id
stringlengths
1
260
contents
stringlengths
1
234k
23373
\section{Way Above Closures Form Basis} Tags: Topological Order Theory, Continuous Lattices \begin{theorem} Let $L = \struct {S, \preceq, \tau}$ be a complete continuous topological lattice with Scott topology. Then $\set {x^\gg: x \in S}$ is an (analytic) basis of $L$. \end{theorem} \begin{proof} Define $B = \set {x^\gg: x \in S}$. Thus by Way Above Closure is Open: :$B \subseteq \tau$ We will prove that: :for all $x \in S$: there exists a local basis $Q$ of $x$: $Q \subseteq B$ Let $x \in S$. By Way Above Closures that Way Below Form Local Basis: :$Q := \set {g^\gg: g \in S \land g \ll x}$ is a local basis at $x$. Thus by definition of subset: :$Q \subseteq B$ {{qed|lemma}} Thus by Characterization of Analytic Basis by Local Bases: :$B$ is an (analytic) basis of $L$. {{qed}} \end{proof}
23374
\section{Way Above Closures that Way Below Form Local Basis} Tags: Way Below Relation, Topological Order Theory, Continuous Lattices \begin{theorem} Let $L = \struct {S, \preceq, \tau}$ be a complete continuous topological lattice with Scott topology. Let $p \in S$. Then $\set {q^\gg: q \in S \land q \ll p}$ is a local basis at $p$. \end{theorem} \begin{proof} Define $B := \set {q^\gg: q \in S \land q \ll p}$ By Way Above Closure is Open: :$B \subseteq \tau$ By definition of way above closure: :$\forall X \in B: p \in X$ Thus by definition: :$B$ is set of open neighborhoods. {{explain|open neighborhoods of what?}} Let $U$ be an open subset of $S$ such that :$p \in U$ By Open implies There Exists Way Below Element: :$\exists u \in U: u \ll p$ Thus by definition of $B$: :$u^\gg \in B$ By definition of Scott topology: :$U$ is upper. We will prove that :$u^\gg \subseteq U$ Let $z \in u^\gg$ By definition of way above closure: :$u \ll z$ By Way Below implies Preceding: :$u \preceq z$ Thus by definition of upper set: :$z \in U$ {{qed}} \end{proof}
23375
\section{Way Below Closure is Directed in Bounded Below Join Semilattice} Tags: Way Below Relation, Join and Meet Semilattices \begin{theorem} Let $\struct {S, \vee, \preceq}$ be a bounded below join semilattice. Let $x \in S$. Then :$x^\ll$ is directed. \end{theorem} \begin{proof} By Bottom is Way Below Any Element: :$\bot \ll x$ By definition of way below closure: :$\bot \in x^\ll$ Thus by definition: :$x^\ll$ is a non-empty set. Let $y, z \in x^\ll$ By definition of way below closure: :$y \ll x$ and $z \ll x$ By Join is Way Below if Operands are Way Below :$y \vee z \ll x$ By definition of way below closure: :$y \vee z \in x^\ll$ By Join Succeeds Operands: :$y \preceq y \vee z$ and $z \preceq y \vee z$ Thus by definition :$x^\ll$ is directed. {{qed}} \end{proof}
23376
\section{Way Below Closure is Ideal in Bounded Below Join Semilattice} Tags: Way Below Relation, Join and Meet Semilattices \begin{theorem} Let $L = \struct {S, \vee, \preceq}$ be a bounded below join semilattice. Let $x \in S$. Then :$x^\ll$ is ideal in $L$. \end{theorem} \begin{proof} By Way Below Closure is Directed in Bounded Below Join Semilattice: :$x^\ll$ is a non-empty directed set. Let $y \in x^\ll, z \in S$ such that :$z \preceq y$ By definition of way below closure: :$y \ll x$ By definition of reflexivity: :$x \preceq x$ By Preceding and Way Below implies Way Below: :$z \ll x$ Thus by definition of way below closure: :$z \in x^\ll$ Thus by definition :$x^\ll$ is a lower set. Thus by definition :$x^\ll$ is an ideal in $L$. {{qed}} \end{proof}
23377
\section{Way Below Closure is Lower Set} Tags: Way Below Relation, Join and Meet Semilattices \begin{theorem} Let $L = \struct {S, \vee, \preceq}$ be an ordered set. Let $x \in S$. Then :$x^\ll$ is a lower set. \end{theorem} \begin{proof} Let $y \in x^\ll, z \in S$ such that: :$z \preceq y$ By definition of way below closure: :$y \ll x$ By definition of reflexivity: :$x \preceq x$ By Preceding and Way Below implies Way Below: :$z \ll x$ Thus by definition of way below closure: :$z \in x^\ll$ Thus by definition: :$x^\ll$ is a lower set. {{qed}} \end{proof}
23378
\section{Way Below Relation is Antisymmetric} Tags: Way Below Relation \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Let $x, y \in S$ such that :$x \ll y$ and $y \ll x$ Then :$x = y$ \end{theorem} \begin{proof} By Way Below implies Preceding: :$x \preceq y$ and $y \preceq x$ Thus by definition of antisymmetry: :$x = y$ {{qed}} \end{proof}
23379
\section{Way Below Relation is Auxiliary Relation} Tags: Way Below Relation, Auxiliary Relations \begin{theorem} Lrt $L = \left({S, \vee, \preceq}\right)$ be a bounded below join semilattice. Then :$\ll$ is auxiliary relation where $\ll$ denotes the way below relation. \end{theorem} \begin{proof} By Way Below implies Preceding: :$\forall x, y \in S: x \ll y \implies x \preceq y$ By Preceding and Way Below implies Way Below: :$\forall x, y, z, u \in S: x \preceq y \ll z \preceq u \implies x \ll u$ By Join is Way Below if Operands are Way Below: :$\forall x, y, z \in S: x \ll z \land y \ll z \implies x \vee y \ll z$ By Bottom is Way Below Any Element: :$\forall x: \bot \ll x$ where $\bot$ denotes the smallest element in $L$. Thus by definition: :$\ll$ is auxiliary relation. {{qed}} \end{proof}
23380
\section{Way Below Relation is Multiplicative implies Pseudoprime Element is Prime} Tags: Way Below Relation, Prime Elements \begin{theorem} Let $L = \struct {S, \vee, \wedge, \preceq}$ be a bounded below continuous lattice such that :$\ll$ is multiplicative relation where $\ll$ denotes the way below relation of $L$. Let $p \in S$. Then $p$ is a pseudoprime element is a prime element. \end{theorem} \begin{proof} Let $p$ be a pseudoprime element. {{AimForCont}} :$p$ is not a prime element. By definition of prime element: :$\exists x, y \in S: x \wedge y \preceq p$ and $x \npreceq p$ and $y \npreceq p$ By definition of continuous: :$\forall z \in S: z^\ll$ is directed. and :$L$ satisfies axiom of approximation. By Axiom of Approximation in Up-Complete Semilattice: :$\exists u \in S: u \ll x \land u \npreceq p$ and :$\exists v \in S: v \ll y \land v \npreceq p$ By Way Below Relation is Auxiliary Relation: :$\ll$ is auxiliary relation. By Multiplicative Auxiliary Relation iff Congruent: :$u \wedge v \ll x \wedge y$ By definition of transitivity: :$u \wedge v \ll p$ By Characterization of Pseudoprime Element when Way Below Relation is Multiplicative: :$u \preceq p$ or $v \preceq p$ This contradicts $u \npreceq p$ and $v \npreceq p$ {{qed}} \end{proof}
23381
\section{Way Below Relation is Transitive} Tags: Way Below Relation \begin{theorem} Let $\left({S, \preceq}\right)$ be an ordered set. Let $x, y, z \in S$ such that :$x \ll y \ll z$ Then :$x \ll z$ \end{theorem} \begin{proof} By Way Below implies Preceding: :$x \preceq y$ By definition of reflexivity: :$z \preceq z$ Thus by Preceding and Way Below implies Way Below: :$x \ll z$ {{qed}} \end{proof}
23382
\section{Way Below has Interpolation Property} Tags: Way Below Relation, Continuous Lattices \begin{theorem} Let $L = \left({S, \vee, \wedge, \preceq}\right)$ be a bounded below continuous lattice. Let $x, z \in S$ such that :$x \ll z$ Then :$\exists y \in S: x \ll y \land y \ll z$ \end{theorem} \begin{proof} Case $x \ne z$: By Way Below has Strong Interpolation Property: :$\exists y \in S: x \ll y \land y \ll z \land x \ne y$ Thus :$\exists y \in S: x \ll y \land y \ll z$ Case $x = z$: Define $y = x$ Thus :$x \ll y \land y \ll z$ {{qed}} \end{proof}
23383
\section{Way Below has Strong Interpolation Property} Tags: Way Below Relation, Continuous Lattices, Approximating Relations \begin{theorem} Let $L = \left({S, \vee, \wedge, \preceq}\right)$ be a bounded below continuous lattice. Let $x, z \in S$ such that :$x \ll z \land x \ne z$ Then :$\exists y \in S: x \ll y \land y \ll z \land x \ne y$ \end{theorem} \begin{proof} By Way Below is Approximating Relation and Way Below Relation is Auxiliary Relation: :$\ll$ is an auxiliary approximating relation on $S$. Thus by Auxiliary Approximating Relation has Interpolation Property: :$\exists y \in S: x \ll y \land y \ll z \land x \ne y$ {{qed}} \end{proof}
23384
\section{Way Below if Between is Compact Set in Ordered Set of Topology} Tags: Way Below Relation, Topology \begin{theorem} Let $T = \struct {S, \tau}$ be a topological space. Let $L = \struct {\tau, \preceq}$ be an ordered set where $\preceq \mathop = \subseteq\restriction_{\tau \times \tau}$ Let $x, y \in \tau$ such that: :$\exists H \subseteq S: x \subseteq H \subseteq y \land H$ is compact Then: :$x \ll y$ \end{theorem} \begin{proof} Let $D$ be a directed subset of $\tau$ such that: :$y \preceq \sup D$ By proof of Topology forms Complete Lattice: :$\ds y \subseteq \bigcup D$ By Subset Relation is Transitive: :$\ds H \subseteq \bigcup D$ By definition: :$D$ is open cover of $H$ By definition of compact: :$H$ has finite subcover $\GG$ of $D$ By Directed iff Finite Subsets have Upper Bounds: :$\exists d \in D: d$ is upper bound for $\GG$ By definitions of upper bound and of $\preceq$: :$\forall z \in \GG: z \subseteq d$ By Union is Smallest Superset/Set of Sets: :$\ds \bigcup \GG \subseteq d$ By definition of cover: :$\ds H \subseteq \bigcup \GG$ By Subset Relation is Transitive: :$x \subseteq d$ Thus by definition of $\preceq$: :$x \preceq d$ Thus by definition of way below relation: :$x \ll y$ {{qed}} \end{proof}
23385
\section{Way Below iff Preceding Finite Supremum} Tags: Way Below Relation \begin{theorem} Let $\left({S, \vee, \wedge, \preceq}\right)$ be a complete lattice. Let $x, y \in S$. Then $x \ll y$ {{iff}} :$\forall X \subseteq S: y \preceq \sup X \implies \exists A \in {\it Fin}\left({X}\right): x \preceq \sup A$ where :$\ll$ denotes the way below relation, :${\it Fin}\left({X}\right)$ denotes the set of all finite subsets of $X$. \end{theorem} \begin{proof} === Sufficient Condition === Let $x \ll y$ Let $X \subseteq S$ such that :$y \preceq \sup X$ Define $F := \left\{ {\sup A: A \in {\it Fin}\left({X}\right)}\right\}$ By definition of union: :$X = \bigcup {\it Fin}\left({X}\right)$ By Supremum of Suprema: :$\sup X = \sup F$ We will prove that :$F$ is directed Let $a, b \in F$ By definition of $F$: :$\exists A \in {\it Fin}\left({X}\right): a = \sup A$ and :$\exists B \in {\it Fin}\left({X}\right): b = \sup B$ By Union of Subsets is Subset: :$A \cup B \subseteq X$ By Finite Union of Finite Sets is Finite: :$A \cup B$ is finite By definition of {\it Fin}: :$A \cup B \in {\it Fin}\left({X}\right)$ By definition of $F$: :$\sup\left({A \cup B}\right) \in F$ By Set is Subset of Union: :$A \subseteq A \cup B$ and $B \subseteq A \cup B$ By Supremum of Subset: :$a \preceq \sup\left({A \cup B}\right)$ and $b \preceq \sup\left({A \cup B}\right)$ Thus by definition :$F$ is directed By definition of way below relation: :$\exists d \in F: x \preceq d$ Thus by definition of $F$: :$\exists A \in {\it Fin}\left({X}\right): x \preceq \sup A$ {{qed|lemma}} \end{proof}
23386
\section{Way Below iff Second Operand Preceding Supremum of Ideal implies First Operand is Element of Ideal} Tags: Way Below Relation \begin{theorem} Let $\mathscr S = \struct {S, \preceq}$ be an up-complete ordered set. Let $x, y \in S$. Then $x \ll y$ {{iff}} :$\forall I \in \map {\operatorname {Ids} } {\mathscr S}: y \preceq \sup I \implies x \in I$ where :$\ll$ denotes the way below relation, :$\map {\operatorname {Ids} } {\mathscr S}$ denotes the set of all ideals in $\mathscr S$. \end{theorem} \begin{proof} === Sufficient Condition === Let $x \ll y$ Let $I \in \map {\operatorname {Ids} } {\mathscr S}$ such that :$y \preceq \sup I$ By definition of ideal: :$I$ is directed and lower. By definition of up-complete: :$I$ admits a supremum. By definition of way below relation: :$\exists i \in I: x \preceq i$ Thus by definition of lower set: :$x \in I$ {{qed|lemma}} \end{proof}
23387
\section{Way Below implies Preceding} Tags: Way Below Relation \begin{theorem} Let $\left({S, \preceq}\right)$ be an ordered set. Let $x, y \in S$ such that :$x \ll y$ where $\ll$ denotes element is way below second element. Then :$x \preceq y$ \end{theorem} \begin{proof} By Singleton is Directed and Filtered Subset: :$\left\{ {y}\right\}$ is directed. By Supremum of Singleton: :$\left\{ {y}\right\}$ admits a supremum and $\sup \left\{ {y}\right\} = y$ By definition of reflexivity: :$y \preceq \sup \left\{ {y}\right\}$ By definition of way below: :$\exists d \in \left\{ {y}\right\}: x \preceq d$ Thus by definition of singleton: :$x \preceq y$ {{qed}} \end{proof}
23388
\section{Way Below implies There Exists Way Below Open Filter Subset of Way Above Closure} Tags: Continuous Lattices \begin{theorem} Let $L = \left({S, \vee, \wedge, \preceq}\right)$ be a bounded below continuous lattice. Let $x, y \in S$ such that :$x \ll y$ where $\ll$ denotes the way below relation. Then there exists a way below open filter in $L$: $y \in F \land F \subseteq x^\gg$ where $x^\gg$ denotes the way above closure of $x$. \end{theorem} \begin{proof} We will prove that :$x^\gg$ is way below open. Let $z \in x^\gg$. By definition of way above closure: :$x \ll z$ By Way Below has Interpolation Property: :$\exists x' \in S: x \ll x' \land x' \ll z$ Thus by definition of way above closure: :$x' \in x^\gg$ Thus :$x' \ll z$ {{qed|lemma}} Then :$\forall z \in x^\gg: \exists y \in x^\gg: y \ll z$ By Axiom of Choice define a mapping $f: x^\gg \to x^\gg$: :$\forall z \in x^\gg: f\left({z}\right) \ll z$ By definition of way above closure: :$y \in x^\gg$ Define $V := \left\{ {z^\succeq: \exists n \in \N: z = f^n\left({y}\right)}\right\}$ We will prove that :$\forall X, Y \in V: \exists Z \in V: X \cup Y \subseteq Z$ Let $X, Y \in V$. By definition of $V$: :$\exists z_1 \in S: X = z_1^\succeq \land \exists n_1 \in \N: z_1 = f^{n_1}\left({y}\right)$ and :$\exists z_2 \in S: Y = z_2^\succeq \land \exists n_2 \in \N: z_2 = f^{n_2}\left({y}\right)$ We will prove that :$\forall n, k \in \N: f^{n+k}\left({y}\right) \preceq f^n\left({y}\right)$ Let $n \in \N$. '''Base Case''': By definition of reflexivity: :$f^{n+0}\left({y}\right) \preceq f^n\left({y}\right)$ '''Induction Hypothesis''': :$f^{n+k}\left({y}\right) \preceq f^n\left({y}\right)$ '''Induction Step''': By definition of $f$: :$f^{n+k+1}\left({y}\right) \ll f^{n+k}\left({y}\right)$ By Way Below implies Preceding: :$f^{n+k+1}\left({y}\right) \preceq f^{n+k}\left({y}\right)$ By Induction Hypothesis and definition of transitivity: :$f^{n+k+1}\left({y}\right) \preceq f^{n}\left({y}\right)$ {{qed|lemma}} WLOG: suppose $n_1 \le n_2$ Then :$\exists k \in \N: n_2 = n_1+k$ Then :$z_2 \preceq z_1$ By Preceding iff Meet equals Less Operand: :$z_1 \wedge z_2 = z_2$ By definition of $V$: :$Z := \left({z_1 \wedge z_2}\right)^\succeq \in V$ By Meet Precedes Operands: :$z_1 \wedge z_2 \preceq z_1$ and $z_1 \wedge z_2 \preceq z_2$ By Upper Closure is Decreasing: :$z_1^\succeq \subseteq Z$ and $z_2^\succeq \subseteq Z$ Thus by Union of Subsets is Subset: :$X \cup Y \subseteq Z$ {{qed|lemma}} Define $F := \bigcup V$. We will prove that :$F$ is way below open. Let $u \in F$. By definition of union: :$\exists Y \in V: u \in Y$ By definition of $V$: :$\exists z \in S: Y = z^\succeq \land \exists n \in \N: z = f^n\left({y}\right)$ By definition of $f$: :$z \in x^\gg$ By definition of $f$: :$f\left({z}\right) \ll z$ Then :$z' := f\left({z}\right) = f^{n+1}\left({y}\right)$ By definition of $V$: :${z'}^\succeq \in V$ By definition of reflexivity: :$z' \preceq z'$ By definition of upper closure of element: :$z' \in {z'}^\succeq$ By definition of union: :$z' \in F$ By definition of upper closure of element: :$z \preceq u$ By Preceding and Way Below implies Way Below: :$z' \ll u$ Hence :$\exists g \in F: g \ll u$ {{qed|lemma}} By Upper Closure of Element is Filter: :$\forall X \in V: X$ is a filtered upper set. By Union of Upper Sets is Upper: :$F$ is an upper set. By Union of Filtered Sets is Filtered: :$F$ is filtered. By definition of filter: :$F$ is a way below open filter in $L$. Then :$f^0\left({y}\right) = y$ By definition of $V$: :$y^\succeq \in V$ By definitions of upper closure of element and reflexivity: :$y \in y^\succeq$ By definition of union: :$y \in F$ It remains to prove that :$F \subseteq x^\gg$ Let $u \in F$. By definition of union: :$\exists Y \in V: u \in Y$ By definition of $V$: :$\exists z \in S: Y = z^\succeq \land \exists n \in \N: z = f^n\left({y}\right)$ By definition of $f$: :$z \in x^\gg$ By definition of way above closure: :$x \ll z$ By definition of upper closure of element: :$z \preceq u$ and $x \preceq x$ By Preceding and Way Below implies Way Below: :$x \ll u$ Thus by definition of way above closure: :$u \in x^\gg$ {{qed}} \end{proof}
