Asymptotic expansion of a partition function with a complex potential
The blog post will be an accessible introduction to my recent paper, joint with A. Guionnet and K. Kozlowski, Asymptotic expansion of the partition function for β-ensembles with complex potentials. This post is based off presentations I have given of this paper. However because of time constraints many “tricks” from the paper are left out from the presentation. This post gives me the opportunity to explain such tricks.
Real integrals
In many problems in mathematics one encounters integrals of the form
IN=∫abeNφ(x)dx
for some smooth (real-valued) function φ∈C∞([a,b]) and a<b being finite real numbers. We are interested in the behaviour of IN for large N. By the triangle inequality we can make the trivial bound
∣IN∣≤(b−a)eNsup[a,b]φ.
Then, in particular, we have
N→+∞limsupNlnIN≤[a,b]supφ.
In fact, the limit exists and we actually have equality here. To see this, observe that, by continuity of φ, for any ϵ>0 there must exist a subinterval Jϵ⊂I, of positive length ∣Jϵ∣>0, such that sup[a,b]φ≤φ(x)+ϵ for all x∈Jϵ. Then we have
IN≥∫JϵeNφ(x)dx≥∣Jϵ∣eNsup[a,b]φ−Nϵ.
Then
N→+∞liminfNlnIN≥[a,b]supφ−ϵ.
Since ϵ>0 was arbitrary, we conclude that NlnIN converges and
N→+∞limNlnIN=[a,b]supφ.
This analysis only gives the leading order of NlnIN. To go further we need further assumptions. Let us suppose that the supremum is attained at a unique point in the interior of the interval, x∗∈(a,b), and suppose that φ′′(x∗)<0. Then there exists a neighbourhood U of x∗ and a smooth function ψ∈C∞(U) such that
φ(x)=φ(x∗)−ψ(x)2.
Indeed, we can take this equation as definingψ and then find a sufficiently small neighbourhood of x∗ such that it is smooth and real valued. Furthermore ψ′(x)>0 for all x∈U. Then
IN=(1+O(e−NC))∫UeNφ(x)dx
for some C>0 and then
IN=(1+O(e−NC))eNφ(x∗)∫ψ−1(U)e−Nu2(ψ−1)′(u)du.
Taylor expanding (ψ−1)′ at 0 and integrating term by term (after taking the limits of integration to ±∞) we find an asymptotic series
IN∼−Nφ′′(x∗)2πeNφ(x∗)(1+NA1+N2A2+…).
Note that that we obtain an asymptotic series in N1 because all the odd integrals vanish. This is the Laplace method in brief.
Remark: Note that we have a kind of “central limit theorem” happening inside the integral, where if we think of the integrand as representing a distribution function, then N(x−x∗) is asymptotically Gaussian with mean 0 and variance −φ′′(x∗)1.
Contour integrals
Let us now repeat the analysis, where now a,b∈C and φ is analytic on some domain containing a and b. Then consider
IN=∫abeNφ(z)dz.
If we choose a path Γ[a,b] between the endpoints, we have
∣IN∣≤∣Γ[a,b]∣eNsupΓ[a,b]ℜφ
where ∣Γ[a,b]∣ is the arc-length of the contour. Hence
N→+∞limsupNln∣IN∣≤z∈Γ[a,b]supℜφ(z).
However by Cauchy’s theorem, the original integral is unchanged under deformations of the contour. Hence we may optimise our bound over some appropriate class of homotopic contours T.
This is the essence of the saddle point method. Suppose we found a curve Γ[a,b] and a point z∗∈Γ[a,b] such that the above inf-sup is actually attained—what would it look like? Well, along the curve ℜφ(z) would attain a maximum at z=z∗. However perpendicular to the curve ℜφ(z) must attain a minimum, since otherwise a lateral deformation of the contour could reduce the value of supz∈Γ[a,b]ℜφ(z). Hence z=z∗ must be a saddle point of ℜφ(z).
