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

Extremal graphs without exponentially-small bicliques

Boris Bukh
Abstract.

The Turán problem asks for the largest number of edges in an nn-vertex graph not containing a fixed forbidden subgraph FF. We construct a new family of graphs not containing Ks,tK_{s,t}, for t=Cst=C^{s}, with Ω(n21/s)\Omega(n^{2-1/s}) edges matching the upper bound of Kövári, Sós and Turán.

2020 Mathematics Subject Classification:
05D40, 14N07
Supported in part by U.S. taxpayers through NSF CAREER grant DMS-1555149.

1. Introduction

The Turán problem.

Let FF be a fixed graph. The Turán problem asks for the value of ex(n,F)\operatorname{ex}(n,F), the largest number of edges in an nn-vertex graph not containing a copy of FF as a subgraph. The classic theorem of Erdős and Stone [12] gives an asymptotic for ex(n,F)\operatorname{ex}(n,F) when FF is not bipartite.

For bipartite FF, much less is known. Even the simplest case when FF is a complete bipartite graph Ks,tK_{s,t} is open. Specifically, Kövári, Sós and Turán [19] proved that

ex(n,Ks,t)=Os,t(n21/s).\operatorname{ex}(n,K_{s,t})=O_{s,t}(n^{2-1/s}).

Obviously, we may reverse the roles of ss and tt to obtain ex(n,Ks,t)=Os,t(n21/t)\operatorname{ex}(n,K_{s,t})=O_{s,t}(n^{2-1/t}), which is superior if t<st<s. So, from now on we discuss only the case tst\geq s. Though the implicit constant in the big-Oh notation has been improved by Füredi [15], the Kövári–Sós–Turán bound remains the only upper bound on ex(n,Ks,t)\operatorname{ex}(n,K_{s,t}). Many researchers conjecture that the Kövári–Sós–Turán bound is tight (e.g., [19, p. 52], [11, p. 6], [14, p. 257], [16, Conjecture 2.24]). However, apart from the numerous results for s=2s=2 and s=3s=3 (see [16, Section 3] for a survey), there are only two constructions attaining the Kövári–Sós–Turán bound for general s>3s>3. The first is due to Alon, Rónyai and Szabó [2] who, improving on the previous construction by Kollár, Rónyai and Szabó [18], showed that

(1) ex(n,Ks,t)=Ωs(n21/s)if t>(s1)!.\operatorname{ex}(n,K_{s,t})=\Omega_{s}(n^{2-1/s})\qquad\text{if }t>(s-1)!.

The construction is a clever use of norms over finite fields. The second, more recent class of constructions originating from [4] uses random varieties. It has hitherto provided inferior dependence of tt on ss. For example, [4] obtains (1) only for ts4st\geq s^{4s}. The advantage of these constructions is their flexibility, see [9, 6, 20, 17, 7, 24] for some of their variations and applications.

In this work, we use a novel version of the random algebraic method to construct graphs that match the Kövári–Sós–Turán bound for tt that is only exponential in ss.

Theorem 1.

Let s2s\geq 2. Then

ex(n,Ks,t)=Ωs(n21/s)if t>9ss4s2/3.\operatorname{ex}(n,K_{s,t})=\Omega_{s}(n^{2-1/s})\qquad\text{if }t>9^{s}\cdot s^{4s^{2/3}}.

The Zarankiewicz problem.

Closely related to the Turán problem for Ks,tK_{s,t}-free graphs is the problem of Zarankiewicz [25]. It is the asymmetric version of the Turán problem. It is well-known that in the study of the growth rate of ex(n,F)\operatorname{ex}(n,F) we may assume that the nn-vertex graph is bipartite. To distinguish the two parts of a bipartite graph, we shall call them the left and right. A copy of Ks,tK_{s,t} in a bipartite graph GG can be situated in two ways: either the ss vertices are in the left part, or the ss vertices are in the right part. In the Zarankiewicz problem, we forbid only the former case. So, we say that GG is a sided graph if it is bipartite with distinguished left and right parts. The Zarankiewicz problem then asks for the estimate on the number of edges in a sided graph not containing Ks,tK_{s,t}, which is regarded as a sided graph with ss vertices on the left and tt vertices on the right.

An important consequence of making the graph bipartite with distinguished parts is that the two parts can (possibly) be of very unequal size. This often occurs in applications (see e.g. [1, 23]). With this in mind, define z(m,n;s,t)z(m,n;s,t) as the largest number of edges in a sided graph with mm vertices on the left and nn vertices on the right that contains no sided Ks,tK_{s,t}. The bound of Kövári, Sós and Turán for the Zarankiewicz problem takes the form

z(m,n;s,t)=Os,t(mn11/s).z(m,n;s,t)=O_{s,t}(mn^{1-1/s}).

In the symmetric case when m=nm=n, the best known constructions for Zarankiewicz problem are the same as the best bipartite constructions for the Turán problem. This is not so for our approach: we are able to take the advantage of the fact that only one orientation of Ks,tK_{s,t} is forbidden to obtain a lower bound on z(n,n;s,t)z(n,n;s,t) that is superior to the corresponding bound for ex(n,Ks,t)\operatorname{ex}(n,K_{s,t}) in Theorem 1.

Theorem 2.
  1. a)

    Suppose s,t,m,n,k3s,t,m,n,k\geq 3 are integers that satisfy the inequalities lognmsk2k!log2ks\log_{n}m\leq\nobreak\frac{s^{k-2}}{k!\log^{2k}s} and ks2log3sk\leq\frac{s}{2\log^{3}s}. Then

    z(m,n;s,t)=Ωs(mn11/s)if t>kse2s/logs.z(m,n;s,t)=\Omega_{s}(mn^{1-1/s})\qquad\text{if }t>k^{s}\cdot e^{2s/\log s}.

    In particular, z(n,n;s,t)=Ωs(n21/s)z(n,n;s,t)=\Omega_{s}(n^{2-1/s}) for t>3s+o(s)t>3^{s+o(s)}.

  2. b)

    For each s3s\geq 3 there is a constant cs>0c_{s}>0 such that if the inequality lognmcst11+2logs\log_{n}m\leq c_{s}t^{\frac{1}{1+2\log s}} holds, then

    z(m,n;s,t)=Ωs(mn11/s).z(m,n;s,t)=\Omega_{s}(mn^{1-1/s}).

Part (b) is an improvement on the result of Conlon [8], who proved the Ωs(mn11/s)\Omega_{s}(mn^{1-1/s}) bound under the condition lognmcst1/(s1)\log_{n}m\leq c_{s}^{\prime}t^{1/(s-1)}. In the same article, Conlon asked if the bound holds for lognmt/s\log_{n}m\leq t/s, which would be tight if true.

Main proof idea, in a nutshell.

The key novelty in our construction is that it is ‘bumpy’. All the previous algebraic constructions, whether random or not, were flat: the vertex of the graph was always an affine or a projective space over 𝔽q\mathbb{F}_{q}, which was occasionally mildly mutilated by having a lower-dimensional subset removed, or stitched to itself via a quotient operation. These are finite field analogues of n\mathbb{R}^{n} with a flat metric. In contrast, the vertex set in our construction is a solution set to a family of random polynomial equations, and cannot be flattened with any change of coordinates.

Proof ideas, in more detail.

To explain our construction, we recall the important ingredients in the previous random algebraic constructions of Ks,tK_{s,t}-free graphs. The key is an estimate on the probability that the variety 𝐕(f1,,fs)\mathbf{V}(f_{1},\dotsc,f_{s}) cut out by ss random polynomials contains many points. This relies on two inputs. The first is Bezout’s theorem, which is used to conclude that if 𝐕(f1,,fs)\mathbf{V}(f_{1},\dotsc,f_{s}) is zero-dimensional, then it contains at most degfi\prod\deg f_{i} many points. The second input is a bound on the probability that 𝐕(f1,,fs)\mathbf{V}(f_{1},\dotsc,f_{s}) is of codimension less than ss.

To prove that bound, the original paper [4] uses Hilbert functions (though it did not use probabilistic language). However, almost all the subsequent works use the approach from [5] relying on the Lang–Weil bounds, the only exception being an elegant argument in [8] which can be described as an implicit use of Hilbert functions.

In this paper, we go back to the explicit use of Hilbert functions. We show that the codimension of 𝐕(f1,,fs)\mathbf{V}(f_{1},\dotsc,f_{s}) is extremely likely equal to ss, unless ss is close to the dimension of the ambient space. That is precisely the situation in the previous works, where the graph’s vertex set was ss-dimensional set 𝔽qs\mathbb{F}_{q}^{s}. To bypass this obstacle we construct the vertex set in two steps: At the start we use affine space 𝔽qs+r\mathbb{F}_{q}^{s+r} of slightly larger dimension. This way the varieties of the form 𝐕(f1,,fs)\mathbf{V}(f_{1},\dotsc,f_{s}) that we obtain have dimension rr with very high probability. We then shrink the vertex set to a random subvariety of codimension rr, thereby cutting all the varieties of the form 𝐕(f1,,fs)\mathbf{V}(f_{1},\dotsc,f_{s}) at once.

Our second innovation concerns the need to control degfi\prod\deg f_{i} in the Bezout’s bound. In the construction of Turán graphs, the random polynomials f1(y),,fs(y)f_{1}(y),\dotsc,f_{s}(y) arise as specializations of a single polynomial g(x,y)g(x,y) in two sets of variables. A simple way to ensure that the random polynomials f1(y),,fs(y)f_{1}(y),\dotsc,f_{s}(y) are mutually independent is to make gg a random polynomial of degree at least ss. That makes degfi\prod\deg f_{i} grow like sss^{s}. To circumvent this, we further replace the (s+r)(s+r)-dimensional affine space by a variety UU that has the property that every specialization of a random polynomial gg of bounded degree to any ss points of UU yields ss mutually independent polynomials. We call these mm-independent varieties. The construction of mm-independent varieties occupies the bulk of the paper (Sections 4 and 5).

Finally, we want to highlight an auxiliary contribution of this work that is of independent interest. The construction of mm-independent varieties depends on a bound on the number of minimal linear dependencies between mm’th powers of linear forms, i.e., 0=αii(x)m0=\sum\alpha_{i}\ell_{i}(x)^{m} where i(x)=ci,0x0++ci,bxb\ell_{i}(x)=c_{i,0}x_{0}+\dotsb+c_{i,b}x_{b}. Here, ‘minimal’ means that no proper subset of the mm’th powers of these linear forms is linearly dependent. Representation of polynomials by sums of mm’th powers of linear forms has been much studied, motivated primarily by the Waring problem. In particular, we adapt the argument which is implicit in the work of Białynicki-Birula and Schinzel [3] to our purposes. Though the argument in [3] suffices to obtain an exponential bound in Theorem 1, we go beyond and obtain a stronger bound that yields the smaller exponent base of 99 in Theorem 1.

Paper organization.

We begin by collecting the algebraic tools we require in Section 2. The concept of an mm-independent set, which is central to the proof of Theorem 1, is introduced in Section 3. The mm-independent varieties used in the proof of Theorem 1 are constructed in Sections 4 and 5. Finally, in Sections 6 and 7 we prove Theorems 1 and 2 respectively.

