1 Introduction

Let \({\mathfrak g}\) be a finite-dimensional reductive Lie algebra over an algebraically closed field \(\Bbbk \) of characteristic zero. The symmetric algebra \({\mathcal {S}}({\mathfrak g})\cong \Bbbk [\mathfrak g^*]\) of \({\mathfrak g}\) is equipped with the standard Poisson structure, i.e., the Lie–Poisson bracket \(\{\,\,,\,\}\). A subalgebra \(A\subset {\mathcal {S}}({\mathfrak g})\) is Poisson commutative if \(\{A,A\}=0\). Poisson commutative subalgebras attract a great deal of attention, because of their relationship to integrable systems and geometric representation theory. If \(\{A,A\}=0\), then \({\mathrm {tr.deg\,}}A\leqslant {\varvec{b}}({\mathfrak g})\), where \({\varvec{b}}({\mathfrak g}):=\frac{1}{2}(\dim {\mathfrak g}+{\mathsf {rk\,}}{\mathfrak g})\) is the dimension of a Borel subalgebra of \({\mathfrak g}\).

The celebrated “argument shift method”, which goes back to Mishchenko–Fomenko [7], produces interesting Poisson commutative subalgebras. Namely, to any \(\gamma \in {\mathfrak g}^*\), one associates the subalgebra \((\mathcal{M}\mathcal{F})_\gamma \subset {\mathcal {S}}({\mathfrak g})\). Following Vinberg [19], we say that \((\mathcal{M}\mathcal{F})_\gamma \) is the Mishchenko–Fomenko subalgebra associated with \(\gamma \). This algebra can be described as follows. Let \({\mathcal {S}}({\mathfrak g})^{\mathfrak g}\) be the Poisson centre of \(({\mathcal {S}}({\mathfrak g}), \{\,\,,\,\})\), i.e.,

$$\begin{aligned} {\mathcal {S}}({\mathfrak g})^{\mathfrak g}=\{H\in {\mathcal {S}}({\mathfrak g})\mid \{H,x\}=0 \ \ \forall x\in {\mathfrak g}\}. \end{aligned}$$

For \(F\in {\mathcal {S}}({\mathfrak g})\), let \(\partial _{\gamma } F\) be the directional derivative of F with respect to \(\gamma \in {\mathfrak g}^*\), i.e.,

$$\begin{aligned} \partial _{\gamma }F(x)=\frac{\textit{d}}{\textit{d}t} F(x+t\gamma )\Big |_{t=0}. \end{aligned}$$

By definition, the algebra \((\mathcal{M}\mathcal{F})_\gamma \) is generated by all \(\partial _\gamma ^k F\) with \(k\geqslant 0\) and \(F\in {\mathcal {S}}({\mathfrak g})^{{\mathfrak g}}\). Clearly, \((\mathcal{M}\mathcal{F})_\gamma \) is a graded subalgebra of \({\mathcal {S}}({\mathfrak g})\). The importance of these subalgebras and their quantum counterparts is demonstrated e.g. in [3, 4, 8, 19]. Suppose that \(\gamma \) is regular. Then \((\mathcal{M}\mathcal{F})_\gamma \) is a maximal Poisson commutative subalgebra of \({\mathcal {S}}({\mathfrak g})\) [11]. For regular semisimple elements, this has been earlier proved by Tarasov [17]. Furthermore, \((\mathcal{M}\mathcal{F})_\gamma \) is freely generated by \({\varvec{b}}({\mathfrak g})\) homogeneous elements, see e.g. [11].

Let \({\mathfrak f}\) be a subalgebra of \({\mathfrak g}\). Suppose that there is an \({\mathfrak f}\)-stable decomposition \({\mathfrak g}={\mathfrak f}\oplus {\mathfrak m}\), i.e., \([{\mathfrak f},{\mathfrak m}]\subset {\mathfrak m}\). This yields a bi-homogeneous decomposition of \({\mathcal {S}}({\mathfrak g})\):

$$\begin{aligned} {\mathcal {S}}({\mathfrak g})=\bigoplus _{i,j \geqslant 0} {\mathcal {S}}^i({\mathfrak f})\otimes {\mathcal {S}}^j({\mathfrak m}) \end{aligned}$$

and for any \(F\in {\mathcal {S}}({\mathfrak g})\) we get the decomposition \(F=\sum _{i,j}F_{i,j}\), where \(F_{i,j}\in {\mathcal {S}}^i({\mathfrak f})\otimes {\mathcal {S}}^j({\mathfrak m})\).

Let \({\mathcal {Z}}_{({\mathfrak g},{\mathfrak f})}\) be the subalgebra of \({\mathcal {S}}({\mathfrak g})\) generated by the bi-homogeneous components \(H_{i,j}\) of all \(H\in {\mathcal {S}}({\mathfrak g})^{{\mathfrak g}}\). Since H is \({\mathfrak g}\)-invariant, each \(H_{i,j}\) is \({\mathfrak f}\)-invariant, hence \({\mathcal {Z}}_{({\mathfrak g},{\mathfrak f})}\subset {\mathcal {S}}({\mathfrak g})^{{\mathfrak f}}\). The subalgebras \({\mathcal {Z}}_{({\mathfrak g},{\mathfrak f})}\) are not necessarily Poisson commutative [13, Example 2.3]. However, it is proved in loc. cit. that \({\mathcal {Z}}_{({\mathfrak g},{\mathfrak f})}\) is Poisson commutative whenever \([{\mathfrak f},{\mathfrak f}]=0\). In particular, if \({\mathfrak f}={\mathfrak t}\) is a Cartan subalgebra of \({\mathfrak g}\), then \({\mathcal {Z}}_{({\mathfrak g},{\mathfrak t})}\) is Poisson commutative. Since \({\mathcal {Z}}_{({\mathfrak g},{\mathfrak t})}\subset {\mathcal {S}}({\mathfrak g})^{\mathfrak t}\) and \({\mathfrak t}\) is commutative, the subalgebra generated by \({\mathcal {Z}}_{({\mathfrak g},{\mathfrak t})}\) and \({\mathfrak t}\) is still Poisson commutative. We denote it by \({\tilde{{\mathcal {Z}}}}=\textsf{alg}\langle {\mathcal {Z}}_{({\mathfrak g},{\mathfrak t})},{\mathfrak t}\rangle \). By [13, Theorem 3.2], \({\tilde{{\mathcal {Z}}}}\) is a polynomial algebra, it is a maximal Poisson commutative subalgebra and \({\mathrm {tr.deg\,}}{\tilde{{\mathcal {Z}}}}={\varvec{b}}({\mathfrak g})\). Results of Section 3 in [13] demonstrate that \({\tilde{{\mathcal {Z}}}}\) is closely related to Mishchenko–Fomenko subalgebras. One of the goals of this paper is to further elaborate on these relations.

Let \({\mathfrak m}\subset {\mathfrak g}\) be the \({\mathfrak t}\)-stable complement of \({\mathfrak t}\) in \({\mathfrak g}\). Then \({\mathfrak t}^*\) (resp. \({\mathfrak m}^*\)) is identified with the annihilator \({\mathrm {Ann\,}}\!({\mathfrak m})\subset {\mathfrak g}^*\) (resp. \({\mathrm {Ann\,}}\!({\mathfrak t})\subset {\mathfrak g}^*\)). Our first result is that, for any \(\gamma \in {\mathfrak t}^*\), the restrictions of \({\tilde{{\mathcal {Z}}}}\) and \((\mathcal{M}\mathcal{F})_\gamma \) to \(\gamma +{\mathfrak m}^*\) coincide, see Theorem 2.1.

There is a Poisson bracket \(\{\,\,,\,\}_{(\gamma )}\) on \(\Bbbk [\gamma +{\mathfrak m}^*]^{\mathfrak t}\cong {\mathcal {S}}({\mathfrak m})^{\mathfrak t}\), inherited from \({\mathcal {S}}({\mathfrak g})\). The rank of this bracket is equal to \(\dim {\mathfrak g}-3{\mathsf {rk\,}}{\mathfrak g}\) and if \(\mathcal {B}\) is a Poisson commutative subalgebra of \(({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}},\{\,\,,\,\}_{(\gamma )})\), then \({\mathrm {tr.deg\,}}\mathcal {B}\leqslant \frac{1}{2}(\dim {\mathfrak g}-{\mathsf {rk\,}}{\mathfrak g})\), see Sect. 2.1. We show that if \(\gamma \) is a regular point of \({\mathfrak g}^*\), then \({\tilde{{\mathcal {Z}}}}|_{\gamma +{\mathfrak m}^*}\) is a maximal Poisson commutative subalgebra of \(({\mathcal {S}}({\mathfrak m})^{\mathfrak t},\{\,\,,\,\}_{(\gamma )})\) of transcendence degree \( \frac{1}{2}(\dim {\mathfrak g}-{\mathsf {rk\,}}{\mathfrak g})\).

If \(\mu \in {\mathfrak t}^*\) is regular in \({\mathfrak g}^*\), then \((\mathcal{M}\mathcal{F})_\mu \subset {\mathcal {S}}({\mathfrak g})^{\mathfrak t}\) and the component of grade 2 in \((\mathcal{M}\mathcal{F})_\mu \) equals \({\mathcal {S}}^2({\mathfrak t})\oplus {\mathbb V}(\mu )\), where \({\mathbb V}(\mu )\) is a certain subspace of dimension \({\mathsf {rk\,}}\![{\mathfrak g},{\mathfrak g}]\). There is a natural choice for such a \({\mathbb V}(\mu )\), which is explicitly described by Vinberg [19]. We recall it in Sect. 2. In [14], Rybnikov proved that if \(\mu \) is generic enough, then \((\mathcal{M}\mathcal{F})_{\mu }\) is equal to the Poisson centraliser of \({\mathbb V}(\mu )\) in \({\mathcal {S}}({\mathfrak g})^{\mathfrak t}\). His proof goes through for all simple types, but does not apply to all regular points. We prove that in type A, the centraliser description holds for all regular \(\mu \), see Proposition 2.8.

An interesting task is to produce a quantisation of \({\tilde{{\mathcal {Z}}}}\), i.e., a commutative subalgebra \(\mathscr {Z}\) of the enveloping algebra \({\mathcal {U}}({\mathfrak g})\) such that \({\mathrm {gr\,}}\!(\mathscr {Z}) \subset {\mathcal {S}}({\mathfrak g})\) coincides with \({\tilde{{\mathcal {Z}}}}\). In case of \((\mathcal{M}\mathcal{F})_{\mu }\), the quantisation problem was raised in [19]. A solution, the quantum Mishchenko–Fomenko subalgebra \(\mathcal {F}_\mu \subset {\mathcal {U}}({\mathfrak g})\), is obtained in [15] and [4]. These subalgebras are studied in [3, 4, 8, 16, 18]. For \({\mathfrak g}=\mathfrak {gl}_n\), we quantise \({\tilde{{\mathcal {Z}}}}\) in Sect. 3, see Theorem 3.2. For any \({\mathfrak g}\), we can lift quadratic in \({\mathfrak m}\) elements of \({\tilde{{\mathcal {Z}}}}\) to commuting elements of \({\mathcal {U}}({\mathfrak g})\). In the context of Sect. 3, we have found many similarities with quantum counterparts of Mishchenko–Fomenko subalgebras. This may be an indication that the algebra \({\tilde{{\mathcal {Z}}}}\) always has a quantisation.

