Mixed cohomology of Lie superalgebras

Mixed cohomology of Lie superalgebras

Journal of Algebra 549 (2020) 1–29 Contents lists available at ScienceDirect Journal of Algebra www.elsevier.com/locate/jalgebra Mixed cohomology o...

529KB Sizes 0 Downloads 104 Views

Journal of Algebra 549 (2020) 1–29

Contents lists available at ScienceDirect

Journal of Algebra www.elsevier.com/locate/jalgebra

Mixed cohomology of Lie superalgebras Yucai Su a , R.B. Zhang b a

Department of Mathematic, Tongji University, Shanghai, China School of Mathematics and Statistics, The University of Sydney, Sydney, NSW 2006, Australia b

a r t i c l e

i n f o

Article history: Received 4 March 2019 Available online 2 January 2020 Communicated by Volodymyr Mazorchuk MSC: primary 17B56, 18G35 secondary 81R05 Keywords: Lie superalgebra cohomology Weyl superalgebras Integral forms

a b s t r a c t Supermanifolds are known to admit both differential forms and integral forms, thus any appropriate super analogue of the de Rham theory should take both types of forms into account. However, the cohomology of Lie superalgebras studied so far in the literature involves only differential forms when interpreted as a de Rham theory for Lie supergroups. Thus a new cohomology theory of Lie superalgebras is needed to fully incorporate differential-integral forms, and we investigate such a theory here. This new cohomology is defined by a BRST complex of Lie superalgebra modules, and includes the standard Lie superalgebra cohomology as a special case. General properties expected of a cohomology theory are established for the new cohomology, and examples of the new cohomology groups are computed. © 2020 Elsevier Inc. All rights reserved.

1. Introduction The de Rham cohomology of a Lie group can be equivalently reformulated in terms of the cohomology of its Lie algebra [7] (also see [21, §7] for Lie algebra cohomology). This fact had profound impact on the development of Lie theory. One may try to cast E-mail addresses: [email protected] (Y. Su), [email protected] (R.B. Zhang). https://doi.org/10.1016/j.jalgebra.2019.11.036 0021-8693/© 2020 Elsevier Inc. All rights reserved.

2

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

a de Rham theory of a Lie supergroup into a similar algebraic setting by using the cohomology of its Lie superalgebra. However, this runs into the difficulty that the Lie superalgebra cohomology [9,10,17] studied so far in the literature is not adequate for this purpose. The problem is rooted in supergeometry, thus is not present in the Lie group context. To see the cause of the problem, we note the following fact about supermanifolds (see, e.g., [8,20] for introductions), which is not widely known. Beside differential forms, there also exist integral forms and mixtures of the two types of forms [1,22] on supermanifolds. They are all necessary for defining integration, and in particular, for establishing a Stokes’ theorem [1]. A generalisation of de Rham theory to supermanifolds should take into account differential-integral forms to capture new features of supergeometry. However, the cohomology of Lie superalgebras [9,10,17] in the literature is a direct generalisation of the cohomology of Lie algebras [7], [21, §7]. It is usually defined by a generalised Chevalley-Eilenberg complex [17] (also see Section 2.2), which corresponds to a complex of differential forms (tensored with a coefficient module). The integral forms are entirely discarded. Our aim is to develop a Lie superalgebra cohomology which will take into full account of differential-integral forms. We will call this a mixed cohomology of Lie superalgebras. We now describe in more detail the background of the current work, and also outline the main ideas and techniques involved. Differential-integral forms on supermanifolds. Let us recall the elementary treatment of differential-integral forms on supermanifolds by Witten [22]. Given a supermanifold M with the tangent bundle T M , denote by ΠT M the tangent bundle with the parity of the fibre space reversed. Thus the odd directions of the fibre space of ΠT M come from the even coordinates of M (in particular, for an ordinary manifold, the fibre of ΠT M is purely odd), and the even directions from the Grassmannian variable coordinates. The forms on M are functions on ΠT M : differential forms are polynomial functions, and integral forms [1] are distributions supported at 0 in the even directions of the fibre space of ΠT M . Mixtures of the two types of (generalised) functions are differential-integral forms. Modules for Weyl superalgebras. A key observation in [22, §3.2] is that differentialintegral forms at a point of a supermanifold M constitute a module over a Weyl superalgebra. This Weyl superalgebra is generated by the coordinates of the fibre of ΠT M and their derivatives, thus is equal to the tensor product of a Clifford algebra of an even degree (i.e., a Clifford algebra over a vector space of which the dimension is even, see Section 3.1) and a bosonic Weyl algebra. While the even degree Clifford algebra has a unique irreducible module (the fermionic Fock space) up to canonical isomorphisms, the Weyl algebra has many non-isomorphic simple modules. Different modules correspond to different functions on ΠT M . To see this, we note that the bosonic Fock space for the Weyl algebra is cyclically generated by a vacuum vector, which is annihilated by all the even derivatives. Vectors in it correspond to differential forms. However, one may con-

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

3

sider a module for the Weyl algebra cyclically generated by a vector which is annihilated by some even variables. Such a vector is a distribution in these variables supported at 0 (see Example 3.1). Applying the corresponding derivatives to the vector produces more distributions. These correspond to integral forms [22, §3.2]. Differential-integral forms in the Lie superalgebra context. Conceptually we may regard a Lie superalgebra g as the Lie superalgebra of left invariant tangent vector fields on the underlying supermanifold of a Lie supergroup G. By [22], the differential-integral forms at a point of G are (generalised) functions on the parity reversed superspace Πg of g. They can be described in terms of modules for the Weyl superalgebra WΠg∗ over Πg∗ = (Πg)∗ , the dual superspace of Πg. We now apply such differential-integral forms to develop a theory of mixed cohomology of Lie superalgebras. The BRST formalism. Useful techniques for doing this are available in the physics literature, which originated from a method for quantising gauge theories known as the BRST formalism (see, e.g., [4]), first introduced by Becchi, Rouet and Stora, and by Tyutin. The BRST method has since developed into a vast theory, which has been applied to many other areas with remarkable success, for example, to string theory, symplectic geometry and semi-infinite cohomology of affine Kac-Moody algebras. We shall adapt the BRST method to implement the ideas of [22] discussed above to Lie superalgebras. As we will see in Section 3, the correspondence between differential-integral forms and modules for Weyl superalgebras [22, §3.2] becomes even more natural within the BRST framework. Mixed cohomology of Lie superalgebras. We define a mixed cohomology of a Lie superalgebra g by a BRST complex of g-modules in Theorem 3.1, which is formulated in terms of modules for the Weyl superalgebra WΠg∗ over Πg∗ . The Weyl superalgebra is the tensor product of a Clifford subalgebra and a bosonic Weyl subalgebra, where the latter is not equal to C if the Lie superalgebra g has a non-trivial odd subspace. In this case, WΠg∗ has non-isomorphic simple modules FΠg∗ (D), which are characterised by g-submodules D of Πg∗ called mixing sets. These non-isomorphic simple modules for WΠg∗ account for the differential-integral forms in the sense of [22, §3.2] (see Section 3.2). The BRST complex defining the mixed cohomology of g with coefficients in a g-module V has the superspace of cochains CD (g, V ) = FΠg∗ (D) ⊗ V . The standard Lie superalgebra cohomology [9,10,17] is recovered in Theorem 3.3 from the mixed cohomology in the special case D = 0, where FΠg∗ (0) is the standard Fock space. Examples. To illustrate how the general theory of mixed cohomology works, we calculate as examples various mixed u-cohomology groups of the general linear Lie superalgebra in Section 4. Some of these cohomology groups are highly non-trivial. It is the BRST method which enables us to carry out such computations explicitly. Some general results on the mixed cohomology of basic classical Lie superalgebras are also obtained.

4

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

Now some comments are in order. While the BRST method is the most direct way to reach Definition 3.1 of the mixed cohomology of Lie superalgebras, it is still very useful to reformulate the definition using more conventional homological algebraic methods [21], e.g., in terms of derived functors in a way analogous to [24]. We also mention that the mixed cohomology of Lie superalgebras is a very natural object to study; one could have discovered it much earlier by simply placing the standard Lie superalgebra cohomology [9,10,17] in the BRST framework. We work over the field C of complex numbers throughout this paper. 2. Standard cohomology of Lie superalgebras We recall the definition of the cohomology of Lie superalgebras [10,9,17] here. As we shall see in Section 3, it can be recovered from the BRST formalism for mixed cohomology. 2.1. Vector superspaces A vector superspace V is a Z2 -graded vector space V = V0 ⊕ V1 , where V0 and V1 are called the even and odd subspaces respectively. Here Z2 := Z/2Z is considered as an additive group. The degree of a homogeneous element v ∈ V0 ∪ V1 will be called the parity of v, and denoted by [v]. Let V and W be any two vector superspaces. The set HomC (V, W ) of homomorphisms is a vector superspace with HomC (V, W )i = ⊕j+k=i Hom(Vj , Wk ) for all i ∈ Z2 . The tensor product V ⊗C W is also a vector superspace with (V ⊗C W )i = ⊕j+k=i Vj ⊗C Wk . The category of vector superspaces is a tensor category equipped with a canonical symmetry τV,W : V ⊗ W −→ W ⊗ V,

v ⊗ w → (−1)[v][w] w ⊗ v.