Next, we note that since φ is analytic, ℜφ is a harmonic function, and so has no local maxima or minima. (Alternatively one could apply the maximum modulus principle to eφ.) That is, the only stationary points of ℜφ (points where the two-dimensional gradient vanishes) are saddle points. And by the Cauchy-Riemann equations the gradient of ℜφ vanishes if and only if the holomorphic derivative vanishes, φ′(z)=0. Hence the inf-sup problem becomes one of finding these stationary points and appropriately traversing them. Indeed, we could choose the contour such that it traverses z∗ such that ℑφ is constant in a neighbourhood of z∗ along the contour, and this would essentially reduce the problem to a real integral discussed previously. Such a contour is called a steepest descent contour since it is the direction of steepest descent of ℜφ. In this way one can show that Nln∣IN∣ really converges to infΓ[a,b]∈Tsupz∈Γ[a,b]ℜφ(z) as N→+∞. (This way of deriving the saddle point method by trying to optimise the triangle inequality I learnt from de Bruijn’s book Asymptotic Methods in Analysis.)
Partition functions of β-ensembles
A β-ensemble is a random collection of N particles on the real line, x1,…,xN∈R, with joint density (with respect to Lebesgue measure on RN)
β is a fixed positive parameter, which can be interpreteted as the inverse of the “temperature” β=T1>0. This is because ϱN can be interpreted as the thermal distribution of N particles with Hamiltonian
V is a confining potential which grows sufficiently fast such that the density is normalisable. For “special” potentials V there is a matrix-model representation of β-ensembles (Dumitriu-Edelman, 2002) and also for more general V if β=1,2,4. The asymptotic expansion of lnZN[V], also known as the free energy, is a central question of interest in statistical mechanics. A full asymptotic expansion was achieved, in a certain off-critical regime, by Borot and Guionnet (see here and here). The integrand of ZN[V] depends on N but so does the number of integrations, so the Laplace method breaks down. Instead Borot and Guionnet use logarithmic potential theory and the method of Dyson-Schwinger equations to obtain their asymptotic expansion. Thus these methods could be regarded as an ∞-dimensional version of the Laplace method.
Let me outline how the logarithmic potential theory arises. Define the empirical measure
We remark that {LN(x)}x∈RNN≥1 forms a dense subset of M1(R), the space of Borel probability measures on R. Thus if we ignore the diagonal we expect that
N→+∞limsupN2lnZN[V]≤−2βμ∈M1(R)infIV[μ]
where
IV[μ]=∫R2(ln∣x−y∣1+V(x)+V(y))dμ(x)⊗dμ(y).
In fact N2lnZN[V]→−2βinfμ∈M1(R)IV[μ].
A complex partition function
We now turn to the problem with which my paper with Guionnet and Kozlowski was concerned. This involves studying a “complexification” of the β-ensemble partition function ZN[V]. The first way we could complexify is to make the potential V complex. However, since we want to develop an analogue of the saddle point/steepest descent method, we want the integrand to be analytic. Thus let us take V to be a polynomial with complex coefficients, say of degree κ≥2. By rescaling we can choose the leading coefficient to be any positive number, e.g. V(z)=κ1zκ+O(zκ−1).
So that the integrand is analytic we should take β to be a positive integer. In fact, it should be an even integer because if β was odd then the interchange of any two integration variables would produce a minus sign, and by symmetry of the integration region ZN,Γ[V]=0. Thus β must be even for the problem to be nontrivial.
Finally, what conditions should we put on Γ?
Definition: Fix α,α′∈[[1,κ−1]] to be two distinct integers. We say a contour Γ is “admissible” (relative to (α,α′)) if
(1) It consists of a finite number of C1 arcs.
(2) It is connected.
(3) There exists an R>0 sufficiently large so that
Γ∖DR(0)=eκ2πiα[R,+∞)∪eκ2πiα′[R,+∞)
where DR(0) is the open disk of radius R centred at 0. We require incoming orientation on eκ2πiα[R,+∞) and outgoing orientation on eκ2πiα′[R,+∞). △
This means that outside a large compact set Γ consists of two parts: an incoming ray and an outgoing ray, and these rays should lie along the line proportional to a κth root of unity. A curve that satisfies all of the above is said to be “admissible.” These properties mean that ∣e−V(z)∣→0 rapidly along these rays. We could, of course, be less restrictive and allow contours that run “close” to the rays eκ2πiαR+∪eκ2πiα′R+, however what happens outside of a sufficiently large compact set will make no contribution to the asymptotic series, and we can also deform our contour to lie exactly upon these rays. Thus we should really think of ZN,Γ[V] as being a function of the homotopy class of Γ, which is labelled by (α,α′). Note that if we took α=α′ then ZN,Γ[V]=0, hence we exclude this trivial case.
