nonparametric regression approach to bayesian estimation · 2017-06-12 · nonparametric regression...

41
ISSN 1440-771X Australia Department of Econometrics and Business Statistics http://www.buseco.monash.edu.au/depts/ebs/pubs/wpapers/ November 2014 Working Paper 25/14 Nonparametric Regression Approach to Bayesian Estimation Jiti Gao and Han Hong

Upload: others

Post on 12-Jul-2020

17 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

ISSN 1440-771X

Australia

Department of Econometrics and Business Statistics

http://www.buseco.monash.edu.au/depts/ebs/pubs/wpapers/

November 2014

Working Paper 25/14

Nonparametric Regression Approach to Bayesian Estimation

Jiti Gao and Han Hong

Page 2: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Nonparametric Regression Approach to Bayesian Estimation 1

Jiti Gao and Han Hong

Monash University and Stanford University

Abstract

Estimation of unknown parameters and functions involved in complex nonlinear econometric

models is a very important issue. Existing estimation methods include generalised method of mo-

ments (GMM) by Hansen (1982) and others, efficient method of moments (EMM) by Gallant and

Tauchen (1997), Markov chain Monte Carlo (MCMC) method by Chernozhukov and Hong (2003),

and nonparametric simulated maximum likelihood estimation (NSMLE) method by Creel and Kris-

tensen (2011), and Kristensen and Shin (2012). Except the NSMLE method, other existing methods

do not provide closed–form solutions. This paper proposes non– and semi–parametric based closed–

form approximations to the estimation and computation of posterior means involved in complex

nonlinear econometric models. We first consider the case where the samples can be independently

drawn from both the likelihood function and the prior density. The samples and observations are

then used to nonparametrically estimate posterior mean functions. The estimation method is also

applied to estimate the posterior mean of the parameter–of–interest on a summary statistic. Both

the asymptotic theory and the finite sample study show that the nonparametric estimate of this

posterior mean is superior to existing estimates, including the conventional sample mean.

This paper then proposes some non– and semi–parametric dimension reductions methods to deal

with the case where the dimensionality of either the regressors or the summary statistics is large.

Meanwhile, the paper develops a nonparametric estimation method for the case where the samples

are obtained from using a resampling algorithm. The asymptotic theory shows that in each case

the rate of convergence of the nonparametric estimate based on the resamples is faster than that

of the conventional nonparametric estimation method by an order of the number of the resamples.

The proposed models and estimation methods are evaluated through using simulated and empirical

examples. Both the simulated and empirical examples show that the proposed nonparametric

estimation based on resamples outperforms existing estimation methods.

Key words: Bayesian method; double asymptotics; Markov chain and Monte Carlo; parametric

regression; nonparametric regression; stationary time series data.

JEL Classification: C12, C14, C22.

Abbreviated Title: Nonparametric Estimation of Bayesian Means.

1The authors acknowledge constructive comments and suggestions from several seminar participants.

Thanks also go to Tingting Cheng and Jiying Yin for their excellent computing assistance. The first author

was supported by an Australian Research Council Professorial Fellowship Award: DP1096374 and an Aus-

tralian Research Council Discovery Projects Scheme under Grant number: DP130104229. The second author

acknowledges financial support by the National Science Foundation (SES 1164589) and SIEPR.

1

Page 3: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

1 Introduction

Beyesian estimation and computation is a complicated issue, particularly when estimation

issues involve computational complexity. The literature basically shows that there are three

stages of the developments. The first stage of the developments is due to the fact that

empirical Bayesian approach has been used to provide some closed–form solutions to various

Bayesian estimation problems. One useful class of models covers a class of exponential

families, in which the Bayes estimate is a ratio of the first–order derivative of the marginal

density and the marginal density itself. Then, a nonparametric kernel density estimation

method may be employed to consistently estimate the posterior mean. Similar approaches

have also been done for a class of uniform families. The literature is summarised and discussed

in Carlin and Louis (1996), Efron (1996), and some other studies.

Mainly due to the fact that most posterior means do not have closed–form relationships

with the marginal density and its functionals, computation of posterior means involves deal-

ing with some non–tractable integrals and possible high dimensionality and therefore becomes

a very difficult issue. This comes to the second stage of the developments that importance

sampling, the Gibbs sampler and other MCMC tools become available and effective for im-

plementing Bayesian estimation and computation. There is a huge literature about such

developments. We refer the reader to Liu (2001), Geweke (2005), and Brooks et al (2011).

Since Bayesian inference basically relies on the full posterior density function and the dimen-

sionality of such posterior density is usually large, both computation and simulation involve

all sorts of difficulties. To partially address such computational issues, the third stage of

the developments is based on the proposal of the so–called “Approximate Bayesian Compu-

tation” (ABC). Recent studies include Beaumont, Zhang and Balding (2002), Blum (2010),

Fearnhead and Prangle (2012), and Blum et al (2013).

This paper proposes some general non– and semi–parametric regression approaches to the

estimation and computation of posterior means involved in complex nonlinear econometric

models. The proposed estimation method provides a simple and useful alternative to existing

estimation methods, such as MCMC (Chernozhukov and Hong 2003), GMM (Hansen 1982),

EMM (Gallant and Tauchen 1997), and NSMLE method proposed recently by Creel and

Kristensen (2011), and Kristensen and Shin (2012). More recently, Gao and Hong (2014)

considerably explore the ABC idea and the NSMLE method for nonparametric implementa-

tion of GMM in practice. As we discuss in Section 4 of this paper, based on direct sampling,

the proposed nonparametric approach makes it possible to provide a closed–form estimate

for a general conditional moment of the form E[ψ(θ)|Tn], where θ is the parameter of inter-

est, ψ(·) is of a known functional form and Tn is a summary statistic, such as the sample

mean of X1, X2, · · · , Xn. As proposed in Section 4 below, moreover, a nonparametric esti-

mation method based on resamples results in asymptotically normal estimates for unknown

2

Page 4: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

conditional moments with rates of convergence faster than those for existing estimates. Such

theoretical findings are evaluated in Sections 6 and 7 through both simulated and real data

examples.

In summary, this paper proposes non– and semi–parametric methods for the establishment

of closed–form estimates for conditional moments. We believe that the newly proposed

estimation method reveals some important findings and has the following theoretical and

computational advantages:

a) it results in closed–form expressions for estimates of unknown parameters and functions

involved in non– and semi–parametric models;

b) it avoids involving numerical approximations to intractable integrals involved in the

computation of Bayesian estimates;

c) it directly and naturally addresses various high–dimensional issues involved in non–

and semi–parametric approximation and estimation;

d) it facilitates both the implementation and the application of Bayesian estimation and

computation for economic and financial models; and

e) it provides a simple and useful alternative to estimating unknown parameters and

functions involved in classes of complex nonlinear econometric models.

The organisation of this paper is given as follows. Section 2 gives some examples and

models to link and motivate the discussion of this paper with the relevant literature before a

nonparametric kernel estimation method is proposed to estimate the posterior mean function.

Section 2 then establishes an asymptotic theory for the estimation method proposed in this

section. Using a resampling algorithm, Section 3 significantly improves the rate of conver-

gence of a nonparametric kernel estimator based on the resamples and its resulting theory is

then established in the end of Section 3. Section 4 proposes to estimate a general posterior

mean of the form E[θ|Tn] before giving a comparison with an existing estimation method.

Estimation problems involving dependent data are discussed respectively in Sections 3 and

4. Section 5 extends the discussion in Sections 2–4 to the case where there are nuisance

parameters involved and then considers a nonparametric estimation issue where the nuisance

parameters involved are consistently estimated. This large sample theory is supported by the

small and finite sample evaluation given in Sections 6 and 7. Section 6 gives some numerical

evidence to support the proposed models and estimation methods. An empirical example dis-

cussing parameter estimation of unknown parameters involved in a GARCH model is given

in Section 7. Some concluding comments are given in Section 8 before the mathematical

technicalities are given in Section 9.

3

Page 5: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

2 Models and Estimation Methods

2.1 Examples and motivation

Before we propose our models and estimation methods, we use some examples to motivate

our discussion.

Example 2.1: Consider a general distributional model of the form

Xt ∼ Ft(x; θ), t = 1, 2, · · · , n, (2.1)

where each Ft(·; θ) is a parametric distributional function indexed by θ, a vector of unknown

parameters. Note that {Xt} can be either independent, stationary or nonstationary time

series.

For model (2.1), the vector of unknown parameters, θ, can be consistently estimated

by classical estimates, such as, the conventional sample moment and MLE. Section 6 below

shows that if we move one–step further by combining simulated samples with a nonparametric

estimation method based on the simulated samples, a nonparametric kernel estimator for a

conditional moment of the form E[ψ(θ)|Tn] is more efficient than such classical estimates,

where Tn is a summary statistic, and ψ(·) is a known function available for computation.

Extensions of model (2.1) are needed to deal with the general conditional mean case

discussed in Chen (2007), in which Ft(·; ·) is allowed to be semiparametric.

Example 2.2 (GARCH model): Consider a GARCH (1,1) model of the form:

yt = σtεt, εt, t = 1, 2, · · · , n,

σ2t = b0 + b1y

2t−1 + b2σ

2t−1, (2.2)

where {εt} is a sequence of errors and θ = (b0, b1, b2)′ denotes a vector of unknown parameters.

Our study in Section 7 below discusses model (2.2) and evaluates the applicability and

practical relevance of the proposed estimation method to be discussed in Sections 3 and 4

below to show that a nonparametric estimator for g(Tn) = E[θ|Tn] is more efficient than Tn

itself when Tn is the MLE of θ.

2.2 Estimation based on simulation

Let f(x|θ) be the conditional density of x given θ and π(·) be the prior density. The Bayesian

estimate of θ given x is defined by

g(x) = E[θ|x] =

∫θf(θ|x)dθ =

∫θf(x|θ)π(θ)dθ∫f(x|θ)π(θ)dθ

≡ q(x)

p(x), (2.3)

where θ = (θ1, · · · , θd)τ is a vector of unknown parameters, p(x) =∫f(x|θ)π(θ)dθ and

q(x) =∫θf(x|θ)π(θ)dθ.

4

Page 6: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Throughout the rest of this paper, we discuss the case where the model is exactly identi-

fied. To present the main idea in this section, we focus on the case of d = 1.

Assume that the functional form of f(x|θ) is available for computation. Suppose that

θi is drawn from a proper probability density λ(θ) and both the forms of π(θ) and λ(θ) are

available for computation. In this case, we may estimate g(x) by

gm(x) =

∑mj=1 θjf(x|θj)π(θj)

λ(θj)∑mj=1 f(x|θj)π(θj)

λ(θj)

, (2.4)

and for the case where Tn is a summary statistic, we have

gm(Tn) =

∑mj=1 θjf(Tn|θj)π(θj)

λ(θj)∑mj=1 f(Tn|θj)π(θj)

λ(θj)

. (2.5)

Such discussions may be found from Geweke (1989), Gelfand and Smith (1990), and

Geweke (2005) for examples. Note that there is no need to draw samples from π(θ) as long

as it is possible to either draw {Xi} from p(·) or to have the data {Xi} available for use.

This section proposes to directly estimate the posterior mean by the nonparametric ker-

nel method. In the rest of this section, we assume that we may draw (xj, θj) jointly from

f(x|θ)λ(θ) when λ(θ) is a proper probability density. For notational simplicity, in the dis-

cussion of the rest of this section and Section 3, we assume that π(θ) is already a proper

prior density available for sampling, and thus choose λ(·) = π(·). This is consistent with the

sampling approach adopted in the ABC literature. In Section 4 below, we consider the case

where λ(·) is the only proper prior density available for sampling and computation. Section

5 proposes a nonparametric estimation method that is based on MCMC samples.

Equation (2.3) implies that we can introduce a regression model of the form

θ = E[θ|x] + (θ − E[θ|x]) ≡ g(x) + e, (2.6)

where e = θ − E[θ|x] satisfies E[e|x] = 0.

Note that the functional of g(x) may not be feasibly available for computation even

though the functional forms of f(x|θ) and π(θ) may be assumed to be either parametrically

or semiparametrically known for sampling and computation. Thus, we propose to estimate

g(x) directly using the samples {(xj, θj)} readily drawn from f(x|θ)π(θ).

The first objective is to estimate g(x). Suppose that we may simulate (xj, θj : j =

1, 2, · · · ,m) directly from f(x|θ)π(θ) and then define

θj = g (xj) + ej, j = 1, 2, · · · ,m, (2.7)

where {ej} is a sequence of independent errors with mean zero and finite variance σ2 = E [e21].

We then estimate g(·) by

gm(x) =m∑j=1

Kmj(x)θj, (2.8)

5

Page 7: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

where Kmj(x) =K(xj−xh

)∑mk=1K(xk−xh )

, in which K(·) is a kernel function and h is a bandwidth.

