组合数学 (Fall 2011)/Extremal graph theory and 组合数学 (Fall 2011)/Extremal set theory: Difference between pages

From TCS Wiki
(Difference between pages)
Jump to navigation Jump to search
imported>Etone
 
imported>Etone
 
Line 1: Line 1:
== Forbidden Cliques ==
== Sperner system ==
Extremal graph theory studies the problems like  "how many edges that a graph <math>G</math> can have, if <math>G</math> has some property?"
A set family <math>\mathcal{F}\subseteq 2^X</math> with the relation <math>\subseteq</math> define a poset. Thus, a '''chain''' is a sequence <math>S_1\subseteq S_2\subseteq\cdots\subseteq S_k</math>.
=== Mantel's theorem ===
We consider a typical extremal problem for graphs: the largest possible number of edges of '''triangle-free''' graphs, i.e. graphs contains no <math>K_3</math>.


{{Theorem|Theorem (Mantel 1907)|
A set family <math>\mathcal{F}\subseteq 2^X</math> is an '''antichain''' (also called a '''Sperner system''') if for all <math>S,T\in\mathcal{F}</math> that <math>S\neq T</math>, we have <math>S\not\subseteq T</math>.
:Suppose <math>G(V,E)</math> is graph on <math>n</math> vertice without triangles. Then <math>|E|\le\frac{n^2}{4}</math>.
}}


We give three different proofs of the theorem. The first one uses induction and an argument based on pigeonhole principle. The second proof uses the famous Cauchy-Schwarz inequality in analysis. And the third proof uses another famous inequality: the inequality of the arithmetic and geometric mean.
The <math>k</math>-uniform <math>{X\choose k}</math> is an antichain. Let <math>n=|X|</math>. The size of <math>{X\choose k}</math> is maximized when <math>k=\lfloor n/2\rfloor</math>. We wonder whether this is also the largest possible size of any antichain <math>\mathcal{F}\subseteq 2^X</math>.


{{Prooftitle|First proof. (pigeonhole principle)|
In 1928, Emanuel Sperner proved a theorem saying that it is indeed the largest possible antichain. This result, called Sperner's theorem today, initiated the studies of extremal set theory.
We prove an equivalent theorem: Any <math>G(V,E)</math> with <math>|V|=n</math> and <math>|E|>\frac{n^2}{4}</math> must have a triangle.


Use induction on <math>n</math>. The theorem holds trivially for <math>n\le 3</math>.
{{Theorem|Theorem (Sperner 1928)|
:Let <math>\mathcal{F}\subseteq 2^X</math> where <math>|X|=n</math>. If <math>\mathcal{F}</math> is an antichain, then
::<math>|\mathcal{F}|\le{n\choose \lfloor n/2\rfloor}</math>.
}}


Induction hypothesis: assume the theorem hold for <math>|V|\le n-1</math>.  
=== First proof (shadows)===
We first introduce the original proof by Sperner, which uses concepts called '''shadows''' and '''shades''' of set systems.


For <math>G</math> with <math>n</math> vertices, without loss of generality, assume that <math>|E|=\frac{n^2}{4}+1</math>, we will show that <math>G</math> must contain a triangle. Take a <math>uv\in E</math>, and let <math>H</math> be the subgraph of <math>G</math> induced by <math>V\setminus \{u,v\}</math>. Clearly, <math>H</math> has <math>n-2</math> vertices.
{{Theorem|Definition|
:'''Case.1:''' If <math>H</math> has <math>>\frac{(n-2)^2}{4}</math> edges, then by the induction hypothesis, <math>H</math> has a triangle.
:Let <math>|X|=n\,</math> and <math>\mathcal{F}\subseteq {X\choose k}</math>, <math>k<n\,</math>.
:'''Case.2:''' If <math>H</math> has <math>\le\frac{(n-2)^2}{4}</math> edges, then at least <math>\left(\frac{n^2}{4}+1\right)-\frac{(n-2)^2}{4}-1=n-1</math> edges are between <math>H</math> and <math>\{u,v\}</math>. By pigeonhole principle, there must be a vertex in <math>H</math> that is adjacent to both <math>u</math> and <math>v</math>. Thus, <math>G</math> has a triangle.
:The '''shade''' of <math>\mathcal{F}</math> is defined to be
::<math>\nabla\mathcal{F}=\left\{T\in {X\choose k+1}\,\,\bigg|\,\, \exists S\in\mathcal{F}\mbox{ such that } S\subset T\right\}</math>.
:Thus the shade <math>\nabla\mathcal{F}</math> of <math>\mathcal{F}</math> consists of all subsets of <math>X</math> which can be obtained by adding an element to a set in <math>\mathcal{F}</math>.
:Similarly, the '''shadow''' of <math>\mathcal{F}</math> is defined to be
::<math>\Delta\mathcal{F}=\left\{T\in {X\choose k-1}\,\,\bigg|\,\, \exists S\in\mathcal{F}\mbox{ such that } T\subset S\right\}</math>.
:Thus the shadow <math>\Delta\mathcal{F}</math> of <math>\mathcal{F}</math> consists of all subsets of <math>X</math> which can be obtained by removing an element from a set in <math>\mathcal{F}</math>.
}}
}}


{{Prooftitle|Second proof. (Cauchy-Schwarz inequality)|(Mantel's original proof)
Next lemma bounds the effects of shadows and shades on the sizes of set systems.
For any edge <math>uv\in E</math>, no vertex can be a neighbor of both <math>u</math> and <math>v</math>, or otherwise there will be a triangle. Thus, for any edge <math>uv\in E</math>, <math>d_u+d_v\le n</math>. It follows that
 
:<math>\sum_{uv\in E}(d_u+d_v)\le n|E|</math>.
{{Theorem|Lemma (Sperner)|
Note that <math>d(v)</math> appears exactly <math>d_v</math> times in the sum, so that
:Let <math>|X|=n\,</math> and <math>\mathcal{F}\subseteq {X\choose k}</math>. Then
:<math>\sum_{uv\in E}(d_u+d_v)=\sum_{v\in V}d_v^2</math>.
::<math>
Applying Chauchy-Schwarz inequality,
\begin{align}
:<math>
&|\nabla\mathcal{F}|\ge\frac{n-k}{k+1}|\mathcal{F}| &\text{ if } k<n\\
n|E|\ge \sum_{uv\in E}(d_u+d_v)=\sum_{v\in V}d_v^2\ge\frac{\left(\sum_{v\in V}d_v\right)^2}{n}=\frac{4|E|^2}{n},
&|\Delta\mathcal{F}|\ge\frac{k}{n-k+1}|\mathcal{F}| &\text{ if } k>0.
\end{align}
</math>
</math>
where the last equation is due to Euler's equality <math>\sum_{v\in V}d_v=2|E|</math>. The theorem follows.
}}
}}


{{Prooftitle|Third proof. (inequality of the arithmetic and geometric mean)|
{{Proof|
Assume that <math>G(V,E)</math> has <math>|V|=n</math> vertices and is triangle-free.
The lemma is proved by double counting. We prove the inequality of <math>|\nabla\mathcal{F}|</math>. Assume that <math>0\le k<n</math>.


Let <math>A</math> be the largest independent set in <math>G</math> and let <math>\alpha=|A|</math>.
Define
Since <math>G</math> is triangle-free, for very vertex <math>v</math>, all its neighbors must form an independent set, thus <math>d(v)\le \alpha</math> for all <math>v\in V</math>.
:<math>\mathcal{R}=\{(S,T)\mid S\in\mathcal{F}, T\in\nabla\mathcal{F}, S\subset T\}</math>.
We estimate <math>|\mathcal{R}|</math> in two ways.  


Take <math>B=V\setminus A</math> and let <math>\beta=|B|</math>.
For each <math>S\in\mathcal{F}</math>, there are <math>n-k</math> different <math>T\in\nabla\mathcal{F}</math> that <math>S\subset T</math>.
Since <math>A</math> is an independent set, all edges in <math>E</math> must have at least one endpoint in <math>B</math>. Counting the edges in <math>E</math> according to their endpoints in <math>B</math>, we obtain <math>|E|\le\sum_{v\in B}d_v</math>. By the inequality of the arithmetic and geometric mean,
:<math>|\mathcal{R}|=(n-k)|\mathcal{F}|</math>.
:<math>|E|\le\sum_{v\in B}d_v\le\alpha\beta\le\left(\frac{\alpha+\beta}{2}\right)^2=\frac{n^2}{4}</math>.
For each <math>T\in\nabla\mathcal{F}</math>, there are <math>k+1</math> ways to choose an <math>S\subset T</math> with <math>|S|=k</math>, some of which may not be in <math>\mathcal{F}</math>.
:<math>|\mathcal{R}|\le (k+1)|\nabla\mathcal{F}|</math>.
 
Altogether, we show that <math>|\nabla\mathcal{F}|\ge\frac{n-k}{k+1}|\mathcal{F}|</math>.
 
The inequality of <math>|\Delta\mathcal{F}|</math> can be proved in the same way.
}}
}}


=== Turán's theorem ===
An immediate corollary of the previous lemma is as follows.
The famous Turán's theorem generalizes the Mantel's theorem for triangles to cliques of any specific size. This theorem is one of the most important results in extremal combinatorics, which initiates the studies of extremal graph theory.
 
{{Theorem|Theorem (Turán 1941)|
{{Theorem|Proposition 1|
:Let <math>G(V,E)</math> be a graph with <math>|V|=n</math>. If <math>G</math> has no <math>r</math>-clique, <math>r\ge 2</math>, then
:If <math>k\le \frac{n-1}{2}</math>, then <math>|\nabla\mathcal{F}|\ge|\mathcal{F}|</math>.
::<math>|E|\le\frac{r-2}{2(r-1)}n^2</math>.
:If <math>k\ge \frac{n-1}{2}</math>, then <math>|\Delta\mathcal{F}|\ge|\mathcal{F}|</math>.
}}
}}


We give an example of graphs with many edges which does not contain <math>K_r</math>.
The idea of Sperner's proof is pretty clear:
* we "push up" all the sets in <math>\mathcal{F}</math> of size <math><\frac{n-1}{2}</math> replacing them by their shades;
* and also "push down" all the sets in <math>\mathcal{F}</math> of size <math>\ge\frac{n+1}{2}</math> replacing them by their shadows.
Repeat this process we end up with a set system <math>\mathcal{F}\subseteq{X\choose \lfloor n/2\rfloor}</math>. We need to show that this process does not decrease the size of <math>\mathcal{F}</math>.
 
{{Theorem|Proposition 2|
:Suppose that <math>\mathcal{F}\subseteq2^X</math> where <math>|X|=n</math>. Let <math>\mathcal{F}_k=\mathcal{F}\ap{X\choose k}</math>. Let <math>k_\min</math> be the smallest <math>k</math> that <math>|\mathcal{F}_k|>0</math>, and let
::<math>
\mathcal{F}'=\begin{cases}
\mathcal{F}\setminus\mathcal{F}_{k_\min}\cup \nabla\mathcal{F}_{k_\min} & \mbox{if }k_\min<\frac{n-1}{2},\\
\mathcal{F} & \mbox{otherwise.}
\end{cases}
</math>
:Similarly, let <math>k_\max</math> be the largest <math>k</math> that <math>|\mathcal{F}_k|>0</math>, and let
::<math>
\mathcal{F}''=\begin{cases}
\mathcal{F}\setminus\mathcal{F}_{k_\max}\cup \Delta\mathcal{F}_{k_\max} & \mbox{if }k_\max\ge\frac{n+1}{2},\\
\mathcal{F} & \mbox{otherwise.}
\end{cases}
</math>
:If <math>\mathcal{F}</math> is an antichain, <math>\mathcal{F}'</math> and <math>\mathcal{F}''</math> are antichains, and we have <math>|\mathcal{F}'|\ge|\mathcal{F}|</math> and <math>|\mathcal{F}''|\ge|\mathcal{F}|</math>.
}}
{{Proof|
We show that <math>\mathcal{F}'</math> is an antichain and <math>|\mathcal{F}'|\ge|\mathcal{F}|</math>.  


