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

Some remarks on associated varieties of vertex operator superalgebras

Hao Li Department of Mathematics and Statistics, SUNY-Albany, Albany 12222, NY, USA [email protected]
Abstract.

We study several families of vertex operator superalgebras from a jet (super)scheme point of view. We provide new examples of vertex algebras which are ”chiralizations” of their Zhu’s Poisson algebras RVR_{V}. Our examples come from affine C(1)C_{\ell}^{(1)}-series vertex algebras (1\ell\geq 1), certain N=1N=1 superconformal vertex algebras, Feigin-Stoyanovsky principal subspaces, Feigin-Stoyanovsky type subspaces, graph vertex algebras WΓW_{\Gamma}, and extended Virasoro vertex algebra. We also give a counterexample to the chiralization property for the N=2N=2 superconformal vertex algebra of central charge 11.

1. Introduction

Beilinson, Feigin and Mazur [BFM91] first introduced the notions of singular support and lisse representation in order to study Virasoro (vertex) algebra. Arakawa later extended these notions to any finitely strongly generated, non-negatively graded vertex algebra VV. More precisely, via a canonical decreasing filtration {Fp(V)}\left\{F_{p}(V)\right\} introduced in [Li05], we can associate to VV a positively graded vertex Possion vertex algebra grF(V)gr^{F}(V). The spectrum of grF(V)gr^{F}(V) is called singular support of VV and is denoted by SS(V)SS(V). With respect to this filtration, V/F1(V)V/F_{1}(V) is the Zhu C2C_{2}-algebra RVR_{V}. The reduced spectrum XV=Specm(RV)X_{V}={\rm Specm}(R_{V}) is a Poisson variety which is called the associated variety of V.V. A large body of work has been devoted to descriptions of associated variety for various vertex operator algebras [Ara12, AM18b, AM18c, AM17]. Certainly the most prominent examples from this point of view are well-known lisse, or C2C_{2}-cofinite vertex algebras characterized by dim(XV)=0{\rm dim}(X_{V})=0. Arakawa and Kawasetsu relaxed this condition to quasi-lisse in [AK18] which requires that XVX_{V} has finitely many symplectic leaves. Associated varieties are important in the geometry of Higgs branches in 4d/2d dualities in physics [BLL+15].

According to [Ara12, Proposition 2.5.1], the embedding

RVgrF(V)R_{V}\hookrightarrow gr^{F}(V)

can be extended to a surjective homomorphism of vertex Poisson algebras

ψ:J(RV)grF(V)\psi:J_{\infty}(R_{V})\twoheadrightarrow gr^{F}(V)

where J(RV)J_{\infty}(R_{V}) is the (infinite) jet algebra of RV.R_{V}. The map ψ\psi induces an injection from the singular support into the jet scheme of the associated scheme of VV, X~V=Spec(RV)\widetilde{X}_{V}={\rm Spec}(R_{V}),

ϕ:SS(V)J(X~V).\phi:SS(V)\hookrightarrow J_{\infty}(\widetilde{X}_{V}).

In [AM18a], authors showed that ϕ\phi is an isomorphism as varieties if VV is quasi-lisse. It was shown in [vEH18] that if the map ψ\psi is an isomorphism, then one can compute Hochschild homology of the Zhu algebra via the chiral homology of elliptic curves. Proving that ψ\psi is an isomorphsim or finding the kernel of ψ\psi turns out to be subtle. In [AL18], authors provided several examples for which ψ\psi is not an isomorphism, including the 2\mathbb{Z}_{2}-orbifold of the rank one Heisenberg algebra. Finding the kernel of ψ\psi even for this example seems quite hard (see also [vEH20]).

For a vertex algebra VV where ψ\psi is an isomorphism we obtain a very interesting (and important) consequence

ch[V](τ)=HSq(J(RV)),{\rm ch}[V](\tau)=HS_{q}(J_{\infty}(R_{V})),

where the left-hand side is the graded dimension of VV and the right-hand side is the Hilbert series. The left-hand side has often combinatorial interpretations which in turn can provide a non-trivial information about the jet scheme.

This work is our modest attempt to try to generalize above notions to vertex superalgebra case. We first generalize the notion of associated variety to vertex superalgebras. Then we investigate the map ψ\psi in the cases of affine vertex algebras, rank one lattice vertex superalgebras including the simple N=2N=2 superconformal vertex algebra at level 1, Feigin-Stoyanovsky principal subspaces, Feigin-Stoyanovsky type subspaces, simple N=1N=1 vertex algebra associated with (2,4k)(2,4k)-minimal model and certain extended Virasoro vertex algebras. Along the way, we get some interesting character fomulas and the bases of vertex algebra. We provide an example which is simple N=2N=2 vertex algebra at level 1, where ψ\psi is not an isomorphism. Moreover we make a conjecture about its kernel. We end the paper with a brief glimpse at our plans for future research.

2. Definitions and Preliminary results

Definition 2.1.

Let VV be a superspace, i.e., a 2\mathbb{Z}_{2}-graded vector space. V=V0¯V1¯V=V_{\overline{0}}\oplus V_{\overline{1}} where {0¯,1¯}=2\{{\overline{0}},\overline{1}\}=\mathbb{Z}_{2}. If aVp(a)a\in V_{p(a)}, we say that the element aa has parity p(a)2p(a)\in\mathbb{Z}_{2}.

A field is a formal series of the form a(z)=na(n)zn1a(z)=\sum_{n\in\mathbb{Z}}a_{(n)}z^{-n-1} where ana_{n}\in End(V)(V) and for each vVv\in V one has

a(n)v=0a_{(n)}v=0

for n0n\gg 0.

We say that a field a(z)a(z) has parity p(a)2p(a)\in\mathbb{Z}_{2} if

a(n)VαVα+p(a)a_{(n)}V_{\alpha}\in V_{{\alpha}+p(a)}

for all α2\alpha\in\mathbb{Z}_{2}, n.n\in\mathbb{Z}.

A vertex superalgebra contains the following data: a vector space of states VV, the vacuum vector 𝟏V0¯,\mathbf{1}\in V_{\overline{0}}, derivation TT, and the state-field correspondence map

aY(a,z)=na(n)zn1,a\longmapsto Y(a,z)=\sum_{n\in\mathbb{Z}}a_{(n)}z^{-n-1},

satisfying the following axioms:

  • (translation coinvariance): [T,Y(a,z)]=Y(a,z)[T,Y(a,z)]=\partial Y(a,z).

  • (vacuum): Y(𝟏,z)=IdVY(\mathbf{1},z)=Id_{V}, Y(a,z)𝟏|z=0=a,Y(a,z)\mathbf{1}|_{z=0}=a,

  • (locality): (zw)NY(a,z)Y(b,w)=(1)p(a)p(b)(zw)NY(b,w)Y(a,z)(z-w)^{N}Y(a,z)Y(b,w)=(-1)^{p(a)p(b)}(z-w)^{N}Y(b,w)Y(a,z) for N0N\gg 0.

In particular, a vertex algebra VV is called supercommutative if a(n)=0a_{(n)}=0 for n0n\geq 0. It is well-known that the category of commutative vertex superalgebras is equivalent with the category of unital commutative associative superalgebra equipped with an even derivation.

We say a vertex algebra V{V} is generated by a subset 𝒰V\mathcal{U}\subset{V} if any element of V{V} can be written as a finite linear combination of terms of the form

b(i1)1b2(i2)bn(in)𝟏\displaystyle b^{1}_{(i_{1})}{b^{2}}_{(i_{2})}\ldots{b^{n}}_{(i_{n})}\mathbf{1}

for bk𝒰b^{k}\in\mathcal{U}, iki_{k}\in\mathbb{Z}, and n0n\geq 0. If every element of VV can be written with ik<0i_{k}<0, we write V=𝒰S{V}=\langle\mathcal{U}\rangle_{S} and say VV is strongly generated by 𝒰\mathcal{U}.

Example 2.2.

(see for instance [Zhe17]) Let 𝔤\mathfrak{g} be a finite dimensional Lie superalgebra with a nondegenerate even supersymmetric invariant bilinear form (,)(\cdot,\cdot). We can associate the affine Lie superalgebra 𝔤^\widehat{\mathfrak{g}} to the pair (𝔤,(,))(\mathfrak{g},(\cdot,\cdot)). Its universal vacuum representation of level kk, V𝔤^(k,0)V_{\widehat{\mathfrak{g}}}(k,0), is a vertex superalgebra. In particular, when 𝔤\mathfrak{g} is a simple Lie algebra, V𝔤^(k,0)V_{\widehat{\mathfrak{g}}}(k,0) has an unique maximal ideal I𝔤^(k,0).I_{\widehat{\mathfrak{g}}}(k,0). And L𝔤^(k,0)=V𝔤^(k,0)/I𝔤^(k,0)L_{\widehat{\mathfrak{g}}}(k,0)=V_{\widehat{\mathfrak{g}}}(k,0)/I_{\widehat{\mathfrak{g}}}(k,0) is also a vertex algebra.

Example 2.3.

[Kac98] To any nn dimensional superspace AA with a non-degenerate anti-supersymmetric bilinear form (,)(\cdot,\cdot), we can associate a Lie superalgebra CAC_{A}. If we fix a basis of AA:

{ϕ1,,ϕn},\left\{\phi^{1},\ldots,\phi^{n}\right\},

the free fermionic vertex algebra \mathcal{F} associated to AA, is a vertex superalgebra strongly generated by ϕ(12)i𝟏\phi^{i}_{(-\frac{1}{2})}\mathbf{1} (1in)(1\leq i\leq n) where Y(ϕ(12)i𝟏,z)=n12+ϕ(n)izn12.Y(\phi^{i}_{(-\frac{1}{2})}\mathbf{1},z)=\displaystyle\sum_{n\in\frac{1}{2}+\mathbb{Z}}\phi_{(n)}^{i}z^{-n-\frac{1}{2}}.

Definition 2.4.

A vertex superalgebra VV is called a vertex operator superalgebra if it is 12\frac{1}{2}\mathbb{Z}-graded,

V=n12V(m),V=\coprod_{n\in\frac{1}{2}\mathbb{Z}}V_{(m)},

with a conformal vector ω\omega such that the set of operators {L(n),idV}n\left\{L_{(n)},id_{V}\right\}_{n\in\mathbb{Z}} with L(n)=ω(n+1)L_{(n)}=\omega_{(n+1)} defines a representation of Virasoro algebra on VV; that is

[L(n),L(m)]=(mn)L(m+n)+m3m12δm+n,0cV[L_{(n)},L_{(m)}]=(m-n)L_{(m+n)}+\frac{m^{3}-m}{12}\delta_{m+n,0}c_{V}

for m,nm,n\in\mathbb{Z}. We call cVc_{V} the central charge of VV. We require that L(0)L_{(0)} is diagonalizible and it defines the 12\frac{1}{2}\mathbb{Z} grading - its eigenvalues are called (conformal) weights. In several examples we will encounter 12\frac{1}{2}\mathbb{Z}-graded vertex superalgebras without a conformal vector. For this reason, we define the character or graded dimension as

ch[V](q)=m12dim(V(m))qm.{\rm ch}[V](q)=\sum_{m\in\frac{1}{2}\mathbb{Z}}{\rm dim}(V_{(m)})q^{m}.

As we do not care about modularity here, we suppress the qc24q^{-\frac{c}{24}} factor and also view qq as a formal variable.

Example 2.5.

[LL12] We let VirVir denote the Virasoro Lie algebra. Then the universal VirVir-module VVir(c,0)V_{Vir}(c,0) has a natural vertex operator algebra with central charge cc.

Example 2.6.

[Kac98] The universal vertex superalgebra associated with the N=1N=1 Neveu-Schwarz Lie superalgebra will be denoted by VcN=1V_{c}^{N=1}, where cc is the central charge. It is a vertex operator superalgebra strongly generated by an odd vector G(32)𝟏G_{(-\frac{3}{2})}\mathbf{1} and the conformal vector L(2)𝟏L_{(-2)}\mathbf{1}.

Example 2.7.

[Kac98], The universal vertex superalgebra associated with the N=2N=2 superconformal Lie algebra will be denoted by VcN=2V_{c}^{N=2}. It is a vertex operator superalgebra strongly generated by two odd vectors G(32)+𝟏G^{+}_{(-\frac{3}{2})}\mathbf{1}, G(32)𝟏G^{-}_{(-\frac{3}{2})}\mathbf{1} and two even vectors L(2)𝟏L_{(-2)}\mathbf{1}, J(1)𝟏J_{(-1)}\mathbf{1}.

Definition 2.8.

A commutative vertex superalgebra VV is called a vertex Poisson superalgebra if it is equipped with a linear operation,

VHom(V,z1V[z1]),aY(a,z)=n0a(n)zn+1,V\rightarrow{\rm Hom}(V,z^{-1}V[z^{-1}]),\quad a\rightarrow Y_{-}(a,z)=\sum_{n\geq 0}a_{(n)}z^{-n+1},

such that

  • (Ta)n=na(n1)(Ta)_{n}=-na_{(n-1)},

  • a(n)b=j0(1)n+j+1(1)p(a)p(b)j!Tj(b(n+j)a),a_{(n)}b=\sum_{j\geq 0}(-1)^{n+j+1}\frac{(-1)^{p(a)p(b)}}{j!}T^{j}(b_{(n+j)}a),

  • [a(m),b(n)]=j0(mj)(a(j)b)(m+nj),[a_{(m)},b_{(n)}]=\sum_{j\geq 0}\binom{m}{j}(a_{(j)}b)_{(m+n-j)},

  • a(n)(bc)=(a(n)b)c+(1)p(a)p(b)b(a(n)c),a_{(n)}(b\cdot c)=(a_{(n)}b)\cdot c+(-1)^{p(a)p(b)}b\cdot(a_{(n)}c),

for a,b,cVa,b,c\in V and n,m0n,m\geq 0.

A vertex Lie superalgebra structure on VV is given by (V,Y,T)(V,Y_{-},T). So we can also say that a vertex Poisson superalgebra is a commutative vertex superalgebra equipped with a vertex Lie superalgebra structure. In fact, we can obtain a vertex Poisson superalgebra from any vertex superalgebra through standard filtration or Li’s filtration. Following [Li05], we can define a decreasing sequence of subspaces {Fn(V)}\left\{F_{n}(V)\right\} of the superalgebra VV, where for nn\in\mathbb{Z}, Fn(V)F_{n}(V) is linearly spanned by the vectors

u(1k1)(1)u(1kr)(r)𝟏u_{(-1-k_{1})}^{(1)}\ldots u_{(-1-k_{r})}^{(r)}\bf{1}

for r1r\geq 1, u(1),,u(r)V,u^{(1)},\ldots,u^{(r)}\in V, k1,,kr0k_{1},\ldots,k_{r}\geq 0 with k1++krn.k_{1}+\ldots+k_{r}\geq n. Then

V=F0(V)F1(V)V=F_{0}(V)\supset F_{1}(V)\supset\ldots

such that

u(n)vFr+sn1(V)foruFr(V),vFs(V),r,s,n,\displaystyle u_{(n)}v\in F_{r+s-n-1}(V)\quad{\rm for}\quad u\in F_{r}(V),v\in F_{s}(V),r,s\in\mathbb{N},n\in\mathbb{Z},
u(n)vFr+sn(V)foruFr(V),vFr(V),r,s,n.\displaystyle u_{(n)}v\in F_{r+s-n}(V)\quad{\rm for}\quad u\in F_{r}(V),v\in F_{r}(V),r,s,n\in\mathbb{N}.

The corresponding associated graded algebra grF(V)=n0Fn(V)/Fn+1(V)gr^{F}(V)=\coprod_{n\geq 0}F_{n}(V)/F_{n+1}(V) is a vertex Poisson superalgebra . Its vertex Lie superalgebra structure is given by:

T(u+Fr+1(V))=Tu+Fr+2(V)T(u+F_{r+1}(V))=Tu+F_{r+2}(V)
Y(u+Fr+1(V),z)(v+Fs+1(V))=n0(u(n)v+Fr+sn+1(V))zn1Y_{-}(u+F_{r+1}(V),z)(v+F_{s+1}(V))=\sum_{n\geq 0}(u_{(n)}v+F_{r+s-n+1}(V))z^{-n-1}

for uFr(z),vFs(z)u\in F_{r}(z),v\in F_{s}(z) with r,sr,s\in\mathbb{N}. For the standard filtration {Gn(V)}\left\{G_{n}(V)\right\}, we also have the associated graded vertex superalgebra grG(V)gr^{G}(V). In [Ara12, Proposition 2.6.1], T.Arakawa showed that

grF(V)grG(V)gr^{F}(V)\cong gr^{G}(V)

as vertex Poisson superalgebras. Thus we sometimes drop the upper index FF or GG for brevity.

According to [Li05], we know that

Fn(V)={u(1i)v|uV,i1,vFni(V)}.F_{n}(V)=\left\{u_{(-1-i)}v|u\in V,i\geq 1,v\in F_{n-i}(V)\right\}.

In particular, F0(V)/F1(V)=V/C2(V)=RVgrF(V)F_{0}(V)/F_{1}(V)=V/C_{2}(V)=R_{V}\subset gr^{F}(V) which is a Poisson superalgebra according to [Zhu96]. Its Poisson structure is given by

u¯v¯=u(1)v¯,\overline{u}\cdot\overline{v}=\overline{u_{(-1)}v},
{u¯,v¯}=u(0)v¯\left\{\overline{u},\overline{v}\right\}=\overline{u_{(0)}v}

for u,vVu,v\in V where u¯=u+C2(V)\overline{u}=u+C_{2}(V). It was shown in [Li05, Corallary 4.3] that grF(V)gr^{F}(V) is generated by RVR_{V} a differential algebra. We compute C2C_{2}-algebra for some simple examples first.

Example 2.9.

Following notation in Example 2.3, let \mathcal{F} be a free fermionic vertex superalgebra associated with an nn-dimensional superspace AA. Clearly, the C2C_{2}-algebra of \mathcal{F} is

R=[ϕ(12)1𝟏¯,,ϕ(12)n𝟏¯]R_{\mathcal{F}}=\mathbb{C}[\overline{\phi^{1}_{(-\frac{1}{2})}\mathbf{1}},\ldots,\overline{\phi^{n}_{(-\frac{1}{2})}\mathbf{1}}]

where ϕ(12)i𝟏¯\overline{\phi^{i}_{(-\frac{1}{2})}\mathbf{1}} is even (resp. odd) if ϕi\phi^{i} is even (resp. odd) in AA.

Example 2.10.

According to [FFL11], for a simple affine vertex algebras L𝔤^(k,0)L_{\widehat{\mathfrak{g}}}(k,0), kk\in\mathbb{N}, where 𝔤\mathfrak{g} is a simple Lie algebra, we have:

RL𝔤^(k,0)=[u(1)1𝟏,u(1)2𝟏,,u(1)n𝟏]/U(𝔤)((eθ)(1))k+1𝟏,R_{L_{\widehat{\mathfrak{g}}}(k,0)}=\mathbb{C}[u^{1}_{(-1)}\mathbf{1},u^{2}_{(-1)}\mathbf{1},\ldots,u^{n}_{(-1)}\mathbf{1}]/\langle U(\mathfrak{g})\circ((e_{\theta})_{(-1)})^{k+1}\mathbf{1}\rangle,

where {u1,u2,un}\left\{u^{1},u^{2}\ldots,u^{n}\right\} is a basis of 𝔤\mathfrak{g}, θ\theta is the highest root of 𝔤\mathfrak{g} and \circ represents the adjoint action. In particular, when 𝔤=sl(2)\mathfrak{g}=sl(2),

RLsl2^(k,0)[e,f,h]/fiek+1|0i2k+2R_{L_{\widehat{{sl_{2}}}}(k,0)}\cong\mathbb{C}[e,f,h]/\langle f^{i}\circ e^{k+1}|0\leq i\leq 2k+2\rangle

where e,f,he,f,h correspond to e(1)𝟏,f(1)𝟏,h(1)𝟏e_{(-1)}\mathbf{1},f_{(-1)}\mathbf{1},h_{(-1)}\mathbf{1}.

Example 2.11.

