To save content items to your account,
please confirm that you agree to abide by our usage policies.
If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account.
Find out more about saving content to .
To save content items to your Kindle, first ensure no-reply@cambridge.org
is added to your Approved Personal Document E-mail List under your Personal Document Settings
on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part
of your Kindle email address below.
Find out more about saving to your Kindle.
Note you can select to save to either the @free.kindle.com or @kindle.com variations.
‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi.
‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.
This volume, like its predecessor, Probability and Statistics by Example, Vol. 1, was initially conceived with the intention of giving Cambridge students an opportunity to check their level of preparation for Mathematical Tripos examinations. And, as with the first volume, in the course of the preparation, another goal became important: to give the general public a clearer picture of how probability- and statistics-related courses are taught in a place like the University of Cambridge, and what level of knowledge is achieved (or aimed for) by the end of these courses. In addition, the specific topic of this volume, Markov chains and their applications, has in recent years undergone a real surge. A number of remarkable theoretical results were obtained in this field which only twenty years or so ago was considered by many probabilists as a ‘dead’ zone. Even more surprisingly, an active part in this exciting development was played by applied research. Motivated by a dramatically increasing number of problems emerging in such diverse areas as computer science, biology and finance, applied people boldly invaded the territory traditionally reserved for the few hardened enthusiasts who until then had continued to improve old results by removing one or another condition in theorems which became increasingly difficult to read, let alone apply. We thus felt compelled to include some of these relatively recent ideas in our book, although the corresponding sections have little to do with current Cambridge courses.
The topic of Markov chains occupies a special place in teaching probability theory. It is named after the Russian mathematician who introduced and developed this elegant concept in the 1900s, 30 years before the notion of probability was shaped in the manner we use it today.
Andrei Andreevich Markov (1856–1922) was born into the family of a Russian civil servant. His father, following the family tradition, began his career by studying at a local seminary, but then moved into a forestry inspection office and later became a private solicitor. Markov's father was well known for his frankness and high principles, qualities inherited by his son, but was also inclined to gamble at card games. Once he lost all the family's possessions, but luckily his opponent was unmasked as a cheat, and the loss was declared void. His son by contrast loved chess and was considered one of the best amateur players of the time. When Mikhail Chigorin, a Russian chess master, was preparing for his 1892 match for the World Chess Championship with the Austrian Wilhelm Steinitz, the reigning World champion, he played a sparring series of four games with Markov; Markov won one and drew another. (Chigorin was later dramatically defeated in the decisive game by Steinitz, to the deep disappointment of numerous chess enthusiasts in Russia who still deplore this loss).
Markov processes specialists like to do it with chains.
(From the series ‘How they do it’.)
Definition 2.1.1 A Q-matrix on a finite or countable state space I is a real-valued matrix (qij, i, j ∈ I) with:
non-positive diagonal entries qii ≤ 0, i ∈ I,
non-negative off-diagonal entries qij ≥ 0, i ≠ j, i, j ∈ I,
the row zero-sum condition: −qii = Σj∈I:j≠iqij, i.e. Σjqij = 0 for all i ∈ I.
For i ≠ j, the value qij represents the jump, or transition rate from state i to j. The value −qii = Σj:j≠iqij is denoted by qi (we will see that it represents the total jump, or exit rate from state i). A Q-matrix will be denoted by Q (a common abuse of notation). As in Chapter 1, we will denote by I the unit matrix.
In a general theory of countable continuous-time Markov chains, the row zerosum condition Σj∈I:j≠iqij = −qii presumes that the series Σj:j≠iqij < ∞. However, a substantial part of the theory can be developed when the equality in this condition is relaxed to the upper bound Σjqij ≤ 0, i.e. qi ≥ Σj:j≠iqij for all i ∈ I. Then a Q-matrix satisfying the row zero-sum condition is called conservative; we will omit this term in the present volume, as we will not consider non-conservative Q-matrices.
The Markov property and its immediate consequences
Mathematics cannot be learned by lectures alone, anymore than piano playing can be learned by listening to a player.
C. Runge (1856–1927), German applied mathematician
Typically, the subject of Markov chains represents a logical continuation from a basic course of probability. We will study a class of random processes describing a wide variety of systems of theoretical and practical interest (and sometimes simply amusing). The fact that deep insight into the subject is possible without using sophisticated mathematical tools may also be an explanation of why Markov chains are popular in so many different disciplines which are seemingly remote from pure mathematics.
