Skip to main content

From Spin Glasses to Branching Brownian Motion—and Back?

  • Chapter

Part of the book series: Lecture Notes in Mathematics ((LNM,volume 2144))

Abstract

I review recent work on the construction of the extremal process of branching Brownian motion. I place this in the context of spin glass theory and in particular the Generalised Random Energy models of Derrida and Gardner. The main emphasis is on a review of the construction of the extremal process of branching Brownian motion, done in collaboration with Louis-Pierre Arguin and Nicola Kistler (Commun Pure Appl Math 64(12):1647–1676, 2011; Ann Appl Probab 22(4):1693–1711, 2012; Probab Theory Relat Fields 157:535–574, 2013). I also present some more recent results on the variable speed Brownian motion, obtained with Lisa Hartung (Electron J Probab 19(18):1–28, 2014; Lat Am J Probab Math Stat 12:261–291, 2015).

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

Buying options

Chapter
USD   29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD   39.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD   54.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Learn about institutional subscriptions

Notes

  1. 1.

    Thanks to Marek Biskup for voicing some doubt about this claim.

  2. 2.

    One should appreciate the beauty of Bramson’s construction: start with a probabilistic model (BBM), derive a pde whose solutions represent quantities of interest, and then use a different probabilistic representation of the solution (in terms of Brownian motion) to analyse these solutions…

References

  1. E. Aïdékon, Convergence in law of the minimum of a branching random walk. Ann. Probab. 41, 1362–1426 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  2. E. Aïdékon, J. Berestycki, E. Brunet, Z. Shi, Branching Brownian motion seen from its tip. Probab. Theory Relat. Fields 157, 405–451 (2013)

    Article  MATH  Google Scholar 

  3. L.-P. Arguin, A. Bovier, N. Kistler, Genealogy of extremal particles of branching Brownian motion. Commun. Pure Appl. Math. 64(12), 1647–1676 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  4. L.-P. Arguin, A. Bovier, N. Kistler, Poissonian statistics in the extremal process of branching Brownian motion. Ann. Appl. Probab. 22(4), 1693–1711 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  5. L.-P. Arguin, A. Bovier, N. Kistler, An ergodic theorem for the frontier of branching Brownian motion. Electron. J. Probab. 18(53), 25 (2013)

    Google Scholar 

  6. L.-P. Arguin, A. Bovier, N. Kistler, The extremal process of branching Brownian motion. Probab. Theory Relat. Fields 157, 535–574 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  7. L.-P. Arguin, A. Bovier, N. Kistler, An ergodic theorem for the extremal process of branching Brownian motion. Ann. Inst. Henri Poincaré Probab. Stat. 51, 557–569 (2015)

    Article  MathSciNet  Google Scholar 

  8. K.B. Athreya, P.E. Ney, Branching Processes. Die Grundlehren der mathematischen Wissenschaften, Band 196 (Springer, New York, 1972)

    Google Scholar 

  9. G. Ben Arous, A. Kuptsov, REM universality for random Hamiltonians, in Spin Glasses: Statics and Dynamics. Progress in Probability, vol. 62 (Birkhäuser, Basel, 2009), pp. 45–84

    Google Scholar 

  10. G. Ben Arous, V. Gayrard, A. Kuptsov, A new REM conjecture, in In and Out of Equilibrium. 2. Progress in Probability, vol. 60 (Birkhäuser, Basel, 2008), pp. 59–96

    Google Scholar 

  11. M. Biskup, O. Louidor, Extreme local extrema of two-dimensional discrete Gaussian free field. ArXiv e-prints (June 2013)

    Google Scholar 

  12. A. Bovier, Statistical Mechanics of Disordered Systems. Cambridge Series in Statistical and Probabilistic Mathematics (Cambridge University Press, Cambridge, 2006)

    Google Scholar 

  13. A. Bovier, L. Hartung, The extremal process of two-speed branching Brownian motion. Electron. J. Probab. 19(18), 1–28 (2014)

    MathSciNet  Google Scholar 

  14. A. Bovier, L. Hartung, Variable speed branching Brownian motion 1. Extremal processes in the weak correlation regime. ALEA, Lat. Am. J. Probab. Math. Stat. 12, 261–291 (2015)

