Skip to main content
Log in

A Correction to a Remark in a Paper by Procacci and Yuhjtman: New Lower Bounds for the Convergence Radius of the Virial Series

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

Abstract

In this note we deduce a new lower bound for the convergence radius of the Virial series of a continuous system of classical particles interacting via a stable and tempered pair potential using the estimates on the Mayer coefficients obtained in the recent paper by Procacci and Yuhjtman (Lett Math Phys 107:31–46, 2017). This corrects the wrongly optimistic lower bound for the same radius claimed (but not proved) in the above cited paper (in Remark 2 below Theorem 1). The lower bound for the convergence radius of the Virial series provided here represents a strong improvement on the classical estimate given by Lebowitz and Penrose in 1964.

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.

References

  1. Basuev, A.G.: A theorem on minimal specific energy for classical systems. Teoret. Mat. Fiz. 37(1), 130–134 (1978)

    MathSciNet  Google Scholar 

  2. Jones, J.E., Ingham, A.E.: On the calculation of certain crystal potential constants, and on the cubic crystal of least potential energy. Proc. R. Soc. Lond. A 107, 636–653 (1925)

    Article  ADS  Google Scholar 

  3. Lebowitz, J.L., Penrose, O.: Convergence of virial expansions. J. Math. Phys. 7, 841–847 (1964)

    Article  ADS  MathSciNet  Google Scholar 

  4. Pathria, R.K., Beale, P.D.: Statistical Mechanics, 3rd edn. Elsevier, Amsterdam (2011)

    MATH  Google Scholar 

  5. Penrose, O.: Convergence of fugacity expansions for fluids and lattice gases. J. Math. Phys. 4, 1312 (1963)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Penrose, O.: Convergence of fugacity expansions for classical systems. In: Bak, A. (ed.) Statistical Mechanics: Foundations and Applications. Benjamin, New York (1967)

    Google Scholar 

  7. Procacci, A., Yuhjtman, S.A.: Convergence of Mayer and Virial expansions and the Penrose tree-graph identity. Lett. Math. Phys. 107, 31–46 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Ruelle, D.: Correlation functions of classical gases. Ann. Phys. 5, 109–120 (1963)

    Article  ADS  MathSciNet  Google Scholar 

  9. Torquato, S., Jiao, Y.: Effect of dimensionality on the percolation thresholds of various d-dimensional lattices. Phys. Rev. E 87, 032149 (2013)

    Article  ADS  Google Scholar 

  10. Ueltschi, D.: An improved tree-graph bound. Oberwolfach Reports (2017) (to appear). arXiv:1705.05353

Download references

Acknowledgements

It is a pleasure to thank Sergio Yuhjtman for his reading of the manuscript and his useful observations and suggestions. This work has been partially supported by the Brazilian agency CNPq (Conselho Nacional de Desenvolvimento Científico e Tecnológico - Bolsa de Produtividade em pesquisa, grant n. 306208/2014-8).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Aldo Procacci.

Appendices

Appendix A: Proof of (2.7)

Estimate (2.7) is basically the bound (1.2) proved in [7] with the unique difference that the factor \(e^{\beta \bar{B}(n-1)}\) replaces the factor \(e^{\beta B n}\). Of course, since \(\bar{B}\ge B\), this is a bad deal for someone interested in an upper bound the Mayer series. On the other hand, as shown in Sec. 3, the use (2.7) instead of (1.2) happens to be a good deal towards an upper bound the Virial series. Inequality (2.7) relies on two lemmas originally proved in [7] (see there Lemma 1 and Lemma 2). For completeness we report these lemmas and their proofs here below.

The first of these two lemmas involves the concept of partition scheme, which, we remind, is a map \(\varvec{M}\) from the set \(T_n\) of the labeled trees in [n] to the set \(G_n\) of the connected graphs in [n] such that \(G_n=\biguplus _{\tau \in {T_n}}[\tau ,{\varvec{M}}(\tau )]\) with \(\biguplus \) disjoint union and \([\tau ,\varvec{M}(\tau )]=\{g\in G_n: \tau \subseteq g\subseteq \varvec{M}(\tau )\}\).

