630 oiocuasuc; piuutjsses 10. Find functions ft{t) and f2{t) = -fl{t) such that a standard Wiener process is between /, and f2 with probability (a) 0.5, (b) 0.95. 11. What is the probability that -Jt< W(t) < jt1 12. Prove that the transition probability density (11.12) of a Wiener process with drift satisfies the heat equation (11.13). 13. Use Theorem 6.5 to find the characteristic function of X(s) = x0 + ps + 0,£>0 find the correlation coefficient p{s,t) (see section 1.3) of W(s) and W{t). Assume s < t and s is fixed. What happens to p as t — co? 12 Diffusion processes, stochastic differential equations and applications 12.1 DIFFUSION PROCESSES AND THE KOLMOGOROV (OR FOKKER-PLANCK) EQUATIONS To introduce a wide class of random processes with properties similar to those of a Wiener process with drift, we generalize the constant drift parameter pt and variance parameter a of such a process, so that they vary with the value of the process and possibly the time. For a general process X, we have that the increment in the small time interval (t, t + At] is AX(t) = X(t + At) - X(t). Now the properties of this increment may depend on the time t and the value x of the process at the beginning of the small time interval. We therefore condition on X(t) = x and define the infinitesimal first moment, or infinitesimal mean, as a(x, t) = lim A(-0 E[AX(t)\X(t) = x] At (12.1) Note that because we have taken the expectation, this is not a random quantity. Thus {y) = exp can be employed - see the exercises. 12.5 STOCHASTIC INTEGRALS AND STOCHASTIC DIFFERENTIAL EQUATIONS We have seen in section 11.4 that a Wiener process with drift can be characterized by the stochastic differential equation dX = ndt + adW. The correct interpretation of this equation is in terms of an integral involving a Wiener process - called a stochastic integral. There are a large number of integrals which one may define in connection with random processes. Mathematical complexities arise when integrals involving W are considered because of the irregular properties of the paths of W. This means that the methods of defining integrals given in real-variable calculus courses cannot be used. We will consider stochastic integrals very briefly and somewhat superficially - there are numerous technical accounts -see for example Gihman and Skorohod (1972), Arnold (1974), Lipster and Shiryayev (1977) or Oksendal (1985). Our main purpose is to enable the reader to understand and know how to use a stochastic differential equation of the general form dX(t) =f(X(t), t)dt + g(X{t),t)dW{t). Equivalently, dropping the reference to t in the random processes, we can write this as dX=f(X,t)dt + g(X,t)dW, (12.20) where / and g are real-valued functions, W is a standard Wiener process and X is a random process which in cases of interest will be a diffusion process. However, it must be stated at the outset that (12.20) does not always have a unique interpretation. This situation arises for the following reason. Equation (12.20) is interpreted correctly as implying the stochastic integral equation X{i) = X(0) + \ 'f(X(t'), t')dt' + !'g(X{t'), t')dW(t'), (12.21) Jo Jo and the process X so defined is called a solution of the stochastic differential equation (12.20). Although the first integral here presents no problems, there are many ways of defining the second one, g(X(t'),t')dW(f), which is called a stochastic integral. Furthermore, the different definitions can lead to various solutions, X, with quite different properties. Despite this apparent ambiguity, there are two useful definitions which are most commonly employed - the Ito stochastic integral and the Stratonovich stochastic integral; and there is a simple relation between these two. A note on notation. It is preferable in (12.20) not to 'divide' throughout by dt, because as we have seen, the derivatives of W and hence of X do not exist in the usual sense. However, as long as we keep that in mind, it is possible to display (12.20) as a stochastic differential equation involving white noise w, the 'derivative' with respect to time t of W (see section 11.3): dX ~dt = f(X,t) + g(X,t)w, or perhaps even dX d7 = f(X,t) + g(X,t)W. Stochastic differential equations written in this form are often called Langevin equations. Let us now make an important observation on the stochastic differential equation (12.20). If the function g is identically zero, the differential equation is deterministic and can be written in the usual way d~=f(x,t). dt Assuming the initial value X{0) = x0 is not random then X(t) is non-random for all t and this equation is solved in the usual way. We expect that the behaviour of solutions of this deterministic equation would be related to those of the stochastic differential equation (12.20), and be close to them when the noise term g is small. We would be correct in believing that, in particular, the expected value £[X(r)] of the solution of f Figure 12.2 A schematic representation of a simple function /. In the present context the constant values on sub-intervals are actually random variables. where a = t0 < tx < • • ■ < t„_ 1 < t„ = b is a partition of the interval [a, b) and the {fj} are a set of n random variables. Now for each sub-interval [tJt tJ+J of [a, fc) we can define the corresponding increment in a standard Wiener process AWj=W(tJ+l)-W(tj), ; = 0,1.....n-1, which defines a further n random variables. We are now ready to define the Ito stochastic integral of a simple random function / with respect to a Wiener process as i j-o We note that we could also write this as f(t)dW(t) - £ /('1) - J-0 (12.23) which will be a useful observation when we distinguish between the Ito and Stratonovich integrals. In order to define the stochastic integral for the general class of random processes we have called M, we state the following lemma without proof - see the references at the beginning of this section. This lemma tells us that for any random process / in M, we can be sure there is a sequence {/„, n — 1,2,...