Skip to main content
Log in

Polology of Superconformal Blocks

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

Abstract

We systematically classify all possible poles of superconformal blocks as a function of the scaling dimension of intermediate operators, for all superconformal algebras in dimensions three and higher. This is done by working out the recently-proven irreducibility criterion for parabolic Verma modules for classical basic Lie superalgebras. The result applies to correlators for external operators of arbitrary spin, and indicates presence of infinitely many short multiplets of superconformal algebras, most of which are non-unitary. We find a set of poles whose positions are shifted by linear in \(\mathcal {N}\) for \(\mathcal {N}\)-extended supersymmetry. We find an interesting subtlety for 3d \(\mathcal {N}\)-extended superconformal algebra with \(\mathcal {N}\) odd associated with odd non-isotropic roots. We also comment on further applications to superconformal blocks.

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.

Fig. 1

Similar content being viewed by others

Notes

  1. The literature is too vast to be exhaustive here. A sample of early references motivated by the works of [49, 51] include [1, 4,5,6,7, 9, 12, 25, 34, 39, 46, 47, 50], see also [38, 40] for more recent computations of superconformal blocks.

  2. The two-dimensional counterpart of this argument, as applied to the Virasoro algebra, goes back to the classic paper by Zamolodchikov [54].

  3. This fact is the basis for the derivation of the recursion relation [14, 29, 35, 36, 45] for conformal blocks in [45].

  4. We do not work out the decompositions into irreducible components. This is related with the question of computing the Kazhdan–Lusztig polynomials [33] for parabolic Verma modules, which seem to be unknown for the cases at hand, at least in general.

  5. See e.g. [27, 32, 42] for mathematical introduction to Lie superalgebras.

  6. We here list the complex form of the Lie superalgebra.

  7. In this paper, we choose the convention that the argument of \(\mathfrak {sp}\) is even, e.g. \(\mathfrak {sp}(2)\simeq \mathfrak {su}(2)\).

  8. Since the irreducibility criterion of [44] applies in general to any contragredient finite-dimensional Lie superalgebra with an indecomposable Cartan matrix, it is straightforward to repeat the computations in this paper for those other such Lie superalgebras which do not appear in the list of SCAs, for example for \(C(n) = \mathfrak {osp}(2| 2n-2)\, (n\ge 2)\), \(D(2,1;\alpha )\, (\alpha \ne 0, 1)\) and G(3). Note that some of these symmetries do appear when we consider defects in SCFTs, so that we break some of the Poincarè symmetries. For example, a 1 / 2-BPS Wilson loop for 5d \(\mathcal {N}=1\) SCFT discussed in [2] preserve a subgroup \(D(2,1;2)\oplus \mathfrak {su}(2)\) of F(4).

  9. 3d SCFT with \(\mathcal {N}=7\) supersymmetry automatically has \(\mathcal {N}=8\) supersymmetry [3, 13].

  10. In the literature these are more often called \(\mathcal {N}=(1,0)\) and \(\mathcal {N}=(2,0)\) SCAs, to emphasize the chirality. \(\mathcal {N}=(1,1)\) case is excluded in Nahm’s classification.

  11. Annihilation by \(K_{\mu }\) and \(S_{\alpha }\) holds only at the origin.

  12. Logically this might better be denoted \(\Delta _{\mathfrak {n}}^{+}\), however we in this paper do not use \(\Delta _{\mathfrak {n}}^{+}\) without the plus sign and hence we simply dropped the plus sign, to simplify the notation.

  13. There are mathematical papers explicitly working out Jantzen criterion for scalar (spin zero) parabolic Verma modules for semisimple Lie algebras. See e.g. [28] for recent discussion.

  14. As will become clear, the overall normalization factor of this pairing is irrelevant for the considerations of this paper, see e.g. (23).

  15. Weyl group for a Lie superalgebra is defined to be the Weyl group for its even part \(\mathfrak {g}_0\).

  16. This is basically how Propositions 4 and 5 of [44] are derived from the irreducibility criterion (Theorem 3 in [44].

  17. As we will see later, the same type of argument does apply to 3d \(\mathcal {N}\)-extended supersymmetry with \(\mathcal {N}\) odd cases, reducing the analysis to the case of 3d \(\mathcal {N}=1\) SCA.

  18. There is an unfortunate crash of notation, where \(\Delta \) is used both for scaling dimension and the root system (both are standard notations, in physics and mathematics, respectively). We hope that context will make clear which we mean. Note that any \(\Delta \) with indices (such as \(+\) and 0) will a subset of the root system.

  19. For tensor representations this was already worked out in [45].

  20. In this paper we choose \(\lambda \) to be the highest weight, rather than the lowest weight, to match the representation theory conventions in [44]. Some physics literature, e.g. [41], rather use lowest weights. The two conventions are related by a replacement \(\lambda \rightarrow -\lambda \).

  21. Of course, for complete analysis one also needs to take into account possible poles (168), as originating from isotropic roots, which are not shifted simply by \(\mathcal {N}\).

  22. It is known that such a relation does not hold for theories with eight supercharges, see [10]. We thank Nikolay Bobev for discussions related to this point.

  23. Please be aware of the notation change: \(\ell _a\) in [45] is denoted as \(\lambda _a\) in this paper.

  24. Compared with [41] we have an extra minus sign in front of \(\ell _2\) in the third line.

References

  1. Alday, L.F., Bissi, A.: Generalized bootstrap equations for \( \cal{N}=4 \) SCFT. JHEP 1502, 101 (2015). arXiv:1404.5864

  2. Assel, B., Estes, J., Yamazaki, M.: Wilson Loops in 5d N=1 SCFTs and AdS/CFT. Ann. Henri Poincare 15, 589 (2014). arXiv:1212.1202

  3. Bashkirov, D.: A Note on \({\cal{N}}\ge 6\) Superconformal Quantum Field Theories in three dimensions, arXiv:1108.4081

  4. Bashkirov, D.: Bootstrapping the \({\cal{N}}= 1\) SCFT in Three Dimensions, arXiv:1310.8255

  5. Beem, C., Lemos, M., Liendo, P., Rastelli, L., van Rees, B.C.: The \( \cal{N}=2 \) superconformal bootstrap. JHEP 1603, 183 (2016). arXiv:1412.7541

  6. Beem, C., Rastelli, L., van Rees, B.C.: The \(\cal{N}=4\) superconformal bootstrap. Phys. Rev. Lett. 111, 071601 (2013). arXiv:1304.1803

  7. Berkooz, M., Yacoby, R., Zait, A.: Bounds on \(\cal{N}= 1\) superconformal theories with global symmetries. JHEP 1408, 008 (2014). arXiv:1402.6068, [Erratum: JHEP01,132(2015)]

  8. Bhattacharya, J., Bhattacharyya, S., Minwalla, S., Raju, S.: Indices for superconformal field theories in 3,5 and 6 dimensions. JHEP 0802, 064 (2008). arXiv:0801.1435

  9. Bobev, N., El-Showk, S., Mazac, D., Paulos, M.F.: Bootstrapping SCFTs with four supercharges. JHEP 1508, 142 (2015). arXiv:1503.02081

  10. Bobev, N., Lauria, E., Mazac, D.: Superconformal blocks for SCFTs with eight supercharges. JHEP 1707, 061 (2017). arXiv:1705.08594

  11. Buican, M., Hayling, J., Papageorgakis, C.: Aspects of Superconformal Multiplets in \(\text{D}>4\), arXiv:1606.00810

  12. Chester, S.M., Lee, J., Pufu, S.S., Yacoby, R.: The \( \cal{N}=8 \) superconformal bootstrap in three dimensions. JHEP 1409, 143 (2014). arXiv:1406.4814

  13. Cordova, C., Dumitrescu, T.T., Intriligator, K.: Multiplets of Superconformal Symmetry in Diverse Dimensions, arXiv:1612.00809

  14. Costa, M.S., Hansen, T., Penedones, J., Trevisani, E.: Radial expansion for spinning conformal blocks. JHEP 1607, 057 (2016). arXiv:1603.05552

  15. Costa, M.S., Penedones, J., Poland, D., Rychkov, S.: Spinning conformal correlators. JHEP 1111, 071 (2011). arXiv:1107.3554

  16. Dobrev, V.K.: Positive energy unitary irreducible representations of D = 6 conformal supersymmetry. J. Phys. A 35, 7079 (2002). arXiv:hep-th/0201076

  17. Dobrev, V., Petkova, V.: All positive energy unitary irreducible representations of extended conformal supersymmetry. Phys. Lett. B 162, 127 (1985)

    Google Scholar 

  18. Dobrev, V., Petkova, V.: On the group theoretical approach to extended conformal supersymmetry: classification of multiplets. Lett. Math. Phys. 9, 287 (1985)

    Google Scholar 

  19. Dobrev, V., Petkova, V.: Group theoretical approach to extended conformal supersymmetry: function space realizations and invariant differential operators. Fortschr. Phys. 35, 537 (1987)

    Google Scholar 

  20. Dolan, F.A., Osborn, H.: Conformal four point functions and the operator product expansion. Nucl. Phys. B 599, 459 (2001). arXiv:hep-th/0011040

  21. Dolan, F., Osborn, H.: Conformal partial waves and the operator product expansion. Nucl. Phys. B 678, 491 (2004). arXiv:hep-th/0309180

