• Nenhum resultado encontrado

As a consequence, we can provide a more profound structure result concerning the group gradings on the incidence algebras, and we can classify their isomorphism classes of group gradings

N/A
N/A
Protected

Academic year: 2022

Share "As a consequence, we can provide a more profound structure result concerning the group gradings on the incidence algebras, and we can classify their isomorphism classes of group gradings"

Copied!
23
0
0

Texto

(1)

ALGEBRAS

EDNEI A. SANTULO JR, JONATHAN P. SOUZA, AND FELIPE Y. YASUMURA

Abstract. In this work, we classify the group gradings on finite-dimensional incidence algebras over a field, where the field has characteristic zero, or the characteristic is greater than the dimension of the algebra, or the grading group is abelian.

Moreover, we investigate the structure of G-graded (D1, D2)-bimodules, whereGis an abelian group, andD1 andD2 are the group algebra of finite subgroups ofG. As a consequence, we can provide a more profound structure result concerning the group gradings on the incidence algebras, and we can classify their isomorphism classes of group gradings.

1. Introduction

When one deals with an algebraic structure, it is usually useful to consider an additional structure on it, when possible. Considering group gradings in algebras has proved to be useful to solve relevant mathematical problems more than once.

That is the case of the classification of finite-dimensional semisimple Lie algebras, which are graded by the root system (see e.g. [8], or [15]). That is also the case of the positive solution for the Specht problem given by Kemer in [12]. Furthermore, group gradings naturally appear in several branches of Mathematics and Physics.

We briefly recall the definition of graded algebras. Let A be any algebra (not necessarily associative, not necessarily with unit), and letGbe any group. We say that A is G-graded if there exists a vector space decomposition A = L

g∈GAg, whereAgare possibly zero subspaces, such thatAgAh⊆Agh, for allg, h∈G. The elements of∪g∈GAgare called homogeneous, and a non-zerox∈Ag is said to have degreeg, denoted by degGx=g, or simply degx=g if there is no ambiguity. An extensive theory concerning graded algebras can be found in the monograph [6].

An interesting question is, given an algebra, determine all the possible group gradings on it up to graded isomorphisms. In this direction, the works [1, 3] gave the answer for the matrix algebras (they investigate a more general situation), and then, in [2], the authors show how to obtain group gradings on some simple Lie and Jordan algebras from the knowledge of the group gradings on matrix algebras.

These works started an extensive research in the subject. The monograph [6] is a complete state-of-art of the theory.

2010Mathematics Subject Classification. 16W50.

Key words and phrases. Group Gradings, Incidence algebras, Associative algebras.

The second author was financed by the Coordena¸ao de Aperfei¸coamento de Pessoal de N´ıvel Superior - Brasil (CAPES) - Finance Code 001.

The third author was supported by S˜ao Paulo Research Foundation (Fapesp), grant number 2013/22.802-1, and by the Coordena¸ao de Aperfei¸coamento de Pessoal de N´ıvel Superior - Brasil (CAPES) - Finance Code 001.

1

(2)

In this paper, we are interested in the non-simple associative algebras called incidence algebras. These are a generalization for the upper triangular matrices U Tn. In short, an incidence algebra is a subalgebra of U Tn, generated by matrix units and containing all diagonal matrices (see a precise definition below). Thus, the classification of group gradings on upper triangular matrices is a particular result of the classification of group gradings on the incidence algebras. The classification of group gradings on the upper triangular matrices was obtained in two works [17, 5].

To state the classification, we give some definitions. We call a grading onU Tn good if all matrix unitseij are homogeneous in the grading; and we call aG-grading on U Tnelementaryif there exists a sequence (g1, g2, . . . , gn)∈Gnsuch that everyeijis homogeneous of degreegigj−1. Clearly every elementary grading is a good grading, and, in the context of upper triangular matrices, the converse holds. Hence, good and elementary gradings are equivalent notions for the algebra of upper triangular matrices. In [17], it is proved that every group grading onU Tn is elementary, up to an isomorphism. In [5], the isomorphism classes of elementary gradings onU Tnare classified, thus obtaining a complete classification of group gradings on this algebra.

Note that the notions of elementary gradings and good gradings can be defined for subalgebras of matrix algebras generated by matrix units. Hence, in particular, one can define these notions for the incidence algebras. It is worth mentioning that these notions were generalized in [4] for a larger class of algebras. It is natural to conjecture that these two notions are equivalent, and that every group grading on an incidence algebra is elementary up to an isomorphism. However, these two notions are not always equivalent: not every group grading on an incidence algebra is elementary or good (see below). Thus, incidence algebra turns out to be a more intricate combinatorial object in the point of view of gradings by a group.

In this paper, we classify group gradings on the incidence algebraI(X) over a field F, where X is finite and one of the following conditions hold: the base field has characteristic zero, charF>dimI(X), orGis abelian.

Some works were dedicated to study the group gradings on the incidence algebras [13, 10, 14], but none of them provide a complete understading of its group gradings.

Some different phenomenon happens in this algebra. The first surprising fact is that not every good grading onI(X) is necessarily elementary, even if X is finite (see, for instance, [10, Example 12]). Moreover, there exist group gradings onI(X) that are not good (see also [13, Example 1]):

Example 1. Let X ={1,2,3} be partially ordered by 1 ≤2 and 1≤3. Then we obtain the algebra

I(X) =

∗ ∗ ∗

∗ 0

 .

LetG=Z2×Z2, and define

A(0,0)= Span{e11, e22+e33}, A(0,1)= Span{e22−e33}, A(1,0)= Span{e12−e13}, A(1,1)= Span{e12+e13}.

This is a well-defined grading on I(X). Moreover, it is not isomorphic to a good grading, since there are not three orthogonal homogeneous idempotents.

(3)

Thus, the study of group gradings on incidence algebras is more complicated than the case of upper triangular matrices. Recall that a grading by a group is equivalent to the action of the group of characters on the algebra, given that the group is abelian and the base field is algebraically closed of characteristic zero (see, for instance, [6, Section 1.4]). The automorphism group of the incidence algebra is known [16, Theorem 7.3.6], and it is more intricate than the automorphism group of the upper triangular matrices (see, for instance, [9]). Hence, the automorphism group ofI(X) suggests the existence of new gradings, other than the gradings on upper triangular matrices.

