Keywords

1 Introduction

This note is based on lectures given at the Centro di Ricerca Matematica Ennio De Giorgi, Pisa, in Winter of 2014–2015. They are aimed as an introduction to W-algebras and their representation theory. Since W-algebras appear in many areas of mathematics and physics there are certainly many other important topics untouched in the note, partly due to the limitation of the space and partly due to the author’s incapability.

The W-algebras can be regarded as generalizations of affine Kac-Moody algebras and the Virasoro algebra. They appeared [34, 70, 78] in the study of the classification of two-dimensional rational conformal field theories. There are several ways to define W-algebras, but it was Feigin and Frenkel [36] who found the most conceptual definition of principal W-algebras that uses the quantized Drinfeld-Sokolov reduction, which is a version of Hamiltonian reduction. There are a lot of works on W-algebras (see [26] and references therein) mostly by physicists in 1980s and 1990s, but they were mostly on principal W-algebras, that is, the W-algebras associated with principal nilpotent elements. It was quite recent that Kac et al. [60] defined the W-algebra \(\mathcal{W}^{k}(\mathfrak{g},f)\) associated with a simple Lie algebra and its arbitrary nilpotent element f by generalizing the method of quantized Drinfeld-Sokolov reduction.

The advantage of the method of quantized Drinfeld-Sokolov reduction is its functoriality, in the sense that it gives rise to a functor from the category of representations of affine Kac-Moody algebras and to the category of representations of W-algebras. Since it is difficult to study W-algebras directly (as no presentation by generators and relations (OPE’s) is known for a general W-algebra), in this note we spend the most of our efforts in understanding this functor.

Although our methods apply to much more general settings [4, 6, 9, 10, 12] we focus on the W-algebras associated with Lie algebras \(\mathfrak{g}\) of type A and its principal nilpotent element that were originally defined by Fateev and Lykyanov [34]. They can be regarded as affinization of the center of the universal enveloping algebra of \(\mathfrak{g}\) via Konstant’s Whittaker model [65] and Kostant-Sternberg’s description [66] of Hamiltonian reduction via BRST cohomolgy, as explained in [36]. For this reason we start with a review of Kostant’s results and proceed to the construction of BRST complex in the finite-dimensional setting in Sect. 2. W-algebras are not Lie algebras, not even associated algebras in general, but vertex algebras. In many cases a vertex algebra can be considered as a quantization of arc spaces of an affine Poisson scheme. In Sect. 3 we study this view point that is useful in understanding W-algebras and their representation theory. In Sect. 4 we study Zhu’s algebras of vertex algebras that connects W-algebras with finite W-algebras [27, 75]. In Sect. 5 we introduce W-algebras and study their basic properties. In Sect. 6 we start studying representation theory of W-algebras. In Sect. 7 we quickly review some fundamental results on irreducible representations of W-algebras obtained in [5]. One of the fundamental problems (at least mathematically) on W-algebras was the conjecture of Frenkel et al. [44] on the existence and construction of so called the minimal models of W-algebras, which give rive to rational conformal field theories as in the case of the integrable representations of affine Kac-Moody algebras and the minimal models of the Virasoro algebra. In Sect. 8 we give an outline of the proof [10] of this conjecture.

2 Review of Kostant’s Results

2.1 Companion Matrices and Invariant Polynomials

Let \(G = GL_{n}(\mathbb{C})\) be the general linear group, and let \(\mathfrak{g} = \mathfrak{g}\mathfrak{l}_{n}(\mathbb{C})\) be the general linear Lie algebra consisting of n × n matrices. The group G acts on \(\mathfrak{g}\) by the adjoint action: x ↦ Ad(g)x = gxg −1, gG. Let \(\mathbb{C}[\mathfrak{g}]^{G}\) be the subring of the ring \(\mathbb{C}[\mathfrak{g}]\) of polynomial functions on \(\mathfrak{g}\) consisting of G-invariant polynomials.

Recall that a matrix

$$\displaystyle{ A = \left (\begin{array}{*{10}c} 0 & 0 & \cdots & 0 & -a_{1} \\ 1 & 0 & \cdots & 0 & -a_{2} \\ 0 & 1 & \cdot & 0 & -a_{3}\\ \vdots & \vdots & \ddots & \vdots & \vdots \\ 0 & 0 & & 1 & -a_{n} \end{array} \right ) }$$
(1)

is called the companion matrix of the polynomial \(a_{1} + a_{2}t + a_{3}t^{2} +\ldots +a_{n}t^{n-1} + t^{n} \in \mathbb{C}[t]\) since

$$\displaystyle{ \mathop{\mathrm{det}}\nolimits (tI - A) = a_{1} + a_{2}t + a_{3}t^{2} +\ldots +a_{ n}t^{n-1} + t^{n}. }$$
(2)

Let be the affine subspace of \(\mathfrak{g}\) consisting of companion matrices of the form (1).

Lemma 1

For \(A \in \mathfrak{g}\) the following conditions are equivalent.

  1. 1.

    .

  2. 2.

    There exists a vector \(v \in \mathbb{C}^{n}\) such that v, Av, A 2 v, , A n−1 v are linearly independent.

Theorem 1

The restriction map gives the isomorphism

Proof

Let \(f \in \mathbb{C}[\mathfrak{g}]\) be a G-invariant polynomial such that . Then clearly . On the other hand it follows from Lemma 1 that is a Zariski open subset in \(\mathfrak{g}\). Therefore f = 0. To see the surjectiveness define \(p_{1},\ldots,p_{n} \in \mathbb{C}[\mathfrak{g}]^{G}\) by

$$\displaystyle{ \mathrm{det}(tI - A) = t^{n} + p_{ 1}(A)t^{n-1} -\ldots +p_{ n}(A),\quad A \in \mathfrak{g}. }$$

By (2), we have . This completes the proof.

Put

(3)

Note that f is a nilpotent element of \(\mathfrak{g}\), that is, (adf)r = 0 for a sufficiently large r. We have

where

$$\displaystyle{ \mathfrak{a} =\Bigg\{ \left (\begin{array}{*{10}c} 0 & \cdots & 0 & {\ast}\\ \vdots & & \vdots & {\ast} \\ \vdots & & \vdots & {\ast}\\ 0 & \cdots & 0 & {\ast} \end{array} \right )\Bigg\}. }$$

Let \(\mathfrak{b}\), \(\mathfrak{n}\) be the subalgebras of \(\mathfrak{g}\) defined by

$$\displaystyle{ \mathfrak{b} =\Bigg\{ \left (\begin{array}{*{10}c} {\ast} & & &\\ & \ddots & {\ast}& \\ & 0 & \ddots &\\ & && {\ast} \end{array} \right )\Bigg\},\quad \mathfrak{n} =\Bigg\{ \left (\begin{array}{*{10}c} 0 & & &\\ & \ddots & {\ast}& \\ & 0 & \ddots &\\ & && 0 \end{array} \right )\Bigg\} \subset \mathfrak{b}, }$$

and let N be the unipotent subgroup of G corresponding to \(\mathfrak{n}\), i.e.,

$$\displaystyle{ N =\Bigg\{ \left (\begin{array}{*{10}c} 1 & & &\\ & \ddots & {\ast}& \\ & 0 & \ddots &\\ & && 1 \end{array} \right )\Bigg\}. }$$
(4)

Let ( |  ) be the invariant inner product of \(\mathfrak{g}\) defined by \((x\vert y) =\mathop{ \mathrm{tr}}\nolimits (xy)\). This gives a G-equivariant isomorphism \(\mathfrak{g}\;\stackrel{_{\sim }}{\rightarrow }\;\mathfrak{g}^{{\ast}}\).

Define \(\chi \in \mathfrak{n}^{{\ast}}\) by

$$\displaystyle{ \chi (x) = (f\vert x)\quad \text{for}\ x \in \mathfrak{n}. }$$

Note that χ is a character of \(\mathfrak{n}\), that is, \(\chi ([\mathfrak{n},\mathfrak{n}]) = 0\). Hence χ defines a one-dimensional representation of N.

Consider the restriction map

$$\displaystyle{ \mu: \mathfrak{g}^{{\ast}}\rightarrow \mathfrak{n}^{{\ast}}. }$$

Then

$$\displaystyle{ \mu ^{-1}(\chi ) =\chi +\mathfrak{n}^{\perp }\cong f + \mathfrak{b}. }$$

Here \(\mathfrak{g}\) is identified with \(\mathfrak{g}^{{\ast}}\) via ( |  ). Since μ is N-equivariant and χ is a one-point orbit of N, it follows that \(f + \mathfrak{b}\) is stable under the action of N.

Theorem 2 (Kostant [65])

The adjoint action gives the isomorphism

of affine varieties.

Proof

It is not difficult to see that the adjoint action gives the bijection . Since it is a morphism of irreducible varieties and \(f + \mathfrak{b}\) is normal, the assertion follows from Zariski’s Main Theorem (see e.g., [76, Corollary 17.4.8]).

Corollary 1

The restriction map gives the isomorphisms

Proof

By Theorem 2, we have

Hence the assertion follows from Theorem 1.

2.2 Transversality of to G-Orbits

Lemma 2

The affine spaces and \(f + \mathfrak{b}\) intersect transversely at f to AdG ⋅ f.

Proof

We need to show that

(5)

But \(T_{f}\mathfrak{g}\cong \mathfrak{g}\), , \(T_{f}(\mathrm{Ad}G \cdot f)\cong [\mathfrak{g},f]\). The assertion follows since \(\mathfrak{g} = \mathfrak{a} + [\mathfrak{g},f]\).

Using the Jacobson-Morozov theorem, we can embed f into an \(\mathfrak{s}\mathfrak{l}_{2}\)-triple {e, f, h} in \(\mathfrak{g}\). Explicitly, we can choose the following elements for e and h:

$$\displaystyle{ e =\sum _{ i=1}^{n-1}i(n - i)e_{ i,i+1},\quad h =\sum _{ i=1}^{n}(n + 1 - 2i)e_{ i,i}, }$$
(6)

where e i, j denotes the standard basis element of \(\mathfrak{g} =\mathop{ \mathrm{Mat}}\nolimits _{n}(\mathbb{C})\).

The embedding \(\mathfrak{s}\mathfrak{l}_{2} =\mathop{ \mathrm{span}}\nolimits _{\mathbb{C}}\{e,h,f\} \rightarrow \mathfrak{g}\) exponents to a homomorphism SL 2G. Restricting it to the torus \(\mathbb{C}^{{\ast}}\) consisting of diagonal matrices we obtain a one-parameter subgroup \(\gamma: \mathbb{C}^{{\ast}}\rightarrow G\). Set

$$\displaystyle{ \rho: \mathbb{C}^{{\ast}}\ni t\longmapsto t^{2}\mathrm{Ad}\gamma (t) \in GL(\mathfrak{g}). }$$
(7)

Then

$$\displaystyle{ \rho (t)(\,f +\sum _{i\leqslant j}c_{ij}e_{i,j}) = f +\sum _{i\leqslant j}t^{2(i-j+1)}c_{ ij}e_{i,j}. }$$

Thus it define a \(\mathbb{C}^{{\ast}}\)-action on \(\mathfrak{g}\) that preserves \(f + \mathfrak{b}\) and . This action on \(f + \mathfrak{b}\) and contracts to f, that is, ρ(t)xf when t → 0.

Proposition 1

The affine space \(f + \mathfrak{b}\) (resp.  ) intersects AdG ⋅ x transversely at any point \(x \in f + \mathfrak{b}\) (resp. ).

Proof

By Lemma 2 the intersection of \(f + \mathfrak{b}\) with AdG-orbits is transversal at each point in some open neighborhood of f in \(f + \mathfrak{b}\). By the contracting \(\mathbb{C}^{{\ast}}\)-action ρ, it follows that the same is true for all points of \(f + \mathfrak{b}\).

2.3 The Transversal Slice as a Reduced Poisson Variety

The affine variety \(\mathfrak{g}^{{\ast}}\) is equipped with the Kirillov-Kostant Poisson structure: the Poisson algebra structure of \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\) is given by

$$\displaystyle{ \{x,y\} = [x,y]\quad \text{for }x,y \in \mathfrak{g} \subset \mathbb{C}[\mathfrak{g}^{{\ast}}]. }$$

Consider the restriction map \(\mu: \mathfrak{g}^{{\ast}}\rightarrow \mathfrak{n}^{{\ast}}\), which is a moment map for the N-action on \(\mathfrak{g}^{{\ast}}\). That is, μ is a regular N-equivariant morphism that gives the following commutative diagram of Lie algebras:

Here \(\mu ^{{\ast}}: \mathfrak{n} \rightarrow \mathfrak{g} \subset \mathbb{C}[\mathfrak{g}^{{\ast}}]\) is the pullback map, the map \(\mathbb{C}[\mathfrak{g}^{{\ast}}] \rightarrow \mathop{\mathrm{Der}}\nolimits \mathbb{C}[\mathfrak{g}^{{\ast}}]\) is given by ϕ ↦ {ϕ, ?}, and \(\mathfrak{n} \rightarrow \mathop{\mathrm{Der}}\nolimits \mathbb{C}[\mathfrak{g}^{{\ast}}]\) is the Lie algebra homomorphism induced by the coadjoint action of G on \(\mathfrak{g}^{{\ast}}\).

The transversality statement of Proposition 1 for \(f + \mathfrak{b}\) is equivalent to that χ is a regular value of μ. By Theorem 2, the action of N on \(\mu ^{-1}(\chi ) =\chi +\mathfrak{n}^{\perp }\) is free and

Therefore has the structure of the reduced Poisson variety, obtained from \(\mathfrak{g}^{{\ast}}\) by the Hamiltonian reduction.

The Poisson structure of is described as follows. Let

$$\displaystyle{ I_{\chi } = \mathbb{C}[\mathfrak{g}^{{\ast}}]\sum _{ x\in \mathfrak{n}}(x -\chi (x)), }$$

so that

$$\displaystyle{ \mathbb{C}[\mu ^{-1}(\chi )] = \mathbb{C}[\mathfrak{g}^{{\ast}}]/I_{\chi }. }$$

Then can be identified as the subspace of \(\mathbb{C}[\mathfrak{g}^{{\ast}}]/I_{\chi }\) consisting of all cosets \(\phi +\mathbb{C}[\mathfrak{g}^{{\ast}}]I_{\chi }\) such that \(\{x,\phi \}\in \mathbb{C}[\mathfrak{g}^{{\ast}}]I_{\chi }\) for all \(x \in \mathfrak{n}\). In this realization, the Poisson structure on is defined by the formula

$$\displaystyle{ \{\phi +\mathbb{C}[\mathfrak{g}^{{\ast}}]I_{\chi },\phi ' + \mathbb{C}[\mathfrak{g}^{{\ast}}]I_{\chi }\} =\{\phi,\phi '\} + \mathbb{C}[\mathfrak{g}^{{\ast}}]I_{\chi } }$$

for ϕ, ϕ′ such that \(\{x,\phi \},\{x,\phi '\} \in \mathbb{C}[\mathfrak{g}^{{\ast}}]I_{\chi }\) for all \(x \in \mathfrak{n}\).

Proposition 2

We have the isomorphism as Poisson algebras. In particular the Poisson structure of is trivial.

Proof

The restriction map (see Corollary 1) is obviously a homomorphism of Poisson algebras.

In the next subsection we shall describe the above Hamiltonian reduction in more factorial way, in terms of the BRST cohomology (where BRST refers to the physicists Becchi, Rouet, Stora and Tyutin) for later purpose.

2.4 BRST Reduction

Let Cl be the Clifford algebra associated with the vector space \(\mathfrak{n} \oplus \mathfrak{n}^{{\ast}}\) and its non-degenerate bilinear form (⋅ | ⋅ ) defined by ( f + x | g + y) = f( y) + g(x) for \(f,g \in \mathfrak{n}^{{\ast}},x,y \in \mathfrak{n}\). Namely, Cl is the unital \(\mathbb{C}\)-superalgebra that is isomorphic to \(\varLambda (\mathfrak{n}) \otimes \varLambda (\mathfrak{n}^{{\ast}})\) as \(\mathbb{C}\)-vector spaces, the natural embeddings \(\varLambda (\mathfrak{n})\hookrightarrow Cl\), \(\varLambda (\mathfrak{n}^{{\ast}})\hookrightarrow Cl\) are homogeneous homomorphism of superalgebras, and

$$\displaystyle{ [x,f] = f(x)\quad x \in \mathfrak{n} \subset \varLambda (\mathfrak{n}),\ f \in \mathfrak{n}^{{\ast}}\subset \varLambda (\mathfrak{n}^{{\ast}}). }$$

(Note that [x, f] = xf + fx since x, f are odd.)

Let \(\{x_{\alpha }\}_{\alpha \in \varDelta _{+}}\) be a basis of \(\mathfrak{n}\), \(\{x_{\alpha }^{{\ast}}\}_{\alpha \in \varDelta _{+}}\) the dual basis of \(\mathfrak{n}^{{\ast}}\), and c α, β γ the structure constants of \(\mathfrak{n}\), that is, [x α , x β ] = α, β γ c αβ γ x γ .

Lemma 3

The following map gives a Lie algebra homomorphism.

$$\displaystyle\begin{array}{rcl} \rho: \mathfrak{n}& \longrightarrow & Cl {}\\ x_{\alpha }& \longmapsto & \sum _{\beta,\gamma \in \varDelta _{+}}c_{\alpha,\beta }^{\gamma }x_{\gamma }x_{\beta }^{{\ast}}. {}\\ \end{array}$$

We have

$$\displaystyle{ [\rho (x),y] = [x,y] \in \mathfrak{n} \subset Cl\quad \quad \mathit{\text{for }}x,y \in \mathfrak{n}. }$$

Define an increasing filtration on Cl by setting \(Cl_{p}:=\varLambda ^{\leqslant p}(\mathfrak{n}) \otimes \varLambda (\mathfrak{n}^{{\ast}})\). We have

$$\displaystyle{ 0 = Cl_{-1} \subset Cl_{0} \subset Cl_{1}\cdots \subset Cl_{N} = Cl, }$$

where \(N = \mathrm{dim}\mathfrak{n} = \frac{n(n-1)} {2}\), and

$$\displaystyle{ Cl_{p} \cdot Cl_{q} \subset Cl_{p+q},\quad [Cl_{p},Cl_{q}] \subset Cl_{p+q-1}. }$$
(8)

Let \(\overline{Cl}\) be its associated graded algebra:

$$\displaystyle{ \overline{Cl}:=\mathop{ \mathrm{gr}}\nolimits Cl =\bigoplus _{p\geqslant 0} \dfrac{Cl_{p}} {Cl_{p-1}}. }$$

By (8), \(\overline{Cl}\) is naturally a graded Poisson superalgebra, called the classical Clifford algebra.

We have \(\overline{Cl} =\varLambda (\mathfrak{n}) \otimes \varLambda (\mathfrak{n}^{{\ast}})\) as a commutative superalgebra. Its Poisson (super)bracket is given by

$$\displaystyle\begin{array}{rcl} \{x,f\}& =& f(x),\quad x \in \mathfrak{n} \subset \varLambda (\mathfrak{n}),\ f \in \mathfrak{n}^{{\ast}}\subset \varLambda (\mathfrak{n}^{{\ast}}), {}\\ \{x,y\}& =& 0,\quad x,y \in \mathfrak{n} \subset \varLambda (\mathfrak{n}),\quad \{f,g\} = 0,\ f,g \in \mathfrak{n}^{{\ast}}\subset \varLambda (\mathfrak{n}^{{\ast}}). {}\\ \end{array}$$

Lemma 4

We have \(\overline{Cl}^{\mathfrak{n}} =\varLambda (\mathfrak{n})\) , where \(\overline{Cl}^{\mathfrak{n}}:=\{ w \in \overline{Cl}\mid \{x,w\} = 0,\forall x \in \mathfrak{n}\}\).

The Lie algebra homomorphism \(\rho: \mathfrak{n}\longrightarrow Cl_{1} \subset Cl\) induces a Lie algebra homomorphism

$$\displaystyle{ \bar{\rho }:=\sigma _{1}\circ \rho: \mathfrak{n}\longrightarrow \overline{Cl}, }$$
(9)

where σ 1 is the projection \(Cl_{1} \rightarrow Cl_{1}/Cl_{0} \subset \mathop{\mathrm{gr}}\nolimits Cl\). We have

$$\displaystyle{ \{\bar{\rho }(x),y\} = [x,y]\quad \text{for }x,y \in \mathfrak{n}. }$$

Set

$$\displaystyle{ \bar{C}(\mathfrak{g}) = \mathbb{C}[\mathfrak{g}^{{\ast}}] \otimes \overline{Cl}. }$$

Since it is a tensor product of Poisson superalgebras, \(\bar{C}(\mathfrak{g})\) is naturally a Poisson superalgebra.

Lemma 5

The following map gives a Lie algebra homomorphism:

$$\displaystyle\begin{array}{rcl} \bar{\theta }_{\chi }: \mathfrak{n}& \longrightarrow & \bar{C}(\mathfrak{g}) {}\\ x& \longmapsto & (\mu ^{{\ast}}(x) -\chi (x)) \otimes 1 + 1 \otimes \bar{\rho } (x), {}\\ \end{array}$$

that is, \(\{\bar{\theta }_{\chi }(x),\bar{\theta }_{\chi }(\,y)\} =\bar{\theta } _{\chi }([x,y])\) for \(x,y \in \mathfrak{n}\).

Let \(\bar{C}(\mathfrak{g}) =\bigoplus _{n\in \mathbb{Z}}\bar{C}^{n}(\mathfrak{g})\) be the \(\mathbb{Z}\)-grading defined by degϕ ⊗ 1 = 0 (\(\phi \in \mathbb{C}[\mathfrak{g}^{{\ast}}]\)), deg1 ⊗ f = 1 (\(f \in \mathfrak{n}^{{\ast}}\)), deg1 ⊗ x = −1 (\(x \in \mathfrak{n}\)). We have

$$\displaystyle{ \bar{C}^{n}(\mathfrak{g}) = \mathbb{C}[\mathfrak{g}^{{\ast}}] \otimes (\bigoplus _{ j-i=n}\varLambda ^{i}(\mathfrak{n}) \otimes \varLambda ^{j}(\mathfrak{n}^{{\ast}})). }$$

Lemma 6 ([20, Lemma 7.13.3])

There exists a unique element \(\bar{Q} \in \bar{ C}^{1}(\mathfrak{g})\) such that

$$\displaystyle{ \{\bar{Q},1 \otimes x\} =\bar{\theta } _{\chi }(x)\quad \mathit{\text{for }}x \in \mathfrak{n}. }$$

We have \(\{\bar{Q},\bar{Q}\} = 0\).

Proof

Existence. It is straightforward to see that the element

$$\displaystyle{ \bar{Q} =\sum _{\alpha }(x_{\alpha } -\chi (x_{\alpha })) \otimes x_{\alpha }^{{\ast}}- 1 \otimes \dfrac{1} {2}\sum _{\alpha,\beta,\gamma }c_{\alpha,\beta }^{\gamma }x_{\alpha }^{{\ast}}x_{\beta }^{{\ast}}x_{\gamma } }$$

satisfies the condition.

Uniqueness. Suppose that \(\bar{Q}_{1},\bar{Q}_{2} \in \bar{ C}^{1}(\mathfrak{g})\) satisfy the condition. Set \(R = Q_{1} - Q_{2} \in \bar{ C}^{1}(\mathfrak{g})\). Then {R, 1 ⊗ x} = 0, and so, \(R \in \mathbb{C}[\mathfrak{g}^{{\ast}}] \otimes \overline{Cl}^{\mathfrak{n}}\). But by Lemma 4, \(\overline{Cl}^{\mathfrak{n}} \cap \overline{Cl}^{1} = 0\). Thus R = 0 as required.

To show that \(\{\bar{Q},\bar{Q}\} = 0\), observe that

$$\displaystyle{ \{1 \otimes x,\{1 \otimes y,\{\bar{Q},\bar{Q}\}\}\} = 0,\quad \forall x,y \in \mathfrak{n} }$$

(note that \(\bar{Q}\) is odd). Applying Lemma 4 twice, we get that \(\{\bar{Q},\bar{Q}\} = 0\).

Since \(\bar{Q}\) is odd, Lemma 6 implies that

$$\displaystyle{ \{\bar{Q},\{\bar{Q},a\}\} = \frac{1} {2}\{\{\bar{Q},\bar{Q}\},a\} = 0 }$$

for any \(a \in \bar{ C}(\mathfrak{g})\). That is, \(\mathrm{ad}\bar{Q}:=\{\bar{ Q},\cdot \}\) satisfies that

$$\displaystyle{ (\mathrm{ad}\bar{Q})^{2} = 0. }$$

Thus, \((\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q})\) is a differential graded Poisson superalgebra. Its cohomology \(H^{\bullet }(\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q}) =\bigoplus \limits _{i\in \mathbb{Z}}H^{i}(\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q})\) inherits a graded Poisson superalgebra structure from \(\bar{C}(\mathfrak{g})\).

According to Kostant and Sternberg [66] the Poisson structure of may be described through the following isomorphism:

Theorem 3 ([66])

We have \(H^{i}(\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q}) = 0\) for i ≠ 0 and

as Poisson algebras.

Proof

Give a bigrading on \(\bar{C}:=\bar{ C}(\mathfrak{g})\) by setting

$$\displaystyle{ \bar{C}^{i,j} = \mathbb{C}[\mathfrak{g}^{{\ast}}] \otimes \varLambda ^{i}(\mathfrak{n}^{{\ast}}) \otimes \varLambda ^{-j}(\mathfrak{n}), }$$

so that \(\bar{C} =\bigoplus \limits _{i\geqslant 0,j\leqslant 0}\bar{C}^{i,j}\).

