Hostname: page-component-745bb68f8f-f46jp Total loading time: 0 Render date: 2025-02-11T05:40:47.746Z Has data issue: false hasContentIssue false

COINTEGRATING SMOOTH TRANSITION REGRESSIONS

Published online by Cambridge University Press:  10 February 2004

Pentti Saikkonen
Affiliation:
University of Helsinki
In Choi
Affiliation:
Hong Kong University of Science and Technology and Cowles Foundation, Yale University
Rights & Permissions [Opens in a new window]

Abstract

This paper studies the smooth transition regression model where regressors are I(1) and errors are I(0). The regressors and errors are assumed to be dependent both serially and contemporaneously. Using the triangular array asymptotics, the nonlinear least squares estimator is shown to be consistent, and its asymptotic distribution is derived. It is found that the asymptotic distribution involves a bias under the regressor-error dependence, which implies that the nonlinear least squares estimator is inefficient and unsuitable for use in hypothesis testing. Thus, this paper proposes a Gauss–Newton type estimator that uses the nonlinear least squares estimator as an initial estimator and is based on regressions augmented by leads and lags. Using leads and lags enables the Gauss–Newton estimator to eliminate the bias and have a mixture normal distribution in the limit, which makes it more efficient than the nonlinear least squares estimator and suitable for use in hypothesis testing. Simulation results indicate that the results obtained from the triangular array asymptotics provide reasonable approximations for the finite-sample properties of the estimators and t-tests when sample sizes are moderately large. The cointegrating smooth transition regression model is applied to the Korean and Indonesian data from the Asian currency crisis of 1997. The estimation results partially support the interest Laffer curve hypothesis. But overall the effects of interest rate on spot exchange rate are shown to be quite negligible in both nations.This paper was partly written while the first author was visiting the Institute of Statistics and Econometrics at Humboldt University, Berlin. This author acknowledges financial support from the Alexander von Humboldt Foundation under a Humboldt Research Award and from the Yrjö Jahnsson Foundation. The second author wrote this paper while visiting the Cowles Foundation for Research in Economics, Yale University. This author thanks the faculty and staff of the Cowles Foundation, especially Don Andrews, John Geanakoplos, David Pearce, Peter Phillips, and Nora Wiedenbach, for their support and hospitality. The second author was financially supported for the research in this paper by Kookmin University. The authors thank Don Andrews, Helmut Lütkepohl, Peter Phillips, Bruce Hansen, and two referees for their valuable comments on this paper. Part of the data studied in this paper was provided by Chi-Young Song, whom we thank.

Type
Research Article
Copyright
© 2004 Cambridge University Press

1. INTRODUCTION

It is often perceived that economic agents may show different behavior depending on which regions some economic variables belong in, though it seems hard to find explicit economic theory supporting such behavior. For example, investors and households may make different decisions regarding their investments and savings, respectively, when interest rates are rising rapidly than when they are stable. Another possible example is that employees under recession may behave differently than under boom. Econometricians and statisticians have developed several methods to study such behavior empirically, which include, among others, switching regression (cf. Goldfeld and Quandt, 1973), threshold autoregression (cf. Tong, 1983), and smooth transition regression (cf. Granger and Teräsvirta, 1993; Teräsvirta, 1998).

In this paper, we focus on the smooth transition regression (STR) model. As argued in Granger and Teräsvirta (1993), the STR model is useful in explaining the aggregate-level economy because the economy is likely to show smooth transition if each economic agent switches sharply at different times. Asymptotic theory for the STR model involving only stationary variables can be inferred from standard theory in nonlinear econometrics (e.g., Newey and McFadden, 1994; Pötscher and Prucha, 1997). The reader is also referred to Franses and van Dijk (2000) and van Dijk, Teräsvirta, and Franses (2002) for detailed discussions on the STR model.

However, general asymptotic theory for the STR model with I(1) variables has not been developed yet. Recent work by Park and Phillips (1999, 2001) provides methods for studying nonstationary and nonlinear time series, but it seems difficult to apply these methods in our context. One of the reasons for this is that we adopt a more general assumption than martingale difference errors as in their work. In addition, Chang and Park (1998) study the STR model by using the methods of Park and Phillips (1999, 2001). However, their model adopts martingale difference errors, and it seems that their conventional asymptotics does not identify threshold parameters (ci's in Section 2 of the current paper).

Therefore, this paper studies asymptotic theory of the nonlinear least squares (NLLS) estimator for the STR model with I(1) regressors and I(0) errors. This model will be called the cointegrating STR model in this paper. As in most cointegration models, the regressors and errors are assumed to be dependent both serially and contemporaneously. Because using the usual asymptotic scheme of sending sample sizes to infinity seems to be difficult in the case of the cointegrating STR model, we will use the triangular array asymptotics. The triangular array asymptotics has been used, among others, in Andrews and McDermott (1995) for nonlinear econometric models with deterministically trending variables.

The asymptotic distribution of the NLLS estimator for the cointegrating STR model involves a bias under the regressor-error dependence, which implies that the NLLS estimator is inefficient and unsuitable for use in hypothesis testing. Therefore, we propose a Gauss–Newton type estimator that uses the NLLS estimator as an initial estimator and is based on nonlinear regressions augmented by leads and lags. Linear cointegrating regressions augmented by leads and lags are studied in Saikkonen (1991), Phillips and Loretan (1991), and Stock and Watson (1993). The Gauss–Newton estimator eliminates the bias and has a mixture normal distribution in the limit, which implies that it is efficient and that standard hypothesis tests can be performed by using the estimator.

Because the triangular array asymptotic methods have not often been used in econometrics, one may rightfully question the finite-sample properties of the tests and estimators using the methods. Therefore, we report some simulation results, which indicate that the results obtained from the triangular array asymptotics provide reasonable approximations for the finite-sample properties of the estimators and tests when sample sizes are moderately large.

The cointegrating STR regression model is applied to the Korean and Indonesian data from the Asian currency crisis of 1997. The estimation results partially support the interest Laffer curve hypothesis, which states that higher interest rates may depreciate a currency when interest rates are too high because excessively high interest rates may increase the default risk by increasing the borrowing cost of corporations, by depressing the economy and by weakening the banking system of an economy (cf. Goldfajn and Baig, 1998). But overall the effects of interest rate on spot rate are shown to be quite negligible in both nations. Considering the ineffectiveness of high interest rates in stabilizing exchange rates and the high economic cost associated with keeping high interest rates, the appropriateness of tight monetary policy during the Asian currency crisis should come into question.

The STR model has been used for some economic applications. The applications are Teräsvirta and Anderson (1992) for modeling business cycle asymmetries; Granger, Teräsvirta, and Anderson (1993) for forecasting gross national product; Sarno (1999) and Lütkepohl, Teräsvirta, and Wolters (1999) for money demand functions; Michael, Nobay, and Peel (1997) and Taylor, Peel, and Sarno (2001) for real exchange rates; and Jansen and Teräsvirta (1996) for consumption. Besides these, Luukkonen, Saikkonen, and Teräsvirta (1988) consider testing linearity against the smooth transition autoregression model.

The rest of the paper is organized as follows. Section 2 introduces the model and basic assumptions. Section 3 studies asymptotic properties of the NLLS and Gauss–Newton efficient estimators. Section 4 reports some simulation results. Section 5 applies the STR model to the data from the Asian currency crisis. Section 6 contains further remarks. The Appendixes include auxiliary results and the proofs of theorems.

A few words on our notation: all limits are taken as T → ∞. Weak convergence is denoted as ⇒. For symmetric matrices the inequality A > B (AB) means that the difference AB is positive definite (semidefinite). For an arbitrary matrix A, ∥A∥ = [tr(AA)]1/2.

2. THE MODEL AND ASSUMPTIONS

Consider the cointegrating STR

where xjt is the jth component of the I(1) vector xt (p × 1), ut is a zero-mean stationary error term, θ = [c1′…cl′ γ′]′, and g(xt;θ) is a smooth real-valued transition function of the process xt and the parameter vector θ.1

Although model (1) assumes that all the regressors have a nonlinear effect on the regressand, our theoretical results can readily be modified to the case where the nonlinearity only appears in some of the regressors. In addition, our setup does not allow for the possibility that different transition functions are used for different regressors. But it would not be difficult to extend our results to that case also. To simplify exposition, we have preferred to work with a single transition function.

Moreover, μ, ν, αj, and δj are scalar parameters.

