Skip to main content
Log in

Revisiting the Replica Trick: Competition Between Spin Glass and Conventional Order

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

Abstract

There is an ambiguity in how to apply the replica trick to spin glass models which have additional order parameters unrelated to spin glass order—with respect to which quantities does one minimize vs maximize the action, and in what sequence? Here we show that the correct procedure is to first maximize with respect to “replica” order parameters, and then minimize with respect to “conventional” order parameters. With this result, we further elucidate the relationship between quenched free energies, annealed free energies, and replica order—it is possible for the quenched and annealed free energies to differ even while all replica order parameters remain zero.

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.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Data Availibility

No datasets were generated during this work, but code and/or further details of calculations are available from the corresponding author upon request.

Notes

  1. We refer to standard textbooks [2, 6, 9] for a full discussion of this point, but briefly, broken ergodicity and the existence of multiple equilibrium states implies that some replicas may lie in the same state while others lie in different states, hence a lack of permutation symmetry among replicas.

  2. In fact, the quantities Q will carry only two replica indices (as opposed to higher numbers as well) only when the couplings \(J_{ij}\) are Gaussian-distributed and enter linearly into the Hamiltonian, such as in Eq. (3). This is by far the situation most considered in the literature, at least for infinite-range models, and we focus on it as well.

  3. When expressed as path integrals, the degrees of freedom in quantum models have imaginary-time dependence, meaning that the (single-replica) imaginary-time correlation function appears as a conventional order parameter once averaging over disorder.

  4. To see this, note for all R and Q, we have by definition that \(S_n(R, Q_c(R)) \ge S_n(R, 0) \ge S_n(R_{\text {A}}, 0)\). Thus if \(Q_c(R_{\text {A}}) = 0\), the minimum of \(S_n(R, Q_c(R))\) must be at \(R_{\text {A}}\), meaning \(f_{\text {Q}} = S_n(R_{\text {A}}, 0) = f_{\text {A}}\).

  5. Strictly speaking, Ref. [14] considers models that consist only of infinite-range Gaussian interactions, but it is clear from the proof technique that the result holds regardless of any other terms in the Hamiltonian as well (as long as those terms are independent of the random interactions). In short, averaging the partition function over the Gaussian couplings always gives a factor with exponent going as \(\beta ^2\) (see Eq. (4) for an example), irrespective of what other terms are in the Hamiltonian. This translates to an upper bound on the annealed free energy going as \(-\beta \), hence \(f_{\text {A}} \rightarrow -\infty \) as \(\beta \rightarrow \infty \), again regardless of any additional terms. Yet in systems with a finite local Hilbert space, such behavior cannot occur in the quenched free energy. See Ref. [14] for further details.

  6. Note that Eqs. (20) and (21) follow directly from Eq. (18), without needing to introduce the effective single-spin Hamiltonian seen in Eq. (19). The saddle-point equations give R and Q additional interpretations as expectation values of \(\sigma ^{\alpha }(\tau ) \sigma ^{\alpha }(\tau ')\) and \(\sigma ^{\alpha }(\tau ) \sigma ^{\alpha '}(\tau ')\) with respect to that single-spin Hamiltonian, but we will not need those interpretations for the present analysis.

  7. This is not an exhaustive list. Barring a rigorous derivation (which this paper does not provide), there is always the possibility that a more complicated prescription may be correct and simply happen to reduce to the min-max prescription in the cases considered here. We cannot rule this out, and encourage further investigation.

  8. One can easily confirm, by reproducing the steps from Eq. (15) to Eq. (26) but with \(n = 1\) throughout, that calculating \(f_{\text {A}}\) is indeed equivalent to minimizing over R while fixing \(m = 1\) and \(Q = 0\), i.e., \(f_{\text {A}} = \min _R S_n(R, 1, 0)\).

  9. We did not run into this issue in Sect. 2 only because the off-diagonal elements of the Hessian happened to be subleading at large p. That said, we found that the prescription of minimizing among stable saddle points gave incorrect results nonetheless.

  10. There is another sense in which one could define local stability, in terms of the eigenvalues of the Hessian with respect to fluctuations in individual elements of the overlap matrix \(Q_{\alpha \alpha '}\). It is known that the dominant saddle point of the replicated action has positive eigenvalues even in the \(n \rightarrow 0\) limit [6, 45], just as a conventional action in terms of conventional order parameters would. We see no reason why the requirement of positive eigenvalues would not continue to hold for an action with simultaneous replica and conventional order parameters. However, this sense of local stability still does not allow one to choose from multiple locally stable saddle points at first-order transitions. Thus the issue of the correct prescription remains.