Observe that \(\mathrm{ad}\bar{Q}\) decomposes as \(\mathrm{ad}\bar{Q} = d_{+} + d_{-}\) such that

$$\displaystyle{ d_{-}(\bar{C}^{i,j}) \subset \bar{ C}^{i,j+1},\quad d_{ +}(\bar{C}^{i,j}) \subset \bar{ C}^{i+1,j}. }$$
(10)

Explicitly, we have

$$\displaystyle\begin{array}{rcl} d_{-}& =& \sum _{i}(x_{i} -\chi (x_{i})) \otimes \mathrm{ ad}x_{i}^{{\ast}}, {}\\ d_{+}& =& \sum _{i}\mathrm{ad}x_{i} \otimes x_{i}^{{\ast}}- 1 \otimes \dfrac{1} {2}\sum _{i,j,k}c_{i,j}^{k}x_{ i}^{{\ast}}x_{ j}^{{\ast}}\mathrm{ad}x_{ k} +\sum _{i}1 \otimes \bar{\rho } (x_{i})\mathrm{ad}x_{i}^{{\ast}}. {}\\ \end{array}$$

Since \(\mathrm{ad}\bar{Q}^{2} = 0\), (10) implies that

$$\displaystyle{ d_{-}^{2} = d_{ +}^{2} = [d_{ -},d_{+}] = 0. }$$

It follows that there exists a spectral sequence

$$\displaystyle{ E_{r}\Longrightarrow H^{\bullet }(\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q}) }$$

such that

$$\displaystyle\begin{array}{rcl} E_{1}^{\bullet,q}& =& H^{q}(\bar{C}(\mathfrak{g}),d_{ -}) = H^{q}(\mathbb{C}[\mathfrak{g}^{{\ast}}] \otimes \varLambda (\mathfrak{n}),d_{ -}) \otimes \varLambda ^{\bullet }(\mathfrak{n}^{{\ast}}), {}\\ E_{2}^{p,q}& =& H^{p}(H^{q}(\bar{C}(\mathfrak{g}),d_{ -}),d_{+}). {}\\ \end{array}$$

Observe that \((\bar{C}(\mathfrak{g}),d_{-})\) is identical to the Koszul complex \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\) associated with the sequence x 1χ(x 1), x 2χ(x 2. ), x N χ(x N ) tensorized with \(\varLambda (\mathfrak{n}^{{\ast}})\). Since \(\mathbb{C}[\mu ^{-1}(\chi )] = \mathbb{C}[\mathfrak{g}^{{\ast}}]/\sum _{i}\mathbb{C}[\mathfrak{g}^{{\ast}}](x_{i} -\chi (x_{i}))\), we get that

$$\displaystyle{ H^{i}(\bar{C}(\mathfrak{g}),d_{ -}) = \left \{\begin{array}{@{}l@{\quad }l@{}} \mathbb{C}[\mu ^{-1}(\chi )] \otimes \varLambda (\mathfrak{n}^{{\ast}}),\quad &\text{if}\quad i = 0 \\ 0, \quad &\text{if}\quad i\neq 0. \end{array} \right. }$$

Next, notice that \((H^{0}(C(\mathfrak{g}),d_{-}),d_{+})\) is identical to the Chevalley complex for the Lie algebra cohomology \(H^{\bullet }(\mathfrak{n}, \mathbb{C}[\mu ^{-1}(\chi )])\). Therefore Theorem 2 gives that

Hence the spectral sequence collapses at E 2 = E and we get that \(H^{i}(\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q}) = 0\) for i ≠ 0. Moreover, there is an isomorphism

This completes the proof.

Theorem 4

The natural map \(\mathbb{C}[\mathfrak{g}^{{\ast}}]^{G}\longrightarrow H^{0}(\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q})\) defined by sending p to p ⊗ 1 is an isomorphism of Poisson algebras.

Proof

It is clear that the map is a well-defined homomorphism of Poisson algebras since \(\mathbb{C}[\mathfrak{g}^{{\ast}}]^{G}\) is the Poisson center of \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\). The assertion follows from the commutativity of the following diagram.

2.5 Quantized Hamiltonian Reduction

We shall now quantize the above construction following [66].

Let \(\{U_{i}(\mathfrak{g})\}\) be the PBW filtration of the universal enveloping algebra \(U(\mathfrak{g})\) of \(\mathfrak{g}\), that is, \(U_{i}(\mathfrak{g})\) is the subspace of \(U(\mathfrak{g})\) spanned by the products of at most i elements of \(\mathfrak{g}\). Then

$$\displaystyle\begin{array}{rcl} 0& =& U_{-1}(\mathfrak{g}) \subset U_{0}(\mathfrak{g}) \subset U_{1}(\mathfrak{g}) \subset \ldots,\quad U(\mathfrak{g}) =\bigcup _{i}U_{i}(\mathfrak{g}), {}\\ & & U_{i}(\mathfrak{g}) \cdot U_{j}(\mathfrak{g}) \subset U_{i+j}(\mathfrak{g}),\quad [U_{i}(\mathfrak{g}),U_{j}(\mathfrak{g})] \subset U_{i+j-1}(\mathfrak{g}). {}\\ \end{array}$$

The associated graded space \(\mathop{\mathrm{gr}}\nolimits U(\mathfrak{g}) =\bigoplus _{i\geqslant 0}U_{i}(\mathfrak{g})/U_{i-1}(\mathfrak{g})\) is naturally a Poisson algebra, and the PBW Theorem states that

$$\displaystyle{ \mathop{\mathrm{gr}}\nolimits U(\mathfrak{g})\cong \mathbb{C}[\mathfrak{g}^{{\ast}}] }$$

as Poisson algebras. Thus, \(U(\mathfrak{g})\) is a quantization of \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\).

Define

$$\displaystyle{ C(\mathfrak{g}) = U(\mathfrak{g}) \otimes Cl. }$$

It is naturally a \(\mathbb{C}\)-superalgebra, where \(U(\mathfrak{g})\) is considered as a purely even subsuperalgebra. The filtration of \(U(\mathfrak{g})\) and Cl induces the filtration of \(C(\mathfrak{g})\): \(C_{p}(\mathfrak{g}) =\sum _{i+j\leqslant p}U_{i}(\mathfrak{g}) \otimes Cl_{j}\), and we have

$$\displaystyle{ \mathop{\mathrm{gr}}\nolimits C(\mathfrak{g})\cong \bar{C}(\mathfrak{g}) }$$

as Poisson superalgebras. Therefore, \(C(\mathfrak{g})\) is a quantization of \(\bar{C}(\mathfrak{g})\).

Define the \(\mathbb{Z}\)-grading \(C(\mathfrak{g}) =\bigoplus \limits _{n\in \mathbb{Z}}C^{n}(\mathfrak{g})\) by setting deg(u ⊗ 1) = 0 (\(u \in U(\mathfrak{g})\)), deg(1 ⊗ f) = 1 (\(f \in \mathfrak{n}^{{\ast}}\)), deg(1 ⊗ x) = −1 (\(x \in \mathfrak{n}\)). Then

$$\displaystyle{ C^{n}(\mathfrak{g}) = U(\mathfrak{g}) \otimes (\bigoplus _{ j-i=n}\varLambda ^{i}(\mathfrak{n}) \otimes \varLambda ^{j}(\mathfrak{n}^{{\ast}})). }$$

Lemma 7

The following map defines a Lie algebra homomorphism.

$$\displaystyle\begin{array}{rcl} \theta _{\chi }: \mathfrak{n}& \longrightarrow & C(\mathfrak{g}) {}\\ x& \longmapsto & (x -\chi (x)) \otimes 1 + 1 \otimes \rho (x) {}\\ \end{array}$$

Lemma 8 ([20, Lemma 7.13.7])

There exists a unique element \(Q \in C^{1}(\mathfrak{g})\) such that

$$\displaystyle{ [Q,1 \otimes x] =\theta _{\chi }(x),\quad \forall x \in \mathfrak{n}. }$$

We have Q 2 = 0.

Proof

The proof is similar to that of Lemma 6. In fact the element Q is explicitly given by the same formula as \(\bar{Q}\):

$$\displaystyle{ Q =\sum _{\alpha }(x_{\alpha } -\chi (x_{\alpha })) \otimes x_{\alpha }^{{\ast}}- 1 \otimes \dfrac{1} {2}\sum _{\alpha,\beta,\gamma }c_{\alpha,\beta }^{\gamma }x_{\alpha }^{{\ast}}x_{\beta }^{{\ast}}x_{\gamma }. }$$

Since Q is odd, Lemma 8 implies that

$$\displaystyle{ (\mathrm{ad}Q)^{2} = 0. }$$

Thus, \((C(\mathfrak{g}),\mathrm{ad}Q)\) is a differential graded algebra, and its cohomology \(H^{\bullet }(C(\mathfrak{g}),\mathrm{ad}Q)\) is a graded superalgebra.

However the operator on \(\mathop{\mathrm{gr}}\nolimits C(\mathfrak{g}) =\bar{ C}(\mathfrak{g})\) induced by adQ does not coincide with \(\mathrm{ad}\bar{Q}\). To remedy this, we introduce the Kazhdan filtration \(K_{\bullet }C(\mathfrak{g})\) of \(C(\mathfrak{g})\) as follows: Defined a \(\mathbb{Z}\)-grading on \(\mathfrak{g}\) by

$$\displaystyle{ \mathfrak{g} =\bigoplus _{j\in \mathbb{Z}}\mathfrak{g}_{j},\quad \mathfrak{g}_{j} =\{ x \in \mathfrak{g}: [h,x] = 2jx\} }$$

where h is defined in (6). Then \(\mathfrak{n} =\bigoplus _{j>0}\mathfrak{g}_{j} \subset \mathfrak{b} =\bigoplus _{j\geqslant 0}\mathfrak{g}_{j}\), and

$$\displaystyle{ \mathfrak{h}:= \mathfrak{g}_{0} }$$

is the Cartan subalgebra of \(\mathfrak{g}\) consisting of diagonal matrices. Extend the basis \(\{x_{\alpha }\}_{\alpha \in \varDelta _{+}}\) of \(\mathfrak{n}\) to the basis \(\{x_{a}\}_{a\in \varDelta _{+}\sqcup I}\) of \(\mathfrak{b}\) by adding a basis {x i } iI of \(\mathfrak{h}\). Let c a, b d denote the structure constant of \(\mathfrak{b}\) with respect to this basis.

Lemma 9

The map \(\rho: \mathfrak{n} \rightarrow Cl\) extends to the Lie algebra homomorphism

$$\displaystyle{ \rho: \mathfrak{b} \rightarrow Cl,\quad x_{a}\mapsto \sum _{\beta,\gamma \in \varDelta _{+}}c_{a,\beta }^{\gamma }x_{\gamma }x_{\beta }^{{\ast}}. }$$

Define the Lie algebra homomorphism

$$\displaystyle{ \theta _{0}: \mathfrak{b} \rightarrow C(\mathfrak{g}),\quad x_{i}\mapsto x_{i} \otimes 1 + 1 \otimes \rho (x_{i}), }$$

and define a \(\mathbb{Z}\)-grading on \(C(\mathfrak{g})\) by

$$\displaystyle{ C(\mathfrak{g}) =\bigoplus _{j\in \mathbb{Z}}C(\mathfrak{g})[j],\quad C(\mathfrak{g})[j] =\{ c \in C(\mathfrak{g})\mid [\theta _{0}(h),c] = 2jx\}. }$$

Set

$$\displaystyle{ K_{p}C(\mathfrak{g}) =\sum _{i-j\leqslant p}C_{i}(\mathfrak{g})[j],\quad \text{where }C_{i}(\mathfrak{g})[j] = C_{i}(\mathfrak{g}) \cap C(\mathfrak{g})[j]. }$$

Then \(K_{\bullet }C(\mathfrak{g})\) defines an increasing, exhaustive, separated filtration of \(C(\mathfrak{g})\) such that \(K_{p}C(\mathfrak{g}) \cdot K_{q}C(\mathfrak{g}) \subset K_{p+q}C(\mathfrak{g})\), \([K_{p}C(\mathfrak{g}),K_{q}C(\mathfrak{g})] \subset K_{p+q-1}C(\mathfrak{g})\), and \(\mathop{\mathrm{gr}}\nolimits _{K}C(\mathfrak{g}) =\bigoplus _{p}K_{p}C(\mathfrak{g})/K_{p-1}C(\mathfrak{g})\) is isomorphic to \(\bar{C}(\mathfrak{g})\) as Poisson superalgebras. Moreover, the complex \((\mathop{\mathrm{gr}}\nolimits _{K}C(\mathfrak{g}),\mathrm{ad}Q)\) is identical to \((\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q})\).

Let be the center of \(U(\mathfrak{g})\).

Theorem 5 ([65])

We have \(H^{i}(C(\mathfrak{g}),\mathrm{ad}Q) = 0\) for i ≠ 0 and the map defined by sending z to [z ⊗ 1] is an isomorphism of algebras. Here [z ⊗ 1] denotes the cohomology class of z ⊗ 1.

Proof

We have the spectral sequence

$$\displaystyle{ E_{r}\Longrightarrow H^{\bullet }(C(\mathfrak{g}),\mathrm{ad}Q) }$$

such that

$$\displaystyle{ E_{1}^{\bullet,i} = H^{i}(\mathop{\mathrm{gr}}\nolimits _{ K}C(\mathfrak{g}),\mathrm{ad}\bar{Q})\cong \left \{\begin{array}{@{}l@{\quad }l@{}} \mathbb{C}[\mathfrak{g}^{{\ast}}]^{G},\quad &\text{if}\quad i = 0 \\ 0, \quad &\text{if}\quad i\neq 0. \end{array} \right. }$$

Therefore the spectral sequence collapses at E 1 = E , so we get

$$\displaystyle{ \mathop{\mathrm{gr}}\nolimits H^{0}(C(\mathfrak{g}),\mathrm{ad}Q)\cong \mathbb{C}[\mathfrak{g}^{{\ast}}]^{G}. }$$

Since the homomorphism , z ↦ [z ⊗ 1], respects the filtration and \(K_{\bullet }H^{0}(C(\mathfrak{g}),\mathrm{ad}Q)\), where , \(K_{p}H^{\bullet }(C(\mathfrak{g}),\mathrm{ad}Q) =\mathop{ \mathrm{Im}}\nolimits (H^{0}(K_{p}C(\mathfrak{g}),\mathrm{ad}Q) \rightarrow H^{0}(C(\mathfrak{g}),\mathrm{ad}Q))\), we get the desired isomorphism.

Remark 1 (See [5, § 2] for the Details)

As in the case of \(\bar{C}(\mathfrak{g})\), \(C(\mathfrak{g})\) is also bigraded, we can also write adQ = d + + d such that d +(C i, j) ⊂ C i+1, j, d (C i, j) ⊂ C i, j+1 and get a spectral sequence

$$\displaystyle{ E_{r}\Longrightarrow H^{\bullet }(C(\mathfrak{g}),\mathrm{ad}Q) }$$

such that

$$\displaystyle\begin{array}{rcl} & & \ E_{2}^{p,q} = H^{p}(H^{q}(C(\mathfrak{g}),d_{ -}),d_{+})\cong \delta _{q,0}H^{p}(\mathfrak{n},U(\mathfrak{g}) \otimes _{ U(\mathfrak{n})}\mathbb{C}_{\chi }) {}\\ & \cong & \delta _{p,0}\delta _{q,0}H^{0}(\mathfrak{n},U(\mathfrak{g}) \otimes _{ U(\mathfrak{n})}\mathbb{C}_{\chi })\cong \mathop{\mathrm{End}}\nolimits _{U(\mathfrak{g})}(U(\mathfrak{g}) \otimes _{U(\mathfrak{n})}\mathbb{C}_{\chi })^{op}. {}\\ \end{array}$$

Where \(\mathbb{C}_{\chi }\) is the one-dimensional representation of \(\mathfrak{n}\) defined by the character χ. Thus we get the Whittaker model isomorphism [65]

2.6 Classical Miura Map

Let \(\mathfrak{n}_{-} =\bigoplus \limits _{j<0}\mathfrak{g}_{j}\) be the subalgebra of \(\mathfrak{g}\) consisting of lower triangular matrices, and set \(\mathfrak{b}_{-} =\bigoplus \limits _{j\leqslant 0}\mathfrak{g}_{j} = \mathfrak{n}_{-}\mathop{\oplus }\mathfrak{h}\). We have

$$\displaystyle{ \mathfrak{g} = \mathfrak{b}_{-}\mathop{\oplus }\mathfrak{n}_{+}. }$$
(11)

Extend the basis \(\{x_{a}\}_{a\in \varDelta _{+}\sqcup I}\) to the basis \(\{x_{a}\}_{a\in \varDelta _{+}\sqcup I\sqcup \varDelta _{-}}\) by adding a basis \(\{x_{\alpha }\}_{\alpha \in \varDelta _{-}}\) of \(\mathfrak{n}_{-}\). Let c a, b d be the structure constant with respect to this basis. Extend \(\theta _{0}: \mathfrak{b} \rightarrow C(\mathfrak{g})\) to the linear map \(\theta _{0}: \mathfrak{g} \rightarrow C(\mathfrak{g})\) by setting

$$\displaystyle{ \theta _{0}(x_{a}) = x_{a} \otimes 1 + 1 \otimes \sum _{\beta,\gamma \in \varDelta _{+}}c_{a,\beta }^{\gamma }x_{\gamma }x_{\beta }^{{\ast}}. }$$

We already know that the restriction of θ 0 to \(\mathfrak{n}\) is a Lie algebra homomorphism and

$$\displaystyle{ [\theta _{0}(x),1 \otimes y] = 1 \otimes [x,y]\quad \text{for }x,y \in \mathfrak{n}. }$$

Although θ 0 is not a Lie algebra homomorphism, we have the following.

Lemma 10

The restriction of θ 0 to \(\mathfrak{b}_{-}\) is a Lie algebra homomorphism. We have [θ 0(x), 1 ⊗ y] = 1 ⊗ ad(x)( y) for \(x \in \mathfrak{b}_{-}\) , \(y \in \mathfrak{n}^{{\ast}}\) , where ad denote the coadjoint action and \(\mathfrak{n}^{{\ast}}\) is identified with \((\mathfrak{g}/\mathfrak{b}_{-})^{{\ast}}\).

Let \(C(\mathfrak{g})_{+}\) denote the subalgebra of \(C(\mathfrak{g})\) generated by \(\theta _{0}(\mathfrak{n})\) and \(\varLambda (\mathfrak{n}) \subset Cl\), and let \(C(\mathfrak{g})_{-}\) denote the subalgebra generated by \(\theta _{0}(\mathfrak{b}_{-})\) and \(\varLambda (\mathfrak{n}^{{\ast}}) \subset Cl\).

Lemma 11

The multiplication map gives a linear isomorphism

$$\displaystyle{ C(\mathfrak{g})_{-}\otimes C(\mathfrak{g})_{+}\;\stackrel{_{\sim }}{\rightarrow }\;C(\mathfrak{g}). }$$

Lemma 12

The subspaces \(C(\mathfrak{g})_{-}\) and \(C(\mathfrak{g})_{+}\) are subcomplexes of \((C(\mathfrak{g}),\mathrm{ad}Q)\) . Hence \(C(\mathfrak{g})\cong C(\mathfrak{g})_{-}\otimes C(\mathfrak{g})_{+}\) as complexes.

Proof

The fact that \(C(\mathfrak{g})_{+}\) is subcomplex is obvious (see Lemma 8). The fact that \(C(\mathfrak{g})_{-}\) is a subcomplex follows from the following formula.

$$\displaystyle\begin{array}{rcl} [Q,\theta _{0}(x_{a})]& =& \sum _{b\in \varDelta _{-}\sqcup I,\alpha \in \varDelta _{+}}c_{\alpha,a}^{b}\theta _{ 0}(x_{b})(1 \otimes x_{\alpha }^{{\ast}}) - 1 \otimes \sum _{\beta,\gamma \in \varDelta _{+}}c_{a,\beta }^{\gamma }\chi (x_{\gamma })x_{\beta }^{{\ast}} {}\\ {}[Q,1 \otimes x_{\alpha }^{{\ast}}]& =& -1 \otimes \frac{1} {2}\sum _{\beta,\gamma \in \varDelta _{+}}c_{\beta,\gamma }^{\alpha }x_{\beta }^{{\ast}}x_{\gamma }^{{\ast}} {}\\ \end{array}$$

(aΔ I, αΔ +).

Proposition 3

\(H^{\bullet }(C(\mathfrak{g})_{-},\mathrm{ad}Q)\cong H^{\bullet }(C(\mathfrak{g}),\mathrm{ad}Q)\).

Proof

By Lemma 12 and Kunneth’s Theorem,

$$\displaystyle{ H^{p}(C(\mathfrak{g}),\mathrm{ad}Q)\cong \bigoplus _{ i+j=p}H^{i}(C(\mathfrak{g})_{ -},\mathrm{ad}Q) \otimes H^{j}(C(\mathfrak{g})_{ +},\mathrm{ad}Q). }$$

On the other hand we have ad(Q)(1 ⊗ x α ) = θ χ (x α ) = θ 0(x α ) −χ(x α ) for αΔ . Hence \(C(\mathfrak{g})_{-}\) is isomorphic to the tensor product of complexes of the form \(\mathbb{C}[\theta _{\chi }(x_{\alpha })] \otimes \varLambda (x_{\alpha })\) with the differential θ χ (x α ) ⊗ x α , where x α denotes the odd derivation of the exterior algebra Λ(x α ) with one variable x α such that x α (x α ) = 1. Each of these complexes has one-dimensional zeroth cohomology and zero first cohomology. Therefore \(H^{i}(C(\mathfrak{g})_{+},\mathrm{ad}Q) =\delta _{i,0}\mathbb{C}\). This completes the proof.

Note that the cohomological gradation takes only non-negative values on \(C(\mathfrak{g})_{-}\). Hence by Proposition 3 we may identify with the subalgebra \(H^{0}(C(\mathfrak{g})_{-},\mathrm{ad}Q) =\{ c \in C(\mathfrak{g})_{-}^{0}\mid \mathrm{ad}Q(c) = 0\}\) of \(C(\mathfrak{g})_{-}\).

Consider the decomposition

$$\displaystyle{ C(\mathfrak{g})_{-}^{0} =\bigoplus _{ j\leqslant 0}C(\mathfrak{g})_{-,j}^{0},\quad C(\mathfrak{g})_{ -,j}^{0} =\{ c \in C(\mathfrak{g})_{ -}^{0}\mid [\theta _{ 0}(h),c] = 2jc\}. }$$

Note that \(C(\mathfrak{g})_{-,0}^{0}\) is generated by \(\theta _{0}(\mathfrak{h})\) and is isomorphic to \(U(\mathfrak{h})\). The projection

$$\displaystyle{ C(\mathfrak{g})_{-}^{0} \rightarrow C(\mathfrak{g})_{ -,0}^{0}\cong U(\mathfrak{h}) }$$

is an algebra homomorphism, and hence, its restriction

is also an algebra homomorphism.

Proposition 4

The map Υ is an embedding.

Let \(K_{\bullet }C(\mathfrak{g})_{\pm }\) be the filtration of \(C(\mathfrak{g})_{\pm }\) induced by the Kazhdan filtration of \(C(\mathfrak{g})\). We have the isomorphism

$$\displaystyle{ \bar{C}(\mathfrak{g}) =\mathop{ \mathrm{gr}}\nolimits _{K}C(\mathfrak{g})\cong \mathop{\mathrm{gr}}\nolimits _{K}C(\mathfrak{g})_{-}\otimes \mathop{\mathrm{gr}}\nolimits _{K}C(\mathfrak{g})_{+} }$$

as complexes. Similarly as above, we have \(H^{i}(\mathop{\mathrm{gr}}\nolimits _{K}C(\mathfrak{g})_{+},\mathrm{ad}\bar{Q}) =\delta _{i,0}\mathbb{C}\), and

$$\displaystyle{ H^{0}(\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q})\cong H^{0}(\mathop{\mathrm{gr}}\nolimits _{ K}C(\mathfrak{g})_{-},\mathrm{ad}\bar{Q}). }$$
(12)

Proof (Proof of Proposition 4 )

The filtration \(K_{\bullet }U(\mathfrak{h})\) of \(U(\mathfrak{h})\cong C(\mathfrak{g})_{-,0}^{0}\) induced by the Kazhdan filtration coincides with the usual PBW filtration. By (12) and Theorem 3, the induced map

$$\displaystyle{ H^{0}(\mathop{\mathrm{gr}}\nolimits _{ K}C(\mathfrak{g})_{-},\mathrm{ad}Q) \rightarrow \mathop{\mathrm{gr}}\nolimits _{K}U(\mathfrak{h}) }$$

can be identified with the restriction map

(13)

It is sufficient to show that \(\bar{\varUpsilon }\) is injective.

If \(\varphi \in \mathbb{C}[\,f + \mathfrak{b}]^{N}\) is in the kernel, φ(g. x) = 0 for all gN and \(x \in f + \mathfrak{h}\). Hence it is enough to show that the image of the the action map

$$\displaystyle{ N \times (\,f + \mathfrak{h}) \rightarrow f + \mathfrak{b},\quad (g,x)\mapsto \mathrm{Ad}(g)x, }$$
(14)

is Zariski dense in \(f + \mathfrak{b}\).

The differential of this morphism at \((1,x) \in N \times (\,f + \mathfrak{h})\) is given by

$$\displaystyle{ \mathfrak{n} \times \mathfrak{h} \rightarrow \mathfrak{b},\quad (\,y,z)\mapsto [\,y,x] + z. }$$

This is an isomorphism if \(x \in f + \mathfrak{h}_{\mathop{\mathrm{reg}}\nolimits }\), where \(\mathfrak{h}_{\mathop{\mathrm{reg}}\nolimits } =\{ x \in \mathfrak{h}\mid \mathfrak{n}^{x} = 0\}\). Hence (14) is a dominant morphism as required, see e.g. [76, Theorem 16.5.7].

Remark 2

The fact that \(\bar{\varUpsilon }\) is injective is in fact well-known. Indeed, under the identifications , \(\mathbb{C}[\,f + \mathfrak{h}]\cong \mathbb{C}[\mathfrak{h}]\), \(\bar{\varUpsilon }\) is identified with the Chevalley restriction map \(\mathbb{C}[\mathfrak{g}]^{G}\;\stackrel{_{\sim }}{\rightarrow }\;\mathbb{C}[\mathfrak{h}]^{W}\), where \(W = \mathfrak{S}_{n}\).

The advantage of the above proof is that it applies to a general finite W-algebra [71], and also, it generalizes to the affine setting, see Sect. 5.9.

The map Υ is called the classical Miura map.

2.7 Generalization to an Arbitrary Simple Lie Algebra

It is clear that the above argument works if we replace \(\mathfrak{g}\mathfrak{l}_{n}\) by \(\mathfrak{s}\mathfrak{l}_{n}\), and \(\mathfrak{a}\) by \(\mathfrak{a} \cap \mathfrak{s}\mathfrak{l}_{n}\).

More generally, let \(\mathfrak{g}\) be an arbitrary simple Lie algebra. Let f be a principal (regular) nilpotent element of \(\mathfrak{g}\), {e, f, h} an associated \(\mathfrak{s}\mathfrak{l}_{2}\)-triple. One may assume that

$$\displaystyle{ f =\sum _{i\in I}f_{i}, }$$

where f i is a root vector of roots α i and {α i } iI is the set of simple roots of \(\mathfrak{g}\). Define the Kostant slice by

where \(\mathfrak{g}^{e}\) is the centralizer of e in \(\mathfrak{g}\).

Then all the statements in previous subsections that make sense hold by replacing the set of companion matrices by the Kostant slice [65].

2.8 Generalization to Finite W-Algebras

In fact, the above argument works in more general setting of Hamiltonian reduction. In particular for Slodowy slices. Namely, for a non-zero nilpotent element f of a finite-dimensional semisimple Lie algebra \(\mathfrak{g}\), we can use Jacobson-Morozov’s theorem to embed f into an sl 2-triple {e, f, h}. The Slodowy slice at f is defined to be the affine subslace

of \(\mathfrak{g}\).

