MTH 411
Abstract Algebra II
Fall 2023
I will not collect daily homework, but the next daily quiz will be based on it. Only look at the solution after a serious effort. You may also hand in problems marked Extra Credit, preferably within a week of the HW date.
+1 extra point to the first person pointing out each significant typo or error on this page. Recent corrections and revisions are in red. Tentative future assignments are in gray.
−2 | 0
0
| 1
| |
√2⁄2 | −√2⁄2
√2⁄2
| √2⁄2
| |
1a. The columns of [L] are L(1,0) = (−1,0) and L(0,1) = (0,1), so [L] = [
−1 | 0
0
| 1
| |
1b. Sketching the line of reflection y = −x shows that L(1,0) = (0,−1) and L(0,1) = (−1,0), so [L] = [
0 | −1
−1
| 0
| |
1c. Rotating 90° clockwise shows that L(1,0) = (0,−1) and L(0,1) = (1,0), so [L] = [
0 | 1
−1
| 0
| |
2a. The columns of the matrix are L(1,0) = (−2,0), so L flips the x-axis and stretches it by a factor of 2; and L(0,1) = (0,1), fixing the y-axis. L(x,y) = (−2x, y).
2b. The values of L(1,0) and L(0,1) are 45° counterclockwise rotations of the x- and y-axes, so L rotates all vectors by 45° counterclockwise. L(x,y) = √2⁄2(x−y, x+y).
3 | 4
4
| −3
| |
5 | 2
1
| 4
| |
√2⁄2 | √2⁄2
√2⁄2
| −√2⁄2.
| |
cos 2α | sin 2α |
sin 2α | −cos 2α |
3−λ | 4
4
| −3−λ
| |
To find the eigenvector v1 = (x,y), we solve (L−λ1I)(v) = (L−5I)(x,y) = (0,0), i.e. the linear system
−2x+4y = 0 |
4x−8y = 0. |
The geometric picture of L stretches direction v1 by a factor of 5, and also stretches the perpendicular vector \v2 by 5 as well as flipping it over the origin. This is a dilation (stretch) by a factor of 5, followed by a reflection across the direction of v1
1b. The characteristic polynomial of L is:
5−λ | 2
1
| 4−λ
| |
To find the eigenvector v1 = (x,y), we solve (L−λ1I)(v) = (L−3I)(x,y) = (0,0), i.e. the linear system
2x+2y = 0 |
x + y = 0. |
The geometric picture of L stretches direction v1 by a factor of 3, and also stretches the (non-perpendicular) vector v2 by 6. This is not directly related to any of our familiar mappings, but you can picture it in terms of its stretching effect on the parallelogram grid generated by v1 and v2.
2a. R is a linear mapping because R(v+w) = R(v) + R(w), since the rotation of a parallelogram of vectors is a parallelogram of rotated vectors; and R(cv) = c R(v), since the rotation of a stretched vector is a stretched rotated vector. R rotates the standard basis vectors e1 = (1,0), e2 = (0,1) by 2π⁄3 radians to v1 = (cos 2π⁄3, sin 2π⁄3), v2 = (−sin 2π⁄3, cos 2π⁄3), giving the matrix:
cos 2π⁄3 | −sin 2π⁄3 |
sin 2π⁄3 | cos 2π⁄3 |
−1⁄2 | −√3⁄2 |
√3⁄2 | −1⁄2 |
2b. The matrix product RT·R gives the matrix of dot products of the column vectors of R = [v1 | v2].
v1 |
v2 |
v1·v1 | v1·v2 |
v2·v1 | v2·v2 |
1⁄21⁄2+√3⁄2√3⁄2 | 1⁄2(−√3⁄2)+√3⁄21⁄2 |
−√3⁄21⁄2+1⁄2√3⁄2 | −√3⁄2(−√3⁄2)+1⁄21⁄2 |
1 | 0 |
0 | 1 |
2c. Wolfram gives v2 = R(2,2) = (−1−√3, −1+√3) and v3 = R(v2) = (−1+√3, −1−√3). Note the symmetry across y = x.
3a. It is easy to check RT·R = I, so R is orthogonal.
3b. Wolfram gives eigenvalue λ1 = 1 with eigenvector v1 = (1+√2, 1) and λ2 = −1 with v2 = (1−√2, 1). Thus, R is the orthogonal reflection across the fixed line in direction v1, having slope 1⁄(1+√2) = −1 + √2.
3c. If (cos 2α, sin 2α) = (√2⁄2, √2⁄2), then 2α = π⁄4 = 45° and α = π⁄8 = 22.5°. The line in this direction has slope tan π⁄8 = 1⁄(1+√2) using the tangent half-angle formula. This agrees with part (b).
(i) R = reflection across an axis (use the y-axis for simplicity); |
(ii) R = 180° rotation. |
1. See next week's text, p. 1.
2a. Denote the x,y,z axes by:
Rx = |
| Ry = |
|
L = Rx·Ry = |
|
2c. Use Gaussian elimination (row reduction) to solve (L−I)v = 0 for v = (x,y,z) to find v = (c,c,c) for any c ∈ R, so we may take v1 = (1,1,1). Thus L is a rotation around axis v1
2d. We can check directly by matrix multiplication that L3 = L·L·L = I. Alternatively, since L cyclically permutes the three coordinate axes e1→e2→e3→e1, we have L3 = I. Since perfoming the rotation three times gives the idenity, it must be a 1⁄3 rotation.
