MTH 411.001
Abstract Algebra II
Spring 2019
I will not collect homework (except problems marked Hand In), but the next daily quiz will be based on it. You may also hand in problems marked Extra Credit, preferably within a week of the HW date. Each problem you give up on is a lost opportunity to learn: only look at the solution after a serious effort.
I will give 1 extra point to the first person pointing out a 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
| |
3 | 4
4
| −3
| |
5 | 2
1
| 4
| |
1a. The columns of [L] are L(1,0) = (−1,0) and L(0,1) = (1,0), 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 over the 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).
3a. Follow the example in the Notes. WolframAlpha says: L fixes the vector (2,1) and flips over the perpendicular vector (−1,2); thus, the motion is an orthogonal reflection across the line y = ½x.
3b. W|A: The motion stretches the vector (−1,1) by a factor of 3, and the vector (2,1) by a factor of 6.
√2⁄2 | √2⁄2
√2⁄2
| −√2⁄2.
| |
cos 2α | sin 2α
sin 2α
| −cos 2α
| |
1a. R is a linear mapping because R(v+w) = R(v) + R(w), since the rotation of a parallelogram is a parallelogram; and R(cv) = c R(v), since the rotation of a stretched vector is a stretched 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 |
1b. 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 |
1c. Wolfram gives v2 = R(2,2) = (−1−√3, −1+√3) and v3 = R(v2) = (−1+√3, −1−√3). Note the symmetry across y = x.
2a. It is easy to check RT·R = I, so R is orthogonal.
2b. 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.
2c. 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).
3a,b. See Notes p.5.
R1 = |
| R2 = | 1⁄3 |
|
1a. According to the Righthand Rule, if you point your thumb along the x-axis and turn ¼ in the direction of your fingers, this takes e1 to e1; e2 to e3; e3 to −e2. Similarly for the y-axis. Thus, the rotation matrices are:
Rx = |
| Ry = |
|
L = Rx·Ry = |
|
Since L cyclically permutes the three coordinate axes e1→e2→e3→e1, we have L3 = I, so 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.
Note that taking the product in the other order would give the opposite rotation:
2a. R1 is reflection of the z-axis across the xy-plane.
2b. Note that R = R2 satisfies R = RT (that is, R2 is a symmetric matrix), so the equation RT·R = I is equivalent to R·R = I. This will happen for any orthogonal twofold symmetry, though not for a general orthogonal matrix.
2c. Wolfram gives R2 = PDP−1 for:
D = J = |
| , P = S = |
| . |
2d. We have:
L = R1·R2 = 1⁄3 |
| , |
The axis must be the intersection of the fixed planes of the two reflections, namely the direction orthogonal to both flipped directions p1 = (0,0,1) and p2 = (1,1,1). This axis can be obtained by solving v·p1 = 0 and v·p2 = 0, i.e. the system z = 0, x + y + z = 0; or equivalently take the 3D vector cross product p1 × p2 = (−1,1,0). Thus the axis is the line R(−1,1,0).
We can find the angle just as for plane reflections: it is twice the angle between the fixed planes, or between the flipped lines p1, p2. This latter angle is θ = arccos (p1·p2)⁄|p1||p2| = arccos 1⁄√3 ≈ 54.74°. Thus the angle of rotation is 2 arccos 1⁄√3 ≈ 109.5°.
0 | −1
1
| 0
| |
1 | 2 | 3 | 4
2
| 3
| 4
| 1
| |
1a,b. 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 x-axis | [
| (
| ||||||||||||
C | reflection across y = x | [
| (
| ||||||||||||
D | reflection across y = −x | [
| (
|
1d. The group table of G:
⚬ | I | R | R2 | R3 | A | B | C | D
I | I | R | R2 | R3 | A | B | C | D
| R | R | R2 | R3 | I | D | C | A | B
| R2 | R2 | R3 | I | R | B | A | D | C
| R3 | R3 | I | R | R2 | C | D | B | A
| A | A | C | B | D | I | R2 | R | R3
| B | B | D | A | C | R2 | I | R3 | R
| C | C | B | D | A | R3 | R | I | R2
| D | D | A | C | B | R | R3 | R2 | I
| |
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. 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: 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.
1 | 2 | 3 | 4 | 5 | 6
1
| 3
| 5
| 2
| 4
| 6
| |
1 | 2 | 3 | 4 | 5 | 6
3
| 2
| 6
| 1
| 5
| 4
| |
1 | 2 | 3 | 4 | 5 | 6
2
| 6
| 3
| 4
| 1
| 5
| |
1a. A cube symmetry T can rotate face 1 to any of the 6 faces, so there are 6 choices for T(1). Then T(2) can be any of the 4 faces adjacent to T(1). Now T(3) must be one of the two faces adjacent to T(1) and T(2), but in fact it must be the one that makes the path T(1) → T(2) → T(3) → T(1) a clockwise circle around their common corner (since the path 1 → 2 → 3 → 1 circles clockwise around their common corner). Finally, T(4) is the face opposite T(3), etc. This makes a total of 6 × 4 = 24 choices, so that |G| ≤ 24. However, this does not prove that all 24 choices lead to symmetries.
1b. The cube has 1 identity symmetry I = (1)(2)...(6). Also 3 axes through opposite pairs of faces, each with 1⁄4, 1⁄2, and 3⁄4 rotations, such as A and A2.
Also 4 axes through opposite pairs of corners, each with 1⁄3 and 2⁄3 rotations, such as:
(
).
Also 6 axes through midpoints of opposite pairs of edges, each with a 1⁄2 rotation, such as:
( 1
2
3
4
5
6
2
3
1
6
4
5
).
Total: 1 + (3)(3) + (4)(2) + (6)(1) = 24, giving all the possible symmetries of G according to part (a).
1
2
3
4
5
6
2
1
4
3
6
5
2a. G = (Z5)× = {1,2,3,4} is a closed under multiplication, since a product of non-zero elements mod 5 (i.e. not divisible by 5) will never be zero mod 5 (i.e. divisible by 5). (This is because 5 is a prime number.) G is associative because integer multiplication is associative: (ab)c = a(bc). G has the identity element 1 (mod 5). We can check that each element of G has a multiplicative inverse (reciprocal): 1−1 = 1, 2−1 = 3, 3−1 = 2, 4−1 = 4, since 2·3 = 6 = 1 (mod 5), etc. (In ring terms, this means Z5 is a field.)