For any simple Virasoro algebras LVir(c(p,p),0)L_{Vir}(c_{(p,p^{\prime})},0), where c(p,p)=16(pp)2ppc_{(p,p^{\prime})}=1-\frac{6(p-p^{\prime})^{2}}{pp^{\prime}} where p>p2p>p^{\prime}\geq 2 and p,pp,p^{\prime} are coprime, according to [BFM91, vEH18] its C2C_{2}-algebra is isomorphic to [x]/x(p1)(p1)2,\mathbb{C}[x]/\langle x^{\frac{(p-1)(p^{\prime}-1)}{2}}\rangle, where xx corresponds to ω=L(2)𝟏\omega=L_{(-2)}\mathbf{1}.

Example 2.12.

The C2C_{2}-algebra of VcN=1V_{c}^{N=1} is RVcN=1=[x,θ]R_{V_{c}^{N=1}}=\mathbb{C}[x,\theta] where xx and θ\theta correspond to even vector L(2)𝟏L_{(-2)}\mathbf{1} and odd vector G(32)𝟏G_{(-\frac{3}{2})}\mathbf{1}, respectively.

Example 2.13.

The C2C_{2}-algebra of VcN=2V_{c}^{N=2} is RVcN=2=[x,y,θ1,θ2]R_{V_{c}^{N=2}}=\mathbb{C}[x,y,\theta_{1},\theta_{2}] where x,y,θ1,θ2x,y,\theta_{1},\theta_{2} correspond to L(2)𝟏L_{(-2)}\mathbf{1}, J(1)𝟏J_{(-1)}\mathbf{1}, G(32)+𝟏G^{+}_{(-\frac{3}{2})}\mathbf{1} and G(32)𝟏,G^{-}_{(-\frac{3}{2})}\mathbf{1}, respectively. Here θ1,θ2\theta_{1},\theta_{2} are odd variables.

3. Affine jet superalgebra

Inspired by the definition of jet algebra, we may give an analogous definition of a jet superalgebra in the affine case. Here we closely follow [Ara12].

Let [x1,x2,,xn,θ1,,θm]\mathbb{C}[x^{1},x^{2},\ldots,x^{n},\theta^{1},\ldots,\theta^{m}] be a polynomial superalgebra where

x1,x2,,xnx^{1},x^{2},\ldots,x^{n}

are ordinary variables and

θ1,,θm\theta^{1},\ldots,\theta^{m}

are odd variables, i.e. (θi)2=0(\theta^{i})^{2}=0 for 1im.1\leq i\leq m. Let f1,f2,,fnf_{1},f_{2},\ldots,f_{n} be 2\mathbb{Z}_{2}-homogeneous elements in the polynomial superalgbera. We will define the jet superalgbra of the quotient superalgebra:

R=[x1,x2,,xn,θ1,,θm]f1,f2,,fr.\displaystyle R=\frac{\mathbb{C}[x^{1},x^{2},\ldots,x^{n},\theta^{1},\ldots,\theta^{m}]}{\langle f_{1},f_{2},\ldots,f_{r}\rangle}.

Firstly, let us introduce new even variables x(Δji)jx^{j}_{(-\Delta_{j}-i)} and odd variables θ(Δji)j\theta^{j^{\prime}}_{(-\Delta_{j^{\prime}}-i)} for i=0,,mi=0,\ldots,m where Δj\Delta_{j} and Δj\Delta_{j^{\prime}} are degrees of xjx^{j} and θj\theta^{j^{\prime}}. In most cases, we will assume that the degree of each variable is 11 although in some cases the odd degree can be shifted by 12\frac{1}{2}. We define an even derivation TT on

[x(Δji)j,θ(Δji))j| 0im, 1jn, 1jm]\mathbb{C}[x^{j}_{(-\Delta_{j}-i)},\theta^{j^{\prime}}_{(-\Delta_{j^{\prime}}-i))}\;|\;0\leq i\leq m,\;1\leq j\leq n,\;1\leq j^{\prime}\leq m]

as

T(x(Δji)j)={(Δji)x(Δji1))jfor im10for i=m,T(x^{j}_{(-\Delta_{j}-i)})=\begin{cases}(-\Delta_{j}-i)x^{j}_{(-\Delta_{j}-i-1))}&\text{for $i\leq m-1$}\\ 0&\text{for $i=m$},\\ \end{cases}

and

T(θ(Δji)j)={(Δji)θ(Δji1))jfor im10for i=m.T(\theta^{j^{\prime}}_{(-\Delta_{j^{\prime}}-i)})=\begin{cases}(-\Delta_{j^{\prime}}-i)\theta^{j^{\prime}}_{(-\Delta_{j^{\prime}}-i-1))}&\text{for $i\leq m-1$}\\ 0&\text{for $i=m$}.\\ \end{cases}

Here we identify xjx^{j} and θj\theta^{j^{\prime}} with x(Δj)jx^{j}_{(-\Delta_{j})} and θ(Δj)j,\theta^{j^{\prime}}_{(-\Delta_{j^{\prime}})}, respectively. Set

Rm=[x(Δji)j,θ(Δji))j| 0im, 1jn, 1jm]Tjfi|i=1,n,j.\displaystyle R_{m}=\frac{\mathbb{C}[x^{j}_{(-\Delta_{j}-i)},\theta^{j}_{(-\Delta_{j^{\prime}}-i))}\;|\;0\leq i\leq m,\;1\leq j\leq n,\;1\leq j^{\prime}\leq m]}{\langle T^{j}f_{i}|i=1,\ldots n,\;j\in\mathbb{N}\rangle}.

Then the mm-jet superscheme VmV_{m} is defined as Spec((Rm)0¯){\rm Spec}((R_{m})_{\overline{0}}) where (Rm)0¯(R_{m})_{\overline{0}} is the even part of the Rm.R_{m}. The infinite jet superalgebra of VV is

J(R)\displaystyle J_{\infty}(R) =lim𝑚Rm\displaystyle=\displaystyle\lim_{\underset{m}{\leftarrow}}R_{m}
=[x(Δji)j,θ(Δji))j| 0i, 1jn, 1jm]Tjfi|i=1,n,j.\displaystyle=\frac{\mathbb{C}[x^{j}_{(-\Delta_{j}-i)},\theta^{j^{\prime}}_{(-\Delta_{j^{\prime}}-i))}\;|\;0\leq i,\;1\leq j\leq n,\;1\leq j^{\prime}\leq m]}{\langle T^{j}f_{i}|i=1,\ldots n,\;j\in\mathbb{N}\rangle}.

We often omit ”infinite” and call it jet superalgebra for brevity. The jet superalgebra is a differential commutative superalgebra. We denote the ideal

Tjfi;i=1,,n,j0\langle T^{j}f_{i};i=1,\ldots,n,j\geq 0\rangle

by f1,,fn.\langle f_{1},\ldots,f_{n}\rangle_{\partial}. Later, we sometimes write x(j)x_{(j)} as x(j).x(j). The infinite jet superscheme, or arc space, is defined as

J(V)=Spec((J(R))0¯).J_{\infty}(V)={\rm Spec}((J_{\infty}(R))_{\overline{0}}).

We define the degree of each variable u(Δj)u_{(-\Delta-j)} to be Δ+j\Delta+j where u=xu=x or θ\theta. Then J(R)=12(J(R))(m)J_{\infty}(R)=\displaystyle\coprod_{\frac{1}{2}\mathbb{Z}}(J_{\infty}(R))_{(m)} where (J(R))(m)(J_{\infty}(R))_{(m)} is the set of all elements in jet superalgebra with degree mm. We define Hilbert series of J(R)J_{\infty}(R) as:

HSq(J(R))=m12dim(J(R)(m))qm.HS_{q}(J_{\infty}(R))=\sum_{m\in\frac{1}{2}\mathbb{Z}}{\rm dim}(J_{\infty}(R)_{(m)})q^{m}.

Following [Ara12], J(R)J_{\infty}(R) has a unique Poisson vertex superalgebra structure such that

u(n)v={{u,v},if n=00,if n>0u_{(n)}v=\begin{cases}\left\{u,v\right\},&\mbox{if }n=0\\ 0,&\mbox{if }n>0\end{cases}

for u,vRJ(R)u,v\in R\in J_{\infty}(R).

Furthermore, we can extend the embedding RVgrF(V)R_{V}\hookrightarrow gr^{F}(V) to a surjective differential superalgebra homomorphism J(RV)grF(V)J_{\infty}(R_{V})\twoheadrightarrow gr^{F}(V). It is obvious that the map is a differential superalgebra homomorphism. It is surjective since grF(V)gr^{F}(V) is generated by RVR_{V} as a differential algebra. Moreover, it was shown in [Ara12] that this map is actually a Poisson vertex superalgebra epimorphism. From now on, we call this map ψ.\psi. The map ψ\psi is not necessarily injective and it is an open problem to characterize rational vertex algebras for which ψ\psi is injective.

3.1. Complete lexicographic ordering

Following [FLK+01], we define the complete lexicographic ordering on a basis or spanning set of the jet superalgebra. Given a jet superalgebra

J([y1,y2,,yn]/I)=[y(Δ1i)1,,y(Δni)n|i]/I\displaystyle J_{\infty}(\mathbb{C}[y^{1},y^{2},\ldots,y^{n}]/I)=\mathbb{C}[y^{1}_{(-\Delta_{1}-i)},\dots,y^{n}_{(-\Delta_{n}-i)}|i\in\mathbb{N}]/I_{\infty}

where Δi\Delta_{i} is the degree of yi,y^{i}, we can first define an ordering of all variables in the following way:

y(Δ1)1<y(Δ1)2<<y(Δn)n<y(Δ11)1<y(Δ21)2<.y^{1}_{(-\Delta_{1})}<y^{2}_{(-\Delta_{1})}<\ldots<y^{n}_{(-\Delta_{n})}<y^{1}_{(-\Delta_{1}-1)}<y^{2}_{(-\Delta_{2}-1)}<\ldots.
Definition 3.1.

A monomial uu of J([y1,y2,,yn]/I)J_{\infty}(\mathbb{C}[y^{1},y^{2},\ldots,y^{n}]/I) is called an ordered monomial if it is of the form:

(y(Δnm)n)am+1n(y(Δ1m)1)am+11(y(Δn)n)a1n(y(Δ1)2)a12(y(Δ1)1)a11(y^{n}_{(-\Delta_{n}-m)})^{a^{n}_{m+1}}\ldots(y^{1}_{(-\Delta_{1}-m)})^{a^{1}_{m+1}}\ldots(y^{n}_{(-\Delta_{n})})^{a^{n}_{1}}\ldots(y^{2}_{(-\Delta_{1})})^{a^{2}_{1}}(y^{1}_{(-\Delta_{1})})^{a^{1}_{1}}

where m+m\in\mathbb{Z}_{+} and ajia^{i}_{j}\in\mathbb{N}.

It should be clear that all ordered monomials form a spanning set of the jet superalgebra. Then let us define the multiplicity of an ordered monomial as

μ(u)=i=1m+1(ai1+ai2++ain).\mu(u)=\sum_{i=1}^{m+1}(a^{1}_{i}+a^{2}_{i}+\ldots+a^{n}_{i}).

Given two arbitrary ordered monomials

u=y(Δnm)n)am+1n(y(Δ1m)1)am+11(y(Δn)n)a1n(y(Δ1)2)a12(y(Δ1)1)a11u=y^{n}_{(-\Delta_{n}-m)})^{a^{n}_{m+1}}\ldots(y^{1}_{(-\Delta_{1}-m)})^{a^{1}_{m+1}}\ldots(y^{n}_{(-\Delta_{n})})^{a^{n}_{1}}\ldots(y^{2}_{(-\Delta_{1})})^{a^{2}_{1}}(y^{1}_{(-\Delta_{1})})^{a^{1}_{1}}

and

v=y(Δnm)n)bm+1n(y(Δ1m)1)bm+11(y(Δn)n)b1n(y(Δ1)2)b12(y(Δ1)1)b11,v=y^{n}_{(-\Delta_{n}-m)})^{b^{n}_{m+1}}\ldots(y^{1}_{(-\Delta_{1}-m)})^{b^{1}_{m+1}}\ldots(y^{n}_{(-\Delta_{n})})^{b^{n}_{1}}\ldots(y^{2}_{(-\Delta_{1})})^{b^{2}_{1}}(y^{1}_{(-\Delta_{1})})^{b^{1}_{1}},

we define a complete lexicographic ordering as following: If μ(u)<μ(v)\mu(u)<\mu(v), we say that u<vu<v. If μ(u)=μ(v)\mu(u)=\mu(v), we compare exponents of

y(Δ1)1,y(Δ1)2,,y(Δn)n,,y(Δ1m)1,y(Δnm)ny^{1}_{(-\Delta_{1})},y^{2}_{(-\Delta_{1})},\ldots,y^{n}_{(-\Delta_{n})},\ldots,y^{1}_{(-\Delta_{1}-m)},y^{n}_{(-\Delta_{n}-m)}

in this order. Namely, we say v<uv<u if a11<b11a^{1}_{1}<b^{1}_{1}; if they are equal, we then compare a12a^{2}_{1} and b12b^{2}_{1}, and so on. Given a polynomial ff, we call the greatest monomial among all its terms with respect to the complete lexicographic ordering the leading term of ff.

4. Affine and lattice vertex algebras

In this section we analyze the Poisson (super)algebra RVR_{V} and the injectivity of the ψ\psi map for some familiar examples of affine and lattice vertex algebras.

Example 4.1.

It was shown in [Ara12, Proposition 2.7.1] that for any simple Lie algebra 𝔤\mathfrak{g}, we have J(RV𝔤^(k,0))grF(V𝔤^(k,0))J_{\infty}(R_{V_{\widehat{\mathfrak{g}}(k,0)}})\cong gr^{F}(V_{\widehat{\mathfrak{g}}(k,0)}).

Proposition 4.2.

For the free fermionic vertex superalgebra, J(R)grF()J_{\infty}(R_{\mathcal{F}})\cong gr^{F}(\mathcal{F}) as vertex Poisson superalgebras.

Proof.

We use T.Arakawa’s argument in [Ara12, Proposition 2.7.1]. We include the proof for completeness. Here we still follow notation from Example 2.3. According to [Kac98, Section 3.6], we can choose a conformal vector such that \mathcal{F} is 120\frac{1}{2}\mathbb{Z}_{\geq 0}-graded. We consider the standard filtration GG on FF. Firstly, we have U(A[t1]t1)\mathcal{F}\cong U(A[t^{-1}]t^{-1}) as super vector spaces. Moreover

Gn()={u(k1)1u(kr)r𝟏|ki120,r0,r2n}G^{n}(\mathcal{F})=\left\{u^{1}_{(-k_{1})}\ldots u^{r}_{(-k_{r})}\mathbf{1}\;|\;k_{i}\in\frac{1}{2}\mathbb{Z}_{\geq 0},\;r\geq 0,\frac{r}{2}\leq n\right\}

where ui{ϕ1,,ϕn}u^{i}\in\left\{\phi^{1},\ldots,\phi^{n}\right\}. So grG()S(A[t1]t1)J(R)gr^{G}(\mathcal{F})\cong S(A[t^{-1}]t^{-1})\cong J_{\infty}(R_{\mathcal{F}}) as Poissson vertex superalgebra. Therefore grG()grF()J(R)gr^{G}(\mathcal{F})\cong gr^{F}(\mathcal{F})\cong J_{\infty}(R_{\mathcal{F}}). ∎

Similarly, we can show that ψ\psi is an isomorphism for vertex superalgebra V𝔤^(k,0)V_{\widehat{{\mathfrak{g}}}}(k,0) where 𝔤\mathfrak{g} is a Lie superalgebra satisfying conditions in Example 2.2 , and for superconformal vertex algebras VcN=1V_{c}^{N=1} and VcN=2V_{c}^{N=2}.

Let

Vp=M(1)[p],V_{\sqrt{p}\mathbb{Z}}=M(1)\otimes\mathbb{C}[\sqrt{p}\mathbb{Z}],

be a rank one lattice vertex algebra (resp. superalgebra) constructed from an integral lattice L=αpL=\mathbb{Z}\alpha\cong\sqrt{p}\mathbb{Z} where α,α=p\langle\alpha,\alpha\rangle=p is even (resp. odd). It has a conformal vector ω=12pα(1)2𝟏\omega=\frac{1}{2p}\alpha_{(-1)}^{2}\mathbf{1}. As usual, we denote the extremal lattice vectors by enαe^{n\alpha}, nn\in\mathbb{Z}.

Proposition 4.3.

For the lattice vertex algebra VpV_{\sqrt{p}\mathbb{Z}} we have

RVp[x,y,z]/x2,y2,xy=zp,xz,yz.R_{V_{\sqrt{p}\mathbb{Z}}}\cong\mathbb{C}[x,y,z]/\langle x^{2},y^{2},xy=z^{p},xz,yz\rangle.

When pp is odd, xx and yy are odd vectors.

Proof.

According to the following calculations,

(eα)(2)(eα)(b(1))p+11(p+1)!RVp\displaystyle(e^{\alpha})_{(-2)}(e^{-\alpha})-\frac{(b_{(-1)})^{p+1}\textbf{1}}{(p+1)!}\in R_{V_{\sqrt{p}\mathbb{Z}}}
(eα)(2)(1)=b(1)eα\displaystyle(e^{\alpha})_{(-2)}(\textbf{1})=b_{(-1)}e^{\alpha}
(eα)(p1)(eα)=e2α\displaystyle(e^{\alpha})_{(-p-1)}(e^{\alpha})=e^{2\alpha}
(eα)(2)(1)=b(1)eα\displaystyle(e^{-\alpha})_{(-2)}(\textbf{1})=-b_{(-1)}e^{-\alpha}
(eα)(p1)(eα)=2e2α,\displaystyle(e^{-\alpha})_{(-p-1)}(e^{-\alpha})=2e^{-2\alpha},

we know that all vectors except for b(1)1b_{(-1)}\textbf{1}, …, b(1)p1b_{(-1)}^{p}\textbf{1}, eαe^{\alpha}, eαe^{-\alpha} and 1 are zero in RVpR_{V_{\sqrt{p}\mathbb{Z}}}.

Then we will show that all those vectors are indeed nonzero in RVpR_{V_{\sqrt{p}\mathbb{Z}}}. Suppose there exist a,bVpa,b\in V_{\sqrt{p}\mathbb{Z}} such that

a(n)beαRVpa_{(-n)}b-e^{\alpha}\in R_{V_{\sqrt{p}\mathbb{Z}}}

where n2-n\geq 2. Then wt(anb)=wt(a)+wt(b)n1=p2wt(a_{n}b)=wt(a)+wt(b)-n-1=\frac{p}{2} which implies that a,bπ0a,b\in\pi_{{0}} where π0\pi_{0} is the Heisenberg subalgebra [α(n)]n<0𝟏.\mathbb{C}[\alpha_{(n)}]_{n<0}\cdot\mathbf{1}. This is a contradiction. So the equivalent class eα¯\overline{e^{\alpha}} is nonzero in RVpR_{V_{\sqrt{p}\mathbb{Z}}}. By using similar weight argument, we can show that equivalence classes

eα¯,1¯,b(1)1¯,,b(1)p1¯\overline{e^{-\alpha}},\overline{\textbf{1}},\overline{b_{(-1)}\textbf{1}},\ldots,\overline{b_{(-1)}^{p}\textbf{1}}

are all nonzero in RVpR_{V_{\sqrt{p}\mathbb{Z}}}. Moreover, we have

(eα)(1)(eα)b(1)p1p!RVp.(e^{\alpha})_{(-1)}(e^{-\alpha})-\frac{b_{(-1)}^{p}\textbf{1}}{p!}\in R_{V_{\sqrt{p}\mathbb{Z}}}.

Then the map ψ\psi is sending eα¯\overline{e^{\alpha}} to xx, eα¯\overline{e^{-\alpha}} to yy, 1¯\overline{\textbf{1}} to 1 and 1p!pb(1)1¯\sqrt[p]{\frac{1}{p!}}\overline{b_{(-1)}\textbf{1}} to zz induces an isomorphism of algebras.

Remark 4.4.

According to the Frenkel-Kac construction, we know that V2Lsl2^(1,0)V_{\sqrt{2}\mathbb{Z}}\cong L_{\widehat{sl_{2}}}(1,0). Following Proposition 4.3, we have RLsl2^(1,0)[e,f,h]/e2,f2,ef=h2,eh,fhR_{L_{\widehat{sl_{2}}}(1,0)}\cong\mathbb{C}[e,f,h]/\langle e^{2},f^{2},ef=h^{2},eh,fh\rangle.

