Skip to main content
Log in

Higher spin \({{\mathfrak {s}}}{{\mathfrak {l}}}_2\)R-matrix from equivariant (co)homology

  • Published:
Letters in Mathematical Physics Aims and scope Submit manuscript

Abstract

We compute the rational \({{\mathfrak {s}}}{{\mathfrak {l}}}_2\)R-matrix acting in the product of two spin-\(\ell \over 2\) (\({\ell \in {\mathbb {N}}}\)) representations, using a method analogous to the one of Maulik and Okounkov, i.e., by studying the equivariant (co)homology of certain algebraic varieties. These varieties, first considered by Nekrasov and Shatashvili, are typically singular. They may be thought of as the higher spin generalizations of \(A_1\) Nakajima quiver varieties (i.e., cotangent bundles of Grassmannians), the latter corresponding to \(\ell =1\).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Alvarez-Gaumé, L., Freedman, D.Z.: Potentials for the supersymmetric nonlinear \(\sigma \)-model. Commun. Math. Phys. 91(1), 87–101 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  2. Brion, M.: Poincaré duality and equivariant (co)homology. Michigan Math. J. 48, 77–92 (2000) Dedicated to William Fulton on the occasion of his 60th birthday. https://doi.org/10.1307/mmj/1030132709

  3. Chari, V., Pressley, A.: A Guide to Quantum Groups. Cambridge University Press, Cambridge (1994)

    MATH  Google Scholar 

  4. Derkachov, S., Chicherin, D.: Matrix factorization for solutions of the Yang-Baxter equation. J. Math. Sci. 213(5), 723–742 (2016)

    Article  MathSciNet  Google Scholar 

  5. Derkachov, S., Chicherin, D.: Matrix factorization for solutions of the Yang-Baxter equation. Zap. Nauchn. Semin. LOMI 433, 156 (2015). https://doi.org/10.1007/s10958-016-2734-0

    Article  MATH  Google Scholar 

  6. Donagi, R., Sharpe, E.: GLSM’s for partial flag manifolds. J. Geom. Phys. 58, 1662–1692 (2008) arXiv:0704.1761, https://doi.org/10.1016/j.geomphys.2008.07.010

  7. Drinfeld, V.G.: Quantum groups. J. Sov. Math. 41, 898–915 (1988)

    Article  Google Scholar 

  8. Drinfeld, V.G.: Quantum groups. Zap. Nauchn. Semin. LOMI 155, 18–49 (1986)

    Google Scholar 

  9. Faddeev, L.D.: How algebraic Bethe Ansatz works for integrable model, Relativistic gravitation and gravitational radiation. In: Proceedings, School of Physics, Les Houches, France, September 26–October 6, 1995, 1996, pp. 149–219. arXiv:hep-th/9605187

  10. Gaiotto, D., Moore, G.W., Neitzke, A.: Wall-crossing, Hitchin systems, and the WKB approximation. Adv. Math. 234, 239–403 (2013) arXiv:0907.3987. https://doi.org/10.1016/j.aim.2012.09.027

  11. Grayson, D., Stillman, M.: Macaulay2, a software system for research in algebraic geometry. Available at http://www.math.uiuc.edu/Macaulay2/. Accessed 22 June 2020

  12. Humphreys, J.: Conjugacy Classes in Semisimple Algebraic Groups, Mathematical Surveys and Monographs, vol. 43. American Mathematical Society, Providence, RI (1995)

    Google Scholar 

  13. Jantzen, J.C.: Nilpotent Orbits in Representation Theory. Lie Theory, Progress in Mathematics, vol. 228, pp. 1–211. Birkhäuser, Boston (2004)

    MATH  Google Scholar 

  14. Kirillov, A.N., Reshetikhin, N.Yu.: The Yangians, Bethe Ansatz and combinatorics. Lett. Math. Phys. 12(3), 199–208 (1986). https://doi.org/10.1007/BF00416510

  15. Kirwan, F.C.: Cohomology of Quotients in Symplectic and Algebraic Geometry. Mathematical Notes. Princeton University Press, Princeton (1984)