  22. Dolan, F., Osborn, H.: Superconformal symmetry, correlation functions and the operator product expansion. Nucl. Phys. B 629, 3 (2002). arXiv:hep-th/0112251

  23. Enright, T., Howe, R., Wallach, N.: A classification of unitaryhighest weight modules. In: Representation Theory of Reductivegroups (Park City, Utah, 1982), pp. 97–143, BirkhäuserBoston, Boston, MA (1983)

  24. Ferrara, S., Grillo, A., Gatto, R.: Tensor representations of conformal algebra and conformally covariant operator product expansion. Ann. Phys. 76, 161 (1973)

    Google Scholar 

  25. Fitzpatrick, A.L., Kaplan, J., Khandker, Z.U., Li, D., Poland, D. et al.: Covariant Approaches to Superconformal Blocks, arXiv:1402.1167

  26. Flato, M., Fronsdal, C.: Representations of conformal supersymmetry. Lett. Math. Phys. 8, 159 (1984)

    Google Scholar 

  27. Frappat, L., Sorba, P., Sciarrino, A.: Dictionary on Lie superalgebras, arXiv:hep-th/9607161

  28. He, H.: On the reducibility of scalar generalized Verma modules of abelian type. Algebras Represent. Theory 19, 147 (2016). https://doi.org/10.1007/s10468-015-9567-2

    Google Scholar 

  29. Iliesiu, L., Kos, F., Poland, D., Pufu, S.S., Simmons-Duffin, D., Yacoby, R.: Fermion-Scalar Conformal Blocks, arXiv:1511.01497

  30. Jantzen, J.C.: Kontravariante Formen auf induzierten Darstellungen halbeinfacher Lie-Algebren. Math. Ann. 226, 53 (1977)

    Google Scholar 

  31. Kac, V.: Representations of classical Lie superalgebras. In: Proceeding Conference of Differential geometrical Methods in Mathematical Physics, II (University Bonn, Bonn, 1977), pp. 597–626, Springer, Berlin (1978)

  32. Kac, V.G.: Lie superalgebras. Adv. Math. 26, 8 (1977)

    Google Scholar 

  33. Kazhdan, D., Lusztig, G.: Representations of Coxeter groups and Hecke algebras. Invent. Math. 53, 165 (1979)

    Google Scholar 

  34. Khandker, Z.U., Li, D., Poland, D., Simmons-Duffin, D.: \( \cal{N} = 1\) superconformal blocks for general scalar operators. JHEP 1408, 049 (2014). arXiv:1404.5300

  35. Kos, F., Poland, D., Simmons-Duffin, D.: Bootstrapping mixed correlators in the 3D Ising model. JHEP 1411, 109 (2014). arXiv:1406.4858

  36. Kos, F., Poland, D., Simmons-Duffin, D.: Bootstrapping the O(N) Vector Models, arXiv:1307.6856

  37. Kravchuk, P., Simmons-Duffin, D.: Counting conformal correlators. JHEP 1802, 096 (2018). arXiv:1612.08987

  38. Lemos, M., Liendo, P., Meneghelli, C., Mitev, V.: Bootstrapping \({\cal{N}} =3\) superconformal theories. JHEP 1704, 032 (2017). arXiv:1612.01536

  39. Li, D., Stergiou, A.: Two-point functions of conformal primary operators in \(\cal{N}=1\) superconformal theories. JHEP 1410, 37 (2014). arXiv:1407.6354

  40. Liendo, P., Meneghelli, C.: Bootstrap equations for \({\cal{N}} = 4\) SYM with defects. JHEP 1701, 122 (2017). arXiv:1608.05126

  41. Minwalla, S.: Restrictions imposed by superconformal invariance on quantum field theories. Adv. Theor. Math. Phys. 2, 781 (1998). arXiv:hep-th/9712074

    Google Scholar 

  42. Musson, I.M.: Lie Superalgebras and Enveloping Algebras. American Mathematical Society, Providence (2012)

    Google Scholar 

  43. Nahm, W.: Supersymmetries and their representations. Nucl. Phys. B 135, 149 (1978)

    Google Scholar 

  44. Oshima, Y., Yamazaki, M.: Determinant Formula for Parabolic Verma Modules of Lie Superalgebras, arXiv:1603.06705

  45. Penedones, J., Trevisani, E., Yamazaki, M.: Recursion Relations for Conformal Blocks, arXiv:1509.00428

  46. Poland, D., Simmons-Duffin, D., Vichi, A.: Carving out the space of 4D CFTs. JHEP 1205, 110 (2012). arXiv:1109.5176

  47. Poland, D., Simmons-Duffin, D.: Bounds on 4D conformal and superconformal field theories. JHEP 1105, 017 (2011). arXiv:1009.2087

  48. Polyakov, A.: Nonhamiltonian approach to conformal quantum field theory. Zh. Eksp. Teor. Fiz. 66, 23 (1974)

    Google Scholar 

  49. Rattazzi, R., Rychkov, V.S., Tonni, E., Vichi, A.: Bounding scalar operator dimensions in 4D CFT. JHEP 0812, 031 (2008). arXiv:0807.0004

  50. Rattazzi, R., Rychkov, S., Vichi, A.: Central charge bounds in 4D conformal field theory. Phys. Rev. D 83, 046011 (2011). arXiv:1009.2725

  51. Rychkov, V.S., Vichi, A.: Universal constraints on conformal operator dimensions. Phys. Rev. D 80, 045006 (2009). arXiv:0905.2211

  52. Schomerus, V., Sobko, E., Isachenkov, M.: Harmony of Spinning Conformal Blocks, arXiv:1612.02479

  53. Yamazaki, M.: Comments on determinant formulas for general CFTs. JHEP 1610, 035 (2016). arXiv:1601.04072

  54. Zamolodchikov, A.: Conformal symmetry in two-dimensions: an explicit recurrence formula for the conformal partial wave amplitude. Commun. Math. Phys. 96, 419 (1984)

    Google Scholar 

Download references

Acknowledgements

We would like to thank Yoshiki Oshima for related discussion. The bulk of this Project was completed in fall 2016–spring 2017, however the publication has been delayed for no good reason. We thank the organizers of the symposium “Bootstrap Approach to Conformal Field Theories and Applications”, OIST, March 2018, for providing motivation to finish up this work. The authors are supported in part by WPI program (MEXT, Japan). MY is also supported by by JSPS Program for Advancing Strategic International Networks to Accelerate the Circulation of Talented Researchers, by JSPS Grant No. 15K17634 and No. 17KK0087, and by JSPS-NRF Joint Research Project. He would like to thank Harvard university for hospitality.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kallol Sen.

Additional information

Communicated by C. Schweigert

Publisher's Note

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

Appendices

ABD Roots and Weights

In this Appendix we summarize some standard facts about the root system and the highest weights for the ABD Lie algebras. We can spare the C case (associated with Lie algebra \(\mathfrak {sp}(2N)\)) for the applications to SCFTs in this paper, since we can use the isomorphism of Lie algebras: \(\mathfrak {sp}(2)=\mathfrak {su}(2)\), \(\mathfrak {sp}(4)=\mathfrak {so}(5)\).

1.1 \(A_{N-1}\)

The \(A_{N-1}\) root system is given by

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm (\delta _a - \delta _b) \bigg \} \ , \end{aligned}$$
(88)

where we introduced a orthonormal basis \(\delta _a\)

$$\begin{aligned} \begin{aligned} (\delta _a, \delta _b)=\delta _{a, b} \ , \quad a, b=1, \ldots , N. \end{aligned} \end{aligned}$$
(89)

Under an ordering \(\delta _1> \ldots >\delta _{\mathcal {N}}\), the positive simple roots are given by \(\alpha _a=\delta _a-\delta _{a+1}\), with \(a=1, \ldots , N-1\).

The fundamental weight is given by

$$\begin{aligned} w_a:=\sum _{b=1}^{a} \delta _b -\frac{a}{N} \sum _{c=1}^{N} \delta _c , \quad \alpha _a:=\delta _a-\delta _{a+1}, \end{aligned}$$
(90)

which are determined by the conditions

$$\begin{aligned}&2\frac{ (w_a, \alpha _b)}{(\alpha _b, \alpha _b)}=\delta _{ab}, \quad a,b=1, \ldots , N, \quad \alpha _a:=\delta _a-\delta _{a+1}, \end{aligned}$$
(91)

as well as the condition that \(w_a\) is in the \(\mathbb {R}\)-span of \(\alpha _b\)’s.

The highest weight for a finite-dimensional representation of the \(\mathfrak {su}(N)\) algebra is given by a dominant integral weight, which is given by a \(\mathbb {Z}_{\ge 0}\)-span of fundamental weights:

$$\begin{aligned} \lambda =\sum _{a=1}^{N-1} \ell _a w_a =\sum _{a=1}^{N-1} \ell _a \left( \sum _{b=1}^{a} \delta _b -\frac{a}{N} \sum _{c=1}^{N} \delta _c \right) , \end{aligned}$$
(92)

where the coefficients \([\ell _1, \ldots , \ell _{\mathcal {N}}]\) are called Dynkin labels. The highest weight (92) can also be written as

$$\begin{aligned} \lambda =\sum _{a=1}^{N-1} \ell _a w_a =\sum _{a=1}^{N} \left( \lambda _a -\frac{|\lambda |}{N} \right) \delta _a , \end{aligned}$$
(93)

where we introduced \(\lambda _a\) and \(|\lambda |\) by

$$\begin{aligned}&\lambda _a=\ell _a+ \ell _{a+1} +\cdots +\lambda _{N-1}, \quad \text {or} \quad \ell _a=\lambda _a-\lambda _{a+1}, \end{aligned}$$
(94)
$$\begin{aligned}&|\lambda |=\sum _a \lambda _a . \end{aligned}$$
(95)

Since this \(\lambda \) satisfies

$$\begin{aligned} \lambda _1 \ge \lambda _2 \ge \cdots \ge \lambda _{N-1} \ge 0, \end{aligned}$$
(96)

\(\{ \lambda _a \}\) define a partition (Young diagram), where the number of boxes at height a is given by \(\lambda _a\); in this language, \(|\lambda |\) is the total number of boxes.