The STR model (1) has been used to describe economic relations that change smoothly depending on the location of some economic variables. In model (1), the relationship between xt and yt may change depending on where xt is located relative to parameters c1,…,cl. Parameter γ in model (1) determines the smoothness of transition in the economic relations. The reader is referred to Granger and Teräsvirta (1993) for more discussions on the STR model, although these authors do not explicitly consider the case of I(1) processes.

We discuss some examples of model (1) by using the following simplified version of model (1) where nonlinearity appears only in the first regressor:

Example 1:

Here the transition function is a logistic function that makes the regression coefficient for x1t vary smoothly between α1 and α1 + δ1. When the value of the regressor x1t is sufficiently far below the value of the parameter c the regression coefficient takes a value close to α1, and when the value of the regressor x1t increases and exceeds the value of the parameter c the value of the regression coefficient changes and approaches α1 + δ1.

Example 2:

This transition function can be used when one wants to allow for the possibility that the regression coefficient changes twice. When |x1t| is large, the function takes a value close to 1 so that the coefficient for x1t approaches α1 + δ1. But when x1t is approximately between c1 and c2, the function takes a value close to zero, which makes the coefficient for x1t approach α1. Instead of function (4), one may also use a linear combination of two logistic functions.

When γ → ∞, functions (3) and (4) approach the indicator functions 1{x1tc} and 1{c1x1tc2}, respectively, and model (2) becomes close to a threshold regression model. Then the change in the regression coefficient of x1t is abrupt and not gradual as assumed in (2). Our results do not apply to threshold models because the transition function is not allowed to be discontinuous. Otherwise our treatment is fairly general and applies to any sufficiently well-behaved transition function.

We shall now discuss assumptions required for model (1). As already mentioned, we make the following assumption.

Assumption 1.

where vt is a zero-mean stationary process and the initial value x0 may be any random variable with the property Ex04 < ∞.

Moreover, it will be convenient to assume that the (p + 1)-dimensional process wt = [ut vt′]′ satisfies the following assumption employed by Hansen (1992) in a somewhat weaker form.

Assumption 2. For some r > 4, wt = [ut vt′]′ is a stationary, zero-mean, strong mixing sequence with mixing coefficients of size −4r/(r − 4) and Ewtr < ∞.

Assumption 2 is fairly general and covers a variety of weakly dependent processes. It also permits the cointegrated system defined by (1) and (5) to have nonlinear short-run dynamics, which is convenient because our cointegrating regression is nonlinear.

Choosing the real number p in Corollary 14.3 of Davidson (1994) as 2r/(r + 2), we find that Assumption 2 implies that the serial covariances of the process wt at lag | j| are of size −2. Thus, we have the summability condition

This implies that the process wt has a continuous spectral density matrix fww(λ) that we assume satisfies the following assumption.

Assumption 3. The spectral density matrix fww(λ) is bounded away from zero or that

Assumption 3 specialized to the case λ = 0 implies that the components of the I(1) process xt are not cointegrated. In addition, it is required for the estimation theory of Section 2 that (7) also holds for other values of λ. Conformably to the partition of the process wt, we write fww(λ) = [fab(λ)] where a,b ∈ {u,v}.

Assumption 2 also implies the multivariate invariance principle

where B(s) is a Brownian motion with covariance matrix Ω = 2πfww(0) (see Hansen, 1992, proof of Theorem 3.1). We partition B(s) = [Bu(s) Bv(s)′]′ and

conformably with the partition of the process wt.

As for the transition function g(x;θ), we make the following assumption.

Assumption 4.

(i) The parameter space Θ of θ is a compact subset of an Euclidean space.

(ii) g(x;θ) is three times continuously differentiable on

where Θ* is an open set containing Θ.

This assumption may not be the weakest possible, but it is satisfied by the most commonly used transition functions and simplifies exposition. Thus, we shall not try to weaken it. The compactness of the parameter space Θ is a standard assumption in nonlinear regression, but no such assumption is needed for other parameters.

3. ESTIMATION PROCEDURES

The cointegrating regression (1) assumes serial and contemporaneous correlation between the I(1) regressor xt and the error term ut. Adverse consequences of this on linear least squares estimation are well known, and various modifications have therefore been devised. In this paper, we extend the leads and lags procedure of Saikkonen (1991) to the STR model discussed in the previous section. Because there are some theoretical difficulties with a direct extension of this procedure, we will first consider the NLLS estimation that can be utilized to develop a Gauss–Newton type leads and lags estimation.

3.1. Triangular Array Asymptotics

Before embarking on the subject of NLLS estimation, we will explain the motivation for the employed asymptotic methods in this section. There are two types of asymptotics that can be considered in nonlinear regressions with I(1) regressors. One is the usual asymptotics, and the other is the so-called triangular array asymptotics in which the actual sample size is fixed at T0, say, and the model is imbedded in a sequence of models depending on a sample size T that tends to infinity. The imbedding is obtained by replacing the I(1) regressor by (T0 /T)1/2xt. This makes the regressand dependent on T and, when T = T0, the original model is obtained. Thus, if T0 is large, the triangular array asymptotics can be expected to give reasonable approximations for finite-sample distributions of estimators and test statistics. The triangular array asymptotics is also used in Andrews and McDermott (1995) for nonlinear econometric models with deterministically trending variables. Related references can also be found in Andrews and McDermott (1995).

We will use the triangular array asymptotics for our cointegrating model, because we expect it to provide quite reasonable approximations for estimators and test statistics and because some parameters cannot be identified when the usual asymptotics is used. The identification issue can be explained intuitively by using a special case of model (1)—the model in Example 1. When the model in Example 1 is applied, a typical situation is that the observations can be divided into three groups with each group containing a reasonably large proportion of the data. In the first and third group the values of the regression coefficient for x1t are essentially α1 and α1 + δ1, respectively, whereas the second group contains part of the sample where the value of the regression coefficient changes between these two values. Because x1t is an I(1) process, the use of conventional asymptotics means that the variation of x1t increases so that the proportion of observations in the first and third groups increases and that in the second group decreases. Eventually the proportion of observations in the second group becomes negligible. This suggests that these parameters are unidentifiable in the limit, because only observations in the second group provide information about the parameters γ and c. This can also be seen by noting that, for T large,

Thus, asymptotically the parameters γ and c vanish from the model and become unidentifiable. This discussion implies that the use of conventional asymptotics leads to a situation that is very different from what happens in the sample where a reasonably large proportion of observations belongs to the second group.

However, the triangular array asymptotics takes the second group and, therefore, the parameters γ and c into account. Recall that g(·;θ) is the logistic function. Basing the asymptotic analysis on g(γ((T0 /T)1/2x1tc)) = g(T−1/2γ(T01/2x1tT1/2c)) instead of g(γ(x1tc)) means that the slope of the logistic function is assumed to decrease so that the proportion of observations in the three groups remains essentially the same even though the variation of x1t increases. In this respect the situation for the triangular array asymptotics remains the same as for the sample. It also makes sense that parameter c has to be of order O(T1/2), because, due to the increasing variation of x1t, a nonzero value of c could otherwise be indistinguishable from zero. Finally, note that when g(γ((T0 /T)1/2x1tc)) is used in asymptotic analysis the process x1t is standardized in such a way that it remains bounded in probability. In this context a possible interpretation is that when T tends to infinity observations of the standardized version of the series x1t are obtained denser and denser within its observed range in the sample, and thereby the proportion of observations in each of the three groups remains essentially the same, which makes information about parameters γ and c retained even asymptotically.

Although the preceding discussion gives a reasonable motivation for using the triangular array asymptotics, it would be imprudent to claim that the triangular array asymptotics would always work well. For instance, we already noted that problems may occur if the value of the parameter γ in model (2) with specification (3) is large so that the model is close to a threshold model.

3.2. NLLS Estimation

This section considers the triangular array asymptotics of the NLLS estimator for model (1). To use the triangular array asymptotics, we imbed model (1) in a sequence of models

where xtT = (T0 /T)1/2xt, f (xtT;θ) = [1 g(xtT;θ) xtTg(xtT;θ)xtT′]′ and φ = [μ ν α′ δ′]′ with α = [α1…αp]′ and δ = [δ1…δp]′.2

In practice we always choose T = T0, so that the transformation xtT is not required. The transformation is made only to facilitate the development of asymptotic analysis.

In what follows we set ϑ = [θ′ φ′]′ and let ϑ0 = [θ0′ φ0′]′ stand for the true value of ϑ. The NLLS estimator of parameter ϑ0 is obtained by minimizing the function

with respect to ϑ.