23389
\section{Way Below in Lattice of Power Set} Tags: Way Below Relation, Power Set \begin{theorem} Let $X$ be a set. Let $L = \struct {\powerset X, \cup, \cap, \preceq}$ be a lattice of power set of $X$ where $\mathord\preceq = \mathord\subseteq \cap \paren {\powerset X \times \powerset X}$ Let $x, y \in \powerset X$. Then $x \ll y$ {{iff}} :for every a set $Y$ of subsets of $X$ such that $\ds y \subseteq \bigcup Y$ ::then there exists a finite subset $Z$ of $Y$: $\ds x \subseteq \bigcup Z$ where $\ll$ denotes the way below relation. \end{theorem} \begin{proof} === Sufficient Condition === Let $x \ll y$ Let $Y$ be a set of subsets of $X$ such that :$\ds y \subseteq \bigcup Y$ By definitions of power set and subset: :$Y \subseteq \powerset X$ By Power Set is Complete Lattice: :$\ds \bigcup Y = \sup Y$ By definition of $\preceq$: :$y \preceq \sup Y$ By Way Below in Complete Lattice: :there exists a finite subset $Z$ of $Y$: $x \preceq \sup Z$ By the proof of Power Set is Complete Lattice: :$\ds \bigcup Z = \sup Z$ Thus by definition of $\preceq$: :there exists a finite subset $Z$ of $Y$: $\ds x \subseteq \bigcup Z$ {{qed|lemma}} \end{proof}
23390
\section{Way Below is Approximating Relation} Tags: Way Below Relation, Continuous Lattices, Approximating Relations \begin{theorem} Let $L = \struct {S, \vee, \wedge, \preceq}$ be a bounded below continuous lattice. Then $\ll$ is an approximating relation on $S$. \end{theorem} \begin{proof} Let $x \in S$. Define $\RR := \mathord \ll$. By definitions of way below closure and $\RR$-segment: :$x^\ll = x^\RR$ where: :$x^\ll$ denotes the way below closure of $x$ :$x^\RR$ denotes the $\RR$-segment of $x$ By definition of continuous: :$L$ satisfies axiom of approximation. Thus by axiom of approximation: :$x = \map \sup {x^\ll} = \map \sup {x^\RR}$ Hence $\ll$ is an approximating relation on $S$. {{qed}} \end{proof}
23391
\section{Way Below is Congruent for Join} Tags: Way Below Relation \begin{theorem} Let $L = \left({S, \vee, \preceq}\right)$ be a join semilattice. Then $\ll$ is congruence relation for $\vee$: :$\forall a, b, x, y \in S: a \ll x \land b \ll y \implies a \vee b \ll x \vee y$ where $\ll$ denotes the way below relation. \end{theorem} \begin{proof} Let $a, b, x, y \in S$ such that $a \ll x$ and $b \ll y$ By Join Succeeds Operands: :$x \preceq x \vee y$ and $y \preceq x \vee y$ By Preceding and Way Below implies Way Below and definition of reflexivity: :$a \ll x \vee y$ and $b \ll x \vee y$ Thus by Join is Way Below if Operands are Way Below: :$a \vee b \ll x \vee y$ {{qed}} \end{proof}
23392
\section{Weak-* Limit in Normed Dual Space is Unique} Tags: Weak-* Convergence (Normed Vector Spaces) \begin{theorem} Let $\struct {X, \norm \cdot}$ be a normed vector space over $\Bbb F$. Let $\struct {X^\ast, \norm \cdot_{X^\ast} }$ be the normed dual space for $\struct {X, \norm \cdot}$. Let $f, g \in X^\ast$. Let $\sequence {f_n}_{n \mathop \in \N}$ be a sequence in $X^\ast$ such that: :$f_n \weakstarconv f$ and: :$f_n \weakstarconv g$ where $\weakstarconv$ denotes weak-$\ast$ convergence. Then: :$f = g$ \end{theorem} \begin{proof} By the definition of weak-$\ast$ convergence, we have: :$\map {f_n} x \to \map f x$ for each $x \in X$ and: :$\map {f_n} x \to \map g x$ for each $x \in X$. Then, from Convergent Complex Sequence has Unique Limit we have: :$\map f x = \map g x$ for each $x \in X$. That is: :$f = g$ {{qed}} \end{proof}
23393
\section{Weak Convergence in Hilbert Space} Tags: Weak Convergence in Hilbert Space, Hilbert Spaces, Weak Convergence (Normed Vector Spaces) \begin{theorem} Let $\struct {\HH, \innerprod \cdot \cdot}$ be a Hilbert space. Let $\sequence {x_n}_{n \mathop \in \N}$ be a sequence in $\HH$. Let $x \in X$. Then: :$\sequence {x_n}_{n \mathop \in \N}$ converges weakly to $x$ {{iff}}: :$\innerprod {x_n} y \to \innerprod x y$ for each $y \in \HH$. \end{theorem} \begin{proof} Let $\struct {\HH^\ast, \norm \cdot_{\HH^\ast} }$ be the normed dual space of $\HH$. \end{proof}
23394
\section{Weak Countable Compactness is not Preserved under Continuous Maps} Tags: Weakly Countably Compact Spaces \begin{theorem} Let $T_A = \struct {S_A, \tau_A}$ be a topological space which is weakly countably compact. Let $T_B = \struct {S_B, \tau_B}$ be another topological space. Let $\phi: T_A \to T_B$ be a continuous mapping. Then $T_B$ is not necessarily weakly countably compact. \end{theorem} \begin{proof} Let $\Z_{>0}$ be the strictly positive integers: :$\Z_{>0} = \set {1, 2, 3, \ldots}$ Let $T_A = \struct {\Z_{>0}, \tau_A}$ be the odd-even topology. Let $T_B = \struct {\Z_{>0}, \tau_B}$ be the discrete topology on $\Z_{>0}$. Let $\phi: T_A \to T_B$ be the mapping: :$\map \phi {2 k} = k, \map \phi {2 k - 1} = k$ Then: :$\map {\phi^{-1} } k = \set {2 k, 2 k - 1} \in \tau_A$ demonstrating that $\phi$ is continuous. Now we have that the Odd-Even Topology is Weakly Countably Compact. But we also have that a Countable Discrete Space is not Weakly Countably Compact. Hence the result. {{qed}} \end{proof}
23395
\section{Weak Existence of Matrix Logarithm} Tags: Matrix Algebra, Matrix Logarithm, Matrix Logarithms \begin{theorem} Let $T$ be a square matrix of order $n$. Let $\norm {T - I} < 1$ in the norm on bounded linear operators, where $I$ the identity matrix. Then there is a square matrix $S$ such that: :$e^S = T$ where $e^S$ is the matrix exponential. \end{theorem} \begin{proof} Define: :$\ds S = \sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n \paren {T - I}^n$ $S$ converges since $\norm {T - I} < 1$. We have that $\ds \sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n \norm {T - I}^n$ is the Newton-Mercator Series. This converges since $\norm {T - I} < 1$. Hence the series for $S$ converges absolutely, and so $S$ is well defined. Using the series definition for the matrix exponential: {{begin-eqn}} {{eqn | l = e^S | r = I + S + \frac 1 {2!} S^2 + \frac 1 {3!} S^3 + \cdots }} {{eqn | r = I + \sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n \paren {T - I}^n + \frac 1 {2!} \paren {\sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n \paren {T - I}^n}^2 + \frac 1 {3!} \paren {\sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n \paren {T - I}^n}^3 + \cdots }} {{eqn | ll= \leadsto | l = e^S | r = I + \paren {T - I} + c_2 \paren {T - I}^2 + c_3 \paren {T - I}^3 + \cdots | c = grouping terms by powers of $T - I$ }} {{eqn | r = T + c_2 \paren {T - I}^2 + c_3 \paren {T - I}^3 + \cdots }} {{end-eqn}} If $c_i = 0$ for $i \ge 2$, then $e^S = T$, and the result is shown. The Newton-Mercator Series is a Taylor expansion for $\map \ln {1 + x}$. When combined with the Power Series Expansion for Exponential Function, it gives: {{begin-eqn}} {{eqn | l = e^{\map \ln {1 + x} } | r = 1 + \map \ln {1 + x} + \frac 1 {2!} \paren {\map \ln {1 + x} }^2 + \frac 1 {3!} \paren {\map \ln {1 + x} }^3 + \cdots }} {{eqn | r = 1 + \sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n x^n + \frac 1 {2!} \paren {\sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n x^n}^2 + \frac 1 {3!} \paren {\sum_{n \mathop = 1}^\infty \frac {\paren {-1}^{n - 1} } n x^n}^3 + \cdots }} {{eqn | r = 1 + x + c_2 x^2 + c_3 x^3 + \cdots | c = grouping terms by powers of $x$ }} {{end-eqn}} But $e^{\map \ln {1 + x} } = 1 + x$. Thus: :$1 + x = 1 + x + c_2 x^2 + c_3 x^3 + \cdots \implies c_i = 0$ for $i \ge 2$. {{qed}} \end{proof}
23396
\section{Weak Inequality of Integers iff Strict Inequality with Integer plus One} Tags: Orderings on Integers, Integers \begin{theorem} Let $a, b \in \Z$ be integers. {{TFAE}} :$(1): \quad a \le b$ :$(2): \quad a < b + 1$ where: :$\le$ is the ordering on the integers :$<$ is the strict ordering on the integers. \end{theorem} \begin{proof} {{ProofWanted}} Category:Orderings on Integers \end{proof}
23397
\section{Weak Law of Large Numbers} Tags: Probability Theory \begin{theorem} Let $P$ be a population. Let $P$ have mean $\mu$ and finite variance. Let $\sequence {X_n}_{n \mathop \ge 1}$ be a sequence of random variables forming a random sample from $P$. Let: :$\ds {\overline X}_n = \frac 1 n \sum_{i \mathop = 1}^n X_i$ Then: :${\overline X}_n \xrightarrow p \mu$ where $\xrightarrow p$ denotes convergence in probability. \end{theorem} \begin{proof} Let $\sigma$ be the standard deviation of $P$. By the definition of convergence in probability, we aim to show that: :$\ds \lim_{n \mathop \to \infty} \map \Pr {\size { {\overline X}_n - \mu} < \varepsilon} = 1$ for all real $\varepsilon > 0$. Let $\varepsilon > 0$ be a real number. By Variance of Sample Mean: :$\var {{\overline X}_n} = \dfrac {\sigma^2} n$ By Chebyshev's Inequality, we have for real $k > 0$: :$\map \Pr {\size {{\overline X}_n - \mu} \ge \dfrac {k \sigma} {\sqrt n}} \le \dfrac 1 {k^2}$ As $\sigma > 0$ and $n > 0$, we can set: :$k = \dfrac {\sqrt n} {\sigma} \varepsilon$ This gives: :$\map \Pr {\size {{\overline X}_n - \mu} \ge \varepsilon} \le \dfrac {\sigma^2} {n \varepsilon^2}$ We therefore have: {{begin-eqn}} {{eqn | l = \map \Pr {\size { {\overline X}_n - \mu} < \varepsilon} | r = 1 - \map \Pr {\size { {\overline X}_n - \mu} \ge \varepsilon} }} {{eqn | o = \ge | r = 1 - \frac {\sigma^2} {n \varepsilon^2} }} {{end-eqn}} So: :$1 - \dfrac {\sigma^2} {n \varepsilon^2} \le \map \Pr {\size { {\overline X}_n - \mu} < \varepsilon} \le 1$ We have: :$\ds \lim_{n \mathop \to \infty} \paren {1 - \dfrac {\sigma^2} {n \varepsilon^2} } = 1$ and: :$\ds \lim_{n \mathop \to \infty} 1 = 1$ So by the Squeeze Theorem: :$\ds \lim_{n \mathop \to \infty} \map \Pr {\size { {\overline X}_n - \mu} < \varepsilon} = 1$ for all real $\varepsilon > 0$. {{qed}} \end{proof}
23398
\section{Weak Limit in Normed Vector Space is Unique} Tags: Weak Convergence (Normed Vector Space), Weak Convergence (Normed Vector Spaces) \begin{theorem} Let $\struct {X, \norm \cdot}$ be a normed vector space. Let $x, y \in X$. Let $\sequence {x_n}_{n \mathop \in \N}$ be a sequence in $X$ such that: :$x_n \weakconv x$ and: :$x_n \weakconv y$ where $\weakconv$ denotes weak convergence. Then: :$x = y$ \end{theorem} \begin{proof} Let $\struct {X^\ast, \norm \cdot_{X^\ast} }$ be the normed dual space of $\struct {X, \norm \cdot}$. Since: :$x_n \weakconv x$ we have: :$\map f {x_n} \to \map f x$ for each $f \in X^\ast$. Since: :$x_n \weakconv x$ we have: :$\map f {x_n} \to \map f y$ for each $f \in X^\ast$. From Convergent Complex Sequence has Unique Limit, we have: :$\map f x = \map f y$ for each $f \in X^\ast$. From Normed Dual Space Separates Points, we therefore have: :$x = y$ {{qed}} \end{proof}
23399
\section{Weak Local Compactness is Preserved under Open Continuous Surjection} Tags: Continuity, Local Compactnesss, Continuous Mappings, Compact Spaces, Weakly Locally Compact Spaces, Open Sets, Local Compactness, Locally Compact Spaces, Open Mappings \begin{theorem} Let $T_A = \struct {S_A, \tau_A}$ and $T_B = \struct {S_B, \tau_B}$ be topological spaces. Let $\phi: T_A \to T_B$ be a continuous mapping which is also an open mapping and a surjection. If $T_A$ is weakly locally compact, then $T_B$ is also weakly locally compact. \end{theorem} \begin{proof} Let $\phi$ be a mapping which is surjective, continuous and open. Let $T_A$ be weakly locally compact. Take $b \in S_B$. Let $V$ be a neighbourhood of $b$. Since $\phi$ is surjective: :$\forall y \in S_B: \exists x \in S_A: x \in \map {\phi^{-1} } y$ From the weak local compactness of $T_A$ and the continuity of $\phi$, there exists a compact neighbourhood $K$ of $x$ such that $\phi \sqbrk K \subseteq V$. Since $K$ is a neighbourhood of $x$, then $x \in K^\circ$ and $y \in \phi \sqbrk {K^\circ} \subseteq \phi \sqbrk K$, where $K^\circ$ is the interior of $K$. $\phi$ is an open mapping and $K^\circ$ is an open set, so $\phi \sqbrk {K^\circ}$ is also open. Finally we get that $y \in \phi \sqbrk K \subseteq V$, where $\phi \sqbrk K$ is a compact neighbourhood. Thus $T_B$ is weakly locally compact. {{qed}} \end{proof}
23400
\section{Weak Lower Closure in Restricted Ordering} Tags: Lower Closures, Order Theory \begin{theorem} Let $\left({S, \preccurlyeq}\right)$ be an ordered set. Let $T \subseteq S$ be a subset of $S$. Let $\preccurlyeq \restriction_T$ be the restricted ordering on $T$. Then for all $t \in T$: :$t^{\preccurlyeq T} = T \cap t^{\preccurlyeq S}$ where: :$t^{\preccurlyeq T}$ is the weak lower closure of $t$ in $\left({T, \preccurlyeq \restriction_T}\right)$ :$t^{\preccurlyeq S}$ is the weak lower closure of $t$ in $\left({S, \preccurlyeq}\right)$. \end{theorem} \begin{proof} Let $t \in T$, and suppose that $t' \in t^{\preccurlyeq T}$. By definition of weak lower closure $t^{\preccurlyeq T}$, this is equivalent to: :$t' \preccurlyeq \restriction_T t$ By definition of $\preccurlyeq \restriction_T$, this comes down to: :$t' \preccurlyeq t \land t' \in T$ as it is assumed that $t \in T$. The first conjunct precisely expresses that $t' \in t^{\preccurlyeq S}$. By definition of set intersection, it also holds that: :$t' \in T \cap t^{\preccurlyeq S}$ {{iff}} $t' \in T$ and $t' \in t^{\preccurlyeq S}$. Thus, it follows that the following are equivalent: :$t' \in t^{\preccurlyeq T}$ :$t' \in T \cap t^{\preccurlyeq S}$ and hence the result follows, by definition of set equality. {{qed}} Category:Lower Closures \end{proof}
23401
\section{Weak Nullstellensatz} Tags: Algebraic Geometry \begin{theorem} Let $K$ be an algebraically closed field. Let $n \ge 0$ be an natural number. Let $K \sqbrk {x_1, \ldots, x_n}$ be the polynomial ring in $n$ variables over $k$. Let $I \subseteq K \sqbrk {x_1,\ldots, x_n}$ be an ideal. {{TFAE}} # $I$ is the unit ideal: $I = (1)$. # Its zero-locus is empty set: $\map V I = \O$. \end{theorem} \begin{proof} {{proof wanted}} Category:Algebraic Geometry \end{proof}
23402
