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

From TCS Wiki
(Difference between pages)
Jump to navigation Jump to search
imported>Etone
 
imported>Etone
 
Line 1: Line 1:
== Cayley's Formula ==
== Forbidden Cliques ==
We now present a theorem of the number of labeled trees on a fixed number of vertices. It is due to [http://en.wikipedia.org/wiki/Arthur_Cayley Cayley] in 1889. The theorem is often referred by the name [http://en.wikipedia.org/wiki/Cayley's_formula Cayley's formula].
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?"
=== 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|Cayley's formula for trees|
{{Theorem|Theorem (Mantel 1907)|
: There are <math>n^{n-2}</math> different trees on <math>n</math> distinct vertices.
:Suppose <math>G(V,E)</math> is graph on <math>n</math> vertice without triangles. Then <math>|E|\le\frac{n^2}{4}</math>.
}}
}}


The theorem has several proofs. Classical methods include the bijection which encodes a tree by a [http://en.wikipedia.org/wiki/Pr%C3%BCfer_sequence Prüfer code], through the [http://en.wikipedia.org/wiki/Kirchhoff's_matrix_tree_theorem Kirchhoff's matrix tree theorem], and by double counting.
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.


===  Prüfer code ===
{{Prooftitle|First proof. (pigeonhole principle)|
The Prüfer code encodes a labeled tree to a sequence of labels. This gives a bijections between trees and tuples.
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.
{{Theorem| Prüfer code (encoder)|
 
:'''Input''': A tree <math>T</math> of <math>n</math> distinct vertices, labeled by <math>1,2,\ldots,n</math>.
Use induction on <math>n</math>. The theorem holds trivially for <math>n\le 3</math>.
:
 
:let <math>T_1=T</math>;
Induction hypothesis: assume the theorem hold for <math>|V|\le n-1</math>.
:for <math>i=1</math> to <math>n-1</math>, do
::let <math>u_i</math> be the leaf in <math>T_i</math> with the smallest label, and <math>v_i</math> be its neighbor;
::let <math>T_{i+1}</math> be the new tree obtained from deleting the leaf <math>u_i</math> from <math>T_i</math>;
:end
:return <math>(v_1,v_2,\ldots,v_{n-2})</math>;
}}


It is trivial to observe the following lemma:
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|Lemma 1|
:'''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.
:For each <math>1\le i\le n-1</math>, <math>T_i</math> is a tree of <math>n-i+1</math> vertices. In particular, the vertices of <math>T_i</math> are  <math>u_i,u_{i+1},\ldots,u_{n-1},v_{n-1}</math>, and the edges of <math>T_i</math> are precisely <math>\{u_j,v_j\}</math>, <math>i\le j\le n-1</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.
}}
}}


And there is a reason that we do not need to store <math>v_{n-1}</math> in the Prüfer code.
{{Prooftitle|Second proof. (Cauchy-Schwarz inequality)|(Mantel's original proof)
{{Theorem|Lemma 2|
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
:It always holds that <math>v_{n-1}=n</math>.
:<math>\sum_{uv\in E}(d_u+d_v)\le n|E|</math>.
}}
Note that <math>d(v)</math> appears exactly <math>d_v</math> times in the sum, so that
{{Proof|
:<math>\sum_{uv\in E}(d_u+d_v)=\sum_{v\in V}d_v^2</math>.
Every tree (of at least two vertices) has at least two leaves. The <math>u_i</math>, <math>1\le i\le n-1</math>, are the leaf of the smallest label in <math>T_i</math>, which can never be <math>n</math>, thus <math>n</math> is never deleted.
Applying Chauchy-Schwarz inequality,
:<math>
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},
</math>
where the last equation is due to Euler's equality <math>\sum_{v\in V}d_v=2|E|</math>. The theorem follows.
}}
}}


Lemma 1 and 2 together imply that given a Prüfer code <math>(v_1,v_2,\ldots,v_{n-2})</math>, the only remaining task to reconstruct the tree <math>T</math> is to figure out those <math>u_i</math>, <math>1\le i\le n-1</math>. The following lemma state how to obtain <math>u_i</math>, <math>1\le i\le n-1</math>, from a Prüfer code <math>(v_1,v_2,\ldots,v_{n-2})</math>.
{{Prooftitle|Third proof. (inequality of the arithmetic and geometric mean)|
Assume that <math>G(V,E)</math> has <math>|V|=n</math> vertices and is triangle-free.


{{Theorem|Lemma 3|
Let <math>A</math> be the largest independent set in <math>G</math> and let <math>\alpha=|A|</math>.  
:For <math>i=1,2,\ldots,n-1</math>, <math>u_i</math> is the smallest element of <math>\{1,2,\ldots,n\}</math> not in <math>\{u_1,\ldots,u_{i-1}\}\cup\{v_i,\ldots,v_{n-1}\}</math>.
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>.
}}
{{Proof|
Note that <math>u_1,u_2,\ldots,u_{n-1},v_{n-1}</math> is a sequence of distinct vertices, because <math>u_1,u_2,\ldots,u_{n-1}</math> are deleted one by one from the tree, and <math>v_{n-1}=n</math> is never deleted. Thus, each vertex <math>v</math> appears among <math>u_1,u_2,\ldots,u_{n-1},v_{n-1}</math> exactly once. And each vertex <math>v</math> appears for <math>deg(v)</math> times among the edges <math>\{u_i,v_i\}</math>, <math>1\le i\le n-1</math>, where <math>deg(v)</math> denotes the degree of vertex <math>v</math> in the original tree <math>T</math>. Therefore, each vertex <math>v</math> appears among  <math>v_1,v_2,\ldots,v_{n-2}</math> for <math>deg(v)-1</math> times.


Similarly, each vertex <math>v</math> of <math>T_i</math> appears among <math>v_i,v_{i+1},\ldots,v_{n-2}</math> for <math>deg_i(v)-1</math> times, where <math>deg_i(v)</math> is the degree of vertex <math>v</math> in the tree <math>T_i</math>. In particular, the leaves of <math>T_i</math> are not among <math>\{v_i,v_{i+1},\ldots,v_{n-2}\}</math>. Recall that the vertices of <math>T_i</math> are <math>u_i,u_{i+1},\ldots,u_{n-1},v_{n-1}</math>. Then the leaves of <math>T_i</math> are the elements of <math>\{1,2,\ldots,n\}</math> not in <math>\{u_1,\ldots,u_{i-1}\}\cup\{v_i,\ldots,v_{n-1}\}</math>. By definition of Prüfer code, <math>u_i</math> is the leaf in <math>T_i</math> of smallest label, hence the smallest element of <math>\{1,2,\ldots,n\}</math> not in <math>\{u_1,\ldots,u_{i-1}\}\cup\{v_i,\ldots,v_{n-1}\}</math>.
Take <math>B=V\setminus A</math> and let <math>\beta=|B|</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>|E|\le\sum_{v\in B}d_v\le\alpha\beta\le\left(\frac{\alpha+\beta}{2}\right)^2=\frac{n^2}{4}</math>.
}}
}}


