Skip to main content
Log in

The Multi-species Mean-Field Spin-Glass on the Nishimori Line

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

In this paper we study a multi-species disordered model on the Nishimori line. The typical properties of this line, a set of identities and inequalities among correlation functions, allow us to prove the replica symmetry i.e. the concentration of the order parameter. When the interaction structure is elliptic we rigorously compute the exact solution of the model in terms of a finite-dimensional variational principle and we study its properties.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1

Similar content being viewed by others

References

  1. Agliari, E., Albanese, L., Barra, A., Ottaviani, G.: Replica symmetry breaking in neural networks: a few steps toward rigorous results. J. Phys. A Math. Theor. 53(41), 415005 (2020). https://doi.org/10.1088/1751-8121/abaf2c

    Article  MathSciNet  Google Scholar 

  2. Agliari, E., Barra, A., Sollich, P., Zdeborova, L.: Machine learning and statistical physics: theory, inspiration, application. J. Phys. A Math. Theor. (2020). https://doi.org/10.1088/1751-8121/abaf2c

    Article  Google Scholar 

  3. Alberici, D., Barra, A., Contucci, P., Mingione, E.: Annealing and replica-symmetry in deep Boltzmann machines. J. Stat. Phys. 180, 1–13 (2020)

    Article  ADS  MathSciNet  Google Scholar 

  4. Alberici, D., Camilli, F., Contucci, P., Mingione, E.: The deep Boltzmann machine on the Nishimori line. (In preparation)

  5. Alberici, D., Contucci, P., Mingione, E.: Deep Boltzmann machines: rigorous results at arbitrary depth. arXiv:2004.04495v1 (2020)

  6. Barbier, J., Dia, M., Macris, N., Krzakala, F., Lesieur, T., Zdeborová, L.: Mutual information for symmetric rank-one matrix estimation: a proof of the replica formula. In: Lee, D.D., Sugiyama, M., Luxburg, U.V., Guyon, I., Garnett, R. (eds.) Advances in Neural Information Processing Systems, vol. 29, pp. 424–432. Curran Associates Inc., New York (2016)

    Google Scholar 

  7. Barbier, J., Dia, M., Macris, N., Krzakala, F., Zdeborová, L.: Rank-one matrix estimation: analysis of algorithmic and information theoretic limits by the spatial coupling method. arXiv:1812.02537 (2018)

  8. Barbier, J., Macris, N.: The adaptive interpolation method for proving replica formulas applications to the curie-weiss and wigner spike models. J. Phys. A 52(29), 294002 (2019). https://doi.org/10.1088/1751-8121/ab2735

    Article  MathSciNet  Google Scholar 

  9. Barbier, J., Macris, N., Miolane, L.: The layered structure of tensor estimation and its mutual information. In: 2017 55th Annual Allerton Conference on Communication, Control, and Computing (Allerton), pp. 1056–1063 (2017). https://doi.org/10.1109/ALLERTON.2017.8262854

  10. Barra, A., Contucci, P., Mingione, E., Tantari, D.: Multi-species mean field spin glasses rigorous results. Ann. Henri Poinc. (2013). https://doi.org/10.1007/s00023-014-0341-5

    Article  MATH  Google Scholar 

  11. Barra, A., Genovese, G., Guerra, F.: Equilibrium statistical mechanics of bipartite spin systems. J. Phys. A Math. Theor. 44, 245002 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  12. Chen W (2014) On the mixed even-spin Sherrington-Kirkpatrick model with ferromagnetic interaction. J. Ann. l’inst. Henri Poinc. B Probab. Stat. 50(1), 63–83

  13. Contucci, P., Giardina, C.: The Ghirlanda-Guerra identities. J. Stat. Phys. (2005). https://doi.org/10.1007/978-3-642-22253-5_5

    Article  MATH  Google Scholar 

  14. Contucci, P., Giardinà, C.: Perspectives on Spin Glasses. Cambridge University Press, Cambridge (2012)

    Book  Google Scholar 

  15. Contucci, P., Morita, S., Nishimori, H.: Surface terms on the Nishimori line of the Gaussian Edwards-Anderson model. J. Stat. Phys. (2005). https://doi.org/10.1007/s10955-005-8020-z