Partition <math>V</math> into <math>r-1</math> disjoint classes <math>V=V_1\cup V_2\cup\cdots\cup V_{r-1}</math>, <math>n_i=|V_i|</math>, <math>n_1+n_2+\cdots+n_{r-1}=n</math>. For every two vertice <math>u,v</math>, <math>uv\in E</math> if and only if <math>u\in V_i</math> and <math>v\in V_j</math> for distinct <math>V_i</math> and <math>V_j</math>. The resulting graph is a '''complete <math>(r-1)</math>-partite graph''', denoted <math>K_{n_1,n_2,\ldots,n_{r-1}}</math>. It is obvious that any <math>(r-1)</math>-partite graph contains no <math>r</math>-clique since only those vertices from different classes can be adjacent.  
First, observe that <math>\nabla\mathcal{F}_k\cap\mathcal{F}=\emptyset</math>, otherwise <math>\mathcal{F}</math> cannot be an antichain, and due to Proposition 1, <math>|\nabla\mathcal{F}_k|\ge|\mathcal{F}_k|</math> when <math>k\le \frac{n-1}{2}</math>, so <math>|\mathcal{F}'|=|\mathcal{F}|-|\mathcal{F}_k|+|\nabla\mathcal{F}_k|\ge |\mathcal{F}|</math>.  


A <math>K_{n_1,n_2,\ldots,n_{r-1}}</math> has <math>\sum_{i<j}n_i n_j\,</math> edges, which is maximized when the numbers <math>n_i</math> are divided as evenly as possible, that is, if <math>n_i\in\left\{\left\lfloor\frac{n}{r-1}\right\rfloor,\left\lceil\frac{n}{r-1}\right\rceil\right\}</math> for every <math>1\le i\le r-1</math>.  
Now we prove that <math>\mathcal{F}'</math> is an antichain . By contradiction, assume that there are <math>S, T\in \mathcal{F}'</math>, such that <math>S\subset T</math>. One of the <math>S,T</math> must be in <math>\nabla\mathcal{F}_{k_\min}</math>, or otherwise <math>\mathcal{F}</math> cannot be an antichain. Recall that <math>k_\min</math> is the smallest <math>k</math> that <math>|\mathcal{F}_k|>0</math>, thus it must be <math>S\in \nabla\mathcal{F}_{k_\min}</math>, and <math>T\in\mathcal{F}</math>. This implies that there is an <math>R\in \mathcal{F}_{k_\min}\subseteq \mathcal{F}</math> such that <math>R\subset S\subset T</math>, which contradicts that <math>\mathcal{F}</math> is an antichain.


{{Theorem|Definition|
The statement for <math>\mathcal{F}''</math> can be proved in the same way.
:We call a complete multipartite graph <math>K_{n_1,n_2,\ldots,n_{r-1}}</math> with <math>n_i\in\left\{\left\lfloor\frac{n}{r-1}\right\rfloor,\left\lceil\frac{n}{r-1}\right\rceil\right\}</math> for every <math>i</math> a ''' Turán graph''', denoted <math>T(n,r-1)</math>.
}}
}}
;Example:Turán graph <math>T(13,4)</math>
[[File:Turan 13-4.svg|center|260px|Turán graph <math>T(13,4)</math>]]


Turán's theorem has been proved for many times by different mathematicians, with different tools. We show just a few.
Applying the above process, we prove the Sperner's theorem.
{{Prooftitle|Proof of Sperner's theorem | (original proof of Sperner)
Let <math>\mathcal{F}_k=\{S\in\mathcal{F}\mid |S|=k\}</math>, where <math>0\le k\le n</math>.


The first proof uses induction;  the second proof uses a technique called "weight shifting"; and the third proof uses the probabilistic method. All of them are very powerful and frequently used proof techniques.
We change <math>\mathcal{F}</math> as follows:
* for the smallest <math>k</math> that <math>|\mathcal{F}_k|>0</math>, if <math>k<\frac{n-1}{2}</math>, replace <math>\mathcal{F}_k</math> by <math>\nabla\mathcal{F}_k</math>.


{{Prooftitle|First proof. (induction)|(Turán's original proof)
Due to Proposition 2, this procedure preserves <math>\mathcal{F}</math> as an antichain and does not decrease <math>|\mathcal{F}|</math>. Repeat this procedure, until <math>|\mathcal{F}_k|=0</math> for all <math>k<\frac{n-1}{2}</math>, that is, there is no member set of <math>\mathcal{F}</math> has size less than <math>\frac{n-1}{2}</math>.


Induction on <math>n</math>. It is easy to verify that the theorem holds for <math>n<r</math>.  
We then define another symmetric procedure:
* for the largest <math>k</math> that <math>|\mathcal{F}_k|>0</math>, if <math>k\ge\frac{n+1}{2}</math>, replace <math>\mathcal{F}_k</math> by <math>\Delta\mathcal{F}_k</math>.
Also due to Proposition 2, this procedure preserves <math>\mathcal{F}</math> as an antichain and does not decrease <math>|\mathcal{F}|</math>. After repeatedly applying this procedure, <math>|\mathcal{F}_k|=0</math> for all <math>k\ge\frac{n+1}{2}</math>.  


Let <math>G</math> be a graph on <math>n</math> vertices without <math>r</math>-cliques where <math>n\ge r</math>. Suppose that <math>G</math> has a maximum number of edges among such graphs. <math>G</math> certainly has <math>(r-1)</math>-cliques, since otherwise we could add edges to <math>G</math>. Let <math>A</math> be an <math>(r-1)</math>-clique and let <math>B=V\setminus A</math>. Clearly <math>|A|=r-1</math> and <math>|B|=n-r+1</math>.
The resulting <math>\mathcal{F}</math> has <math>\mathcal{F}\subseteq{X\choose \lfloor n/2\rfloor}</math>, and since <math>|\mathcal{F}|</math> is never decreased, for the original <math>\mathcal{F}</math>, we have
 
:<math>|\mathcal{F}|\le {n\choose \lfloor n/2\rfloor}</math>.
By the  induction hypothesis, since <math>B</math> has no <math>r</math>-cliques, <math>|E(B)|\le\frac{r-2}{2(r-1)}(n-r+1)^2</math>. And <math>E(A)={r-1\choose 2}</math>. Since <math>G</math> has no <math>r</math>-clique, every <math>v\in B</math> is adjacent to at most <math>r-2</math> vertices in <math>A</math>, since otherwise <math>A</math> and <math>v</math> would form an <math>r</math>-clique. We obtain that the number edges crossing between <math>A</math> and <math>B</math> is <math>|E(A,B)|\le (r-2)|B|=(r-2)(n-r+1)</math>. Combining everything together,
:<math>|E|=|E(A)|+|E(B)|+|E(A,B)|\le {r-1\choose 2}+\frac{r-2}{2(r-1)}(n-r+1)^2+(r-2)(n-r+1)=\frac{r-2}{2(r-1)}n^2</math>.
}}
}}


{{Prooftitle|Second proof. (weight shifting)|(due to Motzkin and Straus)
=== Second proof (counting)===
We now introduce an elegant proof due to Lubell. The proof uses a counting argument, and tells more information than just the size of the set system.


Assign each vertex <math>v\in V</math> a nonnegative weight <math>w_v\ge 0</math>, and assume that <math>\sum_{v\in V}w_v=1</math>. We try to maximize the quantity
{{Prooftitle|Proof of Sperner's theorem | (Lubell 1966)
:<math>S=\sum_{uv\in E}w_uw_v</math>.
Let <math>\pi</math> be a permutation of <math>X</math>. We say that an <math>S\subseteq X</math> '''prefixes''' <math>\pi</math>, if <math>S=\{\pi_1,\pi_2,\ldots, \pi_{|S|}\}</math>, that is, <math>S</math> is precisely the set of the first <math>|S|</math> elements in the permutation <math>\pi</math>.
Let <math>W_u=\sum_{v:v\sim u}w_v\,</math> be the sum of the weights of <math>u</math>'s neighbors.
Note that <math>S</math> can also be computed as <math>S=\frac{1}{2}\sum_{u\in V}w_uW_u</math>.
For any nonadjacent pair of vertices <math>u\not\sim v</math>, supposed that <math>W_u\ge W_v</math>, then for any <math>\epsilon\ge 0</math>,
:<math>(w_u+\epsilon)W_u+(w_v-\epsilon)W_v\ge w_uW_u+w_vW_v</math>.
This means that we do not decrease <math>S</math> by shifting all of the weight of the vertex <math>v</math> to the vertex <math>u</math>. It follows that <math>S</math> is maximized when all of the weight is concentrated on a complete subgraph, i.e., a clique.


Now if <math>w_u>w_v>0</math>, then choose <math>\epsilon</math> with <math>0<\epsilon<w_u-w_v</math> and change <math>w_u'=w_u-\epsilon</math> and <math>w_v'=w_v+\epsilon</math>. This changes <math>S</math> to <math>S'=S+\epsilon(w_u-w_v)-\epsilon^2>S</math>. Thus, the maximal value of <math>S</math> is attained when all nonzero weights are equal and concentrated on a clique.
Fix an <math>S\subseteq X</math>. It is easy to see that the number of permutations <math>\pi</math> of <math>X</math> prefixed by <math>S</math> is <math>|S|!(n-|S|)!</math>.  Also, since <math>\mathcal{F}</math> is an antichain, no permutation <math>\pi</math> of <math>X</math> can be prefixed by more than one members of <math>\mathcal{F}</math>, otherwise one of the member sets must contain the other, which contradicts that <math>\mathcal{F}</math> is an antichain. Thus, the number of permutations <math>\pi</math> prefixed by some <math>S\in\mathcal{F}</math> is
:<math>\sum_{S\in\mathcal{F}}|S|!(n-|S|)!</math>,
which cannot be larger than the total number of permutations, <math>n!</math>, therefore,
:<math>\sum_{S\in\mathcal{F}}|S|!(n-|S|)!\le n!</math>.
Dividing both sides by <math>n!</math>, we have
:<math>\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}=\sum_{S\in\mathcal{F}}\frac{|S|!(n-|S|)!}{n!}\le 1</math>,
where <math>{n\choose |S|}\le {n\choose \lfloor n/2\rfloor}</math>, so
:<math>\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\ge \frac{|\mathcal{F}|}{{n\choose \lfloor n/2\rfloor}}</math>.
Combining this with the above inequality, we prove the Sperner's theorem.
}}


<math>G</math> has at most an <math>(r-1)</math>-clique, thus <math>S\le{r-1\choose 2}\frac{1}{(r-1)^2}=\frac{r-2}{2(r-1)}</math>.
=== The LYM inequality ===
Lubell's proof proves the following inequality:
:<math>\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le 1</math>
which is actually stronger than Sperner's original statement that <math>|\mathcal{F}|\le{n\choose \lfloor n/2\rfloor}</math>.


As we argued above, this inequality hold for any nonnegative weight assignments with <math>\sum_{v\in V}w_v=1</math>. In particular, for the case that all <math>w_v=\frac{1}{n}</math>,
This inequality is independently discovered by Lubell-Yamamoto, Meschalkin, and Bollobás, and is called the LYM inequality today.
:<math>S=\sum_{uv\in E}w_uw_v=\frac{|E|}{n^2}</math>.
 
Thus,
{{Theorem|Theorem (Lubell, Yamamoto 1954; Meschalkin 1963)|
:<math>\frac{|E|}{n^2}\le \frac{r-2}{2(r-1)}</math>,
:Let <math>\mathcal{F}\subseteq 2^X</math> where <math>|X|=n</math>. If <math>\mathcal{F}</math> is an antichain, then
which implies the theorem.
::<math>\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le 1</math>.
}}
}}


{{Prooftitle|Third proof. (the probabilistic method)|(due to Alon and Spencer)
In Lubell's counting argument proves the LYM inequality, which implies the Sperner's theorem. Here we give another proof of the LYM inequality by the probabilistic methoddue to Noga Alon.