  11. To see this, consider the case where f(X) is simply a product of Gaussians: \(f(X) = \prod _k X_{b_k}\). Then Eq. (51) amounts to the statement \({\mathbb {E}} [X_a \prod _k X_{b_k}] = \sum _k {\mathbb {E}}[X_a X_{b_k}] {\mathbb {E}}[\prod _{l \ne k} X_{b_l}]\), which is Wick’s theorem expressed recursively.

  12. One can allow for \(w(\alpha )\) to itself be random and independent of the other quantities, as is the case in Eq. (65). Since \(f \ge {\overline{f}}\) for any individual realization of \(w(\alpha )\), it trivially holds that \(f \ge {\mathbb {E}}_w {\overline{f}}\) as well.

  13. Here the sum over \((i_1 \cdots i_p)\) is over all tuples of p spin indices, i.e., over all \(i_1\) through \(i_p\) such that \(i_1< \cdots < i_p\), whereas the sum over \(\tau _1 \cdots \tau _p\) is over all \(\tau _1\) through \(\tau _p\) without any restriction on the ordering.

  14. To be completely explicit, we are requiring that for any matrices X and Y, and any \(\lambda \in [0, 1]\),

    $$\begin{aligned} V \big ( (1 - \lambda ) X + \lambda Y \big ) \le (1 - \lambda ) V(X) + \lambda V(Y). \end{aligned}$$

    An immediate consequence is that, again for any X and Y,

    $$\begin{aligned} V(X) + \sum _{\tau \tau '} Y_{\tau \tau '} \frac{\partial V(X)}{\partial X_{\tau \tau '}} \le V(X + Y). \end{aligned}$$
  15. As a simple example, consider a spin-1 model: \(\sigma _i \in \{-1, 0, 1\}\). Then \(\sigma \cdot \sigma \equiv N^{-1} \sum _i \sigma _i^2\) can lie anywhere between 0 and 1 depending on the configuration.

  16. Again considering the spin-1 example, R is the value of \(N^{-1} \sum _i \sigma _i^2\), which is precisely the order parameter used in Ref. [28] to analyze the model at large p.

  17. Note that Eq. (100) establishes g(h) as the Legendre transform of s(m). The discussion that follows is really just an explanation of how to invert the Legendre transform.

  18. Since \(\partial m^*(h)/\partial h\) is always positive (as one can explicitly check from Eq. (101)) and \(\lim _{h \rightarrow \pm \infty } m^*(h) = \pm 1\), there is exactly one solution to \(m^*(h) = m\) for all \(m \in (-1, 1)\).

  19. One might wonder why we allow ourselves to use the explicit expression for g(h) when we are pretending to not know the result for s(m). Generically, evaluating s(m) directly will involve a sum over all N degrees of freedom subject to a constraint (here that \(\sum _i \sigma _i = Nm\)). On the other hand, \(Z_0(h)\) is a non-interacting partition function, and thus evaluating g(h) involves a single sum over one degree of freedom. The latter is often significantly simpler, hence the reason to consider g(h).

  20. This is natural to expect—\(\partial S^*(m) / \partial m = \partial \epsilon (m) / \partial m + h^*(m)\) and \(h^*(m) \rightarrow \pm \infty \) as \(m \rightarrow \pm 1\), meaning the minimum cannot lie at either endpoint unless \(\partial \epsilon (m) / \partial m\) diverges there.

  21. For example, suppose \(\epsilon (m) = m^p\) for \(p > 2\). Then \(\partial h^{\times }/\partial m\) does equal 0 at \(m = 0\), but this is a stationary point of \(S^*(m)\) anyway.

