1 Introduction

Intricate connection between supersymmetric gauge theories and integrable systems plays an important role in modern theoretical physics and mathematics. From physics point of view whenever some sector of the gauge theory is related to an integrable system, observables in this sector can be computed using rich variety of integrability techniques. Canonical example of this situation is the integrability of \({\mathcal {N}}=4\) super Yang–Mills (SYM) theory [1, 2]. In particular, it connects calculation of the planar scaling dimensions of \({\mathcal {N}}=4\) SYM to the integrable system of spin chains and allows to compute these scaling dimensions at arbitrary coupling. On the other hand, exploration of such connections can also shed light on some questions about integrable systems and even lead to construction of new classes of such systems making this kind of studies interesting from mathematics point of view.

In our work, we are exploring a particular class of connections between six-dimensional (1, 0) superconformal field theories and elliptic quantum mechanics Hamiltonians in the spirit of Bethe/gauge correspondence of Nekrasov and Shatashvili [3,4,5,6]. In particular, we consider four-dimensional theories with four supercharges obtained by compactification of a 6d SCFT on a punctured Riemann surface. In order for an integrable model to emerge in this setting, we have to introduce surface defects into our 4d theory and study its superconformal index.

This construction was first established in the context of compactifications of 6d (2, 0) of ADE type [7, 8]. In particular, the corresponding 4d superconformal indices with the defect were found to be closely related to the Ruijsenaars–Schneider (RS) elliptic analytic finite difference operators (A\(\Delta \)Os). Later these results were extended to many other cases: class \(S_k\) 4d theories [9], compactifications of \(A_2\) and \(D_4\) minimal 6d SCFTs [10] and compactifications of rank one E-string theories [11]. In some of these cases, the obtained A\(\Delta \)Os were already known in the literature. For example in E-string compactifications, van Diejen (vD) model [12, 13] was observed. But in some cases operators were previously unknown as in the case of the minimal 6d SCFTs compactifications. Study of such novel operators constitutes an interesting field of research with some initial steps already taken in this direction [14, 15].

From the point of view of physics, the connection of superconformal indices in the presence of defects with the integrable quantum mechanics Hamiltonians allows one to bootstrap index of an arbitrary theory obtained by compactifying corresponding 6d theory on a punctured Riemann surface. In particular, if we know the eigenfunctions of the corresponding A\(\Delta \)O, we can compute the index of any theory obtained in such compactifications including non-Lagrangian theories for which there are no other methods of computing the index. As mentioned previously, the first setting where these ideas were tested is 4d class-S theories obtained in the compactifications of 6d (2, 0) theory [7]. In this case, the corresponding integrable system was given by elliptic RS model. Eigenfunctions of the elliptic RS Hamiltonians are not known in general, but some of their limits are well studied in the mathematical literature. These limits were used in order to prove previously established relations [16,17,18] of superconformal indices with the Macdonald and Schur polynomials allowing one to compute indices of class S non-Lagrangian theories.

The construction outlined above relies on the intermediate 5d layer. For things to work there should exist an effective 5d gauge theory obtained by compactifying original 6d SCFT on a circle with a choice of holonomies for its global symmetries. In particular, different 5d compactifications lead to different types of punctures on the Riemann surfaces used to obtain 4d theories. One of the interesting problems related to this fact is the compilation of the dictionary between known compactifications of various 6d SCFTs and elliptic integrable systems. In our previous paper [19], we have considered compactifications of the 6d minimal \((D_{N+3},D_{N+3})\) conformal matter theories [20, 21]. In particular, we have derived A\(\Delta \)Os corresponding to the intermediate 5d \(\textrm{SU}(N+1)\) gauge theory or equivalently the \(A_N\)-type puncture on the compactification surface. These elliptic A\(\Delta \)Os appeared to be previously unknown \(A_N\) generalizations of vD model. On the other hand, there are at least two more 5d effective descriptions of 6d SCFTs corresponding to \(\textrm{USp}(2N)\) and \(\textrm{SU}(2)^N\) gauge theories giving rise to the punctures with the same \(C_N\) and \((A_1)^{\otimes N }\) global symmetries. These two descriptions should lead to additional higher-rank generalizations of the vD model.

In our present paper, we follow this line of research and closely study compactifications of the minimal \((D_{N+3},D_{N+3})\) conformal matter theory on a Riemann surface with the \(C_N\)-type punctures. In particular, we concentrate on the next to simplest case of \(N=2\)Footnote 1 and derive corresponding infinite tower of A\(\Delta \)Os. We also devote part of the paper to the study and proof of some remarkable properties of these operators that follow from the geometry of corresponding compactifications. In particular, we prove the novel kernel property for two operators of different type but same rank. We also discuss commutation property of obtained operators.

The paper is organized as follows. In Sect. 2, we review our previous results for \(A_N\)-type operators. In addition to the previous results, we also derive full tower of such A\(\Delta \)Os which was not obtained previously. In Sect. 3, we derive in details novel generalization of vD operators corresponding to \(C_2\) root system.Footnote 2 In Sect. 4, we discuss properties of derived \(C_2\) A\(\Delta \)Os. In particular we pay special attention to a new kernel function for simultaneously \(A_2\) and \(C_2\) operators and give fully analytic proof of the corresponding kernel equation. We also briefly discuss commutation relations and check them perturbatively in expansion. In Sect. 5, we briefly summarize our results and discuss plans for the future research further developing these results. Finally, the paper has a number of appendices collecting useful formulas as well as technical details of various calculations in our paper.

2 \(A_N\) operators

In this section, we will briefly review derivation of the \(A_N\) generalization of the van Diejen operator. So-called basic versionFootnote 3 of this operator has been derived in our previous paper [19]. Here, we will extend this result to the full tower of operators.

Just as in [19] in order to derive A\(\Delta \)O, we start with the 4d three-punctured sphere theory which was first introduced in [21]. This theory is obtained in the compactification of the 6d minimal \((D_{N+3},D_{N+3})\) conformal matter on a sphere with two maximal \(\textrm{SU}(N+1)\) punctures and one minimal \(\textrm{SU}(2)\) puncture. The quiver of this theory is shown in Fig. 1. In addition to the global symmetry, the punctures are characterized by the moment map operators. In particular, both \(\textrm{SU}(N+1)\) maximal and \(\textrm{SU}(2)\) minimal punctures are characterized by \((2N+4)\) mesonic and 2 baryonic moment maps:

$$\begin{aligned} M_u= & {} \mathbf{N+1}^x\otimes \left( \mathbf{2N+4}_{u^{N+3}v^{-N-1}w^{-2}}\oplus \textbf{1}_{\left( uv^{N+1} \right) ^{2N+4}}\right) \oplus \overline{ \mathbf{N+1} }^x\otimes \textbf{1}_{\left( u^N w^2\right) ^{2N+4}}, \nonumber \\ M_v= & {} \mathbf{N+1}^y\otimes \left( \mathbf{2N+4}_{v^{N+3}u^{-N-1}w^{-2}}\oplus \textbf{1}_{\left( vu^{N+1} \right) ^{2N+4}}\right) \oplus \overline{ \mathbf{N+1} }^y\otimes \textbf{1}_{\left( v^N w^2\right) ^{2N+4}}, \nonumber \\ M_w= & {} \textbf{2}^z\otimes \left( \mathbf{2N+4}_{\left( uvw^{-2} \right) ^{-N-1}}\oplus \textbf{1}_{\left( wv^{N+1}\right) ^{2N+4}} \oplus \textbf{1}_{\left( wu^{N+1}\right) ^{2N+4}}\right) , \end{aligned}$$
(2.1)

where \(a_i,\, u,\, v,\, w\) are fugacities of the Cartans of the 6d global \(\textrm{SO}(4N+8)\) symmetry. Subscripts of the moment maps written above denote their charges w.r.t. to these symmetries.

