Skip to main content
Log in

A class of Minimum Distance Estimators in Markovian Multiplicative Error Models

  • Published:
Sankhya B Aims and scope Submit manuscript

Abstract

This paper proposes a class of minimum distance estimators for the underlying parameters in a Markovian parametric multiplicative error time series model. This class of estimators is based on the integrals of the square of a certain marked residual process. The paper derives the asymptotic distributions of the proposed estimators. In a finite sample comparison, some members of the proposed class of estimators dominate a generalized method of moments estimator in terms of the finite sample bias at a variety of chosen error distributions while neither dominate each other in terms of the mean squared error at these error distributions. A real data example is considered to illustrate the proposed estimation procedures.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Figure 1
Figure 2

Similar content being viewed by others

References

  • Bauwens, L. and Giot, P. (2000). The logarithmic ACD model: an application to the bid-ask quote process of three NYSE stocks. Ann. Econ. Stat. 60, 117–149.

    Google Scholar 

  • Bauwens, L. and Veredas, D. (2004). The stochastic conditional duration model: a latent factor model for the analysis of financial durations. J. Econometrics119, 381–412.

    Article  MathSciNet  MATH  Google Scholar 

  • Brownlees, C.T., Cipollini, F. and Gallo, G.M. (2012). Multiplicative Error Models. Wiley, New Jersey, Bauwens, L., Hafner, C. and Laurent, S (eds.), p. 225–247.

  • Chang, N.M. (1990). Weak convergence of a self-consistent estimator of a survival function with doubly censored data. Ann. Statist. 18, 391–404.

    Article  MathSciNet  MATH  Google Scholar 

  • Chen, Y. -T. and Hsieh, C. -S. (2010). Generalized moment tests for autoregressive conditional duration models. J. Financ. Econom. 8, 345–391.

    Google Scholar 

  • Chou, R.Y. (2005). Forecasting financial volatilities with extreme values: The conditional autoregressive range (CARR) model. J. Money, Credit Banking37, 561–582.

    Article  Google Scholar 

  • Cipollini, F., Engle, R.F. and Gallo, G.M. (2013). Semiparametric vector MEM. J. Appl. Econ. 28, 1067–86.

    Article  MathSciNet  Google Scholar 

  • Donoho, D.L. and Liu, R.C. (1988a). The automatic robustness of minimum distance functionals. Ann. Statist. 16, 552–586.

    Article  MathSciNet  MATH  Google Scholar 

  • Donoho, D.L. and Liu, R.C. (1988b). Pathologies of some minimum distance estimators. Ann. Statist. 16, 587–608.

    Article  MathSciNet  MATH  Google Scholar 

  • Drost, F.C. and Werker, B.J.M. (2004). Semiparametric duration models. J. Bus. Econ. Stat. 22, 40–50.

    Article  MathSciNet  Google Scholar 

  • Engle, R. (2002). New frontiers for ARCH models. J. Appl. Econ. 17, 425–446.

    Article  Google Scholar 

  • Engle, R.F. and Gallo, G.M. (2006). A multiple indicators model for volatility using intra-daily data. J. Econometrics 131, 3–27.

    Article  MathSciNet  MATH  Google Scholar 

  • Engle, R.F. and Russell, J.R. (1998). Autoregressive conditional duration: a new model for irregularly spaced transaction data. Econometrica. J. Econometric Society 66, 1127–1162.

    Article  MathSciNet  MATH  Google Scholar 

  • Fernandes, M. and Grammig, J. (2006). A family of autoregressive conditional duration models. J. Econometrics 130, 1–23.

    Article  MathSciNet  MATH  Google Scholar 

  • Grammig, J. and Maurer, K.-O. (2000). Non-monotonic hazard functions and the autoregressive conditional duration model. Econometrics Journal 3, 16–38.

    Article  MATH  Google Scholar 

  • Grammig, J. and Wellner, M. (2002). Modeling the interdependence of volatility and inter-transaction duration processes. J. Econometrics 106, 369–400.

    Article  MathSciNet  MATH  Google Scholar 

  • Gao, J., Kim, N.H. and Saart, P.W. (2015). A misspecification test for multiplicative error models of non-negative time series processes. J. Econometrics 189, 346–359.

    Article  MathSciNet  MATH  Google Scholar 

  • Guo, B. and Li, S. (2018). Diagnostic checking of markov multiplicative error models. Econ. Lett. 170, 139–142.

    Article  MathSciNet  MATH  Google Scholar 

  • Hall, P. and Heyde, C.C. (1980). Martingale Limit Theory and Its Applications. Academic Press, New York.

    MATH  Google Scholar 

  • Hautsch, N. (2012). Econometrics of Financial High-Frequency Data. Springer, Heidelberg.

    Book  MATH  Google Scholar 

  • Koul, H.L. (1985a). Minimum distance estimation in multiple linear regression. Sankhya, Ser. A. 47 Part 1, 57–74.

    MathSciNet  MATH  Google Scholar 

  • Koul, H.L. (1985b). Minimum distance estimation in linear regression with unknown errors. Statist. Prob. Lett. 3, 1–8.

    Article  MathSciNet  MATH  Google Scholar 

  • Koul, H.L. (1986). Minimum distance estimation and goodness-of-fit tests in first-order autoregression. Ann. Statist. 14, 1194–1213.

    Article  MathSciNet  MATH  Google Scholar 

  • Koul, H.L. (1996). Asymptotics of some estimators and sequential residual empiricals in non-linear time series. Ann. Statist. 24, 380–404.

