ebook img

Equivalence of blocks for the general linear Lie superalgebra PDF

0.19 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Equivalence of blocks for the general linear Lie superalgebra

EQUIVALENCE OF BLOCKS FOR THE GENERAL LINEAR LIE SUPERALGEBRA 3 1 SHUN-JENCHENG, VOLODYMYRMAZORCHUK,ANDWEIQIANGWANG 0 2 Abstract. We develop a reduction procedure which provides an equivalence (as highest c e weight categories) from an arbitrary block (defined in terms of the central character and D the integral Weyl group) of the BGG category O for a general linear Lie superalgebra to an integral block of O for (possibly a direct sum of) general linear Lie superalgebras. We also 8 establish indecomposability of blocks of O. 1 ] T R 1. Introduction . h Recently, Brundan’s Kazhdan-Lusztig-type conjecture [Br] on irreducible and tilting char- t a acters of the general linear Lie superalgebra gl(m|n) in terms of canonical bases of Fock spaces m has been established by the first and third authors jointly with Ngau Lam [CLW]. In these [ papers only the integral part of category O for gl(m|n) was considered. A natural next step is to understand the structure of a block for an arbitrary weight and ask for a possible reduction 2 v via equivalence to an integral weight block. In the setting of semisimple Lie algebras, this was 4 achieved by Soergel [Soe] by establishing an equivalence between a general block and an inte- 0 gral block associated to the so-called integral Weyl group. However, the approach of Soergel 2 1 is not directly applicable to Lie superalgebras, since blocks of O for Lie superalgebras are not . controlled by Coxeter groups in general. 1 0 There is an alternative approach in the case of Lie algebras developed by Mathieu [Mat] 3 and Kashiwara-Tanisaki [KT]. It uses the localization of the universal enveloping algebra with 1 respect to certain root elements and polynomial extensions of conjugation automorphisms for : v this localization. In this note we develop yet another approach to this problem based on twist- i X ing functors from [Ar]. The first step in the definition of twisting functors uses, similarly to r the approach of Mathieu and Kashiwara-Tanisaki, the localization of the universal enveloping a algebra with respect to certain root elements. The difference is in the second step where, instead of polynomial extensions of conjugation automorphisms for the localization, one uses Arkhipov’s twisting trick from [Ar]. These twisting functors admit straightforward general- izations to Lie superalgebras and so allow us to establish, for Lie algebras and superalgebras of type A, an explicit equivalence (as highest weight categories) between a block of O for an arbitrary weight and an integral block of O for a certain subalgebra, where the equivalence is built upon parabolic induction and twisting functors. Let us explain our approach in more detail. Let g be a Lie algebra of ADE type. For a weight λ, consider the corresponding integral subalgebra g of g (see (2.1)). If g is a Levi [λ] [λ] subalgebra, then the usual parabolic induction functor gives the desired equivalence of blocks, Key words and phrases. General linear Lie superalgebra, BGG category, block decomposition. 2010 Mathematics Subject Classification. Primary 17B10. 1 2 CHENG,MAZORCHUK,ANDWANG with the block for (the semisimple part of) g being integral. Next we show that a twisting [λ] functorassociated toasimplerootαsuch thathλ,α∨i 6∈Zis always anequivalence (as highest weight categories) of blocks. It remains to observe that for a type A Lie algebra it is always possible to find a sequence of such twisting functors that connects an arbitrary block with a block for which the integral subalgebra is Levi. Now we are ready to generalize the above approach to the Lie superalgebra g = gl(m|n). One “super phenomenon” is the existence of non-conjugate Borel subalgebras