Another surprising fact is the existence of group gradings on incidence algebras without any homogeneous multiplicative basis (see example in Section 7). We recall that a basisB is multiplicative if for allu, v∈B, there exists a scalarλsuch that λw = uv, for some w ∈B. In contrast, it is clear that any good grading admits a multiplicative homogeneous basis. Thus, every group grading onU Tn admits a multiplicative homogeneous basis.

In the next section, we are going to prove our main result:

Theorem 1. Let F be a field, X a finite poset, and let I(X) be endowed with a G-grading. Assume at least one of the following conditions: charF= 0, charF>

dimI(X), or G is abelian. Then, up to a graded isomorphism, there exist finite abelian subgroupsH1, . . . , Ht⊆G, such that: for eachi= 1, . . . , t,charFdoes not divide |Hi|, andF contains a primitiveexpHi-root of1; and

I(X)∼=

FH1 M1,2 . . . M1,t

FH2 . .. ... . .. Mt−1,t

FHt

 ,

where each Mi,j is a graded (FHi,FHj)-bimodule.

The setE ={1,2, . . . , t}has a natural structure of poset, and it plays an essential role in establishing the isomorphism classes of group gradings on the incidence algebras. The partial order ofE is given byiEj ifMi,j 6= 0 (see Remark 19).

Further, if we assume the grading group abelian, then we can prove a stronger statement concerning the bimodulesMij:

Theorem 2. Assume that G is abelian, and, using the notation of Theorem 1, fix i < j and let Hij = Hi∩Hj. Then, there exist pairwise distinct characters χ(ij)1 , . . . , χ(ij)sij ∈Hˆij, andm1, . . . , msij ∈Mij homogeneous such that

Mij =FHim1FHj⊕ · · · ⊕FHimsijFHj,

where hm` = χ(ij)` (h)m`h, for all h∈ Hij, for ` = 1,2, . . . , sij. Moreover, as a G-graded vector spaces, FHim`FHj ∼= (F(HiHj))[degm`], for each `= 1,2, . . . , sij. Also,sij ≤ |Hij|.

Finally, given that the grading group is abelian, we solve the problem of classify- ing the isomorphism classes of group gradings on the incidence algebras (Theorem 34).

(4)

Example 2. In Example 1, the graded algebra I(X) is isomorphic to F M

0 FZ2

,

whereM is isomorphic toFZ2 as a rightFZ2-module.

2. Preliminaries

2.1. Incidence algebras. We provide the definition of an incidence algebra over a field F. Let (X,≤) be any partially ordered set (poset, for short). Assume that (X,≤) is locally finite, i.e., for allx, y∈X, there exists a finite number ofz∈X such thatx≤z≤y. Define I(X) ={f :X×X →F|f(x, y) = 0,∀x6≤y}. Then I(X) has a natural sum (pointwise sum), and natural scalar multiplication, which givesI(X) a structure ofF-vector space. Forf, g∈I(X), we defineh=f·gas the functionh, such thath(x, y) =P

z∈Xf(x, z)g(z, y). Note that the unique possibly non-zero elements in the previous sum are thez ∈X such that x≤z ≤y; hence, since X is locally finite, the sum is well-defined. So h∈ I(X). It is standard to prove thatI(X), with the defined operations, is an associative algebra. The algebra I(X) is called anincidence algebra.

Note that the defined multiplication onI(X) is similar to the product of matri- ces. Moreover, ifX is totally ordered and containsnelements, then I(X)∼=U Tn. In connection, if (X,≤X) is arbitrary (not necessarilly totally ordered) and finite withnelements, then we can rename the elements ofX ={1,2, . . . , n}in such way thati≤Xj impliesi≤jin the usual ordering of the integers. With this identifica- tion, we see thatI(X)⊂U Tn is a subalgebra. Thus, finite-dimensional incidence algebras are subalgebras of U Tn, generated by matrix units, and containing the diagonal matrices. As we are interested in finite-dimensional incidence algebras, we will assume from now on thatI(X)⊂U Tn.

The incidence algebras are very interesting by its own and, moreover, they are related with other branches of Mathematics. They give rise to interesting and chalenging combinatorial problems. For an extensive theory on incidence algebras see, for instance, the book [16].

2.2. Graded Jacobson radical. It is fundamental in our arguments to guarantee that the Jacobson radical of the incidence algebras are graded ideals.

For this, we have the following result, due to Gordienko:

Lemma 3 (Corollary 3.3 of [7]). Let A be a finite-dimensional algebra, graded by any group G. Suppose charF = 0or charF >dimA. Then the Jacobson radical

J(A)is a graded ideal.

Now, if I(X) has a G-grading, and G is abelian, then the commutator is a homogeneous operation. Moreover, [I(X), I(X)] =J(I(X)) is the Jacobson Radical of the incidence algebraI(X). Hence,J(I(X)) is a graded ideal. Thus, we proved Lemma 4. Let X be finite, and letI(X)be G-graded. Assume at least one of the following conditions: charF = 0, charF > dimI(X), or G is abelian. Then the

Jacobson radical ofI(X)is graded.

(5)

2.3. Group algebras over fields. We include in this subsection a few facts con- cerning the group algebras, which will be essential for our purposes. We do not include the proofs, since the results are classical, or an adaptation of classical proofs when the base field isC, the field of complex numbers.

LetFbe a field,Ga group, and we denote ˆG={χ:G→F group homomorphisms}.

Lemma 5. Assume thatGis abelian finite. The following assertions are equivalent:

(i) G∼= ˆG,

(ii) FG∼=FGˆ (as F-algebras),

(iii) for everyg∈Ghaving orderq,charFdoes not divide q, andFcontains a primitiveq-root of unit.

For instance, we will have the following situation. AG-graded algebraAadmits a subalgebraD⊆Awhich is, simultaneously isomorphic toFmas ordinary algebras, and to FH as graded algebras, for some abelian subgroupH ⊆G. Then, in this case, the fieldFmust satisfy the properties (iii) of the lemma above.

