1 Introduction

Let \(\sigma _{N}(dx)\) denote the uniform probability measure on\(S^N:=\{x\in \mathbb {R}^N\,:\,\Vert x\Vert ^2_2=N\}\), where \(\Vert x\Vert _2\) is the Euclidean norm. For \(x:=(x_1,\ldots x_{N_1})\in \mathbb {R}^{N_{1}}\) and \(y:=(y_1,\ldots ,y_{N_2})\in \mathbb {R}^{N_{2}}\) the bipartite spin glass is defined by the energy function

$$\begin{aligned} H_{N_1,N_2}(x,y;\xi ):=-\frac{1}{\sqrt{N}}\sum _{j=1}^{N_2}\sum _{i=1}^{N_1} \xi _{ij}x_iy_j\,. \end{aligned}$$
(1)

Here \(\{\xi _{ij}\}_{i\in [N_1],j\in [N_{2}]}\) are \({\mathcal {N}}(0,1)\) i.i.d. quenched r.vs. and we set \(N:=N_1+N_2\). The object of interest of this note is the free energy

$$\begin{aligned} A_{N_1,N_2}(\beta ,\xi ):=\frac{1}{N}\log \int \sigma _{N_1}(dx)\sigma _{N_2}(dy)\exp (-\beta H_{N_1,N_2}(x,y;\xi )-b_1(x,1)-b_2(y,1))\, \end{aligned}$$
(2)

in the limit in which \(N_1,N_2\rightarrow \infty \) with \(N_1/N\rightarrow \alpha \in (0,1)\). Here \(\beta \;\geqslant \;0\) is the inverse temperature, \(b_1,b_2\in \mathbb {R}\) are external fields and \((\cdot ,\cdot )\) denotes the Euclidean inner product. By concentration of Lipschitz functions of Gaussian random variables one reduces to study the average free energy \(A_{N_1,N_2}(\beta ):=E[A_{N_1,N_2}(\beta ,\xi )]\), whose limit we denote by \(A(\alpha ,\beta )\).

Auffinger and Chen proved in [1] the following variational formula for \(A(\alpha ,\beta )\) for \(\beta \) small enough

$$\begin{aligned} A(\alpha ,\beta )= & {} \min _{q_1,q_2\in [0,1]^2} P(q_1,q_2) \end{aligned}$$
(3)
$$\begin{aligned} P(q_1,q_2)= & {} \frac{\beta ^2\alpha (1-\alpha )}{2}(1-q_1q_2)+\frac{\alpha }{2}\left( b_1^2(1-q_1)+\frac{q_1}{1-q_1}+\log (1-q_1)\right) \nonumber \\&\quad +\frac{1-\alpha }{2}\left( b_2^2(1-q_2)+\frac{q_2}{1-q_2}+\log (1-q_2)\right) \, \end{aligned}$$
(4)

(the normalisation in (1) leads to different constants w.r.t. [1]). The above formula was successively proved to hold in the whole range of \(\beta \;\geqslant \;0\) in [2, 9]. Yet these proofs are indirect, as in both cases one obtains a formula for the free energy and then verifies a posteriori (analytically for [2] and numerically [14] for [9]) that it coincides with (3). We just mention that the results in [1] have been recently extended in [10, 11] for the complexity and in [5, 6] for the free energy.

The convex variational principle found by Auffinger and Chen appears to be in contrast with the \(\min \max \) characterisation given in [4, 7] for models on the vertices of the hypercube (see also [3] for the Hopfield model). The aim of this note is to show that the Auffinger and Chen formula can be equivalently expressed in terms of a \(\min \max \).

