unnuun-nv- '93;- 0‘"...nun-.nuuauu .,.;.........ou l— IES ’ liiilli\iillililg\\l\\i LIBRARY Michigan State L University } This is to certify that the dissertation entitled Some Numerical Methods For Singular Diffusions Arising In Genetics presented by Nacer Eddine Abrouk has been accepted towards fulfillment of the requirements for PhoDo degree in StatiStiCS W4. 4. 5.“ Major professor Date November 20, 1992 MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 PLACE IN nETunN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE w . u U 9 13%, if: g. 11 2 8 12:94 , ‘1 ') [X'gT I MSU In An Affirmative Action/Equal Opportunity Institution cMMnmS-DJ SOME NUMERICAL METHODS FOR SINGULAR DIFFUSIONS ARISING IN GENETICS By Nacer Eddine Abrouk A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Statistics and Probability 1992 ABSTRACT SOME NUMERICAL METHODS FOR SINGULAR DIFFUSIONS ARISING IN GENETICS By Nacer Eddine Abrouk In this dissertation, several different methods for approximating the laws of some diffusions which are sOlutions of stochastic differential equations are considered. Using Ito’s formula we derive an infinite system of differential equations for the moments of the process. This system of ordinary differential equations is very difficult to solve in most cases. A power series method is proposed for which theoretical and numerical studies are made. Using the power series method, we are able to approximate the moments of the diffusion under consideration and to investigate its steady state behavior. The power series method leads to new numerical algorithms for the study of the steady state analysis of the difl'usions under consideration. The Gauss Galerkin method which is an existing method originally introduced by D. A. Dawson in 1980 is also used. For the Gauss Galerkin method we derive an alternative computational algorithm which has proved to be efficient in practice. A detailed comparison between the power series method and the Gauss Galerkin method is presented. These methods are then applied to some stochastic models arising in population genetics. Numerical results relating to these models are presented and discussed. Suggestions for future work and some open problems, such as the approximation of the transition probability function, are mentioned. TO INGRID AND MICHAEL ACKNOWLEDGMENTS I wish to express my sincere thanks to Professors David Yen and Habib Salehi for the suggestion of the problem and all the helpful direction in solving it. Their ad- vice, their encouragement and the amount of time spent on discussions are greatly appreciated. I would like to thank Professors Roy Erickson and Raoul LePage for their time and interest. A special thanks goes to my wife, Ingrid, for her unwaver- ing support and infinite patience. Finally, I would like to thank the Department of Statistics and Probability and the Department of Mathematics, at Michigan State University, for supporting me and providing access to their computing laboratories. iii Contents 1 Introduction 1.1 Motivation ................................ 1.2 An overview of the dissertation .................... 2 Statement of the Problem 2.1 Background Material from Probability ................ 2.2 Background Material from Numerical Analysis ............ 2.3 Statement of the Problem ....................... 3 Numerical Methods 3.1 Assumptions and Derivation of the Moments System of Equations . 3.2 Summary of Previous Methods .................... 3.2.1 Moment Neglect Method .................... 3.2.2 Fixed Nodes Method ...................... 3.2.3 Gauss-Galerkin Method .................... 3.3 The Power Series Method ....................... 4 Some Models From Genetics 4.1 Description of the Various Models ................... 4.2 Solution of the System of Moments .................. 4.2.1 Moment Neglect Method .................... 4.2.2 Gauss-Galerkin Method .................... 4.2.3 Power Series Method ...................... iv 10 15 15 20 2O 22 24 29 38 38 40 4O 42 42 V 4.2.4 Comparison of the Power Series Method and the Gauss-Galerkin Method ............................. 44 4.3 Steady State Analysis .......................... 45 4.4 Discussion ................................ 50 Appendix 53 Bibliography 71 Chapter 1 Introduction 1.1 Motivation Let X = (X.;t 2 0) be a diffusion process arising from the stochastic differen- tial equation: dX, = a(X,)dt + b(X,)dB, where X0 is given , B, is the standard Brownian motion and the functions a and b are the drift and diffusion coefficients, respectively. Such equations model many problems in biological, physical and engi- neering sciences (for more details see Karlin and Taylor, Chapter 15). For a rigorous mathematical treatment of theses models, we refer to Feller, where a review of the theoretical work on diffusion processes as well as their boundary behavior is thor- oughly carried out. Equations of this type play an important role in population genetics. Wright (1931 and later publications) and Kimura derived the appropri- ate discrete models in population genetics. The continuous counterparts of these discrete models, which have been used extensively, are given by the stochastic dif- ferential equation above (see Ethier and Kurtz). Since, in general, the probability law of X“ t Z 0, is difficult to determine, an important task is the problem of numerically approximating the probability law of X;, for t Z 0. Because most population genetic models have singularities at the boundary of the state space of the process X, it is even more difficult to approximate the probability density (when it exists) of X,, for t 2 0, through the Fokker-Planck 2 equation for general drift and diffusion coefficients a and b. However, when the coefficients a and b2 are polynomials, a system of equations for the moments of the process X can be derived through the Ito formula (see Chapter 3 system 3.4). If a is at most linear and b2 is at most quadratic, then this system is closed in the sense that an explicit solution for the process moments is possible. In any other case, this system of equations forms an ”open hierarchy” in which the equation for the n-th moment may have terms involving the (n+1)-th moment or higher. various truncation schemes have been proposed. The simplest and most frequently used is the moment neglect method in which the (n+l)-th moment and higher mo- ments are set equal to zero. In this way the system of moment equations reduces to a closed system of ordinary differential equations. Unfortunately, as we shall indicate below (see Chapter 3, section 3.2.1) the use of the moment neglect (or any of its variants) can lead to misleading results. The Gauss-Galerkin method was originally proposed by Dawson (1981) and then refined and improved by HajJar far, and HajJafar, Salehi and Yen. It combines elements of the method of Gauss quadratures and Galerkin approximation and yields an approximation in the sense of weak convergence for the law of X t by discrete measures. It transforms the prob- lem of solving the Fokker-Planck equation for the density of the law of X, (when it exists) to one for the ”nodes” and ”weights” of the discrete measures (see Chapter 3, section 3.2.3). This method, when applicable, works for a more general class of models which includes models where a and b2 are polynomials. In the latter case, system 3.15 describing the dynamic behavior of the ”nodes” and ”weights” can be viewed as a finite linear system of equations for the moments of the discrete measure. The latter system is a truncation of the infinite linear system of ordinary differential equations satisfied by the true moments of the process X (see Chapter 3, system 3.9). However, the linear system 3.9, when viewed as a system for the moments of the discrete measure, is not closed in the sense that the equation for the (2n-1)-th moment may involve the 2n-th moment or possibly higher moments. Lemma 3.1 of Chapter 3 provides a procedure, called the Hankel truncation, to close this system. The resulting system (system 3.17) is a nonlinear system of 3 2n equations for the moments of the discrete measure. When this system (system 3.17) has a solution on some t interval of the form [0,T] and T is a positive constant independent of n, the Gauss-Galerkin method has proved to provide good approx- imations in the sense of weak convergence of measures (see HajJ afar, Salehi and Yen, 1992). Unfortunately, there are models in population genetics where T is a function of n, causing the solution of system 3.17 to exist only locally with respect to the variable t. However, numerically the solution seems to be global. Moreover, in the Gauss-Galerkin method the initial distribution is required to be nondegenerate. Our work in this dissertation consists mainly of the following contributions: 1. An improved numerical algorithm for the Gauss-Galerkin method when the the coefficients a and b2 are polynomials (see Chapter 3). 2. A new approximation method, the power series method. The power series method is an alternative to the Gauss-Galerkin method. This method yields a way of replacing the infinite system for the moments of the process by a finite system which is obtained by truncating the power series expansions of the first few moments of the process (see Chapter 3, system 3.22). The resulting system is a finite linear system which has a unique global solution and the initial distribution is allowed to be degenerate. The power series method yields a uniform (in the t variable) approximation for any j-th moment of the process. However, the question of whether the power series method gives rise to convergence in the sense of measures, as was the case for the Gauss- Galerkin method, is not resolved. 3. An analysis of the steady state solution through the power series method. 4. A numerical comparison between some of the existing methods and the power series method. 4 1.2 An overview of the dissertation This dissertation is organized as follows. Chapter 2 contains background mate- rial from both probability theory and numerical analysis, and the statement of the problem. We begin with a review of the notions of diffusion processes, transition probability functions and the associated semigroup. We then present some back- ground materials from numerical analysis. This will be followed by a statement of the problem. We begin Chapter 3 with a detailed discussion of the assumptions needed. We also show how the boundary conditions are important for the proper formulation of the problem. This is followed by the development of the moment system of equations through Ito’s formula. We then formulate the Gauss-Galerkin method and the power series method. Also a summary of other previous methods is given and an improved numerical algorithm is included. A detailed study of the power series method is presented and some convergence theorems are established for the finite interval case. The main convergence result (Chapter 3, Theorem 3.1) gives a uniform approximation for any j-th moment of the process. The other convergence result (Chapter 3, Lemma 3.4) gives a characterization of the construction of the approximating functions. In Chapter 4, we present the various genetic models for which computations have been carried out. Several numerical algorithms for the Gauss-Galerkin method and the power series method are presented. We include a detailed comparison between the Gauss-Galerkin method and the power series method, for these models. We then give a steady state analysis and a discussion concerning our numerical results. We end Chapter 4 by providing concluding remarks and an outline of some suggestions for future work. An open problem is the approximation of joint distributions of the process X. A more difficult problem is the approximation of the transition probability function. An examination of the conditional moments, which is done at the end of Chapter 4, may lead to some useful ideas for the approximation of 5 the transition density function (when it exists). The Appendix contains numerical simulations, results related to several examples, and the boundary classification. According to Feller, the boundary behavior of a diffusion process can be classified in two categories: accessible and inaccessible. Furthermore, an accessible boundary can be either an exit or a regular boundary. An inaccessible boundary can be either a natural or an entrance boundary. Discussions of such classifications as how they relate to our work are given in the Appendix. Chapter 2 Statement of the Problem In this chapter, we introduce some definitions and notations needed for the for- mulation of the problem. Some known results from both stochastic analysis and numerical analysis are included. The stochastic analysis results can be found in Ethier and Kurtz, Wentzell and Arnold. The numerical analysis results can be found in Stoer and Bulirsh, Stroud and Krylov. Then we present the statement of the problem. 