Lemma 6. Let G be a finite abelian group, assume that Gˆ ∼= G, and let Gˆ = {χ1, . . . , χm}. Then the mapψ:FG→Fm, given by

(1) ψ(g) = (χ1(g), . . . , χm(g))

is an isomorphism of algebras. Moreover, every isomorphism of algebrasψ0 :FG→ Fm is given by (1), up to a permutation of the indexes.

3. Group gradings on incidence algebras

Let (X,≤X) be a finite poset, and let n = |X|. We fix a field F and G any group with multiplicative notation and neutral element 1, and we assume one of the conditions of Lemma 4. As mentioned above, we can rename the elements ofX asX ={1,2, . . . , n}, andi≤X jimpliesi≤jin the usual ordering of the integers.

Hence, we can identifyI(X)⊂U Tn. If 1≤i≤j≤nandx∈I(X),x(i, j) denotes the entry (i, j) of the elementx.

By direct computation, it is easy to see that for anyM ∈I(X), which is invertible inU Tn, we haveM−1∈I(X).

Lemma 7. Lete∈I(X)be an idempotent. Then there existsM ∈I(X)such that M−1eM is diagonal.

Proof. The proof is similar to the proof of Lemma 1 of [17].

We assume that I(X) acts on the left on the n×1 matrices. For each i = 1,2, . . . , n, let

ei =

 0 0 ... 1 ... 0

 ,

having non-zero entry only in the position (i,1). One, and only one, betweene·ei or ei −e·ei will have non-zero entry at (i,1), where e·ei is the usual matrix

(6)

multiplication. Let fi be either e·ei, or ei−e·ei, the unique matrix with the non-zero entry in (i,1).

LetM be the matrix such that the columns aref1, f2, . . . , fn. By construction, M ∈I(X). Moreover, since the diagonal entries ofM are non-zero, it is invertible.

Also,f1, f2, . . . , fn are eigenvectors ofe. Thus,M−1eM is diagonal.

A classical result says that a set of diagonalizable matrices, which pairwise com- mutes, is simultaneously diagonalizable. We have a similar result forI(X):

Lemma 8. Let e1, . . . , es ∈ I(X) be orthogonal idempotents. Then there exists M ∈I(X)such thatM−1eiM is diagonal, for eachi= 1,2, . . . , s.

Proof. The proof will be by induction ons. Ifs= 1, then the result follows from the previous lemma. So, assumes >1, andf1, . . . , fn is a basis such thate1, . . . , es−1 are diagonal, and the matrixM having f1, . . . , fn as its columns lies inI(X).

Fixj ∈ {1, . . . , n}. Ifesfj = 0, then fj is an eigenvector of es with eigenvalue 0, and there is nothing to do. If there is i < s such that eifj =fj, then esfj = eseifj = 0, hence we are in the previous case. So, assume that esfj 6= 0 and eifj = 0, for all i < s. As in the previous lemma, we can change fj to esfj or fj−esfj, in such a way that this new fj still form a basis{f1, . . . , fn}of Fn. By construction,eifj = 0, for this new fj, andesfj is either 0 orfj. Continuing the process, we obtainf1, . . . , fs such that the matrixM, having these elements as its columns, is inI(X), and everyei is diagonal in this basis.

Lemma 9. Let x∈ U Tn be such that x=e+b, where e2 =e6= 0 is a diagonal matrix, and b is a strict upper triangular matrix. Then there exists a non-zero idempotent inSpan{x, x2, . . .}. In other words, there exists a non-zero polynomial p(λ)∈F[λ], such that p(0) = 0 andp(x)is a non-zero idempotent.

Proof. Let ¯Fan algebraic closure of F. By Jordan canonical form, there exists an invertible matrixmsuch that

mxm−1=

1 ∗ 0 · · · 0 . .. ... ... 1 0 · · · 0

0 ∗

. .. 0

 .

With an adequates >1, we obtain mxsm−1=

J 0 0 0

,

whereJ is a triangular matrix, with 1’s in the diagonal. Now, there exists a poly- nomialp(λ) such thatp(J) = 1, the identity matrix. Indeed, let q(λ) =P

i≥0aiλi be the characteristic polynomial ofJ. By the Cayley-Hamilton Theorem,q(J) = 0.

Moreover, sinceq(0) =a0= detJ = 1, we obtain thatp(λ) =P

i>0−aiλi satisfies p(J) = 1.

Lete0=p(xs), thenme0m−1=

1 0 0 0

, hencee02=e0, as desired.

(7)

Fix anyG-grading onI(X). Define

E(X) ={e∈I(X) non-zero homogeneous idempotent}.

Of course it is non-empty since 1∈E(X). Fore1, e2∈E(X), lete1e2denote (the non-necessarily partial order)e1I(X)⊆e2I(X) (usually,is called a preordering).

Lete∈E(X) be a minimal element according to(sinceI(X) has finite dimension, we can always find such minimale). Note that dege= 1, sincee2=e. By Lemma 7, we can assumeediagonal, up to a graded isomorphism. The algebraB=eI(X)e has a naturalG-grading, induced from the grading of I(X). We will prove thatB is a (commutative) graded division algebra.

Lemma 10. For every homogeneous x∈B,x /∈J(B),degxhas finite order.

Proof. Note that, ifx /∈J(B), thenxhas at least one non-zero entry in the diagonal.

So,xm∈/ J(B) for allm∈N. In particular, x, x2, x3, . . . are non-zero elements. If degxhas infinite order, then the elementsx, x2, . . .have all different degree, thus they are linearly independent. However, B has finite dimension, so we obtain a contradiction. As a conclusion, degxhas finite order.

Recall, from Lemma 4, that J(B) is graded. Thus B/J(B) has a natural G- grading, induced from the grading onB.

Lemma 11. (i) The algebra B/J(B) is a commutative graded division alge- bra.

(ii) WriteB/J(B) =L

g∈Gg. Thendim ¯Bg≤1, for allg∈G.

