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

thanks: Email: zhouhj@itp.ac.cnthanks: Email: qinyi.liao.phy@gmail.com

Circumventing spin glass traps by microcanonical spontaneous symmetry breaking

Hai-Jun Zhou1,2,3,4    Qinyi Liao1 1CAS Key Laboratory for Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China 2School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China 3MinJiang Collaborative Center for Theoretical Physics, MinJiang University, Fuzhou 350108, China 4Synergetic Innovation Center for Quantum Effects and Applications, Hunan Normal University, Changsha 410081, China
(August 2, 2025)
Abstract

The planted pp-spin interaction model is a paradigm of random-graph systems possessing both a ferromagnetic phase and a disordered phase with the latter splitting into many spin glass states at low temperatures. Conventional simulated annealing dynamics is easily blocked by these low-energy spin glass states. Here we demonstrate that, actually this planted system is exponentially dominated by a microcanonical polarized phase at intermediate energy densities. There is a discontinuous microcanonical spontaneous symmetry breaking transition from the paramagnetic phase to the microcanonical polarized phase. This transition can serve as a mechanism to avoid all the spin glass traps, and it is accelerated by the restart strategy of microcanonical random walk. We also propose an unsupervised learning problem on microcanonically sampled configurations for inferring the planted ground state.

I Introduction

The planted pp-spin interaction model on a finite-connectivity random graph is a representative ferromagnetic system with many coexisting spin glass states. This model system has played an important role in understanding the physics of structural glasses Franz et al. (2001); Montanari and Ricci-Tersenghi (2004). This system is also quite relevant and challenging to the field of statistical inference Zdeborová and Krzakala (2016); Mézard and Montanari (2009). It is equivalent to the generalized Sourlas code problem in information science Sourlas (1989); Huang and Zhou (2009) and the planted maximum XORSAT problem in computer science Watanabe (2013). The large number of low-energy disordered spin glass configurations pose a huge computational challenge to retrieve the planted ground state.

When the environmental temperature slowly decreases, as occurs in a simulated annealing (SA) dynamics Kirkpatrick et al. (1983), the system is predicted to experience an equilibrium phase transition from the disordered paramagnetic phase to the ordered ferromagnetic phase. But this equilibrium transition is extremely difficult to occur in practice when the interactions are many-body in nature (p3p\!\geq\!3), as the free-energy barrier between these two phases is proportional to the system size NN and the crystal-nucleation mechanism then fails. The system instead remains in the paramagnetic phase as temperature further decreases Franz et al. (2001); Montanari and Ricci-Tersenghi (2004).

In typical SA simulations with a small but finite rate of temperature decreasing, the system will decrease its energy with an exceedingly smaller rate, and finally it will reach one of the many spin glass states at the bottom of the energy landscape (Fig. 1) and be trapped there for an exponentially long time (in NN). Besides, we have checked that message-passing algorithms based on belief propagation also fail to retrieve the ferromagnetic ground state. Following the proposal of Ref. Xu et al. (2018), we clamped the objective energy to a value much below those of the spin glass phase to search for ferromagnetic states, but we found that the message-passing iteration process does not converge to the ferromagnetic fixed point. This is because the high gauge symmetry of the planted system makes it statistically indistinguishable from the unplanted purely disordered system Matsuda et al. (2011); Bandeira et al. (2018).

This tremendous computational difficulty is not specific only to the planted pp-spin system but is a common property of many large inference problems such as the planted satisfiability and coloring problems Li et al. (2009); Krzakala et al. (2014); Krzakala and Zdeborova (2009). When the ground state is completely masked by an exponential number of disordered spin glass configurations (for example through quiet planting as in the pp-spin system and the coloring problem Krzakala and Zdeborova (2009)), it becomes evident that searching by SA and other dynamical processes at the bottom of the energy landscape and at low temperatures is not a promising strategy.

Refer to caption
Figure 1: Schematic energy landscape of the planted pp-spin interaction system of size NN (p3p\geq 3). The white region contains all the 2N2^{N} microscopic configurations, the “ocean of configurations”. An exponential number of deep local minima (the spin glass states) exist at the bottom of the energy landscape, which may trap the simulated annealing dynamics. We propose to take advantage of the discontinuous microscopic spontaneous symmetry breaking that occurs at high energies to directly jump to the basin of attraction of the ground state without encountering spin glass configurations.

In this work, we demonstrate the benefit of working at sufficiently high energy values. Especially we point out the existence of an alternative route to the planted ground state, the route of microcanonical spontaneous symmetry breaking (MSSB). This is a discontinuous transition from the paramagnetic (or disordered symmetric, DS) phase to the microcanonical polarized (MP) phase. Because the energy density of the system along the whole DS-to-MP transition trajectory is fixed at a high value much above those of the spin glass states, the difficulty of climbing huge energy barriers between the different spin glass states is completely avoided (Fig. 1). The microcanonically stable MP phase has so far been largely overlooked in the literature (except for a recent detailed analysis concerning the Potts model Zhou (2019a)), but we find that it actually is exponentially more dominant than the paramagnetic phase at intermediate energy density values.

The DS and MP phases are connected by many transition trajectories at energy densities u(ud,umic)u\!\in\!(u_{\rm d},u_{\rm mic}), where udu_{\rm d} is the spin glass ergodicity breaking point and umicu_{\rm mic} is the critical energy density of the equilibrium MSSB phase transition. As a main result of this work, we find that umic>udu_{\rm mic}\!>\!u_{\rm d}. Since ergodicity is preserved for u>udu\!>\!u_{\rm d} and the MP phase is entropically favored than the DS phase for u<umicu\!<\!u_{\rm mic}, starting from an initial configuration residing in the DS phase, the system then is destined to arrive at the MP phase and then reach equilibrium within this phase. The MP phase is stable and meaningful only in the microcanonical ensemble of fixed energy density. Equilibrium MP configurations contain extensive information about the planted ground state as they are already quite similar to it. The ferromagnetic ground state is easily reachable from this intermediate microcanonical phase by gradually decreasing the energy of the system, or it may be inferred based on the statistical properties of the MP configurations. The MSSB route therefore serves as a conceptually straightforward mechanism to circumvent all the low-energy spin glass traps.

The MSSB transition between the DS and MP phases is a discontinuous process. This indicates that the DS-to-MP transition should be an extremely rare event in microcanonical simulations. We find that this process can be accelerated by the restart strategy of microcanonical random walk. We also compare the efficiency of this microcanonical fixed-energy route with that of the conventional canonical fixed-temperature Monte Carlo. We prove that if the two routes start from the same initial state (point A of Fig. 4), the canonical route is more efficient than the microcanonical one; while if they pass through the same intermediate state (point B of Fig. 4), the microcanonical route is more efficient than the canonical one. As a byproduct concerning the conventional SA algorithms, our work indicates that as the mean energy density of the sampled configurations is approaching the critical value udu_{\rm d} of spin glass phase transition, the inverse temperature β\beta of the SA process should stop to be further elevated.

We also suggest the potential possibility of solving the planted pp-spin model by machine-learning methods. We sample a large number of independent microcanonical configurations of the DS phase for the the most studied systems of p=3p\!=\!3 (each interaction involving three vertices).Very interestingly, we find that these DS configurations contain information about the unique planted ground state. It may then be possible to infer the planted solution from the sampled DS configurations through machine-learning techniques. We thus construct an unsupervised learning problem of inferring the planted ground state from the data generated by microcanonical samplings. In the future, we will try to solve this inference problem using machine-learning algorithms, and will also extend this work to other hard planted systems.