The Slodowy slice has the following properties.

  • intersects the G-orbits at any point of , where G is the adjoint group of \(\mathfrak{g}\).

  • admits a \(\mathbb{C}^{{\ast}}\)-action which is contracting to f.

As in the case of the set of companion matrices can be realized by Hamiltonian reduction. Let \(\mathfrak{g}_{j} =\{ x \in \mathfrak{g}\mid [h,x] = 2jx\}\), so that

$$\displaystyle{ \mathfrak{g} =\bigoplus _{j\in \frac{1} {2} \mathbb{Z}}\mathfrak{g}_{j}. }$$

Then the subspace \(\mathfrak{g}_{1/2}\) admits a symplectic form defined by 〈x | y〉 = ( f | [x, y]). Choose a Lagrangian subspace l of \(\mathfrak{g}_{1/2}\) with respect to this form, and set \(\mathfrak{m} = l +\sum _{j\geqslant 1}\mathfrak{g}_{j}\). Then \(\mathfrak{m}\) is a nilpotent subalgebra of \(\mathfrak{g}\) and \(\chi: \mathfrak{m} \rightarrow \mathbb{C}\), x ↦ ( f | x), defines a character. Let M be the unipotent subgroup of G corresponding to \(\mathfrak{m}\), that is, \(\mathop{\mathrm{Lie}}\nolimits M = \mathfrak{m}\). The adjoint action of M on \(\mathfrak{g}\) is Hamiltonian, so we can consider the moment map of this action

$$\displaystyle{ \mu: \mathfrak{g}^{{\ast}}\longrightarrow \mathfrak{m}^{{\ast}}, }$$

which is just a restriction map. Then we have the following realization of the Slodowy slice.

To obtain the BRST realization of this Hamiltonian reduction we simply replace the Clifford algebra Cl by \(Cl_{\mathfrak{m}}\), i.e., the Clifford algebra associated to \(\mathfrak{m} \oplus \mathfrak{m}^{{\ast}}\). Then we can define the operator \(\mathrm{ad}\bar{Q}\) similarly and get a differential cochain complex \((\mathbb{C}[\mathfrak{g}^{{\ast}}] \otimes \overline{Cl}_{\mathfrak{m}},\mathrm{ad}\bar{Q})\). We have

as Poisson algebras.

As above, this construction has a natural quantization and the quantization \(U(\mathfrak{g},f)\) of thus defined is called the finite W-algebra associated to the pair \((\mathfrak{g},f)\) [75]:

$$\displaystyle{ U(\mathfrak{g},f):= H^{0}(U(\mathfrak{g}) \otimes Cl_{\mathfrak{m}},\mathrm{ad}\bar{Q}_{ +})\cong \mathop{\mathrm{End}}\nolimits _{U(\mathfrak{g})}(U(\mathfrak{g}) \otimes _{U(\mathfrak{m})}\mathbb{C}_{\chi })^{op}, }$$

where \(\mathbb{C}_{\chi }\) is the one-dimensional representation of \(\mathfrak{m}\) defined by χ (cf. [5, 29]).

3 Arc Spaces, Poisson Vertex Algebras, and Associated Varieties of Vertex Algebras

3.1 Vertex Algebras

A vertex algebra is a vector space V equipped with | 0〉 ∈ V (the vacuum vector), \(T \in \mathop{\mathrm{End}}\nolimits V\) (the translation operator), and a bilinear product

$$\displaystyle{ V \times V \rightarrow V ((z)),\quad (a,b)\mapsto a(z)b, }$$

where \(a(z) =\sum _{n\in \mathbb{Z}}a_{(n)}z^{-n-1}\), \(a_{(n)} \in \mathop{\mathrm{End}}\nolimits V\), such that

  1. 1.

    ( | 0〉)(z) = id V ,

  2. 2.

    a(z) | 0〉 ∈ V [[z]] and \(\lim \limits _{z\rightarrow 0}a(z)\vert 0\rangle = a\) for all aV,

  3. 3.

    (Ta)(z) = z a(z) for all aV, where z = ddz,

  4. 4.

    for any a, bV, \((z - w)^{N_{a,b}}[a(z),b(w)] = 0\) for some \(N_{a,b} \in \mathbb{Z}_{+} =\{ 0,1,2,\ldots \}\).

The last condition is called the locality, which is equivalent to the fact that

$$\displaystyle{ [a(z),b(w)] =\sum _{ n=0}^{N_{a,b}-1}(a_{ (n)}b)(w) \frac{1} {n!}\partial _{w}^{n}\delta (z - w), }$$
(15)

where \(\delta (z - w) =\sum _{n\in \mathbb{Z}}w^{n}z^{-n-1} \in \mathbb{C}[[z,w,z^{-1},w^{-1}]]\).

A consequence of the definition is the following Borcherds identities:

$$\displaystyle{ [a_{(m)},b_{(n)}] =\sum _{i\geqslant 0}\left (\begin{array}{*{10}c} m\\ i \end{array} \right )(a_{(i)}b)_{(m+n-i)}, }$$
(16)
$$\displaystyle{ (a_{(m)}b)_{(n)} =\sum _{j\geqslant 0}(-1)^{j}\left (\begin{array}{*{10}c} m\\ j \end{array} \right )(a_{(m-j)}b_{(n+j)}-(-1)^{m}b_{ (m+n-j)}a_{(\,j)}). }$$
(17)

We write (15) as

$$\displaystyle{ [a_{\lambda }b] =\sum _{n\geqslant 0} \frac{\lambda ^{n}} {n!}a_{(n)}b \in V [\lambda ], }$$

and call it the λ-bracket of a and b. (We have a (n) b = 0 if (zw)n[a(z), b(w)] = 0.)

Here are some properties of λ-brackets.

$$\displaystyle{ [(Ta)_{\lambda }b] = -\lambda [a_{\lambda }b],\quad [a_{\lambda }(Tb)] = (\lambda +T)[a_{\lambda }b], }$$
(18)
$$\displaystyle{ [b_{\lambda }a] = -[a_{-\lambda -T}b], }$$
(19)
$$\displaystyle{ [a_{\lambda }[b_{\mu }c]] - [b_{\mu }[a_{\lambda }c]] = [[a_{\lambda }b]_{\lambda +\mu }c]. }$$
(20)

The normally ordered product on V is defied as: ab: = a (−1) b. We also write: ab: (z) =: a(z)b(z): . We have

$$\displaystyle{: a(z)b(z):= a(z)_{+}b(z) + b(w)a(z)_{-}, }$$

where a(z)+ = n < 0 a (n) z n−1, \(a(z)_{-} =\sum _{n\geqslant 0}a_{(n)}z^{-n-1}\). We have the following non-commutative Wick formula.

$$\displaystyle{ [a_{\lambda }: bc:] =: [a_{\lambda }b]c: +: [a_{\lambda }c]b: +\int _{0}^{\lambda }[[a_{\lambda }b]_{\mu }c]d\mu, }$$
(21)
$$\displaystyle{ [: ab: _{\lambda }c] =: (e^{T\partial _{\lambda }}a)[b_{\lambda }c]: +: (e^{T\partial _{\lambda }}b)[a_{\lambda }c]: +\int _{ 0}^{\lambda }[b_{\mu }[a_{\lambda -\mu }c]]d\mu. }$$
(22)

3.2 Commutative Vertex Algebras and Differential Algebras

A vertex algebra V is called commutative if

$$\displaystyle{ [a_{\lambda }b] = 0,\quad \forall a,b \in V, }$$

or equivalently, a (n) = 0 for \(n\geqslant 0\) in \(\mathop{\mathrm{End}}\nolimits V\) for all aV. This condition is equivalent to that

$$\displaystyle{ [a_{(m)},b_{(n)}] = 0\quad \forall a,b \in \mathbb{Z},\ m,n \in \mathbb{Z} }$$

by (16).

A commutative vertex algebra has the structure of a unital commutative algebra by the product

$$\displaystyle{ a \cdot b =: ab:= a_{(-1)}b, }$$

where the unite is given by the vacuum vector | 0〉. The translation operator T of V acts on V as a derivation with respect to this product:

$$\displaystyle{ T(a \cdot b) = (Ta) \cdot b + a \cdot (Tb). }$$

Therefore a commutative vertex algebra has the structure of a differential algebra, that is, a unital commutative algebra equipped with a derivation. Conversely, there is a unique vertex algebra structure on a differential algebra R with a derivation T such that

$$\displaystyle{ Y (a,z) = e^{zT}a }$$

for aR. This correspondence gives the following.

Theorem 6 ([24])

The category of commutative vertex algebras is the same as that of differential algebras.

3.3 Arc Spaces

Define the (formal) disc as

$$\displaystyle{ D =\mathop{ \mathrm{Spec}}\nolimits (\mathbb{C}[[t]]). }$$

For a scheme X, a homomorphism α: DX is called an arc of X.

Theorem 7 ([25, 33, 52])

Let X be a scheme of finite type over \(\mathbb{C}\) , Sch the category of schemes of finite type over \(\mathbb{C}\) , Set the category of sets. The contravariant functor

$$\displaystyle{ Sch \rightarrow Set,\quad Y \mapsto \mathop{\mathrm{Hom}}\nolimits _{Sch}(Y \widehat{ \times } D,X), }$$

is represented by a scheme JX, that is,

$$\displaystyle{ \mathop{\mathrm{Hom}}\nolimits _{Sch}(Y,JX)\cong \mathop{\mathrm{Hom}}\nolimits _{Sch}(Y \widehat{ \times } D,X). }$$

for any YSch. Here \(Y \widehat{ \times } D\) is the completion of Y ×D with respect to the subscheme \(Y \widehat{ \times }\{ 0\}\).

By definition, the \(\mathbb{C}\)-points of JX are

$$\displaystyle{ \mathop{\mathrm{Hom}}\nolimits _{Sch}(\mathop{\mathrm{Spec}}\nolimits \mathbb{C},JX) =\mathop{ \mathrm{Hom}}\nolimits _{Sch}(D,X), }$$

that is, the set of arcs of X. The reason we need the completion \(Y \widehat{ \times } D\) in the definition is that \(A \otimes \mathbb{C}[[t]] \subsetneq A[[t]] = A\widehat{ \otimes } \mathbb{C}[[t]]\) in general.

The scheme JX is called the arc space, or the infinite jet scheme, of X.

It is easy to describe JX when X is affine:

First, consider the case \(X = \mathbb{C}^{N} =\mathop{ \mathrm{Spec}}\nolimits \mathbb{C}[x_{1},x_{2},\cdots \,,x_{N}]\). The \(\mathbb{C}\)-points of JX are the arcs \(\mathop{\mathrm{Hom}}\nolimits _{Sch}(D,JX)\), that is, the ring homomorphisms

$$\displaystyle{ \gamma: \mathbb{C}[x_{1},x_{2},\cdots \,,x_{N}] \rightarrow \mathbb{C}[[t]]. }$$

Such a map is determined by the image

$$\displaystyle{ \gamma (x_{i}) =\sum _{n\geqslant 0}\gamma _{i,(-n-1)}t^{n} }$$
(23)

of each x i , and conversely, the coefficients {γ i, (−n−1)} determines a \(\mathbb{C}\)-point of JX. If we choose coordinates x i, (−n−1) of JX as x i, (−n−1)(γ) = γ i, (−n−1), we have

$$\displaystyle{ J\mathbb{C}^{N} =\mathop{ \mathrm{Spec}}\nolimits \mathbb{C}[x_{ i,(n)}\vert i = 1,2,\cdots \,,N,n = -1,-2,\cdots \,]. }$$

Next, let \(X =\mathop{ \mathrm{Spec}}\nolimits R\), with \(R = \mathbb{C}[x_{1},x_{2},\cdots \,,x_{N}]/\langle f_{1},f_{2},\cdots \,,f_{r}\rangle\). The arcs of X are

$$\displaystyle{ \mathop{\mathrm{Hom}}\nolimits _{\mathop{\mathrm{ring}}\nolimits }\big( \frac{\mathbb{C}[x_{1},x_{2},\cdots \,,x_{n}]} {<f_{1},f_{2},\cdots \,,f_{r}>}, \mathbb{C}[[t]]\big) \subset \mathop{\mathrm{Hom}}\nolimits _{\mathop{\mathrm{ring}}\nolimits }(\mathbb{C}[x_{1},x_{2},\cdots \,,x_{n}], \mathbb{C}[[t]]). }$$

An element \(\gamma \in \mathop{\mathrm{Hom}}\nolimits _{\mathop{\mathrm{ring}}\nolimits }(\mathbb{C}[x_{1},x_{2},\cdots \,,x_{n}], \mathbb{C}[[t]])\) is an element of this subset if and only if γ( f i ) = 0 for i = 1, 2, ⋯ , r. By writing

$$\displaystyle{ f_{i}(x_{1}(t),x_{2}(t),\ldots,x_{N}(t)) =\sum _{m\geqslant 0}\frac{f_{i,m}} {m!} t^{m} }$$

with \(f_{i,m} \in \mathbb{C}[x_{i,(-n-1)}]\), where \(x_{i}(t):=\sum _{m\geqslant 0}x_{i,(-m-1)}t^{m}\), we get that

$$\displaystyle{ JX =\mathop{ \mathrm{Spec}}\nolimits \frac{\mathbb{C}[x_{i(n)}\vert i = 1,2,\cdots \,,N;n = -1,-2,\cdots \,]} {\langle f_{i,m}(x_{i(n)}),i = 1,2,\cdots \,,r;m\geqslant 0\rangle }. }$$

Lemma 13

Define the derivation T of \(\mathbb{C}[x_{i(n)}\vert i = 1,2,\cdots \,,N;n = -1,-2,\cdots \,]\) by

$$\displaystyle{ Tx_{i(n)} = -nx_{i(n-1)}. }$$

Then f i, m = T n f i for \(n\geqslant 0\) . Here we identify x i with x i(−1).

With the above lemma, we conclude that for the affine scheme \(X =\mathop{ \mathrm{Spec}}\nolimits R\), \(R = \mathbb{C}[x_{1},x_{2},\cdots \,,x_{n}]/\langle f_{1},f_{2},\cdots \,,f_{r}\rangle\), its arc space JX is the affine scheme \(\mathop{\mathrm{Spec}}\nolimits (JR)\), where

$$\displaystyle{ JR:= \frac{\mathbb{C}[x_{i(n)}\vert i = 1,2,\cdots \,,N;n = -1,-2,\cdots \,]} {\langle T^{n}f_{i},i = 1,2,\cdots \,,r;n\geqslant 0\rangle } }$$

and T is as defined in the lemma.

The derivation T acts on the above quotient ring JR. Hence for an affine scheme \(X =\mathop{ \mathrm{Spec}}\nolimits R\), the coordinate ring \(JR = \mathbb{C}[JX]\) of its arc space JX is a differential algebra, hence is a commutative vertex algebra.

Remark 3

The differential algebra JR has the universal property that

$$\displaystyle{ \mathop{\mathrm{Hom}}\nolimits _{\text{dif.alg.}}(JR,A)\cong \mathop{\mathrm{Hom}}\nolimits _{\mathrm{ring}}(R,A) }$$

for any differential algebra A, where \(\mathop{\mathrm{Hom}}\nolimits _{\text{diff. alg.}}(JR,A)\) is the set of homomorphisms JRA of differential algebras.

For a general scheme Y of finite type with an affine open covering {U i } iI , its arc space JY is obtained by glueing JU i (see [33, 52]). In particular, the structure sheaf is a sheaf of commutative vertex algebras.

There is a natural projection π : JXX that corresponds to the embedding RJR, x i x i, (−1), in the case X is affine. In terms of arcs, π (α) = α(0) for \(\alpha \in \mathop{\mathrm{Hom}}\nolimits _{Sch}(D,X)\), where 0 is the unique closed point of the disc D.

The map from a scheme to its arc space is functorial. i.e., a scheme homomorphism f: XY induces a scheme homomorphism Jf: JXJY that makes the following diagram commutative:

In terms of arcs, Jf(α) = fα for \(\alpha \in \mathop{\mathrm{Hom}}\nolimits _{Sch}(D,X)\).

We also have

$$\displaystyle{ J(X \times Y )\cong JX \times JY. }$$
(24)

Indeed, for any scheme Z,

$$\displaystyle\begin{array}{rcl} \mathop{\mathrm{Hom}}\nolimits (Z,J(X \times Y ))& =& \mathop{\mathrm{Hom}}\nolimits (Z\widehat{ \times } D,X \times Y ) {}\\ & \cong & \mathop{\mathrm{Hom}}\nolimits (Z\widehat{ \times } D,X) \times \mathop{\mathrm{Hom}}\nolimits (Y \widehat{ \times } D,Y ) {}\\ & =& \mathop{\mathrm{Hom}}\nolimits (Z,JX) \times \mathop{\mathrm{Hom}}\nolimits (Y,JY ) {}\\ & \cong & \mathop{\mathrm{Hom}}\nolimits (Z,JX \times JY ). {}\\ \end{array}$$

Lemma 14

The natural morphism \(X_{\mathop{\mathrm{red}}\nolimits } \rightarrow X\) induces an isomorphism \(JX_{\mathop{\mathrm{red}}\nolimits } \rightarrow JX\) of topological spaces, where \(X_{\mathop{\mathrm{red}}\nolimits }\) denotes the reduced scheme of X.

Proof

We may assume that \(X =\mathop{ \mathrm{Spec}}\nolimits R\). An arc α of X corresponds to a ring homomorphism \(\alpha ^{{\ast}}: R \rightarrow \mathbb{C}[[t]]\). Since \(\mathbb{C}[[t]]\) is an integral domain it decomposes as \(\alpha ^{{\ast}}: R \rightarrow R/\sqrt{0} \rightarrow \mathbb{C}[[t]]\). Thus, α is an arc of \(X_{\mathop{\mathrm{red}}\nolimits }\).

If X is a point, then JX is also a point, since \(\mathop{\mathrm{Hom}}\nolimits (D,X) =\mathop{ \mathrm{Hom}}\nolimits (\mathbb{C}, \mathbb{C}[[t]])\) consists of only one element. Thus, Lemma 14 implies the following.

Corollary 2

If X is zero-dimensional then JX is also zero-dimensional.

Theorem 8 ([64])

JX is irreducible if X is irreducible.

Lemma 15

Let Y be irreducible, and let f: XY be a morphism that restricts to a bijection between some open subsets UX and VY. Then Jf: JXJY is dominant.

Proof

Jf restricts to the isomorphism \(JU\;\stackrel{_{\sim }}{\rightarrow }\;JV\), and the open subset JV is dense in JY since JY is irreducible.

3.4 Arc Space of Poisson Varieties and Poisson Vertex Algebras

Let V be a commutative vertex algebra, or equivalently, a differential algebra. V is called a Poisson vertex algebras if it is equipped with a bilinear map

$$\displaystyle{ V \times V \rightarrow V [\lambda ],\quad (a,b)\mapsto \{a_{\lambda }b\} =\sum _{n\geqslant 0} \frac{\lambda ^{n}} {n!}a_{(n)}b,\quad a_{(n)} \in \mathop{\mathrm{End}}\nolimits V, }$$

also called the λ-bracket, satisfying the following axioms:

$$\displaystyle{ \{(Ta)_{\lambda }b\} = -\lambda \{a_{\lambda }b\},\quad \{a_{\lambda }(Tb)\} = (\lambda +T)\{a_{\lambda }b\}, }$$
(25)
$$\displaystyle{ \{b_{\lambda }a\} = -\{a_{-\lambda -T}b\}, }$$
(26)
$$\displaystyle{ \{a_{\lambda }\{b_{\mu }c\}\} -\{ b_{\mu }\{a_{\lambda }c\}\} =\{\{ a_{\lambda }b\}_{\lambda +\mu }c\}, }$$
(27)
$$\displaystyle{ \{a_{\lambda }(bc)\} =\{ a_{\lambda }b\}c +\{ a_{\lambda }c\}b,\quad \{(ab)_{\lambda }c\} =\{ a_{\lambda +T}c\}_{\rightarrow }b +\{ b_{\lambda +T}c\}_{\rightarrow }a, }$$
(28)

where the arrow means that λ + T should be moved to the right, that is, \(\{a_{\lambda +T}c\}_{\rightarrow }b =\sum _{n\geqslant 0}(a_{(n)}c)\frac{(\lambda +T)^{n}} {n!} b\).

The first equation in (28) says that a (n), \(n\geqslant 0\), is a derivation of the ring V. (Do not confuse \(a_{(n)} \in \mathop{\mathrm{Der}}\nolimits (V )\), \(n\geqslant 0\), with the multiplication a (n) as a vertex algebra, which should be zero for a commutative vertex algebra.)

Note that (25), (26), (27) are the same as (18), (19), (20), and (28) is the same with (21) and (22) without the third terms. In particular, by (27), we have

$$\displaystyle{ [a_{(m)},b_{(n)}] =\sum _{i\geqslant 0}\left (\begin{array}{*{10}c} m\\ i \end{array} \right )(a_{(i)}b)_{(m+n-i)},\quad m,n \in \mathbb{Z}_{+}. }$$
(29)

Theorem 9 ([7, Proposition 2.3.1])

Let X be an affine Poisson scheme, that is, \(X =\mathop{ \mathrm{Spec}}\nolimits R\) for some Poisson algebra R. Then there is a unique Poisson vertex algebra structure on \(JR = \mathbb{C}[JX]\) such that

$$\displaystyle{ \{a_{\lambda }b\} =\{ a,b\}\quad \mathit{\text{for }}a,b \in R \subset JR, }$$

where {a, b} is the Poisson bracket in R.

Proof

The uniqueness is clear by (18) since JR is generated by R as a differential algebra. We leave it to the reader to check the well-definedness.

Remark 4

More generally, let X be a Poisson scheme which is not necessarily affine. Then the structure sheaf carries a unique vertex Poisson algebra structure such that {f λ g} = {f, g} for , see [16, Lemma 2.1.3.1].

Example 1

Let G be an affine algebraic group, \(\mathfrak{g} =\mathop{ \mathrm{Lie}}\nolimits G\). The arc space JG is naturally a proalgebraic group. Regarding JG as the \(\mathbb{C}[[t]]\)-points of G, we have JG = G[[t]]. Similarly, \(J\mathfrak{g} = \mathfrak{g}[[t]] =\mathop{ \mathrm{Lie}}\nolimits (JG)\).

The affine space \(\mathfrak{g}^{{\ast}}\) is a Poisson variety by the Kirillov-Kostant Poisson structure, see Sect. 2.3. If {x i }is a basis of \(\mathfrak{g}\), then

$$\displaystyle{ \mathbb{C}[\mathfrak{g}^{{\ast}}] = \mathbb{C}[x_{ 1},x_{2},\cdots \,,x_{n}]. }$$

Thus

$$\displaystyle{ J\mathfrak{g}^{{\ast}} =\mathop{ \mathrm{Spec}}\nolimits \mathbb{C}[x_{ i(-n)}\vert i = 1,2,\cdots \,,l;n\geqslant 1]. }$$
(30)

So we may identify \(\mathbb{C}[J\mathfrak{g}^{{\ast}}]\) with the symmetric algebra \(S(\mathfrak{g}[t^{-1}]t^{-1})\).

Let x = x (−1) | 0〉 = (xt −1) | 0〉, where we denote by | 0〉 the unite element in \(S(\mathfrak{g}[t^{-1}]t^{-1})\). Then (29) gives that

$$\displaystyle{ [x_{(m)},y_{(n)}] = [x,y]_{(m+n)},\quad x,y \in \mathfrak{g},\ m,n \in \mathbb{Z}_{\geqslant 0}. }$$
(31)

So the Lie algebra \(J\mathfrak{g} = \mathfrak{g}[[t]]\) acts on \(\mathbb{C}[J\mathfrak{g}^{{\ast}}]\). This action coincides with that obtained by differentiating the action of JG = G[[t]] on \(J\mathfrak{g}^{{\ast}}\) induced by the coadjoint action of G. In other words, the vertex Poisson algebra structure of \(\mathbb{C}[J\mathfrak{g}^{{\ast}}]\) comes from the JG-action on \(J\mathfrak{g}^{{\ast}}\).

3.5 Canonical Filtration of Vertex Algebras

Haisheng Li [68] has shown that every vertex algebra is canonically filtered: For a vertex algebra V, let F p V be the subspace of V spanned by the elements

$$\displaystyle{ a_{(-n_{1}-1)}^{1}a_{ (-n_{2}-1)}^{2}\cdots a_{ (-n_{r}-1)}^{r}\vert 0\rangle }$$

with a 1, a 2, ⋯ , a rV, \(n_{i}\geqslant 0\), \(n_{1} + n_{2} + \cdots + n_{r}\geqslant p\). Then

$$\displaystyle{ V = F^{0}V \supset F^{1}V \supset \ldots. }$$

It is clear that TF p VF p+1 V.

Set \((F^{p}V )_{(n)}F^{q}V:=\mathop{ \mathrm{span}}\nolimits _{\mathbb{C}}\{a_{(n)}b\vert a \in F^{p}V,b \in F^{q}V \}\).

Lemma 16

We have

$$\displaystyle{ F^{p}V =\sum _{ j\geqslant 0}(F^{0}V )_{ (-j-1)}F^{p-j}V. }$$

Proposition 5

  1. (1)

    (F p V )(n)(F q V ) ⊂ F p+qn−1 V. Moreover, if \(n\geqslant 0\) , we have (F p V )(n)(F q V ) ⊂ F p+qn V.

  2. (2)

    The filtration F V is separated, that is, \(\bigcap _{p\geqslant 0}F^{p}V =\{ 0\}\) , if V is a positive energy representation over itself.

Proof

It is straightforward to check. [(2) also follows from Lemma 17 below.]

In this note we assume that the filtration F V is separated.

Set

$$\displaystyle{ \mathop{\mathrm{gr}}\nolimits V =\bigoplus _{p\geqslant 0}F^{p}V/F^{p+1}V. }$$

We denote by σ p : F p VF p VF p+1 V for \(p\geqslant 0\), the canonical quotient map.

Proposition 5 gives the following.

Proposition 6 ([68])

The space \(\mathop{\mathrm{gr}}\nolimits V\) is a Poisson vertex algebra by

$$\displaystyle{ \sigma _{p}(a) \cdot \sigma _{q}(b):=\sigma _{p+q}(a_{(-1)}b),\quad \sigma _{p}(a)_{(n)}\sigma _{q}(b):=\sigma _{p+q-n}(a_{(n)}b) }$$

for aF p V, bF q V, \(n\geqslant 0\).

Set

$$\displaystyle{ R_{V }:= F^{0}V/F^{1}V \subset \mathop{\mathrm{gr}}\nolimits V. }$$

Note that \(F^{1}V =\mathop{ \mathrm{span}}\nolimits _{\mathbb{C}}\{a_{(-2)}b\mid a,b \in V \}\).

