arXiv:2103.15026v2 [math.RT] 21 Jun 2021 nee;Permutation ; set nmteais(e [7]). (see mathematics in aspects rbe nivratmti algebras. invariant on problem o n field any for n nld h ulne ler of algebra Auslander the include and hn n[] Also, (for others [8]. many in by investigated Zhang are They [15]). (see theory lerial lsdfil then central field as closed studied p algebraically are engineering algebras matrix [14]), invariant (see recently, processes radicals, Markov in their on applications acting subgroups parabolic of tive lasaFoeisetnin u o o lerial clo algebraically not for But extension. Frobenius a always where Let Introduction 1 in on tions R S Md(requivalently (or -Mod mdls ealta w rings two that Recall -modules. ∗ ewrs lmnaysmercplnma;Gets comm Greatest ; symmetric Elementary 16 16W22, Keywords: Primary Classification: Subject Mathematics 2020 hsnt a w upss h rtoei oso that show to is one first The purposes. two has note This If ealta nextension an that Recall h td fivratagba n ig a oghsoya history long a has rings and algebras invariant of study The lal,if Clearly, o subgroup a For Let nain lerso arcsadsmercplnmaso symmetric and matrices of algebras Invariant R S orsodn uhrsEal iccueuc;Fx 0086 Fax: [email protected]; Email: author’s Corresponding mdl and -module C rbnu xeso of extension Frobenius R h smrhs rbe fivratmti lers Supp algebras. matrix invariant of problem the that ymti groups symmetric vra lerial lsdfield closed algebraically an over tnsfrtegets omndvsrof divisor common greatest the for stands stepriinplnma of polynomial partition the is eauiayrn.W eoeby denote We ring. unitary a be ( let , i n ⊆ ) [ n σ o field a For eantrlnme,w set we number, natural a be p ] M by , tnsfrteiaeof image the for stands ∤ S n R λ n ( ( c i R c and σ M and , ) R : h etaie ler of algebra centralizer the , n = ) ( R G R c p R etecnrlzragbaof algebra centralizer the be e R G ) ∤ 1 ∈ S S of ( µ h igof ring the ivratmti lersgnrls etoymti ma centrosymmetric generalise algebras matrix -invariant fcharacteristic of n 1 ≃ Σ R M ( ) j Σ n σ G mdand -mod o 1 for Hom n n + and , ( the , S R R S ··· n ) ) ( ⊆ ≤ Σ h eodoei oivsiaeMrt qiaecsadth and equivalences Morita investigate to is one second The . c : S S = , + m ( S G R R n i S ihccetypes cycle with , n ( λ n R ) ≤ - { e R ( σ frnsi alda called is rings of nain arxring matrix invariant ute,w nrdc e qiaec eaino partit on relation equivalence new a introduce we Further, . i hyaeMrt qiaeti n nyif only and if equivalent Morita are They . C n ( , × S , , hncagXi Changchang a s and under S ( md r equivalent. are -mod) , R 1 , R n ) j i R n R [ ) n ) [ σ p ) as ) X r smrhci n nyif only and if isomorphic are arcsover matrices ] ≤ sawy ellragbaadteextension the and algebra cellular a always is ∈ S stepruainmti of matrix permutation the is ≥ : : ] R = = R / r adt be to said are M σ j Md(or -Mod ( n matrix a and 0 - ≤ C { X S If . { λ n a bmdls hs xesoshv nrgigitrs nm in interest intriguing have extensions These -bimodules. 1 ( n 1 t in , R ) , epoeta h nain algebras invariant the that prove We . ∈ 2 ··· G c ) se[7) hc ly rca oei eciigteorbi the describing in role crucial a plays which [17]), (see , M M | Abstract ··· in = λ , n a n λ ( ( = j i ( M h ∗ s R , R rbnu extension Frobenius R 1 . σ R n = n ib Zhang Jinbi and ) n md h aeoyo l o ntl eeae)left generated) finitely (or all of category the -mod) 5,0A7 eodr 53,1C8 16K99 11C08, 15A30, Secondary 05A17; S50, 68903637. 10 ) i } λ ( n by and ndvsr nain arxagba smrhs;Partit Isomorphism; algebra; matrix Invariant divisor; on sdfie as defined is R over | 1 oiaequivalent Morita a for zrrnsi 1] ti hw hr htif that there shown is It [17]. in rings izer n eoeby denote and , ca ) ( c ··· eso that show We . i e ed,teerslsaesilt eunderstood. be to still are results these fields, sed hl etoymti arcshv significant have matrices centrosymmetric while ) S σ ntefull the in σ olm n unu hsc se[].More [4]). (see physics quantum and roblems n xml,se[01],adrcnl,b Lehrer- by recently, and [10–12]), see example, = , , R ( ( λ ∈ c j ) ac e s fdegree of , σ s that ose ) j i R Σ , , and σ n ) h arxuisof units matrix the m ∀ dcnb akt h lsia invariant classical the to back be can nd ebifl write briefly we , ⊆ ∈ c = σ µ M G ∈ n then , n ( = Σ σ n , × C ( 1 and n n if R and } n d M µ ≤ h ymti ru fpermuta- of group symmetric the . sdfie by defined is ) ftetocategories two the if λ 1 S arxalgebra matrix n S , sawy rbnu extension Frobenius a always is R ε ε i ( ··· , λ n λ τ R safiieygnrtdprojec- generated finitely a is ( ( ( j ) x r emttosi the in permutations are σ 1 ≤ , = ) rxagba se[4 16]) [14, (see algebras trix = ) µ , sawy separable a always is S R t n ) n = ) epciey such respectively, , ( } S ε d σ M , τ n µ , ( ( ε R n S x σ S µ partitions f ( ) osadstudy and ions n ) ( n R where , , ( 1 and ( R σ M ) ) c . ) where , , σ n isomorphism e R for , ( S R R ) m ) ) ⊆ R ( ε S . τ over λ Mdand -Mod n , M ( ( R d x G λ R ) ) n ( , R R o of ion san is any ) ) . ts is Let R be a field and σ,τ Σn. What are the necessary and sufficient conditions for Sn(σ,R) and Sn(τ,R) to be isomorphic? ∈ To state our results specifically, we first introduce a few notions and terminologies. We write σ Σn as a ∈ product of disjoint cycles, say σ = σ1σ2 σs, where σi is a λi-cycle with λi 1. This gives a decomposition ··· ≥ of n. By assuming λ1 λs 1, we get a partition λ = (λ1, ,λs) of n with s parts. This λ is called the ≥···≥ ≥ ··· cycle type of σ. We then associate λ with the so-called partition polynomial ε (x) Z[x], which is defined by λ ∈ elementary symmetric polynomials gi(λ) in λ1, ,λs, over the canonical bisemigroup of natural numbers ··· (see Definition 3.8). Here, the polynomial gi(λ) is closely related to a matrix norm of the triangular divisor matrix of λ, and the multiplication in the bisemigroup is given by taking the greatest common divisors of (see Section 3.2 for details).

Theorem 1.1. If R is a field and c Mn(R), then the extension Sn(c,R) Mn(R) is a separable Frobenius extension. ∈ ⊆ To the isomorphism problem, we have the following answer.

Theorem 1.2. Let R be a field of characteristic p 0, let σ Σn and τ Σm be of cycle types λ = ≥ ∈ ∈ (λ1,λ2, ,λs) and µ = (µ1,µ2, ,µt ), respectively, and let d denote the greatest common divisor of these ··· ··· λ λi’s. Then the following hold. (1) Sn(σ,R) is semisimple if and only if p ∤ λi for 1 i s. ≤ ≤ (2) Assume that R is algebraically closed, p ∤ λi for all 1 i s and p ∤ µ j for all 1 j t. Then ≤ ≤ ≤ ≤ (i) Sn(σ,R) Sm(τ,R) if and only if m = n and ε (x)= εµ(x), where ε (x) Z[x] is the partition polyno- ≃ λ λ ∈ mial of λ. (ii) Sn(σ,R) and Sm(τ,R) are Morita equivalent if and only if dλελ(1)= dµεµ(1). Thus, the isomorphism problem for semisimple, invariant matrix algebras over an algebraically closed field can be read off from the numerical values of the elementary symmetric polynomials evaluated at the corresponding partitions (see Section 3.2.1 for more details). This article is outlined as follows. Section 2 is devoted to some basic facts on invariant algebras of ma- trices, while Section 3.2 provides proofs of the two theorems. During the course of the proofs, we introduce a so-called polynomial equivalence relation and triangular divisor matrices of partitions based on elementary symmetric polynomials of partitions. In Section 4, we investigate real invariant algebras of matrices as an illustrating example. At the end of this section, we formulate a few unsolved problems suggested by the main results.

2 Preliminaries

In this section, we recall basic notions and establish primary facts on invariant matrix algebras. Throughout the paper, R denotes a unitary ring.

2.1 Invariant matrix algebras

Let G be a subgroup of Σn, the symmetric group of all permutations of the set [n] := 1,2, ,n , and σ Σn. { ··· } ∈ Then G acts on [n] and therefore on Mn(R) by

Mn(R) G Mn(R), (ai j) g := (a ), (ai j) Mn(R),g G. × −→ · (i)g,( j)g ∈ ∈ Recall that the G-invariant matrix algebra, denoted Sn(G,R), is the G-fixed points of this action in Mn(R):

G Sn(G,R) := (Mn(R)) := a Mn(R) ai j = a , σ G,i, j [n] = Sn(σ,R), { ∈ | (i)σ,( j)σ ∀ ∈ ∈ } σ\G ∈ where Sn(σ,R) is called the σ-invariant matrix ring over R of degree n.

2 For σ Σn, let cn(σ) := e + e + + e be the in Mn(R), related to σ. ∈ 1,(1)σ 2,(2)σ ··· n,(n)σ If the subscript n is clear from the context, we simply write cσ for cn(σ). In Mn(R), we have cσcτ = cστ, 1 c′ = c 1 = c− and (cσ)i j = δσ i j for i, j [n], where a′ is the of the matrix a and δi j is the σ σ− σ ( ), ∈ Kronecker symbol.