References

  1. Binder, K., Young, A.P.: Spin glasses: experimental facts, theoretical concepts, and open questions. Rev. Mod. Phys. 58, 801 (1986)

    ADS  Google Scholar 

  2. Mezard, M., Parisi, G., Virasoro, M.A.: Spin Glass Theory and Beyond. World Scientific, Singapore (1987)

    MATH  Google Scholar 

  3. Fischer, K.H., Hertz, J.A.: Spin Glasses. CUP, Cambridge (1991)

    Google Scholar 

  4. Mydosh, J.A.: Spin Glasses: An Experimental Introduction. Taylor & Francis, London (1993)

    Google Scholar 

  5. Nishimori, H.: Statistical Physics of Spin Glasses and Information Processing: An Introduction. OUP, Oxford (2001)

    MATH  Google Scholar 

  6. Mezard, M., Montanari, A.: Information, Physics, and Computation. OUP, Oxford (2009)

    MATH  Google Scholar 

  7. Maldacena, J., Stanford, D.: Remarks on the Sachdev-Ye-Kitaev model. Phys. Rev. D 94, 106002 (2016)

    ADS  MathSciNet  Google Scholar 

  8. Chowdhury, D., Georges, A., Parcollet, O., Sachdev, S.: Sachdev-Ye-Kitaev models and beyond: window into non-Fermi liquids. Rev. Mod. Phys. 94, 035004 (2022)

    ADS  MathSciNet  Google Scholar 

  9. Castellani, T., Cavagna, A.: Spin-glass theory for pedestrians. J. Stat. Mech.: Theory Exp. 2005(05), P05012 (2005)

    MathSciNet  MATH  Google Scholar 

  10. Charbonneau, P.: From the replica trick to the replica symmetry breaking technique. arXiv:2211.01802 (2022)

  11. Zdeborová, L., Krzakala, F.: Statistical physics of inference: thresholds and algorithms. Adv. Phys. 65, 453 (2016)

    ADS  Google Scholar 

  12. Foini, L., Kurchan, J.: Annealed averages in spin and matrix models. SciPost Phys. 12, 080 (2022)

    ADS  MathSciNet  Google Scholar 

  13. Saad, P., Shenker, S.H., Stanford, D.: JT gravity as a matrix integral. arXiv:1903.11115 (2019)

  14. Baldwin, C.L., Swingle, B.: Quenched vs annealed: glassiness from SK to SYK. Phys. Rev. X 10, 031026 (2020)

    Google Scholar 

  15. Gur-Ari, G., Mahajan, R., Vaezi, A.: Does the SYK model have a spin glass phase? J. High Energy Phys. 2018(11), 70 (2018)

    MathSciNet  MATH  Google Scholar 

  16. Wang, H., Bagrets, D., Chudnovskiy, A.L., Kamenev, A.: On the replica structure of Sachdev-Ye-Kitaev model. J. High Energy Phys. 2019(9), 57 (2019)

    MathSciNet  MATH  Google Scholar 

  17. Tanaka, T.: Moment problem in replica method. Interdiscip. Inf. Sci. 13(1), 17 (2007)

    MathSciNet  MATH  Google Scholar 

  18. Mourrat, J.-C.: Nonconvex interactions in mean-field spin glasses. Probab. Math. Phys. 2(2), 281 (2021)

    MathSciNet  MATH  Google Scholar 

  19. Sherrington, D., Kirkpatrick, S.: Solvable model of a spin-glass. Phys. Rev. Lett. 35, 1792 (1975)

    ADS  Google Scholar 

  20. Bovier, A.: Statistical Mechanics of Disordered Systems: A Mathematical Perspective. CUP, Cambridge (2006)

    MATH  Google Scholar 

  21. Talagrand, M.: Mean Field Models for Spin Glasses, vol. 1. Springer, New York (2011)

    MATH  Google Scholar 

  22. Talagrand, M.: Mean Field Models for Spin Glasses, vol. 2. Springer, New York (2011)

    MATH  Google Scholar 

  23. Panchenko, D.: The Sherrington-Kirkpatrick Model. Springer, New York (2013)

    MATH  Google Scholar 

  24. Goldschmidt, Y.Y.: Solvable model of the quantum spin glass in a transverse field. Phys. Rev. B 41, 4858 (1990)

    ADS  Google Scholar 

  25. Nieuwenhuizen, T.M., Ritort, F.: Quantum phase transition in spin glasses with multi-spin interactions. Phys. A: Stat. Mech. Appl. 250(1–4), 8 (1998)

    Google Scholar 

  26. Cugliandolo, L.F., Grempel, D.R., da Silva Santos, C.A.: From second to first order transitions in a disordered quantum magnet. Phys. Rev. Lett. 85, 2589 (2000)

    ADS  Google Scholar 

  27. Cugliandolo, L.F., Grempel, D.R., da Silva Santos, C.A.: Imaginary-time replica formalism study of a quantum spherical p-spin-glass model. Phys. Rev. B 64, 014403 (2001)

