id
stringlengths
1
260
contents
stringlengths
1
234k
21173
\section{Set of Finite Subsets of Countable Set is Countable} Tags: Countable Sets, Subsets, Subset, Set of Finite Subsets of Countable Set is Countable \begin{theorem} Let $A$ be a countable set. Then the set of finite subsets of $A$ is countable. \end{theorem} \begin{proof} Let $A^{(n)}$ be the set of subsets of $A$ with no more than $n$ elements. Thus: :$A^{(0)} = \left\{{\varnothing}\right\}$ :$A^{(1)} = A^{(0)} \cup \left\{{\left\{{a}\right\}: a \in A}\right\}$ and $\forall n \ge 0$: :$A^{(n+1)} = \left\{{a^{(n)} \cup a^{(1)}: a^{(n)} \in A^{(n)} \land a^{(1)} \in A^{(1)}}\right\}$ Let us verify by induction that each $A^{(n)}$ is countable. Let $A$ be countable. :$A^{(1)}$ is countable, as its cardinality is $1 + \left|{A}\right|$. Suppose $A^{(n)}$ is countable. Then by Union of Countable Sets of Sets, so $A^{(n+1)}$ also countable. By induction, each $A^{(n)}$ is countable. Denote with $A^f$ the set of finite subsets of $A$. It is apparent that every finite subset is in some $A^{(n)}$, and so: :$A^f = \displaystyle \bigcup_{n \mathop \in \N} A^{(n)}$ The result follows from Countable Union of Countable Sets is Countable. {{WIP|actually, this theorem is provable in ZF}} {{qed}} \end{proof}
21174
\section{Set of Finite Subsets under Induced Operation is Closed} Tags: Subset Products, Power Set \begin{theorem} Let $\struct {S, \circ}$ be a magma. Let $\struct {\powerset S, \circ_\PP}$ be the algebraic structure consisting of the power set of $S$ and the operation induced on $\powerset S$ by $\circ$. Let $T \subseteq \powerset S$ be the set of all finite subsets of $S$. Then the algebraic structure $\struct {T, \circ_\PP}$ is closed. \end{theorem} \begin{proof} Let $X, Y \in T$. Then: {{begin-eqn}} {{eqn | q = | l = X \circ_\PP Y | r = \set {x \circ y: x \in X, y \in Y} | c = {{Defof|Operation Induced on Power Set}} }} {{eqn | ll= \leadsto | l = \card {X \circ_\PP Y} | o = \le | r = \card X \times \card Y | c = }} {{eqn | ll= \leadsto | l = X \circ_\PP Y | o = \in | r = T | c = }} {{end-eqn}} {{handwaving|Between step $1$ and step $2$ we really need a result Order of Subset Product is not Greater than Cardinality of Cartesian Product which is straightforward but tedious, and I lack patience.}} {{qed}} \end{proof}
21175
\section{Set of Finite Suprema is Directed} Tags: Join and Meet Semilattices \begin{theorem} Let $\struct {S, \vee, \preceq}$ be a join semilattice. Let $X$ be a non-empty subset of $S$. Then :$\set {\sup A: A \in \map {\operatorname {Fin} } X \land A \ne \O}$ is directed. where $\map {\operatorname {Fin} } X$ denotes the set of all finite subsets of $X$. \end{theorem} \begin{proof} By Existence of Non-Empty Finite Suprema in Join Semilattice: :for every $A \in \map {\operatorname {Fin} } S$ if $A \ne \O$, then $A$ admits a supremum. By definition of non-empty set: :$\exists a: a \in X$ By definitions of subset and singleton: :$\set x \subseteq X$ By Singleton is Finite: :$\set x$ is finite By definitions of non-empty set and singleton: :$\set x \ne \O$ By definition of $\operatorname {Fin}$: :$\sup {\set x} \in \set {\sup A: A \in \map {\operatorname{Fin} } X \land A \ne \O}$ Thus by definition: :$\set {\sup A: A \in \map {\operatorname {Fin} } X \land A \ne \O}$ is a non-empty set. Let $x, y \in \set {\sup A: A \in \map {\operatorname {Fin} } X \land A \ne \O}$ Then :$\exists A \in \map {\operatorname {Fin} } X: x = \sup A$ and $A \ne \O$ and :$\exists B \in \map {\operatorname {Fin} } X: y = \sup B$ and $B \ne \O$ By Finite Union of Finite Sets is Finite: :$A \cup B$ is finite By Union of Subsets is Subset: :$A \cup B \subseteq X$ By definitions of non-empty set and union: :$A \cup B \ne \O$ By definition of $\operatorname {Fin}$: :$\map \sup {A \cup B} \in \set {\sup A: A \in \map {\operatorname {Fin} } X \land A \ne \O}$ By Set is Subset of Union: :$A \subseteq A \cup B$ and $B \subseteq A \cup B$ Thus by Supremum of Subset: :$x \preceq \map \sup {A \cup B}$ and $y \preceq \map \sup {A \cup B}$ Thus by definition: :$\set {\sup A: A \in \map {\operatorname {Fin} } X \land A \ne \O}$ is directed. {{qed}} \end{proof}
21176
\section{Set of Gödel Numbers of Arithmetic Theorems Not Definable in Arithmetic} Tags: Mathematical Logic \begin{theorem} Let $T$ be the set of theorems of some consistent theory in the language of arithmetic which contains minimal arithmetic. The set of Gödel numbers of the theorems of $T$ is not definable in $T$. \end{theorem} \begin{proof} {{Questionable|Not only the proof is faulty, this theorem is wrong. If we have a recursively enumerable set A of axioms, then the set of theorems (and hence, the Gödel numbers of those) proven by A is recursively enumerable. Minimal arithmetic and PA are both recursively enumerable, and hence their theorems have a recursively enumerable set of Gödel numbers. Any recursively enumerable set is arithmetical with degree Sigma^0_1 in the arithmetical hierarchy. Moreover, there is already a well known provability predicate Pr(x) in the language of Peano arithmetic. Indeed, if we use the diagonal lemma on ~Pr(x) to get a sentence with T proves G iff ~Pr(G), then we will have an unprovable sentence.|Burak}} The proof is by contradiction. Let $\Theta$ be the set of Gödel numbers of the theorems of $T$. Suppose it is defined in $T$ by the formula $\map \theta y$. Since $T$ contains $Q$, we may apply the Diagonal Lemma to $\neg \map \theta y$. This gives us a sentence $G$ such that :$T \vdash G \leftrightarrow \neg \map \theta {\hat G}$ where $\hat G$ is the Gödel number of $G$ (more accurately, it is the term in the language of arithmetic obtained by applying the function symbol $s$ to $0$ this many times). {{Questionable|The below argument is not correct. $T$ does not necessarily prove that $\neg \map \theta {\hat G}$ just because it does not prove $G$. PA has a provability predicate $\map \Pr x$ which satisfies the property that if PA proves G, then PA proves $\map \Pr {\hat G}$). The step used here assumes a somewhat similar property in the opposite direction that our provability predicate satisfies if T does not prove G, then T proves $\neg \map \theta {\hat G}$, which we did not have as an assumption on $\map \Theta x$. |Burak}} Suppose $G$ is not a theorem of $T$. :Then the Gödel number of $G$ is not in $\Theta$. :Since $\theta$ defines $\Theta$ in $T$, this means that: ::$T \vdash \neg \map \theta {\hat G}$ :But, by choice of $G$ (specifically, the bi-implication above), this gives us: ::$T\vdash G$ :which contradicts $G$ not being a theorem of $T$ Thus, $G$ is a theorem of $T$. :But, then the Gödel number of $G$ is in $\Theta$, and :since $\theta$ defines $\Theta$ in $T$, this means that ::$T \vdash \map \theta {\hat G}$ :But, then this gives us ::$T \vdash \neg G$ :which contradicts $G$ being a theorem of $T$, ''since $T$ is consistent''. Since assuming $\Theta$ was definable in $T$ necessarily leads to a contradiction, we conclude that it is impossible. {{qed}} \end{proof}
21177
\section{Set of Homomorphisms to Abelian Group is Subgroup of All Mappings} Tags: Subgroups, Abelian Groups, Group Theory, Morphisms, Homomorphisms \begin{theorem} Let $\struct {S, \circ}$ be an algebraic structure. Let $\struct {T, \oplus}$ be an abelian group. Let $\struct {T^S, \oplus}$ be the algebraic structure on $T^S$ induced by $\oplus$. Then the set of all homomorphisms from $\struct {S, \circ}$ into $\struct {T, \oplus}$ is a subgroup of $\struct {T^S, \oplus}$. \end{theorem} \begin{proof} Let $H$ be the set of all homomorphisms from $\struct {S, \circ}$ into $\struct {T, \oplus}$. \end{proof}
21178
\section{Set of Ideals forms Complete Lattice} Tags: Lattice Theory, Ideal Theory \begin{theorem} Let $\struct {K, +, \circ}$ be a ring. Let $\mathbb K$ be the set of all ideals of $K$. Then $\struct {\mathbb K, \subseteq}$ is a complete lattice. \end{theorem} \begin{proof} Let $\O \subset \mathbb S \subseteq \mathbb K$. By Intersection of Ring Ideals is Largest Ideal Contained in all Ideals: :$\bigcap \mathbb S$ is the largest ideal of $K$ contained in each of the elements of $\mathbb S$. By Intersection of Ring Ideals Containing Subset is Smallest: :The intersection of the set of all ideals of $K$ containing $\bigcup \mathbb S$ is the smallest ideal of $K$ containing $\bigcup \mathbb S$. Thus, not only is $\bigcap \mathbb S$ a lower bound of $\mathbb S$, but also the largest, and therefore an infimum. The supremum of $\mathbb S$ is the join of the set of all ideals of $\mathbb S$. From Sum of Ideals is Ideal: General Result, this supremum is: :$\ds \sum_{S \mathop \in \mathbb S} \struct {S, +, \circ}$ That is: :$\struct {S_1, +, \circ} + \struct {S_2, +, \circ} + \dotsb$ where addition of ideals is as defined in subset product. Therefore $\struct {\mathbb K, \subseteq}$ is a complete lattice. {{qed}} \end{proof}
21179
\section{Set of Infima for Sequence is Directed} Tags: Order Theory \begin{theorem} Let $\left({S, \vee, \wedge, \preceq}\right)$ be a complete lattice. Let $\left({A, \precsim}\right)$ be a non-empty directed set. Let $Z: A \to S$ be a Moore-Smith sequence. Let $D = \left\{ {\inf \left({Z\left[{\precsim \left({j}\right)}\right]}\right): j \in A}\right\}$ be a subset of $S$. Then $D$ is directed. \end{theorem} \begin{proof} By definition of non-empty set: :$\exists j: j \in A$ By definition of $D$: :$\inf \left({Z\left[{\precsim \left({j}\right)}\right]}\right) \in D$ Hence by definition: :$D$ is a non-empty set. Let $x, y \in D$. By definition of $D$: :$\exists j_1 \in A: x = \inf \left({Z\left[{\precsim \left({j_1}\right)}\right]}\right)$ and :$\exists j_2 \in A: y = \inf \left({Z\left[{\precsim \left({j_2}\right)}\right]}\right)$ By definition of directed set: :$\exists j \in A: j_1 \precsim j \land j_2 \precsim j$ By Preceding implies Image is Subset of Image: :$\precsim\left({j}\right) \subseteq \mathord\precsim\left({j_1}\right)$ and $\precsim\left({j}\right) \subseteq \mathord\precsim\left({j_2}\right)$ By Image of Subset under Relation is Subset of Image/Corollary 2 :$Z\left[{\precsim\left({j}\right)}\right] \subseteq Z\left[{\precsim\left({j_1}\right)}\right]$ and $Z\left[{\precsim\left({j}\right)}\right] \subseteq Z\left[{\precsim\left({j_2}\right)}\right]$ Thus by definition of $D$: :$z := \inf \left({Z\left[{\precsim \left({j}\right)}\right]}\right) \in D$ Thus by Infimum of Subset: :$x \preceq z$ and $y \preceq z$ {{qed}} \end{proof}
21180
\section{Set of Infinite Sequences is Uncountable} Tags: Diagonal Arguments, Countable Sets \begin{theorem} Let $S$ be a set which contains more than one element. Let $S^\infty$ denote the set of all sequences of elements of $S$. Then $S^\infty$ is uncountable. \end{theorem} \begin{proof} As $S$ has more than one element, it must have at least two. So, let $a, b \in S$ be those two elements. Let $Z$ be the set of all sequences from $\set {a, b}$. Suppose $S^\infty$ were countable. From Subset of Countably Infinite Set is Countable, $Z$ is likewise countable. So by definition, it would be possible to set up a bijection $\phi: Z \leftrightarrow \N$ between $Z$ and the set $\N$ of natural numbers: {{begin-eqn}} {{eqn | l = 0 | o = \leftrightarrow | r = \sequence {z_0} = \tuple {z_{00}, z_{01}, z_{02}, \ldots, z_{0n}, \ldots} }} {{eqn | l = 1 | o = \leftrightarrow | r = \sequence {z_1} = \tuple {z_{10}, z_{11}, z_{12}, \ldots, z_{1n}, \ldots} }} {{eqn | l = | o = \vdots | r = }} {{eqn | l = n | o = \leftrightarrow | r = \sequence {z_n} = \tuple {z_{n0}, z_{n1}, z_{n2}, \ldots, z_{nn}, \ldots} }} {{eqn | l = | o = \vdots | r = }} {{end-eqn}} where each of $z_{rs}$ is: : an element of $\set {a, b}$ : the $s$'th element of $\sequence {z_r}$, the particular sequence which is in correspondence with $r \in \N$. Now we create the sequence $Q = \sequence {q_{dd} }$ where: :$q_{dd} = \begin{cases} a & : z_{dd} = b \\ b & : z_{dd} = a \end{cases}$ Thus: :$\forall n \in \N: q_{nn} \notin \sequence {z_n}$ and so: :$\forall n \in \N: Q \ne \sequence {z_n}$ So $Q \notin Z$. That is, we have constructed a sequence which has not been placed into correspondence with an element of $\N$. Thus by definition $Z$ is uncountable. Hence, by the Rule of Transposition, $S^\infty$ is likewise uncountable. {{qed}} \end{proof}
21181
\section{Set of Integer Multiples is Integral Ideal} Tags: Sets of Integer Multiples, Integral Ideals \begin{theorem} Let $m \in \Z$ be an integer. Let $m \Z$ denote the set of integer multiples of $m$. Then $m \Z$ is an integral ideal. \end{theorem} \begin{proof} First note that $m \times 0 \in m \Z$ whatever $m$ may be. Thus $m \Z \ne \O$. Let $a, b \in m \Z$. Then: {{begin-eqn}} {{eqn | l = a + b | r = m j + m k | c = for some $j, k \in \Z$ by definition of $m \Z$ }} {{eqn | r = m \paren {j + k} | c = }} {{eqn | o = \in | r = m \Z | c = }} {{end-eqn}} and: {{begin-eqn}} {{eqn | l = a - b | r = m j - m k | c = for some $j, k \in \Z$ by definition of $m \Z$ }} {{eqn | r = m \paren {j - k} | c = }} {{eqn | o = \in | r = m \Z | c = }} {{end-eqn}} Let $r \in \Z$. Then: {{begin-eqn}} {{eqn | l = r a | r = r \paren {m j} | c = for some $j \in \Z$ by definition of $m \Z$ }} {{eqn | r = m \paren {r j} | c = }} {{eqn | o = \in | r = m \Z | c = }} {{end-eqn}} Thus the conditions for $m \Z$ to be an integral ideal are fulfilled. Hence the result. {{qed}} \end{proof}
21182
\section{Set of Integer Multiples of GCD} Tags: Sets of Integer Multiples, Greatest Common Divisor \begin{theorem} Let $m, n \in \Z$. Let $m \Z$ denote the set of integer multiples of $m$ Then: :$m \Z \cup n \Z \subseteq \gcd \set {m, n} \Z$ where $\gcd$ denotes greatest common divisor. \end{theorem} \begin{proof} Let $x \in m \Z \cup n \Z$. Then either: :$m \divides x$ or: :$n \divides x$ In both cases: :$\gcd \set {m, n} \divides x$ and so: :$x \in \gcd \set {m, n} \Z$ Hence by definition of subset: :$m \Z \cup n \Z \subseteq \gcd \set {m, n} \Z$ {{qed}} \end{proof}
21183
\section{Set of Integers Bounded Above by Integer has Greatest Element} Tags: Number Theory, Integers, Set of Integers Bounded Above has Greatest Element \begin{theorem} Let $\Z$ be the set of integers. Let $\le$ be the ordering on the integers. Let $\O \subset S \subseteq \Z$ such that $S$ is bounded above in $\struct {\Z, \le}$. Then $S$ has a greatest element. \end{theorem} \begin{proof} $S$ is bounded above, so $\exists M \in \Z: \forall s \in S: s \le M$. Hence $\forall s \in S: 0 \le M - s$. Thus the set $T = \left\{{M - s: s \in S}\right\} \subseteq \N$. The Well-Ordering Principle gives that $T$ has a smallest element, which we can call $b_T \in T$. Hence: : $\left({\forall s \in S: b_T \le M - s}\right) \land \left({\exists g_S \in S: b_T = M - g_S}\right)$ So: {{begin-eqn}} {{eqn | o= | r=\forall s \in S: M - g_S \le M - s | c= }} {{eqn | o=\implies | r=\forall s \in S: -g_S \le -s | c=Cancellability of elements of $\Z$ }} {{eqn | o=\implies | r=\forall s \in S: g_S \ge s | c= }} {{eqn | o=\implies | r=g_S \in S \land \left({\forall s \in S: g_S \ge s}\right) | c=... which is how the greatest element is defined. }} {{end-eqn}} So $g_S$ is the greatest element of $S$. {{qed}} \end{proof}
21184
\section{Set of Integers Bounded Above by Real Number has Greatest Element} Tags: Number Theory, Set of Integers Bounded Above has Greatest Element \begin{theorem} Let $\Z$ be the set of integers. Let $\le$ be the usual ordering on the real numbers $\R$. Let $\O \subset S \subseteq \Z$ such that $S$ is bounded above in $\struct {\R, \le}$. Then $S$ has a greatest element. \end{theorem} \begin{proof} Let $S$ be bounded above by $x \in \R$. By the Archimedean Principle, there exists an integer $n \ge x$. Then $S$ is bounded above by $n$. By Set of Integers Bounded Above by Integer has Greatest Element, $S$ has a greatest element. {{qed}} \end{proof}