    Article  MATH  Google Scholar 

  16. Genovese, G.: Minimax formula for the replica symmetric free energy of deep restricted Boltzmann machines. arXiv:2005.09424 (2020)

  17. Ghirlanda, S., Guerra, F.: General properties of overlap probability distributions in disordered spin systems. Towards Parisi ultrametricity. J. Phys. A Math. General 31(46), 9149–9155 (1998). https://doi.org/10.1088/0305-4470/31/46/006

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Guerra, F.: Broken replica symmetry bounds in the mean field spin glass model. Commun. Math. Phys. 233(1), 1–12 (2003). https://doi.org/10.1007/s00220-002-0773-5

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Guerra, F., Toninelli, F.L.: The thermodynamic limit in mean field spin glass models. Commun. Math. Phys. 230(1), 71–79 (2002). https://doi.org/10.1007/s00220-002-0699-y

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Morita, S., Nishimori, H., Contucci, P.: Griffiths inequalities for the Gaussian spin glass. J. Phys. A Math. General 37(18), L203–L209 (2004). https://doi.org/10.1088/0305-4470/37/18/L03

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Morita, S., Nishimori, H., Contucci, P.: Griffiths inequalities in the Nishimori line. Progress Theor. Phys. Suppl. 157, 73–76 (2005). https://doi.org/10.1143/PTPS.157.73

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Nishimori, H.: Internal energy, specific heat and correlation function of the bond-random Ising model. Progress Theoret. Phys. 66(4), 1169–1181 (1981). https://doi.org/10.1143/PTP.66.1169

    Article  ADS  MathSciNet  Google Scholar 

  23. Nishimori, H.: Statistical Physics of Spin Glasses and Information Processing: An Introduction. Oxford University Press, Oxford (2001)

    Book  Google Scholar 

  24. Panchenko, D.: The free energy in a multi-speciessherrington-kirkpatrick model. Ann. Probab. 43(6), 3494–3513 (2015). https://doi.org/10.1214/14-AOP967

    Article  MathSciNet  MATH  Google Scholar 

  25. Panchenko, D.: The Sherrington-Kirkpatrick Model. Springer Monographs in Mathematics. Springer, New York (2015)

    MATH  Google Scholar 

  26. Talagrand, M.: Mean Field Models for Spin Glasses: Volume I: Basic Examples. Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge / A Series of Modern Surveys in Mathematics. Springer, Berlin (2010)

Download references

Acknowledgements

The authors thank Jean Barbier, Adriano Barra, Francesco Guerra and Nicolas Macris for interesting discussions. In particular we thank Jean Barbier for pointing out the inferential interpretation of our model. D.A. and E.M. acknowledge support from Progetto Alma Idea 2018, Università di Bologna.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Francesco Camilli.

Additional information

Communicated by Simone Warzel.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

A Appendix: Proof of the Concentration Lemma

Proof of Lemma 3

Let us split the proof into three steps for the sake of clarity. As anticipated, it is convenient to split the total fluctuation of \({\mathcal {L}}_r\) into two parts, thus proving that:

$$\begin{aligned}&{\mathbb {E}}_{\varvec{\epsilon }}{\mathbb {E}} \Big \langle \left( {\mathcal {L}}_r-\langle {\mathcal {L}}_r\rangle _{N,t}^{(\epsilon )}\right) ^2\Big \rangle _{N,t}^{(\epsilon )}\longrightarrow 0\quad \text {as }N\rightarrow \infty \end{aligned}$$
(A.1)
$$\begin{aligned}&{\mathbb {E}}_{\varvec{\epsilon }}{\mathbb {E}} \left( \langle {\mathcal {L}}_r\rangle _{N,t}^{(\epsilon )}-{\mathbb {E}}\langle {\mathcal {L}}_r\rangle _{N,t}^{(\epsilon )}\right) ^2\longrightarrow 0\quad \text {as }N\rightarrow \infty \;. \end{aligned}$$
(A.2)

From this moment on, we neglect sub and superscripts in the Gibbs brackets as well as t-dependencies. We start by proving the inequality (45).

