<<

Statistical Methods for Differential

Mathieu Kessler Alexander Lindner & Michael Sørensen

Contents

1 Estimating functions for -type processes 1 by Michael Sørensen

1.1 Introduction 1 1.2 Lowfrequencyasymptotics 3 1.3 Martingaleestimatingfunctions 7 1.3.1 Asymptotics 8 1.3.2 Likelihoodinference 10 1.3.3 Godambe-Heydeoptimality 12 1.3.4 Small ∆-optimality 22 1.3.5 Simulatedmartingaleestimatingfunctions 27 1.3.6 Explicitmartingaleestimatingfunctions 30 1.3.7 Pearsondiffusions 34 1.3.8 Implementationofmartingaleestimatingfunctions 42 1.4 The likelihood 45 1.5 Non-martingaleestimatingfunctions 49 1.5.1 Asymptotics 49 1.5.2 Explicitnon-martingaleestimatingfunctions 51 1.5.3 Approximatemartingaleestimatingfunctions 54 1.6 High-frequencyasymptotics 56 1.7 High-frequencyasymptoticsinafixedtime-interval 63 1.8 Small-diffusion asymptotics 65 1.9 Non-Markovian models 70

iii iv CONTENTS 1.9.1 Prediction-basedestimatingfunctions 70 1.9.2 Asymptotics 76 1.9.3 Measurementerrors 77 1.9.4 Integrated and hypoelliptic stochastic differ- ential equations 78 1.9.5Sumsofdiffusions 80 1.9.6 Stochastic volatility models 83 1.9.7 Compartmentmodels 84 1.10 General asymptotics results for estimating functions 85 1.11 Optimalestimatingfunctions:generaltheory 88 1.11.1Martingaleestimatingfunctions 92

Bibliography 97 CHAPTER 1

Estimating functions for diffusion-type processes

Michael Sørensen University of Copenhagen

1.1 Introduction

In this chapter we consider parametric inference based on discrete time obser- vations X0,Xt1 ,...,Xtn from a d-dimensional . In most of the chapter the statistical model for the data will be a diffusion model given by a stochastic differential . We shall, however, also consider some examples of non-Markovian models, where we typically assume that the data are partial observations of a multivariate stochastic differential equation. We assume that the statistical model is indexed by a p-dimensional parameter θ. The focus will be on estimating functions.An estimating function is a p-dimen- sional function of the parameter θ and the data:

Gn(θ; X0,Xt1 ,...,Xtn ). Usually we suppress the dependence on the observations in the notation and write Gn(θ). We obtain an estimator by solving the equation

Gn(θ)=0. (1.1) Estimating functions provide a general framework for finding estimators and studying their properties in many different kinds of statistical models. The es- timating function approach has turned out to be very useful for discretely sam- pled parametric diffusion-type models, where the likelihood function is usually not explicitly known. Estimating functions are typically constructed by com- bining relationships (dependent on the unknown parameter) between an obser- vation and one or more of the previous observations that are informative about the parameters.

As an example, suppose the statistical model for the data X0,X∆,X2∆,...,

1 2 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

Xn∆ is the one-dimensional stochastic differential equation dX = θ tan(X )dt + dW , t − t t where θ > 0 and W is a . The state-space is ( π/2,π/2). This model will be considered in more detail in Subsection 1.3.6.−For this process Kessler & Sørensen (1999) proposed the estimating function n (θ+ 1 )∆ Gn(θ)= sin(X(i 1)∆) sin(Xi∆) e− 2 sin(X(i 1)∆) , − − − i=1 X h i which can be shown to be a martingale, when θ is the true parameter. For such martingale estimating functions, asymptotic properties of the estimators as the number of observations tends to infinity can be studied by means of martingale theory, see Subsection 1.3.1. An explicit estimator θˆn of the parameter θ is obtained by solving the estimating equation (1.1): n i=1 sin(X(i 1)∆) sin(Xi∆) 1 ˆ 1 − θn = ∆− log n 2 , i=1 sin(X(i 1)∆) − 2 P −  provided that P n

sin(X(i 1)∆) sin(Xi∆) > 0. (1.2) − i=1 X If this condition is not satisfied, the estimating equation (1.1) has no solution, but fortunately it can be shown that the probability that (1.2) holds tends to one as n tends to infinity. As illustrated by this example, it is quite possible that the estimating equation (1.1) has no solution. We shall give general conditions that ensure the existence of a unique solution as the number of observations tend to infinity. The idea of using estimating equations is an old one and goes back at least to Karl Pearson’s introduction of the method of moments. The term estimat- ing function may have been coined by Kimball (1946). In the econometric literature, the method was introduced by Hansen (1982) and is known as the generalized method of moments (GMM). A general asymptotic theory for estimating functions is presented in Section 1.10, and Section 1.11 reviews the theory of optimal estimating functions. Given a collection of relations between observations at different time points that can be used for estimation, this theory clarifies how to combine the rela- tions in an optimal way, i.e. in such a way that the most efficient estimator is obtained. In Section 1.2 we present conditions ensuring that estimators from estimating functions are consistent and asymptotically normal under the so- called low frequency asymptotics, which is the same as usual large sample asymptotics. In Section 1.3 we present martingale estimating functions for dif- fusion models including asymptotics and two optimality criteria. One of these LOW FREQUENCY ASYMPTOTICS 3 criteria, small ∆-optimality, is particular of diffusion models. Likelihood in- ference is included as a particular case. It is also discussed how to implement martingale estimating functions. There are several methods available for cal- culating approximations to the likelihood function. These a briefly reviewed in Section 1.4, where a particular expansion approach is presented in detail. Non- martingale estimating functions are considered in Section 1.5. In important aspect of the statistical theory of diffusion processes is that a number of alter- native asymptotic scenarios particular to diffusions are available to supplement the traditional large sample asymptotics. High frequency asymptotics, high fre- quency asymptotics in a fixed time-interval and small-diffusion asymptotics are presented in the Sections 1.6, 1.7, and 1.8. A number of non-Markovian mod- els are considered in Section 1.9, including observations with measurement er- rors, integrated diffusions, sums of diffusions, stochastic volatility models and compartment models. A general tool for these models are prediction-based es- timating functions, which generalize the martingale estimating functions and share some of their convenient features.

1.2 Low frequency asymptotics

In this section, we assume that observations have been made at the equidistant time points i∆, i = 1,...,n, and consider the classical large sample asymp- totic scenario, where the time between observations, ∆, is fixed, and the num- ber of observations, n, goes to infinity. Since ∆ is fixed, we will generally suppress ∆ in the notation in this section. We assume that the statistical model is indexed by a p-dimensional parameter θ Θ, which we want to estimate. ∈ The corresponding probability measures are denoted by Pθ. The distribution of the data is given by the true probability , which we denote by P .

Under the true probability measure, P , it is assumed that Xi∆ is a stationary process with state space D IRd. We study the asymptotic{ properties} of an ⊆ estimator, θˆn, obtained by solving the estimating equation (1.1) when Gn is an estimating function of the general form 1 n Gn(θ)= g(X(i r+1)∆,...,Xi∆; θ), (1.3) n − i=r X where r is a fixed integer smaller than n, and g is a suitable function with values in IRp. All estimators discussed in this chapter can be represented in this way. We shall present several useful examples of how g can be chosen in the subsequent sections. A priori there is no guarantee that a unique solution to (1.1) exists, but conditions ensuring this for large sample sizes are given below. By a Gn–estimator, we mean an estimator, θˆn, which solves (1.1) when n the data belongs to a subset An D , and is otherwise given a value δ / Θ. We give results ensuring that, as⊆n , the probability of A tends to one.∈ → ∞ n 4 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

Let Q denote the joint distribution of (X∆,...,Xr∆), and Q(f) the expecta- tion of f(X ,...,X ) for a function f : Dr IR. To obtain our asymptotic ∆ r∆ 7→ results about Gn–estimators, we need to assume that a law of large numbers (an ergodic theorem) as well as a central limit theorem hold. Specifically, we assume that as n → ∞n 1 P f(X(i r+1)∆,...,Xi∆) Q(f) (1.4) n − −→ i=r X for any function f : Dr IR such that Q( f ) < , and that the estimating function (1.3) satisfies 7→ | | ∞ 1 n g(X(i r+1)∆,...,Xi∆; θ) D N(0, V (θ)) (1.5) √n − −→ i=r X under P for any θ Θ for which Q(g(θ)) = 0. Here V (θ) is a positive definite p p-matrix.∈ Moreover, g(θ) denotes the function (x ,...,x ) × 1 r 7→ g(x ,...,x ; θ), convergence in probability under P is indicated by P , and 1 r −→ D denotes convergence in distribution. −→ The following condition ensures the existence of a consistent Gn–estimator. T We denote transposition of matrices by , and ∂ T G (θ) denotes the p p- θ n × matrix, where the ijth entry is ∂θj Gn(θ)i.

Condition 1.2.1 A parameter value θ¯ int Θ and a neighbourhood N of θ¯ in Θ exist such that: ∈

(1) The function g(θ) : (x1,...,xr) g(x1,...,xr; θ) is integrable with respect to Q for all θ N, and 7→ ∈ Q g(θ¯) =0. (1.6)  (2) The function θ g(x1,...,xr; θ) is continuously differentiable on N for all (x ,...,x ) D7→r. 1 r ∈ (3) The function (x1,...,xr) ∂θT g(x1,...,xr; θ) is dominated for all θ N by a function which is integrable7→ k with respect to Qk. ∈ (4) The p p matrix × ¯ W = Q ∂θT g(θ) (1.7) is invertible. 

Here and later Q(g(θ)) denotes the vector (Q(gj(θ)))j=1,...,p, where gj is the jth coordinate of g, and Q (∂ T g(θ)) is the matrix Q ∂ g (θ) . θ { θj i }i,j=1,...,p To formulate the uniqueness result in the following theorem , we need the con- cept of locally dominated integrability. A function f : Dr Θ IRq is × 7→ LOW FREQUENCY ASYMPTOTICS 5 called locally dominated integrable with respect to Q if for each θ Θ ′ ∈ there exists a neighbourhood Uθ′ of θ′ and a non-negative Q-integrable func- r tion hθ′ : D IR such that f(x1,...,xr; θ) hθ′ (x1,...,xr) for all (x ,...,x ,θ)7→ Dr U . | | ≤ 1 r ∈ × θ′

Theorem 1.2.2 Assume Condition 1.2.1 and (1.5). Then a θ¯-consistent Gn– estimator, θˆn, exists, and

1 T 1 √n(θˆ θ ) D N 0, W − V W − (1.8) n − 0 −→ p   under P , where V = V (θ¯). If, moreover, the function g(x1,...,xr; θ) is lo- cally dominated integrable with respect to Q and Q(g(θ)) =0 for all θ = θ,¯ (1.9) 6 6 then the estimator θˆn is the unique Gn–estimator on any bounded subset of Θ containing θ¯ with probability approaching one as n . → ∞ P Remark: By a θ¯-consistent estimator is meant that θˆn θ¯ as n . If the true model belongs to the statistical model, i.e. if P =−→P for some→ ∞θ Θ, θ0 0 ∈ then the estimator θˆn is most useful if Theorem 1.2.2 holds with θ¯ = θ0. Note that because θ¯ int Θ,a θ¯-consistent estimator Gn–estimator θˆn will satisfy G (θˆ )=0 with∈ probability approaching one as n . n n → ∞ In order to prove Theorem 1.2.2, we need the following uniform law of large numbers.

Lemma 1.2.3 Consider a function f : Dr K IRq, where K is a compact × 7→ subset of Θ. Suppose f is a of θ for all (x1,...,xr) Dr, and that there exists a Q-integrable function h : Dr IR such that∈ 7→ f(x1,...,xr; θ) h(x1,...,xr) for all θ K. Then the function θ Qk (f(θ)) is continuous,k ≤ and ∈ 7→ n 1 P sup f(X(i r+1)∆,...,Xi∆; θ) Q(f(θ)) 0. (1.10) k n − − k → θ K i=r ∈ X Proof: That Q(f(θ)) is continuous follows from the dominated convergence theorem. To prove (1.10), define for η > 0:

k(η; x1,...,xr) = sup f(x1,...,xr; θ′) f(x1,...,xr; θ) , θ,θ M: θ θ η k − k ′∈ k ′− k≤ and let k(η) denote the function (x1,...,xr) k(η; x1,...,xr). Since k(η) 2h, it follows from the dominated convergence7→ theorem that Q(k(η)) 0 as≤ η 0. Moreover, Q(f(θ)) is uniformly continuous on the compact→ set K. Hence→ for any given ǫ > 0, we can find η > 0 such that Q(k(η)) ǫ and such that θ θ < η implies that Q(f(θ)) Q(f(θ )) ǫ for θ,θ≤ K. k − ′k k − ′ k≤ ′ ∈ 6 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

Define the balls Bη(θ)= θ′ : θ θ′ < η . Since K is compact, there exists a finite covering { k − k } m K B (θ ), ⊆ η j j=1 [ where θ1,...,θm K, so for every θ K we can find θℓ, ℓ 1,...,m , such that θ B (θ∈). Thus with ∈ ∈ { } ∈ η ℓ n 1 Fn(θ)= f(X(i r+1)∆,...,Xi∆; θ) n − i=r X we have F (θ) Q(f(θ)) k n − k F (θ) F (θ ) + F (θ ) Q(f(θ )) + Q(f(θ )) Q(f(θ)) ≤ k n − n ℓ k k n ℓ − ℓ k k ℓ − k 1 n k(η; X(ν r+1)∆,...,Xν∆)+ Fn(θℓ) Q(f(θℓ)) + ǫ ≤ n − k − k ν=r X n 1 k(η; X(ν r+1)∆,...,Xν∆) Q(k(η)) ≤ n − − ν=r X +Q(k(η)) + F (θ ) Q(f(θ )) + ǫ k n ℓ − ℓ k Z +2ǫ, ≤ n where 1 n Zn = k(η; X(ν r+1)∆,...,Xν∆) Q(k(η)) n − − ν=r X + max Fn(θℓ) Q( f(θℓ)) . 1 ℓ m k − k ≤ ≤ By (1.4), P (Z >ǫ) 0 as n , so n → → ∞

P sup Fn(θ) Q(f(θ)) > 3ǫ 0 θ K k − k →  ∈  for all ǫ> 0. 2

Proof (of Theorem 1.2.2): The existence of a θ¯-consistent Gn–estimator θˆn follows from Theorem 1.10.2. Condition (i) follows from (1.4) and (1.6). De- fine the function W (θ)= Q (∂θT g(θ)). Then condition (iii) in Theorem 1.10.2 is equal to Condition 1.2.1 (4). Finally, let M be a compact subset of N con- ¯ taining θ. Then the conditions of Lemma 1.2.3 are satisfied for f = ∂θT g, so MARTINGALEESTIMATINGFUNCTIONS 7 (1.158) is satisfied. The asymptotic normality, (1.20), follows from Theorem 1.10.4 and (1.5).

In order to prove the last statement, let K be a compact subset of Θ containing θ¯. By the finite covering property of a compact set, it follows from the local dominated integrability of g that g satisfies the conditions of Lemma 1.2.3. Hence (1.159) holds with G(θ)= Q(g(θ)) and M = K. From the local domi- nated integrability of g and the dominated convergence theorem it follows that G(θ) is a continuous function, so (1.9) implies that inf G(θ) > 0, K B¯ǫ(θ¯) | | \ for all ǫ > 0, where B¯ǫ(θ) is the closed ball with radius ǫ centered at θ. By Theorem 1.10.3 it follows that (1.161) holds with M = K for every ǫ> 0. Let ˆ ˆ ˆ ˆ θn′ be a Gn–estimator, and define a Gn–estimator by θn′′ = θn′ 1 θn′ K + ˆ ˆ ˆ { ∈ } θn1 θn′ / K , where 1 denotes an indicator function, and θn is the consistent { ∈ } ˆ Gn–estimator we know exists. By (1.161) the estimator θn′′ is consistent, so by Theorem 1.10.2, P (θˆ = θˆ ) 0 as n . Hence θˆ is eventually the n 6 n′′ → → ∞ n unique Gn–estimator on K. 2

1.3 Martingale estimating functions

In this section we consider observations X0,Xt1 ,...,Xtn of a d-dimensional diffusion process given by the stochastic differential equation

dXt = b(Xt; θ)dt + σ(Xt; θ)dWt, (1.11) where σ is a d d-matrix and W a d-dimensional standard Wiener process. We denote the state× space of X by D. When d = 1, the state space is an interval (ℓ, r), where ℓ could possibly be , and r might be . The drift b and the diffusion matrix σ depend on a parameter−∞ θ which varies∞ in a subset Θ of IRp. The equation (1.11) is assumed to have a weak solution, and the coefficients b and σ are assumed to be smooth enough to ensure, for every θ Θ, the ∈ uniqueness of the law of the solution, which we denote by Pθ. We denote the true parameter value by θ0. We suppose that the transition distribution has a density y p(∆,x,y; θ) with respect to the Lebesgue measure on D, and that p(∆,x,y7→; θ) > 0 for all y D. The transition density is the conditional density under P of X ∈ θ t+∆ given that Xt = x. We shall, in this section, be concerned with statistical inference based on esti- 8 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES mating functions of the form n

Gn(θ)= g(∆i,Xti 1 ,Xti ; θ). (1.12) − i=1 X where g is a p-dimensional function which satisfies that

g(∆,x,y; θ)p(∆,x,y; θ)dy =0 (1.13) ZD for all ∆ > 0, x D and θ Θ. Thus, by the Markov property, the stochastic ∈ ∈ process Gn(θ) n IN is a martingale with respect to n n IN under Pθ. Here { } ∈ {F } ∈ and later n = σ(Xti : i n). An estimating function with this property is called a martingaleF estimating≤ function.

1.3.1 Asymptotics

In this subsection we give asymptotic results for estimators obtained from mar- tingale estimating functions as the number of observations goes to infinity. To simplify the exposition the observation time points are assumed to be equidis- tant, i.e., ti = i∆, i = 0, 1,...,n. Since ∆ is fixed, we will in most cases suppress ∆ in the notation and write for example p(x, y; θ) and g(x, y; θ). It is assumed that the diffusion given by (1.11) is ergodic, that its probability measure has density function µ for all θ Θ, and that X µ θ ∈ 0 ∼ θ under Pθ. Thus the diffusion is stationary. When the observed process, X, is a one-dimensional diffusion, the follow- ing simple conditions ensure ergodicity, and an explicit expression exists for the density of the invariant probability measure. The scale measure of X has Lebesgue density x b(y; θ) (1.14) s(x; θ) = exp 2 2 dy , x (ℓ, r), − # σ (y; θ) ∈  Zx  where x# (ℓ, r) is arbitrary. ∈ Condition 1.3.1 The following holds for all θ Θ: ∈ r x# s(x; θ)dx = s(x; θ)dx = # ∞ Zx Zℓ and r 2 1 [s(x; θ)σ (x; θ)]− dx = A(θ) < . ∞ Zℓ Under Condition 1.3.1 the process X is ergodic with an invariant probability measure with Lebesgue density 2 1 µ (x) = [A(θ)s(x; θ)σ (x; θ)]− , x (ℓ, r); (1.15) θ ∈ MARTINGALEESTIMATINGFUNCTIONS 9 for details see e.g. Skorokhod (1989). For general one-dimensional diffusions, 2 1 the measure with Lebesgue density proportionalto [s(x; θ)σ (x; θ)]− is called the speed measure.

2 Let Qθ denote the probability measure on D given by

Qθ(dx, dy)= µθ(x)p(∆,x,y; θ)dxdy. (1.16)

This is the distribution of two consecutive observations (X∆(i 1),X∆i). Under the assumption of ergodicity the law of large numbers (1.4) is− satisfied for any function f : D2 IR such that Q( f ) < , see e.g. Skorokhod (1989). 7→ | | ∞ We impose the following condition on the function g in the estimating function (1.12)

T Qθ g(θ) g(θ) = (1.17)

T g(y, x; θ) g(y, x; θ)µθ(x)p(x, y; θ)dydx < , 2 ∞ ZD for all θ Θ. By (1.4), ∈ n 1 Pθ g(X∆i,X∆(i 1); θ′) Qθ(g(θ′)), (1.18) n − −→ i=1 X for all θ,θ Θ. Since the estimating function G (θ) is a square integrable ′ ∈ n martingale under Pθ, the asymptotic normality in (1.5) follows without further conditions from the central limit theorem for martingales, see Hall & Heyde (1980). This result for stationary processes goes back to Billingsley (1961). In the martingale case the asymptotic covariance matrix V (θ) in (1.5) is given by T V (θ)= Qθ0 g(θ)g(θ) . (1.19) Thus we have the following particular case of Theorem 1.2.2.

Theorem 1.3.2 Assume Condition 1.2.1 is satisfied with r = 2, θ¯ = θ0, and

Q = Qθ0 , where θ0 is the true parameter value, and that (1.17) holds for θ = θ0. Then a θ0-consistent Gn–estimator θˆn exists, and

1 T 1 √n(θˆ θ ) D N 0, W − V W − (1.20) n − 0 −→ p  ¯  under Pθ0 , where W is given by (1.7) with θ = θ0, and V = V (θ0) with V (θ) given by (1.19). If, moreover, the function g(x, y; θ) is locally dominated integrable with respect to Qθ0 and Q (g(θ)) =0 for all θ = θ , (1.21) θ0 6 6 0 then the estimator θˆn is the unique Gn–estimator on any bounded subset of Θ containing θ with probability approaching one as n . 0 → ∞ 10 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

In practice we do not know the value of θ0, so it is necessary to check that the conditions of Theorem 1.3.2 hold for any value of θ int Θ. 0 ∈ The asymptotic covariance matrix of the estimator θˆn can be estimated consis- tently by means of the following theorem.

Theorem 1.3.3 Under Condition 1.2.1 (2) – (4) (with r = 2, θ¯ = θ0, and

Q = Qθ0 ), n P 1 ˆ θ0 Wn = ∂θT g(X(i 1)∆,Xi∆; θn) W, (1.22) n − −→ i=1 X where θˆn is a θ0-consistent estimator. The probability that Wn is invertible approaches one as n . If, moreover, the function (x, y) g(x, y; θ) is dominated for all →θ ∞N by a function which is square7→ integrable k withk ∈ respect to Qθ0 , then n P 1 ˆ ˆ T θ0 Vn = g(X(i 1)∆,Xi∆; θn)g(X(i 1)∆,Xi∆; θn) V. (1.23) n − − −→ i=1 X

Proof: Let C be a compact subset of N such that θ0 int C. By Lemma 1 n ∈ 1.2.3, i=1 ∂θT g(X(i 1)∆,Xi∆; θ) converges to Qθ0 (∂θT g(θ)) in probabil- n − ity uniformly for θ C. This implies (1.22) because θˆ converges in probabil- P ∈ n ity to θ0. The result about invertibility follows because W is invertible. Also the 1 n uniform convergence in probability for θ C of n i=1 g(X(i 1)∆,Xi∆; θ) T T ∈ − g(X(i 1)∆,Xi∆; θ) to Qθ0 (g(θ)g(θ) ) follows from Lemma 1.2.3. − P 2

1.3.2 Likelihood inference

The diffusion process X is a Markov process, so the likelihood function based on the observations X ,X , ,X , conditional on X , is 0 t1 · · · tn 0 n

Ln(θ)= p(ti ti 1,Xti 1 ,Xti ; θ), (1.24) − − − i=1 Y where y p(s,x,y; θ) is the transition density and t0 = 0. Under weak regularity7→ conditions the maximum likelihood estimator is efficient, i.e. it has the smallest asymptotic variance among all estimators. The transition density is only rarely explicitly known, but several numerical approaches and accurate approximations make likelihood inference feasible for diffusion models. We shall return to the problem of calculating the likelihood function in Subsection 1.4. MARTINGALEESTIMATINGFUNCTIONS 11 The vector of partial of the log-likelihood function with respect to the coordinates of θ, n

Un(θ)= ∂θ log Ln(θ)= ∂θ log p(∆i,Xti 1 ,Xti ; θ), (1.25) − i=1 X where ∆i = ti ti 1, is called the score function (or score vector). Here it is obviously assumed− − that the transition density is a differentiable function of θ. The maximum likelihood estimator usually solves the estimating equation Un(θ)=0. The score function is a martingale with respect to n n IN under {F } ∈ Pθ, which is easily seen provided that the following interchange of differenti- ation and integration is allowed:

Eθ (∂θ log p(∆i,Xti 1 ,Xti ; θ) Xt1 ,...,Xti 1 ) − | −

∂θp(∆i,Xti 1 ,y; θ) − = p(∆i,Xti 1 ,y,θ)dy − D p(∆i,Xti 1 ,y; θ) Z −

= ∂θ p(∆i,Xti 1 ,y; θ)dy =0. − ZD Since the score function is a martingale estimating function, the asymptotic results in the previous subsection applies to the maximum likelihood estimator. Asymptotic results for the maximum likelihood estimator in the low frequency (fixed ∆) asymptotic scenario considered in that subsection were established by Dacunha-Castelle & Florens-Zmirou (1986). Asymptotic results when the observations are made at random time points were obtained by A¨ıt-Sahalia & Mykland (2003).

In the case of likelihood inference, the function Qθ0 (g(θ)) appearing in the identifiability condition (1.21) is related to the Kullback-Leibler between the models. Specifically, if the following interchange of differentiation and integration is allowed, Q (∂ log p(x,y,θ)) = ∂ Q (log p(x,y,θ)) = ∂ K¯ (θ,θ ), θ0 θ θ θ0 − θ 0 where K¯ (θ,θ0) is the average Kullback-Leibler divergence between the tran- sition distributions under Pθ0 and Pθ given by

¯ K(θ,θ0)= K(θ,θ0; x) µθ0 (dx), ZD with

K(θ,θ0; x)= log[p(x, y; θ0)/p(x, y; θ)]p(x, y; θ0) dy. ZD Thus the identifiability condition can be written in the form ∂θK¯ (θ,θ0) = 0 for all θ = θ . The quantity K¯ (θ,θ ) is sometimes referred to as the Kullback-6 6 0 0 12 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES Leibler divergence between the two Markov chain models for the observed process X under P and P . { i∆} θ0 θ A simple approximation to the likelihood function is obtained by approximat- ing the transition density by a Gaussian density with the correctfirst and second conditional moments. For a one-dimensional diffusion we get 1 (y F (∆, x; θ))2 p(∆,x,y; θ) q(∆,x,y; θ)= exp − ≈ − 2φ(∆, x; θ) 2πφ(∆, x; θ)   where p r F (∆, x; θ)= E (X X = x)= yp(∆,x,y; θ)dy. (1.26) θ ∆| 0 Zℓ and φ(∆, x; θ)= (1.27) r Var (X X = x)= [y F (∆, x; θ)]2p(∆,x,y; θ)dy. θ ∆| 0 − Zℓ In this way we obtain the quasi-likelihood n

Ln(θ) QLn(θ)= q(∆i,Xti 1 ,Xti ; θ), ≈ − i=1 Y and by differentiation with respect to the parameter vector, we obtain the quasi- score function n ∂θF (∆i,Xti 1 ; θ) − ∂θ log QLn(θ)= [Xti F (∆i,Xti 1 ; θ)] (1.28) φ(∆ ,X ; θ) − − i=1 i ti 1 X  −

∂θφ(∆i,Xti 1 ; θ) 2 − + (Xti F (∆i,Xti 1 ; θ)) φ(∆i,Xti 1 ; θ) . 2 − − 2φ(∆i,Xti 1 ; θ) − − −    It is clear from (1.26)and (1.27) that ∂θ log QLn(θ) n IN is a martingale with { } ∈ respect to n n IN under Pθ. This quasi-score function is a particular case of the quadratic{F martingale} ∈ estimating functions considered by Bibby & Sørensen (1995) and Bibby & Sørensen (1996). Maximum quasi-likelihood estimation for diffusions was considered by Bollerslev & Wooldridge (1992).

1.3.3 Godambe-Heyde optimality

In this section we present a general way of approximating the score function by means of martingales of a similar form. Suppose we have a collection of real valued functions hj (x, y, ; θ), j =1,...,N, satisfying

hj(x, y; θ)p(x, y; θ)dy =0 (1.29) ZD MARTINGALEESTIMATINGFUNCTIONS 13 for all x D and θ Θ. Each of the functions h could be used separately ∈ ∈ j to define an estimating function of the form (1.3) with g = hj, but a better approximation to the score function, and hence a more efficient estimator, is obtained by combining them in an optimal way. Therefore we consider esti- mating functions of the form n

Gn (θ)= a(X(i 1)∆,θ)h(X(i 1)∆,Xi∆; θ), (1.30) − − i=1 X where h = (h ,...,h )T , and the p N weight matrix a(x, θ) is a function 1 N × of x such that (1.30) is Pθ-integrable. It follows from (1.29) that Gn (θ) is a martingale estimating function, i.e., it is a martingale under P for all θ Θ. θ ∈ The matrix a determines how much weight is given to each of the hj s in the estimation procedure. This weight matrix can be chosen in an optimal way using the theory of optimal estimating functions reviewed in Section 1.11. The optimal weight matrix, a∗, gives the estimating function of the form (1.30) that provides the best possible approximation to the score function (1.25) in a mean square sense. Moreover, the optimal g∗(x, y; θ) = a∗(x; θ)h(x, y; θ) is obtained from ∂θ log p(x, y; θ) by projection in a certain space of square integrable functions, for details see Section 1.11.

The choice of the functions hj , on the other hand,is an art rather thana science. The ability to tailor these functionsto a given model or to particular parameters of interest is a considerable strength of the estimating functions methodology. It is, however, also a source of weakness, since it is not always clear howbest to choose the hj s. In the following and in the Subections 1.3.6 and 1.3.7, we shall present ways of choosing these functions that usually work well in practice. Also the theory in Subsection 1.3.4 and Section 1.6 casts interesting light on this problem.

