Datasets:
AI4M
/

id
stringlengths
17
29
formal
stringlengths
51
351
src_header
stringclasses
10 values
informal
stringlengths
3
474
nl_proof
stringlengths
0
3.93k
Pugh|exercise_2_92
theorem exercise_2_92 {Ξ± : Type*} [topological_space Ξ±] {s : β„• β†’ set Ξ±} (hs : βˆ€ i, is_compact (s i)) (hs : βˆ€ i, (s i).nonempty) (hs : βˆ€ i, (s i) βŠƒ (s (i + 1))) : (β‹‚ i, s i).nonempty :=
import .common open set real filter function ring_hom topological_space open_locale big_operators open_locale filter open_locale topology noncomputable theory
Give a direct proof that the nested decreasing intersection of nonempty covering compact sets is nonempty.
\begin{proof} Let $$ A_1 \supset A_2 \supset \cdots \supset A_n \supset \cdots $$ be a nested decreasing sequence of compacts. Suppose that $\bigcap A_n=\emptyset$. Take $U_n=A_n^c$, then $$ \bigcup U_n=\bigcup A_n^c=\left(\bigcap A_n\right)^c=A_1 . $$ Here, I'm thinking of $A_1$ as the main metric space. Since $\left\{U_n\right\}$ is an open covering of $A_1$, we can extract a finite subcovering, that is, $$ A_{\alpha_1}^c \cup A_{\alpha_2}^c \cup \cdots \cup A_{\alpha_m}^c \supset A_1 $$ or $$ \left(A_1 \backslash A_{\alpha_1}\right) \cup\left(A_1 \backslash A_{\alpha_2}\right) \cup \cdots \cup\left(A_1 \backslash A_{\alpha_m}\right) \supset A_1 . $$ But, this is true only if $A_{\alpha_i}=\emptyset$ for some $i$, a contradiction. \end{proof}
Pugh|exercise_3_1
theorem exercise_3_1 {f : ℝ β†’ ℝ} (hf : βˆ€ x y, |f x - f y| ≀ |x - y| ^ 2) : βˆƒ c, f = Ξ» x, c :=
import .common open set real filter function ring_hom topological_space open_locale big_operators open_locale filter open_locale topology noncomputable theory
Assume that $f \colon \mathbb{R} \rightarrow \mathbb{R}$ satisfies $|f(t)-f(x)| \leq|t-x|^{2}$ for all $t, x$. Prove that $f$ is constant.
\begin{proof} We have $|f(t)-f(x)| \leq|t-x|^2, \forall t, x \in \mathbb{R}$. Fix $x \in \mathbb{R}$ and let $t \neq x$. Then $$ \left|\frac{f(t)-f(x)}{t-x}\right| \leq|t-x| \text {, hence } \lim _{t \rightarrow x}\left|\frac{f(t)-f(x)}{t-x}\right|=0 \text {, } $$ so $f$ is differentiable in $\mathbb{R}$ and $f^{\prime}=0$. This implies that $f$ is constant, as seen in class. \end{proof}
Pugh|exercise_3_63a
theorem exercise_3_63a (p : ℝ) (f : β„• β†’ ℝ) (hp : p > 1) (h : f = Ξ» k, (1 : ℝ) / (k * (log k) ^ p)) : βˆƒ l, tendsto f at_top (𝓝 l) :=
import .common open set real filter function ring_hom topological_space open_locale big_operators open_locale filter open_locale topology noncomputable theory
Prove that $\sum 1/k(\log(k))^p$ converges when $p > 1$.
\begin{proof} Using the integral test, for a set $a$, we see $$ \lim _{b \rightarrow \infty} \int_a^b \frac{1}{x \log (x)^c} d x=\lim _{b \rightarrow \infty}\left(\frac{\log (b)^{1-c}}{1-c}-\frac{\log (a)^{1-c}}{1-c}\right) $$ which goes to infinity if $c \leq 1$ and converges if $c>1$. Thus, $$ \sum_{n=2}^{\infty} \frac{1}{n \log (n)^c} $$ converges if and only if $c>1$. \end{proof}
Pugh|exercise_4_15a
theorem exercise_4_15a {Ξ± : Type*} (a b : ℝ) (F : set (ℝ β†’ ℝ)) : (βˆ€ (x : ℝ) (Ξ΅ > 0), βˆƒ (U ∈ (𝓝 x)), (βˆ€ (y z ∈ U) (f : ℝ β†’ ℝ), f ∈ F β†’ (dist (f y) (f z) < Ξ΅))) ↔ βˆƒ (ΞΌ : ℝ β†’ ℝ), βˆ€ (x : ℝ), (0 : ℝ) ≀ ΞΌ x ∧ tendsto ΞΌ (𝓝 0) (𝓝 0) ∧ (βˆ€ (s t : ℝ) (f : ℝ β†’ ℝ), f ∈ F β†’ |(f s) - (f t)| ≀ ΞΌ (|s - t|)) :=
import .common open set real filter function ring_hom topological_space open_locale big_operators open_locale filter open_locale topology noncomputable theory
A continuous, strictly increasing function $\mu \colon (0, \infty) \rightarrow (0, \infty)$ is a modulus of continuity if $\mu(s) \rightarrow 0$ as $s \rightarrow 0$. A function $f \colon [a, b] \rightarrow \mathbb{R}$ has modulus of continuity $\mu$ if $|f(s) - f(t)| \leq \mu(|s - t|)$ for all $s, t \in [a, b]$. Prove that a function is uniformly continuous if and only if it has a modulus of continuity.
\begin{proof} Suppose there exists a modulus of continuity $w$ for $f$, then fix $\varepsilon>0$, since $\lim _{s \rightarrow 0} w(s)=0$, there exists $\delta>0$ such that for any $|s|<\delta$, we have $w(s)<\varepsilon$, then we have for any $x, z \in X$ such that $d_X(x, z)<\delta$, we have $d_Y(f(x), f(z)) \leq w\left(d_X(x, z)\right)<\varepsilon$, which means $f$ is uniformly continuous. Suppose $f:\left(X, d_X\right) \rightarrow\left(Y, d_Y\right)$ is uniformly continuous. Let $\delta_1>0$ be such that $d_X(a, b)<\delta_1$ implies $d_Y(f(a), f(b))<1$. Define $w:[0, \infty) \rightarrow[0, \infty]$ by $$ w(s)= \begin{cases}\left.\sup \left\{d_Y(f(a), f(b))\right\} \mid d_X(a, b) \leq s\right\} & \text { if } s \leq \delta_1 \\ \infty & \text { if } s>\delta_1\end{cases} $$ We'll show that $w$ is a modulus of continuity for $f \ldots$ By definition of $w$, it's immediate that $w(0)=0$ and it's clear that $$ d_Y(f(a), f(b)) \leq w\left(d_X(a, b)\right) $$ for all $a, b \in X$. It remains to show $\lim _{s \rightarrow 0^{+}} w(s)=0$. It's easily seen that $w$ is nonnegative and non-decreasing, hence $\lim _{s \rightarrow 0^{+}}=L$ for some $L \geq 0$, where $L=\inf w((0, \infty))$ Let $\epsilon>0$. By uniform continuity of $f$, there exists $\delta>0$ such that $d_X(a, b)<\delta$ implies $d_Y(f(a), f(b))<\epsilon$, hence by definition of $w$, we get $w(\delta) \leq \epsilon$. Thus $L \leq \epsilon$ for all $\epsilon>0$, hence $L=0$. This completes the proof. \end{proof}
Herstein|exercise_2_1_18
theorem exercise_2_1_18 {G : Type*} [group G] [fintype G] (hG2 : even (fintype.card G)) : βˆƒ (a : G), a β‰  1 ∧ a = a⁻¹ :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $G$ is a finite group of even order, show that there must be an element $a \neq e$ such that $a=a^{-1}$.
\begin{proof} First note that $a=a^{-1}$ is the same as saying $a^2=e$, where $e$ is the identity. I.e. the statement is that there exists at least one element of order 2 in $G$. Every element $a$ of $G$ of order at least 3 has an inverse $a^{-1}$ that is not itself -- that is, $a \neq a^{-1}$. So the subset of all such elements has an even cardinality (/size). There's exactly one element with order 1 : the identity $e^1=e$. So $G$ contains an even number of elements -call it $2 k$-- of which an even number are elements of order 3 or above -- call that $2 n$ where $n<k$-- and exactly one element of order 1 . Hence the number of elements of order 2 is $$ 2 k-2 n-1=2(k-n)-1 $$ This cannot equal 0 as $2(k-n)$ is even and 1 is odd. Hence there's at least one element of order 2 in $G$, which concludes the proof. \end{proof}
Herstein|exercise_2_1_26
theorem exercise_2_1_26 {G : Type*} [group G] [fintype G] (a : G) : βˆƒ (n : β„•), a ^ n = 1 :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $G$ is a finite group, prove that, given $a \in G$, there is a positive integer $n$, depending on $a$, such that $a^n = e$.
\begin{proof} Because there are only a finite number of elements of $G$, it's clear that the set $\left\{a, a^2, a^3, \ldots\right\}$ must be a finite set and in particular, there should exist some $i$ and $j$ such that $i \neq j$ and $a^i=a^j$. WLOG suppose further that $i>j$ (just reverse the roles of $i$ and $j$ otherwise). Then multiply both sides by $\left(a^j\right)^{-1}=a^{-j}$ to get $$ a^i * a^{-j}=a^{i-j}=e $$ Thus the $n=i-j$ is a positive integer such that $a^n=e$. \end{proof}
Herstein|exercise_2_2_3
theorem exercise_2_2_3 {G : Type*} [group G] {P : β„• β†’ Prop} {hP : P = Ξ» i, βˆ€ a b : G, (a*b)^i = a^i * b^i} (hP1 : βˆƒ n : β„•, P n ∧ P (n+1) ∧ P (n+2)) : comm_group G :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $G$ is a group in which $(a b)^{i}=a^{i} b^{i}$ for three consecutive integers $i$, prove that $G$ is abelian.
\begin{proof} Let $G$ be a group, $a, b \in G$ and $i$ be any integer. Then from given condition, $$ \begin{aligned} (a b)^i & =a^i b^i \\ (a b)^{i+1} & =a^{i+1} b^{i+1} \\ (a b)^{i+2} & =a^{i+2} b^{i+2} \end{aligned} $$ From first and second, we get $$ a^{i+1} b^{i+1}=(a b)^i(a b)=a^i b^i a b \Longrightarrow b^i a=a b^i $$ From first and third, we get $$ a^{i+2} b^{i+2}=(a b)^i(a b)^2=a^i b^i a b a b \Longrightarrow a^2 b^{i+1}=b^i a b a $$ This gives $$ a^2 b^{i+1}=a\left(a b^i\right) b=a b^i a b=b^i a^2 b $$ Finally, we get $$ b^i a b a=b^i a^2 b \Longrightarrow b a=a b $$ This shows that $G$ is Abelian. \end{proof}
Herstein|exercise_2_2_6c
theorem exercise_2_2_6c {G : Type*} [group G] {n : β„•} (hn : n > 1) (h : βˆ€ (a b : G), (a * b) ^ n = a ^ n * b ^ n) : βˆ€ (a b : G), (a * b * a⁻¹ * b⁻¹) ^ (n * (n - 1)) = 1 :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Let $G$ be a group in which $(a b)^{n}=a^{n} b^{n}$ for some fixed integer $n>1$ for all $a, b \in G$. For all $a, b \in G$, prove that $\left(a b a^{-1} b^{-1}\right)^{n(n-1)}=e$.
\begin{proof} We start with the following two intermediate results. (1) $(a b)^{n-1}=b^{n-1} a^{n-1}$. (2) $a^n b^{n-1}=b^{n-1} a^n$. To prove (1), notice by the given condition for all $a, b \in G$ $(b a)^n=b^n a^n$, for some fixed integers $n>1$. Then, $(b a)^n=b^n a^n \Longrightarrow b .(a b)(a b) \ldots .(a b) . a=b\left(b^{n-1} a^{n-1}\right) a$, where $(a b)$ occurs $n-1$ times $\Longrightarrow(a b)^{n-1}=b^{n-1} a^{n-1}$, by cancellation law. Hence, for all $a, b \in G$ $$ (a b)^{n-1}=b^{n-1} a^{n-1} . $$ To prove (2), notice by the given condition for all $a, b \in G$ $(b a)^n=b^n a^n$, for some fixed integers $n>1$. Then we have $$ \begin{aligned} & (b a)^n=b^n a^n \\ \Longrightarrow & b \cdot(a b)(a b) \ldots(a b) \cdot a=b\left(b^{n-1} a^{n-1}\right) a, \text { where }(a b) \text { occurs } n-1 \text { times } \\ \Longrightarrow & (a b)^{n-1}=b^{n-1} a^{n-1}, \text { by cancellation law } \\ \Longrightarrow & (a b)^{n-1}(a b)=\left(b^{n-1} a^{n-1}\right)(a b) \\ \Longrightarrow & (a b)^n=b^{n-1} a^n b \\ \Longrightarrow & a^n b^n=b^{n-1} a^n b, \text { given condition } \\ \Longrightarrow & a^n b^{n-1}=b^{n-1} a^n, \text { by cancellation law. } \end{aligned} $$ Therefore for all $a, b \in G$ we have $$ a^n b^{n-1}=b^{n-1} a^n $$ In order to show that $$ \left(a b a^{-1} b^{-1}\right)^{n(n-1)}=e, \text { for all } a, b \in G $$ it is enough to show that $$ (a b)^{n(n-1)}=(b a)^{n(n-1)}, \quad \forall x, y \in G . $$ Step 3 This is because of $$ \begin{aligned} (a b)^{n(n-1)}=(b a)^{n(n-1)} & \left.\Longrightarrow(b a)^{-1}\right)^{n(n-1)}(a b)^{n(n-1)}=e \\ & \Longrightarrow\left(a^{-1} b^{-1}\right)^{n(n-1)}(a b)^{n(n-1)}=e \\ & \Longrightarrow\left(\left(a^{-1} b^{-1}\right)^n\right)^{n-1}\left((a b)^n\right)(n-1)=e \\ & \Longrightarrow\left((a b)^n\left(a^{-1} b^{-1}\right)^n\right)^{n-1}=e, \text { by }(1) \\ & \Longrightarrow\left(a b a^{-1} b^{-1}\right)^{n(n-1)}=e, \text { ( given condition) } \end{aligned} $$ Now, it suffices to show that $$ (a b)^{n(n-1)}=(b a)^{n(n-1)}, \quad \forall x, y \in G . $$ Now, we have $$ \begin{aligned} (a b)^{n(n-1)} & =\left(a^n b^n\right)^{n-1}, \text { by the given condition } \\ & =\left(a^n b^{n-1} b\right)^{n-1} \\ & =\left(b^{n-1} a^n b\right)^{n-1}, \text { by }(2) \\ & =\left(a^n b\right)^{n-1}\left(b^{n-1}\right)^{n-1}, \text { by }(1) \\ & =b^{n-1}\left(a^n\right)^{n-1}\left(b^{n-1}\right)^{n-1}, \text { by }(1) \\ & =\left(b^{n-1}\left(a^{n-1}\right)^n\right)\left(b^{n-1}\right)^{n-1} \\ & =\left(a^{n-1}\right)^n b^{n-1}\left(b^{n-1}\right)^{n-1}, \text { by }(2) \\ & =\left(a^{n-1}\right)^n\left(b^{n-1}\right)^n \\ & =\left(a^{n-1} b^{n-1}\right)^n, \text { by }(1) \\ & =(b a)^{n(n-1)}, \text { by }(1) . \end{aligned} $$ This completes our proof. \end{proof}
Herstein|exercise_2_3_16
theorem exercise_2_3_16 {G : Type*} [group G] (hG : βˆ€ H : subgroup G, H = ⊀ ∨ H = βŠ₯) : is_cyclic G ∧ βˆƒ (p : β„•) (fin : fintype G), nat.prime p ∧ @card G fin = p :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If a group $G$ has no proper subgroups, prove that $G$ is cyclic of order $p$, where $p$ is a prime number.
