Archive for the ‘Group Algebras’ Category

Throughout this post, k is a field.

Let G be a finite group. If |G| \ge 2, then k[G] is never a domain, a simple fact that we proved here (see Lemma 1). Now, let’s weaken the condition and ask whether or not k[G] can have non-zero nilpotent elements. The answer this time is positive. For example, let p be a prime number, k a field of characteristic p and G a finite group such that p \mid |G|. Let g \in G be an element of order p. Then 0 \ne 1-g \in k[G] is nilpotent because (1-g)^p=1-g^p=0.

A much more interesting question: for which fields k and groups G, does the group algebra k[G] have non-zero nilpotent elements? That is not easy to answer in general but it is not hard to give a useful necessary condition for k[G] to have no non-zero nilpotent elements, which is the subject of this post.

Recall that a ring is said to be reduced if it has no non-zero nilpotent element.

Lemma. Let H be a finite subgroup of a group G. If k[G] is reduced, then |H| is invertible in k.

Proof. That is obvious if the characteristic of k is zero. So suppose that the characteristic of k is p > 0 and assume, to the contrary, that |H| is not invertible in k, i.e. |H| is zero in k or, equivalently, p \mid |H|. Then, since p is a prime number, H has an element h of order p. But then (1-h)^p=1-h^p=0, contradicting our assumption that k[G] is reduced. \ \Box

Proposition. Let G be a group. If k[G] is reduced, then every finite subgroup of G is normal.

Proof. Let H=\{h_1, \cdots , h_n\} be a finite subgroup of G. Note that, by the Lemma, n is invertible in k and so

\displaystyle x:=\frac{1}{n}\sum_{i=1}^nh_i \in k[G].

Now,

\displaystyle \begin{aligned}x^2=\frac{1}{n^2}\left(\sum_{i=1}^nh_i\right)^2=\frac{1}{n^2}\sum_{i=1}^nh_i\sum_{j=1}^nh_j=\frac{1}{n^2}\sum_{i=1}^n\sum_{j=1}^nh_ih_j=\frac{1}{n^2}\sum_{i=1}^n\sum_{j=1}^nh_j =\frac{1}{n^2}\sum_{i=1}^nnx=x.\end{aligned}

So x is an idempotent element of k[G] and hence, since k[G] is reduced, x is central, by Remark 3 in this post. Thus gxg^{-1}=x for all g \in G, which gives

\displaystyle \sum_{i=1}^ngh_ig^{-1}=\sum_{i=1}^nh_i.

So for each i, there exists a j such that gh_ig^{-1} =h_j \in H, proving that H is normal. \ \Box

Example. Let G:=D_6, the dihedral group of order 6. So G is generated by two elements a,b, where the orders of a,b are two and three, respectively, and ba=ab^{-1}. The subgroup of G generated by a is not normal in G and hence, by the Proposition, k[G] is not reduced. We can also prove that directly; just see, for example, that (a+b-b^2-ab)^2=0 and so a+b-b^2-ab is a non-zero nilpotent element of k[G].

Exercise 1. Show that the converse of the above Proposition is not always true.

Exercise 2. Let D_{2n} be the dihedral group of order 2n. Show that k[D_{2n}], \ n \ge 3, always has a non-zero nilpotent element. What about n=1,2 ?

Note. The reference for the proof of the Theorem in this post is the first few pages of this paper. However, I have modified the proof given by the author of the paper quite a bit and also added all the details the reader needs to fully understand the proof.

Throughout this post, k is a field, \text{char}(k) is the characteristic of k, and G is a finite group. Also, as always, J(R), U(R) denote, respectively, the Jacobson radical and the group of units of a ring R. Finally, for a group algebra R:=k[G], we denote by \mathfrak{a}(R) the augmentation ideal of R, i.e. \mathfrak{a}(G) is the ideal of R generated by the set \{g-1: \ g \in G\}. See that \mathfrak{a}(G) equals the k-vector space spanned by \{g-1: \ g \in G\}.