Another way to see this is by looking at the complex roots of p(λ), which are the complex cube roots of 1, namely λ2, λ3 = cos(2π⁄3) ± i sin(2π⁄3). These are the complex eigenvalues of a 1⁄3 rotation.
3. See the text p. 5.
0 | −1
1
| 0
| |
R1 = |
| R2 = | 1⁄3 |
|
1a. The symmetries of the square are:
I | Identity mapping, no motion | [
| ||||
R | 90° rotation counterclockwise | [
| ||||
R2 | 180° rotation | [
| ||||
R3 | 270° = −90° rotation | [
| ||||
A | reflection across y-axis | [
| ||||
B | reflection across y = x | [
| ||||
C | reflection across x-axis | [
| ||||
D | reflection across y = −x | [
|
⚬ | I | R | R2 | R3 | A | B | C | D
I | I | R | R2 | R3
| A | B | C | D
| R | R | R2 | R3 | I
| D | A | B | C
| R2 | R2 | R3 | I | R
| C | D | A | B
| R3 | R3 | I | R | R2
| B | C | D | A
| A | A | B | C | D
| I | R | R2 | R3
| B | B | C | D | A
| R3 | I | R | R2
| C | C | D | A | B
| R2 | R3 | I | R
| D | D | A | B | C
| R | R2 | R3 | I
| |
1 | 2 | 3 | 4
2
| 3
| 4
| 1
| |
1a. The symmetries of the square are:
I | Identity mapping, no motion | [
| (
| ||||||||||||
R | 90° rotation counterclockwise | [
| (
| ||||||||||||
R2 | 180° rotation | [
| (
| ||||||||||||
R3 | 270° = −90° rotation | [
| (
| ||||||||||||
A | reflection across y-axis | [
| (
| ||||||||||||
B | reflection across y = x | [
| (
| ||||||||||||
C | reflection across x-axis | [
| (
| ||||||||||||
D | reflection across y = −x | [
| (
|
1c. To compute compositions, we can multiply the corresponding matrices, or compose permutations. For example, doing A after R:
1 | 2 | 3 | 4
2
| 1
| 4
| 3
| |
1 | 2 | 3 | 4
2
| 3
| 4
| 1
| |
1 | 2 | 3 | 4
1
| 4
| 3
| 2
| |
2 & 3. The coordinates of the vertices are given by repeatedly applying the rotation symmetry R = Rot2π/3 to v1:
I | Identity mapping, no motion | [
| (
| ||||||||||
R | 120° rotation Rot2π/3 | [
| (
| ||||||||||
R2 | 240° rotation Rot4π/3 | [
| (
| ||||||||||
A | reflection across x-axis | [
| (
| ||||||||||
B | reflection across y = −√3 x | [
| (
| ||||||||||
C | reflection across y = √3 x | [
| (
|
Could there be other symmetries of T? No: The three sides of the triangle are the only segments of length 1, and any rigid symmetry S must take these longest segments to each other, so it takes their endpoints (the three vertices) to each other. Now, any element of G can be defined by its permutations of the three vertices, so G ⊂ S3; but |S3| = 3! = 6, so there cannot be more elements of G than the 6 above.
1b. No, (Z, ×) is a not a group. It has identity e = 1, but it does not possess inverses inside the group. For example 2−1 = ½ is not in Z.
1c. Yes, (R, +) is a group.
1d. (R, ×) is almost a group, since it has e = 1 as identity, and most elements have inverse a−1 = 1⁄a in the group; but a = 0 does not have an inverse, so the group property fails.
1e. Yes, (R−{0}, ×) is a group, with identity element e = 1, and inverses inverse a−1 = 1⁄a in the group for all a ∈ R−{0}
1f. (Z−{0}, ×) is not a group, since most elements do not have inverses inside the group. (Again, 2−1 = ½ is not in Z−{0}.)
a |
→ |
b |
→ |
r |
→ |
r |
→ |
a |
→ |
r |
→ |
1. Recall ra = ar2, r3 = a2 = e.
2b. The Cayley graph of C6 = ⟨r=c2, a=c3⟩ is the same as that of D3 = ⟨r,a⟩, except the three-cycles of r-arrows have the same orientation (rather than opposite).
2c. In the geometric model of C6 as the symmetries of a triangular prism, r again acts as a 1⁄3 rotation around a horizontal axis, but a acts as a reflection across a vertical plane. These two mappings move orthogonal subspaces because of the Cartesian product structure of C6 ≅ C3 × C2.
( |
| ). |
1a. The obvious isomorphism is defined by T(ci) = i (mod 6), so that T(e) = 0, T(c) = 1, . . . , T(c5) = 5. The group tables of C6 and (Z6, + ) are:
|
|
After writing the first table, I wrote the second via find-and-replace, relabeling each ci as i.
1b. Surprisingly, there is another isomorphism between C6 and Z6. The mapping T(ci) = −i (mod 6) gives:
Furthermore, T respects the group operations because:
Therefore, T satisfies both requirements of an isomorphism.
2a. The Cayley graph can be drawn as regular hexagon with arrows pointing clockwise, each labeled with the generator c.