Various types of algebras in this category will be called superalgebras of the corresponding types, e.g., associative superalgebras, Hopf superalgebras, and Lie superalgebras (see [12,16] for the theory of Lie superalgebras). We will consider only Z2 -graded modules for any superalgebra. The dual space of V is V ∗ := HomC (V, C), which is also Z2 -graded. Note that V ∗ ⊗W ∗ is embedded in (V ⊗ W )∗ such that for all f ∈ V ∗ and g ∈ W ∗ , (f ⊗ g)(v, w) = (−1)[g][v] f (v)g(W ),

∀ v ∈ V, w ∈ W.

If one of the vector superspaces is finite dimensional, this is an isomorphism. There exists the parity change functor Π on the category of vector superspaces, which acts as identity on morphisms, and reverses the parity on objects, that is, (ΠV )0 = V1 and (ΠV )1 = V0 for any vector superspace V . Now we have a canonical odd isomorphism

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

5

π : V −→ ΠV of vector superspace, which sends any homogeneous element of V to the same element in ΠV but with the opposite parity. Denote by Symr the symmetric group of degree r, and by CSymr its group algebra. The symmetry τV,V : V ⊗ V −→ V ⊗ V extends to a representation νr : CSymr −→ EndC (V ⊗r ) of CSymr on the r-th tensor power of V such that for all the simple reflections si = (i, i + 1) for 1 ≤ i ≤ i − 1, νr : si → τi = idV ⊗ · · · ⊗ idV ⊗τV,V ⊗ idV ⊗ · · · ⊗ idV ,       i−1

∀ i.

(2.1)

r−i−1

Write (σ) for the length of an element σ ∈ Symr , and let Sr =

1 r!



σ,

Σr =

σ∈Symr

1 r!



(−1)(σ) σ.

σ∈Symr

Then the symmetric and skew-symmetric r-th power of V are respectively given by S r (V ) = Sr (V ⊗r ),

r

(V ) = Σr (V ⊗r ).

(2.2)

We have the following Z-graded superspaces. S(V ) =

∞ 

S r (V ),



(V ) =

r=0

∞   r

(V ).

(2.3)

r=0

2.2. Cohomology of Lie superalgebras Let us briefly discuss the standard Lie superalgebra cohomology following [17, §II]. We refer to [12,16] for the theory of Lie superalgebras. Let g = g0 ⊕ g1 be a Lie superalgebra with the super Lie bracket [ , ] : g × g −→ g. Given any Z2 -graded left modules V and W for g (we only consider Z2 -graded modules), their tensor product V ⊗ W is again a g-module with the diagonal action. The canonical symmetry τV,W : V ⊗ W −→ W ⊗ V is a g-homomorphism. This in particular implies r that S r (W ) and (W ) are g-submodules of W ⊗r . The dual vector superspace W ∗ of a g-module W has a g-module structure defined by (Af )(v) = −(−1)[A][f ] f (Av) for all A ∈ g, f ∈ W ∗ and v ∈ W . Fix a g-module V , we consider the family of g-modules C p (g, V ) :=

p

(g∗ ) ⊗ V,

p ∈ Z+ .

For convenience, we also take C q (g, V ) = 0 for all q < 0. The Z2 -gradings of S p (Πg∗ ) and V naturally give rise to a Z2 -grading for C p (g, V ). Define the following contraction maps [17, (II.15)]

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

6

ip : C p (g, V ) ⊗ g −→ C p−1 (g, V ),

g ⊗ A → gA ,

gA (A1 , A2 , . . . , Ap−1 ) = g(A, A1 , A2 , . . . , Ap−1 ), where for g = a ⊗ v ∈ C p (g, V ) with a ∈

p

(g∗ ) and v ∈ V ,

[v]

g(A1 , A2 , . . . , Ap ) = (−1)

∀ Ai ∈ g,

p

[Ai ]

i=1

a(A1 , A2 , . . . , Ap )v.

These maps are g-homomorphisms. We further define maps [17, (II.18)] dp : C p (g, V ) −→ C p+1 (g, V ),

p = 0, 1, 2, . . . ,

(2.4)

inductively by (dp g)A = (−1)[A][g] A.g − dp−1 gA for all A ∈ g. There exists an explicit formula for these maps, which can be described as follows [17, (II.21)]. Define linear maps dp1 , dp2 : C p (g, V ) −→ C p+1 (g, V ) such that for any g ∈ C p (g, V ) and for all A0 , A1 , . . . , Ap ∈ g, (dp1 g)(A0 , A1 , . . . , Ap ) =

p 

(−1)r+Σr−1,r Ar g(A0 , . . . , Aˆr , . . . , Ap ),

(2.5)

r=0

(2.6) (dp2 g)(A0 , A1 , . . . , Ap )  = (−1)s+Σr,s −Σs−1,s g(A0 , . . . , Ar−1 , [Ar , As ], Ar+1 , . . . , Aˆs , . . . , Ap ), r


k with Σk, = [A ] [g] + i=0 [Ai ] for k < . Then one can readily show that dp = dp1 + dp2 ,

∀ p.

(2.7)

The maps dp are g-homomorphisms which are even with respect to the Z2 -grading. Furthermore, they satisfy dp+1 dp = 0 for all p. Definition 2.1. [10,17] The cohomology of a Lie superalgebra g with coefficients in a g-module V is the homology of the differential complex (C(g, V ), d) :

d0

d1

d2

C 0 (g, V ) −→ C 1 (g, V ) −→ C 2 (g, V ) −→ · · · .

Denote the cohomology groups of (C(g, V ), d) by H p (g, V ) =

ker dp im dp−1

for p ∈ Z+ .

To distinguish it from the mixed cohomology to be defined later, we call this the standard cohomology of Lie superalgebras. It has been much studied (see, e.g., [6,9, 11,14,15,18]), and has wide applications, for example, in the study of deformations of universal enveloping superalgebras such as quantum supergroups [3,23,24], extensions of Lie superalgebras and modules [17,18], and support varieties of Lie supergroups [2,13].

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

7

Another important application is in the solution of the character problem [5,11,14,15,19] of classical Lie superalgebras, which relied on results on the super analogue of Kostant’s u-cohomology [6,14,15] for specifically chosen parabolic sub superalgebras. This super u-cohomology is well known to be deeply rooted in the geometry of homogeneous superspaces [14,23]. Remark 2.1. The complex in Definition 2.1 reduces to the usual Chevalley-Eilenberg complex [21, §7.7] if g is an ordinary Lie algebra. 3. Mixed cohomology of Lie superalgebras We present the BRST formulation of the mixed cohomology of Lie superalgebras in this section. As we shall see in Section 3.6.1, the standard cohomology of Lie superalgebras discussed in Section 2 is a special case of the mixed cohomology. 3.1. Weyl superalgebras Given a vector superspace V , the Symr -representation on V ⊗r defined by (2.1) is semi-simple. Thus S r (V ) is a direct summand of V ⊗r as Symr -module, and we may also define S r (V ) as the quotient of V ⊗r by the complement. The skew symmetric power r (V ) can be similarly defined as a quotient. This gives an alternative description of S(V ) and (V ), which has the advantage of making their superalgebraic structures transparent. Denote by T (V ) the tensor algebra of V . Let JS be the two-side ideal generated by v ⊗ v  − τV,V (v ⊗ v  ) for all v, v  ∈ V , and let J∧ be the two-side ideal generated by v ⊗ v  + τV,V (v ⊗ v  ) for all v, v  ∈ V . Then S(V ) = T (V )/JS ,



(V ) = T (V )/J∧ .

Hence both S(V ) and (V ) are associative algebras. The tensor algebra T (V ) has a natural Z+ -grading with V being the degree 1 subspace. As JS and J∧ are homogeneous, the algebras S(V ) and (V ) are both Z+ -graded with the homogeneous subspaces given in (2.3). The Z2 -grading of V induces a Z2 -grading for T (V ). Since both JS and J∧ are Z2 -graded ideals, S(V ) and (V ) are naturally Z2 -graded. Let V ∗ be the dual vector superspace of V . We consider the tensor algebra T (V ⊕ V ∗ ) over V ⊕ V ∗ . Let J be the two-side ideal generated by the following elements v ⊗ w − τV,V (v ⊗ w),

v¯ ⊗ w ¯ − τV ∗ ,V ∗ (¯ v ⊗ w), ¯

v¯ ⊗ w − τV ∗ ,V (¯ v ⊗ w) − v¯(w),

for v, w ∈ V and v¯, w ¯ ∈ V ∗ . Note that this ideal is Z2 -graded. The Weyl superalgebra of V is the associative superalgebra

8

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

WV := T (V ⊕ V ∗ )/J.