In this post, we prove that the group algebra k[G] is local if and only if \text{char}(k)=p > 0 and G is a p-group. This result is not very easy to prove; we need to do some preparation before getting into the proof.

Lemma 1. The group algebra R:=k[G] is a domain if and only if |G| =1.

Proof. i) If |G|=1, then R \cong k and so R is a domain. Now suppose that |G| > 1 and choose an element 1 \ne g \in G of order d. Then d > 1, \ g^d=1, and the elements 1, g, \cdots , g^{d-1} are pairwise distinct. Thus 1 - g \ne 0, \ 1+g + \cdots + g^{d-1} \ne 0 but

(1-g)(1+g+ \cdots + g^{d-1})=1-g^d=0,

which proves that R is not a domain. \ \Box

Lemma 2. For the group algebra R:=k[G], the following statements are equivalent.

i) R is local.

ii) \mathfrak{a}(R) \subseteq J(R).

iii) \mathfrak{a}(R) is nil.

iv) \mathfrak{a}(R)=J(R).

Proof. i) \implies ii). By Theorem 1 in this post, R \setminus U(R)=J(R). Since R \setminus U(R) contains every proper ideal of R, we have \mathfrak{a}(R) \subseteq J(R).

ii) \implies iii). Since R is Artinian, J(R) is nil, by the Example in this post, and so \mathfrak{a}(R) is nil too. In fact J(R) is nilpotent but we don’t need that here.

iii) \implies iv). By the Remark in this post, \mathfrak{a}(R) \subseteq J(R). By the Lemma in this post, R/\mathfrak{a}(R) \cong k. Therefore J(R)/\mathfrak{a}(R) is isomorphic to a proper ideal of the field k hence \mathfrak{a}(R)=J(R).

iv) \implies i). Again, by the Lemma in this post, R/J(R) \cong k. Since k is a field, hence a division ring, R is local, by Theorem 1 in this post. \ \Box

Lemma 3. Let G be a cyclic group of order n. Then

i) k[G] \cong k[x]/\langle x^n-1\rangle,

ii) k[G] is a (commutative) local ring if \text{char}(k)=n.

Proof. i) Let g be a generator of G. By definition, \{1, g, \cdots , g^{n-1}\} is a k-basis for k[G]. Define the map f: k[x] \to k[G] by

f(c_0+c_1x + \cdots c_mx^m)=c_0+c_1g+ \cdots + c_mg^m, \ c_i \in k.

Clearly f is a well-defined onto k-algebra homomorphism. So we just need to show that \ker f = \langle x^n-1\rangle. By definition, f(x^n-1)=g^n-1=0 and so \langle x^n-1\rangle \subseteq \ker f. Conversely, let p(x) \in \ker f. We have p(x)=q(x)(x^n-1)+r(x) for some q(x),r(x) \in k[x] with r(x)=0 or \deg r(x) < n. If r(x)=0, we are done. If r(x)=\sum_{i=0}^m c_ix^i, \ c_i \in k, \ m < n, then

0=f(p(x))=f(q(x))f(x^n-1)+f(r(x)) =f(r(x))=c_0+c_1g+ \cdots + c_mg^m

giving c_i=0 for all i. So again r(x)=0.

ii) Since \text{char}(k)=n, we have in k[x], \ x^n-1=(x-1)^n and so, by i), k[G] \cong k[x]/\langle (x-1)^n \rangle, which is clearly a commutative local ring ring with the unique maximal ideal (x-1)/\langle (x-1)^n\rangle. \ \Box

Lemma 4. Let Z be the center of G and suppose that N \subseteq Z is a subgroup of G. If both k[N] and k[G/N] are local, then k[G] is local too.

