Skip to main content

Section C.23 June 2010

Subsection Section I: Groups

Activity C.193. Problem 1.

Let \(G\) be a group and \(H\) a subgroup of \(G\text{.}\) Recall that the centralizer of \(H\) in \(G\) is
\begin{equation*} C_H(G)=\{g\in G|gh=hg\text{ for all }h\in H\} \end{equation*}
Prove that if \(H\) is normal in \(G\text{,}\) then so is \(C_G(H)\) and that \(G/C_G(H)\) is isomorphic to a subgroup of the automorphism group of \(H\text{.}\)
Solution.
Let \(G\) be a group and \(H\nsg G\text{.}\) Let \(x\in C_H(G)\text{,}\) and consider \(gxg\inv\text{.}\) Let \(h\in H\text{.}\) As \(H\nsg G\) we have \(h=gh'g\inv\) for some \(h\in H\text{,}\) and thus \(g\inv h=h'g\inv\) and \(hg=gh'\text{.}\) Consider \(gxg\inv h\text{.}\) Since \(g\inv h=h'g\inv\text{,}\) we see \(gxg\inv h=gxh'g\inv\text{.}\) As \(x\) commutes with everything in \(H\) we have \(gh'xg\inv\text{,}\) and since \(hg=gh'\) we have \(hgxg\inv\text{.}\) Thus \(C_H(G)\nsg G\text{.}\)
Let \(G\) act on the left cosets of \(C_G(H)\) by left multiplication, giving rise to the permutation representation homomorphism \(\rho:G/C_G(H)\to\Aut(H)\text{.}\) By the First Isomorphism Theorem we see that \(G/C_G(H)\) is isomorphic to a subgroup of the automorphism group of \(H\text{.}\)

Activity C.194. Problem 2 (*).

  1. Suppose \(H\) is a subgroup of a group \(G\) and \([G : H] = 7\text{.}\) Prove \(G\) contains a normal subgroup \(N\) such that \(N ⊂ H\) and \([G : N ] \leq 7!\)
  2. Prove \(7!\) is the best possible bound for the previous part — i.e., prove there is a group \(G\) and a subgroup \(H\) with \([G : H] = 7\) such that for every normal subgroup \(N\) of \(G\) with \(N ⊂ H\text{,}\) we have \([G : N ] ≥ 7!\text{.}\)
Solution.
  1. Let \(G\) act on \(H\) by multiplication on left cosets. Thus \(\rho: G\to\Aut(H)\cong S_7\text{.}\) The kernel of this map is the set of all elements in \(G\) such that \(\rho(g)=\phi_g\) is the identity, meaning that \(\phi_g(xH)=gxH=xH\) for all \(xH\in G/H\text{.}\) Then \(x\inv gxH=H\) and \(x\inv gx\in H\text{.}\)
  2. Coming soon!

Activity C.195. Problem 3.

Suppose \(G\) is a simple group of order \(168 = 2^3 · 3 · 7.\) (Yes, there is such a group.)
  1. How many elements of order \(7\) does \(G\) have?
  2. Show that \(G\) has at least \(14\) elements of order \(3\text{.}\)
Solution.
  1. By Sylow’s Theorems, \(|\Syl_7(G)|\equiv 1\mod{7}\) and divides \(2^3\cdot 3=24\text{.}\) Thus the only options are \(1\) and \(8\text{.}\) However, as \(G\) is simple there cannot be only one Sylow \(7\)-Subgroup, as it would be normal. Thus there are \(8\text{,}\) each having \(7\) unique elements and the identity. Thus there are \(49\) elements of order \(7\text{.}\)
  2. By Sylow’s Theorems, \(|\Syl_3(G)|\equiv 1\mod{3}\) and divides \(2^3\cdot 7=56\text{.}\) As \(G\) is simple there cannot be one, so there must be at least \(7\text{,}\) each with \(2\) non-identity elements. Thus there must be at least \(14\) elements of order \(3\text{.}\)

Subsection Section II: Rings and Fields

Activity C.196. Problem 4.

