This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

Lie algebras of triangular polynomial derivations and an isomorphism criterion for their Lie factor algebras

© 2013 Russian Academy of Sciences (DoM), London Mathematical Society, Turpion Ltd.
, , Citation V. V. Bavula 2013 Izv. Math. 77 1067 DOI 10.1070/IM2013v077n06ABEH002670

1064-5632/77/6/1067

Abstract

We make a detailed study of the Lie algebras $\mathfrak{u}_n$, $n\geqslant 2$, of triangular polynomial derivations, their injective limit $\mathfrak{u}_\infty$, and their completion $\widehat{\mathfrak{u}}_\infty$. We classify the ideals of $\mathfrak{u}_n$ (all of which are characteristic ideals) and use this classification to give an explicit criterion for Lie factor algebras of $\mathfrak{u}_n$ and $\mathfrak{u}_m$ to be isomorphic. We introduce two new dimensions for (Lie) algebras and their modules: the central dimension $\operatorname{c.dim}$ and the uniserial dimension $\operatorname{u.dim}$, and show that $\operatorname{c.dim}(\mathfrak{u}_n)=\operatorname{u.dim}(\mathfrak{u}_n) =\omega^{n-1}+\omega^{n-2}+\dots+\omega +1$ for all $n\geqslant 2$, where $\omega$ is the first infinite ordinal. Similar results are proved for the Lie algebras $\mathfrak{u}_\infty$ and $\widehat{\mathfrak{u}}_\infty$. In particular, $\operatorname{u.dim}(\mathfrak{u}_\infty)=\omega^\omega$ and $\operatorname{c.dim}(\mathfrak{u}_\infty)= 0$.

Export citation and abstract BibTeX RIS

§ 1. Introduction

Throughout, module means a left module; $\mathbb{N}:=\{0, 1, \dots \}$ is the set of non-negative integers; $K$ is a field of characteristic zero and $K^*$ is its group of units; $P_n:= K[x_1,\dots,x_n]=\bigoplus_{\alpha \in \mathbb{N}^n} Kx^{\alpha}$ is the polynomial algebra over $K$, where $x^{\alpha}:=x_1^{\alpha_1}\dotsb x_n^{\alpha_n}$; $\partial_1:=\frac{\partial}{\partial x_1},\dots,\partial_n:=\frac{\partial}{\partial x_n}$ are the partial derivatives ($K$-linear derivations) of $P_n$; $\operatorname{Der}_K(P_n) =\bigoplus_{i=1}^nP_n\partial_i$ is the Lie algebra of $K$-derivations of $P_n$; $A_n\,{:=}\,K\langle x_1, \dots, x_n,\partial_1,\dots,\partial_n\rangle \,{=}\,\bigoplus_{\alpha,\beta\in \mathbb{N}^n} Kx^\alpha \partial^\beta$ is the $n$th Weyl algebra; for every integer $n\geqslant 2$,

is the Lie algebra of triangular polynomial derivations (it is a Lie subalgebra of the Lie algebra $\operatorname{Der}_K(P_n)$), and $U_n:= U(\mathfrak{u}_n)$ is its universal enveloping algebra.

Many properties of the $\mathfrak{u}_n$, where $n\geqslant 2$, are proved in Proposition 2.1. In particular, the Lie algebras $\mathfrak{u}_n$ $(n\geqslant 2)$ are pairwise non-isomorphic, soluble but not nilpotent, and their inner derivations are locally nilpotent. The derived series and lower central series are found for $\mathfrak{u}_n$.

We introduce a new dimension for algebras and modules, the uniserial dimension (see § 4), which has turned out to be a very useful tool in the study of non-Noetherian Lie algebras, their ideals and automorphisms [1], [2].

In § 3 we give a classification of the ideals of $\mathfrak{u}_n$ and find an explicit basis for each. We prove that the Lie algebra $\mathfrak{u}_n$ is uniserial, Artinian but not Noetherian, and its uniserial dimension is equal to

where $\omega$ is the first infinite ordinal (Theorem 3.6). We show that the hypercentral series $\{Z^{(\lambda)} (\mathfrak{u}_n) \}_{\lambda \in \text{W}}$ stabilizes precisely at step $\operatorname{ord}(\Omega_n) \,{:=}\, \omega^{n-1}+\omega^{n-2}+\dots + \omega + 1$. Moreover, $Z^{(\lambda)}(\mathfrak{u}_n) = I_\lambda$ for every $\lambda \in [1,\operatorname{ord}(\Omega_n)]$, where $I_\lambda$ is explicitly given by (3.3). The central dimension of $\mathfrak{u}_n$ is $\operatorname{c.dim}(\mathfrak{u}_n) =\operatorname{ord}(\Omega_n)$ (Theorem 3.10). It is proved that all the ideals of $\mathfrak{u}_n$ are characteristic (Corollary 3.11).

A Lie algebra $\mathcal G$ is said to be locally nilpotent (resp. locally finite-dimensional) if every finitely generated Lie subalgebra of $\mathcal G$ is nilpotent (resp. finite-dimensional). In § 4 we prove that the $\mathfrak{u}_n$ are locally finite-dimensional and locally nilpotent (Theorem 4.2).

In § 5, Theorem 5.1 and Corollary 5.2 give an answer to the following question.

Question. Let $I$ and $J$ be ideals of the Lie algebras $\mathfrak{u}_n$ and $\mathfrak{u}_m$ respectively. When are the Lie factor algebras $\mathfrak{u}_n/I$ and $\mathfrak{u}_m/J$ isomorphic?

The answer is given in explicit terms (via the uniserial dimensions of $I$ and $J$) using the classification of ideals of $\mathfrak{u}_n$ (Theorem 3.6). In particular, there are only countably many ideals $I$ of $\mathfrak{u}_n$ such that $\mathfrak{u}_n/I\simeq \mathfrak{u}_n$. Theorem 3.6(1) shows that every ideal $I$ of $\mathfrak{u}_n$ is uniquely determined by its uniserial dimension

that is, $I=I_\lambda$, where $\omega$ is the first infinite ordinal.

In § 6 we study the Lie algebra $\mathfrak{u}_\infty:= \bigcup_{n\geqslant 2}\mathfrak{u}_n = \bigoplus_{n\geqslant 2} P_{n-1}\partial_n$. Many of the properties of $\mathfrak{u}_\infty$ are similar to those of the $\mathfrak{u}_n$ $(n\geqslant 2)$ but there are several differences. For example, $\mathfrak{u}_\infty$ is not soluble, not Artinian but almost Artinian, and $\operatorname{u.dim} (\mathfrak{u}_\infty) = \omega^\omega$. We obtain a classification of the ideals of $\mathfrak{u}_\infty$ (Theorem 6.2). All the ideals of $\mathfrak{u}_\infty$ are characteristic (Corollary 6.4). Corollary 6.3 is an explicit criterion for when two Lie factor algebras of $\mathfrak{u}_\infty$ are isomorphic.

In § 7 we study the topological Lie algebra $\widehat{\mathfrak{u}}_\infty$, which is the completion of $\mathfrak{u}_\infty$. Its properties diverge further from those of the $\mathfrak{u}_n$ $(n\geqslant 2)$ and $\mathfrak{u}_\infty$. We classify all closed ideals and all open ideals of $\widehat{\mathfrak{u}}_\infty$ (Theorem 7.2(1)).

§ 2. The derived series and lower central series of the Lie algebras $\mathfrak{u}_n$

In this section we prove various properties of the $\mathfrak{u}_n$ (Proposition 2.1, Corollary 2.3) that are widely used in the rest of the paper. At the end of the section we find the image and kernel of the algebra homomorphism $\chi_n\colon U_n\to A_{n-1}\otimes K[\partial_n]$ (see (2.11) and Theorem 2.6). In particular, it is shown that the algebra $\chi_n(U_n)$ is not finitely generated and neither left nor right Noetherian.

Let $\mathcal{G}$ be a Lie algebra over the field $K$, and let $\mathfrak{a}$, $\mathfrak{b}$ be ideals of $\mathcal{G}$. The commutator ${[}\mathfrak{a},\mathfrak{b}]$ of $\mathfrak{a}$ and $\mathfrak{b}$ is the linear span in $\mathcal{G}$ of all the elements ${[}a,b]$, where $a\in\mathfrak{a}$ and $b\in\mathfrak{b}$. The commutator ${[}U,V]$ of subspaces $U$ and $V$ of $\mathcal{G}$ is defined in the same manner. ${[}\mathfrak{a},\mathfrak{b}]$ is an ideal of $\mathcal{G}$, and ${[}\mathfrak{a},\mathfrak{b}] \subseteq\mathfrak{a}\cap \mathfrak{b}$. In particular, $\mathcal{G}_{(1)}:=\mathcal{G}^{(1)}:=[\mathcal{G},\mathcal{G}]$ is called the commutator subalgebra of $\mathcal{G}$. Let us define recursively the following set of ideals of $\mathcal{G}$:

Clearly, $\mathcal{G}_{(i)}\subseteq\mathcal{G}^{(i)}$ for all $i\geqslant 1$. The descending chains of ideals of $\mathcal{G}$,

are called the derived series and the lower central series of $\mathcal{G}$ respectively. Notice that

Thus, the elements

Equation (2.1)

form a $K$-basis $\mathcal{B}_n$ of $\mathfrak{u}_n$ called the canonical basis. For all $i,j$, $1\leqslant i\leqslant j\leqslant n$, $\alpha\in\mathbb{N}^{i-1}$ and $\beta\in\mathbb{N}^{j-1}$ we have

Equation (2.2)

where $e_1:=(1,0,\dots,0),\dots,e_n:=(0,\dots, 0, 1)$ is the canonical free $\mathbb{Z}$-basis of the $\mathbb{Z}$-module $\mathbb{Z}^n$ and $\mathbb{N}^i:=\sum_{k=1}^i\mathbb{N}e_k\subseteq\mathbb{Z}^i:=\sum_{k=1}^i\mathbb{Z} e_k$. In particular, $\mathbb{N}\subseteq\mathbb{N}^2\subseteq\cdots\subseteq\mathbb{N}^n$ and $\mathbb{Z}\subseteq\mathbb{Z}^2\subseteq\dotsb\subseteq\mathbb{Z}^n$. The Lie algebra $\mathfrak{u}_n=\bigoplus_{i=1}^nP_{i-1}\partial_i$ is the direct sum of the Abelian (infinite-dimensional when $i> 1$) Lie subalgebras $P_{i-1}\partial_i$ (that is, ${[}P_{i-1}\partial_i,P_{i-1}\partial_i]= 0$) such that, for all $i<j$,

Equation (2.3)

The inclusion $\subseteq$ in (2.3) is obvious but the equality follows from the fact that ${[}\partial_i,P_{j-1}]=P_{j-1}$. By (2.3), $\mathfrak{u}_n$ admits a finite strictly decreasing chain of ideals

Equation (2.4)

where $\mathfrak{u}_{n,i}:=\sum_{j=i}^nP_{j-1}\partial_j$ for $i=1,\dots,n$. By (2.3), for all $i<j$,

Equation (2.5)

For every $i=1,\dots,n$ there is a canonical isomorphism of Lie algebras

Equation (2.6)

In particular, $\mathfrak{u}_{n-1}\simeq\mathfrak{u}_n/P_{n-1}\partial_n$. Clearly,

is an ascending chain of Lie algebras. The polynomial algebra $P_n$ is an $A_n$-module: for all elements $p\in P_n$,

Clearly, $P_n\simeq A_n/\sum_{i=1}^n A_n\partial_i$, $1\mapsto 1+\sum_{i=1}^n A_n\partial_i$. Since $\mathfrak{u}_n\subseteq A_n$, the polynomial algebra $P_n$ is also a $\mathfrak{u}_n$-module.

Let $V$ be a vector space over $K$. A $K$-linear map $\delta\colon V\to V$ is called a locally nilpotent map if $V\,{=}\,\bigcup_{i\geqslant 1}\operatorname{ker}(\delta^i)$ or, equivalently, for every $v\in V$ we have $\delta^i (v)\,{=}\,0$ for all $i\gg 1$. When $\delta$ is a locally nilpotent map in $V$, we also say that $\delta$ acts locally nilpotently on $V$. Every nilpotent linear map $\delta$ (that is, with $\delta^n= 0$ for some $n\geqslant 1$) is a locally nilpotent map but not vice versa, in general. Let $\mathcal{G}$ be a Lie algebra. Each element $a\in\mathcal{G}$ determines a derivation of $\mathcal{G}$ by the rule $\operatorname{ad}(a)\colon\mathcal{G}\to\mathcal{G}$, $b\mapsto[a,b]$. It is called the inner derivation associated with $a$. The set $\operatorname{Inn}(\mathcal{G})$ of all inner derivations of $\mathcal{G}$ is a Lie subalgebra of the Lie algebra $(\operatorname{End}_K(\mathcal{G}),[\,\cdot\,{,}\,\cdot\,])$, where ${[}f,g]:= fg-gf$. There is a short exact sequence of Lie algebras

that is, $\operatorname{Inn}(\mathcal{G})\,{\simeq}\,\mathcal{G}/Z(\mathcal{G})$, where $Z(\mathcal{G})$ is the centre of $\mathcal{G}$ and $\operatorname{ad}([a,b])\,{=}\,[\operatorname{ad}(a),\operatorname{ad}(b)]$ for all $a, b\in\mathcal{G}$. An element $a\in\mathcal{G}$ is said to be locally nilpotent (resp. nilpotent) if the inner derivation $\operatorname{ad}(a)$ of $\mathcal{G}$ has that property. Let $J$ be a non-empty subset of $\mathcal{G}$. Then $\operatorname{Cen}_{\mathcal{G}}(J)\,{:=}\,\{ a\,{\in}\,\mathcal{G}\mid [a,b]\,{=}\,0$ for all $b\,{\in}\,J\}$ is called the centralizer of $J$ in $\mathcal{G}$. It is a Lie subalgebra of $\mathcal{G}$. Let $A$ be an associative algebra and $I$ a non-empty subset of $A$. Then $\operatorname{Cen}_A(I):=\{a\in A\mid ab=ba$ for all $b\in I\}$ is called the centralizer of $I$ in $A$. It is a subalgebra of $A$.

Proposition 2.1. 

  • 1)  
    The Lie algebra $\mathfrak{u}_n$ is soluble but not nilpotent.
  • 2)  
    The finite chain of ideals (2.4) is the derived series of $\mathfrak{u}_n$, that is, $(\mathfrak{u}_n)_{(i)}=\mathfrak{u}_{n,i+1}$ for all $i\geqslant 0$.
  • 3)  
    The lower central series of $\mathfrak{u}_n$ stabilizes at the first step, that is, $(\mathfrak{u}_n)^{(0)}= \mathfrak{u}_n$ and $(\mathfrak{u}_n)^{(i)}=\mathfrak{u}_{n, 2}$ for all $i\geqslant 1$.
  • 4)  
    Each element $u\in\mathfrak{u}_n$ acts locally nilpotently on the $\mathfrak{u}_n$-module $P_n$.
  • 5)  
    All inner derivations of $\mathfrak{u}_n$ are locally nilpotent.
  • 6)  
    The centre $Z(\mathfrak{u}_n)$ of $\mathfrak{u}_n$ is equal to $K\partial_n$.
  • 7)  
    The $\mathfrak{u}_n$, where $n\geqslant 2$, are pairwise non-isomorphic.

Proof. Part 1 follows from parts 2 and 3, which in their turn follow from the decomposition $\mathfrak{u}_n\,{=}\,\bigoplus_{i=1}^n P_{i-1}\partial_i$ and (2.3). Part 4 follows from the definition of $\mathfrak{u}_n$.

Let us prove part 5. Suppose that $u\in\mathfrak{u}_{n,i}\setminus \mathfrak{u}_{n, i+1}$ for some $i=1,\dots,n$. Then $u\,{=}\,a\partial_i+u'$ for some elements $a\,{\in}\,P_{i-1}$ and $u'\,{\in}\,\mathfrak{u}_{n,i+1}$. Let $\delta\,{=}\,\operatorname{ad}(u)$, $\partial\,{=}\,\operatorname{ad}(a\partial_i)$ and $v\in\mathfrak{u}_n$. We must show that $\delta^s (v)= 0$ for all $s\gg 1$. Applying (2.5) twice, we see that $\delta(v)\in\mathfrak{u}_{n,i}$ and $\delta^2(v)\in\mathfrak{u}_{n,i+1}$. Replacing the element $v$ by $\delta^2(v)$, we may assume without loss of generality that $v\in\mathfrak{u}_{n,i+1}$ and $v\neq 0$. Since $\mathfrak{u}_{n, i+1}=\bigoplus_{j=i+1}^n P_{j-1}\partial_j$, we have $v=b\partial_{i+1}+v'$ for some elements $b\in P_i$, $b\neq 0$, and $v'\in \mathfrak{u}_{n, i+2}$. For all integers $s\geqslant 1$, by (2.5) we have

By part 4, $\partial^s(b)= 0$ for all $s\gg 1$. Then $\delta^s (v)\in\mathfrak{u}_{n,i+2}$ for all $s\gg 1$. Similarly, $\delta^s(v)\in\mathfrak{u}_{n,i+3}$ for all $s\gg 1$. Repeating this argument several times, we see that $\delta^s(v)\in\mathfrak{u}_{n,n+1}= 0$ for all $s\gg 1$. This means that $\delta$ is a locally nilpotent map, as required.

Let us prove part 6. It is well known and easily provable that

Equation (2.7)

It follows that

Equation (2.8)

and, therefore,

The opposite inclusion $Z(\mathfrak{u}_n)\supseteq K\partial_n$ is obvious. Therefore $Z(\mathfrak{u}_n)=K\partial_n$.

Part 7 follows from part 2 since the lengths of the derived series of the $\mathfrak{u}_n$ are distinct (and invariant under isomorphisms). □

Proposition 2.1(5) enables us to produce many automorphisms of the Lie algebra $\mathfrak{u}_n$. For every element $a\in\mathfrak{u}_n$, the inner derivation $\operatorname{ad}(a)$ is locally nilpotent and, therefore,

The automorphism group $\rm{Aut}_K(\mathfrak{u}_n)$ of $\mathfrak{u}_n$ and explicit generators for it were found in [1], and the adjoint group $\langle e^{\operatorname{ad}(a)}\mid a\in\mathfrak{u}_n\rangle$ was shown to be a tiny part of the group $\operatorname{Aut}_K(\mathfrak{u}_n)$.

The next lemma classifies the nilpotent inner derivations of the $\mathfrak{u}_n$.

Lemma 2.2. Suppose that $a\in\mathfrak{u}_n$ and $\delta=\operatorname{ad}(a)$. Then the following assertions are equivalent.

  • 1)  
    The map $\delta$ is a nilpotent derivation of $\mathfrak{u}_n$.
  • 2)  
    $a\in P_{n-1}\partial_n$.
  • 3)  
    $\delta^2=0$.

Proof. The implications $2)\Rightarrow 3)\Rightarrow 1)$ are obvious.

Let us prove $1)\Rightarrow 2)$. We claim that $\delta$ is not nilpotent when $a\in\mathfrak{u}_n\setminus P_{n-1}\partial_n$. Indeed, suppose that $u\in\mathfrak{u}_n\setminus P_{n-1}\partial_n$, that is, $u=p_i\partial_i+p_{i+1}\partial_{i+1}+\dotsb +p_n\partial_n= p_i\partial_i+\dotsb$, where $p_j\in P_{j-1}$ for all $j=i,\dots,n$, $i<n$ and $p_i\neq 0$. Since

the derivation $\delta$ is not nilpotent. □

Let $A$ be a ring. A subset $S$ of $A$ is said to be multiplicative or multiplicatively closed if $1\in S$, $SS\subseteq S$ and $0\not\in S$. Every associative algebra $A$ may be regarded as a Lie algebra $(A,[\,\cdot\,{,}\,\cdot\,])$, where ${[}a,b]=ab-ba$ is the commutator of elements $a,b\in A$. For every $a\in A$, the map $\operatorname{ad}(a) \colon A\to A$, $b\mapsto [a,b]$, is a $K$-derivation of $A$ regarded as an associative algebra and a Lie algebra. It is called the inner derivation of $A$ associated with $a$. Thus the associative algebra $A$ and the Lie algebra $(A,[\,\cdot\,{,}\,\cdot\,])$ have the same set of inner derivations $\operatorname{Inn}(A)$ and the same centre $Z(A)$.

Let $\delta$ be a derivation of a ring $A$. For all elements $a,b\in A$,

Equation (2.9)

Equation (2.10)

Corollary 2.3. 

