Content
Group Theory
🚧

under construction

Group Theory

Group theory is a fascinating area of mathematics. It reveals deep connections between seemingly unrelated fields, from cryptography to quantum mechanics. The concept of a "group" provides a unifying framework to study symmetry and structure.

The most central definition is the definition of a group:

Definition 5.7: A group is an algebra G;,^,e\langle G; *, \hat{}, e \rangle, where:

  1. Associativity: For all a,b,cGa, b, c \in G, (ab)c=a(bc)(a * b) * c = a * (b * c).
  2. Identity Element: There exists an eGe \in G such that for all aGa \in G, ae=ea=aa * e = e * a = a.
  3. Inverse Element: For each aGa \in G, there exists bGb \in G such that ab=ba=ea * b = b * a = e.

Note that we assume that :G×GG* : G \times G \to G, thus we have closure.

Note that these axioms are not minimal.

Theory & Intuition

Direct Product of Groups

Definition 5.9: The direct product of nn groups G1;1,,Gn;n\langle G_1; *_1 \rangle, \dots, \langle G_n; *_n \rangle is the algebra:

G1××Gn;,\langle G_1 \times \cdots \times G_n; \star \rangle,

where the operation \star is defined component-wise:

(a1,,an)(b1,,bn)=(a11b1,,annbn).(a_1, \dots, a_n) \star (b_1, \dots, b_n) = (a_1 *_1 b_1, \dots, a_n *_n b_n).

Lemma 5.4: G1××Gn;\langle G_1 \times \cdots \times G_n; \star \rangle is a group, where the neutral element and the inversion operation are defined component-wise in the respective groups.

Why are these product groups interesting? Well, as we will see in the following subsection, the notion of an isomorphism provides a way to sort of say two groups have the same structure. In that sense, being able to find an equivalent form written as a direct product of smaller groups, can be very helpful.

Homomorphisms & Isomorphisms

Definition 5.10: For two groups G;,^,e\langle G; *, \hat{}, e \rangle and H;,~,e\langle H; \star, \tilde{}, e' \rangle, a function ψ:GH\psi : G \to H is called a group homomorphism if, for all a,bGa, b \in G:

ψ(ab)=ψ(a)ψ(b).\psi(a * b) = \psi(a) \star \psi(b).

If ψ\psi is a bijection from GG to HH, then it is called an isomorphism, and we say that GG and HH are isomorphic and write GHG \simeq H.

What does this actually mean? It sort of means, that it does not matter if we apply the map ψ\psi first and compute the group operation \star (in HH) and then look at the result in HH or if we compute the product * (in GG) and the apply the homomorphism ψ\psi and look at the output in HH. Both ways yield the same result if the equation above holds. So one could intuitively say the role of an element aa when computing the group operation with some other element bb in GG, is preserved through the mapping ψ\psi. While this observation and description is very imprecise, the following section should make the intuition behind this kind of preservation clear. The little property from above actually lets us derive more nice properties of ψ\psi. Namely:

Lemma 5.5: A group homomorphism ψ\psi from G;,^,e\langle G; *, \hat{}, e \rangle to H;,~,e\langle H; \star, \tilde{}, e' \rangle satisfies:

  1. ψ(e)=e\psi(e) = e',
  2. ψ(a^)=ψ(a)~\psi(\hat{a}) = \tilde{\psi(a)} for all aGa \in G.

Intuition about Homomorphisms

Why does a homomorphism map the neutral element in GG to the neutral element in HH? Well, this is exactly what we would expect intuitively. The neutral element in a group is like "doing nothing" to an element - it's the identity operation. If we apply ψ\psi after eGe \in G has acted on aGa \in G, it should produce the same result as letting ψ\psi map aa first and then applying eHe' \in H.

But here's something interesting: does every homomorphism have to be injective? No! For instance, consider an extreme case where the entire group GG maps to just the neutral element eHe' \in H. This mapping satisfies the homomorphism property because applying the neutral element under any group operation still respects the structure.

A nice way to think of this is like projecting a 3D object onto a flat surface: the object's details might be lost, but the structure of relationships between points can still hold.

