<<

1D Three-state mean-field Potts model with first- and second-order phase transitions

Massimo Ostillia, Farrukh Mukhamedovb

aInstituto de F´ısica,Universidade Federal da Bahia, Salvador, Brazil bDepartment of Mathematical Sciences, College of Science, United Arab Emirates University, Al Ain, Abu Dhabi, UAE

Abstract

We analyze a three-state Potts model built over a lattice ring, with coupling J0, and the fully connected graph, with coupling J. This model is effectively mean-field and can be exactly solved by using transfer-matrix method and Cardano formula. When J and J0 are both ferromagnetic, the model has a first-order which turns out to be a smooth modification of the known phase transition of the traditional mean-field Potts model (J0 = 0), despite, as we prove, the connected correlation functions are now non zero, even in the paramagnetic phase. Furthermore, besides the first-order transition, there exists also a hidden continuous transition at a below which the symmetric metastable state ceases to exist. When J is ferromagnetic and J0 antiferromagnetic, a similar antiferromagnetic counterpart phase transition scenario applies. Quite interestingly, differently from the Ising-like two-state case, for large values of the antiferromagnetic coupling J0, the critical temperature of the system tends to a finite value. Similarly, also the latent heat per tends to a finite constant in the limit of J0 → −∞.

1 1. Introduction correlations whenever H0 has short-range interactions . Of course, unlike the traditional mean-field models (where H0 = 0 and there are not short-range correlations), the arbi- trariness of H0 lets it open now a very richer scenario of phase The mean-field concept is a fundamental paradigm in theo- transitions. In particular, it can be shown that, when H0 has retical physics and its interdisciplinary applications. It consists antiferromagnetic interactions, inversion transition phenomena in replacing the interactions acting on a particle with an effec- and first-order phase transitions may set in [4]. More in gen- tive external field to be determined self-consistently. The power eral, the phase transition scenario associated to the term ∆H can of this approach manifests in two ways: on one hand, it allows change drastically when H0 has antiferromagnetic couplings. to face analytically, in a first approximation, any given model; In recent years, a renewed attention toward models hav- on the other hand, it provides a powerful understanding of the ing both short- and long-range interactions, has been drawn physics of the model. In fact, even though very approximate, due to the importance of small-world networks [5], where the mean-field solution is often pedagogically deeper than the a finite-dimensional and an infinite-dimensional character are understanding one would get from a possible exact solution (if both present in the network structure. As expected, such mod- any). In particular, it would be harder to understand the concept els turn out to be mean-field, at least for what concerns their of the collective behavior and the phase transitions of a system critical behavior. However, rather than a theorem, except for without a suitable mean-field theory. the Ising case near the critical point [8, 4], and a few exam- At the mathematical base of the mean-field theory there are ples in one dimension [6, 7], this turns out to be an empir- models which are exactly solvable by a mean-field technique: ical fact. An exact analytically treatment, even not rigorous the mean-field models. These models represent the limit cases and confined to relatively simple models, is still far from being of more realistic models in which one or more parameters are reached when short-range correlations are present, as happens typically send to 0 or to ∞ so that the mean-field approxima- in a small-world network. On the other hand, the models intro- tion becomes exact. Traditionally, the concept of the mean-field duced in [1] can be seen as ideal small-world networks in which models is associated with the absence of correlations in the ther- the random connectivity of the graph goes to the system size N modynamic limit. In [1] (see also [2] and [3]) we have shown and the coupling J associated to the long-range interactions is