    Google Scholar 

  15. A. Bovier, I. Kurkova, Derrida’s generalised random energy models. I. Models with finitely many hierarchies. Ann. Inst. Henri Poincaré Probab. Stat. 40(4), 439–480 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  16. A. Bovier, I. Kurkova, Derrida’s generalized random energy models. II. Models with continuous hierarchies. Ann. Inst. Henri Poincaré Probab. Stat. 40(4), 481–495 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  17. A. Bovier, I. Kurkova, M. Löwe, Fluctuations of the free energy in the REM and the p-spin SK models. Ann. Probab. 30(2), 605–651 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  18. M.D. Bramson, Maximal displacement of branching Brownian motion. Commun. Pure Appl. Math. 31(5), 531–581 (1978)

    Article  MathSciNet  MATH  Google Scholar 

  19. M. Bramson, Convergence of solutions of the Kolmogorov equation to travelling waves. Mem. Am. Math. Soc. 44(285), iv+190 (1983)

    Google Scholar 

  20. M. Bramson, O. Zeitouni, Tightness of the recentered maximum of the two-dimensional discrete Gaussian free field. Commun. Pure Appl. Math. 65(1), 1–20 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  21. M. Bramson, J. Ding, O. Zeitouni, Convergence in law of the maximum of the two-dimensional discrete Gaussian free field (Jan 2013). http://arXiv.org/abs/1301.6669

  22. B. Chauvin, A. Rouault, KPP equation and supercritical branching Brownian motion in the subcritical speed area. Application to spatial trees. Probab. Theory Relat. Fields 80(2), 299–314 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  23. B. Chauvin, A. Rouault, Supercritical branching Brownian motion and K-P-P equation in the critical speed-area. Math. Nachr. 149, 41–59 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  24. D.J. Daley, D. Vere-Jones, An Introduction to the Theory of Point Processes. Springer Series in Statistics (Springer, New York, 1988)

    Google Scholar 

  25. B. Derrida, H. Spohn, Polymers on disordered trees, spin glasses, and traveling waves. J. Stat. Phys. 51(5–6), 817–840 (1988). New directions in statistical mechanics (Santa Barbara, CA, 1987)

    Google Scholar 

  26. J. Ding, Exponential and double exponential tails for maximum of two-dimensional discrete gaussian free field. Probab. Theory Relat. Fields 157, 285–299 (2013)

    Article  MATH  Google Scholar 

  27. M. Fang, O. Zeitouni, Branching random walks in time inhomogeneous environments. Electron. J. Probab. 17(67), 18 (2012)

    Google Scholar 

  28. M. Fang, O. Zeitouni, Slowdown for time inhomogeneous branching Brownian motion. J. Stat. Phys. 149(1), 1–9 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  29. R. Fisher, The wave of advance of advantageous genes. Ann. Eugen. 7, 355–369 (1937)

    Article  Google Scholar 

  30. E. Gardner, B. Derrida, Magnetic properties and function q(x) of the generalised random energy model. J. Phys. C 19, 5783–5798 (1986)

    Article  Google Scholar 

  31. E. Gardner, B. Derrida, Solution of the generalised random energy model. J. Phys. C 19, 2253–2274 (1986)

    Article  Google Scholar 

  32. E. Gardner, B. Derrida, The probability distribution of the partition function of the random energy model. J. Phys. A 22(12), 1975–1981 (1989)

    Article  MathSciNet  Google Scholar 

  33. J.-B. Gouéré, Branching Brownian motion seen from its left-most particle. ArXiv e-prints (May 2013)

    Google Scholar 

  34. N. Ikeda, M. Nagasawa, S. Watanabe, Markov branching processes I. J. Math. Kyoto Univ. 8, 233–278 (1968)

    MathSciNet  MATH  Google Scholar 

  35. N. Ikeda, M. Nagasawa, S. Watanabe, Markov branching processes II. J. Math. Kyoto Univ. 8, 365–410 (1968)

    MathSciNet  MATH  Google Scholar 

  36. N. Ikeda, M. Nagasawa, S. Watanabe, Markov branching processes III. J. Math. Kyoto Univ. 9, 95–160 (1969)

    MathSciNet  MATH  Google Scholar 

  37. O. Kallenberg, Random Measures (Akademie, Berlin, 1983)

    MATH  Google Scholar 

  38. N. Kistler, Derrida’s random energy models from spin glasses to the extremes of correlated random fields, in Correlated Random Systems: Five Different Methods, V. Gayrard, N. Kistler, editors. Lecture Notes in Mathematics, vol. 2143 (Springer, Berlin, 2015), pp. 71–120

