id
stringlengths
1
260
contents
stringlengths
1
234k
22573
\section{Tangent of Half Angle for Spherical Triangles} Tags: Spherical Trigonometry, Half Angle Formulas for Spherical Triangles \begin{theorem} Let $\triangle ABC$ be a spherical triangle on the surface of a sphere whose center is $O$. Let the sides $a, b, c$ of $\triangle ABC$ be measured by the angles subtended at $O$, where $a, b, c$ are opposite $A, B, C$ respectively. Then: :$\tan \dfrac A 2 = \sqrt {\dfrac {\map \sin {s - b} \, \map \sin {s - c} } {\sin s \, \map \sin {s - a} } }$ where $s = \dfrac {a + b + c} 2$. \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \tan \dfrac A 2 | r = \dfrac {\sqrt {\dfrac {\sin \paren {s - b} \sin \paren {s - c} } {\sin b \sin c} } } {\sqrt {\dfrac {\sin s \, \map \sin {s - a} } {\sin b \sin c} } } | c = Sine of Half Angle for Spherical Triangles, Cosine of Half Angle for Spherical Triangles }} {{eqn | r = \dfrac {\sqrt {\sin \paren {s - b} \sin \paren {s - c} } } {\sqrt {\sin s \, \map \sin {s - a} } } | c = simplification }} {{end-eqn}} The result follows. {{qed}} \end{proof}
22574
\section{Tangent of Half Angle plus Quarter Pi} Tags: Trigonometric Identities, Tangent Function, Secant Function \begin{theorem} :$\map \tan {\dfrac x 2 + \dfrac \pi 4} = \tan x + \sec x$ \end{theorem} \begin{proof} Firstly, we have: {{begin-eqn}} {{eqn | n = 1 | l = \tan x | r = \frac {2 \tan \frac x 2} {1 - \tan ^2 \frac x 2} | c = Double Angle Formula for Tangent }} {{end-eqn}} Then: {{begin-eqn}} {{eqn | l = \map \tan {\frac x 2 + \frac \pi 4} | r = \frac {\tan \frac x 2 + \tan \frac \pi 4} {1 - \tan \frac x 2 \tan \frac \pi 4} | c = Tangent of Sum }} {{eqn | r = \frac {\tan \frac x 2 + 1} {1 - \tan \frac x 2} | c = Tangent of $45 \degrees$ }} {{eqn | r = \frac {\paren {\tan \frac x 2 + 1} \paren {\tan \frac x 2 + 1} } {\paren {1 - \tan \frac x 2} \paren {\tan \frac x 2 + 1} } }} {{eqn | r = \frac {\tan^2 \frac x 2 + 2 \tan \frac x 2 + 1} {1 - \tan^2 \frac x 2} | c = Difference of Two Squares, Square of Sum }} {{eqn | r = \frac {2 \tan \frac x 2} {1 - \tan^2 \frac x 2} + \frac {\tan^2 \frac x 2 + 1} {1 - \tan^2 \frac x 2} }} {{eqn | r = \tan x + \frac {\tan^2 \frac x 2 + 1} {1 - \tan^2 \frac x 2} | c = Double Angle Formula for Tangent: see $(1)$ above }} {{eqn | r = \tan x + \frac {\sin^2 \frac x 2 + \cos^2 \frac x 2} {\cos^2 \frac x 2 - \sin^2 \frac x 2} | c = multiplying Denominator and Numerator by $\cos^2 \frac x 2$ }} {{eqn | r = \tan x + \frac {\sin^2 \frac x 2 + \cos^2 \frac x 2} {\cos 2 \frac x 2} | c = Double Angle Formula for Cosine }} {{eqn | r = \tan x + \frac 1 {\cos x} | c = Sum of Squares of Sine and Cosine }} {{eqn | r = \tan x + \sec x | c = Secant is Reciprocal of Cosine }} {{end-eqn}} {{qed}} Category:Trigonometric Identities Category:Tangent Function Category:Secant Function \end{proof}
22575
\section{Tangent of Half Side for Spherical Triangles} Tags: Half Side Formulas for Spherical Triangles \begin{theorem} Let $\triangle ABC$ be a spherical triangle on the surface of a sphere whose center is $O$. Let the sides $a, b, c$ of $\triangle ABC$ be measured by the angles subtended at $O$, where $a, b, c$ are opposite $A, B, C$ respectively. Then: :$\tan \dfrac a 2 = \sqrt {\dfrac {-\cos S \, \map \cos {S - A} } {\map \cos {S - B} \, \map \cos {S - C} } }$ where $S = \dfrac {a + b + c} 2$. \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \tan \dfrac a 2 | r = \dfrac {\sqrt {\dfrac {-\cos S \, \map \cos {S - A} } {\sin B \sin C} } } {\sqrt {\dfrac {\map \cos {S - B} \, \map \cos {S - C} } {\sin B \sin C} } } | c = Sine of Half Angle for Spherical Triangles, Cosine of Half Angle for Spherical Triangles }} {{eqn | r = \sqrt {\dfrac {-\cos S \, \map \cos {S - A} } {\map \cos {S - B} \, \map \cos {S - C} } } | c = simplification }} {{end-eqn}} Hence the result. {{qed}} \end{proof}
22576
\section{Tangent of Right Angle} Tags: Tangent Function \begin{theorem} :$\tan 90^\circ = \tan \dfrac \pi 2$ is undefined where $\tan$ denotes tangent. \end{theorem} \begin{proof} From Tangent is Sine divided by Cosine, $\tan \theta = \dfrac {\sin \theta} {\cos \theta}$. When $\cos \theta = 0$, $\dfrac {\sin \theta} {\cos \theta}$ can be defined only if $\sin \theta = 0$. But there are no such $\theta$ such that both $\cos \theta = 0$ and $\sin \theta = 0$. When $\theta = \dfrac \pi 2$, $\cos \theta = 0$. Thus $\tan \theta$ is undefined at this value. {{qed}} \end{proof}
22577
\section{Tangent of Three Right Angles} Tags: Tangent Function \begin{theorem} :$\tan 270^\circ = \tan \dfrac {3 \pi} 2$ is undefined where $\tan$ denotes tangent. \end{theorem} \begin{proof} We have: {{begin-eqn}} {{eqn | l = \tan 270^\circ | r = \tan \left({360^\circ - 90^\circ}\right) | c = }} {{eqn | r = -\tan 90^\circ | c = Tangent of Conjugate Angle }} {{end-eqn}} But from Tangent of Right Angle, $\tan 90^\circ$ is undefined. Hence so is $\tan 270^\circ$. {{qed}} \end{proof}
22578
\section{Tangent over Secant Plus One} Tags: Trigonometric Identities \begin{theorem} :$\dfrac {\tan x} {\sec x + 1} = \dfrac {\sec x - 1} {\tan x}$ \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \frac {\tan x} {\sec x + 1} | r = \frac {\sin x} {\cos x \paren {\sec x + 1} } | c = Tangent is Sine divided by Cosine }} {{eqn | r = \frac {\sin x} {\cos x \paren {\frac 1 {\cos x} + 1} } | c = Secant is Reciprocal of Cosine }} {{eqn | r = \frac {\sin x} {1 + \cos x} | c = simplifying }} {{eqn | r = \frac {\sin^2 x} {\sin x \paren {1 + \cos x} } | c = Multiply by $1 = \dfrac {\sin x} {\sin x}$ }} {{eqn | r = \frac {1 - \cos^2 x} {\sin x \paren {1 + \cos x} } | c = Sum of Squares of Sine and Cosine }} {{eqn | r = \frac {\paren {1 + \cos x} \paren {1 - \cos x} } {\sin x \paren {1 + \cos x} } | c = Difference of Two Squares }} {{eqn | r = \frac {1 - \cos x} {\sin x} }} {{eqn | r = \frac {\cos x \paren {\frac 1 {\cos x} - 1} } {\sin x} }} {{eqn | r = \frac {\cos x \paren {\sec x - 1} } {\sin x} | c = Secant is Reciprocal of Cosine }} {{eqn | r = \frac {\sec x - 1} {\tan x} | c = Tangent is Sine divided by Cosine }} {{end-eqn}} {{qed}} Category:Trigonometric Identities \end{proof}
22579
\section{Tangent times Tangent plus Cotangent} Tags: Trigonometric Identities, Tangent times Tangent plus Cotangent \begin{theorem} :$\tan x \paren {\tan x + \cot x} = \sec^2 x$ where $\tan$, $\cot$ and $\sec$ denote tangent, cotangent and secant respectively. \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l=\tan x \left({\tan x + \cot x}\right) | r=\tan x \sec x \csc x | c=Sum of Tangent and Cotangent }} {{eqn | r=\frac {\sin x} {\cos^2 x \sin x} | c=by definition of tangent, secant and cosecant }} {{eqn | r=\frac 1 {\cos^2x} | c= }} {{eqn | r=\sec^2x | c=by definition of secant }} {{end-eqn}} {{qed}} Or directly: {{begin-eqn}} {{eqn | l=\tan x \left({\tan x + \cot x}\right) | r=\frac {\sin x} {\cos x} \left({\frac {\sin x} {\cos x} + \frac {\cos x} {\sin x} }\right) | c=by definition of tangent and cotangent }} {{eqn | r=\frac {\sin x} {\cos x} \left({\frac {\sin^2 x + \cos^2 x} {\cos x \sin x} }\right) | c= }} {{eqn | r=\frac {\sin x} {\cos x} \left({\frac 1 {\cos x \sin x} }\right) | c=Sum of Squares of Sine and Cosine }} {{eqn | r=\frac 1 {\cos^2x} | c= }} {{eqn | r=\sec^2x | c=by definition of secant }} {{end-eqn}} {{qed}} Category:Trigonometric Identities 69372 45019 2011-11-09T19:38:42Z Lord Farin 560 69372 wikitext text/x-wiki \end{proof}
22580
\section{Tangent to Astroid between Coordinate Axes has Constant Length} Tags: Hypocycloids \begin{theorem} Let $C_1$ be a circle of radius $b$ roll without slipping around the inside of a circle $C_2$ of radius $a = 4 b$. Let $C_2$ be embedded in a cartesian plane with its center $O$ located at the origin. Let $P$ be a point on the circumference of $C_1$. Let $C_1$ be initially positioned so that $P$ is its point of tangency to $C_2$, located at point $A = \tuple {a, 0}$ on the $x$-axis. Let $H$ be the astroid formed by the locus of $P$. The segment of the tangent to $H$ between the $x$-axis and the $y$-axis is constant, immaterial of the point of tangency. \end{theorem} \begin{proof} :400px From Equation of Astroid, $H$ can be expressed as: :$\begin{cases} x & = a \cos^3 \theta \\ y & = a \sin^3 \theta \end{cases}$ Thus the slope of the tangent to $H$ at $\tuple {x, y}$ is: {{begin-eqn}} {{eqn | l = \frac {\d y} {\d x} | r = \frac {3 a \sin^2 \theta \cos \theta \rd \theta} {-3 a \cos^2 \theta \sin \theta \rd \theta} | c = }} {{eqn | r = -\tan \theta | c = }} {{end-eqn}} Thus the equation of the tangent to $H$ is given by: :$y - a \sin^3 \theta = -\tan \theta \paren {x - a \cos^3 \theta}$ {{explain|Find, or post up, the equation of a line of given tangent passing through point $\tuple {x, y}$ as this is what is needed here}} The $x$-intercept is found by setting $y = 0$ and solving for $x$: {{begin-eqn}} {{eqn | l = x | r = a \cos^3 \theta + a \sin^2 \theta \cos \theta | c = }} {{eqn | r = a \cos \theta \paren {\cos^2 \theta + \sin^2 \theta} | c = }} {{eqn | r = a \cos \theta | c = Sum of Squares of Sine and Cosine }} {{end-eqn}} Similarly, the $y$-intercept is found by setting $x = 0$ and solving for $y$, which gives: :$y = a \sin \theta$ The length of the part of the tangent to $H$ between the $x$-axis and the $y$-axis is given by: {{begin-eqn}} {{eqn | l = \sqrt {a^2 \cos^2 \theta + a^2 \sin^2 \theta} | r = a \sqrt {\cos^2 \theta + \sin^2 \theta} | c = }} {{eqn | r = a | c = Sum of Squares of Sine and Cosine }} {{end-eqn}} which is constant. {{qed}} \end{proof}
22581
\section{Tangent to Cycloid} Tags: Cycloids \begin{theorem} Let $C$ be a cycloid generated by the equations: :$x = a \paren {\theta - \sin \theta}$ :$y = a \paren {1 - \cos \theta}$ Then the tangent to $C$ at a point $\tuple {x, y}$ on $C$ is given by the equation: :$y - a \paren {1 - \cos \theta} = \dfrac {\sin \theta} {1 - \cos \theta} \paren {x - a \theta + a \sin \theta}$ \end{theorem} \begin{proof} From Slope of Tangent to Cycloid, the slope of the tangent to $C$ at the point $\tuple {x, y}$ is given by: :$\dfrac {\d y} {\d x} = \cot \dfrac \theta 2$ This tangent to $C$ also passes through the point $\tuple {a \paren {\theta - \sin \theta}, a \paren {1 - \cos \theta} }$. {{Finish|Find (or write) the result which gives the equation of a line from its slope and a point through which it passes.}} Hence the result. {{qed}} \end{proof}
22582
\section{Tangent to Cycloid is Vertical at Cusps} Tags: Cycloids \begin{theorem} The tangent to the cycloid whose locus is given by: :$x = a \paren {\theta - \sin \theta}$ :$y = a \paren {1 - \cos \theta}$ is vertical at the cusps. \end{theorem} \begin{proof} From Slope of Tangent to Cycloid, the slope of the tangent to $C$ at the point $\tuple {x, y}$ is given by: :$\dfrac {\d y} {\d x} = \cot \dfrac \theta 2$ At the cusps, $\theta = 2 n \pi$ for $n \in \Z$. Thus at the cusps, the slope of the tangent to $C$ is $\cot n \pi$. From Shape of Cotangent Function: :$\ds \lim_{\theta \mathop \to n \pi} \cot \theta \to \infty$ Hence the result by definition of vertical tangent line. {{qed}} \end{proof}
22583
\section{Tangent to Cycloid passes through Top of Generating Circle} Tags: Cycloids \begin{theorem} Let $C$ be a cycloid generated by the equations: :$x = a \paren {\theta - \sin \theta}$ :$y = a \paren {1 - \cos \theta}$ Then the tangent to $C$ at a point $P$ on $C$ passes through the top of the generating circle of $C$. \end{theorem} \begin{proof} From Tangent to Cycloid, the equation for the tangent to $C$ at a point $P = \tuple {x, y}$ is given by: :$(1): \quad y - a \paren {1 - \cos \theta} = \dfrac {\sin \theta} {1 - \cos \theta} \paren {x - a \theta + a \sin \theta}$ From Equation of Cycloid, the point at the top of the generating circle of $C$ has coordinates $\tuple {2 a, a \theta}$. Substituting $x = 2 a$ in $(1)$: {{begin-eqn}} {{eqn | l = y - a \paren {1 - \cos \theta} | r = \dfrac {\sin \theta} {1 - \cos \theta} \paren {a \theta - a \theta + a \sin \theta} | c = }} {{eqn | ll= \leadsto | l = y | r = a \paren {1 - \cos \theta} + \dfrac {\sin \theta} {1 - \cos \theta} a \sin \theta | c = }} {{eqn | r = \frac {a \paren {1 - \cos \theta}^2 + a \sin^2 \theta} {1 - \cos \theta} | c = }} {{eqn | r = \frac {a \paren {1 - 2 \cos \theta + \cos^2 \theta} + a \sin^2 \theta} {1 - \cos \theta} | c = }} {{eqn | r = \frac {a \paren {1 - 2 \cos \theta + 1} } {1 - \cos \theta} | c = }} {{eqn | r = \frac {a \paren {2 - 2 \cos \theta} } {1 - \cos \theta} | c = }} {{eqn | r = 2 a | c = }} {{end-eqn}} That is, the tangent to $C$ passes through $\tuple {a \theta, 2 a}$ as was required. {{qed}} \end{proof}
22584
\section{Tarski's Geometry is Complete/Corollary} Tags: Tarski's Geometry \begin{theorem} Tarski's geometry does not contain minimal arithmetic. {{explain|contain}} \end{theorem} \begin{proof} Immediate from Tarski's Geometry is Complete and the corollary to Gödel's First Incompleteness Theorem. {{qed}} Category:Tarski's Geometry \end{proof}
22585
\section{Tarski's Undefinability Theorem} Tags: Mathematical Logic \begin{theorem} Let $\ZZ$ be the standard structure $\struct {\Z, +, \cdot, s, <, 0}$ for the language of arithmetic. Let $\operatorname {Th}_\ZZ$ be the sentences which are true in $\ZZ$. Let $\Theta$ be the set of Gödel numbers of those sentences in $\operatorname {Th}_\ZZ$. $\Theta$ is not definable in $\operatorname {Th}_\ZZ$. \end{theorem} \begin{proof} $\operatorname {Th}_\ZZ$ is easily seen to be a consistent extension of minimal arithmetic. (In fact, the axioms in minimal arithmetic were selected based on the behavior of standard arithmetic.) Thus, the theorem is a special case of Set of Gödel Numbers of Arithmetic Theorems Not Definable in Arithmetic (and can be seen to follow immediately). {{qed}} {{Namedfor|Alfred Tarski|cat = Tarski}} \end{proof}
22586
\section{Tartaglia's Formula} Tags: Tetrahedra, Geometry, Solid Geometry \begin{theorem} Let $T$ be a tetrahedron with vertices $\mathbf d_1, \mathbf d_2, \mathbf d_3$ and $\mathbf d_4$. For all $i$ and $j$, let the distance between $\mathbf d_i$ and $\mathbf d_j$ be denoted $d_{ij}$. Then the volume $V_T$ of $T$ satisfies: :$V_T^2 = \dfrac {1} {288} \det \ \begin{vmatrix} 0 & 1 & 1 & 1 & 1\\ 1 & 0 & d_{12}^2 & d_{13}^2 & d_{14}^2 \\ 1 & d_{12}^2 & 0 & d_{23}^2 & d_{24}^2 \\ 1 & d_{13}^2 & d_{23}^2 & 0 & d_{34}^2 \\ 1 & d_{14}^2 & d_{24}^2 & d_{34}^2 & 0 \end{vmatrix}$ \end{theorem} \begin{proof} A proof of Tartaglia's Formula will be found in a proof of the Value of Cayley-Menger Determinant as a tetrahedron is a $3$-simplex. {{proof wanted}} {{Namedfor|Niccolò Fontana Tartaglia|cat = Tartaglia}} \end{proof}
22587
\section{Taxicab Metric is Metric} Tags: Taxicab Metric is Metric, Metric Spaces, Taxicab Metric, Examples of Metrics \begin{theorem} The taxicab metric is a metric. \end{theorem} \begin{proof} From the definition, the taxicab metric is as follows: Let $M_{1'} = \left({A_{1'}, d_{1'}}\right), M_{2'} = \left({A_{2'}, d_{2'}}\right), \ldots, M_{n'} = \left({A_{n'}, d_{n'}}\right)$ be a finite number of metric spaces. Let $\mathcal A$ be the Cartesian product $\displaystyle \prod_{i \mathop = 1}^n A_{i'}$. The taxicab metric on $\displaystyle \mathcal A$ is: : $\displaystyle d_1 \left({x, y}\right) = \sum_{i \mathop = 1}^n d_{i'} \left({x_{i'}, y_{i'}}\right)$ for $x = \left({x_1, x_2, \ldots, x_n}\right), y = \left({y_1, y_2, \ldots, y_n}\right) \in \mathcal A$. \end{proof}