2b. The multiplication by g = c2 gives the following permutation of C6:
( |
| ). |
We may relabel this as a permutation of Z6 = {0,1,2,3,4,5}:
( |
| ). |
2d. The possible permutations are all bijections P : {1,2,...,6} → {1,2,...,6}. They are constructed by 6 choices for P(1); 5 choices left for P(2), which can be anything but P(1); 4 choices left for P(3), which can be anything but P(1) or P(2); etc. Combining these choices gives 6! = 6 × 5 × 4 × 3 × 2 × 1 = 720 possible permutations in S6.
Only 6 of these are in Sym(X), because a graph symmetry must not only permute the vertices (group elements), but also move arrows to arrows. This restricts symmetries to be rotations only.
1 | 2 | ··· | n
P(1) | P(2) | ··· | P(n)
| |
Thus, since R(−v) = −R(v):
[R2] = |
| . |
1a. The subgroups are: H = D3 itself (since a subset is allowed be all of a given set); containing the cyclic subgroups
H = {e, r, r2}, {e, a}, {e, ar}, {e, ar2};
and the trivial subgroup H = {e}.
It is easy to see that putting any extra element into one of the above subgroups generates one of the others, so the list is complete.
1b. The subgroups are: the whole cyclic group H = C6; the cyclic subgroups H = {e, r2, r4}, H = {e, r3}; and the trivial subgroup H = {e}.
2a. Like all symmetry groups, G = Sym(R+) is an abstract group. It has a closed operation, since composition of two such functions gives the same kind of function; composition is always associative; the function e(x) = x is the group identity element; and the inverse function is the group inverse.
2b. The functions in H = {e, s, r} satisfy ee = e, er = re = r, es = se = s, rs = sr = e, as they should for a subgroup. But there is one thing missing: s2 and r2 are not in H, so the operation is not closed (or not defined within the given set), and it cannot be a (sub)group.
3a. The diagonal of the multiplication table shows e2 = e and a2 = (ar)2 = (ar2)2 = e. These are all the solutions to x2 = e.
3b. The only elements g2 appearing on the diagonal of the multiplication table are e, r, r2, so there is no solution to x2 = a.
3c. Solving x2a = ar2 gives
1. The only abelian groups we have mentioned so far are:
2. Proposition: For any group G, its center Z(G) is a subgroup.
Proof: To show that the center
To check this, suppose c, c' ∈ G, which means for any g ∈ G, we assume cg = gc and c'g = gc'. Then:
3a. Z(Cn) = Cn.
3b. Z(D4) = {e, r2}, where r2 is a ½ rotation. For each other element, it is easy to find another element which does not commute with it.
3c. Z(Tet+) = {I}, the trivial subgroup. Again, all other elements fail to commute with something.
T5(gh) = T5(g) T5(h) for any g = aℓ, h = am.
1a,b. The diagram of subgroups of C24, written in terms of the isomorphic group (Z24, + ), is on Terras p. 78: each ⟨m⟩, with m a divisor of 24, corresponds to H = ⟨am⟩ ⊂ C24, a subgroup with 24⁄m elements.
1c. For example, if we start with
More generally, if we start with a cyclic subgroup in ⟨am⟩ ⊂ C24 for m | 24, and add a generator aℓ, we get:
1d. We can verify that ⟨aℓ⟩ = ⟨ad⟩ for d = gcd(ℓ, 24) by checking each ℓ which are not divisors of 24. For example,
2a,b. The generators are those ⟨ar⟩ with gcd(24, r) = 1, namely r = 1, 5, 7, 11, 13, 17, 19, 23. These 8 generators are consistent with φ(24) = φ(233) = (2−1)23−1 (3−1) = 8.
2c. We check that T5 preserves multiplication:
2e. The inverse mapping T5−1 is given by reversing the arrows in the diagram. But since they are all bi-directional arrows, we get the same diagram, so T5−1 = T5. We can see this since:
1a.
I | R | R2 | R3 | A | B | C | D |
id | (1234) | (13)(24) | (4321) | (14)(23) | (24) | (12)(34) | (13) |
1b. The order of each symmetry is clear geometrically:
2b. The non-identity elements of Tet+
are: Ri±1 = ⅓ rotations of order 3;
and A, B, C = ½ rotations of order 2.
These again agree with the lcm of the cycle lengths.
1a. The elements of S3 are:
id (zero transpositions, even);
transpositions (12), (13), (23), all odd;
and (123) = (12)(23), (132) = (13)(32), both even.
Thus A3 = {id, (123), (132)}.
1b,c. The permutation (123) has matrix
[ ep(1) | ep(2) | ep(3) ]
=
[ e2 | e3 | e1 ]
=
[
].
This can easily be computed to have determinant 1.
Similarly for the other permutations.
0 0 1
1 0 0
0 1 0
2a. An forms a group, since if σ, τ ∈ An, so that they are each products of an even number of permutations, then σtau; is a product of an even + even = even number, so σtau; ∈ An. Also, σ−1 is a product of the same even number of transpositons in the opposite order (since (gh)−1 = h−1g−1), so σ−1 ∈ An. Since An is closed under multiplication and inverses, it is a subgroup.
2b. The odd permutations are the products of an odd number of permutations, but they are not closed under multiplication, since odd + odd = even.
e | r | r2 | a | b | c |
r | r2 | e | c | a | b |
1a. The two-line notation for each permutation σg is given by the row of g in the multiplication table of G = D3. Writing these in cycle notation:
1b. In the representation D3 ⊂ S6, the elements e, a, b, c are odd, while r, r2 are even (since every 3-cycle is even). In fact, these are the same as the signs in our previous representation from the corners of a triangle, D3 ≅ S3 .
