This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Sofic Lie Algebras

Cameron Cinel Department of Mathematics, University of California, San Diego, La Jolla, CA 92093-0112, USA          Email: ccinel@ucsd.edu
(March 3, 2022)
Abstract

We introduce and study soficity for Lie algebras, modelled after linear soficity in associative algebras. We introduce equivalent definitions of soficity, one involving metric ultraproducts and the other involving almost representations. We prove that any Lie algebra of subexponential growth is sofic. We also prove that a Lie algebra over a field of characteristic 0 is sofic if and only if its universal enveloping algebra is linearly sofic. Finally, we give explicit families of almost representations for the Witt and Virasoro algebras.

Key words and phrases: linearly sofic algebra, universal enveloping algebra, metric ultraproduct, subexponential growth

Mathematics Subject Classification: 17B35, 03C20, 17B68

1 Introduction

Sofic groups were first introduced by Misha Gromov [Gro99] in relation to the Gottschalk surjunctiviy surjecture conjecture (1976). The were later named sofic by Weiss [Wei00]. We recall the definitions of sofic groups via ultraproducts. For more details of ultraproducts in relation to soficity, see §2 of [Pes08].

For nn\in\mathbb{N}, we denote the symmetric group on nn letters by SnS_{n}. We consider the normalized Hamming distance

dn(σ,τ)=1n|{iσ(i)τ(i)}|.d_{n}(\sigma,\tau)=\frac{1}{n}|\{i\mid\sigma(i)\neq\tau(i)\}|.
Definition 1.1.

A group GG is sofic if there exists and ultrafilter ω\omega on \mathbb{N} and a sequence of natural numbers (nk)(n_{k}) such that GG embeds into the metric ultraproduct kω(Snk,dnk)\prod_{k\to\omega}(S_{n_{k}},d_{n_{k}}).

In [ES05], this definition was reformulated without the notion of ultraproducts. It was shown that a group is sofic if and only if it can be locally embedded in some symmetric group almost homomorphically, up to some arbitrarily small error. Similar ultraproducts can be constructed from any family of groups equipped with bi-invariant metrics, such as unitary groups over \mathbb{C} (where subgroups are called hyperlinear).

In particular, for any field FF, the group GLn(F)GL_{n}(F) can be endowed the normalized rank metric

ρn(A,B)=1nrank(AB).\rho_{n}(A,B)=\frac{1}{n}\text{rank}(A-B).

Groups embedable into ultraproducts of GLn(F)GL_{n}(F) with this metric are called linearly sofic. In [AP17] the notion of linear soficity was extended to associative algebras by considering metric ultraproducts of Mn(F)M_{n}(F). In particular, it was shown that a group GG is linearly sofic if and only if its group algebra [G]\mathbb{C}[G] is linearly sofic.

The goal of this paper to do analogous work for Lie algebras. Our first section deals with the ultraproduct construction and definition of soficity for Lie algebras as well as basic properties. We then provide an equivalent characterization for soficity using almost representations for our Lie algebra.

We also provide examples of sofic Lie algebras using families of almost representations. In particular we prove one of our primary results

Theorem 4.2.

Any Lie algebra of subexponential growth is sofic.

We also provide explicit almost representations to show soficity for the Witt and Virasoro algebras. This is analogous to soficity in groups and associative algebras, where amenability, and therefore subexponential growth, implies soficity.

Finally, we show that soficity for Lie algebras and associative algebras is compatible.

Theorem 5.2.

Let LL be a Lie algebra over a field of characteristic 0. Then LL is sofic if and only if its universal enveloping algebra U(L)U(L) is sofic.

This result should be compared to other work on the growth of Lie algebras, notably the theorem in [Smi76] showing that Lie algebras of subexponential growth have universal enveloping algebras of subexponential growth.

2 Ultraproducts

We define our notion of soficity for Lie algebras using the ultra product construction, similar to that of sofic groups and associative algebras. For nn\in\mathbb{N}, we call the function

ρn:Mn(F)\displaystyle\rho_{n}:M_{n}(F) [0,1]\displaystyle\to[0,1]
a\displaystyle a 1nrank(a)\displaystyle\mapsto\frac{1}{n}\text{rank}(a)

the normalized rank function. Let ω\omega be a non-principal ultrafilter on \mathbb{N} and nkn_{k} be a sequence of natural numbers such that limknk=\lim_{k\to\infty}n_{k}=\infty. We can extend the ρnk\rho_{n_{k}}’s to a function

ρnk:Mnk(F)\displaystyle\rho_{n_{k}}:M_{n_{k}}(F) [0,1]\displaystyle\to[0,1]
(ak)\displaystyle(a_{k}) limkωρnk(ak).\displaystyle\mapsto\lim\limits_{k\to\omega}\rho_{n_{k}}(a_{k}).

In [AP17], the pre-image ρω1(0)\rho_{\omega}^{-1}(0) was shown to be an ideal. Thus we can construct the ultra-product of the matrix rings, denoted

kωMnk(F)/kerρω:=k=1Mnk(F)/ρω1(0).\prod_{k\to\omega}M_{n_{k}}(F)/\ker\rho_{\omega}:=\prod_{k=1}^{\infty}M_{n_{k}}(F)/\rho_{\omega}^{-1}(0).

As an associative algebra, this ultra-product has a natural Lie algebra structure via the commuatator bracket, which we denote

kω𝔤𝔩nk(F)/kerρω:=(kωMnk(F)/kerρω)().\prod_{k\to\omega}\mathfrak{gl}_{n_{k}}(F)/\ker\rho_{\omega}:=\left(\prod_{k\to\omega}M_{n_{k}}(F)/\ker\rho_{\omega}\right)^{(-)}.
Definition 2.1.