In order to incorporate the data {Xi : 1 ≤ i ≤ n} into the estimation procedure, we

simulate {θmi : 1 ≤ i ≤ n < m} from a regression model of the form

θmi = gm(Xi) + εmi, (2.9)

where {εmi : 1 ≤ i ≤ n} is available for sampling as a sequence of conditionally independent

random errors given {(xj, θj) : 1 ≤ j ≤ m}, and is independent of {Xi} satisfying

E[εmi|(x1, · · · , xm; θ1, · · · , θm)] = 0 and E[ε2mi|(x1, · · · , xm; θ1, · · · , θm)] = σ2

mx <∞.(2.10)

In practice, εmi can be simulated from εmi = λmi(em1, · · · , emm) ξi, in which emj = θj −gm(xj), {ξi} is a sequence of independent and identically distributed (i.i.d.) random variables,

with E[ξ1] = 0 and E [ξ21 ] = 1, generated from a pre–specified probability distribution, such

as, either the standard normal distribution–N(0, 1) or

P

(η1 = −

√5− 1

2

)=

√5 + 1

2√

5and P

(η1 =

√5 + 1

2

)=

√5− 1

2√

5, (2.11)

and λmi(· · · ) is a sequence of measurable functions. There are many cases one may use in

practice:

• Case I: λmi(em1, · · · , emm) = σm, where σ2m = 1

m

∑mj=1 e

2mj with emj = θj − gm(xj);

• Case II: λmi(em1, · · · , emm) = emi + 1√m−n

∑m−nj=n+1 emj; and

• case III: λmi(em1, · · · , emm) =∑m

j=1 αjiemj, where {αji} is a sequence of real numbers

chosen such that αji ≥ 0 and∑m

j=1 αji = 1.

The construction of equation (2.9) involves some kind of bootstrap idea through com-

pressing the information already available from {(xj, θj) : j = 1, 2, · · · ,m} and then equation

(2.8). We finally estimate g(x) by

gmn(x) =n∑i=1

Lni(x)θmi, (2.12)

where Lni(x) =L(Xi−xb )∑nk=1 L

(Xk−xb

) , in which L(·) is a kernel function and b is a bandwidth.

Our experience shows that the choice of {εmi} does not affect asymptotic consistency of

gmn(·). Before asymptotic properties for gm(x) and gmn(x) are established in Sections 2.3

and 2.4 below, we summarise the estimation procedure as follows:

• Step 1: Simulate {(xj, θj) : j = 1, 2, · · · ,m} from f(x|θ)π(θ);

6

Page 8: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

• Step 2: Estimate g(x) by gm(x) =∑m

j=1 Kmj(x)θj;

• Step 3: Simulate {θmi : 1 ≤ i ≤ n} from θmi = gm(Xi) + εmi; and

• Step 4: Re–estimate g(x) by gmn(x) =∑n

i=1 Lni(x)θmi.

We will establish an asymptotic theory for the univariate case in Section 2.3 and then the

multivariate case in Section 2.4 below.

2.3 Univariate Case

In this section, we assume that the dimensionality of θ is d = 1. Let x = (x1, · · · , xr)τ be

the r–dimensional vector. To establish an asymptotic theory for gm(x) and gmn(x), we now

introduce the following assumptions.

Assumption 2.1: (i) Let the product of f(x|θ) and π(θ) be a proper probability density

function.

(ii) Let f(x|θ) be three times differentiable with respect to x and f(i)x (x|θ) be the i–

th partial derivative of f(x|θ) with respect to x such that∫|θ|∣∣∣∣∣∣f (i)

x (x|θ)∣∣∣∣∣∣ π(θ)dθ < ∞

and∫ ∣∣∣∣∣∣f (i)

x (x|θ)∣∣∣∣∣∣ π(θ)dθ < ∞ for any given x and i = 0, · · · , 3, where || · || denotes the

conventional Euclidean norm.

(iii) Suppose that both p2(x) =∫f

(2)x (x|θ)π(θ)dθ and q2(x) =

∫θf

(2)x (x|θ)π(θ)dθ are

continuous in x.

(iv) Suppose that {θj : j = 1, 2, · · · ,m} is a sequence of i.i.d. random variables drawn

from π(θ) and that {(xj, θj) : j = 1, 2, · · · ,m} is a vector of i.i.d. random vectors drawn

from f(x|θ)π(θ). Let f(x) be the marginal density of {xj}.

Assumption 2.2: (i) Suppose that there is a data set {Xi : i = 1, 2, · · · , n} that is available

as an i.i.d. random variables with p(x) being the density function.

(ii) Suppose that {Xi : i = 1, 2, · · · , n} is independent of {(xj, θj) : j = 1, 2, · · · ,m}. Let

{εmi} satisfy equations (2.9) and (2.10).

Assumption 2.3: (i) Let K(·) be the probability kernel function satisfying∫uK(u)du = 0,

0 <∫||u||2K(u)du < ∞ and 0 <

∫K2(u)du < ∞. Let the bandwidth h satisfy h → 0,

mhr →∞ and mhr+4 → c(r) for some 0 < c(r) <∞.

(ii) Let L(·) be a probability kernel function satisfying∫vL(v)dv = 0, 0 <

∫||v||2L(v)dv <

∞, 0 <∫L2(v)dv < ∞,

∫||v||3L(v)dv < ∞ and

∫||v||4L(v)dv < ∞. Let the bandwidth b

satisfy b→ 0 and nbr →∞.

(iii) Let hb

= o(1), nm

= o(1), nbrh4 = O(1), nbr+4 = O(1) and nbr

mhr= o(1) as (m,n) →

(∞,∞).

7

Page 9: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Assumption 2.1(i) assumes the existence of proper density functions. Assumption 2.1(ii)(iii)

is assumed to ensure that g(x) is twice differentiable and that the second–order derivative

is continuous. In the usual regression setting, such smoothness conditions are imposed di-

rectly on the conditional mean function g(x). Assumption 3.1(iv) implies that (xj, θj) and

ej = θj − g(xj) are i.i.d. random variables. As discussed below, Assumption 2.1(iv) may be

relaxed to the stationary and nonstationary time series case.

Assumption 2.2 imposes that there is a set of data {Xi : i = 1, 2, · · · , n} such that {Xi}has a density function p(x). Assumption 2.3 is a set of standard regularity conditions. Such

conditions are therefore easily verifiable. Assumption 2.3(iii) basically imposes the rate of

convergence on (h, b). When h = C1 ·m−1

4+r and b = C2 · n−1

4+r , Assumption 2.3(iii) reduces

to just nm→ 0.

We now establish the following theorems; their proofs are given in Section 9.1 below.

Theorem 2.1: Let Assumptions 2.1 and 2.3(i) hold. Then as m→∞√∑mj=1K

(xj−xh

)σ2m

(gm(x)− g(x)−

r∑j=1

Bj(x) h2

)→D N(0, σ2(K)), (2.13)

where σ2(K) =∫K2(u)du, Bj(x) =

∫||u||2K(u)du

2·(2f (j)(x)g(j)(x) + f(x)g(jj)(x)

)and σ2

m =1m

∑mj=1 (θj − gm(xj))

2, in which r(j)(x) and r(jj)(x) are the first and second order derivatives

of r(x) = g(x) or f(x), respectively, and f(x) denotes the marginal density function of xj.

Theorem 2.2: Let Assumptions 2.1–2.3 hold. Then as (m,n)→ (∞,∞)√∑ni=1 L

(Xi−xb

)σ2mn

(gmn(x)− g(x)−

r∑j=1

Bj(x) h2

)→D N(0, σ2(L)), (2.14)

where σ2(L) =∫L2(u)du and σ2

mn = 1n

∑ni=1 (θmi − gmn(Xi))

2.

Theorem 2.2 shows that one may use the data set {Xi : 1 ≤ i ≤ n} to re–estimate g(·)and obtain asymptotic consistency. This is mainly because of the following reasoning:

√nbr

(gmn(x)− g(x)−

r∑j=1

Bj(x) h2

)=√nbr (gmn(x)− gm(x))

+

√nbr√mhr

·√mhr

(gm(x)− g(x)−

r∑j=1

Bj(x) h2

)

=√nbr

(gmn(x)− gm(x)−

r∑j=1

Bj(x) h2

)+ oP (1)→D N

(0, σ2(x)

), (2.15)

where σ2(x) > 0 is a variance function.

8

Page 10: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

2.4 Multivariate Case

In this section, we assume that the dimensionality of θ is d > 1. Let x = (x(1), · · · , x(r))τ be

the r–dimensional vector. We impose the following assumptions for the multivariate case.

Assumption 2.4: (i) Assumption 2.1(i) holds.

(ii) Let f(x|θ) be twice differentiable with respect to x such that∫||θ||

∣∣∣∣∣∣f (i)x (x|θ)

∣∣∣∣∣∣ π(θ)dθ <

∞ and∫ ∣∣∣∣∣∣f (i)

x (x|θ)∣∣∣∣∣∣ π(θ)dθ <∞ for any given x and i = 0, 1, 2, where f

(i)x (x|θ) denotes the

i–th partial derivative of f(x|θ) with respect to x.

(iii) Assumption 2.1(iii) holds

(iv) Suppose that {θj : j = 1, 2, · · · ,m} is a vector of i.i.d. random variables drawn from

π(θ) and that {(xj, θj) : j = 1, 2, · · · ,m} is a vector of i.i.d. random variables drawn from

f(x|θ)π(θ).

As θ is now a vector, the conditions corresponding to Assumptions 2.1–2.3 are being

changed. We now establish the following theorems; their proofs are given in Section 9.2

below.

Theorem 2.3: Let Assumptions 2.2, 2.3(i) and 2.4 hold. Then as (m,n)→ (∞,∞)√√√√ m∑j=1

K

(xj − xh

)· Σ−1

m

(gm(x)− g(x)−

r∑j=1

Bjm(x) h2

)→D N

(0, σ2(K) · Id

), (2.16)

where σ2(K) =∫K2(u)du, Σ2

m = 1m

∑mj=1 (θj − gm(xj)) (θj − gm(xj))

τ , Id is the d×d identity

matrix and Bjm(x) is defined in the same way as for Bj(x).

Theorem 2.4: Let Assumptions 2.2–2.4 hold. Then as (m,n)→ (∞,∞) and nbr

mhr→ 0√√√√ n∑

i=1

L

(Xi − xb

)· Σ−1

mn

(gmn(x)− g(x)−

r∑j=1

Bjm(x) h2

)→D N

(0, σ2(L) · Id

), (2.17)

where σ2(L) =∫L2(u)du and Σ2

mn = 1n

∑ni=1 (θmi − gmn(Xi)) (θmi − gmn(Xi))

τ .

It is pointed out that when r, the dimensionality of x, is large, one should use a dimensional–

reduction method, such as, either an additive model or a single–index model, as discussed in

Chapter 2 of Gao (2007), to approximate g(x) and g(Tn) as considered in Section 4 below.

We leave such discussion to future research.

As discussed in Section 3 below, the rate of convergence of gmn(x) can be made faster

than the standard rate when a resampling method is used for generating new samples.

9

Page 11: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

3 Estimation based on Resampling

3.1 Resampling for stationary data

This section considers the case where {Xi : i = 1, 2, · · · , n} is available as a data set and

{Xi} is stationary time series having the same marginal density as p(x). Let f(θ|x) be the

conditional density of θ given Xi = x. Using the Metropolis–Hastings algorithm (see, for

example, Chib and Greenberg 1995), we generate a stationary sequence θi1, · · · , θim from a

proposal density such that as j →∞, the limiting density of θij is f(θ|Xi) (see, for example,

Theorem 3 of Tierney 1994). Note that we need not require {θij; 1 ≤ j ≤ m} to be stationary,

although they may be conditionally stationary given Xi.

Recall g(x) = E[θ|x] =∫θf(θ|x)dθ. The main objective of this section is to estimate

g(x) based on {(θij, Xi) : 1 ≤ i ≤ n; 1 ≤ j ≤ m}. Let eij = θij − g(Xi), θmi = 1m

∑mj=1 θij and

emi = 1m

∑mj=1 eij. Then, we have

θij = g(Xi) + eij, i = 1, 2, · · · , n; j = 1, 2, · · · ,m;

θmi = g(Xi) + emi, i = 1, 2, · · · , n, (3.1)

where {eij : 1 ≤ i ≤ n; 1 ≤ j ≤ m} is assumed to be a stationary sequence in Assumption

3.1 below.

As in Section 2 above, we estimate g(·) by

gmn(x) =n∑i=1

Lni(x)θmi, (3.2)

where Lni(x) =L(Xi−xb )∑nl=1 L

(Xl−xb

) , in which L(·) is a probability kernel function and b is a band-

width parameter.

In order to establish an asymptotic theory for gmn(x), we need to introduce the following

assumptions.

Assumption 3.1: (i) Suppose that {Xi} is a vector of stationary time series data that are

available for generating {θij}. Let eij = θij−E[θij|Xi] and ei = (ei1, · · · , eim)τ . Suppose that

{(ei, Xi)} is a vector of stationary time series satisfying 0 < E[e2ij|Xi = x] = σ2(x) <∞ and

E[e4ij|Xi = x

]= µ4(x) <∞, where σ2(x) is continuous at x.

(ii) Let γj(x) = E[e1,1+je11|X1 = x] satisfy∑∞

j=1 |γj(x)| < ∞ and λ(x) ≡ σ2(x) +

2∑∞

j=1 γj(x) > 0 for each given x. Suppose that {(ei, Xi)} is ρ–mixing with mixing coefficient

ρ(·) satisfying∑∞

k=1 ku√ρ(k) < ∞ for some u > 1

2. In addition, the conditional density of

(X1, Xj) given (em1, emj) is bounded by a positive constant independent of j > 1.