One disadvantage of the spherical prior is that the associated moment generating function

$$\begin{aligned} \Gamma _N(h):=\frac{1}{N}\log \int \sigma _N(dx)e^{(h,x)}\,,\quad h\in \mathbb {R}^N\,, \end{aligned}$$
(5)

is not easy to compute. If h is random with i.i.d. \({\mathcal {N}}(b,q)\) components it is convenient to set

$$\begin{aligned} \Gamma (b,q):=\lim _NE \Gamma _N(h)\,. \end{aligned}$$
(6)

The so-called Crisanti–Sommers variational characterisation of it as \(N\rightarrow \infty \) reads as follows.

Lemma 1

Let \(b\in \mathbb {R}\), \(q>0\), \(h\in \mathbb {R}^N\) with i.i.d \({\mathcal {N}}(b,\sqrt{q})\) components. Then

$$\begin{aligned} \Gamma (b,q)= \frac{1}{2}\min _{r\in [0,1)}\left( (b^2+q)(1-r)+\frac{r}{1-r}+\log (1-r)\right) \, \end{aligned}$$
(7)

At the end of this note we give a simple proof of this statement, based on the method of [8, 9]. We first get a variational characterisation of the moment generating function of a Gaussian distribution (whose variance is Legendre conjugate to q) and then use concentration of measure.

A direct computation shows that the minimum of (7) is attained for

$$\begin{aligned} \frac{r}{(1-r)^2}=q+b^2\,. \end{aligned}$$
(8)

A standard replica symmetric interpolation gives that for any \(q_1,q_2\in [0,1]\)

$$\begin{aligned} A _{N_1,N_2}(\beta )\!\!\!&\,=&\,\!\!\!\frac{\beta ^2\alpha (1-\alpha )}{2}(1-q_1)(1-q_2)+(1-\alpha )\Gamma (b_2,\beta ^2\alpha q_1)\nonumber \\&\quad +\alpha \Gamma (b_1,\beta ^2(1-\alpha ) q_2)+{\text {Error}}_N(q_1,q_2)\,. \end{aligned}$$
(9)

The last summand is an error term whose specific form is not important here. What matters is that by [1, Lemma 1] there is a choice of \((q_1,q_2)\) (see below) for which this remainder goes to zero as \(N\rightarrow \infty \) if \(\beta \) is small enough. Combining (7) and (8) we can rewrite the first line of (9) as

$$\begin{aligned} {{\,\mathrm{RS}\,}}(q_1,q_2):= & {} \frac{\beta ^2\alpha (1-\alpha )}{2}(1-q_1)(1-q_2)\nonumber \\&\quad +\frac{\beta ^2\alpha (1-\alpha )}{2}\left( \left( q_2+\frac{b_1^2}{\beta ^2(1-\alpha )}\right) (1-r_1) +\left( q_1+\frac{b_2^2}{\beta ^2\alpha }\right) (1-r_2)\right) \nonumber \\&\quad +\frac{\alpha }{2}\frac{r_1}{1-r_1}+\frac{\alpha }{2}\log (1-r_1)+\frac{1-\alpha }{2}\frac{r_2}{1-r_2} +\frac{1-\alpha }{2}\log (1-r_2)\,, \end{aligned}$$
(10)

under the condition

$$\begin{aligned} \frac{r_1}{(1-r_1)^2}=\beta ^2(1-\alpha ) q_2+b_1^2\,,\quad \frac{r_2}{(1-r_2)^2}=\beta ^2\alpha q_1+b_2^2\,. \end{aligned}$$
(11)

Here we used that there is a sequence \(o_N\rightarrow 0\) uniformly in \(q_1,q_2,\beta ,\alpha \) such that

$$\begin{aligned} \frac{\beta ^2\alpha (1-\alpha )}{2}(1-q_1)(1-q_2)+(1-\alpha )\Gamma _N(\beta ^2\alpha q_1)+\alpha \Gamma _N(\beta ^2(1-\alpha ) q_2)={{\,\mathrm{RS}\,}}(q_1,q_2)+o_N\,. \end{aligned}$$
(12)

Indeed (12) follows easily once we use Lemma 1 for the limit of the functions \(\Gamma _N\) and we note that (11) are the critical point equations related to the minimisation of (7).

