Skip to main content
Log in

Estimating the Asymptotics of Solid Partitions

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

Abstract

We study the asymptotic behavior of solid partitions using transition matrix Monte Carlo simulations. If \(p_3(n)\) denotes the number of solid partitions of an integer \(n\), we show that \(\lim _{n\rightarrow \infty } n^{-3/4} \log p_3(n)\sim 1.822\pm 0.001\). This shows clear deviation from the value \(1.7898\), attained by MacMahon numbers \(m_3(n)\), that was conjectured to hold for solid partitions as well. In addition, we find estimates for other sub-leading terms in \(\log p_3(n)\). In a pattern deviating from the asymptotics of line and plane partitions, we need to add an oscillatory term in addition to the obvious sub-leading terms. The period of the oscillatory term is proportional to \(n^{1/4}\), the natural scale in the problem. This new oscillatory term might shed some insight into why partitions in dimensions greater than two do not admit a simple generating function.

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. The solid partition graph that we defined is a Bratelli diagrams provided we add the null partition as the zero vertex [23]. For ordinary partitions, the analogous graph (also called the Young’s lattice) is the Bratteli diagram of irreducible representations of the symmetric group [24].

  2. More generally, for a \(d\)-dimensional partition, the natural scale is \(\ell =n^{1/(d+1)}\).

  3. The fits were carried out using Mathematica’s NonlinearModelFit using weights given by the statistical error when possible. We have also used FindFit for fits without weights.

  4. This term corresponds to adding \(f\ n^{-1/4}\) to Eq. (3.7).

  5. The statistical noise in \(n\delta \) is given by \(n\) times the estimated statistical error given in Fig. 1. Thus it is around \(0.13\) for \(n=10^4\) but around \(0.02\) for \(n=6^4\). For \(n^{1/4}<4.5\), the magnitude of the oscillations is larger than the statistical noise.

  6. This term corresponds to adding \(f^{-1/3}\) to Eq. (5.2).

References

  1. Andrews, G.E.: The Theory of Partitions, vol. 2. Cambridge University Press, Cambridge (1998)

    MATH  Google Scholar 

  2. Almkvist, G.: A rather exact formula for the number of plane partitions. Contemp. Math. 143, 21–26 (1993)

    Article  MathSciNet  Google Scholar 

  3. Almkvist, G.: Asymptotic formulas and generalized Dedekind sums. J. Exp. Math. 7, 343–359 (1998)

    Article  MATH  MathSciNet  Google Scholar 

  4. Govindarajan, S., Prabhakar, N.S.: A superasymptotic formula for the number of plane partitions, arXiv preprint arXiv:1311.7227 (2013)

  5. Mutafchiev, L., Kamenov, E.: Asymptotic formula for the number of plane partitions of positive integers. Compt. Rend. Acad. Bulg. Sci. 59, 361–366 (2006)

    MATH  MathSciNet  Google Scholar 

  6. Wright, E.M.: Asymptotic partition formulae I. Plane partitions. Q. J. Math. Oxford, Ser. 2, 177–189 (1931)

    Article  Google Scholar 

  7. Govindarajan, S., Balakrishnan, S.: The solid partitions project, http://boltzmann.wikidot.com/solid-partitions

  8. Mustonen, V., Rajesh, R.: Numerical estimation of the asymptotic behaviour of solid partitions of an integer. J. Phys. A 36, 6651–6659 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  9. MacMahon, P.A.: Combinatory Analysis. Cambridge University Press, Cambridge (1916)

    Google Scholar 

  10. Levine, D., Steinhardt, P.J.: Quasicrystals: a new class of ordered structures. Phys. Rev. Lett. 53, 2477–2480 (1984)

    Article  ADS  Google Scholar 

  11. Elser, V.: Comment on “Quasicrystals: A New Class of Ordered Structures”. Phys. Rev. Lett. 54, 1730 (1985)

    Article  ADS  Google Scholar 

  12. Mosseri, R., Bailly, F.: Configurational entropy in octagonal tiling models. Int. J. Mod. Phys. B 7, 1427–1436 (1993)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  13. Destainville, N., Mosseri, R., Bailly, F.: Fixed-boundary octagonal random tilings: a combinatorial approach. J. Stat. Phys. 102, 147–190 (2001)

    Article  MATH  MathSciNet  Google Scholar 

  14. Widom, M., Mosseri, R., Destainville, N., Bailly, F.: Arctic octahedron in three-dimensional rhombus tilings and related integer solid partitions. J. Stat. Phys. 109, 945–965 (2002)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  15. Destainville, N., Mosseri, R., Bailly, F.: A formula for the number of tilings of an octagon by rhombi. Theor. Comput. Sci. 319, 71–81 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  16. Destainville, N., Widom, M., Mosseri, R., Bailly, F.: Random tilings of high symmetry: I. Mean-field theory. J. Stat. Phys. 120, 799–835 (2005)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  17. Hutchinson, M., Widom, M.: Enumeration of octagonal tilings, arXiv:1306.5977 [math.CO] (2013)

