Skip to main content
Log in

Computing Freidlin’s Cycles for the Overdamped Langevin Dynamics. Application to the Lennard-Jones-38 Cluster

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

Abstract

The large time behavior of a stochastic system with infinitesimally small noise can be described in terms of Freidlin’s cycles. We show that if the system is gradient and the potential satisfies certain non-restrictive conditions, the hierarchy of cycles has a structure of a full binary tree, and each cycle is exited via the lowest saddle adjacent to it. Exploiting this property, we propose an algorithm for computing the asymptotic zero-temperature path and building a hierarchy of Freidlin’s cycles associated with the transition process between two given local equilibria. This algorithm is suitable for systems with a complex potential energy landscape with numerous minima. We apply it to find the asymptotic zero-temperature path and Freidlin’s cycles involved into the transition process between the two lowest minima of the Lennard-Jones cluster of 38 atoms. D. Wales’s stochastic network of minima and transition states of this cluster is used as an input.

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
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. The network available at the Wales group web site is not connected. Its maximal connected component consists of 71887 minima. This component contains FCC and ICO.

  2. There is a typo in [7] in the second paragraph on page 338: \(m_{C}(i)=V_{iJ(i)}-V_{k^{\ast}J(k^{\ast})}\) should be replaced with \(m_{C}(i)=V_{k^{\ast}J(k^{\ast})}-V_{iJ(i)}\).

References

  1. Becker, O.M., Karplus, M.: The topology of multidimensional potential energy surfaces: theory and application to peptide structure and kinetics. J. Chem. Phys. 106, 1495–1517 (1997)

    Article  ADS  Google Scholar 

  2. Cameron, M., Kohn, R.V., Vanden-Eijnden, E.: The string method as a dynamical system. J. Nonlinear Sci. 21(2), 193–230 (2011)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  3. Carr, J.M., Trigubenko, S.A., Wales, D.J.: Finding pathways between distant local minima. J. Chem. Phys. 122, 234903 (2005)

    Article  ADS  Google Scholar 

  4. Dijkstra, E.W.: A note on two problems in connexion with graphs. Numer. Math. 1, 269–271 (1959)

    Article  MathSciNet  MATH  Google Scholar 

  5. Doye, J.P.K., Miller, M.A., Wales, D.J.: The double-funnel energy landscape of the 38-atom Lennard-Jones cluster. J. Chem. Phys. 110, 6896–6906 (1999)

    Article  ADS  Google Scholar 

  6. Freidlin, M.I.: Sublimiting distributions and stabilization of solutions of parabolic equations with small parameter. Sov. Math. Dokl. 18(4), 1114–1118 (1977)

    MATH  Google Scholar 

  7. Freidlin, M.I.: Quasi-deterministic approximation, metastability and stochastic resonance. Physica D 137, 333–352 (2000)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  8. Freidlin, M.I., Wentzell, A.D.: Random Perturbations of Dynamical Systems, 2nd edn. Springer, New York (1998)

    Book  MATH  Google Scholar 

  9. Hamilton, J.C., Siegel, D.J., Uberuaga, B.P., Voter, A.F.: Isometrization rates and mechanisms for the 38-atom Lennard-Jones cluster determined using molecular dynamics and temperature accelerated molecular dynamics. Available via http://www-personal.umich.edu/~djsiege/Energy_Ctorage_Lab/Publications_files/LJ38_v14.eps

  10. Mandelshtam, V.A., Frantsuzov, P.A.: Multiple structural transformations in Lennard-Jones clusters: generic versus size-specific behavior. J. Chem. Phys. 124, 204511 (2006)

    Article  ADS  Google Scholar 

  11. Metzner, P., Schuette, Ch., Vanden-Eijnden, E.: Transition path theory for Markov jump processes. Multiscale Model. Simul. 7(3), 1192–1219 (2008)

    Article  MathSciNet  Google Scholar 

  12. Neirotti, J.P., Calvo, F., Freeman, D.L., Doll, J.D.: Phase changes in 38-atom Lennard-Jones clusters. I. A parallel tempering study in the canonical ensemble. J. Chem. Phys. 112, 10340 (2000)