The assumptions made so far do not ensure that a minimum of function (10) exists, even asymptotically. To be able to introduce further assumptions, we first use the multivariate invariance principle (8) to conclude that

. This fact and a standard application of the continuous mapping theorem show that, for every θ ∈ Θ,

An assumption that together with our previous assumptions ensures that the function QT(ϑ) has a minimum for T large enough is as follows.

Assumption 5. For some ε > 0,

where λmin(·) signifies the smallest eigenvalue of a square matrix.

Assumption 5 guarantees that, with probability approaching one, a minimum of the function QT(ϑ) exists as shown in Appendix B.

3

See Lemma 5 in Appendix A and the proof of Theorem 1 in Appendix B.

Because we are interested in asymptotic results, we may as usual assume that a minimum exists for all values of T and is attained at

.

In addition to Assumption 5, the following assumption is needed for the consistency of the least squares estimator

.

Assumption 6. For some s ∈ [0,1] and all (θ,φ) ≠ (θ00),

This is an identification condition that ensures that the parameters θ and φ can be separated in the product f (xtT;θ)′φ. Taken together, Assumptions 5 and 6 ensure the identifiability of the parameter vector ϑ.

The identification conditions (11) and (12) depend on the sample paths of the Brownian motion Bv0(s) and are therefore different from the identification conditions used by Chang and Park (1998), Chang, Park, and Phillips (1999), and Park and Phillips (2001). However, conditions (11) and (12) are still fairly easy to use. For instance, it can be checked by the conditions that model (2) with specification (3) is identified when δ1 ≠ 0 and γ > 0.

It may also be argued that it makes sense to use identification conditions that depend on the sample paths of the Brownian motion Bv0(s) when the triangular array asymptotics is used. Indeed, in applications of model (2) with specification (3), one can typically divide the observations into three groups in such a way that a fair amount of observations belongs to each group and, when the triangular array asymptotics is used, this state of affairs prevails even asymptotically. Thus, because xtTBv0(s), the triangular array asymptotics in a sense conditions on such sample paths of Bv0(s) for which the shape of the function g(γ(Bv0(s) − c)) is similar to what is observed in the sample. Because of this “conditioning,” it seems quite reasonable to use identification conditions that depend on the sample paths of the Brownian motion Bv0(s) and ensure identifiability when the specified nonlinearity is related to the sample paths of Bv0(s) in the same way as to the observed realizations of xtT within the sample. This means, for instance, that in the case of model (2) with specification (3) we are not interested in identification in cases where sample paths of Bv0(s) are such that the function g(γ(Bv0(s) − c)) is effectively constant and identifiability is very weak although it is still achieved when δ1 ≠ 0 and γ > 0. This point could be made even stronger by replacing the logistic function by a piecewise continuous analog so that for some realizations of Bv0(s) the function g(γ(Bv0(s) − c)) would actually be constant and identifiability would fail. Clearly, such cases would be of no interest if g(γ(xtTc)) is highly nonlinear within the sample.

The following theorem shows the existence and consistency of the least squares estimator

.

THEOREM 1. Suppose that Assumptions 1–6 hold. Then, a NLLS estimator

exists with probability approaching one and is consistent.

Theorem 1 shows the existence and consistency of the least squares estimator

when the triangular array asymptotics is used. The following theorem shows the limiting distribution of the estimator

. For this theorem we need an additional assumption.

Assumption 7.

where

THEOREM 2. Suppose that Assumptions 1–7 hold and that θ0 is an interior point of Θ. Then,

where

.

The limiting distribution given in Theorem 2 depends on nuisance parameters in a complicated way that renders the NLLS estimator inefficient and, in general, makes it unsuitable for hypothesis testing. This difficulty is removed in a special case where the processes vt and ut are totally uncorrelated, because then the limiting distribution becomes mixed normal as can be easily checked.

In its general form, Theorem 2 shows that the NLLS estimator is consistent of order Op(T−1/2). This will be used to obtain an efficient two-step estimator based on the leads and lags modification. The reason why the order of consistency differs from Op(T−1) obtained in previous linear cases is that we employ the triangular array asymptotics in which the regressand is made bounded.

3.3. Efficient Estimation

This section considers efficient estimation of model (1) by using a leads and lags regression. As in Saikkonen (1991), we can express the error term ut as

where et is a zero-mean stationary process such that Eet vtj′ = 0 for all j = 0,±1,…, and

That this summability condition holds follows from condition (6) and Theorem 3.8.3 in Brillinger (1975). Expressions for the spectral density function and long-run variance of the process et can be obtained from the well-known formulas fee(λ) = fuu(λ) − fuv(λ) fvv−1(λ) fvu(λ) and ωe2 = ωu2 − ωuvΩvv−1ωvu, respectively.

Using equations (5) and (14), we can write the cointegrating regression (1) as

where Δ signifies the difference operator and

To eliminate errors caused by truncating the infinite sum in (14) we have to consider asymptotics in which the integer K tends to infinity with T. The condition K = o(T3) used in the linear case by Saikkonen (1991) can also be used here.

Because we continue with the same triangular array asymptotics as in the previous section, we imbed model (16) in a sequence of models

where Vt = [ΔxtK′…Δxt+K′]′ and π = [πK′…πK′]′. Combining the regressors as q(xtT;θ) = [f (xtT;θ)′ Vt′]′ we can write this model more compactly as

where β = [φ′ π′]′.

Instead of proper nonlinear least squares estimators of the parameters in (18) we shall consider two-step estimators based on the NLLS estimator of the previous section. These estimators are defined by

where

with

The latter term on the right-hand side of (19) is obviously the least squares estimator obtained from a regression of

. The estimator defined in (19) will be called the Gauss–Newton estimator.

To see the motivation of the Gauss–Newton estimator, subtract

from both sides of (17) and apply the mean value approximation

to the right-hand side. Thus, after linearization, we get the auxiliary regression model

which in conjunction with standard least squares theory gives estimator (19).

The following theorem describes asymptotic properties of the estimators

. The limiting distribution of the estimator

requires a standardization by the square root of T − 2K, the effective number of observations in the regression of

. For convenience, we denote N = T − 2K.

THEOREM 3. Suppose that the assumptions of Theorem 2 hold and that K → ∞ in such a way that K3/T → 0 and T1/2 [sum ]| j|>K∥πj∥ → 0. Then,

The independence of the Brownian motions Be(s) and Bv(s) implies that the limiting distribution in Theorem 3 is mixed normal. Furthermore, we can conclude from Saikkonen (1991) that the Gauss–Newton estimator

is asymptotically more efficient than the least squares estimator

in general. In the same way as in Saikkonen (1991), we have also here been forced to supplement the previously mentioned condition K = o(T3) by an additional condition that implies that the integer K may not increase too slowly.

Theorem 3 indicates that we can estimate ωe2 consistently (see, e.g., Andrews, 1991) by using the residuals from the regression model (16) with estimator (19). Thus, conventional tests like Wald and t-tests can be constructed in a straightforward manner and shown to have standard distributions in the limit.

4. SIMULATION

Implications of the theoretical results in Section 3 can be summarized as follows. (i) The NLLS and Gauss–Newton estimators are consistent. (ii) In large samples, the Gauss–Newton estimator eliminates the bias in the limiting distribution of the NLLS estimator and is more efficient than the NLLS estimator. (iii) The t-test based on the Gauss–Newton estimator follows a standard normal distribution in the limit. Because these results are based on the triangular array asymptotics where the sample size of the embedding model goes to infinity, it may not seem quite obvious whether these results hold when the sample size T0 is large. Therefore, this section examines the aforementioned results by using simulation.

Data were generated by

Larger ω implies that the regressors and errors are more correlated both serially and contemporaneously. We plotted a typical, simulated data set with ω = 0.5 in Figure 1. It shows that the relation between the regressor and regressand gradually changes as the value of the regressor becomes closer to 5.

Data under smooth transition (T = 150, δ = 1, c = 5).

