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

The upper tail problem for induced 4-cycles in sparse random graphs

Asaf Cohen Antonir
Abstract.

Building on the techniques from the breakthrough paper of Harel, Mousset and Samotij, which solved the upper tail problem for cliques, we compute the asymptotics of the upper tail for the number of induced copies of the 4-cycle in the binomial random graph Gn,pG_{n,p}. We observe a new phenomenon in the theory of large deviations of subgraph counts. This phenomenon is that, in a certain (large) range of pp, the upper tail of the induced 4-cycle does not admit a naive mean-field approximation.

1. Introduction

The study of the random variable counting the number of copies of a given graph in the binomial random graph Gn,pG_{n,p} has a very long history and many things are known about it. In particular, for every graph HH, when the expected number of copies of HH in Gn,pG_{n,p} tends to infinity a ‘law of large numbers’, see Bollobás [6], and a central limit theorem are known, see Ruciński [28].

After establishing such results, it is natural to ask what is the probability of the event that this random variable differs from its expectation by a significant amount. In this paper, we will consider only the ‘upper tail’ problem: what is the probability of a random variable exceeding its expectation by a multiplicative factor 1+δ1+\delta, where δ\delta is some positive real.

We continue by recalling some definitions to make our discussion more rigorous. For every non-negative integer nn and a sequence p=p(n)(0,1)p=p(n)\in(0,1), we let Gn,pG_{n,p} be the binomial random graph with nn vertices and density pp. Furthermore, for any graph HH, let Xn,pHX^{H}_{n,p} be the random variable counting the number of (labeled) copies of HH in Gn,pG_{n,p}. Further, for any δ\delta a positive real and a random variable YY we write 𝐔𝐓(Y,δ)\mathbf{UT}(Y,\delta) for the event {Y(1+δ)𝔼[Y]}\{Y\geq(1+\delta)\mathbb{E}[Y]\}.

The work on the problem of estimating (the logarithm of) the upper tail probability of Xn,pHX_{n,p}^{H} was initiated by the famous works of Kim and Vu [24], Vu [29], Janson and Ruciński [23]. This problem turns out to be so difficult111This problem was called ‘the infamous upper tail’ by Janson and Ruciński [23]. because when HH is a connected graph with at least two edges Xn,pHX_{n,p}^{H} is not a linear function of independent Bernoulli random variables. In the case of linear functions, life is easier and much is known about its large deviation properties, see [15].

In a famous paper [22], Janson, Oleszkiewicz, and Ruciński estimated the logarithm of the upper tail probability for every graph HH up to a multiplicative factor of O(log(1/p))O(\log(1/p)). Seven years later, Chatterjee [8] and DeMarco–Kahn [14] closed this gap up to a multiplicative factor of O(1)O(1) when HH is a triangle and DeMarco–Kahn [13] generalized this for a clique of arbitrary size.

Next, one would like to obtain a first order approximation of the logarithm of the upper tail probability. The first to do that were Chatterjee and Varadhan [10]. They obtained a first order approximation of the upper tail probability for any graph, under the assumption that pp is a constant. Their proof relied on the regularity method and therefore extends only to pp tending to zero very slowly, for more discussion about this see [26].

The general strategy for estimating the logarithm of the upper tail probability, used by Chatterjee and Varadhan as well as all later works on this subject, is to establish a ‘large deviation principle’. That is to prove that the logarithm of the upper tail probability is asymptotically equal to the solution to a minimization problem over a non-trivial set of product probability measures. These minimization problems are called ‘variational problems’. After achieving such large deviation principle one is then left with solving the variational problem.

In a breakthrough paper of Chatterjee–Dembo [9], the authors established a ‘large deviation principle’ not only for Xn,pHX_{n,p}^{H} when p=Ω(nε)p=\Omega(n^{-\varepsilon}), but also for a large class of functions of NN independent Bernoulli random variables with mean p=Ω(Nε)p=\Omega(N^{-\varepsilon}). They proved that for any ‘smooth enough’ function f:{0,1}Nf\colon\{0,1\}^{N}\to\mathbb{R} and a sequence of independent Bernoulli random variables Y=(Y1,Y2,,YN){0,1}NY=(Y_{1},Y_{2},\ldots,Y_{N})\in\{0,1\}^{N}, all with mean pp, we have the following:

log(𝐔𝐓(f(Y),δ))=(1+o(1))inf{i=1NIp(qi):𝔼[f(Y~)](1+δ)𝔼[f(Y)]},-\log\mathbb{P}\left(\mathbf{UT}(f(Y),\delta)\right)=(1+o(1))\inf\left\{\sum_{i=1}^{N}I_{p}(q_{i}):\mathbb{E}[f(\tilde{Y})]\geq(1+\delta)\mathbb{E}[f(Y)]\right\}, (1)

where Ip(q)=qlog(qp)+(1q)log(1q1p)I_{p}(q)=q\log\left(\frac{q}{p}\right)+(1-q)\log\left(\frac{1-q}{1-p}\right) is the relative entropy (the Kullback–Leibler divergence) between Ber(q)\operatorname{Ber}(q) and Ber(p)\operatorname{Ber}(p), and Y~=(Y~1,Y~2,,Y~N){0,1}N\tilde{Y}=(\tilde{Y}_{1},\tilde{Y}_{2},\ldots,\tilde{Y}_{N})\in\{0,1\}^{N} is a sequence of independent Bernoulli random variables with 𝔼[Y~i]=qi\mathbb{E}[\tilde{Y}_{i}]=q_{i} for each ii.

This was further developed by Eldan [16] and Augeri [2, 3]. These methods are completely different from the ones used in the dense regime.

Revisiting the ideas from Chatergee–Varhadan [10], Cook–Dembo [11] developed a decomposition theorem similar to Szemerédi’s regularity lemma and a corresponding counting lemma which are suitable for sparse graphs. Using these they extended the range of sparsity of pp were the variational approximation (1) holds (for every graph HH). Generalizing this method, Cook–Dembo–Pham [12] pushed the bounds even further in the case of subgraph count. They also obtained an approximation of the logarithm of the upper tail probability for the count of uniform sub-hypergraphs in the binomial random uniform hypergraph model, and the induced count of graphs in Gn,pG_{n,p}. Their result combined with the solution to the corresponding variational problem for uniform hypergraph cliques, and one more non-trivial 3-uniform hypergraph which were given by Liu–Zhao in [25], yields estimations for the sub-hypergraph count in the binomial random hypergraph model in the sparse regime. We also note that to the best of our knowledge the solution to the ‘induced variational problem’ is not known apart from the case where HH is a clique. The case of the induced subgraph count will be of main interest in this paper, and will be discussed later.

To estimate the asymptotics of the logarithm of the upper tail probability, one also needs to solve the variational problem in the right-hand side of (1). In the case of the random variable Xn,pHX_{n,p}^{H}, Lubetzky and Zhao [27], and Bhattacharya, Ganguly, Lubetzky, and Zhao [5] solved this variational problem for all H,nH,n and pp satisfying n1/ΔHp1n^{-1/\Delta_{H}}\ll p\ll 1. This then leads to a first order approximation of the logarithm of the upper tail probability for Xn,pHX_{n,p}^{H} for every graph HH and pncHp\geq n^{-c_{H}}. We wish to emphasize that, in all known cases, when ff counts the number of copies of a given HH in Gn,pG_{n,p}, the solution to the corresponding variational problem (the right hand side of (1)) always satisfies qi{p,1}q_{i}\in\{p,1\} when p=o(1)p=o(1).

A recent surprising result due to Gunby [20] studied the upper tail probability for subgraph counts in a random regular graph Gn,dG_{n,d}, where he also proved a large deviation principle. However, for a particular graph, and certain range of dd (the regularity) the answer for this variational problem is achieved with three possible values of qiq_{i} and not two as in the binomial random graph model.

In a breakthrough paper of Harel, Mousset and Samotij [21], using a combinatorial approach, the authors managed to extend the range where the variational problem (the same one as before) bounds from above the logarithm of the upper tail probability for the count of cliques to the optimal range of pp, which is pr12n1(log(n))1r2p^{\frac{r-1}{2}}\gg n^{-1}(\log(n))^{\frac{1}{r-2}} for a clique on rr vertices222That is because below this threshold they also showed a Poisson approximation.. We wish to emphasize that their approach for solving the upper tail problem is completely different from any previous techniques in the papers mentioned above. This, together with the solution to the variational problem, settled the problem of the first order approximation of the logarithm of the upper tail probability for cliques. Moreover, they established the correct range of pp, up to a polylogarithmic factor, where the variational problem bounds from above the logarithm of the upper tail probability for non-bipartite regular graphs333For bipartite regular graphs their result was not optimal, but also a lot of progress was made..

Building on the combinatorial ideas of Harel, Mousset and Samotij, Basak and Basu [4] extended the range of pp to the optimal range for all regular graphs, including bipartite graphs. Again, this with the solution to the variational problem settled the problem of estimating the logarithm of the upper tail probability for regular graphs. We summarize this discussion with the following first order approximation of the upper tail problem for regular graphs in the ‘localized regime’.

Theorem 1.1 (Due to [21, 4]).

For any Δ2\Delta\geq 2 and a Δ\Delta-regular connected graph HH the following holds:

log(Xn,pH(1+δ)𝔼[Xn,pH])n2pΔlog(1/p)=(1+o(1)){min{θH,12δ2/vH} if 1npΔ/21,12δ2/vH if log1vH2(n)npΔ/21n,\frac{-\log\mathbb{P}\left(X_{n,p}^{H}\geq(1+\delta)\mathbb{E}\left[X_{n,p}^{H}\right]\right)}{n^{2}p^{\Delta}\log(1/p)}=(1+o(1))\begin{cases}\min\{\theta_{H},\frac{1}{2}\delta^{2/v_{H}}\}&\text{ if }\frac{1}{\sqrt{n}}\ll p^{\Delta/2}\ll 1,\\ \frac{1}{2}\delta^{2/v_{H}}&\text{ if }\frac{\log^{\frac{1}{v_{H}-2}}(n)}{n}\ll p^{\Delta/2}\ll\frac{1}{\sqrt{n}},\end{cases}

where θH\theta_{H} is the unique solution to the equation PH(θ)=1+δP_{H}(\theta)=1+\delta where PHP_{H} is the independence polynomial of HH444The independent polynomial of HH is PH(X)=k=0α(H)skXkP_{H}(X)=\sum_{k=0}^{\alpha(H)}s_{k}X^{k} where sks_{k} is the number of independent sets in HH of size kk..

In this paper we estimate the logarithm of the upper tail probability for the count of induced copies of C4C_{4} in Gn,pG_{n,p} in the range log9(n)np1\frac{\log^{9}(n)}{n}\ll p\ll 1, which is optimal up to log9(n)\log^{9}(n) in the lower bound. To the best of our knowledge, this is the first exact result for the upper tail probability of a random variable counting the number of induced copies of a given non-complete graph in Gn,pG_{n,p}. Our proofs rely heavily on the combinatorial approach of Harel, Mousset and Samotij.

We now present the main result of this paper. For this, from now on let XX be the random variable counting the number of induced copies of C4C_{4} in Gn,pG_{n,p}.

{restatable}

thmthebesttheorem There is an explicit sequence 0=c1<c2<1/30=c_{1}<c_{2}<\ldots\leq 1/3 such that the following holds for plog9(n)np\gg\frac{\log^{9}(n)}{n}:

log(X(1+δ)𝔼[X])n2p2log(1/p)=(1+o(1)){ρk(n,p)δ2 if n1+ck1pn1+ck,δ2 if n2/3p1nlog(n),δ2+11281128 if n1/2p1,\frac{-\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])}{n^{2}p^{2}\log(1/p)}=(1+o(1))\begin{cases}\rho_{k}(n,p)\sqrt{\frac{\delta}{2}}&\text{ if }n^{-1+c_{k-1}}\leq p\leq n^{-1+c_{k}},\\ \sqrt{\frac{\delta}{2}}&\text{ if }n^{-2/3}\leq p\ll\frac{1}{\sqrt{n}\log(n)},\\ \sqrt{\frac{\delta}{2}+\frac{1}{{128}}}-\frac{1}{\sqrt{128}}&\text{ if }n^{-1/2}\ll p\ll 1,\end{cases}

where ρk(n,p)=kk1(12k+log(n)klog(p))<1\rho_{k}(n,p)=\sqrt{\frac{k}{k-1}}\left(1-\frac{2}{k}+\frac{\log(n)}{k\log(p)}\right)<1.

The sequence above is explicitly defined in Section 6.

There are several interesting phenomena which distinguish Theorem 1.1 from previous results concerning the upper tail problem for subgraph counts.

The first, and most noticeable difference is the infinite number of phase transitions. To the best of our knowledge, there is no earlier example of an upper tail problem exhibiting infinitely many phase transitions in its first order approximation. To understand the reasons for this phenomenon let us first discuss further Theorem 1.1. A strategy to bound from below the upper tail probability is to observe that, conditioned on the event CGn,pC\subseteq G_{n,p} for some CC satisfying 𝔼[Xn,pHCGn,p](1+δ)𝔼[Xn,pH]\mathbb{E}\left[X_{n,p}^{H}\mid C\subseteq G_{n,p}\right]\geq(1+\delta)\mathbb{E}\left[X_{n,p}^{H}\right], the upper tail event holds with ‘decent’ probability. This allows us to bound the lower tail probability from below by the probability of CGn,pC\subseteq G_{n,p} (which is peCp^{e_{C}}) times the ‘decent’ probability above (which is po(eC)p^{o(e_{C})}). We refer to this as ‘planting a copy of CC’. Of course one would like to consider the smallest such CC, to increase the lower bound as much as possible.

In the previous works mentioned earlier, it was shown that for Δ\Delta-regular graphs there are two natural candidates for CC. These candidates are: a clique on Θ(npΔ/2)\Theta(np^{\Delta/2}) vertices or a spanning complete bipartite graph with the smaller side of size Θ(npΔ)\Theta(np^{\Delta}). The second construction is often called a Hub. When pn1/Δp\ll n^{-1/\Delta}, the small side of the Hub construction needs to be smaller than one and hence at that point the Hub construction is no longer valid, and we are left only with the clique construction as a lower bound for the upper tail probability. This is the reason for the two different regimes in Theorem 1.1.

Even though C4C_{4} is a regular graph, the presence of a clique in Gn,pG_{n,p} does not boost the expected number of induced 4-cycles by a significant amount. Therefore, the first construction is no longer valid. The analog of this construction in our case is a balanced and complete bipartite graph with Θ(np)\Theta(np) vertices. Surprisingly, although amongst all graphs with a given number of edges the complete and balanced bipartite graph maximizes the number of induced copies of C4C_{4} (see [7, 19] and Lemma 6.2) planting it does not always give the strongest lower bound on the upper tail probability.

This is because, in a range of densities pp, there is a large family 𝒦p\mathcal{K}_{p} of subgraphs of KnK_{n}, where each C𝒦pC\in\mathcal{K}_{p} satisfies 𝔼[XCGn,p](1+δ)𝔼[X]\mathbb{E}[X\mid C\subseteq G_{n,p}]\geq(1+\delta)\mathbb{E}[X] and has only slightly suboptimal size (amongst all CC with this property), that is so ‘large’ that

log(CGn,p for some C𝒦p)(1Ω(1))minCe(C)log(1/p),\log\mathbb{P}(C\subseteq G_{n,p}\text{ for some }C\in\mathcal{K}_{p})\geq-(1-\Omega(1))\min_{C}e(C)\log(1/p),

where the minimum ranges over all graphs CC with 𝔼[XCGn,p](1+δ)𝔼[X]\mathbb{E}[X\mid C\subseteq G_{n,p}]\geq(1+\delta)\mathbb{E}[X]; moreover, conditioned on the event that CGn,pC\subseteq G_{n,p} for some C𝒦pC\in\mathcal{K}_{p}, the upper tail event still holds with ‘decent’ probability. Note that this is very different from the case of non-induced regular graphs, as here we condition on the appearance of some graph from 𝒦p\mathcal{K}_{p} rather than a single graph (which corresponds to tilting the measure of Gn,pG_{n,p} to Gn,q¯G_{n,\bar{q}} for some q¯{p,1}(n2)\bar{q}\in\{p,1\}^{\binom{n}{2}}). In fact, we will show (in Section 7) that the lower bound obtained with the use of 𝒦p\mathcal{K}_{p} is stronger than any bound corresponding to the more general tilting Gn,pG_{n,p} to Gn,q¯G_{n,\bar{q}} for some q¯[0,1](n2)\bar{q}\in[0,1]^{\binom{n}{2}}.

Let us elaborate on the structure of the graphs in 𝒦p\mathcal{K}_{p}. The family 𝒦p\mathcal{K}_{p} depends on pp in the following way: provided n1+ck1pn1+ckn^{-1+c_{k-1}}\leq p\leq n^{-1+c_{k}} for some integer k2k\geq 2, the set 𝒦p\mathcal{K}_{p} comprises all complete bipartite graphs with sides kk and =2δ𝔼[X]k(k1)\ell=2\sqrt{\frac{\delta\mathbb{E}[X]}{k(k-1)}}. Note that, for any fixed kk and large nn, the graph Kk,K_{k,\ell} has more edges than the complete and balanced bipartite graph with the same number of induced copies of C4C_{4}. However, the key point here is that we are no longer planting a single copy of this graph. Since \ell is large, there are many embeddings of Kk,K_{k,\ell} in KnK_{n} and the probability that Gn,pG_{n,p} contains some such copy is affected by the combinatorial factor of the number of possible embeddings. It turns out that the lower bound obtained by considering these families is actually tight in the range where log9(n)npn2/3o(1)\frac{\log^{9}(n)}{n}\ll p\ll n^{-2/3-o(1)}. Further, between log9(n)n\frac{\log^{9}(n)}{n} and n2/3o(1)n^{-2/3-o(1)} the family 𝒦p\mathcal{K}_{p} changes infinitely many times depending on pp, which explains the infinite number of phase transitions. This will be explained in more details in Section 7.

Organisation of the paper

In Section 2, we prove a general large deviation principle for nonnegative functions of independents Bernoulli variables which is a version of Theorem 9.1 in [21], which was stated in [21] but was not proven. In Section 3, we continue by connecting the general large deviation principle to our large deviation problem. To this end we define special classes of graphs called core graphs. Then we modify the general result proved in Section 2 using this notion of core graphs.

Section 4 is independent of the others and gives lower bounds for the upper tail probability by planting graphs in the denser regimes and families of graphs in the sparser regimes.

In Section 5, we solve the variational problem presented in Section 3. Furthermore, we make a connection between the number of core graphs with mm edges and the maximum number of vertices a core graph with mm edges might have.

Section 6 uses the results from Section 5 to provide upper bounds for the logarithm of the upper tail probability. This section is divided into three parts. First is the dense regime n1/2p1n^{-1/2}\ll p\ll 1, second is the sparse regime log9(n)np1nlog(n)\frac{\log^{9}(n)}{n}\ll p\ll\frac{1}{\sqrt{n}\log(n)} which we also split into two cases: the dense case in the sparse regime n2/3p1nlog(n)n^{-2/3}\ll p\ll\frac{1}{\sqrt{n}\log(n)} and the sparser cases in the sparse regime n1+ck1pn1+ckn^{-1+c_{k-1}}\ll p\ll n^{-1+c_{k}}. Before the second split, we develop some tools used in both cases.

In Section 7 we solve the naive mean field variational problem, showing that it is different the solution to the variational problem we used.

Appendix A was added for completeness, where we reprove results of [27, 5] in a language of graphs rather than graphons.

Acknowledgments

The author thanks his advisor Wojciech Samotij for his guidance through out the write-up of this paper, as well as helpful and fruitful discussions. The author also thanks Arnon Chor and Dor Elboim for helpful discussions. Lastly, the author thanks Eden Kuperwasser for fruitful discussions and for creating the figure in this paper.

2. Main tools - polynomials in the hypercube

We start with some notations and definitions which can also be found in [21]. We work within the probability space ({0,1}N,Ber(p)N)(\{0,1\}^{N},\operatorname{Ber}({p})^{N}). Suppose I[N]I\subseteq[N] and x{0,1}Nx\in\{0,1\}^{N}. We say that

F(I,x)={y{0,1}N:yn=xn for all nI}F(I,x)=\{y\in\{0,1\}^{N}:y_{n}=x_{n}\text{ for all }n\in I\}

is a subcube centered at (I,x)(I,x) with codimension |I||I| denoted by codim(F)=|I|\operatorname{codim}(F)=|I|. Note that every subcube is centered at some pair (I,x)(I,x). Moreover, if a subcube FF is centered at (I,x)(I,x) and it is also centered at (J,y)(J,y) then I=JI=J. Hence, the codimension is well defined. For every subcube FF centered at (I,x)(I,x) we define the one-supcube and the zero-supcube of FF to be F(1)=F(I1,x)F^{(1)}=F(I_{1},x) and F(0)=F(I0,x)F^{(0)}=F(I_{0},x) where Ii={nI:xn=i}I_{i}=\{n\in I:x_{n}=i\} for i=0,1i=0,1. Moreover, for every subcube FF we say that the one-codimension of it is the codimension of F(1)F^{(1)} and the zero-codimension of it is the codimension of F(0)F^{(0)} denoted by codimi(F)\operatorname{codim}_{i}(F) for i=1i=1 and i=0i=0 respectively. A simple observation is that any subcube FF always satisfies F=F(0)F(1)F=F^{(0)}\cap F^{(1)} and codim(F)=codim(F(0))+codim(F(1))\operatorname{codim}(F)=\operatorname{codim}(F^{(0)})+\operatorname{codim}(F^{(1)}). For every X:{0,1}N0X\colon\{0,1\}^{N}\rightarrow\mathbb{R}_{\geq 0} we define the complexity of XX denoted by comp(X)\operatorname{comp}(X) to be the smallest integer dd for which it is possible to represent XX as a linear combination with nonnegative coefficients of indicator functions of subcubes with codimension at most dd. Note that the complexity is well defined as for every such function XX we can write

X=x{0,1}NX(x)𝟙F([N],x).X=\sum\limits_{x\in\{0,1\}^{N}}X(x)\mathbbm{1}_{F([N],x)}.

Moreover, this shows that for any XX we have comp(X)N\operatorname{comp}(X)\leq N. Assume now that YY is a random variable taking values in {0,1}N\{0,1\}^{N} and that X=X(Y)X=X(Y). Given a subcube F{0,1}NF\subseteq\{0,1\}^{N}, we write 𝔼F[X]=𝔼[XYF]\mathbb{E}_{F}[X]=\mathbb{E}[X\mid Y\in F] for the expectation of XX conditioned on YFY\in F. We further define ΦX:00{}\Phi_{X}\colon\mathbb{R}_{\geq 0}\rightarrow\mathbb{R}_{\geq 0}\cup\{\infty\} by

ΦX(δ)=min{\displaystyle\Phi_{X}(\delta)=\min\{ log(YF):F{0,1}N is a subcube with 𝔼F[X](1+δ)𝔼[X]}.\displaystyle-\log\mathbb{P}(Y\in F):F\subseteq\{0,1\}^{N}\text{ is a subcube with }\mathbb{E}_{F}[X]\geq(1+\delta)\mathbb{E}[X]\}.

Our main tool is [21, Theorem 9.1], which was stated but not proved. We prove a slightly different version of the theorem, but the essence remains the same.

Theorem 2.1.

For every positive integer dd and all positive real numbers ε\varepsilon and δ\delta with ε<12\varepsilon<\frac{1}{2}, there is a positive constant K=K(ε,δ,d)K=K(\varepsilon,\delta,d) such that the following holds. Let YY be a sequence of NN independent Ber(p)\operatorname{Ber}(p) random variables for some p(0,1/2)p\in(0,1/2) and assume that X=X(Y)X=X(Y) is nonnegative with complexity at most dd and satisfies ΦX(δε)Klog(1p)\Phi_{X}(\delta-\varepsilon)\geq K\log(\frac{1}{p}). Denote by \mathcal{F} the collection of all subcubes F{0,1}NF\subseteq\{0,1\}^{N} satisfying

  1. (F1F{1})

    𝔼F[X](1+δε)𝔼[X]\mathbb{E}_{F}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X],

  2. (F2F{2})

    codim(F)KΦX(δ+ε)\operatorname{codim}(F)\leq K\cdot\Phi_{X}(\delta+\varepsilon).

Then,

(X(1+δ)𝔼[X])(1+ε)(YF for some F).\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F}).

To prove the theorem we state and prove the following lemmas.

Lemma 2.2.

Let YY be a random variable taking values in {0,1}N\{0,1\}^{N} and let X=X(Y)X=X(Y) be a real-valued function of YY. Suppose that 𝔼[X]>0\mathbb{E}[X]>0 and that XMX\leq M always. Then for all positive ε\varepsilon and δ\delta,

log(X(1+δ)𝔼[X])ΦX(δ+ε)+log(Mε𝔼[X])-\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq\Phi_{X}(\delta+\varepsilon)+\log\bigg{(}\frac{M}{\varepsilon\mathbb{E}[X]}\bigg{)}
Proof.

Let t=(1+δ)𝔼[X]t=(1+\delta)\mathbb{E}[X]. If ΦX(δ+ε)=\Phi_{X}(\delta+\varepsilon)=\infty then the claim is vacuously true. Otherwise, there exists a subcube FF such that log(YF)=ΦX(δ+ε)-\log\mathbb{P}(Y\in F)=\Phi_{X}(\delta+\varepsilon) and 𝔼F[X](1+δ+ε)𝔼[X]=t+ε𝔼[X]\mathbb{E}_{F}[X]\geq(1+\delta+\varepsilon)\mathbb{E}[X]=t+\varepsilon\mathbb{E}[X]. By XMX\leq M we have 𝔼F[X]M(XtYF)+t\mathbb{E}_{F}[X]\leq M\cdot\mathbb{P}(X\geq t\mid Y\in F)+t. Thus, (XtYF)ε𝔼[X]M\mathbb{P}(X\geq t\mid Y\in F)\geq\frac{\varepsilon\mathbb{E}[X]}{M} and hence,

(Xt)(XtYF)(YF)ε𝔼[X]M(YF).\mathbb{P}(X\geq t)\geq\mathbb{P}(X\geq t\mid Y\in F)\mathbb{P}(Y\in F)\geq\frac{\varepsilon\mathbb{E}[X]}{M}\mathbb{P}(Y\in F).

Now taking the negative logarithm gives the assertion of the lemma. ∎

Fact 2.3.

Suppose F1,,FkF_{1},\ldots,F_{k} are subcubes of {0,1}N\{0,1\}^{N}. If F1FkF_{1}\cap\cdots\cap F_{k} is nonempty, then it is also a subcube of {0,1}N\{0,1\}^{N} and, moreover, codim(F1Fk)i=1kcodim(Fi)\operatorname{codim}(F_{1}\cap\cdots\cap F_{k})\leq\sum_{i=1}^{k}\operatorname{codim}(F_{i}).

Proof.

We prove the statement for k=2k=2; the case k>2k>2 follows by a simple inductive argument. Let I1,I2I_{1},I_{2} and x1,x2x_{1},x_{2} be such that FiF_{i} is a subcube centered at (Ii,xi)(I_{i},x_{i}) for i=1,2i=1,2. Define x{0,1}Nx\in\{0,1\}^{N} in the following way: for all iI1i\in I_{1} and jI2j\in I_{2} put (x)i=(x1)i(x)_{i}=(x_{1})_{i} and (x)j=(x2)j(x)_{j}=(x_{2})_{j} and for all iI1I2i\not\in I_{1}\cup I_{2} put (x)i=0(x)_{i}=0. This is indeed a well defined element in {0,1}N\{0,1\}^{N} as for all iI1I2i\in I_{1}\cap I_{2} we have, (x1)i=(x2)i(x_{1})_{i}=(x_{2})_{i}, otherwise it will contradict our assumption that F1F2F_{1}\cap F_{2}\neq\emptyset. It follows from the definition of xx that F1F2=F(I1I2,x)F_{1}\cap F_{2}=F(I_{1}\cup I_{2},x) and hence F1F2F_{1}\cap F_{2} is a subcube. Furthermore, we have

codim(F1F2)=|I1I2||I1|+|I2|=codim(F1)+codim(F2).\operatorname{codim}(F_{1}\cap F_{2})=|I_{1}\cup I_{2}|\leq|I_{1}|+|I_{2}|=\operatorname{codim}(F_{1})+\operatorname{codim}(F_{2}).\qed
Lemma 2.4.

Let YY be a random variable taking values in {0,1}N\{0,1\}^{N} and let X=X(Y)X=X(Y) be a nonnegative real-valued function with complexity bounded by dd. Then for every positive integer \ell and all positive real numbers ε\varepsilon and δ\delta with ε<1+δ\varepsilon<1+\delta,

(X(1+δ)𝔼[X] and YF for all F)(1+δε1+δ),\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]\text{ and }Y\not\in F\text{ for all }F\in\mathcal{F})\leq{\bigg{(}\frac{1+\delta-\varepsilon}{1+\delta}\bigg{)}}^{\ell},

where

={F{0,1}N:F is a subcube and codim(F)d and 𝔼F[X](1+δε)𝔼[X]}.\mathcal{F}=\{F\in\{0,1\}^{N}:F\text{ is a subcube and }\operatorname{codim}(F)\leq d\ell\text{ and }\mathbb{E}_{F}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X]\}.
Proof.

Given S{0,1}NS\subseteq\{0,1\}^{N} a subcube, let ZSZ_{S} be the indicator random variable of the event that YFY\not\in F for all FF\in\mathcal{F} with FSF\supseteq S. Note that SSS\subseteq S^{\prime} implies ZSZSZ_{S}\leq Z_{S^{\prime}} and let Z=ZZ=Z_{\emptyset}. Since XZ0XZ\geq 0 and Z=ZZ^{\ell}=Z, Markov’s inequality gives

(X(1+δ)𝔼[X] and Z=1)=(XZ(1+δ)𝔼[X])𝔼[XZ]((1+δ)𝔼[X]).\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]\text{ and }Z=1)=\mathbb{P}(XZ\geq(1+\delta)\mathbb{E}[X])\leq\frac{\mathbb{E}[X^{\ell}Z]}{((1+\delta)\mathbb{E}[X])^{\ell}}. (2)

To simplify the notation, for every S{0,1}NS\subseteq\{0,1\}^{N} we write 𝟙S\mathbbm{1}_{S} for 𝟙{YS}\mathbbm{1}_{\{Y\in S\}}. Write X=FαF𝟙FX=\sum_{F}\alpha_{F}\mathbbm{1}_{F}, where the sum ranges over all subcubes of {0,1}N\{0,1\}^{N}, each coefficient αF\alpha_{F} is nonnegative, and αF=0\alpha_{F}=0 for all FF with codim(F)>dcodim(F)>d. Then for every k[]k\in[\ell],

𝔼[XkZ]=\displaystyle\mathbb{E}[X^{k}Z]= F1,,FkαF1αFk𝔼[𝟙F1𝟙FkZ]\displaystyle\sum\limits_{F_{1},\ldots,F_{k}}\alpha_{F_{1}}\cdots\alpha_{F_{k}}\mathbb{E}[\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k}}\cdot Z]
\displaystyle\leq F1,,FkαF1αFk𝔼[𝟙F1𝟙FkZF1Fk]\displaystyle\sum\limits_{F_{1},\ldots,F_{k}}\alpha_{F_{1}}\cdots\alpha_{F_{k}}\mathbb{E}[\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k}}\cdot Z_{F_{1}\cap\cdots\cap F_{k}}]
\displaystyle\leq F1,,Fk1αF1αFk1𝔼[𝟙F1𝟙Fk1ZF1Fk1]\displaystyle\sum\limits_{F_{1},\ldots,F_{k-1}}\alpha_{F_{1}}\cdots\alpha_{F_{k-1}}\mathbb{E}[\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}]
𝔼[X𝟙F1𝟙Fk1ZF1Fk1=1],\displaystyle\qquad\qquad\qquad\qquad\;\;\cdot\mathbb{E}[X\mid\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}=1],

where we may let the third sum range only over sequences F1,,Fk1F_{1},\ldots,F_{k-1} for which the event {𝟙F1𝟙Fk1ZF1Fk1=1}\{\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}=1\} has a positive probability of occurring.