To gain intuition about homomorphisms, I think the following image is incredibly helpful:

Von Cronholm144 - Eigenes Werk, CC BY-SA 3.0, Link

So what's going on in this picture? First of all we have to clarify some naming and notation:

  1. h:GHh : G \to H is a homomorphism in this picture
  2. kerh\ker h is the so-called kernel of the map hh. This is simply the set of all elements which map to the neutral element. You might have heard of the concept of kernel or null-space in linear algebra. The concept is related - for every linear map (or more concretely matrix), there is some set of elements in the domain, which maps to 00 in the codomain. So in our case it is simply consider the following set:
kerh={aGh(g)=1H}\ker h = \{a \in G \mid h(g) = 1_H\}

So as you can see in the picture and as pointed out before, there can be more than one element in the kernel. We could actually have the case that kerh=G\ker h = G.

  1. imh\text{im} h is the so-called image of the map hh. Again you might have heard about a concept like this in linear algebra. It is simply the set of all elements we can possibly get as an output:
imh={h(a)aG}\text{im} h = \{h(a) \mid a \in G\}
  1. The set aNa N is a so-called coset. This concept is not explicitly stated in this course, but super useful when you explore more properties. You will find more about it in the Cosets section. We defined N=kerhN = \ker h (i.e. all elements in GG which map to 1H1_H). Now for an arbitrary element aGa \in G, we can define:
aN={aGggN}a N = \{a *_G g \mid g \in N\}

So what we are doing here is simply taking an element and multiplying it (in the sense of the group operation in GG) to all elements which are in NN and making a set of all the outputs. We can do this in arbitrary NGN \subseteq G, however in this case we choose kerh\ker h.

What is interesting about this picture?

  1. The kernel forms the foundation of symmetry in GG - all elements in the kernel behave indistinguishably under hh.

  2. We can sort of see at least in the picture (there is no reason why we should expect/believe this intuitively in the first place!) that there are some sort of classes of elements (e.g. aNa N) which map all to the same value h(a)h(a). These can be built from these ominous constructions we call cosets. There is no reason to believe that this is actually true... well there is - after we proved interesting properties which are more investigated below - also take a look at the exercises.

  3. Think about what would happen in the picture above if hh was injective!

There are so many more connections here which Let's try to look at a specific example to get a better grasp of what we have done until now.

With this deeper understanding, try exploring homomorphisms in different contexts - like how polynomials, matrices, or symmetry groups map to each other. It's all connected!

Intuition of Isomorphisms

A group isomorphism is a special type of group homomorphism that provides a perfect "translation" between two groups. If two groups are isomorphic, they have the same structure, just represented differently. Every operation in one group corresponds exactly to an operation in the other group under the isomorphism. Importantly, the correspondence is bijective, meaning it pairs every element of one group with a unique element of the other group and vice versa.

An isomorphism between two groups GG and HH is a bijective map ϕ:GH\phi: G \to H such that:

  1. ϕ(g1g2)=ϕ(g1)ϕ(g2)\phi(g_1 \cdot g_2) = \phi(g_1) \cdot \phi(g_2) for all g1,g2Gg_1, g_2 \in G (preserves the group operation).
  2. ϕ\phi is a one-to-one and onto mapping (bijective).

If such a map exists, GG and HH are said to be isomorphic, denoted GHG \simeq H.

Analogy: Think of two jigsaw puzzles that look completely different on the surface but are constructed from pieces with identical shapes. You can perfectly match every piece in one puzzle with a piece in the other puzzle, and the way the pieces fit together is preserved. The isomorphism is the "perfect match" function that pairs each piece in one puzzle with the corresponding piece in the other.

In the same way, group isomorphisms preserve the "shape" of the group's structure, ensuring every element and operation in one group corresponds to those in another.

When thinking about a set of all possible groups. It is the case that the \simeq relation is an equivalence relation. Now it really gets interesting. We could now ask questions like how many different groups (in the sense of equivalence classes) are there that have kk elements? There are super interesting results about this question. For example that if kk is prime, there is only one group, i.e. all groups of order kk are isomorphic.