Lemma 2.1. (1) Let a = (ai j) Mn(R). Then a Sn(G,R) if and only if cσa = acσ for σ G. In particular, ∈ ∈ ∈ Z(R)[cσ] Z(Sn(σ,R)), where Z(R) denotes the center of R. ⊆ (2) Sn(G,R) is a subring of the matrix ring Mn(R). (3) The transpose of matrices is an anti-automorphism of order 2 of the ring Sn(G,R). r (4) If G is the subgroup of Σn generated by τ1, ,τr Σn, then Sn(G,R)= Sn(τi,R). ··· ∈ ∩i=1 Proof. (1) This is already noticed in [11]. For the convenience of the reader, we sketch here a proof. By definition, cσa = acσ if and only if (cσa)i j = (acσ)i j for all i, j [n]. But ∈ n n

(cσa)i j = (cσ)ipap j = δ(i)σ,pap j = a(i)σ, j, Xp=1 Xp=1 n n

(ac ) = a (c ) = a δ = a 1 , σ i j ip σ p j ip (p)σ, j i,( j)σ− Xp=1 Xp=1

(c ac 1 ) = a . σ σ− i j (i)σ,( j)σ

Thus cσa = acσ if and only if a i j = a 1 for all i, j if and only if ai j = a i j for all i, j [n]. ( )σ, i,( j)σ− ( )σ,( )σ ∈ The statements remained are clear by definition. 

By Lemma 2.1(1), we may identify Sn(σ,R) with Sn(cσ,R) := a Mn(R) acσ = cσa . The next lemma { ∈ | } shows that Sn(σ,R) depends only upon the cycle type λ of σ. So we may also write Sn(λ,R) for Sn(σ,R). The proof is similar to the one of [17, Lemma 2.1], we leave it to the reader.

Lemma 2.2. Let τ,σ Σn, and let d Mn(R) be an . Then 1 ∈ ∈ (1) Sn(σ− ,R)= Sn(σ,R). In particular, Sn(cσ,R)= Sn(cσ′ ,R). 1 (2) Sn(dcσd ,R) Sn(σ,R). − ≃ (3) Sn(στ,R) Sn(τσ,R). ≃

Let σ Σn be of the cyclic type (λ1, ,λs), G = σ , and Λ := (i, j) 1 i, j n . Then G acts on Λ ∈ ··· h i { | ≤ ≤ } by (i, j) σ = (i)σ,( j)σ . Let O denote the G-orbit of (i, j). Then a Sn(σ,R) if and only if a takes a · (i, j) ∈ constant value on each orbit O . Here, we regard a Mn(R) as a function from Λ to R. Now we describe (i, j) ∈ the number of G-orbits in Λ. We assume n 2 and consider the three cases. ≥ (i) σ = (12 n). In this case, G has order n and acts on Λ by (i, j)σ = (i + 1, j + 1). Thus the stabilizer ··· subgroup of G fixing (i, j) Λ is trivial and the G-orbit of (i, j) has G elements. Hence there are nG-orbits ∈ | | of Λ, they are O , 1 j n. (1, j) ≤ ≤ (ii) σ = (12 p)(p + 1, , p + q) with n = p + q. In this case, G = [p,q], the least common multiple ··· ··· | | of p and q. Let X := (i, j) 1 i p, p + 1 j n with the G-action given by (i, j)σ = (i + 1, j + 1). { | ≤ ≤ ≤ ≤ } Let σ1 = (12 p) Σp and σ2 = (p + 1, , p + q) Σq, here we consider Σq as the symmetric group of ··· ∈ ··· ∈ permutations on p + 1, p + 2, , p + q . If ((i)σl ,( j)σl ) = (i, j) for a natural number 1 l [p,q], then { ··· } ≤ ≤ ((i)σl ,( j)σl ) = (i, j). This implies that p l and q l. Thus [p,q] l and l = [p,q]. Hence the stabilizer subgroup 1 2 | | | of G fixing (i, j) is trivial and the G-orbit of (i, j) has [p,q] elements. Therefore the number of G-orbits of X is pq/[p,q]. (iii) σ = σ1σ2 σs, a product of disjoint cycles, where σi is a cycle of length λi 1 with the context ··· ≥ Xi. Thus G = [λ1, ,λs] and Xi is a G-orbit for i [s]. Therefore Xi Xj = 0/ for 1 i = j s, and s | | ··· ∈ ∩ ≤ 6 ≤ Xj = 1,2, ,n . The set Λ can be identified with ∪ j=1 { ··· } ˙ Xi Xj. 1 i, j s × [ ≤ ≤ 3 For (x,y) Xi Xj, the G-orbit O of (x,y) depends only upon σi and σ j. Thus, if i = j, then we are in ∈ × (x,y) case (i) and the number of G-orbits is λi. If i = j, then we are in the case (ii) and the number of G-orbits of 6 Xi Xj is λiλ j/[λi,λ j]. Thus the number of G-orbits of Λ is × λiλ j ℓn(σ) := = λi  λ j, [λi,λ j] 1 Xi, j s 1 Xi, j s ≤ ≤ ≤ ≤ where λi  λ j stands for gcd(λi,λ j), the greatest common divisor of λi and λ j. We define C(λ) := (λi  λ j)s s Ms(N). This is called the greatest common divisor × ∈ matrix of λ in the literature. In this note, it will be called the matrix of Sn(σ,R). Further, let

fi := e j j, 1 i s, fi j := ep q, (i, j) Λ. ≤ ≤ , ∈ Xj Xi (p,qX) O ∈ ∈ (i, j) s Then f1, , fs are pairwise orthogonal idempotent elements in Sn(σ,R) and f j = 1 in Mn(R). This ··· j=1 yields the of Sn(σ,R) P

f1Sn(σ,R) f1 f1Sn(σ,R) f2 f1Sn(σ,R) fs ···  f2Sn(σ,R) f1 f2Sn(σ,R) f2 f2Sn(σ,R) fs Sn(σ,R)= . . ···. . .  . . .. .     fsSn(σ,R) f1 fsSn(σ,R) f2 fsSn(σ,R) fs  ···  The transpose of matrices is an R-involution on Sn(σ,R) fixing fi for all 1 i s. ≤ ≤ Lemma 2.3. (1) Sn(σ,R) is a free R-module of ℓn(σ). (2) The R-rank of fiSn(σ,R) f j is λi  λ j. (3) The dimension matrix C(λ) of Sn(σ,R) is positive definite if the numbers λi, 1 i s, are pairwise ≤ ≤ s s! distinct. In this case, the detC(λ) of C(λ) satisfies ϕ(λ1) ϕ(λs) detC(λ) λi , ··· ≤ ≤ i=1 − 2 where ϕ(λi) is the Euler’s totient function of λi. Q

Proof. (1) and (2) are consequences of the definition of Sn(σ,R) and the calculation of G-orbits of Λ (see also [12]). (3) is proved in [2, Theorem 2] and [9]. 

Example 2.4. (1) Sn((12 n),R) R[Cn], where R[Cn] is the group algebra of the Cn of order n n··· ≃ over R. Note that g = j=1 e j, j+1 is a generator of Cn, here we understand (n,n + 1) = (n,1). The transpose 1 of matrices in Sn((12 Pn),R) corresponds to the involution defined by g g− . ··· s 7→ (2) If λ = (a, ,a) = (a ) where a appears s times, then Ssa(λ,R) Ms(R[Ca])). In fact, each element of ··· ≃ Sna(λ,R) can be partitioned as an s s block-matrix such that each block has entries in fiSsa(λ,R) fi R[Ca]. × ≃ Thus Ssa(λ,R) Ms(R[Ca]). ≃ By Lemma 2.2(2), up to isomorphism of rings, we may assume σ = (12 λ1)(λ1 +1, ,λ1 +λ2) (λ1 + ··· ··· ··· + λs 1 + 1, ,λ1 + + λs) as a product of disjoint cycles. Let λ = (λ1, ,λs), X1 = 1, ,λ1 , ··· − i 1 ··· ··· i ··· { ··· } Xi = j − λ < j λi for 2 i s, di j := λi  λ j. Then { | ℓ=1 ℓ ≤ ℓ=1 } ≤ ≤ P PP 0 0 0 1 0 λ1 ··· ··· 0 P 0 . . . .  λ2  . . .. . cσ = . . ···. . where Pm :=  . . .. .  0 0 1     P ···  0 0 λs n n 1 0 0m m  ···  ×  ···  × For r0, ,rm 1 R, we denote by C(r0, ,rm 1) (or Cm(r0, ,rm 1)) the circulant matrix ··· − ∈ ··· − ··· − r0 r1 r2 rm 1 ··· − rm 1 r0 r1 rm 2 − ··· − ...... m 1  . . . . .  = r0Im + r1Pm + + rm 1Pm − .   ··· −  .. ..   r2 . . r1   ···   r1 r2 rm 1 r0 m m  ··· −  × 4 Thus, for a cyclic group Cm of order m, we have R[Cm] C(r0, ,rm 1) r0, ,rm 1 R . ≃ { ··· − | ··· − ∈ } Recall that, given σ Σn and τ Σm, an n m matrix a = (ai j) Mn m(R) is called a (σ,τ)-invariant ∈ ∈ × ∈ × matrix if cσa = acτ.