    Google Scholar 

  39. A. Kolmogorov, I. Petrovsky, N. Piscounov, Etude de l’ équation de la diffusion avec croissance de la quantité de matière et son application à un problème biologique. Moscou Universitet Bull. Math. 1, 1–25 (1937)

    Google Scholar 

  40. S.P. Lalley, T. Sellke, A conditional limit theorem for the frontier of a branching Brownian motion. Ann. Probab. 15(3), 1052–1061 (1987)

    Article  MathSciNet  MATH  Google Scholar 

  41. M. Leadbetter, G. Lindgren, H. Rootzén, Extremes and Related Properties of Random Sequences and Processes. Springer Series in Statistics (Springer, New York, 1983)

    Google Scholar 

  42. P. Maillard, O. Zeitouni, Slowdown in branching Brownian motion with inhomogeneous variance. ArXiv e-prints (July 2013)

    Google Scholar 

  43. B. Mallein, Maximal displacement of a branching random walk in time-inhomogeneous environment. ArXiv e-prints (July 2013)

    Google Scholar 

  44. H.P. McKean, Application of Brownian motion to the equation of Kolmogorov-Petrovskii-Piskunov. Commun. Pure Appl. Math. 28(3), 323–331 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  45. C. Newman, D. Stein, Spin Glasses and Complexity (Princeton University Press, Princeton, 2013)

    MATH  Google Scholar 

  46. J. Nolen, J.-M. Roquejoffre, L. Ryzhik, Power-like delay in time inhomogeneous fisher-kpp equations. Commun. Partial Differ. Equ. 40, 475–505 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  47. D. Panchenko, The Sherrington-Kirkpatrick Model (Springer, New York, 2013)

    Book  MATH  Google Scholar 

  48. S. Resnick, Extreme Values, Regular Variation, and Point Processes. Applied Probability. A Series of the Applied Probability Trust, vol. 4 (Springer, New York, 1987)

    Google Scholar 

  49. M.I. Roberts, A simple path to asymptotics for the frontier of a branching Brownian motion. Ann. Probab. 41(5), 3518–3541 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  50. M. Talagrand, Spin Glasses: A Challenge for Mathematicians. Ergebnisse der Mathematik und ihrer Grenzgebiete. (3) [Results in Mathematics and Related Areas (3)], vol. 46 (Springer, Berlin, 2003)

    Google Scholar 

  51. O. Zeitouni, Branching random walks and Gaussian free fields. Lecture notes (2013)

    Google Scholar 

Download references

Acknowledgements

I am deeply grateful to my collaborators on the matters of these lectures, Louis-Pierre Arguin, Nicola Kistler, Irina Kurkova, and, most recently, Lisa Hartung. They did the bulk of the work, and without them none this would have been achieved. I also thank Jiří Černý, Lisa Hartung, and an anonymous referee for pointing out various mistakes and suggesting improvement. I thank Marek Biskup, Jiří Černý, and Roman Kotecký for organising the excellent school in Prague and for the opportunity to present these lectures there.

This work is partially supported through the German Research Foundation in the Collaborative Research Center 1060 “The Mathematics of Emergent Effects”, the Priority Programme 1590 “Probabilistic Structures in Evolution”, the Hausdorff Center for Mathematics (HCM), and the Cluster of Excellence “ImmunoSensation” at Bonn University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Anton Bovier .

Editor information

Editors and Affiliations

Appendix: Point Processes

Appendix: Point Processes

Here we provide some basic background on point processes and in particular Poisson point processes. For more details, see [24, 37, 48].

1.1 Definition and Basic Properties

Point processes are designed to describe the probabilistic structure of point sets in some metric space. The key idea is to associate to a collection of points a point measure. Let us first consider a single point x. We consider the usual Borel-sigma algebra, \(\mathfrak{B} \equiv \mathfrak{B}(\mathbb{R}^{d})\), of \(\mathbb{R}^{d}\), that is generated by the open sets in the Euclidean topology of \(\mathbb{R}^{d}\). Given \(x \in \mathbb{R}^{d}\), we define the Dirac measure, δ x , such that, for any Borel set \(A \in \mathfrak{B}\),

$$\displaystyle{ \delta _{x}(A) = \left \{\begin{array}{@{}l@{\quad }l@{}} 1,\quad &\,\,\text{if}\,x \in A\\ 0,\quad &\,\,\text{if} \,x\not\in A. \end{array} \right. }$$
(260)

A point measure is now a measure, μ, on \(\mathbb{R}^{d}\), such that there exists a countable collection of points, \(\{x_{i} \in \mathbb{R}^{d},i \in \mathbb{N}\}\), such that

$$\displaystyle{ \mu =\sum _{ i=1}^{\infty }\delta _{ x_{i}} }$$
(261)

and, if K is compact, then \(\mu (K) < \infty \).

