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

Quartic Hamiltonians, and higher Hamiltonians at next-to-leading order, for the affine 𝔰𝔩2\bm{\mathfrak{sl}_{2}} Gaudin model

Tommaso Franzini111email t.franzini@herts.ac.uk     Charles Young222email c.young8@herts.ac.uk Department of Physics, Astronomy and Mathematics,
University of Hertfordshire, Hatfield, AL10 9AB, United Kingdom
Abstract

In this work we will use a general procedure to construct higher local Hamiltonians for the affine 𝔰𝔩2\mathfrak{sl}_{2} Gaudin model. We focus on the first non-trivial example, the quartic Hamiltonians. We show by direct calculation that the quartic Hamiltonians commute amongst themselves and with the quadratic Hamiltonians which define the model.

We go on to introduce a certain next-to-leading-order semi-classical limit of the model. In this limit, we are able to write down the full hierarchy of higher local Hamiltonians and prove that they commute.

 

1 Introduction

Gaudin models represent interesting theories that find applications in several contexts in mathematical physics. In particular, they provide a general framework to study rich classes of classical and quantum integrable 1+1 dimensional integrable theories [Gau76, Gau14, Vic18, Lac18]. Recently, it has also been found that classical untwisted Gaudin models provide a dual description of such theories to the one given by 4D holomorphic Chern-Simons theories [Vic21, LV21].

In this paper we will focus on quantum Gaudin models. To define them, one needs a collection of distinct points {z1,,zN}1\{z_{1},\dots,z_{N}\}\in\mathbb{C}\mathbb{P}^{1} on the Riemann sphere and a Kac-Moody algebra 𝔤\mathfrak{g}, which can be of finite or affine type. These models are defined by the quadratic Hamiltonians, namely

i=j=1jiNκabIa(i)Ib(j)zizj,i=1,,N,\mathscr{H}_{i}=\sum_{\begin{subarray}{c}j=1\\ j\neq i\end{subarray}}^{N}\kappa_{ab}\frac{I^{a(i)}I^{b(j)}}{z_{i}-z_{j}},\qquad i=1,\dots,N, (1)

in the (suitably completed, in the affine case) tensor product U(𝔤)NU(\mathfrak{g})^{\otimes N}. Here Ia(i)I^{a(i)} are the generators of 𝔤\mathfrak{g}, defined at site ii, and κ\kappa is the invariant bilinear form on 𝔤\mathfrak{g}.

Finite-type quantum Gaudin models have been deeply studied. In particular, it is known that the quadratic Hamiltonians (1) are part of a larger family of mutually commuting operators U(𝔤N)\mathscr{B}\in U(\mathfrak{g}^{\otimes N}) called the Bethe or Gaudin subalgebra [Fre04, Tal04, Ryb06]. Moreover, these higher Hamiltonians can be diagonalized with an elegant form of Bethe Ansatz, where the eigenvector is the Schechtman-Varchenko vector and the eigenvalues are encoded as functions on a space of 𝔤L\prescript{L}{}{{{\mathfrak{g}}}}-opers, the Langlands dual algebra of 𝔤\mathfrak{g} [FFR94, MTV06, MTV07, MV02, MV04].

On the other hand, affine-type Gaudin models are still far from being fully understood.

One approach to affine Gaudin models was proposed in a recent paper [KL21]. In this work the authors propose an integrable model called generalized affine 𝔰𝔩2\mathfrak{sl}_{2} Gaudin model, which reproduces the usual 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2} model in a certain limit. Their construction is based on a new realization of the subalgebra Uq(𝔟)U_{q}(\mathfrak{b}_{-}) through a new class of vertex operators, and fits affine Gaudin models into the general procedure first given in the seminal work [BLZ97].

Another approach, first proposed in the pioneering work [FF07], is to treat affine Gaudin models in closer analogy with their finite-type cousins. In particular, the spectrum of higher Hamiltonians should be described by suitable functions on a space of opers. Some further conjectures of how this might work, at least for the local Hamiltonians, were made in [LVY18], where it was conjectured that the eigenvalues of higher local Hamiltonians of the affine Gaudin models, as well as the Hamiltonians themselves, are given by hypergeometric-type integrals in the spectral plane, namely

Q^nγ=γ𝒫(z)n/2ςn(z)[0]dz,\widehat{Q}_{n}^{\gamma}=\int_{\gamma}\mathscr{P}(z)^{-n/2}\varsigma_{n}(z)_{[0]}\text{d}z, (2)

where nn lives is a (multi)set of indices given by the exponents of the algebra 𝔤\mathfrak{g}, 𝒫\mathscr{P} is a certain multi-valued function defined by the data of the levels kik_{i} of the modules attached to the marked points ziz_{i}, γ\gamma is a Pochhammer contour in the spectral plane around any two of these points (see e.g. fig. 1) and ςn(z)[0]\varsigma_{n}(z)_{[0]} can be thought as the Hamiltonian density.

Refer to caption
Figure 1: An example of Pochhammer contour γ\gamma around any two marked points.

The key step in computing the higher Hamiltonians is to characterize these Hamiltonian densities, which are obtained by defining a suitable state ςn(z)𝒱\varsigma_{n}(z)\in\mathcal{V} for each given exponent nn. In order to do that, it is possible to exploit the general properties consistent Hamiltonians must obey: they have to commute with the generators {Inα}nα=1,,dim𝔤̊\{I^{\alpha}_{n}\}_{n\in\mathbb{Z}}^{\alpha=1,\dots,\dim\mathring{\mathfrak{g}}} of the algebra 𝔤\mathfrak{g} defining the model as well as amongst themselves (here 𝔤̊\mathring{\mathfrak{g}} denotes the underlying finite algebra). As we will see in section 4.7, this is equivalent to the following requirements

ΔIn0αςm(z)=0modtwisted derivatives,ςn(z)(0)ςm(z)=0modtwisted derivatives and translates,\begin{split}\Delta I^{\alpha}_{n\geq 0}\varsigma_{m}(z)=0&\mod\text{twisted derivatives},\\ \varsigma_{n}(z)_{(0)}\varsigma_{m}(z)=0&\mod\text{twisted derivatives and translates},\end{split} (3)

where ΔIn0α\Delta I^{\alpha}_{n\geq 0} represents the diagonal action of the positive modes of the generators of the algebra 𝔤\mathfrak{g} and the zero mode ςn(z)(0)\varsigma_{n}(z)_{(0)} is intended in the vertex algebra sense (see section 2.4). We give a precise definition of translation and twisted derivative of a state in sections 2.4 and 4.4.

The general expectation is that there exists a state ςm(z)\varsigma_{m}(z) for every exponent mm of the affine algebra 𝔤\mathfrak{g}, and that it takes the following form

ςm(z)=ti1,,im+1I1i1(z)I1im+1(z)|0+quantum corrections,\varsigma_{m}(z)=t_{i_{1},\dots,i_{m+1}}I^{i_{1}}_{-1}(z)\cdots I^{i_{m+1}}_{-1}(z)\ket{0}+\text{quantum corrections}, (4)

where I(z)=iI(i)/(zzi)I(z)=\sum_{i}I^{(i)}/(z-z_{i}) and tt is a certain symmetric invariant tensor of 𝔤̊\mathring{\mathfrak{g}}. This particular structure is justified by the semi-classical counterpart of these models, which have been thoroughly studied [Eva+99, Eva01, LMV17]; in particular, the precise choice of symmetric invariant tensors needed to ensure that the Hamiltonians Poisson-commute is well understood and related to Drinfel’d-Sokolov reduction [Eva01, DS85]. In the very simplest cases, including the cubic Hamiltonian in type 𝔰𝔩^M3\widehat{\mathfrak{sl}}_{M\geq 3}, there are no quantum corrections needed [LVY20]. The first case in which quantum corrections are present appeared in the recent paper [KLT22]: it is the example of the quartic Hamiltonian in type 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}. In this paper, we consider this example in detail. In particular, we carry out the full computation to show that the resulting quartic Hamiltonians commute amongst themselves (see theorem 4.5.1 and proposition 4.7.2).

This direct calculation reveals the following fact: it turns out that any density ς3(z)\varsigma_{3}(z) obeying the first of the two conditions in eq. 3 (i.e. the one which ensures the Hamiltonian commutes with the generators of the affine algebra), automatically also obeys the second condition (which is needed for the Hamiltonians to commute amongst themselves) at least for n=1,3n=1,3, as shown in corollary 4.5.1 below. We should stress that this property is, for the moment, highly non-obvious and suggestive; it would be good to get a more systematic understanding of why it should be true.


Already in this case of the exponent n=3n=3, i.e. of quartic Hamiltonians, the direct computations needed are very lengthy. This is especially true of the computations needed to show the mutual commutativity of the quartic Hamilonians. For higher exponents n5n\geq 5, direct calculations become computationally difficult even with the aid of computer algebra, but we are able to prove a result for all exponents by truncating to the next-to-leading order in \hbar. To introduce the dependence from the formal parameter \hbar, we perform a re-scaling of the generators, namely II~II\to\tilde{I}\equiv\hbar I, kk~kk\to\tilde{k}\equiv k, in such a way that every time we perform a commutation, we introduce a factor of \hbar. This procedure allows us to identify different quantum corrections by their \hbar dependence. In this spirit, we will then work modulo terms of order 2\hbar^{2}. We show that, modulo such terms, the Hamiltonian density for each exponent 2n12n-1, n1n\in\mathbb{Z}_{\geq 1}, of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}, takes the following form

ς~(z)2n1=ti1,,i2nI~1i1(z)I~1i2(z)I~1i2n(z)|0+n(2n+1)(2n2)(2n1)ti1,,i2n4fabcI~2a(z)I~1b(z)I~1(c(z)I~1i1(z)I~1i2n4)(z)|0\begin{split}\tilde{\varsigma}&{}_{2n-1}(z)=t_{i_{1},\dots,i_{2n}}\tilde{I}^{i_{1}}_{-1}(z)\tilde{I}^{i_{2}}_{-1}(z)\cdots\tilde{I}^{i_{2n}}_{-1}(z)\ket{0}\\ &\qquad+\hbar\frac{n(2n+1)(2n-2)}{(2n-1)}t_{i_{1},\dots,i_{2n-4}}f^{abc}\tilde{I}^{a}_{-2}(z)\tilde{I}^{b\prime}_{-1}(z)\tilde{I}^{(c}_{-1}(z)\tilde{I}^{i_{1}}_{-1}(z)\cdots\tilde{I}^{i_{2n-4})}_{-1}(z)\ket{0}\end{split}

and prove that the resulting Hamiltonians commute up to and including terms of order 2\hbar^{2}.

***

The paper is organized as follows.

In section 2 we recall the main ideas behind the theory of Kac-Moody algebras, their local completion and the concept of vacuum Verma module. We also recall the definition and the main features of vertex algebras.

In section 3 we define the algebra of observables of the Gaudin model and we recall the definition of invariant tensor of an algebra 𝔤\mathfrak{g}, focusing in particular on the case of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}.

In section 4 we define the meromorphic states I(z)I(z), which represent the building blocks for the states we want to define. We also present their commutation relations and we describe the gradation they give rise, which will be of fundamental importance to prove the main results of the paper. After that, we characterize the space of states that vanish under the diagonal action of zero modes: this is the first step to define higher Gaudin Hamiltonians. At this point we will present all quantum corrections to the quartic state and we will explain why they appear in this specific example and not in the other known ones. In the following subsection, we restrict the number of possible states by asking they be singular up to twisted derivative under the diagonal action of positive modes. We focus on the quadratic and quartic states. Afterwards, we prove the main result of the paper, where we also ask that the 0th product (in the vertex algebra sense) of these states vanishes modulo twisted derivatives and translates. We will show that there is one quartic state, up to re-scaling and modulo the addition of translates and twisted derivatives, that satisfies this last requirement. At this point, having a precise definition of the quartic state, we describe the general construction of the higher Hamiltonians, in the spirit of [LVY18]. We will explain in detail why we ask for these properties and why they are important at the level of the quantum Hamiltonians.

In section 5 we try to solve the same problem from a different point of view. Instead of focusing on one specific Hamiltonian, trying to work out its explicit definition, we want to characterize all Hamiltonians at arbitrary nn, but working at sub-leading order. In order to do that we introduce the formal parameter \hbar by making a particular re-scaling of the generators of the algebra. We will then prove similar theorems to those of the previous sections, working modulo terms at order 3\hbar^{3}.


Acknowledgements. The authors would like to thank Sylvain Lacroix for helpful discussions and for providing many detailed comments. The research of Charles Young is supported by the Leverhulme Trust, Research Project Grant number RPG-2021-092.

2 Vacuum Verma modules for 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}

2.1 Loop realization of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}

We define the loop algebra 𝔰𝔩2[t,t1]=𝔰𝔩2[t,t1]\mathfrak{sl}_{2}[t,t^{-1}]=\mathfrak{sl}_{2}\otimes\mathbb{C}[t,t^{-1}] as the algebra of Laurent polynomials in a formal variable tt with coefficient in the finite-dimensional Lie algebra 𝔰𝔩2\mathfrak{sl}_{2}. The Lie brackets on this algebra are given by

[af(t),bg(t)]=[a,b]𝔰𝔩2f(t)g(t),[a\otimes f(t),b\otimes g(t)]=[a,b]_{\mathfrak{sl}_{2}}\otimes f(t)g(t), (5)

where f(t)f(t) and g(t)g(t) are arbitrary Laurent polynomials in [t,t1]\mathbb{C}[t,t^{-1}].

Let (|):𝔰𝔩2×𝔰𝔩2(\cdot|\cdot):\mathfrak{sl}_{2}\times\mathfrak{sl}_{2}\rightarrow\mathbb{C} be the canonically normalized symmetric invariant bilinear form on 𝔰𝔩2\mathfrak{sl}_{2}. It is given by taking the trace in the defining representation:

(a|b):=Tr(ab),(a|b)\vcentcolon=\Tr(ab), (6)

It is possible to extend the loop algebra by a one-dimensional central element 𝗄\mathbb{C}\mathsf{k},

0𝗄𝔰𝔩^2𝔰𝔩2[t,t1]0.0\longrightarrow\mathbb{C}\mathsf{k}\longrightarrow\widehat{\mathfrak{sl}}_{2}\longrightarrow\mathfrak{sl}_{2}[t,t^{-1}]\longrightarrow 0. (7)

This extension is called the affine Lie algebra 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}, whose commutation relations are

[af(t),bg(t)]=[a,b]𝔰𝔩2f(t)g(t)(restfdg)(a|b)𝗄,\displaystyle[a\otimes f(t),b\otimes g(t)]=[a,b]_{\mathfrak{sl}_{2}}\otimes f(t)g(t)-(\text{res}_{t}f\text{d}g)(a|b)\mathsf{k}, (8)
[𝗄,]=0.\displaystyle[\mathsf{k},\cdot]=0. (9)

We shall use the notation

an:=atn,for a𝔰𝔩2 and n.a_{n}\vcentcolon=a\otimes t^{n},\quad\text{for $a\in\mathfrak{sl}_{2}$ and $n\in\mathbb{Z}$}. (10)

The commutation relations can be equivalently written as

[am,bn]=[a,b]n+mnδn+m,0(a|b)𝗄.[a_{m},b_{n}]=[a,b]_{n+m}-n\delta_{n+m,0}(a|b)\mathsf{k}. (11)

We can add to this algebra a one-dimensional derivation 𝖽\mathsf{d}, such that [𝖽,𝗄]=0[\mathsf{d},\mathsf{k}]=0 and [𝖽,af(t)]=attf(t)[\mathsf{d},a\otimes f(t)]=a\otimes t\partial_{t}f(t), for all a𝔰𝔩2a\in\mathfrak{sl}_{2} and f(t)[t,t1]f(t)\in\mathbb{C}[t,t^{-1}]. It is possible to show that this algebra is isomorphic to the Kac-Moody algebra over \mathbb{C} of type 𝖠1(1)\mathsf{A}_{1}^{(1)}, see e.g. [Kac90, ch. 7],

𝔤=𝔰𝔩^2𝖽.\mathfrak{g}=\widehat{\mathfrak{sl}}_{2}\oplus\mathbb{C}\mathsf{d}. (12)

2.2 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2} as a Kac-Moody algebra