Assumption 3.2: (i) Let f(x|θ) be twice differentiable with respect to x such that∫|θ|∣∣∣∣f (i)

x (x|θ)∣∣∣∣ π(θ)dθ <∞ and

∫ ∣∣∣∣f (i)x (x|θ)

∣∣∣∣ π(θ)dθ <∞

10

Page 12: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

for any given x and i = 0, 1, 2, where f(i)x (x|θ) denotes the i–th partial derivative of f(x|θ)

with respect to x, and || · || denotes the conventional Euclidean norm.

(ii) Suppose that both p2(x) =∫f

(2)x (x|θ)π(θ)dθ and q2(x) =

∫θf

(2)x (x|θ)π(θ)dθ are

continuous in x.

Assumption 3.3: (i) Let L(·) be a bounded probability kernel function satisfying∫vL(v)dv =

0, 0 <∫||v||2L(v)dv <∞ and 0 <

∫L2(v)dv <∞.

(ii) Let the bandwidth b satisfy b→ 0 and nbr(1+ 22u+1

) = O (nc) for some c > 0, where r is

the dimensionality of Xi. In addition, mnbr →∞ and mnbr+4 = O(1) as (m,n)→ (∞,∞).

There case where {Xi} is a sequence of i.i.d. random variables is covered in Assumption

3.1. The verification of Assumptions 3.1–3.3 may be done similarly to what has been done for

Assumptions 2.1–2.3. While the assumptions may not be the weakest ones, they are easily

verifiable. The stationarity assumption is based on the nature of the MCMC algorithm. The

mixing condition is also standard while the fourth moment condition on E[e4ij] < ∞ may

be weakened to E[|eij|2+c(e)

]< ∞ for some c(e) > 0. Assumption 3.2 is needed to ensure

that the second–order derivative of g(x), g(2)(x), is continuous. The bandwidth conditions

assumed in Assumption 3.3(ii) are also quite standard.

We now establish the following theorem; its proof is given in Section 9.3 below.

Theorem 3.1: Suppose that Assumptions 3.1–3.3 are satisfied. Then, we have as (m,n)→(∞,∞) √√√√m ·

n∑i=1

L

(Xi − xb

)(gmn(x)− g(x)−

r∑j=1

Bj(x) b2

)→D N (0,Σ(x)) , (3.3)

where Bj(x) is the same as defined in Theorem 2.1 and Σ(x) = λ(x) ·∫L2(v)dv, in which

λ(x) = σ2(x) + 2∑∞

j=1 γj(x).

Theorem 3.1 shows that one can achieve a fast rate of convergence of an order of the

form(√

mnbr)−1

= m−12 ·(√

nbr)−1

= o

((√nbr)−1)

as m → ∞, because our estimation

method makes the best use of the availability of the sample (X1, · · · , Xn). The finite sample

evaluation given in Section 6 below supports this fast rate of convergence. In the following

subsection, we consider the case where a summary statistic is available for resampling.

3.2 Resampling for nonstationary data

Since {Xi : i = 1, 2, · · · , n} is available as a nonstationary time series in many practical

situations, this section considers the nonstationary case. Using the Metropolis–Hastings

algorithm (see, for example, Chib and Greenberg 1995) again, we generate an array of random

11

Page 13: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

variables, θi1, · · · , θim, from a proposal density for each given Xi. Once again, we need not

require {θij; 1 ≤ j ≤ m} to be stationary. Consider the case of r = 1 in this subsection.

We then assume that there are an array of martingale differences {εij} and a suitable

function g(·) such that

θij = g(Xi) + εij, i = 1, 2, · · · , n; j = 1, 2, · · · ,m;

θmi = g(Xi) + εmi, i = 1, 2, · · · , n, (3.4)

where θmi = 1m

∑mj=1 θij and εmi = 1

m

∑mj=1 εij.

We then estimate g(·) by

gmn(x) =n∑i=1

Lni(x)θmi, (3.5)

where Lni(x) =L(Xi−xb )∑nl=1 L

(Xl−xb

) , in which L(·) is a probability kernel function and b is a band-

width parameter.

For the case where Xi is nonstationary and Ui = Xi −Xi−1 reduces to be stationary, we

modify Assumptions 3.1–3.3 as follows.

Assumption 3.4: (i) Suppose that {εmi} and Ui are independent of each other. Suppose also

that there is a stochastic process B(r) such that sup0≤r≤1

∣∣∣X[nr]√n−B(r)

∣∣∣ = oP (1) as n→∞.

(ii) Let {εij,Fmi : 1 ≤ i ≤ n} be an array of martingale differences with E [εij|Fm,i−1] = 0

and max1≤j≤mE[ε4ij|Fm,i−1

]<∞ almost surely (a.s.). Moreover, there is some 0 < σ2

ε <∞such that 1

m

∑mj=1E

[ε2ij|Fm,i−1

]→a.s. σ

2ε and 1

m

∑mj1=2

∑j1−1j2=1E [εij1εij2|Fm,i−1] →a.s. 0 as

m→∞.

Assumption 3.5: (i) Let f(x|θ) be twice differentiable with respect to x such that∫|θ|∣∣∣∣f (i)

x (x|θ)∣∣∣∣ π(θ)dθ <∞ and

∫ ∣∣∣∣f (i)x (x|θ)

∣∣∣∣ π(θ)dθ <∞

for any given x and i = 0, 1, 2, where f(i)x (x|θ) denotes the i–th partial derivative of f(x|θ)

with respect to x, and || · || denotes the conventional Euclidean norm.

(ii) Suppose that both p2(x) =∫f

(2)x (x|θ)π(θ)dθ and q2(x) =

∫θf

(2)x (x|θ)π(θ)dθ are

continuous in x.

Assumption 3.6: (i) Let L(·) be a bounded probability kernel function satisfying∫vL(v)dv =

0, 0 <∫||v||2L(v)dv <∞ and 0 <

∫L2(v)dv <∞.

(ii) Let the bandwidth b satisfy b → 0, m√nb → ∞ and m

√nb5 → c(0) for some

0 < c(0) <∞.

The verification of Assumptions 3.4–3.6 may be done in a similar way to those of Theorem

3.2 of Gao and Phillips (2013). We then establish the following theorem; its proof is given in

Section 9.4 below.

12

Page 14: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Theorem 3.2: Suppose that Assumptions 3.4–3.6 are satisfied. Then, we have as (m,n)→(∞,∞) √√√√m

n∑i=1

L

(Xi − xb

)(gmn(x)− g(x)−B(x) b2

)→D N

(0,Σ2

1

), (3.6)

where Σ21 = σ2

ε ·∫L2(v)dv and B(x) =

∫u2L(u)du

2·(2f (1)(x)g(1)(x) + f(x)g2(x)

). Note that we

also have 1√nb

∑ni=1 L

(Xi−xb

)→D LW (1, 0), in which LW (1, 0) is a local–time random variable

driven by a standard Brownian process W (r).

In comparison with Theorem 3.2 of Gao and Phillips (2013), Theorem 3.2 establishes a

much faster rate of (m√nb)− 1

2 than (√nb)− 1

2 when a sampling method is used. Meanwhile,

the multivariate case may be done similarly to Chapter 2 of Gao (2007), and Gao and Phillips

(2013) when a semiparametric reduction method is used. When there is a type of endogeneity,

bias corrections may be done similarly to Phillips and Hansen (1990).

4 Estimation Based on Summary Statistics

We consider the case where we may use a summary statistic based on direct sampling,

importance sampling and resampling in Sections 4.1–4.3, respectively.

4.1 Estimation based on direct sampling

In econometric estimation problems, the parameter–of–interest, θ, is often involved in a

complex model, such as, a structural model of the form ψ(X; θ) = 0. Instead of estimating a

conditional mean of the form E[θ|X], we may make the best use of the availability of some

summary statistics. In this case, we may just be interested in estimating the conditional

mean g(Tn) = E [θ|Tn], where Tn is a one–dimensional summary statistic, such as the sample

mean Tn = 1n

∑ni=1Xi, in which X1, X2, · · · , Xn is a sequence of i.i.d. random variables.

Suppose that we may sample {θj : 1 ≤ j ≤ m} from π(θ) and then (θj, Tnj) from

f(Tn|θ)π(θ). We then estimate g(Tn) by

gkm(Tn) =

∑mj=1 K

(Tnj−Tn

h

)θj∑m

j=1 K(Tnj−Tn

h

) , (4.1)

where K(·) is a probability kernel function and h is a bandwidth parameter.

In some situations, we may estimate g(Tn) by

gam(Tn) =

∑mj=1 f(Tn|θj)θj∑mj=1 f(Tn|θj)

(4.2)

when f(Tn|θ) is available for feasible computation.

13

Page 15: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

We will show that gkm(Tn) is more efficient than Tn and gam(Tn) in terms of their standard

deviations. In order to establish an asymptotic theory for gkm(Tn), we introduce the following

assumptions.

Assumption 4.1: (i) Let the product of f(Tn|θ) and π(θ) be a proper probability density

function for each given n ≥ 1.

(ii) Suppose that (Tnj, θj) is a vector of i.i.d. random variables sampled from f(Tn|θ)π(θ),

and that Tnj and Tn are independent of each other and have the same distribution. For each

given n ≥ 1, let enj = θj − g (Tnj) be a sequence of i.i.d. errors independent of {Tnj} with

E [en1] = 0, 0 < E [e2n1] <∞ and E [e4

n1] <∞.

(iii) Let f(x|θ) be twice differentiable with respect to x such that

supn≥1

∫|θ|∣∣∣∣∣∣f (i)

1 (Tn|θ)∣∣∣∣∣∣ π(θ)dθ <∞ and sup

n≥1

∫ ∣∣∣∣∣∣f (i)1 (Tn|θ)

∣∣∣∣∣∣ π(θ)dθ <∞

for i = 0, 1, 2, where f(i)1 (Tn|θ) denotes the i–th partial derivative of f(Tn|θ) with respect to

Tn, and || · || denotes the conventional Euclidean norm.

(iii) Suppose that both p2(x) =∫f

(2)x (x|θ)π(θ)dθ and q2(x) =

∫θf

(2)x (x|θ)π(θ)dθ are

continuous in x.

Assumption 4.2: (i) Let K(·) be the probability kernel function satisfying∫uK(u)du = 0,

0 <∫u2K(u)du <∞, and 0 <

∫K2(u)du <∞ and 0 <

∫u2K2(u)du <∞.

(ii) Let the bandwidth h satisfy h → 0, mh → ∞, mhλ2n→ ∞,

√mh5

λnγ(Tn) → Cλ for

some 0 < Cλ < ∞ and hλn· γ(Tn) → 0 as (m,n) → (∞,∞), where λ2

n = Var[θ|Tn] and

γ(Tn) = 12g(2)(Tn) + p(1)(Tn)

p(Tn)g(1)(Tn) + g(1)(Tn), in which p(Tn) =

∫f(Tn|θ)π(θ)dθ.

Assumption 4.1 corresponds to Assumption 2.1. Assumption 4.2(i) is similar to Assump-

tion 2.3(i). Assumption 4.2(ii) imposes a set of additional conditions on the relationship

among (m,h, λn) for the establishment of an asymptotic normality for gkm(Tn).

We then introduce the following assumption for the establishment of an asymptotic nor-

mality for gam(Tn).

Assumption 4.3: (i) Suppose that {θj} is a sequence of i.i.d. random variables drawn from

π(θ).

(ii) For given Tn, suppose that σ2(Tn) =∫ (

θf(Tn|θ)− q(Tn)p(Tn)

f(Tn|θ))2

π(θ)dθ satisfies

infn≥1 σ2(Tn) > 0 and m p2(Tn)

σ2(Tn)→∞ as (n,m)→ (∞,∞), where q(Tn) =

∫θf(Tn|θ)π(θ)dθ.

(iii) For given Tn, suppose that supn≥1

∫ ∣∣∣θf(Tn|θ)− q(Tn)p(Tn)

f(Tn|θ)∣∣∣4 π(θ)dθ <∞.

Assumption 4.3(ii)(iii) can be easily verifiable in many cases. For example, we consider

case where Tn = 1n

∑nj=1 Xi ∼ N

(θ, 1

n

)when Xi ∼ N(θ, 1) and θ ∼ N(0, 1). Note that in

this case, we have f(Tn|θ) = 1√2πσn

e− (Tn−θ)2

2σ2n , p(Tn) = 1√2π(1+σ2

n)e− T2

n2(1+σ2n) , q(Tn) = p(Tn) · Tn

1+σ2n

and σ2(Tn) = 12π· 1

2+σ2n

exp(− T 2

n

2+σ2n

)·(

12+σ2

n+ 4T 2

n

(2+σ2n)2

), where σ2

n = 1n.

