Skip to main content
Log in

Kostant, Steinberg, and the Stokes matrices of the tt*-Toda equations

  • Published:
Selecta Mathematica Aims and scope Submit manuscript

Abstract

We propose a Lie-theoretic definition of the tt*-Toda equations for any complex simple Lie algebra \({\mathfrak {g}}\), based on the concept of topological–antitopological fusion which was introduced by Cecotti and Vafa. Our main results concern the Stokes data of a certain meromorphic connection, whose isomonodromic deformations are controlled by these equations. First, by exploiting a framework introduced by Boalch, we show that this data has a remarkable structure. It can be described using Kostant’s theory of Cartan subalgebras in apposition and Steinberg’s theory of conjugacy classes of regular elements, and it can be visualized on the Coxeter Plane. Second, we compute canonical Stokes data for a certain family of solutions of the tt*-Toda equations in terms of their asymptotics. To do this, we compute the Stokes data of an auxiliary meromorphic connection, related to the original meromorphic connection by a loop group Iwasawa factorization.

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
Fig. 2

Similar content being viewed by others

Notes

  1. This implies that the combined connection \(d+\omega +{\hat{\omega }}\) is flat, and hence the monodromy data of \({\hat{\omega }}\) is independent of z, just as the monodromy data of \({\hat{\alpha }}\) was independent of \(z,{\bar{z}}\). However, the isomonodromic deformation of \({\hat{\omega }}\), given by the explicit monomial entries of \(\omega \), is very simple, whereas that of \({\hat{\alpha }}\) is complicated, being given by solutions w of the tt*-Toda equations.

References

  1. Aldrovandi, E., Falqui, G.: Geometry of Higgs and Toda fields on Riemann surfaces. J. Geom. Phys. 17, 25–48 (1995)

    Article  MathSciNet  Google Scholar 

  2. Balan, V., Dorfmeister, J.: Birkhoff decompositions and Iwasawa decompositions for loop groups. Tohoku Math. J. 53, 593–615 (2001)

    Article  MathSciNet  Google Scholar 

  3. Baraglia, D.: Cyclic Higgs bundles and the affine Toda equations. Geometriae Dedicata 174, 25–42 (2015)

    Article  MathSciNet  Google Scholar 

  4. Boalch, P.: Stokes matrices, Poisson Lie groups and Frobenius manifolds. Invent. Math. 146, 479–506 (2001)

    Article  MathSciNet  Google Scholar 

  5. Boalch, P.: G-Bundles, isomonodromy, and quantum Weyl groups. Int. Math. Res. Notices 2002, 1129–1166 (2002)

    Article  MathSciNet  Google Scholar 

  6. Boalch, P., Yamakawa, D.: Twisted wild character varieties. arXiv:1512.08091

  7. Bolton, J., Pedit, F., Woodward, L.: Minimal surfaces and the affine Toda field model. J. Reine Angew. Math. 459, 119–150 (1995)

    MathSciNet  MATH  Google Scholar 

  8. Bridgeland, T., Toledano Laredo, V.: Stokes factors and multilogarithms. J. Reine Angew. Math. 682, 89–128 (2012)

    MathSciNet  MATH  Google Scholar 

  9. Burstall, F.E., Pedit, F.: Harmonic maps via Adler–Kostant–Symes theory. Fordy, A.P., Wood, J.C. (Eds.) Harmonic Maps and Integrable Systems. Aspects of Mathematics, vol. E23, pp. 221–272. Vieweg, Wiesbaden (1994)

    Chapter  Google Scholar 

  10. Casselman, W.: Essays on Coxeter groups: Coxeter elements in finite Coxeter groups. https://www.math.ubc.ca/~cass/research/pdf/Element.pdf (downloaded 24 September 2017)

  11. Cecotti, S., Fendley, P., Intriligator, K., Vafa, C.: A new supersymmetric index. Nuclear Phys. B 386, 405–452 (1992)

