Maschke's theorem

Concerns the decomposition of representations of a finite group into irreducible pieces

In mathematics, Maschke's theorem,[1][2] named after Heinrich Maschke,[3] is a theorem in group representation theory that concerns the decomposition of representations of a finite group into irreducible pieces. Maschke's theorem allows one to make general conclusions about representations of a finite group G without actually computing them. It reduces the task of classifying all representations to a more manageable task of classifying irreducible representations, since when the theorem applies, any representation is a direct sum of irreducible pieces (constituents). Moreover, it follows from the Jordan–Hölder theorem that, while the decomposition into a direct sum of irreducible subrepresentations may not be unique, the irreducible pieces have well-defined multiplicities. In particular, a representation of a finite group over a field of characteristic zero is determined up to isomorphism by its character.

Formulations

Maschke's theorem addresses the question: when is a general (finite-dimensional) representation built from irreducible subrepresentations using the direct sum operation? This question (and its answer) are formulated differently for different perspectives on group representation theory.

Group-theoretic

Maschke's theorem is commonly formulated as a corollary to the following result:

Theorem —  V {\displaystyle V} is a representation of a finite group G {\displaystyle G} over a field F {\displaystyle \mathbb {F} } with characteristic not dividing the order of G {\displaystyle G} . If V {\displaystyle V} has a subrepresentation W {\displaystyle W} , then it has another subrepresentation U {\displaystyle U} such that V = W U {\displaystyle V=W\oplus U} .[4][5]

Then the corollary is

Corollary (Maschke's theorem) — Every representation of a finite group G {\displaystyle G} over a field F {\displaystyle \mathbb {F} } with characteristic not dividing the order of G {\displaystyle G} is a direct sum of irreducible representations.[6][7]

The vector space of complex-valued class functions of a group G {\displaystyle G} has a natural G {\displaystyle G} -invariant inner product structure, described in the article Schur orthogonality relations. Maschke's theorem was originally proved for the case of representations over C {\displaystyle \mathbb {C} } by constructing U {\displaystyle U} as the orthogonal complement of W {\displaystyle W} under this inner product.

Module-theoretic

One of the approaches to representations of finite groups is through module theory. Representations of a group G {\displaystyle G} are replaced by modules over its group algebra  K [ G ] {\displaystyle K[G]} (to be precise, there is an isomorphism of categories between K [ G ] -Mod {\displaystyle K[G]{\text{-Mod}}} and Rep G {\displaystyle \operatorname {Rep} _{G}} , the category of representations of G {\displaystyle G} ). Irreducible representations correspond to simple modules. In the module-theoretic language, Maschke's theorem asks: is an arbitrary module semisimple? In this context, the theorem can be reformulated as follows:

Maschke's Theorem — Let G {\displaystyle G} be a finite group and K {\displaystyle K} a field whose characteristic does not divide the order of G {\displaystyle G} . Then K [ G ] {\displaystyle K[G]} , the group algebra of G {\displaystyle G} , is semisimple.[8][9]

The importance of this result stems from the well developed theory of semisimple rings, in particular, their classification as given by the Wedderburn–Artin theorem. When K {\displaystyle K} is the field of complex numbers, this shows that the algebra K [ G ] {\displaystyle K[G]} is a product of several copies of complex matrix algebras, one for each irreducible representation.[10] If the field K {\displaystyle K} has characteristic zero, but is not algebraically closed, for example if K {\displaystyle K} is the field of real or rational numbers, then a somewhat more complicated statement holds: the group algebra K [ G ] {\displaystyle K[G]} is a product of matrix algebras over division rings over K {\displaystyle K} . The summands correspond to irreducible representations of G {\displaystyle G} over K {\displaystyle K} .[11]

Category-theoretic

Reformulated in the language of semi-simple categories, Maschke's theorem states

Maschke's theorem — If G is a group and F is a field with characteristic not dividing the order of G, then the category of representations of G over F is semi-simple.

Proofs

Group-theoretic

Let U be a subspace of V complement of W. Let p 0 : V W {\displaystyle p_{0}:V\to W} be the projection function, i.e., p 0 ( w + u ) = w {\displaystyle p_{0}(w+u)=w} for any u U , w W {\displaystyle u\in U,w\in W} .