22588
\section{Taxicab Metric is Topologically Equivalent to Chebyshev Distance on Real Vector Space} Tags: Maximum Metric, Taxicab Metric, Chebyshev Distance \begin{theorem} For $n \in \N$, let $\R^n$ be a real vector space. Let $d_1$ be the taxicab metric on $\R^n$. Let $d_\infty$ be the Chebyshev distance on $\R^n$. Then :$\forall x, y \in \R^n: \map {d_\infty} {x, y} \le \map {d_1} {x, y} \le n \cdot \map {d_\infty} {x, y}$ It follows that $d_1$ and $d_\infty$ are Lipschitz equivalent. \end{theorem} \begin{proof} By definition of the Chebyshev distance on $\R^n$, we have: :$\ds \map {d_\infty} {x, y} = \max_{i \mathop = 1}^n {\size {x_i - y_i} }$ where $x = \tuple {x_1, x_2, \ldots, x_n}$ and $y = \tuple {y_1, y_2, \ldots, y_n}$. Let $j$ be chosen so that: :$\ds \size {x_j - y_j} = \max_{i \mathop = 1}^n {\size {x_i - y_i} }$ Then: {{begin-eqn}} {{eqn | l = \map {d_\infty} {x, y} | r = \size {x_j - y_j} | c = {{Defof|Chebyshev Distance on Real Vector Space}} }} {{eqn | o = \le | r = \sum_{i \mathop = 1}^n \size {x_i - y_i} }} {{eqn | r = \map {d_1} {x, y} }} {{eqn | o = \le | r = n \size {x_j - y_j} }} {{eqn | r = n \cdot \map {d_\infty} {x, y} }} {{end-eqn}} {{qed}} \end{proof}
22589
\section{Taxicab Metric on Metric Space Product is Continuous} Tags: Continuous Mappings on Metric Spaces, Taxicab Metric \begin{theorem} Let $M = \struct {A, d}$ be a metric space. Let $\AA$ be the Cartesian product $A \times A$. Let $d_1$ be the taxicab metric on $\AA$: :$\ds \map {d_1} {x, y} = \sum_{i \mathop = 1}^n \map {d_{i'} } {x_{i'}, y_{i'} }$ for $x = \tuple {x_1, x_2, \ldots, x_n}, y = \tuple {y_1, y_2, \ldots, y_n} \in \AA$. Then $d_1: \AA \to \AA$ is a continuous function. \end{theorem} \begin{proof} Recall the definition of continuous mapping in this context. Given metric spaces $M_X = \struct {X, d_X}$ and $M_Y = \struct {Y, d_Y}$, and a mapping $f : X \to Y$, we say that $f$ is $\struct {X, d_X} \to \struct {Y, d_Y}$-continuous {{iff}}: :$\forall x_0 \in X: \forall \epsilon \in \R_{>0}: \exists \delta \in \R_{>0}: \forall x \in X: \map {d_X} {x, x_0} < \delta \implies \map d {\map f x, \map f {x_0} } < \epsilon$ Hence it is necessary to prove that $d: \struct {A \times A, d_1} \to \struct {\R, \size {\,\cdot \,} }$ is continuous. That is, given $\tuple {x_0, y_0} \in A \times A$ and $\epsilon>0$, it is necessary to find $\delta \in \R_{>0}$ such that: :$\forall x, y \in A: \map {d_1} {\tuple {x, y}, \tuple {x_0, y_0} } < \delta \implies \size {\map d {x, y} - \map d {x_0, y_0} } < \epsilon$ Hence, let $\tuple {x_0, y_0} \in A \times A$. If $\map {d_1} {\tuple {x, y}, \tuple {x_0, y_0} } = \map d {x, x_0} + \map d {y, y_0} < \epsilon$, then: :$\size {\map d {x, y} - \map d {x_0, y_0} } \le \size {\map d {x, y} - \map d {x, y_0} } + \size {\map d {x, y_0} - \map d {x_0, y_0} } \le 2 \epsilon$ using {{Metric-space-axiom|2}}. The result follows. {{qed}} \end{proof}
22590
\section{Taxicab Metric on Real Number Plane is Translation Invariant} Tags: Translation Mappings, Taxicab Metric \begin{theorem} Let $\tau_{\mathbf t}: \R^2 \to \R^2$ denote the translation of the Euclidean plane by the vector $\mathbf t = \begin {pmatrix} a \\ b \end {pmatrix}$. Let $d_1$ denote the taxicab metric on $\R^2$. Then $d_1$ is unchanged by application of $\tau$: :$\forall x, y \in \R^2: \map {d_1} {\map \tau x, \map \tau y} = \map {d_1} {x, y}$ \end{theorem} \begin{proof} Let $x = \tuple {x_1, x_2}$ and $y = \tuple {y_1, y_2}$ be arbitrary points in $\R^2$. Then: {{begin-eqn}} {{eqn | l = \map {d_1} {\map \tau x, \map \tau y} | r = \map {d_1} {x - \mathbf t, y - \mathbf t} | c = {{Defof|Translation in Euclidean Space}} }} {{eqn | r = \size {\paren {x_1 - a} - \paren {y_1 - a} } + \size {\paren {x_2 - b} - \paren {y_2 - b} } | c = Definition of $\mathbf t$, {{Defof|Taxicab Metric on Real Number Plane}} }} {{eqn | r = \size {x_1 - y_1} + \size {x_2 - y_2} | c = simplification }} {{eqn | r = \map {d_1} {x, y} | c = {{Defof|Taxicab Metric on Real Number Plane}} }} {{end-eqn}} {{qed}} \end{proof}
22591
\section{Taxicab Metric on Real Number Plane is not Rotation Invariant} Tags: Geometric Rotations, Taxicab Metric \begin{theorem} Let $r_\alpha: \R^2 \to \R^2$ denote the rotation of the Euclidean plane about the origin through an angle of $\alpha$. Let $d_1$ denote the taxicab metric on $\R^2$. Then it is not necessarily the case that: :$\forall x, y \in \R^2: \map {d_1} {\map {r_\alpha} x, \map {r_\alpha} y} = \map {d_1} {x, y}$ \end{theorem} \begin{proof} Proof by Counterexample: Let $x = \tuple {0, 0}$ and $y = \tuple {0, 1}$ be arbitrary points in $\R^2$. Then: {{begin-eqn}} {{eqn | l = \map {d_1} {x, y} | r = \map {d_1} {\tuple {0, 0}, \tuple {0, 1} } | c = Definition of $x$ and $y$ }} {{eqn | r = \size {0 - 0} + \size {0 - 1} | c = {{Defof|Taxicab Metric on Real Number Plane}} }} {{eqn | r = 1 | c = }} {{end-eqn}} Now let $\alpha = \dfrac \pi 4 = 45 \degrees$. {{begin-eqn}} {{eqn | l = \map {d_1} {\map {r_\alpha} x, \map {r_\alpha} y} | r = \map {d_1} {\tuple {0, 0}, \tuple {\dfrac {\sqrt 2} 2, \dfrac {\sqrt 2} 2} } | c = {{Defof|Plane Rotation}} }} {{eqn | r = \size {0 - \dfrac {\sqrt 2} 2} + \size {0 - \dfrac {\sqrt 2} 2} | c = {{Defof|Taxicab Metric on Real Number Plane}} }} {{eqn | r = \sqrt 2 | c = simplification }} {{eqn | o = \ne | r = 1 | c = }} {{end-eqn}} {{qed}} \end{proof}
22592
\section{Taxicab Metric on Real Vector Space is Metric} Tags: Taxicab Metric on Real Vector Space is Metric, Taxicab Metric \begin{theorem} The taxicab metric on the real vector space $\R^n$ is a metric. \end{theorem} \begin{proof} This is an instance of the taxicab metric on the general cartesian product of $A_{1'}, A_{2'}, \ldots, A_{n'}$. This is proved in Taxicab Metric is Metric. {{qed}} \end{proof}
22593
\section{Taxicab Norm is Norm} Tags: Taxicab Norm, Norm Theory, Examples of Norms \begin{theorem} The taxicab norm is a norm on the real and complex numbers. \end{theorem} \begin{proof} By P-Norm is Norm, $\norm {\, \cdot \,}_p$ is a norm. By definition, the taxicab norm is $\norm {\, \cdot \,}_1$. Therefore, the taxicab norm is a norm. {{qed}} \end{proof}
22594
\section{Taylor's Theorem/One Variable with Two Functions} Tags: Taylor's Theorem \begin{theorem} Let $f$ and $g$ be real functions satisfying following conditions: :$(1): \quad f$ is $n + 1$ times differentiable on the open interval $\openint a x$ :$(2): \quad f$ is of differentiability class $C^n$ on the closed interval $\closedint a x$ :$(3): \quad g$ is $k + 1$ times differentiable on the open interval $\openint a x$ :$(4): \quad g$ is of differentiability class $C^k$ on the closed interval $\closedint a x$ :$(5): \quad \map {g^{\paren {k + 1}}} t \ne 0$ for any $t \in \openint a x$ Then the following equation holds for some real number $\xi \in \openint a x$: :$\dfrac {\map {f^{\paren {n + 1} } } \xi /n!} {\map {g^{\paren {k + 1} } } \xi /k!} \paren {x - \xi}^{n - k} = \dfrac {\map f x - \map f a - \map {f'} a \paren {x - a} - \dfrac {\map {f''} a} {2!} \paren {x - a}^2 - \dotsb - \dfrac {\map {f^{\paren n} } a} {n!} \paren {x - a}^n} {\map g x - \map g a - \map {g'} a \paren {x - a} - \dfrac {\map {g''} a} {2!} \paren {x - a}^2 - \dotsb - \dfrac {\map {g^{\paren k} } a} {k!} \paren {x - a}^k}$ or equivalently: {{begin-eqn}} {{eqn | l = \map f x | r = \map f a + \map {f'} a \paren {x - a} + \dfrac {\map {f''} a} {2!} \paren {x - a}^2 + \dotsb + \dfrac {\map {f^{\paren n} } a} {n!} \paren {x - a}^n + R_n }} {{eqn | l = R_n | r = \dfrac {\map {f^{\paren {n + 1} } } \xi / n!} {\map {g^{\paren {k + 1} } } \xi / k!} \paren {x - \xi}^{n - k} \paren {\map g x - \map g a - \map {g'} a \paren {x - a} - \dfrac {\map {g''} a} {2!} \paren {x - a}^2 - \dotsb - \dfrac {\map {g^{\paren k} } a} {k!} \paren {x - a}^k} }} {{end-eqn}} \end{theorem} \begin{proof} We define $F$ and $G$ as follows: {{begin-eqn}} {{eqn | l = \map F t | r = \map f t + \map {f'} t \paren {x - t} + \dfrac {\map {f''} t} {2!} \paren {x - t}^2 + \dotsb + \dfrac {\map {f^{\paren n} } t} {n!} \paren {x - t}^n }} {{eqn | l = \map G t | r = \map g t + \map {g'} t \paren {x - t} + \dfrac {\map {g''} t} {2!} \paren {x - t}^2 + \dotsb + \dfrac {\map {g^{\paren k} } t} {k!} \paren {x - t}^k }} {{end-eqn}} Then $F$ and $G$ are continuous on $\closedint a x$ and differentiable on $\openint a x$. Differentiating {{WRT|Differentiation}} $t$: {{begin-eqn}} {{eqn | l = \map {F'} t | r = \map {f'} t + \paren {\map {f''} t \paren {x - t} - \map {f'} t} + \paren {\frac {\map {f^{\paren 3} } t} {2!} \paren {x - t}^2 - \frac {\map {f^{\paren 2} } t} {1!} \paren {x - t} } + \dotsb + \paren {\frac {\map {f^{\paren {n + 1} } } t} {n!} \paren {x - t}^n - \frac {\map {f^{\paren n} } t} {\paren {n - 1}!} \paren {x - t}^{n - 1} } }} {{eqn | r = \frac {\map {f^{\paren {n + 1} } } t} {n!} \paren {x - t}^n }} {{eqn | l = \map {G'} t | r = \frac {\map {g^{\paren {k + 1} } } t} {k!} \paren {x - t}^k }} {{end-eqn}} By condition $(5)$ of the statement of the theorem, it follows that $G$ does not vanish on $\openint a x$. By Cauchy Mean Value Theorem, there exists a real number $\xi \in \openint a x$ such that: :$\dfrac {\map {F'} \xi} {\map {G'} \xi} = \dfrac {\map F x - \map F a} {\map G x - \map G a}$ That is: :$\dfrac {\map {f^{\paren {n + 1} } } \xi / n!} {\map {g^{\paren {k + 1} } } \xi /k!} \paren {x - \xi}^{n - k} = \dfrac {\map f x - \map f a - \map {f'} a \paren {x - a} - \dfrac {\map {f''} a} {2!} \paren {x - a}^2 - \dotsb - \dfrac {\map {f^{\paren n} } a} {n!} \paren {x - a}^n} {\map g x - \map g a - \map {g'} a \paren {x - a} - \dfrac {\map {g''} a} {2!} \paren {x - a}^2 - \dotsb - \dfrac {\map {g^{\paren k}} a} {k!} \paren {x - a}^k}$ {{qed}} \end{proof}
22595
\section{Taylor Series of Analytic Function has infinite Radius of Convergence} Tags: Taylor Series, Real Analysis \begin{theorem} Let $F$ be a complex function. Let $F$ be analytic everywhere. Let the restriction of $F$ to $\R \to \C$ be a real function $f$. This means: :$\forall x \in \R: \map f x = \map \Re {\map F {x, 0} }, 0 = \map \Im {\map F {x, 0} }$ where $\tuple {x, 0}$ denotes the complex number with real part $x$ and imaginary part $0$. Let $x_0$ be a point in $\R$. Then: :the Taylor series of $f$ about $x_0$ converges to $f$ at every point in $\R$. \end{theorem} \begin{proof} The result follows by Convergence of Taylor Series of Function Analytic on Disk for the case $R = \infty$. {{qed}} \end{proof}
22596
\section{Taylor Series of Holomorphic Function} Tags: Complex Analysis \begin{theorem} Let $a \in \C$ be a complex number. Let $r > 0$ be a real number. Let $f$ be a function holomorphic on some open ball, $D = B \paren {a, r}$. Then: :$\ds \map f z = \sum_{n \mathop = 0}^\infty \frac {\map {f^n} a} {n!} \paren {z - a}^n$ for all $z \in D$. \end{theorem} \begin{proof} In Holomorphic Function is Analytic, it is shown that: :$\ds \map f z = \sum_{n \mathop = 0}^\infty \paren {\frac 1 {2 \pi i} \int_{\partial D} \frac {\map f t} {\paren {t - a}^{n + 1} } \rd t} \paren {z - a}^n$ for all $z \in D$. From Cauchy's Integral Formula: General Result, we have: :$\ds \frac 1 {2 \pi i} \int_{\partial D} \frac {\map f t} {\paren {t - a}^{n + 1} } \rd t = \frac {\map {f^n} a} {n!}$ Hence the result. {{qed}} Category:Complex Analysis \end{proof}
22597
\section{Taylor Series of Logarithm of Gamma Function} Tags: Gamma Function, Natural Logarithm, Natural Logarithms \begin{theorem} Let $\gamma$ denote the Euler-Mascheroni constant. Let $\map \zeta s$ denote the Riemann zeta function. Let $\map \Gamma z$ denote the gamma function. Let $\Log$ denote the natural logarithm. Then $\map \Log {\map \Gamma z}$ has the power series expansion: {{begin-eqn}} {{eqn | l = \map \Log {\map \Gamma z} | r = -\map \gamma {z - 1} + \sum_{k \mathop = 2}^\infty \frac {\paren {-1}^k \map \zeta k} k \paren {z - 1}^k }} {{end-eqn}} which is valid for all $z \in \C$ such that $\cmod {z - 1} < 1$. \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \map \Gamma {x + 1} | r = x \map \Gamma x | c = Gamma Difference Equation }} {{eqn | r = \paren {x } \paren {x - 1 } \map \Gamma {x - 1 } | c = }} {{eqn | r = \paren {x } \paren {x - 1 } \paren {x - 2 } \map \Gamma {x - 2 } | c = }} {{eqn | r = \paren {x } \paren {x - 1 } \paren {x - 2 } \cdots \paren {x - \floor x } \map \Gamma {x -\floor x } | c = For $\floor x$, see the {{Defof|Floor Function}} }} {{end-eqn}} Hence: {{begin-eqn}} {{eqn | l = \map \Log {\map \Gamma {x + 1} } | r = \map \Log { \paren {x } \paren {x - 1 } \paren {x - 2 } \cdots \paren {x - \floor x } \map \Gamma {x - \floor x } } | c = }} {{eqn | r = \map \Log x + \map \Log {x - 1 } + \map \Log {x - 2 } + \cdots + \map \Log {x - \floor x } + \map \Log {\map \Gamma {x - \floor x } } | c = Sum of Logarithms }} {{end-eqn}} Therefore: {{begin-eqn}} {{eqn | l = \frac \d {\d x} \map \Log {\map \Gamma {x + 1} } | r = \frac \d {\d x} \paren {\map \Log x + \map \Log {x - 1} + \map \Log {x - 2 } + \cdots + \map \Log {x - \floor x } + \map \Log {\map \Gamma {x - \floor x} } } | c = }} {{eqn | r = \frac \d {\d x} \map \Log x + \frac \d {\d x} \map \Log {x - 1} + \frac \d {\d x} \map \Log {x - 2} + \cdots + \frac \d {\d x} \map \Log {x - \floor x} + \frac \d {\d x} \map \Log {\map \Gamma {x - \floor x} } | c = Sum Rule for Derivatives }} {{eqn | r = \frac 1 x + \frac 1 {x - 1} + \frac 1 {x - 2} + \cdots + \frac 1 {x - \floor x} + \frac \d {\d x} \map \Log {\map \Gamma {x - \floor x} } | c = Derivative of Natural Logarithm Function }} {{eqn | n = 1 | r = \sum_{k \mathop = 0}^{\floor x} \frac 1 {x - k} + \frac \d {\d x} \map \Log {\map \Gamma {x - \floor x} } | c = }} {{end-eqn}} Setting: :$z = \paren {x - \floor x} \leadsto \d z = \d x$ :$M = \floor x + 1$: we have: :$z + M = x + 1$ Hence: {{begin-eqn}} {{eqn | l = \frac \d {\d x} \map \Log {\map \Gamma {x - \floor x} } | r = \frac \d {\d x} \map \Log {\map \Gamma {x + 1} } - \sum_{k \mathop = 0}^{\floor x} \frac 1 {x - k} | c = rearranging $(1)$ }} {{eqn | l = \frac \d {\d z} \map \Log {\map \Gamma z} | r = \frac \d {\d z} \map \Log {\map \Gamma {z + M} } - \sum_{k \mathop = 0}^{M - 1 } \frac 1 {\paren {z + M - 1} - k} | c = substituting with the values above }} {{eqn | n = 2 | l = \frac \d {\d z} \map \Log {\map \Gamma z} | r = \frac \d {\d z} \map \Log {\map \Gamma {z + M} } - \sum_{k \mathop = 0}^{M - 1 } \frac 1 {z + k} | c = }} {{eqn | ll= \leadsto | l = \frac {\d^n} {\d z^n} \map \Log {\map \Gamma z} | r = \frac {\d^n} {\d z^n} \map \Log {\map \Gamma {z + M} } + \sum_{k \mathop = 0}^{M - 1} \frac {\paren {-1}^n \paren {n - 1}!