Center (algebra): Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>Redrose64
Undid revision 577388259 by 118.149.240.171 (talk) rv per MOS:ENGVAR
en>Daviddwd
Undid revision 623091755 by Daviddwd (talk)
 
(One intermediate revision by one other user not shown)
Line 1: Line 1:
[[File:Venn-diagram-AB.svg|thumbnail|<center>The union of sets ''A'' and ''B''</center>]]
My name is Roseanna McLaren but everybody calls me Roseanna. I'm from Austria. I'm studying at the high school (final year) and I play the Piano for 4 years. Usually I choose songs from the famous films :). <br>I have two sister. I love Musical instruments, watching TV (The Big Bang Theory) and Leaf collecting and pressing.<br><br>Also visit my website [http://www.happybunnies.co.uk/blog/2014/06/chicken-pox/ How To Get Free Fifa 15 Coins]
In [[combinatorics]] (combinatorial mathematics), the '''inclusion–exclusion principle''' is a counting technique which generalizes the familiar method of obtaining the number of elements in the [[union (set theory)|union]] of two finite [[set (mathematics)|set]]s; symbolically expressed as
:<math> |A \cup B| = |A| + |B| - |A \cap B|, </math>
where ''A'' and ''B'' are two finite sets and |''S''| indicates the [[cardinality]] of a set ''S'' (which may be considered as the number of elements of the set, if the set is [[Finite set|finite]]). The formula expresses the fact that the sum of the sizes of the two sets may be too large since some elements may be counted twice. The double-counted elements are those in the [[intersection (set theory)|intersection]] of the two sets and the count is corrected by subtracting the size of the intersection. The principle is more clearly seen in the case of three sets, which for the sets ''A'', ''B'' and ''C'' is given by
:<math>|A \cup B \cup C| = |A| + |B| + |C| - |A \cap B| - |A \cap C| - |B \cap C| + |A \cap B \cap C|.</math>
This formula can be verified by counting how many times each region in the [[Venn diagram]] figure is included in the right-hand side of the formula. In this case, when removing the contributions of over-counted elements, the number of elements in the mutual intersection of the three sets has been subtracted too often, so must be added back in to get the correct total.
[[Image:Inclusion-exclusion.svg|thumb|Inclusion–exclusion illustrated by a Venn diagram for three sets]]
 
Generalizing the results of these examples gives the principle of inclusion–exclusion: to find the cardinality of the union of {{mvar|n}} sets, i. include the cardinalities of the sets, ii. exclude the cardinalities of the pairwise intersections, iii. include the cardinalities of the triple-wise intersections, iv. exclude the cardinalities of the quadruple-wise intersections, v. include the cardinalities of the quintuple-wise intersections, vi. and continue, until the cardinality of the {{mvar|n}}-tuple-wise intersection is included (if {{mvar|n}} is odd) or excluded ({{mvar|n}} even).
 
The name comes from the idea that the principle is based on over-generous ''inclusion'', followed by compensating ''exclusion''.
This concept is attributed to [[Abraham de Moivre]] (1718);<ref name="Roberts 2009 loc=pg. 405">{{harvnb|Roberts|Tesman|2009|loc=pg. 405}}</ref> but it first appears in a paper of [[Daniel da Silva (mathematician)|Daniel da Silva]] (1854),<ref>{{harvnb|Mazur|2010|loc=pg. 94}}</ref> and later in a paper by [[J. J. Sylvester]] (1883).<ref>{{harvnb|van Lint|Wilson|1992|loc=pg. 77}}</ref> Sometimes the principle is referred to as the formula of Da Silva, or Sylvester due to these publications. The principle is an example of the [[Sieve theory|sieve method]] extensively used in [[number theory]] and is sometimes referred to as the ''sieve formula'',<ref>{{harvnb|van Lint|Wilson|1992|loc = pg. 77}}</ref> though Legendre already used a similar device in a sieve context in 1808.
 
As finite probabilities are computed as counts relative to the cardinality of the probability space, the formulas for the principle of inclusion–exclusion remain valid when the cardinalities of the sets are replaced by finite probabilites. More generally, both versions of the principle can be put under the common umbrella of [[measure theory]].
 
In a very abstract setting, the principle of inclusion–exclusion amounts to no more than the calculation of the inverse of a certain matrix.<ref>{{harvnb|Stanley|1986|loc=pg. 64}}</ref> From this point of view, there is nothing mathematically interesting about the principle. However, the wide applicability of the principle makes it an extremely valuable technique in combinatorics and related areas of mathematics. As [[Gian-Carlo Rota]] put it:<ref>{{citation|last=Rota|first=Gian-Carlo|title=On the foundations of combinatoial theory I. Theory of Möbius functions|journal=Zeitschrift fur Wahrscheinlichkeitstheorie|volume=2|year=1964|pages=340–368}}</ref>
<blockquote>"One of the most useful principles of enumeration in discrete probability and combinatorial theory is the celebrated principle of inclusion–exclusion. When skillfully applied, this principle has yielded the solution to many a combinatorial problem."</blockquote>
 
==Statement==
[[Image:inclusion-exclusion-3sets.png|thumb|Each term of the inclusion–exclusion formula gradually corrects the count until finally each portion of the [[Venn diagram]] is counted exactly once.]]
 