2b. H = {0,1,2,3,4} is clearly closed and associative
under the addition operation +.
That is, we write a+b for a combination rather than a·b.
Also, H has the identity element 0, since operating by 0 has
no effect: 0 + a = a + 0 = a. (This is the additive notation for
the axiom 1·a = a·1 = a.)
Finally, each element a has an inverse which cancels it
to give the identiy element: a + (−a) = (−a) + a = 0.
(This is the additive notation for
the axiom a·a−1 = a−1·a = 1.)
2c. The operation tables of H and C5 are:
+ 0 1 2 3 4
0 0 1 2 3 4
1 1 2 3 4 0
2 2 3 4 0 1
3 3 4 0 1 2
4 4 0 1 2 3
· I R R2 R3 R4
I I R R2 R3 R4
R R R2 R3 R4 I
R2 R2 R3 R4 I R
R3 R3 R4 I R R2
R4 R4 I R R2 R3
3a. We show GLn(R) is an abstract group. The matrix A lies in GLn(R) whenever det(A) ≠ 0, i.e. when A represents an invertible linear mapping. Closed: We know from linear algebra that if det(A), det(B) ≠ 0, then det(A·B) = det(A) det(B) ≠ 0; or alternatively, that a composition of invertible mappings is invertible. Associative: One can check algebraically that (A·B)·C = A·(B·C), but this follows immediately from considering · as composition of mappings, since (A·B)(C(x)) = A(B(C(x)) = A((B·C)(x)). Identity element: the n×n identity matrix I. Inverses: By definition, A ∈ GLn(R) is invertible, and its inverse is in GLn(R).
3b. We show O(n) is an abstract group. The matrix A lies in O(n) whenever ATA = I, i.e. when A represents an orthogonal linear mapping. Closed: If ATA = BTB = I, then:
+ | 0 | 1 | 2 0 | 0 | 1 | 2
| 1 | 1 | 2 | 0
| 2 | 2 | 0 | 1
| |
· | e | r | r2
e | e | r | r2
| r | r | r2 | e
| r2 | r2 | e | r
| |
1 | 2 | 3
i
| j
| k
| |
1. Choosing an invertible matrix with entries in the field F = Zp is the same thing as choosing a basis {v1, v2} of the vector space F2. The first vector v1 can be any non-zero element of F2, meaning p2 − 1 choices. The second vector can be any vector linearly independent from v1, i.e. anything outside the line Fv1 having p elements; this means p2 − p choices. The total number of possible bases is |GL2(Zp)| = (p2 − 1)(p2 − p).
2. Proof: By hypothesis, suppose we have two right-inverses ab = ab' = e and two left-inverses ca = c'a = e. We proved above that we can cancel a in the equation ab = ab', so b = b', and similarly c = c'. Now:
3a. The set H = ⟨a⟩ satisfies the four group axioms. It is closed because for any ai, aj ∈ H, we have aiaj = ai+j ∈ H. It is associative because the group G itself is associative. It contains the identity e = a0. Finally, the inverse of ai is a−i ∈ H.
3b. Since G is finite, the elements { . . . , a−2 , a−1, e, a, a2, . . . } cannot all be different: we must have some coincidence ai = aj for some i < j, so that ai = aiaj−i. Cancelling ai from both sides gives e = aj−i, so we can take k = j−i > 0 with ak = e.
3c. Suppose n = min{k > 0 with ak = e}.
Then the elements {e, a, a2, . . . , an−1}
must all be distict. Otherwise we would have
ai = aj for some 0 ≤ i < j ≤ n−1,
so that aj−i = e for 0 < j−i < n,
which contradicts the choice of n as the smallest with an = e.
Now define an isomorphism φ : ⟨a⟩ → Cn
by φ(ai) = ri. This is one-to-one and onto by the above,
and clearly obeys the structural equation
φ(aiaj) = rirj
= φ(ai) φ(aj).
Thus φ is an isomorphism giving ⟨a⟩ ≅ Cn .
1. The subset G×e ⊂ G×H is a subgroup, since it is clearly closed under multiplication and inverses: for example, (g,e)−1 = (g−1, e) ∈ G×e. Define an isomorphism φ : G → G×e by φ(g) = (g,e). This is clearly invertible (with inverse ψ(g,e) = g), and it obeys the stuctural equation:
2a. In the multiplicative group Z9× = {1,2,4,5,7,8}, we find the cyclic subgroup:
2b. In the multiplicative group Z12× = {1,5,7,11}, we find that every element has a2 = e, so that a has order 2, i.e. |a| = 2. Choosing the two elements 5 and 7, we can define an isomorphism
2c. The cyclic group C6 = {e, r, r2, . . . , r5} has r3 as an element of order 2, and r2 as an element of order 3. Thus, writing C2 = {e,s} and C3 = {e, t, t2} we can define an isomorphism
C2×C3 | (e,e) | (e,t) | (e,t2) | (s,e) | (s,t) | (s,t2) |
C6 | e | r2 | r4 | r3 | r5 | r |
2d. The groups C2 × C2 has all
elements of order 2, i.e. g2 = e for all g,
but C4 has elements of order 2 and 4.
But any isomorphism preserves the order of each element,
so there can be no such isomorphism.
More formally, if there were an isomorphism
φ : C4 → C2 × C2,
then the structural equation would force
φ(r2) = φ(r)2 = e = φ(e),
even though r2 ≠ e ∈ C4;
thus φ could not be one-to-one, a contradiction.
3a. As in the previous HW, the dihedral group permutes the three vertices of the triangle, making D3 ≅ S3, with the isomorphism:
1 | 2 | 3
2
| 3
| 1
| |
1 | 2 | 3
1
| 3
| 2
| |
1 | 2 | 3
3
| 2
| 1
| |
1 | 2 | 3
2
| 1
| 3
| |
· | e | r | r2 | a | ar | ar2 |
e | e | r | r2 | a | ar | ar2 |
r | r | r2 | e | ar2 | a | ar |
r2 | r2 | e | r | ar | ar2 | a |
a | a | ar | ar2 | e | r | r2 |
ar | ar | ar2 | a | r2 | e | r |
ar2 | ar2 | a | ar | r | r2 | e |
3b. The cyclic subgroups are:
3d. If we had C2 × C3 ≅ D3,
and by #2c we know C2 × C3 ≅ C6,
this would mean D3 ≅ C6, contradicting #3c;
thus we cannot have C2 × C3 ≅ D3.