Lemma 1

For fixed V and \((x_1,\dots , x_n)\in \mathbb {R}^{dn}\), choose a total order \(\succ \) in the set of edges \(E_n\) of the complete graph \(K_n\) in such a way that \(\{i,j\}\succ \{k,l\}\Longrightarrow V(x_i-x_j)\ge V(x_k-x_l)\) and let \(\varvec{T}:G_n\rightarrow T_n\) be the map that associates to \(g\in G_n\) the tree \(\varvec{T}(g)\in T_n\) constructed by starting from \(\emptyset \) and keeping adding the lowest edge in g that does not create a cycle (Kruskal algorithm).

Let \(\varvec{M}: T_n\rightarrow G_n\) be the map that associates to \(\tau \in T_n\) the graph \(\varvec{M}( \tau )\in G_n\) whose edges are the \(\{i,j\}\in E_n\) such that \(\{i,j\} \succeq \{k,l\} \) for every edge \(\{k,l\} \in E_\tau \) belonging to the path from i to j through \(\tau \).

Then \(\varvec{T}^{-1}(\tau )=\{g \in {\mathcal G}_n :\, \tau \subseteq g \subseteq \varvec{M}(\tau )\}\) and therefore \(\varvec{M}\) is a partition scheme in \({G}_n\).

Proof

Assume first that \(g \in \varvec{T}^{-1}(\tau )\). Then \(\tau =\varvec{T}(g) \subset g\). Now take \(\{i,j\} \in E_g\), and let \(e \in E_\tau \) be any edge belonging to the path from i to j in \(\tau \). Consider the tree \(\tau '\) obtained from \(\tau \) after replacing the edge e by \(\{i,j\}\). By minimality of \(\tau \) we must have \( \{i,j\}\succ e\), i.e. \(\{i,j\} \in E_{\varvec{M}(\tau )}\), whence \(g \subset \varvec{M}(\tau )\). Conversely, let \(\tau \subset g \subset \varvec{M}(\tau )\). We must show \(\varvec{T}(g)=\tau \). By contradiction, take \(\{i,j\} \in E_{\varvec{T}(g)} \setminus E_\tau \). Consider the path \(p^\tau (\{i,j\})\) in \(\tau \) joining i with j. Since \(\varvec{T}(g) \subset \varvec{M}(\tau )\), \(\{i,j\}\) is greater (w.r.t. \(\succ \)) than any edge in the path \(p^\tau (\{i,j\})\). If we remove \(\{i,j\}\) from \(\varvec{T}(g)\), the tree \(\varvec{T}(g)\) splits into two trees. Necessarily, one of the edges in the path \(p^\tau (\{i,j\})\) joins a vertex of one tree with a vertex of the other. Thus, by adding this edge we get a new tree which contradicts the minimality of \(\varvec{T}(g)\). \(\square \)

Remark

The proof of Lemma 1 above, as well as the definition of the partition scheme \(\varvec{M}\), are identical to those given in the homonymous lemma of [7], but the map \(\varvec{T}: G_n\rightarrow T_n\) appearing in the enunciate, based on Kruskal’s algorithm, replaces a similar (but more involved) map constructed in [7] via so-called admissible functions with values in a tomonoid. The idea to use the Kruskal’s algorithm, which eases the definition of the “minimal tree” map \(\varvec{T}\), has been suggested during an Oberwolfach meeting by David Brydges, Tyler Helmuth and Daniel Ueltschi (see [10]).

Using the partition scheme \(\varvec{M}\) defined in Lemma 1 we now state and prove the second key lemma of [7].