} {\paren {z + k}^n} | c = Sum Rule for Derivatives and Nth Derivative of Reciprocal of Mth Power }} {{end-eqn}} Recall Stirling's Formula for Gamma Function: {{begin-eqn}} {{eqn | l = \map \Gamma {z + 1} | r = \sqrt {2 \pi z} \, z^z e^{-z} \paren {1 + \dfrac 1 {12 z} + \dfrac 1 {288 z^2} - \dfrac {139} {51 \, 480 z^3} + \cdots} | c = }} {{eqn | l = \map \Log {\map \Gamma {z + 1} } | r = \frac 1 2 \map \Log {2 \pi } + \frac 1 2 \map \Log z + z \map \Log z - z + \map \log {1 + \dfrac 1 {12 z} + \dfrac 1 {288 z^2} - \dfrac {139} {51 \, 480 z^3} + \cdots} | c = Sum of Logarithms and Logarithm of Power/Natural Logarithm }} {{eqn | l = \frac \d {\d z} \map \Log {\map \Gamma {z + 1} } | r = 0 + \dfrac 1 {2 z} + \paren {\map \Log z + \dfrac z z } - 1 + \map \OO {\frac 1 z} | c = Product Rule for Derivatives and Derivative of Natural Logarithm Function }} {{eqn | r = \map \Log z + \map \OO {\frac 1 z} | c = For $\map \OO {\frac 1 z}$, see {{Defof|Big-O Notation}} }} {{eqn | ll = \leadsto | l = \frac {\d^n} {\d z^n} \map \Log {\map \Gamma {z + 1} } | r = \map \OO {\frac 1 z} | c = For $n > 1$ }} {{end-eqn}} We now have: :$\ds \frac \d {\d z} \map \Log {\map \Gamma {z + 1} } = \map \Log z + \map \OO {\frac 1 z}$ Therefore: {{begin-eqn}} {{eqn | l = \frac \d {\d z} \map \Log {\map \Gamma {z + M } } | r = \map \Log {z + M - 1} + \map \OO {\frac 1 {z + M - 1} } | c = }} {{eqn | l = \frac \d {\d z} \map \Log {\map \Gamma {z + M} } - \sum_{k \mathop = 0}^{M - 1} \frac 1 {z + k} | r = \map \Log {z + M - 1} - \sum_{k \mathop = 0}^{M - 1} \frac 1 {z + k} + \map \OO {\frac 1 {z + M - 1} } | c = subtracting $\ds \sum_{k \mathop = 0}^{M - 1} \frac 1 {z + k}$ from both sides of the equation }} {{eqn | l = \lim_{M \mathop \to \infty} \paren {\frac \d {\d z} \map \Log {\map \Gamma {z + M} } - \sum_{k \mathop = 0}^{M - 1} \frac 1 {z + k} } | r = \lim_{M \mathop \to \infty} \paren {\map \Log {z + M} - \sum_{k \mathop = 0}^{M - 1} \frac 1 {z + k} } + \lim_{M \mathop \to \infty} \map \OO {\frac 1 {z + M} } | c = {{Defof|Limit of Real Function}}, taking the limit as M goes to $\infty$ on both sides of the equation }} {{eqn | r = \lim_{M \mathop \to \infty} \paren {\map \Log {z + M} - \sum_{k \mathop = 0}^{M - 1} \frac 1 {z + k} } | c = $\ds \lim_{M \mathop \to \infty} \map \OO {\frac 1 {z + M} } = 0$ }} {{end-eqn}} We now have: :$\ds \lim_{M \mathop \to \infty} \frac \d {\d z} \map \Log {\map \Gamma {z + M} } = \lim_{M \mathop \to \infty} \map \Log {z + M} $ Setting $z = 1$ into $(2)$: {{begin-eqn}} {{eqn | l = \frac \d {\d z} \map \Log {\map \Gamma 1} | r = \lim_{M \mathop \to \infty} \frac \d {\d z} \map \Log {\map \Gamma {1 + M} } - \sum_{k \mathop = 0}^{M - 1} \frac 1 {1 + k} | c = }} {{eqn | r = \lim_{M \mathop \to \infty} \map \Log {1 + M} - \sum_{k \mathop = 0}^{M - 1} \frac 1 {1 + k} | c = }} {{eqn | r = \lim_{M \mathop \to \infty} \map \Log {1 + M} - \sum_{j \mathop = 1}^{M} \frac 1 j | c = setting $j = 1 + k$ }} {{eqn | r = -\gamma | c = {{Defof|Euler-Mascheroni Constant}} }} {{end-eqn}} Also: :$\ds \frac {\d^{1 + k} } {\d z^{1 + k}} \map \Log {\map \Gamma {z + 1} } = \map \OO {\frac 1 z}$ shows that: :$\ds \lim_{M \mathop \to \infty} \frac {\d^{1 + k} } {\d z^{1 + k} } \map \Log {\map \Gamma {M + 1} } = 0$ thus for $n > 1$: {{begin-eqn}} {{eqn | l = \frac {\d^n} {\d z^n} \map \Log {\map \Gamma 1} | r = \frac {\d^n} {\d z^n} \map \Log {\map \Gamma {1 + M} } + \sum_{k \mathop = 1}^M \frac {\paren {-1}^n \paren {n - 1}!} {k^n} | c = }} {{eqn | l = \frac {\d^n} {\d z^n} \map \Log {\map \Gamma 1} | r = \lim_{M \mathop \to \infty} \paren {\frac {\d^n} {\d z^n} \map \Log {\map \Gamma {1 + M} } + \sum_{k \mathop = 1}^M \frac {\paren {-1}^n \paren {n - 1}!} {k^n} } }} {{eqn | l = \frac {\d^n} {\d z^n} \map \Log {\map \Gamma 1} | r = \paren {-1}^n \paren {n - 1}! \sum_{k \mathop = 1}^{\infty} \frac 1 {k^n} }} {{eqn | r = \paren {-1}^n \paren {n - 1}! \map \zeta n | c = {{Defof|Riemann Zeta Function}} }} {{end-eqn}} Thus by definition of Taylor series: {{begin-eqn}} {{eqn | l = \map f z | r = \sum_{n \mathop = 0}^\infty \map {f^{\paren n} } a \frac {\paren {z - a}^n} {n!} }} {{eqn | r = \map f a + \map {f^{\paren 1} } a \frac {\paren {z - a}^1} {1!} + \map {f^{\paren 2} } a \frac {\paren {z - a}^2} {2!} + \cdots }} {{eqn | l = \map \Log {\map \Gamma z} | r = \map \Log {\map \Gamma 1} - \gamma \paren {z - 1} + \sum_{k \mathop = 2}^\infty \frac {\paren {-1}^k \map \zeta k \paren {k - 1}!} {k!} \paren {z - 1}^k }} {{eqn | r = -\gamma \paren {z - 1} + \sum_{k \mathop = 2}^\infty \frac{\paren {-1}^k \map \zeta k} k \paren {z - 1}^k | c = Gamma of One is One and Natural Logarithm of 1 is 0 }} {{end-eqn}} From Zeroes of Gamma Function, we see that $\map \Gamma z$ is non-zero everywhere. Thus $\map \Log {\map \Gamma z}$ has poles only where $\Gamma$ does, that is, the negative integers. Since the radius of convergence of a power series is equal to the distance of its center to the closest point where the function is not analytic: The radius of convergence of $\map \Log {\map \Gamma z}$ is $\cmod {1 - 0} = 1$. {{qed}} Category:Gamma Function Category:Natural Logarithms \end{proof}
22598
\section{Taylor Series reaches closest Singularity} Tags: Taylor Series, Real Analysis \begin{theorem} Let $F$ be a complex function. Let $F$ be analytic everywhere except at a finite number of singularities. Let a singularity of $F$ be one of the following: :a pole :an essential singularity :a branch point In the latter case $F$ is a restriction of a multifunction to one of its branches. Let $x_0$ be a real number. Let $F$ be analytic at the complex number $\tuple {x_0, 0}$. Let $R \in \R_{>0}$ be the distance from the complex number $\tuple {x_0, 0}$ to the closest singularity of $F$. Let the restriction of $F$ to $\R \to \C$ be a real function $f$. This means: :$\forall x \in \R: \map f x = \map \Re {\map F {x, 0} }, 0 = \map \Im {\map F {x, 0} }$ where $\tuple {x, 0}$ denotes the complex number with real part $x$ and imaginary part $0$. Then: :the Taylor series of $f$ about $x_0$ converges to $f$ at every point $x \in \R$ satisfying $\size {x - x_0} < R$ \end{theorem} \begin{proof} We have that $F$ is analytic everywhere except at its singularities. Also, the distance from the complex number $\tuple {x_0, 0}$ to the closest singularity of $F$ is $R$. Therefore: :$F$ is analytic at every point $z \in \C$ satisfying $\size {z - \tuple {x_0, 0} } < R$ where $\tuple {x_0 , 0}$ denotes the complex number with real part $x_0$ and imaginary part $0$. The result follows by Convergence of Taylor Series of Function Analytic on Disk. {{qed}} \end{proof}
22599
\section{Telescoping Series/Example 1} Tags: Telescoping Series, Series \begin{theorem} Let $\sequence {b_n}$ be a sequence in $\R$. Let $\sequence {a_n}$ be a sequence whose terms are defined as: :$a_k = b_k - b_{k + 1}$ Then: :$\ds \sum_{k \mathop = 1}^n a_k = b_1 - b_{n + 1}$ If $\sequence {b_n}$ converges to zero, then: :$\ds \sum_{k \mathop = 1}^\infty a_k = b_1$ \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \ds \sum_{k \mathop = 1}^n a_k | r = \sum_{k \mathop = 1}^n \paren {b_k - b_{k + 1} } | c = }} {{eqn | r = \sum_{k \mathop = 1}^n b_k - \sum_{k \mathop = 1}^n b_{k + 1} | c = }} {{eqn | r = \sum_{k \mathop = 1}^n b_k - \sum_{k \mathop = 2}^{n + 1} b_k | c = Translation of Index Variable of Summation }} {{eqn | r = b_1 + \sum_{k \mathop = 2}^n b_k - \sum_{k \mathop = 2}^n b_k - b_{n + 1} | c = }} {{eqn | r = b_1 - b_{n + 1} | c = }} {{end-eqn}} If $\sequence {b_k}$ converges to zero, then $b_{n + 1} \to 0$ as $n \to \infty$. Thus: :$\ds \lim_{n \mathop \to \infty} s_n = b_1 - 0 = b_1$ So: :$\ds \sum_{k \mathop = 1}^\infty a_k = b_1$ {{Qed}} \end{proof}
22600
\section{Temperature of Body under Newton's Law of Cooling} Tags: Thermodynamics, Heat \begin{theorem} Let $B$ be a body in an environment whose ambient temperature is $H_a$. Let $H$ be the temperature of $B$ at time $t$. Let $H_0$ be the temperature of $B$ at time $t = 0$. Then: :$H = H_a - \paren {H_0 - H_a} e^{-k t}$ where $k$ is some positive constant. \end{theorem} \begin{proof} By Newton's Law of Cooling: :The rate at which a hot body loses heat is proportional to the difference in temperature between it and its surroundings. We have the differential equation: :$\dfrac {\d H} {\d t} \propto - \paren {H - H_a}$ That is: :$\dfrac {\d H} {\d t} = - k \paren {H - H_a}$ where $k$ is some positive constant. This is an instance of the Decay Equation, and so has a solution: :$H = H_a + \paren {H_0 - H_a} e^{-k t}$ {{qed}} {{Namedfor|Isaac Newton|cat = Newton}} \end{proof}
22601
\section{Tempered Distribution Space is Proper Subset of Distribution Space} Tags: Distributions \begin{theorem} Let $\map {\DD'} \R$ be the distribution space. Let $\map {\SS'} \R$ be the tempered distribution space. Then $\map {\SS'} \R$ is a proper subset of $\map {\DD'} \R$: :$\map {\SS'} \R \subsetneqq \map {\DD'} \R$ \end{theorem} \begin{proof} By Convergence of Sequence of Test Functions in Test Function Space implies Convergence in Schwartz Space we have that $\map {\SS'} \R \subseteq \map {\DD'} \R$. {{Research|how?}} Consider the real function $\map f x = e^{x^2}$. We have that: :Real Power Function for Positive Integer Power is Continuous :Exponential Function is Continuous/Real Numbers :Composite of Continuous Mappings is Continuous Thus, $f$ is a continuous real function. Also: :$\forall x \in \R : e^{x^2} < \infty$ Hence, $f$ is locally integrable. By Locally Integrable Function defines Distribution, $T_f \in \map {\DD'} \R$. {{AimForCont}} $T_f$ is a tempered distribution. We have that $e^{-x^2}$ is a Schwartz test function. Then: {{begin-eqn}} {{eqn | l = \map {T_f} {e^{-x^2} } | r = \int_{- \infty}^\infty e^{x^2} e^{-x^2} \rd x }} {{eqn | r = \int_{- \infty}^\infty 1 \rd x }} {{eqn | r = \infty }} {{end-eqn}} Hence, $\map {T_f} {e^{-x^2} } \notin \R$. This is a contradiction. Therefore, $T_f \notin \map {\SS'} \R$ while at the same time $T_f \in \map {\DD'} \R$. {{qed}} \end{proof}
22602
\section{Tensor Product of Free Modules is Free} Tags: Commutative Algebra, Homological Algebra, Free Modules, Module Theory \begin{theorem} Let $A$ be a commutative ring with unity. Let $F$ and $F'$ be free $A$-modules. Then the tensor product $F \otimes_A F'$ is a free $A$-module. \end{theorem} \begin{proof} By Free Module is Isomorphic to Free Module on Set there are sets $I$ and $I'$ and isomorphisms $\Psi : A^{\paren I} \to F$ and $\Psi' : A^{\paren {I'} } \to F'$. By Tensor Product Distributes over Direct Sum, there is an isomorphism: :$\ds A^{\paren I} \otimes_A A^{\paren {I'} } \cong \bigoplus_{i \mathop \in I} \bigoplus_{i' \mathop \in I'} A$ By Direct Sum of Direct Sums is Direct Sum, there is an isomorphism: :$\ds \bigoplus_{i \mathop \in I} \bigoplus_{i' \mathop \in I'} A \cong \bigoplus_{\tuple {i, i'} \mathop \in I \times I'} A$ Hence $A^{\paren I} \otimes_A A^{\paren {I'} } \cong A^{\paren {I \times I'} }$ by definition of $A^{\paren {I \times I'} }$. By Free Module on Set is Free $A^{\paren {I \times I'} }$ is free. By Module Isomorphic to Free Module is Free $A^{\paren I} \otimes_A A^{\paren {I'} }$ is free. {{qed}} Category:Commutative Algebra Category:Free Modules Category:Homological Algebra Category:Module Theory \end{proof}
22603
\section{Tensor Product of Projective Modules is Projective} Tags: Commutative Algebra \begin{theorem} Let $A$ be a commutative ring with unity. Let $P$ and $Q$ be projective $A$-modules. Then the tensor product $P \otimes_A Q$ is a projective $A$-module. \end{theorem} \begin{proof} By Projective iff Direct Summand of Free Module, there exist $A$-modules $P'$ and $Q'$, such that $P \oplus P'$ and $Q \oplus Q'$ are free. By Tensor Product Distributes over Direct Sum, there is an isomorphism : $\paren {P \oplus P'} \otimes_A \paren {Q \oplus Q'} \cong \paren {P \otimes_A Q} \oplus \paren {P' \otimes_A Q} \oplus \paren {P \otimes_A Q'} \oplus \paren {P' \otimes_A Q'}$ By Tensor Product of Free Modules is Free the {{LHS}} is free. Hence $P \otimes_A Q$ is a direct summand of a free module. By Projective iff Direct Summand of Free Module $P \otimes_A Q$ is projective. {{qed}} Category:Commutative Algebra \end{proof}
22604
\section{Tensor with Zero Element is Zero in Tensor} Tags: Tensor Algebra, Homological Algebra \begin{theorem} Let $R$ be a ring. Let $M$ be a right $R$-module. Let $N$ be a left $R$-module. Let $M \otimes_R N$ denote their tensor product. Then: :$0\otimes_R n = m \otimes_R 0 = 0 \otimes_R 0$ is the zero in $M \otimes_R N$. \end{theorem} \begin{proof} Let $m \in M$ and $n \in N$ Then {{begin-eqn}} {{eqn | l = m \otimes_R n | r = \paren {m + 0} \otimes_R n | c = {{GroupAxiom|2}} }} {{eqn | r = m \otimes_R n + 0 \otimes_R n | c = {{Defof|Tensor Equality}} }} {{eqn | r = m \otimes_R \paren {n + 0} | c = {{GroupAxiom|2}} }} {{eqn | r = m \otimes_R n + m\otimes_R 0 | c = {{Defof|Tensor Equality}} }} {{eqn | r = m \otimes_R n + 0 \otimes_R \paren {n + 0} | c = {{GroupAxiom|2}} }} {{eqn | r = m \otimes_R n + m \otimes_R 0 + 0 \otimes_R n + 0 \otimes_R 0 | c = {{Defof|Tensor Equality}} }} {{eqn | r = m \otimes_R n + 0 \otimes_R 0 | c = {{Defof|Tensor Equality}} }} {{end-eqn}} Hence $0 \otimes_R n$, $m \otimes_R 0$ and $0 \otimes_R 0$ must all be identity elements for $M \otimes_R N$ as a left module. {{qed}} Category:Tensor Algebra Category:Homological Algebra \end{proof}
22605
\section{Termial on Real Numbers is Extension of Integers} Tags: Number Theory, Termial Function \begin{theorem} The termial function as defined on the real numbers is an extension of its definition on the integers $\Z$. \end{theorem} \begin{proof} From the definition of the termial function on the integers: :$\ds n? = \sum_{k \mathop = 1}^n k = 1 + 2 + \cdots + n$ From Closed Form for Triangular Numbers, we have that: :$\ds \forall n \in \Z_{> 0}: \sum_{k \mathop = 1}^n k = \dfrac {n \paren {n + 1} } 2$ This agrees with the definition of the termial function on the real numbers. Hence the result, by definition of extension. {{qed}} \end{proof}
22606
\section{Terminal Object is Unique} Tags: Category Theory \begin{theorem} Let $\mathbf C$ be a metacategory. Let $1$ and $1'$ be two terminal objects of $\mathbf C$. Then there is a unique isomorphism $u: 1 \to 1'$. Hence, terminal objects are unique up to unique isomorphism. \end{theorem} \begin{proof} Consider the following commutative diagram: ::$\begin{xy} <-4em,0em>*+{1} = "M", <0em,0em> *+{1'}= "N", <0em,-4em>*+{1} = "M2", <4em,-4em>*+{1'}= "N2", "M";"N" **@{-} ?>*@{>} ?*!/_.6em/{u}, "M";"M2" **@{-} ?>*@{>} ?*!/^.6em/{\operatorname{id}_1}, "N";"M2" **@{-} ?>*@{>} ?*!/_.6em/{v}, "N";"N2" **@{-} ?>*@{>} ?*!/_1em/{\operatorname{id}_{1'}}, "M2";"N2"**@{-} ?>*@{>} ?*!/^.6em/{u}, \end{xy}$ It commutes as each of the morphisms in it points to a terminal object, and hence is unique. Thus, $v$ is an inverse to $u$, and so $u$ is an isomorphism. {{qed}} \end{proof}
22607