    MATH  Google Scholar 

  16. Kulish, P.P., Reshetikhin, N.Yu., Sklyanin, E.K.: Yang-Baxter equation and representation theory. Lett. Math. Phys. 5, 393–403 (1981)

  17. Kulish, P.P., Sklyanin, E.K.: On the solution of the Yang-Baxter equation. J. Sov. Math. 19, 1596–1620 (1982)

    Article  Google Scholar 

  18. Kulish, P.P., Sklyanin, E.K.: On the solution of the Yang-Baxter equation. Zap. Nauchn. Semin. LOMI 95, 129 (1980). https://doi.org/10.1007/BF01091463

    Article  MathSciNet  Google Scholar 

  19. Mangazeev, V.: On the Yang–Baxter equation for the six-vertex model. Nucl. Phys. B 882, 70–96. arXiv:1401.6494, https://doi.org/10.1016/j.nuclphysb.2014.02.019 (2014)

  20. Maulik, D., Okounkov, A.: Quantum groups and quantum cohomology (2012). arXiv:1211.1287

  21. Nakajima, H.: Instantons on ALE spaces, quiver varieties, and Kac-Moody algebras. Duke Math. J. 76(2), 365–416 (1994). https://doi.org/10.1215/S0012-7094-94-07613-8

    Article  MathSciNet  MATH  Google Scholar 

  22. Nakajima, H.: Quiver varieties and Kac-Moody algebras. Duke Math. J. 91(3), 515–560 (1998). https://doi.org/10.1215/S0012-7094-98-09120-7

    Article  MathSciNet  MATH  Google Scholar 

  23. Nakajima, H.: Quiver varieties and finite-dimensional representations of quantum affine algebras. J. Am. Math. Soc. 14(1), 145–238 (2001). arXiv:math/9912158, https://doi.org/10.1090/S0894-0347-00-00353-2

  24. Nakajima, H.: Reflection functors for quiver varieties and Weyl group actions. Mathematische Annalen 327(4), 671–721 (2003). https://doi.org/10.1007/s00208-003-0467-0

    Article  MathSciNet  MATH  Google Scholar 

  25. Nekrasov, N.A., Shatashvili, S.L.: Supersymmetric vacua and Bethe Ansatz. Nucl. Phys. B Proc. Suppl. 192/193, 91–112 (2009). arXiv:0901.4744, https://doi.org/10.1016/j.nuclphysbps.2009.07.047

  26. Sklyanin, E.K.: Classical limits of the \(SU(2)\)-invariant solutions of the Yang-Baxter equation. J. Math. Sci. 40(1), 93–107 (1988)

    Article  MathSciNet  Google Scholar 

  27. Sklyanin, E.K.: Classical limits of the \(SU(2)\)-invariant solutions of the Yang-Baxter equation. Zap. Nauchn. Semin. LOMI 146, 119–136 (1985)

    MathSciNet  MATH  Google Scholar 

  28. Varagnolo, M.: Quiver varieties and Yangians. Lett. Math. Phys. 53, 273–283 (2000). arXiv:math/0005277

    Article  MathSciNet  Google Scholar 

  29. Yang, Y., Zhao, G.: On two cohomological Hall algebras. Proc Roya Soc Edinb 150, 1581–1607 (2020)

    Article  MathSciNet  Google Scholar 

  30. Zamolodchikov, A.B., Fateev, V.A.: Model factorized \(S\) matrix and an integrable Heisenberg chain with spin 1. Sov. J. Nucl. Phys. 32, 298–303 (1980)

  31. Zamolodchikov, A.B., Fateev, V.A.: Model factorized \(S\) matrix and an integrable Heisenberg chain with spin 1. Yad. Fiz. 32, 581 (1980)

  32. Zinn-Justin, P.: Lectures on geometry, quantum integrability and symmetric functions, lecture notes from a course at HSE, Moscow. http://www.lpthe.jussieu.fr/~pzinn/summary.pdf (2015)

  33. Zinn-Justin, P.: Quiver varieties and the quantum Knizhnik–Zamolodchikov equation. Theor. Math. Phys. 185(3), 1741–1758 (2015). arXiv:1502.01093, https://doi.org/10.1007/s11232-015-0376-x

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Paul Zinn-Justin.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

PZJ was supported by ARC Grant FT150100232. DB wishes to thank D. Lüst and A. A. Slavnov for support. PZJ wishes to thank A. Knutson, Y. Yang, G. Zhao for valuable discussions. Computerized checks of the results of this paper were performed with the help of Macaulay2 [11].