∼ S(V ) ⊗ S(V ∗ ) as It is Z-graded with deg(V ) = 1 and deg(V ∗ ) = −1. Note that WV = vector superspaces. For V = V0 ⊕ V1 , we may regard V0 (resp. V1 ) as a purely even (resp. odd) vector superspace, and consider WV0 and WV1 . Then WV0 is the usual Weyl algebra over V0 , and WV1 is the Clifford algebra generated by V1 ⊕ V1∗ , thus is of degree dim(V1 ⊕ V1∗ ) = 2 dim V1 . We have WV = WV0 ⊗ WV1 as superalgebra. If d1 := dim V1 < ∞, the Clifford algebra WV1 has a unique simple module, the fermionic Fock space, which is 2d1 -dimensional and is given by FV1 (0) =

W

V1 = S(V1 ), WV1 ∂

∂∈V1 ∗

where S(V1 ) is defined by (2.2) with V1 regarded as a purely odd vector superspace. We W may also construct a simple module, say, by taking FV1 (V1 ) := WVV1V1 = S(V1∗ ), but this 1 does not lead to anything new, since FV1 (V1 ) ∼ = FV1 (0) with the isomorphism determined by a vector space isomorphism S 0 (V1∗ ) ∼ = S d1 (V1 ) which is unique up to scalar multiples. On the other hand, there are many non-isomorphic irreducible representations of the Weyl algebra WV0 even when V0 is finite dimensional. Given any subspace D0 ⊂ V0 , we let D0⊥ = {¯ v ∈ V0∗ | v¯(w) = 0, ∀ w ∈ D0 }. Then we have the corresponding simple WV0 -module FV0 (D0 ) =

∂∈D0⊥

WV0

. WV0 ∂ + WV0 v v∈D0

In the extreme case with D0 = 0, the module FV0 (0) is the standard Fock space of the Weyl algebra, where the vacuum vector is 1. Example 3.1. Consider the case V = Cx with x being even. Then WV as an associative d algebra is generated by x and ∂ = dx subject to the relation ∂x − x∂ = 1. We have FV (0) = C[x],

FV (V ) = C[∂].

We can interpret C[∂] as a space of distributions in x supported at 0. To do this, we will regard x as a real variable, and consider complex valued functions f (x) in x. Let δ(x) be Dirac’s delta-function, which satisfies

∞ f (x) −∞

k dk δ(x) k d f (0) dx = (−1) , dxk dxk

for any function f (x) such that its k-order derivative

k = 0, 1, . . . ,

dk f (0) dxk

exists at 0.

(3.1)

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29 k

δ(x) Let D = k≥0 C d dx k , and endow it with a Z-grading such that −k. Then D has the structure of a Z-graded WV -module with



dk δ(x) dk+1 δ(x) = , k dx dxk+1

x

9

dk δ(x) dxk

is at degree

dk δ(x) dk−1 δ(x) = −k , k dx dxk−1

where the first relation is by definition, and the second can be verified by using (3.1). Therefore, the linear map C[∂] −→ D defined by ∂ k →

dk δ(x) , dxk

k = 0, 1, . . . ,

gives rise to an isomorphism of Z-graded WV -modules. This example is a special case of what discussed in [22, §3.2.2, §3.2.3]. The generalisation of this example to mixed Fock spaces of arbitrary Weyl superalgebras is immediate, see [22, §3.2.2, §3.2.3] for a detailed discussion. Now we consider representations of the Weyl superalgebra WV . Let D ⊂ V be a sub superspace, and let D⊥ = {¯ v ∈ V ∗ | v¯(w) = 0, ∀ w ∈ D}. Lemma 3.1. Corresponding to each sub superspace D of V , there exists a simple left WV -module FV (D) defined by FV (D) =

∂∈D ⊥

WV

. WV ∂ + WV v v∈D

This is a Z-graded WV -module which is also Z2 -graded. Furthermore, two modules FV (D) and FV (D ) are not isomorphic if D0 = D0 , where D0 and D0 are the even subspaces of D and D respectively. We call FV (D) a mixed Fock space for WV , and call D the mixing set of FV (D). The usual Fock space for WV is FV (0), and the dual Fock space is FV (V ). 3.1.1. Comments on parity reversal Let ΠV be the parity reversed vector superspace of V . Then as Z-graded associative algebras, S(V ) and (V ) satisfy 

(V ) = S(ΠV ),

S(V ) =



(ΠV ).

(3.2)

Recall from Section 2.1 that S(V ) and (V ) inherit Z2 -gradings from T (V ), thus are superalgebras. There is now a slight complication in that (V ) has a different Z2 -grading from that of S(ΠV ), but this is not a disaster. The Z-graded algebra isomorphism between (V ) and S(ΠV ) can be interpreted as a superalgebra isomorphism, which however is not homogeneous. Let π : V −→ ΠV

10

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

be the canonical odd isomorphism, which sends any homogeneous element of V to the same element in ΠV but with the opposite parity. This induces an isomorphism of superalgebras T (V ) −→ T (ΠV ) which is of degree α on the subspace ⊕∞ T (V )2k+α for k=0 α = 0, 1. It in turn induces the superalgebra isomorphism between (V ) and S(ΠV ), which is what we are after. ˜ where J˜ is the two-sided We may define a Clifford superalgebra by ClV = T (V ⊕V ∗ )/J, ideal generated by the elements v ⊗ w + τV,V (v ⊗ w),

v¯ ⊗ w ¯ + τV ∗ ,V ∗ (¯ v ⊗ w), ¯

v¯ ⊗ w + τV ∗ ,V (¯ v ⊗ w) − v¯(w),

for v, w ∈ V and v¯, w ¯ ∈ V ∗ , and is obviously a Z2 -graded two-sided ideal. Therefore, ClV is a superalgebra. It is also a Z-graded superalgebra with deg(V ) = 1 and deg(V ∗ ) = −1. By noting that (ΠV )∗ = ΠV ∗ , we have the Z-graded superalgebras isomorphisms ClV = WΠV ,

WV = ClΠV ,

(3.3)

where the Z- and Z2 -gradings of WV and WΠV are as described in Section 3.1. However, note that these superalgebra isomorphisms are not homogeneous with respect to the Z2 -grading, similar to the superalgebra isomorphisms (3.2) described above. Remark 3.1. The dualities (3.2) and (3.3) require the introduction of morphisms of associative superalgebras which are more general than those usually allowed. Recall that usually one only allows even morphisms, that is, morphisms which are homogeneous of degree 0, in a category of superalgebras of any given type. 3.2. Differential-integral forms in the Lie superalgebra context Fix a Lie superalgebra g = g0 ⊕g1 , and denote by g∗ its dual space. Let Πg (resp. Πg∗ ) be the image of g (resp. g∗ ) under the parity reversal functor. Note that Πg∗ = (Πg)∗ . We regard g as the Lie superalgebra of left invariant vector fields on a Lie supergroup G. Then forms on G are polynomial functions and distributions on Πg. In particular, differential forms can be identified with elements of S(Πg∗ ). Following [22, §3.2], we can reformulate forms on G using modules over the Weyl superalgebra WΠg∗ over Πg∗ . The differential-integral forms now correspond to elements of mixed Fock spaces FΠg∗ (D) for WΠg∗ . In particular, the vector superspace of differential forms now corresponds to the standard Fock space FΠg∗ (0) = S(Πg∗ ). 3.3. Realisations of Lie superalgebras The dual vector superspace g∗ of the Lie superalgebra g has a g-module structure with the g-action defined, for any A ∈ g and v ∗ ∈ g∗ , by ∗

A.v ∗ (B) = (−1)[A][v ] v ∗ ([A, B]),

∀ B ∈ g.

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

11

There is an odd (i.e., degree 1) vector superspace isomorphism π : g∗ −→ Πg∗ . Now Πg∗ has the following g-module structure: A.π(v ∗ ) = (−1)[A] π(A.v ∗ ),

∀ A ∈ g, v ∗ ∈ g∗ .

(3.4)

Write W = WΠg∗ for the Weyl superalgebra over Πg∗ . To describe W more explicitly, we assume that g has super dimension sdimg = (N |M ), that is, dim g0 = N and dim g1 = M . We choose a basis {Ei | 1 ≤ i ≤ M } for g1 , and a basis {EM +j | 1 ≤ j ≤ N } for g0 . Here M + 1, M + 2, . . . are merely symbols, thus M and N may be infinite. Let Λ = {i | 1 ≤ i ≤ M } ∪ {M + j | 1 ≤ j ≤ N }. Now {Ea | a ∈ Λ} is a homogeneous basis of g. We denote by ad(A) the matrix of any A ∈ g in the adjoint representation of g relative to this basis, that is, [A, Eb ] =



ad(A)cb Ec ,

∀ b.

c∈Λ

Let {Xa | a ∈ Λ} be the homogeneous basis of Πg∗ dual to the basis {Ea | a ∈ Λ} of g. Note that Xi are even for 1 ≤ i ≤ M and XM +j are odd for 1 ≤ j ≤ N . Then the Weyl superalgebra W is generated by Xa , ∂a (:= ∂Xa ) with a ∈ Λ subject to the usual relations, namely, Xa Xb − (−1)[Xa ][Xb ] Xb Xa = 0,

∂a ∂b − (−1)[Xa ][Xb ] ∂b ∂a = 0,

∂a Xb − (−1)[Xa ][Xb ] Xb ∂a = δab

for all a, b ∈ Λ.

Remark 3.2. Note that [Xa ] − [Ea ] ≡ 1 (mod 2) for all a. We write the Z2 -graded commutator in W as [A, B] = AB − (−1)[A][B] BA,

A, B ∈ W .