Still what if we take a,r ∈ D3 of order 2 and 3,
and try defining φ : C2 × C3 → D3 by φ(si, tj) = airj . This is an invertible mapping,
and the structural equation holds in many cases, for example
φ((s,e)(e,t)) = φ(s,t) = ar = φ(s,e) φ(e,t).
But it must fail somewhere!
Namely, φ((e,t)(s,e)) = φ(s,t) = ar,
but φ(e,t) φ(s,e) = ra : this reflects the non-commutativity of D3.
· | e | a | b | ab |
e | e | a | b | ab |
a | a | e | ab | b |
b | b | ab | e | a |
ab | ab | b | a | e |
1a. Proposition: For any group G, the center Z(G) is a subgroup.
Proof: We know that a subset of G is a subgroup
if it is closed under multiplication and inverses
(the other two group axioms follow automatically).
Thus, let c,d ∈ Z(G), meaning
cg = gc and dg = gd for all g ∈ G.
Then:
1b. Considering D4 as symmetries of a square in the plane, we can check that the only matrix which commutes with all the others is the 180° rotation R2 = −I. Thus, Z(D4) = {I, R2}.
2. Proposition: If G is a group with |G| = 5, then G ≅ C5.
Proof: Again let N = max{|a| for a ∈ G}, the size of the largest cyclic subgroup of G.
3. We have seen the following groups with 6 elements in the previous HW:
4a. If ab = e, then
b(ab)b−1 = bb−1 = e,
which simplifies to ba = e.
This also follows from the uniqueness of inverses proved
previously.
5. Prop: If every element of a group G has order 2,
then G must be abelian.
Proof: For a,b ∈ G, suppose by hypothesis
that a2 = b2
= (ab)2 = e, i.e. a = a−1,
b = b−1, and abab = e.
We manipulate abab = e so as to leave only ba on the left side:
6. Let H be any subgroup of (Z, +).
If we do not have H = {0} = ⟨0⟩,
then there is some i ∈ H with i > 0.
(If i < 0, we have the inverse −i ∈ H.)
Let n = min{i ∈ H with i > 0}, and take any i ∈ H.
Then it is clear that ⟨n⟩ ⊂ H.
Applying the division algorithm to i divided by n, we
have i = nk + r for 0 ≤ r < n.
But then r = nk − i = ±(n + ··· + n) − i ∈ H. By the choice of n, we must have r ≥ n,
which is false, or r ≤ 0, which leaves only the possibility r = 0.
Thus i = nk, and i ∈ ⟨n⟩. Therefore
H ⊂ ⟨n⟩.
1. Coset Lemma (ii): For any a ∈ G, we have |H| = |aH|.
Proof: The mapping f : H → aH, h ↦ ah has inverse function
ah ↦ h = a−1 ah. Since it is invertible,
it must be one-to-one and onto, so its domain and codomain have the
same size: |H| = |aH|.
Coset Lemma (iii): G is a disjoint union of cosets,
G = H ∪ a1 ∪ a2 ∪ ··· .
Proof: By Lemma (i), each g ∈ G lies in at most one coset;
but any element is in its own coset, g = ge ∈ gH,
so every element lies in exactly one coset; and the cosets
disjointly cover G.
2a. As for all groups, G = C15 has the trivial subgroup H = {e}, whose cosets are the individual elements of G; and H = G, which has only one coset containing all elements. We have the following non-trivial decompositions of C15 into cosets:
2b. Since (Z15 , +) ≅ C15, this is the same problem as 2a, just in a different notation.
3. If |G| = 5, by Lagrange's Theorem any subgroup H must have 5 = |H| [G : H], so we can only have |H| = 1 or 5 since these are the only factors of the prime number 5. Now, taking H = ⟨a⟩, we find the order |a| = |H| = 1 or 5; hence there must be an element of order 5, and G = ⟨a⟩ ≅ C5. That is, the only group of order 5 is a cyclic group.
This argument goes through the same way for |G| = p, any prime number. We must have G ≅ Cp.
[1] + [2] | = | [3] = [0] |
[−5] + [11] | = | [6] = [0]. |
[r][r2] = [r3] = [e] |
[ar2][ar] = [ar2ar] = [r2]. |
1. Proposition: For a subgroup H ⊂ G, we have HH = H.
Proof: First, if hh' ∈ HH, then hh' ∈ H by the
closure axiom for the subgroup H; thus HH ⊂ H. Second, if h ∈ H, then h = he ∈ HH by the identity axiom for H; thus H ⊂ HH.
Therefore, we have HH = H.
2. The quotient G/N, whose elements are cosets [g] = gN,
satisfies the four group axioms because the group G
satisfies these axioms.
Closure: The multiplication [a][b] = [ab] is well-defined (proved in notes).
Associative: ([a][b])[c] = [(ab)c] = [a(bc)] = [a]([b][c]).
Identity: [a][e] = [ae] = [a].
Inverses: [a][a−1] = [aa−1] = [e].
3a. Proposition: For any subgroup H ⊂ G, the number of left cosets
is equal to the number of right cosets.
Proof: Define the inverse mapping
φ : G → G by φ(g) = g−1,
and denote φ(H) = {φ(h) for h ∈ H}.
For h ∈ H, we have φ(h) = h−1 ∈ H,
so φ(H) ⊂ H. Also, h = (h−1)−1 = φ(h−1),
so H ⊂ φ(H); and therefore φ(H) = H.
Applying φ to any left coset gH = {gh for h ∈ H}
gives:
3b. Proposition: For a subgroup H ⊂ G, if [G : H] = 2,
then H is a normal subgroup.
Proof: Take any g ∈ G. If g ∈ H, then gH = H = Hg.
On the other hand, if g ∉ H, the hypothesis [G : H] = 2 means there are just two left cosets,
and by part (a) there are also just two right cosets:
4. The normal subgroups of G = D4 = {e, r, r2, r3, a, ar, ar2, ar3} with a2 = r4 = e and ra = ar−1, are the trivial subgroups {e} and G, and:
Here the operation on inputs G = Z is + , while that on
the outputs R× is multiplication.