Let \(A\) be the matrix with entries in \(\C\)
\begin{equation*} A=\begin{bmatrix} 3 & 0 & 0 & 0\\ 1 & 3 & 0 & 0\\ 1 & 0 & 3 & 0\\ 0 & 2 & 1 & 3\\ \end{bmatrix} \end{equation*}
  1. Find the Jordan canonical form of \(A\text{.}\)
  2. Is \(A\) similar to
    \begin{equation*} \begin{bmatrix} 0 & -9 & 0 & 0\\ 1 & 6 & 0 & 0\\ 0 & 0 & 0 & -9\\ 0 & 0 & 1& 6\\ \end{bmatrix}? \end{equation*}
Solution.
  1. First, notice that \(A\) is upwards triangular, and thus \(\cp_A(x)=(x-3)^4\text{.}\) By the [cross-reference to target(s) "thm-cayley-hamilton-thm" missing or not unique] the minimum polynomial divides the characteristic polynomial, and from the definition of minimum polynomial we know \(\mp_A(x)\) is the smallest polynomial such that \(\mp_A(A) = 0\text{.}\) Since \(\mp_A(x)\) must be a power of \((x-3)^n\) with \(n\leq 4\text{,}\) we plug in values of \(n\) until we get \(0\text{.}\)
    \begin{equation*} (A-3I)^1=\begin{bmatrix} 0 & 0 & 0 & 0\\ 1 & 0 & 0 & 0\\ 1 & 0 & 0 & 0\\ 0 & 2 & 1 & 0\\ \end{bmatrix} \end{equation*}
    Shucks. Moving on,
    \begin{equation*} (A-3I)^2=\begin{bmatrix} 0 & 0 & 0 & 0\\ 1 & 0 & 0 & 0\\ 1 & 0 & 0 & 0\\ 0 & 2 & 1 & 0\\ \end{bmatrix} \cdot \begin{bmatrix} 0 & 0 & 0 & 0\\ 1 & 0 & 0 & 0\\ 1 & 0 & 0 & 0\\ 0 & 2 & 1 & 0\\ \end{bmatrix}=\begin{bmatrix} 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 0\\ 3 & 0 & 0 & 0\\ \end{bmatrix}, \end{equation*}
    which is also not \(0\text{.}\) However, multiplying one more time we see \(\mp_A(x)=(x-3)^3\text{.}\) By part (2) of the [cross-reference to target(s) "thm-cayley-hamilton-thm" missing or not unique] \((x-3)^3\) is an invariant factor. By part (1) of that same theorem, invariant factors must multiply to \(\cp_A(x)\text{,}\) and so the invariant factors are \(x-3\) and \((x-3)^3\text{.}\) These are also the elementary divisors. So
    \begin{equation*} J_1(3)\oplus J_3(3)=\begin{bmatrix} 3 & 0 & 0 & 0\\ 0 & 3 & 0 & 0\\ 0 & 1 & 3 & 0\\ 0 & 0 & 1 & 3\\ \end{bmatrix}. \end{equation*}
  2. Let
    \begin{equation*} B= \begin{bmatrix} 0 & -9 & 0 & 0\\ 1 & 6 & 0 & 0\\ 0 & 0 & 0 & -9\\ 0 & 0 & 1& 6\\ \end{bmatrix} \end{equation*}
    And notice that \((B-3I)^2=0\text{.}\) Thus \((x-3)^3\) cannot be the minimal polynomial of \(B\text{.}\) Two matrices are only similar if they share the same invariant factors (and thus the same minimum polynomial), so \(A\) and \(B\) are not similar.

Activity C.197. Problem 5.