2 Preliminaries on the coadjoint action and Poisson-commutativity

Let \({\mathfrak q}\) be a Lie algebra over \(\Bbbk \). Let \({\mathfrak q}^\xi \) denote the stabiliser in \({\mathfrak q}\) of \(\xi \in {\mathfrak q}^*\). The index of \({\mathfrak q}\), \({\mathrm {ind\,}}{\mathfrak q}\), is defined by \({\mathrm {ind\,}}{\mathfrak q}=\min _{\xi \in {\mathfrak q}^*} \dim {\mathfrak q}^\xi \). The set of regular elements of \({\mathfrak q}^*\) is

$$\begin{aligned} {\mathfrak q}^*_\textsf{reg}=\{\eta \in {\mathfrak q}^*\mid \dim {\mathfrak q}^\eta ={\mathrm {ind\,}}{\mathfrak q}\} . \end{aligned}$$
(1.1)

Then \({\mathfrak q}^*_\textsf{sing}={\mathfrak q}^*\setminus {\mathfrak q}^*_\textsf{reg}\). Set further \({\varvec{b}}(\mathfrak q)=(\dim {\mathfrak q}+{\mathrm {ind\,}}{\mathfrak q})/2\). If \(\mathfrak q=\mathfrak g\) is reductive, then \({\mathrm {ind\,}}\mathfrak g={\mathsf {rk\,}}\mathfrak g\) and \({\varvec{b}}(\mathfrak g)\) is the dimension of a Borel subalgebra of \(\mathfrak g\).

For any \(\gamma \in {\mathfrak q}^*\), one defines the Poisson bracket \(\{\,,\,\}_{\gamma }\) on \({\mathfrak q}^*\) by \(\{\xi ,\eta \}_{\gamma }=\gamma ([\xi ,\eta ])\) for \(\xi ,\eta \in {\mathfrak q}\). This new bracket is compatible with the standard Lie–Poisson bracket \(\{\,\,,\,\}\) on \({\mathcal {S}}(\mathfrak q)\), i.e., any linear combination of \(\{\,\,,\,\}\) and \(\{\,\,,\,\}_{\gamma }\) is again a Poisson brackets. For more details, see [2, Sect. 1.8.3]. There is a well-known method, the Lenard–Magri scheme, for constructing “large” Poisson commutative subalgebras of \({\mathcal {S}}({\mathfrak q})\), which is related to compatible brackets, see e.g. [5]. In this way, one obtains \((\mathcal{M}\mathcal{F})_\gamma \) from the pair \((\{\,\,,\,\},\{\,\,,\,\}_{\gamma })\).

In [13], starting form an \(\mathfrak f\)-stable decomposition \(\mathfrak q=\mathfrak f\oplus \mathfrak m\) with \([\mathfrak f,\mathfrak f]=0\), a Poisson commutative subalgebra \({\mathcal {Z}}_{({\mathfrak q},{\mathfrak f})}\subset {\mathcal {S}}({\mathfrak q})\) was constructed. From now on, we consider a particularly interesting case, where \(\mathfrak q=\mathfrak g={\mathsf {Lie\,}}G\) is a reductive Lie algebra and \(\mathfrak f={\mathfrak t}\) is a Cartan subalgebra. Set \(l={\mathsf {rk\,}}{\mathfrak g}\).

The algebra \({\mathcal {S}}({\mathfrak g})^{\mathfrak g}\) is a polynomial ring. Let \(\{H_1,\dots ,H_l\}\) be a set of homogeneous algebraically independent generators of \({\mathcal {S}}({\mathfrak g})^{\mathfrak g}\) with \(\deg H_j=d_j\). We have \(\sum _{j=1}^l d_j={\varvec{b}}({\mathfrak g})\). Let \({\mathfrak g}={\mathfrak u}\oplus {\mathfrak t}\oplus {\mathfrak u}^-\) be a fixed triangular decomposition. Set \({\mathfrak m}={\mathfrak u}\oplus {\mathfrak u}^-\). The vector space decomposition \({\mathfrak g}={\mathfrak t}\oplus {\mathfrak m}\) provides the bi-homogeneous decomposition of each \(H_j\):

$$\begin{aligned} H_j=\sum _{i=0}^{\textit{d}_j} (H_j)_{(i,\textit{d}_j-i)}, \end{aligned}$$

where \((H_j)_{(i,\textit{d}_j-i)}\in {\mathcal {S}}^i({\mathfrak t})\otimes {\mathcal {S}}^{\textit{d}_j-i}({\mathfrak m})\subset {\mathcal {S}}^{d_j}({\mathfrak g})\). Then we say that \(d_j-i\) is the \({\mathfrak m}\)-degree of \((H_j)_{(i,d_j-i)}\). Now, \({\mathcal {Z}}:={\mathcal {Z}}_{({\mathfrak g},{\mathfrak t})}\) is the algebra generated by

$$\begin{aligned} \{ (H_j)_{(i,d_j-i)} \mid j=1,\dots ,l; i=0,1,\dots , d_j-3, d_j-2, d_j\}, \end{aligned}$$

see [13]. The total number of functions in this family equals \(\sum _{j=1}^l (d_j+1) - l={\varvec{b}}(\mathfrak g)\) and they are algebraically independent [13]. Replacing the elements \((H_j)_{(d_j,0)}\in {\mathcal {S}}^{d_j}({\mathfrak t})\) with a basis of \({\mathfrak t}\), we obtain a larger subalgebra, denoted \({\tilde{{\mathcal {Z}}}}\), which is still polynomial and Poisson commutative.

2.1 Notation and conventions

For a subalgebra \(A\subset {\mathcal {S}}({\mathfrak g})\) and \(\gamma \in {\mathfrak g}^*\), we set \(\textit{d}_\gamma A=\{\textit{d}_\gamma F \mid F\in A\}\).

Given a Poisson algebra \(\mathcal {A}\) and \(a\in \mathcal {A}\), let \(\mathcal {Z}_a\mathcal {A}=\{F\in \mathcal {A}\mid \{a,F\}=0\}\) denote the Poisson centraliser of a in \(\mathcal {A}\).

Let \(\mathfrak l\subset \mathfrak g\) be a Lie subalgebra. Then \({\mathcal {S}}(\mathfrak g)^{\mathfrak l}=\{F\in {\mathcal {S}}(\mathfrak g)\mid \{\xi ,F\}=0 \ \forall \xi \in \mathfrak l\}\) and \({\mathcal {U}}(\mathfrak g)^{\mathfrak l}\) stands for the centraliser \(\{X\in {\mathcal {U}}(\mathfrak g)\mid [\xi ,X]=0 \ \forall \xi \in \mathfrak l\}\) of \(\mathfrak l\) in \({\mathcal {U}}(\mathfrak g)\).

For an irreducible affine variety Y over \(\Bbbk \), we let \(\Bbbk [Y]\) stand for the ring of regular functions on Y and \(\Bbbk (Y)={\textrm{Quot}\,}\Bbbk [Y]\) for the field of rational functions on Y. A statement that a certain assertion holds for generic points of Y (or for generic orbits on Y) means that this assertion holds for all points of a nonempty open subset \(U\subset Y\) (for all orbits intersecting U).

If \({\mathfrak g}={\mathfrak f}\oplus {\mathfrak m}\) and \(\xi \in {\mathfrak g}\), then \(\xi =\xi _{\mathfrak f}+\xi _{{\mathfrak m}}\), where \(\xi _{\mathfrak f}\in \mathfrak f\) and \(\xi _{{\mathfrak m}}\in {\mathfrak m}\).

Let \(\Delta \) be the set of roots of \(({\mathfrak g},{\mathfrak t})\) and \(\Delta ^+\subset \Delta \) the subset of positive roots corresponding to \({\mathfrak u}\). For \(\alpha \in \Delta \), let \(e_\alpha \in \mathfrak g_{\alpha }\) be a nonzero root vector. We let \(h_\alpha \in [{\mathfrak g}_\alpha ,{\mathfrak g}_{-\alpha }]\subset {\mathfrak t}\) be such that \(\alpha (h_\alpha )=2\).

We say that \({\mathfrak g}\) is of type A, if \({\mathfrak g}=\mathfrak {sl}_{l+1}\). In that case, we fix \({\mathfrak t}\) as the subspace of diagonal matrices and use the standard linear functions \(\varepsilon _i\in {\mathfrak t}^*\) such that \(\varepsilon _i(\textrm{diag}(a_1,\ldots ,a_{l+1}))=a_i\). We fix the standard triangular decomposition with \(\Delta ^+=\{\varepsilon _i-\varepsilon _j\mid i<j\}\). For \(1\leqslant i,j\leqslant n\), let \(E_{ij}\in \mathfrak {gl}_n\) be a matrix unit (elementary matrix).

If \({\mathfrak g}\) is semisimple, then \({\mathfrak g}\) is identified with \({\mathfrak g}^*\) via the Killing form \(\kappa \).

3 Partial localisations and reductions by the action of \({\mathfrak t}\)

Results of [13] show that \({\tilde{{\mathcal {Z}}}}=\textsf{alg}\langle {\mathcal {Z}}_{({\mathfrak g},{\mathfrak t})},{\mathfrak t}\rangle \) is closely related to Mishchenko–Fomenko subalgebras. Our goal now is to elaborate on this relation.