The basic model for the first half of the book will be a system which changes state in discrete time, according to some random mechanism. The collection of states is called a state space and throughout the whole book will be assumed finite or countable; we will denote it by I. Each i ∈ I is called a state; our system will always be in one of these states. Sometimes we will know what state the system occupies and sometimes only that the system is in state i with some probability.
We consider nonlinear parabolic equations of Hamilton–Jacobi– Bellman type. The Lagrangian is assumed to be convex, but with a spatial dependence which is stationary and random. Rescaling in space and time produces a similar equation with a rapidly varying spatial dependence and a small viscosity term. Motivated by corresponding results for the linear elliptic equation with small viscosity, we seek to find the limiting behavior of the solution of the Cauchy (final value) problem in terms of a homogenized problem, described by a convex function of the gradient of the solution. The main idea is to use the principle of dynamic programming to write a variational formula for the solution in terms of solutions of linear problems. We then show that asymptotically it is enough to restrict the optimization to a subclass, one for which the asymptotic behavior can be fully analyzed. The paper outlines these steps and refers to the recently published work of Kosygina, Rezakhanlou and the author for full details.
Homogenization is a theory about approximating solutions of a differential equation with rapidly varying coefficients by a solution of a constant coefficient differential equation of a similar nature. The simplest example of its kind is the solution uϵ of the equation on [0, ∞] ⨯ ℝ. The function a(·) is assumed to be uniformly positive, continuous and periodic of period 1.
The Vlasov equation for the collisionless evolution of the singleparticle probability distribution function (PDF) is a well-known example of coadjoint motion. Remarkably, the property of coadjoint motion survives the process of taking moments. That is, the evolution of the moments of the Vlasov PDF is also a form of coadjoint motion. We find that geodesic coadjoint motion of the Vlasov moments with respect to powers of the single-particle momentum admits singular (weak) solutions concentrated on embedded subspaces of physical space. The motion and interactions of these embedded subspaces are governed by canonical Hamiltonian equations for their geodesic evolution.
1. Introduction
The Vlasov equation. The evolution of N identical particles in phase space with coordinates .(qi ; pi)
i = 1, 2,....,N, may be described by an evolution equation for their joint probability distribution function. Integrating over all but one of the particle phase-space coordinates yields an evolution equation for the single-particle probability distribution function (PDF). This is the Vlasov equation.
The solutions of the Vlasov equation reflect its heritage in particle dynamics, which may be reclaimed by writing its many-particle PDF as a product of delta functions in phase space. Any number of these delta functions may be integrated out until all that remains is the dynamics of a single particle in the collective field of the others. In plasma physics, this collective field generates the total electromagnetic properties and the self-consistent equations obeyed by the single particle PDF are the Vlasov–Maxwell equations.
Integrable partial differential equations have been studied because of their remarkable mathematical structure ever since they were discovered in the 1960s. Some of these equations were originally derived to describe approximately the evolution of water waves as they propagate in shallow water. This paper examines how well these integrable models describe actual waves in shallow water.
1. Introduction Zabusky and Kruskal [1965] introduced the concept of a soliton—a spatially localized solution of a nonlinear partial differential equation with the property that this solution always regains its initial shape and velocity after interacting with another localized disturbance. They were led to the concept of a soliton by their careful computational study of solutions of the Korteweg–de Vries (or KdV) equation, (See [Zabusky 2005] for his summary of this history.) After that breakthrough, they and their colleagues found that the KdV equation has many remarkable properties, including the property discovered by Gardner, Greene, Kruskal and Miura [Gardner et al. 1967]: the KdV equation can be solved exactly, as an initial-value problem, starting with arbitrary initial data in a suitable space. This discovery was revolutionary, and it drew the interest of many people. We note especially the work of Zakharov and Faddeev [1971], who showed that the KdV equation is a nontrivial example of an infinite-dimensional Hamiltonian system that is completely integrable. This means that under a canonical change of variables, the original problem can be written in terms of action-angle variables, in which the action variables are constants of the motion, while the angle variables evolve according to nearly trivial ordinary differential equations (ODEs).
We discuss the relevance of the classical Riccati substitution to the spectral edge statistics in some fundamental models of one-dimensional random Schrödinger and random matrix theory.