Steps to Prove an Isomorphism

To rigorously prove that two groups GG and HH are isomorphic, follow these steps:

  1. Define a Candidate Map: Identify a function ϕ:GH\phi: G \to H that you suspect to be an isomorphism.

  2. Verify the Map is Well-Defined: Ensure that the proposed map is unambiguous and consistent. That is, for any gGg \in G, ϕ(g)\phi(g) is uniquely determined.

  3. Verify the Map is Totally Defined: Confirm that the map is defined for all elements of GG (i.e., ϕ\phi applies to every element of GG).

  4. Verify the Map Maps to the Codomain: Ensure that ϕ(g)H\phi(g) \in H for all gGg \in G, so that the image of ϕ\phi lies entirely within HH.

  5. Check the Homomorphism Property: Verify that ϕ(g1g2)=ϕ(g1)ϕ(g2)\phi(g_1 \cdot g_2) = \phi(g_1) \cdot \phi(g_2) for all g1,g2Gg_1, g_2 \in G. This ensures the map preserves the group operation.

  6. Check Injectivity: Prove that ϕ\phi is one-to-one by showing that if ϕ(g1)=ϕ(g2)\phi(g_1) = \phi(g_2), then g1=g2g_1 = g_2.

  7. Check Surjectivity: Prove that ϕ\phi is onto by demonstrating that for every hHh \in H, there exists gGg \in G such that ϕ(g)=h\phi(g) = h.

  8. Conclude Isomorphism: If the map satisfies the homomorphism property, is well-defined, maps to the codomain, and is bijective, then ϕ\phi is an isomorphism, and GHG \simeq H.

Example

Prove that Z6Z2×Z3\mathbb{Z}_6 \simeq \mathbb{Z}_2 \times \mathbb{Z}_3.

Step 1: Define a Candidate Map

Define ϕ:Z6Z2×Z3\phi: \mathbb{Z}_6 \to \mathbb{Z}_2 \times \mathbb{Z}_3 by:

ϕ(n)=(R2(n),R3(n))\phi(n) = (R_2(n), R_3(n))

where Rk(n)R_k(n) denotes the remainder when nn is divided by kk (i.e., Rk(n){0,1,,k1}R_k(n) \in \{0, 1, \dots, k-1\}).

Step 2: Verify the Map is Well-Defined

The group Z6\mathbb{Z}_6 consists of integers modulo 6. For any nZ6n \in \mathbb{Z}_6, R2(n)R_2(n) and R3(n)R_3(n) are uniquely determined remainders modulo 2 and 3, respectively. Since remainders are unique, ϕ(n)\phi(n) produces unique and consistent pairs (R2(n),R3(n))(R_2(n), R_3(n)). Thus, ϕ\phi is well-defined.

Step 3: Verify the Map is Totally Defined

The map ϕ\phi is defined for all nZ6n \in \mathbb{Z}_6. Since R2(n)R_2(n) and R3(n)R_3(n) exist for any integer nn, ϕ\phi is totally defined.

Step 4: Verify the Map Maps to the Codomain

The codomain is Z2×Z3\mathbb{Z}_2 \times \mathbb{Z}_3, where Z2={0,1}\mathbb{Z}_2 = \{0, 1\} and Z3={0,1,2}\mathbb{Z}_3 = \{0, 1, 2\}. For any nZ6n \in \mathbb{Z}_6, R2(n)Z2R_2(n) \in \mathbb{Z}_2 and R3(n)Z3R_3(n) \in \mathbb{Z}_3, so ϕ(n)Z2×Z3\phi(n) \in \mathbb{Z}_2 \times \mathbb{Z}_3. Thus, ϕ\phi maps into the codomain.

Step 5: Check the Homomorphism Property

Let a,bZ6a, b \in \mathbb{Z}_6. We can compute

ϕ(a+6b)=(R2(a)+2R2(b),R3(a)+3R3(b)).\phi(a +_6 b) = (R_2(a) +_2 R_2(b), R_3(a) +_3 R_3(b)).