Appendices

Appendix A: \({\mathbb {Z}}_2\) invariance of \({\mathfrak {M}}\) and \({\mathfrak {M}}_1\)

In this section, we denote the dual weight by \(k^\vee =n\ell -k\).

Lemma 3

There are two isomorphisms:

\({\mathfrak {M}}(k, n, \ell )\simeq {\mathfrak {M}}(k^\vee , n, \ell )\) and \({\mathfrak {M}}_1(k, n, \ell )\simeq {\mathfrak {M}}_1(k^\vee , n, \ell )\).

Proof

Remembering that \({{\widetilde{Q}}}: {\mathbb {C}}^k\rightarrow {\mathbb {C}}^n\), we form the matrix \({\widetilde{A}}: {\mathbb {C}}^k\rightarrow {\mathbb {C}}^{\ell n}\) as follows:

$$\begin{aligned} {\widetilde{A}}=\begin{pmatrix}{{\widetilde{Q}}}\\ {{\widetilde{Q}}}\Phi \\ \vdots \\ {{\widetilde{Q}}}\Phi ^{\ell -1} \end{pmatrix} \end{aligned}$$
(62)

(We recall that \(\Phi ^\ell =0\).) Earlier, we introduced the stability condition \({{\mathrm {rank}}}\,{\widetilde{A}}=k\). The dimension of the kernel of \({\widetilde{A}}\) is therefore \(k^\vee \). We define a matrix \({\widetilde{B}}: {\mathbb {C}}^{k^\vee }\rightarrow {\mathbb {C}}^{\ell n}\) by the equation

$$\begin{aligned} {\widetilde{B}}^T{\widetilde{A}}=0 \end{aligned}$$
(63)

and the non-degeneracy requirement \({\mathrm {rank}}\,{\widetilde{B}}=k^\vee \). The matrix \({\widetilde{A}}\) was defined up to right multiplication by an element of \(GL(k, {\mathbb {C}})\), and the matrix \({\widetilde{B}}\) is accordingly defined up to right multiplication by the element of \(GL(k^\vee , {\mathbb {C}})\).

Now, due to the structure (62) of the matrix \({\widetilde{A}}\), multiplication by \(\Phi \) from the right essentially amounts to shifting the entries of A upwards by n rows, and filling the lower rows with zeros. Denoting such a shift matrix by \(S_n\), we write

$$\begin{aligned} {\widetilde{A}}\Phi =S_n {\widetilde{A}}\,. \end{aligned}$$
(64)

Now, multiplying (63) by \(\Phi \) from the right, we obtain

$$\begin{aligned} {\widetilde{B}}^T{\widetilde{A}}\Phi ={\widetilde{B}}^T S_n {\widetilde{A}}=0\, \end{aligned}$$
(65)

hence \( S_n^T{\widetilde{B}}\in {\mathrm {Ker}}({\widetilde{A}}^T)\). (Note that \(S_n^T\) acts on \({\widetilde{B}}\) by shifting by n rows downward.) Since \({\mathrm {Ker}}({\widetilde{A}}^T)\) is spanned by the columns of \({\widetilde{B}}\), we may write

$$\begin{aligned} S_n^T{\widetilde{B}}={\widetilde{B}}\Phi ^{\vee } ,\quad \quad \quad \Phi ^{\vee }\in {\mathrm {End}}({\mathbb {C}}^{k^\vee })\,. \end{aligned}$$
(66)

Since \((S_n)^\ell =0\), we find that \((\Phi ^{\vee })^\ell =0\). Moreover, we can rephrase (66) as the requirement that \({\widetilde{B}}\) has the form

$$\begin{aligned} {\widetilde{B}}=\begin{pmatrix}{{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -1} \\ {{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -2}\\ \vdots \\ {{\widetilde{Q}}}^\vee \end{pmatrix}. \end{aligned}$$
(67)

Therefore, we have defined a map

$$\begin{aligned} \{{{\widetilde{Q}}}, \Phi \}^s/GL(k, {\mathbb {C}}) \leftrightarrow \{{{\widetilde{Q}}}^\vee , \Phi ^\vee \}^s/GL(k^\vee , {\mathbb {C}}). \end{aligned}$$
(68)

We have therefore proved the duality of projective retracts \({\mathfrak {P}}(k, n, \ell )\simeq {\mathfrak {P}}(k^\vee , n, \ell )\).