Before we move on, let us briefly recall definition of the associative Zhu algebra. Given a vertex superalgebra V=n12VnV=\displaystyle\coprod_{n\in\frac{1}{2}\mathbb{Z}}V_{n} where V0¯=nVnV_{\overline{0}}=\coprod_{n\in\mathbb{Z}}V_{n} and V1¯=n+12Vn,V_{\overline{1}}=\coprod_{n\in\mathbb{Z}+\frac{1}{2}}V_{n}, there are two binary operations defined as following: for homegeneous aVa\in V and bVb\in V,

ab={i0(wt(a)i)a(i1)b,if a,bV0¯0,if aorbV1¯a\ast b=\begin{cases}\displaystyle\sum_{i\geq 0}\binom{wt(a)}{i}a_{(i-1)}b,&\mbox{if }a,b\in V_{\overline{0}}\\ 0,&\mbox{if }a\;\text{or}\;b\in V_{\overline{1}}\end{cases}

and

ab={i0(wt(a)i)a(i2)b,if aV0¯i0(wt(a)12i)a(i1)b,if aV1¯.a\circ b=\begin{cases}\displaystyle\sum_{i\geq 0}\binom{wt(a)}{i}a_{(i-2)}b,&\mbox{if }a\in V_{\overline{0}}\\ \displaystyle\sum_{i\geq 0}\binom{wt(a)-\frac{1}{2}}{i}a_{(i-1)}b,&\mbox{if }a\in V_{\overline{1}}\end{cases}.

Let O(V)O(V) be the linear span of elements of the form aba\circ b in V.V. Then Zhu’s algebra A(V)A(V) is defined as the quotient space V/O(V)V/O(V) with the mutiplication from \ast. According to [Abe07], there is a filtration {F¯k(A(V))}\left\{\overline{F}_{k}(A(V))\right\} on A(V)A(V) where F¯k(A(V)):=(i=0kVi+O(V))/O(V)\overline{F}_{k}(A(V)):=(\displaystyle\bigoplus_{i=0}^{k}V_{i}+O(V))/O(V). Its associated graded algebra

grF¯(A(V))=i=0F¯k(A(V))/F¯k1(A(V))gr^{\overline{F}}(A(V))=\displaystyle\bigoplus_{i=0}^{\infty}\overline{F}_{k}(A(V))/\overline{F}_{k-1}(A(V))

is a commutative algebra. Now we can prove:

Corollary 4.5.

Let pp be a positive odd integer, then the even part of RVpR_{V_{\sqrt{p}\mathbb{Z}}}, i.e. (RVp)0¯(R_{V_{\sqrt{p}\mathbb{Z}}})_{\overline{0}}, is isomorphic to the associated graded algebra grF¯(AVp)gr^{\overline{F}}(A_{V_{\sqrt{p}\mathbb{Z}}}).

Proof.

According to [Oga00, Theorem 3.3], we know that AVp[x]/(Fp(x))A_{V_{\sqrt{p}\mathbb{Z}}}\cong\mathbb{C}[x]/(F_{p}(x)) where Fp(x)=x(x+1)(x1)(x+(p1)2)(x(p1)2)F_{p}(x)=x(x+1)(x-1)\ldots(x+\frac{(p-1)}{2})(x-\frac{(p-1)}{2}) in which xx corresponds to [α(1)𝟏][\alpha_{(-1)}\mathbf{1}] in AVpA_{V_{\sqrt{p}\mathbb{Z}}}. According to [Abe07], we have an epimorphism

f:RVpgrF¯(AVp)f:R_{V_{\sqrt{p}\mathbb{Z}}}\twoheadrightarrow gr^{\overline{F}}(A_{V_{\sqrt{p}\mathbb{Z}}})

given by f(a¯b¯)=[ab]+F¯k+l1(A(V))f(\overline{a}\cdot\overline{b})=[a\ast b]+\overline{F}^{k+l-1}(A(V)) for aVka\in V_{k} and bVlb\in V_{l}. Then according to Proposition 4.3, we have

(RVp)0¯grF¯(AVp)[x]/xp(R_{V_{\sqrt{p}\mathbb{Z}}})_{\overline{0}}\cong gr^{\overline{F}}(A_{V_{\sqrt{p}\mathbb{Z}}})\cong\mathbb{C}[x]/\langle x^{p}\rangle

via ff.

Remark 4.6.

If L=2kL=\sqrt{2k}\mathbb{Z} (k1)(k\geq 1) is an even lattice, above result is not true. According to [DLM97], we have grF¯(V2k)[x]/x2k1gr^{\overline{F}}(V_{\sqrt{2k}\mathbb{Z}})\cong\mathbb{C}[x]/\langle x^{2k-1}\rangle which is obviously not isomorphic to RV2k.R_{V_{\sqrt{2k}\mathbb{Z}}}.

In [vEH18], authors proved the map ψ\psi is an isomorphism for Lsl2^(k,0)L_{\widehat{sl_{2}}}(k,0) by using a PBW-type basis of Lsl2^(k,0)L_{\widehat{sl_{2}}}(k,0) from [MP87] and Gröbner bases. In [Fei09], author essentially proved the same result by using a technique called ”degeneration procedure”. In the following, we briefly explain how his results proves isomorphism.

Proposition 4.7.

The map ψ:J(RLsl2^(k,0))grF(Lsl2^(k,0))\psi:J_{\infty}(R_{L_{\widehat{sl_{2}}}(k,0)})\cong gr^{F}(L_{\widehat{sl_{2}}}(k,0)) is an isomorphism of vertex Poisson algebras.

Proof.

Let us prove k=1k=1 case. It is clear that ψ(u(i))=u(i)𝟏¯\psi(u_{(-i)})=\overline{u_{(-i)}\mathbf{1}} for u{e,f,h}u\in\left\{e,f,h\right\} and i>0i>0. Let u(z)=n1u(n)zn1u(z)=\sum_{n\leq-1}u_{(n)}z^{-n-1} where u{e,f,h}u\in\left\{e,f,h\right\}. Now we consider e(z)e(z)e(z)e(z). The coefficient of znz^{n} equals Tn(e(1)e(1))T^{n}(e_{(-1)}e_{(-1)}) up to a scalar multiple for n0.n\geq 0. And we have similar results for f2f^{2}, ef=h2ef=h^{2}, eheh and fh.fh. Thus

J(RLsl2^(1,0))[e(1i),f(1i),h(1i)|i]e(z)2,f(z)2,e(z)f(z)=h(z)2,e(z)h(z),f(z)h(z)\displaystyle J_{\infty}(R_{L_{\widehat{sl_{2}}}(1,0)})\cong\frac{\mathbb{C}[e_{(-1-i)},f_{(-1-i)},h_{(-1-i)}\;|\;i\in\mathbb{N}]}{\langle e(z)^{2},f(z)^{2},e(z)f(z)=h(z)^{2},e(z)h(z),f(z)h(z)\rangle}

where

e(z)2,f(z)2,e(z)f(z)=h(z)2,e(z)h(z),f(z)h(z)\langle e(z)^{2},f(z)^{2},e(z)f(z)=h(z)^{2},e(z)h(z),f(z)h(z)\rangle

means the ideal generated by the Fourier coefficients of e(z)2,f(z)2,e(z)f(z)=h(z)2,e(z)h(z),f(z)h(z)e(z)^{2},f(z)^{2},e(z)f(z)=h(z)^{2},e(z)h(z),f(z)h(z). Our result now follows from the above argument and [Fei09, Corollary 2.3]. When k2k\geq 2, the result follows from the same argument and [Fei09, Theorem 3.1].

Before we prove next result, let us fix some notation first. We denote a simple finite-dimensional Lie algebra of type Cn,n2C_{n},n\geq 2 by 𝔤\mathfrak{g}. Here we assume that 𝔤\mathfrak{g} has a basis {bi|1i(2n+1)n}.\left\{b^{i}|1\leq i\leq(2n+1)n\right\}. Let θ\theta be the maximal root of 𝔤\mathfrak{g}, and xθx_{\theta} the corresponding maximal root vector. We let 𝔤^\widehat{\mathfrak{g}} be the affine Lie algebra associated with 𝔤\mathfrak{g} and its universal vacuum representation is V𝔤^(k,0)V_{\widehat{\mathfrak{g}}}(k,0) for k>0.k\in\mathbb{Z}_{>0}. Set

R=U(𝔤)(xθ)(1)k+11,R¯=Span{r(n)|rR,n}R=U(\mathfrak{g})\circ(x_{\theta})_{(-1)}^{k+1}\textbf{1},\quad\overline{R}=\mathbb{C}-{\rm Span}\left\{r_{(n)}|r\in R,\;n\in\mathbb{Z}\right\}

where \circ is the adjoint action. Then 𝔤^\widehat{\mathfrak{g}}-module V𝔤^(k,0)V_{\widehat{\mathfrak{g}}}(k,0) has a maximal submodule I𝔤^(k,0)I_{\widehat{\mathfrak{g}}}(k,0) generated by R¯𝟏.\overline{R}\cdot\bf{1}. And

L𝔤^(k,0)=V𝔤^(k,0)/I𝔤^(k,0).L_{\widehat{\mathfrak{g}}}(k,0)=V_{\widehat{\mathfrak{g}}}(k,0)/I_{\widehat{\mathfrak{g}}}(k,0).

Now we are ready to prove:

Theorem 4.8.

The map ψ\psi is an isomorphism for the affine vertex algebra L𝔤^(1,0).L_{\widehat{\mathfrak{g}}}(1,0).

Proof.

It is clear that the C2C_{2}-algebra of L𝔤^(1,0)L_{\widehat{\mathfrak{g}}}(1,0) is RL𝔤^(1,0)=S(𝔤)/U(𝔤)eθ2R_{L_{\widehat{\mathfrak{g}}}(1,0)}=S(\mathfrak{g})/\langle U(\mathfrak{g})\circ e_{\theta}^{2}\rangle where S(𝔤)S(\mathfrak{g}) is the symmetric algebra of 𝔤\mathfrak{g} and U(𝔤)U(\mathfrak{g}) is the universal enveloping algebra of 𝔤\mathfrak{g}. We denote the algebra

[b(j)i|j>0]/U(𝔤)eθ2(z)\mathbb{C}[b^{i}_{(-j)}|j>0]/\langle U(\mathfrak{g})\circ e_{\theta}^{2}(z)\rangle

by QQ where eθ(z)=n<0(eθ)(n)zn1e_{\theta}(z)=\sum_{n<0}(e_{\theta})_{(n)}z^{-n-1}. Following the similar argument in Proposition 4.7, we see that J(RL(Λ0))QJ_{\infty}(R_{L(\Lambda_{0})})\cong Q. In order to show that ψ\psi is an isomorphism, it is enough to prove that grF(L(Λ0))gr^{F}(L(\Lambda_{0})) and QQ have the same basis. Notice that

I=R¯[b(j)i|j>0]=U(𝔤)eθ2(z).\displaystyle I=\overline{R}\cap\mathbb{C}[b^{i}_{(-j)}|j>0]=\langle U(\mathfrak{g})\circ e_{\theta}^{2}(z)\rangle.

We can define an order on all monomials of [b(j)i|j>0]\mathbb{C}[b^{i}_{(-j)}|j>0] in the sense of [PŠ16, Section 8]. From the same paper, we know that every nonzero homogeneous polynomial [b(j)i|j>0]\mathbb{C}[b^{i}_{(-j)}|j>0] has a unique largest monomial. For an arbitrary nonzero polynomial uu, we define the leading term lt(u)lt(u) as the largest monomial of the nonzero homogeneous component of the smallest degree, which is unique. We denote all monomials in [b(j)i|j>0]\mathbb{C}[b^{i}_{(-j)}|j>0] by 𝒫.\mathcal{P}. We clearly have 𝒫\mathcal{P} as a spanning set of QQ. Since u=0u=0 in QQ if uIu\in I, the leading term lt(u)lt(u) equals the linear combination of other terms. Therefore 𝒫lt(U)\mathcal{P}\setminus\langle lt(U)\rangle is a smaller spanning set of QQ. And we denote it by \mathcal{RR}. Meanwhile according to [PŠ16, Theorem 11.3], we know that ψ()\psi(\mathcal{RR}) is a basis of gr(L(Λ0))gr(L(\Lambda_{0})). Together with the surjectivity of ψ\psi, we have that \mathcal{RR} is a basis of L𝔤^(1,0)L_{\widehat{\mathfrak{g}}}(1,0). Therefore ψ\psi is an isomorphism.

4.1. N=2N=2 vertex superalgebra at c=1c=1

In this section we study the simple N=2N=2 superconformal vertex algebra of central charge c=1c=1, denoted by L1N=2L_{1}^{N=2}. The odd lattice vertex algebra V3V_{\sqrt{3}\mathbb{Z}} is known to be isomorphic to L1N=2L_{1}^{N=2}. Here we identify 13b(1)𝟏\frac{1}{3}b_{(-1)}\mathbf{1} with J(1)𝟏J_{(-1)}\mathbf{1}, 13e±α\frac{1}{\sqrt{3}}e^{\pm\alpha} with G(±32)𝟏G_{(\pm\frac{3}{2})}\mathbf{1} and 16(b(1)b(1)𝟏)(𝟏)𝟏\frac{1}{6}(b_{(-1)}b_{(-1)}\bf{1})_{(-1)}\mathbf{1} with L(2)𝟏L_{(-2)}\mathbf{1}.

According to [Ada99, Ada01], the maximal submodule of V1N=2V_{1}^{N=2} is generated by

G(52)+G(32)+𝟏andG(52)G(32)𝟏.G^{+}_{(-\frac{5}{2})}G^{+}_{(-\frac{3}{2})}\mathbf{1}\quad{\rm and}\quad G^{-}_{(-\frac{5}{2})}G^{-}_{(-\frac{3}{2})}\mathbf{1}.

By identifying G+G^{+} with G(32)+𝟏G^{+}_{(-\frac{3}{2})}\mathbf{1}, GG^{-} with G(32)𝟏G^{-}_{(-\frac{3}{2})}\mathbf{1} and hh with J(1)𝟏,J_{(-1)}\mathbf{1}, we have

RL1N=2[G+,G,h]/(G+)2,(G)2,G+G=h3,G+h,Gh.R_{L^{N=2}_{1}}\cong\mathbb{C}[G^{+},G^{-},h]/\langle(G^{+})^{2},(G^{-})^{2},G^{+}G^{-}=h^{3},G^{+}h,G^{-}h\rangle.

For the jet superalgebra of

[G+,G,h]/(G+)2,(G)2,G+G=h3,G+h,Gh,\mathbb{C}[G^{+},G^{-},h]/\langle(G^{+})^{2},(G^{-})^{2},G^{+}G^{-}=h^{3},G^{+}h,G^{-}h\rangle,

we identify G+,G,hG^{+},G^{-},h with G+(32),G(32),h(1)G^{+}(-\frac{3}{2}),G^{-}(-\frac{3}{2}),h(-1). And

J(RV3)[G+(32i),G(32i),h(1i)|i](G+(z))2,(G(z))2,G+(z)G(z)=h(z)3,G+(z)h(z),G(z)h(z)\displaystyle J_{\infty}(R_{V_{\sqrt{3}\mathbb{Z}}})\cong\frac{\mathbb{C}[G^{+}(-\frac{3}{2}-i),G^{-}(-\frac{3}{2}-i),h(-1-i)|i\in\mathbb{N}]}{\langle(G^{+}(z))^{2},(G^{-}(z))^{2},G^{+}(z)G^{-}(z)=h(z)^{3},G^{+}(z)h(z),G^{-}(z)h(z)\rangle}

where G±(z)=n32G±(n)zn32G^{\pm}(z)=\sum_{n\leq-\frac{3}{2}}G^{\pm}(n)z^{-n-\frac{3}{2}}, h(z)=n1h(n)zn1.h(z)=\sum_{n\leq-1}h(n)z^{-n-1}. The map ψ\psi is not an isomorphism in this case because the images of nonzero elements

G+(52)G+(32)andG(52)G(32)G^{+}(-\frac{5}{2})G^{+}(-\frac{3}{2})\quad{\rm and}\quad G^{-}(-\frac{5}{2})G^{-}(-\frac{3}{2})

in the jet superalgebra under ψ\psi, i.e. G(52)+G(32)+𝟏G^{+}_{(-\frac{5}{2})}G^{+}_{(-\frac{3}{2})}\mathbf{1} and G(52)G(32)𝟏G^{-}_{(-\frac{5}{2})}G^{-}_{(-\frac{3}{2})}\mathbf{1}, are null vectors. Thus

a,b=Ti(G+(52)G+(32)),Ti(G(52)G(32))|i0ker(ψ)\langle a,b\rangle_{\partial}=\langle T^{i}(G^{+}(-\frac{5}{2})G^{+}(-\frac{3}{2})),T^{i}(G^{-}(-\frac{5}{2})G^{-}(-\frac{3}{2}))|i\geq 0\rangle\subset ker(\psi)

where a=G+(52)G+(32)a=G^{+}(-\frac{5}{2})G^{+}(-\frac{3}{2}) and b=G(52)G(32).b=G^{-}(-\frac{5}{2})G^{-}(-\frac{3}{2}).

Let us consider

J(RL1N=2)/a,b.J_{\infty}(R_{L^{N=2}_{1}})/\langle a,b\rangle_{\partial}.

We will write down a spanning set of J(RL1N=2)/a,b.J_{\infty}(R_{L^{N=2}_{1}})/\langle a,b\rangle_{\partial}. We let the ordered monomial be a monomial of the form

G(n12)anh(n)bnG+(n12)cnG(52)a2h(2)b2G+(52)c2\displaystyle{G^{-}(-n-\frac{1}{2})}^{a_{n}}h(-n)^{b_{n}}{G^{+}(-n-\frac{1}{2})}^{c_{n}}\ldots{G^{-}(-\frac{5}{2})}^{a_{2}}h(-2)^{b_{2}}{G^{+}(-\frac{5}{2})}^{c_{2}}
G(32)a1h(1)b1G+(32)c1.\displaystyle{G^{-}(-\frac{3}{2})}^{a_{1}}h(-1)^{b_{1}}{G^{+}(-\frac{3}{2})}^{c_{1}}.

Then we have a complete lexicographic ordering on the set of ordered monomials in the sense of Section 3.1. Now let us find the leading terms of the Fourier coefficients of

G+(z)G(z)=h3(z),G+(z)h(z),G(z)h(z),Ti(a),Ti(b).G^{+}(z)G^{-}(z)=h^{3}(z),\quad G^{+}(z)h(z),\quad G^{-}(z)h(z),\quad T^{i}(a),\quad T^{i}(b).
  • (a)

    Leading term of G+(z)h(z)G^{+}(z)h(z):

    • nn is even, the leading term of the coefficient of znz^{n} is

      h(2n2)G+(3+n2).h(\frac{-2-n}{2})G^{+}(-\frac{3+n}{2}).
    • nn is odd, the leading term of the coefficient of znz^{n} is

      G+(4n2)h(1n2).G^{+}(\frac{-4-n}{2})h(\frac{-1-n}{2}).
  • (b)

    Leading term of G(z)h(z)G^{-}(z)h(z):

    • nn is even, the leading term of the znz^{n}-th coefficient is

      G(3+n2)h(2n2).G^{-}(-\frac{3+n}{2})h(\frac{-2-n}{2}).
    • nn is odd, the leading term of the znz^{n}-th coefficient is

      h(3n2)G(2n2).h(\frac{-3-n}{2})G^{-}(\frac{-2-n}{2}).
  • (c)

    Leading term of G+(z)G(z)=h3(z)G^{+}(z)G^{-}(z)=h^{3}(z):

    • n=1n=1, the leading term of the coefficient of zz is

      h(1)h(1)h(1).h(-1)h(-1)h(-1).
    • nn is even, the leading term of the znz^{n}-th coefficient is

      G(3+n2)G+(3+n2).G^{-}(-\frac{3+n}{2})G^{+}(-\frac{3+n}{2}).
    • nn is odd, the leading term of the znz^{n}-th coefficient is

      G(n42)G+(n22).G^{-}(\frac{-n-4}{2})G^{+}(\frac{-n-2}{2}).
  • (d)

    Leading term of Tn(a)T^{n}(a) or Tn(b)T^{n}(b):

    • nn is even, the leading term of the coefficient of znz^{n} is

      G±(n52)G±(n12).G^{\pm}(\frac{-n-5}{2})G^{\pm}(\frac{-n-1}{2}).
    • nn is odd, the leading term of the coefficient of znz^{n} is

      G±(n42)G±(n22).G^{\pm}(\frac{-n-4}{2})G^{\pm}(\frac{-n-2}{2}).