Further S-gluing two such trinions along the maximal punctures, we can obtain four-punctured sphere with zero flux two maximal and two minimal punctures. To obtain the corresponding 4d gauge theory, we should just take two copies of trinion theories shown in Fig. 1 and then identify and gauge corresponding global symmetries of the maximal punctures. Here and everywhere else in the paper, all operations with gauge theories are expressed in terms of the superconformal indices. In this language, the gluing procedure takes the following form:

$$\begin{aligned} K_4^A(x,{\tilde{x}},z,{\tilde{z}})=\kappa _N\oint \prod \limits _{i=1}^N\frac{dy_i}{2\pi i y_i} \prod \limits _{i\ne j}^{N+1}\frac{1}{\Gamma _e\left( \frac{y_i}{y_j}\right) }{\bar{K}}_3^{A}({\tilde{x}},y,{\tilde{z}})K_3^{A}(x,y,z).\qquad \end{aligned}$$
(2.2)

where \(K^{A}_3(x,y,z)\) is the index of the trinion with y being fugacity of the global \(\textrm{SU}(N+1)\) symmetry of the puncture we glue along and \({\bar{K}}_3^{A}\) is the index of the conjugated trinion. Finally, \(\kappa _N\) is the usual constant given by:

$$\begin{aligned} \kappa _N\equiv \frac{\left( q;q\right) _\infty ^{N}\left( p;p\right) _\infty ^{N}}{\left( N+1\right) !}. \end{aligned}$$
(2.3)
Fig. 1
figure 1

a \(A_N\) three-punctured sphere with two maximal and one minimal puncture. b \(A_N\) four-punctured sphere obtained by S-gluing two three-punctured spheres

Performing this S-gluing operation, we obtain the four-punctured sphere theory shown in Fig. 1 with the corresponding superconformal index specified in (B.1). This theory was previously obtained by us in [19] to derive basic \(A_N\) A\(\Delta \)Os. Now in order to derive the operator, we should close two minimal punctures of this four-punctured sphere. To do it, we should break the global symmetry of the puncture. This can be achieved by giving a non-trivial vev \(\langle \partial _{12}^L \partial _{34}^K M\rangle \ne 0\) to the derivatives of one of the moment map operators. When we close punctures with at least one of K or L not equal to zero, i.e., vev is space-time dependent, we effectively insert defect into the theory [7, 22]. At the level of the superconformal index, closing the puncture amounts to giving a corresponding weight to the fugacity of the puncture’s global symmetry. Once we do it, we hit a pole of the index. Then, computation of the residue of this pole results in the superconformal index of the IR theory that the UV theory flowed to due to the introduction of the vev. In our case, we choose to close \(\textrm{SU}(2)_z\) minimal puncture with the defect and \(\textrm{SU}(2)_{{\tilde{z}}}\) puncture without, i.e., choosing \(L=K=0\) in the vev. In particular, let’s say that we are going to compute A\(\Delta \)O acting on the puncture with the moment maps of charges \({\tilde{h}}_i\) and an overall \(\textrm{U}(1)\) charge

$$\begin{aligned} {\tilde{h}}\equiv \prod \limits _{i=1}^{2N+6}{\tilde{h}}_i. \end{aligned}$$
(2.4)

For example, the moment maps of the \(\textrm{SU}(N+1)_x\) puncture of the trinion shown in Fig. 1 correspond to \(M_u\) operators specified in (2.1) and have the following charges:

$$\begin{aligned}{} & {} {\tilde{h}}_i = u^{N+3}v^{-N-1}w^{-2}a_i,\, \quad i=1,\dots ,2N+4, \quad \prod _{i=1}^{2N+4}a_i=1, \nonumber \\{} & {} {\tilde{h}}_{2N+5} = \left( uv^{N+1}\right) ^{2N+4},\,\quad {\tilde{h}}_{2N+6}= \left( u^{N} w^{2}\right) ^{-2N-4},\quad {\tilde{h}}=\left( uw^{-1}\right) ^{8(N+2)}, \nonumber \\ \end{aligned}$$
(2.5)

where we have flipped the charge \({\tilde{h}}_{2N+6}\) of the last moment map since this parametrization will be more natural for our operators and in this case all of the moment maps of the maximal puncture transform in the fundamental representation of \(\textrm{SU}(N+1)_x\). We can now notice from (2.1) that charges of the minimal puncture moment maps are related to the charges of the moment maps of the maximal punctures by simple relation

$$\begin{aligned} {\tilde{h}}^{\textrm{SU}(2)}_i=\left( {\tilde{h}}^{\textrm{SU}(N+1)}_i\right) ^{-1}\left( {\tilde{h}}^{\textrm{SU}(N+1)}\right) ^{\frac{1}{4}}, \end{aligned}$$
(2.6)

So we can express everything in terms of only the moment maps of the maximal puncture we act on. Now, assume we give vevs to the moment maps (mesonic or baryonic) with charges \({\tilde{h}}_i^{-1}{\tilde{h}}^{1/4}\) of both \(\textrm{SU}(2)_z\) and \(\textrm{SU}(2)_{{\tilde{z}}}\). Moment maps we use should be the same in order to keep total flux of the 6d global symmetries zero. At the level of the index calculations, it corresponds to computing the residue of the index of the four-punctured sphere theory at the pole

$$\begin{aligned} z= & {} Z_{i;L,M}^*=(pq)^{-\frac{1}{2}} {\tilde{h}}_i{\tilde{h}}^{-\frac{1}{4}} q^{-M} p^{-L}, \;\;\;\;\;\;\;\nonumber \\ {\tilde{z}}= & {} {\tilde{Z}}^*_{i;0,0}=(pq)^{-\frac{1}{2}} {\tilde{h}}_i^{-1}{\tilde{h}}^{\frac{1}{4}} q^{-{\tilde{M}}} p^{-{\tilde{L}}}, \end{aligned}$$
(2.7)

where \(L,M,{\tilde{L}},{\tilde{M}}\) are positive integers corresponding to the powers of derivatives inside the vev. As we mentioned previously, it is enough to introduce defect only for one of the two punctures. Hence, we choose \({\tilde{M}}={\tilde{L}}=0\) and keep LM general. Then, we compute corresponding residues of the index of the four-punctured sphere theory and obtain theory for the tube with two maximal \(\textrm{SU}(N+1)\) punctures and a codimension-two defect. Its superconformal index is given byFootnote 4:

$$\begin{aligned} K_{(2;i;L,M)}^A(x,{\tilde{x}})\sim \textrm{Res}_{z\rightarrow Z^*_{i;L,M},{\tilde{z}}\rightarrow {\tilde{Z}}^*_{i;0,0}}K_4^A(x,{\tilde{x}},z,{\tilde{z}}) \end{aligned}$$
(2.8)

Finally in order to obtain desired A\(\Delta \)O, we glue our tube with the defect to an arbitrary Riemann surface with maximal \(\textrm{SU}(N+1)_{{\tilde{x}}}\) puncture. As the result of this gluing, we expect to obtain action of a finite difference operator on the index \( {{\mathcal {I}}}({\tilde{x}})\) of this 4d \({\mathcal {N}}=1\) theory:

$$\begin{aligned} {{\mathcal {O}}}^{(A_N;h_{k};L,M)}_x\cdot {{\mathcal {I}}}(x)=\kappa _N\oint \prod \limits _{j=1}^N\frac{d{\tilde{x}}_j}{2\pi i{\tilde{x}}_{j}}\prod \limits _{i\ne j}^{N+1}\frac{1}{\Gamma _e\left( \frac{{\tilde{x}}_i}{{\tilde{x}}_j}\right) }K_{(2;i;L,M)}^A(x,\,{\tilde{x}}) {{\mathcal {I}}}({\tilde{x}}) \nonumber \\ \end{aligned}$$
(2.9)

Details of all the calculations summarized above are given in Appendix B. They result in the following operator:

figure a

where \(C_{L,M}^{\vec {l},\vec {m},\vec {r},\vec {s}}\) are x-independent constant factors given by,

figure b

Also \(\Delta \)s are shift operators defined as follows:

$$\begin{aligned} \Delta ^m_a(x_i)f(x)\equiv f\left( x_i\rightarrow a^mx_i\right) , \quad a=q,p. \end{aligned}$$
(2.12)