Proposition 7 ([68, 79])

The restriction of the Poisson structure gives R V a Poisson algebra structure, that is, R V is a Poisson algebra by

$$\displaystyle{ \bar{a} \cdot \bar{ b}:= \overline{a_{(-1)}b},\quad \{\bar{a},\bar{b}\} = \overline{a_{(0)}b}, }$$

where \(\bar{a} =\sigma _{0}(a)\).

Proof

It is straightforward from Proposition 6.

In the literature F 1 V is often denoted by C 2(V ) and the Poisson algebra R V is called Zhu’s C 2 -algebra.

A vertex algebra V is called finitely strongly generated if R V is finitely generated as a ring. If the images of vectors a 1, , a N V generate R V , we say that V is strongly generated by a 1, , a N .

Below we always assume that a vertex algebra V is finitely strongly generated.

Note that if ϕ: VW is a homomorphism of vertex algebras, ϕ respects the canonical filtration, that is, ϕ(F p V ) ⊂ F p W. Hence it induces the homomorphism \(\mathop{\mathrm{gr}}\nolimits V \rightarrow \mathop{\mathrm{gr}}\nolimits W\) of Poisson vertex algebra homomorphism which we denote by \(\mathop{\mathrm{gr}}\nolimits \phi\).

3.6 Associated Variety and Singular Support of Vertex Algebras

Definition 1

Define the associated scheme \(\tilde{X}_{V }\) and the associated variety X V of a vertex algebra V as

$$\displaystyle{ \tilde{X}_{V }:=\mathop{ \mathrm{Spec}}\nolimits R_{V },\quad X_{V }:=\mathop{ \mathrm{Specm}}\nolimits R_{V } = (\tilde{X}_{V })_{\mathop{\mathrm{red}}\nolimits }. }$$

It was shown in [68, Lemma4.2] that \(\mathop{\mathrm{gr}}\nolimits V\) is generated by the subring R V as a differential algebra. Thus, we have a surjection \(JR_{V } \rightarrow \mathop{\mathrm{gr}}\nolimits V\) of differential algebras by Remark 3. This is in fact a homomorphism of Poisson vertex algebras:

Theorem 10 ([68, Lemma4.2], [7, Proposition 2.5.1])

The identity map R V R V induces a surjective Poisson vertex algebra homomorphism

$$\displaystyle{ JR_{V } = \mathbb{C}[J\tilde{X}_{V }] \twoheadrightarrow \mathop{\mathrm{gr}}\nolimits V. }$$

Let a 1, , a n be a set of strong generators of V. Since \(\mathop{\mathrm{gr}}\nolimits V \cong V\) as \(\mathbb{C}\)-vector spaces by the assumption that F V is separated, it follows from Theorem 10 that V is spanned by elements

$$\displaystyle{ a_{(-n_{1})}^{i_{1} }\ldots a_{(-n_{r})}^{i_{r} }\vert 0\rangle \quad \text{with }r\geqslant 0,\ n_{i}\geqslant 1. }$$

Definition 2

Define the singular support of a vertex algebra V as

$$\displaystyle{ SS(V ):=\mathop{ \mathrm{Spec}}\nolimits (\mathop{\mathrm{gr}}\nolimits V ) \subset J\tilde{X}_{V }. }$$

Theorem 11

We have dimSS(V ) = 0 if and only if dimX V = 0.

Proof

The “only if” part is obvious sine \(\pi _{\infty }(SS(V )) =\tilde{ X}_{V }\), where \(\pi _{\infty }: J\tilde{X}_{V } \rightarrow \tilde{ X}_{V }\) is the projection. The “if” part follows from Corollary 2.

Definition 3

We call V lisse (or C 2 -cofinite) if dimX V = 0.

Remark 5

Suppose that V is \(\mathbb{Z}_{+}\)-graded, so that \(V =\bigoplus _{i\geqslant 0}V _{i}\), and that \(V _{0} = \mathbb{C}\vert 0\rangle\). Then \(\mathop{\mathrm{gr}}\nolimits V\) and R V are equipped with the induced grading:

$$\displaystyle\begin{array}{rcl} & & \mathop{\mathrm{gr}}\nolimits V =\bigoplus _{i\geqslant 0}(\mathop{\mathrm{gr}}\nolimits V )_{i},\quad (\mathop{\mathrm{gr}}\nolimits V )_{0} = \mathbb{C}, {}\\ & & R_{V } =\bigoplus _{i\geqslant 0}(R_{V })_{i},\quad (R_{V })_{0} = \mathbb{C}. {}\\ \end{array}$$

So the following conditions are equivalent:

  1. 1.

    V is lisse.

  2. 2.

    X V = {0}.

  3. 3.

    The image of any vector aV i for \(i\geqslant 1\) in \(\mathop{\mathrm{gr}}\nolimits V\) is nilpotent.

  4. 4.

    The image of any vector aV i for \(i\geqslant 1\) in R V is nilpotent.

Thus, lisse vertex algebras can be regarded as a generalization of finite-dimensional algebras.

Remark 6

Suppose that the Poisson structure of R V is trivial. Then the Poisson vertex algebra structure of JR V is trivial, and so is that of \(\mathop{\mathrm{gr}}\nolimits V\) by Theorem 10. This happens if and only if

$$\displaystyle{ (F^{p}V )_{ (n)}(F^{q}V ) \subset F^{p+q-n+1}V \quad \text{ for all}\ n\geqslant 0. }$$

If this is the case, one can give \(\mathop{\mathrm{gr}}\nolimits V\) yet another Poisson vertex algebra structure by setting

$$\displaystyle{ \sigma _{p}(a)_{(n)}\sigma _{q}(b):=\sigma _{p+q-n+1}(a_{(n)}b)\quad \text{for }n\geqslant 0. }$$
(32)

(We can repeat this procedure if this Poisson vertex algebra structure is again trivial.)

3.7 Comparison with Weight-Depending Filtration

Let V be a vertex algebra that is \(\mathbb{Z}\)-graded by some Hamiltonian H:

$$\displaystyle{ V =\bigoplus _{\varDelta \in \mathbb{Z}}V _{\varDelta }\quad \mbox{ where}\quad V _{\varDelta }:=\{ v \in V \vert Hv =\varDelta v\}. }$$

Then there is [67] another natural filtration of V defined as follows.

For a homogeneous vector aV Δ , Δ is called the conformal weight of a and is denote by Δ a . Let G p V be the subspace of V spanned by the vectors

$$\displaystyle{ a_{(-n_{1}-1)}^{1}a_{ (-n_{2}-1)}^{2}\cdots a_{ (-n_{r}-1)}^{r}\vert 0\rangle }$$

with \(\varDelta _{a^{1}} +\ldots +\varDelta _{a^{r}}\leqslant p\). Then G V defines an increasing filtration of V:

$$\displaystyle{ 0 = G_{-1}V \subset G_{0}V \subset \ldots G_{1}V \subset \ldots,\quad V =\bigcup _{p}G_{p}V. }$$

Moreover we have

$$\displaystyle\begin{array}{rcl} & & TG_{p}V \subset G_{p}V, {}\\ & & (G_{p})_{(n)}G_{q}V \subset G_{p+q}V \quad \text{for }n \in \mathbb{Z}, {}\\ & & (G_{p})_{(n)}G_{q}V \subset G_{p+q-1}V \quad \text{for }n \in \mathbb{Z}_{+}, {}\\ \end{array}$$

It follows that \(\mathop{\mathrm{gr}}\nolimits _{G}V =\bigoplus G_{p}V/G_{p-1}V\) is naturally a Poisson vertex algebra.

It is not too difficult to see the following.

Lemma 17 ([7, Proposition 2.6.1])

We have

$$\displaystyle{ F^{p}V _{\varDelta } = G_{\varDelta -p}V _{\varDelta }, }$$

where F p V Δ = V Δ F p V, G p V Δ = V Δ G p V. Therefore

$$\displaystyle{ \mathop{\mathrm{gr}}\nolimits V \cong \mathop{\mathrm{gr}}\nolimits _{G}V }$$

as Poisson vertex algebras.

3.8 Example: Universal Affine Vertex Algebras

Let \(\mathfrak{a}\) be a Lie algebra with a symmetric invariant bilinear form κ. Let

$$\displaystyle{ \widehat{\mathfrak{a}} = \mathfrak{a}[t,t^{-1}]\mathop{\oplus }\mathbb{C}\mathbf{1} }$$

be the Kac-Moody affinization of \(\mathfrak{a}\). It is a Lie algebra with commutation relations

$$\displaystyle{ [xt^{m},yt^{n}] = [x,y]t^{m+n} + m\delta _{ m+n,0}\kappa (x,y)\mathbf{1},\quad x,y \in \mathfrak{a},\ m,n \in \mathbb{Z},\quad \quad [\mathbf{1},\widehat{\mathfrak{a}}] = 0. }$$

Let

$$\displaystyle{ V ^{\kappa }(\mathfrak{a}) = U(\hat{\mathfrak{a}}) \otimes _{U(\mathfrak{a}[t]\oplus \mathbb{C}\mathbf{1})}\mathbb{C}, }$$

where \(\mathbb{C}\) is one-dimensional representation of \(\mathfrak{a}[t] \oplus \mathbb{C}\mathbf{1}\) on which \(\mathfrak{a}[t]\) acts trivially and 1 acts as the identity. The space \(V ^{\kappa }(\mathfrak{a})\) is naturally graded: \(V ^{\kappa }(\mathfrak{a}) =\bigoplus _{\varDelta \in \mathbb{Z}_{\geqslant 0}}V ^{\kappa }(\mathfrak{a})_{\varDelta }\), where the grading is defined by setting degxt n = −n, deg | 0〉 = 0. Here | 0〉 = 1 ⊗ 1. We have \(V ^{\kappa }(\mathfrak{a})_{0} = \mathbb{C}\vert 0\rangle\). We identify \(\mathfrak{a}\) with \(V ^{\kappa }(\mathfrak{a})_{1}\) via the linear isomorphism defined by xxt −1 | 0〉.

There is a unique vertex algebra structure on \(V ^{\kappa }(\mathfrak{a})\) such that | 0〉 is the vacuum vector and

$$\displaystyle{ Y (x,z) = x(z):=\sum _{n\in \mathbb{Z}}(xt^{n})z^{-n-1},\quad x \in \mathfrak{a}. }$$

(So x (n) = xt n for \(x \in \mathfrak{a} = V ^{\kappa }(\mathfrak{a})_{1}\), \(n \in \mathbb{Z}\)).

The vertex algebra \(V ^{\kappa }(\mathfrak{a})\) is called the universal affine vertex algebra associated with (\(\mathfrak{a}\), κ).

We have \(F^{1}V ^{\kappa }(\mathfrak{a}) = \mathfrak{a}[t^{-1}]t^{-2}V ^{\kappa }(\mathfrak{a})\), and the Poisson algebra isomorphism

$$\displaystyle\begin{array}{rcl} & \mathbb{C}[\mathfrak{a}^{{\ast}}]\;\stackrel{_{\sim }}{\rightarrow }\;R_{V ^{\kappa }(\mathfrak{a})} = V ^{k}(\mathfrak{a})/\mathfrak{g}[t^{-1}]t^{-2}V ^{\kappa }(\mathfrak{a})& \\ & x_{1}\ldots x_{r}\longmapsto \overline{(x_{1}t^{-1})\ldots (x_{r}t^{-1})\vert 0\rangle }\quad (x_{i} \in \mathfrak{a}). &{}\end{array}$$
(33)

Thus

$$\displaystyle{ X_{V ^{\kappa }(\mathfrak{a})} = \mathfrak{a}^{{\ast}}. }$$

We have the isomorphism

$$\displaystyle{ \mathbb{C}[J\mathfrak{a}^{{\ast}}] \simeq \mathop{\mathrm{gr}}\nolimits V ^{\kappa }(\mathfrak{a}) }$$
(34)

because the graded dimensions of both sides coincide. Therefore

$$\displaystyle{ SS(V ^{\kappa }(\mathfrak{a})) = J\mathfrak{a}^{{\ast}}. }$$

The isomorphism (34) follows also from the fact that

$$\displaystyle{ G_{p}V ^{\kappa }(\mathfrak{a}) = U_{p}(\mathfrak{a}[t^{-1}]t^{-1})\vert 0\rangle, }$$

where \(\{U_{p}(\mathfrak{a}[t^{-1}]t^{-1})\}\) is the PBW filtration of \(U(\mathfrak{a}[t^{-1}]t^{-1})\).

3.9 Example: Simple Affine Vertex Algebras

For a finite-dimensional simple Lie algebra \(\mathfrak{g}\) and \(k \in \mathbb{C}\), we denote by \(V ^{k}(\mathfrak{g})\) the universal affine vertex algebra \(V ^{k\kappa _{0}}(\mathfrak{g})\), where κ 0 is the normalized invariant inner product of \(\mathfrak{g}\), that is,

$$\displaystyle{ \kappa _{0}(\theta,\theta ) = 2, }$$

where θ is the highest root of \(\mathfrak{g}\). Denote by \(V _{k}(\mathfrak{g})\) the unique simple graded quotient of \(V ^{k}(\mathfrak{g})\). As a \(\hat{\mathfrak{g}}\)-module, \(V _{k}(\mathfrak{g})\) is isomorphic to the irreducible highest weight representation L( 0) of \(\widehat{\mathfrak{g}}\) with highest weight 0, where Λ 0 is the weight of the basic representation of \(\widehat{\mathfrak{g}}\).

Theorem 12

The vertex algebra \(V _{k}(\mathfrak{g})\) is lisse if and only if \(V _{k}(\mathfrak{g})\) is integrable as a \(\hat{\mathfrak{g}}\) -module, which is true if and only if \(k \in \mathbb{Z}_{+}\).

Lemma 18

Let (R, ) be a differential algebra over \(\mathbb{Q}\) , I a differential ideal of R, i.e., I is an ideal of R such that ∂II. Then \(\partial \sqrt{I} \subset \sqrt{I}\).

Proof

Let \(a \in \sqrt{I}\), so that a mI for some \(m \in \mathbb{N} =\{ 1,2,\ldots \}\). Since I is -invariant, we have m a mI. But

$$\displaystyle{ \partial ^{m}a^{m} =\sum _{ 0\leqslant i\leqslant m}{m\choose i}a^{m-i}(\partial a)^{i} \equiv m!(\partial a)^{m}\pmod \sqrt{I}. }$$

Hence \((\partial a)^{m} \in \sqrt{I}\), and therefore, \(\partial a \in \sqrt{I}\).

Proof (Proof of the “if” Part of Theorem 12)

Suppose that \(V _{k}(\mathfrak{g})\) is integrable. This condition is equivalent to that \(k \in \mathbb{Z}_{+}\) and the maximal submodule N k of \(V ^{k}(\mathfrak{g})\) is generated by the singular vector (e θ t −1)k+1 | 0〉 [54]. The exact sequence \(0 \rightarrow N_{k} \rightarrow V ^{k}(\mathfrak{g}) \rightarrow V _{k}(\mathfrak{g}) \rightarrow 0\) induces the exact sequence

$$\displaystyle{ 0 \rightarrow I_{k} \rightarrow R_{V ^{k}(\mathfrak{g})} \rightarrow R_{V _{k}(\mathfrak{g})} \rightarrow 0, }$$

where I k is the image of N k in \(R_{V ^{k}(\mathfrak{g})} = \mathbb{C}[\mathfrak{g}^{{\ast}}]\), and so, \(R_{V _{k}(\mathfrak{g})} = \mathbb{C}[\mathfrak{g}^{{\ast}}]/I_{k}\). The image of the singular vector in I k is given by e θ k+1. Therefore, \(e_{\theta } \in \sqrt{I}\). On the other hand, by Lemma 18, \(\sqrt{I_{k}}\) is preserved by the adjoint action of \(\mathfrak{g}\). Since \(\mathfrak{g}\) is simple, \(\mathfrak{g} \subset \sqrt{I}\). This proves that \(X_{V _{k}(\mathfrak{g})} =\{ 0\}\) as required.

The proof of “only if” part follows from [30]. We will give a different proof using W-algebras in Remark 13.

In view of Theorem 12, one may regard the lisse condition as a generalization of the integrability condition to an arbitrary vertex algebra.

4 Zhu’s Algebra

In this section we will introduce and study the Zhu’s algebra of a vertex algebra, which plays an important role in the representation theory.

See [55] in this volume for the definition of modules over vertex algebras.

4.1 Zhu’s C 2-Algebra and Zhu’s Algebra of a Vertex Algebra

Let V be a \(\mathbb{Z}\)-graded vertex algebra. Zhu’s algebra ZhuV [43, 79] is defined as

$$\displaystyle{ \mathrm{Zhu}(V ):= V/V \circ V }$$

where \(V \circ V:=\mathop{ \mathrm{span}}\nolimits \{a \circ b\vert a,b \in V \}\) and

$$\displaystyle{ a \circ b:=\sum _{i\geqslant 0}{\varDelta _{a}\choose i}a_{(i-2)}b }$$

for homogeneous elements a, b and extended linearly. It is an associative algebra with multiplication defined as

$$\displaystyle{ a {\ast} b:=\sum _{i\geqslant 0}{\varDelta _{a}\choose i}a_{(i-1)}b }$$

for homogeneous elements a, bV.

For a simple positive energy representation \(M =\bigoplus _{n\in \mathbb{Z}_{+}}M_{\lambda +n}\), M λ ≠ 0, \(\lambda \in \mathbb{C}\), of V, let M top be the top degree component M λ of M. Also, for a homogeneous vector aV, let \(o(a) = a_{(\varDelta _{a}-1)}\), so that o(a) preserves the homogeneous components of any graded representation of V.

The importance of Zhu’s algebra in vertex algebra theory is the following fact that was established by Yonchang Zhu.

Theorem 13 ([79])

For any positive energy representation M of V, \(\overline{a}\mapsto o(a)\) defines a well-defined representation of Zhu(V ) on M top . Moreover, the correspondence M ↦ M top gives a bijection between the set of isomorphism classes of irreducible positive energy representations of V and that of simple Zhu(V )-modules.

A vertex algebra V is called a chiralization of an algebra A if \(\mathrm{Zhu}(V )\cong A\).

Now we define an increasing filtration of Zhu’s algebra. For this, we assume that V is \(\mathbb{Z}_{+}\)-graded: \(V =\bigoplus _{\varDelta \geqslant 0}V _{\varDelta }\). Then \(V _{\leqslant p} =\bigoplus _{ \varDelta =0}^{p}V _{\varDelta }\) gives an increasing filtration of V. Define

$$\displaystyle{ \mathrm{Zhu}_{p}(V ):=\mathop{ \mathrm{Im}}\nolimits (V _{\leqslant p} \rightarrow \mathrm{ Zhu}(V )). }$$

Obviously, we have

$$\displaystyle{ 0 =\mathrm{ Zhu}_{-1}(V ) \subset \mathrm{ Zhu}_{0}(V ) \subset \mathrm{ Zhu}_{1}(V ) \subset \cdots \,,\quad \mbox{ and}\quad \mathrm{Zhu}(V ) =\bigcup _{p\geqslant -1}\mathrm{Zhu}_{p}(V ). }$$

Also, since \(a_{(n)}b \in V _{\varDelta _{a}+\varDelta _{b}-n-1}\) for \(a \in V _{\varDelta _{a}}\), \(b \in V _{\varDelta _{b}}\), we have

$$\displaystyle{ \mathrm{Zhu}_{p}(V ) {\ast}\mathrm{ Zhu}_{q}(V ) \subset \mathrm{ Zhu}_{p+q}(V ). }$$
(35)

The following assertion follows from the skew symmetry.

Lemma 19

We have

$$\displaystyle{ b {\ast} a \equiv \sum _{i\geqslant 0}{\varDelta _{a} - 1\choose i}a_{(i-1)}b\quad \pmod V \circ V, }$$

and hence,

$$\displaystyle{ a {\ast} b - b {\ast} a \equiv \sum _{i\geqslant 0}{\varDelta _{a} - 1\choose i}a_{(i)}b\quad \pmod V \circ V. }$$

By Lemma 19, we have

$$\displaystyle{ [\mathrm{Zhu}_{p}(V ),\mathrm{Zhu}_{q}(V )] \subset \mathrm{ Zhu}_{p+q-1}(V ). }$$
(36)

By (35) and (36), the associated graded \(\mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(V ) =\bigoplus _{p}\mathrm{Zhu}_{p}(V )/\mathrm{Zhu}_{p-2}(V )\) is naturally a graded Poisson algebra.

Note that \(a \circ b \equiv a_{(-2)}b\pmod \bigoplus _{\varDelta \leqslant \varDelta _{a}+\varDelta _{b}}V _{\varDelta }\) for homogeneous elements a, bV.

Lemma 20 (Zhu, See [29, Proposition 2.17(c)], [17, Proposition 3.3])

The following map defines a well-defined surjective homomorphism of Poisson algebras.

$$\displaystyle\begin{array}{rcl} \eta _{V }: R_{V }& \longrightarrow & \mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(V ) {}\\ \bar{a}& \longmapsto & a\pmod V \circ V +\bigoplus _{\varDelta <\varDelta _{a}}V _{\varDelta }. {}\\ \end{array}$$

Remark 7

The map η V is not an isomorphism in general. For an example, let \(\mathfrak{g}\) be the simple Lie algebra of type E 8 and \(V = V _{1}(\mathfrak{g})\). Then dimR V > dimZhu(V ) = 1.

Corollary 3

If V is lisse then Zhu(V ) is finite dimensional. Hence the number of isomorphic classes of simple positive energy representations of V is finite.

In fact the following stronger facts are known.

Theorem 14 ([1])

Let V be lisse. Then any simple V -module is a positive energy representation. Therefore the number of isomorphic classes of simple V -modules is finite.

Theorem 15 ([31, 73])

Le V be lisse. Then the abelian category of V -modules is equivalent to the module category of a finite-dimensional associative algebra.

4.2 Computation of Zhu’s Algebras

We say that a vertex algebra V admits a PBW basis if R V is a polynomial algebra and the map \(\mathbb{C}[JX_{V }] \twoheadrightarrow \mathop{\mathrm{gr}}\nolimits V\) is an isomorphism.

Theorem 16

If V admits a PBW basis, then \(\eta _{V }: R_{V } \twoheadrightarrow \mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}V\) is an isomorphism.

Proof

We have \(\mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(V ) = V/\mathop{\mathrm{gr}}\nolimits (V \circ V )\), where \(\mathop{\mathrm{gr}}\nolimits (V \circ V )\) is the associated graded space of VV with respect to the filtration induced by the filtration \(V _{\leqslant p}\). We wish to show that \(\mathop{\mathrm{gr}}\nolimits (V \circ V ) = F^{1}V\). Since \(a \circ b \equiv a_{(-2)}b\pmod F_{\leqslant \varDelta _{a}+\varDelta _{b}}V\), it is sufficient to show that ab ≠ 0 implies that a (−2) b ≠ 0.

Suppose that a (−2) b = (Ta)(−1) b = 0. Since V admits a PBW basis, \(\mathop{\mathrm{gr}}\nolimits V\) has no zero divisors. That fact that V admits a PBW basis also shows that Ta = 0 implies that a = c | 0〉 for some constant \(c \in \mathbb{C}\). Thus, a is a constant multiple of | 0〉, in which case ab = 0.

Example 2 (Universal Affine Vertex Algebras)

The universal affine vertex algebra \(V ^{\kappa }(\mathfrak{a})\) (see Sect. 3.8) admits a PBW basis. Therefore

$$\displaystyle{ \eta _{V ^{\kappa }(\mathfrak{a})}: R_{V ^{\kappa }(\mathfrak{a})} = \mathbb{C}[\mathfrak{a}^{{\ast}}]\;\stackrel{_{ \sim }}{\rightarrow }\;\mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}V ^{\kappa }(\mathfrak{a}). }$$

On the other hand, from Lemma 19 one finds that

$$\displaystyle\begin{array}{rcl} U(\mathfrak{a})& \longrightarrow & \mathrm{Zhu}(V ^{\kappa }(\mathfrak{a})) \\ \mathfrak{a}x& \longmapsto & \bar{x} = \overline{x_{(-1)}\vert 0\rangle }{}\end{array}$$
(37)

gives a well-defined algebra homomorphism. This map respects the filtration on both sides, where the filtration in the left-hand-side is the PBW filtration. Hence it induces a map between their associated graded algebras, which is identical to \(\eta _{V ^{\kappa }(\mathfrak{a})}\). Therefore (37) is an isomorphism, that is to say, \(V ^{\kappa }(\mathfrak{a})\) is a chiralization of \(U(\mathfrak{a})\).

Exercise 1

Extend Theorem 16 to the case that \(\mathfrak{a}\) is a Lie superalgebra.

Example 3 (Free Fermions)

Let \(\mathfrak{n}\) be a finite-dimensional vector space. The Clifford affinization \(\hat{Cl}\) of \(\mathfrak{n}\) is the Clifford algebra associated with \(\mathfrak{n}[t,t^{-1}] \oplus \mathfrak{n}^{{\ast}}[t,t^{-1}]\) and its symmetric bilinear form defined by

$$\displaystyle{ (xt^{m}\vert ft^{n}) =\delta _{ m+n,0}f(x),\ (xt^{m}\vert yt^{n}) = 0 = (\,ft^{m}\vert gt^{n}) }$$

for \(x,y \in \mathfrak{n}\), \(f,g \in \mathfrak{n}^{{\ast}}\), \(m,n \in \mathbb{Z}\).

Let \(\{x_{\alpha }\}_{\alpha \in \varDelta _{+}}\) be a basis of \(\mathfrak{n}\), {x α } its dual basis. We write ψ α, m for \(x_{\alpha }t^{m} \in \hat{ Cl}\) and ψ α, m for \(x_{\alpha }^{{\ast}}t^{m} \in \hat{ Cl}\), so that \(\hat{Cl}\) is the associative superalgebra with

  • odd generators: \(\psi _{\alpha,m},\psi _{\alpha,m}^{{\ast}},m \in \mathbb{Z},\alpha \in \varDelta _{+}\).

  • relations: [ψ α, m , ψ β, n ] = [ψ α, m , ψ β, n ] = 0, [ψ α, m , ψ β, n ] = δ α, β δ m+n, 0. 

Define the charged fermion Fock space associated with \(\mathfrak{n}\) as

It is an irreducible \(\hat{Cl}\)-module, and as \(\mathbb{C}\)-vector spaces we have

There is a unique vertex (super)algebra structure on such that the image of 1 is the vacuum | 0〉 and

$$\displaystyle{ Y (\psi _{\alpha,-1}\vert 0\rangle,z) =\psi _{\alpha }(z):=\sum _{n\in \mathbb{Z}}\psi _{\alpha,n}z^{-n-1},Y (\psi _{\alpha,0}^{{\ast}}\vert 0\rangle,z) =\psi _{ \alpha }^{{\ast}}(z):=\sum _{ n\in \mathbb{Z}}\psi _{\alpha,n}^{{\ast}}z^{-n}. }$$

We have , and it follows that there is an isomorphism

as Poisson superalgebras. Thus,

where \(\varPi \mathfrak{n}\) is the space \(\mathfrak{n}\) considered as a purely odd affine space. The arc space \(JT^{{\ast}}\varPi \mathfrak{n}\) is also regarded as a purely odd affine space, such that \(\mathbb{C}[JT^{{\ast}}\varPi \mathfrak{n}] =\varLambda (\mathfrak{n}^{{\ast}}[t^{-1}]) \otimes \varLambda (\mathfrak{n}[t^{-1}]t^{-1})\). The map is an isomorphism and admits a PBW basis. Therefore we have the isomorphism

by Exercise 1. On the other hand the map

gives an algebra homomorphism that respects the filtration. Hence we have

That is, is a chiralization of Cl.