    Article  MathSciNet  MATH  Google Scholar 

  • Koul, H.L. (2002). Weighted Empirical Processes in Dynamic Nonlinear Models, 166, 2nd edn. Springer, USA,.

  • Koul, H.L. (2019). A Uniform Convergence Result for Weighted Residual Empirical Processes RM#719. Department of Statistics and Probability, MSU.

  • Koul, H.L. and Stute, W. (1999). Nonparametric model checks for time series. Ann. Statist. 27, 204–236.

    Article  MathSciNet  MATH  Google Scholar 

  • Koul, H.L. and Perera, I. (2021). A minimum distance lack-of-fit test in a markovian multiplicative error model. J. Stat. Theory Pract. 15, 34.

    Article  MathSciNet  MATH  Google Scholar 

  • Koul, H.L., Perera, I. and Silvapulle, M.J. (2012). Lack-of-fit testing of the conditional mean function in a class of Markov multiplicative error models. Econometric Theory 28, 1283–1312.

    Article  MathSciNet  MATH  Google Scholar 

  • Ljung, G.M. and Box, G.E.P. (1978). On a measure of lack of fit in time series models. Biometrika 65, 297–303.

    Article  MATH  Google Scholar 

  • Lunde, A. (1999). A Generalized Gamma Autoregressive Conditional Duration Model. Discussion paper, Aalborg University.

  • Manganelli, S. (2005). Duration, volume and volatility impact of trades. J. Financ. Mark. 8, 377–399.

    Article  Google Scholar 

  • Meitz, M. and Teräsvirta, T. (2006). Evaluating models of autoregressive conditional duration. J. Bus. Econ. Stat. 24, 104–124.

    Article  MathSciNet  Google Scholar 

  • Pacurar, M. (2008). Autoregressive conditional duration models in finance: a survey of the theoretical and empirical literature. J. of Economic Surveys 22, 711–751.

    Article  Google Scholar 

  • Perera, I., Hidalgo, J. and Silvapulle, M.J. (2016). A goodness-of-fit test for a class of autoregressive conditional duration models. Econ. Rev. 35, 1111–1141.

    Article  MathSciNet  MATH  Google Scholar 

  • Perera, I. and Koul, H.L. (2017). Fitting a two phase threshold multiplicative error model. J. Econometrics 197, 348–367.

    Article  MathSciNet  MATH  Google Scholar 

  • Perera, I. and Silvapulle, M.J. (2021). Bootstrap based probability forecasting in multiplicative error models. J. Econometrics 221, 1–24.

    Article  MathSciNet  MATH  Google Scholar 

  • Stute, W. (1976). On a generalization of the Glivenko-Cantelli theorem. Z. Wahrscheinlichkeitstheorie und Verw. Gebiete 35, 167–175.

    Article  MATH  Google Scholar 

  • Stute, W. (1997). Nonparametric model checks for regression. Ann. Statist. 25, 613–641.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Authors are grateful to the referee and the editor for their useful comments and suggestions on the earlier draft of the paper.

Funding

N. Balakrishna received partial financial suppport from SERB of India under the MATRICS scheme MTR/2018/000195.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Narayana Balakrishna.

Ethics declarations

Conflict of Interest

The authors have no conflicts of interest to declare that are relevant to the content of this article.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A: Main Proofs

Appendix A: Main Proofs

1.1 A.1 Proofs of Theorems 4.1 and 4.2

Proof 1 (Proof of Theorem 4.1).

The proof of (4.4) is given later. Arguing as in the proof of Theorem 5.4.1 of Koul (2002), one verifies that (4.1), (C.3) and (4.4) imply Eq. 4.5 and (4.6)(a).

To prove (4.6)(b), by the positive definiteness of \({\mathcal G}_{1}\) we clearly have \(\widetilde t_{n1}= {\mathcal G}_{1}^{-1}V_{n1}\). Moreover, Vn1 is a vector of the sums of martingale difference arrays satisfying (4.3). By the martingale central limit theorem (CLT), see Hall and Heyde (1980), \(V_{n1}\to _{D} N(0, \sigma ^{2}{\Sigma }_{1}) \) and \(\widetilde t_{n1}={\mathcal G}_{1}^{-1}V_{n1}\to _{D} N(0, \sigma ^{2} {\mathcal G}_{1}^{-1}{\Sigma }_{1} {\mathcal G}_{1}^{-1}). \) This fact together with the Slutsky Theorem, (4.5) and (4.6)(a) imply (4.6)(b).

Proof of (4.4).

Let θnt := θ + n− 1/2t, \(t\in {\mathbb R}^{q}\). Define, for \(z\in {\mathbb R}_{+}^{p}, t\in {\mathbb R}^{q}\),

$$ \begin{array}{@{}rcl@{}} W_{n}(z,t)&:= n^{-1/2}\sum\limits_{i=1}^n \Big[\frac{\psi(Z_{i-1},\theta)}{\psi(Z_{i-1}, \theta_{nt})}- 1 \Big] \varepsilon_{i} I(Z_{i-1}\le z), \end{array} $$
$$ \begin{array}{@{}rcl@{}} S_{n}(z) \!\!\!\!&:=&\!\!\!\! n^{-1}\sum\limits_{i=1}^n \varphi(Z_{i-1}) (\varepsilon_{i} - 1) I(Z_{i-1}\!\le\! z), \quad T_{n11}(t)\! :=\! \int \big(W_{n}(z,t)+t^{\prime}{\Psi}(z)\big)^{2}dL(z), \\ T_{n12}(t) \!\!\!\!&:=&\!\!\!\! \int (W_{n}(z,t)+t^{\prime}{\Psi}(z)) (U_{n}(z) - t^{\prime}{\Psi}(z))dL(z). \end{array} $$