Clearly all ordered monomials constitute a spanning set of J(RL1N=2)/a,bJ_{\infty}(R_{L_{1}^{N=2}})/\langle a,b\rangle_{\partial}. Since all polynomials we considered above equal zero in J(RL1N=2)/a,bJ_{\infty}(R_{L_{1}^{N=2}})/\langle a,b\rangle_{\partial}, the leading term of each can be written as a linear combination of all other terms. Thus if we want to get a ”smaller” spanning set, all above leading terms can not appear as segments of an ordered monomial. Therefore we can impose some difference conditions on ordered monomials by using these leading terms to get a new spanning set.

Definition 4.9.

We call an ordered monomial a GhGh- monomial, if it satisfy the following conditions:

  • (i)(i)

    Either bib_{i} or cic_{i} is 0 and either bib_{i} or ci+1c_{i+1} is 0.0.

  • (ii)(ii)

    Either aia_{i} or bib_{i} is 0 and either aia_{i} or bi+1b_{i+1} is 0.0.

  • (iii)(iii)

    ai+ci+ci+11a_{i}+c_{i}+c_{i+1}\leq 1 and b12b_{1}\leq 2, i2.i\geq 2.

  • (iv)(iv)

    ci+ci+2+ci+11c_{i}+c_{i+2}+c_{i+1}\leq 1,  ai+ai+2+ai+11a_{i}+a_{i+2}+a_{i+1}\leq 1.

Here constraints (i)(i)-(iv)(iv) are coming from leading terms in (a)(a)-(d)(d), respectively. Then we have the following:

Proposition 4.10.

GhGh-monomials form a spanning set of

A=J(RL1N=2)/a,b.A=J_{\infty}(R_{L_{1}^{N=2}})/\langle a,b\rangle_{\partial}.

Let us write down the first few terms of the Hilbert series of AA.

Example 4.11.

For i5i\leq 5, GhGh-monomials give us basis of AiA_{i}:

A1:h(1)\displaystyle A_{1}:h(-1)
A32:G+(32),G(32)\displaystyle A_{\frac{3}{2}}:G^{+}(-\frac{3}{2}),G^{-}(-\frac{3}{2})
A2:h(1)2,h(2)\displaystyle A_{2}:h({-1})^{2},h({-2})
A52:G+(52),G(52)\displaystyle A_{\frac{5}{2}}:G^{+}(-\frac{5}{2}),G^{-}(-\frac{5}{2})
A3:G(32)G+(32),h(1)h(2),h(3)\displaystyle A_{3}:G^{-}(-\frac{3}{2})G^{+}(-\frac{3}{2}),h({-1})h({-2}),h({-3})
A72:G(52)h(1),G+(72),G(72),G+(32)h(2)\displaystyle A_{\frac{7}{2}}:G^{-}({-\frac{5}{2}})h({-1}),G^{+}({-\frac{7}{2}}),G^{-}({-\frac{7}{2}}),G^{+}({-\frac{3}{2}})h({-2})
A4:G+(32)G(52),h(2)2,h(1)2h(2),h(3)h(1),h(4)\displaystyle A_{4}:G^{+}({-\frac{3}{2}})G^{-}({-\frac{5}{2}}),h({-2})^{2},h({-1})^{2}h({-2}),h({-3})h({-1}),h({-4})
A92:G(52)h(1)2,G+(92),G(92),\displaystyle A_{\frac{9}{2}}:G^{-}({-\frac{5}{2}})h({-1})^{2},G^{+}({-\frac{9}{2}}),G^{-}({-\frac{9}{2}}),
G(72)h(1),G+(72)h(1),h(3)G(32),h(3)G+(32)\displaystyle\ \ \ \ \ \ \ G^{-}({-\frac{7}{2}})h({-1}),G^{+}({-\frac{7}{2}})h({-1}),h({-3})G^{-}({-\frac{3}{2}}),h({-3})G^{+}({-\frac{3}{2}})
A5:G(32)G+(72),G+(32)G(72),h(1)h(4),\displaystyle A_{5}:G^{-}({-\frac{3}{2}})G^{+}({-\frac{7}{2}}),G^{+}({-\frac{3}{2}})G^{-}({-\frac{7}{2}}),h({-1})h({-4}),
h(1)h(2)2,h(1)2h(3),h(2)h(3),h(5).\displaystyle\ \ \ \ \ \ \ \ h({-1})h({-2})^{2},h({-1})^{2}h({-3}),h({-2})h({-3}),h({-5}).

We have HSq(A)=1+q+2q32+2q2+2q52+3q3+4q72+5q4+7q92+7q5+O(q112)HS_{q}({A})=1+q+2q^{\frac{3}{2}}+2q^{2}+2q^{\frac{5}{2}}+3q^{3}+4q^{\frac{7}{2}}+5q^{4}+7q^{\frac{9}{2}}+7q^{5}+O(q^{\frac{11}{2}}). Meanwhile

ch[L1N=2](q)=ch[V3](q)=nq32n2n1(1qn)\displaystyle{\rm ch}[L_{1}^{N=2}](q)={\rm ch}[V_{\sqrt{3}\mathbb{Z}}](q)=\frac{\sum_{n\in\mathbb{Z}}q^{\frac{3}{2}n^{2}}}{\prod_{n\geq 1}(1-q^{n})}
=1+q+2q32+2q2+2q52+3q3+4q72+5q4+6q92+7q5+O(q112).\displaystyle=1+q+2q^{\frac{3}{2}}+2q^{2}+2q^{\frac{5}{2}}+3q^{3}+4q^{\frac{7}{2}}+5q^{4}+6q^{\frac{9}{2}}+7q^{5}+O(q^{\frac{11}{2}}).

Since in degree 92\frac{9}{2} dimension of AA is bigger than the dimension of V3V_{\sqrt{3}\mathbb{Z}} by 11, the induced map

ψ¯:J(RL1N=2)/a,bgr(L1N=2)\overline{\psi}:J_{\infty}(R_{L^{N=2}_{1}})/\langle a,b\rangle_{\partial}\rightarrow gr(L_{1}^{N=2})

is not injective. It is not hard to see that the one dimensional kernel of ψ¯\overline{\psi} in degree 92\frac{9}{2} is spanned by

c=G(92)13h(3)G(32)G(72)h(1)+13G(52)h(1)2.c=G^{-}(-\frac{9}{2})-\frac{1}{3}h(-3)G^{-}(-\frac{3}{2})-G^{-}(-\frac{7}{2})h(-1)+\frac{1}{3}G^{-}(-\frac{5}{2})h(-1)^{2}.

We make the following conjecture:

Conjecture 4.12.

The induced map ψ^:J(RL1N=2)/a,b,cgr(L1N=2)\hat{\psi}:J_{\infty}(R_{L_{1}^{N=2}})/\langle a,b,c\rangle_{\partial}\rightarrow gr(L_{1}^{N=2}) is an isomorphism.

5. principal subspaces

Principal subspaces of affine vertex algebras (at least in a special case) were introduced by Feigin and Stoyanovsky [FS93] and further studied by several people; see [But12, BK19, CLM08, CLM06, FFJ+09, FFJ+09] and references therein. In [Pri94, Pri00], M.Primc studied Feigin-Stoyanovsky type subspaces which are analogs of principal subspaces but easier to analyze. They are further investigated for many integral levels and types [BPT16, Jer12, Tru09, Tru11]. Here we follow notation from [MP12] where principal subspaces are defined for general integral lattices (not necessarily positive definite). As in [MP12], we let VL=M(1)[L]V_{L}=M(1)\otimes\mathbb{C}[L] denote a lattice vertex algebra where n=rank(L)n={\rm rank}(L). We fix a \mathbb{Z}-basis ={α1,,αn}\mathcal{B}=\left\{\alpha_{1},...,\alpha_{n}\right\} of LL. Then the principal subspace associated to \mathcal{B} and LL, is defined as

WL():=eα1,,eαn,W_{L}(\mathcal{B}):=\langle e^{\alpha_{1}},...,e^{\alpha_{n}}\rangle,

that is the smallest vertex algebra that contains extremal vectors eαie^{\alpha_{i}}. Once \mathcal{B} is fixed, we shall drop the \mathcal{B} in the parenthesis and write WLW_{L} for brevity.

Let 𝔤\mathfrak{g} be a simple finite dimensional complex Lie algebra of type AA, DD or EE and let 𝔥\mathfrak{h} be a Cartan subalgebra of 𝔤\mathfrak{g}. We choose simple roots {α1,,αn}\left\{\alpha_{1},\ldots,\alpha_{n}\right\} and let Δ+\Delta^{+} denote the set of positive roots. Let (,)(\cdot,\cdot) be a rescaled Killing form on 𝔤\mathfrak{g} such that (αi,αi)=2(\alpha_{i},\alpha_{i})=2 for 1in1\leq i\leq n (as usual we identify 𝔥\mathfrak{h} and 𝔥\mathfrak{h}^{*} via the Killing form). Fundamental weights of 𝔤\mathfrak{g}, {ω1,,ωn}𝔥\left\{\omega_{1},\ldots,\omega_{n}\right\}\subset\mathfrak{h}^{*}, are defined by (ωi,αj)=δi,j(\omega_{i},\alpha_{j})=\delta_{i,j} (1i,jn)(1\leq i,j\leq n).

Let 𝔫+\mathfrak{n}_{+} be αΔ+xα\displaystyle\coprod_{\alpha\in\Delta^{+}}x_{\alpha}, where xαx_{\alpha} is the corresponding root vector and 𝔫+^=𝔫+[t,t1]\widehat{\mathfrak{n}_{+}}=\mathfrak{n}_{+}\otimes\mathbb{C}[t,t^{-1}] its affinization. For an affine vertex algebra L𝔤^(k,0)L_{\widehat{\mathfrak{g}}}(k,0), khk\neq-h^{\vee}, isomorphic to L(kΛ0)L(k\Lambda_{0}) module, we define the (FS)-principal subspace of simple 𝔤^\widehat{\mathfrak{g}}-module L𝔤^(k,0)L_{\widehat{\mathfrak{g}}}(k,0) as

WΛk,0:=U(𝔫^+)𝟙,W_{\Lambda_{k,0}}:=U(\widehat{\mathfrak{n}}_{+})\cdot\mathbb{1},

where 𝟙\mathbb{1} is the vacuum vector. It is easy to see that this is a vertex algebra (without conformal vector). For k=1k=1, we have WLWΛk,0W_{L}\cong W_{\Lambda_{k,0}} where LL is the root lattice spanned by simple roots.

We fix a fundamental weight ω=ωm\omega=\omega_{m} and set

Γ={αR|(ω,α)=1}\Gamma=\left\{\alpha\in R|(\omega,\alpha)=1\right\}

where RR is the root system of 𝔤\mathfrak{g} and 𝔤1:=αΓ\mathfrak{g}_{1}:=\coprod_{\alpha\in\Gamma} 𝔤α\mathfrak{g}_{\alpha} where 𝔤α\mathfrak{g}_{\alpha} is the root vector. This Lie algebra is commutative. We let 𝔤1[t,t1]\mathfrak{g}_{1}\otimes\mathbb{C}[t,t^{-1}] be 𝔤1^\widehat{\mathfrak{g}_{1}}. Then we can define the so-called Feigin-Stoyanovsky type subspace of L𝔤^(k,0)L_{\widehat{\mathfrak{g}}}(k,0) as

WΛk,0:=U(𝔤^1)vkΛ0.W^{\prime}_{\Lambda_{k,0}}:=U(\widehat{\mathfrak{g}}_{1})\cdot v_{k\Lambda_{0}}.

Unlike the FS subspace, this vertex subalgebra is commutative. We denote

Γ~={xγ(r)|γΓ,r>0}.\tilde{\Gamma}^{-}=\left\{x_{\gamma}(-r)|\gamma\in\Gamma,\;r>0\right\}.
Γ~={xγ(r)|γΓ,r}.\tilde{\Gamma}=\left\{x_{\gamma}(-r)|\gamma\in\Gamma,\;r\in\mathbb{Z}\right\}.

Notice that U(𝔤^1)[Γ~]U(\widehat{\mathfrak{g}}_{1})\cong\mathbb{C}[\tilde{\Gamma}]. Therefore we can identify the elements in WΛk,0W^{\prime}_{\Lambda_{k,0}} with the elements in [Γ~].\mathbb{C}[\tilde{\Gamma}^{-}]. For any element in WΛk,0W^{\prime}_{\Lambda_{k,0}},

v=xβ1(m1)xβl(ml),βiΓ,v=x_{\beta_{1}}(m_{1})\ldots x_{\beta_{l}}(m_{l}),\quad\beta_{i}\in\Gamma,

we define colored weight as

cwt(v)=i=1lβicwt(v)=\displaystyle\sum_{i=1}^{l}\beta_{i}

for later use.

5.1. Root lattices of ADE type

Following the notations in [CLM08], we can prove the following result.

Proposition 5.1.

For 𝔤=sl(2)\mathfrak{g}=sl(2), we have WΛk,0gr(WΛk,0)J([x]/xk+1)W_{\Lambda_{k,0}}\cong gr(W_{\Lambda_{k,0}})\cong J_{\infty}(\mathbb{C}[x]/\langle x^{k+1}\rangle) for k1k\geq 1

Proof.

It is clear that RWΛk,0=[x]/xk+1.R_{W_{\Lambda_{k,0}}}=\mathbb{C}[x]/\langle x^{k+1}\rangle. The result follows from Theorem 3.1 in [CLM08]. ∎

Remark 5.2.

When k=1k=1, WΛ1,0W_{\Lambda_{1,0}} is isomorphic to J([x]/x2).J_{\infty}(\mathbb{C}[x]/\langle x^{2}\rangle). By using different methods to calculate the Hilbert-Poincare series, see [BMS13] and [BGK18], one can derive the famous Rogers-Ramanujan identities.

For the rest of this subsection, we let LL be the An1A_{n-1} root lattice with the rescaled Killing form (,)(\cdot,\cdot) such that (α,α)=2(\alpha,\alpha)=2 for any root and the standard \mathbb{Z}-basis α1,,αn1\alpha_{1},...,\alpha_{n-1} of simple roots. We are going to prove that ψ\psi is an isomorphism for the principal subspace WLW_{L} corresponding to this basis. In the following, we will identify WLW_{L} and WΛ1,0W_{\Lambda_{1,0}}. Firstly we prove the following proposition:

Proposition 5.3.

Given elements α\alpha, β\beta, γ\gamma and τ\tau in lattice LL, we have

(1) (eα)(1)eβ=0,if(α,β)>0.\displaystyle(e^{\alpha})_{(-1)}e^{\beta}=0,\quad if\;(\alpha,\beta)>0.
(2) (eα)(1)eβ=ϵ(α,β)ϵ(γ,τ)(eγ)(1)eτ,if(α,β)=(γ,τ)andα+β=γ+τ.\displaystyle(e^{\alpha})_{(-1)}e^{\beta}=\frac{\epsilon(\alpha,\beta)}{\epsilon(\gamma,\tau)}(e^{\gamma})_{(-1)}e^{\tau},\quad if\;\;\;(\alpha,\beta)=(\gamma,\tau)\;\;\;\text{and}\;\;\;\;\alpha+\beta=\gamma+\tau.
Proof.

From the definition of vertex operators from [Kac98], we have

Y(eα,z)eβ=ϵ(α,β)z(α,β)Exp(nα(n)nzn)eα+β,Y(e^{\alpha},z)e^{\beta}=\epsilon(\alpha,\beta)z^{(\alpha,\beta)}{\rm Exp}\left(\sum_{n\in\mathbb{Z}_{-}}\frac{-\alpha_{(n)}}{n}z^{-n}\right)e^{\alpha+\beta},

where ϵ(α,β)\epsilon(\alpha,\beta) is a 22-cocycle constant. Thus

(eα)(1)eβ=Coeffz0(Y(eα,z)eβ)=0(e^{\alpha})_{(-1)}e^{\beta}=\text{Coeff}_{z^{0}}(Y(e^{\alpha},z)e^{\beta})=0

since the minimal power of zz above is greater than 0. The coefficients of z0z^{0} of Y(eα,z)eβY(e^{\alpha},z)e^{\beta} and Y(eγ,z)eτY(e^{\gamma},z)e^{\tau} are ϵ(α,β)eα+β\epsilon(\alpha,\beta)e^{\alpha+\beta} and ϵ(γ,τ)eγ+τ\epsilon(\gamma,\tau)e^{\gamma+\tau}. The identity (2)(2) follows from this fact and given condition. ∎

It is clear that all quotient relations in RLsln^(1,0)R_{L_{\widehat{sl_{n}}}(1,0)} come from (1)(1) and (2)(2). Thus RLsln^(1,0)[Ei,j|1i<jn]=RWL.R_{L_{\widehat{sl_{n}}}}(1,0)\cap\mathbb{C}[E_{i,j}|1\leq i<j\leq n]=R_{W_{L}}.

We be Ei,jE_{i,j} the (i,j)(i,j)-th elementary matrix. Therefore {Ei,j}1i<jn\left\{E_{i,j}\right\}_{1\leq i<j\leq n} is the set of all positive root vectors. It is not hard to see that the C2C_{2}-algebra RWLR_{W_{L}} equals [Ei,j|1i<jn]/I\mathbb{C}[E_{i,j}|1\leq i<j\leq n]/I where we denote the equivalence class of (Ei,j)(1)𝟏(E_{i,j})_{(-1)}\mathbf{1} by Ei,jE_{i,j}. In [FFL11, Corollary 2.7], (see also [FS93] for 𝔤=sl(3)\mathfrak{g}=sl(3)) authors have written down the graded decomposition of RLsln^(1,0).R_{L_{\widehat{sl_{n}}}(1,0)}. By restricting it to its principal subspace, we have

Proposition 5.4.

The C2C_{2}-algebra of WLW_{L} equals

[Ei,j|1i<jn]/σS2Ei1,jσ1Ei2,jσ2|j1>i2\mathbb{C}[E_{i,j}|1\leq i<j\leq n]/\langle\sum_{\sigma\in S_{2}}E_{i_{1},j_{\sigma_{1}}}E_{i_{2},j_{\sigma_{2}}}|j_{1}>i_{2}\rangle

where 1i1i2n1\leq i_{1}\leq i_{2}\leq n and 1j1j2n1\leq j_{1}\leq j_{2}\leq n.

Moreover we have the following combinatorial qq-identity which will be proven in a joint work with Milas [LM], where we also prove more general identities.

Theorem 5.5 (Li-Milas).

Let AA be the Cartan matrix ((αi,αj))1i,jn1((\alpha_{i},\alpha_{j}))_{1\leq i,j\leq n-1} of type An1A_{n-1}, n2n\geq 2, and

𝐧=(n1,2,.,nn1,n)=(ni,j)1i<jn.{\bf n}=(n_{1,2},....,n_{n-1,n})=(n_{i,j})_{1\leq i<j\leq n}.

Then we have

(3) 𝐧0n(n1)/2qB(𝐧)1i<jn(q)ni,j=𝐤=(k1,,kn1)n1q𝐤A𝐤(q)k1(q)k2(q)kn1,\displaystyle\sum_{{\bf n}\in\mathbb{N}_{\geq 0}^{n(n-1)/2}}\frac{{\displaystyle q^{B({\bf n})}}}{\displaystyle\prod_{1\leq i<j\leq n}(q)_{n_{i,j}}}=\displaystyle\sum_{\mathbf{k}=(k_{1},\ldots,k_{n-1})\in\mathbb{N}^{n-1}}\frac{q^{\mathbf{k}A\mathbf{k}^{\top}}}{(q)_{k_{1}}(q)_{k_{2}}\ldots(q)_{k_{n-1}}},

where

B(𝐧)=1i1<j1n1i2<j2n1i1i2n1j1j2nj1>i2ni1,j1ni2,j2B({\bf n})=\sum_{\begin{subarray}{c}1\leq i_{1}<j_{1}\leq n\\ 1\leq i_{2}<j_{2}\leq n\\ 1\leq i_{1}\leq i_{2}\leq n\\ 1\leq j_{1}\leq j_{2}\leq n\\ j_{1}>i_{2}\end{subarray}}n_{i_{1},j_{1}}n_{i_{2},j_{2}}
Example 5.6.