We see easily that an s s a = (Ai j) with Ai j Mλi λ j (R) lies in Sn(σ,R) if and only if × ∈ × Ai j is a (σi,σ j)-invariant matrix for each pair 1 i, j s. Thus a matrix a Sn(σ,R) can be written as a ≤ ≤ ∈ block-matrix form A11 A12 A1s ··· A21 A22 A2s a = (Ai j)= . . ···. .  . . .. .    As1 As2 Asss s  ···  × where the diagonal Aii is a λi λi circulant matrix and the off-diagonal Ai j is a λi λ j matrix over R, having × × the block form:

C(r1, ,rdij ) C(r1, ,rdij ) ···. ··· ···. Ai j =  . .  ··· λ C(r1, ,rdij ) C(r1, ,rdij ) λi j  ··· ··· ···  dij × dij for r1, ,rd R (see [12, p.260] for R = C). In general, we have an isomorphism of R-modules ··· ij ∈

fiSn(σ,R) f j Ai j MXi Xj (R) r1, ,rdij R . ≃ { ∈ × | ··· ∈ }

Lemma 2.5. (1) For 1 i s, fiSn(σ,R) fi R[C ] as rings. ≤ ≤ ≃ λi (2) If λ1 = λ2, then Sn(σ,R) f1 Sn(σ,R) f2 as left Sn(σ,R)-modules. ≃ (3) If m := n λi and f := f1 + + fi 1 + fi+1 + + fs, then f Sn(σ,R) f Sm(σ1 σi 1σi+1 σs,R) as rings. − ··· − ··· ≃ ··· − ···

Proof. (1) If λi = 1, then fiSn(σ,R) fi R and (1) holds. We assume λi 2 and define g := eλi 1+1,λi 1+2 + − − ≃ λi j ≥ + eλi 1+λi 1,λi 1+λi + eλi 1+λi,λi 1+1. Then g = fi and g = fi for all 1 j < λi. One can check that ··· 2− − λ −1 − − 6 ≤ fi,g,g , ,g i is an R-basis of fiSn(σ,R) fi. Thus fiSn(σ,R) fi R[C ]. { ··· − } ≃ λi (2) By definition, f1 = e11 + + e and f2 = e + e + + e . Let a0 := ··· λ1,λ1 λ1+1,λ1+1 λ1+2,λ1+2 ··· 2λ1,2λ1 e + e + + e . Then a0 Sn(σ,R) and f1a0 = a0 f2 = a0. So, we define a map 1,λ1+1 2,λ1+2 ··· λ1,2λ1 ∈

ψ : Sn(σ,R) f1 Sn(σ,R) f2, a f1 a f1a0 for a Sn(σ,R). −→ 7→ ∈

Note that ψ just moves the first λ1 columns of a f1 entirely to the second λ1 columns of a f1a0, and sends other columns of a f1 to zero. Thus ψ is a homomorphism of Sn(σ,R)-modules, and it is in fact an isomorphism of Sn(σ,R)-modules. (3) We may assume i = 2 since the argument below works for any i. If we delete the cycle σ2 from σ and consider τ := σ1σ3 σs as a permutation on Y := [n] X2 = X1 (X3 Xs), then Sm(σ1σ3 σs,R) is ··· \ ∪ ∪···∪ ··· well defined. Observe that σ and τ have the same action on Y and that a matrix a in f Sn(σ,R) f is of the form

A 0 U 0 0 0  V 0 B n n   × where A is a matrix indexed by X1 X1, U is a matrix indexed by X1 (X3 Xs), V is a matrix indexed × × ∪···∪ by (X3 Xs) X1 and B is a matrix indexed by (X3 Xs) (X3 Xs). Since a f Sn(σ,R) f , ∪···∪ × ∪···∪ × ∪···∪ ∈ the block matrix AU b := VB is in Sm(τ,R). Conversely, given a matrix b Sm(τ,R), we partition b into a block form according to Y , ∈ say the above form. Then we get an n n matrix by putting 0-blocks in the middle row and column, and × 5 1 0 0 this matrix is in f Sn(σ,R) f . Note that the identity in f Sn(σ,R) f is f = 0 0 0. In this way, we get 0 0 1 a one-to-one correspondence which preserves identity, addition and multiplication of matrices. This shows f Sn(σ,R) f Sm(τ,R).  ≃ 3 Proofs of main results

This section is devoted to proofs of Theorems 1.1 and 1.2. We first prove Theorem 1.1 and then prove Theorem 1.2 by introducing an equivalence relation on partitions.

3.1 The proof of Theorem 1.1 We starts with the following proposition.

Proposition 3.1. Let k K be an extension of fields with dimk(K) < ∞, and let B A be an extension of ⊆ ⊆ finite-dimensional algebras over k. Then B k K A k K is a Frobenius extension if and only B A is a Frobenius extension. ⊗ ⊆ ⊗ ⊆

To prove Proposition 3.1, we begin with a few lemmas.

Lemma 3.2. (1) Let k K be an extension of fields. Then it is a Frobenius extension if and only if dimk(K) is finite. ⊆ (2) Let A be an algebra over a field, and let M and N be finite-dimensional A-modules. The following are equivalent for M and N: (i) M N. ≃ (ii) There is an integer n 1 such that Mn Nn. ≥ ≃ Proof. (1) is trivial. (2) is not difficult, but for the convenience of the reader, we include here a proof. We s si t t j show that (ii) implies (i). Suppose M i=1 Xi and N j=1 Yi , where Xi and Yj are indecomposable ≃ ≃ n n modules with Xi Xp, Yj Yq for 1 i =Lp s and 1 j = qL t, si 1 and t j 1. It follows from M N 6≃ 6≃ ≤ 6 ≤ ≤ 6 ≤ ≥ ≥ ≃ that s = t and, up to a permutation of these Yj, that Xi Yi and si = ti for all i. Thus M N.  ≃ ≃ For k-algebras A and B, we may form the tensor product A k B of algebras over k. Given an A-module ⊗ AX and a B-module BY , the tensor product X k Y has a left (A k B)-module structure in a natural way, and ⊗ ⊗ the Hom-set Homk(X,Y ) is a B-A-bimodule defined by b f : x b(x f ), f a : x (ax) f for x X,a A,b B 7→ 7→ ∈ ∈ ∈ and f Homk(X,Y). ∈ The first statement (1) of the next lemma is known in [13, Lemma 4] and [5, Corollary 2.8], while the second statement (2) is in [3, Chap. XI, Theorem 3.1, p. 209-210]. In fact, (1) can be proved by applying (2).

Lemma 3.3. (1) Let B A be an extension of algebras over a field k and C be an algebra overk. IfB A ⊆ ⊆ is a Frobenius (or separable) extension, then the extension B k C A k C of tensor products of algebras is a Frobenius (or separable) extension. ⊗ ⊆ ⊗ (2) Let A,B,C,D,E,F be algebras over a field k, and let AX, AY be A-modules and BX ′, BY ′ B-modules.

Then HomA kB(X k X ′,Y k Y ′) HomA(X,X ′) k HomB(Y,Y ′). Moreover, if AXC and AYE are bimodules ⊗ ⊗ ⊗ ≃ ⊗ and if BX , BY ) are bimodules, then the isomorphism is a (C k D)-(E k F)-bimodule homomorphism. D′ F′ ⊗ ⊗ Lemma 3.4. If k K is an extension of fields and c Mn(k), then Sn(c,K) Sn(c,k) k K as k-algebras. ⊆ ∈ ≃ ⊗ Proof. Observe that for a Sn(c,k) and λ K we have aλ = λa Sn(c,K). If we choose a basis of the k- ∈ ∈ ∈ space K, say vi i I , where I is an index set (not necessarily finite), then any matrix a = (ai j) Mn(K) can { | ∈ } s (p) (p) ∈ be expressed uniquely as a = p J a vp, with J a subset of I and a Mn(k) for p J. This is due to the ∈ ∈ ∈ fact that viIn i I is a basisP of the left Mn(k)-module Mn(K). Hence a matrix a = (ai j) Mn(K) belongs { | ∈ } ∈

6 (i) to Sn(c,K) if and only if a Sn(c,k) for all i J. This yields that the restriction of the multiplication map ∈ ∈ µ : Mn(k) k K Mn(K) to Sn(c,k) k K gives rise to a surjective homomorphism µ : Sn(c,k) k K Sn(c,K) ⊗ → ⊗ ′ ⊗ → of k-algebras. Thus Lemma 3.4 follows from the commutative diagram

µ′ / Sn(c,k) k K Sn(c,K)  _⊗  _

  µ / Mn(k) k K Mn(K) ⊗ ∼ with µ clearly an isomorphism of k-algebras. 

Proof of Proposition 3.1. It follows from Lemma 3.3(1) that B k K A k K is a Frobenius extension if ⊗ ⊆ ⊗ B A is a Fribenius extension. Conversely, we postulate that B k K A k K is a Frobenius extension. Then ⊆ ⊗ ⊆ ⊗ B kK(A k K) is a finitely generated projective B k K-module, and HomB kK(A k K, B k K) A kK(A k ⊗ ⊗ ⊗ ⊗ ⊗ ⊗ ≃ ⊗ ⊗ K)B kK. First, we show that BA is finitely generated. Suppose that λi I is a k-basis of k-space K with λ0 = ⊗ { ∈ } 1. Then every element a in A K can be uniquely expressed as a λ with a A. Let x x k i Ia i i i 1, , m ⊗ ∈ ⊗ ∈ ··· ∈ A k K be generators of A k K over B k K and xi = j I ai j Pk λ j with ai j A, where Ii is an index set. ⊗ ⊗ ⊗ i m⊗ ∈ Pick an element a A and consider a 1. We write a 1 ∈ y x where y b λ B K P= j=1 j j, j = p Jj p p k . ∈ ⊗ ⊗ ∈ ⊗ ∈ ⊗ Then a 1 is a summation of bpai j λq. By the unique expressionP of elements in AP k K, we see that a can ⊗ ⊗ m ⊗ be generated over B by all ai j with 1 i m and j Iq. Thus BA is finitely generated. Since A k K is ≤ ≤ ∈∪q=1 ⊗ projective over B k K and B k K is projective over B, we see that A k K is projective over B. This implies ⊗ ⊗ ⊗ that BA is projective over B. Now it follows from Lemma 3.3(2) that HomB kK(BAA k K, BBB k K) ⊗ ⊗ ⊗ ≃ HomB(BA, BB) k K. Thus we have an isomorphism AAB k K AHomB(BAA, BB)B k K as (A k K)-(B k ⊗ ⊗ ≃ ⊗ ⊗ dim (K)⊗ K)-bimodules. Just considering the A-B-bimodule structure of this isomorphism, we then get (AAB) k dim (K) ≃ HomB(BAA, BB) k as A-B-bimodules. By Lemma 3.2(2), we must have AAB HomB(AA, BB) as A-B- ≃ bimodules. Thus B A is a Frobenius extension.  ⊆ Proof of Theorem 1.1. Suppose that R is any field and c Mn(R). Let K¯ be an algebraic closure of R. So all ∈ eigenvalues of the characteristic polynomial of c lie in R¯. We consider the smallest subfield K of k¯ containing R and all eigenvalues of c, that is, K is obtained from R by adding all eigenvalues. Thus K is a finite extension of R, that is, dimR(K) < ∞. We may assume that K is a splitting field of the characteristic polynomial of c. Thus c is similar to a Jordan-block matrix by a matrix in GLn(K). It follows from [17, Theorem 1.2(2)] that Sn(c,K) Mn(K) is a separable Frobenius extension. By Proposition 3.1 and Lemma 3.4, we deduce that ⊆ Sn(c,R) Mn(R) is a separable Frobenius extension.  ⊆ 3.2 Proof of Theorem 1.2 In this section we introduce an equivalence relation on the set of partitions of n with exactly s parts by employing elementary symmetric polynomials. We then study combinatorics of this relation by a matrix norm over the semiring of nonnegative integers. Finally, we give a proof of Theorem 1.2.