    ADS  Google Scholar 

  28. Mottishaw, P.: First-order spin glass transitions: an exact solution. EPL 1, 409 (1986)

    ADS  Google Scholar 

  29. Panchenko, D.: Free energy in the Potts spin glass. Ann. Probab. 46, 829 (2018)

    MathSciNet  MATH  Google Scholar 

  30. Panchenko, D.: Free energy in the mixed p-spin models with vector spins. Ann. Probab. 46, 865 (2018)

    MathSciNet  MATH  Google Scholar 

  31. Camilli, F., Contucci, P., Mingione, E.: An inference problem in a mismatched setting: a spin-glass model with Mattis interactions. SciPost Phys. 12, 125 (2022)

    ADS  MathSciNet  Google Scholar 

  32. Dobrosavljevic, V., Thirumalai, D.: 1/p expansion for a p-spin interaction spin-glass model in a transverse field. J. Phys. A: Math. Gen. 23(15), L767 (1990)

    ADS  Google Scholar 

  33. Baldwin, C.L., Laumann, C.R., Pal, A., Scardicchio, A.: Clustering of nonergodic eigenstates in quantum spin glasses. Phys. Rev. Lett. 118, 127201 (2017)

    ADS  Google Scholar 

  34. Biroli, G., Facoetti, D., Schiró, M., Tarzia, M., Vivo, P.: Out-of-equilibrium phase diagram of the quantum random energy model. Phys. Rev. B 103, 014204 (2021)

    ADS  Google Scholar 

  35. Jörg, T., Krzakala, F., Kurchan, J., Maggs, A.C.: Simple glass models and their quantum annealing. Phys. Rev. Lett. 101, 147204 (2008)

    ADS  Google Scholar 

  36. Baldwin, C.L., Laumann, C.R.: Quantum algorithm for energy matching in hard optimization problems. Phys. Rev. B 97, 224201 (2018)

    ADS  Google Scholar 

  37. Smelyanskiy, V.N., Kechedzhi, K., Boixo, S., Isakov, S.V., Neven, H., Altshuler, B.: Nonergodic delocalized states for efficient population transfer within a narrow band of the energy landscape. Phys. Rev. X 10, 011017 (2020)

    Google Scholar 

  38. Gardner, E.: Spin glasses with p-spin interactions. Nucl. Phys. B 257, 747 (1985)

    ADS  MathSciNet  Google Scholar 

  39. Derrida, B.: Random-energy model: limit of a family of disordered models. Phys. Rev. Lett. 45, 79 (1980)

    ADS  MathSciNet  Google Scholar 

  40. Derrida, B.: Random-energy model: an exactly solvable model of disordered systems. Phys. Rev. B 24, 2613 (1981)

    ADS  MathSciNet  MATH  Google Scholar 

  41. Gross, D., Mezard, M.: The simplest spin glass. Nucl. Phys. B 240(4), 431 (1984)

    ADS  MathSciNet  Google Scholar 

  42. Derrida, B., Gardner, E.: Solution of the generalised random energy model. J. Phys. C: Solid State Phys. 19(13), 2253 (1986)

    ADS  Google Scholar 

  43. Mora, T., Zdeborová, L.: Random subcubes as a toy model for constraint satisfaction problems. J. Stat. Phys. 131(6), 1121 (2008)

    ADS  MathSciNet  MATH  Google Scholar 

  44. Gross, D.J., Kanter, I., Sompolinsky, H.: Mean-field theory of the Potts glass. Phys. Rev. Lett. 55, 304 (1985)

    ADS  Google Scholar 

  45. de Almeida, J.R.L., Thouless, D.J.: Stability of the Sherrington-Kirkpatrick solution of a spin glass model. J. Phys. A: Math. Gen. 11(5), 983 (1978)

    ADS  Google Scholar 

  46. Aizenman, M., Sims, R., Starr, S.L.: Extended variational principle for the Sherrington-Kirkpatrick spin-glass model. Phys. Rev. B 68, 214403 (2003)

    ADS  Google Scholar 

  47. Crisanti, A., Sommers, H.J.: The spherical p-spin interaction spin glass model: the statics. Z. Phys. B 87(3), 341 (1992)

    ADS  Google Scholar 

  48. Crisanti, A., Horner, H., Sommers, H.J.: The spherical p-spin interaction spin-glass model. Z. Phys. B 92(2), 257 (1993)