We now wish to incorporate the variable Q in the above discussion. To this end, we form the \(\Phi \)-tower of Q that we will call \(A: {\mathbb {C}}^{\ell n}\rightarrow {\mathbb {C}}^k\), analogously to the way it was done in (62) for the \({{\widetilde{Q}}}\)-variable:

$$\begin{aligned} A=\{\Phi ^{\ell -1} Q , \ldots , \Phi Q , Q\}\,. \end{aligned}$$
(69)

In this case, the NS-equation \(\sum \nolimits _{i+j=\ell -1}\,\Phi ^{i}\,Q\,{\widetilde{Q}}\,\Phi ^{j}=0\) may be put in the form

$$\begin{aligned} A{\widetilde{A}}=0\,, \end{aligned}$$
(70)

which is the NS-generalization of the equation \(Q{\widetilde{Q}}=0\) in the hyper-Kähler case. To complete the duality, we impose the following condition:

$$\begin{aligned} ({\widetilde{A}}A)^T={\widetilde{B}}B\,. \end{aligned}$$
(71)

Note that in both sides we have matrices of size \(\ell n\times \ell n\). A simple consistency check of the above equations is obtained by multiplying from the left by \({\widetilde{A}}^T\). The l.h.s. is then \(({\widetilde{A}}A {\widetilde{A}})^T=0\) by (70), and the r.h.s. is \({\widetilde{A}}^T{\widetilde{B}}B=0\) by (63).

Since \({\widetilde{B}}\) has already been defined as (67) from (63), one can view Eq. (71) as an equation for B. Multiplying (71) by \({\widetilde{B}}\) from the right and using (63), we get \({\widetilde{B}}B {\widetilde{B}}=0\). Since \(B{\widetilde{B}}\in {\mathrm {End}}({\mathbb {C}}^{k^\vee })\) and \({\mathrm {rank}}({\widetilde{B}})=k^\vee \), it follows that

$$\begin{aligned} B {\widetilde{B}}=0\,, \end{aligned}$$
(72)

which is an analogue of (70).

It remains to show that B, just like A, may be written as a \(\Phi \)-tower of a basic element \(Q^\vee \). To this end, we multiply \(({\widetilde{A}}A)^T\) by \(S_n^T\) from the right and use (64): \(({\widetilde{A}}A)^T S_n^T=(S_n{\widetilde{A}}A)^T=({\widetilde{A}}\Phi A)^T=({\widetilde{A}}A S_n)^T=S_n^T({\widetilde{A}}A )^T\). Here, we have used \(\Phi A=A S_n\). By (71) this implies \({\widetilde{B}}B S_n^T=S_n^T {\widetilde{B}}B\). Using (66), \({\widetilde{B}}(B S_n^T-\Phi ^{\vee } B)=0\). Since \({\widetilde{B}}: {\mathbb {C}}^{k^\vee }\rightarrow {\mathbb {C}}^{\ell n}\) and \({\mathrm {rank}}({\widetilde{B}})=k^\vee \), we get \(B S_n^T-\Phi ^{\vee } B=0\), hence

$$\begin{aligned} B=\{ Q^\vee , \ldots , (\Phi ^\vee )^{\ell -2} Q^\vee , (\Phi ^\vee )^{\ell -1} Q^\vee \}\,. \end{aligned}$$
(73)

Thus, \({\mathfrak {M}}(k, n, \ell )\simeq {\mathfrak {M}}(k^\vee , n, \ell )\).