\section{Weak Solution to Dx u + 3yu = 0 with Heaviside Step Function Boundary Condition} Tags: Examples of Weak Solutions \begin{theorem} Consider the boundary value problem: :$\begin{cases} \dfrac {\partial u} {\partial x} + 3 y u = 0 & : x \in \R_{>0},~ y \in \R \\ & \\ \map u {0, y} = \map H y & : y \in \R \\ \end{cases}$ Then it has a weak solution of the form: :$u = e^{-3 y x} \map H y$ \end{theorem} \begin{proof} Let $u = e^{-3 y x} \map H y$ We have that: :Heaviside Step Function is Locally Integrable :Locally Integrable Function defines Distribution :Multiplication of Distribution induced by Locally Integrable Function by Smooth Function Hence, we can define a distribution $T_u \in \map {\DD'} {\R^2}$ associated with $u$. Then in the distributional sense we have that: {{begin-eqn}} {{eqn | l = \dfrac {\partial u}{\partial x} | r = \dfrac {\partial}{\partial x} \paren {e^{-3 y x} \map H y} }} {{eqn | r = e^{-3 y x} \paren {-3y} \map H y + e^{-3 y x} \dfrac {\partial \map H y}{\partial x} }} {{end-eqn}} That is: {{begin-eqn}} {{eqn | l = \dfrac {\partial T_u}{\partial x} | r = \dfrac {\partial}{\partial x} T_{e^{-3 y x} \map H y} }} {{eqn | r = \dfrac {\partial}{\partial x} \paren {e^{-3 y x} T_{\map H y} } | c = Multiplication of Distribution induced by Locally Integrable Function by Smooth Function }} {{eqn | r = e^{-3 y x} \paren {-3y} T_{\map H y} + e^{-3 y x} \dfrac {\partial}{\partial x} T_{\map H y} | c = Product Rule for Distributional Derivatives of Distributions multiplied by Smooth Functions }} {{end-eqn}} Let $\phi \in \map \DD {\R^2}$ be a test function. Then: {{begin-eqn}} {{eqn | l = \dfrac {\partial}{\partial x} \map {T_{\map H y} } \phi | r = - \map {T_{\map H y} } {\dfrac {\partial \phi}{\partial x} } }} {{eqn | r = - \int_{-\infty}^\infty \int_{-\infty}^\infty \map H y \dfrac {\partial \phi}{\partial x} \rd x \rd y }} {{eqn | r = - \int_0^\infty \int_{-\infty}^\infty \dfrac {\partial \phi}{\partial x} \rd x \rd y }} {{eqn | r = - \int_0^\infty \paren {\bigintlimits {\map \phi {x, y} } {x \mathop = - \infty} {x \mathop = \infty} } \rd y }} {{eqn | r = - \int_0^\infty 0 \rd y | c = {{Defof|Test Function}} }} {{eqn | r = 0 }} {{end-eqn}} Hence, in the distributional sense we have that: :$\dfrac \partial {\partial x} \map H y = \mathbf 0$ where $\mathbf 0$ is the zero distribution. Therefore: {{begin-eqn}} {{eqn | l = \dfrac {\partial u} {\partial x} + 3 y u | r = e^{-3xy} \paren {-3y} \map H y + 3y e^{-3yx} \map H y }} {{eqn | r = \mathbf 0 }} {{end-eqn}} Moreover: {{begin-eqn}} {{eqn | l = \map u {0, y} | r = e^{-0} \map H y }} {{eqn | r = \map H y }} {{end-eqn}} {{qed}} \end{proof}
23403
\section{Weak Solution to Dx u = Heaviside Step Function} Tags: Examples of Weak Solutions \begin{theorem} Let $H: \R \to \closedint 0 1$ be the Heaviside step function. Let $u : \R \to \R$ be such that: :$\map u x = \begin{cases} c & : x < 0 \\ x + c & : x > 0 \end{cases}$ where $c \in \R$. Let $T_u$ be the distribution associated with $u$. Then $u$ is a weak solution of: :$u' = H$ That is, in the distributional sense it holds that: :$\dfrac \d {\d x} T_u = T_H$ \end{theorem} \begin{proof} $u$ is continuous on $\R$ and continously differentiable on $\R \setminus \set 0$. For $x < 0$ we have $\map {u'} x = 0$. For $x > 0$ we have $\map {u'} x = 1$. That is: :$\map {u'} x = \map H x$ Furthermore: :$\ds \lim_{x \mathop \to 0^-} = 0$ :$\ds \lim_{x \mathop \to 0^+} = 1$ By the jump rule: {{begin-eqn}} {{eqn | l = T_u' | r = T_H + 0 \cdot \delta }} {{eqn | r = T_H }} {{end-eqn}} {{qed}} {{Stub|no classical solution; is this obvious?}} \end{proof}
23404
\section{Weak Solution to Dx u = u} Tags: Examples of Distributional Solutions, Examples of Weak Solutions \begin{theorem} Let $H$ be the Heaviside step function. Let $\map u {x, t} = \map H t e^x$ Then $u$ is a weak solution of the partial differential equation $\ds \dfrac {\partial u} {\partial x} = u$. That is, for the distribution $T_u \in \map {\DD'} {\R^2}$ associated with $u$ in the distributional sense it holds that: :$\ds \dfrac {\partial T_u} {\partial x} = T_u$ \end{theorem} \begin{proof} Let $\phi \in \map \DD {\R^2}$ be a test function. Then in the distributional sense we have that: {{begin-eqn}} {{eqn | l = \paren {\dfrac \partial {\partial x} - 1} \map {T_u} \phi | r = -\map {T_u} {\dfrac {\partial \phi} {\partial x} } - \map {T_u} \phi | c = {{Defof|Distributional Partial Derivative}} }} {{eqn | r = -\iint_{\R^2} \paren {\map H t e^x \dfrac {\partial \map \phi {x, t} } {\partial x} + \map H t e^x \map \phi {x, t} }\rd x \rd t | c = {{Defof|Distribution}} }} {{eqn | r = -\int_0^\infty \int_{-\infty}^\infty \paren {e^x \dfrac {\partial \map \phi {x, t} } {\partial x} + e^x \map \phi {x, t} }\rd x \rd t | c = {{Defof|Heaviside Step Function}} }} {{eqn | r = -\int_0^\infty \int_{-\infty}^\infty \dfrac \partial {\partial x} \paren {e^x \map \phi {x, t} }\rd x \rd t | c = Product Rule for Derivatives }} {{eqn | r = -\int_0^\infty \bigintlimits {e^x \map \phi {x, t} } {x \mathop = -\infty} {x \mathop = \infty} \rd t | c = Fundamental Theorem of Calculus }} {{eqn | r = -\int_0^\infty 0 \rd t | c = {{Defof|Test Function}} }} {{eqn | r = 0 }} {{end-eqn}} {{qed}} \end{proof}
23405
\section{Weak Upper Closure in Restricted Ordering} Tags: Upper Closures, Order Theory \begin{theorem} Let $\left({S, \preccurlyeq}\right)$ be an ordered set. Let $T \subseteq S$ be a subset of $S$. Let $\preccurlyeq \restriction_T$ be the restricted ordering on $T$. Then for all $t \in T$: :$t^{\succcurlyeq T} = T \cap t^{\succcurlyeq S}$ where: : $t^{\succcurlyeq T}$ is the weak upper closure of $t$ in $\left({T, \preccurlyeq \restriction_T}\right)$ : $t^{\succcurlyeq S}$ is the weak upper closure of $t$ in $\left({S, \preccurlyeq}\right)$. \end{theorem} \begin{proof} Let $t \in T$. Suppose that: :$t' \in t^{\succcurlyeq T}$ By definition of weak upper closure $t^{\succcurlyeq T}$, this is equivalent to: :$t \preccurlyeq \restriction_T t'$ By definition of $\preccurlyeq \restriction_T$, this comes down to: :$t \preccurlyeq t' \land t' \in T$ as it is assumed that $t \in T$. The first conjunct precisely expresses that $t' \in t^{\succcurlyeq S}$. By definition of set intersection, it also holds that: :$t' \in T \cap t^{\succcurlyeq S}$ {{iff}} $t' \in T$ and $t' \in t^{\succcurlyeq S}$. Thus it follows that the following are equivalent: :$t' \in t^{\succcurlyeq T}$ :$t' \in T \cap t^{\succcurlyeq S}$ and hence the result follows, by definition of set equality. {{qed}} Category:Upper Closures \end{proof}
23406
\section{Weak Whitney Immersion Theorem} Tags: Topology, Manifolds, Named Theorems \begin{theorem} Every $k$-dimensional manifold $X$ admits a one-to-one immersion in $\R^{2 k + 1}$. {{MissingLinks|$k$-dimensional}} \end{theorem} \begin{proof} Let $M > 2 k + 1$ be a natural number such that $f: X \to \R^M$ is an injective immersion. Define a map $h: X \times X \times \R \to \R^M$ by: :$\map h {x, y, t} = \map t {\map f x - \map f y}$ Define a map $g: \map T X \to \R^M$ by: :$\map g {x, v} = \map {\d f_x} v$ where $\map T X$ is the tangent bundle of $X$. Since $M > 2 k + 1$, the Morse-Sard Theorem implies $\exists a \in \R^M$ such that $a$ is in the image of neither function. Let $\pi$ be the projection of $\R^M$ onto the orthogonal complement of $a, H$. Suppose: :$\map {\paren {\pi \circ f} } x = \map {\paren {\pi \circ f} } y$ for some $x, y$. Then: :$\map f x - \map f y = t a$ for some scalar $t$. {{AimForCont}} $x \ne y$. Then as $f$ is injective: :$t \ne 0$ But then $\map h {x, y, 1/t} = a$, contradicting the choice of $a$. By Proof by Contradiction: :$x = y$ and so $\pi \circ f$ is injective. Let $v$ be a nonzero vector in $\map {T_x} X$ (the tangent space of $X$ at $x$) for which: :$\map {\d \paren {\pi \circ f}_x} v = 0$ Since $\pi$ is linear: :$\map {\d \paren {\pi \circ f}_x} = \pi \circ \d f_x$ Thus: :$\map {\pi \circ \d f_x} v = 0$ so: :$\map {\d f_x} v = t a$ for some scalar $t$. Because $f$ is an immersion, $t \ne 0$. Hence: :$\map g {x, 1/t} = a$ contradicting the choice of $a$. Hence $\pi \circ f: X \to H$ is an immersion. $H$ is obviously isomorphic to $\R^{M-1}$. Thus whenever $M > 2 k + 1$ and $X$ admits of a one-to-one immersion in $\R^M$, it follows that $X$ also admits of a one-to-one immersion in $\R^{M-1}$. \end{proof}
23407
\section{Weakly Convergent Sequence in Hilbert Space with Convergent Norm is Convergent} Tags: Hilbert Spaces, Weak Convergence (Normed Vector Spaces) \begin{theorem} Let $\struct {\HH, \innerprod \cdot \cdot}$ be a Hilbert space. Let $\sequence {x_n}_{n \mathop \in \N}$ be a sequence in $\HH$. Let $x \in X$ be such that: :$x_n \weakconv x$ and: :$\norm {x_n} \to \norm x$ where $\weakconv$ denotes weak convergence. Then: :$x_n \to x$ \end{theorem} \begin{proof} Let $\norm \cdot$ be the inner product norm for $\struct {\HH, \innerprod \cdot \cdot}$. From Sequence in Normed Vector Space Convergent to Limit iff Norm of Sequence minus Limit is Null Sequence, we have: :$x_n \to x$ {{iff}}: :$\norm {x_n - x} \to 0$ From Complex Sequence is Null iff Positive Integer Powers of Sequence are Null, it suffices to show that: :$\norm {x_n - x}^2 \to 0$ We have: {{begin-eqn}} {{eqn | l = \norm {x_n - x}^2 | r = \innerprod {x_n - x} {x_n - x} | c = {{Defof|Inner Product Norm}} }} {{eqn | r = \innerprod {x_n} {x_n} - \innerprod x {x_n} - \innerprod {x_n} x + \innerprod x x | c = Inner Product is Sesquilinear }} {{eqn | r = \norm {x_n}^2 - \innerprod x {x_n} - \innerprod {x_n} x + \norm x^2 | c = {{Defof|Inner Product Norm}} }} {{end-eqn}} From Weak Convergence in Hilbert Space, we have: :$\innerprod {x_n} x \to \innerprod x x = \norm x^2$ since $x_n$ converges weakly to $x$. From Weak Convergence in Hilbert Space: Corollary, we also have: :$\innerprod x {x_n} \to \innerprod x x = \norm x^2$ From hypothesis, we have: :$\norm {x_n}^2 \to \norm x$ Then, from Sum Rule for Real Sequences, we have: :$\norm {x_n - x}^2 \to 0$ so: :$x_n \to x$ {{qed}} \end{proof}
23408
\section{Weakly Convergent Sequence in Normed Dual Space is Weakly-* Convergent} Tags: Weak Convergence (Normed Vector Spaces), Weak-* Convergence (Normed Vector Spaces), Weak Convergence (Normed Vector Space), Weak-* Convergence (Normed Vector Space) \begin{theorem} Let $\mathbb F$ be a subfield of $\C$. Let $\struct {X, \norm {\, \cdot \,}_X}$ be a normed vector space over $\mathbb F$. Let $\struct {X^\ast, \norm {\, \cdot \,}_{X^\ast} }$ be the normed dual space of $\struct {X, \norm {\, \cdot \,}_X}$. Let $f \in X^\ast$. Let $\sequence {f_n}_{n \mathop \in \N}$ be a sequence in $X^\ast$ converging weakly to $f$. Then $\sequence {f_n}_{n \mathop \in \N}$ converges weakly-$\ast$ to $f$. \end{theorem} \begin{proof} Let $x \in X$. We aim to show that: :$\map {f_n} x \to \map f x$ Then, since $x \in X$ was arbitrary, we will obtain that $\sequence {f_n}_{n \mathop \in \N}$ converges weakly-$\ast$ to $f$. Since $\sequence {f_n}_{n \mathop \in \N}$ converges weakly to $f$, we have: :$\map F {f_n} \to \map F f$ for each $F \in \paren {X^\ast}^\ast$. Define $x^\wedge : X^\ast \to \mathbb F$ by: :$\map {x^\wedge} f = \map f x$ for each $f \in X^\ast$. From Evaluation Linear Transformation on Normed Vector Space is Linear Transformation from Space to Second Normed Dual, we have: :$x^\wedge \in \paren {X^\ast}^\ast$ So, setting $F = x^\wedge$, we obtain: :$\map {x^\wedge} {f_n} \to \map {x^\wedge} f$ That is: :$\map {f_n} x \to \map f x$ {{qed}} \end{proof}
23409
\section{Weakly Convergent Sequence in Normed Vector Space is Bounded} Tags: Weak Convergence (Normed Vector Spaces) \begin{theorem} Let $\struct {X, \norm {\, \cdot \,}_X}$ be a normed vector space. Let $\sequence {x_n}_{n \mathop \in \N}$ be a weakly convergent sequence in $X$. Then $\sequence {x_n}_{n \mathop \in \N}$ is bounded. \end{theorem} \begin{proof} Let $\struct {X^\ast, \norm {\, \cdot \,}_{X^\ast} }$ be the normed dual of $X$. Let $\struct {X^{\ast \ast}, \norm {\, \cdot \,}_{X^{\ast \ast} } }$ be the second normed dual of $X$. Let $J : X \to X^{\ast \ast}$ be the evaluation linear transformation on $X$. Let: :$x^\wedge = \map J x$ for each $x \in X$. Consider the sequence $\sequence {x_n^\wedge}_{n \mathop \in \N}$ in $X^{\ast \ast}$. Then, for each $f \in X^\ast$, we have: :$\map {x_n^\wedge} f = \map f {x_n}$ Since $\sequence {x_n}_{n \mathop \in \N}$ converges weakly we have that: :$\sequence {\map f {x_n} }_{n \mathop \in \N}$ converges for each $f \in X^\ast$. So, from Convergent Complex Sequence has Unique Limit, $\sequence {\map f {x_n} }_{n \mathop \in \N}$ is bounded. So, there exists a real number $M > 0$ such that: :$\cmod {\map f {x_n} } \le M$ for each $n \in \N$ and $f \in X^\ast$. That is: :$\cmod {\map {x_n^\wedge} f} \le M$ That is: :$\ds \sup_{n \mathop \in \N} \cmod {\map {x_n^\wedge} f}$ is finite for each $f \in X^\ast$. From the Banach-Steinhaus Theorem, we then obtain: :$\ds \sup_{n \mathop \in \N} \norm {x_n^\wedge}_{X^{\ast \ast} }$ is finite. From Evaluation Linear Transformation on Normed Vector Space is Linear Isometry, we have: :$\norm {x_n^\wedge}_{X^{\ast \ast} } = \norm {x_n}_X$ So: :$\ds \sup_{n \mathop \in \N} \norm {x_n}_X$ is finite. So we may conclude that: :$\sequence {x_n}_{n \mathop \in \N}$ is bounded. {{qed}} \end{proof}
23410
\section{Weakly Locally Compact Hausdorff Space is Strongly Locally Compact} Tags: Strongly Locally Compact Spaces, Hausdorff Spaces, Compact Spaces, Weakly Locally Compact Spaces, Local Compactness, Locally Compact Spaces \begin{theorem} Let $T = \struct {S, \tau}$ be a $T_2$ (Hausdorff) space. Let $T$ be weakly locally compact. Then $T$ is strongly locally compact. \end{theorem} \begin{proof} Let $x \in S$. As $T$ is weakly locally compact, $x$ is contained in a compact neighborhood $N_x$. As $T$ is a $T_2$ (Hausdorff) space, we can use the result Compact Subspace of Hausdorff Space is Closed. Thus the interior of $N_x$ has a closure which is compact. Hence the result, from definition of strongly locally compact space. {{qed}} \end{proof}
23411