14

Page 16: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

With this setting, Assumption 4.2(ii) requires that (m,n, h) satisfies m n h→∞, nh2 →0, m n h5 = O(1) and m n h7 → 0. Such conditions are satisfiable when (m,n) is suitably

chosen. In the case n = [mc] for some 0 < c < 1, one may just need to choose c such that

m1+ch5 = O(1) and mch2 → 0. As shown in Theorem 4.1 below, with the possibility to

sample (θj, Tnj) from f(Tn|θ)π(θ), the asymptotic variance of gkm(·) is of an order of the

form 1mnh

, which is smaller than the conventional order of 1nh

, while the bias term remains

the same.

Assumption 4.1(i)(ii) just imposes the i.i.d. structure on (θj, Tnj). Assumptions 4.1(iii)

and 4.3 then impose some moment conditions on f(Tn|θ). In other words, there is no need

to require assuming an explicit or implicit distributional structure or even asymptotically

distributional structure on Tn.

We establish two important asymptotic distributions in Theorems 4.1 and 4.2 below; their

proofs are given in Sections 9.5 and 9.6, respectively.

Theorem 4.1: Let Assumptions 4.1 and 4.2 hold. Then as (m,n)→ (∞,∞)√√√√∑mj=1 K

(Tnj−Tn

h

)λ2n

(gkm(Tn)− g(Tn)−B(Tn) h2

)→D N

(0, σ2(K)

), (4.3)

where σ2(K) =∫K2(u)du, and B(Tn) =

∫u2K(u)du

2·(2p(1)(Tn)g(1)(Tn) + p(Tn)g2(Tn)

).

Theorem 4.2: Let Assumption 4.3 hold. Then as (m,n)→ (∞,∞)√m p2(Tn)

σ2(Tn)(gam(Tn)− g(Tn))→D N(0, 1), (4.4)

where σ2(Tn) is as defined in Assumption 4.3(ii).

Let us now compare the rates of convergence in (4.3) and (4.4). Note that the rate of

convergence in (4.3) of an order of√m ·√nh is faster than the rate of convergence of

√m

involved in (4.4) when nh→∞, n λ2n → C1 > 0 and p(Tn)

σ(Tn)→P C2 > 0.

In general, we conclude that gkm(·) has a faster rate of convergence than that for gam

as long as√h σ(Tn)

λn p(Tn)→ ∞, which is easily verified when nh → ∞, n λ2

n → C1 > 0 andp(Tn)σ(Tn)

→P C2 > 0. Such an asymptotic behaviour is verified by the finite–sample evaluation

in Sections 6 and 7 below.

4.2 Estimation based on importance sampling

In the above discussions, we assume that π(θ) is a proper probability density function and

it is feasible to sample {θj} from π(θ). Let λ(θ) be the importance distribution and θ∗j be

sampled from λ(θ). We also assume that the ratio π(θ)λ(θ)

is available for computation.

15

Page 17: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Let ρ(θ) = π(θ)λ(θ)

. Suppose that we can sample {θ∗j} from λ(θ) and (θ∗j , T∗nj) from f(Tn|θ)λ(θ).

Then, we can replace gkm(·) and gam(·) by

g∗km(Tn) =

∑mj=1K

(T ∗nj−Tn

h

)ρ(θ∗j )θ

∗j∑m

j=1K(T ∗nj−Tn

h

)ρ(θ∗j )

, (4.5)

g∗am(Tn) =

∑mj=1 f(Tn|θ∗j )ρ(θ∗j )θ

∗j∑m

j=1 f(Tn|θ∗j )ρ(θ∗j ). (4.6)

In this case, Assumptions 4.1–4.3 may be replaced by Assumptions 4.1*–4.3* below.

Assumption 4.1*: (i) Let the product of f(Tn|θ) and λ(θ) be a proper probability density

function for each given n ≥ 1.

(ii) Suppose that (T ∗nj, θ∗j ) is a vector of i.i.d. random variables sampled from f(Tn|θ)λ(θ).

Let e∗j = θ∗j − g(T ∗nj)

be a sequence of i.i.d. errors with E[e∗j |T ∗nj

]= 0, 0 < E

[e∗2j |T ∗nj

]<∞

and E[e∗4j |T ∗nj

]<∞.

(iii) Let f(x|θ) be twice differentiable with respect to x such that

supn≥1

∫|θ|∣∣∣∣∣∣f (i)

1 (Tn|θ)∣∣∣∣∣∣ ρ(θ)λ(θ)dθ <∞ and sup

n≥1

∫ ∣∣∣∣∣∣f (i)1 (Tn|θ)

∣∣∣∣∣∣ ρ(θ)λ(θ)dθ <∞

for i = 0, 1, 2, where f(i)1 (Tn|θ) denotes the i–th partial derivative of f(Tn|θ) with respect to

Tn, and || · || denotes the conventional Euclidean norm.

(iii) Suppose that both p2(x) =∫f

(2)x (x|θ)ρ(θ)λ(θ)dθ and q2(x) =

∫θf

(2)x (x|θ)ρ(θ)λ(θ)dθ

are continuous in x.

Assumption 4.2*: Let Assumption 4.2 hold with σ2(Tn) being replaced by σ∗2(Tn) =

Var[θ∗|Tn].

Assumption 4.3*: (i) Suppose that {θ∗j} is a sequence of independent and identically

distributed (i.i.d.) random variables drawn from λ(θ).

(ii) For given Tn, suppose that σ2∗(Tn) =∫ (

θf(Tn|θ)− q∗(Tn)p∗(Tn)

f(Tn|θ))2

ρ2(θ)λ(θ)dθ

satisfies infn≥1 σ2∗(Tn) > 0 and m p∗ 2(Tn)

σ2(Tn)→ ∞, where p∗(Tn) =

∫f(Tn|θ)ρ(θ)λ(θ)dθ and

q∗(Tn) =∫θf(Tn|θ)ρ(θ)λ(θ)dθ.

(iii) For given Tn, suppose that supn≥1

∫ ∣∣∣θf(Tn|θ)− q(Tn)p(Tn)

f(Tn|θ)∣∣∣4 ρ4(θ)λ(θ)dθ <∞.

We then have the following theorems.

Theorem 4.1*: Let Assumptions 4.1* and 4.2* hold. Then, as (m,n)→ (∞,∞)√√√√∑mj=1 K

(T ∗nj−Tn

h

)ρ(θ∗j )

λ2n

(g∗km(Tn)− g(Tn)−B(Tn) h2

)→D N

(0, σ∗2(K)

), (4.7)

whenever∫ρ2(θ)λ(θ)dθ∫ρ(θ)λ(θ)dθ

> 0, where σ∗2(K) =∫K2(u)du ·

∫ρ2(θ)λ(θ)dθ∫ρ(θ)λ(θ)dθ

and B(Tn) is the same as

defined in Theorem 4.1.

16

Page 18: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Theorem 4.2*: Let Assumption 4.3* hold. Then, as (m,n)→ (∞,∞)√m p∗2(Tn)

σ∗2(Tn)(g∗am(Tn)− g(Tn))→D N(0, 1), (4.8)

where σ∗2(Tn) is as defined in Assumption 4.3*(ii).

The proofs and implications of Theorem 4.1* and 4.2* are almost the same as those of

Theorems 4.1 and 4.2, and are therefore omitted.

4.3 Estimation based on resampling of summary statistics

Let Tn be a summary statistic and denote its density function by pn(·) and fn(θ|Tn) be the

conditional density of θ given Tn, where n is the number of observations involved in Tn.

Suppose that one can sample a stationary sequence Tn1, · · · , TnN from the distribution of Tn

and then an array of random variables {θnij : 1 ≤ j ≤ m} from a proposal density such that

as m→∞, the limiting density of θnij is fn(θ|Tni) for each fixed i.

Similarly to equation (3.1), we may write

θnij = g(Tni) + εnij, i = 1, 2, · · · , N ; j = 1, 2, · · · ,m,

θni = g(Tni) + εni, i = 1, 2, · · · , N, (4.9)

where θni = 1m

∑mj=1 θnij and εni = 1

m

∑mj=1 εnij.

In the same way as in equation (4.1), we estimate g(·) by

gmnN(Tn) =

∑Ni=1K

(Tni−Tn

h

)θni∑N

i=1 K(Tni−Tn

h

) , (4.10)

where K(·) is a probability kernel function and h is a bandwidth parameter.

Combining the establishments and the proofs of Theorems 3.1 and 4.1, we have as

(m,n,N)→ (∞,∞,∞)√m∑N

i=1K(Tni−Tn

h

)λ2n

(gmnN(Tn)− g(Tn)−B(Tn) h2

)→D N

(0,Σ2(K)

), (4.11)

where Σ2(K) is defined in the same way as for Σ(x) involved in Theorem 3.1, and λ2n and

B(Tn) are defined in the same way as in Theorem 4.1.

Note that the rate of convergence involved in (4.11) can be as fast as m−12 ·(√

n Nh)− 1

2.

In comparison with Theorem 4.1, the rate of convergence of gmnN(Tn) can be improved by

an order of m−12 when a resampling algorithm is used. The estimation theory is applied to a

stationary GARCH model in Sections 6 and 7 below. The case involving summary statistics

of nonstationary data may be discussed analogously.

17

Page 19: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

5 Estimation with Hyperparameters

In the discussion so far, we assume that the prior density functions π(θ) and λ(θ) do not

involve any hyperparameters. Thus, there is no parameter involved in g(x) and g(Tn). Both

functions are estimated nonparametrically by the kernel method even though the functional

forms of f(x|θ), f(Tn|θ) and π(θ) may be assumed to be known for sampling and computation.

This section considers the case where there is a vector of hyperparameters involved in π(θ),

denoted as π(θ; γ). We still assume that π(θ; γ) is available for sampling and computation

when the value of γ is given. We introduce the following definition:

g(x; γ) =

∫θf(x|θ)π(θ; γ)dθ∫f(x|θ)π(θ; γ)dθ

≡ q(x; γ)

p(x; γ), (5.1)

where p(x; γ) =∫f(x|θ)π(θ; γ)dθ and q(x; γ) =

∫θf(x|θ)π(θ; γ)dθ.

As pointed out before, the functional form of g(x; γ) in most cases is not available for

computation. Since γ is the same involved in q(x; γ) as in p(x; γ), we propose to estimate γ

by a nonparametric maximum likelihood estimation method. Similar ideas have been used

in Kristensen and Shin (2012) for estimating unknown parameters involved in a class of fully

parametric models.

For each given γ, suppose that we can sample {(θj = θj(γ), xj = xj(γ)) : 1 ≤ j ≤ m}from f(x|θ)π(θ; γ). For each given γ, we then estimate p(x; γ) by

pm(x; γ) =1

mhr

m∑j=1

L

(xj(γ)− x

h

), (5.2)

where r is the dimensionality of x = (x1, · · · , xr)τ , L(·) is a probability density function

defined on Rr, and h is a bandwidth parameter.

Assume that the data set {Xi : i = 1, 2, · · · , n} available for us is a sequence of i.i.d.

random variables. Define a normalised log–likelihood function of the form

Ln(γ) =1

νn

n∑i=1

log(p(Xi; γ)), (5.3)

where νn →∞ is a sequence of positive real numbers.

We estimate γ by γn = arg maxγ∈Γ Ln(γ), in which Γ is a subset of Rc with c being the

dimensionality of γ.

The corresponding version for pm(x; γ) is then defined as

Lmn(γ) =1

νn

n∑i=1

log(pm(Xi; γ)), (5.4)

and γ is then estimated by γmn = arg maxγ∈Γ Lmn(γ). Note that in practice, Lmn(γ) may

need to replaced by a truncated form. Such an issue is discussed in Remark A.1 just after

the proof of Theorem 5.1 in the Appendix below.

18

Page 20: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

In order to study asymptotic properties for γn and γmn, we need to introduce the following

notation:

Sn(γ) =∂Ln(γ)

∂γ, Hn(γ) =

∂2Ln(γ)

∂γ∂γ′,

Gni(γ) =∂3Ln(γ)

∂γ∂γ′∂γi, In(γ) =

1

νn

n∑i=1

E

[∂ log(p(X1; γ)

∂γ

∂ log(p(X1; γ)

∂γ′

],

ln(γ) = diag(In(γ)) = the diagonal elements of matrix In(γ),

Un(γ) = l−1n (γ)Sn(γ), Vn(γ) = l

− 12

n (γ)Hn(γ)l− 1

2n (γ) and

Wni(γ) = l− 1

2n (γ)Gni(γ)l

− 12

n (γ). (5.5)

We are now able to introduce the following assumptions.

Assumption 5.1: (i) The parameter space is given by a sequence of local neighbourhoods:

Γn = {γ : ||√ln(γ) (γ − γ0) || ≤ ε} ⊂ Rr, where γ0 is the true value of γ, ε > 0 and

l−1n (γ) = OP (1).

(ii) Let the data set {Xi} be a sequence of i.i.d. random variables having the same density

function as p(x; γ). Let p(x; γ) and p2(x; γ) =∫f

(2)x (x|θ)π(θ; γ)dθ satisfy

supx∈Rr

supγ∈Γn

(p−1(x; γ) [1 + p2(x; γ)]

)<∞.

(iii) Ln(γ) is three times differentiable with its derivatives satisfying:

(a)(√

νnUn(γ0), Vn(γ0))→D (U∞, V∞), where U∞ and V∞ are random variables with

P (V∞ < 0) = 1; and

(b) max1≤i≤r supγ∈Γn ||Wni(γ)|| = OP (1).