In its general form, the principle of inclusion–exclusion states that for finite sets ''A''<sub>1</sub>, ..., ''A<sub>n</sub>'', one has the identity
:<math>
\biggl|\bigcup_{i=1}^n A_i\biggr| = \sum_{i=1}^n\left|A_i\right|\;
-\sum_{1 \le i < j \le n}\left|A_i\cap A_j\right|\;
+ \sum_{1 \le i < j < k \le n}\left|A_i\cap A_j\cap A_k\right|\;-\ \ldots\ +\; \left(-1\right)^{n-1} \left|A_1\cap\cdots\cap A_n\right|.
</math>
This can be compactly written as
:::<math>
\begin{align}
\biggl|\bigcup_{i=1}^n A_i\biggr| &= \sum_{k = 1}^{n} (-1)^{k+1} \left( \sum_{1 \leq i_{1} < \cdots < i_{k} \leq n} \left| A_{i_{1}} \cap \cdots \cap A_{i_{k}} \right| \right).
\end{align}</math>
 
In words, to count the number of elements in a finite union of finite sets, first sum the cardinalities of the individual sets, then subtract the number of elements which appear in more than one set, then add back the number of elements which appear in more than two sets, then subtract the number of elements which appear in more than three sets, and so on. This process naturally ends since there can be no elements which appear in more than the number of sets in the union.
 
In applications it is common to see the principle expressed in its complementary form. That is, letting ''S'' be a finite [[universal set]] containing all of the ''A''<sub>i</sub> and letting <math>\bar{A_i}</math> denote the complement of ''A''<sub>i</sub> in ''S'', by [[De Morgan's laws]] we have
:<math>
\biggl|\bigcap_{i=1}^n \bar{A_i}\biggr| = \biggl|S - \bigcup_{i=1}^n A_i\biggr| = \left| S \right|\; - \sum_{i=1}^n\left|A_i\right|\;
+\sum_{1 \le i < j \le n}\left|A_i\cap A_j\right|\;
-\;  \ldots\ +\; \left(-1\right)^{n} \left|A_1\cap\cdots\cap A_n\right|.
</math>
 
As another variant of the statement, let ''P''<sub>1</sub>, ''P''<sub>2</sub>, ..., ''P''<sub>n</sub> be a list of properties that elements of a set ''S'' may or may not have, then the principle of inclusion–exclusion provides a way to calculate the number of elements of ''S'' which have none of the properties. Just let ''A''<sub>i</sub> be the subset of elements of ''S'' which have the property ''P''<sub>i</sub> and use the principle in its complementary form. This variant is due to [[J.J. Sylvester]].<ref name="Roberts 2009 loc=pg. 405"/>
 
==Examples==
As a simple example of the use of the principle of inclusion–exclusion, consider the question:<ref>{{harvnb|Mazur|2010|loc=pp. 83–4, 88}}</ref>
::How many integers in {1,...,100} are not divisible by 2, 3 or 5?
Let ''S'' = {1,...,100} and ''P''<sub>1</sub> the property that an integer is divisible by 2, ''P''<sub>2</sub> the property that an integer is divisible by 3 and ''P''<sub>3</sub> the property that an integer is divisible by 5. Letting ''A''<sub>i</sub> be the subset of ''S'' whose elements have property ''P''<sub>i</sub> we have by elementary counting: |''A''<sub>1</sub>| = 50, |''A''<sub>2</sub>| = 33, and |''A''<sub>3</sub>| = 20. There are 16 of these integers divisible by 6, 10 divisible by 10 and 6 divisible by 15. Finally, there are just 3 integers divisible by 30, so the number of integers not divisible by any of 2, 3 or 5 is given by:
:::  100 − (50 + 33 + 20) + (16 + 10 + 6) − 3 = 26.
 
A more complex example is the following.
Suppose there is a deck of ''n'' cards, each card is numbered from 1&nbsp;to&nbsp;''n''. Suppose a card numbered ''m'' is in the correct position if it is the ''m''th card in the deck. How many ways, ''W'', can the cards be shuffled with at least 1 card being in the correct position?
 
Begin by defining set ''A''<sub>''m''</sub>, which is all of the orderings of cards with the ''m''th card correct.  Then the number of orders, ''W'', with ''at least'' one card being in the correct position, ''m'', is
 
: <math>W = \biggl|\bigcup_{m=1}^nA_m\biggr|.</math>
 
Apply the principle of inclusion–exclusion,
 
: <math>
\begin{align}
W & = \sum_{m_1=1}^n \left| A_{m_1} \right|  \\
  & {}- \sum_{1 \le m_1 < m_2 \le n} \left|A_{m_1} \cap A_{m_2} \right|  \\
  & {}+ \sum_{1 \le m_1 < m_2 < m_3 \le n} \left|A_{m_1} \cap A_{m_2} \cap A_{m_3} \right|  \\
  & {}- \cdots  \\
  & {}+ (-1)^{p-1} \sum_{1 \le m_1 < \cdots < m_p \le n} \left|A_{m_1} \cap \cdots \cap A_{m_p} \right|  \\
  & \cdots. \\
\end{align}
</math>
 
Each value <math>A_{m_1} \cap \cdots \cap A_{m_p}</math> represents the set of shuffles having ''p'' values ''m''<sub>1</sub>,&nbsp;...,&nbsp;''m''<sub>''p''</sub> in the correct position.  Note that the number of shuffles with ''p'' values correct only depends on ''p'', not on the particular values of <math>m</math>.  For example, the number of shuffles having the 1st, 3rd, and 17th cards in the correct position is the same as the number of shuffles having the 2nd, 5th, and 13th cards in the correct positions.  It only matters that of the n cards, 3 were chosen to be in the correct position.  Thus there are <math>{n \choose p}</math> terms in each summation (see [[combination]]).
 