Example 1.3.4 The martingale estimating function (1.28) is of the type (1.30) with N =2 and h (x, y; θ) = y F (∆, x; θ), 1 − h (x, y; θ) = (y F (∆, x; θ))2 φ(∆,x,θ), 2 − − where F and φ are given by (1.26) and (1.27). The weight matrix is ∂ F (∆, x; θ) ∂ φ(∆, x; θ) θ , θ , (1.31) φ(∆, x; θ) 2φ2(∆, x; θ)∆   which we shall see is an approximation to the optimal weigth matrix. 2

In the econometricsliterature, a popular way of using functions like hj (x, y, ; θ), j = 1,...,N, to estimate the parameter θ is the generalized method of mo- ments (GMM) of Hansen (1982). In practice, the method is often implemented 14 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES as follows, see e.g. Campbell, Lo & MacKinlay (1997). Consider n 1 Fn(θ)= h(X(i 1)∆,Xi∆; θ). n − i=1 X Under weak conditions, cf. Theorem 1.3.3, a consistent estimator of the asymp- totic covariance matrix M of √nFn(θ0) is n 1 ˜ ˜ T Mn = h(X(i 1)∆,Xi∆; θn)h(X(i 1)∆,Xi∆; θn) , n − − i=1 X where θ˜n is a θ0-consistent estimator (for instance obtained by minimizing T Fn(θ) Fn(θ)). The GMM-estimator is obtained by minimizing the function T 1 Hn(θ)= Fn(θ) Mn− Fn(θ). The corresponding estimating function is obtained by differentiation with re- spect to θ 1 ∂θHn(θ)= Dn(θ)Mn− Fn(θ), where by (1.4) n 1 T Pθ0 T Dn(θ)= ∂θh(X(i 1)∆,Xi∆; θ) Qθ0 ∂θh(θ) . n − −→ i=1 X  Hence the estimating function ∂θHn(θ) is asymptotically equivalent to an es- timating function of the form (1.30) with a weight matrix that does not depend on the data T 1 a(x, θ)= Qθ0 ∂θh(θ) M − . We see that the GMM-estimators described here are covered by the theory for martingale estimating functions presented in this section.

We now return to the problemof finding the optimal estimating function Gn∗ (θ), i.e. the estimating functions of the form (1.30) with the optimal weight matrix. We assume that the functions hj satisfy the following condition.

Condition 1.3.5

(1) The functions y hj (x, y; θ), j = 1,...N, are linearly independent for all x D and θ Θ7→. ∈ ∈ (2) The functions y hj(x, y; θ), j = 1,...N, are square integrable with respect to p(x, y; θ) for7→ all x D and θ Θ. ∈ ∈ (3) h(x, y; θ) is differentiable with respect to θ.

(4) The functions y ∂θi hj (x, y; θ) are integrable with respect to p(x, y; θ) for all x D and θ 7→Θ. ∈ ∈ MARTINGALEESTIMATINGFUNCTIONS 15 The class of estimating functions considered here is a particular case of the class treated in detail in Example 1.11.4. By (1.182), the optimal choice of the weight matrix a is given by 1 a∗(x; θ)= Bh(x; θ) Vh(x; θ)− , (1.32) where T Bh(x; θ)= ∂θh(x, y; θ) p(x, y; θ)dy (1.33) ZD and T Vh(x; θ)= h(x, y; θ)h(x, y; θ) p(x, y; θ)dy. (1.34) ZD The matrix Vh(x; θ) is invertible because the functions hj , j = 1,...N, are linearly independent. Compared to (1.182), we have omitted a minus here. This can be done because an optimal estimating function multiplied by an invertible p p-matrix is also an optimal estimating function and yields the same esti- mator.×

The asymptotic variance of an optimal estimator, i.e. a Gn∗ –estimator, is sim- pler than the general expression in (1.20) because in this case the matrices W and V given by (1.7) and (1.19) are identical and given by (1.35). This is a general property of optimal estimating functions as discussed in Section 1.11. The result can easily be verified under the assumption that a∗(x; θ) is a differ- entiable function of θ. By (1.29)

[∂θi a∗(x; θ)] h(x, y; θ)p(x, y; θ)dy =0, ZD so that

W = ∂θT [a∗(x; θ0)h(x, y; θ0)]Qθ0 (dx, dy) D2 Z T 1 T = µθ0 (a∗(θ0)Bh(θ0) )= µθ0 Bh(θ0)Vh(θ0)− Bh(θ0) , and by direct calculation  T T V = Qθ0 (a∗(θ0)h(θ0)h(θ0) a∗(θ0) ) (1.35) 1 T = µθ0 Bh(θ0)Vh(θ0)− Bh(θ0) .  Thuswe have, asa corollaryto Theorem1.3.2,that if g∗(x,y,θ)= a∗(x; θ)h(x, y; θ) satisfies the conditions of Theorem 1.3.2, then the θˆn of consistent Gn∗ –estimators has the asymptotic distribution

1 √n(θˆ θ ) D N 0, V − . (1.36) n − 0 −→ p  Example 1.3.6 Consider the martingale estimating function of form (1.30) with N = 2 and with h1 and h2 as in Example 1.3.4, where the diffusion 16 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES is one-dimensional. The optimal weight matrix has columns given by

∂θφ(x; θ)η(x; θ) ∂θF (x; θ)ψ(x; θ) a∗ (x; θ) = − 1 φ(x; θ)ψ(x; θ) η(x; θ)2 − ∂θF (x; θ)η(x; θ) ∂θφ(x; θ)φ(x; θ) a∗ (x; θ) = − , 2 φ(x; θ)ψ(x; θ) η(x; θ)2 − where η(x; θ)= E ([X F (x; θ)]3 X = x) θ ∆ − | 0 and ψ(x; θ)= E ([X F (x; θ)]4 X = x) φ(x; θ)2. θ ∆ − | 0 − For the square-root diffusion (the a.k.a. the CIR-model) dX = β(X α)dt + τ X dW , X > 0, (1.37) t − t − t t 0 where α,β,τ > 0, the optimal weights canp be found explicitly. For this model β∆ β∆ F (x; θ) = xe− + α(1 e− ) 2 − τ 1 2β∆ β∆ 1 φ(x; θ) = ( α x)e− (α x)e− + α β 2 − − − 2 4  τ β∆ 2β∆ η(x; θ) = α 3(α x)e− + 3(α 2x)e− 2β2 − − − 3β∆ (α 3x)e− − − 6 3τ 4β∆ 3β∆  ψ(x; θ) = (α 4x)e− 4(α 3x)e− 4β3 − − − 2β∆ β∆ 2 + 6(α 2x)e− 4(α x)e− + α +2φ(x; θ) . − − − We give a method to derive these expression in Subsection 1.3.6.

The expressions for a1∗ and a2∗ can for general diffusions be simplified by the approximations η(t, x; θ) 0 and ψ(t, x; θ) 2φ(t, x; θ)2, (1.38) ≈ ≈ which would be exactly true if the transition density were a Gaussian density function. If we insert these Gaussian approximations into the expressions for a1∗ and a2∗ , we obtain the weight functions in (1.28). When ∆ is not large this can be justified, because the transition distribution is not far from Gaussian. In Section 1.4 we present a data transformation after which the transition distri- bution is close to a normal distribution. 2

In Subsections 1.3.6 and 1.3.7 we present martingale estimating functions for which the matrices Bh(x; θ) and Vh(x; θ) can be found explicitly, but for most models these matrices must be found by , a problem considered in Subsection 1.3.5. In situations where a∗ must be determined by a relatively MARTINGALEESTIMATINGFUNCTIONS 17 time consuming numerical method, it might be preferable to use the estimating function n ˜ Gn• (θ)= a∗(X(i 1)∆; θn)h(X(i 1)∆,Xi∆; θ), (1.39) − − i=1 X where θ˜n is a θ0-consistent estimator, for instance obtained by some simple choice of the weight matrix a. In this way a∗ needs to be calculated only once per observationpoint, whereas the weight matrix must be recalculated for every call of Gn∗ (θ) given by (1.30) with a = a∗. Typically, Gn∗ (θ) will be called many times in a numerical search procedure in order to find the Gn∗ -estimator. Under weak regularity conditions, the Gn• -estimator has the same efficiency as the optimal Gn∗ -estimator; see e.g. Jacod & Sørensen (2009). Most martingale estimating functions proposed in the literature are of the form (1.30) with h (x, y; θ)= f (y; θ) πθ (f (θ))(x), (1.40) j j − ∆ j or more specifically, n θ Gn (θ)= a(X(i 1)∆,θ) f(Xi∆; θ) π∆(f(θ))(X(i 1)∆) . (1.41) − − − i=1 X   T N θ Here f = (f1,...,fN ) maps D Θ into IR , and π∆ denotes the transition ×

πθ(f)(x)= f(y)p(s,x,y; θ)dy = E (f(X ) X = x), (1.42) s θ s | 0 ZD applied to each coordinate of f. The polynomial estimating functions given by j fj(y) = y , j = 1,...,N, are an example. For martingale estimating func- tions of the special form (1.41), the expression for the optimal weight matrix simplifies a bit because B (x; θ) = πθ (∂ f (θ))(x) ∂ πθ (f (θ))(x), (1.43) h ij ∆ θi j − θi ∆ j i =1,...p, j =1,...,N, and V (x; θ) = πθ (f (θ)f (θ))(x) πθ (f (θ))(x)πθ (f (θ))(x), (1.44) h ij ∆ i j − ∆ i ∆ j i, j = 1,...,N. If the functions fj are chosen to be independent of θ, then (1.43) simplifies to B (x; θ) = ∂ πθ (f )(x). (1.45) h ij − θi ∆ j

A useful approximation to the optimal weight matrix can be obtained by ap- plying the expansion of conditional moments given in the following lemma. The expansion is expressed in terms of the of the diffusion, which is 18 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES defined by

d d 1 2 (1.46) Aθf(x)= bk(x; θ)∂xk f(x)+ 2 Ckℓ(x; θ)∂xk xℓ f(x), Xk=1 k,ℓX=1 T i where C = σσ . By Aθ we mean i-fold application of Aθ, and in particu- 0 lar, Aθf = f. For an ergodic diffusion with invariant measure with Lebesgue density µθ, let Φθ be the class of real functions f defined on the state space D that are twice continuously differentiable, square intergrable with respect to µθ, and satisfy that

(A f(x))2µ (x)dx < θ θ ∞ ZD d ∂ f(x)∂ f(x)C (x; θ)µ (x)dx < . xi xj i,j θ ∞ i,j=1 D X Z Lemma 1.3.7 Suppose that the diffusion process (1.11) is ergodic with invari- ant measure µθ, that f is 2(k + 1) times continuously differentiable, and that Ai f Φ , i =0,...,k. Then θ ∈ θ k ti πθ(f)(x)= Ai f(x)+ O(tk+1). (1.47) t i! θ i=0 X Proof: We sketch the proof of (1.47), and consider only d =1 to simplify the exposition. First consider k =1. By Ito’s formula t t f(Xt) = f(X0)+ Aθf(Xs)ds + f ′(Xs)σ(Xs; θ)dWs 0 0 Z s Z s 2 Aθf(Xs) = Aθf(X0)+ Aθf(Xu)du + ∂xAθf(Xu)σ(Xu; θ)dWu, Z0 Z0 and by inserting the expression for Aθf(Xs) given by the second equation into the Lebesgue in the first equation, we find that t s 2 f(Xt)= f(X0)+ tAθf(X0)+ Aθf(Xu)duds (1.48) 0 0 t s Z Z t + ∂xAθf(Xu)σ(Xu; θ)dWuds + f ′(Xs)σ(Xs; θ)dWs. Z0 Z0 Z0 Because Ai f Φ , i = 0, 1, the Ito- are proper P -martingales. θ ∈ θ θ Hence by taking the conditional expectation given X0 = x, we obtain θ 2 πt (f)(x)= f(x)+ tAθf(x)+ O(t ). 2 The result for k = 2 is obtained by applying Ito’s formula to Aθf(Xt), in- serting the result into the first Lebesgue integral in (1.48), and finally taking MARTINGALEESTIMATINGFUNCTIONS 19 the conditional expectation given X0 = x. The result for k 3 is obtained i ≥ 2 similarly by applying Ito’s formula to Aθf(Xt), i =3,...,k. Note that (1.47) is an expansion result, so the corresponding power does not necessarily converge. For a fixed k, the sum is a good approximation to the conditional expectation when t is small. The remainder term depends on k and θ. The explicit sufficient conditions in Lemma 1.3.7 for (1.47) to hold for ergodic diffusions was given in Jacobsen (2001). The expansion holds under mild regularity conditions for non-ergodic diffusions too. In a proof similar to that of Lemma 1.3.7, such conditions must essentially ensure, that Ito-integrals are proper martingales and that the remainder term can be controlled.

θ It is often enough to use the approximation π∆(fj )(x) fj (x) + ∆Aθfj(x). When f does not depend on θ this implies that for d =1≈ 1 2 B (x; θ) ∆ ∂ b(x; θ)f ′(x)+ ∂ σ (x; θ)f ′′(x) (1.49) h ≈ θ 2 θ and for d =1 and N =1   2 2 2 V (x; θ) ∆ A (f )(x) 2f(x)A f(x) = ∆ σ (x; θ)f ′(x) . (1.50) h ≈ θ − θ   We will refer to estimating functions obtained by approximating the optimal weight-matrix a∗ in this way as approximately optimal estimating functions. Use of this approximation will save time and improve the numerical performance of the estimation procedure. The approximation will not affect the consistency of the estimators, and if ∆ is not too large, it will just lead to a relatively minor loss of efficiency. The magnitude of this loss of efficiency can be calculated by means of (1.47).

Example 1.3.8 If we simplify the optimal weight matrix found in Example 1.3.6 by the expansion (1.47) and the Gaussian approximation (1.38), we ob- tain the approximately optimal quadratic martingale estimating function n ∂θb(X(i 1)∆; θ) − (1.51) Gn◦ (θ)= 2 [Xi∆ F (X(i 1)∆; θ)] σ (X(i 1)∆; θ) − − i=1  − X 2 ∂θσ (X(i 1)∆; θ) − 2 + 4 (Xi∆ F (X(i 1)∆; θ)) φ(X(i 1)∆; θ) . 2σ (X(i 1)∆; θ)∆ − − − − −    As in Example 1.3.6 the diffusion is assumed to be one-dimensional. Consider a diffusion with linear drift, b(x; θ)= β(x α). Diffusion models with linear drift and a given marginal distribution− were− studied in Bibby, Skov- 2 gaard & Sørensen (2005). If σ (x; θ)µθ(x)dx < , then the Ito-integral in ∞ t R t X = X β(X α)ds + σ(X ; θ)dW t 0 − s − s s Z0 Z0 20 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES is a propermartingalewith mean zero, so the function f(t)= Eθ(Xt X0 = x) satisfies that | t f(t)= x β f(s)ds + βαt − 0 or Z f ′(t)= βf(t)+ βα, f(0) = x. − Hence βt βt f(t)= xe− + α(1 e− ) − or β∆ β∆ F (x; α, β)= xe− + α(1 e− ) − If only estimates of drift parameters are needed, we can use the linear mar- tingale estimating function of the form (1.30) with N = 1 and h1(x, y; θ) = y F (∆, x; θ). If σ(x; θ) = τκ(x) for τ > 0 and κ a positive function, then the− approximately optimal estimating function of this form is n 1 β∆ β∆ 2 Xi∆ X(i 1)∆e− α(1 e− ) κ (X(i 1)∆) − − − −  i=1 −  X   Gn◦ (α, β)= ,  n  X(i 1)∆  − β∆ β∆   2 Xi∆ X(i 1)∆e− α(1 e− )   κ (X(i 1)∆) − − − −   i=1 −   X    where multiplicative constants have been omitted. To solve the estimating equa- tion Gn◦ (α, β)=0 we introduce the weights n κ 2 2 wi = κ(X(i 1)∆)− / κ(X(j 1)∆)− , − − j=1 X ¯ κ n κ ¯ κ n κ and define X = i=1 wi Xi∆ and X 1 = i=1 wi X(i 1)∆. These two − − quantities are conditional precision weighted sample averages of Xi∆ and P P X(i 1)∆, respectively. The equation Gn◦ (α, β)=0 has a unique explicit so- lution− provided that the weighted sample autocorrelation n κ ¯ κ ¯ κ κ i=1 wi (Xi∆ X )(X(i 1)∆ X 1) r = − − − − n n κ ¯ κ 2 i=1 wi (X(i 1)∆ X 1) P − − − is positive. By the law of largeP numbers (1.4) for ergodic processes, the proba- κ bility that rn > 0 tends to one as n tends to infinity. Specifically, we obtain the explicit estimators ¯ κ κ ¯ κ X rnX 1 αˆ = − − n 1 rκ − n 1 βˆ = log (rκ) , n −∆ n see Christensen & Sørensen (2008). A slightly simpler and asymptotically MARTINGALEESTIMATINGFUNCTIONS 21 ¯ κ ¯ κ equivalent estimator may be obtained by substituting X for X 1 everywhere, in which case α is estimated by the precision weighted sample− average X¯ κ. For the square-root process (CIR-model) given by (1.37), where κ(x)= √x,a simulation study and an investigation of the asymptotic variance of these esti- mators in Bibby & Sørensen (1995) show that they are not much less efficient than the estimators from the optimal estimating function. See also the simu- lation study in Overbeck & Ryd´en (1997), who find that these estimators are surprisingly efficient, a finding that can be explained by the results in Section 1.6. To obtain an explicit approximately optimal quadratic estimating function, we need an expression for the conditional variance φ(x; θ). As we saw in Exam- ple 1.3.6, φ(x; θ) is explicitly known for the square-root process (CIR-model) given by (1.37). For this model the approximatelyoptimal quadratic martingale estimating function is n 1 β∆ β∆ Xi∆ X(i 1)∆e− α(1 e− ) X − − − − i=1 (i 1)∆  X −  n    β∆ β∆   Xi∆ X(i 1)∆e− α(1 e− )   − − − −   i=1  .  X     n   1 β∆ β∆ 2   Xi∆ X(i 1)∆e− α(1 e− )   X(i 1)∆ − − − −   i=1 −   X 2 h    τ 2β∆ β∆   α/2 X(i 1)∆ e− (α X(i 1)∆)e− + α/2   − β − − − − −     This expression is obtained from (1.51) after multiplication by an invertible non-random matrix to obtain a simpler expression. This does not change the estimator. From this estimating function explicit estimators can easily be ob- tained: n βˆn∆ 1 e− αˆn = Xi∆ + (Xn∆ X0), n βˆ ∆ − i=1 n 1 e n X − − essentially the sample mean when nis large, and n n n 1 n Xi∆/X(i 1)∆ ( Xi∆)( X− ) βˆn∆ i=1 i=1 i=1 (i 1)∆ e− = − − − 2 n n 1 P n ( i=1 X(i 1)∆P)( i=1 XP(−i 1)∆) − − − ˆ ˆ 2 n 1 P P βn∆ βn i=1 X(−i 1)∆ Xi∆ X(i 1)∆e− αˆn(1 e− ) τˆ2 = − − − − − , n n 1 ˆ  P i=1 X(−i 1)∆ψ(X(i 1)∆;ˆαn, βn) − − where P 1 2β∆ β∆ 1 ψ(x; α, β)= ( α x)e− (α x)e− + α /β. 2 − − − 2  22 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES It is obviously necessary for this solution to the estimating equation to exist βˆn∆ that the expression for e− is strictly positive, an event that happens with a probability tending to one as n . Again this follows from the law of large numbers (1.4) for ergodic processes.→ ∞ 2

When the optimal weight matrix is approximated by means of (1.47), there is a certain loss of efficiency, which as in the previous example is often quite small; see Bibby & Sørensen (1995) and Section 1.6 on high frequency asymptotics below. Therefore the relatively simple estimating function (1.51) is often a good choice in practice.

θ It is tempting to go on to approximate π∆(fj (θ))(x) in (1.41) by (1.47) in or- der to obtain an explicit estimating function, but as will be demonstrated in Subsection 1.5.3, this can be a dangerous procedure. In general the conditional θ expectation in π∆ should therefore be approximated by . Fortu- nately, Kessler & Paredes (2002) have established that, provided the simulation is done with sufficient accuracy, this does not cause any bias, only a minor loss of efficiency that can be made arbitrarily small; see Subsection 1.3.5. More- θ over, as we shall see in the Subsections 1.3.6 and 1.3.7, π∆(fj (θ))(x) can be found explicitly for a quite flexible class of diffusions.

1.3.4 Small ∆-optimality

The Godambe-Heyde optimal estimating functions discussed above are opti- mal within a certain class of estimating functions. In this subsection we present the concept of small ∆-optimality, introduced and studied by Jacobsen (2001) and Jacobsen (2002). A small ∆-optimal estimating function is optimal among all estimating functions satisfying weak regularity conditions, but only for high sampling frequencies, i.e. when the time between observations is small. Thus the advantage of the concept of small ∆-optimality is that the optimality is global, while the advantage of the concept of Godambe-Heyde optimality is that the optimality holds for all sampling frequencies. Fortunately, we do not have to choose between the two, because it turns out that Godambe-Heyde op- timal martingale estimating functions of the form (1.30) and (1.40) are small ∆-optimal. Small ∆-optimality was originally introduced for general estimating functions for multivariate diffusion models, but to simplify the exposition we will con- centrate on martingale estimating functions and on one-dimensional diffusions of the form dXt = b(Xt; α)dt + σ(Xt; β)dWt, (1.52) where θ = (α, β) Θ IR2. This is the simplest model type for which the essential features∈ of the⊆ theory appear. Note that the drift and the diffusion MARTINGALEESTIMATINGFUNCTIONS 23 coefficient depend on different parameters. It is assumed that the diffusion is ergodic, that its invariant probability measure has density function µθ for all θ Θ, and that X µ under P . Thus the diffusion is stationary. ∈ 0 ∼ θ θ Throughout this subsection, we shall assume that the observation times are equidistant, i.e. ti = i∆,i = 0, 1,...,n, where ∆ is fixed, and that the mar- tingale estimating function (1.12) satisfies the conditions of Theorem 1.3.2, so that we know that (eventually) a Gn-estimator θˆn exists, which is asymptoti- 1 T 1 cally normal with covariance matrix M(g)= W − V W − , where W is given by (1.7) with θ¯ = θ0 and V = V (θ0) with V (θ) given by (1.19). The main idea of small ∆-optimality is to expand the asymptotic covariance matrix in powers of ∆ 1 M(g)= v 1 (g)+ v0 (g)+ o(1). (1.53) ∆ − Small ∆-optimal estimating functions minimize the leading term in (1.53). Ja- cobsen (2001) obtained (1.53) by Ito-Taylor expansions, see Kloeden & Platen (1999), of the random matrices that appear in the expressions for W and V under regularity conditions that will be given below. A similar expansion was used in A¨ıt-Sahalia & Mykland (2003) and A¨ıt-Sahalia & Mykland (2004).

To formulate the conditions, we define the θ, θ Θ. Its domain, Γ, is the set of continuous real-valued functions (s,x,y)A ϕ∈(s,x,y) of s 0 and (x, y) (ℓ, r)2 that are continuously differentiable7→ with respect ≥ ∈ to s and twice continuously differentiable with respect to y. The operator θ is given by A ϕ(s,x,y)= ∂ ϕ(s,x,y)+ A ϕ(s,x,y), (1.54) Aθ s θ where Aθ is the generator (1.46), which for fixed s and x is applied to the function y ϕ(s,x,y). The operator acting on functions in Γ that do 7→ Aθ not depend on x is the generator of the space-time process (t,Xt)t 0. We also ∆ ≥ need the probability measure Qθ given by (1.16). Note that in this section the dependence on ∆ is explicit in the notation.

Condition 1.3.9 The function ϕ belongs to Γ and satisfies that r r 2 s ϕ (s,x,y)Qθ0 (dx, dy) < ℓ ℓ ∞ Z r Z r 2 s ( θ0 ϕ(s,x,y)) Qθ0 (dx, dy) < ℓ ℓ A ∞ Z r Z r (∂ ϕ(s,x,y))2σ2(y; β )Qs (dx, dy) < y 0 θ0 ∞ Zℓ Zℓ for all s 0. ≥

As usual θ0 = (α0,β0) denotes the true parameter value. We will say that a 24 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES k k ℓ function with values in IR or IR × satisfies Condition 1.3.9 if each compo- nent of the functions satisfies this condition. Suppose ϕ satisfies Condition 1.3.9. Then by Ito’s formula t

ϕ(t,X0,Xt)= ϕ(0,X0,X0)+ θ0 ϕ(s,X0,Xs)ds (1.55) 0 A tZ + ∂yϕ(s,X0,Xs)σ(Xs; β0)dWs Z0 under Pθ0 . A significant consequence of Condition 1.3.9 is that the Ito-integral in (1.55) is a true Pθ0 -martingale, and thus has expectation zero under Pθ0 .

If the function θ0 ϕ satisfies Condition 1.3.9, a similar result holds for this functions, whichA we can insert in the Lebesgue integral in (1.55). By doing so and then taking the conditional expectation given X0 = x on both sides of (1.55), we obtain πθ0 (ϕ)(t, x)= ϕ(0, x, x)+ t ϕ(0, x, x)+ O(t2), (1.56) t Aθ0 where πθ(ϕ)(t, x)= E (ϕ(t,X ,X ) X = x) . t θ 0 t | 0 If the functions i , satisfy Condition 1.3.9, where i denotes θ0 ϕ i =0,...,k θ0 i-fold applicationA of the operator , we obtain by similar argumentsA that Aθ0 k si πθ0 (ϕ)(t, x)= i ϕ(0, x, x)+ O(tk+1). (1.57) t i! Aθ0 i=0 X 0 0 Note that θ is the identity operator, θϕ = ϕ. The previously used expansion (1.47) is aA particular case of (1.57). InA the case where ϕ does not depend on x (or y) the integrals in Condition 1.3.9 are with respect to the invariant measure

µθ0 . If, moreover, ϕ does not depend on time s, the conditions do not depend on s.

Theorem 1.3.10 Suppose that the function g(∆,x,y; θ0) in (1.12) is such that T T g, ∂θ g, gg and θ0 g satisfy Condition 1.3.9. Assume, moreover, that we have the expansionA

g(∆,x,y; θ0)= g(∆,x,y; θ0) + ∆∂∆g(0,x,y; θ0)+ oθ0,x,y(∆). If the matrix r

S = Bθ0 (x)µθ0 (x)dx (1.58) Zℓ is invertible, where

Bθ(x)= (1.59) 1 2 2 ∂αb(x; α)∂yg1(0, x, x; θ) 2 ∂βσ (x; β)∂y g1(0, x, x; θ) ,  1 2 2  ∂αb(x; α)∂yg2(0, x, x; θ) 2 ∂βσ (x; β)∂y g2(0, x, x; θ)   MARTINGALEESTIMATINGFUNCTIONS 25 then (1.53) holds with 1 r 2 2 − ℓ (∂αb(x; α0)) /σ (x; β0)µθ0 (x)dx 0 v 1 (g) . (1.60) − ≥  R   0 0   There is equality in (1.60) if  2 ∂yg1(0, x, x; θ0) = ∂αb(x; α0)/σ (x; β0), (1.61)

∂yg2(0, x, x; θ0) = 0 (1.62) for all x (ℓ, r). In this case, the second term in (1.53) satisfies that ∈ r 1 2 − v (g) 2 ∂ σ2(x; β ) /σ4(x; β )µ (x)dx 0 22 ≥ β 0 0 θ0 Zℓ  with equality if  2 2 4 ∂y g2(0, x, x; θ0)= ∂βσ (x; β0)/σ (x; β0), (1.63) for all x (ℓ, r). ∈

By ∂ygi(0, x, x; θ) we mean ∂ygi(0,y,x; θ) evaluated at y = x, and similarly for second order partial derivatives. Thus the conditions for small ∆-optimality are (1.61), (1.62) and (1.63). For a proof of Theorem 1.3.10, see Jacobsen (2001). The condition (1.62) ensures that all entries of v 1 (g) involving the − 1 diffusion coefficient parameter, β, are zero. Since v 1(g) is the ∆− -order term in the expansion (1.53) of the asymptotic covariance− matrix, this dramatically decreases the asymptotic variance of the estimator of β when ∆ is small. We refer to the condition (1.62) as Jacobsen’s condition. The reader is reminded of the trivial fact that for any non-singular 2 2 ma- × trix, Mn, the estimating functions MnGn(θ) and Gn(θ) give exactly the same estimator. We call them versions of the same estimating function. The matrix Mn may depend on ∆n. Therefore a given version of an estimating function needs not satisfy (1.61) – (1.63). The point is that a version must exist which satisfies these conditions.

Example 1.3.11 Consider a quadratic martingale estimating function of the form

a1(x, ∆; θ)[y F (∆, x; θ)] g(∆,y,x; θ)= − , (1.64) 2 a2(x, ∆; θ) (y F (∆, x; θ)) φ(∆, x; θ) ! − − where F and φ are given by (1.26) and (1.27). By (1.47), F (∆, x; θ) = x + O(∆) and φ(∆, x; θ)= O(∆), so

a1(x, 0; θ)(y x) g(0,y,x; θ)= − . (1.65) 2 a2(x, 0; θ)(y x) ! − 26 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

Since ∂yg2(0,y,x; θ)=2a2(x, ∆; θ)(y x), the Jacobsen condition (1.62) is satisfied for all quadratic martingale estimating− functions. Using again (1.47),it is not difficult to see that the two other conditions(1.61) and (1.63) are satisfied in three particular cases: the optimal estimating function given in Example 1.3.6 and the approximations (1.28) and (1.51). 2

The following theorem gives conditions ensuring, for given functions f1,..., fN , that a small ∆-optimal estimating function of the form (1.30) and (1.40) exists. This not always the case. We assume that the functions f1( ; θ),..., f ( ; θ) are of full affine rank for all θ, i.e., for any θ Θ, the identity· N · ∈ N aθf (x; θ)+ aθ =0, x (ℓ, r), j j 0 ∈ j=1 X for constants aθ, implies that aθ = aθ = = aθ =0. j 0 1 · · · N Theorem 1.3.12 Suppose that N 2, that the functions fj are twice continu- ously differentiable and satisfies that≥ the matrix 2 ∂xf1(x; θ) ∂xf1(x; θ) D(x)= (1.66) 2 ∂xf2(x; θ) ∂xf2(x; θ) ! is invertible for µθ-almost all x. Moreover, assume that the coefficients b and σ are continuously differentiable with respect to the parameter. Then a spec- ification of the weight matrix a(x; θ) exists such that the estimating function (1.30) satisfies the conditions (1.62), (1.61) and (1.63). When N = 2, these conditions are satisfy for

∂αb(x; α)/v(x; β) c(x; θ) 1 a(x; θ)= D(x)− (1.67)  2  0 ∂βv(x; β)/v(x; β) for any function c(x; θ).  For a proof of Theorem 1.3.12, see Jacobsen (2002). In Section 1.6, we shall see that the Godambe-Heyde optimal choice (1.32) of the weight-matrix in (1.30) gives an estimating function which has a version that satisfies the con- ditions for small ∆-optimality, (1.61) – (1.63). We have focused on one-dimensional diffusions to simplify the exposition. The situation becomes more complicated for multi-dimensional diffusions, as we shall now briefly describe. Details can be found in Jacobsen (2002). For a d- dimensional diffusion, b(x; α) is d-dimensional and v(x; β)= σ(x; β)σ(x; β)T is a d d-matrix. The Jacobsen condition is unchanged(except that ∂yg2(0, x, x; θ0) is now× a d-dimensional vector).The other two conditions for small ∆-optimality are T 1 ∂yg1(0, x, x; θ0)= ∂αb(x; α0) v(x; β0)− MARTINGALEESTIMATINGFUNCTIONS 27 and 2 2 1 vec ∂y g2(0, x, x; θ0) = vec (∂βv(x; β0)) v⊗ (x; β0) − . 2 In the latter equation, vec(M) denotes, for a d d matrix M, the d-dimensional × 2 row vector consisting of the rows of M placed one after the other, and M ⊗ 2 2 is the d d -matrix with (i′, j′), (ij)th entry equal to Mi′iMj′j . Thus if × 2 1 M = ∂βv(x; β) and M • = (v⊗ (x; β))− , then the (i, j)th coordinate of vec(M) M • is Mi j M • . i′j′ ′ ′ (i′j′ ),(i,j) For a d-dimensionalP diffusion process, the conditions analogous to those in Theorem 1.3.12 ensuring the existence of a small ∆-optimal estimating func- tion of the form (1.30)is that N d(d+3)/2, and thatthe N (d+d2)-matrix ≥ × 2 ∂xT f(x; θ) ∂xT f(x; θ) has full rank d(d + 3)/2. 