In order to prove the isomorphism \({\mathfrak {M}}_1(k, n, \ell )\simeq {\mathfrak {M}}_1(k^\vee , n, \ell )\), we will show that, under the map (68), we have \({\mathfrak {M}}^o_1(k, n, \ell )\leftrightarrow {\mathfrak {M}}^o_1(k^\vee , n, \ell )\).

First we set, as earlier, \(k=q\ell +r\), where \(0\le r <\ell \). If \(\Phi \in {\mathcal {O}}_1(k, \ell )\), then \(\Phi \) has the Jordan form

figure c

. This may be reformulated as follows. Consider the truncated matrix \({\widetilde{A}}_j: {\mathbb {C}}^k\rightarrow {\mathbb {C}}^{(\ell -j+1)n}\), defined as

$$\begin{aligned} {\widetilde{A}}_j=\begin{pmatrix}{{\widetilde{Q}}}\Phi ^{j-1}\\ \vdots \\ {{\widetilde{Q}}}\Phi ^{\ell -1} \end{pmatrix}. \end{aligned}$$
(74)

Since \({{\widetilde{Q}}}\) contains the generators of all Jordan towers, \({\mathrm {Span}}(\text {rows of }{\widetilde{A}}_j)\simeq {\mathrm {Im}}(\Phi ^{j-1})\). From the Jordan form of \(\Phi \), it follows that

$$\begin{aligned}&{\mathrm {rank}}\,{\widetilde{A}}_j={\mathrm {dim}}\;{\mathrm {Im}}(\Phi ^{j-1})=(\ell -j+1)q \quad \text {for}\quad j>r\,, \end{aligned}$$
(75)
$$\begin{aligned}&{\mathrm {rank}}\,{\widetilde{A}}_j={\mathrm {dim}}\;{\mathrm {Im}} (\Phi ^{j-1})=(\ell -j+1)q+(r-j+1) \quad \text {for}\quad j\le r\,. \end{aligned}$$
(76)

For the dimension of the null-space, we therefore get \({\mathrm {dim}}\;{\mathrm {Ker}}\,{\widetilde{A}}_j^T=(\ell -j+1)n-{\mathrm {rank}}\,{\widetilde{A}}_j\). From the fact that the matrices \({\widetilde{A}}_j\) are nested inside each other, it follows that \({\mathrm {Ker}}\,{\widetilde{A}}_\ell ^T \subset {\mathrm {Ker}}\,{\widetilde{A}}_{\ell -1}^T\subset \cdots \subset {\mathrm {Ker}}\,{\widetilde{A}}^T\).

Next, we consider the truncated \({\widetilde{B}}\)-matrices \({\widetilde{B}}_j: {\mathbb {C}}^{k^\vee }\rightarrow {\mathbb {C}}^{n j}\)

$$\begin{aligned} {\widetilde{B}}_j=\begin{pmatrix}{{\widetilde{Q}}}^\vee (\Phi ^\vee )^{j-1} \\ \vdots \\ {{\widetilde{Q}}}^\vee \end{pmatrix}. \end{aligned}$$
(77)

The vectors spanning \({\mathrm {Ker}}\,{\widetilde{A}}_j^T\) enter in some of the columns of \({\widetilde{B}}_{\ell -j+1}\), and accordingly the vectors spanning \({\mathrm {Ker}}\,{\widetilde{A}}_{j-1}^T\) enter in some of the columns of the matrix \({\widetilde{B}}_{\ell -(j-1)+1}\), which is obtained from \({\widetilde{B}}_{\ell -j+1}\) by adding the additional rows \({{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -j+1}\) on top. Moreover, we have the embedding \({\mathrm {Ker}}\,{\widetilde{A}}_j^T\subset {\mathrm {Ker}}\,{\widetilde{A}}_{j-1}^T\). The vectors spanning \({\mathrm {Ker}}\,{\widetilde{A}}_{j-1}^T \setminus {\mathrm {Ker}}\,{\widetilde{A}}_j^T\) remain linearly independent even upon truncation to the “upper block” \({{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -j+1}\). Therefore,

$$\begin{aligned} {\mathrm {rank}}({{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -j+1})\ge {\mathrm {dim}}({\mathrm {Ker}}\,{\widetilde{A}}_{j-1}^T \setminus {\mathrm {Ker}}\,{\widetilde{A}}_j^T)\,. \end{aligned}$$
(78)