: <math>
W  = {n \choose 1} \left|A_1 \right|  - {n \choose 2} \left|A_1 \cap A_2 \right|  + {n \choose 3} \left|A_1 \cap A_2 \cap A_3 \right|  - \cdots  + (-1)^{p-1} {n \choose p} \left|A_1 \cap \cdots \cap A_p \right|  \cdots. </math>
 
<math>\left|A_1 \cap \cdots \cap A_p \right|</math> is the number of orderings having ''p'' elements in the correct position, which is equal to the number of ways of ordering the remaining ''n''&nbsp;&minus;&nbsp;''p'' elements, or (''n''&nbsp;&minus;&nbsp;''p'')<nowiki>!</nowiki>.  Thus we finally get:
 
: <math>
\begin{align}
W & = {n \choose 1} (n-1)!  - {n \choose 2} (n-2)!  + {n \choose 3} (n-3)!  - \cdots  + (-1)^{p-1} {n \choose p} (n-p)!  \cdots \\
W & = \sum_{p=1}^n (-1)^{p-1} {n \choose p} (n-p)!. \\
\end{align}
</math>
 
Noting that <math>{n \choose p} = \frac{n!}{p!(n-p)!}</math>, this reduces to
 
: <math>
W  = \sum_{p=1}^n (-1)^{p-1}\, \frac{n!}{p!}.
</math>
 
A permutation where ''no'' card is in the correct position is called a [[derangement]]. Taking ''n''<nowiki>!</nowiki> to be the total number of permutations, the probability ''Q'' that a random shuffle produces a derangement is given by
 
: <math>Q = 1 - \frac{W}{n!} = \sum_{p=0}^n \frac{(-1)^p}{p!}, </math>
 
a truncation to n+1 terms of the [[Taylor series|Taylor expansion]] of ''e''<sup>&minus;1</sup>.  Thus the probability of guessing an order for a shuffled deck of cards and being incorrect about every card is approximately 1/''e'' or 37%.
 
==A special case==
 
The situation that appears in the derangement example above occurs often enough to merit special attention.<ref>{{harvnb|Brualdi|2010|loc=pp. 167–8}}</ref> Namely, when the size of the intersection sets appearing in the formulas for the principle of inclusion–exclusion depend only on the number of sets in the intersections and not on which sets appear. More formally, if the intersection
 
:<math>A_J:=\bigcap_{j\in J} A_j</math>
 
has the same cardinality, say ''α<sub>k</sub>'' = |''A<sub>J</sub>''|, for every ''k''-element subset ''J'' of {1,&nbsp;...,&nbsp;''n''}, then
 
:<math>\biggl|\bigcup_{i=1}^n A_i\biggr|  =\sum_{k=1}^n (-1)^{k-1}\binom nk \alpha_k.</math>
 
Or, in the complementary form, where the universal set ''S'' has cardinality α<sub>0</sub>,
 
:<math>\biggl|S \setminus \bigcup_{i=1}^n A_i\biggr|  =\sum_{k=0}^n (-1)^{k}\binom nk \alpha_k.</math>
 
==A generalization==
 
Given a [[family of sets|family (repeats allowed) of subsets]] ''A''<sub>1</sub>, ''A''<sub>2</sub>, ..., ''A''<sub>n</sub> of a universal set ''S'', the principle of inclusion–exclusion calculates the number of elements of ''S'' in none of these subsets. A generalization of this concept would calculate the number of elements of ''S'' which appear in exactly some fixed ''m'' of these sets.
 
Let ''N'' = <nowiki>[</nowiki>''n''<nowiki>]</nowiki> = {1,2,...,''n''}. If we define <math>A_{\emptyset} = S</math>, then the principle of inclusion–exclusion can be written as, using the notation of the previous section; the number of elements of ''S'' contained in none of the ''A''<sub>i</sub> is:
::<math> \sum_{J \subseteq [n]} (-1)^{|J|} |A_J|.</math>
 
If ''I'' is a fixed subset of the index set ''N'', then the number of elements which belong to ''A''<sub>i</sub> for all ''i'' in ''I'' and for no other values is:<ref>{{harvnb|Cameron|1994|loc=pg. 78}}</ref>
:<math> \sum_{I \subseteq J} (-1)^{|J| - |I|} |A_J|.</math>
Define the sets
:<math>B_k = A_{I \cup \{ k \}}</math> for k in <math>N \setminus I</math>.
We seek the number of elements in none of the ''B''<sub>k</sub> which, by the principle of inclusion–exclusion (with <math>B_{\emptyset} = A_I</math>), is
:<math>\sum_{K \subseteq N \setminus I} (-1)^{|K|}|B_K|.</math>
The correspondence ''K'' ↔ ''J'' = ''I'' ∪ ''K'' between subsets of ''N'' \ ''I'' and subsets of ''N'' containing ''I'' is a bijection and if ''J'' and ''K'' correspond under this map then ''B''<sub>K</sub> = ''A''<sub>J</sub>, showing that the result is valid.
 