arXiv:1205.6777v4 [cond-mat.stat-mech] 7 Jun 2020 that this condition is only a sufficient condition for the system replaced by J/N. Clearly, without a serious understanding of to be mean-field, but in general it is not necessary. There exist the more basic models presented in [1], the analytical study of in fact infinite many models having both non zero correlations the small-world networks and its generalizations (including the and a mean-field character. For example, if H0 is an arbitrary scale-free case [9], which for the Ising case has been analyzed Hamiltonian, the model H = H0 + ∆H, with ∆H a general fully in [10]), will remain impossible. connected interaction, is mean-field, in the sense that we can In this spirit, in the present paper we analyze a simple and yet exactly replace the interactions acting on a particle with an ef- fective external field to be determined self-consistently. How- 1The case of power-law like long-range interactions is more subtle. See the ever, now, the presence of the term H0 gives rise to non zero Conclusions in [1]. rich model: a case in which H0 is a one-dimensional three-state Eq. (3) we find that, if x0;σ(βh1, . . . βhq) is the order param- 2 Potts model [11] and ∆H is the traditional three-state mean- eter for H0 as a function of a q-component external field , field term, i.e., the ordinary fully-connected interaction. The hδ(σi, σ)i0 = x0;σ(βh1, . . . βhq), then, in the thermodynamic mean-field equations in this case are sufficiently simple to be limit, the order parameter for H, xσ = hδ(σi, σ)i, satisfies the exactly solved via the transfer matrix method and the Cardano system formula for cubic equations. As expected, similarly to the ana-   log Ising case [4], the presence of a non zero ferromagnetic xσ = x0;σ βJx1, . . . , βJxq , σ = 1,..., q (4) coupling, J0 > 0, in H0, alters only smoothly the phase dia- gram of the system characterized by a first-order phase tran- and the free energy f is given by sition. The difference with respect to the case without H0 is 2 X βJxσ that, for H0 , 0, the connected correlations functions are now β f = + β f (βJx , . . . , βJx ). (5) 2 0 1 q not zero. Besides the first-order transition, there emerges also σ a second-order transition. This continuous transition is not sta- When J < 0, the approach with the Gaussian variables is not ble (the corresponding free energy being not a local minimum valid since the Gaussian integral diverges. Yet, the saddle point but a saddle point), however it corresponds to a non trivial so- Eqs. (4) are still exact, as derived from the general theorem lution of the mean-field equations and occurs at a temperature (SO) (FO) presented in [1] (while the free energy has a different form with T < T below which the symmetric solution ceases to c c respect to Eq. (5)). exist as a metastable state. When H0 has an antiferromagnetic Concerning the connected C, as a gen- coupling, J0 < 0, a similar phase transition scenario still ap- plies but characterized by an antiferromagnetic order and, quite eral rule we have [1] interestingly, differently from the Ising-like two-state case, in (FO) C = C0(βJx1, . . . , βJxq) + finite size effects, (6) the limit J0 → −∞, Tc tends to a finite value. Moreover, we show that in the same limit also the latent heat per spin tends where C0(βh1, . . . , βhq) is the connected correlation function of to a finite constant. Finally, we prove that the connected cor- the Potts model governed by H0 at the temperature 1/β and in relation functions are not zero and evaluate them in a specific de f the presence of a q-component external field h = (h ,..., h ). case. 1 q

3. The traditional mean-field Potts model (H = 0) 2. Generalized mean-field Potts models 0 Before facing the analysis of our model, we want to briefly In the spirit of [1], we introduce now a model built by us- recall the traditional mean-field Potts model defined as in Eq. ing both finite-dimensional and infinite dimensional Hamilto- (1) with H = 0. nian terms. A generalized mean-field Potts model, i.e., a model 0 where each variable σ can take q values, σ = 1,..., q, can be defined through the following Hamiltonian 3.1. The pure model J X The use of Eqs. (4)-(5) in this case may seem not necessary H = H ({σ }) − δ(σ , σ ), (1) 0 i N i j but it is instructive. To apply Eqs. (4)-(5) to the present case, i< j we need to solve the corresponding pure model, which is a Potts 0 model without interaction but in the presence of a uniform ex- where δ(σ, σ ) is the Kronecker delta function and H0 is any q-states Potts Hamiltonian with no external field. Let us rewrite ternal field, h. We have therefore to calculate the following H as (up to terms negligible for N → ∞) trivial partition function, Z0(h), which differs from Z for the absence of the fully-connected (long-range) interaction:  2 X X J   X β P h N H = H ({σ }) −  δ(σ , σ) . (2) Z (h) = e σ σ σ , (7) 0 i N  i  0 σ i σ1,...,σN P As done in [1], from Eq. (2) we see that, by introducing q in- where Nσ = i δ(σ, σi). We have dependent Gaussian variables xσ, we can evaluate the partition function, Z = P exp(−βH({σ })), as   eβhσ {σi} i x βh , . . . , βh = , (8) 0;σ 1 q P βh 0 σ0 e σ Z q  2  Y −N P βJxσ +β f (βJx ,...,βJx ) Z ∝ dx e σ 2 0 1 q , (3) σ   σ=1 X  −  βhσ  β f0(βh1, . . . , βhq) = log  e  . (9) where f0(βh1, . . . , βhq) is the free energy density of the Potts σ model governed by H0 at the temperature 1/β and in the pres- de f ence of a q-component external field h = (h ,..., hq) via H → 2 P P 1 0 We suppose, for simplicity, that the order parameter associated to H0 (the H0 − σ hσ i δ(σi, σ). By using the saddle point method, from pure model), does not depend on the vertex position. 2 3.2. The mean-field model By plugging Eqs. (8)-(9) in Eqs. (4)-(5) we get immediately 1D 3-state Potts, J =0, J=1 the following system of equations and the free energy density: 0