Recall that \({\mathfrak m}=\mathfrak u\oplus \mathfrak u^-\). Consider \(\mathcal {A}={\mathcal {S}}({\mathfrak g})^{\mathfrak t}\otimes _{{\mathcal {S}}({\mathfrak t})}\Bbbk ({\mathfrak t}^*)\subset \Bbbk ({\mathfrak g}^*)^{{\mathfrak t}}\) as a ring of \({\mathcal {S}}({\mathfrak m})^{\mathfrak t}\)-valued rational functions on \({\mathfrak t}^*\); here \(FM\otimes {\tilde{F}}=M\otimes F{\tilde{F}}\) for \(M\in {\mathcal {S}}({\mathfrak m})^{\mathfrak t}, F\in {\mathcal {S}}({\mathfrak t})\cong \Bbbk [{\mathfrak t}^*], {\tilde{F}}\in \Bbbk ({\mathfrak t}^*)\) and \((FM\otimes {\tilde{F}})(\mu )=F(\mu ){\tilde{F}}(\mu ) M\), if \(\mu \in {\mathfrak t}^*\). Since \(\{{\mathcal {S}}({\mathfrak g})^{\mathfrak t},{\mathcal {S}}({\mathfrak g})^{\mathfrak t}\}\subset {\mathcal {S}}({\mathfrak g})^{\mathfrak t}\) and \(\{{\mathcal {S}}({\mathfrak g})^{\mathfrak t},{\mathfrak t}\}=0\), the ring \(\mathcal {A}\) inherits a Poisson structure from \({\mathcal {S}}({\mathfrak g})\) and this Poisson structure is \(\Bbbk ({\mathfrak t}^*)\)-linear. For \(\mu \in {\mathfrak t}^*\), set

$$\begin{aligned} \mathcal {A}_\mu =\{ A\in \mathcal {A}\mid A(\mu ) \,\text { is well-defined}\,\}. \end{aligned}$$

Let \(\psi _\mu \!: \mathcal {A}_\mu \rightarrow {\mathcal {S}}(\mathfrak m)^{\mathfrak t}\) be the evaluation homomorphism. Then \({\mathcal {S}}({\mathfrak g})^{\mathfrak t}\subset \mathcal {A}_\mu \) for each \(\mu \) and on \({\mathcal {S}}({\mathfrak g})^{{\mathfrak t}}\cong \Bbbk [{\mathfrak g}^*]^{\mathfrak t}\), the map \(\psi _\mu \) coincides with the restriction homomorphism

$$\begin{aligned} \Bbbk [{\mathfrak g}^*]^{\mathfrak t}\rightarrow \Bbbk [\mu +{\mathfrak m}^*]^{\mathfrak t}\cong {\mathcal {S}}({\mathfrak m})^{\mathfrak t}. \end{aligned}$$

We define a bi-linear map \(\{\,\,,\,\}_{(\mu )}\!: {\mathfrak m}\times {\mathfrak m}\rightarrow {\mathfrak m}\oplus \Bbbk \) by

$$\begin{aligned} \{x,y\}_{(\mu )}:=\psi _\mu ([x,y])=[x,y]_{\mathfrak m}+\mu ([x,y]_{{\mathfrak t}}) \ \text { for } \ x,y\in {\mathfrak m}\end{aligned}$$

and then extend it to a bi-linear operation on \({\mathcal {S}}(\mathfrak m)^{\mathfrak t}\) using the Leibniz rule. For XY in \({\mathcal {S}}(\mathfrak m)^{\mathfrak t}\subset \mathcal {A}\), we have

$$\begin{aligned} \psi _{\mu }(\{X,Y\})=\{X,Y\}_{(\mu )}=\{\psi _\mu (X),\psi _\mu (Y)\}_{(\mu )}. \end{aligned}$$

Using the \(\Bbbk ({\mathfrak t}^*)\)-linearity of the Poisson bracket on \(\mathcal {A}\), we conclude that \(\{\,\,,\,\}_{(\mu )}\) is a Poisson structure on \({\mathcal {S}}(\mathfrak m)^{\mathfrak t}\) and \(\psi _\mu \) is a Poisson homomorphism with respect to \(\{\,\,,\,\}_{(\mu )}\).

The ring \(\mathcal {A}=\bigoplus _{N\geqslant 0}\mathcal {A}_N\) is graded by the degree in \({\mathfrak m}\) with \(\mathcal {A}_N={\mathcal {S}}^N({\mathfrak m})^{\mathfrak t}\otimes _{\Bbbk }\Bbbk ({\mathfrak t}^*)\) and \(\psi _\mu (\mathcal {A}_\mu \cap \mathcal {A}_N)={\mathcal {S}}^N({\mathfrak m})^{\mathfrak t}\). Since the Poisson bracket on \(\mathcal {A}\) is \(\Bbbk ({\mathfrak t}^*)\)-linear, for any \(A\in {\mathcal {S}}({\mathfrak g})^{{\mathfrak t}}\) and any \(N\geqslant 0\), the subset \(\mathcal {Z}_A(\mathcal {A}_{\leqslant N})=\mathcal {Z}_A\mathcal {A}\cap (\bigoplus _{i=0}^N \mathcal {A}_{i})\) is a vector space over \(\Bbbk ({\mathfrak t}^*)\). It is a subspace of a finite-dimensional \(\Bbbk ({\mathfrak t}^*)\)-space \(\mathcal {A}_{\leqslant N}\). Evaluating the defining equations of the centraliser \( \mathcal {Z}_A (\mathcal {A}_{\leqslant N})\) at \(\mu \in {\mathfrak t}^*\), we obtain

$$\begin{aligned} \dim _{\Bbbk ({\mathfrak t}^*)} \mathcal {Z}_A (\mathcal {A}_{\leqslant N}) = \dim \left( \mathcal {Z}_{\psi _\mu (A)}({\mathcal {S}}({\mathfrak m})^{\mathfrak t},\{\,\,,\,\}_{(\mu )})\cap {\mathcal {S}}^{\leqslant N}({\mathfrak m})\right) , \end{aligned}$$
(2.1)

whenever \(\mu \) is generic enough.

Recall that we identify \({\mathfrak t}^*\) with \({\mathrm {Ann\,}}\!({\mathfrak m})\subset {\mathfrak g}^*\).

Theorem 2.1

If \(\mu \in {\mathfrak t}^*\), then \(\psi _\mu ({\tilde{{\mathcal {Z}}}})=\psi _\mu ((\mathcal{M}\mathcal{F})_{\mu })\).

Proof

Suppose first that \(\mu \ne 0\). We fix \(h\in {\mathfrak t}\) such that \(\mu (h)=1\) and write \(\mathfrak g =\Bbbk h\oplus \ker \mu \), where \({\mathfrak m}\subset \ker \mu \). Let \(H\in {\mathcal {S}}({\mathfrak g})^{{\mathfrak g}}\) be homogeneous with \(\deg H=d\). We decompose H as a sum

$$\begin{aligned} H=H_{0} h^{d} + H_{1}h^{d-1}+\cdots + H_{k}h^k +\cdots + H_{d}, \end{aligned}$$

where \(H_{k}\in {\mathcal {S}}^k(\ker \mu )\). By the choice of h, we have

$$\begin{aligned} \partial _\mu ^k H=\sum _{r=k}^{d} r(r-1)\ldots (r-k+1)h^{r-k}H_{d-r} \end{aligned}$$

and

$$\begin{aligned} \psi _\mu (\partial _\mu ^k H)= \sum _{r=k}^{d} r(r-1)\ldots (r-k+1) \psi _\mu (H_{d-r}). \end{aligned}$$

Therefore \(\psi _\mu ((\mathcal {M}\mathcal {F})_{\mu })=\textsf {alg}\langle \psi _\mu (H_{d-r}) \mid d\geqslant 1, H\in {\mathcal {S}}^d({\mathfrak g})^{\mathfrak g}, 0\leqslant r < d \rangle \).

Let \(M_{d-i}\in {\mathcal {S}}^{d-i}({\mathfrak m})\) be such that the bi-homogeneous component \(H_{i,d-i}\in {\mathcal {S}}^i({\mathfrak t})\otimes {\mathcal {S}}^{d-i}({\mathfrak m})\) of H lies in \(h^i M_{d-i} + (\ker \mu \cap {\mathfrak t}){\mathcal {S}}({\mathfrak g})\). Then

$$\begin{aligned} \psi _\mu (H_{i,d-i})= M_{d-i} =\psi _\mu (h^i H_{d-i})=\psi _\mu (H_{d-i}) \end{aligned}$$

for all i. Since \({\tilde{{\mathcal {Z}}}}\) is generated by \({\mathfrak t}\) and \(H_{i,d-i}\) with \(H\in {\mathcal {S}}^d({\mathfrak g})^{\mathfrak g}\), we are done for \(\mu \ne 0\).

If \(\mu =0\), then \((\mathcal{M}\mathcal{F})_{0}={\mathcal {S}}(\mathfrak g)^{\mathfrak g}\) and \(\psi _0(H)=\psi _0(H_{0,d})\). For other generators of \({\tilde{{\mathcal {Z}}}}\), we have \(\psi _0({\mathfrak t})=0\) and \(\psi _0(H_{i,d-i})=0\), whenever \(i>0\). \(\square \)

A Mishchenko–Fomenko subalgebra is Poisson commutative and \(\psi _\mu \) is a Poisson homomorphism; hence

$$\begin{aligned} 0\!= & {} \!\psi _\mu (\{ (\mathcal{M}\mathcal{F})_{\mu }, (\mathcal{M}\mathcal{F})_{\mu } \})\! =\!\{ \psi _\mu ((\mathcal{M}\mathcal{F})_{\mu }),\psi _\mu ( (\mathcal{M}\mathcal{F})_{\mu }) \}_{(\mu )}\\= & {} \{\psi _{\mu } ({\tilde{{\mathcal {Z}}}}),\psi _\mu ({\tilde{{\mathcal {Z}}}})\}_{(\mu )}\!=\!\psi _\mu (\{{\tilde{{\mathcal {Z}}}},{\tilde{{\mathcal {Z}}}}\}) \end{aligned}$$

for each \(\mu \in {\mathfrak t}^*\). Since for any \(A\in \mathcal {A}\setminus \{0\}\), the image \(\psi _\mu (A)\) is well-defined and nonzero for almost all \(\mu \), Theorem 2.1 provides a new proof for the fact that \(\{{\tilde{{\mathcal {Z}}}},{\tilde{{\mathcal {Z}}}}\}=0\).

Assume that \(\mu \in {\mathfrak t}^* \cap {\mathfrak g}^*_\textsf{reg}\). Then the polynomials \(\partial _\mu ^k H_j\) with \(1\leqslant j\leqslant l\) and \(0\leqslant k<d_j\) are algebraically independent, cf. [11]. In particular, \(\{\partial _\mu ^{d_j-1}H_j \mid 1\leqslant j\leqslant l\}\) is a basis of \({\mathfrak t}\) and \({\mathfrak t}\subset (\mathcal{M}\mathcal{F})_{\mu }\). This fact follows also from the Kostant regularity criterion for \({\mathfrak g}\) [6, Theorem 9],

$$\begin{aligned} \langle \textit{d}_\xi H_j \mid 1\leqslant j\leqslant l\rangle _{\Bbbk }={\mathfrak g}^\xi \ \text { if and only if }\ \xi \in {\mathfrak g}^*_\textsf{reg}, \end{aligned}$$
(2.2)

since \(\textit{d}_\mu H_j=\frac{1}{(d_j-1)!}\partial _{\mu }^{d_j-1}H_j\).