    ADS  Google Scholar 

  49. Auerbach, A.: Interacting Electrons and Quantum Magnetism. Springer, New York (1994)

    Google Scholar 

  50. den Hollander, F.: Large Deviations. AMS, Providence (2000)

    MATH  Google Scholar 

  51. Barbier, J., Macris, N.: The adaptive interpolation method for proving replica formulas. Applications to the Curie-Weiss and Wigner spike models. J. Phys. A: Math. Theor. 52(29), 294002 (2019)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

It is a pleasure to thank L. Foini for valuable and informative discussions. C.L.B. was supported by the AFOSR, AFOSR MURI, DoE ASCR Quantum Testbed Pathfinder program (award No. DE-SC0019040), DoE ASCR Accelerated Research in Quantum Computing program (award No. DE-SC0020312), DoE QSA, NSF QLCI (award No. OMA-2120757), NSF PFCQC program, ARO MURI, and DARPA SAVaNT ADVENT. B.S. was supported by the DoE (award No. DE-SC0009986).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christopher L. Baldwin.

Ethics declarations

Conflict of interest

The authors have no conflicts of interest to declare.

Additional information

Communicated by Giulio Biroli.

Publisher's Note

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

Appendices

A Extremizing with Respect to Order Parameters vs Lagrange Multipliers

Entirely unrelated to the replica trick and spin glass physics, there are subtleties in how one extremizes over (conventional) order parameters and their associated Lagrange multipliers. This is an old topic and we are certainly not the first to consider it (to the point where it often passes without comment in the literature). Yet since the present work is specifically concerned with the order in which one extremizes an action with respect to various quantities, we feel that it is appropriate to give a clear discussion of the issue here.

As a concrete example, consider a classical spin-1/2 Ising model with “mean-field” interactions:

$$\begin{aligned} H(\sigma ) = N \epsilon \bigg ( N^{-1} \sum _i \sigma _i \bigg ), \end{aligned}$$
(96)

for some function \(\epsilon (m)\). In other words, the energy can be written as a function solely of the magnetization density \(N^{-1} \sum _i \sigma _i\). To evaluate the partition function, we can separate the trace into an outer sum over values of the magnetization m and an inner sum over \(\sigma \) such that \(N^{-1} \sum _i \sigma _i = m\):

$$\begin{aligned} Z \equiv \sum _{\sigma } \exp {\big [ -\beta H(\sigma ) \big ]} = \int _{-1}^1 \text {d}m \exp {\big [ -N \beta \epsilon (m) \big ]} \sum _{\sigma } \delta \bigg ( m - N^{-1} \sum _i \sigma _i \bigg ). \end{aligned}$$
(97)

Defining \(\sum _{\sigma } \delta (m - N^{-1} \sum _i \sigma _i) \equiv \exp {[Ns(m)]}\) and evaluating the integral over m by saddle point, we have that

$$\begin{aligned} -\lim _{N \rightarrow \infty } (N \beta )^{-1} \log {Z} = \min _{m \in [-1, 1]} \Big [ \epsilon (m) - \beta ^{-1} s(m) \Big ]. \end{aligned}$$
(98)

Let us pretend that we do not have an explicit expression for \(\exp {[Ns(m)]}\)—while it is simply a binomial coefficient in the present example, it may not have a closed form more generally. There are then two ways to proceed. One often sees the \(\delta \)-function expressed in integral form as \((2\pi )^{-1} \int _{-i \infty }^{i \infty } N \text {d}h \exp {\big [ -Nhm + h \sum _i \sigma _i \big ]}\) (note that h runs along the imaginary axis). We will discuss this approach momentarily. Alternatively, one can use a method more along the lines of large deviation theory [50] and consider the auxiliary quantity

$$\begin{aligned} Z_0(h) \equiv \sum _{\sigma } \exp {\Big [ \beta h \sum _i \sigma _i \Big ]} = \exp {\big [ -N \beta g(h) \big ]}, \end{aligned}$$
(99)

where \(g(h) \equiv -\beta ^{-1} \log {2 \cosh {\beta h}}\). Since one could again separate the sum over \(\sigma \) into an outer and inner sum just as in Eq. (97), we have thatFootnote 17

$$\begin{aligned} g(h) = \min _{m \in [-1, 1]} \Big [ -hm - \beta ^{-1} s(m) \Big ]. \end{aligned}$$
(100)