The main observation of this note is that (10) under (11) is optimised as a \(\min \max \).

Proposition 1

Assume \(b_1^2+b_2^2>0\). The function \({{\,\mathrm{RS}\,}}(q_1,q_2)\) has a unique stationary point \((\bar{q}_1, \bar{q}_2)\). It solves

$$\begin{aligned} \frac{q_2}{(1-q_2)^2}=\beta ^2\alpha q_1+b_1^2\,,\qquad \frac{q_1}{(1-q_1)^2}=\beta ^2(1-\alpha ) q_2+b_2^2\,. \end{aligned}$$
(13)

Moreover

$$\begin{aligned} {{\,\mathrm{RS}\,}}(\bar{q}_1,\bar{q}_2)=\min _{q_2\in [0,1]}\max _{q_1\in [0,1]} {{\,\mathrm{RS}\,}}(q_1,q_2)\,. \end{aligned}$$
(14)

If \(b_1=b_2=0\) and

$$\begin{aligned} \beta ^4\alpha (1-\alpha )<1\, \end{aligned}$$
(15)

the origin is the unique solution of (13) and

$$\begin{aligned} {{\,\mathrm{RS}\,}}(0,0)=\min _{q_2\in [0,1]}\max _{q_1\in [0,1]} {{\,\mathrm{RS}\,}}(q_1,q_2)\,. \end{aligned}$$
(16)

If \(b_1=b_2=0\) and

$$\begin{aligned} \beta ^4\alpha (1-\alpha )>1\, \end{aligned}$$
(17)

there is a unique \((\bar{q}_1, \bar{q}_2)\ne (0,0)\) which solves (13) and such that (14) holds. Moreover

$$\begin{aligned} {{\,\mathrm{RS}\,}}(0,0)=\max _{q_2\in [0,1]}\max _{q_1\in [0,1]} {{\,\mathrm{RS}\,}}(q_1,q_2)\,. \end{aligned}$$
(18)

The crucial point of [1, Lemma 1] (for us) is that from the Latala argument [13, Sect. 1.4] it follows that the overlaps self-average as \(N\rightarrow \infty \) at a point \((\tilde{q}_1, \tilde{q}_2)\) uniquely given by

$$\begin{aligned} \frac{\tilde{q}_1}{(1-\tilde{q}_1)^2}=\beta ^2(1-\alpha ) \tilde{q}_2+b_1^2\,,\quad \frac{\tilde{q}_2}{(1-\tilde{q}_2)^2}=\beta ^2\alpha \tilde{q}_1+b_2^2\,, \end{aligned}$$
(19)

which (see [12, Lemma 7]) are indeed asymptotically equivalent to

$$\begin{aligned} q_{1,N}:= & {} \frac{1}{N}E\left[ \frac{\int \sigma _{N_1}(dy)\sigma _{N_1}(dy') (y,y') e^{\beta \sqrt{q_2}(y+y',h)}}{\left( \int \sigma _{N_1}(dy) e^{\beta \sqrt{q_2}(y,h)}\right) ^2}\right] \,,\end{aligned}$$
(20)
$$\begin{aligned} q_{2,N}:= & {} \frac{1}{N} E\left[ \frac{\int \sigma _{N_2}(dx)\sigma _{N_2}(dx') (x,x') e^{\beta \sqrt{q_1}(x+x',h)}}{\left( \int \sigma _{N_2}(dx) e^{\beta \sqrt{q_1}(x,h)}\right) ^2}\right] \,, \end{aligned}$$
(21)

naturally arising from the replica symmetric interpolation (here h is random with i.i.d. \(\mathcal N(0,1)\) entries). Comparing (11) and (19) readily implies that we can plug \((r_1,r_2)=(q_1,q_2)\) into (10) and obtain the convex function \(P(q_1,q_2)\) of [1, Theorem 1], optimised by (19).