Write <math>\omega(G)</math> for the number of vertices in a largest clique, called the '''clique number''' of <math>G</math>.  
{{Prooftitle|Third proof (the probabilistic method)| (Due to Alon.)
:'''Claim:''' <math>\omega(G)\ge\sum_{v\in V}\frac{1}{n-d_v}</math>.
Let <math>\pi</math> be a uniformly random permutation of <math>X</math>. Define a random maximal chain by
We prove this by the probabilistic method. Fix a random ordering of vertices in <math>V</math>, say <math>v_1,v_2,\ldots,v_n</math>. We construct a clique as follows:
:<math>\mathcal{C}_\pi=\{\{\pi_i\mid 1\le i\le k\}\mid 0\le k\le n\}</math>.
*for <math>i=1,2,\ldots, n</math>, add <math>v_i</math> to <math>S</math> iff all vertices in current <math>S</math> are adjacent to <math>v_i</math>.
For any <math>S\in\mathcal{F}</math>, let <math>X_S</math> be the 0-1 random variable which indicates whether <math>S\in\mathcal{C}_\pi</math>, that is
It is obvious that an <math>S</math> constructed in this way is a clique. We now show that <math>\mathbf{E}[|S|]=\sum_{v\in V}\frac{1}{n-d_v}</math>.
 
Let <math>X_v</math> be the random variable that indicates whether <math>v\in S</math>, i.e.,
:<math>
:<math>
X_v=\begin{cases}
X_S=\begin{cases}
1 & v\in S,\\
1 & \mbox{if }S\in\mathcal{C}_\pi,\\
0 & \mbox{otherwise.}
0 & \mbox{otherwise.}
\end{cases}
\end{cases}
</math>
</math>
Note that a vertex <math>v\in S</math> if and only if <math>v</math> is ranked before all its <math>n-d_v-1</math> non-neighbors in the random ordering. The probability that this event occurs is <math>\frac{1}{n-d_v}</math>. Thus,
Note that for a uniformly random <math>\pi</math>, <math>\mathcal{C}_\pi</math> has exact one member set of size <math>|S|</math>, uniformly distributed over <math>{X\choose |S|}</math>, thus
:<math>\mathbf{E}[X_v]=\Pr[v\in S]=\frac{1}{n-d_v}.</math>
:<math>\mathbf{E}[X_S]=\Pr[S\in\mathcal{C}_\pi]=\frac{1}{{n\choose |S|}}</math>.
Observe that <math>|S|=\sum_{v\in V}X_v</math>. Due to linearity of expectation,
Let <math>X=\sum_{S\in\mathcal{F}}X_S</math>. Note that <math>X=|\mathcal{F}\cap\mathcal{C}_\pi|</math>. By the linearity of expectation,
:<math>\mathbf{E}[|S|]=\sum_{v\in V}\mathbf{E}[X_v]=\sum_{v\in V}\frac{1}{n-d_v}</math>.
:<math>\mathbf{E}[X]=\sum_{S\in\mathcal{F}}\mathbf{E}[X_S]=\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}</math>.
There must exists a clique of at least such size, so that <math>\omega(G)\ge\sum_{v\in V}\frac{1}{n-d_v}</math>. The claim is proved.
On the other hand, since <math>\mathcal{F}</math> is an antichain, it can never intersect a chain at more than one elements, thus we always have <math>X=|\mathcal{F}\cap\mathcal{C}_\pi|\le 1</math>. Therefore,
:<math>\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le \mathbf{E}[X] \le 1</math>.
}}
 
The Sperner's theorem is an immediate consequence of the LYM inequality.


Apply the Cauchy-Schwarz inequality
{{Theorem|Proposition|
:<math>\left(\sum_{v\in V}a_vb_v\right)^2\le\left(\sum_{v\in V}^na_v^2\right)\left(\sum_{v\in V}^nb_v^2\right)</math>.
:<math>\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le 1</math> implies that <math>|\mathcal{F}|\le{n\choose \lfloor n/2\rfloor}</math>.
Set <math>a_v=\sqrt{n-d_v}</math> and <math>b_v=\frac{1}{\sqrt{n-d_v}}</math>, then <math>a_vb_v=1</math> and so
}}
:<math>n^2\le\sum_{v\in V}(n-d_v)\sum_{v\in V}\frac{1}{n-d_v}\le\omega(G)\sum_{v\in V}(n-d_v).</math>
{{Proof|
By the assumption of Turán's theorem, <math>\omega(G)\le r-1</math>. Recall the handshaking lemma <math>2|E|=\sum_{v\in V}d_v</math>. The above inequality gives us
It holds that <math>{n\choose k}\le {n\choose \lfloor n/2\rfloor}</math> for any <math>k</math>. Thus,
:<math>n^2\le (r-1)(n^2-2|E|)</math>,
:<math>1\ge \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\ge \frac{|\mathcal{F}|}{{n\choose \lfloor n/2\rfloor}}</math>,
which implies the theorem.
which implies that <math>|\mathcal{F}|\le {n\choose \lfloor n/2\rfloor}</math>.
}}
}}


Our last proof uses the idea of vertex duplication. It does not only prove the edge bound of Turán's theorem, but also shows that Turán graphs are the <font color=red>only</font> possible extremal graphs.
== Sunflowers ==
{{Prooftitle|Fourth proof.|
An set system is a '''sunflower''' if all its member sets intersect at the same set of elements.
Let <math>G(V,E)</math> be a <math>r</math>-clique-free graph on <math>n</math> vertices with a maximum number of edges.
{{Theorem|Definition (sunflower)|
:'''Claim:''' <math>G</math> does not contain three vertices <math>u,v,w</math> such that <math>uv\in E</math> but <math>uw\not\in E, vw\not\in E</math>.
: A set family <math>\mathcal{F}\subseteq 2^X</math> is a '''sunflower''' of size <math>r</math> with a '''core''' <math>C\subseteq X</math> if
Suppose otherwise. There are two cases.
::<math>\forall S,T\in\mathcal{F}</math> that <math>S\neq T</math>, <math>S\cap T=C</math>.
* '''Case.1:''' <math>d(w)<d(u)</math> or <math>d(w)<d(v)</math>. Without loss of generality, suppose that <math>d(w)<d(u)</math>. We duplicate <math>u</math> by creating a new vertex <math>u'</math> which has exactly the same neighbors as <math>u</math> (but <math>uu'</math> is not an edge). Such duplication will not increase the clique size. We then remove <math>w</math>. The resulting graph <math>G'</math> is still <math>r</math>-clique-free, and has <math>n</math> vertices. The number of edges in <math>G'</math> is
}}
::<math>|E(G')|=|E(G)|+d(u)-d(w)>|E(G)|\,</math>,
Note that we do not require the core to be nonempty, thus a family of disjoint sets is also a sunflower (with the core <math>\emptyset</math>).
:which contradicts the assumption that <math>|E(G)|</math> is maximal.
* '''Case.2:''' <math>d(w)\ge d(u)</math> and <math>d(w)\ge d(v)</math>. Duplicate <math>w</math> twice and delete <math>u</math> and <math>v</math>. The new graph <math>G'</math> has no <math>r</math>-clique, and the number of edges is
::<math>|E(G')|=|E(G)|+2d(w)-(d(u)+d(v)+1)>|E(G)|\,</math>.
:Contradiction again.


The claim implies that <math>uv\not\in E</math> defines an equivalence relation on vertices (to be more precise, it guarantees the transitivity of the relation, while the reflexivity and symmetry hold directly). Graph <math>G</math> must be a complete multipartite graph <math>K_{n_1,n_2,\ldots,n_{r-1}}</math> with <math>n_1+n_2+\cdots +n_{r-1}=n</math>. Optimize the edge number, we have the Turán graph.
The next result due to Erdős and Rado, called the sunflower lemma, is a famous result in extremal set theory, and has some important applications in Boolean circuit complexity.
{{Theorem|Sunflower Lemma (Erdős-Rado)|
:Let <math>\mathcal{F}\subseteq {X\choose k}</math>. If <math>|\mathcal{F}|>k!(r-1)^k</math>, then <math>\mathcal{F}</math> contains a sunflower of size  <math>r</math>.
}}
}}
{{Proof|
We proceed by induction on <math>k</math>. For <math>k=1</math>, <math>\mathcal{F}\subseteq{X\choose 1}</math>, thus all sets in <math>\mathcal{F}</math> are disjoint. And since <math>|\mathcal{F}|>r-1</math>, we can choose <math>r</math> of these sets and form a sunflower.
Now let <math>k\ge 2</math> and assume the lemma holds for all smaller <math>k</math>. Take a maximal family <math>\mathcal{G}\subseteq \mathcal{F}</math> whose members are disjoint, i.e. for any <math>S,T\in \mathcal{G}</math> that <math>S\neq T</math>, <math>S\cap T=\emptyset</math>.
If <math>|\mathcal{G}|\ge r</math>, then <math>\mathcal{G}</math> is a sunflower of size at least <math>r</math> and we are done.


== Forbidden Cycles ==
Assume that <math>|\mathcal{G}|\le r-1</math>, and let <math>Y=\bigcup_{S\in\mathcal{G}}S</math>. Then <math>|Y|=k|\mathcal{G}|\le k(r-1)</math> (since all members of <math>\mathcal{G}</math>) are disjoint). We claim that <math>Y</math> intersets all members of <math>\mathcal{F}</math>, since if otherwise, there exists an <math>S\in\mathcal{F}</math> such that <math>S\cap Y=\emptyset</math>, then we can enlarge <math>\mathcal{G}</math> by adding <math>S</math> into <math>\mathcal{G}</math> and still have all members of <math>\mathcal{G}</math> disjoint, which contradicts the assumption that <math>\mathcal{G}</math> is the maximum of such families.
Another direction to generalize Mantel's theorem other than Turán's theorem is to see a triangle as a 3-cycle rather than 3-clique. We then ask for the extremal bound for graphs without certain cycle structures.
=== Girth ===
Recall that the '''girth''' of a graph <math>G</math> is the length of the shortest cycle in <math>G</math>. A graph is triangle-free if and only if its girth <math>g(G)\ge 4</math>.
Matel's theorem can be seen as a bound on the edge number of graphs with girth <math>g(G)\ge 4</math>. The next theorem extends this bound to the graphs with <math>g(G)\ge 5</math>, i.e., graphs without triangles and quadrilaterals ("squares").


{{Theorem|Theorem|
By the pigeonhole principle, some elements <math>y\in Y</math> must contained in at least
:Let <math>G(V,E)</math> be a graph on <math>n</math> vertices. If girth <math>g(G)\ge 5</math> then <math>|E|\le\frac{1}{2}n\sqrt{n-1}</math>.
:<math>\frac{|\mathcal{F}|}{|Y|}>\frac{k!(r-1)^k}{k(r-1)}=(k-1)!(r-1)^{k-1}</math>
members of <math>\mathcal{F}</math>. We delete this <math>y</math> from these sets and consider the family
:<math>\mathcal{H}=\{S\setminus\{y\}\mid S\in\mathcal{F}\wedge y\in S\}</math>.
We have <math>\mathcal{H}\subseteq {X\choose k-1}</math> and <math>|\mathcal{H}|>(k-1)!(r-1)^{k-1}</math>, thus by the induction hypothesis, <math>\mathcal{H}</math>contains a sunflower of size <math>r</math>. Adding <math>y</math> to the members of this sunflower, we get the desired sunflower in the original family <math>\mathcal{F}</math>.
}}
}}
{{Proof|
Suppose <math>g(G)\ge 5</math>. Let <math>v_1,v_2,\ldots,v_d</math> be the neighbors of a vertex <math>u</math>, where <math>d=d(u)</math>. Let <math>S_i=\{v\in V\mid v\sim v_i\wedge v\neq u\}</math> be the set of neighbors of <math>v_i</math> other than <math>u</math>.