Acknowledgments.

I am thankful to Chris Cox for extensive feedback on an earlier version of this paper. I am grateful to Jacob Tsimerman for discussions on numerous topics related to this paper, and especially for motivating me to prove Lemma 22 in its current form, and to Anamay Tengse, Mrinal Kumar, Ramprasad Saptharishi for spotting a serious error in the proof of Lemma 22 in the previous version of this paper. Finally, I am grateful to the anonymous referees for a number of constructive comments.

2. Algebraic tools

To make this paper maximally accessible, we tried to keep the use of algebra to the minimum. In particular, we use counting arguments even when similar algebraic arguments could have provided slightly superior numeric constants. Despite this, we require basic familiarity with algebraic geometry on the level of the first chapter of Shafarevich’s book [21]. We collect the other algebraic tools in this section.

Varieties and their 𝔽q\mathbb{F}_{q}-points.

The integer qq will denote a prime power. We shall work exclusively with fields 𝔽q\mathbb{F}_{q} and 𝔽q¯\overline{\mathbb{F}_{q}}. All varieties in this paper are quasi-projective over the field 𝔽q¯\overline{\mathbb{F}_{q}}. We write 𝐕(f1,,ft)\mathbf{V}(f_{1},\dotsc,f_{t}) for the projective variety cut out by homogeneous polynomials f1,,ftf_{1},\dotsc,f_{t}. We denote the vector space of homogeneous polynomials of degree mm in b+1b+1 variables with coefficients in 𝔽q\mathbb{F}_{q} by 𝔽q[x0,,xb]m\mathbb{F}_{q}[x_{0},\dotsc,x_{b}]_{m}. We also work with products of projective spaces a×b\mathbb{P}^{a}\times\mathbb{P}^{b}. The set of bihomogeneous polynomials of bidegrees (m,m)(m,m^{\prime}) on a×b\mathbb{P}^{a}\times\mathbb{P}^{b} is denoted by 𝔽q[x0,,xa]m𝔽q[y0,,yb]m\mathbb{F}_{q}[x_{0},\dotsc,x_{a}]_{m}\otimes\nobreak\mathbb{F}_{q}[y_{0},\dotsc,y_{b}]_{m^{\prime}}.

We write monomials using the multiindex notation: for a multiindex β=(β0,β1,,βb)0b+1\beta=(\beta_{0},\beta_{1},\dotsc,\beta_{b})\in\mathbb{Z}_{\geq 0}^{b+1}, the notation xβx^{\beta} stands for the monomial x0β0x1β1xbβbx_{0}^{\beta_{0}}x_{1}^{\beta_{1}}\dotsb x_{b}^{\beta_{b}}. A general homogeneous polynomial of degree mm is thus written as |β|=mcβxβ\sum_{\lvert\beta\rvert=m}c_{\beta}x^{\beta}.

The graphs we shall construct in the proofs of Theorems 1 and 2 will consist of the 𝔽q\mathbb{F}_{q}-points of certain varieties. We denote the 𝔽q\mathbb{F}_{q}-points of a variety VbV\subseteq\mathbb{P}^{b} by V(𝔽q)V(\mathbb{F}_{q}) or (if the variety VV is a complicated expression) by Vb(𝔽q)V\cap\mathbb{P}^{b}(\mathbb{F}_{q}). We shall use the following bounds on the number of 𝔽q\mathbb{F}_{q}-points.

Lemma 3 (Weakening of [10, Corollary 3.3]).

Suppose VbV\subseteq\mathbb{P}^{b} is any kk-dimensional variety of degree dd. Then |V(𝔽q)|d|k(𝔽q)|\lvert V(\mathbb{F}_{q})\rvert\leq d\lvert\mathbb{P}^{k}(\mathbb{F}_{q})\rvert.

Lemma 4.
  1. a)

    Let m1,,mrm_{1},\dotsc,m_{r} be positive integers, and let Yb(𝔽q)Y\subseteq\mathbb{P}^{b}(\mathbb{F}_{q}) be a non-empty set. Suppose that g1,,gr𝔽q[x0,,xb]g_{1},\dotsc,g_{r}\in\mathbb{F}_{q}[x_{0},\dotsc,x_{b}] are random homogeneous polynomials of degrees deggi=mi\deg g_{i}=m_{i}. Then

    (2) Pr[|Y𝐕(g1,,gr)||Y|2qr]4qr|Y|.\Pr\left[\lvert Y\cap\mathbf{V}(g_{1},\dotsc,g_{r})\rvert\leq\frac{\lvert Y\rvert}{2q^{r}}\right]\leq\frac{4q^{r}}{\lvert Y\rvert}.
  2. b)

    The same holds for bihomogeneous polynomials, i.e., if Ya(𝔽q)×b(𝔽q)Y\subseteq\mathbb{P}^{a}(\mathbb{F}_{q})\times\mathbb{P}^{b}(\mathbb{F}_{q}) is a non-empty set and g1,,grg_{1},\dotsc,g_{r} are random bihomogeneous polynomials of bidegrees deggi=(mi,mi)\deg g_{i}=(m_{i},m_{i}^{\prime}) with mi,mi1m_{i},m_{i}^{\prime}\geq\nobreak 1, then (2) holds.

Proof.

For a point yYy\in Y, let RyR_{y} be the indicator random variable of the event y𝐕(g1,,gr)y\in\mathbf{V}(g_{1},\dotsc,g_{r}). Let y,yYy,y^{\prime}\in Y be two distinct points. We claim that 𝔼[Ry]=𝔼[Ry]=1/qr\mathbb{E}[R_{y}]=\mathbb{E}[R_{y^{\prime}}]=1/q^{r} and that the random variables RyR_{y} and RyR_{y^{\prime}} are independent. To see this in the case (a), apply a change of coordinates so that y=[1:0:0::0]y=[1:0:0:\dotsc:0] and y=[0:1:0::0]y^{\prime}=[0:1:0:\dotsc:0]. The polynomial gig_{i} vanishes at yy if and only if the coefficient of x0mx_{0}^{m} vanishes. Similarly, gi(y)=0g_{i}(y^{\prime})=0 if and only if the coefficient of x1mx_{1}^{m} vanishes. In the case (b), write y=(ya,yb)y=(y_{a},y_{b}) and y=(ya,yb)y^{\prime}=(y_{a}^{\prime},y_{b}^{\prime}) with ya,yaa(𝔽q)y_{a},y_{a}^{\prime}\in\mathbb{P}^{a}(\mathbb{F}_{q}) and yb,ybb(𝔽q)y_{b},y_{b}^{\prime}\in\mathbb{P}^{b}(\mathbb{F}_{q}). Since yyy\neq y^{\prime} we may assume that ybyby_{b}\neq y_{b}^{\prime} (by swapping the roles of a\mathbb{P}^{a} and b\mathbb{P}^{b} if necessary). We can then change the coordinates on b\mathbb{P}^{b} in the same way as in the case (a), and observe that the vanishing of gig_{i} at yy and yy^{\prime} depends on disjoint sets of coefficients. This proves the claim.

Let R=yYRyR=\sum_{y\in Y}R_{y}. From the pairwise independence of the RyR_{y}’s, it follows that

Var[R]=yYVar[Ry]=1qr(11qr)|Y||Y|/qr.\operatorname{Var}[R]=\sum_{y\in Y}\operatorname{Var}[R_{y}]=\tfrac{1}{q^{r}}(1-\tfrac{1}{q^{r}})\lvert Y\rvert\leq\lvert Y\rvert/q^{r}.

Since 𝔼[R]=y𝔼[Ry]=|Y|/qr\mathbb{E}[R]=\sum_{y}\mathbb{E}[R_{y}]=\lvert Y\rvert/q^{r}, Chebyshev’s inequality then implies that

Pr[|R|Y|/qr|λ|Y|/qr]λ2,\Pr\bigl{[}\lvert R-\lvert Y\rvert/q^{r}\rvert\geq\lambda\sqrt{\lvert Y\rvert/q^{r}}\bigr{]}\leq\lambda^{-2},

and the lemma follows upon taking λ=|Y|/4qr\lambda=\sqrt{\lvert Y\rvert/4q^{r}}. ∎

Bézout’s inequality.

For a reducible variety VbV\subseteq\mathbb{P}^{b}, define the total degree deg(V)\deg(V) to be the sum of the degrees of the irreducible components of VV.

Lemma 5 (Bézout’s inequality, [13, p. 228, Example 12.3.1]).

Let VV and WW be two varieties in b\mathbb{P}^{b}. Then

deg(VW)deg(V)deg(W).\deg(V\cap W)\leq\deg(V)\deg(W).

Hilbert functions.

For a homogeneous ideal II, the Hilbert function HI(m)H_{I}(m) is defined as the codimension of Im=defI𝔽q¯[x0,,xb]mI_{m}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}I\cap\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m} in 𝔽q¯[x0,,xb]m\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m}. Equivalently, if I=I(V)I=I(V) is the homogeneous ideal of polynomials vanishing on a variety VV, then HI(m)H_{I}(m) is the dimension of the subspace of functions on VV induced by all homogeneous polynomials of degree mm.

We shall use the following bound on the Hilbert function.

Lemma 6.

Let VV be a variety of dimension kk in b\mathbb{P}^{b}. Then its Hilbert function satisfies

HI(V)(m)(m+km).H_{I(V)}(m)\geq\binom{m+k}{m}.

Though more general bounds are known (e.g. [22, Theorem 2.4]), we give a proof for completeness.

Proof.

By [21, Theorem 1.15 and Corollary 1.6] there is a linear projection π:Vk\pi\colon V\to\mathbb{P}^{k}, which is finite and surjective. Then the pullback of a homogeneous polynomial of degree mm on k\mathbb{P}^{k} is a homogeneous polynomial of degree mm on VV. The result follows since the space of homogeneous degree-mm polynomials on k\mathbb{P}^{k} is of dimension (m+km)\binom{m+k}{m}, and the pullback map has trivial kernel. ∎

We use Hilbert functions to bound the probability that a random polynomial vanishes on a given variety.

Lemma 7.

Let VV be any variety in b\mathbb{P}^{b}. Let g𝔽q[x0,,xb]mg\in\mathbb{F}_{q}[x_{0},\dotsc,x_{b}]_{m} be a random homogeneous degree-mm polynomial. Then the probability that gg vanishes on VV is

Pr[g|V=0]qHI(V)(m).\Pr[g_{|V}=0]\leq q^{-H_{I(V)}(m)}.
Proof.