For sl4,sl_{4}, we have the following qq-series identity:

𝐧𝟎𝟔qn12+n22+n32+n42+n52+n62+n1n4+n1n6+n2n4+n2n5+n3n5+n3n6+n4n6+n5n6+a4a5(q)n1(q)n2(q)n3(q)n4(q)n5(q)n6\displaystyle\displaystyle\sum_{\bf{n}\in\mathbb{N}_{\geq 0}^{6}}\frac{q^{n_{1}^{2}+n_{2}^{2}+n_{3}^{2}+n_{4}^{2}+n_{5}^{2}+n_{6}^{2}+n_{1}n_{4}+n_{1}n_{6}+n_{2}n_{4}+n_{2}n_{5}+n_{3}n_{5}+n_{3}n_{6}+n_{4}n_{6}+n_{5}n_{6}+a_{4}a_{5}}}{(q)_{n_{1}}(q)_{n_{2}}(q)_{n_{3}}(q)_{n_{4}}(q)_{n_{5}}(q)_{n_{6}}}
=𝐤𝟎𝟑qk12k1k2+k22k2k3+k32(q)k1(q)k2(q)k3\displaystyle=\sum_{\bf{k}\in\mathbb{N}^{3}_{\geq 0}}\frac{q^{k_{1}^{2}-k_{1}k_{2}+k_{2}^{2}-k_{2}k_{3}+k_{3}^{2}}}{(q)_{k_{1}}(q)_{k_{2}}(q)_{k_{3}}}

where we use multiindices 𝐧=(n1,n2,,n6)\mathbf{n}=(n_{1},n_{2},\ldots,n_{6}) and 𝐤=(k1,k2,k3).\mathbf{k}=(k_{1},k_{2},k_{3}). By doing following replacement,

n1,2n1,n2,3n2,n3,4n3,n_{1,2}\leftrightarrow n_{1},\;\;n_{2,3}\leftrightarrow n_{2},\;\;n_{3,4}\leftrightarrow n_{3},
n1,3n4,n2,4n5,n1,4n6,n_{1,3}\leftrightarrow n_{4},\;\;n_{2,4}\leftrightarrow n_{5},\;\;n_{1,4}\leftrightarrow n_{6},

we recover the formula in Theorem 5.5.

Now we are ready to prove

Theorem 5.7.

The map ψ\psi is an isomorphism between J(RWL)J_{\infty}(R_{W_{L}}) and gr(WL).gr(W_{L}).

Proof.

From Proposition 5.4, we know that J(RWL)J_{\infty}(R_{W_{L}}) is isomorphic to

[Ei,j(n)|n1, 1i<jn]/σS2Ei1,jσ1(z)Ei2,jσ2(z)|j1>i2\mathbb{C}[E_{i,j}(n)|n\leq-1,\;1\leq i<j\leq n]/\langle\sum_{\sigma\in S_{2}}E_{i_{1},j_{\sigma_{1}}}(z)E_{i_{2},j_{\sigma_{2}}}(z)|j_{1}>i_{2}\rangle

where Ei,j(z)=n1Ei,j(n)zn1E_{i,j}(z)=\sum_{n\leq-1}E_{i,j}(n)z^{-n-1} and 1i1i2n1\leq i_{1}\leq i_{2}\leq n, 1j1j2n1\leq j_{1}\leq j_{2}\leq n. In order to simplify notation, we first order {Ei,j}1i<jn\left\{E_{i,j}\right\}_{1\leq i<j\leq n} as:

E1,2,E1,3,,E1,n,E2,3,,E2,n,,En1,n,\displaystyle E_{1,2},E_{1,3},\ldots,E_{1,n},E_{2,3},\ldots,E_{2,n},\ldots,E_{n-1,n},

and we denote this sequence by {Em}1mn(n1)2\left\{E_{m}\right\}_{1\leq m\leq\frac{n(n-1)}{2}} (i.e. E1=E1,2E_{1}=E_{1,2}, E2=E1,3E_{2}=E_{1,3} etc.). We then have a spanning set of jet algebra with each element of the form:

E1(n11)E1(n1k1)E2(n21)E2(n2k2)\displaystyle E_{1}(-n_{1}^{1})\ldots E_{1}(-n_{1}^{k_{1}})E_{2}(-n_{2}^{1})\ldots E_{2}(-n_{2}^{k_{2}})\dots

where 1nmkmnm11\leq n_{m}^{k_{m}}\leq\ldots\leq n_{m}^{1} for 1mn(n1)21\leq m\leq\frac{n(n-1)}{2}. Here ks=0k_{s}=0 when we don’t have terms involving EsE_{s}. Now we can reduce this spanning set by using quotient relations as following:

  • difference two condition at distance 1: If we have Em(z)2=0E_{m}(z)^{2}=0 in C2C_{2}-algebra, then we can impose a condition: nmpnmp+1+2n_{m}^{p}\geq n_{m}^{p+1}+2 (1pkm1)(1\leq p\leq k_{m}-1) on above spanning set.

  • boundary condition: If we have Es(z)Et(z)+=0E_{s}(z)E_{t}(z)+\ldots=0 (s<t)(s<t), we can impose a condition: nskskt+1.n_{s}^{k_{s}}\geq k_{t}+1.

Therefore we have a reduced spanning set which implies

HSq(J(RWL))𝐧0n(n1)/2qB(𝐧)1i<jn(q)ni,j.HS_{q}(J_{\infty}(R_{W_{L}}))\leq\sum_{{\bf n}\in\mathbb{N}_{\geq 0}^{n(n-1)/2}}\frac{{\displaystyle q^{B({\bf n})}}}{\displaystyle\prod_{1\leq i<j\leq n}(q)_{n_{i,j}}}.

And it is well-known that

ch[gr(WL)](q)=𝐤=(k1,,kn)nq𝐤A𝐤(q)k1(q)k2(q)kn.{\rm ch}[gr(W_{L})](q)=\displaystyle\sum_{\mathbf{k}=(k_{1},\ldots,k_{n})\in\mathbb{N}^{n}}\frac{q^{\mathbf{k}A\mathbf{k}^{\top}}}{(q)_{k_{1}}(q)_{k_{2}}\ldots(q)_{k_{n}}}.

Surjectivity of ψ\psi and identity (2)(2) together imply that ψ\psi is an isomorphism and the image of above spanning set under ψ\psi is a basis of WL.W_{L}.

Remark 5.8.

Using result in [MP12], we can write down a basis of WLW_{L} by using (eαi)(j)(e^{\alpha_{i}})_{(j)} where αi\alpha_{i} is a simple root of slnsl_{n} and jj can be greater than or equal to 0. If we want the subscript jj to be always less than 0,0, we have to include (eβ)(j)(e^{\beta})_{(j)} where β\beta is a positive root. It is clear Em=Eim,jmE_{m}=E_{i_{m},j_{m}} is a root vector of a positive root

βm:=αim+αim+1++αjm1.\beta_{m}:=\alpha_{i_{m}}+\alpha_{i_{m}+1}+\ldots+\alpha_{j_{m}-1}.

Above proposition gives us a new basis of WLW_{L},

(eβ1)(n11)(eβ1)(n1k1)(eβ2)(n21)(eβ2)(n2k2)(eβM)(nMkM)𝟏\displaystyle(e^{\beta_{1}})_{(-n_{1}^{1})}\ldots(e^{\beta_{1}})_{(-n_{1}^{k_{1}})}(e^{\beta_{2}})_{(-n_{2}^{1})}\ldots(e^{\beta_{2}})_{(-n_{2}^{k_{2}})}\ldots(e^{\beta_{M}})_{(-n_{M}^{k_{M}})}\bf{1}

where M=n(n1)2M=\frac{n(n-1)}{2}, nMkM1n_{M}^{k_{M}}\geq 1, nmpnmp+1+2n_{m}^{p}\geq n_{m}^{p+1}+2 (1pkm1)(1\leq p\leq k_{m}-1) and nskskt+1n_{s}^{k_{s}}\geq k_{t}+1 if 1s<tM1\leq s<t\leq M, it<jsjti_{t}<j_{s}\leq j_{t}.

5.2. Feigin-Stoyanovsky type subspaces

In this section, we consider Feigin-Stoyanovsky type subspaces of affine vertex algebra of type AnA_{n} at level 1.1. We first consider the special case when ω=ω1\omega=\omega_{1}. For any element of the AnA_{n} root lattice,

α=m1α1+m2α2++mnαn,\alpha=m_{1}\alpha_{1}+m_{2}\alpha_{2}+\ldots+m_{n}\alpha_{n},

we define a subspace of WΛ1,0W^{\prime}_{\Lambda_{1,0}} as

(WΛ1,0)α:={vWΛ1,0|cwt(v)=α}.(W^{\prime}_{\Lambda_{1,0}})^{{\alpha}}:=\left\{v\in W^{\prime}_{\Lambda_{1,0}}|\;cwt(v)=\alpha\right\}.

It is not hard to see that (WΛ1,0)α(W^{\prime}_{\Lambda_{1,0}})^{\alpha} is nontrivial if and only if m1m2mn0m_{1}\geq m_{2}\geq\ldots\geq m_{n}\geq 0. According to [Tru11, (3.8)], we have

ch[(WΛ1,0)α](q)=qi=1nmi2i=1n1mimi+1(q)mn(q)mn1mn(q)m1m2.{\rm ch}[(W^{\prime}_{\Lambda_{1,0}})^{{\alpha}}](q)=\frac{q^{\sum_{i=1}^{n}m_{i}^{2}-\sum_{i=1}^{n-1}m_{i}m_{i+1}}}{(q)_{m_{n}}(q)_{m_{n-1}-m_{n}}\ldots(q)_{m_{1}-m_{2}}}.

Then

ch[WΛ1,0](q)\displaystyle{\rm ch}[W^{\prime}_{\Lambda_{1,0}}](q) =0mnm1qi=1nmi2i=1n1mimi+1(q)mn(q)mn1mn(q)m1m2\displaystyle=\displaystyle\sum_{0\leq m_{n}\leq\ldots\leq m_{1}}\frac{q^{\sum_{i=1}^{n}m_{i}^{2}-\sum_{i=1}^{n-1}m_{i}m_{i+1}}}{(q)_{m_{n}}(q)_{m_{n-1}-m_{n}}\ldots(q)_{m_{1}-m_{2}}}
=(l1,,ln)nqinli2+1i<jnlilj(q)l1(q)l2(q)ln.\displaystyle=\sum_{(l_{1},\dots,l_{n})\in\mathbb{N}^{n}}\frac{q^{\sum_{i}^{n}l_{i}^{2}+\sum_{1\leq i<j\leq n}{l_{i}l_{j}}}}{(q)_{l_{1}}(q)_{l_{2}}\ldots(q)_{l_{n}}}.

Moreover, in this case,

Γ={β1:=α1,β2:=α1+α2,,βn:=α1++αn}.\Gamma=\left\{\beta_{1}:=\alpha_{1},\beta_{2}:=\alpha_{1}+\alpha_{2},\ldots,\beta_{n}:=\alpha_{1}+\ldots+\alpha_{n}\right\}.

Notice that

L=β1βnL=\mathbb{Z}\beta_{1}\oplus\ldots\oplus\mathbb{Z}\beta_{n}

is a lattice with basis {β1,,βn}\left\{\beta_{1},\ldots,\beta_{n}\right\}. Then we have

WLWΛ1,0.W_{L}\cong W^{\prime}_{\Lambda_{1,0}}.

It is not hard to see that

βi,βi=2,if 1in\displaystyle\langle\beta_{i},\beta_{i}\rangle=2,\quad if\;1\leq i\leq n
βi,βj=1,if 1ijn.\displaystyle\langle\beta_{i},\beta_{j}\rangle=1,\quad if\;1\leq i\neq j\leq n.

According to Proposition 5.3, we have that C2C_{2}-algebra of WLW_{L} is

[x1,,xn]/xixj|1ijn.\mathbb{C}[x_{1},\ldots,x_{n}]/\langle x_{i}x_{j}|1\leq i\leq j\leq n\rangle.

By similar argument in previous section, we get

HSq(J([x1,,xn]/xixj|1ijn)])=ch[WL](q)HS_{q}(J_{\infty}(\mathbb{C}[x_{1},\ldots,x_{n}]/\langle x_{i}x_{j}|1\leq i\leq j\leq n\rangle)])={\rm ch}[W_{{L}}](q)

which implies isomorphism between J(RWΛ1,0)J_{\infty}(R_{W^{\prime}_{\Lambda_{1,0}}}) and gr(WΛ1,0).gr(W^{\prime}_{\Lambda_{1,0}}). Similarly we can also prove the isomorphism in cases where ω=ωi\omega=\omega_{i}, 2in2\leq i\leq n by making use of [Tru11, (3.21)].

5.3. Principal subspaces and jet schemes from graphs

In this part we study principal subspaces and jet algebras coming from graphs. We begin from any graph GG with kk vertices and possibly with loops (and for simplicity we assume no double edges). We denote the vertices of GG by {v1,v2,,vk}\left\{v_{1},v_{2},\ldots,v_{k}\right\}. We denote by Γ:=Γ(G)\Gamma:=\Gamma(G) the (symmetric) incidence matrix of GG and by (L(Γ),,)(L(\Gamma),\langle\ ,\ \rangle) rank kk lattice with basis α1,,αk\alpha_{1},...,\alpha_{k}, such that αi,αj=(Γ)i,j\langle\alpha_{i},\alpha_{j}\rangle=(\Gamma)_{i,j}. The incidence matrix of the graph induces a quadratic form

Γ12Q(x1,,xk),\Gamma\rightarrow\frac{1}{2}Q(x_{1},...,x_{k}),

where

Q(x1,,xk)=i,j=1vivjE(G)kxixjQ(x_{1},...,x_{k})=\sum_{\begin{subarray}{c}i,j=1\\ v_{i}v_{j}\in E(G)\end{subarray}}^{k}x_{i}x_{j}

where we sum over all edges E(G)E(G). Out of monomials appearing in the sum we form the infinite jet scheme JX(Γ)J_{\infty}X(\Gamma) where

RΓ=[x1,,xk]/vivjE(G)xixj.R_{\Gamma}=\mathbb{C}[x_{1},\ldots,x_{k}]/\langle\cup_{v_{i}v_{j}\in E(G)}x_{i}x_{j}\rangle.

We let WL(Γ)VL(Γ)W_{L(\Gamma)}\subset V_{L(\Gamma)} be the principal subspace corresponding to {eαi}i=1k\{e^{\alpha_{i}}\}_{i=1}^{k} inside the lattice vertex algebra VL(Γ)V_{L(\Gamma)}. For simplicity we write WΓW_{\Gamma} for WL(Γ)W_{L(\Gamma)}.

Example 5.9.

Consider the graph \circ-\circ-\circ. Then Γ=[010101010]\Gamma=\left[\begin{array}[]{ccc}0&1&0\\ 1&0&1\\ 0&1&0\end{array}\right], and WΓ=eα1,eα2,eα3W_{\Gamma}=\langle e^{\alpha_{1}},e^{\alpha_{2}},e^{\alpha_{3}}\rangle where L=α1α2α3L=\mathbb{Z}\alpha_{1}\oplus\mathbb{Z}\alpha_{2}\oplus\mathbb{Z}\alpha_{3} with α1,α2=α2,α3=1\langle\alpha_{1},\alpha_{2}\rangle=\langle\alpha_{2},\alpha_{3}\rangle=1 (zero otherwise), RΓ=[x1,x2,x3]/(x1x2,x2x3),R_{\Gamma}=\mathbb{C}[x_{1},x_{2},x_{3}]/(x_{1}x_{2},x_{2}x_{3}), and Q(x1,x2,x3)=x1x2+x2x3Q(x_{1},x_{2},x_{3})=x_{1}x_{2}+x_{2}x_{3}.

Theorem 5.10.

If the bilinear form associated with Γ\Gamma is non-degenerate, that is Γ\Gamma is invertible, then there exists a unique conformal vector in lattice vertex algebra such that eigenvalue of L(0)L_{(0)} defines grading such that:

wt(eαi)=32ifαi,αi=1,wt(e^{\alpha_{i}})=\frac{3}{2}\quad\text{if}\quad\langle\alpha_{i},\alpha_{i}\rangle=1,
wt(eαi)=1ifαi,αi=0.wt(e^{\alpha_{i}})=1\quad\text{if}\quad\langle\alpha_{i},\alpha_{i}\rangle=0.

Moreover, the graded dimension is given by:

ch[WΓ](q)=n1,,nk0qn1+n2++nk+12Q(n1,,nk)(q)n1(q)nk.{\rm ch}[W_{\Gamma}](q)=\sum_{n_{1},\ldots,n_{k}\geq 0}\frac{q^{n_{1}+n_{2}+\ldots+n_{k}+\frac{1}{2}Q(n_{1},\ldots,n_{k})}}{(q)_{n_{1}}\ldots(q)_{n_{k}}}.
Proof.

Clearly, we have the standard conformal vector in lattice vertex algebra given by ωst=12i=1nu(1)(i)u(1)(i)𝟏\omega_{st}=\frac{1}{2}\sum_{i=1}^{n}u^{(i)}_{(-1)}u^{(i)}_{(-1)}\mathbf{1} where u(1),,u(n)u^{(1)},\ldots,u^{(n)} is an orthonormal basis with respect to the bilinear form associated with Γ.\Gamma. We know that

Lst(0)(eαi)=αi,αi2.L_{st}(0)(e^{\alpha_{i}})=\frac{\langle\alpha_{i},\alpha_{i}\rangle}{2}.

It is clear that by adding a linear combination of {(αi)(2)𝟏}i=1n\left\{(\alpha_{i})_{(-2)}\mathbf{1}\right\}_{i=1}^{n}, we will still get a conformal vector. Now assume that ωst+i=1nai(αi)(2)𝟏\omega_{st}+\sum_{i=1}^{n}a_{i}(\alpha_{i})_{(-2)}\mathbf{1} where aia_{i}\in\mathbb{C} would give us expected weights. Then we have a system of linear equations. The non-degeneracy of the bilinear form implies that there is an unique solutions set. Thus we always have a conformal vector with the grading:

wt(eαi)=32ifαi,αi=1,wt(e^{\alpha_{i}})=\frac{3}{2}\quad\text{if}\quad\langle\alpha_{i},\alpha_{i}\rangle=1,
wt(eαi)=1ifαi,αi=0.wt(e^{\alpha_{i}})=1\quad\text{if}\quad\langle\alpha_{i},\alpha_{i}\rangle=0.

By applying [MP12, Corollary 4.14], we can write a combinatorial basis of WΓ.W_{\Gamma}. Now let us use this basis to write down the character. Firstly, the generating function of colored partition into (n1,n2,,nk)(n_{1},n_{2},\ldots,n_{k}) parts is 1(q)n1(q)nk.\frac{1}{(q)_{n_{1}}\ldots(q)_{n_{k}}}. It is clear that

ch[WΓ](q)=k1,,kk0qwt(f(n1,,nk))(q)n1(q)nk,{\rm ch}[W_{\Gamma}](q)=\sum_{k_{1},\ldots,k_{k}\geq 0}\frac{q^{wt(f_{(n_{1},\ldots,n_{k})})}}{(q)_{n_{1}}\ldots(q)_{n_{k}}},

where f(n1,,nk)f_{(n_{1},\ldots,n_{k})} is the vector in WΓW_{\Gamma} of charge (n1,,nk)(n_{1},\ldots,n_{k}) with the minimal weight. For the nin_{i} part, there is an unique element uniu_{n_{i}} of the minimal weight which is

e(1j=1i1αi,αjnj(ni1)αi,αi)αie(1j=1i1αi,αjnj)αi𝟏.e^{\alpha_{i}}_{(-1-\sum_{j=1}^{i-1}\langle\alpha_{i},\alpha_{j}\rangle n_{j}-(n_{i}-1)\langle\alpha_{i},\alpha_{i}\rangle)}\ldots e^{\alpha_{i}}_{(-1-\sum_{j=1}^{i-1}\langle\alpha_{i},\alpha_{j}\rangle n_{j})}\bf{1}.