II Theoretical predictions

Consider a planted pp-spin interaction system in which each vertex j{1,,N}j\!\in\!\{1,\ldots,N\} participates in KK clauses (interactions) and each clause involves pp randomly chosen vertices. For simplicity and to be concrete, we set p=3p\!=\!3 in all the following numerical computations. The total number of clauses is M=KN/pM\!=\!KN/p. The energy of a generic configuration σ(σ1,,σN)\vec{\sigma}\!\equiv\!(\sigma_{1},\ldots,\sigma_{N}) is

E(σ)=a=1M[Jajaσj],E(\vec{\sigma})=\sum\limits_{a=1}^{M}\Bigl{[}-J_{a}\prod\limits_{j\in\partial a}\sigma_{j}\Bigr{]}\;, (1)

where σj±1\sigma_{j}\!\in\!\pm 1 is the spin of vertex jj and a\partial a denotes the set of vertices constrained by clause aa. We denote by uu the energy density of the system, namely

u=E(σ)N.u=\frac{E(\vec{\sigma})}{N}\;. (2)

There is a planted spin configuration σ0(σ10,,σN0)\vec{\sigma}^{0}\!\equiv\!(\sigma_{1}^{0},\ldots,\sigma_{N}^{0}) dictating the coupling constant of clause aa, such that

Ja={+jaσj0(probability 1ε),jaσj0(probabilityε).J_{a}=\left\{\begin{array}[]{ll}+\prod\limits_{j\in\partial a}\sigma_{j}^{0}&\quad\quad(\textrm{probability}\ 1-\varepsilon)\;,\\ &\\ -\prod\limits_{j\in\partial a}\sigma_{j}^{0}&\quad\quad(\textrm{probability}\ \varepsilon)\;.\end{array}\right. (3)

The parameter ε\varepsilon is the noise level of the above planting rule. When ε>0\varepsilon\!>\!0 model (1) can then be interpreted as the Sourlas code of pp-body interactions Sourlas (1989); Huang and Zhou (2009). When ε\varepsilon is sufficiently small the planted configuration σ0\vec{\sigma}^{0} is a ground state of Eq. (1) and it may lie extensively below all the other minimum-energy configurations (Fig. 1). We define the overlap (magnetization) mm of configuration σ\vec{\sigma} with respect to σ0\vec{\sigma}^{0} as

m=1Nj=1Nσjσj0.m=\frac{1}{N}\sum\limits_{j=1}^{N}\sigma_{j}\sigma_{j}^{0}\;. (4)

We introduce an inverse temperature β\beta as the conjugate of the energy E(σ)E(\vec{\sigma}). The partition function ZZ of the system is then defined as

Z=σeβE(σ)=σa=1Mexp(βJajaσj).Z=\sum\limits_{\vec{\sigma}}e^{-\beta E(\vec{\sigma})}=\sum\limits_{\vec{\sigma}}\prod\limits_{a=1}^{M}\exp\Bigl{(}\beta J_{a}\prod\limits_{j\in\partial a}\sigma_{j}\Bigr{)}\;. (5)

The equilibrium Boltzmann distribution of observing a particular configuration σ\vec{\sigma} is then

Peq(σ)=1Z(β)eβE(σ).P_{\textrm{eq}}(\vec{\sigma})=\frac{1}{Z(\beta)}e^{-\beta E(\vec{\sigma})}\;. (6)

In the canonical statistical ensemble the system exchanges energy with the environment. Then β\beta is the inverse temperature of the environment, and the mean energy density uu of the system is determined by β\beta. The free energy density ff of the system is defined as

f=1NlnZβ.f=-\frac{1}{N}\frac{\ln Z}{\beta}\;. (7)

And the entropy density ss is computed as

s=β(uf).s=\beta\bigl{(}u-f\bigr{)}\;. (8)
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2: The DS (solid line), MP (dotted line), and CP (long dashed line) fixed points of the mean-field theory for the planted 33-body model on random graphs of degree K=10K\!=\!10 at noise ε=0\varepsilon\!=\!0. (a) Free energy density ff versus canonical inverse temperature β\beta. (b) Entropy density ss versus energy density uu. (c) Microcanonical inverse temperature β\beta versus uu. (d) Mean overlap mm versus uu. βd\beta_{\rm d} and udu_{\rm d}: the critical inverse temperature and energy density at the dynamical SG phase transition; umicu_{\rm mic}: the critical energy density at the MSSB phase transition; βf\beta_{\rm f}: the critical inverse temperature at the canonical ferromagnetic phase transition. The predicted MSSB phase transition is confirmed by MMC simulations (circles) on a single problem instance of size N=960N\!=\!960.

In the microcanonical statistical ensemble, the system is isolated from the environment and its energy density uu is a fixed value, then β\beta is interpreted as the microcanonical (intrinsic) inverse temperature of the system. With respect to the Boltzmann distribution (6), β\beta in the microcanonical statistical ensemble serves as a Lagrange multiplier such that the mean energy of the sampled configurations according to Eq. (6) is equal to a prescribed value uu. In other words, β\beta is determined by uu in the microcanonical statistical ensemble.

This random-graph system has been intensively studied by the mean-field cavity method of statistical mechanics Mézard and Parisi (2001); Mézard and Montanari (2006); Matsuda et al. (2011); Zhou (2015). For the ensemble of regular random graph of vertex degree KK, the mean-field equations for the mean overlap mm, the mean energy density uu, the free energy density ff, and the entropy density ss are easy to solve. The detailed equations are listed in Appendix A.

The mean-field results obtained at ε=0\varepsilon\!=\!0 and K=10K\!=\!10 are demonstrated in Fig. 2, and the qualitatively identical results obtained for K=4K\!=\!4 are shown in Fig. 7 of Appendix A. Some aspects of these theoretical results are similar to what were found for the Potts model Zhou (2019a), but there is a key qualitative difference to be discussed at the end of this section. First, let us interpret these theoretical predictions from the perspective of microcanonical thermodynamics Gross (2001).

The configuration space has a disordered symmetric phase whose mean magnetization is zero in the thermodynamic limit. This DS phase contains the paramagnetic spin configurations and it corresponds to the paramagnetic phase of the canonical ensemble. As the energy density uu decreases to the spin glass (SG) dynamical transition point udu_{\rm d} this DS phase suddenly divides into an exponential number of ergodicity-broken macro-states, and the system then enters into the SG phase Franz et al. (2001). The value of udu_{\rm d} is independent of ε\varepsilon because of gauge symmetry Matsuda et al. (2011) and can be determined by the tree-reconstruction method Mézard and Montanari (2006) (see also textbooks Mézard and Montanari (2009); Zhou (2015)).

Because of planting, a stable canonical polarized (CP) phase containing configurations similar to σ0\vec{\sigma}^{0} exist in the configuration space at sufficiently low energies. This CP phase is simply the conventional ferromagnetic phase and its overlap mm is close to unity.

The microcanonical polarized phase, corresponding to the unstable fixed point of the mean field theory with higher free energy density [Fig. 2], serves as the watershed between the DS and CP phases [Fig. 2]. Individual configurations of the MP phase have positive overlaps mm and therefore are similar to σ0\vec{\sigma}^{0} and to configurations of the CP phase, but they are unstable in the canonical ensemble due to entropic effect.

The entropy density s(u)s(u) of the MP phase as a function of energy density uu has two branches [Fig. 2]. The higher-entropy branch corresponds to the MP configurations that are stable at fixed uu, and its s(u)s(u) is convex. The MP phase starts to surpass the DS phase in entropy density at the critical energy density umicu_{\rm mic}, leading to a kink of ss and indicating a discontinuous microcanonical spontaneous symmetry-breaking phase transition Zhou (2019a). This phase transition is characterized by a jump in the mean overlap mm and a drop in the microcanonical inverse temperature β\beta, which are verified by our simulation results [Fig. 2 and 2]. The lower-entropy MP branch, on the other hand, always has lower entropy than that of the DS phase and its s(u)s(u) is concave; it marks the watershed between the MP and DS phases in the microcanonical ensemble.

We observe that the MSSB transition energy density umicu_{\rm mic} is located above the SG transition point udu_{\rm d} when the noise level ε\varepsilon is low (Table 1), for all values of degree K4K\!\geq\!4. This means that there is an energy density range u(ud,umic)u\in(u_{\rm d},u_{\rm mic}) within which the MP configurations are exponentially dominating over the DS configurations. In an infinitely slow microcanonical annealing process, the SG phase will then not be encountered, instead the system will jump from the DS phase to the MP phase at the higher energy density point umicu_{\rm mic}. In principle, it is then possible to reach the MP phase by inducing a MSSB transition at fixed value of u(ud,umic)u\in\!(u_{\rm d},u_{\rm mic}), avoiding the spin glass traps of the canonical ensemble. We will continue to discuss the practical feasibility of this proposal in the next section.