  18. Vidal, J., Destainville, N., Mosseri, R.: Quantum dynamics in high codimension tilings: from quasiperiodicity to disorder. Phys. Rev. B 68, 172202 (2003)

    Article  ADS  Google Scholar 

  19. Gopakumar, R., Vafa, C.: M-Theory and Topological Strings-I, ariXiv:hep-th/9809187 (1998)

  20. Gopakumar, R., Vafa, C.: M-Theory and Topological Strings-II, arXiv:hep-th/9812127 (1998)

  21. Behrend, K., Bryan, J., Szendröi, B.: Motivic degree zero Donaldson–Thomas invariants. Invent. Math. 192, 111–160 (2013)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  22. Balakrishnan, S., Govindarajan, S., Prabhakar, N.S.: On the asymptotics of higher-dimensional partitions. J. Phys. A 45, 055001 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  23. Bratteli, O.: Inductive limits of finite dimensional C*-algebras. Trans. Am. Math. Soc. 171, 195–234 (1972)

    MATH  MathSciNet  Google Scholar 

  24. Sagan, B.E.: The Symmetric Group. Wadsworth and Brooks/Cole, Pacific Grove (1991)

    MATH  Google Scholar 

  25. Atkin, A.O.L., Bratley, P., MacDonald, I.G., McKay, K.S.: Some computations for m-dimensional partitions. Proc. Camb. Philos. Soc. 63, 1097–1100 (1967)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  26. Bhatia, D.P., Prasad, M.A., Arora, D.: Asymptotic results for the number of multidimensional partitions of an integer and directed compact lattice animals. J. Phys. A 30, 2281–2285 (1997)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  27. Bratley, P., McKay, J.K.S.: Algorithm 313: Multi-dimensional partition generator. Commun. ACM 10, 666 (1967)

    Article  Google Scholar 

  28. Erdos, P., Lehner, J.: The distribution of the number of summands in the partitions of a positive integer. Duke Math. 8, 335–345 (1941)

    Article  MathSciNet  Google Scholar 

  29. Cerf, R., Kenyon, R.: The low temperature expansion of the Wulff crystal in the 3D Ising model. Comm. Math. Phys. 222, 147–179 (2001)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  30. Kenyon, R., Okounkov, A., Sheffield, S.: Dimers and Amoebae. Ann. Math. 163, 1019–1056 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  31. de Bruijn, N.G.: Algebraic theory of Penrose’s nonperiodic tilings of the plane. I. Neder. Akad. Wetensch. Indag. Math. 43, 39–52 (1981)

    Article  MATH  Google Scholar 

  32. de Bruijn, N.G.: Algebraic theory of Penrose’s nonperiodic tilings of the plane. II. Neder. Akad. Wetensch. Indag. Math. 43, 53–66 (1981)

    Article  MATH  Google Scholar 

  33. Henley, C.L.: Random tiling models. In: Di Vincenzo, D.P., Steingart, P.J. (eds.) Quasicrystals, the State of the Art, p. 429. World Scientific, Singapore (1991)

    Chapter  Google Scholar 

  34. Destainville, N.: Entropy and boundary conditions in random rhombus tilings. J. Phys. A 31, 6123–6139 (1998)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  35. Björner, A., Stanley, R.P.: A combinatorial miscellany, L’enseignement mathématique, Monograph no. 42, Genève, (2010)

  36. Linde, J., Moore, C., Nordahl, M.G.: An \(n\)-dimensional generalization of the rhombus tiling. In: Proceedings of the 1st International conference on Discrete Models: Combinatorics, Computation, and Geometry (DM-CCG+01), M. Morvan, R. Cori, J. Mazoyer and R. Mosseri, eds., Discrete Math. Theo. Comp. Sc. AA:23 (2001)

Download references

Acknowledgments

SG would like to thank Intel India for financially supporting the numerical study of solid partitions. We also thank the High Performance Computing Environment at IIT Madras which provided us access to the Virgo and Vega super clusters where our Monte Carlo simulations were carried out.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nicolas Destainville.

Appendices

Appendix: Asymptotics of Plane Partitions

The asymptotics of plane plane partitions as follows from an application of Meinardus’ formula is [5, 6]

$$\begin{aligned} p_2(n) \sim \frac{(2\zeta (3))^{7/36}}{\sqrt{6\pi }} n^{-25/36} \exp \Big ( \tfrac{3}{2} (2\zeta (3))^{1/3} n^{2/3} + \zeta '(-1) \Big ) \end{aligned}$$

Thus, one has

$$\begin{aligned} n^{-2/3}\log p_2(n)&\sim \tfrac{3}{2} (2\zeta (3))^{1/3} + n^{-2/3}\Big (-\tfrac{25}{36} \log n + \log \tfrac{(2\zeta (3))^{7/36}}{\sqrt{6\pi }} + \zeta '(-1)\Big )\nonumber \nonumber \\&\sim 2.00945 -0.694444 n^{-2/3}\log n -1.4631 n^{-2/3} \end{aligned}$$
(5.1)

The goal of the Monte Carlo simulation is to reproduce the above formula. In particular, we should see if we can match the constant term to one part in \(10^3\).