1.2 \(D_{N}\)

The \(D_N\) root system (corresponding to the Lie algebra \(\mathfrak {so}(2N)\)) is given by

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \delta _a \pm \delta _b \bigg \} \ , \quad a,b=1, \ldots , N, \end{aligned}$$
(97)

and the inner product is given by (89). In dictionary ordering \(\delta _1> \delta _2>\ldots >\delta _N\), the positive simple roots are given by

$$\begin{aligned} \alpha _a=\delta _a-\delta _{a+1} \quad a=1, \ldots , N-1, \quad \alpha _N=\delta _{N-1}+\delta _N. \end{aligned}$$
(98)

The fundamental weights satisfying (91) are obtained as

$$\begin{aligned} \begin{aligned} w_a&=\delta _1 + \cdots + \delta _a, \quad a=1, \ldots , N-2, \\ w_{N-1}&=\frac{\delta _1 + \cdots + \delta _{N-1}-\delta _N}{2},\\ w_{N}&=\frac{\delta _1 + \cdots + \delta _{N-1}+\delta _N}{2}.\\ \end{aligned} \end{aligned}$$
(99)

The highest weight for a finite-dimensional representation is given as

$$\begin{aligned} \begin{aligned} \lambda =\sum _{a=1}^N \ell _a w_a =\sum _{a=1}^N \lambda _a \delta _a \end{aligned} \end{aligned}$$
(100)

where \(\ell _a\in \mathbb {Z}_{\ge 0}\) are integers called Dynkin labels, and we defined half-integers \(\lambda _a\in \mathbb {Z}_{\ge 0}/2\) by

$$\begin{aligned} \begin{aligned}&\lambda _a = \ell _a+ \cdots \ell _{N-2}+\frac{\ell _{N-1}+\ell _N}{2} , \quad a=1, \ldots , N-2, \\&\lambda _{N-1}=\frac{\ell _{N-1}+\ell _N}{2}, \quad \lambda _{N}=\frac{-\ell _{N-1}+\ell _N}{2}. \end{aligned} \end{aligned}$$
(101)

By definition this satisfies \(\lambda _1 \ge \lambda _2 \ge \cdots \ge \lambda _{N-1}\ge |\lambda _N| \ge 0\). Note that \(\lambda _N\) can be negative. The \(\lambda _a\)’s are half-integers for a general representation. If we consider the tensor representation, however, all the \(\lambda _a\)’s takes values in integers; a tensor representation is labeled by a partition and a sign.

1.3 \(B_{N}\)

The \(B_N\) root system (corresponding to the Lie algebra \(\mathfrak {so}(2N+1)\)) is similar to the \(D_N\) root system, but with some extra roots added:

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \delta _a \pm \delta _b, \pm \delta _a \bigg \} , \quad a,b=1, \ldots , N \end{aligned}$$
(102)

and the inner product is given by (89). In dictionary ordering \(\delta _1> \delta _2>\ldots >\delta _N\), the positive simple roots are given by

$$\begin{aligned} \alpha _a=\delta _a-\delta _{a+1} \quad a=1, \ldots , N-1, \quad \alpha _N=\delta _N. \end{aligned}$$
(103)

The fundamental weights satisfying (91) are obtained as

$$\begin{aligned} \begin{aligned} w_a&=\delta _1 + \cdots + \delta _a, \quad a=1, \ldots , N-1, \\ w_{N}&=\frac{\delta _1 + \cdots + \delta _{N-1}+\delta _N}{2}.\\ \end{aligned} \end{aligned}$$
(104)

The highest weight for a finite-dimensional representation is given as

$$\begin{aligned} \begin{aligned} \lambda =\sum _{a=1}^N \ell _a w_a =\sum _{a=1}^N \lambda _a \delta _a \end{aligned} \end{aligned}$$
(105)

where \(\ell _a\in \mathbb {Z}_{\ge 0}\) are Dynkin labels and we defined half-integers \(\lambda _a\in \mathbb {Z}_{\ge 0}/2\) by

$$\begin{aligned} \begin{aligned}&\lambda _a = \ell _a+ \cdots \ell _{N-2}+\frac{\ell _N}{2}, \quad a=1, \ldots , N-1, \\&\lambda _{N}=\frac{\ell _N}{2} . \end{aligned} \end{aligned}$$
(106)

By definition this satisfies \(\lambda _1 \ge \lambda _2 \ge \cdots \ge \lambda _N\ge 0\). In tensor representations all the \(\lambda _a\)’s are integers; a tensor representation is labeled by a partition.

\(\mathcal {N}=0\) Results

In this appendix we present the analysis for the non-supersymmetric cases (i.e. \(\mathcal {N}=0\)) in spacetime dimensions 4, 5 and 6. Note that in these cases the conformal algebra contains no odd elements, and hence \(\Delta =\Delta _{\bar{0}}, \Delta _1=\Delta _{1}^{+}=\varnothing \) and \(\rho =\rho _0\).

Note that such an analysis (for a general spacetime dimension) was already given in [45], except there the spinor representations are not considered there. In this appendix we therefore present the full analysis including the cases of spinor representations. We present this analysis in the same notations as in the rest of this paper, to make the comparison easier.Footnote 23 We will find that while intermediate steps of the analysis changes slightly from [45], the final result is in the end the same as in [45], even for spinor representations.

1.1 4d \(\mathcal {N}=0\)

The conformal algebra in this case is \(\mathfrak {g}=\mathfrak {su}(4)=A(3)\). Let us introduce a basis

$$\begin{aligned} \begin{aligned} (\varepsilon _i, \varepsilon _j)=\delta _{i,j} \ , \quad \end{aligned} \end{aligned}$$
(107)

with \(i,j=1, \ldots , 4\). Then we have

$$\begin{aligned}&\Delta =\bigg \{ \pm ( \varepsilon _i - \varepsilon _j ) \bigg \} \ , \end{aligned}$$
(108)
$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm ( \varepsilon _1 - \varepsilon _2 ), \quad \pm ( \varepsilon _3 - \varepsilon _4) \bigg \}. \end{aligned}$$
(109)

Let us choose an ordering \(\varepsilon _1> \varepsilon _2>-\varepsilon _3>-\varepsilon _4\). We then have

$$\begin{aligned} \Delta ^+&=\bigg \{ \varepsilon _{1}-\varepsilon _{2} , \quad \varepsilon _{4}-\varepsilon _{3} \bigg \}, \end{aligned}$$
(110)
$$\begin{aligned} \Delta _{\mathfrak {n}}&=\bigg \{ \varepsilon _1 - \varepsilon _3 , \quad \varepsilon _1 - \varepsilon _4 , \quad \varepsilon _2 - \varepsilon _3 , \quad \varepsilon _2 - \varepsilon _4 \bigg \}. \end{aligned}$$
(111)

The highest weight vector and the Weyl vector is

$$\begin{aligned} \begin{aligned} \lambda&=-\frac{\Delta }{2}(\varepsilon _1+\varepsilon _2-\varepsilon _3-\varepsilon _4)+\frac{\ell _1}{2}(\varepsilon _1-\varepsilon _2)-\frac{\ell _2}{2}(\varepsilon _3-\varepsilon _4) \\ \rho&=\frac{1}{2} \left( 3 \varepsilon _1 +\varepsilon _2 -3\varepsilon _3 -\varepsilon _4 \right) ,\\ \lambda +\rho&=\frac{-\Delta +3}{2} (\varepsilon _1 -\varepsilon _3 ) +\frac{-\Delta +1}{2}(\varepsilon _2 -\varepsilon _4) +\frac{\ell _1}{2}(\varepsilon _1-\varepsilon _2)-\frac{\ell _2}{2}(\varepsilon _3-\varepsilon _4) \,. \end{aligned} \end{aligned}$$
(112)

Here \(\ell _1\) and \(\ell _2\) are the two angular spins, and take integer values.

