Skip to main content
Log in

Computation of Microcanonical Entropy at Fixed Magnetization Without Direct Counting

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

Abstract

We discuss a method to compute the microcanonical entropy at fixed magnetization without direct counting. Our approach is based on the evaluation of a saddle-point leading to an optimization problem. The method is applied to a benchmark Ising model with simultaneous presence of mean-field and nearest-neighbour interactions for which direct counting is indeed possible, thus allowing a comparison. Moreover, we apply the method to an Ising model with mean-field, nearest-neighbour and next-nearest-neighbour interactions, for which direct counting is not straightforward. For this model, we compare the solution obtained by our method with the one obtained from the formula for the entropy in terms of all correlation functions. This example shows that for general couplings our method is much more convenient than direct counting methods to compute the microcanonical entropy at fixed magnetization.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. We remind that the Legendre–Fenchel transform of any function is automatically concave.

  2. For \(S_i= -1,0,1\) and a function \(U([S_i])\) given by just a term proportional to \(S_i^2\) we would obtain the Hamiltonian of the Blume-Emery-Griffiths model; however, in this computation we are not assuming a specific spin model and a specific function \(U([S_i])\).

  3. This also shows why, as noted above, the use of the representation of the Dirac \(\delta \), instead of that of the Kronecker \(\delta \), has no importance in this computation.

  4. It is also not difficult to realize from Eq. (19) that \(\exp \left[ -N {\hat{\psi }}(0,\varphi ) \right] \) is equal to the partition function of N independent spins subject to a magnetic field h with \(\varphi \) playing the role of \(\beta h\).

  5. It is not difficult to see that for \(\beta > 0\) the conditions (97) and (98), taken together, imply the condition (99); this is not true for \(\beta < 0\).

  6. In a system with only short-range interactions, where ensembles are equivalent and the two entropies are always equal, \({\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m)\) would have the same straight line segment, and for the values of \(\epsilon \) inside the range of the segment the equilibrium states, in both ensembles, would be realized with a phase separation, something that does not occur in long-range systems.

  7. We remind that, because of the properties of min–max extremal problems [3], in general one has \(s_{\mathrm{micr}}(\epsilon ) \le s_{\mathrm{can}}(\epsilon )\).

References

  1. Huang, K.: Statistical Mechanics. Wiley, New York (1987)

    MATH  Google Scholar 

  2. Pathria, R.K.: Statistical Mechanics. Butterworth-Heinemann, Oxford (1996)

    MATH  Google Scholar 

  3. Campa, A., Dauxois, T., Ruffo, S.: Statistical mechanics and dynamics of solvable models with long-range interactions. Phys. Rep. 480(3–6), 57–159 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  4. Parisi, G.: Statistical Field Theory. Wiley, New York (1987)

    Google Scholar 

  5. Misawa, T., Yamaji, Y., Imada, M.: Tricritical behavior in charge-order system. J. Phys. Soc. Jpn. 75(6), 064705 (2006)

    Article  ADS  Google Scholar 

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

    Book  Google Scholar 

  7. Matsubara, T., Matsuda, H.: A lattice model of liquid helium, I. Progr. Theoret. Phys. 16(6), 569–582 (1956)

    Article  ADS  Google Scholar 

  8. Lacroix, C., Mendels, P., Mila, F. (eds.): Introduction to Frustrated Magnetism: Materials, Experiments, Theory. Springer, Heidelberg (2011)

    Google Scholar 

  9. Nagle, J.F.: Ising chain with competing interactions. Phys. Rev. A 2(5), 2124 (1970)

    Article  ADS  Google Scholar 

  10. Kardar, M.: Crossover to equivalent-neighbor multicritical behavior in arbitrary dimensions. Phys. Rev. B 28, 244 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  11. Mukamel, D., Ruffo, S., Schreiber, N.: Breaking of ergodicity and long relaxation times in systems with long-range interactions. Phys. Rev. Lett. 95(24), 240604 (2005)

    Article  ADS  Google Scholar 

  12. Campa, A., Gori, G., Hovhannisyan, V., Ruffo, S., Trombettoni, A.: Ising chains with competing interactions in the presence of long-range couplings. J. Phys. A 52, 344002 (2019)

    Article  MathSciNet  Google Scholar 

  13. Campa, A., Giansanti, A.: Canonical and microcanonical partition functions in long-range systems with order parameter space of arbitrary dimension. Physica A 340, 170–177 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  14. Leyvraz, F., Ruffo, S.: Ensemble inequivalence in systems with long-range interactions. J. Phys. A 35, 285 (2002)