==In probability==
 
In [[probability]], for events ''A''<sub>1</sub>, ..., ''A''<sub>''n''</sub> in a [[probability space]] <math>\scriptstyle(\Omega,\mathcal{F},\mathbb{P})</math>, the inclusion–exclusion principle becomes for ''n''&nbsp;=&nbsp;2
 
:<math>\mathbb{P}(A_1\cup A_2)=\mathbb{P}(A_1)+\mathbb{P}(A_2)-\mathbb{P}(A_1\cap A_2),</math>
 
for ''n''&nbsp;=&nbsp;3
 
:<math>\begin{align}\mathbb{P}(A_1\cup A_2\cup A_3)&=\mathbb{P}(A_1)+\mathbb{P}(A_2)+\mathbb{P}(A_3)\\
&\qquad{}-\mathbb{P}(A_1\cap A_2)-\mathbb{P}(A_1\cap A_3)-\mathbb{P}(A_2\cap A_3)\\
&\qquad{}+\mathbb{P}(A_1\cap A_2\cap A_3)
\end{align}</math>
 
and in general
 
:<math>\begin{align}
\mathbb{P}\biggl(\bigcup_{i=1}^n A_i\biggr) & {} =\sum_{i=1}^n \mathbb{P}(A_i)
-\sum_{i<j}\mathbb{P}(A_i\cap A_j) \\
&\qquad+\sum_{i<j<k}\mathbb{P}(A_i\cap A_j\cap A_k)-\ \cdots\ +(-1)^{n-1}\, \mathbb{P}\biggl(\bigcap_{i=1}^n A_i\biggr),
\end{align}</math>
 
which can be written in closed form as
 
:<math>\mathbb{P}\biggl(\bigcup_{i=1}^n A_i\biggr)  =\sum_{k=1}^n \left((-1)^{k-1}\sum_{\scriptstyle I\subset\{1,\ldots,n\}\atop\scriptstyle|I|=k} \mathbb{P}(A_I)\right),</math>
 
where the last sum runs over all subsets ''I'' of the indices 1, ..., ''n'' which contain exactly ''k'' elements, and
 
:<math>A_I:=\bigcap_{i\in I} A_i</math>
 
denotes the intersection of all those ''A<sub>i</sub>'' with index in ''I''.
 
According to the [[Boole's inequality|Bonferroni inequalities]], the sum of the first terms in the formula is alternately an upper bound and a lower bound for the [[Sides of an equation|LHS]]. This can be used in cases where the full formula is too cumbersome.
 
For a general [[measure space]] (''S'',''Σ'',''μ'') and [[measurable]] subsets ''A''<sub>1</sub>, ..., ''A<sub>n''</sub> of finite measure, the above identities also hold when the probability measure <math>\mathbb{P}</math> is replaced by the measure ''μ''.
 
===Special case===
 
If, in the probabilistic version of the inclusion–exclusion principle, the probability of the intersection ''A<sub>I</sub>'' only depends on the cardinality of ''I'', meaning that for every ''k'' in {1,&nbsp;...,&nbsp;''n''} there is an ''a<sub>k</sub>'' such that
 
:<math>a_k=\mathbb{P}(A_I)\quad\text{for every}\quad I\subset\{1,\ldots,n\}\quad\text{with}\quad |I|=k,</math>
 
then the above formula simplifies to
 
:<math>\mathbb{P}\biggl(\bigcup_{i=1}^n A_i\biggr)  =\sum_{k=1}^n (-1)^{k-1}\binom nk a_k</math>
 
due to the combinatorial interpretation of the [[binomial coefficient]] <math>\scriptstyle\binom nk</math>.
 
An analogous simplification is possible in the case of a general measure space (''S'',''Σ'',''μ'') and measurable subsets ''A''<sub>1</sub>, ..., ''A<sub>n''</sub> of finite measure.
 
==Other forms==
 
The principle is sometimes stated in the form<ref>{{harvnb|Graham|Grotschel|Lovasz|1995|loc=pg. 1049}}</ref> that says that if
 
:<math>g(A)=\sum_{S\subseteq A}f(S)</math>
 
then
 
:<math>f(A)=\sum_{S\subseteq A}(-1)^{\left|A\right|-\left|S\right|}g(S)\qquad(**)</math>
 
::We show now that the combinatorial and the probabilistic version of the inclusion–exclusion principle are instances of (**). Take <math>\underline{m} = \{1,2,\ldots,m\}</math>, <math>f(\underline{m}) = 0</math>, and
 
:::<math>f(S)=\bigg|\bigcap_{i \in \underline{m} \backslash S} A_i \bigg\backslash \bigcup_{i \in S} A_i\bigg| \qquad\text{and}\qquad f(S)=\mathbb{P}\bigg(\bigcap_{i \in \underline{m} \backslash S} A_i \bigg\backslash \bigcup_{i \in S} A_i\bigg)</math>
 
::respectively for all [[set (mathematics)|sets]] <math>S</math> with <math>S \subsetneq \underline{m}</math>. Then we obtain
 
:::<math>g(A)=\bigg|\bigcap_{i \in \underline{m} \backslash A} A_i\bigg|,~~ g(\underline{m}) = \bigg|\bigcup_{i \in \underline{m}} A_i \bigg| \qquad\text{and}\qquad g(A)=\mathbb{P}\bigg(\bigcap_{i \in \underline{m} \backslash A} A_i\bigg),~~ g(\underline{m}) = \mathbb{P}\bigg(\bigcup_{i \in \underline{m}} A_i\bigg)</math>
 
::respectively for all sets <math>A</math> with <math>A \subsetneq \underline{m}</math>. This is  because [[element (mathematics)|elements]] <math>a</math> of <math>\cap_{i \in \underline{m} \backslash A} A_i</math> can be [[element (mathematics)#notation|contained]] in other <math>A_i</math>'s (<math>A_i</math>'s with <math>i \in A</math>) as well, and the <math>\cap\backslash\!\!\cup\!\text{-}</math>formula runs exactly through all possible extensions of the sets <math>\{A_i \mid i \in \underline{m} \backslash A\}</math> with other <math>A_i</math>'s, counting <math>a</math> only for the set that matches the membership behavior of  <math>a</math>, if <math>S</math> runs through all [[subset]]s of <math>A</math> (as in the definition of <math>g(A)</math>).
 