\begin{proof} Case-1: $G=(e), e$ being the identity element in $G$. Then trivially $G$ is cyclic. Case-2: $G \neq(e)$. Then there exists an non-identity element in $G.$ Let us consider an non-identity element in $G$, say $a\neq (e)$. Now look at the cyclic subgroup generated by $a$, that is, $\langle a\rangle$. Since $a\neq (e) \in G,\langle a\rangle$ is a subgroup of $G$. If $G \neq\langle a\rangle$ then $\langle a\rangle$ is a proper non-trivial subgroup of $G$, which is an impossibility. Therfore we must have $$ G=\langle a\rangle . $$ This implies, $G$ is a cyclic group generated by $a$. Then it follows that every non-identity element of $G$ is a generator of $G$. Now we claim that $G$ is finite. \end{proof}
Herstein|exercise_2_5_23
theorem exercise_2_5_23 {G : Type*} [group G] (hG : βˆ€ (H : subgroup G), H.normal) (a b : G) : βˆƒ (j : β„€) , b*a = a^j * b:=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Let $G$ be a group such that all subgroups of $G$ are normal in $G$. If $a, b \in G$, prove that $ba = a^jb$ for some $j$.
\begin{proof} Let $G$ be a group where each subgroup is normal in $G$. let $a, b \in G$. $$ \begin{aligned} \langle a\rangle\triangleright G &\Rightarrow b \cdot\langle a\rangle=\langle a\rangle \cdot b . \\ & \Rightarrow \quad b \cdot a=a^j \cdot b \text { for some } j \in \mathbb{Z}. \end{aligned} $$ (hence for $a_1 b \in G \quad a^j b=b \cdot a$ ). \end{proof}
Herstein|exercise_2_5_31
theorem exercise_2_5_31 {G : Type*} [comm_group G] [fintype G] {p m n : β„•} (hp : nat.prime p) (hp1 : Β¬ p ∣ m) (hG : card G = p^n*m) {H : subgroup G} [fintype H] (hH : card H = p^n) : characteristic H :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Suppose that $G$ is an abelian group of order $p^nm$ where $p \nmid m$ is a prime. If $H$ is a subgroup of $G$ of order $p^n$, prove that $H$ is a characteristic subgroup of $G$.
\begin{proof} Let $G$ be an abelian group of order $p^n m$, such that $p \nmid m$. Now, Given that $H$ is a subgroup of order $p^n$. Since $G$ is abelian $H$ is normal. Now we want to prove that $H$ is a characterestic subgroup, that is $\phi(H)=H$ for any automorphism $\phi$ of $G$. Now consider $\phi(H)$. Clearly $|\phi(H)|=p^n$. Suppose $\phi(H) \neq H$, then $|H \cap \phi(H)|=p^s$, where $s<n$. Consider $H \phi(H)$, this is a subgroup of $G$ as $H$ is normal. Also $|H \phi(H)|=\frac{|H||\phi(H)|}{|H \cap \phi(H)|}=\frac{p^{2 n}}{p^s}=p^{2 n-s}$, where $2 n-s>n$. By lagrange's theorem then $p^{2 n-s}\left|p^n m \Longrightarrow p^{n-s}\right| m \Longrightarrow p \mid m$-contradiction. So $\phi(H)=H$, and $H$ is characterestic subgroup of $G$. \end{proof}
Herstein|exercise_2_5_43
theorem exercise_2_5_43 (G : Type*) [group G] [fintype G] (hG : card G = 9) : comm_group G :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Prove that a group of order 9 must be abelian.
\begin{proof} We use the result from problem 40 which is as follows: Suppose $G$ is a group, $H$ is a subgroup and $|G|=n$ and $n \nmid\left(i_G(H)\right) !$. Then there exists a normal subgroup $\$ K \backslash$ neq $\{$ e $\} \$$ and $K \subseteq H$. So, we have now a group $G$ of order 9. Suppose that $G$ is cyclic, then $G$ is abelian and there is nothing more to prove. Suppose that $G$ s not cyclic,then there exists an element $a$ of order 3 , and $A=\langle a\rangle$. Now $i_G(A)=3$, now $9 \nmid 3$ !, hence by the above result there is a normal subgroup $K$, non-trivial and $K \subseteq A$. But $|A|=3$, a prime order subgroup, hence has no non-trivial subgroup, so $K=A$. So $A$ is normal subgroup. Now since $G$ is not cyclic any non-identity element is of order 3.So Let $a(\neq$ $e) \in G$.Consider $A=\langle a\rangle$. As shown before $A$ is normal. $a$ commutes with any if its powers. Now Let $b \in G$ such that $b \notin A$. Then $b a b^{-1} \in A$ and hence $b a b^{-1}=a^i$.This implies $a=b^3 a b^{-3}=a^{i^3} \Longrightarrow a^{i^3-1}=e$. So, 3 divides $i^3-1$. Also by fermat's little theorem 3 divides $i^2-1$.So 3 divides $i-1$. But $0 \leq i \leq 2$. So $i=1$, is the only possibility and hence $a b=b a$. So $a \in Z(G)$ as $b$ was arbitrary. Since $a$ was arbitrary $G=Z(G)$. Hence $G$ is abelian. \end{proof}
Herstein|exercise_2_5_52
theorem exercise_2_5_52 {G : Type*} [group G] [fintype G] (Ο† : G ≃* G) {I : finset G} (hI : βˆ€ x ∈ I, Ο† x = x⁻¹) (hI1 : (0.75 : β„š) * card G ≀ card I) : βˆ€ x : G, Ο† x = x⁻¹ ∧ βˆ€ x y : G, x*y = y*x :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Let $G$ be a finite group and $\varphi$ an automorphism of $G$ such that $\varphi(x) = x^{-1}$ for more than three-fourths of the elements of $G$. Prove that $\varphi(y) = y^{-1}$ for all $y \in G$, and so $G$ is abelian.
\begin{proof} Let us start with considering $b$ to be an arbitrary element in $A$. 1. Show that $\left|A \cap\left(b^{-1} A\right)\right|>\frac{|G|}{2}$, where $$ b^{-1} A=\left\{b^{-1} a \mid a \in A\right\} $$ First notice that if we consider a map $f: A \rightarrow b^{-1} A$ defined by $f(a)=b^{-1} a$, for all $a \in A$, then $f$ is a 1-1 map and so $\left|b^{-1} A\right| \geq|A|>\frac{3}{4}|G|$. Now using inclusion-exclusion principle we have $$ \left|A \cap\left(b^{-1} A\right)\right|=|A|+\left|b^{-1} A\right|-\left|A \cup\left(b^{-1} A\right)\right|>\frac{3}{4}|G|+\frac{3}{4}|G|-|G|=\frac{1}{2}|G| $$ 2. Argue that $A \cap\left(b^{-1} A\right) \subseteq C(b)$, where $C(b)$ is the centralizer of $b$ in $G$. Suppose $x \in A \cap\left(b^{-1} A\right)$, that means, $x \in A$ and $x \in b^{-1} A$. Thus there exist an element $a \in A$ such that $x=$ $b^{-1} a$, which gives us $x b=a \in A$. Now notice that $x, b \in A$ and $x b \in A$, therefore we get $$ \phi(x b)=(x b)^{-1} \Longrightarrow \phi(x) \phi(b)=(x b)^{-1} \Longrightarrow x^{-1} b^{-1}=b^{-1} x^{-1} \Longrightarrow x b=b x $$ Therefore, we get $x b=b x$, for any $x \in A \cap\left(b^{-1} A\right)$, that means, $x \in C(b)$. 3. Argue that $C(b)=G$. We know that centralizer of an element in a group $G$ is a subgroup (See Page 53). Therefore $C(b)$ is a subgroup of $G$. From statements $\mathbf{1}$ and $\mathbf{2}$, we have $$ |C(b)| \geq\left|A \cap\left(b^{-1} A\right)\right|>\frac{|G|}{2} $$ We need to use the following remark to argue $C(b)=G$ from the above step. Remark. Let $G$ be a finite group and $H$ be a subgroup with more then $|G| / 2$ elements then $H=G$. Proof of Remark. Suppose $|H|=p$ Then by Lagrange Theorem, there exist an $n \in \mathbb{N}$, such that $|G|=n p$, as $|H|$ divide $|G|$. Now by hypothesis $p>\frac{G]}{2}$ gives us, $$ p>\frac{|G|}{2} \Longrightarrow n p>\frac{n|G|}{2} \Longrightarrow n<2 \Longrightarrow n=1 $$ Therefore we get $H=G$. Now notice that $C(b)$ is a subgroup of $G$ with $C(b)$ having more than $|G| / 2$ elements. Therefore, $C(b)=G$. 4. Show that $A \in Z(G)$. We know that $x \in Z(G)$ if and only if $C(a)=G$. Now notice that, for any $b \in A$ we have $C(b)=G$. Therefore, every element of $A$ is in the center of $G$, that means, $A \subseteq Z(G)$. 5. 5how that $Z(G)=G$. As it is given that $|A|>\frac{3|G|}{4}$ and $A \leq|Z(G)|$, therefore we get $$ |Z(G)|>\frac{3}{4}|G|>\frac{1}{2}|G| . $$ As $Z(G)$ is a subgroup of $G$, so by the above Remark we have $Z(G)=G$. Hence $G$ is abelian. 6. Finally show that $A=G$. First notice that $A$ is a subgroup of $G$. To show this let $p, q \in A$. Then we have $$ \phi(p q)=\phi(p) \phi(q)=p^{-1} q^{-1}=(q p)^{-1}=(p q)^{-1}, \quad \text { As } G \text { is abelian. } $$ Therefore, $p q \in A$ and so we have $A$ is a subgroup of $G$. Again by applying the above remark. we get $A=G$. Therefore we have $$ \phi(y)=y^{-1}, \quad \text { for all } y \in G $$ \end{proof}
Herstein|exercise_2_7_7
theorem exercise_2_7_7 {G : Type*} [group G] {G' : Type*} [group G'] (Ο† : G β†’* G') (N : subgroup G) [N.normal] : (map Ο† N).normal :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $\varphi$ is a homomorphism of $G$ onto $G'$ and $N \triangleleft G$, show that $\varphi(N) \triangleleft G'$.
\begin{proof} We first claim that $\varphi(N)$ is a subgroup of $G'$. To see this, note that since $N$ is a subgroup of $G$, the identity element $e_G$ of $G$ belongs to $N$. Therefore, the element $\varphi(e_G) \in \varphi(N)$, so $\varphi(N)$ is a non-empty subset of $G'$. Now, let $a', b' \in \varphi(N)$. Then there exist elements $a, b \in N$ such that $\varphi(a) = a'$ and $\varphi(b) = b'$. Since $N$ is a subgroup of $G$, we have $a, b \in N$, so $ab^{-1} \in N$. Thus, we have $$\varphi(ab^{-1}) = \varphi(a) \varphi(b^{-1}) = a'b'^{-1} \in \varphi(N),$$ which shows that $a', b' \in \varphi(N)$ implies $a'b'^{-1} \in \varphi(N)$. Therefore, $\varphi(N)$ is a subgroup of $G'$. Next, we will show that $\varphi(N)$ is a normal subgroup of $G'$. Let $\varphi(N) = N'$, a subgroup of $G'$. Let $x' \in G'$ and $h' \in N'$. Since $\varphi$ is onto, there exist elements $x \in G$ and $h \in N$ such that $\varphi(x) = x'$ and $\varphi(h) = h'$. Since $N$ is a normal subgroup of $G$, we have $xhx^{-1} \in N$. Thus, $$\varphi(xhx^{-1}) = \varphi(x)\varphi(h)\varphi(x^{-1}) = x'h'x'^{-1} \in \varphi(N),$$ which shows that $x' \in G'$ and $h' \in N'$ implies $x'h'x'^{-1} \in \varphi(N)$. Therefore, $\varphi(N)$ is a normal subgroup of $G'$. This completes the proof. \end{proof}
Herstein|exercise_2_8_15
theorem exercise_2_8_15 {G H: Type*} [fintype G] [group G] [fintype H] [group H] {p q : β„•} (hp : nat.prime p) (hq : nat.prime q) (h : p > q) (h1 : q ∣ p - 1) (hG : card G = p*q) (hH : card G = p*q) : G ≃* H :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Prove that if $p > q$ are two primes such that $q \mid p - 1$, then any two nonabelian groups of order $pq$ are isomorphic.
\begin{proof} For a nonabelian group of order $p q$, the structure of the group $G$ is set by determining the relation $a b a^{-1}=b^{k^{\frac{p-1}{q}}}$ for some generator $k$ of the cyclic group. Here we are using the fact that $k^{\frac{p-1}{q}}$ is a generator for the unique subgroup of order $q$ in $U_p$ (a cyclic group of order $m$ has a unique subgroup of order $d$ for each divisor $d$ of $m$ ). The other possible generators of this subgroup are $k^{\frac{l(p-1)}{q}}$ for each $1 \leq l \leq q-1$, so these give potentially new group structures. Let $G^{\prime}$ be a group with an element $c$ of order $q$, an element $d$ of order $p$ with structure defined by the relation $c d c^{-1}=d^{k^{\frac{l(p-1)}{q}}}$. We may then define $$ \begin{aligned} \phi: G^{\prime} & \rightarrow G \\ c & \mapsto a^l \\ d & \mapsto b \end{aligned} $$ since $c$ and $a^l$ have the same order and $b$ and $d$ have the same order this is a well defined function. Since $$ \begin{aligned} \phi(c) \phi(d) \phi(c)^{-1} & =a^l b a^{-l} \\ & =b^{\left(k^{\frac{p-1}{q}}\right)^l} \\ & =b^{k^{\frac{l(p-1)}{q}}} \\ & =\phi(d)^{k^{\frac{l(p-1)}{q}}} \end{aligned} $$ $\phi\left(c^i d^j\right)=a^{l i} b^j=e$ only if $i=j=0$, so $\phi$ is 1-to-l. Therefore $G$ and $G^{\prime}$ are isomorphic and so up to isomorphism there is only one nonabelian group of order $p q$. \end{proof}
Herstein|exercise_2_10_1
theorem exercise_2_10_1 {G : Type*} [group G] (A : subgroup G) [A.normal] {b : G} (hp : nat.prime (order_of b)) : A βŠ“ (closure {b}) = βŠ₯ :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Let $A$ be a normal subgroup of a group $G$, and suppose that $b \in G$ is an element of prime order $p$, and that $b \not\in A$. Show that $A \cap (b) = (e)$.
\begin{proof} If $b \in G$ has order $p$, then $(b)$ is a cyclic group of order $p$. Since $A$ is a subgroup of $G$, we have $A \cap (b)$ is a subgroup of $G$. Also, $A \cap (b) \subseteq (b)$. So $A \cap (b)$ is a subgroup of $(b)$. Since $(b)$ is a cyclic group of order $p$, the only subgroups of $(b)$ are $(e)$ and $(b)$ itself. Therefore, either $A \cap (b) = (e)$ or $A \cap (b) = (b)$. If $A \cap (b) = (e)$, then we are done. Otherwise, if $A \cap (b) = (b)$, then $A \subseteq (b)$. Since $A$ is a subgroup of $G$ and $A \subseteq (b)$, it follows that $A$ is a subgroup of $(b)$. Since the only subgroups of $(b)$ are $(e)$ and $(b)$ itself, we have either $A = (e)$ or $A = (b)$. If $A = (e)$, then $A \cap (b) = (e)$ and we are done. But if $A = (b)$, then $b \in A$ as $b \in (b)$, which contradicts our hypothesis that $b \notin A$. So $A \neq (b)$. Hence $A \cap (b) \neq (b)$. Therefore, $A \cap (b) = (e)$. This completes our proof. \end{proof}
Herstein|exercise_2_11_7
theorem exercise_2_11_7 {G : Type*} [group G] {p : β„•} (hp : nat.prime p) {P : sylow p G} (hP : P.normal) : characteristic (P : subgroup G) :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $P \triangleleft G$, $P$ a $p$-Sylow subgroup of $G$, prove that $\varphi(P) = P$ for every automorphism $\varphi$ of $G$.