On the other hand, without using the Latala method one might still optimise (10) as a function of four variables, ignoring (11). Taking derivatives first in \(q_1,q_2\), the critical point equations (24), (2425) below select exactly \((q_1,q_2)=(r_1,r_2)\). This procedure is however unjustified a priori and this particular application of Latala’s method legitimises the exchange in the order of the optimisation of the q and the r variables for small \(\beta \), which a posteriori can be extended to all \(\beta \) [2, 9].

We stress that by itself the Latala method is not variational, it only gives the self-consistent equations for the critical points. It is the Crisanti–Sommers formula (7) which makes it implicitly variational. Such a variational representation is not necessary in other cases of interest, for instance for the bipartite SK model (namely Hamiltonian (1) with \(\pm 1\) spins), for which one simply has the \(\log \cosh \). Indeed in this case a direct use of the Latala method yields the validity of the \(\min \max \) formula of [4] for \(\beta \) and \(|b_1|,|b_2|\) small enough. The proof is essentially an exercise after [13, Proposition 1.4.8] and [1, Formula (9)] and will not be reproduced here in details. The replica symmetric sum-rule for the free energy (analogue of formula (9)) reads as

$$\begin{aligned} A _{N_1,N_2}(\beta )= & {} \frac{\beta ^2\alpha (1-\alpha )}{2}(1-q_1)(1-q_2)+(1-\alpha )E\log \cosh (b_2+\beta \sqrt{\alpha q_1}g)\nonumber \\&\quad +\,\alpha E\log \cosh (b_1+\beta \sqrt{(1-\alpha ) q_2}g)\nonumber \\&\quad +\,{\text {Error}}_N(q_1,q_2)\,, \end{aligned}$$
(22)

(here \(g\sim {\mathcal N}(0,1)\)) and the error term can be shown by the Latala method to vanish for small \(\beta ,|b_1|, |b_2|\), if \((q_1,q_2)=(\bar{q}_1, \bar{q}_2)\) are given by

$$\begin{aligned} \bar{q}_1=E[\tanh (b_1+\beta \sqrt{(1-\alpha )\bar{q}_2} g)]\,,\qquad \bar{q}_2=E[\tanh (b_2+\beta \sqrt{\alpha \bar{q}_1} g)]\,. \end{aligned}$$
(23)

Therefore the free energy equals the first two lines on the r.h.s. of (22) evaluated in \((q_1,q_2)=(\bar{q}_1, \bar{q}_2)\), which is the value attained at the \(\min \max \), as shown in [4, 7].

2 Proofs

Proof of Proposition 1

Assume first \(b_1^2+b^2_2>0\). We differentiate (10) and by (11) we get

$$\begin{aligned} \partial _{q_1} {{\,\mathrm{RS}\,}}= & {} \frac{\beta ^2\alpha (1-\alpha )}{2}(q_2-r_2(q_1))\, \end{aligned}$$
(24)
$$\begin{aligned} \partial _{q_2} {{\,\mathrm{RS}\,}}= & {} \frac{\beta ^2\alpha (1-\alpha )}{2}(q_1-r_1(q_2))\,. \end{aligned}$$
(25)

The functions \(r_1,r_2\) write explicitly as

$$\begin{aligned} r_1(q_2)= & {} \frac{\sqrt{1+4(\beta ^2(1-\alpha )q_2+b_1^2)}-1}{\sqrt{1+4(\beta ^2(1-\alpha )q_2+b_1^2)}+1} \end{aligned}$$
(26)
$$\begin{aligned} r_2(q_1)= & {} \frac{\sqrt{1+4(\beta ^2\alpha q_1+b_2^2)}-1}{\sqrt{1+4(\beta ^2\alpha q_1+b_2^2)}+1}\,. \end{aligned}$$
(27)