Unreported simulation results indicate that it is difficult to estimate parameter γ accurately by the NLLS method unless either sample sizes are very large or parameter c is located close to the median of {xt}, and the results also indicate that other parameter estimates are quite adversely affected by poor estimates of γ. No doubt, the occasional poor performance of the NLLS estimator for unknown γ is neither related to the use of triangular array asymptotics nor due to the presence of I(1) variables. It may also occur when a given STR model involves only I(0) variables for which standard asymptotics can be used. Because the purpose of this section is to check the implications of the triangular array asymptotics, we do not want our simulation results to be affected by outliers produced by poor estimates of parameter γ. Therefore, we assume that the value of the transition parameter γ is known to be 1. Also, {xt} were generated such that c is located in between the 15th and 85th percentiles of {xt}. The purpose of this scheme is the same as that of fixing the value of γ. When c is near endpoints of the sample, extremely poor estimates of parameter c are sometimes produced, which affects other parameter estimates to the extent that evaluating their finite-sample performance at different sample sizes becomes meaningless.

The estimators considered are the NLLS, one-step Gauss–Newton, and two-step Gauss–Newton estimators.4

The one-step Gauss–Newton and two-step Gauss–Newton estimators use the NLLS and one-step Gauss–Newton estimators as initial estimators, respectively.

The values of the leads and lags parameter for the Gauss–Newton estimators were set at K = 1,2,3. Table 1 reports the empirical biases and root mean squared errors (RMSEs) of the estimators at sample sizes 150 and 300.5

We do not report the results for the estimators of μ, because these are not the main concern in most applications.

The numbers of replications at T0 = 150 and T0 = 300 were 5,000 and 3,000, respectively. As for the method of minimization, the Polak–Ribiere conjugate gradient method6

It was found that quasi-Newton methods tend to give more outliers. The maximum number of iterations for optimization was set at 100,000.

was used. The results in Table 1 can be summarized as follows.
  • As sample size T0 grows, the RMSEs of all the estimators decrease, which may be interpreted as evidence for consistency.
  • The Gauss–Newton estimators reduce the magnitudes of bias and RMSE substantially in relation to the NLLS estimator as predicted by Theorem 3.
  • As the regressors and errors are more correlated both serially and contemporaneously, the two-step Gauss–Newton estimator tends to improve on the one-step Gauss–Newton estimator in terms of RMSE. But the two-step Gauss–Newton estimator is sometimes more biased than the one-step Gauss–Newton estimator, though the degree of the biases for both the estimators is quite mild.
  • The choice of the parameter K does not seem to affect the results significantly.
  • The nonlinear parameter c is subject to higher RMSE than other linear parameters, which may reflect the computational difficulties associated with estimating the nonlinear parameter.

Biases and root mean squared errors

Table 2 reports empirical sizes of the t-ratios using the Gauss–Newton estimators under the null hypotheses α = 1, δ = 1, and c = 5. Nominal sizes were chosen to be 5% and 10%, and the same experimental format as for Table 1 was used. The results in Table 2 can be summarized as follows.

  • The t-ratios reject more often than they should in part (a). But increasing the sample size T0 to 300 improves the performance of the t-ratios.7

    Increasing the sample size to 500 further improves the empirical size of the t-ratio, though the results are not reported here.

  • When there are fewer serial and contemporaneous correlations between the regressors and errors at T0 = 300, empirical sizes get closer to the corresponding nominal sizes. But this is not noticeable at T0 = 150.
  • The one-step and two-step Gauss–Newton estimators show similar performance.
  • Choosing K = 1 and K = 2 at T0 = 150 and T0 = 300, respectively, tends to provide the best results.

Empirical sizes of the T-ratios

In summary, the simulation results in Tables 1 and 2 seem to confirm that the results from the triangular array asymptotics in Section 2 can provide reasonable approximations for the finite-sample properties of the estimators and tests when the sample size is moderately large.

5. AN EMPIRICAL EXAMPLE

One of the substantial controversies regarding the Asian currency crisis of 1997 has been whether tight monetary policy was effective in stabilizing foreign exchange rates during and in the aftermath of the crisis. See Goldfajn and Baig (1998), Kaminsky and Schmukler (1998), Ghosh and Phillips (1998), Kraay (1998), Dekle, Hsiao, and Wang (1999), Park, Wang, and Chung (1999), and Choi and Park (2000) for empirical results regarding this issue. In fact, tight monetary policy constituted an essential part of the IMF rescue package for Asian countries, because it has conventionally been believed that higher interest rates reduce capital outflows by raising the cost of currency speculation and induce capital inflows by making domestic assets more attractive in the short run and also that they improve current account balance by reducing domestic absorption in the long run.

However, as discussed in Goldfajn and Baig (1998), higher interest rates may depreciate a currency when interest rates are too high because excessively high interest rates may increase the default risk by increasing the borrowing cost of corporations, by depressing the economy and by weakening the banking system of an economy.8

In addition, Feldstein (1998), Furman and Stiglitz (1998), and Radelet and Sachs (1998a, 1998b), among others, argue that tight monetary policy in Asia either was ineffective in stabilizing exchange rates or may have even exacerbated the situation.

This hypothesis may be called the “interest Laffer curve” hypothesis because the effects of interest rates on spot exchange rates are hypothesized to depend on the levels of the interest rates. This section employs the model and asymptotic theory developed in previous sections to study the interest Laffer curve hypothesis and reports the magnitudes of interest elasticity of the spot rate for Korea and Indonesia during the Asian currency crisis.

The uncovered interest rate parity relation predicts that log spot rate is related to the difference of domestic and foreign interest rates and log expected future spot rate.9

The uncovered interest rate parity relation is written as 1 + it = (1 + it*)St+1e/St, where it and it* denote the domestic and the foreign interest rates at date t, respectively, and St and St+1e denote the spot exchange rate at date t and the expected future spot exchange rate at date t + 1, respectively. Taking logs of both sides of the interest parity relation yields ln(St) = ln(1 + it*) − ln(1 + it) + ln(St+1e) ≈ it* − it + ln(St+1e).

Though the relation predicted by the interest rate parity condition is approximately linear, it indicates that the difference of the domestic and foreign interest rates and the log expected future spot rate may be considered as major variables explaining the spot rate. This consideration leads us to employ the difference of the domestic and foreign interest rates and the log expected future spot rate as regressors in our nonlinear regression. But because the expected future spot rate is not observable, forward exchange rate can be used as its substitute. One may wonder at this point why we did not invoke the covered interest parity relation from the beginning. However, in Korea, there has been no well-developed forward exchange market. Without such a market, it would be nonsensical to consider the covered interest parity relation.

More specifically, the STR model we use in this section is

where yt and x1t are the spot and forward rates, respectively, and x2t is the difference between the domestic and foreign interest rates (i.e., itit*). Because we are interested only in the nonlinear relation between the spot rate and the interest rate differential, the transition function includes only the interest rate differential. Equation (21) signifies that the relation between the spot rate and the interest rate differential changes when the latter is well above the level c unless γ is zero. Thus, the model is appropriate for studying the relation between the spot rate and the interest rate differential, which may change depending on the level of the interest rate differential.

The spot exchange rate data that we use are daily nominal exchange rates of Korea and Indonesia vis-à-vis the U.S. dollar. For forward exchange rates, 1-month maturity data are used. For Korea, we use the forward exchange rate from the nondeliverable forward (NDF) market.10

The NDFs are nondeliverable forwards traded in the offshore market. Unlike the onshore forward exchange rates, which have been influenced by direct regulation and heavy intervention of the Korean government, we believe that the NDF rates better reflect expectations of market participants.

For Indonesia, we use data from their onshore forward exchange markets.11

Because Indonesia had already liberalized domestic foreign exchange markets, the Indonesian rupiah was not traded in the NDF market.

For domestic interest rates, we use the overnight call rates of each country. Because the overnight call rates are the main monetary policy instruments of each country, they seem to best reflect monetary policy stances of each country and could be regarded as exogenous policy variables. The U.S. federal funds rate is used as the foreign interest rate.

The whole sample covers the 19-month periods 4/1/1997–10/30/1998 for Korea and 1/3/1997–7/24/1998 for Indonesia. The sample period for each country begins at about 6 to 7 months before the eruption of its own currency crisis. The sample sizes for Korea and Indonesia are 386 and 406, respectively. Figures 2 and 3 plot the Korean and Indonesian data. These figures demonstrate the volatility of the data during the sampling period.

Interest rate differential and spot rate (Korea).

Interest rate differential and spot rate (Indonesia).