1. Introduction
The Riccati map amounts to the observation that the Schr¨odinger eigenvalue problemis transformed into the first order relation
q(x) = ƛ +. p′(x)+p2>′(x)
upon setting. That this simple fact has deep consequences for the problem of characterizing the spectrum of Q with a random potential q has been known for some time. It also turns out to be important for related efforts in random matrix theory (RMT). We will describe some of the recent progress on both fronts.
Random operators of type Q arise in the description of disordered systems. Their use goes back to Schmidt [1957], Lax and Phillips [1958], and Frisch and Lloyd [1960] in connection with disordered crystals, represented by potentials in the form of trains of signed random masses, randomly placed on the line. Consider instead the case of white noise potential, q(x) =b′(x) with a standard brownian motion x ⟼ b(x), which may be viewed as a simplifying caricature of the above.
We discuss recent advances in the theory of turbulent solutions of the Navier–Stokes equations and the existence of their associated invariant measures. The statistical theory given by the invariant measures is described and associated with historically-known scaling laws. These are Hack's law in one dimension, the Batchelor–Kraichnan law in two dimensions and the Kolmogorov's scaling law in three dimensions. Applications to problems in turbulence are discussed and applications to Reynolds Averaged Navier Stokes (RANS) and Large Eddy Simulation (LES) models in computational turbu-lence.
1. Introduction
Everyone is familiar with turbulence in one form or another. Airplane passengers encounter it in wintertime as the plane begins to shake and is jerked in various directions. Thermal currents and gravity waves in the atmosphere create turbulence encountered by low-flying aircraft. Turbulent drag also prevents the design of more fuel-efficient cars and aircrafts. Turbulence plays a role in the heat transfer in nuclear reactors, causes drag in oil pipelines and influence the circulation in the oceans as well as the weather.
In daily life we encounter countless other examples of turbulence. Surfers use it to propel them and their boards to greater velocities as the wave breaks and becomes turbulent behind them and they glide at great speeds down the unbroken face of the wave. This same wave turbulence shapes our beaches and carries enormous amount of sand from the beach in a single storm, sometime to dump it all into the nearest harbor. Turbulence is harnessed in combustion engines in cars and jet engines for effective combustion and reduced emission of pollutants.
We present the integration of the “pair” flows associated to the Camassa–Holm (CH) hierarchy i.e., an explicit exact formula for the update of the initial velocity profile in terms of initial data when run by the flow associated to a Hamiltonian which (up to a constant factor) is given by the sum of the reciprocals of the squares of any two eigenvalues of the underlying spectral problem. The method stems from the integration of “individual” flows of the CH hierarchy described in [Loubet 2006; McKean 2003], and is seen to be more general in scope in that it may be applied when considering more complex flows (e.g., when the Hamiltonian involves an arbitrary number of eigenvalues of the associated spectral problem) up to when envisaging the full CH flow itself which is nothing but a superposition of commuting individual actions. Indeed, by incorporating piece by piece into the Hamiltonian the distinct eigenvalues describing the spectrum associated to the initial profile, we may recover McKean's Fredholm determinant formulas [McKean 2003] expressing the evolution of initial data when acted upon by the full CH flow. We also give account of the large-time (and limiting remote past and future) asymptotics and obtain (partial) confirmation of the thesis about soliton genesis and soliton interaction raised in [Loubet 2006].
1. Introduction
The equation of Camassa and Holm (CH) [1993; 1994] is an approximate one-dimensional description of unidirectional propagation of long waves in shallow water.
We study a stochastic Burgers equation using the geometric point of view initiated by Arnold for the incompressible Euler flow evolution. The geometry is developed as a Cartan-type geometry, using a frame bundle approach (stochastic, in this case) with respect to the infinite-dimensional Lie group where the evolution takes place. The existence of the stochastic Burgers flow is a consequence of the control in the mean of the energy transfer from low modes to high modes during the evolution, together with the use of a Girsanov transformation.
Introduction
Many distinguished authors have made notable contributions to the stochastic Burgers equation, of which a small sample appears in our very short bibliography. It is not our purpose to review those contributions; it is perhaps appropriate that we underline here that which seems to us the novelty of our approach.
We start from the viewpoint of geometrization of inertial evolution initiated in [Arnold 1966] and systematically developed in [Ebin and Marsden 1970; Brenier 2003; Constantin and Kolev 2002], based on infinite-dimensional Riemannian geometry; the classical approach of [Ebin and Marsden 1970] is to use Banach-modeled manifold theory; inherent difficulties appear in the construction of exponential charts and in the introduction of appropriate function spaces. We circumvent these difficulties by using the viewpoint [Malliavin 2007] of Itôcharts, Itô atlas; in short Itô calculus makes it possible to compute any derivative of a smooth function f on the path p of a diffusion from the unique knowledge of its restriction f|p. Then no more function spaces are a priori introduced: the path of the diffusion constructs dynamically its canonical tangent space, built from the evolution of the system.