  • 1)  
    The inner derivations $\{\operatorname{ad}(u)\mid u\in\mathfrak{u}_n\}$ of the universal enveloping algebra $U_n$ of the Lie algebra $\mathfrak{u}_n$ are locally nilpotent derivations.
  • 2)  
    Every multiplicative subset $S$ of $U_n$ generated by an arbitrary set of elements of $\mathfrak{u}_n$ is a (left and right) Ore set in $U_n$. Therefore $S^{-1}U_n\simeq U_nS^{-1}$.

Proof. Part 1 follows from (2.9) and Proposition 2.1(5). Part 2 is an easy corollary of (2.10) and Proposition 2.1(5). □

Example 2.4. The set $S=\{\partial^\alpha\mid\alpha\in\mathbb{N}^n\}$ is a multiplicative subset of $U_n$. By Corollary 2.3(2), the ring of fractions $S^{-1}U_n$ exists.

Let $D$ be a ring, $\sigma=(\sigma_1,\dots,\sigma_n)$ an $n$-tuple of commuting automorphisms of $D$ ($\sigma_i\sigma_j=\sigma_j\sigma_i$ for all $i,j$), and $a=(a_1,\dots,a_n)$ an $n$-tuple of (non-zero) elements of the centre $Z(D)$ of $D$ such that $\sigma_i(a_j)=a_j$ for all $i\neq j$.

The generalized Weyl algebra $A=D(\sigma,a)$ (briefly GWA) of degree $n$ with base ring $D$ is the ring generated by $D$ and $2n$ indeterminates $x_1,\dots,x_n$, $y_1,\dots,y_n$ subject to the following defining relations [3], [4]:

where ${[}x, y]=xy-yx$. We say that $a$ and $\sigma$ are the sets of defining elements and automorphisms of $A$ respectively. Given a vector $k=(k_1,\dots,k_n)\in\mathbb{Z}^n$, we put $v_k=v_{k_1}(1)\dotsb v_{k_n}(n)$, where $v_{m}(i)=x_i^m$, $v_{-m}(i)=y_i^m$, $v_0(i)= 1$ for $1\leqslant i\leqslant n$ and $m\geqslant 0$. It follows from the definition of a GWA that

is a $\mathbb{Z}^n$-graded algebra ($A_kA_e\subset A_{k+e}$ for all $k,e\in\mathbb{Z}^n$), where $A_k=Dv_k=v_kD$.

Let $\mathcal{P}_n$ be the polynomial algebra $K[H_1,\dots,H_n]$ in $n$ indeterminates. We define an $n$-tuple $\sigma=(\sigma_1,\dots,\sigma_n)$ of $n$ commuting automorphisms of $\mathcal{P}_n$ by putting $\sigma_i(H_i)=H_i-1$ and $\sigma_i(H_j)=H_j$ for $i\neq j$. The map

is an isomorphism of $K$-algebras. We identify the Weyl algebra $A_n$ with the GWA above by means of this isomorphism. The Weyl algebra $A_n=\bigoplus_{\alpha\in\mathbb{Z}^n} A_{n,\alpha}$ is a $\mathbb{Z}$-graded algebra ($A_{n,\alpha}A_{n,\beta}\subseteq A_{n,\alpha +\beta }$ for all $\alpha,\beta\in\mathbb{Z}^n$).

The multiplicative sets $S=\{\partial^\alpha\mid\alpha\in\mathbb{N}^n\}$ and $T=\{x^\alpha\mid\alpha\in\mathbb{N}^n\}$ are (left and right) Ore sets of $A_n$, and we have

The Weyl algebra $(A_n,[\,\cdot\,{,}\,\cdot\,])$ is a $\mathbb{Z}^n$-graded Lie algebra, that is, ${[}A_{n,\alpha},A_{n,\beta}]\subseteq A_{n,\alpha +\beta}$ for all $\alpha,\beta\in\mathbb{Z}^n$. By definition, $\mathfrak{u}_n$ and $\operatorname{Der}_K(P_n)$ are $\mathbb{Z}^n$-graded Lie subalgebras of $A_n$.

The following lemma shows that $\mathfrak{u}_n$ contains finite-dimensional and infinite-dimensional maximal Abelian Lie subalgebras.

Lemma 2.5. 

  • 1)  
    $\mathcal{D}_n:=\bigoplus_{i=1}^n K\partial_i$ is a maximal Abelian Lie subalgebra of $\mathfrak{u}_n$ with $\operatorname{Cen}_{\mathfrak{u}_n}(\mathcal{D}_n)=\mathcal{D}_n$.
  • 2)  
    The ideal $P_{n-1}\partial_n$ of $\mathfrak{u}_n$ is a maximal Abelian Lie subalgebra of $\mathfrak{u}_n$ with $\operatorname{Cen}_{\mathfrak{u}_n}(P_{n-1}\partial_n)= P_{n-1}\partial_n$.

Proof. Part 1 follows from the equalities $\operatorname{Cen}_{A_n}(\mathcal{D}_n)=K[\partial_1,\dots,\partial_n]$ (see (2.7)) and $\operatorname{Cen}_{\mathfrak{u}_n}(\mathcal{D}_n)= \mathfrak{u}_n\cap K[\partial_1,\dots,\partial_n]=\mathcal{D}_n$.

Let us prove part 2. It follows from the equality $\operatorname{Cen}_{A_{n-1}}(P_{n-1})=P_{n-1}$ that

Then $\operatorname{Cen}_{\mathfrak{u}_n}(P_{n-1}\partial_n)= \mathfrak{u}_n\cap P_{n-1} [\partial_n]=P_{n-1}\partial_n$. Therefore the ideal $P_{n-1}\partial_n$ is a maximal Abelian Lie subalgebra of $\mathfrak{u}_n$. □

The inclusion $\mathfrak{u}_n\subseteq A_{n-1}\otimes K[\partial_n]$ induces an algebra homomorphism

Equation (2.11)

The image and kernel of $\chi_n$ are found in Theorem 2.6. By Corollary 2.3(2), the homomorphism $\chi_n$ can be extended to an algebra homomorphism

where $S=\{\partial^\alpha\mid\alpha\in\mathbb{N}^\alpha\}$. Clearly, $\chi_n (S^{-1}U_n)=B_{n-1}\otimes K[\partial_n,\partial_n^{-1}]$ (since $X_{\alpha,i}\partial_i^{-1}\mapsto x^\alpha\partial_i\partial_i^{-1}= x^\alpha $ and $x_i=\partial_i^{-1}H_i$).

We define a relation $\prec$ on the set $\mathbb{N}^n$ by writing $\alpha\prec\beta$ for elements $\alpha=(\alpha_i)$ and $\beta=(\beta_i)$ of $\mathbb{N}^n$ if and only if either $\alpha= 0$ and $\beta$ is arbitrary (that is, $0\prec\beta$ for all $\beta\in \mathbb{N}^n$), or $\alpha\neq 0$, $\beta\neq 0$ and $\max\{i\mid\alpha_i\neq 0\} < \max\{ i\mid\beta_i\neq 0\}$. Clearly, $\alpha\prec \beta$ and $\beta\prec\gamma$ imply that $\alpha\prec\gamma$, $\alpha\prec\alpha$ if and only if $\alpha\,{=}\,0$, and for all $\alpha,\beta\in\mathbb{N}^n\setminus \{0\}$, $\alpha\prec \beta$ implies that $\beta\not\prec\alpha$.

Theorem 2.6. 

  • 1)  
    The set $W_n:=\{x^\alpha\partial^\beta\mid\alpha, \beta\in\mathbb{N}^n;\,\alpha\prec\beta\}$ is a $K$-basis for the algebra $U_n':=\chi_n (U_n)$.
  • 2)  
    The kernel of $\chi_n $ is the ideal of $U_n$ generated by the elements $X_{\alpha,i}X_{\beta,j}-X_{0, i}X_{\alpha+\beta, j}$, where $i\leqslant j$, $\alpha\in\mathbb{N}^{i-1}$ and $\beta\in\mathbb{N}^{j-1}$.
  • 3)  
    $\chi_n$ is not surjective.
  • 4)  
    $U_n'$ is neither finitely generated, left Noetherian, nor right Noetherian.

Proof. Let us prove part 1. The elements of $W_n$ are $K$-linearly independent since they are as elements of the algebra $A_{n-1}\otimes K[\partial_n]$. Put $\mathcal{I}_n:=\sum_{w\in W_n}Kw$. We must show that $\chi_n(U_n)=\mathcal{I}_n$. The algebra $\chi_n(U_n)$ is generated by the elements $\chi_n(X_{\alpha,i})=x^\alpha\partial_i$. Using the relations

we see that $\chi_n(U_n)$ is contained in the linear span, call it $\mathcal{I}_n'$, of the elements

Using the commutation relations ${[}\partial_i, x_j]=\delta_{ij}$ (where $\delta_{ij}$ is the Kronecker delta), we can write each of these elements as a linear combination of elements of $W_n$. Therefore, $\chi_n(U_n)\subseteq\mathcal{I}_n'\subseteq\mathcal{I}_n$.

To prove the opposite inclusion $\chi_n(U_n)\supseteq\mathcal{I}_n$ and thus complete the proof of part 1, we must show that every element $w\neq 1$ in $W_n$ belongs to the algebra $\chi_n(U_n)$. The element $w\neq 1$ is a product

The case when $\alpha_1=\dotsb=\alpha_s= 0$ is obvious. Thus, we can assume that $\alpha_s\neq 0$. Every element $x_1^{\alpha_1}\dotsb x_s^{\alpha_s}\partial_1^{\beta_1}\dotsb\partial_s^{\beta_s}$ of the Weyl algebra $A_s$ can be written as a sum $\sum_{u,v\in\mathbb{N}^s}\lambda_{uv}\partial^ux^v$, where $\lambda_{uv}\in K$. Now,

This proves part 1. Moreover, the last step implies that the set

Equation (2.12)

is also a $K$-basis of $\chi_n (U_n)$. In more detail, $\chi_n(U_n)$ is the linear span of the set $W_n'$, whose elements are linearly independent since they are as elements of $A_n$. Therefore $W_n'$ is a $K$-basis of $\chi_n(U_n)$. This basis will be used in the proof of part 2.

We now prove part 2. Let $I$ be the ideal of $U_n$ generated by the elements listed in part 2 (the putative generators for $\operatorname{ker}(\chi_n)$). For every element $w'$ in $W_n'$, that is, for $\partial^\alpha$ and $\partial^\beta x^v\partial_t^i$, we choose an element $w''\in\chi_n^{-1}(w')$ in the following way:

Thus we have $\chi_n(w'')\,{=}\,w'$ for all $w'\in W_n'$. Let $W_n''$ be the set of all elements $w''$. The elements of $W_n''$ are linearly independent in $U_n$, being the pre-images of linearly independent elements. Put $\mathcal{I}_n'':=\sum_{w''\in W_n''}Kw''$. To finish the proof of part 2, it suffices to show that

(Suppose that this equality holds. Since $I\subseteq\operatorname{ker}(\chi_n)$ and the set $W_n''$ is mapped bijectively onto the basis $W_n'$ of $\chi_n(U_n)$, these two facts necessarily imply that $I=\operatorname{ker}(\chi_n)$.) To prove the equality $U_n=\mathcal{I}_n''+I$, we follow the line of the proof of part 1. The relations (2.2) and $X_{\alpha,i}X_{\beta,i}\equiv X_{0,i} X_{\alpha +\beta,i} \ \operatorname{mod}I$ imply that the linear span of the elements

where

generates the algebra $U_n$ modulo the ideal $I$. Using the generators for $I$ and the relations (2.2), we can write each of these elements as a linear combination of the elements $w''$ defined above. Then $U_n=\mathcal{ I}_n''+I$.

Let us prove part 3. Suppose that the homomorphism $\chi_n$ is surjective. We seek a contradiction. The ideal $\mathfrak{a}:= U_n\mathfrak{u}_n U_n$ of $U_n$ contains $\operatorname{ker}(\chi_n)$, $U_n/\mathfrak{a}\simeq K$, $\chi_n(\mathfrak{a})$ is an ideal of the algebra $\chi_n(U_n)=A_{n-1}\otimes K[\partial_n]$, and

The Weyl algebra $A_{n-1}$ is a simple infinite-dimensional algebra. So it is mapped isomorphically onto its image under the algebra homomorphism

a contradiction.

Part 4 follows from Proposition 3.13, where a stronger assertion will be proved. □

Corollary 2.7. The set $W_n'$ (see (2.12)) is a $K$-basis of the algebra $U_n'$.

Proof. This was established in the proof of part 1 of Theorem 2.6. □

§ 3. Classification of ideals of the Lie algebra $\mathfrak{u}_n$

In this section we introduce the uniserial and central dimensions and give a classification of the ideals of the Lie algebra $\mathfrak{u}_n$ (Theorem 3.6(1)). We prove that $\mathfrak{u}_n$ is a uniserial Artinian non-Noetherian Lie algebra of uniserial dimension $\operatorname{u.dim}(\mathfrak{u}_n) =\omega^{n-1}+\omega^{n-2}+\dotsb+\omega+1$ (Theorem 3.6(2)) and every ideal of $\mathfrak{u}_n$ is a characteristic ideal (Corollary 3.11). We also find the hypercentral series of $\mathfrak{u}_n$ and show that the central dimension of $\mathfrak{u}_n$ is given by $\operatorname{c.dim}(\mathfrak{u}_n)= \omega^{n-1}+\omega^{n-2}+\dotsb+\omega+1$ (Theorem 3.10).

3.1. Uniserial dimension.

Let $(S,\leqslant)$ be a partially ordered set, that is, the set $S$ admits a relation $\leqslant$ satisfying the following three conditions for all $a,b,c\in S$:

  • (i)  
    $a\leqslant a$,
  • (ii)  
    $a\leqslant b$ and $b\leqslant a$ imply that $a=b$,
  • (iii)  
    $a\leqslant b$ and $b\leqslant c$ imply that $a\leqslant c$.

A partially ordered set $(S,\leqslant)$ is said to be Artinian if every non-empty subset $T$ of $S$ has a least element, call it $t$, that is, $t\leqslant t'$ for all $t'\in T$. A partially ordered set $(S,\leqslant)$ is totally ordered if for all elements $a,b\in S$, either $a\leqslant b$ or $b\leqslant a$. A bijection $f\colon S\to S'$ between partially ordered sets $(S,\leqslant)$ and $(S',\leqslant)$ is an isomorphism if $a\leqslant b$ in $S$ implies that $f(a)\leqslant f(b)$ in $S'$. We recall that ordinal numbers are the isomorphism classes of totally ordered Artinian sets (that is, well-ordered sets). The ordinal number (isomorphism class) of a totally ordered Artinian set $(S,\leqslant)$ is denoted by $\operatorname{ord}(S)$. The class of all ordinal numbers is denoted by $\text{W}$. The class $\text{W}$ is totally ordered by inclusion $\leqslant$ and is Artinian. It is endowed with an associative addition $+$ and an associative multiplication $\cdot$ that extend the addition and multiplication of the non-negative integers. Every positive integer $n$ is identified with $\operatorname{ord}(1 < 2 < \dotsb < n)$. We put $\omega:= \operatorname{ord}(\mathbb{N},\leqslant)$. (More details on ordinal numbers can be found in [5].)

Definition 3.1. Let $(S,\leqslant)$ be a partially ordered set. The uniserial dimension $\operatorname{u.dim}(S)$ of $S$ is the supremum of the ordinal numbers $\operatorname{ord}(\mathcal{I})$, where $\mathcal{I}$ runs through all Artinian totally ordered subsets of $S$.

For a Lie algebra $\mathcal{G}$, let $\mathcal{J}_0(\mathcal{G})$ (resp. $\mathcal{J}(\mathcal{G})$) be the set of all ideals (resp. all non-zero ideals) of $\mathcal{G}$. Thus $\mathcal{J}_0(\mathcal{G})= \mathcal{J}(\mathcal{G})\cup\{0\}$. The sets $\mathcal{J}_0(\mathcal{G})$ and $\mathcal{J}(\mathcal{G})$ are partially ordered by inclusion. A Lie algebra $\mathcal{G}$ is said to be Artinian (resp. Noetherian) if the partially ordered set $\mathcal{J}(\mathcal{G})$ is Artinian (resp. Noetherian). This means that every descending (resp. ascending) chain of ideals stabilizes. A Lie algebra $\mathcal{G}$ is said to be uniserial if the partially ordered set $\mathcal{J}(\mathcal{G})$ is totally ordered. This means that for any two ideals $\mathfrak{a}$, $\mathfrak{b}$ of the Lie algebra $\mathcal{G}$ we have either $\mathfrak{a}\subseteq\mathfrak{b}$, or $\mathfrak{b}\subseteq\mathfrak{a}$.

Definition 3.2. Let $\mathcal{G}$ be an Artinian uniserial Lie algebra. The ordinal number $\operatorname{u.dim}(\mathcal{G}):= \operatorname{ord}(\mathcal{J}(\mathcal{G}))$ of the Artinian totally ordered set $\mathcal{J}(\mathcal{G})$ of non-zero ideals of $\mathcal{G}$ is called the uniserial dimension of $\mathcal{G}$. For an arbitrary Lie algebra $\mathcal{G}$, the uniserial dimension $\operatorname{u.dim}(\mathcal{G})$ is the supremum of $\operatorname{ord}(\mathcal{I})$, where $\mathcal{I}$ runs through all Artinian totally ordered sets of ideals.

If $\mathcal{G}$ is a Noetherian Lie algebra, then $\operatorname{u.dim}(\mathcal{G})\leqslant\omega$. Thus the uniserial dimension is a measure of the deviation from the Noetherian condition. The concept of uniserial dimension makes sense for any class of algebras (associative, Jordan, etc.).

Let $A$ be an algebra, $M$ an $A$-module, and $\mathcal{J}_l(A)$ (resp. $\mathcal{M}(M)$) the set of all non-zero left ideals of $A$ (resp. all non-zero submodules of $M$). These sets are partially ordered with respect to $\subseteq$. The left uniserial dimension of $A$ is defined as $\operatorname{u.dim}(A):=\operatorname{u.dim}(\mathcal{J}_l(A))$, and the uniserial dimension of $M$ is defined as $\operatorname{u.dim}(M):=\operatorname{u.dim}(\mathcal{M}(M))$.

3.2. An Artinian total ordering on the canonical basis $\mathcal{B}_n$ of $\mathfrak{u}_n$.

We define an Artinian total ordering $\leqslant $ on the canonical basis $\mathcal{B}_n$ of $\mathfrak{u}_n$ by putting $X_{\alpha,i}>X_{\beta,j}$ if and only if either $i<j$, or $i=j$ and $\alpha_{n-1}=\beta_{n-1},\dots,\alpha_{m+1}=\beta_{m+1}$, $\alpha_m > \beta_m$ for some $m$.

Example 3.3. For $n\,{=}\,2$ we have $\partial_2 < x_1\partial_2 < x_1^2\partial_2 < \dotsb < \partial_1 $. For $n\,{=}\,3$ we have

The following lemma is a straightforward consequence of the definition of the ordering $<$. We write $0<X_{\alpha,i}$ for all $X_{\alpha,i}\in\mathcal{B}_n$.

Lemma 3.4. If $X_{\alpha,i}>X_{\beta,j}$, then

  • 1)  
    $X_{\alpha+\gamma,i}>X_{\beta+\gamma,j}$ for all $\gamma\in\mathbb{N}^{i-1}$,
  • 2)  
    $X_{\alpha-\gamma,i}>X_{\beta-\gamma,j}$ for all $\gamma\in\mathbb{N}^{i-1}$ such that $\alpha-\gamma\in\mathbb{N}^{i-1}$ and $\beta-\gamma\in\mathbb{N}^{j-1}$,
  • 3)  
    ${[}\partial_k,X_{\alpha,i}]>[\partial_k,X_{\beta,j}]$ for all $k=1,\dots,i-1$ such that $\alpha_k\neq 0$,
  • 4)  
    ${[}X_{\gamma,k},X_{\alpha,i}]>[X_{\gamma,k},X_{\beta,j}]$ for all $X_{\gamma,k}>X_{\alpha,i}$ such that ${[}X_{\gamma, k},X_{\alpha,i}]\neq 0$, that is, $\alpha_k\neq 0$.