Proof. By the Corollary in this post, k[G/N] \cong k[G]/I, where I is the ideal of k[G] generated by the set \{g-1: \ g \in N\}. Note that since N is central in G, the ideal I is central in k[G]. Since, by definition, \mathfrak{a}(k[N]) is the ideal of k[N] generated by the set \{g-1: \ g \in N\}, we have I=\mathfrak{a}(k[N])k[G]. So, since k[G/N] is local, R:=k[G]/I is local too and hence \mathfrak{a}(R) is nil, by Lemma 2. Now, let x \in \mathfrak{a}(k[G]). Then x+I \in \mathfrak{a}(R) and so x^n \in I for some integer n. Since k[N] is local, \mathfrak{a}(k[N]) is nil, by Lemma 2, and since k[N] is in the center of k[G], the ideal I=\mathfrak{a}(k[N])k[G] is also nil. Hence x is nil, because x^n \in I, and therefore \mathfrak{a}(k[G]) is nil, proving that k[G] is local, by Lemma 2. \ \Box

Theorem. If |G| > 1, then R:=k[G] is local if and only if \text{char}(k)=p > 0 and G is a p-group.

Proof. Suppose first that R is local. If \text{char}(k)=0, then by Maschke‘s theorem J(R)=(0) and so R must be a division ring. But then R would be a domain, contradicting Lemma 1. So \text{char}(k)=p > 0. Now let q be any prime divisor of |G| and let g \in G be an element of order q. Since

(1-g)(1+g+ \cdots + g^{q-1})=1-g^d=0,

we have 1-g \notin U(R) and 1+g+ \cdots + g^{q-1} \notin U(R). Hence, since

1-g^m=(1-g)(1+g+ \cdots + g^{m-1}),

we have 1-g^m \notin U(R) for all positive integers m. Since R is local, R \setminus U(R) is an ideal of R and so

\displaystyle q1_G=1+g+ \cdots +g^{q-1}+\sum_{m=1}^{q-1}(1-g^m) \notin U(R).

But if q \ne 0 in k, then q1_G \in U(R) because q^{-1}1_G would be the inverse of q1_G. Therefore q=0 in k, i.e. p=\text{char}(k) \mid q, which gives q=p. Thus p is the unique prime divisor of |G| and hence G is a p-group.
Conversely, suppose that \text{char}(k)=p > 0 and G is a p-group. We want to show that k[G] is local. The proof is by induction over |G|. If |G|=p, then we are done by Lemma 3, ii). Now suppose that |G| > p and let Z be the center of G. Since G is a p-group, |Z| \ge p. Let N \subseteq Z be a subgroup of order p. Clearly |N| < |G| and |G/N| < |G|. So, by our induction hypothesis, both k[N] and k[G/N] are local. Thus, by Lemma 4, k[G] is local too. \ \Box

Exercise. Show that k[\mathbb{Z}] \cong k[x,x^{-1}], the ring of Laurent polynomials. Conclude that k[\mathbb{Z}] has infinitely many maximal ideals, and so it’s not even semilocal, never mind local.
Hint. To prove that k[\mathbb{Z}] has infinitely many maximal ideals, note that k[x,x^{-1}] is the localization of k[x] by the set \{1, x, x^2, \cdots \}.

Throughout this post, R is a commutative ring with identity.

Let G be a group with a normal subgroup N. What is the relationship between the group algebras R[G] and R[G/N] ? This is particularly important if G is finite and N is proper because then we might be able to use induction on |G| to extend certain properties that we know they hold in R[G/N] to R[G]. The question is fairly easy to answer. Let’s prove a more general result first.

Proposition. Let G_1,G_2 be two groups. Let \alpha: G_1 \to G_2 be a group homomorphism with K:=\ker \alpha. Consider the map \alpha': R[G_1] \to R[G_2] defined by

\alpha'(r_1g_1+ \cdots +r_ng_n)=r_1\alpha(g_1)+ \cdots + r_n\alpha(g_n),

for all r_i \in R, g_i \in G. Then \alpha' is a ring homomorphism and I:=\ker \alpha' is the ideal of R[G_1] generated by the set S:=\{k-1: \ k \in K\}.