We find

$$\begin{aligned} \begin{aligned}&n_{\varepsilon _1-\varepsilon _3}=3-\Delta +\frac{\ell _1+\ell _2}{2}, \ \ n_{\varepsilon _1-\varepsilon _4}=2-\Delta +\frac{\ell _1-\ell _2}{2}, \\&n_{\varepsilon _2-\varepsilon _3}=2-\Delta +\frac{\ell _2-\ell _1}{2}, \ \ n_{\varepsilon _2-\varepsilon _4}=1-\Delta -\frac{\ell _1+\ell _2}{2}\, \end{aligned} \end{aligned}$$
(113)

and \(\Psi _\lambda ^+\ne \varnothing \) in the following cases:

$$\begin{aligned} \Delta -C_{\ell _1+\ell _2} = {\left\{ \begin{array}{ll} 3+\frac{\ell _1+\ell _2}{2}-k, \ k=\{1,\ldots \infty \},\ \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3\}\\ 2-\frac{\ell _1-\ell _2}{2}-k,\ k=\{1,2,\ldots ,1+\ell _2\},\ \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3,\varepsilon _1-\varepsilon _4,\varepsilon _2-\varepsilon _3\},\\ 2+\frac{\ell _1-\ell _2}{2}-k, \ k=\{1,2,\ldots \ell _1-\ell _2\}, \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3,\varepsilon _1-\varepsilon _4\},\\ 1-\frac{\ell _1+\ell _2}{2}-k, \ k=\{1,2,\ldots \}, \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3,\varepsilon _1-\varepsilon _4,\varepsilon _2-\varepsilon _3,\varepsilon _2-\varepsilon _4\},\\ \end{array}\right. } \end{aligned}$$
(114)

for \(\ell _1\ge \ell _2\), and

$$\begin{aligned} \Delta -C_{\ell _1+\ell _2} ={\left\{ \begin{array}{ll}1-\frac{\ell _1+\ell _2}{2}-k, \ k=\{1,2,\ldots \}, \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3,\varepsilon _1-\varepsilon _4,\varepsilon _2-\varepsilon _3,\varepsilon _2-\varepsilon _4\},\\ 2+\frac{\ell _2-\ell _1}{2}-k, \ k=\{1,2,\ldots ,\ell _2-\ell _1\}, \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3,\varepsilon _2-\varepsilon _3\},\\ 2-\frac{\ell _2-\ell _1}{2}-k,\ k=\{1,2,\ldots ,1+\ell _1\} ,\ \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3,\varepsilon _1-\varepsilon _4,\varepsilon _2-\varepsilon _3\},\\ 3+\frac{\ell _1+\ell _2}{2}-k, \ k=\{1,\ldots \infty \},\ \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3\} \end{array}\right. } \end{aligned}$$
(115)

for \(\ell _1< \ell _2\). Here we defined

$$\begin{aligned} C_{\ell _1+\ell _2}= {\left\{ \begin{array}{ll} 0 &{} (\ell _1+\ell _2\text { even})\\ \frac{1}{2} &{} (\ell _1+\ell _2 \text { odd}). \end{array}\right. } \end{aligned}$$
(116)

The final step is to check whether there exists \(\beta \in \Delta _\mathfrak {n}\) such that \((\lambda +\rho ,\beta )=0\). This happens for

$$\begin{aligned} \begin{aligned}&\Delta =3+\frac{\ell _1+\ell _2}{2}, \ n_{\varepsilon _{1}-\varepsilon _{3}}=0, \ \Psi _\lambda ^+=\varnothing \,.\\&\Delta =2+\frac{\ell _1-\ell _2}{2}, \ n_{\varepsilon _{1}-\varepsilon _{4}}=0, \ \Psi _\lambda ^+={\left\{ \begin{array}{ll}\{\varepsilon _1-\varepsilon _3,\varepsilon _2-\varepsilon _3\},\ell _2>\ell _1\\ \{\varepsilon _1-\varepsilon _3\},\ell _2<\ell _1\end{array}\right. }\,.\\&\Delta =2+\frac{\ell _2-\ell _1}{2}, \ n_{\varepsilon _{2}-\varepsilon _{3}}=0,\ \Psi _\lambda ^+={\left\{ \begin{array}{ll}\{\varepsilon _1-\varepsilon _3,\varepsilon _1-\varepsilon _4\},\ell _1>\ell _2\\ \{\varepsilon _1-\varepsilon _3\},\ell _1<\ell _2\end{array}\right. }\,.\\&\Delta =1-\frac{\ell _1+\ell _2}{2},\ n_{\varepsilon _{2}-\varepsilon _{4}}=0, \ \Psi _\lambda ^+=\{\varepsilon _1-\varepsilon _3,\varepsilon _1-\varepsilon _4,\varepsilon _2-\varepsilon _3\}\,. \end{aligned} \end{aligned}$$
(117)

The first line is trivial since \(\Psi _\lambda ^+=\varnothing \). For the remaining cases, we need to apply the final condition (25). In most cases the condition (25) does not hold. The first exception is the obvious case of \(\Delta =3+\frac{\ell _1+\ell _2}{2}\), when \(\Psi _\lambda ^+\) is empy. The other exception, assuming \(\ell _1>\ell _2\), happens for \(\Delta =2+\frac{\ell _1-\ell _2}{2}\). In this case \(\lambda +\rho \) is in the hyperplane orthogonal to \(\varepsilon _1 -\varepsilon _4\), and hence \(s_{\varepsilon _1-\varepsilon _3}(\lambda +\rho )\) is orthogonal to \(\varepsilon _3 -\varepsilon _4\). This means that \(s_{\varepsilon _1-\varepsilon _3}(\lambda +\rho )\) is fixed by an element of the Weyl group \(W_{\mathfrak {l}}\) exchanging \(\varepsilon _3\) and \(\varepsilon _4\). The case of \(\ell _1<\ell _2\) is similar, and we find the reducible points in four dimensions comprises of,

$$\begin{aligned} \Delta =\bigg \{3+\frac{\ell _1+\ell _2}{2}-\mathbb {Z}_{>0}\bigg \}\bigg \backslash \bigg \{2+\frac{|\ell _1-\ell _2|}{2}\bigg \}\, . \end{aligned}$$
(118)

In terms of partitions \(\lambda _1, \lambda _2\in \mathbb {Z}\) with \(\lambda _1\ge |\lambda _2|\ge 0\) (see Appendix A), this becomes

$$\begin{aligned} \Delta =\bigg \{\frac{ \lambda _1}{2}+3-\mathbb {Z}_{>0}\bigg \} \setminus \bigg \{ \frac{|\lambda _2|}{2}+2 \bigg \}. \end{aligned}$$
(119)

1.2 5d \(\mathcal {N}=0\)

For this case, we use the root system for \(\mathfrak {so}_7=B(3)\):

$$\begin{aligned}&\Delta =\bigg \{ \pm \beta _D \pm \beta _{J_1}, \quad \pm \beta _D \pm \beta _{J_2}, \quad \pm \beta _{J_1} \pm \beta _{J_2}, \quad \pm \beta _D , \quad \pm \beta _{J_1}, \quad \pm \beta _{J_2} \bigg \} \ , \end{aligned}$$
(120)

with the inner product \((\beta _i, \beta _j)=\delta _{i,j}\) Under an ordering \(\beta _D> \beta _{J_1}>\beta _{J_2}\),

$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm \beta _{J_1} \pm \beta _{J_2}, \quad \pm \beta _{J_1}, \quad \pm \beta _{J_2} \bigg \}, \end{aligned}$$
(121)
$$\begin{aligned}&\Delta _{\mathfrak {n}}=\bigg \{ \beta _D \pm \beta _{J_1}, \quad \beta _D \pm \beta _{J_2}, \quad \beta _D \bigg \}. \end{aligned}$$
(122)

We have

$$\begin{aligned}&\lambda =-\Delta \beta _D + \left( \ell _1 +\frac{\ell _2}{2}\right) \beta _{J_1}+\frac{\ell _2}{2} \beta _{J_2} + k\delta , \end{aligned}$$
(123)
$$\begin{aligned}&\rho =\frac{5}{2}\beta _D+\frac{3}{2}\beta _{J_1}+\frac{1}{2}\beta _{J_2}+\frac{1}{2}\delta , \end{aligned}$$
(124)
$$\begin{aligned}&\lambda +\rho =\left( -\Delta +\frac{5}{2}\right) \beta _D+\left( \ell _1+\frac{\ell _2}{2}+\frac{3}{2}\right) \beta _{J_1}+\left( \frac{\ell _2}{2}+\frac{1}{2}\right) \beta _{J_2}+ \left( k+\frac{1}{2} \right) \delta . \end{aligned}$$
(125)

In Step 1 we have

$$\begin{aligned} \begin{aligned}&n_{\beta _D}=5-2\Delta ,\\&n_{\beta _D+\beta _{J_1}}=4-\Delta +\ell _1+\frac{\ell _2}{2}, \\&n_{\beta _D-\beta _{J_1}}=1-\Delta -\ell _1-\frac{\ell _2}{2},\\&n_{\beta _D+\beta _{J_2}}=3-\Delta +\frac{\ell _2}{2},\\&n_{\beta _D-\beta _{J_2}}=2-\Delta -\frac{\ell _2}{2}\,. \end{aligned} \end{aligned}$$
(126)