Remark: Note that interchanging α and α′ changes ZN,Γ[V] by a factor of (−1)N. △
The central question of the paper is how lnZN,Γ[V] behaves as N→+∞. Following a similar reasoning to case of the partition function of a real β-ensemble, we have
N→+∞limsupN2ln∣ZN,Γ[V]∣≤−2βμ∈M1(Γ)infIV[μ]
where
IV[μ]=def∫C2(ln∣z−w∣1+φ(z)+φ(w))dμ(z)⊗dμ(w)
for φ=ℜV, and where M1(Γ) is the space of Borel probability measures on Γ. Then, optimising, we find
where T is the space of admissible contours (we suppress the α,α′ dependence since it is fixed throughout). If we believe this bound to be optimal then we have a kind of ∞-dimensional saddle point.
Fact: By potential theoretic arguments one can show that if Γ is admissible then IV has unique minimiser μΓ∈M1(Γ), which we call the equilibrium measure associated to Γ.
Definition:Γeq is said to solve the “max-min energy problem” if
Γ~∈Tsupμ∈M1(Γ~)infIV[μ]=IV[μΓeq].
Theorem (Silva-Kuijlaars, 2015): Let V be a polynomial (with complex coefficients). Then there exists an admissible Γeq solving the max-min energy problem. Γeq is not unique however μΓeq is unique (so that all solutions have the same equilibrium measure). △
Definition: (1) Define the logarithmic potential associated to a probability measure μ as
U[μ](z)=def∫Cln∣z−w∣1dμ(w)
wherever this makes sense. Of course, if z∈suppμ and μ is compactly supported then the integral converges.
(2) Similarly, define the Cauchy transform of probability μ
C[μ](z)=def2πi1∫Cw−z1dμ(w)
wherever this makes sense.
(3) We say that Γ is an S-curve in the external field φ if there is a set of zero capacity E such that for every z∈suppμΓ∖E there is a neighbourhood U of z such that suppμΓ∩U is an analytic arc, and
∂n+∂(U[μΓ]+φ)(z)=∂n−∂(U[μΓ]+φ)(z)
where ∂n+∂ and ∂n−∂ are the directional derivatives normal to the contour (away from the contour). △
Theorem (Silva-Kuijlaars, 2015): Let V be a polynomial (with complex coefficients) and Γeq be an associated solution to the max-min energy problem. Then we have the following.
1) suppμΓeq is a finite collection of (bounded) analytic arcs.
2) suppμΓeq is an S-curve in the external field φ=ℜV.
3) There exists a polynomial R such that
R(z)=(V′(z)+2πiC[μΓeq](z))2.
4) suppμΓeq consists of critical trajectories of the quadratic differential −R(z)dz2. △
Remark: 1) Taking the square root we find R(z)=V′(z)+2πiC[μΓeq](z). The left hand side is analytic everwhere except on its branch cuts. The right hand side is analytic everywhere except on the support of μΓeq. Hence the branch cuts of R(z) are the support of the equilibrium measure.
2) By Plemelj’s formula we have iπ1R(z)dz=dμΓeq(z)>0 whenever the density is positive. Squaring both sides (which amounts to ignoring the orientation of the contour) we find suppμΓeq is a critical trajectory of −R(z)dz2.
3) From the fact that R(z) changes sign across the branch cut we have R(z)++R(z)−=0. This gives
V′(z)+p.v.∫w−z1dμΓeq(w)=0,z∈suppμΓeq.
This equation can be regarded as a kind of “complexified” Euler-Lagrange equation, since taking its perpendicular to the contour one obtains the Euler-Lagrange and S-curve conditions. This again supports the interpretation of an ∞-dimensional saddle point, since it is analogous to the saddle point equation φ′(z)=0, which can also be broken down into two components representing the maximum along the curve and the minimum perpendicular to the curve.
The Euler-Lagrange condition for the minimisation of the energy on the curve states that there is a constant Cφ such that φ(z)+U[μΓeq](z)≥Cφ throughout the curve, with equality when z∈suppμΓeq. Hence let us define the effective potential as
φeff(z)=φ(z)+U[μΓeq](z)−Cφ.