    Article  ADS  Google Scholar 

  13. Norris, J.R.: Markov Chains. Cambridge University Press, Cambridge (1997)

    Book  MATH  Google Scholar 

  14. Picciani, M., Athenes, M., Kurchan, J., Taileur, J.: Simulating structural transitions by direct transition current sampling: the example of LJ38. J. Chem. Phys. 135, 034108 (2011)

    Article  ADS  Google Scholar 

  15. Vanden-Eijnden, E.: Transition Path Theory. Lecture Notes in Physics, vol. 703, pp. 439–478. Springer, Berlin/Heidelberg (2006)

    Google Scholar 

  16. Vanden-Eijnden, E., Heymann, M.: The geometric minimum action method for computing minimum energy paths. J. Chem. Phys. 128, 061103 (2008)

    Article  ADS  Google Scholar 

  17. Wales, D.J.: Discrete path sampling. Mol. Phys. 100, 3285–3306 (2002)

    Article  ADS  Google Scholar 

  18. Wales, D.J.: Some further applications of discrete path sampling to cluster isomerization. Mol. Phys. 102, 891–908 (2004)

    Article  ADS  Google Scholar 

  19. Wales, D.J.: Energy landscapes: calculating pathways and rates. Int. Rev. Phys. Chem. 25(1–2), 237–282 (2006)

    Article  Google Scholar 

  20. Wales, D.J.: Website contains the database for the Lennard-Jones-38 cluster. Available via http://www-wales.ch.cam.ac.uk/examples/PATHSAMPLE

  21. Wales, D.J., Doye, J.P.K.: Global optimization by basin-hopping and the lowest energy structures of Lennard-Jones clusters containing up to 110 atoms. J. Phys. Chem. A 101, 5111–5116 (1997)

    Article  Google Scholar 

  22. Wales, D.J., Miller, M.A., Walsch, T.R.: Archetypal energy landscapes. Nature 394, 758–760 (1998)

    Article  ADS  Google Scholar 

  23. Weinan, E., Ren, W., Vanden-Eijnden, E.: String method for study of rare events. Phys. Rev. B 66, 052301 (2002)

    ADS  Google Scholar 

  24. Weinan, E., Ren, W., Vanden-Eijnden, E.: Simplified and improved string method for computing the minimum energy paths in barrier-crossing events. J. Chem. Phys. 126, 164103 (2007)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

I thank Prof. David Wales for referring me to his LJ38 network and helping to interpret the data files. I am grateful to Prof. Eric Vanden-Eijnden for valuable discussion. I thank Prof. Mark Freidlin for valuable remarks and advice. This work is supported by the Alfred Sloan foundation, by the DARPA YFA Grant N66001-12-1-4220, and the NSF grant 1217118.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. K. Cameron.

Additional information

Alfred Sloan foundation, DARPA YFA N66001-12-1-4220, NSF grant 1217118.

Appendix

Appendix

Proof

(Lemma 3) For zero order cycles consisting of a single state, the main state is the state itself. Hence Lemma 3 holds straightforwardly. For first order cycles, the main state is defined as follows [7]. Let C be a first order cycle. Then its main state is

$$ M(C)=\arg\max_{i\in C}U_{iJ(i)},\quad \mathrm{where}\ J(i)=\arg\min_{j\neq i}U_{ij}. $$
(45)

Using Eq. (3) we obtain

$$ M(C)=\arg\max_{i\in C} (V_{iJ(i)}-V_{i}). $$
(46)

First order cycles are constructed identically in this work and in [68]. Hence Lemma 1 claiming that first order cycles consist of exactly two states holds. Let C={i,j} be a first order cycle. Suppose V j <V i . Then

$$M(C)=\arg\max\{V_{ij}-V_i,V_{ij}-V_j\}=\arg\min\{V_i,V_j\}=j. $$

Hence Lemma 3 holds for first order cycles.

In [68], the main state for higher order cycles is defined via the concept of i-graph.

Definition 6

Let S be a set of states and i be a selected state. Then an i-graph is a directed graph G(S,E) satisfying the following properties.

  1. 1.