eβJxσ 1 x = , σ = 1,..., q, (10) σ P βJx 0 σ0 e σ x1 Leading

0,8 x2, x3 Leading   x , x Unstable A X X 2 1 2  βJx  βJxσ x Unstable A β f = − log  e σ  + . (11) 3   x Metastable A 2 0,6 1 σ σ x , x Metastable A 3 2 3 , x

2 x , x ,x Metastable S Eqs. (10)-(11) give rise to a well known phase transition sce- 1 2 3 , x 1 x nario [11]: a second-order mean-field Ising phase transition sets 0,4 up only for q = 2, while for any q ≥ 3 there is a first-order phase transition at the critical value (see Fig. 1): 0,2 2(q − 1) β(FO) J = log(q − 1). (12) c q − 2 0 0,2 0,25 0,3 0,35 0,4 It is easy however to see that, besides the first-order transition, T there exists also a hidden (unstable) second-order transition tak- ing place when [12] Figure 1: (Color online) Magnetizations for the case J = 1 and J0 = 0 (equiva- lent to the traditional mean-field Potts model). In the figure: “Leading” stands (SO) for the thermodynamic stable state having the lower free-energy, “Metastable βc J = q. (13) A” and “Metastable S” stand for the thermodynamic stable states having the At equilibrium the main role of this second-order transition is higher free-energy, while “Unstable S” and “Unstable A” stand for the ther- to determine the temperature T (SO) below which the metastable modynamic unstable states; A or S stand for asymmetrical or symmetrical, c respectively. For any J ≥ 0 (which includes the present case J = 0), the symmetric state ends to be (locally) stable (see Fig. 2). We shall 0 0 Leading states lie on the subspaces xi1 > xi2 = xi3 (in this figure only the see later that this phase transition scenario holds true (robust) case x1 > x2 = x3 is shown), while the Unstable A states lie on the sub- spaces x = x > x , where i , i , i , is any permutation of the set of indices also in the presence of a positive short-range coupling (H0 , 0). i1 i2 i3 1 2 3 {1, 2, 3} and xi1 + xi2 + xi3 = 1 (in this figure only the case x1 = x2 > x3 is shown). Note that the asymptotic values of the magnetizations toward T = 0

4. The three-state 1D case are xi1 = 1, xi2 = xi3 = 0 and xi1 = xi2 = 1/2, xi3 = 0, for the Leading and Unstable A states, respectively. The first-order phase transition occurs at (FO) (SO) We now specialize the above general result to the case in Tc = 0.3607, while the second-order one at Tc = 1/3. The Metastable A state begins at T = 0.364 and ends at T (FO). which H0 represents a one-dimensional three-state Potts Hamil- c tonian: XN H0({σi}) = −J0 δ(σi, σi+1) (14) i=1 where we have assumed periodic boundary conditions σN+1 = 1D 3-state Potts, J0=0, J=1 σ1, and from now on it is understood that each Potts variable σ can take the values 1, 2, and 3. 1

x Leading 4.1. The pure model 1 0,8 x2, x3 Leading x , x Unstable A To apply Eqs. (4)-(5) to our case we need to solve the cor- 1 2 x3 Unstable A responding pure model, which is a 1D three-state Potts model x Metastable A 0,6 1 x , x Metastable A in the presence of a three-component uniform external field, h, 3 2 3 , x

