ENERGY CORRELATIONS IN THE CRITICAL ON A TORUS

KONSTANTIN IZYUROV, ANTTI KEMPPAINEN AND PETRI TUISKU

Abstract. We compute rigorously the scaling limit of multi-point energy corre- lations in the critical Ising model on a torus. For the one-point function, averaged between horizontal and vertical edges of the square lattice, this result has been known since the 1969 work of Ferdinand and Fischer. We propose an alternative proof, in a slightly greater generality, via a new exact formula in terms of determi- nants of discrete Laplacians. We also compute the main term of the asymptotics of the difference E(V − H ) of the energy density on a vertical and a horizontal 2 edge, which is of order of δ , where δ is the mesh size. The observable V − H has been identified by Kadanoff and Ceva as (a component of) the stress-energy tensor. We then apply the discrete complex analysis methods of Smirnov and Hongler to compute the multi-point correlations. The fermionic observables are only periodic with doubled periods; by anti-symmetrization, this leads to contributions from four “sectors”. The main new challenge arises in the doubly periodic sector, due to the existence of non-zero constant (discrete) analytic functions. We show that some additional input, namely the scaling limit of the one-point function and of relative contribution of sectors to the partition function, is sufficient to overcome this difficulty and successfully compute all correlations.

1. Introduction The Ising model is a very famous and influential model in , mathematical physics, discrete and computer science. Originally in- troduced by Wilhelm Lenz in 1920 and named after Ernst Ising who in 1924 solved the 1D case of the model [33], the model was initially introduced to study the mag- netic phase change at the so-called Curie temperature. For example, a piece of iron loses its ferromagnetic properties and becomes a paramagnet at 770 ◦C. Since its inception, the Ising model has become a very much studied and arche- typal model, a “test laboratory”, of statistical physics. This is because the model is rather simple, but still encompasses a lot of the interesting behavior. Because of the wide applicability of the model, the Ising model has been studied in many arXiv:2104.01084v1 [math-ph] 2 Apr 2021 academic fields ranging from pure mathematics via physics and chemistry to biology and economics. The most interesting aspect of Ising model is the fact that it has an order-disorder phase transition in dimension 2 and higher dimensions. This fact was established by Peierls in 1936 [48]. The temperature of the phase transition, called critical temperature, was predicted by Kramers and Wannier in 1941 [43]. After it was understood that the model does have a phase transition in dimension 2, finding an exact formula for the partition function of the model in this case became a central question of much interest in statistical physics. The feat was achieved by Lars Onsager in a seminal paper in 1944 [46] using transfer matrices. Due to this exact and rigorous formula being established, it is often said that the Ising model 1 is exactly solvable or integrable in 2 dimensions. After 1944, the transfer matrix technique was developed further by Onsager and Kaufman [41, 42]. The spinor analysis of Kaufman [41] underlies much of our work. Besides the algebraic transfer matrix techniques (whose use has continued and developed since the 1940s), Ising model has also been studied by so-called combina- torial approach. This approach was advanced by, among others, Van der Waerden who developed the high temperature expansion in 1941 [54], Kac and Ward who introduced the Kac–Ward determinant in 1952 [36] (see also [49]), Sherman, Hurst- Green, Kasteleyn and Fisher who all developed a technique known as the Pfaffian method in 1960s [52] [32], [40, 38, 39], [26, 27] and McCoy and Wu, who wrote the book [45] summarizing the development of the combinatorial approach in 1970. The results of Van der Waerden and Kac–Ward especially are of key importance to our work. The conjecture of Kac and Ward, namely that the Ising model partition func- tion can be expressed as a determinant of a suitably chosen matrix, the Kac–Ward matrix, was established mathematically in 1999 by Dolbilin, Zinov’ev, Mishchenko, Shtan’ko and Shtogrin [20] and generalized to the form we apply in this paper by David Cimasoni [14], [15]. In 1960s-1970s, another, non-rigorous way to understand Ising model appeared in the physical literature, the renormalization group. This approach postulates that the Ising model has a “continuum limit” described by a “quantum field theory (QFT)” [1]. This theory is thought to be the free fermionic theory. The QFT approach led to multiple seminal hypothesis being proposed by physicists, among them the idea that the continuum limit correlations should be related to the determinants of Laplacians and Dirac operators. We shall show that for the energy density (correlation of 2 neighboring spins) an identity in the discrete setting exists that shows this relation to determinants of Laplacians. It is noteworthy that this identity is exact and rigorous and exists even in the discrete setting, not only in the continuum limit as proposed by QFT approach. The QFT approach was developed further by Belavin, Polyakov, Zamolodchikov [3] who in 1984 suggested that the limiting QFT for the critical Ising model case (when the fermion theory is “massless”) has conformal symmetry (CFT)(see also [1]). This lead to the famous prediction that the critical Ising model possesses conformal covariance in the scaling limit, a conjecture that has inspired a lot of work in the mathematical community in the last decade. In the mathematical community, a lot of work has been done to prove the conjec- tures by physicists, and to further the understanding of the Ising model. Let us here focus only on the study on conformal covariance of the critical Ising model, that is, the conformal covariance of its various correlations. The calculation of the full-plane energy correlations was achieved by Boutillier and de Tilière [4, 5]; this was later extended to some non-integrable models [29, 2]. The full-plane spin correlations were calculated by Palmer [47]. Smirnov [53] introduced a powerful tool to analyze the scaling limit of the Ising model on arbitrary domains, namely, the discrete holo- morphic fermionic observables and associated Riemann boundary value problem. This was applied to the calculation of the energy density in simply connected planar domains by Hongler and Smirnov [31]; later extended by Hongler to multi-point energy correlations in his Ph. D. thesis [30]. The case of spin correlations in planar domains was solved by Chelkak, Hongler and Izyurov [9] and recently these results were extended to all mixed correlations of primary fields in [12]. 2 The success of the discrete holomorphicity techniques in these papers naturally leads to the question of whether they can be applied to the analysis of the model on Riemann surfaces. On the positive side, the existing universality results [11, 10, 7] allow one to extend the techniques to families of graphs that are flexible enough to approximate any . However, there are several known difficulties on this path. First, the fermionic correlations are not well defined on the Riemann surface but have non-trivial monodromy properties. Thus, one needs to consider 22g observables on a genus g surface where just one was sufficient in the planar case. This is similar to the 22g Kac–Ward determinants introduced by Cimasoni [14, 15], and in fact this is no coincidence since the observables are related to the inverses of the Kac–Ward matrices, see [44, 6]. What’s more, the observables do not compute the energy or spin correlations per se, but rather their correlations with one of 22g “topological” observables, normalized by the expectation of that observable, given by the corresponding Kac–Ward determinant. Thus, to recover the correlations of interest, one would need to compute scaling limit of the ratio of the square root of each of the 22g Kac-Ward determinants to their sum. Finally, a degeneracy occurs in that one of the observables happens to be a non-zero constant, whose value is not immediately recoverable by the discrete complex analysis methods. The present paper provides a first step in this program, treating the case of energy correlations on a torus. The advantage of working with the flat torus is that the above-mentioned difficulties can be treated by other methods. Thus, the Kac-Ward determinants can be calculated explicitly, and their limiting asymptotics analyzed. Moreover, as the partition function can be computed for any temperature, one can differentiate it with respect to the temperature to obtain the average energy density at criticality. This has been done by Ferdinand and Fischer [23] and later refined in [50, 51, 34] by using the Onsager–Kaufman expression for the partition function. It turns out that a slight refinement of this result (separating the vertical and the horizontal edges) is sufficient to remedy the above-mentioned degeneracy problem. We propose an alternative computation of the average energy density, based instead on the Kac–Ward solution. The advantage of our approach is that we relate, at the discrete level, the average energy density to a ratio of determinants of the discrete Laplacians. The asymptotics of the determinants of Laplacians has been recently analyzed in great generality [24, 25, 35]. Thus, if a similar “bosonization” relation is found on Riemann surfaces, the asymptotics of the one-points function in the scaling limit would be readily available. It turns out that once the above-mentioned difficulties are treated, or the missing pieces supplied as an input, the rest of the computation of the arbitrary many-point energy correlations can be done by the discrete complex analysis methods, recovering physicists’ predictions in this regard, see Theorem 5 below. While for even number of marked points, the argument in [30, 12] can be extended almost verbatim, for odd number points we had to modify it, as we have found no combinatorial counterpart for the “propagator” ζ(en − em) in (2.5) below. Therefore, rather than reducing everything to the two-point fermionic observables as in [30, 12], we had to work out the convergence result directly for a multi-point observable, and only obtain the Pfaffian formula (2.5) by analysing the resultin scaling limit in the continuum.

3 Acknowledgements. Work supported by the Academy of Finland via Centre of Ex- cellence in Analysis and Dynamics research and the academy project “Critical phe- nomena in dimension two: analytic and probabilistic methods”. We are grateful to Antti Kupiainen and Dmitry Chelkak for useful discussions. We thank David Loeffler for pointing out a quick way to derive the Kronecker limit formula with anti-periodic boundary conditions, used in the proof of Corollary 2.

2. Main results 2.1. Setup and notation. We will study the Ising model on tori Tδ := δZ2/Λδ, where δ > 0 is the mesh size, δ δ δ Λ = {nω1, +mω2 : n, m ∈ Z} δ 2 and ω1,2 ∈ δZ are non-collinear. We will be interested in the scaling limit of the δ model where δ → 0, ω1,2 → ω1,2 ∈ C \{0}. We will assume that the modular parameter τ := ω2/ω1 satisfies =mτ > 0. The (zero magnetic field) Ising model on Tδ is the probability measure on spin configurations σ : Tδ → {±1} given by ! 1 X P [σ] = exp β σxσy Z x∼y where the sum is over all pairs of nearest-neighboring vertices of Tδ, β > 0 is a parameter called the inverse temperature, and ! X X Z = exp β σxσy x∼y σ:Tδ→{±1} is the partition function of the model, which ensures that the configuration prob- abilities sum√ up to 1. Our main results will concern the critical temperature β = 1 βc = 2 log( 2 + 1). We will denote by P and E, respectively, the probability and the expectation with respect to the above measure. The main object of our interest is the expectation of the energy observable. If (xy) is an edge of Tδ, we denote 1 (xy) = σxσy − √ . 2 The constant √1 is the expectation of σ σ in the full-plane, i.e., the thermodynamic 2 x y limit. Thus, our results will measure how the toric boundary conditions affect the δ 2 expectation and the correlations of (xy). Clearly, since T carries an action of δZ by translations, we have Ee = Eeˆ if the edges e, eˆ are either both horizontal, or both vertical. We thus denote by EH and EV the expectation of  on any horizontal and vertical edge, respectively. 2.2. Main results. Our first result concerns the sum of vertical and horizontal δ energy densities. We introduce some notation. Given ω1,2 as above and i, j ∈ Z2, denote ij 2 δ i δ j Vδ := {f : δZ → R : f(v + ω1) ≡ (−1) f(v), f(v + ω2) ≡ (−1) f(v)}. ij δ 00 Each of Vδ is a linear space of dimension |T |; Vδ can be viewed just as the set δ P of functions on T . Note that the lattice Laplacian ∆δf(x) := y∼x(f(y) − f(x)) ij ij preserves each of these spaces. We denote by ∆δ the restriction of ∆ to Vδ . 4 Theorem 1. For the critical Ising model on defined Tδ, we have, for a horizontal edge H and a vertical edge V :