The weight of uniu_{n_{i}} is

ni2(2(j=1i1αi,αjnj+wt((eαi)(1)1)+(ni1)αi,αi)\displaystyle\frac{n_{i}}{2}(2(\sum_{j=1}^{i-1}\langle\alpha_{i},\alpha_{j}\rangle n_{j}+wt((e^{\alpha_{i}})_{(-1)}\textbf{1})+(n_{i}-1)\langle\alpha_{i},\alpha_{i}\rangle)
=j=1i1αi,αjninj+ni22αi,αi+(αi,αi2+wt((eαi)(1)1))ni.\displaystyle=\sum_{j=1}^{i-1}\langle\alpha_{i},\alpha_{j}\rangle n_{i}n_{j}+\frac{n_{i}^{2}}{2}\langle\alpha_{i},\alpha_{i}\rangle+(-\frac{\langle\alpha_{i},\alpha_{i}\rangle}{2}+wt((e^{\alpha_{i}})_{(-1)}\textbf{1}))n_{i}.

Therefore

wt(f(n1,,nk))=i=1kwt(uni)\displaystyle wt(f_{(n_{1},\ldots,n_{k})})=\sum_{i=1}^{k}wt(u_{n_{i}})
=i=1kj=1i1αi,αjninj+ni22αi,αi+(αi,αi2+wt((eαi)(1)1))ni\displaystyle=\sum_{i=1}^{k}\sum_{j=1}^{i-1}\langle\alpha_{i},\alpha_{j}\rangle n_{i}n_{j}+\frac{n_{i}^{2}}{2}\langle\alpha_{i},\alpha_{i}\rangle+(-\frac{\langle\alpha_{i},\alpha_{i}\rangle}{2}+wt((e^{\alpha_{i}})_{(-1)}\textbf{1}))n_{i}
=n1+n2++nk+12Q(n1,,nk).\displaystyle=n_{1}+n_{2}+\ldots+n_{k}+\frac{1}{2}Q(n_{1},\ldots,n_{k}).

Thus we proved the claimed identity. ∎

Remark 5.11.

If the lattice LL is degenerate, then VLV_{L} has no conformal vector which can give us expected weights. But we can still view WLW_{L} as a graded vertex algebra if we define the degree of eαie^{\alpha_{i}} as above. Then the character formula is still valid for singular Γ\Gamma.

Before we prove next result, let us generalize [Pen14, Theorem 4.3.1].

Proposition 5.12.

We have an isomorphism

gr(WΓ)\displaystyle gr(W_{\Gamma})
[xi(p)|1ik,p1]m=0l1(m+αi,αj1)!m!αi,αjxi(αi,αjm)xj(l+m)|1i,jk,l1.\displaystyle\cong\frac{\mathbb{C}[x_{i}(p)|1\leq i\leq k,p\leq-1]}{\langle\sum_{m=0}^{-l-1}\frac{(m+\langle\alpha_{i},\alpha_{j}\rangle-1)!}{m!}\langle\alpha_{i},\alpha_{j}\rangle x_{i}({-\langle\alpha_{i},\alpha_{j}\rangle-m})x_{j}(l+m)|1\leq i,j\leq k,l\leq-1\rangle.}
Proof.

First, we define a map π\pi from

[xi(p)|1ik,p1]\displaystyle\mathbb{C}[x_{i}(p)|1\leq i\leq k,p\leq-1]

to gr(WΓ)gr(W_{\Gamma}) by sending xi(p)x_{i}(p) to e(p)αi𝟏e^{\alpha_{i}}_{(p)}\bf{1}. We denote the ideal

m=0l1(m+αi,αj1)!m!αi,αjxi(αi,αjm)xj(l+m)|1i,jk,l1\langle\sum_{m=0}^{-l-1}\frac{(m+\langle\alpha_{i},\alpha_{j}\rangle-1)!}{m!}\langle\alpha_{i},\alpha_{j}\rangle x_{i}(-\langle\alpha_{i},\alpha_{j}\rangle-m)x_{j}(l+m)|1\leq i,j\leq k,l\leq-1\rangle

by IΓI_{\Gamma}. Next we use an argument from [Pen14] to show that IΓker(π)I_{\Gamma}\subset ker(\pi).

We prove ker(π)IΓker(\pi)\subset I_{\Gamma} by contradiction. Suppose there exists an element a[xi(p)|1ik,p1]a\in\mathbb{C}[x_{i}(p)|1\leq i\leq k,p\leq-1] such that aker(π)a\in ker(\pi) and aIΓ.a\notin I_{\Gamma}. Suppose aa is homogeneous with respect to weight and charge. Choose rr such that aa contains some element xr(p)x_{r}(p) as a factor. We assume that aa has the minimum weight among all elements that satisfy above conditions. Again from the same argument from [Pen14], this aa can be written as uxr(1)ux_{r}(-1) where u[xi(p)|1ik,p1]u\in\mathbb{C}[x_{i}(p)|1\leq i\leq k,p\leq-1]. We prove the case when α,α=0.\langle\alpha,\alpha\rangle=0. For other cases, it is proved in [Pen14]. Firstly we define a map 𝐞α𝐫:WΓWΓ{\bf e^{\alpha_{r}}}:W_{\Gamma}\rightarrow W_{\Gamma} as

𝐞α𝐫((eαj)(m)1)=(eαj)(m)(eαr)(1)1.{\bf e^{\alpha_{r}}}((e^{\alpha_{j}})_{(m)}\textbf{1})=(e^{\alpha_{j}})_{(m)}(e^{\alpha_{r}})_{(-1)}\textbf{1}.

Then we lift this map to

𝐱𝐫:[xi(p)|1ik,p1][xi(p)|1ik,p1]{\bf x_{r}}:\mathbb{C}[x_{i}(p)|1\leq i\leq k,p\leq-1]\rightarrow\mathbb{C}[x_{i}(p)|1\leq i\leq k,p\leq-1]

which is defined as

𝐱𝐫(xi(j))=xi(j)xr(1).{\bf x_{r}}(x_{i}(j))=x_{i}(j)x_{r}(-1).

Since aker(π),a\in ker(\pi), π(a)=π(bxr(1))=π(b)(er)(1)𝟏=0.\pi(a)=\pi(bx_{r}(-1))=\pi(b)(e^{r})_{(-1)}\mathbf{1}=0. Then

𝐞α𝐫(π(b)(er)(1)1)=π(b)1=0{\bf{e^{\alpha_{r}}}}(\pi(b)(e^{r})_{(-1)}\textbf{1})=\pi(b)\textbf{1}=0

which implies that bker(π).b\in ker(\pi). If bIΓ,b\in I_{\Gamma}, then a=𝐱𝐫(b)𝐱𝐫IΓIΓa={\bf x_{r}}(b)\in{\bf x_{r}}I_{\Gamma}\in I_{\Gamma} which contradicts with our assumption. If bIΓ,b\notin I_{\Gamma}, then bb is an element such that bker(π)b\in ker(\pi) and bIΓb\notin I_{\Gamma} but with the weight strictly less than the weight of aa. This also contradicts our assumption. Thus we proved the claim. ∎

Theorem 5.13.

We have that

gr(WΓ)J([y1,y2,,yk]/αi,αjyiyj|1i,jk).gr(W_{\Gamma})\cong J_{\infty}(\mathbb{C}[y_{1},y_{2},\ldots,y_{k}]/\langle\langle\alpha_{i},\alpha_{j}\rangle y_{i}y_{j}|1\leq i,j\leq k\rangle).
Proof.

From the definition of jet superalgebra, we know that

T(l1)(αi,αjyiyj)=m=0l1cmlαi,αjyi(αi,αjm)yj(l+m)T^{(-l-1)}(\langle\alpha_{i},\alpha_{j}\rangle y_{i}y_{j})=\sum_{m=0}^{-l-1}c_{m}^{l}\langle\alpha_{i},\alpha_{j}\rangle y_{i}(-\langle\alpha_{i},\alpha_{j}\rangle-m)y_{j}(l+m)

where cmlc_{m}^{l} is a constant coefficient. Therefore

J([y1,y2,,yk]/αi,αjyiyj|1i,jk)J_{\infty}(\mathbb{C}[y_{1},y_{2},\ldots,y_{k}]/\langle\langle\alpha_{i},\alpha_{j}\rangle y_{i}y_{j}|1\leq i,j\leq k\rangle)

has quotient relation

m=0l1cmlαi,αjyi(αi,αjm)yj(l+m)|1i,jk,l1.\langle\sum_{m=0}^{-l-1}c_{m}^{l}\langle\alpha_{i},\alpha_{j}\rangle y_{i}(-\langle\alpha_{i},\alpha_{j}\rangle-m)y_{j}(l+m)|1\leq i,j\leq k,l\leq-1\rangle.

Together with Proposition 5.12, we get an isomorphism of differential algebras induced from the map ψ:xi(1)yi(1)\psi:x_{i}(-1)\rightarrow y_{i}(-1). ∎

When αi,αi=1\langle\alpha_{i},\alpha_{i}\rangle=1, we increase the degree of yi(1)y_{i}(-1) by 12\frac{1}{2}. Then clearly we have

HSq(J(RΓ))=ch[WΓ](q).HS_{q}(J_{\infty}(R_{\Gamma}))={\rm ch}[W_{\Gamma}](q).

5.4. Positive lattices

Given a lattice LL of rank nn with a \mathbb{Z}-basis {αi}i=1n\left\{\alpha_{i}\right\}_{i=1}^{n}, We say that the basis is positive if we have αi,αj0\langle\alpha_{i},\alpha_{j}\rangle\geq 0 for 1ijn1\leq i\leq j\leq n. In this part, we study principal subspaces associated with positive bases. Examples we studied in previous two sections are such principal subspaces. Now let us prove a more general result about the map ψ\psi and such principal subspaces.

Theorem 5.14.

For a lattice LL of rank nn with a positive basis, the map ψ\psi is an isomorphism for WLW_{L} if and only if its positive basis satisfies αi,αi=a\langle\alpha_{i},\alpha_{i}\rangle=a, where a=0a=0 or 11 or 22 and αi,αj=b\langle\alpha_{i},\alpha_{j}\rangle=b, where b=0b=0 or 11.

Proof.

First let us assume that the positive basis of the lattice LL satisfies given conditions. Notice that according to Theorem 5.13, we know that when αi,αi=a\langle\alpha_{i},\alpha_{i}\rangle=a, where a=0a=0 or 11 and αi,αj=b\langle\alpha_{i},\alpha_{j}\rangle=b, where b=0b=0 or 11, the map ψ\psi is an isomorphism for the principal subspace. Now the only case we need consider is the positive basis for which αi,αj=2δi,j.\langle\alpha_{i},\alpha_{j}\rangle=2\delta_{i,j}. It is not hard to see that J([x]/x2)J_{\infty}(\mathbb{C}[x]/x^{2}) has a basis

{x(m1)x(m2)x(mk)|mj1mj2,k0}.\left\{x_{(m_{1})}x_{(m_{2})}\dots x_{(m_{k})}|m_{j-1}\leq m_{j}-2,k\geq 0\right\}.

Thus J([x1,x2,,xn]/x12,x22,,xn2)J_{\infty}(\mathbb{C}[x_{1},x_{2},\ldots,x_{n}]/\langle x_{1}^{2},x_{2}^{2},\ldots,x_{n}^{2}\rangle) has a basis

{(xi1)(m11)(xi1)(m21)(xi1)(mk11)(xin)(m1n)(xin)(m2n)(xin)(mknn)\displaystyle\{(x_{i_{1}})_{(m^{1}_{1})}(x_{i_{1}})_{(m^{1}_{2})}\dots(x_{i_{1}})_{(m^{1}_{k_{1}})}\ldots(x_{i_{n}})_{(m^{n}_{1})}(x_{i_{n}})_{(m^{n}_{2})}\dots(x_{i_{n}})_{(m^{n}_{k_{n}})}
|mj1imji2, 1jki1}.\displaystyle|m^{i}_{j-1}\leq m^{i}_{j}-2,\;1\leq j\leq k_{i}-1\}.

Notice that the C2C_{2}-algebra of WLW_{L} is

[x1,,xn]/x12,,xn2\mathbb{C}[x_{1},\ldots,x_{n}]/\langle x_{1}^{2},\ldots,x_{{n}}^{2}\rangle

Now the map ψ\psi is sending (xi)(1)(x_{i})_{(-1)} to (eαi)(1)𝟏(e^{\alpha_{i}})_{(-1)}{\bf 1}. According to [MP12, Corollary 4.14], the image of the basis of J(RWL)J_{\infty}(R_{W_{L}}) is the basis of gr(WL)gr(W_{L}). Thus the map ψ\psi is an isomorphism.

Next, let us prove that if the basis does not satisfy given conditions, the map ψ\psi is not an isomorphism. We will consider two cases:

  • Suppose that for one simple root αi\alpha_{i}, we have αi,αi3.\langle\alpha_{i},\alpha_{i}\rangle\geq 3. Without loss generality, we prove that ψ\psi is not an isomorphism when lattice L=αi.L=\mathbb{Z}\alpha_{i}. In this case, from [MP12, Corollary 4.14], the basis of grFWLgr^{F}W_{L} is

    {(eαi)(m1)(eαi)(m2)(eαi)(mk)𝟏|mj1mjαi,αi,mk<0,k0}.\left\{(e^{\alpha_{i}})_{(m^{1})}(e^{\alpha_{i}})_{(m^{2})}\dots(e^{\alpha_{i}})_{(m^{k})}{\bf 1}|m_{j-1}\leq m_{j}-\langle\alpha_{i},\alpha_{i}\rangle,\;m_{k}<0,\;k\geq 0\right\}.

    It is easy to see that neither J([x]/(x2))J_{\infty}(\mathbb{C}[x]/(x^{2})) nor J([x])J_{\infty}(\bigwedge[x]) has the same corresponding basis (here \bigwedge denotes the exterior algebra).

  • Suppose that there exists two distinct roots αi,αj\alpha_{i},\alpha_{j} where i<ji<j such that αi,αj2\langle\alpha_{i},\alpha_{j}\rangle\geq 2. Without loss of generality, we assume L=αiαjL=\mathbb{Z}\alpha_{i}\oplus\mathbb{Z}\alpha_{j}, then the basis of J(WL)J_{\infty}(W_{L}) is

    {(xi)(1m1)(xi)(1m2)(xi)(1mk)(xj)(1n1)(xj)(1n2)(xj)(1nl)\displaystyle\{(x_{i})_{(-1-m_{1})}(x_{i})_{(-1-m_{2})}\ldots(x_{i})_{(-1-m_{k})}(x_{j})_{(-1-n_{1})}(x_{j})_{(-1-n_{2})}\ldots(x_{j})_{(-1-n_{l})}
    |m1m2αi,αi,n1n2αj,αj,mkl,nl0}.\displaystyle|m_{1}-m_{2}\geq\langle\alpha_{i},\alpha_{i}\rangle,\;n_{1}-n_{2}\geq\langle\alpha_{j},\alpha_{j}\rangle,\;m_{k}\geq l,n_{l}\geq 0\}.

    Meanwhile according to [MP12, Corollary 4.14], the image of this basis under ψ\psi strictly contains the basis of WLW_{L}. We do not have isomorphism.

Thus we proved the statement. ∎

5.5. New character formulas for ch[WΓ]{\rm ch}[W_{\Gamma}]

Here we continue from Section 5.3. If the graph Γ\Gamma is of Dynkin type AkA_{k} - path of length k1k-1) or CkC_{k} (cycle of length kk) we expect that the generating series HSq(J(RΓ))HS_{q}(J_{\infty}(R_{\Gamma})) has much better behaved combinatorial and perhaps even mock modular properties. We now present ”sum of tails” formulas for HSq(J(RAk))HS_{q}(J_{\infty}(R_{A_{k}})) for several low ”rank” cases. To simplify notation we let

Ak(q):=HSq(J(RAk)).A_{k}(q):=HS_{q}(J_{\infty}(R_{A_{k}})).

From Theorem 5.10 we have a fermionic formula

(4) Ak(q)=n1,n2,,nk0qn1+n2++nk+n1n2+n2n3++nk1nk(q)n1(q)n2(q)nk,A_{k}(q)=\sum_{n_{1},n_{2},\dotsc,n_{k}\geq 0}\frac{q^{n_{1}+n_{2}+\dotsb+n_{k}+n_{1}n_{2}+n_{2}n_{3}+\dotsb+n_{k-1}n_{k}}}{(q)_{n_{1}}(q)_{n_{2}}\dotsm(q)_{n_{k}}},

Next formulas are recently given by Jennings-Shaffer and Milas [JSM20].

Theorem 5.15.

We have

  • A2(q)=1(1q)(q)\displaystyle{A_{2}(q)=\frac{1}{(1-q)(q)_{\infty}}}

  • A3(q)=q1(1(q)21(q))\displaystyle{A_{3}(q)=q^{-1}\left(\frac{1}{(q)^{2}_{\infty}}-\frac{1}{(q)_{\infty}}\right)}

  • A4(q)=q1(q)2n1qn1qn\displaystyle{A_{4}(q)=\frac{q^{-1}}{(q)_{\infty}^{2}}\sum_{n\geq 1}\frac{q^{n}}{1-q^{n}}}

  • A5(q)=1(q)2n0qn(q)n(1qn+1)2.\displaystyle{A_{5}(q)=\frac{1}{(q)_{\infty}^{2}}\sum_{n\geq 0}\frac{q^{n}}{(q)_{n}(1-q^{n+1})^{2}}}.

  • A6(q)=1(q)2n,m0qn+m+nm(q)n+1(q)m+1.\displaystyle{A_{6}(q)=\frac{1}{(q)_{\infty}^{2}}\sum_{n,m\geq 0}\frac{q^{n+m+nm}}{(q)_{n+1}(q)_{m+1}}}.

Moreover, for cyclic graphs CkC_{k}-graphs we have fermionic formulas for Ck(q):=HSq(J(RCk))C_{k}(q):=HS_{q}(J_{\infty}(R_{C_{k}})) valid for k3k\geq 3

(5) Ck(q)=n1,n2,,nk0qn1+n2++nk+n1n2+n2n3++nk1nk+nkn1(q)n1(q)n2(q)nk.C_{k}(q)=\sum_{n_{1},n_{2},\dotsc,n_{k}\geq 0}\frac{q^{n_{1}+n_{2}+\dotsb+n_{k}+n_{1}n_{2}+n_{2}n_{3}+\dotsb+n_{k-1}n_{k}+n_{k}n_{1}}}{(q)_{n_{1}}(q)_{n_{2}}\dotsm(q)_{n_{k}}}.

Again we have partial results for ”bosonic” representations for 33- and 55-cycle graphs [JSM20].

Proposition 5.16.

We have

C3(q)=1(q)n0qn(qn+1)n+1.\displaystyle{C_{3}(q)=\frac{1}{(q)_{\infty}}\sum_{n\geq 0}\frac{q^{n}}{(q^{n+1})_{n+1}}}.

C5(q)=q1(q)2n1nqn1qn.\displaystyle{C_{5}(q)=\frac{q^{-1}}{(q)_{\infty}^{2}}\sum_{n\geq 1}\frac{nq^{n}}{1-q^{n}}}.

5.6. Combinatorial interpretation

Next we present combinatorial interpretations of formulas in Theorem 5.15 and Proposition 5.16. For simplicity, in several formulas we factored out a (power of) Euler factor which can be easily interpreted as the number of (colored) partitions.

Theorem 5.17.

We have:

  • A2(q)A_{2}(q) counts the number of partitions of 2n2n with all parts either even or equal to 1.

  • qA3(q)qA_{3}(q) counts the number of partitions of n+1n+1 into two kinds of parts with the first kind of parts used in each partition.

  • q(q)A4(q)q(q)_{\infty}A_{4}(q) counts the total number of parts in all partitions of nn, which is also sum of largest parts of all partitions of nn.

  • (q)2A5(q)(q)_{\infty}^{2}A_{5}(q) is the sum of the numbers of times that the largest part appears in each partition of nn.

  • q(q)2A6(q)q(q)_{\infty}^{2}A_{6}(q) counts twice the total number of parts in all partitions of nn minus the number of partitions of nn.

  • (q)C3(q)(q)_{\infty}C_{3}(q) counts the number of partitions of n such that twice the least part is bigger than the greatest part.

  • q(q)C5(q)q(q)_{\infty}C_{5}(q) counts the sum of all parts of all partitions of nn, also known as np(n){\rm np}(n).

Proof.

For A2(q)A_{2}(q), observe that CoeffqnA2(q)=p(1)+p(2)++p(n){\rm Coeff}_{q^{n}}A_{2}(q)=p(1)+p(2)+\cdots+p(n), where p(i)p(i) is the number of partitions of ii. The number of 1s1^{\prime}s must be even, say 2k2k, so we have to compute the number of partitions of 2n2k2n-2k where all parts are even. This is given by p(nk)p(n-k). Then summing over kk gives the claim.