Results from Monte Carlo Simulations

The Monte Carlo simulations for plane partitions are used to estimate the numbers of plane partitions, say, in the range \([1,N_{max}]\). We then compare these numbers to the exact numbers and see how the numbers improve with increasing the statistics by increasing the number of flips. Let \(mc_2(n)\) be the values obtained from the Monte Carlo simulation. We also fit \(n^{-2/3}\log p_2(n)\) in the range \([50,2000]\) with the following formula:

$$\begin{aligned} n^{-2/3} \log p_2(n)\sim \alpha _2\ + \beta _2\ n^{-1/3} + \delta _2\ n^{-2/3} \log n + \epsilon _2\ n^{-2/3}. \end{aligned}$$
(5.2)

Since our Monte Carlo simulations estimate \(N_\pm (n)\), we use

$$\begin{aligned} \log \big [\tfrac{N_+(n-1)}{N_-(n)}\big ] \sim&\ \ \alpha _2\ \big [n^{2/3}\big ]_2 + \beta _2 \ \big [n^{1/3}\big ]_2 + \delta _2\ \big [ \log n\big ]_2 \ , \end{aligned}$$
(5.3)

to extract three of the four parameters \((\alpha _2,\beta _2,\delta _2,\epsilon _2)\) and then determine \(\epsilon _2\) using Eq. (5.2).

  1. 1.

    Fit data in the range \(n\in [50,2000]\) to the three-parameter formula given in Eq. (5.3). This determines the values of \((\alpha _2,\beta _2,\delta _2)\) that we will use below. We obtain

    $$\begin{aligned} (\alpha _2,\beta _2,\delta _2) = (2.00998, -0.0194366, -0.663683). \end{aligned}$$
    (5.4)
  2. 2.

    Next we add the term \(\big [-\tfrac{1}{3} f\ n^{-4/3}\big ]\) to the asymptotic formula Eq. (5.3) and carry out a four-parameter fit to see how the three parameters change.Footnote 6 We obtain

    $$\begin{aligned} (\alpha _2,\beta _2,,\delta _2,f) = (2.00923,+0.0120124,-0.731848,-0.41608). \end{aligned}$$
    (5.5)

    We take the average of the two sets of numbers and use one half of the difference as an estimate of the error.

    $$\begin{aligned} (\alpha _2,\beta _2,\delta _2,f) = (2.0096\pm 0.0004, -0.004\pm 0.02,-0.70\pm 0.03,-0.21\pm 0.21).\nonumber \\ \end{aligned}$$
    (5.6)
  3. 3.

    We substitute the values of \((\alpha _2,\beta _2,\delta _2)\) given in Eq. (5.4) in Eq. (5.2) and then carrying out a one-parameter fit for \(n\in [50,100]\) to determine \(\epsilon _3\). We obtain \(\epsilon _2 = -1.44372\). We then add a term \(f\ n^{-1}\) to Eq. (5.2) and use values given in Eq. (5.5) and carry out a one-parameter fit determine \(\epsilon _2\). We obtain \(\epsilon _2=-1.39798\). We thus obtain the estimate

    $$\begin{aligned} \epsilon _2=-1.423\pm 0.025. \end{aligned}$$
    (5.7)

The error estimates for the four parameters through the Monte Carlo simulations are consistent with the deviation from the exact values. This provides some validation for the methods that we used for solid partitions.

No Oscillations for Plane Partitions

In order to see if the oscillations that we observe are special to solid partitions, we study the residual in the Monte Carlo data for plane partitions. Since our Monte Carlo simulations did not give us estimates for the statistical errors, we used the difference of the exact numbers for plane partitions from the numbers from our Monte Carlo simulations to provide an estimate of the statistical error. In Fig. 7, we observe that both the residual and our estimated statistical errors have similar behaviour that is consistent with no oscillations.

Fig. 7
figure 7

On the left is the plot of \(n\) times the estimated statistical error versus \(n^{1/3}\). On the right we plot the \(n\) times the residual (in blue) to the fit. There are no oscillations to be seen and the residual is comparable to the statistical error (Color figure online)

A similar study on exactly enumerated plane partitions (through Mac Mahon’s formula) led us to the same conclusion.