::Since <math>f(\underline{m}) = 0</math>, we obtain from (**) with <math>A = \underline{m}</math> that
 
:::<math>\sum_{\underline{m} \supseteq T \supsetneq \varnothing}(-1)^{\left|T\right|-1} g(\underline{m} \backslash T) = \sum_{\varnothing \subseteq S \subsetneq \underline{m}}(-1)^{m-\left|S\right|-1}g(S) = g(\underline{m})</math>
 
::and by interchanging sides, the combinatorial and the probabilistic version of the inclusion–exclusion principle follow.
 
If one sees a number <math>n</math> as a set of its prime factors, then (**) is a generalization of [[Möbius inversion formula]] for [[square-free integer|square-free]] [[natural number]]s. Therefore, (**) is seen as the Möbius inversion formula for the [[incidence algebra]] of the [[partially ordered set]] of all subsets of ''A''.
 
For a generalization of the full version of Möbius inversion formula, (**) must be generalized to [[multiset]]s. For multisets instead of sets, (**) becomes
 
:<math>f(A)=\sum_{S\subseteq A}\mu(A - S)g(S)\qquad(***)</math>
 
where <math>A - S</math> is the multiset for which <math>(A - S) \uplus S = A</math>, and
 
* ''μ''(''S'') =  1 if ''S'' is a set (i.e. a multiset without double elements) of [[even and odd numbers|even]] [[cardinality]].
* ''μ''(''S'') =  &minus;1 if ''S'' is a set (i.e. a multiset without double elements) of odd cardinality.
* ''μ''(''S'') =  0 if ''S'' is a proper multiset (i.e. ''S'' has double elements).
 
Notice that ''<math>\mu</math>''<math>(A - S)</math> is just the <math>(-1)^{\left|A\right|-\left|S\right|}</math> of (**) in case <math>A - S</math> is a set.
 
'''Proof of (***):''' Substitute
 
:<math>g(S)=\sum_{T\subseteq S}f(T)</math>
 
on the right hand side of (***). Notice that <math>f(A)</math> appears once on both sides of (***). So we must show that for all <math>T</math> with <math>T\subsetneq A</math>, the terms <math>f(T)</math> cancel out on the right hand side of (***). For that purpose, take a fixed <math>T</math> such that <math>T\subsetneq A</math> and take an arbitrary fixed <math>a \in A</math> such that <math>a \not\in T</math>.
 
Notice that <math>A - S</math> must be a set for each [[Positive number|positive]] or [[negative number|negative]] appearance of <math>f(T)</math> on the right hand side of (***) that is obtained by way of the multiset <math>S</math> such that <math>T \subseteq S \subseteq A</math>. Now each appearance of <math>f(T)</math> on the right hand side of (***) that is obtained by way of <math>S</math> such that <math>A - S</math> is a set that contains <math>a</math> cancels out with the one that is obtained by way of the corresponding <math>S</math> such that <math>A - S</math> is a set that does not contain <math>a</math>. This gives the desired result.
 
==Applications==
 
The inclusion–exclusion principle is widely used and only a few of its applications can be mentioned here.
 
===Counting derangements===
 
{{main|Derangement}}
 
A well-known application of the inclusion–exclusion principle is to the combinatorial problem of counting all [[derangement]]s of a finite set.  A ''derangement'' of a set ''A'' is a [[bijection]] from ''A'' into itself that has no fixed points. Via the inclusion–exclusion principle one can show that if the cardinality of ''A'' is ''n'', then the number of derangements is [''n''!&nbsp;/&nbsp;''e''] where [''x''] denotes the [[nearest integer function|nearest integer]] to ''x''; a detailed proof is available [[Random permutation statistics#Number of permutations that are derangements|here]] and also see [[#Examples|the examples section]] above.
 
The first occurrence of the problem of counting the number of derangements is in an early book on games of chance: ''Essai d'analyse sur les jeux de hazard'' by P. R. de Montmort (1678 &ndash; 1719) and was known as either "Montmort's problem" or by the name he gave it, "''problème des rencontres''."<ref>{{harvnb|van Lint|Wilson|1992|loc=pp. 77-8}}</ref> The problem is also known as the ''hatcheck problem.''
 
The number of derangements is also known as the [[subfactorial]] of ''n'', written !''n''. It follows that if all bijections are assigned the same probability then the probability that a random bijection is a derangement quickly approaches 1/''e'' as ''n'' grows.
 
===Counting intersections===
The principle of inclusion–exclusion, combined with [[De Morgan's law]], can be used to count the cardinality of the intersection of sets as well. Let <math>\scriptstyle\overline{A}_k</math> represent the complement of ''A''<sub>''k''</sub> with respect to some universal set ''A'' such that <math>\scriptstyle A_k\, \subseteq\, A</math> for each ''k''. Then we have
 
:<math>
\bigcap_{i=1}^n A_i = \overline{\bigcup_{i=1}^n \overline{A}_i}
</math>
 
thereby turning the problem of finding an intersection into the problem of finding a union.
 
