1 Introduction

The most famous problem of the perturbative expansion in quantum field theory is the existence of ultraviolet divergences in the amplitudes of Feynman diagrams. This is successfully dealt with using the theory of perturbative renormalization. However, even in one and zero dimensions (quantum mechanics and combinatorial models, respectively), where renormalization is not needed, perturbation theory poses another notorious challenge: in most cases the perturbative series is only an asymptotic series, with zero radius of convergence. Borel resummation is the standard strategy to address this problem, but this comes with its own subtleties. From a practical standpoint, we are often only able to compute just the first few terms in the perturbative expansion. At a more fundamental level, singularities are present in the Borel plane, associated to instantons (and renormalons in higher dimensions). The instanton singularities are not accidental: they stem from the factorial growth of the number of Feynamn diagrams with the perturbation order, which is also the origin of the divergence of the perturbation series.Footnote 1

From the resummation point of view, the most inconvenient feature of perturbation theory is that it does not naively capture contributions from non-analytical terms. For example, it is well known that instanton contributions of the type \(e^{-1/g}\) (\(g>0\) being the coupling constant) can be present in the evaluation of some quantity of interest, but they are missed in the perturbative series as their Taylor expansion at \(g=0\) vanishes identically.

Such exponentially suppressed terms are the archetypal example of nonperturbative effects, and their evaluation poses an interesting challenge. Aiming to include them, but still relying for practical reasons on perturbative methods, one ends up with a more general form of asymptotic expansion, known as transseries, which is roughly speaking a sum of perturbative and nonperturbative sectors, for example:

$$\begin{aligned} F(g) \simeq \sum _{n\ge 0} a_n\, g^n + \sum _i e^{\frac{c_i}{g}} g^{\gamma _i} \sum _{n\ge 0} b_{i,n}\, g^{n} \;. \end{aligned}$$
(1.1)

Over time it became increasingly clear that, in many examples of interest, using the theory of Borel summation for the perturbative sector it is possible to reconstruct some information about the nonperturbative ones. This relation between the perturbative and nonperturbative sectors is known as resurgence, and it was originally developed by Écalle in the context of ordinary differential equations [1] (see [2] for a modern review). Ideas coming from resurgence theory were extensively used in quantum field theory: for recent reviews with a quantum field theory scope, see [3,4,5], and in particular [6], which contains also a comprehensive list of references to applications and other reviews.

Zero-dimensional quantum field theoretical models, which are purely combinatorial models,Footnote 2 are useful toy models for the study of transseries expansions. Most conveniently, they allow one to set aside all the complications arising from the evaluation and renormalization of Feynman diagrams. Moreover, their partition functions and correlations typically satisfy ordinary differential equations, thus fitting naturally in the framework of Écalle’s theory of resurgence. The zero-dimensional \(\phi ^4\), or more generally \(\phi ^{2k}\) with \(k\ge 2\), models in zero dimensions have been exhaustively studied [6, 10]. From a physics perspective, the current mathematical literature on resurgence deals mainly with such zero-dimensional models.

At the opposite end, the rigorous study of the Borel summability in fully fledged quantum field theory is the object of constructive field theory [11,12,13]. It should come as no surprise that the generalization of results on resurgence in zero dimensions to the higher-dimensional setting is very much an open topic: while Écalle’s theory can serve as good inspiration, as in [3,4,5,6], the rigorous study of resurgence in higher-dimensional quantum field theory is much more involved. First of all, in higher dimensions the partition function (and correlators) does not obey an ordinary differential equation, and one cannot simply invoke Écalle’s theory. Moreover, the coefficients of the perturbative series are given by divergent Feynman amplitudes, which need to be renormalized, leading to a running coupling. Incorporating the effects of renormalization in the resurgence analysis is an open question (see, for instance, [14] for an investigation of resurgence in the Callan–Symanzik renormalization group equation). One thing is clear: in order to answer such questions, it is insufficient to simply invoke the current theory of resurgence, and one needs to develop new techniques adapted to the more general context of quantum field theory.

From this perspective, revisiting the resurgence in zero-dimensional models using techniques inspired by constructive field theory can be of great use. Following such route, we consider the zero-dimensional O(N) model with quartic potential.Footnote 3 Denoting \(\phi = (\phi _a)_{a=1,\dots N}\in \mathbb {R}^N\) a vector in \(\mathbb {R}^N\) and \(\phi ^2=\sum _{a=1}^N \phi _a\phi _a\) the O(N) invariant, the partition function of the model isFootnote 4:

$$\begin{aligned} Z(g,N)=\int _{-\infty }^{+\infty } \left( \prod _{a=1}^N\frac{\textrm{d}\phi _a}{\sqrt{2\pi }} \right) \; e^{-S[\phi ] } \;, \qquad S[\phi ]=\frac{1}{2}\phi ^2+\frac{g}{4!} (\phi ^2)^2 \; . \end{aligned}$$
(1.2)

The \(N=1\) case has been extensively studied in [6]. One can analytically continue Z(g, 1), regarded as a function of the coupling constant g, to a maximal domain in the complex plane. Subsequently, one discovers that Z(g, 1) displays a branch cut at the real negative axis and that the nonperturbative contributions to Z(g, 1) are captured by its discontinuity at the branch cut. A resurgent transseries is obtained when one considers g as a point on a Riemann surface with a branch point at \(g=0\). From now on we parameterize this Riemann surface as \(g = |g| e^{\imath \varphi }\) and we choose as principal sheet \(\varphi \in (-\pi ,\pi )\).

An approach to the study of the partition function in Eq. (1.2) in the case \(N=1\) is to use the steepest-descent method [16, 17]. We concisely review this in Appendix A. One notes that on the principal sheet only one Lefschetz thimble contributes. As g sweeps through the principal sheet the thimble is smoothly deformed, but not in the neighborhood of the saddle point: the asymptotic evaluation of the integral is unchanged. When g reaches the negative real axis, there is a discontinuous jump in the relevant thimbles and a pair of thimbles (passing through a pair of conjugated non-trivial saddle points of the action) starts contributing, giving rise to a one-instanton sector in the transseries of Z(g, 1).

Another approach to the transseries expansion of Z(g, 1) is to use the theory of ordinary differential equations [6, 10]. It turns out that Z(g, 1) obeys a second-order homogeneous linear ordinary differential equation for which \(g=0\) is an irregular singular point (e.g., [16]), giving another perspective on why the expansion one obtains is only asymptotic.

More interestingly, one can wonder what can be said about the nonperturbative contributions to the free energy, that is, the logarithm of the partition function \(W(g,1)=\ln Z(g,1)\), or to the connected correlation functions. If we aim to study the free energy, the steepest-descent method does not generalize straightforwardly as we lack a simple integral representation for W(g, 1). One can formally write Z(g, 1) as a transseries and then expand the logarithm in powers of the transseries monomial \(e^{\frac{c}{g}}\), thus obtaining a multi-instanton transseries. However, this is very formal, as the transseries is only an asymptotic expansion, and we would like to have a direct way to obtain the asymptotic expansion of W(g, 1). The closest one can get to an integral formula for the free energy is to use the Loop Vertex Expansion (LVE) [18]. This constructive field theory expansion is a combination of the intermediate field representation with the Brydges–Kennedy–Abdesselam–Rivasseau (BKAR) formula and has successfully been used in higher-dimensional quantum field theory [19] to prove Borel summability results and even study the decay of the correlations. (Note, however, that the results of [19] concern a theory with fixed UV and IR cutoffs, bypassing the issue of the renormalization group flow.) However, deriving directly the transseries expansion of W(g, 1) using the steepest-descent method on the LVE proved so far impractical. One can study the transseries expansion of W(g, 1) using again the theory of ordinary differential equations as W(g, 1) obeys a nonlinear ordinary differential equation [6], but this cannot be directly generalized to higher dimensions.

In this paper, we consider a general N and we revisit both the partition function Z(gN) and the free energy W(gN) from a different angle. We focus, in particular, on the small-N expansion, which provides a natural interpretation for the LVE. Such expansion could be physically interesting in higher dimensions, where the \(N \rightarrow 0\) limit of the model is related to self-avoiding random walks [20], an active area of investigation in modern statistical physics [21]. However, it should be noticed that our results will not be confined to infinitesimal N, as such expansion has a finite radius of convergence (Proposition 2 below).

The aim of our paper is not to “solve the problem” of computing the transseries expansion of Z(gN) or W(gN): this can be done almost immediately using known results about special functions and classical results in resurgence theory. Our aim is to analyze these objects using techniques one can then employ in the interesting case of quantum field theory in higher dimensions. The main results of this paper are the use of the Hubbard–Stratonovich intermediate field formulation to introduce a small-N expansion (Sect. 3.2), the application of the LVE to prove analyticity and Borel summability results for the free energy (Sect. 4.1), in N and g, and the study of the resurgence properties of the LVE. The results of the present paper provide a proof of concept for a set of techniques which can be employed in higher dimensions, starting in the quantum mechanical case and then moving on to quantum field theory.

The paper is organized as follows. In Sect. 2, we review the Borel summability of asymptotic series as well as the notion of Borel summable functions, deriving in the process a slight extension of the Nevanlinna–Sokal theorem.

In Sect. 3, we study Z(gN) in the intermediate field representation. This allows us to quickly prove its Borel summability along all the rays in the cut complex plane \(\mathbb {C}_{\pi } = \mathbb {C}\setminus \mathbb {R}_-\). More importantly, the intermediate field representation provides a new perspective on the origin of the instanton contributions: in this representation, the steepest-descent contour never changes, but when g reaches the negative real axis a singularity traverses it and detaches a Hankel contour around a cut. We insist that this Hankel contour is not a steepest-descent contour, but it does contribute to the asymptotic evaluation of the integral, because the cut is an obstruction when deforming the contour of integration toward the steepest-descent path. It is precisely the Hankel contour that yields the one-instanton contribution. We then build the analytic continuation of Z(gN) to the whole Riemann surface, identify a second Stokes line, compute the Stokes data encoding the jumps in the analytic continuation at the Stokes lines and discuss the monodromy of Z(gN). Next we observe that, because in the intermediate field representation N appears only as a parameter in the action, we can perform a small-N expansion:

$$\begin{aligned} Z(g,N) = \sum _{n\ge 0} \frac{1}{n!} \left( -\frac{N}{2} \right) ^n Z_n(g) \; . \end{aligned}$$
(1.3)

We thus study \(Z_n(g)\) for all integer n, proving its Borel summability in \(\mathbb {C}_\pi \) and computing its transseries expansion in an extended sector of the Riemann surface, with \(\arg (g)\in (-3\pi /2,3\pi /2)\), which we denote \(\mathbb {C}_{3\pi /2}\).