* For any <math>v_i,v_j</math>, <math>v_iv_j\not\in E</math> since <math>G</math> has no triangle. Thus, <math>S_i\cap\{u,v_1,v_2,\ldots,v_d\}=\emptyset</math> for every <math>i</math>.
==The Erdős–Ko–Rado Theorem ==
* No vertex other than <math>u</math> can be adjacent to more than one vertices in <math>v_1,v_2,\ldots,v_d</math> since there is no <math>C_4</math> in <math>G</math>. Thus, <math>S_i\cap S_j=\emptyset</math> for any distinct <math>i</math> and <math>j</math>.
A set family <math>\mathcal{F}\subseteq 2^X</math> is called '''intersecting''', if for any <math>S,T\in\mathcal{F}</math>, <math>S\cap T\neq\emptyset</math>. A natural question of extremal favor is: "how large can an intersecting family be?"


Therefore, <math>\{u,v_1,v_2,\ldots,v_d\}\cap S_1\cup S_2\cup\cdots\cup S_d\subseteq V</math> implies that
Assume <math>|X|=n</math>. When <math>n<2k</math>, every pair of <math>k</math>-subsets of <math>X</math> intersects. So the non-trivial case is when <math>n\le 2k</math>. The famous Erdős–Ko–Rado theorem gives the largest possible cardinality of a nontrivially intersecting family.  
:<math>(d+1)+|S_1|+|S_2|+\cdots+|S_d|=(d+1)+(d(v_1)-1)+(d(v_2)-1)+\cdots+(d(v_d)-1)\le n</math>,
so that <math>\sum_{v:v\sim u}d(v)\le n-1</math>.


By Cauchy-Schwarz inequality,
According to Erdős, the theorem itself was proved in 1938, but was not published until 23 years later.
:<math>n(n-1)\ge \sum_{u\in V}\sum_{v:v\sim u}d(v)=\sum_{v\in V}d(v)^2\ge\frac{\left(\sum_{v\in V}d(v)\right)}{n}=\frac{4|E|^2}{n}</math>,
{{Theorem|Erdős–Ko–Rado theorem (proved in 1938, published in 1961)|
which implies that <math>|E|\le\frac{1}{2}n\sqrt{n-1}</math>.
:Let <math>\mathcal{F}\subseteq {X\choose k}</math> where <math>|X|=n</math> and <math>n\ge 2k</math>. If <math>\mathcal{F}</math> is intersecting, then
::<math>|\mathcal{F}|\le{n-1\choose k-1}</math>.
}}
}}


=== Hamiltonian cycle ===
=== Katona's proof ===
We now look at graphs which does not have large cycles. In particular, we consider graphs without '''Hamiltonian cycles'''.
We first introduce a proof discovered by Katona in 1972. The proof uses double counting.


For a Hamiltonian graph, every vertex must has degree 2. And the graph satisfying this condition with maximum number of edges is the graph composed by a <math>(n-1)</math>-clique and the one remaining vertex is connected to the clique by one edge. This graph has <math>{n-1\choose 2}+1</math> edges, and has no Hamiltonian cycle. It is not very hard to realize that this is the largest possible number of edges that a non-Hamiltonian graph can have.
Let <math>\pi</math> be a '''cyclic permutation''' of <math>X</math>, that is, we think of assigning <math>X</math> in a circle and ignore the rotations of the circle. It is easy to see that there are <math>(n-1)!</math> cyclic permutations of an <math>n</math>-set (each cyclic permutation corresponds to <math>n</math> permutations).
Let
:<math>\mathcal{G}_\pi=\{\{\pi_{(i+j)\bmod n}\mid j\in[k]\}\mid i\in [n]\}</math>.


Since it is not very interesting to bound the number of edges of non-Hamiltonian graphs, we consider a more informative graph invariant, its '''degree sequence'''.
The next lemma states the following observation: in a circle of <math>n</math> points, supposed <math>n\ge 2k</math>, there can be at most <math>k</math> arcs, each consisting of <math>k</math> points, such that every pair of arcs share at least one point.
{{Theorem|Dirac's Theorem|
{{Theorem|Lemma|
:A graph <math>G(V,E)</math> on <math>n</math> vertices has a Hamiltonian cycle if <math>d_v\ge\frac{n}{2}</math> for all <math>v\in V</math>.
:Let <math>\mathcal{F}\subseteq {X\choose k}</math> where <math>|X|=n</math> and <math>n\ge 2k</math>. If <math>\mathcal{F}</math> is intersecting, then for any cyclic permutation <math>\pi</math> of <math>X</math>, it holds that <math>|\mathcal{G}_\pi\cap\mathcal{F}|\le k</math>.
}}
}}
{{Proof|
{{Proof|
Suppose to the contrary, the theorem is not true and there exists a non-Hamiltonian graph with <math>d_v\ge\frac{n}{2}</math> for all <math>v\in V</math>. Let <math>G</math> be such a graph with a maximum number of edges. Then adding any edge to <math>G</math> creates a Hamiltonian cycle. Thus, <math>G</math> must have a Hamiltonian path, say <math>v_1v_2\cdots v_n</math>.
Fix a cyclic permutation <math>\pi</math> of <math>X</math>. Let <math>A_i=\{\pi_{(i+j+n)\bmod n}\mid j\in[k]\}</math>. Then <math>\mathcal{G}_\pi</math> can be written as <math>\mathcal{G}_\pi=\{A_i\mid i\in [n]\}</math>.  


Consider the sets,
Suppose that <math>A_t\in\mathcal{F}</math>. Since <math>\mathcal{F}</math> is intersecting, the only sets <math>A_i</math> that can be in <math>\mathcal{F}</math> other than <math>A_t</math> itself are the <math>2k-2</math> sets <math>A_i</math> with <math>t-(k-1)\le i\le t+k-1, i\neq t</math>. We partition these sets into <math>k-1</math> pairs <math>\{A_i,A_{i+k}\}</math>, where <math>t-(k-1)\le i\le t-1</math>.
*<math>S=\{i\mid v_iv_n\in E\}</math>;
*<math>T=\{i\mid v_{i+1}v_1\in E\}</math>.
Therefore, <math>S\subseteq\{v_1,v_2,\ldots,v_{n-1}\}</math> contains the neighbors of <math>v_n</math>; and <math>T\subseteq\{v_1,v_2,\ldots,v_{n-1}\}</math> contains the predecessors (along the Hamiltonian path) of the neighbors of <math>v_1</math>. It holds that <math>S,T\subseteq\{v_1,v_2,\ldots,v_{n-1}\}</math>.  


Since <math>d_v\ge\frac{n}{2}</math> for all <math>v\in V</math>, <math>|S|,|T|\ge\frac{n}{2}</math>. By the pigeonhole principle, there exists some <math>v_i\in S\cap T</math>. We can construct the following Hamiltonian cycle:
Note that for <math>n\ge 2k</math>, it holds that <math>A_i\cap C_{i+k}=\emptyset</math>. Since <math>\mathcal{F}</math> is intersecting, <math>\mathcal{F}</math> can contain at most one set of each such pair. The lemma follows.
:<math>v_1v_{i+1}v_{i+2}\cdots v_nv_{i}v_{i-1}\cdots v_1\cdots</math>,
which contradict to the assumption that <math>G</math> is non-Hamiltonian.
}}
}}


== Erdős–Stone theorem ==
The Katona's proof of Erdős–Ko–Rado theorem is done by counting in two ways the pairs of member <math>S</math> of <math>\mathcal{F}</math> and cyclic permutation <math>\pi</math> which contain <math>S</math> as a continuous path on the circle (i.e., an arc).
We introduce a notation for the number of edges in extremal graphs with a specific forbidden substructure.
 
{{Theorem|Definition|
{{Prooftitle|Katona's proof of Erdős–Ko–Rado theorem|(double counting)
:Let <math>\mathrm{ex}(n,H)</math> denote the largest number of edges that a graph <math>G\not\supseteq H</math> on <math>n</math> vertices can have.
Let
:<math>\mathcal{R}=\{(S,\pi)\mid \pi \text{ is a cyclic permutation of }X, \text{and }S\in\mathcal{F}\cap\mathcal{G}_\pi\}</math>.
We count <math>\mathcal{R}</math> in two ways.
 
First, due to the lemma, <math>|\mathcal{F}\cap\mathcal{G}_\pi|\le k</math> for any cyclic permutation <math>\pi</math>. There are <math>(n-1)!</math> cyclic permutations in total. Thus,
:<math>|\mathcal{R}|=\sum_{\text{cyclic }\pi}|\mathcal{F}\cap\mathcal{G}_\pi|\le k(n-1)!</math>.
 
Next, for each <math>S\in\mathcal{F}</math>, the number of cyclic permutations <math>\pi</math> in which <math>S</math> is continuous is <math>|S|!(n-|S|)!=k!(n-k)!</math>. Thus,
:<math>|\mathcal{R}|=\sum_{S\in\mathcal{F}}k!(n-k)!=|\mathcal{F}|k!(n-k)!</math>.
 
Altogether, we have
:<math>|\mathcal{F}|\le\frac{k(n-1)!}{k!(n-k)!}=\frac{(n-1)!}{(k-1)!(n-k)!}={n-1\choose k-1}</math>.
}}
}}
With this notation, Turán's theorem can be restated as
 
{{Theorem|Turán's theorem (restated)|
=== Erdős' shifting technique ===
:<math>\mathrm{ex}(n,K_r)\le\frac{r-2}{2(r-1)}n^2</math>.
We now introduce the original proof of the Erdős–Ko–Rado theorem, which uses a technique called '''shifting''' (originally called '''compression''').
 
Without loss of generality, we assume <math>X=[n]</math>, and restate the Erdős–Ko–Rado theorem as follows.
{{Theorem|Erdős–Ko–Rado theorem|
:Let <math>\mathcal{F}\subseteq {[n]\choose k}</math> and <math>n\ge 2k</math>. If <math>\mathcal{F}</math> is intersecting, then <math>|\mathcal{F}|\le{n-1\choose k-1}</math>.
}}
}}


Let <math>K_s^r=K_{\underbrace{s,s,\cdots,s}_{r}}</math> be the complete <math>r</math>-partite graph with <math>s</math> vertices in each class, i.e., the Turán graph <math>T(rs,r)</math>.
We define a '''shift operator''' for the set family.
The Erdős–Stone theorem (also referred as the '''fundamental theorem of extremal graph theory''') gives an asymptotic bound on <math>\mathrm{ex}(n,K-s^r)</math>, i.e., the largest number of edges that an <math>n</math>-vertex graph can have to not contain <math>K_s^r</math>.
{{Theorem|Definition (shift operator)|
: Assume <math>\mathcal{F}\subseteq 2^{[n]}</math>, and <math>0\le i<j\le n-1</math>. Define the '''<math>(i,j)</math>-shift''' <math>S_{ij}</math> as an operator on <math>\mathcal{F}</math> as follows:
:*for each <math>T\in\mathcal{F}</math>, write <math>T_{ij}=(T\setminus\{j\})\cup\{i\} </math>, and let
::<math>S_{ij}(T)=
\begin{cases}
T_{ij} & \mbox{if }j\in T, i\not\in T, \mbox{ and }T_{ij} \not\in\mathcal{F},\\
T & \mbox{otherwise;}
\end{cases}</math>
:* let <math>S_{ij}(\mathcal{F})=\{S_{ij}(T)\mid T\in \mathcal{F}\}</math>.
}}


{{Theorem|Fundamental theorem of extremal graph theory (Erdős–Stone 1946)|
It is easy to verify the following propositions of shifts.
:For any integers <math>r\ge 2</math> and <math>s\ge 1</math>, and any <math>\epsilon>0</math>, if <math>n</math> is sufficiently large then every graph on <math>n</math> vertices and with at least <math>\left(\frac{r-2}{2(r-1)}+\epsilon\right)n^2</math> edges contains <math>K_{r,s}</math> as a subgraph, i.e.,
 
:::<math>\mathrm{ex}(n,K_s^r)= \left(\frac{r-2}{2(r-1)}+o(1)\right)n^2</math>.
{{Theorem|Proposition|
# <math>|S_{ij}(T)|=|T|\,</math> and <math>|S_{ij}(\mathcal{F})|=\mathcal{F}</math>;
# if <math>\mathcal{F}</math> is intersecting, then so is <math>S_{ij}(\mathcal{F})</math>.
}}
{{Proof|
(1) is immediate. Now we prove (2).
 