We denote the metric ultraproduct of the Lie algebras 𝔤𝔩nk(F)\mathfrak{gl}_{n_{k}}(F) as kω𝔤𝔩nk(F)/kerρω\prod_{k\to\omega}\mathfrak{gl}_{n_{k}}(F)/\ker\rho_{\omega}, called a universal sofic Lie algebra.

For any subalgebra L𝔤𝔩n(F)L\subset\mathfrak{gl}_{n}(F), we can restrict the normalized rank map and hence create different metric ultraproducts.

Lemma 2.1.

For any field FF,

kω𝔤𝔩nk(F)/kerρωkω𝔰𝔩nk(F)/kerρω.\prod_{k\to\omega}\mathfrak{gl}_{n_{k}}(F)/\ker\rho_{\omega}\cong\prod_{k\to\omega}\mathfrak{sl}_{n_{k}}(F)/\ker\rho_{\omega}.
Proof.

Clearly, we can embed kω𝔰𝔩nk(F)/kerρωkω𝔤𝔩nk(F)/kerρω\prod_{k\to\omega}\mathfrak{sl}_{n_{k}}(F)/\ker\rho_{\omega}\hookrightarrow\prod_{k\to\omega}\mathfrak{gl}_{n_{k}}(F)/\ker\rho_{\omega}. Moreover, for any sequence (Ak)kω𝔤𝔩nk(F)/kerρω(A_{k})\in\prod_{k\to\omega}\mathfrak{gl}_{n_{k}}(F)/\ker\rho_{\omega}, we can consider the second sequence (AkE11tr(A))(A_{k}-E_{11}\text{tr}(A)). Then

ρω((Ak)(AkE11tr(A)))=limkω1nk=0\rho_{\omega}((A_{k})-(A_{k}-E_{11}\text{tr}(A)))=\lim_{k\to\omega}\frac{1}{n_{k}}=0

so (Ak)=(AkE11tr(A))(A_{k})=(A_{k}-E_{11}\text{tr}(A)) in the ultra-product. Thus the natural embedding is a surjection. ∎

We now define the map object of this paper.

Definition 2.2.

A Lie algebra LL over a field FF is called linearly sofic if there exists a Lie algebra embedding of LL into some metric ultraproduct of some 𝔤𝔩nk(F)\mathfrak{gl}_{n_{k}}(F)’s.

Proposition 2.2.

For an algebraically closed field FF, the associative algebra kωMnk(F)/kerρω\prod_{k\to\omega}M_{n_{k}}(F)/\ker\rho_{\omega} is simple.

Proof.

If ω\omega is a principal ultrafilter, this is trivially true. So we assume that ω\omega is free.

Denote R:=kωMnk(F)/kerρωR:=\prod_{k\to\omega}M_{n_{k}}(F)/\ker\rho_{\omega}. Suppose 0xR0\neq x\in R. Let δ:=ρω(x)\delta:=\rho_{\omega}(x). Since δ>0\delta>0, there exists nn\in\mathbb{N} such that 1n<δ\frac{1}{n}<\delta. Thus, if (Bk)Mnk(F)(B_{k})\in\prod M_{n_{k}}(F) is a representative for xx, there exists some SωS\in\omega such that ρnk(Bk)1n\rho_{n_{k}}(B_{k})\geq\frac{1}{n} for every kSk\in S.

Let JkJ_{k} be the Jordan normal form of BkB_{k}. Then JkJ_{k} row equivalent to

Ak(m):=i=1nknEmnkn+i,mnkn+iA^{(m)}_{k}:=\sum_{i=1}^{\lfloor\frac{n_{k}}{n}\rfloor}E_{\lfloor\frac{mn_{k}}{n}\rfloor+i,\lfloor\frac{mn_{k}}{n}\rfloor+i}

where 0mn10\leq m\leq n-1. Thus, there exists Pk,QkMnk(F)P_{k},Q_{k}\in M_{n_{k}}(F) such that

Bk=PkAk(m)Qk.B_{k}=P_{k}A^{(m)}_{k}Q_{k}.

Define CkMnk(F)C_{k}\in M_{n_{k}}(F) via

Ck={m=0n1Ak(m),kS0,kSC.C_{k}=\begin{cases}\sum_{m=0}^{n-1}A^{(m)}_{k},&k\in S\\ 0,&k\in S^{C}\end{cases}.

Then if y=[Ck]Ry=[C_{k}]\in R, we have that y(x)y\in(x), the ideal generated by xx. Moreover, we notice that for kSk\in S,

rank(CkInk)n1.\text{rank}(C_{k}-I_{n_{k}})\leq n-1.

Let ε>0\varepsilon>0. Since ω\omega is a free ultrafilter, SS must be infinite. Therefore, there exists an infinite subset SSS^{\prime}\subset S such that ρnk(CkInk)<ε\rho_{n_{k}}(C_{k}-I_{n_{k}})<\varepsilon. Notice that SωS^{\prime}\in\omega since otherwise, (S)CSω(S^{\prime})^{C}\cap S\in\omega making ω\omega a principal ultrafilter. Therefore ρω(y[Ink])=0\rho_{\omega}(y-[I_{n_{k}}])=0 and (x)=R(x)=R. ∎

Corollary 2.2.1.

For an algebraically closed field FF, the non-trivial ideals of kω𝔤𝔩nk(F)/kerρω\prod_{k\to\omega}\mathfrak{gl}_{n_{k}}(F)/\ker\rho_{\omega} are contained in its center.