Claim 2.5.

For any such sequence,

F1Fk1 and  1F1𝟙Fk1ZF1Fk1=𝟙F1𝟙Fk1.F_{1}\cap\ldots\cap F_{k-1}\not\in\mathcal{F}\;\text{ and }\;\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}=\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}.
Proof.

To see this, note that if {𝟙F1𝟙Fk1ZF1Fk1=1}\{\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}=1\} has positive probability of occurring, then there exists a yy such that yF1Fk1y\in F_{1}\cap\ldots\cap F_{k-1} and ZF1Fk1(y)=1Z_{F_{1}\cap\ldots\cap F_{k-1}}(y)=1. That means yF1Fk1y\in F_{1}\cap\ldots\cap F_{k-1} and yFy\not\in F for any subcube FF\in\mathcal{F} such that FF1Fk1F\supseteq F_{1}\cap\ldots\cap F_{k-1}. Hence, F1Fk1F_{1}\cap\ldots\cap F_{k-1}\not\in\mathcal{F}. For the second part assume towards a contradiction that 𝟙F1𝟙Fk1ZF1Fk1<𝟙F1𝟙Fk1\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}<\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}, meaning that there is yF1Fk1y^{\prime}\in F_{1}\cap\ldots\cap F_{k-1} such that ZF1Fk1(y)=0Z_{F_{1}\cap\ldots\cap F_{k-1}}(y^{\prime})=0. Therefore, there exists a subcube FF^{*}\in\mathcal{F} such that, FF1Fk1F^{*}\supseteq F_{1}\cap\ldots\cap F_{k-1}. This is a contradiction as now, yF1Fk1Fy\in F_{1}\cap\ldots\cap F_{k-1}\subseteq F^{*} but by our assumption, yFy\not\in F^{*} as FF^{*}\in\mathcal{F}. ∎

Since {𝟙F1Fk1=1}\{\mathbbm{1}_{F_{1}\cap\cdots\cap F_{k-1}}=1\} has a positive probability of occurring, Fact 2.3 asserts that F1Fk1F_{1}\cap\cdots\cap F_{k-1} is a subcube, and codim(F1Fk1)d(k1)d\operatorname{codim}({F_{1}\cap\cdots\cap F_{k-1}})\leq d(k-1)\leq d\ell. Therefore

𝔼[X𝟙F1𝟙Fk1ZF1Fk1=1]=𝔼F1Fk1[X]<(1+δε)𝔼[X],\mathbb{E}[X\mid\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}=1]=\mathbb{E}_{F_{1}\cap\cdots\cap F_{k-1}}[X]<(1+\delta-\varepsilon)\mathbb{E}[X],

as otherwise F1Fk1F_{1}\cap\cdots\cap F_{k-1} would belong to \mathcal{F}. It follows that

F1,,FkαF1αFk𝔼[𝟙F1𝟙FkZF1Fk]\displaystyle\sum\limits_{F_{1},\ldots,F_{k}}\alpha_{F_{1}}\cdots\alpha_{F_{k}}\mathbb{E}[\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k}}\cdot Z_{F_{1}\cap\cdots\cap F_{k}}]
<(1+δε)𝔼[X]F1,,Fk1\displaystyle<(1+\delta-\varepsilon)\mathbb{E}[X]\sum\limits_{F_{1},\ldots,F_{k-1}} αF1αFk1𝔼[𝟙F1𝟙Fk1ZF1Fk1].\displaystyle\alpha_{F_{1}}\cdots\alpha_{F_{k-1}}\mathbb{E}[\mathbbm{1}_{F_{1}}\cdots\mathbbm{1}_{F_{k-1}}\cdot Z_{F_{1}\cap\cdots\cap F_{k-1}}].

By induction, we see that 𝔼[XZ]<((1+δε)𝔼[X])\mathbb{E}[X^{\ell}Z]<((1+\delta-\varepsilon)\mathbb{E}[X])^{\ell}. Substituting this inequality into (2) completes the proof. ∎

Now we use the previous lemmas to prove the theorem.

Proof of Theorem 2.1.

Let KK be a large constant that may depend on ε,δ\varepsilon,\delta and dd. Furthermore, let t=(1+δ)𝔼[X]t=(1+\delta)\mathbb{E}[X] and =K/dΦX(δ+ε)\ell=\left\lfloor{K/d\cdot\Phi_{X}(\delta+\varepsilon)}\right\rfloor. Define

={F{0,1}N:F is a subcube with codim(F)d and 𝔼F[X](1+δε)𝔼[X]}.\mathcal{F}=\{F\subseteq\{0,1\}^{N}:F\text{ is a subcube with }\operatorname{codim}(F)\leq d\ell\text{ and }\mathbb{E}_{F}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X]\}.

It follows from Lemma 2.4 that

(Xt and YF for all F)(1ε1+δ).\mathbb{P}(X\geq t\text{ and }Y\not\in F\text{ for all }F\in\mathcal{F})\leq\bigg{(}1-\frac{\varepsilon}{1+\delta}\bigg{)}^{\ell}.

As comp(X)d\operatorname{comp}(X)\leq d, we can write

X=jJαj𝟙{YFj}X=\sum\limits_{j\in J}\alpha_{j}\mathbbm{1}_{\{Y\in F_{j}\}}

for some JJ so that, for all jJj\in J, the coefficient αj\alpha_{j} is nonnegative and FjF_{j} is a subcube with codim(Fj)d\operatorname{codim}(F_{j})\leq d. Put M=αiM=\sum\alpha_{i} and note that XMX\leq M always. Applying Lemma 2.2 gives

log(Xt)ΦX(δ+ε)+log(Mε𝔼[X]).\displaystyle-\log\mathbb{P}(X\geq t)\leq\Phi_{X}(\delta+\varepsilon)+\log\bigg{(}\frac{M}{\varepsilon\mathbb{E}[X]}\bigg{)}.

Note also that 𝔼[X]Mpd\mathbb{E}[X]\geq Mp^{d} as comp(X)d\operatorname{comp}(X)\leq d and therefore (YFj)min{p,1p}d=pd\mathbb{P}(Y\in F_{j})\geq\min\{p,1-p\}^{d}=p^{d} for every jJj\in J. Therefore,

log(Xt)ΦX(δ+ε)+log(1εpd).\displaystyle-\log\mathbb{P}(X\geq t)\leq\Phi_{X}(\delta+\varepsilon)+\log\bigg{(}\frac{1}{\varepsilon p^{d}}\bigg{)}.

Thus, provided KK is sufficiently large we also have

log(Xt)(1+ε)ΦX(δ+ε),\displaystyle-\log\mathbb{P}(X\geq t)\leq(1+\varepsilon)\Phi_{X}(\delta+\varepsilon),

as we assumed that ΦX(δ+ε)ΦX(δε)Klog(1/p)\Phi_{X}(\delta+\varepsilon)\geq\Phi_{X}(\delta-\varepsilon)\geq K\log(1/p). Putting all of this together gives that for sufficiently large KK, we have

(Xt and YF for all F)\displaystyle\mathbb{P}(X\geq t\text{ and }Y\not\in F\text{ for all }F\in\mathcal{F}) (1ε1+δ)eε1+δK/dΦX(δ+ε)\displaystyle\leq\bigg{(}1-\frac{\varepsilon}{1+\delta}\bigg{)}^{\ell}\leq e^{-\frac{\varepsilon}{1+\delta}\left\lfloor{K/d\cdot\Phi_{X}(\delta+\varepsilon)}\right\rfloor}
(ε/2)(Xt).\displaystyle\leq(\varepsilon/2)\cdot\mathbb{P}(X\geq t).

The assertion of the theorem now follows:

(Xt)(1ε/2)1(YF for some F)(1+ε)(YF for some F)\mathbb{P}(X\geq t)\leq{(1-\varepsilon/2)}^{-1}\cdot\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F})\leq(1+\varepsilon)\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F})

where the last inequality holds for ε<1/2\varepsilon<1/2. ∎

When p1p\ll 1 and comp(X)=O(1)\operatorname{comp}(X)=O(1), the majority of the contribution to both 𝔼F[X]\mathbb{E}_{F}[X] and (YF)\mathbb{P}(Y\in F) comes from the one-supcube F(1)FF^{(1)}\supseteq F. We prove a straightforward corollary of Theorem 2.1 which will be more convenient to work with. We start by proving the following lemma.

Lemma 2.6.

The following holds for every positive integer dd and all positive real numbers ε\varepsilon and δ\delta with ε<1\varepsilon<1. Let YY be a sequence of NN independent Ber(p)\operatorname{Ber}(p) random variables for some p<1(1+δε1+δε/2)1/dp<1-\big{(}{\frac{1+\delta-\varepsilon}{1+\delta-\varepsilon/2}\big{)}^{1/d}} and assume that X=X(Y)X=X(Y) has complexity at most dd. Let FF be a subcube satisfying 𝔼F[X](1+δε/2)𝔼[X]\mathbb{E}_{F}[X]\geq(1+\delta-\varepsilon/2)\mathbb{E}[X]. Then, 𝔼F(1)[X](1+δε)𝔼[X]\mathbb{E}_{F^{(1)}}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X] where F(1)F^{(1)} is the one-supcube of FF.

Proof.

As XX has complexity dd we can write

X=jJαj𝟙{YFj}X=\sum\limits_{j\in J}\alpha_{j}\mathbbm{1}_{\{Y\in F_{j}\}}

for some JJ so that, for all jJj\in J, the coefficient αj\alpha_{j} is nonnegative and FjF_{j} is a subcube with codim(Fj)d\operatorname{codim}(F_{j})\leq d. Moreover, by our assumptions on FjF_{j}, we have codim0(Fj)codim(Fj)d\operatorname{codim}_{0}(F_{j})\leq\operatorname{codim}(F_{j})\leq d. Thus, we have the following inequality,

(YFjYF(1))(1p)d(YFjYF).\mathbb{P}(Y\in F_{j}\mid Y\in F^{(1)})\geq(1-p)^{d}\cdot\mathbb{P}(Y\in F_{j}\mid Y\in F).

It follows that

𝔼F(1)[X]\displaystyle\mathbb{E}_{F^{(1)}}[X] =jJαj𝔼F(1)[𝟙{YFj}]jJαj(1p)d𝔼F[𝟙{YFj}]\displaystyle=\sum\limits_{j\in J}\alpha_{j}\mathbb{E}_{F^{(1)}}[\mathbbm{1}_{\{Y\in F_{j}\}}]\geq\sum\limits_{j\in J}\alpha_{j}(1-p)^{d}\cdot\mathbb{E}_{F}[\mathbbm{1}_{\{Y\in F_{j}\}}]
jJαj(1+δε1+δε/2)𝔼F[𝟙Fj]=(1+δε1+δε/2)𝔼F[X](1+δε)𝔼[X],\displaystyle\geq\sum\limits_{j\in J}\alpha_{j}\Big{(}\frac{1+\delta-\varepsilon}{1+\delta-\varepsilon/2}\Big{)}\mathbb{E}_{F}[\mathbbm{1}_{F_{j}}]=\Big{(}\frac{1+\delta-\varepsilon}{1+\delta-\varepsilon/2}\Big{)}\mathbb{E}_{F}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X], (3)

where the first inequality in (3) follows from the assumption on pp and the second inequality in (3) follows from the assumption on 𝔼F[X]\mathbb{E}_{F}[X]. ∎

Using this lemma we now state and prove a straightforward corollary of Theorem 2.1.

Corollary 2.7.

For every positive integer dd and all positive real numbers ε\varepsilon and δ\delta with ε<1\varepsilon<{1}, there is a positive constant K=K(ε,δ,d)K=K(\varepsilon,\delta,d) such that the following holds. Let YY be a sequence of NN independent Ber(p)\operatorname{Ber}(p) random variables for some p<min{1(1+δε1+δε/2)1/d,1/2}p<\min\{1-\big{(}{\frac{1+\delta-\varepsilon}{1+\delta-\varepsilon/2}\big{)}^{1/d}},1/2\} and assume that X=X(Y)X=X(Y) has complexity at most dd and satisfies ΦX(δε)Klog(1p)\Phi_{X}(\delta-\varepsilon)\geq K\log(\frac{1}{p}). Denote by 1\mathcal{F}_{1} the collection of all subcubes F{0,1}NF\subseteq\{0,1\}^{N} satisfying

  1. (H1H{1})

    𝔼F[X](1+δε)𝔼[X]\mathbb{E}_{F}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X],

  2. (H2H{2})

    codimFKΦX(δ+ε)\operatorname{codim}F\leq K\cdot\Phi_{X}(\delta+\varepsilon),

  3. (H3H{3})

    codim1(F)=codim(F)\operatorname{codim}_{1}(F)=\operatorname{codim}(F).

Then

(X(1+δ)𝔼[X])(1+ε)(YF for some F1).\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F}_{1}).
Proof.

Applying Theorem 2.1 with ε\varepsilon replaced by ε/2\varepsilon/2 we obtain

(X(1+δ)𝔼[X])(1+ε)(YF for some F),\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F}), (4)

where \mathcal{F} is the collection of all subcubes F{0,1}NF\subseteq\{0,1\}^{N} satisfying (F1F{1}) and (F2F{2}) where ε\varepsilon is replaced with ε/2\varepsilon/2. Letting 1={F(1):F}\mathcal{F}_{1}=\{F^{(1)}:F\in\mathcal{F}\}, we obtain (H2H{2}) and (H3H{3}) for every subcube F1F\in\mathcal{F}_{1} due to (F2F{2}) and the definition of one-supcubes. Noting that for every subcube FF we have FF(1)F\subseteq F^{(1)} we obtain also that

(YF for some F)(YF for some F1).\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F})\leq\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F}_{1}). (5)

Combining (4) and (5) we obtain that

(X(1+δ)𝔼[X])(1+ε)(YF for some F1).\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\mathbb{P}(Y\in F\text{ for some }F\in\mathcal{F}_{1}).

Furthermore, Lemma 2.6 implies that 1\mathcal{F}_{1} also satisfies (H1H{1}). ∎

We call a subcube FF a seed if FF is an element of 1\mathcal{F}_{1} defined in Corollary 2.7. A formal definition is given in the following section. A general way one would use Corollary 2.7 to bound upper tails of counts of induced subgraphs is to define a special type of seeds called structured seeds. These structured seeds are subcubes in 1\mathcal{F}_{1} from Corollary 2.7, with a stronger condition than (H1H{1}). This condition is a combinatorial condition involving various supcubes of the subcubes in 1\mathcal{F}_{1} (this will be explained further in Section 3). Then one would define a core subcube to be a subcube containing a structured seed such that ‘every coordinate counts’. In the following section we define these special subcubes. We will also show that Corollary 2.7 can be modified to bound the upper tail probability via the probability of YY being an element in a core subcube.

3. From seeds to modified cores

Note that for any subcube F{0,1}E(Kn)F\subseteq\{0,1\}^{E(K_{n})}, F(1)F^{(1)} the one-supcube of FF corresponds to all subgraphs of KnK_{n} containing a specific subgraph of KnK_{n} with codim1(F)\operatorname{codim}_{1}({F}) edges. Therefore, from this point onwards we think of one-supcubes as a family of subgraphs of KnK_{n} containing a specific subgraph. In particular, instead of writing 𝔼F[X]\mathbb{E}_{F}[X] for some one-supcube (of some subcube) we will write 𝔼G[X]\mathbb{E}_{G}[X] for the graph corresponding to FF. The subgraphs corresponding to the members of 1\mathcal{F}_{1} from Corollary 2.7 will be called seeds. We start with a definition which will be useful in this section as well as the next ones.

Definition.

Suppose HH and GG are graphs and let ee be an edge of GG. We define:

  • Nind(H,G)N_{ind}(H,G) is the number of induced copies of HH in GG.

  • Nind(e,H,G)N_{ind}(e,H,G) is the number of induced copies of HH in GG that contain ee.

Our main concern in this paper is the random variable counting the number of induced copies of C4C_{4} in Gn,pG_{n,p}. Therefore, we let X=Nind(C4,Gn,p)X=N_{ind}(C_{4},G_{n,p}). We will use this notation from this point onward. We now continue by defining seed graphs.

Definition.

Let ε,δ,K\varepsilon,\delta,K be positive reals. In addition let p(0,1)p\in(0,1). Then we define 𝒮(ε,δ,K)\mathcal{S}(\varepsilon,\delta,K) to be the collection of all spanning subgraphs GKnG\subseteq K_{n} satisfying:

  1. (S1S{1})

    e(G)KΦX(δ+ε)e(G)\leq K\cdot\Phi_{X}(\delta+\varepsilon) and

  2. (S2S{2})

    𝔼G[X](1+δε)𝔼[X]\mathbb{E}_{G}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X].

We call the graphs in 𝒮(ε,δ,K)\mathcal{S}(\varepsilon,\delta,K) seeds. Furthermore, for every positive integer mm we define 𝒮m(ε,δ,K)\mathcal{S}_{m}(\varepsilon,\delta,K) to be the set of all seeds with mm edges.

Since 𝔼G[X]\mathbb{E}_{G}[X] is determined by the number of induced copies of various subgraphs of C4C_{4} in GG, it will be convenient to restate (S2S{2}) in the above definition with a graph theoretic condition, giving rise to the notion of structured seeds. In the definition we use the graph obtained from K1,2K_{1,2} by adding an isolated vertex; we denote this graph by K1,2K1K_{1,2}\sqcup K_{1}. Now we define the structured seeds.

Definition.

Let ε,δ,K\varepsilon,\delta,K be positive reals. In addition let p(0,1)p\in(0,1). Then we define 𝒮(ε,δ,K)\mathcal{S}^{\prime}(\varepsilon,\delta,K) to be the collection of all spanning subgraphs GKnG\subseteq K_{n} satisfying:

  1. (S1S^{\prime}{1})

    e(G)KΦX(δ+ε)e(G)\leq K\cdot\Phi_{X}(\delta+\varepsilon) and

  2. (S2S^{\prime}{2})

    Nind(C4,G)+Nind(K1,2K1,G)p2(δε)𝔼[X]N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}\geq(\delta-\varepsilon)\mathbb{E}[X].

We call the graphs in 𝒮(ε,δ,K)\mathcal{S}^{\prime}(\varepsilon,\delta,K) structured seeds. Furthermore, for every positive integer mm we define 𝒮m(ε,δ,K)\mathcal{S}^{\prime}_{m}(\varepsilon,\delta,K) to be the set of all structured seeds with mm edges.

The following is a lemma relating the seeds and the structured seeds, which we prove later.

Lemma 3.1.

Let ε,δ,K\varepsilon,\delta,K be positive reals and suppose also that εδ/2\varepsilon\leq\delta/2 and p1p\ll 1. Then, there exists a positive constant C=C(ε,δ,K)C=C(\varepsilon,\delta,K) such that for any mKΦX(δε)m\leq K\Phi_{X}(\delta-\varepsilon) and large enough nn we have,

𝒮m(ε,δ,C)𝒮m(2ε,δ,C).\mathcal{S}_{m}(\varepsilon,\delta,C)\subseteq\mathcal{S}^{\prime}_{m}(2\varepsilon,\delta,C).

Now we are ready to define what a core graph is. This is given in the following definition.

Definition.

Let ε,δ,K\varepsilon,\delta,K be positive reals. In addition let p(0,1)p\in(0,1). Then we define 𝒞(ε,δ,K)\mathcal{C}(\varepsilon,\delta,K) to be collection of all structured seeds G𝒮(ε,δ,K)G\in\mathcal{S}^{\prime}(\varepsilon,\delta,K) satisfying the following:
For all eE(G)e\in E(G)

Nind(e,C4,G)+Nind(e,K1,2K1,G)p2ε𝔼[X]/(2KΦX(δ+ε)).N_{ind}(e,C_{4},G)+N_{ind}(e,K_{1,2}\sqcup K_{1},G)p^{2}\geq\varepsilon\mathbb{E}[X]/(2K\cdot\Phi_{X}(\delta+\varepsilon)).

We call the graphs in 𝒞(ε,δ,K)\mathcal{C}(\varepsilon,\delta,K) core graphs. Furthermore, for every positive integer mm we define 𝒞m(ε,δ,K)\mathcal{C}_{m}(\varepsilon,\delta,K) to be the set of all cores with mm edges.

In cases where the parameters ε,δ\varepsilon,\delta and KK can be understood from the context we omit them. The main aim of this section is to prove the following theorem which we derive using Corollary 2.7 and several lemmas.

Theorem 3.2.

For all positive real numbers ε,δ,p\varepsilon,\delta,p with ε<δ\varepsilon<{\delta} and p<min{1(1+δε1+δε/2)1/6,1/2}p<\min\{1-\big{(}{\frac{1+\delta-\varepsilon}{1+\delta-\varepsilon/2}\big{)}^{1/6}},1/2\}, there is a positive constant K=K(ε,δ)K=K(\varepsilon,\delta) such that following holds:

(X(1+δ)𝔼[X])(1+ε)(GGn,p for some G𝒞(ε,δ,K)).\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\mathbb{P}(G\subseteq G_{n,p}\text{ for some }G\in\mathcal{C}(\varepsilon,\delta,K)). (6)

In particular,

(X(1+δ)𝔼[X])(1+ε)m|𝒞m(ε,δ,K)|pm.\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\sum\limits_{m}|\mathcal{C}_{m}(\varepsilon,\delta,K)|p^{m}.

We start by proving Lemma 3.1 relating seeds and structured seeds.

Proof of Lemma 3.1.

Suppose G𝒮m(ε,δ,K)G\in\mathcal{S}_{m}(\varepsilon,\delta,K). By (S2S{2}) we have

(δε)𝔼[X]𝔼G[X]𝔼[X]HC4Nind(H,G)p4e(H),\displaystyle(\delta-\varepsilon)\mathbb{E}[X]\leq\mathbb{E}_{G}[X]-\mathbb{E}[X]\leq\sum\limits_{\emptyset\neq H{\subseteq}^{*}C_{4}}N_{ind}(H,G)p^{4-e(H)},

where {\subseteq}^{*} stands for spanning subgraphs. Observe that

Nind(H,G){mn2 if H=K22K1m2 if H=M2 or P4,N_{ind}(H,G)\leq\begin{cases}mn^{2}&\text{ if }H=K_{2}\sqcup 2K_{1}\\ m^{2}&\text{ if }H=M_{2}\text{ or }P_{4},\end{cases}

where K22K1K_{2}\sqcup 2K_{1} is the graph on four vertices and one edge, M2M_{2} is a matching of size two, and P4P_{4} is the path with 3 edges; this indeed holds as two edges of GG span at most one induced M2M_{2} or P4P_{4}. Therefore, we obtain

(δε)𝔼[X]\displaystyle(\delta-\varepsilon)\mathbb{E}[X]\leq Nind(C4,G)+Nind(K1,2K1,G)p2+m2pNind(P4,G)+m2p2Nind(M2,G)+mn2p3Nind(K22K1,G)\displaystyle N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}+\underbrace{m^{2}p}_{N_{ind}(P_{4},G)}+\underbrace{m^{2}p^{2}}_{N_{ind}(M_{2},G)}+\underbrace{mn^{2}p^{3}}_{N_{ind}(K_{2}\sqcup 2K_{1},G)}
\displaystyle\leq Nind(C4,G)+Nind(K1,2K1,G)p2+2pm2+mn2p3.\displaystyle N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}+2pm^{2}+mn^{2}p^{3}. (7)

In Section 5 (Claim 5.1) we prove that ΦX(δ)=Oδ(𝔼[X]log(1/p))\Phi_{X}(\delta)=O_{\delta}(\sqrt{\mathbb{E}[X]}\log(1/p)). Therefore, as 𝔼[X]=Θ(n4p4)\mathbb{E}[X]=\Theta(n^{4}p^{4}), we have m=Oε,δ(n2p2log(1/p))m=O_{\varepsilon,\delta}(n^{2}p^{2}\log(1/p)). Further, as p1p\ll 1 we get from (3) the following inequality for large enough nn,

(δε)𝔼[X]\displaystyle(\delta-\varepsilon)\mathbb{E}[X] Nind(C4,G)+Nind(K1,2K1,G)p2+O(plog2(1/p)𝔼[X])\displaystyle\leq N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}+O(p\log^{2}(1/p)\mathbb{E}[X])
Nind(C4,G)+Nind(K1,2K1,G)p2+ε𝔼[X].\displaystyle\leq N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}+\varepsilon\mathbb{E}[X].

Therefore, we obtain Nind(C4,G)+Nind(K1,2K1,G)p2(δ2ε)𝔼[X]N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}\geq(\delta-2\varepsilon)\mathbb{E}[X] for large enough nn. This is the assertion of the lemma. ∎

Informally, the next claim is that every structured seed contains a core. Formally this is given in the following claim.

Lemma 3.3.

Suppose ε,δ,K\varepsilon,\delta,K are positive reals. Then for every structured seed G𝒮G\in\mathcal{S}^{\prime} and every nonnegative real ss, there exists a subgraph GGG^{*}\subseteq G such that:

  1. (C1C{1})

    e(G)KΦX(δ+ε)e(G)\leq K\cdot\Phi_{X}(\delta+\varepsilon),

  2. (C2C{2})

    Nind(C4,G)+Nind(K1,2K1,G)p2(δε)𝔼[X]sN_{ind}(C_{4},G^{*})+N_{ind}(K_{1,2}\sqcup K_{1},G^{*})p^{2}\geq(\delta-\varepsilon)\mathbb{E}[X]-s, and

  3. (C3C{3})

    Nind(e,C4,G)+Nind(e,K1,2K1,G)p2sKΦX(δ+ε)N_{ind}(e,C_{4},G^{*})+N_{ind}(e,K_{1,2}\sqcup K_{1},G^{*})p^{2}\geq\frac{s}{K\Phi_{X}(\delta+\varepsilon)} for every edge eE(G)e\in E(G^{*}).

Proof.

In this proof, it would be convenient to let

N(G)=Nind(C4,G)+Nind(K1,2K1,G)p2.N(G)=N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}.

This is because then, (C2C{2}) is equivalent to N(G)(δε)𝔼[X]sN(G^{*})\geq(\delta-\varepsilon)\mathbb{E}[X]-s and (C3C{3}) is equivalent to N(G)N(Ge)sKΦX(δ+ε)N(G^{*})-N(G^{*}\setminus e)\geq\frac{s}{K\Phi_{X}(\delta+\varepsilon)} for every ee an edge in GG^{*}.

Define the sequences G=G0G1Gr=GG=G_{0}\supseteq G_{1}\supseteq\cdots\supseteq G_{r}=G^{*} and e1,e2,,erGe_{1},e_{2},\ldots,e_{r}\in G by repeatedly setting Gk+1G_{k+1} to be a subgraph of GkG_{k} obtained by the deletion of an edge eke_{k} such that

N(Gk)N(Gk+1)<se(G),N(G_{k})-N(G_{k+1})<\frac{s}{e(G)},

as long as such edge exists. The graph GG^{*} satisfies (C1C{1}) as it is a subgraph of a structured seed. We also claim that, the subgraph GG^{*} satisfies (C3C{3}). That is because, if there is an edge ee in GG^{*} with N(G)N(Ge)<se(G)N(G^{*})-N(G^{*}\setminus e)<\frac{s}{e(G)} the process would have continued by deleting this edge - a contradiction. Recalling that GG is a structured seed we find that, e(G)KΦX(δ+ε)e(G)\leq K\Phi_{X}(\delta+\varepsilon) and thus,

N(G)se(G)sKΦX(δ+ε).N(G^{*})\geq\frac{s}{e(G)}\geq\frac{s}{K\Phi_{X}(\delta+\varepsilon)}.

Finally, since re(G)r\leq e(G), we have

N(G)N(G)=k=0r1N(Gk)N(Gk+1)rse(G)s.\displaystyle N(G)-N(G^{*})=\sum\limits_{k=0}^{r-1}N(G_{k})-N(G_{k+1})\leq\frac{rs}{e(G)}\leq s.

Rearranging this inequality and recalling that GG is a structured seed we obtain the assertion of the lemma,

N(G)=Nind(C4,G)+Nind(K1,2K1,G)p2(δε)𝔼[X]s.N(G^{*})=N_{ind}(C_{4},G^{*})+N_{ind}(K_{1,2}\sqcup K_{1},G^{*})p^{2}\geq(\delta-\varepsilon)\mathbb{E}[X]-s.\qed

Applying Lemma 3.3 invoked with ε\varepsilon replaced by ε/2\varepsilon/2 and s=ε𝔼[X]/2s=\varepsilon\mathbb{E}[X]/2 yields the following corollary.

Corollary 3.4.

Suppose ε,δ,K\varepsilon,\delta,K are positive reals. Suppose further that G𝒮(ε/2,δ,K)G\in\mathcal{S}^{\prime}(\varepsilon/2,\delta,K) is a structured seed. Then, there exists a core G𝒞(ε,δ,K)G^{*}\in\mathcal{C}(\varepsilon,\delta,K) such that GGG^{*}\subseteq G.

Now we derive Theorem 3.2 from the above claims and Corollary 2.7.

Proof of Theorem 3.2.

Applying Lemma 3.1 with ε\varepsilon replaced by ε/4\varepsilon/4 and Corollary 3.4 with ε\varepsilon replaced by ε/2\varepsilon/2 we find that

(GGn,p for some G𝒮(ε/4,δ,K))\displaystyle\mathbb{P}(G\subseteq G_{n,p}\text{ for some }G\in\mathcal{S}(\varepsilon/4,\delta,K)) (GGn,p for some G𝒮(ε/2,δ,K))\displaystyle\leq\mathbb{P}(G\subseteq G_{n,p}\text{ for some }G\in\mathcal{S}^{\prime}(\varepsilon/2,\delta,K))
(GGn,p for some G𝒞(ε,δ,K)).\displaystyle\leq\mathbb{P}(G\subseteq G_{n,p}\text{ for some }G\in\mathcal{C}(\varepsilon,\delta,K)).

Applying Corollary 2.7 with ε\varepsilon replaced with ε/2\varepsilon/2 we obtain

(X(1+δ)𝔼[X])(1+ε)(GGn,p for some G𝒮(ε/4,δ,K)).\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\mathbb{P}(G\subseteq G_{n,p}\text{ for some }G\in\mathcal{S}(\varepsilon/4,\delta,K)).

Combining the above inequalities we obtain the assertion of the theorem, that is

(X(1+δ)𝔼[X])(1+ε)(GGn,p for some G𝒞(ε,δ,K)).\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1+\varepsilon)\mathbb{P}(G\subseteq G_{n,p}\text{ for some }G\in\mathcal{C}(\varepsilon,\delta,K)).\qed

4. Lower bounds

The aim of this section is to give lower bounds for (X(1+δ)𝔼[X])\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]) for every positive δ\delta. We do so by presenting a family of graphs such that ‘planting’ them in Gn,pG_{n,p} increases the expectation of XX by a multiplicative factor of 1+δ1+\delta. More formally, when we say ‘planting’ we mean changing the probability measure on the hypercube {0,1}(n2)\{0,1\}^{\binom{n}{2}} from the usual product measure of Gn,pG_{n,p} to the probability measure of Gn,pG_{n,p} conditioned on the existence of a subgraph from a predetermined family of graphs. The graphs in the families that we will consider will satisfy 𝔼G[X](1+δ)𝔼[X]\mathbb{E}_{G}[X]\geq(1+\delta)\mathbb{E}[X]. We choose these families as such because, on the event that GGn,pG\subseteq G_{n,p} for GKnG\subseteq K_{n} such that 𝔼G[X](1+δ)𝔼[X]\mathbb{E}_{G}[X]\geq(1+\delta)\mathbb{E}[X] the probability of X(1+δ)𝔼[X]X\geq(1+\delta)\mathbb{E}[X] is pretty ‘large’.