Proof of inequality (45): To begin with, we compute:

$$\begin{aligned}&{\mathbb {E}}[\langle {\mathcal {L}}_r\rangle ]=\frac{1}{N_r}\sum _{i\in \varLambda _r}{\mathbb {E}}\Big \langle \sigma _i+\frac{J_i^r\sigma _i}{2\sqrt{Q_{\epsilon , r}}} \Big \rangle ={\mathbb {E}}\langle m_r\rangle +\frac{1}{2}{\mathbb {E}}[1-\langle m_r\rangle ]=\frac{1}{2}{\mathbb {E}}[1+\langle m_r\rangle ] \end{aligned}$$
(A.3)

where integration by parts has been used.

Then, we proceed with:

$$\begin{aligned} {\mathbb {E}}\langle {\mathcal {L}}_r^2\rangle&=\frac{1}{N_r^2}\sum _{i,j\in \varLambda _r}{\mathbb {E}}\Big \langle \sigma _i\sigma _j+\frac{J_i^r\sigma _i\sigma _j}{\sqrt{Q_{\epsilon ,r}}}+\frac{J_i^rJ_j^r\sigma _i\sigma _j}{4Q_{\epsilon ,r}} \Big \rangle =\underbrace{{\mathbb {E}}\langle m_r^2\rangle }_{R_1}+\underbrace{ \frac{1}{N_r^2}\sum _{i,j\in \varLambda _r}{\mathbb {E}}\Big \langle \frac{J_i^r\sigma _i\sigma _j}{\sqrt{Q_{\epsilon ,r}}}\Big \rangle }_{R_2}+\nonumber \\&\quad + \underbrace{ \frac{1}{N_r^2}\sum _{i,j\in \varLambda _r}{\mathbb {E}}\Big \langle \frac{J_i^rJ_j^r\sigma _i\sigma _j}{4Q_{\epsilon ,r}}\Big \rangle }_{R_3}\,. \end{aligned}$$
(A.4)

We treat the three terms \(R_1\), \(R_2\) and \(R_3\) separately with repeated integrations by parts.

$$\begin{aligned} R_2=\frac{1}{N_r^2}\sum _{i,j\in \varLambda _r}{\mathbb {E}} \left[ \langle \sigma _j\rangle -\langle \sigma _i\sigma _j\rangle \langle \sigma _i\rangle \right] ={\mathbb {E}}\langle m_r\rangle -{\mathbb {E}}\langle m_r\rangle ^2 \end{aligned}$$
(A.5)

where we have used the Nishimori identity (16).

$$\begin{aligned} R_3&=\frac{1}{4N_r^2}\sum _{i,j\in \varLambda _r}{\mathbb {E}}\left[ \frac{\delta _{ij}\overbrace{\langle \sigma _i\sigma _j\rangle }^{=1}}{Q_{\epsilon ,r}} +J_j^r\frac{ (\langle \sigma _j\rangle -\langle \sigma _i\rangle \langle \sigma _i\sigma _j\rangle )}{\sqrt{Q_{\epsilon ,r}}}\right] = \frac{1}{4N_rQ_{\epsilon ,r}}\nonumber \\&\quad +\frac{1}{4N_r^2}\sum _{i,j\in \varLambda _r}{\mathbb {E}}\left[ 1-\langle \sigma _j\rangle ^2-\langle \sigma _i\rangle (\langle \sigma _i\rangle -\langle \sigma _j\rangle \langle \sigma _i\sigma _j\rangle )-\langle \sigma _i\sigma _j\rangle (\langle \sigma _i\sigma _j\rangle -\langle \sigma _i\rangle \langle \sigma _j\rangle ) \right] \nonumber \\&= \frac{1}{4N_rQ_{\epsilon ,r}}+\frac{1}{4}-\frac{1}{2}{\mathbb {E}}\langle m_r\rangle +\frac{1}{2}{\mathbb {E}}\langle m_r\rangle ^2-\frac{1}{4}{\mathbb {E}}\langle m_r^2\rangle \,. \end{aligned}$$
(A.6)