\begin{proof} Let $\phi$ be an automorphism of $G$. Let $P$ be a normal sylow p-subgroup. $\phi(P)$ is also a sylow-p subgroup. But since $P$ is normal, it is unique. Hence $\phi(P)=P$. \end{proof}
Herstein|exercise_3_2_21
theorem exercise_3_2_21 {Ξ± : Type*} [fintype Ξ±] {Οƒ Ο„: equiv.perm Ξ±} (h1 : βˆ€ a : Ξ±, Οƒ a = a ↔ Ο„ a β‰  a) (h2 : Ο„ ∘ Οƒ = id) : Οƒ = 1 ∧ Ο„ = 1 :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $\sigma, \tau$ are two permutations that disturb no common element and $\sigma \tau = e$, prove that $\sigma = \tau = e$.
\begin{proof} Note that $\sigma \tau=e$ can equivalentnly be phrased as $\tau$ being the inverse of $\sigma$. Our statement is then equivalent to the statement that an inverse of a nonidentity permutation disturbs at least one same element as that permutation. To prove this, let $\sigma$ be a nonidentity permutation, then let $\left(i_1 \cdots i_n\right)$ be a cycle in $\sigma$. Then we have that $$ \sigma\left(i_1\right)=i_2, \sigma\left(i_2\right)=i_2, \ldots, \sigma\left(i_{n-1}\right)=i_n, \sigma\left(i_n\right)=i_1, $$ but then also $$ i_1=\tau\left(i_2\right), i_2=\tau\left(i_3\right), \ldots, i_{n-1}=\tau\left(i_n\right), i_n=\tau\left(i_1\right), $$ i.e. its inverse disturbs $i_1, \ldots, i_n$. \end{proof}
Herstein|exercise_4_1_34
theorem exercise_4_1_34 : equiv.perm (fin 3) ≃* general_linear_group (fin 2) (zmod 2) :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Let $T$ be the group of $2\times 2$ matrices $A$ with entries in the field $\mathbb{Z}_2$ such that $\det A$ is not equal to 0. Prove that $T$ is isomorphic to $S_3$, the symmetric group of degree 3.
\begin{proof} The order of $T$ is $2^4-2^3-2^2+2=6$; we now find those six matrices: $$ \begin{array}{ll} A_1=\left(\begin{array}{ll} 1 & 0 \\ 0 & 1 \end{array}\right), & A_2=\left(\begin{array}{ll} 0 & 1 \\ 1 & 0 \end{array}\right) \\ A_3=\left(\begin{array}{ll} 1 & 0 \\ 1 & 1 \end{array}\right), & A_4=\left(\begin{array}{ll} 1 & 1 \\ 0 & 1 \end{array}\right) \\ A_5=\left(\begin{array}{ll} 0 & 1 \\ 1 & 1 \end{array}\right), & A_6=\left(\begin{array}{ll} 1 & 1 \\ 1 & 0 \end{array}\right) \end{array} $$ with orders $1,2,2,2,3,3$ respectively. Note that $S_3$ is composed of elements $$ \text{ id, (1 2), (1 3), (2 3), (1 2 3), (1 3 2)} $$ with orders 1, 2, 2, 2, 3, 3 respectively. Also note that, by Problem 17 of generate $S_3$. We also have that $\left(\begin{array}{llll}1 & 3 & 2\end{array}\right)=\left(\begin{array}{llll}1 & 2 & 3\end{array}\right)\left(\begin{array}{lll}1 & 2 & 3\end{array}\right)$, that $\left(\begin{array}{lll}1 & 3\end{array}\right)=\left(\begin{array}{lll}1 & 2 & 3\end{array}\right)\left(\begin{array}{ll}1 & 2\end{array}\right)$, $\left(\begin{array}{ll}1 & 2\end{array}\right)\left(\begin{array}{lll}1 & 2 & 3\end{array}\right)=\left(\begin{array}{ll}2 & 3\end{array}\right)$ and $\left(\begin{array}{lll}1 & 2\end{array}\right)\left(\begin{array}{ll}1 & 2\end{array}\right)=\mathrm{id}$ Now we can check that $\tau\left(A_2\right)=\left(\begin{array}{ll}1 & 2\end{array}\right), \tau\left(A_5\right)=\left(\begin{array}{lll}1 & 2 & 3\end{array}\right)$ induces an isomorphism. We compute $$ \begin{aligned} & \tau\left(A_1\right)=\tau\left(A_2 A_2\right)=\tau\left(A_2\right) \tau\left(A_2\right)=\mathrm{id} \\ & \tau\left(A_3\right)=\tau\left(A_5 A_2\right)=\tau\left(A_5\right) \tau\left(A_2\right)=\left(\begin{array}{llll} 1 & 2 & 3 \end{array}\right)\left(\begin{array}{lll} 1 & 2 \end{array}\right)=\left(\begin{array}{ll} 1 & 3 \end{array}\right) \\ & \tau\left(A_4\right)=\tau\left(A_2 A_5\right)=\tau\left(A_2\right) \tau\left(A_5\right)=\left(\begin{array}{lll} 1 & 2 \end{array}\right)\left(\begin{array}{lll} 1 & 2 & 3 \end{array}\right)=\left(\begin{array}{ll} 2 & 3 \end{array}\right) \\ & \tau\left(A_6\right)=\tau\left(A_5 A_5\right)=\tau\left(A_5\right) \tau\left(A_5\right)=\left(\begin{array}{lll} 1 & 3 & 2 \end{array}\right) \end{aligned} $$ Thus we see that $\tau$ extendeds to an isomorphism, since $A_2$ and $A_5$ generate $T$, so that $\tau\left(A_i A_j\right)=\tau\left(A_i\right) \tau\left(A_j\right)$ follows from writing $A_i$ and $A_j$ in terms of $A_2$ and $A_5$ and using the equlities and relations shown above. \end{proof}
Herstein|exercise_4_2_6
theorem exercise_4_2_6 {R : Type*} [ring R] (a x : R) (h : a ^ 2 = 0) : a * (a * x + x * a) = (x + x * a) * a :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $a^2 = 0$ in $R$, show that $ax + xa$ commutes with $a$.
\begin{proof} We need to show that $$ a(a x+x a)=(a x+x a) a \text { for } a, x \in R . $$ Now, $$ \begin{gathered} a(a x+x a)=a(a x)+a(x a) \\ =a^2 x+a x a \\ =0+a x a=a x a . \end{gathered} $$ Again, $$ \begin{gathered} (a x+x a) a=(a x) a+(x a) a \\ =a x a+x a^2 \\ =a x a+0=a x a . \end{gathered} $$ It follows that, $$ a(a x+x a)=(a x+x a) a, \text { for } x, a \in R . $$ This shows that $a x+x a$ commutes with $a$. This completes the proof. \end{proof}
Herstein|exercise_4_3_1
theorem exercise_4_3_1 {R : Type*} [comm_ring R] (a : R) : βˆƒ I : ideal R, {x : R | x*a=0} = I :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $R$ is a commutative ring and $a \in R$, let $L(a) = \{x \in R \mid xa = 0\}$. Prove that $L(a)$ is an ideal of $R$.
\begin{proof} First, note that if $x \in L(a)$ and $y \in L(a)$ then $x a=0$ and $y a=0$, so that $$ \begin{aligned} x a-y a & =0 \\ (x-y) a & =0, \end{aligned} $$ i.e. $L(a)$ is an additive subgroup of $R$. (We have used the criterion that $H$ is a subgroup of $G$ if for any $h_1, h_2 \in H$ we have that $h_1 h_2^{-1} \in H$. Now we prove the conclusion. Let $r \in R$ and $b \in L(a)$, then $b a=0$, and so $x b a=0$ which by associativity of multiplication in $R$ is equivalent to $$ (x b) a=0, $$ so that $x b \in L(a)$. Since $R$ is commutative, (1) implies that $(bx)a=0$, so that $b x \in L(a)$, which concludes the proof that $L(a)$ is an ideal. \end{proof}
Herstein|exercise_4_4_9
theorem exercise_4_4_9 (p : β„•) (hp : nat.prime p) : (βˆƒ S : finset (zmod p), S.card = (p-1)/2 ∧ βˆƒ x : zmod p, x^2 = p) ∧ (βˆƒ S : finset (zmod p), S.card = (p-1)/2 ∧ Β¬ βˆƒ x : zmod p, x^2 = p) :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Show that $(p - 1)/2$ of the numbers $1, 2, \ldots, p - 1$ are quadratic residues and $(p - 1)/2$ are quadratic nonresidues $\mod p$.
\begin{proof} To find all the quadratic residues $\bmod p$ among the integers $1,2, \ldots, p-1$, we compute the least positive residues modulo $p$ of the squares of the integers $1,2, \ldots, p-1\}$. Since there are $p-1$ squares to consider, and since each congruence $x^2 \equiv a (\bmod p)$ has either zero or two solutions, there must be exactly $\frac{(p-1)}{2}$ quadratic residues mod $p$ among the integers $1,2, \ldots, p-1$. The remaining $$ (p-1)-\frac{(p-1)}{2}=\frac{(p-1)}{2} $$ positive integers less than $p-1$ are quadratic non-residues of $\bmod p$. \end{proof}
Herstein|exercise_4_5_23
theorem exercise_4_5_23 {p q: polynomial (zmod 7)} (hp : p = X^3 - 2) (hq : q = X^3 + 2) : irreducible p ∧ irreducible q ∧ (nonempty $ polynomial (zmod 7) β§Έ ideal.span ({p} : set $ polynomial $ zmod 7) ≃+* polynomial (zmod 7) β§Έ ideal.span ({q} : set $ polynomial $ zmod 7)) :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Let $F = \mathbb{Z}_7$ and let $p(x) = x^3 - 2$ and $q(x) = x^3 + 2$ be in $F[x]$. Show that $p(x)$ and $q(x)$ are irreducible in $F[x]$ and that the fields $F[x]/(p(x))$ and $F[x]/(q(x))$ are isomorphic.
\begin{proof} We have that $p(x)$ and $q(x)$ are irreducible if they have no roots in $\mathbb{Z}_7$, which can easily be checked. E.g. for $p(x)$ we have that $p(0)=5, p(1)=6, p(2)=6, p(3)=4, p(4)=6$, $p(5)=4, p(6)=4$, and similarly for $q(x)$. We have that every element of $F[x] /(p(x))$ is equal to $a x^2+b x+c+(p(x))$, and likewise for $F[x] /(q(x))$. We consider a map $\tau$ : $F[x] /(p(x)) \rightarrow F[x] /(q(x))$ given by $$ \tau\left(a x^2+b x+c+(p(x))\right)=a x^2-b x+c+(q(x)) . $$ This map is obviously onto, and since $|F[x] /(p(x))|=|F[x] /(q(x))|=7^3$ by Problem 16, it is also one-to-one. We claim that it is a homomorphism. Additivity of $\tau$ is immediate by the linearity of addition of polynomial coefficient, so we just have to check the multiplicativity; if $n=a x^2+b x+$ $c+(p(x))$ and $m=d x^2+e x+f+(p(x))$ then $$ \begin{aligned} \tau(n m) & =\tau\left(a d x^4+(a e+b d) x^3+(a f+b e+c d) x^2+(b f+c e) x+c f+(p(x))\right) \\ & =\tau\left(2 a d x+2(a e+b d)+(a f+b e+c d) x^2+(b f+c e) x+c f+(p(x))\right) \\ & =\tau\left((a f+b e+c d) x^2+(b f+c e+2 a d) x+(c f+2 a e+2 b d)+(p(x))\right) \\ & =(a f+b e+c d) x^2-(b f+c e+2 a d) x+c f+2 a e+2 b d+(q(x)) \\ & =a d x^4-(a e+b d) x^3+(a f+b e+c d) x^2-(b f+c e) x+c f+(q(x)) \\ & =\left(a x^2-b x+c+(q(x))\right)\left(d x^2-e x+f+(q(x))\right) \\ & =\tau(n) \tau(m) . \end{aligned} $$ where in the second equality we used that $x^3+p(x)=2+p(x)$ and in the fifth we used that $x^3+$ $q(x)=-2+q(x)$ \end{proof}
Herstein|exercise_4_6_2
theorem exercise_4_6_2 : irreducible (X^3 + 3*X + 2 : polynomial β„š) :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
Prove that $f(x) = x^3 + 3x + 2$ is irreducible in $Q[x]$.
\begin{proof} Let us assume that $f(x)$ is reducible over $\mathbb{Q}[x]$. Then there exists a rational root of $f(x)$. Let $p / q$ be a rational root of $f(x)$, where $\operatorname{gcd}(p, q)=1$. Then $f(p / q)=0$. Now, $$ \begin{aligned} & f(p / q)=(p / q)^3+3(p / q)+2 \\ \Longrightarrow & (p / q)^3+3(p / q)+2=0 \\ \Longrightarrow & p^3+3 p q^2=-2 q^3 \\ \Longrightarrow & p\left(p^2+3 q^2\right)=-q^3 \end{aligned} $$ It follows that, $p$ divides $q$ which is a contradiction to the fact that $\operatorname{gcd}(p, q)=1$. This implies that $f(x)$ has no rational root. Now we know that, a polynomial of degree two or three over a field $F$ is reducible if and only if it has a root in $F$. Now $f(x)$ is a 3 degree polynomial having no root in $\mathbb{Q}$. So, $f(x)$ is irreducible in $\mathbb{Q}[x]$. This completes the proof. \end{proof}
Herstein|exercise_5_1_8
theorem exercise_5_1_8 {p m n: β„•} {F : Type*} [field F] (hp : nat.prime p) (hF : char_p F p) (a b : F) (hm : m = p ^ n) : (a + b) ^ m = a^m + b^m :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $F$ is a field of characteristic $p \neq 0$, show that $(a + b)^m = a^m + b^m$, where $m = p^n$, for all $a, b \in F$ and any positive integer $n$.
\begin{proof} Since $F$ is of characteristic $p$ and we have considered arbitrary two elements $a, b$ in $F$ we have $$ \begin{aligned} & p a=p b=0 \\ & \Longrightarrow p^n a=p^n b=0 \\ & \Longrightarrow m a=m b=0 \text {. } \\ & \end{aligned} $$ Now we know from Binomial Theorem that $$ (a+b)^m=\sum_{i=0}^m\left(\begin{array}{c} m \\ i \end{array}\right) a^i b^{m-i} $$ Here $$ \left(\begin{array}{c} m \\ i \end{array}\right)=\frac{m !}{i !(m-i) !} . $$ Now we know that for any integer $n$ and any integer $k$ satisfying $1 \leq k<n, n$ always divides $\left(\begin{array}{l}n \\ k\end{array}\right)$. So in our case for $i$ in the range $1 \leq i<m, m$ divides $\left(\begin{array}{c}m \\ i\end{array}\right)$. It follows that $p$ divides $\left(\begin{array}{c}m \\ i\end{array}\right)$, for $i$ satisfying $1 \leq i<m$, since $m=p^n$ for any integer $n$. Therefore other than the terms $a^m$ and $b^m$ in the expansion $\sum_{i=0}^m\left(\begin{array}{c}m \\ i\end{array}\right) a^i b^{m-i}$ will vanish due to char $p$ nature of $F$. Hence we have $$ \sum_{i=0}^m\left(\begin{array}{c} m \\ i \end{array}\right) a^i b^{m-i}=a^m+b^m $$ This follows that, for all $a, b \in F$ $$ (a+b)^m=a^m+b^m . $$ This completes the proof. \end{proof}
Herstein|exercise_5_3_7
theorem exercise_5_3_7 {K : Type*} [field K] {F : subfield K} {a : K} (ha : is_algebraic F (a ^ 2)) : is_algebraic F a :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $a \in K$ is such that $a^2$ is algebraic over the subfield $F$ of $K$, show that a is algebraic over $F$.