We easily see that \(r_1,r_2\) are increasing from \(r_1(0),r_2(0)>0\) (obviously computable by the formulas above) to 1 and concave. Moreover we record for later use that if \(b_1=b_2=0\) we have

$$\begin{aligned} \frac{d}{dq_2}r_1(q_2)\Big |_{q_2=0}=\beta ^2(1-\alpha )\,,\quad \frac{d}{dq_1}r_2(q_1)\Big |_{q_1=0}=\beta ^2\alpha \,. \end{aligned}$$
(28)

Now we take the derivative w.r.t. \(q_1\) and note that the r.h.s. of (24) is decreasing as a function of \(q_1\), thus \(\partial ^2_{q_1} {{\,\mathrm{RS}\,}}<0\). Therefore by the implicit function theorem there is a unique function \(q_1\) such that \(q_2=r_2(q_1)\). As a function of \(q_2\), \(q_1\) is non-negative, increasing and convex and it is \(q_1(r_2(0))=0\). We set

$$\begin{aligned} {{\,\mathrm{RS}\,}}_1(q_2):=\max _{q_1}{{\,\mathrm{RS}\,}}(q_1,q_2)={{\,\mathrm{RS}\,}}(q_1(q_2),q_2)\, \end{aligned}$$
(29)

and compute

$$\begin{aligned} \partial _{q_2}{{\,\mathrm{RS}\,}}_1(q_2)=\frac{\beta ^2\alpha (1-\alpha )}{2}\left( q_1(q_2)-r_1(q_2)\right) \,. \end{aligned}$$
(30)

By the properties of the functions \(q_1\) and \(r_1\) it is clear that there is a unique intersection point \(\bar{q}_2\); moreover \(q_1\;\leqslant \;r_1\) for \(q_2\;\leqslant \;\bar{q}_2\) and otherwise \(q_1\;\geqslant \;r_1\). Therefore \(\partial _{q_2}{{\,\mathrm{RS}\,}}_1(q_2)\) is increasing in a neighbourhood of \(\bar{q}_2\) which allows us to conclude \(\partial ^2_{q_2}{{\,\mathrm{RS}\,}}_1>0\). This finishes the proof if \(b_1^2+b_2^2>0\).

If \(b_1=b_2=0\) the origin is always a stationary point. It is unique if

$$\begin{aligned} \left[ \frac{d}{dq_1}r_2(q_1)\big |_{q_1=0}\right] ^{-1} =\frac{d}{dq_2}q_1(q_2)\big |_{q_2=0}>\frac{d}{dq_2}r_1(q_2)\big |_{q_2=0}\,, \end{aligned}$$
(31)

which, bearing in mind (28), amounts to ask (15).

Since \(r_2\) is increasing around the origin, we have \(\partial ^2_{q_1}{{\,\mathrm{RS}\,}}<0\) and by the implicit function theorem we define locally a function \(q_1(q_2)\) increasing and positive, vanishing at the origin. We set

$$\begin{aligned} {{\,\mathrm{RS}\,}}_1(q_2):=\max _{q_1}{{\,\mathrm{RS}\,}}(q_1,q_2)={{\,\mathrm{RS}\,}}(q_1(q_2),q_2)\, \end{aligned}$$
(32)

and compute

$$\begin{aligned} \partial _{q_2}{{\,\mathrm{RS}\,}}_1(q_2)=\frac{\beta ^2\alpha (1-\alpha )}{2}\left( q_1(q_2)-r_1(q_2)\right) \,. \end{aligned}$$
(33)

By (31) we have \(\partial ^2_{q_2}{{\,\mathrm{RS}\,}}_1\big |_{q_2=0}>0\,\), whence we obtain (16).

If (17) holds, then

$$\begin{aligned} \frac{d}{dq_2}q_1(q_2)\big |_{q_2=0}<\frac{d}{dq_2}r_1(q_2)\big |_{q_2=0}\,, \end{aligned}$$
(34)

which proceeding as before leads to (18).