Proof.

This follows from Theorem 2 of [Her61]. ∎

3 Almost representations

Just as with sofic groups and linearly sofic associative algebras, linearly sofic Lie algebras can be defined via families of maps that are Lie algebra homomorphisms up to a small error. We follow a similar approach to section 11.2 of [AP17].

Definition 3.1.

Let LL be a Lie algebra, WLW\subset L a finite dimensional subspace, VV a finite dimensional vector space, and ε>0\varepsilon>0. A linear map φ:W𝔤𝔩(V)\varphi:W\to\mathfrak{gl}(V) is called and ε\varepsilon-almost representation of WW if there exists a subspace VεVV_{\varepsilon}\subset V such that:

  1. 1.

    for all x,yWx,y\in W such that [x,y]W[x,y]\in W,

    φ([x,y])|Vε=(φ(x)φ(y)φ(y)φ(x))|Vε;\varphi([x,y])|_{V_{\varepsilon}}=(\varphi(x)\varphi(y)-\varphi(y)\varphi(x))|_{V_{\varepsilon}};
  2. 2.

    dimV\dim V-dimVεεdimV\dim V_{\varepsilon}\leq\varepsilon\cdot\dim V.

Clearly a morphism from a Lie algebra to a universal sofic Lie algebra gives rise to a family of almost representations of the Lie algebra. Similarly, a family of almost representations on a covering of a Lie algebra gives a morphism to a universal sofic Lie algebra. However, not all families of almost representations will correspond to embeddings. In particular, some elements of our Lie algebra will have to be mapped to arbitrarily small rank transformations via the family of almost representations. This gives rise to to the following subspace of our Lie algebra.

Definition 3.2.

For a Lie algebra LL, the sofic radical of LL, denote SR(L)SR(L), is defined as follows: pSR(L)p\in SR(L) if for every δ>0\delta>0, there exists a finite dimensional WLW\subset L containing pp and nδ>0n_{\delta}>0 such that if 0<ε<nδ0<\varepsilon<n_{\delta} and φW,ε:W𝔤𝔩(V)\varphi_{W,\varepsilon}:W\to\mathfrak{gl}(V) is an ε\varepsilon-almost representation, then

dimIm φW,ε(p)<δdimV.\dim\text{Im }\varphi_{W,\varepsilon}(p)<\delta\cdot\dim V.

The sofic radical is essentially the collection of ”bad” elements of our Lie algebra when it comes to trying to make embeddings of it into a universal sofic Lie algebra. This view is summarized by the following lemma and corollary.

Lemma 3.1.

For a Lie algebra LL, pSR(L)p\in SR(L) if and only if for every Lie algebra homomorphism

Θ:Lω𝔤𝔩nk(F)/ρω\Theta:L\to\prod_{\omega}\mathfrak{gl}_{n_{k}}(F)/\rho_{\omega}

we have that Θ(p)=0\Theta(p)=0.

Proof.

Suppose pSR(L)p\in SR(L) and let Θ:Lω𝔤𝔩nk(F)/ρω\Theta:L\to\prod_{\omega}\mathfrak{gl}_{n_{k}}(F)/\rho_{\omega} be a Lie algebra homomorphism with lifts θk:L𝔤𝔩nk(F)\theta_{k}:L\to\mathfrak{gl}_{n_{k}}(F). Fix δ>0\delta>0 and choose nδn_{\delta} and finite dimensional WLW\subset L from the definition of the sofic radical. Choose 0<ε<nδ0<\varepsilon<n_{\delta}. Then there exists SωS\in\omega such that θk|W\theta_{k}|_{W} is an ε\varepsilon-almost representation for WW for every kSk\in S. Therefore dimIm θk(p)<δnk\dim\text{Im }\theta_{k}(p)<\delta n_{k}. In other words, ρnk(θk(p))<δ\rho_{n_{k}}(\theta_{k}(p))<\delta for every kHk\in H. Therefore ρω(Θ(p))<δ\rho_{\omega}(\Theta(p))<\delta. Since δ\delta was arbitrary, we get that Θ(p)=0\Theta(p)=0.

Now suppose that pLSR(L)p\in L\setminus SR(L). Then there exists δ>0\delta>0 such that for any finite dimensional WLW\subset L containing pp and n>0n>0, there exists 0<ε<n0<\varepsilon<n and ε\varepsilon-almost representation φ:L𝔤𝔩(V)\varphi:L\to\mathfrak{gl}(V) such that dimIm φ(p)δdimV\dim\text{Im }\varphi(p)\geq\delta\cdot\dim V.

Let WkW_{k} be an increasing sequence of finite dimensional subspaces of LL containing pp that cover LL. Then there exists a sequence of εk\varepsilon_{k} that converges to 0 and εk\varepsilon_{k}-almost representations θk:Wk𝔤𝔩(Vk)\theta_{k}:W_{k}\to\mathfrak{gl}(V_{k}) such that dimIm θk(p)δdimVk\dim\text{Im }\theta_{k}(p)\geq\delta\cdot\dim V_{k}. Define a map Θ^:L𝔤𝔩(Vk)\hat{\Theta}:L\to\prod\mathfrak{gl}(V_{k}) by Θ^(x)=(θ^k(x))\hat{\Theta}(x)=(\hat{\theta}_{k}(x)) where

θ^k(x)={θk(x),xWk0,xLWk.\hat{\theta}_{k}(x)=\begin{cases}\theta_{k}(x),&x\in W_{k}\\ 0,&x\in L\setminus W_{k}\end{cases}.

