This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A simplified Parisi Ansatz II:
Random Energy Model universality

Simone Franchini
( )
Abstract

In a previous work [A simplified Parisi Ansatz, Franchini, S., Commun. Theor. Phys., 73, 055601 (2021)] we introduced a simple method to compute the Random Overlap Structure of Aizenmann, Simm and Stars and the full-RSB Parisi formula for the Sherrington-Kirckpatrick Model without using replica theory. The method consists in partitioning the system into smaller sub-systems that we call layers, and iterate the Bayes rule. A central ansatz in our derivation was that these layers could be approximated by Random Energy Models of the Derrida type. In this paper we analyze the properties of the interface in detail, and show the equivalence with the Random Energy Model at any temperature.

Keywords: Sherrington-Kirkpatrick model, Cavity methods, Random Energy Model, Parisi formula, REM Universality.

Sapienza Universit di Roma, 1 Piazza Aldo Moro, 00185 Roma, Italy

1 Introduction

The Sherrington-Kirkpatrick (SK) model is a well known toy model for complex systems, and plays a central role in the celebrated “Replica Symmetry Breaking” (RSB) theory of spin glasses (SG) by Parisi, Mezard, Virasoro [1] and many others [2]. In previous papers [3, 4] we introduced a generalized cavity method to study the SK model and many other physical systems without relying on the so-called “replica trick”: as is shown in [3], this method allows a natural derivation of the Random Overlap Structure (ROSt) of Aizenmann, Simms and Starr [5], and the full-RSB Parisi functional for computing the free energy per spin [3, 4]. The main steps of our analysis where to define a sequence of SK models of increasing sizes by partitioning the vertices set into subsets, that we call layers, and then show that these layers can be approximated by a simpler noise model, that we call interface, where the Hamiltonian is simply the scalar product between the spin state and some external field (see Lemma 10 of [4]). In [3] a crucial claim was that the interface can be approximated by a Random Energy Model (REM) [6], the simplest toy model for a disordered systems. Introduced by B. Derrida in the 1980s, this model has inspired important mathematical advances in the understanding of spin glasses, particularly trough its relation with the Poisson Point Processes (PPP). Of special interest is the REM universality [7]. After several precursor papers, worth to cite Ebeling and Nadler [8], Mertens [9], Borgs, Chayes and Pittel [10], etc., the REM universality has been finally recognized by Mertens, Franz and Bauke [11], in the context of combinatorial optimization, and further investigated by other authors, see [7] for a survey. As is said in [7], the basic phenomenon is in that the micro-canonical distribution of the energies of a large class of models is close to a REM in distribution for certain energy windows. By implementing a form of REM universality, in Lemma 12 of [4] is shown that the thermal fluctuations of the interface energy near the ground state converge to a REM at near zero-temperature. In this paper we study the interface model in detail, compute the thermodynamic limit, and show the equivalence with the REM at any temperature. We remark that present paper only aims to describe the interface model, ie the one body Hamiltonian of Eq. (2) below, that to best of our knowledge does not have a dedicated paper describing its properties. We do not discuss the SK model here, altough the results can be obviously applied to the SK model following the methods shown in [3, 4]

2 Summary

Let briefly introduce the basic notation. Let V={1,2,,N}V=\left\{1,2,\,...\,,N\right\} be a set of NN vertices and put a spin σiΩ\sigma_{i}\in\Omega of inner states Ω={+,}\Omega=\{+,-\} on each vertex, we denote by

σV:={σiΩ:iV}\sigma_{V}:=\left\{\sigma_{i}\in\Omega:\,i\in V\right\} (1)

the generic magnetization state. The support of σV\sigma_{V} is the product space ΩV\Omega^{V}. We denote by 𝕀(A)\mathbb{I}\left(A\right) the indicator function of the event AA, that is 𝕀(A)=1\mathbb{I}\left(A\right)=1 if AA is verified and zero otherwise. Also, given two ordered sets AA and BB we use notation ABA\otimes B for the tensor product and just ABA\,B for the Cartesian product. The scalar product is denoted by the usual \cdot symbol. The Hadamard product is denoted by the \circ symbol.

2.1 The interface model

Following ideas from Borgs, Chayes [12, 13], Coja-Oghlan [14] and others, in a recent paper we showed [4] that the scalar product of a spin state with some external field, that we call interface model, and formally describe with the Hamiltonian

H(σV|hV):=σVhV,H\left(\sigma_{V}|h_{V}\right):=\sigma_{V}\cdot h_{V}, (2)

can be approximated by the REM in the low temperature phase. The field components are real numbers hih_{i}\in\mathbb{R} indexed by iVi\in V and are assumed to have been independently extracted from some probability distribution pp. We use a braket notation for the average of the test function ff respect to the Gibbs measure ξ\xi (softmax)

f(σV)ξ:=σVΩVf(σV)exp[σVβhV+Nβζ(βhV)],\langle f\left(\sigma_{V}\right)\rangle_{\xi}:=\sum_{\sigma_{V}\in\Omega^{V}}f\left(\sigma_{V}\right)\,\exp\left[-\sigma_{V}\cdot\beta h_{V}+N\beta\zeta\left(\beta h_{V}\right)\right], (3)

where NN is the number of spins and ζ\zeta is the free energy density per spin:

βζ(βhV):=1NiVlog2cosh(βhi).-\beta\zeta\left(\beta h_{V}\right):=\frac{1}{N}\sum_{i\in V}\log 2\cosh\left(\beta h_{i}\right). (4)

We remark that the interface model is closely related to the “Number Partitioning Problem” (NPP) [11, 12, 13], and we may also refer to it as a random field model, or noise model. It correspond for example to the Random Field Ising Model (RIFM) in the limit of zero Ising interactions (or infinite field amplitude) and many other models. In general, the interface could be seen as the zero interaction limit of any lattice field theory of the kind described in [15].

2.2 Thermodynamic limit

In Section 3.1 we study the thermodynamic limit by quantile mechanics [16] and series analysis, and give explicit examples for the binary, uniform and Gaussian cases. The scaling limit of the free energy density for an infinite number of spins will converge almost surely to the following functional:

limNβζ(βhV)=01𝑑qlog2cosh[βx(q)].-\lim_{N\rightarrow\infty}\beta\zeta\left(\beta h_{V}\right)=\int_{0}^{1}dq\,\log 2\cosh\left[\beta x\left(q\right)\right]. (5)

The function x(q)x\left(q\right) is called quantile and is found by inverting the cumulant of pp,

q(x):=0x𝑑zp(z),q\left(x\right):=\int_{0}^{x}dz\ p\left(z\right), (6)

by quantile mechanics [16] the quantile satisfies the differential equation

q2x(q)=ρ[x(q)][qx(q)]2,\partial_{q}^{2}x\left(q\right)=\rho\left[\,x\left(q\right)\right]\left[\partial_{q}\,x\left(q\right)\right]^{2}, (7)

where the function ρ\rho is defined from pp according to the relation

ρ(x):=xlogp(x).\rho\left(x\right):=-\partial_{x}\log p\left(x\right). (8)

We explicitly compute the uniform and Gaussian cases. At high temperature we find that, as expected, the free energy is replica symmetric, and is therefore linear in temperature. At low temperature we find that the convergence toward the ground state energy is quadratic in the temperature, ie., the specific heat is linear like in the Dulong-Petit law. The origin of this quadratic convergence is due to vertices with small field amplitude (see Section 3), and notice that, if we restrict to linear terms, the low temperature modes can be neglected and the free energy is approximately constant in temperature, like in the REM. In Section 5 of [4] the convergence to the REM is actually shown in distribution in the near zero temperature phase.

2.3 REM universality

To this scope we introduce the eigenstates of magnetization

Ω(m):={σVΩV:M(σV)=mN},\Omega\left(m\right):=\left\{\sigma_{V}\in\Omega^{V}:\,M(\sigma_{V})=\left\lfloor mN\right\rfloor\right\}, (9)

where M(σV)M(\sigma_{V}) stands for the total magnetization of the state σV\sigma_{V}. These central objects of our analysis are studied in detail in Section 4. We also introduce the “master direction” ωV\omega_{V}, the flickering state σV\sigma_{V}^{*} and flickering function ff^{*}

ωi:=hi/|hi|,σi:=σiωi,f(σV):=f(σVωV),\omega_{i}:=h_{i}/|h_{i}|,\ \ \ \sigma_{i}^{*}:=\sigma_{i}\,\omega_{i},\ \ \ f^{*}\left(\sigma_{V}\right):=f\left(\sigma_{V}\circ\omega_{V}\right), (10)

where ωi\omega_{i} is the direction of the ground state in the vertex ii, and ff is any test function if not specified otherwise. We can now introduce a fundamental variable, that we call JJ-field: let define the following quantities

ψ:=1NiV|hi|,δ2:=1NiVhi2ψ2,Ji:=|hi|ψδ,\psi:=\frac{1}{N}\sum_{i\in V}|h_{i}|,\ \ \ \delta^{2}:=\frac{1}{N}\sum_{i\in V}h_{i}^{2}-\psi^{2},\ \ \ J_{i}:=\frac{|h_{i}|-\psi}{\delta}, (11)

where ψ\psi denotes the ground state energy density and δ\delta is the amplitude of the fluctuations of |hi||h_{i}| around ψ\psi. As explained in the Sections 5 and 6, it is possible to track the fluctuations around the ground state. This is done by introducing the the vertex set XX,

X(σV):={iV:σi=1},X\left(\sigma_{V}^{*}\right):=\left\{i\in V:\,\sigma_{i}^{*}=-1\right\}, (12)

that collects the vertices in which the spin σi\sigma_{i} is flipped with respect to the direction of the ground state ωi\omega_{i}, and the renormalized field JJ:

J(σV):=1|X(σV)|iX(σV)Ji,J(\sigma_{V}^{*}):=\frac{1}{\sqrt{|X(\sigma_{V}^{*})|}}\sum_{i\in X(\sigma_{V}^{*})}J_{i}, (13)

that is the normalized sum of the flipped local fields. The Hamiltonian of Eq. (2) can be expressed in terms of MM, JJ and XX as follows:

σVhV=ψM(σV)+J(σV)4δ2|X(σV)|.\sigma_{V}\cdot h_{V}=\psi M\left(\sigma_{V}^{*}\right)+J(\sigma_{V}^{*}){\textstyle\sqrt{4\delta^{2}|X(\sigma_{V}^{*})|}}. (14)