Quadratic elements of \((\mathcal{M}\mathcal{F})_{\mu }\) are of particular importance. If not stated otherwise, assume that \({\mathfrak g}\) is semisimple. Set \(\varvec{f}_j:=\varvec{f}_j(\mu ):=\psi _\mu ((H_j)_{(d_j-2,2)})\). Then

$$\begin{aligned} \partial _{\mu }^{d_j-2} H_j \in (d_j-2)!\varvec{f}_j +{\mathcal {S}}^2({\mathfrak t}). \end{aligned}$$
(2.3)

Since \({\mathfrak t}\subset (\mathcal{M}\mathcal{F})_{\mu }\), we have \(\varvec{f}_j \in (\mathcal{M}\mathcal{F})_{\mu }\) for all j. Furthermore, the component of grade 2 in \((\mathcal{M}\mathcal{F})_{\mu }\) is equal to \({\mathcal {S}}^2({\mathfrak t})\oplus {\mathbb V}(\mu )\), where \({\mathbb V}(\mu )=\langle \varvec{f}_j \mid 1\leqslant j\leqslant l\rangle _{\Bbbk }\). If \(\mu \) is generic enough, then \((\mathcal{M}\mathcal{F})_{\mu }\) is equal to the Poisson centraliser

$$\begin{aligned} \mathcal {Z}_{{\mathbb V}(\mu )}:=\{F\in {\mathcal {S}}({\mathfrak g})^{\mathfrak t}\mid \{F,\varvec{f}_j\}=0 \ \forall j\} \end{aligned}$$

by [14, Theorem 1].

Remark 2.2

An explicit description of the elements \(\varvec{f}_j\) is crucial for the considerations in [14, 19] and many others. We present a quick elementary argument that produces such a description. Set \(h_j=\textit{d}_\mu H_j\). Similar to the proof of Theorem 2.1, write \(\mathfrak g=\Bbbk h\oplus \ker \mu \), where \(\mu (h)=1\). Then

$$\begin{aligned} H_j=c h^{d_j}+h^{d_j-1}h' + h^{d_j-2} H'+\sum _{\alpha \in \Delta ^+} C_\alpha h^{d_j-2} e_\alpha e_{-\alpha } +\sum _{k=3}^{d_j} h^{d_j-k} H_{j,k}, \end{aligned}$$

where \(c,C_\alpha \) are scalars, \(h',H',H_{j,k}\in {\mathcal {S}}(\ker \mu )\), and \(h'\in {\mathfrak t}\), \(H'\in {\mathcal {S}}^2({\mathfrak t})\). In this notation, \(h_j=c d_j h+h'\). Since \(H_j\in {\mathcal {S}}({\mathfrak g})^{\mathfrak g}\), we have \(\{e_\alpha ,H_j\}=0\) for each \(\alpha \in \Delta ^+\). Note that \(\{e_\alpha ,e_{-\alpha }\}=\mu (h_\alpha )h+h''\), where \(h''\in (\ker \mu \cap {\mathfrak t})\). Considering the terms of \(\{e_\alpha ,H_j\}\) that lie in \(\Bbbk h^{d_j-1} {\mathfrak m}\), and then necessarily in \(\Bbbk h^{d_j-1} e_{\alpha }\), we obtain

$$\begin{aligned} -\alpha (c d_j h+h') + C_\alpha \mu (h_\alpha )=0 \end{aligned}$$

for each positive root \(\alpha \). Since \(\mu \in {\mathfrak t}^* \cap {\mathfrak g}^*_\textsf{reg}\), we have \(\mu (h_\alpha )\ne 0\) for each \(\alpha \in \Delta ^+\). Hence \(C_\alpha =\dfrac{\alpha (h_j)}{\mu (h_\alpha )}\). Here \(\varvec{f}_j=\sum _{\alpha \in \Delta ^+} C_\alpha e_\alpha e_{-\alpha }\).

The discussion in Remark 2.2 confirms the description obtained by Vinberg in [19] :

$$\begin{aligned} {{\mathbb V}(\mu )} = \left\{ \sum _{\alpha \in \Delta ^+} \frac{\alpha (h)}{\mu (h_\alpha )} e_\alpha e_{-\alpha } \mid h\in {\mathfrak t}\right\} , \end{aligned}$$
(2.4)

whenever \(\mu \in {\mathfrak t}^* \cap {\mathfrak g}^*_\textsf{reg}\).

3.1 Bi-linear operations \(\{\,\,,\,\}_{(\gamma )}\) and \(\{\,\,,\,\}_{\mathfrak m}\)

For any \(\gamma \in {\mathfrak t}^*\), we have defined a Poisson bracket \(\{\,\,,\,\}_{(\gamma )}=\{\,\,,\,\}_{\mathfrak m}+\{\,\,,\,\}_\gamma \) on \({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\) as a sum of two bi-linear operations. The second summand is the restriction to \({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\) of the Poisson bracket “with frozen argument” \(\{\,\,,\,\}_\gamma \), which is defined on \({\mathcal {S}}(\mathfrak g)\). Note that the operation \(\{\,\,,\,\}_{\mathfrak m}\) is not a Poisson bracket on \({\mathcal {S}}({\mathfrak m})\), because it does not satisfy the Jacobi identity. However, in case \(\gamma =0\), we obtain \(\{\,\,,\,\}_{\mathfrak m}=\{\,\,,\,\}_{(0)}\), which is a Poisson bracket on \({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\).

Let \(\widehat{x}\) be a skew-symmetric form on \(\mathfrak g\) associated with \(x\in \mathfrak g^*\), i.e., \(\widehat{x}(\xi ,\eta )=x([\xi ,\eta ])\) if \(\xi ,\eta \in \mathfrak g\). For a Poisson structure, one defines its Poisson tensor (bivector) \(\pi \) by the property that \(\{F,H\}=\pi (\textit{d}F\wedge \textit{d}H)\) for functions F and H. In this terms, \(\widehat{x}=\pi (x)\), if \(\pi \) is the Poisson tensor of \(\{\,,\,\}\). In general, one says that the rank of the Poisson structure is equal to the rank \({\mathsf {rk\,}}\pi \) of its Poisson tensor. Then \({\mathsf {rk\,}}\pi \) is the maximal dimension of a symplectic leave of \(\pi \), see e.g. [2, Chapter 1].

Proposition 2.3

Let \(\mathcal {B}\) be a Poisson commutative subalgebra of \(({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}},\{\,\,,\,\}_{(\gamma )})\). Then for any \(\gamma \), we have \({\mathrm {tr.deg\,}}\mathcal {B}\leqslant \dim \mathfrak u\).

Proof

We identify \({\mathcal {S}}({\mathfrak m})\) with \(\Bbbk [\gamma +{\mathfrak m}^*]\). Then

$$\begin{aligned} \{F,H\}_{(\gamma )}(x) =x([\textit{d}_x F,\textit{d}_x H])=\widehat{x}(\textit{d}_x F,\textit{d}_x H) \ \ \text { for all } \ \ x\in \gamma +{\mathfrak m}^*, \, F,H\in {\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}. \end{aligned}$$

Since \(\{\mathcal {B},\mathcal {B}\}_{(\gamma )}=0\), the subspace \(\textit{d}_x \mathcal {B}\) is isotropic w.r.t. \(\widehat{x}\) for any \(x\in \gamma +{\mathfrak m}^*\). Furthermore \(\widehat{x}(\textit{d}_x {\mathcal {S}}({\mathfrak g})^{{\mathfrak t}},\mathfrak t)=0\).

Let \(T\subset G\) be the torus with \({\mathsf {Lie\,}}T={\mathfrak t}\). Generic T-orbits on \(\gamma +{\mathfrak m}^*\) are closed, hence they are separated by the regular T-invariants \(\Bbbk [\gamma +{\mathfrak m}^*]^T\cong {\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\). Thus \(\textit{d}_x{\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\subset \textit{T}_x^{\,*}(\gamma +{\mathfrak m}^*)\) is the annihilator of the tangent space \(\textit{T}_x(Tx)={\textrm{ad}}^*({\mathfrak t})x\) of the orbit Tx for a generic \(x\in \gamma +{\mathfrak m}^*\). The orthogonal complement \({\mathfrak t}^{\perp _{\widehat{x}}}\) of \({\mathfrak t}\) in \({\mathfrak g}\) w.r.t. \(\widehat{x}\) is the subset

$$\begin{aligned} \{\xi \in \mathfrak g \mid x([\xi ,{\mathfrak t}])=0\}=\{\xi \in \mathfrak g\mid {\textrm{ad}}^*({\mathfrak t})x \ \hbox { annihilates \,}\ \xi \}. \end{aligned}$$

Here we have \({\mathfrak t}^{\perp _{\widehat{x}}}={\mathfrak t}\oplus \textit{d}_x {\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}=\textit{d}_x{\mathcal {S}}(\mathfrak g)^{{\mathfrak t}}\). Note that \({\mathsf {rk\,}}\widehat{x}=\dim {\mathfrak g}- l\).

Keeping the assumption that x is generic, we have \({\mathfrak t}\cap \ker \widehat{x}=0\). Since \(\widehat{x}({\mathfrak t},{\mathfrak t})=0\), the rank of  \(\widehat{x}|_{{\mathfrak t}^{\perp _{\widehat{x}}}}\) is equal to \({\mathsf {rk\,}}\widehat{x} -2\dim {\mathfrak t}\). Thus

$$\begin{aligned} {\mathsf {rk\,}}\!(\widehat{x}|_{\textit{d}_x {\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}})={\mathsf {rk\,}}\!(\widehat{x}|_{\textit{d}_x {\mathcal {S}}({\mathfrak g})^{{\mathfrak t}}})=\dim \mathfrak g-3l=2(\dim \mathfrak u-l). \end{aligned}$$
(2.5)

Since \(\dim \textit{d}_x {\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}=\dim {\mathfrak g}-2l\), we obtain \(\dim \textit{d}_x\mathcal {B}\leqslant (\dim \mathfrak u-l) + (\dim {\mathfrak g}-2\,l - (\dim \mathfrak g-3\,l))= \dim \mathfrak u.\) Thus, \({\mathrm {tr.deg\,}}\mathcal {B}\leqslant \dim \mathfrak u\). \(\square \)

The equality (2.5) shows that the rank of the Poisson bracket \(\{\,\,,\,\}_{(\gamma )}\) on \({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\) is equal to \(2(\dim \mathfrak u-l)=\dim {\mathfrak m}-2\,l\).