Choose a non-principal ultrafilter ω\omega of \mathbb{N} and let Θ:Lω𝔤𝔩(Vk)/ρω\Theta:L\to\prod_{\omega}\mathfrak{gl}(V_{k})/\rho_{\omega} be the composition of Θ^\hat{\Theta} with the quotient map. Then since εk0\varepsilon_{k}\to 0, we have that Θ\Theta is a Lie algebra homomorphism and ρω(Θ(p))δ>0\rho_{\omega}(\Theta(p))\geq\delta>0. ∎

From this lemma, we get the following corollary

Corollary 3.1.1.

For a Lie algebra LL, SR(L)SR(L) is an ideal. Moreover, SR(L/SR(L))=(0)SR(L/SR(L))=(0).

Proof.

The lemma show that

SR(L)=ΘSkerΘSR(L)=\bigcap_{\Theta\in S}\ker\Theta

where

S={Θ:Lω𝔤𝔩nk(F)/ρωΘ is a Lie algebra homomorphism}S=\left\{\Theta:L\to\prod_{\omega}\mathfrak{gl}_{n_{k}}(F)/\rho_{\omega}\mid\Theta\text{ is a Lie algebra homomorphism}\right\}

so it is clear that SR(L)SR(L) is an ideal.

Now suppose 0pL/SR(L)0\neq p\in L/SR(L) and let qLq\in L be a pre-image of pp. Then qSR(L)q\notin SR(L) so there exists a Lie algebra homomorphism Θ:Lω𝔤𝔩nk(F)/ρω\Theta:L\to\prod_{\omega}\mathfrak{gl}_{n_{k}}(F)/\rho_{\omega} such that Θ(q)0\Theta(q)\neq 0. Since SR(L)kerΘSR(L)\subset\ker\Theta, we can get a map Θ^\hat{\Theta} by composing Θ\Theta with the quotient LL/SR(L)L\to L/SR(L). Then we have that Θ^(p)=Θ(q)0\hat{\Theta}(p)=\Theta(q)\neq 0 so pSR(L/SR(L))p\notin SR(L/SR(L)). ∎

We can now use the sofic radical to characterize sofic Lie algebras.

Theorem 3.2.

A Lie algebra LL is sofic if and only if SR(L)=(0)SR(L)=(0).

Proof.

The forward implication is trivial so we only show the reverse implication.

Let LL be a Lie algebra such that SR(L)=0SR(L)=0. Then for every pL(0)p\in L\setminus(0), there exists a Lie algebra homomorphism Θp:Lω𝔤𝔩nk,p(F)/ρω\Theta_{p}:L\to\prod_{\omega}\mathfrak{gl}_{n_{k,p}}(F)/\rho_{\omega} such that Θp(p)0\Theta_{p}(p)\neq 0.

Let {xi}iL\{x_{i}\}_{i\in\mathbb{N}}\subset L be a basis for LL as a vector space. We shall construct maps Ψm:Lω𝔤𝔩nk,m/ρω\Psi_{m}:L\to\prod_{\omega}\mathfrak{gl}_{n_{k,m}}/\rho_{\omega} such that kerΨmspan{x1,,xm}=(0)\ker\Psi_{m}\cap\text{span}\{x_{1},\dots,x_{m}\}=(0) for every mm\in\mathbb{N}.

Let Ψ1=Θx1\Psi_{1}=\Theta_{x_{1}}. Now suppose for m2m\in\mathbb{N}_{\geq 2}, we have a map Ψm1\Psi_{m-1} as above. Then

dim(Ψm1span{x1,,xm})1.\dim(\Psi_{m-1}\cap\text{span}\{x_{1},\dots,x_{m}\})\leq 1.

If the dimension is 0, let Ψm=Ψm1\Psi_{m}=\Psi_{m-1}. Otherwise, choose a non-zero element ymy_{m} in the intersection. In this case, we define Ψm=Ψm1Θym\Psi_{m}=\Psi_{m-1}\oplus\Theta_{y_{m}}. Then if zkerΨmspan{x1,,xm}z\in\ker\Psi_{m}\cap\text{span}\{x_{1},\dots,x_{m}\}, we have that zkerΨm1span{x1,,xm}z\in\ker\Psi_{m-1}\cap\text{span}\{x_{1},\dots,x_{m}\}. Thus z=αymz=\alpha y_{m} for some αF\alpha\in F. However, zkerΘymz\in\ker\Theta_{y_{m}} so z=0z=0.

Now we construct a map Φ:Lω𝔤𝔩nk(F)/ρω\Phi:L\to\prod_{\omega}\mathfrak{gl}_{n_{k}}(F)/\rho_{\omega} such that kerΦkerΨm\ker\Phi\subset\bigcap\ker\Psi_{m}. Let nk=nk,1nk,kn_{k}=n_{k,1}\cdots n_{k,k}. Tensor each component of Ψm\Psi_{m} with an appropriately sized identity matrix gives us maps Ψ^m:Lω𝔤𝔩n^k,m(F)/ρω\hat{\Psi}_{m}:L\to\prod_{\omega}\mathfrak{gl}_{\hat{n}_{k,m}}(F)/\rho_{\omega} where n^k,m=nk\hat{n}_{k,m}=n_{k} if mkm\leq k and nk,mn_{k,m} otherwise. Let ψ^k,m:L𝔤𝔩nk(F)\hat{\psi}_{k,m}:L\to\mathfrak{gl}_{n_{k}}(F) be a lift of Ψ^m\hat{\Psi}_{m} for mkm\leq k.