We find that the DS fixed point of the pp-spin model is always locally stable, down to the minimal energy density at which the entropy density becomes zero [Fig. 2]. Consequently, a gap of the overlap mm always exists between the DS and MP phases down to this minimal energy density [Fig. 2]. This feature is significantly different from what was observed in the Potts model (see Fig. 2(d) of Ref. Zhou (2019a)). The DS phase of the Potts model becomes unstable below certain threshold energy density, and by fixing the energy density below this threshold value, the system will then evolve from the DS phase to the MP phase gradually. Such a gradual evolution will not be possible for the pp-spin model because of the positive gap in the order parameter mm.

Table 1: Overlap jump Δm\Delta m, microcanonical inverse temperature drop Δβ\Delta\beta, critical energy density umicu_{\rm mic}, at the MSSB phase transition of the planted 33-body model on random graphs of degree KK (noise ε=0\varepsilon\!=\!0). The critical energy densities udu_{\rm d} are also listed (based on Table 4.2 of Zhou (2015)).
KK 44 55 66 77 88 99 1010
Δm\Delta m 0.9060.906 0.8210.821 0.7580.758 0.7120.712 0.6760.676 0.6470.647 0.6230.623
Δβ\Delta\beta 0.595-0.595 0.409-0.409 0.324-0.324 0.274-0.274 0.239-0.239 0.214-0.214 0.193-0.193
umicu_{\rm mic} 1.115-1.115 1.167-1.167 1.204-1.204 1.235-1.235 1.264-1.264 1.289-1.289 1.313-1.313
udu_{\rm d} 1.159-1.159 1.316-1.316 1.453-1.453 1.575-1.575 1.687-1.687 1.792-1.792 1.892-1.892

We notice that the special case of degree K=3K\!=\!3 is qualitatively different, for which the DS phase always dominates over the MP and CP phases at all the energy density values (see Fig. 8 of Appendix A). An equilibrium MSSB phase transition is then absent for this special system.

III Numerical simulation results

In this section, we discuss the numerical results obtained by two different microcanonical Monte Carlo (MMC) algorithms, and we also compare the efficiency of the MMC dynamics with that of the conventional canonical Monte Carlo (CMC) dynamics.

III.1 The single-flip MMC algorithm

We first implement a simple MMC algorithm to explore configurations at the vicinity of an objective energy EoNuE_{\rm o}\!\equiv\!Nu, whose energy density u(ud,umic)u\!\in\!(u_{\rm d},u_{\rm mic}). An initial configuration σ\vec{\sigma} with energy E(σ)EoE(\vec{\sigma})\!\leq\!E_{\rm o} can be sampled according to the equilibrium Boltzmann distribution (6) at a suitably chosen inverse temperature β\beta. Starting from this initial configuration, the following single-flip trial is performed at each elementary time interval 1/N1/N of the MMC evolution: (1) a vertex jj is picked uniformly at random from the NN vertices and the flip σjσj\sigma_{j}\!\rightarrow\!-\sigma_{j} is proposed; (2) if the new configuration energy does not exceed EoE_{\rm o}, this flip is accepted, otherwise it is rejected and the old spin σj\sigma_{j} is kept for vertex jj. One sweep (unit time) of this MMC process corresponds to NN consecutive single-flip trials.

This single-flip dynamics obeys detailed balance since the forward and backward transitions between two configurations are equally likely Creutz (1983); Rose and Machta (2019). This microcanonical detailed balance condition ensures that all the visited configurations along the evolution trajectory share the same statistical weight. We sample configurations at fixed time intervals and record their overlap values mm and energies EE. The microcanonical inverse temperature of the system at energy density uu is estimated through

β=14ln(1+4EoE(σ)),\beta=\frac{1}{4}\ln\Bigl{(}1+\frac{4}{E_{\rm o}-\bigl{\langle}E(\vec{\sigma})\bigr{\rangle}}\Bigr{)}\;, (9)

where E(σ)\bigl{\langle}E(\vec{\sigma})\bigr{\rangle} is the averaged energy of the sampled configurations Creutz (1983).

Our MMC dynamics is an unbiased random walk within the microcanonical configuration space. The evolution trajectory is expected to be ergodic since the energy density uu is above the dynamical spin glass transition point udu_{\rm d}. As evolution time goes to infinity every configuration of this space should have equal frequency to be visited, and the system will then certainly be in the MP phase since this phase is exponentially dominant in statistical weight. Empirically, we have observed that this MMC dynamics does indeed achieve MSSB in small-sized systems [Fig. 2 and Fig. 3], while the evolution time needed to witness such a transition increases quickly with system size NN [Fig. 3]. It is easier for the MSSB transition to occur in networks of larger degrees KK. This is consistent with the mean-field theoretical results, which reveal that the entropy barrier and the gap of the order parameter mm both decreases with KK (Table 1). The hardest RR problem instances are those with degree K=4K\!=\!4.

The MSSB transition is an entropy-barrier crossing process [Fig. 3, and the route ADEA\!\rightarrow\!D\!\rightarrow\!E in Fig. 4]. We can regard the MMC dynamics on the coarse-grained level of overlap mm as a one-dimensional random walk under a potential field su(m)-s_{u}(m), where su(m)s_{u}(m) is the entropy density of configurations at fixed energy density uu having overlap mm with the planted configuration σ0\vec{\sigma}^{0}. At each elementary trial, mm may change to m±2/Nm\pm 2/N with probability proportional to (1gm)/2(1\mp g_{m})/2. Here the bias ratio is

gm=1e2su(m)1+e2su(m),g_{m}=\frac{1-e^{2s_{u}^{\prime}(m)}}{1+e^{2s_{u}^{\prime}(m)}}\;, (10)

with su(m)dsu(m)/dms_{u}^{\prime}(m)\!\equiv\!{\rm d}s_{u}(m)/{\rm d}m being the slope of su(m)s_{u}(m) at mm. Because s(m)<0s^{\prime}(m)<0 in m(0,m)m\!\in\!(0,m^{*}) with mm^{*} being the watershed point of su(m)s_{u}(m) [Fig. 3, and point DD of Fig. 4], the occurrence of MSSB is an exponentially rare first-passage event characterized by a waiting time of exponential order O(eNΔs)O(e^{N\Delta s}), where Δs=su(0)su(m)\Delta s\!=\!s_{u}(0)-s_{u}(m^{*}) is the barrier height Weiss (1981); Khantha and Balakrishnan (1983). Consequently, this simple random walk strategy is practical only for systems of sizes

N<20Δs.N<\frac{20}{\Delta s}\;. (11)
Refer to caption
Refer to caption
Figure 3: An entropy barrier has to be crossed to achieve a MSSB transition. (a) A schematic curve of the entropy density su(m)s_{u}(m) of the planted 33-body model at fixed energy density uu and noise ε=0\varepsilon\!=\!0, showing a local maximum at overlap m00m_{0}\!\approx\!0, a global maximum at m1m_{1}, and a minimum at mm^{*}. The MMC dynamics is equivalent to a one-dimensional biased random walk with bias ratio gmg_{m}. The example evolution trajectory shows a MSSB event at u=1.61u\!=\!-1.61 in a random graph of degree K=10K\!=\!10 and size N=960N\!=\!960 (one unit of MMC time tt means NN single-flip trials). (b) The minimal number of sweeps tt needed for a MSSB transition to occur, with each point being the length of the shortest trajectory among 2424 independent single-flip MMC runs on a single network: K=4K\!=\!4, u=1.133u\!=\!-1.133 (circle); K=8K\!=\!8, u=1.6u\!=\!-1.6 (squares); K=10K\!=\!10, u=1.61u\!=\!-1.61 (pluses); K=20K\!=\!20, u=2.2u\!=\!-2.2 (crosses). Curves are power-law (Nb\propto\!N^{b}) relations with b=7,9,12b\!=\!7,9,12 from bottom to top.

The entropy barrier Δs\Delta s decreases with energy density uu [Fig. 4]. Therefore, we can decrease the entropy barrier Δs\Delta s by fixing uu of the MMC dynamics to a lower value. This will help reducing the waiting time of the MSSB transition. On the other hand, at lower energy density values the number of transition paths from point AA to point DD of Fig. 4 becomes much more rarer and then disappears at uudu\!\approx\!u_{\rm d}. This later kinetic effect will lead to an increase in the waiting time. There should be an optimal energy density uu, at which the MSSB transition is easiest to observe. We leave this interesting issue to future studies.