The operator contains all shifts of the form \(q^{m_i}p^{l_i}x_i\) where \(\vec {m}\) and \(\vec {l}\) are all possible partitions of length \(N+1\) of M and L correspondingly. At each level, i.e., fixed M and L, there are \(2N+6\) operators due to \(2N+6\) moment maps with the charges \({\tilde{h}}_k{\tilde{h}}^{-\frac{1}{4}}\). There are also \(2N+6\) other operators obtained by giving vevs to the flipped moment maps of the charge \({\tilde{h}}_k^{-1}{\tilde{h}}^{\frac{1}{4}}\). They have similar form and properties so we do not present them here. All the operators should commute with each other, and we checked this in expansion in pq for a few of the simplest cases. Now, we will refer to the case \(M=1\) and \(L=0\), or vice versa, as basic operators. These basic operators for \(A_N\) generalizations of vD model were derived by us previously in [19]. The operators above also reproduce our previous results when we fix \(M=1,\, L=0\) or \(M=0,\, L=1\). Further in our paper, we will also need the basic operator obtained by closing flipped moment maps. This operator can be found in [19] and has the following form:

figure c

where the shift part of this operator is given by

$$\begin{aligned} A^{(A_N;{\tilde{h}}_i^{-1};1,0)}_{lm}(x)= & {} \frac{\prod \limits _{j=1}^{2N+6}\theta _p\left( (pq)^{\frac{1}{2}} {\tilde{h}}_j^{-1}x_l^{-1}\right) }{\theta _p\left( \frac{x_m}{x_l}\right) \theta _p\left( q\frac{x_m}{x_l}\right) }\nonumber \\{} & {} \prod \limits _{k\ne m\ne l}^{N+1}\frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2} x_k\right) }{\theta _p\left( \frac{x_k}{x_l}\right) \theta _p\left( \frac{x_m}{x_k}\right) },\nonumber \\ \end{aligned}$$
(2.14)

and the constant part is given by:

$$\begin{aligned} W^{(A_N;{\tilde{h}}_i^{-1};1,0)}(x,{\tilde{h}})= & {} \frac{\prod \limits _{j\ne i}^{2N+6}\theta _p\left( q^{-1}{\tilde{h}}_i{\tilde{h}}_j{\tilde{h}}^{-1/2}\right) }{\theta _p\left( q^{-2}{\tilde{h}}_i^{2}{\tilde{h}}^{-1/2}\right) } \prod \limits _{k=1}^{N+1}\frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1} x_k^{-1}\right) }{\theta _p\left( (pq)^{-\frac{1}{2}}{\tilde{h}}_i{\tilde{h}}^{-1/2}q^{-1}x_k^{-1}\right) } \nonumber \\{} & {} +\sum \limits _{m=1}^{N+1}\frac{\prod \limits _{j\ne i}^{2N+6}\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_jx_m\right) }{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2}q x_m\right) } \nonumber \\{} & {} \prod \limits _{k\ne m}^{N+1}\frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_{i}^{-1}{\tilde{h}}^{1/2}x_k\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) }{\theta _p\left( q^{-1}\frac{x_k}{x_m}\right) \theta _p\left( \frac{x_m}{x_k}\right) }. \end{aligned}$$
(2.15)

This constant part is elliptic function in each \(x_i\) variable with periods 1 and p. It has poles in the fundamental domain at the following positions:

$$\begin{aligned} x_i=q^{\pm 1}x_r,\quad x_i=sq^{\pm \frac{1}{2}}P_i^{-\frac{1}{2}},\quad x_i=sq^{\pm \frac{1}{2}}p^{\frac{1}{2}}P_i^{-\frac{1}{2}},\quad s=\pm 1, \end{aligned}$$
(2.16)

where we defined

$$\begin{aligned} P_i\equiv \prod \limits _{j\ne i}^N x_j. \end{aligned}$$
(2.17)

From expression (2.15), it looks like there are extra poles in the constant part, but careful examination shows that residues at these values of x’s are zero so there are no real poles there. At the poles (2.16), we have the following residues:

$$\begin{aligned}&\textrm{Res}_{x_l=qx_r}W^{(A_N;{\tilde{h}}_i^{-1};1,0)}(x,{\tilde{h}})\nonumber \\&\quad =-\frac{qx_r}{\left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) }\prod \limits _{j=1}^{2N+6}\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_jx_r\right) \nonumber \\&\qquad \times \prod \limits _{k\ne l\ne r}^{N+1}\frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2}x_k\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) }{\theta _p\left( q^{-1}\frac{x_k}{x_r}\right) \theta _p\left( \frac{x_r}{x_k}\right) } \nonumber \\&\textrm{Res}_{x_l=q^{-1}x_r}W^{(A_N;{\tilde{h}}_i^{-1};1,0)}(x,h)\nonumber \\&\quad =\frac{q^{-1}x_r}{\left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) }\prod \limits _{j=1}^{2N+6}\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_j^{-1}x_r^{-1}\right) \nonumber \\&\qquad \times \prod \limits _{k\ne l\ne r}^{N+1}\frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2}x_k\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) }{\theta _p\left( q^{-1}\frac{x_r}{x_k}\right) \theta _p\left( \frac{x_k}{x_r}\right) } \nonumber \\&\textrm{Res}_{x_l=sq^{-\frac{1}{2}}P_l^{-\frac{1}{2}}}W^{(A_N;{\tilde{h}}_i^{-1};1,0)}(x,h)\nonumber \\&\quad =s\frac{q^{-\frac{1}{2}}P_l^{-\frac{1}{2} }}{2 \left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) } \prod \limits _{j=1}^{2N+6}\theta _p\left( sp^{1/2}{\tilde{h}}_jP_l^{-1/2}\right) \nonumber \\&\qquad \times \prod \limits _{k\ne l}^N \frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2}x_k\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) }{\theta _p\left( sq^{-1/2}P_l^{-1/2}x_k^{-1}\right) \theta _p\left( sq^{-1/2}P_l^{1/2}x_k\right) }\,, \nonumber \\&\textrm{Res}_{x_l=sq^{\frac{1}{2}}P_l^{-\frac{1}{2}}}W^{(A_N;{\tilde{h}}_i^{-1};1,0)}(x,{\tilde{h}})\nonumber \\&\quad =-s\frac{q^{\frac{1}{2}}P_l^{-\frac{1}{2} }}{2 \left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) } \prod \limits _{j=1}^{2N+6}\theta _p\left( sp^{1/2}{\tilde{h}}_jP_l^{-1/2}\right) \nonumber \\&\qquad \times \prod \limits _{k\ne l}^N \frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2}x_k\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) }{\theta _p\left( sq^{-1/2}P_l^{-1/2}x_k^{-1}\right) \theta _p\left( sq^{-1/2}P_l^{1/2}x_k\right) }\,, \nonumber \\&\textrm{Res}_{x_l=sp^{\frac{1}{2}}q^{-\frac{1}{2}}P_l^{-\frac{1}{2}}}W^{(A_N;{\tilde{h}}_i^{-1};1,0)}(x,{\tilde{h}})\nonumber \\&\quad =s\frac{q^{-\frac{1}{2}}p^{\frac{3}{2}}{\tilde{h}}^{-\frac{1}{2}} P_l^{\frac{1}{2} }}{2 \left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) } \prod \limits _{j=1}^{2N+6}\theta _p\left( s{\tilde{h}}_jP_l^{-1/2}\right) \nonumber \\&\qquad \times \prod \limits _{k\ne l}^N \frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2}x_k\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) }{\theta _p\left( sp^{-1/2}q^{-1/2}P_l^{-1/2}x_k^{-1}\right) \theta _p\left( sp^{1/2}q^{-1/2}P_l^{1/2}x_k\right) }\,, \nonumber \\&\textrm{Res}_{x_l=sp^{\frac{1}{2}}q^{\frac{1}{2}}P_l^{-\frac{1}{2}}}W^{(A_N;{\tilde{h}}_i^{-1};1,0)}(x,{\tilde{h}})\nonumber \\&\quad =-s\frac{q^{\frac{1}{2}}p^{\frac{3}{2}}{\tilde{h}}^{-\frac{1}{2}} P_l^{\frac{1}{2} }}{2 \left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) }\prod \limits _{j=1}^{2N+6}\theta _p\left( s{\tilde{h}}_jP_l^{-1/2}\right) \nonumber \\&\qquad \times \prod \limits _{k\ne l}^N \frac{\theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}{\tilde{h}}^{1/2}x_k\right) \theta _p\left( (pq)^{\frac{1}{2}}{\tilde{h}}_i^{-1}x_k^{-1}\right) }{\theta _p\left( sp^{-1/2}q^{-1/2}P_l^{-1/2}x_k^{-1}\right) \theta _p\left( sp^{1/2}q^{-1/2}P_l^{1/2}x_k\right) }\,, \end{aligned}$$
(2.18)

These are completely general expressions for the whole family of \(A_N\) operators.