By the definition of the Hilbert function, HI(V)(m)H_{I(V)}(m) is the codimension (over 𝔽q¯\overline{\mathbb{F}_{q}}) of I(V)mI(V)_{m} in 𝔽q¯[x0,,xb]m\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m}. This implies that the codimension of I(V)m𝔽q[x0,,xb]I(V)_{m}\cap\mathbb{F}_{q}[x_{0},\dotsc,x_{b}] in 𝔽q[x0,,xb]m\mathbb{F}_{q}[x_{0},\dotsc,x_{b}]_{m} is at least HI(V)(m)H_{I(V)}(m). Since Pr[g|V=0]=Pr[gI(V)m𝔽q[x0,,xb]]\Pr[g_{|V}=0]=\Pr[g\in I(V)_{m}\cap\mathbb{F}_{q}[x_{0},\dotsc,x_{b}]], the lemma follows. ∎

3. Definition and uses of mm-independence

If tt is small compared to bb, then the Hilbert function of tt generic points in b\mathbb{P}^{b} is HI({p1,,pt})(m)=tH_{I(\{p_{1},\dotsc,p_{t}\})}(m)=t for m1m\geq 1. As we shall see shortly, the point sets satisfying HI({p1,,pt})(m)=tH_{I(\{p_{1},\dotsc,p_{t}\})}(m)=t behave random-like with respect to the degree-mm polynomials. Since this property will be key to our construction of Turán graphs, we focus on the sets lacking this pleasant property.

Definition 8.

We say that points p1,,ptbp_{1},\dotsc,p_{t}\in\mathbb{P}^{b} are mm-dependent if HI({p1,,pt})(m)<tH_{I(\{p_{1},\dotsc,p_{t}\})}(m)<t. We furthermore say that they are minimally mm-dependent if no proper subset of these points is mm-dependent.

Definition 9.

A set XbX\subset\mathbb{P}^{b} is called ss-wise mm-independent if no ss distinct points of VV are mm-dependent.

Though we define mm-dependence via Hilbert functions, there are two other ways to think about the concept that will be useful. The first way is to think of a homogeneous degree-mm polynomial ff on b\mathbb{P}^{b} as a linear form in its coefficients. Specialization of ff to f(p)f(p) gives distinct linear forms for distinct pbp\in\mathbb{P}^{b}. It is easy to see from the definition of mm-dependence that the points p1,,ptp_{1},\dotsc,p_{t} are mm-dependent if and only if f(p1),,f(pt)f(p_{1}),\dotsc,f(p_{t}) are linearly dependent as linear forms in (b+mm)\binom{b+m}{m} variables.

The second way to understand mm-dependence is via projective space duality, which we think of as an identification between points of b\mathbb{P}^{b} and linear forms. Namely, to each point pbp\in\mathbb{P}^{b} we associate the linear form \ell in b+1b+1 variables defined by (x)=defx,p\ell(x)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\langle x,p\rangle. Because points in the projective space are defined only up to a multiplication by a non-zero scalar, this linear form too is defined up to a multiplication by a non-zero scalar.

Lemma 10.

Let p1,,ptbp_{1},\dotsc,p_{t}\in\mathbb{P}^{b} be any tt points, and let 1(x),,t(x)\ell_{1}(x),\dotsc,\ell_{t}(x) be the linear forms associated to these points. The following are equivalent:

  1. a)

    points p1,,ptp_{1},\dotsc,p_{t} are mm-dependent,

  2. b)

    there is a linear relation of the form

    (3) c1j=1m1(xj)++ctj=1mt(xj)=0.c_{1}\prod_{j=1}^{m}\ell_{1}(x_{j})+\dotsb+c_{t}\prod_{j=1}^{m}\ell_{t}(x_{j})=0.

Furthermore, if the characteristic of 𝔽q\mathbb{F}_{q} is at least mm, then the following condition is also equivalent to the two above:

  1. c)

    there is a linear relation of the form

    (4) c11(x)m++ctt(x)m=0.c_{1}\ell_{1}(x)^{m}+\dotsb+c_{t}\ell_{t}(x)^{m}=0.
Proof of (a)\iff(b)..

Let k1,,km{0,1,,b}k_{1},\dotsc,k_{m}\in\{0,1,\dotsc,b\} be arbitrary, and let β=(β0,β1,,βb)\beta=(\beta_{0},\beta_{1},\dotsc,\beta_{b}) be the tally vector, i.e., βi\beta_{i} is the number of elements among k1,,kmk_{1},\dotsc,k_{m} that are equal to ii. Then the coefficient of x1,k1xm,kmx_{1,k_{1}}\dotsb x_{m,k_{m}} in ji(xj)\prod_{j}\ell_{i}(x_{j}) is equal to piβp_{i}^{\beta}. Therefore, the linear relation in (3) holds if and only if c1p1β++ctptβ=0c_{1}p_{1}^{\beta}+\dotsb+c_{t}p_{t}^{\beta}=0 for every β0\beta\in\mathbb{Z}_{\geq 0} satisfying |β|=m\lvert\beta\rvert=m. As discussed in the paragraph following Definitions 8 and 9, the latter condition is equivalent to points p1,,ptp_{1},\dotsc,p_{t} being mm-dependent.

Proof of (b)\implies(c)..

Trivial, by setting x1==xmx_{1}=\dotsb=x_{m}.

Proof of (c)\implies(b)..

Set x=x1++xmx=x_{1}+\dotsb+x_{m} in (4), and expand. Let k1,,kmk_{1},\dotsc,k_{m} be arbitrary, and define β\beta as in the proof of (a)\iff(b). The coefficient of x1,k1xm,kmx_{1,k_{1}}\dotsb x_{m,k_{m}} in the expansion is (mβ0,β1,,βb)\binom{m}{\beta_{0},\beta_{1},\dotsc,\beta_{b}} times larger than the coefficient of the same monomial in (3). Since (mβ0,β1,,βb)0\binom{m}{\beta_{0},\beta_{1},\dotsc,\beta_{b}}\neq 0 follows from the assumption on the characteristic, the desired implication follows. ∎

If points p1,,ptp_{1},\dotsc,p_{t} are not mm-dependent, we say that they are mm-independent.

The usefulness of these definitions comes from the combination of two simple observations:

Proposition 11.

Suppose that v1,,vsa(𝔽q)v_{1},\dotsc,v_{s}\in\mathbb{P}^{a}(\mathbb{F}_{q}) are mm-independent points. Pick a random polynomial g𝔽q[x0,,xa]m𝔽q[y0,,yb]mg\in\nobreak\mathbb{F}_{q}[x_{0},\dotsc,x_{a}]_{m}\otimes\mathbb{F}_{q}[y_{0},\dotsc,y_{b}]_{m^{\prime}} uniformly among all bihomogeneous polynomials of bidegree (m,m)(m,m^{\prime}) on a×b\mathbb{P}^{a}\times\mathbb{P}^{b}. Then the ss random polynomials g(v1,y),,g(vs,y)𝔽q[y0,,yb]mg(v_{1},y),\dotsc,g(v_{s},y)\in\mathbb{F}_{q}[y_{0},\dotsc,y_{b}]_{m} are mutually independent.

Proof.

Write gg as g(x,y)=|β|=myβgβ(x)g(x,y)=\sum_{\lvert\beta\rvert=m^{\prime}}y^{\beta}g_{\beta}(x). Since the coefficient of yβy^{\beta} in g(vi,y)g(v_{i},y) is gβ(vi)g_{\beta}(v_{i}), it suffices to show that gβ(v1),,gβ(vs)g_{\beta}(v_{1}),\dotsc,g_{\beta}(v_{s}) are independent for every β\beta.

Think of gβ(vi)g_{\beta}(v_{i}) as a linear function of the coefficients of gβg_{\beta}. By the alternative definition of mm-independence discussed above, the linear functions gβ(v1),,gβ(vs)g_{\beta}(v_{1}),\dotsc,g_{\beta}(v_{s}) are linearly independent. Since the coefficients of gβg_{\beta} are chosen independently and uniformly from 𝔽q\mathbb{F}_{q}, this implies that the random variables gβ(v1),,gβ(vs)g_{\beta}(v_{1}),\dotsc,g_{\beta}(v_{s}) are independent. ∎

Define the function

Mk(t)=defmin{m:(m+kk)t}.M_{k}(t)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\min\Bigl{\{}m:\binom{m+k}{k}\geq t\Bigr{\}}.

Recall that a variety WW is said to be of pure dimension kk if all of its irreducible components are of dimension kk.

Proposition 12.

Let 0skb0\leq s\leq k\leq b be integers, and let WbW\subseteq\mathbb{P}^{b} be a variety of pure dimension kk and degree at most DD.

  1. a)

    If h1,,hsh_{1},\dotsc,h_{s} are independent random homogeneous polynomials of degree mm in b+1b+1 variables with 𝔽q\mathbb{F}_{q}-coefficients, then

    Pr[dim(W𝐕(h1,,hs))>ks]Cq(ks+1+mm),\Pr\bigl{[}\dim(W\cap\mathbf{V}(h_{1},\dotsc,h_{s}))>k-s\bigr{]}\leq Cq^{-\binom{k-s+1+m}{m}},

    where the constant C=C(k,m,D)>0C=C(k,m,D)>0 depends only on k,mk,m and DD.

  2. b)

    Let TT and δ1,,δs\delta_{1},\dotsc,\delta_{s} be positive integers satisfying δiMki+1(T)\delta_{i}\geq M_{k-i+1}(T). If h1,,hsh_{1},\dotsc,h_{s} are independent random homogeneous polynomials of degrees deghi=δi\deg h_{i}=\delta_{i} in b+1b+1 variables with 𝔽q\mathbb{F}_{q}-coefficients, then

    Pr[dim(W𝐕(h1,,hs))>ks]CqT,\Pr\bigl{[}\dim(W\cap\mathbf{V}(h_{1},\dotsc,h_{s}))>k-s\bigr{]}\leq Cq^{-T},

    where the constant C=C(T,s,D,δ1,,δs)>0C=C(T,s,D,\delta_{1},\dotsc,\delta_{s})>0 depends only on T,sT,s,DD and on the polynomial degrees δ1,,δs\delta_{1},\dotsc,\delta_{s}.

Proof.

Part (a) is the special case of part (b) with T=(ks+1+mm)T=\binom{k-s+1+m}{m} and δ1==δs=m\delta_{1}=\dotsb=\delta_{s}=m. So, it suffices to prove part (b).

The proof is by induction on ss. Let 𝒰\mathcal{U} be the set of all irreducible components of W𝐕(h1,,hs1)W\cap\mathbf{V}(h_{1},\dotsc,h_{s-1}). Since 𝐕(h1,,hs1)\mathbf{V}(h_{1},\dotsc,h_{s-1}) is cut out by s1s-1 polynomials and WW is of pure dimension kk, it follows that dim(U)ks+1\dim(U)\geq k-s+1 for each U𝒰U\in\mathcal{U} [21, Corollary 1.14]. So, from Lemmas 6 and 7 we see that, for any fixed U𝒰U\in\mathcal{U},

Pr[hs vanishes on U]q(ks+1+δsks+1)q(ks+1+Mks+1(T)ks+1)qT.\Pr[h_{s}\text{ vanishes on }U]\leq q^{-\binom{k-s+1+\delta_{s}}{k-s+1}}\leq q^{-\binom{k-s+1+M_{k-s+1}(T)}{k-s+1}}\leq q^{-T}.