Let \(R\) be a (not-necessarily commutative) ring let \(M\) be a left \(R\)-module. The annihilator of \(M\) in \(R\) is defined to be
\begin{equation*} \ann_R(M) = \{ r \in R \mid rm = 0 \text{ for all $m \in M$}\}. \end{equation*}
  1. Prove that \(\ann_R(M)\) is a \(2\)-sided ideal of \(R\text{.}\)
  2. Suppose \(M\) is an abelian group (i.e., a \(\Z\)-module) such that \(|M| = 400\) and \(\ann_\Z(M)\) is the ideal generated by \(20\text{.}\) How many possibilities, up to isomorphism, are there for \(M\text{?}\)
Solution.
  1. Let \(a,b\in \ann_R(M)\text{.}\) Consider \((a+b)m=am+bm=0\text{.}\) Thus \(a+b\in \ann_R(M)\text{.}\) Let \(r\in\ann_R(M)\) and consider \(-r\text{.}\) Let \(m\in M\) and suppose \(-rm=n\) for some \(n\in R\text{.}\) Add \(rm\) to both sides to see that \(0=n+rm=n\text{.}\) Thus \(n=0\) and \(-rm=0\) for all \(m\in M\text{.}\) So \(-r\in\ann_R(M)\text{.}\)
    Note that as \(0m=0=m0\) for all \(m\in M\text{,}\) we know that \(0\in\ann_R(M)\text{.}\) As \(rm=0\) for all \(m\in M\text{,}\) we see that \(\ann_R(M)\) is a left sided ideal.
    Suppose \(mr=n\) for some \(m\in R\text{.}\) This time we add \(-rm\) to both sides, but as \(-r\in\ann_R(M)\text{,}\) we once again find that \(mr=0\text{.}\) Notice that this means \(rm=mr\) for all \(r\in \ann_R(M)\) and \(m\in M\text{,}\) and thus that elements of \(\ann_R(M)\) commute with elements of \(M\text{.}\)
    Let \(a,b\in \ann_R(M)\text{.}\) Consider \(m(a+b)=ma+mb\text{.}\) Luckily, we know \(ma=am=0=bm=mb,\) and thus that \(ma+mb=0+0=0\text{.}\) Hence \(a+b\in \ann_R(M)\text{,}\) making \(\ann_R(M)\) is a two sided ideal.
  2. By the [cross-reference to target(s) "thm-ftfgab" missing or not unique] and Sunzi’s Remainder Theorem, there are only so many options we have for \(M\text{:}\)
    • \(\Z_{16}\times \Z_{25}\text{,}\)
    • \(\Z_{8}\times \Z_{2}\times \Z_{25}\text{,}\)
    • \(\Z_{4}\times \Z_{4}\times \Z_{25}\text{,}\)
    • \(\Z_{4}\times \Z_{2}\times \Z_{2}\times \Z_{25}\text{,}\)
    • \(\Z_{2}\times\Z_{2}\times \Z_{2}\times \Z_{2}\times \Z_{25}\text{,}\)
    • \(\Z_{16}\times \Z_{5}\times \Z_{5}\text{,}\)
    • \(\Z_{8}\times \Z_{2}\times \Z_{5}\times \Z_{5}\text{,}\)
    • \(\displaystyle \Z_{4}\times \Z_{5}\times \Z_{20}\)
    • \(\Z_{4}\times \Z_{2}\times \Z_{2}\times \Z_{5}\times \Z_{5}\text{,}\) and
    • \(\Z_{2}\times\Z_{2}\times \Z_{5}\times \Z_{20}\text{.}\)
    However, as ideals are additive subgroups, we know that \(M\) needs to contain a cyclic subgroup of order \(20\text{.}\) Thus we need only consider decompositions with a \(\Z_{20}\) in them, of which there are exactly two:
    1. \(\Z_{4}\times \Z_{5}\times \Z_{20}\) and
    2. \(\displaystyle \Z_{2}\times\Z_{2}\times \Z_{5}\times \Z_{20}\)

Activity C.198. Problem 6.