In Sect. 4, we proceed to study \(W(g,N)=\ln (Z(g,N))\). We first establish its Borel summability along all the rays in \(\mathbb {C}_{\pi }\) using constructive field theory techniques. We then proceed to the small-N expansion of this object:

$$\begin{aligned} W(g,N) = \sum _{n\ge 1} \frac{1}{n!} \left( -\frac{N}{2} \right) ^n W_n(g) \;, \end{aligned}$$
(1.4)

and prove that this is an absolutely convergent series in a subdomain of \(\mathbb {C}_{3\pi /2}\) and that both W(gN) and \(W_n(g)\) are Borel summable along all the rays in \(\mathbb {C}_\pi \). Finally, in order to obtain the transseries expansion of \(W_n(g)\) and W(gN) we note that \(W_n(g)\) can be written in terms of \(Z_n(g)\) using the Möbius inversion formula relating moments and cumulants. Because of the absolute convergence of the small-N series, it makes sense to perform the asymptotic expansion term by term, and thus, we rigorously obtain the transseries for W(gN) in a subdomain of \(\mathbb {C}_{3\pi /2}\). In the Appendices we gather some technical results, and the proofs of our propositions.

Ultimately, we obtain less information on the Stokes data for W(gN) than for Z(gN). While for Z(gN) we are able to maintain analytic control in the whole Riemann surface of g, the constructive field theory techniques we employ here allow us to keep control over W(gN) as an analytic function on the Riemann surface only up to \(\varphi = \pm 3\pi /2\), that is past the first Stokes line, but not up to the second one. The reason for this is that close to \(\varphi = \pm 3\pi /2\) there is an accumulation of Lee–Yang zeros, that is zeros of Z(gN), which make the explicit analytic continuation of W(gN) past this sector highly non-trivial. New techniques are needed if one aims to recover the Stokes data for W(gN) farther on the Riemann surface: an analysis of the differential equation obeyed by W(gN) similar to the one of [22] could provide an alternative way to access it directly.

One can naturally ask what is the interplay between our results at small N and the large-N nonperturbative effects, first studied for the zero-dimensional O(N) model in [23] (see also [3] for a general review, and [24] for a more recent point view). This is a very interesting question: indeed, the relation between the two expansions is a bit more subtle than the relation between the small coupling and the large coupling expansions for instance. The reason is that, when building the large N series, one needs to use the ’t Hooft coupling, which is a rescaling of the coupling constant by a factor of N. This changes the N-dependence of the partition function and free energy, making the relation between small-N and large-N expansions nontrivial. A good news on this front is that the analyticity domains in g becomes uniform in N when recast in terms of the ’t Hooft coupling [25]. But there is still quite some work to do in order to connect the transseries analysis at small N with that at large N. However, we stress once more that, while the large-N expansion is asymptotic, the small-N expansion is convergent.

Main results. Our main results are the following:

  • In Proposition 1, we study Z(gN). While most of the results in this proposition are known for \(N=1\), we recover them using the intermediate field representation (which provides a new point of view) and generalize them to arbitrary \(N\in \mathbb {R}\). In particular, we uncover an interplay between Z(gN) and \(Z(-g,2-N)\) in the transseries expansion of the partition function for general N.

  • Proposition 2 deals with the function \(Z_n(g)\), notably its Borel summability, transseries, and associated differential equation. To our knowledge, \(Z_n(g)\) has not been studied before and all of the results presented here are new.

  • Proposition 3 and 4 generalize previous results in the literature [26] on the analyticity and Borel summability of W(g, 1) to W(gN) and furthermore derive parallel results for \(W_n(g)\).

  • Proposition 5 contains the transseries expansion of \(W_n(g)\), which has not been previously considered in the literature. We also give a closed formula for the transseries expansion of W(gN).

  • Lastly, in Proposition 6, we derive the tower of recursive differential equations obeyed by \(W_n(g)\). This serves as an invitation for future studies of the transseries of \(W_n(g)\) from an ordinary differential equations perspective.

The natural next step is to explore how this picture is altered in higher dimensions. While this is a wide open question, several lessons can be learned from our present work. First, the small N series for the free energy W obtained via the LVE will very likely be convergent; hence, one should be able to study the resurgence properties of W by first studying such properties for the “cumulants” \(W_n\), which can be done by studying the “moments \(Z_n\)” and using Möbius inversion. Second, for the moments \(Z_n\), the steepest-descent contour in the intermediate \(\sigma \) field representation will be insensitive to the coupling constant; hence, in this representation the Stokes phenomenon will correspond to singularities crossing this fixed contour.

2 Borel Summable Series and Borel Summable Functions

When dealing with asymptotic series, a crucial notion is that of Borel summability. Less known, there exists a notion of Borel summability of functions, intimately related to the Borel summability of series. In this section we present a brief review of these notions, which will play a central role in the rest of the paper, as well as a slight generalization of the (optimal) Nevanlinna–Sokal theorem on Borel summability [27]. We will repeatedly use this theorem in this paper.

Notation. We use K as a dustbin notation for irrelevant (real positive) multiplicative constants, and R and \(\rho \) for the important (real positive) constants.

Borel summable series. A formal power series \(A(z)=\sum _{k=0}^\infty a_k z^k\) is called a Borel summable series along the positive real axis if the series

$$\begin{aligned} B (t) = \sum _{k=0}^\infty \frac{a_k}{k!} t^k \;, \end{aligned}$$
(2.1)

is absolutely convergent in some disk \(|t|< \rho \) and B(t) admits an analytic continuation in a strip of width \(\rho \) along the positive real axis such that for t in this strip \(|B(t)| < K e^{|t|/R}\) for some real positive R. The function B(t) is called the Borel transform of A(z), and the Borel sum of A(z) is the Laplace transform of its Borel transform:

$$\begin{aligned} f(z) = \frac{1}{z} \int _0^{\infty } \textrm{d}t \; e^{-t/z} \; B (t) \; . \end{aligned}$$
(2.2)

It is easy to check that the function f is analytic in a disk of diameter R tangent to the imaginary axis at the origin, \(\textrm{Disk}_R = \{z\in \mathbb {C} \mid \textrm{Re }{(1/z)}>1/R \}\).

Clearly, if it exists, the Borel sum of a series is unique. This raises the following question: given a function h(z) whose asymptotic series at zero is the Borel summable series A(z), does the Borel resummation of A(z) reconstruct h(z)? That is, is \(f(z) = h(z)\)? The answer to this question is no in general: for instance the function \(e^{-1/z}\) is asymptotic (along the positive real axis) at 0 to the Borel summable series \(a_k=0\). It turns out that one can formulate necessary and sufficient conditions for h(z) which ensure that it is indeed the Borel sum of its asymptotic series, as we now recall.

Borel summable functions. A function \(f:\mathbb {C} \rightarrow \mathbb {C}\) is called a Borel summable function along the positive real axis if it is analytic in a disk \(\textrm{Disk}_R\) and has an asymptotic series at 0 (which can have zero radius of convergence),

$$\begin{aligned} f(z) = \sum _{k=0}^{q-1} a_k \; z^k + R_q(z) \; , \end{aligned}$$
(2.3)

such that the rest term of order q obeys the bound:

$$\begin{aligned} | R_q(z) | \le K \; q! \; q^\beta \; \rho ^{- q} \; |z|^q \;, \qquad z \in \textrm{Disk}_R \;, \end{aligned}$$
(2.4)

for some fixed \(\beta \in \mathbb {R}_+\). Note that the bound in Eq. (2.4) is slightly weaker that the one in [27]. The positive real axis is selected by the position of the center of \(\textrm{Disk}_R\). We call \(\textrm{Disk}_R = \{z\in \mathbb {C} \mid \textrm{Re }{(1/z)}>1/R \}\) a Sokal disk.

These two notions are intimately related: the Borel sums of Borel summable series are Borel summable functions (this is straightforward to prove). Moreover, the asymptotic series of Borel summable functions are Borel summable series.

Theorem 1

(Nevanlinna–Sokal [27], extended). Let \(f:\mathbb {C} \rightarrow \mathbb {C}\) be a Borel summable function, hence analytic and obeying the bound (2.4) with some fixed \(\beta \). Then:

  • the Borel transform of the asymptotic series of f,

    $$\begin{aligned} B(t) = \sum _{k=0}^{\infty } \frac{1}{k! } \; a_k \; t^k \;, \end{aligned}$$
    (2.5)

    is convergent in a disk of radius \(\rho \) in t, and it defines an analytic function in this domain.

  • B(t) can be analytically continued to the strip \( \{ t\in \mathbb {C} \mid \textrm{dist}(t,\mathbb {R}_+ ) < \rho \}\) and in this strip it obeys an exponential bound \( |B(t)|<K e^{|t|/R}\).

  • for all \(z\in \textrm{Disk}_R\) we can reconstruct the function f(z) by the absolutely convergent Laplace transform:

    $$\begin{aligned} f(z)=\frac{1}{z }\int _{0}^{\infty } \textrm{d}t \; e^{-t/z} \; B(t) \; . \end{aligned}$$
    (2.6)

Proof

See Appendix B. \(\square \)

We emphasize that both for series and for functions, Borel summability is directional:

  • for series, Borel summability along a direction requires the unimpeded analytic continuation of B(t) in a thin strip centered on that direction.

  • for functions, Borel summability along a direction requires analyticity and bound on the Taylor rest terms in a Sokal disk (with 0 on its boundary) centered on that direction.

Clearly, the singularities of the Borel transform B(t) are associated to directions along which the function f(z) ceases to be Borel summable.

3 The Partition Function Z(gN)

In this section, we collect some facts about the asymptotic expansion of the partition function (1.2). Most of them are known, or derivable from the expression of the Z(gN) in terms of special functions, whose asymptotic expansions are to a large degree known [28]. Nonetheless, we present a “path integral–like” derivation and rather explicit formulae for the coefficients that should be useful, as they are more directly generalizable, in applications to proper field theories.