In Section 5 we show that in the thermodynamic limit the average of Eq. (3) is mostly sampled from eigenstates of magnetization with eigenvalue

m0(βψ):=tanh(βψ).m_{0}\left(\beta\psi\right):=\tanh\left(\beta\psi\right). (15)

For any σVΩ(m0)\sigma_{V}\in\Omega\left(m_{0}\right) we can rewrite the Hamiltonian as:

σVhV=ψm0(βψ)N+J(σV)K[m0(βψ)]\sigma_{V}\cdot h_{V}=\psi\,m_{0}\left(\beta\psi\right)N+J(\sigma_{V}^{*}){\textstyle\sqrt{K\left[m_{0}\left(\beta\psi\right)\right]}} (16)

where we introduced the auxiliary function

K(m):=2δ2(1m)N.K(m):=2\delta^{2}(1-m)\,N. (17)

In reference [4] we found that at low temperature the JJ field applied to Ω(m0)\Omega\left(m_{0}\right) actually converges in distribution to a REM. This is done by noticing that when the temperature goes to zero the state align toward the direction of the ground state almost everywhere, and only a small fraction of spins is flipped in the opposite direction. Since the flipped spins are sparse two independent spin configurations will probabily have a small number of common flipped spins, that can be ignored, making the corresponding JJ-fields independent. In Section 6 we show that it is possible to extend the results of [4] to the full temperature range by properly renormalizing the JJ field. In particular, we will introduce the Δ\Delta^{\prime} field, described in detail in the Section 6. It is shown that, when applied to the ensemble Ω(m0)\Omega(m_{0}), this field is distributed like a REM by construction (i.e., a field where the pairwise overlap matrix is zero on average). In Section 6 we show that the JJ is distributed like Δ\Delta^{\prime} up to a constant and a renormalization

J(σV)K(m0)=Δ(σV)K(m0)+const.,K(m):=δ2(1m2)N,J(\sigma_{V}){\textstyle\sqrt{K(m_{0})}}=\Delta^{\prime}(\sigma_{V}){\textstyle\sqrt{K^{\prime}(m_{0})}}+\mathrm{const.},\ \ \ K^{\prime}(m):=\delta^{2}(1-m^{2})\,N, (18)

the renormalized amplitude is found in Section 6. By the averaging properties of PPP, a parameter λ\lambda exists (dependent from β\beta, ψ\psi and δ\delta) such that the Gibbs average satisfy the REM-PPP average formula [3, 4, 5, 6],

limNf(σV)ξ=limNf(σV)λm01/λ,\lim_{N\rightarrow\infty}\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\lim_{N\rightarrow\infty}\langle f^{*}\left(\sigma_{V}\right)^{\lambda}\rangle_{m_{0}}^{1/\lambda}, (19)

that in the subcritical region λ[0,1]\lambda\in\left[0,1\right] interpolates between the geometric and the arithmetic average. The importance of the REM universality to the spin glass physics is now evident in that one could directly apply this formula to Eq. (55) of [3] (or Eq. (6.20) of [4]) and find the full-RSB Parisi functional. This complets the steps to compute the free energy of the SK model (and many other models) with methods and concepts from [3, 4, 15], see Section 6 below and Section 5 of [4] for further details. Notice that, apart from disordered systems and lattice field theory, similar properties have been recently observed also in important neural network models. Of special interest is the relation with Dense Associative Memories (DAM): for example, in [17] has been shown that also in the exponential Hopfield models one can approximate the free energy of each layer with that of a REM.

3 Thermodynamic limit

Let start by formally defining the interface [3, 4]: the Hamiltonian is that of Eq. (2), following the canonical notation we call β\beta the inverse of the temperature. The canonical partition function is defined as follows:

Z(βhV):=σVΩVexp(σVβhV),Z\left(\beta h_{V}\right):=\sum_{\sigma_{V}\in\Omega^{V}}\exp\left(-\sigma_{V}\cdot\beta h_{V}\right), (20)

the associated Gibbs measure is given by the expression

ξ(σV|βhV):=exp(σVβhV)Z(βhV)=exp[σVβhV+Nβζ(βhV)]\xi\left(\sigma_{V}|\beta h_{V}\right):=\frac{\exp\left(-\sigma_{V}\cdot\beta h_{V}\right)}{Z\left(\beta h_{V}\right)}=\exp\left[-\sigma_{V}\cdot\beta h_{V}+N\beta\zeta\left(\beta h_{V}\right)\right] (21)

where the function ζ\zeta is the free energy density per spin

βζ(βhV):=1NlogZ(βhV)=1NiVlog2cosh(βhi).-\beta\zeta\left(\beta h_{V}\right):=\frac{1}{N}\log Z\left(\beta h_{V}\right)=\frac{1}{N}\sum_{i\in V}\log 2\cosh\left(\beta h_{i}\right). (22)

Let ff be a test function of σV\sigma_{V} the Gibbs average is as follows

f(σV)ξ:=σVΩVf(σV)exp[σVβhV+Nβζ(βhV)].\langle f\left(\sigma_{V}\right)\rangle_{\xi}:=\sum_{\sigma_{V}\in\Omega^{V}}f\left(\sigma_{V}\right)\,\exp\left[-\sigma_{V}\cdot\beta h_{V}+N\beta\zeta\left(\beta h_{V}\right)\right]. (23)

3.1 Field fluctuations

Since the free energy depends only on the absolute value of the external field and not on the direction, before proceeding with the computations it is convenient to introduce the following auxiliary variables:

xi=|hi|,ωi=hi/|hi|,σi=σiωix_{i}=\left|h_{i}\right|,\ \ \ \omega_{i}=h_{i}/|h_{i}|,\ \ \ \sigma_{i}^{*}=\sigma_{i}\,\omega_{i} (24)

the flickering state σV\sigma_{V}^{*} is the Hadamard product between the initial state σV\sigma_{V} and the direction of the external field ωV\omega_{V}, that corresponds to the ground state of the system and that we call master direction. Using these variables the Hamiltonian can be rewritten as σVxV\sigma_{V}^{*}\cdot x_{V} where xVx_{V} is a vector with all positive entries, we call it rectified field. To find the scaling limit of the free energy it is convenient to reoreder VV according to the order statistics, a remapping of the index iVi\in V, usually denoted with the symbol (i)\left(i\right), such that x(i)+1x(i)x_{\left(i\right)+1}\geq x_{\left(i\right)}. Then, it is easy to see that if each xix_{i} is independently drawn according to the same probability density pp, then the scaling limit of x(i)x_{\left(i\right)} for (i)/Nq[0,1]\left(i\right)/N\rightarrow q\in\left[0,1\right] will almost surely converge to the quantile function of pp.

3.2 Uniform distribution

There are several important cases that can be treated exactly, one could consider the uniform distribution p(x)=𝕀(x[0,1])p\left(x\right)=\mathbb{I}\left(x\in\left[0,1\right]\right): here the cumulant is q(x)=xq\left(x\right)=x and the quantile is therefore x(q)=qx\left(q\right)=q, then, the free energy density converges to the following integral:

limNβζ(βxV)=01𝑑qlog2cosh(βq)=1β0β𝑑τlog2cosh(τ)-\lim_{N\rightarrow\infty}\beta\zeta\left(\beta x_{V}\right)=\int_{0}^{1}dq\,\log 2\cosh\left(\beta q\right)=\frac{1}{\beta}\int_{0}^{\beta}d\tau\,\log 2\cosh\left(\tau\right) (25)

the primitive of this integral is easily found via computer algebra,

𝑑τlog2cosh(τ)=Li2[exp(2τ)]+τ22\int d\tau\,\log 2\cosh\left(\tau\right)=\frac{\mathrm{Li_{2}\left[-\exp\left(-2\tau\right)\right]}+\tau^{2}}{2} (26)

where Li2\mathrm{Li_{2}} is the dilogarithm [18], or Spence’s function, that is often encountered in particle physics: for z[0,1]z\in\left[0,1\right] the following relations holds

Li2(z)=k1(z)kk2,Li2(1)=π212,Li2(0)=0,\mathrm{Li_{2}}\left(-z\right)=\sum_{k\geq 1}\frac{\left(-z\right)^{k}}{k^{2}},\ \ \ \mathrm{Li_{2}\left(-1\right)}=-\frac{\pi^{2}}{12},\ \ \ \mathrm{Li_{2}}\left(0\right)=0, (27)

the first derivative obeys the following formula

zLi2(z)=k1(z)k1(1)k=z1k1(z)kk=log(1z)z\partial_{z}\,\mathrm{Li_{2}}\left(-z\right)=\sum_{k\geq 1}\frac{\left(-z\right)^{k-1}\left(-1\right)}{k}=z^{-1}\sum_{k\geq 1}\frac{\left(-z\right)^{k}}{k}=\frac{\log\left(1-z\right)}{z} (28)

and notice that at 1-1 the derivative converges to the nontrivial value

zLi2(1)=log2\partial_{z}\mathrm{Li_{2}}\left(-1\right)=-\log 2 (29)

Then, the scaling limit of the free energy density converges to the expression

limNβζ(βxV)=β2+π224β+Li2[exp(2β)]2β,-\lim_{N\rightarrow\infty}\beta\zeta\left(\beta x_{V}\right)=\frac{\beta}{2}+\frac{\pi^{2}}{24\,\beta}+\frac{\mathrm{Li_{2}\left[-\exp\left(-2\beta\right)\right]}}{2\beta}, (30)

where the last term is negative in the whole temperature range. Notice that the convergence to the ground state energy is quadratic in temperature: the specific heat is linear like in the Dulong-Petit law.