Use the model assumption εi = Yi/ψ(Zi− 1, θ) to obtain

$$ \begin{array}{@{}rcl@{}} U_{n}(z,\theta_{nt})\!\!\!&=&\!\!\!n^{-1/2}\sum\limits_{i=1}^n \Big[\frac{Y_{i}}{\psi(Z_{i-1}, \theta_{nt})}-1\Big]I(Z_{i-1}\le z) \\ \!\!\!&=&\!\!\! n^{-1/2}\sum\limits_{i=1}^n \Big[\frac{Y_{i}}{\psi(Z_{i-1}, \theta_{nt})}- \frac{Y_{i}}{\psi(Z_{i-1}, \theta)} +\frac{Y_{i}}{\psi(Z_{i-1}, \theta)} -1\Big]I(Z_{i-1}\le z) \\ \!\!\!&=&\!\!\! n^{-1/2}\sum\limits_{i=1}^n \Big[\frac{\psi(Z_{i-1},\theta)}{\psi(Z_{i-1}, \theta_{nt})}- 1\Big] \varepsilon_{i} I(Z_{i-1}\le z) + n^{-1/2}\sum\limits_{i=1}^n (\varepsilon_{i}-1)I(Z_{i-1}\le z) \\ \!\!\!&=&\!\!\!W_{n}(z,t)+U_{n}(z), \qquad \forall z\in {\mathbb R}_{+}^{p}, t\in {\mathbb R}. \end{array} $$
(A.1)

Hence,

$$ \begin{array}{@{}rcl@{}} \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!T_{n1}(t)\!\!\!&=&\!\!\!\int {U_{n}^{2}}(z,\theta_{nt}) dL(z) = \int \Big(W_{n}(z,t)+t^{\prime}{\Psi}(z)+U_{n}(z)-t^{\prime}{\Psi}(z)\Big)^{2} dL(z) \\ \!\!\!&=&\!\!\!T_{n11}(t)+2 T_{n12}(t) + Q_{n1}(t). \end{array} $$
(A.2)

We shall shortly prove the following facts. For every \(0<b<\infty \),

$$ \begin{array}{@{}rcl@{}} \text{(a)} \quad \sup_{\|t\|\le b} T_{n11}(t)=o_p(1), \qquad \text{(b)} \quad E\Big(\sup_{\|t\|\le b} Q_{n1}(t)\Big)=O(1). \end{array} $$
(A.3)

Then by the Cauchy-Schwarz inequality, \( \sup _{\|t\|\le b}\big |T_{n12}(t)\big |^{2}\le \sup _{\|t\|\le b} T_{n11}\) \((t) \sup _{\|t\|\le b} Q_{n1}(t)\) = op(1). Hence the claim Eq. 4.4.

Proof of (A.3)(a).

Rewrite

$$ \begin{array}{@{}rcl@{}} W_{n}(z,t)\!\!\!&:=&\!\!\! n^{-1/2}\sum\limits_{i=1}^n \Big[\frac{\psi(Z_{i-1},\theta)}{\psi(Z_{i-1}, \theta_{nt})}-1 \Big] \varepsilon_{i} I(Z_{i-1}\le z) \\ \!\!\!&=&\!\!\! - n^{-1/2}\sum\limits_{i=1}^n \frac{1}{\psi(Z_{i-1}, \theta_{nt})} \Big[\psi(Z_{i-1}, \theta_{nt}) - \psi(Z_{i-1},\theta)-n^{-1/2}t^{\prime} \dot \psi(Z_{i-1},\theta)\Big] \\&&\!\!\!\varepsilon_{i} I(Z_{i-1}\le z) \\ && \!\!\!- t^{\prime} n^{-1}\sum\limits_{i=1}^n \Big[\frac{1}{\psi(Z_{i-1}, \theta_{nt})} - \frac{1}{\psi(Z_{i-1},\theta)} \Big] \dot \psi(Z_{i-1},\theta) \varepsilon_{i} I(Z_{i-1}\le z)\\ && \!\!\!- t^{\prime}n^{-1}\sum\limits_{i=1}^n \varphi(Z_{i-1})(\varepsilon_{i}-1) I(Z_{i-1}\le z) \\ && \!\!\!- t^{\prime}n^{-1}\sum\limits_{i=1}^n \big[ \varphi(Z_{i-1}) I(Z_{i-1}\le z) - E\big(\varphi(Z_{0})I(Z_{0}\le z)\big)\big] - t^{\prime}{\Psi}(z). \end{array} $$

Let dit := ψ(Zi− 1, θnt) − ψ(Zi− 1, θ) and \( \delta _{it}:= d_{it} -n^{-1/2}t^{\prime } \dot \psi (Z_{i-1}, \theta ). \) Then the above identity is equivalent to

$$ \begin{array}{@{}rcl@{}} &&\!\!\!\!\!\!\!\!\!\!\!\!W_{n}(z,t) + t^{\prime} {\Psi}(z)\\ &&\!\!\!\!\!\!\!\!\!\!\!\!= - n^{-1/2}\sum\limits_{i=1}^n \frac{\delta_{it}}{\psi(Z_{i-1}, \theta_{nt})} \varepsilon_{i} I(Z_{i-1}\le z) + t^{\prime} n^{-1}\sum\limits_{i=1}^n \frac{d_{it}}{\psi(Z_{i-1}, \theta_{nt})} \varphi(Z_{i-1}) \varepsilon_{i} I(Z_{i-1}\le z) \\ && \!\!\!\!\!\!-t^{\prime}S_{n}(z)-t^{\prime}\big({\Psi}_{n}(z)-{\Psi}(z)\big) \\ &&\!\!\!\!\!\!\!\!\!\!\!\!= A_{1}(z,t)+ A_{2}(z,t)-t^{\prime}S_{n}(z)-t^{\prime}\big({\Psi}_{n}(z)-{\Psi}(z)\big), \qquad \text{say}. \end{array} $$
(A.4)