We study Z(gN) by means of the Hubbard–Stratonovich intermediate field formulation [29, 30], which is crucial to the Loop Vertex Expansion [18] of the free energy W(gN) that we will study below. This is based on rewriting the quartic term of the action as a Gaussian integral over an auxiliary variable \(\sigma \) (or field, in higher dimensions):

$$\begin{aligned} e^{-\frac{g}{4!} (\phi ^2)^2} = \int _{-\infty }^{+\infty } [\textrm{d}\sigma ]\ e^{-\frac{1}{2} \sigma ^2 + \imath \sqrt{\frac{g}{12}} \sigma \phi ^2} \;, \end{aligned}$$
(3.1)

where the Gaussian measure over \(\sigma \) is normalized, i.e., \([\textrm{d}\sigma ]=\textrm{d}\sigma /\sqrt{2\pi }\) and \(\imath = e^{\imath \frac{\pi }{2}}\). Note that \(\sigma \) is a real number, not a vector. With this trick, the integral over \(\phi \) becomes Gaussian and can be performed for \(g>0\), leading to a rewriting of the partition function (1.2) as:

$$\begin{aligned} Z(g,N)=\int _{-\infty }^{+\infty } [\textrm{d}\sigma ] \;e^{-\frac{1}{2} \sigma ^2} \frac{1}{\left( 1 - \imath \sqrt{ \frac{g}{3} } \sigma \right) ^{ N/2} } \;. \end{aligned}$$
(3.2)

Although the original partition function is defined only for integer N and we have assumed that \(g>0\), in the \(\sigma \) representation (3.2) it becomes transparent that Z(gN) can be analytically continued both in N and in g.

3.1 Analytic Continuation and Transseries

As a matter of notation, we denote \(\varphi \equiv \arg (g)\), and, in order to label some sets that will appear repeatedly in the rest of the paper, we define:

$$\begin{aligned} \mathbb {C}_{\psi } \equiv \left\{ g\in \mathbb {C},\; g=|g| e^{\imath \varphi } :\; \varphi \in \left( -\psi ,\psi \right) \right\} \;. \end{aligned}$$
(3.3)

In particular, \(\mathbb {C}_\pi =\mathbb {C} \setminus \mathbb {R}_-\) is the cut complex plane. For \(\psi >\pi \), the set \( \mathbb {C}_{\psi }\) should be interpreted as a sector of a Riemann sheet, extending the principal sheet \(\mathbb {C}_\pi \) into the next sheets.

Our first aim is to understand the analytic continuation of the partition function in the maximal possible domain of the Riemann surface. For later convenience, we introduce the following function, not to be confused with the partition function Z(gN):

$$\begin{aligned} Z^{\mathbb {R}}(g,N) = \int _{-\infty }^{+\infty } [\textrm{d}\sigma ] \;e^{-\frac{1}{2} \sigma ^2} \frac{1}{\left( 1 - \imath \, e^{\imath \frac{\varphi }{2}} \sqrt{ \frac{|g|}{3} } \sigma \right) ^{ N/2} } \;,\quad \forall \varphi \ne (2k+1)\pi \, ,\; k \in \mathbb {Z} \;, \end{aligned}$$
(3.4)

which is an absolutely convergent integral for any |g| and any \(\varphi \ne (2k+1)\pi \). We have used a superscript \(\mathbb {R}\) to distinguish it from Z(gN) and to emphasize that the integral is performed on the real line, irrespective of the value of \(\varphi \ne (2k+1)\pi \). Moreover, for \(N/2\notin \mathbb {Z}\), the integrand is computed using the principal branch of the logarithm:

$$\begin{aligned} \textrm{ln} \left( 1 - e^{\imath \frac{\pi + \varphi }{2}} \sqrt{ \tfrac{|g|}{3} } \sigma \right)= & {} \frac{1}{2} \ln \left[ \left( \cos \tfrac{\varphi }{2} \right) ^2 + \left( \sin \tfrac{\varphi }{2} + \sqrt{ \tfrac{|g|}{3} } \sigma \right) ^2 \right] \nonumber \\ {}{} & {} + \imath \; \textrm{Arg}\left( 1 - e^{\imath \frac{\pi + \varphi }{2}} \sqrt{ \tfrac{|g|}{3} } \sigma \right) \;, \end{aligned}$$
(3.5)

where the \(\textrm{Arg}\) function is the principal branch of the argument, valued in \((-\pi ,\pi )\). In particular, a change of variables \(\sigma \rightarrow -\sigma \) shows that \( Z^{\mathbb {R}}(g,N) = Z^{\mathbb {R}}( e^{2 \pi \imath } g,N)\), which is thus a single-valued function on \(\mathbb {C}_{\pi }\), with a jump at \(\mathbb {R}_-\). The analytic continuation of the partition function Z(gN) will instead be a multi-valued function on \(\mathbb {C}\), and thus require the introduction of a Riemann surface. We can view \(Z^{\mathbb {R}}(g,N)\) as a periodic function on the same Riemann surface, with of course \(Z(g,N) = Z^{\mathbb {R}}(g,N)\) for \(g\in \mathbb {C}_{\pi }\), and \(Z(g,N) \ne Z^{\mathbb {R}}(g,N)\) once one steps out of the principal Riemann sheet.

We collect all the relevant result concerning the partition function in Proposition 1. For now we restrict to N real, but the proposition can be extended to complex N with little effort.Footnote 5 We will drop this assumption later.

Proposition 1

(Properties of Z(gN)). Let \(N \in \mathbb {R}\) be a fixed parameter. The partition function Z(gN) satisfies the following properties:

  1. 1.

    for every \(g\in \mathbb {C}_\pi \), the integral in Eq. (3.2) is absolutely convergent and bounded from above by

    $$\begin{aligned} |Z(g,N)| \le {\left\{ \begin{array}{ll} \left( \cos \frac{\varphi }{2} \right) ^{- N /2} \;, \qquad &{} N \ge 0\\ 2^{|N|/2}+\frac{2^{3|N|/4}}{\sqrt{\pi }} \frac{|g|^{N/4}}{3^{|N|/4}} \Gamma \left( \tfrac{|N|+2}{4} \right) \;, \qquad &{} N < 0 \end{array}\right. } \; ; \end{aligned}$$
    (3.6)

    hence, Z(gN) is analytic in \(\mathbb {C}_\pi \).

  2. 2.

    For \(g\in \mathbb {C}_\pi \), the partition function is \(Z(g,N) = Z^{\mathbb {R}}(g,N) \) and has the perturbative expansion:

    $$\begin{aligned} Z(g,N) \simeq \sum _{n=0}^{\infty } \, \frac{\Gamma (2n+N/2) }{2^{2n}n! \, \Gamma (N/2) } \; \left( - \frac{2g}{3}\right) ^n \;, \end{aligned}$$
    (3.7)

    where \(\simeq \) means that the equation has to be interpreted in the sense of asymptotic series, i.e.,

    $$\begin{aligned} \lim _{\begin{array}{c} g\rightarrow 0 \\ g\in \mathbb {C}_\pi \end{array}} g^{-n_{\textrm{max}}} \left| Z(g,N) - \sum _{n=0}^{n_{\textrm{max}}} \, \frac{\Gamma (2n+N/2) }{2^{2n}n! \, \Gamma (N/2) } \; \left( - \frac{2g}{3}\right) ^n \right| = 0\;, \qquad \forall n_{\textrm{max}}\ge 0\;. \end{aligned}$$
    (3.8)
  3. 3.

    The function Z(gN) is Borel summable along all the directions in \(\mathbb {C}_\pi \).

  4. 4.

    Z(gN) can be continued past the cut, on the entire Riemann surface. However, \(\mathbb {R}_-\) is a Stokes line, that is, the anticlockwise and clockwise analytic continuations \(Z_+(g,N)\) and \(Z_-(g,N)\) are not equal and cease to be Borel summable at \(\mathbb {R}_-\). A second Stokes line is found at \(\mathbb {R}_+\) on the second sheet. For \(\varphi \notin \pi \mathbb {Z}\), the analytic continuation of the partition function to the whole Riemann surface writes:

    $$\begin{aligned}&2k\pi< |\varphi |< (2k+1) \pi : \nonumber \\ {}&\quad Z(g,N ) = \omega _{2k} \; Z(e^{\imath ( 2 k ) \tau \pi } g,N) \nonumber \\ {}&\qquad + \eta _{2k} \; \frac{\sqrt{ 2\pi } }{\Gamma (N/2)} \; e^{\imath \tau \frac{\pi }{2} } \; e^{\frac{3}{2g}} \; \left( e^{\imath (2k + 1) \tau \pi } \frac{ g}{3} \right) ^{\frac{1-N}{2} } \; Z ( e^{\imath (2k + 1) \tau \pi } g , 2-N) \nonumber \\ {}&\quad = \omega _{2k} \; Z^{\mathbb {R}}( g,N) \nonumber \\ {}&\qquad + \eta _{2k} \; \frac{\sqrt{ 2\pi } }{\Gamma (N/2)} \; e^{\imath \tau \frac{\pi }{2} } \; e^{\frac{3}{2g}} \; \left( e^{\imath (2k + 1) \tau \pi } \frac{ g}{3} \right) ^{\frac{1-N}{2} } \; Z^{\mathbb {R}} ( - g , 2-N) \;, \nonumber \\ {}&(2k+1)\pi< |\varphi | < (2k+2) \pi : \nonumber \\ {}&\quad Z(g,N ) = \omega _{2k+1} \; Z(e^{\imath ( 2 k+2) \tau \pi } g,N) \nonumber \\ {}&\qquad + \eta _{2k+1} \; \frac{\sqrt{ 2\pi } }{\Gamma (N/2)} \; e^{\imath \tau \frac{\pi }{2} } \; e^{\frac{3}{2g}} \; \left( e^{\imath (2k + 1) \tau \pi } \frac{ g}{3} \right) ^{\frac{1-N}{2} } \; Z ( e^{\imath (2k + 1) \tau \pi } g , 2-N) \nonumber \\ {}&\quad = \omega _{2k+1} \; Z^{\mathbb {R}}( g,N) \nonumber \\ {}&\qquad + \eta _{2k+1} \; \frac{\sqrt{ 2\pi } }{\Gamma (N/2)} \; e^{\imath \tau \frac{\pi }{2} } \; e^{\frac{3}{2g}} \; \left( e^{\imath (2k + 1) \tau \pi } \frac{ g}{3} \right) ^{\frac{1-N}{2} } \; Z^{\mathbb {R}} ( - g , 2-N) \;, \end{aligned}$$
    (3.9)

    where \(\tau =-\textrm{sgn}(\varphi )\) and the Stokes parameters \((\omega ,\eta )\) are defined recursively as:

    $$\begin{aligned} {}&{} (\omega _0 ,\eta _0) = (1,0) \;,\qquad {\left\{ \begin{array}{ll} \omega _{2k+1} &{}{}= \omega _{2k} \\ \eta _{2k+1} &{}{} = \eta _{2k} + \omega _{2k} \end{array}\right. } \; , \nonumber \\{}&{} {\left\{ \begin{array}{ll} \omega _{2(k+1)} &{}{}= \omega _{2k+1} + {\tilde{\tau }} \; \eta _{2k+1} \\ \eta _{2(k+1)} &{}{} = e^{ \imath \tau \pi (N-1) } \eta _{2k+1} \end{array}\right. } \; ,\end{aligned}$$
    (3.10)

    with \({\tilde{\tau }} = e^{\imath \tau \pi \frac{N +1 }{2} } \; 2 \sin \frac{N\pi }{2}\). The recursion gives:

    $$\begin{aligned} (\omega _{2k},\eta _{2k})= {\left\{ \begin{array}{ll} e^{\imath \tau \pi N\frac{k}{2}}\;(1,0) &{},\,k\ \text {even} \\ e^{\imath \tau \pi N\frac{k+1}{2}}\;(1,-1)&{},\,k\ \text {odd} \end{array}\right. }\;. \end{aligned}$$
    (3.11)

    The monodromy group of Z(gN) is of order 4 if N is odd, and of order 2 if N is even. More generally, we have a monodromy group of finite order if N is a rational number, and an infinite monodromy otherwise.