Bézout’s inequality tells us that |𝒰|Di=1s1δi\lvert\mathcal{U}\rvert\leq D\prod_{i=1}^{s-1}\delta_{i}, and hence the probability that hsh_{s} vanishes on some component of WV(h1,,hs1)W\cap V(h_{1},\dotsc,h_{s-1}) is at most Di=1s1δiqTD\prod_{i=1}^{s-1}\delta_{i}\cdot q^{-T}. By the induction hypothesis, the variety W𝐕(h1,,hs1)W\cap\mathbf{V}(h_{1},\dotsc,h_{s-1}) is of dimension exceeding ks+1k-s+1 with probability at most C(T,s1,D,δ1,,δs1)qTC(T,s-\nobreak 1,D,\delta_{1},\dotsc,\delta_{s-1})q^{-T}, and so the probability that the variety W𝐕(h1,,hs1,hs)W\cap\mathbf{V}(h_{1},\dotsc,h_{s-1},h_{s}) is of dimension exceeding ksk-s is at most

Di=1s1δiqT+C(t,s1,D,δ1,,δs1)qT,D\prod_{i=1}^{s-1}\delta_{i}\cdot q^{-T}+C(t,s-1,D,\delta_{1},\dotsc,\delta_{s-1})q^{-T},

which is at most C(T,s,D,δ1,,δs)qTC(T,s,D,\delta_{1},\dotsc,\delta_{s})q^{-T} for a suitable C()C(\dotsb). ∎

Combining these two observations we obtain the following handy result.

Lemma 13.

Let 0skb0\leq s\leq k\leq b be integers. Suppose that WbW\subseteq\mathbb{P}^{b} is a variety of pure dimension kk, and Xa(𝔽q)X\subset\mathbb{P}^{a}(\mathbb{F}_{q}) is an ss-wise mm-independent set of size |X|cq1s(ks+1+mm)\lvert X\rvert\leq c^{\prime}q^{\frac{1}{s}\binom{k-s+1+m}{m}} where c=c(m,s,degW)>0c^{\prime}=c^{\prime}(m,s,\deg W)>0 is sufficiently small. Let g𝔽q[x0,,xa]m𝔽q[y0,,yb]mg\in\mathbb{F}_{q}[x_{0},\dotsc,x_{a}]_{m}\otimes\mathbb{F}_{q}[y_{0},\dotsc,y_{b}]_{m} be a random bihomogeneous polynomial of bidegree (m,m)(m,m) on a×b\mathbb{P}^{a}\times\mathbb{P}^{b}. Then the following holds with probability at least 45\tfrac{4}{5}: For every ss distinct points v1,,vsXv_{1},\dotsc,v_{s}\in X the variety

{wW:g(v1,w)==g(vs,w)=0}\{w\in W:g(v_{1},w)=\dotsb=g(v_{s},w)=0\}

is of dimension ksk-s.

Proof.

By Proposition 11, polynomials g(v1,w),,g(vs,w)g(v_{1},w),\dotsc,g(v_{s},w) are mutually independent, for every ss distinct points v1,,vsXv_{1},\dotsc,v_{s}\in X. By Item 12(a) and the union bound over all (v1,,vs)Xs(v_{1},\dotsc,v_{s})\in X^{s}, the probability that gg does not satisfy the lemma is at most |X|sC(m,s,degW)q(ks+1+mm)(c)sC(m,s,degW)\lvert X\rvert^{s}\cdot C(m,s,\deg W)q^{-\binom{k-s+1+m}{m}}\leq(c^{\prime})^{s}C(m,s,\deg W). So, choosing cc^{\prime} small enough works. ∎

4. Construction of mm-independent varieties

For positive integers b,m,tb,m,t, let

Φt(b,m)\displaystyle\Phi_{t}(b,m) =def{(p1,,pt)(b)t:p1,,pt are minimally m-dependent},\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(p_{1},\dotsc,p_{t})\in(\mathbb{P}^{b})^{t}:p_{1},\dotsc,p_{t}\text{ are minimally $m$-dependent}\},
ϕt(b,m)\displaystyle\phi_{t}(b,m) =defdimΦt(b,m).\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\dim\Phi_{t}(b,m).

Observe that Φt(b,m)\Phi_{t}(b,m) is a variety. Indeed, for a subset I[t]=def{1,2,,t}I\subseteq[t]\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{1,2,\dotsc,t\}, define

Φ~I(b,m)=def{(p1,,pt)(b)t:(pi)iI are m-dependent}.\widetilde{\Phi}_{I}(b,m)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(p_{1},\dotsc,p_{t})\in(\mathbb{P}^{b})^{t}:(p_{i})_{i\in I}\text{ are $m$-dependent}\}.

The set Φ~I(b,m)\widetilde{\Phi}_{I}(b,m) is a projective variety because of the equivalence of mm-dependence of points and linear dependence of the linear mm-th powers of respective linear forms, which we discussed above. From this, it follows that Φt(b,m)=Φ~[t](b,m)I[t]Φ~I(b,m)\Phi_{t}(b,m)=\widetilde{\Phi}_{[t]}(b,m)\setminus\bigcup_{I\subsetneq[t]}\widetilde{\Phi}_{I}(b,m) is a difference between two projective varieties, and so itself is a (quasiprojective) variety.

We shall upper bound the functions ϕt\phi_{t} in the next section. But first we show how to use these bounds to construct mm-independent varieties.

Lemma 14.

Suppose that b,m,Z,s1b,m,Z,s\geq 1 are integers satisfying

(5) Z>1t1ϕt(b,m) for all t=2,3,,s.Z>\frac{1}{t-1}\phi_{t}(b,m)\qquad\text{ for all }t=2,3,\dotsc,s.

Let f1,,fZf_{1},\dotsc,f_{Z} be generic degree-mm homogeneous polynomials in b+1b+1 variables. Then the variety 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) is ss-wise mm-independent.

As we shall show in Item 22(a), Φt(b,m)=\Phi_{t}(b,m)=\emptyset for tm+1t\leq m+1. Hence, the condition (5) is non-vacuous only for t=m+2,m+3,,st=m+2,m+3,\dotsc,s.

Proof of Lemma 14.

It suffices to show that 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) contains no set of tt minimally mm-dependent points for every t=2,3,,st=2,3,\dotsc,s. Fix tt in this range.

Note that whenever points p1,,ptbp_{1},\dotsc,p_{t}\in\mathbb{P}^{b} are minimally mm-dependent, their Hilbert function satisfies HI({p1,,pt})(m)=t1H_{I(\{p_{1},\dotsc,p_{t}\})}(m)=t-1. Indeed, the inequality HI({p1,,pt})(m)t1H_{I(\{p_{1},\dotsc,p_{t}\})}(m)\leq t-1 follows from the definition of mm-dependence, and HI({p1,,pt})(m)t1H_{I(\{p_{1},\dotsc,p_{t}\})}(m)\geq t-1 follows from minimality. This means that the vector space I({p1,,pt})mI(\{p_{1},\dotsc,p_{t}\})_{m} is of codimension t1t-1 in 𝔽q¯[x0,,xb]m\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m}, and hence I({p1,,pt})mZI(\{p_{1},\dotsc,p_{t}\})_{m}^{Z} is of codimension Z(t1)Z(t-1) in (𝔽q¯[x0,,xb]m)Z(\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m})^{Z}.

Since Z(t1)>ϕt(b,m)Z(t-1)>\phi_{t}(b,m), we can use the algebraic version of the union bound to deduce that for generic degree-mm homogeneous polynomials f1,,fZf_{1},\dotsc,f_{Z}, the variety 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) does not contain any minimally mm-dependent set {p1,,pt}\{p_{1},\dotsc,p_{t}\}. Indeed, write F=def𝔽q¯[x0,,xb]mZF\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m}^{Z} and regard it as a variety with polynomials’ coefficients as indeterminants. Define the variety

V=def{(f1,,fZ,p1,,pt)F×Φt(b,m):fi(pj)=0 for all i,j}.V\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(f_{1},\dotsc,f_{Z},p_{1},\dotsc,p_{t})\in F\times\Phi_{t}(b,m):f_{i}(p_{j})=0\text{ for all }i,j\}.

Consider the projection of VV onto FF. Our claim is that the fiber of a generic fF\vec{f}\in F is empty. If this is not so, then dimVdimF\dim V\geq\dim F. However, for the projection of VV onto the Φt(b,m)\Phi_{t}(b,m) factor, every fiber is of codimension Z(t1)Z(t-1), and so dimVdimΦt(b,m)+(dimFZ(t1))<dimF\dim V\leq\dim\Phi_{t}(b,m)+(\dim F-Z(t-1))<\dim F, which is a contradiction, proving our claim. ∎

Lemma 15.

Suppose that b,m,Z,sb,m,Z,s are integers satisfying the condition (5) and bZ1b-Z\geq 1, and assume that qq0(b,m,Z)q\geq q_{0}(b,m,Z) is sufficiently large in terms of b,m,Zb,m,Z. Then there exist degree-mm polynomials f1,,fZ𝔽q[x0,,xb]mf_{1},\dotsc,f_{Z}\in\mathbb{F}_{q}[x_{0},\dotsc,x_{b}]_{m} such that the variety 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) is ss-wise mm-independent, is of pure dimension bZb-Z, and contains at least 12qbZ\tfrac{1}{2}q^{b-Z} many 𝔽q\mathbb{F}_{q}-points.

Proof.

We shall choose each f1,,fZf_{1},\dotsc,f_{Z} uniformly at random from 𝔽q[x0,,xb]m\mathbb{F}_{q}[x_{0},\dotsc,x_{b}]_{m}. Let

B\displaystyle B =def{(f1,,fZ):𝐕(f1,,fZ) is not s-wise m-independent}\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(f_{1},\dotsc,f_{Z}):\mathbf{V}(f_{1},\dotsc,f_{Z})\text{ is not }s\text{-wise }m\text{-independent}\}
𝔽q¯[x0,,xb]mZ.\displaystyle\subseteq\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m}^{Z}.
The set BB is a variety, for we can think of it as the image of the projection of the bigger variety
B\displaystyle B^{\prime} =def{(f1,,fZ,p1,,pt):fi(pj)=0 for all i,j}\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(f_{1},\dotsc,f_{Z},p_{1},\dotsc,p_{t}):f_{i}(p_{j})=0\text{ for all }i,j\}
𝔽q¯[x0,,xb]mZ×Φt(b,m)\displaystyle\subseteq\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m}^{Z}\times\Phi_{t}(b,m)

onto the first factor.

Furthermore, recall that Φt(b,m)\Phi_{t}(b,m) is defined by the linear dependence of the mm-th powers of the linear forms associated to the points pip_{i}. Since the number of linear forms and the numbers of variables therein do not depend on qq, using Bézout’s inequality we may obtain an upper bound on the degree BB^{\prime} (and hence on the degree of BB) that is independent of qq (but depends on b,mb,m and ZZ).