21185
\section{Set of Integers Bounded Below by Integer has Smallest Element} Tags: Number Theory, Well-Orderings, Integers, Set of Integers Bounded Below has Smallest Element \begin{theorem} Let $\Z$ be the set of integers. Let $\le$ be the ordering on the integers. Let $\O \subset S \subseteq \Z$ such that $S$ is bounded below in $\struct {\Z, \le}$. Then $S$ has a smallest element. \end{theorem} \begin{proof} We have that $S$ is bounded below in $\Z$. So: : $\exists m \in \Z: \forall s \in S: m \le s$ Hence: : $\forall s \in S: 0 \le s - m$ Thus: :$T = \set {s - m: s \in S} \subseteq \N$ The Well-Ordering Principle gives that $T$ has a smallest element, which we can call $b_T \in T$. Hence: :$\paren {\forall s \in S: b_T \le s - m} \land \paren {\exists b_S \in S: b_T = b_S - m}$ So: {{begin-eqn}} {{eqn | o = | r = s \in S: b_S - m \le s - m | c = }} {{eqn | o = \leadsto | r = \forall s \in S: b_S \le s | c = Cancellability of elements of $\Z$ }} {{eqn | o = \leadsto | r = b_S \in S \land \paren {\forall s \in S: b_S \le s} | c = {{Defof|Smallest Element}} }} {{end-eqn}} So $b_S$ is the smallest element of $S$. {{Qed}} \end{proof}
21186
\section{Set of Integers Bounded Below by Real Number has Smallest Element} Tags: Number Theory, Set of Integers Bounded Below has Smallest Element \begin{theorem} Let $\Z$ be the set of integers. Let $\le$ be the usual ordering on the real numbers $\R$. Let $\O \subset S \subseteq \Z$ such that $S$ is bounded below in $\struct {\R, \le}$. Then $S$ has a smallest element. \end{theorem} \begin{proof} Let $S$ be bounded below by $x \in \R$. By the Archimedean Principle, there exists an integer $n \le x$. Then $S$ is bounded below by $n$. By Set of Integers Bounded Below by Integer has Smallest Element, $S$ has a smallest element. {{qed}} \end{proof}
21187
\section{Set of Integers can be Well-Ordered} Tags: Well-Orderings, Integers \begin{theorem} The set of integers $\Z$ can be well-ordered with an appropriately chosen ordering. \end{theorem} \begin{proof} Consider the ordering $\preccurlyeq \subseteq \Z \times \Z$ defined as: :$x \preccurlyeq y \iff \left({\left\vert{x}\right\vert < \left\vert{y}\right\vert}\right) \lor \left({\left\vert{x}\right\vert = \left\vert{y}\right\vert \land x \le y}\right)$ {{finish|It remains to be shown that $\preccurlyeq$ is an ordering, and also that it is a well-ordering.}} \end{proof}
21188
\section{Set of Integers is not Bounded} Tags: Boundedness, Real Analysis, Definitions: Analysis, Analysis \begin{theorem} Let $\R$ be the real number line considered as an Euclidean space. The set $\Z$ of integers is not bounded in $\R$. \end{theorem} \begin{proof} Let $a \in \R$. Let $K \in \R_{>0}$. Consider the open $K$-ball $\map {B_K} a$. By the Archimedean Principle there exists $n \in \N$ such that $n > a + K$. As $\N \subseteq \Z$: :$\exists n \in \Z: a + K < n$ and so: :$n \notin \map {B_K} a$ As this applies whatever $a$ and $K$ are, it follows that there is no $\map {B_K} a$ which contains all the integers. Hence the result, by definition of bounded space. {{qed}} \end{proof}
21189
\section{Set of Integers is not Compact} Tags: Compact Spaces \begin{theorem} Let $\Z$ be the set of integers. Then $\Z$ is not compact. \end{theorem} \begin{proof} Let $\R$ be the real number line considered as an Euclidean space. From Set of Integers is not Bounded, $\Z$ is not bounded in $\R$. The result follows by definition of compact. {{qed}} \end{proof}
21190
\section{Set of Integers is not Well-Ordered by Usual Ordering} Tags: Well-Orderings, Integers \begin{theorem} The set of integers $\Z$ is not well-ordered under the usual ordering $\le$. \end{theorem} \begin{proof} {{AimForCont}} $\Z$ is a well-ordered set. Then by definition, all subsets of $\Z$ has a smallest element. But take $\Z$ itself. Suppose $x \in \Z$ is a smallest element. Then $x - 1 \in \Z$. But $x - 1 < x$, which contradicts the supposition that $x \in \Z$ is a smallest element. Hence there can be no such smallest element. So by Proof by Contradiction, $\Z$ is not well-ordered by $\le$. {{qed}} \end{proof}
21191
\section{Set of Integers under Addition is Isomorphic to Set of Even Integers under Addition} Tags: Additive Group of Integers, Integers, Even Integers \begin{theorem} Let $\struct {\Z, +}$ be the algebraic structure formed by the set of integers under the operation of addition. Let $\struct {2 \Z, +}$ be the algebraic structure formed by the set of even integers under the operation of addition. Then $\struct {\Z, +}$ and $\struct {2 \Z, +}$ are isomorphic. \end{theorem} \begin{proof} Let $f: \Z \to 2 \Z$ be the mapping: :$\forall n \in \Z: \map f n = 2 n$ From Bijection between Integers and Even Integers, $f$ is a bijection. Let $m, n \in \Z$. Then: {{begin-eqn}} {{eqn | l = \map f {m + n} | r = 2 \paren {m + n} | c = Definition of $f$ }} {{eqn | r = 2 m + 2 n | c = Integer Multiplication Distributes over Addition }} {{eqn | r = \map f m + \map f n | c = Integer Multiplication Distributes over Addition }} {{end-eqn}} Thus $f$ is an isomorphism by definition. {{qed}} \end{proof}
21192
\section{Set of Inverse Positive Integers with Zero is Compact} Tags: Compact Spaces, Integers \begin{theorem} Let $K$ be the set of inverse positive integers with zero: :$\ds K := \set {1, \frac 1 2, \frac 1 3, \dots} \cup \set 0$ Let $\struct {\R, \size {\, \cdot \,}}$ be the normed vector space of real numbers. Then $K$ is compact in real numbers. \end{theorem} \begin{proof} We have that $K \subset \closedint 0 1$. Hence, $K$ is bounded. Furthermore: :$\ds \R \setminus K = \openint {-\infty} 0 \cup \paren {\bigcup_{n \mathop = 1}^\infty \openint {\frac 1 {n + 1}} {\frac 1 n}} \cup \openint 1 \infty$ By Union of Open Sets of Normed Vector Space is Open, $\R \setminus K$ is open. By definition, $K$ is closed. By Heine-Borel theorem, $K$ is compact. {{qed}} \end{proof}
21193
\section{Set of Invertible Mappings forms Symmetric Group} Tags: Group of Permutations, Symmetric Groups, Permutation Theory, Group Examples, Permutations \begin{theorem} Let $S$ be a set. Let $\GG$ be the set of all invertible mappings from $S$ to $S$. Then $\struct {\GG, \circ}$ is the symmetric group on $S$. \end{theorem} \begin{proof} Let $\struct {S^S, \circ}$ be the algebraic structure formed from the set of all mappings from $S$ to itself. From Set of all Self-Maps under Composition forms Monoid, $\struct {S^S, \circ}$ is a monoid. By Inverse of Permutation is Permutation, if $f$ is a permutation of $S$, then so is its inverse $f^{-1}$. By Bijection iff Inverse is Bijection, it follows that all the invertible elements of $S^S$ are exactly the permutations on $S$. The result follows from Invertible Elements of Monoid form Subgroup of Cancellable Elements. {{qed}} \end{proof}
21194
\section{Set of Isolated Points of Metric Space is Disjoint from Limit Points} Tags: Isolated Points, Limit Points \begin{theorem} Let $M = \struct {A, d}$ be a metric space. Let $H \subseteq A$ be a subset of $A$. Let $H'$ be the set of limit points of $H$. Let $H^i$ be the set of isolated points of $H$. Then: :$H' \cap H^i = \O$ \end{theorem} \begin{proof} Let $a \in H_i$. Then by definition of isolated point: :$\exists \epsilon \in \R_{>0}: \set {x \in H: \map d {x, a} < \epsilon} = \set a$ But by {{Metric-space-axiom|1}}: :$\map d {a, a} = 0$ and so: :$\set {x \in H: 0 < \map d {x, a} < \epsilon} = \O$ So by definition $a$ is not a limit point of $H$. That is: :$a \notin H'$ or: :$a \in \relcomp A {H'}$ It follows from Intersection with Complement is Empty iff Subset that: :$H' \cap H^i = \O$ {{qed}} \end{proof}
21195
\section{Set of Isometries in Complex Plane under Composition forms Group} Tags: Complex Numbers, Isometries, Examples of Groups \begin{theorem} Let $S$ be the set of all complex functions $f: \C \to \C$ which preserve distance when embedded in the complex plane. That is: :$\size {\map f a - \map f b} = \size {a - b}$ Let $\struct {S, \circ}$ be the algebraic structure formed from $S$ and the composition operation $\circ$. Then $\struct {S, \circ}$ is a group. \end{theorem} \begin{proof} From Complex Plane is Metric Space, $\C$ can be treated as a metric space. Hence it is seen that a complex function $f: \C \to \C$ which preserves distance is in fact an isometry on $\C$. Taking the group axioms in turn: \end{proof}
21196
\section{Set of Isometries in Euclidean Space under Composition forms Group} Tags: Euclidean Metric, Isometries, Examples of Groups \begin{theorem} Let $\struct {\R^n, d}$ be a real Euclidean space of $n$ dimensions. Let $S$ be the set of all mappings $f: \R^n \to \R^n$ which preserve distance: That is: :$\map d {\map f a, \map f b} = \map d {a, b}$ Let $\struct {S, \circ}$ be the algebraic structure formed from $S$ and the composition operation $\circ$. Then $\struct {S, \circ}$ is a group. \end{theorem} \begin{proof} From Euclidean Metric on Real Vector Space is Metric, $\R^n$ is a metric space. Hence it is seen that a complex function $f: \C \to \C$ which preserves distance is in fact an isometry on $\C$. Taking the group axioms in turn: \end{proof}
21197
\section{Set of Linear Combinations of Finite Set of Elements of Principal Ideal Domain is Principal Ideal} Tags: Principal Ideal Domains \begin{theorem} Let $\struct {D, +, \circ}$ be a principal ideal domain. Let $a_1, a_2, \dotsc, a_n$ be non-zero elements of $D$. Let $J$ be the set of all linear combinations in $D$ of $\set {a_1, a_2, \dotsc, a_n}$ Then for some $x \in D$: :$J = \ideal x$ where $\ideal x$ denotes the principal ideal generated by $x$. \end{theorem} \begin{proof} Let the unity of $D$ be $1_D$. By definition of principal ideal: :$\ds \ideal a = \set {\sum_{i \mathop = 1}^n r_i \circ a \circ s_i: n \in \N, r_i, s_i \in D}$ Let $x, y \in J$. By definition of linear combination: {{begin-eqn}} {{eqn | l = x | r = \sum_{i \mathop = 1}^n r_i \circ a_i | c = for some $n \in \N$ and for some $r_i \in D$ where $i \in \set {1, 2, \dotsc, n}$ }} {{eqn | r = r_1 \circ a_1 + r_2 \circ a_2 + \dotsb + r_n \circ a_n | c = for some $r_1, r_2, \dotsc, r_n \in D$ }} {{end-eqn}} and: {{begin-eqn}} {{eqn | l = y | r = \sum_{i \mathop = 1}^n s_i \circ a_i | c = for some $n \in \N$ and for some $s_i \in D$ where $i \in \set {1, 2, \dotsc, n}$ }} {{eqn | r = s_1 \circ a_1 + s_2 \circ a_2 + \dotsb + s_n \circ a_n | c = for some $s_1, s_2, \dotsc, s_n \in R$ }} {{eqn | ll= \leadsto | l = -y | r = -\sum_{i \mathop = 1}^n s_i \circ a_i | c = }} {{eqn | r = -1_D \times \sum_{i \mathop = 1}^n s_i \circ a_i | c = Product with Ring Negative: Corollary }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {-1_D} \times \paren {s_i \circ a_i} | c = {{Ring-axiom|D}} }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {-\paren {s_i \circ a_i} } | c = Product with Ring Negative: Corollary }} {{eqn | r = \sum_{i \mathop = 1}^n s_i \circ \paren {-a_i} | c = Product with Ring Negative }} {{end-eqn}} Thus: {{begin-eqn}} {{eqn | l = x + \paren {-y} | r = \sum_{i \mathop = 1}^n r_i \circ a_i + \sum_{i \mathop = 1}^n s_i \circ \paren {-a_i} | c = }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {r_i \circ a_i + s_i \circ \paren {-a_i} } | c = }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {r_i \circ a_i + \paren {-\paren {s_i \circ a_i} } } | c = }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {r_i \circ a_i + \paren {-s_i} \circ a_i} | c = }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {r_i + \paren {-s_i} } \circ a_i | c = {{Ring-axiom|D}} }} {{eqn | r = \sum_{i \mathop = 1}^n t_i \circ a_i | c = where $t_i - r_1 + \paren {-s_i}$ }} {{eqn | o = \in | r = J | c = as $t_i \in D$ }} {{end-eqn}} Then we have: {{begin-eqn}} {{eqn | l = x \circ y | r = \paren {\sum_{i \mathop = 1}^n r_i \circ a_i} \circ \paren {\sum_{i \mathop = 1}^n s_i \circ a_i } | c = }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {t_i \circ a_i} | c = where $t_i \in D$ for $i \in \set {1, 2, \dotsc, n}$ }} {{eqn | r = \sum_{i \mathop = 1}^n \paren {a_i \circ t_i} | c = as $\circ$ is commutative in an integral domain }} {{end-eqn}} {{explain|There exists (or ought to) some convolution result which proves the above -- I just haven't found it yet.}} Thus by the Test for Ideal, $J$ is an ideal of $D$. As $D$ is a principal ideal domain, it follows that $J$ is a principal ideal. Thus by definition of principal ideal: :$J = \ideal x$ for some $x \in D$. {{qed}} \end{proof}
21198
\section{Set of Linear Subspaces is Closed under Intersection} Tags: Vector Spaces, Vector Subspaces \begin{theorem} Let $\struct {V, +, \circ}_K$ be a $K$-vector space. Let $\family {M_i}_{i \mathop \in I}$ be an $I$-indexed family of subspaces of $V$. Then $M := \ds \bigcap_{i \mathop \in I} M_i$ is also a subspace of $V$. \end{theorem} \begin{proof} It needs to be demonstrated that $M$ is: :$(1): \quad$ a closed algebraic structure under $+$ :$(2): \quad$ closed for scalar product $\circ$. So let $a, b \in M$. By definition of intersection, $a, b \in M_i$ for all $i \in I$. As the $M_i$ are subspaces of $V$, $a + b \in M_i$ for all $i \in I$. That is, by definition of intersection, $a + b \in M$. It follows that $M$ is closed under $+$. Now let $\lambda \in K, a \in M$. By definition of intersection, $a \in M_i$ for all $i \in I$. As the $M_i$ are subspaces of $V$, $\lambda \circ a \in M_i$ for all $i \in I$. That is, by definition of intersection, $\lambda \circ a \in M$. It follows that $M$ is closed under $\circ$. Hence the result, by definition of subspace. {{qed}} Category:Vector Subspaces \end{proof}
21199
\section{Set of Linear Transformations is Isomorphic to Matrix Space} Tags: Linear Algebra, Set of Linear Transformations is Isomorphic to Matrix Space, Matrix Algebra, Linear Transformations \begin{theorem} Let $R$ be a ring with unity. Let $F$, $G$ and $H$ be free $R$-modules of finite dimension $p,n,m>0$ respectively. Let $\sequence {a_p}$, $\sequence {b_n}$ and $\sequence {c_m}$ be ordered bases Let $\map {\LL_R} {G, H}$ denote the set of all linear transformations from $G$ to $H$. Let $\map {\MM_R} {m, n}$ be the $m \times n$ matrix space over $R$. Let $\sqbrk {u; \sequence {c_m}, \sequence {b_n} }$ be the matrix of $u$ relative to $\sequence {b_n}$ and $\sequence {c_m}$. Let $M: \map {\LL_R} {G, H} \to \map {\MM_R} {m, n}$ be defined as: :$\forall u \in \map {\LL_R} {G, H}: \map M u = \sqbrk {u; \sequence {c_m}, \sequence {b_n} }$ Then $M$ is a module isomorphism. \end{theorem} \begin{proof} Let $u, v \in \map {\LL_R} {G, H}$ such that: :$\map M u = \map M v$ We have that the matrix of $u$ relative to $\sequence {b_n}$ and $\sequence {c_m}$ is defined as the $m \times n$ matrix $\sqbrk \alpha_{m n}$ where: :$\ds \forall \tuple {i, j} \in \closedint 1 m \times \closedint 1 n: \map u {b_j} = \sum_{i \mathop = 1}^m \alpha_{i j} \circ c_i$ and it is seen that $\map M u$ and $\map M v$ are the same object. That is: :$\map M u = \map M v \implies u = v$ and $M$ is seen to be injective. {{finish|Surjectivity needs proving, so does homomorphism}} \end{proof}
21200
