id
stringlengths
1
260
contents
stringlengths
1
234k
22373
\section{Summation of Products of n Numbers taken m at a time with Repetitions/Examples/Order 3} Tags: Summation to n of Summation to Index, Summation of Products of n Numbers taken m at a time with Repetitions, Summations \begin{theorem} Let $a, b \in \Z$ be integers such that $b \ge a$. Let $U$ be a set of $n = b - a + 1$ numbers $\set {x_a, x_{a + 1}, \ldots, x_b}$. Then: :$\ds \sum_{i \mathop = a}^b \sum_{j \mathop = a}^i \sum_{k \mathop = a}^j x_i x_j x_k = \dfrac { {S_1}^3} 6 + \dfrac {S_1 S_2} 2 + \dfrac {S_3} 3$ where: :$\ds S_r := \sum_{k \mathop = a}^b {x_k}^r$ \end{theorem} \begin{proof} Let: {{begin-eqn}} {{eqn | n = a | l = A | o = := | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \sum_{k \mathop = 0}^j a_i a_j a_k | c = }} {{eqn | r = \sum_{0 \mathop \le i \mathop \le j \mathop \le k \mathop \le n} a_i a_j a_k | c = }} {{eqn | n = b | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \sum_{k \mathop = j}^n a_i a_j a_k | c = }} {{eqn | n = c | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \sum_{k \mathop = i}^j a_i a_j a_k | c = }} {{eqn | n = d | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \sum_{k \mathop = i}^j a_i a_j a_k | c = }} {{eqn | n = e | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \sum_{k \mathop = i}^n a_i a_j a_k | c = }} {{end-eqn}} Also, let: {{begin-eqn}} {{eqn | l = S_1 | o = := | r = \sum_{i \mathop = 0}^n a_i | c = }} {{eqn | l = S_2 | o = := | r = \sum_{i \mathop = 0}^n {a_i}^2 | c = }} {{eqn | l = S_3 | o = := | r = \sum_{i \mathop = 0}^n {a_i}^3 | c = }} {{end-eqn}} Hence: {{begin-eqn}} {{eqn | l = 2 A | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \sum_{k \mathop = j}^n a_i a_j a_k + \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \sum_{k \mathop = i}^j a_i a_j a_k | c = $(b) + (c)$ }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \left({\sum_{k \mathop = j}^n a_i a_j a_k + \sum_{k \mathop = i}^j a_i a_j a_k}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \left({\sum_{k \mathop = i}^n a_i a_j a_k + a_i a_j a_j}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n a_i \left({\sum_{j \mathop = i}^n a_j}\right)^2 + \sum_{i \mathop = 0}^n a_i \left({\sum_{j \mathop = i}^n {a_j}^2}\right) | c = }} {{end-eqn}} Let: {{begin-eqn}} {{eqn | l = A_1 | o = := | r = \sum_{i \mathop = 0}^n a_i \left({\sum_{j \mathop = i}^n a_j}\right)^2 | c = }} {{eqn | r = \sum_{i \mathop = 0}^n a_i \sum_{j \mathop = i}^n a_j \sum_{k \mathop = i}^n a_k | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \sum_{k \mathop = i}^n a_i a_j a_k | c = }} {{eqn | l = A_3 | o = := | r = \sum_{i \mathop = 0}^n a_i \left({\sum_{j \mathop = i}^n {a_j}^2}\right) | c = }} {{end-eqn}} as calculated above. Thus: :$(1): \quad 2 A = A_1 + A_3$ Similarly: {{begin-eqn}} {{eqn | l = 2 A | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \sum_{k \mathop = 0}^j a_i a_j a_k + \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \sum_{k \mathop = j}^i a_i a_j a_k | c = $(a) + (d)$ }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \left({\sum_{k \mathop = 0}^j a_i a_j a_k + \sum_{k \mathop = j}^i a_i a_j a_k}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \left({\sum_{k \mathop = 0}^i a_i a_j a_k + a_i a_j a_j}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n a_i \left({\sum_{j \mathop = 0}^i a_j}\right)^2 + \sum_{i \mathop = 0}^n a_i \left({\sum_{j \mathop = 0}^i {a_j}^2}\right) | c = }} {{end-eqn}} Then: {{begin-eqn}} {{eqn | l = A_1 + A | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n \sum_{k \mathop = i}^n a_i a_j a_k + \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i \sum_{k \mathop = i}^n a_i a_j a_k | c = using $(e)$ }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{k \mathop = i}^n \left({\sum_{j \mathop = i}^n a_i a_j a_k + \sum_{j \mathop = 0}^i a_i a_j a_k}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{k \mathop = i}^n \left({\sum_{j \mathop = 0}^n a_i a_j a_k + {a_i}^2 a_k}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n a_i \sum_{j \mathop = 0}^n \sum_{k \mathop = j}^n a_j a_k + \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n {a_i}^2 a_j | c = }} {{end-eqn}} Let: {{begin-eqn}} {{eqn | l = A_2 | o = := | r = \sum_{i \mathop = 0}^n a_i \sum_{j \mathop = 0}^n \sum_{k \mathop = j}^n a_j a_k | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^n \sum_{k \mathop = j}^n a_i a_j a_k | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^n \sum_{k \mathop = 0}^j a_i a_j a_k | c = }} {{eqn | l = A_4 | o = := | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n {a_i}^2 a_j | c = }} {{end-eqn}} as calculated above. Thus: :$(2): \quad A_1 + A = A_2 + A_4$ Then: {{begin-eqn}} {{eqn | l = 2 A_2 | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^n \sum_{k \mathop = j}^n a_i a_j a_k + \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^n \sum_{k \mathop = 0}^j a_i a_j a_k | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^n \left({\sum_{k \mathop = j}^n a_i a_j a_k + \sum_{k \mathop = 0}^j a_i a_j a_k}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^n \left({\sum_{k \mathop = 0}^n a_i a_j a_k + a_i {a_j}^2}\right) | c = }} {{eqn | r = \left({\sum_{i \mathop = 0}^n a_i}\right)^3 + \sum_{i \mathop = 0}^n a_i \sum_{i \mathop = 0}^n {a_i}^2 | c = }} {{eqn | n = 3 | r = {S_1}^3 + S_1 S_2 | c = }} {{end-eqn}} Now we have that: {{begin-eqn}} {{eqn | l = A_3 | r = \sum_{i \mathop = 0}^n a_i \left({\sum_{j \mathop = i}^n {a_j}^2}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n {a_i}^2 \left({\sum_{j \mathop = 0}^i a_j}\right) | c = }} {{end-eqn}} and so: {{begin-eqn}} {{eqn | l = A_3 + A_4 | r = \sum_{i \mathop = 0}^n \sum_{j \mathop = 0}^i {a_i}^2 a_j + \sum_{i \mathop = 0}^n \sum_{j \mathop = i}^n {a_i}^2 a_j | c = }} {{eqn | r = \sum_{i \mathop = 0}^n \left({\sum_{j \mathop = 0}^n {a_i}^2 a_j + {a_i}^3}\right) | c = }} {{eqn | r = \sum_{i \mathop = 0}^n {a_i}^2 \sum_{i \mathop = 0}^n a_i + \sum_{i \mathop = 0}^n {a_i}^3 | c = }} {{eqn | n = 4 | r = S_2 S_1 + S_3 | c = }} {{end-eqn}} Finally: {{begin-eqn}} {{eqn | l = 2 A | r = A_1 + A_3 | c = from $(1)$ }} {{eqn | l = A + A_1 | r = A_2 + A_4 | c = from $(2)$ }} {{eqn | l = 2 A_2 | r = {S_1}^3 + S_1 S_2 | c = from $(3)$ }} {{eqn | l = A_3 + A_4 | r = S_1 S_2 + S_3 | c = from $(4)$ }} {{eqn | ll= \leadsto | l = 3 A + A_1 | r = A_1 + A_2 + A_3 + A_4 | c = }} {{eqn | ll= \leadsto | l = 3 A | r = A_2 + A_3 + A_4 | c = }} {{eqn | ll= \leadsto | l = 6 A | r = 2 A_2 + 2 \left({A_3 + A_4}\right) | c = }} {{eqn | r = {S_1}^3 + S_1 S_2 + 2 S_1 S_2 + 2 S_3 | c = }} {{eqn | r = {S_1}^3 + 3 S_1 S_2 + 2 S_3 | c = }} {{eqn | ll= \leadsto | l = A | r = \frac { {S_1}^3} 6 + \frac {S_1 S_2} 2 + \frac {S_3} 3 | c = }} {{end-eqn}} {{qed}} {{Proofread}} \end{proof}
22374
\section{Summation of Products of n Numbers taken m at a time with Repetitions/Examples/Order 4} Tags: Summation to n of Summation to Index, Summation of Products of n Numbers taken m at a time with Repetitions \begin{theorem} Let $a, b \in \Z$ be integers such that $b \ge a$. Let $U$ be a set of $n = b - a + 1$ numbers $\set {x_a, x_{a + 1}, \ldots, x_b}$. Then: :$\ds \sum_{a \mathop \le j_1 \mathop \le j_2 \mathop \le j_3 \mathop \le j_4 \mathop \le b} x_{j_1} x_{j_2} x_{j_3} x_{j_4} = \dfrac { {S_1}^4} {24} + \dfrac { {S_1}^2 S_2} 4 + \dfrac { {S_2}^2} 8 + \dfrac {S_1 S_3} 3 + \dfrac {S_4} 4$ where: :$\ds S_r := \sum_{k \mathop = a}^b {x_k}^r$. \end{theorem} \begin{proof} From Summation of Products of n Numbers taken m at a time with Repetitions: :$\ds \sum_{a \mathop \le j_1 \mathop \le \cdots \mathop \le j_m \mathop \le b} x_{j_1} \cdots x_{j_m} = \sum_{\substack {k_1, k_2, \ldots, k_m \mathop \ge 0 \\ k_1 \mathop + 2 k_2 \mathop + \cdots \mathop + m k_m \mathop = m} } \dfrac { {S_1}^{k_1} } {1^{k_1} k_1 !} \dfrac { {S_2}^{k_2} } {2^{k_2} k_2 !} \cdots \dfrac { {S_m}^{k_m} } {m^{k_m} k_m !}$ where: :$S_j = \ds \sum_{k \mathop = a}^b {x_k}^j$ for $j \in \Z_{\ge 0}$. Setting $m = 4$: {{begin-eqn}} {{eqn | l = \sum_{a \mathop \le j_1 \mathop \le j_2 \mathop \le j_3 \mathop \le j_4 \mathop \le b} x_{j_1} x_{j_2} x_{j_3} x_{j_4} | r = \sum_{\substack {k_1, k_2, k_3, k_4 \mathop \ge 0 \\ k_1 \mathop + 2 k_2 \mathop + 3 k_3 \mathop + 4 k_4 \mathop = 4} } \dfrac { {S_1}^{k_1} } {1^{k_1} k_1 !} \dfrac { {S_2}^{k_2} } {2^{k_2} k_2 !} \dfrac { {S_3}^{k_3} } {3^{k_3} k_3 !} \dfrac { {S_4}^{k_4} } {4^{k_4} k_4 !} | c = }} {{end-eqn}} We need to find all sets of $k_1, k_2, k_3, k_4 \in \Z_{\ge 0}$ such that: :$k_1 + 2 k_2 + 3 k_3 + 4 k_4 = 4$ Thus $\tuple {k_1, k_2, k_3, k_4}$ can be: :$\tuple {4, 0, 0, 0}$ :$\tuple {2, 1, 0, 0}$ :$\tuple {0, 2, 0, 0}$ :$\tuple {1, 0, 1, 0}$ :$\tuple {0, 0, 0, 1}$ Hence: {{begin-eqn}} {{eqn | l = \sum_{a \mathop \le j_1 \mathop \le j_2 \mathop \le j_3 \mathop \le b} x_{j_1} x_{j_2} x_{j_3} | r = \dfrac { {S_1}^4 {S_2}^0 {S_3}^0 {S_4}^0} {\paren {1^4 \times 4!} \paren {2^0 \times 0!} \paren {3^0 \times 0!} \paren {4^0 \times 0!} } | c = }} {{eqn | o = | ro= + | r = \dfrac { {S_1}^2 {S_2}^1 {S_3}^0 {S_4}^0} {\paren {1^2 \times 2!} \paren {2^1 \times 1!} \paren {3^0 \times 0!} \paren {4^0 \times 0!} } | c = }} {{eqn | o = | ro= + | r = \dfrac { {S_1}^0 {S_2}^2 {S_3}^0 {S_4}^0} {\paren {1^0 \times 0!} \paren {2^2 \times 2!} \paren {3^0 \times 0!} \paren {4^0 \times 0!} } | c = }} {{eqn | o = | ro= + | r = \dfrac { {S_1}^1 {S_2}^0 {S_3}^1 {S_4}^0} {\paren {1^1 \times 1!} \paren {2^0 \times 0!} \paren {3^1 \times 1!} \paren {4^0 \times 0!} } | c = }} {{eqn | o = | ro= + | r = \dfrac { {S_1}^0 {S_2}^0 {S_3}^0 {S_4}^1} {\paren {1^0 \times 0!} \paren {2^0 \times 0!} \paren {3^0 \times 0!} \paren {4^1 \times 1!} } | c = }} {{eqn | r = \dfrac { {S_1}^4} {24} + \dfrac { {S_1}^2 S_2} 4 + \dfrac { {S_2}^2} 8 + \dfrac {S_1 S_3} 3 + \dfrac {S_4} 4 | c = }} {{end-eqn}} {{qed}} \end{proof}
22375
\section{Summation of Products of n Numbers taken m at a time with Repetitions/Inverse Formula} Tags: Summation of Products of n Numbers taken m at a time with Repetitions \begin{theorem} Let $a, b \in \Z$ be integers such that $b \ge a$. Let $U$ be a set of $n = b - a + 1$ numbers $\set {x_a, x_{a + 1}, \ldots, x_b}$. Let $m \in \Z_{>0}$ be a (strictly) positive integer. Let: {{begin-eqn}} {{eqn | l = h_m | r = \sum_{a \mathop \le j_1 \mathop \le \cdots \mathop \le j_m \mathop \le b} \paren {\prod_{k \mathop = 1}^m x_{j_k} } | c = }} {{eqn | r = \sum_{a \mathop \le j_1 \mathop \le \cdots \mathop \le j_m \mathop \le b} x_{j_1} \cdots x_{j_m} | c = }} {{end-eqn}} That is, $h_m$ is the product of all $m$-tuples of elements of $U$ taken $m$ at a time. For $r \in \Z_{> 0}$, let: :$S_r = \ds \sum_{j \mathop = a}^b {x_j}^r$ Let $S_m$ be expressed in the form: :$S_m = \ds \sum_{k_1 \mathop + 2 k_2 \mathop + \mathop \cdots \mathop + m k_m \mathop = m} A_m {h_1}^{k_1} {h_2}^{k_2} \cdots {h_m}^{k_m}$ for $k_1, k_2, \ldots, k_m \ge 0$. Then : :$A_m = \paren {-1}^{k_1 + k_2 + \cdots + k_m - 1} \dfrac {m \paren {k_1 + k_2 + \cdots + k_m - 1}! } {k_1! \, k_2! \, \cdots k_m!}$ \end{theorem} \begin{proof} Let $\map G z$ be the generating function for the sequence $\sequence {h_m}$. {{begin-eqn}} {{eqn | l = \sum_{m \mathop \ge 1} \dfrac {S_m z^m} m | r = \map \ln {\map G z} | c = Summation of Products of n Numbers taken m at a time with Repetitions: Lemma 2 }} {{eqn | r = \map \ln {1 + h_1 z + h_2 z^2 + \cdots} | c = }} {{eqn | r = \sum_{k \mathop \ge 1} \dfrac {\paren {-1}^{k - 1} } k \paren {h_1 z + h_2 z^2 + \cdots}^k | c = Power Series Expansion for Logarithm of 1 + x }} {{eqn | ll= \leadsto | l = S_m | r = m \sum_{k \mathop \ge 1} \dfrac {\paren {-1}^{k - 1} } k \paren {\sum_{j \mathop = 1} h_j z^j}^k | c = }} {{end-eqn}} {{finish|Not sure where this is going}} \end{proof}
22376
\section{Summation of Products of n Numbers taken m at a time with Repetitions/Lemma 1} Tags: Summation of Products of n Numbers taken m at a time with Repetitions \begin{theorem} Let $a, b \in \Z$ be integers such that $b \ge a$. Let $U$ be a set of $n = b - a + 1$ numbers $\set {x_a, x_{a + 1}, \ldots, x_b}$. Let $m \in \Z_{>0}$ be a (strictly) positive integer. Let: {{begin-eqn}} {{eqn | l = h_m | r = \sum_{a \mathop \le j_1 \mathop \le \cdots \mathop \le j_m \mathop \le b} \paren {\prod_{k \mathop = 1}^m x_{j_k} } | c = }} {{eqn | r = \sum_{a \mathop \le j_1 \mathop \le \cdots \mathop \le j_m \mathop \le b} x_{j_1} \cdots x_{j_m} | c = }} {{end-eqn}} That is, $h_m$ is the product of all $m$-tuples of elements of $U$ taken $m$ at a time. Let $\map G z$ be the generating function for the sequence $\sequence {h_m}$. Then: {{begin-eqn}} {{eqn | l = \map G z | r = \prod_{k \mathop = a}^b \dfrac 1 {1 - x_k z} | c = }} {{eqn | r = \dfrac 1 {\paren {1 - x_a z} \paren {1 - x_{a + 1} z} \cdots \paren {1 - x_b z} } | c = }} {{end-eqn}} \end{theorem} \begin{proof} For each $k \in \set {a, a + 1, \ldots, b}$, the product of $x_k$ taken $m$ at a time is simply ${x_k}^m$. Thus for $n = 1$ we have: :$h_m = {x_k}^m$ Let the generating function for such a $\sequence {h_m}$ be $\map {G_k} z$. From Generating Function for Sequence of Powers of Constant: :$\map {G_k} z = \dfrac 1 {1 - x_k z}$ By Product of Summations, we have: :$\ds \sum_{a \mathop \le j_1 \mathop \le \cdots \mathop \le j_m \mathop \le b} x_{j_1} \cdots x_{j_m} = \prod_{k \mathop = a}^b \sum_{j \mathop = 1}^m x_j$ Hence: {{begin-eqn}} {{eqn | l = \map G z | r = \sum_{k \mathop \ge 0} h_k z^k | c = {{Defof|Generating Function}} }} {{eqn | r = \prod_{k \mathop = a}^b \dfrac 1 {1 - x_k z} | c = Product of Generating Functions: General Rule }} {{eqn | r = \dfrac 1 {\paren {1 - x_a z} \paren {1 - x_{a + 1} z} \dotsm \paren {1 - x_b z} } | c = }} {{end-eqn}} {{qed}} {{Proofread}} \end{proof}
22377
\section{Summation of Sum of Mappings on Finite Set} Tags: Summations \begin{theorem} Let $\mathbb A$ be one of the standard number systems $\N, \Z, \Q, \R, \C$. Let $S$ be a finite set. Let $f, g: S \to \mathbb A$ be mappings. Let $h = f + g$ be their sum. Then we have the equality of summations on finite sets: :$\ds \sum_{s \mathop \in S} \map h s = \sum_{s \mathop \in S} \map f s + \sum_{s \mathop \in S} \map g s$ \end{theorem} \begin{proof} Let $n$ be the cardinality of $S$. Let $\sigma: \N_{< n} \to S$ be a bijection, where $\N_{< n}$ is an initial segment of the natural numbers. By definition of summation, we have to prove the following equality of indexed summations: :$\ds \sum_{i \mathop = 0}^{n - 1} \map h {\map \sigma i} = \sum_{i \mathop = 0}^{n - 1} \map f {\map \sigma i} + \sum_{i \mathop = 0}^{n - 1} \map g {\map \sigma i}$ By Sum of Mappings Composed with Mapping, $h \circ \sigma = f \circ \sigma + g \circ \sigma$. The above equality now follows from Indexed Summation of Sum of Mappings. {{qed}} \end{proof}
22378
\section{Summation of Summation over Divisors of Function of Two Variables} Tags: Divisors, Divisibility, Summations \begin{theorem} Let $c, d, n \in \Z$. Then: :$\ds \sum_{d \mathop \divides n} \sum_{c \mathop \divides d} \map f {c, d} = \sum_{c \mathop \divides n} \sum_{d \mathop \divides \paren {n / c} } \map f {c, c d}$ where $c \divides d$ denotes that $c$ is a divisor of $d$. \end{theorem} \begin{proof} From Exchange of Order of Summation with Dependency on Both Indices: {{:Exchange of Order of Summation with Dependency on Both Indices}} We have that: :$\map R d$ is the propositional function: ::$d \divides n$ :$\map S {d, c}$ is the propositional function: ::$c \divides d$ Thus $\map {R'} {d, c}$ is the propositional function: ::Both $d \divides n$ and $c \divides d$ This is the same as: ::$c \divides n$ and $\dfrac d c \divides \dfrac n c$ Similarly, $\map {S'} c$ is the propositional function: ::$\exists d$ such that both $d \divides n$ and $c \divides d$ This is the same as: ::$c \divides n$ This gives: :$\ds \sum_{d \mathop \divides n} \sum_{c \mathop \divides d} \map f {c, d} = \sum_{c \mathop \divides n} \sum_{\paren {d / c} \mathrel \divides \paren {n / c} } \map f {c, d}$ Replacing $d / c$ with $d$: :$\ds \sum_{d \mathop \divides n} \sum_{c \mathop \divides d} \map f {c, d} = \sum_{c \mathop \divides n} \sum_{d \mathop \divides \paren {n / c} } \map f {c, c d}$ {{qed}} \end{proof}
22379
\section{Summation of Zero/Finite Set} Tags: Summations \begin{theorem} Let $\mathbb A$ be one of the standard number systems $\N, \Z, \Q, \R, \C$. Let $S$ be a finite set. Let $0 : S \to \mathbb A$ be the zero mapping. {{explain|Presumably the above is a constant mapping on $0$ -- needs to be made explicit.}} Then the summation of $0$ over $S$ equals zero: :$\ds \sum_{s \mathop \in S} 0 \left({s}\right) = 0$ \end{theorem} \begin{proof} At least three proofs are possible: :using the definition of summation and Indexed Summation of Zero :using Indexed Summation of Sum of Mappings :using Summation of Multiple of Mapping on Finite Set. {{ProofWanted}} Category:Summations \end{proof}
22380