Proof. If dim ¯B1 is not 1, then we claim that there would exists an idempotent e0 ∈B¯1 such that e0 e. Indeed, let ¯x∈B¯1 be a non-zero homogeneous element, not a scalar multiple ofe, moduloJ(B). Then there exists a homogeneousx∈Bof degree 1, having non-zero entries in the diagonal. We can writex=α1e12e2+

· · ·+αses+b, whereb∈J(B),e1, e2, . . . , esare diagonals orthogonal idempotents (sum of diagonal matrix units), andα1, . . . , αs∈Fare non-zero elements such that αi 6= αj for all i 6= j. Since x has non-zero entry in the diagonal, s > 0; and since xand e are linearly independent modulo J(B), we have necessarillys > 1.

The powers x, x2, . . . , xsare homogeneous elements of degree 1. By Vandermonde argument, we conclude thaty=e1+b0 is a linear combination ofx, x2, . . . , xs, for some b0 ∈J(B), hence y is homogeneous of degree 1. Using Lemma 9, we obtain a homogeneous idempotent e0, which is a linear combination of y, y2, . . .. Thus e0 ∈ B. Moreover, we have e0I(X) ( eI(X). But e is minimal, so we obtain a contradiction. As a conclusion, dim ¯B1= 1.

Now, let ¯x∈B/J(B) be non-zero and homogeneous. Thus, ¯xhas non-zero entry in the diagonal. Hence, ¯xm6= 0, for allm∈N. By Lemma 10, ¯xm∈B¯1, for some m∈N. This means ¯xm=λ¯e, for someλ6= 0. As a consequence, ¯xis invertible.

Thus, B/J(B) is a graded division algebra. Moreover, since B/J(B) is a sum of copies ofF, as an ordinary algebra, then it is necessarily commutative, proving (i).

Since dim ¯B1= 1 andB/J(B) is a graded division, then dim ¯Bg≤1, for allg∈G,

thus proving (ii).

Corollary 12. The algebraBis semisimple as an ordinary algebra, that isJ(B) = 0. In particular,B is a commutative graded division algebra with unite.

(8)

Proof. Let J = J(B) and ¯B = B/J. Note that J/J2 is a graded ¯B-module.

The right annihilator AnnrB¯(J/J2) is a homogeneous subspace of ¯B. However, no nonzero homogeneous element of ¯B annihilates a nonzero homogeneous element of J/J2, since ¯B is a graded division algebra, by previous lemma. Hence,J(B) = 0.

Since, by previous lemma,B/J(B) is a commutative graded division algebra, then B is a commutative graded division algebra as well.

From now on,ewill always denote a minimal idempotent as before. Denote by D(e) ={i∈X |e(i, i)6= 0}.

Corollary 13. The elements of D(e) are not pairwise comparable.

Proof. If not, supposei < j, withi, j∈D(e). Recall thatB=eI(X)e∼=I(Y), for some posetY (whereY =D(e)). Theneij ∈B. Moreover, one haseij ∈J(B), so

J(B)6= 0, contrary to Corollary 12.

Now, ife6= 1, consider the set

E1(X) ={e0∈I(X) idempotent|e0I(X)⊂(1−e)I(X)}.

We remark that E1(X) 6=∅, since 1−e ∈ E1(X). We claim that we can choose the idempotents belonging to E1(X) so that they commute with e. Indeed, if e0 ∈ E1(X), then ee0 = ee02 ∈ e(1−e)I(X) = 0. Also, let f1 = e0 −e0e. Then f12 =f1, f1e =ef1 = 0 and f1I(X) =e0I(X). Thus,f1 ∈E1(X) is the required element. Let e1 = e, and chose a homogeneous idempotent e2 ∈ E1(X) which is minimal with respect to, and orthogonal toe1.

Proceeding by induction, we assume that we have a set e1, . . . , er of minimal homogeneous orthogonal idempontents. Then we define

Er(X) ={e0∈I(X) idempotent |e0I(X)⊆(1−e1· · · −er)I(X)}.

As before, we can find a minimal homogeneous idempotenter+1∈Er(X) which is orthogonal toe1, . . . , er. We can continue the process, and, since dimI(X) is finite, it ends within finitely many steps. Lete1, e2, . . . , et be the homogeneous idempo- tents, which are minimal with respect to, obtained by the previous construction.

By Lemma 8, we can assumee1, e2, . . . , etdiagonal.

Definition 14. The diagonal homogeneous idempotents e1, . . . , et will be called theminimal idempotents. We denoteE ={e1, . . . , et}.

We denote Di =eiI(X)ei which is, according to Corollary 12, a commutative graded division algebra. Moreover,Mij=eiI(X)ej is a graded (Di, Dj)-bimodule.

The triple (Di, Dj, Mij) constitutes a triangular algebra.

If ei and ej are two minimal idempotents, we denote ei E ej if there exist x∈D(ei) andy∈D(ej) such thatx≤y.

Lemma 15. The relation Eis a partial order inE.

Before we prove Lemma 15, we need to investigate the relation between each Mij and the poset.

Recall that the structure of graded modules over graded division algebras is similar to the structure of vector spaces over division rings (see, for instance, [6, page 29]). More precisely, let D be a G-graded division algebra and let M be a

(9)

graded right D-module. Then M ∼= L

αMα, where each Mα = [hα]D, for some hα∈G. Here, ifD=L

g∈GDg, then

[h]D=M

g∈G

[h]Dg, where[h]Dg=Dh−1g.

In particular, every graded rightD-module is free. A similar discussion is valid for graded leftD-modules.

Let ei and ej be minimal idempotents. For each x ∈ D(ei), we define the link of x to ej as the non-negative integer `(x, ej) = |{y ∈ D(ej) | x ≤ y}|.

Similarly, we define the link of ei to y, denoted `(ei, y). The link of ei to ej is

`(ei, ej) =|{(x, y)|x≤y, x∈D(ei), y∈D(ej)}|.

Lemma 16. Let ei, ej be minimal idempotents. Then, for any x ∈ D(ei), y ∈ D(ej),

`(ei, ej) =|D(ei)|`(x, ej) =`(ei, y)|D(ej)|.

Proof. Let Mij = eiI(X)ej, which is a free G-graded left Di-module. We shall prove that`(x, ej) = dimDiMij, for anyx∈D(ei).