Let $\Omega_n$ be the set of indices $\{(\alpha,i)\}$ that parametrizes the canonical basis $\{X_{\alpha,i}\}$ of $\mathfrak{u}_n$. The set $(\Omega_n,\leqslant)$ is an Artinian totally ordered set, where $(\alpha,i)\geqslant (\beta,j)$ if and only if $X_{\alpha,i}\geqslant X_{\beta,j}$. It is isomorphic to the Artinian totally ordered set $(\mathcal{B}_n,\leqslant)$ via the map $(\alpha,i)\mapsto X_{\alpha,i}$. We identify the partially ordered sets $(\Omega_n,\leqslant)$ and $(\mathcal{B}_n,\leqslant)$ by means of this isomorphism. Clearly,

Equation (3.1)

$\Omega_2\subset\Omega_3\subset\dotsb $ and $\mathcal{B}_2\subset\mathcal{B}_3\subset\dotsb\,$. We put ${[}1,\operatorname{ord}(\Omega_n)]:=\{\lambda\in\text{W}\mid 1\leqslant\lambda \leqslant\operatorname{ord}(\Omega_n)\}$. By (2.2), if ${[}X_{\alpha,i},X_{\beta,j}]\neq 0$, then

Equation (3.2)

By (3.2), the map

Equation (3.3)

is a monomorphism of partially ordered sets ($\rho_n$ is an order-preserving injection). We shall prove that $\rho_n$ is a bijection (Theorem 3.6(1)) and, as a result, we will have a classification of the ideals of $\mathfrak{u}_n$. Each non-zero element $u$ of $\mathfrak{u}_n$ is a finite linear combination

where $\lambda,\mu,\dots,\nu\in K^*$ and $X_{\alpha,i}>X_{\beta,j}>\dotsb>X_{\sigma,k}$. The elements $\lambda X_{\alpha,i}$ and $\lambda\in K^*$ are called the leading term and leading coefficient of $u$ respectively, and the ordinal number denoted by $\operatorname{ord}(X_{\alpha,i})=\operatorname{ord}(\alpha,i) \in[1,\operatorname{ord}(\Omega_n)]$ and defined as the ordinal number that represents the Artinian totally ordered set $\{(\beta,j)\in\Omega_n\mid (\beta,j)\leqslant(\alpha,i)\}$, is called the ordinal degree of $u$ and is denoted by $\operatorname{ord}(u)$ (we hope that this notation will not lead to confusion). The following assertions hold for all non-zero elements $u,v\in\mathfrak{u}_n$ and all $\lambda\in K^*$:

  • (i)  
    $\operatorname{ord}(u+v)\leqslant\max\{\operatorname{ord}(u),\operatorname{ord}(v)\}$ provided that $u+v\neq 0$,
  • (ii)  
    $\operatorname{ord}(\lambda u)=\operatorname{ord}(u)$,
  • (iii)  
    $\operatorname{ord}([u,v])<\min\{\operatorname{ord}(u),\operatorname{ord}(v)\}$ provided that ${[}u,v]\neq 0$,
  • (iv)  
    $\operatorname{ord}(\sigma (u))=\operatorname{ord}(u)$ for all automorphisms $\sigma$ of $\mathfrak{u}_n$ (Corollary 3.12).

3.3. Classification of ideals of the Lie algebra $\mathfrak{u}_n$.

The following lemma plays a crucial role in the proof of Theorem 3.6.

Lemma 3.5. Let $I$ be a non-zero ideal of $\mathfrak{u}_n$. Then $I=\bigcup_{0\neq u\in I}I_{\operatorname{ord}(u)}$. In particular, $(v)=I_{\operatorname{ord}(v)}$ for all non-zero elements $v$ of $\mathfrak{u}_n$, where $(v)$ is the ideal of $\mathfrak{u}_n$ generated by the element $v$.

Proof. It suffices to prove the second assertion (that is, $(v)=I_{\operatorname{ord}(v)}$) because then

To prove that $(v)=I_{\operatorname{ord}(v)}$, we use a double induction: first on $n\geqslant 2$ and then, for a fixed $n$, on $\lambda=\operatorname{ord}(v)$. We may assume without loss of generality that the leading coefficient of $v$ is equal to 1. Notice that $\operatorname{ord}(v)$ is not a limit ordinal.

A. Let $n=2$. If $\lambda= 1$, then $v=\partial_2$ and, therefore, $(v)=K\partial_2=I_1$, as required. Suppose that $\lambda > 1$ and the desired assertion holds for all non-limit ordinals $\lambda'$ such that $\lambda' < \lambda$.

Case 1: $\lambda=\operatorname{ord}(x_1^i\partial_2)$ for some $i\geqslant 1$, that is,

for some scalars $\mu_j\in K$. Put $\delta=\operatorname{ad}(\partial_1)$. Then $I_\lambda\supseteq (v)\supseteq\sum_{j=0}^i K\delta^j(v)=I_\lambda$, that is, $(v)=I_\lambda$.

Case 2: $\lambda=\operatorname{ord}(\partial_1)=\omega+1$, that is, $v=\partial_1+p\partial_2$ for some element $p\in P_1$. For all $i\geqslant 1$ we have $(v)\ni[v,x_1^i\partial_2]=ix_1^{i-1}\partial_2$. By Case 1, $\mathfrak{u}_{2,1}\subseteq(v)$ and, therefore, $\partial_1\in(v)$. Hence $(v)=\mathfrak{u}_2=I_{\omega+1}=I_{\operatorname{ord}(v)}$.

B. Suppose that $n>2$ and the desired assertion holds for all $n'$ such that $n'< n$. If $\lambda= 1$, then $v\in K^*\partial_n$ and, therefore, $(v)=K\partial_n=I_1$, as required. Suppose that $\lambda > 1$ and the assertion holds for all non-limit ordinals $\lambda'$ such that $\lambda' < \lambda$.

Case 1: $\lambda=\operatorname{ord}(x^\alpha\partial_n)$ for some $\alpha$, $0\neq\alpha\in\mathbb{N}^{n-1}$, that is, $v= x^\alpha\partial_n+ \dotsb$, where the dots stand for smaller terms. The following claim is a corollary of (2.2) and the definition of the total ordering on the canonical basis $\mathcal{B}_n$ of $\mathfrak{u}_n$.

For every non-limit ordinal $\lambda'$ such that $\lambda' < \lambda$, there are elements of $\mathcal{B}_n$, say, $a_1,\dots,a_s$, of type $X_{\beta,i}$, $i\neq n$, such that

By induction on $\lambda$, it follows from this claim that $(v)=I_{\operatorname{ord}(v)}$.

Case 2: $v=\partial_i+\dotsb $ for some $i$ such that $i < n$. For all elements $x^\beta\partial_{i+1}$ with $\beta\in\mathbb{N}^i$ and $\beta_i\neq 0$ we have

whence $\operatorname{ord}([v,x^\beta\partial_{i+1}])= \operatorname{ord}(x^{\beta -e_i}\partial_{i+1})$. By induction, $(v)\supseteq\bigcup_{\mu < \lambda}I_\mu=:J$. But $I_\lambda\supseteq(v)=Kv+J=K\partial_i+J=I_\lambda$ and, therefore, $(v)=I_\lambda$.

Case 3: $v=x^\alpha\partial_i+\dotsb$ for some $i$ such that $i<n$ and $\alpha\in\mathbb{N}^{i-1}\setminus\{0\}$. Notice that

By Case 2, $I_{\operatorname{ord}(\partial_i)}\subseteq (v)$. Since $\mathfrak{u}_{n,i+1}\subseteq I_{\operatorname{ord}(\partial_i)}$, we have $\mathfrak{u}_{n,i+1}\subseteq (v)$. By (2.6), $\mathfrak{u}_i\simeq\mathfrak{u}_n/\mathfrak{u}_{n,i+1}$. By assumption, $i < n$. Consider the element $v+\mathfrak{u}_{n,i+1}\in\mathfrak{u}_i$. Then the desired assertion follows by induction on $i$. □

Given any $u,v\in\mathfrak{u}_n\setminus\{0\}$, we have $(u)\subseteq(v)$ if and only if $\operatorname{ord}(u)\leqslant\operatorname{ord}(v)$.

Theorem 3.6. 

  • 1)  
    The map (3.3) is a bijection.
  • 2)  
    The Lie algebra $\mathfrak{u}_n$ is uniserial, Artinian and non-Noetherian. Its uniserial dimension is given by $\operatorname{u.dim}(\mathfrak{u}_n)=\operatorname{ord}(\Omega_n)= \omega^{n-1}+\omega^{n-2}+\dotsb+\omega+1$.

Proof. Part 1 follows at once from Lemma 3.5. Part 2 follows from part 1. □

An ideal $\mathfrak{a}$ of a Lie algebra $\mathcal{G}$ is said to be proper (resp. cofinite) if $\mathfrak{a}\neq 0,\mathcal{G}$ (resp. $\operatorname{dim}_K(\mathcal{G}/\mathfrak{a}) < \infty$).

Corollary 3.7. 

  • 1)  
    The ideal $\mathfrak{u}_{n,2}$ is the largest proper ideal of $\mathfrak{u}_n$.
  • 2)  
    The ideal $\mathfrak{u}_{n,2}$ is the only proper cofinite ideal of $\mathfrak{u}_n$, and $\operatorname{dim}_K(\mathfrak{u}_n/\mathfrak{u}_{n,2})\,{=}\,1$.
  • 3)  
    The centre $Z(\mathfrak{u}_n)=K\partial_n$ of $\mathfrak{u}_n$ is the least non-zero ideal of $\mathfrak{u}_n$.
  • 4)  
    The ideals $I_s:=\sum_{i=0}^{s-1} Kx_1^i\partial_n$, $s=1,2,\dots $, are the only finite-dimensional ideals of $\mathfrak{u}_n$, and $\operatorname{dim}_K(I_s)=s$.

Proof. All these statements are easy corollaries of Theorem 3.6. □

3.4. The centralizers of ideals of $\mathfrak{u}_n$.

In combination with Theorem 3.6, the following proposition describes the centralizers of all ideals of the Lie algebra $\mathfrak{u}_n$. Notice that the centralizer of an ideal of a Lie algebra is also an ideal.

Proposition 3.8. 

  • 1)  
    We have
  • 2)  
    The set $\mathcal{C}(\mathfrak{u}_n)$ of centralizers of all ideals of $\mathfrak{u}_n$ contains precisely $2n\,{-}\,1$ elements, and the map $\mathcal{C}(\mathfrak{u}_n)\,{\to}\,\mathcal{C}(\mathfrak{u}_n)$, $C\,{\mapsto}\,\operatorname{Cen}_{\mathfrak{u}_n}(C)$, is a bijection. Moreover, it is an inclusion-reversing involution, that is, $\operatorname{Cen}_{\mathfrak{u}_n}(\operatorname{Cen}_{\mathfrak{u}_n}(C))\,{=}\,C$ for all $C\,{\in}\,\mathcal{C}(\mathfrak{u}_n)$.
  • 3)  
    We have $\operatorname{Cen}_{\mathfrak{u}_n}(\mathfrak{u}_n)=K\partial_n$, $\operatorname{Cen}_{\mathfrak{u}_n}(K\partial_n)=\mathfrak{u}_n$, $\operatorname{Cen}_{\mathfrak{u}_n}(P_{n-1}\partial_n)=P_{n-1}\partial_n$, $\operatorname{Cen}_{\mathfrak{u}_n}(P_i\partial_n)\,{=}\,\bigoplus_{j=i+1}^nP_{j-1}\partial_j$, $\operatorname{Cen}_{\mathfrak{u}_n}(\bigoplus_{j=i+1}^nP_{j-1}\partial_j)\,{=}\, P_i\partial_n$ for $i\,{=}\,1,\dots,n- 2$.

Proof. Put $C_\lambda:=\operatorname{Cen}_{\mathfrak{u}_n}(I_{\lambda,n})$. If $\lambda=\operatorname{ord}(\Omega_n)$, then $I_{\lambda,n}=\mathfrak{u}_n$ and $C_\lambda=Z(\mathfrak{u}_n)=K\partial_n$ by Proposition 2.1(6). If $\lambda= 1$, then $I_{1, n}=K\partial_n=Z(\mathfrak{u}_n)$ and, therefore, $C_1=\mathfrak{u}_n$. We can assume that $\lambda\neq 1, \operatorname{ord}(\Omega_n)$. To prove part 1, we use induction on $n\geqslant 2$. For $n=2$, the only case to consider is when $\lambda\in(1,\omega]$. In this case,

and, therefore, $C_\lambda=P_1\partial_2$.

Suppose that $n>2$ and part 1 holds for all $n'<n$.

We claim that $C_{\omega^m}= \bigoplus_{i=m+1}^nP_{i-1}\partial_i$ for $m=1,\dots,n-1$.

Indeed, for $m {=} n-1$ we have $I_{\omega^{n-1}\!,n} {=} P_{n{-}1}\partial_n$ and our claim is just Lemma 2.5(2). Suppose that $m<n-1$. Then $I_{\omega^m, n}=P_m\partial_n\subset I_{\omega^{n-1}, n}$, whence $C_{\omega^m}\supseteq C_{\omega^{n-1}}=P_{n-1}\partial_n$. This means that the $\mathfrak{u}_n$-module $I_{\omega^m, n}$ is also a $(\mathfrak{u}_n/P_{n-1}\partial_n)$-module and a $\mathfrak{u}_{n-1}$-module since $\mathfrak{u}_{n-1}\simeq \mathfrak{u}_n/P_{n-1}\partial_n$. The map

is an isomorphism of $\mathfrak{u}_{n-1}$-modules (since $\partial_n\in Z(\mathfrak{u}_n)$ and $\partial_{n-1}\in Z(\mathfrak{u}_{n-1})$), and our claim follows by induction on $n$.

Suppose that $1<\lambda\leqslant\omega^{n-1}$, that is, $\lambda\in (\omega^{m-1},\omega^m]$ for some $m\in\{ 1,\dots,n-1\}$. We must show that $C_\lambda=C_{\omega^m}$ (see the claim). By the claim, the inclusions $I_{\omega^{m-1}, n}\subset I_{\lambda,n} \subseteq I_{\omega^m,n}$ imply that $C_{\omega^{m-1}}\supseteq C_\lambda\supseteq C_{\omega^m}$. Notice that $C_{\omega^{m-1}}= C_{\omega^m}\oplus P_{m-1}\partial_m$, $x_m\partial_n\in I_{\lambda,m}$ (since $\lambda\in(\omega^{m-1},\omega^m]$) and ${[}p\partial_m,x_m\partial_n]= p\partial_n\neq 0$ for all non-zero elements $p\in P_{m-1}$. It follows that $C_\lambda=C_{\omega^m}$, as required.

Suppose that $\omega^{n-1} < \lambda < \operatorname{ord}(\Omega_n)= \omega^{n-1}+\dotsb+\omega+1$, that is, $\lambda\in(\omega^{n-1}+\dotsb+ \omega^{i+1}, \omega^{n-1}+\dotsb+\omega^i]$ for some $i\in\{ 1,\dots,n-2\}$. Put $\lambda_i:=\omega^{n-1}+ \dotsb+ \omega^i$. Then $I_{\lambda,n}\subseteq I_{\lambda_i,n}\,{=}\,\bigoplus_{j=i+1}^nP_{j-1}\partial_j\,{=}\, \operatorname{Cen}_{\mathfrak{u}_n}(I_{\omega^i,n})$ by the claim. Therefore, $C_\lambda\supseteq I_{\omega^i,n}=P_i\partial_n$. It follows from the inclusion $I_{\lambda,n}\supset P_{n-1}\partial_n=I_{\omega^{n-1},n}$ that $C_\lambda\subseteq C_{\omega^{n-1}}=P_{n-1}\partial_n$ by the claim. Then the inclusion $\{\partial_{i+1},\dots, \partial_n\}\subseteq I_{\lambda,n}$ implies that $C_\lambda\subseteq\operatorname{Cen}_{P_{n-1}}(\partial_{i+1},\dots, \partial_n)\partial_n=P_i\partial_n=I_{\omega^i, n}$. This means that $C_\lambda= I_{\omega^i,n}$.

Part 2 follows from part 3. Part 3 follows from part 1. □

3.5. The hypercentral series and central dimension.

For a Lie algebra $\mathcal{G}$ over a field $K$ we define its hypercentral series $\{ Z^{(\lambda)}(\mathcal{G})\}_{\lambda\in \text{W} }$ recursively. We put $Z^{(0)}(\mathcal{G})\,{:=}\,Z(\mathcal{G})$. If $\lambda $ is not a limit ordinal, that is, $\lambda\,{=}\,\mu+1$ for some $\mu\,{\in}\,\text{W}$, then

If $\lambda$ is a limit ordinal, then $Z^{(\lambda)}(\mathcal{G}) {:=} \bigcup_{\mu < \lambda}\!Z^{(\mu)}(\mathcal{G})$. If $\lambda {\leqslant} \mu$, then $Z^{(\lambda)}(\mathcal{G}) {\subseteq} Z^{(\mu)}(\mathcal{G})$. Thus, $\{ Z^{(\lambda)}(\mathcal{G})\} _{\lambda\in\text{W}}$ is an ascending chain of ideals of $\mathcal{G}$. We put $Z^{(\text{W})}(\mathcal{G}):= \bigcup_{\lambda\in\text{W}}Z^{(\lambda)}(\mathcal{G})$.

Definition 3.9. The minimal ordinal number $\lambda$ (if it exists) such that $Z^{(\lambda)}(\mathcal{G})=Z^{(\text{W})}(\mathcal{G})$, is called the central dimension of the Lie algebra $\mathcal{G}$ and is denoted by $\operatorname{c.dim}(\mathcal{G})$. If there is no such $\lambda $, we write $\operatorname{c.dim}(\mathcal{G})=\text{W}$. The concept of central dimension makes sense for any class of algebras (associative, Jordan etc.).

A Lie algebra $\mathcal{G}$ is central (that is, $Z(\mathcal{G})\,{=}\,0$) if and only if $\ \operatorname{c.dim}(\mathcal{G})\,{=}\,0$. Thus the central dimension measures the deviation from 'being central'. The following theorem describes the hypercentral series for the Lie algebra $\mathfrak{u}_n$ and gives $\operatorname{c.dim}(\mathfrak{u}_n)$.

Theorem 3.10. The hypercentral series $\{ Z^{(\lambda)}(\mathfrak{u}_n)\}_{\lambda\in \text{W}}$ stabilizes precisely at step $\operatorname{ord}(\Omega_n)= \omega^{n-1}+\omega^{n-2}+\dotsb+\omega+1$, that is, $\operatorname{c.dim}(\mathfrak{u}_n)=\operatorname{ord}(\Omega_n)$. Moreover, $Z^{(\lambda)}(\mathfrak{u}_n)=I_\lambda$ for every $\lambda\in[1,\operatorname{ord}(\Omega_n)]$. In particular, $Z^{(\operatorname{c.dim}(\mathfrak{u}_n))}(\mathfrak{u}_n)=\mathfrak{u}_n$.

Proof. It suffices to show that $Z^{(\lambda)}=I_\lambda$ for all $\lambda\in[1,\operatorname{ord}(\Omega_n)]$, where $Z^{(\lambda)}:=Z^{(\lambda)}(\mathfrak{u}_n)$. We use induction on $\lambda$. The case $\lambda= 1$ follows from Proposition 2.1(6): $Z^{(1)}=Z(\mathcal{G})=K\partial_n=I_1$. Suppose that $\lambda > 1$ and the desired equality holds for all ordinals $\lambda'<\lambda$.

If $\lambda$ is a limit ordinal, then

Suppose that $\lambda$ is not a limit ordinal, that is, $\lambda=\mu+1$ for some ordinal number $\mu$.

Case 1: $\lambda\leqslant\omega^{n-1}$. Then we necessarily have $\lambda<\omega^{n-1}$ since $\lambda$ is not a limit ordinal but $\omega^{n-1}$ is. Clearly, $I_\lambda\subseteq Z^{(\lambda)}$ by (2.2) (since $Z^{(\mu)}=I_\mu$ by induction). Suppose that $I_\lambda\neq Z^{(\lambda)}$. Then necessarily $I_{\lambda +1}\subseteq Z^{(\lambda)}$ (by Theorem 3.6(1)). If $\lambda=X_{\alpha,i}$ for some $\alpha\in\mathbb{N}^{n-1}$ (recall that we identify $(\Omega_n,\leqslant)$ and $(\mathcal{B}_n,\leqslant)$), then $\lambda +1= X_{\alpha +e_1, n}$ and $ X_{\alpha +e_1, n}\in I_{\lambda+1}$, but ${[}\partial_1,X_{\alpha +e_1,n}]=(\alpha_1+1)X_{\alpha, n}\not\in Z^{(\mu)}$, a contradiction. Therefore, $Z^{(\lambda)}=I_\lambda$.