Proof. It is easy to see that \alpha' is a well-defined ring homomorphism. Also, if s=k-1 \in S, then, by definition, \alpha'(s)=\alpha(k)-\alpha(1)=1-1=0 and so S \subseteq I implying that \langle S \rangle \subseteq I. The non-trivial part of the proof is to show that \langle I \subseteq \langle S \rangle. Let x=r_1g_1+ \cdots +r_ng_n \in I, where r_i \in R, g_i \in G_1, After relabeling, if necessary, we may assume that the cosets Kg_1, \cdots , Kg_m are all the pairwise distinct cosets in the set \{Kg_1, \cdots Kg_n\}. So we can re-write x as

\displaystyle x=\sum_{i=1}^m \sum_{j=1}^{n_i}r_{ij}k_{ij}g_i,

where r_{ij} \in R, k_{ij} \in K. Since x \in I, we have f(x)=0 and so, since f(k_{ij})=1, because k_{ij} \in K, we have

\displaystyle 0=f(x)=\sum_{i=1}^m \sum_{j=1}^{n_i}r_{ij}f(k_{ij}g_i)=\sum_{i=1}^m \sum_{j=1}^{n_i}r_{ij}f(k_{ij})f(g_i)=\sum_{i=1}^m \sum_{j=1}^{n_i}r_{ij}f(g_i).

Now, f(g_i) \ne f(g_j) for all 1 \le i \ne j \le m because, as we assumed, Kg_i \ne Kg_j. Thus, since G_2 is an R-basis for R[G_2], we get that \displaystyle \sum_{j=1}^{n_i}r_{ij}=0 for all 1 \le i \le m, and hence

\displaystyle x=\sum_{i=1}^m \sum_{j=1}^{n_i}r_{ij}k_{ij}g_i=\sum_{i=1}^m \sum_{j=1}^{n_i}r_{ij}(k_{ij}-1)g_i \in \langle S \rangle. \ \Box

Corollary. Let G be a group with a normal subgroup N. Then \displaystyle R[G/N] \cong R[G]/I, where I is the ideal of R[G] generated by the set \{g-1: \ g \in N\}.

Proof. Consider the natural group homomorphism \alpha: G \to G/N. Then the Proposition gives a group homomorphism \alpha': R[G] \to R[G/N] with I:=\ker \alpha'=\langle S \rangle, where S=\{g-1: \ g \in \ker \alpha\}. The result now follows because \ker \alpha=N and, since \alpha is onto, \alpha' is onto too. \ \Box

Remark. In the above Corollary, if we choose N=G, then I will be the augmentation ideal of R[G], which we have already seen in this post.

Let R be a commutative ring with identity and let G be a group. Recall that the group ring, also called group algebra, R[G] is a free R-module which is also a ring. As a free R-module, G is a basis for R[G]. So an element x \in R[G] has the form x=r_1g_1+ \cdots +r_ng_n, for some g_i \in G, \ r_i \in R, and, since G is a basis for the free R-module R[G], this representation of x is unique if g_i \ne g_j for all i \ne j. Elements of R are all central, i.e. they are in the center of R[G], which is why we also call R[G] a group algebra. In this post, we take a look at a very important ideal of R[G] called the augmentation ideal.

Lemma. Let R be a commutative ring and let G be a group. We define the map f : R[G] \longrightarrow R by

f(\sum_{g \in G} r_g g)=\sum_{g \in G} r_g.

Then f is an onto ring homomorphism and hence R[G]/\ker f \cong R. Also, \{g - 1_G: \ g \in G, g \neq 1_G \} is a basis for the free R-module \ker f.