\section{Wedderburn's Theorem} Tags: Division Rings, Fields, Named Theorems, Finite Fields, Galois Fields \begin{theorem} Every finite division ring $D$ is a field. \end{theorem} \begin{proof} Let $D$ be a finite division ring. If $D$ is shown commutative then, by definition, $D$ is a field. Let $\map Z D$ be the center of $D$, that is: :$\map Z D := \set {z \in D: \forall d \in D: z d = d z}$ From Center of Division Ring is Subfield it follows that $\map Z D$ is a Galois field. Thus from Characteristic of Galois Field is Prime the characteristic of $\map Z D$ is a prime number $p$. Let $\Z / \ideal p$ denote the quotient ring over the principal ideal $\ideal p$ of $\Z$. From Field of Prime Characteristic has Unique Prime Subfield, the prime subfield of $\map Z D$ is isomorphic to $\Z / \ideal p$. From Division Ring is Vector Space over Prime Subfield, $\map Z D$ is thus a vector space over $\Z / \ideal p$. From Vector Space over Division Subring is Vector Space, $D$ is a vector space over $\map Z D$. Since $\map Z D$ and $D$ are finite, both vector spaces are of finite dimension. Let $n$ and $m$ be the dimension of the two vector spaces respectively. It now follows from Cardinality of Finite Vector Space that $\map Z D$ has $p^n$ elements and $D$ has $\paren {p^n}^m$ elements. Now the idea behind the rest of the proof is as follows. We want to show $D$ is commutative. By definition, $\map Z D$ is commutative. Hence it is to be shown that $D = \map Z D$. It is shown that: :$\order D = \order {\map Z D}$ Hence $D = \map Z D$, and the proof is complete. $\map Z D$ and $D$ are considered as modules. We have that if $m = 1$ then: :$\order D = \order {\map Z D}$ and the result then follows. Thus it remains to show that $m = 1$. In a finite group, let $x_j$ be a representative of the conjugacy class $\tuple {x_j}$ (the representative does not matter). {{finish|Invoke Normalizer of Conjugate is Conjugate of Normalizer to formalise the independence of representative choice}} Let there be $l$ (distinct) non-singleton conjugacy classes. Let $\map {N_D} x$ denote the normalizer of $x$ with respect to $D$. Then we know by the Conjugacy Class Equation that: :$\ds \order D = \order {\map Z D} + \sum_{j \mathop = 0}^{l - 1} \index D {\map {N_D} {x_j} }$ which by Lagrange's theorem is: :$\ds \order D + \sum_{j \mathop = 1}^l \frac {\order D} {\order {\map {N_D} {x_j} } }$ Consider the group of units $\map U D$ in $D$. Consider what the above equation tells if we start with $\map U D$ instead of $D$. {{explain|We cannot take $D$ in the first place, since $D$ is not a group under multiplication. Doesn't it make sense to start with $\map U D$ directly? --Wandynsky (talk) 16:51, 30 July 2021 (UTC)}} If we centralize a multiplicative unit that is in the center, from Conjugacy Class of Element of Center is Singleton we get a singleton conjugacy class. Bear in mind that the above sum only considers non-singleton classes. Thus choose some element $u$ not in the center, so $\map {N_D} u$ is not $D$. However, $\map Z D \subset \map {N_D} u$ because any element in the center commutes with everything in $D$ including $u$. Then: :$\order {\map {N_D} u} = \paren {p^n}^m$ for $r < m$. Suppose there are $l$ such $u$. Then: {{begin-eqn}} {{eqn | l = \order {\map U D} | r = \order {\map Z {\map U D} } - 1 + \sum_{j \mathop = 1}^l \frac {\order D} {\order {\map {N_D} {u_j} } } | c = }} {{eqn | r = p^n - 1 + \sum_{\alpha_i} \frac {\paren {p^n}^m - 1} {\paren {p^n}^{\alpha_i} - 1} | c = }} {{end-eqn}} {{stub|Clean the following up. This bit is due to {{AuthorRef|Ernst Witt}}.}} We need two results to finish. :$(1):\quad$ If $p^k - 1 \divides p^j - 1$, then $k \divides j$ where $\divides$ denotes divisibility. :$(2)\quad$ If $j \divides k$ then $\Phi_n \divides \dfrac{x^j - 1} {x^k - 1}$ where $\Phi_n$ denotes the $n$th cyclotomic polynomial. {{refactor|The above two results need to be proved, on their own pages.}} {{AimForCont}} $m > 1$. Let $\gamma_i$ be an $m$th primitive root of unity. Then the above used conjugacy class theorem tells us how to compute size of $\map U D$ using non-central elements $u_j$. However, in doing so, we have that: :$\paren {q^n}^{\alpha_i} - 1 \divides \paren {q^n}^m - 1$ Thus by the first result: :$\alpha_i \divides m$ Thus: :$\Phi_m \divides \dfrac {x^m - 1} {x^{\alpha_i} - 1}$ However: :$\size {p^n - \gamma_i} > p^n - 1$ Thus the division is impossible. This contradicts our assumption that $m > 1$. Hence $m = 1$ and the result follows, as determined above. {{qed}} \end{proof}
23412
\section{Weierstrass's Theorem} Tags: Named Theorems: Weierstrass, Named Theorems, Analysis \begin{theorem} There exists a real function $f: \closedint 0 1 \to \closedint 0 1$ such that: :$(1): \quad f$ is continuous :$(2): \quad f$ is nowhere differentiable. \end{theorem} \begin{proof} Let $C \closedint 0 1$ denote the set of all real functions $f: \closedint 0 1 \to \R$ which are continuous on $\closedint 0 1$. By Continuous Function on Closed Interval is Complete, $C \closedint 0 1$ is a complete metric space under the supremum norm $\norm {\,\cdot \,}_\infty$. Let $X$ consist of the $f \in C \closedint 0 1$ such that: :$\map f 0 = 0$ :$\map f 1 = 1$ :$\forall x \in \closedint 0 1: 0 \le \map f x \le 1$ Then we have the following lemma: \end{proof}
23413
\section{Weierstrass-Casorati Theorem} Tags: Complex Analysis, Named Theorems \begin{theorem} Let $f$ be a holomorphic function defined on the open ball $\map B {a, r} \setminus \set a$. Let $f$ have an essential singularity at $a$. Then: :$\forall s < r: f \sqbrk {\map B {a, s} \setminus \set a}$ is a dense subset of $\C$. {{Disambiguate|Definition:Dense}} \end{theorem} \begin{proof} {{WLOG}}, suppose $a = 0$ and $r = 1$. {{AimForCont}} $\exists s < 1$ such that: :$f \sqbrk {\map B {0, s} \setminus \set 0}$ is not a dense subset of $\C$. Then, by definition of dense subset: :$\exists z_0 \in \C: \exists r_0 > 0: \map B {z_0, r_0} \cap f \sqbrk {\map B {0, s} \setminus \set 0} = \O$ {{explain|This definition of dense still needs to be added to {{ProofWiki}}. Whichever definition of denseness (either everywhere dense or dense-in-itself) will need to be expanded for the Complex case so as to make it clear that the above definition follows.}} Hence, the function $\varphi$ defined on $\map B {z_0, r_0}$ by: :$\map \varphi z = \dfrac 1 {\map f z - z_0}$ is analytic on $\map B {0, s} \setminus \set 0$ and bounded near to $0$, because: :$\forall z \in \map B {0, s} \setminus \set 0: \cmod {\map f z - z_0} > r_0 \implies \cmod {\map \varphi z} < \dfrac 1 {r_0}$ Therefore, we can extend the domain of $\varphi$ (using the Analytic Continuation Principle). Let $\map \varphi 0 \ne 0$. Then: :$\map f 0 = z_0 + \dfrac 1 {\map \varphi 0}$ and the singularity of $f$ was removable. Otherwise, let the power series of $\varphi$ be written: :$\ds \map \varphi z = \sum_{n \mathop = 1}^{+\infty} a_n z^n$ Then as $\varphi \ne 0$: :$E = \set {k \in \N: a_k \ne 0} \ne \O$ Let $p = \min E$. Then $0$ is a pole of order $p$ of $f$. In each case, the assumption that: :$\exists s < 1: f \sqbrk {\map B {0, s} \setminus \set 0}$ is not a dense subset of $\C$ contradicts the fact that $0$ is an essential singularity of $f$. Hence the result, by Proof by Contradiction. {{qed}} \end{proof}
23414
\section{Weierstrass Approximation Theorem} Tags: Weierstrass Approximation Theorem, Real Analysis \begin{theorem} Let $f$ be a real function which is continuous on the closed interval $\Bbb I$. Then $f$ can be uniformly approximated on $\Bbb I$ by a polynomial function to any given degree of accuracy. \end{theorem} \begin{proof} Let $\map f t: \Bbb I = \closedint a b \to \R$ be a continuous function. Introduce $\map x t$ with a rescaled domain: :$\map f t \mapsto \map x {a + t \paren {b - a} } : \closedint a b \to \closedint 0 1$ From now on we will work with $x: \closedint 0 1 \to \R$, which is also continuous. Let $n \in \N$. For $t \in \closedint 0 1$ consider the Bernstein polynomial: :$\ds \map {B_n x} t = \sum_{k \mathop = 0}^n \map x {\frac k n} \binom n k t^k \paren {1 - t}^{n - k}$ For $t \in \closedint 0 1$, $0 \le k \le n$, let: :$\map {p_{n, k} } t := \dbinom n k t^k \paren {1 - t}^{n - k}$ By the binomial theorem: :$\ds \sum_{k \mathop = 0}^n \map {p_{n, k} } t = 1$ \end{proof}
23415
\section{Weierstrass Approximation Theorem/Lemma 1} Tags: Weierstrass Approximation Theorem \begin{theorem} Let $\map {p_{n, k} } t : \N^2 \times \closedint 0 1 \to \R$ be a real valued function defined by: :$\map {p_{n, k} } t := \dbinom n k t^k \paren {1 - t}^{n - k}$ where: :$n, k \in \N$ :$t \in \closedint 0 1$ :$\dbinom n k$ denotes the binomial coefficient. Then: :$\ds \sum_{k \mathop = 0}^n k \map {p_{n, k} } t = n t$ \end{theorem} \begin{proof} From the binomial theorem: {{begin-eqn}} {{eqn | l = \paren {x + y}^n | r = \sum_{k \mathop = 0}^n \binom n k y^k x^{n - k} }} {{eqn | ll= \leadsto | l = 1 | r = \sum_{k \mathop = 0}^n \binom n k t^k \paren {1 - t}^{n - k} | c = $y = t, ~x = 1 - t$ }} {{eqn | l = 0 | r = \sum_{k \mathop = 0}^n \binom n k \paren {k t^{k - 1} \paren {1 - t}^{n - k} - t^k \paren{n - k} \paren {1 - t}^{n - k - 1} } | c = Derivative {{WRT|Differentiation}} $t$ }} {{eqn | r = \sum_{k \mathop = 0}^n k p_{n,k} \paren {\frac 1 t + \frac 1 {1 - t} } - \frac n {1 - t} }} {{eqn | ll= \leadsto | l = n t | r = \sum_{k \mathop = 0}^n k p_{n,k} }} {{end-eqn}} {{qed}} \end{proof}
23416
\section{Weierstrass Approximation Theorem/Lemma 2} Tags: Weierstrass Approximation Theorem \begin{theorem} Let $\map {p_{n, k} } t : \N^2 \times \closedint 0 1 \to \R$ be a real valued function defined as: :$\map {p_{n, k} } t := \dbinom n k t^k \paren {1 - t}^{n - k}$ where: :$n, k \in \N$ :$t \in \closedint 0 1$ :$\dbinom n k$ denotes a binomial coefficient. Then: :$\ds \sum_{k \mathop = 0}^n \paren {k - n t}^2 \map {p_{n, k} } t = n t \paren {1 - t}$ \end{theorem} \begin{proof} From the binomial theorem: :$\ds 1 = \sum_{k \mathop = 0}^n \binom n k t^k \paren {1 - t}^{n - k}$ From Lemma 1: {{begin-eqn}} {{eqn | l = n t | r = \sum_{k \mathop = 0}^n \binom n k k t^k \paren {1 - t}^{n - k} }} {{eqn | ll= \leadsto | l = n | r = \sum_{k \mathop = 0}^n \binom n k k \paren {k t^{k - 1} \paren {1 - t}^{n - k} - t^k \paren {n - k} \paren {1 - t}^{n - k - 1} } | c = {{Defof|Derivative of Real Function}} {{WRT|Differentiation}} $t$ }} {{eqn | r = \frac 1 t \sum_{k \mathop = 0}^n k^2 \map {p_{n, k} } t - \frac n {1 - t} \sum_{k \mathop = 0}^n k \map {p_{n, k} } t + \frac 1 {1 - t} \sum_{k \mathop = 0}^n k^2 \map {p_{n, k} } t }} {{eqn | r = \frac 1 {t \paren {1 - t} } \sum_{k \mathop = 0}^n k^2 \map {p_{n, k} } t - \frac n {1 - t} n t }} {{eqn | ll= \leadsto | l = \sum_{k \mathop = 0}^n k^2 \map {p_{n, k} } t | r = n t \paren {1 - t} + n^2 t^2 }} {{end-eqn}} Then: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 0}^n \paren {k - n t}^2 \map {p_{n, k} } t | r = \sum_{k \mathop = 0}^n \paren {k^2 - 2 n t k + n^2 t^2} \map {p_{n, k} } t }} {{eqn | r = n t \paren {1 - t} + n^2 t^2 - 2 n t n t + n^2 t^2 }} {{eqn | r = n t \paren {1 - t} }} {{end-eqn}} {{qed}} \end{proof}
23417
\section{Weierstrass Factorization Theorem} Tags: Complex Analysis, Analysis, Entire Functions, Named Theorems: Weierstrass, Infinite Products \begin{theorem} Let $f$ be an entire function. Let $0$ be a zero of $f$ of multiplicity $m \ge 0$. Let the sequence $\sequence {a_n}$ consist of the nonzero zeroes of $f$, repeated according to multiplicity. \end{theorem} \begin{proof} From Weierstrass Product Theorem, the function: :$\ds \map h z = z^m \prod_{n \mathop = 1}^\infty \map {E_{p_n} } {\frac z {a_n} }$ defines an entire function that has the same zeroes as $f$ counting multiplicity. Thus $f / h$ is both an entire function and non-vanishing. As $f / h$ is both holomorphic and nowhere zero there exists a holomorphic function $g$ such that: :$e^g = f / h$ Therefore: :$f = e^g h$ as desired. {{qed}} \end{proof}
23418
\section{Weierstrass Function is Continuous} Tags: Weierstrass Function \begin{theorem} Let $a \in \openint 0 1$. Let $b$ be a strictly positive odd integer such that: :$\ds a b > 1 + \frac 3 2 \pi$ Let $f: \R \to \R$ be a real function defined by: :$\ds \map f x = \sum_{n \mathop = 0}^\infty a^n \map \cos {b^n \pi x}$ for each $x \in \R$. Then $f$ is well-defined and continuous. \end{theorem} \begin{proof} Note that: :$\ds \sup_{x \mathop \in \R} \size {a^n \map \cos {b^n \pi x} } = a^n$ Since $a \in \openint 0 1$: :$\ds \sum_{n \mathop = 0}^\infty a^n$ converges. So, by the Weierstrass M-Test: :$\ds \sum_{n \mathop = 0}^\infty a^n \map \cos {b^n \pi x}$ converges uniformly on $\R$. That is, $f$ is well-defined. Further, from Uniformly Convergent Series of Continuous Functions Converges to Continuous Function: Corollary: :$f$ is continuous. {{qed}} Category:Weierstrass Function \end{proof}
23419
\section{Weierstrass M-Test} Tags: Named Theorems, Real Analysis, Convergence, Analysis \begin{theorem} Let $f_n$ be a sequence of real functions defined on a domain $D \subseteq \R$. Let $\ds \sup_{x \mathop \in D} \size {\map {f_n} x} \le M_n$ for each integer $n$ and some constants $M_n$ Let $\ds \sum_{i \mathop = 1}^\infty M_i < \infty$. Then $\ds \sum_{i \mathop = 1}^\infty f_i$ converges uniformly on $D$. \end{theorem} \begin{proof} Let: :$\ds S_n = \sum_{i \mathop = 1}^n f_i$ Let: :$\ds f = \lim_{n \mathop \to \infty} S_n$ To show the sequence of partial sums converge uniformly to $f$, we must show that: :$\ds \lim_{n \mathop \to \infty} \sup_{x \mathop \in D} \size {f - S_n} = 0$ But: {{begin-eqn}} {{eqn | l = \sup_{x \mathop \in D} \size {f - S_n} | r = \sup_{x \mathop \in D} \size {\paren {f_1 + f_2 + \dotsb} - \paren {f_1 + f_2 + \dotsb + f_n} } | c = }} {{eqn | r = \sup_{x \mathop \in D} \size {f_{n + 1} + f_{n + 2} + \dotsc} | c = }} {{end-eqn}} By the Triangle Inequality, this value is less than or equal to: :$\ds \sum_{i \mathop = n + 1}^\infty \sup_{x \mathop \in D} \size {\map {f_i} x} \le \sum_{i \mathop = n + 1}^\infty M_i$ We have that: :$\ds 0 \le \sum_{i \mathop = 1}^\infty M_n < \infty$ It follows from Tail of Convergent Series tends to Zero: :$\ds 0 \le \lim_{n \mathop \to \infty} \sum_{i \mathop = n + 1}^\infty \sup_{x \mathop \in D} \size {\map {f_i} x} \le \lim_{n \mathop \to \infty} \sum_{i \mathop = n + 1}^\infty M_i = 0$ So: :$\ds \lim_{n \mathop \to \infty} \sup_{x \mathop \in D} \size {f - S_n} = 0$ Hence the series converges uniformly on the domain. {{qed}} {{expand|Establish how broadly this can be applied - it's defined for real functions but should also apply to complex ones and possibly a general metric space.}} \end{proof}
23420
\section{Weierstrass Product Inequality} Tags: Inequalities, Analysis, Infinite Products \begin{theorem} For $n \ge 1$: :$\ds \prod_{i \mathop = 1}^n \paren {1 - a_i} \ge 1 - \sum_{i \mathop = 1}^n a_i$ where all of $a_i$ are in the closed interval $\closedint 0 1$. \end{theorem} \begin{proof} For $n = 1$ we have: :$1 - a_1 \ge 1 - a_1$ which is clearly true. Suppose the proposition is true for $n = k$, that is: :$\ds \prod_{i \mathop = 1}^k \paren {1 - a_i} \ge 1 - \sum_{i \mathop = 1}^k a_i$ Then: {{begin-eqn}} {{eqn | l = \prod_{i \mathop = 1}^{k + 1} \paren {1 - a_i} | r = \paren {1 - a_{k + 1} } \prod_{i \mathop = 1}^k \paren {1 - a_i} | c = }} {{eqn | o = \ge | r = \paren {1 - a_{k + 1} } \paren {1 - \sum_{i \mathop = 1}^k a_i} | c = as $1 - a_{k + 1} \ge 0$ }} {{eqn | r = 1 - a_{k + 1} - \sum_{i \mathop = 1}^k a_i + a_{k + 1} \sum_{i \mathop = 1}^k a_i | c = }} {{eqn | o = \ge | r = 1 - a_{k + 1} - \sum_{i \mathop = 1}^k a_i | c = as $a_i \ge 0$ }} {{eqn | r = 1 - \sum_{i \mathop = 1}^{k + 1} a_i | c = }} {{end-eqn}} Thus, by Principle of Mathematical Induction, the proof is complete. {{qed}} \end{proof}
23421
\section{Weierstrass Product Theorem} Tags: Complex Analysis, Infinite Products \begin{theorem} Let $\sequence {a_k}$ be a sequence of non-zero complex numbers such that: :$\cmod {a_n} \to \infty$ as $n \to \infty$ Let $\sequence {p_n}$ be a sequence of non-negative integers for which the series: :$\ds \sum_{n \mathop = 1}^\infty \size {\dfrac r {a_n} }^{1 + p_n}$ converges for every $r \in \R_{> 0}$. Let: :$\ds \map f z = \prod_{n \mathop = 1}^\infty \map {E_{p_n} } {\frac z {a_n} }$ where $E_{p_n}$ are Weierstrass elementary factors. Then $f$ is entire and its zeroes are the points $a_n$, counted with multiplicity. \end{theorem} \begin{proof} By: :Locally Uniformly Absolutely Convergent Product is Locally Uniformly Convergent :Infinite Product of Analytic Functions is Analytic :Zeroes of Infinite Product of Analytic Functions it suffices to show that the product $\ds \prod_{n \mathop = 1}^\infty \map {E_{p_n} } {\frac z {a_n} }$ converges locally uniformly absolutely. By Bounds for Weierstrass Elementary Factors and Weierstrass M-Test, this is the case. {{stub}} {{qed}} \end{proof}
23422
\section{Weight of Body at Earth's Surface} Tags: Weight (Physics), Weight \begin{theorem} Let $B$ be a body of mass $m$ situated at (or near) the surface of Earth. Then the weight of $B$ is given by: :$W = m g$ where $g$ is the value of the acceleration due to gravity at the surface of Earth. \end{theorem} \begin{proof} The weight of $B$ is the magnitude of the force exerted on it by the influence of the gravitational field it is in. By Newton's Second Law of Motion, that force is given by: :$\mathbf W = -m g \mathbf k$ where: :$g$ is the value of the acceleration due to gravity at the surface of Earth :$\mathbf k$ is a unit vector directed vertically upwards. Hence the magnitude of $\mathbf W$ is given by: :$W = \size {-m g \mathbf k} = m g$ {{qed}} \end{proof}