Lemma 2

Let V be stable with Basuev stability constant \(\bar{B}\), and let \(\tau \in { T}_n\). Let \((x_1,...,x_n) \in {\mathbb R}^{dn}\) and let \(\varvec{M}\) be the partition scheme given above, then

$$\begin{aligned} \sum _{\{i,j\} \in \varvec{E}_{M(\tau )} \setminus E_\tau ^+} v(x_i-x_j) \ge -\bar{B}(n-1) \end{aligned}$$
(2.20)

Proof

. The set of edges \(E_\tau \setminus E_\tau ^+\) forms the forest \(\{\tau _1,...,\tau _k\}\). Let us denote \({ V}_{\tau _s}\) the vertex set of the tree \(\tau _s\) from the forest. Assume \(i \in { V}_{\tau _a}\), \(j \in {V}_{\tau _b}\). If \(a \ne b\), the path from i to j through \(\tau \) involves an edge \(\{k,l\}\) in \(E_\tau ^+\). Thus, if in addition \(\{i.j\} \in E_{\varvec{M}(\tau )}\), we have \(\{i,j\}\succeq \{k,l\}\) and therefore \(v(x_i-x_j) \ge v(x_k-x_l) \ge 0\). If \(a=b\), the path from i to j through \(\tau \) is contained in \(\tau _a\). Thus, if in addition \(\{i,j\} \notin E_{\varvec{M}(\tau )}\), there must be at least one edge \(\{r,s\}\) in that path such that \(\{i,j\}\prec \{r,s\}\) and therefore \(v(x_i-x_j) \le v(x_r-x_s) < 0\). This allows to bound:

$$\begin{aligned} \sum _{\{i,j\} \in E_{\varvec{M}(\tau )} \setminus E_\tau ^+}v(x_i-x_j) \ge \sum _{s=1}^k \sum _{\{i,j\} \subset { V}_{\tau _s}} v(x_i-x_j) \ge \sum _{s=1}^k -| (V_{\tau _s}|-1)\bar{B} \ge -(n-1)\bar{B} \end{aligned}$$

where to get the last inequality we have used (2.4). \(\square \)

We are now ready to conclude the proof of (2.7). By the so-called Penrose tree-graph identity [6], given a pair potential V and \((x_1,...,x_n) \in {\mathbb R}^{dn}\), one has that

$$\begin{aligned} \sum _{g\in G_{n}}~ \prod _{\{i,j\}\in E_g}\left[ e^{ -\beta V(x_i-x_j)} -1\right] = \sum _{\tau \in T_n} e^{-\beta \sum _{\{i,j\}\in E_{\varvec{M}(\tau )}\backslash E_\tau }V(x_i-x_j)} \prod _{\{i,j\}\in E_\tau }\left( e^{- \beta V(x_i-x_j)}-1\right) \end{aligned}$$
(2.21)

where \(\varvec{M}: T_n\rightarrow G_n\) is any map such that \(G_n=\biguplus _{\tau \in {T_n}}[\tau ,{\varvec{M}}(\tau )]\) with \(\biguplus \) disjoint union and \([\tau ,\varvec{M}(\tau )]=\{g\in G_n: \tau \subseteq g\subseteq \varvec{M}(\tau )\}\) (partition scheme) and \({ T}_n\) is the set of all trees with vertex set [n].

Set now \(E_\tau ^+=\{\{i,j\}\in E_\tau :V(x_i-x_j)\ge 0\}_{}\), then from (2.21) one immediately gets

$$\begin{aligned}&\left| \sum \limits _{g\in G_{n}}~ \prod \limits _{\{i,j\}\in E_g}\left[ e^{ -\beta V(x_i-x_j)} -1\right] \right| \nonumber \\&\quad ~\le ~ \sum _{\tau \in T_n} e^{-\beta \sum \limits _{\{i,j\}\in E_{\varvec{M}(\tau )}\backslash E^+_\tau }V(x_i-x_j)} \prod _{\{i,j\}\in E_\tau }\left( 1-e^{- \beta |V(x_i-x_j)|}\right) \end{aligned}$$
(2.22)