    Article  ADS  MathSciNet  Google Scholar 

  15. Ruelle, D.: Statistical Mechanics: Rigorous Results. Benjamin, New York (1969)

    MATH  Google Scholar 

  16. Barré, J., Mukamel, D., Ruffo, S.: Inequivalence of ensembles in a system with long-range interactions. Phys. Rev. Lett. 87, 030601 (2001)

    Article  ADS  Google Scholar 

  17. Gori, G., Trombettoni, A.: The inverse Ising problem for one-dimensional chains with arbitrary finite-range couplings. J. Stat. Mech. P10021 (2011)

  18. Bouchet, F., Barré, J.: Classification of phase transitions and ensemble inequivalence, in systems with long range interactions. J. Stat. Phys. 118, 1073 (2005)

    Article  ADS  MathSciNet  Google Scholar 

Download references

Acknowledgements

Discussions with N. Defenu, D. Mukamel and N. Ananikian are gratefully acknowledged. A.C. acknowledges financial support from INFN (Istituto Nazionale di Fisica Nucleare) through the projects DYNSYSMATH and ENESMA. G.G. is supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy EXC 2181/1 - 390900948 (the Heidelberg STRUCTURES Excellence Cluster). The authors acknowledge support by the RA MES Science Committee and National Research Council of the Republic of Italy in the frames of the joint research Project No. SCS 19IT-008 and “Statistical Physics of Classical and Quantum Non Local Hamiltonians: Phase Diagrams and Renormalization Group”. This work is part of MUR-PRIN2017 Project “Coarse-grained description for non- equilibrium systems and transport phenomena (CO-NEST)” No. 201798CZL whose partial financial support is acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alessandro Campa.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Communicated by Roberto Livi.

Publisher's Note

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

Appendices

Appendix A: Derivation Using Large Deviation Techniques

The basic expressions of this paper, i.e., Eqs. (21), (22) and (23), or equivalently Eqs. (91), (92) and (93), for the rescaled free energy and the canonical and microcanonical entropies, respectively, have been obtained starting from the formal expressions (8) and (9). The latter have been adapted to the models with Hamiltonian of the type (14) obtaining the expressions (18) and (20). In this Appendix we show how one can arrive at the basic expressions by using an approach based on large deviation techniques. Of course, since one arrives at the same basic expressions for the rescaled free energy and the entropies, the analysis presented in section 3 remains identical. Here, for brevity, we show the procedure by starting directly from models with Hamiltonian of the type (14), therefore, without writing the more general expressions and then adapting them to that kind of Hamiltonian.

Using the definitions (1) and (15) of the magnetization \({\hat{m}}\) and the quadrupole moment \({\hat{q}}\), respectively, and introducing the definition

$$\begin{aligned} {\hat{r}} = \frac{1}{N}\sum _{i=1}^N U([S_i]) \, \end{aligned}$$
(126)

the Hamiltonian (14) can be written as

$$\begin{aligned} H(\{S_i\}) = N \left[ -\frac{J}{2} {\hat{m}}^2 -\frac{K}{2} {\hat{q}}^2 + {\hat{r}} \right] \, . \end{aligned}$$
(127)

At this point one formally defines

$$\begin{aligned} {\widehat{\varOmega }}(m,q,r,N)&\equiv \exp \left[ N {\widehat{s}}_{\mathrm{micr}}(m,q,r) \right] \nonumber \\&= \sum _{\{S_i\}} \delta \left( N{\hat{m}} -Nm \right) \, \delta \left( N{\hat{q}} -Nq \right) \, \delta \left( N{\hat{r}} -Nr \right) \,. \end{aligned}$$
(128)

One also defines the following kind of partition function

$$\begin{aligned} {\widehat{Z}}(\lambda _1,\lambda _2,\lambda _3,N)&\equiv \exp \left[ - N {\widehat{\phi }} (\lambda _1,\lambda _2,\lambda _3) \right] \nonumber \\&= \sum _{\{S_i\}} \exp \left[ -\lambda _1 {\hat{m}} -\lambda _2 {\hat{q}} -\lambda _3 {\hat{r}} \right] \, . \end{aligned}$$
(129)