Proof. Obviously f is well-defined, additive and onto. Now, let x = \sum_{g \in G} r_g g, \ y = \sum_{g \in G} s_g g. Then

f(xy) = f(\sum_{g \in G} (\sum_{g_1g_2=g} r_{g_1}s_{g_2}) g)=\sum_{g \in G} \sum_{g_1g_2=g} r_{g_1}s_{g_2}=(\sum_{g \in G} r_g)(\sum_{g \in G} s_g)

=f(x)f(y).

So f is a ring homomorphism. Therefore \ker f is an ideal of R[G] and hence an R-module. Now, x = \sum_{g \in G} r_g g \in \ker f if and only if \sum_{g \in G}r_ g = 0 if and only if

x = x - (\sum_{g \in G}r_g)1_G=\sum_{g \in G}r_g(g-1_G).

Thus \ker f, as an R-module, is generated by the set \{g - 1_G : \ g \in G \}. Then, obviously, the set B=\{g-1_G: \ g \in G, g \neq 1_G\} still generates \ker f. To show that B is a basis for \ker f, as an R-module, we suppose that \sum_{g \in B}r_g (g-1_G)=0. Then

\sum_{g \in B}r_g g = (\sum_{g \in B}r_g)1_G.

But g \neq 1_G for all g \in B, and so r_g = 0 for all g \in B. \ \Box

Definition. The ring homomorphism f, as defined in the lemma, is called the augmentation map and \ker f is called the augmentation ideal of R[G].

Hurewicz Theorem. Let G be a group with the commutator subgroup G'. Let I be the augmentation ideal of \mathbb{Z}[G] and consider I as an additive group. Then \displaystyle \frac{G}{G'} \cong \frac{I}{I^2}.

Proof. Define the map \displaystyle \varphi : G \longrightarrow \frac{I}{I^2} by \varphi(g)=g-1_G + I^2. Clearly \varphi is well-defined because g - 1_G \in I for all g \in G. Also, since (g_1 - 1_G)(g_2-1_G) \in I^2, we have

\varphi(g_1g_2)=g_1g_2-1_G + I^2 = g_1g_2 -1_G - (g_1-1_G)(g_2-1_G) + I^2=

g_1 - 1_G + g_2 - 1_G + I^2 = \varphi(g_1) + \varphi(g_2).

Thus \varphi is a group homomorphism. So \varphi(g^{-1})=-\varphi(g) and therefore, since I is an abelian group, we have \varphi(g_1g_2g_1^{-1}g_2^{-1})=\varphi(g_1) + \varphi(g_1^{-1})+\varphi(g_2)+\varphi(g_2^{-1})=0. Thus G' \subseteq \ker \varphi and hence we have a group homomorphism \displaystyle \overline{\varphi}: \frac{G}{G'} \longrightarrow \frac{I}{I^2} defined by

\overline{\varphi}(gG')=g - 1_G +I^2.

Now, to show that \overline{\varphi} is an isomorphism, we will find an inverse for it. Define the map \psi : I \longrightarrow G/G' by \psi(\sum_{g \in G} n_g(g-1_G))=(\prod_{g \in G} g^{n_g})G'. Note that since g_1g_2G'=g_2g_1G' for all g_1,g_2 \in G, the map \psi is a well-defined group homomorphism. Now,

\psi((g_1 - 1_G)(g_2-1_G))=\psi(g_1g_2 -1_G -(g_1-1_G)-(g_2-1_G))=g_1g_2g_1^{-1}g_2^{-1}G' =G'.

So I^2 \subseteq \ker \psi because, by the lemma, the set \{(g_1-1_G)(g_2-1_G): \ g_1,g_2 \in G\} generates the additive group I^2. Hence the map \displaystyle \overline{\psi} : \frac{I}{I^2} \longrightarrow \frac{G}{G'} defined by

\overline{\psi}(\sum_{g \in G}n_g (g - 1_G)+ I^2) = (\prod_{g \in G} g^{n_g})G'

is a well-defined group homomorphism. It is now easy to see that both \overline{\varphi} \circ \overline{\psi} and \overline{\psi} \circ \overline{\varphi} are identity maps. \Box

Theorem. Let k be a field and let G be a finite group. Let \mathcal{C}_1, \cdots , \mathcal{C}_m be the conjugacy classes of G. Let

\displaystyle c_i = \sum_{g \in \mathcal{C}_i} g \in k[G], \ \ \ \ \ 1 \leq i \leq m.