3.2.1 A new equivalence relation on partitions Recall that a nonempty set A together with two associative binary operations + and , named addition and · multiplication, respectively, is called a bisemigroup. A bisemigroup (A,+, ) is called a semiring if (A,+) is · commutative and the distributive laws hold: x (y + z)= x y + x z and (x + y) z = x z + y z for x,y,z A. · · · · · · ∈ A commutative semiring is a semiring such that the multiplication is commutative.

Definition 3.5. A bisemigroup (A,+,) is called a quarter-ring if (1) (A,+) and (A,) are commutative, (2) a  a = a for a A, and ∈ (3) (a + b)  a = b  a for a,b A. ∈

7 m In a bisemgroup (A,+, ), for any positive integer m and ai,b,d A, we write i=1 ai for a1 + + am, m · m ∈ ··· particularly, ma for a, and a· for a a a, the product of m copies of a. P i=1 · ····· P m  In a quarter-ring (A,+, ), the following| {z hold.} If a  b = d, then a  d = d. In particular, if ai  b = d for 1 i n, then a1   an  b = d. • ≤ ≤ ··· (ma + b)  a = a  b and (ma)  a = a for any positive integer m. • If (A,+) is additionally a monoid with zero element 0, then 0  a = a for a A. • ∈ a1  a2   an = (a1  a2)  (a2  a3)   (an 1  an). • ··· ··· − (a1 a2  an)b = (a1 b)(a2  an) = (a1 b)(a2 b)(a3  an)= = (a1 b)(a2 b)  • ··· ··· ··· ··· ··· (an  b).

Example 3.6. (1) Let N be the set of natural numbers including 0. Then N with the usual addition and multiplication forms a semiring. The polynomial semiring N[x1, ,xs] over N in variables x1, ,xs is the ··· ··· set of all polynomials over N with the usual addition and multiplication of polynomials. Similarly, we define the matrix semiring Mn(R) over a semiring R. (2) Let a  b := gcd(a,b) for a,b N. Then (N,+,) is a quarter-ring, but not a semiring. It is called the canonical quarter-ring of natural numbers.∈

Recall that a multiset is a collection of elements possibly containing duplicate. For a multiset λ = λ1, ,λs of s elements in N, we define a map ϕ from the polynomial semiring N[x1, ,xs] to (N,+,) { ··· } λ ··· as follows:

i1 is i1   is ϕλ : N[x1, ,xs] N, ai1, ,is x1 xs ai1, ,is λ1 λs . ··· −→ ··· ··· 7→ ··· ··· i1X, ,is i1X, ,is ··· ···

Note that ϕ is additive and can be extended to the matrix semiring over N[x1, ,xs]: λ ···

: Mn(N[x1, ,xs]) N, (ai j)n n (ai j) := (ai j)ϕλ, k − kλ ··· −→ × 7→ k kλ 1 Xi, j n ≤ ≤ where (ai j)ϕ denotes the image of ai j under ϕ . The map is called the norm map. λ λ k − kλ Now we recall the following elementary fact in combinatorics.

Lemma 3.7. Let Vi be a finite set for 1 i s. We define ≤ ≤

V := Vj, gi := Vk Vk , 1 i s, 1 ∩···∩ i ≤ ≤ 1 [j s 1 k1

gi(σ) := gcd(λk ,λk , ,λk ), 1 2 ··· i 1 k1

8 and Hi(σ) to be the subset of V consisting of all elements which belong to exactly i sets of the given V1,V2, ,Vs, namely ···

Hi(σ)= v V 1 k1 < k2 < < ki s,v Vk , v Vr . { ∈ |∃ ≤ ··· ≤ ∈ j 6∈ } 1 \j i r=k j[,1 j i ≤ ≤ 6 ≤ ≤ s i k i Clearly, Hi(σ) Hj(σ)= 0/ if 1 i = j s. By Lemma 3.7(1), hi(σ) := Hi(σ) = − ( 1) C gi k(σ). ∩ ≤ 6 ≤ | | k=0 − i+k + Obviously, gi(σ) and Hi(σ) depend merely on the type of σ. So gi(λ) and hi(λ) areP well-defined functions from the set of all partitions of positive integers to N if we define gi = 0 and hi = 0 for i > s. As in Example 3.6, we define a  b = gcd(a,b) for a,b Z. Then gi(x1, ,xs) is just the i-th elementary ∈ ··· symmetric polynomial over the canonical quarter-ring of natural numbers:

gi(x1, ,xs)= x j  x j   x j ··· 1 2 ··· i 1 j1 s. Clearly, gi(x1, ,xs) does not depend on ≤ ≤ ··· , ··· ··· any order of x1 x2 xs, and gs(x1, ,xs) gi(x1, ,xs) for 1 i s and all xi N. ≥ ≥···≥ ··· | ··· ≤ ≤ ∈ Recall that a partition λ of a positive integer n is an s-tuple (λ1, ,λs) of integers satisfying λ1 λ2 s ··· ≥ ≥ λs 1 and i=1 λi = n. We write λ = (λ1, ,λs) n for simplicity. Each λi is called a part of λ. ···≥ ≥ ··· ⊢ a1 a2 at Sometimes it is alsoP convenient to think of partitions as multisets and to write λ in the form (λ ,λ , ,λt ) 1 2 ··· with λ1 > λ2 > > λt 1, where λi appears ai 1 times in the multiset of λ. ··· ≥ ≥ Let P(n) be the set of all partitions of n, P(s,n) the set of partitions of n with exactly s parts, and Ps∗(n) the set of partitions of n with the largest part s. Then P(s,n) and Ps∗(n) have the same cardinality. This can be seen from Ferrers graphs (or Young diagrams) of partitions. Convention: Given partitions λ P(n) and µ P(m), we write (λ,µ) for the partition γ P(n + m) ∈ ∈ ∈ such that the parts of γ is the disjoint union of the parts of λ with the parts of µ. For example, if λ = (4,3,2,1) and µ = (5,4,4,2,2,1,1), then (λ,µ) = (5,4,4,4,3,2,2,2,1,1,1). For 1 d N, we define dλ = ˆ ≤ ∈ (dλ1, ,dλs) P(s,dn) and λ := (λ1, ,λs 1) = (λ1, ,λs 1,λs) P(s 1,n λs). Inductively, λ j = ··· ∈ − ··· − ··· − ∈ − − − (λ1, ,λs j). If j s, we define λs j = (0, ,0). Then gi(dλ)= d gi(λ), and ··· − ≥ − ···      gi(λ)= gi(λ ) + λ j1 λ ji 1 λs gi(λ ) + λ j1 λ ji 1 , − ··· − ≤ − ··· − 1 j1< X< ji 1 s 1 1 j1< X< ji 1 s 1 ≤ ··· − ≤ − ≤ ··· − ≤ − i 1 and therefore, if λs = 1, then gi(λ)= gi(λ )+Cs−1. In general, we have − − i 1 i 1 gi(λ )+Cs−1 gi(λ) gi(λ )+ min gi 1(λ ),λsCs−1 . − − ≤ ≤ − { − − − } Now, we introduce a new equivalence relation on P(s,n) and define a polynomial ε (x) for each λ n. ∼ λ ⊢ Definition 3.8. (1) The partition polynomial ε (x) of λ P(s,n) is λ ∈

s 1 s 1 i gi+1(λ) i ελ(x) := − ( 1) − − x i=0 − gs(λ) P s 1 gs 1(λ) s 2 s 2 g2(λ) s 1 g1(λ) = x − − x − + + ( 1) − x + ( 1) − N[x] Z[x]. − gs(λ) ··· − gs(λ) − gs(λ) ∈ ⊆ (2) The equivalence relation on P(s,n) (or P(n)) is given by ∼

λ µ if ε (x)= εµ(x). ∼ λ

Equivalently, λ µ if and only if gi(λ)= gi(µ) for 1 i s. This equivalence relation is called the polyno- ∼ ≤ ≤ mial equivalence of partitions in P(n).

9 n 1 n 1 i i+1 i n m n m i i+1 i n m Example 3.9. (1) ε n x − 1 C x while ε n m x − 1 C x 1 n (1 )( )= i=0 ( ) − − n , (m,1 − )( )= i=1 ( ) − − n m+1 +( ) − − − − − for 1 m < n and ε n (x)=P1. P ≤ ( ) (2) In P(2, p) with p a prime, any two-part partitions are polynomial equivalent because the partition polynomial of each such partition is of the form x p. In P(3,11), there hold (8,2,1) (7,2,2) (6,4,1) − ∼ ∼ ∼ (5,4,2), but (8,2,1) ≁ (6,3,2).

s 1 Remark that ( 1) − gs(λ)ελ(1) counts the number of distinct eigenvalues of cσ in C, where σ Σn has the cycle type λ (see− [12, Lemma 6.3]). ∈ Next, we consider when two partitions are polynomial equivalent. We first give a matrix interpretation of the polynomial equivalence.