Refer to caption
Refer to caption
Refer to caption
Figure 4: Two different routes to escape the disordered symmetric phase. (a) Schematic plot showing the canonical route (ABCA\!\rightarrow\!B\!\rightarrow\!C) at fixed inverse temperature β\beta and the microcanonical route (ADEA\!\rightarrow\!D\!\rightarrow\!E) at fixed energy density uu. The black solid line represents the DS phase (containing points AA and AA^{\prime}); the blue long-dashed line is the CP phase (containing point CC); the red dotted line is the MP phase (containing points B,D,EB,D,E, and EE^{\prime}). (b) The rescaled free energy barrier βΔf\beta\Delta f of the canonical route (solid line) and the entropy barrier Δs\Delta s of the microcanonical route (dashed line), as a function of the energy density uu of the initial configuration. These are theoretical results obtained at K=10K\!=\!10 and noise ε=0\varepsilon\!=\!0. The vertical dotted lines mark udu_{\rm d} (left) and umicu_{\rm mic} (right). (c) Cumulative distribution C(W)C(W) on the total number of spin flips WW performed to escape the DS phase, obtained by simulating 40004000 independent hybrid-flip CMC evolution trajectories at β=0.4847\beta\!=\!0.4847 (route ABCA\!\rightarrow\!B\!\rightarrow\!C, solid line), 40004000 independent hybrid-flip MMC evolution trajectories at u=1.50u\!=\!-1.50 (route ADEA\!\rightarrow\!D\!\rightarrow\!E, dashed line) and another 40004000 independent trajectories at u=1.543u\!=\!-1.543 (route ABEA^{\prime}\!\rightarrow\!B\!\rightarrow\!E^{\prime}, dotted line) on a single graph instance of size N=1008N\!=\!1008 and degree K=10K\!=\!10 (noise ε=0\varepsilon\!=\!0).

As a minor point, we notice that the watershed entropy density su(m)s_{u}(m^{*}) of Fig. 3 decreases as the energy density uu decreases and it will reach zero at certain critical energy density value (denoted as umebu_{\rm meb}) which marks the point of strong microcanonical ergodicity-breaking. When u<umebu\!<\!u_{\rm meb} there will be an extended region of the intermediate overlap values mm within which there is no any configuration of energy density uu. In other words, it is impossible to reach from a disordered initial configuration to a highly ordered final configuration through a sequence of small changes of the configuration at fixed energy density u<umebu\!<\!u_{\rm meb}. We find that umedu_{\rm med} is below udu_{\rm d} for all the degrees K3K\!\geq\!3 (see Figs. 2, 7 and 8), so this strong microcanonical ergodicity-breaking is not really relevant for our MSSB mechanism which occurs at u>udu\!>\!u_{\rm d}.

III.2 The hybrid-flip MMC and CMC algorithms

Besides the microcanonical route ADEA\!\rightarrow\!D\!\rightarrow\!E as depicted in Fig. 4, there is also the conventional canonical route (ABCA\!\rightarrow\!B\!\rightarrow\!C) to escape the disordered symmetric phase. Will this canonical route be blocked by the low-energy disordered spin glass states? If the planted ground state could be successfully approached following either the microcanonical route or the canonical route, which of these two routes will be more efficient? We explore these two questions in this and the next subsections.

To be fair in comparing the MMC and CMC routes, we design a hybrid-flip Monte Carlo dynamics which is rejection-free both at fixed energy density uu (MMC) and at fixed inverse temperature β\beta (CMC). Besides the action of flipping a single spin, the hybrid-flip dynamics differs from the single-flip dynamics by allowing the action of flipping a pair of spins in one elementary updating step.

An elementary step of the hybrid-flip MMC dynamics goes as follows: (1) a vertex jj is picked uniformly at random from the NN vertices and its spin σj\sigma_{j} is flipped, causing a change ΔE\Delta E to the energy; (2) if and only if the energy of the new configuration exceeds the objective value EoE_{\rm o}, an additional flip is performed on the spin σk\sigma_{k} of a vertex kk to bring the energy back to its original value (kk is picked uniformly at random from all the vertices capable of causing a change ΔE-\Delta E to the energy). This microcanonical hybrid-flip rule obeys microcanonical detailed balance, so all the visited configurations on the MMC evolution trajectory have the same statistical weight.

An elementary step of the hybrid-flip CMC dynamics is quite similar: (1) a vertex jj is picked uniformly at random from the NN vertices and its spin σj\sigma_{j} is flipped, causing a change ΔE\Delta E to the energy; (2) if ΔE\Delta E is positive, then with probability 1eβΔE1\!-\!e^{-\beta\Delta E} an additional flip is performed on the spin σk\sigma_{k} of a vertex kk to bring the energy back to its original value (kk is picked uniformly at random from all the vertices capable of causing a change ΔE-\Delta E to the energy). This canonical hybrid-flip rule obeys canonical detailed balance and the visited configurations on the CMC evolution trajectory are governed by the Boltzmann distribution (6).

We carry out hybrid-flip MMC and hybrid-flip CMC simulations on a single random graph instance of degree K=10K\!=\!10 and size N=1008N\!=\!1008. A total number of 40004000 independent MMC evolution trajectories and 40004000 independent CMC evolution trajectories are sampled with the help of the TRNG pseudo-random number library Bauke and Mertens (2007), all of these trajectories start from the same initial equilibrium configuration which is located in the DS phase (inverse temperature β=0.4847\beta\!=\!0.4847 and energy density u=1.50u\!=\!-1.50). We regard the DS phase as successfully escaped when the overlap mm with the planted configuration exceeds 0.60.6 for the first time, and the evolution is then terminated and the total number WW of spin flips along the whole evolution trajectory is returned. We refer to WW as the effort to escape DS. Figure 4 shows the cumulative distribution profile C(W)C(W) of the effort WW over the 40004000 samples, obtained at fixed inverse temperature β\beta (CMC) or fixed energy density uu (MMC).

We find that the effort (number of spin flips) WW fluctuates over a wide range from 10610^{6} up to 101210^{12} among repeated runs (all of which start from the same initial configuration). This indicates that restart might be an optimal strategy if the evolution trajectory fails to escape the DS phase within a certain threshold value of WW. The cumulative distribution function C(W)C(W) is concave in shape, both for the MMC dynamics and the CMC dynamics. For example, within an effort W16.071×109W_{1}\!\approx\!6.071\times 10^{9} the success probability of escaping the DS phase is C(W1)=20%C(W_{1})\!=\!20\% by the CMC dynamics, while this success probability only increases to C(W2)=35.6%C(W_{2})\!=\!35.6\% when the effort is doubled to W2=1.214×1010W_{2}\!=\!1.214\times 10^{10} [Fig. 4]. This concave property of C(W)C(W) guarantees that restart is an optimal strategy. For example, we could restart the CMC or the MMC evolution from the fixed initial configuration if the effort of the current trajectory exceeds W1W_{1}, and the resulting cumulative distribution of the effort will change to be approximately linear.

III.3 Entropy barrier and free energy barrier

Our CMC simulation results clearly demonstrate that the low-energy spin glass states are not encountered by the canonical route ABCA\!\rightarrow\!B\!\rightarrow\!C. This may be easy to understand: the canonical route occurs at a much lower inverse temperature than the critical value βd\beta_{\rm d} where the spin glass dynamical phase transition occurs, and the energy density of the watershed point BB separating the DS and the CP phases is considerably higher than the critical value udu_{\rm d} of spin glass dynamical phase transition.

Figure 4 demonstrates that, when starting from the same initial configuration, the CMC dynamics is more efficient than the MMC dynamics. For example, the MMC dynamics needs a larger effort W17.093×109W_{1}\!\approx\!7.093\times 10^{9} to achieve a success probability C(W1)=20%C(W_{1})\!=\!20\%. The empirical observation of CMC being more efficient than MMC is consistent with the theoretical predictions on the entropy-barrier height Δs\Delta s and the rescaled free energy barrier height βΔf\beta\Delta f [Fig. 4]. We could prove that Δs>βΔf\Delta s>\beta\Delta f as follows.