Using

$$\begin{aligned} {\mathrm {dim}}({\mathrm {Ker}}\,{\widetilde{A}}_{1}^T)=\ell (n-q)-r,\quad \quad {\mathrm {dim}}({\mathrm {Ker}}\,{\widetilde{A}}_{2}^T)=(\ell -1)(n-q)-{\mathrm {max}}(r-1,0)\,,\nonumber \\ \end{aligned}$$
(79)

we find from (78) for \(j=2\):

$$\begin{aligned} {\mathrm {rank}}({{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -1})\ge n-q+{\mathrm {max}}(r-1,0)-r={\left\{ \begin{array}{ll} n-q\quad \text {if}\quad r=0 \\ n-q-1 \quad \text {if}\quad r>0 \end{array}\right. } \end{aligned}$$
(80)

Since \(k^\vee =n\ell -k=(n-q-1)\ell +\ell -r\), there can be at most \(n-q-1\) or \(n-q\) Jordan towers of height \(\ell \) for \(r>0\) and \(r=0\), respectively. As the number of such maximal towers is measured by \({\mathrm {rank}}({{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -1})\), the inequality (80) shows that this bound is saturated. This implies that \(\Phi ^\vee \) has the Jordan structure

figure d

, i.e., \(\Phi ^\vee \in {\mathcal {O}}_1(k^\vee , \ell )\).

Now we need to show that \(Q^\vee {{\widetilde{Q}}}^\vee \in T_{\Phi ^\vee } {\mathcal {O}}_1(k^\vee , \ell )\). Since the defining equations of \({\mathcal {O}}_1(k^\vee , \ell )\) are \((\Phi ^\vee )^\ell =0\) and \({{\,\mathrm{tr}\,}}((\Phi ^\vee )^i)=0\) for all i, we need to show that \( {d\over d\epsilon }(\Phi +\epsilon Q^\vee {{\widetilde{Q}}}^\vee )^\ell \big |_{\epsilon =0}=0\) and \({{\,\mathrm{tr}\,}}((\Phi ^\vee )^{i-1}Q^\vee {{\widetilde{Q}}}^\vee )=0\) for all i. The first part follows from (72), if one recalls the definitions (67) and (73). To prove the second part, we look more closely at the relation (71). The \(\ell n\times \ell n\)-matrix \({\widetilde{A}}A\) may be viewed as an \(\ell \times \ell \)-matrix with the entries \(({\widetilde{A}}A)^T_{ij}={{\widetilde{Q}}}\Phi ^{\ell -1+j-i} Q\), analogously \(({\widetilde{B}}B)_{ij}={{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -1+j-i} Q^\vee \). Therefore, (71) is equivalent to \({{\widetilde{Q}}}\Phi ^{\ell -1+j-i} Q={{\widetilde{Q}}}^\vee (\Phi ^\vee )^{\ell -1+j-i} Q^\vee \) for all ij. Since \((Q, {{\widetilde{Q}}}, \Phi )\in {\mathfrak {M}}^o_1(k, n, \ell )\), \({{\,\mathrm{tr}\,}}(Q{{\widetilde{Q}}}\Phi ^{i})=0\) is satisfied for all i. It then follows that \({{\,\mathrm{tr}\,}}(Q^\vee {{\widetilde{Q}}}^\vee (\Phi ^\vee )^{i})=0\) for all i, as required. Therefore, \(Q^\vee {{\widetilde{Q}}}^\vee \in T_\Phi {\mathcal {O}}_1(k^\vee , \ell )\) and \((Q^\vee , {{\widetilde{Q}}}^\vee , \Phi ^\vee )\in {\mathfrak {M}}^o_1(k^\vee , n, \ell )\). \(\square \)

Appendix B: The inverse matrix \(S^{-1}\)

To prove the formula (60) for \(S^{-1}\), we observe that \(S^{-1}S\) is a rational function of z. We will now compute the residues of \((S^{-1}S)_{ij'}\) at the potential poles and show that they vanish.