2.1 Background Material from Probability Let X = (Xht Z 0) be a stochastic process defined on a probability space (9, f, P) with state space E(E C R), and let f}=0{X,,s S t}. Let B(E) be the Borel a- algebra of subsets of E. Then X is a Markov process if the following property holds (almost surely) P(X,+,|}}) = P(X,+, 6 I‘lX,),VI‘ E B(E),Vs,t Z 0. (2.1) Note that 2.1 is equivalent to the following E[f(Xt+.)I-7::] = Elf(Xt+a)lth,Vf 6 B(E),V8,t Z 0, (2-2) where B(E) is the space of all bounded measurable functions on E equipped with the supremum norm. Equation 2.1 is equivalent to saying that ”the past and the future are statistically independent when the present is known”. 6 7 A function P(.,.,.) defined on [0, 00) x E x B(E) is a time homogeneous transition function if 1. P(t,x,.) is a probability measure on B(E), Vs,t 2 0,Va: E E, 2. P(O,x,.)=6, (Dirac’s delta function at x), Va: 6 E, 3. P(., .,I‘) e B([0,oo) x E),VI‘ E B(E), 4. P(t + s,x,I‘) = [E P(s,y,I‘)P(t,x,dy),VI‘ e B(E),Vs,t _>_ 0,Vx 6 E. A transition function P(.,.,.) is a transition function for a time homogeneous Markov process X if P(XH-alft) = P(S’Xh F),V3,t 2 09w 6 B(E)- Let X be a time homogeneous Markov process with transition function P(.,.,.). Then X will be called a diffusion process if X has continuous sample paths and if O Vs 2 0,1: 6 E, e > 0, the following properties are true 5. lim,.., 5-37) f|z_y'>€(:r — y)P(t - 3,1, dy) = 0, 6. lim,_., 5:15;)- flz—yISe(x — y)P(t — s,a:,dy) = a(:r, s), 7. lim,_., “—15 flz_ylzse($ — y)P(t - s, x, dy) = b2(:r), for some measurable functions a and b. The function a is called the ”drift” and the function h is called the ”diffusion”. 2 The standard Brownian motion with mean 0 and variance 0 is a simple example of a difl’usion process with drift 0 = O and diffusion b = a. For a given time homogeneous transition function P(., ., .) and for every f e B(E), we can define a family of operators (Tht 2 0) as follows (inmxx) = (mama) the family (Tht Z 0) defines a measurable contraction semigroup on E (by the Chapman-Kolmogorov property). To each diffusion process with coefficients a and b is assigned the second order differential operator d 1 d2 = — e—2— L_adx+2b dx2’ 8 L( f) can be formally written for every twice differentiable function f and is deter- mined by a and b. L is known as the Kolmogorov backward operator. Its formal adjoint operator L', is known as the Kolmogorov forward operator. The Kolmogorov Backward and Forward equations Let X be a time homogeneous diffusion process with state space E, generated by the stochastic differential equation dX, = a(X,)dt + b(X,)dB,, X0 being given. Let a and b be smooth enough and let p(t,:c, y) denote the probability transition density of X t. Then the right hand side of relations 5 and 6 above coincide with the coefficients a and b involved in the stochastic differential equation above. Moreover, the density p(t, x, y) satisfies the backward equation W = (L).p(t.x.y). (2.3) p(0.w,y) = 6r.-.» (2.4) with suitable boundary conditions on p(t, 2:, y). The density p(t, 2:, 3]) also satisfies the forward equation (Fokker-Planck equation) 9395—5439 = (L‘),p(t,x,y), (2.5) p(0,$,y) = 6(2—y): (2.6) with appropriate boundary conditions on p(t, 2:, y). Equations 2.3, 2.4 plus boundary conditions and 2.5, 2.6 plus boundary conditions form two parabolic initial boundary value problem. Since the process X may ex- hibit various types of behavior at the boundaries of E, it is necessary to impose boundary conditions on the density p(t, x, y). For a more detailed discussion of the appropriate boundary conditions, see Chapter 3 and the Appendix . 9 2.2 Background Material from Numerical Anal- ysis We now present some results from numerical analysis. These results can be found in Stoer and Bulirsh, Stroud, and Krylov. Lemma 2.1 Let ,u be a probability measure on ([0,1], 8) such that its support does not reduce to a finite set. Then, for every n (positive integer) there exist nodes {3”, 1 _<_ k 5 n} and weights {whml _<_ k g n} with the following properties 1. The nodes are unique and lie in (0,1). They are the zeros of p" (where p“ is the n — th orthogonal polynomial associated with the measure u, described below). 2. The weights are unique and positive. 3. V f 6 C2”[0, 1], fol f(:c)dp(a:) = 22:1 wk,nf(xlc,n) + €n(f), where on is an error term given by f‘””(£) 1 2 e.. ° 7.- = -———.2=0.1,---, +1 (Pi—1,1?i—1) ’71 = 0- The measure p..(.) = 22:1 wk,,,6k,n(.) is called the Gauss-Christofi'el measure asso— ciated with the measure p. 10 Lemma 2.2 Let the hypotheses of Lemma 2.1 be satisfied. If we define the sym- metric tridiagonal matrix Jn = Dn + Un + (Un)", where D” and U” are given by (610 o o) 0.520 o Du: I: I . 00...6,,_10 (00... o 6,.) Manama) 0073...0 U“: ... 00...0'y,, (oo...o 0) Then it is known (for a proof, see Stoer [4]) that Jn has n distinct eigenvalues which are the zeros of P" (hence the nodes of the measure an). Moreover, if we let A.- be the i — th eigenvalue of the matrix J,, and v,- be the corresponding normalized eigenvector, then the weight 212,-," is given by 10.3.. = ((0002. where (v,)1 is the first coordinate of the eigenvector v,-. 2.3 Statement of the Problem We are concerned in this dissertation with numerical solutions of the probability law of a class of stochastic differential equations of the type dXt = 0(Xt)dt+b(Xt)dB¢, (2.7) X0 = given, (2.8) where B. is the standard Brownian motion, and the functions a and b are known as the ”drift” and the ”diffusion” coefficients, respectively. Equations 2.7 and 11 2.8 model many problems in the physical, engineering and biological sciences (see Karlin and Taylor, Chapter 15). Under suitable assumptions on the functions a and b, the existence and uniqueness of the solution of the above initial value problem is known and the state space can be one of the following: ...: . (-00, +00), M . [0, +00), . [1, r), 4. (l, r), 09 5. [l, r]. where -00 < l < r < 00. In this thesis, we are concerned with the case where the coefficients a and b2 are polynomials (in the spatial variable 2:). Under more stringent conditions on a and b (see Chapter 3, section 3.1), it can be shown that the state space for the diffusion X can be taken to be the finite interval [1, r]. Without loss of generality, the state space can be taken to be [0, 1]. By Ito’s formula (see Chapter 3, section 3.1), it can be shown that the following basic equation holds dE[¢(X.)l dt = E[(L¢)(Xr)l. W5 E 62W. 11- (2-9) D. Dawson suggested a Galerkin type method for approximating the law of X. by discrete measures. This method transforms the problem of solving the Fokker- Planck equation for the density of the law of X 1 (when it exists) to one for a system of nonlinear ordinary differential equations for the ”nodes” and ”weights” of the discrete measures (see Chapter 3, system 3.15). This method, when applicable, works for a more general class of models which includes models where a and b2 are polynomials. In this latter case, system 3.15 can be viewed as a linear system of 2n equations for the moments of the atomic measure, where n is the number of nodes of the discrete measure. This linear system of equations may not be closed in the sense that these equations may involve the 2n - th or higher moments. Lemma 3.1 of Chapter 3 provides a procedure called the Hankel truncation to close this system. Then the resulting system (see Chapter 3, system 3.17) is a nonlinear 12 system for the first 2n moments of the discrete measure. Systems 3.15 and 3.17 are both nonlinear systems. System 3.15 is more general than system 3.17 in the sense when the coefficients a and (or) b2 are not polynomials, it is still meaningful while system 3.17 is not. It should be pointed out that when system 3.17 can be formulated, the problem of existence and uniqueness of its solution is equivalent to the problem of existence and uniqueness of the solution of system 3.15, although the moment system 3.17 is more tractable (see Chapter 3, section 3.2.3). When system 3.15 has a solution on some t interval of the form [0,T] and T is a positive constant independent of n, the Gauss-Galerkin method provides good approxima- tions (see HajJ afar). Unfortunately, there are models in genetics (see models 1 and 3, in Chapter 4) where T is a function of n causing the solution of system 3.15 to exist only locally. In the latter case the existence of T as a positive constant must be established numerically. Moreover, in the Gauss-Galerkin method the initial distribution is required to be nondegenerate (for more details, see Chapter 3 of the thesis). Our main aim in this dissertation is to present a new method, the power series method, for which theoretical and numerical studies are made (see Chapter 3 of the thesis). The boundary behavior of the process is especially of importance for the complete formulation of the system of equations for the moments of the process which is derived based on equation ?. In Chapter 3 we show how equation 3.4 is more basic than the Fokker-Planck equation, since 3.4 does incorporate the boundary behavior of the process as well as its behavior in the interior of the state space, while the Fokker-Planck equation governs the behavior of the density only in the interior of the state space. When specialized to functions of the type ¢(a:) = 1:", for k E N, provided they are in the domain of L, equation 3.4 yields an infinite system of ordinary differential equations for the moments of the process. This system is open since the first n equations may involve n +1 moments or more. The number of higher moments involved in the equation for any n — th moment depends on the degree of the polynomials a and b2 (see Chapter 3 for more details). The approximation suggested by the power series method is simple in nature. It takes l3 advantage of the power series expansions of the first few moments of the process X. This method is based on the following observation L(x") = lsa(:r:).v‘"l + EUC—g-flb2(x)xk'2, where L is the Kolmogorov backward operator associated