\begin{proof} Since $a^2$ is algebraic over $F$, there exist a non-zero polynomial $f(x)$ in $F[x]$ such that $f\left(a^2\right)=0$. Consider a new polynomial $g(x)$ defined as $g(x)=f\left(x^2\right)$. Clearly $g(x) \in F[x]$ and $g(a)=f\left(a^2\right)= 0$. \end{proof}
Herstein|exercise_5_4_3
theorem exercise_5_4_3 {a : β„‚} {p : β„‚ β†’ β„‚} (hp : p = Ξ» x, x^5 + real.sqrt 2 * x^3 + real.sqrt 5 * x^2 + real.sqrt 7 * x + 11) (ha : p a = 0) : βˆƒ p : polynomial β„‚ , p.degree < 80 ∧ a ∈ p.roots ∧ βˆ€ n : p.support, βˆƒ a b : β„€, p.coeff n = a / b :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $a \in C$ is such that $p(a) = 0$, where $p(x) = x^5 + \sqrt{2}x^3 + \sqrt{5}x^2 + \sqrt{7}x + \sqrt{11}$, show that $a$ is algebraic over $\mathbb{Q}$ of degree at most 80.
\begin{proof} Given $a \in \mathbb{C}$ such that $p(a)=0$, where $$ p(x)=x^5+\sqrt{2} x^3+\sqrt{5} x^2+\sqrt{7} x+\sqrt{11} $$ Here, we note that $p(x) \in \mathbb{Q}(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11})$ and $$ \begin{aligned} {[Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11}): \mathbb{Q}] } & =[Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11}): Q(\sqrt{2}, \sqrt{5}, \sqrt{7})] \cdot[\mathbb{Q}(\sqrt{2}, \sqrt{5}, \sqrt{7}): \mathbb{Q}(\sqrt{2}, \sqrt{5})] \\ & \cdot[\mathbb{Q}(\sqrt{2}, \sqrt{5}): \mathbb{Q}(\sqrt{2})] \cdot[\mathbb{Q}(\sqrt{2}): \mathbb{Q}] \\ & =2 \cdot 2 \cdot 2 \cdot 2 \\ & =16 \end{aligned} $$ Here, we note that $p(x)$ is of degree 5 over $\mathbb{Q}(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11})$. If $a$ is root of $p(x)$, then $$ [Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11}, a): \mathbb{Q}]=[Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11}): Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11})] \cdot 15 $$ and $[Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11}): Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11})] \leq 5$. We get equality if $p(x)$ is irreducible over $Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11})$. This gives $$ [Q(\sqrt{2}, \sqrt{5}, \sqrt{7}, \sqrt{11}, a): \mathbb{Q}] \leq 16 \cdot 5=80 $$ \end{proof}
Herstein|exercise_5_6_14
theorem exercise_5_6_14 {p m n: β„•} (hp : nat.prime p) {F : Type*} [field F] [char_p F p] (hm : m = p ^ n) : card (root_set (X ^ m - X : polynomial F) F) = m :=
import .common open set function nat fintype real subgroup ideal polynomial submodule zsqrtd open char_p mul_aut matrix open_locale big_operators noncomputable theory
If $F$ is of characteristic $p \neq 0$, show that all the roots of $x^m - x$, where $m = p^n$, are distinct.
\begin{proof} Let us consider $f(x)=x^m-x$. Then $f \in F[x]$. Claim: $f(x)$ has a multiple root in some extension of $F$ if and only if $f(x)$ is not relatively prime to its formal derivative, $f^{\prime}(x)$. Proof of the Claim: Let us assume that $f(x)$ has a multiple root in some extension of $F$. Let $y$ be a multiple root of $f(x)$. Then over a splitting field, we have $$ f(x)=(x-y)^n g(x), \text { for some integer } n \geq 2 . $$ Here $g(x)$ is a polynomial such that $g(y) \neq 0$. Now taking derivative of $f$ we get $$ f^{\prime}(x)=n \cdot(x-y)^{n-1} g(x)+(x-y)^n g^{\prime}(x) $$ here $g^{\prime}(x)$ implies derivative of $g$ with respect to $x$. Since we have $n \geq 2$, this implies $(n-1) \geq 1$. Hence, (1) shows that $f^{\prime}(x)$ has $y$ as a root. Therefore, $f(x)$ is not relatively prime to $f^{\prime}(x)$. We now prove the other direction. Conversely, let us assume that $f(x)$ is not relatively prime to $f^{\prime}(x)$. Let $y$ is a root of both $f(x)$ and $f^{\prime}(x)$. Since $y$ is a root of $f(x)$, we can write $$ f(x)=(x-y) \cdot g(x) $$ for some polynomial $g(x)$. then taking derivative of $f(x)$ we have $$ f^{\prime}(x)=g(x)+(x-y) \cdot g^{\prime}(x) $$ where $g^{\prime}(x)$ is the derivative of $g(x)$ with respect to $x$. Since $y$ is a root of $f^{\prime}(x)$ also we have $$ f^{\prime}(y)=0 $$ Then we have $$ \begin{aligned} & f^{\prime}(y)=g(y)+(y-y) \cdot g^{\prime}(y) \\ \Longrightarrow & f^{\prime}(y)=g(y) \\ \Longrightarrow & g(y)=0 . \end{aligned} $$ This implies $y$ is a root of $g(x)$ also. Therefore we have $$ g(x)=(x-y) \cdot h(x) $$ for some polynomial $h(x)$. Now form (2) we have $$ f(x)=(x-y)^2 \cdot h(x) . $$ This follows that $y$ is a multiple root of $f(x)$. Therefore, $f(x)$ has a multiple root in some extension of the field $F$. This completes the proof of the Claim. In our case, $f(x)=x^m-x$, where $m=p^n$. Now we calculate the derivative of $f$. That is $$ f^{\prime}(x)=m x^{m-1}-1=-1(\bmod p) . $$ By the above condition it follows that, $f^{\prime}$ has no root same as $f$, that is, $f(x)$ and $f^{\prime}(x)$ are relatively prime. Hence, $f(x)$ has no multiple root in $F$. Since $f(x)=x^m-x$ is a polynomial of degree $m$, it follows that $f(x)$ has $m$ distinct roots in $F$, where $m=p^n$. This completes the proof. \end{proof}
Artin|exercise_2_3_2
theorem exercise_2_3_2 {G : Type*} [group G] (a b : G) : βˆƒ g : G, b* a = g * a * b * g⁻¹ :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that the products $a b$ and $b a$ are conjugate elements in a group.
\begin{proof} We have that $(a^{-1})ab(a^{-1})^{-1} = ba$. \end{proof}
Artin|exercise_2_8_6
theorem exercise_2_8_6 {G H : Type*} [group G] [group H] : center (G Γ— H) ≃* (center G) Γ— (center H) :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that the center of the product of two groups is the product of their centers.
\begin{proof} We have that $(g_1, g_2)\cdot (h_1, h_2) = (h_1, h_2)\cdot (g_1, g_2)$ if and only if $g_1h_1 = h_1g_1$ and $g_2h_2 = h_2g_2$. \end{proof}
Artin|exercise_3_2_7
theorem exercise_3_2_7 {F : Type*} [field F] {G : Type*} [field G] (Ο† : F β†’+* G) : injective Ο† :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that every homomorphism of fields is injective.
\begin{proof} Suppose $f(a)=f(b)$, then $f(a-b)=0=f(0)$. If $u=(a-b) \neq 0$, then $f(u) f\left(u^{-1}\right)=f(1)=1$, but that means that $0 f\left(u^{-1}\right)=1$, which is impossible. Hence $a-b=0$ and $a=b$. \end{proof}
Artin|exercise_3_7_2
theorem exercise_3_7_2 {K V : Type*} [field K] [add_comm_group V] [module K V] {ΞΉ : Type*} [fintype ΞΉ] (Ξ³ : ΞΉ β†’ submodule K V) (h : βˆ€ i : ΞΉ, Ξ³ i β‰  ⊀) : (β‹‚ (i : ΞΉ), (Ξ³ i : set V)) β‰  ⊀ :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Let $V$ be a vector space over an infinite field $F$. Prove that $V$ is not the union of finitely many proper subspaces.
\begin{proof} If $V$ is the set-theoretic union of $n$ proper subspaces $W_i$ ( $1 \leq i \leq n$ ), then $|F| \leq n-1$. Proof. We may suppose no $W_i$ is contained in the union of the other subspaces. Let $u \in W_i, \quad u \notin \bigcup_{j \neq i} W_j$ and $v \notin W_i$. Then $(v+F u) \cap W_i=\varnothing$ and $(v+F u) \cap W_j(j \neq i)$ contains at most one vector since otherwise $W_j$ would contain $u$. Hence $$ |v+F u|=|F| \leq n-1 . $$ Corollary: Avoidance lemma for vector spaces. Let $E$ be a vector space over an infinite field. If a subspace is contained in a finite union of subspaces, it is contained in one of them. \end{proof}
Artin|exercise_6_4_2
theorem exercise_6_4_2 {G : Type*} [group G] [fintype G] {p q : β„•} (hp : prime p) (hq : prime q) (hG : card G = p*q) : is_simple_group G β†’ false :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that no group of order $p q$, where $p$ and $q$ are prime, is simple.
\begin{proof} If $|G|=n=p q$ then the only two Sylow subgroups are of order $p$ and $q$. From Sylow's third theorem we know that $n_p \mid q$ which means that $n_p=1$ or $n_p=q$. If $n_p=1$ then we are done (by a corollary of Sylow's theorem) If $n_p=q$ then we have accounted for $q(p-1)=p q-q$ elements of $G$ and so there is only one group of order $q$ and again we are done. \end{proof}
Artin|exercise_6_4_12
theorem exercise_6_4_12 {G : Type*} [group G] [fintype G] (hG : card G = 224) : is_simple_group G β†’ false :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that no group of order 224 is simple.
\begin{proof} The following proves there must exist a normal Sylow 2 -subgroup of order 32 , Suppose there are $n_2=7$ Sylow 2 -subgroups in $G$. Making $G$ act on the set of these Sylow subgroups by conjugation (Mitt wrote about this but on the set of the other Sylow subgroups, which gives no contradiction), we get a homomorphism $G \rightarrow S_7$ which must be injective if $G$ is simple (why?). But this cannot be since then we would embed $G$ into $S_7$, which is impossible since $|G| \nmid 7 !=\left|S_7\right|$ (why?) \end{proof}
Artin|exercise_10_1_13
theorem exercise_10_1_13 {R : Type*} [ring R] {x : R} (hx : is_nilpotent x) : is_unit (1 + x) :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
An element $x$ of a ring $R$ is called nilpotent if some power of $x$ is zero. Prove that if $x$ is nilpotent, then $1+x$ is a unit in $R$.
\begin{proof} If $x^n=0$, then $$ (1+x)\left(\sum_{k=0}^{n-1}(-1)^k x^k\right)=1+(-1)^{n-1} x^n=1 . $$ \end{proof}
Artin|exercise_10_6_7
theorem exercise_10_6_7 {I : ideal gaussian_int} (hI : I β‰  βŠ₯) : βˆƒ (z : I), z β‰  0 ∧ (z : gaussian_int).im = 0 :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that every nonzero ideal in the ring of Gauss integers contains a nonzero integer.
\begin{proof} Let $I$ be some nonzero ideal. Then there exists some $z \in \mathbb{Z}[i], z \neq 0$ such that $z \in I$. We know that $z=a+b i$, for some $a, b \in \mathbb{Z}$. We consider three cases: 1. If $b=0$, then $z=a$, so $z \in \mathbb{Z} \cap I$, and $z \neq 0$, so the statement of the exercise holds. 2. If $a=0$, then $z=i b$. Since $z \neq 0$, we conclude that $b \neq 0$. Since $I$ is an ideal in $\mathbb{Z}[i]$, and $i \in \mathbb{Z}[i]$, we conclude that $i z \in I$. Furthermore, $i z=-b \in \mathbb{Z}$. Thus, $i z$ is a nonzero integer which is in $I$. 3. Let $a \neq 0$ and $b \neq 0$. Since $I$ is an ideal and $z \in I$, we conclude that $z^2 \in I$; that is, $$ (a+b i)^2=a^2-b^2+2 a b i \in I $$ Furthermore, since $-2 a \in \mathbb{Z}[i]$, and $z \in I$ and $I$ is an ideal, $-2 a z \in I$; that is, $$ -2 a z=-2 a(a+b i)=-2 a^2-2 a b i \in I $$ Since $I$ is closed under addition, $$ \left(a^2-b^2+2 a b i\right)+\left(-2 a^2-2 a b i\right) \in I \Longrightarrow-a^2-b^2 \in I $$ Notice that $-a^2-b^2 \neq 0$ since $a^2>0$ and $b^2>0$, so $-a^2-b^2<0$. Furthermore, it is an integer. Thus, we have found a nonzero integer in $I$. \end{proof}
Artin|exercise_10_4_7a
theorem exercise_10_4_7a {R : Type*} [comm_ring R] [no_zero_divisors R] (I J : ideal R) (hIJ : I + J = ⊀) : I * J = I βŠ“ J :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Let $I, J$ be ideals of a ring $R$ such that $I+J=R$. Prove that $I J=I \cap J$.
\begin{proof} We have seen that $IJ \subset I \cap J$, so it remains to show that $I \cap J \subset IJ$. Since $I+J = (1)$, there are elements $i \in I$ and $j \in J$ such that $i+j = 1$. Let $k \in I \cap J$, and multiply $i+j=1$ through by $k$ to get $ki+kj = k$. Write this more suggestively as \[ k = ik+kj. \] The first term is in $IJ$ because $k \in J$, and the second term is in $IJ$ because $k \in I$, so $k \in IJ$ as desired. \end{proof}
Artin|exercise_11_2_13
theorem exercise_11_2_13 (a b : β„€) : (of_int a : gaussian_int) ∣ of_int b β†’ a ∣ b :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
If $a, b$ are integers and if $a$ divides $b$ in the ring of Gauss integers, then $a$ divides $b$ in $\mathbb{Z}$.
\begin{proof} Suppose $a|b$ in $\mathbb{Z}[i]$ and $a,b\in\mathbb{Z}$. Then $a(x+yi)=b$ for $x,y\in\mathbb{Z}$. Expanding this we get $ax+ayi=b$, and equating imaginary parts gives us $ay=0$, implying $y=0$. \end{proof}
Artin|exercise_11_4_6a
theorem exercise_11_4_6a {F : Type*} [field F] [fintype F] (hF : card F = 7) : irreducible (X ^ 2 + 1 : polynomial F) :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that $x^2+x+1$ is irreducible in the field $\mathbb{F}_2$.
\begin{proof} If $x^2+x+1$ were reducible in $\mathbb{F}_2$, its factors must be linear. But we neither have that $0^2+0+1=$ nor $1^2+1+1=0$, therefore $x^2+x+1$ is irreducible. \end{proof}
Artin|exercise_11_4_6c
theorem exercise_11_4_6c : irreducible (X^3 - 9 : polynomial (zmod 31)) :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that $x^3 - 9$ is irreducible in $\mathbb{F}_{31}$.
\begin{proof} If $p(x) = x^3-9$ were reducible, it would have a linear factor, since it either has a linear factor and a quadratic factor or three linear factors. We can then verify by brute force that $p(x)\neq 0$ for $x \in \mathbb{F}_31$. \end{proof}
Artin|exercise_11_13_3
theorem exercise_11_13_3 (N : β„•): βˆƒ p β‰₯ N, nat.prime p ∧ p + 1 ≑ 0 [MOD 4] :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Prove that there are infinitely many primes congruent to $-1$ (modulo $4$).