The Cartan matrix for the Kac-Moody algebra of type 𝖠11{}^{1}\mathsf{A}_{1} is defined as A=(ai,j)i,j=01=(2δi,jδi+1,jδi1,j)i,j=01A=(a_{i,j})_{i,j=0}^{1}=(2\delta_{i,j}-\delta_{i+1,j}-\delta_{i-1,j})_{i,j=0}^{1}. The Cartan decomposition is given by 𝔤=𝔫𝔥𝔫+\mathfrak{g}=\mathfrak{n}_{-}\oplus\mathfrak{h}\oplus\mathfrak{n}_{+}. The Chevalley-Serre generators are {ei}i=01𝔫+\{e_{i}\}_{i=0}^{1}\subset\mathfrak{n}_{+}, {fi}i=01𝔫\{f_{i}\}_{i=0}^{1}\subset\mathfrak{n}_{-} while {αˇi}i=01𝔥\{\check{\alpha}_{i}\}_{i=0}^{1}\subset\mathfrak{h} and {αi}i=01𝔥\{\alpha_{i}\}_{i=0}^{1}\subset\mathfrak{h^{\ast}} are respectively a basis for the Cartan subalgebra of simple coroots of 𝔤\mathfrak{g} and a basis for the dual Cartan subalgebra of simple roots of 𝔤\mathfrak{g}. The latter are related via the canonical pairing between the Cartan algebra and its dual, ,:𝔥×𝔥\langle\cdot,\cdot\rangle:\mathfrak{h}^{\ast}\times\mathfrak{h}\rightarrow\mathbb{C}

αi,αˇj=ai,j.\langle\alpha_{i},\check{\alpha}_{j}\rangle=a_{i,j}. (13)

The fundamental commutation relations in 𝔤\mathfrak{g} are

=αi,xei,[x,fi]=αi,xfi,[x,x]=0,[ei,fj]=αˇiδij,\begin{gathered}=\langle\alpha_{i},x\rangle e_{i},\qquad[x,f_{i}]=-\langle\alpha_{i},x\rangle f_{i},\\ [x,x^{\prime}]=0,\qquad[e_{i},f_{j}]=\check{\alpha}_{i}\delta_{ij},\end{gathered} (14)

where x,x𝔥x,x^{\prime}\in\mathfrak{h} and i,j=0,1i,j=0,1, together with the Serre relations

(adei)1aijej=0,(adfi)1aijfj=0.(\text{ad}e_{i})^{1-a_{ij}}e_{j}=0,\qquad(\text{ad}f_{i})^{1-a_{ij}}f_{j}=0. (15)

The Kac-Moody algebra 𝔤\mathfrak{g} has a central element 𝗄=i=01αˇi\mathsf{k}=\sum_{i=0}^{1}\check{\alpha}_{i}, which spans a one-dimensional centre. A basis for the Cartan subalgebra 𝔥\mathfrak{h} is given by the coroots {αˇi}i=01\{\check{\alpha}_{i}\}_{i=0}^{1} together with the derivation element 𝖽\mathsf{d}, which by definition satisfies

αi,𝖽=δi,0.\langle\alpha_{i},\mathsf{d}\rangle=\delta_{i,0}. (16)

If we remove the 0th row and column from AA, we obtain the Cartan matrix corresponding to the finite dimensional Lie algebra 𝔰𝔩2\mathfrak{sl}_{2}. This subalgebra of 𝔤\mathfrak{g} is generated by e1𝔫+e_{1}\in\mathfrak{n}_{+}, f1𝔫f_{1}\in\mathfrak{n}_{-} and αˇ1𝔥\check{\alpha}_{1}\in\mathfrak{h}.

2.3 Local completion and vacuum Verma modules

For any kk\in\mathbb{C}, let us define Uk(𝔰𝔩^2)U_{k}(\widehat{\mathfrak{sl}}_{2}) as the quotient of the universal enveloping algebra U(𝔰𝔩^2)U(\widehat{\mathfrak{sl}}_{2}) of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2} by the two-sided ideal generated by 𝗄k\mathsf{k}-k. For each n0n\in\mathbb{Z}_{\geq 0}, let us introduce the left ideal Jn=Uk(𝔰𝔩^2)(𝔰𝔩2tn[t])J_{n}=U_{k}(\widehat{\mathfrak{sl}}_{2})\cdot(\mathfrak{sl}_{2}\otimes t^{n}\mathbb{C}[t]). The inverse limit

Uk~(𝔰𝔩^2)=lim\faktorUk(𝔰𝔩^2)Jn\widetilde{U_{k}}(\widehat{\mathfrak{sl}}_{2})=\varprojlim\faktor{U_{k}(\widehat{\mathfrak{sl}}_{2})}{J_{n}} (17)

is a complete topological algebra, called the local completion of Uk(𝔰𝔩^2)U_{k}(\widehat{\mathfrak{sl}}_{2}) at level kk. With this definition, the elements of Uk~(𝔰𝔩^2)\widetilde{U_{k}}(\widehat{\mathfrak{sl}}_{2}) are possibly infinite sums of the type m0am\sum_{m\geq 0}a_{m} of elements amUk(𝔰𝔩^2)a_{m}\in U_{k}(\widehat{\mathfrak{sl}}_{2}) which do truncate to finite sums when working modulo any JnJ_{n}.

A module \mathscr{M} over 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2} is said to be smooth if, for all a𝔰𝔩2a\in\mathfrak{sl}_{2} and all vv\in\mathscr{M}, anv=0a_{n}v=0 for sufficiently large nn. A module \mathscr{M} has level kk if 𝗄k\mathsf{k}-k acts as zero on \mathscr{M}. Any smooth module of level kk over 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2} is also a module over the completion Uk~(𝔰𝔩^2)\widetilde{U_{k}}(\widehat{\mathfrak{sl}}_{2}).

We can identify the subalgebra of positive modes 𝔰𝔩2[t]𝗄𝔰𝔩^2\mathfrak{sl}_{2}[t]\oplus\mathbb{C}\mathsf{k}\subset\widehat{\mathfrak{sl}}_{2} and introduce the one-dimensional representation |0k\mathbb{C}\ket{0}^{k} defined by

(𝗄k)|0k=0,an|0k=0for all n0a𝔰𝔩2.(\mathsf{k}-k)\ket{0}^{k}=0,\qquad\qquad a_{n}\ket{0}^{k}=0\quad\text{for all $n\geq 0$, $a\in\mathfrak{sl}_{2}$}. (18)

We define 𝕍0k\mathbb{V}_{0}^{k}, the vacuum Verma module at level kk, as the induced smooth 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}-module

𝕍0k=U(𝔰𝔩^2)U(𝔰𝔩2[t]𝗄)|0k\mathbb{V}_{0}^{k}=U(\widehat{\mathfrak{sl}}_{2})\otimes_{U(\mathfrak{sl}_{2}[t]\oplus\mathbb{C}\mathsf{k})}\mathbb{C}\ket{0}^{k} (19)

This vector space is spanned by monomials of the form apbq|0ka_{p}\cdots b_{q}\ket{0}^{k}, with a,,b𝔰𝔩2a,\dots,b\in\mathfrak{sl}_{2} and strictly negative mode numbers p,,q<0p,\dots,q\in\mathbb{Z}_{<0}. We call these vectors states.

Let us denote by [T,][T,\cdot] the derivation on Uk(𝔰𝔩^2)U_{k}(\widehat{\mathfrak{sl}}_{2}) defined by [T,an]=nan1[T,a_{n}]=-na_{n-1} and [T,1]=0[T,1]=0. By setting T(X|0k)=[T,X]|0kT(X\ket{0}^{k})=[T,X]\ket{0}^{k} for any XUk(𝔰𝔩^2)X\in U_{k}(\widehat{\mathfrak{sl}}_{2}), one can interpret TT as a translation operator T:𝕍0k𝕍0kT:\mathbb{V}_{0}^{k}\rightarrow\mathbb{V}_{0}^{k}: the reason for this identification will be clear in the next section.

2.4 Vertex algebra structure

As we now recall, the vacuum Verma module defined in the previous section has the structure of vertex algebra. Namely, we have the state-field map Y(,x)Y(\cdot,x), which for every state A𝕍0kA\in\mathbb{V}_{0}^{k} associates a formal power series in the variable xx,

Y(,x):𝕍0kHom(𝕍0k,𝕍0k((x)))AY(A,x)=nA(n)xn1\begin{split}Y(\cdot,x):\mathbb{V}_{0}^{k}&\longrightarrow\operatorname{Hom}(\mathbb{V}_{0}^{k},\mathbb{V}_{0}^{k}((x)))\\ A&\longmapsto Y(A,x)=\sum_{n\in\mathbb{Z}}A_{(n)}x^{-n-1}\end{split} (20)

where A(n)End(𝕍)A_{(n)}\in\operatorname{End}(\mathbb{V}) is the nnth mode of AA. By definition, if A=a1|0kA=a_{-1}\ket{0}^{k} for some a𝔰𝔩2a\in\mathfrak{sl}_{2}, then A(n)=anA_{(n)}=a_{n} for all nn\in\mathbb{Z}, i.e.

Y(a1|0k,x)=nanxn1.Y(a_{-1}\ket{0}^{k},x)=\sum_{n\in\mathbb{Z}}a_{n}x^{-n-1}. (21)

The fields for all other states can be obtained with the following properties

Y(TA,x)=zY(A,x),Y(A(1)B,x)=:Y(A,x)Y(B,x):Y(TA,x)=\partial_{z}Y(A,x),\qquad\qquad Y(A_{(-1)}B,x)=:Y(A,x)Y(B,x): (22)

where we have introduced the normal ordered product between fields

:Y(A,x)Y(B,x):=(m<0A(m)zm1)Y(B,x)+Y(B,x)(m0A(m)xm1).:Y(A,x)Y(B,x):\,=\left(\sum_{m<0}A_{(m)}z^{-m-1}\right)Y(B,x)+Y(B,x)\left(\sum_{m\geq 0}A_{(m)}x^{-m-1}\right). (23)

In fact, any state C𝕍0kC\in\mathbb{V}_{0}^{k} can be written as C=anBC=a_{-n}B, and using eqs. 22 and 23 we can always explicitly compute Y(C,x)Y(C,x). Summarising what we have introduced so far, we have

  • a space of states 𝕍0k\mathbb{V}_{0}^{k},

  • a vacuum vector |0k𝕍0k\ket{0}^{k}\in\mathbb{V}_{0}^{k},

  • a translation operator TEnd(𝕍0k)T\in\operatorname{End}(\mathbb{V}_{0}^{k}),

  • the state field map Y(,x)Y(\cdot,x) as in eq. 20.

This structure, together with some additional properties (see e.g. [FB01, §2.4.4]), defines a vertex algebra on 𝕍0k\mathbb{V}_{0}^{k}.

3 Construction of higher Hamiltonians

3.1 The algebra of observables

Let us introduce a set of complex numbers 𝒌={ki}i=1N\bm{k}=\{k_{i}\}_{i=1}^{N}, where N>0N\in\mathbb{Z}_{>0} and ki2k_{i}\neq-2 for all i=1,,Ni=1,\dots,N. Consider the following tensor product of vacuum Verma modules

𝕍0𝒌=𝕍0k1𝕍0kN.\mathbb{V}_{0}^{\bm{k}}=\mathbb{V}_{0}^{k_{1}}\otimes\dots\otimes\mathbb{V}_{0}^{k_{N}}. (24)

This space can be interpreted as a module over the direct sum of NN copies of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}. Let us denote by A(i)𝔰𝔩^2NA^{(i)}\in\widehat{\mathfrak{sl}}_{2}^{\oplus N} the copy of A𝔰𝔩^2A\in\widehat{\mathfrak{sl}}_{2} in the iith direct summand.

Let us denote by |0𝒌\mathbb{C}\ket{0}^{\bm{k}} the one-dimensional vacuum representation of the “positive modes” Lie subalgebra (𝔰𝔩2[t]𝗄)N𝔰𝔩^2N(\mathfrak{sl}_{2}[t]\oplus\mathbb{C}\mathsf{k})^{\oplus N}\subset\widehat{\mathfrak{sl}}_{2}^{\oplus N}, defined by (𝗄(i)ki)|0𝒌=0(\mathsf{k}^{(i)}-k_{i})\ket{0}^{\bm{k}}=0 and an(i)|0𝒌=0a_{n}^{(i)}\ket{0}^{\bm{k}}=0 for all n0n\geq 0, a𝔰𝔩2a\in\mathfrak{sl}_{2} and i=1,,Ni=1,\dots,N. Therefore 𝕍0𝒌\mathbb{V}_{0}^{\bm{k}} is the induced 𝔰𝔩^2N\widehat{\mathfrak{sl}}_{2}^{\oplus N}-module, namely

𝕍0𝒌=U(𝔰𝔩^2N)U(𝔰𝔩2[t]𝗄)N|0𝒌.\mathbb{V}_{0}^{\bm{k}}=U(\widehat{\mathfrak{sl}}_{2}^{\oplus N})\otimes_{U(\mathfrak{sl}_{2}[t]\oplus\mathbb{C}\mathsf{k})^{\oplus N}}\mathbb{C}\ket{0}^{\bm{k}}. (25)

Repeating similar arguments to those of the previous sections, we can define U𝒌(𝔰𝔩^2N)U_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) as the quotient of U(𝔰𝔩^2N)U(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) by the two-sided ideal generated by 𝗄(i)ki\mathsf{k}^{(i)}-k_{i} for all i=1,,Ni=1,\dots,N. We have the isomorphism

U𝒌(𝔰𝔩^2N)Uk1(𝔰𝔩^2)Uk2(𝔰𝔩^2)UkN(𝔰𝔩^2).U_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N})\cong U_{k_{1}}(\widehat{\mathfrak{sl}}_{2})\otimes U_{k_{2}}(\widehat{\mathfrak{sl}}_{2})\otimes\dots\otimes U_{k_{N}}(\widehat{\mathfrak{sl}}_{2}). (26)

Thanks to this fact, A(i)𝔰𝔩^2NU𝒌(𝔰𝔩^2N)A^{(i)}\in\widehat{\mathfrak{sl}}_{2}^{\oplus N}\subset U_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) can be presented as

A(i)=𝟙𝟙A𝟙𝟙,A^{(i)}=\mathds{1}\otimes\dots\otimes\mathds{1}\otimes A\otimes\mathds{1}\otimes\dots\otimes\mathds{1}, (27)

where AA is acting as the iith tensor factor. Again, we can introduce the left ideals JnNU𝒌(𝔰𝔩^2N)J_{n}^{N}\in U_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) generated by ar(i)a_{r}^{(i)} for all rnr\geq n, a𝔰𝔩2a\in\mathfrak{sl}_{2} and i=1,,Ni=1,\dots,N. Let U~𝒌(𝔰𝔩^2N)=limU𝒌(𝔰𝔩^2N)/JnN\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N})=\varprojlim U_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N})/J_{n}^{N} be the inverse limit. This space is a complete topological algebra and

U~𝒌(𝔰𝔩^2N)U~k1(𝔰𝔩^2)^^U~kN(𝔰𝔩^2),\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N})\cong\widetilde{U}_{k_{1}}(\widehat{\mathfrak{sl}}_{2})\widehat{\otimes}\cdots\widehat{\otimes}\widetilde{U}_{k_{N}}(\widehat{\mathfrak{sl}}_{2}), (28)

where ^\hat{\otimes} denotes the completed tensor product. This space U~𝒌(𝔰𝔩^2N)\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) is called the algebra of observables of the Gaudin model.

The tensor product 𝕍0𝒌\mathbb{V}_{0}^{\bm{k}} is again a vertex algebra. The state-field map Y(,x):𝕍0𝒌Hom(𝕍0𝒌,𝕍0𝒌((x)))Y(\cdot,x):\mathbb{V}_{0}^{\bm{k}}\rightarrow\operatorname{Hom}(\mathbb{V}_{0}^{\bm{k}},\mathbb{V}_{0}^{\bm{k}}((x))) is just as in section 2.4 but decorated with the extra index (i).