5 W-Algebras

We are now in a position to define W-algebras. We will construct a differential graded vertex algebra, so that its cohomology algebra is a vertex algebra and that will be our main object to study.

For simplicity, we let \(\mathfrak{g} = \mathfrak{g}\mathfrak{l}_{n}\) and we only consider the principal nilpotent case. However the definition works for any simple Lie algebra. The general definition for an arbitrary nilpotent element will be similar but one does need a new idea (see [60] for the most general definition).

5.1 The BRST Complex

Let \(\mathfrak{g},\mathfrak{n}\) be as in Sect. 2.1. Denote by \(\kappa _{\mathfrak{g}}\) the Killing form on \(\mathfrak{g}\) and \(\kappa _{0} = \frac{1} {2n}\kappa _{\mathfrak{g}}\), so that κ 0(θ, θ) = 2.

Choose any symmetric invariant bilinear form κ on \(\mathfrak{g}\) and let \(V ^{\kappa }(\mathfrak{g})\) be the universal affine vertex algebra associated with \((\mathfrak{g},\kappa )\) (see Sect. 3.8) and let be the fermion Fock space as in Example 3.

We have the following commutative diagrams:

Define

Since it is a tensor product of two vertex algebras, \(C^{\kappa }(\mathfrak{g})\) is a vertex algebra. We have

and

Thus, \(C^{\kappa }(\mathfrak{g})\) is a chiralization of \(C(\mathfrak{g})\) considered in Sect. 2.5. Further we have

So we have the following commutative diagram:

Define a gradation

(38)

by setting \(\mathrm{deg}\psi _{\alpha,m} = -1,\mathrm{deg}\psi _{\alpha,k}^{{\ast}} = 1,\forall i,j \in I,m,k \in \mathbb{Z}\), deg | 0〉 = 0. This induces a \(\mathbb{Z}\)-grading (that is different from the conformal grading) on \(C^{\kappa }(\mathfrak{g})\):

(39)

Let \(V (\mathfrak{n})\) be the the universal affine vertex algebra associated with \(\mathfrak{n}\) and the zero bilinear form, which is identified with the vertex subalgebra of \(V ^{\kappa }(\mathfrak{g})\) generated by x α (z) with αΔ +.

Lemma 21

The following defines a vertex algebra homomorphism.

Remark 8

In the above formula the normally ordered product is not needed because \(\mathfrak{n}\) is nilpotent.

The map \(\hat{\rho }\) induces an algebra homomorphism

and a Poisson algebra homomorphism

that are identical to ρ and \(\bar{\rho }\) (see Lemma 3 and 9), respectively.

Recall the character \(\chi: \mathfrak{n} \rightarrow \mathbb{C}\), x ↦ ( f | x).

Lemma 22

The following defines a vertex algebra homomorphism.

$$\displaystyle\begin{array}{rcl} \hat{\theta }_{\chi }: V (\mathfrak{n})& \longrightarrow & C^{\kappa }(\mathfrak{g}) {}\\ x_{\alpha }(z)& \longmapsto & (x_{\alpha }(z) +\chi (x_{\alpha })) \otimes \mathrm{ id} +\mathrm{ id} \otimes \hat{\rho } (x_{\alpha }(z)). {}\\ \end{array}$$

The map \(\hat{\theta }_{\chi }\) induces an algebra homomorphism

$$\displaystyle{ \mathrm{Zhu}(V (\mathfrak{n})) = U(\mathfrak{n}) \rightarrow \mathrm{ Zhu}C^{\kappa }(\mathfrak{g}) = C(\mathfrak{g}) }$$

and a Poisson algebra homomorphism

that are identical to θ χ and \(\bar{\theta }\), respectively (see Lemmas 5 and 7).

The proof of the following assertion is similar to that of Lemma 6.

Proposition 8

There exists a unique element \(\hat{Q} \in C^{k,1}(g)\) such that

$$\displaystyle{ [\hat{Q}_{\lambda }(1 \otimes \psi _{\alpha })] =\hat{\theta } _{\chi }(x_{\alpha }),\quad \forall \alpha \in \varDelta _{+}. }$$

We have \([\hat{Q}_{\lambda }\hat{Q}] = 0\).

The field \(\hat{Q}(z)\) is given explicitly as

$$\displaystyle{ \hat{Q}(z) =\sum _{\alpha \in \varDelta _{+}}(x_{\alpha } +\chi (x_{\alpha })) \otimes \psi _{\alpha }^{{\ast}}(z) -\mathrm{ id} \otimes \frac{1} {2}\sum _{\alpha,\beta,\gamma \in \varDelta _{+}}c_{\alpha,\beta }^{\gamma }\psi _{ \alpha }^{{\ast}}(z)\psi _{\beta }^{{\ast}}(z)\psi _{\gamma }(z). }$$

Since \(\hat{Q}\) is odd and \([\hat{Q}_{\lambda }\hat{Q}] = 0\), we have

$$\displaystyle{ \hat{Q}_{(0)}^{2} = 0. }$$

(Recall that we write \(\hat{Q}(z) =\sum _{n\in \mathbb{Z}}\hat{Q}_{(n)}z^{-n-1}\).) So \((C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) is a cochain complex.

Lemma 23

If it is nonzero, the cohomology \(H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) inherits the vertex algebra structure from \(C^{\kappa }(\mathfrak{g})\).

Proof

Set \(Z:=\{ v \in C^{\kappa }(\mathfrak{g})\mid \hat{Q}_{(0)}v = 0\}\), \(B =\hat{ Q}_{(0)}C^{\kappa }(\mathfrak{g}) \subset Z\), so that \(H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)}) = Z/B\). From the commutator formula (16), we know that

$$\displaystyle{ [\hat{Q}_{(0)},a_{(m)}] = (\hat{Q}_{(0)}a)_{(m)}\quad \forall a \in C^{\kappa }(\mathfrak{g}),m \in \mathbb{Z}. }$$

Thus, if a, bZ, then \(\hat{Q}_{(0)}(a_{(m)}b) = 0\), that is, a (m) bZ. It follows that Z a vertex subalgebra of \(C^{\kappa }(\mathfrak{g})\). Further, if aZ and \(b =\hat{ Q}_{(0)}b' \in B\), then \(a_{(m)}b = a_{(m)}\hat{Q}_{(0)}b' =\hat{ Q}_{(0)}(a_{(m)}b) \in B\). Hence B is an ideal of Z. This completes the proof.

Definition 4

The W-algebra \(\mathcal{W}^{\kappa }(\mathfrak{g}) =\mathcal{ W}^{\kappa }(\mathfrak{g},f)\) associated to \((\mathfrak{g},f,\kappa )\) is defined to be the zero-th cohomology of the cochain complex \((C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\), that is,

$$\displaystyle{ \mathcal{W}^{\kappa }(\mathfrak{g}):= H^{0}(C^{\kappa }(\mathfrak{g}),\hat{Q}_{ (0)}). }$$

This definition of \(\mathcal{W}^{\kappa }(\mathfrak{g})\) is due to Feigin and Frenkel [36]. In Sect. 5.9 we show that the above \(\mathcal{W}^{\kappa }(\mathfrak{g})\) is identical to the original W-algebra defined by Fateev and Lukyanov [34].

5.2 Cohomology of Associated Graded

We have \(\hat{Q}_{(0)}F^{p}C^{\kappa }(\mathfrak{g}) \subset F^{p}C^{\kappa }(\mathfrak{g})\), so \((\mathop{\mathrm{gr}}\nolimits ^{F}C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) is also a cochain complex. The cohomology \(H^{\bullet }(\mathop{\mathrm{gr}}\nolimits ^{F}C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) inherits a Poisson vertex algebra structure from \(\mathop{\mathrm{gr}}\nolimits ^{F}C^{\kappa }(\mathfrak{g})\).

Theorem 17

We have \(H^{i}(\mathop{\mathrm{gr}}\nolimits ^{F}C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)}) = 0\) for i ≠ 0 and

as Poisson vertex algebras, where is the slice defined in Sect.  2 .

Proof

The proof is an arc space analogue of that of Theorem 3.

The moment map \(\mu: \mathfrak{g}^{{\ast}}\longrightarrow \mathfrak{n}^{{\ast}}\) for the N-action on \(\mathfrak{g}\) induces a JN-equivariant morphism

$$\displaystyle{ J\mu: J\mathfrak{g}^{{\ast}}\longrightarrow J\mathfrak{n}^{{\ast}}. }$$

The pullback \((J\mu )^{{\ast}}: \mathbb{C}[J\mathfrak{n}^{{\ast}}] \rightarrow \mathbb{C}[J\mathfrak{g}^{{\ast}}]\) is an embedding of vertex Poisson algebras.

The point χ = of \(J\mathfrak{n}^{{\ast}}\) corresponds to the arc \(\alpha \in \mathop{\mathrm{Hom}}\nolimits (D,\mathfrak{n}^{{\ast}}) =\mathop{ \mathrm{Hom}}\nolimits (\mathbb{C}[\mathfrak{n}^{{\ast}}], \mathbb{C}[[t]])\) such that α( f) = χ(x) for \(x \in \mathfrak{n} \subset \mathbb{C}[\mathfrak{n}^{{\ast}}]\).

We have

$$\displaystyle{ (J\mu )^{-1}(\chi ) = J(\mu ^{-1}(\chi )) =\chi +J\mathfrak{b} \subset J\mathfrak{g}^{{\ast}}, }$$

and the adjoint action gives the isomorphism

(40)

by Theorem 2 and (24).

Now put

$$\displaystyle{ C:=\mathop{ \mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g}) = \mathbb{C}[J\mathfrak{g}^{{\ast}}] \otimes \varLambda (\mathfrak{n}[t^{-1}]t^{-1}) \otimes \varLambda (\mathfrak{n}^{{\ast}}[t^{-1}]) }$$
(41)

and define a bigrading on C by

$$\displaystyle{ C =\bigoplus _{i\leqslant 0,j\geqslant 0}C^{i,j},\quad \mbox{ where}\quad C^{i,j} = \mathbb{C}[J\mathfrak{g}^{{\ast}}] \otimes \varLambda ^{-i}(\mathfrak{n}[t^{-1}]t^{-1}) \otimes \varLambda ^{j}(\mathfrak{n}^{{\ast}}[t^{-1}]). }$$
(42)

As before, we can decompose the operator \(\hat{Q}_{(0)}\) as the sum of two suboperators such that each of them preserves one grading but increase the other grading by 1. Namely, we have

$$\displaystyle\begin{array}{rcl} & & \hat{Q}_{(0)} =\hat{ d}_{+} +\hat{ d}_{-}, {}\\ & & \hat{d}_{-}: C^{i,j}\longrightarrow C^{i,j+1},\quad \hat{d}_{ +}: C^{i,j}\longrightarrow C^{i+1,j}. {}\\ \end{array}$$

This shows that

$$\displaystyle{ (\hat{d}_{+})^{2} = (\hat{d}_{ -})^{2} = [\hat{d}_{ +},\hat{d}_{-}] = 0. }$$

Thus we can get a spectral sequence \(E_{r}\Longrightarrow H^{\bullet }(C,\hat{Q}_{(0)})\) such that

$$\displaystyle{ E_{1} = H^{\bullet }(C,\hat{d}_{-}),\quad E_{2} = H^{\bullet }(H^{\bullet }(C,\hat{d}_{-}),\hat{d}_{+}). }$$

This is a converging spectral sequence since C is a direct sum of subcomplexes \(F^{p}C^{\kappa }(\mathfrak{g})/F^{p+1}C^{\kappa }(\mathfrak{g})\), and the associated filtration is regular on each subcomplex.

The complex \((C,\hat{d}_{-})\) is the Koszul complex with respect to the sequence

$$\displaystyle{ x_{1}t^{-1} -\chi (x_{ 1}),\ldots,x_{N}t^{-1} -\chi (x_{ N}),x_{1}t^{-2},x_{ 2}t^{-2},\ldots,x_{ N}t^{-2},x_{ 1}t^{-3},x_{ 2}t^{-3},\ldots }$$

where \(N = \mathrm{dim}\mathfrak{n}\). Hence we have

$$\displaystyle{ H^{i}(C,\hat{d}_{ -}) =\delta _{i,0}\mathbb{C}[J\mu ^{-1}(\chi )] \otimes \varLambda (\mathfrak{n}^{{\ast}}[t^{-1}]). }$$
(43)

Next, by (43), the complex \((H^{0}(C,\hat{d}_{-}),\hat{d}_{+})\) is identical to the Chevalley complex for the Lie algebra cohomology \(H^{\bullet }(\mathfrak{n}[t], \mathbb{C}[J\mu ^{-1}(\chi )])\). By (40),

We conclude that

Thus, the spectral sequence E r collapses at E 2 = E , and we get the desired isomorphisms.

Theorem 18 ([36, 40])

We have \(H^{0}(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)}) = 0\) for i ≠ 0 and

In particular, , so , .

The proof of Theorem 18 will be given in Sect. 5.6.

Note that there is a spectral sequence for \(H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) such that \(E_{1}^{\bullet,q} = H^{q}(\mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\). Hence Theorem 18 would immediately follow from Theorem 17 if this spectral sequence converges. However, this is not clear at this point because our algebra is not Noetherian.

Remark 9

The complex \((C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) is identical to Feigin’s standard complex for the semi-infinite \(\mathfrak{n}[t,t^{-1}]\)-cohomology \(H^{\frac{\infty } {2} +\bullet }(\mathfrak{n}[t,t^{-1}],V ^{\kappa }(\mathfrak{g}) \otimes \mathbb{C}_{\hat{\chi }})\) with coefficient in the \(\mathfrak{g}[t,t^{-1}]\)-module \(V ^{\kappa }(\mathfrak{g}) \otimes \mathbb{C}_{\hat{\chi }}\) [35], where \(\mathbb{C}_{\hat{\chi }}\) is the one-dimensional representation of \(\mathfrak{n}[t,t^{-1}]\) defined by the character \(\hat{\chi }: \mathfrak{n}[t,t^{-1}] \rightarrow \mathbb{C}\), xt nδ n, −1 χ(x):

$$\displaystyle{ H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\cong H^{\frac{\infty } {2} +\bullet }(\mathfrak{n}[t,t^{-1}],V ^{\kappa }(\mathfrak{g}) \otimes \mathbb{C}_{\hat{\chi }}). }$$
(44)

5.3 W-Algebra Associated with \(\mathfrak{s}\mathfrak{l}_{n}\)

It is straightforward to generalize the above definition to an arbitrary simple Lie algebra \(\mathfrak{g}\). In particular, by replacing \(V ^{\kappa }(\mathfrak{g}\mathfrak{l}_{n})\) with \(V ^{k}(\mathfrak{s}\mathfrak{l}_{n})\), \(k \in \mathbb{C}\), we define the W-algebra

$$\displaystyle{ \mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{ n}):= H^{0}(C^{k}(\mathfrak{s}\mathfrak{l}_{ n}),\hat{Q}_{(0)}) }$$
(45)

associated with \((\mathfrak{s}\mathfrak{l}_{n},f)\) at level k.

We have \(V ^{\kappa }(\mathfrak{g}\mathfrak{l}_{n}) =\pi _{\kappa } \otimes V ^{\kappa }(\mathfrak{s}\mathfrak{l}_{n})\), where \(\kappa \vert _{\mathfrak{s}\mathfrak{l}_{n}\times \mathfrak{s}\mathfrak{l}_{n}} = k\kappa _{0}\) and π κ is the rank 1 Heisenberg vertex algebra generated by I(z) = i = 1 n e ii (z) with λ-bracket [I λ I] = κ(I, I)λ. It follows that \(C^{k}(\mathfrak{g}\mathfrak{l}_{n}) =\pi _{\kappa } \otimes C^{k}(\mathfrak{s}\mathfrak{l}_{n})\). As easily seen, \(\hat{Q}_{(0)}I = 0\). Hence \(H^{\bullet }(C^{\kappa }(\mathfrak{g}\mathfrak{l}_{n})) =\pi _{\kappa } \otimes H^{\bullet }(C^{k}(\mathfrak{s}\mathfrak{l}_{n}))\), so that

$$\displaystyle{ \mathcal{W}^{\kappa }(\mathfrak{g}\mathfrak{l}_{n}) =\mathcal{ W}^{k}(\mathfrak{s}\mathfrak{l}_{ n}) \otimes \pi _{\kappa }. }$$

In particular if we choose the form κ to be 0, we find that π κ belongs to the center of \(\mathcal{W}^{\kappa }(\mathfrak{g}\mathfrak{l}_{n})\) as π κ belongs to the center of \(C^{\kappa }(\mathfrak{g}\mathfrak{l}_{n})\). Thus, \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\) is isomorphic to the quotient of \(\mathcal{W}^{k\kappa _{0}}(\mathfrak{g}\mathfrak{l}_{n})\) by the ideal generated by I (−1) | 0〉.

5.4 The Grading of \(\mathcal{W}^{\kappa }(\mathfrak{g})\)

The standard conformal grading of \(C^{\kappa }(\mathfrak{g})\) is given by the Hamiltonian H defined by

$$\displaystyle\begin{array}{rcl} & & \quad H\vert 0\rangle = 0,\quad [H,x_{(n)}] = -nx_{(n)}\quad (x \in \mathfrak{g}), {}\\ & & [H,\psi _{\alpha,n}] = -n\psi _{\alpha,n},\quad [H,\psi _{\alpha,n}^{{\ast}}] = -n\psi _{\alpha,n}^{{\ast}}. {}\\ \end{array}$$

However H is not well-defined in \(\mathcal{W}^{\kappa }(\mathfrak{g})\) since H does not commute with the action of

$$\displaystyle{ \hat{Q}_{(0)} =\sum _{\alpha \in \varDelta _{+}}\sum _{k\in \mathbb{Z}}(x_{\alpha })_{(-k)}\psi _{\alpha,k} +\sum _{\alpha \in \varDelta _{+}}\chi (x_{\alpha })\psi _{\alpha,1} -\sum _{\alpha,\beta,\gamma \in \varDelta _{+}}\sum _{k+l+m=0}c_{\alpha,\beta }^{\gamma }\psi _{ \alpha,k}^{{\ast}}\psi _{ \beta,l}^{{\ast}}\psi _{ \gamma,m}. }$$

Here and below we omit the tensor product sign.

To remedy this, define the linear operator \(H_{\mathcal{W}}\) by

$$\displaystyle\begin{array}{rcl} \begin{array}{rr} & H_{\mathcal{W}}\vert 0\rangle = 0,\quad [H_{\mathcal{W}},(x_{i})_{(n)}] = -n(x_{i})_{(n)}\quad (i \in I), \\ & \quad [H_{\mathcal{W}},(x_{\alpha })_{(n)}] = (\alpha (\rho ^{\vee }) - n)(x_{\alpha })_{(n)}\quad (\alpha \in \varDelta ), \\ &\quad [H_{\mathcal{W}},\psi _{\alpha,n}] = (\alpha (\rho ^{\vee }) - n)\psi _{\alpha,n},\quad [H_{\mathcal{W}},\psi _{\alpha,n}^{{\ast}}] = (-\alpha (\rho ^{\vee }) - n)\psi _{\alpha,n}^{{\ast}},\quad (\alpha \in \varDelta _{+}).\end{array} & & {}\\ \end{array}$$

Here ρ = 1∕2h, where h is defined in (6). Set \(C^{\kappa }(\mathfrak{g})_{\varDelta,new} =\{ v \in C^{\kappa }(\mathfrak{g})\mid H_{\mathcal{W}}c =\varDelta c\}\). Then

$$\displaystyle{ C^{\kappa }(\mathfrak{g}) =\bigoplus _{\varDelta \in \mathbb{Z}}C^{\kappa }(\mathfrak{g})_{\varDelta,new}. }$$
(46)

Since \([\hat{Q},H_{\mathcal{W}}] = 0\), \(C^{\kappa }(\mathfrak{g})_{\varDelta,new}\) is a subcomplex of \(C^{\kappa }(\mathfrak{g})\). We have

$$\displaystyle\begin{array}{rcl} H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})& =& \bigoplus _{\varDelta \in \mathbb{Z}}H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})_{\varDelta }, {}\\ \quad H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})_{\varDelta }& =& H^{\bullet }(C^{\kappa }(\mathfrak{g})_{\varDelta,new},\hat{Q}_{(0)}). {}\\ \end{array}$$

In particular \(\mathcal{W}^{\kappa }(\mathfrak{g}) =\bigoplus _{\varDelta \in \mathbb{Z}}\mathcal{W}^{\kappa }(\mathfrak{g})_{\varDelta }\). Note that the grading (46) is not bounded from below.

If k ≠ − n then the action of \(H_{\mathcal{W}}\) on the vertex subalgebra \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\) of \(\mathcal{W}^{k}(\mathfrak{g})\) is inner: Set

where L sug (z) is the Sugawara field of \(V ^{k}(\mathfrak{s}\mathfrak{l}_{n})\):

$$\displaystyle{ L_{sug}(z) = \frac{1} {2(k + n)}\sum _{a}: x_{a}(z)x^{a}(z):, }$$

and

Here {x a } is a basis of \(\mathfrak{s}\mathfrak{l}_{n}\) and {x a} is the dual basis of {x a } with respect to ( |  ). Then \(\hat{Q}_{(0)}L = 0\), and so L defines an element of \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\). It is a conformal vector of \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\), that is to say, \(L_{0} = H_{\mathcal{W}}\) and L −1 = T and

$$\displaystyle{ [L_{m},L_{n}] = (m - n)L_{m+n} + \frac{m^{3} - m} {12} \delta _{m,n}c, }$$

where \(c \in \mathbb{C}\) is the central charge of L, which is in this case given by

$$\displaystyle{ (n - 1)(1 - n(n + 1)(n + k - 1)^{2}/(n + k)). }$$

5.5 Decomposition of BRST Complex

We extend the map in Sect. 2.6 to the linear map \(\widehat{\theta }_{0}: \mathfrak{g}[t,t^{-1}] \rightarrow C^{\kappa }(\mathfrak{g})\) by setting

$$\displaystyle{ \widehat{\theta }_{0}(x_{a}(z)) = x_{a}(z) +\sum _{\beta,\gamma \in \varDelta _{+}}c_{a,\beta }^{\gamma }:\psi _{\gamma }(z)\psi _{\beta }^{{\ast}}(z):. }$$

Proposition 9

  1. 1.

    The correspondence

    $$\displaystyle{ x_{a}(z)\mapsto J_{a}(z):=\widehat{\theta } _{0}(x_{a}(z))\quad (x_{a} \in \mathfrak{b}_{-}) }$$

    defines a vertex algebra embedding \(V ^{\kappa _{\mathfrak{b}}}(\mathfrak{b})\hookrightarrow C^{\kappa }(\mathfrak{g})\) , where \(\kappa _{\mathfrak{b}}\) is the bilinear form on \(\mathfrak{b}\) defined by \(\kappa _{\mathfrak{b}}(x,y) =\kappa (x,y) + \frac{1} {2}\kappa _{\mathfrak{g}}(x,y)\) . We have

    $$\displaystyle{ [J_{a}{}_{\lambda }\psi _{\alpha }^{{\ast}}] =\sum _{\beta \in \varDelta _{+}}c_{a,\beta }^{\alpha }\psi _{ \beta }^{{\ast}}. }$$
  2. 2.

    The correspondence

    $$\displaystyle{ x_{\alpha }(z)\mapsto J_{\alpha }(z):=\widehat{\theta } _{0}(x_{\alpha })\quad (x_{\alpha } \in \mathfrak{n}) }$$

    defines a vertex algebra embedding \(V (\mathfrak{n})\hookrightarrow C^{\kappa }(\mathfrak{g})\) . We have

    $$\displaystyle{ [J_{\alpha }{}_{\lambda }\psi _{\beta }] =\sum _{\beta \in \varDelta _{+}}c_{\alpha,\beta }^{\gamma }\psi _{ \gamma }^{{\ast}}. }$$

Let \(C^{\kappa }(\mathfrak{g})_{+}\) denote the subalgebra of \(C^{\kappa }(\mathfrak{g})\) generated by J α (z) and ψ α (z) with αΔ +, and let \(C^{\kappa }(\mathfrak{g})_{-}\) denote the subalgebra generated by J a (z) and ψ α (z) with aΔ I, αΔ +.

The proof of the following assertions are parallel to that of Lemmas 11, 12 and Proposition 3.

Lemma 24

The multiplication map gives a linear isomorphism

$$\displaystyle{ C^{\kappa }(\mathfrak{g})_{-}\otimes C^{\kappa }(\mathfrak{g})_{+}\;\stackrel{_{\sim }}{\rightarrow }\;C^{\kappa }(\mathfrak{g}). }$$

Lemma 25

The subspaces \(C^{\kappa }(\mathfrak{g})_{-}\) and \(C^{\kappa }(\mathfrak{g})_{+}\) are subcomplexes of \((C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) . Hence \(C^{\kappa }(\mathfrak{g})\cong C^{\kappa }(\mathfrak{g})_{-}\otimes C^{\kappa }(\mathfrak{g})_{+}\) as complexes.

Theorem 19 ([28, 40])

We have \(H^{i}(C^{\kappa }(\mathfrak{g})_{+},\hat{Q}_{(0)}) =\delta _{i,0}\mathbb{C}\) . Hence \(H^{\bullet }(C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)}) = H^{\bullet }(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\) . In particular \(\mathcal{W}^{\kappa }(\mathfrak{g}) = H^{0}(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\).

Since the complex \(C^{\kappa }(\mathfrak{g})_{-}\) has no positive cohomological degree, its zeroth cohomology \(\mathcal{W}^{k}(\mathfrak{g}) = H^{0}(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\) is a vertex subalgebra of \(C^{\kappa }(\mathfrak{g})_{-}\). Observe also that \(C^{\kappa }(\mathfrak{g})_{-}\) has no negative degree with respect to the Hamiltonian \(H_{\mathcal{W}}\), and each homogeneous space is finite-dimensional:

$$\displaystyle{ C^{\kappa }(\mathfrak{g})_{-} =\bigoplus _{\varDelta \in \mathbb{Z}_{-}}C^{\kappa }(\mathfrak{g})_{-,\varDelta,new},\quad \mathrm{dim}C^{\kappa }(\mathfrak{g})_{-,\varDelta,new} <\infty. }$$
(47)

Here \(C^{\kappa }(\mathfrak{g})_{-,\varDelta,new} = C^{\kappa }(\mathfrak{g})_{-}\cap C^{\kappa }(\mathfrak{g})_{\varDelta,new}\).