    Article  MathSciNet  Google Scholar 

  12. Cecotti, S., Vafa, C.: Topological—anti-topological fusion. Nuclear Phys. B 367, 359–461 (1991)

    Article  MathSciNet  Google Scholar 

  13. Cecotti, S., Vafa, C.: On classification of \(N=2\) supersymmetric theories. Commun. Math. Phys. 158, 569–644 (1993)

    Article  MathSciNet  Google Scholar 

  14. Chi, C.Y.: On the Toda systems of VHS type. arXiv:1302.1287

  15. Coxeter, H.S.M.: Regular Complex Polytopes. Cambridge University Press, Cambridge (1974)

    MATH  Google Scholar 

  16. Dorey, P.: Root systems and purely elastic S-matrices. Nuclear Phys. B 358, 654–676 (1991)

    Article  MathSciNet  Google Scholar 

  17. Dorey, P.: Root systems and purely elastic S-matrices (II). Nuclear Phys. B 374, 741–761 (1992)

    Article  MathSciNet  Google Scholar 

  18. Drinfeld, V.G., Sokolov, V.V.: Lie algebras and equations of Korteweg–de Vries type. J. Soviet Math. 30, 1975–2036 (1985)

    Article  Google Scholar 

  19. Dubrovin, B.: Geometry and integrability of topological–antitopological fusion. Commun. Math. Phys. 152, 539–564 (1993)

    Article  MathSciNet  Google Scholar 

  20. Fokas, A.S., Its, A.R., Kapaev, A.A., Novokshenov, V.Y.: Painlevé Transcendents: The Riemann–Hilbert Approach. Mathematical Surveys and Monographs, vol. 128. American Mathematical Society, Providence (2006)

    Chapter  Google Scholar 

  21. Freeman, M.D.: On the mass spectrum of affine Toda field theory. Nuclear Phys. B 261, 57–61 (1991)

    MathSciNet  Google Scholar 

  22. Frenkel, E., Gross, B.: A rigid irregular connection on the projective line. Ann. Math. 170, 1469–1512 (2009)

    Article  MathSciNet  Google Scholar 

  23. Guest, M.A.: Harmonic Maps, Loop Groups, and Integrable Systems. LMS Student Texts 38. Cambridge University Press, Cambridge (1997)

  24. Guest, M.A., Ho, N.-K.: A Lie-theoretic description of the solution space of the tt*-Toda equations. Math. Phys. Anal. Geom. 20(4), Art. 24, 27 pp (2017)

  25. Guest, M.A., Its, A., Lin, C.S.: Isomonodromy aspects of the tt* equations of Cecotti and Vafa I. Stokes data. Int. Math. Res. Not. 2015, 11745–11784 (2015)

    MathSciNet  MATH  Google Scholar 

  26. Guest, M.A., Its, A., Lin, C.S.: Isomonodromy aspects of the tt* equations of Cecotti and Vafa II. Riemann–Hilbert problem. Commun. Math. Phys. 336, 337–380 (2015)

    Article  Google Scholar 

  27. Guest, M.A., Its, A., Lin, C.S.: Isomonodromy aspects of the tt* equations of Cecotti and Vafa III. Iwasawa factorization and asymptotics. Commun. Math. Phys. (to appear)

  28. Guest, M.A., Lin, C.S.: Nonlinear PDE aspects of the tt* equations of Cecotti and Vafa. J. Reine Angew. Math. 689, 1–32 (2014)

    Article  MathSciNet  Google Scholar 

  29. Helgason, S.: Differential Geometry, Lie groups and Symmetric Spaces. Graduate Studies in Mathematics, vol. 34, American Mathematical Society, Providence (2001)

  30. Hitchin, N.J.: Lie groups and Teichmüller space. Topology 31, 449–473 (1992)

    Article  MathSciNet  Google Scholar 

  31. Humphreys, J.E.: Reflection Groups and Coxeter Groups. Cambridge Studies in Advanced Mathematics, vol. 29. Cambridge University Press, Cambridge (1992)

  32. Katzarkov, L., Kontsevich, M., Pantev, T.: Hodge theoretic aspects of mirror symmetry, From Hodge Theory to Integrability and TQFT: tt*-geometry. Donagi, R.Y., Wendland, K. (Eds.) Proceedings of Symposia in Pure Mathematics, vol. 78, pp. 87–174. American Mathematical Society, Providence (2008)