2 x , x ,x Metastable S i.e., we have to calculate the following partition function 1 2 3 , x 1 x X PN P βJ0 δ(σi,σi+1)+β σ hσ Nσ 0,4 Z0(h) = e i=1 . (15)

σ1,...,σN For any finite N, we can express (15) as (“transfer matrix 0,2 method”)

N 0 Z0(h) = Tr T , (16) 0,35 0,36 0,37 T where T is the 3 × 3 matrix whose elements, T(σ, σ0), for any σ, σ0 ∈ {1, 2, 3}, are defined as Figure 2: (Color online) Particular of Fig. (1) near the critical temperature. " # 0 0 1 1 T(σ, σ ) = exp βJ δ(σ, σ ) + βh + βh 0 . (17) 0 2 σ 2 σ 3 ∂a2 For the thermodynamic limit it will be enough to evaluate the = −eβJ0 eβhσ , (32) eigenvalues of T, λ1, λ2, λ3, the free energy density of the pure ∂βhσ model, f0, being given by − 2 ! log (Z (h)) ∂S 1  1  3 ∂R 1 ∂D 0 R D 2 , −β f0(βh) = lim = log(λmax), (18) = + + 1 (33) N→∞ N ∂βhσ 3 ∂βhσ 2D 2 ∂βhσ where λmax is such that |λmax| = max {|λ1|, |λ2|, |λ3|}. From Eq. − 2 ! (17) we see that the eigenvalues equation reads ∂T 1  1  3 ∂R 1 ∂D 2 = R − D − 1 , (34) 3 2 ∂βhσ 3 ∂βhσ 2D 2 ∂βhσ λ + a2λ + a1λ + a0 = 0, (19) where ∂D 2 ∂Q ∂R     = 3Q + 2R , (35) βJ0 3βJ0 βh1+βh2+βh3 ∂βh ∂βh ∂βh a0 = 3e − e − 2 e , (20) σ σ σ

    ∂a1 ∂a2 2βJ0 βh1+βh2 βh2+βh3 βh1+βh3 3 − 2a2 a = e − 1 e + e + e , (21) ∂Q ∂βhσ ∂βhσ 1 = , (36) ∂βhσ 9   βJ0 βh1 βh2 βh3 a2 = −e e + e + e . (22) ∂a1 ∂a2 ∂a0 2 ∂a2 ∂R 9 ∂βh a2 + 9 ∂βh a1 − 27 ∂βh − 6a2 ∂βh Eq. (19) is cubic in λ so that we can solve it explicitly using the = σ σ σ σ . (37) Cardano formula which gives the three roots: ∂βhσ 54 1 λ = − a + (S + T), (23) 4.2. The mean-field model 1 3 2 √ By performing the effective substitutions h → Jx in Eqs. 1 1 i 3 σ σ λ = − a − (S + T) + (S − T), (24) (18)-(37), Eqs. (4)-(5) take the form 2 3 2 2 2 √  0  1 1 i 3 1 ∂λmax βhσ | 0 0 λ3 = − a2 − (S + T) − (S − T), (25) xσ = {hσ=Jxσ}, (38) 3 2 2 λmax ({βJxσ}) ∂βhσ where 2 1 1 X βJxσ  1  3  1  3 β f = − log(λ ({βJx })), (39) S = R + D 2 , T = R − D 2 , (26) 2 max σ σ

2 where we have written the explicit dependence on the argu- 3a1 − a2 D = Q3 + R2, Q = , (27) { } 9 ments βJxσ . For the per spin u we have 3 9a1a2 − 27a0 − 2a2 1 X R = . (28) u = f − x log(x ), (40) 54 β σ σ σ Once the eigenvalues have been calculated, the magnetizations, where the second term corresponds to the per spin. Par- x0;σ, of the pure model in the thermodynamic limit can be cal- culated from Eq. (18) as ticularly simple are the expressions corresponding to the sym- metric solution {xσ = 1/q}. Direct application of the transfer 1 ∂λ x = max . (29) matrix method provides (these formulas are valid for any q) 0;σ λ ∂βh max σ   βJ β f = − log eβJ0 + q − 1 − , (41) Notice that λ1, λ2, and λ3 are functions of the vector-field h. In symm 2q the case of equal external fields, h1 = h2 = h3, which in par- ticular includes the case h = h = h = 0, due to the fact that 1 2 3 1 a0, a1, and a2 are symmetrical in the fields, Eq. (29) provides u = f + log(q), (42) symm symm β always the symmetric solution x0;σ = 1/3, i.e., as expected, in one dimension there is no phase transition. Less trivial is to where fsymm and usymm stand for free energy and energy (per evaluate Eq. (29) for an arbitrary external field h. To this aim, spin) of the symmetric solution. Notice that, whereas f is con- from Eqs. (20)-(28) we see that we need to take into account tinuous at the critical point of a first-order phase transition, u 0 00 the following derivatives, with {σ, σ , σ } = {1, 2, 3} is not. Eqs. (40) and (42) can be used for evaluating the latent