Since for every labeled graph GKnG\subseteq K_{n} the probability of Gn,pG_{n,p} containing GG is pe(G)p^{e(G)} it makes sense to consider such graphs with the smallest e(G)e(G). In contrast, in our case there is a wide range of pp where the strongest lower bound is obtained by planting some member of a large family of graphs; each individual graph in the family is sub-optimal in terms of the number of edges, but the size of the family compensates for the difference in the number of edges between the optimal graph and the graphs in our family. This family is, roughly speaking all embeddings of an unbalanced complete bipartite graph into KnK_{n}.

More precisely, denoting the complete bipartite graph with sides of size ss and tt by Ks,tK_{s,t}, we will define integers m0,m2,m3,m_{0},m_{2},m_{3},\ldots and mm_{*} such that k|mkk|m_{k} for k0k\neq 0 and n|mn|m_{*} and mk2k/(k1)δ𝔼[X]m_{k}\approx 2\sqrt{k/(k-1)\delta\mathbb{E}[X]} for k0k\neq 0, m02δ𝔼[X]m_{0}\approx 2\sqrt{\delta\mathbb{E}[X]}, and m(1+2δ1)2𝔼[X]m_{*}\approx(\sqrt{1+2\delta}-1)\sqrt{2\mathbb{E}[X]}.

Note that provided that n1pn1/2n^{-1}\ll p\ll n^{-1/2} the constructions Kk,mk/kK_{k,m_{k}/k} and Km0,m0K_{\sqrt{m_{0}},\sqrt{m_{0}}} contain δ𝔼[X]\delta\mathbb{E}[X] induced copies of C4C_{4}, up to lower order terms. In addition, we will show later that H=K2m/n,n/2H=K_{2m_{*}/n,n/2} admits 𝔼H[X](1+δ)𝔼[X]\mathbb{E}_{H}[X]\geq(1+\delta)\mathbb{E}[X].

Denote by k\mathcal{E}_{k} the set of all copies of Kk,mk/kK_{k,m_{k}/k} in KnK_{n} when k0,1k\neq 0,1. Denote by 0\mathcal{E}_{0} the set of all Km0,m0K_{\sqrt{m_{0}},\sqrt{m_{0}}} in KnK_{n} and and by \mathcal{E}_{*} denote the set of all copies of HH in KnK_{n}. Planting one of k,0\mathcal{E}_{k},\mathcal{E}_{0} or \mathcal{E}_{*} yields a lower bound for the probability of the upper tail event which is valid for all values of pp. As was mentioned in the introduction the significant different between this work and previous ones is the need to plant a large family of sub-optimal graphs and not a single optimal graph. This is true only for the families k\mathcal{E}_{k} where k0k\neq 0. In the case of 0,\mathcal{E}_{0},\mathcal{E}_{*} we could as well plant a single graph from these sets as |0||\mathcal{E}_{0}| and |||\mathcal{E}_{*}| are negligible.

One can compare these bounds, and see that the best one depends on pp in the following way. There exists an increasing sequence {ck}k=1\{c_{k}\}_{k=1}^{\infty} with c1=0c_{1}=0 and limkck=1/3\lim_{k\rightarrow\infty}c_{k}=1/3 so that, provided n1+ck1pn1+ckn^{-1+c_{k-1}}\ll p\ll n^{-1+c_{k}} for some integer k2k\geq 2 we obtain the strongest lower bound on the upper tail probability by planting k\mathcal{E}_{k}. The best family to plant when n2/3pn1/2n^{-2/3}\ll p\ll n^{-1/2} is 0\mathcal{E}_{0} which should be thought of as the ‘limit’ of k\mathcal{E}_{k} when kk goes to infinity. Lastly, \mathcal{E}^{*} is the best family to plant when n1/2p1n^{-1/2}\ll p\ll 1.

The main result of this section is a formalisation of the above discussion. In order to make things rigorous from now on we fix δ,ε\delta,\varepsilon to be positive reals and let

rk={2 for k=0,2k/(k1) for k2,r_{k}=\Bigg{\{}\begin{array}[]{@{}l@{\thinspace}l}2&\text{ for }k=0,\\ 2\sqrt{k/(k-1)}&\text{ for }k\geq 2,\\ \end{array}

when kk is some non negative integer. Moreover, note the following: For all n1p1n^{-1}\ll p\ll 1 we have 𝔼[X]=Ω(n4p4)\mathbb{E}[X]=\Omega(n^{4}p^{4}). Therefore, for clarity of the presentation we assume that rk(δ+ε)𝔼[X]r_{k}\sqrt{(\delta+\varepsilon)\mathbb{E}[X]} is an integer divisible by kk for any 2kO(np)2\leq k\leq O(np). Furthermore, if pn1/2p\gg n^{-1/2} we have 𝔼[X]/n=Ω(np2)\sqrt{\mathbb{E}[X]}/n=\Omega(np^{2}), hence for clarity of the presentation we assume C(ε,δ)𝔼[X]/n\sqrt{C(\varepsilon,\delta)\mathbb{E}[X]}/n is an integer where C(ε,δ)C(\varepsilon,\delta) will be specified later. The lower bounds given by the following theorem should be thought of as the probability of the appearance of Kk,mk/kK_{k,m_{k}/k} in Gn,pG_{n,p} for some fixed integer 2kO(np)2\leq k\leq O(np) (provided pn1/2p\ll n^{-1/2}) and the appearance of HH in Gn,pG_{n,p}. Where we define:

mk\displaystyle m_{k} =rk(δ+ε)𝔼[X],\displaystyle=r_{k}\sqrt{(\delta+\varepsilon)\mathbb{E}[X]},
m\displaystyle m_{*} =C(ε,δ)𝔼[X]/2,\displaystyle=C(\varepsilon,\delta)\sqrt{\mathbb{E}[X]}/2,

where C(ε,δ)=r+d2dC(\varepsilon,\delta)=\sqrt{r+d^{2}}-d and r=16(δ+3ε/2)r=16(\delta+3\varepsilon/2) and d=2/(1+ε)d=\sqrt{2}/(1+\varepsilon). Now we are ready to state the main result of this section. We wish to emphasize that both mkm_{k} and mm_{*} depend on δ\delta and ε\varepsilon.

Theorem 4.1.

Let ε,δ,C\varepsilon,\delta,C be positive real numbers and let 2kCnp2\leq k\leq Cnp be a positive integer. Then the following holds

(X(1+δ)𝔼[X]){(pmk(nmk/k))1+ε for n1pn1/2p(1+ε)m for n1/2p1\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\geq\begin{cases}\left(p^{m_{k}}\binom{n}{m_{k}/k}\right)^{1+\varepsilon}&\text{ for }n^{-1}\ll p\ll n^{-1/2}\\ p^{({1+\varepsilon})m_{*}}&\text{ for }n^{-1/2}\ll p\ll 1\\ \end{cases}

for large enough nn.

In the following lemma we claim that the expected number of induced copies of C4C_{4} conditioned on one labeled copy of Kk,mk/kK_{k,m_{k}/k} or HH being a subgraph of Gn,pG_{n,p} in a suitable range of pp is at least (1+δ+ε/2)𝔼[X](1+\delta+\varepsilon/2)\mathbb{E}[X] provided nn is large enough. To this end let us introduce the following notations. From now on we assume that the vertex set of Gn,pG_{n,p} is [n][n]. For every positive integer knk\leq n and A[n][k]A\subset[n]\setminus[k] with |A|=mk/k|A|=m_{k}/k define the following events:

  1. (I)

    FA,kF_{A,k} is the event that i=1kN(i)=A\cap_{i=1}^{k}N(i)=A and there are no edges between any i,j[k]i,j\in[k],

  2. (II)

    Fk=A([n][k]mk/k)FA,kF_{k}=\cup_{A\in\binom{[n]\setminus[k]}{m_{k}/k}}F_{A,k}.

Now we are ready to state the lemma.

Lemma 4.2.

Suppose ε,δ,C\varepsilon,\delta,C are positive reals, 2kCnp2\leq k\leq Cnp is some positive integer and A[n][k]A\subseteq[n]\setminus[k] with |A|=mk/k|A|=m_{k}/k. Then the following holds:

  1. (1)

    Suppose n1pn1/2n^{-1}\ll p\ll n^{-1/2}. Then, for large enough nn we have

    𝔼[XFA,k](1+δ+3ε/4)𝔼[X].\mathbb{E}[X\mid F_{A,k}]\geq(1+\delta+3\varepsilon/4)\mathbb{E}[X].
  2. (2)

    Suppose n1/2p1n^{-1/2}\ll p\ll 1. Then, for large enough nn we have

    𝔼H[X](1+δ+ε)𝔼[X].\mathbb{E}_{H}[X]\geq(1+\delta+\varepsilon)\mathbb{E}[X].
Proof.

We start with the first item in the lemma. Assume n1pn1/2n^{-1}\ll p\ll n^{-1/2}. First, note that

E[XFA,k]Nind(C4,Kk,mk/k)(1p)+𝔼[Nind(C4,Gnk,p)].E[X\mid F_{A,k}]\geq N_{ind}(C_{4},K_{k,m_{k}/k})(1-p)+\mathbb{E}[N_{ind}(C_{4},G_{n-k,p})].

Let us compute these two quantities.

Since Ks,tK_{s,t} contains exactly (s2)(t2)\binom{s}{2}\binom{t}{2} induced copies of C4C_{4} we obtain that Kk,mk/kK_{k,m_{k}/k} contains

(k1)mk2/4kO(mk)=(δ+ε)𝔼[X]O(𝔼[X]1/2)(k-1)m_{k}^{2}/4k-O(m_{k})=(\delta+\varepsilon)\mathbb{E}[X]-O(\mathbb{E}[X]^{1/2})

induced copies of C4C_{4} as rk=2k/(k1)r_{k}=2\sqrt{k/(k-1)}. Further, as k=O(np)k=O(np) and for all nonnegative integers a,b,ca,b,c we have (acb)/(ab)(abcab)b\binom{a-c}{b}/\binom{a}{b}\geq\left(\frac{a-b-c}{a-b}\right)^{b}, we deduce

𝔼[Nind(C4,Gnk,p)]=(nk4)(n4)𝔼[X](1kn4)4𝔼[X]=(1o(1))𝔼[X].\mathbb{E}[N_{ind}(C_{4},G_{n-k,p})]=\frac{\binom{n-k}{4}}{\binom{n}{4}}\mathbb{E}[X]\geq\left(1-\frac{k}{n-4}\right)^{4}\mathbb{E}[X]=(1-o(1))\mathbb{E}[X].

Combining the above we obtain,

𝔼[XFA,k](1o(1))(δ+ε)𝔼[X]O(𝔼[X]1/2)+(1o(1))𝔼[X](1+δ+3ε/4)𝔼[X].\mathbb{E}[X\mid F_{A,k}]\geq(1-o(1))(\delta+\varepsilon)\mathbb{E}[X]-O(\mathbb{E}[X]^{1/2})+(1-o(1))\mathbb{E}[X]\geq(1+\delta+3\varepsilon/4)\mathbb{E}[X].

For the second item assume n1/2p1n^{-1/2}\ll p\ll 1. Similar to the first case we have

𝔼H[X](Nind(C4,H)+Nind(K1,2K1,H)p2)(1p)2+𝔼[Nind(C4,Gn2m/n,p)].\mathbb{E}_{H}[X]\geq(N_{ind}(C_{4},H)+N_{ind}(K_{1,2}\sqcup K_{1},H)p^{2})(1-p)^{2}+\mathbb{E}[N_{ind}(C_{4},G_{n-2m_{*}/n,p})].

We now compute these quantities.

Since Ks,tK_{s,t} contains exactly (s2)(t2)\binom{s}{2}\binom{t}{2} induced copies of C4C_{4}, we obtain that HH contains m2/4+O(mn){m_{*}^{2}}/{4}+O(m_{*}n) induced copies of C4C_{4}. Moreover, thinking of HH as a spanning subgraph of KnK_{n} by adding isolated vertices, we see that HH contains at least mn2/8+O(m2)m_{*}n^{2}/8+O(m_{*}^{2}) induced copies of K1,2K1K_{1,2}\sqcup K_{1} as we can choose one vertex from one side, two form the other side and another isolated vertex. Furthermore, as for all nonnegative integers a,b,ca,b,c we have (acb)/(ab)(abcab)b\binom{a-c}{b}/\binom{a}{b}\geq\left(\frac{a-b-c}{a-b}\right)^{b} and 2m/n=O(np2)=o(n)2m_{*}/n=O(np^{2})=o(n), we deduce

𝔼[Nind(C4,Gn2m/n,p)]=(n2m/n4)(n4)𝔼[X](12mn(n4))4𝔼[X](1o(1))𝔼[X],\mathbb{E}[N_{ind}(C_{4},G_{n-2m_{*}/n,p})]=\frac{\binom{n-2m_{*}/n}{4}}{\binom{n}{4}}\mathbb{E}[X]\geq\left(1-\frac{2m_{*}}{n(n-4)}\right)^{4}\mathbb{E}[X]\geq(1-o(1))\mathbb{E}[X],

taking nn large enough we have, 𝔼[Nind(C4,Gn2m/n,p)](1ε/4)𝔼[X]\mathbb{E}[N_{ind}(C_{4},G_{n-2m_{*}/n,p})]\geq(1-\varepsilon/4)\mathbb{E}[X]. Recalling that pn1/2p\gg n^{-1/2} and combining all the above bounds we obtain,

𝔼H[X]m2/4+mn2p2/8+(1ε/4)𝔼[X]+o(m2).\mathbb{E}_{H}[X]\geq{m_{*}^{2}}/{4}+m_{*}n^{2}p^{2}/8+(1-\varepsilon/4)\mathbb{E}[X]+o(m_{*}^{2}).

By this inequality and the definition of mm_{*} one can check that for large enough nn,

𝔼H[X](δ+3ε/2)𝔼[X]+(1ε/4)𝔼[X]ε/4𝔼[X]=(1+δ+ε)𝔼[X].\mathbb{E}_{H}[X]\geq(\delta+3\varepsilon/2)\mathbb{E}[X]+(1-\varepsilon/4)\mathbb{E}[X]-\varepsilon/4\mathbb{E}[X]=(1+\delta+\varepsilon)\mathbb{E}[X].

This completes the proof. ∎

Now we are ready to prove Theorem 4.1. Before proving the theorem we make the following remark. Suppose γ>0\gamma>0 is a real number, k=npk=np and pn1p\gg n^{-1}. Then m0mk(1+γ)m0m_{0}\leq m_{k}\leq(1+\gamma)m_{0} provided nn is sufficiently large. Therefore, Theorem 4.1 invoked with ε=γ\varepsilon=\gamma implies:

(X(1+δ)𝔼[X])(pmk(nmk/np))(1+γ)p(1+γ)2m0\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\geq\left(p^{m_{k}}\binom{n}{m_{k}/np}\right)^{(1+\gamma)}\geq p^{(1+\gamma)^{2}m_{0}}

for large enough nn. Thus, by setting γ=1+ε1\gamma=\sqrt{1+\varepsilon}-1, and letting {ck}k=2\{c_{k}\}_{k=2}^{\infty} be any increasing sequence satisfying c2=0c_{2}=0 and limkck=1/3\lim_{k\rightarrow\infty}c_{k}=1/3, we have the following corollary of Theorem 4.1, which will be shown to be tight in the next sections for some specific sequence ckc_{k}.

Corollary 4.3.

Let ε\varepsilon and δ\delta be positive real numbers and let k2k\geq 2 be positive integer. Then the following holds

log(X(1+δ)𝔼[X]){(1+ε)(mklogp+log(nmk/k)) for n1+ck1pn1+ck,(1+ε)m0log(p) for n2/3pn1/2,(1+ε)mlog(p) for n1/2p1.\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\geq\begin{cases}(1+\varepsilon)(m_{k}\log p+\log\binom{n}{m_{k}/k})&\text{ for }n^{-1+c_{k-1}}\ll p\ll n^{-1+c_{k}},\\ (1+\varepsilon)m_{0}\log(p)&\text{ for }n^{-2/3}\ll p\ll n^{-1/2},\\ {(1+\varepsilon)m_{*}}\log(p)&\text{ for }n^{-1/2}\ll p\ll 1.\\ \end{cases}

for large enough nn.

Note that in the proof of Theorem 4.1 we use a similar method as was used in [21].

Proof of Theorem 4.1.

Through out the proof we assume that the vertex set of Gn,pG_{n,p} is [n][n].

We start with the first item. To this end fix an integer 2kCnp2\leq k\leq Cnp and ε\varepsilon some positive real and assume that n1pn1/2n^{-1}\ll p\ll n^{-1/2}. Let A[n][k]A\subset[n]\setminus[k] with |A|=mk/k|A|=m_{k}/k and recall the definitions of the events FA,kF_{A,k} and FkF_{k}:

  1. (I)

    FA,kF_{A,k} is the event that i=1kN(i)=A\cap_{i=1}^{k}N(i)=A and there are no edges between any i,j[k]i,j\in[k],

  2. (II)

    Fk=A([n][k]mk/k)FA,kF_{k}=\cup_{A\in\binom{[n]\setminus[k]}{m_{k}/k}}F_{A,k}.

Note that for any A,B[n][k]A,B\subseteq[n]\setminus[k] such that ABA\neq B with |A|=|B|=mk/k|A|=|B|=m_{k}/k we have FA,kFB,k=F_{A,k}\cap F_{B,k}=\emptyset. Therefore, we have the following,

(X(1+δ)𝔼[X])\displaystyle\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]) (Fk and X(1+δ)𝔼[X])\displaystyle\geq\mathbb{P}(F_{k}\text{ and }X\geq(1+\delta)\mathbb{E}[X])
=A([n][k]mk/k)(X(1+δ)𝔼[X]FA,k)(FA,k)\displaystyle=\sum\limits_{A\in\binom{[n]\setminus[k]}{m_{k}/k}}\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]\mid F_{A,k})\cdot\mathbb{P}(F_{A,k})
A([n][k]mk/k)(X(1+δ)𝔼[X]FA,k)pmk(1p)k2(1pk)(nk).\displaystyle\geq\sum\limits_{A\in\binom{[n]\setminus[k]}{m_{k}/k}}\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]\mid F_{A,k})\cdot p^{m_{k}}(1-p)^{k^{2}}(1-p^{k})^{(n-k)}. (8)

Lemma 4.2 asserts that provided n1pn1/2n^{-1}\ll p\ll n^{-1/2} we have the following for all A[n][k]A\subset[n]\setminus[k] with |A|=mk/k|A|=m_{k}/k:

𝔼[XFA,k](1+δ+ε/2)𝔼[X].\mathbb{E}[X\mid F_{A,k}]\geq(1+\delta+\varepsilon/2)\mathbb{E}[X].

Note that Xn4X\leq n^{4} always, we can bound 𝔼[XFA,k]\mathbb{E}[X\mid F_{A,k}] from above (similar to the proof of Lemma 2.2) as follows:

𝔼[XFA,k](1+δ)𝔼[X]+(X(1+δ)𝔼[X]FA,k)n4.\mathbb{E}[X\mid F_{A,k}]\leq(1+\delta)\mathbb{E}[X]+\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]\mid F_{A,k})\cdot n^{4}.

Combining the two inequalities we obtain,

(X(1+δ)𝔼[X]FA,k)ε𝔼[X]2n4.\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]\mid F_{A,k})\geq\frac{\varepsilon\cdot\mathbb{E}[X]}{2n^{4}}.

Therefore, for large enough nn we also have,

(X(1+δ)𝔼[X]FA,k)εp4(1p)2/(244)p5.\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]\mid F_{A,k})\geq\varepsilon p^{4}(1-p)^{2}/(2\cdot 4^{4})\geq p^{5}. (9)

Substituting (9) into (4) gives,

(X(1+δ)𝔼[X])\displaystyle\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X]) A([n][k]mk/k)p5pmk(1p)k2(1pk)nk\displaystyle\geq\sum\limits_{A\in\binom{[n]\setminus[k]}{m_{k}/k}}p^{5}p^{m_{k}}(1-p)^{k^{2}}(1-p^{k})^{n-k}
=(nkmk/k)pmk+5(1p)k2(1pk)nk.\displaystyle=\binom{n-k}{m_{k}/k}p^{m_{k}+5}(1-p)^{k^{2}}(1-p^{k})^{n-k}. (10)

We now show that p5(1p)k2(1pk)nkpo(mk)p^{5}(1-p)^{k^{2}}(1-p^{k})^{n-k}\geq p^{o(m_{k})} and (nkmk/k)(nmk/k)1+o(1)po(mk)\binom{n-k}{m_{k}/k}\geq\binom{n}{m_{k}/k}^{1+o(1)}p^{o(m_{k})}.

First, as pn1/2,mk1p\ll n^{-1/2},m_{k}\gg 1 and 2kCnp2\leq k\leq Cnp, we have the following for large nn:

p5(1p)k2(1pk)(nk)\displaystyle p^{5}(1-p)^{k^{2}}(1-p^{k})^{(n-k)} exp(5log(p)k2(p+p2)(nk)(pk+p2k))\displaystyle\geq\exp(5\log(p)-k^{2}(p+p^{2})-(n-k)(p^{k}+p^{2k}))
exp(5log(p)C2n2p3(1+p)np2(1+pk))\displaystyle\geq\exp(5\log(p)-C^{2}n^{2}p^{3}(1+p)-np^{2}(1+p^{k}))
=exp(o(n2p2))\displaystyle=\exp(-o(n^{2}p^{2}))
pεmk/2,\displaystyle\geq p^{\varepsilon m_{k}/2}, (11)

where the first inequality follows as 1xexp(xx2)1-x\geq\exp(-x-x^{2}) for all small enough xx. Second, for all nonnegative integers a,b,ca,b,c we have (acb)/(ab)(abcab)b\binom{a-c}{b}/\binom{a}{b}\geq\left(\frac{a-b-c}{a-b}\right)^{b}, and thus we obtain that

(nkmk/k)/(nmk/k)(1knmk/k)mk/k=eO(mk/n).\binom{n-k}{m_{k}/k}/\binom{n}{m_{k}/k}\geq\left({1-\frac{k}{n-m_{k}/k}}\right)^{m_{k}/k}=e^{-O(m_{k}/n)}.

In addition, as 2kCnp2\leq k\leq Cnp and n1pn1/2n^{-1}\ll p\ll n^{-1/2}, we have

(nmk/k)pmk\displaystyle\binom{n}{m_{k}/k}p^{m_{k}} (enkmk)mk/kpmk(eCn2pmk)mk/kpmk\displaystyle\leq\left(\frac{enk}{m_{k}}\right)^{m_{k}/k}p^{m_{k}}\leq\left(\frac{eCn^{2}p}{m_{k}}\right)^{m_{k}/k}p^{m_{k}}
exp((11k+O(1klog(p)))mklog(p))\displaystyle\leq\exp\left(\left(1-\frac{1}{k}+O\left(\frac{1}{k\log(p)}\right)\right)m_{k}\log(p)\right)
exp((1+o(1))mklog(p)).\displaystyle\leq\exp((1+o(1))m_{k}\log(p)).

Furthermore, as 2kCnp2\leq k\leq Cnp and n1pn1/2n^{-1}\ll p\ll n^{-1/2} we have mk/n=o(mklog(1/p))m_{k}/n=o(m_{k}\log(1/p)) and hence, for sufficiently large nn we have,

(nkmk/k)/(nmk/k)(nmk/k)εpεmk/2.\binom{n-k}{m_{k}/k}/\binom{n}{m_{k}/k}\geq\binom{n}{m_{k}/k}^{\varepsilon}p^{\varepsilon m_{k}/2}. (12)

Combining (4), (4), and (12) gives,

(X(1+δ)𝔼[X])((nmk/k)pmk)1+ε.\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\geq\left(\binom{n}{m_{k}/k}p^{m_{k}}\right)^{1+\varepsilon}.

This finishes the first part of the proof.

For the second item, define the event FF^{*} to be the event that Gn,pG_{n,p} contains K2m/n,n/2K_{2m_{*}/n,n/2} as a subgraph on the vertex set [n/2+2m/n][n/2+2m_{*}/n] with sides [2m/n][2m_{*}/n] and [n/2+2m/n][2m/n][n/2+2m_{*}/n]\setminus[2m_{*}/n]. Note that Xn4X\leq n^{4} always. Further, Lemma 4.2 asserts that provided n1/2p1n^{-1/2}\ll p\ll 1 we have

𝔼[XF](1+δ+ε)𝔼[X],\mathbb{E}[X\mid F^{*}]\geq(1+\delta+\varepsilon)\mathbb{E}[X],

and thus, ΦX(δ+ε)log(F)=log(pm)\Phi_{X}(\delta+\varepsilon)\leq-\log\mathbb{P}(F^{*})=-\log(p^{m_{*}}). Therefore, applying Lemma 2.2 gives,

log(X(1+δ)𝔼[X])\displaystyle-\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq ΦX(δ+ε)+log(n4ε𝔼[X])log(pmn4ε𝔼[X]).\displaystyle\Phi_{X}(\delta+\varepsilon)+\log\left(\frac{n^{4}}{\varepsilon\mathbb{E}[X]}\right)\leq-\log\left(\frac{p^{m_{*}}n^{4}}{\varepsilon\mathbb{E}[X]}\right).

Moreover, for sufficiently large nn we have, ε𝔼[X]n4p5\frac{\varepsilon\mathbb{E}[X]}{n^{4}}\geq p^{5}. Thus, taking negative logarithms we obtain the following for large enough nn:

(X(1+δ)𝔼[X])pm+5p(1+ε)m.\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\geq p^{m_{*}+5}\geq p^{(1+\varepsilon)m_{*}}.

This is as claimed. ∎

5. Counting the number of cores

Recall that XX is the random variable counting the number of induced copies of C4C_{4} in Gn,pG_{n,p}. In this section we prove a general upper bound on the logarithmic probability of the upper tail event of XX. The main tool we use in this section is Theorem 2.7 which is a variation of . There are two major parts in this section. The first is an evaluation of ΦX(δ)\Phi_{X}(\delta) in different regimes of pp. The second is an evaluation of the entropic term |𝒞m||\mathcal{C}_{m}|. The main results of this section are the following lemmas.

Lemma 5.1.

Suppose ε,δ\varepsilon,\delta are positive real numbers with ε\varepsilon being small as a function of δ\delta. Then the following hold:

  1. (i)

    If pn1/2p\ll n^{-1/2} then for large enough nn we have,

    2(1ε)δ𝔼[X]ΦX(δ)/log(1/p)2(1+ε)δ𝔼[X].2(1-\varepsilon)\sqrt{\delta\mathbb{E}[X]}\leq\Phi_{X}(\delta)/\log(1/p)\leq 2(1+\varepsilon)\sqrt{\delta\mathbb{E}[X]}.
  2. (ii)

    If n1/2p1n^{-1/2}\ll p\ll 1 then for large enough nn we have,

    (1ε)(n4p416+4δ𝔼[X]n2p24)ΦX(δ)log(1/p)(1+ε)(n4p416+4δ𝔼[X]n2p24).(1-\varepsilon)\left(\sqrt{\frac{n^{4}p^{4}}{16}+4\delta\mathbb{E}[X]}-\frac{n^{2}p^{2}}{4}\right)\leq\frac{\Phi_{X}(\delta)}{\log(1/p)}\leq(1+\varepsilon)\left(\sqrt{\frac{n^{4}p^{4}}{16}+4\delta\mathbb{E}[X]}-\frac{n^{2}p^{2}}{4}\right).

Before we present the second lemma, let us remind the reader the definition of 𝒞m\mathcal{C}_{m}.

Definition.

Let ε,δ,K\varepsilon,\delta,K be positive reals. In addition let p(0,1)p\in(0,1). Then we define 𝒞(ε,δ,K)\mathcal{C}(\varepsilon,\delta,K) to be collection of all GKnG\subset K_{n} spanning subgraphs satisfying the following:

  1. (C1C{1})

    e(G)KΦX(δ+ε)e(G)\leq K\cdot\Phi_{X}(\delta+\varepsilon),

  2. (C2C{2})

    Nind(C4,G)+Nind(K1,2K1,G)p2(δε)𝔼[X]N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}\geq(\delta-\varepsilon)\mathbb{E}[X], and

  3. (C3C{3})

    For all eE(G)e\in E(G)

    Nind(e,C4,G)+Nind(e,K1,2K1,G)p2ε𝔼[X]/(2KΦX(δ+ε)).N_{ind}(e,C_{4},G)+N_{ind}(e,K_{1,2}\sqcup K_{1},G)p^{2}\geq\varepsilon\mathbb{E}[X]/(2K\cdot\Phi_{X}(\delta+\varepsilon)).

We call the graphs in 𝒞(ε,δ,K)\mathcal{C}(\varepsilon,\delta,K) core graphs. Furthermore, for every positive integer mm we define 𝒞m(ε,δ,K)\mathcal{C}_{m}(\varepsilon,\delta,K) be the set of all cores with mm edges.

Lemma 5.2.

Suppose ε,δ,C,K\varepsilon,\delta,C,K are positive reals with ε<1\varepsilon<1. Furthermore, suppose n1p1n^{-1}\ll p\ll 1 as nn tends to infinity. Then, there exist DD and n0n_{0} such that the following holds for all n>n0n>n_{0}:
Let mm be a positive integer with Cn2p2mKΦX(δ+ε)Cn^{2}p^{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon). Furthermore, let

vm=max{|{vV(G):deg(v)0}|:G𝒞m(ε,δ,K)}.v_{m}=\max\{|\{v\in V(G):\deg(v)\neq 0\}|:G\in\mathcal{C}_{m}(\varepsilon,\delta,K)\}.

Then,

|𝒞m|log(1/p)Dm(nvm).|\mathcal{C}_{m}|\leq\log(1/p)^{Dm}\binom{n}{v_{m}}.

We start with the first lemma. Let us give some motivation and history of the problem.

Definition.

For every graph HH and any positive integer mm define,

N(m,H)=max|{TG:TH}|N(m,H)=\max|\{T\subset G:T\cong H\}|

where the maximum ranges over all graphs with mm edges.

This definition was first presented by Erdős and Hanani in [17]. In their paper they computed the asymptotic of this function where HH is a clique. In [1] Alon generalized this and computed the asymptotic of this function for all HH. Later, in [18] Friedgut and Kahn reproved Alon’s result using entropy methods which they were also able to generalize for the case of hypergraphs.

In [22] Janson, Oleszkiewicz, and Ruciński found a relation between N(m,H)N(m,H) and a related parameter N(n,m,H)N(n,m,H) and the probability that the random variable counting the number of copies of a fixed graph HH in Gn,pG_{n,p} exceeds its expectation by a multiplicative factor. This result used and generalized the machinery developed by Friedgut and Kahn. This led us to generalize the definition of N(m,H)N(m,H) to fit our setting. Let us recall some definitions from Section 3 and introduce some new ones.

Definition.

Suppose HH and GG are graphs. Let n,mn,m be positive integers and let ee be an edge of GG. We define:

  • Nind(H,G)N_{ind}(H,G) is the number of induced copies of HH in GG.

  • Nind(e,H,G)N_{ind}(e,H,G) is the number of induced copies of HH in GG that contain ee.

  • Nind(n,m,H){N}_{ind}(n,m,H) is the maximum of Nind(H,G)N_{ind}(H,G) over all graphs GG such that the number of vertices of GG is at most nn and the number of edges of GG is mm.

Note that similar generalizations have been studied in [7] and [19]. The following is a simple corollary of Lemma 6.2 which we prove later in Section 6.

Corollary 5.3.

For every n,mn,m positive integers such that 3<mn3<m\leq n we have,

Nind(n,m,C4)m(mn+1)4m24.{N}_{ind}(n,m,C_{4})\leq\frac{m(m-n+1)}{4}\leq\frac{m^{2}}{4}.