Note that the points x i need not be all distinct. The set \(S_{\mu } \equiv \{ x \in \mathbb{R}^{d}:\mu (x)\neq 0\}\) is called the support of μ. A point measure such that for all \(x \in \mathbb{R}^{d}\), μ(x) ≤ 1 is called simple.

We denote by \(M_{p}(\mathbb{R}^{d})\) the set of all point measures on \(\mathbb{R}^{d}\). We equip this set with the sigma-algebra \(\mathcal{M}_{p}(\mathbb{R}^{d})\), the smallest sigma algebra that contains all subsets of \(M_{p}(\mathbb{R}^{d})\) of the form \(\{\mu \in M_{m}(\mathbb{R}^{d}):\mu (F) \in B\}\), where \(F \in \mathfrak{B}(\mathbb{R}^{d})\) and \(B \in \mathfrak{B}([0,\infty ))\). \(\mathcal{M}_{p}(\mathbb{R}^{d})\) is also characterised by saying that it is the smallest sigma-algebra that makes the evaluation maps, \(\mu \rightarrow \mu (F)\), measurable for all Borel sets \(F \in \mathfrak{B}(\mathbb{R}^{d})\).

A point process, N, is a random variable taking values in \(M_{p}(\mathbb{R}^{d})\), i.e. a measurable map, \(N: (\varOmega,\mathfrak{F}; \mathbb{P}) \rightarrow M_{p}(\mathbb{R}^{d})\), from a probability space to the space of point measures.

This looks very fancy, but in reality things are quite down-to-earth:

Proposition 6.1

N is a point process, if and only if the map \(N(\cdot,F):\omega \rightarrow N(\omega,F)\) , is measurable from \((\varOmega,\mathfrak{F}) \rightarrow \mathfrak{B}([0,\infty ))\) , for any Borel set F, i.e. if N(F) is a real random variable.

Proof

Let us first prove necessity, which should be obvious. In fact, since \(\omega \rightarrow N(\omega,\cdot )\) is measurable into \((M_{p}(\mathbb{R}^{d}),\mathcal{M}_{p}(\mathbb{R}^{p}))\), and \(\mu \rightarrow \mu (F)\) is measurable from this space into \((\mathbb{R}_{+},\mathfrak{B}(\mathbb{R}_{+}))\), the composition of these maps is also measurable.

Next we prove sufficiency. Define the set

$$\displaystyle{ \mathfrak{G} \equiv \{ A \in \mathcal{M}_{p}(\mathbb{R}^{d}): N^{-1}A \in \mathfrak{F}\} }$$
(262)

This set is a sigma-algebra and N is measurable from \((\varOmega,\mathfrak{F}) \rightarrow (M_{p}(\mathbb{R}^{d}),\mathfrak{G})\) by definition. But \(\mathfrak{G}\) contains all sets of the form \(\{\mu \in M_{p}(\mathbb{R}^{d}):\mu (F) \in B\}\), since

$$\displaystyle{ N^{-1}\{\mu \in M_{ p}(\mathbb{R}^{d}):\mu (F) \in B\} =\{\omega \in \varOmega: N(\omega,F) \in B\} \in \mathfrak{F}, }$$
(263)

since N(⋅ , F) is measurable. Thus \(\mathfrak{G} \supset \mathcal{M}_{p}(\mathbb{R}^{d})\), and N is measurable a fortiori as a map from the smaller sigma-algebra. □ 

We will have need to find criteria for convergence of point processes. For this we recall some notions of measure theory. If \(\mathfrak{B}\) is a Borel-sigma algebra, of a metric space E, then \(\mathcal{T} \subset \mathfrak{B}\) is called a Π-system, if \(\mathcal{T}\) is closed under finite intersections; \(\mathfrak{G} \subset \mathfrak{B}\) is called a \(\lambda\)-system, or a sigma-additive class, if

  1. (i)

    \(E \in \mathfrak{G}\),

  2. (ii)

    If \(A,B \in \mathfrak{G}\), and \(A \supset B\), then \(A\setminus B \in \mathfrak{G}\),

  3. (iii)

    If \(A_{n} \in \mathfrak{G}\) and \(A_{n} \subset A_{n+1}\), then \(\lim _{n\uparrow \infty }A_{n} \in \mathfrak{G}\).