Case 2: $\lambda>\omega^{n-1}$. Notice that $Z^{(\lambda)}\supseteq I_\lambda\varsupsetneq I_{\omega^{n-1}}= \mathfrak{u}_{n,n}$ and $\mathfrak{u}_n/\mathfrak{u}_{n,n}\simeq \mathfrak{u}_{n-1}$ by (2.6). We now complete the argument by induction on $n$. The base of the induction $(n=2)$ is covered by Cases 1, 2 for $n=2$ above. □

An ideal $I$ of a Lie algebra $\mathcal{G}$ is said to be characteristic if it is invariant under all automorphisms of $\mathcal{G}$, that is, $\sigma (I)=I$ for all $\sigma\in\operatorname{Aut}_K(\mathcal{G})$. Clearly, an ideal $I$ is characteristic if and only if $\sigma (I)\subseteq I$ for all $\sigma\in\operatorname{Aut}_K(\mathcal{G})$.

Corollary 3.11. All ideals of the Lie algebra $\mathfrak{u}_n$ are characteristic.

Proof. By definition, the members of the hypercentral series $\{ Z^{(\lambda)}(\mathfrak{u}_n)\}$ are characteristic ideals of $\mathfrak{u}_n$. But Theorem 3.10 shows that these are all the ideals of $\mathfrak{u}_n$. We can also deduce this corollary from Theorem 3.6(2). □

Corollary 3.12. For all non-zero elements $u\in\mathfrak{u}_n$ and all automorphisms $\sigma$ of $\mathfrak{u}_n$ we have $\operatorname{ord}(\sigma(u))=\operatorname{ord}(u)$.

Proof. Notice that

Equation (3.4)

The statement now follows from the fact that all the ideals $I_\lambda$ are characteristic (Corollary 3.11).

3.6. The subalgebra $U_n'$ of the Weyl algebra $A_n$.

Let $\mathfrak{a} $ be an ideal of a Lie algebra $\mathcal{G}$. Then the ideal $U(\mathcal{G})\mathfrak{a} U(\mathcal{G})$ of the universal enveloping algebra $U(\mathcal{G})$ generated by $\mathfrak{a}$ is equal to $U(\mathcal{G})\mathfrak{a}= \mathfrak{a} U(\mathcal{G})$. The chain

of ideals of $\mathfrak{u}_n$ yields a chain of subalgebras and a chain of ideals of $U_n$:

We put $U_{n,\lambda }'{:=}\chi_n(U(I_\lambda))$ and $I_\lambda'{:=} \chi_n(U_nI_\lambda)$, where $\chi_n$ is the algebra homomorphism (2.11). Then

Equation (3.5)

is a chain of subalgebras of $U_n'$, and

Equation (3.6)

is a chain of ideals of $U_n'$.

Proposition 3.13. 

  • 1)  
    All the inclusions in (3.5) are strict. The algebra $U_n'$ is not finitely generated.
  • 2)  
    All the inclusions in (3.6) are strict. In particular, $U_n'$ is neither left Noetherian nor right Noetherian and does not satisfy the ascending chain condition for ideals.

Proof. 1. We use induction on $n$. Suppose that $n=2$. For every $i\in [1,\omega)$, the algebra $U_{2,i}'$ is generated by the commuting elements $x_1^j\partial_2$ $(0\leqslant j\leqslant i-1)$ since

These equalities mean that $U_{2,i}'$ is isomorphic to the monoid algebra $K\mathcal{M}_i$ (via $x_1^j\partial_2\,{\mapsto}\,(j,1)$), where $\mathcal{M}_i$ is the submonoid of $(\mathbb{N}^2, +)$ generated by the elements $(j,1)$, $0\leqslant j\leqslant i-1$. It follows that

This means that the inclusions $U_{2,1}'\subset U_{2,2}'\subset \dotsb\subset U_{2,i}'\subset\dotsb\subset U_{2,\omega}'=\bigcup_{i\geqslant 1}U_{2,i}'$ are strict and, therefore, the algebra $U_{2,\omega}'$ is not finitely generated. We have $\partial_1\in U_2'\setminus U_{2,\omega }'$, $U_2'=U_{2,\omega}'[\partial_1;\operatorname{ad}\ (\partial_1)]$ is the skew polynomial algebra and ${[}\partial_1, U_{2,i}']\subseteq U_{2,i}'$ for all $i\geqslant 1$. Therefore part 1 holds for $n=2$.

Suppose that $n > 2$ and part 1 holds for all $n'<n$. Let $\lambda\in [1,\omega^{n-1})$. Then $\lambda+ 1$ is not a limit ordinal. Notice that ${[}1,\omega^{n-1})\subseteq[1,\operatorname{ord}(\Omega_n))$. Hence $\lambda+1=(\alpha,n)$. The elements $\{ x^{\beta}\partial_n\mid\beta\in\mathbb{N}^{n-1}\}$ commute. Moreover,

Therefore the algebra $U_{n,\lambda}'$ is commutative and is isomorphic to the monoid algebra $K\mathcal{M}_{n,\lambda}$ (via $x^{\beta}\partial_n\mapsto(\beta,1)$), where $\mathcal{M}_{n,\lambda}$ is the submonoid of $(\mathbb{N}^n,+)$ generated by the elements $\{(\beta, 1)\mid(\beta,n)\leqslant\lambda\}$. It follows that

This means that

(the inclusions are strict). Let $\lambda\geqslant\omega^{n-1}$. By Theorem 2.6(1) and (3.3),

Equation (3.7)

By induction on $n$, part 1 holds.

2. We make free use of the facts established in the proof of part 1. Arguing by induction on $n$, we first take $n=2$. For every $i\in [1,\omega)$ the ideal $I_{2,i}'$ of $U_2'$ is generated by the commuting elements $x_1^j\partial_2$, where $0\leqslant j\leqslant i-1$. By Corollary 2.7, the set $W_2'=\{\partial_1^i\partial_2^j,\partial_1^ix_1^k\partial_2\cdot\partial_2^j \mid i,j\in\mathbb{N},k\geqslant 1\}$ is a $K$-basis of $U_2'$. For each $i\in [1,\omega)$,

It follows that

This means that

Since $1\in U_2'\setminus I_{2,\omega +1}'$ and $\partial_1\in I_{2,\omega+1}'\setminus I_{2,\omega}'$, part 2 holds for $n=2$.

Suppose that $n>2$ and part 2 holds for all $n'<n$. Let $\lambda\in [1,\omega^{n-1})$. Then $\lambda+ 1$ is not a limit ordinal. Notice that ${[}1,\omega^{n-1})\subseteq[1,\operatorname{ord}(\Omega_n))$. Hence $\lambda+1=(\alpha,n)$. It follows from the equality $I_{n,\lambda +1}'= U_n'\chi_n(I_{\lambda +1})=\sum_{(\beta,n)\leqslant \lambda +1}U_n'x^\beta\partial_n$ and (2.2) that

Equation (3.8)

Therefore,

This means that

Let $\lambda >\omega^{n-1}$. In view of (3.7) we also have

Equation (3.9)

By induction on $n$, part 2 holds. □

3.7. The Heisenberg Lie subalgebras of $\mathfrak{u}_n$.

Let

be the Lie algebra of $n\times n$ matrices, where $E_{ij}$ are the elementary matrices. For every $n\geqslant 2$, let $\operatorname{UT}_n(K)=\bigoplus_{i\leqslant j}KE_{ij}$ be the Lie algebra of upper triangular $n\times n$ matrices, $\operatorname{UT}_n(K)\subseteq\operatorname{gl}_n(K)$. The $K$-linear map

is a Lie algebra monomorphism. We identify the Lie algebra $\operatorname{UT}_n(K)$ with its image in $\mathfrak{u}_n$.

The Heisenberg Lie algebra $\mathcal{H}_n$ is the $(2n+1)$-dimensional Lie algebra with $K$-basis $X_1,\dots,X_n$, $ Y_1,\dots,Y_n, Z$ such that $Z$ is a central element of $\mathcal{ H}_n$ and

where $\delta_{ij}$ is the Kronecker delta. The $K$-linear map

is a Lie algebra monomorphism. We identify $\mathcal{H}_{n-1}$ with its image in $\mathfrak{u}_n$. Since $\mathcal{H}_{i-1}\subseteq\mathfrak{u}_i\subseteq\mathfrak{u}_n$ for $i=2,\dots,n-1$, $\mathfrak{u}_n$ contains the $\mathcal{H}_i$, $i=1,\dots,n-1$. For all positive integers $i$ and $j$ with $i\neq j$ we have $\mathcal{H}_i\cap\mathcal{H}_j= \sum_{k=1}^{\min\{ i,j\}}K\partial_k$.

Let $A$ be an algebra or a Lie algebra and $M$ an $A$-module. Then $\operatorname{ann}_A(M):=\{a\in A\mid aM=0\}$ is called the annihilator of $M$. This is an ideal of $A$. A module is said to be faithful if its annihilator is 0.

3.8. The $\mathfrak{u}_n$-module $P_n$.

To say that $P_n$ is a $\mathfrak{u}_n$-module is the same as to say that $P_n$ is a $U_n'$-module. This obvious observation is important because it enables us to use the relations of the Weyl algebra $A_n$ in various computations with $U_n'$ (since $U_n'\subseteq A_n$). For every $n\geqslant 2$, the Lie algebra $\mathfrak{u}_n$ is a Lie subalgebra of $\mathfrak{u}_{n+1}=\mathfrak{u}_n\oplus P_n\partial_{n+1}$, where $P_n\partial_{n+1}$ is an ideal of $\mathfrak{u}_{n+1}$ and ${[}P_n\partial_{n+1},P_n\partial_{n+1}]= 0$. In particular, $P_n\partial_{n+1}$ is a left $\mathfrak{u}_n$-module, where the action of $\mathfrak{u}_n$ on $P_n\partial_{n+1}$ is given by the rule $uv:=[u,v]$ for all $u\in\mathfrak{u}_n$ and $v\in P_n\partial_{n+1}$. We recall that the polynomial algebra $P_n$ is a left $\mathfrak{u}_n$-module.

Lemma 3.14. 

  • 1)  
    The $K$-linear map $P_n\to P_n\partial_{n+1}$, $p\mapsto p\partial_{n+1}$, is an isomorphism of $\mathfrak{u}_n$-modules.
  • 2)  
    The $\mathfrak{u}_n$-module $P_n$ is indecomposable and uniserial. Moreover, $\operatorname{u.dim}(P_n) {=}\omega^n$ and $\operatorname{ann}_{\mathfrak{u}_n}(P_n)= 0$.
  • 3)  
    The set $\{ P_{\lambda,n}:=\bigoplus_{\alpha\in\mathbb{N}^n, (\alpha,n+1)\leqslant\lambda}Kx^\alpha\mid\lambda\in[1,\omega^n]\}$ is the set of all non-zero $\mathfrak{u}_n$-submodules of $P_n$. We have $P_{\lambda,n}\subset P_{\mu,n}$ if and only if $\lambda<\mu$. Moreover, $\operatorname{u.dim}(P_{\lambda,n})=\lambda$ for all $\lambda\in [1,\omega^n]$.
  • 4)  
    The $\mathfrak{u}_n$-submodules of $P_n$ are pairwise non-isomorphic indecomposable uniserial $\mathfrak{u}_n$-modules.

Proof. To prove part 1, we notice that the map indicated is a bijection and a $\mathfrak{u}_n$-homomorphism.

Let us prove part 3. By part 1, the $\mathfrak{u}_n$-module $P_n$ can be identified with the ideal $P_n\partial_{n+1}$ of $\mathfrak{u}_{n+1}$. Under this identification, every $\mathfrak{u}_n$-submodule of $P_n$ becomes an ideal of $\mathfrak{u}_{n+1}$ in $P_n\partial_{n+1}$, and vice versa. Part 3 now follows from part 1 and the classification of ideals of $\mathfrak{u}_{n+1}$ (Theorem 3.6(1)).

We now prove part 2. By part 3, the $\mathfrak{u}_n$-module $P_n$ is uniserial and, therefore, indecomposable and $\operatorname{u.dim}(P_n)= \omega^n$. The Weyl algebra $A_n$ is a simple algebra, whence $\operatorname{ann}_{A_n}(P_n)= 0$. Then we have $\operatorname{ann}_{\mathfrak{u}_n}(P_n)= \mathfrak{u}_n\cap\operatorname{ann}_{A_n}(P_n)= 0$ because $\mathfrak{u}_n\subseteq A_n$.

Let us prove part 4. By part 3 we have $\operatorname{u.dim}(P_{\lambda,n})=\lambda$. Hence $P_{\lambda, n}\not\simeq P_{\mu,n}$ for all $\lambda\neq\mu$. The rest is obvious (see part 2). □

The following corollary describes the annihilators of all the $\mathfrak{u}_n$-submodules of $P_n$. In particular, it classifies the faithful ones.

Corollary 3.15. 

  • 1)  
    The $\mathfrak{u}_n$-submodule $P_{\lambda,n}$ of $P_n$ is a faithful submodule if and only if $\lambda\in [\omega^{n-1}+1,\omega^n]$.
  • 2)  
    $\operatorname{ann}_{\mathfrak{u}_n}(P_{\omega^{n-1}, n})= P_{n-1}\partial_n$.
  • 3)  
    $\operatorname{ann}_{\mathfrak{u}_n}(P_{\lambda,n})= \begin{cases} 0&\textit{if}\quad\lambda\in(\omega^{n-1},\omega^n], \\ \displaystyle \bigoplus_{i=m+1}^nP_{i-1}\partial_i&\textit{if}\quad \lambda\in(\omega^{m-1},\omega^m], \\ &\qquad\qquad\qquad m=1,\dots,n-1,\\ \mathfrak{u}_n &\textit{if}\quad\lambda=1.\end{cases}$

Proof. We first prove part 2. The inclusion $\supseteq $ is obvious. The opposite inclusion $\subseteq $ follows since $P_{n-1}$ is a faithful $\mathfrak{u}_{n-1}$-module (Lemma 3.14(2)), $\mathfrak{u}_{n-1}\simeq\mathfrak{u}_n/ P_{n-1}\partial_n$ and $P_{n-1}=P_{\omega^{n-1}, n}$ is a $(\mathfrak{u}_n/P_{n-1}\partial_n)$-module.

To prove part 1, we notice that if $N$ is a submodule of $M$, then $\operatorname{ann}(N)\supseteq\operatorname{ann}(M)$. In view of this fact and part 2, to finish the proof of part 1, it suffices to show that the $\mathfrak{u}_n$-module $P_{\omega^{n-1}+1, n}$ is faithful. Since $P_{\omega^{n-1}, n}\subseteq P_{\omega^{n-1}+1, n}$, we have $\mathfrak{a}:=\operatorname{ann}_{\mathfrak{u}_n}(P_{\omega^{n-1}+1, n})\subseteq \operatorname{ann}_{\mathfrak{u}_n}(P_{\omega^{n-1}, n})=P_{n-1}\partial_n$. Since $x_n\in P_{\omega^{n-1}+1, n}$ and $p\partial_n *(x_n)=p\neq 0$ for all non-zero elements $p\in P_{n-1}$, we have $\mathfrak{a}= 0$.

Let us prove part 3. We use induction on $n\geqslant 2$. The base of the induction is $n=2$, and there are three cases to consider: $\lambda\in (\omega,\omega^2]$, $\lambda=(1,\omega]$ and $\lambda= 1$. The first case is part 1, the last case is obvious, and in the second,

Suppose that $n>2$ and part 3 holds for all $n'<n$. If $\lambda\in(\omega^{n-1},\omega^n]$, then the $\mathfrak{u}_n$-module $P_{\lambda,n}$ is faithful by part 1. If $\lambda= 1$, then $\operatorname{ann}_{\mathfrak{u}_n}(P_{1, n})=\operatorname{ann}_{\mathfrak{u}_n}(K) = \mathfrak{u}_n$. If $1<\lambda\leqslant\omega^{n-1}$, that is, $\lambda\in(\omega^{m-1},\omega^m]$ for some $m\in\{1,\dots,n-1\}$, then $P_{\lambda,n}\subseteq P_{n-1}$. By part 2 we have $\mathfrak{a}:=\operatorname{ann}_{\mathfrak{u}_n}(P_{\lambda,n})\supseteq \operatorname{ann}_{\mathfrak{u}_n}(P_{\omega^{n-1}, n})= P_{n-1}\partial_n$. Since $\mathfrak{u}_{n-1}\simeq\mathfrak{u}_n/P_{n-1}\partial_n$, the inclusion $P_{\lambda,n}\subseteq P_{n-1}$ is an inclusion of $\mathfrak{u}_{n-1}$-modules. Moreover, the $\mathfrak{u}_{n-1}$-submodule $P_{\lambda,n}$ of $P_{n-1}$ can be identified with the $\mathfrak{u}_{n-1}$-submodule $P_{\lambda,n-1}$ of $P_{n-1}$. The result now follows by induction on $n$. □

Corollary 3.16. If $n\geqslant 3$, then the ideal $I_{\omega^{n-2}+1}=Kx_{n-1}\partial_n+ \sum_{\alpha\in\mathbb{N}^{n-2}}Kx^\alpha\partial_n$ is the least ideal $I$ of the Lie algebra $\mathfrak{u}_n$ which is a faithful $\mathfrak{u}_{n-1}$-module (that is, $\operatorname{ann}_{\mathfrak{u}_{n-1}}(I)= 0$).

Proof. By Lemma 3.14(1) and Corollary 3.15(1), the $\mathfrak{u}_{n-1}$-module $I_{\omega^{n-2}+1}$ is faithful, but its predecessor $I_{\omega^{n-2}}$ is not since

The inclusion $\mathfrak{u}_n\subseteq\mathfrak{u}_{n+1}= \mathfrak{u}_{n}\oplus P_{n-1}\partial_n$ of Lie algebras respects the total orderings on the bases $\mathcal{B}_{n}$ and $\mathcal{B}_{n+1}$. The isomorphism $P_n\to P_n\partial_{n+1}$, $p\mapsto p\partial_{n+1}$ of $\mathfrak{u}_n$-modules (Lemma 3.14(1)) induces a total ordering on the monomials $\{x^\alpha\}_{\alpha\in\mathbb{N}^n}$ of the polynomial algebra $P_n$ by the following rule: $x^\alpha >x^\beta $ if and only if $X_{\alpha,n+1}>X_{\beta,n+1}$, that is, if and only if $\alpha_n=\beta_n$, $\alpha_{n-1}=\beta_{n-1}$, $\dots$, $\alpha_{m+1}=\beta_{m+1}$ and $\alpha_m > \beta_m$ for some $m\in\{1,\dots,n\}$. This is the so-called reverse lexicographic ordering on $\{ x^\alpha\}_{\alpha\in\mathbb{N}^n}$ or on $\mathbb{N}^n$ ($\alpha>\beta$ if and only if $x^\alpha>x^\beta$). By Lemma 3.4(3),(4) and Lemma 3.14(1), if $x^\alpha>x^\beta$ (where $\alpha,\beta\in\mathbb{N}^n$), then

  • (i)  
    $\partial_k*x^\alpha >\partial_k*x^\beta$ for all $k=1,\dots,n$ such that $\alpha_k\neq 0$,
  • (ii)  
    $X_{\gamma,k} *x^\alpha >X_{\gamma,k} *x^\beta$ for all $k=1,\dots,n-1$ and $\gamma\in\mathbb{N}^{k-1}$ such that $X_{\gamma,k} *x^\alpha\neq 0$, that is, $\alpha_k\neq 0$.

Let $\mathcal{M}(P_n)$ be the set of all non-zero submodules of the $\mathfrak{u}_n$-module $P_n$. This set is totally ordered with respect to $\subseteq$. By Lemma 3.14(3), the map

Equation (3.10)

is an isomorphism of totally ordered sets. Every non-zero polynomial $p\in P_n$ can be written uniquely as a sum

The elements $\lambda_\alpha x^\alpha$ and $\lambda_\alpha$ are called the leading term and leading coefficient of $p$ respectively (with respect to the total ordering $>$). The ordinal number

(where ${[}1,\alpha]\subseteq[1,\omega^n]$) is called the ordinal degree of $p$. The following assertions hold for all non-zero polynomials $p,q\in P_n$ and all $\lambda\in K^*$:

  • (i)  
    $\operatorname{ord}(p+q)\leqslant\max\{\operatorname{ord}(p), \operatorname{ord}(q)\}$ provided that $p+q\neq 0$,
  • (ii)  
    $\operatorname{ord}(\lambda p)=\operatorname{ord}(p)$,
  • (iii)  
    $\operatorname{ord}(u*p)<\operatorname{ord}(p)$ for all $u\in\mathfrak{u}_n$ such that $u*p\neq 0$.