By Lemma 14, BB is of codimension at least 11 in 𝔽q¯[x0,,xb]mZ\overline{\mathbb{F}_{q}}[x_{0},\dotsc,x_{b}]_{m}^{Z}. By Lemma 3, a random element 𝔽q[x0,,xb]mZ\mathbb{F}_{q}[x_{0},\dotsc,x_{b}]_{m}^{Z} is in BB with probability O(1qdegB)O(\frac{1}{q}\deg B). Since degB\deg B is independent of qq, this probability is O(1/q)O(1/q).

Similarly, since 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) is of codimension ZZ for generic polynomials f1,,fZf_{1},\dotsc,f_{Z}, it follows that 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) is of smaller codimension with probability O(1/q)O(1/q) for random polynomials f1,,fZf_{1},\dotsc,f_{Z}. Furthermore, since 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) is defined by ZZ polynomials, no component of it can have codimension more than ZZ (see [21, Corollary 1.14]), and so 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) is of pure dimension bZb-Z.

Let Y=b(𝔽q)Y=\mathbb{P}^{b}(\mathbb{F}_{q}). Applying Item 4(a) we see that

Pr[|𝐕(f1,,fZ)b(𝔽q)|12qbZ]4qZb.\Pr\bigl{[}\lvert\mathbf{V}(f_{1},\dotsc,f_{Z})\cap\mathbb{P}^{b}(\mathbb{F}_{q})\rvert\leq\tfrac{1}{2}q^{b-Z}\bigr{]}\leq 4q^{Z-b}.

Since bZ1b-Z\geq 1, this is also O(1/q)O(1/q), and so random polynomials f1,,fZf_{1},\dotsc,f_{Z} satisfy the conclusion of the lemma with probability 1O(1/q)1-O(1/q). ∎

5. Upper bound on ϕt(b,m)\phi_{t}(b,m)

For the purpose of proving an exponential bound in Theorem 1, we need only a bound of the form ϕt(b,m)(1ε)bt\phi_{t}(b,m)\leq(1-\varepsilon)bt that is valid for small tt. We go a step further: it can be shown that m+2m+2 points are mm-dependent if and only if they are collinear, and so ϕm+2(b,m)=2(b1)+m\phi_{m+2}(b,m)=2(b-1)+m. The bound on ϕt(b,m)\phi_{t}(b,m) that we prove is sufficiently strong that the maximum of the quantity 1t1ϕt(b,m)\frac{1}{t-1}\phi_{t}(b,m) in our application is achieved at t=m+2t=m+2, and so further improvements in the bound would not lead to a smaller base of exponent in Theorem 1.

Definition 16.

We say that points p1,,ptbp_{1},\dotsc,p_{t}\in\mathbb{P}^{b} are strongly mm-dependent if they span b\mathbb{P}^{b} and the associated linear forms i(x)=x,pi\ell_{i}(x)=\langle x,p_{i}\rangle satisfy a relation of the form

(6) c1j=1m1(xj)++ctj=1mt(xj)=0c_{1}\prod_{j=1}^{m}\ell_{1}(x_{j})+\dotsb+c_{t}\prod_{j=1}^{m}\ell_{t}(x_{j})=0

in which all the coefficients c1,,ctc_{1},\dotsc,c_{t} are non-zero.

Note that Lemma 10 implies that every minimal mm-dependent set is strongly mm-dependent inside whichever space it spans. The key to our upper bound on ϕt(b,m)\phi_{t}(b,m) is a lower bound on the number of points in any strongly mm-dependent set. We begin by proving a relatively weak such bound, adapting the argument which is implicit in the work of Białynicki-Birula and Schinzel [3]. We then bootstrap this weak bound to a stronger bound in Lemma 20.

Lemma 17.

For m2m\geq 2, every strongly mm-dependent set in b\mathbb{P}^{b} has at least 2(b+1)2(b+1) many points.

Proof.

Let p1,,ptp_{1},\dotsc,p_{t} be strongly mm-dependent points in b\mathbb{P}^{b}. By renumbering the points if necessary, we may assume that the points p1,,pb+1p_{1},\dotsc,p_{b+1} span b\mathbb{P}^{b}. If t2b+1t\leq 2b+1, then the remaining points pb+2,,ptp_{b+2},\dotsc,p_{t} do not span b\mathbb{P}^{b} because there are only tbbt-b\leq b of them. Hence, there is an aba\in\mathbb{P}^{b} such that a,pi=0\langle a,p_{i}\rangle=0 for i=b+2,,ti=b+2,\dotsc,t. Plugging aa in for each of x2,x3,,xmx_{2},x_{3},\dotsc,x_{m}, in (6) and writing simply xx for x1x_{1} we obtain a relation of the form

c11(x)++cb+1b+1(x)=0c_{1}^{\prime}\ell_{1}(x)+\dotsb+c_{b+1}^{\prime}\ell_{b+1}(x)=0

where ci=cia,pim1c_{i}^{\prime}=c_{i}\langle a,p_{i}\rangle^{m-1}. Since p1,,pb+1p_{1},\dotsc,p_{b+1} span b\mathbb{P}^{b}, not all the coefficients cic_{i}^{\prime} vanish. Because the points p1,,pb+1p_{1},\dotsc,p_{b+1} are linearly independent, the linear forms 1,,b+1\ell_{1},\dotsc,\ell_{b+1} are linearly independent, and so we reach a contradiction with the previously-made assumption that t2b+1t\leq 2b+1. ∎

To prove the stronger bound, we need to establish a couple of auxiliary results.

Lemma 18.

Let GG be an nn-vertex graph with at most nn edges. Then GG contains an independent set of size n/3n/3.

Proof.

According to Turán’s theorem applied to the complement of GG, the independence number of GG is minimized when GG is a union of cliques whose sizes differ by at most 11. So, assume that GG is of this form. This implies that the largest clique in GG is of size at most 33, for otherwise GG would have had more than nn edges. Hence, GG is a union of at least n/3n/3 disjoint cliques, and so has an independent set of size n/3\lceil n/3\rceil. ∎

Lemma 19.

Suppose that VV is a finite-dimensional vector space (over any field), BB and BB^{\prime} are each a basis for VV, and that no vector in BB is a multiple of any vector in BB^{\prime}, and vice versa. Then there exists a subset CBC\subset B of size |C|13dimV\lvert C\rvert\geq\tfrac{1}{3}\dim V such that Bspan(C)=B^{\prime}\cap\operatorname{span}(C)=\emptyset.

Proof.

Given a vector vBv\in B^{\prime}, express it in the basis BB as v=bBcb,vbv=\sum_{b\in B}c_{b,v}b, and define the set S(v)=def{bB:cb,v0}S(v)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{b\in B:c_{b,v}\neq 0\}. Equivalently, the set S(v)S(v) is the support of vv, when expressed in the basis BB.

Let H=def{S(v):vB}H\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{S(v):v\in B^{\prime}\}. Since no vector of BB^{\prime} is a multiple of any vector in BB, it follows that |S(v)|2\lvert S(v)\rvert\geq 2, and so we may view HH as a (potentially non-uniform) hypergraph all whose edges have at least two elements. Since vspan(C)v\in\operatorname{span}(C) if and only if S(v)CS(v)\subseteq C, our task is to find an independent in HH of size n/3n/3.

Replace each edge in HH by an arbitrary two-element subset thereof, obtaining a graph GG. An independent set in GG is also an independent set in HH. Since GG has dimV\dim V vertices and at most dimV\dim V distinct edges, an appeal to Lemma 18 completes the proof. ∎

With this in place, we are ready to prove the stronger bound on ϕt(b,m)\phi_{t}(b,m).

Lemma 20.

For m2m\geq 2, every strongly mm-dependent set in b\mathbb{P}^{b} has at least m+43(b+1)\frac{m+4}{3}(b+1) many points.

Proof.

The proof is by induction on mm. The base case m=2m=2 is contained in Lemma 17. So, assume that m3m\geq 3.

Let PP be strongly mm-dependent. Let BPB\subset P be any set of b+1b+1 points that span b\mathbb{P}^{b}. We break into two cases.

Suppose that the set PBP\setminus B does not span b\mathbb{P}^{b}. In this case we proceed similarly to the proof of Lemma 17. Namely, we choose an aba\in\mathbb{P}^{b} such that a,p=0\langle a,p\rangle=0 for all pPBp\in P\setminus B. Plugging aa in for each of x2,x3,,xmx_{2},x_{3},\dotsc,x_{m} in (6) and writing xx for x1x_{1}, we obtain a non-trivial linear relation between the b+1b+1 linear forms x,p\langle x,p\rangle with pBp\in B, contradicting the fact that BB spans b\mathbb{P}^{b}.

Suppose that PBP\setminus B spans b\mathbb{P}^{b}. Let BB^{\prime} be any b+1b+1 points of PBP\setminus B that span b\mathbb{P}^{b}. Thinking of BB^{\prime} as points in a vector space of dimension b+1b+1, from Lemma 19 it follows that there is CBC\subset B of size |C|(b+1)/3\lvert C\rvert\geq(b+1)/3 such that Bspan(C)=B^{\prime}\cap\operatorname{span}(C)=\emptyset. Since the field 𝔽q¯\overline{\mathbb{F}_{q}} is infinite, there is an aspan(C)a\in\operatorname{span}(C)^{\bot} such that a,p0\langle a,p\rangle\neq 0 for all pPspan(C)p\in P\setminus\operatorname{span}(C). Setting xm=ax_{m}=a in (6) hence yields a relation

pPspan(C)cpj=1m1p(xj)=0,\sum_{p\in P\setminus\operatorname{span}(C)}c_{p}^{\prime}\prod_{j=1}^{m-1}\ell_{p}(x_{j})=0,

where cp=defcpa,p0c_{p}^{\prime}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}c_{p}\langle a,p\rangle\neq 0. Since the set Pspan(C)P\setminus\operatorname{span}(C) contains BB^{\prime}, this relation shows that Pspan(C)P\setminus\operatorname{span}(C) is a strongly (m1)(m-1)-dependent set. Since |P||Pspan(C)|+|C|\lvert P\rvert\geq\lvert P\setminus\operatorname{span}(C)\rvert+\lvert C\rvert, we are done by induction. ∎

Define

Φt(b,m)\displaystyle\Phi^{\prime}_{t}(b,m) =def{(p1,,pt)Φt(b,m):points p1,,pt span b},\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(p_{1},\dotsc,p_{t})\in\Phi_{t}(b,m):\text{points }p_{1},\dotsc,p_{t}\text{ span }\mathbb{P}^{b}\},
ϕt(b,m)\displaystyle\phi^{\prime}_{t}(b,m) =defdimΦt(b,m).\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\dim\Phi^{\prime}_{t}(b,m).
Lemma 21.

Suppose that m3m\geq 3. Then ϕt(b,m)(tb1)(b+1)\phi_{t}^{\prime}(b,m)\leq(t-b-1)(b+1).

Proof.