1.3.5 Simulated martingale estimating functions

The conditional moments that appear in the martingale estimating functions can for most diffusion models not be calculated explicitly. For a versatile class of one-dimensional diffusions, optimal martingale estimating functions can be found explicitly; see Subsections 1.3.6 and 1.3.7. Estimation and inference is dramatically simplified by using a model for which an explicit optimal martin- gale estimating function is available. However, if for some reason a diffusion from this class is not a suitable model, the conditional moments must be deter- mined by simulation.

∆ The conditional moment πθ f(x) = Eθ(f(X∆) X0 = x) can be found straightforwardly. Simply fix θ and simulate numerically| M independent tra- (i) jectories X , i = 1,...,M of Xt : t [0, ∆] with X0 = x. Of course, a cannot be simulated{ continuously∈ for} all t [0, ∆]. In practice ∈ values of Xjδ : j =0,...,K are simulated, where K is a large integer and δ = ∆/K{. By the law of large} numbers,

M 1 π∆f(x) =. f(X(i)). θ M ∆ i=1 X The variance of the error can be estimated in the traditional way, and by the cental limit theorem, the error is approximately normal distributed. This sim- ple approach can be improved by applying variance reduction methods, for ∆ instance methods that take advantage of the fact that πθ f(x) can be approxi- mated by (1.47). Methods for numerical simulation of diffusion models can be found in Chapter xxxx and Kloeden & Platen (1999). The approach just described is sufficient when calculating the conditional ex- 28 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES pectation appearing in (1.40). Note, however, that when using a search algo- rithm to find a solution to the estimating equation, it is important to use the same random numbers (the same seed) when calculating the estimating func- tions for different values of the parameter θ. More care is needed if the opti- mal weight functions are calculated numerically. The problem is that the opti- mal weight matrix typically contains derivatives with respect to θ of functions that must be determined numerically, see e.g. Example 1.3.6. Pedersen (1994) ∆ proposed a procedure for determining ∂θπθ f(x; θ) by simulations based on results in Friedman (1975). However, it is often preferable to use an approx- imation to the optimal weight matrix obtained by using (1.47), possibly sup- plemented by Gaussian approximations, as explained in Subsection 1.3.3. This is not only much simpler, but also avoids potentially serious problems of nu- merical instability, and by results in Section 1.6 the loss of efficiency is often very small. The approach outlined here, where martingale estimating functions are approximated by simulation, is closely related to the simulated method of moments, see Duffie & Singleton (1993) and Clement (1997). One might be worried that when approximating a martingale estimating func- tion by simulation of conditional moments, the resulting estimator might have considerably smaller efficiency or even be inconsistent. The asymptotic proper- ties of the estimators obtained when the conditional moments are approximated by simulation were investigated by Kessler & Paredes (2002), who found that if the simulations are done with sufficient care, there is no need to worry. How- ever, their results also show that care is needed: if the used in the simulation method is too crude, the estimator behaves badly. Kessler & Paredes (2002) considered martingale estimating functions of the general form n

Gn(θ)= f(Xi∆,X(i 1)∆; θ) F (X(i 1)∆; θ) , (1.68) − − − i=1 X   where f is a p-dimensional function, and F (x; θ)= E (f(X , x; θ)) X = x). θ ∆ | 0 As previously, X is the unique solution of the stochastic differential equation (1.11). For simplicity X is assumed to be one-dimensional,but Kessler & Pare- des (2002) point out that similar results hold for multivariate diffusions. Below the dependence of X on the initial value X0 = x and θ is, when needed, em- phasized in the notation by writing X(x, θ). Let Y (δ, θ, x) be an approximationto the solution X(θ, x) , which is calculated at discrete time points with a step size δ that is much smaller than ∆, and which satisfies that Y0(δ, θ, x)= x. A simple example is the Euler scheme

Yiδ = Y(i 1)δ + b(Y(i 1)δ ; θ)δ + σ(Y(i 1)δ ; θ)Zi, Y0 = x, (1.69) − − − where the Z s are independent and Z N(0,δ). i i ∼ MARTINGALEESTIMATINGFUNCTIONS 29 If the conditional expectation F (x; θ) is approximated by the simple method described above, we obtain the following approximationto the estimating func- tion (1.68)

M,δ Gn (θ)= (1.70) n M 1 (j) f(Xi∆,X(i 1)∆; θ) f(Y∆ (δ,θ,X(i 1)∆),X(i 1)∆; θ) ,  − − M − −  i=1 j=1 X X   where Y (j)(δ, θ, x), j =1,...,M, are independent copies of Y (δ, θ, x).

Kessler & Paredes (2002) assume that the approximation scheme Y (δ, θ, x) is of weak order β > 0 in the sense that E (g(X (x, θ), x; θ)) E(g(Y (δ, θ, x), x; θ)) R(x; θ)δβ (1.71) | θ ∆ − ∆ |≤ for all θ Θ, for all x in the state space of X, and for δ sufficiently small. Here R(x∈; θ) is of polynomial growth in x uniformly for θ in compact sets, i.e., for any compact subset K Θ, there exist constants C1, C2 > 0 such that ⊆C2 supθ K R(x; θ) C1(1 + x ) for all x in the state space of the diffusion. The inequality∈ | (1.71)|≤ is assumed| | to hold for any function g(y, x; θ) which is 2(β + 1) times differentiable with respect to x, and satisfies that g and its par- tial derivatives with respect to x up to order 2(β + 1) are of polynomial growth in x uniformly for θ in compact sets. This definition of weak order is stronger than the definition in Kloeden & Platen (1999)in that control of the polynomial order with respect to the initial value x is added, but Kessler & Paredes (2002) point out that theorems in Kloeden & Platen (1999) that give the order of ap- proximation schemes can be modified in a tedious, but straightforward, way to ensure that the schemes satisfy the stronger condition (1.71). In particular, the Euler scheme (1.69) is of weak order one if the coefficients of the stochastic differential equation (1.11) are sufficiently smooth.

Under a number of further regularity conditions, Kessler & Paredes (2002) M,δ ˆM,δ M,δ showed the following results about a Gn -estimator, θn , with Gn given by (1.70). We shall not go into these rather technical conditions. Not surpris- ingly, they include conditions that ensure the eventual existence of a consistent and asymptotically normal Gn-estimator, cf. Theorem 1.3.2. If δ goes to zero sufficiently fast that √nδβ 0 as n , then → → ∞ M,δ 1 √n θˆ θ D N 0, (1 + M − )Σ , n − 0 −→    where Σ denotes the asymptotic covariance matrix of a Gn-estimator, see The- orem 1.3.2. Thus for δ sufficiently small and M sufficiently large, it does not matter much that the conditional moment F (x; θ) has been determined by sim- ulation in (1.70). Moreover, we can control the loss of efficiency by our choice 30 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES β of M. However, when 0 < limn √nδ < , →∞ ∞ M,δ 1 √n θˆ θ D N m(θ ), (1 + M − )Σ , n − 0 −→ 0   and when √nδβ ,  → ∞ β N,δ δ− θˆ θ m(θ ) n − 0 → 0   in probability. Here the p-dimensional vector m(θ0) depends on f and is gen- erally different from zero. Thus it is essential that a sufficiently small value of δ is used. The discretization problems caused by the choice of δ can be avoided by using the exact simulation methods introduced by Beskos & Roberts (2005) and Beskos, Papaspiliopoulos & Roberts (2006), see Chapter xxxx.

1.3.6 Explicit martingale estimating functions

In this section we consider one-dimensional diffusion models for which esti- mation is particularly easy because an explicit martingale estimating function exists. Kessler & Sørensen (1999) proposed estimating functions of the form (1.41) where the functions fj , i = 1,...,N are eigenfunctions for the generator (1.46), i.e. A f (x; θ)= λ (θ)f (x; θ), θ j − j j where the real number λ (θ) 0 is called the eigenvalue corresponding to j ≥ fj(x; θ). Under weak regularity conditions, fj is also an eigenfunction for the θ transition operator πt defined by (1.42), i.e.

θ λj (θ)t πt (fj (θ))(x)= e− fj (x; θ). for all t> 0. Thus the function hj given by (1.40) is explicit. The following result holds for ergodic diffusions. The density of the stationary distribution is, as usual, denoted by µθ.

Theorem 1.3.13 Let φ(x; θ) be an eigenfunction for the generator (1.46) with eigenvalue λ(θ), and suppose that r [∂ φ(x; θ)σ(x; θ)]2µ (dx) < (1.72) x θ ∞ Zℓ for all t> 0. Then θ λ(θ)t πt (φ(θ))(x)= e− φ(x; θ). (1.73) for all t> 0. MARTINGALEESTIMATINGFUNCTIONS 31 λt Proof: Define Yt = e φ(Xt). We suppress θ in the notation. By Ito’s formula

t λs t λs Yt = Y0 + 0 e [Aφ(Xs)+ λφ(Xs)]ds + 0 e φ′(Xs)σ(Xs)dWs = Y + t eλsφ (X )σ(X )dW , 0 R0 ′ s s s R so by (1.72), Y isR a true martingale, which implies (1.73). 2

Note that if σ(x; θ) and ∂xφ(x; θ) are bounded functions of x (ℓ, r), then (1.72) holds. If φ is a polynomial of order k and σ(x) C(1∈ + xm), then ≤ (1.72) holds if the 2(k + m 1)’th moment of the invariant distribution µθ is finite. −

Example 1.3.14 For the square-root model (CIR-model) defined by (1.37) (ν) 2 with α> 0, β > 0, and τ > 0, the eigenfunctions are φi(x)= Li (2βxτ − ) with ν =2αβτ 2 1, where L(ν) is the ith order Laguerre polynomial − − i i i + ν xm L(ν)(x)= ( 1)m , i − i m m! m=0 X  −  and the eigenvalues are iβ : i =0, 1, . It is easily seen by direct calcula- (ν) { ···} tion that Li solves the differential equation

τxf ′′(x) β(x α)f ′(x)+ iβf(x)=0. − − By Theorem 1.3.13, (1.73) holds, so we can calculate all conditional poly- nomial moments, of which the first four were given in Example 1.3.6. Thus all polynomial martingale estimating functions are explicit for the square-root model. 2

Example 1.3.15 The diffusion given as the solution of dX = θ tan(X )dt + dW , (1.74) t − t t is an ergodic diffusion on the interval ( π/2,π/2) provided that θ 1/2, which implies that Condition 1.3.1 is satisfied.− This process was introduced≥ by Kessler & Sørensen (1999), who called it an Ornstein-Uhlenbeck process on ( π/2,π/2) because tan x x near zero. The generalization to other finite− intervals is obvious. The invariant∼ measure has a density proportional to cos(x)2θ. The eigenfunctions are θ φi(x; θ)= Ci (sin(x)), i =1, 2,..., θ where Ci is the Gegenbauer polynomial of order i, and the eigenvalues are θ i(θ + i/2), i =1, 2,.... This follows because the Gegenbauer polynomial Ci 32 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES solves the differential equation (2θ + 1)y i(2θ + i) f ′′(y)+ f ′(y) f(y)=0, y2 1 − y2 1 − − so that φi(x; θ) solves the equation 1 φ′′(x; θ) θ tan(x)φ′ (x; θ)= i(θ + i/2)φ (x; θ). 2 i − i − i

Hence φi is an eigenfunction for the generator of the model with eigenvalue i(θ + i/2). From equation 8.934-2 in Gradshteyn & Ryzhik (1965) it follows that i θ 1+ m θ 1+ i m φi(x; θ)= − − − cos[(2m i)(π/2 x)]. m i m − − m=0 X    −  Condition (1.72) in Theorem 1.3.13 is obviously satisfied because the state space is bounded, so (1.73) holds. The first non-trivial eigenfunction is sin(x) (a constant is omitted) with eigen- value θ +1/2. From the martingale estimating function n ˇ (θ+1/2)∆ Gn(θ)= sin(X(i 1)∆)[sin(Xi∆)) e− sin(X(i 1)∆))], (1.75) − − − i=1 X we obtain the simple estimator for θ n 1 i=1 sin(X(i 1)∆) sin(Xi∆) ˇ − θn = ∆− log n 2 1/2, (1.76) − i=1 sin (X(i 1)∆) − P − ! which is defined when the numeratorP is positive. An asymmetric generalization of (1.74) was proposed in Larsen & Sørensen (2007) as a model of the logarithm of an exchange rate in a target zone. The diffusion solves the equation 1 sin 2 π(Xt m)/z ϕ dXt = ρ − − dt + σdWt, − cos 1 π(X m)/z 2 t −  where ρ > 0, ϕ ( 1, 1), σ > 0 z > 0, m IR. The process (1.74) is obtained is when ∈ϕ =− 0, m = 0, z = π/2, and∈σ = 1. The state space is 1 2 2 (m z,m + z), and the process is ergodic if ρ 2 σ and 1+ σ /(2ρ) ϕ −1 σ2/(2ρ). The eigenfunctions are ≥ − ≤ ≤ − 2 1 2 1 (ρ(1 ϕ)σ− , ρ(1+ϕ)σ− ) 1 φ (x; ρ,ϕ,σ,m,z)= P − − 2 − 2 sin( πx/z m) , i i 2 − 1 2 (a,b) with eigenvalues λi(ρ, ϕ, σ) = i ρ + 2 nσ , i = 1, 2,.... Here (x) denotes the Jacobi polynomial of order i.  2 MARTINGALEESTIMATINGFUNCTIONS 33 For most diffusion models where explicit expressions for eigenfunctions can be found, including the examples above, the eigenfunctions are of the form

i j φi(y; θ)= ai,j (θ) κ(y) (1.77) j=0 X where κ is a real function defined on the state space and is independent of θ. For martingale estimating functions based on eigenfunctions of this form, the optimal weight matrix (1.32) can be found explicitly too.

Theorem 1.3.16 Suppose 2N eigenfunctions are of the form (1.77) for i = 1,..., 2N, where the coefficients ai,j (θ) are differentiable with respect to θ. If a martingale estimating functions is defined by (1.40) using the first N eigen- functions, then

j λj (θ)∆ B (x, θ) = ∂ a (θ)ν (x; θ) ∂ [e− φ (x; θ)] (1.78) h ij θi j,k k − θi j kX=0   and

Vh(x, θ)i,j = (1.79) i j [λi(θ)+λj (θ)]∆ a (θ)a (θ)ν (x; θ) e− φ (x; θ)φ (x; θ) , i,r j,s r+s − i j r=0 s=0 X X   θ i where νi(x; θ) = π∆(κ )(x), i = 1,..., 2N, solve the following triangular system of linear equations

i λi(θ)∆ e− φi(x; θ)= ai,j (θ)νj (x; θ) i =1,..., 2N, (1.80) j=0 X with ν0(x; θ)=1.

Proof: The expressions for Bh and Vh follow from (1.43) and (1.44) when the eigenfunctions are of the form (1.77), and (1.80) follows by applying the θ transition operator π∆ to both sides of (1.77). 2

Example 1.3.17 Consider again the diffusion (1.74) in Example 1.3.15. We will find the optimal martingale estimating function based on the first non- trivial eigenfunction, sin(x) (where we have omitted a non-essential multi- plicative function of θ) with eigenvalue θ +1/2. It follows from (1.45) that (θ+1/2)∆ Bh(x; θ) = ∆e− sin(x) because sin(x) does not depend on θ. To find Vh we need Theorem 1.3.16. 34 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES The second non-trivial eigenfunction is 2(θ + 1) sin2(x) 1 with eigenvalue 2(θ + 1), so −

2(θ+1)∆ 2 1 1 1 1 ν (x; θ)= e− [sin (x) (θ + 1)− ]+ (θ + 1)− . 2 − 2 2 Hence the optimal estimating function is

n (θ+ 1 )∆ sin(X(i 1)∆)[sin(Xi∆) e− 2 sin(X(i 1)∆)] G∗ (θ)= − − − n 1 (e2(θ+1)∆ 1)/(θ + 1) (e∆ 1) sin2(X ) i=1 2 (i 1)∆ X − − − − where a constant has been omitted. When ∆ is small, it is a good idea to mul- tiply Gn∗ (θ) by ∆ because the denominator is then of order ∆. Note that when ∆ is sufficiently small, we can expand the exponential func- tion in the numerator to obtain (after multiplication by ∆) the approximately optimal estimating function

n (θ+ 1 )∆ sin(X(i 1)∆)[sin(Xi∆) e− 2 sin(X(i 1)∆)] G˜ (θ)= − − − , n cos2(X ) i=1 (i 1)∆ X − which has the explicit solution n i=1 tan(X(i 1)∆) sin(Xi∆))/ cos(X(i 1)∆) 1 ˜ 1 − − θn = ∆− log n 2 . − i=1 tan (X(i 1)∆) − 2 P −  The explicit estimator θ˜ can, forP instance, be used as a starting value when finding the optimal estimator by solving Gn∗ (θ)=0 numerically. Note, how- ˜ ever, that for Gn the square integrability (1.17) under Qθ0 required in Theorem 1.3.2 (to ensure the central limit theorem) is only satisfied when θ0 > 1.5. This 2 problem can be avoided by replacing cos (X(i 1)∆) in the numerator by 1, to which it is close when the process is not near− the boundaries. In that way we arrive at the simple estimating function (1.75), which is thus approximately optimal too. 2

1.3.7 Pearson diffusions

A widely applicable class of diffusion models for which explicit polynomial eigenfunctionsare available is the class of Pearson diffusions, see Wong (1964) and Forman & Sørensen (2008). A Pearson diffusion is a stationary solution to a stochastic differential equation of the form

dX = β(X α)dt + 2β(aX2 + bX + c)dW , (1.81) t − t − t t t where β > 0, and a, b and c are suchq that the square root is well defined when Xt is in the state space. The parameter β > 0 is a scaling of time that deter- MARTINGALEESTIMATINGFUNCTIONS 35 mines how fast the diffusion moves. The parameters α, a, b, and c determine the state space of the diffusion as well as the shape of the invariant distribu- tion. In particular, α is the expectation of the invariant distribution. We define θ = (α,β,a,b,c). In the context of martingale estimating functions, an important property of the Pearson diffusions is that the generator (1.46) maps polynomials into polyno- mials. It is therefore easy to find eigenfunctions among the polynomials n j pn(x)= pn,j x . j=0 X Specifically, the polynomial pn(x) is an eigenfunction if an eigenvalue λn > 0 exist satisfying that 2 β(ax + bx + c)p′′ (x) β(x α)p′ (x)= λ p (x), n − − n − n n or n n 1 n 2 − − λ a p xj + b p xj + c p xj =0. { n − j } n,j j+1 n,j+1 j+2 n,j+2 j=0 j=0 j=0 X X X where a = j 1 (j 1)a β, b = j α + (j 1)b β, and c = j(j 1)cβ j { − − } j { − } j − for j = 0, 1, 2,.... Without loss of generality, we assume pn,n = 1. Thus by equating the coefficients, we find that the eigenvalue is given by λ = a = n 1 (n 1)a β. (1.82) n n { − − } If we define pn,n+1 =0, then the coefficients pn,j j=0,...,n 1 solve the of equations { } − (a a )p = b p + c p . (1.83) j − n n,j j+1 n,j+1 j+2 n,j+2 Equation (1.83) is equivalent to a simple recursive formula if a a =0 for n − j 6 all j = 0, 1,...,n 1. Note that an aj = 0 if and only if there exists an − − 1 integer n 1 m < 2n 1 such that a = m− and j = m n +1. In − ≤ − 1 − particular, an aj = 0 cannot occur if a < (2n 1)− . It is important to notice that λ −is positive if and only if a< (n 1) −1. We shall see below that n − − this is exactly the condition ensuring that pn(x) is integrable with respect to 1 the invariant distribution. If the stronger condition a< (2n 1)− is satisfied, the first n eigenfunctions belong to the space of functions that− are square inte- grable with respect to the invariant distribution, and they are orthogonal with respect to the usual inner product in this space. The space of functions that are square integrable with respect to the invariant distribution (or a subset of this space) is often taken as the domain of the generator. Obviously, the eigen- function pn(x) satisfies the condition (1.72) if pn(x) is square integrable with respect to the invariant distribution, i.e. if a< (2n 1) 1. By Theorem 1.3.13 − − this implies that the transition operator satisfies (1.73), so that pn(x) can be 36 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES used to construct explicit optimal martingale estimating functions as explained 1 in Subsection 1.3.6. For Pearson diffusions with a 0, a< (2n 1)− is au- tomatically satisfied, so there are infinitely many polynomi≤ al eigenfunctions.− In these cases the eigenfunctions are well-known families of orthogonal poly- nomials. When a > 0, there are only finitely many square integrable polyno- mial eigenfunctions. In these cases more complicated eigenfunctions defined in terms of exist too, see Wong (1964). It is of some historical interest that Hildebrandt (1931) derived the polynomials above from the view- point of Gram-Charlier expansions associated with the Pearson system. Some special cases had previously been derived by Romanovsky (1924). From a modeling point of view, it is important that the class of stationary dis- tributions equals the full Pearson system of distributions. Thus a very wide spectrum of marginal distributions is available ranging from distributions with compact support to very heavy-tailed distributions with tails of the Pareto-type. To see that the invariant distributions belong to the Pearson system, note that the scale measure has density x u α s(x) = exp − du , au2 + bu + c Zx0  2 where x0 is a point such that ax0 + bx0 + c> 0, cf. (1.14). Since the density of the invariant probability measure is given by 1 µ (x) , θ ∝ s(x)(ax2 + bx + c) cf. (1.15), it follows that (2a + 1)x α + b µ′ (x)= − µ (x). θ − ax2 + bx + c θ The Pearson system is defined as the class of probability densities obtained by solving a differential equation of this form, see Pearson (1895). In the following we present a full classification of the ergodic Pearson diffu- sions, which shows that all distributions in the Pearson system can be obtained as invariant distributions for a model in the class of Pearson diffusions. We consider six cases according to whether the squared diffusion coefficient is constant, linear, a convex parabola with either zero, one or two roots, or a con- cave parabola with two roots. The classification problem can be reduced by first noting that the Pearson class of diffusions is closed under location and scale- transformations. To be specific, if X is an ergodic Pearson diffusion, then so is X˜ where X˜t = γXt + δ. The parameters of the stochastic differential equa- tion (1.81) for X˜ are a˜ = a, ˜b = bγ 2aδ, c˜ = cγ2 bγδ + aδ2, β˜ = β, and α˜ = γα + δ. Hence, up to transformations− of location− and scale, the er- godic Pearson diffusions can take the following forms. Note that we consider scale transformations in a general sense where multiplication by a negative real MARTINGALEESTIMATINGFUNCTIONS 37 number is allowed, so to each case of a diffusion with state space (0, ) there corresponds a diffusion with state space ( , 0). ∞ −∞ Case 1: σ2(x)=2β. The solution to (1.81) is an Ornstein-Uhlenbeck process. The state space is IR, and the invariant distribution is the normal distribution with mean α and variance 1. The eigenfunctions are the . Case 2: σ2(x)=2βx. The solution to (1.81) is the square root process (CIR process) (1.37) with state space (0, ). Condition 1.3.1 that ensures ergodicity is satisfied if and only if α> 1. If 0∞< α 1, the boundary 0 can with positive probability be reached at a finite time point,≤ but if the boundaryis made instan- taneously reflecting, we obtain a stationary process. The invariant distribution is the gamma distribution with scale parameter 1 and shape parameter α. The eigenfunctions are the Laguerre polynomials. Case 3: a > 0 and σ2(x)=2βa(x2 + 1). The state space is the real line, 1 2 2a α 1 and the scale density is given by s(x) = (x + 1) exp( a tan− x). By Condition 1.3.1, the solution is ergodic for all a > 0 and− all α IR. The 1 2 2a 1 α 1 ∈ invariant density is given by µθ(x) (x + 1)− − exp( a tan− x) If α =0 ∝ 1 the invariant distribution is a scaled t-distribution with ν =1+ a− degrees of 1 freedom and scale parameter ν− 2 . If α = 0 the invariant distribution is skew 6 1 and has tails decayingat the same rate as the t-distribution with 1+a− degrees of freedom. A fitting name for this distribution is the skew t-distribution. It is also known as Pearson’s type IV distribution. In either case the mean is α and 1 the invariant distribution has moments of order k for k< 1+ a− . Because of its skew and heavy tailed marginal distribution, the class of diffusions with α = 0 is potentially very useful in many applications, e.g. finance. It was studied6 and fitted to financial data by Nagahara (1996) using the local method of Ozaki (1985). We consider this process in more detail below. Case 4: a > 0 and σ2(x)=2βax2. The state space is (0, ) and the scale 1 ∞ a α density is s(x)= x exp( ax ). Condition 1.3.1 holds if and only if α> 0. The 1 a 2 α invariant distribution is given by µθ(x) x− − exp( ax ), and is thus an in- ∝ 1− a verse gamma distribution with shape parameter 1+ a and scale parameter α . 1 The invariant distribution has moments of order k for k < 1+ a− . This pro- cess is sometimes referred to as the GARCH diffusion model. The polynomial eigenfunctions are known as the Bessel polynomials. Case 5: a > 0 and σ2(x)=2βax(x + 1).The state space is (0, ) and α+1 α ∞ the scale density is s(x) = (1+ x) a x− a . The ergodicity Condition 1.3.1 α holds if and only if a 1. Hence for all a > 0 and all µ a, a unique ergodic solution to (1.81)≥ exists. If 0 <α< 1, the boundary 0 can≥ be reached at a finite time point with positive probability, but if the boundary is made instantaneously reflecting, a stationary process is obtained. The density of the α+1 1 α 1 invariant distribution is given by µ (x) (1+x) a x a . Thisis a scaled θ ∝ − − − 38 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES 2α 2 α F-distribution with a and a +2 degrees of freedom and scale parameter 1+a . 1 The invariant distribution has moments of order k for k< 1+ a− . Case 6: a < 0 and σ2(x)=2βax(x 1). The state space is (0, ) and 1 α −α ∞ the scale density is s(x) = (1 x) −a x a . Condition 1.3.1 holds if and only α 1 α − if a 1 and −a 1. Hence for all a < 0 and all α > 0 such that min(α,≤1 − α) a, a≤ uniqueergodicsolution − to (1.81)exists. If 0 <α< a, the boundary− 0≥− can be reached at a finite time point with positive probability,− but if the boundary is made instantaneously reflecting, a stationary process is obtained. Similar remarks apply to the boundary 1 when 0 < 1 α < a. 1 α α − 1 − 1 − The invariant distribution is given by µθ(x) (1 x)− a − x− a − and is ∝ α− 1 α thus the Beta distribution with shape parameters a and −a . This class of diffusions will be discussed in more detail below. It− is often− referred to as the Jacobi diffusions because the related eigenfunctions are Jacobi polynomials. Multivariate Jacobi diffusions were considered by Gourieroux& Jasiak (2006).

Example 1.3.18 The skew t-distribution with mean zero, ν degrees of free- dom, and skewness parameter ρ has (unnormalized) density f(z) ∝ 2 (ν+1)/2 1 (z/√ν + ρ) +1 − exp ρ(ν 1) tan− z/√ν + ρ , { } − which is the invariant density of the diffusion Zt = √ν(X t ρ) with ν = 1 − 1+a− and ρ = α, where X is as in Case 3. An expression for the normalizing constant when ν is integer valued was derived in Nagahara (1996). By the transformation result above, the corresponding stochastic differential equation is

2 1 dZ = βZ dt + 2β(ν 1) 1 Z +2ρν 2 Z +(1+ ρ2)ν dW . (1.84) t − t − − { t t } t For ρ = 0 the invariantq distribution is the t-distribution with ν degrees of freedom.

1 The skew t-diffusion (1.84) has the eigenvalues λn = n(ν n)(ν 1)− β for n < ν. The four first eigenfunctions are − − p1(z) = z, 1 2 4ρν 2 (1 + ρ )ν p (z) = z2 z , 2 − ν 3 − ν 2 − 1 −2 2 2 3 12ρν 2 24ρ ν + 3(1+ρ )ν(ν 5) 8ρ(1+ρ )ν 2 p (z) = z3 z2 + − z + , 3 − ν 5 (ν 5)(ν 4) (ν 5)(ν 3) − − − − − and 1 2 2 24ρν 2 144ρ ν 6(1 + ρ )ν(ν 7) p (z) = z4 z3 + − − z2 4 − ν 7 (ν 7)(ν 6) − − − MARTINGALEESTIMATINGFUNCTIONS 39 2 3 2 3 3 3 8ρ(1 + ρ )ν 2 (ν 7)+48ρ(1 + ρ )ν 2 (ν 6) 192ρ ν 2 + − − − z (ν 7)(ν 6)(ν 5) − − − 3(1 + ρ2)2ν(ν 7) 72ρ2(1 + ρ2)ν2 + − − , (ν 7)(ν 6)(ν 4) − − − provided that ν > 4. If ν > 2i, the first i eigenfunctions are square integrable and thus satisfy (1.72). Hence (1.73) holds, and the eigenfunctions can be used to construct explicit martingale estimating functions. 2

Example 1.3.19 The model dX = β[X (m + γz)]dt + σ z2 (X m)2dW , (1.85) t − t − − t − t where β > 0 and γ ( 1, 1), has beenp proposed as a model for the random variation of the logarithm∈ − of an exchange rate in a target zone between realign- ments by De Jong, Drost & Werker (2001) (γ = 0) and Larsen & Sørensen (2007). This is a diffusion on the interval (m z,m + z) with mean reversion around m + γz. It is a Jacobi diffusion obtained− by a location-scale trans- formation of the diffusion in Case 6 above. The parameter γ quantifies the asymmetry of the model. When β(1 γ) σ2 and β(1 + γ) σ2, X is an ergodic diffusion, for which the stationary− distribution≥ is a Beta-distribution≥ on 2 2 (m z,m + z) with parameters κ1 = β(1 γ)σ− and κ2 = β(1 + γ)σ− . If the− parameter restrictions are not satisfied,− one or both of the boundaries can be hit in finite time, but if the boundaries are made instantaneously reflecting, a stationary process is obtained.