23423
\section{Weight of Discrete Topology equals Cardinality of Space} Tags: Discrete Topology \begin{theorem} Let $T = \struct {S, \tau}$ be a discrete topological space. Then: :$\map w T = \size S$ where: :$\map w T$ denotes the weight of $T$ :$\card S$ denotes the cardinality of $S$. \end{theorem} \begin{proof} By Basis for Discrete Topology the set $\BB = \set {\set x: x \in S}$ is a basis of $T$. By Set of Singletons is Smallest Basis of Discrete Space $\BB$ is smallest basis of $T$: :for every basis $\CC$ of $T$, $\BB \subseteq \CC$. Then by Subset implies Cardinal Inequality: :for every basis $\CC$ of $T$, $\card \BB \le \card \CC$. Hence $\card \BB$ is minimal cardinalty of basis of $T$: :$\map w T = \card \BB$ by definition of weight. Thus by Cardinality of Set of Singletons: :$\map w T = \card S$ {{qed}} \end{proof}
23424
\section{Weight of Sorgenfrey Line is Continuum} Tags: Sorgenfrey Line \begin{theorem} Let $T = \struct {\R, \tau}$ be the Sorgenfrey line. Then $\map w T = \mathfrak c$ where :$\map w T$ denotes the weight of $T$ :$\mathfrak c$ denotes continuum, the cardinality of real numbers. \end{theorem} \begin{proof} By definition of Sorgenfrey line, the set: :$\BB = \set {\hointr x y: x, y \in \R \land x < y}$ is a basis of $T$. By definition of weight: :$\map w T \le \card \BB$ where $\card \BB$ denotes the cardinality of $\BB$. By Cardinality of Basis of Sorgenfrey Line not greater than Continuum: :$\card \BB \le \mathfrak c$ Thus :$\map w T \le \mathfrak c$ It remains to show that: :$\mathfrak c \le \map w T$ {{AimForCont}} :$\mathfrak c \not \le \map w T$ Then: :$\map w T < \mathfrak c$ By definition of weight, there exists a basis $\BB_0$ of $T$: :$\map w T = \card {\BB_0}$ Then by Set of Subset of Reals with Cardinality less than Continuum has not Interval in Union Closure: :$\exists x, y \in \R: x < y \land \hointr x y \notin \set {\bigcup A: A \subseteq \BB_0} = \tau$ By definition of $\BB$: :$\hointr x y \in \BB \subseteq \tau$ Thus this contradicts by definition of subset with: :$\hointr x y \notin \tau$ {{qed}} \end{proof}
23425
\section{Well-Founded Induction} Tags: Set Theory, Axiom of Foundation \begin{theorem} Let $\struct {A, \RR}$ be a strictly well-founded relation. Let $\RR^{-1} \sqbrk x$ denote the preimage of $x$ for each $x \in A$. Let $B$ be a class such that $B \subseteq A$. Suppose that: :$(1): \quad \forall x \in A: \paren {\RR^{-1} \sqbrk x \subseteq B \implies x \in B}$ Then: :$A = B$ That is, if a property passes from the preimage of $x$ to $x$, then this property is true for all $x \in A$. \end{theorem} \begin{proof} {{AimForCont}} $A \nsubseteq B$. Then $A \setminus B \ne 0$. By Strictly Well-Founded Relation determines Strictly Minimal Elements, $A \setminus B$ must have some strictly minimal element under $\RR$. Then: :$\exists x \in A \setminus B: \paren {A \setminus B} \cap \RR^{-1} \sqbrk x = \O$ so: :$A \cap \RR^{-1} \sqbrk x \subseteq B$ Since $\RR^{-1} \sqbrk x \subseteq A$: :$\RR^{-1} \sqbrk x \subseteq B$ Thus, by hypothesis $(1)$: :$x \in B$ But this contradicts the fact that: :$x \in A \setminus B$ By Proof by Contradiction it follows that: :$A \setminus B = \O$ and so: :$A \subseteq B$ Therefore: :$A = B$ {{qed}} \end{proof}
23426
\section{Well-Founded Proper Relational Structure Determines Minimal Elements} Tags: Relational Closures, Relational Closure \begin{theorem} Let $A$ and $B$ be classes. Let $\struct {A, \prec}$ be a proper relational structure. Let $\prec$ be a strictly well-founded relation. Suppose $B \subset A$ and $B \ne \O$. Then $B$ has a strictly minimal element under $\prec$. \end{theorem} \begin{proof} $B$ is not empty. So $B$ has at least one element $x$. By Singleton of Element is Subset: :$\set x \subseteq B$ By Relational Closure Exists for Set-Like Relation: :$\set x$ has a $\prec$-relational closure. {{explain|Set-like relation: it needs to be shown that $\prec$ is one of those.}} This $\prec$-relational closure shall be denoted $y$. By Intersection is Subset: :$y \cap B \subseteq A$ As $x \in y$ and $x \in B$: :$y \cap B \ne \O$ By the definition of strictly well-founded relation: :$(1): \quad \exists x \in \paren {y \cap B}: \forall w \in \paren {y \cap B}: w \nprec x$ {{AimForCont}} that $w \in B$ such that $w \prec x$. Since $x \in y$, it follows that $w \in y$, so $w \in \paren {B \cap y}$ by the definition of intersection. This contradicts $(1)$. Therefore: :$w \nprec x$ and so $B$ has a strictly minimal element under $\prec$. {{qed}} \end{proof}
23427
\section{Well-Founded Relation has no Relational Loops} Tags: Well-Founded Relations \begin{theorem} Let $\RR$ be a well-founded relation on $S$. Let $x_1, x_2, \ldots, x_n \in S$. Then: :$\neg \paren {\paren {x_1 \mathrel \RR x_2} \land \paren {x_3 \mathrel \RR x_4} \land \cdots \land \paren {x_n \mathrel \RR x_1} }$ That is, there are no relational loops within $S$. \end{theorem} \begin{proof} Since $x_1, x_2, \ldots, x_n \in S$, there exists a non-empty subset $T$ of $S$ such that: :$T = \set {x_1, x_2, \ldots, x_n}$ By the definition of a well-founded relation: :$(1): \quad \exists z \in T: \forall y \in T \setminus z: \neg y \mathrel \RR z$ {{AimForCont}} $\paren {x_1 \mathrel \RR x_2} \land \paren {x_2 \mathrel \RR x_3} \land \cdots \land \paren {x_n \mathrel \RR x_1}$. We have that the elements of $T$ are $x_1, x_2, \ldots, x_n$. Hence: :$\forall y \in T: \exists z \in T \setminus z: y \mathrel \RR z$ This then contradicts $(1)$. Hence: :$\neg \paren {\paren {x_1 \mathrel \RR x_2} \land \paren {x_3 \mathrel \RR x_4} \land \cdots \land \paren {x_n \mathrel \RR x_1} }$ and so it follows that a well-founded relation has no relational loops. {{qed}} \end{proof}
23428
\section{Well-Founded Relation is not necessarily Ordering} Tags: Well-Founded Relations, Orderings, Order Theory \begin{theorem} Let $\struct {S, \RR}$ be a relational structure. Let $\RR$ be a well-founded relation on $S$. Then it is not necessarily the case that $\RR$ is also either an ordering or a strict ordering. \end{theorem} \begin{proof} Proof by Counterexample: Let $P$ be the set of all polynomials over $\R$ in one variable with real coefficients. Let $\DD$ be a relation on $P$ defined as: :$\forall p_0, p_1 \in P: \tuple {p_0, p_1} \in \DD$ {{iff}} $p_0$ is the derivative of $p_1$. From Differentiation of Polynomials induces Well-Founded Relation, we have that $\DD$ is a well-founded relation on $P$. Let $P_a$ be the polynomial defined as: :$P_a = x^3$ Then the derivative of $P_a$ {{WRT|Differentiation}} $x$ is: :$P_a' = \dfrac \d {\d x} x^3 = 3 x^2$ Then the derivative of $P_a$ {{WRT|Differentiation}} $x$ is: :$P_a'' = \dfrac \d {\d x} 3 x^2 = 6 x$ So we have that: :$\tuple {P_a', P_a} \in \DD$ and: :$\tuple {P_a'', P_a'} \in \DD$ but it is not the case that $\tuple {P_a'', P_a} \in \DD$. That is, $\DD$ is not transitive. It follows that $\DD$ is neither an ordering nor a strict ordering. {{qed}} \end{proof}
23429
\section{Well-Ordered Induction} Tags: Order Theory, Principle of Mathematical Induction, Well-Orderings, Mathematical Induction, Class Theory \begin{theorem} Let $\struct {A, \prec}$ be a strict well-ordering. For all $x \in A$, let the $\prec$-initial segment of $x$ be a small class. Let $B$ be a class such that $B \subseteq A$. Let: :$(1): \quad \forall x \in A: \paren {\paren {A \mathop \cap \map {\prec^{-1} } x} \subseteq B \implies x \in B}$ Then: :$A = B$ That is, if a property passes from the initial segment of $x$ to $x$, then this property is true for all $x \in A$. \end{theorem} \begin{proof} {{AimForCont}} that $A \nsubseteq B$. Then: :$A \setminus B \ne 0$. By Proper Well-Ordering Determines Smallest Elements, $A \setminus B$ must have some $\prec$-minimal element. Thus: :$\ds \exists x \in \paren {A \setminus B}: \paren {A \setminus B} \cap \map {\prec^{-1} } x = \O$ implies that: :$A \cap \map {\prec^{-1} } x \subseteq B$ Hence this fulfils the hypothesis for $(1)$. We have that $x \in A$. so by $(1)$: :$x \in B$ But this contradicts the fact that $x \in \paren {A \setminus B}$. Thus by Proof by Contradiction: :$A \subseteq B$ It follows by definition of set equality that: :$A = B$ {{qed}} \end{proof}
23430
\section{Well-Ordered Transitive Subset is Equal or Equal to Initial Segment} Tags: Well-Orderings, Order Theory \begin{theorem} Let $\struct {\prec, A}$ be a well-ordered set. For every $x \in A$, let every $\prec$-initial segment $A_x$ be a set. Let $B$ be a subclass of $A$ such that :$\forall x \in A: \forall y \in B: \paren {x \prec y \implies x \in B}$. That is, $B$ must be $\prec$-transitive. Then: :$A = B$ or: :$\exists x \in A: B = A_x$ \end{theorem} \begin{proof} Let $A \ne B$. Then $B \subsetneq A$. Therefore, by Set Difference with Proper Subset: :$A \setminus B \ne \O$ Then: {{begin-eqn}} {{eqn | l = A \setminus B | o = \ne | r = \O }} {{eqn | ll= \leadsto | q = \exists x \in A \setminus B | l = \paren {A \setminus B} \cap A_x | r = \O | c = Proper Well-Ordering Determines Smallest Elements }} {{eqn | ll= \leadsto | q = \exists x \in A \setminus B | l = A \cap A_x | o = \subseteq | r = B | c = Set Difference with Superset is Empty Set }} {{end-eqn}} One direction of inclusion is proven. By the hypothesis: {{begin-eqn}} {{eqn | l = x \in A \land x \prec y \land y \in B | o = \implies | r = x \in B }} {{end-eqn}} But $x \in A \land x \notin B$, so: {{begin-eqn}} {{eqn | l = y | o = \in | r = B }} {{eqn | ll= \leadsto | l = \neg x | o = \prec | r = y | c = Modus Tollendo Tollens and other propositional manipulations }} {{eqn | ll= \leadsto | l = y | o = \prec | r = x | c = $\prec$ is totally ordered and $y \ne x$ }} {{end-eqn}} Therefore: :$B \subseteq A_x$ and so :$B \subseteq A \cap A_x$ {{qed}} \end{proof}
23431
\section{Well-Ordering Minimal Elements are Unique} Tags: Well-Orderings \begin{theorem} Let $\struct {S,\preceq}$ be a well-ordered set. Then every non-empty subset of $S$ has a unique minimal element. \end{theorem} \begin{proof} The proof consists of a uniqueness and an existence part. Let $S'$ be a non-empty subset of $S$. \end{proof}
23432
\section{Well-Ordering Principle} Tags: Number Theory, Well-Ordering Principle, Ordering on Natural Numbers, Natural Numbers, Named Theorems, Well-Orderings \begin{theorem} Every non-empty subset of $\N$ has a smallest (or '''first''') element. That is, the relational structure $\struct {\N, \le}$ on the set of natural numbers $\N$ under the usual ordering $\le$ forms a well-ordered set. This is called the '''well-ordering principle'''. \end{theorem} \begin{proof} Consider the natural numbers $\N$ defined as the naturally ordered semigroup $\struct {S, \circ, \preceq}$. From its definition, $\struct {S, \circ, \preceq}$ is well-ordered by $\preceq$. The result follows. As $\N_{\ne 0} = \N \setminus \set 0$, by Set Difference is Subset $\N_{\ne 0} \subseteq \N$. As $\N$ is well-ordered, by definition, every subset of $\N$ has a smallest element. {{qed}} \end{proof}
23433
\section{Well-Ordering Theorem} Tags: Set Theory, Named Theorems, Ordinals, Axiom of Choice \begin{theorem} Every set is well-orderable. \end{theorem} \begin{proof} Let $S$ be a set. Let $\powerset S$ be the power set of $S$. By the Axiom of Choice, there is a choice function $c$ defined on $\powerset S \setminus \set \O$. We will use $c$ and the Principle of Transfinite Induction to define a bijection between $S$ and some ordinal. Intuitively, we start by pairing $\map c S$ with $0$, and then keep extending the bijection by pairing $\map c {S \setminus X}$ with $\alpha$, where $X$ is the set of elements already dealt with. \end{proof}
23434
\section{Well-Ordering Theorem implies Hausdorff Maximal Principle} Tags: Well-Orderings \begin{theorem} Let the Well-Ordering Theorem hold. Then the Hausdorff Maximal Principle holds. \end{theorem} \begin{proof} Let $X$ be a non-empty set. Let $X$ contain at least two elements; otherwise, any non-empty ordering on $X$ is trivially a maximal chain. By the Well-Ordering Theorem, $X$ can be well-ordered. Fix such a well-ordering. Let $\le$ be any ordering on $X$. Let $\map P {a, Y}$ be the predicate: :$a$ is $\le$-comparable with every $y \in Y$. Here, $a$ and $Y$ are bound variables. That is, $\map P {a, Y}$ holds if, for every $y \in Y$, $a \le y$ or $y \le a$. Define the mapping: :$\map \rho {f: S_x \to \powerset X} = \begin {cases} f \sqbrk {S_x} \cup \set x & : \map P {x, S_x} \\ f \sqbrk {S_x} & : \text {otherwise} \end{cases}$ where $S_x$ is the initial segment in $X$ defined by $x$, and $\map f {\,\cdot\,}$ denotes an image set. Using the Principle of Recursive Definition for Well-Ordered Sets, we can use $\rho$ to uniquely define: :$h: X \to \powerset X$: :$\map h \alpha = \map \rho {h {\restriction_{S_\alpha} } } = \begin {cases} h \sqbrk {S_\alpha} \cup \set \alpha & : \map P {\alpha, S_\alpha} \\ h \sqbrk {S_\alpha} & : \text {otherwise} \end{cases}$ Then $\map h \alpha$ is a $\le$-totally ordered set, by virtue of the construction of $h$ in terms of the predicate $P$. Similarly, $h \sqbrk {S_\alpha}$ is totally ordered, for the same reason. Then the union: :$Z = \ds \bigcup_{\alpha \mathop \in X} \map h \alpha$ admits an ordered sum in terms of $\le$ imposed on each $\map h \alpha$. By Ordered Sum of Tosets is Totally Ordered Set, this is a totally ordered set. In particular, it is a chain of $\struct {X, \le}$. We claim this chain is maximal. To see this, consider $x_0 \notin \ds \bigcup_{\alpha \mathop \in X} \map h \alpha$. Then $x_0$ is not $\le$-comparable with the elements of $Z$, because $\neg \map P {x, Z}$. That is, there are no proper supersets of $Z$ that are totally ordered. Therefore the ordered sum on $Z$ is a maximal chain in $\struct {X, \le}$. {{qed}} \end{proof}
23435
\section{Westwood's Puzzle} Tags: Euclidean Geometry, Named Theorems \begin{theorem} :500px Take any rectangle $ABCD$ and draw the diagonal $AC$. Inscribe a circle $GFJ$ in one of the resulting triangles $\triangle ABC$. Drop perpendiculars $IEF$ and $HEJ$ from the center of this incircle $E$ to the sides of the rectangle. Then the area of the rectangle $DHEI$ equals half the area of the rectangle $ABCD$. \end{theorem} \begin{proof} Construct the perpendicular from $E$ to $AC$, and call its foot $G$. Call the intersection of $IE$ and $AC$ $K$, and the intersection of $EH$ and $AC$ $L$. :500px {{begin-eqn}} {{eqn|l=\angle CKI|r=\angle EKG|c=By Two Straight Lines make Equal Opposite Angles}} {{eqn|l=\angle EGK|r=\mbox{Right Angle}|c=By Tangent to Circle is Perpendicular to Radius}} {{eqn|l=\angle KIC|r=\mbox{Right Angle}|c=Because $IF \perp CD$}} {{eqn|l=\angle EGK|r=\angle KIC|c=By Euclid's Fourth Postulate}} {{eqn|l=IC|r=EJ|c=By Parallel Lines are Everywhere Equidistant}} {{eqn|l=EJ|r=EG|c=Because both are radii of the same circle}} {{eqn|l=IC|r=EG|c=By Euclid's First Common Notion}} {{eqn|l=\mbox{Area}\triangle IKC|r=\mbox{Area}\triangle GKE|c=By Triangle Angle-Angle-Side Equality}} {{eqn|l=\angle HLA|r=\angle GLE|c=By Two Straight Lines make Equal Opposite Angles}} {{eqn|l=\angle EGL|r=\mbox{Right Angle}|c=By Tangent to Circle is Perpendicular to Radius}} {{eqn|l=\angle AHL|r=\mbox{Right Angle}|c=Because $HJ \perp AD$}} {{eqn|l=\angle EGL|r=\angle AHL|c=By Euclid's Fourth Postulate}} {{eqn|l=HA|r=EF|c=By Parallel Lines are Everywhere Equidistant}} {{eqn|l=EF|r=EG|c=Because both are radii of the same circle}} {{eqn|l=HA|r=EG|c=By Euclid's First Common Notion}} {{eqn|l=\mbox{Area}\triangle HAL|r=\mbox{Area}\triangle GEL|c=By Triangle Angle-Angle-Side Equality}} {{eqn|l=\mbox{Area}\triangle ADC|r=\frac{AD\cdot CD} 2|c=By Area of a Triangle in Terms of Side and Altitude}} {{eqn|l=\frac{\mbox{Area}\Box ABCD} 2|r=\frac{AD\cdot CD} 2|c=By Area of a Parallelogram}} {{eqn|l=\frac{\mbox{Area}\Box ABCD} 2|r=\mbox{Area}\triangle ADC|c=By Euclid's First Common Notion}} {{eqn|r=\mbox{Area}\triangle HAL + \mbox{Area}\triangle IKC + \mbox{Area}\Box DHLKI}} {{eqn|r=\mbox{Area}\triangle GEL + \mbox{Area}\triangle GKE+ \mbox{Area}\Box DHLKI}} {{eqn|r=\mbox{Area}\Box DHEI}} {{end-eqn}} {{qed}} \end{proof}