\section{Set of Linear Transformations over Commutative Ring forms Submodule of Module of All Mappings} Tags: Linear Transformations \begin{theorem} Let $R$ be a commutative ring. Let $\struct {G, +_G, \circ}_R$ and $\struct {H, +_H, \circ}_R$ be $R$-modules. Let $\map {\LL_R} {G, H}$ be the set of all linear transformations from $G$ to $H$. Then $\map {\LL_R} {G, H}$ is a submodule of the $R$-module $H^G$. \end{theorem} \begin{proof} From Group equals Center iff Abelian, the center of a commutative ring $R$ is $R$ itself. So, we need to show that every $\phi \in \map {\LL_R} {G, H}$ fulfils the conditions to be a linear transformation: :$(1): \quad \forall x, y \in G: \map {\paren {\lambda \circ \phi} } {x +_G y} = \lambda \circ \map \phi x +_H \lambda \circ \map \phi y$ :$(2): \quad \forall x \in G: \forall \mu \in R: \map {\paren {\lambda \circ \phi} } {\mu \circ x} = \mu \circ \map {\paren {\lambda \circ \phi} } x$ for all $\phi \in \map {\LL_R} {G, H}$. From Linear Transformation from Center of Scalar Ring, these conditions are indeed fulfulled. Hence the result. {{qed}} \end{proof}
21201
\section{Set of Linear Transformations over Commutative Ring forms Submodule of Module of All Mappings/Unitary} Tags: Unitary Modules, Linear Transformations \begin{theorem} Let $R$ be a commutative ring with unity whose unity is $1_R$. Let $\struct {G, +_G, \circ}_R$ and $\struct {H, +_H, \circ}_R$ be $R$-modules. Let $\map {\LL_R} {G, H}$ be the set of all linear transformations from $G$ to $H$. Let $\struct {H, +_H, \circ}_R$ be a unitary module. Then $\map {\LL_R} {G, H}$ is also a unitary module. \end{theorem} \begin{proof} From Set of Linear Transformations over Commutative Ring forms Submodule of Module of All Mappings, $\map {\LL_R} {G, H}$ is a module. It remains to be shown that $\map {\LL_R} {G, H}$ is a unitary module, that is: :$\forall \phi \in \map {\LL_R} {G, H}: 1_R \circ \phi = \phi$ So, let $\struct {H, +_H, \circ}_R$ be a unitary $R$-module. Then: :$\forall x \in H: 1_R \circ x = x$ Thus: {{begin-eqn}} {{eqn | l = \map {\paren {1_R \circ \phi} } x | r = 1_R \circ \paren {\map \phi x} | c = }} {{eqn | r = \map \phi x | c = }} {{end-eqn}} Hence the result. {{qed}} \end{proof}
21202
\section{Set of Linear Transformations under Pointwise Addition forms Abelian Group} Tags: Abelian Groups, Linear Transformations \begin{theorem} Let $\struct {G, +_G}$ and $\struct {H, +_H}$ be abelian groups. Let $\struct {R, +_R, \times_R}$ be a ring. Let $\struct {G, +_G, \circ}_R$ and $\struct {H, +_H, \circ}_R$ be $R$-modules. Let $\map {\LL_R} {G, H}$ be the set of all linear transformations from $G$ to $H$. Let $\oplus_H$ be defined as pointwise addition on $\map {\LL_R} {G, H}$: :$\forall u, v \in \map {\LL_R} {G, H}: \forall x \in G: \map {\paren {u \oplus_H v} } x := \map u x +_H \map v x$ Then $\struct {\map {\LL_R} {G, H}, \oplus_H}$ is an abelian group. \end{theorem} \begin{proof} From Structure Induced by Group Operation is Group, $\struct {H^G, \oplus_H}$ is a group Let $\phi, \psi \in \map {\LL_R} {G, H}$. From Addition of Linear Transformations: :$\phi \oplus_H \psi \in \map {\LL_R} {G, H}$ From Negative Linear Transformation: :$-\phi \in \map {\LL_R} {G, H}$ Thus, from the Two-Step Subgroup Test: :$\struct {\map {\LL_R} {G, H}, \oplus_H}$ is a subgroup of $\struct {H^G, \oplus_H}$. It remains to be shown that $\struct {\map {\LL_R} {G, H}, \oplus_H}$ is abelian. Let $u$ and $v$ be arbitrary elements of $\map {\LL_R} {G, H}$. Indeed, we have that: {{begin-eqn}} {{eqn | q = | l = \map {\paren {u \oplus_H v} } x | r = \map u x +_H \map v x | c = {{Defof|Pointwise Addition of Linear Transformations}} }} {{eqn | r = \map v x +_H \map u x | c = as $H$ is abelian }} {{eqn | r = \map {\paren {v \oplus_H u} } x | c = {{Defof|Pointwise Addition of Linear Transformations}} }} {{end-eqn}} {{qed}} \end{proof}
21203
\section{Set of Liouville Numbers is Uncountable} Tags: Transcendental Numbers \begin{theorem} The set of Liouville numbers is uncountable. \end{theorem} \begin{proof} By Corollary to Liouville's Constant is Transcendental, all numbers of the form: {{begin-eqn}} {{eqn | l = \sum_{n \mathop \ge 1} \frac {a_n} {10^{n!} } | r = \frac {a_1} {10^1} + \frac {a_2} {10^2} + \frac {a_3} {10^6} + \frac {a_4} {10^{24} } + \cdots | c = }} {{end-eqn}} where :$a_1, a_2, a_3, \ldots \in \set {1, 2, \ldots, 9}$ are Liouville numbers. Therefore each sequence in $\set {1, 2, \ldots, 9}$ defines a unique Liouville number. By Set of Infinite Sequences is Uncountable, there are uncountable sequences in $\set {1, 2, \ldots, 9}$. As the set of Liouville numbers has an uncountable subset, it is also uncountable by Sufficient Conditions for Uncountability. {{qed}} Category:Transcendental Numbers \end{proof}
21204
\section{Set of Local Minimum is Countable} Tags: Countable Sets, Real Analysis \begin{theorem} Let $X$ be a subset of $\R$. The set: :$\leftset {x \in X: x}$ is local minimum in $\rightset X$ is countable. \end{theorem} \begin{proof} Define: :$Y := \leftset {x \in X: x}$ is local minimum in $\rightset X$ By definition of $Y$ and definition of local minimum in set: :$\forall x \in Y: \exists y \in \R: y < x \land \openint y x \cap X = \O$ By the Axiom of Choice, define a mapping $f: Y \to \powerset \R$ as: :$\forall x \in Y: \exists y \in \R: \map f x = \openint y x \land y < x \land \map f x \cap X = \O$ We will prove that $f$ is an injection by definition: Let $x_1, x_2 \in Y$ such that :$\map f {x_1} = \map f {x_2}$ By definition of $f$: :$\exists y_1 \in \R: \map f {x_1} = \openint {y_1} {x_1} \land y_1 < x_1 \land \map f {x_1} \cap X = \O$ and: :$\exists y_2 \in \R: \map f {x_2} = \openint {y_2} {x_2} \land y_2 < x_2 \land \map f {x_2} \cap X = \O$ Then: :$\openint {y_1} {x_1} = \openint {y_2} {x_2}$ Thus $x_1 = x_2$. This ends the proof of injection. By Cardinality of Image of Injection: :$(1): \quad \card Y = \card {\map {f^\to} Y} = \card {\Img f}$ where :$\card Y$ denotes the cardinality of $Y$, :$\map {f^\to} Y$ denotes the image of $Y$ under $f$, :$\Img f$ denotes the image of $f$. We will prove that $\Img f$ is pairwise disjoint by definition. Let $A, B \in \Img f$ such that :$A \ne B$. Then by definition of image: :$\exists x_1 \in Y: \map f {x_1} = A$ and :$\exists x_2 \in Y: \map f {x_2} = B$. By difference of $A$ and $B$: :$x_1 \ne x_2$ By definition of $f$: :$\exists y_1 \in \R: \map f {x_1} = \openint {y_1} {x_1} \land y_1 < x_1 \land \map f {x_1} \cap X = \O$ and: :$\exists y_2 \in \R: \map f {x_2} = \openint {y_2} {x_2} \land y_2 < x_2 \land \map f {x_2} \cap X = \O$ Aiming at contradiction suppose that :$A \cap B \ne \O$. $x_1 < x_2$ or $x_1 > x_2$. In case when $x_1 < x_2$, $x_1 \in \map f {x_2}$ what contradicts with $\map f {x_2} \cap X = \O$. In case when $x_1 > x_2$, analogically. This ends the proof that $\Img f$ is pairwise disjoint. By Set of Pairwise Disjoint Intervals is Countable: :$\Img f$ is countable. Thus by $(1)$ and Set is Countable if Cardinality equals Cardinality of Countable Set the result: :$Y$ is countable. {{qed}} \end{proof}
21205
\section{Set of Mappings can be Ordered by Inclusion} Tags: Order Theory \begin{theorem} Let $S \times T$ be the product of two sets. Let $\FF$ be a set of mappings on $S \times T$. Then $\FF$ can be ordered by inclusion. \end{theorem} \begin{proof} By the definition of mapping, a mapping is a specific type of relation. The result then follows from Set of Relations can be Ordered by Inclusion. {{qed}} Category:Order Theory \end{proof}
21206
\section{Set of Mappings which map to Same Element induces Equivalence Relation} Tags: Mapping Theory, Equivalence Relations \begin{theorem} Let $X$ and $Y$ be sets. Let $E$ be the set of all mappings from $X$ to $Y$. Let $b \in X$. Let $\RR \subseteq E \times E$ be the relation on $E$ defined as: :$\RR := \set {\tuple {f, g} \in \RR: \map f b = \map g b}$ Then $\RR$ is an equivalence relation. \end{theorem} \begin{proof} Checking in turn each of the criteria for equivalence: \end{proof}
21207
\section{Set of Meet Irreducible Elements Excluded Top is Order Generating} Tags: Order Generating, Continuous Lattices, Meet Irreducible \begin{theorem} Let $L = \left({S, \vee, \wedge, \preceq}\right)$ be a continuous complete lattice. Let $X = \mathit{IRR}\left({L}\right) \setminus \left\{ {\top}\right\}$ where $\mathit{IRR}\left({L}\right)$ denotes the set of all meet irreducible element of $S$, :$\top$ denotes the top of $L$. Then $X$ is order generating. \end{theorem} \begin{proof} We will prove that :$\forall x, y \in S: \left({ y \npreceq x \implies \exists p \in X: x \preceq p \land y \npreceq p }\right)$ Let $x, y \in S$ such that :$y \npreceq x$ By Not Preceding implies There Exists Meet Irreducible Element Not Preceding :$\exists p \in S: p$ is meet irreducible and $x \preceq p$ and $y \npreceq p$ By definition of greatest element: :$p \ne \top$ and $p \in \mathit{IRR}\left({L}\right)$ By definitions of difference and singleton: :$p \in X$ Thus :$\exists p \in X: x \preceq p \land y \npreceq p$ {{qed|lemma}} Hence by Order Generating iff Not Preceding implies There Exists Element Preceding and Not Preceding: :$X$ is order generating. {{qed}} \end{proof}
21208
\section{Set of Monomials is Closed Under Multiplication} Tags: Polynomial_Theory, Polynomial Theory, Monomials \begin{theorem} Let $M$ be the set of all monomials on the set $\set {X_j: j \in J}$, with multiplication $\circ$ defined by: :$\ds \paren {\prod_{j \mathop \in J} X_j^{k_j} } \circ \paren {\prod_{j \mathop \in J} X_j^{k_j'} } = \paren {\prod_{j \mathop \in J} X_j^{k_j + k_j'} }$ Then $M$ is closed under $\circ$. \end{theorem} \begin{proof} Let $\ds m_1 = \prod_{j \mathop \in J} X_j^{k_j}, m_2 = \prod_{j \mathop \in J} X_j^{k_j'}$ be two monomials. Their product is: :$\ds m_1 \circ m_2 = \paren {\prod_{j \mathop \in J} X_j^{k_j + k_j'} }$ If $k_j + k_j' \ne 0$ then either $k_j \ne 0$ or $k_j' \ne 0$ (or both are nonzero). Therefore if $k_j + k_j' \ne 0$ for infinitely many $j$, then either $m_1$ or $m_2$ is not a monomial. {{qed}} Category:Monomials \end{proof}
21209
\section{Set of Natural Numbers Equals Union of its Successor} Tags: Natural Numbers \begin{theorem} Let $\omega$ denote the set of natural numbers as defined by the von Neumann construction on a Zermelo universe $V$. Then: :$\bigcup \omega^+ = \omega$ \end{theorem} \begin{proof} We have that: :$\omega^+ = \omega \cup \set \omega$ and so: :$\omega \subseteq \omega^+$ {{qed|lemma}} By definition: :$\bigcup \omega^+ = \set {x: \exists y \in \omega^+: x \in y}$ Thus: :$x \in \bigcup \omega^+ \implies x \in \omega$ {{qed|lemma}} So by definition of set equality: :$\bigcup \omega^+ = \omega$ {{qed}} \end{proof}
21210
\section{Set of Natural Numbers Equals its Union} Tags: Natural Numbers \begin{theorem} Let $\omega$ denote the set of natural numbers as defined by the von Neumann construction on a Zermelo universe $V$. Then: :$\bigcup \omega = \omega$ \end{theorem} \begin{proof} By hypothesis, $\omega \in V$, where $V$ is a Zermelo universe. By the Axiom of Infinity we have that $\omega$ is a set. By the Axiom of Transitivity, $\omega$ is transitive. Hence: :$\bigcup \omega \subseteq \omega$ {{qed|lemma}} Let $n \in \omega$. Then by definition of the von Neumann construction: :$n + 1 = n \cup \set n$ That is: :$\exists X \in \omega: n \in X$ where in this case $X = n + 1$. Thus we have that: :$n \in \set {x: \exists X \in \omega: x \in X}$ and so by definition of union of class: :$n \in \bigcup \omega$ Thus we have that: :$\omega \subseteq \bigcup \omega$ which, together with: :$\bigcup \omega \subseteq \omega$ gives us, by set equality: :$\bigcup \omega = \omega$ {{qed}} \end{proof}
21211
\section{Set of Natural Numbers is Primitive Recursive} Tags: Natural Numbers, Primitive Recursive Functions \begin{theorem} The set of natural numbers $\N$ is primitive recursive. \end{theorem} \begin{proof} The characteristic function $\chi_\N: \N \to \N$ is defined as: :$\forall n \in \N: \chi_\N \left({n}\right) = 1$. So: : $\chi_\N \left({n}\right) = f^1_1 \left({n}\right)$ The constant function $f^1_1$ is primitive recursive. Hence the result. {{qed}} Category:Primitive Recursive Functions Category:Natural Numbers \end{proof}
21212
\section{Set of Natural Numbers is either Finite or Denumerable} Tags: Natural Numbers \begin{theorem} Let $S$ be a subset of the natural numbers $\N$. Then $S$ is either finite or denumerable. \end{theorem} \begin{proof} Let $\omega$ denote the set of natural numbers as defined by the von Neumann construction. By the Well-Ordering Principle, $\omega$ is well-ordered by the $\le$ relation. Thus from the Well-Ordering Principle, $S$ has a smallest element. Let this smallest element of $S$ be denoted $s_0$. Also from the Well-Ordering Principle, every subset of $S$ also has a smallest element. {{ProofWanted|establish a sequence where $s_k$ is the smallest element of $S_k$ where $S_k {{=}} S_{k - 1} \setminus \set {s_{k - 1} }$, and then build a mapping $f$ such that $\map f {s_k} \to \set {s_0, s_1, \ldots, s_{k - 1} }$ and show it's progressing, or something, but it's tedious.}} \end{proof}
21213
\section{Set of Non-Negative Real Numbers is not Well-Ordered by Usual Ordering} Tags: Real Numbers, Well-Orderings \begin{theorem} The set of non-negative real numbers $\R_{\ge 0}$ is not well-ordered under the usual ordering $\le$. \end{theorem} \begin{proof} {{AimForCont}} $\R_{\ge 0}$ is a well-ordered set. Then by definition, all subsets of $\R_{\ge 0}$ has a smallest element. Take the subset $\R_{> 0}$: :$\R_{> 0} = \left\{ {x \in \R_{\ge 0}: x > 0}\right\} = \R_{\ge 0} \setminus \left\{ {0}\right\}$ Suppose $x \in \R_{> 0}$ is a smallest element. Then $\dfrac x 2 \in \R_{> 0}$. But $\dfrac x 2 < x$, which contradicts the supposition that $x \in \R_{> 0}$ is a smallest element. Hence there can be no such smallest element. So by Proof by Contradiction, $\R_{\ge 0}$ is not well-ordered by $\le$. {{qed}} \end{proof}
21214
\section{Set of Non-Zero Natural Numbers is Primitive Recursive} Tags: Primitive Recursive Functions \begin{theorem} Let $\N^*$ be defined as $\N^* = \N \setminus \set 0$. The subset $\N^* \subset \N$ is primitive recursive. \end{theorem} \begin{proof} We have that the characteristic function $\chi_{\set 0}$ of $\set 0$ is primitive recursive. We note that: :If $n = 0$ then $\map {\chi_{\set 0} } n = 1$ therefore $\map {\chi_{\set 0} } {\map {\chi_{\set 0} } n} = 0$. :If $n > 0$ then $\map {\chi_{\set 0} } n = 0$ therefore $\map {\chi_{\set 0} } {\map {\chi_{\set 0} } n} = 1$. Thus $\map {\chi_{\set 0} } {\map {\chi_{\set 0} } n} = \map {\chi_{\N^*} } n$. So $\chi_{\N^*}$ is obtained by substitution from the primitive recursive function $\chi_{\set 0}$. Hence the result. {{qed}} Category:Primitive Recursive Functions \end{proof}
21215
\section{Set of Normal Subgroups of Group is Subsemigroup of Power Set Semigroup} Tags: Normal Subgroups, Semigroups, Subset Products \begin{theorem} Let $\struct {G, \circ}$ be a group. Let $\circ_\PP$ be the operation induced by $\circ$ on $\powerset G$, the power set of $G$. Let $\HH$ be the set of all normal subgroups of $\struct {G, \circ}$. Then the algebraic structure $\struct {\HH, \circ_\PP}$ is a subsemigroup of the algebraic structure $\struct {\powerset G, \circ_\PP}$. \end{theorem} \begin{proof} From Power Set of Group under Induced Operation is Semigroup, we have that $\struct {\powerset G, \circ_\PP}$ is a semigroup. Note that $\HH \subseteq \powerset G$. From Subset Product of Normal Subgroups is Normal, $\struct {\HH, \circ_\PP}$ is closed. By Subsemigroup Closure Test, $\struct {\HH, \circ_\PP}$ is a subsemigroup of $\struct {\powerset G, \circ_\PP}$. {{qed}} \end{proof}
21216