Hence:

$$\begin{aligned} R_1+R_2+R_3=\frac{1}{4}+\frac{1}{4N_rQ_{\epsilon ,r}}+\frac{3}{4}{\mathbb {E}}\langle m_r^2\rangle +\frac{1}{2}{\mathbb {E}}\langle m_r\rangle -\frac{1}{2}{\mathbb {E}}\langle m_r\rangle ^2\,. \end{aligned}$$
(A.7)

Summing up all the contributions:

$$\begin{aligned}&{\mathbb {E}}\langle {\mathcal {L}}_r^2\rangle -({\mathbb {E}}\langle {\mathcal {L}}_r\rangle )^2=\frac{1}{4N_rQ_{\epsilon ,r}}+\frac{3}{4}{\mathbb {E}}\langle m_r^2\rangle -\frac{1}{2}{\mathbb {E}}\langle m_r\rangle ^2-\frac{1}{4}({\mathbb {E}}\langle m_r\rangle )^2\nonumber \\&\quad =\frac{1}{4N_rQ_{\epsilon ,r}}+\frac{1}{4}({\mathbb {E}}\langle m_r^2\rangle -({\mathbb {E}}\langle m_r\rangle )^2)+\frac{1}{2}({\mathbb {E}}\langle m_r^2\rangle -{\mathbb {E}}\langle m_r\rangle ^2)\ge \frac{1}{4}{\mathbb {E}}\Big \langle (m_r-{\mathbb {E}}\langle m_r\rangle )^2\Big \rangle \,.\nonumber \\ \end{aligned}$$
(A.8)

Proof of (A.1): Notice that:

(A.9)
$$\begin{aligned}&\frac{\partial ^2 {\bar{p}}_{N,\epsilon }}{\partial Q_{\epsilon ,r}^2}=\alpha _r N_r{\mathbb {E}}\langle ({\mathcal {L}}_r-\langle {\mathcal {L}}_r\rangle )^2\rangle -\frac{1}{4NQ_{\epsilon ,r}^{3/2}}\sum _{i\in \varLambda _r}{\mathbb {E}}\langle J_i^r\sigma _i\rangle \,. \end{aligned}$$
(A.10)

From the last one, after an integration by parts and using the regularity of the map \(\epsilon \longmapsto {\mathbf {Q}}_\epsilon (\cdot )\) and Lemma 2 we get:

$$\begin{aligned}&{\mathbb {E}}_{\varvec{\epsilon }}{\mathbb {E}}\langle ({\mathcal {L}}_r-\langle {\mathcal {L}}_r\rangle )^2\rangle \le \frac{1}{N_r\alpha _rs_N^K}\prod _{s=1}^K\int _{Q_{s_N,s}}^{Q_{2s_N,s}}dQ_{\epsilon ,s}\frac{\partial ^2 {\bar{p}}_{N,\epsilon }}{\partial Q_{\epsilon ,r}^2}+{\mathbb {E}}_{\varvec{\epsilon }}\frac{1}{4N_r\epsilon _r}{\mathbb {E}}[1-\langle m_r\rangle ]\nonumber \\&\quad \le \frac{1}{N_r\alpha _rs_N^K}\prod _{s\ne r,1}^K\int _{Q_{s_N,s}}^{Q_{2s_N,s}}dQ_{\epsilon ,s}\left[ \left. \frac{\partial {\bar{p}}_{N,\epsilon }}{\partial Q_{\epsilon ,r}}\right| _{Q_{2s_N,r}}-\left. \frac{\partial {\bar{p}}_{N,\epsilon }}{\partial Q_{\epsilon ,r}}\right| _{Q_{s_N,r}}\right] +\frac{\log 2}{4N_rs_N}\nonumber \\&\quad \le \frac{2K_r(\varDelta )}{N_rs_N^K}+\frac{\log 2}{4N_rs_N}={\mathcal {O}}\left( \frac{1}{N_rs_N^K} \right) \longrightarrow 0 \end{aligned}$$
(A.11)

where:

$$\begin{aligned} \prod _{s\ne r,1}^K(Q_{2s_N,s}-Q_{s_N,s})\le K_r(\varDelta )\,. \end{aligned}$$
(A.12)