This means that we have \(\Psi _\lambda ^+\ne \varnothing \) when

$$\begin{aligned} \ell _2 \text { even}, \ \ \Delta ={\left\{ \begin{array}{ll} 1-\ell _1-\frac{\ell _2}{2}-k, \ \Psi _\lambda ^+=\{\beta _D,\beta _D\pm \beta _{J_1},\beta _D\pm \beta _{J_2}\}\\ 2-\frac{\ell _2}{2}-k, \ k=\{1,\ldots , 1+\ell _1\}\ \Psi _\lambda ^+=\{\beta _D,\beta _D+\beta _{J_1},\beta _D\pm \beta _{J_2}\}\\ \frac{5}{2}-k, \ k=\{\frac{3}{2},\ldots ,\frac{\ell _2+1}{2}\}\ \Psi _\lambda ^+=\{\beta _D,\beta _D+\beta _{J_1},\beta _D+\beta _{J_2}\}\\ 3+\frac{\ell _2}{2}-k, \ k=\{1,\ldots ,\frac{1+\ell _2}{2}\}\ \Psi _\lambda ^+=\{\beta _D+\beta _{J_1},\beta _D+\beta _{J_2}\}\\ 4+\ell _1+\frac{\ell _2}{2}-k, \ k=\{1,\ldots ,1+\ell _1\}\ \Psi _\lambda ^+=\{\beta _D+\beta _{J_1}\} \\ \frac{5}{2}-k, \ k=\{1,\ldots ,\frac{\ell _2}{2}\}\ \Psi _\lambda ^+=\{\beta _D \}\\ \end{array}\right. } \end{aligned}$$
(127)

and,

$$\begin{aligned} \ell _2\text { odd}, \ \ \Delta ={\left\{ \begin{array}{ll} 1-\ell _1-\frac{\ell _2}{2}-k, \ \Psi _\lambda ^+=\{\beta _D,\beta _D\pm \beta _{J_1},\beta _D\pm \beta _{J_2}\}\\ 2-\frac{\ell _2}{2}-k, \ k=\{\frac{1}{2},\ldots , \ell _1+\frac{1}{2}\}\ \Psi _\lambda ^+=\{\beta _D,\beta _D+\beta _{J_1},\beta _D\pm \beta _{J_2}\}\\ \frac{5}{2}-k, \ k=\{\frac{1}{2},\ldots ,\frac{\ell _2}{2}\}\ \Psi _\lambda ^+=\{\beta _D,\beta _D+\beta _{J_1},\beta _D+\beta _{J_2}\}\\ 3+\frac{\ell _2}{2}-k, \ k=\{\frac{1}{2},\ldots ,\frac{\ell _2}{2}\}\ \Psi _\lambda ^+=\{\beta _D+\beta _{J_1},\beta _D+\beta _{J_2}\}\\ 4+\ell _1+\frac{\ell _2}{2}-k, \ k=\{\frac{1}{2},\ldots ,\ell _1+\frac{1}{2}\}\ \Psi _\lambda ^+=\{\beta _D+\beta _{J_1}\}. \end{array}\right. } \end{aligned}$$
(128)

The next step is to identify the walls for which there exists \(\beta \in \Delta _\mathfrak {n}\) such that \((\lambda +\rho ,\beta )=0\),

$$\begin{aligned}&n_{\beta _D}=0, \ \Delta =\frac{5}{2}, \ \Psi _\lambda ^+=\{\beta _D+\beta _{J_1},\beta _D+\beta _{J_2}\}\nonumber \\&n_{\beta _D+\beta _{J_1}}=0, \ \Delta =4+\ell _1+\frac{\ell _2}{2}, \ \Psi _\lambda ^+=\varnothing \nonumber \\&n_{\beta _D-\beta _{J_1}}=0, \ \Delta =1-\ell _1-\frac{\ell _2}{2}, \ \Psi _\lambda ^+=\{\beta _D,\beta _D+\beta _{J_1},\beta _D\pm \beta _{J_2}\}\\&n_{\beta _D+\beta _{J_2}}=0,\ \Delta =3+\frac{\ell _2}{2}, \ \Psi _\lambda ^+=\{\beta _D+\beta _{J_1}\}\nonumber \\&n_{\beta _D-\beta _{J_2}}=0, \ \Delta =2-\frac{\ell _2}{2}, \ \Psi _\lambda ^+=\{\beta _D,\beta _D+\beta _{J_1},\beta _D+\beta _{J_2}\}\,.\nonumber \end{aligned}$$
(129)

Finally we need to check the condition (25) for each of the above cases. In most cases the condition is not satisfied and the representation is reducible. The exception in the case \(n_{\beta _D}=0\), when \(\Psi _\lambda ^+\) has two elements \(\beta _D+\beta _{J_1},\beta _D+\beta _{J_2}\), and \(s_{\beta _D+\beta _{J_1}}(\lambda +\rho )\) and \(s_{\beta _D+\beta _{J_2}}(\lambda +\rho )\) are each in the hyperplane orthogonal to \(\beta _{J_1}\) and \(\beta _{J_2}\). Since these two vectors can be rotated by an element of \(W_{\mathfrak {l}}\) (rotation symmetry) we have an irreducible representation. Hence we get the following set of reducible points,

$$\begin{aligned} \Delta =\bigg (\bigg \{4+\ell _1+\frac{\ell _2}{2}-\mathbb {Z}_{>0}\bigg \}\bigg \backslash \bigg \{3+\frac{\ell _2}{2},2-\frac{\ell _2}{2},1-\ell _1-\frac{\ell _2}{2}\bigg \}\bigg )\bigcup \bigg (\frac{5}{2}-\mathbb {Z}_{>0}\bigg )\,. \end{aligned}$$
(130)

In terms of \(\lambda _1 \ge \lambda _2 \ge 0\) this can be written as

$$\begin{aligned} \begin{aligned}&\Delta =\left( \bigg \{ \lambda _1+4-\mathbb {Z}_{>0}\bigg \} \setminus \bigg \{\lambda _2+3, -\lambda _2+2 , -\lambda _1+1\bigg \}\right) \bigcup \left( \frac{5}{2}-\mathbb {Z}_{>0}\right) . \end{aligned} \end{aligned}$$
(131)

1.3 6d \(\mathcal {N}=0\)

The root system for \(\mathfrak {g}=\mathfrak {so}(8)=D(4)\) is

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \alpha _i \pm \alpha _j, \bigg \} \quad (i,j=D,1,2,3)\; . \end{aligned}$$
(132)

Under an ordering \(\alpha _D>\alpha _1>\alpha _2>\alpha _3\),

$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm \alpha _1 \pm \alpha _2, \quad \pm \alpha _1 \pm \alpha _3, \quad \pm \alpha _2 \pm \alpha _3 \bigg \}, \end{aligned}$$
(133)
$$\begin{aligned}&\Delta _{\mathfrak {n}}\cap \overline{\Delta }_0=\Delta _{\mathfrak {n}}=\bigg \{ \alpha _D\pm \alpha _{1}, \quad \alpha _D\pm \alpha _{2}, \quad \alpha _D\pm \alpha _{3} \bigg \}\; \end{aligned}$$
(134)

with the constraint \(s_1 s_2 s_3=1\).

$$\begin{aligned}&\lambda = -\Delta \alpha _D +\left( \ell _1 +\frac{\ell _2+\ell _3}{2} \right) \alpha _1+\frac{\ell _2+\ell _3}{2}\alpha _2+\frac{-\ell _2+\ell _3}{2} \alpha _3+k \beta , \end{aligned}$$
(135)
$$\begin{aligned}&\rho =3\alpha _D+2 \alpha _1+\alpha _2+\frac{1}{2}\beta , \end{aligned}$$
(136)
$$\begin{aligned}&\lambda +\rho =( -\Delta +3) \alpha _D +\left( \ell _1 +\frac{\ell _2+\ell _3}{2} +2\right) \alpha _1+\left( \frac{\ell _2+\ell _3}{2} +1\right) \alpha _2\nonumber \\&\qquad \qquad + \frac{-\ell _2+\ell _3}{2} \alpha _3+\left( k+\frac{1}{2}\right) \beta . \end{aligned}$$
(137)

We compute

$$\begin{aligned} \begin{aligned}&n_{\alpha _D+\alpha _1}=5-\Delta +\ell _1+\frac{\ell _2+\ell _3}{2},\\&n_{\alpha _D-\alpha _1}=1-\Delta -\ell _1-\frac{\ell _2+\ell _3}{2},\\&n_{\alpha _D+\alpha _2}=4-\Delta +\frac{\ell _2+\ell _3}{2},\\&n_{\alpha _D-\alpha _2}=2-\Delta -\frac{\ell _2+\ell _3}{2},\\&n_{\alpha _D+\alpha _3}=3-\Delta -\frac{\ell _2-\ell _3}{2},\\&n_{\alpha _D-\alpha _3}=3-\Delta +\frac{\ell _2-\ell _3}{2}\,. \end{aligned} \end{aligned}$$
(138)