\section{Set of Normal Subgroups of Group is Subsemigroup of Power Set under Intersection} Tags: Semigroups, Power Set, Power Sets, Set Intersection, Normal Subgroups \begin{theorem} Let $\struct {G, \circ}$ be a group. Let $\HH$ be the set of all normal subgroups of $\struct {G, \circ}$. Then the algebraic structure $\struct {\HH, \cap}$ is a subsemigroup of the algebraic structure $\struct {\powerset G, \cap}$. \end{theorem} \begin{proof} From Power Set with Intersection is Commutative Monoid, we have that $\struct {\powerset G, \cap}$ is ''a fortiori'' a semigroup. Note that $\HH \subseteq \powerset G$. Let $H_1$ and $H_2$ be normal subgroups of $\struct {G, \circ}$. From Intersection of Normal Subgroups is Normal, $H_1 \cap H_2$ is also a normal subgroup of $\struct {G, \circ}$. Hence $\struct {\HH, \cap}$ is closed. By Subsemigroup Closure Test, $\struct {\HH, \cap}$ is a subsemigroup of $\struct {\powerset G, \cap}$. {{qed}} \end{proof}
21217
\section{Set of Numbers of form n - 1 over n is Bounded Above} Tags: Suprema \begin{theorem} Let $S$ be the subset of the set of real numbers $\R$ defined as: :$S = \set {\dfrac {n - 1} n: n \in \Z_{>0} }$ $S$ is bounded above with supremum $1$. $S$ has no greatest element. \end{theorem} \begin{proof} We have that: :$\dfrac {n - 1} n = 1 - \dfrac 1 n$ As $n > 0$ it follows from Reciprocal of Strictly Positive Real Number is Strictly Positive that $\dfrac 1 n > 0$. Thus $1 - \dfrac 1 n < 1$ and so $S$ is bounded above by $1$. Next it is to be shown that $1$ is the supremum of $S$. Suppose $x$ is the supremum of $S$ such that $x < 1$. Then: :$1 - x = \epsilon$ where $\epsilon \in \R_{>0}$. By the Archimedean Principle we have that: :$\exists m \in \Z_{>0}: n > \dfrac 1 \epsilon$ and so from Reciprocal Function is Strictly Decreasing: :$\exists m \in \Z_{>0}: \dfrac 1 n < \epsilon$ Thus: :$1 - \dfrac 1 m > 1 - \epsilon = x$ and so $x$ is not the supremum of $S$ after all. Thus $\sup S$ cannot be less than $1$. It follows that: :$\sup S = 1$ Next it is noted that: :$\forall n \in \Z_{>0}: \dfrac 1 n > 0$ and so: :$\forall x \in S: x < 1$ Thus as $\sup S \notin S$ it follows from Greatest Element is Supremum that $S$ has no greatest element. {{qed}} \end{proof}
21218
\section{Set of Odd Integers is Countably Infinite} Tags: Countable Sets, Odd Integers \begin{theorem} Let $\Bbb O$ be the set of odd integers. Then $\Bbb O$ is countably infinite. \end{theorem} \begin{proof} Let $f: \Bbb O \to \Z$ be the mapping defined as: :$\forall x \in \Bbb O: \map f x = \dfrac {x + 1} 2$ $f$ is well-defined as $x + 1$ is even and so $\dfrac {x + 1} 2 \in \Z$. Let $x, y \in \Bbb O$ such that $\map f x = \map f y$. Then: {{begin-eqn}} {{eqn | l = \map f x | r = \map f y | c = }} {{eqn | ll= \leadsto | l = \dfrac {x + 1} 2 | r = \dfrac {y + 1} 2 | c = Definition of $f$ }} {{eqn | ll= \leadsto | l = x + 1 | r = y + 1 | c = }} {{eqn | ll= \leadsto | l = x | r = y | c = }} {{end-eqn}} Thus $f$ is injective by definition. Consider the inverse $f^{-1}$. By inspection: :$\forall x \in \Z: \map {f^{-1} } x = 2 x - 1$ $f^{-1}$ is well-defined, and $2 x - 1$ is odd. Thus $f^{-1}$ is a mapping from $\Z$ to $\Bbb O$. Then: {{begin-eqn}} {{eqn | l = \map {f^{-1} } x | r = \map {f^{-1} } y | c = }} {{eqn | ll= \leadsto | l = 2 x - 1 | r = 2 y - 1 | c = Definition of $f^{-1}$ }} {{eqn | ll= \leadsto | l = 2 x | r = 2 y | c = }} {{eqn | ll= \leadsto | l = x | r = y | c = }} {{end-eqn}} Thus $f^{-1}$ is injective by definition. It follows by the Cantor-Bernstein-Schröder Theorem that there exists a bijection between $\Z$ and $\Bbb O$. {{qed}} \end{proof}
21219
\section{Set of Orbits forms Partition} Tags: Group Actions \begin{theorem} Let $G$ be a group. Let $X$ be a set. Let $G$ act on $X$. Then the set of orbits of the group action forms a partition of $X$. \end{theorem} \begin{proof} Follows from the Fundamental Theorem on Equivalence Relations. {{qed}} \end{proof}
21220
\section{Set of Order 3 Vectors under Cross Product does not form Ring} Tags: Examples of Rings, Vectors \begin{theorem} Let $S$ be the set of all vectors in a vector space of dimension $3$. Let $\times$ denote the cross product operation. Then the algebraic structure $\struct {S, +, \times}$ is not a ring. \end{theorem} \begin{proof} For $\struct {S, +, \times}$ to be a ring, it is a necessary condition that $\struct {S, \times}$ is a semigroup. For $\struct {S, \times}$ to be a semigroup, it is a necessary condition that $\times$ is associative on $S$. However, from Vector Cross Product is not Associative, this is not the case here. The result follows. {{qed}} \end{proof}
21221
\section{Set of Order m times n Matrices does not form Ring} Tags: Matrices, Examples of Rings \begin{theorem} Let $m, n \in \N_{>0}$ be non-zero natural numbers such that $m > n$. Let $S$ be the set of all matrices of order $m \times n$. Then the algebraic structure $\struct {S, +, \times}$ is not a ring. Note that $\times$ denotes conventional matrix multiplication. \end{theorem} \begin{proof} For $\struct {S, +, \times}$ to be a ring, it is a necessary condition that $\struct {S, \times}$ is a semigroup. For $\struct {S, \times}$ to be a semigroup, it is a necessary condition that $\struct {S, \times}$ is closed. That is: :$\forall x, y \in S: x \times y \in S$ Let $\mathbf A$ and $\mathbf B$ be elements of $S$. The matrix multiplication operation is defined on $\mathbf A$ and $\mathbf B$ {{iff}} $\mathbf A$ is of order $m \times p$ and $\mathbf A$ is of order $p \times n$, for some $m, n, p \in \N_{>0}$. That is, the second dimension of $\mathbf A$ must be equal to the first dimension of $\mathbf B$. But in this case, the second dimension of $\mathbf A$ is $n$, and the first dimension of $\mathbf B$ is $m$. As we have been given that $m \ne n$, it follows that matrix multiplication operation is not defined on $S$. Hence the result. {{qed}} \end{proof}
21222
\section{Set of Pairwise Disjoint Intervals is Countable} Tags: Countable Sets \begin{theorem} Let $X$ be a subset of $\powerset \R$ such that: :$(1): \quad X$ is pairwise disjoint: ::::$\forall A, B \in X: A \ne B \implies A \cap B = \O$. :$(2): \quad$ every element of $X$ contains an open interval: ::::$\forall A \in X: \exists x, y \in \R: x < y \land \openint x y \subseteq A$. Then $X$ is countable. \end{theorem} \begin{proof} By Between two Real Numbers exists Rational Number: :$\forall A \in X: \exists x, y \in \R, q \in \Q: x < y \land q \in \openint x y \subseteq A$ By the Axiom of Choice define a mapping $f: X \to \Q$: :$\forall A \in X: \map f A \in A$ First it needs to be shown that $f$ is an injection by definition. Let $A, B \in X$ such that: :$\map f A = \map f B$ By definition of $f$: :$\map f A \in A$ and $\map f B \in B$ By definition of intersection: :$\map f A \in A \cap B$ Then by definition of empty set: :$A \cap B \ne \O$ Thus by definition of pairwise disjoint: :$A = B$ Hence $f$ is an injection. By Set is Subset of Itself, $X$ is a subset of $X$. Thus by Cardinality of Image of Injection: :$\card X = \card {f^\to \sqbrk X}$ By definition of image: :$f^\to \sqbrk X \subseteq \Q$ By Rational Numbers are Countably Infinite: :$\Q$ is countable. Hence by Subset of Countable Set is Countable: :$f^\to \sqbrk X$ is countable. Thus by Set is Countable if Cardinality equals Cardinality of Countable Set the result: :$X$ is countable. {{qed}} \end{proof}
21223
\section{Set of Points at which Sequence of Measurable Functions does not Converge to Given Measurable Function is Measurable} Tags: Measurable Functions, Convergence \begin{theorem} Let $\struct {X, \Sigma, \mu}$ be a measure space. Let $f : X \to \R$ be a $\Sigma$-measurable function. For each $n \in \N$, let $f_n : X \to \R$ be a $\Sigma$-measurable function. Then: :$\ds \set {x \in X : \sequence {\map {f_n} x}_{n \in \N} \text { does not converge to } \map f x}$ is $\Sigma$-measurable. \end{theorem} \begin{proof} From Expression for Set of Points at which Sequence of Functions does not Converge to Given Function, we have: :$\ds \set {x \in X : \sequence {\map {f_n} x}_{n \in \N} \text { does not converge to } \map f x} = \bigcup_{k \mathop = 1}^\infty \bigcap_{N \mathop = 1}^\infty \bigcup_{n \mathop = N}^\infty \set {x \in X : \size {\map {f_n} x - \map f x} \ge \frac 1 k}$ From Pointwise Difference of Measurable Functions is Measurable, we have: :$f_n - f$ is $\Sigma$-measurable for each $n \in \N$. From Absolute Value of Measurable Function is Measurable, we have: :$\size {f_n - f}$ is $\Sigma$-measurable for each $n \in \N$. From Characterization of Measurable Functions, we have that: :$\ds \set {x \in X : \size {\map {f_n} x - \map f x} \ge \frac 1 k}$ is $\Sigma$-measurable for each $k, n \in \N$. Since $\sigma$-Algebras are closed under countable union, we have: :$\ds \bigcup_{n \mathop = N}^\infty \set {x \in X : \size {\map {f_n} x - \map f x} \ge \frac 1 k}$ is $\Sigma$-measurable for each $k, N \in \N$. Then, from Sigma-Algebra Closed under Countable Intersection: :$\ds \bigcap_{N \mathop = 1}^\infty \bigcup_{n \mathop = N}^\infty \set {x \in X : \size {\map {f_n} x - \map f x} \ge \frac 1 k}$ is $\Sigma$-measurable for each $k \in \N$. Again, since Since $\sigma$-Algebras are closed under countable union, we have: :$\ds \bigcup_{k \mathop = 1}^\infty \bigcap_{N \mathop = 1}^\infty \bigcup_{n \mathop = N}^\infty \set {x \in X : \size {\map {f_n} x - \map f x} \ge \frac 1 k}$ is $\Sigma$-measurable. So: :$\ds \set {x \in X : \sequence {\map {f_n} x}_{n \in \N} \text { does not converge to } \map f x}$ is $\Sigma$-measurable. {{qed}} Category:Measurable Functions Category:Convergence \end{proof}
21224
\section{Set of Points for which Measurable Function is Real-Valued is Measurable} Tags: Measurable Sets, Measurable Functions, Set of Points for which Measurable Function is Real-Valued is Measurable \begin{theorem} Let $\struct {X, \Sigma, \mu}$ be a measure space. Let $f: X \to \overline \R$ be a $\Sigma$-measurable. Then: :$\set {x \in X : \map f x \in \R}$ is $\Sigma$-measurable. \end{theorem} \begin{proof} Since $f$ is $\Sigma$-measurable, we have that: :for all $n \in \N$ the set $\set {x \in X : \map f x \le n}$ is $\Sigma$-measurable and: :for all $n \in \N$ the set $\set {x \in X : -n \le \map f x}$ is $\Sigma$-measurable. From $\sigma$-Algebra Closed under Countable Intersection, we have: :$\set {x \in X : -n \le \map f x \le n} = \set {x \in X : \map f x \le n} \cap \set {x \in X : -n \le \map f x}$ is $\Sigma$-measurable. Since $\sigma$-algebras are closed under countable union, we also have: :$\ds \bigcup_{n \mathop = 1}^\infty \set {x \in X : -n \le \map f x \le n}$ is $\Sigma$-measurable. We will finally show that: :$\ds \set {x \in X : \map f x \in \R} = \bigcup_{n \mathop = 1}^\infty \set {x \in X : -n \le \map f x \le n}$ which gives the claim. If $x \in X$ has $\map f x \in \R$, then we have: :$-\paren {\floor {\size {\map f x} } + 1} \le \map f x \le \floor {\size {\map f x} } + 1$ with: :$\floor {\size {\map f x} } + 1 \in \N$ So: :$\ds x \in \bigcup_{n \mathop = 1}^\infty \set {x \in X : -n \le \map f x \le n}$ Now, let: :$\ds x \in \bigcup_{n \mathop = 1}^\infty \set {x \in X : -n \le \map f x \le n}$ Then: :$-n \le \map f x \le n$ for some $n \in \N$, so certainly: :$\map f x \in \R$ So we have: :$\ds \set {x \in X : \map f x \in \R} = \bigcup_{n \mathop = 1}^\infty \set {x \in X : -n \le \map f x \le n}$ and hence the claim. {{qed}} Category:Measurable Sets Category:Measurable Functions Category:Set of Points for which Measurable Function is Real-Valued is Measurable \end{proof}
21225
\section{Set of Points for which Measurable Function is Real-Valued is Measurable/Corollary} Tags: Set of Points for which Measurable Function is Real-Valued is Measurable \begin{theorem} Let $\struct {X, \Sigma, \mu}$ be a measure space. Let $f: X \to \overline \R$ be a $\Sigma$-measurable. Then: :$\set {x \in X : \size {\map f x} = +\infty}$ is $\Sigma$-measurable. \end{theorem} \begin{proof} We have: :$\set {x \in X : \size {\map f x} = +\infty} = X \setminus \set {x \in X : \map f x \in \R}$ From Set of Points for which Measurable Function is Real-Valued is Measurable, we have: :$\set {x \in X : \map f x \in \R}$ is $\Sigma$-measurable. Since $\sigma$-algebras are closed under complementation, we have that: :$\set {x \in X : \size {\map f x} = +\infty}$ is $\Sigma$-measurable. {{qed}} Category:Set of Points for which Measurable Function is Real-Valued is Measurable \end{proof}
21226
\section{Set of Points on Line Segment is Infinite} Tags: Infinite Sets, Lines \begin{theorem} The set of points on a line segment is infinite. \end{theorem} \begin{proof} Let $S$ denote the set of points on a line segment. {{AimForCont}} $S$ is finite. Then there exists $n \in \N$ such that $S$ has $n$ elements. Let $s_1$ and $s_2$ be two arbitrary adjacent points in $S$. That is, such that there are no points in $S$ between $s_1$ and $s_2$. But there exists (at least) one point on the line segment between $s_1$ and $s_2$ which is not in $S$. Hence there must be more than $n$ elements of $S$. From that contradiction it follows by Proof by Contradiction that $S$ is not finite. {{Qed}} \end{proof}
21227
\section{Set of Polynomials over Infinite Set has Same Cardinality} Tags: Polynomial Theory \begin{theorem} Let $S$ be a set of infinite cardinality $\kappa$. Let $S \sqbrk x$ be the set of polynomial forms over $S$ in the indeterminate $x$. Then $S \sqbrk x$ has cardinality $\kappa$. \end{theorem} \begin{proof} Since $S \sqbrk x$ contains a copy of $S$ as constant polynomials, we have an injection $S \to S \sqbrk x$. We define an injection from $S \sqbrk x$ to the set $\FF$ of finite sequences over $S$ as follows: Each polynomial in $f \in S \sqbrk x$ is of the form: :$f = a_0 + a_1 x + a_2 x^2 + \dotsb + a_n x^n$ where $a_n$ is non-zero and each $a_i$ is in $S$. We send each polynomial $f$ to the sequence of its coefficients $\sequence {a_0, \dotsc, a_n}$. By the definition of equality of polynomials, this is injective. Now the set of finite sequences over $S$ is a countable union of sets of cardinality $\kappa$. From Cardinality of Infinite Union of Infinite Sets, $\FF$ has cardinality $\kappa$. Therefore there is a bijection $\FF \leftrightarrow S$. Composing this with the injection $S \sqbrk x \to \FF$, we have an injection $S \sqbrk x \to S$. So by the Cantor-Bernstein-Schröder Theorem there is a bijection $S \sqbrk x \leftrightarrow S$. Hence, we have: :$\card {S \sqbrk x} = \card S = \kappa$ {{qed}} Category:Polynomial Theory \end{proof}
21228
\section{Set of Polynomials over Integral Domain is Subring} Tags: Subrings, Polynomials, Polynomial Theory \begin{theorem} Let $\struct {R, +, \circ}$ be a commutative ring. Let $\struct {D, +, \circ}$ be an integral subdomain of $R$. Then $\forall x \in R$, the set $D \sqbrk x$ of polynomials in $x$ over $D$ is a subring of $R$. \end{theorem} \begin{proof} By application of the Subring Test: As $D$ is an integral domain, it has a unity $1_D$ and so $x = 1_D x$. Hence $x \in D \sqbrk x$ and so $D \sqbrk x \ne \O$. Let $p, q \in D \sqbrk x$. Then let: :$\ds p = \sum_{k \mathop = 0}^m a_k \circ x^k, q = \sum_{k \mathop = 0}^n b_k \circ x^k$ Thus: :$\ds -q = -\sum_{k \mathop = 0}^n b_k \circ x^k = \sum_{k \mathop = 0}^n \paren {-b_k} \circ x^k$ and so: :$q \in D \sqbrk x$ Thus as Polynomials Closed under Addition, it follows that: :$p + \paren {-q} \in D \sqbrk x$ Finally, from Polynomials Closed under Ring Product, we have that $p \circ q \in D \sqbrk x$. All the criteria of the Subring Test are satisfied. Hence the result. {{qed}} \end{proof}
21229
\section{Set of Positive Integers does not form Ring} Tags: Integers, Examples of Rings \begin{theorem} Let $\Z_{\ge 0}$ denote the set of positive integers. Then the algebraic structure $\struct {\Z_{\ge 0}, +, \times}$ does not form a ring. \end{theorem} \begin{proof} For $\struct {\Z_{\ge 0}, +, \times}$ to be a ring, it is necessary for the algebraic structure $\struct {\Z_{\ge 0}, +}$ to form a group. But from the corollary to Natural Numbers under Addition do not form Group: :$\struct {\Z_{\ge 0}, +}$ is not a group. {{qed}} \end{proof}