Applying Lemma 3, we have the following decoder for the Prüfer code:
=== Turán's theorem ===
{{Theorem| Prüfer code (decoder)|
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.
:'''Input''': A tuple <math>(v_1,v_2,\ldots,v_{n-2})\in\{1,2,\ldots,n\}^{n-2}</math>.
{{Theorem|Theorem (Turán 1941)|
:
: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
:let <math>T</math> be empty graph, and <math>v_{n-1}=n</math>;
::<math>|E|\le\frac{r-2}{2(r-1)}n^2</math>.
:for <math>i=1</math> to <math>n-1</math>, do
::let <math>u_i</math> be the smallest label not in <math>\{u_1,\ldots,u_{i-1}\}\cup\{v_i,\ldots,v_{n-1}\}</math>;
::add an edge <math>\{u_i,v_i\}</math> to <math>T</math>;
:end
:return <math>T</math>;
}}
}}


In other words, the encoding of trees to tuples by the Prüfer code is reversible, thus the mapping is injective (1-1). To see it is also surjective, we need to show that for every possible <math>(v_1,v_2,\ldots,v_{n-2})\in\{1,2,\ldots,n\}^{n-2}</math>, the above decoder recovers a tree from it.  
We give an example of graphs with many edges which does not contain <math>K_r</math>.


It is easy to see that the decoder always returns a graph of <math>n-1</math> edges on the <math>n</math> vertices. The only thing remaining to verify is that the returned graph has no cycle in it, which can be easily proved by a timeline argument (left as an exercise).
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.  


Therefore, the Prüfer code establishes a bijection between the set of trees on <math>n</math> distinct vertices and the tuples from <math>\{1,2,\ldots,n\}^{n-2}</math>. This proves Cayley's formula.
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>.  


=== Double counting ===
{{Theorem|Definition|
We now present a proof of the Cayley's formula by double counting, which is considered by the [http://en.wikipedia.org/wiki/Proofs_from_THE_BOOK Proofs from THE BOOK] "the most beautiful of them all".
: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>.
{{Prooftitle|Proof of Cayley's formula by double counting|
}}
(Due to Pitman 1999)
;Example:Turán graph <math>T(13,4)</math>
[[File:Turan 13-4.svg|center|260px|Turán graph <math>T(13,4)</math>]]


Let <math>T_n</math> be the number of different trees defined on <math>n</math> distinct vertices.
Turán's theorem has been proved for many times by different mathematicians, with different tools. We show just a few.


A '''rooted tree''' is a tree with a special vertex. That is, one of the <math>n</math> vertices is marked as the "root" of the tree. A rooted tree defines a natural direction of all edges, such that an edge <math>uv</math> of the tree is directed from <math>u</math> to <math>v</math> if <math>u</math> is before <math>v</math> along the unique path from the root.
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 count the number of different ''sequences'' of directed edges that can be added to an empty graph on <math>n</math> vertices to form from it a ''rooted'' tree. We note that such a sequence can be formed in two ways:
{{Prooftitle|First proof. (induction)|(Turán's original proof)
# Starting with an unrooted tree, choose one of its vertices as root, and fix an total order of edges to specify the order in which the edges are added.
# Starting from an empty graph, add the edges one by one in steps.


In the first method, we pick one of the <math>T_n</math> unrooted trees, choose one of the <math>n</math> vertices as the root, and pick one of the <math>(n-1)!</math> total orders of the <math>n-1</math> edges. This gives us <math>T_nn(n-1)!=T_nn!</math> ways.
Induction on <math>n</math>. It is easy to verify that the theorem holds for <math>n<r</math>.  


In the second method, we consider the number of choices in one step, and multiply the numbers of choices in all steps. This is done as follows.
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>.


Given a sequence of ''adding'' <math>n-1</math> edges to an empty graph to form a rooted tree, we reverse this sequence and get a sequence of ''removing'' edges one by one from the final rooted tree until no edge left. We observe that:
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,
* At first, we remove an edge from the rooted tree. Suppose that the root of the tree is <math>r</math>, and the removed directed edge is <math>(u,v)</math>. After removing <math>(u,v)</math>, the original rooted tree is disconnected into two rooted trees, one rooted at <math>r</math> and the other rooted at <math>v</math>.
:<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>.
* After removing <math>k-1</math> edges, there are <math>k</math> rooted trees. In the <math>k</math>th step, a directed edge <math>(u,v)</math> in the current forest is removed and the tree containing <math>(u,v)</math> is disconnected into two trees, one rooted at the old root of that tree, and the other rooted at <math>v</math>.
 
We now again reverse the above procedure, and consider the sequence of adding directed edges to an empty graph to form a rooted tree.
* At first, we have <math>n</math> rooted trees, each of 0 edge (<math>n</math> isolated vertices).
* After adding <math>n-k</math> edges, there are <math>k</math> rooted trees. Denoting the directed edge added next as <math>(u,v)</math>. As observed above, <math>u</math> can be any one of the <math>n</math> vertices; but <math>v</math> must be the root of one of the <math>k</math> trees, except the tree which contains <math>u</math>. There are <math>n(k-1)</math> choices of such <math>(u,v)</math>.
Multiplying the numbers of choices in all steps, the number of sequences of adding directed edges to an empty graph to form a rooted tree is given by
:<math>\prod_{k=2}^nn(k-1)=n^{n-2}n!</math>.
 
By the principle of double counting, counting the same thing by different methods yield the same result.
:<math>T_nn!=n^{n-2}n!</math>,
which gives that <math>T_n=n^{n-2}</math>.
}}
}}


== Existence by Counting ==
{{Prooftitle|Second proof. (weight shifting)|(due to Motzkin and Straus)
=== Shannon's circuit lower bound===
This is a fundamental problem in in Computer Science.


A '''boolean function''' is a function in the form <math>f:\{0,1\}^n\rightarrow \{0,1\}</math>.
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
:<math>S=\sum_{uv\in E}w_uw_v</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.


[http://en.wikipedia.org/wiki/Boolean_circuit Boolean circuit] is a mathematical model of computation.
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.
Formally, a boolean circuit is a directed acyclic graph. Nodes with indegree zero are input nodes, labeled <math>x_1, x_2, \ldots , x_n</math>. A circuit has a unique node with outdegree zero, called the output node. Every other node is a gate. There are three types of gates: AND, OR (both with indegree two), and NOT (with indegree one).


Computations in Turing machines can be simulated by circuits, and any boolean function in '''P''' can be computed by a circuit with polynomially many gates. Thus, if we can find a function in '''NP''' that cannot be computed by any circuit with polynomially many gates, then '''NP'''<math>\neq</math>'''P'''.
<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 following theorem due to Shannon says that functions with exponentially large circuit complexity do exist.
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>,
:<math>S=\sum_{uv\in E}w_uw_v=\frac{|E|}{n^2}</math>.
Thus,
:<math>\frac{|E|}{n^2}\le \frac{r-2}{2(r-1)}</math>,
which implies the theorem.
}}


{{Theorem
{{Prooftitle|Third proof. (the probabilistic method)|(due to Alon and Spencer)
|Theorem (Shannon 1949)|
:There is a boolean function <math>f:\{0,1\}^n\rightarrow \{0,1\}</math> with circuit complexity greater than <math>\frac{2^n}{3n}</math>.
}}
{{Proof|
We first count the number of boolean functions <math>f:\{0,1\}^n\rightarrow \{0,1\}</math>. There are <math>2^{2^n}</math> boolean functions <math>f:\{0,1\}^n\rightarrow \{0,1\}</math>.