Let <math>A,B\in\mathcal{F}</math>. All the cases are easy to dealt with except when <math>A\cap B=\{j\}</math>, <math>i\in A</math>, and <math>i\not\in B</math>. Denote <math>A_{ij}=A\setminus\{j\}\cup\{i\}</math>. It holds that <math>A_{ij}\cap B=(A\cap B)\setminus\{j\}=\emptyset</math>. Since <math>\mathcal{F}</math> is intersecting, it must hold that <math>A_{ij}\not\in\mathcal{F}</math>. Thus, <math>S_{ij}(A)=A_{ij}</math> and clearly <math>S_{ij}(B)=B_{ij}=B\setminus\{j\}\cup\{i\}</math>. Therefore, <math>i\in S_{ij}(A)\cap S_{ij}(B)</math> thus <math>S_{ij}(A)\cap S_{ij}(B)\neq \emptyset</math>.
}}
}}


The theorem is called fundamental because of its single most important corollary: it relate the extremal bound for an arbitrary subgraph <math>H</math> to a very natural parameter of <math>H</math>, its chromatic number.
Repeatedly applying <math>S_{ij}(\mathcal{F})</math> for any <math>0\le i<j\le n-1</math>, since we only replace elements by smaller elements, eventually <math>\mathcal{F}</math> will stop changing, that is, <math>S_{ij}(\mathcal{F})=\mathcal{F}</math> for all <math>0\le i<j\le n-1</math>. We call such an <math>\mathcal{F}</math> '''shifted'''.
 
The idea behind the shifting technique is very natural: by applying shifting, all intersecting families are transformed to some ''special forms'', and we only need to prove the theorem for these special form of intersecting families.
 
{{Prooftitle|Proof of Erdős-Ko-Rado theorem| (The original proof of Erdős-Ko-Rado by shifting)
By the above lemma, it is sufficient to prove the Erdős-Ko-Rado theorem holds for shifted <math>\mathcal{F}</math>. We assume that <math>\mathcal{F}</math> is shifted.
 
First, it is trivial to see that the theorem holds for <math>k=1</math> (no matter whether shifted).


Recall that <math>\chi(G)</math> is the '''chromatic number''' of <math>G</math>, the smallest number of colors that one can use to color the vertices so that no adjacent vertices have the same color.
Next, we show that the theorem holds when <math>n=2k</math>  (no matter whether shifted). For any <math>S\in{X\choose k}</math>, both <math>S</math> and <math>X\setminus S</math> are in <math>{X\choose k}</math>, but at most one of them can be in <math>\mathcal{F}</math>. Thus,
:<math>|\mathcal{F}|\le\frac{1}{2}{n\choose k}=\frac{n!}{2k!(n-k)!}=\frac{(n-1)!}{(k-1)!(n-k)!}={n-1\choose k-1}</math>.


{{Theorem|Corollary|
We then apply the induction on <math>n</math>. For <math>n> 2k</math>, the induction hypothesis is stated as:
:For every nonempty graph <math>H</math>,
* the Erdős-Ko-Rado theorem holds for any smaller <math>n</math>.
::<math>\lim_{n\rightarrow\infty}\frac{\mathrm{ex}(n,H)}{{n\choose 2}}=\frac{\chi(H)-2}{\chi(H)-1}</math>.
Define
}}
:<math>\mathcal{F}_0=\{S\in\mathcal{F}\mid n\not\in S\}</math> and <math>\mathcal{F}_1=\{S\in\mathcal{F}\mid n\in S\}</math>.
{{Prooftitle|Proof of corollary|
Clearly, <math>\mathcal{F}_0\subseteq{[n-1]\choose k}</math> and <math>\mathcal{F}_0</math> is intersecting. Due to the induction hypothesis, <math>|\mathcal{F}_0|\le{n-2\choose k-1}</math>.  
Let <math>r=\chi(H)</math>.  
 
In order to apply the induction, we let
:<math>\mathcal{F}_1'=\{S\setminus\{n\}\mid S\in\mathcal{F}_1\}</math>.
Clearly, <math>\mathcal{F}_1'\subseteq{[n-1]\choose k-1}</math>. If only it is also intersecting, we can apply the induction hypothesis, and indeed it is. To see this, by contradiction we assume that <math>\mathcal{F}_1'</math> is not intersecting. Then there must exist <math>A,B\in\mathcal{F}</math> such that <math>A\cap B=\{n\}</math>, which means that <math>|A\cup B|\le 2k-1<n-1</math>. Thus, there is some <math>0\le i\le n-1</math> such that <math>i\not\in A\cup B</math>. Since <math>\mathcal{F}</math> is shifted, <math>A_{in}=A\setminus\{n\}\cup\{i\}\in\mathcal{F}</math>. On the other hand it can be verified that <math>A_{in}\cap B=\emptyset</math>, which contradicts that <math>\mathcal{F}</math> is intersecting.  


Note that <math>T(n,r-1)</math> can be colored with <math>r-1</math> colors, one color for each part. Thus, <math>H\not\subseteq T(n,r-1)</math>, since otherwise <math>H</math> can also be colored with <math>r-1</math> colors, contradicting that <math>\chi(H)=1</math>. By definition, <math>\mathrm{ex}(n,H)</math> is the maximum number of edges that an <math>n</math>-vertex graph <math>G\not\supseteq H</math> can have. Thus,
Thus, <math>\mathcal{F}_1'\subseteq{[n-1]\choose k-1}</math> and <math>\mathcal{F}_1'</math> is intersecting. Due to the induction hypothesis, <math>|\mathcal{F}_1'|\le{n-2\choose k-2}</math>.  
:<math>|T(n,r-1)|\le\mathrm{ex}(n,H)</math>.
It is not hard to see that
:<math>|T(n,r-1)|\ge {r-1\choose 2}\left\lfloor\frac{n}{r-1}\right\rfloor^2\ge{r-1\choose 2}\left(\frac{n}{r-1}-1\right)^2=\left(\frac{r-2}{2(r-1)}-o(1)\right)n^2</math>.


On the other hand, any finite graph <math>H</math> with chromatic number <math>r</math> has that <math>H\subseteq K_s^r</math> for all sufficiently large <math>s</math>. We just connect all pairs of vertices from different color classes. Thus,  
Combining these together,
:<math>\mathrm{ex}(n,H)\le\mathrm{ex}(n,K_s^r)</math>.
:<math>|\mathcal{F}|=|\mathcal{F}_0|+|\mathcal{F}_1|=|\mathcal{F}_0|+|\mathcal{F}_1'|\le {n-2\choose k-1}+{n-2\choose k-2}={n-1\choose k-1}</math>.
Due to Erdős–Stone theorem,
:<math>\mathrm{ex}(n,K_s^r)=\left(\frac{r-2}{2(r-1)}+o(1)\right)n^2</math>.
Altogether, we have
:<math>
\frac{r-2}{r-1}-o(1)\le\frac{|T(n,r-1)|}{{n\choose 2}}\le \frac{\mathrm{ex}(n,H)}{{n\choose 2}} \le \frac{\mathrm{ex}(n,K_s^r)}{{n\choose 2}}=\frac{r-2}{r-1}+o(1)
</math>
The theorem follows.
}}
}}


Line 239: Line 310:
:('''Disclaimer:''' The following copyrighted materials are meant for educational uses only.)
:('''Disclaimer:''' The following copyrighted materials are meant for educational uses only.)


* van Lin and Wilson. ''A course in combinatorics.'' Cambridge Press. Chapter 4.
* van Lin and Wilson. ''A course in combinatorics.'' Cambridge Press. Chapter 6.
* Aigner and Ziegler. ''Proofs from THE BOOK, 4th Edition.'' Springer-Verlag. [[media:PFTB_chap36.pdf| Chapter 36]].
* Aigner and Ziegler. ''Proofs from THE BOOK, 4th Edition.'' Springer-Verlag. [[media:PFTB_chap27.pdf| Chapter 27]].
* Diestel. ''Graph Theory, 3rd Edition''. Springer-Verlag 2000. [[media:Diestel2ed_chap7.pdf|Chapter 7]].

Revision as of 01:19, 10 November 2011

Sperner system

A set family [math]\displaystyle{ \mathcal{F}\subseteq 2^X }[/math] with the relation [math]\displaystyle{ \subseteq }[/math] define a poset. Thus, a chain is a sequence [math]\displaystyle{ S_1\subseteq S_2\subseteq\cdots\subseteq S_k }[/math].

A set family [math]\displaystyle{ \mathcal{F}\subseteq 2^X }[/math] is an antichain (also called a Sperner system) if for all [math]\displaystyle{ S,T\in\mathcal{F} }[/math] that [math]\displaystyle{ S\neq T }[/math], we have [math]\displaystyle{ S\not\subseteq T }[/math].

The [math]\displaystyle{ k }[/math]-uniform [math]\displaystyle{ {X\choose k} }[/math] is an antichain. Let [math]\displaystyle{ n=|X| }[/math]. The size of [math]\displaystyle{ {X\choose k} }[/math] is maximized when [math]\displaystyle{ k=\lfloor n/2\rfloor }[/math]. We wonder whether this is also the largest possible size of any antichain [math]\displaystyle{ \mathcal{F}\subseteq 2^X }[/math].

In 1928, Emanuel Sperner proved a theorem saying that it is indeed the largest possible antichain. This result, called Sperner's theorem today, initiated the studies of extremal set theory.

Theorem (Sperner 1928)
Let [math]\displaystyle{ \mathcal{F}\subseteq 2^X }[/math] where [math]\displaystyle{ |X|=n }[/math]. If [math]\displaystyle{ \mathcal{F} }[/math] is an antichain, then
[math]\displaystyle{ |\mathcal{F}|\le{n\choose \lfloor n/2\rfloor} }[/math].

First proof (shadows)

We first introduce the original proof by Sperner, which uses concepts called shadows and shades of set systems.

Definition
Let [math]\displaystyle{ |X|=n\, }[/math] and [math]\displaystyle{ \mathcal{F}\subseteq {X\choose k} }[/math], [math]\displaystyle{ k\lt n\, }[/math].
The shade of [math]\displaystyle{ \mathcal{F} }[/math] is defined to be
[math]\displaystyle{ \nabla\mathcal{F}=\left\{T\in {X\choose k+1}\,\,\bigg|\,\, \exists S\in\mathcal{F}\mbox{ such that } S\subset T\right\} }[/math].
Thus the shade [math]\displaystyle{ \nabla\mathcal{F} }[/math] of [math]\displaystyle{ \mathcal{F} }[/math] consists of all subsets of [math]\displaystyle{ X }[/math] which can be obtained by adding an element to a set in [math]\displaystyle{ \mathcal{F} }[/math].
Similarly, the shadow of [math]\displaystyle{ \mathcal{F} }[/math] is defined to be
[math]\displaystyle{ \Delta\mathcal{F}=\left\{T\in {X\choose k-1}\,\,\bigg|\,\, \exists S\in\mathcal{F}\mbox{ such that } T\subset S\right\} }[/math].
Thus the shadow [math]\displaystyle{ \Delta\mathcal{F} }[/math] of [math]\displaystyle{ \mathcal{F} }[/math] consists of all subsets of [math]\displaystyle{ X }[/math] which can be obtained by removing an element from a set in [math]\displaystyle{ \mathcal{F} }[/math].

Next lemma bounds the effects of shadows and shades on the sizes of set systems.

