latter two is z2αz+1z^{2}-\alpha\cdot z+1, so it suffices to show that there exists α𝔽\alpha\in\mathbb{F} for which this quadratic polynomial has no root. Notice that in case that z2αz+1z^{2}-\alpha\cdot z+1 has a root, it can be written as (zγ)(zγ1)(z-\gamma)\cdot(z-\gamma^{-1}) for some γ0\gamma\neq 0. Since the number of possible values of α\alpha is larger than the number of possible invertible values of γ\gamma, it follows that the required α\alpha exists.

    • For =4\ell=4, suppose first that the field 𝔽\mathbb{F} is of characteristic larger than 22, and assign to the vertices along the cycle C4C_{4} the subspaces of 𝔽4\mathbb{F}^{4} defined by

      U1=U2=span(e1,e2),U3=span(e1+e4,e2+e3), and U4=span(e1+αe3,e2+e4),U_{1}=U_{2}=\mathop{\mathrm{span}}(e_{1},e_{2}),~{}~{}~{}U_{3}=\mathop{\mathrm{span}}(e_{1}+e_{4},e_{2}+e_{3}),\mbox{~{}~{}~{}and~{}~{}~{}}U_{4}=\mathop{\mathrm{span}}(e_{1}+\alpha\cdot e_{3},e_{2}+e_{4}),

      where α𝔽\alpha\in\mathbb{F} is some nonzero field element to be determined. We claim that for some α0\alpha\neq 0 it is impossible to choose four nonzero vectors xiUix_{i}\in U_{i} (i[4]i\in[4]) that form a valid choice for C4C_{4}. By α0\alpha\neq 0, it is easy to verify, as before, that x1x_{1} cannot be chosen as a scalar multiple of e1e_{1} nor of e2e_{2}, so it can be assumed that it is proportional to e1+ze2e_{1}+z\cdot e_{2} for some z0z\neq 0. If the vectors xix_{i} form a valid choice for C4C_{4}, then x2x_{2} is proportional to ze1e2z\cdot e_{1}-e_{2}, thus x3x_{3} is proportional to e1+e4+ze2+ze3e_{1}+e_{4}+z\cdot e_{2}+z\cdot e_{3}, and x4x_{4} is proportional to ze1+αze3e2e4z\cdot e_{1}+\alpha z\cdot e_{3}-e_{2}-e_{4}. However, the inner product of the latter two is αz21\alpha\cdot z^{2}-1, so it suffices to show that there exists α0\alpha\neq 0 for which αz21\alpha\cdot z^{2}\neq 1 for all values of zz. Since 𝔽\mathbb{F} is of characteristic larger than 22, it has a non-square element, whose choice for α\alpha completes the argument.

      If, however, 𝔽\mathbb{F} is of characteristic 22, one can consider the subspaces of 𝔽5\mathbb{F}^{5} defined by U1=U2=span(e1,e2)U_{1}=U_{2}=\mathop{\mathrm{span}}(e_{1},e_{2}), U3=span(e1+e4,e2+e3+e5)U_{3}=\mathop{\mathrm{span}}(e_{1}+e_{4},e_{2}+e_{3}+e_{5}), U4=span(e1+e3,e2+αe4+e5)U_{4}=\mathop{\mathrm{span}}(e_{1}+e_{3},e_{2}+\alpha\cdot e_{4}+e_{5}), where α𝔽\alpha\in\mathbb{F} is some nonzero field element for which z2+zαz^{2}+z\neq\alpha for all values of zz. Notice that such an α\alpha exists because the function zz2+zz\mapsto z^{2}+z maps both 0 and 11 to 0, so some nonzero element does not belong to its image. It can be verified that for the above subspace assignment, no valid choice of vectors for C4C_{4} exists.

    Finally, observe that for every odd >3\ell>3, one can extend the above subspace assignment for C3C_{3} by adding 3\ell-3 copies of the subspace span(e1,e2)\mathop{\mathrm{span}}(e_{1},e_{2}) between U1U_{1} and U2U_{2} to get a subspace assignment showing that ch-s(C,𝔽)>2\mathop{\mathrm{ch\text{-}s}}(C_{\ell},\mathbb{F})>2. Similarly, for every even >4\ell>4, one can extend the above subspace assignment for C4C_{4} by adding 4\ell-4 copies of the subspace span(e1,e2)\mathop{\mathrm{span}}(e_{1},e_{2}) between U1U_{1} and U2U_{2}.  

Remark 4.1.

As shown in [13], the characterization given in Proposition 1.9 for finite fields holds for the real field \mathbb{R} too. In particular, for every integer 3\ell\geq 3, it holds that ch-s(C,)>2\mathop{\mathrm{ch\text{-}s}}(C_{\ell},\mathbb{R})>2. For an odd \ell, this simply follows by assigning 2\mathbb{R}^{2} to every vertex. For an even \ell, this follows from the construction given above in the proof for fields of characteristic larger than 22, taking α\alpha to be some non-square over \mathbb{R}.

5 Subspace Choosability in Complete Bipartite Graphs

In this section we prove our results on subspace choosability in complete bipartite graphs.

5.1 Complete Balanced Bipartite Graphs

Erdös et al. [8] proved that the choice number of the complete balanced bipartite graph Km,mK_{m,m} exceeds kk for m=(2k1k)m=\binom{2k-1}{k}. We provide here a quick proof for an analogue result for subspace choosability. Note, however, that when the number of vertices is sufficiently large, the lower bound given by Theorem 1.2 is significantly better.

Proposition 5.1.

For every integer kk and for every field 𝔽\mathbb{F}, ch-s(Km,m,𝔽)>k\mathop{\mathrm{ch\text{-}s}}(K_{m,m},\mathbb{F})>k for m=(2k1k)m=\binom{2k-1}{k}.

  • Proof:

    Let kk be an integer and let 𝔽\mathbb{F} be a field. Consider the graph Km,mK_{m,m} for m=(2k1k)m=\binom{2k-1}{k}, and associate with the vertices of every side of the graph all the kk-subsets of [2k1][2k-1]. For a vertex associated with a kk-subset AA of [2k1][2k-1] we assign the kk-subspace of 𝔽2k1\mathbb{F}^{2k-1} spanned by the vectors eie_{i} with iAi\in A, where eie_{i} stands for the vector of 𝔽2k1\mathbb{F}^{2k-1} with 11 on the iith entry and 0 everywhere else. We claim that there is no choice of nonzero vectors from these subspaces such that the vectors of the left side are orthogonal to those of the right side. To see this, suppose in contradiction that such a choice exists, and denote by x1,,xmx_{1},\ldots,x_{m} and y1,,ymy_{1},\ldots,y_{m} the vectors chosen for the vertices of the left and right sides respectively. Letting U=span(x1,,xm)U=\mathop{\mathrm{span}}(x_{1},\ldots,x_{m}) and V=span(y1,,ym)V=\mathop{\mathrm{span}}(y_{1},\ldots,y_{m}), it follows that VUV\subseteq U^{\perp}, and thus

    dim(U)+dim(V)dim(U)+dim(U)=2k1,\dim(U)+\dim(V)\leq\dim(U)+\dim(U^{\perp})=2k-1,

    implying that at least one of UU and VV has dimension at most k1k-1. Without loss of generality, assume that dim(U)k1\dim(U)\leq k-1. Put =dim(U)\ell=\dim(U), fix some \ell vectors from x1,,xmx_{1},\ldots,x_{m} that span UU, and consider the (2k1)×(2k-1)\times\ell matrix whose columns are these vectors. Since the dimension of the subspace spanned by the rows of UU is also \ell, it follows that there exists a set B[2k1]B\subseteq[2k-1] of \ell indices whose rows are linearly independent. It follows that the only vector in UU with zeros in all entries of BB is the zero vector. However, by |B|=k1|B|=\ell\leq k-1, there exists a kk-subset AA of [2k1][2k-1] disjoint from BB, so the vertex associated with this AA in the left side of the graph cannot receive any nonzero vector of UU. This gives us the required contradiction and completes the proof.  

5.2 Asymmetric Subspace Choosability in Complete Bipartite Graphs

We consider now complete bipartite graphs in the asymmetric setting, where the dimensions of the subspaces assigned to the vertices of the right and left sides might be different.

Definition 5.2.

The complete bipartite graph K1,2K_{\ell_{1},\ell_{2}} with the vertex set AA of size 1\ell_{1} on the left side and the vertex set BB of size 2\ell_{2} on the right side is said to be (k1,k2)(k_{1},k_{2})-subspace choosable over a field 𝔽\mathbb{F} if it is ff-subspace choosable over 𝔽\mathbb{F} for the function f:AB{k1,k2}f:A\cup B\rightarrow\{k_{1},k_{2}\} defined by f(u)=k1f(u)=k_{1} for every uAu\in A and f(u)=k2f(u)=k_{2} for every uBu\in B.