Definition 3.10. Given λ = (λ1, ,λs) P(s,n), we associate it with a matrix Λ Ms(N) by setting ··· ∈ λ ∈

0 d12 d13 ... d1s 0 0 d23 ... d2s  . . . . . Λ := ......  λ    ..  0 0 . ds 1,s  ··· −  0 0 0 ... 0    where di j = λi  λ j is the product in the canonical quarter-ring of natural numbers. This matrix is called the triangular divisor matrix of λ.

Now we define a norm Λ of Λ . First, we consider λ j,1 j s, as variables and then regard Λ as k λk λ ≤ ≤ λ a matrix with the entries in the polynomial semiring N[λ1, ,λs], thinking of λi  λ j as the usual monomial ··· λiλ j in N[λ1, ,λs]. By applying λ to this matrix, we get the norm Λλ of the triangular divisor ··· k − k k k i matrix Λλ. That is, Λλ = 1 i< j s λi λ j. Similarly, we consider the i-th power Λλ of Λλ in the s s k k ≤ ≤ i i i × matrix semiring Ms(N[λ1, P,λs]) over N[λ1, ,λs] and then define Λ := Λ = (Λ )ϕ . Note that ··· ··· k λk k λkλ λ λ dimR Sn(λ,R)= g1(λ)+ 2 Λ (see Lemma 2.3(1)). k λk The triangular divisor matrix Λ of λ depends upon the ordering of λ1, ,λs. Nevertheless, we will λ ··· show that the norms of triangular divisor matrices are independent of the choices of orderings. k − k Proposition 3.11. Let λ,µ P(s,n). Then i ∈ (1) Λ = gi 1(λ) for 1 i s 1. k λk + ≤ ≤ − (2) λ µ if and only if Λi = Λi for all 1 i s 1, that is, the norms of Λi and Λi are equal for ∼ k λk k µk ≤ ≤ − λ µ all 1 i s 1. ≤ ≤ − i i i 1 Proof. (1) We check Λλ = gi+1(λ) for 1 i s 1. Since the summands of Λλ = Λλ− Λλ are    k k ≤ ≤ − k k k k of the form λ j1 λ ji λ ji+1 with j1 < < ji < ji+1, these are precisely the summands of gi+1(λ). Hence i ··· ··· Λλ = gi+1(λ). k k i i (2) Clearly, g1(λ)= g1(µ)= n. Hence g j(λ)= g j(µ) for all 1 j s if and only if Λ = Λ for all ≤ ≤ k λk k µk 1 i s 1. . ≤ ≤ − An immediate consequence of Proposition 3.11 is that if µ1, ,µs is a permutation of λ1, ,λs i i { ··· } { ··· } then Λ = Λ for all 1 i s 1 because the i-th elementary symmetric polynomial gi(λ) does not k λk k µk ≤ ≤ − depend on the ordering of its variables λi. Another consequence is the following result which is quite useful for deciding whether two partitions are polynomial equivalent.

Corollary 3.12. Let λ,µ P(s,n) with the corresponding triangular divisor matrices Λ = (di j) and Λµ = ∈ λ (ci j), respectively. If the multisets di j 1 i < j s and ci j 1 i < j s are equal, then λ µ. { | ≤ ≤ } { | ≤ ≤ } ∼

10 Proof. That the multisets di j 1 i < j s and ci j 1 i < j s are equal means g2(λ)= g2(µ).   {  | ≤ ≤ } { | ≤ ≤ }   Since any summand λ j1 λ j2 λ ji of gi(λ) for i 3 is a product d j1, j2 d j2, j3 d ji 1, ji (in the canonical ··· ≥ ··· − quarter-ring (N,+,) of i 1 elements of the mutiset di j 1 i < j s with the increasing indices, the − { | ≤ ≤ } summation of all such products of elements in the multiset of λ equals the one in the multiset of µ. Hence gi(λ)= gi(µ) for 2 i s. Clearly, g1(λ)= g1(µ). Thus λ µ.  ≤ ≤ ∼ Note that the converse of Lemma 3.12 fails. For example, λ = (12,4,3,1) µ = (10,5,3,2), but the ∼ corresponding multisets for λ and µ are 4,3,1,1,1,1 and 5,1,2,1,1,1 , respectively, they are clearly { } { } different. This example also shows that if the common part 3 is removed from λ and µ, then the resulting partitions (12,4,1) and (10,5,2) are no longer polynomial equivalent.

Corollary 3.13. Let λ = (λ1, ,λs),µ = (µ1, ,µs) P(s,n). If λi  λ j = d = µp  µq for all 1 i = j s ··· ··· ∈ ≤ 6 ≤ and 1 p = q s, then λ µ. In particular, if λ1, ,λs are pairwise coprime and if µ1, ,µs are pairwise ≤ 6 ≤ ∼ ··· ··· coprime, then λ µ. ∼ Proof. Under the assumption on λ, the triangular divisor matrix of λ is

0 d d ···. .  0 .. .  Λ = . λ  ..   . d    0  

Thus Λλ = Λµ. Hence Corollary 3.13 follows immediately from Corollary 3.12.  By the definition of partition polynomials ελ(x), we have the following useful fact. Lemma 3.14. Let λ,µ P(s,n). Then λ µ in P(s,n) if and only if dλ dµ in P(s,dn) for some integer ∈ ∼ ∼ d 1 if and only if dλ dµ inP(s,dn) for all integers d 1. ≥ ∼ ≥ Proof. The coefficients of the monic partition polynomial εdλ(x) is given by gi(dλ)/gs(dλ). As we know, gi(dλ)= dgi(λ) for all i. Thus εdλ(x)= ελ(x).  Thus, when considering λ µ, we can always assume that λ1, ,λs are coprime and that µ1, ,µs are ∼ ··· ··· coprime.

Lemma 3.15. Let λ,µ P(s,n) and 1 m N such that λ j m 1 i s = µ j m 1 j s as multisets. ∈ ≤ ∈ { | ≤ ≤ } { | ≤ ≤ } Then λ µ if and only if (λ,m) (µ,m) in P(s + 1,n + m). In particular, suppose λ j  m = d = µ j  m for ∼ ∼ µ 1 j s. Then λ µ if and only if ( λ , m ) ( , m ) in P(s + 1, n+m ). ≤ ≤ ∼ d d ∼ d d d Proof. The last conclusion in Lemma 3.15 follows from Lemma 3.14. We prove the other one. Let λ¯ = (λ,m) n+m P(s+1,n+m). Since the multisets λ j m 1 i s and µ j m 1 j s are equal, the ⊢ ∈ { | ≤ ≤ } { | ≤ ≤ } summations λ  m   λ  m equals the summation µ  m   µ  j1< j2< < ji 1 ( j1 ) ( ji 1 ) j1< j2< < ji 1 ( j1 ) ( ji 1 ··· − ··· − ··· − ··· − m In particular,P g λ¯ g µ¯ . Clearly, for 1 i s, we have g λ¯ Pg λ λ   ). s+1( )= s+1( ) i( )= i( )+ 1 ji< < ji 1 s j1 ≤ ≤ ≤ ··· − ≤ ··· λ  m g λ λ  m  λ  m . Thus, for 1 i s, g λ¯P g µ¯ if and only if ji 1 = i( )+ 1 ji< < ji 1 s( j1 ) ( ji 1 ) i( )= i( ) − ≤ ··· − ≤ ··· − ≤ ≤ gi(λ)= gi(µ). ThisP shows Lemma 3.15. . A special case of Lemma 3.15 is that λ µ if and only if (λ,1) (µ,1). A slightly generalized form of ∼ ∼ the last statement of Lemma 3.15 is as follows.

Lemma 3.16. Assume λ µinP(s,n) and γ δ in P(t,m). If λi γ j = µp δq for all 1 i, p s,1 j,q t, ∼ ∼ ≤ ≤ ≤ ≤ then (λ,γ) (µ,δ) in P(s +t,n + m). ∼ Proof. By assumption, we suppose d := λi  γ j = µp  δq. Now, we calculate gi(λ,γ) and gi(µ,δ) for 1 i s +t. Clearly, g1(λ,γ)= g1(µ,δ) and gs t (λ,γ)= gs t (µ,δ). We may assume s t and consider the ≤ ≤ + + ≤ following cases.

11 (1) 1 < i s t. ≤ ≤ ¯ gi(λ) = λ j   λ j + γk   γk 1 j1< < ji s 1 ··· i 1 k1<

(2) s + 1 i < s +t. ≤ In this case, we first assume i t. ≤ ¯ gi(λ) = γk   γk 1 k1<

Next, we assume t < i. Then

g λ¯ λ   λ  γ   γ i( ) = (p,q):p+q=i,1 p,q i 1 j1< < jp s,1 k1<

Lemma 3.17. Let R be a division ring of characteristic p 0, and let σ Σn be of cycle type (λ1,λ2, ,λs) ≥ ∈ ··· with λi 1 for all 1 i s. Then ≥ ≤ ≤ (1) Sn(σ,R) is semisimple if and only if p ∤ λi for 1 i s. ≤ ≤ h1(σ) h2(σ) (2) If R is an algebraically closed field and p ∤ λi for all i, then Sn(σ,R) R⊕ M2(R)⊕ hs(σ) h1(σ) ≃ × ×···× Ms(R)⊕ as algebras, where R⊕ denotes the direct product of h1(σ) copies of the ring R.

