You signed in with another tab or window. Reload to refresh your session.You signed out in another tab or window. Reload to refresh your session.You switched accounts on another tab or window. Reload to refresh your session.Dismiss alert
A group $G$ acts \emph{triply transitively} on a set $X$ if given $x_1, x_2, x_3\in X$ all distinct and $y_1, y_2, y_3\in X$ all distinct there exists $g \in G$ such that $g(x_i) = y_i,\ i = 1, 2, 3$. \\
159
+
A group $G$ acts \emph{sharply triply transitively} if such a $g$ is unique.
160
+
\end{definition}
161
+
162
+
\begin{theorem} \label{thm:16}
163
+
The action of $\mathcal{M}$ on $\mathbb{C}_\infty$ is sharply triply transitive.
164
+
\end{theorem}
165
+
166
+
\begin{proof}
167
+
Label first triple $\{z_0, z_1, z_\infty\}$ and second triple $\{w_0, w_1, w_\infty\}$.
g f^{-1} f' g^{-1} : 0 &\mapsto 0 \implies b = 0 \\
194
+
: 1 &\mapsto 1 \implies a = d \\
195
+
: \infty &\mapsto\infty\implies c = 0 \\
196
+
\implies g f^{-1} f' g^{-1} &= \text{id} \\
197
+
\implies f^{-1} f' &= \text{id} \\
198
+
\implies f &= f'.
199
+
\end{align*}
200
+
So, the image of just three points determines the map
201
+
\end{proof}
202
+
203
+
\subsection{Conjugacy classes in $\mathcal{M}$}
204
+
Recall $\Phi : \operatorname{GL}_2(\mathbb{C}) \twoheadrightarrow\mathcal{M}$ from proof of \Cref{thm:15} ($\twoheadrightarrow$ means its a surjective homomorphism).
205
+
Suppose $A, B$ are conjugate in $\operatorname{GL}_2(\mathbb{C})$, i.e. $\exists\; P \in\operatorname{GL}_2(\mathbb{C})$ s.t. $PAP^{-1} = B$ then
206
+
\begin{align*}
207
+
\Phi(P) \Phi(A) \Phi(P)^{-1} &= \phi(PAP^{-1}) \\
208
+
&= \Phi(B) \in\mathcal{M}
209
+
\end{align*}
210
+
i.e. $\Phi(A)$ and $\Phi(B)$ are conjugate in $\mathcal{M}$.
211
+
212
+
Use knowledge of conjugacy classes in $\operatorname{GL}_2(\mathbb{C})$.
213
+
214
+
\begin{theorem} \label{thm:17}
215
+
Any non-identity M\"obius map is conjugate to $f(z) = \nu z$ for some $\nu\neq0, 1$ or to $f(z) = z + 1$.
Then $\alpha$ is a fixed point of $f$ (i.e. $f(\alpha) = \alpha$) $\iff$$g(\alpha)$ is a fixed point of $h$ (i.e. $h(g(\alpha)) = g(\alpha)$).\footnote{$f(\alpha) = \alpha\iff gf(\alpha) = g(\alpha) \iff h(g(\alpha)) = gfg^{-1}(g(\alpha)) = g(\alpha)$ or $g f = h g \implies h(g(\alpha)) = g(\alpha)$ if $f(\alpha) = \alpha$ and conversely $g(f(\alpha)) = h(g(\alpha)) \implies g(f(\alpha)) = g(\alpha)$.} \\
252
+
So number of fixed points of $f = $ number of fixed points of $h$ as $g$ is a bijection.
253
+
By \Cref{thm:17} either, $f$ conjugate to $z \mapsto\nu z$ which has two fixed points: $0, \infty$; or $f$ conjugate to $z \mapsto z + 1$ which has one fixed point: $\infty$.
254
+
\end{proof}
255
+
256
+
\subsection{Circles in $\mathbb{C}_\infty$}
257
+
258
+
A Euclidean circle is the set of points in $\mathbb{C}$ given by some equation $|z - z_0| = r,\ r > 0$.
259
+
A Euclidean line is the set of points in $\mathbb{C}$ given by some equation $|z - a| = |z - b|,\ a \neq b$.
260
+
261
+
\begin{definition}[Circle in $\mathbb{C}_\infty$]
262
+
A \emph{circle in $\mathbb{C}_\infty$} is either a Euclidean circle of a set $L \cup\{\infty\}$ where $L$ is a Euclidean line.
263
+
Its general equation is of the form
264
+
\begin{align}
265
+
Az\bar z + \bar Bz + B\bar z + C = 0, \label{eq:circle}
266
+
\end{align}
267
+
where $A, C \in\mathbb{R}$ and $|B|^2 > AC$.
268
+
\end{definition}
269
+
270
+
Where $z = \infty$ is a solution iff $A = 0$. \\
271
+
$A = 0$: line \\
272
+
$C = 0$: goes through the origin.
273
+
274
+
There is a unique circle passing through any 3 distinct points in $\mathbb{C}_\infty$.
275
+
276
+
\begin{theorem} \label{thm:18}
277
+
Let $f \in\mathcal{M}$ and $C$ a circle in $\mathbb{C}_\infty$, then $f(C)$ is a circle in $C_\infty$.