(3.5)

Denote by Der0 (S(Πg∗ )) the sub superspace of W consisting of Z2 -graded derivations of S(Πg∗ ), that is,    Der0 (S(Πg∗ )) = Ξ ∈ W  Ξ(f g) = Ξ(f )g + (−1)[Ξ][f ] f Ξ(g), f, g ∈ S(Πg∗ ) . It is a Lie superalgebra with the super Lie bracket being the graded commutator in W . We have the following result. Lemma 3.2. There exists a homomorphism of Lie superalgebras defined by Γ : g −→ Der0 (S(Πg∗ )),

Γ(A) = −



(−1)[A][Ec ] Xc ad(A)bc ∂b .

b,c∈Λ

(3.6)

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

12



The image Γ(g) acts on ∂g∗ = a∈Λ C∂a (resp. g∗ = a∈Λ CXa ) by the adjoint representation (resp. the dual of the adjoint representation) of g. Explicitly, for all A ∈ g, [Γ(A), Xb ] = −

 (−1)[A][Ec ] Xc ad(A)bc , c∈Λ

[Γ(A), ∂b ] = (−1)[A]



ad(A)cb ∂c ,

(3.7)

b ∈ Λ.

(3.8)

c∈Λ

Proof. The following very easy computation proves (3.7). [Γ(A), Xb ] = −



(−1)[A][Ec ] Xc ad(A)ac [∂a , Xb ]

a,c∈Λ

 (−1)[A][Ec ] Xc ad(A)bc . =− c∈Λ

A similar computation proves (3.8). It follows from (3.7) that [Γ(B), [Γ(A), Xb ]] =



(−1)[A][Ec ]+[B][Ea ] Xa ad(B)ca ad(A)bc

a,c∈Λ



= (−1)[A][B]

(−1)([A]+[B])[Ea ] Xa (ad(A)ad(B))ba ,

a,c∈Λ

and similarly, 

[Γ(A), [Γ(B), Xb ]] = (−1)[A][B]

(−1)([A]+[B])[Ea ] Xa (ad(B)ad(A))ba .

a,c∈Λ

Now [[Γ(A), Γ(B)], Xb ] = [Γ(A), [Γ(B), Xb ]] − (−1)[A][B] [Γ(B), [Γ(A), Xb ]], hence [[Γ(A), Γ(B)], Xb ] = (−1)[A][B] 

(−1)([A]+[B])[Ea ] Xa

a,c∈Λ

(ad(B)ad(A))ba − (−1)[A][B] (ad(A)ad(B))ba



=−





(−1)([A]+[B])[Ea ] Xa [ad(A), ad(B)]ba

a,c∈Λ



=−

(−1)([A]+[B])[Ea ] Xa ad([A, B])ba .

a,c∈Λ

As [Γ(A), Γ(B)] is linear in Xa ’s and in ∂a ’s, this implies that [Γ(A), Γ(B)] = −



(−1)([A]+[B])[Ea ] Xa ad([A, B])ba ∂b = Γ([A, B]),

a,c∈Λ

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

13

proving that Γ is a Lie superalgebra homomorphism. This completes the proof. 2 We note in particular that for all a, b ∈ Λ, Γ(Ea ) = Γa := − [Γa , Xb ] =





(−1)[Ea ][Ec ] Xc ad(Ea )bc ∂b ,

b,c∈Λ

Xc ad(Ec )ba ,

[Γa , ∂b ] = (−1)[Ea ]

c∈Λ

(3.9)



ad(Ea )cb ∂c ,

(3.10)

c∈Λ

and Γa = b,c∈Λ Xc ad(Ec )ba ∂b . The following result is an immediate consequence of Lemma 3.2. Corollary 3.1. The Weyl superalgebra WΠg∗ admits the following g-action g × WΠg∗ −→ WΠg∗ ,

(A, f ) → A(f ) := [Γ(A), f ],

∀ A ∈ g, f ∈ WΠg∗ .

That is, the above defines a g-action on WΠg∗ , which preserves the superalgebra structure of the latter in the sense that A(f g) = A(f )g + (−1)[A][f ] f A(g),

A ∈ g, f, g ∈ WΠg∗ .

3.4. BRST cohomology of Lie superalgebras The Lie superalgebra cohomology given in Definition 2.1 can be reformulated in terms of representations of Weyl superalgebras in the spirit of BRST theory (see, e.g., [4]). Retain notation in the last section, in particular, W = WΠg∗ . For any sub superspace D ⊂ Πg∗ , we write F(D) for the mixed Fock superspace FΠg∗ (D) for simplicity. Recall that F(D) = ⊕p∈Z F(D)p is Z-graded. If the even subspace of D is a proper subspace of V0 , then F(D)p is infinite dimensional for all p ∈ Z. We consider W ⊗ U(g) as a Z-graded superalgebra with 1 ⊗ U(g) in degree 0 and thus (W ⊗ U(g))q = Wq ⊗ U(g) for all q ∈ Z. Let us fix an arbitrary mixing set D ⊂ Πg∗ . For any g-module V , the tensor product CD (g, V ) := F(D) ⊗C V is a W ⊗ U(g) module with a Z-grading given by CD (g, V ) =



p CD (g, V ),

p CD (g, V ) = F(D)p ⊗ V.

(3.11)

p∈Z

It also inherits a Z2 -grading from the Z2 -gradings of F(D) and V . Given a g-module homomorphism ϕ : V −→ V  , we have a corresponding W ⊗U(g) module homomorphism given by ϕ˜D := idFΠg∗ (D) ⊗ ϕ : CD (g, V ) −→ CD (g, V  ).

(3.12)

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

14

Thus we have obtained a functor from the category of g-modules to the category of W ⊗ U(g) modules for each mixing set D. It is clear from (3.12) that the functor is exact. We summarise the discussion into the following lemma. Lemma 3.3. Corresponding to each mixing set D ⊂ Πg∗ , there exists an exact functor from the category of g-modules to the category of W ⊗ U(g) modules defined by (3.11) and (3.12). Define the following elements of W ⊗ U(g) δ1 =



(−1)[Ea ] Xa ⊗ Ea ,

a∈Λ

δ2 =

1 (−1)[Ea ] Xa Γa ⊗ 1, 2

(3.13)

a∈Λ

δ = δ1 + δ2 .

(3.14)

Remark 3.3. One can easily show that the definitions of δ1 , δ2 , and hence of δ, are independent of the choice of basis for g. Note that the operators δ1 , δ2 and δ are homogeneous of degree 1 with respect to the Z-grading, and are odd with respect to the Z2 -grading, of W ⊗ U(g). Thus as linear p p+1 operators on CD (g, V ), they map CD (g, V )ε to CD (g, V )ε+1 for any p ∈ Z and ε ∈ Z2 := Z/2Z. The following result easily follows from Lemma 3.2. Lemma 3.4. The operator δ defined by (3.14) is an odd g-invariant in W ⊗ U(g), that is, it is of degree 1 with respect to the Z2 -grading and satisfies [Γ(A) ⊗ 1 + 1 ⊗ A, δ] = 0,

∀ A ∈ g.

Proof. It is clear that δ is odd. By Lemma 3.2, Γ(g) is isomorphic to the adjoint module for g, and Πg∗ ⊂ W to the dual adjoint module with reversed parity. Therefore, δ1 and δ2 are both g-invariant, and hence so is also δ. We can also verify this by direct computation. We have [Γ(A) ⊗ 1 + 1 ⊗ A, δ1 ]    = (−1)[Ea ] [Γ(A), Xa ] ⊗ Ea + (−1)[A]([Ea ]+1) Xa ⊗ [A, Ea ] a

=−

 (−1)[Ea ]+[A][Ec ] Xc ⊗ ad(A)ac Ea a,c

+

 a,c

= 0.

(−1)[Ea ]+[A]([Ea ]+1) Xa ⊗ ad(A)ca Ec

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

15

This shows the invariance of δ1 . We can also show that [Γ(A), a (−1)[Ea ] Xa Γa ] = 0 by a similar computation. This implies the g-invariance of δ2 . 2 We have the following lemma, which is of crucial importance for the remainder of this paper. Lemma 3.5. The operators δ1 and δ2 satisfy the following relations. [δ1 , δ2 ] + [δ2 , δ1 ] = −[δ1 , δ1 ],

[δ2 , δ2 ] = 0.

Hence it follows that the operator δ = δ1 + δ2 satisfies (δ)2 = 0. Proof. The proof is by direct computation, which is relatively straightforward, but we nevertheless present details of the proof because of the importance of the lemma. Note that we only need to prove the first part of the lemma, as the second part follows from the first. Let us start by showing that [δ1 , δ1 ] = −



(−1)[Eb ] Xa Xb ⊗ [Eb , Ea ].