1b. The cosets of H = {e, a, c, r2} are H and rH = {r, ra = d, rc = b, r3}. Together G = H ∪ rH, the union of two disjoint subsets each of size |H| = 4. This guarantees that |G| = 2 |H|.
2. Since the subgroup H contains e, and different cosets never intersect, no other coset gH can contain e, so it cannot be a subgroup.
3a. The two cosets of H = ⟨r⟩ are H = {e, r, r2, r3} itself and aH = {a, ar, ar2, ar3} = {a, b, c, d}. Note that the second coset could be represented by any of its elements: for example aH = bH since bH = {b, br, br2, br3} = {b, c, d, a}.
In the Cayley graph, the cosets are the cycles formed by the blue r-arrows:
r3H = {r3, r3a} = {r3, b}.
3c. The cosets of H = ⟨b⟩ are H = {e,b}, rH = {r, a}, r2H = {r2, d}, r3H = {r3, c}.
1. The conjugacy classes of G = D4 are:
K(a) = {a,c} (reflections across sides), K(b) = {b,d} (reflections across diagonals).
2. Claim: x ∈ Z(G) if and only if K(x) = {x}.
Proof: First, assume x ∈ Z(G), so that gx = xg for all g. Then the conjugate is (gx)g−1 = (xg)g−1 = x, and the conjugacy class is K(x) = {gxg−1 for all g} = {x}.
Conversely, assume K(x) = {gxg−1 for all g} = {x}. This means gxg−1 = x, so (gxg−1)g = xg, i.e. gx = xg for all g. Thus x ∈ Z(G).
In our case, we have Z(D4) = {e, r2}, since r2a = ar−2 = ar2, and all other non-identity elements clearly fail to commute with some reflection a, b, c, or d. QED
3. The normal subgroups are marked with N:
1a. Cn is the rotation symmetries of a regular n-gon (no reflections), or all geometric symmetries of an n-gon decorated with counterclockwise arrows. Its elements are Cn = {e, r, r2, . . . , rn−1} with the relation rn = e, so that rirj = ri+j mod n.
The subgroups are K = ⟨rk⟩ for k a divisor of n. Since G is commutative, all subgroups are normal, and G/K = Cn / ⟨rk⟩ ≅ Cn/k. The Cayley graph for the single generator r is a circle of n vertices with counterclockwise r-arrows.
1b. (Zn, + ) is isomorphic to Cn, so all structural features are the same, only the notation is different. The elements are Zn = {[0], [1], . . . , [n−1]} where [k] means (k mod n), with the relation [n] = [0]. The subgroup H = ⟨ [k] ⟩ for k a divisor of n has quotient (Zn/k , + ).
1c. (Zn×, · ) = {ℓ with gcd(ℓ, n) = 1} under multiplication. These are the algebraic symmetries of the group (Zn, + ): that is, each ℓ ∈ Zn× corresponds to an isomorphism
1d. Dn is the geometric symmetries of a regular n-gon. Dn = {e, r, r2, . . . , rn−1, a, ar, ar2, . . . , arn−1} with relations rn = a2 = e and ra = ar−1 = arn−1 , so that
The Cayley graph of Dn with generators S = {r, a} consists of a circle of n vertices with counterclockwise r-arrows, another circle with clockwise r-arrows, and bidirectional a-arrows connecting the n pairs of vertices in the two circles.
1e. Sn is the symmetries of any unstructured set of n distinct objects, usually the numbers {1, 2, . . . , n}. A group element is a permutation, a one-to-one and onto mapping σ : {1, 2, . . . , n} → {1, 2, . . . , n}, denoted in 2-line notation with σ(i) written underneath i; or in cycle notation with σ(i) written next to i, except when i closes a cycle beginning with σ(i). The order (number of elements) is |Sn| = n! = n(n−1)··(2)(1). The multiplication σ·τ is computed by composing permutations: (σ·τ)(i) = (σ(τ(i)), so that each output of τ becomes an input of σ. The inverse σ−1 can be computed by switching the rows in 2-line notation, or by reversing the cycles in cycle notation.
Sn has many, many subgroups. In fact, Cayley's Theorem says that any finite group G can be represented by an isomorphic subgroup of permutations in some Sn: G ≅ G' ⊂ Sn. These subgroups are almost never normal; in fact, the only normal subgroup of Sn for n ≥ 5 is the alternating group An below. The Cayley graph of Sn for n ≥ 5 is too complicated to be useful. However, S2 ≅ C2, S3 ≅ D3, and S4 ≅ Tet, the geometric symmetries of a tetrahedron, which has a nice Cayley graph below.
1f. An is the subgroup of even permutations of Sn, namely those permutations σ which can be written as a product of an even number of transpositions τ = (ab), not necessarily disjoint transpositions. For example, any k-cycle can be written: as a product of k−1 transpositions:
An again has many subgroups. Typically, if a geometric symmetry group G = Sym(X) is represented by permutations of n corners, G ≅ G' ⊂ Sn, then its subgroup of even permutations, H' = G' ∩ An, represents the subgroup G+ of rotations only.
1g. Tet is the geometric symmetries (rotations and reflections) of a regular tetrahedron (pyramid with triangular base).