Let us introduce the map Δ:𝔰𝔩^2𝔰𝔩^2N\Delta:\widehat{\mathfrak{sl}}_{2}\hookrightarrow\widehat{\mathfrak{sl}}_{2}^{\oplus N}, which is the diagonal embedding of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2} into 𝔰𝔩^2N\widehat{\mathfrak{sl}}_{2}^{\oplus N}, defined as

Δx=i=1Nx(i),for all x𝔰𝔩^2.\Delta x=\sum_{i=1}^{N}x^{(i)},\qquad\text{for all }x\in\widehat{\mathfrak{sl}}_{2}. (29)

It extends to an embedding Δ:U(𝔰𝔩^2)U(𝔰𝔩^2N)U(𝔰𝔩^2)N\Delta:U(\widehat{\mathfrak{sl}}_{2})\hookrightarrow U(\widehat{\mathfrak{sl}}_{2}^{\oplus N})\cong U(\widehat{\mathfrak{sl}}_{2})^{\otimes N}. It is easy to check that

[ΔXm,ΔYn]=Δ[X,Y]n+mnδn+m,0(X|Y)i=1N𝗄(i),[\Delta X_{m},\Delta Y_{n}]=\Delta[X,Y]_{n+m}-n\delta_{n+m,0}(X|Y)\sum_{i=1}^{N}\mathsf{k}^{(i)}, (30)

where (|)(\cdot|\cdot) is the usual Killing form as in eq. 6.

Therefore Δ\Delta descends to an embedding of the quotients Δ:U|𝒌|(𝔰𝔩^2)U𝒌(𝔰𝔩^2N)\Delta:U_{|\bm{k}|}(\widehat{\mathfrak{sl}}_{2})\hookrightarrow U_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}), where |𝒌|=i=1Nki|\bm{k}|=\sum_{i=1}^{N}k_{i}, and hence of their completions

Δ:U~|𝒌|(𝔰𝔩^2)U~𝒌(𝔰𝔩^2N).\Delta:\widetilde{U}_{|\bm{k}|}(\widehat{\mathfrak{sl}}_{2})\hookrightarrow\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}). (31)

3.2 Invariant tensors on 𝔰𝔩2\mathfrak{sl}_{2}

Let {Ia}a=13\{I^{a}\}_{a=1}^{3} be a basis of 𝔰𝔩2\mathfrak{sl}_{2}, and let {Ia}a=13\{I_{a}\}_{a=1}^{3} be its dual basis with respect to the non-degenerate Killing form of eq. 6. Let fabcf^{ab}{}_{c} denote the structure constants, so that

[Ia,Ib]=fabIcc.[I^{a},I^{b}]=f^{ab}{}_{c}I^{c}. (32)

Here and in what follows we employ the summation convention on Lie algebra indices. Thanks to the non-degeneracy of the bilinear form, we may suppose our basis is chosen in such a way that

(Ia|Ib)=δab.(I^{a}|I^{b})=\delta^{ab}. (33)

By doing this, we no longer have to distinguish between upper and lower indices. The structure constants are then

fab=cfabc=i2ϵabc,f^{ab}{}_{c}=f^{abc}=i\sqrt{2}\epsilon^{abc}, (34)

where ϵabc\epsilon^{abc} is the usual Levi-Civita symbol.

(Concretely, in the defining representation we have

I1=12(0110)I2=12(0ii0)I3=12(1001).I^{1}=\frac{1}{\sqrt{2}}\begin{pmatrix}0&1\\ 1&0\end{pmatrix}\qquad I^{2}=\frac{1}{\sqrt{2}}\begin{pmatrix}0&-i\\ i&0\end{pmatrix}\qquad I^{3}=\frac{1}{\sqrt{2}}\begin{pmatrix}1&0\\ 0&-1\end{pmatrix}. (35)

Using (6), it is easy to check that eq. 33 holds.)

Recall that for any finite-dimensional Lie algebra 𝔤̊\mathring{\mathfrak{g}}, a tensor t:𝔤̊××𝔤̊t:\mathring{\mathfrak{g}}\times\dots\times\mathring{\mathfrak{g}}\rightarrow\mathbb{C} is invariant if

t([a,x],y,,z)+t(x,[a,y],,z)++t(x,y,,[a,z])=0,for all x𝔤̊,t([a,x],y,\dots,z)+t(x,[a,y],\dots,z)+\dots+t(x,y,\dots,[a,z])=0,\qquad\text{for all }x\in\mathring{\mathfrak{g}}, (36)

or equivalently, if its components ta1an:=t(Ia1,,Ian)t^{a_{1}\dots a_{n}}:=t(I^{a_{1}},\dots,I^{a_{n}}) satisfy

fca1tba2anb+fca2ta1banb++fcanta1a2bb=0,f^{ca_{1}}{}_{b}t^{ba_{2}\dots a_{n}}+f^{ca_{2}}{}_{b}t^{a_{1}b\dots a_{n}}+\dots+f^{ca_{n}}{}_{b}t^{a_{1}a_{2}\dots b}=0, (37)

where the indices take values from 11 to dim𝔤̊\dim\mathring{\mathfrak{g}}. In our case of 𝔰𝔩2\mathfrak{sl}_{2}, the ring of invariant tensors is generated by δab\delta^{ab} and fabcf^{abc}. We shall need the following syzygy relations between them:

fabcfcde=2(δaeδbdδadδbe),fabcfabd=4δcd,fabcδdefbcdδae+fcdaδbefdabδce=0.\begin{gathered}f^{abc}f^{cde}=2(\delta^{ae}\delta^{bd}-\delta^{ad}\delta^{be}),\qquad\qquad f^{abc}f^{abd}=-4\delta^{cd},\\ f^{abc}\delta^{de}-f^{bcd}\delta^{ae}+f^{cda}\delta^{be}-f^{dab}\delta^{ce}=0.\end{gathered} (38)

Note in particular the last of these, which will play a crucial role in the explicit calculations of the next sections. It can also be generalized to higher rank tensors (see e.g. [Itō93, §369 F]).

4 Quartic Hamiltonian

4.1 Meromorphic states

Let us introduce a set {z1,,zN}\{z_{1},\dots,z_{N}\} of N>0N\in\mathbb{Z}_{>0} points ziz_{i}\in\mathbb{C} in the complex plane, chosen to be pairwise distinct, zizjz_{i}\neq z_{j} whenever iji\neq j. For any element A𝔰𝔩^2A\in\widehat{\mathfrak{sl}}_{2} we introduce the 𝔰𝔩^2N\widehat{\mathfrak{sl}}_{2}^{\oplus N}-valued meromorphic functions

A(z):=i=1NA(i)zzi.A(z):=\sum_{i=1}^{N}\frac{A^{(i)}}{z-z_{i}}. (39)

We are allowed to take derivatives of such functions, which will be denoted by A(z)A^{\prime}(z) or, in general, for each p0p\geq 0,

A[p](z):=(ddz)pA(z)=i=1N(1)pp!A(i)(zzi)p+1.A^{[p]}(z):=\left(\frac{d}{dz}\right)^{p}A(z)=\sum_{i=1}^{N}(-1)^{p}p!\frac{A^{(i)}}{(z-z_{i})^{p+1}}. (40)

Considering two of these functions with different spectral parameters, we get the following commutation relations

[A[p](z),B[q](w)]=(1)p+1(p+q)![A,B](z)[A,B](w)(zw)p+q+1+k=1p(1)p+1k(pk)(p+qk)![A,B][k](z)(zw)p+q+1kk=1q(1)p+1(qk)(p+qk)![A,B][k](w)(zw)p+q+1k.\begin{split}[A^{[p]}(z),B^{[q]}(w)]=&(-1)^{p+1}(p+q)!\frac{[A,B](z)-[A,B](w)}{(z-w)^{p+q+1}}\\ &+\sum_{k=1}^{p}(-1)^{p+1-k}\binom{p}{k}(p+q-k)!\frac{[A,B]^{[k]}(z)}{(z-w)^{p+q+1-k}}\\ &-\sum_{k=1}^{q}(-1)^{p+1}\binom{q}{k}(p+q-k)!\frac{[A,B]^{[k]}(w)}{(z-w)^{p+q+1-k}}.\end{split} (41)

By taking the limit wzw\to z, we get the commutation relations for the same spectral parameter, namely

[A[p](z),B[q](z)]=[A,B][p+q+1](z).[A^{[p]}(z),B^{[q]}(z)]=-[A,B]^{[p+q+1]}(z). (42)

We see that these A[p](z)A^{[p]}(z), for A𝔰𝔩^2A\in\widehat{\mathfrak{sl}}_{2} and p0p\geq 0, span a Lie algebra of 𝔰𝔩^2N\widehat{\mathfrak{sl}}_{2}^{\oplus N}-valued meromorphic functions of zz with poles at the marked points.

It is helpful to be able to treat this as an abstract Lie algebra. Thus, let 𝔏\mathfrak{L} denote the Lie algebra over \mathbb{C} with basis consisting of Ina[p](z)I_{n}^{a[p]}(z) and 𝗄[p](z)\mathsf{k}^{[p]}(z), for nn\in\mathbb{Z}, p0p\in\mathbb{Z}_{\geq 0} and a{1,2,3}a\in\{1,2,3\} with the non-vanishing Lie brackets given by

[Ima[p](z),Inb[q](z)]=p!q!(p+q+1)!(fcabIm+nc[p+q+1](z)nδabδm+n,0𝗄[p+q+1](z)).[I^{a[p]}_{m}(z),I^{b[q]}_{n}(z)]=-\frac{p!q!}{(p+q+1)!}(f^{ab}_{c}I^{c[p+q+1]}_{m+n}(z)-n\delta^{ab}\delta_{m+n,0}\mathsf{k}^{[p+q+1]}(z)). (43)

Let 𝔏+\mathfrak{L}_{+} denote subalgebra generated by Ina[p](z)I_{n}^{a[p]}(z) for n0n\geq 0, p0p\in\mathbb{Z}_{\geq 0} and a{1,2,3}a\in\{1,2,3\}, and let

𝒱:=U(𝔏)U(𝔏+)|0\mathcal{V}:=U(\mathfrak{L})\otimes_{U(\mathfrak{L}_{+})}\mathbb{C}\ket{0} (44)

denote the module over 𝔏\mathfrak{L} induced from the trivial one-dimensional module |0\mathbb{C}\ket{0} over 𝔏+\mathfrak{L}_{+}.

We call the 𝒱\mathcal{V} the space of meromorphic states. It is again a vertex algebra, with the state-field map as in section 2.4 but decorated with extra indices. For each z{z1,,zN}z\in\mathbb{C}\setminus\{z_{1},\dots,z_{N}\}, one has the homomorphism of Lie algebras 𝔏𝔰𝔩^2N\mathfrak{L}\to\widehat{\mathfrak{sl}}_{2}^{\oplus N} given by evaluating at zz. It gives rise to a map 𝒱𝕍0𝒌\mathcal{V}\to\mathbb{V}_{0}^{\bm{k}} of vertex algebras.

There is a bi-gradation of 𝔏\mathfrak{L} in which Xn[p](z)X_{-n}^{[p]}(z) (for any X𝔰𝔩2X\in\mathfrak{sl}_{2}) has weight (n,p+1)(n,p+1) and 𝗄[p](z)\mathsf{k}^{[p]}(z) has weight (0,p+1)(0,p+1). This yields a bi-gradation of 𝒱\mathcal{V}

𝒱=n0,p0𝒱n,p.\mathcal{V}=\bigoplus_{n\geq 0,p\geq 0}\mathcal{V}_{n,p}. (45)

For each nn, let 𝒱n:=𝒱n,n\mathcal{V}_{n}:=\mathcal{V}_{n,n} denote the subspace of grade (n,n)(n,n). We call elements of 𝒱n\mathcal{V}_{n} homogeneous meromorphic states of degree nn.

4.2 Diagonal action of the zero modes of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}

There is an evident diagonal action of the Lie algebra 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2} on the 𝔏\mathfrak{L}-module 𝒱\mathcal{V}, defined in the same way as the action on 𝕍0𝒌\mathbb{V}_{0}^{\bm{k}} in eq. 29. In particular, for any X𝔰𝔩2X\in\mathfrak{sl}_{2}, the zero modes stabilize each subspace 𝒱n,p\mathcal{V}_{n,p}, namely

ΔX0:𝒱n,p𝒱n,p.\Delta X_{0}:\mathcal{V}_{n,p}\rightarrow\mathcal{V}_{n,p}. (46)

An important fact is that every state in 𝒱n\mathcal{V}_{n} properly contracted with an 𝔰𝔩2\mathfrak{sl}_{2}-invariant tensor vanishes under the diagonal action of the zero modes. This follows directly from the defining property of invariant tensors in eq. 37. Let denote with 𝒱n𝔰𝔩2\mathcal{V}_{n}^{\mathfrak{sl}_{2}} the invariant subspace. We can characterize this space for small nn:

  • for n=0n=0, 𝒱0𝔰𝔩2=𝒱0=|0\mathcal{V}_{0}^{\mathfrak{sl}_{2}}=\mathcal{V}_{0}=\mathbb{C}\ket{0}.

  • for n=1n=1, 𝒱1𝔰𝔩2={0}\mathcal{V}_{1}^{\mathfrak{sl}_{2}}=\{0\}. Indeed, elements of 𝒱1\mathcal{V}_{1} are of the form taI1a(z)|0t_{a}I^{a}_{-1}(z)\ket{0}. Such an element is in 𝒱1𝔰𝔩2\mathcal{V}^{\mathfrak{sl}_{2}}_{1} if and only if tat_{a} are the components of an 𝔰𝔩2\mathfrak{sl}_{2}-invariant tensor of rank 1. But there are no nonzero such tensors.

  • for n=2n=2, 𝒱2𝔰𝔩2\mathcal{V}_{2}^{\mathfrak{sl}_{2}} has dimension 1 and it is spanned by the state

    ς1(z)=δabI1a(z)I1b(z)|0.\varsigma_{1}(z)=\delta_{ab}I^{a}_{-1}(z)I^{b}_{-1}(z)\ket{0}. (47)
  • for n=3n=3, 𝒱3𝔰𝔩2\mathcal{V}^{\mathfrak{sl}_{2}}_{3} has dimension 2 and it is spanned by the states

    fabcI1a(z)I1b(z)I1c(z)|0=fabc12(I1a(z)I1b(z)I1b(z)I1a(z))I1c(z)|0=fabcfabdI2d(z)I1c(z)|0=4I2c(z)I1c(z)|0,\begin{split}f^{abc}I^{a}_{-1}(z)I^{b}_{-1}(z)I^{c}_{-1}(z)\ket{0}=&f^{abc}\frac{1}{2}(I^{a}_{-1}(z)I^{b}_{-1}(z)-I^{b}_{-1}(z)I^{a}_{-1}(z))I^{c}_{-1}(z)\ket{0}\\ =&f^{abc}f^{abd}I^{d\prime}_{-2}(z)I^{c}_{-1}(z)\ket{0}\\ =&-4I^{c\prime}_{-2}(z)I^{c}_{-1}(z)\ket{0},\end{split} (48)

    and

    I2c(z)I1c(z)𝗄(z)|0.I^{c}_{-2}(z)I^{c}_{-1}(z)\mathsf{k}(z)\ket{0}. (49)
  • for n=4n=4, 𝒱4𝔰𝔩2\mathcal{V}_{4}^{\mathfrak{sl}_{2}} has dimension 14. Below, we will make use of the following explicit choice of basis:

    𝗏1:=δ(abδcd)I1a(z)I1b(z)I1c(z)I1d(z),𝗏2:=fabcI2a(z)I1b(z)I1c(z),𝗏3:=I3a′′(z)I1a(z),𝗏4:=I3a(z)I1a′′(z),𝗏5:=I2a′′(z)I2a(z),𝗏6:=I3a(z)I1a(z),𝗏7:=I2a(z)I2a(z),𝗏8:=I3a(z)I1a(z)𝗄(z),𝗏9:=I2a(z)I2a(z)𝗄(z),𝗏10:=I3a(z)I1a(z)𝗄(z),𝗏11:=I3a(z)I1a(z)𝗄(z),𝗏12:=I2a(z)I2a(z)𝗄(z),𝗏13:=I3a(z)I1a(z)𝗄(z)2,𝗏14:=I2a(z)I2a(z)𝗄(z)2.\begin{gathered}\mathsf{v}_{1}:=\delta^{(ab}\delta^{cd)}I^{a}_{-1}(z)I^{b}_{-1}(z)I^{c}_{-1}(z)I^{d}_{-1}(z),\qquad\mathsf{v}_{2}:=f^{abc}I^{a}_{-2}(z)I^{b\prime}_{-1}(z)I^{c}_{-1}(z),\\ \mathsf{v}_{3}:=I^{a\prime\prime}_{-3}(z)I^{a}_{-1}(z),\qquad\mathsf{v}_{4}:=I^{a}_{-3}(z)I^{a\prime\prime}_{-1}(z),\qquad\mathsf{v}_{5}:=I^{a\prime\prime}_{-2}(z)I^{a}_{-2}(z),\\ \mathsf{v}_{6}:=I^{a\prime}_{-3}(z)I^{a\prime}_{-1}(z),\qquad\mathsf{v}_{7}:=I^{a\prime}_{-2}(z)I^{a\prime}_{-2}(z),\qquad\mathsf{v}_{8}:=I^{a}_{-3}(z)I^{a}_{-1}(z)\mathsf{k}^{\prime}(z),\\ \mathsf{v}_{9}:=I^{a}_{-2}(z)I^{a}_{-2}(z)\mathsf{k}^{\prime}(z),\qquad\mathsf{v}_{10}:=I^{a\prime}_{-3}(z)I^{a}_{-1}(z)\mathsf{k}(z),\\ \mathsf{v}_{11}:=I^{a}_{-3}(z)I^{a\prime}_{-1}(z)\mathsf{k}(z),\qquad\mathsf{v}_{12}:=I^{a\prime}_{-2}(z)I^{a}_{-2}(z)\mathsf{k}(z),\\ \mathsf{v}_{13}:=I^{a}_{-3}(z)I^{a}_{-1}(z)\mathsf{k}(z)^{2},\qquad\mathsf{v}_{14}:=I^{a}_{-2}(z)I^{a}_{-2}(z)\mathsf{k}(z)^{2}.\end{gathered} (50)

Note that to write these terms we have to choose an ordering prescription. In this work we sort level first in ascending order from left to right and after that, for a given level, we sort derivatives in descending order from left to right. For example fabcI2aI2b′′I3c=fabcI3cI2b′′I2a+terms obtained from commutationsf^{abc}I^{a}_{-2}I^{b\prime\prime}_{-2}I^{c\prime}_{-3}=f^{abc}I^{c\prime}_{-3}I^{b\prime\prime}_{-2}I^{a}_{-2}+\text{terms obtained from commutations}.