(3.15)

a,b∈Λ

Clearly [δ1 , δ1 ] is equal to 

  (−1)[Ea ]+[Eb ] Xa Xb ⊗ Ea Eb (−1)[Xb ][Ea ] + Xb Xa ⊗ Eb Ea (−1)[Xa ][Eb ] .

a,b∈Λ

Since Xa Xb = (−1)[Xa ][Xb ] Xb Xa and [Xc ] − [Ec ] ≡ 1 (mod 2) for all c, we have [δ1 , δ1 ] =



(−1)[Xb ] Xa Xb ⊗ [Eb , Ea ] = −

a,b∈Λ



(−1)[Eb ] Xa Xb ⊗ [Eb , Ea ].

a,b∈Λ

This proves equation (3.15). We note that [δ1 , δ2 ] = [δ2 , δ1 ]. Now 2[δ1 , δ2 ] =



(−1)[Ea ]+[Eb ] Xa [Γa , Xb ] ⊗ Eb .

a,b

Using the first relation in (3.10), we obtain

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

16



2[δ2 , δ1 ] =

(−1)[Ea ]+[Eb ] Xa Xc ⊗ ad(Ec )ba Eb

a,b,c



=

(−1)[Ec ] Xa Xc ⊗ ad(Ec )ba Eb

a,b,c

=



(−1)[Ec ] Xa Xc ⊗ [Ec , Ea ].

a,c

Hence [δ1 , δ2 ] + [δ2 , δ1 ] = −[δ1 , δ1 ] by (3.15), proving the first relation in the lemma. Now consider [δ2 , δ2 ]. The property [A, BC] = [A, B]C +(−1)[A][B] B[A, C] of the super commutator leads to

[δ2 , δ2 ] =

1 (−1)[Ea ]+[Eb ] Xa [Γa , Xb ]Γb 2 a,b

+

1 (−1)([Ea ]+1)[Eb ] Xa Xb [Γa , Γb ]. 4 a,b

Using Lemma 3.2, we obtain [δ2 , δ2 ] = −

1 (−1)([Ea ]+1)[Eb ] Xa Xb [Γa , Γb ] 4 a,b

=−

1 4

(−1)([Ea ]+1)[Eb ] Xa Xb ad(Ea )cb Γc .

a,b

As this operator is tri-linear in the elements Xf and linear in the elements ∂f , it vanishes if and only if [δ2 , δ2 ](Xf ) = 0 for all f ∈ Λ. Now by Lemma 3.2,

[δ2 , δ2 ](Xf ) =

1  (−1)([Ea ]+1)[Eb ]+[Ec ][Eg ] Xa Xb Xg ad(Ea )cb ad(Ec )f g . 4 a,b,c,g

Introduce the element Θ := only if Θ = 0. We have

Θ=

f ∈Λ

[δ2 , δ2 ](Xf ) ⊗ Ef in W ⊗ g. Then [δ2 , δ2 ] = 0 if and

1  (−1)([Ea ]+1)[Eb ]+([Ea ]+[Eb ])[Eg ] Xa Xb Xg ⊗ [[Ea , Eb ], Eg ] 4 a,b,g

=

1 (−1)[Eb ] Xa Xb Xc ⊗ [Ec , [Eb , Ea ]]. 4 a,b,c

The right hand side can clearly be re-written as

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

17

1  (−1)[Eb ] Xa Xb Xc ⊗ [Ec , [Eb , Ea ]] 12 a,b,c

+ (−1)[Ec ] Xb Xc Xa ⊗ [Ea , [Ec , Eb ]]  + (−1)[Ea ] Xc Xa Xb ⊗ [Eb , [Ea , Ec ]] . By the super commutativity property of Xf , this can be further re-written as  1  (−1)[Eb ] Xa Xb Xc ⊗ [Ec , [Eb , Ea ]] 12 a,b,c

+ (−1)[Ea ]([Eb ]+[Ec ]) [Ea , [Ec , Eb ]]  + (−1)[Ec ]([Ea ]+[Eb ]) [Eb , [Ea , Ec ]] , which vanishes identically by the Jacobian identity of g. This completes the proof. 2 The following result is an obvious corollary of Lemma 3.5. p Theorem 3.1. Given a sub superspace D ⊂ Πg∗ and a g-module V , set CD (g, V ) = F(D)p ⊗ V for all p ∈ Z (cf. (3.11)). Let δ be the operator defined by (3.14). Then there is the following differential complex −2 −1 0 1 · · · −→ CD (g, V ) −→ CD (g, V ) −→ CD (g, V ) −→ CD (g, V ) −→ · · · , δ

δ

δ

δ

δ

p which will be denoted by (CD (g, V ), δ), and its homology groups by HD (g, V ).

Proof. It follows Lemma 3.5 that (CD (g, V ), δ) is indeed a differential complex. 2 Remark 3.4. If g is an ordinary Lie algebra, then Πg∗ is purely odd, and WΠg∗ is a Clifford −p dim g−p algebra. Therefore FΠg∗ (0) = FΠg∗ (Πg∗ ), and hence HΠg (g, V ) for ∗ (g, V ) = H all p = 0, 1, . . . , dim g. 3.5. Basic properties of the BRST cohomology The BRST cohomology of Lie superalgebras defined in Theorem 3.1 has usual properties expected of a cohomology theory. In particular, we have the following result. Lemma 3.6. Retain the notation of Theorem 3.1, and fix an arbitrary mixing set D ⊂ Πg∗ . Then each g-module homomorphism ϕ : V −→ V  induces a homomorphism of differential complexes ϕ˜D : CD (g, V ) −→ CD (g, V  ), and hence a homomorphism of the ∗ ∗ cohomology groups ϕ˜∗D : HD (g, V ) −→ HD (g, V  ).

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

18

Proof. Everything will be clear if we can show that ϕ˜D (which is defined by (3.12)) commutes with the differential operator. Now it follows from the definitions of the operators δ1 and δ2 that they commute with ϕ˜D , and hence so does also δ = δ1 + δ2 . 2 p p Hence given any mixing set D, we have functors HD (g, −) : V → HD (g, V ) (for p ∈ Z) from the category of g-modules to the category of vector superspaces. These functors have the following property.

Theorem 3.2. Retain the notation of Theorem 3.1, and fix a mixing set D ⊂ Πg∗ . Then β α each short exact sequence 0 −→ V  −→ V −→ V  −→ 0 of g-modules induces a long exact sequence of cohomology groups β˜∗

δ˜∗

α ˜∗

β˜∗

D D D D k−1 k k k · · · −→ HD (g, V  ) −→ HD (g, V  ) −→ HD (g, V ) −→ HD (g, V  )

δ˜∗

α ˜∗

β˜∗

D D D k+1 k+1 −→ HD (g, V  ) −→ HD (g, V ) −→ ...,

∗ where δ˜D is the connecting map defined in the proof of the theorem below.

Proof. We regard the exact functor of Lemma 3.3 as a functor from g-modules to Z-graded vector superspaces. Hence the given short exact sequence of g-modules induces the following short exact sequence. β˜D

D 0 −→ CD (g, V  ) −→ CD (g, V ) −→ CD (g, V  ) −→ 0.

α ˜

By Lemma 3.6, this is a short exact sequence of differential complexes. It leads to the long ∗ ∗ exact sequence of cohomology groups by standard arguments, where α ˜D and β˜D are re˜ ˜ D and βD , and the connecting homomorphism spectively the naturally induced maps of α ∗ is defined in the usual way (see, e.g., [21, §1.3]) as follows. δ˜D ∗ k . Given any [c] ∈ HD (g, V  ) with a representative c, there exists b ∈ Definition of δ˜D k (g, V ) such that c = β˜D (b) by the surjectivity of β˜D . Now δD (c) = δD (β˜D (b)) = CD ˜ αD ) by the short exactness of the above sequence βD (δD (b)) = 0, and hence δD (b) ∈ im (˜

k+1 of differential complexes. Thus there exists an a ∈ CD (g, V  ) such hat δD (b) = α ˜ D (a), and hence α ˜ D (δD (a)) = 0. The injectivity of α ˜ D implies that a is closed, and [a] ∈ k+1 ∗ (g, V  ). The connecting morphism δ˜D is defined to be the map which sends [c] to HD [a]. To see that the map is well defined, we note that any representative of [c] must be k−1 k−1 of the form c + δD (c ) for some c ∈ CD (g, V  ). There exists b ∈ CD (g, V ) such     ˜ ˜ that c = βD (b ), and hence c + δD (c ) = βD (b + δD (b )). We have δD (b + δD (b )) = ˜ D (a) for a unique a. Hence the element a is independent of the choice of the δD (b) = α representative for [c]. Given c, the element b is determined up to ker(β˜D ). Replacing b ˜ D (a1 ). Obviously by b + b1 with b1 ∈ ker(β˜D ) will change a to a + δD (a1 ) with b1 = α a + δD (a1 ) and a are in the same homological class. This shows that the connecting

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

19

∗ morphism δ˜D is independent of the choices of representatives c and b, completing the proof of the theorem. 2

3.6. Mixed cohomology of Lie superalgebras 3.6.1. BRST formulation of the standard cohomology We now show that the standard cohomology of Lie superalgebras given in Definition 2.1 is the special case of the cohomology in Theorem 3.1 with D = 0. We write F = F(0), then F = S(Πg∗ ) ∼ = g∗ . By Section 3.1.1, the vector superspace isomor p ∗ g are even (resp., odd) for even (resp., odd) p. Thus we have the phisms Fp ∼ = isomorphisms of vector superspaces ∼ =

ιp : C p (g, V ) −→ Fp ⊗ V,

∀ p,

which are even (resp., odd) for even (resp., odd) p. Theorem 3.3. (1) Define the following linear maps δ p := ιp+1 dp ι−1 p : Fp ⊗ V −→ Fp+1 ⊗ V

for all p.