of g. Given two Borel subalgebras b and b′ that are related by an odd reflection with respect to an isotropic odd simple root, it is known that the categories O relative to b and b′ coincide, but the identity functor does not preserve the structure of highest weight categories in general [CLW, Section 6]. Thisstatement iseasily seentobevalid fortheversion ofcategory Owitharbitrary (i.e., non-integral) weights. We show here that the block equivalence with respect to an odd reflection associated to an isotropic odd simple root α is indeed an equivalence of highest weight categories under the assumption that hλ,α∨i 6∈ Z. In the case of gl(m|n), we show that it is possible to use a sequence of twisting functors and such odd reflections to reduce to the setting of an integral subalgebraof Levi type. Thesame typeof parabolic induction functor as in the case of semisimple Lie algebras completes the reduction process. Finally, we show that a block as defined via its central character and integral Weyl group is indeed indecomposable just as for semisimple Lie algebras. The above strategy works for a large set of, but not all, non-integral weights for other types of simple and Kac-Moody Lie (super)algebras; compare with Fiebig [Fie]. This note is organized as follows. In Section 2, we develop our approach in the Lie algebra setting. The block equivalences via parabolic induction and twisting functors are formulated and established. In Section 3, we generalize it to the Lie superalgebra setting. For basics on category O we refer to the book of Humphreys [Hu], and for basics on Lie superalgebras we refer to the new book [CW]. Acknowledgement. ThefirstauthorispartiallysupportedbyanNSCgrantoftheR.O.C. The second author is partially supported by the Royal Swedish Academy of Sciences and the Swedish Research Council. The third author is partially supported by NSF DMS-1101268, and he thanks T. Tanisaki for some helpful correspondence. We thank Kevin Coulembier for pointing out an oversight in the original formulation of Theorem 2.1. 2. Equivalence of blocks for Lie algebras 2.1. The setup. Let g be a complex semisimple or reductive Lie algebra with a Cartan subalgebra h, root system Φ, and Weyl group W. We denote by U(g) the universal enveloping algebra of g and by Z(g) the center of U(g). Let Π be a simple system of Φ and denote by Φ+ the corresponding set of positive roots. For α ∈ Φ, we let α∨ stand for the corresponding coroot. Denote by ρ the half of the sum of all positive roots, and consider the usual dot- (i.e., ρ-shifted) action of the Weyl group given by w·λ = w(λ+ρ)−ρ, for w ∈W and λ ∈h∗. We have the corresponding triangular decomposition g = n− ⊕h⊕n, where n = g and α∈Φ+ α n− = g . Let gO be the Bernstein-Gelfand-Gelfand (BGG) category of g-modules (see [BGGα∈]−oΦr+[Hαu, Chapter 1]). We shall drop the superscript g and write simLply O, when g L is clear from the context. 3 For each λ ∈ h∗ we have the integral root system Φ = {α ∈ Φ | hλ,α∨i ∈ Z} associated to λ λ and the corresponding integral Weyl group W , which is the subgroup of W generated by λ all reflections s , α ∈Φ . The weight λ is called integral if W = W . When g is of ADE type, α λ λ we define (2.1) g = h⊕ g , [λ] α αM∈Φλ which is a Lie subalgebra of g, called the integral subalgebra associated to λ. If g is not of ADE type, then g may not be closed under the Lie bracket. [λ] For λ ∈ h∗ we denote by M(λ) the Verma module with highest weight λ, by L(λ) the unique simple quotient of M(λ) and by χ : Z(g) → C the central character of M(λ) (see λ [Hu, Chapter 1]). Denote by O the Serre subcategory of O generated by the simple objects χλ L(w · λ), w ∈ W (these are exactly the simple objects of O with central character χ , see λ [Hu, Section 1.10]). Then the action of Z(g) gives rise to a decomposition of O as follows: O = O . If λ is integral, then O is indecomposable (see [Hu, Section 1.13]). λ∈h∗/(W,·) χλ χλ However, O is decomposable in general. DenLote bχyλ O the Serre subcategory of O generated by L(w · λ), w ∈ W . Define an λ λ equivalence relation ∼ on h∗ by declaring λ ∼ µ if µ ∈ W · λ. Let h∗ denote the set λ dom of all dominant weights with respect to the dot-action, that is the set of all λ such that hλ + ρ,α∨i 6∈ Z<0 for all α ∈ Φ+ (see [Hu, Section 3.5]). Each such dominant λ is the maximum element in W · λ with respect to the natural order ≤ on h∗ (given by µ ≤ ν if λ and only if ν −µ ∈ Z≥0Φ+). Then we have a refined decomposition of O into indecomposable blocks (see [Hu, Theorem 4.9]): O = ⊕λ∈h∗ Oλ. dom 2.2. Twisting functors. For α ∈ Π fix a nonzero X ∈ g and let U′ be the (Ore) localiza- −α α tion of U(g) with respect to powers of X. Then U(g) is a subalgebra of the associative algebra U′ and the quotient U := U′/U(g) has the induced structure of a U(g)-U(g)–bimodule. Let α α α ϕ = ϕ be an automorphism of g that maps g to g for all β ∈ Π, where s is the simple α β sα(β) α reflection corresponding to α. Finally, consider the bimodule ϕU , which is obtained from U α α by twisting the left action of U(g) by ϕ (i.e. a·u·b := ϕ(a)ub for all a,b ∈ U(g) and u ∈U ). α Tensoring with ϕU defines an endofunctor T of O, called the twisting functor. This functor α α was originally definedby Arkhipovin [Ar]andfurtherinvestigated inmoredetail in [AS,KM]. Let Db(O) denote the bounded derived category of O and let LT : Db(O) → Db(O) be the α corresponding left derived endofunctor. Furthermore, for i ≥ 0, let L T : O → O denote the i α i-th cohomology of LT . Let us recall some basic properties of T (see, e.g., [AS]): α α (I) T is right exact; α (II) T preserves O , for each λ ∈h∗; α χλ (III) L T = 0 for i > 1; i α (IV) L T is isomorphicto thefunctor of taking themaximal submoduleon which the action 1 α of g is locally nilpotent. −α 2.3. Equivalencesusingtwistingfunctors. Letλ ∈ h∗andα ∈Πbesuchthathλ,α∨i 6∈ Z. Denote by O(α) the Serre subcategory of O generated by all L(w ·λ), w ∈ W, such that χλ χλ hw ·λ,α∨i 6∈ Z (the set of all w having the latter property will be denoted W(λ,α)). Then O(α) is a direct summand of O . χλ χλ 4 CHENG,MAZORCHUK,ANDWANG Theorem 2.1. Let λ ∈ h∗ and α ∈ Π be such that hλ,α∨i 6∈ Z. (a) The functor T :O(α) −→ O(α) is an autoequivalence (of highest weight categories) sending α χλ χλ M(w·λ) to M(s w·λ) and L(w·λ) to L(s w·λ) for all w ∈ W(λ,α). α α (b) For any µ ∈ W(λ,α)·λ, the functor T induces an equivalence between O and O . α µ sα·µ Proof. Since hλ,α∨i 6∈ Z, the classical theory of highest weight modules over sl(2) (see, e.g., [Maz, Chapter 3]) implies that the action of g is injectively on each module from O(α). −α χλ Therefore Properties (III) and (IV) imply that T : O(α) → O(α) is exact. The functor T α χλ χλ α maps Verma modules to Verma modules (this follows from, e.g., [KM, Lemma 3]). Let us now check that T sends L(w·λ) to L(s w·λ). Set M := T L(w·λ). We claim that α α α s w·λ is a highest weight of M. Indeed, if we localize L(w·λ) and then factor out the copy of α L(w·λ) in the localization, the resulting module will have a one-dimensional weight space of weight w·λ+α. The twisting proceduremaps w·λ+αto s w·λ and thus s w·λ ∈supp(M). α α By construction, (s w·λ)+α6∈ supp(M). Sinceϕ leaves the set of all positive roots different α α from α invariant, it follows that (s w·λ)+β 6∈ supp(M) for any β ∈ Φ+\{α}. Hence s w·λ α α is a highest weight of M. In particular, M 6= 0 (i.e., T does not annihilate any modules). α Denote by a the sl(2)-subalgebra of g generated by g . Starting with N ∈ {L(w·λ),M} ±α we can do two things: first apply T to N and then consider the result as an a-module, or first α consider N and as an a-module and then apply the a-version of T to it. By [KM, Lemma 3] α (note that theintegrality assumption inthat lemmajustreflects thegeneral setupof [KM]and