3.3 Half-Gaussian distribution

We could also consider more complex shapes, like the half-normal distribution,

p(x)=2/πexp(x2/2).p\left(x\right)=\sqrt{2/\pi}\ \exp\left(-x^{2}/2\right). (31)

From quantile mechanics [16] one finds

ρ(x)=x(x2/2+log2/π)=x,\rho\left(x\right)=-\partial_{x}(-x^{2}/2+\log\sqrt{2/\pi}\,)=x, (32)

then, the quantile equation and its boundary conditions is as follows:

q2x(q)=x(q)[qx(q)]2,x(0)=0,x(1/2)=2πerf1(1/2),\partial_{q}^{2}x\left(q\right)=x\left(q\right)\left[\partial_{q}\,x\left(q\right)\right]^{2},\ \ \ x\left(0\right)=0,\ \ \ x\left(1/2\right)=\sqrt{2\pi}\ \mathrm{erf^{-1}}\left(1/2\right), (33)

solving the equation with these boundary give us

x(q)=2πerf1(q).x\left(q\right)=\sqrt{2\pi}\ \mathrm{erf^{-1}}\left(q\right). (34)

Then, for the half-normal (and normal) noise model we expect the free energy to converge toward the following limit expression

limNβζ(βxV)=01𝑑qlog2cosh[β2πerf1(q)].-\lim_{N\rightarrow\infty}\beta\zeta\left(\beta x_{V}\right)=\int_{0}^{1}dq\,\log 2\cosh\,[\,\beta\sqrt{2\pi}\ \mathrm{erf^{-1}}\left(q\right)]. (35)

We notice that the differential equation in Eq. (33) before is remarkably similar to the term in parentesis in Eq. (10) of [19], that is the nonlinear antiparabolic equation from which we obtain the Parisi functional. Further investigation on the relation between the Guerra interpolation theory and quantile mechanics would be certainly interesting, we hope to explore this in a future work.

3.4 Series analysis of the Half-Gaussian

To highligt the low temperature features it will be more instructive to rather perform a series analysis. We start from the expression

βζ(βxV)=limN1NiVlog2cosh(βxi),-\beta\zeta\left(\beta x_{V}\right)=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i\in V}\log 2\cosh\left(\beta x_{i}\right), (36)

the ground state in the thermodynamic limit (TL) is

ζ()=limNlimβ1NiVxitanh(βxi)=limN1NiVxi=2π,-\zeta\left(\infty\right)=\lim_{N\rightarrow\infty}\lim_{\beta\rightarrow\infty}\frac{1}{N}\sum_{i\in V}x_{i}\tanh\left(\beta x_{i}\right)=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i\in V}x_{i}=\sqrt{\frac{2}{\pi}}, (37)

where the numeric value is obtained by computing the average of xx with xx Gaussian variable of zero average and unitary variance:

limN1NiVxi=2π0xexp(x2/2)𝑑x=2π0exp(t)𝑑t=2π,\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i\in V}x_{i}=\sqrt{\frac{2}{\pi}}\int_{0}^{\infty}x\,\exp\left(-x^{2}/2\right)\,dx=\sqrt{\frac{2}{\pi}}\int_{0}^{\infty}\exp\left(-t\right)\,dt=\sqrt{\frac{2}{\pi}}, (38)

in the last step we applied the substitution t=x2/2t=x^{2}/2. Now, let consider the equivalent expression

limN1NiVlog2cosh(βxi)=β2π+limN1NiVlog[1+exp(2βxi)],\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i\in V}\log 2\cosh\left(\beta x_{i}\right)=\beta\sqrt{\frac{2}{\pi}}+\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i\in V}\log\left[1+\exp\left(-2\beta x_{i}\right)\right], (39)

the logarithm can be expanded in the limit of large β\beta,

log[1+exp(2kβxi)]=k1(1)k1kexp(2kβxi),\log\left[1+\exp\left(-2k\beta x_{i}\right)\right]=\sum_{k\geq 1}\frac{\left(-1\right)^{k-1}}{k}\exp\left(-2k\beta x_{i}\right), (40)

the average of the exponential term can be computed by Gaussian integration

limN1NiVexp(2kβxi)=2π0𝑑xexp(2kβxx2/2)==exp( 2k2β2)2π0𝑑xexp[(2kβ+x/2)2]==exp( 2k2β2)4π0d(2kβ+x/2)exp[(2kβ+x/2)2]==exp( 2k2β2)4π2kβ𝑑texp(t2)=exp( 2k2β2)erfc(2kβ),\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i\in V}\exp\,(-2k\beta x_{i})=\sqrt{\frac{2}{\pi}}\int_{0}^{\infty}dx\,\exp\,(-2k\beta x-x^{2}/2)=\\ =\exp\,(\,2k^{2}\beta^{2})\,\sqrt{\frac{2}{\pi}}\int_{0}^{\infty}dx\,\exp\,[-(\sqrt{2}k\beta+x/\sqrt{2}\,)^{2}]=\\ =\exp\,(\,2k^{2}\beta^{2})\,\sqrt{\frac{4}{\pi}}\int_{0}^{\infty}d(\sqrt{2}k\beta+x/\sqrt{2}\,)\,\exp\,[-(\sqrt{2}k\beta+x/\sqrt{2}\,)^{2}]=\\ =\exp\,(\,2k^{2}\beta^{2})\,\sqrt{\frac{4}{\pi}}\int_{\sqrt{2}k\beta}^{\infty}dt\,\exp\,(-t^{2})=\exp\,(\,2k^{2}\beta^{2})\ \mathrm{erfc}\,(\sqrt{2}k\beta), (41)

the last substitution is t=x/2+2kβt=x/\sqrt{2}+\sqrt{2}k\beta. We found a series representation for the free energy of the gaussian noise model

βζ(βxV)=β2π+k1(1)k1kexp(2k2β2)erfc(2kβ).-\beta\zeta\left(\beta x_{V}\right)=\beta\sqrt{\frac{2}{\pi}}+\sum_{k\geq 1}\frac{\left(-1\right)^{k-1}}{k}\exp\,(2k^{2}\beta^{2})\,\mathrm{erfc}(\sqrt{2}k\beta). (42)

For large β\beta one can use the following approximation for the Error Function

exp(2k2β2)erfc(2kβ)=12πkβ+O(β3).\exp\,(2k^{2}\beta^{2})\,\mathrm{erfc}(\sqrt{2}k\beta)=\frac{1}{\sqrt{2\pi}k\beta}+O\left(\beta^{-3}\right). (43)

Put the last expression back into the logarithm expansion before and one finds:

limN1NiVlog[1+exp(2βxi)]=K02πβ+O(β3),\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i\in V}\log\left[1+\exp\left(-2\beta x_{i}\right)\right]=\frac{K_{0}}{\sqrt{2\pi}\beta}+O\left(\beta^{-3}\right), (44)

where the constant K0K_{0} is the convergent sum

K0:=k1(1)k1k2=π212K_{0}:=\sum_{k\geq 1}\frac{\left(-1\right)^{k-1}}{k^{2}}=\frac{\pi^{2}}{12} (45)

The asymptotic expansion of ζ\zeta near zero temperature is then found to be

βζ(βxV)=β2π+1βπ348+O(β3),-\beta\zeta\left(\beta x_{V}\right)=\beta\sqrt{\frac{2}{\pi}}+\frac{1}{\beta}\sqrt{\frac{\ \pi^{3}}{48}}+O\left(\beta^{-3}\right), (46)

also in this case the convergence to the ground state is quadratic in temperature.

3.5 Convergence to the ground state

The origin of the quadratic difference in the low temperature behavior is in that for both uniform and Gaussian distributions some couplings may be close to zero for a fraction of spins. To see this, let analyze the energy contribution from the subset of spins

V(δ)={iV:xi>δ},V\left(\delta\right)=\left\{i\in V:\>x_{i}>\delta\right\}, (47)

following the steps before we find

δ𝑑xexp(2kβxx2/2)==exp(2k2β2)erfc(δ/2+2kβ)=O[exp(2δkβ)].\int_{\delta}^{\infty}dx\,\exp\,(-2k\beta x-x^{2}/2)=\\ =\exp\,(2k^{2}\beta^{2})\,\mathrm{erfc}(\delta/\sqrt{2}+\sqrt{2}k\beta)=O\left[\exp\,(-2\delta k\beta)\right]. (48)

and then the limit in Eq. (44) restricted to V(δ)V\left(\delta\right) is

limN1|V(δ)|iV(δ)log[1+exp(2βxi)]=O[exp(2δβ)],\lim_{N\rightarrow\infty}\ \frac{1}{|V\left(\delta\right)|}\sum_{i\in V\left(\delta\right)}\log\left[1+\exp\left(-2\beta x_{i}\right)\right]=O\left[\exp\,(-2\delta\beta)\right], (49)

we can see if we exclude the spins with small coupling the temperature dependence is again exponentially suppressed in β\beta.

4 Eigenstates of magnetization

We introduce a central element in our analysis: the kernel of the eigenstates of magnetization. Let define the microcanonical set

Ω(M):={σVΩV:M(σV)=M},\Omega\left(M\right):=\left\{\sigma_{V}\in\Omega^{V}:\,M\left(\sigma_{V}\right)=M\right\}, (50)

that is equivalent to the kernel of all magnetization states with a given magnetization MM, or a lattice gas with exactly EE particles. Hereafter we denote by |Ω(M)||\Omega\left(M\right)| the number of such eigenstates, that we will call cardinality of Ω(M)\Omega\left(M\right), or “complexity”, as is sometimes found in the spin glass literature. Also, define a notation for the average,

f(σV)Ω(M):=1|Ω(M)|σVΩ(M)f(σV).\langle f\left(\sigma_{V}\right)\rangle_{\Omega\left(M\right)}:=\frac{1}{|\Omega\left(M\right)|}\sum_{\sigma_{V}\in\Omega\left(M\right)}f\left(\sigma_{V}\right). (51)

The eigenstates of magnetization Ω(M)\Omega\left(M\right) can be studied in detail using Large Deviation Theory (LDT) [20], even at the “sample-path” LDT level. For example, by a simple applications of the Varadhan Integral Lemma, the Mogulskii theorem, and other standard LDT methods [20, 21, 22, 23, 24, 25], one can compute the number of the eigenstates with given magnetization mN\left\lfloor mN\right\rfloor, ie the integer part of mNmN, with m[1,1]m\in\left[-1,1\right]: to simplify the notation, hereafter

Ω(mN)=:Ω(m),f(σV)Ω(mN)=:f(σV)m.\Omega\left(\left\lfloor mN\right\rfloor\right)=:\Omega\left(m\right),\ \ \ \langle f\left(\sigma_{V}\right)\rangle_{\Omega\left(\left\lfloor mN\right\rfloor\right)}=:\langle f\left(\sigma_{V}\right)\rangle_{m}. (52)

With some algebraic effort is possible to show that the cardinality of Ω(m)\Omega\left(m\right) is proportional to exp[Nϕ(m)]\exp\left[\,N\phi\left(m\right)\right], with rate function ϕ\phi equal to

ϕ(m)=log212log(1m2)+m2log(1+m1m),\phi\left(m\right)=\log 2-\frac{1}{2}\log\left(1-m^{2}\right)+\frac{m}{2}\log\left(\frac{1+m}{1-m}\right), (53)

this result can be obtained by applying the inverse Legendre transform to the free energy of the binary noise model, studied in Section 5. In Ref. [21, 22] a detailed description of the magnetization eigenstates is achieved by adapting methods from the large deviations theory, in particular, the Varadhan lemma and the Mogulskii theorem. Further details can be found in [21, 22, 23], where a full mathematical derivation is shown for the more generel HLS model (for example, the binary noise model is recovered in the most trivial case of constant urn function).