q ? 00 det ∆δ 1 (2.1) EV + EH = 4 · q q q · , 10 01 11 | δ| det ∆δ + det ∆δ + det ∆δ T where det? denotes the product of all non-zero eigenvalues. The continuous counterpart of this identity has appeared in the CFT literature, see [28], [22], [17], however, the discrete version is, to the best of our knowledge, new. Cimasoni [15] has related the determinants of the discrete Laplacians to the critical Ising partition functions on arbitrary isoradial graphs embedded on a torus. However, we were unable to adapt his methods to the computation of energy densi- ties. A more general approach to bosonization of the Ising model was developed by Dubédat in [21], however, it involves more complicated modifications of the original graph and does not seem to lead to (2.1) either. In Section 8, we provide an analog of (2.1) for triangular lattice; we do not know whether such analogs hold true for other lattices. By combining the above formula with known results on the asymptotics of de- terminants of discrete Laplacians, we recover the asymptotics result of Ferdinand and Fischer [23], in a slightly greater generality of arbitrary torus as compared to diagonal one:

δ Corollary 2. In the limit δ → 0, ω1,2 → ω1,2 with ω2/ω1 = τ, we have

1 2(=mτ) 2 |θ2θ3θ4| 1 EV + EH = · 1 · δ + o(δ). |θ2| + |θ3| + |θ4| |T| 2

Hereinafter, our notation for the theta constants is θi := θi(τ) := θi(0, q), where iπτ q = e and θi(0, q) is as in [19, Chapter 20] Our second result concerns the asymptotics of the difference between energy den- sity on vertical and horizontal edges. This difference happens to be of order δ2; such an observable was identified [37] as a component of stress-energy tensor in the model, see also [8].

δ Theorem 3. In the limit δ → 0, ω1,2 → ω1,2 with ω2/ω1 = τ, we have √ 2π (2.2) (e ) − (e ) = · H(ω , ω ) · δ2 + o(δ2), E V E H 24 1 2 where H(ω1, ω2) is given by Z(01) h i Z(10) h i Z(11) h i

δ Corollary 4. In the limit δ → 0, ω1,2 → ω1,2 with ω2/ω1 = τ, we have

1 (=mτ) 2 |θ2θ3θ4| 1 EH = · 1 · δ + o(δ). |θ2| + |θ3| + |θ4| |T| 2 and similarly for EV . 3 In fact, higher-order terms (up to δ ) of the expansion of EH +EV were computed by Salas and Izmailyan–Hu in [50, 34]. They showed that the δ2 term is absent from the expansion. Therefore, our results in fact give the expansion of EH up to order o(δ2). Chinta, Jorgenson and Karlsson [13] indicate a way to compute the asymptotic expansion of det? ∆00 up to arbitrary order in δ. Combined with our Theorem 3 , this could in principle used to the compute EH + EV up to arbitrary order. We also compute the scaling limit of multi-point correlation functions.

δ Theorem 5. In the scaling limit δ → 0, T → T, as e1, . . . , ek approach distinct points of T, we have, for even k, Z(01) (2.4) πkδ−k [ . . .  ] −→ |Pf [cs (e − e )]|2 E e1 ek Z ω1,ω2 n m Z(10) Z(11) + |Pf [ns (e − e )]|2 + |Pf [ds (e − e )]|2 , Z ω1,ω2 n m Z ω1,ω2 n m where csω1,ω2 , nsω1,ω2 , dsω1,ω2 are Jacobian elliptic functions, see Section 6. For odd k, we have k −k k (2.5) π δ E [e1 . . . ek ] → (−i) · Pf M, where M is the 2k×2k anti-symmetric matrix given by M2n−1,2m−1 = ζω1,ω2 (en −em),

M2n,2m = ζω1,ω2 (en − em), and 1 (=mτ) 2 |θ2θ3θ4| 1 M2n−1,2m ≡ πi · · 1 . |θ2| + |θ3| + |θ4| |T| 2

Here ζω1,ω2 is the Weierstrass ζ–function, see Section 6. The formula (2.4) has been predicted in the physics literature by di Francesco, Saleur and Zuber [18], see also [17, Section 12]. The formula (2.5), on the other hand, appears to be new. We expect it to be related to the prediction of [18] by an appropriate version of Fay’s formula.

3. Partition functions and Kac–Ward determinants In this Section, we record the necessary results involving Kac–Ward solution to the critical Ising model. This approach was originated in [36]; the first complete proof was given in [20], and in the case of surfaces in [14], with simplified proof in [6]. We now recall the required material in detail in the case of a torus, following [6, Section 4]. 6 Let us denote E(Tδ) the set of even subgraphs of Tδ (understood as subsets of edges of Tδ), that is, subgraphs of Tδ such that each vertex is adjacent to an even number of edges. It is well known (see for example [15, Subsection 2.1] for a short exposition) that the partition function Z can be expressed as (the high temperature expansion)

δ δ Z = cosh (β)|Edges(T )| 2|T | X tanh (β)|ξ| . ξ∈E(Tδ) To avoid the factor const appearing in every formula, let us define the partition function of the Ising model without this factor: (3.1) ZI := X α|ξ|. ξ∈E(Tδ) We shall refer to this sum also as “partition function”. δ Pick a point z0 ∈ C such that the two lines γ1,2 := {z0 + tω1,2 : t ∈ R} do not 2 δ δ intersect δZ . We identify γ1,2 with their projection onto T . Given an edge e ∈ T and i, j ∈ Z2, put

(3.2) ϕij(e) = iIe∩γ16=0 + jIe∩γ26=0 mod 2 = iϕ10(e) + jϕ01(e) mod 2 δ These are four Z2-valued flat connections on T (written additively); in fact, these are the only flat connections up to gauge equivalence. As explained in [6, Section 4], this allows one to construct four spin structures λij, which are, roughly speaking, ways to assign a winding number modulo 4π to a closed lattice path. Namely, if the lattice path γ consist of the edges e1, e2, . . . , ek, then windλij (γ) = wind(γ) + 2πϕij(γ), where wind is the winding of the lift of γ to the plane. This, in its turn, allows δ δ one to define four quadratic forms qij on E(T ): given ξ ∈ E(T ), decompose it into a collection of loops ξ1, . . . ξN that do not intersect themselves or each other transversally (to this end, for each vertex of degree 4 in ξ, pick any two incident edges forming a right angle, and declare them belong to the same loop, and also other two qij (ξ) QN i to belong to the same loop). Then, put (−1) := k=1(− exp( 2 windλij (Ck))). 2 Let us calculate qij(ξ) concretely. If a loop ξm lifts to a closed loop on δZ , then it crosses γ1 and γ2 an even number of times, and thus windλij (ξm) = wind(ξm) = 2π. Otherwise, it lifts to a path connecting two distinct points in the plane, and we have wind(ξm) = 0. Thus, we have (−1)qij (ξ) = (−1)N(ξ)+ϕij (ξ), where N(ξ) is a number of non-contractible loops in (the decomposition of) ξ. The δ δ lift of a non-contractible loop connects z and z + m1ω1 + m2ω2; since the loops are simple and non-intersecting, m1 and m2 are relatively prime and the same for all non-lifting loops in ξ. This means that N(ξ) = ϕ01(ξ) + ϕ01(ξ) − ϕ01(ξ)ϕ10(ξ), by exclusion-inclusion: configurations with an odd number of non-contractible loops contributes to ϕ01(ξ) iff m1 is odd (respectively, to ϕ10(ξ) iff m2 is odd). Therefore, (3.3) (−1)qij (ξ) = (−1)(1−i)ϕ10(ξ)+(1−j)ϕ01(ξ)+ϕ10(ξ)·ϕ01(ξ).

Of course, this is a manifestation of the general fact that qij is a quadratic form on δ the homology space H1(T , Z2). To each of the four spin structures, one associates a twisted Kac–Ward matrix ~ δ δ KWij. That is is a matrix indexed by the set E(T ) of oriented edges of T , by 7 putting KWij := Id − KWij, where

  i 0  0 0 ij ϕij (~e) exp 2 wind (~e,~e ) α, if t (~e) = o (~e ) but ~e 6= −~e KW~e,~e0 := 0, otherwise Here t(~e) and o(~e) denote the end and the beginning of e, respectively. The following theorem due to Cimasoni relates the Kac–Ward matrices with Ising partition functions:

Theorem 6. We have, for i, j ∈ Z2, q X (3.4) det KWij = (−1)qij (ξ)α|ξ| =: Z(ij). ξ∈E(Tδ) The following Lemma is a particular case of the equation (4.6) in [6]. It al- lows, in particular, to express the partition function ZI in terms of determinants of Kac–Ward matrices.

Lemma 7. We have, for any ξ ∈ E(Tδ), (3.5) (−1)q10(ξ) + (−1)q01(ξ) + (−1)q11(ξ) − (−1)q00(ξ) = 2 q q q q (3.6) 2ZI = − det KW00 + det KW01 + det KW10 + det KW11

Proof. The first identity easily follows from (3.3): if ϕ01(ξ) = ϕ10(ξ) = 0, then all the terms in the left-hand side are equal to 1, and altering ϕ01(ξ) or ϕ10(ξ) always changes the sign of exactly two terms. The second identity is obtained by |ξ| multiplying the first one by α and summing over ξ.  In the case of a torus, the determinant of Kac–Ward matrix can be calculated explicitly. Let us first denote, for q ∈ C, z (q) = exp (2πi

For the following (purely combinatorial) discussion, we will assume δ = 1. The dual lattice Λ∗ of a lattice Λ ⊂ Z2 is defined as ∗ Λ := {q ∈ C |

1 2 ∗ (3.9) 2 v(αc, q), q ∈ Z /Λ + s(ij), 2αc √ where αc = 2 − 1

Proof. The exponential function fq(z) = exp(2πi(

4. The sum of the horizontal and vertical edge energy densities We continue with a standard Lemma representing energy density in high-temperature expansion.