with the process X. If a is the degree of a and B the degree of b2, then L(x") is a polynomial of degree at most V(a,fl, k)=maa:imum{a + k — l,fl + k — 2}. When the first 11 moments are analytic functions in t, their respective power series are truncated at the same index n, i.e. from the infinite series expansion of any i-th moment of the process, where O _<_ i S V - 1, we retain the first (n + 1) terms. The V finite sums resulting from this truncation involve (n +1)z/ coefficients. These (n +1)V coefficients can be found uniquely, by retaining the first (n + 1)V equations of the original infinite open system. The full details can be found in Chapter 3. Based on the above truncation, the resulting system has a unique global solution (see Chapter 3). Then we establish two main convergence results and give a detailed steady state analysis. Our first convergence result (Lemma 3.4) provides a way of constructing the approximating functions needed for the power series method. The second result (Theorem 3.1) concerns the uniform approximation for any It — th moment of the process X by a sequence of functions. We also provide a more efficient numerical algorithm for the Gauss-Galerkin method. This numerical algorithm is an improvement over the existing one which was suggested by Dawson. The latter scheme uses several inversions of Vandermonde matrices in order to compute the ” nodes” and ” weights”. As the number of ”nodes” and ”weights” increases, this numerical scheme becomes ill-conditioned due to the near singular behavior of the Vandermonde matrices. Our numerical scheme is based on an eigenvalue problem for the ”nodes” and ”weights” and it reduces the ill-conditioning considerably. The proofs of these theorems require techniques from both analysis and probability theory. Specifically, the proofs involve the use of differential inequalities and results from the classical problem of moments. A number of numerical results related to Models 1, 2 and 3 are presented and, whenever possible, compared with the exact solutions. For Model 1 the exact solution for the density of the law of X , is explicitly known. For Model 2, an explicit solution for the moments is available. Moreover, for Model 2 14 an explicit solution for the absolutely continuous part of the law of X t is available. Several computer simulations are also made for all three models. They can be found in the Appendix of this thesis. Chapter 3 Numerical Methods We start this chapter by stating the assumptions needed for the formulation of our problem. This is followed by a detailed derivation of a system of ordinary differential equations for the moments of the process X (system 3.9 below). Then a summary of some existing methods which have been used for the various truncations of this infinite system of equations is stated. These methods are: the moment neglect method, the fixed nodes method and the Gauss-Galerkin method. An improved numerical algorithm for the Gauss-Galerkin method when the coefficients a and b2 are polynomials is then presented. This is followed by a detailed study of the power series method indicating two convergence theorems. The main convergence result (Theorem 3.1) gives a uniform approximation (with respect to the t-variable) for any k-th moment of the process X. The other convergence result (Lemma 3.4) gives a characterization of the closure problem for the power series method. 3.1 Assumptions and Derivation of the Moments System of Equations Consider the stochastic differential equation dX, = a(x,)dt+b(X,)dB, (3.1) 15 16 where X0 is given, B, is the standard Brownian motion and the coefficients a and b satisfy the standard conditions for the existence and uniqueness of the solution. Under the following assumptions, it is known that the solution of equations (3.1) and (3.2) is a Markov process with state space [0,1] (see Ethier and Kurtz) A1. a,b 6 C[0,1], A2. b(:r) > 0,V:r 6 (0,1), A3. X0 is independent of the o-fields o{B¢;t 2 0}, A4. a(0) 2 O,a(1) 5 0 and b(0) = b(l) = 0, A5. D(L) contains all the monomials (where D(L) is the domain of the operator L). Remark; It is easy to show the assumption b(O) = b(1) = 0 implies that neither of the boundaries of the state space is regular. Moreover, if in assumption A4 we have a(0) = a(1) = 0, then it can be shown that assumption A5 holds. Now we derive the basic equations which will be the starting point of our work. Let ¢ 6 C 2[0, 1] fl D(L), then by applying Ito’s lemma to equation (3.1) we have the following equation 1 d¢(Xt) = {a(X.)¢’(X.)) + 552(X:)¢"(X:)}dt + b(X:)dBo this implies t 1 t ax.) - ax.) = /. {a(x..)¢'¢"1 = [E[(L¢)(X.)ldu- (3.3) Using the strong continuity of the semigroup (T,;t _>_ 0) corresponding to the diffusion process X, the function: t ——r E [(L¢)(X t )] can be shown to be continuous. Therefore, by using (3.3), we obtain fiEMx.) = E[(L¢)(X.)1, (3.4) X0 = given. (3.5) 17 This equation always includes the boundary contributions, i.e., contributions from the singular part and the absolutely continuous parts of the law of X,. The deriva- tion of a similar equation is possible starting from the Fokker-Planck equation. In this latter case the boundary contribution may not be accounted for. The following analysis justifies the importance of the boundary conditions. Consider the parabolic initial boundary value problem 2.3 and 2.4 for the function p(t,x,y) (see Chapter 2). We multiply equation 2.4 by ¢ 6 D(L), then the resulting equation is integrated by parts, twice with respect to the variable x. The following equation is then obtained %A1p(t,x)¢(x)dx = ((15, L'p) = (1145.1?) + T, where T is a boundary resulting from integrations by parts. When T=0 we obtain the system %A1P(t,$)¢(fir)dx = /01[(L¢)($)]p(t,x)d:r, (3.6) P(Oflv) = given. (3.7) These two equations represent counterparts of 3.4 and 3.5. Unlike equations (3.4) and (3.5), equations (3.6) and (3.7) generally do not include the boundary behavior of the process X. As an illustration, we give a few examples where T=0 and D(L) contains all the monomials (for more details see Ethier and Kurtz, page 366). Brownian Motion: For the standard Brownian motion with mean 0 and variance t, the corresponding Fokker-Planck equation for the transition density is given by 0100.3. 31) _ 102mm. 31) 6t 2 63/2 pm. :8. y) = 6(x - y)- ,V$,y E (—00,00),Vt 6 (0,00), It is easy to show that this initial value problem has a unique solution given by 1 1. _ 2 130.1331) = fiexpkgT'z/L) 18 and the law of X. is absolutely continuous with respect to the Lebesgue measure on (—00, 00). Its density p(t,x) is obtained as follows p(t, y) = [190.3, y)p(0.$)dx- Therefore, in this case no boundary conditions are required for the formulation of the problem. Other Models: a E 0 and b(:c) = {RH—:5, 7 > 0. When 7 = $3 this model is one of the Wright-Fisher models which arises in population genetics. For this class of models, the state space is [0,1] and the boundary term T depends on the magnitude of the parameter 7 as follows for 7 > £3 T reduces to —%[p(t, 1)¢(1)]. In this case it can be shown that T 76 0. Therefore boundary conditions at the end point 1 must be imposed on the function p(t,x,y). For the case where 7 = % one can show that boundary conditions at both end points 0 and 1 must be imposed on p(t,x,y)- In our studies, we consider equations 3.1 and 3.2 with polynomial coefficients whose state space is the interval [0,1]. Therefore, we set the following conditions a 19 a(:v) = z: Akx", and b2(:c) = Z Skxk, (3.8) k=0 k=l where 0,,6 E N and A055 75 0. It can be shown when a(0)=a(1)=0, the domain of L contains all monomials (see Ethier and Kurtz, Chapter 12, page 367). By specializing equation 3.4 to ¢(x) = xi , where j 6 N U {0}, we get dM- t . ° ‘9 —th—(2 = J z Aij+k_1(t) + 91' z: Sij+k_2(t), (3.9) k=0 k=l Mk(0) = given, (3.10) where Mj(t) = E(th) and 0,- = 312;” . The system of equations 3.9 and 3.10 is an infinite system of linear equations for the hierarchy of the moments of the process 19 X. This system is closed if and only if osa+j-13j and 0_<_fi+j-2_<_j,VjeNU{0}, this is possible only when a 5 l and ,6 5 2, in which case the exact solution of this system is given recursively by t Mj(t) = [Mj(0) - Mj_1(0) + 91'!) Mj_1(8) exp(—0js)ds] exp(0jt). If a Z 2 or fl 2 3 then 3.9 is not a closed system and therefore cannot be solved ex- plicitly. Our task is to close the system 3.9 in such a way that the resulting solution yields a good approximation to the original system 3.9. Various truncation schemes have been proposed to close the system. For some models which arise in Genetics, the derivation of system 3.9 and 3.10 is not possible from the Fokker-Planck equa— tion. This is an important observation because when the boundary behavior is not incorporated, the steady state analysis for the process X is not possible. We em- phasize the importance. of the boundary behavior through the following discussion. Consider the following class of models which arise in population genetics quite often: a(:c) = x’(1 — x)’ and b(:r) = x”(1 — 1:)", where r, 3,2}? and 2q 6 N. Note that a and b2 can then be written in the following form a l3 a(:c) = z Akx", and b2(:r) = X 51.3,”, k=0 k=l witha=r+sandfl=2(p+q). For a E 0, we treat the following cases here. The discussion of the more general case is contained in Appendix. Case 1: p,q 2 1 (this implies that both 0 and 1 are natural boundaries), we have T E 0 and <¢, L'p) = (Lo, p),v¢ e C2[0,1]. 20 Case 2: p = %,q 2 1 (this implies that 0 is an exit boundary and 1 is a natural boundary), T is given by 1 T = —§¢(0)p(ta 0): which implies <¢.L*p> = (lap) — §¢(0)p(t.0).\r¢ 6 070,1]. Case 3: p Z l, q = z],- (this implies that 0 is a natural boundary and 1 is an exit boundary), T is given by l T = -§¢(1)p(ta 1): which implies (¢.L"p) = sap) — grow, an e C2,, 11. Case 4: p = q = i; (this implies that both 0 and 1 is are exit boundaries), T is given by T = -§[¢(1)p(t. 1) + ¢(0)p(t. 0), which implies (45,131?) = — %[¢(1)p(t.1)+ ¢p(t,o>1,v¢ e 62m. 11. Cases 2, 3 and 4 require special analysis because of the boundary behavior of the process. At an exit boundary the density p(t,y) > 0 (see the Appendix) which implies that T 96 0, for some <15 6 CZ[0,1]. 3.2 Summary of Previous Methods 3.2.1 Moment Neglect Method This is the simplest and most frequently used method. If the state space of the stochastic process X is the interval [0,1], then all the moments exist (they are all 21 bounded by 1, uniformly in t). The following observation is the basis for this method: for each fixed t we have 1 1 0 5 M..(t) = E(Xl‘) =/0 x”p(t,:r)dx 5 D(t)]0 as” = D(t);—.1;T —r 0, as n —+ 00, where D(t) is assumed to be finite and is defined by D(t) = sup{p(t,:r);a: 6 [0,1]}. Therefore in system 3.9 it is reasonable to set M..(t) equal to zero for large values of n. The method consists in setting the (n+1)-th and higher moments equal to zero. In this way the infinite system of moment equations for the process X reduces to a closed finite linear system of n equations. There are two basic difficulties with this approach (or variants of it). The method leads to an unbounded solution and it does not lead to convergence. For a more detailed discussion, see Chapter 3. As an illustration of the moment neglect method, we present the following model (this model will be referred to as model 3, see Chapter 4 for this and two additional models) 050 and b(:1:)=:r\/1-:r, for this model system 3.9 yields the following equations M120) = BkleU) - Mk+l(t)l1Vk E N U {0}. for k large enough, MI,“ is neglected leading to the following finite system of linear equations m;(t) = 0,-[mJ-(t) — mj+1(t)],Vj = 0, 1,2, ..., k — 1, "15.0) = 9kmk(t). the last equation implies mk(t) = mk(0)exp(0kt). Consequently, mic—1(t) = gk—llmk—l(t) - mk(t)l. implies 22 t mic—1U) = [mic—l - aka/0 mk(3) eXP(-0k-13)d3l eXP(9k-1t). and in general we have the following recursion valid for j = 2, 3, ..., k + 1 mj-1(t) = [mj_.1 — 01--1A:mj(s)exp(—0,-_1s)ds]exp(0j_1t). This method may give satisfactory results only for very small values of t. It is observed (see Chapter 3) that the first few moments have exponential growth and exceed 1 even for small values of t hence giving misleading results. 