4.3 Top terms

We can see from the above construction that in the case n=2n=2 and n=4n=4, there is a particular state, that we will call top term, which is the state in 𝒱n𝔰𝔩2\mathcal{V}_{n}^{\mathfrak{sl}_{2}} that contains exactly nn generators:

δabI1a(z)I1b(z)|0,δ(abδcd)I1a(z)I1b(z)I1c(z)I1d(z)|0.\delta_{ab}I^{a}_{-1}(z)I^{b}_{-1}(z)\ket{0},\qquad\delta_{(ab}\delta_{cd)}I^{a}_{-1}(z)I^{b}_{-1}(z)I^{c}_{-1}(z)I^{d}_{-1}(z)\ket{0}. (51)

We do not have such state for n=3n=3, because we can always use the commutation relations to reduce the number of generators, as shown in eq. 48. This pattern continues, as we now describe.

Notice that the universal enveloping algebra U(𝔏)U(\mathfrak{L}) has an increasing filtration

F0U(𝔏)F1U(𝔏)U(𝔏),F_{0}U(\mathfrak{L})\subseteq F_{1}U(\mathfrak{L})\subseteq\dots\subseteq U(\mathfrak{L}), (52)

in which the generators Ina[p](z)I^{a[p]}_{n}(z) count as +1+1 and the generators 𝗄[p](z)\mathsf{k}^{[p]}(z) count as 0, cf. the commutation relations of 𝔏\mathfrak{L} in eq. 43. For example I1a(z)I2a(z)F2I^{a}_{-1}(z)I^{a\prime}_{-2}(z)\in F_{2}, and I1a(z)I2a(z)𝗄(z)F2I^{a}_{-1}(z)I^{a\prime}_{-2}(z)\mathsf{k}(z)\in F_{2} as well. It gives rise to a corresponding filtration, F0𝒱F1𝒱𝒱F_{0}\mathcal{V}\subseteq F_{1}\mathcal{V}\subseteq\dots\subseteq\mathcal{V}, on the space 𝒱\mathcal{V} of meromorphic states.

Observe that if 𝗏𝒱N\mathsf{v}\in\mathcal{V}_{N} then 𝗏FN𝒱N\mathsf{v}\in F_{N}\mathcal{V}_{N}. We see that

𝗏\displaystyle\mathsf{v} ti1iNI1i1(z)I1iN(z)|0modFN1𝒱N\displaystyle\equiv t_{i_{1}\dots i_{N}}I^{i_{1}}_{-1}(z)\dots I^{i_{N}}_{-1}(z)\ket{0}\mod F_{N-1}\mathcal{V}_{N}
t(i1iN)I1i1(z)I1iN(z)|0modFN1𝒱N,\displaystyle\equiv t_{(i_{1}\dots i_{N})}I^{i_{1}}_{-1}(z)\dots I^{i_{N}}_{-1}(z)\ket{0}\mod F_{N-1}\mathcal{V}_{N}, (53)

for some 𝔰𝔩2\mathfrak{sl}_{2} tensor ti1,,iNt_{i_{1},\dots,i_{N}}, where the brackets around the indices denote the operation of symmetrization,

t(i1,,in)=1n!σ𝒮ntσ(i1)σ(in)t_{(i_{1},\dots,i_{n})}=\frac{1}{n!}\sum_{\sigma\in\mathcal{S}_{n}}t_{\sigma(i_{1})\dots\sigma(i_{n})} (54)

(and we may symmetrize without loss of generality because the non-symmetric pieces fall into FN1F_{N-1}, as for example in eq. 48). Let us call t(i1iN)I1i1(z)I1iN(z)|0t_{(i_{1}\dots i_{N})}I^{i_{1}}_{-1}(z)\dots I^{i_{N}}_{-1}(z)\ket{0} the top term of the state 𝗏𝒱N\mathsf{v}\in\mathcal{V}_{N}.

If this state 𝗏𝒱\mathsf{v}\in\mathcal{V} is 𝔰𝔩2\mathfrak{sl}_{2}-invariant, 𝗏𝒱𝔰𝔩2\mathsf{v}\in\mathcal{V}^{\mathfrak{sl}_{2}}, then t(i1iN)t_{(i_{1}\dots i_{N})} is a symmetric invariant tensor. Nonzero such tensors exist only in even degrees, and up to rescaling they are, explicitly,

ti1,i2=δi1i2ti1,i2,i3,i4=δ(i1i2δi3i4)ti1,i2,i3,i4,i5,i6=δ(i1i2δi3i4δi5i6) \begin{gathered}t_{i_{1},i_{2}}=\delta_{i_{1}i_{2}}\\ t_{i_{1},i_{2},i_{3},i_{4}}=\delta_{(i_{1}i_{2}}\delta_{i_{3}i_{4})}\\ t_{i_{1},i_{2},i_{3},i_{4},i_{5},i_{6}}=\delta_{(i_{1}i_{2}}\delta_{i_{3}i_{4}}\delta_{i_{5}i_{6})}\\ \dots{}\end{gathered}

In what follows, our interest is in meromorphic states 𝗏𝒱𝔰𝔩2\mathsf{v}\in\mathcal{V}^{\mathfrak{sl}_{2}} that have nonzero top term (in other words states whose principal symbol has maximal degree) and that are 𝔰𝔩2\mathfrak{sl}_{2}-invariant.

4.4 Singular vectors up to twisted derivative

Let us define the twisted derivative operator of degree jj\in\mathbb{Z} with respect to the spectral parameter zz,

Dz(j)=(zj2𝗄(z)).D_{z}^{(j)}=\left(\partial_{z}-\frac{j}{2}\mathsf{k}(z)\right). (55)

Note that this operator sends 𝒱n,p𝒱n,p+1\mathcal{V}_{n,p}\to\mathcal{V}_{n,p+1} in the bigradation we introduced above.

We will say that a vector 𝗏𝒱n𝔰𝔩2\mathsf{v}\in\mathcal{V}^{\mathfrak{sl}_{2}}_{n} is singular up to twisted derivatives if for all x𝔰𝔩2x\in\mathfrak{sl}_{2} we have

Δxm𝗏=0modDz(n1)𝒱nm,n1.\Delta x_{m}\mathsf{v}=0\qquad\textup{mod}\>D^{(n-1)}_{z}\mathcal{V}_{n-m,n-1}. (56)

for all non-negative modes xmx_{m}, m0m\geq 0. This defines a subspace

𝒱nsing𝒱n𝔰𝔩2\mathcal{V}_{n}^{\text{sing}}\subset\mathcal{V}_{n}^{\mathfrak{sl}_{2}} (57)

of vectors singular up to twisted derivatives.

Proposition 4.4.1.

The space of singular vectors 𝒱2sing\mathcal{V}^{\text{sing}}_{2} is spanned by the quadratic state ς1(z)\varsigma_{1}(z) defined in (47).

 
Proof.

We need to show that

ΔIkrς1(z)=0modDz(1)𝖦kr(z),\Delta I^{r}_{k}\varsigma_{1}(z)=0\quad\mod D^{(1)}_{z}\mathsf{G}^{r}_{k}(z), (58)

for some meromorphic states 𝖦kr𝒱2k,1\mathsf{G}_{k}^{r}\in\mathcal{V}_{2-k,1}, for all k0k\geq 0 and r=1,2,3r=1,2,3. For k=0k=0 there is nothing to check since ΔI0rς1(z)=0\Delta I^{r}_{0}\varsigma_{1}(z)=0 identically, by the definition of 𝒱2𝔰𝔩2\mathcal{V}^{\mathfrak{sl}_{2}}_{2}. It is enough to check the action of the first modes I1rI^{r}_{1}, since any higher modes can be expressed in terms of their brackets, i.e. I2r=14frbc[I1b,I1c]I_{2}^{r}=-\frac{1}{4}f^{rbc}[I^{b}_{1},I^{c}_{1}] etc. From direct calculations we get that

ΔI1rς1(z)=Dz(1)𝖦1r(z),\Delta I^{r}_{1}\varsigma_{1}(z)=D^{(1)}_{z}\mathsf{G}^{r}_{1}(z), (59)

where

𝖦1r(z)=4I1r(z)|0.\mathsf{G}^{r}_{1}(z)=-4I^{r}_{-1}(z)\ket{0}. (60)
 

More non-trivially, for n=4n=4 we have the following result.

Proposition 4.4.2.

The space of singular vectors 𝒱4sing\mathcal{V}^{\text{sing}}_{4} is of dimension 7. A choice of basis is given by the state

ς3(z)=[δ(abδcd)I1a(z)I1b(z)I1c(z)I1d(z)+203fabcI2a(z)I1b(z)I1c(z)+409I3a(z)I1a′′(z)203I2a′′(z)I2a(z)+409I3a(z)I1a(z)103I2a(z)I2a(z)203I3a(z)I1a(z)𝗄(z)]|0,\begin{split}\varsigma_{3}(z)=\Big{[}&\delta_{(ab}\delta_{cd)}I^{a}_{-1}(z)I^{b}_{-1}(z)I^{c}_{-1}(z)I^{d}_{-1}(z)+\frac{20}{3}f_{abc}I^{a}_{-2}(z)I^{b\prime}_{-1}(z)I^{c}_{-1}(z)\\ &+\frac{40}{9}I^{a}_{-3}(z)I^{a\prime\prime}_{-1}(z)-\frac{20}{3}I^{a\prime\prime}_{-2}(z)I^{a}_{-2}(z)+\frac{40}{9}I^{a\prime}_{-3}(z)I^{a\prime}_{-1}(z)\\ &\phantom{+}\qquad-\frac{10}{3}I^{a\prime}_{-2}(z)I^{a\prime}_{-2}(z)-\frac{20}{3}I^{a}_{-3}(z)I^{a}_{-1}(z)\mathsf{k}^{\prime}(z)\Big{]}\ket{0},\end{split} (61)

together with the double translate state

T2(I1a′′(z)I1a(z)|0I1a(z)I1a(z)|034I1a(z)I1a(z)𝗄(z)|0)T^{2}\Big{(}I^{a\prime\prime}_{-1}(z)I^{a}_{-1}(z)\ket{0}-I^{a\prime}_{-1}(z)I^{a\prime}_{-1}(z)\ket{0}-\frac{3}{4}I^{a}_{-1}(z)I^{a}_{-1}(z)\mathsf{k}^{\prime}(z)\ket{0}\Big{)} (62)

and the following twisted derivative states

Dz(3)(I3a(z)I1a(z)|0),Dz(3)(I3a(z)I1a(z)|0),Dz(3)(I2a(z)I2a(z)|0),\displaystyle D^{(3)}_{z}\Big{(}I^{a}_{-3}(z)I^{a\prime}_{-1}(z)\ket{0}\Big{)},\quad D^{(3)}_{z}\Big{(}I^{a\prime}_{-3}(z)I^{a}_{-1}(z)\ket{0}\Big{)},\quad D^{(3)}_{z}\Big{(}I^{a\prime}_{-2}(z)I^{a}_{-2}(z)\ket{0}\Big{)}, (63)
Dz(3)(I3a(z)I1a(z)𝗄(z)|0),Dz(3)(I2a(z)I2a(z)𝗄(z)|0).\displaystyle D^{(3)}_{z}\Big{(}I^{a}_{-3}(z)I^{a}_{-1}(z)\mathsf{k}(z)\ket{0}\Big{)},\quad D^{(3)}_{z}\Big{(}I^{a}_{-2}(z)I^{a}_{-2}(z)\mathsf{k}(z)\ket{0}\Big{)}.

 
Proof.

Let s(z)𝒱4𝔰𝔩2s(z)\in\mathcal{V}_{4}^{\mathfrak{sl}_{2}}. We may write it in our basis (50),

s(z)=i=114ξi𝗏i(z)s(z)=\sum_{i=1}^{14}\xi_{i}\mathsf{v}_{i}(z) (64)

for some coefficients ξi\xi_{i}\in\mathbb{C} with i=1,,14i=1,\dots,14, and ask what conditions the requirement of being singular up to twisted derivatives, (56), places on these coefficients. It is enough to demand that

ΔIkrs(z)=0modDz(3)𝖦kr(z)\Delta I^{r}_{k}s(z)=0\mod D^{(3)}_{z}\mathsf{G}^{r}_{k}(z) (65)

for some meromorphic states 𝖦kr(z)𝒱4k,3\mathsf{G}^{r}_{k}(z)\in\mathcal{V}_{4-k,3}, for all k0k\geq 0 and r=1,2,3r=1,2,3. For zero modes there is nothing to check since ΔI0as(z)=0\Delta I^{a}_{0}s(z)=0 exactly, by definition of 𝒱4𝔰𝔩2\mathcal{V}_{4}^{\mathfrak{sl}_{2}}. It is then enough to check the action of first modes, I1rI^{r}_{1}, since any higher modes can be expressed in terms of their brackets, I2r=14frbc[I1b,I1c]I^{r}_{2}=-\frac{1}{4}f^{rbc}[I^{b}_{1},I^{c}_{1}] etc. So we are to check under what conditions

ΔI1rs(z)=Dz(3)𝖦1r(z)\Delta I^{r}_{1}s(z)=D^{(3)}_{z}\mathsf{G}^{r}_{1}(z) (66)

for some 𝖦1a(z)𝒱3,3\mathsf{G}^{a}_{1}(z)\in\mathcal{V}_{3,3}. By direct calculation, one finds that solutions exist precisely if the coefficient ξi\xi_{i} obey the relations

ξ2=203ξ1,ξ3=203ξ1ξ4+2ξ5+ξ62ξ7,ξ9=54ξ138ξ5+38ξ723ξ14,ξ10=53ξ1+32ξ432ξ532ξ6+32ξ7+ξ8,ξ11=553ξ132ξ4+32ξ532ξ7+ξ8,ξ12=152ξ134ξ534ξ743ξ14,ξ13=554ξ198ξ5+98ξ732ξ8.\begin{gathered}\xi_{2}=\frac{20}{3}\xi_{1},\qquad\xi_{3}=\frac{20}{3}\xi_{1}-\xi_{4}+2\xi_{5}+\xi_{6}-2\xi_{7},\qquad\xi_{9}=-\frac{5}{4}\xi_{1}-\frac{3}{8}\xi_{5}+\frac{3}{8}\xi_{7}-\frac{2}{3}\xi_{14},\qquad\\ \xi_{10}=\frac{5}{3}\xi_{1}+\frac{3}{2}\xi_{4}-\frac{3}{2}\xi_{5}-\frac{3}{2}\xi_{6}+\frac{3}{2}\xi_{7}+\xi_{8},\qquad\xi_{11}=\frac{55}{3}\xi_{1}-\frac{3}{2}\xi_{4}+\frac{3}{2}\xi_{5}-\frac{3}{2}\xi_{7}+\xi_{8},\qquad\\ \xi_{12}=-\frac{15}{2}\xi_{1}-\frac{3}{4}\xi_{5}-\frac{3}{4}\xi_{7}-\frac{4}{3}\xi_{14},\qquad\xi_{13}=-\frac{55}{4}\xi_{1}-\frac{9}{8}\xi_{5}+\frac{9}{8}\xi_{7}-\frac{3}{2}\xi_{8}.\end{gathered} (67)