5.6 Proof of Theorem 18

As \(\hat{Q}_{(0)}F^{p}C^{\kappa }(\mathfrak{g})_{-}\subset F^{p}C^{\kappa }(\mathfrak{g})_{-}\), one can consider a spectral sequence for \(H^{\bullet }(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\) such that the E 1-term is \(H^{\bullet }(\mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\). This spectral sequence clearly converges, since \(C^{\kappa }(\mathfrak{g})_{-}\) is a direct sum of finite-dimensional subcomplexes \(C^{\kappa }(\mathfrak{g})_{-,\varDelta,new}\).

We have \(\mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g})_{-}\cong S(\mathfrak{b}_{-}[t^{-1}]t^{-1}) \otimes \varLambda (\mathfrak{n}[t^{-1}]t^{-1})\cong \mathbb{C}[J\mu ^{-1}(\chi )] \otimes \varLambda (\mathfrak{n}[t^{-1}]t^{-1})\), and the complex \((\mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\) is identical to the Chevalley complex for the Lie algebra cohomology \(H^{\bullet }(\mathfrak{n}[t], \mathbb{C}[J\mu ^{-1}(\chi )])\). Therefore

(48)

Thus the spectral sequence collapses at E 1 = E , and we get

Here \(\mathop{\mathrm{gr}}\nolimits ^{G}H^{i}(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\) is the associated graded space with respect to the filtration \(G^{\bullet }H^{i}(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\) induced by the filtration \(F^{\bullet }C^{\kappa }(\mathfrak{g})_{-}\), that is,

$$\displaystyle{ G^{p}H^{i}(C^{\kappa }(\mathfrak{g})_{ -},\hat{Q}_{(0)}) =\mathop{ \mathrm{Im}}\nolimits (H^{i}(F^{p}C^{\kappa }(\mathfrak{g})_{ -},\hat{Q}_{(0)}) \rightarrow H^{i}(C^{\kappa }(\mathfrak{g})_{ -},\hat{Q}_{(0)})). }$$

We claim that the filtration \(G^{\bullet }H^{0}(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)})\) coincides with the canonical filtration of \(H^{0}(C^{\kappa }(\mathfrak{g})_{-},\hat{Q}_{(0)}) =\mathcal{ W}^{\kappa }(\mathfrak{g})\). Indeed, from the definition of the canonical filtration we have \(F^{p}W^{k}(\mathfrak{g}) \subset G^{p}\mathcal{W}^{\kappa }(\mathfrak{g})\) for all p, and hence, there is a Poisson vertex algebra homomorphism

(49)

that restricts to a surjective homomorphism

Since is generated by as differential algebras it follows that (49) is surjective. On the other hand the cohomology vanishing and the Euler-Poincaré principle imply that the graded character of \(\mathcal{W}^{\kappa }(\mathfrak{g})\) and are the same. Therefore (49) is an isomorphism, and thus, \(G^{p}\mathcal{W}^{\kappa }(\mathfrak{g}) = F^{p}\mathcal{W}^{\kappa }(\mathfrak{g})\) for all p.

Finally the embedding \(\mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g})_{-}\rightarrow \mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g})\) induces an isomorphism

$$\displaystyle{ H^{0}(\mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g})_{ -},\hat{Q}_{(0)})\cong H^{0}(\mathop{\mathrm{gr}}\nolimits C^{\kappa }(\mathfrak{g}),\hat{Q}_{ (0)}) }$$

by Theorem 17 and (48). This completes the proof. □

5.7 Zhu’s Algebra of W-Algebra

Let \(\mathrm{Zhu}_{new}(C^{\kappa }(\mathfrak{g}))\) be Zhu’s algebra of \(C^{\kappa }(\mathfrak{g})\) with respect to the Hamiltonian H W , \(\mathrm{Zhu}_{old}(C^{\kappa }(\mathfrak{g}))\) Zhu’s algebra of \(C^{\kappa }(\mathfrak{g})\) with respect to the standard Hamiltonian H. We have

$$\displaystyle{ \mathrm{Zhu}_{new}(C^{\kappa }(\mathfrak{g}))\cong \mathrm{Zhu}_{old}(C^{\kappa }(\mathfrak{g}))\cong C(\mathfrak{g}), }$$

see [10, Proposition 5.1] for the details. Then it is legitimate to write \(\mathrm{Zhu}(C^{\kappa }(\mathfrak{g}))\) for \(\mathrm{Zhu}_{new}(C^{\kappa }(\mathfrak{g}))\) or \(\mathrm{Zhu}_{old}(C^{\kappa }(\mathfrak{g}))\).

By the commutation formula, we have

$$\displaystyle{ \hat{Q}_{(0)}(C^{\kappa }(\mathfrak{g}) \circ C^{\kappa }(\mathfrak{g})) \subset C^{\kappa }(\mathfrak{g}) \circ C^{\kappa }(\mathfrak{g}). }$$

Here the circle ∘ is defined as in the definition of the Zhu algebra (with respect to the grading \(H_{\mathcal{W}}\)). So \((\mathrm{Zhu}_{new}C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) is a differential, graded algebra, which is identical to \((C(\mathfrak{g}),\mathrm{ad}Q)\).

Theorem 20 ([5])

We have

Proof

By Theorem 18, it follows that \(\mathcal{W}^{\kappa }(\mathfrak{g})\) admits a PBW basis. Hence \(\eta _{\mathcal{W}^{\kappa }(\mathfrak{g})}:\mathop{ \mathrm{gr}}\nolimits \mathrm{Zhu}\mathcal{W}^{\kappa }(\mathfrak{g}) \rightarrow R_{\mathcal{W}_{k}(\mathfrak{g})}\) is an isomorphism by Theorem 16. On the other hand we have a natural algebra homomorphism \(\mathrm{Zhu}\mathcal{W}^{\kappa }(\mathfrak{g})\longrightarrow H^{0}(\mathrm{Zhu}C^{\kappa }(\mathfrak{g}),\hat{Q}_{(0)})\) which makes the following diagram commute.

Note that we have the isomorphisms and in the diagram. Now the other three isomorphisms will give the desired isomorphism.

We conclude that we have the following commutative diagram:

Remark 10

The same proof applies for an arbitrary simple Lie algebra \(\mathfrak{g}\). In particular, we have . In fact the same proof applies for the W-algebra associated with a simple Lie algebra \(\mathfrak{g}\) and an arbitrary nilpotent element f of \(\mathfrak{g}\) to show its Zhu’s algebra is isomorphic to the finite W-algebra \(U(\mathfrak{g},f)\) [29].

5.8 Explicit Generators

It is possible to write down the explicit generators of \(\mathcal{W}^{\kappa }(\mathfrak{g}) \subset C^{\kappa }(\mathfrak{g})_{-}\).

Recall that the column-determinant of a matrix A = (a ij ) over an associative algebra is defined by

$$\displaystyle{ \mathop{\mathrm{cdet}}\nolimits A =\sum _{\sigma \in \mathfrak{S}_{n}}\mathop{ \mathrm{sgn}}\nolimits \sigma \cdot a_{\sigma (1)1}^{}a_{\sigma (2)2}^{}\ldots \,a_{\sigma (n)\,n}^{}. }$$

Introduce an extended Lie algebra \(\mathfrak{b}[t^{-1}]t^{-1} \oplus \mathbb{C}\tau\), where the element τ commutes with 1, and

$$\displaystyle{ \big[\tau,x_{(-n)}\big] = nx_{(-n)}\qquad \text{for}\quad x \in \mathfrak{b},n \in \mathfrak{n}, }$$

where x (−n) = xt n. This induces an associative algebra structure on the tensor product space \(U\big(\mathfrak{b}[t^{-1}]t^{-1}\big)\mathop{\oplus }\mathbb{C}[\tau ]\).

Consider the matrix

$$\displaystyle{ B = \left [\begin{array}{*{10}c} \alpha \quad \tau + (e_{11})_{(-1)} & -1\phantom{-} & 0 & \ldots & 0 \\ [0.4em](e_{2\quad 1})_{(-1)} & \alpha \quad \tau + (e_{2\quad 2})_{(-1)} & -1\phantom{-}& \ldots & 0 \\ [0.4em]\vdots & \vdots & \ddots & & \vdots \\ [0.4em](e_{n-1\quad 1})_{(-1)} & (e_{n-1\quad 2})_{(-1)} & \ldots & \alpha \quad \tau + (e_{n-1\,n-1})_{(-1)} & -1\phantom{-} \\ [0.4em](e_{n\quad 1})_{(-1)} & (e_{n\quad 2})_{(-1)} & \ldots & \ldots & \alpha \quad \tau + (e_{n\quad n})_{(-1)}. \end{array} \right ] }$$

with entries in \(U(\mathfrak{b}[t^{-1}]t^{-1}]) \otimes \mathbb{C}[\tau ] \otimes \mathbb{C}[\alpha ]\), where α is a parameter.

For its column-determinantFootnote 1 we can write

$$\displaystyle{ \mathop{\mathrm{cdet}}\nolimits B =\tau ^{\quad n} + W_{\alpha }^{(1)}\tau ^{\quad n-1} +\ldots +W_{\alpha }^{(n)} }$$

for certain coefficients W α (r) which are elements of \(U(\mathfrak{b}[t^{-1}]t^{-1}]) \otimes \mathbb{C}[\alpha ]\). Set

$$\displaystyle{ W^{(i)} = W_{\alpha }^{(i)}\vert _{\alpha =k+n-1}. }$$

This is an element of \(U(\mathfrak{b}[t^{-1}]t^{-1})\), which we identify with \(V ^{\kappa _{\mathfrak{b}}}(\mathfrak{b}) \subset C^{\kappa }(\mathfrak{g})_{ -}\).

Theorem 21 ([14])

\(\mathcal{W}^{k}(\mathfrak{g})\) is strongly generated by W (1), , W (n).

5.9 Miura Map

The Cartan subalgebra \(\mathfrak{h}\) of \(\mathfrak{g}\) acts on \(C^{\kappa }(\mathfrak{g})_{+}\) by x i ↦ (J i )(0), iI, see Proposition 9. Let \(C^{\kappa }(\mathfrak{g})_{+}^{\lambda }\) be the weight space of weight \(\lambda \in \mathfrak{h}^{{\ast}}\) with respect to this action. Then

$$\displaystyle{ C^{\kappa }(\mathfrak{g})_{+} =\bigoplus _{\lambda \leqslant 0}C^{\kappa }(\mathfrak{g})_{+}^{\lambda },\quad C^{\kappa }(\mathfrak{g})_{ +}^{0} = V ^{\kappa _{\mathfrak{b}} }(\mathfrak{h}) \subset V ^{\kappa _{k}}(\mathfrak{b}). }$$

The vertex algebra \(V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h})\) is the Heisenberg vertex algebra associated with \(\mathfrak{h}\) and the bilinear form \(\kappa _{\mathfrak{h}}:=\kappa _{\mathfrak{b}}\vert _{\mathfrak{h}\times \mathfrak{h}}\).

The projection \(C^{\kappa }(\mathfrak{g})_{+} \rightarrow C^{\kappa }(\mathfrak{g})_{+}^{0} = V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h})\) with respect to this decomposition is a vertex algebra homomorphism. Therefore it restriction

$$\displaystyle{ \hat{\varUpsilon }:\mathcal{ W}^{\kappa }(\mathfrak{g}) \rightarrow V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h}) }$$
(50)

is also a vertex algebra homomorphism that is called the Miura map.

Theorem 22

The Miura map is injective for all \(k \in \mathbb{C}\).

Proof

The induced Poisson vertex algebra homomorphism

(51)

is just a restriction map and coincides with \(J\bar{\varUpsilon }\), where \(\bar{\varUpsilon }\) is defined in (13). Clearly, it is sufficient to show that \(J\bar{\varUpsilon }\) is injective.

Recall that the action map gives an isomorphism

$$\displaystyle{ N \times (\,f + \mathfrak{h}_{\mathop{\mathrm{reg}}\nolimits })\;\stackrel{_{\sim }}{\rightarrow }\;U \subset f + \mathfrak{b}, }$$

where U is some open subset of \(f + \mathfrak{b}\), see the proof of Proposition 4. Therefore, by Lemma 15, the action map \(JN \times J(\,f + \mathfrak{h}) \rightarrow J(\,f + \mathfrak{b})\) is dominant. Thus, the induced map \(\mathbb{C}[J(\,f + \mathfrak{b})] \rightarrow \mathbb{C}[JN \times J(\,f + \mathfrak{h})]\) is injective, and so is \(J\bar{\varUpsilon }: C[J(\,f + \mathfrak{b})]^{JN} \rightarrow \mathbb{C}[JN \times J(\,f + \mathfrak{h})]^{JN} = \mathbb{C}[J(\,f + \mathfrak{h})]\).

Remark 11

It is straightforward to generalize Theorem 22 for the W-algebra \(\mathcal{W}^{k}(\mathfrak{g})\) associated with a general simple Lie algebra \(\mathfrak{g}\).

Theorem 23

Let \(x_{i} = E_{ii} \in \mathfrak{h} \subset \mathfrak{g} = \mathfrak{g}\mathfrak{l}_{n}\) , and J i (z) the corresponding field of \(V ^{\kappa _{k}}(\mathfrak{h})\) . The image Υ(W (i)(z)) of W (i)(z) by the Miura map is described by

$$\displaystyle{ \sum _{i=0}^{n}\varUpsilon (W^{(i)})(z)(\alpha \partial _{ z})^{n-i} =: (\alpha \partial _{ z} + J_{1}(z))(\alpha \partial _{z} + J_{2}(z))\ldots (\alpha \partial _{z} + J_{N}(z)):, }$$

where α = k + n − 1, W (0)(z) = 1, \([\partial _{z},J_{i}(z)] = \frac{d} {dz}J_{i}(z)\).

Proof

It is straightforward from Theorem 21.

Note that if we choose κ to be 0 and set i = 1 N J i (z) = 0, we obtain the image of the generators of \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\) by the Miura map \(\hat{\varUpsilon }\). For k + n ≠ 0, this expression can be written in more symmetric manner: Set \(b_{i}(z) = \frac{1} {\sqrt{k+n}}J_{i}(z)\), so that i = 1 n b i (z) = 0, and

$$\displaystyle{ [(b_{i})_{\lambda }b_{j}] = \left \{\begin{array}{@{}l@{\quad }l@{}} (1 - \frac{1} {n})\lambda \quad &\text{if }i = j, \\ -\frac{1} {n}\lambda \quad &\text{if }i\neq j. \end{array} \right. }$$

Then we obtain the following original description of the \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\) due to Fateev and Lukyanov [34].

Corollary 4

Suppose that k + n ≠ 0. Then the image of \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\) by the Miura map is the vertex subalgebra generated by fields \(\tilde{W}_{2}(z)\ldots,\tilde{W}_{n}(z)\) defined by

$$\displaystyle{ \sum _{i=0}^{n}\tilde{W}_{ i}(z)(\alpha _{0}\partial _{z})^{n-i} =: (\alpha _{ 0}\partial _{z} + b_{1}(z))(\alpha _{0}\partial _{z} + b_{2}(z))\ldots (\alpha _{0}\partial _{z} + b_{n}(z)):, }$$

where α 0 = α + + α , \(\alpha _{+} = \sqrt{k + n}\) , \(\alpha _{-} = -1/\sqrt{k + n}\) , \(\tilde{W}_{0}(z) = 1\) , \(\tilde{W}_{1}(z) = 0\).

Corollary 5

Suppose that k + n ≠ 0. We have

$$\displaystyle{ \mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{ n})\cong \mathcal{W}^{^{L} k}(\mathfrak{s}\mathfrak{l}_{ n}), }$$

where L k is defined by (k + n)(L k + n) = 1.

Example 4

Let \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\), k ≠ − 2. Set \(b(z) = \sqrt{2}b_{1}(z) = -\sqrt{2}b_{2}(z)\), so that [b λ b] = λ. Then the right-hand-side of the formula in Corollary 5 becomes

$$\displaystyle\begin{array}{rcl} & &: (\alpha _{0}\partial _{z} + \frac{1} {\sqrt{2}}b(z))(\alpha _{0}\partial _{z} - \frac{1} {\sqrt{2}}b(z)): {}\\ & & =\alpha _{ 0}^{2}\partial _{ z}^{2} - L(z), {}\\ \end{array}$$

where

$$\displaystyle{ L(z) = \frac{1} {2}: b(z)^{2}: + \frac{\alpha _{0}} {\sqrt{2}}\partial _{z}b(z). }$$

It is well-known and is straightforward to check that the field generates the Virasoro algebra of central charge 1 − 6(k + 1)2∕(k + 2). Thus \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{2})\), k ≠ − 2, is isomorphic to the universal Virasoro vertex algebra of central charge 1 − 6(k + 1)2∕(k + 2).

In the case that \(\kappa =\kappa _{c}:= -\frac{1} {2}\kappa _{\mathfrak{g}}\), then it follows from Theorem 22 that \(\mathcal{W}^{\kappa _{c}}(\mathfrak{g}\mathfrak{l}_{n})\) is commutative since \(V ^{(\kappa _{c})_{\mathfrak{h}}}(\mathfrak{h})\) is commutative. In fact the following fact is known: Let \(Z(V ^{\kappa }(\mathfrak{g})) =\{ z \in V ^{\kappa }(\mathfrak{g})\mid [z_{(n)},a_{(n)}] = 0\}\), the center of \(V ^{\kappa }(\mathfrak{g})\).

Theorem 24 ([37])

We have the isomorphism

$$\displaystyle{ Z(V ^{\kappa _{c}}(\mathfrak{g}))\;\stackrel{_{\sim }}{\rightarrow }\;\mathcal{W}^{\kappa _{c}}(\mathfrak{g}),\quad z\mapsto [z \otimes 1]. }$$

This is a chiralization of Kostant’s Theorem 5 in the sense that we recover Theorem 5 from Theorem 24 by considering the induced map between Zhu’s algebras of both sides. The statement of Theorem 24 holds for any simple Lie algebra \(\mathfrak{g}\) [37].

Remark 12

For a general simple Lie algebra \(\mathfrak{g}\), the image of the Miura map for a generic k is described in terms of screening operators, see [40, 15.4]. Theorem 23 for \(\mathfrak{g} = \mathfrak{g}\mathfrak{l}_{n}\) also follows from this description (the proof reduces to the case \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\)). An important application of this realization is the Feigin-Frenkel duality which states

$$\displaystyle{ \mathcal{W}^{k}(\mathfrak{g})\cong \mathcal{W}^{^{L} k}(^{L}\mathfrak{g}), }$$

where \(^{L}\mathfrak{g}\) is the Langlands dual Lie algebra of \(\mathfrak{g}\), r (k + h )(L k +L h ) = 1. Here r is the maximal number of the edges of the Dynking diagram of \(\mathfrak{g}\) andL h is the dual Coxeter number of \(^{L}\mathfrak{g}\). In [37, 40] this isomorphism was stated only for a generic k, but it is not too difficult to see the isomorphism remains valid for an arbitrary k using the injectivity of the Miura map.

The Miura map is defined [60] for the W-algebra \(\mathcal{W}^{k}(\mathfrak{g},f)\) associated with an arbitrary f, which is injective as well since the proof of Theorem 22 applies. Recently Naoki Genra [46] has obtained the description of the image by the Miura map in terms of screening operators for the W-algebra \(\mathcal{W}^{k}(\mathfrak{g},f)\) associated with an arbitrary nilpotent element f.

5.10 Classical W-Algebras

Since the Poisson structure of is trivial, we can give \(\mathop{\mathrm{gr}}\nolimits \mathcal{W}^{\kappa }(\mathfrak{g})\) a Poisson vertex algebra structure by the formula (32). The Poisson structure of \(R_{V ^{k+n}(\mathfrak{h})} = \mathbb{C}[\mathfrak{h}]\) is also trivial, hence \(\mathop{\mathrm{gr}}\nolimits V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h}) = \mathbb{C}[J\mathfrak{h}^{{\ast}}]\) is equipped with the Poisson vertex algebra structure by the formula (32) as well. Then the map \(\mathop{\mathrm{gr}}\nolimits \hat{\varUpsilon }:\mathop{ \mathrm{gr}}\nolimits \mathcal{W}^{\kappa }(\mathfrak{g})\hookrightarrow \mathop{\mathrm{gr}}\nolimits V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h})\) is a homomorphism of Poisson vertex algebras with respect to these structures. Set κ = 0, \(k \in \mathbb{C}\), and consider its restriction \(\mathop{\mathrm{gr}}\nolimits \hat{\varUpsilon }:\mathop{ \mathrm{gr}}\nolimits \mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\hookrightarrow \mathop{\mathrm{gr}}\nolimits V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h}')\), where \(\mathfrak{h}'\) is the Cartan subalgebra of \(\mathfrak{s}\mathfrak{l}_{n}\).

In \(\mathop{\mathrm{gr}}\nolimits V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h}')\) we have

$$\displaystyle{ \{h_{\lambda }h'\} =\kappa _{\mathfrak{h}}(h,h') = (k + n)\kappa _{0}(h,h'), }$$

and this uniquely determines the λ-bracket of \(\mathop{\mathrm{gr}}\nolimits V ^{\kappa _{\mathfrak{h}}}(\mathfrak{h}')\). Hence it is independent of k provided that k ≠ − n. Since the image of \(\mathop{\mathrm{gr}}\nolimits \mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\) is strongly generated by elements of \(\mathbb{C}[(\mathfrak{h}')^{{\ast}}]^{W}\), it follows that the Poisson vertex algebra structure of \(\mathop{\mathrm{gr}}\nolimits \mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\), k ≠ − n, is independent of k. We denote this Poisson vertex algebra by \(\mathcal{W}^{cl}(\mathfrak{s}\mathfrak{l}_{n})\).

The Poisson vertex algebra \(\mathcal{W}^{cl}(\mathfrak{s}\mathfrak{l}_{n})\) is called the classical W-algebra associated with \(\mathfrak{s}\mathfrak{l}_{n}\), which appeared in the works of Adler [3], Gelfand-Dickey [45] and Drinfeld-Sokolov [32]. Thus, the W-algebra \(\mathcal{W}^{k}(\mathfrak{s}\mathfrak{l}_{n})\), k ≠ − n, is a deformation of \(\mathcal{W}^{cl}(\mathfrak{s}\mathfrak{l}_{n})\).

On the other hand the \(\mathcal{W}\)-algebra \(\mathcal{W}^{-n}(\mathfrak{s}\mathfrak{l}_{n})\) at the critical level can be identified with the space of the \(\mathfrak{s}\mathfrak{l}_{n}\) -opers [21] on the disk D. We refer to [39, 40] for more on this subject.

6 Representations of W-Algebras

From now on we set \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{n}\) and study the representations of \(\mathcal{W}^{k}(\mathfrak{g})\) [see (45)].

6.1 Poisson Modules

Let R be a Poisson algebra. Recall that a Poisson R-module is a R-module M in the usual associative sense equipped with a bilinear map

$$\displaystyle{ R \times M \rightarrow M,\quad (r,m)\mapsto \mathrm{ad}r(m) =\{ r,m\}, }$$

which makes M a Lie algebra module over R satisfying

$$\displaystyle{ \{r_{1},r_{2}m\} =\{ r_{1},r_{2}\}m + r_{2}\{r_{1},m\},\quad \{r_{1}r_{2},m\} = r_{1}\{r_{2},m\} + r_{2}\{r_{1},m\} }$$

for r 1, r 2R, mM. Let \(R\mathop{\mathrm{-PMod}}\nolimits\) be the category of Poisson modules over R.

Lemma 26

A Poisson module over \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\) is the same as a \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\) -module M in the usual associative sense equipped with a Lie algebra module structure \(\mathfrak{g} \rightarrow \mathop{\mathrm{End}}\nolimits M\) , x ↦ ad(x), such that

$$\displaystyle{ \mathrm{ad}(x)(\,fm) =\{ x,f\}.m + f.\mathrm{ad}(x)(m) }$$

for \(x \in \mathfrak{g}\) , \(f \in \mathbb{C}[\mathfrak{g}^{{\ast}}]\) , mM.

6.2 Poisson Vertex Modules

Definition 5

A Poisson vertex module over a Poisson vertex algebra V is a V -module M as a vertex algebra equipped with a linear map

$$\displaystyle{ V \mapsto (\mathop{\mathrm{End}}\nolimits M)[[z^{-1}]]z^{-1},\quad a\mapsto Y _{ -}^{M}(a,z) =\sum _{ n\geqslant 0}a_{(n)}^{M}z^{-n-1}, }$$

satisfying

$$\displaystyle{ a_{(n)}^{M}m = 0\quad \text{for }n \gg 0, }$$
(52)
$$\displaystyle{ (Ta)_{(n)}^{M} = -na_{ (n-1)}^{M}, }$$
(53)
$$\displaystyle{ a_{(n)}^{M}(bv) = (a_{ (n)}b)v + b(a_{(n)}^{M}v), }$$
(54)
$$\displaystyle{ [a_{(m)}^{M},b_{ (n)}^{M}] =\sum _{ i\geqslant 0}\left (\begin{array}{*{10}c} m\\ i \end{array} \right )(a_{(i)}b)_{(m+n-i)}^{M}, }$$
(55)
$$\displaystyle{ (ab)_{(n)}^{M} =\sum _{ i=0}^{\infty }(a_{ (-i-1)}^{M}b_{ (n+i)}^{M} + b_{ (-i-1)}^{M}a_{ (n+i)}^{M}) }$$
(56)

for all a, bV, \(m,n\geqslant 0\), vM.

A Poisson vertex algebra R is naturally a Poisson vertex module over itself.

Example 5

Let M be a Poisson vertex module over \(\mathbb{C}[J\mathfrak{g}^{{\ast}}]\). Then by (55), the assignment

$$\displaystyle{ xt^{n}\mapsto x_{ (n)}^{M}\quad x \in \mathfrak{g} \subset \mathbb{C}[\mathfrak{g}^{{\ast}}] \subset \mathbb{C}[J\mathfrak{g}^{{\ast}}],\ n\geqslant 0, }$$

defines a \(J\mathfrak{g} = \mathfrak{g}[[t]]\)-module structure on M. In fact, a Poisson vertex module over \(\mathbb{C}[J\mathfrak{g}^{{\ast}}]\) is the same as a \(\mathbb{C}[J\mathfrak{g}^{{\ast}}]\)-module M in the usual associative sense equipped with an action of the Lie algebra \(J\mathfrak{g}\) such that (xt n)m = 0 for n ≫ 0, \(x \in \mathfrak{g}\), mM, and

$$\displaystyle{ (xt^{n}) \cdot (am) = (x_{ (n)}a)m + a(xt^{n}) \cdot m }$$

for \(x \in \mathfrak{g}\), \(n\geqslant 0\), \(a \in \mathbb{C}[J\mathfrak{g}^{{\ast}}]\), mM.