\begin{proof} First we show a lemma: if $a \equiv 3(\bmod 4)$ then there exists a prime $p$ such that $p \mid a$ and $p \equiv 3(\bmod 4)$. Clearly, all primes dividing $a$ are odd. Suppose all of them would be $\equiv 1(\bmod 4)$. Then their product would also be $a \equiv 1(\bmod 4)$, which is a contradiction. To prove the main claim, suppose that $p_1, \ldots, p_n$ would be all such primes. (In particular, we have $p_1=3$.) Consider $a=4 p_2 \cdots p_n+3$. (Or you can take $a=4 p_2 \cdots p_n-1$.) Show that $p_i \nmid a$ for $i=1, \ldots, n$. (The case $3 \nmid a$ is solved differently than the other primes - this is the reason for omitting $p_1$ in the definition of $a$.) Then use the above lemma to get a contradiction. \end{proof}
Artin|exercise_13_6_10
theorem exercise_13_6_10 {K : Type*} [field K] [fintype Kˣ] : ∏ (x : Kˣ), x = -1 :=
import .common open function fintype subgroup ideal polynomial submodule zsqrtd char_p open_locale big_operators noncomputable theory
Let $K$ be a finite field. Prove that the product of the nonzero elements of $K$ is $-1$.
\begin{proof} Since we are working with a finite field with $q$ elements, anyone of them is a root of the following polynomial $$ x^q-x=0 . $$ In particular if we rule out the 0 element, any $a_i \neq 0$ is a root of $$ x^{q-1}-1=0 . $$ This polynomial splits completely in $\mathbb{F}_q$ so we find $$ \left(x-a_1\right) \cdots\left(x-a_{q-1}\right)=0 $$ in particular $$ x^{q-1}-1=\left(x-a_1\right) \cdots\left(x-a_{q-1}\right) $$ Thus $a_1 \cdots a_{q-1}=-1$. \end{proof}
Dummit-Foote|exercise_1_1_2a
theorem exercise_1_1_2a : βˆƒ a b : β„€, a - b β‰  b - a :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove the the operation $\star$ on $\mathbb{Z}$ defined by $a\star b=a-b$ is not commutative.
\begin{proof} Not commutative since $$ 1 \star(-1)=1-(-1)=2 $$ $$ (-1) \star 1=-1-1=-2 . $$ \end{proof}
Dummit-Foote|exercise_1_1_4
theorem exercise_1_1_4 (n : β„•) : βˆ€ (a b c : β„•), (a * b) * c ≑ a * (b * c) [ZMOD n] :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that the multiplication of residue class $\mathbb{Z}/n\mathbb{Z}$ is associative.
\begin{proof} We have $$ \begin{aligned} (\bar{a} \cdot \bar{b}) \cdot \bar{c} &=\overline{a \cdot b} \cdot \bar{c} \\ &=\overline{(a \cdot b) \cdot c} \\ &=\overline{a \cdot(b \cdot c)} \\ &=\bar{a} \cdot \overline{b \cdot c} \\ &=\bar{a} \cdot(\bar{b} \cdot \bar{c}) \end{aligned} $$ since integer multiplication is associative. \end{proof}
Dummit-Foote|exercise_1_1_15
theorem exercise_1_1_15 {G : Type*} [group G] (as : list G) : as.prod⁻¹ = (as.reverse.map (λ x, x⁻¹)).prod :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that $(a_1a_2\dots a_n)^{-1} = a_n^{-1}a_{n-1}^{-1}\dots a_1^{-1}$ for all $a_1, a_2, \dots, a_n\in G$.
\begin{proof} For $n=1$, note that for all $a_1 \in G$ we have $a_1^{-1}=a_1^{-1}$. Now for $n \geq 2$ we proceed by induction on $n$. For the base case, note that for all $a_1, a_2 \in G$ we have $$ \left(a_1 \cdot a_2\right)^{-1}=a_2^{-1} \cdot a_1^{-1} $$ since $$ a_1 \cdot a_2 \cdot a_2^{-1} a_1^{-1}=1 . $$ For the inductive step, suppose that for some $n \geq 2$, for all $a_i \in G$ we have $$ \left(a_1 \cdot \ldots \cdot a_n\right)^{-1}=a_n^{-1} \cdot \ldots \cdot a_1^{-1} . $$ Then given some $a_{n+1} \in G$, we have $$ \begin{aligned} \left(a_1 \cdot \ldots \cdot a_n \cdot a_{n+1}\right)^{-1} &=\left(\left(a_1 \cdot \ldots \cdot a_n\right) \cdot a_{n+1}\right)^{-1} \\ &=a_{n+1}^{-1} \cdot\left(a_1 \cdot \ldots \cdot a_n\right)^{-1} \\ &=a_{n+1}^{-1} \cdot a_n^{-1} \cdot \ldots \cdot a_1^{-1}, \end{aligned} $$ using associativity and the base case where necessary. \end{proof}
Dummit-Foote|exercise_1_1_17
theorem exercise_1_1_17 {G : Type*} [group G] {x : G} {n : β„•} (hxn: order_of x = n) : x⁻¹ = x ^ (n - 1 : β„€) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $x$ be an element of $G$. Prove that if $|x|=n$ for some positive integer $n$ then $x^{-1}=x^{n-1}$.
\begin{proof} We have $x \cdot x^{n-1}=x^n=1$, so by the uniqueness of inverses $x^{-1}=x^{n-1}$. \end{proof}
Dummit-Foote|exercise_1_1_20
theorem exercise_1_1_20 {G : Type*} [group G] {x : G} : order_of x = order_of x⁻¹ :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
For $x$ an element in $G$ show that $x$ and $x^{-1}$ have the same order.
\begin{proof} Recall that the order of a group element is either a positive integer or infinity. Suppose $|x|$ is infinite and that $\left|x^{-1}\right|=n$ for some $n$. Then $$ x^n=x^{(-1) \cdot n \cdot(-1)}=\left(\left(x^{-1}\right)^n\right)^{-1}=1^{-1}=1, $$ a contradiction. So if $|x|$ is infinite, $\left|x^{-1}\right|$ must also be infinite. Likewise, if $\left|x^{-1}\right|$ is infinite, then $\left|\left(x^{-1}\right)^{-1}\right|=|x|$ is also infinite. Suppose now that $|x|=n$ and $\left|x^{-1}\right|=m$ are both finite. Then we have $$ \left(x^{-1}\right)^n=\left(x^n\right)^{-1}=1^{-1}=1, $$ so that $m \leq n$. Likewise, $n \leq m$. Hence $m=n$ and $x$ and $x^{-1}$ have the same order. \end{proof}
Dummit-Foote|exercise_1_1_22b
theorem exercise_1_1_22b {G: Type*} [group G] (a b : G) : order_of (a * b) = order_of (b * a) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Deduce that $|a b|=|b a|$ for all $a, b \in G$.
\begin{proof} Let $a$ and $b$ be arbitrary group elements. Letting $x=a b$ and $g=a$, we see that $$ |a b|=\left|a^{-1} a b a\right|=|b a| . $$ \end{proof}
Dummit-Foote|exercise_1_1_29
theorem exercise_1_1_29 {A B : Type*} [group A] [group B] : βˆ€ x y : A Γ— B, x*y = y*x ↔ (βˆ€ x y : A, x*y = y*x) ∧ (βˆ€ x y : B, x*y = y*x) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that $A \times B$ is an abelian group if and only if both $A$ and $B$ are abelian.
\begin{proof} $(\Rightarrow)$ Suppose $a_1, a_2 \in A$ and $b_1, b_2 \in B$. Then $$ \left(a_1 a_2, b_1 b_2\right)=\left(a_1, b_1\right) \cdot\left(a_2, b_2\right)=\left(a_2, b_2\right) \cdot\left(a_1, b_1\right)=\left(a_2 a_1, b_2 b_1\right) . $$ Since two pairs are equal precisely when their corresponding entries are equal, we have $a_1 a_2=a_2 a_1$ and $b_1 b_2=b_2 b_1$. Hence $A$ and $B$ are abelian. $(\Leftarrow)$ Suppose $\left(a_1, b_1\right),\left(a_2, b_2\right) \in A \times B$. Then we have $$ \left(a_1, b_1\right) \cdot\left(a_2, b_2\right)=\left(a_1 a_2, b_1 b_2\right)=\left(a_2 a_1, b_2 b_1\right)=\left(a_2, b_2\right) \cdot\left(a_1, b_1\right) . $$ Hence $A \times B$ is abelian. \end{proof}
Dummit-Foote|exercise_1_3_8
theorem exercise_1_3_8 : infinite (equiv.perm β„•) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $\Omega=\{1,2,3, \ldots\}$ then $S_{\Omega}$ is an infinite group
\begin{proof} Recall that the codomain of an injective function must be at least as large (in cardinality) as the domain of the function. With that in mind, define the function $$ \begin{gathered} f: \mathbb{N} \rightarrow S_{\mathbb{N}} \\ f(n)=(1 n) \end{gathered} $$ where $(1 n)$ is the cycle decomposition of an element of $S_{\mathbb{N}}$ (specifically it's the function given by $g(1)=n, g(2)=2, g(3)=3, \ldots)$. The function $f$ maps every natural number to a distinct one of these functions. Hence $f$ is injective. Hence $\infty=|\mathbb{N}| \leq\left|S_{\mathbb{N}}\right|$. \end{proof}
Dummit-Foote|exercise_1_6_11
theorem exercise_1_6_11 {A B : Type*} [group A] [group B] : A Γ— B ≃* B Γ— A :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $A$ and $B$ be groups. Prove that $A \times B \cong B \times A$.
\begin{proof} We know from set theory that the mapping $\varphi: A \times B \rightarrow B \times A$ given by $\varphi((a, b))=(b, a)$ is a bijection with inverse $\psi: B \times A \rightarrow A \times B$ given by $\psi((b, a))=(a, b)$. Also $\varphi$ is a homomorphism, as we show below. Let $a_1, a_2 \in A$ and $b_1, b_2 \in B$. Then $$ \begin{aligned} \varphi\left(\left(a_1, b_1\right) \cdot\left(a_2, b_2\right)\right) &=\varphi\left(\left(a_1 a_2, b_1 b_2\right)\right) \\ &=\left(b_1 b_2, a_1 a_2\right) \\ &=\left(b_1, a_1\right) \cdot\left(b_2, a_2\right) \\ &=\varphi\left(\left(a_1, b_1\right)\right) \cdot \varphi\left(\left(a_2, b_2\right)\right) \end{aligned} $$ Hence $A \times B \cong B \times A$. \end{proof}
Dummit-Foote|exercise_1_6_23
theorem exercise_1_6_23 {G : Type*} [group G] (Οƒ : mul_aut G) (hs : βˆ€ g : G, Οƒ g = 1 β†’ g = 1) (hs2 : βˆ€ g : G, Οƒ (Οƒ g) = g) : βˆ€ x y : G, x*y = y*x :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $G$ be a finite group which possesses an automorphism $\sigma$ such that $\sigma(g)=g$ if and only if $g=1$. If $\sigma^{2}$ is the identity map from $G$ to $G$, prove that $G$ is abelian.
\begin{proof} Solution: We define a mapping $f: G \rightarrow G$ by $f(x)=x^{-1} \sigma(x)$. Claim: $f$ is injective. Proof of claim: Suppose $f(x)=f(y)$. Then $y^{-1} \sigma(y)=x^{-1} \sigma(x)$, so that $x y^{-1}=\sigma(x) \sigma\left(y^{-1}\right)$, and $x y^{-1}=\sigma\left(x y^{-1}\right)$. Then we have $x y^{-1}=1$, hence $x=y$. So $f$ is injective. Since $G$ is finite and $f$ is injective, $f$ is also surjective. Then every $z \in G$ is of the form $x^{-1} \sigma(x)$ for some $x$. Now let $z \in G$ with $z=x^{-1} \sigma(x)$. We have $$ \sigma(z)=\sigma\left(x^{-1} \sigma(x)\right)=\sigma(x)^{-1} x=\left(x^{-1} \sigma(x)\right)^{-1}=z^{-1} . $$ Thus $\sigma$ is in fact the inversion mapping, and we assumed that $\sigma$ is a homomorphism. By a previous example, then, $G$ is abelian. \end{proof}
Dummit-Foote|exercise_2_1_13
theorem exercise_2_1_13 (H : add_subgroup β„š) {x : β„š} (hH : x ∈ H β†’ (1 / x) ∈ H): H = βŠ₯ ∨ H = ⊀ :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $H$ be a subgroup of the additive group of rational numbers with the property that $1 / x \in H$ for every nonzero element $x$ of $H$. Prove that $H=0$ or $\mathbb{Q}$.
\begin{proof} Solution: First, suppose there does not exist a nonzero element in $H$. Then $H=0$. Now suppose there does exist a nonzero element $a \in H$; without loss of generality, say $a=p / q$ in lowest terms for some integers $p$ and $q$ - that is, $\operatorname{gcd}(p, q)=1$. Now $q \cdot \frac{p}{q}=p \in H$, and since $q / p \in H$, we have $p \cdot \frac{q}{p} \in H$. There exist integers $x, y$ such that $q x+p y=1$; note that $q x \in H$ and $p y \in H$, so that $1 \in H$. Thus $n \in H$ for all $n \in \mathbb{Z}$. Moreover, if $n \neq 0,1 / n \in H$. Then $m / n \in H$ for all integers $m, n$ with $n \neq 0$; hence $H=\mathbb{Q}$. \end{proof}
Dummit-Foote|exercise_2_4_16a
theorem exercise_2_4_16a {G : Type*} [group G] {H : subgroup G} (hH : H β‰  ⊀) : βˆƒ M : subgroup G, M β‰  ⊀ ∧ βˆ€ K : subgroup G, M ≀ K β†’ K = M ∨ K = ⊀ ∧ H ≀ M :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
A subgroup $M$ of a group $G$ is called a maximal subgroup if $M \neq G$ and the only subgroups of $G$ which contain $M$ are $M$ and $G$. Prove that if $H$ is a proper subgroup of the finite group $G$ then there is a maximal subgroup of $G$ containing $H$.
\begin{proof} If $H$ is maximal, then we are done. If $H$ is not maximal, then there is a subgroup $K_1$ of $G$ such that $H<K_1<G$. If $K_1$ is maximal, we are done. But if $K_1$ is not maximal, there is a subgroup $K_2$ with $H<K_1<K_2<G$. If $K_2$ is maximal, we are done, and if not, keep repeating the procedure. Since $G$ is finite, this process must eventually come to an end, so that $K_n$ is maximal for some positive integer $n$. Then $K_n$ is a maximal subgroup containing $H$. \end{proof}
Dummit-Foote|exercise_2_4_16c
theorem exercise_2_4_16c {n : β„•} (H : add_subgroup (zmod n)) : βˆƒ p : β„•, nat.prime p ∧ H = add_subgroup.closure {p} ↔ H β‰  ⊀ ∧ βˆ€ K : add_subgroup (zmod n), H ≀ K β†’ K = H ∨ K = ⊀ :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Show that if $G=\langle x\rangle$ is a cyclic group of order $n \geq 1$ then a subgroup $H$ is maximal if and only $H=\left\langle x^{p}\right\rangle$ for some prime $p$ dividing $n$.
\begin{proof} Suppose $H$ is a maximal subgroup of $G$. Then $H$ is cyclic, and we may write $H=\left\langle x^k\right\rangle$ for some integer $k$, with $k>1$. Let $d=(n, k)$. Since $H$ is a proper subgroup, we know by Proposition 6 that $d>1$. Choose a prime factor $p$ of $d$. If $k=p=d$ then $k \mid n$ as required. If, however, $k$ is not prime, then consider the subgroup $K=\left\langle x^p\right\rangle$. Since $p$ is a proper divisor of $k$, it follows that $H<K$. But $H$ is maximal, so we must have $K=G$. Again by Proposition 6 , we must then have $(p, n)=1$. However, $p$ divides $d$ which divides $n$, so $p \mid n$ and $(p, n)=p>1$, a contradiction. Therefore $k=p$ and the left-to-right implication holds. Now, for the converse, suppose $H=\left\langle x^p\right\rangle$ for $p$ a prime dividing $n$. If $H$ is not maximal then the first part of this exercise shows that there is a maximal subgroup $K$ containing $H$. Then $K=\left\langle x^q\right\rangle$. So $x^p \in\left\langle x^q\right\rangle$, which implies $q \mid p$. But the only divisors of $p$ are 1 and $p$. If $q=1$ then $K=G$ and $K$ cannot be a proper subgroup, and if $q=p$ then $H=K$ and $H$ cannot be a proper subgroup of $K$. This contradiction shows that $H$ is maximal. \end{proof}
Dummit-Foote|exercise_3_1_22a
theorem exercise_3_1_22a (G : Type*) [group G] (H K : subgroup G) [subgroup.normal H] [subgroup.normal K] : subgroup.normal (H βŠ“ K) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $H$ and $K$ are normal subgroups of a group $G$ then their intersection $H \cap K$ is also a normal subgroup of $G$.