Let us now choose the partition scheme defined in Lemma 1. Then from Lemma 2, inserting (2.20) into (2.22), one obtains, for any \(n\ge 2\) and any \((x_1,\dots ,x_n)\in \mathbb {R}^{dn}\), the following inequality

$$\begin{aligned} \left| \sum _{g \in {G}_n} \prod _{\{i,j\} \in E_g} (e^{-\beta V(x_i-x_j)}-1)\right| ~\le ~ e^{\beta \bar{B} (n-1)}\sum _{\tau \in {T}_n} \prod _{\{i,j\} \in E_\tau } (1-e^{-\beta | V(x_i-x_j)|}) \end{aligned}$$
(2.23)

Now (2.7) follows easily from (2.23) recalling that \(|T_n|=n^{n-2}\) and observing that, for any \(\tau \in T_n\), it holds

$$\begin{aligned} \int _{{\Lambda } }d{x}_1\dots \int _{{\Lambda } }d{x}_{n} \prod _{\{i,j\}\in E_{\tau }} \left( 1-e^{-\beta |v(x_i-x_j)|}\right) \le |{\Lambda } |\left[ \tilde{C}(\beta )\right] ^{n-1} \end{aligned}$$
(2.24)

Appendix B: Proof of (2.5) and (2.6)

Proof of (2.5)

Let assume that V(|x|) is a stable and tempered pair potential in d dimensions with stability constant B and Basuev stability constant \(\bar{B}\). Let

$$\begin{aligned} B_n =\sup _{(x_1,\dots ,x_n)\in \mathbb {R}^{dn}}\Big \{-{1\over n}\sum _{1\le i<j\le n}V(x_i-x_j)\Big \} \end{aligned}$$

and

$$\begin{aligned} \bar{B}_n =\sup _{(x_1,\dots ,x_n)\in \mathbb {R}^{dn}}\Big \{-{1\over n- 1}\sum _{1\le i<j\le n}V(x_i-x_j)\Big \} \end{aligned}$$

so that \(B=\sup _{n\in \mathbb {N}} B_n\) and \(\bar{B}=\sup _{n\in \mathbb {N}} \bar{B}_n\).

Let us first prove that if

$$\begin{aligned} \bar{B}= \limsup _{n\rightarrow \infty } \bar{B}_n \end{aligned}$$
(2.25)

then

$$\begin{aligned} \bar{B} = B \end{aligned}$$

Suppose by contradiction that \(\bar{B}- B=\delta >0\). If this holds then necessarily there exists a finite \(m \ge 2\) such that \(B=B_m\) otherwise if \(B=\limsup _{n\rightarrow \infty } B_n\) then \(B=\bar{B}\). Now, due to the hypothesis (2.25), for all \(\varepsilon >0\) there exists \(n_0\) such that for infinitely many \(n>n_0\)

$$\begin{aligned} \bar{B}_n> \bar{B}-\varepsilon ~~~~~~~~~~\Longrightarrow ~~~~~~ B_n> {n-1\over n}(\bar{B}-\varepsilon )= {n-1\over n}(B+\delta -\varepsilon )~ \end{aligned}$$

Choose \(\varepsilon ={\delta \over 2}\), then for infinitely many n we have that

$$\begin{aligned} B_n> {n-1\over n}(B+{\delta \over 2})> B~~~~~~~~~~~~ \text{ as } \text{ soon } \text{ as } n> {2B\over \delta }+1 \end{aligned}$$

in contradiction with the assumption that \(B=B_m\). Hence we must have \(B=\limsup B_n= \limsup \bar{B}_n=\bar{B}\).

So let us suppose that the \(\sup \bar{B}_n\) is reached at some finite integer m, i.e. \(\bar{B}= \bar{B}_m\).