21230
\section{Set of Prime Numbers is Primitive Recursive} Tags: Primitive Recursive Functions \begin{theorem} The set $\Bbb P$ of prime numbers is primitive recursive. \end{theorem} \begin{proof} A prime number is defined as an element of $\N$ with '''exactly two''' positive divisors. So, we have that $n > 0$ is prime {{iff}} $\map \tau n = 2$, where $\tau: \N \to \N$ is the divisor counting function. Thus we can define the characteristic function of the set of prime numbers $\Bbb P$ as: :$\forall n > 0: \map {\chi_\Bbb P} n := \map {\chi_{\operatorname {eq} } } {\map \tau n, 2}$ Now we let $g: \N^2 \to \N$ be the function given by: :$\map g {n, z} = \begin{cases} 0 & : z = 0 \\ \ds \sum_{y \mathop = 1}^z \map {\operatorname {div} } {n, y} & : z > 0 \end{cases}$ As: :$\operatorname {div}$ is primitive recursive :$\ds \sum_{y \mathop = 1}^z$ is primitive recursive it follows that $g$ is primitive recursive. Then for $n > 0$: :$\ds \map g {n, n} = \sum_{y \mathop = 1}^n \map {\operatorname {div} } {n, y} = \map \tau n$ and from Divisor Counting Function is Primitive Recursive we have that $g$ is primitive recursive. Then let $h: \N \to \N$ be the function defined as: :$\map h n = \map g {n, n}$ which is also primitive recursive. So we have, for all $n \in \N$: :$\map {\chi_\Bbb P} n = \map {\chi_{\operatorname {eq} } } {\map h n, 2}$ Hence the result. {{qed}} Category:Primitive Recursive Functions \end{proof}
21231
\section{Set of Rational Cuts forms Ordered Field} Tags: Totally Ordered Fields, Cuts \begin{theorem} Let $\RR$ denote the set of rational cuts. Let $\struct {\RR, +, \times, \le}$ denote the ordered structure formed from $\RR$ and: :the operation $+$ of addition of cuts :the operation $\times$ of multiplication of cuts :the ordering $\le$ of cuts. Then $\struct {\RR, + \times, \le}$ is an ordered field. \end{theorem} \begin{proof} We demonstrate that $\struct {\RR, +, \times}$ is a field by showing it is a subfield of the structure $\struct {\CC, +, \times}$, where $\CC$ denotes the set of all cuts. We do this by establishing that all $4$ criteria of the Subfield Test are satisfied. We note that $0^* \in \RR$, where $0^*$ is the rational cut associated with the (rational) number $0$: :$0^* = \set {r \in \Q: r < 0}$ So $\RR \ne \O$. Thus criterion $(1)$ is satisfied. {{qed|lemma}} {{finish}} {{qed}} \end{proof}
21232
\section{Set of Rational Numbers Strictly between Zero and One has no Greatest or Least Element} Tags: Rational Numbers, Infima, Suprema \begin{theorem} Let $S \subseteq \Q$ be the subset of the set of rational numbers defined as: :$S = \set {r \in \Q: 0 < r < 1}$ Then $S$ has no greatest or smallest element. However, $S$ has a supremum $1$ and an infimum $0$. \end{theorem} \begin{proof} We have that: :$\forall r \in S: 0 < r$ and: :$\forall r \in S: r < 1$ Hence $0$ and $1$ are lower and upper bounds of $S$ respectively. Let $s \in S$. Then $s \in \Q: 0 < s < 1$. {{AimForCont}} $s$ is the greatest element of $S$. But then we have: :$0 < s < \dfrac {s + 1} 2 < 1$ and so $\dfrac {s + 1} 2 \in S$ but $s < \dfrac {s + 1} 2$. This contradicts our assertion that $s$ is the greatest element of $S$. {{AimForCont}} $s$ is the smallest element of $S$. But then we have: :$0 < \dfrac s 2 < s < 1$ and so $\dfrac s 2 \in S$ but $\dfrac s 2 < s$. This contradicts our assertion that $s$ is the smallest element of $S$. So $S$ cannot have either a greatest or smallest element. Let $H$ be such that $0 < H < 1$. Suppose $H$ is a smaller upper bound of $S$ than $1$. Then from Between two Real Numbers exists Rational Number there exists a rational number $r$ such that $H < r < 1$. But then $r$ is a rational number between $0$ and $1$ which is greater than $H$. Hence $H$ cannot be an upper bound of $S$. Thus, by definition, $1$ is the supremum of $S$ Suppose $H$ is a greater lower bound of $S$ than $0$. Then from Between two Real Numbers exists Rational Number there exists a rational number $r$ such that $0 < r < H$. But then $r$ is a rational number between $0$ and $1$ which is smaller than $H$. Hence $H$ cannot be a lower bound of $S$. Thus, by definition, $0$ is the infimum of $S$. {{Qed}} \end{proof}
21233
\section{Set of Rational Numbers is not Closed in Reals} Tags: Real Number Line with Euclidean Topology, Rational Numbers, Real Number Space \begin{theorem} Let $\Q$ be the set of rational numbers. Let $\struct {\R, \tau}$ denote the real number line with the usual (Euclidean) topology. Then $\Q$ is not closed in $\R$. \end{theorem} \begin{proof} Let $\alpha \in \R \setminus \Q$. Let $I := \openint a b$ be an open interval in $\R$ such that $\alpha \in I$. By Between two Real Numbers exists Rational Number: :$\exists \beta \in \Q: \beta \in I$. Thus $I$ contains elements of $\Q$ and so $\R \setminus \Q$ is not open in $\R$. Thus by definition, $\Q$ is not closed in $\R$. {{qed}} \end{proof}
21234
\section{Set of Rational Numbers is not G-Delta Set in Reals} Tags: G-Delta Sets, Real Number Line with Euclidean Topology, Rational Numbers, Real Number Space \begin{theorem} Let $\Q$ be the set of rational numbers. Let $\struct {\R, \tau}$ denote the real number line with the usual (Euclidean) topology. Then $\Q$ is not a $G_\delta$ set in $\R$. \end{theorem} \begin{proof} {{AimForCont}} $\Q$ is a $G_\delta$ set in $\R$. Let $\Q = \ds \bigcap_{i \mathop \in \N} V_i$. Since Rational Numbers are Countably Infinite, there exists an enumeration of $\Q$. Write $\Q = \set {q_i}_{i \mathop \in \N}$. Define $U_i = V_i \setminus \set {q_i}$. We show that $U_i$ is dense in $\R$. :Let $A \subseteq \R$ be an open set of $\struct {\R, \tau}$. :From Basis for Euclidean Topology on Real Number Line, the set of all open real intervals of $\R$ form a basis for $\struct {\R, \tau}$. :So there exists $\openint a b \subseteq A$ for some $a < b$. :One of $q_i \le a$, $q_i > a$ must hold. :Suppose $q_i \le a$. :By Between two Real Numbers exists Rational Number we have: ::$\exists r \in Q: a < r < b$. :Since $r > a \ge q_i$: ::$r \in \Q \setminus \set {q_i} \subseteq V_i \setminus \set {q_i} = U_i$. :Suppose $q_i > a$. :By Between two Real Numbers exists Rational Number we have: ::$\exists r \in Q: a < r < \min \set {q_i, b}$. :Since $r < q_i$: ::$r \in \Q \setminus \set {q_i} \subseteq V_i \setminus \set {q_i} = U_i$. :In both cases we see that $r \in \openint a b \cap U_i \subseteq A \cap U_i$. :Therefore $A \cap U_i \ne \O$. :Thus $U_i$ is dense in $\R$. Also we have: {{begin-eqn}} {{eqn | l = \bigcap_{i \mathop \in \N} U_i | r = \bigcap_{i \mathop \in \N} \paren {V_i \setminus \set {q_i} } }} {{eqn | r = \paren {\bigcap_{i \mathop \in \N} V_i} \setminus \paren {\bigcup_{i \mathop \in \N} \set {q_i} } | c = Difference with Union }} {{eqn | r = \Q \setminus \Q }} {{eqn | r = \O | c = Set Difference with Self is Empty Set }} {{end-eqn}} By Real Number Line is Complete Metric Space and Baire Category Theorem, $\struct {\R, \tau}$ is a Baire space. By definition of Baire Space, $\ds \bigcap_{i \mathop \in \N} U_i$ is dense in $\R$. But Empty Set is Nowhere Dense. This is a contradiction. Hence $\Q$ is not a $G_\delta$ set in $\R$. {{qed}} \end{proof}
21235
\section{Set of Rational Numbers whose Numerator Divisible by p is Closed under Addition} Tags: Prime Numbers, Algebraic Closure, Rational Addition, Rational Numbers \begin{theorem} Let $p$ be a prime number. Let $A_p$ be the set of all rational numbers which, when expressed in canonical form has a numerator which is divisible by $p$. Then $A_p$ is closed under rational addition. \end{theorem} \begin{proof} Let $a, b \in A_p$. Then $a = \dfrac {p n_1} {d_1}, b = \dfrac {p n_1} {d_1}$ where: :$n_1, n_2 \in \Z$ :$d_1, d_2 \in \Z_{>0}$ :$p n_1 \perp d_1, p n_2 \perp d_2$ Then: {{begin-eqn}} {{eqn | l = a + b | r = \frac {p n_1} {d_1} + \frac {p n_2} {d_2} | c = }} {{eqn | r = \frac {p n_1 d_2 + p n_2 d_1} {d_1 d_2} | c = {{Defof|Rational Addition}} }} {{eqn | r = \frac {p \paren {n_1 d_2 + n_2 d_1} } {d_1 d_2} | c = }} {{end-eqn}} From Euclid's Lemma for Prime Divisors, if $p \divides d_1 d_2$ then either $p \divides d_1$ or $p \divides d_2$. But neither of these is the case, so $p \nmid d_1 d_2$. Hence by Prime not Divisor implies Coprime: :$p \perp d_1 d_2$ where $\perp$ denotes coprimeness. So when $\dfrac {p \paren {n_1 d_2 + n_2 d_1} } {d_1 d_2}$ is expressed in canonical form, $p$ will still be a divisor of the numerator. Hence the result. {{qed}} \end{proof}
21236
\section{Set of Rational Numbers whose Numerator Divisible by p is Closed under Multiplication} Tags: Prime Numbers, Algebraic Closure, Rational Multiplication, Rational Numbers \begin{theorem} Let $p$ be a prime number. Let $A_p$ be the set of all rational numbers which, when expressed in canonical form has a numerator which is divisible by $p$. Then $A_p$ is closed under rational multiplication. \end{theorem} \begin{proof} Let $a, b \in A_p$. Then $a = \dfrac {p n_1} {d_1}, b = \dfrac {p n_1} {d_1}$ where: :$n_1, n_2 \in \Z$ :$d_1, d_2 \in \Z_{>0}$ :$p n_1 \perp d_1, p n_2 \perp d_2$ Then: {{begin-eqn}} {{eqn | l = a \times b | r = \frac {p n_1} {d_1} \times \frac {p n_2} {d_2} | c = }} {{eqn | r = \frac {p n_1 p n_2} {d_1 d_2} | c = {{Defof|Rational Multiplication}} }} {{eqn | r = \frac {p^2 \paren {n_1 n_2} } {d_1 d_2} | c = }} {{end-eqn}} From Euclid's Lemma for Prime Divisors, if $p \divides d_1 d_2$ then either $p \divides d_1$ or $p \divides d_2$. But neither of these is the case, so $p \nmid d_1 d_2$. Hence by Prime not Divisor implies Coprime: :$p \perp d_1 d_2$ where $\perp$ denotes coprimeness. So when $\dfrac {p^2 \paren {n_1 n_2} } {d_1 d_2}$ is expressed in canonical form, $p$ will still be a divisor of the numerator. Hence the result. {{qed}} \end{proof}
21237
\section{Set of Rationals Greater than Root 2 has no Smallest Element} Tags: Rational Numbers \begin{theorem} Let $B$ be the set of all positive rational numbers $p$ such that $p^2 > 2$. Then $B$ has no smallest element. \end{theorem} \begin{proof} {{AimForCont}} $p \in B$ is the smallest element of $B$. Then by definition of $B$: :$p^2 > 2$ Let $q = p - \dfrac {p_2 - 2} {2 p}$. Then $q > p$, and: {{begin-eqn}} {{eqn | l = q | r = p - \dfrac {p_2 - 2} {2 p} | c = }} {{eqn | r = \dfrac p 2 + \dfrac 1 p | c = }} {{end-eqn}} Hence: :$0 < q < p$ and so: {{begin-eqn}} {{eqn | l = q^2 | r = p^2 - \dfrac {p^2 - 2} + \paren {\dfrac {p^2 - 2} {2 p} }^2 | c = }} {{eqn | o = > | r = p^2 - \paren {p^2 - 2} | c = }} {{eqn | r = 2 | c = }} {{end-eqn}} That means $q \in B$. This contradicts our assertion that $p$ is the smallest element of $B$. It follows that $B$ can have no smallest element. {{qed}} \end{proof}
21238
\section{Set of Rationals Less than Root 2 has no Greatest Element} Tags: Rational Numbers \begin{theorem} Let $A$ be the set of all positive rational numbers $p$ such that $p^2 < 2$. Then $A$ has no greatest element. \end{theorem} \begin{proof} {{AimForCont}} $p \in A$ is the greatest element of $A$. Then by definition of $A$: :$p^2 < 2$ Let $h \in \Q$ be a rational number such that $0 < h < 1$ such that: :$h < \dfrac {2 - p^2} {2 p + 1}$ This is always possible, because by definition $2 - p^2 > 0$ and $2 p + 1 > 0$. Let $q = p + h$. Then $q > p$, and: {{begin-eqn}} {{eqn | l = q^2 | r = p^2 + \paren {2 p + h} h | c = }} {{eqn | o = < | r = p^2 + \paren {2 p + 1} h | c = }} {{eqn | o = < | r = p^2 + \paren {2 - p^2} | c = }} {{eqn | r = 2 | c = }} {{end-eqn}} That means $q \in A$. This contradicts our assertion that $p$ is the greatest element of $A$. It follows that $A$ can have no greatest element. {{qed}} \end{proof}
21239
\section{Set of Reciprocals of Positive Integers is Nowhere Dense in Reals} Tags: Analysis, Real Number Line with Euclidean Metric, Integer Reciprocal Space, Real Number Space, Denseness \begin{theorem} Let $N$ be the set defined as: :$N := \set {\dfrac 1 n: n \in \Z_{>0} }$ where $\Z_{>0}$ is the set of (strictly) positive integers. Let $\R$ denote the real number line with the usual (Euclidean) metric. Then $N$ is nowhere dense in $\R$. \end{theorem} \begin{proof} From Zero is Limit Point of Integer Reciprocal Space, the only limit point of $N$ is $0$. Hence: :$\map \cl N = \set {\dfrac 1 n: n \in \Z_{>0} } \cup \set 0$ where $\map \cl N$ denotes the closure of $N$ in $\R$. Trivially, $\map \cl N$ contains no open real intervals. Hence no subset of $\map \cl N$ is open in $\R$. Hence the union of all the subset of $\map \cl N$ which are open in $\R$ is empty. That is, by definition, the interior of $N$ is empty. That is: :$\paren {\map \cl N}^\circ = \O$ and the result follows by definition of nowhere dense. {{qed}} \end{proof}
21240
\section{Set of Relations can be Ordered by Inclusion} Tags: Subsets, Order Theory \begin{theorem} Let $S \times T$ be the product of two sets. Let $\RR$ be a set of relations on $S \times T$. Then $\RR$ can be ordered by inclusion. \end{theorem} \begin{proof} Let $R$ be a relation on $S \times T$. By the definition of relation, $R$ is associated with a subset $R \subseteq S \times T$. Thus $\RR$ is a subset of the power set $\powerset {S \times T}$. The result follows from Subset Relation is Ordering. {{qed}} \end{proof}
21241
\section{Set of Ring Elements forming Zero Product with given Element is Ideal} Tags: Ideal Theory \begin{theorem} Let $\struct {R, +, \circ}$ be a commutative ring whose zero is $0_R$. Let $a \in R$ be an arbitrary element of $R$. Let $A$ be the subset of $R$ defined as: :$A = \set {x \in R: x \circ a = 0_R}$ Then $A$ is an ideal of $A$. \end{theorem} \begin{proof} By definition of ring zero: :$\forall x \in R: x \circ 0_R = 0_R$ Hence $0_R \in A$ and so $A \ne \O$. Let $a, b \in A$. {{begin-eqn}} {{eqn | q = \forall x \in R | l = x \circ b | r = 0_R | c = }} {{eqn | ll= \leadsto | l = -\paren {x \circ b} | r = 0_R | c = }} {{eqn | ll= \leadsto | l = x \circ \paren {-b} | r = 0_R | c = }} {{end-eqn}} Thus: {{begin-eqn}} {{eqn | q = \forall x \in R | l = x \circ a | r = 0_R | c = }} {{eqn | lo= \land | l = x \circ \paren {-b} | r = 0_R | c = }} {{eqn | ll= \leadsto | l = \paren {x \circ a} + \paren {x \circ \paren {-b} } | r = 0_R | c = as $0_R$ is identity of $\struct {R, +}$ }} {{eqn | ll= \leadsto | l = x \circ \paren {a + \paren {-b} } | r = 0_R | c = {{Ring-axiom|D}} }} {{eqn | ll= \leadsto | l = a + \paren {-b} | o = \in | r = A | c = Definition of $A$ }} {{end-eqn}} Then: {{begin-eqn}} {{eqn | q = \forall x \in R | l = x \circ a | r = 0_R | c = }} {{eqn | lo= \land | l = x \circ b | r = 0_R | c = }} {{eqn | ll= \leadsto | l = \paren {x \circ a} \circ b | r = 0_R | c = {{Defof|Ring Zero}} }} {{eqn | ll= \leadsto | l = x \circ \paren {a \circ b} | r = 0_R | c = {{Ring-axiom|M1}} }} {{eqn | ll= \leadsto | l = a \circ b | o = \in | r = A | c = Definition of $A$ }} {{end-eqn}} Hence the result, from Test for Ideal: {{qed}} \end{proof}
21242