The Dickey–Fuller-GLSμ test of Elliott, Rothenberg, and Stock (1996) and the LM test of Choi (1994) were applied to the spot and forward rates and the interest differentials for both Korea and Indonesia. The results support the presence of a unit root at conventional levels, and hence the theoretical results in previous sections are relevant here. Truly, the interest differentials should be I(0) under normal circumstances. Otherwise, there are unrealistically many arbitrage opportunities. But during the currency crisis period, the test results indicate that they look like I(1) at least in Korea and Indonesia. Probably, market participants during the period did not perceive the diverging interest differentials as arbitrage opportunities in the light of the huge risks involved.

The results of the one-step and two-step Gauss–Newton estimation of model (21) are reported in Table 3.12

Prior to estimating the STR model, it is proper to perform linearity tests. But the linearity tests for models with I(1) variables are not yet available, so we bypass the stage of hypothesis testing.

The Polak–Ribiere conjugate gradient method was used for initial estimation as in Section 4. The results for Korea in Table 3 show that the forward rate and interest rate differential are significant at conventional levels excepting a couple of cases.13

Needless to say, this statement assumes that the given nonlinear model represents the true data generating process.

For the parameters inducing nonlinearity, conventional hypothesis testing is difficult.14

We thank Bruce Hansen for pointing out this problem.

When δ = 0, c and γ are not identified. In addition, when γ = 0, δ and c are not identified. Thus, conventional testing procedures cannot be used for δ and γ. Testing the null hypotheses δ = 0 and γ = 0 is equivalent to testing the null of linearity. The standard errors for the parameter γ are relatively high, which indicates the difficulty of estimating the parameter. The location parameter c is estimated to lie between 13.80 and 16.43 depending on the choice of the leads and lags parameter K and the estimation method.15

These results are based on the assumptions that the error term in equation (21) is I(0) and that regressors are not cointegrated. Formal tests for cointegration for the STR model are not yet available. But fitting the AR(1) regression for the residuals from equation (21) using the parameter values in the first row of Table 3, we obtained AR(1) coefficient estimate 0.557 and corresponding standard error 0.043. Similar results were obtained for other parameter values. Thus, it seems unlikely that the residuals are I(1). In addition, we tested for cointegration between the future rates and interest differentials but found no evidence of cointegration.

Gauss–Newton estimation results

For Indonesia, nonlinear effects of the interest differentials seem to be weak relative to Korea, though coefficients for the forward rate and the interest differential (α1 and α2) are significant at conventional levels. The estimates of the location parameter c are similar in magnitudes to those for Korea, though Indonesia experienced much higher interest rates than Korea during the period of currency crisis.16

Indonesia's maximum call rate during the sample period was 91.5%, and the average was 29.4%. But the maximum and average for Korea were 35% and 15.6%, respectively.

The results in Table 3 indicate that the future rates are quite important in explaining the spot rates given the magnitudes of the coefficient estimates. But the coefficient estimates for the terms involving the interest differentials are close to zero. To visualize the nonlinear effects of the interest differentials, we draw the interest differential elasticity of spot rate in Figures 4 and 5 by assuming that the estimation results in Table 3 (using the one-step estimation method with K = 1) represent the true relation.17

Ignoring the error term in equation (21) and assuming that the parameter estimates are the true parameter values, the elasticity was calculated by using the formula ∂yt/(1/x2t)∂x2t = ∂yt/∂ ln(x2t) . Here the partial derivative is multiplied by x2t, because log was taken for the spot rate but not for the interest differential.

Interest elasticity of spot rate (Korea).

Interest elasticity of spot rate (Indonesia).

Figures 4 and 5 show that when the interest differentials take values lower than approximately 11% and 12% for Korea and Indonesia, respectively, the conventional wisdom that increasing interest rate helps stabilize spot rate seems to be supported. But when the interest differentials take higher values up to approximately 28% and 16% for Korea and Indonesia, respectively, the elasticities become positive, which implies that increasing interest rate has negative effects on stabilizing spot rate. When the interest differentials are above 28% and 16% for Korea and Indonesia, respectively, the elasticities become negative again.

Figures 4 and 5 partially support the interest Laffer curve hypothesis. But they also indicate that tight monetary policy is effective, though very weakly, when interest rates are very high. Notwithstanding this remark, we conclude from the magnitudes of the elasticities shown in Figures 4 and 5 that the effects of interest rate on the spot rate are negligible in either direction.18

Choi and Park (2000) also report that interest differential did not cause spot rate in both the short and long runs during the Asian currency crisis.

For example, when the interest differential is 30% for Korea, the elasticity is only −0.007. This implies that raising the interest differential from 30% to 33% (i.e., a 10% increase) has the effect of appreciating the Korean currency by only 0.07%. Considering the fact that the currency was depreciated by approximately 30% on the average during the sampling period, such a meager effect is certainly unsatisfactory to the Korean economy. This is more so when one considers the negative effects of an interest rate increase of such magnitude on the corporations and banking system of the economy.

6. FURTHER REMARKS

We have analyzed and applied the cointegrated STR model in this paper. However, there are a couple of topics that deserve our attention but were not studied in this paper. First, methods for testing linearity in the presence of I(1) variables are not yet available but are useful for empirical analyses. Because nonlinear models are flexible, they may give a good in-sample fit even when the true model is linear. Thus, testing linearity prior to nonlinear model fitting is important. Second, testing for cointegration for the STR model should precede estimation, but relevant methods are not yet available. We hope that these topics can be studied in the future by the authors and other researchers.

APPENDIX A: AUXILIARY LEMMAS

We shall first prove some auxiliary results that may also have applications elsewhere. Recall the notation N = T − 2K and note that a (possibly) matrix-valued function h(x) defined on

is said to be locally bounded if ∥h(x)∥ is bounded on compact subsets of

.

LEMMA 1. Let h(x) be a locally bounded, vector-valued function defined on

(d < ∞) and lett,

be a square integrable martingale difference sequence such that supt E∥εt2 < ∞. Let ζtT(1) (d × 1) and ζtT(2) (t = 1,…,T) be random vectors defined on the same probability space as εt. Assume that max1≤tT∥ζtT(1)∥ = Op(1) and supt,T E∥ζtT(2)∥ < ∞. Then,

when ζtT(1) is measurable with respect to the σ-algebra

. The third result also holds with ζtT(1) replaced by ζt+j−1,T(1).

Proof. To prove the first assertion, let ε > 0 and use the assumption max1≤tT∥ζtT(1)∥ = Op(1) to choose m > 0 such that P{max1≤tT∥ζtT(1)∥ > m} < ε for all T large. Next, use the assumption that h(x) is locally bounded to conclude that Hm = supx∥≤mh(x)∥ is finite. Then, the desired result follows because for all T large

The second result is an immediate consequence of the first result and the moment condition imposed on ζtT(2). To prove the third assertion, first note that an application of the triangular inequality yields

Now, let ε > 0 and define m and Hm in the same way as in the proof of (i). Then, for every M > 0 and T large

As for A1T, use the assumptions that

is a square integrable martingale difference sequence and that ζtT(1) is measurable with respect to the σ-algebra

to obtain

Hence, P{(N1/2/K)|A1T| > M/2} ≤ 2CHm /M by Markov's inequality, and we can conclude that for every M and T large

For M > 2CHm /ε the last expression is smaller than 2ε, which proves the stated result. A similar proof shows the final assertion. █

Note that the first two results of Lemma 1 obviously hold when h(x) and ζt+j,T(2) are matrix-valued and that the third result improves Lemma A.4(c) of Park and Phillips (2001) by relaxing the exponentially boundedness assumption used therein to local boundedness.

The first two results of Lemma 1 can be applied with the process

where wt is as in Assumption 2 and z0 may be any random vector such that Ez04 < ∞. In this case ζtT(1) = ztT = T−1/2zt and max1≤tTztT∥ = Op(1) is an immediate consequence of the invariance principle (8). This definition of ztT will be assumed in subsequent lemmas. The proofs of these lemmas make use of the fact that, as a result of Assumption 2, we can write

where

with Et the conditional expectation operator with respect to the σ-algebra

(cf. Hansen, 1992). Because {ηt,

is a stationary martingale difference sequence equation (A.2) is analogous to the so-called Beveridge–Nelson decomposition, which has been used extensively in asymptotic analysis of linear processes (see, e.g., Phillips and Solo, 1992). Therefore, we shall refer to equation (A.2) as the Beveridge–Nelson decomposition also in the present context. In our applications of the third result of Lemma 1 the martingale difference sequence εt will be ηt. For these applications, and also for other subsequent derivations, it is worth noting that the (stationary) processes ηt and ξt have finite moments of order 4 (see Hansen, 1992, the proof of Theorem 3.1).