Assumption 5.2: (i) For each given γ, {xj = xj(γ)} is a sequence of i.i.d. random variables.

The function xj(γ) is a differentiable with respect to γ and xj(γ) denotes the first–order

derivative with maxj≥1E [||xj(γ)||2] <∞ for each given γ.

(ii) Define

B1(γ) = supx∈Rd

E [x1(γ)|x1(γ) = x] p(x; γ) and

B2(γ) = supx∈Rd||x||δ0 E [x1(γ)|x1(γ) = x] p(x; γ)

for some δ0 ≥ r. There are some 0 < C1, C2 < ∞ such that supγ∈ΓBi(γ) ≤ Ci < ∞ for

i = 1, 2.

Assumption 5.3: (i) Let L(·) be a symmetric and bounded probability kernel function.

(ii) There are some constants 0 < C,C1, C2 <∞ and µ > 1 such that L(u) is differentiable

with∣∣∣∣∣∣∂L(u)

∂u

∣∣∣∣∣∣ ≤ C1 and∣∣∣∣∣∣∂L(u)

∂u

∣∣∣∣∣∣ ≤ C2 |||u||µ when ||u|| ≥ C. In addition, L(u) ≤ C3 ||u||µ

when ||u|| ≥ C for some 0 < C3 <∞.

(iii) The bandwidth h satisfies h→ 0, mhr →∞ and log(m)m hr

→ 0 as m→∞.

19

Page 21: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

(iv) The sequence νn satisfies nh2 = o(√

νn)

and n2 log(m)mhr

= o(νn) as (m,n)→ (∞,∞).

Assumptions 5.1–5.3 are standard for this kind of problem. For example, Assumption 5.1

is similar to Assumptions C1 and C3 of Kristensen and Shin (2012). Assumption 5.1 can be

simplified to a set of standard conditions that are similar to the conditions of Theorems 4.1.2–

4.1.4 of Amemiya (1985) for the standard maximum likelihood estimation. Assumptions 5.2

and 5.3(i)(ii)(iii) are simplified versions of A.2–A.5 of Kristensen (2009). Assumption 5.3(iv)

imposes some conditions on the rates of convergence of (m,n). For example, when νn = n, it

simply requires nh4 → 0 and n log(m)mhr

→ 0, which can be satisfied with many different choices

of (m,n). While the conditions may not be the weakest possible, they are sufficient for the

establishment of the main theorem; its proof is given in Section 9.7.

Theorem 5.1: Let Assumptions 5.1–5.3 hold. Then, as n→∞

√νn l

12n (γ0) (γmn − γ0)→D ξ, (5.6)

where ξ = −U∞V∞

with (U∞, V∞) being defined in Assumption 5.1(iii).

Theorem 5.1 shows that γmn is still able to achieve the standard rate of convergence of an

order of n−12 when νn = n, even though a kernel estimation is involved in the construction of

the nonparametric maximum likelihood function. This is basically because of the following

derivations:

√νn l

12n (γ0) (γmn − γ0) =

√νn l

12n (γ0) (γmn − γn) +

√νn l

12n (γ0) (γn − γ0)

=√νn l

12n (γ0) (γn − γ0) + oP (1)→D ξ (5.7)

as shown in Section 9.7 below. With a suitable rate of convergence for γmn to γ, the corre-

sponding versions of g(x; γ) or g(Tn; γ) may be consistently estimated in the same way as in

Sections 2–4.

We now conclude the establishment of the theory of this paper. In Sections 6 and 7 below,

we will evaluate the proposed theory and estimation methods using both simulated and real

data examples.

6 Numerical Evidence

This section evaluates the finite sample performance of the theory and the proposed estima-

tion methods. Examples 6.1–6.3 consider the case where {(xj, θj)} can be sampled directly

from f(x|θ)π(θ) and {Xi; 1 ≤ i ≤ n} is drawn from p(x). Example 6.4 consider the case

where an MCMC algorithm is used for sampling.

Example 6.1: Consider the case where f(x|θ) = 1√2πe−

(x−θ)22 for −∞ < x < ∞, and

π(θ) = 1√2πe−

θ2

2 . This implies g(x) = x2.

20

Page 22: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

In this example and Example 6.2 below, we use (m,n) = (2000, 200). Figures 1 and 2

below give the corresponding estimates of g(x) for Example 6.1.

−5 −4 −3 −2 −1 0 1 2 3 4 5−4

−3

−2

−1

0

1

2

3

4

5

x

θ

(xj , θj): Sample data

gm(x) =∑m

j=1 Kmj(x)θj

Figure 1: hcv = 0.2204, gm(x) =∑m

j=1Kmj(x)θj

−5 −4 −3 −2 −1 0 1 2 3 4 5−3

−2

−1

0

1

2

3

x

g(x) = x/2

gm(x) =∑m

j=1 Kmj(x)θj

gmn(x) =∑n

i=1 Lmi (x)θmi

Figure 2: For gmn(x) =∑n

i=1 Lmi(x)θmi, hcv = 0.6720; For gm(x) =∑m

j=1Kmj(x)θj, hcv =

0.453, Xi is drawn from p(x) = N(0, 2).

Example 6.2: Consider the case where f(x|θ) = 1√2πe−

(x−θ)22 for −∞ < x < ∞, and

π(θ) = 16I[−3 ≤ θ ≤ 3]. This implies

p(x) =

∫ ∞−∞

f(x|θ)π(θ)dθ =1

6(Φ(x+ 3)− Φ(x− 3))

q(x) =

∫ ∞−∞

θf(x|θ)π(θ)dθ =1

6√

(e−

(x+3)2

2 − e−(x−3)2

2

)+ x p(x), (6.1)

which implies g(x) = q(x)p(x)

= x+ 1√2π·

(e−

(x+3)2

2 −e−(x−3)2

2

)Φ(x+3)−Φ(x−3)

, in which Φ(u) =∫ u−∞

1√2πe−

v2

2 dv.

21

Page 23: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Figures 3 and 4 below give the corresponding estimates of g(x) for Example 6.2.

−6 −4 −2 0 2 4 6

−3

−2

−1

0

1

2

3

x

θ

(xj , θj): Sample data

gm(x) =∑m

j=1 Kmj(x)θj

Figure 3: hcv = 0.2946, gm(x) =∑m

j=1Kmj(x)θj

−6 −4 −2 0 2 4

−3

−2

−1

0

1

2

3

x

g(x) = q(x)/p(x)

gm(x) =∑m

j=1 Kmj(x)θj

gmn(x) =∑n

i=1 Lmi (x)θmi

Figure 4: For gmn(x) =∑n

i=1 Lmi(x)θmi, hcv = 0.4812; For gm(x) =∑m

j=1Kmj(x)θj, hcv =

0.2946, Xi is drawn from p(x).

Figures 1–4 show that the proposed estimates gm(x) and gmn(x) are both very close to

the true function g(x).

Example 6.3: Let Xi = θ0 + ei, where ei ∼ N(0, 1). Parameter θ can be estimated by

Tn = 1n

∑ni=1Xi, and θ0 = 1. In this case, we have g(Tn) = n

1+nTn.

Let θj ∼ N(0, 1) and generate Tnj by a mean model of the form:

Tnj = θj + σn ξj, j = 1, 2, · · · ,m, (6.2)

where θj ∼ N(0, 1), ξj ∼ N(0, 1) and σ2n = 1

n.

22

Page 24: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

To evaluate the finite sample performance of the proposed estimates, we use the following

estimators:

gm(Tn) =

∑mj=1 θjf(Tn|θj)∑mj=1 f(Tn|θj)

, gm(Tn) =

∑mj=1K

(Tnj−Tn

h

)θj∑m

j=1K(Tnj−Tn

h

) , (6.3)

g∗m(Tn) =

∑mj=1K

(Tnj−Tn

h

)θ∗Nj∑m

j=1K(Tnj−Tn

h

) , gN(Tn) =1

N

N∑i=1

θi(Tn). (6.4)

where the last one denotes the sample mean of the direct MCMC draws, and K(x) = 1√2πe−

x2

2

and h is a bandwidth chosen by the leave–one–out cross–validation method.

The main task is summarised as follows. Consider Case A: (m,n) = (1000, 100); Case

B: (m,n) = (2000, 200). Let the number of replications, M = 1000.

• Calculate bias1 =∣∣T n − θ∣∣ and std1 =

√1M

∑Mi=1

(Tn(i)− T n

)2, where T n = 1

M

∑Mi=1 Tn(i),

and Tn(i) is the value at i–th replication of Tn = 1n

∑ni=1Xi;

• Calculate bias2 =∣∣∣gm(T n)− g(T n)

∣∣∣ and std2 =

√1M

∑Mi=1

(gmi(T n)− gm(T n)

)2

; and

• Calculate bias3 =∣∣∣gm(T n)− g(T n)

∣∣∣ and std3 =

√1M

∑Mi=1

(gmi(T n)− gm(T n)

)2

.

Table 6.1: The biases and standard deviations

case A: (m,n) = (1000, 100) case B: (m,n) = (2000, 200)

bias1 0.0013 std1 0.1761 bias1 0.0007 std1 0.1179

bias2 0.0297 std2 0.0303 bias2 0.0152 std2 0.0159

bias3 0.0189 std3 0.0097 bias3 0.0094 std3 0.0056

Table 6.1 shows that gmk(·) is the best performer in terms of the finite sample performance

of the standard deviations, while Tn has the smallest bias in each case. This is because Tn is

an unbiased estimator of θ.

In Example 6.4 below, we will consider the case where resamples are used in the nonpar-

metric kernel estimation.

Example 6.4: Simulate Xi = θ0 + ei with ei ∼ N(0, 1) for i = 1, 2, · · · , n, where θ0 = 1.

Let Tn = 1n

∑ni=1Xi and generate Tnj as in Example 6.1. Generate θij from f(θ|Tnj) for

i = 1, 2, · · · , N and each fixed j for 1 ≤ j ≤ m.

We use the following proposal density functions for the generations of the resamples.

23

Page 25: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

• Mixture density 1: Proposal density

f(θ|Tn) = 12√

2πexp

(− θ2

2

)+ 1

2√

2πT 2n

exp(− θ2

2T 2n

), implying g(Tn) = 0.

• Mixture density 2: Proposal density

f(θ|Tn) = 12√

2πexp

(− (θ−1)2

2

)+ 1

2√

2πexp

(− (θ−Tn)2

2

), implying g(Tn) = 0.5 + 0.5Tn.

• Mixture density 3: Proposal density

f(θ|Tn) = 1√2π

exp(− θ2

2

)+ 1

2(1+T 2

n) exp (−(1 + T 2n)θ), implying g(Tn) = 1√

2π+ 1

2(1+T 2n)

.

As in Example 6.3, we use K(x) = 1√2πe−

x2

2 and choose h by the leave–one–out cross–

validation method. Let the number of replications, M = 1000. We consider two cases: Case

A: (m,n,N) = (1000, 100, 1000), and Case B: (m,n,N) = (2000, 200, 2000).

Define gNi(Tn(i)) = 1N

∑Nj=1 θj(Tn(i)) and gN(T n) = 1

M

∑Mi=1 gNi(Tn(i)), where T n =

1M

∑Mi=1 Tn(i), and Tn(i) is the value at i–th replication of Tn = 1

n

∑ni=1 Xi.

• Calculate bias4 =∣∣gN(T n)− g(T n)

∣∣ and std4 =

√1M

∑Mi=1

(gNi(Tn(i))− gN(T n)

)2;

• Calculate bias5 =∣∣∣gm(T n)− g(T n)

∣∣∣ and std5 =

√1M

∑Mi=1

(gmi(T n)− gm(T n)

)2

; and

• Calculate bias6 =∣∣∣g∗m(T n)− g(T n)

∣∣∣ and std6 =

√1M

∑Mi=1

(g∗mi(T n)− g∗m(T n)

)2

.

Table 6.2: The biases and standard deviations

Case A Case B

Mixture density 1

bias4 0.0048 std4 0.0317 bias4 0.0022 std4 0.0261

bias5 0.0185 std5 0.0089 bias5 0.0105 std5 0.0053

bias6 0.0011 std6 0.0014 bias6 0.0006 std6 0.0007

Mixture density 2

bias4 0.0069 std4 0.0281 bias4 0.0031 std4 0.0206

bias5 0.0192 std5 0.0092 bias5 0.0115 std5 0.0050

bias6 0.0009 std6 0.0008 bias6 0.0004 std6 0.0005

Mixture density 3

bias4 0.0053 std4 0.0328 bias4 0.0038 std4 0.0239

bias5 0.0178 std5 0.0096 bias5 0.0094 std5 0.0048

bias6 0.0012 std6 0.0011 bias6 0.0007 std6 0.0006

24

Page 26: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

The estimated biases and standard deviations are given in Table 6.2 above. Table 6.2

shows that the nonparametric estimate based on the resamples has the smallest standard

deviation in each case, and gmk(·) is better than the sample average of the resamples in terms

of the standard deviations. In addition, Table 6.2 also shows that the biases and standard

deviations are smaller for the case of (N,m, n) = (2000, 2000, 200) than those for the case of

(N,m, n) = (1000, 1000, 100).