Definition: We say that the potential V is regular if it is possible to choose Γeq such that
1) The polynomial R vanishes nowhere on Γeq except at the endpoints of suppμeq, and at these endpoints R has only simple zeros. (Note this is equivalent to saying that the density of the equilibrium measure as square-root type vanishing.)
2) φeff(z)>0 for all z∈Γeq∖suppμeq.
V is one-cut regular if it is regular and suppμeq is connected. △
It is believed that the regular potentials are “generic”, in the sense of being an open set of full measure within the space of all potentials. However for this max-min energy problem this is unproven (the claim that the regular potentials form an open set is a recent result of Bertola, Bleher, Gharakhloo, McLaughlin and Tovbis, 2022).
Definition: Define the “g-function” as
g[μΓeq](z)"="∫Γeqln(z−w)dμeq(w).
We put quotation marks because some care must be taken with the branch cut. To be more precise, we want to choose the branch cut of g[μΓeq] along Γeq and require that
where g±[μΓeq] are the left and right boundary values of the g-function up to the curve. △
Remark: Note that the real part of the complex energy functional is the real energy functional, which justifies the name, ℜIΓeq[μeqΓ]=IV[μeqΓ]△.
We now have enough tools to state the main theorem
Theorem: (Guionnet, Kozlowski, L., 2025) Let V(z)=κzκ+O(zκ−1) be a polynomial, and one-cut regular as a potential. Then there exists a sequence of complex numbers (Fk(β,V))k=−2∞ such that
Thus, the leading term at scale N2 has the form of a complex energy, whereas the term at scale N involves a complex entropy (note that dzdμeqΓ is not real valued).
Let us sketch the proof of this result. The proof involves a combination of pre-existing theory and novel tricks. The pre-existing theory consists primarily in the method developed by Guionnet and Borot to analyse the partition function of β-ensembles on the real line. However the fact that the integrand is complex-valued introduces new obstacles which must be overcome.
A differential identity
Suppose we have a family of potentials {Vt}t∈[0,1] smoothly parametrised by t, such that the potential we are interested in V=V1. Then a short calculation shows that
If we take V0 to be quadratic then lnZN,Γ[V0] can be computed by a Selberg integral and its asymptotics is then known via the asymptotics of the Barnes G-function. Thus our problem reduces to ab asymptotic expansion of ⟨LN(z)(∂t∂Vt)⟩N,Vt, ensuring that our error terms are sufficiently uniform in t∈[0,1] that we can integrate them. LN(z)(f)=N1∑j=1Nf(zj) is a linear statistic, hence our problem reduces to the asymptotics of moments (in this case the first moment) of a linear statistic.
Dyson-Schwinger equations
The method we use to study the asymptotics of moments of linear statistics is the method of Dyson-Schwinger equations. Define the fluctuation measure as
FluctN=N(LN(z)−μΓeq).
and let f0,…,fk be a collection of analytic test functions, growing sufficiently slowly on Γ such that the associated integrals converge. Then the DS equations read as follows.
These equations can be derived by integration by parts. To make these equations useful we must do two things. Firstly, we must invert the master operator Ξ. When we are in the one-cut regular regime this is possible (up to constants, but a constant gives zero under the fluctuation measure). Hence we make this hypothesis. Replacing f0→Ξ−1(f0) throughout, our DS equations relate moments of FluctN at different orders. As it stands, these equations are not closed, however they are asymptotically closed. By this I mean that FluctN(f)=O(1); more precisely, there are constants Ck such that
⟨∣FluctN(f)∣k⟩N,V≤Ck.
This motivates the original factor of N in the definition of FluctN. Substituting this bound into the Dyson-Schwinger equations in we can recursively derive a N1 expansion. For example
Remark: Note that in the case of β=2, it becomes a N21 expansion.
Remark: Note that if we discard the subleading terms in the DS equations, we essentially have the formula from “Wick’s theorem” which recursively relates Gaussian moments. This means that the linear statistics obey a kind of complex CLT, such that their cumulants of degree 3 and higher all tend to zero.