Proof. (1) Clearly, Mn(R) is a simple artinian ring if R is a division ring. Let rad(A) denote the Ja- cobson radical of a ring A. If Sn(σ,R) is semisimple, that is rad Sn(σ,R) = 0, then rad fiSn(σ,R) fi = firad Sn(σ,R) fi = 0 for all 1 i s. This means that all fiSn(σ,R) fi are semisimple. By Lemma 2.5(1), ≤ ≤ the group algebra R[Cλi ] is semisimple. Hence p ∤ λi for all i. Conversely, if p ∤ λi for all i, then p does not divide the least common multiple of these λi, that is, the order of G = σ is invertible in R. In this case, h i Sn(σ,R) is semisimple by [10, Theorem 1.15, p.15], which says that if a finite group acts on a semisimple artinian ring with its order invertible in the ring, then the fixed ring is semisimple. (2) Let cσ := e + e + + e be the permutation matrix in Mn(R) corresponding to σ. Then 1,(1)σ 2,(2)σ ··· n,(n)σ cσ = diag cσ ,cσ , ,cσ , where cσ M (R) is defined similarly. For 1 i s, the eigenvalues of cσ are { 1 2 ··· s } i ∈ λi ≤ ≤ i distinct and denoted by Vi. Hence the Jordan canonical form of cσi has only Jordan blocks of order 1, namely λi 1 the diag(1,ωi, ,ω − ), where ωi is a primitive root of the unity of order λi in R. This ··· i implies that the Jordan canonical form of cσ is the diagonal matrix

λ1 1 λi 1 λs 1 c := diag(1,ω1, ,ω − , ,1,ωi, ,ω − , ,1,ωs, ,ω − ). ··· 1 ··· ··· i ··· ··· s 12 For 1 i s, we write Hi(σ)= ri1,ri2, ,r (see the notation above Definition 3.8). Then the algebraic ≤ ≤ { ··· ihi(σ)} and geometric multiplicities of the eigenvalue ri j of cσ are just i for 1 j hi(σ). Thus c is similar to ≤ ≤ d := diag(r11I1,r12I1, ,r I1, ,ri1Ii,ri2Ii, ,r Ii, ,rs1Is,rs2Is, ,r Is) ··· 1h1(σ) ··· ··· ihi(σ) ··· ··· shs(σ) s with i=1 ihi(σ)= n. Recall that Ii denotes the in Mi(R). Hence the matrix cσ is similar to the matrixP d. By Lemma 2.2 (see also [17, Lemma 2.1]), Sn(σ,R)= Sn(cσ,R) := x Mn(R) xcσ = cσx { ∈ | }≃ Sn(d,R) := x Mn(R) xd = dx as algebras. Note that ri j = rkl for (i, j) = (k,l) with 1 i,k s,1 { ∈ | } 6 6 ≤ ≤ ≤ j hi(σ),1 l hk(σ). It follows from [17, Lemma 2.2] that the centralizer algebra Sn(d,R) is isomorphic ≤ h (σ) ≤ ≤ h (σ) h (σ) h (σ) h (σ) h (σ) to R 1 M2(R) 2 Ms(R) s . Thus Sn(σ,R) R 1 M2(R) 2 Ms(R) s as ⊕ × ⊕ ×···× ⊕ ≃ ⊕ × ⊕ ×···× ⊕ algebras.  Lemma 3.17(2) shows that the multiplicity of the matrix algebra Mi(R) in a Wedderburn decomposition of the semisimple algebra Sn(λ,R) is given by hi(λ).

h1(σ) h2(σ) hs(σ) Proof of Theorem 1.2. By Lemma 3.17(1), Sn(σ,R) R⊕ M2(R)⊕ Ms(R)⊕ and h (τ) h (τ) h (τ) ≃ × ×···× Sm(τ,R) R 1 M2(R) 2 Mt (R) t . ≃ ⊕ × ⊕ ×···× ⊕ (i) It follows from hi(σ) 0,hi(τ) 0 for 1 i s, hs(σ)= λ1   λs = 0 and ht (τ) = 0 that ≥ ≥ ≤ ≤ ··· 6 6 h1(σ) h2(σ) hs(σ) h1(τ) h2(τ) ht (τ) R⊕ M2(R)⊕ Ms(R)⊕ R⊕ M2(R)⊕ Mt(R)⊕ × ×···× ≃ × ×···× as algebras if and only if s = t and hi(σ)= hi(τ) for all 1 i s. ≤ ≤ Suppose Sn(σ,R) Sm(τ,R), that is, we assume that s = t and hi(σ)= hi(τ) for all i. We show that m = n ≃ s t and ελ(x)= εµ(x). In fact, since i=1 ihi(σ)= n and i=1 ihi(τ)= m, we have m = n. By the definitions of hs(σ) and hs(τ), we get hs(σ)=Pgs(σ) and hs(τ)= gsP(τ). It follows from hs(σ)= hs(τ) that gs(σ)= gs(τ). s 1 Now, by the definitions hs 1(σ) and hs 1(τ) and Lemma 3.7(1), we have hs 1(σ)= gs 1(σ) Cs− gs(σ) s−1 − − − − and hs 1(τ)= gs 1(τ) Cs− gs(τ). It then follows from hs 1(σ)= hs 1(τ) and gs(σ)= gs(τ) that gs 1(σ)= − − − − − − gs 1(τ). Continuing this procedure, we get gi(σ)= gi(τ) for 1 i s 2. Thus, if hi(σ)= hi(τ) for 1 i s, − ≤ ≤ − ≤ ≤ then gi(σ)= gi(τ) for 1 i s and ε (x)= εµ(x). ≤ ≤ λ Conversely, if m = n and ε (x)= εµ(x), then s = t and gs(τ)gi(σ)= gs(σ)gi(τ) for 1 i s 1. As λ ≤ ≤ − g1(λ)= n = m = g1(µ), we get gs(σ)= gs(τ). Thus it follows from ελ(x)= εµ(x) that gi(λ)= gi(τ) for all 1 i s. By Lemma 3.7(2), we obtain hi(σ)= hi(τ) for all i. Hence Sn(σ,R) Sm(τ,R). ≤ ≤ h1(σ) h2(σ) hs(σ) ≃ (ii) Since Sn(σ,R) R⊕ M2(R)⊕ Ms(R)⊕ by Lemma 3.17(2), the basic algebra ≃ × h (σ) h×···×(σ) h (σ) B(σ) of Sn(σ,R) is isomorphic to R 1 R 2 R s . As is known, Sn(σ,R) and B(σ) are ⊕ × ⊕ ×···× ⊕ always Morita equivalent, while Sn(σ,R) and Sm(τ,R) are Morita equivalent if and only if their basic algebras s t B(σ) and B(τ) are isomorphic. Clearly, B(σ) B(τ) if and only if i=1 hi(σ)= i=1 hi(τ). By Lemma s t ≃ s i+1 t i+1 3.7(2), i=1 hi(σ)= i=1 hi(τ) if and only if i=1( 1) gi(σ)= Pi=1( 1) gPi(τ). Therefore Sn(σ,R) −s i+1 −t i+1 and Sm(Pτ,R) are MoritaP equivalent if and onlyP if ( 1) gi(σP)= ( 1) gi(τ) if and only if i=1 − i=1 − gs(λ)ελ(1)= gt (µ)εµ(1) if and only if the matrices Pcσ and cτ have the sameP number of distinct eigenvalues in R. Note that gs(λ) is the greatest common divisor of λ1, ,λs.  ··· Theorem 1.2 tells us that the foregoing consideration on the polynomial equivalence λ µ in P(s,n) ∼ provides us with many isomorphisms of invariant matrix algebras. For example, S38 (17,11,8,2),C ≃ S38 (17,11,6,4),C by repeatedly applying Corollary 3.15 to the polynomial equivalence (8,2) (6,4) in ∼ P(2,10). Also, for any p 5 and two-part partitions λ and µ of p, it follows from Theorem ≥ 1.2 that Sp(λ,C) Sp(µ,C). For instance, S5((4,1),C) S5((3,2),C) C C C M2(C). ≃ ≃ ≃ × × × Remark that the condition m = n in Theorem 1.2(2)(i) is needed. For example, both λ = (4,2,2) and 2 µ = (2,1,1) have the same partition polynomial x 3x + 4, but S8(λ,R) S4(µ,R) for any ring R. − 6≃ As a consequence of Theorem 1.2 and Corollary 3.13. we have immediately the corollary.

Corollary 3.18. Let R be an algebraically closed field of characteristic p 0, and let σ Σn and τ Σm be ≥ ∈ ∈ of cycle types (λ1,λ2, ,λs) and (µ1,µ2, ,µt ), respectively. Assume that p ∤ λi and p ∤ µ j for 1 i s and ··· ··· ≤ ≤ 1 i t. If λ1, ,λs are pairwise coprime and µ1, ,µt are pairwise coprime, then ≤ ≤ { ··· } { ··· } (1) Sn(σ,R) Sm(τ,R) if and only if n = m and s = t. ≃ (2) Sn(σ,R) and Sm(τ,R) are Morita equivalent if and only if n s = m t. − − 13 4 Example of real invariant matrix algebras

As an example of invariant matrix algebras, we consider Sn(c,R) in this section.

Proposition 4.1. Let σ Σn be of cycle type (λ1,λ2, ,λs) with λi 1 for all 1 i s, and l be the ∈ ··· ≥ ≤ ≤ λ number of even parts of λ. Then we have the isomorphism of R-algebras:

h1(σ)/2 h2(σ)/2 (hs(σ) 1)/2 C⊕ M2(C)⊕ Ms(C)⊕ − Ms(R) i f lλ = 0, h1(σ)/2 × h2(σ)/2 ×···× (hs(σ) 2)/2 × 2  C⊕ M2(C)⊕ Ms(C)⊕ − Ms(R)⊕ i f l = s, Sn(σ,R) × ×···× × λ ≃  h j(σ)/2 (h j(σ) 1)/2  ( M j(C) ) ( M j(C) − ) Mlλ (R) Ms(R) i f 0 < lλ < s, 1 j=l s 1 × j l ,s × × ≤ 6 Lλ≤ − ∈{Lλ }  m  where R⊕ denotes the direct product of m copies of a ring R. The proof of this proposition relies on the following lemmas. For (x,y) R R, let ∈ ×

K(x,y) I2  K(x,y) I2  x y .. .. K(x,y) := − , Kn(x,y) :=  . .  M2n(R), y x    ∈  ..   . I2     K(x,y)   where I2 is the identity of M2(R). Let F := K(r,s) M2(R) r,s R . Then there is an algebra isomorphism { ∈ | ∈ } S C by sending K(x,y) x + iy for x,y R, where i is the complex unit √ 1. ≃ 7→ ∈ − Lemma 4.2. Let I = Km(u,v) M2m(R) and J = Kn(x,y) M2n(R), where u,v,x,y R and v > 0,y > 0. ∈ ∈ ∈ Define Γ := a M2m 2n(R) Ia = aJ and Λ := b Mm n(F) Ib = bJ M2m 2n(R). { ∈ × | } { ∈ × | }⊆ × (1) If (u,v) = (x,y), then Γ = 02m 2n . 6 { , } (2) If (u,v) = (x,y), then Γ = Λ.