Then the set \{c_1, \cdots , c_m \} is a k-basis for Z(k[G]), the center of k[G], and thus \dim_k Z(k[G]) is equal to the number of conjugacy classes of G.

Proofc_1, \cdots , c_m are linearly independent over k because if i \neq j, then \mathcal{C}_i \cap \mathcal{C}_j = \emptyset. So we only need to show that \text{span} \{c_1, \cdots , c_m \} = Z(k[G]). Let x = \sum_{g \in G} x_g g \in k[G],  where x_g \in k for all g \in G.

Claim. x \in Z(k[G]) if and only if x_g = x_{h^{-1}gh} for all g,h \in G.

Proof  of the claim. Well, x \in Z(k[G]) if and only if xh = hx for all h \in G. Thus, x \in Z(k[G]) if and only if

\displaystyle \sum_{g \in G} x_g g = \sum_{g \in G} x_g h g h^{-1}. \ \ \ \ \ \ \ \ (1)

Clearly, fixing h \in G, we have \{hgh^{-1}: \ g \in G \} =G. So if we put hgh^{-1}=u, then g = h^{-1}uh and thus (1) becomes

\displaystyle \sum_{g \in G} x_g g = \sum_{g \in G} x_{h^{-1}gh} g. \ \ \ \ \ \ \ \ \ \ \ (2)

The claim now follows from (2).

Now, if 1 \leq i \leq m, then, by definition, g \in \mathcal{C}_i if and only if h^{-1}gh \in \mathcal{C}_i for all h \in G. So for every g,h \in G, the coefficients of g and h^{-1}gh in c_i are either both 1 or both 0. Thus, by the claim, c_i \in Z(k[G]) for all i. So \text{span} \{c_1, \cdots , c_m \} \subseteq Z(k[G]).

Conversely, if x = \sum_{g \in G} x_g g \in Z(k[G]), then, by the claim, x_g = x_{g'} for all g and g' which are in the same conjugacy class. Thus if for every i we choose a g_i \in \mathcal{C}_i, then

\displaystyle x = \sum_{i=1}^m x_{g_i} c_i.

So Z(k[G]) \subseteq \text{span} \{c_1, \cdots , c_m \}. \ \Box

For a ring R, let J(R) be the Jacobson radical of R. Recall that R is called semiprimitive if J(R)=(0). Let \text{char}(k) denote the characteristic of a field k. Maschke’s theorem characterizes semiprimitive group algebras k[G] for finite groups G.

Theorem (Maschke, 1899) Let k be a field and let G be a finite group of order n. Let R:=k[G]. Then R is semiprimitive if and only if \text{char}(k) \nmid n. 

Proof. Let G = \{g_1, \cdots , g_n \} where g_1=1. Suppose first that \text{char}(k) \nmid n and consider the algebra homomorphism \rho : R \longrightarrow \text{End}_k(R) defined by \rho(r)(x)=rx for all r,x \in R. Define \alpha : R \longrightarrow k by \alpha(r) = \text{tr}(\rho(r)), \ r \in R, where \text{tr}(\rho(r)) is the trace of the matrix corresponding to the linear map \rho(r) with respect to the ordered basis \{g_1, \cdots , g_n \}. Let’s make a few points about \alpha.

1) \alpha(1)=n because \rho(1) is the identity map of R.