(where +m+_m means addition modulo mm, i.e. we take remainder modulo mm after doing addition).

Using the modular arithmetic property a+kb=Rk(a+b)=Rk(Rk(a)+Rk(b))a +_k b = R_k(a + b) = R_k(R_k(a) + R_k(b)), we have:

ϕ(a+6b)=(R2(R2(a)+2R2(b)),R3(R3(a)+3R3(b)))\phi(a +_6 b) = (R_2(R_2(a) +_2 R_2(b)), R_3(R_3(a) +_3 R_3(b)))

On the other hand:

ϕ(a)+6ϕ(b)=(R2(a),R3(a))+(R2(b),R3(b))=(R2(a)+2R2(b),R3(a)+3R3(b))\phi(a) +_6 \phi(b) = (R_2(a), R_3(a)) + (R_2(b), R_3(b)) = (R_2(a) +_2 R_2(b), R_3(a) +_3 R_3(b))

Therefore

ϕ(a+6b)=ϕ(a)+ϕ(b)\phi(a +_6 b) = \phi(a) + \phi(b)

Thus, ϕ\phi preserves the group operation.

Step 6: Check Injectivity

Assume ϕ(a)=ϕ(b)\phi(a) = \phi(b). Then:

(R2(a),R3(a))=(R2(b),R3(b)).(R_2(a), R_3(a)) = (R_2(b), R_3(b)).

This implies:

R2(a)=R2(b)andR3(a)=R3(b).R_2(a) = R_2(b) \quad \text{and} \quad R_3(a) = R_3(b).

Hence, a2ba \equiv_2 b and a3ba \equiv_3 b. By the Chinese Remainder Theorem (as gcd(2,3)=1\gcd(2,3)=1), a6ba \equiv_6 b. Thus, a=ba = b in Z6\mathbb{Z}_6, and ϕ\phi is injective.

Step 7: Check Surjectivity

For any (x,y)Z2×Z3(x, y) \in \mathbb{Z}_2 \times \mathbb{Z}_3, where x{0,1}x \in \{0, 1\} and y{0,1,2}y \in \{0, 1, 2\}, the Chinese Remainder Theorem (gcd(2,3)=1\gcd(2,3)=1) guarantees the existence of an integer zz such that:

z2xandz3y.z \equiv_2 x \quad \text{and} \quad z \equiv_3 y .

By construction, zZ6z \in \mathbb{Z}_6, and ϕ(z)=(R2(z),R3(z))=(x,y)\phi(z) = (R_2(z), R_3(z)) = (x, y). Thus, ϕ\phi is surjective.

Step 8: Conclude Isomorphism

Since ϕ\phi is well-defined, totally defined, maps to the codomain, preserves the group operation, and is bijective, ϕ\phi is an isomorphism. Therefore, Z6Z2×Z3\mathbb{Z}_6 \simeq \mathbb{Z}_2 \times \mathbb{Z}_3.

Subgroups

Order

Cyclic Groups

Lagrange Theorem

Lagrange's Theorem states:

If GG is a finite group and HH is a subgroup of GG, then the order (number of elements) of HH divides the order of GG.

We want to prove this theorem here! Because it illustrates interesting concepts and provides deep insights into the structure of groups

Useful Definitions

Definition (Left Coset): Let HH be a subgroup of GG, and let gGg \in G. The left coset of HH with respect to gg is the set:

gH={ghhH}.gH = \{ gh \mid h \in H \}.

Key Properties of Cosets

Lemma (Equal Size of Cosets): Every left coset gHgH has the same number of elements as the subgroup HH. Specifically, gH=H|gH| = |H|.

Proof: Define a map φ:HgH\varphi: H \to gH by φ(h)=gh\varphi(h) = gh. We show that φ\varphi is a bijection:

  1. Injectivity: Assume φ(h1)=φ(h2)\varphi(h_1) = \varphi(h_2). Then:

    gh1=gh2.gh_1 = gh_2.

    Multiplying on the left by g1g^{-1}, we get h1=h2h_1 = h_2. Thus, φ\varphi is injective.