On one hand, asF-linear space, we have

Mij = M

(x,y)∈D(eiD(ej);x≤y

Fexy.

Thus,`(ei, ej) = dimFMij = dimFDidimDiMij. On the other hand, letv1, . . . , v`

be aDi-basis ofMij, where` >0 (the case`= 0 is trivial).

Claim. Fork∈ {1, . . . , `}, letw∈Divk, and asumew(x, y) = 0, butvk(x, y)6= 0.

Thenw(x, y0) = 0, for ally0∈D(ej).

Indeed, in this case,B={exxvk |x∈D(ei)}is a set of |D(ei)| F-linearly inde- pendent elements. Since dimFDivk= dimFDi=|D(ei)|,Bis a (non-homogeneous) F-vector space basis ofDivk. So, writew=P

x∈D(ei)λxexxvk. Ifvk(x, y)6= 0, but w(x, y) = 0, then necessarilyλx= 0. This impliesw(x, y0) = 0, for ally0∈D(ej).

As a conclusion, sincev1, . . . , v`generateMijas aDi-linear space, every element ofMij is a linear combination of elements inDiv1, . . . , Div`. By the claim, in order to be possible to fill all the entries of the coordinates of type (x,·) in Mij, it is necessary that`≥`(x, ej). Hence,`≥max{`(x, ej)|x∈D(ei)}. And, therefore,

dimFDidimDiMij=`(ei, ej) = X

x∈D(ei)

`(x, ej)≤ |D(ei)|`= dimFDidimDiMij which implies`(x, ej) = dimDiMij, for any x∈D(ei) and we are done.

Now, we can prove the lemma:

Proof of Lemma 15. It is clear thatE is reflexive and transitive. So, let ei 6= ej

(thus, by construction,D(ei)∩D(ej) =∅), and assume thatei Eej, andej Eei. Then, by Lemma 16, every x ∈ D(ei) is ≤ to at least one y ∈ D(ej). In the same way, every y ∈ D(ej) is ≤ to at least one x ∈ D(ei). Hence, we can find x1, x2∈D(ei), andy∈D(ej) such thatx1< y < x2, so x1< x2. This contradicts

Corollary 13. Hence,Eis antisymmetric.

As a consequence of Lemma 15, we can rename the elements of X, in such a way that D(e1) = {1,2, . . . , i1}, D(e2) = {i1+ 1, i1+ 2, . . . , i2}, ..., D(et) = {it−1+ 1, . . . , n}. Since eiI(X)ej = Span{ek` | k ∈ D(ei), ` ∈ D(ej), k < `}, we

(10)

have dimeiI(X)ej ≤ dimDidimDj. Moreover, as graded vector space, I(X) = P

i,jeiI(X)ej. In this way, we obtain the proof of the Theorem 1.

Corollary 17. Up to a graded isomorphism, there exists commutative graded di- vision algebras D1, D2, . . . , Dt, all of them having all homogeneous subspaces with dimension at most 1, such that

I(X)∼=

D1 M1,2 . . . M1,t

D2 . .. ... . .. Mt−1,t

Dt

 ,

where eachMi,jis a graded(Di, Dj)-bimodule of dimension at mostdimFDidimFDj. IfD is a commutative graded division algebra, where each homogeneous compo- nent has dimension at most 1 andD∼=FdimD, thenD∼=FH, whereH = SuppD.

Indeed, a set of nonzero homogeneous elements{Xu|u∈H}, where degXu=uis a vector space basis ofD. Also,XuXv =σ(u, v)Xuv, for some (nonzero)σ(u, v)∈F. Since D is associative, we obtain that σ is a 2-cocycle. Thus, σ ∈ Z2(H,F×), and D ∼=FσH (the twisted group algebra). To complete the proof, we need the following result, communicated by M. Kochetov:

Proposition 18. Let H be a finite abelian group and σ∈ Z2(H,F×). IfFσH ∼= F|H|, thenFH ∼=F|H|.

Proof. Given a finite-dimensional unital commutative algebraA overF, let N(A) denote the number of algebra homomorphisms fromAto F.

Claim. Ais a direct sum of copies ofFif and only ifN(A) = dimA.

Indeed, (⇒) is obvious, and (⇐) follows from linear independence of distinct homomorphisms.

Note that bothN and dim are multiplicative over tensor products of algebras.

Now, writeH =hh1i × · · · × hhri, direct product of cyclic groups, where eachhi

has ordermi. Then, we notice that

FσH ∼=F[x1]/(xm11−a1)⊗ · · · ⊗F[xr]/(xmrr−ar), whereai=Qmi−1

j=1 σ(hi, hji)∈F. In the same way,FH ∼=Nr

i=1F[xi]/(xmi i−1).

From the remarks above,FσH is a direct sum of copies ofFif and only if each F[xi]/(xmi i−ai) is a direct sum of copies ofF. The former condition is equivalent to Fcontainsmi distinct mi-roots ofai; and this implies thatF contains a primitive mi-th root of 1. Thus, each F[xi]/(xmi i−1) is a direct sum of F’s. Hence, by the remarks above,FH is a direct sum of copies ofF. As a consequence, FσH ∼=FH as ungraded algebras. From [11, Corollary 1.2], we obtain thatαis a coboundary map. Thus,FσH∼=FH as graded algebras.

Remark 19. Consider the notation of Corollary 17. Letei ∈Di be the unity. Set E ={1,2, . . . , t}. We define the partial orderEonE by the following (equivalent) conditions (see Lemma 15):

(1) iEj, (2) eiEej,

(3) there existsx∈D(ei) andy∈D(ej) such thati≤j,

(11)

(4) Mij 6= 0.

Note that we can identify the present definition ofE with that given in Definition 14. We callE theassociated poset of the graded algebraI(X).

Note that we have the following equality:

|Hi|= dimDi=|D(ei)|.

4. On the structure of graded bi-vector spaces

In this section, we shall investigate graded bimodules. We are interested in the special case whereGis an abelian group,H1, H2⊆Gare finite subgroups, andV is aG-graded (FH1,FH2)-bimodule. We will conclude that, in this situation,V is a FH-vector space, whereH =H1H2. Hence, the bimodules have a nice description in this case.