When they do obey these relations, the required functions 𝖦1r(z)\mathsf{G}^{r}_{1}(z) are given by

𝖦1r(z)=[ρ1I1(a(z)I1r(z)I1a)(z)+frab(ρ2I2a(z)I1b(z)+ρ3I2a(z)I1b(z)+ρ4I2a(z)I1b(z)𝗄(z))+ρ5I3r(z)𝗄(z)2+ρ6I3r(z)𝗄(z)+ρ7I3r(z)𝗄(z)+ρ8I3r′′(z)]|0,\begin{split}\mathsf{G}^{r}_{1}(z)=\Big{[}&\rho_{1}I^{(a}_{-1}(z)I^{r}_{-1}(z)I^{a)}_{-1}(z)\\ &+f^{rab}\left(\rho_{2}I^{a}_{-2}(z)I^{b\prime}_{-1}(z)+\rho_{3}I^{a\prime}_{-2}(z)I^{b}_{-1}(z)+\rho_{4}I^{a}_{-2}(z)I^{b}_{-1}(z)\mathsf{k}(z)\right)\\ &+\rho_{5}I^{r}_{-3}(z)\mathsf{k}(z)^{2}+\rho_{6}I^{r}_{-3}(z)\mathsf{k}^{\prime}(z)+\rho_{7}I^{r\prime}_{-3}(z)\mathsf{k}(z)+\rho_{8}I^{r\prime\prime}_{-3}(z)\Big{]}\ket{0},\end{split} (68)

where

ρ1=83ξ1,ρ2=203ξ1ξ4+ξ5,ρ3=ξ4ξ5ξ6+2ξ7,ρ4=556ξ134ξ5+34ξ7ξ843ξ14ρ5=556ξ1+34ξ534ξ7+ξ8,ρ6=ξ4,ρ7=5ξ1ξ4+32ξ5+ξ632ξ7+83ξ14,ρ8=1009ξ143ξ523ξ7.\begin{gathered}\rho_{1}=-\frac{8}{3}\xi_{1},\qquad\rho_{2}=\frac{20}{3}\xi_{1}-\xi_{4}+\xi_{5},\qquad\rho_{3}=\xi_{4}-\xi_{5}-\xi_{6}+2\xi_{7},\\ \rho_{4}=-\frac{55}{6}\xi_{1}-\frac{3}{4}\xi_{5}+\frac{3}{4}\xi_{7}-\xi_{8}-\frac{4}{3}\xi_{14}\qquad\rho_{5}=\frac{55}{6}\xi_{1}+\frac{3}{4}\xi_{5}-\frac{3}{4}\xi_{7}+\xi_{8},\qquad\rho_{6}=\xi_{4},\\ \rho_{7}=5\xi_{1}-\xi_{4}+\frac{3}{2}\xi_{5}+\xi_{6}-\frac{3}{2}\xi_{7}+\frac{8}{3}\xi_{14},\qquad\rho_{8}=-\frac{100}{9}\xi_{1}-\frac{4}{3}\xi_{5}-\frac{2}{3}\xi_{7}.\end{gathered}

The basis reported in the proposition can be obtained by the one defined by the restrictions (67) by a change of basis. ∎

 

The proposition above is in agreement with the calculation of the quartic Hamiltonian density S4(z)S_{4}(z) (the analogue of our ς3(z)\varsigma_{3}(z)), recently presented in [KLT22]. In the present conventions, the latter is given by

S4(z)=[δ(abδcd)I1a(z)I1b(z)I1c(z)I1d(z)+203fabcI2a(z)I1b(z)I1c(z)409I3a′′(z)I1a(z)1409I2a′′(z)I2a(z)+403I3a(z)I1a(z)103I2a(z)I2a(z)+5I2a(z)I2a(z)𝗄(z)2]|0\begin{split}S_{4}(z)=\Big{[}&\delta^{(ab}\delta^{cd)}I^{a}_{-1}(z)I^{b}_{-1}(z)I^{c}_{-1}(z)I^{d}_{-1}(z)+\frac{20}{3}f^{abc}I^{a}_{-2}(z)I^{b\prime}_{-1}(z)I^{c}_{-1}(z)\\ &-\frac{40}{9}I^{a\prime\prime}_{-3}(z)I^{a}_{-1}(z)-\frac{140}{9}I^{a\prime\prime}_{-2}(z)I^{a}_{-2}(z)+\frac{40}{3}I^{a\prime}_{-3}(z)I^{a\prime}_{-1}(z)\\ &\phantom{+}\qquad-\frac{10}{3}I^{a\prime}_{-2}(z)I^{a\prime}_{-2}(z)+5I^{a}_{-2}(z)I^{a}_{-2}(z)\mathsf{k}(z)^{2}\Big{]}\ket{0}\end{split} (69)

and it does333To match conventions, note that for us δ(abδcd)=13(δabδcd+δacδbd+δadδbc)\delta_{(ab}\delta_{cd)}=\frac{1}{3}\left(\delta_{ab}\delta_{cd}+\delta_{ac}\delta_{bd}+\delta_{ad}\delta_{bc}\right) (70) and in [KLT22] the tensor called τ3abcd\tau_{3}^{abcd} is given by τ3abcd=116(δabδcd+δacδbd+δadδbc).\tau_{3}^{abcd}=\frac{1}{16}\left(\delta_{ab}\delta_{cd}+\delta_{ac}\delta_{bd}+\delta_{ad}\delta_{bc}\right). (71) We thank Sylvain Lacroix for clarifying discussions on this point. indeed lie in the space 𝒱4sing\mathcal{V}^{\text{sing}}_{4}.

4.5 Hamiltonian densities

Now, to state the main result of the paper, we need two reintroduce rational functions of two different spectral parameters, zz and ww, cf. eq. 41 and eq. 42.

Recall the Lie algebra 𝔏𝔏(z)\mathfrak{L}\equiv\mathfrak{L}_{(z)} over \mathbb{C} from section 4.1. Let 𝔏(z,w)\mathfrak{L}_{(z,w)} be the Lie algebra with generators Ina[p](z)I_{n}^{a[p]}(z), Ina[p](w)I_{n}^{a[p]}(w), 𝗄[p](z)\mathsf{k}^{[p]}(z) and 𝗄[p](w)\mathsf{k}^{[p]}(w) for a=1,2,3a=1,2,3, nn\in\mathbb{Z} and p0p\in\mathbb{Z}_{\geq 0} and commutation relations

[Ima[p](z),Inb[q](w)]=(1)p+1(p+q)!fcab(Im+nc(z)Im+nc(w))nδm+n,0δab(𝗄(z)𝗄(w))(zw)p+q+1+j=1p(1)p+1j(pj)(p+qj)!fcabIm+nc[j](z)nδn+m,0δab𝗄[j](z)(zw)p+q+1kk=1q(1)p+1(qk)(p+qk)!fcabIm+nc[j](w)nδn+m,0δab𝗄[j](w)(zw)p+q+1k,\begin{split}[I_{m}^{a[p]}(z),I_{n}^{b[q]}(w)]=&(-1)^{p+1}(p+q)!\frac{f^{ab}_{c}(I^{c}_{m+n}(z)-I^{c}_{m+n}(w))-n\delta_{m+n,0}\delta^{ab}(\mathsf{k}(z)-\mathsf{k}(w))}{(z-w)^{p+q+1}}\\ &+\sum_{j=1}^{p}(-1)^{p+1-j}\binom{p}{j}(p+q-j)!\frac{f^{ab}_{c}I^{c[j]}_{m+n}(z)-n\delta_{n+m,0}\delta^{ab}\mathsf{k}^{[j]}(z)}{(z-w)^{p+q+1-k}}\\ &-\sum_{k=1}^{q}(-1)^{p+1}\binom{q}{k}(p+q-k)!\frac{f^{ab}_{c}I^{c[j]}_{m+n}(w)-n\delta_{n+m,0}\delta^{ab}\mathsf{k}^{[j]}(w)}{(z-w)^{p+q+1-k}},\end{split} (72)

together with eq. 43 for the generators with parameter zz and the analogue with parameter ww. This Lie algebra 𝔏(z,w)\mathfrak{L}_{(z,w)} and its modules are defined over the ground ring [(zw)1]\mathbb{C}[(z-w)^{-1}] of polynomials in (zw)1(z-w)^{-1}. We have the vertex algebra 𝒱(z,w)\mathcal{V}_{(z,w)} defined analogously to eq. 44 and the two obvious embedding maps of vertex algebras 𝒱𝒱(z,w)\mathcal{V}\hookrightarrow\mathcal{V}_{(z,w)}, which we write as 𝗏𝗏(z)\mathsf{v}\mapsto\mathsf{v}(z) and 𝗏𝗏(w)\mathsf{v}\mapsto\mathsf{v}(w).

Moreover, there is a natural notion of “expanding around z=wz=w”. Namely, there is a homomorphism 𝔏(z,w)𝔏(z)((wz))\mathfrak{L}_{(z,w)}\to\mathfrak{L}_{(z)}((w-z)) of Lie algebras over [(zw)1]\mathbb{C}[(z-w)^{-1}] defined by

Ima[p](w)=Ima[p](z)+Ima[p+1](z)(wz)+12Ima[p+2](z)(wz)2+I^{a[p]}_{m}(w)=I^{a[p]}_{m}(z)+I^{a[p+1]}_{m}(z)(w-z)+\frac{1}{2}I^{a[p+2]}_{m}(z)(w-z)^{2}+\dots (73)

which is motivated by considering the Taylor expansion ιwzA(w)\iota_{w-z}A(w) of the function A(w)A(w) from eq. 39. This gives rise to a map 𝒱(z,w)𝒱(z)((wz))\mathcal{V}_{(z,w)}\to\mathcal{V}_{(z)}((w-z)). We say a state 𝗏𝒱(z,w)\mathsf{v}\in\mathcal{V}_{(z,w)} is regular at z=wz=w modulo translates if there exists Z𝒱(z,w)Z\in\mathcal{V}_{(z,w)} such that the image of 𝗏TZ\mathsf{v}-TZ under this map has no singularities in (zw)(z-w).

Recall from eqs. 47 and 61 the definitions of the quadratic state ς1𝒱2sing\varsigma_{1}\in\mathcal{V}_{2}^{\textup{sing}} and of the vector ς3𝒱4sing\varsigma_{3}\in\mathcal{V}_{4}^{\textup{sing}}, respectively.

Theorem 4.5.1.

The elements ς1𝒱2sing\varsigma_{1}\in\mathcal{V}_{2}^{\textup{sing}} and ς3𝒱4sing\varsigma_{3}\in\mathcal{V}_{4}^{\textup{sing}} obey the relations

ς1(z)(0)ς1(w)\displaystyle\varsigma_{1}(z)_{(0)}\varsigma_{1}(w) =(Dz(1)Dw(1))𝖠1,1(z,w)+T𝖡1,1(z,w),\displaystyle=(D_{z}^{(1)}-D_{w}^{(1)})\mathsf{A}_{1,1}(z,w)+T\mathsf{B}_{1,1}(z,w), (74a)
ς1(z)(0)ς3(w)\displaystyle\varsigma_{1}(z)_{(0)}\varsigma_{3}(w) =(3Dz(1)Dw(3))𝖠1,3(z,w)+T𝖡1,3(z,w),\displaystyle=(3D_{z}^{(1)}-D_{w}^{(3)})\mathsf{A}_{1,3}(z,w)+T\mathsf{B}_{1,3}(z,w), (74b)
ς3(z)(0)ς1(w)\displaystyle\varsigma_{3}(z)_{(0)}\varsigma_{1}(w) =(Dz(3)3Dw(1))𝖠3,1(z,w)+T𝖡3,1(z,w),\displaystyle=(D_{z}^{(3)}-3D_{w}^{(1)})\mathsf{A}_{3,1}(z,w)+T\mathsf{B}_{3,1}(z,w), (74c)
ς3(z)(0)ς3(w)\displaystyle\varsigma_{3}(z)_{(0)}\varsigma_{3}(w) =(3Dz(1)3Dw(3))𝖠3,3(z,w)+T𝖡3,3(z,w),\displaystyle=(3D_{z}^{(1)}-3D_{w}^{(3)})\mathsf{A}_{3,3}(z,w)+T\mathsf{B}_{3,3}(z,w), (74d)

where 𝖠i,j(z,w)\mathsf{A}_{i,j}(z,w) and 𝖡i,j(z,w)\mathsf{B}_{i,j}(z,w) are elements of 𝒱(z,w)\mathcal{V}_{(z,w)}. Moreover, 𝖠i,j(z,w)\mathsf{A}_{i,j}(z,w), i,j{1,3}i,j\in\{1,3\}, are regular at z=wz=w modulo translates.


 
Proof.

The two statements of the theorem follow from direct calculations. In particular, when m=n=1m=n=1, we get

𝖠1,1(z,w)=8zwI2a(z)I1a(w)|0,𝖡1,1(z,w)=8(zw)2I1a(z)I1a(w)|0.\mathsf{A}_{1,1}(z,w)=\frac{8}{z-w}I^{a}_{-2}(z)I^{a}_{-1}(w)\ket{0},\qquad\mathsf{B}_{1,1}(z,w)=\frac{8}{(z-w)^{2}}I^{a}_{-1}(z)I^{a}_{-1}(w)\ket{0}. (75)

We have computed 𝖠1,3(z,w)\mathsf{A}_{1,3}(z,w), 𝖡1,3(z,w)\mathsf{B}_{1,3}(z,w), 𝖠3,3(z,w)\mathsf{A}_{3,3}(z,w) and 𝖡3,3(z,w)\mathsf{B}_{3,3}(z,w) explicitly, with the aid of the computer algebra system FORM [Ver13, Kui+13]. The expressions for 𝖠1,3(z,w)\mathsf{A}_{1,3}(z,w) and 𝖡1,3(z,w)\mathsf{B}_{1,3}(z,w), are given in appendix A. The expressions for 𝖠3,3(z,w)\mathsf{A}_{3,3}(z,w) and 𝖡3,3(z,w)\mathsf{B}_{3,3}(z,w) are extremely lengthy (more that 500 terms in total), and we do not reproduce them here.

Once the expression of ς1(z)(0)ς3(w)\varsigma_{1}(z)_{(0)}\varsigma_{3}(w) is known, i.e. the functions 𝖠13(z,w)\mathsf{A}_{13}(z,w) and 𝖡13(z,w)\mathsf{B}_{13}(z,w) are found, it can be shown that the theorem is automatically satisfied for the product ς3(z)(0)ς1(w)\varsigma_{3}(z)_{(0)}\varsigma_{1}(w). This comes from the property of the nnth product between two states a,ba,b of a vertex algebra, namely

a(n)b=k=01k!(1)k+nTk(b(n+k)a).a_{(n)}b=-\sum_{k=0}^{\infty}\frac{1}{k!}(-1)^{k+n}T^{k}(b_{(n+k)}a). (76)

Therefore, by swapping two states in a 0th product, we obtain 𝖠31(z,w)=𝖠13(w,z)\mathsf{A}_{31}(z,w)=-\mathsf{A}_{13}(w,z) and a series of terms which are nothing but translates and therefore can be absorbed in the definition of 𝖡31(z,w)=𝖡13(w,z)+k=0(1)kTk(ς1(w)(k+1)ς3(z))\mathsf{B}_{31}(z,w)=-\mathsf{B}_{13}(w,z)+\sum_{k=0}^{\infty}(-1)^{k}T^{k}(\varsigma_{1}(w)_{(k+1)}\varsigma_{3}(z)).

To prove the second part of the theorem one can expand according to eq. 73 and the result follows from direct calculation. ∎

 

Having established this statement for the particular choice of quartic density ς3\varsigma_{3}, we automatically get the following property for any element of 𝒱4sing\mathcal{V}^{\text{sing}}_{4}. It is a slightly weaker property, because the condition on the twisted derivative terms on the right hand side is less rigid. As we shall see in section 4.7 below, it is sufficient for defining consistent Hamiltonians.

