Skip to main content
Log in

Empirical likelihood for change point detection in autoregressive models

  • Research Article
  • Published:
Journal of the Korean Statistical Society Aims and scope Submit manuscript

Abstract

Change point analysis has become an important research topic in many fields of applications. Several research work have been carried out to detect changes and its locations in time series data. In this paper, a nonparametric method based on the empirical likelihood is proposed to detect structural changes in the parameters of autoregressive (AR) models . Under certain conditions, the asymptotic null distribution of the empirical likelihood ratio test statistic is proved to be Gumbel type. Further, the consistency of the test statistic is verified. Simulations are carried out to show that the power of the proposed test statistic is significant. The proposed method is applied to monthly average soybean sales data to further illustrate the testing procedure.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  • Akashi, F., Dette, H., & Liu, Y. (2016). Change point detection in autoregressive models with no moment assmptions. arXiv:1612.01520v1.

  • Aue, A., & Horvárth, L. (2012). Structural breaks in time series. Journal of Time Series Analysis, 34, 1–16.

    Article  MathSciNet  Google Scholar 

  • Bai, J. S., & Perron, P. (1998). Estimating and testing linear models with multiple structural changes. Econometrica, 66, 47–78.

    Article  MathSciNet  Google Scholar 

  • Balcombe, K., Bailey, A., & Brooks, J. (2007). Threshold effects in price transmissions: the case of Brazilian wheat, maize and soya prices. American Journal of Agricultural Economics, 89, 308–323.

    Article  Google Scholar 

  • Baragona, R., Battaglia, F., & Cucina, D. (2013). Empirical likelihood for break detection in time series. Electronic Journal of Statistics, 7, 3089–3123.

    Article  MathSciNet  Google Scholar 

  • Berkes, I., Horváth, L., Ling, S., & Schauer, J. (2011). Testing for structural change of AR model to threshold AR model. Journal of Time Series Analysis, 32(5), 547–565.

    Article  Google Scholar 

  • Chen, J., & Gupta, A. K. (2000). Parametric Statistical Change Point Analysis. Boston: Birkhäuser.

    Book  Google Scholar 

  • Chernoff, H., & Zacks, S. (1964). Estimating the current mean of a normal distribution which is subjected to changes in time. The Annals of Mathematical Statistics, 35, 999–1018.

    Article  MathSciNet  Google Scholar 

  • Csörgo, M., & Horváth, L. (1997). Limit Theorems in Change-Point Analysis. New York: Wiley & Sons.

    MATH  Google Scholar 

  • Davis, R. A., Huang, D., & Yao, Y. C. (1996). Testing for a change in the parameter values and order of an autoregressive model. The Annals of Statistics, 23, 282–304.

    Article  MathSciNet  Google Scholar 

  • Davis, R. A., Lee, T. C. M., & Rodriguez-Yam, G. A. (2006). Structual break estimation for nonstationary time series models. Journal of the American Statistical Association, 101, 223–239.

    Article  MathSciNet  Google Scholar 

  • Gombay, E., & Horváth, L. (1994). An Application of the maximum likelihood test to the change point problem. Stochastic Processes and Their Applications, 50(1), 161–171.

    Article  MathSciNet  Google Scholar 

  • Hawkins, D. M. (1977). Testing a sequence of observations for a shift in location. Journal of the American Statistical Association, 72, 180–186.

    Article  MathSciNet  Google Scholar 

  • Kitamura, Y. (1997). Empirical likelihood methods with weakly depedent processes. Annals of Statistics, 25(5), 2084–2102.

    Article  MathSciNet  Google Scholar 

  • Kolaczyk, E. D. (1994). Empirical likelihood for generalized linear models. Statistica Sinica, 4, 199–218.

    MathSciNet  MATH  Google Scholar 

  • Lee, S., Ha, J., Na, O., & Na, S. (2003). The Cusum test for parameter change in time series models. Scandinavian Journal of Statistics, 30, 781–796.

    Article  MathSciNet  Google Scholar 

  • Liu, Y., Zou, C., & Zhang, R. (2008). Empirical likelihood ratio test for a change-point in linear regression model. Communication in Statistics-Theory and Methods, 37, 1–13.