Define a map φk:A𝔤𝔩2knk\varphi_{k}:A\to\mathfrak{gl}_{2^{k}n_{k}} by

φk=(ψ^k,1Id2k1(ψ^k,2Id2k2)(ψ^k,kId1)Idnk.\varphi_{k}=(\hat{\psi}_{k,1}\otimes\text{Id}_{2^{k-1}}\oplus(\hat{\psi}_{k,2}\otimes\text{Id}_{2^{k-2}})\oplus\cdots\oplus(\hat{\psi}_{k,k}\otimes\text{Id}_{1})\oplus\text{Id}_{n_{k}}.

Let Φ=ωφk/ρω\Phi=\prod_{\omega}\varphi_{k}/\rho_{\omega}.

A direct calculation gives us that for any xLx\in L

ρω(Φ(x))\displaystyle\rho_{\omega}(\Phi(x)) =limkωrank(φk(x))2knkk\displaystyle=\lim_{k\to\omega}\frac{\text{rank}(\varphi_{k}(x))}{2^{k}n_{k}^{k}}
=limkω12knkki=1krank((ψ^k,1Id2k1)(x))\displaystyle=\lim_{k\to\omega}\frac{1}{2^{k}n_{k}^{k}}\sum_{i=1}^{k}\text{rank}((\hat{\psi}_{k,1}\otimes\text{Id}_{2^{k-1}})(x))
=limkω1nkki=1krank(ψ^k,1(x))2i\displaystyle=\lim_{k\to\omega}\frac{1}{n_{k}^{k}}\sum_{i=1}^{k}\frac{\text{rank}(\hat{\psi}_{k,1}(x))}{2^{i}}
=limkωi=1krank(ψk,1(x))2ink,i\displaystyle=\lim_{k\to\omega}\sum_{i=1}^{k}\frac{\text{rank}(\psi_{k,1}(x))}{2^{i}n_{k,i}}
=i=1k12iρω(Ψi(x))\displaystyle=\sum_{i=1}^{k}\frac{1}{2^{i}}\rho_{\omega}(\Psi_{i}(x))

where ψk,m:L𝔤𝔩nk,m\psi_{k,m}:L\to\mathfrak{gl}_{n_{k,m}} is a lift of Ψm\Psi_{m}.

Therefore kerΦkerΨm=(0)\ker\Phi\subset\bigcap\ker\Psi_{m}=(0), so we have that Φ\Phi is injective. ∎

4 Examples of Sofic Lie Algebras

By using the sofic radical, we can determine if particular Lie algebras is sofic.

Proposition 4.1.

Any abelian Lie algebra is sofic.

Proof.

Suppose LL is an abelian Lie algebra over a field FF and suppose that 0pL0\neq p\in L. Then for any finite dimensional subspace pWLp\in W\subset L and ε>0\varepsilon>0, there exists an ε\varepsilon-almost representation φ:WF𝔤𝔩1(F)\varphi:W\to F\cong\mathfrak{gl}_{1}(F) such that φ(p)=1\varphi(p)=1 and φ(WFp)=0\varphi(W\setminus Fp)=0. Thus dimImφ(p)=1=dimF\dim\text{Im}\varphi(p)=1=\dim F so pSR(L)p\notin SR(L). ∎

For groups, those of subexponential growth are amenable and therefore sofic. We show the same result holds true for Lie alebras.

Theorem 4.2.

Any Lie algebra of subexponential growth is sofic.

Proof.

Let LL be a Lie algebra of subeponential growth generated by the set XLX\subset L. Let VnV_{n} denote the words in LL of length at most nn. We inductively create a basis for LL as follows. Let {x1,,xdimV1}\{x_{1},\dots,x_{\dim V_{1}}\} be a basis for V1V_{1}. For n2n\geq 2, let {xdimVn1+1,,xdimVn}\{x_{\dim V_{n-1}+1},\dots,x_{\dim V_{n}}\} be the preimage in LL of a basis for Vn/Vn1V_{n}/V_{n-1}. We also define WnW_{n} to be the subspace of U(L)U(L) spanned by words of length at most nn and γ(n)=dimWn\gamma(n)=\dim W_{n}.

For m>nm>n, we define a linear map φn,m:Vn𝔤𝔩(Wm)\varphi_{n,m}:V_{n}\to\mathfrak{gl}(W_{m}) where

φn,m(xi)(xj1xjk)={xixj1xjk,xixj1xjkWm0,otherwise.\varphi_{n,m}(x_{i})(x_{j_{1}}\cdots x_{j_{k}})=\begin{cases}x_{i}x_{j_{1}}\cdots x_{j_{k}},&x_{i}x_{j_{1}}\cdots x_{j_{k}}\in W_{m}\\ 0,&\text{otherwise}\end{cases}.

We notice on that for vWmnv\in W_{m-n} and x,yVnx,y\in V_{n},

[φn,m,k(x),φn,m,k(y)](v)=φn,m,k([x,y])(v).[\varphi_{n,m,k}(x),\varphi_{n,m,k}(y)](v)=\varphi_{n,m,k}([x,y])(v).

Let us show that for a fixed nn that

limmγ(mn)γ(m)=1.\lim_{m\to\infty}\frac{\gamma(m-n)}{\gamma(m)}=1.

Since LL is of subexponential growth, we have that γ\gamma is a function of subexponential growth from [Smi76]. To see this suppose that f:f:\mathbb{R}\to\mathbb{R} is a function of subexponential growth and d>0d>0. We have that if

limxf(x)f(x+d)=1L<1\lim_{x\to\infty}\frac{f(x)}{f(x+d)}=\frac{1}{L}<1

Then there exists x0x_{0}\in\mathbb{R} such that

f(x0+nd)Lnf(x0).f(x_{0}+nd)\geq L^{n}f(x_{0}).