In what follows, we provide several conditions that imply subspace choosability and subspace non-choosability in complete bipartite graphs, and in particular prove Theorems 1.41.5, and 1.7.

5.2.1 Upper Bounds

We start with the following simple statement.

Proposition 5.3.

For every field 𝔽\mathbb{F}, the graph K1,2K_{\ell_{1},\ell_{2}} is (k1,k2)(k_{1},k_{2})-subspace choosable over 𝔽\mathbb{F} whenever 1<k2\ell_{1}<k_{2} or 2<k1\ell_{2}<k_{1}.

  • Proof:

    Suppose that 1<k2\ell_{1}<k_{2}, and let U1,,U1U_{1},\ldots,U_{\ell_{1}} and V1,,V2V_{1},\ldots,V_{\ell_{2}} be k1k_{1}-subspaces and k2k_{2}-subspaces, respectively, of 𝔽t\mathbb{F}^{t} for some integer tt. Choose an arbitrary nonzero vector from each UiU_{i} for i[1]i\in[\ell_{1}]. Such a choice poses at most 1\ell_{1} linear constraints on the choice of a vector from each VjV_{j}, and since the dimension of those subspaces is k2>1k_{2}>\ell_{1}, a nonzero choice exists, resulting in a valid choice for the whole graph. By symmetry, the result holds for the case 2<k1\ell_{2}<k_{1} as well.  

We next prove the following result, which confirms Theorem 1.4.

Theorem 5.4.

For every two integers kk and nn and for every field 𝔽\mathbb{F}, the graph Kk,mK_{k,m} is (n,k)(n,k)-subspace choosable over 𝔽\mathbb{F} for m=i=0k1n1kim=\sum_{i=0}^{k-1}{\lfloor\frac{n-1}{k-i}\rfloor}.

We need the following lemma.

Lemma 5.5.

Let WW be a kk-subspace of some finite-dimensional vector space over a field 𝔽\mathbb{F}, let W1,,WtW_{1},\ldots,W_{t} be rr-subspaces of WW, and suppose that tk1krt\leq\frac{k-1}{k-r}. Then, there exists a nonzero vector in the intersection i[t]Wi\bigcap_{i\in[t]}{W_{i}}.

  • Proof:

    Using the standard equality dim(V1V2)=dim(V1)+dim(V2)dim(V1+V2)\dim(V_{1}\cap V_{2})=\dim(V_{1})+\dim(V_{2})-\dim(V_{1}+V_{2}), observe that

    dim(i[t]Wi)\displaystyle\dim\Big{(}\bigcap_{i\in[t]}{W_{i}}\Big{)} =\displaystyle= dim(W1)+dim(i2Wi)dim(W1+i2Wi)\displaystyle\dim(W_{1})+\dim\Big{(}\bigcap_{i\geq 2}{W_{i}}\Big{)}-\dim\Big{(}W_{1}+\bigcap_{i\geq 2}{W_{i}}\Big{)}
    \displaystyle\geq dim(W1)+dim(i2Wi)dim(W).\displaystyle\dim(W_{1})+\dim\Big{(}\bigcap_{i\geq 2}{W_{i}}\Big{)}-\dim(W).

    By repeatedly applying this inequality we obtain that

    dim(i[t]Wi)i[t]dim(Wi)(t1)dim(W)=tr(t1)k1,\dim\Big{(}\bigcap_{i\in[t]}{W_{i}}\Big{)}\geq\sum_{i\in[t]}{\dim(W_{i})}-(t-1)\cdot\dim(W)=t\cdot r-(t-1)\cdot k\geq 1,

    hence there exists a nonzero vector in i[t]Wi\bigcap_{i\in[t]}{W_{i}}, as required.  

We are ready to prove Theorem 5.4.

  • Proof of Theorem 5.4:

    For integers kk and nn, put m=i=0k1n1kim=\sum_{i=0}^{k-1}{\lfloor\frac{n-1}{k-i}\rfloor}. We show that the graph Kk,mK_{k,m} is (n,k)(n,k)-subspace choosable over every field 𝔽\mathbb{F}. Denote the left and right vertices of the graph by u1,,uku_{1},\ldots,u_{k} and v1,,vmv_{1},\ldots,v_{m} respectively, and consider an arbitrary assignment of nn-subspaces and kk-subspaces of 𝔽t\mathbb{F}^{t} to the left and right vertices, respectively, for some integer tt. For every i[k]i\in[k] let UiU_{i} be the subspace assigned to uiu_{i}, and for every j[m]j\in[m] let VjV_{j} be the subspace assigned to vjv_{j}. We will show that it is possible to choose nonzero vectors from these subspaces such that the vectors of the left side are orthogonal over 𝔽\mathbb{F} to those of the right side.

    We first describe how the vectors x1,,xkx_{1},\ldots,x_{k} of the left vertices u1,,uku_{1},\ldots,u_{k} are chosen. We choose them one by one, and to do so we maintain a set J[m]J\subseteq[m] and some subspaces L1,,LmL_{1},\ldots,L_{m} of 𝔽t\mathbb{F}^{t}. Initially, we define J=[m]J=[m] and Lj=VjL_{j}=V_{j}^{\perp} for all j[m]j\in[m]. Note that dim(Lj)=tk\dim(L_{j})=t-k. Then, for every i[k]i\in[k] we act as follows.

    • Pick some set JJJ^{\prime}\subseteq J of size |J|=n1k(i1)|J^{\prime}|=\lfloor\frac{n-1}{k-(i-1)}\rfloor.

    • Let J′′JJ^{\prime\prime}\subseteq J^{\prime} be the set of indices jJj\in J^{\prime} satisfying dim(Lj)=tk+(i1)\dim(L_{j})=t-k+(i-1).

    • Choose xix_{i} to be some nonzero vector of UiU_{i} that belongs to the intersection jJ′′Lj\bigcap_{j\in J^{\prime\prime}}L_{j}.

    • Add the vector xix_{i} to every subspace LjL_{j}, that is, update every subspace LjL_{j} to be the subspace Lj+span(xi)L_{j}+\mathop{\mathrm{span}}(x_{i}).

    • Remove the elements of JJ^{\prime} from JJ.

    Observe that the number of elements removed from JJ during the above kk iterations is i=1kn1k(i1)\sum_{i=1}^{k}{\lfloor\frac{n-1}{k-(i-1)}\rfloor}. Since the latter coincides with our definition of mm, it follows that after the kkth iteration the set JJ is empty.

    We show now that the vectors xix_{i} are well defined, in the sense that in the iith iteration there exists a nonzero vector that belongs to UiU_{i} and to the intersection jJ′′Lj\bigcap_{j\in J^{\prime\prime}}L_{j}. To see this, put W=UiW=U_{i} and consider its subspaces Wj=LjWW_{j}=L_{j}\cap W for jJ′′j\in J^{\prime\prime}. By the definition of J′′J^{\prime\prime}, for every jJ′′j\in J^{\prime\prime} it holds that dim(Lj)=tk+(i1)\dim(L_{j})=t-k+(i-1), hence, using dim(W)=n\dim(W)=n, it follows that

    dim(Wj)\displaystyle\dim(W_{j}) =\displaystyle= dim(Lj)+dim(W)dim(Lj+W)\displaystyle\dim(L_{j})+\dim(W)-\dim(L_{j}+W)
    \displaystyle\geq tk+(i1)+nt=nk+i1.\displaystyle t-k+(i-1)+n-t=n-k+i-1.

    By Lemma 5.5 applied to WW and to its subspaces WjW_{j}, using the fact that |J′′|n1k(i1)|J^{\prime\prime}|\leq\lfloor\frac{n-1}{k-(i-1)}\rfloor, the required vector xix_{i} is guaranteed to exist.

    We finally show that the above choice of vectors for the left vertices can be extended to a valid choice of vectors for the whole graph. Fix some j[m]j\in[m] and observe that if the subspace LjL_{j} obtained at the end of the kkth iteration has dimension strictly smaller than tt then it is possible to choose an appropriate vector yjy_{j} for the vertex vjv_{j}. Indeed, yjy_{j} can be chosen as any nonzero vector orthogonal to this LjL_{j}, because such a vector is orthogonal to VjV_{j}^{\perp}, hence belongs to VjV_{j}, and is orthogonal to all the vectors x1,,xkx_{1},\ldots,x_{k} that were chosen for the left vertices and were added to LjL_{j} during the kk iterations. Since the initial dimension of LjL_{j} is tkt-k it suffices to show that in at least one of the kk iterations, the chosen vector xix_{i} was already inside LjL_{j}. So suppose that the set JJ^{\prime} includes jj in the iith iteration. If jJ′′j\in J^{\prime\prime} then the vector xix_{i} chosen in this iteration belongs to the current LjL_{j}. Otherwise, the dimension of LjL_{j} in this iteration is smaller than tk+(i1)t-k+(i-1), implying that in one of the previous i1i-1 iterations a vector that already belongs to LjL_{j} was chosen, so we are done.  