Since V is stable, it is bounded from below and since V is tempered \(\inf V\) cannot be positive. Let \(\inf _{x\in \mathbb {R}^d}V(x)=-C\) with \(C\ge 0\). Then for any \(\varepsilon >0\) there exists \(r_\varepsilon \) such that \(V(r_\varepsilon )< -(C -\varepsilon )\). Take the configuration \((x_1,x_2,\dots , x_{d+1})\in (\mathbb {R}^d)^{d+1}\) such that \(x_1\), \(x_2,\dots ,x_{d+1}\) are vertices of a d-dimensional hypertetrahedron with sides of length \(r_\varepsilon \). Recall that a d-dimensional hypertetrahedron has \(d+1\) vertices and \(d(d+1)/2\) sides. Then \(U(x_1,x_2,\dots , x_{d+1})< -{d(d+1)\over 2}(C-\varepsilon )\), which implies that \(B> {d\over 2}(C-\varepsilon )\) and by the arbitrariness of \(\varepsilon \) we get \(B\ge {d\over 2}C\). On the other hand we also have, for any \((x_1,\dots ,x_m)\in \mathbb {R}^{dm}\) that \(U(x_1,\dots ,x_m)\ge -{m(m-1)C/2}\) which implies that \(\bar{B}_m=\bar{B}\le {m\over 2}C\). Hence we can write \({d\over 2}C\le B\le \bar{B}=\bar{B}_m\le {m\over 2}C\) which implies \(m\ge d+1\) and so \(\bar{B}= \bar{B}_m= {m\over m-1}B_m\le {m\over m-1}B\le {d+1\over d}B\). \(\square \)

Proof of (2.6)

Let us assume that V(|x|) is a stable pair potential in d dimensions with stability constant B and Basuev stability constant \(\bar{B}\) and that V(|x|) reaches its negative minimum \(-C\) at some \(|x|=r_0\) and it is negative for all \(|x|>r_0\). We want to prove that the inequalities (2.6) hold.

First note that \(d(d-1)C\) is always a lower bound for B when \(d\ge 3\). Just consider a configuration in which n particle (with n as large as we want) are arranged in close-packed configuration at the sites of a d-dimensional face-centered cubic lattice with step \(r_0\). The energy of such configuration is (asymptotically as \(n\rightarrow \infty \)) less than or equal to \(-d(d-1)Cn\) since in a d-dimensional face-centered cubic lattice each site has \(2d(d-1)\) neighbors (see e.g. [9]) and so there are (asymptotically) \(d(d-1)n\) pairs of neighbors in the configuration. On the other hand, for any n-particle configuration \((x_1,\dots ,x_n)\in \mathbb {R}^{dn}\), it holds that \(U(x_1,\dots ,x_n)\ge -n(n-1)C/2\), i.e. \(\bar{B}_n\le nC/2\). Now, if \(\bar{B}= \sup _n\bar{B}_n\) is attained at \(n\rightarrow \infty \) then, as previously seen, we have \(\bar{B}=B\). So let us suppose that \(\bar{B}= \sup _n\bar{B}_n\) is attained at a certain finite m. Then we must have that \(d(d-1)C\le B\le \bar{B}\le mC/2\) and so \(m \ge 2d(d-1)\). Now if \(m=2d(d-1)\) then \(\bar{B}=B\). So when \(\bar{B}>B\) then \(m > 2d(d-1)\) and so we have \(\bar{B}=\frac{m}{m-1}B_m \le {2d(d-1)+1\over 2d(d-1)} B\). The case \(d=1\) and \(d=2\) are treated analogously by just observing that the close-packed arrangement in \(d=1\) is simply the cubic lattice with 2 neighbors for each site while for \(d=2\) is the triangular lattice with 6 neighbors for each site. So \(d(d-1)\) must be replaced by 1 for \(d=1\) and by 3 for \(d=2\) yielding \(m>2\) and \(m>6\) for \(d=1\) and \(d=2\) respectively. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Procacci, A. A Correction to a Remark in a Paper by Procacci and Yuhjtman: New Lower Bounds for the Convergence Radius of the Virial Series. J Stat Phys 168, 1353–1362 (2017). https://doi.org/10.1007/s10955-017-1853-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1853-4

Keywords

Mathematics Subject Classification

Navigation