∂a0 heat per spin L defined as the difference of the energies of the = a0 (30) symmetric solution with the non symmetric one at the critical ∂βhσ point of the first-order phase transition:   ∂a1  2βJ   βh +βh 0 βh +βh 00  = e 0 − 1 e σ σ + e σ σ , (31) L = usymm − u . (43) | (FO) ∂βhσ T=Tc 4 In the case J0 = 0 one can shows that L reduces to [11] of a second-order phase transition which lies in the sub-space 2 L = βJ(q − 2) /(2q(q − 1)). The latent heat is interesting be- xi1 = xi2 > xi3 , where i1, i2, i3, is any permutation of the set of cause it quantifies the amount of energy the system requires to indices {1, 2, 3} and xi1 + xi2 + xi3 = 1. This second-order phase (SO) “transform” a metastable state (i.e. subleading) into a leading transition takes place at a critical temperature Tc where the one along the first-order transition, in close analogy with the metastable state x1 = x2 = x3 = 1/3 (i.e. a stable state with change of phase of fluids, like the gas-liquid transition. an higher free energy with respect to the stable leading state) Note that in the present work we are not going to consider becomes unstable and two components become favored against an additional external field (see Eq. (1)): according to Eqs. a third one, so that their values toward T = 0 are the states (4)-(5), the role that the external field had on the pure model (1/2,1/2,0) (and their permutations). As for any model having a H0, has been now replaced by the effective magnetizations Jxσ fully connected interaction 1, also this second-order transition to be found self-consistently by Eqs. (38). We could easily is mean-field like with classical critical exponents (for each or- consider the presence of an additional external field h by sim- der parameter we have α = 0, β = 1/2, γ = 1, δ = 1/3), ply performing the effective substitutions hσ → Jxσ + hσ in as can be checked directly or by applying the general result of Eqs. (18)-(37), which does not change the structure of the self- Ref. [1]. For J0 = 0 it is easy to see that the critical temperature (SO) consistent Eqs. (38), but its numerical detailed analysis goes of this transition is given by Tc = J/q. Here the symmetry beyond the aim of the present work. to be broken seems to be a two-fold one, as can be seen if we consider that two non zero components are forced to change si-