Then we count the number of boolean circuit with fixed number of gates.
Write <math>\omega(G)</math> for the number of vertices in a largest clique, called the '''clique number''' of <math>G</math>.  
Fix an integer <math>t</math>, we count the number of circuits with <math>t</math> gates. By the [http://en.wikipedia.org/wiki/De_Morgan's_laws De Morgan's laws], we can assume that all NOTs are pushed back to the inputs. Each gate has one of the two types (AND or OR), and has two inputs. Each of the inputs to a gate is either a constant 0 or 1, an input variable <math>x_i</math>, an inverted input variable <math>\neg x_i</math>, or the output of another gate; thus, there are at most <math>2+2n+t-1</math> possible gate inputs. It follows that the number of circuits with <math>t</math> gates is at most <math>2^t(t+2n+1)^{2t}</math>.  
:'''Claim:''' <math>\omega(G)\ge\sum_{v\in V}\frac{1}{n-d_v}</math>.
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:
*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>.
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>.


If <math>t=2^n/3n</math>, then
Let <math>X_v</math> be the random variable that indicates whether <math>v\in S</math>, i.e.,
:<math>
:<math>
\frac{2^t(t+2n+1)^{2t}}{2^{2^n}}=o(1)<1,</math>     thus, <math>2^t(t+2n+1)^{2t} < 2^{2^n}.</math>
X_v=\begin{cases}
1 & v\in S,\\
0 & \mbox{otherwise.}
\end{cases}
</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,
:<math>\mathbf{E}[X_v]=\Pr[v\in S]=\frac{1}{n-d_v}.</math>
Observe that <math>|S|=\sum_{v\in V}X_v</math>. Due to 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>.
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.


Each boolean circuit computes one boolean function. Therefore, there must exist a boolean function <math>f</math> which cannot be computed by any circuits with <math>2^n/3n</math> gates.
Apply the Cauchy-Schwarz inequality
:<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>.
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>
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
:<math>n^2\le (r-1)(n^2-2|E|)</math>,
which implies the theorem.
}}
}}


Note that by Shannon's theorem, not only there exists a boolean function with exponentially large circuit complexity, but ''almost all'' boolean functions have exponentially large circuit complexity.
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.
{{Prooftitle|Fourth proof.|
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.
:'''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>.
Suppose otherwise. There are two cases.
* '''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>,
: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.


=== Double counting ===
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 double counting principle states the following obvious fact: if the elements of a set are counted in two different ways, the answers are the same.
;Handshaking lemma
The following lemma is a standard demonstration of double counting.
{{Theorem|Handshaking Lemma|
:At a party, the number of guests who shake hands an odd number of times is even.
}}
}}


We model this scenario as an undirected graph <math>G(V,E)</math> with <math>|V|=n</math> standing for the <math>n</math> guests. There is an edge <math>uv\in E</math> if <math>u</math> and <math>v</math> shake hands. Let <math>d(v)</math> be the degree of vertex <math>v</math>, which represents the number of times that <math>v</math> shakes hand. The handshaking lemma states that in any undirected graph, the number of vertices whose degrees are odd is even. It is sufficient to show that the sum of odd degrees is even.
== Forbidden Cycles ==
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").


The handshaking lemma is a direct consequence of the following lemma, which is proved by Euler in his 1736 paper on [http://en.wikipedia.org/wiki/Seven_Bridges_of_K%C3%B6nigsberg Seven Bridges of Königsberg] that began the study of graph theory.
{{Theorem|Theorem|
 
: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>.
{{Theorem|Lemma (Euler 1736)|
:<math>\sum_{v\in V}d(v)=2|E|</math>
}}
}}
{{Proof|
{{Proof|
We count the number of '''directed''' edges. A directed edge is an ordered pair <math>(u,v)</math> such that <math>\{u,v\}\in E</math>. There are two ways to count the directed edges.
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>.


First, we can enumerate by edges. Pick every edge <math>uv\in E</math> and apply two directions <math>(u,v)</math> and <math>(v,u)</math> to the edge. This gives us <math>2|E|</math> directed edges.
* 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>.
* 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>.


On the other hand, we can enumerate by vertices. Pick every vertex <math>v\in V</math> and for each of its <math>d(v)</math> neighbors, say <math>u</math>, generate a directed edge <math>(v,u)</math>. This gives us <math>\sum_{v\in V}d(v)</math> directed edges.
Therefore, <math>\{u,v_1,v_2,\ldots,v_d\}\cup S_1\cup S_2\cup\cdots\cup S_d\subseteq V</math> implies that
:<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>.


It is obvious that the two terms are equal, since we just count the same thing twice with different methods. The lemma follows.
By Cauchy-Schwarz inequality,
:<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>,
which implies that <math>|E|\le\frac{1}{2}n\sqrt{n-1}</math>.
}}
}}


The handshaking lemma is implied directly by the above lemma, since the sum of even degrees is even.
=== Hamiltonian cycle ===
We now look at graphs which does not have large cycles. In particular, we consider graphs without '''Hamiltonian cycles'''.


== The Pigeonhole Principle ==
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.
The '''pigeonhole principle''' states the following "obvious" fact:
:''<math>n+1</math> pigeons cannot sit in <math>n</math> holes so that every pigeon is alone in its hole.''
This is one of the oldest '''non-constructive''' principles: it states only the ''existence'' of a pigeonhole with more than one pigeons and says nothing about how to ''find'' such a pigeonhole.


The general form of pigeonhole principle, also known as the '''averaging principle''', is stated as follows.
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'''.
{{Theorem|Generalized pigeonhole principle|
{{Theorem|Dirac's Theorem|
:If a set consisting of more than <math>mn</math> objects is partitioned into <math>n</math> classes, then some class receives more than <math>m</math> objects.
: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>.
}}
}}
{{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>.


=== Monotonic subsequences ===
Consider the sets,
Let <math>(a_1,a_2,\ldots,a_n)</math> be a sequence of <math>n</math> distinct real numbers. A '''subsequence''' is a sequence of distinct terms of <math>(a_1,a_2,\ldots,a_n)</math> appearing in the same order in which they appear in <math>(a_1,a_2,\ldots,a_n)</math>. Formally, a subsequence of <math>(a_1,a_2,\ldots,a_n)</math> is an <math>(a_{i_1},a_{i_2},\ldots,a_{i_k})</math>, with <math>i_1<i_2<\cdots<i_k</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>.  