For notational brevity, let \mathcal{L} denote the vector space of all non-zero linear forms in b+1b+1 variables. Let =def{0}\mathcal{L}_{*}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathcal{L}\setminus\{0\} and define

U\displaystyle U =def{(1,,t)t:span{1,,t}=},\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(\ell_{1},\dotsc,\ell_{t})\in\mathcal{L}_{*}^{t}:\operatorname{span}\{\ell_{1},\dotsc,\ell_{t}\}=\mathcal{L}\},
Wt\displaystyle W_{t}^{\prime} =def{(1,,t)U:i=1tcij=1mi(xj)=0 for unique c1,,ct0},\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(\ell_{1},\dotsc,\ell_{t})\in U:\sum_{i=1}^{t}c_{i}\prod_{j=1}^{m}\ell_{i}(x_{j})=0\text{ for unique }c_{1},\dotsc,c_{t}\neq 0\},
Wt\displaystyle W_{t} =def{(1,,t)U:i=1tj=1mi(xj)=0}.\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{(\ell_{1},\dotsc,\ell_{t})\in U:\sum_{i=1}^{t}\prod_{j=1}^{m}\ell_{i}(x_{j})=0\}.

Since each point in b\mathbb{P}^{b} corresponds to a 11-dimensional family of linear forms (differing up to a multiplication by a non-zero scalar), a collection of tt points corresponds to a tt-dimensional family of linear forms. Using the characterization of mm-dependence in Lemma 10 this implies that dimΦt(b,m)+tdimWt\dim\Phi^{\prime}_{t}(b,m)+t\leq\dim W_{t}^{\prime}. Since we also have dimWtdimWt+t\dim W_{t}^{\prime}\leq\dim W_{t}+t, it follows that dimΦt(b,m)dimWt\dim\Phi^{\prime}_{t}(b,m)\leq\dim W_{t}. We shall upper bound dimWt\dim W_{t} by bounding the dimension of the tangent space at every point of WtW_{t}.

Let (1,,t)Wt(\ell_{1},\dotsc,\ell_{t})\in W_{t}. Since 1,,t\ell_{1},\dotsc,\ell_{t} span \mathcal{L}, by renumbering the forms, we may assume that the forms 1,,b+1\ell_{1},\dotsc,\ell_{b+1} span \mathcal{L}. Furthermore, by applying a linear change of coordinates, we may also assume that i(y0,,yb)=yi1\ell_{i}(y_{0},\dotsc,y_{b})=y_{i-1} for each i=1,2,,b+1i=1,2,\dotsc,b+1. Then (Δ1,,Δt)t(\Delta_{1},\dotsc,\Delta_{t})\in\mathcal{L}^{t} is in the tangent space to WtW_{t} at the point (1,,t)Wt(\ell_{1},\dotsc,\ell_{t})\in W_{t} if and only if

(7) i=1b+1Δi(xk)jkxj,i1+i=b+2tk=1mΔi(xk)jki(xj)=0.\sum_{i=1}^{b+1}\Delta_{i}(x_{k})\prod_{j\neq k}x_{j,i-1}+\sum_{i=b+2}^{t}\sum_{k=1}^{m}\Delta_{i}(x_{k})\prod_{j\neq k}\ell_{i}(x_{j})=0.

Think of this condition as a system of linear equations. Each of the unknowns Δ1,,Δt\Delta_{1},\dotsc,\Delta_{t}\in\mathcal{L} can be thought of as a vector of b+1b+1 many scalar unknowns. There is one linear equation for each monomial in the xx-variables appearing in (7). For ib+1i\leq b+1, all the monomials appearing as coefficients of Δi\Delta_{i} are a product of m1m-1 variables of the form xj,i1x_{j,i-1}, and a single other variable. Since m3m\geq 3, this means the monomials appearing as coefficients of Δi\Delta_{i}, for ib+1i\leq b+1, are disjoint. This means that the system of equations is of rank at least (b+1)2(b+1)^{2}, i.e., the tangent space to WtW_{t} at (1,,t)(\ell_{1},\dotsc,\ell_{t}) is of codimension at least (b+1)2(b+1)^{2}. Therefore,

ϕt(b,m)dimWtt(b+1)(b+1)2=(tb1)(b+1).\phi^{\prime}_{t}(b,m)\leq\dim W_{t}\leq t(b+1)-(b+1)^{2}=(t-b-1)(b+1).\qed
Lemma 22.

Suppose that t,b,m3t,b,m\geq 3 are integers. Then

  1. a)

    Φt(b,m)=\Phi_{t}(b,m)=\emptyset if tm+1t\leq m+1, and

  2. b)

    ϕt(b,m)3m+4t(b+1+m2m+4t)\phi_{t}(b,m)\leq\lfloor\tfrac{3}{m+4}t\rfloor(b+1+\tfrac{m-2}{m+4}t) if m+2tbm+2\leq t\leq b.

Proof.

From Lagrange interpolation it follows that no set of m+1m+1 or fewer points is mm-dependent. Hence Φt(b,m)=\Phi_{t}(b,m)=\emptyset if tm+1t\leq m+1.

Let r=3m+4t1r^{\prime}=\lfloor\tfrac{3}{m+4}t-1\rfloor. By Lemma 20, every minimal mm-dependent set of size tt spans a subspace of projective dimension at most rr^{\prime}. Let Gr(r,b)\operatorname{Gr}(r,\mathbb{P}^{b}) be the Grassmanian of linear subspaces of dimension rr in b\mathbb{P}^{b}. We then have

ϕt(b,m)\displaystyle\phi_{t}(b,m) =dimΦt(b,m)\displaystyle=\dim\Phi_{t}(b,m)
=dimrr{(p1,,pt,H)Φt(b,m)×Gr(r,b):p1,,pt span H},\displaystyle=\dim\bigcup_{r\leq r^{\prime}}\{(p_{1},\dotsc,p_{t},H)\in\Phi_{t}(b,m)\times\operatorname{Gr}(r,\mathbb{P}^{b}):p_{1},\dotsc,p_{t}\text{ span }H\},
=maxrr(ϕt(r,m)+dimGr(r,b))\displaystyle=\max_{r\leq r^{\prime}}\bigl{(}\phi^{\prime}_{t}(r,m)+\dim\operatorname{Gr}(r,\mathbb{P}^{b})\bigr{)}
maxrr((tr1)(r+1)+(r+1)(br))by Lemma 21\displaystyle\leq\max_{r\leq r^{\prime}}\bigl{(}(t-r-1)(r+1)+(r+1)(b-r)\bigr{)}\qquad\qquad\text{by \lx@cref{creftypecap~refnum}{lem:vdtprime}}
=maxrr(t+b2r1)(r+1).\displaystyle=\max_{r\leq r^{\prime}}(t+b-2r-1)(r+1).
Without the restriction on rr, the maximum of the quadratic (t+b2r1)(r+1)(t+b-2r-1)(r+1) is achieved when r=t+b34r=\frac{t+b-3}{4}. Since btb\geq t, the value t+b34\frac{t+b-3}{4} is larger than rr^{\prime}, and so
ϕt(b,m)\displaystyle\phi_{t}(b,m) (t+b2r1)(r+1)(t+b+16m+4t)3m+4t.\displaystyle\leq(t+b-2r^{\prime}-1)(r^{\prime}+1)\leq(t+b+1-\tfrac{6}{m+4}t)\lfloor\tfrac{3}{m+4}t\rfloor.\hskip-60.00009pt

6. Construction for the Turán problem

Lemma 23.

Let r,s,Z1r,s,Z\geq 1 be integers. Set b=r+s+Zb=r+s+Z. Suppose that b,m,Z,rb,m,Z,r satisfy the inequalities (m+1+rm)s2\binom{m+1+r}{m}\geq s^{2}, m3m\geq 3 as well as the condition (5) for all fields of sufficiently large characteristic. Then, for every sufficiently large nn there is a graph with Θ(n)\Theta(n) vertices and Ω(n21/s)\Omega(n^{2-1/s}) edges that is Ks,tK_{s,t}-free for every t>ms+Zi=1rMi(s2)t>m^{s+Z}\prod_{i=1}^{r}M_{i}(s^{2}).

Proof.

Using Bertrand’s postulate, pick a prime qq such that nqs<2snn\leq q^{s}<2^{s}n. Since nn is sufficiently large, we may assume that the condition (5) is satisfied for 𝔽q¯\overline{\mathbb{F}_{q}}.

Let f1,,fZf_{1},\dotsc,f_{Z} be polynomials as in Lemma 15. Let δi=defMri+1(s2)\delta_{i}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}M_{r-i+1}(s^{2}) for i=1,2,,ri=1,2,\dotsc,r. Let h1,,hrh_{1},\dotsc,h_{r} and h1,,hrh_{1}^{\prime},\dotsc,h_{r}^{\prime} be two independent collections of random homogeneous polynomials on b\mathbb{P}^{b} with 𝔽q\mathbb{F}_{q}-coefficients of degrees deghi=deghi=δi\deg h_{i}=\deg h_{i}^{\prime}=\delta_{i}. Let gg be a random bihomogeneous 𝔽q\mathbb{F}_{q}-polynomial on b×b\mathbb{P}^{b}\times\mathbb{P}^{b} of bidegree (m,m)(m,m).

Define

L0\displaystyle L_{0} =def𝐕(f1,,fZ,h1,,hr)b(𝔽q),\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathbf{V}(f_{1},\dotsc,f_{Z},h_{1},\dotsc,h_{r})\cap\mathbb{P}^{b}(\mathbb{F}_{q}),
R0\displaystyle R_{0} =def𝐕(f1,,fZ,h1,,hr)b(𝔽q).\displaystyle\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathbf{V}(f_{1},\dotsc,f_{Z},h_{1}^{\prime},\dotsc,h_{r}^{\prime})\cap\mathbb{P}^{b}(\mathbb{F}_{q}).

Let c=c(m,s,degW)c^{\prime}=c^{\prime}(m,s,\deg W) be the constant from Lemma 13 applied with W=𝐕(f1,,fZ)W=\mathbf{V}(f_{1},\dotsc,f_{Z}), and let C=C(s2,s,ms+Z,δ1,,δr)C=C(s^{2},s,m^{s+Z},\delta_{1},\dotsc,\delta_{r}) be the constant from Item 12(b). Put c=defmin(14,c,(5C)1/s)c\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\min\bigl{(}\tfrac{1}{4},c^{\prime},(5C)^{-1/s}\bigr{)}. Let LL be a set of size min(cqs,|L0|)\min(cq^{s},\lvert L_{0}\rvert) chosen canonically from L0L_{0} (for example, it could be the initial segment of L0L_{0} in some fixed ordering of b\mathbb{P}^{b}). Let RR be a similarly defined subset of R0R_{0} of size min(cqs,|R0|)\min(cq^{s},\lvert R_{0}\rvert).

Define the bipartite graph GG with parts LL and RR by connecting the pair (l,r)L×R(l,r)\in L\times R whenever g(l,r)=0g(l,r)=0.

GG has Θ(qs)\Theta(q^{s}) vertices and Ω(q2s1)\Omega(q^{2s-1}) edges.

By Item 4(a), both L0L_{0} and R0R_{0} have at least 14qs\tfrac{1}{4}q^{s} elements with probability 1O(1/qs)1-O(1/q^{s}). Since c14c\leq\tfrac{1}{4}, this means that

(8) Pr[|L|=|R|=cqs]1O(qs).\Pr\bigl{[}\lvert L\rvert=\lvert R\rvert=cq^{s}\bigr{]}\geq 1-O(q^{-s}).