Therefore f(x)Lxx0d1f(x0)f(x)\geq L^{\frac{x-x_{0}}{d-1}}f(x_{0}) and is therefore not of subexponential growth.

Thus for any nn\in\mathbb{N} and ε>0\varepsilon>0, we can find φn,m\varphi_{n,m} such that

dimWmdimWnmdimWm<1(1ε)=ε.\frac{\dim W_{m}-\dim W_{n-m}}{\dim W_{m}}<1-(1-\varepsilon)=\varepsilon.

Now suppose pL{0}p\in L\setminus\{0\}, VLV\subset L is a finite dimensional subspace containing pp, and ε>0\varepsilon>0. Then there exists nn\in\mathbb{N} such that VVnV\subset V_{n}. Thus for a sufficiently large mm\in\mathbb{N}, we have an ε\varepsilon-almost representation ψ:=φn,m|V\psi:=\varphi_{n,m}|_{V} for VV. On WmnW_{m-n}, we have that ψ(p)\psi(p) works like left multiplication by pp, considered as an element of U(L)U(L). Thus dimIm(ψ(p))dimWmn\dim\text{Im}(\psi(p))\geq\dim W_{m-n}. By our choice of mm to make ψ\psi an ε\varepsilon-almost representation, we have that

dimWmn>(1ε)dimWm.\dim W_{m-n}>(1-\varepsilon)\dim W_{m}.

Hence the normalized rank of ψ(p)\psi(p) is greater than 1ε1-\varepsilon.

Thus for any non-zero pLp\in L, any finite dimensional subspace pVLp\in V\subset L, and ε>0\varepsilon>0, we can find an ε\varepsilon-almost representation ψ\psi for VV such that the normalized rank of ψ(p)\psi(p) is at least 1/2. Specifically, we can choose ψ\psi to be a min{ε,1/2}\min\{\varepsilon,1/2\}-almost representation. Therefore pSR(L)p\notin SR(L) so SR(L)=(0)SR(L)=(0) and LL is sofic. ∎

We now provide explicit families of almost representations for two particular Lie algebras. Though both of the examples are of subexponential growth, their families of almost homomorphisms are not based on their subspaces of words of particular length.

Example 4.3.

The Witt algebra LWL_{W}, which is the Lie algebra of derivations for [t,t1]\mathbb{C}[t,t^{-1}], is sofic.

We can cover LWL_{W} by the finitely dimensional subspaces

Vn=i=nnxiV_{n}=\sum_{i=-n}^{n}\mathbb{C}x_{i}

where xi=tn+1ddtx_{i}=-t^{n+1}\frac{d}{dt}. We also consider the finite dimensional vector spaces

Wn=i=nnti[t,t1].W_{n}=\sum_{i=-n}^{n}\mathbb{C}t^{i}\subset\mathbb{C}[t,t^{-1}].

For mnm\geq n, define a linear map φn,m:Vn𝔤𝔩(Wm)\varphi_{n,m}:V_{n}\to\mathfrak{gl}(W_{m}) via

φn,m(xi)(tj)={jti+j,jmi0,otherwise\varphi_{n,m}(x_{i})(t^{j})=\begin{cases}-jt^{i+j},&j\leq m-i\\ 0,&\text{otherwise}\end{cases}

and

φn,m(xi)(tj)={jtji,jim0,otherwise\varphi_{n,m}(x_{-i})(t^{j})=\begin{cases}-jt^{j-i},&j\geq i-m\\ 0,&\text{otherwise}\end{cases}

for i0i\geq 0. For any nn\in\mathbb{N} and ε>0\varepsilon>0, mm can be made sufficiently large so as to make φn,m\varphi_{n,m} and ε\varepsilon-almost homomorphism.

Example 4.4.

The Virasoro algebra LVL_{V}, which is the unique central extension of LWL_{W}, is sofic.

To construct the family of almost representations, we need to use the Verma modules of LVL_{V}. For more information on the Verma modules, see [WY86].

We have a basis of LVL_{V} given by the xix_{i}’s from LWL_{W} and a central element cc. Define the subspaces 𝔥=x0+c\mathfrak{h}=\mathbb{C}x_{0}+\mathbb{C}c and

𝔫+=k=1xk.\mathfrak{n}_{+}=\sum_{k=1}^{\infty}\mathbb{C}x_{k}.

Given λ𝔥\lambda\in\mathfrak{h}^{*}, we define the Verma module to be the space

M(λ)=U(LV)U(𝔥𝔫+)M(\lambda)=U(L_{V})\otimes_{U(\mathfrak{h}\oplus\mathfrak{n}_{+})}\mathbb{C}

where the action of 𝔥𝔫+\mathfrak{h}\oplus\mathfrak{n}_{+} on \mathbb{C} is given by

(h+x)α=λ(h)α.(h+x)\cdot\alpha=\lambda(h)\alpha.

For m,dm,d\in\mathbb{N}, we consider the subspaces

M(λ)m,d=span{xi1xik10kr,0iki1m}.M(\lambda)_{m,d}=\text{span}\{x_{-i_{1}}\cdots x_{-i_{k}}\otimes 1\mid 0\leq k\leq r,0\leq i_{k}\leq\cdots\leq i_{1}\leq m\}.