\section{Terminal Velocity of Body under Fall Retarded Proportional to Square of Velocity} Tags: Mechanics \begin{theorem} Let $B$ be a body falling in a gravitational field. Let $B$ be falling through a medium which exerts a resisting force of magnitude $k v^2$ upon $B$ which is proportional to the square of the velocity of $B$ relative to the medium. Then the terminal velocity of $B$ is given by: :$v = \sqrt {\dfrac {g m} k}$ \end{theorem} \begin{proof} Let $B$ start from rest. The differential equation governing the motion of $B$ is given by: :$m \dfrac {\d^2 \mathbf s} {\d t^2} = m \mathbf g - k \paren {\dfrac {\d \mathbf s} {\d t} }^2$ Dividing through by $m$ and setting $c = \dfrac k m$ gives: :$\dfrac {\d^2 \mathbf s} {\d t^2} = \mathbf g - c \paren {\dfrac {\d \mathbf s} {\d t} }^2$ By definition of velocity: :$\dfrac {\d \mathbf v} {\d t} = \mathbf g - c \mathbf v^2$ for some constant $c$. and so taking magnitudes of the vector quantities: {{begin-eqn}} {{eqn | l = \int \dfrac {\d v} {g - c v^2} | r = \int \rd t | c = }} {{eqn | ll= \leadsto | l = \dfrac 1 {2 c} \sqrt {\frac c g} \ln \paren {\frac {\sqrt {\frac c g} + v} {\sqrt {\frac c g} - v} } | r = t + c_1 | c = Primitive of $\dfrac 1 {a^2 - x^2}$: Logarithm Form }} {{end-eqn}} We have $v = 0$ when $t = 0$ which leads to $\ln 1 = 0 + c_1$ and thus $c_1 = 0$. Hence: {{begin-eqn}} {{eqn | l = \ln \paren {\frac {\sqrt {\frac c g} + v} {\sqrt {\frac c g} - v} } | r = 2 t \sqrt {c g} | c = }} {{eqn | ll= \leadsto | l = \frac {\sqrt {\frac c g} + v} {\sqrt {\frac c g} - v} | r = e^{2 t \sqrt {c g} } | c = }} {{eqn | ll= \leadsto | l = \sqrt {\frac c g} + v | r = \paren {\sqrt {\frac c g} - v} e^{2 t \sqrt {c g} } | c = }} {{eqn | ll= \leadsto | l = v \paren {e^{2 t \sqrt {c g} } + 1} | r = v \paren {e^{2 t \sqrt {c g} } - 1} \sqrt {\frac g c} | c = }} {{eqn | ll= \leadsto | l = v | r = \sqrt {\frac g c} \paren {\frac {e^{2 t \sqrt {c g} } - 1} {e^{2 t \sqrt {c g} } + 1} } | c = }} {{eqn | r = \sqrt {\frac g c} \paren {\frac {1 - e^{-2 t \sqrt {c g} } } {1 + e^{-2 t \sqrt {c g} } } } | c = }} {{end-eqn}} Since $c > 0$ it follows that $v \to \sqrt {\dfrac g c}$ as $t \to \infty$. Thus in the limit: :$v = \sqrt {\dfrac g c} = \sqrt {\dfrac {g m} k}$ {{qed}} \end{proof}
22608
\section{Terminal Velocity of Body under Fall Retarded Proportional to Velocity} Tags: Mechanics \begin{theorem} Let $B$ be a body falling in a gravitational field. Let $B$ be falling through a medium which exerts a resisting force $k \mathbf v$ upon $B$ which is proportional to the velocity of $B$ relative to the medium. Then the terminal velocity of $B$ is given by: :$v = \dfrac {g m} k$ \end{theorem} \begin{proof} Let $B$ start from rest. From Motion of Body Falling through Air, the differential equation governing the motion of $B$ is given by: :$m \dfrac {\d^2 \mathbf s} {\d t^2} = m \mathbf g - k \dfrac {\d \mathbf s} {\d t}$ Dividing through by $m$ and setting $c = \dfrac k m$ gives: :$\dfrac {\d^2 \mathbf s} {\d t^2} = \mathbf g - c \dfrac {\d \mathbf s} {\d t}$ By definition of velocity: :$\dfrac {\d \mathbf v} {\d t} = \mathbf g - c \mathbf v$ and so: {{begin-eqn}} {{eqn | l = \int \dfrac {\d \mathbf v} {\mathbf g - c \mathbf v} | r = \int \rd t | c = }} {{eqn | ll= \leadsto | l = -\dfrac 1 c \map \ln {\mathbf g - c \mathbf v} | r = t + c_1 | c = }} {{eqn | ll= \leadsto | l = \mathbf g - c \mathbf v | r = \mathbf c_2 e^{-c t} | c = }} {{end-eqn}} When $t = 0$ we have that $\mathbf v = 0$ and so: :$\mathbf c_2 = \mathbf g$ Hence by taking magnitudes: :$v = \dfrac g c \paren {1 - e^{-c t} }$ Since $c > 0$ it follows that $v \to \dfrac g c$ as $t \to \infty$. Thus in the limit: :$v = \dfrac g c = \dfrac {g m} k$ {{qed}} \end{proof}
22609
\section{Terms in Convergent Series Converge to Zero} Tags: Convergent Series, Series \begin{theorem} Let $\sequence {a_n}$ be a sequence in any of the standard number fields $\Q$, $\R$, or $\C$. Suppose that the series $\ds \sum_{n \mathop = 1}^\infty a_n$ converges in any of the standard number fields $\Q$, $\R$, or $\C$. Then: :$\ds \lim_{n \mathop \to \infty} a_n = 0$ {{expand|Expand (on a different page) to Banach spaces}} \end{theorem} \begin{proof} Let $\ds s = \sum_{n \mathop = 1}^\infty a_n$. Then $\ds s_N = \sum_{n \mathop = 1}^N a_n \to s$ as $N \to \infty$. Also, $s_{N - 1} \to s$ as $N \to \infty$. Thus: {{begin-eqn}} {{eqn | l = a_N | r = \paren {a_1 + a_2 + \cdots + a_{N - 1} + a_N} - \paren {a_1 + a_2 + \cdots + a_{N - 1} } | c = }} {{eqn | r = s_N - s_{N - 1} | c = }} {{eqn | o = \to | r = s - s = 0 \text{ as } N \to \infty | c = }} {{end-eqn}} Hence the result. {{qed}} \end{proof}
22610
\section{Test Function/Examples/Exponential of One over x Squared minus One} Tags: Examples of Test Functions \begin{theorem} Let $\phi : \R \to \R$ be a real function with support on $x \in \closedint {-1} 1$ such that: :$\map \phi x = \begin{cases} \ds \map \exp {\frac 1 {x^2 - 1}} & : \size x < 1 \\ 0 & : \size x \ge 1 \end{cases}$ Then $\phi$ is a test function. \end{theorem} \begin{proof} Consider a real function $f : \R \to \R$ such that: :$\map f x = \begin {cases} \map \exp {-\dfrac 1 x} & : x > 0 \\ 0 & : x \le 0 \end {cases}$ From Nth Derivative of Exponential of Minus One over x: :$\dfrac {\d^n} {\d x^n} \map \exp {-\dfrac 1 x} = \dfrac {\map {P_n} x} {x^{2 n} } \map \exp {-\dfrac 1 x}$ where $\map {P_n} x$ is a real polynomial of degree $n$. Then the {{RHS}} can be rewritten in terms of at most $n + 1$ terms of the form $\dfrac 1 {x^m} \map \exp {-\dfrac 1 x}$ where $m \in \N$. Let us take the limit $x \to 0$ from the right: {{begin-eqn}} {{eqn | l = \lim_{x \mathop \to 0^+} \frac 1 {x^m} \map \exp {-\frac 1 x} | r = \lim_{z \mathop \to \infty} \frac {z^m} {\map \exp z} | c = Substitution $\ds z = \frac 1 x$ }} {{eqn | r = 0 | c = Limit at Infinity of Polynomial over Complex Exponential }} {{end-eqn}} Therefore: :$\ds \lim_{x \mathop \to 0^+} \frac {\map {P_n} x} {x^{2 n} } \map \exp {-\frac 1 x} = 0$ By construction: :$\ds \map \phi x = \map f {1 - x^2} = \begin {cases} \map \exp {-\dfrac 1 {1 - x^2} } & : 1 - x^2 > 0 \\ 0 & : 1 - x^2 \le 0 \end {cases}$ where: {{begin-eqn}} {{eqn | l = 1 - x^2 | o = > | r = 0 }} {{eqn | ll= \leadsto | l = 1 | o = > | r = x^2 }} {{eqn | ll= \leadsto | l = \size x | o = < | r = 1 }} {{end-eqn}} Furthermore: {{begin-eqn}} {{eqn | l = \dfrac {\d \map \phi x} {\d x} | r = \dfrac {\d \map f y} {\d y} \paren {-2 x} }} {{eqn | r = \frac {\map {P_1} y} {y^2} \map \exp {-\frac 1 y} \paren {-2 x} }} {{eqn | r = \frac {\map {M_3} x} {y^2} \map \exp {-\frac 1 y} }} {{end-eqn}} where $y = 1 - x^2$ and $\map {M_3} x$ is a real polynomial of degree $3$. Similarly, any higher derivative of $\map \phi x$ will have $\map {M_k} x$ instead of $\map {P_n} x$ with $k \ge n$. Thus: :$\dfrac {\d^n} {\d x^n} \map \phi x = \dfrac {\map {M_k} x} {y^{2 n} } \map \exp {-\dfrac 1 y}$ Consequently: {{begin-eqn}} {{eqn | l = \lim_{x \mathop \to -1^+} \frac {\map {M_k} x} {y^{2 n} } \map \exp {-\frac 1 y} | r = \lim_{x \mathop \to -1^+} \map {M_k} x \lim_{x \mathop \to -1^+} \frac {\map \exp {- \frac 1 y} } {y^{2n} } | c = Product Rule for Limits of Real Functions }} {{eqn | r = C \lim_{y \mathop \to 0^+} \frac {\map \exp {- \frac 1 y} } {y^{2n} } | c = $C \in \R$ }} {{eqn | r = 0 }} {{end-eqn}} Analogously: :$\ds \lim_{x \mathop \to 1^-} \frac {\map {M_k} x} {y^{2 n} } \map \exp {-\frac 1 y} = 0$ Since outside of the support we have that $\map \phi x = 0$, the limit coming from outside is also $0$. Therefore, $\map \phi x$ is smooth at the boundaries of its support. Also: :$\forall x \in \openint {-1} 1 : \map \phi x \in \CC^\infty$ By definition, $\phi$ is a test function. {{qed}} \end{proof}
22611
\section{Test Function Space with Pointwise Addition and Pointwise Scalar Multiplication forms Vector Space} Tags: Functional Analysis, Examples of Vector Spaces \begin{theorem} Let $\map \DD {\R^d}$ be the test function space. Let $\struct {\C, +_\C, \times_\C}$ be the field of complex numbers. Let $\paren +$ be the pointwise addition of test functions. Let $\paren {\, \cdot \,}$ be the pointwise scalar multiplication of test functions over $\C$. Then $\struct {\map \DD {\R^d}, +, \, \cdot \,}_\C$ is a vector space. \end{theorem} \begin{proof} Let $f, g, h \in \map \DD {\R^d}$ be test functions with the compact support $K$. Let $\lambda, \mu \in \C$. Let $\map 0 x$ be a real-valued function such that: :$\map 0 x : \R^d \to 0$. Let us use real number addition and multiplication. $\forall x \in \R^d$ define pointwise addition as: :$\map {\paren {f + g}} x := \map f x +_\C \map g x$. Define pointwise scalar multiplication as: :$\map {\paren {\lambda \cdot f}} x := \lambda \times_\C \map f x$ Let $\map {\paren {-f} } x := -\map f x$. \end{proof}
22612
\section{Test Function with Vanishing Partial Derivative} Tags: Partial Differentiation, Examples of Test Functions \begin{theorem} Let $\phi \in \map \DD {\R^2}$ be a test function such that: :$\tuple {x, y} \stackrel \phi {\longrightarrow} \map \phi {x, y}$ Suppose $\phi$ is a solution to the following partial differential equation: :$\ds \dfrac {\partial \phi}{\partial x} = 0$ Then $\phi$ is identically $0$. \end{theorem} \begin{proof} $\ds \dfrac {\partial \phi}{\partial x} = 0$ implies that: :$\forall x \in \R : \map \phi {x, y} = \map C y$ where $C : \R \to \C$ is a complex-valued function. By definition, $\phi$ is a test function. Hence, $\phi$ must have a compact support $\Omega \subset \R^2$. Let $\map {B^-_\epsilon} 0 \subset \R^2$ be a closed ball in Euclidean space such that: :$\Omega \subseteq \map {B^-_\epsilon} 0$ By Closed Ball in Euclidean Space is Compact, $\map {B^-_\epsilon} 0$ is a compact. Then $\map {B^-_\epsilon} 0$ also qualifies as a compact support of $\phi$. By definition of a test function: :$\forall y \in \R : \size y > \epsilon \implies \map \phi {x, y} = 0$ But for each $y \in \R$ we have that $\map \phi {x, y} = \map C y$ is a constant. By smoothness of test functions, this constant has to be the same for all $y \in \R$. Hence, $\map \phi {x, y} = 0$. {{qed}} \end{proof}
22613
\section{Test for Submonoid} Tags: Abstract Algebra, Monoids \begin{theorem} To show that $\struct {T, \circ}$ is a submonoid of a monoid $\struct {S, \circ}$, we need to show that: :$(1): \quad T \subseteq S$ :$(2): \quad \struct {T, \circ}$ is a magma (that is, that it is closed) :$(3): \quad \struct {T, \circ}$ has an identity. \end{theorem} \begin{proof} From Subsemigroup Closure Test, $(1)$ and $(2)$ are sufficient to show that $\struct {T, \circ}$ is a subsemigroup of $\struct {S, \circ}$. Demonstrating the presence of an identity is then sufficient to show that it is a monoid. {{qed}} Category:Monoids \end{proof}
22614
\section{Tetrahedral Number as Sum of Squares} Tags: Tetrahedral Numbers, Tetrahedral Number as Sum of Squares, Square Numbers \begin{theorem} :$H_n = \ds \sum_{k \mathop = 0}^{n / 2} \paren {n - 2 k}^2$ where $H_n$ denotes the $n$th tetrahedral number. \end{theorem} \begin{proof} Let $n$ be even such that $n = 2 m$. We have: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 0}^{n / 2} \paren {n - 2 k}^2 | r = \sum_{k \mathop = 0}^m \paren {2 m - 2 k}^2 | c = }} {{eqn | r = \sum_{k \mathop = 0}^m \paren {2 k}^2 | c = Permutation of Indices of Summation }} {{eqn | r = \frac {2 m \paren {m + 1} \paren {2 m + 1} } 3 | c = Sum of Sequence of Even Squares }} {{eqn | r = \frac {2 m \paren {2 m + 1} \paren {2 m + 2} } 6 | c = }} {{eqn | r = H_{2 m} | c = Closed Form for Tetrahedral Numbers }} {{end-eqn}} Let $n$ be odd such that $n = 2 m + 1$. We have: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 0}^{n / 2} \paren {n - 2 k}^2 | r = \sum_{k \mathop = 0}^m \paren {2 m + 1 - 2 k}^2 | c = }} {{eqn | r = \sum_{k \mathop = 0}^m \paren {2 k + 1}^2 | c = Permutation of Indices of Summation }} {{eqn | r = \frac {\paren {m + 1} \paren {2 m + 1} \paren {2 m + 3} } 3 | c = Sum of Sequence of Odd Squares: Formulation 1 }} {{eqn | r = \frac {\paren {2 m + 1} \paren {2 m + 2} \paren {2 m + 3} } 6 | c = }} {{eqn | r = H_{2 m + 1} | c = Closed Form for Tetrahedral Numbers }} {{end-eqn}} {{qed}} Category:Tetrahedral Numbers Category:Square Numbers Category:Tetrahedral Number as Sum of Squares \end{proof}
22615
\section{Thabit's Rule} Tags: Amicable Pairs, Amicable Numbers \begin{theorem} Let $n$ be a positive integer such that: {{begin-eqn}} {{eqn | l = a | r = 3 \times 2^n - 1 | c = }} {{eqn | l = b | r = 3 \times 2^{n - 1} - 1 | c = }} {{eqn | l = c | r = 9 \times 2^{2 n - 1} - 1 | c = }} {{end-eqn}} are all prime. Then: :$\tuple {2^n a b, 2^n c}$ forms an amicable pair. \end{theorem} \begin{proof} Let $r = 2^n a b, s = 2^n c$. Let $\map {\sigma_1} k$ denote the divisor sum of an integer $k$. From Divisor Sum of Power of 2: :$\map {\sigma_1} {2^n} = 2^{n + 1} - 1$ From Divisor Sum of Prime Number: {{begin-eqn}} {{eqn | l = \map {\sigma_1} a | r = 3 \times 2^n | c = }} {{eqn | l = \map {\sigma_1} b | r = 3 \times 2^{n - 1} | c = }} {{eqn | l = \map {\sigma_1} c | r = 9 \times 2^{2 n - 1} | c = }} {{end-eqn}} From Divisor Sum Function is Multiplicative: {{begin-eqn}} {{eqn | l = \map {\sigma_1} r | r = \map {\sigma_1} {2^n} \map {\sigma_1} a \map {\sigma_1} b | c = }} {{eqn | r = \paren {2^{n + 1} - 1} \paren {3 \times 2^n} \paren {3 \times 2^{n - 1} } | c = }} {{eqn | r = \paren {2^{n + 1} - 1} \paren {9 \times 2^{2 n - 1} } | c = }} {{end-eqn}} and: {{begin-eqn}} {{eqn | l = \map {\sigma_1} s | r = \map {\sigma_1} {2^n} \map {\sigma_1} c | c = }} {{eqn | r = \paren {2^{n + 1} - 1} \paren {9 \times 2^{2 n - 1} } | c = }} {{eqn | r = \map {\sigma_1} r | c = }} {{end-eqn}} Thus it is seen that: :$\map {\sigma_1} r = \map {\sigma_1} s$ Now we have: {{begin-eqn}} {{eqn | l = r + s | r = 2^n a b + 2^n c | c = }} {{eqn | r = 2^n \paren {\paren {3 \times 2^n - 1} \paren {3 \times 2^{n - 1} - 1} + 9 \times 2^{2 n - 1} - 1 } | c = }} {{eqn | r = 2^n \paren {9 \times 2^{2 n - 1} - 3 \times 2^n - 3 \times 2^{n - 1} + 1 + 9 \times 2^{2 n - 1} - 1} | c = }} {{eqn | r = 2^n \paren {2 \times 9 \times 2^{2 n - 1} - 3 \times 2 \times 2^{n - 1} - 3 \times 2^{n - 1} } | c = }} {{eqn | r = 2^{n + 1} \paren {9 \times 2^{2 n - 1} - 9 \times 2^{n - 2} } | c = }} {{eqn | r = 2^{n + 1} \paren {9 \times 2^{2 n - 1} } - 2^{n + 1} \paren {9 \times 2^{n - 2} } | c = }} {{eqn | r = 2^{n + 1} \paren {9 \times 2^{2 n - 1} } - \paren {9 \times 2^{2 n - 1} } | c = }} {{eqn | r = \paren {2^{n + 1} - 1} \paren {9 \times 2^{2 n - 1} } | c = }} {{end-eqn}} and so it is seen that: :$r + s = \map {\sigma_1} r = \map {\sigma_1} s$ Hence the result, by definition of amicable pair. {{qed}} \end{proof}
22616
\section{Thales' Theorem} Tags: Circles, Euclidean Geometry, Thales' Theorem \begin{theorem} Let $A$ and $B$ be two points on opposite ends of the diameter of a circle. Let $C$ be another point on the circle such that $C \ne A, B$. Then the lines $AC$ and $BC$ are perpendicular to each other. :400px \end{theorem} \begin{proof} :400px Let $O$ be the center of the circle, and define the vectors $\mathbf u = \overrightarrow{OC}$, $\mathbf v = \overrightarrow{OB}$ and $\mathbf w = \overrightarrow{OA}$. If $AC$ and $BC$ are perpendicular, then $\left({ \mathbf u - \mathbf w}\right) \cdot \left({\mathbf u - \mathbf v}\right) = 0$ (where $\cdot$ is the dot product). Notice that since $A$ is directly opposite $B$ in the circle, $\mathbf w = - \mathbf v$. Our expression then becomes :$\left({\mathbf u + \mathbf v}\right) \cdot \left({\mathbf u - \mathbf v}\right)$ From the distributive property of the dot product, :$\left({ \mathbf u + \mathbf v}\right) \cdot \left({\mathbf u - \mathbf v}\right) = \mathbf u \cdot \mathbf u - \mathbf u \cdot \mathbf v + \mathbf v \cdot \mathbf u - \mathbf v \cdot \mathbf v$ From the commutativity of the dot product and Dot Product of a Vector with Itself, we get :$\mathbf u \cdot \mathbf u - \mathbf u \cdot \mathbf v + \mathbf v \cdot \mathbf u - \mathbf v \cdot \mathbf v = \left|{\mathbf u}\right|^2 - \mathbf u \cdot \mathbf v + \mathbf u \cdot \mathbf v - \left|{\mathbf v}\right|^2 = \left|{\mathbf u}\right|^2 - \left|{\mathbf v}\right|^2$ Since the vectors $\mathbf u$ and $\mathbf v$ have the same length (both go from the centre of the circle to the circumference), we have that $|\mathbf u| = |\mathbf v|$, so our expression simplifies to :$\left|{\mathbf u}\right|^2 - \left|{\mathbf v}\right|^2 = \left|{\mathbf u}\right|^2 - \left|{\mathbf u}\right|^2 = 0$ The result follows. {{Qed}} \end{proof}