\begin{proof} Suppose $H$ and $K$ are normal subgroups of $G$. We already know that $H \cap K$ is a subgroup of $G$, so we need to show that it is normal. Choose any $g \in G$ and any $x \in H \cap K$. Since $x \in H$ and $H \unlhd G$, we know $g x g^{-1} \in H$. Likewise, since $x \in K$ and $K \unlhd G$, we have $g x g^{-1} \in K$. Therefore $g x g^{-1} \in H \cap K$. This shows that $g(H \cap K) g^{-1} \subseteq H \cap K$, and this is true for all $g \in G$. By Theorem 6 (5) (which we will prove in Exercise 3.1.25), this is enough to show that $H \cap K \unlhd G$. \end{proof}
Dummit-Foote|exercise_3_2_8
theorem exercise_3_2_8 {G : Type*} [group G] (H K : subgroup G) [fintype H] [fintype K] (hHK : nat.coprime (fintype.card H) (fintype.card K)) : H βŠ“ K = βŠ₯ :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $H$ and $K$ are finite subgroups of $G$ whose orders are relatively prime then $H \cap K=1$.
\begin{proof} Solution: Let $|H|=p$ and $|K|=q$. We saw in a previous exercise that $H \cap K$ is a subgroup of both $H$ and $K$; by Lagrange's Theorem, then, $|H \cap K|$ divides $p$ and $q$. Since $\operatorname{gcd}(p, q)=1$, then, $|H \cap K|=1$. Thus $H \cap K=1$. \end{proof}
Dummit-Foote|exercise_3_2_16
theorem exercise_3_2_16 (p : β„•) (hp : nat.prime p) (a : β„•) : nat.coprime a p β†’ a ^ p ≑ a [ZMOD p] :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Use Lagrange's Theorem in the multiplicative group $(\mathbb{Z} / p \mathbb{Z})^{\times}$to prove Fermat's Little Theorem: if $p$ is a prime then $a^{p} \equiv a(\bmod p)$ for all $a \in \mathbb{Z}$.
\begin{proof} Solution: If $p$ is prime, then $\varphi(p)=p-1$ (where $\varphi$ denotes the Euler totient). Thus $$ \mid\left((\mathbb{Z} /(p))^{\times} \mid=p-1 .\right. $$ So for all $a \in(\mathbb{Z} /(p))^{\times}$, we have $|a|$ divides $p-1$. Hence $$ a=1 \cdot a=a^{p-1} a=a^p \quad(\bmod p) . $$ \end{proof}
Dummit-Foote|exercise_3_3_3
theorem exercise_3_3_3 {p : primes} {G : Type*} [group G] {H : subgroup G} [hH : H.normal] (hH1 : H.index = p) : βˆ€ K : subgroup G, K ≀ H ∨ H βŠ” K = ⊀ ∨ (K βŠ“ H).relindex K = p :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $H$ is a normal subgroup of $G$ of prime index $p$ then for all $K \leq G$ either $K \leq H$, or $G=H K$ and $|K: K \cap H|=p$.
\begin{proof} Solution: Suppose $K \backslash N \neq \emptyset$; say $k \in K \backslash N$. Now $G / N \cong \mathbb{Z} /(p)$ is cyclic, and moreover is generated by any nonidentity- in particular by $\bar{k}$ Now $K N \leq G$ since $N$ is normal. Let $g \in G$. We have $g N=k^a N$ for some integer a. In particular, $g=k^a n$ for some $n \in N$, hence $g \in K N$. We have $[K: K \cap N]=p$ by the Second Isomorphism Theorem. \end{proof}
Dummit-Foote|exercise_3_4_4
theorem exercise_3_4_4 {G : Type*} [comm_group G] [fintype G] {n : β„•} (hn : n ∣ (fintype.card G)) : βˆƒ (H : subgroup G) (H_fin : fintype H), @card H H_fin = n :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Use Cauchy's Theorem and induction to show that a finite abelian group has a subgroup of order $n$ for each positive divisor $n$ of its order.
\begin{proof} Let $G$ be a finite abelian group. We use induction on $|G|$. Certainly the result holds for the trivial group. And if $|G|=p$ for some prime $p$, then the positive divisors of $|G|$ are 1 and $p$ and the result is again trivial. Now assume that the statement is true for all groups of order strictly smaller than $|G|$, and let $n$ be a positive divisor of $|G|$ with $n>1$. First, if $n$ is prime then Cauchy's Theorem allows us to find an element $x \in G$ having order $n$. Then $\langle x\rangle$ is the desired subgroup. On the other hand, if $n$ is not prime, then $n$ has a prime divisor $p$, so that $n=k p$ for some integer $k$. Cauchy's Theorem allows us to find an element $x$ having order $p$. Set $N=\langle x\rangle$. By Lagrange's Theorem, $$ |G / N|=\frac{|G|}{|N|}<|G| . $$ Now, by the inductive hypothesis, the group $G / N$ must have a subgroup of order $k$. And by the Lattice Isomorphism Theorem, this subgroup has the form $H / N$ for some subgroup $H$ of $G$. Then $|H|=k|N|=k p=n$, so that $H$ has order $n$. This completes the inductive step. \end{proof}
Dummit-Foote|exercise_3_4_5b
theorem exercise_3_4_5b {G : Type*} [group G] [is_solvable G] (H : subgroup G) [subgroup.normal H] : is_solvable (G β§Έ H) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that quotient groups of a solvable group are solvable.
\begin{proof} Next, note that $$ H_i=G_i \cap H=\left(G_i \cap G_{i+1}\right) \cap H=G_i \cap H_{i+1} . $$ By the Second Isomorphism Theorem, we then have $$ H_{i+1} / H_i=H_{i+1} /\left(H_{i+1} \cap G_i\right) \cong H_{i+1} G_i / G_i \leq G_{i+1} / G_i . $$ Since $H_{i+1} / H_i$ is isomorphic to a subgroup of the abelian group $G_{i+1} / G_i$, it follows that $H_{i+1} / H_i$ is also abelian. This completes the proof that $H$ is solvable. Next, let $N \unlhd G$. For each $i$, define $$ N_i=G_i N, \quad 0 \leq i \leq n . $$ Now let $g \in N_{i+1}$, where $g=g_0 n_0$ with $g_0 \in G_{i+1}$ and $n_0 \in N$. Also let $x \in N_i$, where $x=g_1 n_1$ with $g_1 \in G_i$ and $n_1 \in N$. Then $$ g x g^{-1}=g_0 n_0 g_1 n_1 n_0^{-1} g_0^{-1} . $$ Now, since $N$ is normal in $G, N g=g N$, so $n_0 g_1=g_1 n_2$ for some $n_2 \in N$. Then $$ g x g^{-1}=g_0 g_1\left(n_2 n_1 n_0^{-1}\right) g_0^{-1}=g_0 g_1 n_3 g_0^{-1} $$ for some $n_3 \in N$. Then $n_3 g_0^{-1}=g_0^{-1} n_4$ for some $n_4 \in N$. And $g_0 g_1 g_0^{-1} \in G_i$ since $G_i \unlhd G_{i+1}$, so $$ g x g^{-1}=g_0 g_1 g_0^{-1} n_4 \in N_i . $$ This shows that $N_i \unlhd N_{i+1}$. So by the Lattice Isomorphism Theorem, we have $N_{i+1} / N \unlhd N_i / N$. This shows that $$ 1=N_0 / N \unlhd N_1 / N \unlhd N_2 / N \unlhd \cdots \unlhd N_n / N=G / N . $$ Moreover, the Third Isomorphism Theorem says that $$ \left(N_{i+1} / N\right) /\left(N_i / N\right) \cong N_{i+1} / N_i, $$ so the proof will be complete if we can show that $N_{i+1} / N_i$ is abelian. Let $x, y \in N_{i+1} / N_i$. Then $$ x=\left(g_0 n_0\right) N_i \quad \text { and } \quad y=\left(g_1 n_1\right) N_i $$ for some $g_0, g_1 \in G_{i+1}$ and $n_0, n_1 \in N$. We have $$ \begin{aligned} x y x^{-1} y^{-1} & =\left(g_0 n_0\right)\left(g_1 n_1\right)\left(g_0 n_0\right)^{-1}\left(g_1 n_1\right)^{-1} N_i \\ & =g_0 n_0 g_1 n_1 n_0^{-1} g_0^{-1} n_1^{-1} g_1^{-1} N_i . \end{aligned} $$ Since $N \unlhd G, g N=N g$ for any $g \in G$, so we can find $n_2 \in N$ such that $$ x y x^{-1} y^{-1}=g_0 g_1 g_0^{-1} g^{-1} n_2 N_i . $$ Now $N_i=G_i N=N G_i$ since $N \unlhd G$ (see Proposition 14 and its corollary). Therefore $$ n_2 N_i=n_2 N G_i=N G_i=G_i N $$ and we get $$ x y x^{-1} y^{-1}=g_0 g_1 g_0^{-1} g^{-1} G_i N=G_i N . $$ So $x y x^{-1} y^{-1}=1 N_i$ or $x y=y x$. This completes the proof that $G / N$ is solvable. \end{proof}
Dummit-Foote|exercise_4_2_8
theorem exercise_4_2_8 {G : Type*} [group G] {H : subgroup G} {n : β„•} (hn : n > 0) (hH : H.index = n) : βˆƒ K ≀ H, K.normal ∧ K.index ≀ n.factorial :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $H$ has finite index $n$ then there is a normal subgroup $K$ of $G$ with $K \leq H$ and $|G: K| \leq n!$.
\begin{proof} Solution: $G$ acts on the cosets $G / H$ by left multiplication. Let $\lambda: G \rightarrow S_{G / H}$ be the permutation representation induced by this action, and let $K$ be the kernel of the representation. Now $K$ is normal in $G$, and $K \leq \operatorname{stab}_G(H)=H$. By the First Isomorphism Theorem, we have an injective group homomorphism $\bar{\lambda}: G / K \rightarrow S_{G / H}$. Since $\left|S_{G / H}\right|=n !$, we have $[G: K] \leq n !$. \end{proof}
Dummit-Foote|exercise_4_2_9a
theorem exercise_4_2_9a {G : Type*} [fintype G] [group G] {p Ξ± : β„•} (hp : p.prime) (ha : Ξ± > 0) (hG : card G = p ^ Ξ±) : βˆ€ H : subgroup G, H.index = p β†’ H.normal :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $p$ is a prime and $G$ is a group of order $p^{\alpha}$ for some $\alpha \in \mathbb{Z}^{+}$, then every subgroup of index $p$ is normal in $G$.
\begin{proof} Solution: Let $G$ be a group of order $p^k$ and $H \leq G$ a subgroup with $[G: H]=p$. Now $G$ acts on the conjugates $g H g^{-1}$ by conjugation, since $$ g_1 g_2 \cdot H=\left(g_1 g_2\right) H\left(g_1 g_2\right)^{-1}=g_1\left(g_2 H g_2^{-1}\right) g_1^{-1}=g_1 \cdot\left(g_2 \cdot H\right) $$ and $1 \cdot H=1 H 1=H$. Moreover, under this action we have $H \leq \operatorname{stab}(H)$. By Exercise 3.2.11, we have $$ [G: \operatorname{stab}(H)][\operatorname{stab}(H): H]=[G: H]=p, $$ a prime. If $[G: \operatorname{stab}(H)]=p$, then $[\operatorname{stab}(H): H]=1$ and we have $H=\operatorname{stab}(H)$; moreover, $H$ has exactly $p$ conjugates in $G$. Let $\varphi: G \rightarrow S_p$ be the permutation representation induced by the action of $G$ on the conjugates of $H$, and let $K$ be the kernel of this representation. Now $K \leq \operatorname{stab}(H)=H$. By the first isomorphism theorem, the induced map $\bar{\varphi}: G / K \rightarrow S_p$ is injective, so that $|G / K|$ divides $p$ !. Note, however, that $|G / K|$ is a power of $p$ and that the only powers of $p$ that divide $p$ ! are 1 and $p$. So $[G: K]$ is 1 or $p$. If $[G: K]=1$, then $G=K$ so that $g H g^{-1}=H$ for all $g \in G$; then $\operatorname{stab}(H)=G$ and we have $[G: \operatorname{stab}(H)]=1$, a contradiction. Now suppose $[G: K]=p$. Again by Exercise $3.2$.11 we have $[G: K]=[G: H][H: K]$, so that $[H: K]=1$, hence $H=K$. Again, this implies that $H$ is normal so that $g H g^{-1}=H$ for all $g \in G$, and we have $[G: \operatorname{stab}(H)]=1$, a contradiction. Thus $[G: \operatorname{stab}(H)] \neq p$ If $[G: \operatorname{stab}(H)]=1$, then $G=\operatorname{stab}(H)$. That is, $g H g^{-1}=H$ for all $g \in G$; thus $H \leq G$ is normal. \end{proof}
Dummit-Foote|exercise_4_4_2
theorem exercise_4_4_2 {G : Type*} [fintype G] [group G] {p q : nat.primes} (hpq : p β‰  q) (hG : card G = p*q) : is_cyclic G :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $G$ is an abelian group of order $p q$, where $p$ and $q$ are distinct primes, then $G$ is cyclic.
\begin{proof} Let $G$ be an abelian group of order $p q$. We need to prove that if $p$ and $q$ are distinct primes than $G$ is cyclic. By Cauchy's theorem there are $a, b \in G$ with $a$ of order $p$ and $b$ of order $q$. Since $(|a|,|b|)=1$ and $a b=b a$ then $|a b|=|a| \cdot|b|=p q$. Therefore $a b$ is an element of order $p q$, the order of $G$, which means $G$ is cyclic. \end{proof}
Dummit-Foote|exercise_4_4_6b
theorem exercise_4_4_6b : βˆƒ (G : Type*) (hG : group G) (H : @subgroup G hG), @characteristic G hG H ∧ Β¬ @subgroup.normal G hG H :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that there exists a normal subgroup that is not characteristic.
\begin{proof} We have to produce a group $G$ and a subgroup $H$ such that $H$ is normal in $G$, but not characterestic. Consider the Klein's four group $G=\{ e, a, b, a b\}$. This is an abelian group with each element having order 2. Consider $H=\{ e, a\}$. $H$ is normal in $G$. Define $\sigma: G \rightarrow G$ as $\sigma(a)=b, \sigma(b)=a, \sigma(a b)=a b$. Clearly $\sigma$ does not fix $H$. So, $H$ is not characterestic. \end{proof}
Dummit-Foote|exercise_4_4_8a
theorem exercise_4_4_8a {G : Type*} [group G] (H K : subgroup G) (hHK : H ≀ K) [hHK1 : (H.subgroup_of K).normal] [hK : K.normal] : H.normal :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $G$ be a group with subgroups $H$ and $K$ with $H \leq K$. Prove that if $H$ is characteristic in $K$ and $K$ is normal in $G$ then $H$ is normal in $G$.
\begin{proof} We prove that $H$ is invariant under every inner automorphism of $G$. Consider a inner automorphism $\phi_g$ of $G$. Now, $\left.\phi_g\right|_K$ is a automorphism of $K$ because $K$ is normal in $G$. But $H$ is a characterestic subgroup of $K$, so $\left.\phi_g\right|_K(H) \subset H$, so in general $\phi_g(H) \subset H$. Hence $H$ is characteretstic in $G$. \end{proof}
Dummit-Foote|exercise_4_5_13
theorem exercise_4_5_13 {G : Type*} [group G] [fintype G] (hG : card G = 56) : βˆƒ (p : β„•) (P : sylow p G), P.normal :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that a group of order 56 has a normal Sylow $p$-subgroup for some prime $p$ dividing its order.