Define p ( x ) = 1 # G g G g p 0 g 1 ( x ) {\textstyle p(x)={\frac {1}{\#G}}\sum _{g\in G}g\cdot p_{0}\cdot g^{-1}(x)} , where g p 0 g 1 {\displaystyle g\cdot p_{0}\cdot g^{-1}} is an abbreviation of ρ W g p 0 ρ V g 1 {\displaystyle \rho _{W}{g}\cdot p_{0}\cdot \rho _{V}{g^{-1}}} , with ρ W g , ρ V g 1 {\displaystyle \rho _{W}{g},\rho _{V}{g^{-1}}} being the representation of G on W and V. Then, ker p {\displaystyle \ker p} is preserved by G under representation ρ V {\displaystyle \rho _{V}} : for any w ker p , h G {\displaystyle w'\in \ker p,h\in G} ,

p ( h w ) = h h 1 1 # G g G g p 0 g 1 ( h w ) = h 1 # G g G ( h 1 g ) p 0 ( g 1 h ) w = h 1 # G g G g p 0 g 1 w = h p ( w ) = 0 {\displaystyle {\begin{aligned}p(hw')&=h\cdot h^{-1}{\frac {1}{\#G}}\sum _{g\in G}g\cdot p_{0}\cdot g^{-1}(hw')\\&=h\cdot {\frac {1}{\#G}}\sum _{g\in G}(h^{-1}\cdot g)\cdot p_{0}\cdot (g^{-1}h)w'\\&=h\cdot {\frac {1}{\#G}}\sum _{g\in G}g\cdot p_{0}\cdot g^{-1}w'\\&=h\cdot p(w')\\&=0\end{aligned}}}

so w ker p {\displaystyle w'\in \ker p} implies that h w ker p {\displaystyle hw'\in \ker p} . So the restriction of ρ V {\displaystyle \rho _{V}} on ker p {\displaystyle \ker p} is also a representation.

By the definition of p {\displaystyle p} , for any w W {\displaystyle w\in W} , p ( w ) = w {\displaystyle p(w)=w} , so W ker   p = { 0 } {\displaystyle W\cap \ker \ p=\{0\}} , and for any v V {\displaystyle v\in V} , p ( p ( v ) ) = p ( v ) {\displaystyle p(p(v))=p(v)} . Thus, p ( v p ( v ) ) = 0 {\displaystyle p(v-p(v))=0} , and v p ( v ) ker p {\displaystyle v-p(v)\in \ker p} . Therefore, V = W ker p {\displaystyle V=W\oplus \ker p} .

Module-theoretic

Let V be a K[G]-submodule. We will prove that V is a direct summand. Let π be any K-linear projection of K[G] onto V. Consider the map

{ φ : K [ G ] V φ : x 1 # G s G s π ( s 1 x ) {\displaystyle {\begin{cases}\varphi :K[G]\to V\\\varphi :x\mapsto {\frac {1}{\#G}}\sum _{s\in G}s\cdot \pi (s^{-1}\cdot x)\end{cases}}}

Then φ is again a projection: it is clearly K-linear, maps K[G] to V, and induces the identity on V (therefore, maps K[G] onto V). Moreover we have

φ ( t x ) = 1 # G s G s π ( s 1 t x ) = 1 # G u G t u π ( u 1 x ) = t φ ( x ) , {\displaystyle {\begin{aligned}\varphi (t\cdot x)&={\frac {1}{\#G}}\sum _{s\in G}s\cdot \pi (s^{-1}\cdot t\cdot x)\\&={\frac {1}{\#G}}\sum _{u\in G}t\cdot u\cdot \pi (u^{-1}\cdot x)\\&=t\cdot \varphi (x),\end{aligned}}}

so φ is in fact K[G]-linear. By the splitting lemma, K [ G ] = V ker φ {\displaystyle K[G]=V\oplus \ker \varphi } . This proves that every submodule is a direct summand, that is, K[G] is semisimple.

Converse statement

The above proof depends on the fact that #G is invertible in K. This might lead one to ask if the converse of Maschke's theorem also holds: if the characteristic of K divides the order of G, does it follow that K[G] is not semisimple? The answer is yes.[12]

Proof. For x = λ g g K [ G ] {\textstyle x=\sum \lambda _{g}g\in K[G]} define ϵ ( x ) = λ g {\textstyle \epsilon (x)=\sum \lambda _{g}} . Let I = ker ϵ {\displaystyle I=\ker \epsilon } . Then I is a K[G]-submodule. We will prove that for every nontrivial submodule V of K[G], I V 0 {\displaystyle I\cap V\neq 0} . Let V be given, and let v = μ g g {\textstyle v=\sum \mu _{g}g} be any nonzero element of V. If ϵ ( v ) = 0 {\displaystyle \epsilon (v)=0} , the claim is immediate. Otherwise, let s = 1 g {\textstyle s=\sum 1g} . Then ϵ ( s ) = # G 1 = 0 {\displaystyle \epsilon (s)=\#G\cdot 1=0} so s I {\displaystyle s\in I} and

s v = ( 1 g ) ( μ g g ) = ϵ ( v ) g = ϵ ( v ) s {\displaystyle sv=\left(\sum 1g\right)\!\left(\sum \mu _{g}g\right)=\sum \epsilon (v)g=\epsilon (v)s}

so that s v {\displaystyle sv} is a nonzero element of both I and V. This proves V is not a direct complement of I for all V, so K[G] is not semisimple.