}, also in M, whose members get closer and closer to / as n -* oo, in the following sense. Lemma Let feM be a random process in the class defined above. Then a sequence of simple functions f„(t)eM exists such that as »—> oo, r That is, this sequence of integrals, which measure the distances between /and /„, converges in probability to zero - recall the definition of this mode of convergence in section 6.6. Definition of Ito stochastic integral for arbitrary feM Since we know how to define the Ito stochastic integral for simple functions, , we extend the definition to arbitrary random processes in M by using a sequence of approximating integrals of simple functions. The limit of this sequence is defined as the required integral. Thus we set f(t)dW{t)= lim L(t)dW(t), where the limit is again in the sense of convergence in probability of a sequence of random variables. The above lemma guarantees that a suitable sequence {/«W} °f approximating random functions exists. We mentioned that other definitions can be given for the Ito stochastic integral according to the different properties which are ascribed to the integrand f(t). Usually these conditions are more restrictive (stronger) than the ones we have employed. However, then, and indeed in most cases of practical interest, we have the following results concerning the mean and the variance of the Ito stochastic integral: (i) Mean (ii) Variance f(t)dW(t)) = 0. f(t)dW(t) = f2(t)dt. (12.24) (12.25) We will not prove (12.25) but (12.24) will be considered in the exercises. Furthermore, if / and g are both in M then f(t)dW(t)x g(t)dW(t) EUmmdt. (12.25') 6iU l^lUUÖlUll JJ1UUCSBCB Thus the differential equations (12.4) and (12.5) satisfied by the transition probability density function can be obtained simply from the stochastic differential equation. In particular, for a time-homogeneous process with stochastic equation dX=f{X)dt + g(X)dW, the forward Kolmogorov equation for the transition density p(y, t\x) is, from Equation (12.6), dp dt' d(fjy)p) l d2(g2(y)p) dy 2 dy2 ' and the corresponding backward equation is, from Equation (12.7), ^=/(X)^+V(XA. dt dx 2 dx2 If a stochastic differential equation is given with Stratonovich's interpretation, it may first be converted to an Ito equation using Equation (12.30); then Theorem 12.5 will yield the infinitesimal moments of the process. Itmight seem that, in modelling a randojn_phenomenon with a stochastic differential equation, an ambiguity arisesin the choice of an Ito or Stratonovich interpretaljfln. This, however, is not the case. A diffusion processls specified once its infinitesimal mean and variance are given. Thus, when deriving a model one must be certain that these infinitesimal moments are correct. Once this is done it matters not which stochastic calculus one adopts. A word of caution is necessary. It is not a good idea to take an existing deterministic differential equation and convert one of its parameters to white noise. The reason is that in most cases there will be ambiguity because there is no a priori reason why a particular interpretation, Ito or Stratonovich, should be correct. A particular choice would only be defensible if the infinitesimal moments could be ascertained to be the correct ones. This will be illustrated in the exercises. However, in certain cases it has been shown that limits of sequences of discrete stochastic equations have solutions corresponding to the Stratonovich differential (Wong and Zakai, 1965). 12.7 APPLICATIONS Diffusion approximation to a random walk Suppose a random walk Xe — {Xt{t\ t>Q) occurs with jumps up or down of magnitude e. Jumps up form a Poisson process with intensity A and jumps down form another Poisson process N2, independent of Nlt but with the same intensity. One may assume for convenience that the processes start Applications Z'l 1 at zero at time r = 0. Note that this random walk, sometimes called a randomized random walk (Feller, 1971), differs from the simple random walk of Chapter 7 in that there the jumps (steps) occurred at fixed time intervals whereas now they occur at random times. We may write the following expression for the value of the process Xe at time t: Xt{t) = sN1(t)-eN2(t). Similarly we may write the following expression for the increment in Xc in the interval (t, t + Ar]: AXt = Xc(t + At)-Xc(t) = eiN^t + At] - NM] - eLN2(t + At) - N2(*)] = eANi - eAN2. One can also write the following stochastic differential equation involving Poisson processes: dXe = eEdjVi - dN2l We will now determine the first two infinitesimal moments of a diffusion process, X, obtained from the ^processes as the jump magnitudes e go to zero and the jump rates X go to infinity - but in such a way that these moments remain finite and non-zero in the limit. We may call this a diffusion approximation to the original discontinuous