First, changing \(r\rightarrow -r\) in the third factor in the denominator of (59), we simplify the expression for \(S_{jj'}\):

$$\begin{aligned} S_{jj'}= \frac{(-1)^{j'-j}{j'\atopwithdelims ()j} \prod _{r=j}^{j'-1}(r\varphi +\varepsilon )}{ \prod _{r=0}^{k-j-1} (r\varphi +\varepsilon +z) \prod _{r=k-j-j',\;r\ne k-2j}^{k-j}(r\varphi +z) } \end{aligned}$$

Using this, as well as (60), we find

$$\begin{aligned} \sum \limits _{j=i}^{j=j'}\,S^{-1}_{ij} S_{jj'}=\prod \limits _{r=i}^{j'-1}(r \varphi +\epsilon )\,\sum \limits _{j\in {\mathbb {Z}}}\,{j' \atopwithdelims ()j} {j\atopwithdelims ()i}\frac{(-1)^{j'-j}((k-2j)\varphi +z)}{\prod _{r=k-j-j'}^{k-j-i} (r\varphi +z)} \end{aligned}$$
(81)

We have extended the sum over j to the integers, as the binomial coefficients suppress all terms outside of \(i\le j\le j'\). Clearly, the potential poles of the r.h.s. are at \(z=-n \varphi , n\in {\mathbb {Z}}\). Note that there are exactly \(j'-i+1\) factors in the denominator of each term: Setting \(r=k-j-j'+r'\), these factors correspond to \(r'=0\ldots j'-i\). The \({\tilde{r}}\)th factor vanishes at \(z=-n \varphi \) when \(k-j-j'+{\tilde{r}}=n\), i.e., this happens in the jth term in the sum, where \(j=k-j'+{\tilde{r}}-n\). The residue at this point is the sum of \(j'-i+1\) terms:

$$\begin{aligned} \underset{z=-n \varphi }{\mathrm {res}}\left( S^{-1}S\right) _{ij'}=\prod \limits _{r=i}^{j'-1}(r \varphi +\epsilon )\,\sum \limits _{{\tilde{r}}=0}^{j'-i}{j' \atopwithdelims ()j} {j\atopwithdelims ()i}\frac{(-1)^{j'-j}(k-2j-n)\varphi }{\prod _{r'=0, r'\ne {\tilde{r}}}^{j'-i} ((r'-{\tilde{r}})\varphi )}\big |_{j=k-j'+{\tilde{r}}-n}=\\ =\prod \limits _{r=i}^{j'-1}(r \varphi +\epsilon )\,\frac{(-1)^{k-n}}{\varphi ^{j'-i-1}} \sum \limits _{{\tilde{r}}\in {\mathbb {Z}}}{j' \atopwithdelims ()j} {j\atopwithdelims ()i}\frac{(k-2j-n)}{{\tilde{r}}! (j'-i-{\tilde{r}})!}\big |_{j=k-j'+{\tilde{r}}-n}. \end{aligned}$$

We have again extended the summation to the integers, due to the presence of the factorials in the denominator. The factor \((k-2j-n)\big |_{j=k-j'+{\tilde{r}}-n}\) in the numerator is skew-symmetric under the change \({\tilde{r}}\rightarrow 2j'-(k-n)-{\tilde{r}}\). At the same time, one can check that the product of factorials and binomial coefficients in the sum is symmetric under this change. Therefore, \(\underset{z=-n \varphi }{\mathrm {res}}\left( S^{-1}S\right) _{ij'}=0\).

It is obvious from (81) that \((S^{-1}S)_{ij'}\) vanishes for \(z\rightarrow \infty \) if \(j'-i>0\) and approaches 1 if \(j'=i\). We conclude that \((S^{-1}S)_{ij'}=\delta _{ij'}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bykov, D., Zinn-Justin, P. Higher spin \({{\mathfrak {s}}}{{\mathfrak {l}}}_2\)R-matrix from equivariant (co)homology. Lett Math Phys 110, 2435–2470 (2020). https://doi.org/10.1007/s11005-020-01302-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11005-020-01302-z

Keywords

Mathematics Subject Classification

Navigation