3.2.2 Fixed Nodes Method In this method 11 nodes are fixed according to a given distribution (the distribution of X0). Then 3.9 can be made closed by expressing higher order moments in terms of lower order moments as follows. If we let 2:1(n,0),:r2(n,0), ...,xn(n,0) be the n nodes of the Gauss-Christofi'el measure corresponding to X0 (see Chapter 2, Lemma 2.1) and let ”(nat) = Z wi(ntt)635(fl,0)t i=1 then in system 3.9 when the moments of the process X are replaced by the moments of the measure u(n,t), it can be shown that the weights of u(n,t) satisfy the following initial value problem VW’ = BW, (3.11) W(n,0) = given, (3.12) where (W(n,t))T = (w1(n,t),w2(n,t),...,wn(n,t)), and u(n,0) being the Gauss- Chfistofiel measure associated with X0, V=v.-, ={z{,.‘=1,2,...,n;j=0,1,...,n— 1}, and B = {bgj} = {[L(1")](x,-);i = 1,2, ...,n;j = 0,1,...,n -1}, 23 this system has a unique solution given by W(n, t) = W(n, 0) exp[t(V"l)B]. To express the higher moments of p(n, t) in terms of the lower moments, we proceed as follows. Define the matrix V(k) by ( (21):" (2:2)“ (33)“ my» \ (2:1)Im+l (x2)kn+l ($3)lm+l . . . (1.")!an V(k) = ($1)lcn+n—2 (x2)kn+n-2 . . . (xn_l)kn+n—2 (xn)lcn+n—2 \ (1'1)""+."‘l ($2),m+n—l . . . (xn_1)k”+“‘1 (xnyflH-n-l } Note that V(0) is the usual Vandermonde matrix. We define the following vector m(k, t) = (mien: mkn-i-l: "'a mkn+n-l)a where m,-(n,t) = 233:, wj(n,t)[xj(n,t)]. Then for each k=0, 1, 2, ..., we obtain the following system m(k,t) = V(k)W(n,t). For k=0, we have W(n,t) = V(0)“M(0,t). which implies m(k,t) = V(Ic)[V(0)‘1M(0,t)]. The fixed nodes method is easy to implement and often yields satisfactory approx— imation of the moments of the process X for small values of t. This method is similar in nature to the well known Newton Cotes method for approximating inte- grals. Unfortunately, we do not have a convergence theorem of the approximating moments to the true moments, as n —> 00. In the following section we present the Gauss-Galerkin method. In this method the following assumption is required: A6. Equations 3.4 and 3.5 above, uniquely determine a measure for all t E [0,T]. 24 3.2.3 Gauss-Galerkin Method We start by reviewing the Gauss-Galerkin method. Then we shall present a new numerical algorithm for the implementation of the method. Let um, be the Gauss- Christoffel measure associated with the law of X, (see Chapter 1). Then system 3.9 can be written as follows 2,1,1; n..‘1 = [21 ...,(r)L(g,,(.)y] + 6...”, for all i=0, 1, 2, ..., n, . This system cannot be used to find the weights 17,, J(t) and the nodes (“J-(t) because of the error term involved in the above system. The Gauss—Galerkin scheme is obtained by dropping the error term in the equation above. This way, as stated in Chapter 2, we obtain a sequence of discrete measures n “mt = Z wnJ(t)61’u.j(‘)’ 1:1 where the weights wad-(t) and the nodes 1:" J(t) are solutions of the nonlinear finite system of dynamic equations i123, WWW”)? = gaseous-(0):], (3.13) wnJ(0) = 77n.j(0) and mud-(0) = 6mm). (3.14) for all i=0, 1, ..., 2n-1. The measure 11...: = z wn,j(t)6:r.,,-(t) i=1 is called the Gauss-Galerkin approximation of the law of X,. If the nodes and the weights are differentiable functions of t, then the dynamic nonlinear finite system 3.13 and 3.14 can be written as follows: A(n,t)X(n,t)' = D(n,t)X(n,t), (3.15) X(n, 0) = given, (3.16) where the initial conditions and matrices A(n,t) and D(n,t) are defined as follows: X(n1 0) = (nn,l(0)a "713(0): "-1 nu,n(0)a £n,1(0)1 £n,2(0)1 "'3 £n,n(0))a 25 A A A(n,t) = 11 12 , A21 A22 D 0 D(n,t) = 1 , D2 0 A11 = {(xnj)i}n>(n: A12 = {wnJKanl‘l'hxm A21 = {(xnj)i+n}nxvn A22 = {wnJ[($n.j)‘+"l'}nxm DI = {L($‘)($n.i)}nxm D2 = {L(x‘+")(xn,j)’}..x... The existence and uniqueness of the solution of system 3.15 has not been settled except for some special cases. In case the coefficients a and b2 are both polynomials, system 3.15 describing the dynamic behavior of the nodes and weights can be viewed as a finite linear system of equations for the moments of the discrete measure. The latter system is a truncation of the infinite linear system of ordinary differential equations 3.9 which is satisfied by the true moments of the process X. However, the system 3.15, when viewed as a system for the moments of the discrete measure, is not closed in the sense that the equation for the (2n-1)-th moment involves 2n-th moment or possibly higher moments. A procedure called the Hankel truncation is used to close this system. The resulting system is a nonlinear system of 2n equations for the moments of the atomic measure. To do this, the Hankel determinants of order higher than n are set equal to zero. The Hankel determinants are defined as follows, for each r 6 N U 0, mo m1 . . . m, m1 m2 . . . m,.+1 mr mr+1 e e 0 m2,- 26 and m1 771.2 . . . mf+1 m2 m3 . . . mr+2 r, = mr+1 mr+2 - ° - m2r+l Lemma 3.1 Given a sequence of numbers {m,-,i Z 0} there exist a unique atomic measure whose support consists of exactly 11 nodes if and only if I. A, > 0,Vr E 0,1,...,n — 1, 2. I‘, > 0,‘v’r E 0,1,...,n — 1, and 3. A, = Pr = 0,Vr 2 11. Both Dawson and HajJafar assumed (i) and (ii). The assumptions are then ver- ified numerically (a theoretical proof is not available). As a consequence of these assumptions the higher moments are expressed in terms of the lower moments as follows An(t) = An-1(t)m2,,(t) — g[mo(t), m1(t), ..., m2"-1(t)], where g is a polynomial. By setting An(t) = 0, we obtain g[mo(t), m1(t), ..., 7712"_1(t)] A.._1(t) m2n(t) = = f[m0(t), m1(t), ..., m2"-1(t)]. Let’s give an illustration of the method through model 3 (i.e., a E 0, b(x) .-= le - x. Then the system of equations for the moments of the process takes the following form m;(t) = 0,-[mj(t) - mj+1(t)],Vj = 0,1, ...,2n - 2, man—1(t) = 62n-llm2n-l(t) - flm0(t)r ml(t)r '"r m2n-l(t)]a which can be written as m'(n,t) = A(n)m(n,t)+F(n,t), (3.17) m(n,0) = given, (3.18) 27 where m(n,t) = (mo(t),m1(t), ...,m2n_1(t)), F(n,t) = (0,0, ...,0,f[mo(t),m1(t), ...,m2n_1(t)]), .401) = {055},i,j = 0,1, ...,2n — I, where 0,- if i=j aij = -0,~ ifj=i+1 0 otherwise. Lemma 3.2 The system 3.17 and 3.18 has a unique solution if and only if there exist a neighborhood [0, T] such that Vr = 0,1, ..., n - 1 and t E [0, T], the following holds 1. A.(t) > 0, 2. I‘,(t) > 0. (The proof is straightforward). Lemma 3.3 Let the 2n functions {m,-(t),0 5 j 5 2n — 1} obtained in Lemma 3.1 be the first 2n moments of an atomic probability measure p(n,t). Then Vt E [0,T],V1,ZI 6 C[0,1], then as n —r 00, the following is true [01 mm t) —> [01¢(X.)dP. The nodes and weights of the discrete measure p(n,t) can be found through the usual inversion of the well known Vandermonde matrix. This matrix is given by V(0) of section 3.2.2 above. It is easy to see that the determinant of V(0) is directly proportional to n H [(121.40 - x..,,-(t)]2. 5J=1.i¢j This quantity may tend to zero as It tends to infinity. As a consequence V(0) may become nearly singular, causing the problem of solving for the nodes and weights of the measure u(n,t) to become ill-conditioned. An alternative method which 28 considerably reduces the ill-conditioning when computing the nodes and weights is the content of the numerical algorithm below. An Improved Numerical Algorithm for the Gauss-Galerkin Method Our numerical algorithm for finding the nodes and weights of the Gauss-Galerkin measure p(n,t) is based on Lemma 2.1 and Lemma 2.2 of Chapter 2. Lemma 2.1 gives a relationship between the orthogonal polynomials associated with p(n, t) and its first 2n moments {mJ-(t), 0 5 j 5 2n - 1}. Lemma 2.2 provides a way of com- puting the entries of the tridiagonal matrix Jn. We shall now present the detailed derivation of the algorithm. Given the 2n moments {m, (t), 0 5 j 5 2n — 1}, the inner products involved in the construction of the matrix J" can be expressed in terms of the given moments as follows (P01P0> = m0) ($130,170) = ml, in general, given 2.‘ , (Path) = Zagz')m2,-_j,where0 5 i 5 n — 1, i=0 and the a?” are known. Then 2.‘ (xpirpi) = Z a§2i)m2i-j+l- j=0 Through the three term recursion formula we also have the following ($191.19.) - 5i+1 (Pi+1.P.'+1) = aligning + - -- + Gigantic - 25i+1(082i)m2i+1 + - - - + Ohiomll + (6.,1)2(a32‘)m2. + . . . + agimo) — (Palms-1) + (’Y.‘+1)4(Pi—1.Pi-1)- 29 We now describe a one step computational algorithm from t = 0 to At. This computational algorithm is used to compute the approximate values of the first 2n moments of the process (the process arising from model 3) at time At using the initial conditions {Mk(0),0 5 k 5 2n - l}. 1. Given {Mk(0),0 5 k 5 2n — 1} we have (for the specific example under consideration) M2“) = 0..[M,.(t) - Mk+1(t)l which implies MUM) = Mk(0) + othtm) - Mk+1(0)l- 2. With the new mk(At), k = 0,1, ...,2n — 1, we compute 6,,(At), ..., 6,,(At) and 7,,(At), ..., 7,,(At) through Lemma 2.2 of Chapter 2. Then we form the matrix J..(At). The eigenvalues of Jn(At) and their corresponding eigenvectors are then found. This yields the nodes {x,-(n, At), j = 1, 2, . . . , n} and the weights {w,-(n,At),j =1,2,...,n}. 3. Using the first iterates, we obtain the second and higher iterates from {mk(At), k = 0,1,. . . , 2n — 1} through the following formula mk((i + I)At) = mk(iAt) + 0k(At)[mk(iAt) - mk+1(iAt)], where i 2 1. This numerical algorithm has proved to perform better than the algorithm proposed by HajJ afar and Dawson in which several matrix inversions are required causing the problem of finding the weights and nodes to become ill-conditioned. Our algorithm does not require any matrix inversion. Therefore, our numerical algorithm is more suitable when higher values of n need to be considered. 3.3 The Power Series Method When a and b2 are polynomial functions, system 3.9 of equations for the moments of the process X is closed if and only if \ 30 OSa+j-1_<_j and 05fl+j-2SJ.VJ'€NU{0}. this is possible only when a 5 1 and fl 5 2. In the latter case the unique solution of 3.9 is easily obtained. From 3.9 it is clear that Vi E N, M.