Below we often write a (n) for a (n) M.

The proofs of the following assertions are straightforward.

Lemma 27

Let R be a Poisson algebra, E a Poisson module over R. There is a unique Poisson vertex JR-module structure on JR R E such that

$$\displaystyle{ a_{(n)}(b \otimes m) = (a_{(n)}b) \otimes m +\delta _{n,0}b \otimes \{ a,m\} }$$

for \(n\geqslant 0\) , aRJR, bJR. mE (Recall that \(JR = \mathbb{C}[J\mathop{\mathrm{Spec}}\nolimits R]\) .)

Lemma 28

Let R be a Poisson algebra, M a Poisson vertex module over JR. Suppose that there exists a R-submodule E of M (in the usual commutative sense) such that a (n) E = 0 for n > 0, aR, and M is generated by E (in the usual commutative sense). Then there exists a surjective homomorphism

$$\displaystyle{ JR \otimes _{R}E \twoheadrightarrow M }$$

of Poisson vertex modules.

6.3 Canonical Filtration of Modules Over Vertex Algebras

Let V be a vertex algebra graded by a Hamiltonian H. A compatible filtration of a V -module M is a decreasing filtration

$$\displaystyle{ M =\varGamma ^{0}M \supset \varGamma ^{1}M \supset \cdots }$$

such that

$$\displaystyle\begin{array}{rcl} & & a_{(n)}\varGamma ^{q}M \subset \varGamma ^{p+q-n-1}M\quad \text{for }a \in F^{p}V,\ \forall n \in \mathbb{Z}, {}\\ & & a_{(n)}\varGamma ^{q}M \subset \varGamma ^{p+q-n}M\quad \text{for }a \in F^{p}V,\ n\geqslant 0, {}\\ & & H.\varGamma ^{p}M \subset \varGamma ^{p}M\quad \text{for all }p\geqslant 0, {}\\ & & \bigcap _{p}\varGamma ^{p}M = 0. {}\\ \end{array}$$

For a compatible filtration Γ M the associated graded space

$$\displaystyle{ \mathop{\mathrm{gr}}\nolimits ^{\varGamma }M =\bigoplus _{p\geqslant 0}\varGamma ^{p}M/\varGamma ^{p+1}M }$$

is naturally a graded vertex Poison module over the graded vertex Poisson algebra \(\mathop{\mathrm{gr}}\nolimits ^{F}V\), and hence, it is a graded vertex Poison module over \(JR_{V } = \mathbb{C}[\tilde{X}_{V }]\) by Theorem 10.

The vertex Poisson JR V -module structure of \(\mathop{\mathrm{gr}}\nolimits ^{\varGamma }M\) restricts to the Poisson R V -module structure of MΓ 1 M = Γ 0 MΓ 1 M, and a (n)(MΓ 1 M) = 0 for aR V JR V , n > 0. It follows that there is a homomorphism

$$\displaystyle{ JR_{V } \otimes _{R_{V }}(M/\varGamma ^{1}M) \rightarrow \mathop{\mathrm{gr}}\nolimits ^{\varGamma }M,\quad a \otimes \bar{ m}\mapsto a\bar{m}, }$$

of vertex Poisson modules by Lemma 28.

Suppose that V is positively graded and so is a V -module M. We denote by F M the Li filtration [68] of M, which is defined by

$$\displaystyle{ F^{p}M =\mathop{ \mathrm{span}}\nolimits _{ \mathbb{C}}\{a_{(-n_{1}-1)}^{1}\ldots a_{ (-n_{r}-1)}^{r}m\mid a^{i} \in V,\ m \in M,\ n_{ 1} +\ldots +n_{r}\geqslant p\}. }$$

It is a compatible filtration of M, and in fact is the finest compatible filtration of M, that is, F p MΓ p M for all p for any compatible filtration Γ M of M. The subspace F 1 M is spanned by the vectors a (−2) m with aV, mM, which is often denoted by C 2(M) in the literature. Set

$$\displaystyle{ \bar{M} = M/F^{1}M(= M/C_{ 2}(M)), }$$
(57)

which is a Poisson module over \(R_{V } =\bar{ V }\). By [68, Proposition 4.12], the vertex Poisson module homomorphism

$$\displaystyle{ JR_{V } \otimes _{R_{V }}\bar{M} \rightarrow \mathop{\mathrm{gr}}\nolimits ^{F}M }$$

is surjective.

Let {a i; iI} be elements of V such that their images generate R V in usual commutative sense, and let U be a subspace of M such that M = U + F 1 M. The surjectivity of the above map is equivalent to that

$$\displaystyle\begin{array}{rcl} & & \qquad F^{p}M \\ & & =\mathop{ \mathrm{span}}\nolimits _{\mathbb{C}}\quad \{a_{(-n_{1}-1)}^{i_{1} }\ldots a_{(-n_{r}-1)}^{i_{r} }m\mid m \in U,\ n_{i}\geqslant 0,n_{1} +\ldots +n_{r}\geqslant p,\ i_{1},\ldots,i_{r} \in I\}.{}\end{array}$$
(58)

Lemma 29

Let V be a vertex algebra, M a V -module. The Poisson vertex algebra module structure of \(\mathop{\mathrm{gr}}\nolimits ^{F}M\) restricts to the Poisson module structure of \(\bar{M}:= M/F^{1}M\) over R V , that is, \(\bar{M}\) is a Poisson R V -module by

$$\displaystyle{ \bar{a} \cdot \bar{ m} = \overline{a_{(-1)}m},\quad \mathrm{ad}(\bar{a})(\bar{m}) = \overline{a_{(0)}m}. }$$

A V -module M is called finitely strongly generated if \(\bar{M}\) is finitely generated as a R V -module in the usual associative sense.

6.4 Associated Varieties of Modules Over Affine Vertex Algebras

A \(\widehat{\mathfrak{g}}\)-module M of level k is called smooth if x(z) is a field on M for \(x \in \mathfrak{g}\), that is, xt n m = 0 for n ≫ 0, \(x \in \mathfrak{g}\), mM. Any \(V ^{k}(\mathfrak{g})\)-module M is naturally a smooth \(\widehat{\mathfrak{g}}\)-module of level k. Conversely, any smooth \(\widehat{\mathfrak{g}}\)-module of level k can be regarded as a \(V ^{k}(\mathfrak{g})\)-module. It follows that a \(V ^{k}(\mathfrak{g})\)-module is the same as a smooth \(\widehat{\mathfrak{g}}\)-module of level k.

For a \(V = V ^{k}(\mathfrak{g})\)-module M, or equivalently, a smooth \(\widehat{\mathfrak{g}}\)-module of level k, we have

$$\displaystyle{ \bar{M} = M/\mathfrak{g}[t^{-1}]t^{-2}M, }$$

and the Poisson \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\)-module structure is given by

$$\displaystyle{ x \cdot \bar{ m} = \overline{xt^{-1}m},\quad \mathrm{ad}(x)\bar{m} = \overline{xm}. }$$

For a \(\mathfrak{g}\)-module E let

$$\displaystyle{ V _{E}^{k}:= U(\widehat{\mathfrak{g}}) \otimes _{ U(\mathfrak{g}[t]\mathop{\oplus }\mathbb{C}K)}E, }$$

where E is considered as a \(\mathfrak{g}[t]\mathop{\oplus }\mathbb{C}K\)-module on which \(\mathfrak{g}[t]\) acts via the projection \(\mathfrak{g}[t] \rightarrow \mathfrak{g}\) and K acts as multiplication by k. Then

$$\displaystyle{ \overline{V _{E}^{k}}\cong \mathbb{C}[\mathfrak{g}] \otimes E, }$$

where the Poisson \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\)-module structure is given by

$$\displaystyle{ f \cdot g \otimes v = (\,fg) \otimes v,\quad \mathrm{ad}x(\,f \otimes v) =\{ x,f\} \otimes v + f \otimes xv, }$$

for \(f,g \in \mathbb{C}[\mathfrak{g}^{{\ast}}]\), vV.

Let be the category of \(\widehat{\mathfrak{g}}\) of level k [53], KL k the full subcategory of consisting of modules M which are integrable over \(\mathfrak{g}\). Note that V E k is a object of KL k for a finite-dimensional representation E of \(\mathfrak{g}\). Thus, \(V ^{k}(\mathfrak{g}) = V _{\mathbb{C}}^{k}\) and its simple quotient \(V _{k}(\mathfrak{g})\) are also objects of KL k .

Both and KL k can be regarded as full subcategories of the category of \(V ^{k}(\mathfrak{g})\)-modules.

Lemma 30

For MKL k the following conditions are equivalent.

  1. 1.

    M is finitely strongly generated as a \(V ^{k}(\mathfrak{g})\) -module,

  2. 2.

    M is finitely generated as a \(\mathfrak{g}[t^{-1}]t^{-1}\) -module,

  3. 3.

    M is finitely generated as a \(\widehat{\mathfrak{g}}\) -module.

For a finitely strongly generated \(V ^{k}(\mathfrak{g})\)-module M define its associated variety X M by

$$\displaystyle{ X_{M} =\mathop{ \mathrm{supp}}\nolimits _{R_{V }}(\bar{M}) \subset X_{V }, }$$

equipped with a reduced scheme structure.

Example 6

\(X_{V _{E}^{k}} = \mathfrak{g}^{{\ast}}\) for a finite-dimensional representation E of \(\mathfrak{g}\).

6.5 Ginzburg’s Correspondence

Let be the full subcategory of the category of Poisson \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\)-modules on which the Lie algebra \(\mathfrak{g}\)-action (see Lemma 26) is integrable.

Lemma 31

For MKL k , the Poisson \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\) -module \(\bar{M}\) belongs to .

By Lemma 31 we have a right exact functor

For , \(M \otimes \overline{Cl}\) is naturally a Poisson module over \(\bar{C}(\mathfrak{g}) = \mathbb{C}[\mathfrak{g}^{{\ast}}] \otimes \overline{Cl}\). (The notation of Poisson modules natural extends to the Poisson supralgebras.) Thus, \((M \otimes \overline{Cl},\mathrm{ad}\bar{Q})\) is a differential graded Poisson module over the differential graded Poisson module \((\bar{C}(\mathfrak{g}),\mathrm{ad}\bar{Q})\). In particular it cohomology \(H^{\bullet }(\bar{M} \otimes \overline{Cl},\mathrm{ad}\bar{Q})\) is a Poisson module over . So we get a functor

The following assertion is a restatement of a result of Ginzburg [47] (see [10, Theorem 2.3]).

Theorem 25

Let . Then H i(M) = 0 for i ≠ 0, and we have an isomorphism

$$\displaystyle{ H^{0}(M)\cong (M/\sum _{ i}\mathbb{C}[\mathfrak{g}^{{\ast}}](x_{ i} -\chi (x_{i}))M)^{N}. }$$

In particular if M is finitely generated H 0(M) is finitely generated over and

Corollary 6

The functor , M ↦ H 0(M), is exact.

Denote by the set of nilpotent elements of \(\mathfrak{g}\), which equals to the zero locus of the augmentation ideal \(\mathbb{C}[\mathfrak{g}^{{\ast}}]_{+}^{G}\) of \(\mathbb{C}[\mathfrak{g}^{{\ast}}]^{G}\) under the identification \(\mathfrak{g} = \mathfrak{g}^{{\ast}}\) via ( |  ). Since the element f (defined in (3)). is regular (or principal), the orbit

$$\displaystyle{ \mathbb{O}_{prin}:= G.f \subset \mathfrak{g} = \mathfrak{g}^{{\ast}} }$$

is dense in :

The transversality of implies that

Theorem 26 ([47])

Let M be a finitely generated object in .

  1. 1.

    H 0(M) 0 if and only if .

  2. 2.

    H 0(M) is nonzero and finite-dimensional if .

Proof

(1) Note that is invariant under the \(\mathbb{C}^{{\ast}}\)-action  (7) on , which contracts the point { f}. Hence is nonempty if and only if . The assertion follows since is G-invariant and closed. (2) Obvious since the assumption implies that .

6.6 Losev’s Correspondence

Let be the category of Harish-Chandra bimodules, that is, the full subcategory of the category of \(U(\mathfrak{g})\)-bimodules on which the adjoint action of \(\mathfrak{g}\) is integrable.

Lemma 32

Every finitely generated object M of admits a good filtration, that is, an increasing filtration 0 = F 0 MF 1 M… such that \(M =\bigcup F_{p}M\) ,

$$\displaystyle{ U_{p}(\mathfrak{g}) \cdot F_{q}M \cdot U_{r}(\mathfrak{g}) \subset F_{p+q+r}M,\quad [U_{p}(\mathfrak{g}),F_{p}M] \subset F_{p+q-1}M, }$$

and \(\mathop{\mathrm{gr}}\nolimits ^{F}M =\bigoplus _{p}F_{p}M/F_{p-1}M\) is finitely generated over \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\).

If and F M is a good filtration, then \(\mathop{\mathrm{gr}}\nolimits ^{F}M\) is naturally a Poisson module over \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\). Therefore, it is an object of .

Let M be a finitely generated object in . It is known since Bernstein that

$$\displaystyle{ \mathop{\mathrm{Var}}\nolimits (M):=\mathop{ \mathrm{supp}}\nolimits _{\mathbb{C}[\mathfrak{g}^{{\ast}}]}(\mathop{\mathrm{gr}}\nolimits ^{F}M) \subset \mathfrak{g}^{{\ast}} }$$

in independent of the choice of a good filtration F M of M.

For , MCl is naturally a bimodule over \(C(\mathfrak{g}) = U(\mathfrak{g}) \otimes Cl\). Thus, (MCl, adQ) is a differential graded bimodule over \(C(\mathfrak{g})\), and its cohomology

$$\displaystyle{ H^{\bullet }(M):= H^{\bullet }(M \otimes Cl,\mathrm{ad}Q) }$$

is naturally a module over \(H^{0}(C(\mathfrak{g}),\mathrm{ad}Q)\) that is identified with \(Z(\mathfrak{g})\) by Theorem 5. Thus, we have a functor

(59)

Let be finitely generated, F M a good filtration. Then F p (MCl): = i+j = p F i MCl j defines a good filtration of MCl, and the associated graded space \(\mathop{\mathrm{gr}}\nolimits _{F}(M \otimes Cl) =\sum _{i}F_{p}(M \otimes Cl)/F_{p-1}(M \otimes Cl) = (\mathop{\mathrm{gr}}\nolimits _{F}M) \otimes \overline{Cl}\) is a Poisson module over \(\mathop{\mathrm{gr}}\nolimits C(\mathfrak{g}) =\bar{ C}(\mathfrak{g})\).

The filtration F (MCl) induces a filtration F H (M) on H (M), and \(\mathop{\mathrm{gr}}\nolimits _{F}H^{\bullet }(M) =\bigoplus _{p}F_{p}H^{\bullet }(M)/F_{p-1}H^{0}(M)\) is a module over .

For a finitely generated -module M, set , \(\mathop{\mathrm{gr}}\nolimits M\) is the associated graded M with respect to a good filtration of M.

The following assertion follows from Theorems 25 and 26.

Theorem 27 ([47, 69])

  1. 1.

    We have H i(M) = 0 for all i ≠ 0, . Therefore the functor  (59) is exact.

  2. 2.

    Let M be a finitely generated object of , F M a good filtration. Then \(\mathop{\mathrm{gr}}\nolimits _{F}H^{0}(M)\cong H^{0}(\mathop{\mathrm{gr}}\nolimits _{F}M)\) . In particular H 0(M) is finitely generated, F H 0(M) is a good filtration of H 0(M).

  3. 3.

    For a finitely generated object M of , .

6.7 Frenkel-Zhu’s Bimodules

Recall that for a graded vertex algebra V, Zhu’s algebra Zhu(V ) = VVV is defined. There is a similar construction for modules due to Frenkel and Zhu [43]. For a V -module M set

$$\displaystyle{ \mathrm{Zhu}(M) = M/V \circ M, }$$

where VM is the subspace of M spanned by the vectors

$$\displaystyle{ a\circ m =\sum _{i\geqslant 0}\left (\begin{array}{*{10}c} \varDelta _{a} \\ i \end{array} \right )a_{(i-2)}m }$$

for \(a \in V _{\varDelta _{a}}\), \(\varDelta _{a} \in \mathbb{Z}\), and mM.

Proposition 10 ([43])

Zhu(M) is a bimodule over Zhu(V ) by the multiplications

$$\displaystyle{ a{\ast}m =\sum _{i\geqslant 0}\left (\begin{array}{*{10}c} \varDelta _{a} \\ i \end{array} \right )a_{(i-1)}b,\quad m{\ast}a =\sum _{i\geqslant 0}\left (\begin{array}{*{10}c} \varDelta _{a} - 1 \\ i \end{array} \right )a_{(i-1)}m }$$

for \(a \in V _{\varDelta _{a}}\) , \(\varDelta _{a} \in \mathbb{Z}\) , and mM.

Thus, we have a right exact functor

$$\displaystyle{ V \mathop{\mathrm{-Mod}}\nolimits \rightarrow \mathrm{ Zhu}(V )\mathop{\mathrm{-biMod}}\nolimits,\quad M\mapsto \mathrm{Zhu}(M). }$$

Lemma 33

Let \(M =\bigoplus _{d\in h+\mathbb{Z}_{+}}M_{d}\) be a positive energy representation of a \(\mathbb{Z}_{+}\) -graded vertex algebra V. Define an increasing filtration {Zhu p (M)} on Zhu(V ) by

$$\displaystyle{ \mathrm{Zhu}_{p}(M) =\mathop{ \mathrm{Im}}\nolimits (\bigoplus _{d=h}^{h+p}M_{ p} \rightarrow \mathrm{ Zhu}(M)). }$$
  1. 1.

    We have

    $$\displaystyle\begin{array}{rcl} & & \mathrm{Zhu}_{p}(V ) \cdot \mathrm{ Zhu}_{q}(M) \cdot \mathrm{ Zhu}_{r}(V ) \subset \mathrm{ Zhu}_{p+q+r}(M), {}\\ & & [\mathrm{Zhu}_{p}(V ),\mathrm{Zhu}_{q}(M)] \subset \mathrm{ Zhu}_{p+q-1}(M). {}\\ \end{array}$$

    Therefore \(\mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(M) =\bigoplus _{p}\mathrm{Zhu}_{p}(M)/\mathrm{Zhu}_{p-1}(M)\) is a Poisson \(\mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(V )\) -module, and hence is a Poisson R V -module through the homomorphism \(\eta _{V }: R_{V } \twoheadrightarrow \mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(V )\).

  2. 2.

    There is a natural surjective homomorphism

    $$\displaystyle{ \eta _{M}:\bar{ M}(= M/F^{1}M) \rightarrow \mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(M) }$$

    of Poisson R V -modules. This is an isomorphism if V admits a PBW basis and \(\mathop{\mathrm{gr}}\nolimits M\) is free over \(\mathop{\mathrm{gr}}\nolimits V\).

Example 7

Let M = V E k. Since \(\mathop{\mathrm{gr}}\nolimits V _{E}^{k}\) is free over \(\mathbb{C}[J\mathfrak{g}^{{\ast}}]\), we have the isomorphism

$$\displaystyle{ \eta _{V _{E}^{k}}: \overline{V _{E}^{k}} = E \otimes \mathbb{C}[\mathfrak{g}^{{\ast}}]\;\stackrel{_{ \sim }}{\rightarrow }\;\mathop{\mathrm{gr}}\nolimits \mathrm{Zhu}(V _{E}^{k}). }$$

On the other hand, there is a \(U(\mathfrak{g})\)-bimodule homomorphism

$$\displaystyle\begin{array}{rcl} E \otimes U(\mathfrak{g})& \rightarrow & \mathrm{Zhu}(V _{E}^{k}), \\ v \otimes x_{1}\ldots x_{r}& \mapsto & (1 \otimes v) {\ast} (x_{1}t^{-1}) {\ast} (x_{ 1}t^{-1}) + V ^{k}(\mathfrak{g}) \circ V _{ E}^{k}{}\end{array}$$
(60)

which respects the filtration. Here the \(U(\mathfrak{g})\)-bimodule structure of \(U(\mathfrak{g}) \otimes E\) is given by

$$\displaystyle{ x(v \otimes u) = (xv) \otimes u + v \otimes xu,\quad (v \otimes u)x = v \otimes (ux), }$$

and the filtration of \(U(\mathfrak{g}) \otimes E\) is given by \(\{U_{i}(\mathfrak{g}) \otimes E\}\). Since the induced homomorphism between associated graded spaces (60) coincides with \(\eta _{V _{E}^{k}}\), (60) is an isomorphism.

Lemma 34

For MKL k we have . If M is finitely generated, then so is Zhu(M).

6.8 Zhu’s Two Functors Commute with BRST Reduction

For a smooth \(\widehat{\mathfrak{g}}\)-module M over level k, is naturally a module over . Thus, (C(M), Q (0)) is a cochain complex, and its cohomology H (M): = H (C(M), Q (0)) is a module over \(\mathcal{W}^{k}(\mathfrak{g}) = H^{\bullet }(C^{k}(\mathfrak{g}),Q_{(0)})\). Thus we have a functor

$$\displaystyle{ V ^{k}(\mathfrak{g})\mathop{\mathrm{-Mod}}\nolimits \rightarrow \mathcal{ W}^{k}(\mathfrak{g})\mathop{\mathrm{-Mod}}\nolimits,\quad M\mapsto H^{0}(M). }$$

Here \(V \mathop{\mathrm{-Mod}}\nolimits\) denotes the category of modules over a vertex algebra V.

Theorem 28

  1. 1.

    [ 9 , 41 ] We have H i(M) = 0 for i ≠ 0, MKL k . In particular the functor

    $$\displaystyle{ \mathbf{KL}_{k} \rightarrow \mathcal{ W}^{k}(\mathfrak{g})\mathop{\mathrm{-Mod}}\nolimits,\quad M\mapsto H^{0}(M), }$$

    is exact.

  2. 2.

    [ 9 ] For a finitely generated object M of KL ,

    $$\displaystyle{ \overline{H^{0}(M)}\cong H^{0}(\bar{M}) }$$

    as Poisson modules over . In particular H 0(M) is finitely strongly generated and

  3. 3.

    [ 10 ] For a finitely generated object M of KL ,

    $$\displaystyle{ \mathop{\mathrm{Zhu}}\nolimits (H^{0}(M))\cong H^{0}(\mathop{\mathrm{Zhu}}\nolimits (M)) }$$

    as bimodules over .

Let \(\mathcal{W}_{k}(\mathfrak{g})\) denote the unique simple graded quotient of \(\mathcal{W}^{k}(\mathfrak{g})\). Then \(X_{\mathcal{W}_{k}(\mathfrak{g})}\) is a \(\mathbb{C}^{{\ast}}\)-invariant subvariety of . Therefore \(X_{\mathcal{W}^{k}(\mathfrak{g})}\) is lisse if and only if \(X_{\mathcal{W}_{k}(\mathfrak{g})} =\{\, f\}\) since the \(\mathbb{C}^{{\ast}}\)-action on contracts to the point f.

Corollary 7

  1. 1.

    \(H^{0}(V _{k}(\mathfrak{g}))\) is a quotient of \(\mathcal{W}^{k}(\mathfrak{g}) = H^{0}(V ^{k}(\mathfrak{g}))\) . In particular \(\mathcal{W}_{k}(\mathfrak{g})\) is a quotient of \(H^{0}(V _{k}(\mathfrak{g}))\) if \(H^{0}(V _{k}(\mathfrak{g}))\) is nonzero.

  2. 2.

    \(H^{0}(V _{k}(\mathfrak{g}))\) is nonzero if and only if .

  3. 3.

    The simple W-algebra \(\mathcal{W}_{k}(\mathfrak{g})\) is lisse if .

Proof

 

  1. 1.

    follows from the exactness statement of Theorem 28.

  2. 2.

    \(H^{0}(V _{k}(\mathfrak{g}))\) is nonzero if and only of is non-empty. This happens if and only if fX M since \(X_{H^{0}(M)}\) is \(\mathbb{C}^{{\ast}}\)-stable. The assertion follows since X M is G-invariant and closed.

  3. 3.

    If \(X_{V _{k}(\mathfrak{g})} = \overline{G.f}\), , and thus, \(H^{0}(V _{k}(\mathfrak{g}))\) is lisse, and thus, so its quotient \(\mathcal{W}_{k}(\mathfrak{g})\).

Remark 13

 

  1. 1.

    The above results hold for W-algebras associated with any \(\mathfrak{g}\) and any without any restriction on the level k [9, 10]. In particular we have the vanishing result

    $$\displaystyle{ H_{f}^{i}(M) = 0\quad \text{for }i\neq 0,\ M \in \mathbf{KL}_{ k}, }$$
    (61)

    for the BRST cohomology H f i(M) of the quantized Drinfeld-Sokolov reduction functor associated with f in the coefficient in an object M of KL k . Thus the functor

    $$\displaystyle{ \mathbf{KL}_{k} \rightarrow \mathcal{ W}^{k}(\mathfrak{g},f)\mathop{\mathrm{-Mod}}\nolimits,\quad M\mapsto H_{ f}^{0}(M), }$$

    is exact, and moreover,

    where S f is the Slodowy slice at f (see Sect. 2.8). In particular

    $$\displaystyle{ H_{f}^{0}(V _{ k}(\mathfrak{g}))\neq 0\;\Longleftrightarrow\;X_{V _{k}(\mathfrak{g})} \supset \overline{G.f}. }$$
    (62)
  2. 2.

    In the case that f = f θ , a minimal nilpotent element of \(\mathfrak{g}\), then we also have the following result [4]:

    $$\displaystyle{ H_{f_{\theta }}(V _{k}(\mathfrak{g})) = \left \{\begin{array}{@{}l@{\quad }l@{}} \mathcal{W}_{k}(\mathfrak{g},f_{\theta })\quad &\text{if }k\not\in \mathbb{Z}_{+}, \\ 0 \quad &\text{if }k \in \mathbb{Z}_{+}.\end{array} \right. }$$

    Here \(\mathcal{W}_{k}(\mathfrak{g},f_{\theta })\) is the simple quotient of \(\mathcal{W}^{k}(\mathfrak{g},f_{\theta })\). Together with (62), this proves the “only if” part of Theorem 12. Indeed, if \(V _{k}(\mathfrak{g})\) is lisse, then \(H_{f_{\theta }}(V _{k}(\mathfrak{g})) = 0\) by (62), and hence, \(k \in \mathbb{Z}_{+}\).

7 Irreducible Representations of W-Algebras

In this section we quickly review results obtained in [5].