Remark 2.4

Set \(Y_\gamma :=\gamma +{\mathfrak m}^*\). The algebra \({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\cong \Bbbk [Y_\gamma ]^T\) is the algebra of regular functions on the affine variety \(Y_\gamma /\!\!/T\)  and the bracket \(\{\,,\,\}_{(\gamma )}\) on \({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\) is obtained by the Hamiltonian reduction w.r.t. the restriction \({\mathfrak g}^*\rightarrow {\mathfrak t}^*\). We have \(\dim Y_\gamma /\!\!/T={\mathrm {tr.deg\,}}{\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}=\dim {\mathfrak g}-2 l\). The bound for \({\mathrm {tr.deg\,}}\mathcal {B}\) given by Proposition 2.3 is of the form \(\dim Y_\gamma /\!\!/T -\frac{1}{2} (\dim {\mathfrak m}-2l)\), where \(\dim {\mathfrak m}-2l\) is the rank of the Poisson structure in question. This is a general upper bound, existing for any Poisson algebra. The equality (2.5) can be deduced from the fact that generic symplectic leaves of \(Y_\gamma /\!\!/T\) are of the form \((Gx\cap Y_\gamma )/\!\!/T\) with \(\dim (Gx\cap Y_\gamma )/\!\!/T = \dim {\mathfrak g}-3\dim {\mathfrak t}\).

Theorem 2.5

Suppose that \(\mu \in {\mathfrak t}^* \cap {\mathfrak g}^*_\textsf{reg}\). Then for any \(\gamma \in {\mathfrak t}^*\), we have \({\mathrm {tr.deg\,}}\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })=\dim \mathfrak u\), and \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\) is a maximal Poisson commutative subalgebra of \(({\mathcal {S}}({\mathfrak m})^{\mathfrak t},\{\,\,,\,\}_{(\gamma )})\).

Proof

Let \(\{e,h,f\}\subset \mathfrak g\) be a principal \(\mathfrak {sl}_2\)-triple such that \(h\in {\mathfrak t}\), \(e\in \mathfrak u\), \(f\in \mathfrak u^-\). Set \(\chi _+=\kappa (e,.\,), \chi _-=\kappa (f,.\,)\in \mathfrak g^*\). Since \(e+({\mathfrak t}\oplus \mathfrak u^-)\) and \(f+({\mathfrak t}\oplus \mathfrak u)\) consist of regular elements [6], we have \((\Bbbk \chi _+\oplus \Bbbk \chi _-\oplus \Bbbk \mu ) \cap \mathfrak g^*_\textsf{sing}=0\). Therefore \(\dim \textit{d}_{x}(\mathcal{M}\mathcal{F})_{\mu }={\varvec{b}}(\mathfrak g)\) for any nonzero \(x\in \Bbbk \chi _+\oplus \Bbbk \chi _-\), see e.g. [12, Cor. 1.6  & Lemma 2.1].

For any \(F\in {\mathcal {S}}(\mathfrak g)\) and \(y\in \mathfrak m^*\), we have \(\textit{d}_y F \in \textit{d}_y(\psi _0(F))+{\mathfrak t}\). Hence \(\dim \textit{d}_x \psi _0((\mathcal{M}\mathcal{F})_{\mu })\) is equal to \(\dim \mathfrak u\). Each \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\) is a Poisson commutative subalgebra of \(({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}},\{\,,\,\}_{(\gamma )})\), thereby \({\mathrm {tr.deg\,}}\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\leqslant \dim \mathfrak u\) by Proposition 2.3. Thus, \({\mathrm {tr.deg\,}}\psi _0((\mathcal{M}\mathcal{F})_{\mu })=\dim \mathfrak u\).

The differentials \(\textit{d}_x(\psi _0(\partial _\mu ^k H_j))\) with \(k<d_j-1\) and \(1\leqslant j\leqslant l\) are linearly independent for each \(x\in (\Bbbk \chi _+ \oplus \Bbbk \chi _-){\setminus }\{0\}\). Thus, \({\mathcal J}\cap (\Bbbk \chi _+ \oplus \Bbbk \chi _-)\subset \{0\}\) for the Jacobian subset

$$\begin{aligned} \mathcal {J}=\{y\in {\mathfrak m}^* \mid \bigwedge _{0\leqslant k<d_j-1,\, 1\leqslant j\leqslant l} \textit{d}_y (\psi _0(\partial _\mu ^k H_j))=0\}. \end{aligned}$$

If \(F\in {\mathcal {S}}(\mathfrak g)\) is homogeneous, then \(\psi _0(F)\) is also homogeneous. This applies to each \(\partial _\mu ^k H_j\) and leads to the conclusion that \(\mathcal {J}\) does not contain divisors. By [10, Theorem 1.1], \(\psi _0((\mathcal{M}\mathcal{F})_{\mu })\) is an algebraically closed subalgebra of \({\mathcal {S}}({\mathfrak m})\), i.e., if \(F\in {\mathcal {S}}({\mathfrak m})\) is algebraic over the quotient field \({\textrm{Quot}\,}\psi _0((\mathcal{M}\mathcal{F})_{\mu })\), then \(F\in \psi _0((\mathcal{M}\mathcal{F})_{\mu })\).

Suppose \(\psi _0((\mathcal{M}\mathcal{F})_{\mu })\subset \mathcal {B}\subset ({\mathcal {S}}({\mathfrak m})^{{\mathfrak t}},\{\,\,,\,\}_{(\gamma )})\), where \(\mathcal {B}\) is a Poisson commutative subalgebra. Then \({\mathrm {tr.deg\,}}\mathcal {B}\leqslant \dim \mathfrak u\) by Proposition 2.3. Thereby the inclusion \(\psi _0((\mathcal{M}\mathcal{F})_{\mu })\subset \mathcal {B}\) is an algebraic extension and \(\psi _0((\mathcal{M}\mathcal{F})_{\mu })=\mathcal {B}\). The argument shows also that \(\psi _0((\mathcal{M}\mathcal{F})_{\mu })\) coincides with its Poisson centraliser in \({\mathcal {S}}({\mathfrak m})^{\mathfrak t}\) w.r.t. \(\{\,,\,\}_{{\mathfrak m}}\) and finishes the case \(\gamma =0\).

For any homogeneous \(F\in {\mathcal {S}}(\mathfrak g)\setminus {\mathfrak t}{\mathcal {S}}({\mathfrak g})\), the image \(\psi _0(F)\) is the highest degree component of any \(\psi _\gamma (F)\). In particular, the equality \({\mathrm {tr.deg\,}}\psi _0((\mathcal{M}\mathcal{F})_{\mu })=\dim \mathfrak u\) leads to \({\mathrm {tr.deg\,}}\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\geqslant \dim \mathfrak u\), thereby \({\mathrm {tr.deg\,}}\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })=\dim \mathfrak u\). Assume that \(F\in {\mathcal {S}}({\mathfrak m})^{\mathfrak t}\) commutes with \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\) w.r.t. \(\{\,\,,\,\}_{(\gamma )}\) and does not lie in \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\). Then the highest degree component of F commutes with \(\psi _0((\mathcal{M}\mathcal{F})_{\mu })\) w.r.t. \(\{\,\,,\,\}_{{\mathfrak m}}\), which means that this component lies in \(\psi _0((\mathcal{M}\mathcal{F})_{\mu })\). Then we can reduce the degree of F by subtracting a suitable element of \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\). This standard reduction argument proves that \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\) is a maximal Poisson commutative subalgebra of \(({\mathcal {S}}({\mathfrak m})^{\mathfrak t},\{\,\,,\,\}_{(\gamma )})\) for each \(\gamma \in {\mathfrak t}^*\). \(\square \)

Corollary 2.6

Both, \(\mathcal {B}_1=(\mathcal{M}\mathcal{F})_{\mu }\otimes _{{\mathcal {S}}({\mathfrak t})}\Bbbk ({\mathfrak t}^*)\) and \(\mathcal {B}_2={\tilde{{\mathcal {Z}}}}\otimes _{{\mathcal {S}}({\mathfrak t})}\Bbbk ({\mathfrak t}^*)\), are maximal Poisson commutative subalgebras of \(\mathcal {A}\).

Proof

By the construction, \(\{\mathcal {B}_i,\mathcal {B}_i\}=0\) for both i. Assume that \(\mathcal {B}_i\) is not maximal. Then there is \(a\in \mathcal {A}\setminus \mathcal {B}_i\) such that \(\{a,\mathcal {B}_i\}=0\). For each \(\gamma \in {\mathfrak t}^*\), we have \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })\subset \psi _\gamma (\mathcal {B}_1)\) and \(\psi _\gamma ({\tilde{{\mathcal {Z}}}})\subset \psi _\gamma (\mathcal {B}_2)\). For any \(\gamma \in {\mathfrak t}^*\) such that \(a(\gamma )\) is well-defined, \(\{\psi _\gamma (a),\psi _\gamma (\mathcal {B}_i)\}_{(\gamma )}\) is zero. If \(\gamma \) is regular in \({\mathfrak g}^*\), then \(\psi _\gamma ((\mathcal{M}\mathcal{F})_{\mu })=\psi _\gamma (\mathcal {B}_1)\),  \(\psi _\gamma ({\tilde{{\mathcal {Z}}}})=\psi _\gamma ((\mathcal{M}\mathcal{F})_{\gamma })=\psi _\gamma (\mathcal {B}_2)\), and \(\psi _\gamma (a)\in \psi _\gamma (\mathcal {B}_i)\) by Theorems 2.52.1.

There is \(N\geqslant 0\) such that \(a\in \mathcal {A}_{\leqslant N}\). Here \(\mathcal {A}_{\leqslant N}\) is a finite-dimensional vector space over \(\Bbbk ({\mathfrak t}^*)\) and \(a\not \in \mathcal {A}_{\leqslant N}\cap \mathcal {B}_i\). Then for almost all \(\gamma \), we have \(\psi _\gamma (a)\not \in \psi _\gamma ( \mathcal {A}_{\leqslant N}\cap \mathcal {B}_i)\). The algebra \({\tilde{{\mathcal {Z}}}}\) is generated by bi-homogeneous elements. Hence it is graded by the \({\mathfrak m}\)-degree, \({\tilde{{\mathcal {Z}}}}=\bigoplus _{k\geqslant 0} (\mathcal {A}_k\cap {\tilde{{\mathcal {Z}}}})\). Recall that \(\psi _\gamma (\mathcal {A}_k)={\mathcal {S}}^k({\mathfrak m})^{\mathfrak t}\). Thus, in case \(i=2\),

$$\begin{aligned} \psi _\gamma (\mathcal {B}_2)=\psi _\gamma ({\tilde{{\mathcal {Z}}}})=\bigoplus _{k\geqslant 0} \psi _\gamma (\mathcal {A}_k\cap {\tilde{{\mathcal {Z}}}}), \end{aligned}$$

and we can conclude that \(\psi _\gamma (a)\not \in \psi _\gamma (\mathcal {B}_2)\) for generic \(\gamma \), which is a contradiction.