  33. Knapp, A.W.: Lie Groups Beyond an Introduction. Progress in Mathematics, vol. 140, Birkhäuser, Basel (2002)

  34. Kostant, B.: The principal three-dimensional subgroup and the Betti numbers of a complex simple Lie group. Am. J. Math. 81, 973–1032 (1959)

    Article  MathSciNet  Google Scholar 

  35. Kostant, B.: The McKay correspondence, the Coxeter element and representation theory. Astérisque, hors série, pp. 209–255 (1985)

  36. Kostant, B.: The Coxeter element and the branching law for the finite subgroups of SU(2). Davis, C., Ellers, E.W. (Eds.) The Coxeter Legacy: Reflections and Projections, pp. 63–70. American Mathematical Society, Providence (2006)

  37. Kostant, B.: Experimental evidence for the occurrence of \(E_8\) in nature and the radii of the Gosset circles. Sel. Math. New Ser. 16, 419–438 (2010)

    Article  Google Scholar 

  38. Labourie, F.: Cyclic surfaces and Hitchin components in rank \(2\). Ann. Math. 185, 1–58 (2017)

    Article  MathSciNet  Google Scholar 

  39. Leznov, A.N., Saveliev, M.V.: Group-Theoretical Methods for Integration of Nonlinear Dynamical Systems, Progress in Mathematical Physics 15. Springer, Berlin (1992)

    Book  Google Scholar 

  40. McIntosh, I.: Global solutions of the elliptic 2D periodic Toda lattice. Nonlinearity 7, 85–108 (1994)

    Article  MathSciNet  Google Scholar 

  41. Mikhailov, A.V., Olshanetsky, M.A., Perelomov, A.M.: Two-dimensional generalized Toda lattice. Commun. Math. Phys. 79, 473–488 (1981)

    Article  MathSciNet  Google Scholar 

  42. Mochizuki, T.: Harmonic bundles and Toda lattices with opposite sign. arXiv:1301.1718

  43. Mochizuki, T.: Harmonic bundles and Toda lattices with opposite sign II. Commun. Math. Phys. 328, 1159–1198 (2014)

    Article  MathSciNet  Google Scholar 

  44. Nirov, Kh.S Razumov, A.V.: Toda equations associated with loop groups of complex classical Lie groups. Nuclear Phys. B 782, 241–275 (2007)

    Article  MathSciNet  Google Scholar 

  45. Olive, D., Turok, N.: The symmetries of Dynkin diagrams and the reduction of Toda field equations. Nuclear Phys. B 215, 470–494 (1983)

    Article  MathSciNet  Google Scholar 

  46. Pressley, A., Segal, G.B.: Loop Groups. Oxford University Press, Oxford (1986)

    MATH  Google Scholar 

  47. Simpson, C.T.: Katz’s middle convolution algorithm. Pure Appl. Math. Q. 5, 781–852 (2009)

    Article  MathSciNet  Google Scholar 

  48. Steinberg, R.: Finite reflection groups. Trans. Am. Math. Soc. 91, 493–504 (1959)

    Article  MathSciNet  Google Scholar 

  49. Steinberg, R.: Conjugacy Classes in Algebraic Groups. Lecture Notes in Mathematics, vol. 366, Springer, Berlin (1974)

    Book  Google Scholar 

  50. Wilson, G.: The modified Lax and two-dimensional generalized Toda lattice equations associated with simple Lie algebras. Ergod. Theory Dyn. Syst. 1, 361–380 (1981)

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank Eckhard Meinrenken for useful conversations, and for suggesting the relevance of the Coxeter Plane and the article [37] by Kostant. They thank Bill Casselman for clarifying the relation between the Coxeter Plane and Coxeter groups ([10] and “Appendix B”). They also thank Philip Boalch for discussions, and for pointing out the article [22] by Frenkel and Gross as well as the relevance of twisted connections in the sense of [6]. The first author was partially supported by JSPS Grant (A) 25247005. He is grateful to the National Center for Theoretical Sciences for excellent working conditions and financial support. The second author was partially supported by MOST Grants 105-2115-M-007-006 and 106-2115-M-007-004.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nan-Kuo Ho.