δ Lemma√ 10. Let e = (xy) ∈ Edges(T ) be any edge. For the critical temperature α = 2 − 1, one has

1 1 X  −1  |ξ| (4.1) E(e) = √ α Ie∈ξ − αIe/∈ξ α . 2 ZI ξ⊂E(Tδ) δ δ Proof. Denote by Ee(T ) the set of configurations ω ⊆ Edges(T ) such that every vertex except the end-vertices of e has even degree and the end-vertices of e have δ δ odd degree in ω. If ξ ∈ E(T ), define ωe(ξ) ∈ Ee(T ) by ωe(ξ) = ξ ∪ e if e∈ / ξ and 9 δ δ ωe(ξ) = ξ \ e if e ∈ ξ. Then, ωe is is a bijection between E(T ) and Ee(T ). By high-temperature expansion, we have

1 X |ω| 1 X |ωe(ξ)| 1 X  |ξ|−1 |ξ|+1 E(σxσy) = I α = I α = I Ie∈ξα + Ie/∈ξα Z δ Z δ Z δ ω∈Ee(T ) ξ∈E(T ) ξ∈E(T ) ! ! ! 1 1 X −1 1 |ξ| 1 |ξ| = √ + α − √ Ie∈ξα + α − √ Ie/∈ξα , 2 ZI 2 2 ξ∈E(Tδ)

−1 and we finish by noting that α−1 − √1 = α√ and α − √1 = − √α . 2 2 2 2 

δ For i, j ∈ Z2 and e and edge of T , define   (ij) X −1 |ξ| qij (ξ) |ξ| B (e) := α Ie∈ξ − αIe/∈ξα (−1) α . ξ∈E(Tδ) In view of Lemma 10 and (3.5), we have

1 1  (01) (10) (11) (00)  (4.2) E(e) = √ B (e) + B (e) + B (e) − B (e) . 2 2 ZI It turns out that at the critical temperature, it is possible to express the sum of B(ij)(e) for horizontal and vertical edge in terms of a determinant of the Laplacian: √ Lemma 11. One has, for α = αc = 2 − 1,  | |+5 √ Tδ ? 1  2 00 (ij) (ij) −2 det ∆ | δ| , i = j = 0; B (eH ) + B (eV ) = T 0, otherwise.

Proof. Denote by Z(ij)(α) the right-hand side of (3.4). We can write

d 1 X 1 X X Z(ij)(α) = |ξ| α|ξ|(−1)qij (ξ) = α|ξ|(−1)qij (ξ) dα α α Ie∈ξ ξ∈E(Tδ) e ξ∈E(Tδ)

δ   T X |ξ| qij (ξ) X |ξ| qij (ξ) =  e ∈ξα (−1) + e ∈ξα (−1)  , α I H I V ξ∈E(Tδ) ξ∈E(Tδ) where we used that Tδ has exactly |Tδ| horizontal edges and |Tδ| vertical edges, and P |ξ| qij (ξ) ξ∈E(Tδ) Ie∈ξα (−1) only depends on whether e is vertical or horizontal. Now, 1 −1 α using that Ie∈ξ = α+α−1 (α Ie∈ξ − αIe/∈ξ) + α+α−1 , we arrive at d 2| δ| | δ|α−1   (4.3) Z(ij)(α) = T Z(ij)(α) + T B(ij)(e ) + B(ij)(e ) dα α + α−1 α + α−1 H V d (ij) We now use Theorem 6 and Theorem 8 to compute dα Z (α) at α = αc. It is straightforward to see from (3.7) that

2  −1 −1 (4.4) v(αc, q) = 2αc 4 − z (q) − z (q) − w (q) − w (q) ; √ d 2  −1 −1 (4.5) v(α, q) = 2 2αc 4 − z (q) − z (q) − w (q) − w (q) . dα α=αc

In particular, since |z(q)| = |w(q)| = 1, the quantity v(αc, q) can only vanish when 2 ? z(q) = w(q) = 1, that is, q ∈ Z . If i = 1 (respectively, j = 1), then, for q ∈ Λ +s(ij), 1 1 we have

X −s ζT(s) := λn , 1 λn6=0 where the sum is over all non-zero eigenvalues of ∆. This ζ-functions can be ana- lytically continued into the vicinity of 0, and the ζ-regularized determinant of the Laplacian is defined by

? 0 log detζ ∆ = −(ζT) (0). We will derive Corollary 2 from the following result relating asymptotics of deter- ? minant of discrete Laplacian and to detζ ∆:

δ Proposition 12. ([13, Theorem 1], see also [35]) As δ → 0 and ω1,2 → ω1,2, one has

? 00 −2  δ  ? (4.8) det ∆δ = δ exp C|T |) detζ ∆ · (1 + o(1)), where C is an explicit constant.

? Another ingredient is the classical computation of detζ ∆, due to Kronecker:

? 4 (4.9) detζ ∆ = =mτ|T||η(τ)| ,

1  1  3 where η(τ) = 2 θ2(τ)θ3(τ)θ4(τ) is the Dedekind eta-function, see e. g. [17, section 10.2].

Proof of Corollary 2. In this proof, we will use several lattices, and hence we will ij ij ij ij use the notation V δ δ and ∆ δ δ for Vδ and ∆δ , respectively, emphasizing the ω1,ω2 ω1,ω2 dependence on the periods. Note that

00 00 10 00 00 01 01 01 11 V δ δ = V δ δ ⊕ V δ δ ; V δ δ = V δ δ ⊕ V δ δ ; V δ δ = V δ δ ⊕ V δ δ , 2ω1,ω2 ω1,ω2 ω1,ω2 ω1,2ω2 ω1,ω2 ω1,ω2 2ω1,ω2 ω1,ω2 ω1,ω2 and the direct summands are invariant subspaces for the discrete Laplacian. Hence, using the asymptotics (4.8), we obtain

? 00 ? det ∆2ωδ,ωδ δ det ∆,2ω ,ω 10 1 2 C|T |+o(1) ζ 1 2 (4.10) det ∆ δ δ = = e ; ω1,ω2 ? 00 ? det ∆ δ δ detζ ∆ω1,ω2 ω1,ω2 ? 00 ? det ∆ωδ,2ωδ δ det ∆ω ,2ω 01 1 2 C|T |+o(1) ζ 1 2 (4.11) det ∆ δ δ = = e ; ω1,ω2 ? 00 ? det ∆ δ δ detζ ∆ω1,ω2 ω1,ω2 01 ? 00 ? 00 det ∆ δ δ det ∆ δ δ det ∆ δ δ 11 2ω1,ω2 2ω1,2ω2 ω1,ω2 (4.12) det ∆ δ δ = = ω1,ω2 ? 01 ? 00 ? 00 det ∆ δ δ det ∆ δ δ det ∆ δ δ ω1,ω2 2ω1,ω2 ω1,2ω2 ? ? δ det ∆ det ∆ C|T |+o(1) ζ 2ω1,2ω2 ζ ω1,ω2 = e ? ? detζ ∆2ω1,ω2 detζ ∆ω1,2ω2 12 By plugging z = 0 into the reduction identities for theta functions [19, 20.7.11- 20.7.12], we get 1 1 1 θ (2τ)θ (2τ)θ (2τ) = θ2(τ)θ 2 (τ)θ 2 (τ) 2 3 4 2 2 3 4 √ 1 1 2 2 2 θ2 (τ/2) θ3 (τ/2) θ4 (τ/2) = 2θ2 (τ)θ3 (τ)θ4(τ). Using these identities and (4.9), we get ? 4 2 detζ ∆,2ω1,ω2 |η(τ/2)| |θ4(τ)| ? = 4 = 2 ; detζ ∆ω1,ω2 |η(τ)| |η(τ)| ? 4 2 detζ ∆ω1,2ω2 |η(2τ)| |θ2(τ)| ? = 4 4 = 2 ; detζ ∆ω1,ω2 |η(τ)| |η(τ)| ? ? 8 2 detζ ∆2ω1,2ω2 detζ ∆ω1,ω2 |η(τ)| |θ3(τ)| ? ? = 4 4 = 2 detζ ∆2ω1,ω2 detζ ∆ω1,2ω2 |η(τ/2)| |η(2τ)| |η(τ)| Combining this with (4.10)–(4.12) and and plugging these asymptotics and (4.9) into the expression obtained in Theorem 3, we get 1 1 3 −1 (=mτ) 2 | | 2 |η(τ)| δ 1 T o E(H ) + E(V ) = 4 δ (1 + (1)) |θ2(τ)| + |θ3(τ)| + |θ4(τ)| |T | 1 3 4(=mτ) 2 |η(τ)| = 1 δ + o(δ), |T| 2 (|θ2(τ)| + |θ3(τ)| + |θ4(τ)|) δ 2 where we have noticed that |T |δ → |T|. This concludes the proof.  5. Discrete holomorphic fermionic observables on a torus To prove Theorems 3 and 5, we employ the discrete holomorphic fermionic ob- servables. We follow√ the definitions and constructions in [12]; note that the lattice there is scaled by 2 and rotated by π/4. By a corner of the lattice δZ2, we mean a midpoint of a segment joining a lattice vertex v ∈ δZ2 with a vertex of the 2 ? 2 δ δ dual lattice (δZ ) = δZ + 2 + i 2 . The corners thus form another square lattice δ 2 δ δ Cδ := 2 Z + 4 + i 4 . We also define the corner graph C(Tδ) and the dual graph δ,? δ ◦ T of the torus T in an obvious way. Given a point z ∈ C(Tδ), we denote by z (respectively, z•) the vertex of Tδ (respectively, Tδ,?) incident to z. By the doubling Tˆ δ of the torus Tδ, we mean the torus ˆ δ 2 δ δ 2 T := δZ /{2mω1 + 2nω2 : m, n ∈ Z }. Note that Tˆ δ naturally forms a 4-sheet cover of Tδ. Given a corner a ∈ C(Tˆ δ), we denote δ δ apq := a + pω1 + pω2, p, q ∈ {0, 1}, the four point that project to the same image in Tδ. Given a subset γ of edges of Tδ,?, we define the disorder observable P −2β σxσy µγ := e (xy)∩γ6=0 , where the sum is over the nearest neighbors x ∼ y ∈ Tδ such that the edge (xy) intersects γ. We view subsets of edges of Tδ,? as chains modulo 2, in particular, γ1 + γ2 = γ1 − γ2 will denote the symmetric difference of γ1 and γ2, and ∂γ is 13 boundary of γ, that is, the set of all vertices of Tδ,? incident to an odd number of edges in γ. It turns out that the correlation of µγ with spins only depends on ∂γ and a homological class of γ modulo 2, as we now describe:

δ,? Lemma 13. Let v1, . . . , v2n be vertices of Tδ, and γ1,2 two subsets of edges of T δ,? with ∂γ1 = ∂γ2. If [γ1] − [γ2] = 0 in H1(T , Z2), then

E[σv1 . . . σv2n µγ1 ] = Sv1,...,v2n (γ1 − γ2)E[σv1 . . . σvn µγ2 ]. N Here Sv1,...,vn (γ) = (−1) , where N = |{v1, . . . , vn} ∩ F | and F is (any) collection δ,? P of faces of T such that f∈F ∂v = γ. ˜ Proof. Using γ1,2 as branch cuts, one can construct two double covers T1,2 of the graph Tδ. To this end, consider two copies ("sheets") of Tδ, remove in each copy the edges crossing γ1,2, and add instead the two edges connecting the corresponding vertices on different sheets. We can write P 1 X β P σ σ −2β σxσy [σ . . . σ µ ] = σ . . . σ e x∼y x y e (xy)∩γ6=0 E v1 vn γ1,2 Z v1 v2n σ:Tδ→{±1} 1 β P σ σ X 2 x∼y x y = σv1 . . . σv2n e , Z ˜ σ:T1,2→{±1} σv≡−σv∗ where v ∈ Tδ is identified with its copy on the first sheet, and v∗ denote its copy P on the second sheet. However, if γ1 − γ2 = v∈F ∂v, then flipping the sheets of all vertices corresponding to v ∈ F gives an isomorphism of the two double covers, under which exactly N of v1, . . . , v2n will move to the second sheet. This gives the desired result. 