In the sequel, the range of the vectors z, 𝜗 in \(\inf _{z,\vartheta }\) is over \({\mathbb R}_{+}^{p} \times {\Theta }\), unless mentioned otherwise. By (C.1), \(\inf _{z, \vartheta }\psi (z,\vartheta )\ge C>0\) and for every \(b<\infty \), \(\sup _{1\le i\le n, \|t\|\le b}n^{1/2}\big |\delta _{it}\big |=o_{p}(1)\). Moreover, because E(ε0) = 1, we have \(n^{-1}{\sum }_{i=1}^n \varepsilon _{i}=O_{p}(1)\). Hence

$$ \begin{array}{@{}rcl@{}} \sup_{z\in {\mathbb R}_+^p, \|t\|\le b}|A_1(z,t)|\le C^{-1}\sup_{1\le i\le n, \|t\|\le b}n^{1/2}\big|\delta_{it}\big|n^{-1}\sum\limits_{i=1}^n \varepsilon_i=o_p(1). \end{array} $$
(A.5)

Next, recall that the stationarity of the process \(Y_{i}, i\in {\mathbb Z}\) and \(E\|\dot \psi (Z_{0},\theta )\|^{2}\) \(<\infty \) imply that \(n^{-1/2} \max \limits _{1\le i\le n}\|\dot \psi (Z_{i-1},\theta )\| =o_{p}(1)\), \(E\big (n^{-1}{\sum }_{i=1}^{n} \|\varphi \) \((Z_{i-1}) \| \varepsilon _{i}\big )= E\|\varphi (Z_{0})\|<\infty \), and by the Markov inequality, \(n^{-1}{\sum }_{i=1}^{n} \|\varphi \) (Zi− 1)∥εi = Op(1). Hence, (C.1) and (C.2) imply that

$$ \begin{array}{@{}rcl@{}} &&D_{n} := \sup_{1\le i\le n, \|t\|\le b} |d_{it}| \le \sup_{1\le i\le n, \|t\|\le b} \big|\delta_{it}\big| + b n^{-1/2} \max_{1\le i\le n}\|\dot \psi(Z_{i-1},\theta)\| =o_{p}(1),\\ &&\sup_{z\in {\mathbb R}_{+}^{p}, \|t\|\le b}|A_{2}(z,t)| \le b C^{-1} D_{n} n^{-1}\sum\limits_{i=1}^n \|\varphi(Z_{i-1})\| \varepsilon_{i} =o_{p}(1). \end{array} $$
(A.6)

Next, consider Sn(z). Observe that Sn(z) is a vector of weighted empirical processes with the summands of each component being stationary and ergodic and ESn(z) ≡ 0. Using a Glivenko-Cantelli Lemma type argument one obtains that \(\sup _{z\in {\mathbb R}_{+}^{p}}\|S_{n}(z)\|=o_{p}(1)\). For details see Stute (1976) and Koul (2019). Similarly, \(\sup _{z\in {\mathbb R}_{+}^{p}}\big \|{\Psi }_{n}(z)-{\Psi }(z)\big \|^{2}=o_{p}(1)\). Upon combining these two facts with (A.4), (A.5) and (A.6) we obtain that for every \(0<b<\infty \),

$$ \begin{array}{@{}rcl@{}} \sup_{z\in {\mathbb R}_+^p, \|t\|\le b}\big|W_n(z,t) + t^{\prime} {\Psi}(z)\big|=o_p(1). \end{array} $$
(A.7)

This fact combined with the definition of Tn11 and L being a d.f. readily yields (A.3)(a).

Next, to prove (A.3)(b), note that

$$ \begin{array}{@{}rcl@{}} Q_{n1}(t)&:=& \int \Big(U_{n}(z,\theta)-t^{\prime}{\Psi}(z)\Big)^{2} dL(z)\le 2 T_{n1}(0) + 2 t^{\prime}{\mathcal G}_{1} t, \\ E\Big(\sup_{\|t\|\le b}Q_{n1}(t)\Big)&\le& 2ET_{n1}(0)+ 2 b^{2} \int \|{\Psi}(z)\|^{2} dL(z)\\ &=& 2 \int G(z)dL(z)\big[1+ b^{2} E\|\varphi(Z_{0})\|^{2}\big]<\infty. \end{array} $$

This also completes the proof of Theorem 4.1. □

Proof 2 (Proof of Theorem 4.2).

The proof is similar to that of Theorem 4.1, with some differences. Let

$$ \begin{array}{@{}rcl@{}} && \tilde V_{n2}:= \int U_n(z){\Psi}(z)dG_n(z), \qquad {\mathcal G}_{n2}:=\int {\Psi}(z){\Psi}(z)' dG_n(z), \\ && \tilde Q_{n2}(t):= \int \big(U_n(z)-t^{\prime}{\Psi}(z)\big)^2 dG_n(z)= M_{n2}(\theta) -2t^{\prime}\tilde V_{n2} +t^{\prime} {\mathcal G}_{n2} t, \\ &&\tilde T_{n2}(t):=M_{n2}(\theta+n^{-1/2}t), \qquad \tilde T_{n21}(t) \!:=\! \int \big(W_n(z,t)+t^{\prime}{\Psi}(z)\big)^2dG_n(z), \\ && \tilde T_{n12}(t) := \int (W_n(z,t)+t^{\prime}{\Psi}(z)) (U_n(z) - t^{\prime}{\Psi}(z))dG_n(z). \end{array} $$