In the following example, we consider estimating unknown parameters involved in a

GARCH model before we use the same model in an empirical evaluation.

7 GARCH model estimation and implementation

Consider a GARCH (1,1) model of the form:

yt = σtεt, εt ∼ N(0, 1), t = 1, 2, · · · , n,

σ2t = b0 + b1y

2t−1 + b2σ

2t−1,

where θ = (b0, b1, b2)′ denotes a vector of unknown parameters. Let Tn denote the maximum

likelihood estimates of θ. Our aim is to estimate the conditional mean E(θ|Tn). Now we

consider the following estimation methods.

We compute the maximum likelihood estimates of b0, b1 and b2 based on a given data

series and denote the estimates as T n = (b0, b1, b2)′.

We then propose the following nonparametric kernel estimation method:

• First, we simulate θj = (b0j , b

1j , b

2j), for j = 1, 2, · · · ,m, from the prior density π(θ). We

assume b0 ∼ Uniform(0, 1), b2 ∼ Uniform(0, 1) and b1 ∼ Uniform(0, 1− b2).

• Second, we simulate Tn,j from a limiting distribution of the maximum likelihood esti-

mator of θ, which is a normal distribution with mean θj and variance 1nΣ, where Σ can

be computed with a closed–form expression provided by Ma (2008). When n is large

enough, we have

Tn,j = (T 0n,j, T

1n,j, T

2n,j) ∼ N

(θj,

1

).

• Define g(Tn) = (g0 (T 0n) , g1 (T 1

n) , g2 (T 2n))

τ, in which g0(T 0

n) = E(b0|Tn), g1(T 1n) =

E(b1|Tn) and g2(T 2n) = E(b2|Tn). Based on (θj, Tn,j), we then estimate g(Tn) by the

nonparametric Nadaraya–Waston kernel estimate, denoted by gNW (·).

The conventional importance sampling estimate is denoted by gIM(·). To distinguish the

nonparametric kernel estimation based on MCMC samples from the conventional Bayesian

sample mean, we use MC1 to denote the latter, while MC2 to denote the former.

25

Page 27: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

MC1 method is summarised as follows.

(i) Assume π(θ|Tn) = (1 + T 2n0)I[1 < b0 <

11+T 2

n0](1 + T 2

n1)I[1 < b2 <1

1+T 2n2

]I[0 < b1 <

1− b2].

(ii) We can sample θj from π(θ|T n), for j = 1, 2, · · · ,m, and then we approximate the

conditional mean of the form gMC1(·) = 1m

∑mj=1 θj.

MC2 method is described as follows.

(i) Given Tn,j, we simulate a sample θj1, θj2, · · · , θjN from π(θj|Tn,j) and then we compute

the average denoted as θj = 1N

∑Ni=1 θji.

(ii) Based on(θj, Tn,j

), we then estimate g(Tn) by the conventional Nadaraya–Waston

method denoted as gMC2(·).We then consider two types of error distributions in Sections 7.1 and 7.2 below.

7.1 Gaussian error density

We simulated 1000 samples from the following GARCH (1,1) model:

yt = σtεt, εt ∼ N(0, 1), t = 1, 2, · · · , n

σ2t = b0 + b1y

2t−1 + b2σ

2t−1,

where θ = (b0, b1, b2)′ = (0.5, 0.15, 0.7)′.

• Case A: n = 500 and m = 5000.

• Case B: n = 1000 and m = 10000.

For each replication, we estimate θ by θ and the bias and standard deviation for Case A

and Case B are given in Tables 7.1 and 7.2, respectively.

7.2 Chi–squared error density

We simulated 1000 samples from the following GARCH (1,1) model:

yt = σtεt, εt =ut − 1√

2, ut ∼ χ2(1), t = 1, 2, · · · , n

σ2t = b0 + b1y

2t−1 + b2σ

2t−1,

where n = 1000 and θ = (b0, b1, b2)′ = (0.5, 0.15, 0.7)′.

For each replication, we estimate θ by θ and the bias and standard deviation for Case A

and Case B are given in Tables 7.3 and 7.4, respectively. For i = 0, 1, 2, define

biasbi =1

1000

1000∑r=1

(bi,r − bi

), biasθ = biasb0 + biasb1 + biasb2 ,

σij =1

1000

1000∑r=1

(bi,r − bi

)(bj,r − bj

), stdθ =

√∑i,j

σ2ij.

26

Page 28: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Tables 7.1–7.4 below give the biases and standard deviations of the maximum likelihood

(ML) estimate, the importance sampling average (IMS) estimate, the simple MC mean (MC1)

estimate, the nonparametric NW kernel (NW) estimate based on direct samples, and the

nonparametric NW kernel estimate based on MCMC samples (MC2).

Table 7.1: The biases and standard deviations for Case A with Normal error

Case A bias std

biasb0 biasb1 biasb2 biasθ stdb0 stdb1 stdb2 stdθ

MLE 0.1117 0.0043 −0.0396 0.0764 0.3540 0.0537 0.1396 0.2456

IMS 0.1174 0.0027 −0.0390 0.0811 0.1389 0.0487 0.0746 0.1010

MC1 −0.1304 0.1760 −0.3520 −0.3064 0.0886 0.0210 0.0418 0.0693

NW 0.0006 −0.0042 −0.1464 −0.1499 0.0042 0.0499 0.0766 0.0248

MC2 −0.1158 0.1632 −0.3265 −0.2791 0.0039 0.0132 0.02630 0.0167

Table 7.2: The biases and standard deviations for Case B with Normal error

Case B bias std

biasb0 biasb1 biasb2 biasθ stdb0 stdb1 stdb2 stdθ

MLE 0.0452 0.0009 −0.0151 0.0310 0.2068 0.0363 0.0850 0.1449

IMS 0.0917 0.0019 −0.0294 0.0642 0.1309 0.0344 0.0624 0.0914

MC1 −0.1143 0.1796 −0.3593 −0.2940 0.0600 0.0135 0.0267 0.0476

NW −0.0080 −0.0072 −0.1542 −0.1694 0.0112 0.0349 0.0469 0.0119

MC2 −0.1091 0.1611 −0.3221 −0.2701 0.0034 0.0083 0.0165 0.0073

Table 7.3: The biases and standard deviations for Case A with Chi–squared

error

Case A bias std

biasb0 biasb1 biasb2 biasθ stdb0 stdb1 stdb2 stdθ

MLE 0.1667 0.0360 −0.0788 0.1238 0.5112 0.1598 0.2236 0.3801

IMS 0.1435 0.0147 −0.0875 0.0707 0.1967 0.1060 0.1860 0.1492

MC1 −0.1404 0.1708 −0.3417 −0.3112 0.1135 0.0314 0.0627 0.0866

NW −0.0048 0.0232 −0.1712 −0.1528 0.0277 0.1253 0.1768 0.0169

MC2 −0.1168 0.1611 −0.3223 −0.2780 0.0120 0.0272 0.0545 0.0097

27

Page 29: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Table 7.4: The biases and standard deviations for Case B with Chi-squared

error

Case B bias std

biasb0 biasb1 biasb2 biasθ stdb0 stdb1 stdb2 stdθ

MLE 0.1128 0.0204 −0.0523 0.0808 0.3790 0.1071 0.1712 0.2731

IMS 0.1269 0.0131 −0.0589 0.0811 0.1923 0.0866 0.1301 0.1458

MC1 −0.1306 0.1742 −0.3484 −0.3048 0.0939 0.0251 0.0501 0.0716

NW −0.0045 0.0126 −0.1725 −0.1644 0.0217 0.0956 0.1290 0.0235

MC2 −0.1103 0.1595 −0.3191 −0.2690 0.0065 0.0205 0.0411 0.0108

Note that the bandwidth used in either the nonparametric NW estimate or the nonpara-

metric NW estimate based on the resample is chosen by the normal reference rule. Note

also that Tables 7.1–7.4 show that in terms of the standard deviation performance, MC2,

the nonparametric estimate based on the resample, in each case, outperforms its natural

competitors.

In the following subsection, using a set of real data, we examine the finite–sample perfor-

mance of NW and MC2 with their natural competitors of the unknown parameters involved

in the GARCH model.

7.3 Real data

We downloaded the S&P 500 daily closing prices, pt from http://finance.yahoo.com. The

date t return is calculated as yt = log(pt/pt−1). We consider the case where S&P 500 daily

returns are from the 4th of January 2007 to the 31st of May 2013 with the rolling period

starting from the 3rd of January 2011. Figure 5 below gives the plot of yt for this first period.

To evaluate whether the inclusion of the period of the main global financial crisis has any

impact on the finite–sample performance, we also consider the returns for the period of the

5th January 2009 to the 2nd September 2014 rolling from 2nd January 2013. Figure 6 below

gives the plot of yt for this second period.

Unlike in simulation, we cannot replicate the data. Instead, we propose a so–called

forward–rolling method to use the consecutive sections of the data for the evaluation of

standard deviations of the estimators proposed above. Let T denote the number of all the

observations and n denote the rolling sample size. We consider the first case where T = 1613,

the size of rolling samples, n = 1007, and the number of rolling samples, R = T − n = 606.

We also consider the second case where T = 1425, the size of rolling samples, n = 1005, and

the number of rolling samples, R = T − n = 420.

28

Page 30: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

−10

−50

510

Year

S&P 5

00 re

turn

2007 2008 2009 2010 2011 2012 2013

Figure 5: Plot of yt for the period of the 4th of January 2007 to the 31st of May 2013.

−10

−50

510

Year

S&P 5

00 re

turn

2009 2010 2011 2012 2013 2014

Figure 6: Plot of yt for the period of the 5th January 2009 to the 2nd September 2014.

For the r–th rolling sample with r = 1, 2, · · · , T − n, we compute parameter estimates

denoted as bi,r, i = 0, 1, 2 using the five methods outlined in the previous section. We then

29

Page 31: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

compute forward–rolling standard deviation (FRSD) by

FRSDbi =

√√√√ 1

T − n

T−n∑r=1

(bi,r − bi

)2

and FRSDθ =

√∑i,j

σ2ij (7.1)

for i = 0, 1, 2, where bi = 1T−n

∑T−nr=1 bi,r and σij is as defined before for i, j = 0, 1, 2.

Table 7.5 below gives the Forward–rolling standard deviations (FRSDs) for the first pe-

riod, while Table 7.6 below reports the FRSDs for the second period. For the nonparametric

kernel based estimates: NW and MC2, the usual normal reference rule was used for the

bandwidth choice.

Table 7.5: FRSDs of S&P 500 returns for the first period

Estimates FRSDb0 FRSDb1 FRSDb2 FRSDθ

MLE 0.0028 0.0069 0.0084 0.0044

IMS 0.1203 0.0046 0.0623 0.0679

MC1 0.0029 0.0026 0.0028 0.0038

NW 0.0005 0.0062 0.0103 0.0036

MC2 0.00008 0.0020 0.0040 0.0013

Table 7.6: FRSDs of S&P 500 returns for the second period

Estimates FRSDb0 FRSDb1 FRSDb2 FRSDθ

MLE 0.0029 0.0063 0.0117 0.0052

IMS 0.0168 0.0053 0.0242 0.0140

MC1 0.0029 0.0029 0.0038 0.0039

NW 0.0005 0.0059 0.0091 0.0027

MC2 0.0001 0.0018 0.0035 0.0009

While Column 4 of Figure 7.5 shows that MC1 has slightly smaller forward–rolling stan-

dard deviation than that of MC2, the last column of Table 7.5 shows that MC2 uniformly

outperforms its natural competitors in terms of the proposed forward–rolling standard devi-

ation. Similar conclusions can be made for Table 7.6. Without including the period of the

main global financial crisis, there are some improvements on the FRSDs, particularly with

the IMS method.

In summary, both the simulation and the real data evaluation show that MC2 has the

smallest standard deviation, which supports that the large–sample theory is verifiable in the

finite–sample situations.

30

Page 32: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

8 Conclusions and Discussions

This paper has proposed some closed–form estimation and approximation methods for some

infeasible computational issues. We have also developed some new asymptotic theory to

support the proposed estimation and approximation methods. The proposed estimation

and approximation theory has been evaluated by the simulated examples. Meanwhile, an

empirical example has also been included to show that the proposed nonparametric estimation

method based on resamples for the unknown parameters involved in the GARCH model has

the smallest forward–rolling standard deviation.

There are several topics that are left for future research. The first topic is that issues

regarding model overidentification will be carefully examined. The second topic is that new

estimation and approximation methods are needed to deal with the case where there are

structural breaks with the conditional distribution of X given θ. In this case, one may need

to extend the work by Andrews and Fair (1988), and then Imbens and Kalyanaraman (2012)

for our case. The third topic is how to deal with various estimation and approximation issues

for the case where there is a type of endogeneity involved in the conditional distribution

of X given θ. Estimation issues for unknown parameters and functions involved in general

structural models will also be considered.