In view of this lemma, the quantity E[σv1 . . . σv2n µγ] can be understood as a "multi-valued function" of v1, . . . , v2n and u1, . . . , u2m, where ∂γ = {u1, . . . , u2m}, meaning that once the initial condition (i. e., the choice of γ for a certain posi- tion of v1, . . . , u2m) is prescribed, there is a natural way to extend its value as the 0 marked points move around in in the lattice. Namely, when moving ui 7→ ui ∼ ui, 0 0 one replaces γ by γ + (uiui), and when moving vi 7→ vi, one adds a − sign whenever 0 (vi, vi) crosses γ. With this convention, if all the points move and come back to their initial positions, say, without leaving a fixed fundamental domain of Tδ, then 1 Q 2 the expression E[σv1 . . . σv2n µγ] changes sign in the same way as (ui − vj) For p, q ∈ {0, 1}, we denote

(5.1) µpq := µγpq ,

δ,? 2 where γpq is a simple loop on T that lifts to a path on δZ connecting a point z δ δ with z + pω1 + qω2. In particular, we can choose µ00 = ∅. Definition 14. We define the Dirac spinor

1 • ◦ !− 2 iπ z − z (5.2) η := e 4 , z |z• − z◦| understood as a two-valued function on the corner lattice C(Tδ), i. e., the function the double cover of that graph ramified at every face. 14 Figure 5.1. The lattice δZ2 as ◦, its dual lattice as •, the corner [e1] lattice as . On the right: the modified graph C[a] (·). Any closed • ◦ path on C[a](·) either encircles both a and a , or none; hence F (a, ·)

is single-valued on the modified graph. The values Fe1 (a, aL) and

Fe1 (a, aR) are proportional to E(eL e1 ) and E(eR e1 ) respectively, see Lemma 16.

The fermionic observables we use are similar to the to multi-point fermionic ob- servables as in [12, Section 2.4]. Since we will only need those for the energy corre- lation, we will restrict our consideration to a very specific configuration. Let C[e1...ek](Tˆ δ) denote the graph obtained by deleting from C(Tˆ δ) the edges cross- δ δ δ δ ing e1, . . . , ek and their shifts by by ω1, ω2, ω1 + ω2. The quantity Fe1...ek (a, ·) de- fined below in (5.3) is a “spinor” on C(Tˆ δ) ramified at a•, a◦ and their shifts by δ δ δ δ ˆ δ ω1, ω2, ω1 +ω2, i. e., a function on the double cover of C(T ) ramified at those points and changing sign between sheets. We prefer to view is as a function rather that spinor, by introducing the corresponding the branch cut [a◦a•] that divides a into two vertices a±, a+ being on the left as seen from a•, and similarly for the shifts. [e1...ek] ˆ δ We denote the resulting graph by C[a] (T ).

δ Definition 15. Given distinct edges e1, . . . , ek in the torus T that pairwise have no common incident corner, and a corner a ∈ C(Tδ), we define the observable

" k # Y ◦ • • • (5.3) Fe1...ek (a, z) := ηzE σz σa µz µa ei . i=1

We fix a choice of a square root in the definition of ηa and set the “initial conditions” +,• • + for Fe1...ek (a, z) of the above observable by µz a = µ∅ and ηa = ηa , so that we have " k # ± Y Fe1...ek (a, a ) = ±ηaE ei . i=1 Recall that a function F defined on a corner graph is called s-holomorphic if it satisfies the phase condition F (z) ∈ ηzR, and for every edge e, one has

(5.4) F (eNE) + F (eSW ) = F (eNW ) + F (eSE), 15 where eNE, eSW , eNW , eSE are the four corners neighboring e. We will, in fact, only need the following simple properties of s-holomorphic functions: if F is s- holomorphic, then its restriction to each of the two sub-lattices {z : ηz ∈ ηz0 R ∪ iηz0 R} are discrete holomorphic in the usual sense, that is, δ ! δ ! 1 δ ! δ !! (5.5) F z + − F z − − F z + i − F z − i = 0, 2 2 i 2 2 for every z ∈ C such that F is s-holomorphic at both edges intersecting the seg- h δ δ i h δ δ i ments z − 2 ; z + 2 and z − i 2 ; z + i 2 . As a consequence, if, for an open set Ω, 2 a sequence Fδ of s-holomorphic functions defined on refining lattices Ω ∩ C(δZ ) is uniformly bounded on every compact subset of Ω, then there is a holomorphic function f :Ω → C and a subsequence δk such that

Fδk (z) = Projηz (f(z)) + o(1), uniformly on compact subsets of Ω.

Lemma 16. The observable Fe1...ek (a, ·) is a well defined s-holomorphic function on [e1...ek] ˆ C[a] (Tδ). Moreover, it has the following special values: " k # ± Y (5.6) Fe1...ek (a, a ) = ±ηaE em , m=1 " k # ± δ δ Y (5.7) Fe1...ek (a, a + pω1 + qω2) = ∓ηaE µpq em , (pq) 6= (00). m=1 • ◦ • ◦ If, moreover, we denote aL := a + i · (a − a ) and aR = a − i · (a − a ), and by eL,R the edges incident to both a and aL,R respectively, then

" k # √ Y (5.8) Fe1...ek (a, aL,R) = iηa 2E eL,R em , m=1 √ " k # δ δ Y (5.9) Fe1...ek (a, aL,R + pω1 + qω2) = ∓iηa 2E µpqeL,R em , (pq) 6= (00). m=1 Proof. First, observe that F (a, ·) is locally well-defined, since when z moves e1...ek ˆ • ◦ around any face of C(T) except one corresponding to aij or aij, the −1 sign acquired • ◦ + by ηz cancels the −1 sign coming from winding of z around z . Moving z from a − ◦ − to a around, say, a results only in the sign change of ηz, hence, F (a, a ) = −ηa. δ δ Moving z "around the torus" to zpq = z + pω1 + qω2 results in replacing γ defining µa• µz• with γ + γpq; hence, doing so twice yields the same value. Finally, a simple + − loop lifting to a path connecting a with apq must start with a and end at apq, hence the − sign in (5.7) when (pq) 6= (00). The proof of s-holomorphicity in [12] is completely local and thus extend verbatim to our case. To prove (5.8), we start + with the value Fe1...ek (a, a ) = ηaE[... ] and move in two steps to z = aL so that z• − z◦ rotates clockwise. This way, z◦ does not cross the newly created disorder • • line [a aL] , and we get

−2βσa◦ σa◦ η σ ◦ σ ◦ µ • µ • = iη σ ◦ σ ◦ e L aL a aL a aL a a aL   √ ◦ ◦ = iηa σa σa cosh(−2β) + sinh(−2β) = 2iηaσa◦ σa◦ L L 16 √ inside the correlation, where we have used that cosh(−2β) = 2 and sinh(−2β) = − −1. The computation for aR is identical, except we start Fe1...ek (a, a ) = −ηaE[... ] and move so that z• − z◦ rotates counterclockwise. The proof of (5.9) is similar, taking into account the disorder line γpq created when moving from a to apq. 

It is convenient to anti-symmetrize Fe1...ek by introducing 1 F (ij) (a, z) := X (−1)ip+jqF (a, z + pωδ + qωδ). e1...ek 1 2 4 p,q∈{0,1} We summarize their properties in the following Lemma: Lemma 17. The observables F (ij) (a, ·) are s-holomorphic functions on C[e1...ek](ˆ δ) e1...ek [a] T that satisfy the anti-periodicity condition

(5.10) F (ij) (a, z + pωδ + qωδ) = (−1)ip+jqF (ij) (a, z). e1...ek 1 2 e1...ek They have special values given by " k # (5.11) F (ij) (a, a±) = ±η µ(ij) Y  , e1...ek aE em m=1 √ " k # (5.12) F (ij) (a, a ) = iη 2 µ(ij) Y  e1...ek L,R a E eL,R em m=1 where (ij) 1 X (1−i)p+(1−j)q+pq (5.13) µ := (−1) µpq. 4 pq∈{0,1} Moreover, there exists a constant c ∈ R (depending on Tδ and a but not on z) such (00) that F (a, z) = Projηa (iηac) . In particular, (00) (00) (5.14) F (a, aL) = F (a, aR). Proof. The anti-periodicity (5.10) is manifest from the construction, s-holomorphicity follows from linearity, and (5.11–5.12) are obtained by simply summing (5.6–5.9). The fact that F (00) is essentially a constant is a discrete analog of the claim that a meromorphic function on a torus with at most one simple pole is constant. In- (00) 0 deed, consider the restriction of F (a, ·) onto the index 2 sub-lattice C (Tδ) := iπ/4 −iπ/4 {z ∈ C(Tδ): ηz ∈ e ηaR ∪ e ηaR} of C(Tδ). As noticed above, this restric- tion is discrete holomorphic, that is, the identity (5.5) holds for every z ∈ C(Tδ) such that ηz ∈ ηaR ∪ iηaR, except, possibly, for z = a. However, summing (5.5) 0 over all z, we see that each value F (z), z ∈ C (Tδ), enters the sum with the co- efficient 1 − 1 + i − i = 0. Therefore, (5.5) also holds for z = a. Since the restriction of a discrete holomorphic function F (00)(a, ·) to each of the two sub- ±iπ/4 lattices {z ∈ C(Tδ): ηz ∈ e ηaR} is discrete harmonic, the maximum prin- ciple implies that these restrictions are constant. The s-holomorphicity condi- tion (5.4) now implies that the restrictions of F (00)(a, ·) to two other sub-lattices {z ∈ C(Tδ): ηz = ηaR} and {z ∈ C(Tδ): ηz = iηaR} are also constant (and it’s easy (00) to see that the s-holomorphicity implies that in that case, F (a, z) = Projηz (C). Finally, since F (a, a+) = −F (a, a−), it is a simple direct check that the only value + consistent with (5.5) vanishing at a is F (a, a ) = 0, hence F (a, z) ≡ 0 if ηz = ±ηa, implying the restriction C ∈ iηaR.  17 1 We now describe the discrete analog of the Cauchy kernel z−a , tailored for our 2 purposes. Given a ∈ C(Z ) and a choice of the square root at ηa this is a unique 2 s-holomorphic function Pa(·) on C[a](Z ) such that + − (5.15) Pa(a ) = ηa,Pa(a ) = −ηa, and, as z → ∞, one has √ 2   (5.16) P (z) = Proj η¯ (z − a)−1 + O(|z − a|−2). a π ηz a Moreover, we have

(5.17) Pa(aL) = Pa(aR) = 0. We also define the re-scaled version of this function, living on C(δZ2): δ −1 −1 Pa (z) = δ Pδ−1a(δ z).