\section{Summation of Zero/Indexed Summation} Tags: Summations \begin{theorem} Let $\mathbb A$ be one of the standard number systems $\N,\Z,\Q,\R,\C$. Let $a, b$ be integers. Let $\closedint a b$ denote the integer interval between $a$ and $b$. Let $f_0 : \closedint a b \to \mathbb A$ be the zero mapping. Then the indexed summation of $0$ from $a$ to $b$ equals zero: :$\ds \sum_{i \mathop = a}^b \map {f_0} i = 0$ \end{theorem} \begin{proof} At least three proofs are possible: * by induction, using Identity Element of Addition on Numbers * using Indexed Summation of Multiple of Mapping * using Indexed Summation of Sum of Mappings {{ProofWanted}} Category:Summations \end{proof}
22381
\section{Summation of Zero/Set} Tags: Summations \begin{theorem} Let $\mathbb A$ be one of the standard number systems $\N, \Z, \Q, \R, \C$. Let $S$ be a set. Let $0: S \to \mathbb A$ be the zero mapping. Then the summation with finite support of $0$ over $S$ equals zero: :$\ds \sum_{s \mathop \in S} \map 0 s = 0$ \end{theorem} \begin{proof} By Support of Zero Mapping, the support of $0$ is empty. By Empty Set is Finite, the support of $0$ is indeed finite. By Summation over Empty Set: :$\ds \sum_{s \mathop \in S} \map 0 s = \sum_{s \mathop \in \O} \map 0 s = 0$ {{qed}} Category:Summations \end{proof}
22382
\section{Summation of i from 1 to n of Summation of j from 1 to i} Tags: Summation of i from 1 to n of Summation of j from 1 to i, Summations \begin{theorem} :$\ds \sum_{i \mathop = 1}^n \sum_{j \mathop = 1}^i a_{i j} = \sum_{j \mathop = 1}^n \sum_{i \mathop = j}^n a_{i j}$ \end{theorem} \begin{proof} :$\displaystyle \sum_{i \mathop = 1}^n \sum_{j \mathop = 1}^i$ can be expressed as: :$\displaystyle \sum_{R \left({i}\right)} \sum_{S \left({i, j}\right)} a_{i j}$ where: :$R \left({i}\right)$ is the propositional function $1 \le i \le n$ :$S \left({i, j}\right)$ is the propositional function $1 \le j \le i$ We wish to find a propositional function $S' \left({j}\right)$ which is to be: :there exists an $i$ such that both $1 \le i \le n$ and $1 \le j \le i$ This is satisfied by the propositional function: :$S' \left({j}\right) := 1 \le j \le n$ Next we wish to find a propositional function $R' \left({i, j}\right)$ which is to be: :both $1 \le i \le n$ and $1 \le j \le i$ This is satisfied by the propositional function: :$R' \left({i, j}\right) := j \le i \le n$ Hence the result, from Exchange of Order of Summation with Dependency on Both Indices. {{qed}} \end{proof}
22383
\section{Summation over Finite Set Equals Summation over Support} Tags: Summations \begin{theorem} Let $\mathbb A$ be one of the standard number systems $\N, \Z, \Q, \R, \C$. Let $S$ be a finite set. Let $f: S \to \mathbb A$ be a mapping. Let $\map \supp f$ be its support. Then we have an equality of summations over finite sets: :$\ds \sum_{s \mathop \in S} \map f s = \sum_{s \mathop \in \map \supp f} \map f s$ \end{theorem} \begin{proof} Note that by Subset of Finite Set is Finite, $\map \supp f$ is indeed finite. The result now follows from: * Sum over Complement of Finite Set * Sum of Zero over Finite Set * Identity Element of Addition on Numbers {{qed}} \end{proof}
22384
\section{Summation over Finite Set is Well-Defined} Tags: Summations \begin{theorem} Let $\mathbb A$ be one of the standard number systems $\N, \Z, \Q, \R, \C$. Let $S$ be a finite set. Let $f: S \to \mathbb A$ be a mapping. Let $n$ be the cardinality of $S$. let $\N_{<n}$ be an initial segment of the natural numbers. Let $g, h: \N_{<n} \to S$ be bijections. Then we have an equality of indexed summations of the compositions $f \circ g$ and $f \circ h$: :$\ds \sum_{i \mathop = 0}^{n - 1} \map f {\map g i} = \sum_{i \mathop = 0}^{n - 1} \map f {\map h i}$ That is, the definition of summation over a finite set does not depend on the choice of the bijection $g: S \to \N_{< n}$. \end{theorem} \begin{proof} By Inverse of Bijection is Bijection, $h^{-1} : \N_{<n} \to S$ is a bijection. By Composite of Bijections is Bijection, the composition $h^{-1}\circ g$ is a permutation of $\N_{<n}$. By Indexed Summation does not Change under Permutation, we have an equality of indexed summations: :$\ds \sum_{i \mathop = 0}^{n - 1} \map {\paren {f \circ h} } i = \sum_{i \mathop = 0}^{n - 1} \map {\paren {f \circ h} \circ \paren {h^{-1} \circ g} } i$ By Composition of Mappings is Associative and Composite of Bijection with Inverse is Identity Mapping, the {{RHS}} equals $\ds \sum_{i \mathop = 0}^{n - 1} \map f {\map g i}$. {{qed}} \end{proof}
22385
\section{Summation over Interval equals Indexed Summation} Tags: Summations \begin{theorem} Let $\mathbb A$ be one of the standard number systems $\N, \Z, \Q, \R, \C$. Let $a, b \in \Z$ be integers. Let $\closedint a b$ be the integer interval between $a$ and $b$. Let $f: \closedint a b \to \mathbb A$ be a mapping. Then the summation over the finite set $\closedint a b$ equals the indexed summation from $a$ to $b$: :$\ds \sum_{k \mathop \in \closedint a b} \map f k = \sum_{k \mathop = a}^b \map f k$ \end{theorem} \begin{proof} By Cardinality of Integer Interval, $\closedint a b$ has cardinality $b - a + 1$. By Translation of Integer Interval is Bijection, the mapping $T : \closedint 0 {b - a} \to \closedint a b$ defined as: :$\map T k = k + a$ is a bijection. By definition of summation: :$\ds \sum_{k \mathop \in \closedint a b} \map f k = \sum_{k \mathop = 0}^{b - a} \map f {k + a}$ By Indexed Summation over Translated Interval: :$\ds \sum_{k \mathop = 0}^{b - a} \map f {k + a} = \sum_{k \mathop = a}^b \map f k$ {{qed}} Category:Summations \end{proof}
22386
\section{Summation over Lower Index of Unsigned Stirling Numbers of the First Kind} Tags: Stirling Numbers, Factorials \begin{theorem} Let $n \in \Z_{\ge 0}$ be a positive integer. Then: :$\ds \sum_k {n \brack k} = n!$ where: :$\ds {n \brack k}$ denotes an unsigned Stirling number of the first kind :$n!$ denotes the factorial of $n$. \end{theorem} \begin{proof} The proof proceeds by induction on $n$. For all $n \in \Z_{\ge 0}$, let $\map P N$ be the proposition: :$\ds \sum_k {n \brack k} = n!$ $\map P 0$ is the case: {{begin-eqn}} {{eqn | l = \sum_k {0 \brack k} | r = \sum_k \delta_{0 k} | c = Unsigned Stirling Number of the First Kind of 0 }} {{eqn | r = 1 | c = all terms vanish but for $k = 0$ }} {{eqn | r = 0! | c = {{Defof|Factorial}} }} {{end-eqn}} Thus $\map P 0$ is seen to hold. \end{proof}
22387
\section{Summation over Lower Index of Unsigned Stirling Numbers of the First Kind with Alternating Signs} Tags: Stirling Numbers, Factorials \begin{theorem} Let $n \in \Z_{\ge 0}$ be a positive integer. Then: :$\ds \sum_k \paren {-1}^k {n \brack k} = \delta_{n 0} - \delta_{n 1}$ where: :$\ds {n \brack k}$ denotes an unsigned Stirling number of the first kind :$\delta_{n 0}$ denotes the Kronecker delta. \end{theorem} \begin{proof} The proof proceeds by induction on $n$. For all $n \in \Z_{\ge 0}$, let $\map P n$ be the proposition: :$\ds \sum_k \paren {-1}^k {n \brack k} = \delta_{n 0} - \delta_{n 1}$ $\map P 0$ is the case: {{begin-eqn}} {{eqn | l = \sum_k \paren {-1}^k {0 \brack k} | r = \sum_k \delta_{0 k} | c = Unsigned Stirling Number of the First Kind of 0 }} {{eqn | r = 1 | c = all terms vanish but for $k = 0$ }} {{eqn | r = \delta_{0 0} - \delta_{0 1} | c = {{Defof|Kronecker Delta}} }} {{end-eqn}} Thus $\map P 0$ is seen to hold. $\map P 1$ is the case: {{begin-eqn}} {{eqn | l = \sum_k \paren {-1}^k {1 \brack k} | r = \sum_k \paren {-1}^k \delta_{1 k} | c = Unsigned Stirling Number of the First Kind of 1 }} {{eqn | r = -1 | c = all terms vanish but for $k = 1$ }} {{eqn | r = \delta_{1 0} - \delta_{1 1} | c = {{Defof|Kronecker Delta}} }} {{end-eqn}} Thus $\map P 1$ is seen to hold. \end{proof}
22388
\section{Summation over k of Ceiling of k over 2} Tags: Ceiling Function, Summations \begin{theorem} :$\ds \sum_{k \mathop = 1}^n \ceiling {\dfrac k 2} = \ceiling {\dfrac {n \paren {n + 2} } 4}$ \end{theorem} \begin{proof} By Permutation of Indices of Summation: :$\ds \sum_{k \mathop = 1}^n \ceiling {\dfrac k 2} = \sum_{k \mathop = 1}^n \ceiling {\dfrac {n + 1 - k} 2}$ and so: :$\ds \sum_{k \mathop = 1}^n \ceiling {\dfrac k 2} = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\ceiling {\dfrac k 2} + \ceiling {\dfrac {n + 1 - k} 2} }$ First take the case where $n$ is even. For $k$ odd: :$\ceiling {\dfrac k 2} = \dfrac k 2 + \dfrac 1 2$ and: :$\ceiling {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2$ Hence: {{begin-eqn}} {{eqn | l = \ceiling {\dfrac k 2} + \ceiling {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 + \dfrac 1 2 + \dfrac {n + 1 - k} 2 | c = }} {{eqn | r = \dfrac {k + 1 + n + 1 - k} 2 | c = }} {{eqn | r = \dfrac {n + 2} 2 | c = }} {{end-eqn}} For $k$ even: :$\ceiling {\dfrac k 2} = \dfrac k 2$ and: :$\ceiling {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2 + \dfrac 1 2 = \dfrac {n - k + 2} 2$ Hence: {{begin-eqn}} {{eqn | l = \ceiling {\dfrac k 2} + \ceiling {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 + \dfrac {n - k + 2} 2 | c = }} {{eqn | r = \dfrac {k + n - k + 2} 2 | c = }} {{eqn | r = \dfrac {n + 2} 2 | c = }} {{end-eqn}} So: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^n \ceiling {\dfrac k 2} | r = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\ceiling {\dfrac k 2} + \ceiling {\dfrac {n + 1 - k} 2} } | c = }} {{eqn | r = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\dfrac {n + 2} 2} | c = }} {{eqn | r = \dfrac 1 2 n \paren {\dfrac {n + 2} 2} | c = }} {{eqn | r = \dfrac {n \paren {n + 2} } 4 | c = }} {{eqn | r = \ceiling {\dfrac {n \paren {n + 2} } 4} | c = as $\dfrac {n \paren {n + 2} } 4$ is an integer }} {{end-eqn}} {{qed|lemma}} Next take the case where $n$ is odd. For $k$ odd: :$\ceiling {\dfrac k 2} = \dfrac k 2 + \dfrac 1 2$ and: :$\ceiling {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2 + \dfrac 1 2$ Hence: {{begin-eqn}} {{eqn | l = \ceiling {\dfrac k 2} + \ceiling {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 + \dfrac 1 2 + \dfrac {n + 1 - k} 2 + \dfrac 1 2 | c = }} {{eqn | r = \dfrac {k + 1 + n + 1 - k + 1} 2 | c = }} {{eqn | r = \dfrac {n + 3} 2 | c = }} {{end-eqn}} For $k$ even: :$\ceiling {\dfrac k 2} = \dfrac k 2$ and: :$\ceiling {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2$ Hence: {{begin-eqn}} {{eqn | l = \ceiling {\dfrac k 2} + \ceiling {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 + \dfrac {n - k + 1} 2 | c = }} {{eqn | r = \dfrac {k + n - k + 1} 2 | c = }} {{eqn | r = \dfrac {n + 1} 2 | c = }} {{end-eqn}} Let $n = 2 t + 1$. Then: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^n \ceiling {\dfrac k 2} | r = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\ceiling {\dfrac k 2} + \ceiling {\dfrac {n + 1 - k} 2} } | c = }} {{eqn | r = \dfrac 1 2 \sum_{k \mathop = 1}^{2 t + 1} \paren {\ceiling {\dfrac k 2} + \ceiling {\dfrac {2 t + 2 - k} 2} } | c = }} {{eqn | r = \dfrac t 2 \dfrac {\paren {2 t + 1} + 1} 2 + \dfrac {t + 1} 2 \dfrac {\paren {2 t + 1} + 3} 2 | c = there are $t$ even terms and $t + 1$ odd terms }} {{eqn | r = \dfrac {2 t^2 + 2 t} 4 + \dfrac {2 t^2 + 6 t + 4} 4 | c = multiplying out }} {{eqn | r = \dfrac {4 t^2 + 8 t + 3} 4 + \dfrac 1 4 | c = }} {{eqn | r = \dfrac {\paren {2 t + 1} \paren {2 t + 3} } + \dfrac 1 4 | c = }} {{eqn | r = \dfrac {n \paren {n + 2} } 4 + \dfrac 1 4 | c = }} {{eqn | r = \ceiling {\dfrac {n \paren {n + 2} } 4} | c = }} {{end-eqn}} {{qed}} \end{proof}
22389
\section{Summation over k of Floor of k over 2} Tags: Summations, Floor Function \begin{theorem} :$\ds \sum_{k \mathop = 1}^n \floor {\dfrac k 2} = \floor {\dfrac {n^2} 4}$ \end{theorem} \begin{proof} By Permutation of Indices of Summation: :$\ds \sum_{k \mathop = 1}^n \floor {\dfrac k 2} = \sum_{k \mathop = 1}^n \floor {\dfrac {n + 1 - k} 2}$ and so: :$\ds \sum_{k \mathop = 1}^n \floor {\dfrac k 2} = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\floor {\dfrac k 2} + \floor {\dfrac {n + 1 - k} 2} }$ First take the case where $n$ is even. For $k$ odd: :$\floor {\dfrac k 2} = \dfrac k 2 - \dfrac 1 2$ and: :$\floor {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2$ Hence: {{begin-eqn}} {{eqn | l = \floor {\dfrac k 2} + \floor {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 - \dfrac 1 2 + \dfrac {n + 1 - k} 2 | c = }} {{eqn | r = \dfrac {k - 1 + n + 1 - k} 2 | c = }} {{eqn | r = \dfrac n 2 | c = }} {{end-eqn}} For $k$ even: :$\floor {\dfrac k 2} = \dfrac k 2$ and: :$\floor {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2 - \dfrac 1 2 = \dfrac {n - k} 2$ Hence: {{begin-eqn}} {{eqn | l = \floor {\dfrac k 2} + \floor {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 + \dfrac {n - k} 2 | c = }} {{eqn | r = \dfrac {k + n - k} 2 | c = }} {{eqn | r = \dfrac n 2 | c = }} {{end-eqn}} So: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^n \floor {\dfrac k 2} | r = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\floor {\dfrac k 2} + \floor {\dfrac {n + 1 - k} 2} } | c = }} {{eqn | r = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\dfrac n 2} | c = }} {{eqn | r = \dfrac 1 2 n \dfrac n 2 | c = }} {{eqn | r = \dfrac {n^2} 4 | c = }} {{eqn | r = \floor {\dfrac {n^2} 4} | c = as $\dfrac {n^2} 4$ is an integer }} {{end-eqn}} {{qed|lemma}} Next take the case where $n$ is odd. For $k$ odd: :$\floor {\dfrac k 2} = \dfrac k 2 - \dfrac 1 2$ and: :$\floor {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2 - \dfrac 1 2$ Hence: {{begin-eqn}} {{eqn | l = \floor {\dfrac k 2} + \floor {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 - \dfrac 1 2 + \dfrac {n + 1 - k} 2 - \dfrac 1 2 | c = }} {{eqn | r = \dfrac {k - 1 + n + 1 - k - 1} 2 | c = }} {{eqn | r = \dfrac {n - 1} 2 | c = }} {{end-eqn}} For $k$ even: :$\floor {\dfrac k 2} = \dfrac k 2$ and: :$\floor {\dfrac {n + 1 - k} 2} = \dfrac {n + 1 - k} 2$ Hence: {{begin-eqn}} {{eqn | l = \floor {\dfrac k 2} + \floor {\dfrac {n + 1 - k} 2} | r = \dfrac k 2 + \dfrac {n - k + 1} 2 | c = }} {{eqn | r = \dfrac {k + n - k + 1} 2 | c = }} {{eqn | r = \dfrac {n + 1} 2 | c = }} {{end-eqn}} Let $n = 2 t + 1$. Then: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^n \floor {\dfrac k 2} | r = \dfrac 1 2 \sum_{k \mathop = 1}^n \paren {\floor {\dfrac k 2} + \floor {\dfrac {n + 1 - k} 2} } | c = }} {{eqn | r = \dfrac 1 2 \sum_{k \mathop = 1}^{2 t + 1} \paren {\floor {\dfrac k 2} + \floor {\dfrac {2 t + 2 - k} 2} } | c = }} {{eqn | r = \dfrac t 2 \dfrac {\paren {2 t + 1} + 1} 2 + \dfrac {t + 1} 2 \dfrac {\paren {2 t + 1} - 1} 2 | c = there are $t$ even terms and $t + 1$ odd terms }} {{eqn | r = \dfrac {2 t^2 + 2 t} 4 + \dfrac {2 t^2 + 2 t} 4 | c = multiplying out }} {{eqn | r = \dfrac {4 t^2 + 4 t} 4 + \dfrac 1 4 - \dfrac 1 4 | c = }} {{eqn | r = \dfrac {\paren {2 t + 1}^2} 4 - \dfrac 1 4 | c = }} {{eqn | r = \dfrac {n^2} 4 - \dfrac 1 4 | c = }} {{eqn | r = \floor {\dfrac {n^2} 4} | c = }} {{end-eqn}} {{qed}} \end{proof}
22390
\section{Summation over k of Floor of mk+x over n} Tags: Summations, Floor Function \begin{theorem} Let $m, n \in \Z$ such that $n > 0$. Let $x \in \R$. Then: :$\ds \sum_{0 \mathop \le k \mathop < n} \floor {\dfrac {m k + x} n} = \dfrac {\paren {m - 1} \paren {n - 1} } 2 + \dfrac {d - 1} 2 + d \floor {\dfrac x d}$ where: :$\floor x$ denotes the floor of $x$ :$d$ is the greatest common divisor of $m$ and $n$. \end{theorem} \begin{proof} By definition of modulo 1: :$\ds \sum_{0 \mathop \le k \mathop < n} \floor {\dfrac {m k + x} n} = \sum_{0 \mathop \le k \mathop < n} \dfrac {m k + x} n - \sum_{0 \mathop \le k \mathop < n} \fractpart {\dfrac {m k + x} n}$ where $\fractpart y$ in this context denotes the fractional part of $y$. First we have: {{begin-eqn}} {{eqn | l = \sum_{0 \mathop \le k \mathop < n} \dfrac {m k + x} n | r = \frac m n \sum_{0 \mathop \le k \mathop < n} k + \sum_{0 \mathop \le k \mathop < n} \dfrac x n | c = }} {{eqn | r = \frac m n \frac {n \paren {n - 1} } 2 + n \dfrac x n | c = Closed Form for Triangular Numbers }} {{eqn | r = \frac {m \paren {n - 1} } 2 + x | c = }} {{end-eqn}} Let $S$ be defined as: :$\ds S := \sum_{0 \mathop \le k \mathop < n} \fractpart {\dfrac {m k + x} n}$ Thus: :$(1): \quad \ds \sum_{0 \mathop \le k \mathop < n} \floor {\dfrac {m k + x} n} = \dfrac {m \paren {n - 1} } 2 + x - S$ Let $d = \gcd \set {m, n}$. Let: {{begin-eqn}} {{eqn | l = t | r = \frac n d }} {{eqn | l = u | r = \frac m d }} {{eqn | ll= \leadsto | l = \frac m n | r = \frac u t | c = }} {{eqn | ll= \leadsto | l = m t | r = u n | c = }} {{eqn | ll= \leadsto | l = u | r = \frac {m t} n | c = }} {{end-eqn}} We have that: {{begin-eqn}} {{eqn | l = \fractpart {\dfrac {m k + x} n} | r = \fractpart {\dfrac {m k + x} n + u} | c = {{Defof|Fractional Part}}: $u$ is an integer }} {{eqn | r = \fractpart {\dfrac {m k + x} n + \frac {m t} n} | c = }} {{eqn | r = \fractpart {\dfrac {m \paren {k + t} + x} n} | c = }} {{end-eqn}} Thus $S$ consists of $d$ copies of the same summation: {{begin-eqn}} {{eqn | l = S | r = \sum_{0 \mathop \le k \mathop < n} \fractpart {\dfrac {m k + x} n} | c = }} {{eqn | r = d \sum_{0 \mathop \le k \mathop < t} \fractpart {\dfrac {m k + x} n} | c = }} {{end-eqn}} and so: {{begin-eqn}} {{eqn | l = \sum_{0 \mathop \le k \mathop < t} \fractpart {\dfrac {m k + x} n} | r = \sum_{0 \mathop \le k \mathop < t} \fractpart {\dfrac x n + \dfrac {u k} t} | c = substituting $\dfrac u t$ for $\dfrac m n$ }} {{eqn | r = \sum_{0 \mathop \le k \mathop < t} \fractpart {\dfrac {x \bmod d} n + \dfrac k t} | c = as $t \perp u$ }} {{eqn | r = \sum_{0 \mathop \le k \mathop < t} \dfrac {x \bmod d} n + \dfrac k t | c = as $\dfrac {x \bmod d} n < \dfrac 1 t$ }} {{eqn | r = t \dfrac {x \bmod d} n + \frac 1 t \sum_{0 \mathop \le k \mathop < t} k | c = }} {{eqn | r = \dfrac {t \paren {x \bmod d} } n + \frac 1 t \frac {t \paren {t - 1} } 2 | c = Closed Form for Triangular Numbers }} {{eqn | r = \dfrac {t \paren {x \bmod d} } n + \frac {t - 1} 2 | c = }} {{eqn | ll= \leadsto | l = S | r = d \paren {\dfrac {t \paren {x \bmod d} } n + \frac {t - 1} 2} | c = }} {{eqn | r = \dfrac {n \paren {x \bmod d} } n + \frac {n - d} 2 | c = as $n = d t$ }} {{eqn | r = x \bmod d + \frac {n - d} 2 | c = }} {{end-eqn}} {{explain|Greater detail needed as to why $\ds \sum_{0 \mathop \le k \mathop < t} \fractpart {\dfrac x n + \dfrac {u k} t} {{=}} \sum_{0 \mathop \le k \mathop < t} \fractpart {\dfrac {x \bmod d} n + \dfrac k t}$}} Thus: {{begin-eqn}} {{eqn | l = \sum_{0 \mathop \le k \mathop < n} \floor {\dfrac {m k + x} n} | r = \frac {m \paren {n - 1} } 2 + x - S | c = from $(1)$ }} {{eqn | r = \frac {m \paren {n - 1} } 2 + x - d \paren {\dfrac {t \paren {x \bmod d} } n + \frac {t - 1} 2} | c = }} {{eqn | r = \frac {m \paren {n - 1} } 2 + x - x \bmod d - \frac {n - d} 2 | c = }} {{eqn | r = \frac {m \paren {n - 1} } 2 + x - x + d \floor {\frac x d} - \frac {n - 1} 2 + \frac {d - 1} 2 | c = {{Defof|Modulo Operation}} and algebra }} {{eqn | r = \frac {\paren {m - 1} \paren {n - 1} } 2 + \frac {d - 1} 2 + d \floor {\frac x d} | c = simplification }} {{end-eqn}} {{qed}} \end{proof}