22617
\section{Theorem of Even Perfect Numbers/Necessary Condition} Tags: Perfect Numbers, Number Theory, Mersenne Numbers, Theorem of Even Perfect Numbers \begin{theorem} Let $a \in \N$ be an even perfect number. Then $a$ is in the form: :$2^{n - 1} \paren {2^n - 1}$ where $2^n - 1$ is prime. \end{theorem} \begin{proof} Let $a \in \N$ be an even perfect number. We can extract the highest power of $2$ out of $a$ that we can, and write $a$ in the form: :$a = m 2^{n - 1}$ where $n \ge 2$ and $m$ is odd. Since $a$ is perfect and therefore $\map {\sigma_1} a = 2 a$: {{begin-eqn}} {{eqn| l = m 2^n | r = 2 a | c = }} {{eqn| r = \map {\sigma_1} a | c = }} {{eqn| r = \map {\sigma_1} {m 2^{n - 1} } | c = }} {{eqn| r = \map {\sigma_1} m \map {\sigma_1} {2^{n - 1} } | c = Divisor Sum Function is Multiplicative }} {{eqn| r = \map {\sigma_1} m {2^n - 1} | c = Divisor Sum of Power of Prime }} {{end-eqn}} So: :$\map {\sigma_1} m = \dfrac {m 2^n} {2^n - 1}$ But $\map {\sigma_1} m$ is an integer and so $2^n - 1$ divides $m 2^n$. From Consecutive Integers are Coprime, $2^n$ and $2^n - 1$ are coprime. So from Euclid's Lemma $2^n - 1$ divides $m$. Thus $\dfrac m {2^n - 1}$ divides $m$. Since $2^n - 1 \ge 3$ it follows that: :$\dfrac m {2^n - 1} < m$ Now we can express $\map {\sigma_1} m$ as: :$\map {\sigma_1} m = \dfrac {m 2^n} {2^n - 1} = m + \dfrac m {2^n - 1}$ This means that the sum of all the divisors of $m$ is equal to $m$ itself plus one other divisor of $m$. Hence $m$ must have exactly two divisors, so it must be prime by definition. This means that the other divisor of $m$, apart from $m$ itself, must be $1$. That is: :$\dfrac m {2^n - 1} = 1$ Hence the result. {{qed}} \end{proof}
22618
\section{Theorem of Even Perfect Numbers/Sufficient Condition} Tags: Perfect Numbers, Number Theory, Mersenne Numbers, Theorem of Even Perfect Numbers \begin{theorem} Let $n \in \N$ be such that $2^n - 1$ is prime. Then $2^{n - 1} \paren {2^n - 1}$ is perfect. {{:Euclid:Proposition/IX/36}} \end{theorem} \begin{proof} Suppose $2^n - 1$ is prime. Let $a = 2^{n - 1} \paren {2^n - 1}$. Then $n \ge 2$ which means $2^{n - 1}$ is even and hence so is $a = 2^{n - 1} \paren {2^n - 1}$. Note that $2^n - 1$ is odd. Since all divisors (except $1$) of $2^{n - 1}$ are even it follows that $2^{n - 1}$ and $2^n - 1$ are coprime. Let $\map {\sigma_1} n$ be the divisor sum of $n$, that is, the sum of all divisors of $n$ (including $n$). From Divisor Sum Function is Multiplicative, it follows that $\map {\sigma_1} a = \map {\sigma_1} {2^{n - 1} } \map {\sigma_1} {2^n - 1}$. But as $2^n - 1$ is prime, $\map {\sigma_1} {2^n - 1} = 2^n$ from Divisor Sum of Prime Number. Then we have that $\map {\sigma_1} {2^{n - 1} } = 2^n - 1$ from Divisor Sum of Power of Prime. Hence it follows that $\map {\sigma_1} a = \paren {2^n - 1} 2^n = 2 a$. Hence from the definition of perfect number it follows that $2^{n - 1} \paren {2^n - 1}$ is perfect. {{qed}} {{Euclid Note|36|IX}} \end{proof}
22619
\section{Theoretical Justification for Cycle Notation} Tags: Permutation Theory \begin{theorem} Let $\N_k$ be used to denote the initial segment of natural numbers: :$\N_k = \closedint 1 k = \set {1, 2, 3, \ldots, k}$ Let $\rho: \N_n \to \N_n$ be a permutation of $n$ letters. Let $i \in \N_n$. Let $k$ be the smallest (strictly) positive integer for which $\map {\rho^k} i$ is in the set: :$\set {i, \map \rho i, \map {\rho^2} i, \ldots, \map {\rho^{k - 1} } i}$ Then: :$\map {\rho^k} i = i$ \end{theorem} \begin{proof} {{AimForCont}} $\map {\rho^k} i = \map {\rho^r} i$ for some $r > 0$. As $\rho$ has an inverse in $S_n$: :$\map {\rho^{k - r} } i = i$ This contradicts the definition of $k$, because $k - r < k$ Thus: :$r = 0$ The result follows. {{qed}} \end{proof}
22620
\section{Theories with Infinite Models have Models with Order Indiscernibles} Tags: Model Theory \begin{theorem} Let $T$ be an $\LL$-theory with infinite models. Let $\struct {I, <$ be an infinite strict linearly ordered set. There is a model $\MM \models T$ containing an order indiscernible set $\set {x_i : i \in I}$. \end{theorem} \begin{proof} We will construct the claimed model using the Compactness Theorem. Let $LL^*$ be the language obtained by adding constant symbols $c_i$ to $\LL$ for each $i \in I$. Let $T^*$ be the $\LL^*$-theory obtained by adding to $T$ the $\LL^*$-sentences: :$c_i \ne c_j$ for each $i \ne j$ in $I$ and: :$\map \phi {c_{i_1}, \dotsc, c_{i_n} } \leftrightarrow \map \phi {c_{j_1}, \dotsc, c_{j_n} }$ for each $n \in \N$, each $\LL$-formula $\phi$ with $n$ free variables, and each pair of chains $i_1 < \cdots < i_n$ and $j_1 < \cdots < j_n$ in $I$. For future reference, we will refer to these last sentences using as those which '''assert indiscernibility with respect to $\phi$'''. If we can find a model of $T^*$, then its interpretations of the constants $c_i$ for $i\in I$ will be order indiscernibles. Suppose $\Delta$ is a finite subset of $T^*$. {{refactor|Extract this section as a lemma}} :We will show that there is a model of $\Delta$ using the Infinite Ramsey's Theorem. :Let $I_\Delta$ be the finite subset of $I$ containing those $i$ for which $c_i$ occurs in some sentence in $\Delta$. :Let $\psi_1, \dots, \psi_k$ be the finitely many $\LL$-formulas in $\Delta$ which assert indiscernibility with respect to $\phi_1, \dots, \phi_k$. :Let $m$ be the maximum number of free variables in the formulas $\phi_1, \dots, \phi_k$. :Let $\MM$ be an infinite model of $T$ (which exists by assumption). :Let $X$ be any infinite subset of the universe of $\MM$ such that $X$ is linearly ordered by some relation $\prec$. :Define a partition $P$ of $X^{\paren m} = \set {X' \subseteq X: \card {X'} = m}$ into $2^k$ components $S_A$ for each $A \subseteq \set {1, 2, \dotsc, k}$ by: ::$\set {x_1, \dots, x_m} \in S_A \iff x_1 \prec \cdots \prec x_m \text { and } A = \set {i: \MM \models \map {\phi_i} {x_1, \dotsc, x_m} }$ :By Infinite Ramsey's Theorem, there is an infinite subset $Y\subseteq X$ such that each element of $Y^{\paren m} = \set {Y' \subseteq Y: \card {Y'} = m}$ is in the same component $S_A$ of $P$. :Note that $Y$ is indexed by some infinite subset $J_Y$ of $J$ and hence still linearly ordered by $\prec$. :We now show that the constants $c_i$ for $i \in I_\Delta$ can be interpreted as elements of $Y$ in $\MM$, and that the sentences $\psi_1, \dots, \psi_k$ will be satisfied using this interpretation. :Since $I_\Delta$ is finite and linearly ordered, and $J_Y$ is infinite and linearly ordered, there is clearly an increasing function $f:I_\Delta \to J_Y$. :For each $h = 1, \dotsc, k$ and any pair of chains $i_1 < \cdots < i_n$ and $j_1 < \cdots < j_n$ in $I_\Delta$, we thus have: ::$\MM \models \map {\phi_h} {y_{\map f {i_1} }, \dotsc, y_{\map f {i_n} } }$ {{iff}} $h \in A$ {{iff}} $\MM \models \map {\phi_h} {y_{\map f {j_1} }, \dotsc, y_{\map f {j_n} } }$ :and hence: ::$\MM \models \map {\phi_h} {y_{\map f {i_1} }, \dotsc, y_{\map f {i_n} } } \iff \map {\phi_h} {y_{\map f {j_1} }, \dotsc, y_{\map f {j_n} } }$ :So, if we interpret each $c_i$ for $i \in I_\Delta$ as $y_{\map f i}$, we will have $\MM \models \psi_h$ for each $h = 1, \dotsc, k$. :Thus, we have shown that $\MM$ can be extended to an $\LL^*$-structure which models $\Delta$. Hence all finite subsets $\Delta$ of $T^*$ are satisfiable. By the Compactness Theorem, we have that $T^*$ is satisfiable. Let $\NN$ be any model of $T^*$. Let $n_i$ be the interpretation of $c_i$ in $\NN$ for each $i \in I$. Then $\set {n_i: i \in I}$ is easily seen to be an order indiscernible set, since $T^*$ was defined to include sentences guaranteeing such. {{qed}} Category:Model Theory \end{proof}
22621
\section{Theory of Algebraically Closed Fields of Characteristic p is Complete} Tags: Mathematical Logic, Model Theory \begin{theorem} Let $p$ be either $0$ or a prime number. Let $ACF_p$ be the theory of algebraically closed fields of characteristic $p$ in the language $\LL_r = \set {0, 1, +, -, \cdot}$ for rings, where: :$0, 1$ are constants and: :$+, -, \cdot$ are binary functions. Then $ACF_p$ is complete. \end{theorem} \begin{proof} By the Łoś-Vaught Test, it suffices to show that $ACF_p$ is satisfiable, has no finite models, and is $\kappa$-categorical for some uncountable $\kappa$. \end{proof}
22622
\section{Theory of Set of Formulas is Theory} Tags: Formal Semantics \begin{theorem} Let $\LL$ be a logical language. Let $\mathscr M$ be a formal semantics for $\LL$. Let $\FF$ be a set of $\LL$-formulas. Let $\map T \FF$ be the $\LL$-theory of $\FF$. Then $\map T \FF$ is a theory. \end{theorem} \begin{proof} Let $\phi$ be a $\LL$-formula. Suppose that $\map T \FF \models_{\mathscr M} \phi$. By definition of $\map T \FF$: :$\FF \models_{\mathscr M} \psi$ for every $\psi \in \map T \FF$. Hence by definition of semantic consequence: :$\FF \models_{\mathscr M} \map T \FF$ By Semantic Consequence is Transitive, it follows that: :$\FF \models_{\mathscr M} \phi$ Finally, by definition of $\map T \FF$: :$\phi \in \map T \FF$. Since $\phi$ was arbitrary, the result follows. {{qed}} \end{proof}
22623
\section{There Exists No Universal Set} Tags: Naive Set Theory, There Exists No Universal Set \begin{theorem} There exists no set which is an absolutely universal set. That is: :$\map \neg {\exists \, \UU: \forall T: T \in \UU}$ where $T$ is any arbitrary object at all. That is, a set that contains ''everything'' cannot exist. \end{theorem} \begin{proof} {{AimForCont}} such a $\UU$ exists. Using the Axiom of Specification, we can create the set: :$R = \set {x \in \UU: x \notin x}$ But from Russell's Paradox, this set cannot exist. Thus: :$R \notin \UU$ and so $\UU$ cannot contain everything. {{qed}} \end{proof}
22624
\section{There are no Odd Unitary Perfect Numbers} Tags: Unitary Perfect Numbers \begin{theorem} No unitary perfect numbers exist which are odd. \end{theorem} \begin{proof} Let $n$ be an odd number with prime decomposition $n = p_1^{a_1} \cdots p_k^{a_k}$. By Sum of Unitary Divisors of Integer, the sum of its unitary divisors is $\ds \prod_{1 \mathop \le i \mathop \le k} \paren {1 + p_i^{a_i} }$. To be a unitary perfect number, this must be equal to $2 n$. Since each $p$ is odd, each $1 + p_i^{a_i}$ is even. Hence $\ds \prod_{1 \mathop \le i \mathop \le k} \paren {1 + p_i^{a_i} }$ is divisible by $2^k$. Since $n$ is odd, $2 n$ is divisible by $2$ but not $4$. Thus $k = 1$. So $n$ is a prime power. By Sum of Unitary Divisors of Power of Prime, the sum of its unitary divisors is $1 + n$. This cannot be equal to $2 n$, since $n > 1$. Therefore there are no unitary perfect numbers which are odd. {{qed}} \end{proof}
22625
\section{Third Hyperoperation is Integer Power Operation} Tags: Hyperoperation \begin{theorem} The '''$3$rd hyperoperation''' is the integer power operation restricted to the positive integers: :$\forall x, y \in \Z_{\ge 0}: H_3 \left({x, y}\right) = x^y$ \end{theorem} \begin{proof} By definition of the hyperoperation sequence: :$\forall n, x, y \in \Z_{\ge 0}: H_n \left({x, y}\right) = \begin{cases} y + 1 & : n = 0 \\ x & : n = 1, y = 0 \\ 0 & : n = 2, y = 0 \\ 1 & : n > 2, y = 0 \\ H_{n - 1} \left({x, H_n \left({x, y - 1}\right)}\right) & : n > 0, y > 0 \end{cases}$ Thus the $3$rd hyperoperation is defined as: :$\forall x, y \in \Z_{\ge 0}: H_3 \left({x, y}\right) = \begin{cases} 1 & : y = 0 \\ H_2 \left({x, H_3 \left({x, y - 1}\right)}\right) & : y > 0 \end{cases}$ From Second Hyperoperation is Multiplication Operation: :$(1): \quad \forall x, y \in \Z_{\ge 0}: H_3 \left({x, y}\right) = \begin{cases} 1 & : y = 0 \\ x \times H_3 \left({x, y - 1}\right) & : y > 0 \end{cases}$ The proof proceeds by induction. For all $y \in \Z_{\ge 0}$, let $P \left({y}\right)$ be the proposition: :$\forall x \in \Z_{\ge 0}: H_3 \left({x, y}\right) = x^y$ \end{proof}
22626
\section{Third Isomorphism Theorem/Groups} Tags: Isomorphism Theorems, Quotient Groups, Third Isomorphism Theorem, Normal Subgroups, Isomorphisms, Group Isomorphisms, Group Homomorphisms \begin{theorem} Let $G$ be a group, and let: :$H, N$ be normal subgroups of $G$ :$N$ be a subset of $H$. Then: :$(1): \quad H / N$ is a normal subgroup of $G / N$ :::where $H / N$ denotes the quotient group of $H$ by $N$ :$(2): \quad \dfrac {G / N} {H / N} \cong \dfrac G H$ :::where $\cong$ denotes group isomorphism. \end{theorem} \begin{proof} We define a mapping: :$\phi: G / N \to G / H$ by $\map \phi {g N} = g H$ Since $\phi$ is defined on cosets, we need to check that $\phi$ is well-defined. Suppose $x N = y N$. Then: :$y^{-1} x \in N$ Then: :$N \le H \implies y^{-1} x \in H$ and so: :$x H = y H$ So: :$\map \phi {x N} = \map \phi {y N}$ and $\phi$ is indeed well-defined. Now $\phi$ is a homomorphism, from: {{begin-eqn}} {{eqn | l = \map \phi {x N} \map \phi {y N} | r = \paren {x H} \paren {y H} | c = }} {{eqn | r = x y H | c = }} {{eqn | r = \map \phi {x y N} | c = }} {{eqn | r = \map \phi {x N y N} | c = }} {{end-eqn}} Also, since $N \subseteq H$, it follows that: :$\order N \le \order H$ So: :$\order {G / N} \ge \order {G / H}$, indicating $\phi$ is surjective. So: {{begin-eqn}} {{eqn | l = \map \ker \phi | r = \set {g N \in G / N: \map \phi {g N} = e_{G / H} } | c = }} {{eqn | r = \set {g N \in G / N: g H = H} | c = }} {{eqn | r = \set {g N \in G / N: g \in H} | c = }} {{eqn | r = H / N | c = }} {{end-eqn}} The result follows from the First Isomorphism Theorem. {{qed}} \end{proof}
22627
\section{Third Isomorphism Theorem/Groups/Corollary} Tags: Isomorphism Theorems, Normal Subgroups, Quotient Groups, Group Isomorphisms \begin{theorem} Let $G$ be a group. Let $N$ be a normal subgroup of $G$. Let $q: G \to \dfrac G N$ be the quotient epimorphism from $G$ to the quotient group $\dfrac G N$. Let $K$ be the kernel of $q$. Then: :$\dfrac G N \cong \dfrac {G / K} {N / K}$ \end{theorem} \begin{proof} From Kernel is Normal Subgroup of Domain we have that $K$ is a normal subgroup of $G$. Thus the Third Isomorphism Theorem for Groups can be applied directly. {{qed}} \end{proof}
22628
\section{Third Isomorphism Theorem/Rings} Tags: Isomorphism Theorems, Isomorphisms, Ring Isomorphisms, Ideal Theory, Ring Homomorphisms \begin{theorem} Let $R$ be a ring, and let: :$J, K$ be ideals of $R$ :$J$ be a subset of $K$. Then: :$(1): \quad K / J$ is an ideal of $R / J$ :where $K / J$ denotes the quotient ring of $K$ by $J$ :$(2): \quad \dfrac {R / J} {K / J} \cong \dfrac R K$ :where $\cong$ denotes ring isomorphism. This result is also referred to by some sources as the '''first isomorphism theorem''', and by others as the '''second isomorphism theorem'''. \end{theorem} \begin{proof} In Ring Homomorphism whose Kernel contains Ideal, take $\phi: R \to R / K$ to be the quotient epimorphism. Then (from the same source) its kernel is $K$. Thus we have that: :$\phi = \psi \circ \nu$ where $\psi : R / J \to R / K$ is a homomorphism. This can be illustrated by means of the following commutative diagram: ::$\begin{xy}\xymatrix@L+2mu@+1em{ R \ar[dr]_*{\phi} \ar[r]^*{\nu} & R / J \ar@{-->}[d]^*{\psi} \\ & R / K }\end{xy}$ As $\phi$ is an epimorphism then from Surjection if Composite is Surjection we have that $\psi$ is a surjection. So $\Img \psi = \Img \phi = R / K$ and the First Isomorphism Theorem applies. {{qed}} \end{proof}