Tet+ is the subgroup of rotation symmetries only, represented by even permutations: Tet+ ≅ A4. It has 12 elements generated by R1 = (243), R2 = (134), where Ri is the right-handed ⅓ rotation around the corner i.
The Cayley graph of Tet with respect to the transposition generators (12), (23), (34) is drawn below, as the 24 corners of the permutahedron solid X obtained by rotating and reflecting a fundamental domain Xo with one corner by the 24 permutations (in one-line notation), so that X = ⋃σ∈S4 σ(Xo).
1h. GLn(R) is the symmetries of the n-dimensional vector space Rn, and its elements are linear mappings L, which preserve vector addition and scalar multiplication:
2a. See HW 9/1 Text.
2b. Proposition: Let a,b,c be elements of a group G. If ab = ac = e, then b = c. If ab = ca = e, then b = c.
Proof: Suppose ab = ac = e. By the definition of a group, there is an element d = a−1 with da = e, so b = (da)b = d(ab) = d(ac) = (da)c = c; that is, b = c.
Now suppose ab = ca = e. Then b = eb = cab = ce = c; that is, b = c.
2c. Proposition: If z ∈ Z(G), then z−1 ∈ Z(G).
Proof: z ∈ Z(G), the center of G, means that gz = zg for all g ∈ G. Then for any g ∈ G:
2d. See HW 9/15.
2e. Proposition: If H ⊂ G is a subgroup with |H| = ½ |G|, then H is a normal subgroup.
First proof. The subgroup H is normal means that gHg−1 = H for any g ∈ G, where gHg−1 = {ghg−1 for all h ∈ H}. Thus, we must show that ghg−1 ∈ H for any g ∈ G and h ∈ H. Clearly for g ∈ H, we must have ghg−1 ∈ H since H is closed under multiplication and inverses.
Now suppose g ∉ H, so that the coset gH is disjoint from eH = H. Since |H| = |gH| = ½ |G|, we have G = H ∪ gH. Which one of these contains ghg−1? If we had ghg−1 ∈ gH, so that ghg−1 = gh' for h' ∈ H, then multiplying both sides by g−1 would give hg−1 = h', and g = (h')−1h ∈ H. But we chose g ∉ H. The contradiction shows we cannot have ghg−1 ∈ gH, so ghg−1 must be in the other coset, ghg−1 ∈ H. This shows H is normal.
Second proof. An equivalent definition of a normal subgroup H is that its left and right cosets are equal: gH = Hg for all g ∈ G. For g ∈ H, we have gH = Hg = H. For g ∉ H, the cosets gH and Hg must both be disjoint from H, so gH and Hg lie inside G−H = {g ∈ G with g ∉ H}. But gH, Hg, and G−H all have the same size |H| = ½|G| elements, so they are all the same set, and gH = Hg. Therefore H is is normal.
2f. See HW 10/6.
2g. Claim: For any group element g ∈ G, the conjugation mapping T : G → G is defined as T(x) = gxg−1. Then Tg is a group isomorphism.
Proof. An isomorphism means a one-to-one and onto mapping which preserves multiplication.
To check that Tg is one-to-one, suppose that Tg(x) = Tg(y). Then
To check that Tg is onto, take any y ∈ G, and note that for x = g−1yg,
To check that Tg(x) preserves multiplication, note that:
a =
(
)
=
(BD).
A
B
C
D
A
D
C
B
1a. This is a homomorphism because
1b. This is a homomorphism because
even and odd numbers add together
in the same way as +1 and −1 multiply together,
as you can prove by examining
each case of even/odd plus even/odd.
Im(sgn) = {±1},
K = Ker(T) = An even permutations,
and G/K = Sn/An
≅ {±1} ≅ C2.
1c. This is a homomorphism because det(AB) = det(A)det(B). Im(det) = R×, meaning det is onto. Ker(det) = SL2(R) = {A ∈ GL2(R) with det(A) = 1}, the special linear group of determinant-one matrices: these denote linear mappings of the plane which move any shape X to a distorted shape A(X) having the same area. G/K ≅ R×.
2a. Im(P) = {8 permutations in HW 9/8}, K = Ker(P) = {e}, G/K = G, so T is an isomorphism between G and Im(T).
2b. Im(Q) = {Q(e) = I, Q(a) = (BD), Q(b) = (AC), Q(r) = (AC)(BD)}
K = Ker(Q) = {e, r2} = ⟨r2⟩,
a normal subgroup: in fact, K = Z(G) the center of G.
G/K = D4/⟨r2⟩ = {K, aK, bK, rK}
≅ K4 symmetries of the rectangle.
1. The cyclic subgroup of g contains all powers gi including negative powers and g0 = e, so that ⟨g⟩ = {e, g, g2, . . . , gn−1}, where the order n is the smallest positive number with gn = e:
2. We have |g| = 1 exactly when g = g1 = e. Any g ≠ e has |g| > 1, and G has 7 such elements, so the maximum |g| is N > 1.
3. Lagrange's Theorem says that for any subgroup H ⊂ G, the number of elements |H| is a divisor of |G|. This is because G is split into distinct cosets gH, each with |H| elements; the number of cosets is denoted [G : H] = |G|⁄|H| .
Now, a cyclic subgroup H = ⟨g⟩ has |H| = |g| elements, so the maximum order N = |g| is a divisor of |G| = 8, namely 1,2,4,8, but we know N > 1.