22391
\section{Summation over k of Floor of x plus k over y} Tags: Summations, Floor Function \begin{theorem} Let $x, y \in \R$ such that $y > 0$. Then: :$\ds \sum_{0 \mathop \le k \mathop < y} \floor {x + \dfrac k y} = \floor {x y + \floor {x + 1} \paren {\ceiling y - y} }$ \end{theorem} \begin{proof} When $x$ increases by $1$, both sides increase by $\ceiling y$. So we can assume $0 \le x < 1$. When $x = 0$, both sides are equal to $0$. When $x$ increases past $1 - \dfrac k y$ for $0 \le k < y$, both sides increase by $1$. Hence the result. {{qed}} \end{proof}
22392
\section{Summation over k to n of Product of kth with n-kth Fibonacci Numbers} Tags: Fibonacci Numbers \begin{theorem} :$\ds \sum_{k \mathop = 0}^n F_k F_{n - k} = \dfrac {\paren {n - 1} F_n + 2n F_{n - 1} } 5$ where $F_n$ denotes the $n$th Fibonacci number. \end{theorem} \begin{proof} From Generating Function for Fibonacci Numbers, a generating function for the Fibonacci numbers is: :$\map G z = \dfrac z {1 - z - z^2}$ Then: {{begin-eqn}} {{eqn | l = \map G z | r = \dfrac z {1 - z - z^2} | c = }} {{eqn | r = \dfrac 1 {\sqrt 5} \paren {\dfrac 1 {1 - \phi z} - \dfrac 1 {1 - \hat \phi z} } | c = Partial Fraction Expansion }} {{end-eqn}} where: :$\phi = \dfrac {1 + \sqrt 5} 2$ :$\hat \phi = \dfrac {1 - \sqrt 5} 2$ Hence: {{begin-eqn}} {{eqn | l = \paren {\map G z}^2 | r = \paren {\dfrac 1 {\sqrt 5} \paren {\dfrac 1 {1 - \phi z} - \dfrac 1 {1 - \hat \phi z} } }^2 | c = }} {{eqn | r = \dfrac 1 5 \paren {\paren {\dfrac 1 {1 - \phi z} }^2 + \paren {\dfrac 1 {1 - \hat \phi z} }^2 - 2 \dfrac 1 {1 - \phi z} \dfrac 1 {1 - \hat \phi z} } | c = }} {{eqn | r = \dfrac 1 5 \paren {\dfrac 1 {\paren {1 - \phi z}^2} + \dfrac 1 {\paren {1 - \hat \phi z}^2} - 2 \dfrac {1 - \hat \phi z + 1 - \phi z} {\paren {1 - \phi z} \paren {1 - \hat \phi z} } } | c = }} {{eqn | r = \dfrac 1 5 \paren {\dfrac 1 {\paren {1 - \phi z}^2} + \dfrac 1 {\paren {1 - \hat \phi z}^2} - \dfrac 2 {\paren {1 - \phi z} \paren {1 - \paren {1 - \phi} z} } } | c = Definition of $\hat \phi$ }} {{eqn | r = \dfrac 1 5 \paren {\dfrac 1 {\paren {1 - \phi z}^2} + \dfrac 1 {\paren {1 - \hat \phi z}^2} - \dfrac 2 {1 - z - z^2} } | c = after algebra }} {{eqn | r = \dfrac 1 5 \paren {\sum_{n \mathop = 0}^\infty \paren {n + 1} \paren {\phi z}^n + \sum_{n \mathop = 0}^\infty \paren {n + 1} \paren {\hat \phi z}^n - \dfrac 2 z \map G z} | c = Power Series Expansion for Square of Reciprocal of 1-z }} {{eqn | r = \dfrac 1 5 \paren {\sum_{n \mathop = 0}^\infty \paren {n + 1} \paren {\phi^n + \hat \phi^n} z^n - \dfrac 2 z \sum_{n \mathop = 0}^\infty F_n z^n} | c = Generating Function for Fibonacci Numbers }} {{eqn | r = \dfrac 1 5 \paren {\sum_{n \mathop = 0}^\infty \paren {n + 1} \paren {\phi^n + \hat \phi^n} z^n - 2 \sum_{n \mathop = 0}^\infty F_n z^{n - 1} } | c = }} {{eqn | r = \dfrac 1 5 \paren {\sum_{n \mathop = 0}^\infty \paren {n + 1} \paren {\phi^n + \hat \phi^n} z^n - 2 \sum_{n \mathop = 1}^\infty F_{n + 1} z^n} | c = Translation of Index Variable of Summation }} {{eqn | r = \dfrac 1 5 \sum_{n \mathop = 0}^\infty \paren {\paren {n + 1} \paren {\phi^n + \hat \phi^n} - 2 F_{n + 1} } z^n | c = }} {{end-eqn}} Since the coefficient of $z^n$ in $\paren {\map G z}^2$ is $\ds \sum_{k \mathop = 0}^n F_k F_{n - k}$, by comparing coefficients: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 0}^n F_k F_{n - k} | r = \dfrac 1 5 \paren {\paren {n + 1} \paren {\phi^n + \hat \phi^n} - 2 F_{n + 1} } | c = }} {{eqn | r = \dfrac 1 5 \paren {\paren {n + 1} \paren {\dfrac {\phi^{2 n} - \hat \phi^{2 n} } {\sqrt 5} \times \dfrac {\sqrt 5} {\phi^n - \hat \phi^n} } - 2 F_{n + 1} } | c = Difference of Two Squares }} {{eqn | r = \dfrac 1 5 \paren {\paren {n + 1} \dfrac {F_{2 n} } {F_n} - 2 F_{n + 1} } | c = Euler-Binet Formula }} {{eqn | r = \dfrac 1 5 \paren {\paren {n + 1} \dfrac {F_{n - 1} F_n + F_n F_{n + 1} } {F_n} - 2 F_{n + 1} } | c = Fibonacci Number in terms of Smaller Fibonacci Numbers }} {{eqn | r = \dfrac 1 5 \paren {\paren {n + 1} \paren {F_{n - 1} + F_{n + 1} } - 2 F_{n + 1} } | c = }} {{eqn | r = \dfrac 1 5 \paren {\paren {n + 1} \paren {F_n + 2 F_{n - 1} } - 2 \paren {F_n + F_{n - 1} } } | c = {{Defof|Fibonacci Numbers}} }} {{eqn | r = \dfrac {\paren {n - 1} F_n + 2n F_{n - 1} } 5 | c = }} {{end-eqn}} {{qed}} \end{proof}
22393
\section{Summation to n of Power of k over k} Tags: Summation to n of Power of k over k, Harmonic Numbers, Binomial Coefficients \begin{theorem} :$\ds \sum_{k \mathop = 1}^n \dfrac {x^k} k = H_n + \sum_{k \mathop = 1}^n \dbinom n k \dfrac {\paren {x - 1}^k} k$ where: :$H_n$ denotes the $n$th harmonic number :$\dbinom n k$ denotes a binomial coefficient. \end{theorem} \begin{proof} The proof proceeds by induction over $n$. For all $n \in \Z_{\ge 1}$, let $P \left({n}\right)$ be the proposition: :$\displaystyle \sum_{k \mathop = 1}^n \dfrac {x^k} k = H_n + \displaystyle \sum_{k \mathop = 1}^n \dbinom n k \dfrac {\left({x - 1}\right)^k} k$ \end{proof}
22394
\section{Summation to n of Square of kth Harmonic Number} Tags: Harmonic Numbers \begin{theorem} :$\ds \sum_{k \mathop = 1}^n {H_k}^2 = \paren {n + 1} {H_n}^2 - \paren {2 n + 1} H_n + 2 n$ where $H_k$ denotes the $k$th harmonic number. \end{theorem} \begin{proof} Consider: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^{n - 1} \dfrac k {k + 1} | r = \sum_{k \mathop = 1}^{n - 1} \dfrac {\paren {k + 1} - 1} {k + 1} | c = factorizing }} {{eqn | r = \sum_{k \mathop = 1}^{n - 1} 1 - \sum_{k \mathop = 1}^{n - 1} \dfrac 1 {k + 1} | c = Sum of Summations equals Summation of Sum }} {{eqn | r = n - 1 - \sum_{k \mathop = 1}^{n - 1} \dfrac 1 {k + 1} | c = Summation of Unity over Elements }} {{eqn | r = n - 1 - \paren {H_n - 1} | c = Harmonic Number and compensate for $\dfrac 1 1$ term }} {{eqn | n = 1 | r = n - H_n | c = simplification }} {{end-eqn}} {{refactor|The above result is worth publishing on its own page as a result in its own right. I'm surprised we don't already have it.}} Using Summation by Parts: {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^n {H_k}^2 | r = H_n \sum_{k \mathop = 1}^n H_k - \sum_{k \mathop = 1}^{n - 1} \paren {\paren {\sum_{i \mathop = 1}^k {H_i} } \paren {H_{k + 1} - H_k} } | c = }} {{eqn | r = H_n \paren {\paren {n + 1} H_n - n} - \sum_{k \mathop = 1}^{n - 1} \paren {\paren {k + 1} H_k - k} \paren {H_{k + 1} - H_k} | c = Sum of Sequence of Harmonic Numbers twice }} {{eqn | r = \paren {n + 1} {H_n}^2 - n H_n - \sum_{k \mathop = 1}^{n - 1} \dfrac {\paren {k + 1} H_k - k} {k + 1} | c = $H_{k + 1} - H_k = \dfrac 1 {k+1}$ }} {{eqn | r = \paren {n + 1} {H_n}^2 - n H_n - \sum_{k \mathop = 1}^{n - 1} H_k + \sum_{k \mathop = 1}^{n - 1} \dfrac k {k + 1} | c = Sum of Summations equals Summation of Sum }} {{eqn | r = \paren {n + 1} {H_n}^2 - n H_n - \paren { \sum_{k \mathop = 1}^n H_k - H_n} + \sum_{k \mathop = 1}^{n - 1} \dfrac k {k + 1} | c = increase range of summation and compensate }} {{eqn | r = \paren {n + 1} {H_n}^2 - n H_n - \paren {\paren {n + 1} H_n - n - H_n } + \sum_{k \mathop = 1}^{n - 1} \dfrac k {k + 1} | c = Sum of Sequence of Harmonic Numbers }} {{eqn | r = \paren {n + 1} {H_n}^2 - n H_n - \paren {n + 1} H_n + n + H_n + n - H_n | c = $(1)$ }} {{eqn | r = \paren {n + 1} {H_n}^2 - \paren {2 n + 1} H_n + 2 n | c = simplification }} {{end-eqn}} {{qed}} \end{proof}
22395
\section{Summation to n of kth Harmonic Number over k} Tags: Harmonic Numbers, General Harmonic Numbers \begin{theorem} :$\ds \sum_{k \mathop = 1}^n \dfrac {H_k} k = \dfrac { {H_n}^2 + H_n^{\paren 2} } 2$ where: :$H_n$ denotes the $n$th harmonic number :$H_n^{\paren 2}$ denotes a general harmonic number. \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^n \dfrac {H_k} k | r = \sum_{k \mathop = 1}^n \sum_{j \mathop = 1}^k \dfrac 1 {j k} | c = {{Defof|Harmonic Number}} }} {{eqn | r = \sum_{k \mathop = 1}^n \sum_{j \mathop = 1}^k \dfrac 1 j \dfrac 1 k | c = }} {{eqn | r = \dfrac 1 2 \paren {\paren {\sum_{k \mathop = 1}^n \dfrac 1 k}^2 + \paren {\sum_{k \mathop = 1}^n \dfrac 1 {k^2} } } | c = Summation of Products of n Numbers taken m at a time with Repetitions }} {{eqn | r = \dfrac { {H_n}^2 + H_n^{\paren 2} } 2 | c = {{Defof|Harmonic Number}} and {{Defof|General Harmonic Numbers}} }} {{end-eqn}} Hence the result. {{qed}} \end{proof}
22396
\section{Summation to n of kth Harmonic Number over k+1} Tags: Harmonic Numbers, General Harmonic Numbers \begin{theorem} :$\ds \sum_{k \mathop = 1}^n \dfrac {H_k} {k + 1} = \dfrac { {H_{n + 1} }^2 - H_{n + 1}^{\paren 2} } 2$ where: :$H_n$ denotes the $n$th harmonic number :$H_n^{\paren 2}$ denotes a general harmonic number. \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \sum_{k \mathop = 1}^n \dfrac {H_k} {k + 1} | r = \sum_{k \mathop = 1}^n \paren {\dfrac {H_{k + 1} } {k + 1} - \dfrac 1 {\paren {k + 1} \paren {k + 1} } } | c = {{Defof|Harmonic Number}} }} {{eqn | r = \sum_{k \mathop = 1}^n \dfrac {H_{k + 1} } {k + 1} - \sum_{k \mathop = 1}^n \dfrac 1 {\paren {k + 1}^2} | c = }} {{eqn | r = \sum_{k \mathop = 2}^{n + 1} \dfrac {H_k} k - \sum_{k \mathop = 2}^{n + 1} \dfrac 1 {k^2} | c = Translation of Index Variable of Summation }} {{eqn | r = \sum_{k \mathop = 1}^{n + 1} \dfrac {H_k} k - \dfrac {H_1} 1 - \paren{\sum_{k \mathop = 1}^{n + 1} \dfrac 1 {k^2} - \dfrac 1 {1^2} } | c = }} {{eqn | r = \sum_{k \mathop = 1}^{n + 1} \dfrac {H_k} k - 1 - \paren{\sum_{k \mathop = 1}^{n + 1} \dfrac 1 {k^2} - 1} | c = Harmonic Number $H_1$ }} {{eqn | r = \sum_{k \mathop = 1}^{n + 1} \dfrac {H_k} k - \sum_{k \mathop = 1}^{n + 1} \dfrac 1 {k^2} | c = simplifying }} {{eqn | r = \sum_{k \mathop = 1}^{n + 1} \dfrac {H_k} k - H_{n + 1}^{\paren 2} | c = {{Defof|General Harmonic Numbers}} }} {{eqn | r = \dfrac { {H_{n + 1} }^2 + H_{n + 1}^{\paren 2} } 2 - H_{n + 1}^{\paren 2} | c = Summation to n of kth Harmonic Number over k }} {{eqn | r = \dfrac { {H_{n + 1} }^2 + H_{n + 1}^{\paren 2} - 2 H_{n + 1}^{\paren 2} } 2 | c = }} {{eqn | r = \dfrac { {H_{n + 1} }^2 - H_{n + 1}^{\paren 2} } 2 | c = }} {{end-eqn}} Hence the result. {{qed}} \end{proof}
22397
\section{Sums of Consecutive Sequences of Squares that equal Squares} Tags: Sums of Sequences, Sum of Sequence of Squares, Square Numbers \begin{theorem} The $24$th square pyramidal number is the only one which is square: :$1^2 + 2^2 + 3^2 + \cdots + 24^2 = 70^2$ while there are several Sum of Sequence of Squares which are square, for example: :$18^2 + 19^2 + \cdots + 28^2 = 77^2$ and: :$25^2 + 26^2 + \cdots + 624^2 = 9010^2$ \end{theorem} \begin{proof} We have: {{begin-eqn}} {{eqn | l = 1^2 + 2^2 + 3^2 + \cdots + 24^2 | r = \dfrac {24 \times \paren {24 + 1} \times \paren {2 \times 24 + 1} } 6 | c = Sum of Sequence of Squares }} {{eqn | r = \dfrac {24 \times 25 \times 49} 6 | c = }} {{eqn | r = \dfrac {2^3 \times 3 \times 5^2 \times 7^2} {2 \times 3} | c = }} {{eqn | r = 2^2 \times 5^2 \times 7^2 | c = }} {{eqn | r = \paren {2 \times 5 \times 7}^2 | c = }} {{eqn | r = 70^2 | c = }} {{end-eqn}} and: {{begin-eqn}} {{eqn | l = 18^2 + 19^2 + \cdots + 28^2 | r = \dfrac {28 \times \paren {28 + 1} \times \paren {2 \times 28 + 1} } 6 - \dfrac {17 \times \paren {17 + 1} \times \paren {2 \times 17 + 1} } 6 | c = Sum of Sequence of Squares }} {{eqn | r = \dfrac {28 \times 29 \times 57 - 17 \times 18 \times 35} 6 | c = }} {{eqn | r = \dfrac {\paren {2^2 \times 7} \times 29 \times \paren {3 \times 19} - 17 \times \paren {2 \times 3^2} \times \paren {5 \times 7} } {2 \times 3} | c = }} {{eqn | r = 2 \times 7 \times 29 \times 19 - 17 \times 3 \times 5 \times 7 | c = }} {{eqn | r = 7 \times \paren {1102 - 255} | c = }} {{eqn | r = 7 \times 847 | c = }} {{eqn | r = 7 \times 7 \times 11^2 | c = }} {{eqn | r = 77^2 | c = }} {{end-eqn}} and: {{begin-eqn}} {{eqn | l = 25^2 + 26^2 + \cdots + 624^2 | r = \dfrac {624 \times \paren {624 + 1} \times \paren {2 \times 624 + 1} } 6 - \dfrac {24 \times \paren {24 + 1} \times \paren {2 \times 24 + 1} } 6 | c = Sum of Sequence of Squares }} {{eqn | r = \dfrac {624 \times 625 \times 1249 - 24 \times 25 \times 49} 6 | c = }} {{eqn | r = \dfrac {\paren {2^4 \times 3 \times 13} \times 5^4 \times 1249 - \paren {2^3 \times 3} \times 5^2 \times 7^2} {2 \times 3} | c = }} {{eqn | r = 2^3 \times 5^4 \times 13 \times 1249 - 2^2 \times 5^2 \times 7^2 | c = }} {{eqn | r = 2^2 \times 5^2 \times \paren {2 \times 5^2 \times 13 \times 1249 - 7^2} | c = }} {{eqn | r = 2^2 \times 5^2 \times \paren {811 \, 850 - 49} | c = }} {{eqn | r = 2^2 \times 5^2 \times \paren {811 \, 801} | c = }} {{eqn | r = 2^2 \times 5^2 \times \paren {17^2 \times 53^2} | c = }} {{eqn | r = 9010^2 | c = }} {{end-eqn}} {{qed}} \end{proof}
22398
\section{Sums of Partial Sequences of Squares} Tags: Sums of Sequences, Sums of Partial Sequences of Squares, Square Numbers \begin{theorem} Let $n \in \Z_{>0}$. Consider the odd number $2 n + 1$ and its square $\paren {2 n + 1}^2 = 2 m + 1$. Then: :$\ds \sum_{j \mathop = 0}^n \paren {m - j}^2 = \sum_{j \mathop = 1}^n \paren {m + j}^2$ That is: :the sum of the squares of the $n + 1$ integers up to $m$ equals: :the sum of the squares of the $n$ integers from $m + 1$ upwards. \end{theorem} \begin{proof} The proof proceeds by induction. For all $n \in \Z_{> 0}$, let $\map P n$ be the proposition: :$\ds \sum_{j \mathop = 0}^n \paren {m - j}^2 = \sum_{j \mathop = 1}^n \paren {m + j}^2$ where $\paren {2 n + 1})^2 = 2 m + 1$. First it is worth rewriting this so as to eliminate $m$. {{begin-eqn}} {{eqn | l = \paren {2 n + 1}^2 | r = 4 n^2 + 4 n + 1 | c = }} {{eqn | r = 2 \paren {2 n^2 + 2 n} + 1 | c = }} {{end-eqn}} Thus the statement to be proved can be expressed: :$\ds \sum_{j \mathop = 0}^n \paren {2 n^2 + 2 n - j}^2 = \sum_{j \mathop = 1}^n \paren {2 n^2 + 2 n + j}^2$ \end{proof}
22399
\section{Sums of Sequences of Consecutive Squares which are Square} Tags: Sums of Sequences, Square Numbers \begin{theorem} The sums of the following sequences of successive squares are themselves square: {{begin-eqn}} {{eqn | l = \sum_{i \mathop = 7}^{29} k^2 | r = 7^2 + 8^2 + \cdots + 29^2 | c = }} {{eqn | l = \sum_{i \mathop = 7}^{39} k^2 | r = 7^2 + 8^2 + \cdots + 39^2 | c = }} {{eqn | l = \sum_{i \mathop = 7}^{56} k^2 | r = 7^2 + 8^2 + \cdots + 56^2 | c = }} {{eqn | l = \sum_{i \mathop = 7}^{190} k^2 | r = 7^2 + 8^2 + \cdots + 190^2 | c = }} {{end-eqn}} \end{theorem} \begin{proof} From Sum of Sequence of Squares: {{:Sum of Sequence of Squares}} Thus: {{begin-eqn}} {{eqn | l = \sum_{i \mathop = 7}^{29} i^2 | r = \sum_{i \mathop = 1}^{29} i^2 - \sum_{i \mathop = 1}^6 i^2 | c = }} {{eqn | r = \frac {29 \left({29 + 1}\right) \left({2 \times 29 + 1}\right)} 6 - \frac {6 \left({6 + 1}\right) \left({2 \times 6 + 1}\right)} 6 | c = }} {{eqn | r = \frac {29 \times 30 \times 59} 6 - \frac {6 \times 7 \times 13} 6 | c = }} {{eqn | r = 8 \, 555 - 91 | c = }} {{eqn | r = 8 \, 464 | c = }} {{eqn | r = 92^2 | c = }} {{end-eqn}} {{begin-eqn}} {{eqn | l = \sum_{i \mathop = 7}^{39} i^2 | r = \sum_{i \mathop = 1}^{39} i^2 - \sum_{i \mathop = 1}^6 i^2 | c = }} {{eqn | r = \frac {39 \left({39 + 1}\right) \left({2 \times 39 + 1}\right)} 6 - \frac {6 \left({6 + 1}\right) \left({2 \times 6 + 1}\right)} 6 | c = }} {{eqn | r = \frac {39 \times 40 \times 79} 6 - \frac {6 \times 7 \times 13} 6 | c = }} {{eqn | r = 20 \, 540 - 91 | c = }} {{eqn | r = 20 \, 449 | c = }} {{eqn | r = 143^2 | c = }} {{end-eqn}} {{begin-eqn}} {{eqn | l = \sum_{i \mathop = 7}^{56} i^2 | r = \sum_{i \mathop = 1}^{56} i^2 - \sum_{i \mathop = 1}^6 i^2 | c = }} {{eqn | r = \frac {56 \left({56 + 1}\right) \left({2 \times 56 + 1}\right)} 6 - \frac {6 \left({6 + 1}\right) \left({2 \times 6 + 1}\right)} 6 | c = }} {{eqn | r = \frac {56 \times 57 \times 113} 6 - \frac {6 \times 7 \times 13} 6 | c = }} {{eqn | r = 60 \, 116 - 91 | c = }} {{eqn | r = 60 \, 025 | c = }} {{eqn | r = 245^2 | c = }} {{end-eqn}} {{begin-eqn}} {{eqn | l = \sum_{i \mathop = 7}^{190} i^2 | r = \sum_{i \mathop = 1}^{190} i^2 - \sum_{i \mathop = 1}^6 i^2 | c = }} {{eqn | r = \frac {190 \left({190 + 1}\right) \left({2 \times 190 + 1}\right)} 6 - \frac {6 \left({6 + 1}\right) \left({2 \times 6 + 1}\right)} 6 | c = }} {{eqn | r = \frac {190 \times 191 \times 381} 6 - \frac {6 \times 7 \times 13} 6 | c = }} {{eqn | r = 2 \, 304 \, 415 - 91 | c = }} {{eqn | r = 2 \, 304 \, 324 | c = }} {{eqn | r = 1 \, 518^2 | c = }} {{end-eqn}} {{qed}} \end{proof}