Proof of (A.2): Let \(p_{N,\epsilon }\) be the random interpolating pressure, such that \({\mathbb {E}}p_{N,\epsilon }={\bar{p}}_{N,\epsilon }\). Define:

$$\begin{aligned}&\hat{p}_{N,\epsilon }=p_{N,\epsilon }-\alpha _r\sqrt{Q_{\epsilon ,r}}\sum _{i\in \varLambda _r}\frac{|J_i^r|}{N_r},\;\quad \hat{{\bar{p}}}_{N,\epsilon }={\mathbb {E}}\hat{p}_{N,\epsilon } \end{aligned}$$
(A.13)
$$\begin{aligned}&\frac{\partial ^2 \hat{p}_{N,\epsilon }}{\partial Q_{\epsilon ,r}^2}=\alpha _r N_r\langle ({\mathcal {L}}_r-\langle {\mathcal {L}}_r\rangle )^2\rangle +\frac{\alpha _r}{4Q_{\epsilon ,r}^{3/2}}\sum _{i\in \varLambda _r}\frac{|J_i^r|- J_i^r\langle \sigma _i\rangle }{N_r} \ge 0\,. \end{aligned}$$
(A.14)

Let us evaluate:

$$\begin{aligned} \left| \frac{\partial \hat{p}_{N,\epsilon }}{\partial Q_{\epsilon ,r}}-\frac{\partial \hat{{\bar{p}}}_{N,\epsilon }}{\partial Q_{\epsilon ,r}} \right| \ge \alpha _r\left| \langle {\mathcal {L}}_r\rangle -{\mathbb {E}}\langle {\mathcal {L}}_r\rangle \right| -\frac{\alpha _r|A_r|}{2\sqrt{Q_{\epsilon ,r}}} \end{aligned}$$
(A.15)

where:

$$\begin{aligned} A_r:=\frac{1}{N_r}\sum _{i\in \varLambda _r}[|J_i^r|-{\mathbb {E}}|J_i^r|]. \end{aligned}$$
(A.16)

Thanks to the independence of the \(J_i^r\) it is immediate to verify that \(\exists \, a\ge 0\) s.t.:

$$\begin{aligned} {\mathbb {E}}[A_r^2]\le \frac{a}{N_r}\,. \end{aligned}$$
(A.17)

Using Lemma 3.2 in [25], with notations used in [8]:

$$\begin{aligned}&\left| \frac{\partial \hat{p}_{N,\epsilon }}{\partial Q_{\epsilon ,r}}-\frac{\partial \hat{{\bar{p}}}_{N,\epsilon }}{\partial Q_{\epsilon ,r}} \right| \le \frac{1}{\delta }\sum _{u\in \{Q_{\epsilon ,r}+\delta ,Q_{\epsilon ,r},Q_{\epsilon ,r}-\delta \}}[|\hat{p}_{N,\epsilon }-\hat{{\bar{p}}}_{N,\epsilon }|+\alpha _r\sqrt{u}|A_r|]\nonumber \\&\quad +\;C_\delta ^+(Q_{\epsilon ,r})+C_\delta ^-(Q_{\epsilon ,r}) \end{aligned}$$
(A.18)

with:

$$\begin{aligned}&C_\delta ^\pm (Q_{\epsilon ,r})=|\hat{{\bar{p}}}'_{N,\epsilon }(Q_{\epsilon ,r}\pm \delta )-\hat{{\bar{p}}}'_{N,\epsilon }(Q_{\epsilon ,r})| \end{aligned}$$
(A.19)
$$\begin{aligned}&|\hat{{\bar{p}}}'_{N,\epsilon }|=\left| \frac{\alpha _r}{2}{\mathbb {E}}[1+\langle m_r\rangle ]-\frac{\alpha _r{\mathbb {E}}|J_1^r|}{2\sqrt{Q_{\epsilon ,r}}}\right| \le \alpha _r\left( 1+\frac{C}{2\sqrt{s_N}}\right) \end{aligned}$$
(A.20)
$$\begin{aligned}&C_\delta ^\pm (Q_{\epsilon ,r})\le \alpha _r\left( 2+\frac{C}{\sqrt{s_N}}\right) \end{aligned}$$
(A.21)