For every ordinal $\lambda\in[1,\omega^n)$ there is a unique representation

where $\alpha_i\in\mathbb{N}$ and not all $\alpha_j$ are equal to zero (notice the shift by 1 in the subscripts of the coefficients $\alpha_i$). Then

Equation (3.11)

where $x^\alpha=\prod_{i=1}^n x_i^{\alpha_i}$. Notice that $\alpha_1\neq 0$ if and only if $\lambda $ is not a limit ordinal. The vector space $P_{\lambda,n}$ possesses a largest monomial if and only if $\lambda$ is not a limit ordinal, and in this case $x_1^{-1}x^\alpha$ is the largest monomial in $P_{\lambda,n}$. The ordinal number $\lambda\in [1,\omega^n)$ can be written uniquely as a sum

where $\alpha_i\in\mathbb{N}$, $1\leqslant m\leqslant n$ and $1\leqslant j\leqslant m$. The positive integers $\alpha_m$ and $\alpha_j$ are called the multiplicity and the comultiplicity of the ordinal number $\lambda$. The non-negative integers $m-1$ and $j-1$ are called the degree and the codegree of $\lambda$.

The following lemma is obvious.

Lemma 3.17. Suppose that $\lambda\in[1,\omega^n)$, that is, $\lambda=\alpha_m\omega^{m-1}+ \alpha_{m-1}\omega^{m-2}+\dotsb\dotsb+ \alpha_j\omega^{j-1}$ with $\alpha_i\in\mathbb{N}$, $\alpha_m\neq 0$, $\alpha_j\neq 0$, $m\leqslant n$ and $j\geqslant 1$. Then

In particular, ${P_{\alpha_m\omega^{m-1},n}=\sum_{i=0}^{\alpha_m-1}x_m^iP_{m-1}}$.

Remark 3.18. If $\alpha_k= 0$, then the corresponding summand is absent.

Let $V$ be a vector space over $K$. A linear map $\varphi\colon V\to V$ is called a Fredholm map/operator if it has finite-dimensional kernel and cokernel. Then the number

is called the index of $\varphi$. Let $\mathcal{F}(V)$ be the set of all Fredholm linear maps on $V$. It is actually a monoid since

Equation (3.12)

Let $V$ be a vector space with a countable $K$-basis $\{e_i\}_{i\in\mathbb{N}}$ and let $\partial $ be a $K$-linear map on $V$ given by the rule $\partial e_i=e_{i-1}$ for all $i\in\mathbb{N}$, where $e_{-1}:= 0$. For example, $V=K[x]$, $e_i:=\frac{x^i}{i!}$ and $\partial=\frac{d}{dx}$. The subalgebra of $\operatorname{End}_K(V)$ generated by the map $\partial$ is the polynomial algebra $K[\partial]$. Since $\partial$ is locally nilpotent, the algebra $\operatorname{End}_K(V)$ contains the algebra $K[[\partial]]=\{\sum_{i=0}^\infty\lambda_i\partial^i\mid\lambda_i\in K\}$ of formal power series in $\partial $. The set $K[[\partial]]^*= \{\sum_{i=0}^\infty\lambda_i\partial^i\in K[[\partial]]\mid\lambda_0\in K^*\}$ is the group of units of the algebra $K[[\partial]]$. The vector space $V$ is a $K[\partial]$-module.

The following lemma is trivial.

Lemma 3.19. 

  • 1)  
    We have
  • 2)  
    In particular, the non-zero elements of $\operatorname{End}_{K[\partial]}(V)$ are surjective Fredholm maps.
  • 3)  
    If $a=\sum_{i\geqslant d}\lambda_i\partial^i\in K[[\partial]]$ and $\lambda_d\neq 0$, then $d=\operatorname{ind}(a)= \operatorname{dim}_K(\operatorname{ker}(a))$.

Let $V$ be a vector space with $K$-basis $\{e_i\}_{i\in\mathbb{N}}$. We define a $K$-linear map $x$ on $V$ by putting $xe_i:=e_{i+1}$ for all $i\in\mathbb{N}$. For example, $V=K[x]$, $e_i:= x^i$ and $x\colon K[x]\to K[x]$, $p\mapsto xp$. The polynomial algebra $K[x]$ is a subalgebra of the algebra $\operatorname{End}_K(V)$.

The following lemma is obvious.

Lemma 3.20. 

  • 1)  
    We have $\operatorname{End}_{K[x]}(V)=\{p\colon V\to V, \,a\mapsto pa\mid p\in K[x]\}\simeq K[x]$ (via $p\mapsto p$). In particular, the non-zero elements of $\operatorname{End}_{K[x]}(V)$ are injective maps.
  • 2)  
    We have $\operatorname{Aut}_{K[x]}(V)=K^*$.
  • 3)  
    For all $p\in\operatorname{End}_{K[x]}(V)=K[x]$ we have $\deg (p)=-\operatorname{ind}(p)= \operatorname{dim}_K(\operatorname{coker}(p))$.

The following proposition describes the algebra of all $\mathfrak{u}_n$-homomorphisms (and its group of units) of the $\mathfrak{u}_n$-module $P_n$. Clearly, $K[x_n]\subseteq P_n$ is an embedding of $K[\partial_n]$-modules and $K[[\frac{d}{dx_n}]]=\operatorname{End}_{K[\partial_n]}(K[x_n])$ (Lemma 3.19(1)). The $K$-derivation $\frac{d}{dx_n}$ of the polynomial algebra $K[x_n]$ is also denoted by $\partial_n$.

Proposition 3.21. 

  • 1)  
    The map $\operatorname{End}_{\mathfrak{u}_n}(P_n)\to \operatorname{End}_{K[\partial_n]}(K[x_n])=K[[\frac{d}{d x_n}]]$, $\varphi\mapsto\varphi|_{K[x_n]}$, is an isomorphism of $K$-algebras with inverse map $\varphi'\mapsto\varphi$, where $\varphi(x^\beta x_n^i):=X_{\beta,n}\varphi'(\frac{x_n^{i+1}}{i+1})$ for all $\beta\in\mathbb{N}^{n-1}$ and $i\in\mathbb{N}$.
  • 2)  
    The map $\operatorname{Aut}_{\mathfrak{u}_n}(P_n)\to \operatorname{Aut}_{K[\partial_n]}(K[x_n])=K[[\frac{d}{dx_n}]]^*$, $\varphi\mapsto\varphi|_{K[x_n]}$, is an isomorphism of groups with the same inverse map as in part 1.
  • 3)  
    Every non-zero element $\varphi\in \operatorname{End}_{\mathfrak{u}_n}(P_n)$ is a surjective map with kernel $\operatorname{ker}(\varphi)= \bigoplus_{i=0}^{d-1}P_{n-1}x_n^i$, where $d=\operatorname{ind}(\varphi|_{K[x_n]})= \operatorname{dim}_K(\operatorname{ker}(\varphi|_{K[x_n]}))$.
  • 4)  
    For all integers $d\geqslant 1$ we have $(\frac{\partial} {\partial x_n})^d\in\operatorname{End}_{\mathfrak{u}_n}(P_n)$ and $\operatorname{ker}_{P_n}(\frac{\partial}{\partial x_n})^d= \bigoplus_{i=0}^{d-1}P_{n-1}x_n^i$. In particular, $(\frac{\partial}{\partial x_n})^d\colon P_n/\bigoplus_{i=0}^{d-1} P_{n-1}x_n^i\to P_n$ is an isomorphism of $\mathfrak{u}_n$-modules.

Proof. We first prove part 1. By Lemma 3.19, $\operatorname{End}_{K[\partial_n]}(K[x_n])=K[[\frac{d}{dx_n}]]$. Since $K[x_n]=\bigcap_{i=1}^{n-1}\operatorname{ker}_{P_n}(\partial_i)$ and the maps $\partial_1,\dots,\partial_n$ commute, the restriction map $\varphi\mapsto \varphi|_{K[x_n]}$ is a well-defined homomorphism of $K$-algebras. For all $\beta\in\mathbb{N}^{n-1}$ and $i\in\mathbb{N}$,

Therefore the restriction map is a monomorphism. To prove that it is surjective, we must show that for any map $\varphi'\in\operatorname{End}_{K[\partial_n]}(K[x_n])$, its extension $\varphi$ defined as in the statement of part 1 is a $\mathfrak{u}_n$-homomorphism. The map $\varphi$ is $K$-linear. Hence we must only check that $\varphi X_{\alpha,i }=X_{\alpha,i}\varphi$ for all $i=1,\dots,n$ and $\alpha\in \mathbb{N}^{i-1}$.

Case 1: $i<n$. For all $\beta\in\mathbb{N}^{n-1}$ and $j\in\mathbb{N}$ we have

Case 2: $i=n$. Let $\beta\in\mathbb{N}^{n-1}$ and $j\in\mathbb{N}$. Suppose that $j\geqslant 1$. Then

Suppose that $j= 0$. Then

(since $\varphi'(1)\in K$).

Part 2 follows from part 1. Part 3 follows from part 1 and Lemma 3.19. Part 4 follows from part 1. □

3.9. Monomial subspaces of the polynomial algebra $P_n$.

Let $S$ be a subset of $\mathbb{N}^n$. The vector space $P_n(S):=\bigoplus_{\alpha\in S}Kx^\alpha$ is called a monomial subspace of the polynomial algebra $P_n$ with support $S$. We put $P_n(\varnothing):= 0$ by definition. Clearly, $\bigcap_{i\in I}P_n(S_i)=P_n(\bigcap_{i\in I}S_i)$.

Example 3.22. For every integer $i=1,\dots,n$ and any ordinal number $\lambda\in[1,\omega^n)$, the vector space $\{p\in P_n\mid\frac{\partial p}{\partial x_i}\in P_{\lambda,n}\}$ is a monomial subspace of $P_n$. Hence so is their intersection

Clearly, $P_{\lambda,n}\subseteq P_{\lambda,n}'$. The following theorem describes the vector space $P_{\lambda,n}'$ and shows that the inclusion is always strict and $\operatorname{dim}_K(P_{\lambda,n}'/P_{\lambda,n})<\infty$.

Theorem 3.23. Suppose that $\lambda\in[1,\omega^n)$, that is, $\lambda=\alpha_n\omega^{n-1}+\alpha_{n-1}\omega^{n-2} +\dotsb\dots+ \alpha_i\omega^{i-1}+\dotsb+\alpha_j\omega^{j-1}$, where $\alpha_i\in\mathbb{N}$, $\alpha_j\neq 0$ and $j\geqslant 1$. Then the following assertions hold.

  • 1)  
    $P_{\lambda,n}'=P_{\lambda,n}\oplus\bigoplus_{i=j}^nK\theta_i $, where $\theta_i=\begin{cases}x^\alpha=\prod_{k=1}^nx_k^{\alpha_k}&\textit{if}\quad i=j,\\ x_i\prod_{k=i}^nx_k^{\alpha_k}&\textit{if}\quad j<i\leqslant n.\end{cases}$
  • 2)  
    $1\leqslant\operatorname{dim}_K(P_{\lambda,n}'/P_{\lambda,n})=n-j+1\leqslant n$.
  • 3)  
    $\lambda$ is a non-limit ordinal (that is, $\alpha_1\,{\neq}\,0$) if and only if $\operatorname{dim}_K(P_{\lambda,n}'/P_{\lambda,n})\,{=}\,n$.
  • 4)  
    We have $1\leqslant\operatorname{dim}_K(P_{\lambda +1,n}'/P_{\lambda,n}')= j\leqslant n-1$, and the set $\{x_ix^\alpha+P_{\lambda,n}'\mid i=1,\dots,j\}$ is a basis of the vector space $P_{\lambda +1,n}'/P_{\lambda,n}'$.
  • 5)  
    The vector spaces $\{P_{\lambda,n}'\mid\lambda\in[1,\omega^n)\}$ are distinct. In particular, if $\lambda<\mu$, then $P_{\lambda,n}'\varsubsetneq P_{\mu,n}'$.

Proof. We first prove part 1. Let $R$ be the right-hand side of the equality stated in part 1. Then $P_{\lambda,n}'\supseteq R$ (by Lemma 3.17). In particular, the set $P_{\lambda,n}'\setminus P_{\lambda,n}$ is non-empty.

Case 1: $\lambda$ is a non-limit ordinal, that is, $j= 1$ $(\alpha_1\neq 0)$. In this case,

(see (3.11)), where $x^{\alpha'}:= x_1^{-1}x^\alpha$, that is, $\alpha_1'=\alpha_1-1$ and $\alpha_i'=\alpha_i$ for $i=2,\dots,n$. Suppose that $x^\beta\in P_{\lambda,n}'\setminus P_{\lambda,n}$. Then $x^\beta > x^{\alpha'}$ and there is an integer $i$, $1\leqslant i\leqslant n$, such that $\beta_j=\alpha_j$ for all $j>i$ and $\beta_i=\alpha_i'+1$. Then necessarily $\beta_k= 0$ for all $k< i$ (since $\frac{\partial x^\beta}{\partial x_k}\leqslant x^{\alpha'}$ for all $k<i$) and, therefore, $x^\beta=\theta_i$.

Case 2: $\lambda$ is a limit ordinal, that is, $j> 1$ $(\alpha_1= 0)$. In this case,

(see (3.11)). Suppose that $x^\beta\not\in P_{\lambda,n}$. Then $x^\beta\in P_{\lambda,n}'$ if and only if $\frac{\partial x^\beta}{\partial x_i} < x^\alpha$ for all $i=1,\dots,n$ (we recall that $P_{\lambda,n}=\oplus\{Kx^\gamma\mid x^\gamma<x^\alpha\}$) or, in other words, if and only if the following conditions hold:

  • (i)  
    $\frac{\partial x^\beta}{\partial x_i}\leqslant x^\alpha$ for all $i=1,\dots,n$,
  • (ii)  
    $\frac{\partial x^\beta}{\partial x_i}\not\in K^* x^\alpha$ for all $i=1,\dots,n$.

Condition (i) is Case 1 for $\lambda +1$. Therefore,

where

Notice that $\theta_j=x^\alpha$ and $\theta_i'=\theta_i\in P_{\lambda,n}'$ for $j < i\leqslant n$. Condition (ii) excludes precisely the elements $\{\theta_i'\mid i\leqslant j\}$, that is, $\bigoplus_{i=1}^n(P_{\lambda,n}'\cap K\theta_i')= \bigoplus_{j<i\leqslant n}K\theta_i'=\bigoplus_{j<i\leqslant n}K\theta_i$. The proof of part 1 is complete.

Part 2 follows from part 1. Part 3 follows from part 2.

We now prove part 4. The ordinal number $\lambda +1$ is not a limit ordinal. By part 3 we have $\operatorname{dim}_K(P_{\lambda+1,n}'/P_{\lambda +1,n})=n$. Notice that $P_{\lambda,n}\subset P_{\lambda+1,n}\subset P_{\lambda+1,n}'$ and $\operatorname{dim}_K(P_{\lambda+1,n}/P_{\lambda,n})= 1$. By part 1,

Finally, by part 2,

and $1\leqslant j=\operatorname{dim}_K(P_{\lambda+1,n}'/P_{\lambda,n}')\leqslant n-1$. The elements $\theta_i':=x_ix^\alpha$, $i=1,\dots,j$, are the elements $\theta_i$, $i=1,\dots,j$, in part 1, but for the ordinal $\lambda+1$ instead of $\lambda$. Clearly, $\theta_i'\in P_{\lambda+1, n}' \setminus P_{\lambda,n}'$ for $i=1,\dots,j$. Therefore the elements $\{\theta_i'+P_{\lambda,n}\mid i=1,\dots,j\}$ are $K$-linearly independent in the vector space $P_{\lambda+1,n}'/P_{\lambda,n}'$ since the vector spaces $P_{\lambda+1,n}'$ and $P_{\lambda,n}'$ are monomial. These elements form a basis of the vector space $P_{\lambda+1,n}'/P_{\lambda,n}'$ since $j=\operatorname{dim}_K(P_{\lambda+1,n}'/P_{\lambda,n}')$.

Part 5 follows from part 4. □

For example, for all positive integers $i$ and $j$ we have

Example 3.24. Consider the following conditions:

Equation (3.13)

For every integer $i=1,\dots,n$, any element $\alpha\in\mathbb{N}^{i-1}$ and any ordinal number $\lambda\in [1,\omega^n)$, the vector space $\{p\in P_n\mid x^\alpha\frac{\partial p}{\partial x_i}\in P_{\lambda,n}\}$ is a monomial subspace of $P_n$. Hence so is their intersection

For all ordinal numbers $\lambda\in [1,\omega^n)$,

The first inclusion follows since $P_{\lambda,n}$ is a $\mathfrak{u}_n$-module. If $\lambda\,{\leqslant}\,\mu$, then $P_{\lambda,n}'' {\subseteq} P_{\mu,n}''$. The following corollary shows that these inclusions are strict and gives a $K$-basis for every vector space $P_{\lambda,n}''$.

Corollary 3.25. Suppose that $\lambda\in [1,\omega^n)$, that is, $\lambda=\alpha_m\omega^{m-1} +\alpha_{m-1}\omega^{m-2}+\dotsb\dotsb+ \alpha_j\omega^{j-1}$, where $\alpha_i\in\mathbb{N}$, $\alpha_m\neq 0$, $\alpha_j\neq 0$, $m\leqslant n$ and $j\geqslant 1$. Then the following assertions hold.

  • 1)  
    $P_{\lambda,n}''=P_{\lambda,n}\oplus Kx^\alpha=P_{\lambda+1,n}$, where $x^\alpha=\prod_{k=j}^mx_k^{\alpha_k}$.
  • 2)  
    The vector spaces $\{ P_{\lambda,n}''\mid\lambda\in[1,\omega^n)\}$ are distinct. In particular, if $\lambda<\mu$, then $P_{\lambda,n}''\varsubsetneq P_{\mu,n}''$.
  • 3)  
    $\operatorname{dim}_K(P_{\lambda,n}''/P_{\lambda,n})= 1$.
  • 4)  
    $\operatorname{dim}_K(P_{\lambda+1,n}''/P_{\lambda,n}'')= 1$.

Proof. Part 1 follows at once from Lemma 3.17, the inclusions $P_{\lambda,n}\subseteq P_{\lambda,n}''\subseteq P_{\lambda,n}'$ and Theorem 3.23.

Parts 2–4 follow from part 1. □

§ 4. The Lie algebras $\mathfrak{u}_n$ are locally finite-dimensional and locally nilpotent

The aim of this section is to prove Theorem 4.2. The key ideas are to use the fact that the algebra $\mathfrak{u}_n$ is uniserial, induction on the ordinals $\lambda\in[1,\operatorname{ord}(\Omega_n)]$ and Theorem 4.1, which gives sufficient conditions for a Lie algebra to be nilpotent.

Theorem 4.1. Let $J$ be an ideal of a Lie algebra $\mathcal{G}$ such that the Lie factor algebra $\overline{\mathcal{G}}:=\mathcal{G}/J$ is a finite-dimensional nilpotent Lie algebra, $J$ is a nilpotent Lie algebra, and every element $a\in\mathcal{G}$ acts nilpotently on $J$ (that is, $(\operatorname{ad}a)^n(J)= 0$ for some positive integer $n=n(a)$). Then $\mathcal{G}$ is a nilpotent Lie algebra.

Proof. We use induction on $d=\operatorname{dim}_K(\overline{\mathcal{G}})$. The case $d=0$ (that is, $\mathcal{G}=J$) is obvious.

Suppose that $d= 1$. This is the most important case since we will reduce the general case to this one. Then $\mathcal{G}=Ka\oplus J$, where $a\in\mathcal{G}\setminus J$. Put $\delta:=\operatorname{ad}(a)\in\operatorname{Inn}(\mathcal{G})$ and $\partial :=\operatorname{ad}(J):=\{\operatorname{ad}(j)\colon\mathcal{G}\to \mathcal{G}\mid j\in J\}$. The Lie algebra $J$ is nilpotent, that is,

for some integer $n\geqslant 1$. For all elements $b\in J$ we have

whence $\partial\delta\subseteq\delta\partial+\partial$. By hypothesis, the map $\delta$ acts nilpotently on $J$. Thus, increasing $n$ if necessary, we can assume that $\delta^n (J)= 0$. To prove that $\mathcal{G}$ is a nilpotent Lie algebra, we must show that $(K\delta+\partial)^{m+1}(\mathcal{G})= 0$ for some non-negative integer $m$ (the equality $\mathcal{G}=Ka\oplus J$ implies that $\operatorname{ad}(\mathcal{G})=K\delta+\partial$). It suffices to show that

for some non-negative integer $m$ since $(K\delta+\partial)(\mathcal{G})\subseteq[\mathcal{G},\mathcal{G}]=[Ka+J,Ka+ J]\subseteq J$. We claim that it suffices to take $m=n(n+1)$. Indeed, take $m=n(n+1)$. Then

where $W_s$ is the set of finite linear combinations of the elements $w_s$, where $w_s$ is a word of length $m$ in the alphabet $\{\delta,\partial\}$ containing precisely $s$ elements $\partial$ and $m- s$ elements $\delta$. It suffices to show that $w_s(J)= 0$ for all $s=0,1,\dots,m$.