  5. 5.

    From Properties 2 and 4, we obtain that for g in the sector \(k\pi< |\varphi | < (k+1) \pi \) of the Riemann surface the partition function has the following transseries expansion:

    $$\begin{aligned} Z(g,N)&\simeq \omega _k \sum _{n=0}^{\infty } \, \frac{\Gamma (2n+N/2) }{2^{2n}n! \, \Gamma (N/2) } \; \left( - \frac{2g}{3}\right) ^n + \eta _k \; e^{\imath \tau \pi (1- \frac{N}{2} ) } \; \sqrt{2\pi }\nonumber \\&\quad \times \left( \frac{ g}{3} \right) ^{\frac{1-N}{2} } e^{\frac{3}{2g}} \sum _{q\ge 0} \frac{1}{ 2^{2q} q! \; \Gamma (\frac{N}{2} -2q ) } \left( \frac{2 g}{3} \right) ^q \; , \end{aligned}$$
    (3.12)

    where we used:

    $$\begin{aligned}&\Gamma ( 2q+ 1 - N/2) \Gamma (N/2 - 2q)\nonumber \\&\quad = \frac{ \pi }{ \sin ( \pi \frac{N}{2} - 2\pi q ) } = \Gamma (1-N/2) \Gamma (N/2) \;. \end{aligned}$$
    (3.13)

    The transseries displays an additional property: the instanton series is obtained from the perturbative one by substituting \(N\rightarrow 2-N\) and \(g\rightarrow -g\) and vice versa.

  6. 6.

    From Property 5, the discontinuity of the partition function at the negative real axis:

    $$\begin{aligned} \textrm{disc}_\pi \big (Z(g,N)\big )\equiv \lim _{ g \rightarrow \mathbb {R}_-} \bigg (Z_-(g,N) - Z_+(g,N) \bigg ) \;, \end{aligned}$$
    (3.14)

    has the following asymptotic expansion:

    $$\begin{aligned} \text {disc}_\pi \big (Z_n(g)\big )&\simeq \frac{e^{-\frac{3}{2|g|}}}{\sqrt{2\pi }}\sqrt{\frac{|g|}{3}} \sum _{q=0}^{\infty } \sum _{p=0}^{n} \frac{1}{q!} \left( -\frac{|g|}{6}\right) ^q \left( {\begin{array}{c}n\\ p\end{array}}\right) \frac{d^p \Gamma (z)}{\text {d}z^p}\Big |_{z=2q+1}\nonumber \\ {}&\quad \times \left[ \left( \ln |\frac{g}{3}| -\imath \pi \right) ^{n-p} -\left( \ln |\frac{g}{3}| +\imath \pi \right) ^{n-p}\right] \; . \end{aligned}$$
    (3.15)

    where for N even integer the sum truncates at \(q=N/4-1\), if N is a multiple of 4, and at \(q=\lfloor N/4 \rfloor \) otherwise.

  7. 7.

    The partition function obeys an homogenous linear ordinary differential equation:

    $$\begin{aligned}{} & {} 16 g^2 Z^{\prime \prime }(g,N)+\left( (8N+ 24)g+24 \right) Z^\prime (g,N) \nonumber \\{} & {} \quad +N(N+2)Z(g,N)=0 \; , \end{aligned}$$
    (3.16)

    which can be used to reconstruct the resurgent transseries expansion of Z(gN).

Proof

See Appendix C.1. \(\square \)

The proof of this proposition is quite technical. The most interesting points come at Property 4. While the full details can be found in Appendix C.1, we discuss here how the Stokes phenomenon arises in the intermediate field representation.

In order to obtain the asymptotic approximation of an integral, we need to deform the integration contour to steepest-descent contours (or Lefschetz thimbles) where the Laplace method can be applied. An integration contour will in principle intersect several steepest-ascent (upwards) paths of several saddle points and it must then be deformed (i.e., relaxed under the gradient flow) to run along the thimbles of these saddle points [6, 17]. When varying some parameter continuously, the relevant thimbles can collide and change discontinuously leading to discontinuous changes of the asymptotic regimes at Stokes lines. This is exactly the picture in the \(\phi \) representation of the partition function, which we recall in Appendix A.

In the \(\sigma \) representation the picture is different. Let us go back to Eq. (3.2) expressing Z(gN) as an integral over the real line. The partition function Z(gN) is analytically continued to the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\) by tilting the integration contour to \(e^{-\imath \theta }\sigma \), respectively \(e^{\imath \theta }\sigma \), with \(\theta > 0\) for its anticlockwise, respectively clockwise, analytic continuations \(Z_+(g,N)\) and \(Z_-(g,N)\).

In this representation, the Lefshetz thimble is always the real axis, irrespective of g. In fact, as g goes to zero, the Laplace method instructs us to look for the saddle point of the exponent in the integrand (the function f(x) in Eq. (A.1)), which in this case is a simple quadratic function,Footnote 6 while the subexponential function (the function a(x) in Eq. (A.1)) is irrelevant for the determination of the saddle points. However, what happens is that the integrand has a branch point (or pole for N a positive even integer) and this point crosses the thimble at the Stokes line. This is depicted in Fig. 1.

Fig. 1
figure 1

As \(\arg (g)\) increases, the branch cut moves clockwise in the complex \(\sigma \)-plane. When g crosses the negative real axis, the tilted contour is equivalent to a Hankel contour C plus the original contour along the real line (3.17)

In detail, as long as \(g = |g| e^{\imath \varphi }\in \mathbb {C}_{\pi }\), the integral converges because the branch point \(\sigma _\star = -\imath \sqrt{ 3/g } \) with branch cut \(\sigma _\star \times (1,+\infty )\) lies outside the integration contour. As g approaches \(\mathbb {R}_-\), the branch point hits the contour of integration: for \(\varphi \nearrow \pi \) the branch point hits the real axis at \( - \sqrt{3/|g|}\). The analytic continuation \(Z_+(g,N)\) consists in tilting the contour of integration in \(\sigma \) to avoid the collision with the branch point. However, in order to derive the asymptotic behavior of \(Z_+(g,N)\), we need to deform the integration contour back to the thimble, which is always the real axis. Once g passes on the second Riemann sheet (\( \varphi > \pi \)), when deforming the tilted contour to the real axis we generate an additional Hankel contour (see Fig. 1):

$$\begin{aligned} Z_\pm (g,N)\big |_{\varphi \genfrac{}{}{0.0pt}{}{>\pi }{< -\pi }}&=\int _{ e^{ \mp \imath \theta } \mathbb {R}} [\textrm{d}\sigma ] \;\frac{e^{-\frac{1}{2} \sigma ^2} }{\left( 1 - \imath \sqrt{ \frac{g}{3} } \sigma \right) ^{N/2} } = Z^{\mathbb {R}}(g,N) + Z^C_\pm (g,N) \nonumber \\ Z^{\mathbb {R}}(g,N)&= \int _{ \mathbb {R}} [\textrm{d}\sigma ] \;e^{-\frac{1}{2} \sigma ^2} \frac{1}{\left( 1 - \imath \sqrt{ \frac{g}{3} } \sigma \right) ^{N/2} } \;,\nonumber \\ Z^C_\pm (g,N)&= \int _{C} [\textrm{d}\sigma ] \;e^{-\frac{1}{2} \sigma ^2} \frac{1}{\left( 1 - \imath \sqrt{ \frac{g}{3} } \sigma \right) ^{N/2} } \;, \end{aligned}$$
(3.17)

where the Hankel contour C turns clockwise around the cut \(\sigma _\star \times (1,+\infty )\), i.e., starting at infinity with argument \( \frac{3\pi }{2} -\frac{\varphi }{2}\) and going back with argument \( -\frac{\pi }{2} -\frac{\varphi }{2}\) after having encircled the branch point \(\sigma _\star \). We kept a subscript ± for the contribution of the Hankel contour, because, even though the definition of \(Z^C_\pm (g,N)\) and C might suggest that it is one single function of g, in fact the integral around the cut is divergent for \(|\varphi |<\pi /2\), and therefore, the integrals at \(\pi<\varphi <3\pi /2\) and at \(-\pi>\varphi >-3\pi /2\) are not the analytic continuation of each other. This fact is reflected in the \(\tau \)-dependence of the asymptotic expansion in Eq. (3.12).

The appearance of the Hankel contour marks a discontinuity of the contour of integration in \(\sigma \) as a function of the argument of g, which translates into a discontinuity of the asymptotic expansion of Z(gN), that is, a Stokes phenomenon. We insist that the Hankel contour is not a thimble for the integral in Eq. (3.2), but it contributes to the asymptotic evaluation of the integral, providing the one-instanton contribution in the transseries of Z(gN).