Denote the location of the minimum, which will be a function of h, by \(m^*(h)\). There is an explicit expression for \(m^*(h)\): just as Eq. (99) is dominated by \(\sigma \) with magnetizations close to \(m^*(h)\), so is

$$\begin{aligned} \frac{\partial \log {Z_0(h)}}{\partial h} = \sum _{\sigma } \bigg ( \beta \sum _i \sigma _i \bigg ) \frac{\exp {\big [ \beta h \sum _i \sigma _i \big ]}}{\sum _{\sigma '} \exp {\big [ \beta h \sum _i \sigma '_i \big ]}} \sim N \beta m^*(h), \end{aligned}$$
(101)

i.e., \(m^*(h) = -\partial g(h) / \partial h\). At this point, note that if one chooses h so thatFootnote 18\(m^*(h) = m\), then from Eq. (100), one has \(\beta ^{-1} s(m) = -hm - g(h)\) and

$$\begin{aligned} -\lim _{N \rightarrow \infty } (N \beta )^{-1} \log {Z} = \min _{m \in [-1, 1]} \Big [ \epsilon (m) + hm + g(h) \Big ]. \end{aligned}$$
(102)

Since we have an explicit expressionFootnote 19 for g(h), Eq. (102) can readily be evaluated (keeping in mind that h is a function of m defined by \(m^*(h) = m\)).

In fact, since h solves the equation \(m = m^*(h) = -\partial g(h) / \partial h\), we can view h as being determined by extremizing the “action” \(\epsilon (m) + hm + g(h)\) at fixed m. Thus the free energy is determined by extremizing with respect to both m and h. However, note that the second derivative with respect to h is

$$\begin{aligned} \frac{\partial ^2 g(h)}{\partial h^2} = -N \beta \left[ \Big< \bigg ( N^{-1} \sum _i \sigma _i \bigg )^2 \Big> - \Big < N^{-1} \sum _i \sigma _i \Big >^2 \right] , \end{aligned}$$
(103)

where \(\langle \, \cdot \, \rangle \) denotes a thermal expectation value with respect to \(h \sum _i \sigma _i\). Thus the second derivative is automatically negative, and the free energy is maximized with respect to h. Since h is really a function of m in Eq. (102), the maximization occurs inside the minimization, meaning we can write

$$\begin{aligned} -\lim _{N \rightarrow \infty } (N \beta )^{-1} \log {Z} = \min _{m \in [-1, 1]} \max _h \Big [ \epsilon (m) + hm + g(h) \Big ]. \end{aligned}$$
(104)

Interestingly, this is another “min-max” prescription, albeit one unrelated to that of the main text (although see Ref. [51] for an alternate derivation using the interpolation techniques of Sect. 4).

The same min-max prescription is hidden within the approach to calculating s(m) based on the integral representation of \(\delta (m - N^{-1} \sum _i \sigma _i)\). In this approach, starting from Eq. (97), we have that

$$\begin{aligned} Z \sim \int _{-1}^1 \text {d}m \int _{-i \infty }^{i \infty } \text {d}h \exp {\Big [ -N \beta \big [ \epsilon (m) + hm + g(h) \big ] \Big ]}, \end{aligned}$$
(105)

with the same g(h) as defined in Eq. (99). The right-hand side can be evaluated by saddle point, but since h initially runs along the imaginary axis, its contour must be deformed to pass through the (real) solution to \(m = -\partial g(h) / \partial h\). We do need that the action be minimized with respect to h along the trajectory of the contour, but this is fully consistent with the fact that \(\partial ^2 g(h) / \partial h^2 < 0\) for real h since the contour passes through the solution vertically. The second derivative being negative in the real direction implies that it is positive in the imaginary direction, as required. Thus we are in fact maximizing the action with respect to real h after all.

Regardless of the approach, it is clear that m and h play different roles. m is undeniably an order parameter—from the beginning, we use it to decompose the original partition function (Eq. (97)). h is instead a Lagrange multiplier—we use it to enforce the constraint that \(N^{-1} \sum _i \sigma _i = m\). The conclusion here is that one should first maximize the effective action with respect to Lagrange multipliers, and then minimize with respect to order parameters.