LEMMA 2. Let h(x;θ) be a (possibly) vector-valued continuously differentiable function defined on

where Θ* is an open set in an Euclidean space. Suppose thath(x;θ)/∂x is also continuously differentiable and let Θ ⊂ Θ* be a compact set containing the point θ0 in its interior. Then, as K2/T → 0,

where

is a random vector such that

.

Proof. We shall first prove the latter assertion and then note how the first one can be obtained from the employed arguments. Without loss of generality, assume that h(x;θ) is real-valued and use the Beveridge–Nelson decomposition (A.2) in conjunction with the triangular inequality to obtain

First, consider

and use partial summation to obtain

Hence, using the triangular inequality we find that

Because supθ∈Θh(x;θ)∥ is locally bounded, the first two terms on the right-hand side are easily seen to be of order Op(K/N). For the third term we can use a standard mean value expansion to get

where H1(x;θ) = ∂h(x;θ)/∂x′ and ∥zt−1,TztT∥ ≤ ∥zt−1,TztT∥ = T−1/2wt∥. Thus, we can write

Here the latter inequality is justified by the triangular inequality whereas the equality follows from Lemma 1(ii) because supθ∈ΘH1(x;θ)∥ is locally bounded, max1≤tTzt−1,T∥ = Op(1), and Ewtξt−1+j′∥ is a finite constant. For later purposes we note that we actually showed that A4T(θ) = Op(K/N1/2) holds uniformly in θ ∈ Θ.

Next, consider

. Because θ0 is an interior point of Θ and

, we can use the mean value expansion

where

. Thus, using the triangular inequality one obtains

The first term on the right-hand side is A3T0), and the second term can be bounded by

Here the equality is again obtained from Lemma 1(ii) because supθ∈ΘH2(x;θ)∥ is locally bounded, max1≤tTztT∥ = Op(1), and E∥ηt+j∥ is constant. Thus, to complete the proof, we have to show that A3T0) = Op(K/N1/2).

By the definition of A3T0),

Lemma 1(iii) implies that A31T0) = Op(K/N1/2), so we need to show that the same holds true for A32T0). To this end, use the Beveridge–Nelson decomposition (A.2) and the definition of ztT to give

where

. Thus, a mean value expansion yields

where

. This identity and the triangular inequality imply

Here the equality is obtained from Lemma 1(iii), which obviously applies despite the differences in subscripts. Next note that, because the function H1(x;θ) is continuously differentiable by assumption and because ∥stj−1,Tstj−1,T∥ ≤ 2T−1/2rtj∥, we have ∥H1(stj−1,T0) − H1(stj−1,T0)∥ ≤ T−1/2H1T0)∥rtj∥ where H1T0) is determined by the partial derivatives of the function H1(x0) and, by Lemma 1(i), H1T0) = Op(1). Combing these facts with the preceding upper bound of |A32T0)| it is straightforward to show that

Consider the second term on the right-hand side. By the Cauchy–Schwarz inequality, Ertj2∥ηtj∥ ≤ (Ertj4E∥ηtj2)1/2c1(j + 1) where c1 is a finite constant. To justify the latter inequality here, observe that, for some finite constants c2, c3, and c4,

where the inequalities can be obtained from the definitions and Theorem 3.7.8(i) of Stout (1974). Thus,

and, because H1T0) = Op(1), it follows that the second term on the right-hand side of (A.4) is of order Op(K/N1/2).

To complete the proof of the first assertion, we still need to show that the first term on the right-hand side of (A.4) is of order Op(K/N1/2). It suffices to replace rjt in turn by each of the four components in its definition. Thus, consider the quantity

Arguments similar to those used for

in (A.3) show that the first term on the right-hand side is of order Op(K/N1/2). These arguments also apply when the last three terms in the definition of rtj are considered. Thus, we only need to show that the latter term in the last expression is of order Op(K/N1/2). Using the triangular inequality, one obtains

To show that the last quantity is of order Op(K/N1/2), we can make use of a similar truncation argument as in the proof of Lemma 1(iii) and replace the function H1(x0) by 1{∥x∥ ≤ m}H1(x0) with an appropriately chosen real number m. Thus, because H1(x0) is locally bounded 1{∥x∥ ≤ m}H1(x0) is bounded. To simplify notation we proceed by assuming that the function H1(x0) itself is bounded. Assuming this shows that for i ≥ 1

where the equality follows because the terms in the preceding sum are uncorrelated with bounded second moments. Thus, the right-hand side of (A.5) is of order Op(K2/N), which proves the desired result and completes the proof of the second assertion.

To prove the first assertion, notice that we need to show that A3T(θ) and A4T(θ) are of order Op(K/N1/2) for every fixed θ. For A4T(θ) we showed that this holds even uniformly in θ. As for A3T(θ), it suffices to consider A31T(θ) and A32T(θ) separately. In the preceding proof we showed that A31T0) and A32T0) are of order Op(K/N1/2), and an inspection of the proof reveals that θ0 can be replaced by any θ ∈ Θ without changing the result. This completes the proof of Lemma 2. █

It would be useful to be able to show that the pointwise result of Lemma 2(i) also holds uniformly in θ, but we have been unable to obtain this extension. The following result is not difficult to obtain, however.

LEMMA 3. Suppose the assumptions of Lemma 2 hold and let RT = [RKT′…RKT′]′ be a (possibly) stochastic matrix such that each RjT has p + 1 rows and, for some finite constant c, ∥RT∥ ≤ c (a.s.). Then,

Proof. Without loss of generality assume that c = 1 and that h(x;θ) is real-valued. Because ∥RjT∥ ≤ 1 for all j, we have for every fixed θ ∈ Θ

where the equality is due to Lemma 2(i). Thus, the problem is to strengthen this pointwise convergence in probability to uniform convergence in probability. Because Θ is a compact set it suffices to show that the quantity whose norm is taken is stochastically equicontinuous (see, e.g., Davidson, 1994, p. 337). To this end, let θ1 and θ2 be arbitrary points of Θ and consider the quantity

where the inequality follows from the Cauchy–Schwarz inequality. For the difference in the last expression we can use the mean value expansion

where H2(x;θ) = ∂h(x;θ)/∂θ′ and ∥θ − θ1∥ ≤ ∥θ1 − θ2∥. Thus,

where the equality is justified by Lemma 1(i) because supθ∈ΘH2(x;θ)∥2 is locally bounded and max1≤tTztT∥ = Op(1). Hence, the desired stochastic equicontinuity follows in a straightforward manner from (A.6) if we show that the latter factor in the last expression therein is of order Op(1). To see this, define the matrix

and let λmax(·) denote the largest eigenvalue of the indicated matrix. With these definitions we have

Here the last relation is a straightforward consequence of the fact that the spectral density matrix of the process wt is bounded, and the preceding one follows from the assumption ∥RT∥ ≤ 1 (a.s.). Thus, the proof is complete. █

The results of Lemmas 2 and 3 also hold with a fixed value of K. In that case RjT in Lemma 3 may be replaced by an identity matrix, as can easily be checked from the given proofs.

In the following lemma we use the notation C(Θ)a×b to signify the space of all continuous functions from the compact set

endowed with the uniform metric. In

the usual euclidean metric is assumed.

LEMMA 4. Let H(x,θ) (a × b) be a matrix-valued continuous function defined on

. Then, if K/T → 0

where the convergence holds in the function space C(Θ)a×b.

Proof. Because ztTB(s) by (8) the proof can be obtained in the same way as the first result in Theorem 3.1 of Park and Phillips (2001). █

Lemma 4 can be used to prove the following lemma.

LEMMA 5. Let f (x;θ), θ ∈ Θ, and xtT be as in Section 3.2. Then there exists an ε > 0 such that with probability approaching one

Proof. The stated result follows from condition (11), Lemma 4, and the continuity of eigenvalues and the infimum function. █

LEMMA 6. Let h(x) be a vector-valued twice continuously differentiable function defined on

. Then,

where

. Moreover, this weak convergence holds jointly with that in (8).