4. Assume N = 8, which means there is some g with |g| = 8. Then G contains ⟨g⟩ = {e, g, g2, . . . , g7} ≅ C8, but since G only has 8 elements, we have G = ⟨g⟩ ≅ C8.
5. In any group, if g2 = e for all g ∈ G, or equivalently g = g−1, then G is commutative. This is because, for any g1, g2 ∈ G, we have g1g2 = (g1g2)−1 = g2−1g1−1 = g2g1 .
For our G, assume N = 2, which means g2 = e for all g ∈ G. Take any a ∈ G with a ≠ e, and any b ≠ a,e. Since a2 = b2 = e and ba = ab, we have a subgroup ⟨a,b⟩ = {e, a, b, ab} with two cosets: G = ⟨a,b⟩ ∪ c⟨a,b⟩ for any c ∉ ⟨a,b⟩. Then G = {e, a, b, ab, c, ac, bc, abc}, which is clearly isomorphic to:
Specifically, we have the isomorphism T : C2×C2×C2 → G with T(ti, tj, tk) = aibjck. This clearly respects multiplication. To check it is one-to-one, it is enough to show Ker(T) = {(e,e,e)}: this holds since T(ti, tj, tk) = aibjck = e implies ck = a−i b−j ∈ ⟨a,b⟩; but we chose c1 = c ∉ ⟨a,b⟩, so ck = c0 = e. Similarly bj = e and ai = e, so (ti, tj, tk) = (e, e, e). Since T is one-to-one, the images of its 8 elements must fill all of G, i.e. G is onto.
6. Lemma: If H ⊂ G is a subgroup with |H| = ½|G|,
then H is a normal subgroup.
Proof: G has half its elements inside H,
and half outside H,
forming the complement G−H.
For any g ∉ H,
the left coset gH ⊂ G−H,
with |gH| = |H| = ½|G| elements,
must fill the entire complement G−H,
and the same is true of the right coset Hg.
Thus gH = G−H = Hg, and H is normal.
7a. For r ∈ G with |r| = 4, the number of cosets of H = ⟨r⟩ = {e, r, r2, r3} is [G : H] = |G|⁄|H| = 8⁄4 = 2.
H is a subgroup with |H| = ½|G|, so H is normal by the previous problem. Choosing any a ∉ H = ⟨r⟩, we have
If we had a2 = r, this would mean a8 = r4 = e and |a| = 8, contradicting N = 4. (We can easily check ai ≠ e for i = 1,2,3,4,5,6,7.) Thus a2 = r is impossible. Similarly, a2 = r3 = r−1, with |r−1| = 4, is also impossible. The remaining possiblilities are a2 = e and a2 = r2.
7c. Since ⟨r⟩ is normal by part (a), we have ⟨r⟩a = a⟨r⟩, so that ra ∈ {a, ar, ar2, ar3}. Clearly ra ≠ a.
If we had ra = ar2, then rra = rar2 = ar2r2 = a, but then canceling a would give r2 = e, which is untrue since |r| = 4. Thus ra = ar2 is impossible. The remaining possiblities are ra = ar or ar3
7d. Suppose ar = ra, so that G is commutative. Define C4 × C2 with elements (si, tj) with i mod 4 and j mod 2.
If we have a2 = e, then we get an isomorphism T : C4 × C2 → G defined by T(si, tj) = riaj.
If we have a2 = r2, then |a| = |r| = 4. We have an isomorphism T : C4 × C2 → G defined by T(si, tj) = ri(ar)j. That is, r corresponds to T−1(r) = (s,e), and a corresponds to T−1(a) = (s,t). Both groups are commutative, and r4 = e corresponds to (s,e)4 = (e,e). Finally, a2 = r2 corresponds to (s,t)2 = (s,e)2. These show that the multiplication tables correspond.
In either case, G ≅ C4 × C2 .
7e. Suppose ra = ar3 and a2 = r4 = e. Then G = {e, r, r2, r3, a, ar, ar2, ar3} is our usual presentation of D4, which means they have the same multiplication table: G ≅ D4.
8a,b. See Wikipedia.
8c. Suppose ra = ar3 and a2 = r2, along with r4 = a2 = e. The quaternion group is Q8 = {±1, ±i, ±j, ±k} with i2 = j2 = k2 = −1, which is a central element with −g = (−1)g = g(−1) for g = i,j,k, and ij = k = −ji, jk = i = −kj, ki = j = −ik. Then we can define an isomorphism T : Q8 → G by T(i) = r, T(j) = a, T(k) = ar3 = ar−1, T(−1) = r2. This respects multiplication because T(i), T(j), T(k), T(−1) in G obey all the defining relations among i, j, k, −1 in Q. For example, note that ra = ar−1 implies r−1a = ar, so that:
s |
→ |
1. The cycle graphs of the groups of order |G| ≤ 8 are:
These are given in coordinates as follows. Let the orgin be at the center of the block, the x & y axes through the middle square cross-section, and the z-axis along the length of the block. Then the 3×3 matrices of the 8 symmetries are the same as the 2×2 matrices of the corresponding symmetries of the square (in the HW on Rectangle Symmetries), with an extra ±1 diagonal entry in the z-column, making the determinant equal to 1.