3 Derivation of \(C_2\) operators

In this section, we will discuss derivation of the C-type rank-2 analytic finite difference operators A\(\Delta \)O. In our previous paper [19], we have already derived the basic operator for \(A_N\) generalization of van Diejen model and the calculation of the full tower of these operators is summarized in Sect. 2. Unfortunately full \(C_N\) generalization is still out of reach for us due to technical complications but we can concentrate on the study of rank-2 generalizations with \(C_2\) root system.

We will derive \(C_2\) operator corresponding to the insertion of the codimension-two defect due to the non-trivial vev of the holomorphic derivatives of the moment maps \(\langle \partial _{\pm }^K M \rangle \ne 0\). To simplify our calculations of A\(\Delta \)O, we will start with the derivation of four- and three-punctured spheres with \(C_2\)-type punctures.

The basic \(C_N\) trinion was already derived by us in Appendix B.5 of [19] and is shown in Fig. 2b. Generalizations of this trinion to the cases of three maximal punctures and higher numbers of minimal punctures can also be found in [23]. In order to derive this trinion, we used a tube theory with one \(\textrm{USp}(2N)\) and one \(\textrm{SU}(N+1)\) maximal punctures shown in Fig. 2a. This theory was first discussed in [24]. In order to obtain trinion theory shown in Fig. 2b, we start with the three-punctured sphere with two maximal \(\textrm{SU}(N+1)\) and one minimal \(\textrm{SU}(2)\) punctures and glue two \(A_NC_N\) tubes to the maximal punctures. The moment maps of the punctures of the resulting trinion are given by \((2N+6)\) mesons both for maximal and minimal punctures:

$$\begin{aligned} M_x= & {} \textbf{2N}_x\otimes \left( \mathbf{2N+6}\right) _{w^{2\frac{(N+2)^2}{N+3}}}, \qquad M_y=\textbf{2N}_y\otimes \left( \mathbf{2N+6}\right) _{w^{2\frac{(N+2)^2}{N+3}}}, \nonumber \\ M_z= & {} \textbf{2}_z\otimes \left( \mathbf{2N+6}\right) _{w^{2\frac{(N+2)^2}{N+3}}}, \end{aligned}$$
(3.1)

where w is one of the parameters of Cartans of the global \(\textrm{SO}(4N+8)\) symmetry of the 6d minimal conformal matter theory. In general, we will use two types of parametrization of 6d global symmetry in this paper. First one is natural to use in compactifications with \(C_N\)-type punctures, and it is given by

$$\begin{aligned} {\tilde{a}}_i,\quad i=1,\dots ,2N+6,\quad \prod \limits _{i=1}^{2N+6}{\tilde{a}}_i=1, \quad w. \end{aligned}$$
(3.2)

Another parametrization is useful when we work with \(A_N\) expressions and is given by:

$$\begin{aligned} a_i,\quad i=1,\dots ,2N+4,\quad \prod \limits _{i=1}^{2N+4}a_i=1, \quad u,v,w. \end{aligned}$$
(3.3)

These two parametrizations are related by the following map that we will often need in our calculations:

$$\begin{aligned} {\tilde{a}}_l= & {} \left( uv \right) ^{-N-1}w^{-\frac{2}{N+3}}a_l,~l=1,\dots ,2N+4, \nonumber \\ {\tilde{a}}_{2N+6}= & {} u^{2(N+1)(N+2)}w^{2\frac{N+2}{N+3}},\quad {\tilde{a}}_{2N+5}=v^{2(N+1)(N+2)}w^{2\frac{N+2}{N+3}}. \end{aligned}$$
(3.4)

Now in order to derive finite difference operators, we should proceed in the same way as in the case of \(A_N\) operators summarized in Sect. 2. We start by taking two \(C_N\) trinion theories \( {{\mathcal {T}}}^{{\mathcal {A}}}_{x,y,z}\) where \({\mathcal {A}}\) is the nonzero flux of the trinion and xyz in the subscript are fugacities of the global symmetries of two maximal and one minimal punctures correspondingly. Then, we take an arbitrary \( {{\mathcal {N}}}=1\) theory obtained in the compactification of the minimal \((D_{N+3},D_{N+3})\) conformal matter theory on the Riemann surface of genus g with s punctures denoted as \({{\mathcal {C}}}_{g,s}\) with at least one \(\textrm{USp}(2N)_{{\tilde{x}}}\) maximal puncture. Gluing all three surfaces together along the maximal punctures results in a Riemann surface of the same genus and global symmetry fluxes but two extra minimal punctures. We will be performing all these operations at the level of superconformal indices. There gluing amounts to identifying global symmetries of the punctures we glue and gauging it. So for the indices we can write down the following identity after gluing:

$$\begin{aligned}{} & {} {{\mathcal {I}}}\left[ {{\mathcal {C}}}_{g,s}({\tilde{x}})\oplus {\bar{ {{\mathcal {T}}}}}^{{\mathcal {A}}}_{{\tilde{x}},y,z_1}\oplus {{\mathcal {T}}}^{{\mathcal {A}}}_{y,x,z_2} \right] \nonumber \\{} & {} \quad =\frac{\left( q;q\right) _\infty ^4\left( p;p\right) _\infty ^4}{2^4\cdot 2!\cdot 2!}\oint \prod \limits _{i=1}^2\frac{d{\tilde{x}}_i}{2\pi i {\tilde{x}}_i}\frac{dy_i}{2\pi i y_i} \prod \limits _{i=1}^N\frac{1}{\Gamma _e\left( y_i^{\pm 2}\right) }\nonumber \\{} & {} \qquad \times \prod \limits _{i<j}^N\frac{1}{\Gamma _e\left( y_i^{\pm 1}y_j^{\pm 1}\right) } \prod \limits _{i=1}^N\frac{1}{\Gamma _e\left( {\tilde{x}}_i^{\pm 2}\right) }\nonumber \\{} & {} \quad \qquad \prod \limits _{i<j}^N\frac{1}{\Gamma _e\left( {\tilde{x}}_i^{\pm 1}{\tilde{x}}_j^{\pm 1}\right) } K_3^{C}(x,y,z_2){\bar{K}}_3^{C}({\tilde{x}},y,z_1) {{\mathcal {I}}}\left[ {{\mathcal {C}}}_{g,s}({\tilde{x}})\right] , \end{aligned}$$
(3.5)
Fig. 2
figure 2

Tube and trinion with maximal \(C_N\) punctures

where \( {{\mathcal {I}}}\left[ {{\mathcal {C}}}_{g,s}\right] \) is the index of a theory obtained by compactifications of 6d theory on the Riemann surface \({{\mathcal {C}}}_{g,s}\) and \(K_3^{C}(x,y,z)\) is the index of the trinion theory shown in Fig. 2b, while \({\bar{K}}_3^{C}(x,y,z)\) is its conjugate. Definition and properties of the elliptic \(\Gamma \)-function \(\Gamma _e\left( x\right) \) are given in Appendix A. Then, closing \(\textrm{SU}(2)_z\) and \(\textrm{SU}(2)_{{\tilde{z}}}\) minimal punctures with or without introduction of defects we can obtain finite difference operators. As discussed previously in order to close minimal punctures, we have to give certain weights to corresponding global symmetry fugacities \(z_1\) and \(z_2\). As a result, we get the following identity

$$\begin{aligned} \lim _{z_1\rightarrow Z_1^*,z_2\rightarrow Z_2^* } {{\mathcal {I}}}\left[ {{\mathcal {C}}}_{g,s}(x)\oplus {\bar{ {{\mathcal {T}}}}}^{{\mathcal {A}}}_{x,y,z_1}\oplus {{\mathcal {T}}}^{{\mathcal {A}}}_{y,{\tilde{x}},z_2} \right] \sim {{\mathcal {O}}}\cdot {{\mathcal {I}}}\left[ {{\mathcal {C}}}_{g,s}(x)\right] , \end{aligned}$$
(3.6)

where \( {{\mathcal {O}}}\) is some finite difference operator, and in case we close punctures without introducing defects we just obtain the identity operator.