In [7] the authors determined the asymptotic behaviour of max{Nind(n,m,Ks,s):n}\max\{N_{ind}(n,m,K_{s,s}):n\in\mathbb{N}\} for any ss\in\mathbb{N} (and actually obtained optimal bounds), and gave tight bounds in some ranges of mm (in the above we consider only K2,2K_{2,2}). In [19] the authors considered the problem of determining the asymptotic in nn and mm of the maximum number of copies of a fixed bipartite graph HH in a bipartite host graph with nn vertices and mm edges. They solved this problem for some class of graphs which includes all complete bipartite graphs Ks,tK_{s,t}. Here we prove a similar statement to the ones in [7, 19].

We wish to emphasize that there is a significant difference in the behaviour of the problem depending on whether n1pn1/2n^{-1}\ll p\ll n^{-1/2} or n1/2p1n^{-1/2}\ll p\ll 1. This can be seen for example in the first result of this section which we now start to prove. To this end we start with an observation. Recall that we denote by K1,2K1K_{1,2}\sqcup K_{1} the disjoint union of a star with two leaves and an isolated vertex.

Observation 5.4.

Suppose n,mn,m are positive integers with m(n2)m\leq\binom{n}{2}. Then,

Nind(n,m,K1,2K1)mn2/8.N_{ind}(n,m,K_{1,2}\sqcup K_{1})\leq mn^{2}/8.
Proof.

Let GG be a graph achieving the maximum in the definition of Nind(n,m,K1,2K1)N_{ind}(n,m,K_{1,2}\sqcup K_{1}). Let uv=eE(G)uv=e\in E(G) and denote x=|N(u)N(v){u,v}|x=|N(u)\cup N(v)\setminus\{u,v\}|. Then,

Nind(e,K1,2K1,G)x(nx)n2/4.\displaystyle N_{ind}(e,K_{1,2}\sqcup K_{1},G)\leq x(n-x)\leq n^{2}/4.

Moreover we have,

Nind(K1,2K1,G)12eE(G)Nind(e,K1,2K1,G)mn28.N_{ind}(K_{1,2}\sqcup K_{1},{G})\leq\frac{1}{2}\sum_{e\in E({G})}N_{ind}(e,K_{1,2}\sqcup K_{1},G)\leq\frac{mn^{2}}{8}.\qed
Proof of Lemma 5.1.

We start with the upper bounds both in (i) and in (ii).

For the upper bound in (i), assume that n1pn1/2n^{-1}\ll p\ll n^{-1/2}. Since limn𝔼[X]=\lim_{n\rightarrow\infty}\mathbb{E}[X]=\infty we may treat m=4(1+ε)δ𝔼[X]4m^{\prime}=\sqrt[4]{4(1+\varepsilon)\delta\mathbb{E}[X]} as an integer and let H=Km,mH=K_{m^{\prime},m^{\prime}}. Note that HH has 2(1+ε)δ𝔼[X]2\sqrt{(1+\varepsilon)\delta\mathbb{E}[X]} edges and provided nn is large enough HH contains at least (1+ε/2)δ𝔼[X]>δ𝔼[X](1+\varepsilon/2)\delta\mathbb{E}[X]>\delta\mathbb{E}[X] induced copies of C4C_{4}. Note further that,

𝔼H[X]Nind(C4,H)(1p)2+𝔼[Nind(C4,Gn2m,p)](1+ε/2)δ𝔼[X]+(n2m4)(n4)𝔼[X].\mathbb{E}_{H}[X]\geq N_{ind}(C_{4},H)(1-p)^{2}+\mathbb{E}[N_{ind}(C_{4},G_{n-2m^{\prime},p})]\geq(1+\varepsilon/2)\delta\mathbb{E}[X]+\frac{\binom{n-2m^{\prime}}{4}}{\binom{n}{4}}\mathbb{E}[X].

As mnm^{\prime}\ll n we also have the following for sufficiently large nn,

(n2m4)/(n4)(n2m3n)41δε/2.\binom{n-2m^{\prime}}{4}/{\binom{n}{4}}\geq\left(\frac{n-2m^{\prime}-3}{n}\right)^{4}\geq 1-\delta\varepsilon/2.

Combining the above inequalities we find the following for large enough nn,

𝔼H[X](1+ε/2)δ𝔼[X]+(1εδ/2)𝔼[X]=(1+δ)𝔼[X].\mathbb{E}_{H}[X]\geq(1+\varepsilon/2)\delta\mathbb{E}[X]+(1-\varepsilon\delta/2)\mathbb{E}[X]=(1+\delta)\mathbb{E}[X].

Therefore, for large enough nn we have

𝔼H[X]𝔼[X]δ𝔼[X].\mathbb{E}_{H}[X]-\mathbb{E}[X]\geq\delta\mathbb{E}[X].

Hence, by the definition of ΦX(δ)\Phi_{X}(\delta),

ΦX(δ)log(pe(H))=2(1+ε)δ𝔼[X]log(1/p)2(1+ε)δ𝔼[X]log(1/p).\Phi_{X}(\delta)\leq-\log(p^{e(H)})=2\sqrt{(1+\varepsilon)\delta\mathbb{E}[X]}\log(1/p)\leq 2(1+\varepsilon)\sqrt{\delta\mathbb{E}[X]}\log(1/p).

For the upper bound in (ii), assume n1/2p1n^{-1/2}\ll p\ll 1. Letting

m~=(1+ε)(n4p416+4δ𝔼[X]n2p24),\tilde{m}=(1+\varepsilon)\left(\sqrt{\frac{n^{4}p^{4}}{16}+4\delta\mathbb{E}[X]}-\frac{n^{2}p^{2}}{4}\right),

and recalling that limn𝔼[X]/n=\lim_{n\rightarrow\infty}\sqrt{\mathbb{E}[X]}/n=\infty we may treat m~/n\tilde{m}/n as an integer and consider the following graph. Let HH be a graph on the vertex set [n][n] and such that H[[2m~/n+n/2]]K2m~/n,n/2H\left[[2\tilde{m}/n+n/2]\right]\cong K_{2\tilde{m}/n,n/2} and all vertices in [n][2m~/n+n/2][n]\setminus[2\tilde{m}/n+n/2] are isolated. Note that HH contains m~\tilde{m} edges. Let ξ\xi be a small positive real. Assuming nn is large enough, HH contains at least (1ξ)m~2/4(1-\xi)\tilde{m}^{2}/4 induced copies of C4C_{4}. Furthermore, if nn is large enough, HH contains at least (1ξ)m~n28(1-\xi)\frac{\tilde{m}n^{2}}{8} induced copies of K1,2K1K_{1,2}\sqcup K_{1}. Let η\eta be a small positive real (that might depend on ξ\xi). Then provided nn is large enough,

𝔼H[X]\displaystyle\mathbb{E}_{H}[X]\geq (Nind(C4,H)+Nind(K1,2K1,H)p2)(1p)2+𝔼[Nind(C4,Gn2m~/n,p)]\displaystyle\left(N_{ind}(C_{4},H)+N_{ind}(K_{1,2}\sqcup K_{1},H)p^{2}\right)(1-p)^{2}+\mathbb{E}[N_{ind}(C_{4},G_{n-2\tilde{m}/n,p})]
\displaystyle\geq (1η)(1ξ)(m~24+m~n2p28)+(n2m~/n4)(n4)𝔼[X].\displaystyle(1-\eta)(1-\xi)\left(\frac{\tilde{m}^{2}}{4}+\frac{\tilde{m}n^{2}p^{2}}{8}\right)+\frac{\binom{n-2\tilde{m}/n}{4}}{\binom{n}{4}}\mathbb{E}[X].

Where the first inequality holds as the first term is at most the expected number of induced copies C4C_{4} with at least one vertex in [2m~/n][2\tilde{m}/n], and the second term is the expected number of induced copies of C4C_{4} with no vertices in [2m~/n][2\tilde{m}/n].

By the definition of m~\tilde{m} and the choices of ξ,η\xi,\eta being small enough we obtain,

(1η)(1ξ)(m~24+m~n2p28)(1+ε/2)δ𝔼[X].\displaystyle(1-\eta)(1-\xi)\left(\frac{\tilde{m}^{2}}{4}+\frac{\tilde{m}n^{2}p^{2}}{8}\right)\geq(1+\varepsilon/2)\delta\mathbb{E}[X].

Further, for large enough nn we have,

(n2m~/n4)/(n4)(n2m~/n3n)4(1δε/2).\binom{n-2\tilde{m}/n}{4}/\binom{n}{4}\geq\left(\frac{n-2\tilde{m}/n-3}{n}\right)^{4}\geq{(1-\delta\varepsilon/2)}.

Combining all of these inequalities we have,

𝔼H[X]𝔼[X]δ𝔼[X].\mathbb{E}_{H}[X]-\mathbb{E}[X]\geq\delta\mathbb{E}[X].

Hence, by the definition of ΦX(δ)\Phi_{X}(\delta) we have,

ΦX(δ)log(pe(H))=(1+ε)(n4p416+4δ𝔼[X]n2p24)log(1/p).\Phi_{X}(\delta)\leq-\log(p^{e(H)})=(1+\varepsilon)\left(\sqrt{\frac{n^{4}p^{4}}{16}+4\delta\mathbb{E}[X]}-\frac{n^{2}p^{2}}{4}\right)\log(1/p).

For the lower bounds of both (i) and (ii) we start with a claim.

Claim 5.5.

Suppose GG is a spanning subgraph of KnK_{n} achieving the minimum in the definition of ΦX(δ)\Phi_{X}(\delta) and let m=e(G)m=e(G). Then for any small enough γ>0\gamma>0 and large enough nn

δ𝔼[X]m2/4+min{mn2p2/8,m2np2/2}+γn4p4.\displaystyle\delta\mathbb{E}[X]\leq m^{2}/4+\min\{mn^{2}p^{2}/8,m^{2}np^{2}/2\}+\gamma n^{4}p^{4}.
Proof.

Let GG be a spanning subgraph of KnK_{n} achieving the minimum in the definition of ΦX(δ)\Phi_{X}(\delta) and put m=e(G)m=e(G). By the definition of ΦX(δ)\Phi_{X}(\delta) we have 𝔼G[X]𝔼[X]δ𝔼[X]\mathbb{E}_{G}[X]-\mathbb{E}[X]\geq\delta\mathbb{E}[X]. Note that we can bound 𝔼G[X]𝔼[X]\mathbb{E}_{G}[X]-\mathbb{E}[X] from above in the following way:

𝔼G[X]𝔼[X]HC4Nind(H,G)p4e(H),\displaystyle\mathbb{E}_{{G}}[X]-\mathbb{E}[X]\leq\sum\limits_{\emptyset\neq H\subseteq^{*}C_{4}}N_{ind}(H,{G})p^{4-e(H)}, (13)

where \subseteq^{*} stands for spanning subgraphs. Let P4P_{4} be the path with three edges, M2M_{2} be a matching of size two and K2I2K_{2}\sqcup I_{2} be a disjoint union of an edge and an independent set of size two. Since any matching of size two span at most one induced copy of P4P_{4} or M2M_{2} we have

Nind(P4,G),Nind(M2,G)m2.N_{ind}(P_{4},G),N_{ind}(M_{2},G)\leq m^{2}.

Therefore we obtain

𝔼G[X]𝔼[X]\displaystyle\mathbb{E}_{G}[X]-\mathbb{E}[X]\leq Nind(C4,G)+Nind(K1,2K1,G)p2\displaystyle N_{ind}(C_{4},{G})+N_{ind}(K_{1,2}\sqcup K_{1},{G})p^{2}
+m2pNind(P4,G)+m2p2Nind(M2,G)+mn2p3Nind(K2I2,G)\displaystyle+\underbrace{m^{2}p}_{N_{ind}(P_{4},{G})}+\underbrace{m^{2}p^{2}}_{N_{ind}(M_{2},{G})}+\underbrace{mn^{2}p^{3}}_{N_{ind}(K_{2}\sqcup I_{2},{G})}
\displaystyle\leq Nind(C4,G)+Nind(K1,2K1,G)p2+mp(m+mp+n2p2).\displaystyle N_{ind}(C_{4},{G})+N_{ind}(K_{1,2}\sqcup K_{1},{G})p^{2}+mp(m+mp+n^{2}p^{2}).

By Observation 5.4 we have

Nind(K1,2K1,G)Nind(m,n,K1,2K1)mn28.N_{ind}(K_{1,2}\sqcup K_{1},{G})\leq N_{ind}(m,n,K_{1,2}\sqcup K_{1})\leq\frac{mn^{2}}{8}.

Furthermore we always have,

Nind(K1,2K1,G)(m2)nm2n2.N_{ind}(K_{1,2}\sqcup K_{1},{G})\leq\binom{m}{2}\cdot n\leq\frac{m^{2}n}{2}.

Combining the two we find that,

Nind(K1,2K1,G)p2min{mn2p2/8,m2np2/2}.\displaystyle N_{ind}(K_{1,2}\sqcup K_{1},{G})p^{2}\leq\min\{mn^{2}p^{2}/8,m^{2}np^{2}/2\}.

Hence, by the definition of ΦX(δ)\Phi_{X}(\delta) and Corollary 5.3, for large enough nn we have,

δ𝔼[X]\displaystyle\delta\mathbb{E}[X] Nind(C4,G)+Nind(K1,2K1,G)p2+mp(m+mp+n2p2)\displaystyle\leq N_{ind}(C_{4},{G})+N_{ind}(K_{1,2}\sqcup K_{1},{G})p^{2}+mp(m+mp+n^{2}p^{2})
m2/4+min{mn2p2/8,m2np2/2}+mp(m+mp+n2p2).\displaystyle\leq m^{2}/4+\min\{mn^{2}p^{2}/8,m^{2}np^{2}/2\}+mp(m+mp+n^{2}p^{2}). (14)

As p1p\ll 1 and the upper bound achieved before, we have m=O(n2p2)m=O(n^{2}p^{2}). Therefore, for any γ>0\gamma>0 and sufficiently large nn we obtain from (5) the following inequality

δ𝔼[X]m2/4+min{mn2p2/8,m2np2/2}+γn4p4.\delta\mathbb{E}[X]\leq m^{2}/4+\min\{mn^{2}p^{2}/8,m^{2}np^{2}/2\}+\gamma n^{4}p^{4}.\qed

We now prove the lower bound in (i). For this assume that n1pn1/2n^{-1}\ll p\ll n^{-1/2} and note that this implies that min{mn2p2/8,m2np2/2}=o(m2)\min\{mn^{2}p^{2}/8,m^{2}np^{2}/2\}=o(m^{2}). Thus, for any γ>0\gamma>0 and large enough nn we deduce the following from Claim 5.5,

δ𝔼[X]m2/4+min{mn2p2/8,m2np2/2}+γn4p4(1/4+γ)m2+γn4p4.\delta\mathbb{E}[X]\leq m^{2}/4+\min\{mn^{2}p^{2}/8,m^{2}np^{2}/2\}+\gamma n^{4}p^{4}\leq(1/4+\gamma)m^{2}+\gamma n^{4}p^{4}.

Taking γ\gamma sufficiently small we obtain the following bound on mm,

(1ε)δ𝔼[X]14(1ε)m2,(1-\varepsilon)\delta\mathbb{E}[X]\leq\frac{1}{4(1-\varepsilon)}m^{2},

implying

m2(1ε)δ𝔼[X].m\geq 2(1-\varepsilon)\sqrt{\delta\mathbb{E}[X]}.

This proves the lower bound in (i) as for sufficiently large nn we obtain,

ΦX(δ)=log(pm)2(1ε)δ𝔼[X]log(1/p).\Phi_{X}(\delta)=-\log(p^{m})\geq 2(1-\varepsilon)\sqrt{\delta\mathbb{E}[X]}\log(1/p).

Lastly, we prove the lower bound in (ii). For this assume n1/2p1n^{-1/2}\ll p\ll 1. Let γ>0\gamma>0 be some small constant, then for large enough nn Claim 5.5 implies,

δ𝔼[X]m2/4+mn2p2/8+γ𝔼[X].\displaystyle\delta\mathbb{E}[X]\leq m^{2}/4+mn^{2}p^{2}/8+\gamma\mathbb{E}[X].

We conclude the following as the above is a quadratic inequality in mm and the fact that 𝔼[X]=(1+o(1))n4p48\mathbb{E}[X]=(1+o(1))\frac{n^{4}p^{4}}{8},

mn4p416+4(δγ)𝔼[X]n2p24(1ε)(n4p416+4δ𝔼[X]n2p24).m\geq\sqrt{\frac{n^{4}p^{4}}{16}+4(\delta-\gamma)\mathbb{E}[X]}-\frac{n^{2}p^{2}}{4}\geq(1-\varepsilon)\left(\sqrt{\frac{n^{4}p^{4}}{16}+4\delta\mathbb{E}[X]}-\frac{n^{2}p^{2}}{4}\right).

This implies (ii) as follows,

ΦX(δ)=log(pm)(1ε)(n4p416+4δ𝔼[X]n2p24)log(1/p).\Phi_{X}(\delta)=-\log(p^{m})\geq(1-\varepsilon)\left(\sqrt{\frac{n^{4}p^{4}}{16}+4\delta\mathbb{E}[X]}-\frac{n^{2}p^{2}}{4}\right)\log(1/p).\qed

This finishes the first evaluation we prove in this section. The second evaluation is the evaluation of the entropic term |𝒞m||\mathcal{C}_{m}| given by Lemma 5.2. Roughly speaking Lemma 5.2 shows that the number of core graphs with mm edges is determined by vmv_{m} the maximum number of non-isolated vertices over all graphs in 𝒞m\mathcal{C}_{m}. As seen already, there is a big difference in the behaviour of the problem depending on whether pn1/2p\ll n^{-1/2} or pn1/2p\gg n^{-1/2}. Later, we will derive two corollaries from Lemma 5.2, corresponding to these regimes, these corollaries will be very important in Section 6.

Proof of Lemma 5.2.

Let G𝒞mG\in\mathcal{C}_{m} be a core graph with mm edges and let uvuv be an edge in GG. Then by the definition of a core graph we have,

Nind(uv,C4,G)+Nind(uv,K1,2,G)np2ε𝔼[X]/(2KΦX(δ+ε)).N_{ind}(uv,C_{4},G)+N_{ind}(uv,K_{1,2},G)np^{2}\geq\varepsilon\mathbb{E}[X]/(2K\cdot\Phi_{X}(\delta+\varepsilon)).

Therefore,

deg(u)deg(v)Nind(uv,C4,G)ε𝔼[X]/(4KΦX(δ+ε)),\deg(u)\deg(v)\geq N_{ind}(uv,C_{4},G)\geq\varepsilon\mathbb{E}[X]/(4K\cdot\Phi_{X}(\delta+\varepsilon)), (15)

or

deg(u)+deg(v)Nind(uv,K1,2,G)ε𝔼[X]/(4KΦX(δ+ε)np2),\deg(u)+\deg(v)\geq N_{ind}(uv,K_{1,2},G)\geq\varepsilon\mathbb{E}[X]/(4K\cdot\Phi_{X}(\delta+\varepsilon)np^{2}), (16)

call this property ()(*). To bound |𝒞m||\mathcal{C}_{m}| it is enough to bound the number of subgraphs of KnK_{n} with mm edges satisfying property ()(*). This can be bounded from above by multiplying the following quantities:

  1. (1)

    The number of possible ways to choose the set of non-isolated vertices of such graph.

  2. (2)

    The number of possible choices of the edges of the graph such that property ()(*) is being satisfied.

The first item can be bounded from above by the number of ways to choose a set of at most vmv_{m} vertices which is at most 2vm(nvm)2^{v_{m}}\binom{n}{v_{m}}.

We now bound the second item. Let HH be a graph with vv vertices and mm edges that satisfies property ()(*). For every integer 0tlog2(m)0\leq t\leq\log_{2}(m) define Vt(H)={vV(H):2tdeg(v)<2t+1}V_{t}(H)=\{v\in V(H):2^{t}\leq\deg(v)<2^{t+1}\}. Furthermore, for every integer 0tlog2(m)0\leq t\leq\log_{2}(m) define Ut(H)=tVU_{t}(H)=\cup_{\ell\geq t}V_{\ell}. Using a standard double counting argument and the fact that Ut(H)V(H)U_{t}(H)\subseteq V(H) we have

|Vt(H)||Ut(H)|min{m2t1,vm}|V_{t}(H)|\leq|U_{t}(H)|\leq\min\left\{\frac{m}{2^{t-1}},v_{m}\right\}

for all t0t\geq 0. By property ()(*) edges can be placed between the sets VtV_{t} in two ways.

The first option is when the edges satisfy (15). Since, the degree of each vertex is always bounded from above by nn, we obtain that the degrees of the endpoints of each edge satisfying (15) are bounded from below by dd^{*} defined as

d=ε𝔼[X]4KnΦX(δ+ε)=C0np2log(1/p)d^{*}=\frac{\varepsilon\mathbb{E}[X]}{4K\cdot n\Phi_{X}(\delta+\varepsilon)}=\frac{C_{0}np^{2}}{\log(1/p)}

where C0C_{0} is some positive real that might depend on ε,δ\varepsilon,\delta and KK. Thus, such edges can only connect a vertex in VtV_{t} and a vertex in Ulog2(t)U_{\lfloor\log_{2}\ell(t)\rfloor} for integers log2(d)tlog2(n)\lfloor\log_{2}(d^{*})\rfloor\leq t\leq\log_{2}(n) and (t)=ε𝔼[X]/(4KΦX(δ+ε)2t)\ell(t)=\varepsilon\mathbb{E}[X]/(4K\cdot\Phi_{X}(\delta+\varepsilon)2^{t}). We denote the number of such options by TmT_{m}.

The second option is that the edge has an endpoint in Ulog2(r)(H)U_{\lfloor\log_{2}(r)\rfloor}(H) where rr is defined as r=ε𝔼[X]/(8KΦX(δ+ε)np2)r=\varepsilon\mathbb{E}[X]/\left(8K\cdot\Phi_{X}(\delta+\varepsilon)np^{2}\right). That happens when the edge satisfies (16). We denote the number of such options by SmS_{m}.

Furthermore, note that provided nn is large enough there are at most

(log2(n)log2(d)+2)vm=(log2(log(1/p)C0p2)+2)vm3vmlog(1/p)vm{\left(\log_{2}(n)-\log_{2}(d^{*})+2\right)^{v_{m}}}={\left(\log_{2}\left(\frac{\log(1/p)}{C_{0}p^{2}}\right)+2\right)^{v_{m}}}\leq 3^{v_{m}}\log(1/p)^{v_{m}}

partitions of the non-isolated vertices of our graph into sets VtV_{t} where log2(d)tlog2(n)\lfloor\log_{2}(d^{*})\rfloor\leq t\leq\log_{2}(n) is an integer and t=0log2(d)1Vt\cup_{t=0}^{\lfloor\log_{2}(d^{*})\rfloor-1}V_{t}. We will think of the vertices in VtV_{t} as vertices which will have degrees between 2t2^{t} and 2t+12^{t+1} where log2(d)tlog2(n)\lfloor\log_{2}(d^{*})\rfloor\leq t\leq\log_{2}(n) is an integer.

We conclude that given the vertex set and a partition as explained above, there are at most Tm+SmT_{m}+S_{m} pairs of vertices that can be edges of HH. Hence, the number of graphs with property ()(*) is bounded from above by

(3log(1/p))vm2vm(nvm)(Sm+Tmm)(6log(1/p))vm(nvm)(Sm+Tmm).{\left(3\log(1/p)\right)^{v_{m}}}2^{v_{m}}\binom{n}{v_{m}}\binom{S_{m}+T_{m}}{m}\leq(6\log(1/p))^{v_{m}}\binom{n}{v_{m}}\binom{S_{m}+T_{m}}{m}.

Let us now estimate SmS_{m} and TmT_{m}.

Tm=\displaystyle T_{m}= t=log2(d)log2(n)|Vt||Ulog2((t))|t=log2(d)log2(n)m2t14m(t)=t=log2(d)log2(n)m22t34KΦX(δ+ε)2tε𝔼[X]\displaystyle\sum\limits_{t=\lfloor\log_{2}(d^{*})\rfloor}^{\log_{2}(n)}|V_{t}||U_{\lfloor{\log_{2}(\ell(t))}\rfloor}|\leq\sum\limits_{t=\lfloor\log_{2}(d^{*})\rfloor}^{\log_{2}(n)}\frac{m}{2^{t-1}}\frac{4m}{\ell(t)}=\sum\limits_{t=\lfloor\log_{2}(d^{*})\rfloor}^{\log_{2}(n)}\frac{m^{2}}{2^{t-3}}\frac{4K\cdot\Phi_{X}(\delta+\varepsilon)2^{t}}{\varepsilon\mathbb{E}[X]}
\displaystyle\leq (log2(n)log2(d)+1)64KΦX(δ+ε)m2ε𝔼[X]192KΦX(δ+ε)m2log(1/p)ε𝔼[X].\displaystyle(\log_{2}(n)-\lfloor\log_{2}(d^{*})\rfloor+1)\frac{64K\Phi_{X}(\delta+\varepsilon)m^{2}}{\varepsilon\mathbb{E}[X]}\leq\frac{192K\Phi_{X}(\delta+\varepsilon)m^{2}\log(1/p)}{\varepsilon\mathbb{E}[X]}.

Recalling the assumption that mKΦX(δ+ε)m\leq K\Phi_{X}(\delta+\varepsilon) we obtain,

Tm192K3ΦX3(δ+ε)log(1/p)ε𝔼[X].T_{m}\leq\frac{192K^{3}\Phi^{3}_{X}(\delta+\varepsilon)\log(1/p)}{\varepsilon\mathbb{E}[X]}.

Applying Lemma 5.1 we find that there is C1>0C_{1}>0 such that,

ΦX(δ+ε)C1𝔼[X]n2p2log(1/p).\Phi_{X}(\delta+\varepsilon)\leq\frac{C_{1}\mathbb{E}[X]}{n^{2}p^{2}}\log(1/p).

By our assumptions, Cn2p2mΦX(δ+ε)=O(n2p2log(1/p))Cn^{2}p^{2}\leq m\leq\Phi_{X}(\delta+\varepsilon)=O(n^{2}p^{2}\log(1/p)). This implies there are C2,C3>0C_{2},C_{3}>0 such that for large enough n0n_{0} we have,

Tm\displaystyle T_{m}\leq C2n2p2log4(1/p)\displaystyle C_{2}n^{2}p^{2}\log^{4}(1/p)
\displaystyle\leq C3mlog5(1/p).\displaystyle C_{3}m\log^{5}(1/p).

For SmS_{m} recall that there are at most 4m/r4m/r vertices in Ulog2(r)(H)U_{\lfloor\log_{2}(r)\rfloor}(H) and further recall that ΦX(δ+ε)=O(n2p2log(1/p))\Phi_{X}(\delta+\varepsilon)=O(n^{2}p^{2}\log(1/p)). Hence,

Smn4mlog(1/p)rC4mlog(1/p),S_{m}\leq n\frac{4m\log(1/p)}{r}\leq C_{4}m\log(1/p),

where C4=C4(ε,δ,K)>0C_{4}=C_{4}(\varepsilon,\delta,K)>0. This implies that for large enough n0n_{0} we have Sm,Tm=O(mlog5(1/p))S_{m},T_{m}=O(m\log^{5}(1/p)). Putting it all together there exists C5>0C_{5}>0 such that provided n0n_{0} is large enough we have,

|Cm|/(nvm)\displaystyle|C_{m}|/\binom{n}{v_{m}}\leq (6log(1/p))vm(Tm+Smm)\displaystyle{(6\log(1/p))}^{v_{m}}\binom{T_{m}+S_{m}}{m}
\displaystyle\leq (6log2(1/p))2m(C5mlog5(1/p)m)m\displaystyle{(6\log_{2}(1/p))}^{2m}\left(\frac{C_{5}m\log^{5}(1/p)}{m}\right)^{m}
\displaystyle\leq log(1/p)O(m).\displaystyle{\log(1/p)}^{O(m)}.

This proves the lemma. ∎

We now deduce two corollaries and discuss how one would use them in order to obtain an upper bound on the upper tail probability of XX. The first corollary will be used when n1pn1/2n^{-1}\ll p\ll n^{-1/2} and the second when n1/2p1n^{-1/2}\ll p\ll 1.

Corollary 5.6.

Suppose ε,δ,C,K\varepsilon,\delta,C,K are positive reals with ε<1\varepsilon<1. Furthermore, suppose n1pn1/2n^{-1}\ll p\ll n^{-1/2} as nn tends to infinity. Then, there exist D>0D>0 and n0n_{0} such that the following holds for all n>n0n>n_{0}:
Let mm be a positive integer with Cn2p2mKΦX(δ+ε)Cn^{2}p^{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon). Furthermore, let

vm=max{|{vV(G):deg(v)0}|:G𝒞m(ε,δ,K)}.v_{m}=\max\{|\{v\in V(G):\deg(v)\neq 0\}|:G\in\mathcal{C}_{m}(\varepsilon,\delta,K)\}.

Then,

log(|𝒞m|)vmlog(np2)+ηmlog(n).\log\left(|\mathcal{C}_{m}|\right)\leq-v_{m}\log\left(np^{2}\right)+\eta m\log(n).
Proof.

By Lemma 5.2 and the assumption that pn1p\gg n^{-1} we have the following provided n0n_{0} is sufficiently large,

log(|𝒞m|)log(nvm)+Dmloglog(1/p)vmlog(en/vm)+ηmlog(n)/3.\log(|\mathcal{C}_{m}|)\leq\log\binom{n}{v_{m}}+Dm\log\log(1/p)\leq v_{m}\log(en/v_{m})+\eta m\log(n)/3.

Since, vm2mv_{m}\leq 2m we deduce that

log(|𝒞m|)vmlog(n/vm)+2ηmlog(n)/3.\log(|\mathcal{C}_{m}|)\leq v_{m}\log(n/v_{m})+2\eta m\log(n)/3.

We now split into two cases. First, assume vmηm/3v_{m}\leq\eta m/3. This assumption implies that

vmlog(n/vm)ηmlog(n)/3.v_{m}\log(n/v_{m})\leq\eta m\log(n)/3.

On the other hand, if we assume vmηm/3v_{m}\geq\eta m/3 we obtain

vmlog(n/vm)vmlog(3n/ηm)=vmlog(np2)+vmlog(O(1)),v_{m}\log(n/v_{m})\leq v_{m}\log(3n/\eta m)=-v_{m}\log(np^{2})+v_{m}\log(O(1)),

where the equality is due to the assumption that m=Ω(n2p2)m=\Omega(n^{2}p^{2}). Since vm2mv_{m}\leq 2m, in both cases we obtain that, provided nn is large enough, the following holds:

log(|𝒞m|)vmlog(np2)+ηmlog(n).\log\left(|\mathcal{C}_{m}|\right)\leq-v_{m}\log\left(np^{2}\right)+\eta m\log(n).\qed

Corollary 5.6 suggests the following ‘plan of attack’ which we implement in Section 6. Assuming pn1/2p\ll n^{-1/2} the term log(np2)\log(np^{2}) is negative. Therefore in order to estimate |𝒞m||\mathcal{C}_{m}| estimate the maximum number of vertices that a core with mm edges can have for each mm. This estimation and Corollary 5.6 then yield an upper bound on the number of cores with mm edges, denote this bound by βm\beta_{m}. Then we compare pmβmp^{m}\beta_{m} for every m0mKΦX(δ)m_{0}\leq m\leq K\Phi_{X}(\delta), where m0m_{0} is the minimum number of edges in a core graph and denote this maximum by β\beta^{*}. This implies,

m=m0KΦδ(X)pm|𝒞m|KΦX(δ)β.\sum\limits_{m=m_{0}}^{K\Phi_{\delta}(X)}p^{m}|\mathcal{C}_{m}|\leq K\Phi_{X}(\delta)\beta^{*}.