Asymptotics of MacMahon Numbers

Recall that the formula that MacMahon guessed for the generating function of solid partitions gives rise to a series of numbers that we call MacMahon numbers, \(m_{3}(n)\). One has

$$\begin{aligned} {\prod _{n=1}^\infty (1-q^n)^{-\tfrac{n(n+1)}{2}}}:=\sum _{n=0}^\infty m_3(n)\ q^n. \end{aligned}$$
(5.8)

In this section, we briefly discuss the asymptotics of these numbers in order to compare with our results for solid partitions. While these numbers do not have any relation to solid partitions, we can derive their asymptotics from the above product formula using Meinardus’ method. One obtains [22]

$$\begin{aligned} n^{-3/4}\log m_3(n) \sim 1.78982 + \frac{0.333546}{n^{ 1/4}} - \frac{0.0414393}{\sqrt{n}} + \frac{(-1.54436 - 0.635417 \log n)}{n^{3/4}} \end{aligned}$$
(5.9)

Assuming that their asymptotic behavior is similar to that of solid partitions, one can see how well fits to, say the first 1000 MacMahon numbers, agree with the exact asymptotic formula. The accuracy of these fits will provide some hints towards the quality of the fits that one may expect for fits using Monte Carlo data. The fit has five parameters given by the formula

$$\begin{aligned} n^{-3/4}\log m_3(n) \sim a+ \frac{b}{n^{ 1/4}} + \frac{c}{\sqrt{n}} + \frac{(e + d \log n)}{n^{3/4}} \end{aligned}$$
(5.10)

We observe that the the first 2,000 numbers reproduce the leading constant in the asymptotic formula with an error of \(0.0005\) (Table 2). Our numerical study of solid partitions has used data corresponding to the first 10,200 numbers but achieves a slightly lower accuracy.

Table 2 Results of the fit to the exact MacMahon numbers \(m_3(n)\) in the range \([50,N_{max}]\)

Oscillations of Boxed Solid Partitions Near the Entropy Maximum

So far we essentially focussed on unbounded partitions of an integer, in other words partitions restricted to a box of lateral size \(B\gg n^{1/4}\). We needed to have a box for computational reasons, but we have discussed that the box has essentially no incidence. In this context, we have identified original oscillations of the residuals \(\delta \), with a period proportional to the natural scale \(n^{1/4}\), which we have related to the unexpected sensitivity to the underlying lattice. Anticipating that the same phenomenon might have a similar signature for bounded partitions [i.e. partitions restricted to a box of size \(B=\mathcal {O}(n^{1/4})\)], we have re-examined the numerical data from Ref. [14] as follows. For given values of \(B\) and of \(n\), there are \(p_3(B,n)\) partitions of \(n\) such that all coordinates of all nodes are \(\le B\). We then define the partial entropy \(S(n) \equiv \log p_3(B,n)\) and we explore its behavior near its maximum (at \(n=B^4/2\)), by opposition to the limit \(n \ll B^4\) (or by symmetry \(B^4-n \ll B^4\)) studied so far. Near this maximum, it has been conjectured that the amoeba of Fig. 5 (Right) becomes a regular “arctic” octahedron [14, 36].

The asymptotic expansion, Eq. (3.7), is replaced by the fit of \(S(n)\) near its maximum by a (somewhat arbitrarily) fourth-order polynomial, and the resulting residual is displayed in Fig. 8 for \(B=10\). Even though it will have to be confirmed in future studies, this figure suggests they there exist oscillations. By analogy with our above findings, we anticipate that there period should be \(2B^3/3\). Indeed, as discussed in [14], near the entropy maximum, the values of \(k\) for which the integers \(X_k\) (as defined in Sects. 2 and 3.2) are generically non-vanishing define a finite subset of \([0,B-1]^3\) (called the “slab” in Ref. [14]) of cardinality \(2B^3/3\). This “slab” is the projection of the “arctic” octahedron on \([0,B-1]^3\). Increasing (resp. decreasing) all the integers \(X_k\) in the slab by 1 thus leads to an increase (resp. decrease) of \(n\equiv \sum _{k\in [0,B-1]^3} X_k\) by \(2B^3/3\). The signature of the underlying lattice is thus expected in this case to lead to a sub-sub-dominant correction of period \(2B^3/3\) to the entropy \(S(n)\), which is indeed suggested by the figure.

Fig. 8
figure 8

Residuals between the numerical entropy \(S(n)\) and its fourth-order fit on the interval \([3000,7000]\) for \(B=10\); the black sinusoid is a guide for eyes and has period \(2 B^3/3\simeq 667\). Numerical data were kindly provided by the authors of Ref. [14]

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Destainville, N., Govindarajan, S. Estimating the Asymptotics of Solid Partitions. J Stat Phys 158, 950–967 (2015). https://doi.org/10.1007/s10955-014-1147-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-014-1147-z

Keywords

Navigation