Since , by Zhu’s theorem irreducible positive energy representations of \(\mathcal{W}^{k}(\mathfrak{g})\) are parametrized by central characters of . For a central character , let \(\mathbb{L}(\gamma )\) be the corresponding irreducible positive energy representations of \(\mathcal{W}^{k}(\mathfrak{g})\). This is a simple quotient of the Verma module \(\mathbb{M}(\gamma )\) of \(\mathcal{W}^{k}(\mathfrak{g})\) with highest weight γ, which has the character

$$\displaystyle{ \mathop{\mathrm{ch}}\nolimits \mathbb{M}(\gamma ):=\mathop{ \mathrm{tr}}\nolimits _{\mathbb{M}(\gamma )}q^{L_{0} } = \frac{q^{ \frac{\gamma (\varOmega )} {2(k+h^{\vee })} }} {\prod _{j\geqslant 1}(1 - q^{j})^{\mathop{\mathrm{rk}}\nolimits \mathfrak{g}}} }$$

in the case that k is non-critical, where Ω is the Casimir element of \(U(\mathfrak{g})\).

In Theorem 28 we showed that the functor \(\mathbf{KL}_{k} \rightarrow \mathcal{ W}^{k}(\mathfrak{g})\mathop{\mathrm{-Mod}}\nolimits\), MH 0(M), is exact. However in order to obtain all the irreducible positive energy representation we need to extend this functor to the whole category . However the functor , MH 0(M), is not exact in general except for the case \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\) [4]. Nevertheless, we can [44] modify the functor to obtain the following result.

Theorem 29 ([5])

There exists an exact functor

(called the ˆˆ ˆˆ”-reduction functor in [ 44 ]), which enjoys the following properties.

  1. 1.

    \(H_{-}^{0}(M(\lambda ))\cong \mathbb{M}(\gamma _{\bar{\lambda }})\) , where M(λ) is the Verma module of \(\widehat{\mathfrak{g}}\) with highest weight λ, and \(\gamma _{\bar{\lambda }}\) is the evaluation of at the Verma module \(M_{\mathfrak{g}}(\bar{\lambda })\) of \(\mathfrak{g}\) with highest weight \(\bar{\lambda }\).

  2. 2.

    \(H_{-}^{0}(L(\lambda ))\cong \left \{\begin{array}{@{}l@{\quad }l@{}} \mathbb{L}(\gamma _{\bar{\lambda }})\quad &\mathit{\text{if }}\bar{\lambda }\mathit{\text{ is anti-dominant}}\ (\mathit{\text{that is,}}\ M_{\mathfrak{g}}(\bar{\lambda })\ \mathit{\text{is simple}}),\\ 0 \quad &\mathit{\text{otherwise.}}\end{array} \right.\)

Corollary 8

Write \(\mathop{\mathrm{ch}}\nolimits L(\lambda ) =\sum _{\mu }c_{\lambda,\mu }\mathop{ \mathrm{ch}}\nolimits M(\mu )\) with \(c_{\lambda,\mu } \in \mathbb{Z}\) . If \(\bar{\lambda }\) is anti-dominant, we have

$$\displaystyle{ \mathop{\mathrm{ch}}\nolimits \mathbb{L}(\gamma _{\bar{\lambda }}) =\sum _{\mu }c_{\lambda,\mu }\mathop{ \mathrm{ch}}\nolimits \mathbb{M}(\gamma _{\bar{\mu }}). }$$

In the case that k is non-critical, then it is known by Kashiwara and Tanisaki [62] that the coefficient c λ, μ is expressed in terms of Kazhdan-Lusztig polynomials. Since any central character of can be written as \(\gamma _{\bar{\lambda }}\) with anti-dominant \(\bar{\lambda }\), Corollary 8 determines the character of all the irreducible positive energy representations of \(\mathcal{W}^{k}(\mathfrak{g})\) for all non-critical k.

On the other hand, in the case that k is critical, all \(\mathbb{L}(\gamma _{\bar{\lambda }})\) are one-dimensional since \(\mathcal{W}^{-n}(\mathfrak{g})\) is commutative. This fact with Theorem 29 can be used in the study of the critical level representations of \(\widehat{\mathfrak{g}}\), see [13].

The results in this section hold for arbitrary simple Lie algebra \(\mathfrak{g}\).

Remark 14

The condition \(\bar{\lambda }\in \mathfrak{h}^{{\ast}}\) is anti-dominant does not imply that \(\lambda \in \widehat{ \mathfrak{h}}^{{\ast}}\) is anti-dominant. In fact this condition is satisfied by all non-degenerate admissible weights λ (see below) which are regular dominant.

Remark 15

Theorem 29 has been generalized in [6]. In particular the character of all the simple ordinary representations (=simple positive energy representations with finite-dimensional homogeneous spaces) has been determined for W-algebras associated with all nilpotent elements f in type A.

8 Kac-Wakimoto Admissible Representations and Frenkel-Kac-Wakimoto Conjecture

We continue to assume that \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{n}\), but the results in this section holds for arbitrary simple Lie algebra \(\mathfrak{g}\) as well with appropriate modification unless otherwise stated.

8.1 Admissible Affine Vertex Algebras

Let \(\widehat{\mathfrak{h}}\) be the Cartan subalgebra \(\mathfrak{h}\mathop{\oplus }\mathbb{C}K\) of \(\widehat{\mathfrak{g}}\), \(\tilde{\mathfrak{h}} = \mathfrak{h}\mathop{\oplus }\mathbb{C}K\mathop{\oplus }\mathbb{C}D\) the extended Cartan subalgebra, \(\widehat{\varDelta }\) the set of roots of \(\widehat{\mathfrak{g}}\) in \(\tilde{\mathfrak{h}}^{{\ast}} = \mathfrak{h}^{{\ast}}\mathop{\oplus }\mathbb{C}\varLambda _{0}\mathop{\oplus }\mathbb{C}\delta\), where Λ 0(K) = 1 = δ(d), \(\varLambda _{0}(\mathfrak{h} + \mathbb{C}D) =\delta (\mathfrak{h}\mathop{\oplus }\mathbb{C}K) = 0\), \(\widehat{\varDelta }_{+}\) the set of positive roots. \(\widehat{\varDelta }^{re} \subset \widehat{\varDelta }\) the set of real roots, \(\widehat{\varDelta }_{+}^{re} =\widehat{\varDelta } ^{re} \cap \widehat{\varDelta }_{+}\). Let \(\widehat{W}\) be the affine Weyl group of \(\widehat{\mathfrak{g}}\).

Definition 6 ([57])

A weight \(\lambda \in \widehat{ \mathfrak{h}}^{{\ast}}\) is called admissible if

  1. 1.

    λ is regular dominant, that is,

    $$\displaystyle{ \langle \lambda +\rho,\alpha ^{\vee }\rangle \not\in - \mathbb{Z}_{ +}\quad \text{for all }\alpha \in \widehat{\varDelta }_{+}^{re}, }$$
  2. 2.

    \(\mathbb{Q}\widehat{\varDelta }(\lambda ) = \mathbb{Q}\widehat{\varDelta }^{re}\), where \(\widehat{\varDelta }(\lambda ) =\{\alpha \in \widehat{\varDelta }^{re}\mid \langle \lambda +\rho,\alpha ^{\vee }\rangle \in \mathbb{Z}\}\).

The irreducible highest weight representation L(λ) of \(\widehat{\mathfrak{g}}\) with highest weight \(\lambda \in \widehat{ \mathfrak{h}}^{{\ast}}\) is called admissible if λ is admissible. Note that an irreducible integrable representations of \(\widehat{\mathfrak{g}}\) is admissible.

Clearly, integrable representations of \(\widehat{\mathfrak{g}}\) are admissible.

For an admissible representation L(λ) we have [56]

$$\displaystyle{ \mathop{\mathrm{ch}}\nolimits L(\lambda ) =\sum _{w\in \widehat{W}(\lambda )}(-1)^{\ell_{\lambda }(w)}\mathop{ \mathrm{ch}}\nolimits M(w\circ \lambda ) }$$
(63)

since λ is regular dominant, where \(\widehat{W}(\lambda )\) is the integral Weyl group [61, 74] of λ, that is, the subgroup of \(\widehat{W}\) generated by the reflections s α associated with \(\alpha \in \widehat{\varDelta }\) and wλ = w(λ + ρ) −ρ. Further the condition (2) implies that \(\mathop{\mathrm{ch}}\nolimits L(\lambda )\) is written in terms of certain theta functions. Kac and Wakimoto [57] showed that admissible representations are modular invariant, that is, the characters of admissible representations form an \(SL_{2}(\mathbb{Z})\) invariant subspace.

Let λ, μ be distinct admissible weights. Then the condition (1) implies that

$$\displaystyle{ \mathop{\mathrm{Ext\,}}\nolimits _{\widehat{\mathfrak{g}}}^{1}(L(\lambda ),L(\mu )) = 0. }$$

Further, the following fact is known by Gorelik and Kac [49].

Theorem 30 ([49])

Let λ be admissible. Then \(\mathop{\mathrm{Ext}}\nolimits _{\widehat{\mathfrak{g}}}(L(\lambda ),L(\lambda )) = 0\).

Therefore admissible representations form a semisimple fullsubcategory of the category of \(\widehat{\mathfrak{g}}\)-modules.

Recall that the simple affine vertex algebra \(V _{k}(\mathfrak{g})\) is isomorphic to L( 0) as an \(\widehat{\mathfrak{g}}\)-module.

Lemma 35

The following conditions are equivalent.

  1. 1.

    0 is admissible.

  2. 2.

    0 is regular dominant and \(k \in \mathbb{Q}\).

  3. 3.

    k + h = pq, \(p,q \in \mathbb{N}\) , ( p, q) = 1, \(p\geqslant h^{\vee } = n\).

If this is the case, the level k is called admissible for \(\widehat{\mathfrak{g}}\) , and \(V _{k}(\mathfrak{g})\) is called an admissible affine vertex algebra.

For an admissible number k let Pr k be the set of admissible weights of \(\widehat{\mathfrak{g}}\) of level k. (For \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{n}\), Pr k is the same as the set of principal admissible weights of level k.)

8.2 Feigin-Frenkel Conjecture and Adamović-Milas Conjecture

The following fact was conjectured by Feigin and Frenkel and proved for the case that \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\) by Feigin and Malikov [38].

Theorem 31 ([9])

The associated variety \(X_{V _{k}(\mathfrak{g})}\) is contained in if k is admissible.

In fact the following holds.

Theorem 32 ([9])

Let k be admissible, and let qN be the denominator of k, that is, k + h = pq, pN, ( p, q) = 1. Then

$$\displaystyle{ X_{V _{k}(\mathfrak{g})} =\{ x \in \mathfrak{g}\mid (\mathrm{ad}x)^{2q} = 0\} = \overline{\mathbb{O}_{ q}}, }$$

where \(\mathbb{O}_{q}\) is the nilpotent orbit corresponding to the partition

$$\displaystyle{ \left \{\begin{array}{@{}l@{\quad }l@{}} (n) \quad &\mathit{\text{if }}q\geqslant n,\\ (q, q,\ldots, q, s)\quad (0\leqslant s\leqslant n - 1)\quad &\mathit{\text{if }}q <n. \end{array} \right. }$$

The following fact was conjectured by Adamović and Milas [2].

Theorem 33 ([11])

Let k be admissible. Then an irreducible highest weight representation L(λ) is a \(V _{k}(\mathfrak{g})\) -module if and only if kPr k . Hence if M is a finitely generated \(V _{k}(\mathfrak{g})\) -module on which \(\widehat{\mathfrak{n}}_{+}\) acts locally nilpotently and \(\widehat{\mathfrak{h}}\) acts locally finitely then M is a direct sum of L(λ) with λPr k .

8.3 Outline of Proofs of Theorems 31, 32 and 33

The idea of the proofs of Theorems 31 and 33 is to reduce to the \(\widehat{\mathfrak{s}\mathfrak{l}_{2}}\)-cases.

Let \(\mathfrak{s}\mathfrak{l}_{2,i} \subset \mathfrak{g}\) be the copy of \(\mathfrak{s}\mathfrak{l}_{2}\) spanned by e i : = e i, i+1, h i : = e i, i e i+1, i+1, f i : = e i+1, i , and let \(\mathfrak{p}_{i} = \mathfrak{s}\mathfrak{l}_{2,i} + \mathfrak{b} \subset \mathfrak{g}\), the associated minimal parabolic subalgebra. Then

$$\displaystyle{ \mathfrak{p}_{i} = \mathfrak{l}_{i}\mathop{\oplus }\mathfrak{m}_{i}, }$$

where \(\mathfrak{l}_{i}\) is the Levi subalgebra \(\mathfrak{s}\mathfrak{l}_{2,i} + \mathfrak{h}\), and \(\mathfrak{m}_{i}\) is the nilradical \(\bigoplus \limits _{{ 1\leqslant p<q\leqslant n \atop (\,p,q)\neq (i,i+1)} }\mathbb{C}e_{p,q}\).

Consider the semi-infinite cohomology \(H^{\frac{\infty } {2} +0}(\mathfrak{m}_{i}[t,t^{-1}],M)\). It is defined as a cohomology of Feigin’s complex \((C(\mathfrak{m}_{i}[t,t^{-1}],M),d)\) [35]. There is a natural vertex algebra homomorphism

$$\displaystyle{ V ^{k_{i} }(\mathfrak{s}\mathfrak{l}_{2}) \rightarrow H^{\frac{\infty } {2} +0}(\mathfrak{m}_{i}[t,t^{-1}],M), }$$
(64)

where k i = k + n − 2, see, e.g. [50]. Note that if k is an admissible number for \(\widehat{\mathfrak{g}}\) then k i is an admissible number for \(\widehat{\mathfrak{s}\mathfrak{l}_{2}}\).

Theorem 34 ([8])

Let k be an admissible number. The map  (64) factors through the vertex algebra embedding

$$\displaystyle{ V _{k_{i}}(\mathfrak{s}\mathfrak{l}_{2})\hookrightarrow H^{\frac{\infty } {2} +0}(\mathfrak{m}_{i}[t,t^{-1}],V _{k}(\mathfrak{g})). }$$

Proof (Outline of Proof of Theorem 31)

First, consider the case that \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\). Let N k be the maximal submodule of \(V ^{k}(\mathfrak{g})\), and let I k be the image of N k in \(R_{V ^{k}(\mathfrak{g})} = \mathbb{C}[\mathfrak{g}^{{\ast}}]\), so that \(R_{V _{k}(\mathfrak{g})} = \mathbb{C}[\mathfrak{g}^{{\ast}}]/I_{k}\). It is known by Kac and Wakimoto [56] that N k is generated by a singular vector, say v k . The projection formula [72] implies that the image [v k ] of v k in I k is nonzero. Since [v k ] is a singular vector of \(\mathbb{C}[\mathfrak{g}^{{\ast}}]\) with respect to the adjoint action of \(\mathfrak{g}\), Kostant’s Separation Theorem implies that

$$\displaystyle{ [v_{k}] = e^{m}\varOmega ^{n} }$$

for some \(m,n \in \mathbb{N}\) up to constant multiplication, where \(\varOmega = ef + fe + \frac{1} {2}h^{2}\). Now suppose that and let , so that Ω(λ) ≠ 0. Then e(λ) = 0. Since \(X_{V _{k}(\mathfrak{g})}\) is G-invariant this implies that x(λ) = 0 for any nilpotent element x of \(\mathfrak{g}\). Because any element of \(\mathfrak{g}\) can be written as a sum of nilpotent elements we get that λ = 0. Contradiction.

Next, consider the case that \(\mathfrak{g}\) is general. Note that since \(X_{V _{k}(\mathfrak{g})}\) is G-invariant and closed, the condition is equivalent to that \(X_{V _{k}(\mathfrak{g})} \cap \mathfrak{h}^{{\ast}} =\{ 0\}\). Now the complex structure of \(C(\mathfrak{m}_{i}[t,t^{-1}],V _{k}(\mathfrak{g}))\) induces the complex structure on Zhu’s C 2-algebra \(R_{C(\mathfrak{m}_{i}[t,t^{-1}],V _{k}(\mathfrak{g}))}\). The embedding in Theorem 34 induces a homomorphism

$$\displaystyle{ R_{V _{k_{ i}}(\mathfrak{s}\mathfrak{l}_{2})} \rightarrow H^{0}(R_{ C(\mathfrak{m}_{i}[t,t^{-1}],V _{k}(\mathfrak{g}))},d) }$$

of Poisson algebra. Since Ω is nilpotent in \(R_{V _{k_{ i}}(\mathfrak{s}\mathfrak{l}_{2})}\), so is its image \(\varOmega _{i} = e_{i}f_{i} + f_{i}e_{i} + \frac{1} {2}h_{i}^{2}\) in \(H^{0}(R_{C(\mathfrak{m}_{i}[t,t^{-1}],V _{k}(\mathfrak{g}))},d)\). It follows that \(h_{i}^{N} \equiv 0\pmod \mathfrak{n}_{+}R_{V _{k}(\mathfrak{g})} + \mathfrak{n}_{-}R_{V _{k}(\mathfrak{g})}\) in \(R_{V _{k}(\mathfrak{g})}\) for all i = 1, , nq, and we get that \(X_{V _{k}(\mathfrak{g})} \cap \mathfrak{h}^{{\ast}} =\{ 0\}\) as required.

Proof (Outline of Proof of Theorem 32)

The proof is done by determining the variety \(X_{V _{k}(\mathfrak{g})}\). By Theorem 31, \(X_{V _{k}(\mathfrak{g})}\) is a finite union of nilpotent orbits. Thus it is enough to know which nilpotent element orbits is contained in \(X_{V _{k}(\mathfrak{g})}\). On the other hand, (62) says \(X_{V _{k}(\mathfrak{g})} \supset \overline{G.f}\) if and only \(H_{f}^{0}(V _{k}(\mathfrak{g}))\neq 0\). Thus, it is sufficient to compute the character of \(H_{f}^{0}(V _{k}(\mathfrak{g}))\). This is in fact possible since we know the explicit formula (63) of the character of \(V _{k}(\mathfrak{g})\), thanks to the vanishing theorem (61) and the Euler-Poincaré principle.

Proof (Outline of Proof of Theorem 33)

Let L(λ) be a \(V _{k}(\mathfrak{g})\)-module. Then, the space \(H^{\frac{\infty } {2} +i}(\mathfrak{m}_{i}[t,t^{-1}],L(\lambda ))\), \(i \in \mathbb{Z}\), is naturally a \(H^{\frac{\infty } {2} +i}(\mathfrak{m}_{i}[t,t^{-1}],V _{k}(\mathfrak{g}))\)-module. By Theorem 34, this means that \(H^{\frac{\infty } {2} +i}(\mathfrak{m}_{i}[t,t^{-1}],L(\lambda ))\) is in particular a module over the admissible affine vertex algebra \(V _{k_{i}}(\mathfrak{s}\mathfrak{l}_{2})\). Therefore Theorem 33 for \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\) that was established by Adamović and Milas [2] implies that \(H^{\frac{\infty } {2} +i}(\mathfrak{m}_{i}[t,t^{-1}],L(\lambda ))\) must be a direct sum of admissible representations of \(\widehat{\mathfrak{s}\mathfrak{l}}_{2}\). This information is sufficient to conclude that L(λ) is admissible.

Conversely, suppose that L(λ) is an admissible representation of level k. If L(λ) is integrable over \(\mathfrak{g}\), then it has been already proved by Frenkel and Malikov [42] that L(λ) is a \(V _{k}(\mathfrak{g})\)-module. But then an affine analogue of Duflo-Joseph Lemma [11, Lemma 2.6] implies that this is true for a general admissible representation as well.

8.4 Lisse Property of W-Algebras

An admissible number k is called non-degenerate if . By Theorem 32, this condition is equivalent to that

$$\displaystyle{ k + n = \frac{p} {q},\ p,q \in \mathbb{N},\ (\,p,q) = 1,\ p\geqslant n,\ q\geqslant n. }$$

The following assertion follows immediately from Corollary 7.

Theorem 35 ([9])

Let k be a non-degenerate admissible number. Then the W-algebra \(\mathcal{W}_{k}(\mathfrak{g})\) is lisse.

8.5 Minimal Models of W-Algebras

A vertex algebra V is called rational if any V -module is completely reducible. To a lisse and rational conformal vertex algebra V one can associate rational 2d conformal field theory, and in particular, the category \(V \mathop{\mathrm{-Mod}}\nolimits\) of V -modules forms [51] a modular tensor category [19], as in the case of the category of integrable representation of \(\widehat{\mathfrak{g}}\) at a positive level and the category of minimal series representations [23] of the Virasoro algebra.

An admissible weight λ is called non-degenerate if \(\bar{\lambda }\) is anti-dominant. Let Pr k non-deg be the set of non-degenerate admissible weights of level k of \(\widehat{\mathfrak{g}}\). It is known [44] that Pr k non-deg is non-empty if and only if k is non-degenerate.

By Theorem 29, for λPr k, H 0(L(λ)) is a (non-zero) simple \(\mathcal{W}^{k}(\mathfrak{g})\)-module if and only of λPr k non-deg, and \(H_{-}^{0}(L(\lambda ))\cong H_{-}^{0}(L(\mu ))\) if and only if μWλ for λ, μPr k non-deg.

Let [Pr k non-deg] = Pr k non-deg∕ ∼, where λμ ⇔ μWλ. It is known [44] that we have a bijection

$$\displaystyle{ (\widehat{P}_{+}^{p-n} \times \widehat{ P}_{ +}^{q-n})/\mathbb{Z}_{ n}\;\stackrel{_{\sim }}{\rightarrow }\;[Pr_{k}^{non\mathrm{-}deg}],\quad [(\lambda,\mu )]\mapsto [\bar{\lambda }-\frac{p} {q}(\bar{\mu }+\rho ) + k\varLambda _{0}]. }$$

Here k + n = pq as before, \(\widehat{P}_{+}^{k}\) is the set of integral dominant weights of level k of \(\widehat{\mathfrak{g}}\), the cyclic group \(\mathbb{Z}_{n}\) acts diagonally on \(\widehat{P}_{+}^{p-n} \times \widehat{ P}_{+}^{q-n}\) as the Dynkin automorphism, and \(\rho = \frac{1} {2}\sum _{\alpha \in \varDelta _{+}}\alpha\).

The following assertion was conjectured by Frenkel et al. [44].

Theorem 36 ([10])

Let k be a non-degenerate admissible number. Then the simple W-algebra \(\mathcal{W}_{k}(\mathfrak{g})\) is rational, and \(\{\mathbb{L}(\gamma _{\bar{\lambda }}) = H_{-}^{0}(L(\lambda ))\mid \lambda \in [Pr_{k}^{non\mathrm{-}deg}]\}\) forms the complete set of isomorphism classes of simple \(\mathcal{W}_{k}(\mathfrak{g})\) -modules.

In the case that \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\), Theorems 35 and 36 have been proved in [22, 77], and the above representations are exactly the minimal series representations of the Virasoro algebra.

The representations

$$\displaystyle{ \{\mathbb{L}(\gamma _{\bar{\lambda }})\mid \lambda \in [Pr_{k}^{non\mathrm{-}deg}]\} }$$

are called the minimal series representations of \(\mathcal{W}^{k}(\mathfrak{g})\), and if k + n = pq, \(p,q \in \mathbb{N}\), ( p, q) = 1, \(p,q,\geqslant n\), then the rational W-algebra \(\mathcal{W}_{k}(\mathfrak{g})\) is called the ( p, q)-minimal model of \(\mathcal{W}^{k}(\mathfrak{g})\). Note that the ( p, q)-minimal model and the (q, p)-minimal model are isomorphic due to the duality, see Corollary 5.

Proof (Outline of the Proof of Theorem 36)

Let k be a non-degenerate admissible number. We have

$$\displaystyle{ H^{0}(V _{ k}(\mathfrak{g}))\cong \mathcal{W}_{k}(\mathfrak{g}) }$$

by [5]. Hence by Theorem 28 (3)

$$\displaystyle{ \mathop{\mathrm{Zhu}}\nolimits (\mathcal{W}_{k}(\mathfrak{g})) =\mathop{ \mathrm{Zhu}}\nolimits (H^{0}(V _{ k}(\mathfrak{g}))) = H^{0}(\mathop{\mathrm{Zhu}}\nolimits (V _{ k}(\mathfrak{g})). }$$

From this together with Theorem 33, it is not too difficult to obtain the classification is the simple \(\mathcal{W}_{k}(\mathfrak{g})\)-modules as stated in Theorem 36. One sees that the extensions between simple modules are trivial using the linkage principle that follows from Theorem 29.

Remark 16

 

  1. 1.

    We have \(\mathcal{W}_{k}(\mathfrak{g}) = \mathbb{L}(\gamma _{-(k+n)\rho })\) for a non-degenerate admissible number k. (Note that 0Pr k nondeg.)

  2. 2.

    Let λPr k . From Corollary 8 and (63), we get

    $$\displaystyle{ \mathop{\mathrm{ch}}\nolimits \mathbb{L}(\gamma _{\bar{\lambda }}) =\sum _{w\in \widehat{W}(\lambda )}\epsilon (w)\mathop{\mathrm{ch}}\nolimits \mathbb{M}(\gamma _{\overline{w\circ \lambda }}). }$$
    (65)

    This was conjectured by [44].

  3. 3.

    When it is trivial (that is, equals to \(\mathbb{C}\)), \(\mathcal{W}_{k}(\mathfrak{g})\) is obviously lisse and rational. This happens if and only if \(\mathcal{W}_{k}(\mathfrak{g})\) is the (n, n + 1)-minimal model (=the (n + 1, n)-minimal model). In this case the character formula (65) for \(\mathcal{W}_{k}(\mathfrak{g}) = \mathbb{L}(\gamma _{\bar{\lambda }})\), λ = −(k + n)ρ + 0, gives the following denominator formula:

    $$\displaystyle{ \sum _{w\in \widehat{W}(\lambda )}\epsilon (w)q^{\frac{(\overline{w\circ \lambda },\overline{w\circ \lambda }+2\rho )} {2(k+n)} } =\prod _{ j=1}^{n-1}(1 - q^{j})^{n-1}. }$$

    In the case that \(\mathfrak{g} = \mathfrak{s}\mathfrak{l}_{2}\), we get the denominator formula for the Virasoro algebra, which is identical to Euler’s pentagonal identity.

  4. 4.

    As a generalization of the GKO construction [48] it has been conjectured [58] that the ( p, q)-minimal model of \(\mathcal{W}^{k}(\mathfrak{g})\), with p > q, is isomorphic to the commutant of \(V _{l+1}(\mathfrak{g})\) inside \(V _{l}(\mathfrak{g}) \otimes V _{1}(\mathfrak{g})\), where l + n = q∕( pq). (Note that \(V _{l}(\mathfrak{g})\) and \(V _{l+1}(\mathfrak{g})\) are admissible.) This conjecture has been proved in [18] for the special case that ( p, q) = (n + 1, n).

    A similar conjecture exists in the case that \(\mathfrak{g}\) is simply laced.

  5. 5.

    The existence of rational and lisse W-algebras has been conjectured for general W-algebras \(\mathcal{W}^{k}(\mathfrak{g},f)\) by Kac and Wakimoto [59]. This has been proved in [12] in part including all the cases in type A. See [15, 63] for a recent development in the classification problem of rational and lisse W-algebras.