Additional information

Publisher's Note

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

Appendices

Appendix A: Formal solutions

This section is included primarily as motivation for Lemma 4.4. We sketch a proof of the existence of a formal solution to the equation

$$\begin{aligned} \frac{d\Psi }{d\xi }= \left( \sum _{k=-r-1}^\infty A_{k+1} \xi ^k \right) \Psi \end{aligned}$$
(A.1)

following [20], Proposition 1.1, but using Lie-theoretic notation as far as possible. Here \(r\in {\mathbb {N}}\) and all coefficients \(A_k\) belong to \({\mathfrak {g}}\), which we take to be the Lie algebra of a (simple) matrix Lie group G.

We assume that the leading coefficient \(A_{-r}\) is regular, hence contained in a unique Cartan subalgebra \({\mathfrak {h}}_1\) (thus we have \(\beta (A_{-r})\ne 0\) for all roots \(\beta \in \Delta _1\) with respect to \({\mathfrak {h}}_1\)).

Let \(P\in G\). Then \(\Lambda _{-r}=\mathrm {Ad}(P^{-1})A_{-r}\) is also regular, and contained in a unique Cartan subalgebra \({\mathfrak {h}}_2\) (namely \(\mathrm {Ad}(P^{-1}){\mathfrak {h}}_1\)). Let \({\mathfrak {g}}={\mathfrak {h}}_2\oplus (\oplus _\alpha {\mathfrak {g}}_\alpha )\) be the eigenspace decomposition for the adjoint action of \({\mathfrak {h}}_2\).

Proposition A.1

There is a unique formal fundamental solution \(\Psi _f\) of equation (A.1) of the form

$$\begin{aligned} \Psi _f(\xi )=P \left( \sum _{k=0}^\infty \psi _k\xi ^k \right) \exp ^{\Lambda (\xi )}, \end{aligned}$$
(A.2)

where \(\psi _0=I\) and \(\displaystyle \Lambda (\xi )=\Lambda _0\log \xi +\sum _{k=-r}^{-1}\frac{\Lambda _{k}}{k}\xi ^{k}\) with all \(\Lambda _k\in {\mathfrak {h}}_2\).

Proof

Consider

$$\begin{aligned} \Psi _f(\xi )=P \left( \sum _{k=0}^\infty Y_k \xi ^k \right) \exp ^{\Upsilon (\xi )}. \end{aligned}$$
(A.3)

We shall show that it is possible to find \(Y_k\) and \(\Upsilon (\xi )=\Lambda (\xi )+\sum _{k=1}^\infty \frac{\Lambda _k}{k}\xi ^k\) with \(Y_0=I\), all \(Y_k\in \oplus _\alpha {\mathfrak {g}}_\alpha \), and all \(\Lambda _k\in {\mathfrak {h}}_2\), such that (A.3) formally satisfies (A.1). As (A.3) can be rewritten in the required form (A.2), this will be sufficient.

Substitution into (A.1) leads to the recurrence relations

$$\begin{aligned} \Lambda _{-r+k}+[Y_k,\Lambda _{-r}]=F_{-r+k} \end{aligned}$$
(A.4)

for \(k\in {\mathbb {N}}\), where \(F_{-r+1}=P^{-1}A_{-r+1}P\) and, for \(k\ge 2\),

$$\begin{aligned} F_{-r+k}= & {} P^{-1}A_{-r+k}P+\sum _{n=1}^{k-1}(P^{-1}A_{-r+k-n}PY_n-Y_n\Lambda _{-r+k-n})\\&+\left\{ \begin{array}{ll}0&{}k=2,3,\ldots ,r \\ -(k-r)Y_{-r+k} &{} k=r+1,r+2,\ldots \end{array}\right. \end{aligned}$$

Taking components of (A.4) in \({\mathfrak {g}}^\prime ={\mathfrak {h}}_2\), \({\mathfrak {g}}^{\prime \prime }=\oplus _\alpha {\mathfrak {g}}_\alpha \) we obtain

$$\begin{aligned} \Lambda _{-r+k}=\text {Proj}^\prime (F_{-r+k}),\quad [Y_k,\Lambda _{-r}]=\text {Proj}^{\prime \prime }(F_{-r+k}). \end{aligned}$$

As \(\mathrm {Ad}\,\Lambda _{-r}\) is invertible on \(\oplus _\alpha {\mathfrak {g}}_\alpha \), all of the \(\Lambda _k\), \(Y_k\) can be determined recursively. \(\square \)

Appendix B: Singular directions and the (enhanced) Coxeter Plane

Let \({\mathfrak {g}}\) be a complex simple Lie algebra with \(\mathrm {rank}\,{\mathfrak {g}}= l>1\). Let \({\mathfrak {h}}\) be a Cartan subalgebra, and \(\Delta _+\) a choice of positive roots. The vectors \(H_\alpha \in {\mathfrak {h}}\) are defined as in Sect. 2 by \(B(h,H_\alpha )=\alpha (h)\). They span a polyhedron in the vector space \({\mathfrak {h}}\). The Coxeter Plane consists of a certain real two-dimensional subspace of \({\mathfrak {h}}\) together with the orthogonal projections of all points \(H_\alpha \in {\mathfrak {h}}\) onto this plane. As a visual aid, the rays (or “spokes”) from the origin to these points, and the concentric circles (or “wheels”) passing through these points, may be drawn. For the Lie algebra \({\mathfrak {e}}_8\) this picture is very well known and appears as the frontispiece of Coxeter’s book [15]. The version in Fig. 3 is taken from [10]. There are 60 spokes and 8 wheels in this case.

Fig. 3
figure 3

Coxeter Plane with root projections for \({\mathfrak {e}}_8\)

We shall sketch three descriptions of the Coxeter Plane, all of them rather indirect. Then we shall explain how the Stokes data of the meromorphic o.d.e. of Sect. 4 (or that of Sect. 6) provides a more direct description of the Coxeter Plane, which also illustrates its properties effectively.

The first (and, as far as we know, original) description is aesthetic: the two-dimensional plane is chosen to give the “most symmetrical projection”.

This was made more precise by Kostant [37], and it is this second description that we shall make use of. It depends on the choice of the Lie algebra element \(E_+\), hence the choice of a Cartan subalgebra \({\mathfrak {h}}^\prime ={\mathfrak {g}}^{E_+}\) in apposition to \({\mathfrak {h}}\). We have the real subspace

$$\begin{aligned} {\mathfrak {h}}^\prime _\sharp =\{ X\in {\mathfrak {h}}^\prime \ \vert \ \beta (X)\in {\mathbb {R}}\ \text {for all}\ \beta \in \Delta ^\prime \}. \end{aligned}$$

Recall that the Coxeter number of \({\mathfrak {g}}\) is \(s=1+\sum _{i=1}^l q_i = \sum _{i=0}^l q_i\), where \(\psi =\sum _{i=1}^l q_i \alpha _i\) and \(q_0=1\), and that \(\tau =\mathrm {Ad}P_0\) acts on \({\mathfrak {h}}^\prime \) as a Coxeter element, with \(\tau ^s=1\). Let us choose (for simplicity) the specific coefficients \(E_+=\sum _{i=0}^l \sqrt{q_i} e_{\alpha _i}\). Then \(E_+\in {\mathfrak {h}}^\prime ={\mathfrak {h}}^\prime _\sharp \otimes {\mathbb {C}}\) is isotropic with respect to B, hence defines an oriented real two-dimensional subspace Y of \({\mathfrak {h}}^\prime _\sharp \). Let \(Q:{\mathfrak {h}}^\prime _\sharp \rightarrow Y\) be orthogonal projection with respect to the (positive definite) inner product \(B|_{{\mathfrak {h}}^\prime _\sharp }\). Then [37, section 0.2] Y is the essentially unique plane with the property that the projection Q commutes with the action of the Coxeter element \(\tau \). This gives a precise meaning to “most symmetrical projection”.