The eigenfunctions for the generator of the diffusion (1.85) are

(κ1 1,κ2 1) φ (x; β,γ,σ,m,z)= P − − ((x m)/z), i =1, 2,... i i − (a,b) where Pi (x) denotes the Jacobi polynomial of order i given by

i (a,b) j n + a a + b + n + j j P (x)= 2− (x 1) , 1

Explicit formulae for the conditional moments of a Pearson diffusion can be obtained from the eigenfunctions by means of (1.73). Specifically,

n n n λℓt k E(Xt X0 = x)= qn,k,ℓe− x , (1.86) | ! kX=0 Xℓ=0 40 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES where q = p , q =0 for ℓ n 1, and n,k,n n,k n,n,ℓ ≤ − n 1 − q = p q n,k,ℓ − n,j j,k,ℓ j=k ℓ X∨ for k,ℓ =0,...,n 1 with λℓ and pn,j given by (1.82) and (1.83). For details see Forman & Sørensen− (2008). Also the moments of the Pearson diffusions can, when they exist, be found explicitly by using the fact that the integral of the eigenfunctionswith respect to κ the invariant probability measure is zero.We have seen above that E( Xt ) < 1 | | if and only if a < (κ 1)− . Thus if a 0 all moments exist, while for ∞ − ≤ 1 a > 0 only the moments satisfying that κ

n 1 n 1 n 2 E(X )= a− b E(X − )+ c E(X − ) , n =2, 3,..., (1.87) t n { n · t n · t } where an = n 1 (n 1)a β, bn = n α+(n 1)b β, and cn = n(n 1)cβ. { − − } { 0 − } − The initial conditions are given by E(Xt )=1, and E(Xt) = α. This can be found from the expressions for the eigenfunctions, but is more easily seen as follows. By Ito’s formula

n n 1 n 2 2 dXt = βnXt − (Xt µ)dt + βn(n 1)Xt − (aXt + bXt + c)dt − − − n 1 +nXt − σ(Xt)dWt, 2n 1 and if E(Xt ) is finite, i.e. if a < (2n 1)− , the last term is a martingale with expectation zero. −

Example 1.3.20 Equation (1.87) allows us to find the moments of the skewed t-distribution, in spite of the fact that the normalizing constant of the density is unknown. In particular, for the diffusion (1.84),

E(Zt) = 0, (1 + ρ2)ν E(Z2) = , t ν 2 − 2 3 4ρ(1 + ρ )ν 2 E(Z3) = , t (ν 3)(ν 2) − − 24ρ2(1 + ρ2)ν2 + 3(ν 3)(1 + ρ2)2ν2 E(Z4) = − . t (ν 4)(ν 3)(ν 2) − − − 2

For a diffusion T (X) obtained from a solution X to (1.81) by a twice differen- tiable and invertible transformation T , the eigenfunctions of the generator are MARTINGALEESTIMATINGFUNCTIONS 41 1 pn T − (x) , where pn is an eigenfunction of the generator of X. The eigen- values{ are the} same as for the original eigenfunctions. Since the original eigen- functions are polynomials, the eigenfunctions of T (X) are of the form (1.77) 1 with κ = T − . Hence explicit optimal martingale estimating functions are also available for transformations of Pearson diffusions, which is a very large and flexible class of diffusion processes. Their stochastic differential equations can, of course, be found by Ito’s formula.

Example 1.3.21 For the Jacobi-diffusion (case 6) with µ = a = 1 , i.e. − 2 dX = β(X 1 )dt + βX (1 X )dW t − t − 2 t − t t the invariant distribution is the uniform distributionp on (0, 1) for all β > 0. For any strictly increasing and twice differentiable distribution function F , we 1 therefore have a class of diffusions given by Yt = F − (Xt) or 1 2 1 (F (Yt) 2 )f(Yt) + 2 F (Yt) 1 F (Yt) dYt = β − 3 { − }dt − f(Yt)

βF (Yt) 1 F (Yt) + { − }dWt, f(Yt) which has invariant distribution with density f = F ′. A particular example is the logistic distribution ex F (x)= , x IR, 1+ ex ∈ for which dY = β sinh(x) +8cosh4(x/2) dt +2 β cosh(x/2)dW . t − t 1 If the same transformation F − (y) = log(y/ (1 yp)) is applied to the general Jacoby diffusion (case 6), then we obtain −

Yt 1 4 dY = β 1 2µ + (1 µ)e µe− 8a cosh (Y /2) dt t − − − − − t  +2 aβ cosh(Y t/2)dWt, − a diffusion for which the invariant distribution is thep generalized logistic dis- tribution with density eκ1x f(x)= x κ +κ , x IR, (1 + e ) 1 2 B(κ1,κ2) ∈ where κ1 = (1 α)/a, κ2 = α/a and B denotes the Beta-function. This distribution was− introduced− and studied in Barndorff-Nielsen,Kent & Sørensen (1982).

2 42 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES Example 1.3.22 Let again X be a general Jacobi-diffusion (case 6). If we ap- 1 ply the transformation T (x) = sin− (2x 1) to X , we obtain the diffusion − t sin(Yt) ϕ dYt = ρ − dt + aβ/2dWt, − cos(Yt) − p where ρ = β(1 + a/4) and ϕ = (2α 1)/(1 + a/4). The state space is ( π/2,π/2). Note that Y has dynamics− that are very different from those of the− Jacobi diffusion: the drift is highly non-linear and the diffusion coefficient is constant. This model was considered in Example 1.3.15.

2

1.3.8 Implementation of martingale estimating functions

An R-package, where a number of methods for calculating estimators for dif- fusion models are implemented, has been developed by Stefano Iacus and is described in the book Iacus (2008), which outlines the underlying theory too. The R-package also contains implementations of methods for simulating so- lutions to stochastic differential equations. It is, however, useful to notice that for many martingales estimating functions the estimators, or asymptotically equivalent estimators, can be calculated by means of standard statistical soft- ware packages. Specifically, they can be calculated as weighted least squares estimators for non-linear regression models. To see this, consider the weighted least squares estimator obtained by mini- mizing

Cn(θ)= (1.88) n θ T 1 θ f(Xti ) π∆(f)(Xti 1 ) Vi− f(Xti ) π∆(f)(Xti 1 ) , − − − − i=1 X     with f(x) = (f1(x),...,fN (x)) and ˜ Vi = Vh(Xti 1 ; θn), (1.89) − where Vh is the N N-matrix given by (1.44), and θ˜n is a consistent estimator θ × of θ. As usual, π∆ denotes the transition operator (1.42). The consistent esti- mator can, for instance, be the non-weighted least squares estimator obtained by minimizing (1.88) with V = I , where I is the N N identity matrix. i N N × The weighted least squares estimator obtained from (1.88) with Vi given by (1.89) solves the estimating equation n ˜ 1 θ Bh(Xti 1 ; θ)Vh(Xti 1 ; θn)− f(Xti ) π∆(f)(Xti 1 ) =0 (1.90) − − − − i=1 X   MARTINGALEESTIMATINGFUNCTIONS 43 with Bh given by (1.45). Therefore this estimator has the same asymptotic variance as the optimal Gn∗ -estimator with h given by (1.40); see e.g. Jacod & Sørensen (2009). The estimating function (1.90) is similar in spirit to (1.39). The estimators obtained by minimizing (1.88) is the weighted least squares estimator for a regression model for the data f(Xti ), i =1,...,n with Xti 1 , i = 1,...,n, as explanatory regression variables, the non-linear regression− θ function π∆(f)(Xti 1 ), and the weight matrix Vi. In some particularly nice cases, the regression− function is linear in the parameters, and the estimator is a linear regression estimator.

Example 1.3.23 Let X be the square root process (1.37),and suppose we have 2 the observations Xi∆, i =0,...,n. Let us think of (Xi∆,Xi∆), i =1,...,n, as data with explanatory regression variables X(i 1)∆, i = 1,...,n, and with the non-linear regression function −

θ F (X(i 1)∆; θ) π∆(f)(X(i 1)∆)= − 2 , − φ(X(i 1)∆; θ)+ F (X(i 1)∆; θ)  − −  where F and φ are as in Example 1.3.6. Then we obtain a weighted least 2 squares estimator for θ, by minimizing (1.88) with f1(x) = x, f2(x) = x , and

Vh(x; θ)=

φ(x; θ) η(x; θ)+2F (x; θ)φ(x; θ)2 , η(x; θ)+2F (x; θ)φ(x; θ)2 ψ(x; θ)+4F (x; θ)2φ(x; θ)+4F (x; θ)η(x; θ)   where η and ψ are as in Example 1.3.6. This estimator has the same efficiency as the estimator obtained from the opti- mal martingale estimating function of form (1.30) with N =2 and h (x, y; θ) = y F (x; θ) 1 − h (x, y; θ) = y2 φ(x; θ) F (x; θ)2. 2 − − The optimal estimating function of this form is equivalent to the optimal es- timating function in Example 1.3.6. For the square root process some simpli- fication can be achieved by using the Gaussian approximation (1.38) in the definition of the matrix Vh.

2

Example 1.3.24 Consider the process (1.74), and suppose we have the obser- vations Xi∆, i =0,...,n. Let us think of sin(Xi∆), i =1,...,n, as data with explanatory regression variables X(i 1)∆, i =1,...,n and with the non-linear θ − (θ+1/2)∆ regression function π∆(sin)(X(i 1)∆)= e− sin(X(i 1)∆). Again we − − 44 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES can obtain a weighted least squares estimator for θ, by minimizing (1.88) with f(x) = sin(x) and ˜ 1 2(θn+1)∆ ˜ ∆ 2 Vh(x; θ)= 2 (e 1)/(θn + 1) (e 1) sin (X(i 1)∆), − − − − where θ˜n is a consistent estimator, for instance the simple estimator (1.76). Note that the non-linear regression is, in fact, a linear regression in the param- θ∆ eter ξ = e− . The regression estimator equals the estimator obtained from the estimating function

n (θ+ 1 )∆ sin(X(i 1)∆)[sin(Xi∆) e− 2 sin(X(i 1)∆)] − − Gn• (θ)= ˜ − , 1 2(θn+1)∆ ˜ ∆ 2 i=1 2 (e 1)/(θn + 1) (e 1) sin (X(i 1)∆) X − − − − which has the same efficiency as the optimal estimator obtained in Exam- ple 1.3.17. If instead we minimize (1.88) with the approximation Vh(x; θ) = 2 cos (x), then we obtain the estimator θ˜n from Example 1.3.17, and if we min- imize (1.88) with the more crude approximation Vh(x; θ)=1, then we obtain the simple estimator (1.76) from Example 1.3.15.

2

More generally, an estimator with the same efficiency as the optimal estimator from (1.30) with optimal weights (1.32) is obtained by minimizing the objec- tive function n T 1 h(Xti 1 ,Xti ; θ) Vi− h(Xti 1 ,Xti ; θ) (1.91) − − i=1 X with Vi defined as in (1.89), but here with Vh given by (1.34). This estimator can be found by applying standard software for minimizing objective functions to (1.91).

Example 1.3.25 Let again X be the square root process (1.37), and consider the martingale estimating function of form (1.30) with N =2 and h1 and h2 as in Example 1.3.4. In this case an optimal estimator is obtained by minimizing (1.91) with φ(x; θ) η(x; θ) V (x; θ)= , h η(x; θ) ψ(x; θ)   where φ, η and ψ are as in Example 1.3.6. Here a considerable simplification can be obtained by the Gaussian approximation (1.38). With this approxima- tion φ(x; θ) 0 V (x; θ)= . h 0 2φ(x; θ)2  

2 THE LIKELIHOOD FUNCTION 45 1.4 The likelihood function

The likelihood function for a discretely observed diffusion model, (1.24) is a product of transitions densities. Unfortunately, the transition density of a diffu- sion process is only rarely explicitly known, but several numerical approaches make likelihood inference feasible for diffusion models. Pedersen (1995) proposed a method for obtaining an approximation to the like- lihood function by rather extensive simulation. Pedersen’s method was very considerably improved by Durham & Gallant (2002), whose method is com- putationally much more efficient. Poulsen (1999) obtained an approximation to the transition density by numerically solving a partial differential equation, whereas A¨ıt-Sahalia (2002) and A¨ıt-Sahalia (2008) proposed to approximate the transition density by means of expansions. A Gaussian approximation to the likelihood function obtained by local linearization of (1.11) was proposed by Ozaki (1985), while Forman & Sørensen (2008) proposed to use an approx- imation in terms of eigenfunctions of the generator of the diffusion. Bayesian estimators with the same asymptotic properties as the maximum likelihood es- timator can be obtained by Markov chain Monte Carlo methods, see Elerian, Chib & Shephard (2001), Eraker (2001), and Roberts & Stramer (2001). Fi- nally, exact and computationally efficient likelihood-based estimation methods were presented by Beskos et al. (2006). The latter approach is presented in Chapter XXX. In the following we will outline the expansion approach of A¨ıt- Sahalia (2002) for scalar diffusion models. The various other approaches to calculation of the likelihood function will not be considered further in this chapter. Assume that the diffusion process (1.11) is one-dimensional and that the state space is either ( , ) or (0, ), i.e. r = and ℓ is either or 0. The coefficients b and−∞σ are∞ assumed∞ to satisfy the∞ following condition.−∞

Condition 1.4.1 (i) The functions b(x; θ) and σ(x; θ) are infinitely often differentiable w.r.t. x and three times continuously differentiable w.r.t. θ for all x (ℓ, r) and θ Θ. ∈ ∈ (ii-a) If ℓ = , there exists a constant c > 0 such that σ(x; θ) > c for all x (ℓ, r) and−∞ all θ Θ. ∈ ∈ (ii-b) If ℓ = 0, then σ is non-degenerate on (0, ) in the sense that for each ξ > 0 there exists a constant c > 0 such that σ∞(x; θ) c for all x ξ and ξ ≥ ξ ≥ all θ Θ. Moreover, if limx 0 σ(x; θ)=0, then constants ξ0, ω and ρ exist such that∈ σ(x; θ) ωxρ for all→ x (0, ξ ) and all θ Θ. ≥ ∈ 0 ∈ The idea is to make an expansion of the transition density. However, the distri- bution of X∆ given X0 can be so far from a normal distribution that a conver- gent expansion with the normal density as leading term is not possible. This 46 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES is possible for a diffusion with constant diffusion coefficient. Therefore the standard transformation x 1 h(x; θ)= du, σ(u; θ) Zx∗ where x∗ is arbitrary, is applied to obtain the diffusion process

Yt = h(Xt; θ). Since σ > 0, the transformation h is increasing, and by Ito’s formula

dYt = a(Yt; θ)dt + dWt, (1.92) where 1 b(h− (y; θ); θ) 1 1 a(y; θ)= 1 2 σ′(h− (y; θ); θ) σ(h− (y; θ); θ) − with σ′(x; θ) = ∂xσ(x; θ). The state space of Y , (ℓY , rY ) could in principle depend on θ, but we assume that this is not the case. If only one of the bound- aries ℓY and rY is finite, it can always be arranged that the finite boundary equals zero by choosing x suitably. For instance if r = and ℓ is finite, ∗ Y ∞ Y then we can choose x∗ = ℓ to obtain ℓY =0. We will assume that ℓY is either or 0, and that rY is either 0 or . It is further assumed that a satisfies the −∞following condition (which can be translated∞ into a condition on b and σ).

Condition 1.4.2 (i) For all θ Θ, the drift coefficient a(y; θ) and its derivatives w.r.t. y and θ have at most∈ polynomial growth near the boundaries, and lim[a(y; θ)2 + ∂ a(y; θ)] > as y ℓ and y r . y −∞ ↓ Y ↑ Y (ii-a) If ℓY =0, then there exist constants ǫ0 > 0, κ and α such that a(y; θ) κy α for all y (0,ǫ ) and all θ Θ, where either α > 1 and κ > 0, or≥ − ∈ 0 ∈ α =1 and κ 1. If ℓY = , then there exists constants E0 > 0 and K > 0 such that a(y≥; θ) Ky for−∞ all y E and all θ Θ. ≥ ≤− 0 ∈ (ii-b) If rY =0, then there exist constants ǫ0 > 0, κ and α such that a(y; θ) κ y α for all y ( ǫ , 0) and all θ Θ, where either α> 1 and κ> 0, or≤ − | |− ∈ − 0 ∈ α =1 and κ 1/2. If rY = , then there exist constants E0 > 0 and K > 0 such that a(y;≥θ) Ky for all∞y E and all θ Θ. ≤ ≥ 0 ∈ A real function f is said to be of polynomial growth near a boundary at or if there exist constants C > 0, K > 0 and p> 0 such that f(x) C∞x p for−∞x>K or x < K. If the boundary is at zero, polynomial| growth|≤ means| | − p that there exist constants C > 0, ǫ > 0 and p > 0 such that f(x) C x − for x ǫ. | | ≤ | | | |≤ Under the assumptions imposed, a solution exists to (1.92) with a transition density that is sufficiently regular for likelihood inference. This is the contents of the following proposition from A¨ıt-Sahalia (2002). THE LIKELIHOOD FUNCTION 47 Proposition 1.4.3 Under the Conditions 1.4.1 and 1.4.2, the stochastic differ- ential equation (1.92) has a unique weak solution for every initial distribu- tion. The boundaries are unattainable. The solution Y has a transition density pY (∆,y0,y; θ) that is continuously differentiable w.r.t. ∆, infinitely often dif- ferentiable w.r.t. y (ℓY , rY ), and three times continuously differentiable w.r.t. θ Θ. ∈ ∈ This result implies that the original stochastic differential equation (1.11) has a unique weak solution, and by the transformation theorem for density functions, it has a similarly regular transition density given by

p(∆, x0, x; θ)= pY (∆,h(x0; θ),h(x; θ); θ)/σ(x; θ). (1.93) Instead of expanding the transition density of Y , i.e. the conditional density function of Y∆ given Y0 = y0, we expand the conditional density of the nor- malized increment 1 Z = ∆− (Y y ) ∆ − 0 given Y0 = y0. This is because pY gets peaked around y0 as ∆ gets close to zero, whereas the distribution of Z is sufficiently close to the N(0, 1)- distribution to make it the appropriate transformation of X∆ to obtain a con- vergent expansion of the conditional density function with the standard normal density function as the leading term. Obviously, 1/2 1/2 p (∆,y ,y; θ) = ∆− p (∆, ∆− (y y ) y ; θ), (1.94) Y 0 Z − 0 | 0 where p (∆,z y ; θ) is the conditional density of Z given that Y = y . Z | 0 0 0 We can now obtain an approximation to the transition density of X, and hence an approximation to the likelihood function, by expanding the conditional den- sity, pZ , of Z given Y0 = y0 in terms of Hermite polynomials up to order J: J pJ (∆,z y ; θ)= ϕ(z) η (∆,y ; θ)H (z), (1.95) Z | 0 j 0 j j=0 X where ϕ denotes the density function of the standard normal distribution, and Hj is the jth Hermite polynomial, which is defined by j j x2/2 d x2/2 H (x) = ( 1) e e− , j =0, 1,.... j − dxj The Hermite polynomials up to order 4 are

H0(x) = 1

H1(x) = x H (x) = x2 1 2 − H (x) = x3 3x 3 − H (x) = x4 6x2 +3. 4 − 48 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

The coefficients ηj (∆,y0; θ) can be found by using that the Hermite polyno- mials are orthogonal in the space L2(ϕ):

∞ 0 if i = j H (x)H (x)ϕ(x)dx = i j i! if i =6 j. Z−∞  Hence if ∞ p (∆,z y ; θ)= ϕ(z) η (∆,y ; θ)H (z), Z | 0 j 0 j j=0 X it follows that

∞ ∞ ∞ H (z)p (∆,z y ; θ)dz = η (∆,y ; θ) H (z)H (z)ϕ(z)dz i Z | 0 j 0 i j j=0 Z−∞ X Z−∞ = i! ηi(∆,y0; θ).

By inserting the expansion (1.95) in (1.94) and (1.93), we obtain the following approximations to the transitions densities pY and p

J J 1/2 1/2 1/2 p (∆,y ,y; θ) = ∆− ϕ(∆− (y y )) η (∆,y ; θ)H (∆− (y y )) Y 0 − 0 j 0 j − 0 j=0 X (1.96) and

J p (∆, x0, x; θ)= (1.97)

h(x;θ) h(x0;θ) ϕ − J √∆ h(x; θ) h(x ; θ) η (∆,h(x ; θ); θ)H − 0 . √  j 0 j √ ∆ σ(x; θ) j=0 ∆ X   A¨ıt-Sahalia (2002) gave the following theorem about the convergence of the approximation pJ to the exact transition density p.

Theorem 1.4.4 Under the Conditions 1.4.1 and 1.4.2, there exists ∆¯ > 0 such that J lim p (∆, x0, x; θ)= p(∆, x0, x; θ) J →∞ for all ∆ (0, ∆)¯ , θ Θ and (x , x) (ℓ, r)2. ∈ ∈ 0 ∈

If rY = and a(y; θ) 0 near rY , and if a(y; θ) 0 near ℓY (which is either 0 or∞ ), then ∆=¯ ≤ , see Proposition 2 in A¨ıt-Sahalia≥ (2002). −∞ ∞ In order to use the expansions of the transition densities to calculate likelihood functions in practice, it is necessary to determine the coefficients ηj (∆,y0; θ). NON-MARTINGALEESTIMATINGFUNCTIONS 49

Note that by inserting (1.94) in the expression above for ηi(∆,y0; θ) we find that

1 ∞ η (∆,y ; θ) = H (z)∆1/2p (∆,y , ∆1/2z + y ; θ)dz i 0 i! i Y 0 0 Z−∞ 1 ∞ 1/2 = H (∆− (y y ))p (∆,y ,y; θ)dy i! i − 0 Y 0 Z−∞ 1/2 = E H (∆− (Y y )) Y = y . θ i ∆ − 0 | 0 0   Thus the coefficients ηi(∆,y0; θ), i =0, 1,..., are conditional moments of the process Y , and can therefore be found by simulation of Y or X. An approxi- mation to ηi(∆,y0; θ) can be obtained by applying the expansion (1.57) to the functions (y x)i, i =1,...,J. For instance, we find that − 1/2 1 3/2 1 2 η1(∆,y0; θ) = ∆ a(y0; θ)+ 2 ∆ a(y0; θ)∂ya(y0; θ)+ 2 ∂y a(y0; θ) 5/2 + O(∆ )  2 2 η2(∆,y0; θ) = ∆ a(y0; θ) + ∂ya(y0; θ) + O(∆ ).

By expanding the coefficients ηi(∆,y0; θ) suitably and collecting terms of the same order in ∆, A¨ıt-Sahalia (2002) found the following approximation to pY K p˜Y (∆,y0,y; θ)= y K k 1/2 y y0 ∆ ∆− ϕ − exp a(w,θ)dw ck(y0,y; θ), √∆ y0 k!   Z  Xk=0 where c0(y0,y; θ)=1, and

ck(y0,y; θ)= y k k 1 1 2 k(y y0)− (w y0) − λ(w; θ)ck 1(y0, w; θ)+ 2 ∂wck 1(y0, w; θ) dw, − − − − Zy0   for k 1, where ≥ λ(w; θ)= 1 a(w; θ)2 + ∂ a(w; θ) . − 2 w  1.5 Non-martingale estimating functions

1.5.1 Asymptotics

When the estimating function n

Gn(θ)= g(X(i r+1)∆,...,Xi∆; θ) − i=r X 50 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES is not a martingale under Pθ, further conditions on the diffusion process must be imposed to ensure the asymptotic normality in (1.5). A sufficient condi- tion that (1.5) holds under Pθ0 with V (θ) given by (1.98) is that the diffusion process is stationary and geometrically α-mixing, that T V (θ) = Qθ0 g(θ)g(θ) (1.98)

∞  T + Eθ0 g(X∆,...,Xr∆)g(X(k+1)∆,...,X(k+r)∆) k=1 X  T + Eθ0 g(X(k+1)∆,...,X(k+r)∆)g(X∆,...,Xr∆) , 2+ǫ converges and is strictly positive definite, and that Qθ0 (gi(θ) ) < ,i = ∞ 1,...,p for some ǫ> 0, see e.g. Doukhan (1994). Here gi is the ith coordinate of g, and Qθ is the joint distribution of X∆,...,Xr∆ under Pθ. To define the concept of α-mixing, let t denote the σ-field generated by Xs s t , and let t denote the σ-field generatedF by X s t . A stochastic{ | process≤ } X is F { s | ≥ } said to be α-mixing under Pθ0 , if

sup Pθ0 (A)Pθ0 (B) Pθ0 (A B) α(u) t+u A t,B | − ∩ |≤ ∈F ∈F for all t > 0 and u> 0, where α(u) 0 as u . This means that X and → → ∞ t Xt+u are almost independent, when u is large. If positive constants c1 and c2 exist such that c2u α(u) c e− , ≤ 1 for all u > 0, then the process X is called geometrically α-mixing. For one- dimensional diffusions there are simple conditions for geometric α-mixing. If all non-zero eigenvalues of the generator (1.46) are larger than some λ > 0, then the diffusion is geometrically α-mixing with c2 = λ. This is for instance the case if the spectrum of the generator is discrete. Ergodic diffusions with a linear drift β(x α), β > 0, for instance the Pearson diffusions, are geomet- − − rically α-mixing with c2 = β; see Hansen, Scheinkman & Touzi (1998). Genon-Catalot, Jeantheau & Lar´edo (2000) gave the following simple suffi- cient condition for the one-dimensional diffusion that solves (1.11) to be geo- metrically α-mixing, provided that it is ergodic with invariant probability den- sity µθ.

Condition 1.5.1 (i) The function b is continuously differentiable with respect to x, and σ is twice continuously differentiable with respect to x, σ(x; θ) > 0 for all x (ℓ, r), 2∈ and a constant Kθ > 0 exists such that b(x; θ) Kθ(1+ x ) and σ (x; θ) K (1 + x2) for all x (ℓ, r). | |≤ | | ≤ θ ∈ (ii) σ(x; θ)µ (x) 0 as x ℓ and x r. θ → ↓ ↑ NON-MARTINGALEESTIMATINGFUNCTIONS 51

(iii) 1/γ(x; θ) has a finite limit as x ℓ and x r, where γ(x; θ)= ∂xσ(x; θ) 2b(x; θ)/σ(x; θ). ↓ ↑ −

Other conditions for geometric α-mixing were given by Veretennikov (1987), Hansen & Scheinkman (1995), and Kusuoka & Yoshida (2000).

For geometrically α-mixing diffusions processes and estimating functions Gn satisfying Condition 1.2.1, the existence of a θ¯-consistent and asymptotically normal Gn-estimator follows from Theorem 1.2.2, which also contains a result about eventual uniqueness of the estimator.

1.5.2 Explicit non-martingale estimating functions

Explicit martingale estimating functions are only available for the relatively small, but versatile, class of diffusions for which explicit eigenfunctions of the generator are available; see the Subsections 1.3.6 and 1.3.7. Explicit non- martingale estimating functions can be found for all diffusions, but cannot be expected to approximate the score functions as well as martingale estimating functions, and therefore usually give less efficient estimators. As usual we con- sider ergodic diffusion processs with invariant probability density µθ. First we consider estimating functions of the form n Gn(θ)= h(X∆i; θ), (1.99) i=1 X where h is a p-dimensional function. We assume that the diffusion is geomet- rically α-mixing, so that a central limit theorem holds (under an integrability condition), and that Condition 1.2.1 holds for r = 1 and θ¯ = θ0. The latter condition simplifies considerably, because for estimating functions of the form (1.99), it does not involve the transition density, but only the invariant probabil- ity density µθ, which for one-dimensional ergodic diffusions is given explicitly by (1.15). In particular, (1.6) and (1.7) simplifies to r

µθ0 (h(θ0)) = h(x; θ0)µθ0 (x)dx =0 (1.100) Zℓ and r

W = µθ0 (∂θT h(θ0)) = ∂θT h(x; θ0)µθ0 (x)dx. Zℓ The condition for eventual uniqueness of the Gn-estimator (1.9) is here that θ0 is the only root of µθ0 (h(θ)). Kessler (2000) proposed

h(x; θ)= ∂θ log µθ(x), (1.101) 52 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES which is the score function (the of the log-likelihood function) if we pretend that the observations are an i.i.d. sample from the stationary distribu- tion. If ∆ is large, this might be a reasonable approximation. That (1.100) is satisfied for this specification of h follows under standard conditions that allow the interchange of differentiation and integration. r r r (∂θ log µθ(x)) µθ(x)dx = ∂θµθ(x)dx = ∂θ µθ(x)dx =0. Zℓ Zℓ Zℓ A modificationof the simple estimating function (1.101)was shown by Kessler, Schick & Wefelmeyer (2001) to be efficient in the sense of semiparamet- ric models. The modified version of the estimating function was derived by Kessler & Sørensen (2005) in a completely different way. Hansen & Scheinkman (1995) and Kessler (2000) proposed and studied the generally applicable specification

hj (x; θ)= Aθfj (x; θ), (1.102) where Aθ is the generator (1.46), and fj, j =1,...,p, are twice differentiable functions chosen such that Condition 1.2.1 holds. The estimating function with h given by (1.102) can easily be applied to multivariate diffusions, because an explicit expression for the invariant density µθ is not needed. The following lemma for one-dimensional diffusions shows that only weak conditions are needed to ensure that (1.100) holds for hj given by (1.102).

Lemma 1.5.2 Suppose f C2((ℓ, r)), A f L1(µ ) and ∈ θ ∈ θ 2 2 lim f ′(x)σ (x; θ)µθ(x) = lim f ′(x)σ (x; θ)µθ(x). (1.103) x r x ℓ → → Then r (Aθf)(x)µθ(x)dx =0. Zℓ 1 2 Proof: Note that by (1.15), the function ν(x; θ) = 2 σ (x; θ)µθ(x) satisfies that ν′(x; θ)= b(x; θ)µθ(x). In this proof all derivatives are with respect to x. It follows that r (Aθf)(x)µθ(x)dx ℓ Z r 1 2 = b(x; θ)f ′(x)+ 2 σ (x; θ)f ′′(x) µθ(x)dx Zℓ r  r = (f ′(x)ν′(x; θ)+ f ′′(x)ν(x; θ)) dx = (f ′(x)ν(x; θ))′ dx ℓ ℓ Z 2 2 Z = lim f ′(x)σ (x; θ)µθ(x) lim f ′(x)σ (x; θ)µθ(x)=0. x r − x ℓ → → 2 NON-MARTINGALEESTIMATINGFUNCTIONS 53

Example 1.5.3 Consider the square-rootprocess (1.37) with σ =1. For f1(x)= 2 x and f2(x)= x , we see that β(x α) A f(x)= , θ 2β−(x −α)x + x  − −  which gives the simple estimators

1 n X n n i∆ 1 i=1 αˆn = Xi∆, βˆn = X . n n n 2 i=1 1 1 X 2 X2 X n i∆ − n i∆  i=1 i=1 ! X X   The condition (1.103) is obviously satisfied because the invariant distribution is a normal distribution. 2

Conley et al. (1997) proposed a model-based choice of the fjs in (1.102): fj =

∂θj log µθ(x), i.e. the i.i.d. score function used in (1.101). Thus they obtained an estimating function of the form (1.99) with

h(x; θ)= Aθ∂θ log µθ(x). (1.104) Sørensen (2001) independently derived the same estimating function as an ap- proximation to the score function for continuous-time observation of the diffu- sion process. Jacobsen (2001) showed that this estimating function is small ∆- optimal. This result was later rediscovered by A¨ıt-Sahalia & Mykland (2008) who obtained a similar result for estimating functions given by (1.105).