- 6 C°°[0,1],VT > 0. In the Gauss-Galerkin approach the existence and uniqueness of the truncation of 3.9 is known to hold only on a t interval of the form [0,T(n)], where T(n) is a positive function of n. For more details we refer to Chapter 4. The power series method is based on a different kind of approximation. The two approximations (the Gauss- Galerkin and the power series) are not in the same sense. The Gauss-Galerkin method, when applicable, yields an approximation in the sense of convergence in measure. The power series method gives an approximation in the sense of uniform convergence and is applicable only when the coefficients a and b2 are polynomials. For a more detailed comparison of the methods, we refer to Chapter 4 where several examples are illustrated. If a and fl are integers such that a 2 2 or fl 2 3 then 3.9 is not a closed system and therefore cannot be solved explicitly. Our task is to close the system 3.9 in such a way that the resulting solution yields a good approximation. For each i 6 N U 0, the differential equation involved in system 3.9 is 4M“) a [3 dt = j Z Aij+k_1(t) + 0]: 2 Sij+k_2(t). (3.19) le=0 Ir=1 When a 2 2 or fl 2 3, system 3.19 is not closed. It is impossible to explicitly solve equations 3.19. It is assumed that all the moments have power series expansions in t. In fact, as we shall show below, it suffices to assume the first few moments to be analytic in t. We then approximate these moments by polynomials. The details are given below. Depending on the parameters a and ,6, the following cases are considered: (i) ,6 at a +1, (ii) ,8 = a +1. In case (i) there are two possibilities, either fl > a +1 or fl < a + 1. For simplicity, we shall treat the situation where fl > a + 1. The case where 3 < 01 + 1 can be treated similarly. We now consider case (i). 31 Equations 3.19 above can be written as follows: . 544-2 flit“) = _Z Bs.M(t) (3.20) where Bid = B,-J(i,0.-,A,-, 5,) is given by B _ iAj...'+1+ 0,512.5.” If! -' 1 _<_j S a +1— I h] 0;Sj-5+2 ifa+i5j5fl+i—2. Suppose that the moments M1, M2, ..., M54 are analytic functions on [0,T]. Then for i=1 equation 3.20 gives All“ )' 21—2 BIJMJ' Bl’p- ’ Mrs-Kt): which implies M34 is an analytic function. Note that 81,5..1 ¢ 0. In fact B{,fi+i_2 = 0,3,3 9f 0, Vi E N. For i=2, equation 3.20 leads to M50) - 2f;1132,ij MW) = 32x . which implies Mp is an analytic function (since M54 is analytic). More generally, equation 3.20 implies M,- is an analytic function for all i 2 B — 1. Therefore the following expansions are possible M,(t) = Ecj(i)tj,Vi = 1,2,3, ...,n - 2, j=0 where j!c,-(i) = Mij)(0), for allj 2 0. Polynomial Approximation Fix any integer n and define the following polyno- mials n . tj m,(n,t) = 29(1):, (3.21) i=0 3' where i = 1, 2, ..., fl — 2. We then consider the initial value problem d , t 5+"? "'(n )= 2.: B,,,m,-(t) (3.22) j=i- -l 32 subject to the condition "11(0) = M'(0)1 for all i = 1, 2, ..., u, where u = n(fl - 2). The number of unknowns cj(i) involved in power is (n +1)(fl - 2) (note that we have co(i) = M,(0), for all i = 1,2, . . . ,fi - 2. System 3.22 is not a closed system. In the following lemma we suggest a way to close it. Lemma 3.4 Using the system 3.22 the coeflicients in (3.21) can be determined by the the initial conditions M,-(0), i = 1,2, . . . , (n + 1)(fl — 2). 3:991; The coefficients involved in 3.21 are 61(1) 61(2) 61(3—2) (32(1) 62(2) 02(3—2) cn(1) c,,(2) cum—2). The coefficients co(1), co(2), . . . , Com — 2) are uniquely determined since by defini- tion we have co(i) = M.-(0), for all i = 1, 2,... ,fl — 2. To determine the coefficients c1(1),c1(2), . . . ,c1(fl — 2) we consider the first fl — 2 equations involved in sys- tem 3.22. These equations are then evaluated at t = 0 to yield the following fl-i-i-2 61(1) = Z B.ij(n,0), j=i-l for all i = 1,2, . . . ,fl - 2. To determine the coefficients c2(1),c2(2), . . . ,c2(,6 - 2) the first (fl — 2) as well as the next (,6 — 2) equations are needed. This is clear from the following equation which is obtained by differentiating the i — th equation in system 3.22 c12m__,-__(n,t)= “"2 dm (n, t) cit—T 2 B‘”— —_J__d ’ j=i- l which can be written as 13+i-2 [NJ-2 2 Bid 2 Bjkmk(n,),t j=i-—l k=j- -1 33 then “1,4 2lc2(i) = X j = i — 13+‘-2B.- J 2 B-,,.M,.(0). k=j-l Continuing in this fashion, all the coefficients {c,-(i),1 5 j 5 n, 1 5 i 5 fl — 2} are uniquely determined through the relation m2+2_5(n,t) - l:=li+2-B Bi+2—fl,lcmk(na t) Bin—M The total number of equations required is n(fl - 2). The number of initial condi- tions used is (n + 1)(fl - 2). This completes the proof. mi(nrt) = Now we turn to the question of approximation of the moments of the process X. The following Theorem provides a uniform approximation of any I — th moment of the process X. Theorem 3.1 Suppose assumptions A1 through A5 hold and assume that the first fl - l moments of the process X are analytic. Fix any integer l and consider the sequence of functions {m,~(n,t), n _>_ 55—2}. Then for any To 6 (0,T), this sequence converges uniformly (in t) to M;(t), on [0,To], as n -r 00. Emf; The following elementary result from the theory of power series (see Smirnov, page 150) is needed. This result, when applied to the analytic moments {M,~(t), 1 5 i5 fl—2}, states the following for each i = 1, 2, . . . ,fl—2, VI 6 NUO, VTo E (0,T), HM! - mi(n.t)ll -: 0." -* 00 (I) where [I I] is the supremum norm on [0,T]. Let I 6 N, then we have 1. If 1 5 l 5 fl — 2, then by the definition of the functions m,(n, t), the theorem holds by applying relation (I). 2. If I _>_ fl — 1, then the (l — (B — 2)) - th equation from system 3.22 above gives 1 l—l mi—a+2(n,t)= Z BI—a+2.jmj(",t)= Z Bl—B+2.jmj(nvtl+Bl—B+2Jml(”at) j:i-B+2 j:l—fi+2 34 which implies mi— s+2(n t) 2j=l—fl+2 B1_,9+2ij(n t) Bl—fi+2.l m1(n,t)- — Consider the first l— 3 + 2 equations from system 3.22. The first of these equations leads to :fi_2 m’1(n,t)— oBlJmJ-(n, t) 31).- mfi-1(n1t) = which implies m’ n ’ - 2 1J mJ-n, j lMa—1(t)-ma—1(t)l=|[ 1”) Mm] 2;? l ( t) Mun, therefore m’ n,t M’( t()l m n,t M-t 1( B)- l l‘l'lZlBldl 1(B)- J()[—r0,n-+OO, lfi-l Bl,B-l 9 [M5_1(t)—m3_1(t)| S I by part 1 above. From the second equation (of the first I — fl + 2 equations) we have m2(nrt) - Egg-:11 B2ij(nrt) = 32,8 ma(n.t) = m’2(n,t) — 233:21B2ij(n, t) _m’1(n, t)— 2%: BlJm,(n, t) Again because the moments of the process satisfy system (3.9) we can write m2(nrt) _ M2(t)|+ B2 ,. WW) - W(tll S l 5‘2 m-(n t) - M-(t) B . J i J + 12:; l 2d B2,fl l ,m’1(n.t) - M10) B2’fllBl’fi_l $31M, 0- M.- 00. This completes the proof. The case where [i = a +1 requires the following assump- tion iAa + 0‘55 ¢ 0,Vi E N U {0}, and a convergence theorem like Theorem 3.1 can be proved in a similar way. Now we give sufficient conditions for analyticity of the moments of the process X and provide some examples where the moments are analytic. Lemma 3.5 Suppose the law of X, is absolutely continuous with respect to the Lebesgue measure on [0,1] for allt E [0,T] and let {((bi, A;),i 6 N} be the eigenpairs corresponding to the differential operator L‘. Moreover, assume p(t, y) the density of X, has the expansion Egon-(Md where i!p.~(y) = Who, and the initial density p(0,y) lies in the span of finitely many eigenfunctions of L‘. Then the following are true 1. The formal series above for p(t, y) converges everywhere. 2. The [C - th moment of the process X is an analytic function on [0,T]. £1991; Consider the Fokker-Planck equation for the density p(t, y), ap(t’ y) _ t at '— L p(t, y), where p(O, y) is given. Then from the expansion for p(t, y) and the Fokker-Planck equation we must have the following relation i ... (i) Pi(y) = (zyllmo = (L) if(0,y),i = 0,1,... Since by our assumption p(O, y) = 2:520 AJ-qSJ-(y), then for all i = 0,1, 2, . . ., we get = ((L'):§;oAj¢j 0. i=0 i=0 Hence the density p(t, y) is an analytic function of the variable t. Now we shall prove part 2) of the Lemma. Let k E N, then Mkt()=/p(ty)y"dy=/219431)? "=Z/oy"dy, i=0 by the Legesgue dominated convergence Theorem. Moreover, I [01p1(y)y"dyli _<_ {/01 lpe(y)ldy}i, which implies JlggfiupU/y My )yikdz/I} < 11m sup{/ Ips(y)|dy} < {/0 (RWY-ids} =R+€ for all 1 large, where R 18 the radius of convergence for the series corresponding to the density p(t, y) and e is any arbitrary positive number. Thus by letting e —+ 00, we can conclude that Mk(t) is analytic. This completes the proof. Examples of Models which possess Analytic Moments A few examples where all the moments of the process X involved in the model are analytic functions are as follows: 37 Example 1: a(a:) E 0 and b(x) = ‘/z(1 — :c),Vx 6 [0,1]. It is easy to see that system 3.9 for the moments yields Mi“) = 0,-[M.-_1(t)—M,-(t)], M;(O) = given. This system can easily be solved explicitly. The solution has the form t Mm) = {[M,(0) — M._1(0)] + 9,. [0 141-1(3) exp(a,,s)} exp(—6kt). The solution clearly shows that all the moments are analytic in t. Example 2: It is equally easy to show that the moments of the process involved in model 4.1 are all analytic functions in t. An explicit solution for the moments system of equations is available. Chapter 4 Some Models From Genetics In this Chapter we present the various genetic models for which numerical compu- tations have been carried out. Several numerical algorithms for the Gauss-Galerkin method and the power series method are presented. For each model considered, we solve the truncated system of moment equations (system 3.9, Chapter 3), as suggested by the power series method. This is followed by a detailed analysis of the steady state of the process through the power series method. We then present a detailed comparison between the Gauss-Galerkin method and the power series method. We end Chapter 4 with a discussion concerning the numerical results and future work. These numerical results can be found in the Appendix. In the following section we present a brief description of the various genetic models under consideration. For a more detailed description of how these models are derived we refer to Ewens, Karlin and Taylor, and Kimura and Crow. 4.1 Description of the Various Models The following models where b(x) = Ax”(1 - x)? are considered: 1. ModellzaEOandp=q=A=L 2. Model2: aEOandp=%,q=%,/\ l 3. Model3: a50andp=1,q=%,/\ 1 38 39 4. Model 4: a(:1:)= sa:(1 — :r)[h + (1 — 2h):c] — 112: + v(1— 55),]; = q = (,\)2 =% where s, h, u, and v are parameters. We shall consider two special cases of Model 4. Model 4.1 corresponds to the case where s = 0, h = 0.5, u = 0.5, and v = 0.5. Model 4.2 corresponds to the case where s = 4.8, h = 1.3, u = 0, and v = 0. Our numerical results concerning these two models will be compared to the numerical results obtained by J. Keller and R. Voronka. The four models above are important stochastic processes which describe the change of gene frequency in a Mendelian population. These models are of interest in pop- ulation genetics. The models describe the behavior of the gene frequency of a given population, as it fluctuates in time. Consider a pair of alleles A and a with respective frequencies 3 and 1 — 2:. Then Model 1 describes the case of random fluctuation of selection intensities. Model 2 describes the simplest situation. It describes the case which is known in population genetics as the process of random genetic drift due to random sampling of gametes. Model 3 is a model of mathe- matical importance. In this model, the moments system of equations is not closed. Model 4 includes the effects of selection, mutation and migration as well as random drift due to the finite size of the population (note that through a time change, the population size can be considered to be equal to 1). In this model the parameter 3 denotes the selection advantage of the zygote AA over the zygote Aa. The param- eter h is such that the quantity sh denotes the advantage of the zygote Aa over the zygote aa. Moreover, the parameter u represents the effective mutation rate of A to a and the parameter v represents the effective mutation rate of a to A (the effective mutation rates include the effects of both mutation and migration). In our considerations, Model 4.1 (which corresponds to s = 0, h = 0.5, v = 0.5, and v = 0.5), and Model 4.2 (which corresponds to s = 4.8, h = 1.3, u = 0, and v = 0) are well known models in population genetics. For more details see Ewens, Kimura and Crow, and J. Keller and R. Voronka. We note that all these models satisfy our assumptions. In the next sections of this Chapter, we shall describe the numerical results of the theory presented in Chapter 3 as it relates to the four models above. 