This means that if

$$\begin{aligned} \Delta =5+\ell _1+\frac{\ell _2+\ell _3}{2}-\mathbb {Z}_{>0}, \end{aligned}$$
(139)

then we will have \(\Psi _\lambda ^+\ne \varnothing \) with varying elements depending on what \(n_\beta \) are non-zero. We will directly go to the next step which is to analyze the wall condition for \(\beta \in \Delta _\mathfrak {n}\) such that \((\lambda +\rho ,\beta )=0\):

$$\begin{aligned} \begin{aligned}&n_{\alpha _D+\alpha _1}=0,\ \Delta =5+\ell _1+\frac{\ell _2+\ell _3}{2}, \ \Psi _\lambda ^+=\varnothing ,\\&n_{\alpha _D-\alpha _1}=0,\ \Delta =1-\ell _1-\frac{\ell _2+\ell _3}{2}, \ \Psi _\lambda ^+=\{\alpha _D+\alpha _1,\alpha _D\pm \alpha _2,\alpha _D\pm \alpha _3\},\\&n_{\alpha _D+\alpha _2}=0,\ \Delta =4+\frac{\ell _2+\ell _3}{2}, \ \Psi _\lambda ^+=\{\alpha _D+\alpha _1\},\\&n_{\alpha _D-\alpha _2}=0,\ \Delta =2-\frac{\ell _2+\ell _3}{2}, \ \Psi _\lambda ^+=\{\alpha _D+\alpha _1,\alpha _D+\alpha _2,\alpha _D\pm \alpha _3\},\\&n_{\alpha _D+\alpha _3}=0,\ \Delta =3-\frac{\ell _2-\ell _3}{2},\ \Psi _\lambda ^+={\left\{ \begin{array}{ll}\{\alpha _D+\alpha _1,\alpha _D+\alpha _2,\alpha _D-\alpha _3\}, \ \ell _2>\ell _3\\ \{\alpha _D+\alpha _1,\alpha _D+\alpha _2\}, \ \ell _2<\ell _3\end{array}\right. },\\&n_{\alpha _D-\alpha _3}=0,\ \Delta =3+\frac{\ell _2-\ell _3}{2},\ \Psi _\lambda ^+={\left\{ \begin{array}{ll}\{\alpha _D+\alpha _1,\alpha _D+\alpha _2\}, \ \ell _2>\ell _3\\ \{\alpha _D+\alpha _1,\alpha _D+\alpha _2,\alpha _D+\alpha _3\},\ \ell _2<\ell _3\end{array}\right. }\,. \end{aligned} \end{aligned}$$
(140)

After working out the condition (25) we find that the reducible points are

$$\begin{aligned} \Delta =\bigg (\bigg \{5+\ell _1+\frac{\ell _2+\ell _3}{2}-\mathbb {Z}_{>0}\bigg \}\bigg \backslash \bigg \{4+\frac{\ell _2+\ell _3}{2},3+\frac{|\ell _2-\ell _3|}{2} \bigg \}\bigg )\,. \end{aligned}$$
(141)

In the language of the parametrization \(\lambda _1\ge \lambda _2\ge |\lambda _3|\ge 0\) in Appendix A this becomes

$$\begin{aligned} \Delta =\bigg \{ \lambda _1+5-\mathbb {Z}_{>0}\bigg \} \setminus \bigg \{\lambda _2+4, |\lambda _3|+3 \bigg \}. \end{aligned}$$
(142)

Root System for SCAs

1.1 3d \(\mathcal {N}\) Even

The root system for \(\mathfrak {g}=D\left( \frac{\mathcal {N}}{2}, 2\right) \) is given by

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \beta _D , \quad \pm \beta _J, \quad \pm \beta _D \pm \beta _J , \quad \pm \delta _i \pm \delta _j \bigg \} \ , \end{aligned}$$
(143)
$$\begin{aligned}&\Delta _{\bar{1}}=\bigg \{ \pm \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _J \pm \delta _i \bigg \} \ , \end{aligned}$$
(144)

with \(i=1, \ldots , \frac{\mathcal {N}}{2}\) and the inner product is given by

$$\begin{aligned} (\beta _a, \beta _b)=2\delta _{ab}, \quad (\delta _i, \beta _a)=0, \quad (\delta _i, \delta _j)=-\delta _{ij}, \end{aligned}$$
(145)

with \(a,b=D,J\). In dictionary ordering \(\beta _D> \beta _J> \delta _1> \delta _2>\ldots >\delta _{\frac{\mathcal {N}}{2}}\), we find

$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm \beta _J, \quad \pm \delta _i \pm \delta _j \bigg \}, \end{aligned}$$
(146)
$$\begin{aligned}&\Delta _{1}^{+}=\overline{\Delta }_{1}^{+}=\bigg \{ \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _J\pm \delta _i \bigg \}, \end{aligned}$$
(147)
$$\begin{aligned}&\Delta _{\mathfrak {n}}\cap \overline{\Delta }_0=\Delta _{\mathfrak {n}}=\bigg \{ \beta _D+\beta _J, \quad \beta _D, \quad \beta _D-\beta _J \bigg \}. \end{aligned}$$
(148)

1.2 3d \(\mathcal {N}\) Odd

We have \(\mathfrak {g}=B\left( \frac{\mathcal {N}-1}{2}, 2\right) \). We have the root system

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \beta _D , \quad \pm \beta _J, \quad \pm \beta _D \pm \beta _J,\quad \pm \delta _i \pm \delta _j, \quad \pm \delta _i \bigg \} \ , \end{aligned}$$
(149)
$$\begin{aligned}&\Delta _{\bar{1}}=\bigg \{ \pm \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _J \pm \delta _i, \quad \pm \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _J \bigg \} \ , \end{aligned}$$
(150)

with \(i=1,\ldots , \frac{\mathcal {N}-1}{2}\). For \(\mathcal {N}=1\) the vectors \(\delta _i\)’s are absent, and correspondingly we disregard those roots containing these vectors. The inner product is given by (145), and with dictionary ordering \(\beta _D> \beta _J> \delta _1> \delta _2>\cdots >\delta _{\frac{\mathcal {N}-1}{2}}\) we obtain

$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm \beta _J, \quad \pm \delta _i \pm \delta _j, \quad \pm \delta _i, \bigg \}, \end{aligned}$$
(151)
$$\begin{aligned}&\overline{\Delta }_{1}^{+}=\bigg \{\pm \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _J \pm \delta _i \bigg \}, \end{aligned}$$
(152)
$$\begin{aligned}&\Delta _{1}^{+}\setminus \overline{\Delta }_{1}^{+}=\bigg \{ \pm \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _J \bigg \}, \end{aligned}$$
(153)
$$\begin{aligned}&\Delta _{\mathfrak {n}}\cap \overline{\Delta }_0=\Delta _{\mathfrak {n}}=\bigg \{ \beta _D+\beta _J, \quad \beta _D, \quad \beta _D-\beta _J \bigg \}. \end{aligned}$$
(154)

1.3 4d \(\mathcal {N}\ge 1\)

For \(\mathfrak {g}=\mathfrak {su}(4|\mathcal {N})\) (\(\mathfrak {g}=\mathfrak {psu}(4|4)\) for \(\mathcal {N}=4\)), it is useful to introduce a basis

$$\begin{aligned} \begin{aligned} (\varepsilon _i, \varepsilon _j)=\delta _{i,j} \ , \quad (\varepsilon _i, \delta _b)=0 \ , \quad (\delta _a, \delta _b)=-\delta _{a, b} \ , \end{aligned} \end{aligned}$$
(155)

with \(i,j=1, \ldots , 4\) and \(a, b=1, \ldots , \mathcal {N}\). Then we have

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm ( \varepsilon _i - \varepsilon _j ), \quad \pm (\delta _a - \delta _b) \bigg \} \ , \quad \Delta _{\bar{1}}=\bigg \{ \pm ( \varepsilon _i + \delta _a) \bigg \}, \end{aligned}$$
(156)
$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm ( \varepsilon _1 - \varepsilon _2 ), \quad \pm ( \varepsilon _3 - \varepsilon _4), \quad \pm (\delta _a - \delta _b) \bigg \}. \end{aligned}$$
(157)

Following [19, 41] we choose an ordering

$$\begin{aligned} \varepsilon _1> \varepsilon _2>-\varepsilon _3>-\varepsilon _4>-\delta _1> \cdots >-\delta _{\mathcal {N}}. \end{aligned}$$
(158)

We have positive simple roots

$$\begin{aligned}&\bigg \{ \varepsilon _{1}-\varepsilon _{2} , \quad \varepsilon _{4}-\varepsilon _{3} , \quad \varepsilon _{2}+\delta _{1} , \quad -\varepsilon _{4}-\delta _{\mathcal {N}}, \quad -\delta _{a}+\delta _{a+1} \quad (a=1, \ldots , \mathcal {N}-1) \bigg \}. \end{aligned}$$
(159)

We have

$$\begin{aligned}&\Delta _{1}^{+}=\overline{\Delta }_{1}^{+}=\bigg \{ \varepsilon _1 - \delta _a , \varepsilon _2 - \delta _a , -\varepsilon _3 + \delta _a , -\varepsilon _4 + \delta _a \bigg \}, \end{aligned}$$
(160)
$$\begin{aligned}&\Delta _{\mathfrak {n}}\cap \overline{\Delta }_0=\Delta _{\mathfrak {n}}=\bigg \{ \varepsilon _1 - \varepsilon _3 , \quad \varepsilon _1 - \varepsilon _4 , \quad \varepsilon _2 - \varepsilon _3 , \quad \varepsilon _2 - \varepsilon _4 \bigg \}. \end{aligned}$$
(161)

The highest weight vector is

$$\begin{aligned} \begin{aligned} \lambda&=-\frac{\Delta }{2}(\varepsilon _1+\varepsilon _2-\varepsilon _3-\varepsilon _4)+\frac{\ell _1}{2}(\varepsilon _1-\varepsilon _2)-\frac{\ell _2}{2}(\varepsilon _3-\varepsilon _4)\\&\quad + \frac{\mathcal {N}-4}{8 \mathcal {N}} R \left( \sum _{i=1}^4 \varepsilon _i- \sum _{a=1}^{\mathcal {N}} \delta _a \right) -\sum _{b=1}^{\mathcal {N}} \left( \lambda _b -\frac{|\lambda |}{\mathcal {N}} \right) \delta _b \,. \end{aligned} \end{aligned}$$
(162)

Here \(\ell _1\) and \(\ell _2\) are the two angular spins, and take integer values. The set of integers \(\{\lambda _a\}\) define a partition, see Appendix A.1. Note the factor with R is absent for the special case of \(\mathcal {N}=4\), where there is a reduction of \(\mathfrak {u}(1)\) symmetry, from \(\mathfrak {sl}(4|4)\) into \(\mathfrak {psl}(4|4)\).