Proof. Let T := b M2(R) K(u,v)b = bK(x,y) . We show T = 0 if (u,v) = (x,y), and T = F if { ∈ | } 6 (u,v) = (x,y). In fact, take b0 = (b jk) T, that is K(u,v)b0 = b0K(x,y). This implies ∈

u x y v 0 b11 − − −  y u x 0 v b12 ( ) − − = 0. ∗ v 0 u x y b21  − −    0 v y u xb22  −   The determinant of the coefficient matrix of ( ) is (u x)2((u x)2 + 2v2 + y2) + (v2 y2)2. Let w := (u ∗ − − − − x)2((u x)2 + 2v2 + y2)+ (v2 y2)2. It follows from v > 0 and y > 0 that w = 0 if and only if (u,v) = (x,y). − − Thus if (u,v) = (x,y), then w = 0 and b jk = 0 for 1 j k. This implies b0 = 0 and T = 0. If (u,v) = (x,y), 6 6 ≤ ≤ then it follows from ( ) that b11 = b22 and b12 = b21. Therefore b0 F and T F. On the other hand, ∗ − ∈ ⊆ since F is a field, we have K(r,s) T for r,s R, and therefore F T. Hence T = F. In particular, ∈ ∈ ⊆ b M2(R) K(0,v)b = bK(0,v) = F. { ∈ | } Next, we show that if we write a = (A jk) M2m 2n(R) as an m n block matrix with A jk M2(R), then ∈ × × ∈ Ia = (K(u,v)A jk + A j+1,k) M2m 2n(R) and aJ = (A j,k 1 + A jkK(x,y)) M2m 2n(R). Hence a Γ if and ∈ × − ∈ × ∈ only if the following holds

(⋆) K(u,v)A jk A jkK(x,y)= A j,k 1 A j+1,k for all 1 j m,1 k n. − − − ≤ ≤ ≤ ≤

Here we understand A j0 = Am 1 k = 0 for 1 j m, 1 k n. + , ≤ ≤ ≤ ≤ 14 (1) If (u,v) = (x,y), then T = 0. We shall show Γ = 0. Let a = (A jk) Γ. Due to (⋆), K(u,v)Am1 6 ∈ − Am1K(x,y)= Am 0 Am 1 1 = 0. Thus Am1 T and Am1 = 0. It follows from K(u,v)A j1 A j1K(x,y)= A j 0 , − + , ∈ − , − A j 1 1 = A j 1 1 that A j1 = 0 for 1 j m 1. Similarly, K(u,v)A j2 A j2K(x,y)= A j 1 A j 1 2 = A j 1 2 + , − + , ≤ ≤ − − , − + , − + , that A j2 = 0 for 1 j m. Continuing this argument, we get A jk = 0 for 1 j m and 3 k n. Thus ≤ ≤ ≤ ≤ ≤ ≤ a = 0 and Γ = 0 . { } (2) If (u,v) = (x,y), then T = F. Clearly, Mm n(F) M2m 2n(R) and Λ Γ. To prove Λ = Γ, we have × ⊆ × ⊆ to prove Λ Γ, that is, if a = (Ai j) Γ, we have to show A jk F for all 1 j m,1 k n. ⊇ ∈ ∈ ≤ ≤ ≤ ≤ Indeed, we have K(u,v)= I2 + K(0,v) with I2 the identity matrix in M2(R). Thus (⋆) holds true if and only if (⋆⋆) K(0,v)A jk A jkK(0,v)= A j,k 1 A j+1,k for all 1 j m,1 k n. − − − ≤ ≤ ≤ ≤ Notice also the fact (†) If c M2(R) satisfies K(0,v)c cK(0,v) F, then c F. ∈ − ∈ ∈ This can be seen as follows: we write d := K(0,v)c cK(0,v) F. Then K(0,v)d = K2(0,v)c 2 − ∈ 2 2 2 − K(0,v)cK(0,v) and dK(0,v)= K(0,v)cK(0,v) cK (0,v). It follows from K (0,v)= v I2 that K (0,v)c = − − cK2(0,v) and K(0,v)d = dK(0,v). Due to d F, there holds K(0,v)d = dK(0,v). Thus K(0,v)d = 0 and − ∈ K2(0,v)d = v2K(0,v)d = v2d = 0. It then follows from v > 0 that d = 0 and c F. − − ∈ Suppose a = (Ai j) Γ. By (⋆⋆), K(0,v)Am1 Am1K(0,v)= Am 0 Am 1 1 = 0 F. Thus Am1 F. ∈ − , − + , ∈ ∈ Note that K(0,v)A j1 A j1K(0,v)= A j,0 A j+1,1 = A j+1,1 for 1 j m 1. By (†), we have Am 1,1 F. − − − ≤ ≤ − − ∈ By repeating this argument, we get A j1 F for 1 j m 1. Similarly, it follows from K(0,v)Am2 ∈ ≤ ≤ − − Am2K(0,v)= Am 1 Am 1 2 = Am1 F and K(0,v)A j2 A j2K(0,v)= A j1 A j 1 2 that A j2 F for all 1 , − + , ∈ − − + , ∈ ≤ j m. Continuing the procedure, we finally obtain A jk F for all 1 j m and 3 k n. This means ≤ ∈ ≤ ≤ ≤ ≤ a Λ. Hence Λ Γ.  ∈ ⊇ Lemma 4.3. For x,y,r R and y > 0, let I = Km(x,y) M2m(R), and let J = Jn(r) Mn(R) be the Jordan block with the eigenvalue∈ r. ∈ ∈ (1) If a M2m n(R) satisfies Ia = aJ, then a = 0. ∈ × (2) If b Mn 2m(R) satisfies bI = Jb, then b = 0. ∈ × Proof. (1) Let a = (a jk) M2m n(R) with Ia = aJ. As a convention, we set a j0 = a2m+1,k = a2m+2,k = 0 ∈ × for 1 j 2m, 1 k n. Then ≤ ≤ ≤ ≤ 2m m 2m 2 2m n − Ia = xe j j + y(e2 j,2 j 1 e2 j 1,2 j)+ e j, j+2 ( apqepq) − − − Xj=1 Xj=1 Xj=1  Xp=1 Xq=1 2m n m n 2m 2 n − = xa jqe jq + y(a2 j 1,qe2 j,q a2 j,qe2 j 1,q)+ a j+2,qe jq − − − Xj=1 Xq=1 Xj=1 Xq=1 Xj=1 Xq=1 2m n m n 2m n = xa jqe jq + y(a2 j 1,qe2 j,q a2 j,qe2 j 1,q)+ a j+2,qe jq − − − Xj=1 Xq=1 Xj=1 Xq=1 Xj=1 Xq=1 m n = (xa2p,q + ya2p 1,q + a2p+2,q)e2p,q + (xa2p 1,q ya2p,q + a2p+1,q)e2p 1,q . − − − − Xp=1 Xq=1  m n Similarly, aJ = p=1 q=1 (a2p,qr+a2p,q 1)e2p,q +(a2p 1,qr+a2p 1,q 1)e2p 1,q . Thus Ia = aJ if and − − − − − only if (x r)a2p,q +Pya2p P1,q = a2p,q 1 a2p+2,q and (x r)a2p 1,q ya2p,q = a2p 1,q 1 a2p+1,q for all 1 − − − − − − − − − − ≤ p m,1 q n. In particular, there holds ≤ ≤ ≤ (x r)a2m,1 + ya2m 1,1 = a2m,0 a2m+2,1 = 0, ( ) − − − ∗  (x r)a2m 1,1 ya2m,1 = a2m 1,0 a2m+1,1 = 0. − − − − − 2 2 Since the determinant of the coefficients of the linear equations ( ) is (x r) + y > 0, we obtain a2m 1 = ∗ − , a2m 1,1 = 0. Similarly, it follows from (x r)a2p,1 +ya2p 1,1 = a2p,0 a2p+2,1 = a2p+2,1 and (x r)a2p 1,1 − − − − − − − − 15 ya2p,1 = a2p 1,0 a2p+1,q = a2p+1,q for 1 p m 1 that a2p,1 = a2p 1,1 = 0 for 1 p m 1. Therefore − − − ≤ ≤ − − ≤ ≤ − a j1 = 0 for 1 j 2m. In the same way, we can deduce from (x r)a2p,2 + ya2p 1,2 = a2p,1 a2p+2,2 = ≤ ≤ − − − a2p+2,2 and (x r)a2p 1,2 ya2p,2 = a2p 1,1 a2p+1,2 = a2p+1,2 for 1 p m that a2p,2 = a2p 1,2 = 0 − − − − − − − ≤ ≤ − for 1 p m. Then a j2 = 0 for 1 j 2m. Pursuing this process, we finally get a jk = 0 for all 1 j 2m ≤ ≤ ≤ ≤ ≤ ≤ and 3 k n. Thus a = 0. ≤ ≤ (2) The proof is similar to the one of (1), we ommit here the details. 