Then akin to (A.2),

$$ \begin{array}{@{}rcl@{}} \tilde T_{n2}(t)=\int {U_{n}^{2}}(z,\theta_{nt}) dG_{n}(z) =\tilde T_{n21}(t)+2 \tilde T_{n22}(t) + \tilde Q_{n2}(t). \end{array} $$

Now, \( \sup _{\|t\|\le b} \tilde T_{n21}(t)\le \sup _{z\in {\mathbb R}_{+}^{p}, \|t\|\le b}\big (W_{n}(z,t)+t^{\prime }{\Psi }(z)\big )^{2}=o_{p}(1), \) by (A.7). We shall shortly prove that for every \(0<b<\infty \),

$$ \begin{array}{@{}rcl@{}} \sup_{\|t\|\le b} \big|\tilde Q_{n2}(t) - Q_{n2}(t)\big|=o_p(1). \end{array} $$
(A.8)

Argue as for (A.3)(b) to conclude that \(E\big (\sup _{\|t\|\le b} Q_{n2}(t)\big )=O(1). \) Hence \(\sup _{\|t\|\le b} \tilde Q_{n2}(t)=O_{p}(1)\) and \( \sup _{\|t\|\le b}\big |\tilde T_{n22}(t)\big |^{2}\le \sup _{\|t\|\le b} \tilde T_{n21}(t) \sup _{\|t\|\le b}\) \(\tilde Q_{n2}(t) \) = op(1). These facts together yield the claim Eq. 4.7.

Next, to prove (A.8), note that the left hand side of (A.8) is bounded from the above by

$$ \begin{array}{@{}rcl@{}} \big|\int {U_{n}^{2}}\big[dG_{n} -dG\big]\big| +b \big\|\int U_{n} {\Psi}\big[dG_{n}- dG\big]\big\|+b^{2} \big\|\int {\Psi} {\Psi}^{\prime}\big[dG_{n}-dG\big] \big\|. \end{array} $$

By the Ergodic Theorem, \(\big \|{\int \limits } {\Psi } {\Psi }^{\prime }\big [dG_{n}-dG\big ]\big \|\to 0\), a.s. while by Lemmas A.1 and A.2 below, the first two terms tend to zero in probability. This completes the proof of (4.7). The proofs of the other two claims of this theorem are similar to those of (4.5) and (4.6)(a), (b) of Theorem 4.1. □

1.2 A.2 Proof of Lemma 4.1

Proof 3 (Proof of the claim about (C.3)).

Let h be a positive function on \({\mathbb R}_{+}^{p}\) with \(0<{\int \limits } h^{2} dL<\infty \) and let \({\mathbf {\infty }}\) (0) denote the vector of p infinities (zeros). Let \(\varphi (z):= {\int \limits }_{x\le z} h(x) dL(x)\). Note that φ is nondecreasing in each coordinate and \(\varphi ({\mathbf {\infty }})<\infty \). Moreover, \(\gamma (z):={\int \limits } I(z\le x) h(x) dL(x)={\int \limits } I(z\le x) d\varphi (x)\ge 0\), for all \(z\in {\mathbb R}_{+}^{p}\). Define

$$ \begin{array}{@{}rcl@{}} {\mathcal V}_{n1}(t)&:=& \int U_{n}(z,\theta_{nt}) h(z)dL(z)= \int U_{n}(z,\theta_{nt}) d\varphi(z) \end{array} $$
$$ \begin{array}{@{}rcl@{}} &=&n^{-1/2}\sum\limits_{i=1}^n \Big[\frac{Y_{i}}{\psi(Z_{i-1}, \theta+n^{-1/2}t)} -1\Big]\int I(Z_{i-1}\le z)d \varphi(z) \\ &=&n^{-1/2} \sum\limits_{i=1}^n \Big[\frac{Y_{i}}{\psi(Z_{i-1}, \theta+n^{-1/2}t)} -1\Big] \gamma(Z_{i-1}). \end{array} $$

By the Cauchy-Schwarz inequality,

$$ \begin{array}{@{}rcl@{}} M_{n1}(\theta+n^{-1/2}t)\ge \Big(\int U_n(z,\theta_{nt}) h(z)dL(z)\Big)^2 \Big /\int h^2 dL, \quad \forall t\in {\mathbb R}^q. \end{array} $$

Because γ(z) ≥ 0 for all z, by (4.8), \(\forall e\in R^{q}, \|e\|=1, {\mathcal V}_{n1}(r e)\) is monotonic in r. Hence

$$ \begin{array}{@{}rcl@{}} \inf_{\|t\|>b}M_{n1}(\theta + n^{1/2}t) = \inf_{e \in {\mathbb R}^{q}, \|e\|=1, |r|>b}M_{n}(\theta + n^{-1/2}re) \!\ge\! \inf_{e \in {\mathbb R}^{q}, \|e\|=1, |r|=b}{\mathcal V}_{n1}^{2}(re)\Big / \int h^{2} dL. \end{array} $$

Let

$$ \begin{array}{@{}rcl@{}} \widehat {\mathcal V}_{n1}(t) \!\!\!&:=&\!\!\! \int \big(U_{n}(z)-t^{\prime}{\Psi}(z)\big) h(z)dL(z)= \int U_{n}(z) d\varphi(z) - t^{\prime} \int {\Psi}(z)d\varphi(z) \\ \!\!\!&=&\!\!\! {\mathcal V}_{n1}(0) - t^{\prime} \int {\Psi}(z)d\varphi(z). \end{array} $$