(3.16)

Then there is a differential complex (F ⊗ V, δ) :

δ0

δ1

δ2

F0 ⊗ V −→ F1 ⊗ V −→ F2 ⊗ V −→ · · · ,

which is isomorphic to the differential complex (C(g, V ), d) in Definition 2.1. (2) The restriction of δi (i = 1, 2) to Fp ⊗ V coincides with ιp di ι−1 p for all p, and hence δ p = δ|Fp ⊗V ,

∀ p.

(3.17)

Proof. Part (1) is clear, thus we only need to prove part (2). By inspecting the structure of the operators δ1 and δ2 , one can see that the theorem is valid if g is an ordinary Lie algebra, that is, a purely even Lie superalgebra. This is a well known fact from the study of BRST cohomology of Lie algebras in the physics literature. To generalise this to Lie superalgebras with non-trivial odd parts, we only need to check that the operators lead to the correct signs in (2.5) and (2.6), and it is indeed true. 2 We may restate the above theorem as follows, obtaining a BRST reformulation of the standard cohomology of Lie superalgebras [10,17] given in Definition 2.1. Corollary 3.2. When D = 0, the differential complex (C0 (g, V ), δ) of Theorem 3.1 is isomorphic to the complex of g-modules in Definition 2.1.

20

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

Note that the superspace isomorphism of co-chains of degree p ∈ Z is even (resp., odd) if p is even (resp., odd) with respect to the Z2 -grading. Remark 3.5. Lemma 3.4 and also Lemma 3.7 below describe explicitly the g-module structure of the differential complex (C0 (g, V ), δ). 3.6.2. Mixed cohomology of Lie superalgebras Recall the g-action on the Weyl superalgebra W given in Lemma 3.1. If D ⊂ Πg∗ ⊂ W is a g-submodule, then so is also D⊥ = {∂ ∈ (Πg∗ )∗ | ∂(w) = 0, ∀ w ∈ D}. It follows that the left ideal J(D) :=

 ∂∈D ⊥

W∂+



Wx

(3.18)

x∈D

is also a g-submodule of W , and hence F(D) = W /J(D) is a quotient module. The following result is immediate. Lemma 3.7. Assume that the mixing set D of a W -module F(D) is a g-submodule of Πg∗ . Then for any g-module V , the superspace of mixed co-chains CD (g, V ) = F(D) ⊗ V has the structure of a Z-graded g-module. The following result is an obvious corollary of Lemma 3.7 and Lemma 3.4. Theorem 3.4. Assume that the mixing set D of a W -module F(D) is a g-submodule of Πg∗ . Then δ : CD (g, V ) −→ CD (g, V ) is an odd g-module homomorphism which has degree 1 with respect to the Z-grading of CD (g, V ), and in this sense the differential complex (CD (g, V ), δ) of Theorem 3.1 is a complex of g-modules. This enables us to make the following definition. Definition 3.1. Retain the setting of the theorem above. Call (CD (g, V ), δ) a mixed com∗ plex of g-modules, and its homology groups, denoted by HD (g, V ), mixed cohomology groups of g with coefficients in V . Remark 3.6. As we have already seen, in the special case with D = 0, we have C0 (g, V ) = F(0) ⊗ V , and the complex (CD (g, V ), δ) is the generalised Chevalley-Eilenberg complex for the standard cohomology of the Lie superalgebra g. 3.6.3. Completion Note that if D = 0 (resp. D = Πg∗ ), then homogeneous subspaces of F(D) are all finite dimensional if g is finite dimensional, and there exist no homogeneous subspaces of negative (resp. positive) degrees. However, if the even subspace of D is a proper subspace of V0 , then F(D)p is infinite dimensional for all p ∈ Z. Thus there is the possibility to complete each F(D)p by introducing a topology.

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

21

Each F(D)p is filtered by the degree of Πg∗ . Explicitly, we let F(D)kp be the subspace of F(D)p spanned by monomials of Xa ∈ / D and ∂b ∈ / D⊥ which are of order ≤ k in the Xa ’s (cf. (3.18)). Then F(D)0p ⊂ F(D)1p ⊂ F(D)2p ⊂ · · · . k

We take the complements F(D)p of F(D)kp in F(D)p as the fundamental system of open k

neighbourhoods at 0. Then arbitrary unions of finite intersections of F(D)p are the open sets of the topology which we wish to define.  the completion of F(D)p in this topology. Denote by F(D) p Lemma 3.8. Retain the notation above. Assume that the mixing set D ⊂ Πg∗ is a  ⊗ V . Then there exists the following differ p (g, V ) = F(D) g-submodule, and let C D p ential complex of g-modules. δ δ δ δ δ  −2 (g, V ) −→  −1 (g, V ) −→  0 (g, V ) −→  1 (g, V ) −→ · · · −→ C C C C ··· . D D D D

 p (g, V ), and called them comDenote the cohomology groups of this complex by H D pleted mixed cohomology groups of g with coefficients module V . 4. Examples: mixed u-cohomologies of glm|n Let g = glm|n , and let V = C m|n be the natural module for glm|n . Choose the standard basis vi , i ∈ I := I0 ∪ I1 for V , where I0 = {1, ..., m}, I1 = {m + 1, ..., m + n}, with vi being even for i ∈ I0 and odd for i ∈ I1 . Let {Eij | i, j ∈ I} be the set of matrix units relative to this basis, which forms a homogeneous basis of glm|n . We will consider mixed u-cohomologies of g for two choices of parabolic subalgebras p = l ⊕ u with coefficients in V . It is essential to understand u-cohomology with coefficients in simple g-modules in order to understand parabolic induction. Let W = WΠu∗ be the Weyl superalgebra over Πu∗ , the dual superspace of u with parity reversed. We study the mixed u-cohomologies in the setting of the Weyl superalgebra W . 4.1. The case with Levi subalgebra glm ⊕ gln Consider the parabolic subalgebra p := glm ⊕ gln ⊕ u with the nilpotent ideal u = ⊕i∈I0 , j∈I1 CEij . Let xij , i ∈ I0 , j ∈ I1 be basis elements of Πu∗ ⊂ W such that π −1 (xij )(Ek ) = δik δj . We denote the corresponding derivations by ∂xij = ∂x∂ij . The differential operator for the u-cohomology is the negative of δ=

 i∈I0 ,j∈I1

xij ⊗ Eij .

(4.1)

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

22

4.1.1. Standard u-cohomology The superspace of cochains is C(u, V ) = A ⊗ V , where A = C[xij | i ∈ I0 , j ∈ I1 ] is

the space of polynomials in xij ’s. Let f = i∈I fi vi be an arbitrary element in C(u, V ),



where fi ∈ A. We have δ(f ) = i∈I0 j∈I1 xij fj vi . Thus  ⎫   ⎬  ker δ = Av1 + · · · + Avm + K1 , K1 = fj vj  xij fj = 0, ∀ i ∈ I0 , ⎭ ⎩  j∈I1 j∈I1 ⎧  ⎫ ⎨   ⎬      im δ = xij fj vi  fj ∈ A, ∀ j ∈ I1 = A xij vi ; ⎩ ⎭  i∈I0 j∈I1 j∈I1 i∈I0 ⎧ ⎨

Av1 + · · · + Avm  + K1 , H ∗ (u, V ) = 

A xij vi j∈I1

i∈I0

where we note that K1 = 0 if and only if m ≥ n. These u-cohomology groups were calculated in [6] using a different method. 4.1.2. The u-cohomology on the dual Fock space We take D = Πu∗ . The space of co-chains is given by CD (u, V ) = B ⊗ V , where B = C[∂xij | i ∈ I0 , j ∈ I1 ] is the dual Fock space such that the vacuum vector 1 satisfies

xij 1 = 0 for i ∈ I0 , j ∈ I1 . Let f = i∈I fi vi be in CD (u, V ), where fi ∈ B. We have



δ(f ) = − i∈I0 ¯ij = ∂xij and ∂x¯ij = ∂ x¯∂ij , ¯ij fj vi , where we have denoted x j∈I1 ∂x thus ∂x¯ij fj is the partial derivative with respect to x ¯ij = ∂xij of the polynomial fj . Hence for D = Πu∗ ,  ⎫  ⎬  ker δ = Bv1 + · · · + Bvm + K2 , K2 = fj vj  ∂x¯ij fj = 0, ∀ i ∈ I0 , ⎭ ⎩  j∈I1 j∈I1 ⎧  ⎫ ⎨   ⎬   im δ = ∂x¯ij fj vi  fj ∈ B, ∀ j ∈ I1 ; ⎩ ⎭  ⎧ ⎨

i∈I0

j∈I1

∗ HD (u, V ) =  

i∈I0

Bv1 + · · · + Bvm    + K2 .  ∂x¯ij fj vi  fj ∈ B, ∀ j ∈ I1

j∈I1

4.1.3. Mixed u-cohomology We take D = ⊕m1 +1≤i≤m, j∈I1 Cxi,j for any m1 ∈ Z with 1 ≤ m1 < m. Then the superspace of co-chains is CD (u, V ) = C ⊗ V , where C = C[xij , x ¯kj | 1 ≤ i ≤ m1 < k ≤ m, j ∈ I1 ] is the space of polynomials in xij ’s and x ¯kj ’s with x ¯ij = ∂xij . For any

f = i∈I fi vi with fi ∈ C, we have

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

δ(f ) =

m1    i=1

 xij fj vi −

m   i=m1 +1

j∈I1

23

 ∂x¯ij fj vi .