With the proper ordering in mind, let us lastly consider the solution to the saddle point equations. We have the pair

$$\begin{aligned} \frac{\partial g(h)}{\partial h} = -m, \qquad \frac{\partial \epsilon (m)}{\partial m} = -h, \end{aligned}$$
(106)

where the former is to be solved for h, and then the latter is to be solved for m. Nonetheless, it is tempting to interpret the latter equation as determining h and then use that expression in the former to obtain an equation for m. In fact, this is what we do in Sect. 2 of the main text—we use Eq. (22) to solve for the Lagrange multipliers K and \(\Lambda \). Although decidedly not the procedure we have derived thus far, the substitution \(h = -\partial \epsilon (m) / \partial m\) turns out to be justified, as we now show.

To be precise, let \(h^*(m)\) be the solution to \(\partial g(h) / \partial h = -m\), and let \(h^{\times }(m)\) denote the function \(-\partial \epsilon (m) / \partial m\). We have already established that the correct free energy is obtained by minimizing \(S^*(m) \equiv \epsilon (m) + h^* m + g(h^*)\) with respect to m—taking a derivative (assuming the minimum lies in the interiorFootnote 20 of \([-1, 1]\)) leads to the equation \(\partial \epsilon (m) / \partial m = -h^*(m)\). Now instead consider minimizing \(S^{\times }(m) \equiv \epsilon (m) + h^{\times } m + g(h^{\times })\)—taking a derivative gives \([m + \partial g(h^{\times }) / \partial h^{\times }] \partial h^{\times } / \partial m = 0\). Thus unless \(\partial h^{\times } / \partial m = 0\) (a case that can often be treated separatelyFootnote 21), the extrema of \(S^{\times }\) occur where \(m = -\partial g(h^{\times }) / \partial h^{\times }\). Either way — whether minimizing \(S^*\) or \(S^{\times }\)—the same equations are being solved (namely Eq. (106)) and the same two-parameter action is being evaluated (namely \(\epsilon (m) + hm + g(h)\)). Thus the correct global minimum is identified (except for points at which \(\partial h^{\times } / \partial m = 0\)). This is true even though \(S^*(m) \ne S^{\times }(m)\) for general values of m.

B Consequences of Sub-additivity

We demonstrated in Sect. 4.1 of the main text that the disorder-averaged free energy of the classical p-spin model is sub-additive, \(F^{(N+L)} \le F^{(N)} + F^{(L)}\) (where the superscript indicates the system size). This implies both that \(f \equiv \lim _{N \rightarrow \infty } F^{(N)}/N\) exists and that it can be written as in Eq. (47), reproduced here:

$$\begin{aligned} f = \lim _{N \rightarrow \infty } \limsup _{L \rightarrow \infty } \frac{F^{(N+L)} - F^{(L)}}{N}. \end{aligned}$$
(107)

For completeness, we prove this statement here (following Ref. [23]).

The fact that sub-additivity implies the existence of \(\lim _{N \rightarrow \infty } F^{(N)}/N\) goes by the name of Fekete’s lemma. To prove it, pick integers M and P, and note that we inductively have

$$\begin{aligned} \frac{F^{(KM+P)}}{KM+P} \le \frac{K F^{(M)}}{KM+P} + \frac{F^{(P)}}{KM+P}. \end{aligned}$$
(108)

Taking \(K \rightarrow \infty \) gives

$$\begin{aligned} \limsup _{K \rightarrow \infty } \frac{F^{(KM+P)}}{KM+P} \le \frac{F^{(M)}}{M}. \end{aligned}$$
(109)

This holds for all \(P \in \{0, 1, \cdots , M-1\}\), thus \(\limsup _{N \rightarrow \infty } F^{(N)}/N \le F^{(M)}/M\). Taking the liminf as \(M \rightarrow \infty \) then gives

$$\begin{aligned} \limsup _{N \rightarrow \infty } \frac{F^{(N)}}{N} \le \liminf _{M \rightarrow \infty } \frac{F^{(M)}}{M}, \end{aligned}$$
(110)

i.e., the two must be equal and the limit exists.