In order to go beyond \(|\varphi |=3\pi /2\), one notices that we can analytically continue separately \(Z^{\mathbb {R}}(g,N)\) and \(Z^C_\pm (g,N)\). The first is analytic in the range \(|\varphi |\in (\pi ,3\pi )\), where its asymptotic expansion is just the standard perturbative series. The analytic continuation of \(Z^C_\pm (g,N)\) is not immediate in the Hankel contour representation, which is only convergent for \(\pi /2<|\varphi |<3\pi /2\), but after resolving the discontinuity at the cut and using again the Hubbard–Stratonovich trick (as detailed in Appendix C.1), it turns out that it can be rewritten as:

$$\begin{aligned} Z^C_\pm (g,N) = e^{\imath \tau \pi (1-\frac{N}{2}) }\left( \frac{ g}{3} \right) ^{\frac{1-N}{2} } e^{\frac{3}{2g}} \frac{\sqrt{ 2\pi } }{\Gamma (N/2)} \; Z ( e^{\imath \tau \pi } g , 2-N) \;. \end{aligned}$$
(3.18)

with \(\tau = -\textrm{sgn}(\varphi )\). In this form it is manifest that \( Z^C_\pm (g,N)\) is analytic in g as long as \( e^{\imath \tau \pi } g \) belongs to the principal sheet of the Riemann surface, that is for \(\pi<|\varphi |<2 \pi \). We have thus shown that, when going from \(|\varphi | <\pi \) to \(\pi< |\varphi | < 2\pi \), our analytic continuation of Z(gN) switches:

$$\begin{aligned} \begin{aligned}&Z(g,N) \xrightarrow []{ |\varphi | \nearrow \pi _+} Z^{\mathbb {R}}(g,N) + \frac{\sqrt{ 2\pi } }{\Gamma (N/2)} \; e^{\imath \tau \frac{\pi }{2} } \; e^{\frac{3}{2g}} \; \left( e^{\imath \tau \pi } \frac{ g}{3} \right) ^{\frac{1-N}{2} } Z^{\mathbb {R}} ( - g , 2-N) \\&\quad = Z( e^{\imath (2\tau \pi ) } g,N) + \frac{\sqrt{ 2\pi } }{\Gamma (N/2)} \; e^{\imath \tau \frac{\pi }{2} } \; e^{\frac{3}{2g}} \; \left( e^{\imath \tau \pi } \frac{ g}{3} \right) ^{\frac{1-N}{2} } \; Z ( e^{\imath \tau \pi } g , 2-N) \;, \end{aligned} \end{aligned}$$
(3.19)

where we explicitly exhibited the argument at which the switching takes place. In the second line above, for \(\pi< |\varphi | < 2\pi \), both arguments \( e^{\imath (2\tau \pi ) } g\) and \( e^{\imath \tau \pi } g\) belong to the principal sheet of the Riemann surface, where Z(gN) has already been constructed and proven to be analytic. The first term in Eq. (3.19) is regular up to \(|\varphi | = 3\pi \), but the second one has a problem when \(e^{\imath \tau \pi } g \) reaches the negative real axis and retracing our steps we conclude that the analytic continuation switches again:

$$\begin{aligned} \begin{aligned}&Z ( e^{\imath \tau \pi } g , 2-N) \xrightarrow []{ |\varphi | \nearrow 2\pi _+} Z ( e^{\imath (3\tau \pi )} g , 2-N) \\&\quad + \frac{\sqrt{ 2\pi } }{\Gamma (1-N/2)} \; e^{\imath \tau \frac{\pi }{2} } \; e^{\frac{3}{2ge^{\imath \tau \pi }}} \; \left( e^{\imath (2\tau \pi ) } \frac{ g}{3} \right) ^{\frac{N-1}{2} } \; Z ( e^{\imath (2 \tau \pi ) } g , N) \; , \end{aligned} \end{aligned}$$
(3.20)

where this time the arguments at which Z is evaluated on the right hand side stay in the principal sheet for \(2\pi< |\varphi | < 3\pi \). Iterating, one obtains the analytic continuation to the whole Riemann surface.

Remark 1

The differential equation (3.16) can be solved in terms of special functions. For \(N=1\), setting \(Z(g,1)=\sqrt{\frac{3}{2\pi g}} e^{\frac{3}{4g}} f(\frac{3}{4g})\), we find that the equation reduces to a modified Bessel’s equation for f(z):

$$\begin{aligned} z^2 f^{\prime \prime }(z) + z f^{\prime }(z) - \left( z^2 +\frac{1}{16}\right) f(z)=0 \;. \end{aligned}$$
(3.21)

Its two linearly independent solutions are the modified Bessel functions of the first and second kind of order 1/4. However, only the second, \(K_{1/4}(z)\), decays for \(z\rightarrow \infty \); hence, the initial condition \(Z(0,1)=1\) fixes \(f(z)=K_{1/4}(z)\).

Similarly, for general N, we find that setting \(Z(g,N)=(\frac{3}{2\,g})^{N/4} f(\frac{3}{2\,g})\) the differential equation reduces to Kummer’s equation for f(z):

$$\begin{aligned} z f^{\prime \prime }(z) + \left( \frac{1}{2}-z\right) f^{\prime }(z) - \frac{N}{4} f(z)=0 \;. \end{aligned}$$
(3.22)

With the addition of the initial condition \(Z(0,N)=1\), we find that the solution is given by the Tricomi confluent hypergeometric function \(f(z)=U(N/4,1/2,z)\), whose transseries expansion can easily be obtained order by order. However, such expressions of the partition function in terms of special functions do not generalize to quantum field theory. Similarly, even the more general resurgence theory for formal solutions of ordinary differential equations will be of limited use in that context, as the partition function of a quantum field theory model does not satisfy an ordinary differential equation. For this reason, while displaying for completeness the relevant ordinary differential equations, in this work we do not make use of them and rather refer to the literature (see references in [6]).

Our result (3.12) provides a useful repackaging of the transseries expansion of the partition function and an alternative derivation that should be more easily generalizable to quantum field theory.

3.2 Convergent Small-N Series of Z(gN) and Transseries of Its Coefficients \(Z_n(g)\)

We will now study the discontinuity of Z(gN) from a different perspective. We expand the integrand of Eq. (3.2) in powers of N and exchange the order of summation and integration:

$$\begin{aligned} Z(g,N) = \sum _{n\ge 0} \frac{1}{n!} \left( -\frac{N}{2}\right) ^n Z_n(g)\; ,\quad Z_n(g)= \int _{-\infty }^{+\infty } [\textrm{d}\sigma ] e^{-\frac{1}{2} \sigma ^2} \left( \ln \left( 1-\imath \sqrt{\frac{g}{3}} \sigma \right) \right) ^n \;. \end{aligned}$$
(3.23)

Unlike the usual perturbative expansions in g, this is a convergent expansion: from the bound in Property 1 of Proposition 2 below, the Gaussian integral and the sum can be commuted due to Fubini’s Theorem. As a function of N, we can regard Z(gN) as a generating function of “moments”: unlike the usual moments, we are dealing with expectations of powers of the logarithm.

Proposition 2

(Properties of \(Z_n(g)\)). The \(Z_n(g)\), \(n\in \mathbb {N}_{\ge 0}\) satisfy the following properties:

  1. 1.

    \(Z_n(g)\) is analytic in the cut plane \(\mathbb {C}_\pi \). Indeed, for every \(g\in \mathbb {C}_\pi \), the integral (3.23) is absolutely convergent and bounded from above by:

    $$\begin{aligned} | Z_n(g) | \le K^n \frac{ \bigg ( |\ln ( \cos \frac{\varphi }{2}) |+1 \bigg )^n }{ \varepsilon ^n } \bigg (1 + |g|^{\frac{n\varepsilon }{2}} \Gamma \left( \tfrac{n\varepsilon + 1}{2} \right) \bigg ) \; , \end{aligned}$$
    (3.24)

    for any \(\varepsilon >0\) and with K some g-independent constant. Using this bound with some fixed \(\varepsilon <2\) shows that, \(\forall g \in \mathbb {C}_\pi \) the series in Eq. (3.23) has infinite radius of convergence in N.

  2. 2.

    For \(g\in \mathbb {C}_\pi \), \(Z_n(g)\) has the perturbative expansion:

    $$\begin{aligned} Z_n(g)&\simeq \sum _{m\ge n/2} \left( -\frac{2g}{3}\right) ^m \, \frac{(2m)!}{2^{2m} m!} \sum _{\begin{array}{c} m_1,\ldots , m_{2m-n+1} \ge 0 \\ \sum k m_k=2m , \; \sum m_k=n \end{array}} \frac{(-1)^n n!}{\prod _k k^{m_k}m_k!} \nonumber \\&\equiv \, Z_n^{\mathrm{pert.}}(g) \; . \end{aligned}$$
    (3.25)
  3. 3.

    The functions \(Z_n(g)\) are Borel summable along all the directions in \(\mathbb {C}_\pi \).

  4. 4.

    \(Z_n(g)\) can be continued past the cut on the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\), and the small-N series has infinite radius of convergence in N in this domain. However, \(\mathbb {R}_-\) is a Stokes line and the anticlockwise and clockwise analytic continuations \(Z_{n+}(g)\) and \(Z_{n-}(g)\) are not equal and cease to be Borel summable at \(\mathbb {R}_-\).

  5. 5.

    For \(g\in \mathbb {C}_{3\pi /2}\), \(Z_n(g)\) has the following transseries expansion:

    $$\begin{aligned} Z_n(g) \simeq Z_n^{\mathrm{pert.}}(g) + \eta \, e^{\frac{3}{2g}} Z_n^{(\eta )}(g) \;, \end{aligned}$$
    (3.26)

    with \(Z_n^{\mathrm{pert.}}(g)\) as in Eq. (3.25), and:

    $$\begin{aligned} Z_n^{(\eta )}(g)&= \frac{\imath }{\sqrt{2\pi }}\sqrt{ \frac{ g}{3}} \sum _{q=0}^{\infty } \sum _{p=0}^{n} \frac{1}{q!} \left( \frac{g}{6}\right) ^q \left( {\begin{array}{c}n\\ p\end{array}}\right) \frac{d^p \Gamma (z)}{\textrm{d}z^p}\Big |_{z=2q+1} \nonumber \\&\quad \times \left[ \left( \ln \left( e^{\imath \tau \pi } \frac{g}{3}\right) -\imath \pi \right) ^{n-p} -\left( \ln \left( e^{\imath \tau \pi } \frac{g}{3}\right) +\imath \pi \right) ^{n-p}\right] \; , \end{aligned}$$
    (3.27)

    with \(\tau = -\textrm{sgn}(\varphi )\) and \(\eta \) a transseries parameter which is zero on the principal Riemann sheet and one if \(|\varphi |>\pi \). Proceeding in parallel to Proposition 1, one can study the full monodromy of \(Z_n(g)\).

  6. 6.

    The discontinuity on the negative axis has the following asymptotic expansion:

    $$\begin{aligned} \textrm{disc}_\pi \big (Z_n(g)\big )&\simeq \frac{e^{-\frac{3}{2|g|}}}{\sqrt{2\pi }}\sqrt{\frac{|g|}{3}} \sum _{q=0}^{\infty } \sum _{p=0}^{n} \frac{1}{q!} \left( -\frac{|g|}{6}\right) ^q \left( {\begin{array}{c}n\\ p\end{array}}\right) \frac{d^p \Gamma (z)}{\textrm{d}z^p}\Big |_{z=2q+1}\nonumber \\&\quad \times \left[ \left( \ln |\frac{g}{3}| -\imath \pi \right) ^{n-p} -\left( \ln |\frac{g}{3}| +\imath \pi \right) ^{n-p}\right] \; . \end{aligned}$$
    (3.28)

    Summing over n, the discontinuity of the partition function (3.15) is recovered.