Corollary 4.5.1.

For any element 𝗏3𝒱4sing\mathsf{v}_{3}\in\mathcal{V}^{\text{sing}}_{4}, one has

ς1(z)(0)𝗏3(w)\displaystyle\varsigma_{1}(z)_{(0)}\mathsf{v}_{3}(w) =Dz(1)𝖠1,3\@slowromancapi@(z,w)+Dw(3)𝖠1,3\@slowromancapii@(z,w)+T𝖡1,3(z,w),\displaystyle=D_{z}^{(1)}\mathsf{A}^{\@slowromancap i@}_{1,3}(z,w)+D_{w}^{(3)}\mathsf{A}^{\@slowromancap ii@}_{1,3}(z,w)+T\mathsf{B}_{1,3}(z,w), (77)
𝗏3(z)(0)ς1(w)\displaystyle\mathsf{v}_{3}(z)_{(0)}\varsigma_{1}(w) =Dz(3)𝖠3,1\@slowromancapi@(z,w)+Dw(1)𝖠3,1\@slowromancapii@(z,w)+T𝖡3,1(z,w),\displaystyle=D_{z}^{(3)}\mathsf{A}^{\@slowromancap i@}_{3,1}(z,w)+D_{w}^{(1)}\mathsf{A}^{\@slowromancap ii@}_{3,1}(z,w)+T\mathsf{B}_{3,1}(z,w), (78)
𝗏3(z)(0)𝗏3(w)\displaystyle\mathsf{v}_{3}(z)_{(0)}\mathsf{v}_{3}(w) =Dz(3)𝖠3,3\@slowromancapi@(z,w)+Dw(3)𝖠3,3\@slowromancapii@(z,w)+T𝖡3,3(z,w).\displaystyle=D_{z}^{(3)}\mathsf{A}^{\@slowromancap i@}_{3,3}(z,w)+D_{w}^{(3)}\mathsf{A}^{\@slowromancap ii@}_{3,3}(z,w)+T\mathsf{B}_{3,3}(z,w). (79)

where 𝖠i,j\@slowromancapi@,\@slowromancapii@(z,w)\mathsf{A}^{\@slowromancap i@,\@slowromancap ii@}_{i,j}(z,w) and 𝖡i,j(z,w)\mathsf{B}_{i,j}(z,w) are elements of 𝒱(z,w)\mathcal{V}_{(z,w)}. Moreover, 𝖠ij\@slowromancapi@,\@slowromancapii@(z,w)\mathsf{A}^{\@slowromancap i@,\@slowromancap ii@}_{ij}(z,w), i,j{1,3}i,j\in\{1,3\}, are regular at z=wz=w modulo translates.


 
Proof.

We already know from theorem 4.5.1 that there exists an element, ς3(z)\varsigma_{3}(z), satisfying these relations. But we saw in proposition 4.4.2 that every element 𝗏3(z)\mathsf{v}_{3}(z) of 𝒱4sing\mathcal{V}^{\text{sing}}_{4} is proportional to ς3(z)\varsigma_{3}(z) up to the addition of certain translates and twisted derivatives.

It follows from the property (76) that if we add to ς3(z)\varsigma_{3}(z) any translate then the statement of theorem 4.5.1 still holds, the only difference being a re-definition of the states 𝖡(z,w)\mathsf{B}(z,w). And it is evident that, if we add to ς3(z)\varsigma_{3}(z) any linear combination of the twisted derivatives in eq. 63 then the resulting vector 𝗏3(z)\mathsf{v}_{3}(z) still obeys the weaker relations given above. (One might worry about introducing singularities at z=wz=w, but note that for any meromorphic states a(z)a(z) and b(z)b(z), the product a(z)(0)b(w)a(z)_{(0)}b(w) is regular at z=wz=w, as is manifest if we expand b(w)b(w) about w=zw=z in the spectral plane before taking the vertex-algebra product: a(z)(0)b(w)=a(z)(0)(b(z)+(wz)b(z)+)=a(z)(0)b(z)+(wz)a(z)(0)b(z)+a(z)_{(0)}b(w)=a(z)_{(0)}\left(b(z)+(w-z)b^{\prime}(z)+\dots\right)=a(z)_{(0)}b(z)+(w-z)a(z)_{(0)}b^{\prime}(z)+\dots.) ∎

 

4.6 Gaudin Hamiltonian

Let us define the following state at non-critical level, i.e. ki2k_{i}\neq-2,

s1(z)=12(ς1(z)+4Dz(1)ω(z))𝕍0𝒌,s_{1}(z)=\frac{1}{2}\Big{(}\varsigma_{1}(z)+4D^{(1)}_{z}\omega(z)\Big{)}\in\mathbb{V}_{0}^{\bm{k}}, (80)

where ς1(z)\varsigma_{1}(z) is now the image in 𝕍0𝒌\mathbb{V}_{0}^{\bm{k}} of the density defined in (47) and where

ω(z):=i=1N1zzi(12(ki+2)I1a(i)I1a(i)|0𝒌),\omega(z):=\sum_{i=1}^{N}\frac{1}{z-z_{i}}\left(\frac{1}{2(k_{i}+2)}I^{a(i)}_{-1}I^{a(i)}_{-1}\ket{0}^{\bm{k}}\right), (81)

the term in the brackets being the Segal-Sugawara vector at site ii.

It is possible to show (see [LVY20]), that the operator (s1(z))(0)(s_{1}(z))_{(0)} is the image in U~𝒌(𝔰𝔩^2N)\tilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) of

i=1N𝒞(i)(zzi)2+i=1NizziU~(𝔤N),\sum_{i=1}^{N}\frac{\mathscr{C}^{(i)}}{(z-z_{i})^{2}}+\sum_{i=1}^{N}\frac{\mathscr{H}_{i}}{z-z_{i}}\in\tilde{U}(\mathfrak{g}^{\oplus N}), (82)

where

𝒞(i):=(𝗄(i)+2)𝖽(i)+12I0a(i)I0a(i)+n>0Ina(i)Ina(i)\mathscr{C}^{(i)}:=(\mathsf{k}^{(i)}+2)\mathsf{d}^{(i)}+\frac{1}{2}I^{a(i)}_{0}I^{a(i)}_{0}+\sum_{n>0}I^{a(i)}_{-n}I^{a(i)}_{-n} (83)

is the iith copy of the quadratic Casimir operator of 𝔤\mathfrak{g} in U~(𝔤N)\tilde{U}(\mathfrak{g}^{\oplus N}) and i\mathscr{H}_{i} are the Hamiltonians in (1).

Theorem 4.6.1.

Given the images in 𝕍0𝐤\mathbb{V}_{0}^{\bm{k}} of the densities ςi\varsigma_{i}, i{1,3}i\in\{1,3\}, we have

s1(z)(0)ςi(w)=12Dw(1)𝖠1,i(z,w)+T(12𝖡1,i(z,w)+2Dz(1)ςi(w)zw),s_{1}(z)_{(0)}\varsigma_{i}(w)=-\frac{1}{2}D^{(1)}_{w}\mathsf{A}_{1,i}(z,w)+T\Big{(}\frac{1}{2}\mathsf{B}_{1,i}(z,w)+2D^{(1)}_{z}\frac{\varsigma_{i}(w)}{z-w}\Big{)}, (84)

with 𝖠1,i(z,w)\mathsf{A}_{1,i}(z,w) and 𝖡1,i(z,w)\mathsf{B}_{1,i}(z,w) being the images in 𝕍0𝐤\mathbb{V}_{0}^{\bm{k}} of the meromorphic states in theorem 4.5.1.


 
Proof.

The result follows from direct calculations, using the definitions of 𝖠1,1\mathsf{A}_{1,1}, 𝖡1,1\mathsf{B}_{1,1}, 𝖠1,3\mathsf{A}_{1,3}, 𝖡1,3\mathsf{B}_{1,3} in (75) and appendix A, respectively. ∎

 

As we will see in the next section, this requirement is sufficient to ensure the commutativity of local Hamiltonians, arising from the densities ς1(z)\varsigma_{1}(z) and ς3(z)\varsigma_{3}(z), with the usual quadratic Gaudin Hamiltonians which define the model.

4.7 Commuting Hamiltonians

In this section, we will simply recall the ideas presented in [LVY18]. Consider two states X,Y𝕍0𝒌X,Y\in\mathbb{V}_{0}^{\bm{k}} and their formal zero modes X(0),Y(0)U~𝒌(𝔰𝔩^2N)X_{(0)},Y_{(0)}\in\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}). It is possible to show the following vertex algebra identity (see e.g. [FB01])

[X(0),Y(0)]=(X(0)Y)(0).[X_{(0)},Y_{(0)}]=(X_{(0)}Y)_{(0)}. (85)

This means that if one is able to find a family of operators whose 0th product vanishes (or that can be expressed as a translation, since (TZ)(0)=0(TZ)_{(0)}=0 by definition), then their formal zero modes form a commutative subalgebra of the algebra of observables U~𝒌(𝔰𝔩^2N)\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}^{\oplus N}_{2}).

The meromorphic function which is obtained by acting with 𝗄(z)\mathsf{k}(z) on the module 𝕍0𝒌\mathbb{V}_{0}^{\bm{k}}, i.e. setting the central elements to numbers,

k(z)=i=1Nkizzi,k(z)=\sum_{i=1}^{N}\frac{k_{i}}{z-z_{i}}, (86)

has a special role and it is called the twist function of the model. Let us define also

𝒫(z):=j=1N(zzj)kj.\mathscr{P}(z)\vcentcolon=\prod_{j=1}^{N}(z-z_{j})^{k_{j}}. (87)

The function 𝒫1/2\mathscr{P}^{1/2} is multi-valued. It becomes single-valued on a certain multi-sheeted cover of {z1,,zN}\mathbb{C}\setminus\{z_{1},\dots,z_{N}\}. Let γ\gamma be any closed contour in this cover. For example, γ\gamma could be the lift to this cover of a Pochhammer contour in {z1,,zN}\mathbb{C}\setminus\{z_{1},\dots,z_{N}\} around any two of the marked points. Then 𝒫n/2\mathscr{P}^{n/2}, for any integer nn, is single-valued along γ\gamma, and one can introduce the integral

γ𝒫(z)n/2f(z)dz\int_{\gamma}\mathscr{P}(z)^{-n/2}f(z)\text{d}z (88)

which has the fundamental property that, for any meromorphic function f(z)f(z) which is non-singular along γ\gamma

γ𝒫(z)n/2(ddzn2k(z))f(z)dz=γddz(𝒫n/2f(z))dz=0,\int_{\gamma}\mathscr{P}(z)^{-n/2}\left(\frac{d}{dz}-\frac{n}{2}k(z)\right)f(z)\text{d}z=\int_{\gamma}\frac{d}{dz}(\mathscr{P}^{-n/2}f(z))\text{d}z=0, (89)

cf. eq. 55. (See e.g. [LVY18] for the details.)

Let us now define the following object in U~𝒌(𝔰𝔩^2N)\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}),

Qnγ=γ𝒫(z)n/2ςn(z)(0)dz,Q_{n}^{\gamma}=\int_{\gamma}\mathscr{P}(z)^{-n/2}\varsigma_{n}(z)_{(0)}\text{d}z, (90)

for n=1,3n=1,3, where ςn(z)\varsigma_{n}(z) are now the images in 𝕍0𝒌\mathbb{V}_{0}^{\bm{k}} of the densities we have defined in the previous section.

Proposition 4.7.1.

The operators QnγU~𝐤(𝔰𝔩^2N)Q_{n}^{\gamma}\in\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) commute amongst themselves, with the generators of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}, and with the quadratic Hamiltonians i\mathscr{H}_{i}.


 
Proof.

We can use eq. 85 to compute

[Qmγ,Qnη]=γη𝒫(z)m/2𝒫(w)n/2[ςm(z)(0),ςn(w)(0)]dzdw=γη𝒫(z)m/2𝒫(w)n/2(ςm(z)(0)ςn(w))(0)dzdw,\begin{split}[Q_{m}^{\gamma},Q_{n}^{\eta}]=&\int_{\gamma}\int_{\eta}\mathscr{P}(z)^{-m/2}\mathscr{P}(w)^{-n/2}[\varsigma_{m}(z)_{(0)},\varsigma_{n}(w)_{(0)}]\text{d}z\text{d}w\\ =&\int_{\gamma}\int_{\eta}\mathscr{P}(z)^{-m/2}\mathscr{P}(w)^{-n/2}(\varsigma_{m}(z)_{(0)}\varsigma_{n}(w))_{(0)}\text{d}z\text{d}w,\end{split} (91)

but we know from theorem 4.5.1 that the 0th product between those states can be expressed as a sum of twisted derivatives and translations. It is now straightforward to check that the result of the commutator is zero: on one side because (TX)(0)=0(TX)_{(0)}=0 for every state X𝕍0𝒌X\in\mathbb{V}_{0}^{\bm{k}}, on the other because of the property (89).

To prove the second statement, we recall that Ina=(I1a|0𝒌)(n)I^{a}_{n}=(I^{a}_{-1}\ket{0}^{\bm{k}})_{(n)} and the general property that for any two states A,BA,B of a vertex algebra we have

[A(m),B(n)]=k0(mk)(A(k)B)(m+nk),m,n[A_{(m)},B_{(n)}]=\sum_{k\geq 0}\binom{m}{k}(A_{(k)}B)_{(m+n-k)},\qquad m,n\in\mathbb{Z} (92)

where

(mk)=m(m1)(mk+1)k!,k>0;(m0)=1.\binom{m}{k}=\frac{m(m-1)\dots(m-k+1)}{k!},\quad k\in\mathbb{Z}_{>0};\qquad\binom{m}{0}=1. (93)

Note that this represents a generalization of eq. 85, see e.g. [VY17, FB01]. At this point we can consider the generators {Ina}a=13\{I_{n}^{a}\}_{a=1}^{3} with nn\in\mathbb{Z} and we get

[Ina,Qmγ]=γ𝒫(z)m/2[Ina,ςm(z)(0)]dz=γk0(nk)𝒫(z)m/2(Ikaςm(z))(nk)dz.\begin{split}[I^{a}_{n},Q_{m}^{\gamma}]=&\int_{\gamma}\mathscr{P}(z)^{-m/2}[I^{a}_{n},\varsigma_{m}(z)_{(0)}]\text{d}z\\ =&\int_{\gamma}\sum_{k\geq 0}\binom{n}{k}\mathscr{P}(z)^{-m/2}(I^{a}_{k}\varsigma_{m}(z))_{(n-k)}\text{d}z.\end{split} (94)

It is now straightforward to check that the result is zero, by using property (56), described in the relevant cases in propositions 4.4.1 and 4.4.2, and the property (89). ∎

 

4.7.1 Fourier modes

Even though the operators QmγQ^{\gamma}_{m} we have just defined have all the right characteristics to be well-defined Hamiltonians as pointed out in proposition 4.7.1, there is one last subtlety about these objects, related to the fact that we want their action on highest weight modules to be diagonalisable. In fact, considering Xn(i)U~(𝔰𝔩^2N)X^{(i)}_{n}\in\widetilde{U}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) such that degXn(i)=n\deg{X^{(i)}_{n}}=n and setting deg(|0𝒌)=0\deg(\ket{0}^{\bm{k}})=0, induces a 0\mathbb{Z}_{\leq 0}-gradation on the product of vacuum Verma modules, called the homogeneous gradation. Therefore if X𝕍0𝒌X\in\mathbb{V}_{0}^{\bm{k}} with degree deg(X)=k\deg(X)=k, the degree of its modes is deg(X(m))=1+k+m\deg(X_{(m)})=1+k+m. The objects we have constructed ςn(z)\varsigma_{n}(z), by definition, have deg(ςn(z))=n1\deg(\varsigma_{n}(z))=-n-1, therefore deg(ςn(z)(0))=n\deg(\varsigma_{n}(z)_{(0)})=-n.

This shows that in the homogeneous gradation these operators have degree 0\neq 0: this means that if we consider a module over U(𝔰𝔩^2N)U(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) which has a trivial subspace of grade nn for large nn, then the operator γn𝒫(z)n/2ςn(z)(0)𝑑z\int_{\gamma_{n}}\mathscr{P}(z)^{-n/2}\varsigma_{n}(z)_{(0)}dz has a non-zero eigenvalue.