The Weyl vector is given by

$$\begin{aligned}&\rho _0=\frac{1}{2} \left( 3 \varepsilon _1 +\varepsilon _2 -3\varepsilon _3 -\varepsilon _4 \right) -\sum _{a=1}^{\mathcal {N}}\frac{\mathcal {N}+1-2a}{2}\delta _a, \end{aligned}$$
(163)
$$\begin{aligned}&\rho _1=\frac{\mathcal {N}}{2}\left( \varepsilon _1+ \varepsilon _2 - \varepsilon _3- \varepsilon _4 \right) , \end{aligned}$$
(164)
$$\begin{aligned}&\rho = \left( \frac{3-\mathcal {N}}{2} \varepsilon _1 +\frac{1-\mathcal {N}}{2}\varepsilon _2 -\frac{3-\mathcal {N}}{2}\varepsilon _3 -\frac{1-\mathcal {N}}{2}\varepsilon _4 \right) -\sum _{a=1}^{\mathcal {N}} \frac{\mathcal {N}+1-2a}{2} \delta _a\,. \end{aligned}$$
(165)

We therefore obtain

$$\begin{aligned} \begin{aligned} \lambda +\rho&=\frac{-\Delta +3-\mathcal {N}}{2} \varepsilon _1 +\frac{-\Delta +1-\mathcal {N}}{2}\varepsilon _2 -\frac{-\Delta +3-\mathcal {N}}{2}\varepsilon _3 -\frac{-\Delta +1-\mathcal {N}}{2}\varepsilon _4 \\&\quad +\frac{\ell _1}{2}(\varepsilon _1-\varepsilon _2)-\frac{\ell _2}{2}(\varepsilon _3-\varepsilon _4)\\&\quad + \frac{\mathcal {N}-4}{8 \mathcal {N}} R \left( \sum _{i=1}^4 \varepsilon _i- \sum _{a=1}^{\mathcal {N}} \delta _a \right) -\sum _{a=1}^{\mathcal {N}} \frac{2\lambda _b -2\frac{|\lambda |}{\mathcal {N}} +\mathcal {N}+1-2a}{2} \delta _a\,. \end{aligned} \end{aligned}$$
(166)

For Step 1, we compute

$$\begin{aligned} \begin{aligned} (\lambda +\rho , \varepsilon _1)&=\frac{1}{2} \left( -\Delta +\ell _1+3-\mathcal {N}+\frac{\mathcal {N}-4}{4\mathcal {N}} R\right) , \\ (\lambda +\rho , \varepsilon _2)&=\frac{1}{2} \left( -\Delta -\ell _1+1-\mathcal {N}+\frac{\mathcal {N}-4}{4\mathcal {N}} R\right) , \\ (\lambda +\rho , \varepsilon _3)&=\frac{1}{2} \left( \Delta -\ell _2-3+\mathcal {N}+\frac{\mathcal {N}-4}{4\mathcal {N}} R\right) , \\ (\lambda +\rho , \varepsilon _4)&=\frac{1}{2} \left( \Delta +\ell _2-1+\mathcal {N}+\frac{\mathcal {N}-4}{4\mathcal {N}} R\right) , \\ (\lambda +\rho , \delta _a)&=\frac{1}{2} \left( \frac{\mathcal {N}-4}{4\mathcal {N}}R+2\lambda _a -2\frac{|\lambda |}{\mathcal {N}} +\mathcal {N}+1-2a \right) . \end{aligned} \end{aligned}$$
(167)

From this we can easily see that the module is reducible atFootnote 24

$$\begin{aligned} \Delta = {\left\{ \begin{array}{ll} \ell _1+\frac{\mathcal {N}-4}{2\mathcal {N}} R +2 \lambda _a-2\frac{|\lambda |}{\mathcal {N}} -2a +4&{} (\varepsilon _1-\delta _a),\\ -\ell _1+\frac{\mathcal {N}-4}{2\mathcal {N}}R +2 \lambda _a-2\frac{|\lambda |}{\mathcal {N}} -2a+2&{}(\varepsilon _2-\delta _a),\\ -\ell _2-\frac{\mathcal {N}-4}{2\mathcal {N}} R -2 \lambda _a+2\frac{|\lambda |}{\mathcal {N}} +2a+2-2\mathcal {N}&{} (\varepsilon _3-\delta _a),\\ \ell _2-\frac{\mathcal {N}-4}{2\mathcal {N}} R -2 \lambda _a+2\frac{|\lambda |}{\mathcal {N}} +2a-2\mathcal {N}&{}(\varepsilon _4-\delta _a).\\ \end{array}\right. } \end{aligned}$$
(168)

We next come to Step 2\(^{\prime }\).

A care is needed in this step since in the expression for \(\lambda +\rho \) in (166) the coefficients of the \(\varepsilon _i\) do depend non-trivially on the R-charge R. However, such a R-dependence drops out when we consider irreducibility in this step, since \(\sum _i \varepsilon _i -\sum _a \delta _a\) is orthogonal to all the roots corresponding to momentum generators. Indeed, we can compute

$$\begin{aligned} \begin{aligned}&n_{\varepsilon _1-\varepsilon _3}= -\Delta +\frac{\ell _1-\ell _2}{2}+3 -\mathcal {N},\quad n_{\varepsilon _1-\varepsilon _4}= -\Delta +\frac{\ell _1+\ell _2}{2}+2 -\mathcal {N},\\&n_{\varepsilon _2-\varepsilon _3}= -\Delta +\frac{-\ell _1-\ell _2}{2}+2 -\mathcal {N},\quad n_{\varepsilon _2-\varepsilon _4}= -\Delta +\frac{-\ell _1+\ell _2}{2}+1 -\mathcal {N}\,. \end{aligned} \end{aligned}$$
(169)

From this we can work out when the set \(\Psi _{\lambda , \mathrm{non-iso}}\) is non-empty. We again learn that only the effect of \(\mathcal {N}\) in the rest of the analysis is to shift \(\Delta \rightarrow \Delta +\mathcal {N}\).

The result is then obtained by combining (168) and (118), where \(\Delta \) in the latter is shifted by \(\mathcal {N}\).

1.4 5d \(\mathcal {N}=1\)

For this case, we use the root system for \(\mathfrak {f}_4\):

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \beta _D \pm \beta _{J_1}, \quad \pm \beta _D \pm \beta _{J_2}, \quad \pm \beta _{J_1} \pm \beta _{J_2}, \quad \pm \beta _D , \quad \pm \beta _{J_1}, \quad \pm \beta _{J_2}, \quad \pm \delta \bigg \} \ , \end{aligned}$$
(170)
$$\begin{aligned}&\Delta _{\bar{1}}=\bigg \{ \pm \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _{J_1} \pm \frac{1}{2}\beta _{J_2}\pm \frac{1}{2}\delta \bigg \} \ , \end{aligned}$$
(171)

with the inner product

$$\begin{aligned} \begin{aligned} (\beta _i, \beta _j)=\delta _{i,j}, \quad (\beta _i, \delta )=0, \quad (\delta , \delta )=-3. \end{aligned} \end{aligned}$$
(172)

Under an ordering \(\beta _D> \beta _{J_1}>\beta _{J_2}>\delta \),

$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm \beta _{J_1} \pm \beta _{J_2}, \quad \pm \beta _{J_1}, \quad \pm \beta _{J_2}, \quad \pm \delta \bigg \}, \end{aligned}$$
(173)
$$\begin{aligned}&\Delta _{1}^{+}=\overline{\Delta }_{1}^{+}=\bigg \{ \frac{1}{2}\beta _D \pm \frac{1}{2}\beta _{J_1} \pm \frac{1}{2}\beta _{J_2} \pm \frac{1}{2}\delta \bigg \}, \end{aligned}$$
(174)
$$\begin{aligned}&\Delta _{\mathfrak {n}}\cap \overline{\Delta }_0=\Delta _{\mathfrak {n}}=\bigg \{ \beta _D \pm \beta _{J_1}, \quad \beta _D \pm \beta _{J_2}, \quad \beta _D \bigg \}. \end{aligned}$$
(175)

We have

$$\begin{aligned}&\lambda =-\Delta \beta _D + \left( \ell _1 +\frac{\ell _2}{2}\right) \beta _{J_1}+\frac{\ell _2}{2} \beta _{J_2} + k\delta , \end{aligned}$$
(176)
$$\begin{aligned}&\rho _0=\frac{5}{2}\beta _D+\frac{3}{2}\beta _{J_1}+\frac{1}{2}\beta _{J_2}+\frac{1}{2}\delta ,\quad \rho _1=2 \beta _D, \end{aligned}$$
(177)
$$\begin{aligned}&\rho =\frac{1}{2}\beta _D+\frac{3}{2}\beta _{J_1}+\frac{1}{2}\beta _{J_2}+\frac{1}{2}\delta , \end{aligned}$$
(178)
$$\begin{aligned}&\lambda +\rho =\left( -\Delta +\frac{1}{2}\right) \beta _D+\left( \ell _1+\frac{\ell _2}{2}+\frac{3}{2}\right) \beta _{J_1}+\left( \frac{\ell _2}{2}+\frac{1}{2}\right) \beta _{J_2}+ \left( k+\frac{1}{2} \right) \delta . \end{aligned}$$
(179)

Step 1 gives

$$\begin{aligned} \Delta =\frac{1}{2}+s_1\left( \ell _1+\frac{\ell _2}{2}+\frac{3}{2}\right) +s_2\left( \frac{\ell _2}{2}+\frac{1}{2}\right) -3\sigma \left( k+\frac{1}{2} \right) , \end{aligned}$$
(180)

with \(s_1, s_2, \sigma =\pm 1\). In Step 2\(^{\prime }\) the shift of the value of \(\Delta \) is 2, as originating from the coefficient of \(\beta _D\) in \(\rho _1\). Hence the module is reducible at these values, as well at at (130) with shift of \(\Delta \) by 2.