    MathSciNet  MATH  Google Scholar 

  • Ning, W. (2012). Empirical likelihood ratio test for a mean change point model with a linear trend followed by an abrupt change. Journal of Applied Statistics, 39(5), 947–961.

    Article  MathSciNet  Google Scholar 

  • Nordman, D. J., & Lahiri, S. N. (2014). A review of empirical likelihood methods for time series. Journal of Statistical Planning and Inference, 155, 1–18.

    Article  MathSciNet  Google Scholar 

  • Ogata, H. (2005). Empirical Llikelihood for non Gaussian stationary processes. Scientiae Mathematicae Japonicae Online, e–2005, 465–473.

    Google Scholar 

  • Owen, A. B. (1988). Empirical likelihood ratio confidence Intervals for a single functional. Biometrika, 75, 237–49.

    Article  MathSciNet  Google Scholar 

  • Owen, A. B. (1990). Empirical likelihood ratio confidence regions. The Annals of Statistics, 18, 90–120.

    Article  MathSciNet  Google Scholar 

  • Owen, A. B. (1991). Empirical likelihood for linear models. The Annals of Statistics, 19, 1725–47.

    Article  MathSciNet  Google Scholar 

  • Owen, A. B. (2001). Empirical Likelihood. New York: Chapman & Hall/CRC.

    Book  Google Scholar 

  • Page, E. S. (1954). Continuous inspection schemes. Biometrika, 41, 100–116.

    Article  MathSciNet  Google Scholar 

  • Page, E. S. (1955). A test for a change in a parameter occurring at an unknown point. Biometrika, 42, 523–527.

    Article  MathSciNet  Google Scholar 

  • Perron, P., & Vogelsang, T. J. (1992). Testing for a unit root in a time series with a changing mean: corrections and extensions. Journal of Business & Economic Statistics, 10(4), 467–470.

    Google Scholar 

  • Qin, J., & Lawless, J. (1994). Empirical likelihood and general estimating equations. The Annals of Statistics, 22, 300–325.

    Article  MathSciNet  Google Scholar 

  • Robbins, M., Gallagher, C., Lund, R., & Aue, A. (2011). Mean shift testing in correlated data. Journal of Time Series Analysis, 32, 498–511.

    Article  MathSciNet  Google Scholar 

  • Sen, A. K., & Srivastava, M. S. (1975). On tests for detecting changes in mean. The Annals of Statistics, 3, 98–108.

    Article  MathSciNet  Google Scholar 

  • Vostrikova, L. J. (1981). Detecting disorder in multidimenstional random processes. Soviet Mathematics Doklady, 24, 55–59.

    MATH  Google Scholar 

  • Wang, L., Aue, A., & Paul, D. (2017). Spectral analysis of sample autocovariance matrices of a class of linear time series in moderately high dimensions. Bernoulli, 23, 2181–2209.

    Article  MathSciNet  Google Scholar 

  • Worsley, K. J. (1986). Confidence regions and test for a change point in a sequence of exponential family random variables. Biometrika, 73, 91–104.

    Article  MathSciNet  Google Scholar 

  • Yu, H. (2007). High moment partial sum processes of residuals in ARMA models and their applications. Journal of Time Series Analysis, 28, 72–91.

    Article  MathSciNet  Google Scholar 

  • Zou, C. L., Liu, Y. K., Qin, P., & Wang, Z. (2007). Empirical likelihood ratio test for the change-point problem. Statistics & Probability Letters, 77, 374–382.

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors would like to thank two anonymous referees for their constructive comments and suggestions which helped to improve this manuscript significantly.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wei Ning.

Additional information

Publisher's Note

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

Appendix

Appendix

In order to prove Theorem 1, we need following Lemmas.