2) If 1 \neq g \in G, then \alpha(g)=0. The reason is that \rho(g)(g_i)=gg_i \neq g_i for all i and thus the diagonal entries of the matrix of \rho(g) are all zero and so \alpha(g)=\text{tr}(\rho(g))=0.

3) If r \in R is nilpotent, then \alpha(r)=0 because then r^m=0 for some m hence (\rho(r))^m = \rho(r^m)=0. So \rho(r) is nilpotent and we know that the trace of a nilpotent matrix is zero. 

Now let r \in J(R). Since R is finite dimensional over k, it is Artinian and hence r is nilpotent. Thus, by 3), \alpha(r)=0. Let r= \sum_{i=1}^n c_i g_i, where c_i \in k. Then

0=\alpha(r)=\sum_{i=1}^n c_i \alpha(g_i)=c_1n,

by 1) and 2). It follows that c_1=0 because \text{char}(k) \nmid n. So the coefficient of g_1=1 of every element in J(R) is zero. But for every i, the coefficient of g_1=1 of the element g_i^{-1}r \in J(R) is c_i and so we must have c_i=0. Hence r = 0 and so J(R)=(0).

Conversely, suppose that \text{char}(k) \mid n and put x = \sum_{i=1}^n g_i. Clearly xg_j = x for all j and hence

x^2=x\sum_{j=1}^n g_j = \sum_{j=1}^n xg_j=\sum_{j=1}^n x = nx = 0.

Thus kx is a nilpotent ideal of R and so kx \subseteq J(R). Therefore J(R) \ne (0) because kx \ne (0). \ \Box 

Let k be a field or \mathbb{Z} and let G_1,G_2 be two groups. It is clear that if G_1 \cong G_2, then k[G_1] \cong k[G_2], as k algebras. The group ring isomorphism problem is this question that whether or not k[G_1] \cong k[G_2], as k algebras, implies G_1 \cong G_2. Another version of the group ring isomorphism problem is this: given a group G_1 and a field k find all groups G_2 such that k[G_1] \cong k[G_2], as k algebras. These questions have been answered in special cases only. For example, an old result due to Perlis and Walker states that if G_1,G_2 are finite abelian groups and \mathbb{Q}[G_1] \cong \mathbb{Q}[G_2], as \mathbb{Q} algebras, then G_1 \cong G_2. In 2001 Hertweek found two non-isomorphic groups G_1, G_2 of order 2^{21}97^{28} such that \mathbb{Z}[G_1] \cong \mathbb{Z}[G_2].

We now show that it is quite possible to have \mathbb{C}[G_1] \cong \mathbb{C}[G_2] and G_1 \ncong G_2:

Theorem. If G_1, G_2 are two finite abelian groups of order n, then \mathbb{C}[G_1] \cong \mathbb{C}[G_2] \cong \mathbb{C}^n as \mathbb{C}-algebras.

Proof. We have already seen in this post that J(\mathbb{C}[G])=(0) for any group G. So if G is a finite abelian group of order n, then \mathbb{C}[G] is a commutative semisimple algebra. Thus, since \mathbb{C} is algebraically closed, the Arrin-Wedderburn theorem gives \mathbb{C}[G] \cong \mathbb{C}^n as \mathbb{C}-algebras. \Box

Example 1. Let G_1 be the Klein four-group and let G_2 be the cyclic group of order four. Then G_1 \ncong G_2 but, by the theorem, \mathbb{C}[G_1] \cong \mathbb{C}[G_2] \cong \mathbb{C}^4 as \mathbb{C}-algebras.

Example 2. Let D_8, Q_8 be, respectively, the dihedral group of order 8 and the quaternion group. Then \mathbb{C}[D_8] \cong \mathbb{C}[Q_8] but D_8 \ncong Q_8 (see the Remark in this post!).

Notation. For a ring R let J(R) be the Jacobson radical of R.