Proof. Using the Beveridge–Nelson decomposition (A.2), one obtains

First, consider the latter term on the right-hand side. By partial summation,

where the latter equality is an immediate consequence of the assumptions. Thus, a standard mean value expansion and the fact ΔztT = T−1/2wt yield

where the notation is as before so that H1(zt−1,T) signifies a matrix each row of which is evaluated at a possibly different intermediate point in the line segment between ztT and zt−1,T. Because the function H1(x) is continuously differentiable by assumption, we have ∥H1(zt−1,T) − H1(ztT)∥ ≤ T−1/2H1Twt∥ where H1T is determined by the second partial derivatives of the function h(x) and, as a straightforward consequence of Lemma 1(i), H1T = Op(1). Hence, because Ewt∥∥wtξt−1′∥ is a finite constant, we can write

where the latter equality follows from the Beveridge–Nelson decomposition (A.2). Theorems 3.2 and 3.3 of Hansen (1992) imply that replacing wtξt′ in the first term of the last expression by its expectation causes an error of order op(1). To see that a similar replacement can be done in the second term of the last expression, observe that, by Assumption 2 and the mixing inequality in Davidson (1994, p. 211), wt wt′ − Ewt wt′ is a stationary L1-mixingale. Hence, the desired result follows from Theorem 3.3 of Hansen (1992). As a whole we can thus conclude that

where the result Ewtξt′ + Ewt wt′ = Λ is a simple consequence of the definition of the matrix Λ and the process ξt (cf. Hansen, 1992, proof of Theorem 4.1).

Now consider the first term on the right-hand side of (A.7) and use the same mean value expansion as before to write

In the same way as previously, we can also here replace H1(zt−1,T) by H1(ztT) and combine equations (A.8) and (A.9) with (A.7). This gives

To complete the proof, notice that ηt is a stationary square integrable martingale difference sequence and that an invariance principle holds jointly for the processes

(see Hansen, 1992, proof of Theorem 3.1). Hence, the stated result is obtained from Theorem 2.1 of Hansen (1992). █

APPENDIX B: PROOFS OF MAIN RESULTS

Proof of Theorem 1. We shall first demonstrate the existence of the estimators

. For any fixed value of θ, the least squares estimator of φ, denoted by

, exists and is unique with probability approaching one. This is an immediate consequence of the definition of the estimator

and Lemma 5. Thus, we have

It is straightforward to check that, when the estimator

exists and is unique,

is a continuous function of θ so that, by the assumed compactness of the parameter space Θ, there exists

such that

equals the preceding infimum. Thus,

are the desired least squares estimators.

The next step is to show that

is bounded in probability. To this end, notice that

Lemma 5 implies that the largest eigenvalue of the inverse on the right-hand side is of order Op(1). Thus, we have to show that the latter factor on the right-hand side is of order Op(1). To see this, note that the assumptions imply that supθ∈Θf (x;θ)∥ is locally bounded. Therefore, by Lemma 1(i) we have

and similarly with

replaced by θ0. Hence, it follows that

. Moreover, because

holds trivially by the compactness of the parameter space Θ, we have

, which means that the sequence of estimators

is tight.

To prove the consistency of the estimators

, use the definitions to write

Because

the latter equality follows from Lemma 3 with K = 0. Now suppose that

does not hold. Then, by the tightness of the sequence

, we can find a subsequence

that converges weakly to ϑ* = [θ*′ φ*′]′, say, and ϑ* ≠ ϑ0 with a positive probability (see Billingsley, 1968, Theorem 6.1). Thus, we can conclude that

where the weak convergence is justified by Lemma 4 and Lemma A.2 of Saikkonen (2001). (The latter lemma requires that the relevant quantities converge jointly, which can be guaranteed by redefining the subsequence if necessary.) When ϑ* ≠ ϑ0 it follows from condition (12) that the difference in the weak limit in the preceding expression is nonzero for some value of s and, by continuity, in an open interval. Thus, the last expression is positive with a positive probability. This gives a contradiction, so we must have ϑ* = ϑ0. This completes the proof. █

Proof of Theorem 2. For simplicity, denote h(xtT;ϑ) = f (xtT;θ)′φ so that

. Because θ0 is assumed to be an interior point of Θ, the consistency of the estimator

justifies the mean value expansion

where the notation is as before so that ∂2QT(ϑT)/∂ϑ∂ϑ′ signifies a matrix each row of which is evaluated at a possibly different intermediate point in the line segment between

. The partial derivatives can be expressed as

Next, note that

Because the function f (x;θ) is three times continuously differentiable by assumption, it follows from the consistency of the estimator

and Lemma 2(ii) with K = 0 that the first term on the right-hand side is of order op(1). It can be seen that the same is true for the second term by taking a mean value expansion of the difference in the brackets and using the local boundedness of the resulting summands in conjunction with Lemma 1(i) and the consistency of the estimator

. Thus, we can write

Here the weak convergence can be justified by using the consistency of the estimator

, Lemma 4, and Lemma A.2 of Saikkonen (2001). The expression of the limit follows from the definitions.

To complete the proof, use Lemma 6 and the definitions to conclude that

where the weak convergence holds jointly with that in (B.2). Thus, because the weak limit in (B.2) is positive definite (a.s.) by assumption the result of the theorem is an immediate consequence of (B.1)–(B.3) and the continuous mapping theorem. █

Proof of Theorem 3. Denote again f (xtT;θ)φ = h(xtT;ϑ) and conclude from the definitions that

where

. For simplicity, denote

Then,

The latter equality is obtained by replacing H2(xtT;ϑT) in the second expression by

and observing that

. We shall show next that

To this end, notice that, because the function H2(x;θ) is continuously differentiable by assumption, a mean value expansion and an application of Lemma 1(i) show that

where the latter equality is due to the T1/2-consistency of the estimator

obtained from Theorem 2. Thus, because

, the local boundedness of supθ∈ΘH2(x;ϑ)∥ and Lemma 1(i) similarly yield

. Hence, (B.5) holds with

replaced by

, and we need to show that it also holds with

replaced by Vt. This can be seen by observing that

where the last relation follows from Lemma 2(ii) and the T1/2-consistency of the estimator

. Thus, we have established (B.5).

The next step is to observe that

where, denoting λmax(A) as the largest eigenvalue of matrix A, ∥A1 = (λmax(AA))1/2 and

To see this, first note that

again by Lemma 2(ii) and the T1/2-consistency of the estimator

. This and the well-known fact ∥·∥1 ≤ ∥·∥ imply that

, and we need to show that a similar result holds for the corresponding inverses. By Lemma A.2 of Saikkonen and Lütkepohl (1996), this holds true if ∥MT−11 = Op(1) or if

. The former requirement can be obtained from condition (13), the consistency of the estimator

, Lemma 5, and Lemma A.2 of Saikkonen (2001) whereas the latter can be deduced from Lemmas A2–A4 of Saikkonen (1991). Because the assumptions used in Saikkonen (1991) were slightly different from the present ones we note that these lemmas, and also Lemmas A5 and A6 of that paper, can also be proved under the present assumptions. For Lemmas A3 and A5 the previous proofs apply, whereas Lemma A2 and, consequently, Lemmas A4 and A6 can be proved by using Lemma 2(i) of this paper and the fact that, for some finite constant C independent of j = −K,…,K,

This follows from Assumption 2 and Lemma 6.19 of White (1984). For later purposes we also note that the preceding discussion implies that

.

Next, note that

. The former result will become evident subsequently, whereas the latter is obtained from Lemmas A5 and A6 of Saikkonen (1991). Because

we can use (B.5), (B.6), and the T1/2-consistency of the estimator

to conclude from (B.4) that

Because K3/2/N1/2 → 0 by assumption this implies that

Here the last equality follows from results obtained in the Appendix of Saikkonen (1991) and already used earlier. To show that the limiting distribution of

is as stated in the theorem and thereby to complete the proof, first note that the arguments used for (B.2) in the proof of Theorem 2 show that the inverse on the right-hand side of (B.7) converges weakly to the inverse in the theorem. Thus, we need to consider

where the equalities can be justified as follows. First, recall that

and note that EaKt2 = op(T−1) for all t, as shown in the proof of Lemma A5 of Saikkonen (1991). Thus, the first equality in (B.8) follows because

, as already noted. To justify the second equality, recall that

, take a mean value expansion of

about ϑ0, and use the T1/2-consistency of the estimator

in conjunction with Lemma 3 with K = 0.