The notion of tensor product will play a fundamental role in this section. Recall that, ifR is anyF-algebra, M, N areF-vector spaces,M is a rightR-module, and N is a leftR-module, thenM⊗RN is a uniquely definedF-vector space, satisfying a universal property. It is spanned bym⊗n, wherem∈M, andn∈N, and they satisfymr⊗n=m⊗rn,r∈R. For the special case whereM andN are algebras, thenM⊗RN is an algebra as well.

We are interested in the following special case: letR, A1, A2 beF-algebras, and assume ϕi : R →Ai homomorphism of algebras, for i = 1,2. Then A1 becomes a rightR-module by a·r =aϕ1(r), and, similarly, A2 becomes a left R-module.

So we can construct the tensor product A1RA2. Since ϕ1 and ϕ2 will play a somewhat significant role, we use the notation (A1RA2, ϕ1, ϕ2). In this case, A1RA2 is spanned bya1⊗a2, for ai ∈βi, whereβi is any basis of Ai; also, we haveb1ϕ1(r)⊗b2=b1⊗ϕ2(r)b2, for anyr∈R,b1∈A1, b2∈A2.

Now, assume further that R, A1, A2 are G-graded algebras, and ϕ1, ϕ2 are G- graded homomorphisms of algebras. SoA1 andA2 areG-graded R-modules. The algebraA1RA2 admits a natural structure ofG-grading, where the homogeneous component of degreeg∈Gis spanned bya1⊗a2,ai∈Ai homogeneous satisfying dega1dega2 = g. Moreover, A1RA2 becomes a G-graded algebra. In what follows, we will see that, in the case of group algebras, the just defined grading has a nicer way to be constructed.

Lemma 20. Let G be an abelian group, and H1, H2⊆Gbe finite subgroups. Let H =H1H2, andH=H1∩H2. Let ιj :FH →FHj be G-graded monomorphisms, forj= 1,2. Then, as graded algebras,

(FH1FHFH2, ι1, ι2)∼=FH.

Proof. First we note thatFH1FHFH2has a natural well-definedG-grading, given by degh1⊗h2=h1h2, where h1∈H1, h2∈H2.

Now, since every homogeneous component ofFHj has dimension at most 1, we see that an inclusionιj(h) =χj(h)h, for eachh∈ H, for someχj(h)∈F. Moreover, h 7→ χj(h) ∈ F is a character of H. Consider the character given by χ−1j . It is known that every character ofHcan be extended to a character of Hj. So, let ¯χj be an extension ofχ−1j .

(12)

Letϕ:FH1FHFH2→FH be defined by ϕ(h1⊗h2) = ¯χ1(h1) ¯χ2(h2)h1h2. If h∈ H, then

ϕ(h1ι1(h)⊗h2) =χ1(h) ¯χ1(h1h) ¯χ2(h2)h1hh21(h) ¯χ1(h) ¯χ1(h1) ¯χ2(h2)h1hh2, ϕ(h1⊗ι2(h)h2) =χ2(h) ¯χ1(h1) ¯χ2(hh2)h1hh22(h) ¯χ2(h) ¯χ1(h1) ¯χ2(h2)h1hh2. Sinceχ1(h) ¯χ1(h) =χ2(h) ¯χ2(h) = 1, we obtain that ϕis well-defined.

SinceGis abelian,ϕis an onto homomorphism of graded algebras.

LetPm

i=1λigi⊗hi∈Kerϕ. As Kerϕis a graded subspace, we can suppose that g1h1 = . . . = gmhm. Moreover, we can assume that h1, . . . , hm are FH-linearly independent. Thus, gihi = gjhj implies hih−1j ∈ H, which is a contradiction if m >1. Hence,m≤1, and this is sufficient to conclude that Kerϕ= 0.

Lemma 21. LetGbe an abelian group,H1, H2⊆Gfinite subgroups,H=H1∩H2. Let V be a finite-dimensional G-graded (FH1,FH2)-bimodule. Then, there exist m1, . . . , ms∈V homogeneous, and χ1, . . . , χs∈H, such thatˆ

V =FH1m1FH2⊕ · · · ⊕FH1msFH2, andhmii(h)mih, for alli= 1, . . . , s,h∈ H.

Proof. Let Wg be the homogeneous component of degree g of V. Given h ∈ H, letLh, Rh :V →V denote the maps Lh(m) = hm, Rh(m) =mh. Then Lh is a linear isomorphism fromWg to Whg. Also,Rh−1 : Wgh →Wg is an isomorphism as well. Thus,LhRh−1 ∈GL(Wg). It is elementary to see that this defines a linear representationh∈ H →LhRh−1 ∈GL(Wg).

Since H is abelian, the representation is completely reducible, and each irre- ducible representation has degree 1. So, we can find a vector space basisv1, . . . , vr∈ Wg, and characters χ1, . . . , χr ∈ H such that LhRh−1(vi) = χi(h)vi, for all i = 1, . . . , r. In particular, this giveshvii(h)vih, for allh∈ H.

It is clear that every componentWg0, whereg0∈H1gH2belongs toFH1v1FH2

· · · ⊕FH1vrFH2. Thus, we can takeg2 ∈/ H1gH2 and repeat the process. Since

dimV <∞, the process ends.

In the notation of the previous lemma, giveng1∈H1 andg2∈H2, we have g1hmig2i(h)g1mihg2,∀h∈ H.

Lemma 22. Assuming the same hypotheses of the previous lemma, let χ ∈ H,ˆ and V = FH1mFH2, where hm = χ(h)mh, for all h ∈ H. Then V is a graded FH1FHFH2-module.

Proof. Consider the inclusions

ι1:h∈FH →h∈FH1, ι2:h∈FH →χ(h)h∈FH2.

Define the left action of FH1FHFH2 on V by g1⊗g2m =g1mg2. It suffices to prove that this action is well-defined. For, assumeg1∈H1,g2∈H2,h∈ H. Then g1h⊗g2=χ(h)g1⊗hg2, and

(g1h⊗g2)m=g1hmg2=χ(h)g1mhg2= (g1⊗χ(h)hg2)m.