Here for convenience, we will not compute the full expression (3.5) directly right away. Just as for \(A_N\) operators derived in Sect. 2, it appears to be much easier to perform calculation in a slightly different way. First we derive the index of the four-punctured sphere with zero flux, two maximal \(\textrm{USp}(2N)\) and two minimal \(\textrm{SU}(2)\) punctures. For this purpose, we perform S-gluing of two trinions \( {{\mathcal {T}}}^{AC}_{x,y,z}\) with \(\textrm{SU}(N+1)\) and \(\textrm{USp}(2N)\) maximal punctures and \(\textrm{SU}(2)\) minimal punctures each. This kind of trinions can be derived by appropriate gluing of \(A_NC_N\) tube theory shown in Fig. 2a, to one of the maximal punctures of \(A_N\) trinion introduced in [21]. Superconformal index of the resulting four-punctured sphere theory is given by:

$$\begin{aligned} K_4^C(x,{\tilde{x}},z,{\tilde{z}})=\kappa _N\oint \prod \limits _{i=1}^N\frac{dy_i}{2\pi i y_i} \prod \limits _{i\ne j}^{N+1}\frac{1}{\Gamma _e\left( \frac{y_i}{y_j}\right) } {\bar{K}}_3^{AC}({\tilde{x}},y,{\tilde{z}})K_3^{AC}(x,y,z). \end{aligned}$$
(3.7)

where \(K^{AC}_3(x,y,z)\) is the index of the trinion with y being \(\textrm{SU}(N+1)\) fugacity and \(\kappa _N\) is the usual constant given in (2.3). Geometrically this operation is shown in Fig. 5a.

Next we close one of the two minimal punctures by giving nonzero vev to one of the \(M_z\) moment maps given in (3.1). At the level of the index computations, this means that the corresponding global symmetry fugacity should be consistent with the vev, i.e., in our case we should fix \({\tilde{z}}\) fugacity to

$$\begin{aligned} {\tilde{z}}={\tilde{Z}}_{i;K,M}^*\equiv (pq)^{-\frac{1}{2}}w^{2\frac{(N+2)^2}{N+3}}{\tilde{a}}_i q^{-K}p^{-M}, \end{aligned}$$
(3.8)

where \({\tilde{a}}_i\) can be chosen arbitrary since all the expressions are symmetric w.r.t. permutations of \({\tilde{a}}_i\). Integers K and M correspond to an order of the derivative of the moment map in 34 and 12 planes correspondingly, i.e., we give vev to \(\langle \partial _{12}^M\partial _{34}^K{\widehat{M}}_i \rangle \ne 0\). Physically this corresponds to introducing various codimension two defects into 4d theory. For the first \(\textrm{SU}(2)_{{\tilde{z}}}\) minimal puncture, we choose to close it without introducing any defect, which in turn corresponds to \(K=M=0\) choice in (3.8).

At this value of \({\tilde{z}}\), the superconformal index (3.7) of the four-punctured sphere has a pole. Computing the residue at this pole, we obtain the index of the three-punctured sphere:

$$\begin{aligned} K_{(3;i,0)}^C(x,y,z)\sim \textrm{Res}_{{\tilde{z}}\rightarrow {\tilde{Z}}_{i;0,0 }^*}K_4^C(x,y,z,{\tilde{z}}) \end{aligned}$$
(3.9)

Here, subscript (3; i, 0) refers to the fact that we obtain three-punctured sphere by closing minimal puncture of the four-punctured sphere choosing \({\tilde{a}}_i\) for the vev in (3.8) and not introducing a defect, which corresponds to the choice \(K=M=0\) in the same equation.

Calculation of (3.7) and (3.9) results in the theory shown in Fig. 3. Detailed derivations of this section can be found in Appendix C.

Fig. 3
figure 3

Three-punctured sphere theory with two maximal \(\textrm{USp}(2N)\) punctures and one minimal \(\textrm{SU}(2)\) puncture obtained after gluing (3.7) and closing one minimal \(\textrm{SU}(2)_{{\tilde{z}}}\) puncture without defect. Solid and dashed line attached to the gauge node denotes multiplet in AS and \({\overline{AS}}\) representation correspondingly

Now that we have obtained the desired three-punctured sphere theory we are ready to derive A\(\Delta \)O. For this purpose, we should close the remaining minimal \(\textrm{SU}(2)_z\) puncture by giving non-trivial vev \(\langle \partial _{12}^M\partial _{34}^KM \rangle \ne 0\) to a derivative of one of its moment maps. In order to get zero total flux through the resulting two-punctured sphere, we should choose the same moment map as we did closing \(\textrm{SU}(2)_{{\tilde{z}}}\) minimal puncture previously. The difference is that now we have to choose non-trivial derivative that is at least one of K and M numbers is not zero. Physically this corresponds to introducing codimension-two defect into the tube theory with two maximal punctures as shown in Fig. 5b. At the level of the index, we should give the following value to the z-fugacity:

$$\begin{aligned} z=Z_{i;K,M}^*\equiv (pq)^{-\frac{1}{2}}w^{-2\frac{(N+2)^2}{N+3}}{\tilde{a}}_i^{-1} q^{-K}p^{-M}, \end{aligned}$$
(3.10)

Then, according to (3.6), capturing the corresponding pole and gluing the resulting tube with two maximal \(\textrm{USp}(2N)\) punctures to an arbitrary theory with at least one puncture of this type as shown in Fig. 5c, we obtain an A\(\Delta \)O. However, technically it sometimes appears not to be as straightforward. In particular, if we perform these operations with the expression (C.12) instead of A\(\Delta \)O, we obtain some integral-finite difference operator. It is highly possible that in fact this operator can be written in the form of A\(\Delta \)O. However, technical issues make it too difficult task and we leave it for future investigation.

Here, we will instead concentrate on the derivation of next to the lowest rank \(C_2\) operator. This case is simpler to analyze. For example, in our trinion theory shown in Fig. 3, when \(N=2\), the \({\overline{AS}}\) multiplet of the right \(\textrm{SU}(3)\) gauge node becomes just anti-fundamental, and a chain of duality transformations can be used to simplify the theory. These calculations are summarized in Appendix C. As a result, we obtain single-node \(\textrm{SU}(6)\) gauge theory shown in Fig. 4 with the corresponding index specified in (C.13).