The following useful observation is called Dynkin’s theorem.

Theorem 6.2

If \(\mathcal{T}\) is a Π-system and \(\mathfrak{G}\) is a \(\lambda\) -system, then \(\mathfrak{G} \supset \mathcal{T}\) implies that \(\mathfrak{G}\) contains the smallest sigma-algebra containing \(\mathcal{T}\) .

The most useful application of Dynkin’s theorem is the observation that, if two probability measures are equal on a Π-system that generates the sigma-algebra, then they are equal on the sigma-algebra (since the set on which the two measures coincide forms a \(\lambda\)-system containing \(\mathcal{T}\)).

As a consequence we can further restrict the criteria to be verified for N to be a Point process. In particular, we can restrict the class of F’s for which N(⋅ , F) need to be measurable to bounded rectangles.

Proposition 6.3

Suppose that \(\mathcal{T}\) are relatively compact sets in \(\mathfrak{B}\) satisfying

  1. (i)

    \(\mathcal{T}\) is a Π-system,

  2. (ii)

    The smallest sigma-algebra containing \(\mathcal{T}\) is \(\mathfrak{B}\) ,

  3. (iii)

    Either, there exists \(E_{n} \in \mathcal{T}\) , such that E n ↑ E, or there exists a partition, {E n }, of E with \(\cup _{n}E_{n} = E\) , with \(E_{n} \subset \mathcal{T}\) .

Then N is a point process on \((\varOmega,\mathfrak{F})\) in \((E,\mathfrak{B})\) , if and only if the map \(N(\cdot,I):\omega \rightarrow N(\omega,I)\) is measurable for any \(I \in \mathcal{T}\) .

Corollary 6.4

Let \(\mathcal{T}\) satisfy the hypothesis of Proposition 6.3 and set

$$\displaystyle{ \mathfrak{G} \equiv \left \{\{\mu:\mu (I_{j}) = n_{j},1 \leq j \leq k\},k \in \mathbb{N},I_{j} \in \mathcal{T},n_{j} \geq 0\right \}. }$$
(264)

Then the smallest sigma-algebra containing \(\mathfrak{G}\) is \(\mathcal{M}_{p}(\mathbb{R}^{d})\) and \(\mathfrak{G}\) is a Π-system.

Next we show that the law, P N , of a point process is determined by the law of the collections of random variables N(F n ), \(F_{n} \in \mathfrak{B}(\mathbb{R}^{d})\).

Proposition 6.5

Let N be a point process in \((\mathbb{R}^{d},\mathfrak{B}(\mathbb{R}^{d})\) and suppose that \(\mathcal{T}\) is as in Proposition 6.3 . Define the mass functions

$$\displaystyle{ P_{I_{1},\ldots,I_{k}}(n_{1},\ldots,n_{k}) \equiv \mathbb{P}\left [N(I_{j}) = n_{j},\forall 1 \leq j \leq k\right ] }$$
(265)

for \(I_{j} \in \mathcal{T}\) , n j ≥ 0. Then P N is uniquely determined by the collection

$$\displaystyle{ \{P_{I_{1},\ldots,I_{k}},k \in \mathbb{N},I_{j} \in \mathcal{T}\} }$$
(266)

We need some further notions. First, if N 1, N 2 are point processes, we say that they are independent, if and only if, for any collection \(F_{j} \in \mathfrak{B}\), \(G_{j} \in \mathfrak{B}\), the vectors

$$\displaystyle{ (N_{1}(F_{j}),1 \leq j \leq k)\quad \text{and}\quad (N_{2}(G_{j}),1 \leq j \leq \ell) }$$
(267)

are independent random vectors.

The intensity measure, \(\lambda\), of a point process N is defined as

$$\displaystyle{ \lambda (F) \equiv \mathbb{E}N(F) =\int _{M_{p}(\mathbb{R}^{d})}\mu (F)P_{N}(\mathrm{d}\mu ) }$$
(268)

for \(F \in \mathfrak{B}\).