5.2.2 Lower Bounds

We start with the following simple statement.

Proposition 5.6.

For every two integers n,k2n,k\geq 2 and for every field 𝔽\mathbb{F}, the graph Kk,nkK_{k,n^{k}} is not (n,k)(n,k)-subspace choosable over 𝔽\mathbb{F}.

  • Proof:

    Denote the vertices of the left side of Kk,nkK_{k,n^{k}} by u1,,uku_{1},\ldots,u_{k}. For every i[k]i\in[k], assign to the vertex uiu_{i} the nn-subspace of 𝔽nk\mathbb{F}^{nk} spanned by the vectors eie_{i} with i[(i1)n+1,in]i\in[(i-1)\cdot n+1,i\cdot n], where eie_{i} stands for the vector in 𝔽nk\mathbb{F}^{nk} with 11 on the iith entry and 0 everywhere else. Then, associate with each of the nkn^{k} vertices of the right side a distinct kk-tuple (a1,,ak)[n]k(a_{1},\ldots,a_{k})\in[n]^{k}, and assign to it the kk-subspace of 𝔽nk\mathbb{F}^{nk} spanned by the vectors e(i1)n+aie_{(i-1)n+a_{i}} for i[k]i\in[k].

    We claim that there is no choice of nonzero vectors from these subspaces such that the vectors of the left side are orthogonal over 𝔽\mathbb{F} to those of the right side. To see this, consider any choice of a nonzero vector xix_{i} for each vertex uiu_{i} for i[k]i\in[k]. For every i[k]i\in[k], consider the restriction x~i𝔽n\widetilde{x}_{i}\in\mathbb{F}^{n} of the vector xix_{i} to the support of its subspace, that is, to the entries with indices in [(i1)n+1,in][(i-1)\cdot n+1,i\cdot n]. Since xix_{i} is nonzero, it follows that there exists some ai[n]a_{i}\in[n] such that the vector x~i\widetilde{x}_{i} is nonzero in its aia_{i}th entry. However, the only vector in the subspace of the vertex (a1,,ak)(a_{1},\ldots,a_{k}) of the right side which is orthogonal to all the vectors xix_{i} (i[k]i\in[k]) is the zero vector. This implies that no choice of nonzero vectors for the left side can be extended to a valid choice of vectors for the whole graph, so we are done.  

We next prove the following result.

Theorem 5.7.

For every integers nn, tt, kk and for every field 𝔽\mathbb{F}, the following holds. If there exists a collection of m=k(t1)+1m=k\cdot(t-1)+1 nonzero vectors in 𝔽n\mathbb{F}^{n} satisfying that every tt of them span the entire space 𝔽n\mathbb{F}^{n}, then the graph Kk,mK_{k,m} is not (n,k)(n,k)-subspace choosable over 𝔽\mathbb{F}.

  • Proof:

    Suppose that there exists a collection of m=k(t1)+1m=k\cdot(t-1)+1 nonzero vectors b1,,bmb_{1},\ldots,b_{m} in 𝔽n\mathbb{F}^{n} satisfying that every tt of them span the space 𝔽n\mathbb{F}^{n}. To prove that Kk,mK_{k,m} is not (n,k)(n,k)-subspace choosable, we have to show that it is possible to assign nn-subspaces and kk-subspaces over 𝔽\mathbb{F} to the left and right vertices of the graph Kk,mK_{k,m} respectively, so that no choice of a nonzero vector from each subspace satisfies that the vectors of the left vertices are orthogonal over 𝔽\mathbb{F} to the vectors of the right vertices.

    Let u1,,uku_{1},\ldots,u_{k} be the vertices of the left side, and let v1,,vmv_{1},\ldots,v_{m} be the vertices of the right side. For every i[k]i\in[k], we assign to the vertex uiu_{i} the subspace UiU_{i} of 𝔽kn\mathbb{F}^{kn} that includes all the vectors whose support is contained in the entries indexed by [(i1)n+1,in][(i-1)\cdot n+1,i\cdot n]. In other words, viewing the vectors of 𝔽kn\mathbb{F}^{kn} as a concatenation of kk parts of length nn, UiU_{i} is the nn-subspace of all the vectors that have zeros in all the parts but the iith one. Then, for every j[m]j\in[m], we assign to the vertex vjv_{j} the subspace VjV_{j} spanned by the kk vectors e1bj,,ekbje_{1}\otimes b_{j},\ldots,e_{k}\otimes b_{j} of 𝔽kn\mathbb{F}^{kn}. Here, eie_{i} stands for the vector in 𝔽k\mathbb{F}^{k} with 11 on the iith entry and 0 everywhere else, and \otimes stands for the tensor product operation of vectors. Hence, VjV_{j} is the kk-subspace of all the vectors in 𝔽kn\mathbb{F}^{kn} consisting of kk parts, each of which is equal to the vector bjb_{j} multiplied by some element of 𝔽\mathbb{F}.

    Assume for the sake of contradiction that there exist nonzero vectors xiUix_{i}\in U_{i} (i[k]i\in[k]) and yjVjy_{j}\in V_{j} (j[m]j\in[m]) such that xi,yj=0\langle x_{i},y_{j}\rangle=0 for all ii and jj. For any i[k]i\in[k], let x~i𝔽n\tilde{x}_{i}\in\mathbb{F}^{n} be the (nonzero) restriction of the vector xix_{i} to the iith part. For any j[m]j\in[m], write yj=i[k]αi,jeibjy_{j}=\sum_{i\in[k]}{\alpha_{i,j}\cdot e_{i}\otimes b_{j}} for some coefficients αi,j𝔽\alpha_{i,j}\in\mathbb{F}. Since all the vectors yjy_{j} are nonzero, it clearly follows that at least mm of the coefficients αi,j\alpha_{i,j} are nonzero. Now, observe that for all i[k]i\in[k] and j[m]j\in[m], xi,yj=0\langle x_{i},y_{j}\rangle=0 implies that x~i,αi,jbj=0\langle\tilde{x}_{i},\alpha_{i,j}\cdot b_{j}\rangle=0. However, combining the facts that x~i\tilde{x}_{i} is nonzero and that every tt vectors among b1,,bmb_{1},\ldots,b_{m} span 𝔽n\mathbb{F}^{n}, it follows that for every i[k]i\in[k], at most t1t-1 of the coefficients αi,j\alpha_{i,j} with j[m]j\in[m] are nonzero. This yields that the total number of nonzero coefficients αi,j\alpha_{i,j} is at most k(t1)<mk\cdot(t-1)<m, providing the desired contradiction.  

We derive the following.

Corollary 5.8.

Let kk be an integer, and let 𝔽\mathbb{F} be a field.

  1. 1.

    For an integer nn, set m=k(n1)+1m=k\cdot(n-1)+1. If |𝔽|m|\mathbb{F}|\geq m then the graph Kk,mK_{k,m} is not (n,k)(n,k)-subspace choosable over 𝔽\mathbb{F}.

  2. 2.

    For integers nn and qq, set m=kqn11q1+1m=k\cdot\frac{q^{n-1}-1}{q-1}+1. If 𝔽\mathbb{F} is a finite field of size qkq\geq k then the graph Kk,mK_{k,m} is not (n,k)(n,k)-subspace choosable over 𝔽\mathbb{F}.

  • Proof:

    For Item 1, set m=k(n1)+1m=k\cdot(n-1)+1, and let γ1,,γm\gamma_{1},\ldots,\gamma_{m} be some distinct elements of the field 𝔽\mathbb{F}. For each i[m]i\in[m], let bib_{i} be the vector in 𝔽n\mathbb{F}^{n} defined by bi=(1,γi,γi2,,γin1)b_{i}=(1,\gamma_{i},\gamma_{i}^{2},\ldots,\gamma_{i}^{n-1}). As follows from standard properties of the Vandermonde matrix, every nn of the vectors b1,,bmb_{1},\ldots,b_{m} are linearly independent and thus span the space 𝔽n\mathbb{F}^{n}. By Theorem 5.7 applied to these vectors with t=nt=n, it follows that Kk,mK_{k,m} is not (n,k)(n,k)-subspace choosable over 𝔽\mathbb{F}, as required.