Kernel: trivial, since φ(i) = 2i = 1 only for i = 0; Ker(φ) = {0}.
Image: the set of integer powers of 2, which is the cyclic subgroup generated by 2
Note that x,y are real numbers with their usual commutative multiplication.
Kernel: the solutions of φ(x) = x2 = 1; Ker(φ) = {1, −1} = ⟨−1⟩.
Image: the real numbers of the form φ(x) = x2, i.e. the positive numbers;
Im(φ) = R>0 .
1a. You only need to check the structure equation:
1b. The kernel is:
1c. The theorem says R>0 ≅ C×/Cir. In fact, any complex number can be written in polar form as a+bi = r (cos(θ) + i sin(θ)), with r = √(a2+b2). Taking this mod Cir, the second factor becomes irrelevant, and [a+bi] = (a+bi)Cir = r Cir, and clearly this is isomorphic to multiplication of positive reals.
2. Proposition: Any homomorphism has φ(e) = e' and φ(g−1) = φ(g)−1.
Proof: First, for any g ∈ G, we have g = ge, so φ(g) = φ(ge) = φ(g)φ(e). Multiplying both sides by φ(g)−1 gives e' = φ(e).
Second, φ(g)φ(g−1)
= φ(gg−1) = φ(e) = e', so φ(g) and φ(g−1) multiply to e', and are inverses.
3. Proposition: Ker(φ) = {e} if and only if φ is injective.
Proof: (⇒) Suppose Ker(φ) = {e}, i.e. φ(k) = e' only for k = e. Suppose φ(g1) = φ(g2); then e' = φ(g1)−1φ(g2)
= φ(g1−1g2), and g1−1g' ∈ Ker(φ), so g1−1g2 = e, so g1 = g2.
Thus two different elements of G cannot have the same output under φ.
(⇐) Suppose φ is one-to-one. Since φ(e) = e',
there can be no other k ∈ G with φ(k) = e',
so Ker(φ) = {e}.
4. Given a group homomorphism φ : G → G'.
Proposition: Ker(φ) is a normal subgroup of G.
Proof: First we check that Ker(φ) is a subgroup.
If k1,k2 ∈ Ker(φ), then φ(k1k2) = φ(k1) φ(k2) = e'e' = e',
so k1k2 ∈ Ker(φ).
Hence Ker(φ) is closed under multiplication.
Second, if k ∈ Ker(φ), then φ(k−1)
= φ(k)−1 = (e')−1 = e',
so k−1 ∈ Ker(φ). Hence Ker(φ) has inverses.
These two axioms imply Ker(φ) is a subgroup.
Next, we show Ker(φ) is normal by checking g Ker(φ) g−1 = Ker(φ).
Let k ∈ Ker(φ) and g ∈ G. Then:
Proposition: Im(φ) is a subgroup of G'.
Proof: Take any two elements of the image,
φ(g1), φ(g2) ∈ Im(φ).
Then φ(g1) φ(g1) = φ(g1g2) ∈ Im(φ),
and Im(φ) is closed under multiplication.
Also, if φ(g) ∈ Im(φ), then φ(g)−1
= φ(g−1) ∈ Im(φ), so
Im(φ) has inverses. Hence Im(φ) is a subgroup.
a |
→ |
b |
→ |
r |
→ |
r |
→ |
a |
→ |
r |
← |
1. Recall ra = ar2, r3 = a3 = 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. The Cayley graph is a product because the group is isomorphic to a product: C6 ≅ C3 × C2, where r = c2 and a = c3 map to the generators of the two factors.
2d. 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 described in (c).
1. Since a2 = b2 = c2 = e,
all edges are bi-directional ↔.
In the picture, the bottom blue edge is
e = 1234 ↔ 2143 = (12) = a,
so a-edges are blue ↔a,
and similarly green b-edges ↔b
and red c-edges ↔c.
To check that the edges are correct, note that that the edge f ↔a fa is written in one-line notation as:
a |
↔ |
so that multiplying on the right by a = (12) switches the first two entries; and similarly for b,c.
2. The vertices correspond to group elements, so |S4| = 4! = 24. The edges of each color pair off the 24 vertices, so there are 12 of each color, representing multiplication by a generator. There are three types of faces: 6 squares corresponding to the relation ac = ca; 4 hexagons corresponding to aba = bab; and 4 hexagons corresponding to bcb = cbc.
3. Let X be a regular tetrahedron, a pyramid with three sides and triangular bottom, all of them equilateral triangles. This has 4 vertices, and
the geometric symmetries permute them in all possible ways,
so Sym(X) = S4.
We can obtain a permutohedron by slicing off the tip of each corner of the tetrahedron, exposing 4 new triangles and chopping the original 4 triangle faces into hexagons;
then shaving off a sliver from each of the original 6 edges, exposing
6 rectangle faces and chopping the 4 new triangle faces into hexagons.
It is clearer in this picture of the permutahedron, with unequal edge-lengths.
1a. The vertices are P1 = (1,0) and P2, P3 = (cos(2π⁄3), ±sin(2π⁄3)). See W|A.
1b. See first week HW.
cos 2π⁄3 | −sin 2π⁄3 |
sin 2π⁄3 | cos 2π⁄3 |
−1⁄2 | −√3⁄2 |
√3⁄2 | −1⁄2 |
a | c |
b | d |
d | −c |
−b | a |
1 | 0 |
0 | −1 |
−1⁄2 | −√3⁄2 |
−√3⁄2 | 1⁄2 |
−1⁄2 | √3⁄2 |
√3⁄2 | 1⁄2 |
2a.Taking each 1⁄4 rotation according to the right-hand rule, with thumb along the axis, we get:
P = R1 =
(
|
|
Q = R2 =
(
|
|
2b. By inputing i = 1, 2, . . . on the right side of the composed permutation R1 ∘ R2, we compute:
2c. An orthogonal transformation R is a linear mapping which preserves distances, angles, and the dot product. Its matrix [R] has column vectors which form an orthonormal basis, or equivalently [R]T R = I. Any orthogonal matrix has determinant ±1; and in 3 dimensions, det[R] = 1 means R is a rotation, and det[R] = −1 means R is a roto-reflection (reflection possibly composed with a rotation whose axis is the reflection direction).