For measurable functions \(f: \mathbb{R}^{d} \rightarrow \mathbb{R}_{+}\), we define

$$\displaystyle{ N(\omega,f) \equiv \int _{\mathbb{R}^{d}}f(x)N(\omega,\mathrm{d}x) }$$
(269)

Then N(⋅ , f) is a random variable. We have that

$$\displaystyle{ \mathbb{E}N(f) =\lambda (f) =\int _{\mathbb{R}^{d}}f(x)\lambda (\mathrm{d}x). }$$
(270)

1.2 Laplace Functionals

If Q is a probability measure on \((M_{p},\mathcal{M}_{p})\), the Laplace transform of Q is a map, ψ from non-negative Borel functions on \(\mathbb{R}^{d}\) to \(\mathbb{R}_{+}\), defined as

$$\displaystyle{ \psi (f) \equiv \int _{M_{p}}\exp \left (-\int _{R^{d}}f(x)\mu (\mathrm{d}x)\right )Q(\mathrm{d}\mu ). }$$
(271)

If N is a point process, the Laplace functional of N is

$$\displaystyle\begin{array}{rcl} \psi _{N}(f)& \equiv & \mathbb{E}\mathrm{e}^{-N(f)} =\int \mathrm{ e}^{-N(\omega,f)}\mathbb{P}(\mathrm{d}\omega ) \\ & =& \int _{M_{p}}\exp \left (-\int _{R^{d}}f(x)\mu (\mathrm{d}x)\right )P_{N}(\mathrm{d}\mu ){}\end{array}$$
(272)

Proposition 6.6

The Laplace functional, ψ N , of a point process, N, determines N uniquely.

Proof

For k ≥ 1, and \(F_{1},\ldots,F_{k} \in \mathfrak{B}\), \(c_{1},\ldots,c_{k} \geq 0\), let \(f =\sum _{ i=1}^{k}c_{i}\mathbb{1}_{F_{i}}(x)\). Then

$$\displaystyle{ N(\omega,f) =\sum _{ i=1}^{k}c_{ i}N(\omega,F_{i}) }$$
(273)

and

$$\displaystyle{ \psi _{N}(f) = \mathbb{E}\exp \left (-\sum _{i=1}^{k}c_{ i}N(F_{i})\right ). }$$
(274)

This is the Laplace transform of the vector (N(F i ), 1 ≤ i ≤ k), that determines uniquely its law. Hence the proposition follows from Proposition 6.5 □ 

1.3 Poisson Point Processes

The most important class of point processes for our purposes will be Poisson point processes.

Definition 6.7

Let \(\lambda\) be a \(\sigma\)-finite, positive measure on \(\mathbb{R}^{d}\). Then a point process, N, is called a Poisson point process with intensity measure \(\lambda\) (\(PPP(\lambda )\)), if

  1. (i)

    For any \(F \in \mathfrak{B}(\mathbb{R}^{d})\), and \(k \in \mathbb{N}\),

    $$\displaystyle{ \mathbb{P}\left [N(F) = k\right ] = \left \{\begin{array}{@{}l@{\quad }l@{}} \mathrm{e}^{-\lambda (F)}\frac{(\lambda (F))^{k}} {k!},\quad &\quad \text{if}\,\,\lambda (F) < \infty \\ 0, \quad &\quad \text{if}\,\,\lambda (F) = \infty,\end{array} \right. }$$
    (275)
  2. (ii)

    If \(F,G \in \mathfrak{B}\) are disjoint sets, then N(F) and N(G) are independent random variables.

In the next theorem we will assert the existence of a Poisson point process with any desired intensity measure. In the proof we will give an explicit construction of such a process.

Proposition 6.8

  1. (i)

    \(PPP(\lambda )\) exists and its law is uniquely determined by the requirements of the definition.