40 4.2 Solution of the System of Moments In this section we illustrate the various numerical methods we considered earlier. We shall apply these numerical methods to Models 1, 2, 3, and 4.1 and 4.2. This is followed by a discussion of the relative merits of these methods as suggested by the numerical results. The numerical methods considered are the moment neglect method, the Gauss-Galerkin method, and the power series method. 4.2.1 Moment Neglect Method The moments system of equations for the four models are as follows Modem; M1,:(t) = ohm/11:0)’2Mk+1(t)+Mk+2(t)l, (4-1) Mk(0) = given. (4.2) M93121}; ML“) = 9kle-1(t) — Mk(t)]. (4-3) Mk(0) = given. (4.4) Model 3: Ml’c(t) = 9kle(t)‘Mk+1(t)l, (4-5) M1,,(0) = given. (4.6) MQQM M},(t) = k{(% + v)M,,_1(t) + (£0: — 1) — v — u + sh)Mk(t) + (4.7) 8(1 - 3h)Mk+1(t) - 3(1 - 2h)Mk+2(-t), (4.8) Mk(0) = given. (4.9) As an illustration, we apply the moment neglect method (for more details, see Chapter 3) to Model 3. The resulting system of equations has the following form m5“) = 92lm2(t)-ma(t)l, 41 m3“) = 93[m3(t) - 77140)], mix—1U) = on-llmn-l(t) - mn(t)la mn(t) E 0, this system can be written in matrix form as follows m'(n, t) = Nnm(n, t), m(n,0) = given, where (02—02 0 0... 0) 0 0 —0 0 0 N": o o3 3 o o o \0 o woos.-.) and ( m2(t) ) t m(n,t): "'3,” , l "in—10) / the unique solution is given by m(n,t) = m(n,0) exp(tN,,). When the initial distribution of the process X is uniform on (0, 1), the initial con- dition is given by #IH wlh‘ m(n, 0) = If 3'... 00. \ For n=20, the approximations for M2,M3,...,M8, are summarized in the Ap- pendix. Since the true moments are bounded by 1, it can be seen from the graphs 42 in the Appendix that the moment neglect method leads to misleading results even for small values of t. When applied to models 1, 2 and 4 the moment neglect method also gives misleading results. Higher values of 11 gave similar results. Based on this observation, we do not recommend the use of the moment neglect method or any of its variants. 4.2.2 Gauss-Galerkin Method With the initial density being uniform, and with n=5 we compute the initial nodes $5,1(0), 353(0), . . . , 355(0) and weights w5,1(0), w5,2(0), . . . , w5,5(0) through the tridiagonal matrix method described in Chapter 3. With these initial values we solve the moment problem using our improved algorithm which is described in Chapter 3. The results are summarized in Tables g.2, g.3, g.4.1, and g.4.2, and Figures 1.1, 1.2, 1.3, and 1.4. Our computation results show that the power series method and the Gauss-Galerkin method are in close agreement. For models 4.1 and 4.2, both methods support the results obtained by J. Keller and R. Voronka. However, the method used by J. Keller and R. Voronka is valid only for small values of t. The Gauss-Galerkin method and the power series method have the advantage of making the steady state analysis possible. 4.2.3 Power Series Method In this section we describe the numerical results of the power series method. The actual numerical results can be found in the Appendix. We have proved the fol- lowing result (see Theorem 3.1 of Chapter 3), VT > 0, and We 6 N, lim sup IMk(t) — mk(n,t)| = 0. "“°° 05:57 Therefore, given any k E N we can approximate (uniformly) the moment M], by the polynomial mk(n, .) on [0,T], for large values of n. We note that when n gets larger, more and more initial conditions are needed. Since mk(n, .) is a polynomial in t, no matter how large n is , we have mk(n,t) -» 00 as t -> 00. Therefore to analyze the steady state solution, a single polynomial of degree n (no matter 43 how large n is) will not provide useful information when t -> 00. To illustrate the approximation resulting from the power series method for finite values of t, we choose the initial distribution to be uniform and fix n = 7. Using the power series method described earlier, the following results are derived for model 3 (see Appendix for results concerning all three models), 7 m2(7,t) = 20¢... i=0 the coefficients c,~, for i = 0,1,..., 7 are found to be — — 1 — .1. — l. CO - 313361 "' '13-’62 "' _30903 - 729 _ _a - _2§_ _ __1_9__ _ .2231; c4 - 315,65 " 7200’C6 — 10800’67 — 2116800' The rest of the functions m3(7,t), m4(7,t), . . ., m9(7,t) are easily obtained from the function m2(7,t). They are tabulated in the Appendix. This results are very close to the results provided by the Gauss-Galerkin method as shown below. 44 4.2.4 Comparison of the Power Series Method and the Gauss-Galerkin Method In this section we give a detailed comparison between the power series method and the Gauss-Galerkin method. We present advantages and disadvantages of both methods from a theoretical as well as a numerical point of view. Advantages of the Gauss-Galerkin method 1. When the truncated system of equations for the nodes and the weights admits a solution on a t interval of the form [0,T], where T 6 (0,00), then the approximate solution converges to the law of X, in distribution. 2. The approximation in part 1 allows any functional of the process X to be approximated. 3. Using the numerical algorithm we proposed in Chapter 4, the Gauss-Galerkin method proved to be numerically efficient. Disadvantages of the Gauss-Galerkin method 1. The Gauss-Galerkin method fails to work when the initial distribution is degenerate. 2. The approximate solution for the process moments has no known rate of convergence. 3. The existence of a unique solution for the truncated system of equations for the nodes and weights (equivalently for the truncated system of moments when the drift and the square of the diffusion are polynomials) is only true on a t interval of the form [0,T(n)] where T(.) is a function of n. There is no available theoretical proof for a global solution, i.e. T(n) E T, Vn E N. However, for some models, this can be established numerically. Advantages of the power series method 45 1. The truncated system of equations for the moments has a unique solution on any fixed t interval. 2. The approximate solution converges to moments of X uniformly on any in- terval [0,T], for all T 6 (0, oo). 3. The expression of the approximate solution can be explicitly found as a func- tion of t, for all t 6 (0,00). Disadvantages of the power series method 1. It is not clear that the truncated system of equations for the moments yields functions which are moments of a probability measure. This does not make approximations of functionals of the process possible. 2. As in the Gauss-Galerkin method, there are no available rates of convergence to the steady state solution of the stochastic differential equations. 4.3 Steady State Analysis In this section we study the behavior of the process X as t —» 00. According to Theorem 3.1 (Chapter 3), we have VT > 0, and We 6 N, lim sup IMk(t) - mk(n,t)| = 0. ""°° 05:57" For any given k E N we can uniformly approximate the k-th moment by the sequence {mk(n, .), n 2 1} on [0,T]. When we let t increase to infinity, we have mk(n,t) —r 00. Therefore to analyze the steady state solution, a single polynomial of degree n (no matter how large n is) will not provide useful information when t —; 00. To illustrate this we choose the initial distribution to be uniform and fix n = 7. Using the power series method described earlier, the following results are derived for Model 3, 7 m2(7, t) = 2 c,-t', i=0 46 the coefficients c,, for i = 0, 1, ..., 7 are as above The rest of the functions m3(7, t), m4(7, t), ..., m9(7, t) are easily obtained from the function m2(7, t). Moreover, it is clear that 111.12, m2(7,t) = 00. Based on the power series method, the following computational algorithm yields useful information about the steady state solution of the process X. The results are summarized in the appendix. We now give an outline of this algorithm designed for the study of the steady state solution. Computational Algorithm 1. Given any T 6 (0,00), let h = %, where N E N. 2. Divide the interval [0, T] into N sub-intervals of length h, to get the partition [O,h1)[h1,h2), . . ., [hN-1,hN], where h,- =1h, 0 S13 N — l. 3. Fix a low degree p (say p = 1 or p = 2). 4. Define the following functions - p (i) m(k, 1,t) = z cj (k)(t — h,-_1), i=0 where t E [h,-_1,h,-) and 1 S i _<_ N — 1, and m(k,i,t) will represent the piecewise linear (if p = 1) or quadratic (if p = 2) approximation of the k-th moment in the i-th sub-interval. For p = l, the functions {m(k, i,t),0 5 i g N — 1} are computed as follows. 5. For the sub—interval [0, h), we have m(2.1,t) = of.” + c‘.”t. where c9) = M2(0) and c9) = Mi’,(0). This yields an approximation for M2(t), t E [0, h). Note that we used two initial conditions. In general, to calculate m(k, l, t), for each k, 2 initial conditions are needed. Computing 0 functions will require oz+1 initial conditions. 47 6. For the sub-interval [h, 2h), we have m(2, 2,.) = ct” + 4% — h), where 682) = m(2,1,h) and C(12) = 02(m(2,1,h) —- m(3, 1, h)). Note that com- puting more and more functions will involve more and more initial conditions. As we march on from sub-interval to sub-interval, more initial conditions are needed. 7. To study the steady state solution for M2(t) for example, we observe that (lim CI = 0, this yields the steady state value (for M2) lim c0. l—boo A similar behavior is observed for higher moments. Recall that we considered the operator L over the interval [0,1] with its domain containing all the monomials. This assumption holds for all three models. The boundary 0 and 1 may be classified as ”regular”, ”exit”, ”entrance” or ” natural” (see Appendix), depending on the behavior of the functions a and b involved in the stochastic differential equation. Using the boundary classification, we can easily show the following: In Model 1, both 0 and l are ”natural” boundaries. In Model 2, both 0 and 1 are ”exit” boundaries. In model 3, 0 is a ”natural” boundary and 1 is an ”exit” boundary. In model 4.1, both 0 and l are ”entrance” boundaries. In model 4.2, both 0 and 1 are ”exit” boundaries. According to Ethier and Kurtz, the appropriate domain of the operator L is given by D(L) .—. {<15 6 C[0,1] 0 02(0, 1) : L(qb) e C[0,1]}, at an entrance boundary or at a natural boundary, and D(L) = {as e 010.11n 010,1) = Les) 6 C[O. 11.12:; Lab) = 0}. 