Notice that $\delta(J)\subseteq J$ and $\delta\partial^i(J)\subseteq\partial^i(J)$ for all $i\geqslant 1$. For any $s$ with $ n\leqslant s\leqslant m$ we have $w_s(J)\subseteq\partial^s(J)\,{=}\,0$ since $s\geqslant n$ and $\partial^n(J)\,{=}\,0$. For any $s$ with $0\leqslant s<n$, the word $w_s$ is of the form $\delta^{n_1}\partial\delta^{n_2}\partial\dotsb\delta^{n_s}\partial\delta^{n_{s+1}}$, where $n_1+\dotsb+n_{s+1}= m-s$. At least one of the numbers $n_i$ is greater than or equal to $n$ since otherwise we have

a contradiction. Then $\delta^{n_i}\partial\dotsb\delta^{n_s}\partial \delta^{n_{s+1}}(J)\subseteq\delta^{n_i}(J)= 0$ since $ n_i\geqslant n$ and $\delta^n(J)= 0$. Therefore $w_s (J)= 0$.

Suppose that $d> 1$. The Lie algebra $\overline{\mathcal{G}}$ is nilpotent. We fix an ideal $\overline{\mathcal{J}}$ of $\overline{\mathcal{G}}$ such that $\operatorname{dim}_K(\overline{\mathcal{G}}/\overline{\mathcal{J}})= 1$. Let $\pi\colon\mathcal{G}\to\overline{\mathcal{G}}$ be the canonical Lie algebra epimorphism $g\mapsto\overline{g}:=g+J$. The ideal $\mathcal{J}:= \pi^{-1}(\overline{\mathcal{J}})$ of $\mathcal{G}$ has codimension 1 $(\operatorname{dim}_K(\mathcal{G}/\mathcal{J})= \operatorname{dim}_K(\overline{\mathcal{G}}/\overline{\mathcal{J}})= 1)$ and $J$ is an ideal of the Lie algebra $\mathcal{J}$. The pair $(\mathcal{J},J)$ satisfies the hypotheses of the theorem and $\operatorname{dim}_K(\mathcal{J}/J)= \operatorname{dim}_K(\overline{\mathcal{G}})-1<d$. By induction on $d$, the Lie algebra $\mathcal{J}$ is nilpotent. Now the pair $(\mathcal{G},\mathcal{J})$ is as in the case $d= 1$ considered above. Indeed, $\operatorname{dim}_K(\mathcal{G}/\mathcal{J})= 1$, and the element $a\in\mathcal{G}\setminus\mathcal{J}$ in the decomposition $\mathcal{G}= Ka\oplus\mathcal{J}$ acts nilpotently on $\mathcal{J}$ since $\overline{\mathcal{G}}$ is a finite-dimensional nilpotent Lie algebra. In more detail, put $\delta:=\operatorname{ad}(a)$. Then $\delta^s(\mathcal{G})\subseteq J$ for some $s$ since $\overline{\mathcal{G}}$ is a finite-dimensional nilpotent Lie algebra; $\delta^t (J)= 0$ for some $t$ by hypothesis. Hence $\delta^{s+t}(\mathcal{J})\subseteq \delta^{s+t}(\mathcal{G})\subseteq\delta^t(J)= 0$. Therefore $\mathcal{G}$ is a nilpotent Lie algebra. □

Theorem 4.2. The Lie algebras $\mathfrak{u}_n$ are locally finite-dimensional and locally nilpotent.

Proof. We must show that the Lie subalgebra $\mathcal{G}$ of $\mathfrak{u}_n$ generated by a finite set of elements, say, $a_1,\dots,a_\nu$, is a finite-dimensional nilpotent Lie algebra. The case $\nu= 1$ is trivial. We may assume without loss of generality that $\nu\geqslant 2$, the elements $a_i$ are $K$-linearly independent and

We use induction on the ordinal number $\lambda:=\operatorname{ord}(a_\nu) \in[1,\operatorname{ord}(\Omega_n)]$. By definition, $\lambda$ is a non-limit ordinal. The initial case $\lambda= 1$ is obvious since $\mathcal{G}=I_1=K\partial_n$. Thus, let $\lambda$ be a non-limit ordinal such that $\lambda > 1$ and assume that the result holds for all non-limit ordinals $\lambda'$ such that $\lambda'<\lambda$. Then $\lambda'':=\operatorname{ord}(a_{\nu -1})$ is a non-limit ordinal and $\operatorname{ord}(a_i)\leqslant\lambda''<\lambda$ for all $i=1,\dots,\nu -1$. Let $V$ be the Lie subalgebra of $\mathfrak{u}_n$ generated by the elements $a_1,\dots,a_{\nu -1}$. By induction, $V$ is finite-dimensional and nilpotent. The inner derivation $\delta :=\operatorname{ad}(a_\nu)$ of $\mathfrak{u}_n$ is locally nilpotent (Proposition 2.1(5)) and $\operatorname{dim}_K(V)< \infty$. It follows that $\delta^{s+1}(V)= 0$ for some non-negative integer $s$. The vector space $U:=V+\delta (V)+\dotsb+\delta^s(V)$ is a finite-dimensional $\delta$-invariant subspace of the ideal $I_{\lambda''}$ (see (3.3)). Let $J$ be the Lie subalgebra of $\mathfrak{u}_n$ generated by $U$. By induction, $J$ is finite-dimensional and nilpotent. Then $\delta^t(J)= 0$ for some integer $t\geqslant 1$. Clearly, $\delta (J)\subseteq J$ since $\delta (U)\subseteq U$ and $\delta$ is a derivation. We see that $\mathcal{G}=Ka_\nu+J=Ka_\nu\oplus J$ (clearly, $\mathcal{G}\supseteq Ka_\nu +J$; on the other hand, $Ka_\nu + J$ is a Lie subalgebra of $\mathfrak{u}_n$ that contains the elements $a_1,\dots,a_\nu $, and so $\mathcal{G}\subseteq Ka_\nu + J$). We claim that the hypotheses of Theorem 4.1 hold for the pair $(\mathcal{G},J)$, whence $\mathcal{G}$ is a nilpotent Lie algebra (by Theorem 4.1) and $\operatorname{dim}_K(\mathcal{G})=1+\operatorname{dim}_K(J)< \infty$. Indeed, to prove the claim, it suffices to show that $(\delta+\partial)^m(J)= 0$ for all $m\gg 0$, where $\partial= \operatorname{ad}(J)$. We fix a number $n$ such that $\delta^n(J)= 0$ and $\partial^n(J)= 0$. For the same reason as in the proof of Theorem 4.1, it suffices to take $m=n(n+ 1)$:

§ 5. The isomorphism problem for factor algebras of the Lie algebras $\mathfrak{u}_n$

We already know that the Lie algebras $\mathfrak{u}_n$ and $\mathfrak{u}_m$ are not isomorphic for $n\neq m$. The aim of this section is to answer the following question.

Question. Let $I$ and $J$ be ideals of the Lie algebras $\mathfrak{u}_n$ and $\mathfrak{u}_m$ respectively. When are the Lie algebras $\mathfrak{u}_n/I$ and $\mathfrak{u}_m/J$ isomorphic?

We first consider the case when $n=m$ (Theorem 5.1) and then deduce the general case from this one (Corollary 5.2). Put $\omega^0:= 1$.

Theorem 5.1. 

  • 1)  
    Let $I$ be an ideal of $\mathfrak{u}_n$, that is, $I{=}I_\lambda$ for some $\lambda {\in} [1,\operatorname{ord}(\Omega_n)] {\cup} \{0\}$ (Theorem 3.6), where $I_0\,{:=}\,\{ 0\}$. Then the Lie algebras $\mathfrak{u}_n$ and $\mathfrak{u}_n/I_\lambda$ are isomorphic if and only if $\lambda=i\omega^{n-2}$, where $i\in\mathbb{N} $.
  • 2)  
    Let $I$ and $J$ be ideals of $\mathfrak{u}_n$, that is, $I=I_\lambda$ and $J=I_\mu$ for some $\lambda,\mu\in [1,\operatorname{ord}(\Omega_n)]\cup\{0\}$ (Theorem 3.6(1)). Then the Lie algebras $\mathfrak{u}_n/I_\lambda$ and $\mathfrak{u}/I_\mu$ are isomorphic if and only if one of the following conditions holds:
    • (a)  
      $\lambda=i\omega^{n-2}+\nu$ and $\mu=j\omega^{n-2}+\nu$, where $i,j\in\mathbb{N}$ and $\nu\in [1,\omega^{n-2})\cup\{ 0\}$;
    • (b)  
      $\lambda=\omega^{n-1}+\omega^{n-2}+\dotsb +\omega^s+i\omega^{s-2}+\nu $ and $\mu=\omega^{n-1}+\omega^{n-2}+\dotsb +\omega^s+j\omega^{s-2}+\nu $, where $2\leqslant s\leqslant n-1$, with $i,j\in\mathbb{N}$ and $\nu\in [1,\omega^{s-2})\cup\{0\}$;
    • (c)  
      $\lambda=\mu=\omega^{n-1}+\omega^{n-2}+\dotsb +\omega+\varepsilon $, where $\varepsilon=0,1$.

For all integers $n$ and $m$ with $2\leqslant n<m$ there is a natural isomorphism of Lie algebras

Equation (5.1)

where $X_{\alpha,i}\in\mathcal{B}_n$, $\nu_{mn}:=\omega^{m-1}+\omega^{m-2}+\dotsb+\omega^n$.

In more detail, $I_{\nu_{mn}}=\mathfrak{u}_{m,n+1}= \bigoplus_{j=n+1}^mP_{j-1}\partial_j$.

Corollary 5.2. Let $n$ and $m$ be integers with $2\leqslant n <m$, and let $I, J$ be ideals of $\mathfrak{u}_n$, $\mathfrak{u}_m$ respectively. Then $\mathfrak{u}_n/I$ and $\mathfrak{u}_m/J$ are isomorphic if and only if

We recall that $\mathfrak{u}_n=\mathfrak{u}_{n-1}\oplus P_{n-1}\partial_n$, where $\mathfrak{u}_{n-1}$ is a Lie subalgebra of $\mathfrak{u}_n$ and $P_{n-1}\partial_n$ is an Abelian ideal of $\mathfrak{u}_n$. We introduced the $K$-basis $\mathcal{B}_n=\{ X_{\alpha,i}\}$ for $\mathfrak{u}_n$ in (2.1). Define a $K$-linear map $f_n\colon\mathfrak{u}_n\to\mathfrak{u}_n$ in one of the two equivalent ways:

Equation (5.2)

Lemma 5.3. For every integer $i\geqslant 1$, the $K$-linear map $f_n^i\colon \mathfrak{u}_n\to\mathfrak{u}_n$ is an epimorphism of Lie algebras and $\operatorname{ker}(f_n^i)= \sum_{j=0}^{i-1}P_{n-2}x_{n-1}^j\partial_n=I_{i\omega^{n-2}}$.

Proof. By the definition of $f_n$ we have

To finish the proof of the lemma, it suffices to show that the $K$-linear map $f_n$ is a homomorphism of Lie algebras, that is,

Equation (5.3)

for all elements $u$ and $v$ of the basis $\mathcal{B}_n$ of $\mathfrak{u}_n$ (see (2.1)). Since $\mathfrak{u}_n=\mathfrak{u}_{n-1}\oplus P_{n-1}\partial_n$, we have $\mathcal{B}_n=\mathcal{ B}_{n-1}\cup \{ X_{\alpha,n}\}_{\alpha\in\mathbb{N}^{n-1}}$. The equality (5.3) is obvious if either $u,v\in\mathcal{B}_{n-1}$ (since $f_n(a)=a$ for all $a\in\mathfrak{u}_{n-1}$), or $u,v\in \{ X_{\alpha,n}\}_{\alpha\in\mathbb{N}^{n-1}}$ (since $f_n(P_{n-1}\partial_n)\subseteq P_{n-1}\partial_n$ and ${[}P_{n-1}\partial_n,P_{n-1}\partial_n]= 0$). In the remaining cases when $u\in\mathcal{B}_{n-1}$ and $v=X_{\alpha,n}$ for some $\alpha\in\mathbb{N}^{n-1}$, the equality follows from the following three facts: $\partial_{n-1}$ is a central element of $\mathfrak{u}_{n-1}$ (Proposition 2.1(6)), $f_n(P_{n-1}\partial_n) \subseteq P_{n-1}\partial_n$ and $f_n|_{P_{n-1}\partial_n }= \operatorname{ad}(\partial_{n-1})$. Indeed, applying the inner derivation $\delta:=\operatorname{ad}(\partial_{n-1})$ of $\mathfrak{u}_n$ to the equality ${[}u,X_{\alpha,n}]= w$ (where $w\in P_{n-1}\partial_n$), we obtain the required equality

Corollary 5.4. If $n\geqslant 3$, then the ideal $P_{n-1}\partial_n$ is the least ideal $I$ of $\mathfrak{u}_n$ such that $\mathfrak{u}_n/I$ is isomorphic to $\mathfrak{u}_{n-1}$.

Proof. Clearly, $\mathfrak{u}_n/P_{n-1}\partial_n\simeq\mathfrak{u}_{n-1}$ and $P_{n-1}\partial_n=I_{\omega^{n-1}}$. We suppose that $I\neq P_{n-1}\partial_n$ and seek a contradiction. Then $I=I_\lambda$ for some $\lambda<\omega^{n-1}$. Fix an integer $i\geqslant 1$ such that $\lambda <i\omega^{n-2}$. Then $I_\lambda\subset I_{i\omega^{n-2}}$. Since there is a natural epimorphism of Lie algebras $\mathfrak{u}_n/I\to\mathfrak{u}_n/I_{i\omega^{n-2}}$, we have $\operatorname{u.dim}(\mathfrak{u}_n/I_{i\omega^{n-2}})\leqslant \operatorname{u.dim}(\mathfrak{u}_n/I)=\operatorname{u.dim}(\mathfrak{u}_{n-1})$. By Lemma 5.3, $\mathfrak{u}_n/I_{i\omega^{n-2}}\simeq\mathfrak{u}_n$. Hence,

a contradiction. □

Proof of Theorem 5.1. Let us prove part 1. $(\Leftarrow)$ By Lemma 5.3, for every integer $i\geqslant 1$ the map $f_n^i\colon\mathfrak{u}_n\to\mathfrak{u}_n$ is a Lie algebra epimorphism with kernel $I_{i\omega^{n-2}}$. Therefore $\mathfrak{u}_n\simeq\mathfrak{u}_n/I_{i\omega^{n-2}}$. The implication $(\Rightarrow)$ follows from part 2.

To prove part 2, we use induction on $n\geqslant 2$. The initial step $n=2$ is a direct corollary of Lemma 5.3 and the classification of the ideals of $\mathfrak{u}_2$. By Theorem 3.6(1), the proper ideals of $\mathfrak{u}_2$ are $I_n$, $n\geqslant 1$, and $I_\omega=P_1\partial_2$. By Lemma 5.3, $\mathfrak{u}_2/I_n\simeq \mathfrak{u}_2$ for all $n\,{\geqslant}\,1$, and $\operatorname{dim}_K(\mathfrak{u}_2/I_\omega)\,{=} \operatorname{dim}_K(K\partial_1)\,{=}\,1 \,{<}\,\operatorname{dim}_K(\mathfrak{u}_2)\,{=}\,\infty$, and part 2 follows.

Suppose that $n>2$ and the result holds for all $n'<n$. We recall that $\mathfrak{u}_{n-1}\simeq\mathfrak{u}_n/I_{\omega^{n-1}}$, $I_{\omega^{n-1}}=P_{n-1}\partial_n$ and $\operatorname{u.dim}(\mathfrak{u}_{n-1})<\operatorname{u.dim}(\mathfrak{u}_n)$. Since $\mathfrak{u}_n$ is uniserial and Artinian, so are all its factor algebras. Thus, if $\mathfrak{u}_n/I_\lambda\simeq\mathfrak{u}_n/I_\mu$, then $\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)= \operatorname{u.dim}(\mathfrak{u}_n/I_\mu)$.

Step 1: $\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)=\operatorname{u.dim}(\mathfrak{u}_n)$ for all $\lambda < \omega^{n-1}$.

Indeed, $\lambda<\omega^{n-1}$ implies that $\lambda\leqslant i\omega^{n-2}$ for some $i\geqslant 1$. Then $\mathfrak{u}_n\simeq\mathfrak{u}_n/I_{i\omega^{n-2}}$ (Lemma 5.3) and, therefore,

whence $\operatorname{u.dim}(\mathfrak{u}_n)=\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)$.

Suppose that $\mathfrak{u}_n/I_\lambda\simeq\mathfrak{u}_n/I_\mu$ for some $\lambda $ and $\mu$.

Step 2: It suffices to consider the case when $\lambda,\mu <\omega^{n-1}$.

Indeed, if $\lambda,\mu\geqslant\omega^{n-1}$, then part 2 follows by induction on $n$ because $I_{\omega^{n-1}}\subseteq I_\lambda$, $I_{\omega^{n-1}} \subseteq I_\mu$ and $\mathfrak{u}_n/I_{\omega^{n-1}}\simeq\mathfrak{u}_{n-1}$. We can assume without loss of generality that $\lambda\leqslant \mu$. The case when $\lambda<\omega^{n-1}\leqslant\mu$ is impossible by Step 1 since

a contradiction. This finishes the proof of step 2.

Step 3: In view of Lemma 5.3 we may assume that $\lambda,\mu<\omega^{n-2}$.

The idea of the proof of the theorem is to introduce, for every Lie factor algebra $\mathfrak{u}_n/I_\lambda$ with $\lambda < \omega^{n-2}$, a quantity which is invariant under isomorphisms and takes distinct values for the ordinal numbers $\lambda <\omega^{n-2}$. This invariant is the uniserial dimension of certain ideals of $\mathfrak{u}_n/I_\lambda$. For each ordinal number $\nu\in[1,\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)]$ let $I_\nu'$ be the unique ideal of the uniserial Artinian Lie algebra $\mathfrak{u}_n/I_\lambda$ with uniserial dimension $\nu$. Every ideal of a uniserial Artinian Lie algebra is characteristic. This fact is crucial for our arguments below. Clearly,

We put $\mathfrak{a}:=I_{\omega^{n-1}+\omega^{n-2}}'$ and $\theta:=x_{n-1}\partial_n+I_\lambda\in\mathfrak{u}_n/I_\lambda$. Notice that $I_{\omega^{n-2}+1}'=K\theta\oplus I_{\omega^{n-2}}'$. For every element $v\in\mathfrak{u}_n/I_\lambda$ we put $\operatorname{Cen}_\mathfrak{a}(v):=\{a\in\mathfrak{a}\mid[a,v]=0\}$.

Step 4: For all $\sigma\in\operatorname{Aut}_K(\mathfrak{u}_n/ I_\lambda)$ and $ u\in S:=I_{\omega^{n-2}+1}'\setminus I_{\omega^{n-2}}'$, we have $\operatorname{Cen}_\mathfrak{a}(\sigma (u))= \operatorname{Cen}_\mathfrak{a}(\theta)=I_{\omega^{n-1}+\lambda}'$.

Indeed, the second equality follows from two facts: ${[}x_{n-1}\partial_n, I_{\omega^{n-1}}]=[x_{n-1}\partial_n,P_{n-1}\partial_n]= 0$ and, for all $\alpha\in\mathbb{N}^{n-2}$, ${[}x^\alpha\partial_{n-1},x_{n-1}\partial_n]=x^\alpha\partial_n$.