The interpretation for the A3(q)A_{3}(q) series, is clear because we can also write

qA3(q)=1(q)(1(q)1).qA_{3}(q)=\frac{1}{(q)_{\infty}}\left(\frac{1}{(q)_{\infty}}-1\right).

Extracting the coefficient on the right-hand side gives p2(n)p(n)p_{2}(n)-p(n), where p2(i)p_{2}(i) denotes the number of two colored partitions.

For A4(q)A_{4}(q), this can be seen from identity n1qn1qn(q)=n1nqn(q)n,\frac{\sum_{n\geq 1}\frac{q^{n}}{1-q^{n}}}{(q)_{\infty}}=\sum_{n\geq 1}\frac{nq^{n}}{(q)_{n}}, which follows by taking the (xdd)(x\frac{d}{d}) derivative of 1(xq;q)=n0xnqn(q)n\frac{1}{(xq;q)_{\infty}}=\sum_{n\geq 0}\frac{x^{n}q^{n}}{(q)_{n}}. This clearly counts the total number of parts in all partitions of nn.

The (q)2A5(q)(q)_{\infty}^{2}A_{5}(q) case was already discussed in [JSM20].

For (q)2A6(q)(q)_{\infty}^{2}A_{6}(q), this follows from another identity given in [JSM20]:

1(q)2n,m0qn+m+nm(q)n+1(q)m+1=q1(q)2(2n1qn(1qn)(q)+11(q)),\frac{1}{(q)_{\infty}^{2}}\sum_{n,m\geq 0}\frac{q^{n+m+nm}}{(q)_{n+1}(q)_{m+1}}=\frac{q^{-1}}{(q)^{2}_{\infty}}\left(2\sum_{n\geq 1}\frac{q^{n}}{(1-q^{n})(q)_{\infty}}+1-\frac{1}{(q)_{\infty}}\right),

together with a previous observation that n1qn1qn(q)\frac{\sum_{n\geq 1}\frac{q^{n}}{1-q^{n}}}{(q)_{\infty}} counts the total number of parts in all partitions of nn.

For (q)C3(q)(q)_{\infty}C_{3}(q) we use a well-known interpretation for the fifth order mock theta function, and finally for (q)C5(q)(q)_{\infty}C_{5}(q) we observe the formula

(qddq)1(q)=1(q)n1nqn1qn=n1np(n)qn\left(q\frac{d}{dq}\right)\frac{1}{(q)_{\infty}}=\frac{1}{(q)_{\infty}}\sum_{n\geq 1}\frac{nq^{n}}{1-q^{n}}=\sum_{n\geq 1}np(n)q^{n}

as claimed.

Remark 5.18.

It is interesting to observe that the numerators of C3(q)C_{3}(q) and C5(q)C_{5}(q) are mock modular forms, and thus C3(q)C_{3}(q) and C5(q)C_{5}(q) are mixed mock. Completion of the Ramanujan fifth order mock theta n0qn(qn+1)n+1\sum_{n\geq 0}\frac{q^{n}}{(q^{n+1})_{n+1}} function is well-documented [BFOR17] . For n1nqn1qn\sum_{n\geq 1}\frac{nq^{n}}{1-q^{n}} we only have to observe that adding 124-\frac{1}{24} to the numerator gives E2(τ)E_{2}(\tau), the weight two quasimodular Eisenstein series, which is known to be mock.

6. N=1N=1 superconformal vertex algebra

In this section we consider the rational N=1N=1 vertex superalgebra Lc2,4kN=1L_{c_{2,4k}}^{N=1} (k1)(k\geq 1) associated to N=1N=1 superconformal (2,4k)(2,4k)-minimal models [Ada97]. Here the central charge is c2,4k=32(12(4k1)28k)c_{2,4k}=\frac{3}{2}(1-\frac{2(4k-1)^{2}}{8k}).

According to [Mel94, Mil07], we know that the normalized character of Lc2,4kN=1L_{c_{2,4k}}^{N=1} (without the qc/24q^{-c/24} factor) is:

ch[Lc2,4kN=1](q)\displaystyle{\rm ch}[L_{c_{2,4k}}^{N=1}](q) =n=1n2(mod 2)n0,±1(mod 4k)1(1qn2)\displaystyle=\displaystyle\prod^{\infty}_{\begin{subarray}{c}n=1\\ n\not\equiv 2(\text{mod}\;2)\\ n\not\equiv 0,\;\pm 1(\text{mod}\;4k)\end{subarray}}\frac{1}{(1-q^{\frac{n}{2}})}
=m1,,mk10(q12)N1q12N12+N22++Nk+12+N(s+1)/2+N(s+3)/2++Nk1(q)m1(q)m2(q)mk1.\displaystyle=\displaystyle\sum_{m_{1},\ldots,m_{k-1}\geq 0}\frac{(-q^{\frac{1}{2}})_{N_{1}}q^{\frac{1}{2}N_{1}^{2}+N_{2}^{2}+\ldots+N_{k+1}^{2}+N_{(s+1)/2}+N_{(s+3)/2}+\ldots+N_{k-1}}}{(q)_{m_{1}}(q)_{m_{2}}\ldots(q)_{m_{k-1}}}.

And the fermionic character formula is the generating function (cf. [Mel94])

ch[Lc2,4kN=1](q)=n=0Dk,1(n)qn2{\rm ch}[L_{c_{2,4k}}^{N=1}](q)=\sum_{n=0}^{\infty}D_{k,1}(n)q^{\frac{n}{2}}

of the number of partitions of Dk,1(n)D_{k,1}(n) of n2\frac{n}{2} in the form n2=b1++bm\frac{n}{2}=b_{1}+\ldots+b_{m} (bj121)(b_{j}\in\frac{1}{2}\mathbb{Z}_{\geq 1}) where b1,,bmb_{1},\ldots,b_{m} satisfy the following conditions:

  • no half-odd integer is repeated.

  • bjbj+1,b_{j}\geq b_{j+1}, bm32b_{m}\geq\frac{3}{2},

  • bjbj+k11b_{j}-b_{j+k-1}\geq 1 if bj+12,b_{j}\in\mathbb{Z}+\frac{1}{2},

  • bjbj+k1>1b_{j}-b_{j+k-1}>1 if bj.b_{j}\in\mathbb{Z}.

Since N=1N=1 vertex superalgebra Lc2,4N=1L_{c_{2,4}}^{N=1} is isomorphic to \mathbb{C}, we only need consider Lc2,4kN=1L_{c_{2,4k}}^{N=1} where k>1k>1. First let us find the C2C_{2}-algebra of Lc2,4kN=1L_{c_{2,4k}}^{N=1}. According to [Mil07, Section 4], we know that the null vector in universal algebra which survives inside the C2C_{2}-algebra is L(2)k1G(32)𝟏.L_{(-2)}^{k-1}G_{(-\frac{3}{2})}\mathbf{1}. Moreover if we let G(12)G_{(-\frac{1}{2})} act on the null vector, we get another null vector which survives in C2C_{2}-algebra, i.e L(2)k𝟏.L_{(-2)}^{k}\mathbf{1}. These two null vectors in the vacuum algebra generate the whole quotient ideal of RLc2,4kN=1R_{L_{c_{2,4k}}^{N=1}}. Thus RLc2,4kN=1R_{L_{c_{2,4k}}^{N=1}} is isomorphic to superalgebra [l,g]/lk,lk1g\mathbb{C}[l,g]/\langle l^{k},l^{k-1}g\rangle where gg is an odd element.

We are going to prove that ψ\psi is an isomorphism. We identify ll, gg with l(2)l(-2), g(32)g(-\frac{3}{2}), respectively, inside the jet superalgebra.

It is clear that J([l,g]/lk,lk1g)J_{\infty}(\mathbb{C}[l,g]/\langle l^{k},l^{k-1}g\rangle) is isomorphic to

[l(2i),g(32j)|i,j0]/l(z)k,l(z)k1g(z)\mathbb{C}[l(-2-i),g(-\frac{3}{2}-j)|i,j\geq 0]/\langle l(z)^{k},l(z)^{k-1}g(z)\rangle

where l(z)=nl(2n)znl(z)=\displaystyle\sum_{n\in\mathbb{N}}l(-2-n)z^{n}, g(z)=ng(32n)zng(z)=\displaystyle\sum_{n\in\mathbb{N}}g(-\frac{3}{2}-n)z^{n} and l(z)k,l(z)k1g(z)\langle l(z)^{k},l(z)^{k-1}g(z)\rangle is the ideal generated by the Fourier coefficients of l(z)k,l(z)k1g(z)l(z)^{k},l(z)^{k-1}g(z). We define ordered monomial in J([l,g]/lk,lk1g)J_{\infty}(\mathbb{C}[l,g]/\langle l^{k},l^{k-1}g\rangle) to be a monomial of the form

l(2n)a1g(32n))b1l(1n)a2g(32n+1))b2l(2)an+1g(32)bn+1l(-2-n)^{a_{1}}g(-\frac{3}{2}-n))^{b_{1}}l(-1-n)^{a_{2}}g(-\frac{3}{2}-n+1))^{b_{2}}\ldots l(-2)^{a_{n+1}}g(-\frac{3}{2})^{b_{n+1}}

where n0.n\geq 0. Then we have a complete lexicographic ordering on all ordered monomials according to Section 3.1.

We know that all ordered monomials constitute a spanning set of the jet superalgebra. Following the similar argument in Section 4.1, we can make use of the quotient relation to impose some conditions on the spanning set to get a smaller spanning set. Firstly since all variables g(k)sg(k)^{\prime}s are odd, no two g(k)g(k) can appear in the ordered monomial. The leading term of any coefficient of znkz^{nk} in l(z)kl(z)^{k} is l(2n)kl(-2-n)^{k}. Thus l(2n)kl(-2-n)^{k} should not appear as a segment of any element in spanning set. Similarly we can list further leading terms in the quotient:

  • Leading term of the coefficient of znkz^{nk} in l(z)k1g(z)l(z)^{k-1}g(z):

    l(2n)k1g(32n).l(-2-n)^{k-1}g(-\frac{3}{2}-n).
  • Leading term of the coefficient of zn(k1i)+(n1)i+nz^{n(k-1-i)+(n-1)i+n} in l(z)k1g(z)l(z)^{k-1}g(z):

    l(2n)k1ig(32n)l(2n+1)i(0ik1).l(-2-n)^{k-1-i}g(-\frac{3}{2}-n)l(-2-n+1)^{i}\quad(0\leq i\leq k-1).

Now we obtain a smaller spanning set where above three type leading terms can not appear inside any ordered monomial. More precisely, any element in this spanning set is of the form

w(b1)w(b2)w(bm)w(b_{1})w(b_{2})\ldots w(b_{m})

where bibi+1b_{i}\geq b_{i+1}, w(a)=l(a)w(a)=l(a) if aa\in\mathbb{Z} and w(a)=g(a)w(a)=g(a) if a12+.a\in\frac{1}{2}+\mathbb{Z}. And the fact that g(a)g(a) is odd implies that no half-odd-integer is repeated in {b1,b2,,bm}.\left\{b_{1},b_{2},\ldots,b_{m}\right\}. Moreover we have the condition

bjbjk+1>1,ifbj,b_{j}-b_{j-k+1}>1,\quad\text{if}\quad b_{j}\in\mathbb{Z},

because

l(2n)k,l(2n)k1g(32n),\displaystyle l(-2-n)^{k},\quad l(-2-n)^{k-1}g(-\frac{3}{2}-n),
l(2n)k1ig(32n)l(2n+1)i(1ik1)\displaystyle l(-2-n)^{k-1-i}g(-\frac{3}{2}-n)l(-2-n+1)^{i}\quad(1\leq i\leq k-1)

are leading terms of some elements in the quotient ideal. We also have a condition

bjbjk+21ifbj12,b_{j}-b_{j-k+2}\geq 1\quad\text{if}\quad b_{j}\in\frac{1}{2}\mathbb{Z},

because

g(32n)l(2n+1)k1g(-\frac{3}{2}-n)l(-2-n+1)^{k-1}

is the leading term of some element in the quotient ideal. So we have

HSq(J([l,g]/lk,lk1g)n=0Dk,1(n)qn2=ch[gr(Lc2,4kN=1)](q).HS_{q}(J_{\infty}(\mathbb{C}[l,g]/\langle l^{k},l^{k-1}g\rangle)\leq\sum_{n=0}^{\infty}D_{k,1}(n)q^{\frac{n}{2}}={\rm ch}[gr(L_{c_{2,4k}}^{N=1})](q).

Meanwhile the surjectivity of ψ\psi implies that

HSq(J([l,g]/lk,lk1g)ch[gr(Lc2,4kN=1)](q).HS_{q}(J_{\infty}(\mathbb{C}[l,g]/\langle l^{k},l^{k-1}g\rangle)\geq{\rm ch}[gr(L_{c_{2,4k}}^{N=1})](q).

Thus HSq(J([l,g]/lk,lk1g)=ch[gr(Lc2,4kN=1)](q)HS_{q}(J_{\infty}(\mathbb{C}[l,g]/\langle l^{k},l^{k-1}g\rangle)={\rm ch}[gr(L_{c_{2,4k}}^{N=1})](q) and ψ\psi is an isomorphism. It implies that above spanning set is a basis of the jet superalgebra. The image of basis of jet superalgebra under map ψ\psi is a basis of gr(Lc2,4kN=1).gr(L_{c_{2,4k}}^{N=1}).

We have following result which is a super-analog of [vEH18, Theorem 16.13]:

Theorem 6.1.

Let p>p2p^{\prime}>p\geq 2 satisfy pp2\frac{p^{\prime}-p}{2} and pp are coprime positive integers. We let Lcp,pN=1L_{c_{p,p^{\prime}}}^{N=1} denote the simple N=1N=1 vertex superalgebra associated with N=1N=1 superconformal (p,p)(p,p^{\prime})-minimal model of central charge cp,p=32(12(pp)2pp)c_{p,p^{\prime}}=\frac{3}{2}(1-\frac{2(p^{\prime}-p)^{2}}{pp^{\prime}}). Then the map ψ\psi is an isomorphism if and only if (p,p)=(2,4k)(p,p^{\prime})=(2,4k), (k1).(k\geq 1).

Proof.

We first consider C2C_{2}-algebra of Lcp,pN=1L_{c_{p,p^{\prime}}}^{N=1}. We let

|cp,p|=(p1)(p1)4+1+(1)pp8.|c_{p,p^{\prime}}|=\frac{(p-1)(p^{\prime}-1)}{4}+\frac{1+(-1)^{pp^{\prime}}}{8}\in\mathbb{N}.

When pp and pp^{\prime} are both even, according to [Mil07, Section 4], there are two null vectors which survive in RVcp,pN=1R_{V_{c_{p,p^{\prime}}}^{N=1}}, i.e. L(2)|cp,p|𝟏L_{(-2)}^{|c_{p,p^{\prime}}|}\mathbf{1} and L(2)|cp,p|1G32𝟏L_{(-2)}^{|c_{p,p^{\prime}}|-1}G_{-\frac{3}{2}}\mathbf{1}. They generate the quotient ideal of RVcp,pN=1R_{V_{c_{p,p^{\prime}}}^{N=1}} in the vacuum algebra. In this case, the C2C_{2}-algebra RLcp,pN=1R_{L_{c_{p,p^{\prime}}}^{N=1}} is isomorphic to

[l,g]/l|cp,p|,l|cp,p|1g.\mathbb{C}[l,g]/\langle l^{|c_{p,p^{\prime}}|},l^{|c_{p,p^{\prime}}|-1}g\rangle.

When pp and pp^{\prime} are both odd, again from [Mil07, Section 4], the null vector L(2)|cp,p|𝟏L_{(-2)}^{|c_{p,p^{\prime}}|}\bf{1} generates the quotient ideal of RLcp,pN=1.R_{L_{c_{p,p^{\prime}}}^{N=1}}. The C2C_{2}-algebra is isomorphic to

[l,g]/l|cp,p|.\mathbb{C}[l,g]/\langle l^{|c_{p,p^{\prime}}|}\rangle.

It is clear that HSq(J([l,g]/l|cp,p|)HS_{q}(J_{\infty}(\mathbb{C}[l,g]/\langle l^{|c_{p,p^{\prime}}|}\rangle) does not equal to ch[Lcp,pN=1](q){\rm ch}[L_{c_{p,p^{\prime}}}^{N=1}](q) when pp and pp^{\prime} are both odd. Thus ψ\psi is not an isomorphism in this case. Let pp and pp^{\prime} be both even. Suppose (p,p){(2,4k)|(k1)}(p,p^{\prime})\notin\left\{(2,4k)\;|\;(k\geq 1)\right\} and ψ\psi is an isomorphism for Lcp,pN=1L_{c_{p,p^{\prime}}}^{N=1}. Then

HSq(J([l,g]/l|cp,p|,l|cp,p|1g)=ch[Lcp,pN=1](q).HS_{q}(J_{\infty}(\mathbb{C}[l,g]/\langle l^{|c_{p,p^{\prime}}|},l^{|c_{p,p^{\prime}}|-1}g\rangle)={\rm ch}[L_{c_{p,p^{\prime}}}^{N=1}](q).

On the other hand, we have shown that

HSq(J([l,g]/lk,lk1g)=ch[Lc2,4kN=1](q),(k1).HS_{q}(J_{\infty}(\mathbb{C}[l,g]/\langle l^{k},l^{k-1}g\rangle)={\rm ch}[L_{c_{2,4k}}^{N=1}](q),\;(k\geq 1).

Therefore the character of Lcp,pN=1L_{c_{p,p^{\prime}}}^{N=1} must coincide with the character of Lc2,4kN=1L_{c_{2,4k}}^{N=1} for some kk. But according to [Mel94], the character of Lcp,pN=1L_{c_{p,p^{\prime}}}^{N=1} is

ch[Lcp,pN=1](q)=i1(1+qi1/2)i1(1qi)j(qj(jpp+pp)2q(jp+1)(jp+1)2),{\rm ch}[L_{c_{p,p^{\prime}}}^{N=1}](q)=\frac{\prod_{i\geq 1}(1+q^{i-1/2})}{\prod_{i\geq 1}(1-q^{i})}\sum_{j\in\mathbb{Z}}\left(q^{\frac{{j(jpp^{\prime}+p^{\prime}-p)}}{2}}-q^{\frac{{(jp+1)(jp^{\prime}+1)}}{2}}\right),

and it is easy to verify from the numerator that no two N=1N=1 minimal vertex algebras have the same character. This is a contradiction. Thus the statement is proved.

7. Extended Virasoro vertex algebras

For a simple Virasoro vertex algebra LVir(c2,2k+1,0)L_{Vir}(c_{2,2k+1},0) coming from (2,2k+1)(2,2k+1)-minimal model, according to [FF93], we know that RLVir(c2,2k+1,0)[x]/(xk)R_{L_{Vir}(c_{2,2k+1},0)}\cong\mathbb{C}[x]/(x^{k}) and ψ\psi is an isomorphism. Let pp and pp^{\prime} be two positive coprime integers satisfying p>p2p>p^{\prime}\geq 2. It is easy to see that ψ\psi is an isomorphism if and only if (p,p)=(2,2k+1)(p,p^{\prime})=(2,2k+1) (see [vEH18, Theorem 16.13]). Recently, the authors displayed the kernel of ψ\psi [vEH20, Theorem 2] for the c=12c=\frac{1}{2} Ising model vertex algebra LVir(c3,4,0)L_{Vir}(c_{3,4},0), based on a new fermionic character formula of LVir(c3,4,0)L_{Vir}(c_{3,4},0).

If we consider extended Virasoro vertex algebras associated with minimal model which is not necessarily a (2,2k+1)(2,2k+1)-minimal model, we might still have that ψ\psi is isomorphism. Our discussion is heavily motivated by [JM06] where the combinatorics of (super)extensions of (3,p)(3,p)-minimal vertex algebras were discussed.

Example 7.1.

For the free fermion model =LVir(c(3,4),0)LVir(c(3,4),12)\mathcal{F}=L_{Vir}(c_{(3,4)},0)\oplus L_{Vir}(c_{(3,4)},\frac{1}{2}), ψ\psi is clearly an isomorphism as discussed in Proposition 4.2.

Example 7.2.