22629
\section{Third Partial Derivatives of x^y} Tags: Examples of Partial Derivatives \begin{theorem} Let: :$u = x^y$ Then: :$\dfrac {\partial^3 u} {\partial x^2 \partial y} = \dfrac {\partial^3 u} {\partial x \partial y \partial x}$ \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \dfrac {\partial u} {\partial y} | r = x^y \ln x | c = Derivative of General Logarithm Function keeping $x$ constant }} {{eqn | ll= \leadsto | l = \dfrac {\partial^2 u} {\partial x \partial y} | r = \map {\dfrac \partial {\partial x} } {\dfrac {\partial u} {\partial y} } | c = {{Defof|Second Partial Derivative}} }} {{eqn | r = y x^{y - 1} \ln x + \dfrac 1 x x^y | c = Power Rule for Derivatives, Derivative of Natural Logarithm keeping $y$ constant }} {{eqn | r = x^{y - 1} \paren {y \ln x + 1} | c = rearranging }} {{eqn | ll= \leadsto | l = \dfrac {\partial^3 u} {\partial x^2 \partial y} | r = \map {\dfrac \partial {\partial x} } {\dfrac {\partial^2 u} {\partial x \partial y} } | c = {{Defof|Third Partial Derivative}} }} {{eqn | r = x^{y - 1} \map {\dfrac \partial {\partial x} } {y \ln x + 1} + \paren {y \ln x + 1} \map {\dfrac \partial {\partial x} } {x^{y - 1} } | c = Product Rule for Derivatives }} {{eqn | r = x^{y - 1} \paren {\dfrac y x} + \paren {y \ln x + 1} \paren {y - 1} x^{y - 2} | c = Derivative of Natural Logarithm, Power Rule for Derivatives keeping $y$ constant }} {{eqn | r = x^{y - 2} \paren {y + \paren {y \ln x + 1} \paren {y - 1} } | c = simplifying }} {{eqn | r = x^{y - 2} \paren {y + y^2 \ln x + y -y \ln x - 1} | c = multiplying out }} {{eqn | r = x^{y - 2} \paren {2 y + y \paren {y - 1} \ln x - 1} | c = simplifying }} {{end-eqn}} {{qed|lemma}} {{begin-eqn}} {{eqn | l = \dfrac {\partial u} {\partial x} | r = y x^{y - 1} | c = Power Rule for Derivatives keeping $y$ constant }} {{eqn | ll= \leadsto | l = \dfrac {\partial^2 u} {\partial y \partial x} | r = \map {\dfrac \partial {\partial y} } {\dfrac {\partial u} {\partial x} } | c = {{Defof|Second Partial Derivative}} }} {{eqn | r = x^{y - 1} \map {\dfrac \partial {\partial y} } y + y \map {\dfrac \partial {\partial y} } {x^{y - 1} } | c = Product Rule for Derivatives }} {{eqn | r = x^{y - 1} + y x^{y - 1} \ln x | c = Derivative of Identity Function, Derivative of General Logarithm Function keeping $x$ constant }} {{eqn | ll= \leadsto | l = \dfrac {\partial^3 u} {\partial x \partial y \partial x} | r = \map {\dfrac \partial {\partial x} } {\dfrac {\partial^2 u} {\partial y \partial x} } | c = {{Defof|Third Partial Derivative}} }} {{eqn | r = \map {\dfrac \partial {\partial x} } {x^{y - 1} + y x^{y - 1} \ln x} | c = }} {{eqn | r = \paren {y - 1} x^{y - 2} + y \paren {x^{y - 1} \map {\dfrac \partial {\partial x} } {\ln x} + \ln x \map {\dfrac \partial {\partial x} } {x^{y - 1} } } | c = Power Rule for Derivatives, Product Rule for Derivatives }} {{eqn | r = \paren {y - 1} x^{y - 2} + y x^{y - 1} \paren {\dfrac 1 x} + y \ln x \paren {y - 1} x^{y - 2} | c = Derivative of Natural Logarithm, Power Rule for Derivatives }} {{eqn | r = x^{y - 2} \paren {\paren {y - 1} + y + y \ln x \paren {y - 1} } | c = factoring }} {{eqn | r = x^{y - 2} \paren {2 y + y \paren {y - 1} \ln x - 1} | c = simplifying }} {{end-eqn}} {{qed|lemma}} The two expressions are equal. Hence the result. {{qed}} \end{proof}
22630
\section{Third Principle of Mathematical Induction} Tags: Number Theory, Proofs by Induction, Mathematical Induction, Named Theorems, Principle of Mathematical Induction, Proof Techniques \begin{theorem} Let $\map P n$ be a propositional function depending on $n \in \N$. If: :$(1): \quad \map P n$ is true for all $n \le d$ for some $d \in \N$ :$(2): \quad \forall m \in \N: \paren {\forall k \in \N, m \le k < m + d: \map P k} \implies \map P {m + d}$ then $\map P n$ is true for all $n \in \N$. \end{theorem} \begin{proof} Let $A = \set {n \in \N: \map P n}$. We show that $A$ is an inductive set. By $(1)$: :$\forall 1 \le i \le d: i \in A$ Let: :$\forall x \ge d: \set {1, 2, \dotsc, x} \subset A$ Then by definition of $A$: :$\forall k \in \N: x - \paren {d - 1} \le k < x + 1: \map P k$ Thus $\map P {x + 1} \implies x + 1 \in A$ Thus $A$ is an inductive set. Thus by the fifth axiom of Peano: :$\forall n \in \N: A = \N \implies \map P n$ {{qed}} {{Proofread}} \end{proof}
22631
\section{Third Sylow Theorem} Tags: P-Groups, Sylow Theorems, Subgroups, Third Sylow Theorem, Group Theory, Named Theorems \begin{theorem} All the Sylow $p$-subgroups of a finite group are conjugate. \end{theorem} \begin{proof} Suppose $P$ and $Q$ are Sylow $p$-subgroups of $G$. By the Second Sylow Theorem, $Q$ is a subset of a conjugate of $P$. But since $\order P = \order Q$, it follows that $Q$ must equal a conjugate of $P$. {{qed}} \end{proof}
22632
\section{Thirteen Catalan Polyhedra} Tags: Catalan Polyhedra \begin{theorem} There exist exactly $13$ distinct Catalan polyhedra: :Triakis tetrahedron :Triakis octahedron :Disdyakis dodecahedron :Tetrakis hexahedron :Triakis icosahedron :Disdyakis triacontahedron :Pentakis dodecahedron :Rhombic dodecahedron :Rhombic triacontahedron :Deltoidal icositetrahedron :Deltoidal hexecontahedron :Pentagonal icositetrahedron :Pentagonal hexecontahedron \end{theorem} \begin{proof} By definition, the Catalan polyhedra are the dual polyhedra of the Archimedean polyhedra. There are $13$ Archimedean polyhedra, and so there are $13$ Catalan polyhedra. {{ProofWanted|Include links to specific results linking the Archimedean polyhedra with their duals.}} \end{proof}
22633
\section{Three-Way Exclusive Or and Equivalence} Tags: Propositional Logic, Biconditional, Exclusive Or, Logic \begin{theorem} Let $p \iff q$ be the biconditional operator, and $p \oplus q$ be the exclusive or operator. Then: : $p \iff q \iff r \dashv \vdash p \oplus q \oplus r$ \end{theorem} \begin{proof} We apply the Method of Truth Tables to the proposition. As can be seen by inspection, in each case, the truth values under the main connectives match for all boolean interpretations. $\begin{array}{|ccccc||ccccc|} \hline (p & \iff & q) & \iff & r & (p & \oplus & q) & \oplus & r \\ \hline F & T & F & F & F & F & F & F & F & F \\ F & T & F & T & T & F & F & F & T & T \\ F & F & T & T & F & F & T & T & T & F \\ F & F & T & F & T & F & T & T & F & T \\ T & F & F & T & F & T & T & F & T & F \\ T & F & F & F & T & T & T & F & F & T \\ T & T & T & F & F & T & F & T & F & F \\ T & T & T & T & T & T & F & T & T & T \\ \hline \end{array}$ {{qed}} From Biconditional is Associative and Exclusive Or is Associative, we have that both $\iff$ and $\oplus$ are associative, which justifies the rendition of this result without parentheses. \end{proof}
22634
\section{Three Eighths as Pandigital Fraction} Tags: Pandigital Fractions \begin{theorem} $\dfrac 3 8$ cannot be expressed as a pandigital fraction. \end{theorem} \begin{proof} Can be verified by brute force. Category:Pandigital Fractions \end{proof}
22635
\section{Three Fifths as Pandigital Fraction} Tags: Pandigital Fractions \begin{theorem} $\dfrac 3 5$ cannot be expressed as a pandigital fraction. \end{theorem} \begin{proof} Can be verified by brute force. Category:Pandigital Fractions \end{proof}
22636
\section{Three Non-Collinear Planes have One Point in Common} Tags: Projective Geometry \begin{theorem} Three planes which are not collinear have exactly one point in all three planes. \end{theorem} \begin{proof} Let $A$, $B$ and $C$ be the three planes in question. From Two Planes have Line in Common, $A$ and $B$ share a line, $p$ say. From Propositions of Incidence: Plane and Line, $p$ meets $C$ in one point. {{qed}} \end{proof}
22637
\section{Three Points Describe a Circle} Tags: Circles \begin{theorem} Let $A$, $B$ and $C$ be points which are not collinear. Then there exists exactly one circle whose circumference passes through all $3$ points $A$, $B$ and $C$. \end{theorem} \begin{proof} As $A$, $B$ and $C$ are not collinear, the triangle $ABC$ can be constructed by forming the lines $AB$, $BC$ and $CA$. The result follows from Circumscribing Circle about Triangle. {{qed}} \end{proof}
22638
\section{Three Points in Ultrametric Space have Two Equal Distances} Tags: Metric Spaces \begin{theorem} Let $\struct {X, d}$ be an ultrametric space. Let $x, y, z \in X$ with $\map d {x, z} \ne \map d {y, z}$. Then: :$\map d {x, y} = \max \set {\map d {x, z}, \map d {y, z} }$ \end{theorem} \begin{proof} {{WLOG}}, let $\map d {x, z} > \map d {y, z}$. Then: {{begin-eqn}} {{eqn | l = \map d {x, y} | o = \le | r = \max \set {\map d {x, z}, \map d {y, z} } | c = {{Defof|Non-Archimedean Metric}} }} {{eqn | r = \map d {x, z} | c = since $\map d {x, z} > \map d {y, z}$ }} {{end-eqn}} On the other hand: {{begin-eqn}} {{eqn | l = \map d {x, z} | o = \le | r = \max \set {\map d {x, y}, \map d {y, z} } | c = {{Defof|Non-Archimedean Metric}} }} {{eqn | r = \map d {x, y} | c = since $\map d {x, z} > \map d {y, z}$ }} {{end-eqn}} Putting the two inequalities together then: :$\map d {x, y} = \map d {x, z} = \max \set {\map d {x, z}, \map d {y, z} }$ {{qed}} \end{proof}
22639
\section{Three Points in Ultrametric Space have Two Equal Distances/Corollary} Tags: Metric Spaces \begin{theorem} Let $\struct {X, d}$ be an ultrametric space. Let $x, y, z \in X$. Then: :at least two of the distances $\map d {x, y}$, $\map d {x, z}$ and $\map d {y, z}$ are equal. \end{theorem} \begin{proof} Either: :$\map d {x, z} = \map d {y, z}$ or: :$\map d {x, z} \ne \map d {y, z}$ By Three Points in Ultrametric Space have Two Equal Distances: :$\map d {x, z} = \map d {y, z}$ or $\map d {x, y} = \max \set {\map d {x, z}, \map d {y, z} }$ In either case two of the distances are equal. {{Qed}} \end{proof}
22640
\section{Three Points in Ultrametric Space have Two Equal Distances/Corollary 2} Tags: Normed Division Rings, Definitions: Normed Division Rings \begin{theorem} Let $\struct {R, \norm {\,\cdot\,} }$ be a normed division ring with non-Archimedean norm $\norm{\,\cdot\,}$, Let $x, y \in R$ and $\norm x \ne \norm y$. Then: :$\norm {x + y} = \norm {x - y} = \norm {y - x} = \max \set {\norm x, \norm y}$ \end{theorem} \begin{proof} Let $d$ be the metric induced by the norm $\norm {\,\cdot\,}$. By Non-Archimedean Norm iff Non-Archimedean Metric then $d$ is a non-Archimedean metric and $\struct {R, d}$ is an ultrametric space. Let $x, y \in R$ and $\norm x \ne \norm y$. By the definition of the non-Archimedean metric $d$ then: :$\norm x = \norm {x - 0} = \map d {x, 0}$ and similarly: :$\norm y = \map d {y, 0}$ By assumption then: :$\map d {x, 0} \ne \map d {y, 0}$ By Three Points in Ultrametric Space have Two Equal Distances then: :$\norm {x - y} = \map d {x, y} = \max \set {\map d {x, 0}, \map d {y, 0} } = \max \set {\norm x, \norm y}$ By Norm of Negative then: :$\norm {y - x} = \norm {x - y} = \max \set {\norm x, \norm y}$ By the definition of the non-Archimedean metric $d$ then: :$\norm y = \norm {0 - \paren {-y} } = \map d {0, -y} = \map d {-y, 0}$ By assumption then: :$\map d {x, 0} \ne \map d {-y, 0}$ By Three Points in Ultrametric Space have Two Equal Distances then: :$\map d {x, -y} = \max \set {\map d {x, 0}, \map d {-y, 0} } = \max \set {\norm x, \norm y}$ By the definition of the non-Archimedean metric $d$ then: :$\map d {x, -y} = \norm {x - \paren {-y} } = \norm {x + y}$ So $\norm {x + y} = \max \set {\norm x, \norm y}$. The result follows. {{qed}} \end{proof}
22641
\section{Three Points in Ultrametric Space have Two Equal Distances/Corollary 3} Tags: Normed Division Rings \begin{theorem} Let $\struct {R, \norm {\,\cdot\,} }$ be a normed division ring with non-Archimedean norm $\norm{\,\cdot\,}$, Let $x, y \in R$ and $\norm x \lt \norm y$. Then: :$\norm {x + y} = \norm {x - y} = \norm {y - x} = \norm y$ \end{theorem} \begin{proof} By Corollary 2 then: :$\norm {x + y} = \norm {x - y} = \norm {y - x} = \max \set {\norm x, \norm y} = \norm y$ {{qed}} \end{proof}
22642
\section{Three Points in Ultrametric Space have Two Equal Distances/Corollary 4} Tags: Normed Division Rings \begin{theorem} Let $\struct {R, \norm {\,\cdot\,} }$ be a normed division ring with non-Archimedean norm $\norm{\,\cdot\,}$, Let $x, y \in R$. Then: :* $\norm {x + y} \lt \norm y \implies \norm x = \norm y$ :* $\norm {x - y} \lt \norm y \implies \norm x = \norm y$ :* $\norm {y - x} \lt \norm y \implies \norm x = \norm y$ \end{theorem} \begin{proof} The contrapositive statements are proved. Let $\norm x \ne \norm y$ By Corollary 2 then: :$\norm {x + y} = \norm {x - y} = \norm {y - x} = \max \set {\norm x, \norm y} \ge \norm y$ The result follows. {{qed}} \end{proof}
22643
\section{Three Points in Ultrametric Space have Two Equal Distances/Corollary 5} Tags: Normed Division Rings, P-adic Number Theory \begin{theorem} Let $\norm {\, \cdot \,}$ be a non-trivial non-Archimedean norm on the rational numbers $\Q$. Let $a, b \in \Z_{\ne 0}$ be coprime, $a \perp b$ Then: :$\norm a = 1$ or $\norm b = 1$ \end{theorem} \begin{proof} By Bézout's Identity then: :$\exists n, m \in \Z : m a + n b = 1$ By Norm of Unity then: :$\norm {m a + n b} = 1$ By Corollary 5 of Characterisation of Non-Archimedean Division Ring Norms then: :$\norm a, \norm b, \norm n, \norm m \le 1$ Let $\norm a \lt 1$. By Norm axiom $(\text N 2)$: Multiplicativity: :$\norm {m a} = \norm m \norm a \lt 1$ Hence: :$\norm {m a} < \norm {m a + n b}$ By Corollary 4: :$\norm {n b} = \norm {m a + n b} = 1$ By Norm axiom $(\text N 2)$: Multiplicativity: :$\norm n \norm b = 1$. Hence $\norm b = 1$. The result follows. {{qed}} \end{proof}
22644
\section{Three Points on Sphere in Same Hemisphere} Tags: Spherical Geometry \begin{theorem} Let $S$ be a sphere. Let $A$, $B$ and $C$ be points on $S$ which do not all lie on the same great circle. Then it is possible to divide $S$ into two hemispheres such that $A$, $B$ and $C$ all lie on the same hemisphere. \end{theorem} \begin{proof} Because $A$, $B$ and $C$ do not lie on the same great circle, no two of these points are the endpoints of the same diameter of $S$. Otherwise it would be possible to construct a great circle passing through all $3$ points $A$, $B$ and $C$. Let a great circle $E$ be constructed through $A$ and $B$. Then as $C$ is not on $E$, it is on either one side or the other. Hence there is a finite spherical angle $\phi$ between $E$ and $C$. Let a diameter $PQ$ of $S$ be constructed whose endpoints are on $E$ such that: :neither $P$ nor $Q$ coincides with $A$ or $B$ :both $A$ and $B$ are on the same semicircle of $E$ into which $PQ$ divides $E$. Let two great circles $F$ and $G$ be constructed through $PQ$ which are at a spherical angle $\dfrac \phi 2$ from $E$, one in one direction and one in the other. Then $F$ and $G$ both divide $S$ into two hemispheres: :one such division has a hemisphere which contains $A$ and $B$ but does not contain $E$ :the other such division has a hemisphere which contains $A$, $B$ and $E$. Hence the result. {{qed}} \end{proof}
22645