Non-examples

The theorem can not apply to the case where G is infinite, or when the field K has characteristics dividing #G. For example,

  • Consider the infinite group Z {\displaystyle \mathbb {Z} } and the representation ρ : Z G L 2 ( C ) {\displaystyle \rho :\mathbb {Z} \to \mathrm {GL} _{2}(\mathbb {C} )} defined by ρ ( n ) = [ 1 1 0 1 ] n = [ 1 n 0 1 ] {\displaystyle \rho (n)={\begin{bmatrix}1&1\\0&1\end{bmatrix}}^{n}={\begin{bmatrix}1&n\\0&1\end{bmatrix}}} . Let W = C [ 1 0 ] {\displaystyle W=\mathbb {C} \cdot {\begin{bmatrix}1\\0\end{bmatrix}}} , a 1-dimensional subspace of C 2 {\displaystyle \mathbb {C} ^{2}} spanned by [ 1 0 ] {\displaystyle {\begin{bmatrix}1\\0\end{bmatrix}}} . Then the restriction of ρ {\displaystyle \rho } on W is a trivial subrepresentation of Z {\displaystyle \mathbb {Z} } . However, there's no U such that both W, U are subrepresentations of Z {\displaystyle \mathbb {Z} } and C 2 = W U {\displaystyle \mathbb {C} ^{2}=W\oplus U} : any such U needs to be 1-dimensional, but any 1-dimensional subspace preserved by ρ {\displaystyle \rho } has to be spanned by an eigenvector for [ 1 1 0 1 ] {\displaystyle {\begin{bmatrix}1&1\\0&1\end{bmatrix}}} , and the only eigenvector for that is [ 1 0 ] {\displaystyle {\begin{bmatrix}1\\0\end{bmatrix}}} .
  • Consider a prime p, and the group Z / p Z {\displaystyle \mathbb {Z} /p\mathbb {Z} } , field K = F p {\displaystyle K=\mathbb {F} _{p}} , and the representation ρ : Z / p Z G L 2 ( F p ) {\displaystyle \rho :\mathbb {Z} /p\mathbb {Z} \to \mathrm {GL} _{2}(\mathbb {F} _{p})} defined by ρ ( n ) = [ 1 n 0 1 ] {\displaystyle \rho (n)={\begin{bmatrix}1&n\\0&1\end{bmatrix}}} . Simple calculations show that there is only one eigenvector for [ 1 1 0 1 ] {\displaystyle {\begin{bmatrix}1&1\\0&1\end{bmatrix}}} here, so by the same argument, the 1-dimensional subrepresentation of Z / p Z {\displaystyle \mathbb {Z} /p\mathbb {Z} } is unique, and Z / p Z {\displaystyle \mathbb {Z} /p\mathbb {Z} } cannot be decomposed into the direct sum of two 1-dimensional subrepresentations.

Notes

  1. ^ Maschke, Heinrich (1898-07-22). "Ueber den arithmetischen Charakter der Coefficienten der Substitutionen endlicher linearer Substitutionsgruppen" [On the arithmetical character of the coefficients of the substitutions of finite linear substitution groups]. Math. Ann. (in German). 50 (4): 492–498. doi:10.1007/BF01444297. JFM 29.0114.03. MR 1511011.
  2. ^ Maschke, Heinrich (1899-07-27). "Beweis des Satzes, dass diejenigen endlichen linearen Substitutionsgruppen, in welchen einige durchgehends verschwindende Coefficienten auftreten, intransitiv sind" [Proof of the theorem that those finite linear substitution groups, in which some everywhere vanishing coefficients appear, are intransitive]. Math. Ann. (in German). 52 (2–3): 363–368. doi:10.1007/BF01476165. JFM 30.0131.01. MR 1511061.
  3. ^ O'Connor, John J.; Robertson, Edmund F., "Heinrich Maschke", MacTutor History of Mathematics Archive, University of St Andrews
  4. ^ Fulton & Harris 1991, Proposition 1.5.
  5. ^ Serre 1977, Theorem 1.
  6. ^ Fulton & Harris 1991, Corollary 1.6.
  7. ^ Serre 1977, Theorem 2.
  8. ^ It follows that every module over K [ G ] {\displaystyle K[G]} is a semisimple module.
  9. ^ The converse statement also holds: if the characteristic of the field divides the order of the group (the modular case), then the group algebra is not semisimple.
  10. ^ The number of the summands can be computed, and turns out to be equal to the number of the conjugacy classes of the group.
  11. ^ One must be careful, since a representation may decompose differently over different fields: a representation may be irreducible over the real numbers but not over the complex numbers.
  12. ^ Serre 1977, Exercise 6.1.

References