(13)

Recall that, ifR andS are any rings,ϕ:R→S is a ring homomorphism, and M is an S-module; then M becomes an R-module if we set m·r = mϕ(r), for m∈M andr∈R. We summarize the results of this section.

Corollary 23. Let G be an abelian group, H1, H2 ⊆ G finite subgroups, H = H1∩H2,H =H1H2. LetV be a finite-dimensionalG-graded(FH1,FH2)-bimodule.

Then, there exist homogeneousm1, . . . , ms∈V and charactersχ1, . . . , χs∈Hˆ such that

(2) V =FH1m1FH2⊕ · · · ⊕FH1msFH2,

wherehm``(h)m`h, for eachh∈ H. Moreover, for eachi, one hasFH1miFH2∼= (FH)[degmi], as graded vector spaces.

Proof. Lemma 21 says that we can find homogeneousm1, . . . , ms∈V such that V =FH1m1FH2⊕ · · · ⊕FH1msFH2.

By Lemma 22, everyFH1miFH2has a structure of graded (FH1FHFH2, ιi1, ιi2)- module. By Lemma 20, there exists a graded isomorphismϕi :FH →(FH1FH

FH2, ιi1, ιi2). Hence, FH1miFH2 has a structure of graded FH-module, and it is 1-dimensional, since it isFH-spanned bymi. This concludes the proof.

Remark 24. It should be noted that Corollary 23 is no longer valid when the grading group is not abelian. Indeed, letH be a finite abelian group,C2={1, w}the cyclic group of order 2, and letG=H∗C2 be the free product ofH andC2. Consider V =FH⊗FFH, and define theG-grading onV byV =L

g∈GVg, where Vh1wh2 = Span{h1⊗h2}, h1, h2∈H.

ThenV is aG-graded vector space, andV has a structure ofG-graded (FH,FH)- bimodule. For every nonzero homogeneousv∈V, we haveFHvFH =V; and it is not isomorphic to a shift of FH. Moreover,V is not a G-graded left FH⊗FFH- module.

Now, we are ready to prove Theorem 2.

Proof of Theorem 2. Fixi, j. We consider the triangular algebra A=

Fmi Vij

Fmj

,

where m`=|H`|, which is graded isomorphic to

FHi Mij

FHj

. Lete`∈H` be the unit. ThenA = (ei+ej)I(X)(ei+ej), and it is the incidence algebra of the subposetY =D(ei)∪D(ej)⊆X.

LetH=Hi∩Hj, ˆH={χ1, . . . , χm}, and assume ˆH= ˆHi∩Hˆj. Write Hˆ`(`)1 Hˆ∪˙. . .∪χ˙ (`)n

`

H,ˆ wherem`=mn`, for`∈ {i, j}(we can assumeχ(`)1 = 1).

By Lemma 6, up to renaming the entries, the graded isomorphism φ : FHi⊕ Mij⊕FHj→A is such that, forh1∈Hi,h2∈Hj,

(3)

φ(h1+h2) = (χ(i)1 χ1(h1), . . . , χ(i)niχm(h1)) 0

(j)1 χ1(h2), . . . , χ(j)njχm(h2))

! .

(14)

LetMij=FHim1FHj⊕ · · · ⊕FHimsFHj, with charactersµ1, . . . , µs, as in (2). We only need to prove that µ1, . . . , µs are pairwise distinct characters. By Corollary 23, dimFHim`FHj =|HiHj|, so φ(m`)∈Vij has at least|HiHj|non-zero entries.

To make notation easier, we will identify the elements ofD(ei)∪D(ej) with the respective characterχ(`)r χs, as in (3). So, for instance, to denote the entry (1,1) of φ(h1+h2) in (3), we will writeφ(h1+h2)(χ(i)1 χ1, χ(i)1 χ1).

Claim. For each χ0 ∈ Hˆi, we can find unique χ0k

1, . . . , χ0k

nj ∈ Hˆ such that φ(m`)(χ0, χ(j)0r χ0k

r)6= 0, r= 1,2, . . . , nj.

Indeed, let µ ∈ Hˆi and let µ0 ∈ Hˆj be such that φ(m`)(µ, µ0) 6= 0. Now, if µ00 ∈ Hˆj is such that φ(m`)(µ, µ00) 6= 0, then µ0 and µ00 cannot be in the same left coset, unless they coincide, as we will prove now. For eachh∈ H,hm`h−1= µ`(h)m`. On the other hand, hφ(m`)h−1(µ, µ0) =µ(h)φ(m`)(µ, µ00(h−1), and a similar relation holds forµ00. This impliesµ0|H= (µ`µ−1)|H00|H. Hence, ifµ0 andµ00 are in the same coset of ˆH, thenµ000. So, for eachµ∈Hˆi, there exist at most|Hj|/|H|elementsµ0 such thatφ(m`)(µ, µ0)6= 0. Thus,φ(m`) has at most

|Hi||Hj|/|H|non-zero entries.

This implies that φ(m`) has exactly |HiHj| non-zero entries. Moreover, the non-zero entries of theφ(m1), . . . , φ(ms) are disjoint.

Now, let`1, `2 withµ`1`2, and fixχ0∈H ⊆ˆ Hˆi, and letχ01, χ02∈H ⊆ˆ Hˆj be such thatφ(m`k)(χ0, χ0k)6= 0,k= 1,2. For eachh∈ H, again we havehm`kh−1= µ`k(h)m`k. On the other hand, (recall thatχ(i)1 = 1,χ(j)1 = 1)

(hφ(m`k)h−1)(χ0, χ0k) =χ0(h)χ0−1k (h)φ(m`k)(χ0, χ0k).

Hence,χ010µ−1`

k02. Thus, this impliesm`1 =m`2. We also have the following immediate consequence of Corollary 23:

Corollary 25. Let Gbe abelian, and leti < j. Then:

(i) ifHi∩Hj={1}, thenMij is either 0 orFHij, whereHij ∼=Hi×Hj. (ii) if Hi ⊆ Hj (in particular, Hi = Hj), then Mij is a graded FHj-vector

space, and 0≤dimFHjMij ≤ |Hi|.