4.1 Lattice gas

Notice that the set Ω(M)\Omega\left(M\right) is equivalent to a self-avoiding lattice gas of E=N/2M/2E=N/2-M/2 particles on a lattice of size NN. Let define the particle displacements

X(σV):={iV:σi=1}VX\left(\sigma_{V}\right):=\left\{i\in V:\,\sigma_{i}=-1\right\}\subset V (54)

in the magnetic representation would be the subsets of VV where the flickering state σV\sigma_{V}^{*} is flipped with respect to the master direction 1V1_{V} (that indicates a vector with all 11 entries). Inside Ω(M)\Omega\left(M\right) the size of XX is fixed, and related to the magnetization by

|X(σV)|=(NM)/2=:E,σVΩ(M).|X\left(\sigma_{V}\right)|=\left(N-M\right)/2=:E,\ \ \forall\sigma_{V}\in\Omega\left(M\right). (55)

We interpret EE as the number of particles in our self-avoiding gas. Then, we introduce the set of all possible displacements of E=ϵNE=\left\lfloor\epsilon N\right\rfloor particles

𝒱(E):={XV:|X|=E},\mathcal{V}\left(E\right):=\left\{X\subset V:\,|X|=E\right\}, (56)

as for the magnetic rapresentation before we use the notation

𝒱(ϵN)=:𝒱(ϵ),f(σV)𝒱(ϵN)=:f(σV)ϵ.\mathcal{V}\left(\left\lfloor\epsilon N\right\rfloor\right)=:\mathcal{V}\left(\epsilon\right),\ \ \ \langle f\left(\sigma_{V}\right)\rangle_{\mathcal{V}\left(\left\lfloor\epsilon N\right\rfloor\right)}=:\langle f\left(\sigma_{V}\right)\rangle_{\epsilon}. (57)

that is in fact the exact image of Ω(m)\Omega\left(m\right) if one takes ϵ=(1m)/2\epsilon=\left(1-m\right)/2. We represent the spin states in terms of XX as follows: let σV=σVXσX\sigma_{V}=\sigma_{V\setminus X}\cup\sigma_{X}, then for the flipped spins we have σX=1X\sigma_{X}=-1_{X}, vector with all negative entries, for the others σVX=1VX\sigma_{V\setminus X}=1_{V\setminus X} with all positive entries. The spin state σV\sigma_{V} can be reconstructed from the flipped vertices

σV=1VX(1X),\sigma_{V}=1_{V\setminus X}\cup(-1_{X}), (58)

this representation in terms of particle displacements allows to easily explore the overlap structure. Consider two configurations X,Y𝒱(ϵ)X,Y\in\mathcal{V}\left(\epsilon\right), corresponding to

σV=1VX(1X),τV=1VY(1Y).\sigma_{V}=1_{V\setminus X}\cup(-1_{X}),\ \ \tau_{V}=1_{V\setminus Y}\cup(-1_{Y}). (59)

Within 𝒱(ϵ)\mathcal{V}\left(\epsilon\right) the number of particles is fixed |X|=|Y|=ϵN\left|X\right|=\left|Y\right|=\epsilon N. Now, let XYX\cap Y be the set of points of VV at which the two particle configurations XX and YY overlap, and define the non-overlapping components

X:=X(XY),Y:=Y(XY),X^{\prime}:=X\setminus\left(X\cap Y\right),\ \ Y^{\prime}:=Y\setminus\left(X\cap Y\right), (60)

that correspond to the non overlapping points of the particle displacements. By definition, their intersection is void, ie XY= X^{\prime}\cap Y^{\prime}=\textrm{ }, moreover, the following equalities hold for of the union of XX and YY

XY=XY(XY),XY=(XY)(XY).X\cup Y=X^{\prime}\cup Y^{\prime}\cup\left(X\cap Y\right),\ \ X^{\prime}\cup Y^{\prime}=\left(X\cup Y\right)\setminus\left(X\cap Y\right). (61)

It follows that |XY||X\cup Y| total fraction of VV occupied by the particles is

|XY|=|X|+|Y||XY|,\left|X\cup Y\right|=\left|X\right|+\left|Y\right|-\left|X\cap Y\right|, (62)

while the total non-overlapping volume is

|XY|=|XY||XY|=|X|+|Y|2|XY|.\left|X^{\prime}\cup Y^{\prime}\right|=\left|X\cup Y\right|-\left|X\cap Y\right|=\left|X\right|+\left|Y\right|-2\left|X\cap Y\right|. (63)

The overlap between the corresponding spin states can be expressed as

σVτV=[1VX(1X)][1VY(1Y)]==N2|X|2|Y|=N2|X|2|Y|+4|XY|,\sigma_{V}\cdot\tau_{V}=\left[1_{V\setminus X}\cup(-1_{X})\right]\cdot\left[1_{V\setminus Y}\cup(-1_{Y})\right]=\\ =N-2|X^{\prime}|-2|Y^{\prime}|=N-2\left|X\right|-2\left|Y\right|+4\left|X\cap Y\right|, (64)

since we are considering magnetizations eigenstates with fixed eigenvalue, the volumes of |X|\left|X\right| and |Y|\left|Y\right| are also fixed at ϵN\epsilon N, and the overlap of the spin states can be expressed in terms of the overlap between the particle configurations:

σVτV=(14ϵ)N+4|XY|.\sigma_{V}\cdot\tau_{V}=\left(1-4\epsilon\right)N+4\left|X\cap Y\right|. (65)

The overlap size is 0|XY|ϵN0\leq\left|X\cap Y\right|\leq\epsilon N, but it can be shown (e.g., see the next section) that for large NN the overlap concentrates on ϵ2N\epsilon^{2}N, with fluctuations of order N\sqrt{N} (it converges to a gaussian), then from m=12ϵm=1-2\epsilon and the Eq. (64) follows that the spin overlap concentrates almost surely on the mean value 14ϵ+4ϵ2=m21-4\epsilon+4\epsilon^{2}=m^{2}.

4.2 Entropy of the overlap

We compute the probability that two configuration randomly extracted from 𝒱(ϵ)\mathcal{V}\left(\epsilon\right) have an intersection of size xN\left\lfloor xN\right\rfloor, with x[0,ϵ]x\in\left[0,\epsilon\right]. The limit entropy density (rate function) of such event is defined as follows

η(x|ϵ):=limN1NlogP(|XY|=xN|X,Y𝒱(ϵ))\eta\left(x|\epsilon\right):=-\lim_{N\rightarrow\infty}\frac{1}{N}\log P\left(\,\left|X\cap Y\right|=\left\lfloor xN\right\rfloor\,|\,X,Y\in\mathcal{V}\left(\epsilon\right)\right) (66)

that gives the shape of the distribution for large number of spins

P(|XY|=xN|X,Y𝒱(ϵ))exp[Nη(x|ϵ)].P\left(\,\left|X\cap Y\right|=\left\lfloor xN\right\rfloor\,|\,X,Y\in\mathcal{V}\left(\epsilon\right)\right)\sim\exp\left[-N\eta\left(x|\epsilon\right)\right]. (67)

The first step is to notice that due to the uniformity of the distribution of the flipped spins the intersection size does not depend on the special realization of both states, then we can fix one of the two states: let’s fix YY and call it ‘target’ set, then

P(|XY|=xN|X,Y𝒱(ϵ))==P(|XY|=xN|X𝒱(ϵ)),Y𝒱(ϵ).P\left(\,\left|X\cap Y\right|=\left\lfloor xN\right\rfloor\,|\,X,Y\in\mathcal{V}\left(\epsilon\right)\right)=\\ =P\left(\,\left|X\cap Y\right|=\left\lfloor xN\right\rfloor\,|\,X\in\mathcal{V}\left(\epsilon\right)\right),\ \forall Y\in\mathcal{V}\left(\epsilon\right). (68)

Since only the size of the target set actually matters, to highligth its internal components it will be convenient to chose a special configuration of the target (see Figure 1)

Y0:={1kϵN}Y_{0}:=\left\{1\leq k\leq\left\lfloor\epsilon N\right\rfloor\right\} (69)

where the vertices of the flipped spins are placed at the beginning of the set VV (ie, the labels kVY0k\in V\setminus Y_{0} are all larger than ϵN\left\lfloor\epsilon N\right\rfloor), formally holds that

P(|XY|=xN|X,Y𝒱(ϵ))=P(|XY0|=xN|X𝒱(ϵ)),P\left(\,\left|X\cap Y\right|=\left\lfloor xN\right\rfloor\,|\,X,Y\in\mathcal{V}\left(\epsilon\right)\right)=P\left(\,\left|X\cap Y_{0}\right|=\left\lfloor xN\right\rfloor\,|\,X\in\mathcal{V}\left(\epsilon\right)\right), (70)

The entropy density is given by the limit

η(x|ϵ):=limN1NlogP(|XY0|=xN|X𝒱(ϵ)),\eta\left(x|\epsilon\right):=-\lim_{N\rightarrow\infty}\frac{1}{N}\log P\left(\,\left|X\cap Y_{0}\right|=\left\lfloor xN\right\rfloor\,|\,X\in\mathcal{V}\left(\epsilon\right)\right), (71)

and it can be computed in many ways by Varadhan Lemma, the Mogulskii theorem and other large deviations techniques.

Refer to caption
Figure 1: Two configurations XX (first row) and Y0Y_{0} (second row) extracted from 𝒱(ϵ)\mathcal{V}\left(\epsilon\right), reordered in such way that both XX and XY0X\cap Y_{0} are compact sets. The last row shows the partition into the disjoint non-overlapping components XX^{\prime}, Y0Y^{\prime}_{0} and the common component XY0X\cap Y_{0} projected on VV (last row).