23436
\section{Westwood's Puzzle/Proof 1} Tags: Euclidean Geometry, Named Theorems \begin{theorem} :500px Take any rectangle $ABCD$ and draw the diagonal $AC$. Inscribe a circle $GFJ$ in one of the resulting triangles $\triangle ABC$. Drop perpendiculars $IEF$ and $HEJ$ from the center of this incircle $E$ to the sides of the rectangle. Then the area of the rectangle $DHEI$ equals half the area of the rectangle $ABCD$. \end{theorem} \begin{proof} Construct the perpendicular from $E$ to $AC$, and call its foot $G$. Let $K$ be the intersection of $IE$ and $AC$. Let $L$ be the intersection of $EH$ and $AC$. :500px First we have: {{begin-eqn}} {{eqn | n = 1 | l = \angle CKI | r = \angle EKG | c = Two Straight Lines make Equal Opposite Angles }} {{eqn | l = \angle EGK | r = \text {Right Angle} | c = Tangent to Circle is Perpendicular to Radius }} {{eqn | l = \angle KIC | r = \text {Right Angle} | c = as $IF \perp CD$ }} {{eqn | n = 2 | ll= \therefore | l = \angle EGK | r = \angle KIC | c = Euclid's Fourth Postulate }} {{eqn | l = IC | r = EJ | c = Opposite Sides and Angles of Parallelogram are Equal }} {{eqn | l = EJ | r = EG | c = as both are radii of the same circle }} {{eqn | n = 3 | ll= \therefore | l = IC | r = EG | c = Euclid's First Common Notion }} {{eqn | ll= \therefore | l = \Area \triangle IKC | r = \Area \triangle GKE | c = Triangle Angle-Angle-Side Equality: $(1)$, $(2)$ and $(3)$ }} {{end-eqn}} Similarly: {{begin-eqn}} {{eqn | n = 4 | l = \angle HLA | r = \angle GLE | c = Two Straight Lines make Equal Opposite Angles }} {{eqn | l = \angle EGL | r = \text {Right Angle} | c = Tangent to Circle is Perpendicular to Radius }} {{eqn | l = \angle AHL | r = \text {Right Angle} | c = as $HJ \perp AD$ }} {{eqn | n = 5 | ll= \therefore | l = \angle EGL | r = \angle AHL | c = Euclid's Fourth Postulate }} {{eqn | l = HA | r = EF | c = Opposite Sides and Angles of Parallelogram are Equal }} {{eqn | l = EF | r = EG | c = as both are radii of the same circle }} {{eqn | n = 6 | ll= \therefore | l = HA | r = EG | c = Euclid's First Common Notion }} {{eqn | ll= \therefore | l = \Area \triangle HAL | r = \Area \triangle GEL | c = Triangle Angle-Angle-Side Equality: $(4)$, $(5)$ and $(6)$ }} {{end-eqn}} Finally: {{begin-eqn}} {{eqn | l = \frac {\Area \Box ABCD} 2 | r = \frac {AD \cdot CD} 2 | c = Area of Parallelogram }} {{eqn | r = \Area \triangle ADC | c = Area of Triangle in Terms of Side and Altitude }} {{eqn | r = \Area \triangle HAL + \Area \triangle IKC + \Area \Box DHLKI }} {{eqn | r = \Area \triangle GEL + \Area \triangle GKE + \Area \Box DHLKI }} {{eqn | r = \Area \Box DHEI }} {{end-eqn}} {{qed}} {{Namedfor|Matt Westwood}} Category:Euclidean Geometry \end{proof}
23437
\section{Westwood's Puzzle/Proof 2} Tags: Euclidean Geometry, Named Theorems \begin{theorem} :500px Take any rectangle $ABCD$ and draw the diagonal $AC$. Inscribe a circle $GFJ$ in one of the resulting triangles $\triangle ABC$. Drop perpendiculars $IEF$ and $HEJ$ from the center of this incircle $E$ to the sides of the rectangle. Then the area of the rectangle $DHEI$ equals half the area of the rectangle $ABCD$. \end{theorem} \begin{proof} The crucial geometric truth to note is that: :$CJ = CG, AG = AF, BF = BJ$ This follows from the fact that: :$\triangle CEJ \cong \triangle CEG$, $\triangle AEF \cong \triangle AEG$ and $\triangle BEF \cong \triangle BEJ$ This is a direct consequence of the point $E$ being the center of the incircle of $\triangle ABC$. Then it is just a matter of algebra. Let $AF = a, FB = b, CJ = c$. {{begin-eqn}} {{eqn | l = \paren {a + b}^2 + \paren {b + c}^2 | r = \paren {a + c}^2 | c = Pythagoras's Theorem }} {{eqn | ll= \leadsto | l = a^2 + 2 a b + b^2 + b^2 + 2 b c + c^2 | r = a^2 + 2 a c + c^2 | c = }} {{eqn | ll= \leadsto | l = a b + b^2 + b c | r = a c | c = }} {{eqn | ll= \leadsto | l = a b + b^2 + b c + a c | r = 2 a c | c = }} {{eqn | ll= \leadsto | l = \paren {a + b} \paren {b + c} | r = 2 a c | c = }} {{end-eqn}} {{qed}} {{Namedfor|Matt Westwood}} Category:Euclidean Geometry \end{proof}
23438
\section{Whitney Embedding Theorem} Tags: Topology, Manifolds \begin{theorem} Every smooth $m$-dimensional manifold admits a smooth embedding into Euclidean space $\R^{2m+1}$. \end{theorem} \begin{proof} {{ProofWanted}} {{Namedfor|Hassler Whitney|cat = Whitney}} Category:Manifolds \end{proof}
23439
\section{Whitney Immersion Theorem} Tags: Differential Topology \begin{theorem} Let $m > 1$ be a natural number. Every smooth $m$-dimensional manifold can be immersed in Euclidean $\left({2m-1}\right)$-space. \end{theorem} \begin{proof} {{ProofWanted}} {{Namedfor|Hassler Whitney|cat = Whitney}} Category:Differential Topology \end{proof}
23440
\section{Whole Space is Open in Neighborhood Space} Tags: Neighborhood Spaces \begin{theorem} Let $\struct {S, \NN}$ be a neighborhood space. Then $S$ itself is an open set of $\struct {S, \NN}$. \end{theorem} \begin{proof} Let $x \in S$. Then by neighborhood space axiom $\text N 1$ there exists a neighborhood $N$ of $x$. As $N \subseteq S$ it follows from neighborhood space axiom $\text N 3$ that $S$ is a neighborhood of $x$. As this holds for all $x \in S$ it follows that $S$ is an open set of $\struct {S, \NN}$. {{qed}} \end{proof}
23441
\section{Wholly Real Number and Wholly Imaginary Number are Linearly Independent over the Rationals} Tags: Complex Analysis \begin{theorem} Let $z_1$ be a non-zero wholly real number. Let $z_2$ be a non-zero wholly imaginary number. Then, $z_1$ and $z_2$ are linearly independent over the rational numbers $\Q$, where the group is the complex numbers $\C$. \end{theorem} \begin{proof} From Rational Numbers form Subfield of Complex Numbers, the unitary module $\struct {\C, +, \times}_\Q$ over $\Q$ satisfies the unitary module axioms: * $\Q$-Action: $\C$ is closed under multiplication, so $\Q \times \C \subset \C$. * Distributive: $\times$ distributes over $+$. * Associativity: $\times$ is associative. * Multiplicative Identity: $1$ is the multiplicative identity in $\C$. Let $a, b \in \Q$ such that: :$a z_1 + b z_2 = 0$ By the definition of wholly imaginary, there is a real number $c$ such that: :$c i = z_2$ where $i$ is the imaginary unit. Therefore: :$a z_1 + b c i = 0$ Equating real parts and imaginary parts: :$a z_1 = 0$ :$b c = 0$ Since $z_1$ and $c$ are both non-zero, $a$ and $b$ must be zero. The result follows from the definition of linear independence. {{qed}} Category:Complex Analysis \end{proof}
23442
\section{Wilson's Theorem} Tags: Factorials, Number Theory, Wilson's Theorem, Named Theorems, Prime Numbers, Modulo Arithmetic \begin{theorem} A (strictly) positive integer $p$ is a prime {{iff}}: :$\paren {p - 1}! \equiv -1 \pmod p$ \end{theorem} \begin{proof} If $p = 2$ the result is obvious. Therefore we assume that $p$ is an odd prime. \end{proof}
23443
\section{Wilson's Theorem/Corollary 1} Tags: Wilson's Theorem \begin{theorem} Let $p$ be a prime number. Then $p$ is the smallest prime number which divides $\paren {p - 1}! + 1$. \end{theorem} \begin{proof} From Wilson's Theorem, $p$ divides $\paren {p - 1}! + 1$. Let $q$ be a prime number less than $p$. Then $q$ is a divisor of $\paren {p - 1}!$ and so does not divide $\paren {p - 1}! + 1$. {{qed}} \end{proof}
23444
\section{Wilson's Theorem/Corollary 2} Tags: Prime Numbers, Factorials, Modulo Arithmetic, Wilson's Theorem \begin{theorem} Let $n \in \Z_{>0}$ be a (strictly) positive integer. Let $p$ be a prime factor of $n!$ with multiplicity $\mu$. Let $n$ be expressed in a base $p$ representation as: {{begin-eqn}} {{eqn | l = n | r = \sum_{j \mathop = 0}^m a_j p^j | c = where $0 \le a_j < p$ }} {{eqn | r = a_0 + a_1 p + a_2 p^2 + \cdots + a_m p^m | c = for some $m > 0$ }} {{end-eqn}} Then: :$\dfrac {n!} {p^\mu} \equiv \paren {-1}^\mu a_0! a_1! \dotsb a_m! \pmod p$ \end{theorem} \begin{proof} Proof by induction: Let $\map P n$ be the proposition: :$\dfrac {n!} {p^\mu} \equiv \paren {-1}^\mu a_0! a_1! \dotsm a_k! \pmod p$ where $p, a_0, \dots, a_k, \mu$ are as defined above. \end{proof}
23445
\section{Wilson's Theorem/Necessary Condition} Tags: Wilson's Theorem \begin{theorem} Let $p$ be a prime number. Then: :$\paren {p - 1}! \equiv -1 \pmod p$ \end{theorem} \begin{proof} If $p = 2$ the result is obvious. Therefore we assume that $p$ is an odd prime. Let $p$ be prime. Consider $n \in \Z, 1 \le n < p$. As $p$ is prime, $n \perp p$. From Law of Inverses (Modulo Arithmetic), we have: :$\exists n' \in \Z, 1 \le n' < p: n n' \equiv 1 \pmod p$ By Solution of Linear Congruence, for each $n$ there is '''exactly one''' such $n'$, and $\paren {n'}' = n$. So, provided $n \ne n'$, we can pair any given $n$ from $1$ to $p$ with another $n'$ from $1$ to $p$. We are then left with the numbers such that $n = n'$. Then we have $n^2 \equiv 1 \pmod p$. Consider $n^2 - 1 = \paren {n + 1} \paren {n - 1}$ from Difference of Two Squares. So either $n + 1 \divides p$ or $n - 1 \divides p$. Observe that these cases do not occur simultaneously, as their difference is $2$, and $p$ is an odd prime. From Congruence Modulo Negative Number, we have that $p - 1 \equiv -1 \pmod p$. Hence $n = 1$ or $n = p - 1$. So, we have that $\paren {p - 1}!$ consists of numbers multiplied together as follows: :in pairs whose product is congruent to $1 \pmod p$ :the numbers $1$ and $p - 1$. The product of all these numbers is therefore congruent to $1 \times 1 \times \cdots \times 1 \times p - 1 \pmod p$ by modulo multiplication. From Congruence Modulo Negative Number we therefore have that $\paren {p - 1}! \equiv -1 \pmod p$. {{Namedfor|John Wilson}} \end{proof}
23446
\section{Wilson's Theorem/Sufficient Condition} Tags: Wilson's Theorem \begin{theorem} Let $p$ be a (strictly) positive integer such that: :$\paren {p - 1}! \equiv -1 \pmod p$ Then $p$ is a prime number. \end{theorem} \begin{proof} Assume $p$ is composite, and $q$ is a prime such that $q \divides p$. Then both $p$ and $\paren {p - 1}!$ are divisible by $q$. If the congruence $\paren {p - 1}! \equiv -1 \pmod p$ were satisfied, we would have $\paren {p - 1}! \equiv -1 \pmod q$. However, this amounts to $0 \equiv -1 \pmod q$, a contradiction. Hence for $p$ composite, the congruence $\paren {p - 1}! \equiv -1 \pmod p$ cannot hold. {{qed}} \end{proof}
23447
\section{Word Metric is Metric} Tags: Group Theory, Word Metric, Examples of Metric Space, Examples of Metric Spaces \begin{theorem} Let $\struct {G, \circ}$ be a group. Let $S$ be a generating set for $G$ which is closed under inverses (that is, $x^{-1} \in S \iff x \in S$). Let $d_S$ be the associated word metric. Then $d_S$ is a metric on $G$. \end{theorem} \begin{proof} Let $g, h \in G$. It is given that $S$ is a generating set for $G$. It follows that there exist $s_1, \ldots, s_n \in S$ such that $g^{-1} \circ h = s_1 \circ \cdots \circ s_n$. Therefore $\map {d_S} {g, h} \le n$, establishing that $\R$ is a valid codomain for the mapping $d_S$ with domain $G \times G$. This is the form a mapping must have to be able to be a metric. Now checking the other defining properties for a metric in turn: \end{proof}
23448
\section{Woset is Isomorphic to Set of its Initial Segments} Tags: Order Isomorphisms, Well-Orderings, Orderings, Order Morphisms \begin{theorem} Let $\struct {S, \preceq}$ be a well-ordered set. Let: :$A = \set {a^\prec: a \in S}$ where $a^\prec$ is the strict lower closure of $S$ determined by $a$. Then: :$\struct {S, \preceq} \cong \struct {A, \subseteq}$ where $\cong$ denotes order isomorphism. \end{theorem} \begin{proof} Define $f: S \to A$ as: :$\forall a \in S: \map f a = a^\prec$ where $a^\prec$ is the initial segment determined by $a$. \end{proof}
23449
\section{Wosets are Isomorphic to Each Other or Initial Segments} Tags: Well-Orderings, Wosets are Isomorphic to Each Other or Initial Segments \begin{theorem} Let $\struct {S, \preceq_S}$ and $\struct {T, \preceq_T}$ be well-ordered sets. Then precisely one of the following hold: :$\struct {S, \preceq_S}$ is order isomorphic to $\struct {T, \preceq_T}$ or: :$\struct {S, \preceq_S}$ is order isomorphic to an initial segment in $\struct {T, \preceq_T}$ or: :$\struct {T, \preceq_T}$ is order isomorphic to an initial segment in $\struct {S, \preceq_S}$ \end{theorem} \begin{proof} We assume $S \ne \varnothing \ne T$; otherwise the theorem holds vacuously. Define: :$S' = S \cup \text{ initial segments in } S$ :$T' = T \cup \text{ initial segments in } T$ :$\mathcal F = \left\{ { f:S' \to T' \ \vert \ f \text{ is an order isomorphism} } \right\}$ We note that $\mathcal F$ is not empty, because it at least contains a trivial order isomorphism between singletons: :$f_0: \left\{ { \text{smallest element in } S} \right\} \to \left\{ { \text{smallest element in } T} \right\}$ Such smallest elements are guaranteed to exist by virtue of $S$ and $T$ being well-ordered. By the definition of initial segment, the initial segments of $S$ are subsets of $S$. For any initial segment $I_{\alpha}$ of $S$, such a segment has an upper bound by definition, namely, $\alpha$. Also, no initial segment of $S'$ is the entirety of $S$. Thus every initial segment of $S'$ has an upper bound, $S$ itself. Every chain in $S'$ has an upper bound, because it defines an initial segment. The previous reasoning also applies to $T'$. By Ordered Product of Tosets is Totally Ordered Set, $S' \times T'$ is itself a totally ordered set, with upper bound $S \times T$. Thus the hypotheses of Zorn's Lemma are satisfied for $\mathcal F$. Let $f_1$ be a maximal element of $\mathcal F$. Call its domain $A$ and its codomain $B$. Suppose $A$ is an initial segment $I_a$ in $S$ and $B$ is an initial segment $I_b$ in $T$. Then $f_1$ can be extended by defining $f_1\left({a}\right) = b$. This would contradict $f_1$ being maximal, so it cannot be the case that both $A$ and $B$ are initial segments. Then precisely one of the following hold: :$A = S$, with $S$ order isomorphic to an initial segment in $T$ or: :$B = T$, with $T$ order isomorphic to an initial segment in $S$ or: *both $A = S$ and $B = T$, with $S$ order isomorphic to $T$. {{qed}}{{improve|I suspect this proof can be adjusted to not use choice}} {{proofread}} {{AoC|Zorn's Lemma}} \end{proof}
23450