A way to overcome this issue is to consider the notion of Fourier mode X[n]U~(𝔰𝔩^2N)X_{[n]}\in\widetilde{U}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) of the state X𝕍0𝒌X\in\mathbb{V}_{0}^{\bm{k}}: they have the property that we are looking for, namely deg(X[n])=n\deg(X_{[n]})=n. Additionally, they satisfy a similar relation to (85),

[X[0],Y[0]]=(X(0)Y)[0],[X_{[0]},Y_{[0]}]=(X_{(0)}Y)_{[0]}, (95)

with (TX)[0]=0(TX)_{[0]}=0. One has (x1(i)|0𝒌)[n]=xn(x^{(i)}_{-1}\ket{0}^{\bm{k}})_{[n]}=x_{n} for x𝔰𝔩2x\in\mathfrak{sl}_{2} and it is possible to show that the following recursive formula holds:

(A(n)B)[m]=((Af(t))B)[m]+k>0ckA[k]B[k+m]+k0ckB[k+m]A[k],(A_{(-n)}B)_{[m]}=((A\otimes f(t))B)_{[m]}+\sum_{k>0}c_{k}A_{[-k]}B_{[k+m]}+\sum_{k\leq 0}c_{k}B_{[k+m]}A_{[-k]}, (96)

where f(t)f(t) is the Taylor series in t:=uvt:=u-v given by

f=1(n1)!(u)n1(1uvιuveveuev).f=\frac{1}{(n-1)!}(-\partial_{u})^{n-1}\Big{(}\frac{1}{u-v}-\iota_{u-v}\frac{e^{v}}{e^{u}-e^{v}}\Big{)}. (97)

and where the coefficients ckc_{k} are defined by the requirement that k>0ck(zw)k\sum_{k>0}c_{k}(\frac{z}{w})^{k} and k0ck(zw)k-\sum_{k\leq 0}c_{k}(\frac{z}{w})^{k} are the expansions, for |z|<|w||z|<|w| and |w|<|z||w|<|z| respectively, of the function

1(n1)!(ww)n1zwz.\frac{1}{(n-1)!}(-w\partial_{w})^{n-1}\frac{z}{w-z}. (98)

The first relevant examples are

(A(1)B)[m]=12(A(0)B)[m]112(A(1)B)[m]++k>0A[k]B[k+m]+k0B[k+m]A[k]\displaystyle(A_{(-1)}B)_{[m]}=\begin{aligned} \frac{1}{2}(A_{(0)}B)_{[m]}-\frac{1}{12}(A_{(1)}&B)_{[m]}+\dots\\ &+\sum_{k>0}A_{[-k]}B_{[k+m]}+\sum_{k\leq 0}B_{[k+m]}A_{[-k]}\end{aligned} (99a)
(A(2)B)[m]=112(A(0)B)[m]1240(A(2)B)[m]++k>0kA[k]B[k+m]+k0(k)B[k+m]A[k]\displaystyle(A_{(-2)}B)_{[m]}=\begin{aligned} \frac{1}{12}(A_{(0)}B)_{[m]}-\frac{1}{240}&(A_{(2)}B)_{[m]}+\dots\\ +&\sum_{k>0}kA_{[-k]}B_{[k+m]}+\sum_{k\leq 0}(-k)B_{[k+m]}A_{[-k]}\end{aligned} (99b)

where A,B𝕍0𝒌A,B\in\mathbb{V}_{0}^{\bm{k}}. These formulae are the Fourier-analog of the normal ordered product formula eq. 23, and they allow one to compute by recursion the Fourier modes of a general state X𝕍0𝒌X\in\mathbb{V}_{0}^{\bm{k}}.

Property (95) means that if the vertex algebra 0th product of XX and YY vanishes their Fourier zero-modes generate a commutative subalgebra, in homogeneous degree zero, of U~(𝔰𝔩^2N)\widetilde{U}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}). We let

Q^nγ=γ𝒫(z)n/2ςn(z)[0]dz,\widehat{Q}_{n}^{\gamma}=\int_{\gamma}\mathscr{P}(z)^{-n/2}\varsigma_{n}(z)_{[0]}\text{d}z, (100)

for n=1,3n=1,3. By the same logic as for proposition 4.7.1, we have the following.

Proposition 4.7.2.

The operators Q^nγU~𝐤(𝔰𝔩^2N)\widehat{Q}_{n}^{\gamma}\in\widetilde{U}_{\bm{k}}(\widehat{\mathfrak{sl}}_{2}^{\oplus N}) have homogeneous degree 0 and they commute amongst themselves, with the generators of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}, and with the quadratic Hamiltonians i\mathscr{H}_{i}.

5 Higher local Hamiltonians to sub-leading order

In the previous section, we have shown that it is possible to define quartic local Hamiltonians which commute among themselves and with the quadratic ones, together with the generators of 𝔰𝔩2\mathfrak{sl}_{2}. Following the same steps, one could in principle try to construct the Hamiltonians for every exponent of 𝔰𝔩^2\widehat{\mathfrak{sl}}_{2}. However, the direct calculation (already lengthy in the case of ς3(z)(0)ς3(w)\varsigma_{3}(z)_{(0)}\varsigma_{3}(w), as we noted above) becomes computationally very demanding. What we shall do instead is work to next-to-leading order in a certain semiclassical limit, which will at least give a strong consistency check on the existence of the expected Hamiltonian densities.

Thus, let us introduce a formal parameter \hbar and work over [[]]\mathbb{C}[[\hbar]]. So in particular, all vector spaces above are now to be regarded as modules over [[]]\mathbb{C}[[\hbar]]. We consider the following rescaled generators:

IaI~a\displaystyle I^{a}\longrightarrow\tilde{I}^{a} :=Ia,\displaystyle:=\hbar I^{a},
𝗄𝗄~\displaystyle\mathsf{k}\longrightarrow\widetilde{\mathsf{k}} :=𝗄.\displaystyle:=\mathsf{k}. (101)

With this re-scaling the commutation relations become

[I~ma[p](z),I~nb[q](z)]=p!q!(p+q+1)!(fcabI~m+nc[p+q+1](z)2nδabδm+n,0𝗄~[p+q+1](z)).[\tilde{I}^{a[p]}_{m}(z),\tilde{I}^{b[q]}_{n}(z)]=-\frac{p!q!}{(p+q+1)!}(\hbar f^{ab}_{c}\tilde{I}^{c[p+q+1]}_{m+n}(z)-\hbar^{2}n\delta^{ab}\delta_{m+n,0}\widetilde{\mathsf{k}}^{[p+q+1]}(z)). (102)

At this point we can identify the various quantum corrections by their \hbar dependence and work grade by grade. We shall work at next-to-leading order, i.e. the next order beyond the usual semi-classical calculation of Poisson brackets. Thus, we consider the densities of Hamiltonians up to and including the leading quantum corrections at order \hbar, and we compute commutators up to and including terms of order 2\hbar^{2}.

Remark.

It is worth remarking that the classical limit, section 5, that we take is not quite the standard one which recovers the usual classical Gaudin model (cf. [KLT22] for a very complete discussion of that limit), because for us the central charges remain O(1)O(1) in the limit. From our present perspective this is simply for computational convenience – this limit produces the simplest possible non-trivial check, and had we rescaled the central charges there would be more potential quantum correction terms already at next-to-leading order. But it might be interesting to consider this classical limit in its own right.

Having introduced the formal parameter \hbar, there is a gradation on the enveloping algebras in which I~a\tilde{I}^{a} and \hbar have grade one and 𝗄~\widetilde{\mathsf{k}} has grade zero. Recall that 𝒱n\mathcal{V}_{n} denotes the space of homogeneous meromorphic states of degree nn, eq. 45. Let now 𝒱~n𝒱n\widetilde{\mathcal{V}}_{n}\subset\mathcal{V}_{n} denote the subspace consisting of states that are also of grade nn in this new gradation (i.e. which are sums of terms having exactly nn factors of I~a\tilde{I}^{a} or \hbar).

Proposition 5.0.1.

Modulo terms of order 2\hbar^{2} there is, up to rescaling, exactly one state ς~2n1𝒱~2n𝔰𝔩2\tilde{\varsigma}_{2n-1}\in\widetilde{\mathcal{V}}_{2n}^{\mathfrak{sl}_{2}}, n1n\in\mathbb{Z}_{\geq 1}, such that, for all x𝔰𝔩2x\in\mathfrak{sl}_{2} and m0m\in\mathbb{Z}_{\geq 0},

Δxmς~2n1=0mod2Dz(2n1)𝒱2nm,2n1,mod3𝒱2nm,2n.\Delta x_{m}\tilde{\varsigma}_{2n-1}=0\qquad\textup{mod}\>\hbar^{2}D^{(2n-1)}_{z}\mathcal{V}_{2n-m,2n-1},\quad\textup{mod}\>\hbar^{3}\mathcal{V}_{2n-m,2n}. (103)

Explicitly, modulo terms in 2𝒱2n\hbar^{2}\mathcal{V}_{2n},

ς~(z)2n1=ti1,,i2nI~1i1(z)I~1i2(z)I~1i2n(z)|0+n(2n+1)(2n2)(2n1)ti1,,i2n4fabcI~2a(z)I~1b(z)I~1(c(z)I~1i1(z)I~1i2n4)(z)|0.\begin{split}\tilde{\varsigma}&{}_{2n-1}(z)=t_{i_{1},\dots,i_{2n}}\tilde{I}^{i_{1}}_{-1}(z)\tilde{I}^{i_{2}}_{-1}(z)\cdots\tilde{I}^{i_{2n}}_{-1}(z)\ket{0}\\ &\qquad+\hbar\frac{n(2n+1)(2n-2)}{(2n-1)}t_{i_{1},\dots,i_{2n-4}}f^{abc}\tilde{I}^{a}_{-2}(z)\tilde{I}^{b\prime}_{-1}(z)\tilde{I}^{(c}_{-1}(z)\tilde{I}^{i_{1}}_{-1}(z)\cdots\tilde{I}^{i_{2n-4})}_{-1}(z)\ket{0}.\end{split} (104)

 
Proof.

Given the basis {Ir}r=13\{I^{r}\}_{r=1}^{3} for 𝔰𝔩2\mathfrak{sl}_{2}, we need to show that there exist a function 𝖦m(z)\mathsf{G}_{m}(z) such that

ΔImrς~2n1(z)=Dz(2n1)𝖦mr(z)+𝒪(3).\Delta I^{r}_{m}\tilde{\varsigma}_{2n-1}(z)=D^{(2n-1)}_{z}\mathsf{G}^{r}_{m}(z)+\mathcal{O}(\hbar^{3}). (105)

For m=0m=0, this is always true thanks to the invariance of the tensor (37). For the same reasons explained in the previous section, the only relevant check that one needs to make is the one for m=1m=1. From direct calculation, we get

𝖦1r(z)=4n(2n1)2ti1,,i2n2I~1(r(z)I~1i1(z)I~1i2n2)(z)|0.\mathsf{G}^{r}_{1}(z)=-\frac{4n}{(2n-1)}\hbar^{2}t_{i_{1},\dots,i_{2n-2}}\tilde{I}^{(r}_{-1}(z)\tilde{I}^{i_{1}}_{-1}(z)\cdots\tilde{I}^{i_{2n-2})}_{-1}(z)\ket{0}. (106)

Note that this result is in accordance with the exact ones obtained for the quadratic (n=1n=1) and the leading order of quartic (n=2n=2) states, cf. eqs. 60 and 68. ∎

 

(Observe that this is consistent with proposition 4.4.2 because the vectors in eqs. 62 and 63 all come with factors of 2\hbar^{2} in the limit.)

We can now state the following theorem

Theorem 5.0.1.

Let ς~n(z)\tilde{\varsigma}_{n}(z) be as in eq. 104 above, for all odd m,n1m,n\in\mathbb{Z}_{\geq 1}. We have

ς~m(z)(0)ς~n(w)=(nDz(m)mDw(n))𝖠m,n(z,w)+T𝖡m,n(z,w)+𝒪(3),\begin{split}\tilde{\varsigma}_{m}(z)_{(0)}\tilde{\varsigma}_{n}(w)=(nD^{(m)}_{z}-mD^{(n)}_{w})\mathsf{A}_{m,n}(z,w)+T\mathsf{B}_{m,n}(z,w)+\mathcal{O}(\hbar^{3}),\end{split} (107)

where 𝖠m,n(z,w),𝖡m,n(z,w)𝒱~(z,w)\mathsf{A}_{m,n}(z,w),\mathsf{B}_{m,n}(z,w)\in\widetilde{\mathcal{V}}_{(z,w)} are given by

𝖠m,n(z,w)=ζm,n2zwti1,,im1tj1,,jn1×T(I~1(a(z)I~1i1(z)I~1im1)(z))I~1(a(w)I~1j1(w)I~1jn1)(w)|0+𝒪(3),\begin{split}\mathsf{A}_{m,n}(z,w)=&\zeta_{m,n}\frac{\hbar^{2}}{z-w}t_{i_{1},\dots,i_{m-1}}t_{j_{1},\dots,j_{n-1}}\\ &\times T\Big{(}\tilde{I}_{-1}^{(a}(z)\tilde{I}_{-1}^{i_{1}}(z)\dots\tilde{I}_{-1}^{i_{m-1})}(z)\Big{)}\tilde{I}_{-1}^{(a}(w)\tilde{I}_{-1}^{j_{1}}(w)\dots\tilde{I}_{-1}^{j_{n-1})}(w)\ket{0}+\mathcal{O}(\hbar^{3}),\end{split} (108)
𝖡m,n(z,w)=ζm,n2(zw)2ti1,,im1tj1,,jn1×I~1(a(z)I~1i1(z)I~1im1)(z)I~1(a(w)I~1j1(w)I~1jn1)(w)|0+𝒪(3),\begin{split}\mathsf{B}_{m,n}(z,w)=&\zeta_{m,n}\frac{\hbar^{2}}{(z-w)^{2}}t_{i_{1},\dots,i_{m-1}}t_{j_{1},\dots,j_{n-1}}\\ &\times\tilde{I}_{-1}^{(a}(z)\tilde{I}_{-1}^{i_{1}}(z)\dots\tilde{I}_{-1}^{i_{m-1})}(z)\tilde{I}_{-1}^{(a}(w)\tilde{I}_{-1}^{j_{1}}(w)\dots\tilde{I}_{-1}^{j_{n-1})}(w)\ket{0}+\mathcal{O}(\hbar^{3}),\end{split} (109)

where

ζm,n=2(m+1)(n+1)mn.\zeta_{m,n}=\frac{2(m+1)(n+1)}{mn}. (110)

Moreover, 𝖠m,n(z,w)\mathsf{A}_{m,n}(z,w) is a regular function for z=wz=w modulo translates and modulo terms proportional to 3\hbar^{3}.


 
Proof.