Proof of Proposition 4.1. Let cσ := e + e + + e be the permutation matrix in Mn(R) 1,(1)σ 2,(2)σ ··· n,(n)σ corresponding to σ. Then cσ = diag cσ ,cσ , ,cσ with cσ M (R). For 1 j s, the eigenvalues in { 1 2 ··· s } j ∈ λ j ≤ ≤ C of cσ are distinct and denoted by Vj. Suppose Hj(σ)= r j1,r j2, ,r C for 1 j s. Then the j { ··· jh j (σ)}⊂ ≤ ≤ algebraic and geometric multiplicities of the eigenvalue r jk of cσ are just j for 1 k h j(σ). Clearly, if n is ≤ ≤ odd, then 1 is the only real n-th roots of the unity; if n is even, then 1 and 1 are the all real n-th roots of the − unity. We have to consider three cases: lλ = s, 0 < lλ < s and lλ = 0. (i) l = s. Then 1 and 1 are the all real eigenvalues of cσ and 1, 1 Hs(σ). Since non-real complex λ − − ∈ eigenvalues of a real matrix always occur in complex-conjugate pairs, we must have 2 h j(σ) for 1 j s 1 | ≤ ≤ − and 2 (hs 2). Without loss of generality, we may assume that r j,2k 1 = x jk +iy jk and r j,2k = x jk iy jk with | − − − x jk,y jk R and y jk > 0 for 1 j s 1 and 1 k h j(σ)/2, and that rs,2k 1 = xsk +iysk and rs,2k = xsk iysk ∈ ≤ ≤ − ≤ ≤ − − with xsk,ysk R and ysk > 0 for 1 k (hs(σ) 2)/2 and rs,hs(σ) 1 = xs,hs(σ) 1 + iys,hs(σ) 1 = 1, and ∈ ≤ ≤ − − − − − r = x + iy = 1. It follows from [6, Theorem 3.4.5, p.152] that cσ Mn(R) is similar via real s,hs(σ) s,hs(σ) s,hs(σ) ∈ similarity to a block diagonal real matrix of the form c := diag(K,J), where

K :=diag K(x11,y11),K(x12,y12), ,K(x ,y ), , ··· 1,h1(σ)/2 1,h1(σ)/2 ··· s 1 s 1 s 1 K(xs 1,1,ys 1,1) − ,K(xs 1,2,ys 1,2) − , ,K(xs 1,hs 1(σ)/2,ys 1,hs 1(σ)/2) − , − − − − ··· − − − − s s s K(xs1,ys1) ,K(xs2,ys2) ,K(xs,(hs(σ) 2)/2,ys,(hs(σ) 2)/2) Mn 2s(R), ··· − − ∈ − s s  J :=diag(J1( 1) ,J1(1) ) − j Here L means that the matrix L appears j times in the diagonal. According to Lemma 2.2, Sn(σ,R) ≃ Sn(c,R) as algebras. It follows from Lemma 4.3 that Sn(c,R) Sn 2s(K,R) S2s(J,R) as algebras. For ≃ − × 1 j, p s,1 k h j(σ) and 1 q hp(σ), since r jk = rpq for ( j,k) = (p,q), we get (x jk,y jk) = (xpq,ypq) ≤ ≤ ≤ ≤ ≤ ≤ 6 6 6 for ( j,k) = (p,q). By Lemma 4.2(1), we obtain 6

s 1 h j(σ)/2 (hs(σ) 2)/2 − j − s Sn 2s(K,R) S2 j(K(x jk,y jk) ,R) S2s(K(xsk,ysk) ,R) − ≃ j⊕=1 k⊕=1 × k⊕=1   j as algebras. Further, by Lemma 4.2(2), we have S2 j(K(x jk,y jk) ,R) M j(F) as algebras. As F C as fields, h1(σ)/2 h2(σ)/≃2 (hs(σ) 2)/2≃ we see that Sn 2s(K,R) is isomorphic to C⊕ M2(C)⊕ Ms(C)⊕ − as algebras. − s × s ×···× Thanks to [17, Lemma 2.2], S2s(J,R) Ss(J1( 1) ,R) Ss(J1(1) ,R)= Ms(R) Ms(R) as algebras. Thus h (σ)/2 h (σ)/2 ≃ − (h (×σ) 2)/2 2 × Sn(c,R) C 1 M2(C) 2 Ms(C) s Ms(R) as algebras. Hence ≃ ⊕ × ⊕ ×···× ⊕ − × ⊕ h1(σ)/2 h2(σ)/2 (hs(σ) 2)/2 2 Sn(σ,R) C⊕ M2(C)⊕ Ms(C)⊕ − Ms(R)⊕ . ≃ × ×···× ×

(ii) 0 < lλ < s and (iii) lλ = 0. These two cases can be proved similarly as in (i), we leave the details to the interested reader.  For real invariant matrix algebras, we have

Corollary 4.4. Let σ Σn and τ Σm be of cycle types λ = (λ1,λ2, ,λs) and µ = (µ1,µ2, ,µt ), respec- ∈ ∈ ··· ··· tively, and let lλ be the number of even parts of λ and dλ the the greatest common divisor of these λ j’s. Then (1) Sn(σ,R) Sm(τ,R) if and only if m = n, l = lµ and ε (x)= εµ(x). ≃ λ λ (2) Sn(σ,R) and Sm(τ,R) are Morita equivalent if and only if lλ = 0 is equivalent to lµ = 0 and dλελ(1)= dµεµ(1).

16 Finally, we mention that there are a few issues worthy of further consideration, which have directly been suggested in this section but not answered. We formalize some of them as the following questions.

Questions. (1) When are Sn(λ,R) and Sn(µ,R) isomorphic (or derived equivalent) for λ,µ P(s,n) and ∈ R a unitary ring? s 1 s 2 s i 1 i s 2 s 1 (2) Given a polynomial f (x)= x − as 2x − + + ( 1) − − aix + + ( 1) − a1x + ( 1) − a0 i+1 − − ··· − ··· − − ∈ Z[x] with s 3 and ai C for 1 i s 2, is there a partition λ P(s,n) for some n such that ε (x)= ≥ ≥ s ≤ ≤ − ∈ λ f (x)? In other words, what are the necessary and sufficient conditions for a monic polynomial f (x) to be a partition polynomial? (3) Are ελ(x)εµ(x) and ελ(εµ(x)) again partition polynomials? Or more generally, how do the operations of partitions correspond to the ones of partition polynomials? (4) Let I(s,n) be the set of equivalence classes of P(s,n) with respect to the polynomial equivalence relation. Then I(s,n) is the union of distinct equivalence classes. Let E(s,n, j) be the set of those equivalence classes that contain j elements of P(s,n). Define p(s,n), i(s,n) and e(s,n, j) to be the numbers of elements in P(s,n), I(s,n) and E(s,n, j), respectively. What are the generating functions of i(s,n) and e(s,n, j)? ∞ n qs Note that the generating function for p(s,n) is n=0 p(s,n)q = (1 q)(1 q2) (1 qs) (see [1, Chapter 6]). (5) Describe partitions λ P(s,n) that are self-equivalent,P that is, if µ− P−(s,n···) such− that µ λ, then µ = λ. ∈ ∈ ∼ For instance, λ = (4,4,1) is self-equivalent with ε (x) = (x 3)2, but µ = (5,2,2) is not self-equivalent. (4,4,1) − (6) For which partitions λ are ελ(x) irreducible polynomials over Q? (7) What are the necessary and sufficient conditions for λ µ in P(s,n) in terms of the conjugate partitions ∼ of λ and µ? Acknowledgement. The research work of both authors was partially supported by the National Natural Science Foundation of China (12031014) and Beijing Natural Science Foundation (1192004). The corre- sponding author CCX thanks Professor Hourong Qin from Nanjing University for discussion on greatest common divisor matrices.

References

[1] G.E.ANDREWS and K. ERIKSSON, Integer partitions, Cambridge University Press, Cambridge, 2004.

[2] S. BESLIN and S. LIGH, Greatest common divisor matrices. Appl. 118 (1989) 69-76.

[3] H. CARTAN and S. EILENBERG, Homological Algebra, Princeton Univ. Press, Princeton, NJ, 1956.

[4] L. DATTA and S. D. MORGERA, On the reducibility of centrosymmetric matrices-Applications in engineering problems, Circuits Systems Sig. Proc. 8 (1) (1989) 71-96.

[5] K. HIRATA and K. SUGANO, On semisimple extensions and separable extensions over non commuta- tive rings, J. Math. Soc. Japan 18(4)(1968) 360-373.

[6] R.A.HORN and C. R. JOHNSON, Matrix Analysis, Cambridge University Press, 1985.

[7] L. KADISON, New examples of Frobenius extensions, University Lecture Series 14, Amer. Math. Soc., Providence, RI, 1999.

[8] G. I. LEHRER and R. B. ZHANG, Cellular algebras, The second fundamental theorem of invariant theory for the orthogonal group, Ann. Math. 176 (3) (2012) 2031-2054.

[9] Z.S.LI, The of GCD matrices, Linear Algebra Appl. 134 (1990) 137-143.

[10] S. MONTGOMERY, Fixed rings of finite automorphism groups of associative rings. Lecture Notes in Mathematics 818. Springer, Berlin, 1980.

17 [11] J. L. STUART, Inflation matrices and ZME-matrices that commute with a permutation matrix. SIAM J. Matrix Anal. Appl. 9 (3) (1988) 408-418.

[12] J. L. STUART and J. R. WEAVER, Matrices that commute with a permutation matrix. Linear Algebra Appl. 150 (1991) 255-265.

[13] K. SUGANO, Separable extensions and Frobenius extensions, Osaka J. Math. 7 (1970) 291-299.

[14] J. R. WEAVER, Centrosymmetric (cross-symmetric) matrices, their basic properties, eigenvalues, and eigenvectors. Amer. Math. Monthly 92 (10) (1985) 711-717.

[15] H. WEYL, The classical groups. Their invariants and representations, Princeton University Press, Princeton, N.J., 1939.

[16] C. C. XI and S. J. YIN, Cellularity of algebras and Frobenius extensions, Linear Algebra Appl. 590 (2020) 317-329.

[17] C. C. XI and J. B. ZHANG, Structure of centralizer matrix algebras. Linear Algebra Appl. 622 (2021) 215-246.

Changchang Xi, School of Mathematical Sciences, Capital Normal University, 100048 Beijing, P.R. China; and School of Mathematics and , Central China Normal University, 430079 Wuhan, P.R. China Email: [email protected] Jinbi Zhang, School of Mathematical Sciences, Capital Normal University, 100048 Beijing, P.R. China Email: [email protected]

18