The function Pa is a multiple of the discrete Cauchy kernel on the square√ lattice; see ([12, Lemma 4.9]) for the details. Since in ([12]), the lattice is scaled by 2 and √ π iπ  i π  rotated by , one must re-define P (z) := e 8 P√ i π 2e 4 z , which is reflected in 4 a 2ae 4 the pre-factor in (5.16). The following lemma elucidates the singularity structure of F (ij) (a, z ). e1...ek pq Lemma 18. The function " k # (5.18) F˜ (·) = F (ij) (a, ·) − δ · µ(ij) Y  P δ(·) a e1...ek E ei a i=1 δ extends to an s-holomorphic function on C(T ) in a neighborhood of a. If z2i−1, z2i are two opposite corners incident to ei, then (5.19) F˜ (·) = F (ij) (a, ·) m e1...ek

δ (ij) δ δ (ij) δ − √ · iη¯z F (a, z2m)P (·) − √ · iη¯z F (a, z2m−1)P (·) 2 2m−1 e1...eˆm...ek z2m−1 2 2m e1...eˆm...ek z2m δ extends to an s-holomorphic function on C(T ) in a neighborhood of em, where the δ δ choice of square root in ηz2m−1 and Pz2m−1 (·) (resp. ηz2m and Pz2m (·)) are related by (5.15). Moreover, after that extension, ˜ ˜ ˜ (5.20) F (a) = Fm(z2m−1) = Fm(z2m) = 0. ˜ ± Proof. It follows from (5.11) and (5.15) that Fa(a ) = 0, and hence these two corners [e1...ek] δ of C[a] (T ) can be glued back together, yielding the first claim. For the second claim, we use that

−2βσz◦ σz◦ σz◦ σz◦ µz• µz• = σz◦ σz◦ e 2m−1 2m 2m−1 2m 2m−1 2m 2m−1 2m √ = σ ◦ σ ◦ cosh(−2β) + sinh(−2β) = 2 , z2m−1 z2m em − 1 so that we can replace  by 2 2 σ ◦ σ ◦ µ • µ • in the definition of F (a, ·). m z2m−1 z2m z2m−1 z2m e1...ek ˆ δ ◦ • ◦ • With this substitution, it becomes a spinor on C[a](T ) ramified at z2m−1, z2m−1, z2m, z2m, δ δ δ δ and their shifts by ω1, ω2, ω1 + ω2. We modify the graph by introducing branch cuts ◦ • ◦ • [z2m−1, z2m−1] and [z2m, z2m] (and their shifts) which splits the vertices z2m−1, z2m 18 ± ± into z2m−1 and z2m, so that Fe1...ek (a, ·) becomes a function on the resulting graph + − obeying the property Fe1...ek (a, zm) = −Fe1...ek (a, zm). More concretely,   k + 1  Y  √ + ◦ • • • Fe1...ek (a, z2m−1) = ηz E σz µz σa µa ei  2m−1  2m 2m  2 i=1 i6=m

1 (ij) + = √ ηz η¯z2m Fe ...eˆ ...e (a, z2m); 2 2m−1 1 m k

  k + 1  Y  √ + ◦ • • • Fe1...ek (a, z2m) = ηz E σz µz σa µa ei  2m  2m−1 2m−1  2 i=1 i6=m

1 (ij) + = √ ηz η¯z2m Fe ...eˆ ...e (a, z2m−1). 2 2m−1 1 m k

+ + + Here the signs are related in such a way that η η¯z2m = η η¯z2m = i if z2m−1 z2m−1 z2m−1 + and z2m are on the outer side from the edge em. Therefore, taking into account δ δ ˜ ± ˜ ± that Pz2m−1 (z2m) = 0 and Pz2m (z2m−1) = 0, we see that Fm(z2m−1) = Fm(z2m) = 0, + − ± so that z2m−1 and z2m−1 can be glued back together, and similarly for z2m. These δ δ δ δ results hold with the same proof for shifts of em by ω1, ω2, ω1 + ω2 and hence extend (ij) ± to F (a, z ) by linearity, thus proving the claim.  Remark 19. For (ij) 6= (00), assuming Eµ(ij) are known, the formulae (5.18–5.19) identify the functions F (ij) uniquely by recursion. They do not identify uniquely e1...ek F (00) , because of a possibility of adding a “constant” s-holomorphic function Proj (c) , e1...ek ηz c ∈ C. However, if k ≥ 1, (5.20) removes this degree of freedom. We we were unable to identify this constant and its asymptotics for k = 0 based on discrete holomor- phicity considerations only, hence the input from Corollary 2 is needed to start the induction. For the energy difference of Theorem 3, the value of this is constant does h (00) i h (00) i not matter as it cancels out, since E µ V = E µ H by (5.12, 5.14).

We conclude this section by identifying Eµ(ij) with Kac-Ward determinants from Section 3, and deducing their asymptotics: Lemma 20. We have Z(ij) (µ(ij)) = , E Z(01) + Z(10) + Z(11) where Z(ij) is defined in (3.4). Therefore, we have, as δ → 0,

(10) (10) (10) Z |θ4| (5.21) Eµ → E := = , Z |θ2| + |θ3| + |θ4| (01) (01) (01) Z |θ2| (5.22) Eµ → E := = , Z |θ2| + |θ3| + |θ4| (11) (11) (11) Z |θ3| (5.23) Eµ → E := = . Z |θ2| + |θ3| + |θ4| 19 Proof. We prove the identities using high-temperature expansion; since Tδ is self- dual, we can freely pass between primal and dual lattice. For x, y ∈ (Tδ)?, x ∼ y, recall the notation (3.2). The high-temperature expansion reads   P P ϕpq(xy) σ exp β x∼y(−1) σxσy µ = E pq P  P  σ exp β x∼y σxσy   P Q ϕpq(xy) P ϕpq(ξ) |ξ| 1 + (−1) σxσyα δ σ x∼y ξ∈E(T )(−1) α = P Q = P |ξ| . σ x∼y (1 + σxσyα) ξ∈E(Tδ) α (ij) To compute Eµ , we recall that ϕpq(ξ) = pϕ10(ξ) + qϕ01(ξ) mod 2 and note that X (−1)(1−i)p+(1−j)q+pq+pϕ10(ξ)+qϕ01(ξ) p,q X = (−1)qij (ξ)+(1−i)(1−j) (−1)(p+ϕ01(ξ)+1−j)(q+ϕ10(ξ)+1−i) p,q X = (−1)qij (ξ)+(1−i)(1−j) (−1)pq = 2(−1)qij (ξ)(−1)(1−i)(1−j). p,q

Therefore, plugging the above formula for Eµpq into (5.13) and taking into account (3.5) yields

P qij (ξ) |ξ| (ij) 1 2 δ (−1) α 1 2Z µ(ij) = (−1)(1−i)(1−j) ξ∈E(T ) = , E P |ξ| 1 (01) (10) (11) 4 ξ∈E(Tδ) α 4 2 (Z + Z + Z ) where we have used that Z(00) = 0 and (−1)(1−i)(1−j) = 1 unless i = j = 0. For √ δ q (ij) ij −|T |/2 −|Tδ| ij the asymptotics, we use Z = KW = 2 αc det ∆δ , and then use asymptotics of these determinants computed in the course of the proof of Corollary 2. 

6. Scaling limits of the fermionic observables In what follow, we define the continuous limits of the observable F (ij) (a, z). e1...ek These limits will depend on ηa, i. e., the orientation of the corner a and the choice of the sign of the square root in (5.2). Since there are only 8 options for ηa, we will from now on assume it fixed. It will be convenient to use physics notation for the Pfaffian: for symbols O1,. . . ,O2N , and a label  to distinguish between different anti-periodicity “sectors”, if a 2N ×2N  antisymmetric matrix M is given whose entries are denoted by Mn,m = hOnOmi , we denote   hO1 ... O2ni := Pf M = Pf hOnOmi1≤n,m≤2N .

Recall that, given ω1,2, the Weierstrass ζ-function ζω1,ω2 is the unique odd function that has a simple pole of residue 1 at the origin, and such that its derivative is doubly periodic (in fact, −ζ0(z) = ℘(z), where ℘(z) is the Weierstrass ℘ function). Thus, ζ(z) is not an elliptic function, but has periodicity property ζ(z + ω12) = ζ(z) + c12 P for some constants c12. However, a linear combination αiζ(z − βi) is an elliptic P function provided that αi = 0. We also denote by csω1,ω2 (z), nsω1,ω2 (z), dsω1,ω2 (z) the unique meromorphic function on Tˆ with simple pole of residue 1 at the origin and satisfying the anti-periodicity relations (6.7) below for (ij) = (01), (10), (11) 2K  2K  respectively. We have csω ,ω (z) = · cs (z − a) , k in the notation of [19, 1 2 ω1 ω1 20 Section 22], and similarly for ns, ds, where the elliptic modulus k and the complete elliptic integral K are given by !2 θ2(τ) π 2 k := ,K = θ3(τ). θ3(τ) 2

Definition 21. Given a continuous torus T and distinct points a, e1, . . . , ek ∈ T, for (ij) 6= (00), we define  Z(ij) k ? ? (ij)  ηa(−i) hψe ψ . . . ψe ψ ψzψai k even, (6.1) f (ij) (a, z) = Z 1 e1 k ek e1...ek Z(ij) η (−i)khψ ψ? . . . ψ ψ? ψ ψ?i(ij) k odd,  Z a e1 e1 ek ek z a

? ? ? where hψwψwˆi ≡ 0, hψwψwˆi = hψwψwˆi, and (01) (6.2) hψwψwˆi = csω1,ω2 (w − wˆ), (10) (6.3) hψwψwˆi = nsω1,ω2 (w − wˆ), (11) (6.4) hψwψwˆi = dsω1,ω2 (w − wˆ). Definition 22. Given data as above, we also define  k ? ? (00) ηa(−i) hψe ψ . . . ψe ψ ψzψai , k odd, (6.5) f (00) (a, z) = 1 e1 k ek e1...ek η¯ (−i)khψ ψ? . . . ψ ψ? ψ ψ?i(00), k even,  a e1 e1 ek ek z a

? ? (00) where hψen ψem i = ζ(en − em),

1 ? (00) (=mτ) 2 |θ2θ3θ4| 1 (6.6) hψen ψem i ≡ πi · 1 , |θ2| + |θ3| + |θ4| |T| 2 (00) and hψen ψem i = ζ(en − em); here ζ denotes the Weierstrass ζ-function. Definition 23. Given data as above, we define for (ij) 6= (00),

 (ij)  Z (−i)khψ ψ? . . . ψ ψ? i(ij) k even, (ij) Z e1 e1 ek ek Ee ...e := 1 k 0 k odd, and  (−i)khψ ψ? . . . ψ ψ? i(00), k odd, (00) e1 e1 ek ek Ee ...e = 1 k 0 k even.