The first equality follows since $\sigma (u)=\xi\theta +v+I_\lambda$ for some elements $\xi\in K^*$, $v\in P_{n-2}\partial_n$ and

Alternatively, using the equality $\operatorname{Cen}_\mathfrak{a}(\theta)= I_{\omega^{n-1}+\lambda}'$ and the fact that $I_{\omega^{n-1}+\lambda}'$ and $\mathfrak{a}$ are characteristic ideals of $\mathfrak{u}_n/I_\lambda$, we see that

Since $\sigma (\mathfrak{a})=\mathfrak{a}$ and $\sigma (S)=S$ for all automorphisms $\sigma\in\operatorname{Aut}_K (\mathfrak{u}_n/I_\lambda) $, Step 4 means that the ordinal number

is an invariant (under isomorphisms) of the algebra $\mathfrak{u}_n/I_\lambda$, where $\operatorname{u.dim} (I_{\omega^{n-1}+\lambda})$ is the uniserial dimension of the $\mathfrak{u}_n$-module $I_{\omega^{n-1}+\lambda}$. If $\mathfrak{u}_n/I_\lambda\simeq\mathfrak{u}_n/I_\mu$ for some ordinals $\lambda,\mu<\omega^{n-2}$, then $\omega^{n-1}+\lambda=\omega^{n-1}+\mu$, whence $\lambda=\mu$. □

Corollary 5.5. Let $I_\lambda$ be an ideal of $\mathfrak{u}_n$, where $\lambda\in[1,\operatorname{ord}(\Omega_n)]\cup\{0\}$, and $I_0:= 0$. Then the following assertions hold.

  • 1)  
    $\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)= \operatorname{u.dim}(\mathfrak{u}_n)$ for all $\lambda\in [1,\omega^{n-1})\cup\{ 0\}$.
  • 2)  
    $\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)= \operatorname{u.dim}(\mathfrak{u}_s)$ for all $\lambda\in [\omega^{n-1}+\omega^{n-2}+\dotsb+\omega^s,\omega^{n-1}+ \omega^{n-2}+\dotsb +\omega^{s-1})$ and $s$ such that $2\leqslant s\leqslant n-1$.
  • 3)  
    $\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)=1-\varepsilon$ for all $\lambda=\omega^{n-1}+\omega^{n-2}+\dotsb+\omega+\varepsilon$, where $\varepsilon=0,1$.

Proof. Part 1 is Step 1 in the proof of Theorem 5.1. Part 2 follows from part 1 and (5.1). Part 3 is trivial. □

Proof of Corollary 5.2. Notice that the uniserial dimensions $\operatorname{u.dim}(\mathfrak{u}_i)$, $i\geqslant 2$, are distinct, and if $\mathfrak{u}_n/I_\lambda\simeq \mathfrak{u}_m/I_\mu$, then $\operatorname{u.dim}(\mathfrak{u}_n/I_\lambda)= \operatorname{u.dim}(\mathfrak{u}_m/I_\mu)$. The isomorphism of Lie algebras $\mathfrak{u}_n\simeq\mathfrak{u}_m/I_{\nu_{mn}}$ (see (5.1)) induces a bijection between the set $\mathcal{J}(\mathfrak{u}_n)$ of all non-zero ideals of the Lie algebra $\mathfrak{u}_n$ and the set $\mathcal{J}(\mathfrak{u}_m, I_{\nu_{mn}})$ of all ideals of the Lie algebra $\mathfrak{u}_m$ that properly contain the ideal $I_{\nu_{mn}}$:

Equation (5.4)

The corollary now follows from Theorem 5.1, Corollary 5.5 and (5.4). □

§ 6. The Lie algebra $\mathfrak{u}_\infty$

In this section we study the Lie algebra $\mathfrak{u}_\infty$ in detail. Many properties of $\mathfrak{u}_\infty$ are similar to those of $\mathfrak{u}_n$, $n\geqslant 2$, but there are several differences. For example, $\mathfrak{u}_\infty$ is insoluble and non-Artinian (but almost Artinian) with $\operatorname{u.dim}(\mathfrak{u}_\infty)= \omega^\omega$. We obtain a classification of the ideals of $\mathfrak{u}_\infty$ (Theorem 6.2). All of them are characteristic ideals (Corollary 6.4(2)). We give an isomorphism criterion for the Lie factor algebras of $\mathfrak{u}_\infty$ (Corollary 6.3).

Let $P_\infty\,{:=}\,\bigcup_{n\geqslant 1} P_n\,{=}\,K[x_1, x_2,\dots]$ be the polynomial algebra in countably many variables, and let $A_\infty:=\bigcup_{n\geqslant 1}A_n= K\langle x_1,x_2,\dots,\partial_1,\partial_2,\dots\rangle$ be the infinite Weyl algebra. The Lie algebra $\mathfrak{u}_\infty$ is a Lie subalgebra of the Lie algebra $(A_\infty,[\,\cdot\,{,}\,\cdot\,])$. The polynomial algebra $P_\infty$ is an $A_\infty$-module. In particular, $P_\infty$ is a $\mathfrak{u}_\infty$-module. The Lie algebra

is the direct sum of the Abelian (infinite-dimensional for $i\geqslant 2$) Lie subalgebras $P_{i-1}\partial_i$. For every integer $i\geqslant 1$,

is an ideal of $\mathfrak{u}_\infty$ by (2.3). Clearly, $\mathfrak{u}_{i,i}\subset\mathfrak{u}_{i+1,i}\subset\dotsb\subset \mathfrak{u}_{\infty,i}=\bigcup_{n\geqslant i}\mathfrak{u}_{n,i}$ for all $i\geqslant 2$. $\mathfrak{u}_\infty$ contains a strictly descending chain of ideals

Equation (6.1)

and $\mathfrak{u}_\infty/\mathfrak{u}_{\infty,n+1}\simeq\mathfrak{u}_n$ for all $n\geqslant 2$.

Proposition 6.1. 

  • 1)  
    $\mathfrak{u}_\infty$ is insoluble.
  • 2)  
    $\mathfrak{u}_\infty$ is locally nilpotent and locally finite-dimensional.
  • 3)  
    Each element $u\in\mathfrak{u}_\infty$ acts locally nilpotently on the $\mathfrak{u}_\infty$-module $P_\infty$.
  • 4)  
    The chain of non-zero ideals in (6.1) is the derived series of $\mathfrak{u}_\infty$, that is, $(\mathfrak{u}_\infty)_{(i)}=\mathfrak{u}_{\infty,i+1}$ for all $i\geqslant 0$.
  • 5)  
    The lower central series of $\mathfrak{u}_\infty$ stablilizes at the first step, that is, $(\mathfrak{u}_{\infty})^{(0)}{=}\, \mathfrak{u}_{\infty}$ and $(\mathfrak{u}_{\infty})^{(i)}= \mathfrak{u}_{\infty,2}$ for all $i\geqslant 1$.
  • 6)  
    All inner derivations of $\mathfrak{u}_\infty$ are locally nilpotent.
  • 7)  
    The centre $Z(\mathfrak{u}_\infty)$ of $\mathfrak{u}_\infty$ is 0. In particular, $\operatorname{c.dim}(\mathfrak{u}_\infty)= 0$.
  • 8)  
    The Lie algebras $\mathfrak{u}_\infty$ and $\mathfrak{u}_n$, where $n\geqslant 2$, are not isomorphic.
  • 9)  
    $\mathfrak{u}_\infty$ contains a copy of every nilpotent finite-dimensional Lie algebra.
  • 10)  
    The inner derivation $\operatorname{ad}(a)$, $a\in\mathfrak{u}_\infty$, is a nilpotent derivation of $\mathfrak{u}_\infty$ if and only if $a= 0$.

Proof. Part 1 follows from part 4. Part 2 follows from Theorem 4.2 since $\mathfrak{u}_\infty {=}\bigcup_{n\geqslant 2}\mathfrak{u}_n$. Part 3 follows from Proposition 2.1(4) since $P_\infty {=}\bigcup_{n\geqslant 1}P_n$ and $\mathfrak{u}_\infty {=} \bigcup_{n\geqslant 2}\mathfrak{u}_n$.

Parts 4, 5 follow from (2.3) and the decomposition $\mathfrak{u}_\infty=\bigoplus_{i\geqslant 1}P_{i-1}\partial_i$. Part 6 follows from Proposition 2.1(5) since $\mathfrak{u}_\infty=\bigcup_{n\geqslant 2}\mathfrak{u}_n$ and $\mathfrak{u}_2\subseteq\mathfrak{u}_3\subseteq\dotsb$.

Let us prove part 7. If $z\in Z(\mathfrak{u}_\infty)$, then $z\in\mathfrak{u}_n$ for some $n$, whence $z\in Z(\mathfrak{u}_n)= K\partial_n$ (Proposition 2.1(6)). Since ${[}\partial_n,x_n\partial_{n+1}]=\partial_{n+1}$, we must have $z= 0$.

Let us prove part 8. The Lie algebra $\mathfrak{u}_n$ is soluble (Proposition 2.1(1)), but $\mathfrak{u}_\infty$ is not (part 1). Therefore $\mathfrak{u}_\infty\not\simeq\mathfrak{u}_n$ for all $n\geqslant 2$. By Proposition 2.1(7), the $\mathfrak{u}_n$, $n\geqslant 2$, are pairwise non-isomorphic.

Let us prove part 9. Any finite-dimensional nilpotent Lie algebra is a subalgebra of the Lie algebra $\operatorname{UT}_n(K)$ for some $n\geqslant 2$. The result now follows from the inclusions $\operatorname{UT}_n(K)\subseteq\mathfrak{u}_n\subseteq\mathfrak{u}_\infty$.

We finally prove part 10. It suffices to show that if $a\neq 0$, then the derivation $\delta$ is not nilpotent. We suppose that this is not the case for some $a$ and seek a contradiction. We can write the element $a$ as a sum $p_n\partial_n+p_{n+1}\partial_{n+1}+\dotsb$, where $p_i\in P_{i-1}$ for all $i\geqslant n$ and $p_n\neq 0$. In view of the Lie algebra isomorphism $\mathfrak{u}_\infty/\mathfrak{u}_{\infty,n+2}\simeq\mathfrak{u}_{n+1}$, the inner derivation $\delta'=\operatorname{ad} (p_n\partial_n+p_{n+1}\partial_{n+1})$ induced by $\delta$ is a nilpotent derivation of $\mathfrak{u}_{n+1}$. By Lemma 2.2, $p_n= 0$, a contradiction. □

The following theorem gives a list of all ideals of $\mathfrak{u}_\infty $.

Theorem 6.2. The set $\mathcal{J}(\mathfrak{u}_\infty)$ of all non-zero ideals of $\mathfrak{u}_\infty$ is equal to the set $\{\mathfrak{u}_\infty, \mathfrak{u}_{\infty,n},I_{\lambda,n}:=I_\lambda (n)+ \mathfrak{u}_{\infty,n+1}\mid n\geqslant 2,\,\lambda\in [1,\omega^{n-1})\}$, where $I_\lambda(n)$ is the ideal $I_\lambda$ of $\mathfrak{u}_n$ as defined in (3.3), that is, $I_\lambda (n)= \bigoplus_{(\alpha,n)\leqslant\lambda}KX_{\alpha,n}$. In particular, every non-zero ideal $I$ of $\mathfrak{u}_\infty$ contains an ideal $\mathfrak{u}_{\infty,n'+1}$ for some $n'\geqslant 2$, and if $I=I_{\lambda,n}$, then $n=\min\{n'\geqslant 2\mid\mathfrak{u}_{\infty,n'+1}\subseteq I\}$.

Proof. Let $I$ be a non-zero ideal of $\mathfrak{u}_\infty$, and let $a$ be a non-zero element of $I$. Then

for some integers $l\geqslant 1$, $a_i\in P_{i-1}$ and $a_l\neq 0$. Let $d$ be the total degree of the polynomial $a_l=\sum_{\alpha\in\mathbb{N}^{l-1}}\lambda_\alpha x^\alpha\in P_{l-1}$, where $\lambda_\alpha\in K$. Fix an $\alpha=(\alpha_i)\in\mathbb{N}^{l-1}$ such that $|\alpha|:=\alpha_1+\dotsb +\alpha_{l-1}=d$. Applying $\operatorname{ad}(\partial)^\alpha:=\prod_{i=1}^{l-1} \operatorname{ad}(\partial_i)^{\alpha_i}$ to the element $a$, we get an element of $I$ of type $\alpha!\lambda_\alpha\partial_l+\dotsb $. Thus, without loss of generality, we may assume from the very beginning that $a_l= 1$, that is, $a=\partial_l+a_{l+1}\partial_{l+1} + \dotsb\dotsb+a_m\partial_m$. Then

Hence $I\supseteq[\partial_{m+1},P_{s-1}\partial_s]= [\partial_{m+1},P_{s-1}]\partial_s=P_{s-1}\partial_s$ for all $s>m+1$. This means that $\mathfrak{u}_{\infty,m+2}\subseteq I$. Consider the following epimorphism of Lie algebras (where $n=\min\{n'\geqslant 2\mid\mathfrak{u}_{\infty,n'+1}\subseteq I\}$):

Equation (6.2)

The image $\pi_n (I)$ of $I$ is an ideal of $\mathfrak{u}_n$ such that $\pi_n(I)\subset\mathfrak{u}_{n,n}= P_{n-1}\partial_n$, by the definition of the number $n$. We may assume without loss of generality that $I\neq\mathfrak{u}_\infty,\mathfrak{u}_{\infty,2},\dots$. Then $I=I_{\lambda,n}$ for some $\lambda\in[1,\omega^{n-1})$ by Theorem 3.6(1). The following two facts are obvious:

  • (i)  
    $I_{\lambda,n}\supseteq I_{\mu,m}$ if and only if either $n < m$, or $n=m$ and $\lambda\geqslant\mu$,
  • (ii)  
    $\mathfrak{u}_{\infty,n-1}\supset I_{\lambda,n}\supset \mathfrak{u}_{\infty,n}$ for all $n\geqslant 2$.

They yield that $\mathfrak{u}_\infty$ is uniserial (that is, for any distinct ideals $I$, $J$ of $\mathfrak{u}_\infty$, either $I\subset J$ or $I\supset J$). Hence the chain

Equation (6.3)

contains all the non-zero ideals of $\mathfrak{u}_\infty$. □

In combination with Theorem 5.1 and Corollary 5.2, the following corollary gives an isomorphism criterion for the Lie factor algebras of $\mathfrak{u}_\infty$.

Corollary 6.3. 

  • 1)  
    An ideal $I$ of $\mathfrak{u}_\infty$ satisfies $\mathfrak{u}_\infty/I\simeq\mathfrak{u}_\infty$ if and only if $I= 0$.
  • 2)  
    Let $I$, $J$ be non-zero ideals of $\mathfrak{u}_\infty$. We put $n=\min\{n'\geqslant 2\mid\mathfrak{u}_{\infty,n'+1} \subseteq I,\mathfrak{u}_{\infty,n'+1}\subseteq J\}$ ($n\,{<}\,\infty$ by Theorem 6.2). Then $\mathfrak{u}_\infty/I\simeq\mathfrak{u}_\infty/J$ if and only if $\mathfrak{u}_n/ I'\simeq\mathfrak{u}_n/J'$, where $I':=I/\mathfrak{u}_{\infty,n+1}$ and $J':=J/\mathfrak{u}_{\infty,n+1}\subseteq \mathfrak{u}_n=\mathfrak{u}_\infty/\mathfrak{u}_{\infty,n+1}$.

In contrast to the Lie algebras $\mathfrak{u}_n$, $n\geqslant 2$, no proper Lie factor algebra of $\mathfrak{u}_\infty$ is isomorphic to $\mathfrak{u}_\infty$.

A Lie algebra $\mathcal{G}$ is said to be almost Artinian if all its proper factor algebras are Artinian Lie algebras (that is, for every non-zero ideal $I$ of $\mathcal{G}$, the factor algebra $\mathcal{G}/I$ is Artinian).

Corollary 6.4. 

  • 1)  
    The Lie algebra $\mathfrak{u}_\infty$ is uniserial, non-Artinian, non-Noetherian, and almost Artinian. Its uniserial dimension is given by $\operatorname{u.dim}(\mathfrak{u}_\infty)=\omega^{\omega}$.
  • 2)  
    All the ideals of $\mathfrak{u}_\infty$ are characteristic.

Proof. We first prove part 1. We have just seen that $\mathfrak{u}_\infty$ is uniserial (see (6.3)). By Theorem 6.2, $\mathfrak{u}_\infty$ is neither Artinian, nor Noetherian, but almost Artinian (since all the $\mathfrak{u}_n$ are Artinian; see Theorem 3.6(2)). By Theorem 6.2, every non-zero ideal of $\mathfrak{u}_\infty$ contains the ideal $\mathfrak{u}_{\infty,n+1}$ for some $n$, and the factor algebra $\mathfrak{u}_\infty/\mathfrak{u}_{\infty,n+1}\simeq\mathfrak{u}_n$ is a uniserial Artinian Lie algebra, whence

by Theorem 3.6(2).

We now prove part 2. The ideals $\mathfrak{u}_{\infty, n}$, $n\geqslant 1$, are characteristic because they form the derived series of $\mathfrak{u}_\infty$ (Proposition 6.1(4)). By Theorem 6.2 it remains to show that every ideal $I_{\lambda,n}$ is characteristic. Let $\sigma$ be an automorphism of $\mathfrak{u}_\infty$. Since $\sigma(\mathfrak{u}_{\infty,n+1})=\mathfrak{u}_{\infty,n+1}$, $\sigma$ induces an automorphism $\sigma_n$ of $\mathfrak{u}_{\infty}/\mathfrak{u}_{\infty,n+1}\simeq \mathfrak{u}_n$. Then $\sigma_n(\pi_n(I_{\lambda,n}))=\pi_n(I_\lambda)$ by Corollary 3.11, where $\pi_n$ is as in (6.2). Therefore, $\sigma(I_{\lambda,n})=I_{\lambda,n}$, as required. □

Let $U_\infty:=U(\mathfrak{u}_\infty)$ be the universal enveloping algebra of $\mathfrak{u}_\infty$. The chain $\mathfrak{u}_2\subset\mathfrak{u}_3\subset\dotsb\subset \mathfrak{u}_\infty=\bigcup_{n\geqslant 2}\mathfrak{u}_n$ of Lie algebras gives a chain $U_2\subset U_3\subset\dotsb\subset U_\infty=\bigcup_{n\geqslant 2}U_n$ of universal enveloping algebras.

Corollary 6.5. 

  • 1)  
    All the inner derivations $\{\operatorname{ad}(u)\mid u\in\mathfrak{u}_\infty\}$ of the universal enveloping algebra $U_\infty $ of $\mathfrak{u}_\infty $ are locally nilpotent.
  • 2)  
    Every multiplicative subset $S$ of $U_\infty$ which is generated by an arbitrary set of elements of $\mathfrak{u}_\infty$ is a left Ore set and a right Ore set in $U_\infty $. Therefore, $S^{-1}U_\infty\simeq U_\infty S^{-1}$.

Proof. Both parts follow from Corollary 2.3 since $U_\infty=\bigcup_{n\geqslant 2}U_n$. □

§ 7. The Lie algebra $\widehat{\mathfrak{u}}_\infty$

In this section we study the completion $\widehat{\mathfrak{u}}_\infty$ of $\mathfrak{u}_\infty$ with respect to the ideal topology. Its properties diverge further from those of the $\mathfrak{u}_n$, $n\geqslant 2$, and $\mathfrak{u}_\infty$. For example, none of the non-zero inner derivations of $\widehat{\mathfrak{u}}_\infty$ is locally nilpotent, and $\widehat{\mathfrak{u}}_\infty$ is neither locally nilpotent nor locally finite-dimensional. The main result of this section is a classification of the closed and open ideals of the topological Lie algebra $\widehat{\mathfrak{u}}_\infty$ (Theorem 7.2(1)). As a result, we prove that all the open ideals and all the closed ideals of $\widehat{\mathfrak{u}}_\infty$ are topologically characteristic (Corollary 7.3(3)).

$\mathfrak{u}_\infty$ is a topological Lie algebra. Its topology is the ideal topology, that is, the topology with basis $\{u+I\mid u\in\mathfrak{u}_\infty,\,I\in\mathcal{J} (\mathfrak{u}_\infty)\}$, where $\mathcal{J}(\mathfrak{u}_\infty)$ is the set of non-zero ideals of $\mathfrak{u}_\infty$. We recall that this means that the maps $K\times\mathfrak{u}_\infty\to\mathfrak{u}_\infty$, $(\lambda,v)\mapsto\lambda v$, $\mathfrak{u}_\infty\times \mathfrak{u}_\infty\to\mathfrak{u}_\infty$, $(u,v)\mapsto u-v$, and $\mathfrak{u}_\infty\times\mathfrak{u}_\infty\to\mathfrak{u}_\infty$, $(u,v)\mapsto[u,v]$, are continuous, where the topologies on $K\times\mathfrak{u}_\infty$ and $\mathfrak{u}_\infty\times\mathfrak{u}_\infty$ are the product topologies, and the topology on $K$ is the discrete topology. By Theorem 6.2, every non-zero ideal of $\mathfrak{u}_\infty$ contains the ideal $\mathfrak{u}_{\infty,n}$ for some $n\geqslant 1$. Hence we can take the ideals $\{\mathfrak{u}_{\infty,n}\}_{n\geqslant 1}$ instead of all non-zero ideals in the definition of the topology on $\mathfrak{u}_\infty$.