random walk - and this approximation will have, as we have seen, continuous sample paths. The situation is illustrated in Fig. 12.3. Now as we have seen in Equation (9.2), the increments in the Poisson process Nt are such that: ' 1, with prob. XAt + o(At); 0, with prob. 1 - XAt + o(At), and the probabilities of other values are o(At); similarly for increments AN2 in the process N2. Thus, utilizing the fact that the increments are independent of previous values and that the two Poisson processes are independent, we have £[AXJXc(t) = x] = o(A£) (12.31) and Var[AXJX£(£) = x] = £2[Var(AN1) + Var(AN2)] " (12.32) = 2e2XAt + o(At). The first two infinitesimal moments, ae and ßc, of Xt are thus, from (12.31) When the Ito definition of stochastic integral is used in the interpretation of equation (12.20) we call it an Ito stochastic differential or Ito stochastic differential equation. The process X is called the solution of the equation. There are many accounts of existence and uniqueness theorems and properties of the solution - see for example the references at the beginning of this section. We will avoid technicalities and simply assume that the functions / and g are suitable. Ito's formula When one uses Ito's definition of stochastic integral, the usual rules of calculus sometimes must be modified. This is not the case with the Stratonovich definition considered below. In particular, one must be careful when changing variables as the following result, called Ito's formula, shows. We will not give a detailed proof, which can be found for example in Liptser and Shiryayev (1977). We will simply utilize the fact that (dW2) acts like At. (12.26) This statement can be understood in terms of the following probability one relation (see, for example the previous reference) for the so-called quadratic variation of W over the interval [0, t): lim £ lWh*i)-wW «-*co j'=0 ■t, where {0 = t0 < frj < ■ ■ ■ < t„ = t} is a partition of [0, t). This suggests we can write (dW)2 = t. and the latter can be written as (12.26). The significance of this result for our purposes is that whereas second order differentials such as (dt)2 can usually be ignored, (d W)2 is of the same order as dt and hence cannot be neglected. We are now ready to discuss Ito's formula. Notation. In the following we will use subscripts to denote partial differentiation with respect to a variable. Thus, for example, if h = h(x, t) then h - — dtdx Theorem 12.4 Let the random process 0, it will be seen in the exercises that all the higher order infinitesimal moments of X vanish as required by (12.3) - although some are zero regardless of the value of e. We can now use Theorem 12.5. Since (3(x, t) = l we have g2{x, t) = 1, so, choosing the positive root, g{x,t)= l. Hence the diffusion process approximating the random walk has the stochastic differential equation dX = dW. That is, X is a standard Wiener process. One may also illustrate this convergence to a Wiener process using characteristic functions - see the exercises. A mathematical model for the activity of a nerve cell (neuron) In Fig. 12.4 we depict a nerve cell or neuron - a component of the nervous system. The one depicted is a pyramidal cell of the part of the human brain called the cerebral cortex. Such cells are specialized to receive signals from other cells at the junctions called- synapses. If no signals are received, the electrical potential difference VM across the target cell membrane tends to a fixed value called the resting potential, which we designate by VR. Usually VR is about 70 millivolts (the inside being negative) or about ^th of the voltage of the familiar AA battery. Let us denote the difference between VM and VR by V: v=vM-vR, - see the left part of Fig. 12.5. V is called the depolarization. As explained at the end of Chapter 7, the incoming signals may excite the target cell: this makes VM less negative, in which case V is increased. We assume that each such incoming signal increases V by an amount aE and that such signals occur as a Poisson process NE with intensity XE. Alternatively an incoming signal may inhibit the target cell, thereby decreasing V (which can go negative), by an amount at. Let such inhibitory effects be generated by a Poisson process N{ with intensity Xj. These events result in K's executing a random walk as depicted on the right in Fig. 12.5. The state of the neuron at time t can thus be characterized by V{t) = aENE{t)-aINI{t), and we assume V(0) = 0, so the cell is initially at resting level. The reader can show that the mean and variance of the change in V during (t, t + At] are £[AK] = aEXEAt - a1XIAt + o(At), Incoming fibre Dendrites Cell body or soma Axon Figure 12.4 Here we depict a class of neuron called a pyramidal cell, showing the main components of axon, dendrites and soma. The neuron is shown connected into a network through synapses at which other cells transmit signals to the target cell Here a (+) indicates an excitatory junction whereas a (-) denotes an inhibitory one. and Var[A K] = a\XEAt + ajA, At + o(At). Dividing by At and taking the limit we obtain the first and second infinitesimal moments. We may construct a diffusion approximation V*(t) to V with the same infinitesimal mean and variance, viz., a(x, t) = aEXE — atXf, P(x, t) = a2EXE + a)kv Using Theorem 12.5, we find the following stochastic differential for V*: d V* = (a A - Mj) dt + d W. This will be recognized as the stochastic