In our case, we are composing rotations, i.e. multiplying orthogonal matrices of determinant 1, so the result is also orthogonal of det 1, and is a rotation.
Looking at how the faces move on the cube, we see that P = (124)(365) is a 1⁄3 rotation around the diagonal axis (1,1,1), which is the common point of faces 1,2,4. Similarly, Q = (12)(34)(56) is a 1⁄2 rotation around the partial diagonal axis (1,0,1), which is the midpoint of the common edge of faces 1,2. These are somewhat surprising symmetries of the cube: try them out on dice or some other physical model.
2d. By multiplying matrices, we get:
P = R1R2 = |
| Q = R12 R2 = |
|
2e. From the geometric description of P, we see that its rotation axis is an eigenvector of eigenvalue 1, and there are no other real eigenvectors. Applying the general method, we find the eigenvalues as roots of the characteristic polynomial p(λ) = det(R−λI) = 1 − λ3. The only real number solution is λ = 1, and we find its eigenvector v = (x,y,z) by solving the linear system given by (P−I)(x,y,z) = (0,0,0). This gives the axis (x,y,z) = (1,1,1), as in (c).
P also has complex eigenvalues and eigenvectors. The non-real roots of p(λ) = 1 − λ3 are the cube roots of unity λ = cos(2π⁄3) ± i sin(2π⁄3). These indicate that P is a rotation by angle 2π⁄3 .
Similarly for Q, the rotation axis (1,0,1) is an eigenvector of eigenvalue 1, and the vectors perpendicular to this are are all eigenvectors of eigenvalue −1 = cos(π) ± i sin(π). Since sin(π) = 0, all the roots of the characteristic polynomial are real: p(λ) = −λ3 − λ2 + λ + 1 = −(λ−1)(λ+1)2.
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. If |g| = 8, then G contains ⟨g⟩ = {e, g, g2, . . . , g7} ≅ C8, but since G only has 8 elements, we have G = ⟨g⟩ ≅ C8.
5. We saw in HW 1/25 #5 that if g2 = e for all g ∈ G, or equivalently g = g−1, then G is abelian. This is because, for any a,b ∈ G, we have ab = (ab)−1 = b−1a−1 = ba.
Suppose this is the case for G, with N = 2. Take any e ≠ a ∈ G, 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:
6a. For r ∈ G with |r| = 4, the number of cosets of H = ⟨r⟩ = {e, r, r2, r3} is [G : ⟨r⟩] = |G|⁄|⟨r⟩| = 8⁄4 = 2.
We saw in HW 2/4 #3b that if H is any subgroup with [G : H] = 2, then H is normal. This is because the complement set G − H is equal to the coset gH for any g ∉ H, and also equal to Hg, so gH = Hg. Thus H = ⟨r⟩ is normal.
Choosing any a ∉ ⟨r⟩, we have G = {e, r, r2, r3, a, ar, ar2, ar3}.
6b. We have the normal subgroup H = ⟨r⟩ with the quotient group G/H = {H, aH} ≅ C2, so a2H = (aH)2 = H. That is, that a2H = H, which means a2 ∈ H = {e, r, r2, r3}.
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.
6c. 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 = ar4 = 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
6d. Define C4 × C2 = ⟨r,s⟩ with r4 = s2 = e and rs = sr. In G, assume ar = ra, so that G is abelian. If we have a2 = e, then we get an obvious isomorphism φ : C4 × C2 → G defined by φ(r) = r and φ(s) = a.
The other possibility is a2 = r2, so that |a| = |r| = 4. Then we can define an isomorphism φ : C4 × C2 → G by φ(r) = r and φ(s) = ar. We check the crucial case of the structure equation:
6e. If G = {e, r, r2, r3, a, ar, ar2, ar3} with r4 = a2 = e and ra = ar3, then this is just our usual presentation of D4, which means they have the same multiplication table: G ≅ D4
6f. We defined the quaternion group as Q = {±1, ±i, ±j, ±k} with i2 = j2 = k2 = 1 and ij = k = −ji, jk = i = −kj, ki = j = −ik. In G, assume that ra = ar3 and a2 = r2, along with r4 = e. Then we can define an isomorphism φ : Q → G by φ(i) = r, φ(j) = a, φ(k) = ar3 = ar−1, φ(−1) = r2. This obeys the structure equation because φ(i), φ(j), φ(k), φ(−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:
1a. The subset gHg−1 is a subgroup, since it is closed under multiplication and inverses:
1b. H is normal ⇔ gH = Hg ∀g∈G ⇔ (gH)g−1 = (Hg)g−1 ∀g∈G ⇔ gHg−1 = H ∀g∈G.
2a. In G = D3, the trivial subgroups {e} and G are normal, equal to all their conjugates. We also have the normal subgroup N = {e,r,r2}, since any subgroup of index [G : N] = 2 is normal. Also, we have the conjugate subgroups:
2b. Because S3 ≅ D3, this is just a different notation for the previous problem. The normal subgroup is N = {(1), (123), (132)}, and the conjugate subgroups are:
3. Prop: For subgroups H,K ⊂ G, we have HK a subgroup if and only if HK = KH.
Pf: (⇒) Suppose HK is a subgroup. Then k = ek and h = he are both elements of the subgroup HK, which is closed under multiplication, so k·h ∈ HK. Thus KH ⊂ HK, and similarly HK ⊂ KH, so HK = KH.
(⇐) Conversely, suppose HK = KH. Then any kh ∈ KH is also in HK, and can be written as kh = h'k', where h,h' ∈ H and k,k' ∈ K. Thus:
4a. Since D3 is not commutative, it is clearly not isomorphic to H × K = C3 × C2. We have H ∩ K = {e} and |G| = |H| |K| = 6. Also H is a normal subgroup, but K is not, as we can see from its conjugate subgroups in #2a.
4b. Assume that H,K are normal subgroups with H ∩ K = {e}. Since K is closed, we have KK = K, and since it is normal, we have gKg−1 = K. Now consider the element:
4c. The mapping φ : H × K → G defined by φ(h,k) = hk satisfies:
4d. To show φ is one-to-one, suppose φ(h,k) = φ(h',k'). Then hk = h'k' and (h')−1h = k'k−1. The left side of the last equation lies in H, while the right side lies in K, so both sides must lie in H ∩ K = {e}. That is, (h')−1h = k'k−1 = e, so h = h' and k = k'. This means φ is one-to-one.