We also show that KΦX(δ)K\Phi_{X}(\delta) is negligible compared to β\beta^{*} and therefore we obtain,

log(m=m0KΦX(δ)pm|𝒞m|)(1η)log(β).\log\left(\sum\limits_{m=m_{0}}^{K\Phi_{X}(\delta)}p^{m}|\mathcal{C}_{m}|\right)\leq(1-\eta)\log(\beta^{*}).

Then Theorem 3.2 will imply

log(X(1+δ)𝔼[X])(1ε)log(β).\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1-\varepsilon)\log(\beta^{*}).

Note that β\beta^{*} depends on pp. We will also show that this upper bound is matched to the lower bounds that we obtained in Section 4.

Next we present a corollary of Lemma 5.2 which will be used in the range where pn1/2p\gg n^{-1/2}.

Corollary 5.7.

Suppose ε,δ,C,K\varepsilon,\delta,C,K are positive reals with ε<1\varepsilon<1. Furthermore, suppose n1/2p1n^{-1/2}\ll p\ll 1 as nn tends to infinity. Then, there exists n0n_{0} such that the following holds for all n>n0n>n_{0}:
Let mm be a positive integer with Cn2p2mKΦX(δ+ε)Cn^{2}p^{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon). Then,

|𝒞m|(1p)εm.|\mathcal{C}_{m}|\leq\left(\frac{1}{p}\right)^{\varepsilon m}.

This is also an immediate corollary of Lemma 5.2. To see this note that in this range of pp we have nmn\ll m and thus (nvm)2nlog(1/p)m\binom{n}{v_{m}}\leq 2^{n}\leq\log(1/p)^{m}. Hence Lemma 5.2 implies

log|𝒞m|O(mloglog(1/p))=o(mlog(1/p)).\log|\mathcal{C}_{m}|\leq O(m\log\log(1/p))=o(m\log(1/p)).

Similar to before ΦX(δ+ε)\Phi_{X}(\delta+\varepsilon) will be shown to be negligible and therefore Corollary 5.7 will yield the following provided pn1/2p\gg n^{-1/2},

log(m=mKΦX(δ)pm|𝒞m|)(1ε)mlog(p)\log\left(\sum_{m=m_{*}}^{K\Phi_{X}(\delta)}p^{m}|\mathcal{C}_{m}|\right)\leq(1-\varepsilon)m_{*}\log\left(p\right)

where mm_{*} is the minimum number of edges in a core graph. Which then by Theorem 3.2 implies

log(X(1+δ)𝔼[X])(1ε)mlog(p).\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1-\varepsilon)m_{*}\log(p).

In the next section we will compute this mm_{*} and show that the above matches the lower bounds given in Section 4.

6. Upper bounds

As can be seen in previous sections there is a big difference in the behaviour of the problem depending on the regime pp lies at. As explained in Section 5, in order to obtain quantitative bounds on the upper tail probability we need to bound the entropic term |𝒞m||\mathcal{C}_{m}| when log9(n)np1nlog(n)\frac{\log^{9}(n)}{n}\ll p\ll\frac{1}{\sqrt{n}\log(n)}, and when pn1/2p\gg n^{-1/2} we only need to compute the minimum number of edges in a core graph. Actually, the situation is more involved in the sparse regime where we see a surprising change in the behaviour of the problem. To present the main result of this section let us define a sequence ckc_{k} for k2k\geq 2:

ck={0 for k=1,12+k+1k1 for k2.c_{k}=\Bigg{\{}\begin{array}[]{@{}l@{\thinspace}l}0&\text{ for }k=1,\\ \frac{1}{2+\sqrt{\frac{k+1}{k-1}}}&\text{ for }k\geq 2.\end{array}

We note that ckc_{k} is an increasing sequence and limkck=1/3\lim_{k\rightarrow\infty}c_{k}=1/3. This is the sequence promised in the our main theorem 1.1 and in Section 4. Furthermore, recall the definitions of mkm_{k} and mm_{*} which were given at the beginning of Section 4. For the convenience of the reader we give these definitions here as well. First, we define mk=rk(δ+ε)𝔼[X]m_{k}=r_{k}\sqrt{(\delta+\varepsilon)\mathbb{E}[X]}, where

rk={2 for k=0,2k/(k1) for k2.r_{k}=\Bigg{\{}\begin{array}[]{@{}l@{\thinspace}l}2&\text{ for }k=0,\\ 2\sqrt{k/(k-1)}&\text{ for }k\geq 2.\\ \end{array}

Second, we define

m=(r+d2d)𝔼[X]/2,m_{*}=(\sqrt{r+d^{2}}-d)\sqrt{\mathbb{E}[X]}/2,

where r=16(δ+3ε/2)r=16(\delta+3\varepsilon/2) and d=2/(1+ε)d=\sqrt{2}/(1+\varepsilon). We are now ready to state the main theorem of this section which is the following.

Theorem 6.1.

Suppose kk is an integer greater than 11 and suppose that ε,δ\varepsilon,\delta are positive reals with ε\varepsilon small enough. Then there exists n0n_{0} such that for any n>n0n>n_{0} the following holds:

  1. (1)

    If max{log9(n)n,n1+ck1}pn1+ck\max\left\{\frac{\log^{9}(n)}{n},n^{-1+c_{k-1}}\right\}\leq p\leq n^{-1+c_{k}}, then

    log(X(1+δ)𝔼[X])(1ε)log(pmk(nmk/k)).\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq{(1-\varepsilon)}\log\left(p^{m_{k}}\binom{n}{m_{k}/{k}}\right).
  2. (2)

    If n2/3p1nlog(n)n^{-2/3}\leq p\leq{\frac{1}{\sqrt{n}\log(n)}}, then

    log(X(1+δ)𝔼[X])(1ε)m0log(p).\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq{(1-\varepsilon)}m_{0}\log\left(p\right).
  3. (3)

    If n1/2p1n^{-1/2}\ll p\ll 1, then

    log(X(1+δ)𝔼[X])(1ε)mlog(p).\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq{(1-\varepsilon)}m_{*}\log(p).

The proof naturally splits into three parts. We will prove each part in a different subsection. Before we split into cases let us prove a lemma which makes a connection between the number of induced copies of C4C_{4} in a graph to the numbers of edges and vertices of that graph. This lemma will be important for us both in the sparse regime and the dense regime.

Lemma 6.2.

Suppose n,mn,m are positive integers such that m>3m>3 and let γ0\gamma\geq 0 be real. Let GG be a graph with nn vertices and mm edges and assume δ(G)2\delta(G)\geq 2. Then,

Nind(C4,G)m(mn+1)4.N_{ind}(C_{4},G)\leq\frac{m(m-n+1)}{4}.
Remark.

The bound from the lemma is sharp when n=mk+kn=\frac{m}{k}+k for some kk\in\mathbb{N} as the graph Kk,mkK_{k,\frac{m}{k}} has nn vertices and mm edges and contains (k2)(mk2)=m4(mkmk+1)\binom{k}{2}\binom{\frac{m}{k}}{2}=\frac{m}{4}(m-k-\frac{m}{k}+1) induced copies of C4C_{4}.

Proof of Lemma 6.2.

Since an nn-vertex graph with minimum degree at least 22 has at least nn and at most (n2)\binom{n}{2} edges we may assume that nm(n2)n\leq m\leq\binom{n}{2}. Since,

Nind(C4,G)=14eE(G)Nind(e,C4,G)m4maxeE(G)Nind(e,C4,G).N_{ind}(C_{4},G)=\frac{1}{4}\sum_{e\in E(G)}N_{ind}(e,C_{4},G)\leq\frac{m}{4}\max_{e\in E(G)}N_{ind}(e,C_{4},G).

We only need to prove that Nind(uv,C4,G)mn+1N_{ind}(uv,C_{4},G)\leq m-n+1 for all uvE(G)uv\in E(G). Note that every C4C_{4} containing uvuv is determined by the ‘parallel’ edge to uvuv meaning the edge between the two other vertices in that C4C_{4}. This is because we are considering induced copies. In other words, the number of induced copies containing uvuv is bounded by the number of edges between N(u){v}N(u)\setminus\{v\} and N(v){u}N(v)\setminus\{u\}. Let XX be the set of vertices not in N(u)N(v)N(u)\cup N(v) and note that |X|ndeg(u)deg(v)|X|\geq n-\deg(u)-\deg(v). The number of edges between N(u){v}N(u)\setminus\{v\} and N(v){u}N(v)\setminus\{u\} is bounded from above by me(X)(deg(u)+deg(v)1)m-e(X)-(\deg(u)+\deg(v)-1) where e(X)e(X) is the number of edges with an endpoint in XX. Since δ(G)2\delta(G)\geq 2 we may estimate e(X)e(X) as follows,

e(X)12xXdeg(x)|X|.e(X)\geq\frac{1}{2}\sum\limits_{x\in X}\deg(x)\geq|X|.

Thus, the number of edges between N(u){v}N(u)\setminus\{v\} and N(v){u}N(v)\setminus\{u\} is bounded from above by

m(|X|+deg(v)+deg(u)1)mn+1m-(|X|+\deg(v)+\deg(u)-1)\leq m-n+1

and the lemma follows. ∎

In order to prove Theorem 6.1 we now split into cases, each dealing with each item of the theorem.

6.1. The dense regime

Assume that n1/2p1n^{-1/2}\ll p\ll 1 and assume that ε\varepsilon is any positive real which is small enough as a function of δ\delta. Corollary 5.7 asserts that in this regime of pp the number of core graphs is negligible. Therefore, as explained at the end of Section 5, to bound the upper tail probability using Theorem 3.2 we are only left with showing that mm_{*} is a lower bound on the minimum number of edges in a core. This is given in the following lemma.

Lemma 6.3.

Suppose ε,δ\varepsilon,\delta are positive real numbers with ε\varepsilon small enough as a function of δ\delta. Further suppose pn1/2p\gg n^{-1/2}. Then, there exists n0n_{0} such that for all n>n0n>n_{0}

min{e(G):G𝒞}m.\min\{e(G):G\in\mathcal{C}\}\geq m_{*}.

Before proving the claim let us recall Observation 5.4. For the convenience of the reader we state it here as well:

Observation 6.4.

Suppose n,mn,m are positive integers with m(n2)m\leq\binom{n}{2}. Then,

Nind(m,n,K1,2K1)mn2/8.N_{ind}(m,n,K_{1,2}\sqcup K_{1})\leq mn^{2}/8.
Proof of Lemma 6.3.

Suppose G𝒞mG\in\mathcal{C}_{m} is a core graph with mm edges. Then GG is also a structured seed and thus

Nind(C4,G)+Nind(K1,2K1,G)p2(δε)𝔼[X].N_{ind}(C_{4},G)+N_{ind}(K_{1,2}\sqcup K_{1},G)p^{2}\geq(\delta-\varepsilon)\mathbb{E}[X].

Lemma 6.2 and Observation 6.4 assert that

Nind(C4,G)Nind(m,C4)m2/4,N_{ind}(C_{4},G)\leq N_{ind}(m,C_{4})\leq m^{2}/4,

and

Nind(K1,2K1,G)Nind(m,n,K1,2K1)mn2/8.N_{ind}(K_{1,2}\sqcup K_{1},G)\leq N_{ind}(m,n,K_{1,2}\sqcup K_{1})\leq mn^{2}/8.

Therefore,

m2+mn2p2/24(δε)𝔼[X]0.m^{2}+mn^{2}p^{2}/2-4(\delta-\varepsilon)\mathbb{E}[X]\geq 0.

Taking n0n_{0} large enough we have 𝔼[X](1ε)2n4p4/8\mathbb{E}[X]\geq(1-\varepsilon)^{2}n^{4}p^{4}/8 and thus,

m2+m2𝔼[X](1ε)4(δε)𝔼[X]0.m^{2}+\frac{m\sqrt{2\mathbb{E}[X]}}{(1-\varepsilon)}-4(\delta-\varepsilon)\mathbb{E}[X]\geq 0.

This implies that for large enough n0n_{0} and small enough ε\varepsilon we have

m(2(1ε)2+16(δε)2(1ε))𝔼[X]/2=m.m\geq\left(\sqrt{\frac{2}{(1-\varepsilon)^{2}}+16(\delta-\varepsilon)}-\frac{\sqrt{2}}{(1-\varepsilon)}\right)\sqrt{\mathbb{E}[X]}/2=m_{*}.\qed

As explained after Corollary 5.7, this implies the following corollary which is the third item in Theorem 6.1.

Corollary 6.5.

Suppose ε,δ\varepsilon,\delta are positive real numbers with ε\varepsilon small enough as a function of δ\delta. Further suppose n1/2p1n^{-1/2}\ll p\ll 1. Then, for large enough nn we have

log(X(1+δ)𝔼[X])(1ε)mlog(p).\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq{(1-\varepsilon)}m_{*}\log(p).

6.2. The sparse regime

Throughout this subsection, assume log9(n)np1nlog(n)\frac{\log^{9}(n)}{n}\leq p\ll\frac{1}{\sqrt{n}\log(n)}. As mentioned earlier, there is another change in the behaviour of the problem when log9(n)npn2/3ε\frac{\log^{9}(n)}{n}\leq p\leq n^{-2/3-\varepsilon} and n2/3p1nlog(n)n^{-2/3}\leq p\ll\frac{1}{\sqrt{n}\log(n)}. Before splitting into these two cases we show a reduction and develop some tools which we use in both cases.

We start with the following fact which plays a major role in the later proofs.

Lemma 6.6.

Suppose ε,δ,s,K\varepsilon,\delta,s,K are reals and positive with ε<1\varepsilon<1. Furthermore, suppose log(n)np1nlog(n)\frac{\log(n)}{n}\leq p\ll\frac{1}{\sqrt{n}\log(n)}. Then, there exist n0n_{0} and ξ>0\xi>0 such that the following holds for all n>n0n>n_{0}:
Suppose G𝒞mG\in\mathcal{C}_{m} where Cn2p2mKΦX(δ+ε)Cn^{2}p^{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon) and uvE(G)uv\in E(G). Then,

deg(u)deg(v)ξn2p2log(1/p)s,\deg(u)\deg(v)\geq\frac{\xi n^{2}p^{2}}{\log(1/p)}\geq s,

and furthermore, deg(u),deg(v)2\deg(u),\deg(v)\geq 2.

Proof.

Let G𝒞mG\in\mathcal{C}_{m} be a core graph with Cn2p2mKΦX(δ+ε)Cn^{2}p^{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon) edges and let uvE(G)uv\in E(G) be an edge of GG. By the definition of 𝒞m\mathcal{C}_{m} we have the following:

Nind(uv,C4,G)+Nind(uv,K1,2,G)np2ε𝔼[X]2KΦX(δ+ε).N_{ind}(uv,C_{4},G)+N_{ind}(uv,K_{1,2},G)np^{2}\geq\frac{\varepsilon\mathbb{E}[X]}{2K\Phi_{X}(\delta+\varepsilon)}.

Note that Nind(uv,K1,2,G)mKΦX(δ+ε)N_{ind}(uv,K_{1,2},G)\leq m\leq K\Phi_{X}(\delta+\varepsilon). Lemma 5.1 asserts that there exists D>0D>0 such that provided n0n_{0} is large enough we have

ΦX(δ+ε)D𝔼[X]log(1/p).\Phi_{X}(\delta+\varepsilon)\leq D\sqrt{\mathbb{E}[X]}\log(1/p).

Hence as p1nlog(n)p\ll\frac{1}{\sqrt{n}\log(n)} we deduce the following for large enough n0n_{0}:

Nind(uv,K1,2,G)np2KD𝔼[X]np2log(1/p)𝔼[X]/log(1/p),N_{ind}(uv,K_{1,2},G)np^{2}\leq KD\sqrt{\mathbb{E}[X]}np^{2}\log(1/p)\ll\sqrt{\mathbb{E}[X]}/\log(1/p),

and

Nind(uv,C4,G)ε𝔼[X]2DKlog(1/p)Nind(uv,K1,2,G)np2ε𝔼[X]4DKlog(1/p).N_{ind}(uv,C_{4},G)\geq\frac{\varepsilon\sqrt{\mathbb{E}[X]}}{2DK\log(1/p)}-N_{ind}(uv,K_{1,2},G)np^{2}\geq\frac{\varepsilon\sqrt{\mathbb{E}[X]}}{4DK\log(1/p)}. (17)

Note also that deg(u)deg(v)Nind(C4,uv,G)\deg(u)\deg(v)\geq N_{ind}(C_{4},uv,G) and thus the assertion follows for small enough ξ\xi,

deg(u)deg(v)ε𝔼[X]4DKlog(1/p)ξn2p2log(1/p)\deg(u)\deg(v)\geq\frac{\varepsilon\sqrt{\mathbb{E}[X]}}{4DK\log(1/p)}\geq\frac{\xi n^{2}p^{2}}{\log(1/p)}

and if deg(u)=1\deg(u)=1 or deg(v)=1\deg(v)=1 then Nind(C4,uv,G)=0N_{ind}(C_{4},uv,G)=0 which is a contradiction to (17). ∎

The above lemma will be used many times throughout this subsection, mostly the fact that when log9(n)np1nlog(n)\frac{\log^{9}(n)}{n}\leq p\ll\frac{1}{\sqrt{n}\log(n)} the minimum degree of a core graph is at least 22. We will omit the reference to this lemma and keep in mind that the minimum degree of all graphs considered is at least 22.

The following is a corollary which bounds vm=max|{vV(G):G𝒞m and deg(v)>0}|v_{m}=\max|\{v\in V(G):G\in\mathcal{C}_{m}\text{ and }\deg(v)>0\}|. This will be used afterwards to formalize the discussion after Corollary 5.6.

Lemma 6.7.

Suppose ε,δ,K,γ\varepsilon,\delta,K,\gamma are positive reals. Suppose further that log9(n)np1nlog(n)\frac{\log^{9}(n)}{n}\leq p\ll\frac{1}{\sqrt{n}\log(n)}. Then, there exists an integer n0n_{0} such that for all integers n>n0n>n_{0} the following holds:
Suppose mm is an integer with m0mKΦX(δ+ε)m_{0}\leq m\leq K\Phi_{X}(\delta+\varepsilon). Then,

vmm2+m3/4.v_{m}\leq\frac{m}{2}+m^{3/4}.
Proof.

Fix an arbitrary G𝒞mG\in\mathcal{C}_{m} and define, A={vV(G):deg(v)<m/log2(n)}A=\{v\in V(G):\deg(v)<\sqrt{m}/\log^{2}(n)\}. We have

2m=vV(G)deg(v)=vAdeg(v)+vAcdeg(v)|Ac|m/log2(n).2m=\sum_{v\in V(G)}\deg(v)=\sum_{v\in A}\deg(v)+\sum_{v\in A^{c}}\deg(v)\geq|A^{c}|\sqrt{m}/\log^{2}(n).

Thus, |Ac|2mlog2(n)|A^{c}|\leq 2\sqrt{m}\cdot\log^{2}(n).

By Theorem 5.1 we have ΦX(δ+ε)=O(n2p2log(1/p))\Phi_{X}(\delta+\varepsilon)=O(n^{2}p^{2}\log(1/p)) and hence,

(mlog2(n))2O(n2p2log(1/p))log4(n)n2p2log(1/p).\left(\frac{\sqrt{m}}{\log^{2}(n)}\right)^{2}\leq\frac{O(n^{2}p^{2}\log(1/p))}{\log^{4}(n)}\ll\frac{n^{2}p^{2}}{\log(1/p)}.

Thus, by Lemma 6.6 for sufficiently large n0n_{0} no two vertices of AA can be connected by an edge. Therefore,

mvAdeg(v).m\geq\sum_{v\in A}\deg(v).

Since deg(v)2\deg(v)\geq 2 for all of the vertices of GG, we obtain |A|m2|A|\leq\frac{m}{2}.

Therefore for large enough n0n_{0},

|V(G)|=|AAc|=|A|+|Ac|m2+2mlog2(n).|V(G)|=|A\sqcup A^{c}|=|A|+|A^{c}|\leq\frac{m}{2}+2\sqrt{m}\cdot\log^{2}(n).

Provided n0n_{0} is large enough, 2mlog2(n)m3/42\sqrt{m}\log^{2}(n)\leq m^{3/4}. Hence, for large enough n0n_{0} the assertion of the lemma holds. ∎

We now formalize the discussion after Corollary 5.6 in the following proposition.

Proposition 6.8.

Suppose δ\delta is a positive real and ε,η\varepsilon,\eta are sufficiently small positive reals. Furthermore, suppose log9(n)np1nlog(n)\frac{\log^{9}(n)}{n}\leq p\ll\frac{1}{\sqrt{n}\log(n)}. Then, there exists n0n_{0} such that for all n>n0n>n_{0} we have,

log(X(1+δ)𝔼[X])(1η)maxm0mm2{mlog(p)vmlog(np2)}.\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1-\eta)\max_{m_{0}\leq m\leq m_{2}}\{m\log(p)-v_{m}\log(np^{2})\}.

To prove this proposition we start with a lemma reducing the problem of estimating mlog(p)vmlog(np2)m\log(p)-v_{m}\log(np^{2}) only when m0mKΦX(δ+ε)m_{0}\leq m\leq K\Phi_{X}(\delta+\varepsilon).

Lemma 6.9.

Suppose δ\delta is a positive real and ε,η\varepsilon,\eta are sufficiently small positive reals. Furthermore, suppose n1pn1/2/log(n)n^{-1}\ll p\ll n^{-1/2}/\log(n). Then, there exists K,n0>0K,n_{0}>0 such that for all n>n0n>n_{0} we have,

log(X(1+δ)𝔼[X])(1η)maxm0mKΦX(δ+ε){mlog(p)vmlog(np2)}.\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1-\eta)\max_{m_{0}\leq m\leq K\Phi_{X}(\delta+\varepsilon)}\left\{m\log(p)-v_{m}\log(np^{2})\right\}.
Proof.

For all η1,K>0\eta_{1},K^{\prime}>0 Corollary 5.6 asserts that provided n0n_{0} is large enough we have the following for all m0mKΦX(δ+ε)m_{0}\leq m\leq K^{\prime}\Phi_{X}(\delta+\varepsilon):

log(pm|𝒞m|)mlog(p)vmlog(np2)+η1mlog(n),\log(p^{m}|\mathcal{C}_{m}|)\leq m\log(p)-v_{m}\log(np^{2})+\eta_{1}m\log(n),

where vm=max|{vV(G):G𝒞m and deg(v)>0}|v_{m}=\max|\{v\in V(G):G\in\mathcal{C}_{m}\text{ and }\deg(v)>0\}|.

Since n1p1nlog(n)n^{-1}\ll p\ll\frac{1}{\sqrt{n}\log(n)}, provided nn is sufficiently large, we have vmm/2+m3/4v_{m}\leq m/2+m^{3/4} for all m0mKΦX(δ+ε)m_{0}\leq m\leq K^{\prime}\Phi_{X}(\delta+\varepsilon) by Lemma 6.7. Thus,

mlog(p)vmlog(np2)Ω(mlog(n)).m\log(p)-v_{m}\log(np^{2})\leq-\Omega\left(m\log(n)\right).

For any positive η2\eta_{2} we may choose η1\eta_{1} small enough such that we have the following for sufficiently large nn,

mlog(p)vmlog(np2)+η1mlog(n)(1η2)(mlog(p)vmlog(np2)).m\log(p)-v_{m}\log(np^{2})+\eta_{1}m\log(n)\leq(1-\eta_{2})(m\log(p)-v_{m}\log(np^{2})).

Therefore, provided η2\eta_{2} is sufficiently small Theorem 3.2 and Theorem 5.1 then yield that there exist K,n0>0K,n_{0}>0 such that for all n>n0n>n_{0}:

log(X(1+δ)𝔼[X])(1η)maxm0mKΦX(δ+ε){mlog(p)vmlog(np2)}.\log\mathbb{P}(X\geq(1+\delta)\mathbb{E}[X])\leq(1-\eta)\max_{m_{0}\leq m\leq K\Phi_{X}(\delta+\varepsilon)}\{m\log(p)-v_{m}\log(np^{2})\}.\qed

We now use the above lemma to prove Proposition 6.8

Proof of Proposition 6.8.

By Lemma 6.9 it is sufficient to prove that for all m2mKΦX(δ+ε)m_{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon) we have the following provided η\eta^{\prime} is sufficiently small

mlog(p)vmlog(np2)(1η)(m2log(p)vm2log(np2)).m\log(p)-v_{m}\log(np^{2})\leq(1-\eta^{\prime})(m_{2}\log(p)-v_{m_{2}}\log(np^{2})).

Provided n0n_{0} is large enough Lemma 6.7 gives vmm/2+m3/4m/2+ηmv_{m}\leq m/2+m^{3/4}\leq m/2+\eta^{\prime}m for all m2mKΦX(δ+ε)m_{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon). Therefore,

mlog(p)vmlog(np2)\displaystyle m\log(p)-v_{m}\log(np^{2}) mlog(p)(1/2+η)mlog(np2)\displaystyle\leq m\log(p)-(1/2+\eta^{\prime})m\log\left(np^{2}\right)
(1/22η)mlog(n)(1/22η)m2log(n).\displaystyle\leq(-1/2-2\eta^{\prime})m\log(n)\leq(-1/2-2\eta^{\prime})m_{2}\log(n).

Noting that each copy of Km2/2,2K_{m_{2}/2,2} in KnK_{n} is a core graph with m2m_{2} edges we find that vm2m2/2v_{m_{2}}\geq m_{2}/2. Therefore,

m2log(p)vm2log(np2)m2log(p)m22log(np2)=m22log(n).m_{2}\log(p)-v_{m_{2}}\log(np^{2})\geq m_{2}\log(p)-\frac{m_{2}}{2}\log(np^{2})=-\frac{m_{2}}{2}\log(n).

These inequalities imply the following for all m2mKΦX(δ+ε)m_{2}\leq m\leq K\Phi_{X}(\delta+\varepsilon) and sufficiently small η′′\eta^{\prime\prime}

mlog(p)vmlog(np2)(1η′′)(m2log(p)vm2log(np2)).m\log(p)-v_{m}\log(np^{2})\leq(1-\eta^{\prime\prime})(m_{2}\log(p)-v_{m_{2}}\log(np^{2})).

This establish the proposition provided η′′\eta^{\prime\prime} is small enough. ∎

Proposition 6.8 and Corollary 5.6 show that to prove Theorem 6.1 when log9(n)np1nlog(n)\frac{\log^{9}(n)}{n}\leq p\ll\frac{1}{\sqrt{n}\log(n)} we only need to bound mlog(p)vmlog(np2)m\log(p)-v_{m}\log(np^{2}) for m0mm2m_{0}\leq m\leq m_{2}. Therefore, to prove the first item in Theorem 6.1 it is enough to prove that, when n1+ck1pn1+ckn^{-1+c_{k-1}}\leq p\leq n^{-1+c_{k}} we have

mlog(p)vmlog(np2)(1η)(mklog(p)vmklog(np2)),m\log(p)-v_{m}\log(np^{2})\leq(1-\eta)(m_{k}\log(p)-v_{m_{k}}\log(np^{2})),

for all m0mm2m_{0}\leq m\leq m_{2}. Moreover, to prove the second item in Theorem 6.1 it is enough to prove that, when n2/3p1nlog(n)n^{-2/3}\leq p\ll\frac{1}{\sqrt{n}\log(n)} we have

mlog(p)vmlog(np2)(1η)m0log(p),m\log(p)-v_{m}\log(np^{2})\leq(1-\eta)m_{0}\log(p),

for all m0mm2m_{0}\leq m\leq m_{2}. We now split into these two cases mentioned above and prove them. First we deal with the denser case where n2/3p1nlog(n)n^{-2/3}\leq p\ll\frac{1}{\sqrt{n}\log(n)}.

6.2.1. The denser case in the sparse regime

Assume n2/3p1nlog(n)n^{-2/3}\leq p\ll\frac{1}{\sqrt{n}\log(n)}. Proposition 6.8 implies that we need to evaluate the term mlog(p)vmlog(np2)m\log(p)-v_{m}\log(np^{2}) only for m0mm2m_{0}\leq m\leq m_{2}. To do this we use Lemma 6.2 which implies a connection between the number of induced copies of C4C_{4} in a graph and the number of edges and vertices in it. We start with a simple corollary of Lemma 6.2.

Corollary 6.10.

Suppose ε,δ,K\varepsilon,\delta,K are positive reals. Suppose further that log(n)np1\frac{\log(n)}{n}\leq p\ll 1 and mm is an integer. Then, for any G𝒞mG\in\mathcal{C}_{m} the number of vertices of GG is at most

m4(δε)𝔼[X]/m+1.m-4(\delta-\varepsilon)\mathbb{E}[X]/m+1.
Proof.

Since the minimum degree of GG is at least 22, we may apply Lemma 6.2 and obtain,

(δε)𝔼[X]N(C4,G)Nind(v(G),m,C4)m(mv(G)+1)4(\delta-\varepsilon)\mathbb{E}[X]\leq N(C_{4},G)\leq N_{ind}(v(G),m,C_{4})\leq\frac{m(m-v(G)+1)}{4} (18)

By algebraic manipulations we see that, v(G)m4(δε)𝔼[X]/m+1v(G)\leq m-4(\delta-\varepsilon)\mathbb{E}[X]/m+1 which is the assertion of the corollary. ∎

Now we obtain the desired bound for mlog(p)vmlog(np2)m\log(p)-v_{m}\log(np^{2}) in this regime of pp.

Lemma 6.11.

Suppose that ε,δ,γ\varepsilon,\delta,\gamma are positive reals with ε\varepsilon small enough, and γ\gamma small enough as a function of ε\varepsilon. Furthermore, suppose n2/3p1nlog(n)n^{-2/3}\leq p\ll\frac{1}{\sqrt{n}\log(n)} and K(ε,δ)K(\varepsilon,\delta) is some constant that might depends on ε\varepsilon and δ\delta. Then there exists n0n_{0} such that for any n>n0n>n_{0} and any m0mm2m_{0}\leq m\leq m_{2} we have,

mlog(p)vmlog(np2)(1γ)m0log(p).m\log(p)-v_{m}\log(np^{2})\leq(1-\gamma)m_{0}\log(p).
Proof.

Let c[1/3,1/2]c\in[1/3,1/2] be such that p=n1+cp=n^{-1+c} and recall that m=Θ(n2p2)m=\Theta(n^{2}p^{2}). Applying Corollary 6.10 we obtain the following for any m0mm2m_{0}\leq m\leq m_{2} and sufficiently large n0n_{0}:

mlog(p)vmlog(np2)\displaystyle m\log(p)-v_{m}\log(np^{2}) (1+c)mlog(n)+(12c)(m4(δε)𝔼[X]/m+ηm)log(n)\displaystyle\leq(-1+c)m\log(n)+(1-2c)(m-4(\delta-\varepsilon)\mathbb{E}[X]/m+\eta m)\log(n)
cmlog(n)4(12c)(δε)𝔼[X]log(n)/m+ηmlog(n)/3.\displaystyle\leq-cm\log(n)-4(1-2c)(\delta-\varepsilon)\mathbb{E}[X]\log(n)/m+{\eta m\log(n)}/{3}.

Recalling that m0=2(δε)𝔼[X]m_{0}=2\sqrt{(\delta-\varepsilon)\mathbb{E}[X]} we may rewrite the above as follows provided n0n_{0} is large enough:

mlog(p)vmlog(np2)\displaystyle m\log(p)-v_{m}\log(np^{2}) (cm(12c)m02/m))log(n)+ηmlog(n)/3.\displaystyle\leq(-cm-(1-2c)m_{0}^{2}/m))\log(n)+\eta m\log(n)/3.