Fig. 4
figure 4

Three-punctured sphere theory with two maximal \(\textrm{USp}(4)\) and one minimal \(\textrm{SU}(2)\) punctures. Dashed line starting and ending on the gauge \(\textrm{SU}(6)\) node as previously corresponds to the matter in the \({\overline{AS}}\) representation

Finally, we can close \(\textrm{SU}(2)_z\) puncture. In particular, we give the following weight consistent with (3.8) to the fugacity z:

$$\begin{aligned} z=Z^*_{i;K,0}=(pq)^{-\frac{1}{2}}w^{-\frac{32}{5}}a_i^{-1}q^{-K}. \end{aligned}$$
(3.11)
Fig. 5
figure 5

On the figures above we summarize all the steps we go through in order to derive \(C_2\) A\(\Delta \)O given in (3.14). First as shown on the (a), we glue two trinions with two maximal \(\textrm{USp}(4)\) (shown with green) and one minimal \(\textrm{SU}(2)\) (shown with orange) punctures each. For one of the trinions we conjugate all the charges in order to perform S-gluing. At the level of the superconformal index, this operation is expressed in (3.7). Next, as shown on the (b), we close two minimal punctures of the four-punctured sphere. This operation is performed by giving vev \(\langle \partial _+^K{\widehat{M}}_i\rangle \ne 0\) to holomorphic derivative of one of the moment map operators with the U(1) charge \(h_i\). As the result, we obtain tube theory with two maximal \(\textrm{USp}(4)\) punctures and codimension-two defect introduced. On the Figure we denote this defect with the red ring. Finally, as the last step of our algorithm shown on the (c), we glue this tube with the defect to an arbitrary surface with at least one maximal \(\textrm{USp}(4)\) puncture. This results in the action of the set of certain A\(\Delta \)Os on the index of the original theory (color figure online)

Performing this closure, we obtain the index of the theory for two-punctured sphere with two \(\textrm{USp}(4)\) maximal punctures

$$\begin{aligned} K_{(2;i;K,0)}^C(x,{\tilde{x}})\sim \textrm{Res}_{z\rightarrow Z_{i;K,0 }^*}K_{(3;i,0)}^C(x,{\tilde{x}},z) \end{aligned}$$
(3.12)

The index itself is specified in (C.18). This time it does not have natural gauge theory interpretation in case of general K due to the presence of the codimension-two defect.

As a final step of our derivation, we glue the obtained tube with the defect to an arbitrary \({\mathcal {N}}=1\) theory with at least one maximal \(\textrm{USp}(4)\) puncture and obtain A\(\Delta \)O as follows:

$$\begin{aligned} {{\mathcal {O}}}^{(C_2;h_{k};K,0)}_x\cdot {{\mathcal {I}}}(x)\sim & {} \frac{\left( q;q\right) _\infty ^2\left( p;p\right) _\infty ^2}{2^2\cdot 2!}\oint \frac{d{\tilde{x}}_{1,2}}{2\pi i{\tilde{x}}_{1,2}}\frac{1}{\Gamma _e\left( {\tilde{x}}_{1,2}^{\pm 2}\right) \Gamma _e\left( {\tilde{x}}_1^{\pm 1}{\tilde{x}}_2^{\pm 1}\right) }\nonumber \\{} & {} \times K_{(2;i;K,0)}^C(x,\,{\tilde{x}}) {{\mathcal {I}}}({\tilde{x}}) \end{aligned}$$
(3.13)

Details of this calculation can be found in Appendix C. It leads to the following expression for A\(\Delta \)O:

figure d

where the constant \({\tilde{C}}^k_{\vec {K}}\) is given by

figure e

We have also introduced notation \(h_i\), \(i=1,\dots ,10\) for U(1) charges of the moment maps we act on. Notice that according to (3.1) moment maps of both minimal and maximal punctures have the very same charges, so that in fact

$$\begin{aligned} h_i=w^{\frac{32}{5}}{\tilde{a}}_i,\qquad h\equiv \prod \limits _{j=1}^{10} h_j=w^{64}, \end{aligned}$$
(3.16)

for both charges of the maximal punctures we act on and minimal punctures we close. We have also introduced here an overall U(1) charge h. The operator itself is labeled by an index k according to the choice of U(1) charge \(h_k\) of the moment map we give vev to in order to close the minimal puncture and to introduce defects into the theory.

The first sum in the expression is performed over all possible partitions \(\vec {K}=(k_1,\cdots ,k_6)\) of the integer K. Finally, the products in (3.14) should be understood as ordered ones, i.e., one should assume \(\prod _{i=n_1}^{n_2}\) is just 1 if \(n_2<n_1\) since there are no terms in the product. The coefficient \({\tilde{C}}^k_{\vec {K}}\) is chosen so that the x-independent factors of the highest order shift terms \(\Delta _q^{\pm K}(y_{1,2})\) are just one.

As expected obtained operators differ from the canonical \(BC_n\) vD. In order to obtain the latter one, we should instead consider compactifications of 6d rank-Q E-string theory using the very same approach described in details on our paper. One of the most important difference from the \(BC_n\) vD model is the number of parameters it depends on. Higher-rank (i.e., rank higher than one) vD model depends on p, q and another 9 parameters. This number of parameters is independent of the rank of the BC-type root system. Meanwhile, our \(C_2\) operators depend on p, q and another 10 parameters. Moreover, in the present paper due to technical difficulties we failed to derive higher-rank \(C_N\) generalizations. But from our construction we can expect them to depend on \(2N+6\) parameters corresponding to U(1) charges of the moment maps.Footnote 5

Notice that the expression above is not the most general since we could have closed \(\textrm{SU}(2)_z\) minimal puncture using general K and M integers in (3.10), but we will concentrate here only on the case of non-trivial K while putting \(M=0\). As we see, the operator appears to be quite complicated and it is hard to analyze it. Instead it is useful to write down basic operator corresponding to the choice of \(K=1\), i.e., the simplest non-trivial case. Taking into account six possible partitions \(\vec {K}\), we can write down explicit form of A\(\Delta \)O:

figure f

One thing to be noticed is that just as in the case of vD model, the constant term choice is not unique. What really defines this constant term are the following properties. First of all it can be noticed that \(W^{(h_k;1,0)}(x,h)\) is an elliptic function in both \(x_1\) and \(x_2\) with periods 1 and p. In the fundamental domain, this elliptic function has poles located at:

$$\begin{aligned} x_i=sq^{\pm \frac{1}{2}},\qquad x_i=sq^{\pm \frac{1}{2}}p^{\frac{1}{2}}, \quad s=\pm 1, \end{aligned}$$
(3.18)

which are in fact exactly the same as the poles of vD model. Corresponding residues are given by:

$$\begin{aligned} \textrm{Res}_{x_i=sq^{\frac{1}{2}}}W^{(C_2;h_l;1,0)}(x,h)&=-s\frac{q^{\frac{1}{2}}\prod \limits _{k=1}^{10}\theta _p\left( (pq)^{\frac{1}{2}}sq^{-\frac{1}{2}}h_k\right) }{2\left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) }\prod \limits _{j\ne i}^2 \frac{\theta _p\left( (pq)^{\frac{1}{2}}h_lx_j^{\pm 1}\right) }{\theta _p\left( sq^{-\frac{1}{2}}x_j^{\pm 1} \right) }\,, \nonumber \\ \textrm{Res}_{x_i=sq^{-\frac{1}{2}}}W^{(C_2;h_l;1,0)}(x,h)&=s\frac{q^{-\frac{1}{2}}\prod \limits _{k=1}^{10}\theta _p\left( (pq)^{\frac{1}{2}}sq^{-\frac{1}{2}}h_k\right) }{2\left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) }\prod \limits _{j\ne i}^2 \frac{\theta _p\left( (pq)^{\frac{1}{2}}h_lx_j^{\pm 1}\right) }{\theta _p\left( sq^{-\frac{1}{2}}x_j^{\pm 1} \right) }\,, \nonumber \\ \textrm{Res}_{x_i=sp^{\frac{1}{2}}q^{\frac{1}{2}}}W^{(C_2;h_l;1,0)}(x,h)&=-s\frac{ph^{-\frac{1}{2}}\prod \limits _{k=1}^{10}\theta _p\left( sh_k\right) }{2\left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) }\prod \limits _{j\ne i}^2 \frac{\theta _p\left( (pq)^{\frac{1}{2}}h_lx_j^{\pm 1}\right) }{\theta _p\left( (pq)^{-\frac{1}{2}}sx_j^{\pm 1} \right) }\,, \nonumber \\ \textrm{Res}_{x_i=sp^{\frac{1}{2}}q^{-\frac{1}{2}}}W^{(C_2;h_l;1,0)}(x,h)&=s\frac{pq^{-1}h^{-\frac{1}{2}}\prod \limits _{k=1}^{10}\theta _p\left( sh_k\right) }{2\left( p;p\right) _\infty ^2\theta _p\left( q^{-1}\right) }\prod \limits _{j\ne i}^2 \frac{\theta _p\left( (pq)^{\frac{1}{2}}h_lx_j^{\pm 1}\right) }{\theta _p\left( (pq)^{-\frac{1}{2}}sx_j^{\pm 1} \right) }\,. \end{aligned}$$
(3.19)

This concludes our derivation of the \(C_2\) A\(\Delta \)O. All possible details of these derivations can be found in Appendix C.