22400
\section{Sums of Squares in Lines of Order 3 Magic Square} Tags: Magic Squares \begin{theorem} Consider the order 3 magic square: {{:Magic Square/Examples/Order 3}} : The sums of the squares of the top and bottom rows are equal, and differ by $18$ from the sums of the squares of the middle row : The sums of the squares of the left and right columns are equal , and differ by $18$ from the sums of the squares of the middle column. \end{theorem} \begin{proof} For the rows: {{begin-eqn}} {{eqn | l = 2^2 + 7^2 + 6^2 | r = 4 + 49 + 36 | c = }} {{eqn | r = 89 | c = }} {{eqn | l = 4^2 + 3^2 + 8^2 | r = 16 + 9 + 64 | c = }} {{eqn | r = 89 | c = }} {{eqn | l = 9^2 + 5^2 + 1^2 | r = 81 + 25 + 1 | c = }} {{eqn | r = 107 | c = }} {{eqn | r = 89 + 18 | c = }} {{end-eqn}} For the colums: {{begin-eqn}} {{eqn | l = 2^2 + 9^2 + 4^2 | r = 4 + 81 + 16 | c = }} {{eqn | r = 101 | c = }} {{eqn | l = 6^2 + 1^2 + 8^2 | r = 36 + 1 + 64 | c = }} {{eqn | r = 101 | c = }} {{eqn | l = 7^2 + 5^2 + 3^2 | r = 49 + 25 + 9 | c = }} {{eqn | r = 83 | c = }} {{eqn | r = 101 - 18 | c = }} {{end-eqn}} {{qed}} \end{proof}
22401
\section{Sums of Squares of Diagonals of Order 3 Magic Square} Tags: Magic Squares \begin{theorem} Consider the order 3 magic square: {{:Magic Square/Examples/Order 3}} The sums of the squares of the diagonals, when expressed as $3$-digit decimal numbers, are equal to the sums of the squares of those same diagonals of that same order 3 magic square when reversed. {{improve|Find a way to describe the "diagonals" accurately, as what is being demonstrated here does not match the description.}} \end{theorem} \begin{proof} For the top-left to bottom-right diagonals: {{begin-eqn}} {{eqn | l = 258^2 + 714^2 + 693^2 | r = 66564 + 509796 + 480249 | c = }} {{eqn | r = 1056609 | c = }} {{eqn | l = 852^2 + 417^2 + 396^2 | r = 725904 + 173889 + 156816 | c = }} {{eqn | r = 1056609 | c = }} {{end-eqn}} For the bottom-left to top-right diagonals: {{begin-eqn}} {{eqn | l = 456^2 + 312^2 + 897^2 | r = 207936 + 97344 + 804609 | c = }} {{eqn | r = 1109889 | c = }} {{eqn | l = 654^2 + 213^2 + 798^2 | r = 427716 + 45369 + 636804 | c = }} {{eqn | r = 1109889 | c = }} {{end-eqn}} {{qed}} \end{proof}
22402
\section{Sums of Squares of Lines of Order 3 Magic Square} Tags: Magic Squares \begin{theorem} Consider the order 3 magic square: {{:Magic Square/Examples/Order 3}} : The sums of the squares of the rows, when expressed as $3$-digit decimal numbers, are equal to the sums of the squares of those same rows of that same order 3 magic square when reflected in a vertical axis: :$\begin{array}{|c|c|c|} \hline 6 & 7 & 2 \\ \hline 1 & 5 & 9 \\ \hline 8 & 3 & 4 \\ \hline \end{array}$ Similarly: : The sums of the squares of the columns, when expressed as $3$-digit decimal numbers, are equal to the sums of the squares of those same columns of that same order 3 magic square when reflected in a horizontal axis: :$\begin{array}{|c|c|c|} \hline 4 & 3 & 8 \\ \hline 9 & 5 & 1 \\ \hline 2 & 7 & 6 \\ \hline \end{array}$ \end{theorem} \begin{proof} For the rows: {{begin-eqn}} {{eqn | l = 276^2 + 951^2 + 438^2 | r = 76176 + 904401 + 191844 | c = }} {{eqn | r = 1172421 | c = }} {{eqn | l = 672^2 + 159^2 + 834^2 | r = 451584 + 25281 + 695556 | c = }} {{eqn | r = 1172421 | c = }} {{end-eqn}} For the columns: {{begin-eqn}} {{eqn | l = 492^2 + 357^2 + 816^2 | r = 242064 + 127449 + 665856 | c = }} {{eqn | r = 1035369 | c = }} {{eqn | l = 294^2 + 753^2 + 618^2 | r = 86436 + 567009 + 381924 | c = }} {{eqn | r = 1035369 | c = }} {{end-eqn}} {{qed}} \end{proof}
22403
\section{Superabundant Numbers are Infinite in Number} Tags: Superabundant Numbers \begin{theorem} There are infinitely many superabundant numbers. \end{theorem} \begin{proof} {{AimForCont}} the set $S$ of superabundant numbers is finite. Let $m$ be the greatest element of $S$. By definition of superabundant, $m$ has the largest abundancy index of all the elements of $S$. Consider the integer $2 m$. From Abundancy Index of Product is greater than Abundancy Index of Proper Factors, $2 m$ has a higher abundancy index than $m$. There are two possibilities: :$(1): \quad 2 m$ is the smallest integer greater that $n$ which has a higher abundancy index than $m$. By definition, that would make $m$ superabundant. :$(2) \quad$ There exists a finite set $T := \set {n \in \Z: m < n < 2 m: \map A n > \map A m}$, where $\map A n$ denotes the abundancy index of $n$. The smallest element $t$ of $T$ therefore has an abundancy index greater than all smaller positive integers. Thus by definition $t$ is superabundant. In either case, there exists a superabundant number not in $S$. Thus $S$ cannot contain all superabundant numbers. But this contradicts our initial assumption that the set $S$, containing all superabundant numbers is finite. It follows by Proof by Contradiction that $S$ is infinite. Hence the result. {{qed}} \end{proof}
22404
\section{Superset Relation is Compatible with Subset Product} Tags: Compatible Relations \begin{theorem} Let $\struct {S, \circ}$ be a magma. Let $\circ_\PP$ be the subset product on $\powerset S$, the power set of $S$. Then the superset relation $\supseteq$ is compatible with $\circ_\PP$. \end{theorem} \begin{proof} By Subset Relation is Compatible with Subset Product, the subset relation $\subseteq$ is compatible with $\circ_\PP$. From Inverse of Subset Relation is Superset, the inverse of $\subseteq$ is $\supseteq$. The result follows from Inverse of Relation Compatible with Operation is Compatible. {{qed}} Category:Compatible Relations \end{proof}
22405
\section{Superset of Co-Countable Set} Tags: Set Theory \begin{theorem} Every superset of a co-countable set is co-countable. \end{theorem} \begin{proof} Let $S$ be a set. Let $A$ be co-countable in $S$, and let $B$ be such that $A \subseteq B \subseteq S$. From Relative Complement inverts Subsets, it follows that: :$\complement_S \left({B}\right) \subseteq \complement_S \left({A}\right)$ As $A$ is co-countable, $\complement_S \left({A}\right)$ is countable. By Subset of Countably Infinite Set is Countable, it follows that $\complement_S \left({B}\right)$ is also countable. Therefore, $B$ is also co-countable, and the result follows. {{qed}} \end{proof}
22406
\section{Superset of Dependent Set is Dependent} Tags: Matroid Theory \begin{theorem} Let $M = \struct {S, \mathscr I}$ be a matroid. Let $A, B \subseteq S$ such that $A \subseteq B$ If $A$ is a dependent subset then $B$ is a dependent subset. \end{theorem} \begin{proof} From the contrapositive statement of matroid axiom $(\text I 2)$: :$A \notin \mathscr I \implies B \notin \mathscr I$ By the definition of a dependent subset: :If $A$ is not an dependent subset then $B$ is not an dependent subset. {{qed}} Category:Matroid Theory \end{proof}
22407
\section{Superset of Dependent Set is Dependent/Corollary} Tags: Matroid Theory \begin{theorem} Let $M = \struct {S, \mathscr I}$ be a matroid. Let $A \subseteq S$. Let $x \in A$. If $x$ is a loop then $A$ is dependent. \end{theorem} \begin{proof} Let $x$ be a loop. By definition of a loop: :$\set x \notin \mathscr I$ By definition of a dependent subset: :$\set x$ is a dependent subset From Singleton of Element is Subset: :$\set x \subseteq A$ From Superset of Dependent Set is Dependent: :$A$ is a dependent subset {{qed}} \end{proof}
22408
\section{Superset of Infinite Set is Infinite} Tags: Set Theory \begin{theorem} Let $S$ be an infinite set. Let $T \supseteq S$ be a superset of $S$. Then $T$ is also infinite. \end{theorem} \begin{proof} Suppose $T$ were finite. Then by Set Finite iff Injection to Initial Segment of Natural Numbers, there is an injection $f: T \to \N_{<n}$ for some $n \in \N$. But then by Restriction of Injection is Injection, also the restriction of $f$ to $S$: :$f {\restriction_S}: S \to \N_{<n}$ is an injection. Again by Set Finite iff Injection to Initial Segment of Natural Numbers, this contradicts the assumption that $S$ is infinite. Hence $T$ is infinite. {{qed}} \end{proof}
22409
\section{Superset of Linearly Dependent Set is Linearly Dependent} Tags: Unitary Modules, Linear Algebra, Proofs by Contraposition, Modules \begin{theorem} Let $S$ be a set of elements of a unitary module. Let $S$ contain a subset $T$ which is linearly dependent. Then $S$ is also linearly dependent. \end{theorem} \begin{proof} Let $S$ be a linearly independent set. Let $T$ be a subset of $S$. By Subset of Linearly Independent Set is Linearly Independent, $T$ is also linearly independent. Thus by Proof by Contraposition, if $T$ is linearly dependent, then so must $S$ be. {{qed}} \end{proof}
22410
\section{Superset of Neighborhood in Metric Space is Neighborhood} Tags: Neighborhoods \begin{theorem} Let $M = \struct {A, d}$ be a metric space. Let $a \in A$ be a point in $M$. Let $N$ be a neighborhood of $a$ in $M$. Let $N \subseteq N' \subseteq A$. Then $N'$ is a neighborhood of $a$ in $M$. \end{theorem} \begin{proof} By definition of neighborhood: :$\exists \epsilon \in \R_{>0}: \map {B_\epsilon} a \subseteq N$ where $\map {B_\epsilon} a$ is the open $\epsilon$-ball of $a$ in $M$. By Subset Relation is Transitive: :$\map {B_\epsilon} a \subseteq N'$ The result follows by definition of neighborhood of $a$. {{qed}} \end{proof}
22411
\section{Superset of Neighborhood in Topological Space is Neighborhood} Tags: Neighborhoods \begin{theorem} Let $T = \struct {S, \tau}$ be a topological space. Let $x \in S$. Let $N$ be a neighborhood of $x$ in $T$. Let $N \subseteq N' \subseteq S$. Then $N'$ is a neighborhood of $x$ in $T$. That is: :$\forall x \in S: \forall N \in \NN_x: N' \supseteq N \implies N' \in \NN_x$ where $\NN_x$ is the neighborhood filter of $x$. \end{theorem} \begin{proof} By definition of neighborhood: :$\exists U \in \tau: x \in U \subseteq N \subseteq S$ where $U$ is an open set of $T$. By Subset Relation is Transitive: :$U \subseteq N'$ The result follows by definition of neighborhood of $x$. {{qed}} \end{proof}
22412
\section{Superset of Order Generating is Order Generating} Tags: Order Generating, Complete Lattices \begin{theorem} Let $L = \struct {S, \vee, \wedge, \preceq}$ be a complete lattice. Let $X, Y$ be subsets of $S$ such that :$X$ is order generating and :$X \subseteq Y$ Then $Y$ is order generating. \end{theorem} \begin{proof} Let $x \in S$. Thus by definition of complete lattice: :$x^\succeq \cap Y$ admits an infimum. By definition of complete lattice: :$x^\succeq$ admits an infimum and :$x^\succeq \cap X$ admits an infimum. By Intersection is Subset: :$x^\succeq \cap Y \subseteq x^\succeq$ By Infimum of Subset: :$\map \inf {x^\succeq} \preceq \map \inf {x^\succeq \cap Y}$ By Infimum of Upper Closure of Element: :$x \preceq \map \inf {x^\succeq \cap Y}$ By Set Intersection Preserves Subsets/Corollary: :$x^\succeq \cap X \subseteq x^\succeq \cap Y$ By Infimum of Subset: :$\map \inf {x^\succeq \cap Y} \preceq \map \inf {x^\succeq \cap X}$ By definition of order generating: :$\map \inf {x^\succeq \cap Y} \preceq x$ Thus by definition of antisymmetry: :$x = \map \inf {x^\succeq \cap Y}$ {{qed}} \end{proof}
22413
\section{Superset of Unsatisfiable Set is Unsatisfiable} Tags: Formal Semantics \begin{theorem} Let $\LL$ be a logical language. Let $\mathscr M$ be a formal semantics for $\LL$. Let $\FF$ be an $\mathscr M$-unsatisfiable set of formulas from $\LL$. Let $\FF'$ be a superset of $\FF$. Then $\FF'$ is also $\mathscr M$-unsatisfiable. \end{theorem} \begin{proof} By assumption, $\FF$ is unsatisfiable. Suppose now $\FF'$ were satisfiable. Then it would follow from Subset of Satisfiable Set is Satisfiable that $\FF$ were also satisfiable. We conclude that $\FF'$ must be unsatisfiable. {{qed}} Category:Formal Semantics \end{proof}
22414
\section{Superspace of Homeomorphic Subspaces may not have Homeomorphism to Itself containing Subspace Homeomorphism} Tags: Homeomorphisms \begin{theorem} Let $T_1 = \struct {S_1, \tau_1}$ and $T_2 = \struct {S_2, \tau_2}$ be topological spaces. Let $H_1 \subseteq S_1$ and $H_2 \subseteq S_2$. Let $H_1$ and $H_2$ be a homeomorphic. Then it may be the case that there does not exist a homeomorphism $g: T_1 \to T_2$ such that: :$g \restriction_{H_1} = f$ where: :$g \restriction_{H_1}$ is the restriction of $g$ to $H_1$ :$f: H_1 \to H_2$ is a homeomorphism. \end{theorem} \begin{proof} Proof by Counterexample: Let $\struct {\R, \tau_d}$ be the real number line with the usual (Euclidean) topology. Let $H_1 := \set 0 \cup \closedint 1 2 \cup \set 3$, where $\closedint 1 2$ denotes the closed interval from $1$ to $2$. Let $H_2 := \closedint 0 1 \cup \set 2 \cup \set 3$. $H_1$ and $H_2$ are homeomorphic, as can be demonstrated by the mapping $\phi: H_1 \to \H_2$ defined as: :$\forall x \in H_1: \map \phi x = \begin {cases} 2 & : x = 0 \\ 3 & : x = 3 \\ x - 1 & : x \in \closedint 1 2 \end {cases}$ which is trivially a homeomorphism. So, let $g$ be a homeomorphism from $H_1$ and $H_2$. Each of the singletons in $H_1$ has to map to one of the singletons in $H_2$. As a result, $g$ is not monotone. Let $f: \R \to \R$ be a homeomorphism such that $g$ is a restriction of $f$. From Continuous Real Function on Closed Interval is Bijective iff Strictly Monotone, $f$ is monotone. {{explain|We need another result for the above over the whole range $\R$ -- for some reason this does not exist. We have Surjective Monotone Function is Continuous but not in the other direction.}} Hence $f$ cannot be a bijection. Hence $f$ cannot be a homeomorphism. {{qed}} \end{proof}
22415
\section{Supplementary Interior Angles implies Parallel Lines} Tags: Parallel Lines, Angles, Lines, Transversals (Geometry) \begin{theorem} Given two infinite straight lines which are cut by a transversal, if the interior angles on the same side of the transversal are supplementary, then the lines are parallel. {{:Euclid:Proposition/I/28}} \end{theorem} \begin{proof} :200px Let $AB$ and $CD$ be infinite straight lines. Let $EF$ be a transversal that cuts them. Let at least one pair of interior angles on the same side of the transversal be supplementary. {{WLOG}}, let those interior angles be $\angle BGH$ and $\angle DHG$. So, by definition, $\angle DHG + \angle BGH$ equals two right angles. Also, from Two Angles on Straight Line make Two Right Angles, $\angle AGH + \angle BGH$ equals two right angles. Then from Euclid's first and third common notion and Euclid's fourth postulate: :$\angle AGH = \angle DHG$ Finally, by Equal Alternate Angles implies Parallel Lines: :$AB \parallel CD$ {{qed}} {{Euclid Note|28|I|{{EuclidNoteConverse|prop = 29|title = Parallelism implies Supplementary Interior Angles|part = third}}|part = second}} \end{proof}
22416
\section{Suprema Preserving Mapping on Ideals Preserves Directed Suprema} Tags: Order Theory \begin{theorem} Let $\left({S, \preceq}\right)$, $\left({T, \precsim}\right)$ be ordered sets. Let $f: S \to T$ be a mapping. Let every filter $F$ in $\left({S, \preceq}\right)$, $f$ preserve the infimum on $F$. Then $f$ preserves directed suprema. \end{theorem} \begin{proof} This follows by mutatis mutandis of the proof of Infima Preserving Mapping on Filters Preserves Filtered Infima. {{qed}} \end{proof}
22417
\section{Suprema Preserving Mapping on Ideals is Increasing} Tags: Increasing Mappings, Order Theory \begin{theorem} Let $\struct {S, \preceq}$ and $\struct {T, \precsim}$ be ordered sets. Let $f: S \to T$ be a mapping. For every ideal $I$ in $\struct {S, \preceq}$, let $f$ preserve the supremum on $I$. Then $f$ is increasing. \end{theorem} \begin{proof} Let $x, y \in S$ such that: :$x \preceq y$ By Supremum of Singleton: :$\set x$ and $\set y$ admit suprema in $\struct {S, \preceq}$ By Supremum of Lower Closure of Set: :$\set x^\preceq$ and $\set y^\preceq$ admit suprema in $\struct {S, \preceq}$ where $\set x^\preceq$ denotes the lower closure of $\set x$ By Lower Closure of Singleton: :$x^\preceq$ and $y^\preceq$ admit suprema in $\struct {S, \preceq}$ By Lower Closure of Element is Ideal: :$x^\preceq$ and $y^\preceq$ are ideals in $\struct {S, \preceq}$ By assumption and definition of mapping preserves the supremum on subset: :$\map {f^\to} {x^\preceq}$ and $\map {f^\to} {y^\preceq}$ admit suprema in $\struct {T, \precsim}$ and: :$\map \sup {\map {f^\to} {x^\preceq} } = \map f {\map \sup {x^\preceq} }$ and: :$\map \sup {\map {f^\to} {y^\preceq} } = \map f {\map \sup {y^\preceq} }$ By Supremum of Lower Closure of Element: :$\map \sup {x^\preceq} = x$ and $\map \sup {y^\preceq} = y$ By Lower Closure is Increasing: :$x^\preceq \subseteq y^\preceq$ By Image of Subset under Relation is Subset of Image/Corollary 2: :$\map {f^\to} {x^\preceq} \subseteq \map {f^\to} {y^\preceq}$ Thus by Supremum of Subset: :$\map f x \precsim \map f y$ Thus by definition: :$f$ is increasing. {{qed}} \end{proof}