The rescaled free energy barrier (βΔf\beta\Delta f) of the canonical route ABCA\!\rightarrow\!B\!\rightarrow\!C is

βΔf=βA(fBfA).\beta\Delta f=\beta_{A}(f_{B}-f_{A})\;. (12)

Here βX\beta_{X}, uXu_{X}, fXf_{X}, and sXs_{X} denote the inverse temperature, energy density, free energy density, and entropy density at the phase point X{A,B,}X\in\{A,B,\ldots\} of Fig. 4. The entropy barrier (Δs\Delta s) of the microcanonical route ADEA\!\rightarrow\!D\!\rightarrow\!E) is

Δs=sAsD.\Delta s=s_{A}-s_{D}\;. (13)

Becuase of the relationship (8) and the fact that βB=βA\beta_{B}=\beta_{A} and uD=uAu_{D}=u_{A}, we know that

Δs\displaystyle\Delta s =\displaystyle= βΔf+[sBsD+βB(uDuB)]\displaystyle\beta\Delta f+\bigl{[}s_{B}-s_{D}+\beta_{B}(u_{D}-u_{B})\bigr{]} (14)
\displaystyle\approx βΔf+12(βBβD)(uDuB)\displaystyle\beta\Delta f+\frac{1}{2}(\beta_{B}-\beta_{D})(u_{D}-u_{B})
>\displaystyle> βΔf.\displaystyle\beta\Delta f\;.

In deriving this inequality, we have used the following approximation

sD\displaystyle\hskip-7.11317pts_{D} \displaystyle\approx sB+dsdu|uB(uDuB)+12d2sdu2|uB(uDuB)2\displaystyle s_{B}+\left.\frac{{\rm d}s}{{\rm d}u}\right|_{u_{B}}(u_{D}-u_{B})+\frac{1}{2}\left.\frac{{\rm d}^{2}s}{{\rm d}u^{2}}\right|_{u_{B}}(u_{D}-u_{B})^{2} (15)
=\displaystyle= sB+βB(uDuB)+12dβdu|uB(uDuB)2\displaystyle s_{B}+\beta_{B}(u_{D}-u_{B})+\frac{1}{2}\left.\frac{{\rm d}\beta}{{\rm d}u}\right|_{u_{B}}(u_{D}-u_{B})^{2}
\displaystyle\approx sB+βB(uDuB)+12(βDβB)(uDuB),\displaystyle s_{B}+\beta_{B}(u_{D}-u_{B})+\frac{1}{2}(\beta_{D}-\beta_{B})(u_{D}-u_{B})\;,

assuming that uDu_{D} is quite close to uBu_{B}.

On the other hand, if we compare the CMC and MMC routes passing through the same intermediate state (say point BB of Fig. 4), namely the CMC route ABCA\!\rightarrow\!B\!\rightarrow\!C and the MMC route ABEA^{\prime}\!\rightarrow B\!\rightarrow\!E^{\prime} (points A,B,EA^{\prime},B,E^{\prime} have the same energy density), we find that the entropy barrier

Δs=sAsB,\Delta s^{\prime}=s_{A^{\prime}}-s_{B}\;, (16)

is lower than the rescaled free energy barrier βBΔf\beta_{B}\Delta f [Eq. (12)]. To prove this, we notice that

Δs\displaystyle\Delta s^{\prime} =\displaystyle= βBΔf+[sAsA+βA(uAuB)]\displaystyle\beta_{B}\Delta f+\bigl{[}s_{A^{\prime}}-s_{A}+\beta_{A}(u_{A}-u_{B})\bigr{]} (17)
\displaystyle\approx βBΔf12(βAβA)(uAuA)\displaystyle\beta_{B}\Delta f-\frac{1}{2}(\beta_{A^{\prime}}-\beta_{A})(u_{A}-u_{A^{\prime}})
<\displaystyle< βBΔf,\displaystyle\beta_{B}\Delta f\;,

as βA>βA\beta_{A^{\prime}}\!>\!\beta_{A} and uA>uAu_{A}\!>\!u_{A^{\prime}}. This indicates that for all the transition trajectories passing through the same intermediate state BB, the MMC trajectories are more efficient than the CMC trajectories. Our simulation results confirm this prediction, see the dotted line and the solid line of Fig. 4.

III.4 Simulated energy annealing (SEA)

Refer to caption
Refer to caption
Figure 5: Simulated energy annealing on a single regular random graph instance of size N=1008N\!=\!1008 and degree K=10K\!=\!10. The energy density uu decreases in step of 2/N2/N, and τ\tau sweeps are simulated at uu, with τ=5×104\tau\!=\!5\times 10^{4} (pluses), 10510^{5} (diamonds), or 10610^{6} (circles). The microcanonical inverse temperature β\beta (a) and the mean overlap mm (b) are computed using τ\tau configuration samples. The vertical dashed line marks the spin glass transition energy density udu_{\rm d}.

We do not need to fix the energy or the inverse temperature in the Monte Carlo simulations. Instead the energy or the inverse temperature can slowly change with time. Following some previous numerical efforts Alava et al. (2008); Ma and Zhou (2011); Rose and Machta (2019), we implement SEA (simulated energy annealing) as follows: (1) Start from an initial configuration of high energy value EE; (2) perform MMC simulation with at objective energy Eo=EE_{\rm o}=E for a total number τ\tau of sweeps (each sweep corresponds to NN contiguous hybrid-flip updates); (2) decrease the objective energy by a small amount (say, Eo2EoE_{\rm o}\!-\!2\rightarrow E_{\rm o}) and repeat step (2).

Some SEA evolution trajectories are shown in Fig. 5 in the (u,β)(u,\beta) and (u,m)(u,m) phase planes. As the waiting time τ\tau at each energy density uu increases the MSSB transition occurs at higher energy values uu. This is consistent with expectations.

Similar results are also observed in the conventional simulated temperature annealing process, where the canonical inverse temperature β\beta gradually increases (but does not exceed βd\beta_{\rm d}). The energy density uu will change dramatically during this annealing process.

In comparison with the SEA algorithm, the MMC algorithm (with restart) at a suitably chosen fixed energy density uu is more efficient.

III.5 A machine-learning challenge

Since it is exceedingly difficult to escape the DS phase by both the MMC dynamics and the CMC dynamics as the system size NN increases, might it be possible to retrieve the planted ground state without actually leaving the DS phase? Here we report a preliminary exploration of this interesting question.

Given a pp-spin problem instance, it is straightforward to sample a large set of independent equilibrium configurations of the DS phase at a given energy density u(ud,umic)u\!\in\!(u_{\rm d},u_{\rm mic}) by the hybrid-flip MMC algorithm. The overlaps mm of these configurations are of order 1/N1/\sqrt{N}, so a more convenient quantity is the rescaled overlap λmN\lambda\equiv m\sqrt{N}, namely

λ=1Ni=1Nσiσi0.\lambda=\frac{1}{\sqrt{N}}\sum\limits_{i=1}^{N}\sigma_{i}\sigma_{i}^{0}\;. (18)

An empirical probability distribution P(λ)P(\lambda) could be constructed as

P(λ)=1𝒩=1𝒩δ(λ1Ni=1Nσi()σi0),P(\lambda)=\frac{1}{\mathcal{N}}\sum\limits_{\ell=1}^{\mathcal{N}}\delta\biggl{(}\lambda-\frac{1}{\sqrt{N}}\sum\limits_{i=1}^{N}\sigma_{i}^{(\ell)}\sigma_{i}^{0}\biggr{)}\;, (19)

where 𝒩\mathcal{N} denotes the total number of samples, σ()=(σ1(),,σN())\vec{\sigma}^{(\ell)}=(\sigma_{1}^{(\ell)},\ldots,\sigma_{N}^{(\ell)}) is the \ell-th sample, and δ(x)=1\delta(x)\!=\!1 for x=0x\!=\!0 and δ(x)=0\delta(x)\!=\!0 for x0x\!\neq\!0.

For the planted 33-spin systems of degree K4K\!\geq\!4 at zero noise (ε=0\varepsilon\!=\!0), we find that the overlap distribution P(λ)P(\lambda) is not symmetric in λ\lambda, see Fig. 6 for some representative examples. When the magnitude |λ||\lambda| is small the rescaled overlap λ\lambda is more likely to be negative than to be positive, while it is more likely to be positive than to be negative as |λ||\lambda| increases beyond certain moderate value [Fig. 6]. Overall, the rescaled overlap λ\lambda is more likely to be negative than to be positive. Our extensive empirical results suggest that the probabilities of λ\lambda being non-positive (P0P_{\leq 0}) and non-negative (P0P_{\geq 0}) scale with NN as

P0=12+γ2N,P0=12γ2N,P_{\leq 0}=\frac{1}{2}+\frac{\gamma}{2\sqrt{N}}\;,\quad P_{\geq 0}=\frac{1}{2}-\frac{\gamma}{2\sqrt{N}}\;, (20)

where the asymmetry parameter γ\gamma is estimated to be γ1.0\gamma\approx 1.0 through the bootstrap method Efron (1979). (Very surprisingly, our empirical results on the first moment λ\langle\lambda\rangle suggest that it is identical to zero.)