Proposition 24. The quantity f (ij) (a, ·) is a meromorphic function on ˆ satisfy- e1...ek T ing, for p, q ∈ {0, 1}, (6.7) f (ij) (a, z + pω + qω ) = (−1)ip+jqf (ij) (a, z) e1...ek 1 2 e1...ek

Its poles are simple and located at e1, . . . , ek, a, and

(ij) f (a, em) (6.8) f (ij) (a, ·) = −i e1...eˆm...ek + o(1), z → e , e1...ek m z − em η E (ij) (ij) a e1...ek (ij) (6.9) f (a, z) = + iηa ·E + o(1), z → a. e1...ek z − a e1...eka 21 Remark 25. The equations (6.7–6.9) give an overdetermined set of conditions that identify f (ij) (a, z) and E (ij) uniquely by induction, given E (ij) for (ij) 6= (00) e1...ek e1...ek and a constant E (00). Indeed, to see that f (ij) (a, ·) is uniquely determined, note e e1...ek that if two functions f, g both satisfy (6.7–6.9), then, by induction hypothesis, their ˆ difference is holomorphic everywhere on T and vanishes, say, at e1, hence it is zero. In its turn, f (ij) uniquely determines E (ij) by (6.9). For (ij) 6= (00), the vanishing e1...ek e1...ek+1 of constant terms in (6.8) is not needed to prove uniqueness; for (ij) = (00), we (00) (00) need it just for one m. We note that f (a, z) = ηaiEe (it is an elliptic function with at most one simple pole, hence a constant), and

(ij) (ij) (ij) f =η ¯aE hψzψai , where the latter is given by (6.2–6.4)

 −1 Proof. In all cases, hψzψwi = (z − w) + o(1), since it is an odd function of z − w with a simple pole of residue 1 at the origin; also, complex conjugating the Pfaffian ? amounts to replacing ψ ←→ ψ . Expanding the Pfaffian, we obtain as z → e1 and  using the convention hψe1 ψe1 i = 0, we have

hψ ψ? . . . ψ ψ? ψ ψ i e1 e1 ek ek z a = hψ ψ ihψ? . . . ψ ψ? ψ i + hψ ψ? . . . ψ ψ? ψ ψ i + o(1) z e1 e1 ek ek a e1 e1 ek−1 ek−1 e1 a = (z − e )−1hψ? . . . ψ ψ? ψ i + o(1) 1 e1 ek ek a = (z − e )−1(−1)k−1hψ . . . ψ ψ? ψ?i + o(1), 1 e1 ek ek a ? and similarly for m = 2, . . . , k and with ψa replaced with ψa. This proves (6.8). Similarly, as z → a, we have hψ ψ? . . . ψ ψ? ψ ψ i = hψ ψ ihψ? . . . ψ ψ? i + o(1), e1 e1 ek ek z a z a e1 ek ek hψ ψ? . . . ψ ψ? ψ ψ?i = hψ ψ? . . . ψ ψ? ψ ψ?i + o(1). e1 e1 ek ek z a e1 e1 ek ek a a The first identity proves (6.9) for (ij) 6= (00), k even and for (ij) = 0, k odd, while the second identity proves it for (ij) 6= 00, k odd and for (ij) = (00), k even. To prove (6.7) for (ij) 6= (00), simply expand the Pfaffian in ψz and note that (ij) hψzψwi satisfies (6.7) for each w = e1, . . . , ek, a. For (ij) = (00), the Weierstrass ζ-function does not satisfy (6.7); as discussed above, we need to check that the sum of the residues of f (00) (a, ·) is zero. We claim that this result follows by induction e1...ek (00) from the existence of any solution to (6.7–6.9) with iEe given by (6.6). Indeed, suppose f˜(00) (a, z) is such a solution, and suppose by induction hypothesis that e1...ek f (00) (a, z) = f˜(00) (a, z) for any e , . . . , e , a. Then, by (6.8–6.9), the residues e1...ek−1 e1...ek−1 1 k−1 of f (00) (a, z) and f˜(00) are the same; since the latter function satisfies (6.7), their e1...ek e1...ek sum vanishes; hence also f (00) (a, z) satisfies (6.7), and we have f (00) ≡ f˜(00) , e1...ek e1...ek e1...ek completing the induction step. The solution to (6.7–6.9) exists since it is constructed as the limit of discrete observables in the proof of Theorem 26 below.  At the heart of our analysis is the following convergence result: Theorem 26. One has, as δ → 0,   (6.10) δ−(k+1)F (ij) (a, z) = C · Proj η(k)f (ij) (a, z) + o(1), e1...ek k ηz e1...ek 22 √ 2 uniformly in z, e1, . . . , ek, a away from each other, where Ck = πk+1 . Moreover, " k # (6.11) δ−k µ(ij) Y  → π−kE (ij) . E em e1...ek m=1

Proof. We prove this result by induction: (6.11)k =⇒(6.10)k =⇒(6.11)k+1, first separately for each (ij) 6= (00) and then for (ij) = (00). For the base of induction, for (ij) 6= 0, we have (5.21–5.23), and for (ij) = (00), the induction starts with k = 1. Namely, we have −1 h (00) i −1 X −1 h (ij) i −1 (00) −1 X (ij) δ E µ e = δ Ee − δ E µ e → π Ee − π Ee , (ij)6=(00) (ij)6=(00) where the convergence of the first term is by Corollary (4) and the definition of (00) Ee , and the convergence of other three terms will by that point have been proven. (ij) It remains to notice that Ee ≡ 0 for (ij) 6= 0. Note that the proof of Corollary (4) relies on Theorem (3) whose proof below does use (6.10), but only in the case (ij) 6= (00). Thus, the base of induction is established. To prove (6.10), we follow a general scheme used in ([12]), where we first identify the scaling limit assuming precompactness, and then justify precompactness. Let us first assume that for all r > 0, there is a constant C such that the functions r δ−(k+1)F (ij) (a, ·) are bounded by C on the set := \ B (a) ∪ B (e ) ∪ · · · ∪ e1...ek r Tr T r r 1 Br(ek), uniformly in δ. We claim that in this case,   (6.12) δ−(k+1)F (ij) (a, ·) = C · Proj f (ij) (a, z) + o(1) e1...ek k ηz e1...ek ˆ uniformly on compact subsets of T\{apq}. Indeed, as noted above, the s-holomorphicity of F (ij) (a, z), together with uniform boundedness, implies that (6.12) holds along e1...ek ˆ a subsequence, and with some holomorphic function f : T \{apq} → C instead of f (ij) (a, z). Therefore, it is enough to show that any sub-sequential limit f must e1...ek be equal to f (ij) (a, z). Clearly, f must satisfy the (anti)-periodicity condition e1...ek ip+jq f(z + pω1 + qω2) = (−1) f(z), because of (5.10). Denote U = f (ij) (a, e ). We have by induction hypothesis e1...eˆm...ek m −k (ij) δ F (a, z2m) = Ck−1Proj (U) + o(1), e1...eˆm...ek ηz2m −k (ij) δ F (a, z2m−1) = Ck−1Proj (U) + o(1), e1...eˆm...ek ηz2m−1 1 2 so that, taking into account (5.16) and the identity Projη (U) = 2 (U + η U), in a fixed annulus around BR(em) \ Br(em), we have, for the “corrective” term in the ˜ definition of Fk(·):

−k−1 δ (ij) δ −k−1 (ij) δ iη¯z δ √ F (a, z2m)P (z)+iη¯z δ F (a, z2m−1)P (z) 2m−1 2 e1...eˆm...ek z2m−1 2m e1...eˆm...ek z2m   1 η¯z2m−1 = Ck−1 · iη¯z2m−1 Projηz (U) Projηz π 2m z − em   1 η¯z2m + Ck−1 · iη¯z2m Projηz (U) Projηz + o(1) π 2m−1 z − em 1 iU¯ ! = Ck−1 · · Projηz + o(1). π z − em 23 −(k+1) ˜ Recall δ Fm(·) (5.19) extends to an s-holomorphic function near em. We can express its values by discrete Cauchy integral formula with contour in BR(em) \ B (e ) and pass to the limit in that formula; this shows that f(z) − C · iU¯ r m k z−em extends to a holomorphic function at B2r(em); moreover, (5.20) shows that this function vanishes at em. Hence, f must satisfy (6.8). h i The analysis near a is similar. Since δ−k · µ(ij) Qk  = E (ij) + o(1) by E i=1 ei e1...ek induction hypothesis, we have the expansion √ " k # 2  η¯  δ−k µ(ij) Y  P δ(·) = E (ij) · Proj a + o(1). E ei a k+1 e1...ek ηz i=1 π z − a uniformly in BR(em) \ Br(em). Therefore, by the same argument as above, f(z) − E(ij) e1...ek ηa z−a can be analytically continued in BR(a), and thus satisfies (6.9). In other words, f(·) satisfies the defining conditions of f (ij) (a, ·), and therefore, due to e1...ek Remark 25, we have f(·) ≡ f (ij) (a, ·). e1...ek We now turn to justifying the uniform (in δ) boundedness away form a,e1, . . . , ek. Let us assume towards a contradiction that there exist a (small fixed) R > 0 such δ −1−k (ij) that C := max ˆδ δ F (a, ·) tends to infinity (at least along some se- R z∈T e1...ek R quence of δ). We claim that in that case, the functions (Cδ )−1δ−1−kF (ij) (a, ·) are R e1...ek ˆ δ uniformly bounded on any Tr with r < R. Indeed, " k # ! (Cδ)−1 δ−1−kF (ij) (a, ·) − δ−k−1 µ(ij) Y  P δ(·) r e1...ek E ei a i=1 is uniformly bounded on B(a, 2R) \ B(a, R) and discrete holomorphic in B(a, 2R), and hence, by maximum principle, it is uniformly bounded in the whole of B(a, 2R). δ As Pa (z) is also uniformly bounded on compact subsets of B(a, 2R) \{a} and δ −1 (Cr ) → 0, we get the claim. Similarly, we justify the uniform boundedness on each BR(em) \ Br(em) for m = 1, . . . , k. Therefore, the above convergence argument can be applied verbatim to (Cδ )−1δ−1−kF (ij) (a, ·) R e1...ek with the conclusion that it converges uniformly on compact subsets of Tˆ to a function ˆ that is analytic in the whole T and vanishing, say, at e1, that is, to 0. This contradicts δ the choice of Cr . ˜ To derive (6.11)k+1, note that we have shown above that Fa(·) extends to an s-holomorphic function in a neighborhood of a, and that   −(k+1) ˜ (ij) (ij) η¯a δ Fa(z) = Ck · Proj f (a, z) − E + o(1), ηz e1...ek e1...ek z − a uniformly in a neighborhood of a, where the function inside the projection is an- alytically continued. It remains to take into account (5.8–5.9) and (5.17) to note that √ " k # ˜ (ij) Y Fa(aL,R) = iηa · 2 · E µ eL,R m . m=1 .  24 Remark 27. The planar domain counterpart of this theorem is a particular case of ([12, Theorem 2]), asserting that

−k−1 ηa k 2 ηa (6.13) δ E [e1 . . . ek ψa ψz] = C CψηaProjηz (he1 . . . ek ψa ψzi) + o(1).