1.5 6d \(\mathcal {N}=(0,1)\)

The root system for \(\mathfrak {g}=\mathfrak {osp}(8|2)=D(4,1)\) is

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \alpha _i \pm \alpha _j, \quad \pm \beta , \bigg \} \quad (i,j=D,1,2,3), \end{aligned}$$
(181)
$$\begin{aligned}&\Delta _{\bar{1}}=\bigg \{ \frac{s_D}{2}\alpha _D + \frac{s_1}{2}\alpha _1 + \frac{s_2}{2}\alpha _2 + \frac{s_3}{2}\alpha _3 + \frac{\sigma }{2}\beta \bigg \} \ , \end{aligned}$$
(182)

where \(s_D, s_1, \ldots , s_3, \sigma =\pm 1\) with the constraint \(s_D s_1 s_2 s_3=1\), and we have

$$\begin{aligned} (\alpha _i, \alpha _j)=\delta _{ij}, \quad (\beta , \beta )=-2. \end{aligned}$$
(183)

Under an ordering \(\alpha _D>\alpha _1>\alpha _2>\alpha _3>\beta \),

$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm \alpha _1 \pm \alpha _2, \quad \pm \alpha _1 \pm \alpha _3, \quad \pm \alpha _2 \pm \alpha _3, \quad \pm \beta \bigg \}, \end{aligned}$$
(184)
$$\begin{aligned}&\Delta _{1}^{+}=\overline{\Delta }_{1}^{+}=\bigg \{ \frac{1}{2}\alpha _D +\frac{s_1}{2}\alpha _1 + \frac{s_2}{2}\alpha _2 +\frac{s_3}{2}\alpha _3 +\frac{\sigma }{2}\beta \bigg \}, \end{aligned}$$
(185)
$$\begin{aligned}&\Delta _{\mathfrak {n}}\cap \overline{\Delta }_0=\Delta _{\mathfrak {n}}=\bigg \{ \alpha _D\pm \alpha _{1}, \quad \alpha _D\pm \alpha _{2}, \quad \alpha _D\pm \alpha _{3} \bigg \}, \end{aligned}$$
(186)

with the constraint \(s_1 s_2 s_3=1\).

$$\begin{aligned}&\lambda = -\Delta \alpha _D +\left( \ell _1 +\frac{\ell _2+\ell _3}{2} \right) \alpha _1+\frac{\ell _2+\ell _3}{2}\alpha _2+\frac{-\ell _2+\ell _3}{2} \alpha _3+\frac{k}{2} \beta , \end{aligned}$$
(187)
$$\begin{aligned}&\rho =\alpha _D+2 \alpha _1+\alpha _2+\frac{1}{2}\beta , \end{aligned}$$
(188)
$$\begin{aligned}&\lambda +\rho =( -\Delta +1) \alpha _D +\left( \ell _1 +\frac{\ell _2+\ell _3}{2} +2\right) \alpha _1+\left( \frac{\ell _2+\ell _3}{2} +1\right) \alpha _2\nonumber \\&\qquad \qquad \,\, + \frac{-\ell _2+\ell _3}{2} \alpha _3+\left( \frac{k}{2}+\frac{1}{2}\right) \beta . \end{aligned}$$
(189)

Then \((\lambda +\rho , \alpha )=0\) for an odd isotropic root \(\alpha \) from (185) gives

$$\begin{aligned} \Delta&=1+\left( \ell _1 +\frac{\ell _2+\ell _3}{2} +2\right) s_1+\left( \frac{\ell _2+\ell _3}{2} +1\right) s_2\nonumber \\&\quad +\frac{-\ell _2+\ell _3}{2} s_3 +4\sigma \left( \frac{k}{2}+\frac{1}{2}\right) . \end{aligned}$$
(190)

The module is reducible either at these values or at values (141), with \(\Delta \) shifted by 2.

1.6 6d \(\mathcal {N}=(0,2)\)

We have \(\mathfrak {g}=\mathfrak {osp}(8|4)=D(4,2)\), which has the root system

$$\begin{aligned}&\Delta _{\bar{0}}=\bigg \{ \pm \alpha _i \pm \alpha _j, \quad \pm \beta _1 \pm \beta _2, \quad \pm \beta _1 , \quad \pm \beta _2 \bigg \} \ , \quad (i,j=D,1,2,3) \end{aligned}$$
(191)
$$\begin{aligned}&\Delta _{\bar{1}}=\bigg \{ \frac{s_D}{2}\alpha _D + \frac{s_1}{2}\alpha _1 + \frac{s_2}{2}\alpha _2 + \frac{s_3}{2}\alpha _3 + \frac{\sigma _1}{2}\beta _1 +\frac{\sigma _2}{2}\beta _2 \bigg \}, \end{aligned}$$
(192)

where \(s_D, s_1, \ldots , s_3, \sigma _1, \sigma _2=\pm 1\) with the constraint \(s_D s_1 s_2 s_3=1\), and with

$$\begin{aligned} (\alpha _i, \alpha _j)=\delta _{ij}, \quad (\beta _a, \beta _b)=-\delta _{ij}. \end{aligned}$$
(193)

Under an ordering \(\alpha _D>\alpha _1>\alpha _2>\alpha _3>\beta _1>\beta _2\),

$$\begin{aligned}&\Delta _{\mathfrak {l}}=\bigg \{ \pm \alpha _1 \pm \alpha _2, \quad \pm \alpha _1 \pm \alpha _3, \quad \pm \alpha _2 \pm \alpha _3, \quad \pm \beta _1 \pm \beta _2, \quad \pm \beta _1 , \quad \pm \beta _2\bigg \}, \end{aligned}$$
(194)
$$\begin{aligned}&\Delta _{1}^{+}=\overline{\Delta }_{1}^{+}=\bigg \{ \frac{1}{2}\alpha _D + \frac{s_1}{2}\alpha _1 + \frac{s_2}{2}\alpha _2 +\frac{s_3}{2}\alpha _3 + \frac{\sigma _1}{2}\beta _ 1+ \frac{\sigma _2}{2}\beta _ 2 \bigg \}, \end{aligned}$$
(195)
$$\begin{aligned}&\Delta _{\mathfrak {n}}\cap \overline{\Delta }_0=\Delta _{\mathfrak {n}}=\bigg \{ \alpha _D\pm \alpha _{1}, \quad \alpha _D\pm \alpha _{2}, \quad \alpha _D\pm \alpha _{3} \bigg \}. \end{aligned}$$
(196)

with the constraint \(s_1 s_2 s_3=1\). We have

$$\begin{aligned}&\lambda = -\Delta \alpha _D +\left( \ell _1 +\frac{\ell _2+\ell _3}{2} \right) \alpha _1+\frac{\ell _2+\ell _3}{2} \alpha _2+\frac{-\ell _2+\ell _3}{2} \alpha _3+k_1 \beta _1+k_2 \beta _2, \end{aligned}$$
(197)
$$\begin{aligned}&\rho =-\alpha _D+2 \alpha _1+\alpha _2+\frac{3}{2}\beta _1+\frac{1}{2}\beta _2, \end{aligned}$$
(198)
$$\begin{aligned}&\lambda +\rho =( -\Delta -1) \alpha _D +\left( \ell _1 +\frac{\ell _2+\ell _3}{2}+2\right) \alpha _1+ \left( \frac{\ell _2+\ell _3}{2}+1 \right) \alpha _2+\frac{-\ell _2+\ell _3}{2} \alpha _3 \nonumber \\&\qquad \qquad +\left( k_1 +\frac{3}{2}\right) \beta _1+\left( k_2 +\frac{1}{2}\right) \beta _2. \end{aligned}$$
(199)

Then \((\lambda +\rho , \alpha )=0\) for an odd isotropic root \(\alpha \) from (195) gives

$$\begin{aligned} \Delta&=-1+\left( \ell _1 +\frac{\ell _2+\ell _3}{2}+2\right) s_1+ \left( \frac{\ell _2+\ell _3}{2}+1 \right) s_2+\frac{-\ell _2+\ell _3}{2} s_3 \nonumber \\&\quad +2\sigma _1 \left( k_1 +\frac{3}{2}\right) +2\sigma _2 \left( k_2 +\frac{1}{2}\right) . \end{aligned}$$
(200)

The module is reducible either at these values or at (141) with \(\Delta \) shifted by minus 4.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sen, K., Yamazaki, M. Polology of Superconformal Blocks. Commun. Math. Phys. 374, 785–821 (2020). https://doi.org/10.1007/s00220-019-03572-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03572-8

Navigation