where for simplicity we have kept the dependence on \(Q_{\epsilon ,r}\) only, \(\hat{{\bar{p}}}'_{N,\epsilon }\) is the derivative w.r.t. it and \(\delta >0\). Notice that \(\delta \) will be chosen strictly smaller than \(s_N\), so that \(Q_{\epsilon ,r}-\delta \ge \epsilon -\delta \ge s_N-\delta >0\).

Then, using (A.15), (A.16) and (A.18), and thanks to the fact that \((\sum _{i=1}^p\nu _i)^2\le p\sum _{i=1}^p\nu _i^2\), we get:

$$\begin{aligned} \frac{\alpha ^2_r}{9}\left| \langle {\mathcal {L}}_r\rangle -{\mathbb {E}}\langle {\mathcal {L}}_r\rangle \right| ^2&\le \frac{1}{\delta ^2}\sum _{u\in \{Q_{\epsilon ,r}+\delta ,Q_{\epsilon ,r},Q_{\epsilon ,r}-\delta \}}[|\hat{p}_{N,\epsilon }-\hat{{\bar{p}}}_{N,\epsilon }|^2+\alpha _r^2u|A_r|^2]\nonumber \\&\quad +C_\delta ^+(Q_{\epsilon ,r})^2+C_\delta ^-(Q_{\epsilon ,r})^2+\frac{\alpha _r^2A_r^2}{4\epsilon _r}\,. \end{aligned}$$
(A.22)

We first evaluate the two terms containing \(C_\delta ^\pm \):

$$\begin{aligned}&{\mathbb {E}}_{\varvec{\epsilon }}[C_\delta ^+(Q_{\epsilon ,r})^2+C_\delta ^-(Q_{\epsilon ,r})^2]\le 2\alpha _r\left( 2+\frac{C}{\sqrt{s_N}} \right) {\mathbb {E}}_{\varvec{\epsilon }}[C_\delta ^+(Q_{\epsilon ,r})+C_\delta ^-(Q_{\epsilon ,r})] \nonumber \\&\quad \le \frac{2\alpha _r}{s_N^K}\left( 2+\frac{C}{\sqrt{s_N}} \right) \prod _{s=1}^K\int _{Q_{s_N,s}}^{Q_{2s_N,s}}dQ_{\epsilon ,s}[\hat{{\bar{p}}}'_{N,\epsilon }(Q_{\epsilon ,r}+\delta )-\hat{{\bar{p}}}'_{N,\epsilon }(Q_{\epsilon ,r}-\delta )]\nonumber \\&\quad = \frac{2\alpha _r}{s_N^K}\left( 2+\frac{C}{\sqrt{s_N}} \right) \prod _{s\ne r,1}^K \int _{Q_{s_N,s}}^{Q_{2s_N,s}}dQ_{\epsilon ,s}[\hat{{\bar{p}}}_{N,\epsilon }(Q_{2s_N,r}+\delta )-\hat{{\bar{p}}}_{N,\epsilon }(Q_{2s_N,r}-\delta )+\nonumber \\&\quad -\hat{{\bar{p}}}_{N,\epsilon }(Q_{s_N,r}+\delta )+\hat{{\bar{p}}}_{N,\epsilon }(Q_{s_N,r}-\delta )]\le \frac{8\alpha _r^2 K_r(\varDelta )}{s_N^K}\delta \left( 2+\frac{C}{\sqrt{s_N}} \right) ^2\,. \end{aligned}$$
(A.23)