Lemma 1

Assume that for \(i=1,2, E[g_i(X, \beta _0)g_i^{\prime }(X, \beta _0)]\) is positive definite, \(\frac{\partial g_i(X, \beta )}{\partial \beta }\) is continuous in a neighborhood of the true value \(\beta _0\), \(E \left[ \left( \frac{\partial g_i(X, \beta _0)}{\partial \beta ^{\prime }}\right) \left( \frac{\partial g_i(X, \beta _0)}{\partial \beta ^{\prime }}\right) ^{\prime } \right]\), \(E \left[ \frac{\partial ^2 g_i(X, \beta )}{\partial \beta \partial \beta ^{\prime }}\right]\), \(E\left[ \left( \frac{\partial g_i(X, \beta )}{\partial \beta ^{\prime }}\right) ^{\prime } g_i(X, \beta ) \right]\) and \(E\parallel g_i(X, \beta ) \parallel ^3\) are all bounded in the neighborhood of the true value \(\beta _0\). Then, as \(n \rightarrow \infty\), \(\exists \tilde{\beta }\), \(\tilde{\lambda }=\lambda (\tilde{\beta })\) with probability 1 satisfying,

$$\begin{aligned} Q_{1n}(\tilde{\beta }, \tilde{\lambda }) = 0, \quad Q_{2n}(\tilde{\beta }, \tilde{\lambda }) = 0 \text { and } \parallel \tilde{\beta }-\beta _0 \parallel = O_p\left( m^{-\frac{1}{2}}\right) , \end{aligned}$$

where

$$\begin{aligned} Q_{1n}(\beta , \lambda )&=\sum _l \frac{1}{1 + \lambda ^{\prime }(\beta ) \theta _l^{-1}g(x_l, \beta )}\theta _l^{-1}g(x_l, \beta ),\\ Q_{2n}(\beta ,\lambda )&=\sum _l \frac{1}{1 + \lambda ^{\prime }(\beta ) \theta _l^{-1}g(x_l, \beta )} \theta _l^{-1} \left( \frac{\partial g(x_l, \beta )}{\partial \beta }\right) ^{\prime } \lambda (\beta ). \end{aligned}$$

Proof

First we will show

$$\begin{aligned} \lambda (\beta )&= \epsilon _k O_p\left( m^{-\frac{1}{2}}\right) \\&= \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta ) g^{\prime }(x_l, \beta ) \right] ^{-1} \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] +\epsilon _k o_p\left( m^{-\frac{1}{2}}\right) , \end{aligned}$$

where \(\epsilon _k = \min \{\theta _k, 1- \theta _k\}\) and \(m=n\epsilon _k = \min \{k, n-k\}\).

Let \(\beta -\beta _0 = um^{-\frac{1}{2}}\) for \(\beta \in \{\beta :\parallel \beta -\beta _0\parallel =m^{-\frac{1}{2}}\}\) where \(\parallel u\parallel =1\). Let \(\lambda\) be the solution of the function \(f(\lambda )\) given by the first score function defined in Sect. 3.

$$\begin{aligned} f(\lambda ) = \frac{1}{n} \sum _{l=1}^n \frac{\theta _l^{-1}}{1 + \lambda ^{\prime }(\beta ) \theta _l^{-1} g(x_l,\beta )} g(x_l,\beta ) = 0. \end{aligned}$$
(A.1)

Let \(\lambda =\rho u\) where \(u=(\beta -\beta _0)m^\frac{1}{2}\) and \(\parallel u\parallel =1\).

$$\begin{aligned} 0&=\,\parallel f(\rho u)\parallel \\&\ge |u^{\prime } f(\rho u)| \\&= \frac{1}{n} \left| u^{\prime } \left( \sum _l \theta _l^{-1} g(x_l,\beta ) -\rho \sum _l \frac{\theta _l^{-2} g(x_l, \beta )u^{\prime } g(x_l, \beta )}{1 + \rho u^{\prime } \theta _l^{-1} g(x_l, \beta )} \right) \right| \\&\ge \frac{\rho }{n} u^{\prime } \sum _l \frac{\theta _l^{-2} g(x_l, \beta ) u^{\prime } g(x_l, \beta )}{1 + \rho u^{\prime } \theta _l^{-1}} u -\frac{1}{n} \left| \sum _{j=1}^p e_j \sum _l \theta _l^{-1}g(x_l, \beta )\right| \\&\qquad (\text {where}\ e_j\ \text {is the unit vector in the} \ j^{th}\ \text {coordinate direction.}) \\&\ge \frac{\rho u^{\prime } Su}{1 + \rho \theta _l g^*} - O_p\left( m^{-\frac{1}{2}}\right) ,\\&\qquad \left( \text {where}\ g^*={\displaystyle \max _{l}} g(x_l, \beta )\ \text {and} \ S=\frac{1}{n}\sum _l \theta _l^{-2} g(x_l,\beta ) g^{\prime }(x_l,\beta ).\right) \end{aligned}$$