We then define linear maps φn,m,d:Vn+c𝔤𝔩(M(λ)m,d)\varphi_{n,m,d}:V_{n}+\mathbb{C}c\to\mathfrak{gl}(M(\lambda)_{m,d}) via φn,m,d(c)=idM(λ)m,d\varphi_{n,m,d}(c)=\text{id}_{M(\lambda)_{m,d}} and

φn,m,d(xr)(xi1xik)={0,k=d or ri1mxkxi1xik,otherwise.\varphi_{n,m,d}(x_{r})(x_{-i_{1}}\cdots x_{-i_{k}})=\begin{cases}0,&k=d\text{ or }r\geq i_{1}-m\\ x_{k}x_{-i_{1}}\cdots x_{-i_{k}},&\text{otherwise}\end{cases}.

Just as in the case with LWL_{W}, for a fixed n,dn,d\in\mathbb{N} and ε>0\varepsilon>0, we can choose mm large enough to make φn,m,d\varphi_{n,m,d} an ε\varepsilon-almost representation.

5 Soficity of Universal Enveloping Algebras

In [AP17], it was shown that a group is linearly sofic if and only if its corresponding group algebra is itself linearly sofic. For Lie algebras, we have a similar construction with universal enveloping algebras. This section is spent on showing equivalence of soficity for Lie algebras and universal enveloping algebras.

We first need the following result from [AP17].

Proposition 5.1.

Let AA be a unital algebra and ω𝒫()\omega\subset\mathscr{P}(\mathbb{N}) an ultrafilter. If we have a set {Θi}iI\{\Theta^{i}\}_{i\in I} of unital algebra morphisms from AA to ωMnk,i(F)/ρω\prod_{\omega}M_{n_{k,i}}(F)/\rho_{\omega}, then there exists a unital algebra homomorphism ψ:AωMm(F)/ρω\psi:A\to\prod_{\omega}M_{m}(F)/\rho_{\omega} such that

kerψiIkerΘi.\ker\psi\subset\bigcap_{i\in I}\ker\Theta^{i}.

We can now tackle our main result:

Theorem 5.2.

Let LL be a Lie algebra over a field FF of characteristic 0. Then LL is sofic if and only if U(L)U(L) is sofic.

Proof.

Since LL embeds in U(L)U(L) the reverse direction is trivial. Thus we only prove the forward direction.

Since LL is sofic, there exists a Lie algebra homomorphism

Θ:Lω𝔤𝔩nk(F)/ρω=:A.\Theta:L\to\prod_{\omega}\mathfrak{gl}_{n_{k}}(F)/\rho_{\omega}=:A.

Let Θ1=Θ\Theta_{1}=\Theta and

Θj=11j1 timesΘ\Theta_{j}=\underbrace{1\otimes\cdots\otimes 1}_{j-1\text{ times}}\otimes\Theta

for j2j\geq 2. Define a new Lie algebra homomorphism Θi:Lj=1iA\Theta^{i}:L\to\bigotimes_{j=1}^{i}A by Θ0=0\Theta^{0}=0, Θ1=Θ\Theta^{1}=\Theta, and Θi=Θi11+1Θi1\Theta^{i}=\Theta^{i-1}\otimes 1+1\otimes\Theta_{i-1}. Since finite tensor product of matrix rings are again matrix rings, we use the proposition to get a unital algebra homomorphism

ψ:U(L)ωMm(F)/ρω\psi:U(L)\to\prod_{\omega}{M_{m}(F)}/\rho_{\omega}

such that

kerψi=0Θ~i.\ker\psi\subset\bigcap_{i=0}^{\infty}\widetilde{\Theta}^{i}.

We claim that ψ\psi is injective.

Let {xj}jJL\{x_{j}\}_{j\in J}\subset L be a basis. First suppose that xj1xjdkerψx_{j_{1}}\cdots x_{j_{d}}\in\ker\psi is a monomial for d2d\geq 2. Then by construction

Θ~i(xj1xjd)=0\widetilde{\Theta}^{i}(x_{j_{1}}\cdots x_{j_{d}})=0

for all i1i\geq 1. Notice that Θd(xj1xjd)\Theta^{d}(x_{j_{1}}\cdots x_{j_{d}}) will consist of a linear sum of pure tensors of 11 and products of the Θ(xji)\Theta(x_{j_{i}})’s. In fact, we have two sets of pure tensors: one set consisting of those with at least one 1 and at least one degree 2 or greater monomial and the other set consisting of pure tensors with only linear monomials and no 1’s, i.e. pure tensors of the form xjσ(1)xjσ(d)x_{j_{\sigma(1)}}\otimes\cdots\otimes x_{j_{\sigma(d)}} for some σSd\sigma\in S_{d}. In fact we get exactly one such pure tensor for every σSd\sigma\in S_{d}. Call the first set the lower order tensors and the second set the high order tensors.

Now notice that Θ~d1(xj1xjd)\widetilde{\Theta}^{d-1}(x_{j_{1}}\cdots x_{j_{d}}) is a sum of low order tensors of Θ~d(xj1xjd)\widetilde{\Theta}^{d}(x_{j_{1}}\cdots x_{j_{d}}) with one of the 1’s removed. Thus, viewing 0 as 0000\otimes 0\otimes\cdots\otimes 0 (d1d-1 times), we can tensor the equation

Θ~d1(xj1xjd)=0\widetilde{\Theta}^{d-1}(x_{j_{1}}\cdots x_{j_{d}})=0

by 1 in dd spots and adding up all the new equations shows that the low order tensors of Θ~d(xj1xjd)\widetilde{\Theta}^{d}(x_{j_{1}}\cdots x_{j_{d}}) sum to 0. Therefore, we get that

σSdΘ(xjσ(1))Θ(xjσ(d))=0.\sum_{\sigma\in S_{d}}\Theta(x_{j_{\sigma(1)}})\otimes\cdots\otimes\Theta(x_{j_{\sigma(d)}})=0.

