Skip to main content
Log in

On the Riemann-Hilbert Problem for a q-Difference Painlevé Equation

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

Abstract

A Riemann-Hilbert problem for a q-difference Painlevé equation, known as \(q{\text {P}}_{{\text {IV}}}\), is shown to be solvable. This yields a bijective correspondence between the transcendental solutions of \(q{\text {P}}_{{\text {IV}}}\) and corresponding data on an associated q-monodromy surface. We also construct the moduli space of \(q{\text {P}}_{{\text {IV}}}\) explicitly.

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. This equation has alternative names in the literature and is also referred to as \(q{\text {P}}_{{\text {IV}}}(A_5^{(1)})\), for its initial value space, or \(q{\text {P}}_{{\text {IV}}}\bigl ((A_2+A_1)^{(1)}\bigr )\), for its symmetry group – see [22].

References

  1. Adams, C.R.: On the linear ordinary \(q\)-difference equation. Ann. Math. 30(1/4), 195–205 (1928)

  2. Birkhoff, G.D.: The generalized Riemann problem for linear differential equations and the allied problems for linear difference and \(q\)-difference equations. Proc. Am. Acad. Arts Sci. 49, 521–568 (1913)

    Article  Google Scholar 

  3. Borodin, A.: Isomonodromy transformations of linear systems of difference equations. Ann. Math. 2(160), 1141–1182 (2005)

    MathSciNet  MATH  Google Scholar 

  4. Buckingham, R.: Large-degree asymptotics of rational Painlevé-IV functions associated to generalized Hermite polynomials. Int. Math. Res. Not. (2018)

  5. Carmichael, R.D.: The general theory of linear \(q\)-difference equations. Am. J. Math. 34, 147–168 (1912)

    Article  MathSciNet  Google Scholar 

  6. Chekhov, L., Mazzocco, M., Rubtsov, V.: Quantised Painlevé monodromy manifolds, Sklyanin and Calabi-Yau algebras. Adv. Math. 376, 107442 (2021)

    Article  MathSciNet  Google Scholar 

  7. Chekhov, L., Mazzocco, M., Rubtsov, V.: Painlevé monodromy manifolds, decorated character varieties, and cluster algebras. Int. Math. Res. Not. IMRN 24, 7639–7691 (2017)

    MATH  Google Scholar 

  8. Deift, P.A.: Orthogonal polynomials and random matrices: a Riemann-Hilbert approach, 3, Courant Lecture Notes in Mathematics, New York University, Courant Institute of Mathematical Sciences, New York and American Mathematical Society, Providence, RI (1999)

  9. Deift, P.A., Zhou, X.: A steepest descent method for oscillatory Riemann-Hilbert problems. Asymptotics for the MKdV equation. Ann. Math. 2(137), 295–368 (1993)

    Article  MathSciNet  Google Scholar 

  10. Fokas,A., Its, A., Kapaev, A., Novokshenov, A.: Painlevé transcendents, the Riemann-Hilbert approach. Math. Surv. Monograph. Am. Math. Soc. 128 (2006)

  11. Fokas, A., Its, A., Kitaev, A.: Discrete Painlevé equations and their appearance in quantum gravity. Commun. Math. Phys. 142, 313–344 (1991)

    Article  ADS  Google Scholar 

  12. Fokas, A., Muğan, U., Zhou, X.: On the solvability of Painlevé \({\rm I},\;{\rm III}\) and \({\rm V}\). Inverse Prob. 8, 757–785 (1992)

    Article  ADS  Google Scholar 

  13. Gardner, C.S., Greene, J.M., Kruskal, M.D., Miura, R.M.: Method for solving the Korteweg-deVries equation. Phys. Rev. Lett. 19, 1095–1097 (1967)

    Article  ADS  Google Scholar 

  14. Its, A.R., Kitaev, A.V., Fokas, A.S.: Matrix models of two-dimensional quantum gravity and isomonodromic solutions of “discrete Painlevé” equations. J. Math. Sci. 73, 415–429 (1995)

    Article  MathSciNet  Google Scholar 

  15. Its, A., Kitaev, A., Fokas, A.: An isomonodromy approach to the theory of two-dimensional quantum gravity. Uspekhi Mat. Nauk 45, 135–136 (1990)

    MathSciNet  Google Scholar 

  16. Jimbo, M., Miwa, T., Mori, T., Sato, M.: Density matrix of an impenetrable Bose gas and the fifth Painlevé transcendent. Phys. D 1, 80–158 (1980)

  17. Jimbo, M., Miwa, T.: Monodromy preserving deformation of linear ordinary differential equations with rational coefficients. II. Phys. D 2, 407–448 (1981)

    Article  MathSciNet  Google Scholar 

  18. Jimbo, M., Miwa, T.: Monodromy preserving deformation of linear ordinary differential equations with rational coefficients. III. Phys. D, 4, 26–46 (1981/82)

  19. Jimbo, M., Miwa, T., Ueno, K.: Monodromy preserving deformation of linear ordinary differential equations with rational coefficients. I. General theory and \(\tau \)-function. Phys. D 2, 306–352 (1981)

    Article  MathSciNet  Google Scholar 

  20. Jimbo, M., Nagoya, H., Sakai, H.: CFT approach to the \(q\)-Painlevé VI equation. J. Integr. Syst. 2, (2017)

  21. Jimbo, M., Sakai, H.: A \(q\)-analogue of the sixth Painlevé equation. Lett. Math. Phys. 38, 145–154 (1996)

    Article  ADS  MathSciNet  Google Scholar 

  22. Joshi, N.: Discrete Painlevé Equations. CBMS Regional Conference Series in Mathematics, 131. American Mathematical Society. Providence, Rhode Island (2019)

  23. Joshi, N., Nakazono, N.: Lax pairs of discrete Painlevé equations: \((A_2+A_1)^{(1)}\) case. Proc. Roy Soc. A. 472, 20160696 (2016)

    Article  ADS  Google Scholar 

  24. Joshi, N., Roffelsen, P.: Analytic solutions of \(q\)-\(P(A_1)\) near its critical points. Nonlinearity 29, 3696 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  25. Kajiwara, K., Noumi, M., Yamada, Y.: Geometric aspects of Painlevé equations. J. Phys. A Math. Theor. 50, 073001 (2017)

  26. Kapaev, A.A.: Global asymptotics of the fourth Painlevé transcendent. (1996), ftp://www.pdmi.ras.ru/pub/publicat/preprint/1996/06-96.ps.gz

  27. Kapaev, A.A.: Connection formulae for degenerated asymptotic solutions of the fourth Painlevé equation. (1998), arXiv:solv-int/9805011 [nlin.SI]

  28. Mano, T.: Asymptotic behaviour around a boundary point of the \(q\)-Painlevé VI equation and its connection problem. Nonlinearity 23, 1585–1608 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  29. Masoero, D., Roffelsen, P.: Poles of Painlevé IV Rationals and their Distribution. SIGMA Symmetry Integrability Geom. Methods Appl., 14, Paper No. 002, 49, (2018)

  30. Masoero, D., Roffelsen, P.: Roots of generalised Hermite polynomials when both parameters are large. ArXiv e-prints, arXiv:1907.08552, (2019)

  31. Murata, M.: Lax forms of the \(q\)-Painlevé equations. J. Phys. A: Math. Theor. 42, 115201 (2009)

    Article  ADS  Google Scholar 

  32. Ormerod, C.M., Witte, N.S., Forrester, P.J.: Connection preserving deformations and q-semi-classical orthogonal polynomials. Nonlinearity 24, 2405 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  33. Rains, E.M.: An isomonodromy interpretation of the hypergeometric solution of the elliptic Painlevé equation (and generalizations). SIGMA Symmetry Integrability Geom. Methods Appl., 7, Paper 088, 24 (2011)

  34. Ramis, J.P., Sauloy, J., Zhang, C.: Local analytic classification of \(q\)-difference equations. Astérisque 355 (2013)

  35. Roffelsen, P.: On the global asymptotic analysis of a \(q\)-discrete Painlevé equation. PhD thesis, The University of Sydney, https://ses.library.usyd.edu.au/handle/2123/16601, (2017)

  36. Sakai, H.: Rational surfaces associated with affine root systems and geometry of the Painlevé equations. Commun. Math. Phys. 220(1), 165–229 (2001)

    Article  ADS  Google Scholar 

  37. Sauloy, J.: Galois theory of \(q\)-difference equations: the “analytical” approach. Differential equations and the Stokes phenomenon, pp. 277–292, World Sci. Publ., River Edge, NJ, (2002)

  38. Shohat, J.: A differential equation for orthogonal polynomials. Duke Math. J. 5, 401–417 (1939)

    Article  MathSciNet  Google Scholar 

  39. van der Put, M., Singer,M.F.: Galois theory of difference equations. 1666, Lecture Notes in Mathematics. Springer, Berlin (1997)

  40. van der Put, M., Saito, M.: Moduli spaces for linear differential equations and the Painlevé equations. Annales de l’Institut Fourier 59, 2611–2667 (2009)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nalini Joshi.

Additional information

Communicated by K. Johansson.

Publisher's Note

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

N. Joshi: Her research was supported by an Australian Research Council Georgina Sweet Laureate Fellowship #FL120100094 and Discovery Projects #DP130100967 and #DP200100210.

P. Roffelsen: PR acknowledges the support of the H2020-MSCA-RISE-2017 PROJECT No. 778010 IPADEGAN.

Appendices

Appendix A: A Birational Transformation and Singularities

Define

$$\begin{aligned} g_1&=qf_2^{-1}+a_0a_2f_1^{-1}f_2^{-1}+a_0f_1^{-1}, \end{aligned}$$
(A.1a)
$$\begin{aligned} g_2&=qf_2+a_0a_2f_1f_2+a_0f_1,\end{aligned}$$
(A.1b)
$$\begin{aligned} g_3&=qa_0a_2f_1+qa_0f_1f_2^{-1}+a_0^2a_2f_2^{-1},\end{aligned}$$
(A.1c)
$$\begin{aligned} g_4&=qa_0a_2f_1^{-1}+qa_0f_2f_1^{-1}+a_0^2a_2f_2, \end{aligned}$$
(A.1d)

then \(g=(g_1,g_2,g_3,g_4)\) satisfies the algebraic equations (2.9) and the rational inverse of (A.1) is given by

$$\begin{aligned} f_1&=\frac{a_0^2+g_3}{a_0(qa_2+g_1)}=\frac{a_0(qa_2+g_2)}{a_0^2+g_4},\end{aligned}$$
(A.2a)
$$\begin{aligned} f_2&=\frac{q^2+g_4}{a_0a_2(a_0+a_1g_1)}=\frac{a_0a_2(a_0+a_1g_2)}{q^2+g_3}. \end{aligned}$$
(A.2b)

We denote the algebraic surface obtained by cutting \(\{g\in {\mathbb {C}}^4\}\) with respect to (2.9) by \({\mathcal {G}}(a)\). The f and g variables are bi-rationally equivalent, and in particular \(q{\text {P}}_{{\text {IV}}}(a)\) induces the time-evolution given by Eq. (2.10) on \({\mathcal {G}}(a)\).

While the forward iteration of Eq. (2.10) is singular on \({\mathcal {G}}(a)\), only when \(g_3=0\), we show its continuation is possible by means of singularity confinement. It is also possible to regularize these singularities by lifting to the initial value space \((A_2+A_1)^{(1)}\) following Sakai [36].

Namely, if \(g_3=0\), then \(\overline{g}\) and \(\overline{\overline{g}}\) do not exist whereas \(\overline{\overline{\overline{g}}}\) does and is given explicitly by

$$\begin{aligned} \overline{\overline{\overline{g}}}_1=&\frac{(1-q^3t^2)g_1+(1-q^2)g_2}{1-q^5t^2},\end{aligned}$$
(A.3a)
$$\begin{aligned} \overline{\overline{\overline{g}}}_2=&\frac{qt^2(1-q^2)g_1+(1-q^3t^2)g_2}{1-q^5t^2},\end{aligned}$$
(A.3b)
$$\begin{aligned} \overline{\overline{\overline{g}}}_3=&g_4+(q^{-2}-1)g_1g_2+q^{-4}(q^2-1)g_1^2+\frac{(1-q^2)(g_1-q^2g_2)((2-q^2)g_1-q^2g_2)}{q^4(1-q^5t^2)}\nonumber \\&+\frac{(1-q^2)^2(g_1-q^2g_2)^2}{q^4(1-q^5t^2)^2},\end{aligned}$$
(A.3c)
$$\begin{aligned} \overline{\overline{\overline{g}}}_4=&0. \end{aligned}$$
(A.3d)

Similarly the inverse time-evolution is singular only when \(g_4=0\), in which case the first and second inverse iterates do not exist, whereas the third one does. We say that g(t) is singular at \(t_0\) when it does not exist at \(t=t_0\). The continuation formulae (A.3) of \(q{\text {P}}_{{\text {IV}}}^{\text {mod}}(a)\) can be obtained by means of direct calculation.

Considering the forward iteration, \(\overline{g}\) is ill-defined if and only if \(g_3=0\). So let us take any \(g^*\in {\mathcal {G}}(a)\) with \(g_3^*=0\) and perturb around it within \({\mathcal {G}}(a)\), setting

$$\begin{aligned} g_1=g_1^*+{\mathcal {O}}(\epsilon ),\quad g_2=g_2^*+{\mathcal {O}}(\epsilon ),\quad g_3=\epsilon +{\mathcal {O}}(\epsilon ^2),\quad g_4=g_4^*+{\mathcal {O}}(\epsilon ), \end{aligned}$$

in particular \(g=g^*+{\mathcal {O}}(\epsilon )\), as \(\epsilon \rightarrow 0\). Then direct calculation gives

$$\begin{aligned} \overline{g}_1&=a_0^2a_2(qt^2-1)t^{-2}\epsilon ^{-1}+{\mathcal {O}}(1),\\ \overline{g}_2&=-a_0^2a_2(qt^2-1)\epsilon ^{-1}+{\mathcal {O}}(1),\\ \overline{g}_3&=a_0^4a_2^2q(qt^2-1)^2t^{-2}\epsilon ^{-2}+{\mathcal {O}}(\epsilon ^{-1}),\\ \overline{g}_4&={\mathcal {O}}(\epsilon ), \end{aligned}$$

which diverges, as \(\epsilon \rightarrow 0\). Similarly

$$\begin{aligned} \overline{\overline{g}}_1&=-a_0^2a_2(qt^2-1)q^{-2}t^{-2}\epsilon ^{-1}+{\mathcal {O}}(1),\\ \overline{\overline{g}}_2&=a_0^2a_2q^3(qt^2-1)\epsilon ^{-1}+{\mathcal {O}}(1),\\ \overline{\overline{g}}_3&={\mathcal {O}}(\epsilon ),\\ \overline{\overline{g}}_4&=-a_0^4a_2^2q(qt^2-1)^2t^{-2}\epsilon ^{-2}+{\mathcal {O}}(\epsilon ^{-1}), \end{aligned}$$

which diverges, as \(\epsilon \rightarrow 0\). However, upon calculating the third iteration, we find

$$\begin{aligned} \overline{\overline{\overline{g}}}_1=&\frac{(1-q^3t^2)g_1^*+(1-q^2)g_2^*}{1-q^5t^2}+{\mathcal {O}}(\epsilon ),\\ \overline{\overline{\overline{g}}}_2=&\frac{qt^2(1-q^2)g_1^*+(1-q^3t^2)g_2^*}{1-q^5t^2}+{\mathcal {O}}(\epsilon ),\\ \overline{\overline{\overline{g}}}_3=&g_4^*+(q^{-2}-1)g_1^*g_2^*+q^{-4}(q^2-1)(g_1^*)^2\\&+\frac{(1-q^2)(g_1^*-q^2g_2^*)((2-q^2)g_1^*-q^2g_2^*)}{q^4(1-q^5t^2)} +\frac{(1-q^2)^2(g_1^*-q^2g_2^*)^2}{q^4(1-q^5t^2)^2}+{\mathcal {O}}(\epsilon ),\\ \overline{\overline{\overline{g}}}_4=&{\mathcal {O}}(\epsilon ). \end{aligned}$$

which converges to (A.3), as \(\epsilon \rightarrow 0\). We conclude that the singularity is confined within three iterations. The singularity analysis of the inverse time evolution follows by similar arguments.

Appendix B: Cubic Surface Calculations

Recall the definition of the determinants \(\Delta _1\) and \(\Delta _2\) in Eqs. (5.9) and (5.1). Each of these determinants defines a cubic equation in the variables \(\rho _{1,2,3}^{x,y}\). The aim of this section is to show that these cubics are proportional to each other

$$\begin{aligned} \Delta _2=-\lambda _0^3 \Delta _1, \end{aligned}$$
(B.1)

and that they are proportional to the cubic defined in equation (2.16),

$$\begin{aligned} \Delta _2=\frac{1}{a_1^2a_2^2}\theta _q(a_0,a_1,a_2)\theta _q(-\lambda _0)^2T_{hom}(\rho _1^x,\rho _1^y,\rho _2^x,\rho _2^y,\rho _3^x,\rho _3^y;\lambda _0,a). \end{aligned}$$
(B.2)

Firstly, we derive Eq. (B.1). To this end, let us note that we may alternatively write the functions \(v_{1,2,3}\) as

$$\begin{aligned} v_1(z)&=\frac{z}{\lambda _0x_1}\theta _q(z/x_2,z/x_3,-x_1/z\lambda _0),\\ v_2(z)&=\frac{z}{\lambda _0x_2}\theta _q(z/x_1,z/x_3,-x_2/z\lambda _0),\\ v_3(z)&=\frac{z}{\lambda _0x_3}\theta _q(z/x_1,z/x_2,-x_3/z\lambda _0), \end{aligned}$$

due to the symmetries (1.6) of the q-theta function.

Using these alternative expressions for \(v_{1,2,3}\) and expanding the difference of both sides of equation (B.1), i.e.

$$\begin{aligned} \Delta =\Delta _2+\lambda _0^3 \Delta _1, \end{aligned}$$

in terms of the variables \(\rho _{1,2,3}^{x,y}\), we find that all terms cancel except for

$$\begin{aligned} \Delta = \rho _1^x\rho _2^x\rho _3^x(U_1-U_2)(V_1-V_2), \end{aligned}$$

where

$$\begin{aligned} U_1&=\theta _q\left( -\frac{x_1}{x_2},-\frac{x_2}{x_3},-\frac{x_3}{x_1}\right) ,&V_1&=\theta _q\left( \frac{x_1}{x_2}\lambda ,\frac{x_2}{x_3}\lambda ,\frac{x_3}{x_1}\lambda \right) ,\\ U_2&=\theta _q\left( -\frac{x_2}{x_1},-\frac{x_3}{x_2},-\frac{x_1}{x_3}\right) ,&V_2&=\theta _q\left( \frac{x_2}{x_1}\lambda ,\frac{x_3}{x_2}\lambda ,\frac{x_1}{x_3}\lambda \right) . \end{aligned}$$

However, it follows directly from the symmetries (1.6) of the q-theta function that \(U_1=U_2\) and thus \(\Delta =0\), which proves equality (B.1).

Next, we derive Eq. (B.2). To this end, we first expand \(\Delta _2\) in terms of the variables \(\rho _{1,2,3}^{x,y}\), yielding the cubic

$$\begin{aligned} \Delta _2=&\,\delta _{123}\rho _1^x\rho _2^x\rho _3^x +\delta _{23}\rho _1^y\rho _2^x\rho _3^x +\delta _{13}\rho _1^x\rho _2^y\rho _3^x +\delta _{12}\rho _1^x\rho _2^x\rho _3^y\nonumber \\&+\delta _{1}\rho _1^x\rho _2^y\rho _3^y +\delta _{2}\rho _1^y\rho _2^x\rho _3^y +\delta _{3}\rho _1^y\rho _2^y\rho _3^x +\delta _{0}\rho _1^y\rho _2^y\rho _3^y, \end{aligned}$$
(B.3)

with coefficients \(\delta _0,\delta _1,\delta _2\) and \(\delta _3\) given by

$$\begin{aligned} \delta _0&=\theta _q\left( +\frac{x_1}{x_2},+\frac{x_2}{x_1},+\frac{x_1}{x_3},+\frac{x_3}{x_1},+\frac{x_2}{x_3},+\frac{x_3}{x_2}\right) \theta _q(-\lambda _0)^3,\\ \delta _1&=\theta _q\left( -\frac{x_1}{x_2},+\frac{x_2}{x_1},-\frac{x_1}{x_3},+\frac{x_3}{x_1},+\frac{x_2}{x_3},+\frac{x_3}{x_2}\right) \theta _q(-\lambda _0)^2\theta _q(\lambda _0),\\ \delta _2&=\theta _q\left( +\frac{x_1}{x_2},-\frac{x_2}{x_1},+\frac{x_1}{x_3},+\frac{x_3}{x_1},-\frac{x_2}{x_3},+\frac{x_3}{x_2}\right) \theta _q(-\lambda _0)^2\theta _q(\lambda _0),\\ \delta _3&=\theta _q\left( +\frac{x_1}{x_2},+\frac{x_2}{x_1},+\frac{x_1}{x_3},-\frac{x_3}{x_1},+\frac{x_2}{x_3},-\frac{x_3}{x_2}\right) \theta _q(-\lambda _0)^2\theta _q(\lambda _0), \end{aligned}$$

\(\delta _{23},\delta _{13}\) and \(\delta _{12}\) given by

$$\begin{aligned} \delta _{23}&=\theta _q\left( +\frac{x_1}{x_2},-\frac{x_2}{x_1},+\frac{x_1}{x_3},-\frac{x_3}{x_1},-\lambda _0\right) h(\lambda _0;x_2,x_3),\\ \delta _{13}&=\theta _q\left( +\frac{x_1}{x_2},-\frac{x_2}{x_1},+\frac{x_2}{x_3},-\frac{x_3}{x_2},-\lambda _0\right) h(\lambda _0;x_1,x_3),\\ \delta _{12}&=\theta _q\left( -\frac{x_1}{x_3},+\frac{x_3}{x_1},+\frac{x_3}{x_2},-\frac{x_2}{x_3},-\lambda _0\right) h(\lambda _0;x_1,x_2), \end{aligned}$$

where

$$\begin{aligned} h(\lambda ;z_1,z_2):=\theta _q(-\frac{z_1}{z_2},-\frac{z_2}{z_1})\theta _q(\lambda )^2-\theta _q(-1)^2\theta _q(-\frac{z_1}{z_2}\lambda ,-\frac{z_2}{z_1}\lambda ), \end{aligned}$$

and the coefficient \(\delta _{123}\) given by

$$\begin{aligned} \delta _{123}=&H(\lambda _0), \end{aligned}$$

where

$$\begin{aligned} H(\lambda ):=&+\theta _q\left( -\frac{x_1}{x_2},-\frac{x_2}{x_1},-\frac{x_1}{x_3},-\frac{x_3}{x_1},-\frac{x_2}{x_3},-\frac{x_3}{x_2}\right) \theta _q(\lambda )^3\\&-\theta _q\left( \lambda \frac{x_1}{x_2},\lambda \frac{x_2}{x_1},-\frac{x_1}{x_3},-\frac{x_3}{x_1},-\frac{x_2}{x_3},-\frac{x_3}{x_2}\right) \theta _q(\lambda )\theta _q(-1)^2\\&-\theta _q\left( -\frac{x_1}{x_2},-\frac{x_2}{x_1},\lambda \frac{x_1}{x_3},\lambda \frac{x_3}{x_1},-\frac{x_2}{x_3},-\frac{x_3}{x_2}\right) \theta _q(\lambda )\theta _q(-1)^2\\&-\theta _q\left( -\frac{x_1}{x_2},-\frac{x_2}{x_1},-\frac{x_1}{x_3},-\frac{x_3}{x_1},\lambda \frac{x_2}{x_3},\lambda \frac{x_3}{x_2}\right) \theta _q(\lambda )\theta _q(-1)^2\\&+\theta _q\left( \lambda \frac{x_1}{x_2},-\frac{x_2}{x_1},-\frac{x_1}{x_3},\lambda \frac{x_3}{x_1},\lambda \frac{x_2}{x_3},-\frac{x_3}{x_2}\right) \theta _q(-1)^3\\&+\theta _q\left( -\frac{x_1}{x_2},\lambda \frac{x_2}{x_1},\lambda \frac{x_1}{x_3},-\frac{x_3}{x_1},-\frac{x_2}{x_3},\lambda \frac{x_3}{x_2}\right) \theta _q(-1)^3. \end{aligned}$$