Since \(u^{\prime } Su \ge \sigma _p + o_p(1)\), where \(\sigma _p>0\) is the smallest eigen value of \(\Sigma\), then

$$\begin{aligned} \frac{\rho }{1 + \rho \theta _l g^*} = O_p\left( m^{-\frac{1}{2}}\right) \end{aligned}$$

So, \(\parallel \lambda \parallel = \rho = O_p(m^{-\frac{1}{2}})\).

Let \(\gamma _l = \lambda ^{\prime }(\beta )\theta _l^{-1} g(x_l, \beta )\). Then, \({\displaystyle \max _{l}} |\gamma _l| =O_p(m^{-\frac{1}{2}})o(m^{\frac{1}{2}}) = o_p(1)\).

Expanding (A.1),

$$\begin{aligned} 0 = f(\lambda )&= \frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta ) \left[ 1 - \gamma _l + \frac{\gamma _l^2}{1+\gamma _l} \right] \nonumber \\&= \frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta ) - \frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta )\cdot \gamma + \frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta ) \frac{\gamma _l^2}{1+\gamma _l} \nonumber \\&= E(\theta _l^{-1} g(x_l, \beta )) - S \lambda + \frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta ) \frac{\gamma _l^2}{1+\gamma _l}. \end{aligned}$$
(A.2)

The last equality is since \(\frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta )\cdot \gamma = \frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta ) \theta _l^{-1} g^{\prime }(x_l, \beta ) \lambda = S \lambda\).

By substituting \(\gamma _l\), we have the final term of (A.2);

$$\begin{aligned} \frac{1}{n} \sum _l \parallel \theta _l^{-1} g(x_l, \beta )\parallel ^3 \parallel \lambda \parallel ^2 |1+\gamma _l|^{-1} = o_p\left( m^{\frac{1}{2}}\right) O_p(m^{-1}) o_p(1) = o_p\left( m^{-\frac{1}{2}}\right) . \end{aligned}$$

Therefore,

$$\begin{aligned} 0&= E(\theta _l^{-1} g(x_l, \beta )) - S \lambda + o_p\left( m^{-\frac{1}{2}}\right) \nonumber \\&\Rightarrow \lambda = S^{-1} E(\theta _l^{-1} g(x_l, \beta )) + o_p\left( m^{-\frac{1}{2}}\right) \nonumber \\&\Rightarrow \lambda = \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta ) g^{\prime }(x_l, \beta ) \right] ^{-1} \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] + o_p\left( m^{-\frac{1}{2}}\right) . \end{aligned}$$
(A.3)

Now, denote \(V_n(\beta )=\frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta )g^{\prime }(x_l, \beta )\), \(\bar{g}(\beta ) =\frac{1}{n}\sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\), and \(\varepsilon =\epsilon _k o_p(m^{-\frac{1}{2}})\). So (A.2) can be rewritten as,

$$\begin{aligned} \lambda (\beta ) = V_n(\beta )^{-1} \bar{g}(\beta ) + \varepsilon . \end{aligned}$$

Since \(\gamma _l = \lambda ^{\prime }(\beta )\theta _l^{-1} g(x_l, \beta ),\) so \(\sum _{l=1}^n |r_l|^3 = o_p(1)\).