equation for a Wiener process with drift. c CD s 0_ Outside OmV Resting level —70mV Inhib Ex. Ex. Figure 12.5 On the left we show the relation between V, VR and VM. On the right are shown the variations which occur in V as signals come in from other cells. Two excitatory inputs (Ex.) and one inhibitory input (Inhib.) are depicted. The solid line is a trajectory for the process aENE(t) — atNused in the text to model V. The dashed line represents the more physiologically realistic trajectory involving decay to resting level (V=0) between input events. The mean and variance, as well as the transition density of the state of a nerve cell can be estimated using this approximation. In addition, the time taken for V or V* to reach a threshold value, 9, say, can be determined as a first passage time. When V reaches this threshold value, which is usually about 10 millivolts (i.e., 0.01 volts) from resting level, our target cell itself sends out a signal along its axon (see Fig. 12.3) to excite or inhibit other neurons. The whole collection of neurons is sometimes called a neural network. Details of such calculations can be found in Tuckwell (1988) where it is also seen that the Ornstein-Uhlenbeck process (see section 12.4) is a somewhat more realistic mathematical model for a neuron than the Wiener process with drift. In the absence of inputs the OUP model correctly predicts that the electrical potential across the neuron's membrane will return exponentially to resting level. A model for population growth in a random environment As we have seen in section 9.1, when a population grows in an unrestrained way, the population size x(r) at time t may be approximated by the solution of the deterministic differential equation dx dt = rx, (12.34) where the assumed constant growth rate r is the difference between the birth and death rates. This is the Malthusian law with exponential solutions x(t) = x(0)er', x(0) being the initial population size. There are often factors, particularly environmental ones, which make the birth and death rates chop and change from generation to generation. For example, in very cold or very dry seasons we might expect a higher death rate in populations of many species, so r would drop below some long-term average value. The same might happen in very hot and very wet seasons. On the other hand, if climatic conditions lead to an abundance of food, r might increase above its average value, but this would be complicated if an organism's predators also benefited. This suggests that a more accurate mathematical model of population growth would result if the growth rate r and hence the population size x are random processes. The simplest (but not necessarily accurate) assumption is that the growth rate r(t) is the sum of its constant mean value r and a fluctuating white noise, r(t) = f + cnv(t), where a is a variance parameter and w(t) is standard white noise, defined in section 11.3. Then we may write the model for the random population size X(t), ~ = (r + aw(t))X(t), dt or, more satisfactorily as the stochastic differential equation, dX = FX dt + ffX dW. (12.35) We let the initial value be X(0) = xo. As mentioned in the last section, the approach we have used has the apparent drawback of necessitating a choice of stochastic integral. Fortunately, however, in this case the derivation of a diffusion approximation to Malthusian growth has been carried out (Tuckwell and Walsh, 1983a) from first principles. This approximation satisfies Equation (12.35) interpreted as a Stratonovich equation. Note that this does not imply that the Stratonovich integral is better or more correct than the Ito integral. If we proceed with Equation (12.35) as a Stratonovich equation we can use the usual rules of calculus. In particular, we find that the change of variable Y=ln(X) leads to áY dt Thus dY dX dX dt 1 cLY I dt' dY dW — = r + a- dt di or, equivalently, dY=fdt + odW. Thus the transformed process Y is simply a Wiener process with drift. This is a particular case of a general transformation method (Tuckwell, 1973; 1974) in which the equation dX=f(X)[fidt + cdW] is transformed to a Wiener process with drift using the transformation "* 1 f(X') -dX' - see the exercises. Note that in our present application the possible values of X are in (0, oo) whereas the possible values of Y are in (—oo, oo). In addition, if X(0) = x0 and Y(0) = y0, then y0 = ln(x„). The transition density of a Wiener process with drift was given by Equation (12.15). If, in conjunction with that result, we use Equation (1.6) for the transformation of a density under a monotonie change of variable, we are immediately able to find the transition probability density of the population size X as p(x,t\x0) = 1 x^fhiaH exp -[ln(x/x0)-řt]: 2oh~ Now, we can see that the probability PE that the population eventually becomes extinct can be estimated from PE = lim Pt{X(t) < e|X(0) = x0} where e > 0 is arbitrarily small. Note that X can never reach exactly zero in this model - for we know that Y can never reach - oo. However, an extremely small population size implies extinction in a continuous model. hiti uniusion processes Applications c i a It will be seen in the exercises that fo, ifr>0; PE = U, if r = 0; [l, if r < 0. Thus, if the mean growth rate is negative, the population will become extinct with certainty, regardless of either the initial population size or how great the fluctuations (