  7. 7.

    The functions \(Z_n(g)\) obey a tower of linear, inhomogeneous ordinary differential equations:

    $$\begin{aligned}&Z_0(g)=1 \;,\nonumber \\&4 g^2 Z_1^{\prime \prime }(g)+6 \left( g+1 \right) Z_1^{\prime }(g) = 1 \;,\nonumber \\&4 g^2 Z_n^{\prime \prime }(g)+6 \left( g+1 \right) Z_n^{\prime }(g) = n \left( 4 g Z_{n-1}^{\prime }(g) + Z_{n-1}(g) \right) \nonumber \\&\quad - n (n-1) Z_{n-2}(g) \;. \end{aligned}$$
    (3.29)

    which can be used to reconstruct the resurgent transseries expansion of \(Z_n(g)\).

Proof

See Appendix C.2\(\square \)

As with Proposition 1, the most interesting points are Properties 4 and 5. Again the analytic continuations \(Z_{n\pm }(g)\) of \(Z_n(g)\) to the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\) are obtained by tilting the integration contour to \(e^{\mp \imath \theta }\sigma \) with \(\theta > 0\). The branch point \(\sigma _{\star }\) of the integrand in Eq. (3.23) crosses the real axis when g reaches \(\mathbb {R}_-\), and deforming the tilted contours back to the real axis detaches Hankel contours around the cut \( \sigma _\star \times (1,+\infty )\):

$$\begin{aligned} Z_{n\pm }(g)\big |_{\varphi \genfrac{}{}{0.0pt}{}{>\pi }{< -\pi }}&=\int _{ e^{ \mp \imath \theta } \mathbb {R}} [\textrm{d}\sigma ] \; e^{-\frac{1}{2} \sigma ^2} \left( \ln \left( 1 - \imath \sqrt{ \frac{g}{3} } \sigma \right) \right) ^{n} = Z^{\mathbb {R}}_n(g) + Z^C_{n\pm }(g) \nonumber \\ Z^{\mathbb {R}}_n(g)&= \int _{ \mathbb {R}} [\textrm{d}\sigma ] \;e^{-\frac{1}{2} \sigma ^2} \left( \ln \left( 1 - \imath \sqrt{ \frac{g}{3} } \sigma \right) \right) ^{n} \;,\nonumber \\ Z^C_{n\pm }(g)&= \int _{C} [\textrm{d}\sigma ] \;e^{-\frac{1}{2} \sigma ^2} \left( \ln \left( 1 - \imath \sqrt{ \frac{g}{3} } \sigma \right) \right) ^{n} \; . \end{aligned}$$
(3.30)

The transseries of \(Z_n(g)\) is obtained by summing the asymptotic expansions of the two pieces:

$$\begin{aligned} Z_{n}^{\mathbb {R}}(g) \simeq Z_n^{\mathrm{pert.}}(g) ,\qquad Z_{n\pm }^C(g) \simeq e^{\frac{3}{2g}} Z^{(\eta )}_{n}(g)\big |_{\tau = \mp 1} . \end{aligned}$$

Notice that the homogeneous equation in Property 7 in Proposition 2 is the same for all \(n\ge 1\), and it admits an exact solution in the form of a constant plus an imaginary error function:

$$\begin{aligned} 4g^2 Z_1^{\prime \prime }(g) + 6(g+1)Z_1^\prime (g)=0 \qquad \Rightarrow \qquad Z_1(g)=c_1+ c_2 \int _0^{\imath \sqrt{ \frac{3}{2g} } }e^{-t^2} \textrm{d}t \; . \end{aligned}$$
(3.31)

The asymptotic expansion of the error function reproduces the one-instanton contribution of Eq. (3.26) for \(n=1\). For \(n>1\), instead, this is only part of the instanton contribution, the rest being generated by the recursive structure of the inhomogeneous equations. Similarly, the perturbative expansion comes from the special solution to the inhomogeneous equation, even at \(n=1\), as for those we cannot match exponential terms with the right-hand side. For \(n>1\) the homogeneous equation remains the same, but the inhomogeneous part depends on the solutions to previous equations, and thus, it can also contain exponential terms.

4 The Free Energy W(gN)

We now turn to the free energy \(W(g,N) = \ln Z(g,N)\). Our aim is find the equivalent of the results listed in Proposition 1, in the case of W(gN). Taking the logarithm has drastic effects: the nonperturbative effects encountered in W(gN) are significantly more complicated than the ones encountered for Z(gN). One can understand this from the fact that the linear differential equation satisfied by Z(gN) translates into a nonlinear one for W(gN), leading to an infinite tower of multi-instanton sectors in the transseries [6]. Here we will follow a different route, based on the small-N expansion.

Much like the partition function Z(gN), its logarithm W(gN) can also be expanded in N:

$$\begin{aligned} W(g,N) = \ln (Z(g,N)) \equiv \sum _{n\ge 1}\frac{1}{n!} \left( -\frac{N}{2}\right) ^n W_n(g) \;. \end{aligned}$$
(4.1)

The coefficients \(W_n(g)\) can be computed in terms of \(Z_n(g)\). As already mentioned, \(Z_n(g)\) are the moments of the random variable \(\ln (1-\imath \sqrt{g/ 3} \, \sigma )\); hence, \(W_n\) are the cumulants of the same variable and can be computed in terms of \(Z_n(g)\) by using the Möbius inversion formula (which in this case becomes the moments-cumulants formula). Let us denote \(\pi \) a partition of the set \(\{1,\dots n\}\), \(b\in \pi \) the parts in the partition, \(|\pi |\) the number of parts of \(\pi \) and |b| the cardinal of b. Then:

$$\begin{aligned} Z_n (g)= \sum _{\pi } \prod _{b\in \pi } W_{|b|} (g) \;,\qquad W_n (g) = \sum _{\pi } \lambda _{\pi } \prod _{b\in \pi } Z_{|b|} (g) \;, \end{aligned}$$
(4.2)

where \(\lambda _{\pi } = (-1)^{|\pi |-1} (|\pi |-1)!\) is the Möbius function on the lattice of partitions. Grouping together the partitions having the same number \(n_i\) of parts of size i, this becomesFootnote 7

$$\begin{aligned} Z_n(g)= & {} \sum _{\begin{array}{c} n_1,\dots ,n_n \ge 0\\ \sum in_i = n \end{array} } \frac{n!}{\prod _i n_i! (i!)^{n_i}} \prod _{i=1}^{n} W_i(g)^{n_i} \;,\nonumber \\ W_n(g)= & {} \sum _{k= 1}^{n} (-1)^{k-1} (k-1)! \sum _{ \begin{array}{c} n_1,\dots , n_{n-k+1}\ge 0 \\ \sum in_i = n ,\, \sum n_i = k \end{array}} \frac{n!}{\prod _i n_i! (i!)^{n_i}} \prod _{i=1}^{n-k+1} Z_i(g)^{n_i} \;.\nonumber \\ \end{aligned}$$
(4.3)

Equation (4.3) relates \(W_n(g)\) and \(Z_n(g)\) as analytic functions of g. However, this translates into a relation between W(gN) and Z(gN) which holds only in the sense of formal power series in N. Even though Z(gN) is analytic in some domain, one cannot conclude that W(gN) is also analytic in the same domain: convergence of the series defining Z(gN) does not imply convergence of the series defining W(gN) in Eq. (4.1). This can most readily be seen at the zeros of the partition function, the so-called Lee–Yang zeros, which are singular points for the free energy. In order to study the analyticity properties of W(gN) one needs to use a completely different set of tools. However, as we will see below, the Möbius inversion has its own uses: it is the most direct way to access the transseries expansion of W(gN).

4.1 Constructive Expansion

The following Proposition 3 is a slight variation on the Loop Vertex Expansion (LVE) introduced in [18] (see also [26] for more details). It gives an integral representation for \(W_n(g)\) in Eq. (4.1) which allows us to prove that W(gN) is convergent (hence analytic) in a bounded domain on the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\), wrapping around the branch point at the origin.

In Proposition 1 we fixed N to be a real parameter. However, Eq. (3.23) writes Z(gN) as an expansion in N with a nonzero (infinite!) radius of convergence in N, as long as \(|\varphi | < 3\pi /2\) (note that the bound in Property 1 of Proposition 2 suffices only for \(|\varphi |< \pi \); in order to reach \(|\varphi | < 3\pi /2\), one needs to use the improved bound in Eq. (C.42)). We can therefore extend N to a larger domain in the complex plane. As the following proposition shows, something similar applies also to W(gN), but with a finite radius of convergence.

Notation. Let us denote \(T_n\) the set of combinatorial trees with n vertices labeled \(1,\dots n\). There are \(\frac{ (n-2)! }{ \prod _{i=1}^n (d_i-1)!}\) trees over n labeled vertices with coordination \(d_i\) at the vertex i and \(\sum _i d_i = 2(n-1)\). The total number of trees in \(T_n\) is \(n^{n-2}\). Let \(\mathcal {T}\in T_n\) be such a tree. We denote \(P_{k-l}^{\mathcal {T}} \) the (unique) path in the tree \(\mathcal {T}\) connecting the vertices k and l. If we associate to each edge \((k,l)\in \mathcal {T}\) a variable \(u_{kl}\) between 0 and 1, we can define the \(n\times n\) matrix \(w^\mathcal {T}\):

$$\begin{aligned} w^\mathcal {T}_{kl} \equiv {\left\{ \begin{array}{ll} 1 \;,\qquad &{} \text {if} \;\; k=l \\ \inf _{(i,j)\in P_{k-l}^{\mathcal {T}} } \{ u_{ij} \} \;, \qquad &{}\text {else} \end{array}\right. } \; . \end{aligned}$$
(4.4)

The matrix \(w^\mathcal {T}\) is a positive matrix for any choice of u parameters and is strictly positive outside a set of measure 0 (see Appendix D for more details). Of course the matrix w depends on u, but we suppress this in order to simplify the notation.