The expansions of the right-hand side at z = em and z = a can be read off the fusion rules [12, (6.2–6.4)]. In the case (ij) 6= (00), the same convergence proof as in [12], by expressing F (ij) (a, z) as a Pfaffian of two-point correlations, and then passing to e1...ek the limit term-by-term, could have been carried on. We were unable to do it for the (00) (00) sector, due to the lack of combinatorial analog of hψzψai . By the formalism ηa ? of [12, Section 5.2], in the RHS of (6.13) we can expand ψa =η ¯aψa + ηaψa; in the torus case, only one of these terms will contribute for each sector depending on pairity of k, which is reflected in (6.1, 6.5)

7. Proofs of Theorem 3 and Theorem 5. In order to prove Theorem 3, we need the one more Lemma. Denote E (ij)η¯ g(ij)(a, z) := f (ij)(a, z) − a ; z − a (ij) −1 (ij) h (ij)i δ G (a, z) := δ F (a, z) − E µ Pa (z). It is a standard fact that if discrete holomorphic functions converge uniformly in a ball B(a, r), then so do their discretized derivatives. This is what it means for our case: Lemma 28. One has, as δ → 0, for (ij) 6= (00),

−2  (ij) (ij)  2 h 3 (ij) i δ F (a, aL) − F (a, aR) → iηa

(ij) (ij) G (a, aL) − G (a, aR) √ 2   (ij)   (ij)   = aLR ∂zProj g (a, z) + aLR ∂zProj g (a, z) + o(δ). π iηa z=a iηa z=a √ 1 2 2 (ij) Since Projη (x) = 2 (x + η x¯) , aLR = − 2δη¯a, and g (a, ·) is holomorphic, using (ij) the notation X = ∂zg (a, z) , this can be further simplified as z=a     a ∂ Proj g(ij)(a, z) + a ∂ Proj g(ij)(a, z) LR z iηa LR z iηa √ z=a z=a √ 2  2 4 ¯  1  3 ¯ 3  = − δ η¯aX − ηaX = 2δiηa ηaX − η¯aX 2 √ 2i √ 3 ¯ 3 = 2δiηa=m[ηaX] = 2δiηa

We are in the position to prove Theorem 3:

(ij) Proof of Theorem 3. Note that the correlations E[µ (V − H )] can be expressed, using (5.12), as

h (ij) i 1  (ij) (ij)  (7.1) E µ (eL − eR ) = √ F (a, aL) − F (a, aR) . i 2ηa

a•−a◦ π i 4 In fact, this is zero for (ij) = (00) by (5.14). Choose a so that |a•−a◦| = e , i. e., π i 8 P (1−i)p+(1−j)q+pq ηa = e ; then eL = V , eR = H . Since i,j∈{0,1}(−1) = 4Ip=q=0 we have

X (ij) (7.2) µ = µ00 = 1. (ij)

Therefore, summing (7.1) and applying Lemma 28 yields

−2 2 X δ  (ij) (ij)  (7.3) E(eV ) − E(eH ) = δ √ F (a, aL) − F (a, aR) . (ij)6=(00) i 2ηa √ 2 2 X  h 3 (ij) i  = δ · ·

It remains to compute explicitly the derivatives. Jacobian elliptic functions have series expansions

1 −1 1    cs (u, k) = + + k2 u + O u3 , u 3 6 1 1     ns (u, k) = + 1 + k2 u + O u3 , u 6 1 1 1    ds (u, k) = + − k2 u + O u3 u 6 3 and plugging them into the formula for f (ij)(a, ·) yields

 1 2 1  4 4  k − θ − 2θ , (ij) = (01), 4K2  6 3 π2  2 3 (ij) (ij) 1 2 (ij) 4 4 ∂zg (a, z) =η ¯aE 2 · (1 + k ) =η ¯aE 2 · θ2 + θ3, (ij) = (10), z=a ω  6 24ω  1  1 1 2 1  4 4 6 − 3 k θ3 − 2θ2, (ij) = (11).

Plugging into (7.3) concludes the proof. 

Proof of Theorem 5. By (7.2) and (6.11), we have

" k # " k # δ−k Y  = X δ−k µ(ij) Y  → π−k X E (ij) . E m E m e1...ek m=1 (ij) m=1 (ij) 26 By Definition 23, only E (00) contributes for odd k; chasing the definitions leads e1...ek directly to (2.5). For even k, only E (01) , E (10) , E (11) contribute, and we have e1...ek e1...ek e1...ek

(ij) (ij) Z k ? ? (ij) E = (−1) 2 hψe ψ . . . ψe ψ i e1...ek Z 1 e1 k ek (ij) Z k k(k−1) ? ? (ij) = (−1) 2 (−1) 2 hψe . . . ψe ψ . . . ψ i Z 1 k e1 ek (ij) (ij) Z (ij) ? ? (ij) Z (ij) 2 = hψe . . . ψe i hψ . . . ψ i = hψe . . . ψe i . Z 1 k e1 ek Z 1 k  8. Theorem 1 for the triangular lattice In this section, we prove an analog of Theorem 1 for triangular lattice H := iπ {m + e 3 n : m, n ∈ Z}; a similar argument (or duality) can be applied to treat the hexagonal lattice. We follow the argument for the case of the square lattice. The energy observable on the triangular lattice is defined by

(xy) = σxσy − ,¯ 2 where ¯ = 3 . There are three types of edges on the lattice, of which we choose πi k representatives e0, e1, e2, where ek is aligned with e 3 . Our result reads as follows: Theorem 29. For the critical Ising model on defined Tδ = δH/Λδ, we have q ? 00 √ det ∆δ 1 Ee0 + Ee1 + Ee2 = 6 2q q q . 10 01 11 | δ| det ∆δ + det ∆δ + det ∆δ T where det? denotes the product of all non-zero eigenvalues. Remark 30. Since the asymptotics of the determinant of the Laplacian has been worked out for arbitrary doubly-periodic lattices [35], this does lead immediately to the analog of Corollary 2, obtained by other methods in [51]. Other part of the paper extend to H as well: the discrete holomorphicity techniques of Section 5 are known to extend to a larger class of isoradial lattices with critical weights, see [11, 10]. As in the square lattice case, we start with the high-temperature expansion iden- tity 1 (8.1)  = X b(e, ξ)α|ξ|, E e ZI ξ∈E(Tδ) −1 where b(e, ξ) = (α − ¯)Ie∈ξ + (α − ¯)Ie/∈ξ. Denote the triangular lattice by H, and define vtri(α, q) := (1 + 3α2 + 8α3 + 3α4 + α6)  1 1 z w  + (3α3 − α − α5) z + w + + + + , z w w z where, as in section 3, z (q) = exp (2πi

Lemma 31. One has, for α = αtri,  δ √ 1−|T | √  2 ? 00 1 (ij) (ij) (ij) 12(11 − 5 3)ctri det ∆ , i = j = 0; B (e0) + B (e1) + B (e2) = |Tδ| 0, otherwise. Proof. As in the square lattice case, we have δ d | | X (8.2) Z(ij) = T ( + + ) α|ξ|(−1)qij (ξ) dα α Ie0∈ξ Ie1∈ξ Ie2∈ξ ξ∈E(Tδ) The following identities are straightforward to check: ! d tri 1 tri (8.3) v (α, q) ≡ 2 + √ v (αtri, q). dα 3 α=αtri

d2 √ (8.4) vtri(α, 0) = 144(2 − 3). dα2 α=αtri on the other hand, for (ij) 6= (00), we have KWij 6= 0, and therefore (8.3) gives

(8.5) q d (ij) d ij 1 d ij c δ Z = KW = √ KW = | |Z ij T ij dα dα dα 2 α=αtri α=αtri α=αtri 2 KW α=αtri where c = 2 + √1 . Subtracting (8.2) and (8.5), we get 3   X ˆ ˆ ˆ |ξ| qij (ξ) b(e0, ξ) + b(e1, ξ) + b(e2, ξ) α (−1) = 0, ξ∈E(Tδ) ˆ  c 1  c where b(e, ξ) = − Ie∈ξ + Ie/∈ξ. It is now a matter of simple algebra to check 6 √α 6 that b(e, ξ) = (6 − 4 3)ˆb(e, ξ), concluding the proof in the case (i, j) 6= (0, 0). (00) c δ (00) For (i, j) = (0, 0), we have Z = 0, therefore, subtracting 0 = 2 |T |Z from the right-hand side of (8.2) yields

d   |ξ| X ˆ ˆ ˆ qij (ξ) Zij = − b(e0, ξ) + b(e1, ξ) + b(e2, ξ) αtri(−1) . dα δ α=αtri ξ∈E(T )

q 00 Now, since Z00(α) = KW (α), and because of (8.3–8.4), we have 1 √ KW00(α) = 144(2 − 3)(α − α )2 Y v(α, q) + o((α − α )2), α → α , 2 c c c q∈H/Λ∗\{0} 28 therefore, we arrive at