Taking the expectation \({\mathbb {E}}_{\varvec{\epsilon }}{\mathbb {E}}\) in (A.22), and defining \(W_r\) s.t. \(Q_{\epsilon ,r}\le W_r\), we get:

$$\begin{aligned}&\frac{\alpha ^2_r}{9}{\mathbb {E}}_{\varvec{\epsilon }}{\mathbb {E}}\left| \langle {\mathcal {L}}_r\rangle -{\mathbb {E}}\langle {\mathcal {L}}_r\rangle \right| ^2\le \frac{3}{\delta ^2}\left[ \frac{S}{N}+\frac{\alpha _r^2W_ra}{N_r}\right] \nonumber \\&\quad +\frac{8\alpha _r^2 K_r(\varDelta )}{s_N^K}\delta \left( 2+\frac{C}{\sqrt{s_N}} \right) ^{2}+\frac{\alpha _r^2a\log 2}{4N_rs_N}\,. \end{aligned}$$
(A.24)

We can make the r.h.s. vanish by choosing for example: \(\delta =s_N^{2K/3}N^{-1/3}\). The choice \(s_N\propto N^{-1/16K}\) makes the r.h.s. (A.24) behave like \({\mathcal {O}}(N^{-1/4})\). \(\square \)

B Appendix: the SK Case

In the case \(K=1\) the equation (46) reduces to:

$$\begin{aligned} \lim _{N\rightarrow \infty }{\bar{p}}_N(\mu ,h)=\sup _{x\in {\mathbb {R}}_{\ge 0}}\left\{ \mu \frac{(1-x)^2}{4}-\frac{\mu x^2}{2}+\psi (\mu x+h) \right\} \end{aligned}$$
(B.1)

while (48) simply becomes:

$$\begin{aligned} x={\mathbb {E}}_z\tanh \left( z\sqrt{\mu x+h}+\mu x+h\right) :=T(x;\mu ,h)\;. \end{aligned}$$
(B.2)

We collect the main results on this model in the following proposition.

Proposition 4

Define:

$$\begin{aligned} {\bar{p}}_{var}(x;\mu ,h)=\mu \frac{(1-x)^2}{4}-\frac{\mu x^2}{2}+\psi (\mu x+h)\;. \end{aligned}$$
(B.3)

The following hold:

  1. 1.

    if \(\mu <1\) then \({\bar{p}}_{var}\) is concave in x. Equivalently if \(\mu <1\) then \(T(x;\mu ,h)\) is a contraction, and if further \(h=0\) then \(x=0\) is its fixed point;

  2. 2.

    the stable solution of the consistency equation (B.2) is continuous at \((\mu ,h)=(1,0)\):

    $$\begin{aligned} \lim _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h)=0={\bar{x}}(1,0)\;; \end{aligned}$$
    (B.4)
  3. 3.

    for fixed \(h=0\), the magnetization goes to 0 linearly with \(\mu -1\) as \(\mu \rightarrow 1_+\), more precisely:

    $$\begin{aligned} {\bar{x}}=(1+o(1))\frac{\mu -1}{\mu ^2} \end{aligned}$$
    (B.5)

    where o(1) goes to 0 when \(\mu \rightarrow 1_+\). Therefore the critical exponent \(\beta \) (in the Landau classification) is 1, which means that the derivative of the magnetization w.r.t. \(\mu \) does not diverge at the critical point, it only jumps from 0 to 1 and then decreases;

  4. 4.

    Along the line \((\mu ,\lambda (\mu -1)),\;\lambda >0\) in the plane \((\mu ,h)\) the magnetization goes to 0 as follows:

    $$\begin{aligned} {\bar{x}}=\sqrt{\frac{\lambda (\mu -1)}{\mu ^2}}(1+o(1))\; \end{aligned}$$
    (B.6)

    when \(\mu \rightarrow 1_+\), therefore with a critical exponent 1/2;

  5. 5.

    For fixed \(\mu =1\) and \(h\rightarrow 0_+\) the magnetization behaves as:

    $$\begin{aligned} {\bar{x}}^2=h(1+o(1)) \end{aligned}$$
    (B.7)

    where \(o(1)\rightarrow 0\) when \(h\rightarrow 0_+\). Therefore we have a critical exponent \(\delta =2\) (according to Landau’s classification).