I am simplifying somewhat since in the paper it is useful to restrict to a compact subset of the contour, but doing so introduces boundary terms which must be dealt with. Moving forward, we immediately can obtain the N1 expansion of the partition function. However I have not explained how to obtain the bound ⟨∣FluctN(f)∣k⟩N,V≤Ck and indeed this is the chief difficulty of the paper. The challenge to bounding ⟨⋅⟩N,V is that it involves dividing by the complex partition function, and this is a complex (and so oscillatory) integral, so in principle it could be very small.
Let γ:R⟶C be a parametrisation of the contour Γeq. Define the “real model” as
Thus this reduces to the usual β-ensemble but with some smooth 2-body and 1-body interactions, which has been studied by Borot, Guionnet and Kozlowski, 2015. This is done by the method of Dyson-Schwinger equations, where now the “master operator” is more complicated and so more involved to invert (we remark that this is not an “integrable” model, which shows that DS equations are quite robust to the form of the interaction). The key conclusion is that one can compute asymptotically the moments of linear statistics like
k=1∑Nf(xk)
and show that they are bounded in N. These techniques also allow one to prove a CLT for such linear statistics under the real model. This means that our problem reduces to bounding the ratio of the partition functions.
The ratio of the partition functions
In this final section we sketch how to bound ZN,γ[V]∣ZN,Γ[V]∣ from below. The idea is to separate the integrand of ZN,Γ[V] into a modulus and a phase, and then treat the integral as an expectation of the oscillatory part with respect to the modulus. Given z∈C define the phase ph(z)=∣z∣z. Notice that ph(rz)=ph(z) for any r>0 and ph(zw)=ph(z)ph(w). Hence, using that ∏i<j(xi−xj)β>0 for β>0 even,
Now, let us recall that the equilibrium measure for the real model is just the pullback of the equilibrium measure μeq under γ. Let us call this equilibrium measure νeq. Let us now centre our empirical measure, LN(x)=νeq+(LN(x)−νeq). Then
This vanishes by the S-curve condition! By itself, this isn’t quite enough. We need to do two things. Firstly, our particles range over all of R, but it is convenient to condition them to lie in some bounded open set K⊃suppνeq, which can be done up to exponentially small error. Secondly, γ can be taken to be an analytic function on suppνeq, and it natural to choose our contour to be the analytic extension of this to a neighbourhood K of suppνeq. When you do this, you find the S-curve condition also extends. Thus Q will be constant on K also. Since LN(x)−νeq has zero net mass,
where FluctN=N(LN(x)−νeq). We know from an analysis of the real model that FluctN(f) is asymptotically Gaussian for a smooth bounded function f. Thus, morally, what we have is
E[eixTAx+iyTx]
where A is a deterministic matrix and x is normally distributed vector in Rn and y is a deterministic vector in Rn (in reality we are dealing with an infinite dimensional case). There is an explicit formula (which can be found in the paper) for the above in terms of A, y, and the mean and covariance of x, and crucially this formula is non-zero. This formula permits an infinite dimensional extension to when A is an operator on a Hilbert space. In this way ZN,γ[V]∣ZN,Γ[V]∣ may be bounded from below by a constant. This illustrates what was meant at the beginning when it was said that the S-curve minimises the oscillations and hence optimises the triangle inequality. This essentially concludes the proof.
Remark: This choosing of the contour to be an analytic extension of suppμeq should be regarded as the infinite dimensional analogue of the steepest descent contour, since it stabilises the oscillations by eliminating the potentially dangerous term eiβN2∫KQd(LN(x)−νeq).
Remark: Before finishing something must be said about the interpolation. Everything said above is for a fixed potential, but in fact we need to find a smooth family of potentials for which all our hypotheses (in particular, one cut regular) are true. This is not entirely trivial, unlike for the case of real potentials where there is a nice interpolation. The solution we found is the following. Let γ:[0,1]→C be an analytic curve such that γ(0)=0 and γ(1)=1. Then let
γt(x)=γ(t)γ(tx).
Clearly γ1(x)=γ(x) and limt↓0γ(t)γ(tx)=x. Furthermore γt(0)=0 and γt(1)=1, and γt is injective since γ is. Hence γt is an interpolating family of curves from γ to a straight line, with the same endpoints. The key observation, now, is that for a quadratic potential the S-curve is a straight line. Hence if we find the potential Vt associated to the S-curve γt, we will have found our interpolating family. This construction easily generalises to curves where the endpoints are not 0 and 1. △