This volume is dedicated to Henry McKean, on the occasion of his seventyfifth birthday. His wide spectrum of interests within mathematics is reflected in the variety of theory and applications in these papers, discussed in the Tribute on page xv. Here we comment briefly on the papers that make up this volume, grouping them by topic. (The papers appear in the book alphabetically by first author.)
Since the early 1970s, the subject of completely integrable systems has grown beyond all expectations. The discovery that the Kortweg – de Vries equation, which governs shallow-water waves, has a complete system of integrals of motion has given rise to a search for other such evolution equations. Two of the papers in this volume, one by Boutet de Monvel and Shepelsky and the other by Loubet, deal with the completely integrable system discovered by Camassa and Holm. This equation provides a model describing the shallow-water approximation in inviscid hydrodynamics. The unknown function u.(x; t) refers to the horizontal fluid velocity along the x-direction at time t . The first authors show that the solution of the CH equation in the case of no breaking waves can be expressed in parametric form in terms of the solution of an associated Riemann– Hilbert problem. This analysis allows one to conclude that each solution within this class develops asymptotically into a train of solitons.
Loubet provides a technical tour de force, extending previous results of McKean on the Camassa–Holm equation. More specifically, he gives an explicit formula for the velocity profile in terms of its initial value, when the dynamics are defined by a Hamiltonian that is the sum of the squares of the reciprocals of a pair of eigenvalues of an associated acoustic equation.
We review some of the recent progress on the scaling limit of two-dimensional critical percolation; in particular, the convergence of the exploration path to chordal SLE6 and the full scaling limit of cluster interface loops. The results given here on the full scaling limit and its conformal invariance extend those presented previously. For site percolation on the triangular lattice, the results are fully rigorous. We explain some of the main ideas, skipping most technical details.
1. Introduction
In the theory of critical phenomena it is usually assumed that a physical system near a continuous phase transition is characterized by a single length scale (the correlation length) in terms of which all other lengths should be measured. When combined with the experimental observation that the correlation length diverges at the phase transition, this simple but strong assumption, known as the scaling hypothesis, leads to the belief that at criticality the system has no characteristic length, and is therefore invariant under scale transformations. This suggests that all thermodynamic functions at criticality are homogeneous functions, and predicts the appearance of power laws.
It also implies that if one rescales appropriately a critical lattice model, shrinking the lattice spacing to zero, it should be possible to obtain a continuum model, known as the scaling limit. The scaling limit is not restricted to a lattice and may possess more symmetries than the original model. Indeed, the scaling limits of many critical lattice models are believed to be conformally invariant and to correspond to Conformal Field Theories (CFTs).
This is a survey article dedicated mostly to the theory of real regular finite-gap (algebro-geometrical) periodic and quasiperiodic sine-Gordon solutions. Long period this theory remained unfinished and ineffective, and by that reason practically had no applications. Even for such simple physical quantity as topological charge no formulas existed expressing it through inverse spectral data. A few years ago the present authors solved this problem and made this theory effective. This article contains description of the history and recent achievements. It describes also the reality problems for several other fundamental soliton systems.
1. Introduction
The most powerful method for constructing explicit periodic and quasiperiodic solutions of soliton equations is based on the finite-gap or algebro-geometric approach, developed by Novikov [1974], Dubrovin et al. [1976b], Its and Matveev [1975], Lax [1975], and McKean and van Moerbeke [1975] for 1+1 systems, and extended by Krichever in 1976 for 2+1 systems like KP. Already in 1976 new ideas were formulated on how to extend this approach to the 2 + 1 systems associated with the spectral theory of the 2D Schr¨odinger operator restricted to one energy level; see [Manakov 1976; Dubrovin et al. 1976a]. These ideas were developed in 1980s by several people in Moscow's Novikov Seminar, as discussed see below. The “spectral data” characterizing the associated Lax-type operators consist of a Riemann surface (spectral curve) equipped with a selected set of points (divisor of poles, infinities). In the finite gap case this Riemann surface has finite genus, and the number of selected point is also finite.