4.3 Urn methods

We can adapt methods from the urn process theory (see [22, 23]) to compute the shape of the overlap entropy density. The method consists in defining a nested set sequence that start from the null set and converges to XX in exactly EE steps,

Xn:=nE{in},X_{n}:=\bigcup_{n\leq E}\left\{i_{n}\right\}, (72)

that is a markov chain with transition matrix

P(in+1=k)=𝕀(kVXn)|VXn|.P\left(i_{n+1}=k\right)=\frac{\mathbb{I}\left(\,k\in V\setminus X_{n}\right)}{\left|V\setminus X_{n}\right|}. (73)

We indicate the overlap between XnX_{n} and the target set Y0Y_{0} with

Rn:=|XnY0|,R_{n}:=\left|X_{n}\cap Y_{0}\right|, (74)

the final conditions of the processes are fixed at XE=XX_{E}=X and RE=|XY0|R_{E}=\left|X\cap Y_{0}\right| respectively, reached in E=ϵNE=\left\lfloor\epsilon N\right\rfloor steps. It can be shown that the overlap between XnX_{n} and the target set follows a urn process [21, 22, 23, 24, 25] in the step variable nn

Rn+1={Rn+1πn(Rn)Rn1πn(Rn)R_{n+1}=\left\{\begin{array}[]{ccc}R_{n}+1&&\pi_{n}\left(R_{n}\right)\\ R_{n}&&1-\pi_{n}\left(R_{n}\right)\end{array}\right. (75)

the urn function πn\pi_{n} at step nn is the ratio between the number of vertices in Y0Y_{0} that have not been occupied in the preceeding nn extractions (that are ERnE-R_{n}) and the number of vertices of VV that have not been occupied (ie, NnN-n),

πn(Rn):=ERnNn.\pi_{n}\left(R_{n}\right):=\frac{E-R_{n}}{N-n}. (76)

adapting large-deviation methods [21, 22, 23, 24, 25] from generalized urn models is possible to show that the distribution of the overlap is aproximately Gaussian. It is also possible to compute the parameters by solving the difference equation

𝔼(Rn+1)=𝔼[(Rn+1)πn(Rn)]+𝔼{Rn[1πn(Rn)]}\mathbb{E}\left(R_{n+1}\right)=\mathbb{E}\left[\left(R_{n}+1\right)\pi_{n}\left(R_{n}\right)\right]+\mathbb{E}\left\{R_{n}\left[1-\pi_{n}\left(R_{n}\right)\right]\right\} (77)

where 𝔼()\mathbb{E}\left(\,\cdot\,\right) indicates the average respect to the urn process. Substituting the expression of the urn function we find

𝔼(Rn+1)𝔼(Rn)=1Nn𝔼(Rn)+ENn,\mathbb{E}\left(R_{n+1}\right)-\mathbb{E}\left(R_{n}\right)=-\frac{1}{N-n}\mathbb{E}\left(R_{n}\right)+\frac{E}{N-n}, (78)

solving with null initial condition brings to the linear average solution 𝔼(Rn)=nE/N\mathbb{E}\left(R_{n}\right)=nE/N from which follows that the average overlap converges to

limN|XY0|ϵN=limN𝔼(RE)N=ϵ2.\lim_{N\rightarrow\infty}\frac{\langle\left|X\cap Y_{0}\right|\rangle_{\epsilon}}{N}=\lim_{N\rightarrow\infty}\frac{\mathbb{E}\left(R_{E}\right)}{N}=\epsilon^{2}. (79)

We can show that the fluctuations are small: consider

𝔼(Rn+12)=𝔼[(Rn+1)2πn(Rn)]+𝔼{Rn2[1πn(Rn)]},\mathbb{E}\left(R_{n+1}^{2}\right)=\mathbb{E}[\left(R_{n}+1\right)^{2}\pi_{n}\left(R_{n}\right)]+\mathbb{E}\left\{R_{n}^{2}\left[1-\pi_{n}\left(R_{n}\right)\right]\right\}, (80)

substituting the urn function and the formula for the linear we find

𝔼(Rn+12)𝔼(Rn2)=2Nn𝔼(Rn2)+ENn[1+n(2E1N)],\mathbb{E}\left(R_{n+1}^{2}\right)-\mathbb{E}\left(R_{n}^{2}\right)=-\frac{2}{N-n}\,\mathbb{E}\left(R_{n}^{2}\right)+\frac{E}{N-n}\left[1+n\left(\frac{2E-1}{N}\right)\right], (81)

solving again for null initial condition gives another linear solution

𝔼(Rn2)=nE[E(n1)n+N]N(N1)\mathbb{E}\left(R_{n}^{2}\right)=\frac{nE\left[E\left(n-1\right)-n+N\right]}{N\left(N-1\right)} (82)

Let now compute the variance of RnR_{n}: the variance is defined by

𝔼(Rn2)𝔼(Rn)2=nE[E(n1)n+N]N(N1)n2E2N2\mathbb{E}\left(R_{n}^{2}\right)-\mathbb{E}\left(R_{n}\right)^{2}=\frac{nE\left[E\left(n-1\right)-n+N\right]}{N\left(N-1\right)}-\frac{n^{2}E^{2}}{N^{2}} (83)

and after some algebra it can be shown that for the variance holds

limN|XY0|2ϵ|XY0|ϵ2N=limN𝔼(RE2)𝔼(RE)2N=ϵ2(1ϵ)2.\lim_{N\rightarrow\infty}\frac{\langle\left|X\cap Y_{0}\right|^{2}\rangle_{\epsilon}-\langle\left|X\cap Y_{0}\right|\rangle_{\epsilon}^{2}}{N}=\lim_{N\rightarrow\infty}\frac{\mathbb{E}\left(R_{E}^{2}\right)-\mathbb{E}\left(R_{E}\right)^{2}}{N}=\epsilon^{2}\left(1-\epsilon\right)^{2}. (84)

The entropy density η\eta can thus be expanded at second order in the variable

Λ(x|ϵ):=xϵ2ϵ(1ϵ),\Lambda\left(x|\epsilon\right):=\frac{x-\epsilon^{2}}{\epsilon\left(1-\epsilon\right)}, (85)

in the limit of large NN it can be shown that

η(x|ϵ)=12Λ(x|ϵ)2+O[Λ(x|ϵ)3],\eta\left(x|\epsilon\right)=\frac{1}{2}\,\Lambda\left(x|\epsilon\right)^{2}+O[\,\Lambda\left(x|\epsilon\right)^{3}], (86)

and since m=12ϵm=1-2\epsilon, according to Eq. (64), the corresponding spin overlap concentrates almost surely on the average value, ie., 14ϵ+4ϵ2=m21-4\epsilon+4\epsilon^{2}=m^{2}. Notice that the spin overlap concentrates on the same value of the correlation matrix σiσjm=m2\langle\sigma_{i}\sigma_{j}\rangle_{m}=m^{2}. This means that the kernel of the magnetization eigenstates commutes in distribution, ie., that the correlation matrix converges to the overlap matrix, see in Section 2 of [4] for further details on kernel commutation and its implications.

5 Binary noise model

Let now consider the simplest situation where the field has two states only (binary noise model), the absolute value is a delta function centered on one, that is p(x)=δ(x1)p\left(x\right)=\delta\left(x-1\right). We study the Hamiltonian σVωV\sigma_{V}\cdot\omega_{V}, scalar product between σV\sigma_{V} and the input ωV\omega_{V},

H(σV|ωV)=iVσiωi.H\left(\sigma_{V}|\omega_{V}\right)=\sum_{i\in V}\sigma_{i}\omega_{i}. (87)

Since |ωi|=1|\omega_{i}|=1 the canonical analysis here is very simple: notice that due to parity of the cosh\cosh function the partition function does not depend on the input state ωV\omega_{V},

Z(β)=σVΩVexp(βσVωV)=[2cosh(β)]N,Z\left(\beta\right)=\sum_{\sigma_{V}\in\Omega^{V}}\exp\,\left(-\beta\sigma_{V}\cdot\omega_{V}\right)=\left[2\cosh\left(\beta\right)\right]^{N}, (88)

the free energy per spin and the ground state energy are

βζ(β)=log2cosh(β),ψ=limβtanh(β)=1.-\beta\zeta\left(\beta\right)=\log 2\cosh\left(\beta\right),\ \ \ \psi=-\lim_{\beta\rightarrow\infty}\tanh\left(\beta\right)=-1. (89)

5.1 Free energy phases

In the low temperature limit the free energy is

βζ(β)=β+exp(2β)+O[exp(4β)],-\beta\zeta\left(\beta\right)=\beta+\exp\left(-2\beta\right)+O\left[\exp\,(-4\beta)\right], (90)

then, the free energy per spin converges to the ground state energy ζ()\zeta\left(\infty\right) exponentially fast in β\beta. Moreover, we find at high temperature the free energy converges to the replica symmetric (RS) free energy of the spin glass theory: Taylor expansion of the logcosh\log\cosh function for small β\beta gives

βζ(β)=log2+β22+O(β4).-\beta\zeta\left(\beta\right)=\log 2+\frac{\beta^{2}}{2}+O\left(\beta^{4}\right). (91)

It can be shown that in the zero temperature limit the Gibbs measure can be approximated by a random energy model: this will be discussed later.