Proposition 3

(The LVE, analyticity). Let N be a fixed complex parameter and let us denote \(g= |g|e^{\imath \varphi }\). The cumulants \(W_n(g)\) can be written as:

$$\begin{aligned} W_1(g)&=Z_1(g)= \int _{-\infty }^{+\infty } [\text {d}\sigma ] \; e^{-\frac{1}{2} \sigma ^2 } \ln \Big [ 1 -\imath \sqrt{\tfrac{g}{3}} \sigma \Big ] \;, \nonumber \\ W_n(g)&= - \left( \frac{g}{3} \right) ^{n-1} \sum _{ \mathcal {T}\in T_n } \int _{0}^1 \prod _{ (i,j) \in \mathcal {T}} \text {d}u_{ij} \int _{-\infty }^{+\infty } \frac{\prod _i [\text {d}\sigma _i] }{\sqrt{ \det w^\mathcal {T}}} \; e^{ - \frac{1}{2} \sum _{i,j} \sigma _i ( w^{\mathcal {T}})_{ij}^{-1} \sigma _j }\nonumber \\ {}&\quad \times \prod _{i} \frac{(d_i-1)!}{ \left( 1 -\imath \sqrt{\tfrac{g}{3}} \sigma _i \right) ^{d_i} } \;, \end{aligned}$$
(4.5)

where we note that the Gaussian integral over \(\sigma \) is well defined, as \(w^\mathcal {T}\) is positive, and normalized. Furthermore:

  1. 1.

    The functions \(W_n(g), n\ge 2\) are bounded by:

    $$\begin{aligned} |W_n(g)| \le \frac{(2n-3)!}{(n-1)!} \left| \frac{g}{3 \left( \cos \frac{\varphi }{2} \right) ^{2} } \right| ^{n-1} \;. \end{aligned}$$
    (4.6)

    Therefore, they are analytic in the cut plane \(\mathbb {C}_{\pi }\).

  2. 2.

    The series

    $$\begin{aligned} W(g,N) = \sum _{n\ge 1} \frac{1}{n!} \left( -\frac{N}{2} \right) ^{n} W_n(g) \end{aligned}$$
    (4.7)

    is absolutely convergent in the following cardioid domain:

    $$\begin{aligned} \mathbb {D}_0 = \left\{ g\in \mathbb {C},\; g=|g| e^{\imath \varphi } :\; |g|<\frac{1}{|N|} \; \frac{3}{2} (\cos \frac{\varphi }{2} )^{2}\right\} \;. \end{aligned}$$
    (4.8)
  3. 3.

    \(W_n(g)\) can be analytically continued to a subdomain of the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\) by tilting the integration contours to \(\sigma \in e^{-\imath \theta } \mathbb {R}\):

    $$\begin{aligned} W_{1\theta }(g)&= e^{-\imath \theta }\int _{-\infty }^{+\infty } [\text {d}\sigma ] \; e^{-\frac{1}{2} e^{-2\imath \theta } \sigma ^2 } \ln \Big ( 1 -\imath \sqrt{\tfrac{g}{3}} e^{-\imath \theta }\sigma \Big ) \;, \nonumber \\ W_{n\theta }(g)&= - \left( \frac{g}{3} \right) ^{n-1} \sum _{ \mathcal {T}\in T_n } \int _{0}^1 \prod _{ (i,j) \in \mathcal {T}} \text {d}u_{ij} \nonumber \\ {}&\quad \times \int _{\mathbb {R}} \frac{\prod _i e^{-\imath \theta } [ \text {d}\sigma _i] }{\sqrt{ \det w^\mathcal {T}}} \; e^{ - \frac{1}{2} e^{-2\imath \theta }\sum _{i,j} \sigma _i ( w^\mathcal {T})_{ij}^{-1} \sigma _j } \;\; \prod _{i} \frac{(d_i-1)!}{ \left( 1 -\imath \sqrt{\tfrac{g}{3}} e^{-\imath \theta }\sigma _i \right) ^{d_i} } \;. \end{aligned}$$
    (4.9)
  4. 4.

    For \(n\ge 2\) we have the following bound:

    $$\begin{aligned} |W_{n\theta }(g)| \le \frac{(2n-3)!}{(n-1)!} \frac{1}{\sqrt{ \cos (2\theta ) }} \left| \frac{g}{3 \sqrt{\cos (2\theta )} \left( \cos \frac{\varphi -2\theta }{2} \right) ^2 } \right| ^{n-1} \;. \end{aligned}$$
    (4.10)
  5. 5.

    The series

    $$\begin{aligned} W_{\theta }(g,N) = \sum _{n\ge 1} \frac{1}{n!} \left( -\frac{N}{2} \right) ^{n} W_{n\theta }(g) \;, \end{aligned}$$
    (4.11)

    is absolutely convergent in the following domain:

    $$\begin{aligned} \mathbb {D}_{\theta } = \left\{ g\in \mathbb {C},\; g=|g| e^{\imath \varphi } :\; |g| < \frac{1}{|N|} \; \frac{3}{2} \; \left( \cos \frac{\varphi -2 \theta }{2} \right) ^{2} \; \sqrt{ \cos (2\theta ) } \right\} \;. \end{aligned}$$
    (4.12)

Consequently, \(W_n(g)\) and W(gN) can be analytically extended to the following respective domains:

$$\begin{aligned} W_{n}(g) : \qquad&|2\theta |< \frac{\pi }{2} \; , \;\; |\varphi - 2\theta |< \pi \;, \nonumber \\ W(g,N) : \qquad&|2\theta |< \frac{\pi }{2} \; , \;\; |\varphi - 2\theta |< \pi \;, \qquad |g| < \frac{1}{|N|} \; \frac{3}{2} \; \left( \cos \frac{\varphi -2 \theta }{2} \right) ^{2} \; \sqrt{ \cos (2\theta ) }. \end{aligned}$$
(4.13)

Pushing \(\theta \rightarrow \pm \pi /4\) allows us to write a convergent expansion for all \(|\varphi |< \frac{3\pi }{2}\).

Proof

See Appendix C.3\(\square \)

The main point of the proposition is that by constructive methods we can prove analyticity of W(gN) in a nontrivial domain. In a first step, without touching the integration contours, we prove that such domain is the cardioid of Eq.(4.8). However, the cardioid does not allow us to reach (and cross) the branch cut. Tilting the integration contours by \(\theta \), we are able to extend the original cardioid domain to the larger domain of Eq. (4.12) (see Fig. 2), going beyond the cut on a subdomain of the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\). The optimal domain \(\mathbb {D}_{\textrm{opt}}\) can be found by maximizing the right-hand side of Eq. (4.12) with respect to \(\theta \), at fixed \(\varphi \), but a simpler and qualitatively similar choice is to take \(\theta =\varphi /6\).

Fig. 2
figure 2

The cardioid domain \(\mathbb {D}_0\) of Eq. (4.8) (dotted blue line) and the extended cardioid \(\mathbb {D}_{\theta }\) of Eq. (4.12) (red line), for \(\theta =\varphi /6\), which is similar to the optimal domain \(\mathbb {D}_{\textrm{opt}}\), in the complex g-plane. The branch cut is on the negative real axis; thus, the portions of \(\mathbb {D}_{\theta }\) going beyond it are to be understood as being on different Riemann sheets

Note that the domain of analyticity of W(gN), Eq. (4.12), depends on N and shrinks to zero for \(N\rightarrow \infty \). Results uniform in N can only be established if one keeps the ’t Hooft coupling \(g_{t} = g N\) fixed [26]. On the other hand, for any g on the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\), the radius of convergence of the LVE in N is nonzero.

Remark 2

It is also worth noticing that the explicit expressions for the partition function in terms of special functions, discussed around Eq. (3.21), provide us with some useful information about the zeros of Z(gN) (Lee–Yang zeros), and hence about the singularities of W(gN). For example, in the case \(N=1\), the partition function is expressed in terms of a modified Bessel function of the second kind, whose zeros have been studied in some depth. In particular, from what is known about \(K_{\nu }(z)\) (e.g., [31]) we deduce that \(Z(g,1)=\sqrt{\frac{3}{2\pi g}} e^{\frac{3}{4\,g}} K_{1/4}(\frac{3}{4\,g})\) has no zeros in the principal sheet \(\mathbb {C}_{\pi }\), while on each of the two following sheets it has an infinite sequence of zeros approaching the semiaxis at \(|\varphi |=3\pi /2\) from the left, and accumulating toward \(g=0\) (see Fig. 3). Therefore, it should come as no surprise that W(gN) cannot be analytically continued around the origin beyond \(|\varphi |=3\pi /2\).

Fig. 3
figure 3

Approximate location (see [31]) of the Lee–Yang zeros of Z(g, 1) (blue dots) in the quadrant \(\pi<\varphi <3\pi /2\) of \(\mathbb {C}_{3\pi /2}\), together with the boundary of the domain \(\mathbb {D}_{\theta }\) (in red)

Remark 3

Integrating out the u parameters and performing the sum over trees, one should be able to prove that integral expressions (4.9) reproduce the moment-cumulant relation in Eq. (4.3). In particular this would provide an alternative proof that the moment cumulant relation holds in the sense of analytic functions on the Riemann surface. The proof that this indeed happens is involved as the summation over trees requires the use of combinatoiral techniques similar to the ones discussed in Appendix D. We postpone this for future work.

In [18] the LVE is used to prove the Borel summability of W(g, 1) along the positive real axis. Building on the techniques introduced in [18], we now generalize this result.

Proposition 4

(Borel summability of \(W_n(g)\) and W(gN) in \(\mathbb {C}_\pi \)). The cumulants \(W_n(g)\) and the free energy W(gN) at any fixed complex N are Borel summable along all the directions in the cut complex plane \(\mathbb {C}_\pi \).

Proof

See Appendix C.4. \(\square \)

4.2 Transseries Expansion

It is well known that the (perturbative) asymptotic series of W(gN) at \(g=0\) is a sum over connected Feynman graphs. The connection between the LVE of W(gN) presented in Proposition 3 and the Feynman graphs is discussed in Appendix D.1. On the other hand, the power series in each multi-instanton sector of the transseries of W(gN) has no simple diagrammatic interpretation; they can be constructed from the nonlinear differential equation obeyed by W(gN), or more formally by expanding the logarithm of the transseries expansion of Z(gN) in powers of the transseries monomial \(\exp \{3/(2g)\}\) (e.g., [6]). The latter is, however, only a meaningful operation in the sense of formal power series.

In this section we take a different route and derive rigorously the transseries expansion of W(gN) by exploiting the analytical control we have on the small-N expansion. We first notice that, from Propositions 2 and 3, \(W_n(g)\) and \(Z_n(g)\) are analytic functions on the extended Riemann sheet \(\mathbb {C}_{3\pi /2}\). Next, we use Eq. (4.3) to construct \(W_n(g)\) as a finite linear combination of finite products of \(Z_i(g)\)’s. Each such product is in fact a (factored) multidimensional integral; hence, we can apply to it the steepest descent method to obtain its asymptotic expansion. In \(\mathbb {C}_\pi \), the asymptotic expansion of each factor \(Z_i(g)\) is of the perturbative type, Eq. (3.25), and \(W_n(g)\) is just a finite linear combination of Cauchy products of such series.