    For Item 2, set t=qn11q1+1t=\frac{q^{n-1}-1}{q-1}+1 and m=k(t1)+1m=k\cdot(t-1)+1. Consider the equivalence relation on the nonzero vectors of 𝔽n\mathbb{F}^{n} defined by calling two vectors equivalent if one is a multiple of the other by an element of 𝔽\mathbb{F}. Let BB be a collection of vectors in 𝔽n\mathbb{F}^{n} that consists of one vector from every equivalence class, and note that |B|=qn1q1|B|=\frac{q^{n}-1}{q-1}. We observe that every tt vectors of BB span the space 𝔽n\mathbb{F}^{n}. Indeed, every strict subspace of 𝔽n\mathbb{F}^{n} has dimension at most n1n-1, so it includes at most qn11q^{n-1}-1 nonzero vectors, and thus at most t1t-1 vectors that represent different equivalence classes. The assumption qkq\geq k implies that

    m=kqn11q1+1qn1q1=|B|,m=k\cdot\frac{q^{n-1}-1}{q-1}+1\leq\frac{q^{n}-1}{q-1}=|B|,

    so by applying Theorem 5.7 to mm of the vectors of BB, we get that Kk,mK_{k,m} is not (n,k)(n,k)-subspace choosable over 𝔽\mathbb{F}, and we are done.  

Theorem 1.5 and Corollary 1.6 follow, respectively, from Theorem 5.7 and Corollary 5.8.

We note that the approach proposed by Theorem 5.7 for proving subspace non-choosability results seems to be more beneficial for large fields. This is justified by the following lemma that relates the size of the collection needed in Theorem 5.7 to the size of the field. Its proof is inspired by an argument given in [5].

Lemma 5.9.

Let 𝔽\mathbb{F} be a finite field of size qq, and let mtnm\geq t\geq n be integers. If there exists a collection of mm nonzero vectors in 𝔽n\mathbb{F}^{n} satisfying that every tt of them span the space 𝔽n\mathbb{F}^{n}, then

mn2+(q+1)(tn+1).m\leq n-2+(q+1)\cdot(t-n+1).
  • Proof:

    Let S𝔽nS\subseteq\mathbb{F}^{n} be a set of mm nonzero vectors in 𝔽n\mathbb{F}^{n} satisfying that every tt of them span the space 𝔽n\mathbb{F}^{n}. Let x1,,xn2x_{1},\ldots,x_{n-2} be n2n-2 linearly independent vectors of SS, and consider all the (n1)(n-1)-subspaces of 𝔽n\mathbb{F}^{n} that include all of these vectors. Observe that the number of such subspaces is q+1q+1, and that these subspaces cover together the entire space 𝔽n\mathbb{F}^{n}. Since every tt vectors of SS span 𝔽n\mathbb{F}^{n}, it follows that each of these q+1q+1 subspaces includes less than t(n2)t-(n-2) of the vectors of S{x1,,xn2}S\setminus\{x_{1},\ldots,x_{n-2}\}. We thus conclude that m=|S|(n2)+(q+1)(t(n2)1)m=|S|\leq(n-2)+(q+1)\cdot(t-(n-2)-1), and we are done.  

5.2.3 Two Vertices on the Left Side

We next consider the particular case of the complete bipartite graph K2,mK_{2,m} with two vertices on the left side and mm vertices on the right side. It will be convenient to use the following definition.

Definition 5.10.

The complete bipartite graph K2,mK_{2,m} with the vertex set A={u1,u2}A=\{u_{1},u_{2}\} on the left side and the vertex set BB of size mm on the right side is said to be (k1;k2,k3)(k_{1};k_{2},k_{3})-subspace choosable over a field 𝔽\mathbb{F} if it is ff-subspace choosable over 𝔽\mathbb{F} for the function f:AB{k1,k2,k3}f:A\cup B\rightarrow\{k_{1},k_{2},k_{3}\} defined by f(u1)=k1f(u_{1})=k_{1}, f(u2)=k2f(u_{2})=k_{2} and f(u)=k3f(u)=k_{3} for every uBu\in B.

In what follows, we prove Theorem 1.7. We start with its first item, restated and proved below.

Proposition 5.11.

For every integer nn and for every field 𝔽\mathbb{F}, the graph K2,n1K_{2,n-1} is (n;2,2)(n;2,2)-subspace choosable over 𝔽\mathbb{F}.

  • Proof:

    For an integer nn, consider the graph K2,n1K_{2,n-1}. To prove that it is (n;2,2)(n;2,2)-subspace choosable over a field 𝔽\mathbb{F}, consider for some integer tt arbitrary subspaces U1,U2U_{1},U_{2} and V1,,Vn1V_{1},\ldots,V_{n-1} of 𝔽t\mathbb{F}^{t} whose dimensions satisfy dim(U1)=n\dim(U_{1})=n, dim(U2)=2\dim(U_{2})=2, and dim(Vj)=2\dim(V_{j})=2 for j[n1]j\in[n-1]. Choose an arbitrary nonzero vector x2U2x_{2}\in U_{2}, and for every j[n1]j\in[n-1] choose a nonzero vector yjVjy_{j}\in V_{j} such that x2,yj=0\langle x_{2},y_{j}\rangle=0. Note that this is possible since dim(Vj)=2\dim(V_{j})=2. Finally, choose a vector x1U1x_{1}\in U_{1} satisfying x1,yj=0\langle x_{1},y_{j}\rangle=0 for all j[n1]j\in[n-1], which is possible by dim(U1)=n\dim(U_{1})=n. This gives us the required choice of vectors.  

By the above proposition, K2,n1K_{2,n-1} is (n;2,2)(n;2,2)-subspace choosable over every field. We consider the question of whether this holds even after adding another vertex to the right side of the graph. Under certain conditions the answer is positive, as shown by the following result, confirming Item 2 and the “if” part of Item 3 in Theorem 1.7.

Proposition 5.12.

The graph K2,nK_{2,n} is (n;2,2)(n;2,2)-subspace choosable for every integer nn over \mathbb{C} and for every odd integer nn over \mathbb{R}.

We need the following lemma, which is essentially given in [13].

Lemma 5.13 ([13, Lemma 2.9]).

Let t2t\geq 2 be an integer, and let 𝔽\mathbb{F} be either \mathbb{R} or \mathbb{C}. Let U,VU,V be two 22-subspaces of 𝔽t\mathbb{F}^{t} such that for every nonzero vector xUx\in U there exists a nonzero vector yVy\in V such that x,y0\langle x,y\rangle\neq 0. Then, for every basis u(1),u(2)u^{(1)},u^{(2)} of UU satisfying u(i),u(j)0\langle u^{(i)},u^{(j)}\rangle\neq 0 if and only if i=ji=j, there exists a basis v(1),v(2)v^{(1)},v^{(2)} of VV satisfying v(i),v(j)0\langle v^{(i)},v^{(j)}\rangle\neq 0 if and only if i=ji=j and, in addition, u(i),v(j)=0\langle u^{(i)},v^{(j)}\rangle=0 if and only if i=ji=j.

  • Proof of Proposition 5.12:

    Let nn be an integer, and let 𝔽\mathbb{F} be either \mathbb{R} or \mathbb{C}. Consider the graph K2,nK_{2,n} with the vertex set A={u1,u2}A=\{u_{1},u_{2}\} on the left side and the vertex set B={v1,,vn}B=\{v_{1},\ldots,v_{n}\} on the right side. To prove that the graph is (n;2,2)(n;2,2)-subspace choosable over 𝔽\mathbb{F}, consider some subspaces U1,U2,V1,,VnU_{1},U_{2},V_{1},\ldots,V_{n} of 𝔽t\mathbb{F}^{t} for some integer tt, where dim(U1)=2\dim(U_{1})=2, dim(U2)=n\dim(U_{2})=n, and dim(Vj)=2\dim(V_{j})=2 for all j[n]j\in[n]. We will show now that there exist nonzero vectors xiUix_{i}\in U_{i} (i[2]i\in[2]) and yjVjy_{j}\in V_{j} (j[n]j\in[n]) such that xi,yj=0\langle x_{i},y_{j}\rangle=0 over 𝔽\mathbb{F} for all ii and jj.

    Suppose first that there exists a nonzero vector x1U1x_{1}\in U_{1} such that x1x_{1} is orthogonal to the subspace VjV_{j^{\prime}} for some j[n]j^{\prime}\in[n]. In this case, choose x1x_{1} for the vertex u1u_{1}, and for every j[n]{j}j\in[n]\setminus\{j^{\prime}\} let yjVjy_{j}\in V_{j} be a nonzero choice for the vertex vjv_{j} satisfying x1,yj=0\langle x_{1},y_{j}\rangle=0. Note that such a choice exists because dim(Vj)=2\dim(V_{j})=2. These choices pose at most n1n-1 linear constraints on the choice for u2u_{2}, so by dim(U2)=n\dim(U_{2})=n, there exists a nonzero vector x2U2x_{2}\in U_{2} that is orthogonal to all the vectors yjy_{j} with j[n]{j}j\in[n]\setminus\{j^{\prime}\}. Finally, choose yjVjy_{j^{\prime}}\in V_{j^{\prime}} as a nonzero vector orthogonal to x2x_{2}, whose existence is guaranteed by dim(Vj)=2\dim(V_{j^{\prime}})=2. The assumption on x1x_{1} implies that x1,yj=0\langle x_{1},y_{j^{\prime}}\rangle=0, so we obtain the required choice of vectors.