===Graph coloring===
The inclusion exclusion principle forms the basis of algorithms for a number of NP-hard graph partitioning problems, such as graph coloring.<ref name="bhk">{{harvnb|Björklund|Husfeldt|Koivisto|2009}}</ref>
 
A well known application of the principle is the construction of the [[chromatic polynomial]] of a graph.<ref>{{harvnb|Gross|2008|loc=pp. 211–13}}</ref>
 
===Bipartite graph perfect matchings===
The number of [[perfect matching]]s of a [[bipartite graph]] can be calculated using the principle.<ref>{{harvnb|Gross|2008|loc=pp. 208–10}}</ref>
 
===Number of onto functions===
Given finite sets ''A'' and ''B'', how many [[surjective function]]s (onto functions) are there from ''A'' to ''B''? Without any loss of generality we may take ''A'' = {1,2,...,''k''} and ''B'' = {1,2,...,''n''}, since only the cardinalities of the sets matter. By using ''S'' as the set of all [[Function (mathematics)|functions]] from ''A'' to ''B'', and defining, for each ''i'' in ''B'', the property ''P''<sub>i</sub> as "the function misses the element ''i'' in ''B''" (''i'' is not in the [[Image (mathematics)|image]] of the function), the principle of inclusion&ndash;exclusion gives the number of onto functions between ''A'' and ''B'' as:<ref>{{harvnb|Mazur|2008|loc=pp.84-5, 90}}</ref>
::<math>\sum_{j=0}^{n} \binom{n}{j} (-1)^j (n-j)^k.</math>
 
===Permutations with forbidden positions===
A [[permutation]] of the set ''S'' = {1,2,...,n} where each element of ''S'' is restricted to not being in certain positions (here the permutation is considered as an ordering of the elements of ''S'') is called a ''permutation with forbidden positions''. For example, with ''S'' = {1,2,3,4}, the permutations with the restriction that the element 1 can not be in positions 1 or 3, and the element 2 can not be in position 4 are: 2134, 2143, 3124, 4123, 2341, 2431, 3241, 3421, 4231 and 4321. By letting ''A''<sub>i</sub> be the set of positions that the element ''i'' is not allowed to be in, and the property ''P''<sub>i</sub> to be the property that a permutation puts element ''i'' into a position in ''A''<sub>i</sub>, the principle of inclusion–exclusion can be used to count the number of permutations which satisfy all the restrictions.<ref>{{harvnb|Brualdi|2010|loc=pp. 177&ndash;81}}</ref>
 
In the given example, there are 12 = 2(3!) permutations with property ''P''<sub>1</sub>, 6 = 3! permutations with property ''P''<sub>2</sub> and no permutations have properties ''P''<sub>3</sub> or ''P''<sub>4</sub> as there are no restrictions for these two elements. The number of permutations satisfying the restrictions is thus:
::4! - (12 + 6 + 0 + 0) + (4) = 24 - 18 + 4 = 10.
The final 4 in this computation is the number of permutations having both properties ''P''<sub>1</sub> and ''P''<sub>2</sub>. There are no other non-zero contributions to the formula.
 
===Stirling numbers of the second kind===
{{main|Stirling numbers of the second kind}}
 
The [[Stirling numbers of the second kind]], ''S''(''n'',''k'') count the number of [[Partition of a set|partitions]] of a set of ''n'' elements into ''k'' non-empty subsets (indistinguishable ''boxes''). An explicit formula for them can be obtained by applying the principle of inclusion–exclusion to a very closely related problem, namely, counting the number of partitions of an ''n''-set into ''k'' non-empty but distinguishable boxes ([[Ordered set|ordered]] non-empty subsets). Using the universal set consisting of all partitions of the ''n''-set into ''k'' (possibly empty) distinguishable boxes, ''A''<sub>1</sub>, ''A''<sub>2</sub>, ..., ''A''<sub>k</sub>, and the properties ''P''<sub>i</sub> meaning that the partition has box ''A''<sub>i</sub> empty, the principle of inclusion–exclusion gives an answer for the related result. Dividing by ''k''! to remove the artificial ordering gives the Stirling number of the second kind:<ref>{{harvnb|Brualdi|2010|loc=pp. 282&ndash;7}}</ref>
::<math>S(n,k) = \frac{1}{k!}\sum_{t=0}^{k} (-1)^t \binom{k}{t} (k-t)^n.</math>
 
===Rook polynomials===
{{main|Rook polynomial}}
 
A rook polynomial is the [[generating polynomial|generating function]] of the number of ways to place non-attacking [[rook (chess)|rooks]] on a ''board B'' that looks like a subset of the squares of a [[checkerboard]]; that is, no two rooks may be in the same row or column.  The board ''B'' is any subset of the squares of a rectangular board with ''n'' rows and ''m'' columns; we think of it as the squares in which one is allowed to put a rook. The [[coefficient]], ''r''<sub>''k''</sub>(''B'') of ''x''<sup>''k''</sup> in the rook polynomial ''R''<sub>''B''</sub>(''x'') is the number of ways ''k'' rooks, none of which attacks another, can be arranged in the squares of ''B''. For any board ''B'', there is a complementary board ''B' '' consisting of the squares of the rectangular board that are not in ''B''. This complementary board also has a rook polynomial  ''R''<sub>''B' ''</sub>(''x'') with coefficients ''r''<sub>''k''</sub>(''B''').
 