When turning g past the negative real axis, each integration contour in this multidimensional integral must be deformed past a cut and each \(Z_{i\pm }(g) = Z_i^{\mathbb {R}}(g) + Z^C_{i\pm }(g)\) (see the discussion below Proposition 2). It follows that \(W_i(g)\) is a linear combination of products involving \(Z_i^{\mathbb {R}}(g)\)’s and \(Z^C_{i\pm }(g)\), and this representation holds in the sense of analytic functions on the Riemann surface. In order to obtain the transseries of \(W_n(g)\), one needs to build the transseries expansion of each of the terms in the linear combinations. As the multidimensional integrals are factored, this is just the Cauchy product of the transseries \(Z_i^\mathrm{pert.}(g)\) and \(e^{\frac{3}{2g}}Z_i^{(\eta )}(g)\) corresponding to \(Z_i^{\mathbb {R}}(g)\) and \(Z^C_{i\pm }(g)\), respectively.

The summation over n is more delicate, as it is an infinite series. As we have seen in Proposition 3, the small-N series of W(gN) converges in the domain \(\mathbb {D}_0\) of Eq. (4.12), thus yielding W(gN) in terms of \(W_n(g)\) as an analytic function on such domain. Therefore, we can apply the steepest-descent method term by term to the small-N series, and hence, the transseries of W(gN) is rigorously reconstructed by substituting the transseries for \(W_n(g)\) in Eq. (4.1).

Unsurprisingly, at the end we recover the formal transseries of W(gN) which can be obtained by direct substitution of the transseries expansion of Z(gN), taking formally its logarithm, and then expanding in powers of \(Z_i^{(\eta )}(g)\) and \(Z_i^\mathrm{pert.}(g)-1\). What we gained in the process is that we replaced a formal manipulation on transseries with a rigorous manipulation on analytic functions.

Proposition 5

The cumulant \(W_n(g)\) and the full free energy W(gN) have transseries expansions that can be organized into instanton sectors. The instanton counting parameter is denoted by p.

  1. 1.

    For \(g\in \mathbb {C}_{3\pi /2}\),, the cumulant \(W_n(g)\) has the transseries expansion:

    $$\begin{aligned} W_n(g)= \sum _{p=0}^n e^{\frac{3}{2g}p} \;\Bigg (\eta \sqrt{2\pi } \sqrt{\frac{g}{3}}\Bigg )^{p}\; \sum _{l'= 0}^{n-p} \left( \ln \left( \tfrac{g}{3}\right) \right) ^{l'} \sum _{l\ge 0} g^l\, W^{(p)}_{n;l,l'} \,, \end{aligned}$$
    (4.14)

    where \(\mathbb {R}_-\) is a Stokes line, \(\tau = -\textrm{sgn}(\varphi )\) and \(\eta \) is a transseries parameter which is zero on the principal Riemann sheet and is one when \(|\varphi |>\pi \). The g-independent coefficient \(W^{(p)}_{n;l,l'}\) is given by the following nested sum:

    $$\begin{aligned} {}&{} W^{(p)}_{n;l,l'} = \sum _{\begin{array}{c} k = p \\ k+p\ge 1 \end{array}}^{n} (-1)^{k-1} (k-1)! \sum _{ \begin{array}{c} n_1, \ldots , n_{n-k+1} \ge 0 \\ \sum in_i = n ,\, \sum n_i = k \end{array}} \ \sum _{\begin{array}{c} \{0\le p_i\le n_i\} \\ _{i=1,\dots ,n-k+1} \\ \sum p_i=p \end{array}} \frac{n!}{\prod _i (n_i-p_i)!p_i! (i!)^{n_i}} \nonumber \\ {}{}&{} \quad \times \sum _{\begin{array}{c} \{a^i_j\ge 0\}^{i=1,\dots ,n-k+1}_{j=1,\dots ,n_i} \\ \sum _i\sum _j a^i_j=l \end{array}} \ \sum _{\begin{array}{c} \{0\le c^i_j\le i-1\}^{i=1,\dots ,n-k+1}_{j=1,\dots ,p_i} \\ \sum _i\sum _j c^i_j=l' \end{array}} \left( \frac{1}{6}\right) ^{\sum _{i=1}^{n-k+1}\sum _{j=1}^{p_i}a^i_j} \left( -\frac{2}{3}\right) ^{\sum _{i=1}^{n-k+1}\sum _{j=p_i+1}^{n_i}a^i_j} \nonumber \\ {}{}&{} \qquad \prod _{i=1}^{n-k+1} \Bigg (\prod _{j=1}^{p_i} G(a^i_j,c^i_j;i)\Bigg )\Bigg (\prod _{j=p_i+1}^{n_i} G(a^i_j;i)\Bigg ) \; , \end{aligned}$$
    (4.15)

    with

    $$\begin{aligned} G(a;i)&= \frac{(2a)!}{2^{2a} a!} \sum _{\begin{array}{c} a_1,\ldots ,a_{2a-i+1} \ge 0 \\ \sum k a_k=2a , \; \sum a_k=i \end{array}} \frac{(-1)^i i!}{\prod _k k^{a_k}a_k!} \; , \nonumber \\ G(a,c;i)&= \sum _{b=0}^{i-1} \left( \imath \tau 2\pi \right) ^{i-1-b-c} \frac{i!}{a!\,b!\,c!\,(i-b-c)!} \frac{d^b \Gamma (z)}{\textrm{d}z^b}\Big |_{z=2a+1} \;. \end{aligned}$$
    (4.16)
  2. 2.

    For \(g\in \mathbb {D}_{\theta }\), the full free energy W(gN) has the transseries expansion:

    $$\begin{aligned} W(g,N)&= \sum _{n\ge 1} \frac{1}{n!} \left( -\frac{N}{2}\right) ^n W_n(g) \nonumber \\&= \sum _{p\ge 0} e^{\frac{3}{2g}p} \; \Bigg (\eta \sqrt{2\pi }e^{\imath \tau \frac{\pi }{2}} \left( \frac{e^{\imath \tau \pi } g}{3}\right) ^{\frac{1-N}{2}}\Bigg )^{p} \; \sum _{l\ge 0} \left( - \frac{2g}{3}\right) ^{l} W^{(p)}_{l}(N) \; , \end{aligned}$$
    (4.17)

    where

    $$\begin{aligned} W^{(p)}_{l}(N)&= \sum _{\begin{array}{c} q\ge 0\\ p+q\ge 1 \end{array}} (-1)^{p+q-1} \frac{(p+q-1)!}{p!q!} \nonumber \\ {}&\quad \times \sum _{\begin{array}{c} n_1,\dots ,n_q\ge 1 \\ m_1,\dots ,m_p\ge 0 \\ \sum n_i+\sum m_j=l \end{array}} \Bigg ( \prod _{i=1}^{q} \frac{\Gamma (2n_i+N/2) }{2^{2n_i}n_i! \, \Gamma (N/2) }\Bigg ) \Bigg ( \prod _{j=1}^{p} \frac{(-1)^{m_j}}{ 2^{2m_j} m_j! \; \Gamma \left( \frac{N}{2} -2m_j\right) }\Bigg ) \;. \end{aligned}$$
    (4.18)

Proof

See Appendix C.5\(\square \)

While the expressions in Proposition 5 are not the most amenable to computations, one feature is striking. Expanding the cumulant \(W_n(g)\) into p instanton sectors, we observe that only the first n instantons contribute to \(W_n\), that is the sum in Eq. (4.14) truncates to \(p=n\). The n instanton contribution to \(W_n\) comes from \(n=p=k\) in Eq. (4.15) which implies \(n_1=n\) and all the others 0, henceFootnote 8:

$$\begin{aligned} W_n^{(n)}(g) \simeq e^{\frac{3}{2g} n } \left( \eta \sqrt{2\pi } \sqrt{\frac{g}{3}}\right) ^n (-1)^{n-1} (n-1)! \left( \sum _{q=0}^{\infty } \frac{(2q)!}{q!} \left( \frac{g}{6}\right) ^q \right) ^n \;. \end{aligned}$$
(4.19)

This is genuinely new phenomenon. Usually, for quantities that are interesting for physics, one either deals with functions having only one instanton, like Z(gN) (or \(Z_n(g)\)) or with function receiving contributions from all the instanton sectors, like W(gN). This is, to our knowledge, the first instance when some physically relevant quantity receiving contributions from a finite number of instantons strictly larger than one is encountered. The n instanton contribution comes from \(n_1=n\) and all the others 0, such that effectively:

$$\begin{aligned} W_n^{(n)}(g) \approx (Z_1(g))^n \;, \end{aligned}$$

and, for all n, \(Z_n(g)\) has just a single instanton.

4.3 Differential Equations

The exotic behavior of \(W_n(g)\) can also be understood in terms of differential equations. By rewriting the partition function as \(Z(g,N)=e^{W(g,N)}\), it is straightforward to turn (3.16) into a differential equation for W(gN), which in turn implies a tower of equations for \(W_n(g)\).

Proposition 6

The function W(gN) obeys the nonlinear differential equation:

$$\begin{aligned} 16 g^2 W^{\prime \prime }(g,N) + 16g^2\left( W^{\prime }(g,N)\right) ^2 + \left( (8N+24)g+24 \right) W^\prime (g,N)+N(N+2)=0 . \end{aligned}$$
(4.20)

The functions \(W_n(g)\) obey the tower of differential equations:

$$\begin{aligned}&4 g^2 W_1^{\prime \prime }(g)+6(g+1)W_1^\prime (g)-1=0, \nonumber \\&4 g^2 W_2^{\prime \prime }(g)+6(g+1)W_2^\prime (g)+8g^2\left( W_1^\prime (g)\right) ^2 -8g W_1^\prime (g)+2=0, \nonumber \\&4 g^2 W_n^{\prime \prime }(g)+6(g+1)W_n^\prime (g)+4 g^2 \sum _{k=1}^{n-1}\left( {\begin{array}{c}n\\ k\end{array}}\right) W_{n-k}^\prime W_{k}^{\prime }-4n g W_{n-1}^\prime (g)=0 \;. \end{aligned}$$
(4.21)

The differential equation for \(W_{1}(g)\) is, unsurprisingly, identical to the one for \(Z_1(g)\) (the connected \(1-\)point function equals to the full \(1-\)point function). Note that although the differential equation for W(gN) is nonlinear, the one for \(W_n(g)\) is linear. In fact, since \(W_0(g)=0\), the nonlinear term \(\left( W^\prime (g,N)\right) ^2\) produces only source terms in (4.21). The linearity of the equations provides another point of view on why only a finite number of instantons arise in each \(W_n(g)\).