22418
\section{Suprema and Infima of Combined Bounded Functions/Bounded Above} Tags: Suprema and Infima of Combined Bounded Functions \begin{theorem} Let $f$ and $g$ be real functions. Let $c$ be a constant. Let both $f$ and $g$ be bounded above on $S \subseteq \R$. Then: :$\ds \map {\sup_{x \mathop \in S} } {\map f x + c} = c + \map {\sup_{x \mathop \in S} } {\map f x}$ :$\ds \map {\sup_{x \mathop \in S} } {\map f x + \map g x} \le \map {\sup_{x \mathop \in S} } {\map f x} + \map {\sup_{x \mathop \in S} } {\map g x}$ where $\ds \map \sup {\map f x}$ is the supremum of $\map f x$. \end{theorem} \begin{proof} First we show that: :$\ds \map {\sup_{x \mathop \in S} } {\map f x + c} = c + \map {\sup_{x \mathop \in S} } {\map f x}$ Let $T = \set {\map f x: x \in S}$. Then: {{begin-eqn}} {{eqn | l = \map {\sup_{x \mathop \in S} } {\map f x + c} | r = \map {\sup_{y \mathop \in T} } {y + c} | c = }} {{eqn | r = c + \map {\sup_{y \mathop \in T} } y | c = Supremum Plus Constant }} {{eqn | r = c + \map {\sup_{x \mathop \in S} } {\map f x} | c = }} {{end-eqn}} Next we show that $\ds \map {\sup_{x \mathop \in S} } {\map f x + \map g x} \le \map {\sup_{x \mathop \in S} } {\map f x} + \map {\sup_{x \mathop \in S} } {\map g x}$: Let: :$\ds H = \map {\sup_{x \mathop \in S} } {\map f x}$ :$\ds K = \map {\sup_{x \mathop \in S} } {\map g x}$ Then: :$\forall x \in S: \map f x + \map g x \le H + K$ Hence $H + K$ is an upper bound for $\set {\map f x + \map g x: x \in S}$. The result follows. {{qed}} \end{proof}
22419
\section{Suprema and Infima of Combined Bounded Functions/Bounded Below} Tags: Suprema and Infima of Combined Bounded Functions \begin{theorem} Let $f$ and $g$ be real functions. Let $c$ be a constant. Let both $f$ and $g$ be bounded below on $S \subseteq \R$. Then: :$\ds \map {\inf_{x \mathop \in S} } {\map f x + c} = c + \map {\inf_{x \mathop \in S} } {\map f x}$ :$\ds \map {\inf_{x \mathop \in S} } {\map f x + \map g x} \ge \map {\inf_{x \mathop \in S} } {\map f x} + \map {\inf_{x \mathop \in S} } {\map g x}$ where $\ds \map \inf {\map f x}$ is the infimum of $\map f x$. \end{theorem} \begin{proof} First we show that: :$\ds \map {\inf_{x \mathop \in S} } {\map f x + c} = c + \map {\inf_{x \mathop \in S} } {\map f x}$ Let $T = \set {\map f x: x \in S}$. Then: {{begin-eqn}} {{eqn | l = \map {\inf_{x \mathop \in S} } {\map f x + c} | r = \map {\inf_{y \mathop \in T} } {y + c} | c = }} {{eqn | r = c + \map {\sup_{y \mathop \in T} } y | c = Infimum Plus Constant }} {{eqn | r = c + \map {\inf_{x \mathop \in S} } {\map f x} | c = }} {{end-eqn}} Next we show that: :$\ds \map {\inf_{x \mathop \in S} } {\map f x + \map g x} \ge \map {\inf_{x \mathop \in S} } {\map f x} + \map {\inf_{x \mathop \in S} } {\map g x}$ Let: :$\ds H = \map {\inf_{x \mathop \in S} } {\map f x}$ :$\ds K = \map {\inf_{x \mathop \in S} } {\map g x}$ Then: :$\forall x \in S: \map f x + g \map x \ge H + K$ Hence $H + K$ is a lower bound for $\set {\map f x + \map g x: x \in S}$. The result follows. {{qed}} \end{proof}
22420
\section{Supremum Inequality for Ordinals} Tags: Ordinals \begin{theorem} Let $A \subseteq \On$ and $B \subseteq \On$ be ordinals. Then: :$\ds \forall x \in A: \exists y \in B: x \le y \implies \bigcup A \le \bigcup B$ \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = x | o = < | r = \bigcup A }} {{eqn | ll= \leadsto | q = \exists z | l = x | o = < | r = z | c = {{Defof|Set Union}} }} {{eqn | lo= \land | l = z < A | o = < | r = A | c = }} {{eqn | ll= \leadsto | q = \exists y \in B: \exists z | l = x | o = < | r = z | c = {{hypothesis}} }} {{eqn | lo= \land | l = z | o = \le | r = y | c = }} {{eqn | ll= \leadsto | q = \exists y \in B | l = x | o = < | r = y }} {{eqn | ll= \leadsto | l = x | o = < | r = \bigcup B | c = Union of Ordinals is Least Upper Bound }} {{end-eqn}} Therefore: :$\ds \bigcup A \subseteq \bigcup B$ and: :$\ds \bigcup A \le \bigcup B$ {{qed}} \end{proof}
22421
\section{Supremum Metric and L1 Metric on Closed Real Intervals are not Topologically Equivalent} Tags: L1 Metric, Supremum Metric \begin{theorem} Let $S$ be the set of all real functions which are continuous on the closed interval $\closedint a b$. Let $d_1$ be the $L^1$ metric on $S$: :$\ds \forall f, g \in S: \map {d_1} {f, g} := \int_a^b \size {\map f t - \map g t} \rd t$ Let $d_\infty$ be the supremum metric on $S$: :$\ds \forall f, g \in S: \map {d_\infty} {f, g} := \sup_{x \mathop \in \closedint a b} \size {\map f x - \map g x}$ Then $d_1$ and $d_\infty$ are not topologically equivalent. \end{theorem} \begin{proof} Let $f, g \in S$. Then by definition of supremum metric: :$\forall x \in \closedint a b: \size {\map f x - \map g x} \le \map {d_\infty} {f, g}$ Hence by ... :$\map {d_1} {f, g} = \ds \int_a^b \size {\map f x - \map g x} \rd x \le \paren {b - a} \map {d_\infty} {f, g}$ {{finish|Find that link}} and so: :$\map {B_\epsilon} {f; d_\infty} \subseteq \map {b_{\paren {b - a} } } {f; d_1}$ Now let $f_0$ denote the constant mapping: :$\forall x: \map {f_0} x = 0$ We show that $\map {B_1} {f_0; d_\infty}$ is not $d_1$-open. {{AimForCont}} instead that $\map {B_1} {f_0; d_\infty}$ is $d_1$-open. Then we should have: :$\map {B_1} {f_0; d_1} \subseteq \map {B_1} {f_0; d_\infty}$ for some $\epsilon \in \R_{>0}$. But for $\epsilon > 0$ there exists a continuous function on $\closedint a b$ such that: :$\map {d_1} {f, f_0} < \epsilon$ yet: :$\map {d_\infty} {f, f_0} = 1$ So: :$f \in \map {B_1} {f_0; d_1}$ but: :$f \notin \map {B_1} {f_0; d_\infty}$ which contradicts our deduction that $\map {B_1} {f_0; d_1} \subseteq \map {B_1} {f_0; d_\infty}$. Hence it cannot be the case that $\map {B_1} {f_0; d_\infty}$ is $d_1$-open. The result follows. {{qed}} \end{proof}
22422
\section{Supremum Metric is Metric} Tags: Supremum Metric \begin{theorem} Let $S$ be a set. Let $M = \struct {A', d'}$ be a metric space. Let $A$ be the set of all bounded mappings $f: S \to M$. Let $d: A \times A \to \R$ be the supremum metric on $A$. Then $d$ is a metric. \end{theorem} \begin{proof} We have that the supremum metric on $A \times A$ is defined as: :$\ds \forall f, g \in A: \map d {f, g} := \sup_{x \mathop \in S} \map {d'} {\map f x, \map g x}$ where $f$ and $g$ are bounded mappings. First note that we have: {{begin-eqn}} {{eqn | l = \size {\map f x - \map g x} | r = \size {\map f x + \paren {-\map g x} } | c = }} {{eqn | o = \le | r = \size {\map f x} + \size {\paren {-\map g x} } | c = Triangle Inequality for Real Numbers }} {{eqn | r = \size {\map f x} + \size {\map g x} | c = {{Defof|Absolute Value}} }} {{eqn | o = \le | r = K + L | c = }} {{end-eqn}} and so the {{RHS}} exists. \end{proof}
22423
\section{Supremum Metric on Bounded Continuous Mappings is Metric} Tags: Supremum Metric \begin{theorem} Let $M_1 = \struct {A_1, d_1}$ and $M_2 = \struct {A_2, d_2}$ be metric spaces. Let $A$ be the set of all continuous mappings $f: M_1 \to M_2$ which are also bounded. Let $d: A \times A \to \R$ be the supremum metric on $A$. Then $d$ is a metric. \end{theorem} \begin{proof} The set $A$ is a subset of the set $A'$ of all bounded mappings $f: M_1 \to M_2$. Let $d': A' \times A' \to \R$ be the supremum metric on $A'$. From Supremum Metric is Metric, $\struct {A', d'}$ is a metric space. By definition, $A$ is a metric subspace of $A'$. Hence the result. {{qed}} \end{proof}
22424
\section{Supremum Metric on Bounded Real-Valued Functions is Metric} Tags: Supremum Metric on Bounded Real-Valued Functions is Metric, Supremum Metric \begin{theorem} Let $X$ be a set. Let $A$ be the set of all bounded real-valued functions $f: X \to \R$. Let $d: A \times A \to \R$ be the supremum metric on $A$. Then $d$ is a metric. \end{theorem} \begin{proof} We have that the supremum metric on $A \times A$ is defined as: :$\displaystyle \forall f, g \in A: d \left({f, g}\right) := \sup_{x \mathop \in X} \left\vert{f \left({x}\right) - g \left({x}\right)}\right\vert$ where $f$ and $g$ are bounded real-valued functions. So: :$\exists K, L \in \R: \left\vert{f \left({x}\right)}\right\vert \le K, \left\vert{g \left({x}\right)}\right\vert \le L$ for all $x \in X$. First note that we have: {{begin-eqn}} {{eqn | l = \left\vert{f \left({x}\right) - g \left({x}\right)}\right\vert | r = \left\vert{f \left({x}\right) + \left({- g \left({x}\right)}\right)}\right\vert | c = }} {{eqn | o = \le | r = \left\vert{f \left({x}\right)}\right\vert + \left\vert{\left({- g \left({x}\right)}\right)}\right\vert | c = Triangle Inequality for Real Numbers }} {{eqn | r = \left\vert{f \left({x}\right)}\right\vert + \left\vert{g \left({x}\right)}\right\vert | c = Definition of Absolute Value }} {{eqn | o = \le | r = K + L | c = }} {{end-eqn}} and so the RHS exists. \end{proof}
22425
\section{Supremum Metric on Bounded Real Functions on Closed Interval is Metric} Tags: Supremum Metric, Supremum Metric on Bounded Real Functions on Closed Interval is Metric \begin{theorem} Let $\closedint a b \subseteq \R$ be a closed real interval. Let $A$ be the set of all bounded real functions $f: \closedint a b \to \R$. Let $d: A \times A \to \R$ be the supremum metric on $A$. Then $d$ is a metric. \end{theorem} \begin{proof} We have that the supremum metric on $A \times A$ is defined as: :$\displaystyle \forall f, g \in A: d \left({f, g}\right) := \sup_{x \mathop \in \left[{a \,.\,.\, b}\right]} \left\vert{f \left({x}\right) - g \left({x}\right)}\right\vert$ where $f$ and $g$ are bounded real functions. So: :$\exists K, L \in \R: \left\vert{f \left({x}\right)}\right\vert \le K, \left\vert{g \left({x}\right)}\right\vert \le L$ for all $x \in \left[{a \,.\,.\, b}\right]$. First note that we have: {{begin-eqn}} {{eqn | l = \left\vert{f \left({x}\right) - g \left({x}\right)}\right\vert | r = \left\vert{f \left({x}\right) + \left({- g \left({x}\right)}\right)}\right\vert | c = }} {{eqn | o = \le | r = \left\vert{f \left({x}\right)}\right\vert + \left\vert{\left({- g \left({x}\right)}\right)}\right\vert | c = Triangle Inequality for Real Numbers }} {{eqn | r = \left\vert{f \left({x}\right)}\right\vert + \left\vert{g \left({x}\right)}\right\vert | c = Definition of Absolute Value }} {{eqn | o = \le | r = K + L | c = }} {{end-eqn}} and so the RHS exists. \end{proof}
22426
\section{Supremum Metric on Bounded Real Sequences is Metric} Tags: Supremum Metric on Bounded Real Sequences is Metric, Supremum Metric \begin{theorem} Let $A$ be the set of all bounded real sequences. Let $d: A \times A \to \R$ be the supremum metric on $A$. Then $d$ is a metric. \end{theorem} \begin{proof} We have that the supremum metric on $A \times A$ is defined as: :$\displaystyle \forall x, y \in A: d \left({x, y}\right) := \sup_{n \mathop \in \N} \left\vert{x_n - y_n}\right\vert$ where $x = \left\langle{x_i}\right\rangle$ and $y = \left\langle{y_i}\right\rangle$ are bounded real sequences. So: :$\exists K, L \in \R: \left\vert{x_n}\right\vert \le K, \left\vert{y_n}\right\vert \le L$ for all $n \in \N$. First note that we have: {{begin-eqn}} {{eqn | l = \left\vert{x_n - y_n}\right\vert | r = \left\vert{x_n + \left({- y_n}\right)}\right\vert | c = }} {{eqn | o = \le | r = \left\vert{x_n}\right\vert + \left\vert{\left({- y_n}\right)}\right\vert | c = Triangle Inequality for Real Numbers }} {{eqn | r = \left\vert{x_n}\right\vert + \left\vert{y_n}\right\vert | c = Definition of Absolute Value }} {{eqn | o = \le | r = K + L | c = }} {{end-eqn}} and so the RHS exists. \end{proof}
22427
\section{Supremum Metric on Continuous Real Functions is Subspace of Bounded} Tags: Supremum Metric \begin{theorem} Let $\closedint a b \subseteq \R$ be a closed real interval. Let $\mathscr C \closedint a b$ be the set of all continuous functions $f: \closedint a b \to \R$. Let $\map {\mathscr B} {\closedint a b, \R}$ be the set of all bounded real functions $f: \closedint a b \to \R$. Let $d$ be the supremum metric on $\map {\mathscr B} {\closedint a b, \R}$. Then $\struct {\mathscr C \closedint a b, d_{\mathscr C} }$ is a subspace of $\struct {\map {\mathscr B} {\closedint a b, \R}, d}$. \end{theorem} \begin{proof} Let $f \in \mathscr C \closedint a b$. Then by Image of Closed Real Interval is Bounded, $f$ is bounded on $\closedint a b$. Thus $f \in \map {\mathscr B} {\closedint a b, \R}$ and the result follows. {{qed}} \end{proof}
22428
\section{Supremum Metric on Differentiability Class is Metric} Tags: Supremum Metric \begin{theorem} Let $\closedint a b \subseteq \R$ be a closed real interval. Let $r \in \N$ be a natural number. Let $A := \mathscr D^r \closedint a b$ be the set of all continuous functions $f: \closedint a b \to \R$ which are of differentiability class $r$. Let $d: A \times A \to \R$ be the supremum metric on $A$. Then $d$ is a metric. \end{theorem} \begin{proof} We have that the supremum metric on $A \times A$ is defined as: :$\ds \forall f, g \in A: \map d {f, g} := \sup_{\substack {x \mathop \in \closedint a b \\ i \mathop \in \set {0, 1, 2, \ldots, r} } } \size {\map {f^{\paren i} } x - \map {g^{\paren i} } x}$ where $f$ and $g$ are continuous functions on $\closedint a b$ which are of differentiability class $r$. \end{proof}
22429
\section{Supremum Norm is Norm} Tags: Supremum Norm, Examples of Norms, Supremum Norm is Norm, Norm Examples, Functional Analysis \begin{theorem} Let $S$ be a set. Let $\struct {X, \norm {\, \cdot \,} }$ be a normed vector space over $K \in \set {\R, \C}$. Let $\BB$ be the set of bounded mappings $S \to X$. Let $\norm {\, \cdot \,}_\infty$ be the supremum norm on $\BB$. Then $\norm {\, \cdot \,}_\infty$ is a norm on $\BB$. {{MissingLinks|Add a link that establishes that $\BB$ is a vector space}} \end{theorem} \begin{proof} First: {{begin-eqn}} {{eqn | l = \norm f_\infty | r = 0 }} {{eqn | ll= \leadstoandfrom | l = \sup_{x \mathop \in S} \norm {\map f x} | r = 0 }} {{eqn | ll= \leadstoandfrom | q = \forall x \in S | l = \norm {\map f x} | r = 0 | c = since $\norm {\, \cdot \,}$ is a norm, and hence non-negative }} {{eqn | ll= \leadstoandfrom | q = \forall x \in S | l = \map f x | r = 0 | c = since $\norm x = 0 \iff x = 0$ }} {{eqn | ll= \leadstoandfrom | l = f | r = 0 | c = }} {{end-eqn}} Now let $\lambda \in K, f \in \BB$ We have: {{begin-eqn}} {{eqn | l = \norm {\lambda f}_\infty | r = \sup_{x \mathop \in S} \norm {\lambda \map f x} | c = }} {{eqn | r = \size \lambda_K \sup_{x \mathop \in S} \norm {\map f x} | c = Multiple of Supremum, and because $\norm {\, \cdot \,}$ is a norm }} {{eqn | r = \size \lambda_K \, \norm f_\infty | c = }} {{end-eqn}} Finally let $f, g \in \BB$. We have: {{begin-eqn}} {{eqn | l = \norm {f + g}_\infty | r = \sup_{x \mathop \in S} \norm {\map f x + \map g x} | c = }} {{eqn | o = \le | r = \sup_{x \mathop \in S} \paren {\norm {\lambda \map f x} + \norm {\lambda \map g x} } | c = because $\norm {\, \cdot \,}$ is a norm }} {{eqn | o = \le | r = \sup_{x \mathop \in S} \norm {\lambda \map f x} + \sup_{x \mathop \in S} \norm {\lambda \map g x} | c = Supremum of Sum }} {{eqn | r = \norm f_\infty + \norm g_\infty | c = }} {{end-eqn}} Thus $\norm {\, \cdot \,}_\infty$ has the defining properties of a norm. {{qed}} Category:Examples of Norms Category:Supremum Norm Category:Supremum Norm is Norm \end{proof}
22430
\section{Supremum Norm on Vector Space of Real Matrices is Norm} Tags: Examples of Norms \begin{theorem} Supremum Norm forms a norm on the vector space of real matrices. \end{theorem} \begin{proof} Let $M \in \R^{m \times n}: m, n \in \N_{>0}$ be a real matrix. Denote the $\paren {i, j}$-th entry of $M$ by $a_{i j}$. Note that the set of matrix elements of $M$ is a finite set of real numbers. We have that: :Real Numbers form Ordered Field :Finite Non-Empty Subset of Ordered Set has Maximal and Minimal Elements Therefore, $M$ has the greatest element. \end{proof}
22431
\section{Supremum Plus Constant} Tags: Real Analysis, Analysis \begin{theorem} Let $S$ be a subset of the set of real numbers $\R$. Let $S$ be bounded above. Let $\xi \in \R$. Then: :$\ds \map {\sup_{x \mathop \in S} } {x + \xi} = \xi + \map {\sup_{x \mathop \in S} } x$ where $\sup$ denotes supremum. \end{theorem} \begin{proof} Let $B = \sup S$. Let $T = \set {x + \xi: x \in S}$. Since $\forall x \in S: x \le B$ it follows that: :$\forall x \in S: x + \xi \le B + \xi$ Hence $\xi + B$ is an upper bound for $T$. If $C$ is the supremum for $T$ then $C \le \xi + B$. On the other hand: :$\forall y \in T: y \le C$ Therefore: :$\forall y \in T: y - \xi \le C - \xi$ Since $S = \set {y - \xi: y \in T}$ it follows that $C - \xi$ is an upper bound for $S$ and so $B \le C - \xi$. So we have shown that $C \le \xi + B$ and $C \ge \xi + B$, hence the result. {{qed}} \end{proof}
22432