Given a graded (FH1,FH2)-bimoduleMwith decomposition given by (2), denote [M] = [(χ1, h1), . . . ,(χs, hs)], whereh` = degm`. We finish this section showing that the knowledge of those characters determines the isomorphism classes of the bimodules, or certain triangular algebras.

Lemma 26. Let V and V0 be G-graded (FH1,FH2)-bimodules, and denote[V] = [(χ1, h1), . . . ,(χs, hs)], [V0] = [(χ01, h01), . . . ,(χ0s0, h0s0)]. Then, as G-graded bimod- ules, V ∼=V0 if, and only ifs=s0 and there existsσ∈Sssuch that

h`≡h0σ(`) (modH1H2), χ`0σ(`), ∀`= 1,2, . . . , s.

Proof. Write V = FH1m1FH2 ⊕ · · · ⊕ FH1msFH2, V0 = FH1m01FH2 ⊕ · · · ⊕ FH1m0s0FH2, as in (2). Let ψ : V → V0 be a graded isomorphism of bimod- ules. Then ψ is an isomorphism of H1∩H2 representations, where the action is given by Lh◦Rh−1. Let Vχ =P

χiFH1miFH2. Note that Vχ is the sum of all subrepresentations of V isomorphic toχ. So, ψ(Vχ) = P

χ0iFH1m0iFH2 =:Vχ0.

(15)

Moreover,ψ:Vχ →Vχ0 is a graded isomorphism ofF(H1H2)-modules. This proves one implication.

The converse is immediate.

Lemma 27. Let V and V0 be G-graded (FH1,FH2)-bimodules, and denote[V] = [(χ1, h1), . . . ,(χs, hs)],[V0] = [(χ01, h01), . . . ,(χ0s0, h0s0)]. Then, asG-graded algebras,

A:=

FH1 V FH2

∼=

FH1 V0 FH2

=:A0 if and only if s=s0 and there existχ∈H\1∩H2 andσ∈Ss such that

h`≡h0σ(`) (mod H1H2), χ`=χχ0σ(`), ∀`= 1,2, . . . , s.

Proof. The converse is an elementary computation.

Letψ:A→A0 be a graded isomorphism of algebras. Then we have an induced graded isomorphism ¯ψ : A/V → A0/V0, moreover, ¯ψ(FHi) = FHi. So, there exists χi ∈ Hˆi such that ¯ψ(hi) = χi(hi)hi, for each hi ∈ Hi, for i = 1,2. Write V0=FH1m01FH2⊕ · · · ⊕FH1m0s0FH2, and consider the mapϕ:A0 →A0 given by

ϕ

h1 h01m0ih02 h2

=

χ−11 (h1)h1 χ−11 (h01−12 (h02)h01m0ih02 χ−12 (h2)h2

.

It is elementary to prove that ϕis a graded isomorphism of algebras. Moreover, givenhi∈Hi andm∈V, we have

ϕψ(h1mh2) =h1ϕψ(m)h2.

Thus, ϕψ : V → ϕ(V0) is a graded isomorphism of bimodules. Consider the product of the restriction of the characters, χ =χ1|H1∩H2 ·χ2|H1∩H2. Note that, if [V0] = [(χ01, h01), . . . ,(χ0s0, h0s0)], then the decomposition of ϕ(V0) is [ϕ(V0)] = [(χχ01, h01), . . . ,(χχ0s0, h0s0)]. Thus, the result follows from Lemma 26.

5. The isomorphism problem

Assume that the conditions of Theorem 1 hold. Consider aG-grading onI(X) and let E be the associated poset. Denote bye /ce0 ife, e0 ∈ E, and e0 is a cover fore; that is,e6=e0,eEe0, and ifeEf Ee0, then eitherf =eorf =e0.

Note that the Jacobson RadicalJ =J(I(X)) coincides withP

eEe0,e6=e0eI(X)e0. Also,J2 is the sum of alleI(X)e00 where there exists a chain of different elements e/e0/e00;J3is the sum of alleI(X)e000, wheree/e0/e00/e000, and so on. Givene, e0 ∈ E, e E e0, we can always construct a chain of elements containing only subsequent covers, starting withe, and ending withe0; however, we can have two different such chains having different number of elements. Nonetheless, ife /ce0, thene / e0 is the unique chain of subsequent covers linking eande0. LetJ1=L

e/ce0eI(X)e0. The previous discussion is the proof of the following:

Lemma 28. Let I(X)be G-graded, and letJ1 be as above.

(i) The restriction of the natural projection π: J1 →J/J2 is a graded linear isomorphism.

(ii) J =J1⊕P

m>1J1m.

Referências

Documentos relacionados

Table 6 – TTAs extraction yield and TTAs content of the Soxhlet and supercritical extractions. The poor performance of SFE assays for E. globulus leaves clearly contrast with the

A Universidade Aberta de Portugal, a Escola Superior de Educação do Instituto Politécnico de Bragança, a Universidade de Salamanca e a Faculdade de Educação da Universidade

É preocupante que cidades relativamente grandes não tenham ainda dado à devida importância a criação da comissão. Outra questão que a entrevistada 3 enfatiza o caráter

Em sua tese, Andrade Filho (2011) exprime que as experiências de movimento corporal das crianças tendem a ser interditadas pela cultura institucional, evidenciando que há

Os autores encontraram uma redução nos escores de problemas atencionais do CBCL – Child Behavior Checklist - assim como nos escores de interação disfuncional entre pais e filhos,

Para conhecer a seleção completa, com todas cores e materiais, visite a área de IKEA das Crianças e a exposição de Móveis de Exterior, bem como www.IKEA.pt.. *Este artigo pode

São Leopoldo era um centro comercial, com “contatos e experiências sociais e étnicas diversas” (MOREIRA; MUGGE, 2014, p. Ambos pararam na casa de comércio do imigrante

Falo também para todos aqueles que não tendo votado Aliança Democrática no dia 5 de Outubro querem em Portugal uma democracia plena, sem Conselho da Revolução, sem os comunistas