Definition. Recall that if k is a field and G is a group, then the group algebra k[G] has two structures. Firstly, as a vector space over k, it has G as a basis, i.e. every element of k[G] is uniquely written as \sum_{g \in G} x_g g, where x_g \in k. In particular, \dim_k k[G]=|G|, as cardinal numbers. Secondly, multiplication is also defined in k[G]. If x = \sum_{g \in G} x_g g and y = \sum_{g \in G} y_g g are two elements of k[G], then we just multiply xy in the ordinary fashion using distribution law. To be more precise, we define xy = \sum_{g \in G} z_g g, where z_g = \sum_{rs=g} x_r y_s.

We are going to prove that J(\mathbb{C}[G])=0, for every group G.

Lemma. J(\mathbb{C}[G]) is nil, i.e. every element of J(\mathbb{C}[G]) is nilpotent.

Proof. If G is countable, we are done by this theorem. For the general case, let \alpha \in J(\mathbb{C}[G]). So \alpha =\sum_{i=1}^n c_ig_i, for some c_i \in \mathbb{C}, \ g_i \in G. Let H=\langle g_1,g_2, \cdots , g_n \rangle. Clearly \alpha \in H and H is countable. So to complete the proof, we only need to show that \alpha \in J(\mathbb{C}[H]). Write G = \bigcup_i x_iH, where x_iH are the distinct cosets of H in G. Then \mathbb{C}[G]=\bigoplus_i x_i \mathbb{C}[H], which means \mathbb{C}[G]=\mathbb{C}[H] \oplus K, for some right \mathbb{C}[H] module K. Now let \beta \in \mathbb{C}[H]. Since  \alpha \in J(\mathbb{C}[G]), there exists \gamma \in \mathbb{C}[G] such that \gamma (1 - \beta \alpha ) = 1. We also have \gamma = \gamma_1 + \gamma_2, for some \gamma_1 \in \mathbb{C}[H], \ \gamma_2 \in K. That now gives us \gamma_1(1 - \beta \alpha)=1. \ \Box

Theorem. J(\mathbb{C}[G])=0, for any group G.

Proof. For any x =\sum_{i=1}^n c_i g_i\in \mathbb{C}[G] define

x^* = \sum_{i=1}^n \overline{c_i} g_i^{-1}.

It’s easy to see that xx^*=0 if and only if x=0 and for all x,y \in \mathbb{C}[G]: \ (xy)^*=y^*x^*. Now suppose that J(\mathbb{C}[G]) \neq 0 and let 0 \neq \alpha \in J(\mathbb{C}[G]). Put \beta = \alpha \alpha^* \in J(\mathbb{C}[G]). By what I just mentioned \beta \neq 0 and (\beta^m)^* = (\beta^*)^m=\beta^m, for all positive integers m. By the lemma, there exists k \geq 2 such that \beta^k = 0 and \beta^{k-1} \neq 0. Thus \beta^{k-1} (\beta^{k-1})^* = \beta^{2k-2} = 0, which implies that \beta^{k-1} = 0. Contradiction! \Box

Corollary. If G is finite, then \mathbb{C}[G] is semisimple.

Proof. We just proved that J(\mathbb{C}[G])=(0). So we just need to show that \mathbb{C}[G] is Artinian. Let

I_1 \supset I_2 \supset \cdots \ \ \ \ \ \ \ \ \ \ \ \ \ (*)

be a descending chain of left ideals of \mathbb{C}[G]. Obviously each I_j is a \mathbb{C}-subspace of \mathbb{C}[G]. Thus each I_j is finite dimensional because \dim_{\mathbb{C}} \mathbb{C}[G]=|G| < \infty. Hence (*) will stablize at some point because \dim_{\mathbb{C}} I_1 < \infty and \dim_{\mathbb{C}}I_1 > \dim_{\mathbb{C}} I_2 > \cdots . Thus \mathbb{C}[G] is (left) Artinian and the proof is complete because we know a ring is semisimple if and only if it is (left) Artinian and its Jacobson radical is zero. \Box