We claim that gc(m)cm(12c)m02/m(1+c)m0g_{c}(m)\coloneqq-cm-(1-2c)m_{0}^{2}/m\leq(-1+c)m_{0} for all m0mm2m_{0}\leq m\leq m_{2}. Indeed, gc(m)=c+(12c)(m0/m)213c0g^{\prime}_{c}(m)=-c+(1-2c)(m_{0}/m)^{2}\leq 1-3c\leq 0 for mm0m\geq m_{0} and c1/3c\geq 1/3. This concludes the proof as now provided γ\gamma is small enough we have the following for all m0mm2m_{0}\leq m\leq m_{2}:

mlog(p)vmlog(np2)\displaystyle m\log(p)-v_{m}\log(np^{2}) (1+c)m0log(n)+ηmlog(n)/3\displaystyle\leq(-1+c)m_{0}\log(n)+\eta m\log(n)/3
=m0log(p)+ηmlog(n)/3m0log(p)+ηm2log(n)/3\displaystyle=m_{0}\log(p)+\eta m\log(n)/3\leq m_{0}\log(p)+\eta m_{2}\log(n)/3
(1γ)m0log(p).\displaystyle\leq(1-\gamma)m_{0}\log(p).\qed

This together with Proposition 6.8 conclude the proof for the second item in Theorem 6.1.

6.2.2. The sparser case in the sparse regime

Assume log9(n)npn2/3ε\frac{\log^{9}(n)}{n}\leq p\ll n^{-2/3-\varepsilon}. It was already seen that there is a big difference in the behaviour of the upper tail probability depending on the regime pp lies at. In this section we present a surprising change of the behaviour in the sparser regime. This phenomenon comes from the fact that the number of core graphs is significant and matters a lot in the evaluation of the upper tail probability. This is completely different from the previous cases where the number of cores was negligible. This will be explained in more details later on.

The essence of the proofs in this subsection is still to make use of the connection between the number of induced copies of C4C_{4} in a core graph and the number of vertices in it. To state the claims in this section it is useful to introduce the following notations.

Notation.

Suppose GG is a graph and kk is a nonnegative integer. Then we denote the following sets accordingly:

  • Denote by Xk(G)X_{k}(G) the set of all edges of GG with an endpoint of degree kk and denote its cardinality by xk(G)x_{k}(G).

  • Denote by X>k(G)X_{>k}(G) the set of all edges of GG whose both endpoints have degree greater than kk; similarly we denote by x>k(G)x_{>k}(G) the cardinality of X>k(G)X_{>k}(G).

In most cases we omit ‘GG’ for brevity.

The following lemma bounds from above the number of induced copies of C4C_{4} in any core graph with mm edges where p1nlog(n)p\ll\frac{1}{\sqrt{n}\log(n)}.

Lemma 6.12.

Suppose ε,δ,R,K\varepsilon,\delta,R,K are positive reals and log(n)np1nlog(n)\frac{\log(n)}{n}\leq p\ll\frac{1}{\sqrt{n}\log(n)}. Then there exists n0n_{0} such that for any n>n0n>n_{0} the following holds:
Let GG be a core graph with mm edges. Then

Nind(C4,G)i=2Rxii(i2)(i<jxjj+xi2i)+i=2Rxie(G)(i1)R+x>R24,N_{ind}(C_{4},G)\leq\sum^{R}_{i=2}\frac{x_{i}}{i}\binom{i}{2}\left(\sum_{i<j}\frac{x_{j}}{j}+\frac{x_{i}}{2i}\right)+\sum^{R}_{i=2}\frac{x_{i}e(G)(i-1)}{R}+\frac{x_{>R}^{2}}{4}, (19)

and

v(G)i=2Rxii+2e(G)R.v(G)\leq\sum_{i=2}^{R}\frac{x_{i}}{i}+\frac{2e(G)}{R}. (20)
Proof.

Let GG be a core graph. First, let us make some notations for the proof. For every nonnegative integer kk let Vk(G)V_{k}(G) be set of all vertices of GG with degree kk and denote its cardinality by vk(G)v_{k}(G). Furthermore, let V>k(G)V_{>k}(G) be the set of all vertices of GG with degree greater than kk and denote its cardinality by v>k(G)v_{>k}(G).

We now proceed with the proof. Note that taking n0n_{0} large enough according to Lemma 6.6 applied with s=R2+1s=R^{2}+1, we are guaranteed that no two vertices of degree less or equal to RR can be connected by an edge. This implies that {Xi:iR}{X>R}\{X_{i}:i\leq R\}\cup\{X_{>R}\} is a partition of the edges of GG. Furthermore, v>R(G)2e(G)/Rv_{>R}(G)\leq 2e(G)/R and for any iRi\leq R we have vi=xi/iv_{i}=x_{i}/i. This follows by a standard double counting argument and as the set of all vertices of degree at most RR is an independent set. We thus obtain (20) as

v(G)=i=2Rvi(G)+v>R(G)i=2Rxii+2e(G)R.v(G)=\sum_{i=2}^{R}v_{i}(G)+v_{>R}(G)\leq\sum_{i=2}^{R}\frac{x_{i}}{i}+\frac{2e(G)}{R}.

To prove (19) we note that there are three types of induced copies of C4C_{4} in core graphs:

  1. (i)

    Induced copies of C4C_{4} with exactly two vertices in i=2RVi\cup_{i=2}^{R}V_{i}.

  2. (ii)

    Induced copies of C4C_{4} with exactly one vertex in i=2RVi\cup_{i=2}^{R}V_{i}.

  3. (iii)

    Induced copies of C4C_{4} with no vertices in i=2RVi\cup_{i=2}^{R}V_{i} i.e. all vertices in V>RV_{>R}.

Let us bound from above the number of induced copies of C4C_{4} of each type. We claim that the number of induced copies of C4C_{4} of the first type is at most

i=2Rvi(i2)(i<jvj+vi2)=i=2Rxii(i2)(i<jxjj+xi2i).\displaystyle\sum^{R}_{i=2}v_{i}\binom{i}{2}\left(\sum_{i<j}v_{j}+\frac{v_{i}}{2}\right)=\sum^{R}_{i=2}\frac{x_{i}}{i}\binom{i}{2}\left(\sum_{i<j}\frac{x_{j}}{j}+\frac{x_{i}}{2i}\right).

Indeed, for every ii, there are at most (vi2)(i2)\binom{v_{i}}{2}\binom{i}{2} copies of C4C_{4} with two vertices in viv_{i} and, for all i<ji<j there are at most vivj(i2)v_{i}v_{j}\binom{i}{2} copies of C4C_{4} with one vertex in ViV_{i} and one vertex in VjV_{j}.

We now bound the number of induced copies of C4C_{4} of the second kind. Note that each such induced copy of C4C_{4} is determined by choosing its vertex of degree less or equal to RR, two of its neighbours and another vertex of degree larger than RR. Therefore, the number of induced copies of C4C_{4} of the second kind it at most

i=2Rviv>R(i2)=i=2Rxiv>R(i1)2i=2Rxie(G)(i1)R.\sum_{i=2}^{R}v_{i}v_{>R}\binom{i}{2}=\sum_{i=2}^{R}\frac{x_{i}v_{>R}(i-1)}{2}\leq\sum_{i=2}^{R}\frac{x_{i}e(G)(i-1)}{R}.

To bound from above the number of induced copies of C4C_{4} of the third kind we observe that each such induced copy of C4C_{4} is determined by each of the two perfect matchings in it. Since each induced copy of C4C_{4} of the third type contains only edges of X>RX_{>R}, there are at most 12(x>R2)x>R24\frac{1}{2}\binom{x_{>R}}{2}\leq\frac{x_{>R}^{2}}{4} such copies. Summing all of these bounds we obtain the assertion of the lemma. ∎

Our only use of Lemma 6.12 is when R=1/εR=\lceil 1/\varepsilon\rceil. Therefore, form now on we set R=1/εR=\left\lceil 1/\varepsilon\right\rceil. We will only consider core graphs GG with at most m2m_{2} edges, thus we may bound from above the right-hand side of (19) by

f(x1,x2,,xR,x>R)=\displaystyle f(x_{1},x_{2},\ldots,x_{R},x_{>R})= i=2R(i2)xii(i<jxjj+xi2i)+εm2i=2R(i1)xi+x>R24\displaystyle\sum^{R}_{i=2}\binom{i}{2}\frac{x_{i}}{i}\left(\sum_{i<j}\frac{x_{j}}{j}+\frac{x_{i}}{2i}\right)+\varepsilon m_{2}\sum^{R}_{i=2}{(i-1)x_{i}}+\frac{x_{>R}^{2}}{4}
=\displaystyle= i=2R(i1)xi(i<jRxj2j+xi4i+εm2)+x>R24.\displaystyle\sum^{R}_{i=2}{(i-1)x_{i}}\left(\sum_{i<j\leq R}\frac{x_{j}}{2j}+\frac{x_{i}}{4i}+\varepsilon m_{2}\right)+\frac{x_{>R}^{2}}{4}.

In particular, for every core graph GG with at most m2m_{2} edges,

(δε)𝔼[X]Nind(C4,G)f(x1(G),x2(G),,xR(G),x>R(G)).(\delta-\varepsilon)\mathbb{E}[X]\leq N_{ind}(C_{4},G)\leq f(x_{1}(G),x_{2}(G),\ldots,x_{R}(G),x_{>R}(G)).

Recall that in this range of pp we have log(np2)<0\log(np^{2})<0 and note that by Lemma 6.12 for any core graph GG the term e(G)log(p)v(G)log(np2)e(G)\log(p)-v(G)\log(np^{2}) can be bounded from above using xi(G)x_{i}(G) by

i=2R(xi(G)log(p)xi(G)log(np2)/i)+x>R(G)log(p)2e(G)log(np2)/R.\displaystyle\sum_{i=2}^{R}\left(x_{i}(G)\log(p)-x_{i}(G)\log(np^{2})/i\right)+x_{>R}(G)\log(p)-2e(G)\log(np^{2})/R.

Furthermore, for ε<η/2\varepsilon<\eta/2 we have,

2e(G)/R2e(G)εηe(G).2e(G)/R\leq 2e(G)\varepsilon\leq\eta e(G).

Therefore, letting n0n_{0} sufficiently large we find that,

maxm0mm2mlog(p)vmlog(np2)\displaystyle\max_{m_{0}\leq m\leq m_{2}}m\log(p)-v_{m}\log(np^{2})\leq maxGm=m0m2𝒞mi=2R(xi(G)log(p)xi(G)log(np2)/i)\displaystyle\max_{G\in\bigcup_{m=m_{0}}^{m_{2}}\mathcal{C}_{m}}\sum_{i=2}^{R}\left(x_{i}(G)\log(p)-x_{i}(G)\log(np^{2})/i\right)
+x>R(G)log(p)+ηm2log(n).\displaystyle+x_{>R}(G)\log(p)+\eta m_{2}\log(n). (21)

We now restate this bound in a more compact way. To this end we introduce the following notations. For every graph GG define the following:

  • x(G)R+1x(G)\in\mathbb{R}^{R+1} is the vector defined by xi(G)x_{i}(G) in its ii-th coordinate for 1iR1\leq i\leq R and x>R(G)x_{>R}(G) in the last coordinate.

  • uR+1u\in\mathbb{R}^{R+1} is defined by 0 as the first coordinate, log(p)log(np2)/i\log(p)-\log(np^{2})/i as the ii-th coordinate for 2iR2\leq i\leq R, and log(p)\log(p) in the last coordinate.

Using these notations we may rewrite (6.2.2) as follows,

maxm0mm2mlog(p)vmlog(np2)maxGm=m0m2𝒞mx(G),u+ηm2log(n).\displaystyle\max_{m_{0}\leq m\leq m_{2}}m\log(p)-v_{m}\log(np^{2})\leq\max_{G\in\cup_{m=m_{0}}^{m_{2}}\mathcal{C}_{m}}\langle x(G),u\rangle+\eta m_{2}\log(n). (22)

From now on we let t=(δε)𝔼[X]t=(\delta-\varepsilon)\mathbb{E}[X]. Since tf(x(G))t\leq f(x(G)) for every Gm=m0m2𝒞mG\in\bigcup_{m=m_{0}}^{m_{2}}\mathcal{C}_{m},

maxGm=m0m2𝒞mx(G),umax{x,u:x0R+1f(x)t}.\displaystyle\max_{G\in\cup_{m=m_{0}}^{m_{2}}\mathcal{C}_{m}}\langle x(G),u\rangle\leq\max\{\langle x,u\rangle:x\in\mathbb{R}_{\geq 0}^{R+1}\land f(x)\geq t\}.

Note that in the above we dropped the condition that xx came from a graph. By this we only consider more possible cases and thus obtain an upper bound on the maximization problem.

Now we are ready to state the main technical proposition in order to bound from above the term mlog(p)vmlog(np2)m\log(p)-v_{m}\log(np^{2}) when log9(n)npn2/3ε\frac{\log^{9}(n)}{n}\leq p\leq n^{-2/3-\varepsilon}.

Proposition 6.13.

Suppose k>1k>1 is an integer and suppose that η,ε,δ\eta,\varepsilon,\delta are positive reals with ε\varepsilon small enough as a function of η\eta and kk. Then, there exists n0n_{0} such that for any n>n0n>n_{0} and any n1+ck1pn1+ckn^{-1+c_{k-1}}\leq p\leq n^{-1+c_{k}} we have

max{x,u:x0R+1f(x)t}(1η)mkuk.\max\{\langle x,u\rangle:x\in\mathbb{R}_{\geq 0}^{R+1}\land f(x)\geq t\}\leq(1-\eta)m_{k}u_{k}.

This proposition and Proposition 6.8 imply the following corollary. This corollary yields the first item in Theorem 6.1 as explained before.

Corollary 6.14.

Suppose k>1k>1 is an integer and suppose that η,ε,δ,K\eta,\varepsilon,\delta,K are positive reals with ε\varepsilon small as a function of η\eta and kk. Then, there exists n0n_{0} such that for any n>n0n>n_{0} and any max{log9(n)n,n1+ck1}pn1+ck\max\left\{\frac{\log^{9}(n)}{n},n^{-1+c_{k-1}}\right\}\leq p\leq n^{-1+c_{k}} we have the following:

mlog(p)vmlog(np2)(1η)mkuk=(1η)mk(log(p)log(np2)/k)m\log(p)-v_{m}\log(np^{2})\leq(1-\eta)m_{k}u_{k}=(1-\eta)m_{k}(\log(p)-\log(np^{2})/k)

for every m0mKΦX(δ+ε)m_{0}\leq m\leq K\Phi_{X}(\delta+\varepsilon).

We now turn to the proof of Proposition 6.13. We show that for max{log9(n)n,n1+ck1}pn1+ck\max\left\{\frac{\log^{9}(n)}{n},n^{-1+c_{k-1}}\right\}\leq p\leq n^{-1+c_{k}} the maximum of x,u\langle x,u\rangle is achieved by a vector xx of the form αek\alpha\cdot e_{k} where eke_{k} is the kk-th element in the standard basis. The strategy of the proof is to think of the vector xx as a distribution vector and push its mass towards the ‘center of mass’ while keeping x,u\langle x,u\rangle constant and ensuring that ff evaluated on the vector after the transformation is still greater than tt. To be more precise, the center of mass will be the kk-th coordinate such that kk satisfies max{log9(n)n,n1+ck1}pn1+ck\max\left\{\frac{\log^{9}(n)}{n},n^{-1+c_{k-1}}\right\}\leq p\leq n^{-1+c_{k}}. This will be done iteratively by showing that if xx is a vector achieving the maximum of x,u\langle x,u\rangle and xi=0x_{i}=0 for all ij<ki\leq j<k then we may define a vector x~\tilde{x} which also achieves this maximum and satisfies xi~=0\tilde{x_{i}}=0 for all ij+1i\leq j+1. A similar statement will be shown for the other direction i.e. pushing the mass from the right to the left.

Let us introduce the following notations of pushing mass. Let 2kR2\leq k\leq R be some integer (should be thought of as the center of mass) and suppose that xR+1x\in\mathbb{R}^{R+1}. Then for any integers 2i<k<jR+12\leq i<k<j\leq R+1 define the following vectors,

x(i,k)=xxiei+uixiui+1ei+1,x^{(i,k)}=x-x_{i}e_{i}+\frac{u_{i}x_{i}}{u_{i+1}}e_{i+1},
x(j,k)=xxjej+ujxjuj1ej1.x^{(j,k)}=x-x_{j}e_{j}+\frac{u_{j}x_{j}}{u_{j-1}}e_{j-1}.

One should think of these operations as pushing the mass towards the kk-th coordinate while staying on the hyperplane defined by x,u=s\langle x,u\rangle=s for some constant ss.

Proof of Proposition 6.13.

As explained earlier we prove this proposition iteratively. We do so in several claims. First, we show that we may assume that the distribution vector xx satisfies xR+1=0x_{R+1}=0. This is given in the following claim.

Claim 6.15.

If pn2/3ηp\leq n^{-2/3-\eta} for some positive η\eta, then provided ε\varepsilon is small enough there exists n0n_{0} such that for any n>n0n>n_{0} the following holds:

max{x,u:x0R+1f(x)t}=max{x,u:x0R+1xR+1=0f(x)t}.\max\{\langle x,u\rangle:x\in\mathbb{R}_{\geq 0}^{R+1}\land f(x)\geq t\}=\max\{\langle x,u\rangle:x\in\mathbb{R}_{\geq 0}^{R+1}\land x_{R+1}=0\land f(x)\geq t\}.
Proof.

Let

s=max{x,u:x0R+1f(x)t}.s=\max\{\langle x,u\rangle:x\in\mathbb{R}_{\geq 0}^{R+1}\land f(x)\geq t\}.

Furthermore, let x0R+1x\in\mathbb{R}_{\geq 0}^{R+1} be such that s=x,us=\langle x,u\rangle and f(x)tf(x)\geq t. By the definition of x(R+1,k)x^{(R+1,k)} we have,

x,u=x(R+1,k),u\langle x,u\rangle=\langle x^{(R+1,k)},u\rangle

and therefore, s=x(R+1,k),us=\langle x^{(R+1,k)},u\rangle. In order to prove the lemma it is sufficient to prove f(x(R+1,k))f(x)f(x^{(R+1,k)})\geq f(x). To do so let

f1(x1,x2,,xR+1)=i=2R1(i1)xi(i<jR1xj2j+xi4i+εm2),f_{1}(x_{1},x_{2},\ldots,x_{R+1})=\sum^{R-1}_{i=2}(i-1){x_{i}}\left(\sum_{i<j\leq R-1}\frac{x_{j}}{2j}+\frac{x_{i}}{4i}+\varepsilon m_{2}\right),

which depends only on x1,x2,,xR1x_{1},x_{2},\ldots,x_{R-1}, and

f2(x1,x2,,xR+1)=i=2R1(i1)xixR2R+(R1)xR(xR4R+εm2)+xR+124.f_{2}(x_{1},x_{2},\ldots,x_{R+1})=\sum^{R-1}_{i=2}\frac{(i-1)x_{i}x_{R}}{2R}+(R-1)x_{R}\left(\frac{x_{R}}{4R}+\varepsilon m_{2}\right)+\frac{x_{R+1}^{2}}{4}.

Note that for all yR+1y\in\mathbb{R}^{R+1} we have

f(y)=f1(y)+f2(y).f(y)=f_{1}(y)+f_{2}(y).

Note also that for all 1iR11\leq i\leq R-1 we have xi(R+1,k)=xix^{(R+1,k)}_{i}=x_{i}. Therefore, we obtain that f(x(R+1,k))f(x)=f2(x(R+1,k))f2(x)f(x^{(R+1,k)})-f(x)=f_{2}(x^{(R+1,k)})-f_{2}(x). As xR(R+1,k)=xR+uR+1uRxR+1x_{R}^{(R+1,k)}=x_{R}+\frac{u_{R+1}}{u_{R}}x_{R+1} and xR+1(R+1,k)=0x^{(R+1,k)}_{R+1}=0 a straightforward computation gives

f(x(R+1,k))f(x)\displaystyle f(x^{(R+1,k)})-f(x) =(i=2R(i1)xi2R+εm2(R1))uR+1uRxR+1\displaystyle=\left(\sum^{R}_{i=2}\frac{(i-1)x_{i}}{2R}+\varepsilon m_{2}(R-1)\right)\frac{u_{R+1}}{u_{R}}x_{R+1}
+((R1)uR+12RuR21)xR+124.\displaystyle\quad+\left(\frac{(R-1)u^{2}_{R+1}}{Ru^{2}_{R}}-1\right)\frac{x_{R+1}^{2}}{4}.

Note that uR+1=log(p)<0u_{R+1}=\log(p)<0 and uR<0u_{R}<0. This holds as uR=log(p)log(np2)/Ru_{R}=\log(p)-\log(np^{2})/R and we assume ε\varepsilon is small enough so that pR2np^{R-2}\leq n, as then the left-hand side decays to zero and the right-hand side approaches infinity. This implies that

uR+1uR>0.\frac{u_{R+1}}{u_{R}}>0.

Therefore,

f(x(R+1,k))f(x)((R1)uR+12RuR21)xR+124.f(x^{(R+1,k)})-f(x)\geq\left(\frac{(R-1)u^{2}_{R+1}}{Ru^{2}_{R}}-1\right)\frac{x^{2}_{R+1}}{4}.

Proving that the above is nonnegative provided n0n_{0} is sufficiently large will conclude the proof. Indeed, letting y=log(np2)Rlog(p)y=\frac{\log(np^{2})}{R\log(p)}

(R1)uR+12RuR2=(R1)log2(p)R(log(p)log(np2)/R)21\frac{(R-1)u^{2}_{R+1}}{Ru^{2}_{R}}=\frac{(R-1)\log^{2}(p)}{R(\log(p)-\log(np^{2})/R)^{2}}\geq 1

is equivalent to,

y2+2y1/R0,-y^{2}+2y-1/R\geq 0,

which is also equivalent to

111/Ry1+11/R.1-\sqrt{1-1/R}\leq y\leq 1+\sqrt{1-1/R}.

Since 11xx/2+x21-\sqrt{1-x}\leq x/2+x^{2} and 1+1x2x1+\sqrt{1-x}\geq 2-x for all 0x10\leq x\leq 1, it is enough to prove that

2/R+1/R2log(np2)Rlog(p)21/R.2/R+1/R^{2}\leq\frac{\log(np^{2})}{R\log(p)}\leq 2-1/R.

First, as log(p)<0\log(p)<0, the right inequality is equivalent to

log(np2)(2R1)log(p).\log(np^{2})\geq(2R-1)\log(p).

Equivalently, log(n)(2R3)log(p)\log(n)\geq(2R-3)\log(p), which holds for ε\varepsilon satisfying 2R302R-3\geq 0 as then the left-hand side is positive for n>1n>1 and the right-hand side tends to minus infinity as pp tends to infinity. Second, as np21np^{2}\ll 1, the left inequality is equivalent to

(1/2+1/R)log(p)log(np2).(1/2+1/R)\log(p)\geq\log(np^{2}).

This holds if and only if pn1/(3/21/R)p\leq n^{-1/(3/2-1/R)}. Finally, assuming that ε\varepsilon is small enough so that 1/R3/21/(2/3+η)1/R\leq 3/2-1/(2/3+\eta) implying the claim. ∎

We now continue to the more technical part of the proof which is to push the mass to the kk-th coordinate provided that the last coordinate is 0, which we may assume due to the above claim.

Claim 6.16.

If n1+ck1pn1+ckn^{-1+c_{k-1}}\leq p\leq n^{-1+c_{k}} then there exists n0n_{0} such that for any n>n0n>n_{0} the following holds:

max{x,u:x0R+1f(x)t}=max{αuk:f(αek)t}.\max\{\langle x,u\rangle:x\in\mathbb{R}_{\geq 0}^{R+1}\land f(x)\geq t\}=\max\{\alpha u_{k}:f(\alpha e_{k})\geq t\}.
Proof.

Let

s=max{x,u:x0R+1f(x)t}.s=\max\{\langle x,u\rangle:x\in\mathbb{R}_{\geq 0}^{R+1}\land f(x)\geq t\}.

Furthermore, let x0R+1x\in\mathbb{R}_{\geq 0}^{R+1} be such that s=x,us=\langle x,u\rangle and f(x)tf(x)\geq t. By Claim 6.15 we may assume xR+1=0x_{R+1}=0. As xx satisfies x,u=s\langle x,u\rangle=s we also have x(i,k),u=x(j,k),u=s\langle x^{(i,k)},u\rangle=\langle x^{(j,k)},u\rangle=s for any 2i<k<jR2\leq i<k<j\leq R.

We will start with the push of the mass from the left to the right. To this end we prove iteratively the following: Let x0R+1x\in\mathbb{R}_{\geq 0}^{R+1} and suppose that xr=0x_{r}=0 for all 1r<i<k1\leq r<i<k. Then f(x(i,k))f(x)0f\left(x^{(i,k)}\right)-f(x)\geq 0 provided n0n_{0} is large enough. To see this let

f1(x1,x2,,xR+1)==i+2R(1)x(<jRxj2j+x4+εm2),\displaystyle f_{1}(x_{1},x_{2},\ldots,x_{R+1})=\sum^{R}_{\ell=i+2}{(\ell-1)x_{\ell}}\left(\sum_{\ell<j\leq R}\frac{x_{j}}{2j}+\frac{x_{\ell}}{4\ell}+\varepsilon m_{2}\right),

which depends only on xi+2,xi+3,,xRx_{i+2},x_{i+3},\ldots,x_{R}, and let

f2(x1,x2,,xR+1)==ii+1(1)x(<jRxj2j+x4+εm2),\displaystyle f_{2}(x_{1},x_{2},\ldots,x_{R+1})=\sum^{i+1}_{\ell=i}{(\ell-1)x_{\ell}}\left(\sum_{\ell<j\leq R}\frac{x_{j}}{2j}+\frac{x_{\ell}}{4\ell}+\varepsilon m_{2}\right),

which depends only on xi,xi+1,,xRx_{i},x_{i+1},\ldots,x_{R}. Note that for all yy with yr=0y_{r}=0 for r<ir<i we have f(y)=f1(y)+f2(y)f(y)=f_{1}(y)+f_{2}(y). This implies that f(x(i,k))f(x)=f2(x(i,k))f2(x)f(x^{(i,k)})-f(x)=f_{2}(x^{(i,k)})-f_{2}(x). By the definition of x(i,k)x^{(i,k)} we have

f(x(i,k))f(x)=\displaystyle f\left(x^{(i,k)}\right)-f(x)= (i+1jxixj2j+εm2xi)(iuiui+1(i1))+xi24(iui2(i+1)ui+12i1i).\displaystyle\left(\sum_{i+1\leq j}\frac{x_{i}x_{j}}{2j}+\varepsilon m_{2}x_{i}\right)\left(\frac{iu_{i}}{u_{i+1}}-(i-1)\right)+\frac{x_{i}^{2}}{4}\left(\frac{iu^{2}_{i}}{(i+1)u^{2}_{i+1}}-\frac{i-1}{i}\right).

Thus, to show that f(x(i,k))f(x)0f\left(x^{(i,k)}\right)-f(x)\geq 0 it is suffices to show the following:

iuiui+1(i1)0,\displaystyle\frac{iu_{i}}{u_{i+1}}-(i-1)\geq 0, (23)

and

iui2(i+1)ui+12i1i0.\displaystyle\frac{iu^{2}_{i}}{(i+1)u^{2}_{i+1}}-\frac{i-1}{i}\geq 0. (24)

As np2=o(1)np^{2}=o(1) we have 0>ui>ui+10>u_{i}>u_{i+1}. Therefore, inequality (23) is equivalent to

ilog(p)log(np2)(i1)log(p)i1i+1log(np2).\displaystyle i\log(p)-\log(np^{2})\leq(i-1)\log(p)-\frac{i-1}{i+1}\log(np^{2}).

By algebraic manipulation the above is equivalent to

log(p)2i+1log(np2).\displaystyle\log(p)\leq\frac{2}{i+1}\log(np^{2}).

This always holds in our settings provided nn is sufficiently large. Indeed, the above is equivalent to pi3n2p^{i-3}\leq n^{2} which holds as n1p1n^{-1}\leq p\ll 1. It is left to show inequality (24). We prove this in the following claim.

Claim 6.17.
i2ui2(i+1)(i1)ui+12pn1+ci.i^{2}u_{i}^{2}\geq(i+1)(i-1)u^{2}_{i+1}\iff p\geq n^{-1+c_{i}}.
Proof.

The inequality in the left-hand side is equivalent to iui(i+1)(i1)ui+1iu_{i}\leq\sqrt{(i+1)(i-1)}u_{i+1} as 0>ui>ui+10>u_{i}>u_{i+1}. Recall that u=log(p)log(np2)/u_{\ell}=\log(p)-\log(np^{2})/\ell and plug this in the above inequality to obtain

ilog(p)log(np2)(i1)(i+1)log(p)i1i+1log(np2).i\log(p)-\log(np^{2})\leq\sqrt{{(i-1)}{(i+1)}}\log(p)-\sqrt{\frac{i-1}{i+1}}\log(np^{2}).

Equivalently,

(i2(i+1)(i1)+2i1i+1)log(p)(1i1i+1)log(n)\left(i-2-\sqrt{(i+1)(i-1)}+2\sqrt{\frac{i-1}{i+1}}\right)\log(p)\leq\left(1-\sqrt{\frac{i-1}{i+1}}\right)\log(n)

Thus we conclude that our assumption is equivalent to the following:

log(p)i+1i1(i2)i+1(i1)3/2log(n)=(12+i+1i11)log(n)=(1+ci)log(n).\log(p)\geq\frac{\sqrt{i+1}-\sqrt{i-1}}{(i-2)\sqrt{i+1}-(i-1)^{3/2}}\log(n)=\left(\frac{1}{2+\sqrt{\frac{i+1}{i-1}}}-1\right)\log(n)=(-1+c_{i})\log(n).\qed

Since we assumed pn1+ck1p\geq n^{-1+c_{k-1}} and for all i<ki<k we have cick1c_{i}\leq c_{k-1} we also have pn1+ck1n1+cip\geq n^{-1+c_{k-1}}\geq n^{-1+c_{i}}. Therefore, for any xR+1x\in\mathbb{R}^{R+1} we may iterate this process for 2ik12\leq i\leq k-1 and obtain that f(x(2,k)(3,k)(k1,k))f(x)f\left(x^{(2,k)(3,k)\ldots(k-1,k)}\right)\geq f(x). This conclude the push of the mass of xx from the ii-th coordinates for i<ki<k to the kk-th coordinate.

Now provided what we showed so far we prove a similar statement for x(j,k)x^{(j,k)}. More specifically, we prove iteratively the following: Let xR+1x\in\mathbb{R}^{R+1} and suppose that xr=0x_{r}=0 for all k<j<rR+1k<j<r\leq R+1 and x=0x_{\ell}=0 for all <k\ell<k. Then f(x(j,k))f(x)0f\left(x^{(j,k)}\right)-f(x)\geq 0 provided n0n_{0} is large enough. To see this let

f3(x1,x2,,xR+1)=i=kj2(i1)xi(i<j2x2+xi4i+εm2),\displaystyle f_{3}(x_{1},x_{2},\ldots,x_{R+1})=\sum^{j-2}_{i=k}{(i-1)x_{i}}\left(\sum_{i<\ell\leq j-2}\frac{x_{\ell}}{2\ell}+\frac{x_{i}}{4i}+\varepsilon m_{2}\right),

which depends only on xix_{i} for kij2k\leq i\leq j-2, and let

f4(x1,x2,,xR+1)=\displaystyle f_{4}(x_{1},x_{2},\ldots,x_{R+1})= i=kj2(i1)xi=j1jx2+i=j1j(i1)xi(i<jx2+xi4i+εm2),\displaystyle\sum^{j-2}_{i=k}{(i-1)x_{i}}\sum_{\ell=j-1}^{j}\frac{x_{\ell}}{2\ell}+\sum^{j}_{i=j-1}{(i-1)x_{i}}\left(\sum_{i<\ell\leq j}\frac{x_{\ell}}{2\ell}+\frac{x_{i}}{4i}+\varepsilon m_{2}\right),

which depends only on xix_{i} for kijk\leq i\leq j.