A sequence <math>(a_1,a_2,\ldots,a_n)</math> is '''increasing''' if <math>a_1<a_2<\cdots<a_n</math>, and '''decreasing''' if <math>a_1>a_2>\cdots>a_n</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:
 
:<math>v_1v_{i+1}v_{i+2}\cdots v_nv_{i}v_{i-1}\cdots v_1\cdots</math>,
We are interested in the ''longest'' increasing and decreasing subsequences of an <math>a_1<a_2<\cdots<a_n</math>. It is intuitive that the length of both the longest increasing subsequence and the longest decreasing subsequence cannot be small simultaneously. A famous result of Erdős and Szekeres formally justifies this intuition. This is one of the first results in extremal combinatorics, published in the influential 1935 paper of Erdős and Szekeres.
which contradict to the assumption that <math>G</math> is non-Hamiltonian.
}}


{{Theorem|Theorem (Erdős-Szekeres 1935)|
== Erdős–Stone theorem ==
:A sequence of more than <math>mn</math> different real numbers must contain either an increasing subsequence of length <math>m+1</math>, or a decreasing subsequence of length <math>n+1</math>.
We introduce a notation for the number of edges in extremal graphs with a specific forbidden substructure.
{{Theorem|Definition|
: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.
}}
}}
{{Proof|(due to Seidenberg 1959)
With this notation, Turán's theorem can be restated as
Let <math>(a_1,a_2,\ldots,a_{N})</math> be the original sequence of <math>N>mn</math> distinct real numbers. Associate each <math>a_i</math> a pair <math>(x_i,y_i)</math>, defined as:
{{Theorem|Turán's theorem (restated)|
*<math>x_i</math>: the length of the longest ''increasing'' subsequence ''ending'' at <math>a_i</math>;
:<math>\mathrm{ex}(n,K_r)\le\frac{r-2}{2(r-1)}n^2</math>.
*<math>y_i</math>: the length of the longest ''decreasing'' subsequence ''starting'' at <math>a_i</math>.
A key observation is that <math>(x_i,y_i)\neq (x_j,y_j)</math> whenever <math>i\neq j</math>. This is proved as follows:
: '''Case 1:''' If <math>a_i<a_j</math>, then the longest increasing subsequence ending at <math>a_i</math> can be extended by adding on <math>a_j</math>, so <math>x_i<x_j</math>.
: '''Case 2:'''  If <math>a_i>a_j</math>, then the longest decreasing subsequence starting at <math>a_j</math> can be preceded by <math>a_i</math>, so <math>y_i>y_j</math>.
Now we put <math>N</math> "pigeons" <math>a_1,a_2,\ldots,a_N</math> into "pigeonholes" <math>\{1,2,\ldots,N\}\times\{1,2,\ldots,N\}</math>, such that <math>a_i</math> is put into hole <math>(x_i,y_i)</math>, with at most one pigeon per each hole (since different <math>a_i</math> has different <math>(x_i,y_i)</math>).
 
The number of pigeons is <math>N>mn</math>. Due to pigeonhole principle, there must be a pigeon which is outside the region <math>\{1,2,\ldots,m\}\times\{1,2,\ldots,n\}</math>, which implies that there exists an <math>a_i</math> with either <math>x_i>m</math> or <math>y_i>n</math>. Due to our definition of <math>(x_i,y_i)</math>, there must be either an increasing subsequence of length <math>m+1</math>, or a decreasing subsequence of length <math>n+1</math>.
}}
}}


=== Dirichlet's approximation ===
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>.
Let <math>x</math> be an irrational number. We now want to approximate <math>x</math> be a rational number (a fraction).
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>.


Since every real interval <math>[a,b]</math> with <math>a<b</math> contains infinitely many rational numbers, there must exist rational numbers arbitrarily close to <math>x</math>. The trick is to let the denominator of the fraction sufficiently large.
{{Theorem|Fundamental theorem of extremal graph theory (Erdős–Stone 1946)|
: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>.
}}


Suppose however we restrict the rationals we may select to have denominators bounded by <math>n</math>. How closely we can approximate <math>x</math> now?
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.


The following important theorem is due to Dirichlet and his ''Schubfachprinzip'' ("drawer principle"). The theorem is fundamental in numer theory and real analysis, but the proof is combinatorial.
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.


{{Theorem|Theorem (Dirichlet 1879)|
{{Theorem|Corollary|
:Let <math>x</math> be an irrational number. For any natural number <math>n</math>, there is a rational number <math>\frac{p}{q}</math> such that <math>1\le q\le n</math> and
:For every nonempty graph <math>H</math>,
::<math>\left|x-\frac{p}{q}\right|<\frac{1}{nq}</math>.
::<math>\lim_{n\rightarrow\infty}\frac{\mathrm{ex}(n,H)}{{n\choose 2}}=\frac{\chi(H)-2}{\chi(H)-1}</math>.
}}
}}
{{Proof|
{{Prooftitle|Proof of corollary|
Let <math>\{x\}=x-\lfloor x\rfloor</math> denote the '''fractional part''' of the real number <math>x</math>. It is obvious that <math>\{x\}\in[0,1)</math> for any real number <math>x</math>.
Let <math>r=\chi(H)</math>.  


Consider the <math>n+1</math> numbers <math>\{kx\}</math>, <math>k=1,2,\ldots,n+1</math>. These <math>n+1</math> numbers (pigeons) belong to the following <math>n</math> intervals (pigeonholes):
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,
:<math>\left(0,\frac{1}{n}\right),\left(\frac{1}{n},\frac{2}{n}\right),\ldots,\left(\frac{n-1}{n},1\right)</math>.
:<math>|T(n,r-1)|\le\mathrm{ex}(n,H)</math>.
Since <math>x</math> is irrational, <math>\{kx\}</math> cannot coincide with any endpoint of the above intervals.
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>.


By the pigeonhole principle, there exist <math>1\le a<b\le n+1</math>, such that <math>\{ax\},\{bx\}</math> are in the same interval, thus
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,  
:<math>|\{bx\}-\{ax\}|<\frac{1}{n}</math>.
:<math>\mathrm{ex}(n,H)\le\mathrm{ex}(n,K_s^r)</math>.
Therefore,
Due to Erdős–Stone theorem,
:<math>|(b-a)x-\left(\lfloor bx\rfloor-\lfloor ax\rfloor\right)|<\frac{1}{n}</math>.
:<math>\mathrm{ex}(n,K_s^r)=\left(\frac{r-2}{2(r-1)}+o(1)\right)n^2</math>.
Let <math>q=b-a</math> and <math>p=\lfloor bx\rfloor-\lfloor ax\rfloor</math>. We have <math>|qx-p|<\frac{1}{n}</math> and <math>1\le q\le n</math>. Dividing both sides by <math>q</math>, the theorem is proved.
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 217: Line 239:
:('''Disclaimer:''' The following copyrighted materials are meant for educational uses only.)
:('''Disclaimer:''' The following copyrighted materials are meant for educational uses only.)


* Aigner and Ziegler. ''Proofs from THE BOOK, 4th Edition.'' Springer-Verlag. [[media:PFTB_chap25.pdf| Chapter 25]] and [[media:PFTB_chap30.pdf| Chapter 30]].  
* van Lin and Wilson. ''A course in combinatorics.'' Cambridge Press. Chapter 4.
* Alon and Spencer. ''The Probabilistic Method, 3rd Edition.'' Wiley, 2008. [[media:TPM_Chap1.pdf|Chapter 1]], [[media:TPM_Chap2.pdf|Chapter 2]], and [[media:TPM_Chap3.pdf|Chapter 3]].
* Aigner and Ziegler. ''Proofs from THE BOOK, 4th Edition.'' Springer-Verlag. [[media:PFTB_chap36.pdf| Chapter 36]].  
* Diestel. ''Graph Theory, 3rd Edition''. Springer-Verlag 2000. [[media:Diestel2ed_chap7.pdf|Chapter 7]].