Lemma (Sperner)
Let [math]\displaystyle{ |X|=n\, }[/math] and [math]\displaystyle{ \mathcal{F}\subseteq {X\choose k} }[/math]. Then
[math]\displaystyle{ \begin{align} &|\nabla\mathcal{F}|\ge\frac{n-k}{k+1}|\mathcal{F}| &\text{ if } k\lt n\\ &|\Delta\mathcal{F}|\ge\frac{k}{n-k+1}|\mathcal{F}| &\text{ if } k\gt 0. \end{align} }[/math]
Proof.

The lemma is proved by double counting. We prove the inequality of [math]\displaystyle{ |\nabla\mathcal{F}| }[/math]. Assume that [math]\displaystyle{ 0\le k\lt n }[/math].

Define

[math]\displaystyle{ \mathcal{R}=\{(S,T)\mid S\in\mathcal{F}, T\in\nabla\mathcal{F}, S\subset T\} }[/math].

We estimate [math]\displaystyle{ |\mathcal{R}| }[/math] in two ways.

For each [math]\displaystyle{ S\in\mathcal{F} }[/math], there are [math]\displaystyle{ n-k }[/math] different [math]\displaystyle{ T\in\nabla\mathcal{F} }[/math] that [math]\displaystyle{ S\subset T }[/math].

[math]\displaystyle{ |\mathcal{R}|=(n-k)|\mathcal{F}| }[/math].

For each [math]\displaystyle{ T\in\nabla\mathcal{F} }[/math], there are [math]\displaystyle{ k+1 }[/math] ways to choose an [math]\displaystyle{ S\subset T }[/math] with [math]\displaystyle{ |S|=k }[/math], some of which may not be in [math]\displaystyle{ \mathcal{F} }[/math].

[math]\displaystyle{ |\mathcal{R}|\le (k+1)|\nabla\mathcal{F}| }[/math].

Altogether, we show that [math]\displaystyle{ |\nabla\mathcal{F}|\ge\frac{n-k}{k+1}|\mathcal{F}| }[/math].

The inequality of [math]\displaystyle{ |\Delta\mathcal{F}| }[/math] can be proved in the same way.

[math]\displaystyle{ \square }[/math]

An immediate corollary of the previous lemma is as follows.

Proposition 1
If [math]\displaystyle{ k\le \frac{n-1}{2} }[/math], then [math]\displaystyle{ |\nabla\mathcal{F}|\ge|\mathcal{F}| }[/math].
If [math]\displaystyle{ k\ge \frac{n-1}{2} }[/math], then [math]\displaystyle{ |\Delta\mathcal{F}|\ge|\mathcal{F}| }[/math].

The idea of Sperner's proof is pretty clear:

  • we "push up" all the sets in [math]\displaystyle{ \mathcal{F} }[/math] of size [math]\displaystyle{ \lt \frac{n-1}{2} }[/math] replacing them by their shades;
  • and also "push down" all the sets in [math]\displaystyle{ \mathcal{F} }[/math] of size [math]\displaystyle{ \ge\frac{n+1}{2} }[/math] replacing them by their shadows.

Repeat this process we end up with a set system [math]\displaystyle{ \mathcal{F}\subseteq{X\choose \lfloor n/2\rfloor} }[/math]. We need to show that this process does not decrease the size of [math]\displaystyle{ \mathcal{F} }[/math].

Proposition 2
Suppose that [math]\displaystyle{ \mathcal{F}\subseteq2^X }[/math] where [math]\displaystyle{ |X|=n }[/math]. Let [math]\displaystyle{ \mathcal{F}_k=\mathcal{F}\ap{X\choose k} }[/math]. Let [math]\displaystyle{ k_\min }[/math] be the smallest [math]\displaystyle{ k }[/math] that [math]\displaystyle{ |\mathcal{F}_k|\gt 0 }[/math], and let
[math]\displaystyle{ \mathcal{F}'=\begin{cases} \mathcal{F}\setminus\mathcal{F}_{k_\min}\cup \nabla\mathcal{F}_{k_\min} & \mbox{if }k_\min\lt \frac{n-1}{2},\\ \mathcal{F} & \mbox{otherwise.} \end{cases} }[/math]
Similarly, let [math]\displaystyle{ k_\max }[/math] be the largest [math]\displaystyle{ k }[/math] that [math]\displaystyle{ |\mathcal{F}_k|\gt 0 }[/math], and let
[math]\displaystyle{ \mathcal{F}''=\begin{cases} \mathcal{F}\setminus\mathcal{F}_{k_\max}\cup \Delta\mathcal{F}_{k_\max} & \mbox{if }k_\max\ge\frac{n+1}{2},\\ \mathcal{F} & \mbox{otherwise.} \end{cases} }[/math]
If [math]\displaystyle{ \mathcal{F} }[/math] is an antichain, [math]\displaystyle{ \mathcal{F}' }[/math] and [math]\displaystyle{ \mathcal{F}'' }[/math] are antichains, and we have [math]\displaystyle{ |\mathcal{F}'|\ge|\mathcal{F}| }[/math] and [math]\displaystyle{ |\mathcal{F}''|\ge|\mathcal{F}| }[/math].
Proof.

We show that [math]\displaystyle{ \mathcal{F}' }[/math] is an antichain and [math]\displaystyle{ |\mathcal{F}'|\ge|\mathcal{F}| }[/math].

First, observe that [math]\displaystyle{ \nabla\mathcal{F}_k\cap\mathcal{F}=\emptyset }[/math], otherwise [math]\displaystyle{ \mathcal{F} }[/math] cannot be an antichain, and due to Proposition 1, [math]\displaystyle{ |\nabla\mathcal{F}_k|\ge|\mathcal{F}_k| }[/math] when [math]\displaystyle{ k\le \frac{n-1}{2} }[/math], so [math]\displaystyle{ |\mathcal{F}'|=|\mathcal{F}|-|\mathcal{F}_k|+|\nabla\mathcal{F}_k|\ge |\mathcal{F}| }[/math].

Now we prove that [math]\displaystyle{ \mathcal{F}' }[/math] is an antichain . By contradiction, assume that there are [math]\displaystyle{ S, T\in \mathcal{F}' }[/math], such that [math]\displaystyle{ S\subset T }[/math]. One of the [math]\displaystyle{ S,T }[/math] must be in [math]\displaystyle{ \nabla\mathcal{F}_{k_\min} }[/math], or otherwise [math]\displaystyle{ \mathcal{F} }[/math] cannot be an antichain. Recall that [math]\displaystyle{ k_\min }[/math] is the smallest [math]\displaystyle{ k }[/math] that [math]\displaystyle{ |\mathcal{F}_k|\gt 0 }[/math], thus it must be [math]\displaystyle{ S\in \nabla\mathcal{F}_{k_\min} }[/math], and [math]\displaystyle{ T\in\mathcal{F} }[/math]. This implies that there is an [math]\displaystyle{ R\in \mathcal{F}_{k_\min}\subseteq \mathcal{F} }[/math] such that [math]\displaystyle{ R\subset S\subset T }[/math], which contradicts that [math]\displaystyle{ \mathcal{F} }[/math] is an antichain.

The statement for [math]\displaystyle{ \mathcal{F}'' }[/math] can be proved in the same way.

[math]\displaystyle{ \square }[/math]

Applying the above process, we prove the Sperner's theorem.

Proof of Sperner's theorem
(original proof of Sperner)

Let [math]\displaystyle{ \mathcal{F}_k=\{S\in\mathcal{F}\mid |S|=k\} }[/math], where [math]\displaystyle{ 0\le k\le n }[/math].

We change [math]\displaystyle{ \mathcal{F} }[/math] as follows:

  • for the smallest [math]\displaystyle{ k }[/math] that [math]\displaystyle{ |\mathcal{F}_k|\gt 0 }[/math], if [math]\displaystyle{ k\lt \frac{n-1}{2} }[/math], replace [math]\displaystyle{ \mathcal{F}_k }[/math] by [math]\displaystyle{ \nabla\mathcal{F}_k }[/math].

Due to Proposition 2, this procedure preserves [math]\displaystyle{ \mathcal{F} }[/math] as an antichain and does not decrease [math]\displaystyle{ |\mathcal{F}| }[/math]. Repeat this procedure, until [math]\displaystyle{ |\mathcal{F}_k|=0 }[/math] for all [math]\displaystyle{ k\lt \frac{n-1}{2} }[/math], that is, there is no member set of [math]\displaystyle{ \mathcal{F} }[/math] has size less than [math]\displaystyle{ \frac{n-1}{2} }[/math].

We then define another symmetric procedure:

  • for the largest [math]\displaystyle{ k }[/math] that [math]\displaystyle{ |\mathcal{F}_k|\gt 0 }[/math], if [math]\displaystyle{ k\ge\frac{n+1}{2} }[/math], replace [math]\displaystyle{ \mathcal{F}_k }[/math] by [math]\displaystyle{ \Delta\mathcal{F}_k }[/math].

Also due to Proposition 2, this procedure preserves [math]\displaystyle{ \mathcal{F} }[/math] as an antichain and does not decrease [math]\displaystyle{ |\mathcal{F}| }[/math]. After repeatedly applying this procedure, [math]\displaystyle{ |\mathcal{F}_k|=0 }[/math] for all [math]\displaystyle{ k\ge\frac{n+1}{2} }[/math].

The resulting [math]\displaystyle{ \mathcal{F} }[/math] has [math]\displaystyle{ \mathcal{F}\subseteq{X\choose \lfloor n/2\rfloor} }[/math], and since [math]\displaystyle{ |\mathcal{F}| }[/math] is never decreased, for the original [math]\displaystyle{ \mathcal{F} }[/math], we have

[math]\displaystyle{ |\mathcal{F}|\le {n\choose \lfloor n/2\rfloor} }[/math].
[math]\displaystyle{ \square }[/math]

Second proof (counting)

We now introduce an elegant proof due to Lubell. The proof uses a counting argument, and tells more information than just the size of the set system.

Proof of Sperner's theorem
(Lubell 1966)

Let [math]\displaystyle{ \pi }[/math] be a permutation of [math]\displaystyle{ X }[/math]. We say that an [math]\displaystyle{ S\subseteq X }[/math] prefixes [math]\displaystyle{ \pi }[/math], if [math]\displaystyle{ S=\{\pi_1,\pi_2,\ldots, \pi_{|S|}\} }[/math], that is, [math]\displaystyle{ S }[/math] is precisely the set of the first [math]\displaystyle{ |S| }[/math] elements in the permutation [math]\displaystyle{ \pi }[/math].

Fix an [math]\displaystyle{ S\subseteq X }[/math]. It is easy to see that the number of permutations [math]\displaystyle{ \pi }[/math] of [math]\displaystyle{ X }[/math] prefixed by [math]\displaystyle{ S }[/math] is [math]\displaystyle{ |S|!(n-|S|)! }[/math]. Also, since [math]\displaystyle{ \mathcal{F} }[/math] is an antichain, no permutation [math]\displaystyle{ \pi }[/math] of [math]\displaystyle{ X }[/math] can be prefixed by more than one members of [math]\displaystyle{ \mathcal{F} }[/math], otherwise one of the member sets must contain the other, which contradicts that [math]\displaystyle{ \mathcal{F} }[/math] is an antichain. Thus, the number of permutations [math]\displaystyle{ \pi }[/math] prefixed by some [math]\displaystyle{ S\in\mathcal{F} }[/math] is

[math]\displaystyle{ \sum_{S\in\mathcal{F}}|S|!(n-|S|)! }[/math],

which cannot be larger than the total number of permutations, [math]\displaystyle{ n! }[/math], therefore,

[math]\displaystyle{ \sum_{S\in\mathcal{F}}|S|!(n-|S|)!\le n! }[/math].