9 Appendix

9.1 Proofs of Theorems 2.1 and 2.2

Let Amj(x) =K(xj−xh

)√∑m

k=1K2(xk−xh

) . We have

√√√√ m∑j=1

K

(xj − xh

)(gm(x)− g(x)) =

√√√√√∑mj=1K

2(xj−xh

)∑m

j=1K(xj−xh

) · m∑j=1

Amj(x)ej

+

∑mj=1K

(xj−xh

)(g(xj)− g(x))√∑m

j=1K(xj−xh

) . (10.1)

The converge of the first term of (10.1) in distribution to a Normal random variable follows

directly from Assumptions 2.1 and 2.3(i). Assumption 2.1(ii)(iii) implies that g(x) = E[θ|x] is twice

differentiable and the second–order derivative is continuous. Thus, the second term of equation

(10.1) is of an order of O(√

mhr+4)

.

31

Page 33: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

To prove Theorem 2.2, we have a look at the following decomposition:

gmmn(x)− g(x) = gmn(x)− gm(x) + gm(x)− g(x)

=

∑ni=1 L

(Xi−xb

)εmi∑n

i=1 L(Xi−xb

) +

∑ni=1 L

(Xi−xb

)(gm(Xi)− gm(x))∑n

i=1 L(Xi−xb

) + gm(x)− g(x)

≡ B1mn(x) +B2mn(x) + gm(x)− g(x), (10.2)

where B1mn(x) =

∑ni=1 L

(Xi−xb

)εmi∑n

i=1 L(Xi−xb

) and B2mn(x) =

∑ni=1 L

(Xi−xb

)(gm(Xi)−gm(x))∑n

i=1 L(Xi−xb

) .

Let us first deal with B1mn(x). Under Assumptions 2.1–2.3, the standard conditions required

for establishing the central limit theorem are satisfied. Thus, we have as n→∞√√√√∑ni=1 L

(Xi−xb

)σ2mn

√√√√√∑ni=1 L

2(Xi−xb

)∑n

i=1 L(Xi−xb

) · n∑i=1

L(Xi−xb

)√∑n

i=1 L2(Xi−xb

) εmi

→D N(0, σ2(L)

). (10.3)

Let us then deal with B2mn(x). Observe that

B2mn(x) =

∑ni=1 L

(Xi−xb

)(gm(Xi)− gm(x))∑n

i=1 L(Xi−xb

) =

∑ni=1 L

(Xi−xb

)(gm(Xi)− g(Xi))∑n

i=1 L(Xi−xb

)+

∑ni=1 L

(Xi−xb

)(g(Xi)− g(x))∑n

i=1 L(Xi−xb

) +

∑ni=1 L

(Xi−xb

)(g(x)− gm(x))∑n

i=1 L(Xi−xb

)≡ C1mn(x) + C2mn(x) + C3mn(x), (10.4)

where we have used the decomposition: gm(Xi)− gm(x) = gm(Xi)− g(Xi) + g(Xi)− g(x) + g(x)−gm(x), and C3mn(x) = g(x)− gm(x).

Meanwhile, using the following decomposition:

gm(x)− g(x) =m∑j=1

Kmj(x)θj − g(x) =m∑j=1

Kmj(x)ej +m∑j=1

Kmj(x) (g(xj)− g(x)) (10.5)

we have

C1mn(x) =

∑ni=1 L

(Xi−xb

)(gm(Xi)− g(Xi))∑n

i=1 L(Xi−xb

) =

∑ni=1 L

(Xi−xb

)(∑mj=1Kmj(Xi)ej

)∑n

i=1 L(Xi−xb

)+

∑ni=1 L

(Xi−xb

)(∑mj=1Kmj(Xi) (g(xj)− g(Xi))

)∑n

i=1 L(Xi−xb

) ≡ D1mn(x) +D2mn(x), (10.6)

where Kmj(x) =K(xj−xh

)∑mk=1K

(xk−xh

) .

32

Page 34: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Define D3mn(x) = 1mhnb

∑mj=1

(∑ni=1 L

(Xi−xb

)K(xj−Xih

))ej and let E[e2

1] = 1 for notational

simplicity. We then have

E[D2

3mn(x)]

=1

m2h2n2b2

m∑j=1

n∑i=1

K2

(xj −Xi

h

)L2

(Xi − xb

)+

2

m2h2rn2b2r

m∑j=1

n∑i=2

i−1∑k=1

K

(xj −Xi

h

)K

(xj −Xk

h

)L

(Xi − xb

)L

(Xk − x

b

)≡ 1

m2h2rn2b2r(D4mn(x) + 2D5mn(x)) . (10.7)

Simple calculation implies as (m,n)→ (∞,∞)

E [D4mn(x)] =

m∑j=1

n∑i=1

E

[K2

(xj −Xi

h

)L2

(Xi − xb

)]

= mn

∫ ∫K2

(u− vh

)L2

(v − xb

)p(v)f(u)dudv

= mnhrbr (1 + o(1)) ·∫ ∫

K2(x)L2(y)dydx · p(x) · f(x), (10.8)

where p(·) and f(·) denote the density functions of Xi and xj , respectively.

Similarly, we have as hb → 0

E [D5mn(x)] =m∑j=1

n∑i=2

i−1∑k=1

E

[K

(xj −Xi

h

)K

(xj −Xk

h

)L

(Xi − xb

)L

(Xk − x

b

)]

= mn2(1 + o(1)) ·∫ ∫ ∫

K

(u− vh

)K

(u− wh

)L

(v − xb

)L

(w − xb

)q(v)q(w)f(u)

× dudvdw = mn2h2rbr (1 + o(1)) ·∫L2(u)du · q2(x) · p(x). (10.9)

Equations (10.7)–(10.9) then imply

D3mn(x) = OP

(1√mbr

)and D1mn(x) = OP

(1√mbr

). (10.10)

We now deal with D2mn(x) involved in (10.6). By Assumption 2.1(ii), using the standard

derivation for the bias term (see Theorem 2.2 of Li and Racine 2007), we have

m∑j=1

Kmj(x) (g(xj)− g(x)) = (1 + oP (1)) Q(x) · h2,

where

Q(x) =

∫u2K(u)du (1 + o(1))

r∑j=1

(g(jj)(x) +

2g(j)(x)f (j)(x)

f(x)

),

in which f(·) is the density function of {xj}.

33

Page 35: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

We then define D6mn(x) = h2

nbr∑n

i=1 L(Xi−xb

)Q(Xi). As (m,n)→ (∞,∞), we then have

E[D2

6mn(x)]

=h4

n2b2r

(n∑i=1

E

[L2

(Xi − xb

)Q2(Xi)

]

+ 2n∑i=2

i−1∑j=1

E

[L

(Xi − xb

)L

(Xj − xb

)Q(Xi)Q(Xj)

]=

h4

n2b2r(1 + o(1))Q2(x)

(nbr

∫L2(u)du · p(x) + n(n− 1)b2rp2(x)

)= h4Q2(x)p(x) (1 + o(1))

(1

nbr

∫L2(u)du+ p(x)

), (10.11)

which implies

D6mn(x) = OP (h2) and D2mn(x) = OP (h2). (10.12)

Similarly, we have as n→∞

C2mn(x) =

∑ni=1 L

(Xi−xb

)(g(Xi)− g(x))∑n

i=1 L(Xi−xb

) = b2 (1 +OP (1)) ·Rn(x), (10.13)

where Rn(x) is a continuous function.

Therefore, under Assumptions 2.1–2.3, equations (10.2)–(10.4), (10.10) and (10.12)–(10.13) com-

plete the proof of Theorem 2.2.

9.2 Proofs of Theorems 2.3 and 2.4

The main ideas for the proofs of Theorems 2.3 and 2.4 are very similar to those for Theorems 2.1

and 2.2. In addition, the idea dealing with the multivariate case for Theorem 2.3 is also very similar

to that for Theorem 2.4. Therefore, we just give the main idea for the proof of Theorem 2.3.

Observe that√√√√ n∑j=1

K

(xj − xh

)(gm(x)− g(x)) =

√√√√√∑mj=1K

2(xj−xh

)∑m

j=1K(xj−xh

) · m∑j=1

Wmj(x)ej

+

∑mj=1K

(xj−xh

)(g(xj)− g(x))√∑n

j=1K(xj−xh

)

√√√√√∑mj=1K

2(xj−xh

)∑m

j=1K(xj−xh

) · Jm1(x) +

√√√√ n∑j=1

K

(xj − xh

)· Jm2(x), (10.14)

whereWmj(x) =K(xj−xh

)√∑m

j=1K2(xj−xh

) , Jm1(x) =∑m

j=1Wmj(x)ej and Jm2(x) =

∑mj=1K

(xj−xh

)(g(xj)−g(x))∑n

j=1K(xj−xh

) .

In order to show that Jm1(x)→D N (0,Σ), it suffices to show that AτJm1(x)→D N (0, AτΣA)

for any constant vector A = (a1, · · · , ad)τ satisfying AτA = 1. The proof of the latter follows

trivially.

34

Page 36: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Meanwhile, the bias term Jm2(x) may also be dealt with through computing the bias term

AτJm2(x), and this follows trivially. This therefore completes the proof of Theorem 2.3.

9.3 Proof of Theorem 3.1

Observe that

√√√√mn∑i=1

L

(Xi − xh

)(gmn(x)− g(x)) =

√∑ni=1 L

2(Xi−xh

)√∑n

i=1 L(Xi−xh

∑ni=1 L

(Xi−xh

)emi√∑n

i=1 L2(Xi−xh

) +

√m∑n

i=1 L(Xi−xh

)(g(Xi)− g(x))√∑n

i=1 L(Xi−xh

) , (10.15)

where emi = 1√m

∑mj=1 eij .

By Peligrad (1987), we have for some 0 < C <∞

E[e4mi

]≤ 1

m2

m2 E[e4ij ] +

E m∑j=1

eij

22 ≤ C <∞, (10.16)

which, along with the conditions of Theorem 3.1 of this paper and Theorem 2.22 of Fan and Yao

(2003), implies

1√nhrσm(x)

n∑i=1

L

(Xi − xh

)emi →D N

(0,Σ2(x)

), (10.17)

where σ2m(x) = 1

m

∑mj=1E

[e2ij |Xi = x

]+ 2

m

∑mj1=2

∑j1−1j2=1E [eij1eij2 |Xi = x]→ γ(x) + 2

∑∞j=1 γj(x),

and Σ2(x) = p(x) ·∫K2(u)du, in which p(x) is the density of X1.

The second term of equation (10.15) can be dealt with as before. Therefore, we have completed

the proof of Theorem 3.1.

9.4 Proof of Theorem 3.2

The proof of the first part follows in the same way from equation (10.15). In a similar fashion to

equation (10.16), by Assumption 6.4(ii), we have almost surely,

E[e2mi|Fm,i−1

]=

1

m

m∑j=1

E[ε2ij |Fm,i−1

]+

2

m

m∑j1=2

j1−1∑j2=1

E [εi,j1εi,j2 |Fm,i−1]

=1

m

m∑j=1

E[ε2ij |Fm,i−1

]→ σ2

ε , (10.18)

E[e4mi|Fm,i−1

]≤ 1

m2

m2 E[e4ij |Fm,i−1] +

E m∑

j=1

eij

2

|Fm,i−1

2 ≤ C <∞,

35

Page 37: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

which, along with an application of Theorem 3.2 of Gao and Phillips (2013), imply√√√√√∑ni=1 L

2(Xi−xh

)∑n

i=1 L(Xi−xh

) · ∑ni=1 L

(Xi−xh

)emi√∑n

i=1 L2(Xi−xh

) →D N(0,Σ2

1

), (10.19)

where Σ21 is the same as defined in Theorem 3.2. The second term can be dealt with in a similar

way to what has been done before. This completes the proof of Theorem 3.2.

9.5 Proof of Theorem 4.1

Let enj = θj − E[θj |Tnj ]. Observe that

√√√√ n∑j=1

K

(Tnj − Tn

h

)(gkm(Tn)− g(Tn)) =

√√√√√∑mj=1K

2(Tnj−Tn

h

)∑m

j=1K(Tnj−Tn

h

) · m∑j=1

Pmj(Tn)enj

+

∑mj=1K

(Tnj−Tn

h

)(g(Tnj)− g(Tn))√∑n

j=1K(Tnj−Tn

h

)

√√√√√∑mj=1K

2(Tnj−Tn

h

)∑m

j=1K(Tnj−Tn

h

) ·Qm1(Tn) +

√√√√ n∑j=1

K

(Tnj − Tn

h

)·Qm2(Tn), (10.20)

where Pmj(Tn) =K(Tnj−Tn

h

)√∑m

j=1K2(Tnj−Tn

h

) , Qm1(Tn) =∑m

j=1 Pmj(Tn)enj and

Qm2(Tn) =

∑mj=1K

(Tnj−Tn

h

)(g(Tnj)− g(Tn))∑n

j=1K(Tnj−Tn

h

) . (10.21)

Let fn(·) be the density of Tnj . Simple calculation implies for i = 1, 2,∑m

j=1E[Ki(Tnj−Tn

h

)]=

(1+o(1)) mh∫Ki(v)dv ·fn(Tn) when Tn is given. Such results, along with the law of large numbers,

imply that 1mhfn(Tn)

∑mj=1K

i(Tnj−Tn

h

)= (1 + oP (1))

∫Ki(v)dv for i = 1, 2.