\section{Three Quarters as Pandigital Fraction} Tags: Pandigital Fractions \begin{theorem} $\dfrac 3 4$ cannot be expressed as a pandigital fraction. \end{theorem} \begin{proof} Can be verified by brute force. Category:Pandigital Fractions \end{proof}
22646
\section{Three Regular Tessellations} Tags: Tessellations \begin{theorem} There exist exactly $3$ regular tessellations of the plane. \end{theorem} \begin{proof} Let $m$ be the number of sides of each of the regular polygons that form the regular tessellation. Let $n$ be the number of those regular polygons which meet at each vertex. From Internal Angles of Regular Polygon, the internal angles of each polygon measure $\dfrac {\paren {m - 2} 180^\circ} m$. The sum of the internal angles at a point is equal to $360^\circ$ by Sum of Angles between Straight Lines at Point form Four Right Angles So: {{begin-eqn}} {{eqn | l = n \paren {\dfrac {\paren {m - 2} 180^\circ} m} | r = 360^\circ | c = }} {{eqn | ll= \leadsto | l = \dfrac {m - 2} m | r = \dfrac 2 n | c = }} {{eqn | ll= \leadsto | l = 1 - \dfrac 2 m | r = \dfrac 2 n | c = }} {{eqn | ll= \leadsto | l = \dfrac 1 m + \dfrac 1 n | r = \dfrac 1 2 | c = }} {{end-eqn}} But $m$ and $n$ are both greater than $2$. So: :if $m = 3$, $n = 6$. :if $m = 4$, $n = 4$. :if $m = 5$, $n = \dfrac {10} 3$, which is not an integer. :if $m = 6$, $n = 3$. Now suppose $m > 6$. We have: {{begin-eqn}} {{eqn | l = \dfrac 1 m + \dfrac 1 n | r = \dfrac 1 2 | c = }} {{eqn | ll= \leadsto | l = \dfrac 1 6 + \dfrac 1 n | o = > | r = \dfrac 1 2 | c = }} {{eqn | ll= \leadsto | l = \dfrac 1 n | o = > | r = \dfrac 1 3 | c = }} {{eqn | ll= \leadsto | l = 3 | o = > | r = n | c = }} {{end-eqn}} But there are no integers between $2$ and $3$, so $m \not > 6$. There are $3$ possibilities in all. Therefore all regular tessellations have been accounted for. {{qed}} \end{proof}
22647
\section{Three Regular Tessellations/Squares} Tags: Tessellations, Squares \begin{theorem} Squares form a regular tessellation: :400px \end{theorem} \begin{proof} {{proof wanted}} Category:Tessellations Category:Squares \end{proof}
22648
\section{Three Regular Tessellations/Triangles} Tags: Equilateral Triangles, Tessellations, Triangles \begin{theorem} Equilateral triangles form a regular tessellation: :400px \end{theorem} \begin{proof} {{proof wanted}} Category:Tessellations Category:Equilateral Triangles \end{proof}
22649
\section{Three Sevenths as Pandigital Fraction} Tags: Pandigital Fractions \begin{theorem} $\dfrac 3 7$ cannot be expressed as a pandigital fraction. \end{theorem} \begin{proof} Can be verified by brute force. Category:Pandigital Fractions \end{proof}
22650
\section{Three Times Sum of Cubes of Three Indeterminates Plus 6 Times their Product} Tags: Cube Roots of Unity, Algebra \begin{theorem} :$3 \paren {a^3 + b^3 + c^3 + 6 a b c} = \paren {a + b + c}^3 + \paren {a + b \omega + c \omega^2}^3 + \paren {a + b \omega^2 + c \omega}^3$ where: : $\omega = -\dfrac 1 2 + \dfrac {\sqrt 3} 2$ \end{theorem} \begin{proof} Multiplying out: {{begin-eqn}} {{eqn | l = \paren {a + b + c}^3 | r = \paren {a + b + c} \paren {a^2 + b^2 + c^2 + 2 a b + 2 a c + 2 b c} | c = }} {{eqn | n = 1 | r = a^3 + b^3 + c^3 + 3 a^2 b + 3 a^2 c + 3 a b^2 + 3 b^2 c + 3 a c^2 + 3 b c^2 + 6 a b c | c = }} {{end-eqn}} Replacing $b$ with $b \omega$ and $c$ with $c \omega^2$ in $(1)$: {{begin-eqn}} {{eqn | l = \paren {a + b \omega + c \omega^2}^3 | r = a^3 + b^3 \omega^3 + c^3 \omega^6 + 3 a^2 b \omega + 3 a^2 c \omega^2 + 3 a b^2 \omega^2 + 3 b^2 c \omega^4 + 3 a c^2 \omega^4 + 3 b c^2 \omega^5 + 6 a b c \omega^3 | c = }} {{eqn | n = 2 | r = a^3 + b^3 + c^3 + 3 a^2 b \omega + 3 a^2 c \omega^2 + 3 a b^2 \omega^2 + 3 b^2 c \omega + 3 a c^2 \omega + 3 b c^2 \omega^2 + 6 a b c | c = as $\omega^3 = 1$ }} {{end-eqn}} Replacing $b$ with $b \omega^2$ and $c$ with $c \omega$ in $(1)$: {{begin-eqn}} {{eqn | l = \paren {a + b \omega^2 + c \omega}^3 | r = a^3 + b^3 \omega^6 + c^3 \omega^3 + 3 a^2 b \omega^2 + 3 a^2 c \omega + 3 a b^2 \omega^4 + 3 b^2 c \omega^5 + 3 a c^2 \omega^2 + 3 b c^2 \omega^4 + 6 a b c \omega^3 | c = }} {{eqn | n = 3 | r = a^3 + b^3 + c^3 + 3 a^2 b \omega^2 + 3 a^2 c \omega + 3 a b^2 \omega + 3 b^2 c \omega^2 + 3 a c^2 \omega^2 + 3 b c^2 \omega + 6 a b c | c = as $\omega^3 = 1$ }} {{end-eqn}} Adding together $(1)$, $(2)$ and $(3)$: {{begin-eqn}} {{eqn | r = \paren {a + b + c}^3 + \paren {a + b \omega + c \omega^2}^3 + \paren {a + b \omega^2 + c \omega}^3 | o = | c = }} {{eqn | r = a^3 + b^3 + c^3 + 3 a^2 b + 3 a^2 c + 3 a b^2 + 3 b^2 c + 3 a c^2 + 3 b c^2 + 6 a b c | c = }} {{eqn | o = | ro= + | r = a^3 + b^3 + c^3 + 3 a^2 b \omega + 3 a^2 c \omega^2 + 3 a b^2 \omega^2 + 3 b^2 c \omega + 3 a c^2 \omega + 3 b c^2 \omega^2 + 6 a b c | c = }} {{eqn | o = | ro= + | r = a^3 + b^3 + c^3 + 3 a^2 b \omega^2 + 3 a^2 c \omega + 3 a b^2 \omega + 3 b^2 c \omega^2 + 3 a c^2 \omega^2 + 3 b c^2 \omega + 6 a b c | c = }} {{eqn | r = 3 \paren {a^3 + b^3 + c^3 + 6 a b c} | c = }} {{eqn | o = | ro= + | r = \paren {3 a^2 b + 3 a^2 c + 3 a b^2 + 3 b^2 c + 3 a c^2 + 3 b c^2} \paren {1 + \omega + \omega^2} | c = }} {{eqn | r = 3 \paren {a^3 + b^3 + c^3 + 6 a b c} | c = Sum of Cube Roots of Unity: $1 + \omega + \omega^2 = 0$ }} {{end-eqn}} {{qed}} \end{proof}
22651
\section{Three Tri-Automorphic Numbers for each Number of Digits} Tags: Tri-Automorphic Numbers \begin{theorem} Let $d \in \Z_{>0}$ be a (strictly) positive integer. Then there exist exactly $3$ tri-automorphic numbers with exactly $d$ digits. These tri-automorphic numbers all end in $2$, $5$ or $7$. \end{theorem} \begin{proof} Let $n$ be a tri-automorphic number with $d$ digits. Let $n = 10 a + b$. Then: :$3 n^2 = 300a^2 + 60 a b + 3 b^2$ As $n$ is tri-automorphic, we have: :$(1): \quad 300 a^2 + 60 a b + 3 b^2 = 1000 z + 100 y + 10 a + b$ and: :$(2): \quad 3 b^2 - b = 10 x$ where $x$ is an integer. This condition is only satisfied by $b = 2$, $b = 5$, or $b = 7$ {{ProofWanted|Guess: Try proving for $n {{=}} 10 a + b$ and then by induction.}} Substituting $b = 2$ in equation $(1)$: :$a = 9$ Substituting $b = 5$ in equation $(1)$: :$a = 7$ Substituting $b = 7$ in equation $(1)$: :$a = 6$ {{qed}} <!-- Corollary: For di-automorphic numbers, equation (5) becomes 2b^2-b=10x and b=8 For tetra-automorphic numbers, equation (5) becomes 4b^2-b=10x and b=4 For penta-automorphic numbers, equation (5) becomes 5b^2-b=10x and b=5 --> \end{proof}
22652
\section{Three times Number whose Divisor Sum is Square} Tags: Numbers whose Sigma is Square, Three times Number whose Sigma is Square, Sigma Function, Three times Number whose Divisor Sum is Square, Numbers whose Divisor Sum is Square \begin{theorem} Let $n \in \Z_{>0}$ be a positive integer. Let the divisor sum $\map {\sigma_1} n$ of $n$ be square. Let $3$ not be a divisor of $n$. Then the divisor sum of $3 n$ is square. \end{theorem} \begin{proof} Let $\sigma \left({n}\right) = k^2$. We have from Numbers whose $\sigma$ is Square: :{{:Numbers whose Sigma is Square/Examples/3}} As $3$ is not a divisor of $n$, it follows that $3$ and $n$ are coprime. Thus: {{begin-eqn}} {{eqn | l = \sigma \left({3 n}\right) | r = \sigma \left({3 n}\right) \sigma \left({3 n}\right) | c = Sigma Function is Multiplicative }} {{eqn | r = 2^2 k^2 | c = from above }} {{eqn | r = \left({2 k}\right)^2 | c = from above }} {{end-eqn}} {{qed}} Category:Sigma Function Category:Numbers whose Sigma is Square 286280 286276 2017-02-25T08:51:28Z Prime.mover 59 286280 wikitext text/x-wiki \end{proof}
22653
\section{Time Taken for Body to Fall at Earth's Surface} Tags: Gravity, Mechanics \begin{theorem} Let an object $m$ be released above ground from a point near the Earth's surface and allowed to fall freely. Let $m$ fall a distance $s$ in time $t$. Then: :$s = \dfrac 1 2 g t^2$ or: :$t = \sqrt {\dfrac {2 s} g}$ where $g$ is the Acceleration Due to Gravity at the height through which $m$ falls. It is supposed that the distance $s$ is small enough that $g$ can be considered constant throughout. \end{theorem} \begin{proof} From Body under Constant Acceleration: Distance after Time: :$\mathbf s = \mathbf u t + \dfrac {\mathbf a t^2} 2$ Here the body falls from rest, so: :$\mathbf u = \mathbf 0$ Thus: :$\mathbf s = \dfrac {\mathbf g t^2} 2$ and so taking magnitudes: :$s = \dfrac {g t^2} 2$ It follows by multiplying by $\dfrac 2 g$ that: :$t^2 = \dfrac {2 s} g$ whence: :$t = \sqrt {\dfrac {2 s} g}$ {{qed}} \end{proof}
22654
\section{Time of Travel down Brachistochrone} Tags: Cycloids \begin{theorem} Let a wire $AB$ be curved into the shape of a brachistochrone. Let $AB$ be embedded in a constant and uniform gravitational field where Acceleration Due to Gravity is $g$. Let a bead $P$ be released at $A$ to slide down without friction to $B$. Then the time taken for $P$ to slide from $A$ to $B$ is: :$T = \pi \sqrt {\dfrac a g}$ where $a$ is the radius of the generating circle of the cycloid which forms $AB$. \end{theorem} \begin{proof} That the curve $AB$ is indeed a cycloid is demonstrated in Brachistochrone is Cycloid. Let $A$ be located at the origin of a cartesian plane. We have the equations of the cycloid: {{begin-eqn}} {{eqn | l = x | r = a \paren {\theta - \sin \theta} }} {{eqn | l = y | r = a \paren {1 - \cos \theta} }} {{end-eqn}} Differentiating with respect to theta: {{begin-eqn}} {{eqn | l = \frac {\d x} {\d \theta} | r = \paren {1 - \cos \theta} | c = }} {{eqn | l = \frac {\d y} {\d \theta} | r = a \sin \theta }} {{end-eqn}} Let $s$ be the distance along $AB$ from the origin. Then: {{begin-eqn}} {{eqn | l = \paren {\frac {\d s} {\d \theta} }^2 | r = \paren {\frac {\d x} {\d \theta} }^2 + \paren {\frac {\d y} {\d \theta} }^2 | c = }} {{eqn | r = a^2 \paren {1 - \cos \theta}^2 + a^2 \sin^2 \theta | c = }} {{eqn | r = a^2 \paren {1 - 2 \cos \theta + \cos^2 \theta + \sin^2 \theta} | c = }} {{eqn | r = 2 a^2 \paren {1 - \cos \theta} | c = }} {{eqn | r = 4 a^2 \map {\sin^2} {\frac \theta 2} | c = }} {{end-eqn}} Thus: :$\dfrac {\d s} {\d \theta} = 2 a \map \sin {\dfrac \theta 2}$ Now: {{begin-eqn}} {{eqn | l = v | r = \dfrac {\d s} {\d t} | c = }} {{eqn | ll= \leadsto | l = \int \rd t | r = t | c = }} {{eqn | r = \int \frac {\d s} v | c = }} {{end-eqn}} We also have, from the Principle of Conservation of Energy, that: {{begin-eqn}} {{eqn | l = m g y | r = \frac {m v^2} 2 | c = }} {{eqn | ll= \leadsto | l = v | r = \dfrac {\d s} {\d t} | c = }} {{eqn | r = \sqrt {2 g y} | c = }} {{eqn | r = \sqrt {2 g a \paren {1 - \cos \theta} } | c = }} {{end-eqn}} Thus: {{begin-eqn}} {{eqn | l = \frac {\d t} {\d \theta} | r = \frac {\d t} {\d s} \theta | c = }} {{eqn | r = \frac {a \sqrt {2 \paren {1 - \cos \theta} } } {\sqrt {2 g a \paren {1 - \cos \theta} } } | c = }} {{eqn | r = \sqrt {\frac a g} | c = }} {{end-eqn}} Let $T$ be the time taken to slide down the brachistochrone. At the top: :$\theta = 0$ and at the bottom: :$\theta = \pi$ Then: {{begin-eqn}} {{eqn | l = T | r = \int \rd t | c = }} {{eqn | r = \sqrt {\frac a g} \int_0^\pi \rd \theta | c = }} {{eqn | r = \pi \sqrt {\frac a g} | c = }} {{end-eqn}} So the time to slide down this brachistochrone from the top to the bottom is: :$T = \pi \sqrt {\dfrac a g}$ {{qed}} \end{proof}
22655
\section{Time of Travel down Brachistochrone/Corollary} Tags: Cycloids \begin{theorem} Let a wire $AB$ be curved into the shape of a brachistochrone. Let $AB$ be embedded in a constant and uniform gravitational field where Acceleration Due to Gravity is $g$. Let a bead $P$ be released from anywhere on the wire between $A$ and $B$ to slide down without friction to $B$. Then the time taken for $P$ to slide to $B$ is: :$T = \pi \sqrt{\dfrac a g}$ where $a$ is the radius of the generating circle of the cycloid which forms $AB$. \end{theorem} \begin{proof} That the curve $AB$ is indeed a cycloid is demonstrated in Brachistochrone is Cycloid. Let $A$ be located at the origin of a cartesian plane. Let the bead slide from an intermediate point $\theta_0$. We have: :$v = \dfrac {\d s} {\d t} = \sqrt {2 g \paren {y - y_0} }$ which leads us, via the same route as for Time of Travel down Brachistochrone, to: {{begin-eqn}} {{eqn | l = T | r = \int_{\theta_0}^\pi \sqrt {\frac {2 a^2 \paren {1 - \cos \theta} } {2 g a \paren {\cos \theta_0 - \cos \theta} } } \rd \theta | c = }} {{eqn | r = \sqrt {\frac a g} \int_{\theta_0}^\pi \sqrt {\frac {1 - \cos \theta} {\cos \theta_0 - \cos \theta} } \rd \theta | c = }} {{end-eqn}} Using the Half Angle Formula for Cosine and Half Angle Formula for Sine, this gives: :$\ds T = \sqrt {\frac a g} \int_{\theta_0}^\pi \frac {\map \sin {\theta / 2} } {\sqrt {\map \cos {\theta_0 / 2} - \map \cos {\theta / 2} } } \rd \theta$ Now we make the substitution: {{begin-eqn}} {{eqn | l = u | r = \frac {\map \cos {\theta / 2} } {\map \cos {\theta_0 / 2} } | c = }} {{eqn | ll= \leadsto | l = \frac {\d u} {\d \theta} | r = -\frac {\map \sin {\theta / 2} } {2 \map \cos {\theta_0 / 2} } | c = }} {{end-eqn}} Recalculating the limits: :when $\theta = \theta_0$ we have $u = 1$ :when $\theta = \pi$ we have $u = 0$. So: {{begin-eqn}} {{eqn | l = T | r = -2 \sqrt {\frac a g} \int_1^0 \frac {\d u} {\sqrt {1 - u^2} } | c = }} {{eqn | r = \intlimits {2 \sqrt {\frac a g} \sin^{-1} u} 0 1 | c = }} {{eqn | r = \pi \sqrt {\frac a g} | c = }} {{end-eqn}} Thus the time to slide down a brachistochrone from any arbitrary point $\theta_0$ is: :$T = \pi \sqrt {\dfrac a g}$ {{qed}} \end{proof}
22656
\section{Time when Hour Hand and Minute Hand at Right Angle} Tags: Clocks \begin{theorem} Let the time of day be such that the hour hand and minute hand are at a right angle to each other. Then the time happens $22$ times in every $12$ hour period: :when the minute hand is $15$ minutes ahead of the hour hand :when the minute hand is $15$ minutes behind of the hour hand. In the first case, this happens at $09:00$ and every $1$ hour, $5$ minutes and $27 . \dot 2 \dot 7$ seconds after In the second case, this happens at $03:00$ and every $1$ hour, $5$ minutes and $27 . \dot 2 \dot 7$ seconds after. Thus the times are, to the nearest second: $\begin {array} 09:00:00 & 03:00:00 \\ 10:05:27 & 04:05:27 \\ 11:10:55 & 05:10:55 \\ 12:16:22 & 06:16:22 \\ 13:21:49 & 07:21:49 \\ 14:27:16 & 08:27:16 \\ 15:32:44 & 09:32:44 \\ 16:38:11 & 10:38:11 \\ 17:43:38 & 11:43:38 \\ 18:49:05 & 12:49:05 \\ 19:54:33 & 13:54:33 \\ \end{array}$ and times $12$ hours different. \end{theorem} \begin{proof} Obviously the hands are at right angles at $3$ and $9$ o'clock. Thus we only need to show that the angle between the hands will be the same after every $1$ hour, $5$ minutes and $27 . \dot 2 \dot 7$ seconds. Note that: {{begin-eqn}} {{eqn | l = 1 h \ 5 m \ 27. \dot 2 \dot 7 s | r = 65 m \ 27 \tfrac {27}{99} s }} {{eqn | r = 65 m \ 27 \tfrac 3 {11} s }} {{eqn | r = 65 m \ \frac {300} {11} s }} {{eqn | r = 65 m + \frac 5 {11} m }} {{eqn | r = \frac {720} {11} m }} {{eqn | r = \frac {12} {11} h }} {{end-eqn}} In $\dfrac {12} {11}$ hours: :The minute hand has rotated $\dfrac {12} {11} \times 360^\circ$ :The hour hand has rotated $\dfrac {12} {11} \times 30^\circ$ Thus the angle between the hands has changed by: {{begin-eqn}} {{eqn | l = \frac {12} {11} \times 360^\circ - \frac {12} {11} \times 30^\circ | r = \frac {12} {11} \times 330^\circ }} {{eqn | r = 360^\circ }} {{end-eqn}} which is a full rotation. Hence after $\dfrac {12} {11}$ hours the angle between the hands would remain unchanged. {{qed}} Category:Clocks \end{proof}