\section{Set of Sequence Codes is Primitive Recursive} Tags: Primitive Recursive Functions \begin{theorem} Let $\operatorname{Seq}$ be the set of all code numbers of finite sequences in $\N$. Then $\operatorname{Seq}$ is primitive recursive. \end{theorem} \begin{proof} By the definition of a primitive recursive set, it is sufficient to show that the characteristic function $\chi_{\operatorname{Seq}}$ of $\operatorname{Seq}$ is primitive recursive. Let $p: \N \to \N$ be the prime enumeration function. Let $\map {\operatorname{len} } n$ be the length of $n$. We note that $\map {\chi_{\operatorname{Seq}} } n = 1$ {{iff}} $\map p y$ divides $n$ for $1 \le y \le \map {\operatorname{len} } n$. That is, {{iff}} $\map {\operatorname{div} } {n, \map p y} = 1$ for $1 \le y \le \map {\operatorname{len} } n$, where $\operatorname{div}$ is the divisor relation. We then see that $\map {\operatorname{div} } {n, \map p y} = 1$ for $1 \le y \le \map {\operatorname{len} } n$ {{iff}} their product equals $1$. So we can define $\chi_{\operatorname{Seq}}$ by: :$\map {\chi_{\operatorname{Seq}} } n = \begin{cases} \ds \prod_{y \mathop = 1}^{\map {\operatorname{len} } n} \map {\operatorname{div} } {n, \map p y} & : n > 1 \\ 0 & : \text{otherwise} \end{cases}$ Then we define $g: \N^2 \to \N$ as: :$\map g {n, z} = \begin{cases} 1 & : z = 0 \\ \ds \prod_{y \mathop = 1}^z \map {\operatorname{div} } {n, \map p y} & : z \ne 0 \end{cases}$ We then apply Bounded Product is Primitive Recursive to the primitive recursive function $\map {\operatorname{div} } {n, \map p y}$, and see that $g$ is primitive recursive. Finally, we have that: :$\map {\chi_{\operatorname{Seq}} } n = \begin{cases} \map g {n, \map {\operatorname{len} } n} & : n > 1 \\ 0 & : \text{otherwise} \end{cases}$ is obtained by substitution from: * the primitive recursive function $\operatorname{len}$ * the primitive recursive function $g$ * the primitive recursive relation $>$ * the constant $1$ So $\chi_{\operatorname{Seq}}$ is primitive recursive. Hence the result. {{qed}} Category:Primitive Recursive Functions \end{proof}
21243
\section{Set of Sets can be Defined as Family} Tags: Indexed Families, Set Systems \begin{theorem} Let $\Bbb S$ be a set of sets. Then $\Bbb S$ can be defined as an indexed family of sets. \end{theorem} \begin{proof} Let $S: \Bbb S \to \Bbb S$ denote the identity mapping on $\Bbb S$: :$\forall i \in \Bbb S: S_i = i$ where we use $S_i$ to mean the image of $i$ under $S$: :$S_i := \map S i$ Then we can consider $S$ as an indexing function from $\Bbb S$ to $\Bbb S$. Hence in this case $\Bbb S$ is at the same time both: :an indexing set and: :the set indexed by itself. It follows that each of the sets $i \in \Bbb S$ is both: :an index and: :a term $S_i$ of the family of elements of $\Bbb S$ indexed by $\Bbb S$. Thus we would write $\Bbb S$ as: :$\family {S_i}_{i \mathop \in \Bbb S}$ {{qed}} \end{proof}
21244
\section{Set of Singletons is Smallest Basis of Discrete Space} Tags: Discrete Spaces, Discrete Topology \begin{theorem} Let $T = \struct {S, \tau}$ be a discrete topological space. Let $\BB = \set {\set x : x \in S}$. Then $\BB$ is the smallest basis of $T$. That is: :$\BB$ is a basis of $T$ and: :for every basis $\CC$ of $T$, $\BB \subseteq \CC$. \end{theorem} \begin{proof} By Basis for Discrete Topology $\BB$ is a basis of $T$. It remains to be shown that $\BB$ is the smallest basis of $T$. Let $\CC$ be a basis of $T$. Let $A \in \BB$. By definition of the set $\BB$: :$\exists x \in S: A = \set x$ By definition of basis: :$\exists B \in \CC: x \in B \subseteq A$ Then by Singleton of Element is Subset: :$\set x \subseteq B$ Hence $B = A$ by definition of set equality. Thus $A \in \CC$. {{qed}} \end{proof}
21245
\section{Set of Strictly Positive Real Numbers has no Smallest Element} Tags: Real Analysis \begin{theorem} Let $\R_{>0}$ denote the set of strictly positive real numbers. Then $\R_{>0}$ has no smallest element. \end{theorem} \begin{proof} {{AimForCont}} $\R_{>0}$ has a smallest element. Let $m$ be that smallest element. Then we have that: :$0 < \dfrac m 2 < m$ But as $0 < \dfrac m 2$ it follows that $\dfrac m 2 \in \R_{>0}$. This contradicts our assertion that $m$ is the smallest element of $\R_{>0}$. Hence the result by Proof by Contradiction. {{qed}} \end{proof}
21246
\section{Set of Subfields forms Complete Lattice} Tags: Lattice Theory, Rings, Fields, Subfields \begin{theorem} Let $\struct {F, +, \circ}$ be a field. Let $\mathbb F$ be the set of all subfields of $F$. Then $\struct {\mathbb F, \subseteq}$ is a complete lattice. \end{theorem} \begin{proof} Let $\O \subset \mathbb S \subseteq \mathbb F$. By Intersection of Subfields is Largest Subfield Contained in all Subfields: :$\bigcap \mathbb S$ is the largest subfield of $F$ contained in each of the elements of $\mathbb S$. By Intersection of Subfields Containing Subset is Smallest: :The intersection of the set of all subfields of $F$ containing $\bigcup \mathbb S$ is the smallest subfield of $F$ containing $\bigcup \mathbb S$. Thus: :Not only is $\bigcap \mathbb S$ a lower bound of $\mathbb S$, but also the largest, and therefore an infimum. :The supremum of $\mathbb S$ is the intersection of the set of all subfields of $R$. Therefore $\struct {\mathbb F, \subseteq}$ is a complete lattice. {{qed}} \end{proof}
21247
\section{Set of Subgroups forms Complete Lattice} Tags: Lattice Theory, Examples of Complete Lattices, Subgroups \begin{theorem} Let $\struct {G, \circ}$ be a group, and let $\mathbb G$ be the set of all subgroups of $G$. Then $\struct {\mathbb G, \subseteq}$ is a complete lattice. \end{theorem} \begin{proof} Let $\O \subset \mathbb H \subseteq \mathbb G$. By Intersection of Subgroups: General Result, $\bigcap \mathbb H$ is the largest subgroup of $G$ contained in each of the elements of $\mathbb H$. Thus, not only is $\bigcap \mathbb H$ a lower bound of $\mathbb H$, but also the largest, and therefore an infimum. The supremum of $\mathbb H$ is the smallest subgroup of $G$ containing $\bigcup \mathbb H$. Therefore $\struct {\mathbb G, \subseteq}$ is a complete lattice. {{qed}} \end{proof}
21248
\section{Set of Subgroups of Abelian Group form Subsemigroup of Structure Induced on Power Set} Tags: Abelian Groups, Subset Products \begin{theorem} Let $\struct {G, \circ}$ be an abelian group. Let $\circ_\PP$ be the operation induced on $\powerset G$, the power set of $G$. Let $\SS$ be the set of all subgroups of $G$. Then $\struct {\SS, \circ_\PP}$ is a subsemigroup of $\struct {\powerset G, \circ_\PP}$. \end{theorem} \begin{proof} A fortiori, an abelian group is a commutative semigroup. Also a fortiori, a subgroup of $G$ is also a subsemigroup of $G$. The result then follows directly from Set of Subsemigroups of Commutative Semigroup form Subsemigroup of Structure Induced on Power Set. {{qed}} \end{proof}
21249
\section{Set of Subrings forms Complete Lattice} Tags: Lattice Theory, Rings, Subrings \begin{theorem} Let $\struct {K, +, \circ}$ be a ring. Let $\mathbb K$ be the set of all subrings of $K$. Then $\struct {\mathbb K, \subseteq}$ is a complete lattice. \end{theorem} \begin{proof} Let $\P \subset \mathbb S \subseteq \mathbb K$. By Intersection of Subrings is Largest Subring Contained in all Subrings: :$\bigcap \mathbb S$ is the largest subring of $K$ contained in each of the elements of $\mathbb S$. By Intersection of Subrings Containing Subset is Smallest: :The intersection of the set of all subrings of $K$ containing $\bigcup \mathbb S$ is the smallest subring of $K$ containing $\bigcup \mathbb S$. Thus: :Not only is $\bigcap \mathbb S$ a lower bound of $\mathbb S$, but also the largest, and therefore an infimum. :The supremum of $\mathbb S$ is the intersection of the set of all subrings of $K$. Therefore $\struct {\mathbb K, \subseteq}$ is a complete lattice. {{qed}} \end{proof}
21250
\section{Set of Subsemigroups of Commutative Semigroup form Subsemigroup of Structure Induced on Power Set} Tags: Commutative Semigroups, Subset Products \begin{theorem} Let $\struct {S, \circ}$ be a commutative semigroup. Let $\circ_\PP$ be the operation induced on $\powerset S$, the power set of $S$. Let $\TT$ be the set of all subsemigroups of $S$. Then $\struct {\TT, \circ_\PP}$ is a subsemigroup of $\struct {\powerset S, \circ_\PP}$. \end{theorem} \begin{proof} First we establish that from Power Set of Semigroup under Induced Operation is Semigroup: :$\struct {\powerset S, \circ_\PP}$ is a semigroup. From Subset Product within Commutative Structure is Commutative: :$\struct {\powerset S, \circ_\PP}$ is a commutative semigroup. Let $A$ and $B$ be arbitrary subsemigroups of $S$. As $A$ and $B$ are subsemigroups of $S$, they themselves are closed for $\circ$. That is: :$\forall x, y \in A: x \circ y \in A$ and: :$\forall x, y \in B: x \circ y \in B$ By definition of operation induced on $\powerset S$: :$A \circ_\PP B = \set {a \circ b: a \in A, b \in B}$ We are to show that: :$\forall x, y \in A \circ_\PP B: x \circ y \in A \circ_\PP B$ Let $x, y \in A \circ_\PP B$ such that $x = a_x \circ b_x$, $y = a_y \circ b_y$. We have: {{begin-eqn}} {{eqn | l = x \circ y | r = \paren {a_x \circ b_x} \circ \paren {a_y \circ b_y} | c = }} {{eqn | r = a_x \circ \paren {b_x \circ a_y} \circ b_y | c = {{SemigroupAxiom|1}} }} {{eqn | r = a_x \circ \paren {a_y \circ b_x} \circ b_y | c = $\circ$ is commutative }} {{eqn | r = \paren {a_x \circ a_y} \circ \paren {b_x \circ b_y} | c = {{SemigroupAxiom|1}} }} {{eqn | o = \in | r = A \circ_\PP B | c = {{SemigroupAxiom|0}}: both $A$ and $B$ are closed for $\circ$ }} {{end-eqn}} That is: :$x, y \in A \circ_\PP B \implies x \circ y \in A \circ_\PP B$ and $A \circ_\PP B$ is seen to be $\struct {\TT, \circ_\PP}$. Hence the result. {{qed}} \end{proof}
21251
\section{Set of Subset of Reals with Cardinality less than Continuum has not Interval in Union Closure} Tags: Infinite Sets \begin{theorem} Let $\BB$ be a set of subsets of $\R$, the set of all real numbers. Let: :$\card \BB < \mathfrak c$ where :$\card \BB$ denotes the cardinality of $\BB$ :$\mathfrak c = \card \R$ denotes continuum. Let $\FF = \set {\bigcup \GG: \GG \subseteq \BB}$. Then: :$\exists x, y \in \R: x < y \land \hointr x y \notin \FF$ \end{theorem} \begin{proof} Define: :$ Z = \leftset {x \in \R: \exists U \in \FF: x}$ is local minimum in $\rightset U$ By Set of Subsets of Reals with Cardinality less than Continuum Cardinality of Local Minimums of Union Closure less than Continuum: :$\card Z < \mathfrak c$ Then by Cardinalities form Inequality implies Difference is Nonempty: :$\R \setminus Z \ne \O$ Hence by definition of empty set: :$\exists z: z \in \R \setminus Z$ By definition of difference: :$z \in \R \land z \notin Z$ Thus $z < z + 1$. We will show that $z$ is local minimum in $\hointr z {z + 1}$. Thus: :$z \in \hointr z {z + 1}$ Hence: :$z - 1 < z$ Thus: :$\openint {z - 1} z \cap \hointr z {z + 1} = \O$ Then by definition $z$ is a local minimum in $\hointr z {z + 1}$. Because $z \notin Z$: :$\hointr z {z + 1} \notin \FF$ {{qed}} \end{proof}
21252
\section{Set of Subsets is Cover iff Set of Complements is Free} Tags: Set Theory, Definition Equivalences \begin{theorem} Let $S$ be a set. Let $\CC$ be a set of sets. Then $\CC$ is a cover for $S$ {{iff}} $\set {\relcomp S X: X \in \CC}$ is free. \end{theorem} \begin{proof} Let $S$ be a set. \end{proof}
21253
\section{Set of Subsets of Element of Minimally Inductive Class is Finite} Tags: Minimally Inductive Classes, Finite Sets \begin{theorem} Let $M$ be a class which is minimally inductive under a progressing mapping $g$. Let $x \in M$. Let $S$ be the set of all $y \in M$ such that $y \subseteq x$. Then $S$ is finite. \end{theorem} \begin{proof} The proof proceeds by general induction. For all $x \in M$, let $\map P x$ be the proposition: :$\set {y \in M: y \subseteq x}$ is finite. \end{proof}
21254
\section{Set of Subsets of Reals with Cardinality less than Continuum Cardinality of Local Minimums of Union Closure less than Continuum} Tags: Infinite Sets \begin{theorem} Let $\BB$ be a set of subsets of $\R$. Let: :$\size \BB < \mathfrak c$ where :$\size \BB$ denotes the cardinality of $\BB$ :$\mathfrak c = \size \R$ denotes continuum. Let :$X = \leftset {x \in \R: \exists U \in \set {\bigcup \GG: \GG \subseteq \BB}: x}$ is local minimum in $\rightset U$ Then: :$\size X < \mathfrak c$ \end{theorem} \begin{proof} We will prove that: :$(1): \quad \size \BB \aleph_0 < \mathfrak c$ where $\aleph_0 = \size \N$ by Aleph Zero equals Cardinality of Naturals. In the case when $\size \BB = \mathbf 0$ we have by Zero of Cardinal Product is Zero: :$\size \BB \aleph_0 = \mathbf 0 < \mathfrak c$ In the case when $\mathbf 0 < \size \BB < \aleph_0$: {{begin-eqn}} {{eqn | l = \size \BB \aleph_0 | r = \aleph_0 \size \BB | c = Product of Cardinals is Commutative }} {{eqn | r = \size {\N \times \BB} | c = {{Defof|Product of Cardinals}} }} {{eqn | r = \map \max {\size \N, \size \BB} | c = Cardinal Product Equal to Maximum }} {{eqn | r = \aleph_0 | c = because $\size \BB < \aleph_0$ }} {{eqn | o = < | r = \mathfrak c | c = Aleph Zero is less than Continuum }} {{end-eqn}} In the case when $\size \BB \ge \aleph_0$ we have: {{begin-eqn}} {{eqn | l = \size \BB \aleph_0 | r = \size {\BB \times \N} | c = {{Defof|Product of Cardinals}} }} {{eqn | r = \map \max {\size \BB, \size \N} | c = Cardinal Product Equal to Maximum }} {{eqn | r = \size \BB | c = because $\size \BB \ge \aleph_0$ }} {{eqn | o = < | r = \mathfrak c | c = {{hypothesis}} }} {{end-eqn}} Define: :$Y = \leftset {x \in \R: \exists U \in \BB: x}$ is local minimum in $\rightset U$ We will show that $X \subseteq Y$ by definition of subset. Let $x \in X$. By definition of $X$: :$\exists U \in \leftset {\bigcup \GG: \GG \subseteq \BB}: x$ is local minimum in $\rightset U$ :$\exists \GG \subseteq \BB: U = \bigcup \GG$ By definition of local minimum: :$x \in U$ By definition of union: :$\exists V \in \GG: x \in V$ By definition of subset :$V \in \BB$ By definition of local minimum :$\exists y \in \R: y < x \land \openint y x \cap U = \O$ By Set is Subset of Union: :$V \subseteq U$ Then: :$\exists y \in \R: y < x \land \openint y x \cap V = \O$ By definition: :$x$ is local minimum in $V$ Thus by definition of $Y$ :$x \in Y$ So :$(2): \quad X \subseteq Y$ Define $\family {Z_A}_{A \mathop \in \BB}$ as: :$Z_A = \leftset {x \in \R: x}$ is local minimum in $\rightset A$ We will prove that: :$(3): \quad Y \subseteq \ds \bigcup_{A \mathop \in \BB} Z_A$ Let $x \in Y$. By definition of $Y$: :$\exists U \in \BB: x$ is local minimum in $U$ By definition of $Z_U$: :$x \in Z_U$ Thus by definition of union: :$x \in \ds \bigcup_{A \mathop \in \BB} Z_A$ This ends the proof of inclusion. {{qed|lemma}} By Set of Local Minimum is Countable: :$\forall A \in \BB: Z_A$ is countable By Countable iff Cardinality not greater Aleph Zero: :$\forall A \in \BB: \size {Z_A} \le \aleph_0$ By Cardinality of Union not greater than Product: :$(4): \quad \ds \size {\bigcup_{A \mathop \in \BB} Z_A} \le \size \BB \aleph_0$ Thus: {{begin-eqn}} {{eqn | l = \size X | o = \le | r = \size Y | c = $(2)$ and Subset implies Cardinal Inequality }} {{eqn | o = \le | r = \size {\bigcup_{A \mathop \in \BB} Z_A} | c = $(3)$ and Subset implies Cardinal Inequality }} {{eqn | o = \le | r = \size \BB \aleph_0 | c = $(4)$ }} {{eqn | o = < | r = \mathfrak c | c = $(1)$ }} {{end-eqn}} {{qed}} \end{proof}
21255
\section{Set of Successive Numbers contains Unique Multiple} Tags: Number Theory \begin{theorem} Let $m \in \Z_{\ge 1}$. Then $\set {m, m + 1, \ldots, m + n - 1}$ contains a unique integer that is a multiple of $n$. That is, in any set containing $n$ successive integers, $n$ divides exactly one of those integers. \end{theorem} \begin{proof} Let $S_m = \set {m, m + 1, \ldots, m + n - 1}$ be a set containing $n$ successive integers. The proof proceeds by induction on $m$, the smallest number in $S_m$. \end{proof}
21256