    Otherwise, let u1(1),u1(2)u_{1}^{(1)},u_{1}^{(2)} be a basis of U1U_{1} satisfying u1(i),u1(j)0\langle u_{1}^{(i)},u_{1}^{(j)}\rangle\neq 0 if and only if i=ji=j. Since no nonzero vector of U1U_{1} is orthogonal to some VjV_{j}, we can apply Lemma 5.13 to obtain for every j[n]j\in[n] a basis vj(1),vj(2)v_{j}^{(1)},v_{j}^{(2)} of VjV_{j} that satisfies the assertion of the lemma. Note that it can be assumed that u1(1),vj(2)=u1(2),vj(1)=1\langle u_{1}^{(1)},v_{j}^{(2)}\rangle=\langle u_{1}^{(2)},v_{j}^{(1)}\rangle=1 for all j[n]j\in[n]. Let M1M_{1} and M2M_{2} be the n×tn\times t matrices over 𝔽\mathbb{F} whose jjth rows are vj(1)v_{j}^{(1)} and vj(2)v_{j}^{(2)} respectively.

    Now, to obtain the required choice of nonzero vectors, let x1=αu1(1)+βu1(2)x_{1}=\alpha\cdot u_{1}^{(1)}+\beta\cdot u_{1}^{(2)} be our nonzero choice for the vertex u1u_{1} for some α,β𝔽\alpha,\beta\in\mathbb{F} to be determined. Observe that this choice forces us to choose, up to a multiplicative constant, the vector yj=αvj(1)βvj(2)y_{j}=\alpha\cdot v_{j}^{(1)}-\beta\cdot v_{j}^{(2)} for the vertex vjv_{j} for each j[n]j\in[n]. For the vertex u2u_{2}, let U𝔽t×nU\in\mathbb{F}^{t\times n} denote a matrix whose columns form a basis of the subspace U2U_{2}, and denote its choice by x2=Uγx_{2}=U\cdot\gamma for γ𝔽n\gamma\in\mathbb{F}^{n}. We consider the question of whether there exist α,β\alpha,\beta as above and a nonzero γ\gamma such that x2,yj=0\langle x_{2},y_{j}\rangle=0 for all j[n]j\in[n]. Observe that this condition is equivalent to

    (αM1βM2)(Uγ)=0.(\alpha\cdot M_{1}-\beta\cdot M_{2})\cdot(U\cdot\gamma)=0.

    Letting M1M^{\prime}_{1} and M2M^{\prime}_{2} be the n×nn\times n matrices defined by M1=M1UM^{\prime}_{1}=M_{1}\cdot U and M2=M2UM^{\prime}_{2}=M_{2}\cdot U, we ask whether there exist α,β𝔽\alpha,\beta\in\mathbb{F}, that are not both zeros, and a nonzero vector γ𝔽n\gamma\in\mathbb{F}^{n} satisfying

    (αM1βM2)γ=0.(\alpha\cdot M^{\prime}_{1}-\beta\cdot M^{\prime}_{2})\cdot\gamma=0.

    If det(M1)=0\det(M^{\prime}_{1})=0 then we can take α=1\alpha=1 and β=0\beta=0, for which a nonzero γ\gamma is guaranteed to exist. Otherwise, if det(M1)0\det(M^{\prime}_{1})\neq 0, we take, say, β=1\beta=-1, and show that for some α𝔽\alpha\in\mathbb{F}, the matrix αM1+M2\alpha\cdot M^{\prime}_{1}+M^{\prime}_{2} is singular, implying the existence of the required vector γ\gamma. To see this, observe that αM1+M2\alpha\cdot M^{\prime}_{1}+M^{\prime}_{2} is singular if and only if αIn+N\alpha\cdot I_{n}+N is singular as well, where N=M2(M1)1N=M^{\prime}_{2}\cdot(M^{\prime}_{1})^{-1}. This reduces our question to whether for some α𝔽\alpha\in\mathbb{F} it holds that det(αIn+N)=0\det(\alpha\cdot I_{n}+N)=0 . This determinant is a degree nn polynomial in α\alpha. Over 𝔽=\mathbb{F}=\mathbb{C}, this polynomial clearly has a root, and over 𝔽=\mathbb{F}=\mathbb{R}, assuming that nn is odd, it has a root as well. This completes the proof.  

We end this section by proving that adding a vertex to the right side of K2,n1K_{2,n-1} for an even integer nn results in a graph which is no longer (n;2,2)(n;2,2)-subspace choosable over the real field \mathbb{R} and over every finite field. This, in particular, gives us the “only if” part of Item 3 of Theorem 1.7.

Proposition 5.14.

Let 𝔽\mathbb{F} be either \mathbb{R} or any finite field. Then, for every even integer nn, the graph K2,nK_{2,n} is not (n;2,2)(n;2,2)-subspace choosable over 𝔽\mathbb{F}.

  • Proof:

    For a field 𝔽\mathbb{F} as above, Proposition 1.9 and Remark 4.1 imply that ch-s(K2,2,𝔽)>2\mathop{\mathrm{ch\text{-}s}}(K_{2,2},\mathbb{F})>2. Hence, for some integer tt, there exist 22-subspaces L1,L2,R1,R2𝔽tL_{1},L_{2},R_{1},R_{2}\subseteq\mathbb{F}^{t} such that no choice of nonzero vectors x~iLi\widetilde{x}_{i}\in L_{i} and y~jRj\widetilde{y}_{j}\in R_{j} for i,j[2]i,j\in[2] satisfies x~i,y~j=0\langle\widetilde{x}_{i},\widetilde{y}_{j}\rangle=0 for all i,ji,j.

    For an even integer n=2kn=2k, we define a subspace assignment to the vertices of K2,nK_{2,n} that lies in the vector space 𝔽tk\mathbb{F}^{t\cdot k} as follows. To the left vertices we assign the subspaces U1,U2𝔽tkU_{1},U_{2}\subseteq\mathbb{F}^{t\cdot k} defined by

    U1=span(e1L1,,ekL1)andU2=(i=1kei)L2,U_{1}=\mathop{\mathrm{span}}(e_{1}\otimes L_{1},\ldots,e_{k}\otimes L_{1})~{}~{}\mbox{and}~{}~{}U_{2}=\Big{(}\sum_{i=1}^{k}{e_{i}}\Big{)}\otimes L_{2},

    and to the right vertices we assign the subspaces V1,,Vn𝔽tkV_{1},\ldots,V_{n}\subseteq\mathbb{F}^{t\cdot k}, defined by

    V2j1=ejR1andV2j=ejR2V_{2j-1}=e_{j}\otimes R_{1}~{}~{}\mbox{and}~{}~{}V_{2j}=e_{j}\otimes R_{2}

    for each j[k]j\in[k]. Note that dim(U1)=n\dim(U_{1})=n, dim(U2)=2\dim(U_{2})=2, and dim(Vj)=2\dim(V_{j})=2 for all j[n]j\in[n]. Intuitively, the assignment is designed so that the tt-dimensional restriction of U1,U2,V2j1,V2jU_{1},U_{2},V_{2j-1},V_{2j} to the jjth block is the assignment L1,L2,R1,R2L_{1},L_{2},R_{1},R_{2}.

    To complete the proof, we show that there is no choice of nonzero vectors xiUix_{i}\in U_{i} and yjVjy_{j}\in V_{j} for i[2]i\in[2] and j[n]j\in[n] that satisfies xi,yj=0\langle x_{i},y_{j}\rangle=0 for all i,ji,j. So suppose for contradiction that such a choice exists, and let j[k]j\in[k] be an integer for which the restriction of x1x_{1} to the jjth block is nonzero. Denote by x~1,x~2,y~1,y~2\widetilde{x}_{1},\widetilde{x}_{2},\widetilde{y}_{1},\widetilde{y}_{2} the restrictions of the vectors x1,x2,y2j1,y2jx_{1},x_{2},y_{2j-1},y_{2j} to the jjth block. Observe that these are nonzero vectors that satisfy x~iLi\widetilde{x}_{i}\in L_{i}, y~jRj\widetilde{y}_{j}\in R_{j}, and x~i,y~j=0\langle\widetilde{x}_{i},\widetilde{y}_{j}\rangle=0 for all i,j[2]i,j\in[2], in contradiction.  

6 Hardness Result

In this section we prove our hardness result, given in Theorem 1.8. We start by presenting a gadget graph that will be used in the proof.

6.1 Gadget Graph

The main component of our hardness proof is the \exists-graph defined as follows.

Definition 6.1 (\exists-graph).