  2. Surjectivity: Let xgHx \in gH. By definition of gHgH, x=ghx = gh for some hHh \in H. Thus, there exists hHh \in H such that φ(h)=x\varphi(h) = x. Hence, φ\varphi is surjective.

Since φ\varphi is both injective and surjective, it is a bijection, and gH=H|gH| = |H|.

Lemma (Disjoint or Equal Cosets): If g1Hg2Hg_1H \cap g_2H \neq \emptyset, then g1H=g2Hg_1H = g_2H.

Proof: Assume g1Hg2Hg_1H \cap g_2H \neq \emptyset. Then there exists xg1Hg2Hx \in g_1H \cap g_2H, so xg1Hx \in g_1H and xg2Hx \in g_2H. Thus:

x=g1h1=g2h2,x = g_1h_1 = g_2h_2,

for some h1,h2Hh_1, h_2 \in H. Rearranging, we get:

g21g1=h2h11.g_2^{-1}g_1 = h_2h_1^{-1}.

Since h2h11Hh_2h_1^{-1} \in H (because HH is a subgroup), g21g1Hg_2^{-1}g_1 \in H. Let h=g21g1Hh = g_2^{-1}g_1 \in H, so:

g1=g2h.g_1 = g_2h.

Now consider any element xg1Hx \in g_1H. By definition, x=g1hx = g_1h' for some hHh' \in H. Substituting g1=g2hg_1 = g_2h, we get:

x=g2(hh)g2H.x = g_2(hh') \in g_2H.

Thus, g1Hg2Hg_1H \subseteq g_2H. By symmetry, g2Hg1Hg_2H \subseteq g_1H. Hence, g1H=g2Hg_1H = g_2H.

Lemma (Cosets Partition GG): The cosets gHgH form a partition of GG, meaning:

  1. Every element of GG is in some coset gHgH.
  2. Two cosets g1Hg_1H and g2Hg_2H are either disjoint or identical.

Proof:

  1. By definition of a coset, for any gGg \in G, ggHg \in gH. Thus, every element of GG is in at least one coset.
  2. By the previous lemma, if g1Hg2Hg_1H \cap g_2H \neq \emptyset, then g1H=g2Hg_1H = g_2H. Otherwise, g1Hg2H=g_1H \cap g_2H = \emptyset. Hence, the cosets are either disjoint or identical.

Proof of Lagrange's Theorem

Let GG be a finite group and HH a subgroup of GG. By the lemmas above, the cosets gHgH partition GG. Let kk be the number of distinct cosets. Since each coset has size H|H| and the cosets are disjoint, the total number of elements in GG is:

G=kH,|G| = k \cdot |H|,

where k=[G:H]k = [G : H] is the number of cosets (the index of HH in GG). Thus, H|H| divides G|G|.

Euler's Totient Function and multiplicative groups

RSA

Digressions

Cosets

Finite Abelian Groups

A fundamental result for finite abelian groups is the Fundamental Theorem of Finite Abelian Groups, which states that every finite abelian group is isomorphic to a direct product of cyclic groups of prime power order.

Representation Theory

Representation theory studies groups by representing their elements as matrices or linear transformations. This approach is widely used in quantum mechanics, crystallography, and other fields.

Exercises

Finite group of even orderDifficulty: 2/5Relevance: 🎓🎓🎓Source:  Abstract Algebra: Theory and Applications

Show that if GG is a finite group of even order, then there is an aGa \in G such that aa is not the identity and

a2=ea^2 = e
Subgroups of a Cyclic GroupDifficulty: 3/5Relevance: 🎓🎓🎓🎓Source:  Tobias Steinbrecher

Let G;,^,e\langle G; *, \hat{}, e \rangle be a cyclic group. Prove that every subgroup of GG is cyclic.

The Square RootDifficulty: 2/5Relevance: 🎓🎓🎓🎓Source:  Yannick Funke

Let GG be a (finite) group of odd order. Define the operation :GG\sqrt{} : G \to G such that

a2=a\sqrt{a}^2 = a

Show that \sqrt{} is a bijective, well- and totally defined function.