\section{Supremum by Suprema of Directed Set in Simple Order Product} Tags: Simple Order Product, Up-Complete Semilattices \begin{theorem} Let $\struct {S, \preceq}$ be an up-complete meet semilattice. Let $\struct {S \times S, \precsim}$ be the simple order product of $\struct {S, \preceq}$ and $\struct {S, \preceq}$. Let $D$ be a directed subset of $S \times S$. Then: :$\sup D = \tuple {\map \sup {\map {\pr_1^\to} D}, \map \sup {\map {\pr_2^\to} D} }$ where :$\pr_1$ denotes the first projection on $S \times S$ :$\pr_2$ denotes the second projection on $S \times S$ :$\map {\pr_1^\to} D$ denotes the image of $D$ under $\pr_1$ \end{theorem} \begin{proof} By Up-Complete Product: :$\struct {S \times S, \precsim}$ is up-complete. By definition of up-complete: :$D$ admits a supremum. By definition of Cartesian product: :$\exists d_1, d_2 \in S: \sup D = \tuple {d_1, d_2}$ By Up-Complete Product/Lemma 2: :$D_1 := \map {\pr_1^\to} D$ is directed and :$D_2 := \map {\pr_2^\to} D$ is directed. By definition of up-complete: :$D_1$ admits a supremum and :$D_2$ admits a supremum We will prove that :$d_2$ is upper bound for $D_2$ Let $x \in D_2$. By definition of image of set: :$\exists \tuple {a, b} \in D: \map {\pr_2} {a, b} = x$ By definition of second projection: $b = x$ By definition of supremum: :$\tuple {d_1, d_2}$ is upper bound for $D$. By definition of upper bound: :$\tuple {a, x} \precsim \tuple {d_1, d_2}$ Thus by definition of simple order product: :$x \preceq d_2$ {{qed|lemma}} Analogically we have that :$d_1$ is upper bound for $D_1$ By definition of supremum: :$\sup D_1 \preceq d_1$ and $\sup D_2 \preceq d_2$ By definition of simple order product: :$\tuple {\sup D_1, \sup D_2} \precsim \sup D$ By Up-Complete Product/Lemma 1: :$D_1 \times D_2$ is directed. By definition of up-complete: :$D_1 \times D_2$ admits a supremum. By definition of subset: :$D \subseteq D_1 \times D_2$ By Supremum of Subset: :$\sup D \precsim \map \sup {D_1 \times D_2}$ By Supremum of Simple Order Product: :$\sup D \precsim \tuple {\sup D_1, \sup D_2}$ Thus by definition of antisymmetry: :$\sup D = \tuple {\sup D_1, \sup D_2}$ {{qed}} \end{proof}
22433
\section{Supremum does not Precede Infimum} Tags: Order Theory, Orderings, Infima, Suprema \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Let $T \subseteq S$ admit both a supremum $M$ and an infimum $m$. Then $m \preceq M$. \end{theorem} \begin{proof} By definition of supremum: :$\forall a \in T: a \preceq M$ By definition of infimum: :$\forall a \in T: m \preceq a$ The result follows from transitivity of ordering. {{qed}} \end{proof}
22434
\section{Supremum in Ordered Subset} Tags: Order Theory \begin{theorem} Let $L = \left({S, \preceq}\right)$ be an ordered set. Let $R = \left({T, \preceq'}\right)$ be an ordered subset of $L$. Let $X \subseteq T$ such that :$X$ admits an supremum in $L$. Then $\sup_L X \in T$ {{iff}} :$X$ admits an supremum in $R$ and $\sup_R X = \sup_L X$ \end{theorem} \begin{proof} This follows by mutatis mutandis of the proof of Infimum in Ordered Subset. {{qed}} \end{proof}
22435
\section{Supremum is Coproduct in Order Category} Tags: Poset Categories, Order Categories \begin{theorem} Let $\mathbf P$ be an order category with ordering $\preceq$. Let $p, q \in P_0$, and suppose they have some supremum $r = \sup \left\{{p, q}\right\}$. Then $r$ is the coproduct of $p$ and $q$ in $\mathbf P$. \end{theorem} \begin{proof} Let $\mathbf P^{\text{op}}$ be the dual category of $\mathbf P$. From Dual of Order Category, it is the order category corresponding to the dual ordering $\succeq$. From Dual Pairs (Order Theory), it follows that in $\mathbf P^{\text{op}}$: :$r = \inf \left\{{p, q}\right\}$ where $\inf$ denotes infimum. By Infimum is Product in Order Category, $r$ is the product of $p$ and $q$ in $\mathbf P^{\text{op}}$. By Dual Pairs (Category Theory), $r$ is the coproduct of $p$ and $q$ in $\mathbf P$. {{qed}} \end{proof}
22436
\section{Supremum is Dual to Infimum} Tags: Order Theory, Infima, Suprema \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Let $a \in S$ and $T \subseteq S$. The following are dual statements: :$a$ is a supremum for $T$ :$a$ is an infimum for $T$ \end{theorem} \begin{proof} By definition, $a$ is a supremum for $T$ {{iff}}: :$a$ is an upper bound for $T$ :$a \preceq b$ for all upper bounds $b$ of $T$ The dual of this statement is: :$a$ is a lower bound for $T$ :$b \preceq a$ for all lower bounds $b$ of $T$ by Dual Pairs (Order Theory). By definition, this means $a$ is an infimum for $T$. The converse follows from Dual of Dual Statement (Order Theory). {{qed}} \end{proof}
22437
\section{Supremum is Increasing relative to Product Ordering} Tags: Increasing Mappings, Order Theory \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Let $I$ be a set. Let $f, g: I \to S$. Let $f \sqbrk I$ denote the image of $I$ under $f$. Let: :$\forall i \in I: \map f i \preceq \map g i$ That is, let $f \preceq g$ in the product ordering. Let $f \sqbrk I$ and $g \sqbrk I$ admit suprema. Then: :$\sup f \sqbrk I \preceq \sup g \sqbrk I$ \end{theorem} \begin{proof} Let $x \in f \sqbrk I$. Then: :$\exists j \in I: \map f j = x$ Then: :$\map f j \prec \map g j$ By the definition of supremum: :$\sup g \sqbrk I$ is an upper bound of $g \sqbrk I$ Thus: :$\map g j \preceq \sup g \sqbrk I$ Since $\preceq$ is transitive: :$x = \map f j \preceq \sup g \sqbrk I$ Since this holds for all $x \in f \sqbrk I$, $\sup g \sqbrk I$ is an upper bound of $f \sqbrk I$. Thus by the definition of supremum: :$\sup f \sqbrk I \preceq \sup g \sqbrk I$ {{qed}} Category:Order Theory Category:Increasing Mappings \end{proof}
22438
\section{Supremum is Unique} Tags: Order Theory \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Let $T$ be a non-empty subset of $S$. Then $T$ has at most one supremum in $S$. \end{theorem} \begin{proof} Let $c$ and $c'$ both be suprema of $T$ in $S$. From the definition of supremum, $c$ and $c'$ are upper bounds of $T$ in $S$. By that definition: :$c$ is an upper bound of $T$ in $S$ and $c'$ is a supremum of $T$ in $S$ implies that $c' \preceq c$ :$c'$ is an upper bound of $T$ in $S$ and $c$ is a supremum of $T$ in $S$ implies that $c \preceq c'$. So: :$c' \preceq c \land c \preceq c'$ and thus by the antisymmetry of the ordering $\preceq$: :$c = c'$ {{Qed}} \end{proof}
22439
\section{Supremum is not necessarily Greatest Element} Tags: Suprema, Order Theory, Supremum is not necessarily Greatest Element \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Let $T$ admit a supremum in $S$. Then the supremum of $T$ in $S$ is not necessarily the greatest element of $T$. \end{theorem} \begin{proof} Consider the subset $T$ of the set of real numbers $\R$: :$T := \set {x \in \R: 1 \le x < 2}$ The number $2$ cannot be the greatest element of $T$ as $2 \notin T$. However, $2$ is the supremum of $T$ in $S$. Indeed, by definition: :$\forall x \in T: x < 2$ So, let $x < 2$. Then consider $y = \dfrac x 1 + \dfrac 2 1 = \dfrac {x + 2} 2$. We have that $x < y < 2$ by Mediant is Between. Thus $y \in T$ but $y > x$ and so $x$ cannot be the greatest element of $T$. Neither can $y$ be the supremum of $T$ in $S$. The conclusion is that there is no greatest element of $T$. Hence the result. {{qed}} \end{proof}
22440
\section{Supremum of Absolute Value of Difference equals Difference between Supremum and Infimum} Tags: Real Analysis \begin{theorem} Let $f$ be a real function. Let $S$ be a subset of the domain of $f$. Let $\ds \sup_{x \mathop \in S} \set {\map f x}$ and $\ds \inf_{x \mathop \in S} \set {\map f x}$ exist. Then $\ds \sup_{x, y \mathop \in S} \set {\size {\map f x - \map f y} }$ exists and: :$\ds \sup_{x, y \mathop \in S} \set {\size {\map f x - \map f y} } = \sup_{x \mathop \in S} \set {\map f x} - \inf_{x \mathop \in S} \set {\map f x}$ \end{theorem} \begin{proof} {{begin-eqn}} {{eqn | l = \sup_{x \mathop \in S} \set {\map f x} - \inf_{x \mathop \in S} \set {\map f x} | r = \sup_{x \mathop \in S} \set {\map f x} + \sup_{x \mathop \in S} \set {-\map f x} | c = Negative of Infimum is Supremum of Negatives }} {{eqn | r = \sup_{x, y \mathop \in S} \set {\map f x + \paren {-\map f y} } | c = Supremum of Sum equals Sum of Suprema }} {{eqn | r = \sup_{x, y \mathop \in S} \set {\map f x - \map f y} }} {{eqn | r = \sup_{x, y \mathop \in S} \set {\size {\map f x - \map f y} } | c = Supremum of Absolute Value of Difference equals Supremum of Difference }} {{end-eqn}} {{qed}} Category:Real Analysis \end{proof}
22441
\section{Supremum of Absolute Value of Difference equals Supremum of Difference} Tags: Suprema, Real Analysis, Absolute Value Function \begin{theorem} Let $S$ be a non-empty real set. Let $\ds \sup_{x, y \mathop \in S} \paren {x - y}$ exist. Then $\ds \sup_{x, y \mathop \in S} \size {x - y}$ exists and: :$\ds \sup_{x, y \mathop \in S} \size {x - y} = \sup_{x, y \mathop \in S} \paren {x - y}$ \end{theorem} \begin{proof} Consider the set $\set {x - y: x, y \in S, x - y \le 0}$. There is a number $x'$ in $S$ as $S$ is non-empty. Therefore, $0 \in \set {x - y: x, y \in S, x - y \le 0}$ as $x = y = x'$ implies that $x - y = 0$, $x, y \in S$, and $x - y \le 0$. Also, $0$ is an upper bound for $\set {x - y: x, y \in S, x - y \le 0}$ by definition. Accordingly: :$\ds \sup_{x, y \mathop \in S, x − y \mathop \le 0} \paren {x - y} = 0$ Consider the set $\left\{{x - y: x, y \in S, x - y \ge 0}\right\}$. There is a number $x'$ in $S$ as $S$ is non-empty. Therefore, $0 \in \left\{{x - y: x, y \in S, x - y \ge 0}\right\}$ as $x = y = x'$ implies that $x - y = 0$, $x, y \in S$, $x - y \ge 0$. Accordingly: :$\ds \sup_{x, y \mathop \in S, x − y \mathop \ge 0} \paren {x - y} \ge 0$ {{improve|I can't immediately think of how it would be done, but it would be good if we could devise a neater and more compact notation that what is used here. All the complicated mathematics is being done in the underscript, which makes it not easy to follow. (Improved Dec. 2016.)}} {{begin-eqn}} {{eqn | l = \sup_{x, y \mathop \in S} \paren {x - y} | r = \sup_{x, y \mathop \in S, x − y \mathop \ge 0 \text { or } x − y \mathop \le 0} \paren {x - y} | c = as ($x - y \ge 0$ or $x - y \le 0$) is true }} {{eqn | r = \max \set {\sup_{x, y \mathop \in S, x − y \mathop \ge 0} \paren {x - y}, \sup_{x, y \mathop \in S, x − y \mathop \le 0} \paren {x - y} } | c = by Supremum of Set Equals Maximum of Suprema of Subsets }} {{eqn | r = \max \set {\sup_{x, y \mathop \in S, x − y \mathop \ge 0} \paren {x - y}, 0} | c = as $\ds \sup_{x, y \mathop \in S, x − y \mathop \le 0} \paren {x - y} = 0$ }} {{eqn | r = \sup_{x, y \mathop \in S, x − y \mathop \ge 0} \paren {x - y} | c = as $\ds \sup_{x, y \mathop \in S, x − y \mathop \ge 0} \paren {x - y} \ge 0$ }} {{eqn | r = \sup_{x, y \mathop \in S, x − y \mathop \ge 0} \size {x - y} | c = as $\size {x − y} = x − y$ since $x − y \ge 0$ }} {{eqn | r = \max \set {\sup_{x, y \mathop \in S, x − y \mathop \ge 0} \size {x - y}, \sup_{x, y \mathop \in S, x − y \mathop \ge 0} \size {x - y} } | c = as the two arguments of max are equal }} {{eqn | r = \max \set {\sup_{x, y \mathop \in S, x − y \mathop \ge 0} \size {x - y}, \sup_{y, x \mathop \in S, y − x \mathop \ge 0} \size {y - x} } | c = by renaming variables $x \leftrightarrow y$ }} {{eqn | r = \max \set {\sup_{x, y \mathop \in S, x − y \mathop \ge 0} \size {x - y}, \sup_{x, y \mathop \in S, x − y \mathop \le 0} \size {x - y} } | c = }} {{eqn | r = \sup_{x, y \mathop \in S, x − y \mathop \ge 0 \text { or } x − y \mathop \le 0} \size {x - y} | c = by Supremum of Set Equals Maximum of Suprema of Subsets }} {{eqn | r = \sup_{x, y \mathop \in S} \size {x - y} | c = as ($x - y \ge 0$ or $x - y \le 0$) is true }} {{end-eqn}} {{qed}} Category:Suprema Category:Absolute Value Function \end{proof}
22442
\section{Supremum of Bounded Above Set of Reals is in Closure} Tags: Set Closures, Real Analysis, Analysis \begin{theorem} Let $\R$ be the real number line under the Euclidean metric. Let $H \subseteq \R$ be a bounded above subset of $\R$ such that $H \ne \O$. Let $u = \sup H$ be the supremum of $H$. Then: :$u \in \map \cl H$ where $\map \cl H$ denotes the closure of $H$ in $\R$. \end{theorem} \begin{proof} Let $\epsilon \in \R_{>0}$ be a strictly positive real number. Let $\map {B_\epsilon} u$ be the open $\epsilon$-ball of $u$ in $\R$. From Distance from Subset of Real Numbers: :$\map d {u, H} = 0$ Thus by definition of distance from subset: :$\exists x \in H: \map d {u, x} < \epsilon$ Thus $x \in \map {B_\epsilon} u$. As $x \in H$ and $x \in \map {B_\epsilon} u$, from the definition of intersection: :$x \in H \cap \map {B_\epsilon} u$ The result follows from Condition for Point being in Closure. {{qed}} \end{proof}
22443
\section{Supremum of Elements of Sublattice not necessarily Same as for Lattice} Tags: Sublattices, Suprema \begin{theorem} Let $\struct {S, \preceq}$ be a '''lattice'''. Let $\struct {T, \preceq_T}$ be a '''sublattice''' of $S$. Let $a, b \in T$. Then it is not necessarily the case that: :$\sup_S \set {a, b}$ is the same as: :$\sup_T \set {a, b}$ \end{theorem} \begin{proof} Proof by Counterexample: Let $\struct {G, \circ}$ be a group. Let $\mathbb G$ be the set of all subgroups of $G$. Let $\powerset G$ denote the power set of $G$. Let $\struct {\powerset G, \subseteq}$ be the complete lattice formed by $\powerset G$ and $\subseteq$. From Power Set is Complete Lattice, $\struct {\powerset G, \subseteq}$ indeed forms a complete lattice. Let $\struct {\mathbb G, \subseteq}$ be the complete lattice formed by $\mathbb G$ and $\subseteq$. From Set of Subgroups forms Complete Lattice, $\struct {\mathbb G, \subseteq}$ indeed forms a complete lattice. By definition, $\struct {\mathbb G, \subseteq}$ is a sublattice of $\struct {\powerset G, \subseteq}$. Let $H, K \in \mathbb G$. We have that: :$H \cup K$ is the supremum of $H$ and $K$ in $\struct {\powerset G, \subseteq}$. But from Union of Subgroups, $H \cup K$ is not a subgroup of $G$ element of $\mathbb G$. However, from Supremum of Subgroups in Lattice, we do have that the subset product $H K$ in $G$ is a subgroup of $G$ such that: :$\sup_{\mathbb G} \set {H, K} = H K$ {{qed}} \end{proof}
22444
\section{Supremum of Empty Set is Smallest Element} Tags: Empty Set, Order Theory, Suprema \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Then: :the supremum of the empty set exists {{iff}} the smallest element of $S$ exists in which case: :$\map \sup \O$ is the smallest element of $S$ \end{theorem} \begin{proof} Observe that, vacuously, any $s \in S$ is an upper bound for $\O$. \end{proof}
22445
\section{Supremum of Function is less than Supremum of Greater Function} Tags: Real Analysis, Suprema \begin{theorem} Let $f$ and $g$ be real functions. Let $S$ be a subset of $\Dom f \cap \Dom g$. Let $\map f x \le \map g x$ for every $x \in S$. Let $\ds \sup_{x \mathop \in S} \map g x$ exist. Then $\ds \sup_{x \mathop \in S} \map f x$ exists and: :$\ds \sup_{x \mathop \in S} \map f x \le \sup_{x \mathop \in S} \map g x$. \end{theorem} \begin{proof} We have: {{begin-eqn}} {{eqn | l = \sup g | r = \map \sup {f + \paren {g - f} } }} {{eqn | r = \sup f + \sup \left({g - f}\right) | c = Supremum of Sum equals Sum of Suprema }} {{end-eqn}} Supremum of Sum equals Sum of Suprema also gives that $\sup f$ and $\sup \paren {g - f}$ exist. We have: {{begin-eqn}} {{eqn | q = \forall x \in S | l = \map g x | o = \ge | r = \map f x | c = }} {{eqn | ll= \leadsto | q = \forall x \in S | l = \map g x - \map f x | o = \ge | r = 0 | c = }} {{eqn | ll= \leadsto | q = \forall x \in S | l = \map \sup {g - f} | o = \ge | r = \map g x - \map f x \ge 0 | c = as $\map \sup {g - f}$ is an upper bound for $\set {\map g x - \map f x: x \in S}$ }} {{eqn | ll= \leadsto | l = \map \sup {g - f} | o = \ge | r = 0 }} {{eqn | ll= \leadsto | l = \sup f + \map \sup {g - f} | o = \ge | r = \sup f }} {{eqn | ll= \leadsto | l = \sup g | o = \ge | r = \sup f | c = as $\sup g = \sup f + \map \sup {g - f}$ }} {{end-eqn}} {{qed}} Category:Real Analysis Category:Suprema \end{proof}
22446
\section{Supremum of Ideals is Increasing} Tags: Supremum, Increasing Mappings \begin{theorem} Let $L = \left({S, \preceq}\right)$ be an up-complete ordered set. Let $\mathit{Ids}\left({L}\right)$ be the set of all ideals in $L$. Let $P = \left({\mathit{Ids}\left({L}\right), \precsim}\right)$ be an ordered set where $\mathord \precsim = \subseteq\restriction_{\mathit{Ids}\left({L}\right)\times \mathit{Ids}\left({L}\right)}$ Let $f: \mathit{Ids}\left({L}\right) \to S$ be a mapping such that :$\forall I \in \mathit{Ids}\left({L}\right): f\left({I}\right) = \sup I$ Then $f$ is an increasing mapping. \end{theorem} \begin{proof} Let $I, J \in \mathit{Ids}\left({L}\right)$ such that :$I \precsim J$ By definition of $\precsim$: :$I \subseteq J$ By definition of up-complete: :$I$ and $J$ admit suprema in $L$. By Supremum of Subset: :$\sup I \preceq \sup J$ Thus by definition of $f$: :$f\left({I}\right) \preceq f\left({J}\right)$ Hence $f$ is an increasing mapping. {{qed}} \end{proof}
22447