Dividing both sides by [math]\displaystyle{ n! }[/math], we have

[math]\displaystyle{ \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}=\sum_{S\in\mathcal{F}}\frac{|S|!(n-|S|)!}{n!}\le 1 }[/math],

where [math]\displaystyle{ {n\choose |S|}\le {n\choose \lfloor n/2\rfloor} }[/math], so

[math]\displaystyle{ \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\ge \frac{|\mathcal{F}|}{{n\choose \lfloor n/2\rfloor}} }[/math].

Combining this with the above inequality, we prove the Sperner's theorem.

[math]\displaystyle{ \square }[/math]

The LYM inequality

Lubell's proof proves the following inequality:

[math]\displaystyle{ \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le 1 }[/math]

which is actually stronger than Sperner's original statement that [math]\displaystyle{ |\mathcal{F}|\le{n\choose \lfloor n/2\rfloor} }[/math].

This inequality is independently discovered by Lubell-Yamamoto, Meschalkin, and Bollobás, and is called the LYM inequality today.

Theorem (Lubell, Yamamoto 1954; Meschalkin 1963)
Let [math]\displaystyle{ \mathcal{F}\subseteq 2^X }[/math] where [math]\displaystyle{ |X|=n }[/math]. If [math]\displaystyle{ \mathcal{F} }[/math] is an antichain, then
[math]\displaystyle{ \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le 1 }[/math].

In Lubell's counting argument proves the LYM inequality, which implies the Sperner's theorem. Here we give another proof of the LYM inequality by the probabilistic method, due to Noga Alon.

Third proof (the probabilistic method)
(Due to Alon.)

Let [math]\displaystyle{ \pi }[/math] be a uniformly random permutation of [math]\displaystyle{ X }[/math]. Define a random maximal chain by

[math]\displaystyle{ \mathcal{C}_\pi=\{\{\pi_i\mid 1\le i\le k\}\mid 0\le k\le n\} }[/math].

For any [math]\displaystyle{ S\in\mathcal{F} }[/math], let [math]\displaystyle{ X_S }[/math] be the 0-1 random variable which indicates whether [math]\displaystyle{ S\in\mathcal{C}_\pi }[/math], that is

[math]\displaystyle{ X_S=\begin{cases} 1 & \mbox{if }S\in\mathcal{C}_\pi,\\ 0 & \mbox{otherwise.} \end{cases} }[/math]

Note that for a uniformly random [math]\displaystyle{ \pi }[/math], [math]\displaystyle{ \mathcal{C}_\pi }[/math] has exact one member set of size [math]\displaystyle{ |S| }[/math], uniformly distributed over [math]\displaystyle{ {X\choose |S|} }[/math], thus

[math]\displaystyle{ \mathbf{E}[X_S]=\Pr[S\in\mathcal{C}_\pi]=\frac{1}{{n\choose |S|}} }[/math].

Let [math]\displaystyle{ X=\sum_{S\in\mathcal{F}}X_S }[/math]. Note that [math]\displaystyle{ X=|\mathcal{F}\cap\mathcal{C}_\pi| }[/math]. By the linearity of expectation,

[math]\displaystyle{ \mathbf{E}[X]=\sum_{S\in\mathcal{F}}\mathbf{E}[X_S]=\sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}} }[/math].

On the other hand, since [math]\displaystyle{ \mathcal{F} }[/math] is an antichain, it can never intersect a chain at more than one elements, thus we always have [math]\displaystyle{ X=|\mathcal{F}\cap\mathcal{C}_\pi|\le 1 }[/math]. Therefore,

[math]\displaystyle{ \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le \mathbf{E}[X] \le 1 }[/math].
[math]\displaystyle{ \square }[/math]

The Sperner's theorem is an immediate consequence of the LYM inequality.

Proposition
[math]\displaystyle{ \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\le 1 }[/math] implies that [math]\displaystyle{ |\mathcal{F}|\le{n\choose \lfloor n/2\rfloor} }[/math].
Proof.

It holds that [math]\displaystyle{ {n\choose k}\le {n\choose \lfloor n/2\rfloor} }[/math] for any [math]\displaystyle{ k }[/math]. Thus,

[math]\displaystyle{ 1\ge \sum_{S\in\mathcal{F}}\frac{1}{{n\choose |S|}}\ge \frac{|\mathcal{F}|}{{n\choose \lfloor n/2\rfloor}} }[/math],

which implies that [math]\displaystyle{ |\mathcal{F}|\le {n\choose \lfloor n/2\rfloor} }[/math].

[math]\displaystyle{ \square }[/math]

Sunflowers

An set system is a sunflower if all its member sets intersect at the same set of elements.

Definition (sunflower)
A set family [math]\displaystyle{ \mathcal{F}\subseteq 2^X }[/math] is a sunflower of size [math]\displaystyle{ r }[/math] with a core [math]\displaystyle{ C\subseteq X }[/math] if
[math]\displaystyle{ \forall S,T\in\mathcal{F} }[/math] that [math]\displaystyle{ S\neq T }[/math], [math]\displaystyle{ S\cap T=C }[/math].

Note that we do not require the core to be nonempty, thus a family of disjoint sets is also a sunflower (with the core [math]\displaystyle{ \emptyset }[/math]).

The next result due to Erdős and Rado, called the sunflower lemma, is a famous result in extremal set theory, and has some important applications in Boolean circuit complexity.

Sunflower Lemma (Erdős-Rado)
Let [math]\displaystyle{ \mathcal{F}\subseteq {X\choose k} }[/math]. If [math]\displaystyle{ |\mathcal{F}|\gt k!(r-1)^k }[/math], then [math]\displaystyle{ \mathcal{F} }[/math] contains a sunflower of size [math]\displaystyle{ r }[/math].
Proof.

We proceed by induction on [math]\displaystyle{ k }[/math]. For [math]\displaystyle{ k=1 }[/math], [math]\displaystyle{ \mathcal{F}\subseteq{X\choose 1} }[/math], thus all sets in [math]\displaystyle{ \mathcal{F} }[/math] are disjoint. And since [math]\displaystyle{ |\mathcal{F}|\gt r-1 }[/math], we can choose [math]\displaystyle{ r }[/math] of these sets and form a sunflower.

Now let [math]\displaystyle{ k\ge 2 }[/math] and assume the lemma holds for all smaller [math]\displaystyle{ k }[/math]. Take a maximal family [math]\displaystyle{ \mathcal{G}\subseteq \mathcal{F} }[/math] whose members are disjoint, i.e. for any [math]\displaystyle{ S,T\in \mathcal{G} }[/math] that [math]\displaystyle{ S\neq T }[/math], [math]\displaystyle{ S\cap T=\emptyset }[/math].

If [math]\displaystyle{ |\mathcal{G}|\ge r }[/math], then [math]\displaystyle{ \mathcal{G} }[/math] is a sunflower of size at least [math]\displaystyle{ r }[/math] and we are done.

Assume that [math]\displaystyle{ |\mathcal{G}|\le r-1 }[/math], and let [math]\displaystyle{ Y=\bigcup_{S\in\mathcal{G}}S }[/math]. Then [math]\displaystyle{ |Y|=k|\mathcal{G}|\le k(r-1) }[/math] (since all members of [math]\displaystyle{ \mathcal{G} }[/math]) are disjoint). We claim that [math]\displaystyle{ Y }[/math] intersets all members of [math]\displaystyle{ \mathcal{F} }[/math], since if otherwise, there exists an [math]\displaystyle{ S\in\mathcal{F} }[/math] such that [math]\displaystyle{ S\cap Y=\emptyset }[/math], then we can enlarge [math]\displaystyle{ \mathcal{G} }[/math] by adding [math]\displaystyle{ S }[/math] into [math]\displaystyle{ \mathcal{G} }[/math] and still have all members of [math]\displaystyle{ \mathcal{G} }[/math] disjoint, which contradicts the assumption that [math]\displaystyle{ \mathcal{G} }[/math] is the maximum of such families.

By the pigeonhole principle, some elements [math]\displaystyle{ y\in Y }[/math] must contained in at least

[math]\displaystyle{ \frac{|\mathcal{F}|}{|Y|}\gt \frac{k!(r-1)^k}{k(r-1)}=(k-1)!(r-1)^{k-1} }[/math]

members of [math]\displaystyle{ \mathcal{F} }[/math]. We delete this [math]\displaystyle{ y }[/math] from these sets and consider the family

[math]\displaystyle{ \mathcal{H}=\{S\setminus\{y\}\mid S\in\mathcal{F}\wedge y\in S\} }[/math].

We have [math]\displaystyle{ \mathcal{H}\subseteq {X\choose k-1} }[/math] and [math]\displaystyle{ |\mathcal{H}|\gt (k-1)!(r-1)^{k-1} }[/math], thus by the induction hypothesis, [math]\displaystyle{ \mathcal{H} }[/math]contains a sunflower of size [math]\displaystyle{ r }[/math]. Adding [math]\displaystyle{ y }[/math] to the members of this sunflower, we get the desired sunflower in the original family [math]\displaystyle{ \mathcal{F} }[/math].

[math]\displaystyle{ \square }[/math]

The Erdős–Ko–Rado Theorem

A set family [math]\displaystyle{ \mathcal{F}\subseteq 2^X }[/math] is called intersecting, if for any [math]\displaystyle{ S,T\in\mathcal{F} }[/math], [math]\displaystyle{ S\cap T\neq\emptyset }[/math]. A natural question of extremal favor is: "how large can an intersecting family be?"

Assume [math]\displaystyle{ |X|=n }[/math]. When [math]\displaystyle{ n\lt 2k }[/math], every pair of [math]\displaystyle{ k }[/math]-subsets of [math]\displaystyle{ X }[/math] intersects. So the non-trivial case is when [math]\displaystyle{ n\le 2k }[/math]. The famous Erdős–Ko–Rado theorem gives the largest possible cardinality of a nontrivially intersecting family.

According to Erdős, the theorem itself was proved in 1938, but was not published until 23 years later.

Erdős–Ko–Rado theorem (proved in 1938, published in 1961)
Let [math]\displaystyle{ \mathcal{F}\subseteq {X\choose k} }[/math] where [math]\displaystyle{ |X|=n }[/math] and [math]\displaystyle{ n\ge 2k }[/math]. If [math]\displaystyle{ \mathcal{F} }[/math] is intersecting, then
[math]\displaystyle{ |\mathcal{F}|\le{n-1\choose k-1} }[/math].

Katona's proof

We first introduce a proof discovered by Katona in 1972. The proof uses double counting.

Let [math]\displaystyle{ \pi }[/math] be a cyclic permutation of [math]\displaystyle{ X }[/math], that is, we think of assigning [math]\displaystyle{ X }[/math] in a circle and ignore the rotations of the circle. It is easy to see that there are [math]\displaystyle{ (n-1)! }[/math] cyclic permutations of an [math]\displaystyle{ n }[/math]-set (each cyclic permutation corresponds to [math]\displaystyle{ n }[/math] permutations). Let

[math]\displaystyle{ \mathcal{G}_\pi=\{\{\pi_{(i+j)\bmod n}\mid j\in[k]\}\mid i\in [n]\} }[/math].

The next lemma states the following observation: in a circle of [math]\displaystyle{ n }[/math] points, supposed [math]\displaystyle{ n\ge 2k }[/math], there can be at most [math]\displaystyle{ k }[/math] arcs, each consisting of [math]\displaystyle{ k }[/math] points, such that every pair of arcs share at least one point.

Lemma
Let [math]\displaystyle{ \mathcal{F}\subseteq {X\choose k} }[/math] where [math]\displaystyle{ |X|=n }[/math] and [math]\displaystyle{ n\ge 2k }[/math]. If [math]\displaystyle{ \mathcal{F} }[/math] is intersecting, then for any cyclic permutation [math]\displaystyle{ \pi }[/math] of [math]\displaystyle{ X }[/math], it holds that [math]\displaystyle{ |\mathcal{G}_\pi\cap\mathcal{F}|\le k }[/math].
Proof.