An estimating function of the simple form (1.99) cannot be expected to yield as efficient estimators as an estimating function that depends on pairs of con- secutive observations, and therefore can use the information contained in the transitions. Hansen & Scheinkman (1995) proposed non-martingale estimating functions of the form (1.12) with g given by g (∆,x,y; θ)= h (y)A f (x) f (x)Aˆ h (y), (1.105) j j θ j − j θ j where the functions fj and hj satisfy weak regularity conditions ensuring that (1.6) holds for θ¯ = θ0. The differential operator Aˆθ is the generator of the time reversal of the observed diffusion X. For a multivariate diffusion it is given by

d d ˆ ˆ 1 2 Aθf(x)= bk(x; θ)∂xk f(x)+ 2 Ckℓ(x; θ)∂xk xℓ f(x), Xk=1 k,ℓX=1 54 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES where C = σσT and d ˆ 1 bk(x; θ)= bk(x; θ)+ ∂xℓ (µθCkl) (x; θ). − µθ(x) Xℓ=1 For one-dimensional ergodic diffusions, Aˆθ = Aθ. That ˆb = b for a one- dimensional diffusion follows from (1.15). Obviously, the estimating function of the form (1.99) with hj (x; θ)= Aθfj(x) is a particular case of (1.105) with hj (y)=1.

1.5.3 Approximate martingale estimating functions

For martingale estimating functions of the form (1.30) and (1.40), we can al- ways, as discussed in Subsection 1.3.3, obtain an explicit approximation to the optimal weight matrix by means of the expansion (1.47). For diffusion models where there is no explicit expression for the transition operator, it is tempting θ to go on and approximate the conditional moments π∆(fj (θ))(x) using (1.47), and thus, quite generally, obtain explicit approximate martingale estimating function. Such estimators were the first type of estimators for discretely ob- served diffusion processes to be studied in the literature. They have been con- sidered by Dorogovcev (1976), Prakasa Rao (1988), Florens-Zmirou (1989), Yoshida (1992), Chan et al. (1992), Kloeden et al. (1996), Kessler (1997), Kelly, Platen & Sørensen (2004), and many othes. It is, however, important to note that there is a dangerous pitfall when using these simple approximate martingale estimating functions. They do not satisfy the condition that Qθ0 (g(θ0)) = 0, and hence the estimators are inconsistent. To illustrate the problem, consider an estimating function of the form (1.12) with g(x, y; θ)= a(x, θ)[f(y) f(x) ∆A f(x)], (1.106) − − θ θ where Aθ is the generator (1.46), i.e., we have replaced π∆f(x) by a first order expansion. To simplify the exposition, we assume that θ, a and f are one- dimensional. We assume that the diffusion is geometrically α-mixing, that the other conditions mentioned above for the weak convergence result (1.5) hold, and that Condition 1.2.1 is satisfied. Then by Theorem 1.2.2, the estimator obtained using (1.106) converges to the solution, θ¯, of ¯ Qθ0 (g(θ))=0, (1.107) where, as usual, θ0 is the true parameter value. We assume that the solution is unique. Using the expansion (1.47), we find that

Q (g(θ)) = µ a(θ)[πθ0 f f ∆A f] θ0 θ0 ∆ − − θ = ∆µ  a(θ)[A f A f + 1 ∆A2 f] + O(∆3) θ0 θ0 − θ 2 θ0  NON-MARTINGALEESTIMATINGFUNCTIONS 55

= (θ θ)∆µ (a(θ )∂ A f)+ 1 ∆2µ a(θ )A2 f 0 − θ0 0 θ θ0 2 θ0 0 θ0 +O(∆ θ θ 2)+ O(∆2 θ θ )+ O(∆3). | − 0| | − 0|  If we neglect all O-terms, we obtain that ¯ . 1 2 θ = θ0 + ∆ 2 µθ0 a(θ0)Aθ0 f /µθ0 (a(θ0)∂θAθ0 f) , which indicates that when ∆ is small, the asymptotic bias is of order ∆. How- ever, the bias can be huge when ∆ is not sufficiently small as the following example shows.

Example 1.5.4 Consider again a diffusion with linear drift, b(x; θ)= β(x α). − − In this case (1.106) with f(x)= x gives the estimating function n

Gn(θ)= a(X∆(i 1); θ)[X∆i X∆(i 1) + β X∆(i 1) α ∆], − − − − − i=1 X  where a is 2-dimensional. For a diffusion with linear drift, we found in Exam- ple 1.3.8 that β∆ β∆ F (x; α, β)= xe− + α(1 e− ). − Using this, we obtain that

β0∆ Q (g(θ)) = c (e− 1+ β∆) + c β(α α), θ0 1 − 2 0 − where

c = a(x)xµ (dx) µ (a)α , c = µ (a)∆. 1 θ0 − θ0 0 2 θ0 ZD Thus α¯ = α0 and 1 e β0∆ 1 β¯ = − − . ∆ ≤ ∆ We see that the estimator of α is consistent, while the estimator of β will tend to be small if ∆ is large, whatever the true value β0 is. We see that what deter- mines how well βˆ works is the magnitude of β0∆, so it is not enough to know that ∆ is small. Moreover, we cannot use βˆ∆ to evaluate whether there is a problem, because this quantity will always tend to be smaller than one. If β0∆ actually is small, then the bias is proportional to ∆ as expected β¯ = β 1 ∆β2 + O(∆2). 0 − 2 0 We get an impression of how terribly misled we can be when estimating the 56 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES parameter β by means of the dangerous estimating function given by (1.106) from a simulation study in Bibby & Sørensen (1995) for the square root pro- cess (1.37). The result is given in Table 1.1. For the weight function a, the approximately optimal weight function was used, cf. Example 1.3.8. For dif- ferent values of ∆ and the sample size, 500 independent datasets were simu- lated, and the estimators were calculated for each dataset. The expectation of the estimator βˆ was determined as the average of the simulated estimators. The true parameter values were α0 = 10, β0 =1 and τ0 =1, and the initial value was x0 = 10. When ∆ is large, the behaviour of the estimator is bizarre. 2

∆ #obs. mean ∆ #obs. mean 0.5 200 0.81 1.5 200 0.52 500 0.80 500 0.52 1000 0.79 1000 0.52 1.0 200 0.65 2.0 200 0.43 500 0.64 500 0.43 1000 0.63 1000 0.43

Table 1.1 Empirical mean of 500 estimates of the parameter β in the CIR model. The true parameter values are α0 = 10, β0 = 1, and τ0 = 1.

The asymptotic bias given by (1.107) is small when ∆ is sufficiently small, and the results in the following section on high frequency asymptotics show that in this asymptotic scenario the approximate martingale estimating func- tions work well. However, how small ∆ needs to be depends on the parameter values, and without prior knowledge about the parameters, it is safer to use an exact martingale estimating function, which gives consistent estimators at all sampling frequencies.

1.6 High-frequency asymptotics

An expression for the asymptotic variance of estimators was obtained in Theo- rem 1.3.2 using a low frequency asymptotic scenario, where the time between observations is fixed. This expression is rather complicated and is not easy to use for comparing the efficiency of different estimators. Therefore the relative merits of estimators have often been investigated by simulation studies, and the general picture has been rather confusing. A much simpler and more manage- able expression for the asymptotic variance of estimators can be obtained by considering the high frequency scenario, n , ∆ 0, n∆ . (1.108) → ∞ n → n → ∞ HIGH-FREQUENCY ASYMPTOTICS 57

The assumption that n∆n is needed to ensure that parameters in the drift coefficient can be consistently→ ∞ estimated. For this type of asymptotics Sørensen (2007) obtained simple conditions for rate optimality and efficiency for ergodic diffusions, which allow identification of estimators that work well when the time between observations, ∆n, is not too large. How small ∆n needs to be for the high frequency scenario to be relevant, depends on the speed with which the diffusion moves. For financial data the speed of reversion is usually slow enough that this type of asymptotics works for daily, sometimes even weekly observations. A main result of the the- ory in this section is that under weak conditions optimal martingale estimating functions give rate optimal and efficient estimators. It is also interesting that the high frequency asymptotics provides a very clear statement of the important fact that parameters in the diffusion coefficient can be estimated more exactly than drift parameters when the time between ob- servations is small. A final advantage of high frequency asymptotics is that it also gives useful results about the approximate martingale estimating functions discussed in Subsection 1.5.3, in situations where they work. To simplify the exposition, we restrict attention to a one-dimensional diffusion given by dXt = b(Xt; α)dt + σ(Xt; β)dWt, (1.109) where θ = (α, β) Θ IR2. The results below can be generalized to mul- tivariate diffusions∈ and parameters⊆ of higher dimension. We consider estimat- ing functions of the general form (1.3), where the two-dimensional function g = (g ,g ) for some κ 2 and for all θ Θ satisfies 1 2 ≥ ∈

Eθ(g(∆n,X∆ni,X∆n(i 1); θ) X∆n(i 1)) (1.110) − − | κ = ∆nR(∆n,X∆n(i 1); θ). − Martingale estimating functions obviously satisfy (1.110) with R = 0, but for instance the approximate martingale estimating functions discussed at the end of the previous section satisfy (1.110) too. Here and later R(∆,y,x; θ) denotes a function such that R(∆,y,x; θ) F (y, x; θ), where F is of poly- nomial growth in y and x uniformly| for θ in|≤ compact sets. This means that for any compact subset K Θ, there exist constants C1, C2, C3 > 0 such that ⊆ C2 C3 supθ K F (y, x; θ) C1(1 + x + y ) for all x and y in the state space of the∈ diffusion.| |≤ | | | | The main results in this section are simple conditions on the function g(∆,y,x; θ) that ensure rate optimality and efficiency of estimators. The condition for rate optimality is

Condition 1.6.1

∂yg2(0, x, x; θ)=0 (1.111) 58 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES for all x (ℓ, r) and all θ Θ. ∈ ∈

By ∂yg2(0, x, x; θ) we mean ∂yg2(0,y,x; θ) evaluated at y = x. This condi- tion is called the Jacobsen condition,because it was first found in the theory of small ∆-optimal estimation developed in Jacobsen (2001), cf. (1.62) in Sub- section 1.3.4. The condition for efficiency is

Condition 1.6.2

2 ∂yg1(0, x, x; θ)= ∂αb(x; α)/σ (x; β) (1.112) and 2 2 4 ∂y g2(0, x, x; θ)= ∂βσ (x; β)/σ (x; β), (1.113) for all x (ℓ, r) and all θ Θ. ∈ ∈ Also (1.112) and (1.113) were found as conditions for small ∆-optimality in Jacobsen (2002), cf. (1.61) and (1.63). This is not surprising. The following theorem provides an interpretation of small ∆-optimality in terms of the clas- sical statistical concepts rate optimality and efficiency. As usual, θ0 = (α0,β0) denotes the true parameter value.

Theorem 1.6.3 Assume that the diffusion is ergodic, that θ0 int Θ, and that the technical regularity Condition 1.6.4 given below holds. Denote∈ the density function of the invariant probability measure by µθ. Suppose that g(∆,y,x; θ) satisfies Condition 1.6.1. Assume, moreover, that the following identifiability condition is satisfied r [b(x, α ) b(x, α)]∂ g (0, x, x; θ)µ (x)dx = 0 when α = α , 0 − y 1 θ0 6 6 0 Zℓ r [σ2(x, β ) σ2(x, β)]∂2g (0, x, x; θ)µ (x)dx = 0 when β = β , 0 − y 2 θ0 6 6 0 Zℓ and that r S = ∂ b(x; α )∂ g (0, x, x; θ )µ (x)dx =0, 1 α 0 y 1 0 θ0 6 Zℓ r S = 1 ∂ σ2(x; β )∂2g (0, x, x; θ )µ (x)dx =0. 2 2 β 0 y 2 0 θ0 6 Zℓ Then a consistent Gn–estimator θˆn = (ˆαn, βˆn) exists and is unique in any compact subset of Θ containing θ0 with probability approaching one as n . If, moreover, → ∞ 2 ∂α∂y g2(0, x, x; θ)=0, (1.114) HIGH-FREQUENCY ASYMPTOTICS 59 then for a martingale estimating function, and for more general estimating functions if n∆2(κ 1) 0, − → W1 √n∆n(ˆαn α0) S2 0 − 0 11 D N2 , (1.115) ˆ −→  0  W2  √n(βn β0) ! ! 0 S2 − 22    where r 2 2 W1 = σ (x; β0)[∂yg1(0, x, x; θ0)] µθ0 (x)dx ℓ Z r 1 4 2 2 W2 = 2 σ (x; β0)[∂y g2(0, x, x; θ0)] µθ0 (x)dx. Zℓ Note that the estimator of the diffusion coefficient parameter, β, converges faster than the estimator of the drift parameter, α, and that the two estima- tors are asymptotically independent. Gobet (2002) showed, under regularity conditions, that a discretely sampled diffusion model is locally asymptotically normal under high frequency asymptotics, and that the optimal rate of conver- gence for a drift parameter is 1/√n∆n, while it is 1/√n for a parameter in the diffusion coefficient. Thus under the conditions of Theorem 1.6.3 the es- timators αˆn and βˆn are rate optimal. More precisely, Condition 1.6.1 implies rate optimality. If this condition is not satisfied, the estimator of the diffusion coefficient parameter, β, does not use the information about the diffusion coef- ficient contained in the quadratic variation and therefore converges at the same relatively slow rate 1/√n∆n as estimators of α, see Sørensen (2007). Gobet gave the following expression for the Fisher information matrix

W1 0 = , (1.116) I 0 W2 ! where r (∂ b(x; α ))2 W = α 0 µ (x)dx, (1.117) 1 σ2(x; β ) θ0 Zℓ 0 r ∂ σ2(x; β ) 2 W = β 0 µ (x)dx. (1.118) 2 σ2(x; β ) θ0 Zℓ  0  By comparing the covariance matrix in (1.115) to (1.116), we see that Con- dition 1.6.2 implies that S1 = W1 and S2 = W2, with W1 and W2 given by (1.117) and (1.118), and that hence the asymptotic covariance matrix of (ˆαn, βˆn) under Condition 1.6.2 equals the inverse of the Fisher information matrix (1.116). Thus Condition 1.6.2 ensures efficiency of (ˆαn, βˆn). Under the conditions of Theorem 1.6.3 and Condition 1.6.2, we see that for a martingale 60 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES estimating function, and more generally if n∆2(κ 1) 0, − → √n∆n(ˆαn α0) 0 − D N , 1 . (1.119) ˆ 2 0 − √n(βn β0) ! −→ I −    Note that condition (1.114) is automatically satisfied under the efficiency Con- dition 1.6.2. Proof of Theorem 1.6.3: Only a brief outline the proof is given; for details see Sørensen (2007). Consider the normalized estimating function n 1 G (θ)= g(∆ ,X n ,X n ; θ). n n ti ti 1 n∆ − n i=1 X First the conditions of Theorem 1.10.2 must be checked. Using Lemma 9 in Genon-Catalot & Jacod (1993), it can be shown that G (θ ) 0 in P - n 0 → θ0 probability, and that ∂θT Gn(θ) under Pθ0 converges pointwise to a matrix, which for θ = θ0 is upper triangular and has diagonal elements equal to S1 and S2, and thus is invertible. In order to prove that the convergenceis uniform for θ in a compact set K, we show that the sequence n 1 ζ ( )= g(∆ ,X n ,X n , ) n n ti ti 1 · n∆ − · n i=1 X converges weakly to the limit γ( ,θ0) in the space, C(K), of continuous func- tions on K with the supremum norm.· Since the limit is non-random, this im- plies uniform convergence in probability for θ K. We have proved point- wise convergence, so the weak convergence result∈ follows because the family of distributions of ζn( ) is tight. The tightness is shown by checking the condi- tions in Corollary 14.9· in Kallenberg (1997). Thus the conditions of Theorem 1.10.2 are satisfied, and we conclude the existence of a consistent and even- tually unique Gn-estimator. The uniqueness on compact subsets follows from Theorem 1.10.3 because the identifiability condition in Theorem 1.6.3 implies (1.160). The asymptotic normality of the estimators follows from Theorem 1.10.4 with √∆ n 0 A = n . n 0 √n   The weak convergence of AnGn(θ0) follows from a central limit theorem for martingales, e.g. Corollary 3.1 in Hall & Heyde (1980). The uniform conver- 1 gence of An∂θT Gn(θ)An− was proved for three of the entries when the con- ditions of Theorem 1.10.2 were checked. The result for the last entry is proved in a similar way using (1.114). 2 HIGH-FREQUENCY ASYMPTOTICS 61 The reader is reminded of the trivial fact that for any non-singular 2 2 ma- × trix, Mn, the estimating functions MnGn(θ) and Gn(θ) have exactly the same roots and hence give the same estimator(s). We call them versions of the same estimating function. The matrix Mn may depend on ∆n. The point is that a version must exist which satisfies the conditions (1.111) – (1.113), but not all versions of an estimating function satisfy these conditions. It follows from results in Jacobsen (2002) that to obtain a rate optimal and efficient estimator from an estimating function of the form (1.41), we need that N 2 and that the matrix ≥ 2 ∂xf1(x; θ) ∂xf1(x; θ) D(x)= 2 ∂xf2(x; θ) ∂xf2(x; θ) !

is invertible for µθ-almost all x. Under these conditions,Sørensen (2007)showed that Godambe-Heyde optimal martingale estimating functions give rate opti- mal and efficient estimators. For a d-dimensional diffusion, Jacobsen (2002) gave the conditions N d(d + 3)/2, and that the N (d + d2)-matrix 2 ≥ × D(x) = ∂xf(x; θ) ∂xf(x; θ) has full rank d(d + 3)/2, which are needed to ensure the existence of a rate optimal and efficient estimator from an esti- mating function of the form (1.41). We conclude this section by an example, but first we state technical conditions under which the results in this section hold. The assumptions about polynomial growth are far too strong, but simplify the proofs. These conditions can most likely be weakened considerably.

Condition 1.6.4 The diffusion is ergodic with invariant probability density µθ, and the following conditions hold for all θ Θ: ∈ (1) r xkµ (x)dx < for all k IN. ℓ θ ∞ ∈ (2) sup E ( X k) < for all k IN. R t θ | t| ∞ ∈ (3) b, σ C ((ℓ, r) Θ). ∈ p,4,1 × (4) There exists a constant C such that for all x, y (ℓ, r) θ ∈ b(x; α) b(y; α) + σ(x; β) σ(y; β) C x y | − | | − |≤ θ| − | 2 (5) g(∆,y,x; θ) Cp,2,6,2(IR+ (ℓ, r) Θ) and has an expansion in powers of ∆: ∈ × × g(∆,y,x; θ)= (1) 1 2 (2) 3 g(0,y,x; θ) + ∆g (y, x; θ)+ 2 ∆ g (y, x; θ) + ∆ R(∆,y,x; θ), where g(0,y,x; θ) C ((ℓ, r)2 Θ), ∈ p,6,2 × g(1)(y, x; θ) C ((ℓ, r)2 Θ), ∈ p,4,2 × 62 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES g(2)(y, x; θ) C ((ℓ, r)2 Θ). ∈ p,2,2 × 2 We define Cp,k1,k2,k3 (IR+ (ℓ, r) Θ) as the class of real functions f(t,y,x; θ) satisfying that × ×

(i) f(t,y,x; θ) is k1 times continuously differentiable with respect t, k2 times continuously differentiable with respect y, and k3 times continuously dif- ferentiable with respect α and with respect to β

i1 i2 i3 i4 (ii) f and all partial derivatives ∂t ∂y ∂α ∂β f, ij = 1,...kj , j = 1, 2, i3 + i4 k3, are of polynomial growth in x and y uniformly for θ in a compact set≤ (for fixed t).

2 The classes Cp,k1,k2 ((ℓ, r) Θ) and Cp,k1,k2 ((ℓ, r) Θ) are defined similarly for functions f(y; θ) and f×(y, x; θ), respectively. ×

Example 1.6.5 We can now interpret the findings in Example 1.3.11 as fol- lows. The general quadratic martingale estimating function (1.64) gives rate optimal estimators in the high frequency asymptotics considered in this sec- tion. Moreover, the estimators are efficient in three particular cases: the opti- mal estimating function given in Example 1.3.6 and the approximations (1.28) and (1.51). Kessler (1997) considered an approximation to the Gaussian quasi-likelihood presented in Subsection 1.3.2, where the conditional mean F and the condi- tional variance Φ are approximated as follows. The conditional mean is re- placed by the expansion

k k 1 ∆i − ∆i r (∆, x; θ)= Ai f(x)= x + ∆ Ai b(x; α), k i! θ (i + 1)! θ i=0 i=0 X X 2 where f(x)= x, cf. (1.47). For fixed x, y and θ the function (y rk(∆, x; θ)) is a polynomial in ∆ of order 2k. Define gj (y), j =0, 1, −, k by x,θ · · · k (y r (∆, x; θ))2 = ∆j gj (y)+ O(∆k+1). − k x,θ j=0 X For instance, for k =2 (y r (∆, x; θ))2 = − 2 (y x)2 2(y x)b(x; α)∆+ (y x)A b(x; α)+b(x; α)2 ∆2 +O(∆3), − − − − θ j from which we can see the expressionsfor gx,θ(y), j =0, 1, 2. The conditional variance can be approximated by

k k j − ∆r Γ (∆, x; θ)= ∆j Argj (x). k r! θ x,θ j=0 r=0 X X HIGH-FREQUENCY ASYMPTOTICS IN A FIXED TIME-INTERVAL 63 In particular, Γ (∆, x; θ) = ∆σ2(x; β)+ 1 ∆2 A σ2(x; β) σ2(x; β)∂ b(x; α) . 2 2 θ − x By inserting these approximations in (1.28), we obtain the approximate mar- tingale estimating function n (k) ∂θrk(∆i,Xti 1 ; θ) − Hn (θ) = [Xti rk(∆i,Xti 1 ; θ)] (1.120) − Γk+1(∆i,Xti 1 ; θ) − i=1 − n X ∂θΓk+1(∆i,Xti 1 ; θ) 2 − + 2 [(Xti rk(∆i,Xti 1 ; θ)) Γk+1(∆i,Xti 1 ; θ)]. 2Γ (∆ ,X ; θ) − − − − i=1 k+1 i ti 1 X − Kessler (1997) (essentially) showed that for ergodic diffusions satisfying Con- (k) dition 1.6.4 (1) – (4), the estimator obtained from Hn (θ) satisfies (1.119) provided that n∆2k+1 0. → 2

1.7 High-frequency asymptotics in a fixed time-interval

We will now briefly consider a more extreme type of high-frequence asymp- totics, where the observation times are restricted to a bounded interval, which, without loss of generality, we can take to be [0, 1]. Suppose that the d-dimen- sional diffusion X which solves (1.109) has been observed at the time points ti = i/n, i = 0,...,n. Note that in this section W in equation (1.109) is a d- dimensional standard Wiener process, and σ is a d d-matrix. We assume that the matrix C(x; β)= σ(x; β)σ(x; β)T is invertible× for all x in the state space, D, of X. Because the observation times are bounded, the drift parameter, α, cannot be consistently estimated as n , so in the following we consider es- timation of β only, and concentrate on→ the ∞ following Gaussian quasi-likelihood function:

Qn(β)= (1.121) n T 1 log det C(Xti 1 ; β)+ n(Xti Xti 1 ) C(Xti 1 ; β)− (Xti Xti 1 ) . − − − − − − i=1 X   This is an approximation to a multivariate version of the Gaussian quasi-likeli- hood in Subsection 1.3.2 with b = 0, where the conditional mean F (x; θ) is approximated by x, and the conditional covariance matrix Φ is approximated 1 by n− C. An estimator is obtained by minimizing Qn(β). This estimator can also be obtained from the approximate martingale estimating function which we get by differentiating Qn(β) with respect to β. The drift may be known, but in general we allow it to depend on an unknown parameter α. We assume 64 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES that θ = (α, β) A B = Θ, and we denote the true parameter value by ∈ × θ0 = (α0,β0). Genon-Catalot & Jacod (1993) showed the following theorem under the as- sumption that β B, where B is a compact subset of IRq, which ensures that a βˆ B that minimizes∈ Q (β) always exists. The results in the theorem hold n ∈ n for any βˆn that minimizes Qn(β).

Theorem 1.7.1 Assume that Condition 1.7.2 given below holds. Then the esti- mator βˆ is consistent, and provided that β int B, n 0 ∈ √n(βˆ β ) D Z, n − 0 −→ where the distribution of Z is a normal variance mixture with characteristic function 1 T 1 s E exp s W (β )− s 7→ θ0 − 2 0 with W (β) given by (1.122). Conditional on W (β0), the asymptotic distri- bution of √n(βˆn β0) is a centered q-dimensional normal distribution with − 1 covariate matrix W (β0)− .

We will not prove Theorem 1.7.1 here. Note, however, that to do so we need the full generality of the Theorems 1.10.2, 1.10.3 and 1.10.4, where the ma- trix W (θ) (equal to W0(θ) in Theorem 1.10.4) is random. Only if the matrix B(x; β) defined below does not depend on x, is W (β) non-random, in which case the limit distribution is simply the centered q-dimensional normal distri- 1 bution with covariate matrix W (β0)− . A simple example of a non-random W (β) is when β is one-dimensional and a q q-matrix F (x) exists such that C(x; β)= βF (x). So for instance for the Ornstein-Uhlenbeck× process and the square-root diffusion (1.37), W (β) is non-random, and the limit distribution is normal.

Condition 1.7.2 The stochastic differential equation (1.109) has a non-exploding, unique strong solution for t [0, 1], and the following conditions hold for all θ = (α, β) Θ: ∈ ∈ 2 (1) b(x; α) is a continuous function of x, and the partial derivatives ∂xσ(x; β), 2 ∂x∂βσ(x; β), ∂βσ(x; β) exist and are continuous functions of (x, β) D B. ∈ ×

(2) With Pθ-probability one it holds that for all β1 = β, the functions t C(X ; β ) and t C(X ; β) are not equal. 6 7→ t 1 7→ t (3) The random q q- matrix × 1 W (β)= B(Xt; β)dt, (1.122) Z0 SMALL-DIFFUSIONASYMPTOTICS 65 where the ijth entry of B(x; β) is given by 1 1 B(x; β)ij =2 tr ∂βi C(x; β)C(x; β)− ∂βj C(x; β)C(x; β)− ,

is invertible Pθ-almost surely. 

The Condition 1.7.2 (2) can be difficult to check because it depends on the path of the process X. It is implied by the stronger condition that for all β1 = β, C(x; β ) = C(x; β) for almost all x D. 6 1 6 ∈ Gobet (2001) showed, under regularity conditions, that for the high-frequency asymptotics in a fixed time-interval considered in this section, the diffusion model is locally asymptotically mixed normal (LAMN) with rate √n and con- ditional variance given by W (β); see e.g. Le Cam & Yang (2000) for the def- inition of LAMN. Therefore the estimator discussed above is efficient in the sense of Jeganathan (1982) and Jeganathan (1983).

Example 1.7.3 Consider the one-dimensional model given by

dX = (X α)dt + β + X2dW , t − t − t t q where α> 0 and β > 0. In this case c(x; β)= β + x2, so 1 2X4 W (β)= t dt, (β + X2)2 Z0 t which is random.

2

1.8 Small-diffusion asymptotics

Under the high-frequency asymptotics with bounded observation times consi- dered in the previous section, drift parameters could not be consistently esti- mated. Here we combine the high-frequency asymptotics with small-diffusion asymptotics to show that if the diffusion coefficient is small, we can find accu- rate estimators of drift parameters even when we have only observations in a bounded time-interval, which we again take to be [0, 1].

We consider observations that the time points ti = i/n, i = 1,...,n, of a d-dimensional diffusion process that solves the stochastic differential equation

dXt = b(Xt, α)dt + εσ(Xt,β)dWt, X0 = x0, (1.123) with ε > 0 and (α, β) A B, where A IRq1 and B IRq2 are convex, compact subsets.∈ It is assumed× that ǫ is⊆ known, while the⊆ parameter θ = (α, β) Θ= A B must be estimated. In (1.123) W is a d-dimensional ∈ × 66 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES standard Wiener process, and σ is a d d-matrix. We assume that the matrix C(x; β)= σ(x; β)σ(x; β)T is invertible× for all x in the state space, D, of X. In this section the asymptotic scenario is that n and ε 0 with a suitable balance between the of→ the ∞ two. Small→ diffusion asymptotics, where ε 0, has been widely studied and has proved fruitful in applied problems, see→ e.g. Freidlin & Wentzell (1998). Applications to contin- gent claim pricing and other financial problems can be found in Takahashi & Yoshida (2004) and Uchida & Yoshida (2004a), and applications to filtering problems in Picard (1986) and Picard (1991). The estimation problem outlined above was studied by Genon-Catalot (1990), Sørensen & Uchida (2003), and Gloter & Sørensen (2009). Here we follow Gloter & Sørensen (2009), which generalize results in the other papers, and consider the asymptotic scenario: ρ n εn 0 lim inf εnn > 0 (1.124) n → ∞ → →∞ for some ρ> 0. When ρ is large, ǫ can go faster to zerothan when ρ is relatively small. The value of ρ depends on the quasi-likelihood, as we shall see below. The solution to (1.123) for ǫ = 0 plays a crucial role in the theory. It is obvi- ously non-random. More generally, we define the flow ξt(x, α) as the solution to the equation

∂tξt(x, α)= b(ξt(x, α), α), ξ0(x, α)= x, (1.125) for all x D. The solution to (1.123) for ǫ =0 is given by ξt(x0, α). A related function∈ of central importance is δ˜ (x, α)= ξ (x, α) x. (1.126) n 1/n − ˜ When ε is small, δn(Xti 1 , α)+ Xti 1 approximates the conditional expecta- − − tion of Xti given Xti 1 , and can be used to define a Gaussian quasi-likelihood. However, equation (1.125)− does not generally have en explicit solution, so ξt(x, α) is usually not explicitly available. Therefore we replace it by an ap- proximation δ(x, α) that satisfies Condition 1.8.2 (5) given below. Using this approximation, we define a Gaussian quasi-log-likelihood by n 2 T 1 Uε,n(θ)= log det Ck 1(β)+ ε− nPk(α) Ck 1(β)− Pk(α) , − − k=1 X  (1.127) where

Pk(α) = Xk/n X(k 1)/n δn(X(k 1)/n, α) − − − − T Ck(β) = σ(Xk/n,β)σ(Xk/n,β) . This is the log-likelihood function that would have been obtained if the con- ditional distribution of Xti given Xti 1 were a normal distribution with mean − 2 δn(Xti 1 , α)+ Xti 1 and covariance matrix (ti ti 1)ε Ck 1(β). − − − − − SMALL-DIFFUSIONASYMPTOTICS 67

When ξ is explicitly available, a natural choice is δn(x, α)= δ˜n(x, α). Other- wise, simple useful approximations to δ˜n(x, α) are given by

k j k n− j 1 δ (x, α)= ( ) − (b( , α))(x), n j! Lα · j=1 X k =1, 2 . . . , where the operator is defined by Lα d (f)(x)= b (x, α)∂ f(x). Lα i xi i=1 X j By ( α) we denote j-fold application of the operator α. The approxima- Lk L tion δn satisfies Conditions 1.8.2 (5)-(6), when k 1/2 ρ. The first two approximations are − ≥ 1 1 δn(x, α)= n− b(x, α), for which the quasi-likelihood studied in Sørensen & Uchida (2003) is ob- tained, and d 2 1 1 2 δn(x, α)= n− b(x, α)+ 2 n− bi(x, α)∂xi b(x, α). i=1 X

Since the parameter space Θ is compact, a θˆε,n = (ˆαε,n, βˆε,n) that minimizes the Gaussian quasi- log-likelihood Uε,n(θ) always exists. The results in the following theorem hold for any θˆε,n that minimizes Uε,n(θ). As usual, θ0 = (α0,β0) denotes the true parameter value.