The completion $\widehat{\mathfrak{u}}_\infty$ of the topological space $\mathfrak{u}_\infty$ is a topological Lie algebra:

where ${\sum_{i\geqslant 1}a_i\partial_i}$ is an infinite sum which is uniquely determined by its coefficients $a_i$. The inclusion $\mathfrak{u}_\infty\subseteq\widehat{\mathfrak{u}}_\infty$ is an inclusion of topological Lie algebras, and the topology on $\widehat{\mathfrak{u}}_\infty$ is (by definition) the strongest topology on $\widehat{\mathfrak{u}}_\infty$ such that the map $\mathfrak{u}_\infty\to\widehat{\mathfrak{u}}_\infty$, $a\mapsto a$, is continuous.

The Lie algebra $\widehat{\mathfrak{u}}_\infty$ contains a strictly descending chain of ideals

Equation (7.1)

and $\widehat{\mathfrak{u}}_\infty/\widehat{\mathfrak{u}}_{\infty,n+1}\simeq \mathfrak{u}_\infty/\mathfrak{u}_{\infty,n+1}\simeq \mathfrak{u}_{n+1}/P_n\partial_{n+1}\simeq\mathfrak{u}_n$ for all $n\geqslant 2$. One can also regard $\widehat{\mathfrak{u}}_\infty$ as the projective limit of the projective system of Lie algebra epimorphisms

where the epimorphism $\mathfrak{u}_{n+1}\to\mathfrak{u}_n$ is the composite of the natural epimorphism $\mathfrak{u}_{n+1}\to\mathfrak{u}_{n+1}/P_n\partial_{n+1}$ and the isomorphism $\mathfrak{u}_{n+1}/P_n\partial_{n+1}\simeq\mathfrak{u}_n$. Each closed ideal $\widehat{\mathfrak{u}}_{\infty,n}$ of $\widehat{\mathfrak{u}}_\infty$ is the completion/closure of the ideal $\mathfrak{u}_{\infty, n}$ of $\mathfrak{u}_\infty$. For every element $a\in\mathfrak{u}_\infty$, the set $\{a+\widehat{\mathfrak{u}}_{\infty,n}\}_{n\geqslant 1}$ is a basis of open neighbourhoods of $a$. These sets $a+\widehat{\mathfrak{u}}_{\infty,n}$, $n\geqslant 1$, are open and closed in $\widehat{\mathfrak{u}}_\infty$.

The polynomial algebra $P_\infty=\bigcup_{n\geqslant 1}P_n=K[x_1, x_2,\dots]$ is a left $\widehat{\mathfrak{u}}_\infty$-module: $a\cdot p= \sum_{i=1}^n a_i\partial_i(p)$ for all $p\in P_n$ and $a=\sum_{i\geqslant 1} a_i\partial_i\in\widehat{\mathfrak{u}}_\infty$.

Proposition 7.1. 

  • 1)  
    The Lie algebra $\widehat{\mathfrak{u}}_\infty$ is insoluble.
  • 2)  
    $\widehat{\mathfrak{u}}_\infty$ is neither locally nilpotent nor locally finite-dimensional.
  • 3)  
    Each element $u\in\widehat{\mathfrak{u}}_\infty$ acts locally nilpotently on the $\widehat{\mathfrak{u}}_\infty$-module $P_\infty$.
  • 4)  
    The chain of non-zero ideals in (7.1) is the derived series of $\widehat{\mathfrak{u}}_\infty$, that is, $(\widehat{\mathfrak{u}}_\infty)_{(i)}=\widehat{\mathfrak{u}}_{\infty,i+1}$ for all $i\geqslant 0$.
  • 5)  
    The lower central series of $\widehat{\mathfrak{u}}_\infty$ stabilizes at the first step, that is, $(\widehat{\mathfrak{u}}_{\infty})^{(0)}= \widehat{\mathfrak{u}}_{\infty}$ and $(\widehat{\mathfrak{u}}_{\infty})^{(i)}= \widehat{\mathfrak{u}}_{\infty,2}$ for all $i\geqslant 1$.
  • 6)  
    None of the non-zero inner derivations of $\widehat{\mathfrak{u}}_\infty$ are locally nilpotent.
  • 7)  
    The centralizer $\operatorname{Cen}_{\widehat{\mathfrak{u}}_\infty} (\partial_1,\partial_2,\dots)=\{\sum_{i\geqslant 1}\lambda_i\partial_i\in \widehat{\mathfrak{u}}_\infty\mid\lambda_i\in K\}$ is a maximal Abelian Lie subalgebra of $\widehat{\mathfrak{u}}_\infty$.
  • 8)  
    The centre $Z(\widehat{\mathfrak{u}}_\infty)$ of $\widehat{\mathfrak{u}}_\infty$ is equal to 0.
  • 9)  
    The Lie algebras $\mathfrak{u}_n$, where $n\geqslant 2$, $\mathfrak{u}_\infty$ and $\widehat{\mathfrak{u}}_\infty$ are pairwise non-isomorphic.

Proof. Part 1 follows from part 4.

To prove part 2, we put $a=\sum_{n\geqslant 1}\frac{x_1^n}{n!}\partial_n$. Then

Hence the elements $b_1, b_2,\dots$ are $K$-linearly independent. Therefore the Lie subalgebra $\mathcal{G}$ of $\widehat{\mathfrak{u}}_\infty$ generated by the elements $\partial_1$ and $a$ is neither nilpotent nor finite-dimensional.

Part 3 follows from Proposition 6.1(3).

To prove part 4, we put $\delta_i:=\operatorname{ad}(\partial_i)$ for $i\geqslant 1$. The assertion follows from the following obvious facts: $\partial_i\in\widehat{\mathfrak{u}}_{\infty,i}$ and ${[}\partial_i,\widehat{\mathfrak{u}}_{\infty,i}]=\widehat{\mathfrak{u}}_{\infty,i+1}$ for all $i\geqslant 1$. They immediately yield that $(\widehat{\mathfrak{u}}_\infty)_{(i)}\supseteq \widehat{\mathfrak{u}}_{\infty,i+1}$. The reverse inclusion follows from Proposition 6.1(4) since $\widehat{\mathfrak{u}}_\infty/\widehat{\mathfrak{u}}_{\infty,i+1}\simeq\mathfrak{u}_n$.

Part 5 follows since ${[}\partial_1,\widehat{\mathfrak{u}}_\infty] =\widehat{\mathfrak{u}}_{\infty,2}$ and ${[}\partial_1,\widehat{\mathfrak{u}}_{\infty,2}]=\widehat{\mathfrak{u}}_{\infty,2}$. The first equality yields that ${[}\widehat{\mathfrak{u}}_\infty,\widehat{\mathfrak{u}}_\infty]=\widehat{\mathfrak{u}}_{\infty,2}$, and the second that $(\widehat{\mathfrak{u}}_\infty)^{(i)}=\widehat{\mathfrak{u}}_{\infty,2}$ for $i\geqslant 1$.

Let us prove part 7. We have $C\,{:=}\,\operatorname{Cen}_{\widehat{\mathfrak{u}}_\infty}(\partial_1,\partial_2,\dots)\,{=}\,\{\sum_{i\geqslant 1} a_i\partial_i\,{\in}\, \widehat{\mathfrak{u}}_\infty\!\mid a_1\,{\in}\,K,a_i\in\bigcap_{j=1}^{i-1}\operatorname{ker}_{P_{i-1}}(\partial_j)=K$ for all $i\geqslant 2\}$. Since the centralizer $C=\operatorname{Cen}_{\widehat{\mathfrak{u}}_\infty}(C)$ is an Abelian Lie subalgebra of $\widehat{\mathfrak{u}}_\infty$, it is automatically maximal Abelian.

To prove part 8, let $z=\sum_{i\geqslant 1}\lambda_i\partial_i$ be a central element of $\widehat{\mathfrak{u}}_\infty$. By part 7, $\lambda_i\in K$ for all $i\geqslant 1$. For all $i\geqslant 1$ we have $\lambda_i\partial_{i+1}=[z,x_i\partial_{i+1}]= 0$, that is, $z= 0$.

Let us prove part 9. We already know that the $\mathfrak{u}_n$, where $n\geqslant 2$, and $\mathfrak{u}_\infty$ are pairwise non-isomorphic (Proposition 6.1(8)). The $\mathfrak{u}_n$ are soluble (Proposition 2.1(1)), but $\widehat{\mathfrak{u}}_\infty$ is not (part 1). Hence $\widehat{\mathfrak{u}}_\infty\not\simeq\mathfrak{u}_n$ for all $n\geqslant 2$. $\mathfrak{u}_\infty$ is locally nilpotent (Proposition 6.1(2)), but $\widehat{\mathfrak{u}}_\infty$ is not (part 2). Hence $\widehat{\mathfrak{u}}_\infty\not\simeq \mathfrak{u}_\infty$.

We now prove part 6. Since the centre of $\widehat{\mathfrak{u}}_\infty$ is equal to 0 (part 8), we must show that if $a$ is a non-zero element of $\widehat{\mathfrak{u}}_\infty$, then the inner derivation $\delta=\operatorname{ad}(a)$ is not a locally nilpotent map. Write $a=\sum_{i\in I}p_i\partial_i=p_n\partial_n+\dotsb $, where $p_i\in P_{i-1}$, $I=\{i\geqslant 1\mid p_i\neq 0\}\neq\varnothing$ and $n$ is the smallest number in $I$. If $b=\sum_{m\geqslant 0} \frac{x_n^m}{m!}\partial_{n+m}$, then

Hence $\delta$ is not a nilpotent map. □

The following theorem gives a classification of the open ideals and closed ideals of the topological Lie algebra $\widehat{\mathfrak{u}}_\infty$. Every non-zero closed ideal of $\widehat{\mathfrak{u}}_\infty$ is open and vice versa.

Theorem 7.2. 

  • 1)  
    The set $\mathcal{J}(\widehat{\mathfrak{u}}_\infty)$ of all open ideals of $\widehat{\mathfrak{u}}_\infty$ is equal to the set of all non-zero closed ideals of $\widehat{\mathfrak{u}}_\infty$ and also to the set $\{\widehat{\mathfrak{u}}_\infty,\widehat{\mathfrak{u}}_{\infty,n}, I_{\lambda,n}:=I_\lambda(n)+\widehat{\mathfrak{u}}_{\infty,n+1}\mid n\geqslant 2, \,\lambda\in [1,\omega^{n-1})\}$, where $I_\lambda(n)$ is the ideal $I_\lambda$ of $\mathfrak{u}_n$ defined in (3.3), that is, $I_\lambda (n)=\bigoplus_{(\alpha,n)\leqslant\lambda}KX_{\alpha,n}$. In particular, every non-zero closed ideal $I$ of $\widehat{\mathfrak{u}}_\infty$ contains the ideal $\widehat{\mathfrak{u}}_{\infty,n'+1}$ for some $n'\geqslant 2$, and if $I=I_{\lambda,n}$, then $n=\min\{n'\geqslant 2\mid\widehat{\mathfrak{u}}_{n'+1} \subseteq I\}$.
  • 2)  
    A non-zero ideal $I$ of $\widehat{\mathfrak{u}}_\infty$ is open if and only if it is closed.
  • 3)  
    The topological Lie algebras $\mathfrak{u}_\infty$ and $\widehat{\mathfrak{u}}_\infty$ are Hausdorff.
  • 4)  
    The zero ideal is a closed non-open ideal of $\widehat{\mathfrak{u}}_\infty$.

Proof. We prove part 1 for open ideals only. For closed ideals, it will follow from part 2. Every open ideal of $\widehat{\mathfrak{u}}_\infty$ is necessarily non-zero. Clearly, the ideals $\widehat{\mathfrak{u}}_\infty$, $\widehat{\mathfrak{u}}_{\infty,n}$ and $I_{\lambda,n}$ in the statement of part 1 are open. Let $I$ be an open ideal of $\widehat{\mathfrak{u}}_\infty$ such that $I\neq\widehat{\mathfrak{u}}_{\infty,n}$ for all $n\geqslant 1$. We must show that $I=I_{\lambda, n}$ for some $\lambda$ and $n$ as in the statement of part 1. The ideal $I$ contains an open neighbourhood of the zero element, that is, $\widehat{\mathfrak{u}}_{\infty,n+1}\subseteq I$ for some integer $n\geqslant 2$. We can assume that $n=\min\{n'\mid n'\geqslant 2,\,\widehat{\mathfrak{u}}_{\infty, n'+1}\subseteq I\}$. In view of the Lie algebra isomorphism $\widehat{\mathfrak{u}}_\infty/\widehat{\mathfrak{u}}_{\infty, n+1}\simeq \mathfrak{u}_n$, our claim and part 1 follow from Theorem 3.6(1) and the minimality of $n$.

We now prove part 2. $(\Rightarrow)$ Let $I$ be an open ideal. By part 1 for open ideals, $I$ is one of the ideals listed in part 1. All the ideals listed in part 1, that is, $\widehat{\mathfrak{u}}_\infty$, $\widehat{\mathfrak{u}}_{\infty,n}$ and $I_{\lambda,n}$, are obviously non-zero closed ideals. Hence $I$ is a non-zero closed ideal.

$(\Leftarrow)$ Let $I$ be a non-zero closed ideal in $\widehat{\mathfrak{u}}_\infty$. It suffices to show that $\widehat{\mathfrak{u}}_{\infty,n+1}\subseteq I$ for some $n\geqslant 1$. This will automatically imply that $I$ is open. The ideal $\widehat{\mathfrak{u}}_{\infty,n+1}$ is the closure of the ideal $\mathfrak{u}_{\infty, n+1}$ of $\mathfrak{u}_\infty$. We fix a non-zero element of $I$, say $a=\sum_{m\geqslant n} a_m\partial_m$, where $n\geqslant 1$, $a_m\in P_{m-1}$ for all $m\geqslant n$ and $a_n\neq 0$. Let $d$ be the total degree of the polynomial $a_n=\sum_{\alpha\in\mathbb{N}^{n-1}}\lambda_\alpha x^\alpha\in P_{n-1}$, where $\lambda_\alpha\in K$. We fix an $\alpha=(\alpha_i)\in\mathbb{N}^{n-1}$ such that $|\alpha|:=\alpha_1+\dotsb +\alpha_{n-1}=d$. Applying $\operatorname{ad}(\partial)^\alpha:=\prod_{i=1}^{n-1} \operatorname{ad}(\partial_i)^{\alpha_i}$ to the element $a$, we get an element of $I$ of type $\alpha !\lambda_\alpha\partial_n+ \dotsb$. Thus, without loss of generality, we may assume from the outset that $a_n= 1$, that is, $a=\partial_n+a_{n+1}\partial_{n+1}+\dotsb$. For every $\alpha\in\mathbb{N}^n$ we have

whence $P_n\partial_{n+1}\subseteq I$. For all integers $l\geqslant n+1$, the ideal of $\mathfrak{u}_l$ generated by the subspace $P_n\partial_{n+1}$ is equal to $\bigoplus_{i=n+1}^lP_{i-1}\partial_i$. Therefore, $\mathfrak{u}_{\infty, n+1}=\bigoplus_{i\geqslant n+1}P_{i-1} \partial_i\subseteq I$. It follows that the closure $\widehat{\mathfrak{u}}_{\infty,n+1}$ of the ideal $\mathfrak{u}_{\infty, n+1}$ belongs to $I$, as required.

We now prove part 3. Let $a=\sum_{i\geqslant 1} a_i\partial_i$ and $b=\sum_{i\geqslant 1} b_i\partial_i$ be distinct elements of $\mathfrak{u}_\infty$ (resp. $\widehat{\mathfrak{u}}_\infty$), where $a_i, b_i\in P_{i-1}$ for all $i\geqslant 1$. Then $a_n\neq b_n$ for some $n$, whence $(a+\mathfrak{u}_{\infty, n+1})\cap(b+\mathfrak{u}_{\infty,n+1})=\varnothing$ (resp. $(a+\widehat{\mathfrak{u}}_{\infty, n+1})\cap (b+\widehat{\mathfrak{u}}_{\infty, n+1})=\varnothing$). This means that $\mathfrak{u}_\infty$ and $\widehat{\mathfrak{u}}_\infty$ are Hausdorff.

Part 4 is obvious. □

A topological Lie algebra $\mathcal{G}$ is said to be closed uniserial (resp. open uniserial) if the set of all closed (resp. open) ideals of $\mathcal{G}$ is totally ordered by inclusion. By Theorem 7.2(1), the Lie algebra $\widehat{\mathfrak{u}}_\infty$ is closed uniserial and open uniserial. Hence the chain

Equation (7.2)

contains all the open ideals and non-zero closed ideals of $\widehat{\mathfrak{u}}_\infty$.

A topological Lie algebra $\mathcal{G}$ is said to be open Artinian (resp. closed Artinian) if the set of open (resp. closed) ideals of $\mathcal{G}$ satisfies the descending chain condition. A topological Lie algebra $\mathcal{G}$ is said to be open Noetherian (resp. closed Noetherian) if the set of open (resp. closed) ideals of $\mathcal{G}$ satisfies the ascending chain condition. A topological Lie algebra $\mathcal{G}$ is said to be open almost Artinian (resp. closed almost Artinian) if, for every non-zero open (resp. closed) ideal $I$ of $\mathcal{G}$, the Lie factor algebra $\mathcal{G}/I$ is Artinian. For a topological Lie algebra $\mathcal{G}$, let

be its group of automorphisms. Every element of $\operatorname{Aut}_c(\mathcal{G})$ is an isomorphism both of the Lie algebra $\mathcal{G}$ and of the topological space $\mathcal{G}$. An ideal $I$ of a topological Lie algebra $\mathcal{G}$ is said to be topologically characteristic if $\sigma(I)=I$ for all $\sigma\in \operatorname{Aut}_c(\mathcal{G})$.

Corollary 7.3. 

  • 1)  
    The topological Lie algebra $\widehat{\mathfrak{u}}_\infty$ is open uniserial, closed uniserial, open almost Artinian and closed almost Artinian. It is neither open nor closed Artinian and neither open nor closed Noetherian.
  • 2)  
    The uniserial dimensions of the sets of open and closed ideals of the topological Lie algebra $\widehat{\mathfrak{u}}_\infty$ coincide and are equal to $\operatorname{u.dim}(\widehat{\mathfrak{u}}_\infty) =\omega^\omega$.
  • 3)  
    All the open/closed ideals of the Lie algebra $\widehat{\mathfrak{u}}_\infty$ are topologically characteristic.

Proof. Part 1 follows from Theorem 7.2(1). Part 2 follows from Theorem 7.2(1) and Corollary 6.4(1).

Let us prove part 3. The ideals $\widehat{\mathfrak{u}}_{\infty,n}$ $(n\geqslant 1)$ are topologically characteristic since they form the derived series of $\widehat{\mathfrak{u}}_\infty$ (Proposition 7.1(4)). By Theorem 7.2(1), it remains to show that every ideal $I_{\lambda,n}$ of $\widehat{\mathfrak{u}}_\infty$ is topologically characteristic. Take any $\sigma\in\operatorname{Aut}_c(\widehat{\mathfrak{u}}_\infty)$. Since $\sigma(\widehat{\mathfrak{u}}_{\infty,n+1})= \widehat{\mathfrak{u}}_{\infty,n+1}$, $\sigma$ induces an automorphism $\sigma$ of the Lie factor algebra $\widehat{\mathfrak{u}}_\infty/\widehat{\mathfrak{u}}_{\infty, n+1}\simeq \mathfrak{u}_n$. Since $\widehat{\mathfrak{u}}_{\infty,n+1}\subseteq I_{\lambda,n}$ and all the ideals of $\mathfrak{u}_n$ are characteristic (Corollary 3.11), we must have $\sigma (I_{\lambda,n})= I_{\lambda,n}$, that is, $I_{\lambda,n}$ is a topologically characteristic ideal of $\widehat{\mathfrak{u}}_\infty$. □

Please wait… references are loading.
10.1070/IM2013v077n06ABEH002670