Since the polynomial gg is independent of L0L_{0} and R0R_{0}, it follows, by Item 4(b) applied with Y=L×RY=L\times R to the single random polynomial gg, that the edge set E(G)=(L×R)𝐕(g)E(G)=(L\times R)\cap\mathbf{V}(g) is of size at least |L||R|/2q\lvert L\rvert\lvert R\rvert/2q with probability at least 1O(q2s+1)1-O(q^{-2s+1}). In view of (8), it then follows that

(9) Pr[E(G)12c2q2s1]1O(qs).\Pr[E(G)\geq\tfrac{1}{2}c^{2}q^{2s-1}]\geq 1-O(q^{-s}).

GG is Ks,tK_{s,t}-free.

Because of the symmetry between the two parts of GG, it suffices to show that GG is very unlikely to contain Ks,tK_{s,t} with the part of size ss embedded into LL. Since L𝐕(f1,,fZ)L\subset\mathbf{V}(f_{1},\dotsc,f_{Z}) and 𝐕(f1,,fZ)\mathbf{V}(f_{1},\dotsc,f_{Z}) is mm-independent, the set LL is mm-independent. Let W=def𝐕(f1,,fZ)W\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathbf{V}(f_{1},\dotsc,f_{Z}). Since (m+1+rm)s2\binom{m+1+r}{m}\geq s^{2}, Lemma 13 applies, telling us that with probability 45\tfrac{4}{5} every variety of the form

Wl1,,ls=def{wW:g(l1,w)==g(ls,w)=0},W_{l_{1},\dotsc,l_{s}}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{w\in W:g(l_{1},w)=\dotsb=g(l_{s},w)=0\},

for distinct l1,,lsl_{1},\dotsc,l_{s}, is of dimension dimWs=r\dim W-s=r. Let 𝒲\mathcal{W} be the set of all varieties of the form Wl1,,lsW_{l_{1},\dotsc,l_{s}} for distinct l1,,lsLl_{1},\dotsc,l_{s}\in L. The set 𝒲\mathcal{W} is random: it depends on the random choice of polynomials h1,,hrh_{1},\dotsc,h_{r} (because LL depends on these polynomials), and it depends on the polynomial gg. Crucially, 𝒲\mathcal{W} does not depend on the polynomials h1,,hrh_{1}^{\prime},\dotsc,h_{r}^{\prime}.

So, we may apply Item 12(b) to each variety in 𝒲\mathcal{W} and polynomials h1,,hrh_{1}^{\prime},\dotsc,h_{r}^{\prime}. Note that deg(Wl1,,ls)deg(W)msms+Z\deg(W_{l_{1},\dotsc,l_{s}})\leq\deg(W)m^{s}\leq m^{s+Z} by Bézout’s inequality. Therefore, combined with the union bound, the proposition tells us that

Pr[\displaystyle\Pr[\exists distinct l1,,lsL s.t. ​dim(Wl1,,ls𝐕(h1,,hr))>0]\displaystyle\text{ distinct }l_{1},\dotsc,l_{s}\in L\text{ s.t.\ \!}\dim\bigl{(}W_{l_{1},\dotsc,l_{s}}\cap\mathbf{V}(h_{1}^{\prime},\dotsc,h_{r}^{\prime})\bigr{)}>0]
Pr[dim(Wl1,,ls)r for some l1,,ls]\displaystyle\leq\Pr[\dim(W_{l_{1},\dotsc,l_{s}})\neq r\text{ for some }l_{1},\dotsc,l_{s}]
+Pr[|L|cqs]+(cqs)sC(s2,s,ms+Z)qs2\displaystyle\qquad+\Pr\bigl{[}\lvert L\rvert\neq cq^{s}\bigr{]}+(cq^{s})^{s}\cdot C(s^{2},s,m^{s+Z})q^{-s^{2}}
15+O(qs)+csC(s2,s,ms+Z).\displaystyle\leq\tfrac{1}{5}+O(q^{-s})+c^{s}C(s^{2},s,m^{s+Z}).

Because the constant cc satisfies c(5C)1/sc\leq(5C)^{-1/s}, this probability is at most 25+O(qs)\tfrac{2}{5}+O(q^{-s}). Note that the variety Wl1,,ls𝐕(h1,,hr)W_{l_{1},\dotsc,l_{s}}\cap\nobreak\mathbf{V}(h_{1}^{\prime},\dotsc,h_{r}^{\prime}) contains all common neighbors of the vertices l1,,lsl_{1},\dotsc,l_{s} in the graph GG. By Bézout’s inequality, the degree of this variety is at most deg(W)msi=1rδi<t\deg(W)m^{s}\prod_{i=1}^{r}\delta_{i}<t, and hence

Pr[some s vertices in L have t common neighbors in G]25+O(qs).\Pr[\text{some }s\text{ vertices in }L\text{ have }t\text{ common neighbors in }G]\leq\tfrac{2}{5}+O(q^{-s}).

By symmetry, we may derive the same bound with the roles of LL and RR reversed, and so

(10) Pr[G contains Ks,t]45+O(qs).\Pr[G\text{ contains }K_{s,t}]\leq\tfrac{4}{5}+O(q^{-s}).

Putting (8), (9) and (10) together, it follows that graph GG has Θ(qs)=Θ(n)\Theta(q^{s})=\Theta(n) vertices, at least Ω(q2s1)=Ω(n21/s)\Omega(q^{2s-1})=\Omega(n^{2-1/s}) edges, and contains no Ks,tK_{s,t} with probability at least 15O(qs)>0\tfrac{1}{5}-O(q^{-s})>0. In particular, such a graph GG exists. ∎

Lemma 24.

For every r1r\geq 1 and every TT, we have k=1rMk(T)T1+logrr!\prod_{k=1}^{r}M_{k}(T)\leq T^{1+\log r}r!.

Proof.

Since (m+km)(m+1k)k\binom{m+k}{m}\geq\bigl{(}\frac{m+1}{k}\bigr{)}^{k}, the function MkM_{k} satisfies Mk(T)kT1/k.M_{k}(T)\leq\lfloor kT^{1/k}\rfloor. The lemma then follows from the inequality 1+12++1r1+logr1+\tfrac{1}{2}+\dotsb+\tfrac{1}{r}\leq 1+\log r. ∎

Proof of Theorem 1.

We may assume that s100s\geq 100, for otherwise the theorem follows from the result of Alon, Rónyai and Szabó [2] that we mentioned in the introduction. Let m=def3m\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}3, r=def(6s2)1/3r\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\lfloor(6s^{2})^{1/3}\rfloor, and Z=defs+r+3Z\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}s+r+3.

Observe that

1t137t{12 for t=5,6,7,715 for t=8,9,10,11,.\tfrac{1}{t-1}\lfloor\tfrac{3}{7}t\rfloor\leq\begin{cases}\tfrac{1}{2}&\text{ for }t=5,6,7,\\ \tfrac{7}{15}&\text{ for }t=8,9,10,11,\dotsc.\\ \end{cases}

Hence, for 5t75\leq t\leq 7, Item 22(b) implies that

1t1ϕt(b,m)\displaystyle\tfrac{1}{t-1}\phi_{t}(b,m) 12(s+r+Z+1+17t)12(s+r+2)+12Z<Z.\displaystyle\leq\tfrac{1}{2}(s+r+Z+1+\tfrac{1}{7}t)\leq\tfrac{1}{2}(s+r+2)+\tfrac{1}{2}Z<Z.
Similarly, for 8ts8\leq t\leq s we have
1t1ϕt(b,m)\displaystyle\tfrac{1}{t-1}\phi_{t}(b,m) 715(s+r+Z+1+17t)715(87s+r+1)+715Z\displaystyle\leq\tfrac{7}{15}(s+r+Z+1+\tfrac{1}{7}t)\leq\tfrac{7}{15}(\tfrac{8}{7}s+r+1)+\tfrac{7}{15}Z
<815(s+r+2)+715Z<Z.\displaystyle<\tfrac{8}{15}(s+r+2)+\tfrac{7}{15}Z<Z.

This shows that the condition (5) holds for t=5,6,,st=5,6,\dotsc,s. It also holds for t=2,3,4t=2,3,4 by Item 22(a).

We have (m+1+rm)=(r+43)(r+1)3/6s2\binom{m+1+r}{m}=\binom{r+4}{3}\geq(r+1)^{3}/6\geq s^{2}, and so the result follows from Lemmas 23 and 24, the estimation

r!(s2)1+logr2s1/2(r/e)r(s2)1+logrs23r+2+2logrsrfor s23,r!(s^{2})^{1+\log r}\leq 2s^{1/2}(r/e)^{r}\cdot(s^{2})^{1+\log r}\leq s^{\tfrac{2}{3}r+2+2\log r}\leq s^{r}\qquad\text{for }s\geq 23,

and the inequality 3r+3srs3r/2s4s2/33^{r+3}s^{r}\leq s^{3r/2}\leq s^{4s^{2/3}} that is valid for s100s\geq 100. ∎

7. Construction for the Zarankiewicz problem

This is similar, but simpler than the construction for the Turán problem from the preceding section because we do not need Lemmas 15 and 22.

Lemma 25.

Let r,s,T1r,s,T\geq 1 be integers satisfying T(r+1+mm)T\leq\binom{r+1+m}{m}. Then, for every sufficiently large nn there is a sided graph with Θ(nT/s2)\Theta(n^{T/s^{2}}) vertices on the left, Θ(n)\Theta(n) vertices on the right, and Ω(nT/s2+11/s)\Omega(n^{T/s^{2}+1-1/s}) edges that contains no sided Ks,tK_{s,t} for every t>msi=1rMi(T)t>m^{s}\prod_{i=1}^{r}M_{i}(T).

Proof.

Let qq be the power of 22 satisfying nqs<2snn\leq q^{s}<2^{s}n. Let b=defr+sb\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}r+s.

Set δi=defMri+1(T)\delta_{i}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}M_{r-i+1}(T) for i=1,,ri=1,\dotsc,r. Let c=c(m,s,1)c^{\prime}=c^{\prime}(m,s,1) be the constant from Lemma 13, and let C=C(T,s,ms,δ1,,δr)C=C\bigl{(}T,s,m^{s},\delta_{1},\dotsc,\delta_{r}\bigr{)} be the constant from Item 12(b). Put c=defmin(c,(5C)1/s)c\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\min(c^{\prime},(5C)^{-1/s}). Pick a1a\geq 1 and an ss-wise mm-independent set LL in a(𝔽q)\mathbb{P}^{a}(\mathbb{F}_{q}) of size cqT/scq^{T/s} arbitrarily. For example, we may choose a=cqT/sa=cq^{T/s} and let LL consist of the standard basis vectors.