is not used in the proof), both these ways produce isomorphic a-modules. In the sl(2)-case it is straightforward to verify (cf. [Maz, Chapter 5]) that the characters of T M and L(w ·λ) α coincide. It follows that the characters of T M and L(w·λ) coincide in the genral case and α hence T M ∼= L(w·λ), as any simple module in O is uniquely determined by its character. α Now, by the above we know that L(s w·λ) is a simple subquotient of M. We claim that M α is simple. Indeed, assume that this were not the case. Then T M cannot be simple, since T α α is exact and does not annihilate any modules. This contradicts the conclusion in the previous paragraph. Thus, we know that T is an exact endofunctor of O(α) which sends simple modules to α χλ simple modules. Let K denote the right adjoint of T . By [AS, Theorem 4.1], we have α α K ∼= dT d, where d : O(α) → O(α) is the usual duality (i.e., a contravariant autoequivalence α α χλ χλ which preserves isoclasses of simple modules, see [Hu, 3.2]). Since d is exact, it follows that K is exact as well. Furthermore, using the above we compute: α K L(w·λ) ∼= dT dL(w·λ) ∼= dT L(w·λ) ∼= dL(s w·λ) ∼= L(s w·λ), α α α α α which means that K also sends simple modules to simple modules. Now the fact that T α α and K are mutually inverse equivalences of categories follows by standard arguments using α induction on the length of a module and the Short Five Lemma. This proves (a), and (b) follows from (a) by restriction. (cid:3) Example 2.2. If g = gl(2), then every non-integral block of O is semisimple. For each non- integral central characters there existexactly twonon-isomorphic simple highestweightmodules (see e.g. [Maz, Section 3.2]) and T swaps them. α 2.4. Parabolic induction from a Levi subalgebra. For a weight module M and a weight µ we denote by M the µ-weight space in M. For Π′ ⊆ Π consider the corresponding Levi µ 5 subalgebra l= l(Π′) (which, by definition, contains h). Let u be the nilpotent radical of l+b, where b is the Borel subalgebra corresponding to Φ+. For M ∈ O the maximal subspace Mu of M on which uacts trivially inherits the natural structureof an l-moduleby restriction. The following proposition is standard (see e.g. [J, Lemma 2]). Proposition 2.3. Let g be a semisimple Lie algebra, λ ∈ h∗, and suppose that g = l is a [λ] Levi subalgebra of g. Then the parabolic induction functor Indg : lO → gO is an equivalence l+u λ λ (of highest weight categories) with inverse Resg :gO → lO given by M 7→ Mu. l,λ λ λ Denote by g′ = [g ,g ] thederived subalgebraof g . Forgetting the action of thecenter [λ] [λ] [λ] [λ] of g[λ] we see that lOλ is equivalent to a block of g′[λ]O which is integral by construction. 2.5. A sequence of reductions. Remark 2.4. Let g = gl(m) with simple system {ǫ −ǫ ,...,ǫ −ǫ }. Let λ = m λ ǫ 1 2 m−1 m i=1 i i be a weight and let α = ǫ −ǫ be a simple root. We have i i+1 P s ·λ = λ ǫ +...+λ ǫ +(λ −1)ǫ +(λ +1)ǫ +λ ǫ +...+λ ǫ . α 1 1 i−1 i−1 i+1 i i i+1 i+2 i+2 m m Hence s has the effect of interchanging the i-th and the i+1-st coefficients of λ modulo Z. α Proposition 2.5. Let g = gl(m) and λ = λ ∈ h∗. There exists α ,...,α ∈ Π such that for 0 1 n λ := (s s ···s )·λ, 1≤ k ≤ n, we have: k αk αk−1 α1 (i) hλ ,α∨i 6∈ Z for k = 1,...,n; k−1 k (ii) g is a Levi subalgebra of g. [λn] Proof. Write λ = (λ ,λ ,...,λ ) if λ = m λ ǫ . Define an equivalence relation ≈ on the 1 2 m i=1 i i multiset {λ ,λ ,...,λ } via λ ≈ λ if and only if λ −λ ∈ Z. Modulo integral shifts, the 1 2 m i j i j P claim of the proposition is equivalent to the assertion that, using a sequence of elementary transpositions, where at each step we can swap two neighboring elements belonging to dif- ferent ≈-equivalence classes, the sequence (λ ,λ ,...,λ ) can be transformed to a sequence 1 2 m (µ ,µ ,...,µ ) in which all equivalence classes are connected in the sense that µ ≈ µ for 1 2 m i j some i< j implies µ ≈ µ for all