In analogy with what we have noted for \(\beta \) in the main text, the fact that the energy is upper bounded allows to consider both signs for the parameters \(\lambda _1\), \(\lambda _2\) and \(\lambda _3\). It is not difficult to see that in the thermodynamic limit the function \({\widehat{s}}_{\mathrm{micr}}(m,q,r)\) and \({\widehat{\phi }}(\lambda _1,\lambda _2,\lambda _3)\) are related by the Legendre-Fenchel transformation

$$\begin{aligned} {\widehat{\phi }}(\lambda _1,\lambda _2,\lambda _3) = \min _{m,q,r} \left[ \lambda _1 m + \lambda _2 q + \lambda _3 r - {\widehat{s}}_{\mathrm{micr}}(m,q,r) \right] \, . \end{aligned}$$
(130)

In principle this transformation is not invertible. However, if \({\widehat{\phi }}\) is everywhere differentiable, the inversion is possible, so to have

$$\begin{aligned} {\widehat{s}}_{\mathrm{micr}}(m,q,r) = \min _{\lambda _1,\lambda _2,\lambda _3} \left[ \lambda _1 m + \lambda _2 q + \lambda _3 r - {\widehat{\phi }}(\lambda _1,\lambda _2,\lambda _3) \right] \, . \end{aligned}$$
(131)

It can be seen, in analogy with the function \({\hat{\psi }}\) defined in Eq. (19), that the function \({\widehat{\phi }}(\lambda _1,\lambda _2,\lambda _3)\) is differentiable (basically, this is assured from the fact that the expressions in the exponent in the right hand side of Eq. (129) are sums of one-particle functions). Then, Eq. (131) is verified. Once \({\widehat{s}}_{\mathrm{micr}}(m,q,r)\) is given, our function of interest, \({\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m)\) is obtained from

$$\begin{aligned} {\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m) = \max _{\left[ q,r | - \frac{J}{2}m^2 - \frac{K}{2} q^2 + r = \epsilon \right] } {\widehat{s}}_{\mathrm{micr}}(m,q,r) \, . \end{aligned}$$
(132)

It remains to see that from the last expression we can derive Eq. (93). Substituting Eq. (131) and expressing r as a function of \(\epsilon \), m and q, we have

$$\begin{aligned}&{\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m) \nonumber \\&\quad = \max _q \left\{ \min _{\lambda _1,\lambda _2,\lambda _3} \left[ \lambda _1 m + \lambda _2 q + \lambda _3 \left( \epsilon + \frac{J}{2}m^2 + \frac{K}{2}q^2\right) - {\widehat{\phi }}(\lambda _1,\lambda _2,\lambda _3) \right] \right\} .\nonumber \\ \end{aligned}$$
(133)

The four stationarity conditions of this problem are:

$$\begin{aligned} m - \frac{\partial {\widehat{\phi }}}{\partial \lambda _1}= & {} 0 \end{aligned}$$
(134)
$$\begin{aligned} q - \frac{\partial {\widehat{\phi }}}{\partial \lambda _2}= & {} 0 \end{aligned}$$
(135)
$$\begin{aligned} \epsilon + \frac{J}{2}m^2 + \frac{K}{2}q^2 - \frac{\partial {\widehat{\phi }}}{\partial \lambda _3}= & {} 0 \end{aligned}$$
(136)
$$\begin{aligned} \lambda _2 + \lambda _3 Kq= & {} 0 \, . \end{aligned}$$
(137)

Eliminating \(\lambda _2\) from the problem by implementing immediately the last stationarity condition, we have

$$\begin{aligned}&{\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m) \nonumber \\&\quad = \max _q \left\{ \min _{\lambda _1,\lambda _3} \left[ \lambda _1 m + \lambda _3 \left( \epsilon + \frac{J}{2}m^2 - \frac{K}{2}q^2\right) - {\widehat{\phi }}(\lambda _1,-\lambda _3 Kq,\lambda _3) \right] \right\} \, . \end{aligned}$$
(138)

After the notation changes \(\lambda _3 \rightarrow \beta \) and \(q \rightarrow y\), and defining \(\varphi \) by \(\lambda _1 = - \beta J m -\varphi \), the last expression becomes

$$\begin{aligned}&{\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m) \nonumber \\&\quad = \max _y \left\{ \min _{\beta ,\varphi } \left[ \lambda _1 m + \beta \epsilon -\beta \frac{J}{2}m^2 - \beta \frac{K}{2}q^2 - {\widehat{\phi }}(-\beta Jm -\varphi ,-\beta K y,\beta ) \right] \right\} \, . \nonumber \\ \end{aligned}$$
(139)

This is recognized to be the same as Eq. (93), once one realizes from the definitions (19) and (129) that \({\widehat{\phi }}(-\beta Jm -\varphi ,-\beta K y,\beta ) = {\hat{\psi }}(\beta ,\beta J m + \varphi ,\beta K y)\).

Appendix B: General Expressions for the Entropy Obtained with Direct Counting

In this Appendix we provide a summary of the general expressions, obtained with direct counting, for the entropy of one-dimensional Ising spin models having the form \(H=H_{LR}+H_{SR}\), where \(H_{LR}=-\frac{J}{2N}\left( \sum _i S_i \right) ^2\) is the long-range, mean-field term, and \(H_{SR}\) is the short-range part of the Hamiltonian. The interested reader can find full details in Ref. [17]. We emphasize that we are referring to the expressions of the entropy defined as the logarithm of the number of configurations for given values of the magnetization and of the spin correlations, like, e.g., the function \(s(m,g_1,g_2,t)\) in (119). From these expressions one can obtain the microcanonical entropy \({\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m)\) with the long optimization procedure described in Sect. 4.1.

In the main text we confined ourselves to the form (111), where only two-spin terms are included. The short-range Hamiltonian (111) is a sub-case of the general one-dimensional short-range Ising model with multispin interactions defined by

$$\begin{aligned} H_{SR}= & {} -\sum _{i} j^{(1)}_{i} S_{i} - \sum _{i,j} j^{(2)}_{i,j} S_{i} S_{j} - \sum _{i,j,k} j^{(3)}_{i,j,k} S_{i} S_{j} S_{k} \nonumber \\&- \sum _{i,j,k,l} j^{(4)}_{i,j,k,l} S_{i} S_{j} S_{k} S_{l} - \ldots \end{aligned}$$
(140)

where the sums run over distinct couples, triples, quartets and so on up to a certain finite range. Periodic boundary conditions are assumed and the couplings \(j^{(n)}\) are assumed to be invariant under translation by \(\rho \) spins:

$$\begin{aligned} j^{(n)}_{i_1,i_2,\ldots ,i_n}= j^{(n)}_{i_1+\rho ,i_2+\rho ,\ldots ,i_n+\rho } . \end{aligned}$$
(141)

As in the main text, N denotes the number of sites and R the finite-range of the interaction. For example, the Hamiltonian \(H_{SR}=-(K_1/2) \sum _i S_i S_{i+1}\) has \(\rho =1\) and \(R=2\), while \(H_{SR}=-(K_1/2) \sum _i S_i S_{i+1}-(K_2/2) \sum _i S_i S_{i+2}\) has \(\rho =1\) and \(R=3\). In the general case, for simplicity we assume \(N/\rho \) is an integer.

Following the notation and the procedure presented in [17], let us start from the case \(J=0\), for which \(H=H_{SR}\). To simplify the notation let us rewrite (140) as

$$\begin{aligned} H_{SR} \equiv - \sum _{\mathrm {Rg}(\mu )\le R}' \sum _{n=1}^{N/\rho } j_{\mu } O_{\mu +n \rho } (\{S_i\}) \, , \end{aligned}$$
(142)

where \(\mu \) is a subset of \(\{1,\ldots , R\}\). The notation \(\mathrm {Rg}(\mu )\le R\), stands for “the range of the interaction is less than or equal to R”. Moreover, \(O_{\mu +n \rho }\) is an operator associated to the subset \(\mu \equiv \{n_{1}, n_{2}, \ldots n_{|\mu |}\}\), where \(|\mu |\) is the number of elements of \(\mu \), and translated by \( n\rho \) so that it acts on the spins as

$$\begin{aligned} O_{\mu + n\rho } (\{S_i\}) =S_{n_{1} + n \rho } S_{n_{2} + n \rho } \ldots S_{n_{|\mu |} + n \rho }. \end{aligned}$$
(143)

For the null subset \(\varnothing \) we define \(O_{\varnothing }=1\) and the prime \('\) in the sum over \(\mu \) in (142) denotes that the null subset is not included and that the terms related by a translation of a multiple of \(\rho \) are counted only once. The correlation functions are denoted by \(g_{\mu }\). They are associated to the operator \(O_{\mu }\) and defined according to

$$\begin{aligned} g_{\mu }=\langle O_{\mu }(\{S_i\})\rangle = \langle S_{n_{1}} S_{n_{2}} \ldots S_{n_{|\mu |}} \rangle \, \end{aligned}$$
(144)

(by definition, \(g_{\varnothing }=1\)). For example, for the Hamiltonian considered in Sect. 4.1 we would have the correlation functions \(g_{\{1\}}\), \(g_{\{1,2\}}\), \(g_{\{1,3\}}\) and \(g_{\{1,2,3\}}\), that in the lighter notation of the main text were denoted, respectively, with m, \(g_1\), \(g_2\) and t.