The algebra \((\mathcal{M}\mathcal{F})_{\mu }\) is not homogeneous in \({\mathfrak m}\). However, the highest \({\mathfrak m}\)-degree components of the generators \(\partial ^k_\mu H_j\) with \(k<d_j-1\) lie in \({\mathcal {S}}({\mathfrak m})\) and are algebraically independent by Theorem 2.5. Therefore, if \(\psi _\gamma (a)\not \in \psi _\gamma ( \mathcal {A}_{\leqslant N}\cap \mathcal {B}_1)\), then \(\psi _\gamma (a)\not \in \psi _\gamma (\mathcal {B}_1)\). \(\square \)

3.2 Poisson centraliser of the quadratic part

For any \(F\in {\mathcal {S}}({\mathfrak g})\), let \(F_\bullet \) be the component of F of the highest degree in \({\mathfrak t}\). If \(F\in {\mathcal {S}}({\mathfrak g})\) is homogeneous and \(\psi _\gamma (F_\bullet )\ne 0\) for \(\gamma \in \mathfrak t^*\), then \(\psi _\gamma (F_\bullet )\) is the lowest degree component of \(\psi _\gamma (F)\). Let \(F\in \mathcal {Z}_{{\mathbb V}(\mu )}\) be homogeneous. Since \(\varvec{f}_j=\varvec{f}_j(\mu )\in {\mathcal {S}}({\mathfrak m})\) for each j, we may write \(\psi _\gamma (\varvec{f}_j)=\varvec{f}_j\). Then \(\{\varvec{f}_j,\psi _\gamma (F)\}_{(\gamma )}=0\) and \(\{\varvec{f}_j,\psi _\gamma (F_\bullet )\}_{\gamma }=0\). A computation of \(\{\varvec{f}_j,\psi _\gamma (F_\bullet )\}_{\gamma }\) is not difficult, cf. [14], because

$$\begin{aligned} \{e_\beta e_{-\beta },\prod _{\alpha \in \Delta ^+} (e_\alpha ^{r_\alpha } e_{-\alpha }^{r_\alpha ^-})\}_{\gamma } = \gamma (h_\beta )(r^-_{\beta }-r_{\beta })\prod _{\alpha \in \Delta ^+} (e_\alpha ^{r_\alpha } e_{-\alpha }^{r_\alpha ^-}), \end{aligned}$$

if \(\beta \in \Delta ^+\). Note that the centraliser \(\mathcal {Z}_{{\mathbb V}(\mu )}\) is a homogeneous subalgebra of \({\mathcal {S}}({\mathfrak g})\).

We write \(\psi _\gamma (F_\bullet )\) in the basis \(\{e_{\pm \alpha }\mid \alpha \in \Delta ^+\}\). Let \(M=c_{{\bar{r}},\gamma } \prod _{\alpha \in \Delta ^+} (e_\alpha ^{r_\alpha } e_{-\alpha }^{r_\alpha ^-})\) be a summand of \(\psi _\gamma (F_\bullet )\) with \(c_{{\bar{r}},\gamma }\ne 0\). Then the explicit description of \({\mathbb V}(\mu )\), see (2.4), implies that

$$\begin{aligned} \sum _{\alpha \in \Delta ^+} \frac{\alpha ({\tilde{h}})}{\mu (h_\alpha )} (r_\alpha -r_\alpha ^-)\gamma (h_\alpha )=0 \end{aligned}$$
(2.6)

for each \({\tilde{h}}\in {\mathfrak t}\). The equality holds for all points \(\gamma \) such that \(c_{{\bar{r}},\gamma }\ne 0\). These points form a dense open subset of \({\mathfrak t}^*\). Thereby

$$\begin{aligned} \sum _{\alpha \in \Delta ^+} \frac{\alpha ({\tilde{h}})}{\mu (h_\alpha )} (r_\alpha -r_\alpha ^-) h_\alpha =0. \end{aligned}$$
(2.7)

If the numbers \(\frac{1}{\mu (h_\alpha )}\) are linearly independent over \(\mathbb Q\), then \(r_\alpha =r_\alpha ^-\) for each positive root \(\alpha \) [14]. In the same paper, it is shown that indeed the numbers \(\frac{1}{\mu (h_\alpha )}\) are linearly independent over \(\mathbb Q\) for generic \(\mu \in {\mathfrak t}^*\). We keep the assumption \(\mu \in {\mathfrak g}^*_\textsf{reg}\).

Lemma 2.7

(cf. [14]) Suppose that \(\psi _\gamma (F_\bullet )\) lies in \(\Bbbk [e_{\alpha } e_{-\alpha } \mid \alpha \in \Delta ^+]\) for any \(F\in \mathcal {Z}_{{\mathbb V}(\mu )}\) and any \(\gamma \in {\mathfrak t}^*\). Then \(\mathcal {Z}_{{\mathbb V}(\mu )}=(\mathcal{M}\mathcal{F})_{\mu }\).

Proof

If \(F_\bullet \) does not lie in \(\Bbbk [{\mathfrak t}, e_\alpha e_{-\alpha }\mid \alpha \in \Delta ^+]\) for some \(F\in \mathcal {Z}_{{\mathbb V}(\mu )}\), then there is \(\gamma \in {\mathfrak t}^*\), such that \(\psi _\gamma (F_\bullet )\not \in \Bbbk [e_{\alpha } e_{-\alpha } \mid \alpha \in \Delta ^+]\), a contradiction. Thus \(F_\bullet \in \Bbbk [{\mathfrak t}, e_\alpha e_{-\alpha }\mid \alpha \in \Delta ^+]\) for each \(F\in \mathcal {Z}_{{\mathbb V}(\mu )}\) and \({\mathrm {tr.deg\,}}\Bbbk [F_\bullet \mid F\in \mathcal {Z}_{{\mathbb V}(\mu )}]\leqslant |\Delta ^+|+l={\varvec{b}}(\mathfrak g)\). The algebras \(\mathcal {Z}_{{\mathbb V}(\mu )}\) and \(\Bbbk [F_\bullet \mid F\in \mathcal {Z}_{{\mathbb V}(\mu )}]\) are homogeneous and their Poincaré series coinside. Hence they have one and the same transcendence degree by [1, Satz 4.5]. Thus \({\mathrm {tr.deg\,}}\mathcal {Z}_{{\mathbb V}(\mu )}\leqslant {\varvec{b}}(\mathfrak g)\). By the construction, \((\mathcal{M}\mathcal{F})_{\mu }\subset \mathcal {Z}_{{\mathbb V}(\mu )}\). Moreover, \({\mathrm {tr.deg\,}}(\mathcal{M}\mathcal{F})_{\mu }={\varvec{b}}(\mathfrak g)\) and the algebra \((\mathcal{M}\mathcal{F})_{\mu }\) is algebraically closed in \({\mathcal {S}}(\mathfrak g)\) [11, Sect. 3]. Since \((\mathcal{M}\mathcal{F})_{\mu }\subset \mathcal {Z}_{{\mathbb V}(\mu )}\) is an algebraic extension, we have \((\mathcal{M}\mathcal{F})_{\mu }=\mathcal {Z}_{{\mathbb V}(\mu )}\). \(\square \)

Proposition 2.8

Suppose that \({\mathfrak g}\) is of type A\(_{n-1}\). Then \((\mathcal{M}\mathcal{F})_{\mu }=\mathcal {Z}_{{\mathbb V}(\mu )}\) for any \(\mu \in {\mathfrak t}^*\cap {\mathfrak g}^*_\textsf{reg}\).

Proof

Let \(M=c_{{\bar{r}},\gamma } \prod _{\alpha \in \Delta ^+} (e_\alpha ^{r_\alpha } e_{-\alpha }^{r_\alpha ^-})\) be a summand of \(\psi _\gamma (F_\bullet )\) with \(c_{{\bar{r}},\gamma }\ne 0\) for some homogeneous element \(F\in \mathcal {Z}_{{\mathbb V}(\mu )}\) and some \(\gamma \in {\mathfrak t}^*\). For \({\tilde{h}}=\frac{1}{n}\textrm{diag}(n{-}1,-1,\ldots ,-1)\), the equality (2.7) reads

$$\begin{aligned} \sum _{k=2}^n \frac{1}{\mu (E_{11}-E_{kk})} (r_{\varepsilon _1-\varepsilon _k}-r_{\varepsilon _1-\varepsilon _k}^-) (E_{11}-E_{kk})=0. \end{aligned}$$

Since the matrices \(E_{11}-E_{kk}\) with \(2\leqslant k\leqslant n\) are linearly independent, we conclude that \(r_{\varepsilon _1-\varepsilon _k}=r_{\varepsilon _1-\varepsilon _k}^-\) for each k. Then inserting \(\frac{1}{n-1}\textrm{diag}(0,n{-}2,-1,\ldots ,-1)\) as \({\tilde{h}}\) brings \(r_{\varepsilon _2-\varepsilon _k}=r_{\varepsilon _2-\varepsilon _k}^-\) for all \(k\geqslant 3\). Continuing in this way, we prove that the assumptions of Lemma 2.7 are satisfied for F and hence also for all elements of \(\mathcal {Z}_{{\mathbb V}(\mu )}\). \(\square \)

4 Remarks on quantisation

A commutative subalgebra \(\mathcal {Q}\subset {\mathcal {U}}({\mathfrak g})\) is a quantisation of a Poisson commutative subalgebra \(\mathcal {B}\subset {\mathcal {S}}({\mathfrak g})\), if \({\mathrm {gr\,}}\!(\mathcal {Q})=\langle {\mathrm {gr\,}}\!(a) \mid a\in \mathcal {Q}\rangle _{\Bbbk }\) coincides with \(\mathcal {B}\).

Keep the assumption \(\mu \in {\mathfrak t}^* \cap {\mathfrak g}^*_\textsf{reg}\). Then \({\mathfrak t}\subset (\mathcal{M}\mathcal{F})_{\mu }\). Assume that \(\mathcal {Q}\subset {\mathcal {U}}({\mathfrak g})\) is a quantisation of \((\mathcal{M}\mathcal{F})_{\mu }\). Then \({\mathfrak t}\subset \mathcal {Q}\) and \(\mathcal {Q}\subset {\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}\). We regard \(\varvec{f}_j=\sum _{\alpha \in \Delta ^+} C_\alpha e_{\alpha }e_{-\alpha }\) as an element of \({\mathcal {U}}(\mathfrak g)\) without adding new summands or changing the order of the factors. Since \(\mathfrak g^{{\mathfrak t}}={\mathfrak t}\), the subalgebra \(\mathcal {Q}\) must contain all \(\varvec{f}_j\in {\mathcal {U}}({\mathfrak g})\). In [19], it is shown that \([\varvec{f}_j,\varvec{f}_i]=0\) in \({\mathcal {U}}({\mathfrak g})\) for all ij. Set \(\widetilde{{\mathbb V}}(\mu )=\langle \varvec{f}_j\mid 1\leqslant j\leqslant l\rangle _{\Bbbk }\subset {\mathcal {U}}({\mathfrak g})\).