2c. The Cayley graph encodes the multiplication table of G = Q8. Draw two 4-cycles of r-edges around ⟨r⟩ and its coset a⟨r⟩. Then compute the appropriate 4-cycles of a-edges g → ga. Remember each vertex will have one incoming and one out-going a-arrow.
2. Mq turns out to be ⅓ rotation matrix around the diagonal axis u: it permutes the x,y,z axes in a 3-cycle:
3. One way to label the vertices:
4. The matrices are given here: they are the 4×4 matrices with coefficients a,b,c,d.
1. For A ⇒ B, the following mean the same:
For B ⇒ A, the following mean the same:
2. Versions of Lagrange's Theorem. For any group G of order |G| = n with a subgroup H ⊂ G:
There can be no other subgroups, because none of the above can be enlarged without generating the entire group. Any two rotations generate G, since this is clearly true for R1, R2, and any other rotations are conjugate to these. Same for R1 along with any of A = R1R2−1, B = R1R2R1, or C = R1−1R2, and similarly for any conjugate Ri.
Thus, there is no subgroup with |H| = 6, a divisor of |G| = 12.
4a. If an element of Tet+ ≅ A4 is represented by a permutation P of vertices {1,2,3,4}, then:
4b. Besides the trivial subgroup and the whole group, clearly the only subgroup K that can be formed from joining up conjugacy classes is {I,A,B,C}. All other unions of conjugacy classes generate the whole group.
1a,b. Inverses in Z10: 1•1 = 3•7 = 9•9 = 1. All other a ∈ Z10 have a common divisor d with 10, which makes an inverse impossible. For example a = 5: if we could write 5b ≡ 1 mod 10, this would mean 5b + 10k = 1 with b,k ∈ Z (Bezout's equation); but this would mean 5(b+2k) = 1, meaning 5 is a divisor of 1. The contradiction means there could not have been any such b.
1c,d. The multiplicative group G =(Z10)×:
× | 1 | 3 | 7 | 9 |
1 | 1 | 3 | 7 | 9 |
3 | 3 | 9 | 1 | 7 |
7 | 7 | 1 | 9 | 3 |
9 | 9 | 7 | 3 | 1 |
2a. We perform division-with-remainder repeatedly, starting with 100÷17:
2b,c. The numbers relatively prime to n = 100 = 2252 are:
x · b = m · a · b = m · 1 = m mod n
xb = mαβ = m1mk φ(n) = m1(mφ(n))k = m.
m, x ∈ (Zn)× = (Z55)× = |
|
1b. R = Z13 is commutative. An element k ≠ 0 is a zero-divisor when it has a common factor with 13, but since 13 is prime there is no such k. In fact Z13 is a field, with every non-zero element invertible. Indeed, for any k ≠ 0, use the Euclidean algorithm to solve ks + 13t = 1 for integers s,t; then ks = 1 ∈ R.
1c. R = M2(R) does not have commutative multiplication, so it is not a commutative ring. Also, any singular matrix X with det(X) = 0 is a zero-divisor, for example, XY = 0 for X = [
1 | 0
0 | 0
| |
0 | 0
0 | 1
| |
2. R = {a+b√2 for a,b ∈ Q} is a field. First, (R, + ) is a commutative group: it is closed under addition, and since the addition is the same as that of real numbers, it is associative and commutative. R contains 0 = 0 + 0√2, and −(a+b√2) = (−a)+(−b)√2 ∈ R, so R contains negatives.
Second, R is closed under multiplication:
3a. Proposition: If R is a ring and a ∈ R is a zero-divisor, then a is not invertible.
First Proof: By definition, the zero-divisor a has ab = 0 for some b ≠ 0. Then we have (ca)b = c(ab) = c(0) = 0 ≠ b. Since (ca)b ≠ b, by definition ca ≠ 1 for any c ∈ R, so a has no inverse.
Second Proof: If an element a were both a zero-divisor and invertible, we would have some b, c ≠ 0 with ab = 0 and ac = ca = 1. Then we would have 0 = c0 = cab = 1b = b, so b = 0 which contradicts our assumptions. Therefore a could not be both a zero-divisor and invertible; and if a is a zero-divisor, it is not invertible.
3b. It is not true in general that a non-invertible element is a zero-divisor. A counter-example is the ring R = Z, which contains no zero-divisors since ab ≠ 0 for any integers a,b ≠ 0. However, the only invertible elements are ±1, since otherwise 1⁄a ≠ Z.
4. See here #1.
1b. Corrected. α9 = (α3)3 = (1−α)3 = 1 − 3α + 3α2 − α3 = 1 − 3α + 3α2 − (1 − α) = −2α + 3α2.
2a,b,c. See Wikipedia.
2d. The additive group (F, + ) is isomorphic to the Klein 4-group C2 × C2, not to C4. The multiplicative group F× = F−{0} has 3 elements, whereas Z4× = {1, 3} has only 2 elments; they cannot be isomorphic. Consequently, the ring F (a field) cannot be isomorphic to the ring Z4 (having zero-divisors).
2e. The Quadratic Formula makes no sense over the base ring Z2, since it divides by 2 = 0. However, we can solve each equation by guess-and-check, since there are only 4 elements altogether. In fact:
3a,b. See Wikipedia.