4 Properties of the \(C_2\) operators

There is a number of interesting properties that the operators we derived should possess by construction. In this section, we discuss checks and, where it is possible, give proofs of these properties using explicit expressions we have derived.

The most important but also most complicated property to discuss is the so-called kernel property of our operators. The main idea behind it is that the superconformal index of any \({\mathcal {N}}=1\) theory obtained in the compactifications of 6d minimal \((D_5,D_5)\) conformal matter theory is a kernel function of our operators. The kernel function is defined by the following mathematical identity:

$$\begin{aligned} {{\mathcal {O}}}_z^{(G_1;h_i;r,m)}\cdot {{\mathcal {I}}}\left[ {{\mathcal {C}}}_{g,s}[z,u]\right] = {{\mathcal {O}}}_u^{(G_2;h_i;r,m)}\cdot {{\mathcal {I}}}\left[ {{\mathcal {C}}}_{g,s}[z,u]\right] . \end{aligned}$$
(4.1)

Physically we consider the superconformal index of the theory obtained in the 6d compactification on the Riemann surface \({{\mathcal {C}}}_{g,s}[z,u]\) with at least two maximal punctures parametrized by the fugacities z and u correspondingly. In general, there can be as many punctures as we want. The claim here is that any such superconformal index plays the role of the kernel function for the derived A\(\Delta \)Os according to (4.1). Namely we can act with our operators on different punctures and we should always obtain the same result. This property is expected to hold due to an argument coming from the geometry of compactification shown in Fig. 6. Equivalently, it can be understood from the invariance of the superconformal indices under S duality transformations.

In principle, we can choose any sphere with multiple punctures to check the kernel property. Natural simplest candidate would be WZW model for the two-punctured sphere with two maximal \(\textrm{USp}(2N)\) punctures for which the index is given in (C.18) with \(K=0\). However, we will check a more interesting and trickier case where two punctures are of different types. Namely we take the tube theory shown in Fig. 2a which has one \(\textrm{SU}(N+1)\) and one \(\textrm{USp}(2N)\) maximal puncture. Then for \(N=2\), the kernel property (4.1) in this case reads:

$$\begin{aligned} {{\mathcal {O}}}^{(A_2;{\tilde{h}}_i^{-1};r,m)}_{y}K_2^{AC}(x,y)= {{\mathcal {O}}}^{(C_2;h_i;r,m)}_{x}K_2^{AC}(x,y), \end{aligned}$$
(4.2)

where \(K_2^{AC}(x,y)\) is the index of the tube specified in (C.5) with x and y being fugacities of \(\textrm{USp}(4)\) and \(\textrm{SU}(3)\) punctures correspondingly. Operators \( {{\mathcal {O}}}^{(A_2;{\tilde{h}}_i;1,0)}_{y}\) and \( {{\mathcal {O}}}^{(C_2;h_i;1,0)}_{x}\) are given in (2.13) and (3.17) correspondingly. Notice that in order for the kernel property to work we need to close minimal punctures in the same way on two sides of the equation, i.e., we have to give vev to the same moment map of the minimal puncture. In particular, in \(A_2\) case we should give vevs to the minimal puncture moment map of the charge \({\tilde{h}}_i^{-1}{\tilde{h}}^{\frac{1}{4}}\) where \({\tilde{h}}_i\) are given in (2.5). At the same time, in \(C_2\) case minimal puncture is closed by giving vev to the moment map of charge \(h_i\) specified in (3.16). Using the map (3.4), we can see that the only map that matches between the two is

$$\begin{aligned} {\tilde{h}}_{10}^{-1}{\tilde{h}}^{\frac{1}{4}}=v^{-24}w^{-8}=w^{-\frac{32}{5}}{\tilde{a}}_{10}^{-1}=h_{10}. \end{aligned}$$
(4.3)

If we would like to check the kernel property for other ways of closing minimal punctures, we should consider \(A_2\) operators (2.10) with the non-flipped moment maps so that in this case we close the minimal puncture with the vevs of the moment map with charge \({\tilde{h}}_i{\tilde{h}}^{-\frac{1}{4}}\). Because of this complication, here we consider only closing with the vev given to the moment map of charge (4.3). The proofs for other operators should work identically. Also in our proof we restrict ourselves to only basic operators (3.17) and (2.13), i.e., we fix \(r=1,m=0\) (or equivalently \(m=1,r=0\) on both sides of the kernel equation (4.2). We have to do it since higher operators specified in (3.14) and (2.10) are too complicated for the analysis. However, the kernel property (4.2) should also work for these higher operators as well.

Fig. 6
figure 6

On this figure, we represent the argument in favor of the kernel property of \({\mathcal {N}}=1\) indices. For this, we consider an index of a theory obtained in the compactification of the minimal \((D_5,D_5)\) conformal matter theory on a generic Riemann surface with at least two maximal punctures (shown as colored disks on the figure) of any types with fugacities x and y as well as one minimal \(\textrm{SU}(2)\) puncture (shown as red crosses) with the fugacity z. For example in case of the \(A_2C_2\) tube and kernel property (4.2), one maximal puncture has \(\textrm{USp}(4)\) global symmetry, while the second one has \(\textrm{SU}(3)\) symmetry. Then, we close the minimal puncture by giving space-dependent vev \(\langle \partial _{12}^m\partial _{34}^k M \rangle \ne 0\) to one of the moment maps of this puncture. As discussed in the paper, this corresponds to the introduction of the codimension-two defect into 4d effective description and results in the action of the A\(\Delta \)O on one of the maximal punctures of the index of the theory corresponding to the Riemann surface with only two maximal punctures. This operation can be performed in different duality frames. In each frame, we obtain operators acting on one of the two punctures. Since the result of the calculation should not depend on the duality frame we choose, we conclude that the action of two, possibly different operators, on two different punctures leads to two expressions equal to each other. Hence, we arrive to the kernel property Eq. (4.1) (color figure online)

Since we precisely know explicit expressions for both operators (2.13) and (3.17) as well as supposed kernel function (C.5), it is straightforward to check this kernel identity. Acting with finite difference operators of two kinds on the kernel, we get two algebraic functions. Equality of these functions implies validity of the kernel property (4.2). In Appendix D, we give details of the analytic proof of this identity.

Another important property of the derived A\(\Delta \)Os is their commutation with each other which also directly follows from the S duality of our compactification construction as shown in Fig. 7. Since it does not depend which duality frame we close the minimal punctures in, it also does not matter in which order two different A\(\Delta \)Os act on the index. Hence, we conclude that all of operators we derived should commute with each other:

$$\begin{aligned}{} & {} \left[ {{\mathcal {O}}}_x^{(C_2;h_a;K_1,M_1)},\, {{\mathcal {O}}}_x^{(C_2;h_b;K_2,M_2)}\right] =0,\quad \nonumber \\{} & {} \quad \quad \forall ~ ~a,b=1,\ldots ,10;~ K_{1,2}\ge 0;~M_{1,2}\ge 0; \end{aligned}$$
(4.4)
Fig. 7
figure 7

On this figure, we represent the argument in favor of commutation of A\(\Delta \)Os derived in the present paper. For this, we start with a theory obtained in compactification of 6d minimal \(\left( D_5,D_5\right) \) conformal matter on an arbitrary Riemann surface. This surface has one maximal \(\textrm{USp}(4)\) puncture (orange disk) with fugacity x and two minimal \(\textrm{SU}(2)\) punctures (red crosses) with fugacities \(z_1\) and \(z_2\) correspondingly. Then, we introduce codimension-two defects into this theory giving space-dependent vevs to some of the moment maps \(M_{h_1}\) and \(M_{h_2}\) correspondingly. Giving this vev is equivalent to closing puncture, and at the level of the index computations leads to an action of two different A\(\Delta \)Os. We can perform this operation in any duality frame, and the result should not depend on the choice of a particular frame. Hence, the action of operators on the index does not depend on their order. Since the compactification surface, and hence, the effective 4d descriptions were chosen arbitrarily, we conclude that the operators themselves also commute (color figure online)