\section{Set of Transpositions is not Subgroup of Symmetric Group} Tags: Symmetric Groups \begin{theorem} Let $S$ be a finite set with $n$ elements such that $n > 2$. Let $G = \struct {\map \Gamma S, \circ}$ denote the symmetric group on $S$. Let $H \subseteq G$ denote the set of all transpositions of $S$ along with the identity mapping which moves no elements of $S$. Then $H$ does not form a subgroup of $G$. \end{theorem} \begin{proof} First it is noted that $H \subseteq G$, and that the identity mapping is an element of $H$. Hence to demonstrate that $H$ is a subgroup of $G$, one may use the Two-Step Subgroup Test. Let $\phi \in H$ be a transposition. Then from Transposition is Self-Inverse, $\phi^{-1} \in H$. So $H$ is closed under inversions. Let $a, b, c \in S$. Let $\pi, \phi \in H$ such that: :$\pi = \tuple {a \ b}$ :$\phi = \tuple {b \ c}$ Then by inspection: :$\phi \circ \pi = \tuple {a \ c \ b}$ Thus $\phi \circ \pi$ is a $3$-cycle and so not a transposition. Hence $H$ is not closed under composition. So, by the Two-Step Subgroup Test, $H$ is not a subgroup of $G$. {{Qed}} \end{proof}
21257
\section{Set of Upper Closures of Compact Elements is Basis implies Complete Scott Topological Lattice is Algebraic} Tags: Topological Order Theory, Continuous Lattices \begin{theorem} Let $L = \struct {S, \preceq, \tau}$ be a complete Scott Definition:Topological Lattice. Let $\BB = \left\{ {x^\succeq: x \in K\left({L}\right)}\right\}$ be a basis of $L$ where :$x^\succeq$ denotes the upper closure of $x$, :$K\left({L}\right)$ denotes the compact subset of $L$. Then $L$ is algebraic. \end{theorem} \begin{proof} Thus by Compact Closure is Directed: :$\forall x \in S:x^{\mathrm{compact} }$ is directed where $x^{\mathrm{compact} }$ denotes the compact closure of $x$. Thus by definition of complete lattice: :$L$ is up-complete. Let $x \in S$. By definition of lower closure of element: :$x$ is upper closure for $x^\preceq$ By definition of supremum: :$\sup \left({x^\preceq}\right) \preceq x$ By Compact Closure is Intersection of Lower Closure and Compact Subset: :$x^{\mathrm{compact} } = x^\preceq \cap K\left({L}\right)$ By Intersection is Subset: :$x^{\mathrm{compact} } \subseteq x^\preceq$ By Supremum of Subset: :$\sup \left({x^{\mathrm{compact} } }\right) \preceq \sup \left({x^\preceq}\right)$ By definition of transitivity: :$\sup \left({x^{\mathrm{compact} } }\right) \preceq x$ {{AimForCont}} :$x \ne \sup \left({x^{\mathrm{compact} } }\right)$ We will prove that :$x \notin \left({\left({x^{\mathrm{compact} } }\right) }\right)^\preceq$ {{AimForCont}} :$x \in \left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq$ By definition of lower closure of element: :$x \preceq \sup \left({x^{\mathrm{compact} } }\right)$ Thus by definition of antisymmetry: :this contrasicts $x \ne \sup \left({x^{\mathrm{compact} } }\right)$ {{qed|lemma}} By definition of relative complement: :$x \in \complement_S\left({\left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq}\right)$ By Complement of Lower Closure of Element is Open in Scott Topological Ordered Set: :$\complement_S\left({\left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq}\right)$ is open. By definition of basis: :$\complement_S\left({\left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq}\right) = \bigcup\left\{ {G \in \BB: G \subseteq \complement_S\left({\left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq}\right)}\right\}$ By definition of union :$\exists X \in \left\{ {G \in \BB: G \subseteq \complement_S\left({\left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq}\right)}\right\}: x \in X$ Then :$X \in \BB \land X \subseteq \complement_S\left({\left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq}\right)$ By definition of $\BB$: :$\exists k \in K\left({L}\right): X = k^\succeq$ By definition of compact subset: :$k$ is compact. By definition of upper closure of element: :$k \preceq x$ By definition of compact closure: :$k \in x^{\mathrm{compact} }$ By definitions of supremum and upper bound: :$k \preceq \sup \left({x^{\mathrm{compact} } }\right)$ By definition of upper closure of element: :$\sup \left({x^{\mathrm{compact} } }\right) \in X$ By gefinition of subset: :$\sup \left({x^{\mathrm{compact} } }\right) \in \complement_S\left({\left({\sup \left({x^{\mathrm{compact} } }\right) }\right)^\preceq}\right)$ By definition of difference: :$\sup \left({x^{\mathrm{compact} } }\right) \notin \left({\sup \left({x^{\mathrm{compact} } }\right)}\right)^\preceq$ Thus by definition of lower closure of element: :this contradicts $\sup \left({x^{\mathrm{compact} } }\right) \in \left({\sup \left({x^{\mathrm{compact} } }\right)}\right)^\preceq$ {{qed}} \end{proof}
21258
\section{Set of Words Generates Group} Tags: Set of Words Generates Group, Generated Subgroups, Group Theory \begin{theorem} Let $S \subseteq G$ where $G$ is a group. Let $\hat S$ be defined as $S \cup S^{-1}$, where $S^{-1}$ is the set of all the inverses of all the elements of $S$. Then $\gen S = \map W {\hat S}$, where $\map W {\hat S}$ is the set of words of $\hat S$. \end{theorem} \begin{proof} Let $H = \gen S$ where $S \subseteq G$. $H$ must certainly include $\hat S$, because any group containing $s \in S$ must also contain $s^{-1}$. Thus $\hat S \subseteq H$. By {{GroupAxiom|0}}, $H$ must also contain all products of a finite number of elements of $\hat S$. Thus $\map W {\hat S} \subseteq H$. Now we prove that $\map W {\hat S} \le G$. By the Two-Step Subgroup Test: Let $x, y \in \map W {\hat S}$. As $x$ and $y$ are both products of a finite number of elements of $\hat S$, it follows that so is their product $x y$. Thus $x y \in \map W {\hat S}$ and {{GroupAxiom|0}} is satisfied. Let $x = s_1 s_2 \ldots s_n \in \map W {\hat S}$. Then $x^{-1} = s_n^{-1} \ldots s_2^{-1} s_1^{-1} \in \map W {\hat S}$. Thus the conditions of the Two-Step Subgroup Test are fulfilled, and $\map W {\hat S} \le G$. Thus $\map W {\hat S}$ is the subgroup of $G$ generated by $S$. {{qed}} \end{proof}
21259
\section{Set of all Self-Maps under Composition forms Monoid} Tags: Examples of Monoids, Abstract Algebra, Monoids, Mappings, Composite Mappings, Mapping Theory \begin{theorem} Let $S$ be a set. Let $S^S$ be the set of all mappings from $S$ to itself. Let the operation $\circ$ represent composition of mappings. Then the algebraic structure $\struct {S^S, \circ}$ is a monoid whose identity element is the identity mapping on $S$. \end{theorem} \begin{proof} By Set of all Self-Maps under Composition forms Semigroup, $\struct {S^S, \circ}$ is a semigroup. By Identity Mapping is Left Identity and Identity Mapping is Right Identity the identity mapping on $S$ is the identity element of $\struct {S^S, \circ}$. Since $\struct {S^S, \circ}$ is a semigroup with an identity element, it is a monoid by definition. {{qed}} \end{proof}
21260
\section{Set of all Self-Maps under Composition forms Semigroup} Tags: Semigroups, Mapping Theory, Examples of Semigroups, Composite Mappings \begin{theorem} Let $S$ be a set. Let $S^S$ be the set of all mappings from $S$ to itself. Let the operation $\circ$ represent composition of mappings. Then the algebraic structure $\struct {S^S, \circ}$ is a semigroup. \end{theorem} \begin{proof} Let $f, g \in S^S$. As the domain of $g$ and codomain of $f$ are the same, the composition $f \circ g$ is defined. By the definition of composition, $f \circ g$ is a mapping from the domain of $g$ to the codomain of $f$. Thus $f \circ g: S \to S$, so $f \circ g \in S^S$. Since this holds for all $f, g \in S^S$, $\struct {S^S, \circ}$ is closed. By Composition of Mappings is Associative, $\circ$ is associative. Since $\struct {S^S, \circ}$ is closed and $\circ$ is associative: :$\struct {S^S, \circ}$ is a semigroup. {{qed}} \end{proof}
21261
\section{Set together with Condensation Points is not necessarily Closed} Tags: Condensation Points, Set Closures, Omega-Accumulation Points \begin{theorem} Let $T = \struct {S, \tau}$ be a topological space. Let $H \subseteq S$. Let $\CC$ denote the set of condensation points of $H$. Then it is not necessarily the case that $H \cup \CC$ is a closed set of $T$. \end{theorem} \begin{proof} Proof by Counterexample: Let $T = \struct {\R, \tau}$ denote the right order topology on $\R$. Let $H \subseteq \R$ be a finite subset of $\R$. Let $\CC$ denote the set of condensation points of $H$. From Finite Set of Right Order Topology with Condensation Points is not Closed, $H \cup \CC$ is not a closed set of $T$. {{qed}} \end{proof}
21262
\section{Set together with Omega-Accumulation Points is not necessarily Closed} Tags: Set Closures, Omega-Accumulation Points \begin{theorem} Let $T = \struct {S, \tau}$ be a topological space. Let $H \subseteq S$. Let $\Omega$ denote the set of $\omega$-accumulation points of $H$. Then it is not necessarily the case that $H \cup \Omega$ is a closed set of $T$. \end{theorem} \begin{proof} Proof by Counterexample: Let $T = \struct {\R, \tau}$ denote the right order topology on $\R$. Let $H \subseteq \R$ be a finite subset of $\R$. Let $\Omega$ denote the set of $\omega$-accumulation points of $H$. From Finite Set of Right Order Topology with Omega-Accumulation Points is not Closed, $H \cup \Omega$ is not a closed set of $T$. {{qed}} \end{proof}
21263
\section{Set with Complement forms Partition} Tags: Set Partitions, Set Complement, Universe \begin{theorem} Let $\O \subset S \subset \mathbb U$. Then $S$ and its complement $\map \complement S$ form a partition of the universal set $\mathbb U$. \end{theorem} \begin{proof} Follows directly from Set with Relative Complement forms Partition: If $\O \subset T \subset S$, then $\set {T, \relcomp S T}$ is a partition of $S$. {{Qed}} Category:Set Complement Category:Universe Category:Set Partitions \end{proof}
21264
\section{Set with Dispersion Point is Biconnected} Tags: Dispersion Points, Biconnectedness, Connectedness, Biconnected Sets \begin{theorem} Let $T = \struct {S, \tau}$ be a topological space. Let $H \subseteq S$ be a connected set in $T$. Let $p \in H$ be a dispersion point of $H$. Then $H$ is biconnected. \end{theorem} \begin{proof} {{AimForCont}} $H$ is not biconnected. Then by definition there exist disjoint non-degenerate connected sets $U, V$ such that $H = U \cup V$. {{WLOG}}, let $p \in U$. Then $V \subset H \setminus \set p$. As $p$ is a dispersion point of $H$, $H \setminus \set p$ is totally disconnected. Thus, by definition, $H \setminus \set p$ contains no non-degenerate connected sets. But this contradicts our definition of $V$, as $V \subset H \setminus \set p$. It follows by Proof by Contradiction that $H$ is biconnected. {{qed}} \end{proof}
21265
\section{Set with Two Parallel Elements is Dependent} Tags: Matroid Theory \begin{theorem} Let $M = \struct{S, \mathscr I}$ be a matroid. Let $A \subseteq S$. Let $x, y \in S$. Let $x, y$ be parallel elements. If $x, y \in A$ then $A$ is dependent. \end{theorem} \begin{proof} Let $x, y \in A$. From Doubleton of Elements is Subset: :$\set{x, y} \subseteq A$ By the definition of parallel elements: :$\set {x, y}$ is dependent From Superset of Dependent Set is Dependent: :$A$ is dependent {{qed}} \end{proof}
21266
\section{Sets in Modified Fort Space are Disconnected} Tags: Modified Fort Space, Connectedness, Disconnected Spaces \begin{theorem} Let $T = \struct {S, \tau_{a, b}}$ be a modified Fort space. Let $H$ be a subset of $S$ with more than one point. Then $H$ is disconnected. \end{theorem} \begin{proof} By Isolated Points in Subsets of Modified Fort Space: :$\exists x \in H: x$ is isolated By Point in Topological Space is Open iff Isolated, $\set x$ is open in $T$. By Modified Fort Space is $T_1$ and definition of $T_1$ space, $\set x$ is closed in $T$. Therefore $\relcomp S {\set x}$ is open in $T$. Then we have: {{begin-eqn}} {{eqn | l = H | o = \subseteq | r = S }} {{eqn | r = \set x \cup \relcomp S {\set x} | c = Union with Relative Complement }} {{eqn | l = H \cap \set x \cap \relcomp S {\set x} | o = \subseteq | r = \set x \cap \relcomp S {\set x} | c = Intersection is Subset }} {{eqn | r = \O | c = Intersection with Relative Complement is Empty }} {{eqn | ll = \leadsto | l = H \cap \set x \cap \relcomp S {\set x} | r = \O | c = Subset of Empty Set }} {{eqn | l = H \cap \set x | r = \set x }} {{eqn | o = \ne | r = \O }} {{eqn | l = H \cap \relcomp S {\set x} | r = H \setminus \set x | c = Set Difference as Intersection with Relative Complement }} {{eqn | o = \ne | r = \O | c = $H$ has more than one point }} {{end-eqn}} This shows that $H$ is disconnected. {{qed}} \end{proof}
21267
\section{Sets of Operations on Set of 3 Elements/Automorphism Group of A} Tags: Sets of Operations on Set of 3 Elements \begin{theorem} Let $S = \set {a, b, c}$ be a set with $3$ elements. Let $\AA$ be the set of all operations $\circ$ on $S$ such that the group of automorphisms of $\struct {S, \circ}$ is the symmetric group on $S$, that is, $\map \Gamma S$. Then: :$\AA$ has $3$ elements. \end{theorem} \begin{proof} Recall the definition of (group) automorphism: :$\phi$ is an automorphism on $\struct {S, \circ}$ {{iff}}: ::$\phi$ is a permutation of $S$ ::$\phi$ is a homomorphism on $\struct {S, \circ}$: $\forall a, b \in S: \map \phi {a \circ b} = \map \phi a \circ \map \phi b$ Hence $\AA$ can be defined as the set of operations $\circ$ on $S$ such that every permutation on $S$ is an automorphism of $\struct {S, \circ}$. The set $\map \Gamma S$ of all permutations on $S = \set {a, b, c}$ has $6$ elements. From Identity Mapping is Group Automorphism, $I_S$ is always an automorphism on $\struct {S, \circ}$. Hence it is not necessary to analyse the effect of $I_S$ on the various elements of $S$. Let us denote each of the remaining permutations on $S$ as follows: {{begin-eqn}} {{eqn | l = p | o = : | r = \map p a = b, \map p b = c, \map p c = a }} {{eqn | l = q | o = : | r = \map q a = c, \map q b = a, \map q c = b }} {{eqn | l = r | o = : | r = \map r a = a, \map r b = c, \map r c = b }} {{eqn | l = s | o = : | r = \map s a = c, \map s b = b, \map s c = a }} {{eqn | l = t | o = : | r = \map t a = b, \map t b = a, \map t c = c }} {{end-eqn}} So, let $\circ \in \AA$. From Permutation of Set is Automorphism of Set under Left Operation and Permutation of Set is Automorphism of Set under Right Operation: :${\to} \in \AA$ :${\gets} \in \AA$ where $\to$ and $\gets$ are the right operation and left operation respectively. Next we note that from Structure of Cardinality 3+ where Every Permutation is Automorphism is Idempotent: :$\forall x \in S: x \circ x = x$ Let us explore various options for $a \circ b$. Let $a \circ b = a$. Then by definition of the above mappings $p$, $q$, $r$, $s$ and $t$, and the definition of a homomorphism: {{begin-eqn}} {{eqn | l = \map p a \circ \map p b | r = \map p a | c = }} {{eqn | ll= \leadsto | l = b \circ c | r = b | c = }} {{eqn | l = \map q a \circ \map q b | r = \map q a | c = }} {{eqn | ll= \leadsto | l = c \circ a | r = c | c = }} {{eqn | l = \map r a \circ \map r b | r = \map r a | c = }} {{eqn | ll= \leadsto | l = a \circ c | r = a | c = }} {{eqn | l = \map s a \circ \map s b | r = \map s a | c = }} {{eqn | ll= \leadsto | l = c \circ b | r = c | c = }} {{eqn | l = \map t a \circ \map t b | r = \map t a | c = }} {{eqn | ll= \leadsto | l = b \circ a | r = b | c = }} {{end-eqn}} This is none other than the left operation, which has already been counted. {{qed|lemma}} Let $a \circ b = b$. Then by definition of the above mappings $p$, $q$, $r$, $s$ and $t$, and the definition of a homomorphism: {{begin-eqn}} {{eqn | l = \map p a \circ \map p b | r = \map p b | c = }} {{eqn | ll= \leadsto | l = b \circ c | r = c | c = }} {{eqn | l = \map q a \circ \map q b | r = \map q b | c = }} {{eqn | ll= \leadsto | l = c \circ a | r = a | c = }} {{eqn | l = \map r a \circ \map r b | r = \map r b | c = }} {{eqn | ll= \leadsto | l = a \circ c | r = c | c = }} {{eqn | l = \map s a \circ \map s b | r = \map s b | c = }} {{eqn | ll= \leadsto | l = c \circ b | r = b | c = }} {{eqn | l = \map t a \circ \map t b | r = \map t b | c = }} {{eqn | ll= \leadsto | l = b \circ a | r = a | c = }} {{end-eqn}} This is none other than the right operation, which has already been counted. {{qed|lemma}} Let $a \circ b = c$. Then by definition of the above mappings $p$, $q$, $r$, $s$ and $t$, and the definition of a homomorphism: {{begin-eqn}} {{eqn | l = \map p a \circ \map p b | r = \map p c | c = }} {{eqn | ll= \leadsto | l = b \circ c | r = a | c = }} {{eqn | l = \map q a \circ \map q b | r = \map q c | c = }} {{eqn | ll= \leadsto | l = c \circ a | r = b | c = }} {{eqn | l = \map r a \circ \map r b | r = \map r c | c = }} {{eqn | ll= \leadsto | l = a \circ c | r = b | c = }} {{eqn | l = \map s a \circ \map s b | r = \map s c | c = }} {{eqn | ll= \leadsto | l = c \circ b | r = a | c = }} {{eqn | l = \map t a \circ \map t b | r = \map t c | c = }} {{eqn | ll= \leadsto | l = b \circ a | r = c | c = }} {{end-eqn}} Thus we have identified another operation on $S$ for which all permutations on $S$ are automorphisms of $\struct {S, \circ}$. {{qed|lemma}} We now note that once the product element $a \circ b$ has been selected to be either $a$, $b$ or $c$, the full structure of $\struct {S, \circ}$ is forced. Hence there are no other operations $\circ$ on $S$ but these ones we have counted. That is, there is a total of $3$ such operations. {{qed}} \end{proof}