multaneously by the constrain xi1 = xi2 . However, only the state 5. Numerical analysis and physical interpretation of the coming from the first-order transition is stable, while the other self-consistent equations turns out to be unstable, as can be seen from the fact that the initial conditions giving rise to the second-order phase transi- In this Section we analyze numerically Eqs. (38) and (39) tion live in a subspace of (x , x , x ) of the kind x = x which and provide the corresponding physical explanation. It turns 1 2 3 i1 i2 has zero volume in 3 dimensions. A basin of attraction of zero out that λ coincides always with λ . We find it convenient to max 1 volume corresponds to a unstable state. In fact, a control of distinguish the cases J ≥ 0 and J < 0 both for J > 0. Later 0 0 the Hessian of the Landau free energy (39) 3 gives, for the so- on we will consider also the case J < 0. As we have seen in lution corresponding to the second-order phase transition, al- the previous Section, the pure model, in one dimension, does ways one negative eigenvalue. The presence of a second-order not undergo a spontaneous symmetry breaking. However, from phase transition in a non disordered three-state mean-field Potts Eqs. (38) and (39) we see that, for any positive value of J, the model, even if hidden in a subspace, is a quite non obvious and model governed by H turns out to be a mean-field model so interesting fact: in the subspace x = x > x , as the temper- that a phase transition is always expected. Therefore, in our nu- i1 i2 i3 ature is decreased, the system is forced to favor x and x not merical experiments, it will be enough to keep the value of the i1 i2 by a jump, but continuously with a two-fold broken symmetry long-range coupling fixed at J = 1 and obverse what happens mechanism which in turn sets the end of the metastable state. by changing the short-range coupling J0. In the next paragraph we will see that for J0 < 0 this scenario In general, the trivial and symmetric solution x1 = x2 = x3 = (FO) is somehow reversed. 1/3 is stable even below the critical temperature Tc within (SO) (FO) a finite range of [Tc , Tc ], though in general (SO) 5.0.2. The case J > 0,J0 < 0 is not leading (i.e., it is a metastable state), and below Tc becomes unstable. When J0 < 0 the model is still mean field, so that a first- order phase transition similar to the case J0 > 0 is also present, but with a corresponding lower value for T (FO), as confirmed 5.0.1. The case J > 0,J ≥ 0 c 0 by Fig. (4). Now, however, due to the fact that J < 0, we When we set J → 0, our model coincides with the tra- 0 0 must to take into account that two spins that are consecutive ditional mean-field model governed by Eqs. (10) and (11). along the 1D chain, at low enough temperatures, cannot have When J = 0, from Eq. (12) with q = 3, we see that for 0 the same value, so that, among the three components (1, 2, 3), J = 1 a first-order phase transition develops at the critical point the favored one(s), if any, along the 1D chain must be alter- T (FO) = 0.3607, see Fig. (1). The phase transition is triggered c nated with another one (others). There are two ways to real- by a broken symmetry mechanism according to which one of ize this alternation. If for example we look for situations in the three components (1, 2, 3) becomes favored in spite of the which, at low enough temperatures, the component 1 is favored, other two that remain equal to each other so that, for T → 0, along the 1D chain we can look for configurations of the kind two components go to 0 and the favored one reaches the value (1, 2, 1, 3, 1, 2, 1, 3,...). But we can also look for situations in 1. As expected, when J > 0 we observe a smooth modification 0 which, for example, both the components 1 and 2 are favored of such a scenario, as reported in Figs. (2)-(3). Fig. 5 shows and, at low enough temperatures, the configurations are of the that T (FO) is an increasing function of J for J > 0. Similarly, c 0 0 kind (1, 2, 1, 2, 1, 2, 1, 2,...). Notice, for both the situations, Fig. 6 shows that the latent heat per particle L is an increasing function of J0 for J0 > 0. Besides the first-order phase transition (“dominant”, or 3 The true free energy is given by (39) calculated in the solutions of the “leading”), as shown in Figs. (1-3), we observe the existence system (38), while the Landau free energy is represented by (39) alone. 5 1D 3-state Potts, J0=1, J=1

1

x Leading 1 1D 3-states Potts, J =-0.5, J=1 0,8 0 x2,x3 Leading x , x Unstable A 1 2 0,5 x3 Unstable A x Metastable A 0,6 1 3 x2, x3 Metastable A , x