If \(\mu \) is generic enough, then \((\mathcal{M}\mathcal{F})_{\mu }\) is equal to the Poisson centraliser \(\mathcal {Z}_{{\mathbb V}(\mu )}\) [14]. Therefore \(\mathcal {Q}\) is the centraliser of \(\widetilde{{\mathbb V}}(\mu )\) in \({\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}\). The quantisation of this \((\mathcal{M}\mathcal{F})_{\mu }\) is unique. In type A, it was shown earlier in [18], that the quantisation of \((\mathcal{M}\mathcal{F})_{\mu }\) is unique for any \(\mu \in {\mathfrak t}^* \cap {\mathfrak g}^*_\textsf{reg}\). Proposition 2.8 provides a new proof for the uniqueness result in case \({\mathfrak g}=\mathfrak {sl}_{n}\).

The existence of \(\mathcal {Q}\) is proven in [9] for classical \({\mathfrak g}\), in [17] for type A, and in [4, 15] for any \({\mathfrak g}\). Let \(\mathcal {F}_\gamma \subset {\mathcal {U}}({\mathfrak g})\) with \(\gamma \in {\mathfrak g}^*\) be the quantum Mishchenko–Fomenko subalgebra, which is a commutative algebra constructed in [15] and [4]. Then \({\mathrm {gr\,}}\!(\mathcal {F}_\gamma )=(\mathcal{M}\mathcal{F})_{\gamma }\) for any \(\gamma \in {\mathfrak g}^*_\textsf{reg}\) [4, 15] and for any \(\gamma \in {\mathfrak g}^*\) in case \({\mathfrak g}\) is of type A or C [8].

Now we lift elements \(F_j=(H_j)_{(d_j-2,2)}\in {\tilde{{\mathcal {Z}}}}\) to \({\mathcal {U}}({\mathfrak g})\). Write \(F_j=\sum _{\alpha \in \Delta ^+} F_{j,\alpha } e_\alpha e_{-\alpha }\), where \(F_{j,\alpha }\in {\mathcal {S}}^{d_j-2}({\mathfrak t})\), and regard this sum as an element of \({\mathcal {U}}(\mathfrak g)\). Recall that in \({\mathcal {S}}({\mathfrak g})\), we have \(\varvec{f}_j\in (\mathcal{M}\mathcal{F})_{\mu }\) and \(\psi _\mu (F_j)=\psi _\mu (\varvec{f}_j)=\varvec{f}_j\).

We work with \({\mathcal {C}}={\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}\otimes _{{\mathcal {S}}({\mathfrak t})}\Bbbk ({\mathfrak t}^*)\) as with a non-commutative algebra over \(\Bbbk ({\mathfrak t}^*)\) generated by the monomials

$$\begin{aligned} M=e_{\alpha _1}^{r_1}\ldots e_{\alpha _N}^{r_N}e_{-\alpha _N}^{r_N^-}\ldots e_{-\alpha _1}^{r_1^-}, \end{aligned}$$
(3.1)

where \(N=|\Delta ^+|\), some numbering of the positive roots is fixed, and \(\sum \nolimits _{i=1}^N(r_i-r_i^-)\alpha _i=0\). These monomials form a basis of \({\mathcal {S}}(\mathfrak m)^{\mathfrak t}\). We say that \(\sum _{i=1}^N(r_i+r_i^-)=:\deg _{\mathfrak m}M\) is the degree (or the \(\mathfrak m\)-degree) of M. The algebra structure of \({\mathcal {C}}\) is given by the coefficients \(Q_M^{M',M''}\in \Bbbk ({\mathfrak t}^*)\) of \(M' M''= \sum _{M} Q_M^{M',M''} M\), where \(M,M',M''\) are of the form (3.1). In these terms, one can extend the map \(\psi _\mu \!:\mathcal {A}_\mu \rightarrow {\mathcal {S}}({\mathfrak m})^{{\mathfrak t}}\) to a rational map from \({\mathcal {C}}\) by evaluating at \(\mu \) the coefficients \(Q_M^{M',M''}\). Formally, set \({\mathcal {C}}_\mu =\{ A\in {\mathcal {C}}\mid A(\mu ) \,\text { is well-defined}\,\}\). As a vector space, the image \(\psi _\mu ({\mathcal {C}}_\mu )=:{\tilde{{\mathcal {U}}}}({\mathfrak m})^{{\mathfrak t}}\) is isomorphic to \({\mathcal {S}}({\mathfrak m})^{\mathfrak t}\). We let \([\,\,,\,]_{(\mu )}\) stand for the Lie algebra structure on it. The algebra \(({\tilde{{\mathcal {U}}}}({\mathfrak m})^{{\mathfrak t}},[\,,\,]_{(\mu )})\) should not be regarded as a subset of \({\mathcal {U}}({\mathfrak m})^{\mathfrak t}\) in whatever sense! Note that a similar construction exists for any \(\Bbbk ({\mathfrak t}^*)\)-basis of \({\mathcal {C}}\).

Example 3.1

We check that \([F_j,F_s]=0\) in \({\mathcal {U}}(\mathfrak g)\) for all j and s. By definition

$$\begin{aligned}{}[F_j,F_s]{} & {} =\sum \limits _{\alpha ,\beta \in \Delta ^+} F_{j,\alpha }F_{s,\beta } e_\alpha e_{-\alpha } e_{\beta }e_{-\beta } - \sum \limits _{\alpha ,\beta \in \Delta ^+} F_{j,\alpha }F_{s,\beta } e_\beta e_{-\beta } e_{\alpha }e_{-\alpha } \\{} & {} \quad = \sum _{\alpha \ne \beta } F_{j,\alpha }F_{s,\beta } (e_\alpha e_{-\alpha } e_{\beta }e_{-\beta }- e_\beta e_{-\beta } e_{\alpha }e_{-\alpha }) \\{} & {} \quad = \sum _{\alpha \ne \beta } F_{j,\alpha }F_{s,\beta } ( [e_{\alpha },e_\beta ] e_{-\alpha } e_{-\beta }+e_\alpha [e_{-\alpha }, e_{\beta }] e_{-\beta } + e_{\beta } [e_{\alpha }, e_{-\beta }] e_{-\alpha }\\{} & {} \qquad + e_\beta e_\alpha [e_{-\alpha }, e_{-\beta }]). \end{aligned}$$

In this particular case, further straightening of the sums in brackets does not involve elements of \({\mathfrak t}\). For each \(\gamma \in {\mathfrak t}^*\cap \mathfrak g^*_\textsf{reg}\), the elements \(\psi _\gamma (F_j)\) and \(\psi _\gamma (F_s)\) belong to the quantum Mishchenko–Fomenko subalgebra \(\mathcal {F}_\gamma \subset {\mathcal {U}}({\mathfrak g})\) associated with \(\gamma \). If we replace each \(F_{j,\alpha }F_{s,\beta }\) with \(\psi _\gamma (F_{j,\alpha }), \psi _{\gamma }(F_{s,\beta })\), then the total sum is zero for each \(\gamma \in {\mathfrak t}^*\cap \mathfrak g^*_\textsf{reg}\). This implies that the initial sum is zero in \({\mathcal {U}}(\mathfrak g)\).

The algebra \({\mathcal {C}}=\bigcup _{N\geqslant 0}{\mathcal {C}}_N\) is filtered by the degree in \({\mathfrak m}\). Here \({\mathcal {C}}_0={\mathcal {C}}_1=\Bbbk ({\mathfrak t}^*)\) and the \(\Bbbk ({\mathfrak t}^*)\)-space \({\mathcal {C}}_2\) has a basis \(\{1, e_\alpha e_{-\alpha } \mid \alpha \in \Delta ^+\}\). More generally, any \({\mathcal {C}}_N\) has a monomial basis consisting of the monomials M, of the form (3.1), with \(\deg _{\mathfrak m}M\leqslant N\). Since the commutator in \({\mathcal {C}}\) is \(\Bbbk ({\mathfrak t}^*)\)-linear, for any \(A\in {\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}\) and any \(N\geqslant 0\), the centraliser \(\mathcal {Z}_A({\mathcal {C}}_{N})=\mathcal {Z}_A {\mathcal {C}}\cap {\mathcal {C}}_N\) is a vector space over \(\Bbbk ({\mathfrak t}^*)\). It is a subspace of the finite-dimensional space \({\mathcal {C}}_N\). Evaluating the defining equations of \(\mathcal {Z}_A({\mathcal {C}}_{N})\) at \(\mu \in {\mathfrak t}^*\), we obtain

$$\begin{aligned} \dim _{\Bbbk ({\mathfrak t}^*)} \mathcal {Z}_A ({\mathcal {C}}_{N}) = \dim \left( \mathcal {Z}_{\psi _\mu (A)}({\tilde{{\mathcal {U}}}}({\mathfrak m})^{\mathfrak t},[\,\,,\,]_{(\mu )})\cap {\tilde{{\mathcal {U}}}}_{N}({\mathfrak m})\right) , \end{aligned}$$
(3.2)

whenever \(\mu \) is generic enough. This applies also to centralisers of finite subsets of elements. Since \(\psi _\mu (\mathcal {F}_\mu )\) commutes with all \(\psi _{\mu }(\varvec{f}_j)=\psi _\mu (F_j)\) w.r.t. \([\,,\,]_{(\mu )}\), the equality (3.2) shows that the coefficients of the Poincaré series of the centraliser

$$\begin{aligned} \mathcal {Z}_{\widetilde{{\mathbb V}}(\mu )}{\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}=\{ \Xi \in {\mathcal {U}}({\mathfrak g})^{{\mathfrak t}} \mid [\Xi ,F_j]=0 \ \forall j\} \end{aligned}$$

are large, \(\dim _{\Bbbk ({\mathfrak t}^*)} \mathcal {Z}_{\widetilde{{\mathbb V}}(\mu )}({\mathcal {C}}_N)\geqslant D_N\), where

$$\begin{aligned} D_N=\dim ({\mathcal {S}}^{\leqslant N}({\mathfrak m}){\mathcal {S}}({\mathfrak t})\cap \Bbbk [\partial _\mu ^k H_j \mid 1\leqslant j\leqslant l, 0\leqslant k<d_j-1]). \end{aligned}$$

In other words, enough elements of \({\mathcal {U}}({\mathfrak g})^{\mathfrak t}\) commute with \(\widetilde{{\mathbb V}}(\mu )\). It is not known, whether the centraliser \(\mathcal {Z}_{\widetilde{{\mathbb V}}(\mu )}{\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}\) is commutative or not. In order to solve the quantisation problem for \({\tilde{{\mathcal {Z}}}}\), one may try to obtain upper bounds for dimensions related to \(\mathcal {Z}_{\widetilde{{\mathbb V}}(\mu )}{\mathcal {U}}({\mathfrak g})^{\mathfrak t}\) or to \(\mathcal {Z}_{\widetilde{{\mathbb V}}(\mu )}{\mathcal {C}}\).