\begin{proof} Since $|G|=56=2^{3}.7$, $G$ has $2-$Sylow subgroup of order $8$, as well as $7-$Sylow subgroup of order $7$. Now, we count the number of such subgroups. Let $n_{7}$ be the number of $7-$Sylow subgroup and $n_{2}$ be the number of $2-$Sylow subgroup. Now $n_{7}=1+7k$ where $1+7k|8$. The choices for $k$ are $0$ or $1$. If $k=0$, there is only one $7-$Sylow subgroup and hence normal. So, assume now, that there are $8$ $7-$Sylow subgroup(for $k=1$). Now we look at $2-$ Sylow subgroups. $n_{2}=1+2k| 7$. So choice for $k$ are $0$ and $3$. If $k=0$, there is only one $2-$Sylow subgroup and hence normal. So, assume now, that there are $7$ $2-$Sylow subgroup (for $k=3$). Now we claim that simultaneously, there cannot be $8$ $7-$Sylow subgroup and $7$ $2-$Sylow subgroup. So, either $7-$Sylow subgroup is normal being unique, or the $2-$Sylow subgroup is normal. Now, to prove the claim, we observe that there are 48 elements of order $7$. Let $H_{1}$ and $H_{2}$ be two distinct $2-$Sylow subgroup. Now $|H_{1}|=8$. So we already get $48+8=56$ distinct elements in the group. Now $H_{2}$ being distinct from $H_{1}$, has at least one element which is not in $H_{1}$. This adds one more element in the group, at the least. Now already we have number of elements in the group exceeding the number of element in $G$. This gives a contradiction and proves the claim. \end{proof}
Dummit-Foote|exercise_4_5_15
theorem exercise_4_5_15 {G : Type*} [group G] [fintype G] (hG : card G = 351) : βˆƒ (p : β„•) (P : sylow p G), P.normal :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that a group of order 351 has a normal Sylow $p$-subgroup for some prime $p$ dividing its order.
\begin{proof} Since $|G|=351=3^{2}.13$, $G$ has $3-$Sylow subgroup of order $9$, as well as $13-$Sylow subgroup of order $13$. Now, we count the number of such subgroups. Let $n_{13}$ be the number of $13-$Sylow subgroup and $n_{3}$ be the number of $3-$Sylow subgroup. Now $n_{13}=1+13k$ where $1+13k|9$. The choices for $k$ is $0$. Hence, there is a unique $13-$Sylow subgroup and hence is normal. \end{proof}
Dummit-Foote|exercise_4_5_17
theorem exercise_4_5_17 {G : Type*} [fintype G] [group G] (hG : card G = 105) : nonempty(sylow 5 G) ∧ nonempty(sylow 7 G) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $|G|=105$ then $G$ has a normal Sylow 5 -subgroup and a normal Sylow 7-subgroup.
\begin{proof} Since $|G|=105=3.5.7$, $G$ has $3-$Sylow subgroup of order $3$, as well as $5-$Sylow subgroup of order $5$ and, $7-$Sylow subgroup of order 7. Now, we count the number of such subgroups. Let $n_{3}$ be the number of $3-$Sylow subgroup, $n_{5}$ be the number of $5-$Sylow subgroup, and $n_{7}$ be the number of $7-$Sylow subgroup. Now $n_{7}=1+7k$ where $1+7k|15$. The choices for $k$ are $0$ or $1$. If $k=0$, there is only one $7-$Sylow subgroup and hence normal. So, assume now, that there are $15$ $7-$Sylow subgroup(for $k=1$). Now we look at $5-$ Sylow subgroups. $n_{5}=1+5k| 21$. So choice for $k$ are $0$ and $4$. If $k=0$, there is only one $5-$Sylow subgroup and hence normal. So, assume now, that there are $24$ $5-$Sylow subgroup (for $k=4$). Now we claim that simultaneously, there cannot be $15$ $7-$Sylow subgroup and $24$ $5-$Sylow subgroup. So, either $7-$Sylow subgroup is normal being unique, or the $5-$Sylow subgroup is normal. Now, to prove the claim, we observe that there are 90 elements of order $7$. Also, see that there are $24\times 4=96$ number of elements of order 5. So we get $90+94=184$ number of elements which exceeds the order of the group. This gives a contradiction and proves the claim. So, now we have proved that there is either a normal $5-$Sylow subgroup or a normal $7-$Sylow subgroup. Now we prove that indeed both $5-$ Sylow subgroup and 7 -Sylow subgroup are normal. Assume that 7 -Sylow subgroup is normal. So, there is a unique 7 -Sylow subgroup, say $H$. Now assume that there are 245 -Sylow subgroups. So, we get again $24 \times 4=96$ elements of order 5 . From $H$ we get 7 elements which gives us total of $96+7=103$ elements. Now consider the number of 3 -Sylow subgroups. $n_3=1+3 k \mid 35$. Then the possibilities for $k$ are 0 and 2 . But we can rule out $k=2$ because having 73 -Sylow subgroup, will mean we have 14 elements of order 3 . So we get $103+14=117$ elements in total which exceeds the order of the group. So we have now that there is a unique 3 -Sylow subgroup and hence normal. Call that subgroup $K$. Now take any one 5 -Sylow subgroup, call it $L$. Now observe $L K$ is a subgroup of $G$ with order 15 . We know that a group of order 15 is cyclic by an example in Page-143 of the book. So, there is an element of order 15. Actually we have $\phi(15)=8$ number of elements of order 15. But then again we already had 103 elements and then we actually get at least $103+8=111$ elements which exceeds the order of the group. So, there can't be 24 5-Sylow subgroups, and hence there is a unique 5-Sylow subgroup, and hence normal. \end{proof}
Dummit-Foote|exercise_4_5_19
theorem exercise_4_5_19 {G : Type*} [fintype G] [group G] (hG : card G = 6545) : Β¬ is_simple_group G :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $|G|=6545$ then $G$ is not simple.
\begin{proof} Since $|G|=132=2^{2}.3.11$, $G$ has $2-$Sylow subgroup of order $4$, as well as $11-$Sylow subgroup of order $11$, and $3-$Sylow subgroup of order $3$. Now, we count the number of such subgroups. Let $n_{11}$ be the number of $11-$Sylow subgroup and $n_{3}$ be the number of $3-$Sylow subgroup. Now $n_{11}=1+11k$ where $1+11k|12$. The choices for $k$ are $0$ or $1$. If $k=0$, there is only one $11-$Sylow subgroup and hence normal. So, assume now, that there are $12$ $11-$Sylow subgroup(for $k=1$). Now we look at $3-$ Sylow subgroups. $n_{3}=1+3k| 44$. So choice for $k$ are $0$, $1$, and $7$. If $k=0$, there is only one $3-$Sylow subgroup and hence normal. So, assume now, that there are $4$ $2-$Sylow subgroup (for $k=3$). Now we claim that simultaneously, there cannot be $12$ $11-$Sylow subgroup and $4$ $3-$Sylow subgroups provided there is more than one $2-$Sylow subgroups. So, either $2-$Sylow subgroup is normal or if not, then, either $11-$Sylow subgroup is normal being unique, or the $3-$Sylow subgroup is normal(We don't consider the possibility of $22$ $3-$Sylow subgroup because of obvious reason). Now, to prove the claim, we observe that there are $120$ elements of order $11$. Also there are $8$ elements of order $3$. So we already get $120+8+1=129$ distinct elements in the group. Let us count the number of $2-$Sylow subgroups in $G$. $n_{2}=1+2k|33$. The possibilities for $k$ are $0$, $1$, $5$, $16$. Now, assume there is more than one $2-$Sylow subgroups. Let $H_{1}$ and $H_{2}$ be two distinct $2-$Sylow subgroup. Now $|H_{1}|=4$. So we already get $129+3=132$ distinct elements in the group. Now $H_{2}$ being distinct from $H_{1}$, has at least one element which is not in $H_{1}$. This adds one more element in the group, at the least. Now already we have number of elements in the group exceeding the number of element in $G$. This gives a contradiction and proves the claim. Hence $G$ is not simple. \end{proof}
Dummit-Foote|exercise_4_5_21
theorem exercise_4_5_21 {G : Type*} [fintype G] [group G] (hG : card G = 2907) : Β¬ is_simple_group G :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $|G|=2907$ then $G$ is not simple.
\begin{proof} Since $|G|=2907=3^{2}.17.19$, $G$ has $19-$Sylow subgroup of order $19$. Now, we count the number of such subgroups. Let $n_{19}$ be the number of $19-$Sylow subgroup. Now $n_{19}=1+19k$ where $1+19k|3^{2}.17$. The choices for $k$ is $0$. Hence, there is a unique $19-$Sylow subgroup and hence is normal. so $G$ is not simple. \end{proof}
Dummit-Foote|exercise_4_5_23
theorem exercise_4_5_23 {G : Type*} [fintype G] [group G] (hG : card G = 462) : Β¬ is_simple_group G :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $|G|=462$ then $G$ is not simple.
\begin{proof} Let $G$ be a group of order $462=11 \cdot 42$. Note that 11 is a prime not dividing 42 . Let $P \in$ $S y l_{11}(G)$. [We know $P$ exists since $S y l_{11}(G) \neq \emptyset$]. Note that $|P|=11^1=11$ by definition. The number of Sylow 11-subgroups of $G$ is of the form $1+k \cdot 11$, i.e., $n_{11} \equiv 1$ (mod 11) and $n_{11}$ divides 42 . The only such number that divides 42 and equals 1 (mod 11) is 1 so $n_{11}=1$. Hence $P$ is the unique Sylow 11-subgroup. Since $P$ is the unique Sylow Il-subgroup, this implies that $P$ is normal in $G$. \end{proof}
Dummit-Foote|exercise_4_5_33
theorem exercise_4_5_33 {G : Type*} [group G] [fintype G] {p : β„•} (P : sylow p G) [hP : P.normal] (H : subgroup G) [fintype H] : βˆ€ R : sylow p H, R.to_subgroup = (H βŠ“ P.to_subgroup).subgroup_of H ∧ nonempty (sylow p H) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $P$ be a normal Sylow $p$-subgroup of $G$ and let $H$ be any subgroup of $G$. Prove that $P \cap H$ is the unique Sylow $p$-subgroup of $H$.
\begin{proof} Let $G$ be a group and $P$ is a normal $p$-Sylow subgroup of $G .|G|=p^a . m$ where $p \nmid m$. Then $|P|=p^a$. Let $H$ be a subgroup of $G$. Now if $|H|=k$ such that $p \nmid k$. Then $P \cap H=\{e\}$. There is nothing to prove in this case. Let $|H|=p^b . n$, where $b \leq a$, and $p \nmid n$. Now consider $P H$ which is a subgroup of $G$, as $P$ is normal. Now $|P H|=\frac{|P||H|}{|P \cap H|}=\frac{p^{a+b} \cdot n}{|P \cap H|}$. Now since $P H \leq G$, so $|P H|=p^a$.l, as $P \leq P H$. This forces $|P \cap H|=p^b$. So by order consideration we have $P \cap H$ is a sylow $-p$ subgroup of $H$. Now we know $P$ is unique $p$ - Sylow subgroup. Suppose $H$ has a sylow-p subgroup distinct from $P \cap H$, call it $H_1$. Now $H_1$ is a p-subgroup of $G$. So, $H_1$ is contained in some Sylow-p subgroup of $G$, call it $P_1$. Clearly $P_1$ is distinct from $P$, which is a contradiction. So $P \cap H$ is the only $p$-Sylow subgroup of $H$, and hence normal in $H$ \end{proof}
Dummit-Foote|exercise_7_1_2
theorem exercise_7_1_2 {R : Type*} [ring R] {u : R} (hu : is_unit u) : is_unit (-u) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that if $u$ is a unit in $R$ then so is $-u$.
\begin{proof} Solution: Since $u$ is a unit, we have $u v=v u=1$ for some $v \in R$. Thus, we have $$ (-v)(-u)=v u=1 $$ and $$ (-u)(-v)=u v=1 . $$ Thus $-u$ is a unit. \end{proof}
Dummit-Foote|exercise_7_1_12
theorem exercise_7_1_12 {F : Type*} [field F] {K : subring F} (hK : (1 : F) ∈ K) : is_domain K :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that any subring of a field which contains the identity is an integral domain.
\begin{proof} Solution: Let $R \subseteq F$ be a subring of a field. (We need not yet assume that $1 \in R$ ). Suppose $x, y \in R$ with $x y=0$. Since $x, y \in F$ and the zero element in $R$ is the same as that in $F$, either $x=0$ or $y=0$. Thus $R$ has no zero divisors. If $R$ also contains 1 , then $R$ is an integral domain. \end{proof}
Dummit-Foote|exercise_7_2_2
theorem exercise_7_2_2 {R : Type*} [ring R] (p : polynomial R) : p ∣ 0 ↔ βˆƒ b : R, b β‰  0 ∧ b β€’ p = 0 :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $p(x)=a_{n} x^{n}+a_{n-1} x^{n-1}+\cdots+a_{1} x+a_{0}$ be an element of the polynomial ring $R[x]$. Prove that $p(x)$ is a zero divisor in $R[x]$ if and only if there is a nonzero $b \in R$ such that $b p(x)=0$.
\begin{proof} Solution: If $b p(x)=0$ for some nonzero $b \in R$, then it is clear that $p(x)$ is a zero divisor. Now suppose $p(x)$ is a zero divisor; that is, for some $q(x)=\sum_{i=0}^m b_i x^i$, we have $p(x) q(x)=0$. We may choose $q(x)$ to have minimal degree among the nonzero polynomials with this property. We will now show by induction that $a_i q(x)=0$ for all $0 \leq i \leq n$. For the base case, note that $$ p(x) q(x)=\sum_{k=0}^{n+m}\left(\sum_{i+j=k} a_i b_j\right) x^k=0 . $$ The coefficient of $x^{n+m}$ in this product is $a_n b_m$ on one hand, and 0 on the other. Thus $a_n b_m=0$. Now $a_n q(x) p(x)=0$, and the coefficient of $x^m$ in $q$ is $a_n b_m=0$. Thus the degree of $a_n q(x)$ is strictly less than that of $q(x)$; since $q(x)$ has minimal degree among the nonzero polynomials which multiply $p(x)$ to 0 , in fact $a_n q(x)=0$. More specifically, $a_n b_i=0$ for all $0 \leq i \leq m$. For the inductive step, suppose that for some $0 \leq t<n$, we have $a_r q(x)=0$ for all $t<r \leq n$. Now $$ p(x) q(x)=\sum_{k=0}^{n+m}\left(\sum_{i+j=k} a_i b_j\right) x^k=0 . $$ On one hand, the coefficient of $x^{m+t}$ is $\sum_{i+j=m+t} a_i b_j$, and on the other hand, it is 0 . Thus $$ \sum_{i+j=m+t} a_i b_j=0 . $$ By the induction hypothesis, if $i \geq t$, then $a_i b_j=0$. Thus all terms such that $i \geq t$ are zero. If $i<t$, then we must have $j>m$, a contradiction. Thus we have $a_t b_m=0$. As in the base case, $$ a_t q(x) p(x)=0 $$ and $a_t q(x)$ has degree strictly less than that of $q(x)$, so that by minimality, $a_t q(x)=0$. By induction, $a_i q(x)=0$ for all $0 \leq i \leq n$. In particular, $a_i b_m=0$. Thus $b_m p(x)=0$. \end{proof}
Dummit-Foote|exercise_7_3_16
theorem exercise_7_3_16 {R S : Type*} [ring R] [ring S] {Ο† : R β†’+* S} (hf : surjective Ο†) : Ο† '' (center R) βŠ‚ center S :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $\varphi: R \rightarrow S$ be a surjective homomorphism of rings. Prove that the image of the center of $R$ is contained in the center of $S$.