Theorem 1.8.1 Assume that Condition 1.8.2 given below holds, that θ0 int Θ, and that the matrix ∈ I (θ ) 0 I(θ )= 1 0 0 0 I (θ )  2 0  is invertible, where the ijth entries of the q q matrix I and of the q q 1 × 1 1 2 × 2 matrix I2 are given by i,j I1 (θ0)= 1 T 1 ∂αi b(ξs(x0, α0), α0) C− (ξs(x0, α0),β0)∂αj b(ξs(x0, α0), α0)ds Z0 and 1 i,j 1 1 1 Iσ (θ0)= 2 tr (∂βi C)C− (∂βj C)C− (ξs(x0, α0),β0) ds. Z0   Then, under the asymptotic scenario (1.124), the estimator θˆε,n is consistent, and 1 ε− (ˆαε,n α0) 1 − D N 0, I(θ0)− . √n(βˆ β ) −→  ε,n 0  −  68 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES We do not prove the theorem here, but a similar result for the estimating func- tion obtained by differentiation of Uε,n(θ) with respect to θ can be proved using the asymptotic results in Section 1.10. Note that the estimators of the drift and diffusion coefficient parameters are asymptotically independent. The 1 two parametersare not estimated at the same rate. For the approximation δn the 1 conditions below are satisfied if ǫ− converges at a rate smaller than or equal to √n, so in this case the rate of convergenceof αˆε,n is slower than or equal to that of βˆ . For the approximations δk, k 2, the rate of convergenceof αˆ ε,n n ≥ ε,n can be slower than or faster than that of βˆε,n, dependent on how fast ǫ goes to zero.

The matrix I1 equals the Fisher information matrix when the data is a continu- ous sample path in [0, 1] and ε 0, cf. Kutoyants (1994), so αˆ is efficient. → ε,n Probably βˆε,n is efficient too, but this cannot be seen in this simple way and has not yet been proved. We now give the technical conditions that imply Theorem 1.8.1.

Condition 1.8.2 The following holds for all ε> 0:

(1) The stochastic differential equation (1.123) has a unique strong solution for t [0, 1] for all θ = (α, β) Θ. ∈ ∈ (2) b(x; α) is a smooth (i.e. ∞) function of (x, α), and a constant c exists such that for all x, y D andC all α , α A: ∈ 1 2 ∈ b(x; α ) b(y; α ) c( x y + α α ). | 1 − 2 |≤ | − | | 1 − 2| (3) σ(x; β) is continuous, and there exists an open convex subset D such that ξ (x , α ) for all t [0, 1], and σ(x; β) is smooth onU ⊆ B. t 0 0 ∈U ∈ U × (4) If α = α , then the two functions t b(ξ (x , α ); α) and t b(ξ (x , α ); α ) 6 0 7→ t 0 0 7→ t 0 0 0 are not equal. If β = β0, then the two functions t C(ξt(x0, α0); β) and t C(ξ (x , α ); β6 ) are not equal. 7→ 7→ t 0 0 0 (5) The function δn(x; α) is smooth, and for any compact subset K D, a constant c(K) exists such that ⊆

3/2 sup δn(x; α) δ˜n(x; α) c(K)εn− . x K,α A − ≤ ∈ ∈

Similar bounds hold for the first two derivatives of δn w.r.t. α. (6) For any compact subset K D A, there exists a constant c(K), inde- pendent of n, such that ⊆ × nδ (x; α ) nδ (x; α ) c(K) α α | n 1 − n 2 |≤ | 1 − 2| for all (x, α ), (x, α ) K and for all n IN. The same holds for deriva- 1 2 ∈ ∈ tives of any order w.r.t. α of nδn. SMALL-DIFFUSIONASYMPTOTICS 69

It can be shown that δn(x, α) = δ˜n(x, α) satisfies Condition 1.8.2 (6), under Condition 1.8.2 (2). This choice of δn trivially satisfies Condition 1.8.2 (5).

Example 1.8.3 Consider the two dimensional diffusion X = (Y, R) given by 1 dYt = (Rt + µ1)dt + εκ1dWt dR = µ (R m)dt + εκ R ρdW 1 + 1 ρ2dW 2 , t − 2 t − 2 t t − t p  p  where (Y0, R0) = (y0, r0) with r0 > 0. This model was used in finance by Longstaff & Schwartz (1995). In their mode, the second component represents the short term interest rate, while Y is the logarithm of the price of some asset. The second component is the square-root diffusion. The parameters are θ = 2 2 (α, β), where α = (µ1, µ2,m) and β = (κ1,κ2,ρ). The parameter ρ allows correlation between the innovation terms of the two coordinates. The diffusion process (Y, R) satisfies Condition 1.8.2 (1) – (3), and (4) is holds if r0 = m0. The equation (1.125) is linear and has the solution 6

1 µ2t y + (µ1 + m)t + µ2− (r m)(1 e− ) ξt(y, r, µ1, µ2,m)= − − . µ2t m + (r m)e− ! − Therefore we can choose δn(x, α)= δ˜n(x, α), which satisfies Condition 1.8.2 (5) – (6). The matric I(θ ) is invertible when r = m and is given by 0 0 6 0 2 κ1− 0 µ m(µ2+log(q))+(m r0)(e− 2 1)) µ1+log(q) 2 1 0 − 2 − − − 2 I1(θ)=(1 ρ )−  κ µ2 κ  , − 2 2 µ1+log(q) µ2 log(q) −  0 κ2 mκ2 ,   2 − 2    where q = r /(r + m(eµ2 1)), and 0 0 − 2κ4 2ρ2κ2κ2 ρ(1 ρ2)κ2 1 1 2 − 1 2 2 2 4 2 2 I2(θ)=  ρ κ κ 2κ ρ(1 ρ )κ  . 1 2 2 − 2  ρ(1 ρ2)κ2 ρ(1 ρ2)κ2 (1 ρ2)2   − 1 − 2 −  Note that the asymptotic variance of the estimators of the drift parameter goes to zero, as the correlation parameter ρ goes to one.

2

Several papers have studied other aspects of small diffusion asymptotics for estimators of parameters in diffusion models. First estimation of the parameter α based on a continuously observed sample path of the diffusion process was considered by Kutoyants (1994). Semiparametric estimation for the same type of data was studied later by Kutoyants (1998) and Iacus & Kutoyants (2001). Information criteria were investigated by Uchida & Yoshida (2004b). Uchida 70 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES (2004) and Uchida (2008) studied approximations to martingale estimating functions for discretely sampled diffusions under small diffusion asymptotics. Martingale estimating functions were studied by Sørensen (2000b) under an extreme type of small diffusion asymptotics where n is fixed.

1.9 Non-Markovian models

In this section we consider estimating functions that can be used when the ob- served process is not a Markov process. In this situation, it is usually not easy to find a tractable martingale estimating function. For instance a simple estimat- ing function of the form (1.41) is not a martingale. To obtain a martingale, the conditional expectation given X(i 1)∆ in (1.41) must be replaced by the condi- tional expectation given all previous− observations, which can only very rarely be found explicitly, and which it is rather hopeless to find by simulation. In- stead we will consider a generalization of the martingale estimating functions, called the prediction-based estimating functions, which can be interpreted as approximations to martingale estimating functions.

To clarify our thoughts, we will consider a concrete model type. Let the D- dimensional process X be the stationary solution to the stochastic differential equation

dXt = b(Xt; θ)dt + σ(Xt; θ)dWt, (1.128) where b is D-dimensional, σ is a D D-matrix, and W a D-dimensional standard Wiener process. As usual the parameter× θ varies in a subset Θ of IRp. However, we do not observed X directly. What we observe is

Yi = k(Xti )+ Zi, i =1,...,n, (1.129) D d where k maps IR into IR (d

1.9.1 Prediction-based estimating functions

In the following we will outline the method of prediction-based estimating functions introduced in Sørensen (2000a). Assume that fj , j = 1,...,N, are s+1 2 functions that map IR Θ into IR such that Eθ(fj (Ys+1,...,Y1; θ) ) < θ × θ ∞ for all θ Θ. Let i 1,j be a closed linear subset of the L2-space, Li 1, of all ∈ P − θ − functions of Y1,...,Yi 1 with finite variance under Pθ. The set i 1,j can be − P − interpreted as a set of predictors of fj(Yi,...,Yi s; θ) based on Y1,...,Yi 1. − − NON-MARKOVIAN MODELS 71 A prediction-based estimating function has the form

n N (i 1) (i 1) Gn(θ)= Πj − (θ) fj (Yi,...,Yi s; θ) π˘j − (θ) , − − i=s+1 j=1 X X h i (i 1) where Πj − (θ) is a p-dimensional vector, the coordinates of which belong to θ (i 1) θ i 1,j , and π˘j − (θ) is the minimum mean square error predictor in i 1,j of P − θ P − fj(Yi,...,Yi s; θ) under Pθ. When s =0 and i 1,j is the set of all functions − (i P1) − of Y1,...,Yi 1 with finite variance, then π˘j − (θ) is the conditional expec- − tation under Pθ of fj (Yi; θ) given Y1,...,Yi 1, so in this case we obtain a martingale estimating function. Thus for a Markov− process, a martingale es- timating function of the form (1.41) is a particular case of a prediction-based estimating function.

θ The minimum mean square error predictor in i 1,j of fj(Yi,...,Yi s; θ) is θ P − − the projection in Li 1 of fj (Yi,...,Yi s; θ) onto the subspace i 1,j . There- (i 1) − − P − fore π˘j − (θ) satisfies the normal equation

(i 1) (i 1) Eθ πj − fj(Yi,...,Yi s; θ) π˘j − (θ) =0 (1.130) − −  h i (i 1) θ for all πj − i 1,j . This implies that a prediction-based estimating func- tion satisfies that∈ P −

Eθ (Gn(θ))=0. (1.131) We can interpret the minimum mean square error predictor as an approxima- tion to the conditional expectation of fj (Yi,...,Yi s; θ) given X1,...,Xi 1, − − which is the projection of fj(Yi,...,Yi s; θ) onto the subspace of all functions − of X1,...,Xi 1 with finite variance. − To obtain estimators that can relatively easily be calculated in practice, we will θ from now on restrict attention to predictor sets, i 1,j , that are finite dimen- P − r sional. Let hjk, j = 1,...,N, k = 0,...,qj be functions from IR into IR (r s), and define (for i r +1) random variables by ≥ ≥ (i 1) Z − = hjk(Yi 1, Yi 2,...,Yi r). jk − − − (i 1) 2 We assume that Eθ((Z − ) ) < for all θ Θ, and let i 1,j denote the jk − (i 1) ∞(i 1) ∈ P subspace spanned by − − . We set and make the natural Zj0 ,...,Zjqj hj0 =1 assumption that the functions hj0,...,hjqj are linearly independent. We write (i 1) T − T the elements of i 1,j in the form a Zj , where a = (a0,...,aqj ) and P − T (i 1) (i 1) (i 1) − − − Zj = Zj0 ,...,Zjqj   are (qj + 1)-dimensional vectors. With this specification of the predictors, the 72 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES estimating function can only include terms with i r +1: ≥ n N (i 1) (i 1) Gn(θ)= Πj − (θ) fj (Yi,...,Yi s; θ) π˘j − (θ) . (1.132) − − i=r+1 j=1 X X h i

(i 1) It is well-known that the minimum mean square error predictor, π˘j − (θ), is found by solving the normal equations (1.130). Define Cj (θ) as the covariance matrix of (r) (r) T under , and as the vector for which the (Zj1 ,...,Zjqj ) Pθ bj (θ) ith coordinate is (r) bj(θ)i = Covθ(Zji ,fj(Yr+1,...,Yr+1 s; θ)), (1.133) − i =1,...,qj . Then we have

(i 1) T (i 1) π˘j − (θ)=˘aj (θ) Zj − , (1.134) T T where a˘j (θ) = (˘aj0(θ), a˘j (θ) ) with ∗ 1 a˘j (θ)= Cj (θ)− bj (θ) (1.135) ∗ and

qj a˘ (θ)= E (f (Y ,...,Y ; θ)) a˘ (θ)E (Z(r)). (1.136) j0 θ j s+1 1 − jk θ jk Xk=1 That Cj (θ) is invertible follows from the assumption that the functions hjk are linearly independent. If fj(Yi,...,Yi s; θ) has mean zero under Pθ for all θ Θ, we need not include a constant in− the space of predictors, i.e. we need ∈ (i 1) (i 1) only the space spanned by − − . Zj1 ,...,Zjqj

j Example 1.9.1 An important particular case when d = 1 is fj (y) = y , j = (i 1) 1,...,N. For each i = r +1,...,n and j = 1,...,N, we let Z − k = { jk | κ (i 1) 0,...,qj be a subset of Yi ℓ ℓ = 1,...,r,κ = 0,...,j , where Zj0− is } { − | 2N } always equal to 1. Here we need to assume that Eθ(Yi ) < for all θ Θ. (i 1) ∞ ∈ To find π˘j − (θ), j = 1,...,N, by means of (1.135) and (1.136), we must calculate moments of the form E (Y κY j ), 0 κ j N, k =1,...,r. (1.137) θ 1 k ≤ ≤ ≤ To avoid the matrix inversion in (1.135), the vector of coefficients a˘j can be found by means of the N-dimensional Durbin-Levinson applied to 2 N the process (Yi, Yi ,...,Yi ) i IN, see Brockwell & Davis (1991). Suppose the diffusion{ process X is exponentially} ∈ ρ-mixing, see Doukhan (1994) for a definition. This is for instance the case for a Pearson diffusion (see Subsection 1.3.7) or for a one-dimensional diffusion that satisfies Condition 1.5.1. Then the observed process Y inherits this property, which implies that constants NON-MARKOVIAN MODELS 73 j j λk K > 0 and λ > 0 exist such that Covθ(Y1 , Yk ) Ke− . Therefore a small value of r can usually be used. | | ≤ In many situations it is reasonable to choose N =2 with the following simple predictor sets where q1 = r and q2 = 2r. The predictor sets are generated by (i 1) (i 1) (i 1) 2 Zj0− = 1, Zjk− = Yi k, k = 1,...,r,j = 1, 2 and Z2k− = Yi+r k, k = r +1,..., 2r. In this− case the minimum mean square error predictor− of Yi can be found using the Durbin-Levinson algorithm for real processes, while 2 the predictor of Yi can be found by applying the two-dimensional Durbin- 2 Levinson algorithm to the process (Yi, Yi ). Including predictors in the form of lagged terms Yi kYi k l for a number of lags l’s might also be of relevance. − − − We illustrate the use of the Durbin-Levinson algorithm in the simplest possible (i 1) (i 1) case, where N = 1, f(x) = x, Z0 − = 1, Z − = Yi k, k = 1,...,r. k − We suppress the superfluous j in the notation. Let Kℓ(θ) denote the covari- ance between Y1 and Yℓ+1 under Pθ, and define φ1,1(θ)) = K1(θ)/K0(θ) and v0(θ)= K0(θ). Then the Durbin-Levinson algorithm works as follows

ℓ 1 − 1 φℓ,ℓ(θ)= Kℓ(θ) φℓ 1,k(θ)Kℓ k(θ) vℓ 1(θ)− , − − − ! − kX=1

φℓ,1(θ) φℓ 1,1(θ) φℓ 1,ℓ 1(θ) . − . − .− . = . φ (θ) .  .   .  − ℓ,ℓ  .  φℓ,ℓ 1(θ)) φℓ 1,ℓ 1(θ)) φℓ 1,1(θ))  −   − −   −        and 2 vℓ(θ)= vℓ 1(θ) 1 φℓ,ℓ(θ) . − − The algorithm is run for ℓ =2,...,r. Then 

a˘ (θ) = (φr,1(θ),...,φr,r(θ)), ∗ while a˘0 can be found from (1.136), which here simplifies to r a˘0(θ)= Eθ(Y1) 1 φr,k(θ) . − ! kX=1 (i 1) 2 The quantity vr(θ) is the prediction error Eθ (Yi π˘ − ) . Note that if we want to include a further lagged value of Y in the− predictor, we just iterate the algorithm once more.  2

We will now find the optimal prediction-based estimating function of the form (1.132) in the sense explained in Section 1.11. First we express the estimating 74 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES (i 1) function in a more compact way. The ℓth coordinate of the vector Πj − (θ) can be written as

qj (i 1) (i 1) πℓ,j− (θ)= aℓjk(θ)Zjk− , ℓ =1,...,p. Xk=0 With this notation, (1.132) can be written in the form n (i) Gn(θ)= A(θ) H (θ), (1.138) i=r+1 X where a (θ) a (θ) a (θ) a (θ) 110 · · · 11q1 ··· ··· 1N0 · · · 1NqN . . . . A(θ)=  . . . .  , a (θ) a (θ) a (θ) a (θ)  p10 · · · p1q1 ··· ··· pN0 · · · pNqN  and   (i) (i 1) (i 1) H (θ)= Z − F (Yi,...,Yi s; θ) π˘ − (θ) , (1.139) − − T (i 1) (i 1) (i 1) T with F = (f1,...,fN ) , π˘ − (θ)=(˘π1 − (θ),..., π˘N− (θ)) , and

(i 1) Z − 0 0 1 q1 · · · q1  (i 1)  (i 1) − Z = 0q2 Z2 0q2 . (1.140) −  · · ·   . . .   . . .   (i 1)   0 0 Z −   qN qN · · · N    Here 0qj denotes the qj -dimensional zero-vector. When we have chosen the (i) functions fj and the predictor spaces, the quantities H (θ) are completely determined, whereas we are free to choose the matrix A(θ) in an optimal way, i.e. such that the asymptotic variance of the estimators is minimized.

We will find en explicit expression for the optimal weight matrix, A∗(θ), under the following condition, in which we need one further definition: T a˘(θ)=(˘a10(θ),..., a˘1q1 (θ),..., a˘N0(θ),... a˘NqN (θ)) , (1.141) where the quantities a˘jks define the minimum mean square error predictors, cf. (1.134).

Condition 1.9.2 (1) The function F (y1,...,ys+1; θ) and the coordinates of a˘(θ) are continu- ously differentiable functions of θ.

(2) p p¯ = N + q + + q . ≤ 1 · · · N NON-MARKOVIAN MODELS 75

(3) The p¯ p-matrix ∂ T α˘(θ) has rank p. × θ

(4) The functions 1,f1,...,fN are linearly independent (for fixed θ) on the support of the conditional distribution of (Yi,...,Yi s) given (Xi 1,...,Xi r). − − − (5) The p p-matrix × T (i 1) U(θ) = Eθ Z − ∂θT F (Yi,...,Yi s; θ) (1.142) − exists.  

If we denote the optimal prediction-based estimating function by Gn∗ (θ), then T T E G (θ)G∗ (θ) = (n r)A(θ)M¯ (θ)A∗ (θ) , θ n n − n n where  (r+1) (r+1) T M¯ n(θ)= Eθ H (θ)H (θ) (1.143) n r 1  − − (n r k) + − − E H(r+1)(θ)H(r+1+k)(θ)T (n r) θ Xk=1 − n   (r+1+k) (r+1) T + Eθ H (θ)H (θ) ,

n  (i) o which is the covariance matrix of i=r+1 H (θ)/√n r. The sensitivity function (1.166) is given by − P T S (θ) = (n r)A(θ) U(θ) D(θ)∂ T a˘(θ) , Gn − − θ where D(θ) is the p¯ p¯-matrix   × (i 1) (i 1) T D(θ)= Eθ Z − (Z − ) . (1.144)

  T It follows from Theorem 1.11.1 that An∗ (θ) is optimal if Eθ Gn(θ)Gn∗ (θ) = ¯ SGn (θ). Under Condition 1.9.2 (4) the matrix Mn(θ) is invertible, see Sørensen (2000a), so it follows that  T 1 A∗ (θ) = (U(θ) ∂ a˘(θ) D(θ))M¯ (θ)− , (1.145) n − θ n so that the estimating function n (i 1) (i 1) Gn∗ (θ)= An∗ (θ) Z − F (Yi,...,Yi s; θ) π˘ − (θ) , (1.146) − − i=s+1 X   is Godambe optimal. When the function F does not depend on θ, the expres- sion for An∗ (θ) simplifies slightly as in this case U(θ)=0.

Example 1.9.3 Consider again the type of prediction-based estimating func- tion discussed in Example 1.9.1. In order to calculate (1.143), we need mixed 76 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES moments of the form

k1 k2 k3 k4 Eθ[Yt1 Yt2 Yt3 Yt4 ], (1.147) for t1 t2 t3 t4 and k1 + k2 + k3 + k4 4N, where ki, i = 1,..., 4 are non-negative≤ ≤ integers.≤ ≤ 2

1.9.2 Asymptotics

A prediction-based estimating function of the form (1.138) gives consistent and asymptotically normal estimators under the following condition, where θ0 is the true parameter value.

Condition 1.9.4 (1) The diffusion process X is stationary and geometrically α-mixing.

(2) There exists a δ > 0 such that 2+δ (r) Eθ0 Z fj (Xr+1,...,Xr+1 s; θ0) < jk − ∞   and 2+δ E Z(r)Z(r) < , θ0 jk jℓ ∞   for j =1,...,N, k,ℓ =0,...q j .

(3) The function F (y1,...,ys+1; θ) and the components of A(θ) and a˘(θ), given by (1.141) are continuously differentiable functions of θ.

(4) The matrix W = A(θ0)(U(θ0) D(θ0)∂θT a˘(θ0)) has full rank p. The matrices U(θ) and D(θ) are given by−(1.142) and (1.144). (5)

(i 1) A(θ) Eθ0 Z − F (Yi,...,Yi s; θ) D(θ0)∂θT a˘(θ)) =0 − − 6 for all θ = θ.    6 0 Condition 1.9.4 (1) and (2) ensures that the central limit theorem (1.5) holds and that M¯ (θ ) M(θ ), where n 0 → 0 (r+1) (r+1) T M(θ)= Eθ H (θ)H (θ)   NON-MARKOVIAN MODELS 77

∞ (r+1) (r+1+k) T + Eθ H (θ)H (θ) kX=1 n   (r+1+k) (r+1) T + Eθ H (θ)H (θ) .  o The asymptotic covariance matrix in (1.5) is V (θ) = A(θ)M(θ)A(θ)T . The concept of geometric α-mixing was explained in Subsection 1.5.1, where also conditions for geometric α-mixing were discussed. It is not difficult to see that if the basic diffusion process X is geometrically α-mixing, then the observed process Y inherits this property. We only need to check Condition 1.2.1 with θ¯ = θ0 to obtain asymptotic results for prediction-based estimators. The con- dition (1.6) is satisfied because of (1.131). It is easy to see that Condition 1.9.4 (3) and (4) implies that θ g(y ,...,y ) is continuously differentiable 7→ 1 r+1 and that g as well as ∂θT g are locally dominated integrable under Pθ0 . Finally, for a prediction-based estimating function, the condition (1.9) is identical to Condition 1.9.4 (5). Therefore it follows from Theorem 1.2.2 that a consistent Gn–estimator θˆn exists and is the unique Gn–estimator on any bounded subset of Θ containing θ0 with probability approaching one as n . The estimator satisfies that → ∞ 1 T T 1 √n(θˆ θ ) D N 0, W − A(θ )M(θ )A(θ ) W − n − 0 −→ p 0 0 0 as n .   → ∞

1.9.3 Measurement errors

Suppose a one-dimensional diffusion has been observed with measurement er- rors so that the data are

Yi = Xti + Zi, i =1,...,n, where X solves (1.11), and the measurement errors Zi are independent and identically distributed and independent of X. Since the observed process (Yi) is not a Markov process, it is usually not possible to find a feasible martingale estimating function. Instead we can use a prediction-based estimating function of the type considered Example 1.9.1. To find the minimum mean square error predictor, we must find mixed moments of the form (1.137). By the binomial formula,

k1 k2 k1 k2 Eθ(Y1 Y2 )= Eθ (Xt1 + Z1) (Xt2 + Z2)

k1 k2 k1 k2 i1 i2 k1 i1 k2 i2 = E (X X )E (Z − )E (Z − ). i i θ t1 t2 θ 1 θ 2 i =0 i =0 1 2 X1 X2    Note that the distribution of the measurement error Zi can depend on com- ponents of the unknown parameter θ. We need to find the mixed moments 78 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

i1 i2 Eθ(Xt1 Xt2 ),(t1 < t2). If expressions for the moments and conditional mo- ments of Xt are available, these mixed moments can be found explicitly. As an example, consider the Pearson diffusions discussed in Subsection 1.3.7, for which the conditional moments are given by (1.86). Thus E (Xi1 Xi2 ) = E (Xi1 E (Xi2 X )) (1.148) θ t1 t2 θ t1 θ t2 | t1 i2 i2 λℓ(t2 t1) i1+k = qi2,k,ℓe− − Eθ(Xt1 ), ! kX=0 Xℓ=0 i1+k where Eθ(Xt1 ) can be found by (1.87), provided, of course, that it exists. For stationary, ergodic one-dimensional diffusions, the polynomial moments can usually be found because we have an explicit expression for the marginal density functions, at least up to a multiplicative constant, cf. (1.15). In order to find the optimal prediction-based estimating functions of the form considered in Example 1.9.1, we must find the mixed moments of the form (1.147), which can be calculated in a similar way.

1.9.4 Integrated diffusions and hypoelliptic stochastic differential equations

Sometimes a diffusion process, X, cannot be observed directly, but data of the form 1 i∆ Yi = Xs ds, i =1,...,n, (1.149) ∆ (i 1)∆ Z − are available for some fixed ∆. Such observations might be obtained when the process X is observed after passage through an electronic filter. Another example is provided by ice-core records. The isotope ratio 18O/16O in the ice is a proxy for paleo-temperatures. The average isotope ratio is measured in pieces of ice, each of which represent a time interval. The variation of the paleo-temperature can be modelled by a stochastic differential equation, and hence the ice-core data can be modelled as an integrated diffusion process, see Ditlevsen, Ditlevsen & Andersen (2002). Estimation based on this type of data was considered by Gloter (2000), Bollerslev & Wooldridge (1992), Ditlevsen & Sørensen (2004), Gloter (2006), and Baltazar-Larios & Sørensen (2009). Non-parametric inference was studied in Comte, Genon-Catalot & Rozenholc (2009). The model for data of the type (1.149) is a particular case of (1.128) with X b(X ; θ) σ(X ; θ) d 1,t = 1,t dt + 1,t dW , X X 0 t  2,t   1,t    with X2,0 = 0, where W and the two components are one-dimensional, and NON-MARKOVIAN MODELS 79 only the second coordinate, X2,t, is observed.The second coordinate is not sta- tionary, but if the first coordinate is a stationary process, then the observed in- crements Yi = (X2,i∆ X2,(i 1)∆)/∆ form a stationary sequence. A stochas- tic differential equation− of the− form (1.150) is called hypoelliptic. Hypoelliptic stochastic differential equations are, for instance, used to model molecular dy- namics, see e.g. Pokern, Stuart & Wiberg (2009). The unobserved component, X1,t, can more generally be multivariate and have coefficients that depend on the observed component X2,t too. The observed smooth component can also be multivariate. The drift is typically minus the derivative of a potential. A simple example is the stochastic dX = (β X + β X ) dt + γdW 1,t − 1 1,t 2 2,t t dX2,t = X1,t dt,

β1,β2,γ > 0. Observations of the form (1.149), or more generally discrete time observations of the smooth componentsof a hypoelliptic stochastic differ- ential equation, do not form a Markov process, so usually a feasible martingale estimating function is not available, but prediction-based estimating functions can be used instead. For instance, the stochastic harmonic oscillator above is a Gaussian process. Therefore all the mixed moments needed in the optimal prediction-based estimating function of the form considered in Example 1.9.1 can be found explicitly.

In the following we will again denote the basic diffusion by X (rather than X1), and assume that the data are given by (1.149). Suppose that 4N’th moment of Xt is finite. The moments (1.137) and (1.147) can be calculated by

k1 k2 k3 k4 E Y1 Yt1 Yt2 Yt3 =

h E[Xv Xiv Xu Xu Xs Xs Xr Xr ] dt A 1 · · · k1 1 · · · k2 1 · · · k3 1 · · · k4 k1+k2+k3+k4 R ∆ where 1 t t t , A = [0 , ∆]k1 [(t 1)∆ , t ∆]k2 [(t ≤ 1 ≤ 2 ≤ 3 × 1 − 1 × 2 − 1)∆ , t ∆]k3 [(t 1)∆ , t ∆]k4 , and dt = dr dr ds ds du 2 × 3− 3 k4 · · · 1 k3 · · · 1 k2 · · · du1 dvk1 dv1. The domain of integration can be reduced considerably by · ·arguments, · but the point is that we need to calculate mixed moments κ1 κk of the type E(Xt1 Xtk ), where t1 < < tk. For the Pearson diffusions discussed in Subsection· · · 1.3.7, these mixed· · · moments can be calculated by a simple iterative formula obtained from (1.86) and (1.87), as explained in the previous subsection. Moreover, for the Pearson diffusions, E(Xκ1 Xκk ) t1 · · · tk depends on t1,...,tk through sums and products of exponential functions, cf. (1.86) and (1.148). Therefore the integral above can be explicitly calculated, and thus explicit optimal estimating functions of the type considered in Exam- ple 1.9.1 are available for observations of integrated Pearson diffusions.