We find that the empirical P(λ)P(\lambda) distribution could be precisely fitted by

P(λ)exp(λ22U(λ)),P(\lambda)\propto\exp\Bigl{(}-\frac{\lambda^{2}}{2}-U(\lambda)\Bigr{)}\;, (21)

with U(λ)U(\lambda) being an odd function of the following form

U(λ)=c1λc3λ3.U(\lambda)=c_{1}\lambda-c_{3}\lambda^{3}\;. (22)

Both the linear-order coefficient c1c_{1} and the third-order coefficient c3c_{3} are positive [Fig. 6]. According to the pioneering theoretical work in the 1990s Van den Broeck and Reimann (1996); Reimann and Van den Broeck (1996), the anti-symmetric “potential energy” U(λ)U(\lambda) may make it possible to infer the planted ground state σ0\vec{\sigma}^{0} by unsupervised learning.

Refer to caption
Refer to caption
Figure 6: The probability distribution function P(λ)P(\lambda) of the rescaled overlap λ\lambda is not symmetric for the planted 33-spin system. The empirical P(λ)P(\lambda) is obtained by sampling 𝒩=4.8×108\mathcal{N}\!=\!4.8\!\times\!10^{8} configurations of energy density u=1.61u\!=\!-1.61 through the hybrid-flip MMC algorithm at interval of 100100 sweeps, for a single random graph instance of degree K=10K\!=\!10 (noise ε=0\varepsilon\!=\!0). The system size is N=3840N\!=\!3840 (crosses) or 76807680 (pluses). (a) The logarithm of P(λ)P(\lambda) can be fitted by ln[P(λ)]=c0λ2/2c1λ+c3λ3\ln[P(\lambda)]\!=c_{0}\!-\!\lambda^{2}/2\!-\!c_{1}\lambda\!+\!c_{3}\lambda^{3} with c1=0.06963c_{1}\!=\!0.06963 and c3=0.02436c_{3}\!=\!0.02436 (solid curve) or with c1=0.04939c_{1}\!=\!0.04939 and c3=0.01766c_{3}\!=\!0.01766 (dashed curve). (b) Difference of empirical probability P(λ)P(\lambda) between positive and negative overlaps of the same magnitude.

We can formulate an unsupervised learning task for the planted pp-spin model as follows: (1) Assume all the observed configurations σ()\vec{\sigma}^{(\ell)} are generated by the following generative model

Prob(σ())exp(U(1Ni=1Nσi()σi0)),{\rm Prob}(\vec{\sigma}^{(\ell)})\propto\exp\Bigl{(}-U\bigl{(}\frac{1}{\sqrt{N}}\sum\limits_{i=1}^{N}\sigma_{i}^{(\ell)}\sigma_{i}^{0}\bigr{)}\Bigr{)}\;, (23)

with NN Ising parameters {σi0}\{\sigma_{i}^{0}\}. (2) Determine the optimal assignment for this set of model parameters so that these 𝒩\mathcal{N} observed configurations will be most likely to be observed Engel and Van den Broeck (2001).

We expect the planted ground state to be the true optimal solution of this unsupervised learning problem. But it is also likely that this problem has many sub-optimal solutions, some of which being the spin glass minimal-energy configurations of Eq. (1). Whether the optimal solution could be efficiently achieved for this machine-learning problem is an open issue and deserves thorough investigations in the future.

Other statistical inference strategies, such as perceptron learning and principal components analysis may also help for the planted pp-spin model Engel and Van den Broeck (2001); Kabashima and Uda (2004); Braunstein and Zecchina (2006); Huang and Toyoizumi (2015); Zhou (2019b); Hu et al. (2019). For example, the asymmetric property (20) of the sampled configurations indicates the existence of an optimal hyperplane separating the 𝒩\mathcal{N} configuration samples in the most asymmetric manner.

IV Summary

In summary, we studied the planted pp-spin interaction model as an example to explore the feasibility of microcanonical spontaneous symmetry breaking for solving hard inference problems. We confirmed the existence of a discontinuous MSSB phase transition in the regular random graph ensemble [Fig. 2]. Our numerical results demonstrated that both the canonical route and microcanonical route [Fig. 4] could circumvent the low-energy spin glass states to reach the planted ground state, and that the restart strategy is superior to the persistent random walk strategy [Fig. 4]. We also pointed out that inferring the planted ground state for this model system could be projected into an unsupervised learning problem [Fig. 6], which may serve as a hard bench-mark problem for various machine-learning algorithms.

The finite-size scaling behavior (20) and the potential function (22) are quite fascinating and they call for a thorough theoretical understanding. Numerical implementation of the machine-learning strategy (23) needs to be done in the future. We are also starting to work on other challenging planted ensembles of optimization problems, such as KK-satisfiability and QQ-coloring Li et al. (2009); Krzakala and Zdeborova (2009); Krzakala et al. (2014), to get more insights on the MSSB mechanism.

Very recently, two preprints on solving the planted random 33-spin model of degree K=3K\!=\!3 were posted Bellitti et al. (2021); Bernaschi et al. (2021). The numerical results of the quasi-greedy algorithm of these two studies are consistent with the physical picture offered in our present work. To be concrete, we include the theoretical results obtained for this special system in Fig. 8 of Appendix A.

Acknowledgements.
One of the authors (H.-J.Z.) thanks Erik Aurell, Yuliang Jin, Federico Ricci-Tersenghi, and Pan Zhang for helpful conversations. This work was supported by the National Natural Science Foundation of China Grants No. 11975295, No. 11947302 and No. 12047503, and the Chinese Academy of Sciences Grant No. QYZDJ-SSW-SYS018. Numerical simulations were carried out at the Tianwen clusters of ITP-CAS and the Tianhe-2 platform of the National Supercomputer Center in Guangzhou.

Appendix A Mean field theoretical equations

We employed the replica-symmetric mean field theory of spin glasses to predict the MSSB phase transition for the planted pp-spin model. To serve as a quick reference, here we list the main theoretical equations of this theory Mézard and Parisi (2001); Mézard and Montanari (2006); Matsuda et al. (2011); Zhou (2015).

First, the marginal probability qiσiq_{i}^{\sigma_{i}} of vertex ii being in spin state σi\sigma_{i} is evaluated through

qiσi=1ziaipaiσi,q_{i}^{\sigma_{i}}=\frac{1}{z_{i}}\prod\limits_{a\in\partial i}p_{a\rightarrow i}^{\sigma_{i}}\;, (24)

where ziz_{i} is a normalization constant to ensure qi+1+qi1=1q_{i}^{+1}\!+\!q_{i}^{-1}\!=\!1, and i\partial i contains all the clauses that are connected to vertex ii, and paiσip_{a\rightarrow i}^{\sigma_{i}} is defined as the probability of vertex ii being in spin state σi\sigma_{i} if it is affected only by clause aa but not by all the other attached clauses. We define a complementary quantity qiaσiq_{i\rightarrow a}^{\sigma_{i}} as the probability of vertex ii being in spin state σi\sigma_{i} if it is affected by all the attached clauses except clause aa. In the literature, both qiaσiq_{i\rightarrow a}^{\sigma_{i}} and paiσip_{a\rightarrow i}^{\sigma_{i}} are referred to as cavity probability distributions.