Here, \(h_a\) and \(h_b\) are moment maps we give vev to in order to close two minimal punctures and \(K_{1,2}, M_{1,2}\) are numbers of holomorphic/anti-holomorphic derivatives. Unfortunately it is very hard to prove or even to check these relations for general operators (3.14) of the full tower. However, we can perform checks of the commutation relations (4.4) for the basic operators, i.e., for the case when \(K_{1,2}\) and \(M_{1,2}\) are taking values 0 and 1. In particular, we show that:

$$\begin{aligned} \left[ {{\mathcal {O}}}_x^{(C_2;h_a;1,0)},\, {{\mathcal {O}}}_x^{(C_2;h_b;1,0)}\right]= & {} \left[ {{\mathcal {O}}}_x^{(C_2;h_a;0,1)},\, {{\mathcal {O}}}_x^{(C_2;h_b;0,1)}\right] \nonumber \\= & {} \left[ {{\mathcal {O}}}_x^{(C_2;h_a;1,0)},\, {{\mathcal {O}}}_x^{(C_2;h_b;0,1)}\right] \nonumber \\= & {} 0,\qquad \forall ~ a,b =1,\dots ,10. \end{aligned}$$
(4.5)

In Appendix E, we check these relations. In particular, using ellipticity properties of \(\theta _p\left( x\right) \) function we prove that the action of the third type of the commutators on an arbitrary trial function is zero. Hence, these commutators are zero themselves. For the first and second type of commutators, it is hard to prove analytically that their action on a trial function is zero. Instead we check these identities perturbatively in p and q expansion. These checks performed up to a sufficiently high order suggest that both of the commutators are indeed zero.

5 Discussion and outlook

In this paper, we have considered 4d description of the compactification of the minimal \((D_5,D_5)\) conformal matter theory on a Riemann surface with \(\textrm{USp}(4)\) maximal punctures. Using this 4d description and introducing codimension-two defect in corresponding theory, we derive \(C_2\) generalization of the van Diejen model. In particular, we obtain an infinite set of analytic difference operators acting on the maximal puncture of \(C_2\) type. Different operators of our set correspond to different ways of closing minimal punctures on the compactification surfaces. They are organized in a decuplet of basic operators and an infinite tower of operators on top of each of the basic ones.

The operators we obtain are supposed to satisfy interesting and important properties following directly from the geometry of compactifications. First such property is the commutation of all operators. In our paper, using combination of residue computations and perturbative expansions we show that at least all basic operators indeed commute with each other. Second important property is that the superconformal index of any compactification of the minimal \((D_5,D_5)\) conformal matter theory on a surface with several \(\textrm{USp}(4)\) punctures is the kernel function of our difference operators. As an example of such kernel function, we considered a tube theory corresponding to the compactification on a sphere with one \(\textrm{USp}(4)\) and one \(\textrm{SU}(3)\) puncture. In order to prove the kernel property, we should act on two punctures with two different operators: \(A_2\) operator derived in [19] from one side and \(C_2\) operator derived in the present paper from the other. This fact makes corresponding kernel function very interesting for the study. In the paper, we managed to prove this kernel property fully analytically for the basic \(A_2\) and \(C_2\) operators.

Another result of our paper is derivation of the full tower of \(A_N\) generalization of van Diejen operator originating from the compactification of the minimal \((D_{N+3},D_{N+3})\) conformal matter theory on a Riemann surface with \(\textrm{SU}(N+1)\) maximal punctures. We have already derived basic operators of this kind in our previous paper [19]. However, the tower of operators was missing and we filled this gap in the present paper.

The present paper is another brick in our program of establishing dictionary between compactifications of 6d SCFTs and integrable analytic difference operators and studying this dictionary in details. There are plenty of ways this research can be continued in. In particular, there should be separate difference operator for each possible 5d compactification of 6d theory, or equivalently for each possible puncture type from 4d prospective. So far several results have been observed in this frame. First of all \(BC_1\) van Diejen model was observed in E-string compactifications in several ways [11, 19]. Also previously unknown \(A_2\) and \(A_3\) operators were derived using compactifications of the minimal conformal matter theories of types \(\textrm{SU}(3)\) and \(\textrm{SO}(8)\) [10]. Finally, we have considered compactifications of the minimal \((D_{N+3},D_{N+3})\) conformal matter theories on a Riemann surfaces with \(A_N\)- and \(C_2\) (in case of \((D_5,D_5)\) 6d theory)-type punctures. However, there are plenty of examples of 6d SCFT compactifications that are known up to date [25,26,27,28,29,30,31,32,33,34]. Our main goal is to extend our results to more examples of compactifications and ideally establish a dictionary between them and integrable models. Such dictionary can from one point of view shed light on physics of 6d, 5d and 4d supersymmetric gauge theories. From the other point of view, this research program can lead to important results in the field of integrable systems.

First of all as a continuation of the present paper, it is natural to consider compactifications of the minimal \((D_{N+3},D_{N+3})\) conformal matter on Riemann surfaces with other types of punctures. In first place, these are of course punctures of \(C_N\) type for general N. In the present work, we failed to derive explicit form of the corresponding A\(\Delta \)O because of the technical difficulties but it is worth trying to overcome them. Second candidate for the study is \(\left( A_1\right) ^N\) puncture. Corresponding 4d description was derived in [20]. It would be interesting to derive explicit form of the corresponding difference operators and study their properties. For example, we immediately know examples of the kernel functions of these operators even before deriving them explicitly.

Second candidate for the studies are non-minimal \((D_{N+3},D_{N+3})\) conformal matter theories, which are obtained as worldvolume theories of the stack of k M5 branes probing \(D_{N+3}\) singularity [26]. In this case, the known 5d gauge theory, and hence the puncture type on Riemann surface, is a direct generalization of \(\left( A_1 \right) ^N\) case of minimal conformal matter compactification discussed above. It corresponds to the \(\left( A_{k-1}\right) ^2\times \left( A_{2k-1}\right) ^N\times \left( A_{k-1}\right) ^2\) gauge theory in 5d and hence the puncture type with the same global symmetry. Trinion theories with two maximal and one minimal punctures, which are building blocks in the construction of A\(\Delta \)Os, are also known [35]. So construction of the corresponding difference operators should be straightforward though can happen to be technically complicated.

Finally, one more thing to be made in this direction is the study of the rank-Q E-string theory compactifications along the same lines of research. Unfortunately so far corresponding trinion theories were not obtained but the tube theories were already derived in [36, 37]. Although we cannot derive corresponding A\(\Delta \)O using only tube theories, in this case we can naturally conjecture it to be \(BC_n\) van Diejen model. We can at least test this conjecture by checking corresponding kernel property.

Another possible direction of the research is related to the study of the properties of the operators we derived. Most interesting questions here are related to the eigenfunctions of these operators. Deriving full eigenfunctions is of course difficult and most probably impossible. Realistically we can try to derive certain limits of these eigenfunction. In case we succeed these eigenfunctions can play the same role for indices of \({\mathcal {N}}=1\) theories as Macdonald and Schur polynomials played for the indices of \({\mathcal {N}}=2\) theories [16,17,18].

It would also be interesting to establish connections of our research program with other integrable models emerging in supersymmetric gauge theories. For example, integrable systems derived using compactifications of the worldvolume theory of k M5 branes probing \(A_{N-1}\) singularity were shown to be related to the set of transfer matrices [9, 38,39,40]. It would be interesting to understand if similar connection emerges in case of \(D_{N+3}\) singularity.

Also, a relation of E-string theory to the van Diejen model was observed in [41] using quantization of the corresponding Seiberg–Witten curve. Later, authors of this paper generalized their result to 6d (1, 0) SCFTs with \(\textrm{SO}(N)\) gauge group and \((N-8)\) fundamental flavors obtaining yet another set of difference operators [42]. It would be interesting to clarify precise relation of our construction with the methods and results of these papers.