2 0,4 x1, x2, x3 Metastable S , x 1 x 0,4

0,3 3 , x 0,2 2 , x 1 x1,x2 Leading x 0,2 x3 Leading

x1 Unstable A

0 x2, x3 Unstable A 0,6 0,7 0,8 0,9 Metastable S T 0,1 x1, x2 Metastable A

x3 Metastable A Figure 3: (Color online) As in Fig. (1) but with J0 = 1. 0 0,08 0,1 0,12 0,14 T with respect to the case J0 ≥ 0, the necessary modification of the values (x1, x2, x3) for T → 0 due to the alternations: now Figure 4: (Color online) As in Fig. 1 with J0 = −0.5. For J0 < 0 the Leading states lie on the subspaces x = x > x (in this figure only the case x = x > the asymptotic values are either (1/2, 1/4, 1/4) or (1/2, 1/2, 0) i1 i2 i3 1 2 x is shown), while the Unstable A states lie on the subspaces x = x < x (and their permutations) for the above former and latter case, 3 i1 i2 i3 (in this figure only the case x1 > x2 = x3 is shown), where i1, i2, i3, is any respectively. In both the cases, we have a translational broken permutation of the set of indices {1, 2, 3}, and xi1 + xi2 + xi3 = 1. Note that, symmetry (similarly to an antiferromagnetic ) but for J0 < 0, the asymptotic values of the magnetizations toward T = 0 are x = x = 1/2, x = 0 and x = 1/2, x = x = 1/4, for the Leading and in the latter case, we have also a further two-fold broken sym- i1 i2 i3 i1 i2 i3 Unstable A states, respectively. metry (since one state, either the state 2 or the state 3, must be excluded). Interestingly, while the latter phase-transition mech- anism corresponds to a first-order phase transition, which turns out to be, in shape, quite similar to the phase transition that occurred for J0 > 0, the former phase-transition mechanism corresponds to a second-order phase transition. However, as in the case J0 ≥ 0, only the state coming from the first-order transition is stable, while the other turns out to be unstable, as (FO) seen from the fact that the initial conditions giving rise to the Tc vs J0, J=1 second-order phase transition live in a subspace of (x1, x2, x3) of the kind xi1 = xi2 which has zero volume in 3 dimensions or, alternatively, by controlling the Hessian of the Landau free 8 0,4 energy (39). Again, we stress that the presence a second-order 0,2 phase transition in a non disordered three-state mean-field Potts 0 6 -8 -6 -4 -2 0 2 model, even if hidden in a subspace, is a quite non obvious and (FO) Tc interesting fact: in the subspace xi1 = xi2 < xi3 , as the tem- perature is decreased, the system is forced to favor xi not by a T 3 4 jump, but continuously and, in turn, this transition sets the end of the symmetric metastable state. (FO) As shown in Fig. 5, the behavior of Tc as a function of J0 2 for J0 < 0 is quite interesting. Differently from the Ising-like two-state case, where the critical temperature obeys the equa- tion exp(2βJ0)βJ = 1 [4] (so that it tends to zero for J0 → −∞), 0 (FO) 0 10 20 30 40 J in the present 3-state case Tc tends to a finite constant for 0 J0 → −∞ (see Inset of Fig. 5). Similarly, as shown in Fig. 6, the latent heat per spin also tends to a finite constant for Figure 5: (Color online) Behavior of the critical temperature of the first-order (FO) transition Tc as a function of J0 with J = 1 (filled dots; the line is a guide J0 → −∞. for the eyes). Inset: particular in the region of J0 negative. 5.0.3. The case J < 0 As anticipated in Sec. II, when J < 0, the saddle point Eqs. (38) are still exact, as derived from the general theorem pre- 6 L vs J0, J=1 {xi = 1/q}, which, according to Eq. 6, amounts to have a con- 0,8 stant effective external field h = (h,..., h) for which we have

L trivially C = C0(βJx1, . . . , βJxq) = C0(0,..., 0), i.e., as Eq. (45). For finite size effects see Sec. 3 of Ref. [13].

0,6 7. Conclusions

0,4 On the base of a general result [1], we have considered a Energy three-state Potts model built over a lattice ring, with coupling J0, and the fully connected graph, with coupling J. This is a non 0,2 trivial exactly solvable effective mean-field model where new phenomena emerge as a consequence of the interplay between its finite- and infinite-dimensional character. A similar analysis