3c. The monic cubic polynomials are x3, x(x + 1)2, x(x2 + x + 1), (x + 1)3, which clearly have solutions x = 0 or 1, and two irreducible polynomials (unfactorable in Z2[x]) x3 + x + 1 and x3 + x2 + 1, which each have three solutions in F, accounting for the 2 + 3 + 3 = 8 total elements.
cosets [a] = a + (n) = {a + kn for k ∈ Z}.
1. Clearly 1, 2, α, 2α are not generators. The remaining elements must be the 4 possible generators: δ = α+1, α+2, 2α+1, 2α+2. For example:
2. For any γ = δj, we have
3. The Frobenius mapping preserves addition and multiplication because:
4. A quadratic polynomial is reducible if and only if it has linear factors:
x9 − x | = | x (x−1) (x−2) | (x2+1) | (x2+x+2) | (x2+2x+2) |
= | x (x−1) (x−2) | (x−α)(x−2α) | (x−(α+1))(x−(2α+1)) | (x−(α+2))(x−(2α+2)). |
a | c
b
| d
| |
a | c
b
| d
| |
0 | −1
1
| 0
| |
1 | 2 | 3 | 4
2
| 1
| 4
| 3
| |
1a. For a matrix A ∈ GL2(F) with F = Z2, there are |F2| − 1 = 3 choices for the first column, times |F2| − |F| = 2 choices for the second column, giving |G| = 6. there are |F2| − 1 = 3 choices for the first column, times |F2| − |F| = 2 choices for the second column, giving |G| = 6.
1b. We classified groups of order 6 in HW 3: G ≅ D3 ≅ S3 or G ≅ C6. Matrix multiplication is not commutative, so G cannot be C6, and we must have G ≅ S3.
1c. The explicit isomorphism is given below, labeling
the non-zero points of F2 as
v1 = (1,0),
v2 = (0,1).
v3 = (1,1).
[
| (
| ||||||||||
[
| (
| ||||||||||
[
| (
| ||||||||||
[
| (
| ||||||||||
[
| (
| ||||||||||
[
| (
|
2a. Proposition: Any group homomorphism T : G → G'
satisfies T(e) = e' and T(g−1) = T(g)−1
for all g ∈ G.
Proof: By definition, the homomorphism satisfies
T(g1g2) =
T(g1) T(g2).
Now,
2b. Proposition: For a group homomorphism T : G → G',
the image H' = Im(T) is a subgroup of G'.
Proof: For H' to be a subgroup of G',
it must be closed under multiplication and inverses.
By definition, elements of H' = Im(T) are of the form
T(g1), T(g2), and
their product stays in H'
because T preserves multiplication:
2d. Proposition: A group homomorphism T : G → G'
is one-to-one if and only if Ker(T) = {e}.
Proof: If T is one-to-one, this means
T(g1) = T(g2)
only when g1 = g2.
Thus g ∈ Ker(T) only when
T(g) = e' = T(e), hence only when g = e.
That is, Ker(T) contains only e.
For the other direction, assume Ker(T) = {e}, and take T(g1) = T(g2): we must show that g1 = g2. Indeed, we have:
3a. As before, for a matrix A ∈ GL2(F) with F = Z3, there are |F2| − 1 = 8 choices for the first column, times |F2| − |F| = 6 choices for the second column, giving |GL2(F)| = 48 and |G| = |SL2(F)| = ½|GL2(F)| = 24.
3b. A permutation σ ∈ Sn has n choices for σ(1), times n−1 choices for σ(2), etc., so |Sn| = n(n−1)···(2)(1) = n!. Here |G| = 24 = 4!, so G has the same number of elements as S4.
3c. Any group element g generates a cyclic subgroup ⟨g⟩ = {e, g, g2, . . . , gk−1} with gk = e, and the size of this subgroup is the order |g| = k. Isomorphic groups must have corresponding isomorphic cyclic subgroups.
Now, the permutations in σ ∈ S4 can have order |σ| = 2 for σ of type (12) or (12)(34); or |σ| = 3 for type (123); or |σ| = 4 for type (1234). Thus the maximum order of any element (the maximum size of a cyclic subgroup) is N = 4.
However, our G has the matrix A =
0 | −1
1
| 1
| |
4a. Recall that an even permutation σ ∈ Sn means σ can be written as a product of an even number of transpositions (ij); an odd-length cycle is an even permutation, like (123) = (12)(23). Otherwise, the permutation is odd, and an even-length cycle is an odd permutation, such as (1234) = (12)(23)(34). The even permutations form the alternating group An of order n!⁄2.
The homomorphism T : SL2(Z3) → S4 takes any linear mapping to an even permutation, and in fact covers all 4!⁄2 = 12 of them: Im(T) = A4 .
4b. Consider A ∈ Ker(T), which means a linear mapping A : F2 → F2, with F = Z3, with A(L) = L for every line L = Fv, meaning A(v) = ±v. Thus the colmumns of the matrix [A] are given by A(1,0) = ±(1,0) and A(0,1) = ±(0,1) which gives 4 matrices; but only A = ±I have det(A) = 1. Thus Ker(T) = {±I}.
4c. The matrices ±I commute with any other matrix: AI = IA = A and A(−I) = (−I)A = −A. In fact, these are the only matrices in G that commute with all A, and Ker(T) = {±I} is the center:
The center is always a normal subgroup, since gZ(G) = Z(G)g for all g.
4d. The Homomorphism Theorem implies that the quotient of G by the kernel of T is isomorphic to the image of T: