Skip to main content
Log in

The Z-invariant Ising model via dimers

  • Published:
Probability Theory and Related Fields Aims and scope Submit manuscript

Abstract

The Z-invariant Ising model (Baxter in Philos Trans R Soc Lond A Math Phys Eng Sci 289(1359):315–346, 1978) is defined on an isoradial graph and has coupling constants depending on an elliptic parameter k. When \(k=0\) the model is critical, and as k varies the whole range of temperatures is covered. In this paper we study the corresponding dimer model on the Fisher graph, thus extending our papers (Boutillier and de Tilière in Probab Theory Relat Fields 147:379–413, 2010; Commun Math Phys 301(2):473–516, 2011) to the fullZ-invariant case. One of our main results is an explicit, local formula for the inverse of the Kasteleyn operator. Its most remarkable feature is that it is an elliptic generalization of Boutillier and de Tilière (2011): it involves a local function and the massive discrete exponential function introduced in Boutillier et al. (Invent Math 208(1):109–189, 2017). This shows in particular that Z-invariance, and not criticality, is at the heart of obtaining local expressions. We then compute asymptotics and deduce an explicit, local expression for a natural Gibbs measure. We prove a local formula for the Ising model free energy. We also prove that this free energy is equal, up to constants, to that of the Z-invariant spanning forests of Boutillier et al.  (2017), and deduce that the two models have the same second order phase transition in k. Next, we prove a self-duality relation for this model, extending a result of Baxter to all isoradial graphs. In the last part we prove explicit, local expressions for the dimer model on a bipartite graph corresponding to the XOR version of this Z-invariant Ising model.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Notes

  1. In Equation (7.8.4), Baxter actually uses the complementary parameter \(k'=\sqrt{1-k^2}\) and the parametrization, \(\sinh (2\mathsf {J}_e)=-i{{\mathrm{sn }}}(i\theta _e\vert k')\). The latter is equal to \({{\mathrm{sc}}}(\theta _e\vert k)\) by [40, (2.6.12)].

  2. When indicating sectors on the circle, the convention we adopt is that \((\alpha ,\beta )\) represents the sector where the horizontal coordinate increases from \(\alpha \) to \(\beta \).

  3. Such an orientation always exists by Theorem 3.1 of [14], using that the number of vertices of the fundamental domain of \({\mathsf {G}}^{\mathrm F}\) is even.

  4. If on the torus, there is a unique train-track with a given homology class, making it move across the torus through every pairwise intersection once by Y\(\,\Delta \) moves yields the same rhombic graph, but with the role of primal and dual vertices exchanged. If there are several of them, then moving them all to put the first one instead of the second one, the second instead of the third, etc., yields the result. Note though that this is not possible in the case when there are just two different homology classes, like in the case of the square lattice.

References

  1. Abramowitz, M., Stegun, I.A.: Handbook of mathematical functions with formulas, graphs, and mathematical tables. National Bureau of Standards Applied Mathematics Series, vol. 55. U.S. Government Printing Office, Washington, DC (1964)

  2. Au-Yang, H., Perk, J.H.H.: Q-dependent susceptibilities in ferromagnetic quasiperiodic \(Z\)-invariant ising models. J. Stat. Phys. 127, 265–286 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  3. Baxter, R.J.: Solvable eight-vertex model on an arbitrary planar lattice. Philos. Trans. R. Soc. Lond. A Math. Phys. Eng. Sci. 289(1359), 315–346 (1978)

    Article  MathSciNet  Google Scholar 

  4. Baxter, R.J.: Free-fermion, checkerboard and \({Z}\)-invariant lattice models in statistical mechanics. Proc. R. Soc. Lond. Ser. A 404(1826), 1–33 (1986)

    Article  MathSciNet  Google Scholar 

  5. Baxter, R.J.: Exactly solved models in statistical mechanics. Academic Press Inc. [Harcourt Brace Jovanovich Publishers], London, 1989. Reprint of the 1982 original

  6. Bostan, A., Boukraa, S., Hassani, S., van Hoeij, M., Maillard, J.-M., Weil, J.-A., Zenine, N.: The Ising model: from elliptic curves to modular forms and Calabi-Yau equations. J. Phys. A 44(4), 045204, 44 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  7. Boutillier, C., de Tilière, B.: The critical \(Z\)-invariant Ising model via dimers: the periodic case. Probab. Theory Relat. Fields 147, 379–413 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  8. Boutillier, C., de Tilière, B.: The critical \(Z\)-invariant Ising model via dimers: locality property. Commun. Math. Phys. 301(2), 473–516 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  9. Boutillier, C., de Tilière, B.: Height representation of XOR-Ising loops via bipartite dimers. Electron. J. Probab. 19(80), 33 (2014)

    MathSciNet  MATH  Google Scholar 

  10. Boutillier, C., de Tilière, B., Raschel, K.: The \(Z\)-invariant massive Laplacian on isoradial graphs. Invent. Math. 208(1), 109–189 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  11. Chan, Y., Guttmann, A.J., Nickel, B.G., Perk, J.H.H.: The Ising susceptibility scaling function. J. Stat. Phys. 145(3), 549–590 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  12. Chelkak, D., Smirnov, S.: Universality in the 2D Ising model and conformal invariance of fermionic observables. Inv. Math. 189, 515–580 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  13. Cimasoni, D., Duminil-Copin, H.: The critical temperature for the Ising model on planar doubly periodic graphs. Electron. J. Probab. 18(44), 1–18 (2013)

    MathSciNet  MATH  Google Scholar 

  14. Cimasoni, D., Reshetikhin, N.: Dimers on surface graphs and spin structures. I. Commun. Math. Phys. 275(1), 187–208 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  15. Cohn, H., Kenyon, R., Propp, J.: A variational principle for domino tilings. J. Am. Math. Soc. 14(2), 297–346 (2001). (electronic)

    Article  MathSciNet  MATH  Google Scholar 

  16. Costa-Santos, R.: Geometrical aspects of the \(Z\)-invariant Ising model. EPJ B 53(1), 85–90 (2006)

    Article  Google Scholar 

  17. de Tilière, B.: Quadri-tilings of the plane. Probab. Theory Relat. Fields 137(3–4), 487–518 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  18. de Tilière, B.: Critical Ising model and spanning trees partition functions. Ann. Inst. Henri Poincaré Probab. Stat. 52(3), 1382–1405 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  19. Dubédat, J.: Exact bosonization of the Ising model, pp. 1–35 (2011). arXiv:1112.4399

  20. Duffin, R.J.: Potential theory on a rhombic lattice. J. Comb. Theory 5, 258–272 (1968)

    Article  MathSciNet  MATH  Google Scholar 

  21. Duminil-Copin, H., Hongler, C., Nolin, P.: Connection probabilities and RSW-type bounds for the two-dimensional FK Ising model. Commun. Pure Appl. Math. 64(9), 1165–1198 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  22. Duminil-Copin, H., Li, J., Manolescu, I.: Universality for the random-cluster model on isoradial graphs (2017). arXiv:1711.02338

  23. Fan, C., Wu, F.Y.: General lattice model of phase transitions. Phys. Rev. B 2, 723–733 (1970)

    Article  Google Scholar 

  24. Fisher, M.E.: On the dimer solution of planar Ising models. J. Math. Phys. 7, 1776–1781 (1966)

    Article  Google Scholar 

  25. Grimmett, G.R., Manolescu, I.: Universality for bond percolation in two dimensions. Ann. Probab. 41(5), 3261–3283 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  26. Kadanoff, L.P., Brown, A.C.: Correlation functions on the critical lines of the Baxter and Ashkin-Teller models. Ann. Phys. 121(1–2), 318–342 (1979)

    Article  Google Scholar 

  27. Kadanoff, L.P., Wegner, F.J.: Some critical properties of the eight-vertex model. Phys. Rev. B 4, 3989–3993 (1971)

    Article  Google Scholar 

  28. Kasteleyn, P.W.: The statistics of dimers on a lattice: I. The number of dimer arrangements on a quadratic lattice. Physica 27, 1209–1225 (1961)

    Article  MATH  Google Scholar 

  29. Kasteleyn, P.W.: Graph theory and crystal physics. In: Graph Theory and Theoretical Physics. Academic Press, London, pp. 43–110 (1967)

  30. Kennelly, A.E.: The equivalence of triangles and three-pointed stars in conducting networks. Electr. World Eng. 34, 413–414 (1899)

    Google Scholar 

  31. Kenyon, R.: The Laplacian and Dirac operators on critical planar graphs. Invent. Math. 150(2), 409–439 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  32. Kenyon, R.: An introduction to the dimer model. In: School and Conference on Probability Theory, ICTP Lect. Notes, XVII. Abdus Salam Int. Cent. Theoret. Phys., Trieste, pp. 267–304 (2004) (electronic)

  33. Kenyon, R., Okounkov, A.: Planar dimers and harnack curves. Duke Math. J. 131(3), 499–524 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  34. Kenyon, R., Okounkov, A., Sheffield, S.: Dimers and amoebae. Ann. Math. (2) 163(3), 1019–1056 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  35. Kenyon, R., Schlenker, J.-M.: Rhombic embeddings of planar quad-graphs. Trans. Am. Math. Soc. 357(9), 3443–3458 (2005). (electronic)

    Article  MathSciNet  MATH  Google Scholar 

  36. Kramers, H.A., Wannier, G.H.: Statistics of the two-dimensional ferromagnet. Part I. Phys. Rev. 60(3), 252–262 (1941)

    Article  MathSciNet  MATH  Google Scholar 

  37. Kramers, H.A., Wannier, G.H.: Statistics of the two-dimensional ferromagnet. Part II. Phys. Rev. 60(3), 263–276 (1941)

    Article  MathSciNet  MATH  Google Scholar 

  38. Krieger, M.H.: Constitutions of matter. University of Chicago Press, Chicago, IL (1996). Mathematically modeling the most everyday of physical phenomena

  39. Kuperberg, G.: An exploration of the permanent-determinant method. Electron. J. Comb. 5(1), R46 (1998)

    MathSciNet  MATH  Google Scholar 

  40. Lawden, D.F.: Elliptic functions and applications. Applied Mathematical Sciences, vol. 80. Springer, New York (1989)

  41. Li, Z.: Critical temperature of periodic Ising models. Commun. Math. Phys. 315, 337–381 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  42. Lis, M.: Phase transition free regions in the Ising model via the Kac-Ward operator. Commun. Math. Phys. 331(3), 1071–1086 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  43. Maillard, J.-M., Boukraa, S.: Modular invariance in lattice statistical mechanics. Ann. Fond. Louis de Broglie 26(Special Issue 2), 287–328 (2001)

    MathSciNet  MATH  Google Scholar 

  44. Martìnez, J.R.R.: Correlation functions for the \(Z\)-invariant Ising model. Phys. Lett. A 227(3–4), 203–208 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  45. Martìnez, J.R.R.: Multi-spin correlation functions for the \(Z\)-invariant Ising model. Physica A 256(3–4), 463–484 (1998)

    Article  Google Scholar 

  46. McCoy, B., Wu, F.: The Two-Dimensional Ising Model. Harvard Univ. Press, Cambridge (1973)

    Book  MATH  Google Scholar 

  47. McCoy, B.M., Maillard, J.M.: The anisotropic Ising correlations as elliptic integrals: duality and differential equations. J. Phys. A 49(43), 434004 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  48. Mercat, C.: Discrete Riemann surfaces and the Ising model. Commun. Math. Phys. 218(1), 177–216 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  49. Mercat, C.: Exponentials form a basis of discrete holomorphic functions on a compact. Bull. Soc. Math. France 132(2), 305–326 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  50. Nienhuis, B.: Critical behavior of two-dimensional spin models and charge asymmetry in the Coulomb gas. J. Stat. Phys. 34(5–6), 731–761 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  51. Onsager, L.: Crystal statistics. I. A two-dimensional model with an order-disorder transition. Phys. Rev. 65(3–4), 117–149 (1944)

    Article  MathSciNet  MATH  Google Scholar 

  52. Perk, J.H.H., Au-Yang, H.: Yang Baxter equations. In: Françoise, J.-P., Naber, G.L., Tsun, T.S. (eds.) Encyclopedia of Mathematical Physics, pp. 465–473. Academic Press, Oxford (2006)

    Chapter  Google Scholar 

  53. Temperley, H.N.V., Fisher, M.E.: Dimer problem in statistical mechanics-an exact result. Philos. Mag. 6(68), 1061–1063 (1961)

    Article  MathSciNet  MATH  Google Scholar 

  54. Wannier, G.H.: The statistical problem in cooperative phenomena. Rev. Mod. Phys. 17(1), 50–60 (1945)

    Article  Google Scholar 

  55. Wilson, D.B.: XOR-Ising loops and the Gaussian free field (2011). arXiv: 1102.3782

  56. Wu, F.W.: Ising model with four-spin interactions. Phys. Rev. B 4, 2312–2314 (1971)

    Article  Google Scholar 

  57. Wu, F.Y., Lin, K.Y.: Staggered ice-rule vertex model—the Pfaffian solution. Phys. Rev. B 12, 419–428 (1975)

    Article  Google Scholar 

Download references

Acknowledgements

We acknowledge support from the Agence Nationale de la Recherche (Projet MAC2: ANR-10-BLAN-0123) and from the Région Centre-Val de Loire (Projet MADACA). We are grateful to the referee for his/her many insightful comments.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Cédric Boutillier.

Appendices

Appendix A: Useful identities involving elliptic functions

In this section we list required identities satisfied by elliptic functions.

1.1 Identities for Jacobi’s elliptic functions

Change of argument Jacobi’s elliptic functions satisfy various addition formulas by quarter-periods and half-periods, among which (cf. [1, Table 16.8]):

Table 2 Addition formulas by quarter-periods and half-periods, taken from [1, 16.8]

1.2 Jacobi’s epsilon, zeta and related functions

The explicit expressions of the inverse operators of Theorems 11 and 37 involve the function \(H\) defined as follows. In the remarks following Theorem 11, we also use the function \(V\) mentioned below.

For \(0<k^2<1\), let

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle H(u\vert k)=\displaystyle \frac{K'}{\pi }\Bigl \{\mathrm {E}\Bigl (\frac{u}{2}\Big \vert k\Bigr )+\frac{E'-K'}{K'}\frac{u}{2}\Bigr \},\\ \displaystyle V(u\vert k)=\displaystyle \frac{iK}{\pi }\Bigl \{\mathrm {E}\Bigl (\frac{u}{2}\Big \vert k\Bigr )-\frac{E}{K}\frac{u}{2}\Bigl \}, \end{array}\right. \end{aligned}$$
(66)

where \(\mathrm {E}\) is Jacobi’s epsilon function: \(\mathrm {E}(u)=\int _{0}^u {{\mathrm{ dn }}}^2(v\vert k)\mathrm {d}v\), see [1, 16.26.3]. For \(k^2<0\), let

$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle H(u\vert k)=H(k'u\vert k^*),\\ \displaystyle V(u\vert k)=V(k'u\vert k^*). \end{array}\right. \end{aligned}$$
(67)

Properties of these functions are presented below.

Lemma 42

The functions \(H\) and \(V\) admit jumps in the horizontal and vertical directions, respectively:

$$\begin{aligned} \left\{ \begin{array}{l} H(u+4K\vert k)-H(u\vert k)=1,\\ H(u+4i\mathfrak {R}K'\vert k)-H(u\vert k)=0, \end{array}\right. \qquad \left\{ \begin{array}{l} V(u+4K\vert k)-V(u\vert k)=0,\\ V(u+4i\mathfrak {R}K'\vert k)-V(u\vert k)=1. \end{array}\right. \end{aligned}$$
(68)

In the fundamental rectangle \([0,4K]+[0,4i\mathfrak {R}K']\), the function \(H\) (resp. \(V\)) has a simple pole, at \(2i\mathfrak {R}K'\), with residue \(\frac{2\mathfrak {R}K'}{\pi }\) (resp. \(\frac{2iK}{\pi }\)). Moreover,

$$\begin{aligned} \lim _{k\rightarrow 0}H(u\vert k)=\frac{u}{2\pi },\qquad \lim _{k\rightarrow 0}V(u\vert k)=0. \end{aligned}$$

The following addition formulas hold:

$$\begin{aligned} H(v-u\vert k)= & {} H(v\vert k)-H(u\vert k)\nonumber \\&+\frac{k^2 \mathfrak {R}K'}{\pi } \left\{ \begin{array}{ll} \displaystyle {{\mathrm{sn }}}\Bigl (\frac{u}{2}\Big \vert k\Bigr ){{\mathrm{sn }}}\Bigl (\frac{v}{2}\Big \vert k\Bigr ){{\mathrm{sn }}}\Bigl (\frac{v-u}{2}\Big \vert k\Bigr ) &{}\quad \text {if }\; 0<k^2<1,\\ \displaystyle (-{k'}^2) {{\mathrm{sd}}}\Bigl (\frac{u}{2}\Big \vert k\Bigr ){{\mathrm{sd}}}\Bigl (\frac{v}{2}\Big \vert k\Bigr ){{\mathrm{sd}}}\Bigl (\frac{v-u}{2}\Big \vert k\Bigr ) &{}\quad \text {if }\; k^2<0, \end{array}\right. \end{aligned}$$
(69)
$$\begin{aligned} V(v-u\vert k)= & {} V(v\vert k)-V(u\vert k)\nonumber \\&+\frac{ik^2 K}{\pi } \left\{ \begin{array}{ll} \displaystyle {{\mathrm{sn }}}\Bigl (\frac{u}{2}\Big \vert k\Bigr ){{\mathrm{sn }}}\Bigl (\frac{v}{2}\Big \vert k\Bigr ){{\mathrm{sn }}}\Bigl (\frac{v-u}{2}\Big \vert k\Bigr ) &{}\quad \text {if }\; 0<k^2<1,\\ \displaystyle (-{k'}^2) {{\mathrm{sd}}}\Bigl (\frac{u}{2}\Big \vert k\Bigr ){{\mathrm{sd}}}\Bigl (\frac{v}{2}\Big \vert k\Bigr ){{\mathrm{sd}}}\Bigl (\frac{v-u}{2}\Big \vert k\Bigr ) &{}\quad \text {if }\; k^2<0. \end{array}\right. \nonumber \\ \end{aligned}$$
(70)

Proof

We first prove the lemma in the case \(0<k^2<1\). All properties concerning \(H\) are proved in [10, Lemmas 44 and 45]. The statements for \(V\) follow similarly.

A slightly different proof would consist in using a reformulation of H and V in terms of Jacobi’s zeta functionZ:

$$\begin{aligned} H(u\vert k)=\frac{K'}{\pi }Z\Bigl (\frac{u}{2}\Big \vert k\Bigr )+\frac{u}{4K},\quad V(u\vert k)=\frac{iK}{\pi }Z\Bigl (\frac{u}{2}\Big \vert k\Bigr ). \end{aligned}$$
(71)

(Eq. (71) is a consequence of (66) and [1, 17.4.28 and 17.3.13].) Then we could use the numerous properties satisfied by Z (see in particular the addition formula [1, 17.4.35]) to derive Lemma 42.

In the case \(k^2<0\), we use the definition (67) of \(H\) and \(V\), allowing to transfer all properties from the case \(k^2>0\). In doing so we have to: consider \(\mathfrak {R}K'\) in (68) because in the case \(k^2<0\) the quarter-period \(K'\) is non-real, see (46), and use the transformation of the \({{\mathrm{sn }}}\) function into the \({{\mathrm{sd}}}\) one under the dual transformation, see [1, 16.10]. \(\square \)

The Laplacian operator (6) uses the function \(\mathrm {A}\), which satisfies the following properties.

Lemma 43

(Lemma 44 in [10]) The function \(\mathrm {A}(\cdot \vert k)\) is odd and satisfies the following identities:

$$\begin{aligned} \bullet&\quad \mathrm {A}(v-u\vert k)=\mathrm {A}(v\vert k)-\mathrm {A}(u\vert k)-k'{{\mathrm{sc}}}(u\vert k){{\mathrm{sc}}}(v\vert k){{\mathrm{sc}}}(v-u\vert k), \end{aligned}$$
(72)
$$\begin{aligned} \bullet&\quad \mathrm {A}(u+2K\vert k)=\mathrm {A}(u\vert k). \end{aligned}$$
(73)

Appendix B: Some explicit integral computations

We gather here computations of some contour integrals appearing in the expression of the Kasteleyn operator, in the Fisher case of Sect. 3.

1.1 An important contour integral

The following result has been used when proving Theorem 11.

Lemma 44

One has

$$\begin{aligned} \frac{1}{2i\pi } \int _\Gamma \mathsf {f}(u+2K)\mathsf {f}(u)\mathrm {d}u=-\frac{1}{k'}, \end{aligned}$$

where \(\Gamma \) is a vertical contour on \({\mathbb T}(k)\) winding once vertically on the torus and crossing the horizontal axis in the interval \([\alpha ,\alpha +2K]=\{x:\alpha \leqslant x \leqslant \alpha +2K\}\) and \(\mathsf {f}(u)= {{\mathrm{nc}}}(\frac{u-\alpha }{2})\).

On the rectangle, the contour \(\Gamma \) is supposed to cross the horizontal axis inside of the interval \([\alpha ,\alpha +2K]\). If the vertical contour crosses the horizontal axis in the other interval \([\alpha +2K,\alpha (+4K)]\), the integral is equal to \(+\frac{1}{k'}\), as it corresponds to changing \(\alpha \) into \(\alpha +2K\) and \({{\mathrm{nc}}}(\frac{u-(\alpha +4K)}{2})=-{{\mathrm{nc}}}(\frac{u-\alpha }{2})\). Note further that Lemma 44 is independent of the choice of the angle \(\alpha \) mod 4K, and the integrand \(\mathsf {f}(u+2K)\mathsf {f}(u)\) is.

Proof

First, it follows from the change of variable \(u\rightarrow u+2K\) and the above-mentioned property of \({{\mathrm{nc}}}\) that

$$\begin{aligned} \frac{1}{2i\pi } \int _\Gamma \mathsf {f}(u+2K)\mathsf {f}(u)\mathrm {d}u=-\frac{1}{2i\pi } \int _{\Gamma -2K} \mathsf {f}(u+2K)\mathsf {f}(u)\mathrm {d}u=\frac{1}{2i\pi } \int _{\widetilde{\Gamma }} \mathsf {f}(u+2K)\mathsf {f}(u)\mathrm {d}u, \end{aligned}$$

where \(\widetilde{\Gamma }\) is the contour \(\Gamma -2K\) crossed in the opposite direction of \(\Gamma \). Further, using the \(4iK'\)-periodicity of the integrand, we deduce that

$$\begin{aligned} \frac{1}{2i\pi } \int _\Gamma \mathsf {f}(u+2K)\mathsf {f}(u)\mathrm {d}u=\frac{1}{2}\frac{1}{2i\pi } \int _{\mathscr { C}} \mathsf {f}(u+2K)\mathsf {f}(u)\mathrm {d}u, \end{aligned}$$
(74)

where \(\mathscr { C}\) is the closed contour \((\Gamma -2iK')\bigcup (\widetilde{\Gamma }-2iK')\bigcup S_1\bigcup S_2\), \(S_1\) and \(S_2\) being the horizontal segments joining \((\Gamma -2iK')\) and \((\widetilde{\Gamma }-2iK')\), see Fig. 19.

Fig. 19
figure 19

The contour \(\mathscr { C}\) in (74) used in the proof of Lemma 44

The main point is that the contour integral in the right-hand side of (74) can be computed with the residue theorem: the only pole (of order 1) is at \(\alpha \) and has residue \(\frac{-2}{k'}\), see Table 2. Lemma 44 follows. \(\square \)

1.2 Inverse Kasteleyn operator at an edge

Here we compute the probability that a given edge \(e={\mathbf x}{\mathbf y}\) of the isoradial graph \(\mathsf {G}\) with rhombus half-angle \(\theta _e\) appears in the high temperature contour expansion of the Ising model. In terms of dimers, it corresponds to the probability that the corresponding edge \(\mathsf {e}=\mathsf {v}_j({\mathbf x})\mathsf {v}_\ell ({\mathbf y})=\mathsf {v}_j\mathsf {v}_\ell \) belongs to a dimer configuration of \({\mathsf {G}}^{\mathrm F}\). By Theorem 19 this probability is given by \(\mathbb {P}_{\mathrm {dimer}}(\mathsf {e})=\mathsf {K}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}\mathsf {K}^{-1}_{\mathsf {v}_{\ell },\mathsf {v}_{j}}\).

Lemma 45

One has

$$\begin{aligned} \mathsf {K}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}\mathsf {K}^{-1}_{\mathsf {v}_{\ell },\mathsf {v}_{j}}=\frac{1}{2}-\frac{1-2H(2\theta _e)}{2{{\mathrm{cn }}}\theta _e}. \end{aligned}$$

Proof

Instead of \(\mathsf {K}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}\mathsf {K}^{-1}_{\mathsf {v}_{\ell },\mathsf {v}_{j}}\) we compute \(\mathsf {K}_{\mathsf {v}_{\ell },\mathsf {v}_{j}}\mathsf {K}^{-1}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}\). Both quantities are obviously equal, but the second one happens to be more convenient when applying the results of Sect. 3.

We start from the expression of \(\mathsf {K}^{-1}\) given in (20) of Theorem 11 (note that by (19), \(C_{\mathsf {v}_{j},\mathsf {v}_{\ell }}=0\)):

$$\begin{aligned} \mathsf {K}^{-1}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}= & {} \frac{i k'}{8\pi }\oint _{\mathscr { C}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}}\mathsf {g}_{(\mathsf {v}_{j},\mathsf {v}_{\ell })}(u)H(u)\mathrm {d}u\\= & {} \frac{i k'}{8\pi }\oint _{\mathscr { C}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}}\mathsf {f}_{\mathsf {v}_{j}}(u+2K)\mathsf {f}_{\mathsf {v}_{\ell }}(u){{\mathrm{\mathsf {e}}}}_{({\mathbf x},{\mathbf y})}(u)H(u)\mathrm {d}u. \end{aligned}$$

Using the harmonicity property (16) of \(\mathsf {g}_{(\mathsf {z},\cdot )}(u)\) enables us to rewrite

$$\begin{aligned} \mathsf {K}_{\mathsf {v}_j,\mathsf {v}_\ell }\mathsf {f}_{\mathsf {v}_{\ell }}(u){{\mathrm{\mathsf {e}}}}_{({\mathbf x},{\mathbf y})}(u)= -[\mathsf {K}_{\mathsf {v}_j,{\mathsf {w}_j}}\mathsf {f}_{\mathsf {w}_j}(u)+\mathsf {K}_{\mathsf {v}_j,{\mathsf {w}_{j+1}}}\mathsf {f}_{\mathsf {w}_{j+1}}(u)]. \end{aligned}$$

By definition of the function \(\mathsf {f}_{\mathsf {v}_j}\), see (13), we thus have

$$\begin{aligned}&\mathsf {K}_{\mathsf {v}_j,\mathsf {v}_\ell }\mathsf {f}_{\mathsf {v}_{j}}(u+2K)\mathsf {f}_{\mathsf {v}_{\ell }}(u){{\mathrm{\mathsf {e}}}}_{({\mathbf x},{\mathbf y})}(u)\\&\quad =-\,[\mathsf {K}_{\mathsf {v}_j,\mathsf {w}_j}\mathsf {f}_{\mathsf {w}_j}(u+2K)+\mathsf {K}_{\mathsf {w}_{j+1},\mathsf {v}_j}\mathsf {f}_{\mathsf {w}_{j+1}}(u+2K)] \\&\qquad \times [\mathsf {K}_{\mathsf {v}_j,{\mathsf {w}_j}}\mathsf {f}_{\mathsf {w}_j}(u)+\,\mathsf {K}_{\mathsf {v}_j,{\mathsf {w}_{j+1}}}\mathsf {f}_{\mathsf {w}_{j+1}}(u)]. \end{aligned}$$

Recalling that the orientation of the triangle \((\mathsf {w}_j,\mathsf {v}_j,\mathsf {w}_{j+1})\) is admissible, we moreover have \(\mathsf {K}_{\mathsf {v}_j,{\mathsf {w}_j}}\mathsf {K}_{\mathsf {v}_j,\mathsf {w}_{j+1}}=-\mathsf {K}_{\mathsf {w}_j,\mathsf {w}_{j+1}}\), and since \(\mathsf {K}_{\mathsf {v}_j,\mathsf {v}_\ell }=-\mathsf {K}_{\mathsf {v}_\ell ,\mathsf {v}_j}\), we have

$$\begin{aligned}&\mathsf {K}_{\mathsf {v}_\ell ,\mathsf {v}_j}\mathsf {f}_{\mathsf {v}_{j}}(u+2K)\mathsf {f}_{\mathsf {v}_{\ell }}(u){{\mathrm{\mathsf {e}}}}_{({\mathbf x},{\mathbf y})}(u)\nonumber \\&\quad = \mathsf {f}_{\mathsf {w}_j}(u+2K)\mathsf {f}_{\mathsf {w}_j}(u)-\mathsf {f}_{\mathsf {w}_{j+1}}(u+2K)\mathsf {f}_{\mathsf {w}_{j+1}}(u)\nonumber \\&\qquad +\,\mathsf {K}_{\mathsf {w}_j,\mathsf {w}_{j+1}}(\mathsf {f}_{\mathsf {w}_{j+1}}(u+2K)\mathsf {f}_{\mathsf {w}_{j}}(u)-\mathsf {f}_{\mathsf {w}_j}(u+2K)\mathsf {f}_{\mathsf {w}_{j+1}}(u)). \end{aligned}$$
(75)

With (75) the quantity \(\mathsf {K}_{\mathsf {v}_{\ell },\mathsf {v}_{j}}\mathsf {K}^{-1}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}\) is a sum of four terms. The first two ones are computed thanks to Lemma 44: with the choice of contour \(\mathscr { C}_{\mathsf {v}_j,\mathsf {v}_\ell }\), we have

$$\begin{aligned} \frac{i k'}{8\pi }\oint _{\mathscr { C}_{\mathsf {v}_j,\mathsf {v}_\ell }}[\mathsf {f}_{\mathsf {w}_j}(u+2K)\mathsf {f}_{\mathsf {w}_j}(u)-\mathsf {f}_{\mathsf {w}_{j+1}}(u+2K)\mathsf {f}_{\mathsf {w}_{j+1}}(u)]H(u)\mathrm {d}u= 2\frac{i k'}{8\pi }\frac{-2\pi i}{k'}=\frac{1}{2}. \end{aligned}$$

To compute the last two terms in (75), namely

$$\begin{aligned} \frac{i k'}{8\pi }\oint _{\mathscr { C}_{\mathsf {v}_j,\mathsf {v}_\ell }}\mathsf {K}_{\mathsf {w}_j,\mathsf {w}_{j+1}}[\mathsf {f}_{\mathsf {w}_{j+1}}(u+2K)\mathsf {f}_{\mathsf {w}_{j}}(u)-\mathsf {f}_{\mathsf {w}_{j}}(u+2K)\mathsf {f}_{\mathsf {w}_{j+1}}(u)]H(u)\mathrm {d}u, \end{aligned}$$
(76)

recall that by (11)

$$\begin{aligned} \mathsf {f}_{\mathsf {w}_{j+1}}(u)= \textstyle {{\mathrm{nc}}}\Big (\frac{u-\alpha _{j+1}}{2}\Big )= \mathsf {K}_{\mathsf {w}_j,\mathsf {w}_{j+1}}{{\mathrm{nc}}}\Big (\frac{u-\alpha _j-2\theta _e}{2}\Big ), \end{aligned}$$

so that

$$\begin{aligned} \mathsf {K}_{\mathsf {w}_j,\mathsf {w}_{j+1}}\mathsf {f}_{\mathsf {w}_{j+1}}(u+2K)\mathsf {f}_{\mathsf {w}_{j}}(u)= {{\mathrm{nc}}}\Big (\frac{u+2K-\alpha _j-2\theta _e}{2}\Big ){{\mathrm{nc}}}\Big (\frac{u-\alpha _j}{2}\Big ). \end{aligned}$$

We therefore focus on the term

$$\begin{aligned} \int _{\mathscr { C}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}}{{\mathrm{nc}}}\Bigl (\frac{u+2K-\alpha _j-2\theta _e}{2} \Bigr ){{\mathrm{nc}}}\Bigl (\frac{u-\alpha _j}{2} \Bigr )H(u)\frac{\mathrm {d}u}{2\pi i}, \end{aligned}$$

in which we set, without loss of generality, \(\alpha _j=0\). This is equivalent to replace the function \(H(u)\) by \(H(u-\alpha _j)\), which is possible because both functions satisfy the same jump conditions stated in Lemma 42. There are three residues, at \(2\theta _e\), 2K and \(2i\mathfrak {R}K'\). We thus have

$$\begin{aligned} \int _{\mathscr { C}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}}&{{\mathrm{nc}}}\Bigl (\frac{u+2K-2\theta _e}{2} \Bigr ){{\mathrm{nc}}}\Bigl (\frac{u}{2} \Bigr )H(u)\frac{\mathrm {d}u}{2\pi i}\nonumber \\&\qquad =\frac{-2}{k'}\frac{H(2\theta _e)}{{{\mathrm{cn }}}\theta _e}+\frac{-2}{k'}\frac{H(2K)}{{{\mathrm{cn }}}(2K-\theta _e)}+\frac{2\mathfrak {R}K'}{\pi }{{\mathrm{nc}}}(i\mathfrak {R}K'+K-\theta _e){{\mathrm{nc}}}(i\mathfrak {R}K')\nonumber \\&\qquad =\frac{2}{k'}\frac{H(2K)-H(2\theta _e)}{{{\mathrm{cn }}}\theta _e}, \end{aligned}$$
(77)

since \({{\mathrm{nc}}}(i\mathfrak {R}K')=0\). The same reasoning as above gives

$$\begin{aligned}&\int _{\mathscr { C}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}}{{\mathrm{nc}}}\Bigl (\frac{u-2\theta _e}{2} \Bigr ){{\mathrm{nc}}}\Bigl (\frac{u+2K}{2} \Bigr )H(u)\frac{\mathrm {d}u}{2\pi i}\nonumber \\&\qquad =\frac{2}{k'}\frac{H(2\theta _e-2K)-H(0)}{{{\mathrm{cn }}}\theta _e}+\frac{2\mathfrak {R}K'}{\pi }{{\mathrm{nc}}}(i\mathfrak {R}K'+K){{\mathrm{nc}}}(i\mathfrak {R}K'-\theta _e), \end{aligned}$$
(78)

which may be slightly simplified, using that \(H(0)=0\). Thanks to (77), (78) and Table 2, we obtain that (76) equals

$$\begin{aligned} \mathsf {K}_{\mathsf {v}_{\ell },\mathsf {v}_{j}}\mathsf {K}^{-1}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}=\frac{1}{2}+\frac{H(2\theta _e)-H(2K)+H(2\theta _e-2K)}{2{{\mathrm{cn }}}\theta _e}+\frac{\mathfrak {R}K' k^2}{2\pi }{{\mathrm{sd}}}\theta _e. \end{aligned}$$

The addition formula (69) results in

$$\begin{aligned} H(2\theta _e-2K)-&H(2K)\\&=H(2\theta _e-4K)-\frac{\mathfrak {R}K'k^2}{\pi }\left\{ \begin{array}{ll} {{\mathrm{sn }}}K {{\mathrm{sn }}}(\theta _e-K){{\mathrm{sn }}}(\theta _e-2K) &{} \text {if } k^2>0\\ (-k'^2){{\mathrm{sd}}}K {{\mathrm{sd}}}(\theta _e-K){{\mathrm{sd}}}(\theta _e-2K)&{} \text {if } k^2<0\end{array}\right. \\&=H(2\theta _e)-1-\frac{\mathfrak {R}K'k^2}{\pi }{{\mathrm{cn }}}\theta _e{{\mathrm{sd}}}\theta _e. \end{aligned}$$

The proof is complete. \(\square \)

Fig. 20
figure 20

Compatibility around a rhombus

Using that \(\lim _{k\rightarrow 0}H(u)=\frac{u}{2\pi }\), see Lemma 42, we find that

$$\begin{aligned} \lim _{k\rightarrow 0}\mathsf {K}_{\mathsf {v}_{j},\mathsf {v}_{\ell }}\mathsf {K}^{-1}_{\mathsf {v}_{\ell },\mathsf {v}_{j}}=\frac{1}{2}-\frac{\pi -2\theta _e}{2\pi \cos \theta _e}. \end{aligned}$$

This is in accordance with the computation of [8, Appendix A] for the case \(k=0\). Indeed, the dimer model considered in the latter paper corresponds to complementary polygon configurations meaning that the probability (30) is 1 minus the probability of the paper [8].

Appendix C: Proof of Lemma 7

This section consists in the proof of Lemma 7 stating that the angles \((\overline{\alpha }_j({\mathbf x}))\) defined in Eqs. (11) and (12) are indeed well defined mod \(4\pi \).

Proof

We first need to check that angles around a rhombus corresponding to two neighboring decorations are well defined, see Fig. 20.

We want to prove that the following is equal to 0 mod \(4\pi \):

$$\begin{aligned} (\overline{\alpha }_{\ell +1}-\overline{\alpha }_{\ell })- (\overline{\alpha }_{j+1}-\overline{\alpha }_{j})+(\overline{\alpha }_{\ell }-\overline{\alpha }_{j})- (\overline{\alpha }_{\ell +1}-\overline{\alpha }_{j+1}). \end{aligned}$$

By definition of the angles within a decoration we have, mod \(4\pi \),

$$\begin{aligned} (\overline{\alpha }_{\ell +1}-\overline{\alpha }_{\ell })-(\overline{\alpha }_{j+1}-\overline{\alpha }_{j})= {\left\{ \begin{array}{ll} 0 &{} \text {if }{{\mathrm{co}}}(\mathsf {w}_\ell ,\mathsf {w}_{\ell +1})+{{\mathrm{co}}}(\mathsf {w}_j,\mathsf {w}_{j+1}) \text { is even,}\\ 2\pi &{} \text {if }{{\mathrm{co}}}(\mathsf {w}_\ell ,\mathsf {w}_{\ell +1})+{{\mathrm{co}}}(\mathsf {w}_j,\mathsf {w}_{j+1}) \text { is odd}. \end{array}\right. } \end{aligned}$$

By definition of the angles in neighboring decorations we have, mod \(4\pi \),

$$\begin{aligned} (\overline{\alpha }_{\ell }-\overline{\alpha }_{j})-(\overline{\alpha }_{\ell +1}-\overline{\alpha }_{j+1})= {\left\{ \begin{array}{ll} 0&{}\text {if } {{\mathrm{co}}}(\mathsf {w}_{j+1},\mathsf {v}_j,\mathsf {w}_j)+{{\mathrm{co}}}(\mathsf {w}_{\ell +1},\mathsf {v}_\ell ,\mathsf {w}_\ell ) \text { is even,}\\ 2\pi &{}\text {if } {{\mathrm{co}}}(\mathsf {w}_{j+1},\mathsf {v}_j,\mathsf {w}_j)+{{\mathrm{co}}}(\mathsf {w}_{\ell +1},\mathsf {v}_\ell ,\mathsf {w}_\ell ) \text { is odd}. \end{array}\right. } \end{aligned}$$

This implies that, mod \(4\pi \), we have:

$$\begin{aligned}&(\overline{\alpha }_{\ell +1}-\overline{\alpha }_{\ell })- (\overline{\alpha }_{j+1}-\overline{\alpha }_{j})+ (\overline{\alpha }_{\ell }-\overline{\alpha }_{j})- (\overline{\alpha }_{\ell +1}-\overline{\alpha }_{j+1})\\&\qquad = {\left\{ \begin{array}{ll} 0&{}\text {if }{{\mathrm{co}}}(\mathsf {w}_{j+1},\mathsf {v}_j,\mathsf {w}_j,\mathsf {w}_{j+1})+{{\mathrm{co}}}(\mathsf {w}_{\ell +1},\mathsf {v}_\ell ,\mathsf {w}_\ell ,\mathsf {w}_{\ell +1}) \text { is even,} \\ 2\pi &{} \text {if }{{\mathrm{co}}}(\mathsf {w}_{j+1},\mathsf {v}_j,\mathsf {w}_j,\mathsf {w}_{j+1})+{{\mathrm{co}}}(\mathsf {w}_{\ell +1},\mathsf {v}_\ell ,\mathsf {w}_\ell ,\mathsf {w}_{\ell +1}) \text { is odd}. \end{array}\right. } \end{aligned}$$

But since the orientation of the graph is admissible, we have that \({{\mathrm{co}}}(\mathsf {w}_{j+1},\mathsf {v}_j,\mathsf {w}_j,\mathsf {w}_{j+1})\) and \({{\mathrm{co}}}(\mathsf {w}_{\ell +1},\mathsf {v}_\ell ,\mathsf {w}_\ell ,\mathsf {w}_{\ell +1})\) are odd, implying that the sum is even, thus concluding the proof for angles around a rhombus.

Fig. 21
figure 21

Notation for a cycle C arising from the boundary of a face of \(\mathsf {G}\)

We now need to prove that when doing the inductive procedure around a cycle of \({\mathsf {G}}^{\mathrm F}\), we recover the same angle mod \(4\pi \). There are two types of cycles to consider: inner cycles of decorations and cycles arising from boundary of faces of \(\mathsf {G}\).

Consider a cycle of a decoration corresponding to a vertex \({\mathbf x}\) of \(\mathsf {G}\) of degree d, and let n be the number of edges of the inner cycle oriented clockwise. By definition of the angles, we have:

$$\begin{aligned} \overline{\alpha }_{d+1}- \overline{\alpha }_{1}=\sum _{j=1}^d (\overline{\alpha }_{j+1}-\overline{\alpha }_{j}) =\sum _{j=1}^d 2\overline{\theta }_{j}+2\pi n=2\pi (n+1). \end{aligned}$$

Since the orientation of the edges is admissible, n is odd, thus proving that \(\overline{\alpha }_{d+1}-\overline{\alpha }_{1}=0\ [4\pi ]\).

Now consider a cycle C arising from the boundary \({\mathbf x}_1,\dots ,{\mathbf x}_m\) of a face of \(\mathsf {G}\), with vertices labeled in clockwise order, see also Fig. 21. Up to a relabeling of vertices of the decorations, this cycle can be written as

$$\begin{aligned} C=(\mathsf {w}_{2}({\mathbf x}_1),\mathsf {v}_2({\mathbf x}_1),\mathsf {v}_{1}({\mathbf x}_2),\mathsf {w}_{2}({\mathbf x}_2),\dots ,\mathsf {v}_{1}({\mathbf x}_1)). \end{aligned}$$

Using the definition of the angles within a decoration and in neighboring ones we deduce that, mod \(4\pi \), we have:

$$\begin{aligned} \overline{\alpha }_{2}({\mathbf x}_j)-\overline{\alpha }_{2}({\mathbf x}_{j+1})= {\left\{ \begin{array}{ll} -\pi -2\theta _1({\mathbf x}_j) &{}\text {if } {{\mathrm{co}}}(\mathsf {w}_2({\mathbf x}_j),\mathsf {v}_2({\mathbf x}_j),\mathsf {v}_1({\mathbf x}_{j+1}),\mathsf {w}_2({\mathbf x}_{j+1})) \text { is odd,}\\ \pi -2\theta _1({\mathbf x}_j) &{}\text {if } {{\mathrm{co}}}(\mathsf {w}_2({\mathbf x}_j),\mathsf {v}_2({\mathbf x}_j),\mathsf {v}_1({\mathbf x}_{j+1}),\mathsf {w}_2({\mathbf x}_{j+1})) \text { is even}. \end{array}\right. } \end{aligned}$$

Let n(C) denote the number of portions of the cycle C where

$$\begin{aligned} {{\mathrm{co}}}(\mathsf {w}_2({\mathbf x}_j),\mathsf {v}_2({\mathbf x}_j),\mathsf {v}_1({\mathbf x}_{j+1}),\mathsf {w}_2({\mathbf x}_{j+1})) \end{aligned}$$

is odd. Then, writing \({\mathbf x}_{m+1}={\mathbf x}_1\), we have:

$$\begin{aligned} \alpha _2({\mathbf x}_{1})-\alpha _2({\mathbf x}_{m+1})=\sum _{j=1}^m \alpha _2({\mathbf x}_j)-\alpha _2({\mathbf x}_{j+1}) =\sum _{j=1}^m(\pi -2\theta _1({\mathbf x}_j))-2\pi n(C). \end{aligned}$$

Since \(\sum _{j=1}^m(\pi -2\theta _1({\mathbf x}_j))\) is the sum of angles at the center of the cycle, it is equal to \(2\pi \). The orientation of the cycle being admissible, n(C) is odd, thus concluding the proof. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Boutillier, C., Tilière, B.d. & Raschel, K. The Z-invariant Ising model via dimers. Probab. Theory Relat. Fields 174, 235–305 (2019). https://doi.org/10.1007/s00440-018-0861-x

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00440-018-0861-x

Mathematics Subject Classification

Navigation