It is sometimes convenient to be able to calculate the highest coefficient of a rook polynomial in terms of the coefficients of the rook polynomial of the complementary board. Without loss of generality we can assume that ''n'' ≤ ''m'', so this coefficient is ''r''<sub>''n''</sub>(''B''). The number of ways to place ''n'' non-attacking rooks on the complete ''n'' × ''m'' "checkerboard" (without regard as to whether the rooks are placed in the squares of the board ''B'') is given by the [[falling factorial]]:
::<math>(m)_n = m(m-1)(m-2) \cdots (m-n+1).</math>
Letting ''P''<sub>i</sub> be the property that an assignment of ''n'' non-attacking rooks on the complete board has a rook in column ''i'' which is not in a square of the board ''B'', then by the principle of inclusion–exclusion we have:<ref>{{harvnb|Roberts|Tesman|2009|loc=pp.419&ndash;20}}</ref>
::<math> r_n(B) = \sum_{t=0}^n (-1)^t (m-t)_{n-t}\; r_t(B'). </math>
 
===Euler's phi function===
{{main|Euler's totient function}}
 
Euler's totient or phi function, φ(''n'') is an [[arithmetic function]] that counts the number of positive integers less than or equal to ''n'' that are [[relatively prime]] to ''n''. That is, if ''n'' is a [[positive integer]], then φ(''n'') is the number of integers ''k'' in the range 1 ≤  ''k'' ≤  ''n'' which have no common factor with ''n'' other than 1. The principle of inclusion–exclusion is used to obtain a formula for φ(''n''). Let ''S'' be the set {1,2,...,''n''} and define the property ''P''<sub>i</sub> to be that a number in ''S'' is divisible by the prime number ''p''<sub>i</sub>, for 1 ≤ ''i'' ≤ ''r'', where the [[prime factorization]] of
:<math>n = p_1^{a_1} p_2^{a_2} \cdots p_r^{a_r}.</math>
Then,<ref>{{harvnb|van Lint|Wilson|1992|loc=pg. 73}}</ref>
::<math>\phi(n) = n - \sum_{i=1}^r \frac{n}{p_i} + \sum_{1 \leq i < j \leq r} \frac{n}{p_i p_j} - \cdots = n \prod_{i=1}^r \left (1 - \frac{1}{p_i} \right ).</math>
 
==Diluted inclusion–exclusion principle==
{{See also|Bonferroni inequalities}}
 
In many cases where the principle could give an exact formula (in particular, counting [[prime number]]s using the [[sieve of Eratosthenes]]), the formula arising doesn't offer useful content because the number of terms in it is excessive. If each term individually can be estimated accurately, the accumulation of errors may imply that the inclusion–exclusion formula isn't directly applicable. In [[number theory]], this difficulty was addressed by [[Viggo Brun]]. After a slow start, his ideas were taken up by others, and a large variety of [[sieve theory|sieve methods]] developed. These for example may try to find upper bounds for the "sieved" sets, rather than an exact formula.
 
Let ''A''<sub>1</sub>, ..., ''A<sub>n</sub>'' be arbitrary sets and ''p''<sub>1</sub>, ..., ''p<sub>n</sub>'' real numbers in the closed unit interval <nowiki>[</nowiki>0,1<nowiki>]</nowiki>. Then, for every even number ''k'' in {0, ..., ''n''}, the [[indicator function]]s satisfy the inequality:<ref>{{harv|Fernández|Fröhlich|Alan D.|1992|loc=Proposition&nbsp;12.6}}</ref>
 
:<math>1_{A_1\cup\cdots\cup A_n}\ge\sum_{j=1}^k (-1)^{j-1}\sum_{1\le i_1<\cdots<i_j\le n} p_{i_1}\dots p_{i_j}\,1_{A_{i_1}\cap\cdots\cap A_{i_j}}.</math>
 
==Proof==
Let ''A'' denote the union <math>\scriptstyle \cup_{i=1}^n A_i</math> of the sets ''A''<sub>1</sub>, ..., ''A<sub>n</sub>''.  To prove the inclusion–exclusion principle in general, we first have to verify the identity
 
:<math>1_A  =\sum_{k=1}^n (-1)^{k-1}\sum_{\scriptstyle I\subset\{1,\ldots,n\}\atop\scriptstyle|I|=k} 1_{A_I}\qquad(*)</math>
 
for [[indicator function]]s, where
:<math>A_I = \bigcap_{i\in I} A_i.</math>
There are at least two ways to do this:
 
'''First possibility:''' It suffices to do this for every ''x'' in the union of ''A''<sub>1</sub>, ..., ''A<sub>n''</sub>. Suppose ''x'' belongs to exactly ''m'' sets with 1&nbsp;≤&nbsp;''m''&nbsp;≤&nbsp;''n'', for simplicity of notation say ''A''<sub>1</sub>, ..., ''A<sub>m''</sub>. Then the identity at ''x'' reduces to
 
:<math>1 =\sum_{k=1}^m (-1)^{k-1}\sum_{\scriptstyle I\subset\{1,\ldots,m\}\atop\scriptstyle|I|=k} 1.</math>
 
The number of subsets of cardinality ''k'' of an ''m''-element set is the combinatorical interpretation of the [[binomial coefficient]] <math>\textstyle\binom mk</math>&nbsp;. Since <math>\textstyle1=\binom m0</math>&nbsp;, we have
 
:<math>\binom m0 =\sum_{k=1}^m (-1)^{k-1}\binom mk.</math>
 
Putting all terms to the left-hand side of the equation, we obtain the expansion for (1&nbsp;–&nbsp;1)<sup>''m''</sup> given by the [[binomial theorem]], hence we see that (*) is true for ''x''.
 
'''Second possibility:'''  The following function is identically zero
 
:<math>(1_A-1_{A_1})(1_A-1_{A_2})\cdots(1_A-1_{A_n})\,=\,0,</math>
 
because: if ''x'' is not in ''A'', then all factors are 0&nbsp;&minus;&nbsp;0 = 0; and otherwise, if ''x'' does belong to some  ''A<sub>m</sub>'', then the corresponding ''m''th factor is&nbsp;1&nbsp;&minus;&nbsp;1&nbsp;=&nbsp;0. By expanding the product on the left-hand side, equation (*) follows.
 