    Every state mS∖{i} is the initial point of exactly one arrow.

  2. 2.

    For any state mS∖{i} there exists a sequence of arrows leading from m to i.

Let C be a kth order cycle where k≥2. Then the main state M(C) is given by

$$ M(C)=\arg\min_{j\in C} \biggl(\min _{g\in G_j(C)}\sum_{(m\rightarrow n)\in g} U_{mn} \biggr). $$
(47)

Here G j (C) is the collection of all j-graphs defined on the cycle C and ∑(mn)∈g is the sum over all edges of the j-graph g. Using Eq. (3) we rewrite Eq. (47) as

(48)

We have split the sum into two sums, the sum over all saddles and the sum over all minima associated with the j-graph g. The sum in Eq. (48) is minimized over all i-graphs that may be defined on the set of states C. Next we make the following the following two observations.

Observation 1

The union

$$\bigcup_{j\in C}G_j(C) $$

can be obtained as follows. Let T(C) be the collection of all trees on C. We remind that a tree is an undirected graph in which any two vertices are connected by exactly one simple path (i.e., a path with no repeated vertices). Then for every tree tT(C) and every state jC we can define a j-graph. We note that

$$ \sum_{(m\rightarrow n)\in g\in G_j(C)}2V_{mn} = \sum _{(m\rightarrow n)\in t(g)}2V_{mn}, $$
(49)

where the tree t(g) is obtained from the directed graph g by making all its edges undirected. This means that the minimum value of the sum over all saddles associated with a j-graph g is completely determined by the tree t(g) corresponding to this graph. Hence the sum over the saddles does not influence argmin jC in Eq. (48). Figure 10 illustrates this phenomenon for the second order cycle {A,B,C} in the 3-well potential example in Sect. 2.1.

Fig. 10
figure 10

Illustration for the proof of Lemma 3. The collections of i-graphs for each state of the second order cycle C={A,B,C} in the 3-well potential example in Sect. 2.1. Obviously, there exist three different trees for the set of vertices {A,B,C}. For each of the states A, B, and C, the collection of i-graphs is obtained from these trees

Observation 2

Next we observe that the second sum in Eq. (48) is completely determined by the state j, i.e., it is independent of the j-graph g. Indeed, since g is connected, the second sum in Eq. (48) contains the values of the potential at all states but one: the state j.

Taking into account Observations 1 and 2 we rewrite Eq. (48) as

(50)

Therefore, the main state of a cycle C is its deepest minimum. □

Proof

(Lemma 4) For a first order cycle C, the stationary distribution rate is defined by [7]Footnote 2

$$ m_C(i)=U_{M(C)J(M(C))}-U_{iJ(i)}, \quad \mathrm{where}\ J(i)=\arg\min_{j}U_{ij}. $$
(51)

By Lemma 1, first order cycles consist of two states: C=ij. Let V j <V i , i.e., j is the main state of the cycle C. Then Eq. (51) becomes

(52)

Hence the claim holds for the first order cycles.

For higher order cycles, the stationary distribution rate is defined as

$$ m_C(i)=\min_{g\in G_i(C)}\sum _{(m\rightarrow n)\in g} U_{mn}-\min_{g\in G_{M(C)}(C)}\sum _{(m\rightarrow n)\in g} U_{mn}. $$
(53)

Using Eq. (3) we rewrite Eq. (53) as

(54)

Taking into account Observations 1 and 2 made in the proof of Lemma 3 we get

(55)

Thus, Lemma 4 holds for cycles of order k≥2 as well. □

Proof

(Lemma 5) The exit rate [7] of a cycle C is defined by

$$ \mathcal{E}(C)=\min_{i\in C,~l\notin C} \bigl(m_C(i)+U_{il} \bigr). $$
(56)

Using Eq. (3) and Lemma 4 we rewrite Eq. (56) as

(57)

This completes the proof. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Cameron, M.K. Computing Freidlin’s Cycles for the Overdamped Langevin Dynamics. Application to the Lennard-Jones-38 Cluster. J Stat Phys 152, 493–518 (2013). https://doi.org/10.1007/s10955-013-0770-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-013-0770-4

Keywords

Navigation