\begin{proof} Suppose $r \in \varphi[Z(R)]$. Then $r=\varphi(z)$ for some $z \in Z(R)$. Now let $x \in S$. Since $\varphi$ is surjective, we have $x=\varphi y$ for some $y \in R$. Now $$ x r=\varphi(y) \varphi(z)=\varphi(y z)=\varphi(z y)=\varphi(z) \varphi(y)=r x . $$ Thus $r \in Z(S)$. \end{proof}
Dummit-Foote|exercise_7_4_27
theorem exercise_7_4_27 {R : Type*} [comm_ring R] (hR : (0 : R) β‰  1) {a : R} (ha : is_nilpotent a) (b : R) : is_unit (1-a*b) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $R$ be a commutative ring with $1 \neq 0$. Prove that if $a$ is a nilpotent element of $R$ then $1-a b$ is a unit for all $b \in R$.
\begin{proof} $\mathfrak{N}(R)$ is an ideal of $R$. Thus for all $b \in R,-a b$ is nilpotent. Hence $1-a b$ is a unit in $R$. \end{proof}
Dummit-Foote|exercise_8_2_4
theorem exercise_8_2_4 {R : Type*} [ring R][no_zero_divisors R] [cancel_comm_monoid_with_zero R] [gcd_monoid R] (h1 : βˆ€ a b : R, a β‰  0 β†’ b β‰  0 β†’ βˆƒ r s : R, gcd a b = r*a + s*b) (h2 : βˆ€ a : β„• β†’ R, (βˆ€ i j : β„•, i < j β†’ a i ∣ a j) β†’ βˆƒ N : β„•, βˆ€ n β‰₯ N, βˆƒ u : R, is_unit u ∧ a n = u * a N) : is_principal_ideal_ring R :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $R$ be an integral domain. Prove that if the following two conditions hold then $R$ is a Principal Ideal Domain: (i) any two nonzero elements $a$ and $b$ in $R$ have a greatest common divisor which can be written in the form $r a+s b$ for some $r, s \in R$, and (ii) if $a_{1}, a_{2}, a_{3}, \ldots$ are nonzero elements of $R$ such that $a_{i+1} \mid a_{i}$ for all $i$, then there is a positive integer $N$ such that $a_{n}$ is a unit times $a_{N}$ for all $n \geq N$.
\begin{proof} Let $I \leq R$ be a nonzero ideal and let $I / \sim$ be the set of equivalence classes of elements of $I$ with regards to the relation of being associates. We can equip $I / \sim$ with a partial order with $[x] \leq[y]$ if $y \mid x$. Condition (ii) implies all chains in $I / \sim$ have an upper bound, so By Zorn's lemma $I / \sim$ contains a maximal element, i.e. $I$ contains a class of associated elements which are minimal with respect to divisibility. Now let $a, b \in I$ be two elements such that $[a]$ and $[b]$ are minimal with respect to divisibility. By condition (i) $a$ and $b$ have a greatest common divisor $d$ which can be expressed as $d=$ $a x+b y$ for some $x, y \in R$. In particular, $d \in I$. Since $a$ and $b$ are minimal with respect to divisibility, we have that $[a]=[b]=[d]$. Therefore $I$ has at least one element $a$ that is minimal with regard to divisibility and all such elements are associate, and we have $I=\langle a\rangle$ and so $I$ is principal. We conclude $R$ is a principal ideal domain. \end{proof}
Dummit-Foote|exercise_8_3_5a
theorem exercise_8_3_5a {n : β„€} (hn0 : n > 3) (hn1 : squarefree n) : irreducible (2 :zsqrtd $ -n) ∧ irreducible (⟨0, 1⟩ : zsqrtd $ -n) ∧ irreducible (1 + ⟨0, 1⟩ : zsqrtd $ -n) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $R=\mathbb{Z}[\sqrt{-n}]$ where $n$ is a squarefree integer greater than 3. Prove that $2, \sqrt{-n}$ and $1+\sqrt{-n}$ are irreducibles in $R$.
\begin{proof} Suppose $a=a_1+a_2 \sqrt{-n}, b=b_1+b_2 \sqrt{-n} \in R$ are such that $2=a b$, then $N(a) N(b)=4$. Without loss of generality we can assume $N(a) \leq N(b)$, so $N(a)=1$ or $N(a)=2$. Suppose $N(a)=2$, then $a_1^2+n a_2^2=2$ and since $n>3$ we have $a_2=0$, which implies $a_1^2=2$, a contradiction. So $N(a)=1$ and $a$ is a unit. Therefore 2 is irreducible in $R$. Suppose now $\sqrt{-n}=a b$, then $N(a) N(b)=n$ and we can assume $N(a)<$ $N(b)$ since $n$ is square free. Suppose $N(a)>1$, then $a_1^2+n a_2^2>1$ and $a_1^2+n a_2^2 \mid n$, so $a_2=0$, and therefore $a_1^2 \mid n$. Since $n$ is squarefree, $a_1=\pm 1$, a contradiction. Therefore $N(a)=1$ and so $a$ is a unit and $\sqrt{-n}$ is irreducible. Suppose $1+\sqrt{-n}=a b$, then $N(a) N(b)=n+1$ and we can assume $N(a) \leq N(b)$. Suppose $N(a)>1$, then $a_1^2+n a_2^2>1$ and $a_1^2+n a_2^2 \mid n+1$. If $\left|a_2\right| \geq 2$, then since $n>3$ we have a contradiction since $N(a)$ is too large. If $\left|a_2\right|=1$, then $a_1^2+n$ divides $1+n$ and so $a_1=\pm 1$, and in either case $N(a)=n+1$ which contradicts $N(a) \leq N(b)$. If $a_2=0$ then $a_1^2\left(b_1^2+n b_2^2\right)=\left(a_1 b_1\right)^2+n\left(a_1 b_2\right)^2=n+1$. If $\left|a_1 b_2\right| \geq 2$ we have a contradiction. If $\left|a_1 b_2\right|=1$ then $a_1=\pm 1$ which contradicts $N(a)>1$. If $\left|a_1 b_2\right|=0$, then $b_2=0$ and so $a_1 b_1=\sqrt{-n}$, a contradiction. Therefore $N(a)=1$ and so $a$ is a unit and $1+\sqrt{-n}$ is irreducible. \end{proof}
Dummit-Foote|exercise_8_3_6b
theorem exercise_8_3_6b {q : β„•} (hq0 : q.prime) (hq1 : q ≑ 3 [ZMOD 4]) {R : Type*} [ring R] (hR : R = (gaussian_int β§Έ ideal.span ({q} : set gaussian_int))) : is_field R ∧ βˆƒ finR : fintype R, @card R finR = q^2 :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Let $q \in \mathbb{Z}$ be a prime with $q \equiv 3 \bmod 4$. Prove that the quotient ring $\mathbb{Z}[i] /(q)$ is a field with $q^{2}$ elements.
\begin{proof} The division algorithm gives us that every element of $\mathbb{Z}[i] /\langle q\rangle$ is represented by an element $a+b i$ such that $0 \leq a, b<q$. Each such choice is distinct since if $a_1+b_1 i+\langle q\rangle=a_2+b_2 i+\langle q\rangle$, then $\left(a_1-a_2\right)+\left(b_1-b_2\right) i$ is divisible by $q$, so $a_1 \equiv a_2 \bmod q$ and $b_1 \equiv b_2 \bmod q$. So $\mathbb{Z}[i] /\langle q\rangle$ has order $q^2$. Since $q \equiv 3 \bmod 4, q$ is irreducible, hence prime in $\mathbb{Z}[i]$. Therefore $\langle q\rangle$ is a prime ideal in $\mathbb{Z}[i]$, and so $\mathbb{Z}[i] /\langle q\rangle$ is an integral domain. So $\mathbb{Z}[i] /\langle q\rangle$ is a field. \end{proof}
Dummit-Foote|exercise_9_1_10
theorem exercise_9_1_10 {f : β„• β†’ mv_polynomial β„• β„€} (hf : f = Ξ» i, X i * X (i+1)): infinite (minimal_primes (mv_polynomial β„• β„€ β§Έ ideal.span (range f))) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that the ring $\mathbb{Z}\left[x_{1}, x_{2}, x_{3}, \ldots\right] /\left(x_{1} x_{2}, x_{3} x_{4}, x_{5} x_{6}, \ldots\right)$ contains infinitely many minimal prime ideals.
\begin{proof} Let $R=\mathbb{Z}\left[x_1, x_2, \ldots, x_n\right]$ and consider the ideal $K=\left(x_{2 k+1} x_{2 k+2} \mid k \in \mathbb{Z}_{+}\right)$in $R$. Consider the family of subsets $X=\left\{\left\{x_{2 k+1}, x_{2 k+2}\right\} \mid k \in \mathbb{Z}_{+}\right\}$, and $Y$ the set of choice function on $X$, ie the set of functions $\lambda: \mathbb{Z}_{+} \rightarrow \cup_{\mathbb{Z}_{+}}\left\{x_{2 k+1}, x_{2 k+2}\right\}$ with $\lambda(a) \in$ $\left\{x_{2 a+1}, x_{2 a+2}\right\}$ For each $\lambda \in Y$ we have the ideal $I_\lambda=(\lambda(0), \lambda(1), \ldots)$. All these ideals are distinct, ie for $\lambda \neq \lambda^{\prime}$ we have $I_\lambda \neq I_{\lambda^{\prime}}$. We also have that by construction $K \subset I_\lambda$ for all $\lambda \in Y$. By the Third Isomorphism Treorem $$ (R / K) /\left(I_\lambda / K\right) \cong R / I_\lambda $$ Note also that $R / I_\lambda$ is isomorphic to the polynomial ring over $R$ with indeterminates the $x_i$ not in the image of $\lambda$, and since there is a countably infinite number of them we can conclude $R / I_\lambda \cong R$, an integral domain. Therefore $I_\lambda / K$ is a prime ideal of $R / K$ We prove now that $I_\lambda / K$ is a minimal prime ideal. Let $J / K \subseteq I_\lambda / K$ be a prime ideal. For each pair $\left(x_{2 k+1}, x_{2 k+2}\right)$ we have that $x_{2 k+1} x_{2 k+2} \in K$ so $x_{2 k+1} x_{2 k+2} \bmod K \in J / K$ so $J$ must contain one of the elements in $\left\{x_{2 k+1}, x_{2 k+2}\right\}$. But since $J / K \subseteq I_\lambda / K$ it must be $\lambda(k)$ for all $k \in \mathbb{Z}_{+}$. Therefore $J / K=I_\lambda / K$ \end{proof}
Dummit-Foote|exercise_9_4_2a
theorem exercise_9_4_2a : irreducible (X^4 - 4*X^3 + 6 : polynomial β„€) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that $x^4-4x^3+6$ is irreducible in $\mathbb{Z}[x]$.
\begin{proof} $$ x^4-4 x^3+6 $$ The polynomial is irreducible by Eisenstiens Criterion since the prime $2$ doesnt divide the leading coefficient 2 divide coefficients of the low order term $-4,0,0$ but 6 is not divided by the square of 2. \end{proof}
Dummit-Foote|exercise_9_4_2c
theorem exercise_9_4_2c : irreducible (X^4 + 4*X^3 + 6*X^2 + 2*X + 1 : polynomial β„€) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that $x^4+4x^3+6x^2+2x+1$ is irreducible in $\mathbb{Z}[x]$.
\begin{proof} $$ p(x)=x^4+6 x^3+4 x^2+2 x+1 $$ We calculate $p(x-1)$ $$ \begin{aligned} (x-1)^4 & =x^4-4 x^3+6 x^2-4 x+1 \\ 6(x-1)^3 & =6 x^3-18 x^2+18 x-6 \\ 4(x-1)^2 & =4 x^2-8 x+4 \\ 2(x-1) & =2 x-2 \\ 1 & =1 \end{aligned} $$ $$ \begin{aligned} & p(x-1)=(x-1)^4+6(x-1)^3+4(x-1)^2+2(x-1)+1=x^4+2 x^3-8 x^2+ \\ & 8 x-2 \\ & q(x)=x^4+2 x^3-8 x^2+8 x-2 \end{aligned} $$ $q(x)$ is irreducible by Eisenstiens Criterion since the prime $\$ 2 \$$ divides the lower coefficient but $\$ 2^{\wedge} 2 \$$ doesnt divide constant $-2$. Any factorization of $p(x)$ would provide a factor of $p(x)(x-1)$ Since: $$ \begin{aligned} & p(x)=a(x) b(x) \\ & q(x)=p(x)(x-1)=a(x-1) b(x-1) \end{aligned} $$ We get a contradiction with the irreducibility of $p(x-1)$, so $p(x)$ is irreducible in $Z[x]$ \end{proof}
Dummit-Foote|exercise_9_4_9
theorem exercise_9_4_9 : irreducible (X^2 - C sqrtd : polynomial (zsqrtd 2)) :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that the polynomial $x^{2}-\sqrt{2}$ is irreducible over $\mathbb{Z}[\sqrt{2}]$. You may assume that $\mathbb{Z}[\sqrt{2}]$ is a U.F.D.
\begin{proof} $Z[\sqrt{2}]$ is an Euclidean domain, and so a unique factorization domain. We have to prove $p(x)=x^2-\sqrt{2}$ irreducible. Suppose to the contrary. if $p(x)$ is reducible then it must have root. Let $a+b \sqrt{2}$ be a root of $x^2-\sqrt{2}$. Now we have $$ a^2+2 b^2+2 a b \sqrt{2}=\sqrt{2} $$ By comparing the coefficients we get $2 a b=1$ for some pair of integers $a$ and $b$, a contradiction. So $p(x)$ is irredicible over $Z[\sqrt{2}]$. \end{proof}
Dummit-Foote|exercise_11_1_13
theorem exercise_11_1_13 {ΞΉ : Type*} [fintype ΞΉ] : (ΞΉ β†’ ℝ) ≃ₗ[β„š] ℝ :=
import .common open set function nat int fintype real polynomial mv_polynomial open subgroup ideal submodule zsqrtd gaussian_int char_p mul_aut matrix open_locale pointwise open_locale big_operators noncomputable theory
Prove that as vector spaces over $\mathbb{Q}, \mathbb{R}^n \cong \mathbb{R}$, for all $n \in \mathbb{Z}^{+}$.
\begin{proof} Since $B$ is a basis of $V$, every element of $V$ can be written uniquely as a finite linear combination of elements of $B$. Let $X$ be the set of all such finite linear combinations. Then $X$ has the same cardinality as $V$, since the map from $X$ to $V$ that takes each linear combination to the corresponding element of $V$ is a bijection. We will show that $X$ has the same cardinality as $B$. Since $B$ is countable and $X$ is a union of countable sets, it suffices to show that each set $X_n$, consisting of all finite linear combinations of $n$ elements of $B$, is countable. Let $P_n(X)$ be the set of all subsets of $X$ with cardinality $n$. Then we have $X_n \subseteq P_n(B)$. Since $B$ is countable, we have $\mathrm{card}(P_n(B)) \leq \mathrm{card}(B^n) = \mathrm{card}(B)$, where $B^n$ is the Cartesian product of $n$ copies of $B$. Thus, we have $\mathrm{card}(X_n) \leq \mathrm{card}(P_n(B)) \leq \mathrm{card}(B)$, so $X_n$ is countable. It follows that $X$ is countable, and hence has the same cardinality as $B$. Therefore, we have shown that the cardinality of $V$ is equal to the cardinality of $B$. Since $F$ is countable, it follows that the cardinality of $V$ is countable as well. Now let $Q$ be a countable field, and let $R$ be a vector space over $Q$. Let $n$ be a positive integer. Then any basis of $R^n$ over $Q$ has the same cardinality as $R^n$, which is countable. Since $R$ is a direct sum of $n$ copies of $R^n$, it follows that any basis of $R$ over $Q$ has the same cardinality as $R$. Hence, the cardinality of $R$ is countable. Finally, since $R$ is a countable vector space and $Q$ is a countable field, it follows that $R$ and $Q^{\oplus \mathrm{card}(R)}$ are isomorphic as additive abelian groups. Therefore, we have $R \cong_Q Q^{\oplus \mathrm{card}(R)}$, and in particular $R \cong_Q R^n$ for any positive integer $n$. \end{proof}