Now, assuming |H| |K| = |G|, this means φ is a one-to-one (injective) mapping between two sets with the same number of elements: thus it must also be onto (surjective). Hence φ is a bijective homomorphism, i.e. an isomorphism.
1. Abelian p-groups: any products of cyclic groups Cpk's, such as C4 × C2.
Non-abelian p-groups: dihedral groups D2k such as D8; quaternion group Q; and any product of an abelian p-group with a non-abelian one. Note: A symmetric group Sn is never a p-group, except for n = 2.
2. Consider G = D3 = ⟨r,a⟩ with r3 = a2 = e and ra = ar−1. The unique Sylow 3-subgroup ⟨r⟩ is normal. We have the three conjugate Sylow 2-subgroups ⟨a⟩, ⟨ar⟩ = r⟨a⟩r−1, ⟨ar2⟩ = r2⟨a⟩r−2, verifying Sylow Theorem II. Also, 3 divides |G| = 6 and 3 ≡ 1 mod 2, verifying Sylow Theorem III.
For G = C12 = ⟨r⟩ with r12 = e, the unique Sylow 3-subgroup is ⟨r4⟩, and the unique Sylow 2-subgroup is ⟨r3⟩. We also have the 2-subgroup ⟨r6⟩ of order 2, as guaranteed by Sylow Theorem I.
Consider G = D6 = ⟨r,a⟩ with r6 = a2 = e and ra = ar−1; also the central element r3 with r3a = ar3. The unique Sylow 3-subgroup ⟨r2⟩ is normal. The Sylow 2-subgroups have order 22 = 4; there are three such subgroups, all conjugate:
3. G = S4 has order |G| = 4! = 24 = 3·8. The Sylow 3-subgroups have order 3, which must be cyclic, generated by the 3-cycles. By Sylow III, the number s of these must divide 24, and s ≡ 1 mod 3, so it must be 1 or 4. In fact, we have:
There are two types of 4-subgroups, the three cyclic ones such as ⟨(1234)⟩ = {(1), (1234), (13)(24), (4321)}, all conjugate to each other; and the normal subgroup {(1), (12)(34), (13)(24), (14)(23)}. I turns out that adding one 2-cycle to a 4-cycle produces a subgroup of order 8:
4. Consider |G| = n ≤ 20. We know G must be abelian for n = 1,2,3,4. For even n = 2m ≥ 6, we always have the non-abelian dihedral group G = Dm. For odd n = p, or pq with p > q and q ∤ p− 1, our classification theorems guarantee G must be cyclic: that is, n = 5,7,11,13,15,17,19. The only uncertain case is n = 9, but we will show later that G is abelian for n = p2 for prime p > 2.
5. Proposition: If |G| = 2p for a prime p > 2, then either:
Proof: By Sylow I there exists a Sylow p-subgroup H ⊂ G. Since |H| = p is prime, it must be a cyclic group: H = ⟨x⟩ with xp = 1. Let s be the number of Sylow p-subgroups. By Sylow III, s divides 2p, so s = 1, 2, p, 2p, but also s = 1, p+1, 2p+1, . . . , which can only mean s = 1. Thus all the conugates of H are the same, and H is normal.
Now take a Sylow 2-subgroup {e, y} with y2 = e. Since H = ⟨x⟩ is normal, we have yxy−1 = xk for some k = 0, 1, . . . , p−1. We thus get:
Therefore we may take a non-identity central element c ∈ Z(G). This generates a cyclic subgroup ⟨c⟩ whose order divides p2. If |c| = p2, we have G = ⟨c⟩ ≅ Cp2.
Otherwise, |c| = p, and we can take the quotient group of order |G/⟨c⟩| = p. This must be cyclic, generated by a coset [a] = a⟨c⟩ with [a]p = [e] = ⟨c⟩, meaning ap = ci for some i = 0, 1, . . . , p−1. If i > 0, I claim |a| = p2. Indeed, suppose aj+pk = ajcik = e for 0 ≤ j, k ≤ p−1. Since [a]j = [e] only for p | j, we must have j = 0. We are left with cik = e, which means p | ik, so that p | k and k = 0, since 1 ≤ i ≤ p−1. Thus, am ≠ e for 0 < m = j+pk < p2.
Finally if i = 0, we have ap = cp = e with ac = ca, and we may define an isomorphism φ : Cp × Cp → G by φ(ri, rj) = aicj.
n |
k |
pem |
pe |
Now, by definition, we have H·S = S, so we have a new action of the group H on the pe elements g ∈ S. Thus S splits into H-orbits, which are actually the right cosets H·g = Hg of g ∈ S. Since the size of each coset is |Hg| = |H|, this means |H| divides |S| = pe. Combining with the previous conclusion, we must have |H| = pe, and we have found a Sylow p-subgroup.
For H = K, we clearly have NK(H) = K, so that |K∗K| = 1. However, suppose we had some other H ∈ ℋ with NK(H) = K. Then H and K are both Sylow p-subgroups of N = NG(H), so by Sylow II, they are conjugate: K = nHn−1 for n ∈ N. However, H is normal in N, so K = nHn−1 = H. This shows the orbit sizes:
|K*H| = | { |
|
r |
→ |
a |
→ |
6 |
2 |
1. The conjugacy classes (orbits of the conjugation action) are {e}, {r,r2}, {a,ar,ar2}. The only orbit of size 1 is {e}, so the center is Z(G) = {e}. The class equation thus reads: |G| = |Z(G)| + |C(r)| + |C(a)|, i.e. 6 = 1 + 2 + 3.
2. There are 15 sets S ∈ 𝒮: {e,r}, {e,r2}, . . . , {ar,ar2}. The orbits are: G·{e,r} (6 elements), G·{e,a} (3 elements, stabilizer is Sylow 2-subgroup {e,a}); G·{e,ar} (3 elements, stabilizer is Sylow {e,ar}); G·{e,ar2} (3 elements, stabilizer is Sylow {e,ar2}).