For any integers n1,n2n_{1},n_{2}, define the \exists-graph H=Hn1,n2H=H_{n_{1},n_{2}} and the function fH:V(H){2,3}f_{H}:V(H)\to\{2,3\} as follows. The graph consists of a vertex labelled 𝖨𝖭\mathsf{IN} with degree 2, whose two neighbors serve as the starting points of two subgraphs to which we will refer as the top and bottom branches. Each branch is composed of a sequence of 44-cycles connected by edges, as described in the figure below. In each branch, the vertex of largest distance from 𝖨𝖭\mathsf{IN} in every 44-cycle but the first has a neighbor labelled 𝖮𝖴𝖳\mathsf{OUT} and another neighbor separating it from the next 44-cycle (except for the last 44-cycle). The numbers of 𝖮𝖴𝖳\mathsf{OUT} vertices in the top and bottom branches are n1n_{1} and n2n_{2} respectively. The function fHf_{H} is defined on the vertices of HH as indicated in the figure.

2𝖨𝖭\mathsf{IN}322332232𝖮𝖴𝖳\mathsf{OUT}232232𝖮𝖴𝖳\mathsf{OUT}2\ldots32232𝖮𝖴𝖳\mathsf{OUT}322332232𝖮𝖴𝖳\mathsf{OUT}232232𝖮𝖴𝖳\mathsf{OUT}2\ldots32232𝖮𝖴𝖳\mathsf{OUT}

We need the following two claims.

Claim 6.2.

Let 𝔽\mathbb{F} be any field. Let AA denote a neighbor of 𝖨𝖭\mathsf{IN} in the \exists-graph, and let BB denote another vertex adjacent to AA. Then, for every fHf_{H}-subspace assignment for HH over 𝔽\mathbb{F}, there exists a choice of nonzero vectors for 𝖨𝖭\mathsf{IN} and BB which poses a single linear constraint on the choice for AA.

  • Proof:

    Let W𝖨𝖭,WA,WBW_{\mathsf{IN}},W_{A},W_{B} denote the subspaces assigned to the vertices 𝖨𝖭,A,B\mathsf{IN},A,B respectively, and recall that dim(W𝖨𝖭)=2\dim(W_{\mathsf{IN}})=2, dim(WA)=3\dim(W_{A})=3, and dim(WB)=2\dim(W_{B})=2. If there exists some nonzero vector in W𝖨𝖭WBW_{\mathsf{IN}}\cap W_{B}, then choosing it for both 𝖨𝖭\mathsf{IN} and BB completes the proof. Otherwise, it must hold that dim(W𝖨𝖭+WB)=4>dim(WA)\dim(W_{\mathsf{IN}}+W_{B})=4>\dim(W_{A}), hence there exists some nonzero vector x(W𝖨𝖭+WB)WAx\in(W_{\mathsf{IN}}+W_{B})\cap W_{A}^{\perp}. Write x=x1+x2x=x_{1}+x_{2} for x1W𝖨𝖭x_{1}\in W_{\mathsf{IN}} and x2WBx_{2}\in W_{B}. If both of x1x_{1} and x2x_{2} are nonzero, choose them for 𝖨𝖭\mathsf{IN} and BB. Since every vector yWAy\in W_{A} satisfies y,x=0\langle y,x\rangle=0, it follows that if y,x1=0\langle y,x_{1}\rangle=0 then y,x2=0\langle y,x_{2}\rangle=0. This implies that the only linear constraint that this choice poses on the vector of AA is the orthogonality to x1x_{1}. If, however, x1x_{1} is zero, then we have that x2WAx_{2}\in W_{A}^{\perp}, so one can choose an arbitrary nonzero vector from W𝖨𝖭W_{\mathsf{IN}} for 𝖨𝖭\mathsf{IN} and x2x_{2} for BB. Similarly, if x2x_{2} is zero, we have that x1WAx_{1}\in W_{A}^{\perp}, so one can choose x1x_{1} for 𝖨𝖭\mathsf{IN} and an arbitrary nonzero vector from WBW_{B} for BB, completing the proof.  

Claim 6.3.

Let 𝔽\mathbb{F} be either \mathbb{R} or any finite field, and let xx be either e6e_{6} or e7e_{7} in 𝔽7\mathbb{F}^{7}. Then, there exists a subspace assignment W1,,W4𝔽7W_{1},\ldots,W_{4}\subseteq\mathbb{F}^{7} to the vertices u1,,u4u_{1},\ldots,u_{4} of C4C_{4}, with dim(W1)=3\dim(W_{1})=3 and dim(Wi)=2\dim(W_{i})=2 for i{2,3,4}i\in\{2,3,4\}, for which any valid choice of vectors assigns to u1u_{1} a vector proportional to xx.

  • Proof:

    The proof of Proposition 1.9 (see also Remark 4.1) describes for every field 𝔽\mathbb{F} as above, a 22-subspace assignment for C4C_{4} in 𝔽5\mathbb{F}^{5} that admits no valid choice of vectors. Let W1,,W4𝔽7W_{1},\ldots,W_{4}\subseteq{\mathbb{F}^{7}} be the subspaces obtained from the subspaces of this assignment by adding two additional entries with values zero to their vectors. Define W1=W1+span(x)W^{\prime}_{1}=W_{1}+\mathop{\mathrm{span}}(x), and observe that any valid choice of vectors from the subspace assignment W1,W2,W3,W4W^{\prime}_{1},W_{2},W_{3},W_{4} assigns to u1u_{1} a vector proportional to xx, as otherwise, the restriction of such a choice to the first five entries would provide a valid choice for the given 22-subspace assignment for C4C_{4}.  

The following lemma summarizes some properties of the \exists-graph.

Lemma 6.4.

The \exists-graph HH and the function fHf_{H} given in Definition 6.1 satisfy the following.

  1. 1.

    The graph HH is bipartite, and every bipartition of HH puts all 𝖮𝖴𝖳\mathsf{OUT} vertices in the same part.

  2. 2.

    For every fHf_{H}-subspace assignment for HH over any field 𝔽\mathbb{F}, any choice of a nonzero vector for 𝖨𝖭\mathsf{IN} can be extended to all vertices of each of the branches.

  3. 3.

    For every fHf_{H}-subspace assignment for HH over any field 𝔽\mathbb{F} and for each of the branches of HH, there exists a choice of a nonzero vector for 𝖨𝖭\mathsf{IN} which is compatible with any choice of vectors for the 𝖮𝖴𝖳\mathsf{OUT} vertices of that branch.

  4. 4.

    Let 𝔽\mathbb{F} be either \mathbb{R} or any finite field, and let t8t\geq 8 and j[t]j\in[t] be some integers. Then, there exists an fHf_{H}-subspace assignment for HH in 𝔽t\mathbb{F}^{t} such that for every valid choice of vectors for HH there exists a branch all of whose 𝖮𝖴𝖳\mathsf{OUT} vertices are assigned vectors proportional to eje_{j}.

  • Proof:

    For Item 1, it can be easily seen that the graph defined in Definition 6.1 is bipartite. Since the distance between every two 𝖮𝖴𝖳\mathsf{OUT} vertices is even, it follows that every bipartition puts all of them in the same part.

    For Item 2, consider some fHf_{H}-subspace assignment for HH over a field 𝔽\mathbb{F}, and notice that any choice of a vector for 𝖨𝖭\mathsf{IN} reduces the dimension of the subspaces available to its neighbors by at most 11. So given any choice for 𝖨𝖭\mathsf{IN}, one can choose, in each branch, an available nonzero vector for 𝖨𝖭\mathsf{IN}’s neighbor, reducing the dimension of the subspaces available to its other neighbors to not less than 11, allowing us to choose for them nonzero vectors as well. Their common neighbor has a subspace of dimension 33, so the two chosen vectors of its neighbors reduce the dimension of the subspace available to it to not less than 11, again allowing us to choose a nonzero vector. Proceeding this way for vertices with increasing distances from 𝖨𝖭\mathsf{IN} allows us to choose vectors for all vertices of each of the branches of HH.

    For Item 3, consider some fHf_{H}-subspace assignment for HH over a field 𝔽\mathbb{F} and an arbitrary branch of HH. Let AA denote the neighbor of 𝖨𝖭\mathsf{IN} in this branch, let BB and CC denote the other neighbors of AA, and let DD denote the remaining vertex of their 44-cycle. By Claim 6.2, there exists a choice of nonzero vectors for 𝖨𝖭\mathsf{IN} and BB which poses a single linear constraint on the choice for AA. We claim that this choice for 𝖨𝖭\mathsf{IN} and BB is compatible with any choice of vectors for the 𝖮𝖴𝖳\mathsf{OUT} vertices of that branch. To see this, consider an arbitrary choice of nonzero vectors for these 𝖮𝖴𝖳\mathsf{OUT} vertices. The single neighbor of each 𝖮𝖴𝖳\mathsf{OUT} vertex is assigned a 33-subspace, so having made our choice for the 𝖮𝖴𝖳\mathsf{OUT} vertices, each of these must still have a 22-subspace from which its vector can be chosen. Starting from the neighbor of the 𝖮𝖴𝖳\mathsf{OUT} vertex of largest distance from 𝖨𝖭\mathsf{IN}, we choose an arbitrary nonzero vector from its available 22-subspace, allowing us to choose a nonzero vector for each of its two neighbors. Their other common neighbor has a 33-subspace, so it includes a nonzero vector orthogonal to the vectors chosen for its neighbors. We proceed this way along the branch until we arrive to the 44-cycle closest to 𝖨𝖭\mathsf{IN}. Given the vectors chosen for the previous 44-cycle and the choice for BB, it is possible to choose from the 33-subspace of DD some nonzero vector orthogonal to the vectors already chosen for its neighbors. Given this choice, we choose a nonzero vector orthogonal to it from the 22-subspace of CC, and since the choice for 𝖨𝖭\mathsf{IN} and BB poses a single linear constraint on AA, it is possible to choose a nonzero vector for AA from its 33-subspace. By Item 2 of the lemma, our choice can be extended to the other branch, and we are done.