278
+
\end{theorem}
279
+
280
+
\begin{proof}
281
+
By \Cref{prp:13}, just need to consider
282
+
$f(z) = az,\ z + b \text{ or } \frac{1}{z}.$
283
+
Let $S_{A, B, C}$ be the circle defined by \Cref{eq:circle}.
284
+
\begin{align*}
285
+
f(z) &= az : S_{A, B, C} \mapsto S_{A / a \bar{a}, B / \bar{a}, C} \\
286
+
f(z) &= z + b: S_{A, B, C} \mapsto S_{A, B - Ab, C + A b \bar{b} - \bar{B}b - B \bar{b}} \\
287
+
f(z) &= \frac{1}{z} = w: S_{A, B, C} \mapsto A + B w + \bar{B} \bar{w} + C w \bar{w} = 0 = S_{C, \bar{B}, A}.
288
+
\end{align*}
289
+
\end{proof}
290
+
291
+
\begin{example}
292
+
Consider the image of $\mathbb{R} \cup\{\infty\}$ (a circle in $\mathbb{C}_\infty$) under
293
+
\begin{align*}
294
+
f(z) &= \frac{z - i}{z + i}.
295
+
\end{align*}
296
+
It is thus a circle in $\mathbb{C}_\infty$ containing $f(0) = -1, f(\infty) = 1, f(1) = -i$ so $f(\mathbb{R} \cup\{\infty\}) = $ unit circle.
297
+
Furthermore, complimentary components are mapped to complementary components.
Copy file name to clipboardExpand all lines: Numbers and Sets/05.rmd
+1-1Lines changed: 1 addition & 1 deletion
Original file line number
Diff line number
Diff line change
@@ -344,7 +344,7 @@ No, e.g. $X'' = X' \cup \mathcal{P}(X') \cup \mathcal{P}(\mathcal{P}(X')) \cup \
344
344
Does every set $Y$ inject into $X, X', X'', \dots$? \
345
345
No, e.g. $Y = X \cup X' \cup X'' \cup \dots$
346
346
347
-
# Panorama
347
+
##Panorama
348
348
349
349
- II Logic and Set Theory - We will explore the cardinalities of infinite sets in more details and will learn about the axiomatic system upon which modern mathematics is built.
Bertrand postulated in 1845 that for every $n \in \mathbb{N}$ there is always a prime between $n$ and $2n$ ($n \leq p < 2n$). \
4
+
The primes $2, 5, 11, 23, 47, 89, 179, 359, 719, 1439, 2879$ show it to be true for $n \leq 2^{11}$.
5
+
Bertrand checked it for $n < 3\,000\,000$.
6
+
Chebychev (1850) gave a proof.
7
+
Erd\H{o}s (1932) gave an elementary proof based on the properties of $\binom{2n}{n}$.
8
+
9
+
::: {.lemma #sone}
10
+
\begin{align*}
11
+
\binom{2n}{n} \geq \frac{2^{2n}}{2n + 1}.
12
+
\end{align*}
13
+
:::
14
+
::: {.proof}
15
+
Since $\binom{n}{k + 1} / \binom{n}{k} = \frac{n - k}{k + 1}$, it is evident that $\binom{n}{k}$ increases for $k < \frac{n}{2}$, and decreases for $k > \frac{n}{2}$.
16
+
In particular, $\binom{2n}{n} \geq \frac{2^{2n}}{2n + 1}$, the maximum element is at least as big as the average ($2^{2n}$ is the sum and we have $2n + 1$ elements).
17
+
:::
18
+
19
+
::: {.lemma #stwo}
20
+
If $p \leq n$ is a prime dividing $\binom{2n}{n}$, then $p \leq \frac{2n}{3}$.
21
+
:::
22
+
::: {.proof}
23
+
Suppose $\frac{2n}{3} < p \leq n$, then $p \leq n < \frac{3}{2}p < 2p \leq 2n < 3p$, so the numerator and denominator of
are divisible by exactly one copy of $p$. ⨳ as it can then not divide it.
28
+
:::
29
+
30
+
::: {.lemma #sthree}
31
+
If $p$ is a prime and $p^k \mid \binom{2n}{n}$, then $p^k \leq 2n$.
32
+
:::
33
+
::: {.proof}
34
+
The greatest power of $p$ dividing $n! = n (n-1) \dots 3 \cdot 2 \cdot 1$.
35
+
$\left \lfloor \frac{n}{p} \right \rfloor$ is the no. of multiples of $p$ upto $n$, $\left \lfloor \frac{n}{p^2} \right \rfloor$ is the no. of multiples of $p^2$ upto $n$.
36
+
So the greatest power is $\left \lfloor \frac{n}{p} \right \rfloor +\left \lfloor \frac{n}{p^2} \right \rfloor +\left \lfloor \frac{n}{p^3} \right \rfloor \dots = \sum_{i \geq 1}\left \lfloor \frac{n}{p^i} \right \rfloor$.
37
+
:::
38
+
Hence, if $k$ is a power of $p$ dividing $\binom{2n}{n} = \frac{2n !}{(n!)^2}$, then
0 commit comments