j∈I1

Thus ker δ = Cv1 + · · · + Cvm + K3 , ⎧  ⎫  ⎨ ⎬    K3 = fj vj  xij fj = ∂x¯kj fj = 0, 1 ≤ i ≤ m1 < k ≤ m , ⎩ ⎭  j∈I1 j∈I1 j∈I1  ⎫ ⎧ m1   m ⎬ ⎨      xij fj vi − ∂x¯ij fj vi  fj ∈ C, ∀ j ∈ I1 ; im δ = ⎭ ⎩  i=1

i=m1 +1

j∈I1

∗ HD (u, V ) =  m1 



i=1

 xij fj

j∈I1

Cv1 + · · · + Cvm 

   + K3 . m

 vi − ∂x¯ij fj vi  fj ∈ C, ∀ j ∈ I1 i=m1 +1

j∈I1

j∈I1

4.2. The case with Levi subalgebra glm1 ⊕ glm−m1 |n Consider the parabolic subalgebra p = l ⊕ u with l = glm1 ⊕ glm−m1 |n and u = ⊕1≤i≤m1 ,m1 +1≤j≤m+n CEij for any m1 ∈ Z with 1 ≤ m1 < m. Then Πu∗ has basis θij , xik with 1 ≤ i ≤ m1 < j ≤ m < k ≤ m + n. The differential operator is given by δ=

m1   m  i=1

θij ⊗ Eij −

j=m1 +1



 xik ⊗ Eik .

k∈I1

4.2.1. Standard u-cohomology The superspace of co-chains is D ⊗ V , where D = C[θij , xik | 1 ≤ i ≤ m1 < j ≤ m < k ≤ m + n] is the polynomial ring in θij ’s and xik ’s with θij being odd satisfying

m1 m 2 θij = 0. For any f = i∈I fi vi with fi ∈ D, we have δ(f ) = i=1 j=m1 +1 θij fj −

k∈I1 xik fk vi . Hence ker δ = Dv1 + · · · + Dvm1 + K4 , ⎧  ⎫   m ⎨ m+n ⎬    K4 = fi vi  θij fj − xik fk = 0, 1 ≤ i ≤ m1 , ⎩ ⎭  j=m1 +1 i=m1 +1 k∈I1 ⎧  ⎫ m1   m ⎨ ⎬    im δ = θij fj − xik fk vi  fj ∈ D, m1 < j ≤ m + n , ⎩ ⎭  i=1

j=m1 +1

H ∗ (u, V ) =  m1 

i=1

m

j=m1 +1

k∈I1

Dv1 + · · · + Dvm1    + K4 .

 θij fj − xik fk vi  fj ∈ D, m1 < j ≤ m + n k∈I1

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

24

4.2.2. The u-cohomology on dual Fock space The superspace of co-chains is E ⊗V , where E = C[θ¯ij , x ¯ik | 1 ≤ i ≤ m1 < j ≤ m < k ≤ ¯ij ’s with θ¯ij = ∂θij and x ¯ik = ∂xik . For any m + n] is the polynomial ring on θ¯ij ’s and x

m1 m

f = i∈I fi vi with fi ∈ E, we have δ(f ) = i=1 ∂ f ¯ik fk vi , k∈I1 ∂x j=m1 +1 θ¯ij j − where ∂θ¯ij = ∂ θ∂¯ , ∂x¯ik = ∂ x¯∂ik . Hence ij

ker δ = Dv1 + · · · + Dvm1 + K5 ,  ⎫ ⎧   m ⎬ ⎨ m+n    K5 = fi vi  ∂θ¯ij fj − ∂x¯ik fk = 0, 1 ≤ i ≤ m1 , ⎭ ⎩  j=m1 +1 i=m1 +1 k∈I1  ⎫ ⎧ m1   m ⎬ ⎨    ∂θ¯ij fj − ∂x¯ik fk vi  fj ∈ D, m1 < j ≤ m + n , im δ = ⎭ ⎩  i=1

j=m1 +1

H ∗ (u, V ) =  m1 

i=1

m

j=m1 +1

k∈I1

Dv1 + · · · + Dvm1  + K5 .  

 ∂θ¯ij fj − ∂x¯ik fk vi  fj ∈ D, m1 < j ≤ m + n k∈I1

4.3. The u-cohomologies for Kac modules In this section we take the parabolic subalgebra of glm|n given in Section 4.1, and let K = Kλ = U(glm|n ) ⊗U(p) L0λ be the Kac module with a typical highest weight λ, where L0λ is the simple p-module with highest weight λ. We consider the mixed u-cohomologies 0 with coefficients in K. Let K λ = {v ∈ K | Eji v = 0, ∀i ∈ I0 , j ∈ I1 }, the lowest weight component of K. Since λ is typical, K is a free module over U(u) and we have 0 K = U(u) ⊗ K λ . Note that U(u) = C[Eij | i ∈ I0 , j ∈ I1 ], where Eij ’s are regarded as Grassmannian variables. The differential is still given by (4.1). 4.3.1. The case of the standard Fock space The u-cohomology in this case can be deduced from the generalised Kazhdan-Lusztig theory for parabolic induction of glm|n [5,6,15]; this is not difficult, but is very indirect. Within the BRST framework adopted here, we can read off the cohomology quite directly from a gl1|1 -action on the superspace of co-chains. This gl1|1 -structure is interesting in itself. Now the superspace of co-chains is given by 0

C(u, K) = C[xij , Eij | i ∈ I0 , j ∈ I1 ] ⊗ K λ . Let us introduce the following operators on C[xij , Eij | i ∈ I0 , j ∈ I1 ]:

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

e=



xij Eij ,



f=

i∈I0 ,j∈I1

i∈I0 ,j∈I1



h1 = mn +

xij

i∈I0 ,j∈I1

∂ , ∂xij

25

∂ ∂ , ∂xij ∂Eij 

h2 =

Eij

i∈I0 ,j∈I1

∂ . ∂Eij

They span the Lie superalgebra gl1|1 ; note in particular that [e, f ] = ef + f e = h1 − h2 .

(4.2)

The superspace C[xij , Eij | i ∈ I0 , j ∈ I1 ] is a weight module for this gl1|1 with respect to the Cartan subalgebra spanned by h1 , h2 . We have the following weight space decomposition:  C[xij , Eij | i ∈ I0 , j ∈ I1 ] = C[xij , Eij | i ∈ I0 , j ∈ I1 ](mn+k,α) . (4.3) k∈Z+ , 0≤α≤mn

To consider the cohomology, we note that the differential operator (4.1) can be rewritten as δ = e ⊗ idK 0 . λ

0

Thus ker δ = ker e ⊗ K λ and H ∗ (u, K) = (4.2) we have

ker e im e

0

⊗ K λ . For any p ∈ (ker e)(mn+k,α) , by

ef (p) = [e, f ](p) = (mn + k − α)p. This shows that p ∈ im e unless mn + k − α = 0, i.e., k = 0 and α = mn. For p ∈ (ker e)(mn,mn) , if p ∈ im e, then there exists q such that p = e(q). This requires that q has weight (mn − 1, mn − 1), which is not a weight of C[xij , Eij | i ∈ I0 , j ∈ I1 ] by (4.3). This shows that (im e)(mn,mn) = 0. Hence H 0 (u, K) ∼ = L0λ ,

H i (u, K) = 0 for all i > 0.

4.3.2. The case of the dual Fock space We write yij = ∂x∂ij . Then the superspace of co-chains is 0

CΠu∗ (u, K) = C[yij , Eij | i ∈ I0 , j ∈ I1 ] ⊗ K λ . Define the following operators on C[yij , Eij | i ∈ I0 , j ∈ I1 ]: 

e=

Eij

i∈I0 ,j∈I1

h1 =

 i∈I0 ,j∈I1

∂ , ∂yij

∂ Eij , ∂Eij



f=

yij

i∈I0 ,j∈I1

h2 =

 i∈I0 ,j∈I1

∂ , ∂Eij yij

∂ , ∂yij

26

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

which give rise to a gl1|1 -action. We note in particular the relation [e, f ] = ef + f e = h1 + h2 . The Cartan subalgebra Ch1 + Ch2 is clearly diagonalisable; we have the weight space decomposition  C[yij , Eij | i ∈ I0 , j ∈ I1 ] = C[yij , Eij | i ∈ I0 , j ∈ I1 ](α,k) . 0≤α≤mn, k∈Z+ ∗ Now δ = −e ⊗ idK 0 , thus HΠu ∗ (u, K) = λ the last section lead to 0 ∼ 0 HΠu ∗ (u, K) = K λ ,

ker e im e

0

⊗ K λ . Similar considerations as those in

−i HΠu ∗ (u, K) = 0 for all i > 0.