To complete the proof we have to show that the first term in the last expression of (B.8) converges weakly to the stochastic integral in the theorem and that this holds jointly with the weak convergence of the inverse on the right-hand side of (B.7). If the process [vtet′]′ fulfilled the conditions of Assumption 2 this would follow from Lemma 6, but, because the process et is not guaranteed to be strong mixing, this reasoning does not apply directly. However, using L to denote the usual lag operator we may write et = a(L)′wt where

. In view of the summability condition (15) and Lemma 6 we can use Theorem 4.2 of Saikkonen (1993) and obtain the needed weak convergence results. The assumptions required to apply this theorem are straightforward consequences of Assumption 2, which, in addition to the summability condition (6) and the invariance principle (8), also implies that the first and second sample moments of wt are consistent estimators of their theoretical counterparts. This completes the proof. █

References

REFERENCES

Andrews, D.W.K. (1991) Heteroskedasticity and autocorrelation consistent covariance matrix estimation. Econometrica 59, 817858.Google Scholar
Andrews, D.W.K. & C.J. McDermott (1995) Nonlinear econometric models with deterministically trending variables. Review of Economic Studies 62, 343360.Google Scholar
Billingsley, P. (1968) Convergence of Probability Measures. Wiley.
Brillinger, D.R. (1975) Time Series: Data Analysis and Theory. Holt, Rinehart, and Winston.
Chang, Y. & J.Y. Park (1998) Nonlinear Index Models. Mimeo, Rice University.
Chang, Y., J.Y. Park, & P.C.B. Phillips (1999) Nonlinear Econometric Models with Cointegrated and Deterministically Trending Regressors. Mimeo, Rice University.
Choi, I. (1994) Residual based tests for the null of stationarity with applications to U.S. macroeconomic time series. Econometric Theory 19, 720746.Google Scholar
Choi, I. & D.K. Park (2000) Causal Relation between Interest and Exchange Rates in the Asian Currency Crisis. Mimeo, University of Southern California.
Davidson, J. (1994) Stochastic Limit Theory. Oxford University Press.
Dekle, R., C. Hsiao, & S. Wang (1999) Interest Rate Stabilization of Exchange Rates and Contagion in the Asian Crisis Countries. Mimeo, University of Southern California.
Elliott, G., T.J. Rothenberg, & J.H. Stock. (1996) Efficient tests for an autoregressive unit root. Econometrica 64, 813836.Google Scholar
Feldstein, M. (1998) Refocusing the IMF. Foreign Affairs 77, 2033.Google Scholar
Franses, P.H. & D. van Dijk (2000) Nonlinear Time Series Models in Empirical Finance. Cambridge University Press.
Furman, J. & J.E. Stiglitz (1998) Economic crises: Evidence and insights from East Asia. Brookings Papers on Economic Activity 2, 1135.Google Scholar
Ghosh, A. & S. Phillips (1998) Interest Rates, Stock Market Prices, and Exchange Rates in East Asia. Mimeo, International Monetary Fund.
Goldfajn, I. & T. Baig (1998) Monetary Policy in the Aftermath of Currency Crises: The Case of Asia. IMF Working paper 98-170.
Goldfeld, S.M. & R. Quandt (1973) The estimation of structural shifts by switching regressions. Annals of Economic and Social Measurement 2, 475485.Google Scholar
Granger, C.W.J. & T. Teräsvirta (1993) Modelling Non-linear Economic Relationships. Oxford University Press.
Granger, C.W.J., T. Teräsvirta, & H. Anderson (1993) Modelling non-linearity over the business cycle. In J.H. Stock and M.W. Watson (eds.), New Research on Business Cycles, Indicators and Forecasting, pp. 311325. University of Chicago Press.
Hansen, B.E. (1992) Convergence to stochastic integrals for dependent heterogeneous processes. Econometric Theory 8, 489500.Google Scholar
Jansen, E.S. & T. Teräsvirta (1996) Testing parameter constancy and super exogeneity in econometric equations. Oxford Bulletin of Economics and Statistics 58, 735763.Google Scholar
Kaminsky, G. & S. Schmukler (1998) The Relationship between Interest Rates and Exchange Rates in Six Asian Countries. Mimeo, World Bank.
Kraay, A. (1998) Do High Interest Rates Defend Currencies During Speculative Attacks? Mimeo, World Bank.
Lütkepohl, H., T. Teräsvirta, & J. Wolters (1999) Investigating stability and linearity of a German M1 demand function. Journal of Applied Econometrics 14, 511525.Google Scholar
Luukkonen, R., P. Saikkonen, & T. Teräsvirta (1988) Testing linearity against smooth transition autoregressive models. Biometrika 75, 491499.Google Scholar
Michael, P., A.R. Nobay, & D.A. Peel (1997) Transaction costs and nonlinear adjustment in real exchange rates: An empirical investigation. Journal of Political Economy 105, 862879.Google Scholar
Newey, W. & D.L. McFadden (1994) Large sample estimation and hypothesis testing. In R.F. Engle and D.L. McFadden (eds.), Handbook of Econometrics, vol. 4, pp. 21132445. Elsevier.
Park, J. & P.C.B. Phillips (1999) Asymptotics for nonlinear transformations of time series. Econometric Theory 15, 269298.Google Scholar
Park, J. & P.C.B. Phillips (2001) Nonlinear regressions with integrated time series. Econometrica 69, 117161.Google Scholar
Park, Y.C., Y. Wang, & C.-S. Chung (1999) Exchange Rate Policies in Korea: Has Exchange Rate Volatility Increased after the Crisis? Mimeo, Korea Institute for Economic Policy.
Phillips, P.C.B. & M. Loretan (1991) Estimating long run economic equilibria. Review of Economic Studies 58, 407436.Google Scholar
Phillips, P.C.B. & V. Solo (1992) Asymptotics for linear processes. Annals of Statistics 20, 9711001.Google Scholar
Pötscher, B. & I.R. Prucha (1997) Dynamic Nonlinear Econometric Models. Springer-Verlag.
Radelet, S. & J. Sachs (1998a) The East Asian financial crisis: Diagnosis, remedies, prospects. Brookings Papers on Economic Activity, 174.
Radelet, S. & J. Sachs (1998b) The Onset of the East Asian Financial Crises. Mimeo, Harvard Institute for International Development.
Saikkonen, P. (1991) Asymptotically efficient estimation of cointegration regressions. Econometric Theory 7, 121.Google Scholar
Saikkonen, P. (1993) Continuous weak convergence and stochastic equicontinuity results for integrated processes with an application to the estimation of a regression model. Econometric Theory 9, 155188.Google Scholar
Saikkonen, P. (2001) Consistent estimation in cointegrated vector autoregressive processes with nonlinear time trends in cointegrating relations. Econometric Theory 17, 296326.Google Scholar
Saikkonen, P. & H. Lütkepohl (1996) Infinite order cointegrated vector autoregressive processes: Estimation and inference. Econometric Theory 12, 814844.Google Scholar
Sarno, L. (1999) Adjustment costs and nonlinear dynamics in the demand for money: Italy, 1861–1991. International Journal of Finance and Economics 4, 155177.Google Scholar
Stock, J.H. & M.W. Watson (1993) A simple estimator of cointegrating vectors in higher order integrated systems. Econometrica 61, 783820.Google Scholar
Stout, W.F. (1974) Almost Sure Convergence. Academic Press.
Taylor, M.P., D.A. Peel, & L. Sarno (2001) Nonlinear mean-reversion in exchange rates: Towards a solution to the purchasing power parity puzzle. International Economic Review 42, 10151042.Google Scholar
Teräsvirta, T. (1998) Modeling economic relationships with smooth transition regressions. In A. Ullah and D.E.A. Giles (eds.), Handbook of Applied Economic Statistics, pp. 507552. Dekker.
Teräsvirta, T. & H.M. Anderson (1992) Characterizing nonlinearities in business cycles using smooth transition autoregressive models. Journal of Applied Econometrics 7, S119S136.Google Scholar
Tong, H. (1983) Threshold Models in Non-Linear Time Series Analysis. Springer-Verlag.
van Dijk, D., T. Teräsvirta, & P.H. Franses (2002) Smooth transition autoregressive models: A survey of recent developments. Econometric Review 21, 147.Google Scholar
White, H. (1984) Asymptotic Theory for Econometricians. Academic Press.
Figure 0

Data under smooth transition (T = 150, δ = 1, c = 5).

Figure 1

Biases and root mean squared errors

Figure 2

Empirical sizes of the T-ratios

Figure 3

Interest rate differential and spot rate (Korea).

Figure 4

Interest rate differential and spot rate (Indonesia).

Figure 5

Gauss–Newton estimation results

Figure 6

Interest elasticity of spot rate (Korea).

Figure 7

Interest elasticity of spot rate (Indonesia).