The 0th mode of ςm(z)𝒱~m+1\varsigma_{m}(z)\in\widetilde{\mathcal{V}}_{m+1} can be inferred from a purely combinatorial reasoning. Let us start with the top term ς~mTT(z)\tilde{\varsigma}_{m}^{\textup{TT}}(z) of (104). We know that computing the 0th mode, the number of generators in any term we get does not change, but the result will be a state of total depth mm and therefore we know there must be at least one generator with a positive mode. We can also use the fact that we are working at leading order in \hbar, therefore we could get at least one I~0\tilde{I}_{0}, one I~1\tilde{I}_{1} or a term with two I~0\tilde{I}_{0}, every other term will be 𝒪(3)\mathcal{O}(\hbar^{3}). The only thing to fix is the combinatorial factor describing the number of possible ways to write such terms. The result is

ς~mTT(z)(0)=ti1,,im+1[(m+1)!(m1)!I~2i1(z)I~1i2(z)I~1im(z)I~1im+1(z)+(m+1)!(m)!I~1i1(z)I~1im(z)I~0im+1(z)+(m+1)!(m1)2(m1)!I~2i1(z)I~1i2(z)I~1im1(z)I~0im(z)I~0im+1(z)]+𝒪(3).\begin{split}\tilde{\varsigma}_{m}^{\textup{TT}}(z)_{(0)}=t_{i_{1},\dots,i_{m+1}}\Big{[}&\frac{(m+1)!}{(m-1)!}\tilde{I}^{i_{1}}_{-2}(z)\tilde{I}^{i_{2}}_{-1}(z)\dots\tilde{I}^{i_{m}}_{-1}(z)\tilde{I}^{i_{m+1}}_{1}(z)\\ &+\frac{(m+1)!}{(m)!}\tilde{I}^{i_{1}}_{-1}(z)\dots\tilde{I}^{i_{m}}_{-1}(z)\tilde{I}^{i_{m+1}}_{0}(z)\\ &+\frac{(m+1)!(m-1)}{2(m-1)!}\tilde{I}^{i_{1}}_{-2}(z)\tilde{I}^{i_{2}}_{-1}(z)\dots\tilde{I}^{i_{m-1}}_{-1}(z)\tilde{I}^{i_{m}}_{0}(z)\tilde{I}^{i_{m+1}}_{0}(z)\Big{]}+\mathcal{O}(\hbar^{3}).\end{split}

With similar arguments we can compute the 0th mode of the correction term ς~mC(z)\tilde{\varsigma}_{m}^{\textup{C}}(z) of (104), the result reads

ς~mC(z)(0)=ξ(m2)!ti1,,im3fabcI~2a(z)I~1(b(z)I~1i1(z)I~1im3)(z)I~0c(z)+ξ2(m2)(m3)!ti1,,im3fabcI~2a(z)I~1b(z)I~1(i1(z)I~1im3)(z)I~0c(z)+ξ(m2)(m3)(m3)!ti1,,im5fabcI~2a(z)I~1b(z)×I~1(c(z)I~1d(z)I~1i1(z)I~1im5)I~0d(z)ξ(m2)(m3)(m3)!ti1,,im5fabcI~2d(z)I~1a(z)×I~1(b(z)I~1d(z)I~1i1(z)I~1im5)(z)I~0c(z)+𝒪(3)\begin{split}\tilde{\varsigma}_{m}^{\textup{C}}(z)_{(0)}=&-\hbar\xi(m-2)!t_{i_{1},\dots,i_{m-3}}f^{abc}\tilde{I}_{-2}^{a}(z)\tilde{I}_{-1}^{(b}(z)\tilde{I}_{-1}^{i_{1}}(z)\dots\tilde{I}_{-1}^{i_{m-3})}(z)\tilde{I}^{c}_{0}(z)\\ &+\hbar\xi 2(m-2)(m-3)!t_{i_{1},\dots,i_{m-3}}f^{abc}\tilde{I}_{-2}^{a}(z)\tilde{I}_{-1}^{b\prime}(z)\tilde{I}_{-1}^{(i_{1}}(z)\dots\tilde{I}_{-1}^{i_{m-3})}(z)\tilde{I}_{0}^{c}(z)\\ &+\hbar\xi(m-2)(m-3)(m-3)!t_{i_{1},\dots,i_{m-5}}f^{abc}\tilde{I}_{-2}^{a}(z)\tilde{I}_{-1}^{b\prime}(z)\\ &\hskip 128.0374pt\times\tilde{I}_{-1}^{(c}(z)\tilde{I}_{-1}^{d}(z)\tilde{I}^{i_{1}}_{-1}(z)\dots\tilde{I}_{-1}^{i_{m-5})}\tilde{I}_{0}^{d}(z)\\ &-\hbar\xi(m-2)(m-3)(m-3)!t_{i_{1},\dots,i_{m-5}}f^{abc}\tilde{I}_{-2}^{d}(z)\tilde{I}_{-1}^{a\prime}(z)\\ &\hskip 128.0374pt\times\tilde{I}_{-1}^{(b}(z)\tilde{I}_{-1}^{d}(z)\tilde{I}^{i_{1}}_{-1}(z)\dots\tilde{I}_{-1}^{i_{m-5})}(z)\tilde{I}_{0}^{c}(z)+\mathcal{O}(\hbar^{3})\\ \end{split}

where ξ=(m+2)(m+1)(m1)2m\xi=\frac{(m+2)(m+1)(m-1)}{2m}. At this point, acting with what we have obtained on ςn(w)\varsigma_{n}(w) and using repeatedly the commutation relations (41), we obtain eq. 107. ∎

 

Appendix A Full expression for 𝖠1,3(z,w)\mathsf{A}_{1,3}(z,w) and 𝖡1,3(z,w)\mathsf{B}_{1,3}(z,w)

The explicit expressions for 𝖠1,3(z,w)\mathsf{A}_{1,3}(z,w) and 𝖡1,3(z,w)\mathsf{B}_{1,3}(z,w) in 𝒱(z,w)\mathcal{V}_{(z,w)} obtained by direct calculations are

𝖠1,3(z,w)=[831zwI2a(z)I1(a(w)I1b(w)I1b)(w)+fabc(8091zwI2a(z)I2b(w)I1c(w)8091(zw)2I2a(z)I2b(w)I1c(w)+16091zwI3a(z)I1b(w)I1c(w))8091zwI2a(z)I3a(w)𝗄(w)8031zwI4a(z)I1a(w)𝗄(w)+320271(zw)3I2a(z)I3a(w)+16091zwI4a(z)I1a′′(w)+32091(zw)3I4a(z)I1a(w)+160271(zw)2I2a(z)I3a(w)16091zwI3a(z)I2a′′(w)16091(zw)2I3a(z)I2a(w)16091(zw)2I4a(z)I1a(w)]|0.\begin{split}\mathsf{A}_{1,3}(z,w)&=\Big{[}\frac{8}{3}\frac{1}{z-w}I^{a}_{-2}(z)I^{(a}_{-1}(w)I^{b}_{-1}(w)I^{b)}_{-1}(w)\\ &+f^{abc}\Big{(}\frac{80}{9}\frac{1}{z-w}I^{a}_{-2}(z)I^{b}_{-2}(w)I^{c\prime}_{-1}(w)-\frac{80}{9}\frac{1}{(z-w)^{2}}I^{a}_{-2}(z)I^{b}_{-2}(w)I^{c}_{-1}(w)\\ &\hskip 56.9055pt+\frac{160}{9}\frac{1}{z-w}I^{a}_{-3}(z)I^{b\prime}_{-1}(w)I^{c}_{-1}(w)\Big{)}\\ &-\frac{80}{9}\frac{1}{z-w}I^{a}_{-2}(z)I^{a}_{-3}(w)\mathsf{k}^{\prime}(w)-\frac{80}{3}\frac{1}{z-w}I^{a}_{-4}(z)I^{a}_{-1}(w)\mathsf{k}^{\prime}(w)\\ &+\frac{320}{27}\frac{1}{(z-w)^{3}}I^{a}_{-2}(z)I^{a}_{-3}(w)+\frac{160}{9}\frac{1}{z-w}I^{a}_{-4}(z)I^{a\prime\prime}_{-1}(w)\\ &+\frac{320}{9}\frac{1}{(z-w)^{3}}I^{a}_{-4}(z)I^{a}_{-1}(w)+\frac{160}{27}\frac{1}{(z-w)^{2}}I^{a}_{-2}(z)I^{a\prime}_{-3}(w)\\ &-\frac{160}{9}\frac{1}{z-w}I^{a}_{-3}(z)I^{a\prime\prime}_{-2}(w)-\frac{160}{9}\frac{1}{(z-w)^{2}}I^{a}_{-3}(z)I^{a\prime}_{-2}(w)\\ &-\frac{160}{9}\frac{1}{(z-w)^{2}}I^{a}_{-4}(z)I^{a\prime}_{-1}(w)\Big{]}\ket{0}.\end{split}
𝖡1,3(z,w)=[831(zw)2I1a(z)I1(a(w)I1b(w)I1b)(w)+fabc(16091(zw)3I1a(z)I2b(w)I1c(w)8091(zw)2I1a(z)I1b(w)I2c(w)16091(zw)2I2a(z)I1b(w)I1c(w))16091(zw)2I1a(z)I3a(w)𝗄(w)+1120271(zw)2I1a(z)I3a′′(w)+640271(zw)3I1a(z)I3a(w)32091(zw)2I2a(z)I2a(w)+32091(zw)4I1a(z)I3a(w)64091(zw)3I3a(z)I1a(w)+32031(zw)4I3a(z)I1a(w)]|0.\begin{split}\mathsf{B}_{1,3}(z,w)=&\Big{[}\frac{8}{3}\frac{1}{(z-w)^{2}}I^{a}_{-1}(z)I^{(a}_{-1}(w)I^{b}_{-1}(w)I^{b)}_{-1}(w)\\ &+f^{abc}\Big{(}-\frac{160}{9}\frac{1}{(z-w)^{3}}I^{a}_{-1}(z)I^{b}_{-2}(w)I^{c}_{-1}(w)-\frac{80}{9}\frac{1}{(z-w)^{2}}I^{a}_{-1}(z)I^{b\prime}_{-1}(w)I^{c}_{-2}(w)\\ &\hskip 56.9055pt-\frac{160}{9}\frac{1}{(z-w)^{2}}I^{a}_{-2}(z)I^{b\prime}_{-1}(w)I^{c}_{-1}(w)\Big{)}\\ &-\frac{160}{9}\frac{1}{(z-w)^{2}}I^{a}_{-1}(z)I^{a}_{-3}(w)\mathsf{k}^{\prime}(w)+\frac{1120}{27}\frac{1}{(z-w)^{2}}I^{a}_{-1}(z)I^{a\prime\prime}_{-3}(w)\\ &+\frac{640}{27}\frac{1}{(z-w)^{3}}I^{a}_{-1}(z)I^{a\prime}_{-3}(w)-\frac{320}{9}\frac{1}{(z-w)^{2}}I^{a}_{-2}(z)I^{a\prime}_{-2}(w)\\ &+\frac{320}{9}\frac{1}{(z-w)^{4}}I^{a}_{-1}(z)I^{a}_{-3}(w)-\frac{640}{9}\frac{1}{(z-w)^{3}}I^{a}_{-3}(z)I^{a\prime}_{-1}(w)\\ &+\frac{320}{3}\frac{1}{(z-w)^{4}}I^{a}_{-3}(z)I^{a}_{-1}(w)\Big{]}\ket{0}.\end{split}

References

  • [BLZ97] V.. Bazhanov, S.. Lukyanov and Al.. Zamolodchikov “Integrable structure of conformal field theory. 2. Q operator and DDV equation” In Commun. Math. Phys. 190, 1997, pp. 247–278 DOI: 10.1007/s002200050240
  • [DS85] V.. Drinfel’d and V.. Sokolov “Lie algebras and equations of Korteweg-de Vries type” In Journal of Soviet Mathematics 30.2, 1985, pp. 1975–2036 DOI: 10.1007/BF02105860
  • [Eva+99] J.M. Evans, M. Hassan, N.J. MacKay and A.J. Mountain “Local conserved charges in principal chiral models” In Nuclear Physics B 561.3 Elsevier BV, 1999, pp. 385–412 DOI: 10.1016/s0550-3213(99)00489-7
  • [Eva01] J.. Evans “Integrable sigma-models and Drinfel’d–Sokolov hierarchies” In Nuclear Physics B 608.3 Elsevier BV, 2001, pp. 591–609 DOI: 10.1016/s0550-3213(01)00157-2
  • [FB01] E. Frenkel and D. Ben-Zvi “Vertex Algebras and Algebraic Curves”, Mathematical surveys and monographs American Mathematical Society, 2001 URL: https://books.google.co.uk/books?id=mRgqy1f1QkIC
  • [FF07] B. Feigin and E. Frenkel “Quantization of soliton systems and Langlands duality” arXiv, 2007 DOI: 10.48550/ARXIV.0705.2486
  • [FFR94] B. Feigin, E. Frenkel and N. Reshetikhin “Gaudin model, Bethe Ansatz and critical level” In Communications in Mathematical Physics 166.1, 1994, pp. 27–62 DOI: 10.1007/BF02099300
  • [Fre04] E. Frenkel “Gaudin model and opers” In Workshop on Infinite Dimensional Algebras and Quantum Integrable Systems, 2004 arXiv:math/0407524
  • [Gau14] M. Gaudin “The Bethe Wavefunction” Cambridge University Press, 2014 DOI: 10.1017/CBO9781107053885
  • [Gau76] M. Gaudin “Diagonalisation d’une classe d’hamiltoniens de spin” In Journal De Physique 37, 1976, pp. 1087–1098
  • [Itō93] Kiyosi Itō “Encyclopedic dictionary of mathematics” Cambridge, Mass: MIT Press, 1993
  • [Kac90] V.. Kac “Infinite-Dimensional Lie Algebras” Cambridge University Press, 1990 DOI: 10.1017/CBO9780511626234
  • [KL21] G.. Kotousov and S.. Lukyanov “ODE/IQFT correspondence for the generalized affine 𝔰𝔩\mathfrak{sl}(2) Gaudin model” In JHEP 09, 2021, pp. 201 DOI: 10.1007/JHEP09(2021)201
  • [KLT22] G.. Kotousov, S. Lacroix and J. Teschner “Integrable sigma models at RG fixed points: quantisation as affine Gaudin models” arXiv, 2022 DOI: 10.48550/arxiv.2204.06554
  • [Kui+13] J. Kuipers, T. Ueda, J… Vermaseren and J. Vollinga “FORM version 4.0” In Comput. Phys. Commun. 184.5, 2013, pp. 1453–1467 DOI: 10.1016/j.cpc.2012.12.028
  • [Lac18] S. Lacroix “Integrable models with twist function and affine Gaudin models”, 2018 arXiv:1809.06811 [hep-th]
  • [LMV17] S. Lacroix, M. Magro and B. Vicedo “Local charges in involution and hierarchies in integrable sigma-models” In Journal of High Energy Physics 2017.9 Springer ScienceBusiness Media LLC, 2017 DOI: 10.1007/jhep09(2017)117
  • [LV21] S. Lacroix and B. Vicedo “Integrable E-Models, 4d Chern-Simons Theory and Affine Gaudin Models. I. Lagrangian Aspects” In Symmetry, Integrability and Geometry: Methods and Applications SIGMA (Symmetry, IntegrabilityGeometry: MethodsApplication), 2021 DOI: 10.3842/sigma.2021.058
  • [LVY18] S. Lacroix, B. Vicedo and C.A.S. Young “Affine Gaudin models and hypergeometric functions on affine opers” In Advances in Mathematics 350, 2018 DOI: 10.1016/j.aim.2019.04.032
  • [LVY20] S. Lacroix, B. Vicedo and C.A.S. Young “Cubic hypergeometric integrals of motion in affine Gaudin models” In Adv. Theor. Math. Phys. 24.1, 2020, pp. 155–187 DOI: 10.4310/ATMP.2020.v24.n1.a5
  • [MTV06] E. Mukhin, V. Tarasov and A. Varchenko “Bethe eigenvectors of higher transfer matrices” In Journal of Statistical Mechanics: Theory and Experiment 2006.8, 2006, pp. 08002 DOI: 10.1088/1742-5468/2006/08/P08002
  • [MTV07] E. Mukhin, V. Tarasov and A. Varchenko “Schubert calculus and representations of general linear group” arXiv, 2007 DOI: 10.48550/ARXIV.0711.4079
  • [MV02] E. Mukhin and A. Varchenko “Critical points of master functions and flag varieties” arXiv, 2002 DOI: 10.48550/ARXIV.MATH/0209017
  • [MV04] E. Mukhin and A. Varchenko “Norm of a Bethe Vector and the Hessian of the Master Function” arXiv, 2004 DOI: 10.48550/ARXIV.MATH/0402349
  • [Ryb06] L. Rybnikov “Uniqueness of higher Gaudin Hamiltonians” In Reports on Mathematical Physics 61, 2006 DOI: 10.1016/S0034-4877(08)80013-4
  • [Tal04] D. Talalaev “Quantization of the Gaudin system”, 2004 arXiv:hep-th/0404153
  • [Ver13] J. Vermaseren “FORM”, 2013 URL: https://www.nikhef.nl/~form/maindir/documentation/documentation.html
  • [Vic18] B. Vicedo “On Integrable Field Theories as Dihedral Affine Gaudin Models” In International Mathematics Research Notices 2020.15, 2018, pp. 4513–4601 DOI: 10.1093/imrn/rny128
  • [Vic21] B. Vicedo “4D Chern–Simons theory and affine Gaudin models” In Letters in Mathematical Physics 111, 2021 DOI: 10.1007/s11005-021-01354-9
  • [VY17] B. Vicedo and C.A.S. Young “Vertex Lie algebras and cyclotomic coinvariants” In Communications in Contemporary Mathematics 19.02, 2017, pp. 1650015 DOI: 10.1142/S0219199716500152