Example 1.9.5 Consider observationof an integrated square root process (1.37) 80 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES 2 and a prediction-based estimating function with f1(x) = x and f2(x) = x (i 1) (i 1) with predictors given by π1 − = a1,0 + a1,1Yi 1 and π2 − = a2,0. Then the minimum mean square error predictors are − (i 1) π˘1 − (Yi 1; θ) = µ (1 a˘(β))+˘a(β)Yi 1, − − − (i 1) 2 2 3 2 β∆ π˘ − (θ) = α + ατ β− ∆− (e− 1+ β∆) 2 − with β∆ 2 (1 e− ) a˘(β)= − . 2(β∆ 1+ e β∆) − − The optimal prediction-based estimating function is

n 1 n 0 (i 1) 2 (i 1) Yi 1 [Yi π¯1 − (Yi 1; θ)] + 0 [Yi π¯2 − (θ)],  −  − −   − i=1 0 i=1 1 X X from which we obtain the estimators   1 n a(βˆ)Y Y αˆ = Y + n − 1 n i ˆ i=1 (n 1)(1 a(β)) X − − n n n ˆ ˆ 2 Yi 1Yi =α ˆ(1 a(β)) Yi 1 + a(β) Yi 1 − − − − i=2 i=2 i=2 X X X βˆ3∆2 n Y 2 αˆ2 σˆ2 = i=2 i − . (n 1)ˆα(e βˆ∆ 1+ βˆ∆) − P − −  The estimators are explicit apart from βˆ, which can easily be found numerically by solving a non- in one . For details, see Ditlevsen & Sørensen (2004). 2

1.9.5 Sums of diffusions

An autocorrelation function of the form ρ(t)= φ exp( β t)+ . . . + φ exp( β t), (1.150) 1 − 1 D − D D where i=1 φi = 1 and φi,βi > 0, is found in many observed time series. Examples are financial time series, see Barndorff-Nielsen & Shephard (2001), and turbulence,P see Barndorff-Nielsen, Jensen & Sørensen (1990) and Bibby, Skovgaard & Sørensen (2005). A simple model with autocorrelation function of the form (1.150) is the sum of diffusions Yt = X1,t + . . . + XD,t, (1.151) NON-MARKOVIAN MODELS 81 where the D diffusions dX = β (X α )+ σ (X )dW , i =1,...,D, i,t − i i,t − i i i,t i,t are independent. In this case Var(X ) φ = i,t . i Var(X )+ + Var(X ) 1,t · · · D,t Sums of diffusions of this type with a pre-specified marginal distribution of Y were considered by Bibby & Sørensen (2003) and Bibby, Skovgaard & Sørensen (2005), while Forman & Sørensen (2008) studied sums of Pearson diffusions. The same type of autocorrelation function is obtained for sums of independent Ornstein-Uhlenbeck processes driven by L´evy processes. This class of models was introduced and studied in Barndorff-Nielsen, Jensen & Sørensen (1998).

2 Example 1.9.6 Sum of square root processes. If σi (x)=2βibx and αi = κib for some b > 0, then the stationary distribution of Yt is a gamma-distribution with shape parameter κ1 + + κD and scale parameter b. The weights in the autocorrelation function are·φ · · = κ /(κ + + κ ). i i 1 · · · D 2

For sums of the Pearson diffusions presented in Subsection 1.3.7, we have ex- plicit formulae that allow calculation of (1.137) and (1.147), provided these mixed moments exists. Thus for sums of Pearson diffusions we have explicit optimal prediction-based estimating functions of the type considered in Exam- ple 1.9.1. By the multinomial formula, κ ν E(Yt1 Yt2 )= κ ν E(Xκ1 Xν1 ) ...E(XκD XνD ) 1,t1 1,t2 D,t1 D,t2 κ1,...,κD ν1,...,νD XX    where κ κ! = κ ,...,κ κ ! κ !  1 D 1 · · · D is the multinomial coefficient, and where the first sum is over 0 κ ,...,κ ≤ 1 D such that κ1 +...κD = κ, and the second sum is analogousfor the νis. Higher order mixed moments of the form (1.147) can be found by a similar formula with four sums and four multinomial coefficients. Such formulae may appear daunting, but are easy to program. For a Pearson diffusion, mixed moments κ1 κk of the form E(Xt1 Xtk ) can be calculated by a simple iterative formula obtained from (1.86)· and· · (1.87), as explained in Subsection 1.9.3.

Example 1.9.7 Sum of two skew t-diffusions. If 2 1 2 2 σ (x)=2β (ν 1)− x +2ρ√ν x +(1+ ρ )ν , i =1, 2, i i i − { i } 82 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES the stationary distribution of Xi,t is a skew t-distribution, ν ν Var(Y )=(1+ ρ2) 1 + 2 , ν 2 ν 2  1 − 2 −  1 1 and in (1.150) the weights are given by φi = νi(νi 2)− / ν1(ν1 2)− + ν (ν 2) 1 . To simplify the exposition we assume− that the{ correlation− pa- 2 2 − − } rameters β1, β2, φ1, and φ2 are known or have been estimated in advance, for instance by fitting (1.150) with D = 2 to the empirical autocorrelation func- tion. We will find the optimal estimating function in the simple case where pre- 2 (i 1) dictions of Yi are made based on predictors of the form π − = a0 +a1Yi 1. The estimating equations take the form −

n 2 2 Yi σ ζYi 1 2 − 2 − − 2 =0, (1.152) Yi 1Y σ Yi 1 ζY i=2 i i 1 X  − − − − −  with ν ν σ2 = Var(Y )=(1+ ρ2) 1 + 2 , i ν 2 ν 2  1 2  2 − − Cov(Yi 1, Yi ) √ν1 β1∆ √ν2 β2∆ ζ = − =4ρ φ e− + φ e− . Var(Y ) ν 3 1 ν 3 2 i  1 − 2 −  Solving equation (1.152) for ζ and σ2 we get 1 n 2 1 n 1 n 2 n 1 i=2 Yi 1Yi ( n 1 i=2 Yi 1)( n 1 i=2 Yi ) ζˆ = − − − − − − , 1 n 2 1 n 2 n 1 i=2 Yi 1 ( n 1 i=2 Yi 1) P − − −P− − P 2 1 n 2 + ˆ 1 n σˆ = n 1 i=2 Yi Pζ n 1 i=2 Yi 1. P − − − In order to estimatePρ we restate ζ asP 9(1 + ρ2) φ σ2 2 1 β1∆ ζ = 32(1 + ρ ) ρ 2 − 2 φ1e− · · (p3(1 + ρ ) φ1σ p − 2 2 9(1 + ρ ) φ2σ β2∆ + − φ e− 3(1 + ρ2) φ σ2 2 p − 2 ) and insert σˆ2 for σ2. Thus, we get a one-dimensional estimating equation, ζ(β, φ, σˆ2,ρ) = ζˆ, which can be solved numerically. Finally by inverting 2 2 1+ρ νi 2φiσˆ φ = 2 , we find the estimates νˆ = 2 2 , i =1, 2. i σ νi 2 i φiσˆ (1+ˆρ ) − − 2

A more complex model is obtained if the observations are integrals of the pro- cess Y given by (1.151). In this case the data are 1 i∆ 1 i∆ i∆ Zi = Ys ds = X1,sds + + XD,sds , (1.153) ∆ (i 1)∆ ∆ (i 1)∆ · · · (i 1)∆ Z − Z − Z −  NON-MARKOVIAN MODELS 83 i = 1,...,n. Also here the moments of form (1.137) and (1.147), and hence optimal prediction-basedestimating functions, can be found explicitly for Pear- son diffusions. This is because each of the observations Zi is a sum of pro- k1 k2 k3 k4 cesses of the type consideredin Subsection 1.9.4.To calculate E(Zt1 Zt2 Zt3 Zt4 ), first apply the multinomial formula as above to express this quantity in terms of moments of the form ℓ1 ℓ2 ℓ3 ℓ4 , where E(Yj,t1 Yj,t2 Yj,t3 Yj,t4 ) 1 i∆ Yj,i = Xj,s ds. ∆ (i 1)∆ Z − Now proceed as in Subsection 1.9.4.

1.9.6 Stochastic volatility models

A stochastic volatility model is a generalization of the Black-Scholes model 2 for the logarithm of an asset price dXt = (κ + βσ )dt + σdWt, that takes into account the empirical finding that the volatility σ2 varies randomly over time:

dXt = (κ + βvt)dt + √vtdWt. (1.154)

Here the volatility vt is a stochastic process that cannot be observed directly. If the data are observations at the time points ∆i, i = 0, 1, 2,...,n, then the returns Yi = Xi∆ X(i 1)∆ can be written in the form − −

Yi = κ∆+ βSi + SiAi, (1.155) where p i∆ Si = vtdt, (1.156) (i 1)∆ Z − and where the Ai’s are independent, standard normal distributed random vari- ables. Prediction-based estimating functions for stochastic volatility models were considered in detail in Sørensen (2000a). Here we consider the case where the volatility process v is a sum of inde- pendent Pearson diffusions with state-space (0, ) (the cases 2, 4 and 5). Barndorff-Nielsen & Shephard (2001) demonstrated∞ that an autocorrelation function of the type (1.150) fits empirical autocorrelation functions of volatility well, while an autocorrelation function like that of a single Pearson diffusion is too simple to obtain a good fit. Stochastic volatility models where the volatil- ity process is a sum of independent square root processes were considered by Bollerslev & Zhou (2002) and Bibby & Sørensen (2003). We assume that v and W are independent, so that the Ai and Si are independent. By the multinomial formula we find that { } { }

k1 k2 k3 k4 E Y1 Yt1 Yt2 Yt3 =   84 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

k12+k13/2 k22+k23/2 k32+k33/2 k42+k43/2 Kk11,...,k43 E S1 St1 St2 St3   X E Ak13 E Ak23 E Ak33 E Ak43 , · 1 t1 t2 t3         where the sum is over all non-negative integers kij , i = 1, 2, 3, 4, j = 1, 2, 3 such that ki1 + ki2 + ki3 = ki (i =1, 2, 3, 4), and where

Kk11,...,k43 =

k1 k2 k3 k4 k 1 k 2 (κ∆) · β · k , k , k k , k , k k , k , k k , k , k  11 12 13 21 22 23 31 32 33 41 42 43 ki3 with k j = k1j + k2j + k3j + k4j . The moments E(Ai ) are the well-known · moments of the standard normal distribution. When ki3 is odd, these moments are zero. Thus we only need to calculate the mixed moments of the form ℓ1 ℓ2 ℓ3 ℓ4 E(S1 St1 St2 St3 ), where ℓ1,...,ℓ4 are integers. When the volatility process is a sum of independent Pearson diffusions, Si of the same form as Zi in (1.153) (apart from the factor 1/∆), so we can proceed as in the previoussubsection to calculate the necessary mixed moments. Thus also for the stochastic volatility models defined in terms of Pearson diffusions, we can explicitly find optimal estimating functions based on prediction of powers of returns, cf. Example 1.9.1.

1.9.7 Compartment models

Diffusion compartment models are D-dimensional diffusions with linear drift, dX = [B(θ)X b(θ)] dt + σ(X ; θ)dW , (1.157) t t − t t where only a subset of the coordinates are observed. Here B(θ) is a D D- matrix, b(θ) is a D-dimensional vector, σ(x; θ) is a D D-matrix, and× W a D-dimensional standard Wiener process. Compartment× models are used to model the dynamics of the flow of a substance between different parts (com- partments) of, for instance, an ecosystem or the body of a human being or an animal. The process Xt is the concentration in the compartments, and flow from a given compartment into other compartments is proportional to the con- centration in the given compartment, but modified by the random perturbation given by the diffusion term. The vector b(θ) represents input to or output from the system, for instance infusion or degradation of the substance. The compli- cation is that only a subset of the compartments can be observed, for instance the first compartment, in which case the data are Yi = X1,ti .

Example 1.9.8 The two-compartment model given by β β 0 τ 0 B = 1 2 , b = , σ = 1 , −β (β + β ) 0 0 τ  1 − 1 2     2  GENERAL ASYMPTOTICS RESULTS FOR ESTIMATING FUNCTIONS 85 where all parameters are positive, was used by Bibby (1995) to model how a radioactive tracer moved between the water and the biosphere in a certain ecosystem. Samples could only be taken from the water, the first compartment, so Yi = X1,ti . Likelihood inference, which is feasible because the model is Gaussian, was studied by Bibby (1995). All mixed moments of the form (1.137) and (1.147) can be calculated explicitly, again because the model is Gaussian. Therefore also explicit optimal prediction-based estimating func- tions of the type considered in Example 1.9.1 are available to estimate the parameters and were studied by D¨uring (2002). 2

Example 1.9.9 A non-Gaussian diffusion compartment model is obtained by the specification σ(x, θ) = diag(τ1√x1,...,τD√xD). This multivariate ver- sion of the square root process was studied by D¨uring (2002), who used meth- ods in Down, Meyn & Tweedie (1995) to show that the D-dimensional pro- cess is geometrically α-mixing and established the asymptotic normality of prediction-based estimators of the type considered in Example 1.9.1. As in the previous example, only the first compartment is observed, i.e. the data are

Yi = X1,ti . For the multivariate square root model, the mixed moments (1.137) and (1.147) must be calculated numerically. 2

1.10 General asymptotics results for estimating functions

In this section we review some general asymptotic results for estimators ob- tained from estimating functions for stochastic process models. Proofs can be found in Jacod & Sørensen (2009).

Suppose as a statistical model for the data X1,X2,...,Xn that they are ob- servations from a stochastic process. The corresponding probability measures (P ) are indexed by a p-dimensional parameter θ Θ. An estimating function θ ∈ is a function of the parameter and the observations, Gn(θ; X1,X2,...,Xn), with values in IRp. Usually we suppress the dependence on the observations in the notation and write Gn(θ). We get an estimator by solving the equa- tion (1.1) and call such an estimator a Gn-estimator. It should be noted that n might indicate more than just the sample size: the distribution of the data X1,X2,...,Xn might depend on n. For instance, the data might be obser- vations of a diffusion process at time points i∆n, i = 1,...,n, where ∆n decreases as n increases; see Sections 1.6 and 1.7. Another example is that the diffusion coefficient might depend on n; see Section 1.8. We will not necessarily assume that the data are observations from one of the 86 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES probability measures (Pθ)θ Θ. We will more generally denote the true prob- ability measure by P . If the∈ statistical model contains the true model, in the sense that there exists a θ0 Θ such that P = Pθ0 , then we call θ0 the true parameter value. ∈ A priory, there might be more than one solution or no solution at all to the estimating equation (1.1), so conditions are needed to ensure that a unique solution exists when n is sufficiently large. Moreover, we need to be careful when formally defining our estimator. In the following definition, δ denotes a “special” point, which we take to be outside Θ and Θ =Θ δ . δ ∪{ }

Definition 1.10.1 a) The domain of Gn-estimators (for a given n) is the set An of all observations x = (x1,...,xn) for which Gn(θ)=0 for at least one value θ Θ. ∈ b) A Gn-estimator, θˆn(x), is any function of the data with values in Θδ, such that for P –almost all observations we have either θˆ (x) Θ and G (θˆ (x), x)= n ∈ n n 0 if x A , or θˆ (x)= δ if x / A . ∈ n n ∈ n We usually suppress the dependence on the observations in the notation and write θˆn. The following theorem gives conditions which ensure that, for n large enough, the estimating equation (1.1) has a solution that converges to a particular pa- rameter value θ¯. When the statistical model contains the true model, the esti- mating function should preferably be chosen such that θ¯ = θ0. To facilitate the following discussion, we will refer to an estimator that converges to θ¯ in probability as a θ¯–consistent estimator, meaning that it is a (weakly) consistent estimator of θ¯. We assume that Gn(θ) is differentiable with respect to θ and denote by ∂ T G (θ) the p p-matrix, where the ijth entry is ∂ G (θ) . θ n × θj n i Theorem 1.10.2 Suppose the existence of a parameter value θ¯ int Θ (the interior of Θ), a connected neighbourhood M of θ¯, and a (possibly∈ random) function W on M taking its values in the set of p p matrices, such that the following holds: ×

P (i) Gn(θ¯) 0 (convergence in probability, w.r.t. the true measure P ) as n . → → ∞ (ii) Gn(θ) is continuously differentiable on M for all n, and P sup ∂θT Gn(θ) W (θ) 0. (1.158) θ M k − k → ∈ (iii) The matrix W (θ¯) is non-singular with P –probability one.

Then a sequence (θˆn) of Gn-estimators exists which is θ¯-consistent. Moreover GENERAL ASYMPTOTICS RESULTS FOR ESTIMATING FUNCTIONS 87 ˆ ¯ this sequence is eventually unique, that is if (θn′ ) is any other θ–consistent sequence of G –estimators, then P (θˆ = θˆ ) 0 as n . n n 6 n′ → → ∞ Note that the condition (1.158) implies the existence of a subsequence n { k} T such that ∂θ Gnk (θ) converges uniformly to W (θ) on M with probability one. Hence W is a continuous function of θ (up to a null set), and it follows from elementary that outside some P –null set there exists a unique continuously differentiable function G satisfying ∂θT G(θ) = W (θ) for all θ M and G(θ¯)=0. When M is a bounded set, (1.158) implies that ∈ P sup Gn(θ) G(θ) 0. (1.159) θ M | − | → ∈ This observation casts light on the result of Theorem 1.10.2. Since Gn(θ) can be made arbitrarily close to G(θ) by choosing n large enough, and since G(θ) has a root at θ¯, it is intuitively clear that Gn(θ) must have a root near θ¯ when n is sufficiently large. If we impose an identifiability condition, we can give a stronger result on any sequence of Gn–estimators. By B¯ǫ(θ) we denote the closed ball with radius ǫ centered at θ.

Theorem 1.10.3 Assume (1.159) for some subset M of θ containing θ¯, and that P inf G(θ) > 0 =1 (1.160) M B¯ǫ(θ¯) | |  \  for all ǫ> 0. Then for any sequence (θˆn) of Gn–estimators P (θˆ M B¯ (θ¯)) 0 (1.161) n ∈ \ ǫ → as n for every ǫ> 0 → ∞

If M =Θ, we see that any sequence (θˆn) of Gn–estimators is θ¯–consistent. If the conditions of Theorem 1.10.3 hold for any compact subset M of Θ, then a sequence (θˆn) of Gn–estimators is θ¯–consistent or converges to the boundary of Θ.

Finally, we give a result on the asymptotic distribution of a sequence (θˆn) of θ¯–consistent Gn–estimators.

Theorem 1.10.4 Assume the estimating function Gn satisfies the conditions of Theorem 1.10.2 and that there is a sequence of invertible matrices An such 1 that each entry of An− tends to zero,

AnGn(θ¯) Z ¯ 1 D ¯ , (1.162) An∂θT Gn(θ)A− −→ W0(θ)  n    88 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES and there exists a connected neighbourhood M of θ¯ such that

1 P sup An∂θT Gn(θ)An− W0(θ) 0. (1.163) θ M k − k → ∈ Here Z is a non-degenerate random variable, and W0 is a random function taking values in the set of p p-matrices satisfying that W (θ¯) is invertible. × 0 Under these conditions, we have for any θ¯–consistent sequence (θˆn) of Gn– estimators that 1 A (θˆ θ¯) D W (θ¯)− Z. (1.164) n n − −→ − 0 When Z is normal distributed with expectation zero and covariance matrix V , and when Z is independent of W0(θ¯), then the limit distribution is the normal variance-mixture with characteristic function

1 T 1 T 1 s E exp s W (θ¯)− V W (θ¯) − s . (1.165) 7→ − 2 0 0    If, moreover, W0(θ¯) is non-random, then the limit distribution is a normal dis- 1 T 1 tribution with expectation zero and covariance matrix W0(θ¯)− V W0(θ¯) − .

In the often occurring situation, where W0(θ¯) is non-random, joint conver- ¯ 1 ¯ gence of An∂θT Gn(θ)An− and AnGn(θ) is not necessary – marginal conver- gence of AnGn(θ¯) is enough.

1.11 Optimal estimating functions: general theory

The modern theory of optimal estimating functions dates back to the papers by Godambe (1960) and Durbin (1960), however the basic idea was in a sense already used in Fisher (1935). The theory was extended to stochastic processes by Godambe (1985), Godambe & Heyde (1987), Heyde (1988), and several others; see the references in Heyde (1997). Important particular instances are likelihood inference, the quasi-likelihood of Wedderburn (1974) and the gen- eralized estimating equations developed by Liang & Zeger (1986) to deal with problems of longitudinal data analysis, see also Prentice (1988) and Li (1997). A modern review of the theory of optimal estimating functionscan be found in Heyde (1997). The theory is very closely related to the theory of the general- ized method of moments developed independently in parallel in the economet- rics literature, where the foundation was laid by Hansen (1982), who followed Sagan (1958) by using selection matrices. Important extensions to the theory were made by Hansen (1985), Chamberlain (1987), Newey & West (1987), and Newey (1990); see also the discussion and references in Hall (2005). Par- ticular attention is given to the time series setting in Hansen (1985), Hansen (1993), West (2001), and Kuersteiner (2002). A discussion of links between the econometrics and literature can be found in Hansen (2001). In the OPTIMALESTIMATINGFUNCTIONS:GENERALTHEORY 89 following we present the theory as it was developed in the statistics literature by Godambe and Heyde. The general setup is as in the previous section. We will only consider unbiased estimating functions, i.e., estimating functions satisfying that Eθ(Gn(θ)) = 0 for all θ Θ. This natural requirement is also called Fisher consistency. It ∈ often implies condition (i) of Theorem 1.10.2 for θ¯ = θ0, which is an essential part of the condition for existence of a consistent estimator. Suppose we have a class of unbiased estimating functions. How do we choose the best member Gn in n? And in what sense are some estimating functions better than others? TheseG are the main problems in the theory of estimating functions. To simplify the discussion, let us first assume that p =1. The quantity

T SGn (θ)= Eθ(∂θ Gn(θ)) (1.166) is called the sensitivity function for Gn. As in the previous section, it is as- sumed that Gn(θ) is differentiable with respect to θ. A large absolute value of the sensitivity implies that the equation Gn(θ)=0 tends to have a solu- tion near the true parameter value, where the expectation of Gn(θ) is equal to zero. Thus a good estimating function is one with a large absolute value of the sensitivity.

Ideally, we would base the statistical inference on the likelihood function Ln(θ), and hence use the score function Un(θ) = ∂θ log Ln(θ) as our estimating function. This usually yields an efficient estimator. However, when Ln(θ) is not available or is difficult to calculate, we might prefer to use an estimating function that is easier to obtain and is in some sense close to the score function. Suppose that both Un(θ) and Gn(θ) have finite variance. Then it can be proven under usual regularity conditions that S (θ)= Cov (G (θ),U (θ)). Gn − θ n n Thus we can find an estimating function Gn(θ) that maximizes the absolute value of the correlation between Gn(θ) and Un(θ) by finding one that maxi- mizes the quantity 2 2 2 KGn (θ)= SGn (θ) /Varθ(Gn(θ)) = SGn (θ) /Eθ(Gn(θ) ), (1.167) which is known as the Godambe information. This makes intuitive sense: the ratio KGn (θ) is large when the sensitivity is large and when the variance of Gn(θ) is small. The Godambe information is a natural generalization of the

Fisher information. Indeed, KUn (θ) is the Fisher information. For a discussion of information quantities in a stochastic process setting, see Barndorff-Nielsen & Sørensen (1991) and Barndorff-Nielsen & Sørensen (1994). In a short while, we shall see that the Godambe information has a large sample interpretation too. An estimating function G is called Godambe-optimal in if n∗ ∈ Gn Gn K (θ) K (θ) (1.168) Gn∗ ≥ Gn 90 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES for all θ Θ and for all G . ∈ n ∈ Gn When the parameter θ is multivariate (p > 1), the sensitivity function is the p p-matrix × T SGn (θ)= Eθ(∂θ Gn(θ)). (1.169) For a multivariate parameter, the Godambe information is the p p-matrix × 1 T T − KGn (θ)= SGn (θ) Eθ Gn(θ)Gn(θ) SGn (θ), (1.170) and an optimal estimating function Gn∗ can be defined by (1.168) with the in- equality referring to the partial ordering of the set of positive semi-definite p p-matrices. Whether an Godambe-optimal estimating function exists and × whether it is unique depends on the class n. In any case, it is only unique up to multiplication by a regular matrix thatG might depend on θ. Specifically, if Gn∗ (θ) satisfies (1.168), then so does MθGn∗ (θ) where Mθ is an invertible deterministic p p-matrix. Fortunately, the two estimating functions give rise to the same estimator(s),× and we refer to them as versions of the same estimat- ing function. For theoretical purposes a standardized version of the estimating functions is useful. The standardized version of Gn(θ) is given by 1 G(s)(θ)= S (θ)T E G (θ)G (θ)T − G (θ). n − Gn θ n n n (s) The rationale behind this standardization is that Gn (θ) satisfies the second Bartlett-identity

(s) (s) T (s) E G (θ)G (θ) = E (∂ T G (θ)), (1.171) θ n n − θ θ n an identity usually satisfied by the score function. The standardized estimating (s) function Gn (θ) is therefore more directly comparable to the score function. Note that when the second Bartlett identity is satisfied, the Godambe informa- tion equals minus the sensitivity matrix.

An Godambe-optimal estimating function is close to the score function Un in an L2-sense. Suppose Gn∗ is Godambe-optimal in n. Then the standardized (s) G version G∗n (θ) satisfies the inequality

E (G(s)(θ) U (θ))T (G(s)(θ) U (θ)) θ n − n n − n  (s) T (s)  E (G∗ (θ) U (θ)) (G∗ (θ) U (θ)) ≥ θ n − n n − n for all θ Θ and for all G , see Heyde (1988). In fact, if  is a closed ∈ n ∈ Gn Gn subspace of the L2-space of all square integrable functions of the data, then the optimal estimating function is the orthogonal projection of the score function onto n. For further discussion of this Hilbert space approach to estimating functions,G see McLeish & Small (1988). The interpretation of an optimal es- timating function as an approximation to the score function is important. By choosing a sequence of classes that, as n , converges to a subspace Gn → ∞ OPTIMALESTIMATINGFUNCTIONS:GENERALTHEORY 91 containing the score function Un, a sequence of estimators that is asymptoti- cally fully efficient can be constructed. The following result by Heyde (1988) can often be used to find the optimal estimating function.

Theorem 1.11.1 If G satisfies the equation n∗ ∈ Gn 1 T 1 T SGn (θ)− Eθ Gn(θ)Gn∗ (θ) = SGn∗ (θ)− Eθ Gn∗ (θ)Gn∗ (θ) (1.172) for all θ Θ and for all Gn  n, then it is Godambe-optimal in n. When ∈ ∈ G G n is closed under addition, any Godambe-optimal estimating function Gn∗ satisfiesG (1.172) .

T The condition(1.172)can often be verified by showing that Eθ(Gn(θ)Gn∗ (θ) )= Eθ(∂θT Gn(θ)) for all θ Θ and for all Gn n. In such situations, Gn∗ satisfies− the second Bartlett-identity∈ , (1.171), so that∈ G T KGn∗ (θ)= Eθ Gn∗ (θ)Gn∗ (θ) .  Example 1.11.2 Suppose we have a number of functions hij (x1,...,xi; θ), j =1,...,N, i =1,...n satisfying that

Eθ(hij (X1,...,Xi; θ))=0. Such functions define relationships (dependent on θ) between an observation Xi and the previous observations X1,...,Xi 1 (or some of them) that are on average equal to zero. It is natural to use such− relationships to estimate θ by n solving the equations i=1 hij (X1,...,Xi; θ)=0. In order to estimate θ it is necessary that N p, but if N >p we have too many equations. The theory of optimal estimating≥ P functions tells us how to combine the N relations in an optimal way.

T Let hi denote the N-dimensional vector (hi1,...,hiN ) , and define an N- n dimensional estimating function by Hn(θ) = i=1 hi(X1,...,Xi; θ). First we consider the class of p-dimensional estimating functions of the form P Gn(θ)= An(θ)Hn(θ), where A (θ) is a non-random p N-matrix that is differentiable with respect n × to θ. By An∗ (θ) we denote the optimal choice of An(θ). It is not difficult to see that

SGn (θ)= An(θ)SHn (θ) and T T T Eθ Gn(θ)Gn∗ (θ) = An(θ)Eθ Hn(θ)Hn(θ) An∗ (θ) , where T . If we choose SHn (θ)= Eθ(∂θ Hn(θ))  T T 1 A∗ (θ)= S (θ) E H (θ)H (θ) − , n − Hn θ n n  92 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

then (1.172) is satisfied for all Gn n, so that Gn∗ (θ) = An∗ (θ)Hn(θ) is Godambe optimal. ∈ G

Sometimes there are good reasons to use functions hij satisfying that

Eθ(hij (X1,...,Xi; θ)hi′ j′ (X1,...,Xi′ ; θ))=0 (1.173) for all j, j = 1,...,N when i = i . For such functions the random variables ′ 6 ′ hij (X1,...,Xi; θ),i = 1, 2,... are uncorrelated, and in this sense the “new” random variation of hij (X1,...,Xi; θ) depends only on the innovation in the ith observation. This is for instance the case for martingale estimating func- tions, see (1.180). In this situation it is natural to consider the larger class of estimating functions given by n Gn(θ)= ai(θ)hi(X1,...,Xi; θ), i=1 X where ai(θ), i = 1,...n, are p N matrices that do not depend on the data and are differentiable with respect× to θ. Here n

T SGn (θ)= ai(θ)Eθ(∂θ hi(X1,...,Xi; θ)) i=1 X and T Eθ Gn(θ)Gn∗ (θ) = n  T T ai(θ)Eθ hi(X1,...,Xi; θ)hi(X1,...,Xi; θ) ai∗(θ) , i=1 X  where ai∗(θ) denotes the optimal choice of ai(θ). We see that with

ai∗(θ)= T T 1 E (∂ T h (X ,...,X ; θ)) E h (X ,...,X ; θ)h (X ,...,X ; θ) − − θ θ i 1 i θ i 1 i i 1 i the condition (1.172) is satisfied.  2

1.11.1 Martingale estimating functions

More can be said about martingale estimating functions, i.e. estimating func- tions Gn satisfying that

Eθ(Gn(θ) n 1)= Gn 1(θ), n =1, 2,..., |F − − where n 1 is the σ-field generated by the observations X1,...,Xn 1 (G0 = F − − 0 and 0 is the trivial σ-field). In other words, the stochastic process Gn(θ) : n = 1F, 2,... is a martingale under the model given by the parameter{ value θ. Since the score} function is usually a martingale (see e.g. Barndorff-Nielsen OPTIMALESTIMATINGFUNCTIONS:GENERALTHEORY 93 & Sørensen (1994)), it is natural to approximate it by families of martingale estimating functions. The well-developed martingale limit theory allows a straightforward discus- sion of the asymptotic theory, and motivates an optimality criterion that is particular to martingale estimating functions. Suppose the estimating function Gn(θ) satisfies the conditions of the central limit theorem for martingales and let θˆn be a solution of the equation Gn(θ)=0. Under the regularity conditions of the previous section, it can be proved that

1 2 G(θ) n− G¯ (θ)(θˆ θ ) D N(0, I ). (1.174) h i n n − 0 −→ p Here G(θ) is the quadratic characteristic of G (θ) defined by h in n n T G(θ) n = Eθ (Gi(θ) Gi 1(θ))(Gi(θ) Gi 1(θ)) i 1 , h i − − − − |F − i=1 X  and ∂θT Gn(θ) has been replaced by its compensator n ¯ Gn(θ)= Eθ (∂θT Gi(θ) ∂θT Gi 1(θ) i 1) , − − |F − i=1 X ¯ 1 Pθ using the extra assumption that Gn(θ)− ∂θT Gn(θ) Ip. Details can be found in Heyde (1988). We see that the inverse of the data-dep−→ endent matrix T 1 I (θ)= G¯ (θ) G(θ) − G¯ (θ) (1.175) Gn n h in n estimates the co-variance matrix of the asymptotic distribution of the estimator ˆ θn. Therefore IGn (θ) can be interpreted as an information matrix, called the Heyde-information. It generalizes the incremental expected information of the likelihood theory for stochastic processes; see Barndorff-Nielsen & Sørensen (1994). Since G¯ (θ) estimates the sensitivity function, and G(θ) estimates n h in the variance of the asymptotic distribution of Gn(θ), the Heyde-information has a heuristic interpretation similar to that of the Godambe-information. In fact, E G¯ (θ) = S (θ) and E ( G(θ) )= E G (θ)G (θ)T . θ n Gn θ h in θ n n We can thus think of the Heyde-information as an estimated version of the Godambe information.