Now suppose that i>di>d. Then we can see that in the sum of pure tensors in Θ~i(xj1xjd)\widetilde{\Theta}^{i}(x_{j_{1}}\cdots x_{j_{d}}), every pure tensor has at least one 1 in it. Thus repeating the same process of the tensor the equation Θ~i1(xj1xjd)=0\widetilde{\Theta}^{i-1}(x_{j_{1}}\cdots x_{j_{d}})=0 with 11 in ii spots and summing up, we get back to the equation Θ~i(xj1xjd)=0\widetilde{\Theta}^{i}(x_{j_{1}}\cdots x_{j_{d}})=0.

Now we are ready to prove injectivity of ψ\psi. Suppose that akerψ{0}a\in\ker\psi\setminus\{0\}. Put an order on the indexing set JJ of the basis for LL. Then by the Poincaré-Birkoff-Witt Theorem, we can uniquely write aa as

a=i=1nαixj1(i)e1(i)xjdi(i)edi(i)+α0a=\sum\limits_{i=1}^{n}\alpha_{i}x_{j^{(i)}_{1}}^{e^{(i)}_{1}}\cdots x_{j^{(i)}_{d_{i}}}^{e^{(i)}_{d_{i}}}+\alpha_{0}

where αiF\alpha_{i}\in F, αi0\alpha_{i}\neq 0 for i1i\geq 1, ej(i)1e^{(i)}_{j}\geq 1 for every i,ji,j, di1d_{i}\geq 1, and jk(i)<jk+1(i)j^{(i)}_{k}<j^{(i)}_{k+1}. By applying Θ~0\widetilde{\Theta}^{0}, we get rid of every term but α0\alpha_{0}, so we see that α0=0\alpha_{0}=0. Let

Di=k=1diek(i)D_{i}=\sum\limits_{k=1}^{d_{i}}e^{(i)}_{k}

be the degree of the ii-th monomial and let D=max{Di}1inD=\max\{D_{i}\}_{1\leq i\leq n}. Since we know that Θ\Theta is injective on LL, we get a contradiction if D=1D=1. Thus assume that D2D\geq 2.

Repeating the same process we did on monomials, take the equation Θ~D1(a)=0\widetilde{\Theta}^{D-1}(a)=0, tensor it with 1 in the DD possible spots and add up all the equations. Then we subtract that sum from the equation Θ~D(a)=0\widetilde{\Theta}^{D}(a)=0. For every ii such that Di<DD_{i}<D, the terms coming from the ii-th monomial will vanish in the new equation. Thus we are on left with the monomials of degree exactly DD. The equation we are left with is a linear combination of pure tensors of the form Θ(xj1)Θ(xjD)\Theta(x_{j_{1}})\otimes\cdots\otimes\Theta(x_{j_{D}}). Since our monomials are ordered, the only way to have such pure tensors are equal is if they come from the same original monomial. Therefore we have a non-trivial linear combination. However, {Θ(xj)}jJA\{\Theta(x_{j})\}_{j\in J}\subset A is linearly independent since Θ\Theta is injective, so all the weights in the linear combination must be 0. Since FF is characteristic, we get that αi=0\alpha_{i}=0 for every ii such that Di=DD_{i}=D. This is a contradiction so no such aa exists and kerψ=(0)\ker\psi=(0). ∎

Remark 5.2.1.

This equivalence of soficity is particularly useful when combined with Proposition 11.6 from [AP17] stating amenable algebras without zero divisors are sofic. A particular application is a simplification of Theorem 4.2 for Lie algebras over fields of characteristic 0. Indeed, if LL is such a Lie algebra, by Theorem 7 of [Smi76], U(L)U(L) is also of subexponential growth. From [Ele03], we see that U(L)U(L) is amenable algebra with no zero divisors. Therefore U(L)U(L), and thus LL, is sofic.

References

  • [AP17] Goulnara Arzhantseva and Liviu Paŭnescu. Linear sofic groups and algebras. Transactions of The American Mathematical Society, 369(4):2285–2310, April 2017.
  • [Ele03] Gábor Elek. The amenability of affine algebras. Journal of Algebra, 264:469–478, 2003.
  • [ES05] Gábor Elek and Endre Szabó. Hyperlinearity, essentially free actions and l2-invariants. the sofic property. Mathematische Annalen, 332(2):421–441, apr 2005.
  • [Gro99] Misha Gromov. Endomorphism of symbolic algebraic varieties. Journal of the European Mathematical Society, 1(2):109–197, 1999.
  • [Her61] Isreal Herstein. Lie and jordan structures in simple associative algebras. Bulletin of the American Mathematical Society, 67(6):517–531, 1961.
  • [Pes08] Vladimir Pestov. Hyperlinear and sofic groups: a brief guide. Bulletin of Symbolic Logic, 14(4):449–480, December 2008.
  • [Smi76] Martha K. Smith. Universal enveloping algebras with subexponential but not polynomially bounded growth. Proceedings of the American Mathematical Society, 60:22–24, October 1976.
  • [Wei00] Benjamin Weiss. Sofic groups and dynamical systems. The Indian Journal of Statistics, 62(3):350–359, 2000.
  • [WY86] Minoru Wakimoto and Hirofumi Yamada. The fock representations of the virasoro algebra and the hirota equations of the modified kp hierarchies. Hiroshima Math Journal, 16:427–441, 1986.