Fix a cyclic permutation [math]\displaystyle{ \pi }[/math] of [math]\displaystyle{ X }[/math]. Let [math]\displaystyle{ A_i=\{\pi_{(i+j+n)\bmod n}\mid j\in[k]\} }[/math]. Then [math]\displaystyle{ \mathcal{G}_\pi }[/math] can be written as [math]\displaystyle{ \mathcal{G}_\pi=\{A_i\mid i\in [n]\} }[/math].

Suppose that [math]\displaystyle{ A_t\in\mathcal{F} }[/math]. Since [math]\displaystyle{ \mathcal{F} }[/math] is intersecting, the only sets [math]\displaystyle{ A_i }[/math] that can be in [math]\displaystyle{ \mathcal{F} }[/math] other than [math]\displaystyle{ A_t }[/math] itself are the [math]\displaystyle{ 2k-2 }[/math] sets [math]\displaystyle{ A_i }[/math] with [math]\displaystyle{ t-(k-1)\le i\le t+k-1, i\neq t }[/math]. We partition these sets into [math]\displaystyle{ k-1 }[/math] pairs [math]\displaystyle{ \{A_i,A_{i+k}\} }[/math], where [math]\displaystyle{ t-(k-1)\le i\le t-1 }[/math].

Note that for [math]\displaystyle{ n\ge 2k }[/math], it holds that [math]\displaystyle{ A_i\cap C_{i+k}=\emptyset }[/math]. Since [math]\displaystyle{ \mathcal{F} }[/math] is intersecting, [math]\displaystyle{ \mathcal{F} }[/math] can contain at most one set of each such pair. The lemma follows.

[math]\displaystyle{ \square }[/math]

The Katona's proof of Erdős–Ko–Rado theorem is done by counting in two ways the pairs of member [math]\displaystyle{ S }[/math] of [math]\displaystyle{ \mathcal{F} }[/math] and cyclic permutation [math]\displaystyle{ \pi }[/math] which contain [math]\displaystyle{ S }[/math] as a continuous path on the circle (i.e., an arc).

Katona's proof of Erdős–Ko–Rado theorem
(double counting)

Let

[math]\displaystyle{ \mathcal{R}=\{(S,\pi)\mid \pi \text{ is a cyclic permutation of }X, \text{and }S\in\mathcal{F}\cap\mathcal{G}_\pi\} }[/math].

We count [math]\displaystyle{ \mathcal{R} }[/math] in two ways.

First, due to the lemma, [math]\displaystyle{ |\mathcal{F}\cap\mathcal{G}_\pi|\le k }[/math] for any cyclic permutation [math]\displaystyle{ \pi }[/math]. There are [math]\displaystyle{ (n-1)! }[/math] cyclic permutations in total. Thus,

[math]\displaystyle{ |\mathcal{R}|=\sum_{\text{cyclic }\pi}|\mathcal{F}\cap\mathcal{G}_\pi|\le k(n-1)! }[/math].

Next, for each [math]\displaystyle{ S\in\mathcal{F} }[/math], the number of cyclic permutations [math]\displaystyle{ \pi }[/math] in which [math]\displaystyle{ S }[/math] is continuous is [math]\displaystyle{ |S|!(n-|S|)!=k!(n-k)! }[/math]. Thus,

[math]\displaystyle{ |\mathcal{R}|=\sum_{S\in\mathcal{F}}k!(n-k)!=|\mathcal{F}|k!(n-k)! }[/math].

Altogether, we have

[math]\displaystyle{ |\mathcal{F}|\le\frac{k(n-1)!}{k!(n-k)!}=\frac{(n-1)!}{(k-1)!(n-k)!}={n-1\choose k-1} }[/math].
[math]\displaystyle{ \square }[/math]

Erdős' shifting technique

We now introduce the original proof of the Erdős–Ko–Rado theorem, which uses a technique called shifting (originally called compression).

Without loss of generality, we assume [math]\displaystyle{ X=[n] }[/math], and restate the Erdős–Ko–Rado theorem as follows.

Erdős–Ko–Rado theorem
Let [math]\displaystyle{ \mathcal{F}\subseteq {[n]\choose k} }[/math] and [math]\displaystyle{ n\ge 2k }[/math]. If [math]\displaystyle{ \mathcal{F} }[/math] is intersecting, then [math]\displaystyle{ |\mathcal{F}|\le{n-1\choose k-1} }[/math].

We define a shift operator for the set family.

Definition (shift operator)
Assume [math]\displaystyle{ \mathcal{F}\subseteq 2^{[n]} }[/math], and [math]\displaystyle{ 0\le i\lt j\le n-1 }[/math]. Define the [math]\displaystyle{ (i,j) }[/math]-shift [math]\displaystyle{ S_{ij} }[/math] as an operator on [math]\displaystyle{ \mathcal{F} }[/math] as follows:
  • for each [math]\displaystyle{ T\in\mathcal{F} }[/math], write [math]\displaystyle{ T_{ij}=(T\setminus\{j\})\cup\{i\} }[/math], and let
[math]\displaystyle{ S_{ij}(T)= \begin{cases} T_{ij} & \mbox{if }j\in T, i\not\in T, \mbox{ and }T_{ij} \not\in\mathcal{F},\\ T & \mbox{otherwise;} \end{cases} }[/math]
  • let [math]\displaystyle{ S_{ij}(\mathcal{F})=\{S_{ij}(T)\mid T\in \mathcal{F}\} }[/math].

It is easy to verify the following propositions of shifts.

Proposition
  1. [math]\displaystyle{ |S_{ij}(T)|=|T|\, }[/math] and [math]\displaystyle{ |S_{ij}(\mathcal{F})|=\mathcal{F} }[/math];
  2. if [math]\displaystyle{ \mathcal{F} }[/math] is intersecting, then so is [math]\displaystyle{ S_{ij}(\mathcal{F}) }[/math].
Proof.

(1) is immediate. Now we prove (2).

Let [math]\displaystyle{ A,B\in\mathcal{F} }[/math]. All the cases are easy to dealt with except when [math]\displaystyle{ A\cap B=\{j\} }[/math], [math]\displaystyle{ i\in A }[/math], and [math]\displaystyle{ i\not\in B }[/math]. Denote [math]\displaystyle{ A_{ij}=A\setminus\{j\}\cup\{i\} }[/math]. It holds that [math]\displaystyle{ A_{ij}\cap B=(A\cap B)\setminus\{j\}=\emptyset }[/math]. Since [math]\displaystyle{ \mathcal{F} }[/math] is intersecting, it must hold that [math]\displaystyle{ A_{ij}\not\in\mathcal{F} }[/math]. Thus, [math]\displaystyle{ S_{ij}(A)=A_{ij} }[/math] and clearly [math]\displaystyle{ S_{ij}(B)=B_{ij}=B\setminus\{j\}\cup\{i\} }[/math]. Therefore, [math]\displaystyle{ i\in S_{ij}(A)\cap S_{ij}(B) }[/math] thus [math]\displaystyle{ S_{ij}(A)\cap S_{ij}(B)\neq \emptyset }[/math].

[math]\displaystyle{ \square }[/math]

Repeatedly applying [math]\displaystyle{ S_{ij}(\mathcal{F}) }[/math] for any [math]\displaystyle{ 0\le i\lt j\le n-1 }[/math], since we only replace elements by smaller elements, eventually [math]\displaystyle{ \mathcal{F} }[/math] will stop changing, that is, [math]\displaystyle{ S_{ij}(\mathcal{F})=\mathcal{F} }[/math] for all [math]\displaystyle{ 0\le i\lt j\le n-1 }[/math]. We call such an [math]\displaystyle{ \mathcal{F} }[/math] shifted.

The idea behind the shifting technique is very natural: by applying shifting, all intersecting families are transformed to some special forms, and we only need to prove the theorem for these special form of intersecting families.

Proof of Erdős-Ko-Rado theorem
(The original proof of Erdős-Ko-Rado by shifting)

By the above lemma, it is sufficient to prove the Erdős-Ko-Rado theorem holds for shifted [math]\displaystyle{ \mathcal{F} }[/math]. We assume that [math]\displaystyle{ \mathcal{F} }[/math] is shifted.

First, it is trivial to see that the theorem holds for [math]\displaystyle{ k=1 }[/math] (no matter whether shifted).

Next, we show that the theorem holds when [math]\displaystyle{ n=2k }[/math] (no matter whether shifted). For any [math]\displaystyle{ S\in{X\choose k} }[/math], both [math]\displaystyle{ S }[/math] and [math]\displaystyle{ X\setminus S }[/math] are in [math]\displaystyle{ {X\choose k} }[/math], but at most one of them can be in [math]\displaystyle{ \mathcal{F} }[/math]. Thus,

[math]\displaystyle{ |\mathcal{F}|\le\frac{1}{2}{n\choose k}=\frac{n!}{2k!(n-k)!}=\frac{(n-1)!}{(k-1)!(n-k)!}={n-1\choose k-1} }[/math].

We then apply the induction on [math]\displaystyle{ n }[/math]. For [math]\displaystyle{ n\gt 2k }[/math], the induction hypothesis is stated as:

  • the Erdős-Ko-Rado theorem holds for any smaller [math]\displaystyle{ n }[/math].

Define

[math]\displaystyle{ \mathcal{F}_0=\{S\in\mathcal{F}\mid n\not\in S\} }[/math] and [math]\displaystyle{ \mathcal{F}_1=\{S\in\mathcal{F}\mid n\in S\} }[/math].

Clearly, [math]\displaystyle{ \mathcal{F}_0\subseteq{[n-1]\choose k} }[/math] and [math]\displaystyle{ \mathcal{F}_0 }[/math] is intersecting. Due to the induction hypothesis, [math]\displaystyle{ |\mathcal{F}_0|\le{n-2\choose k-1} }[/math].

In order to apply the induction, we let

[math]\displaystyle{ \mathcal{F}_1'=\{S\setminus\{n\}\mid S\in\mathcal{F}_1\} }[/math].

Clearly, [math]\displaystyle{ \mathcal{F}_1'\subseteq{[n-1]\choose k-1} }[/math]. If only it is also intersecting, we can apply the induction hypothesis, and indeed it is. To see this, by contradiction we assume that [math]\displaystyle{ \mathcal{F}_1' }[/math] is not intersecting. Then there must exist [math]\displaystyle{ A,B\in\mathcal{F} }[/math] such that [math]\displaystyle{ A\cap B=\{n\} }[/math], which means that [math]\displaystyle{ |A\cup B|\le 2k-1\lt n-1 }[/math]. Thus, there is some [math]\displaystyle{ 0\le i\le n-1 }[/math] such that [math]\displaystyle{ i\not\in A\cup B }[/math]. Since [math]\displaystyle{ \mathcal{F} }[/math] is shifted, [math]\displaystyle{ A_{in}=A\setminus\{n\}\cup\{i\}\in\mathcal{F} }[/math]. On the other hand it can be verified that [math]\displaystyle{ A_{in}\cap B=\emptyset }[/math], which contradicts that [math]\displaystyle{ \mathcal{F} }[/math] is intersecting.

Thus, [math]\displaystyle{ \mathcal{F}_1'\subseteq{[n-1]\choose k-1} }[/math] and [math]\displaystyle{ \mathcal{F}_1' }[/math] is intersecting. Due to the induction hypothesis, [math]\displaystyle{ |\mathcal{F}_1'|\le{n-2\choose k-2} }[/math].

Combining these together,

[math]\displaystyle{ |\mathcal{F}|=|\mathcal{F}_0|+|\mathcal{F}_1|=|\mathcal{F}_0|+|\mathcal{F}_1'|\le {n-2\choose k-1}+{n-2\choose k-2}={n-1\choose k-1} }[/math].
[math]\displaystyle{ \square }[/math]

References

(声明: 资料受版权保护, 仅用于教学.)
(Disclaimer: The following copyrighted materials are meant for educational uses only.)
  • van Lin and Wilson. A course in combinatorics. Cambridge Press. Chapter 6.
  • Aigner and Ziegler. Proofs from THE BOOK, 4th Edition. Springer-Verlag. Chapter 27.