Thus, we have as n→∞

E

1

mh

m∑j=1

K

(Tnj − Tn

h

)(g(Tnj)− g(Tn))

= (1 + o(1))h2C(Tn) ·∫u2K(u)du, (10.22)

where C(Tn) = 12p(Tn)g(2)(Tn) + g(1)(Tn)p(1)(Tn).

Let D(Tn) = p(Tn)(g(1)(Tn)

)2. Similarly, we have as n→∞

Var

1√mh

m∑j=1

K

(Tnj − Tn

h

)(g(Tnj)− g(Tn))

=1

mh

m∑j=1

Var

(K

(Tnj − Tn

h

)(g(Tnj)− g(Tn))

)

≤ 1

mh

m∑j=1

E

[K2

(Tnj − Tn

h

)(g(Tnj)− g(Tn))2

]= (1 + o(1)) h2 D(Tn) ·

∫u2K2(u)du,

36

Page 38: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

which, along with equation (10.22), implies as n→∞√mh

λnQm2(Tn) =

(OP

(h

λn

)+OP

(√mh5

λn

))· γ(Tn) = OP (1) (10.23)

by Assumption 4.2(ii), where γ(Tn) = 12g

(2)(Tn) + p(1)(Tn)p(Tn) g

(1)(Tn) + g(1)(Tn).

Using the result that E[e21|Tn1] = Var[θ|Tn] = λ2

n as well as Assumptions 4.1 and 4.2, the

standard conditions required for establishing the central limit theorem are satisfied. The proof of

Theorem 4.1 is therefore completed.

9.6 Proof of Theorem 4.2

Recall that p(Tn) =∫f(Tn|θ)π(θ)dθ, q(Tn) =

∫θf(Tn|θ)π(θ)dθ and p(Tn) = 1

m

∑mi=1 f(Tn|θi). By

the definition of gam(Tn), we have

√m p(Tn) (gam(Tn)− g(Tn)) =

1√m

m∑i=1

(θif(Tn|θi)−

q(Tn)

p(Tn)f(Tn|θi)

)

=

m∑i=1

1√m·(θif(Tn|θi)−

q(Tn)

p(Tn)f(Tn|θi)

)≡

m∑i=1

Ymi, (10.24)

where Ymi = 1√m·(θif(Tn|θi)− q(Tn)

p(Tn)f(Tn|θi))

.

By Assumption 4.3, the standard conditions required for establishing the central limit theorem

of a sum of i.i.d. random variables are satisfied. Therefore, we have for fixed n and as m→∞√m p2(Tn)

σ2(Tn)· p(Tn)

p(Tn)(gam(Tn)− g(Tn))

=1√

m σ2(Tn)

m∑i=1

(θif(Tn|θi)−

q(Tn)

p(Tn)f(Tn|θi)

)→D N(0, 1), (10.25)

which completes the proof of Theorem 4.2.

9.7 Proof of Theorem 5.1

9.7.1 Lemmas

In order to prove Theorem 5.1, we need to introduce the following lemmas.

Lemma 9.1: Suppose that Assumption 5.1 holds. Let supγ∈Γn |Lmn(γ)− Ln(γ)| = oP

(ν− 1

2n

)for

m = m(n)→∞. Then, we have√νn l

12n (γ0) (γmn − γn) = oP (1) and

√νn l

12n (γ0) (γmn − γ0)→D ξ,

where ξ is as defined in Theorem 5.1.

The proof of Lemma 9.1 follows from that of Theorem A.5 of Kristensen and Shin (2012).

Lemma 9.2: Let Assumptions 5.2 and 5.3 hold. Then as (m,n)→ (∞,∞)

supx∈Rr

supγ∈Γn

|pm(x; γ)− E [pm(x; γ)]| = OP (√sm) . (10.26)

37

Page 39: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

where Γ ⊂ Rr is the parameter space of γ, and sm = log(m)m hr .

The proof of Lemma 9.2 follows immediately from Theorem 1(ii) of Kristensen (2009) because

the setting here is all i.i.d.

9.7.2 Proof of Theorem 5.1

The main idea here is to apply Lemma 9.2 to verify supγ∈Γn |Lmn(γ)− Ln(γ)| = oP

(ν− 1

2n

)for

m = m(n)→∞.

Observe that

Lmn(γ)− Ln(γ) =1

νn

n∑i=1

((log(pm(Xi; γ))− (log(p(Xi; γ)))

=1

νn

n∑i=1

log

(1 +

pm(Xi; γ)− p(Xi; γ)

p(Xi; γ)

)

=1

νn

n∑i=1

1

p(Xi; γ)· (pm(Xi; γ)− p(Xi; γ)) +

1

νn

n∑i=1

∆mn(Xi; γ)

≡ Rmn1(γ) +Rmn2(γ), (10.27)

where ∆mn(x; γ) is a function of terms with orders higher than 1p(Xi;γ) · (pm(Xi; γ)− p(Xi; γ)), and

such higher order terms are negligible in the evaluation of the order of Lmn(γ)− Ln(γ).

Using Lemma 9.2, we have

|Rmn1(γ)| = 1

νn

∣∣∣∣∣n∑i=1

1

p(Xi; γ)· (pm(Xi; γ)− p(Xi; γ))

∣∣∣∣∣ (10.28)

≤ 1

νn

n∑i=1

1

p(Xi; γ)· |pm(Xi; γ)− E1 [p(Xi; γ)]|+ 1

νn

n∑i=1

1

p(Xi; γ)· |E1 [pm(Xi; γ)]− p(Xi; γ)|

≤ 1

νn

n∑i=1

supx∈Rr

supγ∈Γn

(1

p(x; γ)· |pm(x; γ)− E [p(x; γ)]|

)

+1

νn

n∑i=1

supx∈Rr

supγ∈Γn

(1

p(x; γ)· |E [pm(x; γ)]− p(x; γ)|

)≤ C n

νn

(√log(m)

mhr+ h2

)= oP (

√νn)

using the standard result that |E[pm(x, γ)]− p(x; γ)| ≤ C(x, γ)h2 as well as Assumptions 5.1(ii) and

5.3(iv) in particular, where E1[U ] denotes the conditional expectation of U given Xi, and C(x, γ) is

a function involving p2(x; γ) that is the second–order partial derivative of p(x; γ) with respect to x.

This shows that supγ∈Γn |Lmn(γ)− Ln(γ)| = oP

(ν− 1

2n

)required in Lemma 9.1 is satisfied.

Therefore, the proof of Theorem 5.1 then follows from Lemma 9.1.

Remark 9.1: As mentioned in Section 5, one may replace Lmn(γ) by a truncated version of the

form

Lmnc(γ) =1

νn

n∑i=1

wmn(Xi) log(pm(Xi; γ)), (10.29)

where wmn(Xi) = 1 if pm(Xi; γ) > cmn and wmn(Xi) = 0 if pm(Xi; γ) < cmn2 , where cmn > 0 and

cmn → 0 as (m,n)→ (∞,∞).

38

Page 40: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Let fm(x; γ) = p−1(x; γ). We then have

Lmnc(γ)− Lmn(δ) = − 1

νn

n∑i=1

I[cmn

2≤ pm(Xi; γ) ≤ cmn

]log(pm(Xi; γ))

=1

νn

n∑i=1

I [log(cmn) ≤ log(fm(Xi; γ)) ≤ log(2cmn)] log(fm(Xi; γ)), (10.30)

which implies

√νn E

[supγ∈Γn

|Lmnc(γ)− Lmn(γ)|

]

≤ 1√νn

n∑i=1

E

[supγ∈Γn

(|I [log(cmn) ≤ log(fm(Xi; γ)) ≤ log(2cmn)]| · |log(fm(Xi; γ))|)

]

≤ |log(cmn)|−δ · n√νnE

[supγ∈Γn

(|log(fm(X1; γ))|1+δ

)]= o(1) (10.31)

when n√νn |log(cmn)|δ

= o(1) and∫

supγ∈Γn

(|log(p(x; γ))|1+δ p(x; γ)

)dx < ∞ for some δ > 0. Note

that it is possible to choose cmn such that n√νn |log(cmn)|δ

= o(1) for a suitable δ > 0. When

|log(cmn)| = (mn)c for some c > 0 and νn = n, for example, the condition is satisfied if√n

(mn)cδ→ 0.

REFERENCES

Amemiya, T. (1985), Advanced Econometrics (Cambridge: Harvard University Press).

Andrews, D. W. K. and Fair, R. C. (1988), “Inference in Nonlinear Econometric Models with Structural

Change”, Review of Economic Studies, LV, 615–640.

Beaumont, M., Zhang, W. and Balding, D. (2002), “Approximate Bayesian Computation in Popula-

tion Genetics”, Genetics, 162, 2025–2035.

Blum, M. G. B. (2010), “Approximate Bayesian Computation: a Nonparametric Perspective”, Journal of

the American Statistical Association, 105, 1178–1187.

Blum, M. G. B. (2013), “A Comparative Review of Dimension Reduction Methods in Approximate

Bayesian Computation”, Statistical Science, 28, 189–208.

Brooks, S., Gelman, A., Jones, G. and Meng, X. (2011), Handbook of Markov Chain Monte Carlo

(London:. Chapman & Hall/CRC).

Carlin, B. and Louis, T. (1996), Bayes and Empirical Bayes Methods for Data Analysis (London:

Chapman and Hall/CRC).

Chen, X. (2007), “Large Sample Sieve Estimation of Semi–Nonparametric Models”, Handbook of Econo-

metrics, 6, 5549–5632.

Chernozhkov, V. and Hong, H. (2003), “An MCMC Approach to Classical Estimation”, Journal of

Econometrics, 115, 293–346.

Chib, S. and Greenberg, E. (1995), “Understanding the Metropolis–Hastings Algorithm”, Journal of

the American Statistical Association, 49, 327–335.

39

Page 41: Nonparametric Regression Approach to Bayesian Estimation · 2017-06-12 · Nonparametric Regression Approach to Bayesian Estimation 1 Jiti Gao and Han Hong Monash University and Stanford

Creel, M. and Kristensen, D. (2011), “Indirect Likelihood Inference”, Working paper is available at

the website: http://econpapers.repec.org/paper/aubautbar/931.13.htm.

Efron, B. (1996), “Empirical Bayes Methods for Combining Likelihoods (with discussion)”, Journal of the

American Statistical Association, 91, 538–565.

Fan, J. and Yao, Q. (2003), Nonlinear Time Series: Non– and Parametric Methods (New York: Springer).

Fearnhead, P. and Prangle, D. (2012), “Constructing Summary Statistics for Approximate Bayesian

Computation: Semi-automatic Approximate Bayesian Computation (with discussion)”, Journal of the

Royal Statistical Society Series B, 74, 419–474.

Gallant, R. and Tauchen, G. (1996), “Which Moments to Match ?”, Econometric Theory, 12, 363–390.

Gao, J. (2007), Nonlinear Time Series: Semi– and Non–Parametric Methods (London: Chapman &

Hall/CRC).

Gao, J. and Hong H. (2014), “A Computational Implementation of GMM”, Working paper available at

http://ssrn.com/abstract=2503199.

Gao, J. and Phillips, P. C. B. (2013), “Semiparametric Estimation in Triangular System Equations

with Nonstationrity”, Journal of Econometrics, 176, 5979.

Gelfand, A. and Smith, A. F. M. (1990), “Sampling–based Approaches to Calculating Marginal Den-

sities”, Journal of the American Statistical Association, 85, 398–409.

Geweke, J. (1989), “Bayesian Inference in Econometric Models using Monte Carlo Integration”, Econo-

metrica, 57, 1317–1339.

Geweke, J. (2005), Contemporary Bayesian Econometrics and Statistics (New jersey: John Wiley & Sons).

Hansen, L. P. (1982), “Large Sample Properties of Generalized Method of Moments Estimators”, Econo-

metrica, 50, 1029–1054.

Imbens, G. and Kalyanaraman, K. (2012), “Optimal Bandwidth Choice for the Regression Discontinuity

Estimator”, Review of Economic Studies, 79, 933–959.

Kristensen, D. (2009), “Uniform Convergence Rates of Kernel Estimators with Heterogeneous Dependent

Data”, Econometric Theory, 25, 1433–1445.

Kristensen, D. and Shin, Y. (2012), “Estimation of Dynamic Models with Nonparametric Simulated

Maximum Likelihood”, Journal of Econometrics, 167, 76–94.

Li, Q. and Racine, J. (2007), Nonparametric Econometrics: Theory and Practice (New Jersey: Princeton

University Press).

Liu, J. (2001), Monte Carlo Strategies in Scientific Computing (New York: Springer).

Ma, J. (2008), “A Closed–form Asymptotic Variance–Covariance Matrix for the Quasi–Maximum Likeli-

hood Estimator of the GARCH(1,1) Model”

(working paper available at http://papers.ssrn.com/sol3/papers.cfm?abstract-id=889461).

Peligrad, M. (1987), “On Central Limit Theorem for ρ–Mixing Sequences”, Annals of Probability, 15,

1387–1394.

Phillips, P. C. B. and Hansen, B. (1990), “Statistical Inference in Instrumental Variables Regression

with I(1) Processes”, Review of Economic Studies, 57, 99-125.

Tierney, L. (1994), “Markov Chains for Exploring Posterior Distributions”, Annals of Statistics, 22,

1701–1762.

40