48 at an exit boundary r. Moreover, all of our assumptions (assumptions A1 through A5) are satisfied by these models. In Tables 1 through 3, numerical results for the moments mo, m1, . . . , m5 are pre- sented for several t values until the steady state solution is reached. Table 1 is for model 1, Table 2 is for model 2, and Table 3 is for model 3. The initial distribu- tion considered is the uniform on the interval (0,1). Since the function a vanishes identically on the state space [0, 1] in all three models, it follows using system (3.9) that both mo and m, are constants in t, and according to the initial distribution (uniform) their values are 1 and 0.5, respectively. In Model 1, both boundaries are natural and the law of X, is absolutely contin- uous with respect to the Lebesgue measure on [0,1]. In this model the density p(oo, y) E 0, Vy 6 (0,1), is clearly a steady state solution with the density escaping to the boundary as t becomes arbitrary large. Our computations show that the steady state solution is reached when t is approximately 50. The results in Table 1 show that the steady state values for the moments are approximately equal to 0.5. This implies that the steady state distribution measure for the process X is the measure $760 + 61). This is in agreement with the exact solution of the model (see Appendix). Table 1.1 describes the approximate values of the first few moments through the power series method until steady state is reached. In Model 2, both boundaries are exit and the law of X, is not absolutely con- tinuous with respect to the Lebesgue measure on [0,1]. However, the absolutely continuous part of the law of X, has support (0, 1). The singular part of the law of X, has a point mass at 0 and a point mass at 1. It is important to note that the Radon-Nikodym derivative p(t, y) of the absolutely continuous part satisfies the Fokker-Planck equation. In this model, for each y 6 (0,1), the function p(t, y) —i 0, as t —» 00. The steady state solution for this model can be deduced from the be- havior of the moments as t -> 00. Our computations show that the steady state solution is reached when t is approximately 12. The results in Table 2 show that the steady state values for the moments are approximately equal to 0.5. This im- 49 plies that the steady state distribution measure for the process X is the measure -;-(60 + 61). This is in agreement with the exact solution. Table 2.1 describes the approximate values of the first few moments through the power series method until steady state is reached. In Model 3, 0 is a natural boundary and 1 is an exit boundary. In this model the law of X, is again not absolutely continuous with respect to the Lebesgue measure on [0,1]. The singular part of the law of X, has a point mass at 1. It is important to note that the Radon-Nikodym derivative p(t, y) of the absolutely continuous part satisfies the Fokker-Planck equation. In this model the function p(t,y) -» 0, as t -+ 00. The steady state solution for this model can be deduced from the behavior of the moments as t -> 00. The results in Table 3 show that the steady state values for the moments are approximately equal to 0.5. This suggests that the steady state distribution measure for the process X is the measure lz-(bo + 61). Our compu- tations show that the steady state solution is reached when t is approximately 20. The steady state analysis for this model cannot be checked against an exact solu- tion since no exact solution is known. Table 3.1 describes the approximate values of the first few moments through the power series method until steady state is reached. In Model 4.1, both 0 and 1 are entrance boundaries. In this model the law of X, is absolutely continuous with respect to the Lebesgue measure on [0,1]. The Radon-Nikodym derivative p(t, y) of the law of X, satisfies the Fokker-Planck equa- tion. The steady state solution for this model can be found by solving the equa- tion L*p(oo, y) = 0. Using both the Gauss-Galerkin method and the power series method, we can recover the steady state moments. By applying the power series method to this model we observe that the steady state solution is reached at ap- proximately t = 6. By applying the Gauss-Galerkin method, we concluded that the steady state solution is achieved at approximately t = 5. The exact steady state moments are very close to both the approximate steady state moments obtained through the power series method and to the approximate steady state moments ob- tained through the Gauss-Galerkin method. Table 4.1 describes the approximate 50 values of the first few moments through the power series method until steady state is reached. Table 4.1.1 is a summary of the true steady state moments and Table 4.1.2 shows the approximate steady state moments. In Model 4.2, both 0 and l are exit boundaries. In this model the law of X, is not absolutely continuous with respect to the Lebesgue measure on [0,1]. The singular part of the law of X, has a point mass at 0 and 1. It is important to note that the Radon-Nikodym derivative p(t, y) of the absolutely continuous part satisfies the Fokker-Planck equation. In this model the function p(t,y) —> 0, as t —> 00. The results in Table 3 show the approximate values of the moments for several values of t. The exact solution is not known. Therefore we did not compare our results with the values of the true moments. Table 4.2.1 is a summary of the values of the first few moments obtained through the power series method. Table g.4.2 shows the approximate values of the first few moments through the power series method. The analysis of the steady state behavior of this process is work in progress and will be included in a later publication. 4.4 Discussion This section provides concluding remarks and an outline of some suggestions for future work. One open problem is the approximation of joint distributions of the process X. A more difficult task is the approximation of the transition probabil- ity function. An examination of the conditional moments may lead to some useful ideas for the approximation of the transition density function. We also outline some problems which are still unresolved for both the Gauss-Galerkin method and the power series method. An important task is the development of a numerical method which provides an approximation of the transition probability function and there- fore leading to an approximation of the joint distribution of the diffusion (since any joint distribution of the process X is in terms of the transition probability function). We propose an idea that might lead to some new development in this direction. We 51 note that neither the power series method nor the Gauss-Galerkin method offers this kind of an approximation. An outline of our idea is as follows. The stochastic differential equation 3.1 has a unique solution X, which is a time homogeneous Markov process. Moreover, assume the density p(t, 2:, y) of the tran- sition probability function satisfies the following initial boundary value problem Eat—5,5439- = L.(p). (4.10) p(0,w.y) = 6(x—y), (4.11) p(th) = ¢(t.y). (4.12) p(tmy) = 1120.21). (4.13) where the state space of the process X is one of the following (l,r), [l,r], (l,r], or [1, r). The functions a) and 1b are given. Because of the lack of smoothness of the initial function p(0, r, y), the above initial boundary value problem may be an ill-posed problem and hence difficult to handle. A somewhat easier problem for the conditional moments of the process X can be derived as follows. The i - th conditional moment of the process is defined by 0.0.30) = [,y‘p(t,x,y)dy,w 6 W U 0. (assuming the integrals exist). In Equation 4.10 we multiply both sides of the equation by y‘, then integrate with respect to the variable y. Then the following initial boundary problem can be derived 0C5“, 27) T L,(C,-(t, r)), (4.14) C,‘(0,£L‘) = (1}., (4.15) 52 C,(t, I) [1' W. y)y‘dy. (4-16) C.(t. r) = [Ir W. y)y‘dy- (4-17) Note that this system does not suffer the lack of smoothness of the initial condi- tion. Moreover, the system for the conditional moments may be easier to solve. Moreover, the process moments may be obtained by integration of the conditional moments. A much more important task is the approximation of the distribution of X, be a sequence of discrete probability measures for each value of t. The knowledge of the first it conditional moments may lead to the construction of probability measures which we conjecture will converge to the transition probability measure. More work along the above ideas is in progress. Another problem which more complicated is the two dimensional problem. In some areas of engineering sciences, stochastic differential equations where the solution is a two dimensional stochastic process present some challenging questions. One of the most difficult problems is the approximation of the two dimensional distributions involved. More work needs to be done in this direction. APPENDIX 53 APPENDIX In this appendix we present the following items. Section A.1 contains a summary of the figures mentioned in Chapter 4. Section A.2 contains a summary of the tables mentioned in Chapter 4. Section A.3 contains a discussion on boundary classifications. A.1: Figures 1 and 2 are results of an application of the moment neglect method to Model 3. Figure 1 shows a plot of the initial moments M2(0), M3(0), . . . , M3(0) against the order of the moments. Figure 2 shows a plot of the approximate mo- ments (according to the moment neglect method) against the order of the moments, for t = 0.001. It can be seen from Figure 2 that the results are misleading since the approximate value of M6(0.001) clearly exceeds 1. This contradicts the fact that the state space of the diffusion under consideration is [0, 1]. Figures 1.1, 1.2, 1.3, 1.4, 1.5, 1.6, and 1.7 represent the exact transition probability density for model 1 plotted against the forward variable y, for several values of t until the exact steady state is reached. Figures 1.2 through 1.4 show that for t between 0 and 10 the density is unimodal. For t = 20, approximately, the density is shown in Figure 1.5 to be bimodal. The density eventually accumulates on the boundaries 0 and 1 at approximately t = 50. Figure 1.7 shows that both accumulations at 0 and 1 are approximately equal to 0.5. This is in agreement with both the Gauss-Galerkin method and the power series method. Figures 2.1, 2.2, 2.3, 2.4, and 2.5 represent the exact Radon-Nikodym derivative of the absolutely continuous part of the tran- sition probability measure with respect to the Lebesgue measure. This function is plotted against the forward variable y, for several values of t until the steady state is reached. Figure 2.5 shows for t = 12, approximately, the steady state value of the Radon-Nikodym derivative of the transition completely leaks out through both boundaries of the state space [0,1]. This ”numerical leakage” is due to the fact that both boundaries 0 and 1 are exit boundaries. A.2: Table g.2 represents the approximate moments for Model 2 according to the Gauss-Galerkin method. In Table g.2, m1,m2, . . . ,m5 represent the approximate 54 values of the exact moments, M1,M2, . . . ,M5, of the process. The approximate steady state moments stabilize around the value 0.5 suggesting that the steady state density has a point mass at 0 and a point mass at 1, both equal to 0.5. Table g.3 represents the approximate moments for Model 3 according to the Gauss-Galerkin method. A similar behavior to the one in Table g.2 is observed. Since the exact moments are not known for this model, we cannot make quantitative comments on the accuracy of the results. However, the power series method is in agreement with these results (see Table 3.1). Table g.4.1, g.4.2 represent the approximate moments for Model 4.1 and Model 4.2, respectively, according to the Gauss-Galerkin method. Both tables are in agreement with the corresponding results obtained by using the power series method (see Table 4.1 and 4.2.1). Table 1.1, 2.1, 3.1, 4.1, and 4.2.1 represent the power series approximate moments for Models 1, 2, 3, 4.1, and 4.2, respectively. In these tables, the values of the functions m1, m2, . . . , m5 are given for several values of t. The functions m1,m2, . . . ,m5 represent the approximate values of the exact moments, M1, M2, . . . , M5, of the process under consideration. From Table 1.1 it can be seen that the steady state solution for Model 1 is reached for t = 55, approximately. We also note that the approximate values of the mo- ments stabilize around 0.5 suggesting that the steady state density has has a point mass at 0 and a point mass at 1, both equal to 0.5. This is in agreement with the Gauss-Galerkin method. A similar behavior can be observed through Tables 2.1, and 3.1 for models 2, and 3, respectively. Table 4.1 shows that the approximate steady state solution is reached at t = 20, approximately. This agrees with the exact steady state solution. Tables 4.1.1, and 4.1.2 show this agreement. Table 4.2.1 is for Model 4.2. It shows the power series approximate values for several values of t. For this model we do not have a steady state analysis at this point. We have compared our results (both through the Gauss-Galerkin method and the power series method) with the results obtained by J. Keller and R. Voronka and we observed a close agreement. 