i ≤ s ≤ j. The latter assertion is evident. (cid:3) i s Corollary 2.6. Let g =gl(m) and λ ∈ h∗. Then O is equivalent to an integral block of O for λ some semisimple Lie algebra (of a perhaps smaller rank). Proof. Set λ = λ , and let α ,...,α ∈ Π be a sequence given by Proposition 2.5. By 0 1 n Theorem2.1,applyingtoO firstthefunctorT ,thenT andsoon,givesanequivalencefrom λ α1 α2 O to O . As g is a Levi subalgebra, the proof is completed applying Proposition 2.3. (cid:3) λ λn [λn] Corollary 2.6 makes Soergel’s equivalence much more explicit in the special case of the Lie algebra gl(m). Soergel’s approach uses the coinvariant algebra and works for all semisimple Lie algebras, see [Soe]. Example 2.7. Let g = gl(3) with simple roots Π = {ǫ −ǫ ,ǫ −ǫ }. Let λ = (1, 1,1) and 1 2 2 3 2 we have Φ = {ǫ −ǫ } and g ∼= gl(2)+h (the latter is not a Levi subalgebra of g). The λ 1 3 [λ] functor T induces an equivalence from O to O . Furthermore, Φ = {ǫ −ǫ } and g ∼=αg2l(2)⊕gl(1) is now a Levi subalgeλbra. sα2·λ sα2·λ 1 2 [sα ·λ] 2 6 CHENG,MAZORCHUK,ANDWANG 2.6. Beyond type A. Theorem 2.1 and Proposition 2.3 remain valid in the case when g is a Kac-Moody algebra. So for a weight λ which satisfies the Conditions (i) and (ii) of Proposition 2.5, we still have an equivalence from O to some integral block (for a Kac- λ Moody Lie algebra of, possibly, lower rank). However, for general Lie or Kac-Moody algebras, Proposition 2.5 holds only for a proper subset of weights in h∗. Here are some examples. Example 2.8. Let {ǫ |1 ≤ i ≤ n} be the standard orthonormal basis of Rn, n ≥ 4. The root i system Φ of type D in Rn has roots {±ǫ ±ǫ |i 6= j}. The standard simple system for this n i j root system is given below with its Dynkin diagram: (cid:13) ǫn−1−ǫn (cid:0) (cid:0) (cid:0) (cid:13) (cid:13) ··· (cid:13) (cid:13) ǫn−2−ǫn−1 ❅ ǫ1−ǫ2 ǫ2−ǫ3 ǫn−3−ǫn−2 ❅ ❅ (cid:13) ǫn−1+ǫn Let λ = 2ǫ +ǫ + n 2n−2i+1ǫ . Then Φ has type A ⊕A ⊕D , where the simple systems 1 2 i=3 2 i λ 1 1 n−2 for A , A , and D are given by 1 1 n−2 P {ǫ −ǫ },{ǫ +ǫ } and {ǫ −ǫ ,ǫ −ǫ ,...,ǫ −ǫ ,ǫ +ǫ }, 1 2 1 2 3 4 4 4 n−1 n n−1 n respectively. Thus, Φ has rank n and is a proper subset of Φ. Therefore Φ cannot coincide λ λ with the set of roots for any Levi subalgebra of a Lie algebra of type D . This means that n Proposition 2.5 does not hold in type D , n ≥ 4, in general. n Example 2.9. Let {ǫ |1 ≤ i ≤ 8} be the standard orthonormal basis of R8. The root system i of type E has roots {±ǫ ±ǫ |i 6= j}∪{1 8 a ǫ |a = ±1 and a = 1}. The standard 8 i j 2 i=1 i i i i simple system for this root system is given below with its Dynkin diagram: P Q ǫ7+ǫ8 (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) ǫ2−ǫ3 ǫ3−ǫ4 ǫ4−ǫ5 ǫ5−ǫ6 ǫ6−ǫ7 ǫ7−ǫ8 21(ǫ1+ǫ8−P7i=2ǫi) Let λ = 33ǫ +2ǫ +ǫ + 9ǫ + 7ǫ + 5ǫ + 3ǫ + 1ǫ . One checks that Φ has type A ⊕E , 2 1 2 3 2 4 2 5 2 6 2 7 2 8 λ 1 7 where the simple systems for A and E are {ǫ −ǫ } and 1 7 2 3 7 1 {ǫ −ǫ ,ǫ −ǫ ,ǫ −ǫ ,ǫ −ǫ ,ǫ +ǫ , (ǫ +ǫ − ǫ ),ǫ +ǫ }, 4 5 5 6 6 7 7 8 7 8 1 8 i 2 3 2 i=2 X respectively. Similarly to the previous example, this implies that Proposition 2.5 does not hold in type E in general. 8 Remark 2.10. Similarly to Example 2.9 one shows that Proposition 2.5 does not hold for E 6 and E in general. In particular, the proof of [Mat, Proposition A.4] is valid only in type A. 7 7 3. Equivalence of blocks for Lie superalgebras In this section, we generalize the above results to Lie superalgebras of type A. 3.1. Preliminaries. Let Cm|n be the complex superspace of dimension (m|n). The general linear Lie superalgebra g = gl(m|n) is the Lie superalgebra of linear operators on Cm|n. Let {e ,...,e } and {e ,...,e } be the standard bases for the even subspace Cm|0 and the 1 m m+1 m+n odd subspace C0|n, respectively. Their union is then a homogeneous basis for Cm|n and we can use it to identify the space of (m +n)×(m +n) matrices with gl(m+n). We let e , ij 1 ≤ i,j ≤ m + n, denote the (i,j)-th elementary matrix unit. The Cartan subalgebra of diagonal matrices is denoted by h = h , and it is spanned by {e |1 ≤ i ≤ m + n}. We m|n ii denote by {δ |1 ≤ i ≤ m+n} the basis in h∗ = h∗ dual to {e |1 ≤ i ≤ m+n}. We let i m|n ii Φ = {δ −δ |1 ≤ i 6= j ≤ m+n} be the root system and W = S ×S be the Weyl group i j m n (it acts by permuting the respective coordinates). Set [1,k] = {1,2,...,k}. Define the parity function [1,m+n]→ Z = {0,1}, i 7→¯i, where 2 0, if 1≤ i ≤ m; ¯i= (1, if m < i≤ m+n. The supertrace form on gl(m|n) induces a non-degenerate symmetric bilinear form (·|·) on h∗ given by (−1)¯i, if 1 ≤i = j ≤ m+n; (δ |δ ) = i j (0, if 1 ≤i 6= j ≤ m+n. The subalgebra of upper triangular matrices with respect to this standard basis is called the standard Borel subalgebra and is denoted by bst. There exist Borel subalgebras of gl(m|n) that are not W-conjugate to bst. However, applying W-conjugation if necessary, we may assume that a general Borel subalgebra b satisfies b = bst, and in such a case b is related to bst by a ¯0 ¯0 sequence of odd reflections (see, e.g., [CW, Proposition 1.32]). The set of positive roots of b is denoted by Φ+. b The BGG category Ob for g with respect to b is defined similarly as for Lie algebras. We let O = Obst denote the BGG category with respect to bst. As abstract categories, the categories Ob and O coincide (and hence are related by the identity functor). The identity functor obviously sends simple objects to simple objects and projective objects to projective objects. Thecategory Ob isahighestweightcategory. Standardobjectsforthehighestweightstructure areVermamodulesandtheidentityfunctordoesnotsendstandardobjectstostandardobjects in general. Another structural family of modules for a highest weight category is formed by tilting modules. For λ ∈ h∗ denote by Mb(λ), Tb(λ), and Lb(λ) the Verma, tilting, and simple modules in Ob with highest weight λ, respectively. When b = bst, we will usually drop the superscript. All tilting modules are direct summands of modules induced from tilting gl(m|n) -modules, and this implies that a b-tilting module remains a b′-tilting for any other ¯0 b′ (which was first proved in [CLW, Proposition 6.9]). Therefore, the identity functor sends tilting modules to tilting modules. One can associate to a Borel subalgebra b with b = bst a sequence b = (b ,b ,...,b ) ¯0 ¯0 1 2 m+n consisting of m 0s and n 1s, called an 0m1n-sequence as follows [CLW, Section 6.1]. Indeed it is well-known that such a Borel is completely determined by a total ordering of the basis 8 CHENG,MAZORCHUK,ANDWANG {e |1 ≤ i ≤ m+n} such that the total ordering induced on basis elements of the same parity i is the standard one. We then attach to b the sequence b by letting b to be the parity of the j jth basis element in this total ordering. For example, the sequence bst =(0m,1n) corresponds to bst. Our next step is to define the b-standard partial order on h∗. This is a straightforward generalization of the integral case dealt with in [Br] and [CLW, Section 2.3]. Let P denote the free abelian group with basis {ε |r ∈ C}. We define a partial order on P by declaring ν ≥ µ, r for ν,µ ∈ P, if ν −µ is a non-negative integral linear combination of ε −ε , r ∈ C. r r+1 Definition 3.1. Fix a 0m1n-sequence b = (b ,...,b ). For f : [1,m + n] −→ C and 1 m+n 1 ≤ j ≤ m+n, we define wtjb(f):= (−1)biεf(i) ∈ P, wtb(f):= wt1b(f)∈ P. j≤i X Define the standard partial order (cid:22)b of type b on the set of C-valued functions on [1,m+n] j j as follows: g (cid:22)b f if and only if wtb(g) = wtb(f) and wtb(g) ≤ wtb(f) for all j. Now the b-standard partial order (cid:22) on h∗ is definedas follows: Let b bethe 0m1n-sequence b associated with b, and let ρ be the Weyl vector normalized as in [CLW, (6.5)]. We have b a bijection between h∗ and the C-valued functions on [1,m + n] given by λ 7→ f , where λ fλ(i) = (λ+ρb|δi). For λ,µ ∈ h∗ we set λ (cid:22)b µ if fλ (cid:22)b fµ. Remark 3.2. Let g = gl(m|n) and λ,µ ∈ h∗ be such that λ (cid:22) µ. Suppose that α is an even b simple root for b such that hλ+ρ ,α∨i 6∈ Z. Then s ·λ (cid:22) s ·µ. b α b α 3.2. A block decomposition. As in the case of Lie algebras, the category Ob has a block decomposition,whichwewilldescribebelow. FirstwenotethatthedefinitionsofΦ andg in λ [λ] Section 2.1alsomakesensewheng = gl(m|n). ThegroupW isthenthesubgroupoftheWeyl λ group W generated by {s |α ∈ Φ ,α even}. For λ ∈ h∗, we let Ob be the Serre subcategory of α λ λ Ob generated by simple objects of the form Lb(µ) with µ = w·(λ− k α ), where w ∈W , j j j λ {α } is a set of mutually orthogonal odd isotropic roots satisfying hλ+ρ ,α∨i = 0 (and hence j P b j α ∈ Φ ), and k ∈ Z. Denote the set of all such µ above by [λ] so that we have an equivalence j λ i relation ∼ on h∗ with equivalence classes [λ]. Note that µ (cid:22) λ implies that µ ∈ [λ]. b Proposition 3.3. We have the following decomposition: (3.1) Ob = Ob. λ λ∈h∗/∼ M Proof. Clearly HomOb(Lb(λ),Lb(µ)) 6= 0 implies that µ = λ. Hence we only need to check that Ext1 (Lb(λ),Lb(µ)) 6= 0 implies that µ ∈ [λ]. Ob First, for λ = m+nλ δ ∈ h∗ we define an equivalence relation ≈ on the set [1,m + n] i=1 i i similarly as before: for i,j ∈ [1,m+n], P (3.2) i ≈ j, if (−1)¯iλ −(−1)¯jλ ∈ Z. i j Denote the equivalence classes by Iλ,Iλ,...,Iλ, where the numbering is determined by the 1 2 ℓ condition that min{i|i ∈ Iλ}< min{i|i ∈ Iλ }, for 1 ≤ j ≤ ℓ−1. j j+1 9 Supposethat Ext1 (Lb(λ),Lb(µ)) 6=0. Clearly, λ and µ musthave the same central charac- Ob ter. We have the following description of the central characters for g, e.g. [CW, Section 2.2.6], which is a consequence of the description of the center of U(g) (see [K, Sv]): Two weights λ and µ have the same central character if and only if there exist w ∈ W and a set of mutually orthogonal odd roots {α } such that µ = w·(λ− k α ) with α satisfying hλ+ρ ,α∨i= 0 j j j j j j b j and kj ∈ C. We observe that wtb(fλ) = wtb(fµ), where b is the 0m1n-sequence corresponding P to b. Looking at the supports of our modules, we obtain that λ−µ ∈ ZΦ ⊆ m+nZδ , and k=1 k thusIλ = Iµ,for all1 ≤ j ≤ ℓ. From this itis readily seenthat onecan findw′ ∈ W , β ∈Φ , j j P λ j λ and l ∈Z such that hλ+ρ ,β∨i = 0 and j b j µ+ρ = w′(λ+ρ − l β ). b b j j j X The proposition is proved. (cid:3) Remark 3.4. Theorem 3.12 below shows that all summands in (3.1) are indecomposable. 3.3. On integral weights. A weight λ ∈ h∗ is called integral if hλ,α∨i ∈ Z for all α ∈ Φ. Set 1 = (−1)¯iδ . m|n i 1≤i≤m+n X Then for any integral weight λ ∈ h∗, we can write λ = λ′ + a1 , where a = a(λ) ∈ C m|n and λ′ = λ′δ ∈ h∗ satisfies that λ′ ∈ Z for all i. Denote by V the one-dimensional i i i i a gl(m|n)-moduleofweighta1 . TensoringwithV (andrespectively, withV intheopposite m|n a −a P direction) we have the following. Lemma 3.5. There is an equivalence of blocks between Ob and Ob . λ λ′ While λ is integral in the sense of this paper, λ′ (but not λ in general) is an integral weight in the sense of [Br, CLW]. This lemma assures us that we can work with either version of integral weights. 3.4. Parabolic induction functor. We have the following generalization of Proposition 2.3. Proposition 3.6. Let b be a Borel subalgebra of g = gl(m|n) and let λ ∈ h∗. Suppose that g = l is a Levi subalgebra of g and let u be the nilradical of l+b. Then the parabolic induction [λ] functor Indg :lO → Ob is an equivalence of blocks, with inverse equivalence Resg :Ob → lO l+u λ λ l λ λ defined by M 7→ Mu, where Mu is the maximal trivial u-submodule of M. Furthermore, this equivalence maps Verma modules to Verma modules and hence is an equivalence of highest weight categories. Proof. Clearly, the parabolic induction functor sends Verma modules to Verma modules. We shall now prove that it sends irreducible