1 1 s √   2   2 d 3 √ Y Y Zij = 12 1 −  v(α, q) = 6( 3−1)  v(α, q) . dα 2 ∗ ∗ α=αtri q∈H/Λ \{0} q∈H/Λ \{0} √ δ 1−|T | q ? 2 00 = 6( 3 − 1)c det ∆δ . Putting everything together, we arrive at √ √ δ X |ξ| 1−|T | q ? qij (ξ) 2 00 (b(e0, ξ) + b(e1, ξ) + b(e2, ξ)) αtri(−1) = (4 3−6)6( 3−1)c det ∆δ , ξ∈E(Tδ) as required.  Proof of Theorem 29. As in the the case of the square lattice, (8.1) implies that 1  = (B01(e) + B(10)(e) + B(11)(e) − B(00)(e)), E e 2ZI and summing this over e0, e1, e2 and applying Lemma 31 yields δ 1 √ 1−|T | √ 1 2 ? 00 Ee0 + Ee1 + Ee2 = I 12(11 − 5 3)ctri det ∆ . 2Z |Tδ| Now, we can recall that q q q I 01 10 11 2Z = Z01 + Z10 + Z11 = det KW + det KW + det KW q q q  −|Tδ| 01 10 11 = ctri det ∆δ + det ∆δ + det ∆δ , and putting all together, after some tedious algebra to simplify the constant in front, yields the result.  References [1] Luis Alvarez-Gaumé, Gregory Moore, and Cumrun Vafa. Theta functions, modular invariance, and strings. Communications in Mathematical Physics, 106(1):1–40, 1986. 1 [2] Giovanni Antinucci, Alessandro Giuliani, and Rafael Leon Greenblatt. Energy correla- tions of non-integrable ising models: The scaling limit in the cylinder. arXiv preprint arXiv:2006.04458, 2020. 1 [3] A.A. Belavin, A.M. Polyakov, and A.B. Zamolodchikov. Infinite conformal symmetry in two- dimensional quantum field theory. Nuclear Physics B, 241(2):333–380, 1984. 1 [4] Cédric Boutillier and Béatrice de Tilière. The critical Z-invariant Ising model via dimers: the periodic case. Probability Theory and Related Fields, 147(3-4):379–413, 2010. 1 [5] Cédric Boutillier and Béatrice de Tilière. The Critical Z-Invariant Ising Model via Dimers: Locality Property. Communications in Mathematical Physics, 301(2):473–516, 2011. 1 [6] D. Chelkak, D. Cimasoni, and A. Kassel. Revisiting the combinatorics of the 2D Ising model. ArXiv e-prints, July 2015. 1, 3, 3, 3 [7] Dmitry Chelkak. Ising model and s-embeddings of planar graphs. arXiv preprint arXiv:2006.14559, 2020. 1 [8] Dmitry Chelkak, Alexander Glazman, and Stanislav Smirnov. Discrete stress-energy tensor in the loop O(n) model. arXiv:1604.06339, 2016. 2.2 [9] Dmitry Chelkak, Clément Hongler, and Konstantin Izyurov. Conformal invariance of spin correlations in the planar ising model. Annals of Mathematics, 181:1087–1138, 2015. 1 [10] Dmitry Chelkak, Konstantin Izyurov, and Remy Mahfouf. Universality of spin correlations in the ising model on isoradial graphs. in preparation, 2021. 1, 30 [11] Dmitry Chelkak and Stanislav Smirnov. Universality in the 2D Ising model and conformal invariance of fermionic observables. Inventiones mathematicae, 189(3):515–580, 2012. 1, 30 29 [12] Chelkak, Dmitry and Hongler, Clément and Izyurov, Konstantin. Correlations of primary fields in the crutical Ising model. 1, 5, 5, 5, 5, 6, 27, 27 [13] G. Chinta, J. Jorgenson, and A. Karlsson. Complexity and heights of tori. In L. Bowen, R. Grigorchuk, and Y. Vorobets, editors, Dynamical Systems and Group Actions, volume 567 of Contemporary Mathematics, pages 89–98. American Mathematical Society, 2012. 2.2, 12 [14] David Cimasoni. A generalized Kac-Ward formula. Journal of Statistical Mechanics: Theory and Experiment, 2010(07):P07023, 2010. 1, 3 [15] David Cimasoni. The Critical Ising Model via Kac-Ward Matrices. Communications in Math- ematical Physics, 316(1):99–126, 2012. 1, 2.2, 3 [16] David Cimasoni and Hugo Duminil-Copin. The critical temperature for the ising model on planar doubly periodic graphs. Electron. J. Probab., 18:18 pp., 2013. 3, 4 [17] P. Di Francesco, P. Mathieu, and D. Senechal. Conformal Field Theory. Graduate Texts in Contemporary Physics. Springer-Verlag, New York, 1997. 2.2, 2.2, 4 [18] Ph Di Francesco, H Saleur, and JB Zuber. Critical ising correlation functions in the plane and on the torus. Nuclear Physics B, 290:527–581, 1987. 2.2 [19] NIST Digital Library of Mathematical Functions. F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds. 2, 4, 6 [20] N. P. Dolbilin, Yu. M. Zinov’ev, A. S. Mishchenko, M. A. Shtan’ko, and Mikhail I. Shtogrin. The two-dimensional Ising model and the Kac-Ward determinant. Izvestiya: Mathematics, 63(4):707, 1999. 1, 3 [21] Julien Dubédat. Exact bosonization of the ising model. arXiv preprint arXiv:1112.4399, 2011. 2.2 [22] Giovanni Felder. BRST approach to minimal models. Nuclear Physics B, 317(1):215–236, 1989. 2.2 [23] Arthur E. Ferdinand and Michael E. Fisher. Bounded and inhomogeneous ising models. i. specific-heat anomaly of a finite lattice. Phys. Rev., 185:832–846, Sep 1969. 1, 2.2, 2.2 [24] Siarhei Finski. Finite difference method on flat surfaces with a flat unitary vector bundle. arXiv preprint arXiv:2001.04862, 2020. 1 [25] Siarhei Finski. Spanning trees, cycle-rooted spanning forests on discretizations of flat surfaces and analytic torsion. arXiv preprint arXiv:2001.05162, 2020. 1 [26] Michael E. Fisher. Statistical mechanics of dimers on a plane lattice. Phys. Rev., 124:1664– 1672, Dec 1961. 1 [27] Michael E. Fisher. On the Dimer Solution of Planar Ising Models. Journal of Mathematical Physics, 7(10):1776–1781, 1966. 1 [28] P. Di Francesco, H. Saleur, and J.B. Zuber. Critical Ising correlation functions in the plane and on the torus. Nuclear Physics B, 290:527–581, 1987. 2.2 [29] Alessandro Giuliani, Rafael L Greenblatt, and Vieri Mastropietro. The scaling limit of the energy correlations in non-integrable ising models. Journal of mathematical physics, 53(9):095214, 2012. 1 [30] Clement Hongler. Conformal invariance of Ising model correlations. PhD thesis, University of Geneva, 2010. 1 [31] Clément Hongler and Stanislav Smirnov. The energy density in the planar Ising model. Acta Mathematica, 211(2):191–225, 2013. 1 [32] C. A. Hurst and H. S. Green. New Solution of the Ising Problem for a Rectangular Lattice. The Journal of Chemical Physics, 33(4):1059–1062, 1960. 1 [33] Ernst Ising. Beitrag zur Theorie des Ferromagnetismus. Zeitschrift für Physik, 31(1):253–258, 1925. 1 [34] N Sh Izmailian and Chin-Kun Hu. Exact amplitude ratio and finite-size corrections for the m× n square lattice ising model. Physical Review E, 65(3):036103, 2002. 1, 2.2, 2.2 [35] Konstantin Izyurov and Mikhail Khristoforov. Asymptotics of the determinant of discrete laplacians on triangulated and quadrangulated surfaces. arXiv preprint arXiv:2007.08941, 2020. 1, 12, 30 [36] M. Kac and J. C. Ward. A combinatorial solution of the two-dimensional ising model. Phys. Rev., 88:1332–1337, Dec 1952. 1, 3, 3

30 [37] Leo P. Kadanoff and Horacio Ceva. Determination of an operator algebra for the two- dimensional Ising model. Phys. Rev. B (3), 3:3918–3939, 1971. 2.2 [38] P. W. Kasteleyn. Dimer Statistics and Phase Transitions. Journal of Mathematical Physics, 4(2):287–293, 1963. 1 [39] P. W. Kasteleyn. Graph theory and crystal physics. In Graph Theory and Theoretical Physics, pages 43–110. Academic Press, London, 1967. 1 [40] P.W. Kasteleyn. The statistics of dimers on a lattice. Physica, 27(12):1209–1225, 1961. 1 [41] Bruria Kaufman. Crystal statistics. ii. partition function evaluated by spinor analysis. Phys. Rev., 76:1232–1243, Oct 1949. 1 [42] Bruria Kaufman and Lars Onsager. Crystal statistics. iii. short-range order in a binary ising lattice. Phys. Rev., 76:1244–1252, Oct 1949. 1 [43] H. A. Kramers and G. H. Wannier. Statistics of the two-dimensional ferromagnet. part i. Phys. Rev., 60:252–262, Aug 1941. 1 [44] Marcin Lis. The fermionic observable in the ising model and the inverse kac–ward operator. In Annales Henri Poincaré, volume 15, pages 1945–1965. Springer, 2014. 1 [45] Barry M. McCoy and Tai Tsun Wu. The two-dimensional Ising model. Courier Corporation, 2014. 1 [46] Lars Onsager. Crystal statistics. i. a two-dimensional model with an order-disorder transition. Phys. Rev., 65:117–149, Feb 1944. 1 [47] John Palmer. Planar Ising correlations, volume 49 of Progress in Mathematical Physics. Birkhäuser Boston, Inc., Boston MA, 2007. 1 [48] R. Peierls. On ising’s model of ferromagnetism. Mathematical Proceedings of the Cambridge Philosophical Society, 32:477–481, 10 1936. 1 [49] R. B. Potts and J. C. Ward. The combinatorial method and the two-dimensional ising model. Progress of Theoretical Physics, 13(1):38–46, 1955. 1 [50] Jesús Salas. Exact finite-size-scaling corrections to the critical two-dimensional ising model on a torus. Journal of Physics A: Mathematical and General, 34(7):1311–1331, feb 2001. 1, 2.2, 2.2 [51] Jesús Salas. Exact finite-size-scaling corrections to the critical two-dimensional ising model on a torus: II. triangular and hexagonal lattices. Journal of Physics A: Mathematical and General, 35(8):1833–1869, feb 2002. 1, 2.2, 30 [52] S. Sherman. Combinatorial Aspects of the Ising Model for Ferromagnetism. I. A Conjecture of Feynman on Paths and Graphs. Journal of Mathematical Physics, 1(3):202–217, 1960. 1 [53] Stanislav Smirnov. Conformal invariance in random cluster models. i. holomorphic fermions in the ising model. Annals of Mathematics, 172:1435–1467, 2010. 1 [54] B. L. van der Waerden. Die lange Reichweite der regelmäßigen Atomanordnung in Mis- chkristallen. Zeitschrift für Physik, 118(7):473–488, 1941. 1

31