Proof

1. The fist assertion follows immediately from (56), since \({\hat{\alpha }}\equiv 1\) and \(\varDelta \equiv \mu \). Then, by (51):

$$\begin{aligned} \frac{dT}{dx}(x;\mu ,h)=\mu {\mathbb {E}}_z\left[ \left( 1-\tanh ^2\left( z\sqrt{\mu x+h}+\mu x+h\right) \right) ^2\right] \le \mu <1 \end{aligned}$$

that implies T is a contraction. It is easy to see that if \(h=0\) then \(x=0\) is a solution of the fixed point equation which must be unique by Banach’s fixed point theorem.

2. Using continuity and monotonicity of T (see (51)):

$$\begin{aligned}&\limsup _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h)=T(\limsup _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h);1,0)\\&\liminf _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h)=T(\liminf _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h);1,0) \end{aligned}$$

hence both \(\limsup _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h)\) and \(\liminf _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h)\) satisfy the consistency equation:

$$\begin{aligned} m={\mathbb {E}}_z\tanh (z\sqrt{m}+m) \end{aligned}$$

whose solution \(m=0\) is unique, since the derivative of T(m; 1, 0) is \(\le 1\) and equality holds only at \(m=0\). We conclude that there exists

$$\begin{aligned} \lim _{(\mu ,h)\rightarrow (1,0)}{\bar{x}}(\mu ,h)=0={\bar{x}}(1,0)\;. \end{aligned}$$
(B.8)

3. We first notice that

$$\begin{aligned} {\mathbb {E}}\tanh ^2\left( z\sqrt{Q}+Q\right) ={\mathbb {E}}\tanh \left( z\sqrt{Q}+Q\right) \,,\quad Q\ge 0\,,z\sim {\mathcal {N}}(0,1) \end{aligned}$$
(B.9)

which simply follows from the third relation in (51) and the identity (14). Indeed, the quantity in (B.9) is nothing but the quenched average magnetization of a free system. Now, by computing the first and second derivatives of the map \(T(x;\mu ,0)\) and using (B.9) we get:

$$\begin{aligned}&T'(0;\mu ,0)=\mu \,,\quad T''(0;\mu ,0)=-2\mu ^2\\&{\bar{x}}=\mu {\bar{x}}-\mu ^2 {\bar{x}}^2(1+o(1))\quad \Rightarrow \quad {\bar{x}}=(1+o(1))\left( \frac{\mu -1}{\mu ^2}\right) \;, \end{aligned}$$

which implies that, in proximity of \(\mu =1\), the magnetization goes to 0 with a critical exponent \(\beta =1\) (not to be confused with inverse absolute temperature) and with slope 1.

4. An analogous expansion of T yields:

$$\begin{aligned} {\bar{x}}=T({\bar{x}};\mu ,\lambda (\mu -1))=\mu {\bar{x}}+\lambda (\mu -1)-(\mu ^2{\bar{x}}^2+o(\mu -1))(1+o(1)) \end{aligned}$$

which in turn entails:

$$\begin{aligned} {\bar{x}}^2=\frac{\lambda (\mu -1)}{\mu ^2}(1+o(1))\;. \end{aligned}$$

5. Here by o(1) we mean a quantity that approaches 0 as \(h\rightarrow 0_+\). As in the previous steps:

$$\begin{aligned} {\bar{x}}=T({\bar{x}};1,h)={\bar{x}}+h-({\bar{x}}^2+o(h))(1+o(1))\;, \end{aligned}$$

then we get:

$$\begin{aligned} {\bar{x}}^2=h(1+o(1))\quad \Rightarrow \quad \delta =2\;. \end{aligned}$$

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Alberici, D., Camilli, F., Contucci, P. et al. The Multi-species Mean-Field Spin-Glass on the Nishimori Line. J Stat Phys 182, 2 (2021). https://doi.org/10.1007/s10955-020-02684-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-020-02684-z

Keywords

Mathematics Subject Classification

Navigation