We next pick several random polynomials with coefficients in 𝔽q\mathbb{F}_{q}: Let gg be a random bihomogeneous polynomial on a×b\mathbb{P}^{a}\times\mathbb{P}^{b} of bidegree (m,m)(m,m), and let h1,,hrh_{1}^{\prime},\dotsc,h_{r}^{\prime} be random independently-chosen homogeneous polynomials on b\mathbb{P}^{b} of degrees deghi=δi\deg h_{i}^{\prime}=\delta_{i}. Let

R=def𝐕(h1,,hr)b(𝔽q).R\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathbf{V}(h_{1}^{\prime},\dotsc,h_{r}^{\prime})\cap\mathbb{P}^{b}(\mathbb{F}_{q}).

Define the sided graph GG with the left part LL and the right part RR by connecting (l,r)L×R(l,r)\in L\times R whenever g(l,r)=0g(l,r)=0.

Invoking Lemma 13 with W=bW=\mathbb{P}^{b} we see that, with probability 45\tfrac{4}{5}, every variety of the form

Wl1,,ls=def{wb:g(l1,w)==g(ls,w)=0},W_{l_{1},\dotsc,l_{s}}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\{w\in\mathbb{P}^{b}:g(l_{1},w)=\dotsb=g(l_{s},w)=0\},

for distinct l1,,lsLl_{1},\dotsc,l_{s}\in L, has dimension bs=rb-s=r. By Bézout’s inequality, degWl1,,lsms\deg W_{l_{1},\dotsc,l_{s}}\leq m^{s}, and so the union bound and Item 12(b) together imply that

(11) Pr[G contains a sided Ks,t]25+O(qs),\Pr[G\text{ contains a sided }K_{s,t}]\leq\tfrac{2}{5}+O(q^{-s}),

where we used that c(5C)1/sc\leq(5C)^{-1/s}, similarly to the corresponding step in the proof of Lemma 23.

Also, from Item 12(b) applied to W=bW=\mathbb{P}^{b} we obtain the inequality Pr[dim𝐕(h1,,hr)>s]=O(qs)\Pr[\dim\mathbf{V}(h_{1}^{\prime},\dotsc,h_{r}^{\prime})>s]=\nobreak O(q^{-s}). In view of Lemma 3, this implies that

(12) Pr[|R|=Os(qs)]1O(qs).\Pr\bigl{[}\lvert R\rvert=O_{s}(q^{s})\bigr{]}\geq 1-O(q^{-s}).

Finally, by applying Items 4(a) and 4(b), it follows that

Pr[|E(G)|14cqT/s+s1]\displaystyle\Pr\bigl{[}\lvert E(G)\rvert\!\geq\!\tfrac{1}{4}cq^{T/s+s-1}\bigr{]} Pr[|R|12qs]Pr[|E(G)|12cqT/s12qs1||R|12qs]\displaystyle\mkern-2.1mu\geq\mkern-2.1mu\Pr\bigr{[}\lvert R\rvert\!\geq\!\tfrac{1}{2}q^{s}\bigr{]}\!\Pr\bigl{[}\lvert E(G)\rvert\!\geq\!\tfrac{1}{2}cq^{T/s}\!\cdot\!\tfrac{1}{2}q^{s-1}\!\bigm{|}\!\lvert R\rvert\!\geq\!\tfrac{1}{2}q^{s}\bigr{]}
(1O(qs))(1O(qs+1qT/s))\displaystyle\mkern-2.1mu\geq\mkern-2.1mu\bigl{(}1-O(q^{-s})\bigr{)}\bigl{(}1-O(q^{-s+1}q^{-T/s})\bigr{)}
(13) =1O(qs+1).\displaystyle\mkern-2.1mu=\mkern-2.1mu1-O(q^{-s+1}).

From (11), (12), (13) we see that there exists a sided graph with cqT/scq^{T/s} vertices on the left, Os(qs)O_{s}(q^{s}) vertices on the right, and at least 14cqT/s+s1\tfrac{1}{4}cq^{T/s+s-1} edges. ∎

Proof of Item 2(a).

Let T=defs2lognmT\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}s^{2}\log_{n}m, and r=defs/log2sr\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\lfloor s/\log^{2}s\rfloor. Note that these constants satisfy (r+1+kr)(s/log2s)k/k!s2lognm\binom{r+1+k}{r}\geq(s/\log^{2}s)^{k}/k!\geq s^{2}\log_{n}m. We then use Lemma 25 with kk in place of mm, and appeal to Lemma 24 to bound

i=1rMi(T)\displaystyle\prod_{i=1}^{r}M_{i}(T) T1+logrr!s(1+logr)krr\displaystyle\leq T^{1+\log r}r!\leq s^{(1+\log r)k}r^{r}
s(1+logs)s2log3s(s/log2s)s/log2se2s/logs\displaystyle\leq s^{(1+\log s)\frac{s}{2\log^{3}s}}(s/\log^{2}s)^{s/\log^{2}s}\leq e^{2s/\log s}

to obtain the stated result. ∎

Proof of Item 2(b).

Let k1k\geq 1 be an integer to be chosen shortly. Let r=defs/logsr\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\lceil s/\log s\rceil, and T=def(r+1+kk)T\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\binom{r+1+k}{k} Note that there are constants cs′′c_{s}^{\prime\prime} and cs′′′c_{s}^{\prime\prime\prime} such that

cs′′krT/s2cs′′′kr,c_{s}^{\prime\prime}k^{r}\leq T/s^{2}\leq c_{s}^{\prime\prime\prime}k^{r},

and choose kk to be the smallest integer so that cs′′krlognmc_{s}^{\prime\prime}k^{r}\geq\log_{n}m.

Applying Lemma 25 with kk in place of mm, we obtain a sided graph whose left part is of size Ω(nT/s2)=Ω(ncs′′kr)=Ω(m)\Omega(n^{T/s^{2}})=\nobreak\Omega(n^{c_{s}^{\prime\prime}k^{r}})=\Omega(m) and the right part is of size Θ(n)\Theta(n), matching the Kövári, Sós, Turán bound for Ks,tK_{s,t}-free graphs for t>ksi=1rMi(r)t>k^{s}\prod_{i=1}^{r}M_{i}(r). Since

ksi=1rMi(r)ksrrT1+logr=Os(kr+2rlogs)=Os(lognm)1+2logs,k^{s}\prod_{i=1}^{r}M_{i}(r)\leq k^{s}r^{r}T^{1+\log r}=O_{s}(k^{r+2r\log s})=O_{s}(\log_{n}m)^{1+2\log s},

this completes the proof. ∎

References

  • [1] Noga Alon, Keith E. Mellinger, Dhruv Mubayi, and Jacques Verstraëte. The de Bruijn–Erdős theorem for hypergraphs. Des. Codes Cryptogr., 65(3):233–245, 2012. arXiv:1007.4150.
  • [2] Noga Alon, Lajos Rónyai, and Tibor Szabó. Norm-graphs: variations and applications. J. Combin. Theory Ser. B, 76(2):280–290, 1999.
  • [3] A. Białynicki-Birula and A. Schinzel. Representations of multivariate polynomials by sums of univariate polynomials in linear forms. Colloq. Math., 112(2):201–233, 2008.
  • [4] Pavle V. M. Blagojević, Boris Bukh, and Roman Karasev. Turán numbers for Ks,tK_{s,t}-free graphs: topological obstructions and algebraic constructions. Israel J. Math., 197(1):199–214, 2013. arXiv:1207.0705.
  • [5] Boris Bukh. Random algebraic construction of extremal graphs. Bull. Lond. Math. Soc., 47(6):939–945, 2015. arXiv:1409.3856.
  • [6] Boris Bukh and David Conlon. Rational exponents in extremal graph theory. J. Eur. Math. Soc. (JEMS), 20(7):1747–1757, 2018. arXiv:1506.06406.
  • [7] Boris Bukh and Michael Tait. Turán numbers of theta graphs. Combin. Probab. Comput., 29(4):495–507, 2020. arXiv:1804.10014.
  • [8] David Conlon. Some remarks on the Zarankiewicz problem. arXiv:2007.12816.
  • [9] David Conlon. Graphs with few paths of prescribed length between any two vertices. Bull. Lond. Math. Soc., 51(6):1015–1021, 2019.
  • [10] Alain Couvreur. An upper bound on the number of rational points of arbitrary projective varieties over finite fields. Proc. Amer. Math. Soc., 144(9):3671–3685, 2016. arXiv:1409.7544.
  • [11] P. Erdős. Some recent progress on extremal problems in graph theory. In Proceedings of the Sixth Southeastern Conference on Combinatorics, Graph Theory and Computing (Florida Atlantic Univ., Boca Raton, Fla., 1975), pages 3–14. Congressus Numerantium, No. XIV, 1975.
  • [12] P. Erdös and A. H. Stone. On the structure of linear graphs. Bull. Amer. Math. Soc., 52:1087–1091, 1946.
  • [13] William Fulton. Intersection theory, volume 2 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, second edition, 1998.
  • [14] Zoltán Füredi. Turán type problems. In Surveys in combinatorics, 1991 (Guildford, 1991), volume 166 of London Math. Soc. Lecture Note Ser., pages 253–300. Cambridge Univ. Press, Cambridge, 1991.
  • [15] Zoltán Füredi. An upper bound on Zarankiewicz’ problem. Combin. Probab. Comput., 5(1):29–33, 1996.
  • [16] Zoltán Füredi and Miklós Simonovits. The history of degenerate (bipartite) extremal graph problems. In Erdös centennial, volume 25 of Bolyai Soc. Math. Stud., pages 169–264. János Bolyai Math. Soc., Budapest, 2013. arXiv:1306.5167.
  • [17] Zhiyang He and Michael Tait. Hypergraphs with few Berge paths of fixed length between vertices. SIAM J. Discrete Math., 33(3):1472–1481, 2019.
  • [18] János Kollár, Lajos Rónyai, and Tibor Szabó. Norm-graphs and bipartite Turán numbers. Combinatorica, 16(3):399–406, 1996.
  • [19] T. Kövari, V. T. Sós, and P. Turán. On a problem of K. Zarankiewicz. Colloq. Math., 3:50–57, 1954.
  • [20] Jie Ma, Xiaofan Yuan, and Mingwei Zhang. Some extremal results on complete degenerate hypergraphs. J. Combin. Theory Ser. A, 154:598–609, 2018.
  • [21] Igor R. Shafarevich. Basic algebraic geometry. 1. Springer-Verlag, Berlin, second edition, 1994. Varieties in projective space, Translated from the 1988 Russian edition and with notes by Miles Reid.
  • [22] Martín Sombra. Bounds for the Hilbert function of polynomial ideals and for the degrees in the Nullstellensatz. J. Pure Appl. Algebra, 117/118:565–599, 1997. arXiv:alg-geom/9610006.
  • [23] Miguel N. Walsh. The polynomial method over varieties. Invent. Math., 222(2):469–512, 2020. arXiv:1811.07865.
  • [24] Zixiang Xu, Tao Zhang, and Gennian Ge. Some tight lower bounds for Turán problems via constructions of multi-hypergraphs. European J. Combin., 89:103161, 11, 2020. arXiv:1907.11909.
  • [25] Kazimierz Zarankiewicz. Problem 101. Colloq. Math., 2:301, 1951.