\section{X + y + z equals 1 implies xy + yz + zx less than Half} Tags: Inequalities, X + y + z equals 1 implies xy + yz + zx less than Half \begin{theorem} Let $x$, $y$ and $z$ be real numbers such that: :$x + y + z = 1$ Then: :$x y + y z + z x < \dfrac 1 2$ \end{theorem} \begin{proof} We have: {{begin-eqn}} {{eqn | l = 1 | r = \paren {x + y + z}^2 }} {{eqn | r = \paren {\paren {x + y} + z}^2 }} {{eqn | r = \paren {x + y}^2 + 2 z \paren {x + y} + z^2 | c = Square of Sum }} {{eqn | r = x^2 + 2 x y + y^2 + 2 x z + 2 y z + z^2 | c = Square of Sum }} {{eqn | r = 2 \paren {x y + y z + z x} + \paren {x^2 + y^2 + z^2} }} {{end-eqn}} So: :$2 \paren {x y + y z + z x} = 1 - \paren {x^2 + y^2 + z^2}$ We have: :$x^2 + y^2 + z^2 \ge 0$ for all real numbers $x$, $y$, $z$ with equality only if: :$x = y = z = 0$ This cannot be the case since $x + y + z = 1$, so we have: :$x^2 + y^2 + z^2 > 0$ so: :$2 \paren {x y + y z + z x} < 1$ giving: :$x y + y z + z x < \dfrac 1 2$ {{qed}} \end{proof}
23451
\section{X Choose n leq y Choose n + z Choose n-1 where n leq y leq x leq y+1 and n-1 leq z leq y} Tags: Binomial Coefficients \begin{theorem} Let $n \in \Z_{\ge 0}$ be a positive integer. Let $x, y \in \R$ be real numbers which satisfy: :$n \le y \le x \le y + 1$ Let $z$ be the unique real number $z$ such that: :$\dbinom x {n + 1} = \dbinom y {n + 1} + \dbinom z n$ where $n - 1 \le z \le y$. Its uniqueness is proved at Uniqueness of Real $z$ such that $\dbinom x {n + 1} = \dbinom y {n + 1} + \dbinom z n$. Then: :$\dbinom x n \le \dbinom y n + \dbinom z {n - 1}$ \end{theorem} \begin{proof} If $z \ge n$, then from Ordering of Binomial Coefficients: :$\dbinom z {n + 1} \le \dbinom y {n + 1}$ Otherwise $n - 1 \le z \le n$, and: :$\dbinom z {n + 1} \le 0 \le \dbinom y {n + 1}$ In either case: :$(1): \quad \dbinom z {n + 1} \le \dbinom y {n + 1}$ Therefore: {{begin-eqn}} {{eqn | l = \dbinom {z + 1} {n + 1} | r = \dbinom z {n + 1} + \dbinom z n | c = Pascal's Rule }} {{eqn | o = \le | r = \dbinom y {n + 1} + \dbinom z n | c = }} {{eqn | r = \dbinom x {n + 1} | c = by hypothesis }} {{end-eqn}} and so $x \ge z + 1$. Now we are to show that every term of the summation: :$\ds \binom x {n + 1} - \binom y {n + 1} = \sum_{k \mathop \ge 0} \dbinom {z - k} {n - k} t_k$ where: :$t_k = \dbinom {x - z - 1 + k} {k + 1} - \dbinom {y - z - 1 + k} {k + 1}$ is negative. Because $z \ge n - 1$, the binomial coefficient $\dbinom {z - k} {n - k}$ is non-negative. Because $x \ge z + 1$, the binomial coefficient $\dbinom {x - z - 1 + k} {k + 1}$ is also non-negative. Therefore: :$z \le y \le x$ implies that: :$\dbinom {y - z - 1 + k} {k + 1} \le \dbinom {x - z - 1 + k} {k + 1}$ When $x = y$ and $z = n - 1$ the result becomes: :$\dbinom x n \le \dbinom x n + \dbinom {n - 1} {n - 1}$ which reduces to: :$\dbinom x n \le \dbinom x n + 1$ which is true. Otherwise: {{begin-eqn}} {{eqn | l = \dbinom x n - \dbinom y n - \dbinom z {n - 1} | r = \sum_{k \mathop \ge 0} \dbinom {z - k} {n - 1 - k} \left({t_k - \delta_{k 0} }\right) | c = where $\delta_{k 0}$ is the Kronecker delta }} {{eqn | r = \sum_{k \mathop \ge 0} \dfrac {n - k} {z - n + 1} \dbinom {z - k} {n - k} \left({t_k - \delta_{k 0} }\right) | c = Factors of Binomial Coefficient }} {{end-eqn}} This is less than or equal to: {{begin-eqn}} {{eqn | l = \sum_{k \mathop \ge 0} \dfrac {n - 1} {z - n + 1} \dbinom {z - k} {n - k} \left({t_k - \delta_{k 0} }\right) | r = \dfrac {n - 1} {z - n + 1} \left({\dbinom x {n + 1} - \dbinom y {n + 1} - \dbinom z n}\right) | c = }} {{eqn | r = 0 | c = because $t_0 - 1 = x - y - 1 \le 0$ }} {{end-eqn}} Hence the result. {{qed}} \end{proof}
23452
\section{Yff's Conjecture} Tags: Triangles \begin{theorem} Let $\triangle ABC$ be a triangle. Let $\omega$ be the Brocard angle of $\triangle ABC$. Then: :$8 \omega^3 < ABC$ where $A, B, C$ are measured in radians. \end{theorem} \begin{proof} The Abi-Khuzam Inequality states that :$\sin A \cdot \sin B \cdot \sin C \le \paren {\dfrac {3 \sqrt 3} {2 \pi} }^3 A \cdot B \cdot C$ The maximum value of $A B C - 8 \omega^3$ occurs when two of the angles are equal. So taking $A = B$, and using $A + B + C = \pi$, the maximum occurs at the maximum of: :$\map f A = A^2 \paren {\pi - 2 A} - 8 \paren {\map \arccot {2 \cot A - \cot 2 A} }^3$ which occurs when: :$2 A \paren {\pi - 3 A} - \dfrac {48 \paren {\map \arccot {\frac 1 2 \paren {3 \cot A + \tan A} } }^2 \paren {1 + 2 \cos 2 A} } {5 + 4 \cos 2 A} = 0$ {{finish|Needs expanding and completing}} \end{proof}
23453
\section{Yoneda Embedding Theorem} Tags: Category Theory \begin{theorem} Let $C$ be a locally small category. Let $\mathbf {Set}$ be the category of sets. Let $\sqbrk {C^{\operatorname {op} }, \mathbf {Set} }$ be the contravariant functor category. Then the Yoneda embedding $h_- : C \to \sqbrk {C^{\operatorname {op} }, \mathbf {Set} }$ is a fully faithful embedding. \end{theorem} \begin{proof} {{proof wanted|use Yoneda Lemma for Contravariant Functors}} Category:Category Theory \end{proof}
23454
\section{Yoneda Lemma for Covariant Functors} Tags: Category theory, Category Theory \begin{theorem} Let $C$ be a locally small category. Let $\mathbf{Set}$ be the category of sets. \end{theorem} \begin{proof} {{proof wanted}} Category:Category Theory 333203 333199 2017-12-28T16:38:51Z Barto 3079 333203 wikitext text/x-wiki \end{proof}
23455
\section{Young's Inequality for Convolutions} Tags: Measure Theory, Inequalities, Young's Inequality for Convolutions \begin{theorem} Let $p, q, r \in \R_{\ge 1}$ satisfy: :$1 + \dfrac 1 r = \dfrac 1 p + \dfrac 1 q$ Let $\map {L^p} {\R^n}$, $\map {L^q} {\R^n}$, and $\map {L^r} {\R^n}$ be Lebesgue spaces with seminorms $\norm {\, \cdot \,}_p$, $\norm {\, \cdot \,}_q$, and $\norm {\, \cdot \,}_r$ respectively. Let $f \in \map {L^p} {\R^n}$ and $g \in \map {L^q} {\R^n}$. Then the convolution $f * g$ is in $\map {L^r} {\R^n}$ and the following inequality is satisfied: :$\norm {f * g}_r \le \norm f_p \cdot \norm g_q$ \end{theorem} \begin{proof} {{Proofread}} We begin by seeking to bound $\size {\map {\paren {f * g} } x}$: {{begin-eqn}} {{eqn | l = \map {\paren {f * g} } x | r = \int \map f {x - y} \map g y \rd y }} {{eqn | l = \size {\map {\paren {f * g} } x} | o = \le | r = \int \size {\map f {x - y} } \cdot \size {\map g y} \rd y }} {{eqn | r = \int \size {\map f {x - y} }^{1 + p/r - p/r} \cdot \size {\map g y}^{1 + q/r - q/r} \rd y }} {{eqn | r = \int \size {\map f {x - y} }^{p/r} \cdot \size {\map g y}^{q/r} \cdot \size {\map f {x - y} }^{1 - p / r} \cdot \size {\map g y}^{1-q/r} \rd y }} {{eqn | r = \int \paren {\size {\map f {x - y} }^p \cdot \size {\map g y}^q}^{1/r} \cdot \size {\map f {x - y} }^{\paren {r - p} / r} \cdot \size {\map g y}^{\paren {r - q} / r} \rd y }} {{eqn | o = \le | r = \underset {(1)} {\underbrace {\norm {\paren {\size {\map f {x - y} }^p \cdot \size {\map g y}^q}^{1/r} }_r} } \cdot \underset {(2)} {\underbrace {\norm {\size {\map f {x - y} }^{\paren {r - p} / r} }_{\textstyle \frac {p r} {r - p} } } } \cdot \underset {(3)} {\underbrace {\norm {\size {\map g y}^{\paren {r - q} / r} }_{\textstyle \frac {q r} {r - q} } } } }} {{end-eqn}} where the last inequality is via the Generalized Hölder Inequality applied to three functions. Note that the relation of conjugate exponents in the Generalized Hölder Inequality is satisfied: {{begin-eqn}} {{eqn | l = \frac 1 r + \frac {r - p} {p r} + \frac {r - q} {q r} | r = \frac 1 r + \frac 1 p - \frac 1 r + \frac 1 q - \frac 1 r }} {{eqn | r = \frac 1 p + \frac 1 q - \frac 1 r }} {{eqn | r = 1 | c = {{hypothesis}} on $p$, $q$, $r$ }} {{end-eqn}} We now analyze terms $(1)$, $(2)$ and $(3)$ in turn: {{begin-eqn}} {{eqn | n = 1 | l = \norm {\paren {\size {\map f {x - y} }^p \cdot \size {\map g y}^q}^{1/r} }_r | r = \paren {\int \paren {\size {\map f {x - y} }^p \cdot \size {\map g y}^q}^{\textstyle \frac 1 r \times r} \rd y}^{1/r} }} {{eqn | r = \paren {\int \size {\map f {x - y} }^p \cdot \size {\map g y}^q \rd y}^{1/r} }} {{eqn | n = 2 | l = \norm {\size {\map f {x - y} }^{\paren {r - p} /r } }_{\textstyle \frac {p r} {r - p} } | r = \paren {\int \size {\map f {x - y} }^{\textstyle \frac {r - p} r \times \frac {p r} {r - p} } \rd y}^{\textstyle\frac {r - p} {p r} } }} {{eqn | r = \paren {\int \size {\map f {x - y} }^p \rd y}^{\textstyle\frac 1 p \times \frac {r - p} r} }} {{eqn | r = {\norm f_p}^{\paren {r - p} / r} }} {{eqn | n = 3 | l = \norm {\size {\map g y}^{\paren {r - q} / r} }_{\textstyle\frac {q r} {r - q} } | r = \paren {\int \size {\map g y}^{\textstyle \frac {r - q} r \times \frac {q r} {r - q} } \rd y}^{\textstyle \frac {r - q} {q r} } }} {{eqn | r = \paren {\int \size {\map g y}^{\textstyle \frac {r - q} r \times \frac {q r} {r - q} } \rd y}^{\textstyle \frac {r - q} {q r} } }} {{eqn | r = \paren {\int \size {\map g y}^q \rd y}^{\textstyle \frac 1 q \times \frac {r - q} r} }} {{eqn | r = {\norm g_q}^{\paren {r - q} / r} }} {{end-eqn}} With these preliminary calculations out of the way, we turn to the main proof: {{begin-eqn}} {{eqn | l = {\norm {f * g}_r}^r | r = \int \size {\map {\paren {f * g} } x}^r \rd x }} {{eqn | o = \le | r = \int \paren {\int \paren {\size {\map f {x - y} }^p \cdot \size {\map g y}^q}^{1/r} \cdot \size {\map f {x - y} }^{\paren {r - p} / r} \cdot \size {\map g y}^{\paren {r - q} / r} \rd y}^r \rd x }} {{eqn | r = \int \paren {\paren {\int \paren {\size {\map f {x - y} }^p \cdot \size {\map g y}^q} \rd y}^{1/r} \cdot {\norm f_p}^{\paren {r - p} / r} \cdot {\norm g_q}^{\paren {r - q} / r} }^r \rd x | c = from $(1)$, $(2)$, $(3)$ }} {{eqn | r = \int \paren {\int \paren {\size {\map f {x - y} }^p \cdot \size {\map g y}^q} \rd y} \cdot {\norm f_p}^{r - p} \cdot {\norm g_q}^{r - q} \rd x }} {{eqn | r = {\norm f_p}^{r - p} \ {\norm g_q}^{r - q} \iint \size {\map g y}^q \size {\map f {x - y} }^p \rd y \rd x }} {{eqn | r = {\norm f_p}^{r - p} \ {\norm g_q}^{r - q} \int \size {\map g y}^q \paren {\int \size {\map f {x - y} }^p \rd x} \rd y | c = Fubini's Theorem }} {{eqn | r = {\norm f_p}^{r - p} \ {\norm g_q}^{r - q} \int \size {\map g y}^q \rd y \int \size {\map f x}^p \rd x }} {{eqn | r = {\norm f_p}^{r - p} \ {\norm g_q}^{r - q} \ {\norm g_q}^q \ {\norm f_p}^p }} {{eqn | r = {\norm f_p}^r \ {\norm g_q}^r }} {{eqn | ll= \leadsto | l = \norm {f * g}_r | o = \le | r = \norm f_p \norm g_q }} {{end-eqn}} {{qed}} {{Namedfor|William Henry Young|cat = Young}} Category:Measure Theory Category:Inequalities Category:Young's Inequality for Convolutions \end{proof}
23456
\section{Young's Inequality for Increasing Functions} Tags: Integral Calculus \begin{theorem} Let $a_0$ and $b_0$ be strictly positive real numbers. Let $f: \closedint 0 {a_0} \to \closedint 0 {b_0}$ be a strictly increasing bijection. Let $a$ and $b$ be real numbers such that $0 \le a \le a_0$ and $0 \le b \le b_0$. Then: :$\ds ab \le \int_0^a \map f u \rd u + \int_0^b \map {f^{-1} } v \rd v$ where $\ds \int$ denotes the definite integral. \end{theorem} \begin{proof} 200pxthumbrightThe blue colored region corresponds to $\ds \int_0^a \map f u \rd u$ and the red colored region to $\ds \int_0^b \map {f^{-1} } v \rd v$. {{ProofWanted}} {{Namedfor|William Henry Young|cat = Young}} \end{proof}
23457
\section{Young's Inequality for Products} Tags: Real Analysis, Inequalities, Analysis, Young's Inequality for Products, Named Theorems \begin{theorem} Let $p, q \in \R_{> 0}$ be strictly positive real numbers such that: :$\dfrac 1 p + \dfrac 1 q = 1$ Then, for any $a, b \in \R_{\ge 0}$: :$a b \le \dfrac {a^p} p + \dfrac {b^q} q$ Equality occurs {{iff}}: :$b = a^{p - 1}$ \end{theorem} \begin{proof} The result is obvious if $a=0$ or $b=0$, so assume WLOG that $a > 0$ and $b > 0$. Then: {{begin-eqn}} {{eqn | l = ab | r = \exp \left({\ln\left({ab}\right)}\right) | c = exponential and logarithm functions are inverses }} {{eqn | l = | r = \exp \left({ \ln a + \ln b }\right) | c = "sum property" of logarithms }} {{eqn | l = | r = \exp \left({ \frac 1 p p \ln a + \frac 1 q q \ln b }\right) | c = definitions of the multiplicative identity and inverse }} {{eqn | l = | r = \exp \left({ \frac 1 p \ln \left( {a^p} \right) + \frac 1 q \ln \left( {b^q} \right) }\right) | c = "power property" of logarithms }} {{eqn | l = | r = \frac 1 p \exp \left( {\ln \left( {a^p} \right)} \right) + \frac 1 q \exp \left( {\ln \left( {b^q} \right)} \right) | o = \le | c = exponential function is convex and the hypothesis that $\dfrac 1 p + \dfrac 1 q = 1$ }} {{eqn | l = | r = \frac{a^p} p + \frac{b^q} q | c = exponential and logarithm functions are inverses }} {{end-eqn}} {{qed}} {{namedfor|William Henry Young}} Category:Analysis 83328 83326 2012-03-12T22:08:16Z Prime.mover 59 83328 wikitext text/x-wiki \end{proof}
23458
\section{Z-Module Associated with Abelian Group is Unitary Z-Module} Tags: Abelian Groups, Z-Module Associated with Abelian Group, Examples of Unitary Modules, Group Theory, Unitary Modules, Modules \begin{theorem} Let $\struct {G, *}$ be an abelian group with identity $e$. Let $\struct {G, *, \circ}_\Z$ be the $Z$-module associated with $G$. Then $\struct {G, *, \circ}_\Z$ is a unitary $Z$-module. \end{theorem} \begin{proof} The notation $*^n x$ can be written as $x^n$. Let us verify that $\struct {G, *, \circ}_\Z$ is a unitary $\Z$-module by verifying the axioms in turn. \end{proof}
23459
\section{Z/(m)-Module Associated with Ring of Characteristic m} Tags: Unitary Modules, Group Theory \begin{theorem} Let $\left({R,+,*}\right)$ be a ring with unity whose zero is $0_R$ and whose unity is $1_R$. Let the characteristic of $R$ be $m$. Let $\left({\Z_m, +_m, \times_m}\right)$ be the ring of integers modulo $m$. Let $\circ$ be the mapping from $\Z_m \times R$ to $R$ defined as: :$\forall \left[\!\left[a\right]\!\right]_m \in \Z_m: \forall x \in R: \left[\!\left[a\right]\!\right]_m \circ x = a \cdot x$ where $\left[\!\left[a\right]\!\right]_m$ is the residue class of $a$ modulo $m$ and $a \cdot x$ is the $a$th power of $x$. Then $\left({R, +, \circ}\right)_{\Z_m}$ is a unitary $\Z_m$-module. \end{theorem} \begin{proof} Let us verify that the definition of $\circ$ is well-defined. Let $\left[\!\left[a\right]\!\right]_m=\left[\!\left[b\right]\!\right]_m$. Then we need to show that: :$\forall x \in R : \left[\!\left[a\right]\!\right]_m \circ x = \left[\!\left[b\right]\!\right]_m \circ x$ By the definition of congruence: :$\left[\!\left[a\right]\!\right]_m = \left[\!\left[b\right]\!\right]_m \iff \exists k \in \Z : a = b + k m$ Then: {{begin-eqn}} {{eqn | l = \left[\!\left[a\right]\!\right]_m \circ x | r = a \cdot x | c = Definition of $\circ$ }} {{eqn | r = \left(b+km\right) \cdot x }} {{eqn | r = b \cdot x + km \cdot x | c = Powers of Group Elements: Sum of Indices }} {{eqn | r = b \cdot x + k \cdot \left( m \cdot x \right) | c = Powers of Group Elements: Product of Indices }} {{eqn | r = b \cdot x + k \cdot 0_R | c = Characteristic times Ring Element is Ring Zero }} {{eqn | r = b \cdot x + 0_R | c = Power of Identity is Identity }} {{eqn | r = b \cdot x }} {{eqn | r = \left[\!\left[b\right]\!\right]_m \circ x | c = Definition of $\circ$ }} {{end-eqn}} Thus, the definition of $\circ$ is well-defined. {{qed|lemma}} Let us verify that $\left({R, +, \circ}\right)_{\Z_m}$ is a unitary $\Z_m$-module by verifying the axioms in turn. \end{proof}
23460