  2. (ii)

    The Laplace functional of \(PPP(\lambda )\) is given, for f ≥ 0, by

    $$\displaystyle{ \varPsi _{N}(f) =\exp \left (\int _{\mathbb{R}^{d}}(\mathrm{e}^{-f(x)} - 1)\lambda (\mathrm{d}x)\right ). }$$
    (276)

Proof

Since we know that the Laplace functional determines a point process, in order to prove that the conditions of the definition uniquely determine the \(PPP(\lambda )\), we show that they determine the form (276) of the Laplace functional. Thus suppose that N is a \(PPP(\lambda )\). Let \(f = c\mathbb{1}_{F}\). Then

$$\displaystyle\begin{array}{rcl} \varPsi _{N}(f)& =& \mathbb{E}\exp \left (-N(f)\right ) = \mathbb{E}\exp \left (-cN(F)\right ) \\ & =& \sum _{k=0}^{\infty }\mathrm{e}^{-ck}\mathrm{e}^{-\lambda (F)}\frac{(\lambda (F))^{k}} {k!} =\mathrm{ e}^{(\mathrm{e}^{-c}-1)\lambda (F) } \\ & =& \exp \left (\int (\mathrm{e}^{-f(x)}-)\lambda (\mathrm{d}x)\right ), {}\end{array}$$
(277)

which is the desired form. Next, if F i are disjoint, and \(f =\sum _{ i=1}^{k}c_{i}\mathbb{1}_{F-i}\), it is straightforward to see that

$$\displaystyle{ \varPsi _{N}(f) = \mathbb{E}\exp \left (-\sum _{i=1}^{k}c_{ i}N(F_{i})\right ) =\prod _{ i=^{1}}^{k}\mathbb{E}\exp \left (-c_{ i}N(F_{i})\right ) }$$
(278)

due to the independence assumption (ii); a simple calculations shows that this yields again the desired form. Finally, for general f, we can choose a sequence, f n , of the form considered, such that f n ↑ f. By monotone convergence then N(f n ) ↑ N(f). On the other hand, since eN(g) ≤ 1, we get from dominated convergence that

$$\displaystyle{ \varPsi _{N}(f_{n}) = \mathbb{E}\mathrm{e}^{-N(f_{n})} \rightarrow \mathbb{E}\mathrm{e}^{-N(f)} =\varPsi _{ N}(f). }$$
(279)

But, since \(1 -\mathrm{ e}^{-f_{n}(x)} \uparrow 1 -\mathrm{ e}^{-f(x)}\), and monotone convergence gives once more

$$\displaystyle{ \varPsi _{N}(f_{n}) =\exp \left (\int (1 -\mathrm{ e}^{-f_{n}(x)})\lambda (\mathrm{d}x)\right ) \uparrow \exp \left (\int (1 -\mathrm{ e}^{-f(x)})\lambda (\mathrm{d}x)\right ) }$$
(280)

On the other hand, given the form of the Laplace functional, it is trivial to verify that the conditions of the definition hold, by choosing suitable functions f.

Finally we turn to the construction of \(PPP(\lambda )\). Let us first consider the case \(\lambda (\mathbb{R}^{d}) < \infty \). Then construct

  1. (i)

    A Poisson random variable, τ, of parameter \(\lambda (\mathbb{R}^{d})\).

  2. (ii)

    A family, X i , \(i \in \mathbb{N}\), of independent, \(\mathbb{R}^{d}\)- valued random variables with common distribution \(\lambda\). This family is independent of τ.

Then set

$$\displaystyle{ N^{{\ast}}\equiv \sum _{ i=1}^{\tau }\delta _{ X_{i}} }$$
(281)

It is not very hard to verify that N is a \(PPP(\lambda )\).

To deal with the case when \(\lambda (\mathbb{R}^{d})\) is infinite, decompose \(\lambda\) into a countable sum of finite measures, \(\lambda _{k}\), that are just the restriction of \(\lambda\) to a finite set F k , where the F k form a partition of \(\mathbb{R}^{d}\). Then N is just the sum of independent \(PPP(\lambda _{k})\ N_{k}^{{\ast}}\). □ 

1.4 Convergence of Point Processes

Before we turn to applications to extremal processes, we still have to discuss the notion of convergence of point processes. As point processes are probability distributions on the space of point measures, we will naturally think about weak convergence. This means that we will say that a sequence of point processes, N n , converges weakly to a point process, N, if for all continuous functions, f, on the space of point measures,

$$\displaystyle{ \mathbb{E}f(N_{n}) \rightarrow \mathbb{E}f(N). }$$
(282)

However, to understand what this means, we must discuss what continuous functions on the space of point measures are, i.e. we must introduce a topology on the set of point measures. The appropriate topology for our purposes will be that of vague convergence.