However also in the case \(b_1=b_2=0\) we can repeat all the steps done in the case \(b_1^2+b_2^2>0\), showing the existence of a point \((\bar{q}_1, \bar{q}_2)\) in which a \(\min \max \) of \({{\,\mathrm{RS}\,}}\) is attained. If (31) (i.e. (15)) holds then it must be \((\bar{q}_1, \bar{q}_2)=(0,0)\). If (17) holds, then (34) enforces

$$\begin{aligned} q_1(q_2)-r_1(q_2)\;\leqslant \;0 \end{aligned}$$

in a neighbourhood of the origin (as \(q_1(0)=r_1(0)=0\)), which implies that the critical point \((\bar{q}_1, \bar{q}_2)\) must fall elsewhere. \(\square \)

Proof of Lemma 1

We will prove that for all \(u\in \sqrt{q}S^N\)

$$\begin{aligned} \Gamma ^{(\sigma )} (q):=\lim _N \Gamma _N (u)=\frac{1}{2}\min _{r\in [0,1)}\left( q(1-r)+\frac{r}{1-r}+\log (1-r)\right) \,. \end{aligned}$$
(35)

We show first that (35) implies the assertion. Let h be a random vector with i.i.d. \(\mathcal N(0,q)\) entries. (As customary we write \(X\simeq Y\) if there are constants \(c,C>0\) such that \(cY\;\leqslant \;X\;\leqslant \;CY\)). The classical estimates

$$\begin{aligned} \Gamma _N(h)\;\leqslant \;\frac{\Vert h\Vert _2}{\sqrt{N}}\,,\quad P\left( \left| \frac{\Vert h\Vert _2}{\sqrt{N}}-\sqrt{q}\right| \;\geqslant \;t\right) \simeq e^{-\frac{t^2N}{2}}\, \end{aligned}$$
(36)

permit us to write for all \(t>0\) (small)

$$\begin{aligned} |E[\Gamma _N]-\Gamma ^{(\sigma )} (q)|&\;\leqslant \;|E[\Gamma _N1_{\left\{ \left| \frac{\Vert h\Vert }{\sqrt{N}}-\sqrt{q}\right|< t\right\} }]-\Gamma ^{(\sigma )} (q)|+\left| E\left[ \frac{\Vert h\Vert _2}{\sqrt{N}}1_{\left\{ \left| \frac{\Vert h\Vert }{\sqrt{N}}-\sqrt{q}\right| \;\geqslant \;t\right\} }\right] \right| \nonumber \\&\quad \simeq \left| \Gamma _N(u^*)P\left( \left| \frac{\Vert h\Vert _2}{\sqrt{N}}-\sqrt{q}\right| < t\right) -\Gamma ^{(\sigma )} (q)\right| +o(t)+e^{-t^2N/2}\,\nonumber \\&\quad \simeq \left| \Gamma _N(u^*)-\Gamma ^{(\sigma )} (q)\right| +o(t)+e^{-t^2N/2}\,, \end{aligned}$$
(37)

for some \(u^*\in \sqrt{q}S^{N}\) and \(o(t)\rightarrow 0\) as \(t\rightarrow 0\). Since \(t>0\) is arbitrary we obtain

$$\begin{aligned} |E[\Gamma _N]-\Gamma ^{(\sigma )} (q)|\;\leqslant \;\left| \Gamma _N(u^*)-\Gamma ^{(\sigma )} (q)\right| \,. \end{aligned}$$