In view of (A.1),

$$ {\mathcal V}_{n1}(t) - \widehat {\mathcal V}_{n1}(t)= \int \big\{U_{n}(z,\theta_{nt}) - U_{n}(z)+t^{\prime}{\Psi}(z)\big\} d\varphi(z) = \int \big\{W_{n}(z,\theta_{nt})+t^{\prime}{\Psi}(z)\big\}d\varphi(z) . $$

Because \(\varphi ({\mathbf {\infty }})={\int \limits } h dL<\infty \), by (A.7),

$$ \begin{array}{@{}rcl@{}} \sup_{\|t\|\le b} \Big|{\mathcal V}_{n1}(t) - \widehat {\mathcal V}_{n1}(t)\Big|\le \sup_{\|t\|\le b} \big|W_{n}(z,\theta_{nt})+t^{\prime}{\Psi}(z)\big| \varphi({\mathbf{\infty}}) =o_{p}(1). \end{array} $$

Therefore,

$$ \begin{array}{@{}rcl@{}} \Big|\inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b}{\mathcal V}_{n1}(re) - \inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b} \widehat {\mathcal V}_{n1}(re)\Big|=o_p(1). \end{array} $$
(A.9)

Let \(\tau _{2}:={\int \limits } h^{2} dL\). Fix an 𝜖 > 0 and \(0<\eta <\infty \). By (A.9), there exists N𝜖, η such that

$$ \begin{array}{@{}rcl@{}} && P\Big(\inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b}\Big|{\mathcal V}_{n1}(re)\Big| \ge (\tau_2 \eta)^{1/2} \Big) \\ && \hskip .5in \ge P\Big(\inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b}\Big|\widehat {\mathcal V}_{n1}(re)\Big| \ge (\tau_2 \eta)^{1/2} \Big ) - \frac{\epsilon}{2}, \qquad \forall n>N_{\epsilon,\eta}. \end{array} $$

Moreover, by the Cauchy-Schwarz inequality, \( E{\mathcal V}_{n1}^{2}(0)=\sigma ^{2} E\gamma ^{2}(Z_{0}) \le \sigma ^{2}{\int \limits } h^{2} dL<\infty . \) Hence by the Markov inequality, for every 𝜖 > 0 there is a b𝜖 such that

$$ \begin{array}{@{}rcl@{}} P\big(|{\mathcal V}_{n1}(0)|\le b_{\epsilon}\big)\ge 1-(\epsilon/2), \quad \forall n\ge 1. \end{array} $$
(A.10)

Recall the elementary fact that for any real numbers a, b, ||a|−|b||≤|a ± b|. Also let \(K:= {\int \limits } \|{\Gamma }\| d \varphi \). Let b𝜖 be as in (A.10). Choose b ≥ (b𝜖 + (v2η)1/2)K− 1. Then, ∀nN𝜖, η,

$$ \begin{array}{@{}rcl@{}} P\Big(\inf_{|t|> b} M_{n1}(\theta+n^{-1/2}t)\ge \eta \Big ) &\ge& P\Big(\inf_{e \in {\mathbb R}^{q}, \|e\|=1, |r|=b}\Big|{\mathcal V}_{n1}(re)\Big| \ge (\tau_{2} \eta)^{1/2} \Big )\\ &=&P\Big(\Big|{\mathcal V}_{n1}(\pm be)\Big| \ge (\tau_{2} \eta)^{1/2}, \forall \|e\|=1 \Big ) \\ &\ge& P\Big(\Big|\widehat {\mathcal V}_{n1}(\pm be)\Big| \ge (\tau_{2} \eta)^{1/2}, \forall \|e\|=1\Big ) - \epsilon/2 \\ &\ge &P\Big(\ \Big | \big|{\mathcal V}_{n1}(0 )|- bK\big |\ \Big| \ge (\tau_{2}\eta)^{1/2} \Big ) - \epsilon/2 \\ &\ge &P\Big(\ |{\mathcal V}_{n1}(0)|\le bK - (\tau_{2}\eta)^{1/2} \Big ) - \epsilon/2 \\ &\ge &P\Big(\ |{\mathcal V}_{n1}(0)|\le b_{\epsilon} \Big ) - \epsilon/2 \ge 1-\epsilon, \end{array} $$
(A.11)

thereby proving the claim about (C.3).

Proof of the claim about (C.4)

The proof of this claim is similar to that of the previous claim, but with some differences. In particular we need the following two preliminary lemmas.

Lemma A.1.

Under the above set up, Un converges weakly to a continuous Gaussian process \({\mathcal Z}(x), x\in {\mathbb R}_{+}^{p}\) in Skorokhod space \(D([0,\infty ]^{p})\) and uniform metric.

The proof of this lemma is similar to that of the main theorem in Stute (1976).

Lemma A.2.

Let \({\mathcal U}\) be a relatively compact subset of \(D[0,\infty ]^{p}\). Let μn, μ be a sequence of possibly random multivariate distribution functions on \([0,\infty )^{p}\) such that \(\sup _{x\in {\mathbb R}_{+}^{p}} \big |\mu _{n}(x)-\mu (x)\big |\to 0\), a.s. Then \( \sup _{y\in {\mathbb R}_{+}^{p}, \alpha \in {\mathcal U}}\Big | {\int \limits }_{x\le y} \alpha (x)\) [dμn(x) − dμ(x)]| →p0.

The proof of this lemma is similar to that of Lemma 3.1 of Chang (1990) and Lemma 4.2 of Koul and Stute (1999). Details are left out for the sake of brevity.