21268
\section{Sets of Operations on Set of 3 Elements/Automorphism Group of A/Commutative Operations} Tags: Sets of Operations on Set of 3 Elements \begin{theorem} Let $S = \set {a, b, c}$ be a set with $3$ elements. Let $\AA$ be the set of all operations $\circ$ on $S$ such that the group of automorphisms of $\struct {S, \circ}$ is the symmetric group on $S$, that is, $\map \Gamma S$. Then: :Exactly $1$ of the operations of $\AA$ is commutative. \end{theorem} \begin{proof} Recall from Automorphism Group of $\AA$ the elements of $\AA$, expressed in Cayley table form: :$\begin {array} {c|ccc} \to & a & b & c \\ \hline a & a & b & c \\ b & a & b & c \\ c & a & b & c \\ \end {array} \qquad \begin {array} {c|ccc} \gets & a & b & c \\ \hline a & a & a & a \\ b & b & b & b \\ c & c & c & c \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & a & c & b \\ b & c & b & a \\ c & b & a & c \\ \end {array}$ From Cayley Table for Commutative Operation is Symmetrical about Main Diagonal, it is apparent by inspection that of the above: :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & a & c & b \\ b & c & b & a \\ c & b & a & c \\ \end {array}$ is the only Cayley table of a commutative operation. {{qed}} \end{proof}
21269
\section{Sets of Operations on Set of 3 Elements/Automorphism Group of A/Isomorphism Classes} Tags: Sets of Operations on Set of 3 Elements \begin{theorem} Let $S = \set {a, b, c}$ be a set with $3$ elements. Let $\AA$ be the set of all operations $\circ$ on $S$ such that the group of automorphisms of $\struct {S, \circ}$ is the symmetric group on $S$, that is, $\map \Gamma S$. Then: :The elements of $\AA$ are each in its own isomorphism class. \end{theorem} \begin{proof} Recall from Automorphism Group of $\AA$ the elements of $\AA$, expressed in Cayley table form: :$\begin {array} {c|ccc} \to & a & b & c \\ \hline a & a & b & c \\ b & a & b & c \\ c & a & b & c \\ \end {array} \qquad \begin {array} {c|ccc} \gets & a & b & c \\ \hline a & a & a & a \\ b & b & b & b \\ c & c & c & c \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & a & c & b \\ b & c & b & a \\ c & b & a & c \\ \end {array}$ We have from Algebraic Structures formed by Left and Right Operations are not Isomorphic for Cardinality Greater than 1 that $\struct {S, \to}$ and $\struct {S, \gets}$ are not isomorphic. Observe from the Cayley table $\circ$ that: :$(1): \quad \forall x, y \in S: x \ne y \implies x \circ y \ne x \land x \circ y \ne y$ {{AimForCont}} there exists an isomorphism $\phi$ from $\struct {S, \to}$ to $\struct {S, \circ}$. We have: Then: {{begin-eqn}} {{eqn | l = \map \phi {a \to b} | r = \map \phi a \circ \map \phi b | c = {{Defof|Isomorphism (Abstract Algebra)|Isomorphism}} }} {{eqn | r = \map \phi c | c = from $(1)$ }} {{eqn | r = \map \phi b | c = {{Defof|Right Operation}} }} {{end-eqn}} Hence: :$\map \phi c = \map \phi b$ which means $\phi$ is not an injection. That is, $\phi$ is not a bijection and hence not an isomorphism. Hence by Proof by Contradiction there exists no isomorphism from $\struct {S, \to}$ to $\struct {S, \circ}$. In a similar way: {{AimForCont}} there exists an isomorphism $\phi$ from $\struct {S, \gets}$ to $\struct {S, \circ}$. We have: Then: {{begin-eqn}} {{eqn | l = \map \phi {a \gets b} | r = \map \phi a \circ \map \phi b | c = {{Defof|Isomorphism (Abstract Algebra)|Isomorphism}} }} {{eqn | r = \map \phi c | c = from $(1)$ }} {{eqn | r = \map \phi a | c = {{Defof|Left Operation}} }} {{end-eqn}} Hence: :$\map \phi c = \map \phi a$ which means $\phi$ is not an injection. That is, $\phi$ is not a bijection and hence not an isomorphism. Hence by Proof by Contradiction there exists no isomorphism from $\struct {S, \gets}$ to $\struct {S, \circ}$. It has been shown that none of the elements of $\AA$ is isomorphism to any of the others. The result follows by definition of isomorphism class. {{qed}} \end{proof}
21270
\section{Sets of Operations on Set of 3 Elements/Automorphism Group of A/Operations with Identity} Tags: Sets of Operations on Set of 3 Elements \begin{theorem} Let $S = \set {a, b, c}$ be a set with $3$ elements. Let $\AA$ be the set of all operations $\circ$ on $S$ such that the group of automorphisms of $\struct {S, \circ}$ is the symmetric group on $S$, that is, $\map \Gamma S$. Then: :None of the operations of $\AA$ has an identity element. \end{theorem} \begin{proof} Recall from Automorphism Group of $\AA$ the elements of $\AA$, expressed in Cayley table form: :$\begin{array}{c|ccc} \to & a & b & c \\ \hline a & a & b & c \\ b & a & b & c \\ c & a & b & c \\ \end{array} \qquad \begin{array}{c|ccc} \gets & a & b & c \\ \hline a & a & a & a \\ b & b & b & b \\ c & c & c & c \\ \end{array} \qquad \begin{array}{c|ccc} \circ & a & b & c \\ \hline a & a & c & b \\ b & c & b & a \\ c & b & a & c \\ \end{array}$ The result is apparent by inspection. {{qed}} \end{proof}
21271
\section{Sets of Operations on Set of 3 Elements/Automorphism Group of B} Tags: Sets of Operations on Set of 3 Elements \begin{theorem} Let $S = \set {a, b, c}$ be a set with $3$ elements. Let $\BB$ be the set of all operations $\circ$ on $S$ such that the group of automorphisms of $\struct {S, \circ}$ forms the set $\set {I_S, \tuple {a, b, c}, \tuple {a, c, b} }$, where $I_S$ is the identity mapping on $S$. Then: :$\BB$ has $3^3 - 3$ elements. \end{theorem} \begin{proof} Recall the definition of (group) automorphism: :$\phi$ is an automorphism on $\struct {S, \circ}$ {{iff}}: ::$\phi$ is a permutation of $S$ ::$\phi$ is a homomorphism on $\struct {S, \circ}$: $\forall a, b \in S: \map \phi {a \circ b} = \map \phi a \circ \map \phi b$ From Identity Mapping is Group Automorphism, $I_S$ is always an automorphism on $\struct {S, \circ}$. Hence it is not necessary to analyse the effect of $I_S$ on $S$. Let us denote each of the remaining elements of $\set {I_S, \tuple {a, b, c}, \tuple {a, c, b} }$ as follows: {{begin-eqn}} {{eqn | l = p | o = : | r = \map p a = b, \map p b = c, \map p c = a }} {{eqn | l = q | o = : | r = \map q a = c, \map q b = a, \map q c = b }} {{end-eqn}} We select various product elements $x \circ y \in S$ and determine how $p$ and $q$ constrain other product elements as follows: Then by definition of the above mappings $p$ and $q$, and the definition of a homomorphism, we obtain as follows: ;$(1): \quad a \circ a$ {{Begin-table}}|- | align="left" | {{begin-eqn}} {{eqn | l = a \circ a | r = a | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p a | r = \map p a | c = }} {{eqn | ll= \leadsto | l = b \circ b | r = b | c = }} {{eqn | l = \map q a \circ \map q a | r = \map q a | c = }} {{eqn | ll= \leadsto | l = c \circ c | r = c | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & a & & \\ b & & b & \\ c & & & c \\ \end {array}$ |- | align="left" | {{begin-eqn}} {{eqn | l = a \circ a | r = b | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p a | r = \map p b | c = }} {{eqn | ll= \leadsto | l = b \circ b | r = c | c = }} {{eqn | l = \map q a \circ \map q a | r = \map q b | c = }} {{eqn | ll= \leadsto | l = c \circ c | r = a | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & b & & \\ b & & c & \\ c & & & a \\ \end {array}$ |- | align="left" | {{begin-eqn}} {{eqn | l = a \circ a | r = c | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p a | r = \map p c | c = }} {{eqn | ll= \leadsto | l = b \circ b | r = a | c = }} {{eqn | l = \map q a \circ \map q a | r = \map q c | c = }} {{eqn | ll= \leadsto | l = c \circ c | r = b | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & c & & \\ b & & a & \\ c & & & b \\ \end {array}$ {{End-table}} So selecting $a \circ a$ fixes $b \circ b$ and $c \circ c$. ;$(2): \quad a \circ b$ {{Begin-table}}|- | align="left" | {{begin-eqn}} {{eqn | l = a \circ b | r = a | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p b | r = \map p a | c = }} {{eqn | ll= \leadsto | l = b \circ c | r = b | c = }} {{eqn | l = \map q a \circ \map q b | r = \map q a | c = }} {{eqn | ll= \leadsto | l = c \circ a | r = c | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & a & \\ b & & & b \\ c & c & & \\ \end {array}$ |- | align="left" | {{begin-eqn}} {{eqn | l = a \circ b | r = b | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p b | r = \map p b | c = }} {{eqn | ll= \leadsto | l = b \circ c | r = c | c = }} {{eqn | l = \map q a \circ \map q b | r = \map q b | c = }} {{eqn | ll= \leadsto | l = c \circ a | r = a | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & b & \\ b & & & c \\ c & a & & \\ \end {array}$ |- | align="left" | {{begin-eqn}} {{eqn | l = a \circ b | r = c | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p b | r = \map p c | c = }} {{eqn | ll= \leadsto | l = b \circ c | r = a | c = }} {{eqn | l = \map q a \circ \map q b | r = \map q c | c = }} {{eqn | ll= \leadsto | l = c \circ a | r = b | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & c & \\ b & & & a \\ c & b & & \\ \end {array}$ {{End-table}} So selecting $a \circ b$ fixes $b \circ c$ and $c \circ a$. ;$(3): \quad a \circ c$ {{Begin-table}}|- | align="left" | {{begin-eqn}} {{eqn | l = a \circ c | r = a | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p c | r = \map p a | c = }} {{eqn | ll= \leadsto | l = b \circ a | r = b | c = }} {{eqn | l = \map q a \circ \map q c | r = \map q a | c = }} {{eqn | ll= \leadsto | l = c \circ b | r = c | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & & a \\ b & b & & \\ c & & c & \\ \end {array}$ |- | align="left" | {{begin-eqn}} {{eqn | l = a \circ c | r = b | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p c | r = \map p b | c = }} {{eqn | ll= \leadsto | l = b \circ a | r = c | c = }} {{eqn | l = \map q a \circ \map q c | r = \map q b | c = }} {{eqn | ll= \leadsto | l = c \circ b | r = a | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & & b \\ b & c & & \\ c & & a & \\ \end {array}$ |- | align="left" | {{begin-eqn}} {{eqn | l = a \circ c | r = c | c = }} {{eqn | ll= \leadsto | l = \map p a \circ \map p c | r = \map p c | c = }} {{eqn | ll= \leadsto | l = b \circ a | r = a | c = }} {{eqn | l = \map q a \circ \map q c | r = \map q c | c = }} {{eqn | ll= \leadsto | l = c \circ b | r = b | c = }} {{end-eqn}} | align="left" | :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & & c \\ b & a & & \\ c & & b & \\ \end {array}$ {{End-table}} So selecting $a \circ c$ fixes $b \circ a$ and $c \circ b$. There are $3$ elements $x$ of $S$ with which $a \circ x$ can be made. For each of these $3$, there are $3$ different elements $y$ such that $a \circ x = y$. These collectively fix all the possible values of $b \circ x$ and $c \circ x$. Hence they exhaust all possible operations $\circ$ on $S$ for which $p$ and $q$ are automorphisms. Thus there are independently: :$3$ different options for $a \circ a$ :$3$ different options for $a \circ b$ :$3$ different options for $a \circ c$ and therefore $3 \times 3 \times 3 = 3^3$ operations $\circ$ on $S$ for which $p$ and $q$ are automorphisms. Note that from Automorphism Group of $\AA$, $3$ of these are also such that the group of automorphisms of $\struct {S, \circ}$ is the symmetric group on $S$. So these are excluded from our count. The result follows. {{qed}} \end{proof}
21272
\section{Sets of Operations on Set of 3 Elements/Automorphism Group of B/Commutative Operations} Tags: Sets of Operations on Set of 3 Elements \begin{theorem} Let $S = \set {a, b, c}$ be a set with $3$ elements. Let $\BB$ be the set of all operations $\circ$ on $S$ such that the group of automorphisms of $\struct {S, \circ}$ forms the set $\set {I_S, \tuple {a, b, c}, \tuple {a, c, b} }$, where $I_S$ is the identity mapping on $S$. Then: :Exactly $8$ of the operations of $\BB$ is commutative. \end{theorem} \begin{proof} Recall Automorphism Group of $\BB$. Consider each of the categories of $\BB$ induced by each of $a \circ a$, $a \circ b$ and $a \circ c$, illustrated by the partially-filled Cayley tables to which they give rise: ;$(1): \quad a \circ a$ :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & a & & \\ b & & b & \\ c & & & c \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & b & & \\ b & & c & \\ c & & & a \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & c & & \\ b & & a & \\ c & & & b \\ \end {array}$ ;$(2): \quad a \circ b$ :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & a & \\ b & & & b \\ c & c & & \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & b & \\ b & & & c \\ c & a & & \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & c & \\ b & & & a \\ c & b & & \\ \end {array}$ ;$(3): \quad a \circ c$ :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & & c \\ b & a & & \\ c & & b & \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & & b \\ b & c & & \\ c & & a & \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & & c \\ b & a & & \\ c & & b & \\ \end {array}$ With a view to Cayley Table for Commutative Operation is Symmetrical about Main Diagonal, we inspect the various combinations of these partial Cayley tables. It is apparent that the result of $x \circ x$ makes no difference to whether an operation is commutative. However, note that each of the partial operations in $(2)$ can be conjoined with exactly $1$ of the partial operations in $(3)$ to make a commutative operation: :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & a & c \\ b & a & & b \\ c & c & b & \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & b & a \\ b & b & & c \\ c & a & c & \\ \end {array} \qquad \begin {array} {c|ccc} \circ & a & b & c \\ \hline a & & c & b \\ b & c & & a \\ c & b & a & \\ \end {array}$ Hence by the Product Rule for Counting there are $3 \times 3 = 9$ commutative operations which can be constructed thus. However, note that one of these: :$\begin {array} {c|ccc} \circ & a & b & c \\ \hline a & a & c & b \\ b & c & b & a \\ c & b & a & c \\ \end {array}$ has already been accounted for in Automorphism Group of $\AA$: Commutative Operations. This commutative operations is such that the group of automorphisms of $\struct {S, \circ}$ forms the complete symmetric group on $S$, not just the given permutations. Hence the result. {{qed}} \end{proof}