Note that for all yy with yi=0y_{i}=0 for all iki\leq k or ij+1i\geq j+1 we have f(y)=f3(y)+f4(y)f(y)=f_{3}(y)+f_{4}(y). Note also that xi(j,k)=xix^{(j,k)}_{i}=x_{i} for all ij2i\leq j-2 thus, f(x(j,k))f(x)=f4(x(j,k))f4(x)f(x^{(j,k)})-f(x)=f_{4}(x^{(j,k)})-f_{4}(x). This implies that

f(x(j,k))f(x)=\displaystyle f\left(x^{(j,k)}\right)-f(x)= i=kj1(i1)xixj2(ujuj1(j1)1j)+xj24((j2)uj2(j1)uj12j1j).\displaystyle\sum_{i=k}^{j-1}\frac{(i-1)x_{i}x_{j}}{2}\left(\frac{u_{j}}{u_{j-1}(j-1)}-\frac{1}{j}\right)+\frac{x_{j}^{2}}{4}\left(\frac{(j-2)u_{j}^{2}}{(j-1)u_{j-1}^{2}}-\frac{j-1}{j}\right).

Thus, to show that f(x(j,k))f(x)0f\left(x^{(j,k)}\right)-f(x)\geq 0 it is suffices to show the following:

ujuj1j1j,\frac{u_{j}}{u_{j-1}}\geq\frac{j-1}{j},

and

uj2uj12(j1)2j(j2).\displaystyle\frac{u^{2}_{j}}{u^{2}_{j-1}}\geq\frac{(j-1)^{2}}{j(j-2)}.

Note that these two inequalities are respectively equivalent to the following inequalities:

jlog(p)log(np2)(j1)log(p)log(np2),\displaystyle j\log(p)-\log(np^{2})\leq(j-1)\log(p)-\log(np^{2}), (25)
uj12uj2j(j2)(j1)2.\displaystyle\frac{u^{2}_{j-1}}{u^{2}_{j}}\leq\frac{j(j-2)}{(j-1)^{2}}. (26)

Inequality (25) holds always as p<1p<1. Moreover, by Claim 6.17 inequality (26) holds if and only if pn1+cj1p\leq n^{-1+c_{j-1}}.This indeed holds as we assume pn1+ckp\leq n^{-1+c_{k}} and we also have ckcj1c_{k}\leq c_{j-1} for all j>kj>k. This conclude the proof of the lemma. ∎

Now we deduce the proposition from the above claims. Assuming n0n_{0} is large enough and using Claim 6.16 it is sufficient to bound from above

smax{αuk:f(αek)t}.s\coloneqq\max\{\alpha u_{k}:f(\alpha e_{k})\geq t\}.

Let α\alpha\in\mathbb{R} be a witness for that. We have,

tf(αek)\displaystyle t\leq f(\alpha e_{k}) =α2k14k+εα(k1)m2\displaystyle=\alpha^{2}\frac{k-1}{4k}+\varepsilon\alpha(k-1)m_{2}
=(δ+ε)𝔼[X]α2mk2+2εα(k1)(δ+ε)𝔼[X]\displaystyle=\frac{(\delta+\varepsilon)\mathbb{E}[X]\alpha^{2}}{m_{k}^{2}}+2\varepsilon\alpha(k-1)\sqrt{(\delta+\varepsilon)\mathbb{E}[X]}
=δ+εδε((αmk)2t+4εk(k1)(αmk)t).\displaystyle=\frac{\delta+\varepsilon}{\delta-\varepsilon}\cdot\left(\left(\frac{\alpha}{m_{k}}\right)^{2}t+4\varepsilon\sqrt{k(k-1)}\left(\frac{\alpha}{m_{k}}\right)t\right).

Denoting x=α/mkx=\alpha/m_{k} implies

x2εk(k1)+2ε2k(k1)+δεδ+ε or x2εk(k1)2ε2k(k1)+δεδ+ε.x\geq-2\varepsilon\sqrt{k(k-1)}+2\sqrt{\varepsilon^{2}k(k-1)+\frac{\delta-\varepsilon}{\delta+\varepsilon}}\text{ or }x\leq-2\varepsilon\sqrt{k(k-1)}-2\sqrt{\varepsilon^{2}k(k-1)+\frac{\delta-\varepsilon}{\delta+\varepsilon}}.

Hence, for every positive η\eta and sufficiently small ε\varepsilon we obtain

x1η or x1+η.x\geq 1-\eta\text{ or }x\leq-1+\eta.

The second option is not possible as α\alpha is nonnegative. Thus we have, α/mk1η\alpha/m_{k}\geq 1-\eta and hence

s=αuk(1η)mkuk.s=\alpha u_{k}\leq(1-\eta)m_{k}u_{k}.

This is as claimed. ∎

7. The solution to the variational problem

In this section we solve the naive mean field variational problem (when log(n)/npn1/2ε\sqrt{\log(n)}/n\ll p\ll n^{-1/2-\varepsilon}) for the upper tail of the number of induced copies of C4C_{4} in Gn,pG_{n,p}. Let us recall the definition of the variational problem.

Let NN be a positive integer and let Y=(Y1,Y2,,YN)Y=(Y_{1},Y_{2},\ldots,Y_{N}) be a sequence of independent Bernoulli random variables with mean pp. Further, let XX be a function from the hypercube to \mathbb{R}. Then the naive mean field variational problem associated to the upper tail of X(Y)X(Y) is the function ΨX:00{}\Psi_{X}\colon\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0}\cup\{\infty\} such that for every δ0\delta\geq 0:

ΨX(δ)=inf{iNIp(zi):𝔼[X(Z)](1+δ)𝔼[X(Y)]},\Psi_{X}(\delta)=\inf\left\{\sum_{i\in N}I_{p}(z_{i}):\mathbb{E}[X(Z)]\geq(1+\delta)\mathbb{E}[X(Y)]\right\},

where Ip(q)=DKL(Ber(q)Ber(p))=qlog(qp)+(1q)log(1q1p)I_{p}(q)=D_{KL}(\operatorname{Ber}(q)\|\operatorname{Ber}(p))=q\log\left(\frac{q}{p}\right)+(1-q)\log\left(\frac{1-q}{1-p}\right), and the infimum is taken over all Z=(Z1,Z2,,ZN)Z=(Z_{1},Z_{2},\ldots,Z_{N}) sequences of NN Bernoulli random variables with means ziz_{i} respectively.

We will be interested in the case where N=(n2)N=\binom{n}{2}, the random variables YiY_{i} correspond to the edges of Gn,pG_{n,p}, and XX is the random variable counting the number of induced copies of C4C_{4}. The main proposition of this section is the following.

Proposition 7.1.

Suppose ε,δ\varepsilon,\delta are positive reals. Further, suppose that log(n)/npn1/2ε\sqrt{\log(n)}/n\ll p\ll n^{-1/2-\varepsilon}. Then, for large enough nn, we have

(1ε)δ2ΨX(δ)n2p2log(1/p)(1+ε)δ2.(1-\varepsilon)\sqrt{\frac{\delta}{2}}\leq\frac{\Psi_{X}(\delta)}{n^{2}p^{2}\log(1/p)}\leq(1+\varepsilon)\sqrt{\frac{\delta}{2}}.

The upper bound follows immediately from

ΨX(δ)log(1/p)min{e(C):𝔼[XCGn,p](1+δ)𝔼[X]}.\Psi_{X}(\delta)\leq\log(1/p)\min\{e(C):\mathbb{E}[X\mid C\subseteq G_{n,p}]\geq(1+\delta)\mathbb{E}[X]\}.

In Section 4 we this minimum. It is attained when CC is the complete bipartite graph Km0,m0K_{m_{0},m_{0}} where m02δ2n2p2m_{0}^{2}\approx\sqrt{\frac{\delta}{2}}n^{2}p^{2}. For more details the reader is referred to Section 4.

For the lower bound we start with the following reduction.

Lemma 7.2.

Suppose ε,δ\varepsilon,\delta are positive reals. Furthermore, suppose p1p\ll 1. Then, for large enough nn, we have

ΨX(δ)infq¯{i([n]2)Ip(qi):𝔼[Nind(C4,Gn,q¯)](1+δε)𝔼[X] and i([n]2)qip}.\displaystyle\Psi_{X}(\delta)\geq\inf_{\bar{q}}\left\{\sum_{i\in\binom{[n]}{2}}I_{p}(q_{i}):\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}})]\geq(1+\delta-\varepsilon)\mathbb{E}[X]\text{ and }\forall i\in\binom{[n]}{2}\,q_{i}\geq p\right\}.

To prove the lemma we define the set 𝒞4\mathcal{C}_{4} as follows: Let 𝒞4\mathcal{C}_{4} be a set of representatives for the set of all tuples (x,y,z,w)[n]4(x,y,z,w)\in[n]^{4} with distinct coordinates and the equivalence relation, (x,y,z,w)(x,y,z,w)(x,y,z,w)\sim(x^{\prime},y^{\prime},z^{\prime},w^{\prime}) if and only if {xy,yz,zw,wx}={xy,yz,zw,wx}\{xy,yz,zw,wx\}=\{x^{\prime}y^{\prime},y^{\prime}z^{\prime},z^{\prime}w^{\prime},w^{\prime}x^{\prime}\}; i.e. both (x,y,z,w)(x,y,z,w) and (x,y,z,w)(x^{\prime},y^{\prime},z^{\prime},w^{\prime}) induce the same 4-cycle in KnK_{n}.

Proof.

Let q¯[0,1](n2)\bar{q}\in[0,1]^{\binom{n}{2}} be such that 𝔼[Nind(C4,Gn,q¯)](1+δ)𝔼[X]\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}})]\geq(1+\delta)\mathbb{E}[X]. Define r¯\bar{r} by setting ri=max{p,qi}r_{i}=\max\{p,q_{i}\}. Since Ip(x)I_{p}(x) is monotone decreasing between 0 and pp, we obtain that

Ip(qi)Ip(ri).\sum I_{p}(q_{i})\geq\sum I_{p}(r_{i}).

To conclude the proof, we prove that 𝔼[Nind(C4,Gn,r¯)](1+δε)𝔼[X]\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{r}})]\geq(1+\delta-\varepsilon)\mathbb{E}[X]. Indeed, provided nn is sufficiently large, we have

𝔼[Nind(C4,Gn,r¯)]\displaystyle\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{r}})] =(x,y,z,w)𝒞4rxyryzrzwrwx(1rxz)(1ryw)\displaystyle=\sum_{(x,y,z,w)\in\mathcal{C}_{4}}r_{xy}r_{yz}r_{zw}r_{wx}(1-r_{xz})(1-r_{yw})
(x,y,z,w)𝒞4qxyqyzqzwqwx(1qxz)(1qyw)(1p)2\displaystyle\geq\sum_{(x,y,z,w)\in\mathcal{C}_{4}}q_{xy}q_{yz}q_{zw}q_{wx}(1-q_{xz})(1-q_{yw})(1-p)^{2}
(1+δ)(1p)2𝔼[X](1+δε)𝔼[X].\displaystyle\geq(1+\delta)(1-p)^{2}\mathbb{E}[X]\geq(1+\delta-\varepsilon)\mathbb{E}[X].\qed

Next we present a lemma about the structure of Gn,q¯G_{n,\bar{q}} when Ip(qi)\sum I_{p}(q_{i}) is close to the infimum in Lemma 7.2.

{restatable}

lemmalemmaone Suppose ε,δ\varepsilon,\delta are positive reals with ε\varepsilon sufficiently small. Suppose also that log(n)/npn1/2\sqrt{\log(n)}/n\ll p\ll n^{-1/2}. Then, for any sequence q¯\bar{q} satisfying:

  1. (1)

    qi=p+uiq_{i}=p+u_{i} for some ui0u_{i}\geq 0,

  2. (2)

    i([n]2)Ip(qi)2δn2p2log(1/p)\sum_{i\in\binom{[n]}{2}}I_{p}(q_{i})\leq\sqrt{2\delta}n^{2}p^{2}\log(1/p), and

  3. (3)

    𝔼[Nind(C4,Gn,q¯)](1+δε)𝔼[X]\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}})]\geq(1+\delta-\varepsilon)\mathbb{E}[X],

we have 𝔼[Nind(C4,Gn,u¯)](δ2ε)𝔼[X]\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})]\geq(\delta-2\varepsilon)\mathbb{E}[X] provided nn is large enough.

This lemma is a variant of the methods in the papers of Lubetzky and Zhao [27] and in Bhattacharya, Ganguly, Lubetzky and Zhao [5, Section 5.2]. In these papers the authors use the language of graph limits or ‘graphons’; we recreate their proof in the language of graphs in the Appendix. Further, Lubetzky and Zhao [27] and Bhattacharya, Ganguly, Lubetzky and Zhao [5], analysed the function Ip(p+x)I_{p}(p+x) and proved several lemmas in the language of graphons. We use these lemmas in the proof of Proposition 7.1, once again in the language of graphs and not graphons. For a proof of the first lemma we refer the reader to [27] and [5]; the second lemma will be proven in the Appendix with some extensions.

{restatable}

[[27]]lemmalemmatwo The following holds,

Ip(p+x)={(1+o(1))x22p,0xp,(1+o(1))xlog(x/p),px1p.I_{p}(p+x)=\begin{cases}(1+o(1))\frac{x^{2}}{2p},&0\leq x\ll p,\\ (1+o(1))x\log(x/p),&p\ll x\leq 1-p.\end{cases}
{restatable}

[[27]]lemmalemmathree Suppose ε,δ,C\varepsilon,\delta,C are positive reals and log(n)/npn1/2\sqrt{\log(n)}/n\ll p\ll n^{-1/2}. Suppose also that u¯[0,1p](n2)\bar{u}\in[0,1-p]^{\binom{n}{2}} such that i([n]2)Ip(p+ui)Cn2p2log(1/p)\sum_{i\in\binom{[n]}{2}}I_{p}(p+u_{i})\leq Cn^{2}p^{2}\log(1/p). Let b=b(n)b=b(n) be such that max{np2,plog(1/p)}b1ε\max\{np^{2},\sqrt{p\log(1/p)}\}\ll b\leq 1-\varepsilon. Then, provided nn is large enough there is a constant D>0D>0 such that the following holds:

  1. (i)

    x[n](yxuxy)2Dn3p2b\sum_{x\in[n]}\left(\sum_{y\neq x}u_{xy}\right)^{2}\leq Dn^{3}p^{2}b and

  2. (ii)

    i([n]2)uiDn2p3/2log(1/p)\sum_{i\in\binom{[n]}{2}}u_{i}\leq Dn^{2}p^{3/2}\sqrt{\log(1/p)}.

Now we are ready to prove Proposition 7.1.

Proof of Proposition 7.1.

We wish to prove the following:

i([n]2)Ip(qi)(1η)δ2n2p2log(1/p)\sum_{i\in\binom{[n]}{2}}I_{p}(q_{i})\geq(1-\eta)\sqrt{\frac{\delta}{2}}n^{2}p^{2}\log(1/p)

for sufficiently small η>0\eta>0 and for q¯[0,1](n2)\bar{q}\in[0,1]^{\binom{n}{2}} such that

𝔼[Nind(C4,Gn,q¯)](1+δ)𝔼[X].\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}})]\geq(1+\delta)\mathbb{E}[X].

We may assume that

i([n]2)Ip(qi)Cn2p2log(1/p),\sum_{i\in\binom{[n]}{2}}I_{p}(q_{i})\leq Cn^{2}p^{2}\log(1/p),

for some absolute constant C>0C>0 as otherwise there is nothing to prove. By Lemma 7.2 and Lemma 7.2 we may assume that qi=p+uiq_{i}=p+u_{i} for some ui0u_{i}\geq 0 which satisfies

𝔼[N(C4,Gn,u¯)]𝔼[Nind(C4,Gn,u¯)](δε)𝔼[X]\mathbb{E}[N(C_{4},G_{n,\bar{u}})]\geq\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})]\geq(\delta-\varepsilon)\mathbb{E}[X]

where N(C4,Gn,u¯)N(C_{4},G_{n,\bar{u}}) is the number of copies of C4C_{4} in Gn,u¯G_{n,\bar{u}}. Note that

𝔼[N(C4,Gn,u¯)]\displaystyle\mathbb{E}[N(C_{4},G_{n,\bar{u}})] =(x,y,z,w)𝒞4uxyuyzuzwuwx(δε)𝔼[X].\displaystyle={\sum_{(x,y,z,w)\in\mathcal{C}_{4}}u_{xy}u_{yz}u_{zw}u_{wx}}\geq(\delta-\varepsilon)\mathbb{E}[X].

In the next three claims we collect some properties of (x,y,z,w)𝒞4(x,y,z,w)\in\mathcal{C}_{4} with non-negligible contribution to the sum above. Afterwards, we use these properties to show that the set of all 4-cycles with non-negligible contribution to the sum above is a subset of

𝒞4γ{(a0,a1,a2,a3)𝒞4:uaiai+1>γ for all i=0,1,2,3},\mathcal{C}_{4}^{\gamma}\coloneqq\{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4}:u_{a_{i}a_{i+1}}>\gamma\text{ for all }i=0,1,2,3\},

where the summation is modulo 44 and γ\gamma is some positive constant depending on ε\varepsilon. We now explain briefly the proof strategy: First, we show that to have a non-negligible contribution a 4-cycle (a0,a1,a2,a3)𝒞4(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4} must contain a K1,2K_{1,2} with both edge weights at least p/log(1/p)\sqrt{p}/\log(1/p); for convenience assume that these edges are a0a1,a1a2a_{0}a_{1},a_{1}a_{2}. Afterwards, we show that, among such 4-cycles, the only ones contributing a non-negligible amount must additionally satisfy ua2a3p12εu_{a_{2}a_{3}}\geq p^{1-2\varepsilon} or ua3a0p12εu_{a_{3}a_{0}}\geq p^{1-2\varepsilon}, assume ua3a0p12εu_{a_{3}a_{0}}\geq p^{1-2\varepsilon}. Then, we prove that a 4-cycle with the above properties contributes a non-negligible amount only if ua0a1,ua2a3γu_{a_{0}a_{1}},u_{a_{2}a_{3}}\geq\gamma (in the complementary case we show ua1a2,ua3a1γu_{a_{1}a_{2}},u_{a_{3}a_{1}}\geq\gamma). Using symmetries of the 4-cycle, we may use this argument again from a different point of view. We then obtain that the only 4-cycles contributing a non-negligible amount to the sum are ones with uaiai+1γu_{a_{i}a_{i+1}}\geq\gamma for all ii. This explanation can also be viewed in Figure 1.

Densities:γ\geq\gammaplog(1/p)\geq\frac{\sqrt{p}}{\log(1/p)}p12ε\geq p^{1-2\varepsilon}
a0a_{0}a1a_{1}a2a_{2}a3a_{3} Claim 7.3 a0a_{0}a1a_{1}a2a_{2}a3a_{3} Claim 7.4 a0a_{0}a1a_{1}a2a_{2}a3a_{3} Claim 7.5 Claim 7.5 a0a_{0}a1a_{1}a2a_{2}a3a_{3} Claim 7.5 a0a_{0}a1a_{1}a2a_{2}a3a_{3}
Figure 1.

We now execute this plan rigorously.

Claim 7.3.

For all η>0\eta>0 the following holds for large enough nn

(x,y,z,w)𝒜uxyuyzuzwuwxη2n4p4,\sum_{(x,y,z,w)\in\mathcal{A}}u_{xy}u_{yz}u_{zw}u_{wx}\leq\eta^{2}n^{4}p^{4},

where 𝒜\mathcal{A} is the set of all tuples (x,y,z,w)𝒞4(x,y,z,w)\in\mathcal{C}_{4} with

uxy,uzwp/log(1/p)oruxw,uyzp/log(1/p).u_{xy},u_{zw}\leq\sqrt{p}/\log(1/p)\quad\text{or}\quad u_{xw},u_{yz}\leq\sqrt{p}/\log(1/p).
Proof.

Suppose otherwise, meaning,

(x,y,z,w)𝒜uxyuyzuzwuwx>η2n4p4.\sum_{(x,y,z,w)\in\mathcal{A}}u_{xy}u_{yz}u_{zw}u_{wx}>\eta^{2}n^{4}p^{4}.

By the definition of 𝒜\mathcal{A} we have

uxy,uzwp/log(1/p)oruxz,uywp/log(1/p).\displaystyle u_{xy},u_{zw}\leq\sqrt{p}/\log(1/p)\quad\text{or}\quad u_{xz},u_{yw}\leq\sqrt{p}/\log(1/p).

This implies that

η2n4p4<(x,y,z,w)𝒜uxyuyzuzwuwx(x,yuxy)2plog2(1/p),\eta^{2}n^{4}p^{4}<\sum_{(x,y,z,w)\in\mathcal{A}}u_{xy}u_{yz}u_{zw}u_{wx}\leq\left(\sum_{x,y}u_{xy}\right)^{2}\frac{p}{\log^{2}(1/p)},

and therefore,

ηn2p3/2log(1/p)x,yuxy.\eta n^{2}p^{3/2}\log(1/p)\leq\sum_{x,y}u_{xy}.

By Lemma 7.2 we have x,yuxy=O(n2p3/2log(1/p))\sum_{x,y}u_{xy}=O\left(n^{2}p^{3/2}\sqrt{\log(1/p)}\right). This implies,

ηn2p3/2log(1/p)O(n2p3/2log(1/p)),\eta n^{2}p^{3/2}\log(1/p)\leq O\left(n^{2}p^{3/2}\sqrt{\log(1/p)}\right),

contradicting the assumption that p=o(1)p=o(1). ∎

Claim 7.4.

For all η>0\eta>0 the following holds for large enough nn

(a0,a1,a2,a3)𝒜ua0a1ua1a2ua2a3ua3a0η2n4p4,\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{B}\setminus\mathcal{A}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}\leq\eta^{2}n^{4}p^{4},

where =i=03i\mathcal{B}=\cup_{i=0}^{3}\mathcal{B}_{i}, and i\mathcal{B}_{i} is the set of all (a0,a1,a2,a3)𝒞4(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4} with

uaiai+1,uai+1ai+2p12ε.u_{a_{i}a_{i+1}},u_{a_{i+1}a_{i+2}}\leq p^{1-2\varepsilon}.
Proof.

Assume towards contradiction that

(a0,a1,a2,a3)𝒜ua0a1ua1a2ua2a3ua3a0>η2n4p4.\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{B}\setminus\mathcal{A}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}>\eta^{2}n^{4}p^{4}.

We claim that for any (a0,a1,a2,a3)𝒜(a_{0},a_{1},a_{2},a_{3})\in\mathcal{B}\setminus\mathcal{A} there is ii satisfying

uaiai+1,uai+1ai+2>p/log(1/p)anduai+2ai+3,uai+3aip12ε,u_{a_{i}a_{i+1}},u_{a_{i+1}a_{i+2}}>\sqrt{p}/\log(1/p)\quad\text{and}\quad u_{a_{i+2}a_{i+3}},u_{a_{i+3}a_{i}}\leq p^{1-2\varepsilon},

where summation is taken modulo 44. Indeed, let (a0,a1,a2,a3)𝒜(a_{0},a_{1},a_{2},a_{3})\in\mathcal{B}\setminus\mathcal{A}. Since (a0,a1,a2,a3)𝒜(a_{0},a_{1},a_{2},a_{3})\not\in\mathcal{A} we have

ua0a1>p/log(1/p)orua2a3>p/log(1/p)u_{a_{0}a_{1}}>\sqrt{p}/\log(1/p)\quad\text{or}\quad u_{a_{2}a_{3}}>\sqrt{p}/\log(1/p)

and

ua0a3>p/log(1/p)orua1a2>p/log(1/p).u_{a_{0}a_{3}}>\sqrt{p}/\log(1/p)\quad\text{or}\quad u_{a_{1}a_{2}}>\sqrt{p}/\log(1/p).

Without loss of generality, assume that ua0,a1,ua1,a2>p/log(1/p)u_{a_{0},a_{1}},u_{a_{1},a_{2}}>\sqrt{p}/\log(1/p). Note that as (a0,a1,a2,a3)(a_{0},a_{1},a_{2},a_{3})\in\mathcal{B} there is ii such that uaiai+1,uai+1ai+2p12εu_{a_{i}a_{i+1}},u_{a_{i+1}a_{i+2}}\leq p^{1-2\varepsilon}. For ε<1/4\varepsilon<1/4 we have p/log(1/p)p12ε\sqrt{p}/\log(1/p)\geq p^{1-2\varepsilon} implying that i0,1,3i\neq 0,1,3 and therefore, i=2i=2. We continue by letting 1\mathcal{L}_{1} be the set of all {x,y}([n]2)\{x,y\}\in\binom{[n]}{2} such that uxy>p/log(1/p)u_{xy}>\sqrt{p}/\log(1/p), and we claim that

(a0,a1,a2,a3)𝒜ua0a1ua1a2ua2a3ua3a0np2(12ε)({x,y}1uxy)2.\displaystyle\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{B}\setminus\mathcal{A}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}\leq np^{2(1-2\varepsilon)}\left(\sum_{\{x,y\}\in\mathcal{L}_{1}}u_{xy}\right)^{2}.

This follows from the fact that any K1,2K_{1,2} in KnK_{n} can be extended to a C4C_{4} in at most nn ways. Therefore, we obtain the following:

{x,y}1uxy>ηn3/2p1+2ε.\sum_{\{x,y\}\in\mathcal{L}_{1}}u_{xy}>\eta n^{3/2}p^{1+2\varepsilon}.

By Lemma 7.2 we have,

x,yIp(p+uxy){x,y}1Ip(p+uxy)={x,y}1(1+o(1))uxylog(uxy/p)=Ω(n3/2p1+2εlog(1/p)).\sum_{x,y}I_{p}(p+u_{xy})\geq\sum_{\{x,y\}\in\mathcal{L}_{1}}I_{p}(p+u_{xy})=\sum_{\{x,y\}\in\mathcal{L}_{1}}(1+o(1))u_{xy}\log(u_{xy}/p)=\Omega\left(n^{3/2}p^{1+2\varepsilon}\log(1/p)\right).

This is a contradiction to our assumption that x,yIp(p+uxy)=O(n2p2log(1/p))\sum_{x,y}I_{p}(p+u_{xy})=O(n^{2}p^{2}\log(1/p)) as pn1/2εp\ll n^{-1/2-\varepsilon} and therefore, n3/2p1+2ε=ω(n2p2)n^{3/2}p^{1+2\varepsilon}=\omega(n^{2}p^{2}). ∎

Claim 7.5.

For any η>0\eta>0 there exists γ>0\gamma>0 such that the following holds provided nn is large enough

(a0,a1,a2,a3)𝒟(γ)ua0a1ua1a2ua2a3ua3a0η2n4p4,\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{D(\gamma)}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}\leq\eta^{2}n^{4}p^{4},

where 𝒟(γ)=𝒟0(γ)𝒟1(γ)\mathcal{D}(\gamma)=\mathcal{D}_{0}(\gamma)\cup\mathcal{D}_{1}(\gamma), and 𝒟i(γ)\mathcal{D}_{i}(\gamma) is the set of all tuples (a0,a1,a2,a3)𝒞4(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4} with

uaiai+1γoruai+2ai+3γ,u_{a_{i}a_{i+1}}\leq\gamma\quad\text{or}\quad u_{a_{i+2}a_{i+3}}\leq\gamma,

and

uai+1ai+2,uaiai+3p12ε.u_{a_{i+1}a_{i+2}},u_{a_{i}a_{i+3}}\geq p^{1-2\varepsilon}.
Proof.

Let γ=(ηε/C)2\gamma=\left({\eta\varepsilon}/{C}\right)^{2} and assume towards contradiction that

(a0,a1,a2,a3)𝒟(γ)ua0a1ua1a2ua2a3ua3a0>η2n4p4.\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{D(\gamma)}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}>\eta^{2}n^{4}p^{4}.

Note that any (a0,a1,a2,a3)𝒟(γ)(a_{0},a_{1},a_{2},a_{3})\in\mathcal{D}(\gamma) satisfies

uaiai+1γoruai+2ai+3γu_{a_{i}a_{i+1}}\leq\gamma\quad\text{or}\quad u_{a_{i+2}a_{i+3}}\leq\gamma

and

uaiai+3,uai+1ai+2p12ε,u_{a_{i}a_{i+3}},u_{a_{i+1}a_{i+2}}\geq p^{1-2\varepsilon},

for some ii and where summation is taken modulo 44. Letting 2\mathcal{L}_{2} be the set of all {x,y}([n]2)\{x,y\}\in\binom{[n]}{2} such that uxyp12εu_{xy}\geq p^{1-2\varepsilon} we have the following by the definition of 𝒟(γ)\mathcal{D}(\gamma):

(a0,a1,a2,a3)𝒟(γ)ua0a1ua1a2ua2a3ua3a0γ({x,y}2uxy)2.\displaystyle\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{D}(\gamma)}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}\leq\gamma\left(\sum_{\{x,y\}\in\mathcal{L}_{2}}u_{xy}\right)^{2}.

This implies that

{x,y}2uxy>ηn2p2/γ.\sum_{\{x,y\}\in\mathcal{L}_{2}}u_{xy}>\eta n^{2}p^{2}/\sqrt{\gamma}.

Therefore, by the definition of 2\mathcal{L}_{2} and by Lemma 7.2 we have,

x,yIp(p+uxy)\displaystyle\sum_{x,y}I_{p}(p+u_{xy}) {x,y}2Ip(p+uxy){x,y}2uxylog(uxy/p)2ηεγn2p2log(1/p).\displaystyle\geq\sum_{\{x,y\}\in\mathcal{L}_{2}}I_{p}(p+u_{xy})\geq\sum_{\{x,y\}\in\mathcal{L}_{2}}u_{xy}\log(u_{xy}/p)\geq\frac{2\eta\varepsilon}{\sqrt{\gamma}}n^{2}p^{2}\log(1/p).

This is a contradiction to the choice of γ\gamma as by our assumptions we have x,yIp(p+uxy)Cn2p2log(1/p)\sum_{x,y}I_{p}(p+u_{xy})\leq Cn^{2}p^{2}\log(1/p). ∎

We now proceed with the proof of Proposition 7.1. By the our assumptions and the above three claims we obtain the following for any η>0\eta>0 and small enough γ>0\gamma>0:

(a0,a1,a2,a3)𝒞4(𝒜𝒟(γ))ua0a1ua1a2ua2a3ua3a0(1ua0a2)(1ua1a3)(δη)𝔼[X].\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4}\setminus\left(\mathcal{A}\cup\mathcal{B}\cup\mathcal{D}(\gamma)\right)}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}(1-u_{a_{0}a_{2}})(1-u_{a_{1}a_{3}})\geq\left(\delta-\eta\right)\mathbb{E}[X].

We claim that provided nn is large enough we have:

𝒞4(𝒜𝒟(γ))𝒞4γ.\mathcal{C}_{4}\setminus\left(\mathcal{A}\cup\mathcal{B}\cup\mathcal{D}(\gamma)\right)\subseteq\mathcal{C}_{4}^{\gamma}.

Indeed, as (a0,a1,a2,a3)𝒜(a_{0},a_{1},a_{2},a_{3})\not\in\mathcal{A}, there exists ii such that provided ε<1/4\varepsilon<1/4 and nn is large enough we have:

uaiai+1,uai+1ai+2>p/log(1/p)>p12ε.u_{a_{i}a_{i+1}},u_{a_{i+1}a_{i+2}}>\sqrt{p}/\log(1/p)>p^{1-2\varepsilon}.

Further, as (a0,a1,a2,a3)i+2(a_{0},a_{1},a_{2},a_{3})\not\in\mathcal{B}_{i+2} we have

uai+2ai+3>p12εoruaiai+3>p12ε.u_{a_{i+2}a_{i+3}}>p^{1-2\varepsilon}\quad\text{or}\quad u_{a_{i}a_{i+3}}>p^{1-2\varepsilon}.

Without loss of generality, assume the latter holds. In particular we have,

uaiai+3,uai+1ai+2>p12ε.u_{a_{i}a_{i+3}},u_{a_{i+1}a_{i+2}}>p^{1-2\varepsilon}.