Now, let h be a positive function on \({\mathbb R}_{+}^{p}\) with \(0<{\int \limits } h^{2} dG<\infty \) and \({\int \limits } h dG=1\). Let

$$ \begin{array}{@{}rcl@{}} \beta_{nk}(z)&:=&{\int}_{x\le z} h^{k}(x) dG_{n}(x) =n^{-1}\sum\limits_{j=1}^n h^{k}(Z_{j-1})I(Z_{j-1}\le z), \\ \beta_{k}(z)&:=&{\int}_{x\le z} h^{k}(x) dG(x), \quad k=1, 2, \ z\in {\mathbb R}_{+}^{p}, \qquad B_{n2}:=\beta_{n2}({\mathbf{\infty}}), \quad B_{2}:=\beta_{2}({\mathbf{\infty}}). \end{array} $$

Note that βnk(0) = 0 = βk(0), \(0< \beta _{k}(z)\le \beta _{k}(\mathbf {\infty })={\int \limits } h^{k} dG<\infty , \) for all \(z\in {\mathbb R}_{+}^{p}\) and Eβnk(z) ≡ βk(z), k = 1, 2. By the Ergodic Theorem and a Glivenko-Cantelli type argument, see Stute (1976) and Koul (2019),

$$ \begin{array}{@{}rcl@{}} \sup_{z\in[0,\infty]^p} \big|\beta_{nk}(z)-\beta_k(z)\big|\to 0, \quad \text{a.s.}, k=1, 2; \quad |B_{n2}-B_2|\to 0, \text{ a.s.} \end{array} $$
(A.12)

Thus for all sufficiently large n, βnk(z) > 0, for all \(z\in (0,\infty ]^{p}, k=1, 2\), Bn2 > 0 (a.s.). The arguments below are carried out on the event \(B_{n2}={\int \limits } h^{2} dG_{n}>0\).

Recall θnt = θ + n− 1/2t. By the Cauchy-Schwarz inequality,

$$ \begin{array}{@{}rcl@{}} M_{n2}(\theta+n^{-1/2}t)\ge \Big(\int U_n(z,\theta_{nt}) h(z)dG_n(z)\Big)^2 \Big / B_{n2}, \quad \forall t\in {\mathbb R}^q. \end{array} $$

Let \(\alpha (z):= {\int \limits }_{x\ge z} d\beta _{1}(x)\), \(\alpha _{n}(z):= {\int \limits }_{x\ge z} d\beta _{n1}(x)\), \(z\in {\mathbb R}_{+}^{p}\) and

$$ \begin{array}{@{}rcl@{}} {\mathcal V}_{n2}(t)&\!\!\!:=&\!\!\!\! \int U_{n}(z,\theta_{nt}) h(z)dG_{n}(z) =n^{-1/2}\sum\limits_{i=1}^n \Big[\frac{Y_{i}}{\psi(Z_{i-1}, \theta_{nt})} -1\Big]\int I(Z_{i-1}\le z)d\beta_{n1}(z)\\ &=&\!\!\!n^{-1/2} \sum\limits_{i=1}^n \Big[\frac{Y_{i}}{\psi(Z_{i-1}, \theta+n^{-1/2}t)} -1\Big] \alpha_{n}(Z_{i-1}). \end{array} $$

Because αn(z) ≥ 0 for all \(z\in {\mathbb R}_{+}^{p}, n\ge 1\), w.p.1, by (4.8), \({\mathcal V}_{n2}(r e)\) is monotonic in r for all eRq, ∥e∥ = 1. Hence

$$ \begin{array}{@{}rcl@{}} \inf_{\|t\|>b}M_{n2}(\theta+n^{1/2}t) = \inf_{e \in {\mathbb R}^{q}, \|e\|=1, |r|>b}M_{n2}(\theta+n^{-1/2}re)\ge \inf_{e \in {\mathbb R}^{q}, \|e\|=1, |r|=b}{\mathcal V}_{n2}^{2}(re)\Big / B_{n2}. \end{array} $$

Let

$$ \begin{array}{@{}rcl@{}} \widehat {\mathcal V}_{n2}(t) &:=& \int \big(U_{n}(z)-t^{\prime}{\Psi}(z)\big) h(z)dG_{n}(z)= \int U_{n}(z) d\beta_{n1}(z) - t^{\prime} \int {\Psi}(z)d\beta_{n1}(z) \\ &=& {\mathcal V}_{n2}(0) - t^{\prime} \int {\Psi}(z)d\beta_{n1}(z), \\ {\mathcal V}_{n}^{*} &:=& \int U_{n}(z) d\beta_{1}(z)=n^{-1/2} \sum\limits_{i=1}^n (\varepsilon_{i}-1)\alpha(Z_{i-1}). \end{array} $$

By Lemma A.1, the process \(U_{n}(z), z\in [0,\infty ]^{p}\) is relatively compact with respect to the uniform metric. Hence (A.12) and Lemma A.2 applied with \(\mu _{n}=\beta _{n1}/\beta _{n1}({\mathbf {\infty }}), \mu =\beta _{1}\) yield

$$ \begin{array}{@{}rcl@{}} \big|{\mathcal V}_{n2}(0)-{\mathcal V}_n^{*}\big|=\Big|\int U_n(z)[ d\beta_{n1}(z)-d\beta_1(z)]\Big|=o_p(1). \end{array} $$

Here we have used the fact \(\beta _{1}({\mathbf {\infty }})={\int \limits } h dG=1\) and \(\beta _{n1}({\mathbf {\infty }})\to _{p} \beta _{1}({\mathbf {\infty }})=1\).