5.2 Flickering states and thermal average

Let study the formula for the average:

f(σV)ξ=1Z(β)σVΩVf(σV)exp(βσVωV).\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\frac{1}{Z\left(\beta\right)}\sum_{\sigma_{V}\in\Omega^{V}}f\left(\sigma_{V}\right)\,\exp\left(-\beta\sigma_{V}\cdot\omega_{V}\right). (92)

Given the independence of the partition function from ωV\omega_{V} it will be convenient to introduce some notation. Define the flickering state σV:=σVωV\sigma_{V}^{*}:=\sigma_{V}\circ\omega_{V} such that the resulting vector has the following components σi:=σiωiΩ\sigma_{i}^{*}:=\sigma_{i}\,\omega_{i}\in\Omega. Notice that since ωi2=1\omega_{i}^{2}=1 a further multiplication of σV\sigma_{V}^{*} by ωV\omega_{V} gives back the original vector σV\sigma_{V}, ie., σVωV=σV\sigma_{V}^{*}\circ\omega_{V}=\sigma_{V}. Then we introduce the flickering function

f(σV):=f(σVωV),f(σV)=f(σVωV)=f(σV).f^{*}\left(\sigma_{V}\right):=f\left(\sigma_{V}\circ\omega_{V}\right),\ \ \ f^{*}\left(\sigma_{V}^{*}\right)=f\left(\sigma_{V}^{*}\circ\omega_{V}\right)=f\left(\sigma_{V}\right). (93)

Finally, we consider the scalar product (overlap) of σV\sigma_{V} with the input state ωV\omega_{V}, that is equivalent to the total magnetization of σV\sigma_{V}^{*},

σVωV=(σVωV)1V=σV1V=:M(σV).\sigma_{V}\cdot\omega_{V}=\left(\sigma_{V}\circ\omega_{V}\right)\cdot 1_{V}=\sigma_{V}^{*}\cdot 1_{V}=:M\left(\sigma_{V}^{*}\right). (94)

Putting together, the sum of ff weighted with the Gibbs weights satisfies the following chain of equivalences

σVΩVf(σV)exp(βσVωV)==σVΩVf(σVωV)exp[βM(σV)]=σVΩVf(σV)exp[βM(σV)]==σVΩVf(σV)exp[βM(σV)],\sum_{\sigma_{V}\in\Omega^{V}}f\left(\sigma_{V}\right)\exp\left(-\beta\sigma_{V}\cdot\omega_{V}\right)=\\ =\sum_{\sigma_{V}\in\Omega^{V}}f\left(\sigma_{V}^{*}\circ\omega_{V}\right)\exp\left[-\beta M\left(\sigma_{V}^{*}\right)\right]=\sum_{\sigma_{V}\in\Omega^{V}}f^{*}\left(\sigma_{V}^{*}\right)\exp\left[-\beta M\left(\sigma_{V}^{*}\right)\right]=\\ =\sum_{\sigma_{V}\in\Omega^{V}}f^{*}\left(\sigma_{V}\right)\exp\left[-\beta M\left(\sigma_{V}\right)\right], (95)

where in the last step we used that σV\sigma_{V} and σV\sigma_{V}^{*} are in a bijective relation, this implies that we can change the sum index to σV\sigma_{V} as the dependence on the input state affects only ff^{*}. Then, the formula for the average is as follows:

f(σV)ξ=1Z(β)σVΩVexp[βM(σV)]f(σV)==1Z(β)Mexp(βM)σVΩ(M)f(σV)==1Z(β)M|Ω(M)|exp(βM)f(σV)Ω(M).\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\frac{1}{Z\left(\beta\right)}\sum_{\sigma_{V}\in\Omega^{V}}\exp\left[-\beta M\left(\sigma_{V}^{*}\right)\right]f^{*}\left(\sigma_{V}\right)=\\ =\frac{1}{Z\left(\beta\right)}\sum_{M}\exp\left(-\beta M\right)\sum_{\sigma_{V}\in\Omega\left(M\right)}f^{*}\left(\sigma_{V}\right)=\\ =\frac{1}{Z\left(\beta\right)}\sum_{M}|\,\Omega\left(M\right)|\exp\left(-\beta M\right)\langle f^{*}\left(\sigma_{V}\right)\rangle_{\Omega\left(M\right)}. (96)

5.3 Average in thermodynamic limit

Assuming that ff exists in the thermodynamic limit NN\rightarrow\infty, we can write also a continuous representation. From [20, 21, 22, 23, 24, 25] it can be shown that

limNf(σV)ξ=limN11𝑑mexp{N[ϕ(m)+βm]}f(σV)m11𝑑mexp{N[ϕ(m)+βm]}\lim_{N\rightarrow\infty}\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\lim_{N\rightarrow\infty}\frac{\int_{-1}^{1}dm\,\exp\left\{-N\left[\phi\left(m\right)+\beta m\right]\right\}\langle f^{*}\left(\sigma_{V}\right)\rangle_{m}}{\int_{-1}^{1}dm\,\exp\left\{-N\left[\phi\left(m\right)+\beta m\right]\right\}} (97)

and it can be also shown that the probability mass concentrates on the value m0(β)m_{0}\left(\beta\right) that maximize βm+ϕ(m)\beta m+\phi\left(m\right). Putting together

limNf(σV)ξ=limNf(σV)m0(β)\lim_{N\rightarrow\infty}\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\lim_{N\rightarrow\infty}\langle f^{*}\left(\sigma_{V}\right)\rangle_{m_{0}\left(\beta\right)} (98)

and after some manipulations one can prove that m0(β)=tanh(β)m_{0}\left(\beta\right)=\tanh\left(\beta\right). Then, it is possible to compute the average in terms of the eigenstates of magnetization and their effects on the flickering function ff^{*}.

6 Relation with the Random Energy Model

It can be shown that at low temperature the Gibbs measure converges in distribution to a Random Energy Model (REM) of the Derrida type [4]. Define

ψ=1NiVxi,φi:=xiψ,\psi=\frac{1}{N}\sum_{i\in V}x_{i},\ \ \ \varphi_{i}:=x_{i}-\psi, (99)

where ψ\psi is the ground state energy and φi\varphi_{i} describes the field fluctuations. The Hamiltonian can be rewritten once again as follows

σVhV=ψM(σV)+σVφV.\sigma_{V}\cdot h_{V}=\psi\,M(\sigma_{V}^{*})+\sigma_{V}^{*}\cdot\varphi_{V}. (100)

As in previous section, we recall the special notation for the composition between the master direction and the test function, we called it flickering function

f(σV)=f(σVωV),f^{*}\left(\sigma_{V}\right)=f\left(\sigma_{V}\circ\omega_{V}\right), (101)

and notice that it does not depend on the external field xVx_{V}. Then, the average is rewritten in terms of the flickering variables only

f(σV)ξ=σVΩVf(σV)exp[βψM(σV)βσVφV+Nβζ(βxV)]\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\sum_{\sigma_{V}\in\Omega^{V}}f^{*}\left(\sigma_{V}\right)\,\exp\left[-\beta\psi\,M(\sigma_{V})-\beta\sigma_{V}\cdot\varphi_{V}+N\beta\zeta\left(\beta x_{V}\right)\right] (102)

so that the dependence on ωV\omega_{V} is all inside the flickering function ff^{*} and both the free energy density and the Gibbs measure depend only on the rectified field xVx_{V}.

6.1 Field fluctuations revisited

By the lattice gas representation described before, the following holds:

σVφV2=1X(σV)φX(σV),\frac{\sigma_{V}\cdot\varphi_{V}}{2}=1_{X(\sigma_{V})}\cdot\,\varphi_{X(\sigma_{V})}, (103)

in fact, consider the chain of identities

iVσiφi=iVφiiV(1σi)φi=iVφiiX(σV)2φi,\sum_{i\in V}\sigma_{i}\varphi_{i}=\sum_{i\in V}\varphi_{i}-\sum_{i\in V}\left(1-\sigma_{i}\right)\varphi_{i}=\sum_{i\in V}\varphi_{i}-\sum_{i\in X(\sigma_{V})}2\varphi_{i}, (104)

by definition we have that the first sum is zero,

iVφi=iVxiiVψ=0.\sum_{i\in V}\varphi_{i}=\sum_{i\in V}x_{i}-\sum_{i\in V}\psi=0. (105)

Let now introduce a notation for the variance inside φV\varphi_{V}, that we denote by δ2\delta^{2}, and the variance over the vertex set γ2\gamma^{2}

δ2=1NiVφi2,γ2=1NiVxi2\delta^{2}=\frac{1}{N}\sum_{i\in V}\varphi_{i}^{2},\ \ \ \gamma^{2}=\frac{1}{N}\sum_{i\in V}x_{i}^{2} (106)

this quantity is related to ground state and variance of xi2x_{i}^{2} by the relation δ2=γ2ψ2\delta^{2}=\gamma^{2}-\psi^{2} where γ\gamma is the average variance over the vertex set.

6.2 The JJ-field

We can now introduce a fundamental variable, that we call JJ-field

JX(σV):={Ji:=φi/δ:iX(σV)},J_{X(\sigma_{V})}:=\left\{J_{i}:=\varphi_{i}/\delta:\,i\in X(\sigma_{V})\right\}, (107)

from which we define the normalized field amplitude

J(σV):=1X(σV)JX(σV)|X(σV)|.J(\sigma_{V}):=\frac{1_{X(\sigma_{V})}\cdot\,J_{X(\sigma_{V})}}{\sqrt{|X(\sigma_{V})|}}. (108)

This variable converges to a Gaussian with zero mean and unitary variance in the thermodynamic limit, moreover, given two states σV\sigma{}_{V} and τV\tau{}_{V} independently extracted from Ω(m)\Omega\left(m\right) the average overlap converges to (1m)/2(1-m)/2. From previous considerations the Hamiltonian can be rewritten as follows

σVhV=ψM(σV)+J(σV)K[M(σV)],\sigma_{V}\cdot h_{V}=\psi\,M(\sigma_{V}^{*})+\,J(\sigma_{V}^{*})\sqrt{K\left[\,M(\sigma_{V})\right]}, (109)

where we introduced a notation for the normalization of the JJ-amplitude

K[M(σV)]:=2δ2N[ 1M(σV)/N].K\left[\,M(\sigma_{V})\right]:=2\delta^{2}N\left[\,1-M(\sigma_{V})/N\,\right]. (110)

The formula for the average is rewritten in terms of the new variables

f(σV)ξ==M|Ω(M)|exp(βψM)f(σV)exp[βJ(σV)K(M)]Ω(M)M|Ω(M)|exp(βψM)exp[βJ(σV)K(M)]Ω(M).\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\\ =\frac{\sum_{\,M}|\Omega\left(M\right)|\exp\left(-\beta\psi\,M\right)\langle f^{*}\left(\sigma_{V}\right)\,\exp\,[-\beta J(\sigma_{V})\sqrt{K(M)}]\rangle_{\Omega\left(M\right)}}{\sum_{\,M}|\Omega\left(M\right)|\exp\left(-\beta\psi\,M\right)\langle\,\exp\,[-\beta J(\sigma_{V})\sqrt{K(M)}]\rangle_{\Omega\left(M\right)}}. (111)