\section{Supremum of Ideals is Upper Adjoint} Tags: Galois Connections, Continuous Lattices \begin{theorem} Let $L = \left({S, \vee, \preceq}\right)$ be a bounded below continuous join semilattice. Let $\mathit{Ids}\left({L}\right)$ be the set of all ideals in $L$. Let $P = \left({\mathit{Ids}\left({L}\right), \precsim}\right)$ be an ordered set where $\mathord \precsim = \subseteq\restriction_{\mathit{Ids}\left({L}\right)\times \mathit{Ids}\left({L}\right)}$ Let $f: \mathit{Ids}\left({L}\right) \to S$ be a mapping such that :$\forall I \in \mathit{Ids}\left({L}\right): f\left({I}\right) = \sup I$ Then $f$ is an upper adjoint of Galois connection. \end{theorem} \begin{proof} Define $d: S \to \mathit{Ids}\left({L}\right)$ :$\forall t \in S: d\left({t}\right) = \inf \left({f^{-1}\left[{t^\succeq}\right]}\right)$ where :$t^\succeq$ denotes the upper closure of $t$, :$f^{-1}\left[{t^\succeq}\right]$ denotes the image of $t^\succeq$ over $f^{-1}$. We will prove that :$\forall t \in S: d\left({t}\right) = \min \left({f^{-1}\left[{t^\succeq}\right]}\right)$ Let $t \in S$. By Continuous iff For Every Element There Exists Ideal Element Precedes Supremum: :there exists an ideal $J$ in $L$ such that ::$t \preceq \sup J$ and for every ideal $K$ in $L$: $t \preceq \sup K \implies J \subseteq K$ We will prove that :$\forall K \in \mathit{Ids}\left({L}\right): K$ is lower bound for $f^{-1}\left[{t^\succeq}\right] \implies K \precsim J$ Let $K \in \mathit{Ids}\left({L}\right)$ such that :$K$ is lower bound for $f^{-1}\left[{t^\succeq}\right]$ By definition of $f$: :$t \preceq f\left({J}\right)$ By definition of upper closure of element: :$f\left({J}\right) \in t^\succeq$ By definition of image of set: :$J \in f^{-1}\left[{t^\succeq}\right]$ Thus by definition of lower bound: :$K \precsim J$ {{qed|lemma}} We will prove that :$J$ is lower bound for $f^{-1}\left[{t^\succeq}\right]$ Let $K \in \mathit{Ids}\left({L}\right)$ such that :$K \in f^{-1}\left[{t^\succeq}\right]$ By definition of image of set: :$f\left({K}\right) \in t^\succeq$ By definition of upper closure of element: :$t \preceq f\left({K}\right)$ By definition of $f$: :$t \preceq \sup K$ Then :$J \subseteq K$ Thus by definition of $\precsim$: :$J \precsim K$ {{qed|lemma}} By definition of supremum: :$t \preceq \sup \left({\inf \left({f^{-1}\left[{t^\succeq}\right]}\right)}\right)$ By definition of $f$: :$t \preceq f\left({\inf \left({f^{-1}\left[{t^\succeq}\right]}\right)}\right)$ By definition of upper closure of element: :$f\left({\inf \left({f^{-1}\left[{t^\succeq}\right]}\right)}\right) \in t^\succeq$ By definition of image of set: :$\inf \left({f^{-1}\left[{t^\succeq}\right]}\right) \in f^{-1}\left[{t^\succeq}\right]$ Thus by definition of smallest element: :$d\left({t}\right) = \min \left({f^{-1}\left[{t^\succeq}\right]}\right)$ {{qed|lemma}} By Supremum of Ideals is Increasing: :$f$ is an increasing mapping. By Galois Connection is Expressed by Minimum: :$\left({f, d}\right)$ is a Galois connection. Hence $f$ is an upper adjoint of Galois connection. {{qed}} \end{proof}
22448
\section{Supremum of Ideals is Upper Adjoint implies Lattice is Continuous} Tags: Galois Connections, Continuous Lattices \begin{theorem} Let $L = \struct {S, \vee, \wedge, \preceq}$ be a bounded below up-complete lattice. Let $\map {\mathit {Ids} } L$ be the set of all ideals in $L$. Let $P = \struct {\map {\mathit {Ids} } L, \precsim}$ be an ordered set where $\mathord \precsim = \subseteq \restriction_{\map {\mathit {Ids} } L \times \map {\mathit {Ids} } L}$ Let $f: \map {\mathit {Ids} } L \to S$ be a mapping such that :$\forall I \in \map {\mathit {Ids} } L: f \sqbrk I = \sup I$ Let $f$ be an upper adjoint of a Galois connection. Then $L$ is continuous. \end{theorem} \begin{proof} We will prove that :$\forall x \in S: \exists I \in \map {\mathit {Ids} } L: x \preceq \sup I \land \forall J \in \map {\mathit {Ids} } L: x \preceq \sup J \implies I \subseteq J$ Let $x \in S$. Define $I := \map \inf {f^{-1} \sqbrk {x^\succeq} }$. By definition of $P$: :$I \in \map {\mathit {Ids} } L$ We will prove that :$\forall J \in \map {\mathit {Ids} } L: x \preceq \sup J \implies I \subseteq J$ Let $J \in \map {\mathit {Ids} } L$ such that :$x \preceq \sup J$ By definition of $f$: :$x \preceq f \sqbrk J$ By definition of upper closure of element: :$f \sqbrk J \in x^\succeq$ By definition of image of set: :$J \in f^{-1} \sqbrk {x^\succeq}$ By definition of infimum: :$I \precsim J$ Hence by definition of $\precsim$: :$I \subseteq J$ {{qed|lemma}} By definition of upper adjoint of a Galois connection: :there exists a mapping $d: S \to \map {\mathit {Ids} } L$: $\struct {f, d}$ is a Galois connection. By Galois Connection is Expressed by Minimum :$\map d x = \map \min {f^{-1} \sqbrk {x^\succeq} }$ By definition of smallest element: :$I \in f^{-1} \sqbrk {x^\succeq}$ By definition of image of set: :$f \sqbrk I \in x^\succeq$ By definition of upper closure of element: :$x \preceq f \sqbrk I$ Thus by definition of $f$: :$x \preceq \sup I$ Hence :$\exists I \in \map {\mathit {Ids} } L: x \preceq \sup I \land \forall J \in \map {\mathit {Ids} } L: x \preceq \sup J \implies I \subseteq J$ {{qed|lemma}} Hence by Continuous iff For Every Element There Exists Ideal Element Precedes Supremum: :$L$ is continuous. {{qed}} \end{proof}
22449
\section{Supremum of Lower Closure of Set} Tags: Lower Closures, Order Theory \begin{theorem} Let $\left({S, \preceq}\right)$ be an ordered set. Let $T \subseteq S$. Let $L = T^\preceq$ be the lower closure of $T$ in $S$. Let $s \in S$ Then $s$ is the supremum of $T$ {{iff}} it is the supremum of $L$. \end{theorem} \begin{proof} By Supremum and Infimum are Unique we need only show that $s$ is a supremum of $L$ {{iff}} it is a supremum of $T$. \end{proof}
22450
\section{Supremum of Lower Sums Never Greater than Upper Sum} Tags: Real Analysis \begin{theorem} Let $\closedint a b$ be a closed real interval . Let $f$ be a bounded real function defined on $\closedint a b$. Let $S$ be a finite subdivision of $\closedint a b$. Let $\map U S$ be the upper sum of $f$ on $\closedint a b$ with respect to $S$. Let $\map L P$ be the lower sum of $f$ on $\closedint a b$ with respect to a finite subdivision $P$. Then: :$\sup_P \map L P \le \map U S$ \end{theorem} \begin{proof} From Upper Sum Never Smaller than Lower Sum for any Pair of Subdivisions, $\map U S$ is an upper bound for the real set: :$T = \leftset {\map L P: P}$ is a finite subdivision of $\rightset {\closedint a b}$ Since $\sup_P \map L P$ is the supremum of $T$: :$\sup_P \map L P \le \map U S$ Hence the result. {{qed}} Category:Real Analysis \end{proof}
22451
\section{Supremum of Meet Image of Directed Set} Tags: Up-Complete Semilattices \begin{theorem} Let $\struct {S, \preceq}$ be an up-complete meet semilattice. Let $f: S \times S \to S$ be a mapping such that: :$\forall \tuple {x, y} \in S \times S: \map f {x, y} = x \wedge y$ Let $D$ be directed subset of $S \times S$ in the simple order product $\struct {S \times S, \precsim}$ of $\struct {S, \preceq}$ and $\struct {S, \preceq}$. Then: :$\map \sup {\map {f^\to} D} = \sup \set {x \wedge y: x \in \map {\pr_1^\to} D, y \in \map {\pr_2^\to} D}$ where :$\pr_1$ denotes the first projection on $S \times S$ :$\pr_2$ denotes the second projection on $S \times S$ :$\map {\pr_1^\to} D$ denotes the image of $D$ under $\pr_1$. \end{theorem} \begin{proof} By definition of image of set: :$\map {f^\to} D = \set {x \wedge y: \tuple {x, y} \in D}$ By definition of subset: :$\map {f^\to} D \subseteq \set {x \wedge y: x \in \map {\pr_1^\to} D, y \in \map {\pr_2^\to} D}$ By Up-Complete Product/Lemma 2: :$D_1 := \map {\pr_1^\to} D$ is directed and :$D_2 := \map {\pr_2^\to} D$ is directed. By Meet of Directed Subsets is Directed: :$\set {x \wedge y: x \in D_1, y \in D_2}$ is directed. By definition of up-complete: :$\set {x \wedge y: x \in D_1, y \in D_2}$ admits a supremum. By Meet is Increasing: :$f$ is increasing mapping. By Image of Directed Subset under Increasing Mapping is Directed: :$\map {f^\to} D$ is directed. By definition of up-complete: :$\map {f^\to} D$ admits a supremum. By Supremum of Subset: :$\map \sup {\map {f^\to} D} \preceq \sup \set {x \wedge y: x \in D_1, y \in D_2}$ We will prove that :$\set {x \wedge y: x \in D_1, y \in D_2} \subseteq \paren {\map {f^\to} D}^\preceq$ where :$\paren {\map {f^\to} D}^\preceq$ denotes the lower closure of $\map {f^\to} D$. Let $z \in \set {x \wedge y: x \in D_1, y \in D_2}$. Then :$\exists x \in D_1, y \in D_2: z = x \wedge y$ By definition of image of set: :$\exists \tuple {a, b} \in D: \map {\pr_1} {a, b} = x$ and :$\exists \tuple {c, d} \in D: \map {\pr_2} {c, d} = y$: By definition of first projection and second projection: :$a = x$ and $d = y$ By definition of directed subset: :$\exists \tuple {g, h} \in D: \tuple {x, b} \precsim \tuple {g, h} \land \tuple {c, y} \precsim \tuple {g, h}$ By definition of simple order product: :$x \preceq g$ and $y \preceq h$ By Meet Semilattice is Ordered Structure: :$x \wedge y \preceq g \wedge h \in \set {x \wedge y: \tuple {x, y} \in D}$ Thus by definition of lower closure: :$z \in \paren {\map {f^\to} D}^\preceq$ {{qed|lemma}} By Supremum of Lower Closure of Set: :$\paren {\map {f^\to} D}^\preceq$ admits a supremum and :$\map \sup {\map {f^\to} D}^\preceq = \map \sup {\map {f^\to} D}$ By Supremum of Subset: :$\sup \set {x \wedge y: x \in D_1, y \in D_2} \preceq \map \sup {\map {f^\to} D}$ Thus by definition of antisymmetry: :$\map \sup {\map {f^\to} D} = \sup \set {x \wedge y: x \in D_1, y \in D_2}$ {{qed}} \end{proof}
22452
\section{Supremum of Power Set} Tags: Power Set, Orderings, Order Theory \begin{theorem} Let $S$ be a set. Let $\powerset S$ be the power set of $S$. Let $\struct {\powerset S, \subseteq}$ be the relational structure defined on $\powerset S$ by the relation $\subseteq$. (From Subset Relation on Power Set is Partial Ordering, this is an ordered set.) Then the supremum of $\struct {\powerset S, \subseteq}$ is the set $S$. \end{theorem} \begin{proof} By the definition of the power set: :$\forall X \in \powerset S: X \subseteq S$ The result then follows from the definition of supremum. {{qed}} Category:Power Set Category:Order Theory \end{proof}
22453
\section{Supremum of Product} Tags: Max and Min Operations, Suprema \begin{theorem} Let $\struct {G, \circ, \preceq}$ be an ordered group. Suppose that subsets $A$ and $B$ of $G$ admit suprema in $G$. Then: :$\sup \paren {A \circ_\PP B} = \sup A \circ \sup B$ where $\circ_\PP$ denotes subset product. \end{theorem} \begin{proof} Let $a \in A$, $b \in B$. Then: {{begin-eqn}} {{eqn | l = a \circ b | o = \preceq | r = \sup A \circ b | c = {{Defof|Supremum of Set}} }} {{eqn | o = \preceq | r = \sup A \circ \sup B | c = {{Defof|Supremum of Set}} }} {{end-eqn}} Hence $\sup A \circ \sup B$ is an upper bound for $A \circ_\PP B$. Suppose that $u$ is an upper bound for $A \circ_\PP B$. Then: {{begin-eqn}} {{eqn | q = \forall b \in B: \forall a \in A | l = a \circ b | o = \preceq | r = u }} {{eqn | ll= \leadsto | q = \forall b \in B: \forall a \in A | l = a | o = \preceq | r = u \circ b^{-1} }} {{eqn | ll= \leadsto | q = \forall b \in B | l = \sup A | o = \preceq | r = u \circ b^{-1} | c = {{Defof|Supremum of Set}} }} {{eqn | ll= \leadsto | q = \forall b \in B | l = b | o = \preceq | r = \paren {\sup A}^{-1} \circ u }} {{eqn | ll= \leadsto | l = \sup B | o = \preceq | r = \paren {\sup A}^{-1} \circ u | c = {{Defof|Supremum of Set}} }} {{eqn | ll= \leadsto | l = \sup A \circ \sup B | o = \preceq | r = u }} {{end-eqn}} Therefore: :$\sup \paren {A \circ_\PP B} = \sup A \circ \sup B$ {{qed}} \end{proof}
22454
\section{Supremum of Set of Integers equals Greatest Element} Tags: Integers \begin{theorem} Let $S \subset \Z$ be a non-empty subset of the set of integers. Let $S$ be bounded above in the set of real numbers $\R$. Then $S$ has a greatest element, and it is equal to the supremum $\sup S$. \end{theorem} \begin{proof} By Set of Integers Bounded Above by Real Number has Greatest Element, $S$ has a greatest element, say $n \in S$. By Greatest Element is Supremum, $n$ is the supremum of $S$. {{qed}} \end{proof}
22455
\section{Supremum of Set of Integers is Integer} Tags: Integers \begin{theorem} Let $S \subset \Z$ be a non-empty subset of the set of integers. Let $S$ be bounded above in the set of real numbers. Then its supremum $\sup S$ is an integer. \end{theorem} \begin{proof} By Supremum of Set of Integers equals Greatest Element, $S$ has a greatest element $n \in \Z$, that is equals to the supremum of $S$. {{qed}} \end{proof}
22456
\section{Supremum of Set of Real Numbers is at least Supremum of Subset} Tags: Supremum of Set of Real Numbers is at least Supremum of Subset, Supremum of Subset is Less Than or Equal to Supremum of Set, Real Analysis \begin{theorem} Let $S$ be a set of real numbers. Let $S$ have a supremum. Let $T$ be a non-empty subset of $S$. Then $\sup T$ exists and: :$\sup T \le \sup S$ \end{theorem} \begin{proof} The number $\sup S$ is an upper bound for $S$. Therefore, $\sup S$ is an upper bound for $T$ as $T$ is a non-empty subset of $S$. Accordingly, $T$ has a supremum by the continuum property. The number $\sup S$ is an upper bound for $T$. Therefore, $\sup S$ is greater than or equal to $\sup T$ as $\sup T$ is the least upper bound of $T$. {{qed}} Category:Real Analysis 239707 239703 2015-12-01T20:29:22Z Prime.mover 59 239707 wikitext text/x-wiki \end{proof}
22457
\section{Supremum of Simple Order Product} Tags: Simple Order Product, Order Theory, Suprema \begin{theorem} Let $\struct {S_1, \preceq_1}$ and $\struct {S_2, \preceq_2}$ be ordered sets. Let $\struct {S_1 \times S_2, \precsim}$ be the simple order product of $\struct {S_1, \preceq_1}$ and $\struct {S_2, \preceq_2}$. Let $X_1$ be a non-empty subset of $S_1$, $X_2$ be a non-empty subset of $S_2$ such that :$X_1$ and $X_2$ admit suprema. Then: :$X_1 \times X_2$ admits a supremum and: :$\map \sup {X_1 \times X_2} = \tuple {\sup X_1, \sup X_2}$ \end{theorem} \begin{proof} We will prove that: :$\tuple {\sup X_1, \sup X_2}$ is upper bound for $X_1 \times X_2$ Let $\tuple {a, b} \in X_1 \times X_2$. By definition of Cartesian product: :$a \in X_1$ and $b \in X_2$ By definitions of supremum and upper bound: :$a \preceq_1 \sup X_1$ and $b \preceq_2 \sup X_2$ Thus by definition of simple order product: :$\tuple {a, b} \precsim \tuple {\sup X_1, \sup X_2}$ {{qed|lemma}} We will prove that: :$\forall \tuple {a, b} \in S1 \times S_2: \tuple {a, b}$ is upper bound for $X_1 \times X_2 \implies \tuple {\sup X_1, \sup X_2} \precsim \tuple {a, b}$ Let $\tuple {a, b} \in S1 \times S_2$ such that: :$\tuple {a, b}$ is upper bound for $X_1 \times X_2$ We will prove as sublemma that: :$a$ is upper bound for $X_1$ Let $c \in X_1$. By definition of non-empty set: :$\exists d: d \in X_2$ By definition of Cartesian product: :$\tuple {c, d} \in X_1 \times X_2$ By definition of upper bound: :$\tuple {c, d} \precsim \tuple {a, b}$ Thus by definition of simple order product: :$c \preceq_1 a$ This ends the proof of sublemma. Analogically we have that: :$b$ is upper bound for $X_2$ By definition of supremum: :$\sup X_1 \preceq_1 a$ and $\sup X_2 \preceq_2 b$ Thus by definition of simple order product: :$\tuple {\sup X_1, \sup X_2} \precsim \tuple {a, b}$ {{qed|lemma}} Thus by definition: :$X_1 \times X_2$ admits a supremum and: :$\map \sup {X_1 \times X_2} = \tuple {\sup X_1, \sup X_2}$ {{qed}} \end{proof}
22458
\section{Supremum of Singleton} Tags: Singletons, Order Theory, Suprema \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Then for all $a \in S$: :$\sup \set a = a$ where $\sup$ denotes supremum. \end{theorem} \begin{proof} Since $a \preceq a$, $a$ is an upper bound of $\set a$. Let $b$ be another upper bound of $\set a$. Then necessarily $a \preceq b$. It follows that indeed: :$\sup \set a = a$ as desired. {{qed}} \end{proof}
22459
\section{Supremum of Subgroups in Lattice} Tags: Lattice Theory, Subgroups, Suprema, Complete Lattices \begin{theorem} Let $\struct {G, \circ}$ be a group. Let $\mathbb G$ be the set of all subgroups of $G$. Let $\struct {\mathbb G, \subseteq}$ be the complete lattice formed by $\mathbb G$ and $\subseteq$. Let $H, K \in \mathbb G$. Let either $H$ or $K$ be normal in $G$. Then: :$\sup \set {H, K} = H \circ K$ where $H \circ K$ denotes subset product. \end{theorem} \begin{proof} Recall that Set of Subgroups forms Complete Lattice. Let $L = \sup \set {H, K}$. Let either $H$ or $K$ be normal in $G$. Since $L$ contains $H$ and $K$, then $L$ contains $H \circ K$. The smallest subgroup of $G$ containing $H$ and $K$ is: :$\gen {H, K}$ the subgroup generated by $H$ and $K$. From Subset Product with Normal Subgroup as Generator: :$\gen {H, K} = H \circ K$ when either $H$ or $K$ is normal. The result follows. {{qed}} \end{proof}
22460
\section{Supremum of Subset} Tags: Orderings, Order Theory \begin{theorem} Let $\left({U, \preceq}\right)$ be an ordered set. Let $S \subseteq U$. Let $T \subseteq S$. Let $S$ admit a supremum (in $U$). If $T$ also admits a supremum (in $U$), then $\sup \left({T}\right) \preceq\sup \left({S}\right)$. \end{theorem} \begin{proof} Let $B = \sup \left({S}\right)$. Then $B$ is an upper bound for $S$. As $T \subseteq S$, it follows by the definition of a subset that $x \in T \implies x \in S$. Because $x \in S \implies x \preceq B$ (as $B$ is an upper bound for $S$) it follows that $x \in T \implies x \preceq B$. So $B$ is an upper bound for $T$. Therefore $B$ succeeds the supremum of $T$ in $S$. Hence the result. {{qed}} Category:Order Theory \end{proof}