Let \(a_m\) be any constant sequence such that \(a_m \rightarrow \infty\), and \(a_m m^{-\frac{1}{2}}\rightarrow 0\). Denote the ball \(B(\beta _0, a_m) = \{\beta |\parallel \beta -\beta _0 \parallel \le a_m m^{-\frac{1}{2}}\}\) and the surface of the ball \(\partial B(\beta _0, a_m) = \{\beta |\parallel \beta -\beta _0 \parallel = \phi a_m m^{-\frac{1}{2}}, \parallel \phi \parallel =1\}\). For any \(\beta \in \partial B(\beta _0, a_m)\), we have

$$\begin{aligned} V_n(\beta )&= \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta )g^{\prime }(x_l, \beta )\\&= \frac{n}{k} \frac{1}{k}\sum _{l=1}^k g_1(x_l, \beta _0)g_1^{\prime }(x_l, \beta _0) +\frac{n}{n-k} \frac{1}{n-k} \sum _{l=k+1}^n g_2(x_l, \beta _0) g_2^{\prime }(x_l, \beta _0)+o_p(\epsilon _k^{-1})\\&= \frac{n}{k} E g_1(x_l, \beta _0)g_1^{\prime }(x_l, \beta _0) +\frac{n}{n-k} E g_2(x_l, \beta _0)g_2^{\prime }(x_l, \beta _0) + o_p(\epsilon _k^{-1})\\&\le \epsilon _k^{-1}\left[ E g_1(x_l, \beta _0) g_1^{\prime }(x_l, \beta _0) + E g_2(x_l, \beta _0) g_2^{\prime }(x_l, \beta _0)\right] + o_p(\epsilon _k^{-1}), \end{aligned}$$

and

$$\begin{aligned} \bar{g}(\beta _0)&= \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta ) \\&= \frac{1}{k} \sum _{l=1}^k g_1(x_l, \beta _0) + \frac{1}{n-k} \sum _{l=k+1}^n g_2(x_l, \beta _0)\\&= \frac{1}{k} o_p\left( k^{\frac{1}{2}}\right) + \frac{1}{n-k} o_p\left( (n-k)^{\frac{1}{2}}\right) \\&= o_p\left( k^{-\frac{1}{2}}\right) + o_p\left( (n-k)^{-\frac{1}{2}}\right) \\&= o_p\left( m^{-\frac{1}{2}}\right) . \end{aligned}$$

By the Taylor expansion, for any \(\beta \in \partial B(\beta _0, a_m)\), we have

$$\begin{aligned} l_E(\beta ) = \sum _l \lambda ^{\prime }(\beta )\theta _l^{-1} g(x_l, \beta ) -\frac{1}{2} \sum _l \left[ \lambda ^{\prime }(\beta )\theta _l^{-1} g(x_l, \beta )\right] ^2+ o_p(1). \end{aligned}$$
(A.4)

The first term of (A.4) is;

$$\begin{aligned} \sum _l \lambda ^{\prime }(\beta )\theta _l^{-1} g(x_l, \beta )&= \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] ^{\prime } \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta ) g^{\prime }(x_l, \beta ) \right] ^{-1}\nonumber \\&\quad \left[ \frac{1}{n}\sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] \nonumber \\&\quad + o_p(1). \end{aligned}$$
(A.4.1)

The second term of (A.4) is:

$$\begin{aligned}&\frac{1}{2} \sum _l \left[ \lambda ^{\prime }(\beta ) \theta _l^{-1} g(x_l, \beta ) \right] ^2\nonumber \\&\quad = \frac{1}{2} \sum _l \lambda ^{\prime }(\beta ) \theta _l^{-2} g(x_l, \beta )g^{\prime }(x_l, \beta )\nonumber \\&\quad = \frac{n}{2}\left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] ^{\prime } \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta ) g^{\prime }(x_l, \beta ) \right] ^{-1}\nonumber \\&\qquad \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta ) g^{\prime }(x_l, \beta ) \right] \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta ) g^{\prime }(x_l, \beta ) \right] ^{-1}\left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] + o_p(1)\nonumber \\&\quad = \frac{n}{2}\left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] ^{\prime } \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-2} g(x_l, \beta ) g^{\prime }(x_l, \beta ) \right] ^{-1} \left[ \frac{1}{n} \sum _{l=1}^n \theta _l^{-1} g(x_l, \beta )\right] + o_p(1). \end{aligned}$$
(A.4.2)