Let us assume that all the vertices in the set a\partial a will be mutually independent if the energetic effect associated with clause aa is absent, then the following belief-propagation equation for the cavity probabilities could be derived:

paiσi\displaystyle p_{a\rightarrow i}^{\sigma_{i}} =1zai{σj:ja\i}eβEaja\iqjaσj,\displaystyle=\frac{1}{z_{a\rightarrow i}}\sum\limits_{\{\sigma_{j}\mathrel{\mathop{\ordinarycolon}}j\in\partial a\backslash i\}}e^{-\beta E_{a}}\prod\limits_{j\in\partial a\backslash i}q_{j\rightarrow a}^{\sigma_{j}}\;, (25a)
qiaσi\displaystyle q_{i\rightarrow a}^{\sigma_{i}} =1ziabi\apbiσi.\displaystyle=\frac{1}{z_{i\rightarrow a}}\prod\limits_{b\in\partial i\backslash a}p_{b\rightarrow i}^{\sigma_{i}}\;. (25b)

In these expressions, zaiz_{a\rightarrow i} and ziaz_{i\rightarrow a} are two normalization constants, EaE_{a} (Jajaσj\equiv\!-J_{a}\prod_{j\in\partial a}\sigma_{j}) is the energy of clause aa, a\partial a contains all the vertices which are attached to clause aa, a\i\partial a\backslash i denotes the residual vertex set obtained by deleting ii from a\partial a, and similarly i\a\partial i\backslash a denotes the residual clause set obtained by deleting clause aa from i\partial i.

The total free energy F(β)F(\beta) is computed as

F(β)=i=1Nfi+a=1Mfa(i,a)f(i,a),F(\beta)=\sum\limits_{i=1}^{N}f_{i}+\sum\limits_{a=1}^{M}f_{a}-\sum\limits_{(i,a)}f_{(i,a)}\;, (26)

where fif_{i}, faf_{a}, and f(i,a)f_{(i,a)} are, respectively, the free energy contribution of vertex ii, clause aa, and link (i,a)(i,a) between vertex ii and clause aa. The corresponding expressions are

fi\displaystyle f_{i} =1βln[σiaipaiσi],\displaystyle=-\frac{1}{\beta}\ln\Bigl{[}\sum\limits_{\sigma_{i}}\prod\limits_{a\in\partial i}p_{a\rightarrow i}^{\sigma_{i}}\Bigr{]}\;, (27a)
fa\displaystyle f_{a} =1βln[{σj:ja}eβEajaqjaσj],\displaystyle=-\frac{1}{\beta}\ln\Bigl{[}\sum\limits_{\{\sigma_{j}\mathrel{\mathop{\ordinarycolon}}j\in\partial a\}}e^{-\beta E_{a}}\prod\limits_{j\in\partial a}q_{j\rightarrow a}^{\sigma_{j}}\Bigr{]}\;, (27b)
f(i,a)\displaystyle f_{(i,a)} =1βln[σipaiσiqiaσi].\displaystyle=-\frac{1}{\beta}\ln\Bigl{[}\sum\limits_{\sigma_{i}}p_{a\rightarrow i}^{\sigma_{i}}q_{i\rightarrow a}^{\sigma_{i}}\Bigr{]}\;. (27c)

The free energy density is f=F/Nf\!=\!F/N. The mean value of the energy density E(σ)/NE(\vec{\sigma})/N is evaluated to be

u=1Na=1M{σj:ja}EaeβEajaqjaσj{σj:ja}eβEajaqjaσj.u=\frac{1}{N}\sum\limits_{a=1}^{M}\frac{\sum\limits_{\{\sigma_{j}\mathrel{\mathop{\ordinarycolon}}j\in\partial a\}}E_{a}e^{-\beta E_{a}}\prod\limits_{j\in\partial a}q_{j\rightarrow a}^{\sigma_{j}}}{\sum\limits_{\{\sigma_{j}\mathrel{\mathop{\ordinarycolon}}j\in\partial a\}}e^{-\beta E_{a}}\prod\limits_{j\in\partial a}q_{j\rightarrow a}^{\sigma_{j}}}\;. (28)
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7: Same as Fig. 2, but for the regular random graph ensemble of vertex degree K=4K\!=\!4 and clause degree p=3p\!=\!3 (noise ε=0\varepsilon\!=\!0). (a) Free energy density ff versus canonical inverse temperature β\beta. (b) Entropy density ss versus energy density uu. (c) Microcanonical inverse temperature β\beta versus uu. (d) Mean overlap mm versus uu.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 8: Same as Figs. 2 and  7, but for the regular random graph ensemble of vertex degree K=3K\!=\!3 and clause degree p=3p\!=\!3 (noise ε=0\varepsilon\!=\!0). (a) Free energy density ff versus canonical inverse temperature β\beta. (b) Entropy density ss versus energy density uu. (c) Microcanonical inverse temperature β\beta versus uu. (d) Mean overlap mm versus uu.

In a regular random graph each clause aa is connected to pp randomly chosen vertices (|a|=p|\partial a|\!=\!p) and each vertex ii is involved in KK randomly chosen clauses (|i|=K|\partial i|\!=\!K). For such a uniform system, when the noise level ε=0\varepsilon\!=\!0, the cavity probability distribution functions qiaσiq_{i\rightarrow a}^{\sigma_{i}} will all be independent of the link indices (i,a)(i,a) and could be described by a single parameter m~\tilde{m} as

qiaσi=1+m~2δ(σiσi0)+1m~2δ(σi+σi0).q_{i\rightarrow a}^{\sigma_{i}}=\frac{1+\tilde{m}}{2}\delta(\sigma_{i}-\sigma_{i}^{0})+\frac{1-\tilde{m}}{2}\delta(\sigma_{i}+\sigma_{i}^{0})\;. (29)

The parameter m~\tilde{m} is the mean value of σiσi0\sigma_{i}\sigma_{i}^{0} when vertex ii is affected only by K1K\!-\!1 (instead of KK) attached clauses. The self-consistent equation for m~\tilde{m} is derived from Eq. (25) as

m~=tanh[(K1)atanh(m~p1tanhβ)].\tilde{m}=\tanh\Bigl{[}(K-1)\textrm{atanh}\bigl{(}\tilde{m}^{p-1}\tanh\beta\bigr{)}\Bigr{]}\;. (30)

The mean value of the overlap mm [Eq. (4)], the mean energy density uu, and the free energy density ff, are computed at a fixed point of Eq. (30) as

m\displaystyle m =tanh[Katanh(m~p1tanhβ)],\displaystyle=\tanh\Bigl{[}K\textrm{atanh}\bigl{(}\tilde{m}^{p-1}\tanh\beta\bigr{)}\Bigr{]}\;, (31a)
u\displaystyle u =Kp[m~p+tanhβ1+m~ptanhβ],\displaystyle=-\frac{K}{p}\Bigl{[}\frac{\tilde{m}^{p}+\tanh\beta}{1+\tilde{m}^{p}\tanh\beta}\Bigr{]}\;, (31b)
f\displaystyle f =1β[ln(1+(1m~p1tanhβ1+m~p1tanhβ)K)\displaystyle=-\frac{1}{\beta}\biggl{[}\ln\Bigl{(}1+\bigl{(}\frac{1-\tilde{m}^{p-1}\tanh\beta}{1+\tilde{m}^{p-1}\tanh\beta}\bigr{)}^{K}\Bigr{)}
+Kln(1+m~p1tanhβ1+m~ptanhβ)\displaystyle\quad\quad\quad\quad+K\ln\Bigl{(}\frac{1+\tilde{m}^{p-1}\tanh\beta}{1+\tilde{m}^{p}\tanh\beta}\Bigr{)}
+Kpln(coshβ+m~p1sinhβ)].\displaystyle\quad\quad\quad\quad+\frac{K}{p}\ln\bigl{(}\cosh\beta+\tilde{m}^{p-1}\sinh\beta\bigr{)}\biggr{]}\;. (31c)

And the entropy density ss is computed following Eq. (8). The relationship between ss and uu can be obtained by eliminating β\beta from the u(β)u(\beta) and s(β)s(\beta) curves.

Equation (30) always has a stable fixed point m~=0\tilde{m}\!=\!0 which corresponds to the disordered symmetric phase with zero overlap (m=0m\!=\!0). As the inverse temperature β\beta increases, another stable fixed point appears, which corresponds to the canonical polarized phase with mm quite close to unity. These two stable fixed points are separated by an unstable fixed point, which corresponds to the higher-entropy branch of the microcanonical polarized phase (when β\beta is relatively small) or to the lower-entropy branch of the microcanonical polarized phase (when β\beta is relatively large).

The mean-field theoretical results obtained for the planted 33-body spin model on a regular random network of degree K=4K\!=\!4 are summarized in Fig. 7. Qualitatively speaking, these theoretical results are identical to the results shown in Fig. 2 for K=10K\!=\!10. But the interval between umicu_{\rm mic} and udu_{\rm d} is much smaller and the entropy barrier [defined by Eq. (13)] is much higher at K=4K\!=\!4.