Now, let take the themodynamic limit: it can be shown by simple saddle point methods [22, 23] that the average admit the following integral representation

limNf(σV)ξ==limN11𝑑mexp{N[ϕ(m)+βψm]}f(σV)exp[βJ(σV)K(m)]m11𝑑mexp{N[ϕ(m)+βψm]}exp[βJ(σV)K(m)]m\lim_{N\rightarrow\infty}\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\\ =\lim_{N\rightarrow\infty}\frac{\int_{-1}^{1}dm\,\exp\left\{-N[\phi\left(m\right)+\beta\psi m]\right\}\,\langle f^{*}\left(\sigma_{V}\right)\,\exp\,[-\beta J(\sigma_{V})\sqrt{K\left(m\right)}]\rangle_{m}}{\int_{-1}^{1}dm\,\exp\left\{-N[\phi\left(m\right)+\beta\psi m]\right\}\,\langle\exp\,[-\beta J(\sigma_{V})\sqrt{K\left(m\right)}]\rangle_{m}} (112)

introducing the auxiliary functions

m0(βψ):=tanh(βψ),K(m):=2δ2(1m)N,m_{0}\left(\beta\psi\right):=\tanh\left(\beta\psi\right),\ \ \ K\left(m\right):=2\delta^{2}\left(1-m\right)N, (113)

we arrive to the final form for our average formula, that is

limNf(σV)ξ=limNf(σV)exp[βJ(σV)K(m0)]m0exp[βJ(σV)K(m0)]m0==limNσVΩ(m0)f(σV)exp[βJ(σV)K(m0)]σVΩ(m0)exp[βJ(σV)K(m0)].\lim_{N\rightarrow\infty}\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\lim_{N\rightarrow\infty}\frac{\langle f^{*}\left(\sigma_{V}\right)\,\exp\,[-\beta J(\sigma_{V})\sqrt{K\left(m_{0}\right)}]\rangle_{m_{0}}}{\langle\,\exp\,[-\beta J(\sigma_{V})\sqrt{K\left(m_{0}\right)}]\rangle_{m_{0}}}=\\ =\lim_{N\rightarrow\infty}\frac{\sum_{\,\sigma_{V}\in\Omega\left(m_{0}\right)}f^{*}\left(\sigma_{V}\right)\,\exp\,[-\beta J(\sigma_{V})\sqrt{K\left(m_{0}\right)}]}{\sum_{\,\sigma_{V}\in\Omega\left(m_{0}\right)}\,\exp\,[-\beta J(\sigma_{V})\sqrt{K\left(m_{0}\right)}]}. (114)

In [4] is shown that in the low temperature limit the Gaussian amplitude JJ converges in the bulk to a random energy model of the Derrida type [6] (ie., with Gaussian energies). This is done by noticing that when the temperature goes to zero the state aligns toward the direction of the ground state almost everywhere, and only a small fraction of spins get flipped in the opposite direction. Since the flipped spins are sparse, any two independent configurations will most probabily have a negligible number of common flips. The number of this common flips (see Figure 1) converges to zero faster than the size of the whole flipped set when the temperature is lowered to near zero (ie., net of quadratic terms) and can be therefore ignored in that limit: see Section 5 of [4] for further details. The crucial fact is in that the field JiJ_{i} is sampled independently for each vertex ii, then for any two disjoint subsets of VV the corresponding JJ-fields are independent like in a REM. The argument works also for multiple replicas if temperature is low enough.

6.3 REM at all temperatures

In this last sub-section we show how is possible to correct the fromulas of [4] in order to make it valid also at higher temperatures. Let consider two subsets of X,YVX,Y\subset V of same size E and their non-overlapping components XX^{\prime} and YY^{\prime} as defined in Eq. (60) of Section 4. Now notice that the following holds:

X=X{XY},Y=Y{XY}.X=X^{\prime}\cup\{X\cap Y\},\ \ \ Y=Y^{\prime}\cup\{X\cap Y\}. (115)

The the REM contribution comes only from the non-overlapping components, then we would like to get rid of the overlapping component (ie, the energy of the spins placed on the vertices in XYX\cap Y ) and write everything in terms of the sets XX^{\prime} and YY^{\prime} . This is made possible by considering the difference between the corresponding JJ-fields

1XJX1YJY=(1XJX+1XYJXY)(1YJY+1XYJXY)==1XJX1YJY1_{X}\cdot J_{X}-1_{Y}\cdot J_{Y}=\left(1_{X^{\prime}}\cdot J_{X^{\prime}}+1_{X\cap Y}\cdot J_{X\cap Y}\right)-\left(1_{Y^{\prime}}\cdot J_{Y^{\prime}}+1_{X\cap Y}\cdot J_{X\cap Y}\right)=\\ =1_{X^{\prime}}\cdot J_{X^{\prime}}-1_{Y^{\prime}}\cdot J_{Y^{\prime}} (116)

Therefore, let consider two independent replicas σV\sigma_{V} and τV\tau_{V} and let indicate with X(σV)X(\sigma_{V}) and Y(τV)Y(\tau_{V}) the associated flipped components. Let introduce the auxiliary Δ\Delta-field, that is the difference between the JJ-fields of the two replicas

Δ(σV|τV):=J(σV)J(τV),\Delta\left(\sigma_{V}|\tau_{V}\right):=J\left(\sigma_{V}\right)-J\left(\tau_{V}\right), (117)

by multiplying both numerator and denominator of the average formula in Eq. (114) by the proper τV\tau_{V}-dependent amplitude: we find

σVΩ(m0)f(σV)exp[βJ(σV)K(m0)]σVΩ(m0)exp[βJ(σV)K(m0)]==σVΩ(m0)f(σV)exp{β[J(σV)J(τV)]K(m0)}σVΩ(m0)exp{β[J(σV)J(τV)]K(m0)}==σVΩ(m0)f(σV)exp[βΔ(σV|τV)K(m0)]σVΩ(m0)exp[βΔ(σV|τV)K(m0)]\frac{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}f^{*}\left(\sigma_{V}\right)\exp\,[-\beta J\left(\sigma_{V}\right)\sqrt{K\left(m_{0}\right)}]}{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}\exp\,[-\beta J\left(\sigma_{V}\right)\sqrt{K\left(m_{0}\right)}]}=\\ =\frac{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}f^{*}\left(\sigma_{V}\right)\exp\,\{-\beta\left[J\left(\sigma_{V}\right)-J\left(\tau_{V}\right)\right]\sqrt{K\left(m_{0}\right)}\}}{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}\exp\{-\beta[J\left(\sigma_{V}\right)-J\left(\tau_{V}\right)]\sqrt{K\left(m_{0}\right)}\}}=\\ =\frac{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}f^{*}\left(\sigma_{V}\right)\exp\,[-\beta\Delta\left(\sigma_{V}|\tau_{V}\right)\sqrt{K\left(m_{0}\right)}]}{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}\exp\,[-\beta\Delta\left(\sigma_{V}|\tau_{V}\right)\sqrt{K\left(m_{0}\right)}]} (118)

from previous considerations and Eq. (116) is easy to verify that the overlapping component cancels out and

Δ(σV|τV)=1E[1X(σV)JX(σV)1Y(τV)JY(τV)]==1E[1X(σV)JX(σV)1Y(τV)JY(τV)]==EE[J(σV)J(τV)]=EEΔ(σV|τV),\Delta\left(\sigma_{V}|\tau_{V}\right)=\frac{1}{\sqrt{E}}[1_{X\left(\sigma_{V}\right)}\cdot J_{X\left(\sigma_{V}\right)}-1_{Y\left(\tau_{V}\right)}\cdot J_{Y\left(\tau_{V}\right)}]=\\ =\frac{1}{\sqrt{E}}[1_{X^{\prime}\left(\sigma_{V}\right)}\cdot J_{X^{\prime}\left(\sigma_{V}\right)}-1_{Y^{\prime}\left(\tau_{V}\right)}\cdot J_{Y^{\prime}\left(\tau_{V}\right)}]=\\ =\sqrt{\frac{E^{\prime}}{E}}[J^{\prime}\left(\sigma_{V}\right)-J^{\prime}\left(\tau_{V}\right)]=\sqrt{\frac{E^{\prime}}{E}}\,\Delta^{\prime}\left(\sigma_{V}|\tau_{V}\right), (119)

Now, since E/EE^{\prime}/E converges to 1ϵ01-\epsilon_{0} in the thermodynamic limit we have

Δ(σV)=1ϵ0Δ(σV)\Delta\left(\sigma_{V}\right)=\sqrt{1-\epsilon_{0}}\ \Delta^{\prime}\left(\sigma_{V}\right) (120)

Most important: notice that the Δ\Delta^{\prime}-amplitude is distributed like a REM by construction since we obtained it by removing the “non-REM” component from Δ\Delta. Then, let define one last auxiliary function

K(m0):=(1ϵ0)K(m0)=[1(1m0)/2][2δ2(1m0)N]==δ2(1+m0)(1m0)N=δ2(1m02)NK^{\prime}\left(m_{0}\right):=\left(1-\epsilon_{0}\right)K\left(m_{0}\right)=\left[1-\left(1-m_{0}\right)/2\,\right][2\delta^{2}\left(1-m_{0}\right)N]=\\ =\delta^{2}\left(1+m_{0}\right)\left(1-m_{0}\right)N=\delta^{2}\left(1-m_{0}^{2}\right)N (121)

and put everything together, the average formula can be transformed into

limNf(σV)ξ=σVΩ(m0)f(σV)exp[βΔ(σV)K(m0)]σVΩ(m0)exp[βΔ(σV)K(m0)]=limNf(σV)exp[βΔ(σV)K(m0)]m0exp[βΔ(σV)K(m0)m0.\lim_{N\rightarrow\infty}\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\frac{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}f^{*}\left(\sigma_{V}\right)\exp\,[-\beta\Delta^{\prime}\left(\sigma_{V}\right)\sqrt{K^{\prime}\left(m_{0}\right)}]}{\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}\exp\,[-\beta\Delta^{\prime}\left(\sigma_{V}\right)\sqrt{K^{\prime}\left(m_{0}\right)}]}\\ =\lim_{N\rightarrow\infty}\frac{\langle f^{*}\left(\sigma_{V}\right)\,\exp\,[-\beta\Delta^{\prime}(\sigma_{V})\sqrt{K^{\prime}\left(m_{0}\right)}]\rangle_{m_{0}}}{\langle\exp\,[-\beta\Delta^{\prime}(\sigma_{V})\sqrt{K^{\prime}\left(m_{0}\right)}\rangle_{m_{0}}}. (122)

We can immediately verify that after this change of variable the average is done with respect to a REM of some type at any temperature.