3a. G = D3 acts by conjugation on its Sylow 2-subgroups as follows, with gnerator r indicated by black arrows, generator a by yellow arrows:
3b,c. The yellow a-arrows give conjugacy orbits of K = {e,a}. Since the total number of elements s is not divisible by p = 2, there must be some odd-size K-orbit. The proof of Sylow III shows that only K∗K has size 1, with the other K-orbits divisible by p = 2, so that the number of orbits splits as s = 3 = 1 + 2·1 = 1 + pk.
1. Wolfram shows the unique real solution to f(x) = x3 + x − 1 = 0 to be α ≈ 0.682. Since f '(x) = 3x2 + 1 > 0 for all x, the function is increasing, and y = f(x) can only have one x-intercept.
2. The long division of polynomials shows that any polynomial g(x) ∈ Q[x] can be divided by f(x) = x3 + x − 1 to give g(x) = q(x)f(x) + r(x) where the remainder has small degree: deg r(x) < deg f(x) = 3; that is, r(x) = a + bx + cx2. This allows us to write g(α) in standard form:
3. One way to do this is to divide x4 by f(x), getting: x4 = x f(x) + x − x2, so that α4 = 0 + α − α2.
4. Long division (using variable coefficients instead of numbers) gives:
Proof: (i) Among all non-zero polynomials having α as a root, take one of minimal degree. Divide by its leading coefficient an to make it monic, and call the result p(x), with deg p(x) = n. If p(x) were reducible, we would have p(x) = a(x)b(x) with deg a(x), deg b(x) < n. But then p(α) = a(α)b(α) = 0, so either a(α) = 0 or b(α) = 0, which is impossible since we assumed α is not the root of any polynomial of degree less than n. Thus p(x) is irreducible, proving the first part except for uniqueness.
(ii) If f(x) ∈ F[x] is any polynomial with f(α) = 0, perform long division to find f(x) = q(x)p(x) + r(x) with deg r(x) < n. Thus r(α) = f(α) − q(α)p(α) = 0; but as before, α cannot be the root of any non-zero polynomial of degree less than n; thus we must have r(x) = 0 and f(x) = q(x)p(x). This proves the second part.
Finally, to prove the uniqueness of p(x), suppose p1(x) and p2(x) both have the required properties. By (ii), we get p1(x) | p2(x) and by the same reasoning p2(x) | p1(x), so p1(x) = c p2(x) for a constant c; but since both have leading term xn, we must have c = 1. This completes the proof of (i).
Proof: To show that the set spans, consider any element f(α) ∈ F[α]. Since f(x) = q(x)p(x) + r(x) for deg r(x) < n, we have:
To show linear independence, suppose a01 + a1α + ··· + an−1αn−1 = 0, where not all ai = 0. This means r(α) = 0 for a non-zero polynomial with deg r(x) < n, contradicting the assumption that deg p(x) = n is the minimal polynomial of α. Also, linear independence is equivalent to the uniqueness of the coefficients ai .
1a. f(x) = (x−1)3 − 2 = x3 − 3x2 - 1
1b. f(x) = (x−3)2 + 25
1c. f(x) = (x2−1)2 − 5
1d. f(x) = (x−√2−√5)(x−√2+√5)(x+√2−√5)(x+√2+√5) = x4 − 14x2 + 9
2. If α2 is the root of f(x), then α is the root of g(x) = f(x2).
3. Clearly R[a+bi] ⊂ C. Also, for b ≠ 0, we have: i = ((a+bi) − a)⁄b = −a⁄b + 1⁄b(a+bi), so i ∈ R[a+bi], and C = R[i] ⊂ R[a+bi].
4. Suppose dimQ(K) = 2, and take α ∈ K, α ∉ Q. Then {1, α} are linearly independent over Q, and make a basis of K. Thus α2 = c + bα for b,c ∈ Q, and α is the root of x2 − bx − c. The Quadratic Formula thus gives: α = ½(b ± √(b2+4ac)), so α ∈ Q[√d] for d = b2 + 4ac. Therefore K = Q[α] ⊂ Q[√d]. But Q[α] = Q[√d] because both have dimension 2 over Q.
Now, we could have [L : K] < 2 if x2 − 5 becomes reducible in K[x]; that is, if √5 ∈ K = Q[3√2]: could we write √5 = a + b3√2 + c3√4 for rational a,b,c? However, K' = Q[√5] clearly has [K' : F] = 2, so [L : F] = [L : K'] [K' : F] = 2[L : K']. Thus [L : F] is divisible by 2 and 3, and so [L : F] = 6. (That is, √5 ∉ K.) This illustrates how considering fields can easily resolve difficult questions about individual numbers.
Furthermore, for a number like γ = 3√2 + √5, it is not clear that it is the root of a polynomial in Q. However the Finite Extension Theorem tells us it is: since the ambient field L has dimension 6 over F = Q, the 7 powers {1, γ, γ2, . . . , γ6} have a linear dependence over Q, and γ satisfies a polynomial of order at most 6. In fact, by the Degree Product Theorem, the degree of the minimal polynomial must be a divisor or 6, namely 1,2,3, or 6.
1a,b. Wolfram Alpha is useful to compute powers of γ. The matrix of {1, γ, γ2, γ3, γ4} with respect to the basis {1, √2, √3, √6} is:
|
|
1c. To find a linear depenency among {1, γ, γ2, γ3, γ4}, we must solve the linear system given by the matrix equation:
|
| = |
|
2. We are given F ⊂ D ⊂ K with D a ring and F ⊂ K a finite extension. For any δ ∈ D, we have F[δ] ⊂ D, since F ⊂ D, and D is closed under addition and multiplication. Now, δ ⊂ K is algebraic by the Finite Extension Theorem, so we have the simple extension F[δ] = F(δ), meaning the reciprocal 1⁄δ is a polynomial in δ. Thus 1⁄δ ∈ D, which is a field.
1a. W|A
1c. If φ is to be multiplicative, we must have φ(3√4) = φ(3√2)2 = (±3√2)2 = 3√4. But then 1 = φ(1) = φ(3√2·3√4) = ±3√2 3√4 = ±1; and hence φ(3√2) = 3√2. Thus φ must be the identity mapping.
2a. Use the Binomial Theorem (a+b)3 = a3 + 32b + 3ab2 + b3. Once you know ζ3 = 1, it follows that (ζ2)3 = (ζ3)2 = 1, and (ζj 3√2)3 = 2.