Latest revision as of 01:04, 3 November 2011

Forbidden Cliques

Extremal graph theory studies the problems like "how many edges that a graph [math]\displaystyle{ G }[/math] can have, if [math]\displaystyle{ G }[/math] has some property?"

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]\displaystyle{ K_3 }[/math].

Theorem (Mantel 1907)
Suppose [math]\displaystyle{ G(V,E) }[/math] is graph on [math]\displaystyle{ n }[/math] vertice without triangles. Then [math]\displaystyle{ |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.

First proof. (pigeonhole principle)

We prove an equivalent theorem: Any [math]\displaystyle{ G(V,E) }[/math] with [math]\displaystyle{ |V|=n }[/math] and [math]\displaystyle{ |E|\gt \frac{n^2}{4} }[/math] must have a triangle.

Use induction on [math]\displaystyle{ n }[/math]. The theorem holds trivially for [math]\displaystyle{ n\le 3 }[/math].

Induction hypothesis: assume the theorem hold for [math]\displaystyle{ |V|\le n-1 }[/math].

For [math]\displaystyle{ G }[/math] with [math]\displaystyle{ n }[/math] vertices, without loss of generality, assume that [math]\displaystyle{ |E|=\frac{n^2}{4}+1 }[/math], we will show that [math]\displaystyle{ G }[/math] must contain a triangle. Take a [math]\displaystyle{ uv\in E }[/math], and let [math]\displaystyle{ H }[/math] be the subgraph of [math]\displaystyle{ G }[/math] induced by [math]\displaystyle{ V\setminus \{u,v\} }[/math]. Clearly, [math]\displaystyle{ H }[/math] has [math]\displaystyle{ n-2 }[/math] vertices.

Case.1: If [math]\displaystyle{ H }[/math] has [math]\displaystyle{ \gt \frac{(n-2)^2}{4} }[/math] edges, then by the induction hypothesis, [math]\displaystyle{ H }[/math] has a triangle.
Case.2: If [math]\displaystyle{ H }[/math] has [math]\displaystyle{ \le\frac{(n-2)^2}{4} }[/math] edges, then at least [math]\displaystyle{ \left(\frac{n^2}{4}+1\right)-\frac{(n-2)^2}{4}-1=n-1 }[/math] edges are between [math]\displaystyle{ H }[/math] and [math]\displaystyle{ \{u,v\} }[/math]. By pigeonhole principle, there must be a vertex in [math]\displaystyle{ H }[/math] that is adjacent to both [math]\displaystyle{ u }[/math] and [math]\displaystyle{ v }[/math]. Thus, [math]\displaystyle{ G }[/math] has a triangle.
[math]\displaystyle{ \square }[/math]
Second proof. (Cauchy-Schwarz inequality)
(Mantel's original proof)

For any edge [math]\displaystyle{ uv\in E }[/math], no vertex can be a neighbor of both [math]\displaystyle{ u }[/math] and [math]\displaystyle{ v }[/math], or otherwise there will be a triangle. Thus, for any edge [math]\displaystyle{ uv\in E }[/math], [math]\displaystyle{ d_u+d_v\le n }[/math]. It follows that

[math]\displaystyle{ \sum_{uv\in E}(d_u+d_v)\le n|E| }[/math].

Note that [math]\displaystyle{ d(v) }[/math] appears exactly [math]\displaystyle{ d_v }[/math] times in the sum, so that

[math]\displaystyle{ \sum_{uv\in E}(d_u+d_v)=\sum_{v\in V}d_v^2 }[/math].

Applying Chauchy-Schwarz inequality,

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

where the last equation is due to Euler's equality [math]\displaystyle{ \sum_{v\in V}d_v=2|E| }[/math]. The theorem follows.

[math]\displaystyle{ \square }[/math]
Third proof. (inequality of the arithmetic and geometric mean)

Assume that [math]\displaystyle{ G(V,E) }[/math] has [math]\displaystyle{ |V|=n }[/math] vertices and is triangle-free.

Let [math]\displaystyle{ A }[/math] be the largest independent set in [math]\displaystyle{ G }[/math] and let [math]\displaystyle{ \alpha=|A| }[/math]. Since [math]\displaystyle{ G }[/math] is triangle-free, for very vertex [math]\displaystyle{ v }[/math], all its neighbors must form an independent set, thus [math]\displaystyle{ d(v)\le \alpha }[/math] for all [math]\displaystyle{ v\in V }[/math].

Take [math]\displaystyle{ B=V\setminus A }[/math] and let [math]\displaystyle{ \beta=|B| }[/math]. Since [math]\displaystyle{ A }[/math] is an independent set, all edges in [math]\displaystyle{ E }[/math] must have at least one endpoint in [math]\displaystyle{ B }[/math]. Counting the edges in [math]\displaystyle{ E }[/math] according to their endpoints in [math]\displaystyle{ B }[/math], we obtain [math]\displaystyle{ |E|\le\sum_{v\in B}d_v }[/math]. By the inequality of the arithmetic and geometric mean,

[math]\displaystyle{ |E|\le\sum_{v\in B}d_v\le\alpha\beta\le\left(\frac{\alpha+\beta}{2}\right)^2=\frac{n^2}{4} }[/math].
[math]\displaystyle{ \square }[/math]

Turán's theorem

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 (Turán 1941)
Let [math]\displaystyle{ G(V,E) }[/math] be a graph with [math]\displaystyle{ |V|=n }[/math]. If [math]\displaystyle{ G }[/math] has no [math]\displaystyle{ r }[/math]-clique, [math]\displaystyle{ r\ge 2 }[/math], then
[math]\displaystyle{ |E|\le\frac{r-2}{2(r-1)}n^2 }[/math].

We give an example of graphs with many edges which does not contain [math]\displaystyle{ K_r }[/math].

Partition [math]\displaystyle{ V }[/math] into [math]\displaystyle{ r-1 }[/math] disjoint classes [math]\displaystyle{ V=V_1\cup V_2\cup\cdots\cup V_{r-1} }[/math], [math]\displaystyle{ n_i=|V_i| }[/math], [math]\displaystyle{ n_1+n_2+\cdots+n_{r-1}=n }[/math]. For every two vertice [math]\displaystyle{ u,v }[/math], [math]\displaystyle{ uv\in E }[/math] if and only if [math]\displaystyle{ u\in V_i }[/math] and [math]\displaystyle{ v\in V_j }[/math] for distinct [math]\displaystyle{ V_i }[/math] and [math]\displaystyle{ V_j }[/math]. The resulting graph is a complete [math]\displaystyle{ (r-1) }[/math]-partite graph, denoted [math]\displaystyle{ K_{n_1,n_2,\ldots,n_{r-1}} }[/math]. It is obvious that any [math]\displaystyle{ (r-1) }[/math]-partite graph contains no [math]\displaystyle{ r }[/math]-clique since only those vertices from different classes can be adjacent.