\section{Zariski's Lemma} Tags: Commutative Algebra, Field Extensions \begin{theorem} Let $L / k$ be a field extension. Let $L$ be finitely generated as an algebra over $k$. Then $L / k$ is a finite field extension. \end{theorem} \begin{proof} {{MissingLinks}} By Noether Normalization Lemma, we find a finite and injective morphism: :$\alpha: k \sqbrk {x_1, \dotsc, x_n} \to L$ If we can prove that $n = 0$, the proof is complete. {{AimForCont}} $n > 0$. Then: :$x_1 \in k \sqbrk {x_1, \dotsc, x_n}$ and: :$\map \alpha {x_1} \ne 0$ We have that $\map \alpha {x_1}^{-1}$ is integral over $k \sqbrk {x_1, \dotsc, x_n}$. Thus there exists a $m \in \N$ and $a_0, \dotsc, a_{m - 1} \in k \sqbrk {x_1, \dotsc, x_n}$ such that: :$\ds \map \alpha {x_1}^{-m} + \sum_{i \mathop = 0}^{m - 1} \map \alpha {a_i} \map \alpha {x_1}^{-i} = 0$ Multiplying by $\map \alpha {x_1}^m$: {{begin-eqn}} {{eqn | l = 0 | r = 1 + \sum_{i \mathop = 0}^{m - 1} \map \alpha {a_i} \map \alpha {x_1}^{m - i} }} {{eqn | r = \map \alpha {1 + x_1 \paren {\sum_{i \mathop = 0}^{m - 1} a_i x_1^{m - i - 1} } } }} {{end-eqn}} and thus, since $\alpha$ is injective, we find that: :$\ds 1 = x_1 \paren {-\sum_{i \mathop = 0}^{m - 1} a_i x_1^{m - i - 1} }$ which means that $x_1$ is invertible. This contradiction shows that $n = 0$. {{qed}} {{Namedfor|Oscar Zariski|cat = Zariski}} Category:Commutative Algebra Category:Field Extensions \end{proof}
23461
\section{Zeckendorf's Theorem} Tags: Zeckendorf Representation, Named Theorems, Fibonacci Numbers \begin{theorem} Every positive integer has a unique Zeckendorf representation. That is: Let $n$ be a positive integer. Then there exists a unique increasing sequence of integers $\left\langle{c_i}\right\rangle$ such that: :$\forall i \in \N: c_i \ge 2$ :$c_{i + 1} > c_i + 1$ :$\ds n = \sum_{i \mathop = 0}^k F_{c_i}$ where $F_m$ is the $m$th Fibonacci number. For any given $n$, such a $\left\langle{c_i}\right\rangle$ is unique. \end{theorem} \begin{proof} First note that every Fibonacci number $F_n$ is itself a '''Zeckendorf representation''' of itself, where the sequence $\left\langle{c_i}\right\rangle$ contains $1$ term. \end{proof}
23462
\section{Zeckendorf Representation of Integer shifted Left} Tags: Zeckendorf Representations, Zeckendorf Representation \begin{theorem} Let $\map f x$ be the real function defined as: :$\forall x \in \R: \map f x = \floor {x + \phi^{-1} }$ where: :$\floor {\, \cdot \,}$ denotes the floor function :$\phi$ denotes the golden mean. Let $n \in \Z_{\ge 0}$ be a positive integer. Let $n$ be expressed in Zeckendorf representation: :$n = F_{k_1} + F_{k_2} + \cdots + F_{k_r}$ with the appropriate restrictions on $k_1, k_2, \ldots, k_r$. Then: :$F_{k_1 + 1} + F_{k_2 + 1} + \cdots + F_{k_r + 1} = \map f {\phi n}$ \end{theorem} \begin{proof} We have: {{begin-eqn}} {{eqn | l = F_{k_j + 1} \hat \phi + F_{k_j} | r = \hat \phi^{k_j + 1} | c = Fibonacci Number by One Minus Golden Mean plus Fibonacci Number of Index One Less }} {{eqn | ll= \leadsto | l = F_{k_j + 1} | r = \hat \phi^{k_j} - \frac {F_{k_j} } {\hat \phi} | c = }} {{eqn | r = \hat \phi^{k_j} + \phi F_{k_j} | c = Reciprocal Form of One Minus Golden Mean }} {{end-eqn}} Hence: {{begin-eqn}} {{eqn | l = F_{k_1 + 1} + F_{k_2 + 1} + \cdots + F_{k_r + 1} | r = \phi \paren {F_{k_1} + F_{k_2} + \cdots + F_{k_r} } + \paren {\hat \phi^{k_1} + \hat \phi^{k_2} + \cdots + \hat \phi^{k_r} } | c = }} {{eqn | r = \phi n + \paren {\hat \phi^{k_1} + \hat \phi^{k_2} + \cdots + \hat \phi^{k_r} } | c = }} {{end-eqn}} We have that: :$\hat \phi^3 + \hat \phi^5 + \hat \phi^7 + \cdots \le \hat \phi^{k_1} + \hat \phi^{k_2} + \cdots + \hat \phi^{k_r} \le \hat \phi^2 + \hat \phi^4 + \hat \phi^6 + \cdots$ Then: {{begin-eqn}} {{eqn | l = \hat \phi^3 + \hat \phi^5 + \hat \phi^7 + \cdots | r = \hat \phi^3 \paren {1 + \hat \phi^2 + \hat \phi^4 + \cdots} | c = }} {{eqn | r = \dfrac {\hat \phi^3} {1 - \hat \phi^2} | c = Sum of Infinite Geometric Sequence }} {{eqn | r = \dfrac {\phi^2 \hat \phi^3} {\phi^2 - \phi^2 \hat \phi^2} | c = }} {{eqn | r = \dfrac {\paren {-1}^2 \hat \phi} {\phi^2 - \paren {-1}^2} | c = Golden Mean by One Minus Golden Mean equals Minus 1 }} {{eqn | r = \dfrac {\hat \phi} {\phi^2 - 1} | c = }} {{eqn | r = \dfrac {1 - \phi} {\phi^2 - 1} | c = {{Defof|One Minus Golden Mean}} }} {{eqn | r = \dfrac {1 - \phi} {1 + \phi - 1} | c = Square of Golden Mean equals One plus Golden Mean }} {{eqn | r = \phi^{-1} - 1 | c = }} {{end-eqn}} Then: {{begin-eqn}} {{eqn | l = \hat \phi^2 + \hat \phi^4 + \hat \phi^6 + \cdots | r = \frac 1 {\hat \phi} \paren {\hat \phi^3 + \hat \phi^5 + \hat \phi^7 + \cdots} | c = }} {{eqn | r = \frac 1 {\hat \phi} \paren {\phi^{-1} - 1} | c = from above }} {{eqn | r = -\phi \paren {\frac 1 \phi - 1} | c = Reciprocal Form of One Minus Golden Mean }} {{eqn | r = -1 + \phi | c = }} {{eqn | r = \frac 1 {\phi} | c = {{Defof|Golden Mean|index = 3}} }} {{eqn | r = \phi^{-1} | c = }} {{end-eqn}} Thus: :$\phi^{-1} - 1 \le \hat \phi^{k_1} + \hat \phi^{k_2} + \cdots + \hat \phi^{k_r} \le \phi^{-1}$ and the result follows. {{qed}} \end{proof}
23463
\section{Zeckendorf Representation of Integer shifted Right} Tags: Zeckendorf Representation \begin{theorem} Let $f \left({x}\right)$ be the real function defined as: :$\forall x \in \R: f \left({x}\right) = \left\lfloor{x + \phi^{-1} }\right\rfloor$ where: :$\left\lfloor{\, \cdot \,}\right\rfloor$ denotes the floor function :$\phi$ denotes the golden mean. Let $n \in \Z_{\ge 0}$ be a positive integer. Let $n$ be expressed in Zeckendorf representation: :$n = F_{k_1} + F_{k_2} + \cdots + F_{k_r}$ with the appropriate restrictions on $k_1, k_2, \ldots, k_r$. Then: :$F_{k_1 - 1} + F_{k_2 - 1} + \cdots + F_{k_r - 1} = f \left({\phi^{-1} n}\right)$ \end{theorem} \begin{proof} Follows directly from Zeckendorf Representation of Integer shifted Left, substituting $F_{k_j - 1}$ for $F_{k_j}$ throughout. {{qed}} \end{proof}
23464
\section{Zenith Distance is Complement of Celestial Altitude} Tags: Spherical Astronomy \begin{theorem} Let $X$ be the position of a star (or other celestial body) on the celestial sphere. The zenith distance $z$ of $X$ is the complement of the altitude $a$ of $X$: :$z = 90 \degrees - a$ \end{theorem} \begin{proof} The vertical circle through $X$ is defined as the great circle that passes through $Z$. By definition, the angle of the arc from $Z$ to the horizon is a right angle. Hence $z + a = 90 \degrees$. The result follows. {{qed}} \end{proof}
23465
\section{Zenith Distance of North Celestial Pole equals Colatitude of Observer} Tags: Spherical Astronomy \begin{theorem} Let $O$ be an observer of the celestial sphere. Let $P$ be the position of the north celestial pole with respect to $O$. Let $z$ denote the zenith distance of $P$. Let $\psi$ denote the (terrestrial) colatitude of $O$. Then: :$z = \psi$ \end{theorem} \begin{proof} Let $Z$ denote the zenith. The zenith distance $z$ of $P$ is by definition the length of the arc $PZ$ of the prime vertical. This in turn is defined as the angle $\angle POZ$ subtended by $PZ$ at $O$. This is equivalent to the angle between the radius of Earth through $O$ and Earth's axis. This is by definition the (terrestrial) colatitude of $O$. Hence the result. {{qed}} \end{proof}
23466
\section{Zermelo's Theorem (Set Theory)} Tags: Set Theory, Named Theorems, Axiom of Choice \begin{theorem} Every set of cardinals is well-ordered with respect to $\le$. \end{theorem} \begin{proof} Let $S_1$ and $S_2$ be sets which are not empty. Suppose there exists an injection $f: S_1 \to S_2$ and another injection $g: S_2 \to S_1$. Then by the Cantor-Bernstein-Schröder Theorem there exists a bijection between $S_1$ and $S_2$ and by definition $S_1$ is equivalent to $S_2$. Let $\AA$ be the set of invertible mappings $\phi: A \to B$ where $A \subseteq S_1$ and $B \subseteq S_2$. Since $S_1$ and $S_2$ are not empty, $\exists s_1 \in S_1$ and $\exists s_2 \in S_2$. Thus we can construct the mapping $\alpha: \set {s_1} \to \set {s_2}$ such that $\map \alpha {s_1} = s_2$. This is trivially an invertible mapping, so $\AA$ is not empty. We can impose an ordering $\le$ on $\AA$ by letting $\phi_1 \le \phi_2 \iff \phi_1 \subseteq \phi_2$, that is, if $\phi_2$ is an extension of $\phi_1$. Let $\CC$ be a chain in $\struct {\AA, \le}$. Then $\ds \bigcup \set {\phi \in \CC}$ is an upper bound of every $\phi \in \CC$, and it lies in $\AA$. The conditions of Zorn's Lemma are satisfied, so we can find a maximal element $M$ in $\AA$. Let: :$M_1 = \set {s_1: \tuple {s_1, s_2} \in M}$ :$M_2 = \set {s_2: \tuple {s_1, s_2} \in M}$ We have that $M_1 \subsetneq S_1$ and $M_2 \subsetneq S_2$ both together contradict the fact that $M$ is maximal element. Thus either $M_1 = S_1$ or $M_2 = S_2$, and possibly both. Thus either: :$M$ is an injection of $S_1$ into $S_2$ or: :$M^{-1}$ is an injection of $S_2$ into $S_1$. Thus either $S_1 \le S_2$ or $S_2 \le S_1$, and the result follows. {{qed}} {{AoC||Zorn's Lemma}} \end{proof}
23467
\section{Zero (Category) is Initial Object} Tags: Category of Categories \begin{theorem} Let $\mathbf{Cat}$ be the category of categories. Let $\mathbf 0$ be the zero category. Then $\mathbf 0$ is an initial object of $\mathbf{Cat}$. \end{theorem} \begin{proof} Let $\mathbf C$ be an object of $\mathbf{Cat}$, i.e. a small category. By Empty Mapping is Unique, there are unique mappings: :$F_0: \mathbf 0_0 = \O \to \mathbf C_0$ :$F_1: \mathbf 0_1 = \O \to \mathbf C_1$ making $F: \mathbf 0 \to \mathbf C$ a functor by vacuous truth. That $F_0$ and $F_1$ are actually mappings follows from $\mathbf C$ being a small category. Hence the result, by definition of initial object. {{qed}} \end{proof}
23468
\section{Zero Choose n} Tags: Binomial Coefficients, Examples of Binomial Coefficients \begin{theorem} :$\dbinom 0 n = \delta_{0 n}$ where: :$\dbinom 0 n$ denotes a binomial coefficient :$\delta_{0 n}$ denotes the Kronecker delta. \end{theorem} \begin{proof} By definition of binomial coefficient: :$\dbinom m n = \begin{cases} \dfrac {m!} {n! \paren {m - n}!} & : 0 \le n \le m \\ & \\ 0 & : \text { otherwise } \end{cases}$ Thus when $n > 0$: :$\dbinom 0 n = 0$ and when $n = 0$: :$\dbinom 0 0 = \dfrac {0!} {0! \paren {0 - 0}!} = 1$ by definition of factorial. Hence the result by definition of the Kronecker delta. {{qed}} \end{proof}
23469
\section{Zero Complement is Not Empty} Tags: Naturally Ordered Semigroup \begin{theorem} Let $\struct {S, \circ, \preceq}$ be a naturally ordered semigroup. Let $S^*$ be the zero complement of $S$. Then $S^*$ is not empty. \end{theorem} \begin{proof} From {{NOSAxiom|4}}, we have: :$\exists m, n \in S: m \ne n$ That is, there are at least two distinct elements in $S$. Therefore, there must be at least one element in $S^* = S \setminus \set 0$. So: :$S^* = S \setminus \set 0 \ne \O$ {{qed}} \end{proof}
23470
\section{Zero Definite Integral of Nowhere Negative Function implies Zero Function} Tags: Integral Calculus, Definite Integrals \begin{theorem} Let $\closedint a b \subseteq \R$ be a closed real interval. Let $h: \closedint a b \to \R$ be a continuous real function such that: :$\forall x \in \closedint a b: \map h x \ge 0$ Let: :$\ds \int_a^b \map h x \rd x = 0$ Then: :$\forall x \in \closedint a b: \map h x = 0$ \end{theorem} \begin{proof} {{AimForCont}} that: :$\exists c \in \closedint a b: \map h c > 0$ As $h$ is continuous, there exists some closed real interval $\closedint r s \subseteq \closedint a b$ where $r < s$ such that: :$\exists \epsilon \in \R_{>0}: \forall x \in \closedint r s: \map h x > \dfrac {\map h c} 2$ From Sign of Function Matches Sign of Definite Integral: :$\ds \int_r^s \map h x \rd x > 0$ This makes a strictly positive contribution to the integral. $h$ is still continuous on $\closedint a r$ and $\closedint s b$. So, again from Sign of Function Matches Sign of Definite Integral: :$\forall x \in \closedint a r: \map f x \ge 0 \implies \ds \int_a^r \map f x \rd x \ge 0$ :$\forall x \in \closedint s b: \map f x \ge 0 \implies \ds \int_s^b \map f x \rd x \ge 0$ Thus as: :$\ds \int_a^b \map h x \rd x = \int_a^r \map f x \rd x + \int_r^s \map f x \rd x + \int_s^b \map f x \rd x$ it follows that: :$\ds \int_a^b \map h x \rd x > 0$ But by hypothesis: :$\ds \int_a^b \map h x \rd x = 0$ Thus by contradiction, there can be no $c \in \closedint a b$ such that $\map h c > 0$. Hence the result. {{qed}} \end{proof}
23471
\section{Zero Derivative implies Constant Complex Function} Tags: Complex Analysis \begin{theorem} Let $D \subseteq \C$ be a connected domain of $\C$. Let $f: D \to \C$ be a complex-differentiable function. For all $z \in D$, let $\map {f'} z = 0$. Then $f$ is constant on $D$. \end{theorem} \begin{proof} Let $u, v: \set {\tuple {x, y} \in \R^2: x + i y = z \in D} \to \R$ be the two real-valued functions defined in the Cauchy-Riemann Equations: :$\map u {x, y} = \map \Re {\map f z}$ :$\map v {x, y} = \map \Im {\map f z}$ By the Cauchy-Riemann Equations: :$f' = \dfrac {\partial f} {\partial x} = \dfrac {\partial u} {\partial x} + i \dfrac {\partial v} {\partial x}$ :$f' = -i \dfrac {\partial f} {\partial y} = \dfrac {\partial v} {\partial y} - i \dfrac {\partial u} {\partial y}$ As $f' = 0$ by assumption, this implies: :$0 = \dfrac {\partial u} {\partial x} = \dfrac {\partial u} {\partial y} = \dfrac {\partial v} {\partial x} = \dfrac {\partial v} {\partial y}$ Then Zero Derivative implies Constant Function shows that $\map u {x + t, y} = \map u {x, y}$ for all $t \in \R$. Similar results apply for the other three partial derivatives. Let $z, w \in D$. From Connected Domain is Connected by Staircase Contours, it follows that there exists a staircase contour $C$ in $D$ with endpoints $z$ and $w$. The contour $C$ is a concatenation of directed smooth curves that can be parameterized as line segments on one of these two forms: :$(1): \quad \map \gamma t = z_0 + t r$ :$(2): \quad \map \gamma t = z_0 + i t r$ for some $z_0 \in D$ and $r \in \R$ for all $t \in \closedint 0 1$. If $z_1 \in D$ lies on the same line segment as $z_0$, it follows that for parameterizations of type $(1)$: {{begin-eqn}} {{eqn | l = \map f {z_1} | r = \map u {z_0 + t r} + \map v {z_0 + t r} | c = for some $t \in \closedint 0 1$ }} {{eqn | r = \map u {z_0} + \map v {z_0} | c = Zero Derivative implies Constant Function }} {{eqn | r = \map f {z_0} }} {{end-eqn}} Similarly, for parameterizations of type $(2)$: {{begin-eqn}} {{eqn | l = \map f {z_1} | r = \map u {z_0 + i t r} + \map v {z_0 + i t r} | c = for some $t \in \closedint 0 1$ }} {{eqn | r = \map f {z_0} | c = Zero Derivative implies Constant Function }} {{end-eqn}} As the image of $C$ is connected by these line segments, it follows that for all $z_0$ and $z_1$ in the image of $C$: :$\map f {z_0} = \map f {z_1}$ In particular: :$\map f z = \map f w$ so $f$ is constant on $D$. {{qed}} \end{proof}
23472
\section{Zero Derivative implies Constant Function} Tags: Differential Calculus, Constant Mappings \begin{theorem} Let $f$ be a real function which is continuous on the closed interval $\closedint a b$ and differentiable on the open interval $\openint a b$. Suppose that: :$\forall x \in \openint a b: \map {f'} x = 0$ Then $f$ is constant on $\closedint a b$. \end{theorem} \begin{proof} When $x = a$ then $\map f x = \map f a$ by definition of mapping. Otherwise, let $x \in \hointl a b$. We have that: :$f$ is continuous on the closed interval $\closedint a b$ :$f$ is differentiable on the open interval $\openint a b$ Hence it satisfies the conditions of the Mean Value Theorem on $\closedint a b$. Hence: :$\exists \xi \in \openint a x: \map {f'} \xi = \dfrac {\map f x - \map f a} {x - a}$ But by our supposition: :$\forall x \in \openint a b: \map {f'} x = 0$ which means: :$\forall x \in \openint a b: \map f x - \map f a = 0$ and hence: :$\forall x \in \openint a b: \map f x = \map f a$ {{qed}} \end{proof}