Now,

$$\begin{aligned}&(\mathrm{A}.4.1)\,\text {--}\,(\mathrm{A}.4.2)\\&\quad =\frac{n}{2} \left( \frac{1}{n}\sum _l \theta _l^{-1} g(x_l, \beta ) \right) ^{\prime } \left( \frac{1}{n}\sum _l \theta _l^{-2} g(x_l, \beta )g^{\prime }(x_l, \beta ) \right) ^{-1} \left( \frac{1}{n}\sum _l \theta _l^{-1} g(x_l, \beta ) \right) + o_p(1). \end{aligned}$$

So we can rewrite (A.4) as,

$$\begin{aligned} l_E(\beta )&= \frac{n}{2} \left( \frac{1}{n}\sum _l \theta _l^{-1} g(x_l, \beta ) \right) ^{\prime } \left( \frac{1}{n}\sum _l \theta _l^{-2} g(x_l, \beta )g^{\prime }(x_l, \beta ) \right) ^{-1} \left( \frac{1}{n} \sum _l \theta _l^{-1} g(x_l, \beta ) \right) + o_p(1)\\&= \frac{n}{2} \bar{g}^{\prime } (\beta ) (V_n(\beta ))^{-1}\bar{g}(\beta ) + o_p(1) \\&= \frac{n}{2} \left\{ \bar{g}(\beta _0) + \frac{1}{n} \sum _l \theta _l^{-1} \frac{\partial g(x_l, \beta _0)}{\partial \beta ^{\prime }}\phi a_m m^{-\frac{1}{2}} + O\left[ \left( a_m m^{-\frac{1}{2}}\right) ^2\right] \right\} ^{\prime } \times \left( V_n(\beta )\right) ^{-1} \\&\quad \times \left\{ \bar{g}(\beta _0) + \frac{1}{n} \sum _l \theta _l^{-1} \frac{\partial g(x_l, \beta _0)}{\partial \beta ^{\prime }}\phi a_m m^{-\frac{1}{2}} + O\left[ \left( a_m m^{-\frac{1}{2}}\right) ^2\right] \right\} +o_p(1)\\&\quad \qquad (\text {By Taylor expansion of each term}.)\\&\ge \frac{n \epsilon _k}{2} \left\{ \bar{g}(\beta _0) +\frac{1}{n} \sum _l \theta _l^{-1} \frac{\partial g(x_l, \beta _0)}{\partial \beta ^{\prime }}\phi a_m m^{-\frac{1}{2}} + O\left[ \left( a_m m^{-\frac{1}{2}}\right) ^2\right] \right\} ^{\prime } \times \left( V_n(\beta )\right) ^{-1} \\&\quad \times \left\{ \bar{g}(\beta _0) + \frac{1}{n} \sum _l \theta _l^{-1} \frac{\partial g(x_l, \beta _0)}{\partial \beta ^{\prime }}\phi a_m m^{-\frac{1}{2}} +O\left[ \left( a_m m^{-\frac{1}{2}}\right) ^2\right] \right\} + o_p(1). \end{aligned}$$

As \(n\rightarrow \infty\), \(l_E(\beta ) \rightarrow \infty\).

Similarly,

$$\begin{aligned}&l_E(\beta _0) = \frac{n}{2} \bar{g}^{\prime }(\beta _0) V_n(\beta _0)^{-1} \bar{g}(\beta _0) + o_p(1),\\&V_n(\beta _0) = \frac{n}{k} E g_1(x_l, \beta _0)g_1^{\prime } (x_l, \beta _0) + \frac{n}{n-k} E g_2(x_l, \beta _0)g_2^{\prime } (x_l, \beta _0) + o_p(\epsilon _k^{-1}). \end{aligned}$$