Let \(F\) be a field and \(V\) a vector space (not necessarily finite-dimensional) over \(F\text{.}\) Prove that every linearly independent subset of \(V\) is contained in a basis for \(V\text{.}\)
Solution.
Let \(I\) be linearly independent in \(V\text{.}\) Add elements to it until it spans \(V\text{,}\) and denote this new set \(S\text{.}\) Let \(\cP\) denote the collection of all subsets \(X\) of \(V\) such that \(I\subseteq X \subseteq S\) and \(X\) is linearly independent. We make \(\cP\) into a poset by the order relation \(\subseteq\text{,}\) set containment.
We note that \(I \in \cP\text{.}\)
Let \(\cT\) be any totally ordered subset of \(\cP\text{.}\) If \(\cT\) is empty, then \(\cT\) is (vacuously) bounded above by \(I\text{.}\) Assume \(\cT\) is non-empty. Let \(Z = \bigcup_{Y \in \cT} Y\text{.}\) I claim \(Z \in \cP\text{.}\)
Given \(z_1, \dots, z_m \in Z\text{,}\) for each \(i\) we have \(z_i \in Y_i\) for some \(Y_i \in\cT\text{.}\) Since \(\cT\) is totally ordered, one of \(Y_1, \dots, Y_m\) contains all the others and hence it contains all the \(z_i\)’s. Since each \(Y_i\) is linearly independent, this shows \(z_1, \dots, z_m\) are linearly independent. We have shown that every finite subset of \(Z\) is linearly independent, and hence \(Z\) is linearly independent. Since \(\cT\) is non-empty, \(Z \supseteq I\text{.}\) Since each member of \(\cT\) is contained in \(S\text{,}\) \(Z \subseteq S\text{.}\) Thus, \(Z \in \cP\text{,}\) and it is clearly an upper bound for \(\cT\text{.}\)
We may thus apply Zorn’s Lemma to conclude that \(\cP\) has at least one maximal element, \(B\text{.}\) I claim \(B\) is a basis of \(V\text{.}\)
Note that \(B\) is linearly independent and \(I \subseteq B \subseteq S\) by construction. We need to show that it spans \(V\text{.}\) Suppose not. Since \(S\) spans \(V\text{,}\) if \(S \subseteq \Span(B)\text{,}\) then \(\Span(B)\) would have to be all of \(V\text{.}\) (For note that if \(\Span(S) = V\) and \(S \subseteq \Span(B)\text{,}\) then for any \(v \in V\) we may write \(v = \sum_i c_i s_i\) for \(s_i \in S\) and \(s_i = \sum_j a_{i,j} b_{i,j}\) with \(b_{i,j} \in B\) and hence \(v= \sum_{i,j} c_i a_{i,j} b_{i,j}\text{,}\) which implies \(\Span(B) = V\text{.}\))
Since we are assuming \(\Span(B) \ne V\text{,}\) there must be at least one \(v \in S\) such that \(v \notin \Span(B)\text{.}\) Set \(X := B \cup \{v\}\text{.}\) Notice, \(I \subset X \subseteq S\) and, by CITEX , \(X\) is linearly independent. This shows that \(X\) is an element of \(\cP\) that is strictly bigger than \(B\text{,}\) contrary to the maximality of \(B\text{.}\) So, \(B\) must span \(V\) and hence it is a basis.
Thus \(I\) is contained in the basis \(B\text{.}\)

Subsection Section III: Linear Algebra and Modules

Activity C.199. Problem 7 (*).