55 A.3. Boundary Classification Let X be a diffusion process with state space [0,1]. According to Feller [15], the boundaries 0 and 1 can be classified as follows. 1. Regular. 2. Exit. 3. Entrance. 4. Natural (or Feller). We shall present the formal definitions of the different types of boundaries. These definitions can be found in Karlin and Taylor, and Feller. The scale function is defined by 3 0(4) = j, q(y)dy, 0 where (M) = epr- A: 20(n)b‘2(n)dnl.Vy 6 (0.1). The numbers 1:0 and yo are arbitrary, but fixed. Their particular choice is of no relevance to the analysis below. The scale measure is defined by u(d:r) = q(a:)da:. The speed measure is defined by duh?) = m($)dx, where m(r) = W' Using the scale and the speed measures, the following important quantities are defined. For the left boundary point 0, we have 2(0) = [4((4,4))q(e)44, and 9(0) = f,” 441:, 41)m(c)44. The following summarizes the boundary classifications. 56 0 is regular if 2(0) < 00 and 9(0) < oo. 0 is an exit if 2(0) < 00 and 9(0) = 00. 0 is an entrance if 2(0) = 00 and 9(0) < oo. 0 is natural if 2(0) = 00 and 9(0) < 00. The boundary behavior at 1 is defined similarly. 57 Moment Neglect Method (Model 3) Figure 1: Moment Neglect, Time=0 10 Figure 2: Moment Neglect, Time=0.001 N U10U'1U'IU'I -7.S' A A A A A A - Model 1 (exact solution) 58 Figure 1.1, Model 1 O O N (11 O O ..s U1 P(0.1.0.S.y) 0.005 O O N 0.01) D ># M v 1 L 1 A A 0 0.2 :4 o: 6 0 Figure 1.2, Model 1 .8 1 ONIhO‘mt-‘Nlb f 59 Figure 1.3. Model 1 P(3.0 5.y) c: c: c: c: c: c: OHlvlaJcbuim Figure 1.4, Model 1 T v f 1 0.145 0.12» .y) 0 H -o.08» ‘0.06' P(10 0 S O O A 0.02- Figure 1.5, Model 1 v r r v V P(20,0.5.Yl <3 :1 <3 c: <3 w 60 Figure 1.6. Model 1 N ...: OU'Il-‘UINUIW P(35,0.5.yl O V V v Or 0:2 0:4 0:6 0: Figure 1.7, Model 1 h) lb 0 O N O P(S0.0.S.y) ..s O a v M Y 0:2 0:4 0:6 ole 61 Model 2 (exact solution) Figure 2.1, Model 2, Exact Solution f i .S.y) O P(l, Figure 2.2, Model 2, Exact Solution 17 T f 1 0.2) 0.15» .5,y) 0.1) P(4, 0.05' 62 Figure 2.3, Model 2, Exact Solution 0.005) r v v v *v Figure 2.4, Model 2, Exact Solution 0.01v 0.008) 50.006» 30.004. 0.002) ——v v w 0:2 0:4 0:6 0:8 1 Figure 2.5, Model 2, Exact Solution 0.4) 0.2) .5,y) P(12. 0:2 0:4 0:6 ole i 63 Gauss Galerkin Method, Model 2 Approx. Moments, Model 2, Table g.2 t m1 m2 m3 m4 m5 0.000000 2.000000 8.000000 30.00000 0.500000 0.497900 0.497890 0.497690 0.333333 0.475130 0.497830 0.497680 0.250000 0.467500 0.497800 0.497680 0.250000 0.456920 0.497780 0.497680 0.166667 0.452350 0.497770 0.497680 Gauss Galerkin Method, Model 3 Approx. Moments, Model 3, Table g.3 t m1 m2 m3 m4 m5 0.000000 2.000000 8.000000 30.00000 0.500000 0.501030 0.501000. 0.500840 0.333333 0.427030 0.484760 0.498700 0.250000 0.401640 0.481690 0.498600 0.250000 0.389110 0.480380 0.498560 0.166667 0.381730 0.479660 0.498540 64 Gauss Galerkin Method, Model 4.1 Approx. Moments, Model 4.1, Table g.4.1 t m, m2 m3 m4 m5 0.000000 2.000000 4.000000 6.000000 10.00000 20.00000 0.500000 0.501100 0.490003 0.498879 0.498888 0.502010 0.333333 0.290225 0.299002 0.301100 0.301230 0.300400 0.250000 0.200337 0.200022 0.202300 0.199050 0.203201 0.167000 0.143231 0.142860 0.142863 0.142863 0.142863 0.200000 0.107517 0.107139 0.107300 0.107566 0.107102 Gauss Galerkin Method, Model 4.2 Approx. Moments, Model 4.2, Table g.4.2 t m1 m2 m3 m4 m5 0.000000 1 .000000 2.000000 3.000000 4.000000 5.000000 6.000000 0.500000 0.520564 0.599000 0.616667 0.645500 0.670098 0.703976 0.333333 0.377767 0.412280 0.442779 0.478420 0.511572 0.541934 0.250000 0.279300 0.319802 0.344444 0.379263 0.412804 0.444842 0.20000 0.225088 0.251222 0.284367 0.314702 0.348146 0.378972 0.166667 0.189048 0.214008 0.241379 0.270711 0.305273 0.335410 65 Power Series Method, Model 1 Approx. Moments, Model 1, Table 1.1 t m1 m2 m3 m4 m5 0.000000 0.100000 0.200000 0.300000 0.400000 0.500000 0.600000 0.700000 0.800000 0.900000 1.000000 5.000000 12.00000 35.00000 55.00000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.333333 0.336667 0.339905 0.343052 0.346114 0.349093 0.351994 0.354819 0.357573 0.360257 0.362875 0.394726 0.474390 0.492492 0.495014 0.250000 0.255000 0.259857 0.264579 0.269171 0.273640 0.277991 0.282229 0.286359 0.290386 0.294313 0.353642 0.456862 0.489186 0.494468 0.250000 0.205714 0.21 1286 0.216719 0.222021 0.227194 0.232243 0.237174 0.241989 0.246693 0.251289 0.325734 0.443768 0.486220 0.492990 0.166667 0.172619 0.178452 0.184168 0.189767 0.195251 0.200623 0.205886 0.211040 0.216089 0.221035 0.304407 0.432859 0.483435 0.491564 66 Power Series Method, Model 2 Approx. Moments, Model 2, Table 2.1 t m1 m2 m3 m4 m5 0.000000 0.100000 0.200000 0.300000 0.400000 0.500000 0.600000 0.700000 0.800000 0.900000 1.000000 2.000000 3.000000 4.000000 5.000000 6.000000 7.000000 8.000000 10.00000 12.00000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0,500000 0.500000 5.000000 5.000000 5.000000 5.00000 5.000000 5.000000 5.000000 0.333333 0.349194 0.363545 0.376530 0.388280 0.398912 0.408531 0.417236 0.425112 0.432238 0.438687 0.477444 0.491702 0.496947 0.498877 0.499587 0.4998480 0.499944 0.499973 0.499838 0.250000 0.273791 0.295317 0.314795 0.332420 0.348367 0.362797 0.375854 0.387668 0.398358 0.408030 0.466166 0.487553 0.495421 0.498316 0.499380 0.499772 0.499916 0.499960 0.499757 0.200000 0.228549 0.254381 0.277755 0.298904 0.318041 0.335357 0.351024 0.365201 0.378029 0.389636 0.459399 0.485064 0.494505 0.497979 0.499256 0.499726 0.499899 0.499952 0.499709 0.166667 0.198388 0.227090 0.253061 0.276560 0.297823 0.317063 0.334472 0.350224 0.364477 0.377374 0.454888 0.483404 0.493895 0.497754 0.499174 0.499696 0.499888 0.499947 0.499677 67 Power Series Method, Model 3 Approximate Moments, Model 3, Table 3.1 t m1 m2 m3 m4 m5 0.000000 0.100000 0.200000 0.300000 0.400000 0.500000 0.600000 0.700000 0.800000 0.900000 1 .000000 3.000000 12.00000 20.00000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.333333 0.341667 0.349333 0.356417 0.362985 0.369094 0.374793 0.380122 0.380122 0.389807 0.394220 0.420367 0.482793 0.499200 0.250000 0.265000 0.278500 0.290736 0.301889 0.312106 0.321503 0.330179 0.330179 0.345678 0.352630 0.404972 0.480233 0.498951 0.167000 0.220000 0.237714 0.253557 0.267834 0.280781 0.292584 0.303394 0.303394 0.322506 0.330997 0.382445 0.480088 0.498515 0.200000 0.190476 0.211310, 0.229762 0.246256 0.261109 0.274568 0.286828 0.286828 0.308355 0.317859 0.384063 0.479000 0.498095 68 Power Series Method, Model 4.1 Approx. Moments, Model 4.1, Table 4.1, u=v=h=0.5, and s=0 t m1 m2 m3 m4 m5 0.000000 0.100000 0.300000 0.500000 0.700000 0.900000 1.100000 1.300000 1.700000 1.900000 2.000000 4.000000 6.000000 8.000000 10.00000 12.00000 14.00000 16.00000 18.00000 20.00000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.500000 0.333333 0.325960 0.315746 0.309550 0.305792 0.303513 0.302131 0.301292 0.300475 0.300288 0.300225 0.300002 0.300000 0.300000 0.300000 0.300000 0.300000 0.300000 0.300000 0.300000 0.250000 0.238940 0.223618 0.214325 0.208689 0.205270 0.203196 0.201939 0.200713 0.200433 0.200337 0.200002 0.200000 0.200000 0.200000 0.200000 0.200000 0.200000 0.200000 0.200000 0.167000 0.186912 0.169294 0.158822 0.152523 0.148716 0.146409 0.145011 0.143650 0.143338 0.143231 0.142860 0.142863 0.142863 0.142863 0.142863 0.142863 0.142863 0.142863 0.142863 0.200000 0.152380 0.133871 0.123180 0.116827 0.113006 0.110696 0.109297 0.107935 0.107624 0.107517 0.107145 0.107142 0.107142 0.107142 0.107142 0.107142 0.107142 0.107142 0.107142 69 Steady State, Model 4.1 Steady State, Model 4.1, Figure 4.1 6y(1-y) O O O O o 0 O O o O O N ab 0‘ (D H N 1b #1 v7 True Steady State Moments, Model 4.1, Table 4.1.1 m, 7H2 1123 m4 m5 0.500000 0.300000 0.200000 0.142857 0.107143 Approx. Steady State, Model 4.1, Table 4.1.2, t=20 772, m2 m3 m4 m5 0.500000 0.300000 0.200000 0.142863 0.107142 70 Power Series Method, Model 4.2 Approximate Moments, Model 4.2, Table 4.2.1 t m, m2 m3 m4 .m5 0.000000 1.000000 2.000000 3.000000 4.000000 5.000000 6.000000 0.500000 0.540000 0.578000 0.612767 0.643648 0.670598 0.693976 0.333333 0.368867 0.405886 0.442871 0.478420 0.511572 0.541934 0.250000 0.279300 0.311312 0.345041 0.379263 0.412804 0.444842 0.20000 0.225057 0.252947 0.283167 0.314902 0.347146 0.378972 0.166667 0.189048 0.214008 0.241379 0.270711 0.301273 0.332219 BIBLIOGRAPHY 71 BIBLIOGRAPHY Arnold L. (1974). Stochastic Differential Equations, Theory and Applications. Wi- ley Interscience. Dawson, D. A. (1980). Galerkin Approximation of Nonlinear Markov Processes, Statistics and Related Topics. Proceedings of the International Symposium on Statistics and Related Topics held in Ottawa, Canada, May 5—7, 1980. North Hol- land. Ethier, Stewart N. and Kurtz, Thomas G., (1986). Markov Processes, Characteri- zation and Convergence. John Wiley and Sons, New York. Ewens, Warren J. (1979). Mathematical Population Genetics. Biomathematics, Volume 9, Springer-Verlag, New York. Feller, W. (1952). The Parabolic Differential Equation and the Associated Semi- Group of Transformation, Ann. Math., Vol. 55, pp. 468-519. HajJafar, A. (1984). On the Convergence of the Gauss Galerkin Method for the Density of Some Markov Processes. Ph.D. Thesis, Department of Mathematics, Michigan State University. HajJ afar, A., Salehi, H., and Yen, D. (1992). On a Class of Gauss-Galerkin Methods for Markov Processes, Proceedings of the Work Shop on Nonstationary Stochastic Processes and Their Applications, Edited by Miamee, A. G., Hampton, Virginia, August 1-2, 1991, pp. 126—146. Karlin, S. and Taylor, H. M. (1981). A Second Course in Stochastic Processes. Academic Press. Keller, J. and Voronka, R. (1975). Asymptotic Analysis of Stochastic Models in Population Genetics. Mathematical Biostatistics, Vol. 25, pp. 331-362. Kimura, M. and Crow, J. (1970). An Introduction to Population Genetics. Harper and Row, New York. Krylov, V. I. (1962). Approximation Calculation of Integrals. Macmillan, New York. (Translated from the first Russian edition (1959) by Stroud, A. H.) Shohat, J. A. and J. D. Tamarkin (1943). The Problem of Moments, A.M.S., New 72 York. Smirnov, V. I. (1964). A Course of Higher Mathematics. Pergamon Press, Oxford, New York. Stoer, J. and Bulirsch, R. (1980). Introduction to Numerical Analysis. Springer- Verlag, New York. Stroud, A. H. ( 1974). Numerical Quadrature and Solution of Ordinary Differential Equations. Springer-Verlag, New York. Szarski, J. (1965). Differential Inequalities. Warszawa. Wentzell, A. D. (1981). A Course in the Theory of Stochastic Processes. McGraw Hill, New York. 111011111111)“