The third (more abstract) description is based on the theory of Coxeter groups [10, 31, 48], and in our situation this means the Weyl group of \({\mathfrak {g}}\) with respect to a Cartan subalgebra \({\mathfrak {h}}\). We sketch this theory, referring to [10] and section 3.19 of [31] for further details. Any product

$$\begin{aligned} \tau =s_1\cdots s_l \end{aligned}$$

of reflections in simple roots is called a Coxeter element. We may assume that

$$\begin{aligned} \tau = xy,\quad x=s_1\cdots s_k, y=s_{k+1}\cdots s_l \end{aligned}$$

for some k, where \(s_1,\ldots ,s_k\) commute and \(s_{k+1},\ldots ,s_l\) commute. Thus \(x^2=1\) and \(y^2=1\) and xy generate a dihedral group, in fact the unique dihedral subgroup of the Weyl group which contains \(\tau \). The subspaces

$$\begin{aligned} \mathrm {Ker}\,\alpha _1\cap \cdots \cap \mathrm {Ker}\,\alpha _k, \mathrm {Ker}\,\alpha _{k+1}\cap \cdots \cap \mathrm {Ker}\,\alpha _l \end{aligned}$$

contain distinguished real lines \(l_x,l_y\) which can be constructed explicitly from a certain “Perron–Frobenius eigenvalue” of the Cartan matrix. The Coxeter Plane is the real span of \(l_x,l_y\). The lines of the Coxeter Plane are given by taking the orthogonal complements of the intersections of all root hyperplanes \(\mathrm {Ker}\,\alpha _i\) with this plane.

It can be shown that \(\tau \) acts on this plane as rotation through \(2\pi /s\), and xy act by reflection in adjacent lines (given by the intersections of \(\mathrm {Ker}\,\alpha _1\cap \cdots \cap \mathrm {Ker}\,\alpha _k, \mathrm {Ker}\,\alpha _{k+1}\cap \cdots \cap \mathrm {Ker}\,\alpha _l\) with the plane). It follows that there are exactly s equally spaced lines (thus 2s spokes) in the Coxeter Plane, whose symmetry group is the dihedral group generated by x and y.

As explained in [10], the relation with the Cartan matrix leads to an efficient algorithm for drawing this version of the Coxeter Plane. We are grateful to Bill Casselman for the current version of [10] which contains these pictures.

It turns out that the Stokes data of the meromorphic o.d.e. of Sect. 4 provides a fourth description of the Coxeter Plane. To see this, we observe that the diagram of singular directions is related to Kostant’s plane Y, as follows. First we introduce the notation

$$\begin{aligned} E_+=\mathfrak {R}(E_+)+ {\scriptstyle \sqrt{-1}}\, \mathfrak {I}(E_+),\quad \mathfrak {R}(E_+),\mathfrak {I}(E_+)\in {\mathfrak {h}}^\prime _\sharp \end{aligned}$$

for the decomposition of \(E_+\) with respect to the real subspace \({\mathfrak {h}}^\prime _\sharp \).

Proposition B.1

Let us identify Kostant’s plane Y with \({\mathbb {C}}\) by identifying the orthonormal basis \((2/s)^{\frac{1}{2}}\mathfrak {R}(E_+), (2/s)^{\frac{1}{2}}\mathfrak {I}(E_+)\) with \(1,\sqrt{-1}\). Then the points \(Q(H_{\beta })\) are identified with the points \((2/s)^{\frac{1}{2}}\beta (E_+)\), for all \(\beta \in {\mathfrak {h}}^\prime \).

Proof

We use the key fact that the complex conjugate of \(E_+\) with restect to \({\mathfrak {h}}^\prime _\sharp \) is given by \({\bar{E}}_+ = E_-\) ([37], Theorems 1.11 and 1.12). Then by direct calculation we obtain

$$\begin{aligned}&B(\mathfrak {R}(E_+),\mathfrak {R}(E_+))=\tfrac{s}{2} \\&B(\mathfrak {I}(E_+),\mathfrak {I}(E_+))=\tfrac{s}{2} \\&B(\mathfrak {R}(E_+),\mathfrak {I}(E_+))=0 \end{aligned}$$

(so \(E_+\) is isotropic, as stated earlier). The projection to Y of a vector \(X\in {\mathfrak {h}}^\prime _\sharp \) is

$$\begin{aligned} Q(X)=\tfrac{2}{s}\, B(X,\mathfrak {R}(E_+))\mathfrak {R}(E_+)+ \tfrac{2}{s}\, B(X,\mathfrak {I}(E_+))\mathfrak {I}(E_+). \end{aligned}$$