Let \(\bar {\mathcal V}_{n2}(t):={\mathcal V}_{n}^{*} - t^{\prime }{\int \limits } {\Psi } d\beta _{1}\). Because by the Ergodic Theorem, \({\int \limits } {\Psi } d\beta _{n1}\to _{p} {\int \limits } {\Psi } d\beta _{1}\) and

$$ \begin{array}{@{}rcl@{}} \sup_{\|t\|\le b}\big|\widehat {\mathcal V}_{n2}(t) - \bar {\mathcal V}_{n2}(t)\big|=o_p(1), \end{array} $$
(A.13)

by (A.1), we obtain

$$ {\mathcal V}_{n2}(t) - \widehat {\mathcal V}_{n2}(t)= \int \big\{U_{n}(z,\theta_{nt}) - U_{n}(z)+t^{\prime}{\Psi}(z)\big\} d\beta_{n1}(z)\\ = \int \big[W_{n}(z,\theta_{nt})+t^{\prime}{\Psi}(z)\big] d\beta_{n1}(z) . $$

Therefore, by (A.7), (A.12) and (A.13)

$$ \begin{array}{@{}rcl@{}} \sup_{\|t\|\le b} \Big|{\mathcal V}_{n2}(t) - \bar {\mathcal V}_{n2}(t)\Big| \!\!\!&\le&\!\!\! \sup_{\|t\|\le b} \Big|{\mathcal V}_{n2}(t) - \widehat {\mathcal V}_{n2}(t)\Big| + \sup_{\|t\|\le b} \Big|\widehat {\mathcal V}_{n2}(t) - \bar {\mathcal V}_{n2}(t)\Big|\\ &\le &\!\!\! \sup_{\|t\|\le b} \big|W_n(z,\theta_{nt})+t^{\prime}{\Psi}(z)\big| \beta_{n1}({\mathbf{\infty}})+o_p(1)=o_p(1). \end{array} $$

Also note that \(\sup _{\|t\|\le b} |\bar {\mathcal V}_{n2}(t)|=O_{p}(1)= \sup _{\|t\|\le b} | {\mathcal V}_{n2}(t)|\) and by (A.12),

$$ \begin{array}{@{}rcl@{}} \sup_{\|t\|\le b} \Big|\frac{{\mathcal V}_{n2}(t)}{B_{n2}^{1/2}} - \frac{\bar {\mathcal V}_{n2}(t)}{B_{2}^{1/2}}\Big| \le \sup_{\|t\|\le b}\big|{\mathcal V}_{n2}(t)\big| \Big| \frac{1}{B_{n2}^{1/2}} - \frac{1}{B_{2}^{1/2}}\Big| + \frac{1}{B_{2}^{1/2}} \sup_{\|t\|\le b} \big|{\mathcal V}_{n2}(t) - \bar {\mathcal V}_{n2}(t)\big| =o_{p}(1). \end{array} $$

This fact in turn implies that

$$ \begin{array}{@{}rcl@{}} \Big|\inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b}\frac{{\mathcal V}_{n2}(re)}{B_{n2}^{1/2}} - \inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b} \frac{\bar {\mathcal V}_{n2}(re)}{B_2^{1/2}}\Big|=o_p(1). \end{array} $$
(A.14)

Fix an 𝜖 > 0 and η > 0. By (A.14), there exists N𝜖, η such that

$$ \begin{array}{@{}rcl@{}} && P\Big(\inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b}\frac{\big|{\mathcal V}_{n2}(re)\big|}{B_{n2}^{1/2}} \ge \eta^{1/2} \Big) \\ & \ge& P\Big(\inf_{e \in {\mathbb R}^q, \|e\|=1, |r|=b}\big|\bar {\mathcal V}_{n2}(re)\big| \ge (B_2 \eta)^{1/2} \Big ) - \frac{\epsilon}{2}, \qquad \forall n>N_{\epsilon,\eta}. \end{array} $$

Moreover, \(E\big ({\mathcal V}_{n}^{*}\big )^{2}= \sigma ^{2} E\alpha ^{2}(Z_{0})\le \sigma ^{2} B_{2}<\infty \). Hence, by the Markov inequality, for any 𝜖 > 0 there exists b𝜖 such that

$$ \begin{array}{@{}rcl@{}} P\big(|{\mathcal V}_n^{*}|\le b_{\epsilon}\big)\ge 1-(\epsilon/2), \qquad \forall n\ge 1. \end{array} $$
(A.15)

Let \(K:= {\int \limits } \|{\Gamma }\| d\beta \) and b𝜖 be as in (A.15). Choose b ≥ (b𝜖 + (B2η)1/2)K− 1 and argue as for Eq. A.11 to obtain that ∀nN𝜖, η,

$$ \begin{array}{@{}rcl@{}} P\Big(\inf_{|t|> b} M_{n2}(\theta_0+n^{-1/2}t)\ge \eta \Big ) &\ge& P\Big(\ |{\mathcal V}_n^{*}|\le bK - (B_2\eta)^{1/2} \Big ) - \epsilon/2 \\ &\ge& P\Big(\ |{\mathcal V}_{n1}^{*}|\le b_{\epsilon} \Big ) - \epsilon/2 \ge 1-\epsilon, \end{array} $$

thereby proving the claim about (C.4). This also completes the proof of Lemma 4.1. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Koul, H.L., Perera, I. & Balakrishna, N. A class of Minimum Distance Estimators in Markovian Multiplicative Error Models. Sankhya B 85 (Suppl 1), 87–115 (2023). https://doi.org/10.1007/s13571-021-00274-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13571-021-00274-x

Keywords

PACS Nos

Navigation