The Lc2,8N=1L_{c_{2,8}}^{N=1} minimal vertex superalgebra has the following realization:

Lc(2,8)N=1LVir(c(3,8),0)LVir(c(3,8),32).L_{c_{(2,8)}}^{N=1}\cong L_{Vir}(c_{(3,8)},0)\oplus L_{Vir}(c_{(3,8)},\frac{3}{2}).

This realization is called extended algebra and was studied in [JM06]. The map ψ\psi is not an isomorphism in the case of LVir(c(3,8),0)L_{Vir}(c_{(3,8)},0). But we have shown that for the extended algebra of LVir(c(3,8),0)L_{Vir}(c_{(3,8)},0), the map ψ\psi is an isomorphism. This model was analyzed from a different perspective in [LMJ20].

Example 7.3.

Next let us consider V=LVir(c(3,10),0)LVir(c(3,10),2)V=L_{Vir}(c_{(3,10)},0)\oplus L_{Vir}(c_{(3,10)},2). It is well-known that

LVir(c(2,5),0)LVir(c(2,5),0)LVir(c(3,10),0)LVir(c(3,10),2).L_{Vir}(c_{(2,5)},0)\otimes L_{Vir}(c_{(2,5)},0)\cong L_{Vir}(c_{(3,10)},0)\oplus L_{Vir}(c_{(3,10)},2).

We let ω1\omega_{1} and ω2\omega_{2} be conformal vectors of the first factor and the second factor of LVir(c(2,5),0)LVir(c(2,5),0)L_{Vir}(c_{(2,5)},0)\otimes L_{Vir}(c_{(2,5)},0). Then the isomorphism map ff sends ω1+ω2\omega_{1}+\omega_{2} to the conformal vector ω\omega of LVir(c(3,10),0)L_{Vir}(c_{(3,10)},0) and ω1ω2\omega_{1}-\omega_{2} to the lowest weight vector ϕ\phi of LVir(c(3,10),2)L_{Vir}(c_{(3,10)},2). Since we know that

J(RLVir(c(2,5),0))J([x]/x2)gr(LVir(c2,5,0)),J_{\infty}(R_{L_{Vir}(c_{(2,5)},0)})\cong J_{\infty}(\mathbb{C}[x]/\langle x^{2}\rangle)\cong gr(L_{Vir}(c_{2,5},0)),

the map ψ\psi is an isomorphism for VV, i.e.

J(RV)=J(RLVir(c(2,5),0)RLVir(c(2,5),0))J([x,y]/x2,y2)gr(V).J_{\infty}(R_{V})=J_{\infty}(R_{L_{Vir}(c_{(2,5)},0)}\otimes R_{L_{Vir}(c_{(2,5)},0)})\cong J_{\infty}(\mathbb{C}[x,y]/\langle x^{2},y^{2}\rangle)\cong gr(V).

For LVir(c(3,10),0)LVir(c(3,10),2)L_{Vir}(c_{(3,10)},0)\oplus L_{Vir}(c_{(3,10)},2), its C2C_{2}-algebra is isomorphic to

[u,v]/uv,u2+v2,u3,v3\mathbb{C}[u,v]/\langle uv,u^{2}+v^{2},u^{3},v^{3}\rangle

after we identify x+yx+y, xyx-y in [x,y]/x2,y2\mathbb{C}[x,y]/\langle x^{2},y^{2}\rangle with uu and vv, respectively.

Remark 7.4.

We also know from [JM06] that the normalized parafermionic character of V=LVir(c(3,10),0)LVir(c(3,10),2)V=L_{Vir}(c_{(3,10)},0)\oplus L_{Vir}(c_{(3,10)},2) is given by

ch[V](q)=n1,n2,m10q(n1+n2+m1)(n1+n2)+n2(n2+m1)+m12+m1+n1+2n2(q)n1(q)n2(q)m1.{\rm ch}[V](q)=\sum_{n_{1},n_{2},m_{1}\geq 0}\frac{q^{(n_{1}+n_{2}+m_{1})(n_{1}+n_{2})+n_{2}(n_{2}+m_{1})+m_{1}^{2}+m_{1}+n_{1}+2n_{2}}}{(q)_{n_{1}}(q)_{n_{2}}(q)_{m_{1}}}.

Next let us consider the jet algebra

J([u,v]/u2,v3,uvJ_{\infty}(\mathbb{C}[u,v]/\langle u^{2},v^{3},uv\rangle

where degrees of uu and vv are both 2. Clearly, it has the following spanning set:

u(n1)u(nN)v(m1)v(mM)u_{(-n_{1})}\ldots u_{(-n_{N})}v_{(-m_{1})}\ldots v_{(-m_{M})}

subject to constraints:

  • (a)

    difference two condition at distance 1: nini+1+2n_{i}\geq n_{i+1}+2.

  • (b)

    difference two condition at distance 2: mimi+2+2.m_{i}\geq m_{i+2}+2.

  • (c)

    boundary condition: nN2+Mn_{N}\geq 2+M

where conditions (a),(a), (b),(b), (c)(c) are coming from (u2)(u^{2})_{\partial}, (v3)(v^{3})_{\partial}, (uv)(uv)_{\partial} in the quotient ideal of the jet algebra. Meanwhile according to Proposition 5.1 and Theorem 5.13 , we know that

J([u]/u2)gr(WΛ1,0),\displaystyle J_{\infty}(\mathbb{C}[u]/\langle u^{2}\rangle)\cong gr(W_{\Lambda_{1,0}}),
J([v]/v3)gr(WΛ2,0),\displaystyle J_{\infty}(\mathbb{C}[v]/\langle v^{3}\rangle)\cong gr(W_{\Lambda_{2,0}}),
J([u,v]/uv)gr(WΓ),\displaystyle J_{\infty}(\mathbb{C}[u,v]/\langle uv\rangle)\cong gr(W_{\Gamma}),

where Γ\Gamma is the graph \circ-\circ. Using three realizations of jet algebras and Gordon-Andrews character formulas from [FS93, CLM08], it is not hard to see that the above spanning set, subject to constraints (a)(a)-(c)(c), would produce a basis of the jet algebra J([u,v]/u2,v3,uv)J_{\infty}(\mathbb{C}[u,v]/\langle u^{2},v^{3},uv\rangle) whose Hilbert series is given by

n1,n2,m10q(n1+n2+m1)(n1+n2)+n2(n2+m1)+m12+m1+n1+2n2(q)n1(q)n2(q)m1.\displaystyle\sum_{n_{1},n_{2},m_{1}\geq 0}\frac{q^{(n_{1}+n_{2}+m_{1})(n_{1}+n_{2})+n_{2}(n_{2}+m_{1})+m_{1}^{2}+m_{1}+n_{1}+2n_{2}}}{(q)_{n_{1}}(q)_{n_{2}}(q)_{m_{1}}}.

One the other hand, the normalized character formula for V=LVir(c(2,5),0)LVir(c(2,5),0)V=L_{Vir}(c_{(2,5)},0)\otimes L_{Vir}(c_{(2,5)},0) is

ch[V](q)=n1,n20qn12+n22+n1+n2(q)n1(q)n2.{\rm ch}[V](q)=\sum_{n_{1},n_{2}\geq 0}\frac{q^{n_{1}^{2}+n_{2}^{2}+n_{1}+n_{2}}}{(q)_{n_{1}}(q)_{n_{2}}}.

Thus we have Hilbert series identities:

HSq(J([x,y]/(x2,y2))=HSq(J([u,v]/uv,u2+v2,u3,v3)\displaystyle HS_{q}(J_{\infty}(\mathbb{C}[x,y]/(x^{2},y^{2}))=HS_{q}(J_{\infty}(\mathbb{C}[u,v]/\langle uv,u^{2}+v^{2},u^{3},v^{3}\rangle)
=HSq(J([u,v]/u2,v3,uv))\displaystyle=HS_{q}(J_{\infty}(\mathbb{C}[u,v]/\langle u^{2},v^{3},uv)\rangle)

and

n1,n2,m10q(n1+n2+m1)(n1+n2)+n2(n2+m1)+m12+m1+n1+2n2(q)n1(q)n2(q)m1=n1,n20qn12+n22+n1+n2(q)n1(q)n2.\displaystyle\sum_{n_{1},n_{2},m_{1}\geq 0}\frac{q^{(n_{1}+n_{2}+m_{1})(n_{1}+n_{2})+n_{2}(n_{2}+m_{1})+m_{1}^{2}+m_{1}+n_{1}+2n_{2}}}{(q)_{n_{1}}(q)_{n_{2}}(q)_{m_{1}}}=\sum_{n_{1},n_{2}\geq 0}\frac{q^{n_{1}^{2}+n_{2}^{2}+n_{1}+n_{2}}}{(q)_{n_{1}}(q)_{n_{2}}}.

8. Conclusion and future work

In this work, we examined the injectivity of the ψ\psi map for several types of vertex algebras by making use of variety of methods, including qq-series identities. More precisely,

  • (1)

    The concept of jet algebra for vertex algebras can be extended to vertex superalgebras extending ideas of Arakawa in the super setup. We showed on several familiar examples, e.g. (2,4k)(2,4k) superconformal vertex algebras, that the ψ\psi map is an isomorphism. However, we also gave counterexamples coming from other N=1N=1 minimal models and a certain odd rank one lattice vertex algebra.

  • (2)

    We analyzed in great depth principal subspaces of lattice vertex algebras and affine vertex algebras and showed that the ψ\psi-map is isomorphism for many examples.

  • (3)

    We investigated the jet algebras coming from graphs. Interestingly, in some examples their Hilbert series are (mixed) mock modular forms.

Remark 8.1.

If LL is a root lattice of a Lie algebra of type DD or E6E_{6}, E7E_{7} and E8E_{8}, we expect that the ψ\psi map is an isomorphism for the FS principal subspace WLW_{L}. We will address this in our future work [LM].

Remark 8.2.

For the simple affine vertex algebra L𝔤^(k,0)L_{\widehat{\mathfrak{g}}}(k,0), we know that ψ\psi is an isomorphism if 𝔤\mathfrak{g} is of type CnC_{n} (n2)(n\geq 2) for k=1k=1 and 𝔤=sl2\mathfrak{g}=sl_{2} for any kk\in\mathbb{N}. But we expect that ψ\psi is isomorphism for any 𝔤\mathfrak{g} and any kk\in\mathbb{N}, and moreover, we expect that ψ\psi is isomorphism for the FS-principal subspaces thereof.

References

  • [Abe07] Toshiyuki Abe. Orbifold model of the symplectic fermionic vertex operator superalgebra. Mathematische Zeitschrift, 255(4):755–792, 2007.
  • [Ada97] Dražen Adamović. Rationality of Neveu-Schwarz vertex operator superalgebras. International mathematics research notices, 17:865–874, 1997.
  • [Ada99] Dražen Adamović. Rationality of unitary N=2N=2 vertex operator superalgebras. arXiv preprint math/9909055, 1999.
  • [Ada01] Dražen Adamović. Vertex algebra approach to fusion rules for N= 2 superconformal minimal models. Journal of Algebra, 239(2):549–572, 2001.
  • [AK18] Tomoyuki Arakawa and Kazuya Kawasetsu. Quasi-lisse vertex algebras and modular linear differential equations. In Lie Groups, Geometry, and Representation Theory, pages 41–57. Springer, 2018.
  • [AL18] Tomoyuki Arakawa and Andrew R Linshaw. Singular support of a vertex algebra and the arc space of its associated scheme. arXiv preprint arXiv:1804.01287, 2018.
  • [AM17] Tomoyuki Arakawa and Anne Moreau. Sheets and associated varieties of affine vertex algebras. Advances in Mathematics, 320:157–209, 2017.
  • [AM18a] Tomoyuki Arakawa and Anne Moreau. Arc spaces and chiral symplectic cores. arXiv preprint arXiv:1802.06533, 2018.
  • [AM18b] Tomoyuki Arakawa and Anne Moreau. Joseph ideals and lisse minimal ww-algebras. Journal of the Institute of Mathematics of Jussieu, 17(2):397–417, 2018.
  • [AM18c] Tomoyuki Arakawa and Anne Moreau. On the irreducibility of associated varieties of ww-algebras. Journal of Algebra, 500:542–568, 2018.
  • [Ara12] Tomoyuki Arakawa. A remark on the C2C_{2}-cofiniteness condition on vertex algebras. Mathematische Zeitschrift, 270(1-2):559–575, 2012.
  • [BFM91] Alexander Beilinson, Boris Feigin, and Barry Mazur. Introduction to algebraic field theory on curves. preprint, 395:462–463, 1991.
  • [BFOR17] Kathrin Bringmann, Amanda Folsom, Ken Ono, and Larry Rolen. Harmonic Maass forms and mock modular forms: theory and applications, volume 64. American Mathematical Soc., 2017.
  • [BGK18] Yuzhe Bai, Eugene Gorsky, and Oscar Kivinen. Quadratic ideals and Rogers–Ramanujan recursions. The Ramanujan Journal, pages 1–23, 2018.
  • [BK19] Marijana Butorac and Slaven Kožić. Principal subspaces for the affine Lie algebras in types DD, EE and FF. arXiv preprint arXiv:1902.10794, 2019.
  • [BLL+15] Christopher Beem, Madalena Lemos, Pedro Liendo, Wolfger Peelaers, Leonardo Rastelli, and Balt C van Rees. Infinite chiral symmetry in four dimensions. Communications in Mathematical Physics, 336(3):1359–1433, 2015.
  • [BMS13] Clemens Bruschek, Hussein Mourtada, and Jan Schepers. Arc spaces and the Rogers–Ramanujan identities. The Ramanujan Journal, 30(1):9–38, 2013.
  • [BPT16] Ivana Baranovic, Mirko Primc, and Goran Trupcevic. Bases of Feigin-Stoyanovsky’s type subspaces for c(1)c_{\ell}^{(1)}. arXiv preprint arXiv:1603.04594, 2016.
  • [But12] Marijana Butorac. Combinatorial bases of principal subspaces of standard modules for affine Lie algebra of type B2(1)B_{2}^{(}1). PhD thesis, Prirodoslovno-matematički fakultet-Matematički odsjek, Sveučilište u Zagrebu, 2012.
  • [CLM06] Stefano Capparelli, James Lepowsky, and Antun Milas. The Rogers–Selberg recursions, the Gordon–Andrews identities and intertwining operators. The Ramanujan Journal, 12(3):379–397, 2006.
  • [CLM08] Corina Calinescu, James Lepowsky, and Antun Milas. Vertex-algebraic structure of the principal subspaces of certain A1(1)A_{1}^{(1)}-modules, II: Higher-level case. Journal of Pure and Applied Algebra, 212(8):1928–1950, 2008.
  • [DLM97] Chongying Dong, Haisheng Li, and Geoffrey Mason. Certain Associative Algebras Similar to U(sl2)U(sl_{2}) and Zhu’s Algebra A(VL)A(VL). Journal of Algebra, 196(2):532–551, 1997.
  • [Fei09] Evgeny Feigin. The PBW filtration. Representation Theory of the American Mathematical Society, 13(9):165–181, 2009.
  • [FF93] Boris Feigin and Edward Frenkel. Coinvariants of nilpotent subalgebras of the Virasoro algebra and partition identities. Adv. Sov. Math, 16(1003):139–148, 1993.
  • [FFJ+09] Boris Feigin, Evgeny Feigin, Michio Jimbo, Tetsuji Miwa, Evgeny Mukhin, et al. Principal sl^3\widehat{sl}_{3} subspaces and quantum Toda Hamiltonian. In Algebraic Analysis and Around: In honor of Professor Masaki Kashiwara’s 60th birthday, pages 109–166. Mathematical Society of Japan, 2009.
  • [FFL11] Boris Feigin, Evgeny Feigin, and Peter Littelmann. Zhu’s algebras, C2C_{2}-algebras and abelian radicals. Journal of Algebra, 329(1):130–146, 2011.
  • [FLK+01] B Feigin, S Loktev, Rinat Kedem, T Miwa, and E Mukhin. Combinatorics of the spaces of coinvariants. Transformation groups, 6(1):25–52, 2001.
  • [FS93] Boris Feigin and AV Stoyanovsky. Quasi-particles models for the representations of Lie algebras and geometry of flag manifold. arXiv preprint hep-th/9308079, 1993.
  • [Jer12] Miroslav Jerković. Character formulas for Feigin–Stoyanovsky?s type subspaces of standard sl(3,C)^\widehat{sl(3,C)}-modules. The Ramanujan Journal, 27(3):357–376, 2012.
  • [JM06] P Jacob and P Mathieu. Embedding of bases: From the (2,2κ+1)\mathcal{M}(2,2\kappa+1) to the (3,4κ+2δ)\mathcal{M}(3,4\kappa+2-\delta) Models. Physics Letters B, 635(5-6):350–354, 2006.
  • [JSM20] Chris Jennings-Shaffer and Antun Milas. Further qq-series identities and conjectures relating false theta functions and characters. arXiv preprint arXiv:2005.13620, 2020.
  • [Kac98] Victor G Kac. Vertex algebras for beginners. Number 10. American Mathematical Soc., 1998.
  • [Li05] Haisheng Li. Abelianizing vertex algebras. Communications in mathematical physics, 259(2):391–411, 2005.
  • [LL12] James Lepowsky and Haisheng Li. Introduction to vertex operator algebras and their representations, volume 227. Springer Science & Business Media, 2012.
  • [LM] Hao Li and Antun Milas. Quantum dilogarithm and characters of FS-principal sub- spaces. in preparation.
  • [LMJ20] Hao Li, Antun Milas, and Wauchope Josh. S2{S}_{2} permutation orbifolds of N=1 and N=2 vertex superalgebras and W-algebras. submitted, 2020.
  • [Mel94] Ezer Melzer. Supersymmetric analogs of the Gordon-Andrews identities, and related TBA systems. arXiv preprint hep-th/9412154, 1994.
  • [Mil07] Antun Milas. Characters, supercharacters and Weber modular functions. Journal für die reine und angewandte Mathematik (Crelles Journal), 2007(608):35–64, 2007.
  • [MP87] Arne Meurman and Mirko Primc. Annihilating ideals of standard modules of sl(2,C)~\widetilde{sl(2,C)} and combinatorial identities. Advances in mathematics (New York, NY. 1965), 64(3):177–240, 1987.
  • [MP12] Antun Milas and Michael Penn. Lattice vertex algebras and combinatorial bases: general case and W-algebras. New York J. Math, 18:621–650, 2012.
  • [Oga00] Akihiko Ogawa. Zhu’s algebra of rank one lattice vertex operator superalgebras. Osaka Journal of Mathematics, 37(4):811–822, 2000.
  • [Pen14] Michael Penn. Lattice vertex superalgebras, I: Presentation of the principal subalgebra. Communications in Algebra, 42(3):933–961, 2014.
  • [Pri94] Mirko Primc. Vertex operator construction of standard modules for An(1)A_{n}(1). Pacific Journal of Mathematics, 162(1):143–187, 1994.
  • [Pri00] Mirko Primc. Basic representations for classical affine Lie algebras. Journal of Algebra, 228(1):1–50, 2000.
  • [PŠ16] Mirko Primc and Tomislav Šikić. Combinatorial bases of basic modules for affine Lie algebras Cn(1)C_{n}^{(1)}. Journal of Mathematical Physics, 57(9):091701, 2016.
  • [Tru09] Goran Trupčević. Combinatorial bases of Feigin–Stoyanovsky’s type subspaces of higher-level standard sl(l+ 1, c)-modules. Journal of Algebra, 322(10):3744–3774, 2009.
  • [Tru11] Goran Trupčević. Characters of Feigin-Stoyanovsky’s type subspaces of level one modules for affine Lie algebras of types Al(1)A_{l}(1) and D4(1)D_{4}(1). Glasnik matematički, 46(1):49–70, 2011.
  • [vEH18] Jethro van Ekeren and Reimundo Heluani. Chiral homology of elliptic curves and Zhu’s algebra. arXiv preprint arXiv:1804.00017, 2018.
  • [vEH20] Jethro van Ekeren and Reimundo Heluani. The singular support of the Ising model. arXiv preprint arXiv:2005.10769, 2020.
  • [Zhe17] Lisun Zheng. Vertex operator superalgebras associated with affine Lie superalgebras. Communications in Algebra, 45(6):2417–2434, 2017.
  • [Zhu96] Yongchang Zhu. Modular invariance of characters of vertex operator algebras. Journal of the American Mathematical Society, 9(1):237–302, 1996.