The main result of Ref. [17] concerns the entropy \(s(\{g_{\mu }\})\) for the model with interactions up to range R; it is given by:

$$\begin{aligned} s(\{g_{\mu }\})=s^{(R)}(\{g_{\mu }\})-s^{(R-\rho )}(\{g_{\mu }\}) \, , \end{aligned}$$
(145)

where in the left side and in the first term in the right hand side \(\{g_{\mu }\}\) stands for the set of all possible correlations of range up to R, while in the seond term in the right hand side it stands for all correlations of range up to \((R-\rho )\). It is written in terms of the functions \(s^{(Q)}(\{g_{\mu }\})\). The quantity \(s^{(Q)}\) can be seen as the “entropy at range Q”, and is given by

$$\begin{aligned} s^{(Q)}(\{g_{\mu }\})=-\sum _{\tau _{Q}} p(\tau _{Q}) \ln p( \tau _{Q}) \, , \end{aligned}$$
(146)

where

$$\begin{aligned} p(\tau _{Q})=2^{-Q}\sum _{\mathrm {Rg}(\mu )\le Q} g_{\mu } O_{\mu } ( \tau _{Q}) \, , \end{aligned}$$
(147)

and \(\tau _Q \equiv \{t_1, t_2, \ldots , t_Q\}\) denotes the configuration of Q Ising spins, with the sum over \(\mu \) is on every subset (including the null one). Finally, from (142) one gets the energy per unit cell as:

$$\begin{aligned} e(\{g_{\mu }\})=-\sum _{1 \le \mathrm {Rg}(\mu )\le R} g_{\mu } j_{\mu } \, . \end{aligned}$$
(148)

When the mean-field term \(H_{LR}\) is turned on, only the energy e is affected, while the dependence of the entropy \(s(\{g_{\mu }\})\) on the correlations \(g_{\mu }\) is not. One then has to add the corresponding contribution to e. In this way one finds the results (114) and (116), respectively for s and for e, for the model (110) with \(K_2=0\); and the results (119) and (125) for the same model with \(K_2 \ne 0\). In particular, specializing (145) to our model with \(K_2=0\) one has that Eq. (114) is obtained from \(s(m,g_1) = s^{(2)}(m,g_1) - s^{(1)}(m)\), while Eq. (119) for the model with \(K_2 \ne 0\) is obtained from \(s(m,g_1,g_2,t) = s^{(3)}(m,g_1,g_2,t) - s^{(2)}(m,g_1)\).

Appendix C: Comparison of the Microcanonical Entropy at Fixed Magnetization for \(K_2 \ne 0\)

In this Appendix we consider an example of explicit determination of the microcanonical entropy at fixed magnetization directly from the entropy \(s=s(m,g_1,g_2,t)\), given in Eq. (119), for the model (110). To compare the findings with the results obtained with the method presented in section 3, we choose the same values of \(K_1\) and \(K_2\) used in Fig. 5: \(K_1=-0.4\), \(K_2=-0.16\) (with \(J=1\)). The energy is chosen as \(\epsilon =-0.107\), as in the bottom right panel of Fig. 5.

As discussed in the main text, one has to determine t via Eqs. (120)–(121). Once this is done, one has to express \(g_2\) as a function of \(m,g_1\) using the energy expression (125). One has then s as a function of m and \(g_1\) and it is possible to plot in the \(m-g_1\) plane the allowed regions. Of course the same procedure can be performed by studying the entropy in the \(m-g_2\) plane. The final point in both cases is to find the maximum of the microcanonical entropy maximizing with respect to, respectively, \(g_1\) or \(g_2\).