Having established that \(f \equiv \lim _{N \rightarrow \infty } F^{(N)}/N\) exists, now turn to Eq. (107). Since \(F^{(N)} \ge F^{(N+L)} - F^{(L)}\) for all L, we certainly have that \(F^{(N)} \ge \limsup _{L \rightarrow \infty } [F^{(N+L)} - F^{(L)}]\). Dividing by N and taking \(N \rightarrow \infty \) then gives

$$\begin{aligned} f \ge \limsup _{N \rightarrow \infty } \limsup _{L \rightarrow \infty } \frac{F^{(N+L)} - F^{(L)}}{N}. \end{aligned}$$
(111)

At the same time, we have that for all N,

$$\begin{aligned} \begin{aligned} f = \lim _{j \rightarrow \infty } \frac{F^{(jN)}}{jN}&= \lim _{j \rightarrow \infty } \frac{1}{j} \sum _{i=0}^{j-1} \frac{F^{((i+1)N)} - F^{(iN)}}{N} \\&\quad \le \limsup _{j \rightarrow \infty } \frac{F^{((j+1)N)} - F^{(jN)}}{N} \le \limsup _{L \rightarrow \infty } \frac{F^{(N+L)} - F^{(L)}}{N}. \end{aligned} \end{aligned}$$
(112)

Taking \(N \rightarrow \infty \) then gives

$$\begin{aligned} f \le \liminf _{N \rightarrow \infty } \limsup _{L \rightarrow \infty } \frac{F^{(N+L)} - F^{(L)}}{N}, \end{aligned}$$
(113)

and Eq. (107) follows.

Lastly, we needed analogues of these results in Sect. 4.4, where we considered the partition function \(Z^{(N)}(R)\) and corresponding free energy \(F^{(N)}(R)\) of states restricted to have a certain value R of the self-overlap (technically a matrix \(R_{\tau \tau '}\)). We proved in Sect. 4.4 the following analogue of sub-additivity: \(F^{(N,L)}(R, R) \le F^{(N)}(R) + F^{(L)}(R)\), where \(F^{(N,L)}(R, R)\) is the free energy of states in a size-\((N+L)\) system restricted to separately have \(\sigma \cdot \sigma = R\) and \(\alpha \cdot \alpha = R\) — recall that we divided the spins into \(\{ \vec {\sigma }_i \}_{i=1}^N\) and \(\{ \vec {\alpha }_j \}_{j=1}^L\), and defined

$$\begin{aligned} \sigma (\tau ) \cdot \sigma (\tau ') \equiv \frac{1}{N} \sum _i \sigma _i(\tau ) \sigma _i(\tau '), \qquad \alpha (\tau ) \cdot \alpha (\tau ') \equiv \frac{1}{L} \sum _j \alpha _j(\tau ) \alpha _j(\tau '). \end{aligned}$$
(114)

In fact, sub-additivity of the sequence \(F^{(N)}(R)\) follows from this result simply by observing that the set of states with \(\sigma \cdot \sigma = \alpha \cdot \alpha = R\) is a subset of the states with total self-overlap R (note that the total self-overlap can be written \((N \sigma \cdot \sigma + L \alpha \cdot \alpha )/(N+L)\)). Thus \(Z^{(N+L)}(R) \ge Z^{(N,L)}(R, R)\) since the sum that is \(Z^{(N+L)}(R)\) includes every term of \(Z^{(N,L)}(R, R)\), and \(F^{(N+L)}(R) \le F^{(N,L)}(R, R)\). Fekete’s lemma then proves that \(f(R) \equiv \lim _{N \rightarrow \infty } F^{(N)}(R)/N\) exists, and a straightforward generalization of Eqs. (111) through (113) gives

$$\begin{aligned} \begin{aligned}&\limsup _{N \rightarrow \infty } \limsup _{L \rightarrow \infty } \frac{F^{(N,L)}(R, R) - F^{(L)}(R)}{N} \le f(R) \le \liminf _{N \rightarrow \infty } \limsup _{L \rightarrow \infty } \frac{F^{(N+L)}(R) - F^{(L)}(R)}{N} \\&\quad \le \liminf _{N \rightarrow \infty } \limsup _{L \rightarrow \infty } \frac{F^{(N,L)}(R, R) - F^{(L)}(R)}{N}. \end{aligned} \end{aligned}$$
(115)

Thus

$$\begin{aligned} f(R) = \lim _{N \rightarrow \infty } \limsup _{L \rightarrow \infty } \frac{F^{(N,L)}(R, R) - F^{(L)}(R)}{N}, \end{aligned}$$
(116)

which is Eq. (84) from the main text.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Baldwin, C.L., Swingle, B. Revisiting the Replica Trick: Competition Between Spin Glass and Conventional Order. J Stat Phys 190, 125 (2023). https://doi.org/10.1007/s10955-023-03135-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-023-03135-1

Keywords

Navigation