6.4 REM-PPP average

We can integrate the REM variable Δ\Delta^{\prime} by applying the well known REM-PPP average formula [3, 4, 5, 6]. The final result is the relation in Eq. (19)

limNf(σV)ξ=limNf(σV)λm01/λ\lim_{N\rightarrow\infty}\langle f\left(\sigma_{V}\right)\rangle_{\xi}=\lim_{N\rightarrow\infty}\langle f^{*}\left(\sigma_{V}\right)^{\lambda}\rangle_{m_{0}}^{1/\lambda} (123)

with λ\lambda depending on β\beta, ψ\psi and δ\delta. Notice that the REM-PPP average formula interpolates between arithmetic and geometric average, in fact,

limλ1f(σV)λm01/λ=f(σV)m0=1|Ω(m0)|σVΩ(m0)f(σV),\lim_{\lambda\rightarrow 1}\langle f\left(\sigma_{V}\right)^{\lambda}\rangle_{m_{0}}^{1/\lambda}=\langle f\left(\sigma_{V}\right)\rangle_{m_{0}}=\frac{1}{|\Omega\left(m_{0}\right)|}\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}f\left(\sigma_{V}\right), (124)

and with little more work it is possible to show that

limλ0f(σV)λm01/λ=limλ0[1|Ω(m0)|σVΩ(m0)f(σV)λ]1/λ==limλ0[1|Ω(m0)|σVΩ(m0)expλlogf(σV)]1/λ==limλ0[1+λ|Ω(m0)|σVΩ(m0)logf(σV)]1/λ==limλ0[1+λlogσVΩ(m0)f(σV)1|Ω(m0)|]1/λ==limλ0[expλlogσVΩ(m0)f(σV)1|Ω(m0)|]1/λ==σVΩ(m0)f(σV)1|Ω(m0)|,\lim_{\lambda\rightarrow 0}\langle f\left(\sigma_{V}\right)^{\lambda}\rangle_{m_{0}}^{1/\lambda}=\lim_{\lambda\rightarrow 0}\left[\frac{1}{|\Omega\left(m_{0}\right)|}\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}f\left(\sigma_{V}\right)^{\lambda}\right]^{1/\lambda}=\\ =\lim_{\lambda\rightarrow 0}\left[\frac{1}{|\Omega\left(m_{0}\right)|}\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}\exp\,\lambda\log f\left(\sigma_{V}\right)\right]^{1/\lambda}=\\ =\lim_{\lambda\rightarrow 0}\left[1+\frac{\lambda}{|\Omega\left(m_{0}\right)|}\sum_{\sigma_{V}\in\Omega\left(m_{0}\right)}\log f\left(\sigma_{V}\right)\right]^{1/\lambda}=\\ =\lim_{\lambda\rightarrow 0}\left[1+\lambda\log\prod_{\sigma_{V}\in\Omega\left(m_{0}\right)}f\left(\sigma_{V}\right)^{\frac{1}{|\Omega\left(m_{0}\right)|}}\right]^{1/\lambda}=\\ =\lim_{\lambda\rightarrow 0}\left[\exp\,\lambda\log\prod_{\sigma_{V}\in\Omega\left(m_{0}\right)}f\left(\sigma_{V}\right)^{\frac{1}{|\Omega\left(m_{0}\right)|}}\right]^{1/\lambda}=\\ =\prod_{\sigma_{V}\in\Omega\left(m_{0}\right)}f\left(\sigma_{V}\right)^{\frac{1}{|\Omega\left(m_{0}\right)|}}, (125)

that is the geometric average. These formulas allows the computation of the average with respect to the thermal fluctuations, although notice the dependence of ff^{*} from the ground state still remains. See Lemma 13, Section 5 of Ref. [4] for further details on how to actually compute λ\lambda in terms of β\beta, ψ\psi and δ\delta in the Gaussian case (or in the low temperature limit). Anyway, notice that the Δ\Delta^{\prime} field is only approximately Gaussian, and its rate function [20] could be different from a quadratic form when the field fluctuations are large. The reason why at low temperatures one can actually consider the bulk (wich makes the arguments relatively elementary) is in that the contribution from spins with near zero external field is only quadratic in termperature, as shown in Section 3 for the Gaussian and uniform cases. This remarkable fact guarantees that the approximate gaussianity of Δ\Delta^{\prime} works up to the linear order in temperature and then Δ\Delta^{\prime} is properly approximated by a REM of the Derrida type (ie., with Gaussian energies) in that limit. More general formulations of REM should be considered if we are interested to extend the computation of λ\lambda shown in Section 5 of [4] to the whole temperature range, like those studied by N. K. Jana in his PhD thesis [26], that consider random energies with an arbitrary large deviation profile. This will be addressed elsewhere.

7 Acknowledgments

I would like to thank Giampiero Bardella and Riccardo Balzan (Sapienza Universit di Roma), Pan Liming (USTC) and Giorgio Parisi (Accademia Nazionale dei Lincei) for interesting discussions. I would also like to thank an anonymous referee, for noticing an error in Eq. (85), and another anonymous referee, for bringing to my attention Ref. [17]. This project has been partially funded by the European Research Council (ERC), under the European Union’s Horizon 2020 research and innovation programme (grant agreement No [694925]).

References

  • [1] Spin Glass Theory and beyond: An Introduction to the Replica Method and Its Applications, Parisi, G., Mezard, M., Virasoro, M., World Scientific, 1-476 (1986).
  • [2] Spin Glass Theory and Far Beyond: Replica Symmetry Breaking After 40 Years, Charbonneau, P., Marinari, E., Mezard, M., Parisi, G., Ricci-Tersenghi, F., Sicuro, G., Zamponi, F. (eds), World Scientific, 1-740 (2023).
  • [3] A simplified Parisi Ansatz, Franchini, S., Commun. Theor. Phys., 73, 055601 (2021).
  • [4] Replica Symmetry Breaking without replicas, Franchini, S., Annals of Physics, 450, 169220 (2023).
  • [5] Mean-field Spin Glass models from the Cavity-ROSt perspective, Aizenmann, M., Sims, R., Starr, S., AMS Contemporary Mathematics Series, 437, 1-30 (2007).
  • [6] Derrida’s generalised random energy models 1: models with finitely many hierarchies, Bovier, A., Kurkova, I., Annales de l’I.H.P. Probabilit s et statistiques, 40 (4), 439-480 (2004).
  • [7] REM Universality for Random Hamiltonians. Arous, G. B., Kuptsov, A., In: Spin Glasses: Statics and Dynamics, de Monvel, A., Bovier, A. (eds), Progress in Probability, 62, 45–84 (2009).
  • [8] On constructing folding heteropolymers, Ebeling, M., Nadler, W., PNAS, 92 (19), 8798-8802 (1995).
  • [9] Random costs in combinatorial optimization, Mertens, S., Phys. Rev. Lett. 84 (6), 1347–1350 (2000).
  • [10] Phase transition and finite-size scaling for the integer partitioning problem. Borgs, C., Chayes, J., Pittel, B., Random Struct. Algorithms, 19 (3–4), 247–288 (2001). Analysis of algorithms, Krynica Morska, (2000).
  • [11] Number partitioning as a random energy model, Bauke, H., Franz, S., Mertens, S., J. Stat. Mech. Theory Exp., 2004, P04003 (2004).
  • [12] Proof of the local REM conjecture for number partitioning. I. Constant energy scales. Borgs, C., Chayes, J., Mertens, S., Nair, C., Random Struct. Algorithms, 34 (2), 217–240 (2009).
  • [13] Proof of the local REM conjecture for number partitioning. II. Growing energy scales. Borgs, C., Chayes, J., Mertens, S., Nair, C., Random Struct. Algorithms, 34 (2), 241–284 (2009).
  • [14] Harnessing the Bethe Free Energy, Bapst, V., Coja-Oghlan, A., Random Struct. Algorithms, 49, 694-741 (2016).
  • [15] Neural activity in quarks language: Lattice Field Theory for a natwork of real neurons, Bardella, G., Franchini, S., Pan, L., Balzan, R., Ramawat, S., Brunamonti, E., Pani, P., Ferraina, S., Entropy, 26 (6), 495 (2024).
  • [16] Quantile mechanics, Steinbrecher, G., Shaw, W. T., Eur. J. Appl. Math., 19 (2), 87-112 (2008).
  • [17] The Exponential Capacity of Dense Associative Memories, Lucibello, C., Mezard M., Phys. Rev. Lett., 132, 077301 (2024).
  • [18] The Dilogarithm Function, Zagier, D., In: Frontiers in Number Theory, Physics, and Geometry. II, Cartier, P., Moussa, P., Julia, B., Vanhove, P. (eds), Springer, Berlin, Heidelberg (2007).
  • [19] Broken Replica Symmetry Bounds in the Mean Field Spin Glass Model, Guerra, F., Comm. Math. Phys., 233 (1), 1-12 (2003).
  • [20] Large Deviations Techniques and Applications, Dembo, A., Zeitouni, O., Springer Berlin, 1-399 (1998).
  • [21] Large deviations for generalized Polya urns with general urn functions,  Franchini, S., PhD thesis, Universit Roma 3 (2015). http://hdl.handle.net/2307/5212
  • [22] Large deviations for generalized Polya urns with arbitrary urn function, Franchini, S., Stoch. Proc. Appl., 127 (10), 3372-3411 (2017).
  • [23] Large-deviation theory of increasing returns, Franchini, S., Balzan, R., Phys. Rev. E, 107, 064142 (2023).
  • [24] Random polymers and generalized urn processes, Franchini, S., Balzan, R., Phys. Rev. E, 98, 042502 (2018).
  • [25] Large deviations in models of growing clusters with symmetry-breaking transitions, Jack, R. L., Phys. Rev. E, 100, 012140 (2019).
  • [26] Contributions to Random Energy Models, Jana, N. K. arXiv:0711.1249 (2007).