A [math]\displaystyle{ K_{n_1,n_2,\ldots,n_{r-1}} }[/math] has [math]\displaystyle{ \sum_{i\lt j}n_i n_j\, }[/math] edges, which is maximized when the numbers [math]\displaystyle{ n_i }[/math] are divided as evenly as possible, that is, if [math]\displaystyle{ 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]\displaystyle{ 1\le i\le r-1 }[/math].

Definition
We call a complete multipartite graph [math]\displaystyle{ K_{n_1,n_2,\ldots,n_{r-1}} }[/math] with [math]\displaystyle{ 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]\displaystyle{ i }[/math] a Turán graph, denoted [math]\displaystyle{ T(n,r-1) }[/math].
Example
Turán graph [math]\displaystyle{ T(13,4) }[/math]
Turán graph [math]\displaystyle{ T(13,4) }[/math]
Turán graph [math]\displaystyle{ T(13,4) }[/math]

Turán's theorem has been proved for many times by different mathematicians, with different tools. We show just a few.

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.

First proof. (induction)
(Turán's original proof)

Induction on [math]\displaystyle{ n }[/math]. It is easy to verify that the theorem holds for [math]\displaystyle{ n\lt r }[/math].

Let [math]\displaystyle{ G }[/math] be a graph on [math]\displaystyle{ n }[/math] vertices without [math]\displaystyle{ r }[/math]-cliques where [math]\displaystyle{ n\ge r }[/math]. Suppose that [math]\displaystyle{ G }[/math] has a maximum number of edges among such graphs. [math]\displaystyle{ G }[/math] certainly has [math]\displaystyle{ (r-1) }[/math]-cliques, since otherwise we could add edges to [math]\displaystyle{ G }[/math]. Let [math]\displaystyle{ A }[/math] be an [math]\displaystyle{ (r-1) }[/math]-clique and let [math]\displaystyle{ B=V\setminus A }[/math]. Clearly [math]\displaystyle{ |A|=r-1 }[/math] and [math]\displaystyle{ |B|=n-r+1 }[/math].

By the induction hypothesis, since [math]\displaystyle{ B }[/math] has no [math]\displaystyle{ r }[/math]-cliques, [math]\displaystyle{ |E(B)|\le\frac{r-2}{2(r-1)}(n-r+1)^2 }[/math]. And [math]\displaystyle{ E(A)={r-1\choose 2} }[/math]. Since [math]\displaystyle{ G }[/math] has no [math]\displaystyle{ r }[/math]-clique, every [math]\displaystyle{ v\in B }[/math] is adjacent to at most [math]\displaystyle{ r-2 }[/math] vertices in [math]\displaystyle{ A }[/math], since otherwise [math]\displaystyle{ A }[/math] and [math]\displaystyle{ v }[/math] would form an [math]\displaystyle{ r }[/math]-clique. We obtain that the number edges crossing between [math]\displaystyle{ A }[/math] and [math]\displaystyle{ B }[/math] is [math]\displaystyle{ |E(A,B)|\le (r-2)|B|=(r-2)(n-r+1) }[/math]. Combining everything together,

[math]\displaystyle{ |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].
[math]\displaystyle{ \square }[/math]
Second proof. (weight shifting)
(due to Motzkin and Straus)

Assign each vertex [math]\displaystyle{ v\in V }[/math] a nonnegative weight [math]\displaystyle{ w_v\ge 0 }[/math], and assume that [math]\displaystyle{ \sum_{v\in V}w_v=1 }[/math]. We try to maximize the quantity

[math]\displaystyle{ S=\sum_{uv\in E}w_uw_v }[/math].

Let [math]\displaystyle{ W_u=\sum_{v:v\sim u}w_v\, }[/math] be the sum of the weights of [math]\displaystyle{ u }[/math]'s neighbors. Note that [math]\displaystyle{ S }[/math] can also be computed as [math]\displaystyle{ S=\frac{1}{2}\sum_{u\in V}w_uW_u }[/math]. For any nonadjacent pair of vertices [math]\displaystyle{ u\not\sim v }[/math], supposed that [math]\displaystyle{ W_u\ge W_v }[/math], then for any [math]\displaystyle{ \epsilon\ge 0 }[/math],

[math]\displaystyle{ (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]\displaystyle{ S }[/math] by shifting all of the weight of the vertex [math]\displaystyle{ v }[/math] to the vertex [math]\displaystyle{ u }[/math]. It follows that [math]\displaystyle{ S }[/math] is maximized when all of the weight is concentrated on a complete subgraph, i.e., a clique.

Now if [math]\displaystyle{ w_u\gt w_v\gt 0 }[/math], then choose [math]\displaystyle{ \epsilon }[/math] with [math]\displaystyle{ 0\lt \epsilon\lt w_u-w_v }[/math] and change [math]\displaystyle{ w_u'=w_u-\epsilon }[/math] and [math]\displaystyle{ w_v'=w_v+\epsilon }[/math]. This changes [math]\displaystyle{ S }[/math] to [math]\displaystyle{ S'=S+\epsilon(w_u-w_v)-\epsilon^2\gt S }[/math]. Thus, the maximal value of [math]\displaystyle{ S }[/math] is attained when all nonzero weights are equal and concentrated on a clique.

[math]\displaystyle{ G }[/math] has at most an [math]\displaystyle{ (r-1) }[/math]-clique, thus [math]\displaystyle{ S\le{r-1\choose 2}\frac{1}{(r-1)^2}=\frac{r-2}{2(r-1)} }[/math].

As we argued above, this inequality hold for any nonnegative weight assignments with [math]\displaystyle{ \sum_{v\in V}w_v=1 }[/math]. In particular, for the case that all [math]\displaystyle{ w_v=\frac{1}{n} }[/math],

[math]\displaystyle{ S=\sum_{uv\in E}w_uw_v=\frac{|E|}{n^2} }[/math].

Thus,

[math]\displaystyle{ \frac{|E|}{n^2}\le \frac{r-2}{2(r-1)} }[/math],

which implies the theorem.

[math]\displaystyle{ \square }[/math]
Third proof. (the probabilistic method)
(due to Alon and Spencer)

Write [math]\displaystyle{ \omega(G) }[/math] for the number of vertices in a largest clique, called the clique number of [math]\displaystyle{ G }[/math].

Claim: [math]\displaystyle{ \omega(G)\ge\sum_{v\in V}\frac{1}{n-d_v} }[/math].

We prove this by the probabilistic method. Fix a random ordering of vertices in [math]\displaystyle{ V }[/math], say [math]\displaystyle{ v_1,v_2,\ldots,v_n }[/math]. We construct a clique as follows:

  • for [math]\displaystyle{ i=1,2,\ldots, n }[/math], add [math]\displaystyle{ v_i }[/math] to [math]\displaystyle{ S }[/math] iff all vertices in current [math]\displaystyle{ S }[/math] are adjacent to [math]\displaystyle{ v_i }[/math].

It is obvious that an [math]\displaystyle{ S }[/math] constructed in this way is a clique. We now show that [math]\displaystyle{ \mathbf{E}[|S|]=\sum_{v\in V}\frac{1}{n-d_v} }[/math].