Since (a0,a1,a2,a3)𝒟i(γ)(a_{0},a_{1},a_{2},a_{3})\not\in\mathcal{D}_{i}(\gamma), we have

uaiai+1,uai+2ai+3>γ.u_{a_{i}a_{i+1}},u_{a_{i+2}a_{i+3}}>\gamma.

Finally, as (a0,a1,a2,a3)𝒟i+1(γ)(a_{0},a_{1},a_{2},a_{3})\not\in\mathcal{D}_{i+1}(\gamma), for large enough nn we have p12ε<γp^{1-2\varepsilon}<\gamma implying that

uaiai+3,uai+1ai+2>γ.u_{a_{i}a_{i+3}},u_{a_{i+1}a_{i+2}}>\gamma.

Letting 3\mathcal{L}_{3} be the set of all {x,y}([n]2)\{x,y\}\in\binom{[n]}{2} with uxy>γu_{xy}>\gamma, we claim the following holds:

(a0,a1,a2,a3)𝒞4γua0a1ua1a2ua2a3ua3a0(1ua0a2)(1ua1a3)({a0,a1}3ua0a1)24.\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4}^{\gamma}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}(1-u_{a_{0}a_{2}})(1-u_{a_{1}a_{3}})\leq\frac{\left(\sum_{\{a_{0},a_{1}\}\in\mathcal{L}_{3}}u_{a_{0}a_{1}}\right)^{2}}{4}. (27)

Indeed,

2(a0,a1,a2,a3)𝒞4γua0a1ua1a2ua2a3ua3a0(1ua0a2)(1ua1a3)\displaystyle 2\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4}^{\gamma}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}(1-u_{a_{0}a_{2}})(1-u_{a_{1}a_{3}})

is at most

ua0a1ua2a3(ua1a2ua3a0(1ua0a2)(1ua1a3)+ua0a2ua1a3(1ua1a2)(1ua3a0)()),\sum u_{a_{0}a_{1}}u_{a_{2}a_{3}}(\underbrace{u_{a_{1}a_{2}}u_{a_{3}a_{0}}(1-u_{a_{0}a_{2}})(1-u_{a_{1}a_{3}})+u_{a_{0}a_{2}}u_{a_{1}a_{3}}(1-u_{a_{1}a_{2}})(1-u_{a_{3}a_{0}})}_{(\star)}),

where the sum ranges over unordered pairs, {{a0,a1},{a2,a3}}3\{\{a_{0},a_{1}\},\{a_{2},a_{3}\}\}\subseteq\mathcal{L}_{3}, such that a0,a1,a2,a3a_{0},a_{1},a_{2},a_{3} are all distinct. We also have

()(1ua0a2)+ua0a2=1.(\star)\leq(1-u_{a_{0}a_{2}})+u_{a_{0}a_{2}}=1.

This implies that

(δη)𝔼[X](a0,a1,a2,a3)𝒞4γua0a1ua1a2ua2a3ua3a0(1ua0a2)(1ua1a3)ua0a1ua2a32,(\delta-\eta)\mathbb{E}[X]\leq\sum_{(a_{0},a_{1},a_{2},a_{3})\in\mathcal{C}_{4}^{\gamma}}u_{a_{0}a_{1}}u_{a_{1}a_{2}}u_{a_{2}a_{3}}u_{a_{3}a_{0}}(1-u_{a_{0}a_{2}})(1-u_{a_{1}a_{3}})\leq\frac{\sum u_{a_{0}a_{1}}u_{a_{2}a_{3}}}{2},

where the summation in the right-hand side is over unordered pairs, {{a0,a1},{a2,a3}}3\{\{a_{0},a_{1}\},\{a_{2},a_{3}\}\}\subseteq\mathcal{L}_{3}, such that a0,a1,a2,a3a_{0},a_{1},a_{2},a_{3} are all distinct. This implies (27) as

({a0,a1}3ua0a1)22={a0,a1}3ua0a12+2ua0a1ua2a32ua0a1ua2a3,\frac{\left(\sum_{\{a_{0},a_{1}\}\in\mathcal{L}_{3}}u_{a_{0}a_{1}}\right)^{2}}{2}=\frac{\sum_{\{a_{0},a_{1}\}\in\mathcal{L}_{3}}u^{2}_{a_{0}a_{1}}+2\sum u_{a_{0}a_{1}}u_{a_{2}a_{3}}}{2}\geq{\sum u_{a_{0}a_{1}}u_{a_{2}a_{3}}},

where the unlabeled summations are over unordered pairs {{a0,a1},{a2,a3}}3\{\{a_{0},a_{1}\},\{a_{2},a_{3}\}\}\subseteq\mathcal{L}_{3}. We conclude that:

2(δη)𝔼[X]{a0,a1}3ua0a1.2\sqrt{(\delta-\eta)\mathbb{E}[X]}\leq\sum_{\{a_{0},a_{1}\}\in\mathcal{L}_{3}}u_{a_{0}a_{1}}.

Now we can bound from below the ‘cost’ using Lemma 7.2 as follows:

{x,y}([n]2)Ip(p+uxy)\displaystyle\sum_{\{x,y\}\in\binom{[n]}{2}}I_{p}(p+u_{xy}) {x,y}3Ip(p+uxy)=(1+o(1)){x,y}3uxylog(uxy/p)\displaystyle\geq\sum_{\{x,y\}\in\mathcal{L}_{3}}I_{p}(p+u_{xy})=(1+o(1))\sum_{\{x,y\}\in\mathcal{L}_{3}}u_{xy}\log(u_{xy}/p)
(1+o(1)){x,y}3uxylog(1/p)(1+o(1))2(δη)𝔼[X]log(1/p).\displaystyle\geq(1+o(1))\sum_{\{x,y\}\in\mathcal{L}_{3}}u_{xy}\log(1/p)\geq(1+o(1))2\sqrt{(\delta-\eta)\mathbb{E}[X]}\log(1/p).

This finishes the proof of the proposition as 𝔼[X]=(1+o(1))n2p28.\mathbb{E}[X]=(1+o(1))\frac{n^{2}p^{2}}{8}.

References

  • [1] Noga Alon, On the number of subgraphs of prescribed type of graphs with a given number of edges, Israel J. Math. 38 (1981), no. 1-2, 116–130. MR 599482
  • [2] Fanny Augeri, Nonlinear large deviation bounds with applications to Wigner matrices and sparse Erdős-Rényi graphs, Ann. Probab. 48 (2020), no. 5, 2404–2448. MR 4152647
  • [3] by same author, A transportation approach to the mean-field approximation, Probability Theory and Related Fields 180 (2021), no. 1, 1–32.
  • [4] Anirban Basak and Riddhipratim Basu, Upper tail large deviations of regular subgraph counts in Erdős–Rényi graphs in the full localized regime, arXiv preprint arXiv:1912.11410 (2019).
  • [5] Bhaswar B Bhattacharya, Shirshendu Ganguly, Eyal Lubetzky, and Yufei Zhao, Upper tails and independence polynomials in random graphs, Advances in Mathematics 319 (2017), 313–347.
  • [6] Béla Bollobás, Random graphs, Combinatorics (Swansea, 1981), London Math. Soc. Lecture Note Ser., vol. 52, Cambridge Univ. Press, Cambridge-New York, 1981, pp. 80–102. MR 633650
  • [7] Béla Bollobás, Chiê Nara, and Shun-ichi Tachibana, The maximal number of induced complete bipartite graphs, Discrete Math. 62 (1986), no. 3, 271–275. MR 866942
  • [8] Sourav Chatterjee, The missing log in large deviations for triangle counts, Random Structures & Algorithms 40 (2012), no. 4, 437–451.
  • [9] Sourav Chatterjee and Amir Dembo, Nonlinear large deviations, Advances in Mathematics 299 (2016), 396–450.
  • [10] Sourav Chatterjee and SR Srinivasa Varadhan, The large deviation principle for the Erdős–Rényi random graph, European Journal of Combinatorics 32 (2011), no. 7, 1000–1017.
  • [11] Nicholas Cook and Amir Dembo, Large deviations of subgraph counts for sparse Erdős–Rényi graphs, Advances in Mathematics 373 (2020), 107289.
  • [12] Nicholas A Cook, Amir Dembo, and Huy Tuan Pham, Regularity method and large deviation principles for the Erdős–Rényi hypergraph, arXiv preprint arXiv:2102.09100 (2021).
  • [13] Robert DeMarco and Jeff Kahn, Tight upper tail bounds for cliques, Random Structures & Algorithms 41 (2012), no. 4, 469–487.
  • [14] by same author, Upper tails for triangles, Random Structures Algorithms 40 (2012), no. 4, 452–459. MR 2925307
  • [15] Amir Dembo and Ofer Zeitouni, Large deviations techniques and applications, Stochastic Modelling and Applied Probability, vol. 38, Springer-Verlag, Berlin, 2010, Corrected reprint of the second (1998) edition. MR 2571413
  • [16] Ronen Eldan, Gaussian-width gradient complexity, reverse log-Sobolev inequalities and nonlinear large deviations, Geometric and Functional Analysis 28 (2018), no. 6, 1548–1596.
  • [17] Paul Erdős, On the number of complete subgraphs contained in certain graphs, Magyar Tud. Akad. Mat. Kutató Int. Közl 7 (1962), no. 3, 459–464.
  • [18] Ehud Friedgut and Jeff Kahn, On the number of copies of one hypergraph in another, Israel J. Math. 105 (1998), 251–256. MR 1639767
  • [19] Dániel Gerbner, Dániel T. Nagy, Balázs Patkós, and Máté Vizer, On the maximum number of copies of HH in graphs with given size and order, J. Graph Theory 96 (2021), no. 1, 34–43. MR 4191113
  • [20] Benjamin Gunby, Upper tails of subgraph counts in sparse regular graphs, arXiv preprint arXiv:2010.00658 (2020).
  • [21] Matan Harel, Frank Mousset, and Wojciech Samotij, Upper tails via high moments and entropic stability, arXiv:1904.08212 (2019).
  • [22] Svante Janson, Krzysztof Oleszkiewicz, and Andrzej Ruciński, Upper tails for subgraph counts in random graphs, Israel J. Math. 142 (2004), 61–92. MR 2085711
  • [23] Svante Janson and Andrzej Ruciński, The infamous upper tail, Random Structures & Algorithms 20 (2002), no. 3, 317–342.
  • [24] Jeong Han Kim and Van H Vu, Divide and conquer martingales and the number of triangles in a random graph, Random Structures & Algorithms 24 (2004), no. 2, 166–174.
  • [25] Yang P Liu and Yufei Zhao, On the upper tail problem for random hypergraphs, Random Structures & Algorithms 58 (2021), no. 2, 179–220.
  • [26] Eyal Lubetzky and Yufei Zhao, On replica symmetry of large deviations in random graphs, Random Structures & Algorithms 47 (2015), no. 1, 109–146.
  • [27] by same author, On the variational problem for upper tails in sparse random graphs, Random Structures & Algorithms 50 (2017), no. 3, 420–436.
  • [28] Andrzej Ruciński, When are small subgraphs of a random graph normally distributed?, Probab. Theory Related Fields 78 (1988), no. 1, 1–10. MR 940863
  • [29] Van H. Vu, A large deviation result on the number of small subgraphs of a random graph, Combin. Probab. Comput. 10 (2001), no. 1, 79–94. MR 1827810

Appendix A Translating graphons’ language into graphs’ language

The aim of this Appendix is to add a proof (in the language of graphs instead of graphons) for Lemmas 7.2 and 7.2 which were proven by Lubetzky–Zhao [27] and Bhattacharya, Ganguly, Lubetzky and Zhao [5] in the language of graphons.

We start with a discussion of how one would prove Lemma 7.2, then we state some lemmas and an extension of Lemma 7.2. To prove Lemma 7.2 we will use Lemma 7.2, and therefore it will be proven before Lemma 7.2. Recall Lemma 7.2:

\lemmaone

*

We start with by analysing the term 𝔼[Nind(C4,Gn,q¯)]\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}})]. Suppose q¯[0,1](n2)\bar{q}\in[0,1]^{\binom{n}{2}} satisfies the conditions of Lemma 7.2. Note that

𝔼[Nind(C4,Gn,q¯))]=(x,y,z,w)𝒞4qxyqyzqzwqwx(1qxz)(1qyw).\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}}))]={\sum_{(x,y,z,w)\in\mathcal{C}_{4}}q_{xy}q_{yz}q_{zw}q_{wx}(1-q_{xz})(1-q_{yw})}.

Since qi=p+uiuiq_{i}=p+u_{i}\geq u_{i} we have the following for large enough nn:

𝔼[Nind(C4,Gn,q¯))]\displaystyle\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}}))] (x,y,z,w)𝒞4qxyqyzqzwqwx(1uxz)(1uyw)\displaystyle\leq{\sum_{(x,y,z,w)\in\mathcal{C}_{4}}q_{xy}q_{yz}q_{zw}q_{wx}(1-u_{xz})(1-u_{yw})}
=(x,y,z,w)𝒞4(p+uxy)(p+uyz)(p+uzw)(p+uwx)(1uxz)(1uyw)\displaystyle={\sum_{(x,y,z,w)\in\mathcal{C}_{4}}(p+u_{xy})(p+u_{yz})(p+u_{zw})(p+u_{wx})(1-u_{xz})(1-u_{yw})}
=𝔼[N(C4,Gn,p)]+𝔼[Nind(C4,Gn,u¯)]+HC4𝔼[N(H,Gn,u¯)]p4eH\displaystyle=\mathbb{E}[N(C_{4},G_{n,p})]+\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})]+\sum_{\emptyset\not=H\subsetneq^{*}C_{4}}\mathbb{E}[N(H,G_{n,\bar{u}})]p^{4-e_{H}}
=𝔼[X]/(1p)2+𝔼[Nind(C4,Gn,u¯)]+HC4𝔼[N(H,Gn,u¯)]p4eH\displaystyle=\mathbb{E}[X]/(1-p)^{2}+\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})]+\sum_{\emptyset\not=H\subsetneq^{*}C_{4}}\mathbb{E}[N(H,G_{n,\bar{u}})]p^{4-e_{H}}
=(1+ε/2)𝔼[X]+𝔼[Nind(C4,Gn,u¯)]+HC4𝔼[N(H,Gn,u¯)]p4eH,\displaystyle=(1+\varepsilon/2)\mathbb{E}[X]+\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})]+\sum_{\emptyset\not=H\subsetneq^{*}C_{4}}\mathbb{E}[N(H,G_{n,\bar{u}})]p^{4-e_{H}},

where \subsetneq^{*} stands for a spanning subgraph which is not equal to the host graph. Recalling that 𝔼[Nind(C4,Gn,q¯))]𝔼[X](δε)𝔼[X]\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{q}}))]-\mathbb{E}[X]\geq(\delta-\varepsilon)\mathbb{E}[X], we obtain,

(δ3ε/2)𝔼[X]\displaystyle(\delta-3\varepsilon/2)\mathbb{E}[X]\leq 𝔼[Nind(C4,Gn,u¯)]+𝔼[N(P4,Gn,u¯)]p+𝔼[N(M2,Gn,u¯)]p2\displaystyle\,\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})]+\mathbb{E}[N(P_{4},G_{n,\bar{u}})]p+\mathbb{E}[N(M_{2},G_{n,\bar{u}})]p^{2}
+𝔼[N(K1,2K1,Gn,u¯)]p2+𝔼[N(K22K1,Gn,u¯)]p3.\displaystyle\,+\mathbb{E}[N(K_{1,2}\sqcup K_{1},G_{n,\bar{u}})]p^{2}+\mathbb{E}[N(K_{2}\sqcup 2K_{1},G_{n,\bar{u}})]p^{3}.
\displaystyle\leq 𝔼[Nind(C4,Gn,u¯)]+𝔼[N(P4,Gn,u¯)]p+𝔼[N(M2,Gn,u¯)]p2\displaystyle\,\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})]+\mathbb{E}[N(P_{4},G_{n,\bar{u}})]p+\mathbb{E}[N(M_{2},G_{n,\bar{u}})]p^{2}
+𝔼[N(K1,2,Gn,u¯)]np2+𝔼[N(K2,Gn,u¯)]n2p3,\displaystyle\,+\mathbb{E}[N(K_{1,2},G_{n,\bar{u}})]np^{2}+\mathbb{E}[N(K_{2},G_{n,\bar{u}})]n^{2}p^{3},

where P4P_{4} is the path with three edges, M2M_{2} is the matching of size two, K1,2K1K_{1,2}\sqcup K_{1} is the complete bipartite graph with sides of size 11 and 22 and an extra isolated vertex and K22K1K_{2}\sqcup 2K_{1} is the disjoint union of an edge and an independent set of size two. Therefore, to prove the lemma, it is enough to show that all of the terms in the right-hand side of the above inequality except for 𝔼[Nind(C4,Gn,u¯)]\mathbb{E}[N_{ind}(C_{4},G_{n,\bar{u}})] are negligible in comparison to n4p4n^{4}p^{4}.

For this we remind the reader Lemma 7.2 and cite two useful lemmas from [27] and [5]:

\lemmatwo

*

Lemma A.1 ([27]).

There exists p0>0p_{0}>0 such that for all 0<pp00<p\leq p_{0} and 0xb1p1/log(1/p)0\leq x\leq b\leq 1-p-1/\log(1/p) we have,

Ip(p+x)x2Ip(p+b)b2.I_{p}(p+x)\geq\frac{x^{2}I_{p}(p+b)}{b^{2}}.
Corollary A.2 ([27]).

There exists p0>0p_{0}>0 such that for all 0<pp00<p\leq p_{0} and all 0x1p0\leq x\leq 1-p we have,

Ip(p+x)x2Ip(11/log(1/p))=(1+o(1))x2Ip(1)I_{p}(p+x)\geq x^{2}I_{p}(1-1/\log(1/p))=(1+o(1))x^{2}I_{p}(1)

where the o(1)o(1)-term goes to zero as p0p\to 0.

Another useful tool we use is the generalized Hölder inequality:

Lemma A.3 (Generalized Hölder inequality [5]).

Let μ1,μ2,,μn\mu_{1},\mu_{2},\ldots,\mu_{n} be probability measures on Ω1,Ω2,,Ωn\Omega_{1},\Omega_{2},\ldots,\Omega_{n} respectively, and let μ=i=1nμi\mu=\prod_{i=1}^{n}\mu_{i}. Let A1,A2,,AmA_{1},A_{2},\ldots,A_{m} be non-empty subsets of [n][n] and for A[n]A\subseteq[n] put μA=jAμj\mu_{A}=\prod_{j\in A}\mu_{j} and ΩA=jAΩj\Omega_{A}=\prod_{j\in A}\Omega_{j}. Let fiLpi(ΩAi,μAi)f_{i}\in L^{p_{i}}(\Omega_{A_{i}},\mu_{A_{i}}) for each i[m]i\in[m], and suppose that i:Aij1pi1\sum_{i:A_{i}\ni j}\frac{1}{p_{i}}\leq 1 for all j[n]j\in[n]. Then,

i=1m|fi|dμi=1m(|fi|pi𝑑μAi)1/pi.\int\prod_{i=1}^{m}|f_{i}|d\mu\leq\prod_{i=1}^{m}\left(\int|f_{i}|^{p_{i}}d\mu_{A_{i}}\right)^{1/p_{i}}.

In particular, if every element of [n][n] is contained in at most two sets AjA_{j}, then we can take all pi=2p_{i}=2 and obtain:

i=1mfidμi=1m(|fi|2𝑑μAi)1/2.\int\prod_{i=1}^{m}f_{i}d\mu\leq\prod_{i=1}^{m}\left(\int|f_{i}|^{2}d\mu_{A_{i}}\right)^{1/2}.

Next, we recreate a proof of Lubetzky–Zhao [27] and Bhattacharya, Ganguly, Lubetzky and Zhao [5] of an extension of Lemma 7.2. Then, we will derive the required bounds.

Lemma A.4.

Suppose ε,δ,C\varepsilon,\delta,C are positive reals and log(n)/npn1/2\sqrt{\log(n)}/n\ll p\ll n^{-1/2}. Suppose also that u¯[0,1p](n2)\bar{u}\in[0,1-p]^{\binom{n}{2}} such that i([n]2)Ip(p+ui)Cn2p2log(1/p)\sum_{i\in\binom{[n]}{2}}I_{p}(p+u_{i})\leq Cn^{2}p^{2}\log(1/p). Let b=b(n)b=b(n) be such that max{np2,plog(1/p)}b1ε\max\{np^{2},\sqrt{p\log(1/p)}\}\ll b\leq 1-\varepsilon. Then, provided nn is large enough there is a constant D>0D>0 such that the following holds:

  1. (i)

    x[n](yxuxy)2Dn3p2b,\sum_{x\in[n]}\left(\sum_{y\neq x}u_{xy}\right)^{2}\leq Dn^{3}p^{2}b,

  2. (ii)

    i([n]2)uiDn2p3/2log(1/p)\sum_{i\in\binom{[n]}{2}}u_{i}\leq Dn^{2}p^{3/2}\sqrt{\log(1/p)}, and

  3. (iii)

    i([n]2)ui2Dn2p2.\sum_{i\in\binom{[n]}{2}}u_{i}^{2}\leq Dn^{2}p^{2}.

Proof.

We first show that the set B={x[n]:yxuxybn}B=\{x\in[n]:\sum_{y\neq x}u_{xy}\geq bn\} is empty, provided nn is large enough. If this were not true, then, as Ip(p+x)I_{p}(p+x) is a convex function of xx, we have the following by Jensen’s inequality

x,yIp(p+uxy)\displaystyle\sum_{x,y}I_{p}(p+u_{xy}) (n1)xIp(p+yxuxyn1)(n1)xBIp(p+yxuxyn1)\displaystyle\geq(n-1)\sum_{x}I_{p}\left(p+\frac{\sum_{y\neq x}u_{xy}}{n-1}\right)\geq(n-1)\sum_{x\in B}I_{p}\left(p+\frac{\sum_{y\neq x}u_{xy}}{n-1}\right)
(n1)xBIp(p+b)(1+o(1))(n1)|B|blog(b/p)\displaystyle\geq(n-1)\sum_{x\in B}I_{p}(p+b)\geq{(1+o(1))(n-1)|B|b\log(b/p)}

where the last inequality follows from Lemma 7.2. Since bplog(1/p)b\gg\sqrt{p\log(1/p)} and i([n]2)Ip(p+ui)Cp2n2log(1/p)\sum_{i\in\binom{[n]}{2}}I_{p}(p+u_{i})\leq Cp^{2}n^{2}\log(1/p) we obtain the following:

|B|(1+o(1))Cp2n2log(1/p)(n1)blog(log(1/p)/p)1,|B|\leq(1+o(1))\frac{Cp^{2}n^{2}\log(1/p)}{(n-1)b\log(\sqrt{\log(1/p)/p})}\ll 1,

and therefore, BB is empty. To prove (i), we use the convexity of Ip(p+x)I_{p}(p+x) and Lemma A.1,

x,yIp(p+uxy)\displaystyle\sum_{x,y}I_{p}(p+u_{xy}) (n1)xIp(p+yxuxyn1)=(n1)xBIp(p+yxuxyn1)\displaystyle\geq(n-1)\sum_{x}I_{p}\left(p+\frac{\sum_{y\neq x}u_{xy}}{n-1}\right)=(n-1)\sum_{x\not\in B}I_{p}\left(p+\frac{\sum_{y\neq x}u_{xy}}{n-1}\right)
(n1)Ip(p+b)xB((yxuxy)2b2n2).\displaystyle\geq(n-1)I_{p}(p+b)\sum_{x\not\in B}\left(\frac{\left(\sum_{y\neq x}u_{xy}\right)^{2}}{b^{2}n^{2}}\right).

Since, i([n]2)Ip(p+ui)Cp2n2log(1/p)\sum_{i\in\binom{[n]}{2}}I_{p}(p+u_{i})\leq Cp^{2}n^{2}\log(1/p) we obtain:

x(yxuxy)2Cp2n4b2log(1/p)(n1)Ip(p+b).\sum_{x}\left(\sum_{y\neq x}u_{xy}\right)^{2}\leq\frac{Cp^{2}n^{4}b^{2}\log(1/p)}{(n-1)I_{p}(p+b)}.

As bplog(1/p)pb\gg\sqrt{p\log(1/p)}\gg p we may use Lemma 7.2 and we obtain:

x[n](yxuxy)2Cn4p2b2log(1/p)(1+o(1))(n1)blog(b/p)2Cn3p2b.\sum_{x\in[n]}\left(\sum_{y\neq x}u_{xy}\right)^{2}\leq\frac{Cn^{4}p^{2}b^{2}\log(1/p)}{(1+o(1))(n-1)b\log(b/p)}\leq 2Cn^{3}p^{2}b.

Now we prove (ii). By convexity we have,

(n2)Ip(p+i([n]2)ui(n2))x,yIp(p+uxy)Cn2p2log(1/p).\binom{n}{2}I_{p}\left(p+\frac{\sum_{i\in\binom{[n]}{2}}u_{i}}{\binom{n}{2}}\right)\leq\sum_{x,y}I_{p}(p+u_{xy})\leq Cn^{2}p^{2}\log(1/p).

Since pp3log(1/p)p\gg\sqrt{p^{3}\log(1/p)} we may use Lemma 7.2, and obtain the following for large enough nn:

Ip(p+12Cp3log(1/p))12Cp3log(1/p)3p=4Cp2log(1/p).I_{p}\left(p+\sqrt{12Cp^{3}\log(1/p)}\right)\geq\frac{12Cp^{3}\log(1/p)}{3p}=4Cp^{2}\log(1/p).

This implies the following provided nn is large enough,

Ip(p+i([n]2)ui(n2))4Cp2log(1/p)Ip(p+12Cp3log(1/p)).I_{p}\left(p+\frac{\sum_{i\in\binom{[n]}{2}}u_{i}}{\binom{n}{2}}\right)\leq 4Cp^{2}\log(1/p)\leq I_{p}\left(p+\sqrt{12Cp^{3}\log(1/p)}\right).

By the monotonicity of Ip(p+x)I_{p}(p+x) for x>0x>0 we obtain (ii). That is

i([n]2)ui3Cn4p3log(1/p).\sum_{i\in\binom{[n]}{2}}u_{i}\leq\sqrt{3Cn^{4}p^{3}\log(1/p)}.

Lastly, we prove (iii). By Corollary A.2,

i([n]2)Ip(p+ui)(1+o(1))i([n]2)ui2log(1/p).\sum_{i\in\binom{[n]}{2}}I_{p}(p+u_{i})\geq(1+o(1))\sum_{i\in\binom{[n]}{2}}u_{i}^{2}\log(1/p).

This and the assumption of the lemma implies (iii). ∎

Now we can finish the proof of Lemma 7.2.

Lemma A.5.

Suppose δ,C\delta,C are positive reals and log(n)/npn1/2\sqrt{\log(n)}/n\ll p\ll n^{-1/2}. Suppose also that u¯\bar{u} is a sequence of (n2)\binom{n}{2} reals between 0 and 11 such that i([n]2)Ip(p+ui)Cp2n2log(1/p)\sum_{i\in\binom{[n]}{2}}I_{p}(p+u_{i})\leq Cp^{2}n^{2}\log(1/p). Then,

𝔼[N(M2,Gn,u¯)]p2=o(n4p4),\mathbb{E}[N(M_{2},G_{n,\bar{u}})]p^{2}=o(n^{4}p^{4}),
𝔼[N(K2,Gn,u¯)]n2p3=o(n4p4),\mathbb{E}[N(K_{2},G_{n,\bar{u}})]n^{2}p^{3}=o(n^{4}p^{4}),
𝔼[N(P4,Gn,u¯)]p=o(n4p4),\mathbb{E}[N(P_{4},G_{n,\bar{u}})]p=o(n^{4}p^{4}),
𝔼[N(K1,2,Gn,u¯)]np2=o(n4p4).\mathbb{E}[N(K_{1,2},G_{n,\bar{u}})]np^{2}=o(n^{4}p^{4}).
Proof.

By item (ii) in Lemma A.4 we have 𝔼[e(Gn,u¯)]=o(n2p2)\mathbb{E}[e(G_{n,\bar{u}})]=o(n^{2}p^{2}), and therefore,

𝔼[Nind(M2,Gn,u¯)]p2𝔼[N(M2,Gn,u¯)]p2𝔼[e(Gn,u¯)]2p2=o(n4p4),\mathbb{E}[N_{ind}(M_{2},G_{n,\bar{u}})]p^{2}\leq\mathbb{E}[N(M_{2},G_{n,\bar{u}})]p^{2}\leq\mathbb{E}[e(G_{n,\bar{u}})]^{2}p^{2}=o(n^{4}p^{4}),
𝔼[Nind(K2,Gn,u¯)]n2p3=𝔼[e(Gn,u¯)]n2p3=o(n4p4),\mathbb{E}[N_{ind}(K_{2},G_{n,\bar{u}})]n^{2}p^{3}=\mathbb{E}[e(G_{n,\bar{u}})]n^{2}p^{3}=o(n^{4}p^{4}),

where N(M2,Gn,u¯)N(M_{2},G_{n,\bar{u}}) is the number of M2M_{2} in Gn,u¯G_{n,\bar{u}}. Let max{np2,plog(1/p)}b=b(n)1\max\{np^{2},\sqrt{p\log(1/p)}\}\ll b=b(n)\ll 1. Note that each P4KnP_{4}\subseteq K_{n} with vertices a0,a1,a2,a3a_{0},a_{1},a_{2},a_{3} can be represented in exactly two ways as a tuple (a0,a1,a2,a3)(a_{0},a_{1},a_{2},a_{3}) where edges are consecutive vertices in this tuple. Let 𝒫4\mathcal{P}_{4} be a collection of exactly one such representative for each P4P_{4} in KnK_{n}. Observe the following:

𝔼[N(P4,Gn,u¯)]\displaystyle\mathbb{E}[N(P_{4},G_{n,\bar{u}})] =(x,y,z,w)𝒫4uxyuyzuzwy,z,w[n](xyuxy)uyzuzw,\displaystyle=\sum_{(x,y,z,w)\in\mathcal{P}_{4}}u_{xy}u_{yz}u_{zw}\leq\sum_{y,z,w\in[n]}\left(\sum_{x\neq y}u_{xy}\right)u_{yz}u_{zw},

where N(P4,Gn,u¯)N(P_{4},G_{n,\bar{u}}) is the number of P4P_{4} in Gn,u¯G_{n,\bar{u}}. By the generalized Hölder inequality we have:

y,z,w[n](xyuxy)uyzuzw(y𝔼[deg(y)]2)1/2(x,yuxy2).\sum_{y,z,w\in[n]}\left(\sum_{x\neq y}u_{xy}\right)u_{yz}u_{zw}\leq\left(\sum_{y}\mathbb{E}[\deg(y)]^{2}\right)^{1/2}\left(\sum_{x,y}u_{xy}^{2}\right).

Applying items (i) and (iii) in Lemma A.4 we obtain

𝔼[N(P4,Gn,u¯)]pD3/2n7/2p4b=o(n4p4).\mathbb{E}[N(P_{4},G_{n,\bar{u}})]p\leq D^{3/2}n^{7/2}p^{4}\sqrt{b}=o(n^{4}p^{4}).

This establishes the third assertion of the lemma. The fourth assertion of the lemma follows immediately from Lemma A.4 item (i) and the definition of bb:

𝔼[N(K1,2,Gn,u¯)]x,y,z[n]uxyuyzy[n](x[n]uxy)2Dn2p3/2b=o(n3p2).\mathbb{E}[N(K_{1,2},G_{n,\bar{u}})]\leq\sum_{x,y,z\in[n]}u_{xy}u_{yz}\leq\sum_{y\in[n]}\left(\sum_{x\in[n]}u_{xy}\right)^{2}\leq Dn^{2}p^{3/2}b=o(n^{3}p^{2}).\qed