It remains to show (35). Given \(\varepsilon >0\) we introduce the spherical shell

$$\begin{aligned} S^{N,\varepsilon }:=S^N+\sqrt{\frac{ \varepsilon }{N}} S^N \end{aligned}$$

and the measure \(\sigma _N^{(\varepsilon )}\) as the uniform probability on it. For any \(\theta >0\) we have

$$\begin{aligned} \int \sigma ^{(\varepsilon )}_{N}(dx) e^{(u,x)}&\;\leqslant \;e^{\frac{\theta (N+\varepsilon )}{2}}\int \sigma ^{(\varepsilon )}_{N}(dx) e^{-\frac{\theta }{2}\Vert x\Vert _2^2+(u,x)}\nonumber \\&\;\leqslant \;e^{\frac{\theta (N+\varepsilon )}{2}}\frac{\sqrt{2\pi }^N}{\theta ^{\frac{N}{2}}|S^{N,\varepsilon }|}\int e^{-\frac{\theta }{2}\Vert x\Vert _2^2+(u,x)}\frac{dx}{\sqrt{2\pi }^N}\nonumber \\&\quad =e^{\frac{\theta (N+\varepsilon )}{2}+\frac{q N}{2\theta }}\frac{\sqrt{2\pi }^N}{\theta ^{\frac{N}{2}}|S^{N,\varepsilon }|}\,. \end{aligned}$$
(38)

Therefore for \(C>0\) large enough

$$\begin{aligned} \frac{1}{N}\log \int \sigma ^{(\varepsilon )}_{N}(dx) e^{(u,x)}\;\leqslant \;\frac{\theta }{2}+\frac{q}{2\theta }-\frac{1}{2}(\log \theta +1)+C\theta \frac{\varepsilon }{N}\,. \end{aligned}$$
(39)

Since this inequality holds for all \(\theta >0\) and \(\varepsilon >0\) we have

$$\begin{aligned} \limsup _N \Gamma _N (u)\;\leqslant \;\inf _{\theta >0}\left( \frac{q}{2\theta }+\frac{\theta -1}{2}-\frac{1}{2}\log \theta \right) \,. \end{aligned}$$
(40)

We set for brevity

$$\begin{aligned} \Gamma _1(\theta ):=\frac{q}{2\theta }+\frac{\theta -1}{2}-\frac{1}{2}\log \theta \, \end{aligned}$$

and notice that \(\Gamma _1\) is uniformly convex in all the intervals \((0,\theta _0)\) for finite \(\theta _0>0\).

For the reverse bound, again we let \(\theta >0\) and write

$$\begin{aligned} \int \sigma ^{(\varepsilon )}_{N}(dx) e^{(u,x)}\;\geqslant \;e^{\frac{\theta }{2} N}\int _{\mathbb {R}^N} \frac{dx}{|S^{N,\varepsilon }|} e^{-\frac{\theta }{2}\Vert x\Vert _2^2+(u,x)}-e^{\frac{\theta }{2} N}\int _{(S^{N,\varepsilon })^c} \frac{dx}{|S^{N,\varepsilon }|} e^{-\frac{\theta }{2}\Vert x\Vert _2^2+(u,x)}\,. \end{aligned}$$
(41)

The first summand on the r.h.s. can be written as before

$$\begin{aligned} e^{\frac{\theta }{2} N}\int _{\mathbb {R}^N} \frac{dx}{|S^{N,\varepsilon }|} e^{-\frac{\theta }{2}\Vert x\Vert _2^2+(u,x)}=e^{\frac{\theta N}{2}+\frac{q N}{2\theta }}\frac{\sqrt{2\pi }^N}{\theta ^{\frac{N}{2}}|S^{N,\varepsilon }|}\,. \end{aligned}$$
(42)