Let \(R\) be an integral domain with field of fractions \(Q\text{.}\) Let \(P\) be a prime ideal of \(R\) and let
\begin{equation*} S = \bigg\{ \frac rd\in Q|d\not\in P\bigg\}. \end{equation*}
  1. Show that \(S\) is a subring of \(Q\text{.}\)
  2. Show that
    \begin{equation*} I = \bigg\{ \frac pd|p\in P, d\not\in P\bigg\} \end{equation*}
    is a prime ideal of \(S\text{.}\)
Solution.
  1. First, notice that since \(P\) is prime, \(R\sm P\) is multiplicatively closed. Let \(x,y\in S\text{.}\) Thus the denominators of \(x\) and \(y\) will still not be in \(P\text{,}\) making \(S\) multiplicatively closed.
  2. Coming soon!

Activity C.200. Problem 8 (*).

Prove \(Z[2i] = \{a + 2bi | a \text{ and } b \text{ are integers}\}\) is not a PID.
Hint.
One method is to use (with proof) the fact that \(2 + 2i\) is irreducible in this ring.
Solution.
Coming soon!

Activity C.201. Problem 9.

Consider \(f(x) = x^6 + 3 \in \Q[x]\text{.}\)
  1. Let \(a\) be a root of \(f(x)\) and prove \(\Q(a)\) is a Galois field extension of \(\Q\text{.}\)
  2. Find the Galois group \(\Gal(\Q(a)/\Q)\text{.}\)
Hint.
First show \(\frac{a^3+1}2\) is primitive \(6\)-th root of unity.
Solution.
Consider \(f(x) = x^6 + 3 \in \Q[x]\text{.}\)
  1. Let \(a\) be a root of \(f(x)\text{.}\) Note that \((\frac{a^3+1}{2})^2=\frac{a^3-1}{2}\text{,}\) and so
    \begin{equation*} (\frac{a^3+1}{2})^6=((\frac{a^3+1}{2})^2)^3=(\frac{a^3-1}{2})^3=1, \end{equation*}
    so yay! It’s primitive. Using one \(6\)th primitive root we can obtain all the others, specifically \(e^{2\pi i/6}\text{.}\) So we multiply each root by this to get all the others. So we have our splitting field.
  2. Since \(\Q(a)\) is Galois, we see that \([\Q(a):\Q]=6\text{.}\) So \(\Gal(\Q(a)/\Q)\) is either \(S_3\) or \(\Z/6.\) The roots of \(f\) are \(\alpha_i=\sqrt[6]{3}e^{{(\pi i/6)}^{2i-1}}\) for \(1\leq i\leq 6\text{.}\)
    By the Porism there exists a \(\sigma\in\Gal(\Q(a)/\Q)\) such that \(\sigma(\alpha_1)=\alpha_2\text{.}\) Let \(\zeta=e^{\pi i/3}\text{,}\) and note that \(\alpha_i=\alpha_{i-1}\zeta\text{.}\) Additionally, note that \(\frac{\alpha_i^3+1}{2}=\zeta\) when \(i=1,3,5\) and \(\frac{\alpha_i^3+1}{2}=\zeta^5\) when \(i=2,4,6\text{.}\)
    Observe then that
    \begin{equation*} \sigma(\alpha_2)=\sigma(\alpha_1)\sigma(\zeta)=\alpha_2\frac{\sigma(\alpha_1)^3+1}{2}=\alpha_2\zeta^5=\alpha_1. \end{equation*}
    Similarly, we see that \(\sigma(\alpha_3)=\alpha_6\) and \(\sigma(\alpha_4)=\alpha_5\text{.}\) Thus \(\sigma\) corresponds to the permutation \((12)(36)(45)\text{.}\)
    Using the Porism again we see there exists a \(\tau\in\Gal(\Q(a)/\Q)\) such that \(\tau(\alpha_2)=\alpha_3\text{.}\) Using a similar process as above we see that \(\tau\) corresponds to \((14)(23)(46)\text{.}\) However, observe that \(\sigma\tau=(153)(264)\text{,}\) while \(\tau\sigma=(135)(246)\text{.}\) Thus these elements do not commute, so we cannot be in \(\Z/6\text{.}\) Thus \(\Gal(\Q(a)/\Q)\cong S_3\text{.}\)