22461
\section{Supremum of Subset of Bounded Above Set of Real Numbers} Tags: Boundedness \begin{theorem} Let $A$ and $B$ be sets of real numbers such that $A \subseteq B$. Let $B$ be bounded above. Then: :$\sup A \le \sup B$ where $\sup$ denotes the supremum. \end{theorem} \begin{proof} Let $B$ be bounded above. By the Continuum Property, $B$ admits a supremum. By Subset of Bounded Above Set is Bounded Above, $A$ is also bounded above. Hence also by the Continuum Property, $A$ also admits a supremum. {{AimForCont}} $\sup A > \sup B$. Then: :$\exists y \in A: y > \sup B$ Thus by definition of supremum, $y \notin B$. That is: :$A \nsubseteq B$ which contradicts our initial assumption that $A \subseteq B$. Hence the result by Proof by Contradiction. {{qed}} \end{proof}
22462
\section{Supremum of Subset of Real Numbers May or May Not be in Subset} Tags: Suprema \begin{theorem} Let $S \subset \R$ be a proper subset of the set $\R$ of real numbers. Let $S$ admit a supremum $M$. Then $M$ may or may not be an element of $S$. \end{theorem} \begin{proof} Consider the subset $S$ of the real numbers $\R$ defined as: :$S = \set {\dfrac 1 n: n \in \Z_{>0} }$ It is seen that: :$S = \set {1, \dfrac 1 2, \dfrac 1 3, \ldots}$ and hence $\sup S = 1$. Thus $\sup S \in S$. Consider the subset $T$ of the real numbers $\R$ defined as: :$T = \set {-\dfrac 1 n: n \in \Z_{>0} }$ It is seen that: :$T = \set {-1, -\dfrac 1 2, -\dfrac 1 3, \ldots}$ and hence $\sup T = 0$. Thus $\sup T \notin T$. {{Qed}} \end{proof}
22463
\section{Supremum of Subset of Real Numbers is Arbitrarily Close} Tags: Real Analysis \begin{theorem} Let $A \subseteq \R$ be a subset of the real numbers. Let $b$ be a supremum of $A$. Let $\epsilon \in \R_{>0}$. Then: :$\exists x \in A: b − x < \epsilon$ \end{theorem} \begin{proof} Note that $A$ is non-empty as the empty set does not admit a supremum (in $\R$). Suppose $\epsilon \in \R_{>0}$ such that: :$\forall x \in A: b − x \ge \epsilon$ Then: :$\forall x \in A: b − \epsilon \ge x$ and so $b − \epsilon$ would be an upper bound of $A$ which is less than $b$. But since $b$ is a supremum of $A$ there can be no such $b − \epsilon$. From that contradiction it follows that: :$\exists x \in A: b − x < \epsilon$ {{qed}} \end{proof}
22464
\section{Supremum of Subset of Union Equals Supremum of Union} Tags: Real Analysis \begin{theorem} Let $S$ be a non-empty real set. Let $S$ have a supremum. Let $\set {S_i: i \in \set {1, 2, \ldots, n} }$, $n \in \N_{>0}$, be a set of non-empty subsets of $S$. Let $\bigcup S_i = S$. Then there exists a $j$ in $\set {1, 2, \ldots, n}$ such that: :$\sup S_j = \sup S$ \end{theorem} \begin{proof} If $S$ equals $S_j$ for a $j$ in $\set {1, 2, \ldots, n}$, it is trivially true that $\sup S = \sup S_j$. Now assume that $S$ is unequal to $S_i$ for every $i$ in $\left\{{1, 2, \ldots, n}\right\}$. By Supremum of Set of Real Numbers is at least Supremum of Subset, $\sup S \ge \sup S_i$ for every $i$ in $\set{1, 2, \ldots, n}$. There are two alternatives; either: :$\sup S > \sup S_i$ for every $i$ in $\set {1, 2, \ldots, n}$ or: :$\sup S = \sup S_j$ for at least one $j$ in $\set {1, 2, \ldots, n}$. Suppose that: :$\sup S > \sup S_i$ for every $i$ in $\set {1, 2, \ldots, n}$ Let $\epsilon = \sup S - \map \max {\sup S_1, \sup S_2, \ldots, \sup S_n}$. We note that $\epsilon > 0$. By Supremum of Subset of Real Numbers is Arbitrarily Close, $S$ has an element $x$ that satisfies: :$x > \sup S - \epsilon$ We have: {{begin-eqn}} {{eqn | o = > | l = x | r = \sup S - \epsilon }} {{eqn | r = \sup S - \paren {\sup S - \map \max {\sup S_1, \sup S_2, \ldots, \sup S_n} } | c = definition of $\epsilon$ }} {{eqn | r = \map \max {\sup S_1, \sup S_2, \ldots, \sup S_n} }} {{end-eqn}} Therefore: :$x > \map \max {\sup S_1, \sup S_2, \ldots, \sup S_n}$ This means that $x > \sup S_i$ for every $i$ in $\set {1, 2, \ldots, n}$. However, $x$ must be an element of $S_j$ for some $j$ in $\set {1, 2, \ldots, n}$ as $x \in S$ and $S = \bigcup S_i$. Accordingly, it is not true that $\sup S > \sup S_i$ for every $i$ in $\set {1, 2, \ldots, n}$. We just concluded that the alternative: :$\sup S > \sup S_i$ for every $i$ in $\set {1, 2, \ldots, n}$ is not true. Therefore, the other alternative: :$\sup S = \sup S_j$ for a $j$ in $\set {1, 2, \ldots, n}$ is true. {{qed}} Category:Real Analysis \end{proof}
22465
\section{Supremum of Sum equals Sum of Suprema} Tags: Real Analysis \begin{theorem} Let $A$ and $B$ be non-empty sets of real numbers. Let $A + B$ be $\set {x + y: x \in A, y \in B}$. Let either $A$ and $B$ have suprema or $A + B$ have a supremum. Then all $\sup A$, $\sup B$, and $\sup \paren {A + B}$ exist and: :$\sup \paren {A + B} = \sup A + \sup B$ \end{theorem} \begin{proof} Assume first that $A$ and $B$ have suprema. We have: :$x \le \sup A$ for an arbitrary $x$ in $A$ :$y \le \sup B$ for an arbitrary $y$ in $B$ Adding these inequalities, we get: :$x + y \le \sup A + \sup B$ The number $x + y$ is an arbitrary element of $A + B$ as $x$ and $y$ are arbitrary elements of $A$ and $B$ respectively. Therefore, $\sup A + \sup B$ is an upper bound for $A + B$. $A + B$ is non-empty as $A$ and $B$ are non-empty. Accordingly, $A + B$ has a supremum by the Continuum Property. Next, assume that $\sup \paren {A + B}$ has a supremum. We need to prove that $A$ and $B$ have suprema. Let $y$ be a point in $B$. We have: {{begin-eqn}} {{eqn | l = x + y | o = \le | r = \sup \paren {A + B} | c = for every $x$ in $A$ as $x + y$ is a point in $A + B$ }} {{eqn | ll= \leadstoandfrom | l = x | o = \le | r = \sup \paren {A + B} - y }} {{end-eqn}} Therefore, $A$ has an upper bound as $x$ is an arbitrary point in $A$. Also, we have that $A$ is non-empty. Accordingly, $A$ has a supremum by the Continuum Property. A similar argument gives that $B$ has a supremum. So, we have shown that all $\sup A$, $\sup B$, and $\sup \paren {A + B}$ exist. We proceed to show that $\sup \paren {A + B} = \sup A + \sup B$. We have $\sup \paren {A + B} \le \sup A + \sup B$ as $\sup A + \sup B$ is an upper bound for $A + B$. Accordingly, either: :$\sup \paren {A + B} < \sup A + \sup B$ or: :$\sup \paren {A + B} = \sup A + \sup B$. {{AimForCont}}: :$\sup \paren {A + B} < \sup A + \sup B$. Let $\epsilon = \sup A + \sup B - \sup \paren {A + B}$. We note that $\epsilon > 0$. Since $\sup A$ is the least upper bound of $A$, there is an element $x$ in $A$ such that: :$x > \sup A - \dfrac \epsilon 2$ by Supremum of Subset of Real Numbers is Arbitrarily Close Since $\sup B$ is the least upper bound of $B$, there is an element $y$ in $B$ such that: :$y > \sup B - \dfrac \epsilon 2$ by Supremum of Subset of Real Numbers is Arbitrarily Close Adding these inequalities, we get: {{begin-eqn}} {{eqn | l = x + y | o = > | r = \sup A - \frac \epsilon 2 + \sup B - \frac \epsilon 2 }} {{eqn | ll= \leadstoandfrom | l = x + y | o = > | r = \sup A + \sup B - \epsilon }} {{eqn | ll= \leadstoandfrom | l = x + y | o = > | r = \sup A + \sup B - \paren {\sup A + \sup B - \sup \paren {A + B} } | c = definition of $\epsilon$ }} {{eqn | ll= \leadstoandfrom | l = x + y | o = > | r = \sup \paren {A + B} }} {{end-eqn}} which is impossible since the number $x + y$ is an element of $A + B$ as $x \in A$ and $y \in B$. We have found that: :$\sup \paren {A + B} < \sup A + \sup B$ is not true. Therefore: :$\sup \paren {A + B} = \sup A + \sup B$ as $\sup \paren {A + B} \le \sup A + \sup B$. {{qed}} Category:Real Analysis \end{proof}
22466
\section{Supremum of Suprema} Tags: Order Theory \begin{theorem} Let $\struct {S, \preceq}$ be an ordered set. Let $\mathbb T \subseteq \powerset S$, where $\powerset S$ is the power set of $S$. Suppose all $T \in \mathbb T$ admit a supremum $\sup T$ in $S$. Then: :$\sup \bigcup \mathbb T = \sup {\set {\sup T: T \in \mathbb T} }$ if one of these two quantities exists (in $S$). \end{theorem} \begin{proof} Suppose that $s = \sup \bigcup \mathbb T \in S$. By Set is Subset of Union, $T \subseteq \bigcup \mathbb T$ for all $T \in \mathbb T$. Hence by Supremum of Subset: :$\forall T \in \mathbb T: \sup T \preceq s$ Suppose now that $a \in S$ satisfies: :$\forall T \in \mathbb T: \sup T \preceq a$ Then by transitivity of $\preceq$: :$\forall t \in T: t \preceq a$ Since this holds for any $T \in \mathbb T$, also: :$\forall t \in \bigcup \mathbb T: t \preceq a$ Hence $s \preceq a$, by definition of supremum. That is, $s = \sup {\set {\sup T: T \in \mathbb T} }$. {{qed|lemma}} Suppose now that $r = \sup {\set {\sup T: T \in \mathbb T} } \in S$. By definition of supremum, for all $T \in \mathbb T$ and $t \in T$: :$t \preceq \sup T$ By transitivity of $\preceq$: :$\forall T \in \mathbb T: \forall t \in T: t \preceq r$ Hence for all $t \in \bigcup \mathbb T$: :$t \preceq r$ Suppose that $a \in S$ satisfies: :$\forall t \in \bigcup \mathbb T: t \preceq a$ In particular, for any $T \in \mathbb T$, since $T \subseteq \bigcup \mathbb T$: :$\sup T \preceq a$ and therefore by definition of supremum, also: :$r \preceq a$ That is, $r = \sup \bigcup \mathbb T$. {{qed}} Category:Order Theory \end{proof}
22467
\section{Supremum of Union of Bounded Above Sets of Real Numbers} Tags: Set Union, Boundedness, Suprema \begin{theorem} Let $A$ and $B$ be sets of real numbers. Let $A$ and $B$ both be bounded above. Then: :$\map \sup {A \cup B} = \max \set {\sup A, \sup B}$ where $\sup$ denotes the supremum. \end{theorem} \begin{proof} Let $A$ and $B$ both be bounded above. By the Continuum Property, $A$ and $B$ both admit a supremum. Let $x \in A \cup B$. Then either $x \le \sup A$ or $x \le \sup B$ by definition of supremum. Hence: :$x \le \max \set {\sup A, \sup B}$ and so $\max \set {\sup A, \sup B}$ is certainly an upper bound of $A \cup B$. It remains to be shown that $\max \set {\sup A, \sup B}$ is the smallest upper bound. {{AimForCont}} there exists $m \in \R$ such that: :$m < \max \set {\sup A, \sup B}$ and: :$\forall x \in A \cup B: x \le m$ {{WLOG}}, let $\sup A \ge \sup B$. Then: :$\max \set {\sup A, \sup B} = \sup A$ and so: :$m < \sup A$ But then by definition of supremum: :$\exists a \in A: a > m$ and so $m$ is not an upper bound of $A \cup B$. This contradicts our assumption that $m$ is a supremum of $A \cup B$. It follows by Proof by Contradiction that $\max \set {\sup A, \sup B}$ is the supremum of $A \cup B$. The same argument shows, mutatis mutandis, that if $\sup A \le \sup B$, $\max \set {\sup A, \sup B}$ is the supremum of $A \cup B$. {{qed}} \end{proof}
22468
\section{Surgery for Rings} Tags: Ring Theory, Ring Isomorphisms, Ring Monomorphisms \begin{theorem} Let $R$ and $S$ be commutative rings with unity, and $\phi: R \to S$ a ring monomorphism. Then there is a ring $T$ isomorphic to $S$ that contains $R$ as a subring. \end{theorem} \begin{proof} Let $T$ be the disjoint union $T = R \cup \paren {S \setminus \Img \phi}$. Define $\theta : T \to S$ as follows: :If $x \in R$, then $\map \theta x = \map \phi x$ :If $x \in \paren {S \setminus \Img \phi}$ then $\map \theta x = x$ We claim that $\theta$ is an isomorphism. ''Injectivity'': Let $\map \theta x = \map \theta y$. Since $\theta \sqbrk R = \phi \sqbrk R$, we have $\theta \sqbrk R \cap \paren {S \setminus \Img \phi} = \O$. Therefore either $x, y \in R$ or $x, y \in \paren {S \setminus \Img \phi}$. If $x, y \in R$ then $\map \phi x = \map \phi y$, so $\theta$ is injective because $\phi$ is. If $x, y \in \paren {S \setminus \Img \phi}$ then $\theta$ acts as the identity and $x = y$. ''Surjectivity'': Under $\theta$, $R$ surjects onto $\Img \phi = \phi \sqbrk R$ by definition. Moreover: :$\theta \sqbrk {S \setminus \Img \phi} = I \sqbrk {S \setminus \Img \phi} = \paren {S \setminus \Img \phi}$ where $I$ denotes the identity mapping. Therefore $\theta$ is surjective everywhere. Now we endow $T$ with addition and multiplication: for $x, y \in T$, let :$x + y = \map {\theta^{-1} } {\map \theta x + \map \theta y}$ :$x y = \map {\theta^{-1} } {\map \theta x \, \map \theta y}$ so trivially we have :$\map \theta {x + y} = \map \theta x + \map \theta y$ :$\map \theta {x y} = \map \theta x \, \map \theta y$ Therefore $T$ is the isomorphic image of a ring, and an isomorphism is precisely a map that preserves the ring axioms. Thus $T$ is a ring isomorphic to $S$ containing $R$ as a subring. {{qed}} {{proof wanted|I wonder whether this page might be no more than an application of the Embedding Theorem.}} Category:Ring Isomorphisms Category:Ring Monomorphisms \end{proof}
22469
\section{Surjection Induced by Powerset is Induced by Surjection} Tags: Power Set, Induced Mappings, Direct Image Mappings, Relation Theory, Named Theorems, Mappings, Surjections, Relations \begin{theorem} Let $\RR \subseteq S \times T$ be a relation. Let $\RR^\to: \powerset S \to \powerset T$ be the direct image mapping $\RR$. Let $\RR^\to$ be a surjection. Let $X = \Preimg \RR$, that is, the preimage of $\RR$. Then $\RR {\restriction_X} \subseteq X \times T$, that is, the restriction of $\RR$ to $X$, is a surjection. \end{theorem} \begin{proof} Let $X$ be the preimage of $\RR$. Suppose $\RR {\restriction_X} \subseteq X \times T$ is a mapping, but not a surjection. Then: :$\exists y \in T: \neg \exists x \in S: \tuple {x, y} \in \RR$ Because no element of $S$ relates to $y$, no subset of $S$ contains any element of $S$ that relates to the subset $\set y \subseteq T$. Thus: :$\exists \set y \in \powerset T: \neg \exists X \in \powerset S: \map {\RR^\to} X = \set y$ So we see that $\RR^\to: \powerset S \to \powerset T$ is not a surjection. Now, suppose $\RR {\restriction_X} \subseteq X \times T$ is not even a mapping. This could happen, according to the definition of a mapping, in one of two ways: :$(1): \quad \exists x \in X: \neg \exists y \in T: \tuple {x, y} \in \RR$ :$(2): \quad \exists x \in S: \tuple {x, y_1} \in \RR \land \tuple {x, y_2} \in \RR: y_1 \ne y_2$ Because $X$ is already the preimage of $\RR$, the first of these cannot happen here. For the second, it can be seen that neither $\set {y_1}$ nor $\set {y_2}$ can be in $\Rng {\map {\RR^\to} {\powerset S} }$. Therefore $\RR^\to: \powerset S \to \powerset T$ can not be a surjection. {{questionable|See talk page}} Thus, by the Rule of Transposition, the result follows. {{qed}} Category:Surjections Category:Direct Image Mappings \end{proof}
22470
\section{Surjection by Free Module} Tags: Homological Algebra \begin{theorem} Let $A$ be a ring. Let $M$ be a left $A$-module. Then there exists a free $A$-module $F$ and a surjective $A$-module homomorphism $f : F \to M$. \end{theorem} \begin{proof} Let $F = A^{\paren M}$ be the free $A$-module on the set $M$. Let $c : M \to A^{\paren M}$ be the canonical mapping. Let $f : F \to M$ be the $A$-module homomorphism induced the by the Universal Property of Free Modules applied to the identity $\operatorname {id}_M$ of $M$. We have :$f \circ c = \operatorname {id}_M$ Thus $f$ is a split epimorphism in the category of sets. By Split Epimorphism is Epic and Surjection iff Epimorphism in Category of Sets $f$ is surjective. {{qed}} Category:Homological Algebra \end{proof}
22471
\section{Surjection from Aleph to Ordinal} Tags: Aleph Function, Aleph Mapping \begin{theorem} Let $x$ and $y$ be ordinals. Suppose that: :$0 < y < \aleph_{x+1}$ Then there is a surjection: :$f : \aleph_x \to y$ \end{theorem} \begin{proof} :$y < \aleph_{x+1}$, then $y < \aleph_x \lor y \sim \aleph_x$ by Ordinal Less than Successor Aleph. In either case, $\left|{ y }\right| \le \aleph_x$ by Ordinal in Aleph iff Cardinal in Aleph and Equivalent Sets have Equal Cardinal Numbers. The existence of the surjection follows from Surjection iff Cardinal Inequality. {{qed}} \end{proof}
22472
\section{Surjection from Class to Proper Class} Tags: Gödel-Bernays Class Theory, Class Mappings \begin{theorem} Let $A$ be a class. Let $\mathrm P$ be a proper class. Let $f: A \to \mathrm P$ be a surjection. Then $A$ is proper. \end{theorem} \begin{proof} {{NotZFC}} {{AimForCont}} $A$ is not proper. Then $A$ must be a set. By the Axiom of Powers, $\powerset A$ is also a set. Let $g: f \sqbrk A \to \powerset A$ be defined as: :$\map g {\map f a} = f^{-1} \sqbrk {\set {\map f a} }$ It should be noted that: :$\forall a \in A: \map {\paren {g \circ f} } a \ne \O$: Suppose that $\map g {\map f a} = \map g {\map f b}$. Let $x \in f^{-1} \sqbrk {\set {\map f a} }$. Then by the definition of a singleton: :$x \in f^{-1} \sqbrk {\set {\map f a} } \implies \map f x \in \set {\map f a} \implies \map f x = \map f a$ By the same argument: :$x \in f^{-1} \sqbrk {\set {\map f b} } \implies \map f x = \map f b$ Hhence by the properties of equality: :$\map f a = \map f b$ It has been shown that $g$ is an injection. Because $f$ is a surjection, it follows by definition of surjection that: :$f \sqbrk A = \mathrm P$ But this contradicts Injection from Proper Class to Class. Thus by contradiction $A$ must be proper. Hence the result. {{qed}} Category:Gödel-Bernays Class Theory Category:Class Mappings \end{proof}