'''Use of (*):''' To prove the inclusion–exclusion principle for the cardinality of sets, sum the equation (*) over all ''x'' in the union of ''A''<sub>1</sub>, ..., ''A''<sub>''n''</sub>. To derive the version used in probability, take the [[expected value|expectation]] in (*). In general, [[Lebesgue integral|integrate]] the equation (*) with respect to ''μ''. Always use linearity.
 
===Alternative proof===
 
Pick an element contained in the union of all sets and let <math>A_1, A_2, \dots, A_t</math> be the individual sets containing it. (Note that ''t'' &gt; 0.) Since the element is counted precisely once by the left-hand side of the equation, we need to show that it is counted precisely once by the right-hand side. By the binomial theorem,
 
:<math> (1-1)^t= \binom{t}{0} - \binom{t}{1} + \binom{t}{2} - \cdots + (-1)^{t}\binom{t}{t}</math>.
 
Using the fact that <math>\binom{t}{0} = 1</math> and rearranging terms, we have
 
: <math>
\begin{align}
1 & = \binom{t}{1} - \binom{t}{2} + \cdots + (-1)^{t+1}\binom{t}{t}\\
  & = |\{A_i \mid 1 \leq i \leq t\}| - |\{A_i \cap A_j \mid 1 \leq i < j \leq t\}| + \cdots + (-1)^{t+1}|\{A_1 \cap A_2 \cap \cdots  \cap A_t\}|,
\end{align}
</math>
and so the chosen element is indeed counted only once by the right-hand side of the proposed equation.
 
==See also==
* [[Combinatorial principles]]
* [[Boole's inequality]]
* [[Necklace problem]]
* [[Schuette–Nesbitt formula]]
* [[Maximum-minimums identity]]
 
==Notes==
{{Reflist|2}}
 
==References==
* {{Citation
| last1 = Allenby  | first1 = R.B.J.T.
| last2 = Slomson | first2 = Alan
| title = How to Count: An Introduction to Combinatorics
| publisher = CRC Press
| year = 2010
| series = Discrete Mathematics and Its Applications
| pages = 51–60
| isbn = 9781420082609
| url = http://www.crcpress.com/product/isbn/9781420082609
| edition = 2
}}
* {{Citation
| last1=Björklund | first1=A.
| last2=Husfeldt | first2=T.
| last3=Koivisto | first3=M.
| journal=[[SIAM Journal on Computing]]
| pages=546–563
| title=Set partitioning via inclusion–exclusion
| volume= 39
| year=2009
| doi=10.1137/070683933| issue=2
}}
* {{Citation|last=Brualdi|first=Richard A.|title=Introductory Combinatorics|edition=5th|publisher=Prentice–Hall|year=2010|isbn=9780136020400}}
* {{Citation|last=Cameron|first=Peter J.|title=Combinatorics: Topics, Techniques, Algorithms|publisher=Cambridge University Press|year=1994|isbn=0-521-45761-0}}
* {{Citation
  | last = Fernández
  | first = Roberto
  | last2 = Fröhlich
  | first2 = Jürg
  | author2-link = Jürg Fröhlich
  | last3 = Alan D.
  | first3 = Sokal
  | author3-link = Alan Sokal
  | title = Random Walks, Critical Phenomena, and Triviality in Quantum Field Theory
  | place = Berlin
  | publisher = [[Springer-Verlag]]
  | series = Texts an Monographs in Physics
  | year = 1992
  | pages = xviii+444
  | isbn = 3-540-54358-9
  | mr = 1219313
  | zbl = 0761.60061}}
* {{Citation
| last1 = Graham  | first1 = R.L.
| last2 = Grotschel | first2 = M.
| last3 = Lovasz | first3 = L.
| title = Hand Book of Combinatorics (volume-2)
| publisher = MIT Press – North Holland
| year = 1995
| isbn = 9780262071710
| url = http://www.amazon.com/Handbook-Combinatorics-Vol-Ronald-Graham/dp/0262071711
}}
* {{citation|last=Gross|first=Jonathan L.|title=Combinatorial Methods with Computer Applications|publisher=Chapman&Hall/CRC|year=2008|isbn=9781584887430}}
* {{springer|title=Inclusion-and-exclusion principle|id=p/i050430}}
* {{citation|last=Mazur|first=David R.|title=Combinatorics A Guided Tour|publisher=The Mathematical Association of America|year=2010|isbn=9780883857625}}
* {{citation|last1=Roberts|first1=Fred S.|last2=Tesman|first2=Barry|title=Applied Combinatorics|edition=2nd|publisher=CRC Press|year=2009|isbn=9781420099829}}
* {{citation|last=Stanley|first=Richard P.|title=Enumerative Combinatorics Volume I|publisher=Wadsworth & Brooks/Cole|year=1986|isbn=0534065465}}
* {{citation|last1=van Lint|first1=J.H. |last2=Wilson|first2=R.M. |title=A Course in Combinatorics|publisher=Cambridge University Press|year= 1992|isbn=0521422604}}
{{PlanetMath attribution|id=2803|title=principle of inclusion–exclusion}}
 
{{DEFAULTSORT:Inclusion-exclusion principle}}
[[Category:Enumerative combinatorics]]
[[Category:Probability theory]]
[[Category:Articles containing proofs]]
[[Category:Mathematical principles]]

Latest revision as of 21:42, 14 September 2014

My name is Roseanna McLaren but everybody calls me Roseanna. I'm from Austria. I'm studying at the high school (final year) and I play the Piano for 4 years. Usually I choose songs from the famous films :).
I have two sister. I love Musical instruments, watching TV (The Big Bang Theory) and Leaf collecting and pressing.

Also visit my website How To Get Free Fifa 15 Coins