2b. Clearly K ⊂ Q[3√2, ζ]. Also ζ = (3√2)2 (ζ 3√2) ∈ L, so Q[3√2, ζ] ⊂ K.
2c. Each ζj 3√2 is clearly a root of the polyonomial, and this means (x − ζj 3√2) is a linear factor; but the degree 3 polyomial p(x) cannot be any more factors than these three.
2d. K has Q-basis given by the powers of α = 3√2 up to 2 (since its minimal polynomial p(x) has degree 3); and the powers of ζ up to 1 (since its minimal polynomial has degree 2). Thus a basis is 1, α, α2, ζ, ζα, ζα2, and [K : Q] = 6.
2e. It is clear that the conjugate field K is equal to K, i.e. K is closed under conjugation, and this is the only property missing to show the conjugation mapping σ is a symmetry of K.
2f. We have ρ(αjζk) = ζj αj ζk = αj ζj+k. Now check ρ(ab) = ρ(a) ρ(b) for each pair of basis elements, or do it on a product of standard forms with W|A.
2g. In cycle notation for S3, we have σ = (23) and ρ = (123). These two elements clearly generate all permutations in S3. Alternatively, one could check the defining relations of D3, namely σ2 = ρ3 = 1 (the identity symmetry) and ρσ = σρ−1.
H | → | L = KH = {α ∈ K with φ(α) = α for all φ ∈ H} |
Gal(K/L) = H | ← | L |
1. The smallest polynomial whose roots include both √2 and √3 is f(x) = (x2−2)(x2−3), and K = Q[±√2, ±√3]. Note f(x) does not need to be irreducible to produce a splitting field.
2. G = {e, σ, τ, στ = τσ} ≅ C2 × C2, where σ(√2) = −√2, σ(√3) = √3; and τ(√2) = √2, τ(√3) = −√3. As we saw in HW 4/1, [K : F] = 4, and this is equal to |G| as stated by Fundamental Theorem I.
3. As in HW 4/1, you can check that 1, γ, γ2, γ3 are linearly independent in the 4-dimensional space K ≅ Q4: that is, the give an invertible 4×4 matrix with respect to the basis {1, √2, √4, √6}. This shows that γ generates K.
As in Notes 3/27, the minimal polynomial is g(x) = (x−√2−√3)(x−√2+√3)(x−+2−√3)(x+√2+√3) = x4 − 10x2 + 1. This must be minimal polynomial (and hence irreducible) since every polynomial h(x) ∈ Q[x] with γ as a root must also have all the conjugates φ(γ) as roots, so that each x−φ(γ) is a linear factor, and so g(x) divides h(x).
4. Since G is abelian, all subgroups are normal, so Fundmental Theorems II and III both apply.
1a. Plot
2. Newton's Method obtains 5 correct decimal places after only 3 steps.
3.
4. Long division gives f(x) = (x − α)(x2 + αx + α2 − 4) + (α3−4 α−1), where the remainder vanishes. Thus β2 + αβ + α2 − 4
5. The relation β2 = −αβ − α2 + 4 allows us to reduce powers βn for n ≥ 2; then α3 = 4α + 1 allows us to reduce all powers αn with n ≥ 3. Thus a Q-basis is {1, α, α2, β, βα, βα2}.
6. We reduce using the relations:
We get ρ similarly. The most complicated computation uses αβ + αγ + βγ = −4.
7. The diagram has H = {e} at the top, under it {e, σ}, {e, σρ}, {e, σρ2}, and {e, ρ, ρ2}; then H = G at the bottom. The corresponding diagram of fixed fields has K at the top, three fields Q[α], Q[β], Q[γ] with degree 3 over Q; one field E with degree 2 over Q; and the base field F = Q at the bottom.
8. To obtain a generator λ of the mysterious field E above with [E : Q] = 2, consider the element:
9. An especially nice element of E is:
1. Tracing the rotation on the picure, we see S(1,0,0) = (1,0,0), S(0,1,0) = (0,−1,0), etc., giving the matrices:
S = |
| R = |
|
2 & 3. Wolfram gives:
S' = RSR−1 = |
| . |
The full group of symmetries is the dihedral group G = {I,R,R2,S,SR,SR2} ≅ D3.
4. The symmetry A must take corners to corners, so A(v1) = vi for i = 1, . . . , 6. Then the vertex counterclockwise from v1 must go to the vertex counterclockwise from vi, the vertex vertical from v1 must go to the vertex vertical from vi, etc., so there is no further choice needed to define A. Hence there are at most 6 symmetries.
5. The trivial subgroup {I} can be obtained by adding any asymmetric decoration to X, for example marking one of the corners. Each of the order 2 subgroups ⟨S⟩, ⟨SR⟩, ⟨SR2⟩ can obtained by marking one of the vertical faces, or two vertices switched by the subgroup. The normal subgroup ⟨R⟩ can be obtained by marking one of the triangular faces, or three vertices around a triangular face.
6. The fixed set of the trivial subgroup {I} is all of X; the fixed sets of the three two-element subgroups are their respective horizontal axes of rotation; the fixed set of the normal 3-element subgroup is its vertical axis of rotation; and the fixed set of all of G is the origin {0}. We clearly have Sym(X ∪ XH) = H, except for the case of H = {I}, since Sym(X ∪ XI) = Sym(X) = G ≠ H. Also, for the three normal subgroups, we have: G/{I} ≅ G = Sym(XI), since XI = X; G/⟨R⟩ ≅ {I, S}, = Sym(XR), since the only non-trivial geometric linear symmetry mapping of the z-axis is a vertical flip; G/G ≅ {I} = Sym(XG) = Sym(0), since a single point has no non-trivial symmetry mappings.
This setup exactly parallels the Main Theorems of Galois Theory, in which G = Gal(K/F), and each subgroup H ⊂ G corresponds to its fixed field KH with Gal(K/KH) = H. Also, for normal subgroups, we have G/H ≅ Gal(KH/F).
We can make the analogy perfect by defining the relative symmetry group of a pair of objects Y ⊂ X to be symmetries of X which fix every point of Y: that is, Sym(X,Y) = {symmetries σ : X → X such that σ(y) = y for all y ∈ Y}. Then we have Sym(X,XH) = H for all subgroups H ⊂ G.