Let [math]\displaystyle{ X_v }[/math] be the random variable that indicates whether [math]\displaystyle{ v\in S }[/math], i.e.,

[math]\displaystyle{ X_v=\begin{cases} 1 & v\in S,\\ 0 & \mbox{otherwise.} \end{cases} }[/math]

Note that a vertex [math]\displaystyle{ v\in S }[/math] if and only if [math]\displaystyle{ v }[/math] is ranked before all its [math]\displaystyle{ n-d_v-1 }[/math] non-neighbors in the random ordering. The probability that this event occurs is [math]\displaystyle{ \frac{1}{n-d_v} }[/math]. Thus,

[math]\displaystyle{ \mathbf{E}[X_v]=\Pr[v\in S]=\frac{1}{n-d_v}. }[/math]

Observe that [math]\displaystyle{ |S|=\sum_{v\in V}X_v }[/math]. Due to linearity of expectation,

[math]\displaystyle{ \mathbf{E}[|S|]=\sum_{v\in V}\mathbf{E}[X_v]=\sum_{v\in V}\frac{1}{n-d_v} }[/math].

There must exists a clique of at least such size, so that [math]\displaystyle{ \omega(G)\ge\sum_{v\in V}\frac{1}{n-d_v} }[/math]. The claim is proved.

Apply the Cauchy-Schwarz inequality

[math]\displaystyle{ \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].

Set [math]\displaystyle{ a_v=\sqrt{n-d_v} }[/math] and [math]\displaystyle{ b_v=\frac{1}{\sqrt{n-d_v}} }[/math], then [math]\displaystyle{ a_vb_v=1 }[/math] and so

[math]\displaystyle{ 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]

By the assumption of Turán's theorem, [math]\displaystyle{ \omega(G)\le r-1 }[/math]. Recall the handshaking lemma [math]\displaystyle{ 2|E|=\sum_{v\in V}d_v }[/math]. The above inequality gives us

[math]\displaystyle{ n^2\le (r-1)(n^2-2|E|) }[/math],

which implies the theorem.

[math]\displaystyle{ \square }[/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 only possible extremal graphs.

Fourth proof.

Let [math]\displaystyle{ G(V,E) }[/math] be a [math]\displaystyle{ r }[/math]-clique-free graph on [math]\displaystyle{ n }[/math] vertices with a maximum number of edges.

Claim: [math]\displaystyle{ G }[/math] does not contain three vertices [math]\displaystyle{ u,v,w }[/math] such that [math]\displaystyle{ uv\in E }[/math] but [math]\displaystyle{ uw\not\in E, vw\not\in E }[/math].

Suppose otherwise. There are two cases.

  • Case.1: [math]\displaystyle{ d(w)\lt d(u) }[/math] or [math]\displaystyle{ d(w)\lt d(v) }[/math]. Without loss of generality, suppose that [math]\displaystyle{ d(w)\lt d(u) }[/math]. We duplicate [math]\displaystyle{ u }[/math] by creating a new vertex [math]\displaystyle{ u' }[/math] which has exactly the same neighbors as [math]\displaystyle{ u }[/math] (but [math]\displaystyle{ uu' }[/math] is not an edge). Such duplication will not increase the clique size. We then remove [math]\displaystyle{ w }[/math]. The resulting graph [math]\displaystyle{ G' }[/math] is still [math]\displaystyle{ r }[/math]-clique-free, and has [math]\displaystyle{ n }[/math] vertices. The number of edges in [math]\displaystyle{ G' }[/math] is
[math]\displaystyle{ |E(G')|=|E(G)|+d(u)-d(w)\gt |E(G)|\, }[/math],
which contradicts the assumption that [math]\displaystyle{ |E(G)| }[/math] is maximal.
  • Case.2: [math]\displaystyle{ d(w)\ge d(u) }[/math] and [math]\displaystyle{ d(w)\ge d(v) }[/math]. Duplicate [math]\displaystyle{ w }[/math] twice and delete [math]\displaystyle{ u }[/math] and [math]\displaystyle{ v }[/math]. The new graph [math]\displaystyle{ G' }[/math] has no [math]\displaystyle{ r }[/math]-clique, and the number of edges is
[math]\displaystyle{ |E(G')|=|E(G)|+2d(w)-(d(u)+d(v)+1)\gt |E(G)|\, }[/math].
Contradiction again.

The claim implies that [math]\displaystyle{ 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]\displaystyle{ G }[/math] must be a complete multipartite graph [math]\displaystyle{ K_{n_1,n_2,\ldots,n_{r-1}} }[/math] with [math]\displaystyle{ n_1+n_2+\cdots +n_{r-1}=n }[/math]. Optimize the edge number, we have the Turán graph.

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

Forbidden Cycles

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]\displaystyle{ G }[/math] is the length of the shortest cycle in [math]\displaystyle{ G }[/math]. A graph is triangle-free if and only if its girth [math]\displaystyle{ g(G)\ge 4 }[/math]. Matel's theorem can be seen as a bound on the edge number of graphs with girth [math]\displaystyle{ g(G)\ge 4 }[/math]. The next theorem extends this bound to the graphs with [math]\displaystyle{ g(G)\ge 5 }[/math], i.e., graphs without triangles and quadrilaterals ("squares").

Theorem
Let [math]\displaystyle{ G(V,E) }[/math] be a graph on [math]\displaystyle{ n }[/math] vertices. If girth [math]\displaystyle{ g(G)\ge 5 }[/math] then [math]\displaystyle{ |E|\le\frac{1}{2}n\sqrt{n-1} }[/math].
Proof.

Suppose [math]\displaystyle{ g(G)\ge 5 }[/math]. Let [math]\displaystyle{ v_1,v_2,\ldots,v_d }[/math] be the neighbors of a vertex [math]\displaystyle{ u }[/math], where [math]\displaystyle{ d=d(u) }[/math]. Let [math]\displaystyle{ S_i=\{v\in V\mid v\sim v_i\wedge v\neq u\} }[/math] be the set of neighbors of [math]\displaystyle{ v_i }[/math] other than [math]\displaystyle{ u }[/math].

  • For any [math]\displaystyle{ v_i,v_j }[/math], [math]\displaystyle{ v_iv_j\not\in E }[/math] since [math]\displaystyle{ G }[/math] has no triangle. Thus, [math]\displaystyle{ S_i\cap\{u,v_1,v_2,\ldots,v_d\}=\emptyset }[/math] for every [math]\displaystyle{ i }[/math].
  • No vertex other than [math]\displaystyle{ u }[/math] can be adjacent to more than one vertices in [math]\displaystyle{ v_1,v_2,\ldots,v_d }[/math] since there is no [math]\displaystyle{ C_4 }[/math] in [math]\displaystyle{ G }[/math]. Thus, [math]\displaystyle{ S_i\cap S_j=\emptyset }[/math] for any distinct [math]\displaystyle{ i }[/math] and [math]\displaystyle{ j }[/math].

Therefore, [math]\displaystyle{ \{u,v_1,v_2,\ldots,v_d\}\cup S_1\cup S_2\cup\cdots\cup S_d\subseteq V }[/math] implies that

[math]\displaystyle{ (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]\displaystyle{ \sum_{v:v\sim u}d(v)\le n-1 }[/math].