Fig. 7
figure 7

Three-dimensional plot of the allowed region in the \((m,g_1,g_2)\) space (left). On the right the allowed region has been cut with the constant energy surface (a parabolic cylinder) given by \(\epsilon = -\frac{1}{2}(J m^2+K_1 g_1+K_2 g_2)\). The green surface is thus the accessible region in the microcanonical ensemble. The chosen values are \(\epsilon =-0.107\), \(K_1=-0.4\) and \(K_2=-0.16\). Please note that in order to improve visibility the \(g_2\) axis has been reversed (Color figure online)

Notice that in this procedure finding the maximum with respect to \(g_1\), \(g_2\) and t is the easier part since the entropy is concave along these directions on the constant energy surface. In the remaining variable m instead, within the constant energy surface, the entropy is not concave and many entropy maxima can and do appear and compete resulting in the emergence of the different phases and transitions among them. This is obviously to be traced to the special role of m in the Hamiltonian, in which it appears nonlinearly and thus can spontaneously break the \(m\rightarrow -m\) symmetry. Restricting to fixed magnetization indeed relieves many of the difficulties. The procedure is described in Figs. 7 and 8, where we consider the value \(m=0.55\), and one finds that the maximum entropy is \(s\simeq 0.47828\), in agreement with the results presented in the bottom right panel of Fig. 5, obtained with the procedure described in section 3. Note that in the course of the process we also determine the macroscopic observables fully characterizing the thermodynamic state. Fig. 9 also shows that if we decide to eliminate \(g_2\) in favour of \((m,g_1)\) or, alternatively, \(g_1\) in favour of \((m,g_2)\), we obtain the same result for the microcanonical entropy at fixed magnetization when the maximum in, respectively, \(g_1\) or \(g_2\) is taken, as of course it has to be. For completeness we also plot the entropy as a function of the correlation t after maximizing with respect to one among \(g_1\) and \(g_2\).

The direct counting method outlined shows some technical difficulties due to the already moderately large number of variables over which the entropy has to be optimized. Of course these extra variables are interesting in their own right being macroscopic observables fully characterising the thermodynamic state. On the other hand direct counting possesses the virtue of making very clear the geometric origin of (microcanonical) phase transitions in long-range systems as the study of the maxima of the entropy restricted on the nonlinear energy surface. This makes interesting also short-range one-dimensional systems, whose entropy is concave in all variables, yielding normally no phase transition. The gained insight could prove useful in the understanding of and the hunt for the many exotic critical points expected in the microcanonical ensemble [18].

Fig. 8
figure 8

Entropy as a function of m and \(g_1\) (left) and m and \(g_2\) (right) for the parameter values \(\epsilon =-0.107\), \(K_1=-0.4\) and \(K_2=-0.16\). The line with magnetization \(m=0.55\) (chosen to have an example of comparison with Fig. 5) is denoted with a black line. The maximum of the entropy s in this fixed magnetization sector is denoted with a black dot. It is characterised by the observables \(m=0.55\), \(g_1 \simeq 0.13235\), \(g_2 \simeq 0.22226\) and \(t \simeq -0.16384\). The global maxima of s are also shown with a red dot. These points are characterized by the following observables: \(m \simeq \pm 0.39839\), \(g_1 \simeq -0.15375\), \(g_2 \simeq 0.03885\), and \(t\simeq -0.46475\) (Color figure online)

Fig. 9
figure 9

Entropy in the fixed energy (\(\epsilon =-0.107\)) and magnetization (\(m=0.55\)) sector as a function of the three independent correlations \(g_1\) (blue), \(g_2\) (green) and t (red) when the other two are eliminated. Parameter values are \(K_1=-0.4\) and \(K_2=-0.16\). The maximum \({\widetilde{s}}_\mathrm {micr}\) is denoted with a dashed line and it occurs at \(g_1 \simeq 0.13235\), \(g_2 \simeq 0.22226\), and \(t\simeq -0.16384\). The value of the maximum, \({\widetilde{s}}_\mathrm {micr} \simeq 0.47828\), corresponds to the value of \({\widetilde{s}}_{\mathrm{micr}}(\epsilon ,m)\) at \(m=0.55\) in the right bottom panel of Fig. 5 (Color figure online)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Campa, A., Gori, G., Hovhannisyan, V. et al. Computation of Microcanonical Entropy at Fixed Magnetization Without Direct Counting. J Stat Phys 184, 21 (2021). https://doi.org/10.1007/s10955-021-02809-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-021-02809-y

Keywords

Mathematics Subject Classification

Navigation