modules to irreducible modules. Wehavehλ,β∨i 6∈ Zforallrootsβ ofuandsoβ 6∈ Φ . LetL0(λ)betheirreduciblel-module λ of highest weight λ and consider the correspondingparabolically induced module Indg L0(λ). l+u If µ is the weight for a non-zero singular vector in a subquotient of the highest weight module Indg L0(λ),thenwehaveµ ∈ [λ]byProposition3.3. SinceΦ isinvariantunderW ,itfollows l+u λ λ that we can write µ = λ− k α, k ∈ Z . Hence any subquotient of Indg L0(λ) α∈Φ+b∩Φλ α α + l+u intersects L0(λ), the latter being a simple l-module. This yields that Indg L0(λ) is simple. P l+u 10 CHENG,MAZORCHUK,ANDWANG The module E0 = ResgL(λ) = L(λ)u is simple by a highest weight argument. Indeed, l assume that E0 contains some simple submodule L0(µ) different from L0(λ). Then from the previous paragraph we have Indg L0(λ) = L(λ) and Indg L0(µ) = L(µ) and, by adjunction, l+u l+u Hom (L0(µ),ResgIndg L0(λ)) = Hom (Indg L0(µ),Indg L0(λ)) = Hom (L(µ),L(λ)) 6= 0, l+u l l+u g l+u l+u g which implies λ = µ. We can now conclude that ResgIndg L0(λ) ∼= L0(λ) and Indg ResgL(λ) ∼= L(λ). By l l+u l+u l g induction on the length of the modules and the Short Five Lemma we conclude that Res and l Indg are mutually inverse equivalences of categories. (cid:3) l+u 3.5. Equivalence via odd reflections. Suppose that α is an isotropic odd simple root of b. Let b′ be the Borel subalgebra obtained from b by applying the odd reflection with respect to α. The BGG categories Ob and Ob′ are identical, although they are equipped with different highest weight structures, see e.g. [CLW, Section 6]. The following lemma follows from the definition of the equivalence class [λ] in Section 3.2. Lemma 3.7. If α is an isotropic simple root and hλ,α∨i 6∈ Z, then for any µ ∈ [λ] we have hµ,α∨i 6∈ Z. The following proposition describes the relation between standard (or simple) modules for the different highest weight structures on O corresponding to b and b′. Proposition 3.8. [PS, Lemma1] Assume that α is an isotropic simple root of b and hλ,α∨i6∈ Z. Then for every µ ∈ [λ] we have Mb(µ) = Mb′(µ−α) and Lb(µ) = Lb′(µ−α). Proof. Denote by v+ the b-highest weight vector in Mb(µ) or Lb(µ), and by f a root vector µ α for −α. Note that e f v+ is a nonzero multiple of v+ thanks to hµ,α∨i6∈ Z (see Lemma 3.7). α α µ µ Then it is straightforward to verify that f v+ is a b′-highest weight vector in both Mb(µ) and α µ Lb(µ). Hence the irreducibility implies that Lb(µ)= Lb′(µ−α). Moreover, we have a natural surjective gl(m|n)-homomorphism Mb′(µ−α) → Mb(µ), which must be an isomorphism by a character comparison. (cid:3) 3.6. Twisting functors for Lie superalgebras. For an even simple root α of g we can define the corresponding twisting endofunctor T of O in exactly the same way as for Lie α algebras, see Section 2.2. There is an obvious analogue of [KM, Lemma 3] for T , saying that α this functor is essentially defined on the level of the sl(2)-subalgebra of g. It follows that the Properties (I)–(IV) in Section 2.2 and [AS, Theorem 4.1] transfer mutatis mutandis to the superalgebra case. In particular, similarly to Theorem 2.1, we have the following. Proposition 3.9. Let λ ∈ h∗ and α be an even simple root such that hλ,α∨i 6∈ Z. Then the functor T : O → O is an equivalence (as highest weight categories) sending M(w·λ) to α λ sα·λ M(s w·λ) and L(w·λ) to L(s w·λ), for each w ∈ W. α α 3.7. Reduction to integral blocks. Now consider a block O with respect to the standard λ Borel bst for an arbitrary weight λ ∈ h∗. We can first apply a sequence of suitable twisting functors (see Propositions 2.5 and 3.9) to Oλ and obtain an equivalent block Oλe such that the congruence classes modulo Z of λ for 1 ≤ i ≤ m, and respectively for m+1 ≤ i ≤ m+n, i are connected intervals, where λ = m+nλ δ . Next we can apply a sequence of suitable odd i=1 i i e P e e

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.