In particular,

$$\begin{aligned} Q(H_\beta )=\tfrac{2}{s}\, \beta (\mathfrak {R}(E_+))\mathfrak {R}(E_+)+ \tfrac{2}{s}\, \beta (\mathfrak {I}(E_+)) \mathfrak {I}(E_+) \end{aligned}$$
(B.1)

for any \(\beta \in \Delta ^\prime \). Note that \(\beta (\mathfrak {R}(E_+))\) and \(\beta (\mathfrak {I}(E_+))\) are both real. It follows that the vector \(Q(H_\beta )\) in the plane Y—under the identification of \((2/s)^{\frac{1}{2}}\mathfrak {R}(E_+),(2/s)^{\frac{1}{2}}\mathfrak {I}(E_+)\) in Y with \(1,\sqrt{-1}\) in \({\mathbb {C}}\)—corresponds to the complex number \((2/s)^{\frac{1}{2}}\beta (\mathfrak {R}(E_+))+ {\scriptstyle \sqrt{-1}}\, (2/s)^{\frac{1}{2}}\beta (\mathfrak {I}(E_+))\), and this is just \((2/s)^{\frac{1}{2}}\beta (E_+)\). \(\square \)

In the proposition and its proof we have assumed that \(E_+=\sum _{i=0}^l \sqrt{q_i} e_{\alpha _i}\). In the general case \(E_+=\sum _{i=0}^l c^+_i e_{\alpha _i}\), (B.1) becomes

$$\begin{aligned} Q(H_\beta )=\tfrac{2t}{s}\, \beta (\mathfrak {R}(E_+))\mathfrak {R}(E_+)+ \tfrac{2t}{s}\, \beta (\mathfrak {I}(E_+)) \mathfrak {I}(E_+) \end{aligned}$$

where \(t\in {\mathbb {C}}^*\) is defined by \(t( \mathfrak {R}(E_+)-\sqrt{-1}\mathfrak {I}(E_+) ) = \sum _{i=0}^l ({q_i}/{c^+_i}) e_{-\alpha _i}\) [37, Theorem 1.11]. From this we obtain \(Q(H_{\beta })=((2t)/s)^{\frac{1}{2}}\beta (E_+)\). Thus the diagram of 2s singular directions, together with the points \(\beta (E_+)\) (or \(\beta (-E_+)\), as in Sect. 4) marked with their corresponding roots \(\beta \in \Delta ^\prime \), is—up to scaling and rotation—simply the Coxeter Plane. In particular, using the fact that the symmetry group of the Coxeter Plane is a dihedral group of order 2s, we see that Theorem 4.6 holds for any such \({\mathfrak {g}}\), not just for the classical Lie algebras.

Our diagram of singular directions may in fact be regarded an “enhanced Coxeter Plane”, because of the additional choice of \(E_+\), which fixes a Cartan subalgebra in apposition and a particular Coxeter element. This facilitates the assignment of a root \(\beta \) to a ray in the Coxeter Plane: one simply takes the ray through the point \(\beta (E_+)\). Having made this assignment, the Coxeter element acts on the roots by clockwise rotation through \(2\pi /s\). There are 2s systems of positive roots corresponding to this Coxeter element (in the sense of Proposition 5.4), given by the 2s choices of “positive sectors”. The head and tail of a positive sector give the associated simple roots. The roots on any two consecutive rays generate all the roots. While this information may be implicit in the (usual) Coxeter Plane, the choice of \(E_+\) makes it explicit.

In the other direction, as our diagram of singular directions arises independently of Lie theory from the Stokes data of the meromorphic connection \(d+{\hat{\alpha }}\) (or \(d+{\hat{\omega }}\)), the possibility of reproving some classical results on root systems arises. This would be very much in the spirit of [4], where the Stokes data of a meromorphic connection (similar to \(d+{\hat{\omega }}\)) was used rather unexpectedly to establish a result in symplectic geometry, the Ginzburg-Weinstein isomorphism.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guest, M.A., Ho, NK. Kostant, Steinberg, and the Stokes matrices of the tt*-Toda equations. Sel. Math. New Ser. 25, 50 (2019). https://doi.org/10.1007/s00029-019-0494-7

Download citation

  • Published:

  • DOI: https://doi.org/10.1007/s00029-019-0494-7

Mathematics Subject Classification

Navigation