1.4.1 Vague Convergence

We consider the space \(\mathbb{R}^{d}\) equipped with its natural Euclidean metric. Clearly \(\mathbb{R}^{d}\) is a complete, separable metric space. We will denote by \(C_{0}(\mathbb{R}^{d})\) the set of continuous real-valued functions on \(\mathbb{R}^{d}\) that have compact support; \(C_{0}^{+}(\mathbb{R}^{n})\) denotes the subset of non-negative functions. We consider \(\mathcal{M}_{+}(\mathbb{R}^{d})\) the set of all \(\sigma\)-finite, positive measures on \((\mathbb{R}^{d},\mathfrak{B}(\mathbb{R}^{d}))\). We denote by \(\mathcal{M}_{+}(\mathbb{R}^{d})\) the smallest sigma-algebra of subsets of \(M_{+}(\mathbb{R}^{d})\) that makes the maps \(m \rightarrow m(f)\) measurable for all \(f \in C_{0}^{+}(\mathbb{R}^{d})\).

We will say that a sequence of measures, \(\mu _{n} \in M_{+}(\mathbb{R}^{d})\) converges vaguely to a measure \(\mu \in M_{+}(\mathbb{R}^{d})\), if, for all \(f \in C_{0}^{+}(\mathbb{R}^{d})\),

$$\displaystyle{ \mu _{n}(f) \rightarrow \mu (f) }$$
(283)

Note that for this topology, typical open neighbourhoods are of the form

$$\displaystyle{ B_{f_{1},\ldots,f_{k},\epsilon }(\mu ) \equiv \{\nu \in M_{+}(\mathbb{R}^{d}): \forall _{ i=1}^{k}\,\vert \nu (f_{ i}) -\mu (f_{i})\vert <\epsilon \}, }$$
(284)

i.e. to test the closeness of two measures, we test it on their expectations on finite collections of continuous, positive functions with compact support. Given this topology, on can of course define the corresponding Borel sigma algebra, \(\mathfrak{B}(M_{+}(\mathbb{R}^{d}))\), which (fortunately) turns out to coincide with the sigma algebra \(\mathcal{M}_{+}(\mathbb{R}^{d})\) introduced before.

The following properties of vague convergence are useful.

Proposition 6.9

Let μ n , \(n \in \mathbb{N}\) be in \(M_{+}(\mathbb{R}^{d})\) . Then the following statements are equivalent:

  1. (i)

    μ n converges vaguely to μ, \(\mu _{n}\buildrel v \over \rightarrow \mu \) .

  2. (ii)

    \(\mu _{n}(B) \rightarrow \mu (B)\) for all relatively compact sets, B, such that μ(∂B) = 0.

  3. (iii)

    \(\limsup _{n\uparrow \infty }\mu _{n}(K) \leq \mu (K)\) and \(\limsup _{n\uparrow \infty }\mu _{n}(G) \geq \mu (G)\) , for all compact K, and all open, relatively compact G.

1.4.2 Weak Convergence

Having established the space of \(\sigma\)-finite measures as a complete, separable metric space, we can think of weak convergence of probability measures on this space just as if we were working on an Euclidean space.

Definition 6.10

A sequence of point processes N n converges weakly with respect to the vague topology to a point process, N, iff for all functions \(F: \mathcal{M}_{+}(\mathbb{R}^{d}) \rightarrow \mathbb{R}\) that are continuous with respect to the vague topology,

$$\displaystyle{ \lim _{n\uparrow \infty }\mathbb{E}F(N_{n}) = \mathbb{E}F(N). }$$
(285)

An important convergence criterion is given by the following theorem (see [24, Chap. 11]).

Theorem 6.11

A sequence of point processes N n converges weakly to a point process N, if for all positive continuous functions, ϕ, with compact support,

$$\displaystyle{ \lim _{n\uparrow \infty }\varPsi _{N_{n}}(\phi ) =\varPsi _{N}(\phi ). }$$
(286)

Rights and permissions

Reprints and permissions

Copyright information

© 2015 Springer International Publishing Switzerland

About this chapter

Cite this chapter

Bovier, A. (2015). From Spin Glasses to Branching Brownian Motion—and Back?. In: Biskup, M., Černý, J., Kotecký, R. (eds) Random Walks, Random Fields, and Disordered Systems. Lecture Notes in Mathematics, vol 2144. Springer, Cham. https://doi.org/10.1007/978-3-319-19339-7_1

Download citation

Publish with us

Policies and ethics