For the second summand we introduce \(\eta \in (0,\frac{\theta }{2})\) and bound

$$\begin{aligned} e^{\frac{\theta }{2} N}\int _{\Vert x\Vert ^2\;\leqslant \;N-\varepsilon } \frac{dx}{|S^{N,\varepsilon }|} e^{-\frac{\theta }{2}\Vert x\Vert _2^2+(u,x)}&\;\leqslant \;&e^{\frac{\theta }{2} N+(N-\varepsilon )\frac{\eta }{2}+\frac{q N}{2(\theta +\eta )}} \frac{\sqrt{2\pi }^N}{\theta ^{\frac{N}{2}}|S^{N,\varepsilon }|} \end{aligned}$$
(43)
$$\begin{aligned} e^{\frac{\theta }{2} N}\int _{\Vert x\Vert ^2\;\geqslant \;N+\varepsilon } \frac{dx}{|S^{N,\varepsilon }|} e^{-\frac{\theta }{2}\Vert x\Vert _2^2+(u,x)}&\;\leqslant \;&e^{\frac{\theta }{2} N-(N+\varepsilon )\frac{\eta }{2}+\frac{q N}{2(\theta -\eta )}}\frac{\sqrt{2\pi }^N}{\theta ^{\frac{N}{2}}|S^{N,\varepsilon }|} \,. \end{aligned}$$
(44)

Thus

$$\begin{aligned} \liminf _N\frac{1}{N}\log \int \sigma ^{(\varepsilon )}_{N}(dx) e^{(u,x)}\;\geqslant \;\max (\Gamma _1, \Gamma _2,\Gamma _3) \end{aligned}$$
(45)

with

$$\begin{aligned} \Gamma _2(\eta ,\theta ):= & {} \frac{q}{2(\theta -\eta )}+\frac{\eta (1-\frac{\varepsilon }{N})}{2}+\frac{\theta -1}{2}-\frac{1}{2}\log \theta \,,\\ \Gamma _3(\eta ,\theta ):= & {} \frac{q}{2(\theta +\eta )}-\frac{\eta (1+\frac{\varepsilon }{N})}{2}+\frac{\theta -1}{2}-\frac{1}{2}\log \theta . \end{aligned}$$

Now we define

$$\begin{aligned} \Delta _{12}(\eta ,\theta ):=\Gamma _1(\theta )-\Gamma _2(\eta ,\theta )\,,\quad \Delta _{13}(\eta ,\theta ):=\Gamma _1(\theta )-\Gamma _3(\eta ,\theta )\,, \end{aligned}$$
(46)

and we seek \(\bar{\theta }>0\) for which \(\Delta _{12},\Delta _{13}\;\geqslant \;0\) for sufficiently small \(\eta \). Since \(\Delta _{12}(0,\theta )=\Delta _{13}(0,\theta )=0\) it suffices to study

$$\begin{aligned} \frac{d}{d\eta } \Delta _{12}\Big |_{\eta =0}\,,\quad \frac{d}{d\eta } \Delta _{13}\Big |_{\eta =0}\,. \end{aligned}$$
(47)

A direct computation shows

$$\begin{aligned} \frac{d}{d\eta } \Delta _{12}\Big |_{\eta =0}= & {} \frac{\varepsilon }{2N}-\partial _\theta \Gamma _1(\theta )\,,\end{aligned}$$
(48)
$$\begin{aligned} \frac{d}{d\eta } \Delta _{13}\Big |_{\eta =0}= & {} \frac{\varepsilon }{2N}+\partial _\theta \Gamma _1(\theta )\,. \end{aligned}$$
(49)

Combining (47), (48) and (49) we see that plugging \(\bar{\theta }=\arg \min \Gamma _1\) into (45) we arrive to

$$\begin{aligned} \lim \inf _N \Gamma _N (u)\;\geqslant \;\min _{\theta >0}\left( \frac{q}{2\theta }+\frac{\theta -1}{2}-\frac{1}{2}\log \theta \right) \,. \end{aligned}$$
(50)

Therefore (40) and (50) give

$$\begin{aligned} \lim _N \Gamma _N (u)= \min _{\theta >0}\left( \frac{q}{2\theta }+\frac{\theta -1}{2}-\frac{1}{2}\log \theta \right) \nonumber \\ \end{aligned}$$

and changing variable \(\theta =(1-r)^{-1}\) we obtain (35). \(\square \)