In order to derive Eq. (B.2), we factorise the functions \(h(\lambda ;z_1,z_2)\) and \(H(\lambda )\) into simple factors of q-theta functions. We start with \(h(\lambda ;z_1,z_2)\). Note that this function is an element of the vector space \(V_2(1)\), namely, it is analytic in \(\lambda \) on \({\mathbb {C}}^*\) and satisfies

$$\begin{aligned} h(q\lambda ;z_1,z_2)=\lambda ^{-2}h(\lambda ;z_1,z_2). \end{aligned}$$

Furthermore, it is easy to see that \(h(-1;z_1,z_2)=0\), hence, by Lemma 5.1,

$$\begin{aligned} h(\lambda ;z_1,z_2)=c(z_1,z_2)\theta _q(-\lambda )^2, \end{aligned}$$
(B.4)

for some yet to be determined coefficient \(c(z_1,z_2)\) which is independent of \(\lambda \). Then, by evaluating equation (B.4) at \(\lambda =1\), we obtain

$$\begin{aligned} c(z_1,z_2)=-\theta _q\left( \frac{z_1}{z_2},\frac{z_2}{z_1}\right) \end{aligned}$$

and thus

$$\begin{aligned} h(\lambda ;z_1,z_2)=-\theta _q\left( \frac{z_1}{z_2},\frac{z_2}{z_1}\right) \theta _q(-\lambda )^2. \end{aligned}$$
(B.5)

We follow the same procedure to compute a factorisation of \(H(\lambda )\). We note that it is an element of the vector space \(V_3(-1)\), i.e. it is analytic on \({\mathbb {C}}^*\) and

$$\begin{aligned} H(q\lambda )=-\lambda ^{-3}H(\lambda ). \end{aligned}$$

Furthermore, it is easy to check by direct calculation that \(H(1)=H(-1)=0\), and hence

$$\begin{aligned} H(\lambda )=c_H\theta _q(\lambda )\theta _q(-\lambda )^2, \end{aligned}$$
(B.6)

for some yet to be determined coefficient \(c_H\), which is independent of \(\lambda \). To determine this constant, we evaluate both sides of Eq. (B.6) at \(\lambda =x_1\). By means of analogous calculations as above, we can simplify \(H(x_1)\) to obtain

$$\begin{aligned} H(x_1)=\frac{\theta _q(x_1)^3\theta _q(-x_1)^2\theta _q(x_2)^2\theta _q(\frac{x_2}{x_1})^2}{q^2x_1^2x_2^4}, \end{aligned}$$

leading to

$$\begin{aligned} H(\lambda )=\frac{\theta _q(x_1)^2\theta _q(x_2)^2\theta _q(\frac{x_2}{x_1})^2}{q^2x_1^2x_2^4}\theta _q(\lambda )\theta _q(-\lambda )^2. \end{aligned}$$
(B.7)

Now that we have explicit factorisations of the functions \(h(\lambda ;z_1,z_2)\) and \(H(\lambda )\) (respectively given by Eqs. (B.5) and (B.7)), and thus factorised expressions for the coefficients of the cubic (B.3), Eq. (B.2) follows immediately, upon using the symmetries (1.6) of the q-theta function.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Joshi, N., Roffelsen, P. On the Riemann-Hilbert Problem for a q-Difference Painlevé Equation. Commun. Math. Phys. 384, 549–585 (2021). https://doi.org/10.1007/s00220-021-04024-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-021-04024-y

Navigation