    For Item 4, let 𝔽\mathbb{F} be either \mathbb{R} or any finite field, and let t8t\geq 8 and j[t]j\in[t] be some integers. Assume without loss of generality that j8j\geq 8. We define an fHf_{H}-subspace assignment for HH in 𝔽t\mathbb{F}^{t} as follows. The vertex 𝖨𝖭\mathsf{IN} is assigned the subspace span(e6,e7)\mathop{\mathrm{span}}(e_{6},e_{7}). By Claim 6.3, for xx being either e6e_{6} or e7e_{7} in 𝔽7\mathbb{F}^{7}, there exists a subspace assignment W1,,W4𝔽7W_{1},\ldots,W_{4}\subseteq\mathbb{F}^{7} to the vertices u1,,u4u_{1},\ldots,u_{4} of C4C_{4}, with dim(W1)=3\dim(W_{1})=3 and dim(Wi)=2\dim(W_{i})=2 for i{2,3,4}i\in\{2,3,4\}, for which any valid choice of vectors assigns to u1u_{1} a vector proportional to xx. By extending these subspaces to 𝔽t\mathbb{F}^{t} with zeros in the last t7t-7 entries, one can get such a subspace assignment in 𝔽t\mathbb{F}^{t}. We put this subspace assignment with x=e6x=e_{6} on the 44-cycle closest to 𝖨𝖭\mathsf{IN} in each branch, where the 33-subspace is assigned to the vertex with largest distance from 𝖨𝖭\mathsf{IN}. To the subspace of the top neighbor of 𝖨𝖭\mathsf{IN}, we add the vector e6e_{6}, and to the one of the bottom, we add the vector e7e_{7}. For all remaining 44-cycles in the graph, we assign the subspaces of 𝔽t\mathbb{F}^{t} given by Claim 6.3 with x=e7x=e_{7}, again with the 33-subspace assigned to the vertex of largest distance from 𝖨𝖭\mathsf{IN}, and add the vector e6e_{6} to the subspace of the vertex closest to 𝖨𝖭\mathsf{IN}. Finally, to all 𝖮𝖴𝖳\mathsf{OUT} vertices we assign the subspace span(e7,ej)\mathop{\mathrm{span}}(e_{7},e_{j}), and to the remaining vertices separating the 4-cycles, we assign the subspace span(e6,e7)\mathop{\mathrm{span}}(e_{6},e_{7}).

    We claim that this fHf_{H}-subspace assignment for HH satisfies that for every valid choice of vectors there exists a branch all of whose 𝖮𝖴𝖳\mathsf{OUT} vertices are assigned vectors proportional to eje_{j}. To see this, consider such a valid choice of vectors, and recall that it assigns to 𝖨𝖭\mathsf{IN} a nonzero vector from span(e6,e7)\mathop{\mathrm{span}}(e_{6},e_{7}). In such a vector, at least one of the sixth and seventh entries is nonzero. We show that in the former case all the vectors of the 𝖮𝖴𝖳\mathsf{OUT} vertices of the top branch are proportional to eje_{j}. A similar argument shows that in the latter case, the same holds for the bottom branch. Our assumption on the vector of 𝖨𝖭\mathsf{IN} implies that its neighbor in the top branch is orthogonal to e6e_{6}. This essentially restricts its 44-cycle to the subspace assignment given by Claim 6.3, thus ensuring that the vertex of largest distance from 𝖨𝖭\mathsf{IN} in this 44-cycle is assigned a vector proportional to e6e_{6}. Applying this argument again to the next 44-cycle yields that its vertex of largest distance from 𝖨𝖭\mathsf{IN} is assigned a vector proportional to e7e_{7}. This ensures that the vector of its 𝖮𝖴𝖳\mathsf{OUT} neighbor is proportional to eje_{j} and that the vector of its neighbor that separates its cycle from the next one is proportional to e6e_{6}. By repeating this argument for all the following 44-cycles, the proof is completed.  

6.2 Proof of Theorem 1.8

To prove Theorem 1.8, we first prove the following.

Theorem 6.5.

Let 𝔽\mathbb{F} be either \mathbb{R} or any finite field. It is 𝖭𝖯\mathsf{NP}-hard to decide given a bipartite graph G=(V,E)G=(V,E) and a function f:V{2,3}f:V\rightarrow\{2,3\} whether GG is ff-subspace choosable over 𝔽\mathbb{F}.

  • Proof:

    Let 𝔽\mathbb{F} be a field as in the statement of the theorem. Given a 33SAT formula ϕ\phi with clauses C1,,CmC_{1},\ldots,C_{m} over the variables x1,,xnx_{1},\ldots,x_{n}, we efficiently construct a graph Gϕ=(V,E)G_{\phi}=(V,E) and a function f:V{2,3}f:V\to\{2,3\} such that ϕ\phi is satisfiable if and only if GϕG_{\phi} is ff-subspace choosable over 𝔽\mathbb{F}. Note that it can be assumed that each clause of ϕ\phi contains three literals involving three distinct variables.

    First, for each variable xjx_{j}, construct an \exists-graph Hn1,n2H_{n_{1},n_{2}} (see Definition 6.1), where n1n_{1} and n2n_{2} are, respectively, the numbers of occurrences of the literals xjx_{j} and xj¯\overline{x_{j}} in ϕ\phi. Label the 𝖮𝖴𝖳\mathsf{OUT} vertices of the top branch of Hn1,n2H_{n_{1},n_{2}} by xjx_{j}, and the 𝖮𝖴𝖳\mathsf{OUT} vertices of its bottom branch by xj¯\overline{x_{j}}. Define the function ff on the vertices of this graph as in Definition 6.1. Next, for each clause CiC_{i} of ϕ\phi, add a vertex representing CiC_{i} and define its ff value to be 33. For each literal xjx_{j} occurring in a clause CiC_{i}, add an edge between the vertex representing CiC_{i} and a previously unchosen vertex labelled xjx_{j}, and likewise for the literals of the form xj¯\overline{x_{j}}. Observe that GϕG_{\phi} is bipartite, as Item 1 of Lemma 6.4 implies that there exists a bipartition placing all 𝖮𝖴𝖳\mathsf{OUT} vertices of all \exists-graphs in the same part, thus the clause vertices may all belong to the opposite part. Note that GϕG_{\phi} can be constructed in polynomial running time.

    We prove now the correctness of the reduction. Suppose first that there exists a satisfying assignment for ϕ\phi, and consider an arbitrary ff-subspace assignment for GϕG_{\phi} over 𝔽\mathbb{F}. Then, for each variable xjx_{j} with value 𝖳𝗋𝗎𝖾\mathsf{True}, choose for the 𝖨𝖭\mathsf{IN} vertex of its \exists-graph a vector, promised by Item 3 of Lemma 6.4, which is compatible with any choice of vectors for the 𝖮𝖴𝖳\mathsf{OUT} vertices labelled xjx_{j}. If, however, xjx_{j} has value 𝖥𝖺𝗅𝗌𝖾\mathsf{False}, choose instead a vector for 𝖨𝖭\mathsf{IN} which is compatible with any choice of vectors for the 𝖮𝖴𝖳\mathsf{OUT} vertices labelled xj¯\overline{x_{j}}. By Item 2 of the lemma, such a choice can be extended to all the vertices in the opposite branch. Now, since every clause has at most two literals which evaluate to 𝖥𝖺𝗅𝗌𝖾\mathsf{False} under the given satisfying assignment, we find that, so far, vectors have been chosen for at most two of the neighbors of each clause vertex. Since each clause vertex has a subspace of dimension 33, we can make a choice for it which is compatible with all of its neighbors whose vectors have already been chosen. Observe that this choice can be extended to all the 𝖮𝖴𝖳\mathsf{OUT} vertices for which no vectors have been chosen so far, because their subspaces have dimension 22 whereas a vector has been chosen only for one of their neighbors. Finally, by our choice of the vectors of the 𝖨𝖭\mathsf{IN} vertices, using Item 3 of Lemma 6.4, one can properly choose vectors for the rest of the graph. This implies that GϕG_{\phi} is ff-subspace choosable over 𝔽\mathbb{F}.