Let \(\textsf{symm}\!:{\mathcal {S}}({\mathfrak g})\rightarrow {\mathcal {U}}({\mathfrak g})\) be the canonical symmetrisation map.

Theorem 3.2

Suppose that \({\mathfrak g}\) is of type A\(_l\). Then there is a commutative subalgebra \(\mathscr {Z}\subset {\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}\) such that \({\mathrm {gr\,}}\!(\mathscr {Z})={\tilde{{\mathcal {Z}}}}\) and \(F_j=\sum _{\alpha \in \Delta ^+} F_{j,\alpha } e_\alpha e_{-\alpha }\in \mathscr {Z}\) for each \(1\leqslant j\leqslant l\).

Proof

For convenience, we work with \({\mathfrak g}=\mathfrak {gl}_n\) instead of \(\mathfrak {sl}_{l+1}\). Let \(H_j=\Delta _j\) be coefficients of the characteristic polynomial, here \(\deg \Delta _j=j\). We write \(\Delta _j\) in the basis \(\{E_{ik}\mid i,k\leqslant n\}\). Let X be a monomial appearing in \(\Delta _j\) with a nonzero coefficient. If \(E_{ii}\) is a factor of X, then for all other factors \(E_{sk}\) of X, we have \(i\not \in \{s,k\}\). For \({\mathfrak t}=\langle E_{ii}\mid 1\leqslant i\leqslant n\rangle _{\Bbbk }\), the \({\mathfrak t}\)-factors of X commute with all factors of X. Another feature of the set \(\{\Delta _j\mid 1\leqslant j\leqslant n\}\) is that

$$\begin{aligned} \mathcal {F}_\xi =\textsf {alg}\langle \textsf {symm}(\partial _\xi ^k \Delta _j) \mid 1\leqslant j\leqslant n, 0\leqslant k< j\rangle \ \text{ for } \text{ any } \ \xi \in {\mathfrak g}^*, \end{aligned}$$

see [16, 18] and [8, Theorem 3.1]. Set

$$\begin{aligned} \mathscr {Z}=\textsf {alg}\langle {\mathfrak t}, \textsf {symm}((\Delta _j)_{(i,j-i)}) \mid 1\leqslant j\leqslant n, 0\leqslant i\leqslant j-2\rangle \subset {\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}. \end{aligned}$$

We have \({\mathrm {gr\,}}\!(\mathscr {Z})={\tilde{{\mathcal {Z}}}}\), because a basis of \({\mathfrak t}\) and the elements \((\Delta _j)_{(i,j-i)}\) form an algebraically independent set of generators of \({\tilde{{\mathcal {Z}}}}\). In order to prove that \(\mathscr {Z}\) is commutative, we use maps \(\psi _\mu \!:{\mathcal {C}}_\mu \rightarrow {\tilde{{\mathcal {U}}}}({\mathfrak m})^{{\mathfrak t}}\) with \(\mu \in {\mathfrak t}^*\), working now with the \(\Bbbk ({\mathfrak t}^*)\)-basis

$$\begin{aligned} \{\textsf{symm}(M) \mid M\in {\mathcal {S}}({\mathfrak m})^{\mathfrak t}\text { monomial in } E_{ij} \text { with } i,j\leqslant n\} \end{aligned}$$

of \({\mathcal {C}}\).

We decompose \((\Delta _j)_{(i,j-i)}=\sum _s P_{j,i}^{(s)}M_{j,j-i}^{(s)}\), where \(P_{j,i}^{(s)}\in {\mathcal {S}}^{i}({\mathfrak t})\) are pairwise different monomials in elements \(E_{kk}\) and \(M_{j,u}^{(s)}\in {\mathcal {S}}^u({\mathfrak m})\) are nonzero. Since the factors of \(P_{j,i}^{(s)}\) commute with \(M_{j,j-i}^{(s)}\), we have

$$\begin{aligned} \textsf{symm}(P_{j,i}^{(s)}M_{j,j-i}^{(s)})=P_{j,i}^{(s)} \textsf{symm}(M_{j,j-i}^{(s)}) \end{aligned}$$

for each s. Furthermore, \(\partial _\mu ^k (P_{j,i}^{(s)}M_{j,j-i}^{(s)})=(\partial _\mu ^k P_{j,i}^{(s)})M_{j,j-i}^{(s)}\) and

$$\begin{aligned} \textsf{symm}(\partial _\mu ^k (P_{j,i}^{(s)}M_{j,j-i}^{(s)}))=(\partial _\mu ^k P_{j,i}^{(s)})\textsf{symm}(M_{j,j-i}^{(s)}). \end{aligned}$$

Finally \(\psi _\mu {\circ }\textsf{symm}(\partial _\mu ^k (P_{j,i}^{(s)}M_{j,j-i}^{(s)}))=i(i-1)\ldots (i-k+1)P_{j,i}^{(s)}\!(\mu )\textsf{symm}(M_{j,j-i}^{(s)})\). Therefore \(\psi _\mu (\mathcal {F}_\mu )\) is generated by \(\sum _{s} P_{i,j}^{(s)}\!(\mu )\textsf{symm}(M_{j,j-i}^{(s)})\) with \(1\leqslant j\leqslant n\) and \(0\leqslant i \leqslant j-2\). Recall that \(\psi _\mu (\mathcal {F}_\mu )\) is a commutative subalgebra of \(({\tilde{{\mathcal {U}}}}({\mathfrak m})^{{\mathfrak t}},[\,\,,\,]_{(\mu )})\) by the construction.

Next we observe that \(\psi _\mu {\circ }\textsf{symm}((\Delta _j)_{(i,j-i)})=\sum _s P_{j,i}^{(s)}\!(\mu )\textsf{symm}(M_{j,j-i}^{(s)})\in \psi _\mu (\mathcal {F}_\mu )\). Thus, \(\psi _\mu (\mathscr {Z})\) is commutative for any \(\mu \in {\mathfrak t}^*\). By a general principle already used in Sect. 2, this implies \([\mathscr {Z},\mathscr {Z}]=0\). In case of

$$\begin{aligned} F_j=(\Delta _j)_{(j-2,2)}=\sum _{\alpha \in \Delta ^+} F_{j,\alpha } e_\alpha e_{-\alpha } \ \text { with } \ j\geqslant 2, \end{aligned}$$

the coefficients \(F_{j,\alpha }\) are monomials in \(E_{kk}\) and \(\textsf{symm}(F_j)=\sum _{\alpha \in \Delta ^+} F_{j,\alpha } (e_\alpha e_{-\alpha } - \frac{1}{2}h_\alpha )\in \mathscr {Z}\). Therefore also \(\sum _{\alpha \in \Delta ^+} F_{j,\alpha } e_\alpha e_{-\alpha }\in \mathscr {Z}\).

In order to return from \(\mathfrak {gl}_n\) to \(\mathfrak {sl}_{l+1}\) with \(l+1=n\), we restrict the invariants \(\Delta _j\) to \(\mathfrak {sl}_n^*\). This can be achieved by writing first \(E_{ii}={\tilde{E}}_{ii}+z\) for each i with \(z=\frac{1}{n}{\textrm{diag}}(1,\ldots ,1)\) and then by setting \(z=0\). If \({\tilde{\Delta }}_j=\Delta _j|_{\mathfrak {sl}_n^*}\) and \({\tilde{\mathscr {Z}}}\subset {\mathcal {U}}(\mathfrak {sl}_n)\) is generated by \({\mathfrak t}\) together with the elements \(\textsf{symm}( ({\tilde{\Delta }}_j)_{(i,j-i)})\), where \(j\geqslant 2\), then \({\tilde{\mathscr {Z}}}\) is a required commutative subalgebra. \(\square \)

The quantisation \(\mathscr {Z}\subset {\mathcal {U}}({\mathfrak g})^{{\mathfrak t}}\) of \({\tilde{{\mathcal {Z}}}}\) described in the proof of Theorem 3.2 is a curious subalgebra. Let \(V_\lambda \) be an irreducible finite-dimensional \({\mathfrak g}\)-module with \({\mathfrak g}=\mathfrak {sl}_n\) and \((V_\lambda )_{\mu }\) the subspace of \(V_\lambda \) corresponding to a \({\mathfrak t}\)-weight \(\mu \). Then \(\mathscr {Z}\) acts on \((V_\lambda )_{\mu }\) as \(\mathcal {F}_\mu \). In particular, the action of \(\mathscr {Z}\) on \(V_\lambda \) is diagonalisable, since \(\mu \) takes real values on the standard real form of \({\mathfrak t}\), see [3]. Furthermore, if \(\mu \in {\mathfrak t}^*\cap {\mathfrak g}^*_\textsf{reg}\), then also by results of [3], \(\mathcal {F}_\mu \), and hence \(\mathscr {Z}\), acts on \((V_\lambda )_{\mu }\) with a simple spectrum. However, the action of \(\mathscr {Z}\) on \((V_\lambda )_{\mu }\) may not have a simple spectrum if \(\mu \not \in {\mathfrak t}^*\cap {\mathfrak g}^*_\textsf{reg}\). For instance, \(\mathscr {Z}\) acts via scalars on the zero weight subspace \(V_\lambda ^{\mathfrak t}\).

Example 3.3

Suppose that \(\mathfrak g=\mathfrak {sl}_3\). Then \({\tilde{{\mathcal {Z}}}}\) is generated by a basis of \({\mathfrak t}\), two invariants \(H_1,H_2\), and \(F_2=(H_2)_{(1,2)}=\sum _{\alpha \in \Delta ^+} F_{2,\alpha } e_\alpha e_{-\alpha }\). Here \(F_{2,\alpha } \in {\mathfrak t}\) for each \(\alpha \). We regard \(F_2\) as an element of \({\mathcal {U}}(\mathfrak g)\). Then \(\mathscr {Z}\) is generated by \({\mathfrak t}\), \(F_2\), and \({\tilde{H}}_1,{\tilde{H}}_2\), where \(\tilde{H}_1,{\tilde{H}}_2\in {\mathcal {U}}({\mathfrak g})\) are independent central elements. On an irreducible finite-dimensional \(\mathfrak g\)-module \(V_\lambda \), the last two generators act via scalar multiplication. The actions of \({\mathfrak t}\) and \(F_2\) annihilate \(V_\lambda ^{{\mathfrak t}}\).