The planted 33-body spin model with degree K=3K\!=\!3 is special in that the minimum energy density of the ferromagnetic CP phase is identical to that of the paramagnetic DS phase. We show in Fig. 8 the mean-field theoretical results obtained for this system. The DS phase experiences a spin glass phase transition at the critical energy density ud=0.961u_{\rm d}\!=\!-0.961 Zhou (2015). At any fixed energy density u>1u\!>\!-1 the entropy density of the DS phase is higher than that of the MP or CP phase, therefore there is no MSSB phase transition in this special system. The MP phase and the CP phase do exist but they are only metastable as compared with the DS phase. The quasi-greedy algorithm of Refs. Bellitti et al. (2021); Bernaschi et al. (2021) relaxed the detailed balance condition to ensure that, once the CP phase is reached, the search process will not jump back to the DS phase.

References

  • Franz et al. (2001) S. Franz, M. Mézard, F. Ricci-Tersenghi, M. Weigt,  and R. Zecchina, “A ferromagnet with a glass transition,” Europhys. Lett. 55, 465–471 (2001).
  • Montanari and Ricci-Tersenghi (2004) A. Montanari and F. Ricci-Tersenghi, “Cooling-schedule dependence of the dynamics of mean-field glasses,” Phys. Rev. B 70, 134406 (2004).
  • Zdeborová and Krzakala (2016) L. Zdeborová and F. Krzakala, “Statistical physics of inference: thresholds and algorithms,” Adv. Phys. 65, 453–552 (2016).
  • Mézard and Montanari (2009) M. Mézard and A. Montanari, Information, Physics, and Computation (Oxford Univ. Press, New York, 2009).
  • Sourlas (1989) N. Sourlas, “Spin-glass models as error-correcting codes,” Nature 339, 693–695 (1989).
  • Huang and Zhou (2009) H. Huang and H. J. Zhou, “Cavity approach to the sourlas code system,” Phys. Rev. E 80, 056113 (2009).
  • Watanabe (2013) O. Watanabe, “Message passing algorithms for mls-3lin problems,” Algorithmica 66, 848–868 (2013).
  • Kirkpatrick et al. (1983) S. Kirkpatrick, C. D. Gelatt Jr.,  and M. P. Vecchi, “Optimization by simulated annealing,” Science 220, 671–680 (1983).
  • Xu et al. (2018) Y.-Z. Xu, C. H. Yeung, H.-J. Zhou,  and D. Saad, “Entropy inflection and invisible low-energy states: Defensive alliance example,” Phys. Rev. Lett. 121, 210602 (2018).
  • Matsuda et al. (2011) Y. Matsuda, H. Nishimori, L. Zdevorová,  and F. Krzakala, “Random-field pp-spin-glass model on regular random graphs,” J. Phys. A: Math. Theor. 44, 185002 (2011).
  • Bandeira et al. (2018) A. S. Bandeira, A. Perry,  and A. S. Wein, “Notes on computational-to-statistical gaps: Predictions using statistical physics,” Portugaliae Mathematica 75, 159–186 (2018).
  • Li et al. (2009) K. Li, H. Ma,  and H. J. Zhou, “From one solution of a 33-satisfiability formula to a solution cluster: Frozen variables and entropy,” Phys. Rev. E 79, 031102 (2009).
  • Krzakala et al. (2014) F. Krzakala, M. Mézard,  and L. Zdeborová, “Reweighted belief propagation and quiet planting for random kk-sat,” Journal on Satisfiability, Boolean Modeling and Computation 8, 149–171 (2014).
  • Krzakala and Zdeborova (2009) F. Krzakala and L. Zdeborova, “Hiding quiet solutions in random constraint satisfaction problems,” Phys. Rev. Lett. 102, 238701 (2009).
  • Zhou (2019a) H.-J. Zhou, “Kinked entropy and discontinuous microcanonical spontaneous symmetry breaking,” Phys. Rev. Lett. 122, 160601 (2019a).
  • Mézard and Parisi (2001) M. Mézard and G. Parisi, “The bethe lattice spin glass revisited,” Eur. Phys. J. B 20, 217–233 (2001).
  • Mézard and Montanari (2006) M. Mézard and A. Montanari, “Reconstruction on trees and spin glass transition,” J. Stat. Phys. 124, 1317–1350 (2006).
  • Zhou (2015) H.-J. Zhou, Spin Glass and Message Passing (Science Press, Beijing, China, 2015).
  • Gross (2001) D. H. E. Gross, Microcanonical Thermodynamics: Phase Transitions in “Small” Systems (World Scientific, Singapore, 2001).
  • Creutz (1983) M. Creutz, “Microcanonical monte carlo simulation,” Phys. Rev. Lett. 50, 1411–1414 (1983).
  • Rose and Machta (2019) N. Rose and J. Machta, “Equilibrium microcanonical annealing for first-order phase transitions,” Phys. Rev. E 100, 063304 (2019).
  • Weiss (1981) G. H. Weiss, “First passage time problems for one-dimensional random walks,” J. Stat. Phys. 24, 587–594 (1981).
  • Khantha and Balakrishnan (1983) M. Khantha and V. Balakrishnan, “First passage time distribution for finite one-dimensional random walks,” Pramana 21, 111–122 (1983).
  • Bauke and Mertens (2007) H. Bauke and S. Mertens, “Random numbers for large-scale distributed monte carlo simulations,” Phys. Rev. E 75, 066701 (2007).
  • Alava et al. (2008) M. Alava, J. Ardelius, E. Aurell, P. Kaski, S. Krishnamurthy, P. Orponen,  and S. Seitz, “Circumspect descent prevails in solving random constraint satisfaction problems,” Proc. Natl. Acad. Sci. USA 105, 15253–15257 (2008).
  • Ma and Zhou (2011) H. Ma and H. Zhou, “Approaching the ground states of the random maximum two-satisfiability problem by a greedy single-spin flipping process,” Phys. Rev. E 83, 052101 (2011).
  • Efron (1979) B. Efron, “Computers and the theory of statistics: Thinking the unthinkable,” SIAM Rev. 21, 460–480 (1979).
  • Van den Broeck and Reimann (1996) C. Van den Broeck and P. Reimann, “Unsupervised learning by examples: On-line wersus off-line,” Phys. Rev. Lett. 76, 2188–2191 (1996).
  • Reimann and Van den Broeck (1996) P. Reimann and C. Van den Broeck, “Learning by examples from a nonuniform distribution,” Phys. Rev. E 53, 3989–3998 (1996).
  • Engel and Van den Broeck (2001) A. Engel and C. Van den Broeck, Statistical Mechanics of Learning (Cambridge University Press, Cambridge, UK, 2001).
  • Kabashima and Uda (2004) Y. Kabashima and S. Uda, “A bp-based algorithm for performing bayesian inference in large perceptron-type networks,” Lect. Notes Comput. Sci. 3244, 479–493 (2004).
  • Braunstein and Zecchina (2006) A. Braunstein and R. Zecchina, “Learning by message passing in networks of discrete synapses,” Phys. Rev. Lett. 96, 030201 (2006).
  • Huang and Toyoizumi (2015) H. Huang and T. Toyoizumi, “Advanced mean-field theory of the restricted boltzmann machine,” Phys. Rev. E 91, 050101(R) (2015).
  • Zhou (2019b) H.-J. Zhou, “Active online learning in the binary perceptron problem,” Commun. Theor. Phys. 71, 243–252 (2019b).
  • Hu et al. (2019) G.-K. Hu, T. Liu, M.-X. Liu, W. Chen,  and X.-S. Chen, “Condensation of eigen microstate in statistical ensemble and phase transition,” Science China Phys. Mech. Astron. 62, 990511 (2019).
  • Bellitti et al. (2021) M. Bellitti, F. Ricci-Tersenghi,  and A. Scardicchio, “Entropic barriers as a reason for hardness in both classical and quantum algorithms,” e-print arXiv:2021.00182 [cond-mat.dis-nn]  (2021).
  • Bernaschi et al. (2021) M. Bernaschi, M. Bisson, M. Fatica, E. Marinari, V. Martin-Mayor, G. Parisi,  and F. Ricci-Tersenghi, “How we are leading a 3-xorsat challenge: from energy landscape to the algorithm and its efficient implementation on gpus,” e-print arXiv:2021.09510 [cond-mat.dis-nn]  (2021).