By Cauchy-Schwarz inequality,

[math]\displaystyle{ 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],

which implies that [math]\displaystyle{ |E|\le\frac{1}{2}n\sqrt{n-1} }[/math].

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

Hamiltonian cycle

We now look at graphs which does not have large cycles. In particular, we consider graphs without Hamiltonian cycles.

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]\displaystyle{ (n-1) }[/math]-clique and the one remaining vertex is connected to the clique by one edge. This graph has [math]\displaystyle{ {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.

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.

Dirac's Theorem
A graph [math]\displaystyle{ G(V,E) }[/math] on [math]\displaystyle{ n }[/math] vertices has a Hamiltonian cycle if [math]\displaystyle{ d_v\ge\frac{n}{2} }[/math] for all [math]\displaystyle{ v\in V }[/math].
Proof.

Suppose to the contrary, the theorem is not true and there exists a non-Hamiltonian graph with [math]\displaystyle{ d_v\ge\frac{n}{2} }[/math] for all [math]\displaystyle{ v\in V }[/math]. Let [math]\displaystyle{ G }[/math] be such a graph with a maximum number of edges. Then adding any edge to [math]\displaystyle{ G }[/math] creates a Hamiltonian cycle. Thus, [math]\displaystyle{ G }[/math] must have a Hamiltonian path, say [math]\displaystyle{ v_1v_2\cdots v_n }[/math].

Consider the sets,

  • [math]\displaystyle{ S=\{i\mid v_iv_n\in E\} }[/math];
  • [math]\displaystyle{ T=\{i\mid v_{i+1}v_1\in E\} }[/math].

Therefore, [math]\displaystyle{ S\subseteq\{v_1,v_2,\ldots,v_{n-1}\} }[/math] contains the neighbors of [math]\displaystyle{ v_n }[/math]; and [math]\displaystyle{ T\subseteq\{v_1,v_2,\ldots,v_{n-1}\} }[/math] contains the predecessors (along the Hamiltonian path) of the neighbors of [math]\displaystyle{ v_1 }[/math]. It holds that [math]\displaystyle{ S,T\subseteq\{v_1,v_2,\ldots,v_{n-1}\} }[/math].

Since [math]\displaystyle{ d_v\ge\frac{n}{2} }[/math] for all [math]\displaystyle{ v\in V }[/math], [math]\displaystyle{ |S|,|T|\ge\frac{n}{2} }[/math]. By the pigeonhole principle, there exists some [math]\displaystyle{ v_i\in S\cap T }[/math]. We can construct the following Hamiltonian cycle:

[math]\displaystyle{ 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]\displaystyle{ G }[/math] is non-Hamiltonian.

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

Erdős–Stone theorem

We introduce a notation for the number of edges in extremal graphs with a specific forbidden substructure.

Definition
Let [math]\displaystyle{ \mathrm{ex}(n,H) }[/math] denote the largest number of edges that a graph [math]\displaystyle{ G\not\supseteq H }[/math] on [math]\displaystyle{ n }[/math] vertices can have.

With this notation, Turán's theorem can be restated as

Turán's theorem (restated)
[math]\displaystyle{ \mathrm{ex}(n,K_r)\le\frac{r-2}{2(r-1)}n^2 }[/math].

Let [math]\displaystyle{ K_s^r=K_{\underbrace{s,s,\cdots,s}_{r}} }[/math] be the complete [math]\displaystyle{ r }[/math]-partite graph with [math]\displaystyle{ s }[/math] vertices in each class, i.e., the Turán graph [math]\displaystyle{ T(rs,r) }[/math]. The Erdős–Stone theorem (also referred as the fundamental theorem of extremal graph theory) gives an asymptotic bound on [math]\displaystyle{ \mathrm{ex}(n,K-s^r) }[/math], i.e., the largest number of edges that an [math]\displaystyle{ n }[/math]-vertex graph can have to not contain [math]\displaystyle{ K_s^r }[/math].

Fundamental theorem of extremal graph theory (Erdős–Stone 1946)
For any integers [math]\displaystyle{ r\ge 2 }[/math] and [math]\displaystyle{ s\ge 1 }[/math], and any [math]\displaystyle{ \epsilon\gt 0 }[/math], if [math]\displaystyle{ n }[/math] is sufficiently large then every graph on [math]\displaystyle{ n }[/math] vertices and with at least [math]\displaystyle{ \left(\frac{r-2}{2(r-1)}+\epsilon\right)n^2 }[/math] edges contains [math]\displaystyle{ K_{r,s} }[/math] as a subgraph, i.e.,
[math]\displaystyle{ \mathrm{ex}(n,K_s^r)= \left(\frac{r-2}{2(r-1)}+o(1)\right)n^2 }[/math].

The theorem is called fundamental because of its single most important corollary: it relate the extremal bound for an arbitrary subgraph [math]\displaystyle{ H }[/math] to a very natural parameter of [math]\displaystyle{ H }[/math], its chromatic number.

Recall that [math]\displaystyle{ \chi(G) }[/math] is the chromatic number of [math]\displaystyle{ G }[/math], the smallest number of colors that one can use to color the vertices so that no adjacent vertices have the same color.

Corollary
For every nonempty graph [math]\displaystyle{ H }[/math],
[math]\displaystyle{ \lim_{n\rightarrow\infty}\frac{\mathrm{ex}(n,H)}{{n\choose 2}}=\frac{\chi(H)-2}{\chi(H)-1} }[/math].
Proof of corollary

Let [math]\displaystyle{ r=\chi(H) }[/math].

Note that [math]\displaystyle{ T(n,r-1) }[/math] can be colored with [math]\displaystyle{ r-1 }[/math] colors, one color for each part. Thus, [math]\displaystyle{ H\not\subseteq T(n,r-1) }[/math], since otherwise [math]\displaystyle{ H }[/math] can also be colored with [math]\displaystyle{ r-1 }[/math] colors, contradicting that [math]\displaystyle{ \chi(H)=1 }[/math]. By definition, [math]\displaystyle{ \mathrm{ex}(n,H) }[/math] is the maximum number of edges that an [math]\displaystyle{ n }[/math]-vertex graph [math]\displaystyle{ G\not\supseteq H }[/math] can have. Thus,

[math]\displaystyle{ |T(n,r-1)|\le\mathrm{ex}(n,H) }[/math].

It is not hard to see that

[math]\displaystyle{ |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]\displaystyle{ H }[/math] with chromatic number [math]\displaystyle{ r }[/math] has that [math]\displaystyle{ H\subseteq K_s^r }[/math] for all sufficiently large [math]\displaystyle{ s }[/math]. We just connect all pairs of vertices from different color classes. Thus,

[math]\displaystyle{ \mathrm{ex}(n,H)\le\mathrm{ex}(n,K_s^r) }[/math].

Due to Erdős–Stone theorem,

[math]\displaystyle{ \mathrm{ex}(n,K_s^r)=\left(\frac{r-2}{2(r-1)}+o(1)\right)n^2 }[/math].

Altogether, we have

[math]\displaystyle{ \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.

[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 4.
  • Aigner and Ziegler. Proofs from THE BOOK, 4th Edition. Springer-Verlag. Chapter 36.
  • Diestel. Graph Theory, 3rd Edition. Springer-Verlag 2000. Chapter 7.