Let n be a class of martingale estimating functions with finite variance. We say thatG a martingale estimating function G is Heyde-optimal in if n∗ Gn I (θ) I (θ) (1.176) Gn∗ ≥ Gn P -almost surely for all θ Θ and for all G . θ ∈ n ∈ Gn The following useful result from Heyde (1988) is similar to Theorem 1.11.1. In 94 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES order to formulate it, we need the concept of the quadratic co-characteristic of two martingales, G and G˜, both of which are assumed to have finite variance: n T G, G˜ n = E (Gi Gi 1)(G˜i G˜i 1) i 1 . (1.177) h i − − − − |F − i=1 X   Theorem 1.11.3 If G satisfies that n∗ ∈ Gn 1 1 G¯ (θ)− G(θ), G∗(θ) = G¯∗ (θ)− G∗(θ) (1.178) n h in n h in for all θ Θ and all G , then it is is Heyde-optimal in . When ∈ n ∈ Gn Gn n is closed under addition, any Heyde-optimal estimating function Gn∗ sat- G ¯ 1 isfies (1.178). Moreover, if Gn∗ (θ)− G∗(θ) n is non-random, then Gn∗ is also Godambe-optimal in . h i Gn Since in many situations condition (1.178) can be verified by showing that G(θ), G∗(θ) n = G¯n(θ) for all Gn n, it is in practice often the case that Heyde-optimalityh i implies− Godambe-optimality.∈ G

Example 1.11.4 Let us discuss a often occurringtype of martingaleestimating functions. To simplify the exposition we assume that the observed process is Markovian. For Markov processes it is natural to base martingale estimating functions on functions hij (y, x; θ), j =1,...,N, i =1,...,n satisfying that

Eθ(hij (Xi,Xi 1; θ) i 1)=0. (1.179) − |F − As in Example 1.11.2, such functions define relationships (dependent on θ) between consecutive observation Xi and Xi 1 that are, on average, equal to zero and can be used to estimate θ. We consider− the class of p-dimensional estimating functions of the form n Gn(θ)= ai(Xi 1; θ)hi(Xi,Xi 1; θ), (1.180) − − i=1 X T where hi denotes the N-dimensional vector (hi1,...,hiN ) , and ai(x; θ) is a function from IR Θ into the set of p N-matrices that are differentiable with × × respect to θ. It follows from (1.179) that Gn(θ) is a p-dimensional unbiased martingale estimating function.

We will now find the matrices ai that combine the N functions hij in an op- timal way. Let n be the class of martingale estimating functions of the form (1.180) that haveG finite variance. Then n ¯ Gn(θ)= ai(Xi 1; θ)Eθ(∂θT hi(Xi,Xi 1; θ) i 1) − − |F − i=1 X OPTIMALESTIMATINGFUNCTIONS:GENERALTHEORY 95 and n T G(θ), G∗(θ) n = ai(Xi 1; θ)Vhi (Xi 1; θ)ai∗(Xi 1; θ) , h i − − − i=1 X where n

Gn∗ (θ)= ai∗(Xi 1; θ)hi(Xi,Xi 1; θ), (1.181) − − i=1 X and T Vhi (Xi 1; θ)= Eθ hi(Xi,Xi 1; θ)hi(Xi,Xi 1; θ) i 1 − − − |F − is the conditional covariancematrix of the random vector hi(Xi,Xi 1; θ) given − i 1. If we assume that Vhi (Xi 1; θ) is invertible and define F − − T 1 T ai∗(Xi 1; θ)= Eθ(∂θ hi(Xi,Xi 1; θ) i 1) Vhi (Xi 1; θ)− , (1.182) − − − |F − − then the condition (1.178) is satisfied. Hence by Theorem 1.11.3 the estimat- ing function Gn∗ (θ) with ai∗ given by (1.182) is Heyde-optimal - provided, of course, that it has finite variance. Since G¯ (θ) 1 G (θ) = I is non- n∗ − h ∗ in − p random, the estimating function Gn∗ (θ) is also Godambe-optimal. If ai∗ were defined without the minus, Gn∗ (θ) would obviously also be optimal. The rea- son for the minus will be clear in the following. We shall now see, in exactly what sense the optimal estimating function (1.181) approximatesthe score function.The following result was first given by Kessler (1996). Let pi(y; θ x) denote the conditional density of Xi given that Xi 1 = | − x. Then the likelihood function for θ based on the data (X1,...,Xn) is n Ln(θ)= pi(Xi; θ Xi 1) | − i=1 Y (with p1 denoting the unconditional density of X1). If we assume that all pis are differentiable with respect to θ, the score function is n Un(θ)= ∂θ log pi(Xi; θ Xi 1). (1.183) | − i=1 X Let us fix i, xi 1 and θ and consider the L2-space i(xi 1,θ) of functions − 2 K − f : IR IR for which f(y) pi(y; θ xi 1)dy < . We equip i(xi 1,θ) with the7→ usual inner product | − ∞ K − R

f,g = f(y)g(y)pi(y; θ xi 1)dy, h i | − Z and let i(xi 1,θ) denote the N-dimensional subspace of i(xi 1,θ) spanned H − K − by the functions y hij (y, xi 1; θ), j =1,...,N. That the functions are lin- 7→ − early independent in i(xi 1,θ) follows from the earlier assumption that the K − covariance matrix Vhi (xi 1; θ) is regular. − 96 ESTIMATING FUNCTIONS FOR DIFFUSION-TYPE PROCESSES

Now, assume that ∂θj log pi(y xi 1; θ) i(xi 1,θ) for j = 1,...,p, de- | − ∈ K − note by gij∗ the orthogonal projection with respect to , of ∂θj log pi onto h· ·i T i(xi 1,θ), and define a p-dimensional function by gi∗ = (gi∗1,...,gip∗ ) . ThenH (under− weak regularity conditions)

gi∗(xi 1, x; θ)= ai∗(xi 1; θ)hi(xi 1, x; θ), (1.184) − − − where ai∗ is the matrix defined by (1.182). To see this, note that g∗ must have the form (1.184) with ai∗ satisfying the normal equations

∂ log p g∗, h =0, h θj i − j iki j = 1,...,p and k = 1,...,N. These equations can be expressed in the form B = a V , where B is the p p-matrix whose (j, k)th element is i i∗ hi i × ∂θj log pi, hik . The main regularity condition needed to prove (1.184) is that weh can interchangei differentiation and integration so that

∂θj [hik(y, xi 1; θ)p(y, xi 1; θ)] dy = − − Z

∂θj hik(y, xi 1; θ)p(xi 1,y; θ)dy =0, − − Z from which it follows that

Bi = ∂θT hi(y, xi 1; θ)p(xi 1,y; θ)dy. − − − Z Thus ai∗ is given by (1.182). 2

Acknowledgements

The research was supported by the Danish Center for Accounting and Finance funded by the Danish Research Council and by the Center for Research in Econometric Analysis of Time Series funded by the Danish Na- tional Research Foundation. Bibliography

A¨ıt-Sahalia, Y. (2002). “Maximum likelihood estimation of discretely sam- pled diffusions: a closed-form approximation approach”. Econometrica, 70:223–262. A¨ıt-Sahalia, Y. (2008). “Closed-form likelihood expansions for multivariate diffusions”. Ann. Statist., 36:906–937. A¨ıt-Sahalia, Y. & Mykland, P. (2003). “The effects of random and discrete sampling when estimating continuous-time diffusions”. Econometrica, 71:483–549. A¨ıt-Sahalia, Y. & Mykland, P. A. (2004). “Estimators of diffusions with ran- domly spaced discrete observations: a general theory”. Ann. Statist., 32:2186–2222. A¨ıt-Sahalia, Y. & Mykland, P. A. (2008). “An analysis of Hansen-Scheinkman moment estimators for discretely and randomly sampled diffusions”. Journal of Econometrics, 144:1–26. Baltazar-Larios, F. & Sørensen, M. (2009). “Maximum likelihood estimation for integrated diffusion processes”. Preprint, Department of Mathematical Sciences, University of Copenhagen. Barndorff-Nielsen, O. E.; Jensen, J. L. & Sørensen, M. (1990). “Parametric Modelling of Turbulence”. Phil. Trans. R. Soc. Lond. A, 332:439–455. Barndorff-Nielsen, O. E.; Jensen, J. L. & Sørensen, M. (1998). “Some Station- ary Processes in Discrete and Continuous Time”. Advances in Applied Probability, 30:989–1007. Barndorff-Nielsen, O. E.; Kent, J. & Sørensen, M. (1982). “Normal variance- mean mixtures and z-distributions”. International Statistical Review, 50:145–159. Barndorff-Nielsen, O. E. & Shephard, N. (2001). “Non-Gaussian Ornstein- Uhlenbeck-Based Models and some of their Uses in Financial Econo- metrics (with discussion)”. Journal of the Royal Statistical Society B, 63:167–241. Barndorff-Nielsen, O. E. & Sørensen, M. (1991). “Information quantities in non-classical settings”. Computational Statistics and Data Analysis,

97 98 BIBLIOGRAPHY 12:143–158. Barndorff-Nielsen, O. E. & Sørensen, M. (1994). “A review of some aspects of asymptotic likelihood theory for stochastic processes”. International Statistical Review, 62:133–165. Beskos, A.; Papaspiliopoulos, O. & Roberts, G. O. (2006). “Retrospective exact simulation of diffusion sample paths with applications”. Bernoulli, 12:1077–1098. Beskos, A.; Papaspiliopoulos, O.; Roberts, G. O. & Fearnhead, P. (2006). “Exact and computationally efficient likelihood-based estimation for dis- cretely observed diffusion processes”. J. Roy. Statist. Soc. B, 68:333–382. Beskos, A. & Roberts, G. O. (2005). “Exact simulation of diffusions”. Ann. Appl. Prob., 15:2422–2444. Bibby, B. M. (1995). Inference for diffusion processes with particular em- phasis on compartmental diffusion processes. PhD thesis, University of Aarhus. Bibby, B. M.; Skovgaard, I. M. & Sørensen, M. (2005). “Diffusion-type models with given marginals and autocorrelation function”. Bernoulli, 11:191–220. Bibby, B. M. & Sørensen, M. (1995). “Martingale estimation functions for discretely observed diffusion processes”. Bernoulli, 1:17–39. Bibby, B. M. & Sørensen, M. (1996). “On estimation for discretely observed diffusions: a review”. Theory of Stochastic Processes, 2:49–56. Bibby, B. M. & Sørensen, M. (2003). “Hyperbolic processes in finance”. In Rachev, S., editor, Handbook of Heavy Tailed Distributions in Finance, pages 211–248. Elsevier Science. Billingsley, P. (1961). “The Lindeberg-L´evy theorem for martingales”. Proc. Amer. Math. Soc., 12:788–792. Bollerslev, T. & Wooldridge, J. (1992). “Quasi-maximum likelihood estima- tors and inference in dynamic models with time-varying covariances”. Econometric Review, 11:143–172. Bollerslev, T. & Zhou, H. (2002). “Estimating stochastic volatility diffusion using conditional moments of integrated volatility”. Journal of Econo- metrics, 109:33–65. Brockwell, P. J. & Davis, R. A. (1991). Time Series: Theory and Methods. Springer-Verlag, New York. Campbell, J. Y.; Lo, A. W. & MacKinlay, A. C. (1997). The Econometrics of Financial Markets. Princeton University Press, Princeton. Chamberlain, G. (1987). “Asymptotic efficiency in estimation with conditional moment restrictions”. Journal of Econometrics, 34:305–34. BIBLIOGRAPHY 99 Chan, K. C.; Karolyi, G. A.; Longstaff,F. A. & Sanders, A. B. (1992). “An em- pirical comparison of alternative models of the short-term interest rate”. Journal of Finance, 47:1209–1227. Christensen, B. J. & Sørensen, M. (2008). “Optimal inference in dynamic moels with conditional moment restrictions”. Working paper, CREATES, Center for Research in Econometric Analysis of Time Series. Clement, E. (1997). “Estimation of diffusion processes by simulated moment methods”. Scand. J. Statist., 24:353–369. Comte, F.; Genon-Catalot, V. & Rozenholc, Y. (2009). “Nonparametric adap- tive estimation for integrated diffusions”. Stochastic processes and their applications, 119:811–834. Conley, T. G.; Hansen, L. P.; Luttmer, E. G. J. & Scheinkman, J. A. (1997). “Short-term interest rates as subordinated diffusions”. Review of Finan- cial Studies, 10:525–577. Dacunha-Castelle, D. & Florens-Zmirou, D. (1986). “Estimation of the coef- ficients of a diffusion from discrete observations”. , 19:263– 284. De Jong, F.; Drost, F. C. & Werker, B. J. M. (2001). “A jump-diffusion model for exchange rates in a target zone”. Statistica Neerlandica, 55:270–300. Ditlevsen, P. D.; Ditlevsen, S. & Andersen, K. K. (2002). “The fast climate fluctuations during the stadial and interstadial climate states”. Annals of Glaciology, 35:457–462. Ditlevsen, S. & Sørensen, M. (2004). “Inference for observations of integrated diffusion processes”. Scand. J. Statist., 31:417–429. Dorogovcev, A. J. (1976). “The consistency of an estimate of a parameter of a stochastic differential equation”. Theor. Probability and Math. Statist., 10:73–82. Doukhan, P. (1994). Mixing, Properties and Examples. Springer, New York. Lecture Notes in Statistics 85. Down, D.; Meyn, S. & Tweedie, R. (1995). “Exponential and uniform ergod- icity of Markov processes”. Annals of Probability, 23:1671–1691. Duffie, D. & Singleton, K. (1993). “Simulated moments estimation of Markov models of asset prices”. Econometrica, 61:929–952. Durbin, J. (1960). “Estimation of parameters in time-series regression models”. J. Roy. Statist. Soc. B, 22:139–153. Durham, G. B. & Gallant, A. R. (2002). “Numerical techniques for maximum likelihood estimation of continuous-time diffusion processes”. J. Business & Econom. Statist., 20:297–338. D¨uring, M. (2002). “Den prediktions-baserede estimationsfunktion for dif- 100 BIBLIOGRAPHY fusions puljemodeller”. Master’s thesis, University of Copenhagen. In Danish. Elerian, O.; Chib, S. & Shephard, N. (2001). “Likelihood inference for dis- cretely observed non-linear diffusions”. Econometrica, 69:959–993. Eraker, B. (2001). “MCMC Analysis of Diffusion Models with Application to Finance”. J. Business & Econom. Statist., 19:177–191. Fisher, R. A. (1935). “The logic of inductive inference”. J. Roy. Statist. Soc., 98:39–54. Florens-Zmirou, D. (1989). “Approximate discrete-time schemes for statistics of diffusion processes”. Statistics, 20:547–557. Forman, J. L. & Sørensen, M. (2008). “The Pearson diffusions: A class of statistically tractable diffusion processes”. Scand. J. Statist., 35:438–465. Freidlin, M. I. & Wentzell, A. D. (1998). Random pertubations of dynamical systems.2nd Edition. Springer, New York. Friedman, A. (1975). Stochastic Differential Equations and Applications, Vol- ume 1. Academic Press, New York. Genon-Catalot, V. (1990). “Maximum contrast estimation for diffusion pro- cesses from discrete observations”. Statistics, 21:99–116. Genon-Catalot, V. & Jacod, J. (1993). “On the estimation of the diffusion coefficient for multi-dimensional diffusion processes”. Ann. Inst. Henri Poincare,´ Probabilites´ et Statistiques, 29:119–151. Genon-Catalot, V.; Jeantheau, T. & Lar´edo, C. (2000). “Stochastic volatility models as hidden Markov models and statistical applications”. Bernoulli, 6:1051–1079. Gloter, A. (2000). “Parameter estimation for a discrete sampling of an inte- grated Ornstein-Uhlenbeck process”. Statistics, 35:225–243. Gloter, A. (2006). “Parameter estimation for a discretely observed integrated diffusion process”. Scand. J. Statist., 33:83–104. Gloter, A. & Sørensen, M. (2009). “Estimation for stochastic differential equa- tions with a small diffusion coefficient”. Stoch. Proc. Appl., 119:679–699. Gobet, E. (2001). “Local asymptotic mixed normality property for elliptic diffusion: a Malliavin calculus approach”. Bernoulli, 7:899–912. Gobet, E. (2002). “LAN property for ergodic diffusions with discrete obser- vations”. Ann. Inst. Henri Poincare,´ Probabilites´ et Statistiques, 38:711– 737. Godambe, V. P. (1960). “An optimum property of regular maximum likelihood estimation”. Ann. Math. Stat., 31:1208–1212. Godambe, V. P. (1985). “The foundations of finite sample estimation in stochastic processes”. Biometrika, 72:419–428. BIBLIOGRAPHY 101 Godambe, V. P. & Heyde, C. C. (1987). “Quasi likelihood and optimal estima- tion”. International Statistical Review, 55:231–244. Gourieroux, C. & Jasiak, J. (2006). “Multivariate Jacobi process and with application to smooth transitions”. Journal of Econometrics, 131:475– 505. Gradshteyn, I. S. & Ryzhik, I. M. (1965). Table of Integrals, Series, and Prod- ucts, 4th Edition. Academic Press, New-York. Hall, A. R. (2005). Generalized Method of Moments. Oxford University Press, New York. Hall, P. & Heyde, C. C. (1980). Martingale Limit Theory and Its Applications. Academic Press, New York. Hansen, L. P. (1982). “Large sample properties of generalized method of mo- ments estimators”. Econometrica, 50:1029–1054. Hansen, L. P. (1985). “A method for calculating bounds on the asymptotic co- variance matrices of generalized method of moments estimators”. Journal of Econometrics, 30:203–238. Hansen, L. P. (1993). “Semiparametric efficiency bounds for linear time-series models”. In Phillips, P. C. B., editor, Models, Methods and Applications of Econometrics: Esseys in Honor of A. R. Bergstrom, pages 253–271. Cambridge, MA: Blackwell. Hansen, L. P. (2001). “Method of moments”. In International Encyclopedia of the Social and Behavior Sciences. New York: Elsevier. Hansen, L. P. & Scheinkman, J. A. (1995). “Back to the future: generating moment implications for continuous-time Markov processes”. Econo- metrica, 63:767–804. Hansen, L. P.; Scheinkman, J. A. & Touzi, N. (1998). “Spectral methods for identifying scalar diffusions”. Journal of Econometrics, 86:1–32. Heyde, C. C. (1988). “Fixed sample and asymptotic optimality for classes of estimating functions”. Contemporary Mathematics, 80:241–247. Heyde, C. C. (1997). Quasi-Likelihood and Its Application. Springer-Verlag, New York. Hildebrandt, E. H. (1931). “Systems of polynomials connected with the Char- lier expansions and the Pearson differential and difference equations”. Ann. Math. Statist., 2:379–439. Iacus, S. M. (2008). Simulation and Inference for Stochastic Differential Equa- tions. Springer, New York. Iacus, S. M. & Kutoyants, Y. (2001). “Semiparametric hypotheses testing for dynamical systems with small noise”. Math. Methods Statist., 10:105– 120. 102 BIBLIOGRAPHY Jacobsen, M. (2001). “Discretely observed diffusions; classes of estimating functions and small ∆-optimality”. Scand. J. Statist., 28:123–150. Jacobsen, M. (2002). “Optimality and small ∆-optimality of martingale esti- mating functions”. Bernoulli, 8:643–668. Jacod, J. & Sørensen, M. (2009). “Aspects of asymptotic statistical theory for stochastic processes.”. Preprint, Department of Mathematical Sciences, University of Copenhagen. In preparation. Jeganathan, P. (1982). “On the asymptotic theory of estimation when the limit of the log-likelihood ratios is mixed normal”. Sankhya¯ A, 44:173–212. Jeganathan, P. (1983). “Some asymptotic properties of risk functions when the limit of the experiment is mixed normal”. Sankhya¯ A, 45:66–87. Kallenberg, O. (1997). Foundations of Modern Probability. Springer-Verlag, New York. Kelly, L.; Platen, E. & Sørensen, M. (2004). “Estimation for discretely ob- served diffusions using transform functions”. J. Appl. Prob., 41:99–118. Kessler, M. (1996). Estimation parametrique´ des coefficients d’une diffusion ergodique a` partir d’observations discretes` . PhD thesis, Laboratoire de Probabilit´es, Universit´eParis VI. Kessler, M. (1997). “Estimation of an ergodic diffusion from discrete observa- tions”. Scand. J. Statist., 24:211–229. Kessler, M. (2000). “Simple and explicit estimating functions for a discretely observed diffusion process”. Scand. J. Statist., 27:65–82. Kessler, M. & Paredes, S. (2002). “Computational aspects related to martin- gale estimating functions for a discretely observed diffusion”. Scand. J. Statist., 29:425–440. Kessler, M.; Schick, A. & Wefelmeyer, W. (2001). “The information in the marginal law of a Markov chain”. Bernoulli, 7:243–266. Kessler, M. & Sørensen, M. (1999). “Estimating equations based on eigenfunc- tions for a discretely observed diffusion process”. Bernoulli, 5:299–314. Kessler, M. & Sørensen, M. (2005). “On time-reversibility and estimating functions for Markov processes”. Statistical Inference for Stochastic Pro- cesses, 8:95–107. Kimball, B. F. (1946). “Sufficient statistical estimation functions for the pa- rameters of the distribution of maximum values”. Ann. Math. Statist., 17:299–309. Kloeden, P. E. & Platen, E. (1999). Numerical Solution of Stochastic Differen- tial Equations. 3rd revised printing. Springer-Verlag, New York. Kloeden, P. E.; Platen, E.; Schurz, H. & Sørensen, M. (1996). “On effects of discretization on estimates of drift parameters for diffusions processes”. BIBLIOGRAPHY 103 J. Appl. Prob., 33:1061–1076. Kuersteiner, G. (2002). “Efficient instrumental variables estimation for au- toregressive models with conditional heteroskedasticity”. Econometric Theory, 18:547–583. Kusuoka, S. & Yoshida, N. (2000). “Malliavin calculus, geometric mixing, and expansion of diffusion functionals”. and Related Fields, 116:457–484. Kutoyants, Y. (1994). Identification of dynamical systems with small noise. Kluwer. Kutoyants, Y. (1998). “Semiparametric estimation for dynamical systems with small noise”. Math. Methods Statist., 7:457–465. Larsen, K. S. & Sørensen, M. (2007). “A diffusion model for exchange rates in a target zone”. , 17:285–306. Le Cam, L. & Yang, G. L. (2000). Asymptotics in Statistical. Some Basic Concepts. 2nd Edition. Springer, New York. Li, B. (1997). “On the consistency of generalized estimating equations”. In Ba- sawa, I. V.; Godambe, V. P. & Taylor, R. L., editors, Selected Proceedings of the Symposium on Estimating Functions, pages 115–136. Hayward: Institute of . IMS Lecture Notes – Monograph Series, Vol. 32. Liang, K.-Y. & Zeger, S. L. (1986). “Longitudinal data analysis using general- ized linear model”. Biometrika, 73:13–22. Longstaff, F. & Schwartz, E. (1995). “A simple approach to valuing risky fixed and floating rate debt”. J. Finance, 1:789–819. McLeish, D. L. & Small, C. G. (1988). The Theory and Applications of Statis- tical Inference Functions. Springer-Verlag, New York. Lecture Notes in Statistics 44. Nagahara, Y. (1996). “Non-Gaussian distribution for stock returns and related stochastic differential equation”. Financial and the Japanese Markets, 3:121–149. Newey, W. K. (1990). “Efficient instrumental variables estimation of nonlinear models”. Econometrica, 58:809–837. Newey, W. K. & West, K. D. (1987). “Hypothesis testing with efficient method of moments estimation”. International Economic Review, 28:777–787. Overbeck, L. & Ryd´en, T. (1997). “Estimation in the Cox-Ingersoll-Ross model”. Econometric Theory, 13:430–461. Ozaki, T. (1985). “Non-linear time series models and dynamical systems”. In Hannan, E. J.; Krishnaiah, P. R. & Rao, M. M., editors, Handbook of Statistics, Vol. 5, pages 25–83. Elsevier Science Publishers. 104 BIBLIOGRAPHY Pearson, K. (1895). “Contributionsto the Mathematical Theoryof Evolution II. Skew Variation in Homogeneous Material”. Philosophical Transactions of the Royal Society of London. A, 186:343–414. Pedersen, A. R. (1994). “Quasi-likelihood inference for discretely observed diffusion processes”. Research Report No. 295, Department of Theoreti- cal Statistics, Institute of Mathematics, University of Aarhus. Pedersen, A. R. (1995). “A new approach to maximum likelihood estima- tion for stochastic differential equations based on discrete observations”. Scand. J. Statist., 22:55–71. Picard, J. (1986). “Nonlinear filtering of one-dimensional diffusions in the case of a high -to-noise ratio”. SIAM J. Appl. Math., 46:1098–1125. Picard, J. (1991). “Efficiency of the extended Kalman filter for nonlinear sys- tem with small noise”. SIAM J. Appl. Math., 51:843–885. Pokern, Y.; Stuart, A. M. & Wiberg, P. (2009). “Parameter estimation for par- tially observed hypoelliptic diffusions”. J. Roy. Statist. Soc. B, 71:49–73. Poulsen, R. (1999). “Approximate maximum likelihood estimation of dis- cretely observed diffusion processes”. Working Paper 29, Centre for An- alytical Finance, Aarhus. Prakasa Rao, B. L. S. (1988). “Statistical inference from sampled data for stochastic processes”. Contemporary Mathematics, 80:249–284. Prentice, R. L. (1988). “Correlated binary regression with covariates specific to each binary observation”. Biometrics, 44:1033–1048. Roberts, G. O. & Stramer, O. (2001). “On inference for partially ob- served nonlinear diffusion models using Metropolis-Hastings algo- rithms”. Biometrika, 88:603–621. Romanovsky,V.(1924). “Generalization of some types of the frequency curves of Professor Pearson”. Biometrika, 16:106–117. Sagan, J. D. (1958). “The estimation of economic relationsships using instru- mental variables”. Econometrica, 26:393–415. Skorokhod, A. V. (1989). Asymptotic Methods in the Theory of Stochastic Dif- ferential Equations. American Mathematical Society, Providence, Rhode Island. Sørensen, H. (2001). “Discretely observed diffusions: Approximation of the continuous-time score function”. Scand. J. Statist., 28:113–121. Sørensen, M. (2000a). “Prediction-Based Estimating Functions”. Economet- rics Journal, 3:123–147. Sørensen, M. (2000b). “Small dispersion asymptotics for diffusion martingale estimating functions”. Preprint 2, Department of Statistics and Operation Research, University of Copenhagen. BIBLIOGRAPHY 105 Sørensen, M. (2007). “Efficient estimation for ergodic diffusions sampled at high frequency”. Preprint, Department of Mathematical Sciences, Uni- versity of Copenhagen. Sørensen, M. & Uchida, M. (2003). “Small-diffusion asymptotics for dis- cretely sampled stochastic differential equations”. Bernoulli, 9:1051– 1069. Takahashi, A. & Yoshida, N. (2004). “An asymptotic expansion scheme for optimal investment problems”. Statistical Inference for Stochastic Pro- cesses, 7:153–188. Uchida, M. (2004). “Estimation for discretely observed small diffusions based on approximate martingale estimating functions”. Scand. J. Statist., 31:553–566. Uchida, M. (2008). “Approximate martingale estimating functions for stochastic differential equations with small noises”. Stoch. Proc. Appl., 118:1706–1721. Uchida, M. & Yoshida, N. (2004a). “Asymptotic expansion for small diffusions applied to option pricing”. Stat. Inference Stoch. Process., 7:189–223. Uchida, M. & Yoshida, N. (2004b). “Information criteria for small diffusions via the theory of Malliavin-Watanabe”. Stat. Inference Stoch. Process., 7:35–67. Veretennikov, A. Y. (1987). “Bounds for the mixing rate in the theory of stochastic equations”. Theory of Probability and its Applications, 32:273– 281. Wedderburn, R. W. M. (1974). “Quasi-likelihood functions, generalized linear models, and the Gauss-Newton method”. Biometrika, 61:439–447. West, K. D. (2001). “On optimal instrumental variables estimation of station- ary time series models”. International Economic Review, 42:1043–1050. Wong, E. (1964). “The construction of a class of stationary Markoff pro- cesses”. In Bellman, R., editor, Stochastic Processes in Mathematical and Engineering, pages 264–276. American Mathematical Soci- ety, Rhode Island. Yoshida, N. (1992). “Estimation for diffusion processes from discrete obser- vations”. Journal of Multivariate Analysis, 41:220–242.