0 was done in [4] for the 1D mean-field Ising model (equivalent -4 -2 0 2 4 6 8 to a two-state Potts model). The three-state Potts model, how- J0 ever, shows dramatic differences with respect to the Ising case. Figure 6: (Color online) Behavior of the latent heat per spin L defined in Eq. In particular, for given J > 0, we have found that, unlike the (43) as a function of J with J = 1 (filled dots; the line is a guide for the eyes). (FO) 0 1D mean-field Ising model [4], the critical temperature Tc tends to a finite constant when J0 → −∞. Similarly, also the latent heat per spin tends to a finite constant for J → −∞. sented in [1] (while the free energy has a different form with 0 Moreover, we have found the existence of a hidden continuous respect to Eq. (39)). When J < 0, Eqs. (38) have only the phase transition for both the ferromagnetic, J ≥ 0, and the trivial symmetric solution and no phase transition sets in. A 0 antiferromagnetic case, J < 0, taking place at a temperature more interesting scenario can emerge under a dynamical ap- 0 T (SO) < T (FO), confirming the robustness of the scenario found proach as done in [12] for the case J = 0 where a dynamical c c 0 in [12] for J = 0 and J > 0. However, for J < 0, the system second-order phase transition takes place. The analysis of the 0 0 has an antiferromagnetic feature, the ground state being charac- dynamical approach for J0 , 0 will be reported elsewhere. terized by a totally different symmetry with respect to the case J0 > 0 (compare the asymptotic values toward T = 0 of Figs. 6. Correlation functions 1 and 4). Concerning the case J < 0, at equilibrium the only possible stable state is the symmetric one, while a more inter- We want to prove now that, as anticipated, the connected esting scenario emerges under a dynamical approach as done correlation function of the effective mean-field model (1) are in [12] for J0 = 0, where the system undergoes only second- not zero and evaluate them in a specific case. Let us consider order phase transitions (stable). The extension of the dynami- for simplicity open boundary conditions and the two point con- cal analysis for J0 , 0 will be reported elsewhere. Finally, we nected correlation function of two consecutive spins. It is easy have evaluated the connected correlation functions of nearest to see that, for the pure q-state Potts model at zero external field, spins and proven that they are not zero, even in the paramag- we have netic phase. eβJ0 δσi,σi+1 0 = , (44) eβJ0 + q − 1 Acknowledgments which implies M. O. acknowledges Grant CNPq 09/2018 - PQ (Brazil). F. M. thanks UAEU UPAR Grant No. 31S391.

de f X   − C0(0,..., 0) = δσi,σδσ,σi+1 0 δσi,σ 0 δσ,σi+1 0 References σ [1] M. Ostilli, EPL 97, 50008 (2012). eβJ0 1 = − . (45) [2] N. N. Bogoliubov, jr., A Method for Studying Model Hamiltonians, Perg- eβJ0 + q − 1 q amon Press (Oxford, 1972). [3] L. W. J. den Ouden, H. W. Capel, and J. H. H. Perk, Physica A 85 425 Equation (45) shows that, in the pure 1D model, the connected (1976); J. H. H. Perk, H. W. Capel, and L. W. J. den Ouden, Physica A correlation functions are zero only in the limit of infinite tem- 89 555 (1977); H. W. Capel, J. H. H. Perk, and L. W. J. den Ouden, Phys. Lett. 66A 437 (1978). perature and, as expected, they are strictly positive or strictly [4] M. Ostilli and J. F. F. Mendes, Phys. Rev. E 78, 031102 (2008). negative according to the sign of J0, respectively. Now, on ap- [5] D. J. Watts, S. H. Strogatz, Nature, 393, 440 (1998). plying Eq. 6 to Eq. (45), we see that, even above the critical [6] N. S. Skantzos and A. C. C. Coolen, J. Phys. A: Math. Gen. 33, 5785 temperature, the connected correlation functions of the effec- (2000). (FO) [7] Skantzos, Nikos S. and Castillo, Isaac Perez´ and Hatchett, Jonathan P. L., tive mean-field model (1) are not zero. In fact, for T > Tc , Phys. Rev. E 72, 066127 (2005). the equilibrium state corresponds to the symmetric solution, [8] M. B. Hastings, Phys. Rev. Lett. 96, 148701 (2006). 7 [9] R. Albert, A.L. Barbasi, Rev. Mod. Phys. 74 47 (2002); S.N. Dorogovtsev, J.F.F. Mendes, Evolution of Networks (University Press: Oxford, 2003); M. E. J. Newman, SIAM Review 45, 167 (2003); S. N. Dorogovtsev, Lec- tures on Complex Networks (Oxford Master Series in Statistical, Compu- tational, and Theoretical Physics, 2010). [10] M. Ostilli, A. L. Ferreria, and J. F. F. Mendes, Phys. Rev. E 83, 061149 (2011). [11] F. Y. Wu, Rev. Mod. Phys., 54, 235 (1982). [12] M. Ostilli and F. Mukhamedov EPL, 101, 60008 (2013). [13] L. Nicolao and M. Ostilli, Physica A 533, 121920 (2019).

8