    For the other direction, suppose that GϕG_{\phi} is ff-subspace choosable over 𝔽\mathbb{F}. Put t=n+7t=n+7, and apply Item 4 of Lemma 6.4 to obtain an fHf_{H}-subspace assignment in 𝔽t\mathbb{F}^{t} for each \exists-gadget, such that, for each j[n]j\in[n], every valid choice of vectors assigns vectors proportional to eje_{j} either to all vertices labelled xjx_{j} or to all vertices labelled xj¯\overline{x_{j}}. Finally, to the vertex of a clause CiC_{i} that involves the three variables xj1,xj2,xj3x_{j_{1}},x_{j_{2}},x_{j_{3}}, assign the subspace spanned by ej1,ej2,ej3e_{j_{1}},e_{j_{2}},e_{j_{3}}. Since GϕG_{\phi} is ff-subspace choosable over 𝔽\mathbb{F}, there exists a valid choice for GϕG_{\phi} from these subspaces. By our definition of the subspace assignment, for every j[n]j\in[n], this choice assigns vectors proportional to eje_{j} to all vertices labelled xjx_{j} or to all vertices labelled xj¯\overline{x_{j}}. In the former case assign xjx_{j} to 𝖥𝖺𝗅𝗌𝖾\mathsf{False}, and in the latter to 𝖳𝗋𝗎𝖾\mathsf{True}. We claim that this assignment satisfies ϕ\phi. To see this, observe that each vertex representing a clause CiC_{i} must have for some j[n]j\in[n] a neighbor labelled xjx_{j} or xj¯\overline{x_{j}} whose chosen vector is not proportional to eje_{j}. This neighbor corresponds to a literal whose value is 𝖳𝗋𝗎𝖾\mathsf{True} according to our assignment, as desired.  

We also need the following simple lemma, whose proof employs ideas from [11].

Lemma 6.6.

For every field 𝔽\mathbb{F} and for every integer k3k\geq 3, the following holds. There exists a polynomial-time reduction from the problem of deciding for a given input of a bipartite graph G=(V,E)G=(V,E) and a function f:V{2,3}f:V\to\{2,3\} whether GG is ff-subspace choosable over 𝔽\mathbb{F}, to the problem of deciding whether a given bipartite graph is kk-subspace choosable over 𝔽\mathbb{F}.

  • Proof:

    We start by proving the statement of the lemma for k=3k=3. Given a bipartite graph G=(V,E)G=(V,E) with bipartition V=V1V2V=V_{1}\cup V_{2} and given a function f:V{2,3}f:V\rightarrow\{2,3\}, consider the graph GG^{\prime} that consists of nine copies of GG, labelled Gi,jG_{i,j} for i,j[3]i,j\in[3], and two additional vertices v1,v2v_{1},v_{2} such that, for each {1,2}\ell\in\{1,2\}, the vertex vv_{\ell} is adjacent to all vertices uu with f(u)=2f(u)=2 in the copies of VV_{\ell}. It is easy to see that GG^{\prime} is bipartite and that it can be constructed in polynomial running time.

    For correctness, suppose first that GG is ff-subspace choosable over 𝔽\mathbb{F}, and consider an arbitrary assignment of 33-subspaces over 𝔽\mathbb{F} to the vertices of GG^{\prime}. Any choice of nonzero vectors for v1v_{1} and v2v_{2} will reduce the dimensions of the subspaces of the vertices of the graphs Gi,jG_{i,j} to not less than their original values under ff. Since each Gi,jG_{i,j} is ff-subspace choosable over 𝔽\mathbb{F}, it follows that there exists a valid choice of vectors for the vertices of GG^{\prime}, as required. For the other direction, suppose that for some integer tt, there exists an ff-subspace assignment for GG such that no choice of nonzero vectors from the subspaces is valid. To the vertices of each subgraph Gi,jG_{i,j} in GG^{\prime} we assign the subspaces of 𝔽t+3\mathbb{F}^{t+3} obtained by adding three zeros to the head of all vectors of those subspaces. To the subspaces of dimension 22 in Gi,jG_{i,j}, we add the vector eie_{i} for the vertices adjacent to v1v_{1} and the vector eje_{j} for the vertices adjacent to v2v_{2}. To each of the vertices v1v_{1} and v2v_{2} we assign the subspace of 𝔽t+3\mathbb{F}^{t+3} spanned by e1,e2,e3e_{1},e_{2},e_{3}. Now, for any choice of nonzero vectors for v1,v2v_{1},v_{2}, the subspaces of at least one of the graphs Gi,jG_{i,j} will be restricted to their initial ff-subspace assignment, and will thus admit no valid choice of vectors for its vertices.

    It remains to consider the case of k>3k>3. It suffices to show a polynomial-time reduction from the problem of deciding whether a given bipartite graph is (k1)(k-1)-subspace choosable over 𝔽\mathbb{F} to that of deciding whether a given bipartite graph is kk-subspace choosable over 𝔽\mathbb{F}. Here, given a bipartite graph G=(V,E)G=(V,E) with bipartition V=V1V2V=V_{1}\cup V_{2}, consider the bipartite graph that consists of k2k^{2} copies of GG and two additional vertices v1,v2v_{1},v_{2} such that, for each {1,2}\ell\in\{1,2\}, the vertex vv_{\ell} is adjacent to all the vertices in the copies of VV_{\ell}. The correctness proof is similar to the one given above, so we omit the details.  

By combining Theorem 6.5 with Lemma 6.6, the proof of Theorem 1.8 is completed.

Acknowledgements

We thank the anonymous referees for their very helpful suggestions.

References

  • [1] N. Alon. Restricted colorings of graphs. In K. Walker, editor, Proc. 14th British Combinatorial Conference, London Mathematical Society Lecture Notes Series 187, pages 1–33. Cambridge University Press, Cambridge, 1993.
  • [2] N. Alon. Degrees and choice numbers. Random Struct. Algorithms, 16(4):364–368, 2000.
  • [3] N. Alon, S. Cambie, and R. J. Kang. Asymmetric list sizes in bipartite graphs. Annals of Combinatorics, 25(4):913–933, 2021.
  • [4] N. Alon and J. H. Spencer. The Probabilistic Method. Wiley Publishing, 4th edition, 2016.
  • [5] S. Ball. On sets of vectors of a finite vector space in which every subset of basis size is a basis. J. Eur. Math. Soc., 14(3):733–748, 2012.
  • [6] P. J. Cameron. Combinatorics: Topics, Techniques, Algorithms. Cambridge University Press, Cambridge, NY, 1994.
  • [7] D. Chawin and I. Haviv. Vector choosability in bipartite graphs. Extended Abstracts EuroComb 2021, 14:404–410, 2021.
  • [8] P. Erdös, A. L. Rubin, and H. Taylor. Choosability in graphs. In Proc. West Coast Conf. on Combinatorics, Graph Theory and Computing, Congressus Numerantium XXVI, pages 125–157, 1979.
  • [9] A. Golovnev and I. Haviv. The (generalized) orthogonality dimension of (generalized) Kneser graphs: Bounds and applications. In 36th Computational Complexity Conference (CCC’21), pages 8:1–8:15, 2021.
  • [10] S. Gutner. The complexity of planar graph choosability. Discret. Math., 159(1–3):119–130, 1996.
  • [11] S. Gutner and M. Tarsi. Some results on (a:b)(a:b)-choosability. Discret. Math., 309(8):2260–2270, 2009.
  • [12] W. H. Haemers. On some problems of Lovász concerning the Shannon capacity of a graph. IEEE Trans. Inform. Theory, 25(2):231–232, 1979.
  • [13] G. Haynes, C. Park, A. Schaeffer, J. Webster, and L. H. Mitchell. Orthogonal vector coloring. Electron. J. Comb., 17(1), 2010.
  • [14] L. Lovász. On the Shannon capacity of a graph. IEEE Trans. Inform. Theory, 25(1):1–7, 1979.
  • [15] L. Lovász. Graphs and Geometry, volume 65. Colloquium Publications, 2019.
  • [16] R. Peeters. Orthogonal representations over finite fields and the chromatic number of graphs. Combinatorica, 16(3):417–431, 1996.
  • [17] D. Saxton and A. Thomason. Hypergraph containers. Inventiones Mathematicae, 201(3):925––992, 2015.
  • [18] W. M. Schmidt. Equations over Finite Fields – An Elementary Approach, volume 536 of Lecture Notes in Mathematics. Springer, Berlin, Heidelberg, 1976.
  • [19] V. G. Vizing. Coloring the vertices of a graph in prescribed colors (in Russian). Metody Diskretn. Anal., 29:3–10, 1976.