4.4. An example of completed cohomology groups We consider the completed u-cohomology in the case gl1|2 ⊃ p = gl1 ⊕ gl2 ⊕ u ⊃ u = CE12 + CE13 and V = C 1|2 . Note that u is purely odd. The differential can be written as the negative of δ = x ⊗E12 +y ⊗E13 , where x = x12 , y = x13 are both even variables. We take D = Cy, and write y¯ = ∂y . Denote the completion of F(D) (cf. Section 3.6.3) by  which is generated by C[x, y¯] and formal power series in x¯ y . We have the differential D, complex  ⊗ V −→ D  ⊗ V. δ:D  we have δ(f ) = (xf2 − For any f = f1 v1 + f2 v2 + f3 v3 with fi ∈ D,  1 . Now f ∈ ker δ if and only if to see that im δ = Dv xf2 −

∂ ∂ y¯ f3 )v1 .

∂ f3 = 0. ∂ y¯

It is easy

(4.4)

We can solve this at each fixed degree, obtaining the following complete set of solutions f2 = xp Pp (x¯ y ),

f3 = xp Pp (x¯ y ),

  (x¯ f2 = y¯−p Q y −p p y ) − p¯

 p (x¯ Q y) , x¯ y

p ≥ 0;  p (x¯ f3 = y¯−p Q y ),

p < 0;

 p is any power series without a constant term. We where Pp is any power series, and Q  the C-span of all f2 v2 + f3 v3 with such f2 and f3 . Then denote by K  1 + K,  ker δ = Dv

 1, im δ = Dv

 ∗ (u, V ) = K.  H D

Remark 4.1. In order for the mixed u-cohomology groups to be l-modules, we need to take D to be an l-submodule of Πg∗ .

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

27

5. Examples: BRST cohomology of classical Lie superalgebras 5.1. BRST cohomology of classical Lie superalgebras We return to the setting of Theorem 3.1. If U(g) has a non-trivial centre, then the result of [17, Proposition 2.2] generalises to the present context in a straightforward manner. Theorem 5.1. Let C be a homogeneous central element of U(g) (i.e., an even or odd Casimir element of g), which is not linear in g and does not contain a constant term. Given a U(g)-module V , denote by νV : U(g) −→ EndC (V ) the corresponding representation of U(g). If the linear operator νV (C) is invertible, then for any mixing set p D ⊂ Πg∗ , the BRST cohomology groups satisfy HD (g, V ) = 0 for all p ∈ Z. Proof. The proof is essentially the same as the proof of [17, Proposition 2.2]. There

exists a set of elements Ca ∈ U(g) (a ∈ Λ) such that C = a∈Λ Ca Ea , see [17, (2.77)].

Define the operator δ¯ = a∈Λ (−1)([Ea ]+1)[Ca ] ∂a ⊗ Ca . Then by direct computation one can verify that ¯ δ] = 1 ⊗ νV (C). [δ,

(5.1)

¯ ) = (−1)[C] (1 ⊗ νV (C))f . Clearly 1 ⊗ C super For any f ∈ ker δ ⊂ CD (g, V ), we have δ δ(f commutes with δ, thus ¯ ). f = δ(1 ⊗ νV (C)−1 )δ(f This shows that f ∈ im δ, proving the theorem. 2 The universal enveloping algebras of the simple basic classical Lie superalgebras (i.e., the A(m | n), B(m|n), C(n), D(m|n) series, and D(2, 1; α), F (4) and G(3)) have nontrivial centres, and so does also the universal enveloping algebra of the general linear superalgebra glm+1|n+1 . If g is one of these Lie superalgebras, we let Og,b denote the category O of g-modules defined with respect to the choice of any Borel subalgebra b of 0 g. Then Og,b decomposes into a direct sum of blocks. We denote by Og,b the principal block, that is, the block contains the module C such that all elements of g act by zero. Corollary 5.1. Let g be a simple basic classical Lie superalgebra or the general linear superalgebra, and fix an arbitrary mixing set D ⊂ Πg∗ . For any object V in Og,b , the q BRST cohomology groups HD (g, V ) vanish for all q ∈ Z unless V has a summand 0 belonging to the principal block Og,b . Proof. If V does not contain a summand in the principal block, there exists a Casimir element C satisfying the given conditions of Theorem 5.1 such that the action of C on V

28

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

q is invertible. Hence it immediately follows Theorem 5.1 that HD (g, V ) = 0 for all q ∈ Z, completing the proof. 2

5.2. Concluding remarks As already mentioned, the standard cohomology of Lie superalgebras [10,17] has many important applications, and we also expect it to be the case for the mixed cohomology of Lie superalgebras. In particular, we hope that the mixed cohomology will enrich the theory of support varieties of Lie supergroups, and help to remedy failures in the context of Lie supergroups and algebraic supergroups of major classical theorems such as the Bott-Borel-Weil theorem and Beilinson-Bernstein localisation theorem. Acknowledgments Both authors were supported by Grant DP170104318 of the Australian Research Council; and Su was also supported by Grants 11971350 and 11431010 of the National Natural Science Foundation of China. References [1] I.N. Bernstein, D.A. Leites, Integral forms and the Stokes formula on supermanifolds, Funkc. Anal. Prilozh. 11 (1) (1977) 55–56 (in Russian). [2] Brian D. Boe, Jonathan R. Kujawa, Daniel K. Nakano, Cohomology and support varieties for Lie superalgebras, Trans. Am. Math. Soc. 362 (12) (2010) 6551–6590. [3] A.J. Bracken, M.D. Gould, R.B. Zhang, Quantum supergroups and solutions of the Yang-Baxter equation, Mod. Phys. Lett. A 5 (11) (1990) 831–840. [4] F. Brandt, G. Barnich, M. Henneaux, Local BRST cohomology in gauge theories, Phys. Rep. 338 (5) (2000) 439–569. [5] J. Brundan, Kazhdan–Lusztig polynomials and character formulae for the Lie superalgebra gl(m|n), J. Am. Math. Soc. 16 (2003) 185–231. [6] Shun-Jen Cheng, R.B. Zhang, Analogue of Kostant’s u-cohomology formula for the general linear superalgebra, Int. Math. Res. Not. (1) (2004) 31–53. [7] C. Chevalley, S. Eilenberg, Cohomology theory of Lie groups and Lie algebras, Trans. Am. Math. Soc. 63 (1948) 85–124. [8] Pierre Deligne, John W. Morgan, Notes on supersymmetry (following Joseph Bernstein), in: Quantum Fields and Strings: A Course for Mathematicians, vols. 1, 2, Princeton, NJ, 1996/1997, Amer. Math. Soc., Providence, RI, 1999, pp. 41–97. [9] D.B. Fuks, Cohomology of Infinite-Dimensional Lie Algebras, Consultants Bureau, New York, 1986. [10] D.B. Fuchs, D.A. Leites, Cohomology of Lie superalgebras, C. R. Acad. Bulgare Sci. 37 (12) (1984) 1595–1596. [11] Caroline Gruson, Vera Serganova, Cohomology of generalized supergrassmannians and character formulae for basic classical Lie superalgebras, Proc. Lond. Math. Soc. (3) 101 (3) (2010) 852–892. [12] V.G. Kac, Lie superalgebras, Adv. Math. 26 (1977) 8–96. [13] Gustav I. Lehrer, Daniel K. Nakano, Ruibin Zhang, Detecting cohomology for Lie superalgebras, Adv. Math. 228 (4) (2011) 2098–2115. [14] Ivan Penkov, Vera Serganova, Cohomology of G/P for classical complex Lie supergroups G and characters of some atypical G-modules, Ann. Inst. Fourier 39 (4) (1989) 845–873. [15] Vera Serganova, Kazhdan-Lusztig polynomials and character formula for the Lie superalgebra gl(m|n), Sel. Math. New Ser. 2 (4) (1996) 607–651. [16] M. Scheunert, The Theory of Lie Superalgebras: An Introduction, Lecture Notes in Mathematics, vol. 716, Springer-Verlag, Berlin, Heidelberg, New York, 1979.

Y. Su, R.B. Zhang / Journal of Algebra 549 (2020) 1–29

29

[17] M. Scheunert, R.B. Zhang, Cohomology of Lie superalgebras and their generalizations, J. Math. Phys. 39 (9) (1998) 5024–5061. [18] Yucai Su, R.B. Zhang, Cohomology of Lie superalgebras slm|n and osp2|2n , Proc. Lond. Math. Soc. (3) 94 (1) (2007) 91–136. [19] Yucai Su, R.B. Zhang, Character and dimension formulae for general linear superalgebra, Adv. Math. 211 (2007) 1–33. [20] V.S. Varadarajan, Supersymmetry for Mathematicians: An Introduction, Courant Lectures Notes, American Mathematical Society, Providence, 2004. [21] C.A. Weibel, An Introduction to Homological Algebra, Cambridge University Press, 1994. [22] E. Witten, Notes on supermanifolds and integration, arXiv:1209.2199. [23] R.B. Zhang, Quantum enveloping superalgebras and link invariants, J. Math. Phys. 43 (4) (2002) 2029–2048. [24] R.B. Zhang, Quantum superalgebra representations on cohomology groups of non-commutative bundles, J. Pure Appl. Algebra 191 (3) (2004) 285–314.