22657
\section{Titanic Prime consisting of 111 Blocks of each Digit plus Zeroes} Tags: Titanic Primes \begin{theorem} The integer defined as: :$\paren {123456789}_{111} \paren 0_{2284} 1$ where $\paren a_b$ means $b$ instances of $a$ in a string, is a titanic prime. \end{theorem} \begin{proof} It is noted that it has $9 \times 111 + 2284 + 1 = 3284$ digits, making it titanic. It can be expressed arithmetically as: :$123456789 \times \dfrac {10^{999} - 1} {10^9 - 1} \times 10^{2285} + 1$ {{Alpertron-factorizer|date = $6$th March $2022$|time = $50.2$ seconds}} \end{proof}
22658
\section{Titanic Prime whose Digits are all 0 or 1} Tags: Titanic Primes \begin{theorem} The integer defined as: :$10^{641} \times \dfrac {10^{640} - 1} 9 + 1$ is a titanic prime all of whose digits are either $0$ or $1$. That is: :$\paren 1_{640} \paren 0_{640} 1$ where $\paren a_b$ means $b$ instances of $a$ in a string. \end{theorem} \begin{proof} It is clear that the digits are instances of $0$ and $1$. It is also noted that it has $640 \times 2 + 1 = 1281$ digits, making it titanic. {{Alpertron-factorizer|date = $6$th March $2022$|time = $2.5$ seconds}} \end{proof}
22659
\section{Titanic Prime whose Digits are all 9 except for one 1} Tags: Titanic Primes \begin{theorem} The integer defined as: :$2 \times 10^{3020} - 1$ is a titanic prime all of whose digits are $9$ except one, which is $1$. That is: :$1 \left({9}\right)_{3020}$ where $\left({a}\right)_b$ means $b$ instances of $a$ in a string. \end{theorem} \begin{proof} It is clear that the digits are instances of $9$ except for the first $1$. It is also noted that it has $3020 + 1 = 3021$ digits, making it titanic. {{Alpertron-factorizer|date = $6$th March $2022$|time = $25.8$ seconds}} \end{proof}
22660
\section{Titanic Prime whose Digits are all Odd} Tags: Titanic Primes \begin{theorem} The integer defined as: :$1358 \times 10^{3821} - 1$ is a titanic prime all of whose digits are odd. That is: :$1357 \paren 9_{3821}$ where $\paren a_b$ means $b$ instances of $a$ in a string. \end{theorem} \begin{proof} It is clear that the digits are all instances of $9$ except for the initial $1357$, all of which are odd. It is also noted that it has $4 + 3821 = 3825$ digits, making it titanic. {{Alpertron-factorizer|date = $6$th March $2022$|time = $34.4$ seconds}} \end{proof}
22661
\section{Titanic Prime whose Digits are all Prime} Tags: Prime Numbers, Titanic Primes \begin{theorem} The integer defined as: :$7352 \times \dfrac {10^{1104} - 1} {10^4 - 1} + 1$ is a titanic prime all of whose digits are themselves prime. That is: :$\underbrace{7352}_{275} 7353$ \end{theorem} \begin{proof} It is clear that the digits are instances of $7$, $3$, $5$ and $2$, and so are all prime. It is also noted that it has $275 \times 4 + 4 = 1104$ digits, making it titanic. {{Alpertron-factorizer|date = $6$th March $2022$|time = $2.1$ seconds}} \end{proof}
22662
\section{Titanic Sophie Germain Prime} Tags: Titanic Primes, Sophie Germain Primes \begin{theorem} The integer defined as: :$39 \, 051 \times 2^{6001} - 1$ is a titanic prime which is also a Sophie Germain prime: {{begin-eqn}} {{eqn | o = | r = 11820 \, 50794 \, 19125 \, 52383 \, 74423 \, 53078 \, 56017 \, 05024 \, 84819 \, 01689 }} {{eqn | o = | r = 74975 \, 95139 \, 68621 \, 89553 \, 48654 \, 81137 \, 72841 \, 27658 \, 52217 \, 40999 }} {{eqn | o = | r = 04778 \, 71896 \, 78015 \, 63535 \, 94741 \, 82340 \, 68638 \, 88011 \, 18130 \, 14219 }} {{eqn | o = | r = 81435 \, 50235 \, 73607 \, 51980 \, 74200 \, 04306 \, 58030 \, 53360 \, 79821 \, 16678 }} {{eqn | o = | r = 32541 \, 21729 \, 72493 \, 53731 \, 27605 \, 59447 \, 95967 \, 46064 \, 11137 \, 07858 }} {{eqn | o = | r = 37078 \, 27755 \, 33462 \, 32179 \, 66482 \, 80947 \, 33386 \, 65681 \, 87582 \, 11189 }} {{eqn | o = | r = 25630 \, 83169 \, 50526 \, 70023 \, 66301 \, 83449 \, 99960 \, 25913 \, 90035 \, 61496 }} {{eqn | o = | r = 03726 \, 62661 \, 50693 \, 56343 \, 90085 \, 30468 \, 46645 \, 69888 \, 03202 \, 50070 }} {{eqn | o = | r = 38139 \, 19172 \, 69637 \, 71838 \, 13812 \, 48256 \, 38384 \, 37787 \, 83423 \, 06357 }} {{eqn | o = | r = 09062 \, 96393 \, 13908 \, 65400 \, 30048 \, 07291 \, 64958 \, 29772 \, 97828 \, 35273 }} {{eqn | o = | r = 02603 \, 73947 \, 05739 \, 46904 \, 93564 \, 50661 \, 00172 \, 36892 \, 20285 \, 60354 }} {{eqn | o = | r = 58830 \, 25332 \, 20848 \, 80128 \, 32451 \, 94645 \, 21648 \, 78503 \, 66425 \, 73281 }} {{eqn | o = | r = 55405 \, 94426 \, 29476 \, 00573 \, 05011 \, 86259 \, 25148 \, 08537 \, 31389 \, 24832 }} {{eqn | o = | r = 90593 \, 45279 \, 70389 \, 89332 \, 87614 \, 90279 \, 77417 \, 70009 \, 37843 \, 56718 }} {{eqn | o = | r = 78965 \, 55090 \, 40413 \, 05491 \, 45610 \, 39734 \, 55313 \, 36378 \, 82326 \, 51747 }} {{eqn | o = | r = 26323 \, 96872 \, 58800 \, 36097 \, 85595 \, 50576 \, 58179 \, 78961 \, 56439 \, 38001 }} {{eqn | o = | r = 61356 \, 42993 \, 82918 \, 89157 \, 64818 \, 24068 \, 61810 \, 98754 \, 13407 \, 25598 }} {{eqn | o = | r = 81076 \, 88939 \, 65566 \, 79970 \, 94454 \, 12508 \, 20606 \, 03037 \, 82723 \, 11003 }} {{eqn | o = | r = 86445 \, 85147 \, 95431 \, 68421 \, 48123 \, 63910 \, 96321 \, 63833 \, 76594 \, 77873 }} {{eqn | o = | r = 36044 \, 25100 \, 46756 \, 76942 \, 21197 \, 98655 \, 69863 \, 08993 \, 13991 \, 54810 }} {{eqn | o = | r = 29955 \, 71299 \, 30916 \, 19908 \, 66968 \, 53268 \, 78801 \, 17165 \, 95377 \, 09390 }} {{eqn | o = | r = 12417 \, 99779 \, 38952 \, 06419 \, 62790 \, 94932 \, 21996 \, 15477 \, 09894 \, 18755 }} {{eqn | o = | r = 79741 \, 05192 \, 62661 \, 21081 \, 92384 \, 45257 \, 78675 \, 87928 \, 74768 \, 12218 }} {{eqn | o = | r = 63148 \, 68786 \, 76854 \, 53862 \, 69957 \, 63612 \, 71978 \, 31119 \, 74476 \, 86496 }} {{eqn | o = | r = 45065 \, 87748 \, 91053 \, 15072 \, 63384 \, 65410 \, 90174 \, 27502 \, 19115 \, 20006 }} {{eqn | o = | r = 99485 \, 86281 \, 23536 \, 18641 \, 48374 \, 90557 \, 49920 \, 15285 \, 92211 \, 19416 }} {{eqn | o = | r = 75209 \, 57766 \, 75409 \, 22211 \, 29543 \, 79999 \, 81129 \, 89523 \, 59262 \, 62800 }} {{eqn | o = | r = 46942 \, 15484 \, 08243 \, 63610 \, 64351 \, 53563 \, 01617 \, 42451 \, 12051 \, 59183 }} {{eqn | o = | r = 34354 \, 13049 \, 42449 \, 46301 \, 59875 \, 51181 \, 09280 \, 53716 \, 57952 \, 29658 }} {{eqn | o = | r = 01206 \, 92006 \, 20396 \, 63689 \, 45859 \, 75910 \, 58626 \, 38955 \, 88424 \, 79023 }} {{eqn | o = | r = 70325 \, 29477 \, 90965 \, 29020 \, 39505 \, 24422 \, 75678 \, 32327 \, 27410 \, 18290 }} {{eqn | o = | r = 15226 \, 89958 \, 01677 \, 48481 \, 42430 \, 49977 \, 81717 \, 47239 \, 67104 \, 08734 }} {{eqn | o = | r = 21063 \, 13953 \, 69197 \, 18416 \, 66197 \, 78782 \, 49199 \, 73757 \, 81152 \, 15777 }} {{eqn | o = | r = 88246 \, 98396 \, 88365 \, 29090 \, 59197 \, 96301 \, 79613 \, 87838 \, 71578 \, 75079 }} {{eqn | o = | r = 17192 \, 38121 \, 06694 \, 45136 \, 51899 \, 17332 \, 26537 \, 65466 \, 92624 \, 57805 }} {{eqn | o = | r = 18650 \, 91862 \, 60159 \, 38818 \, 25424 \, 40894 \, 26520 \, 87364 \, 29048 \, 52293 }} {{eqn | o = | r = 88924 \, 40043 \, 51 }} {{end-eqn}} \end{theorem} \begin{proof} At $1812$ digits, it is clear by definition that this prime is titanic. {{Alpertron-factorizer|date = $6$th March $2022$|time = $4.7$ seconds}} To show that it is in fact a Sophie Germain prime, we also need to check that: :$2 \paren {39 \, 051 \times 2^{6001} - 1} + 1 = 39 \, 051 \times 2^{6002} - 1$ is also prime. {{Alpertron-factorizer|date = $6$th March $2022$|time = $4.5$ seconds}} \end{proof}
22663
\section{Tonelli's Theorem} Tags: Measure Theory, Tonelli's Theorem \begin{theorem} Let $\struct {X, \Sigma_X, \mu}$ and $\struct {Y, \Sigma_Y, \nu}$ be $\sigma$-finite measure spaces. Let $\struct {X \times Y, \Sigma_X \otimes \Sigma_Y, \mu \times \nu}$ be the product measure space of $\struct {X, \Sigma_X, \mu}$ and $\struct {Y, \Sigma_Y, \nu}$. Let $f: X \times Y \to \overline \R_{\ge 0}$ be a positive $\Sigma_X \otimes \Sigma_Y$-measurable function. Then: :$\ds \int_{X \times Y} f \map \rd {\mu \times \nu} = \int_Y \int_X \map f {x, y} \map {\d \mu} x \map {\d \nu} y = \int_X \int_Y \map f {x, y} \map {\d \nu} y \map {\d \mu} x$ \end{theorem} \begin{proof} We rewrite the demand as: :$\ds \int_{X \times Y} f \map \rd {\mu \times \nu} = \int_Y \paren {\int_X f^y \rd \mu} \rd \nu = \int_X \paren {\int_Y f_x \rd \nu} \rd \mu$ where $f_x$ is the $x$-vertical section of $f$, and $f^y$ is the $y$-horizontal section of $f$. From Horizontal Section of Measurable Function is Measurable, we have: :$f^y$ is $\Sigma_X$-measurable. From Vertical Section of Measurable Function is Measurable, we have: :$f_x$ is $\Sigma_Y$-measurable. From Integral of Horizontal Section of Measurable Function gives Measurable Function, we then have: :$\ds y \mapsto \int_X f^y \rd \mu$ is $\Sigma_Y$-measurable and from Integral of Vertical Section of Measurable Function gives Measurable Function, we have: :$\ds x \mapsto \int_Y f_x \rd \nu$ is $\Sigma_X$-measurable. So, both: :$\ds \int_Y \paren {\int_X f^y \rd \mu} \rd \nu$ and: :$\ds \int_X \paren {\int_Y f_x \rd \nu} \rd \mu$ are well-defined. We first prove the case of: :$f = \chi_E$ where $E$ is a $\Sigma_X \otimes \Sigma_Y$-measurable set. \end{proof}
22664
\section{Top in Filter} Tags: Order Theory \begin{theorem} Let $\left({S, \preceq}\right)$ be a bounded above ordered set. Let $F$ be a filter on $S$. Then $\top \in F$ where $\top$ denotes the greatest element of $S$. \end{theorem} \begin{proof} By definition of filter in ordered set: :$F$ is non-empty and upper. By definition of non-empty set: :$\exists x: x \in F$ By definition of greatest element: :$x \preceq \top$ Thus by definition of upper set: :$\top \in F$ {{qed}} \end{proof}
22665
\section{Top in Ordered Set of Topology} Tags: Topology, Order Theory \begin{theorem} Let $T = \left({S, \tau}\right)$ be a topological space. Let $P = \left({\tau, \subseteq}\right)$ be an inclusion ordered set of $\tau$. Then $P$ is bounded above and $\top_P = S$ where $\top_P$ denotes the greatest element in $P$. \end{theorem} \begin{proof} By definition of topological space: :$S \in \tau$ By definition of subset: :$\forall A \in \tau: A \subseteq S$ Hence $P$ is bounded above. Thus by definition of the greatest element: :$\top_P = S$ {{qed}} \end{proof}
22666
\section{Top is Complete} Tags: Category Theory, Topological Spaces \begin{theorem} The category of topological spaces is complete. \end{theorem} \begin{proof} Let $\II$ be a small category. Let $D : \II \to \mathbf {Top}$ be a diagram in the category of topological spaces $\mathbf {Top}$. Let $\family {\lim D, \family {\pi_i}_{i \mathop \in \II}}$ be the limit of topological spaces of $D$. By Limit of Topological Spaces is Limit $\family {\lim D, \family {\pi_i}_{i \mathop \in \II}}$ is a categorical limit of $D$. {{qed}} Category:Topological Spaces Category:Category Theory \end{proof}
22667
\section{Top is Meet Irreducible} Tags: Order Theory, Meet Irreducible \begin{theorem} Let $\left({S, \wedge, \preceq}\right)$ be a bounded above meet semilattice. Then $\top$ is meet irreducible where $\top$ denotes the greatest element in $S$. \end{theorem} \begin{proof} Let $x, y \in S$ such that :$\top = x \wedge y$ By Meet Precedes Operands :$\top \preceq x$ and $\top \preceq y$ By definition of greatest element: :$x \preceq \top$ Thus by definition of antisymmetry: :$\top = x$ {{qed}} \end{proof}
22668
\section{Top is Prime Element} Tags: Prime Elements \begin{theorem} Let $L = \struct {S, \wedge, \preceq}$ be a bounded above meet semilattice. Then $\top$ is a prime element where $\top$ denotes the greatest element of $L$. \end{theorem} \begin{proof} Let $x, y \in S$ such that :$x \wedge y \preceq \top$ Thus by definition of greatest element: :$x \preceq \top$ or $y \preceq \top$ {{qed}} \end{proof}
22669
\section{Top is Unique} Tags: Lattice Theory \begin{theorem} Let $\struct {S, \vee, \wedge, \preceq}$ be a lattice. Then $S$ has at most one top. \end{theorem} \begin{proof} By definition, a top for $S$ is a greatest element. The result follows from Greatest Element is Unique. {{qed}} Category:Lattice Theory \end{proof}
22670
\section{Topological Closure is Closed} Tags: Topology, Set Closures, Closed Sets \begin{theorem} Let $T$ be a topological space. Let $H \subseteq T$. Then the closure $\map \cl H$ of $H$ is closed in $T$. \end{theorem} \begin{proof} From Closure of Topological Closure equals Closure: :$\map \cl {\map \cl H} = \map \cl H$ From Set is Closed iff Equals Topological Closure, it follows that $\map \cl H$ is closed. {{qed}} \end{proof}
22671
\section{Topological Closure of Singleton is Irreducible} Tags: Irreducible Spaces, Topology, Topological Closure of Singleton is Irreducible \begin{theorem} Let $T = \struct {S, \tau}$ be a topological space. Let $x$ be a point of $T$. Then: :$\set x^-$ is irreducible where $\set x^-$ denotes the topological closure of $\set x$. \end{theorem} \begin{proof} {{AimForCont}} that :$\left\{ {x}\right\}^-$ is not irreducible. By Set is Subset of its Topological Closure: :$\left\{ {x}\right\} \subseteq \left\{ {x}\right\}^-$ By definitions of singleton and Subset: :$x \in \left\{ {x}\right\}^-$ By definition of irreducible: :$\exists X_1, X_2 \subseteq S: X_1, X_2$ are closed $\land \left\{ {x}\right\}^- = X_1 \cup X_2 \land \left\{ {x}\right\}^- \ne X_1 \land \left\{ {x}\right\}^- \ne X_2$ By definition of union: :$x \in X_1$ or $x \in X_2$ By definitions of singleton and subset: :$\left\{ {x}\right\} \subseteq X_1$ or $\left\{ {x}\right\} \subseteq X_2$ By definition of Definition:Closure (Topology)/Definition 3: :$\left\{ {x}\right\}^- \subseteq X_1$ or $\left\{ {x}\right\}^- \subseteq X_2$ By Set is Subset of Union: :$X_1 \subseteq \left\{ {x}\right\}$ abd $X_2 \subseteq \left\{ {x}\right\}$ Thus by definition of set equality: :this contradicts $\left\{ {x}\right\}^- \ne X_1 \land \left\{ {x}\right\}^- \ne X_2$ {{qed}} \end{proof}
22672
\section{Topological Closure of Subset is Subset of Topological Closure} Tags: Set Closures, Topological Closure of Subset is Subset of Topological Closure, Subsets, Topology, Subset \begin{theorem} Let $T = \struct {S, \tau}$ be a topological space. Let $H \subseteq K$ and $K \subseteq S$. Then: :$\map \cl H \subseteq \map \cl K$ where $\map \cl H$ denotes the closure of $H$. \end{theorem} \begin{proof} From the definition of closure: : $\operatorname{cl}\left({H}\right)$ is the union of $H$ and its limit points. Let $x \in \operatorname{cl}\left({H}\right)$. If $x \in H$ then $x \in K \implies x \in \operatorname{cl}\left({K}\right)$. Otherwise $x$ is a limit point of $H$. That is, every open set $U$ of $T$ such that $x \in U$ contains $y \in H$ such that $y \ne x$. But as $y \in H$ it follows that $y \in K$. So every open set $U$ of $T$ such that $x \in U$ contains $y \in K$ such that $y \ne x$. This is the definition for a limit point of $K$. Thus $x \in \operatorname{cl}\left({K}\right)$. {{WIP|Extract the above paragraph into something like "Limit Point of Subset is Limit Point" (which proves that $H \subseteq K \implies H' \subseteq K'$). Then use Set Union Preserves Subsets to prove this result.}} {{qed}} \end{proof}