Thus, \(l_E(\beta _0) = O_p(1)\) implies that for any \(\beta \in \partial B(\beta _0, a_m)\), \(l_E(\beta )\) can not arrive its minimum value with the probability approaching to 1. Since \(l_E(\beta )\) is a continuous function about \(\beta\), as \(\beta \in B(\beta _0, a_m)\), \(l_E(\beta )\) has a minimum value in the interior of this ball satisfying,

$$\begin{aligned} 0 = \left. \frac{\partial l_E(\beta )}{\partial \beta }\right| _{\beta =\tilde{\beta }}&= \sum _l \left. \frac{\left( \frac{\partial \lambda ^{\prime }(\beta )}{\partial \beta }\right) \theta _l^{-1}g(x_l, \beta ) + \theta _l^{-1}\left( \frac{\partial g(x_l, \beta )}{\partial \beta }\right) ^{\prime } \lambda (\beta )}{1 + \lambda ^{\prime }(\beta ) \theta _l^{-1}g(x_l, \beta )} \right| _{\beta =\tilde{\beta }}\\&= \left. \frac{\partial \lambda ^{\prime }(\beta )}{\partial \beta } \sum _l \frac{\theta _l^{-1} g(x_l, \beta )}{1 + \lambda ^{\prime }(\beta ) \theta _l^{-1}g(x_l, \beta )} \right| _{\beta =\tilde{\beta }} + \sum _l \frac{\theta _l^{-1} \left( \frac{\partial g(x_l, \beta )}{\partial \beta }\right) ^{\prime } \lambda (\beta )}{1 + \lambda ^{\prime }(\beta )\theta _l^{-1} g(x_l, \beta )}\\&= \sum _l \frac{\theta _l^{-1}\left( \frac{\partial g(x_l, \beta )}{\partial \beta }\right) ^{\prime } \lambda (\beta )}{1 + \lambda ^{\prime }(\beta ) \theta _l^{-1}g(x_l, \beta )}\\&\qquad \left( \text {Since}\ \sum _l \left. \frac{\theta _l^{-1}g(x_l, \beta )}{1 + \lambda ^{\prime }(\beta ) \theta _l^{-1}g(x_l, \beta )} \right| _{\beta =\tilde{\beta }}= Q_{1n}(\tilde{\beta }, \tilde{\lambda }) = 0\right) \\&= Q_{2n}(\tilde{\beta }, \tilde{\lambda }). \end{aligned}$$

Hence, \(Q_{1n}(\tilde{\beta }, \tilde{\lambda }) = 0\) and \(Q_{2n}(\tilde{\beta }, \tilde{\lambda }) = 0\). That is, \(\parallel \tilde{\beta }-\beta _0 \parallel = O_p(a_m m^{-\frac{1}{2}}).\) But \(a_m\) is arbitrary, hence \(\parallel \tilde{\beta }-\beta _0 \parallel = O_p(m^{-\frac{1}{2}})\). \(\square\)

Remark 1

If the partitioned matrix

$$\begin{aligned} \begin{pmatrix} A &{} B \\ B^{\prime } &{} 0 \end{pmatrix} \end{aligned}$$

is non-singular, then

$$\begin{aligned} \begin{pmatrix} A &{} B \\ B^{\prime } &{} 0 \end{pmatrix}^{-1} =\begin{pmatrix} P &{} Q \\ Q^{\prime } &{} R \end{pmatrix} \end{aligned}$$

where

$$\begin{aligned}&P=A^{-1} -A^{-1}B(B^{\prime } A^{-1} B)^{-1} B^{\prime } A^{-1}, \qquad Q= A^{-1}B(B^{\prime } A^{-1} B)^{-1},\\&Q^{\prime } = (B^{\prime } A^{-1} B)^{-1}B^{\prime } A^{-1}, \qquad R=-(B^{\prime } A^{-1} B)^{-1} \end{aligned}$$

Remark 2

$$\begin{aligned} \text {If } \begin{pmatrix} A &{} B \\ C &{} D \end{pmatrix} \text {is a}\ n\times n\ \text {symmetric positive definite matrix, and the partitioned matrices}\ A\in {\mathbb {R}}^{m\times m}, \end{aligned}$$

\(B\in {\mathbb {R}}^{m\times n-m}\), and \(D\in {\mathbb {R}}^{(n-m)\times (n-m)}\), then

  1. 1.

    the matrix \((D-CA^{-1}B)\) is symmetric and positive definite,

  2. 2.
    $$\begin{aligned} \begin{pmatrix} A &{} B \\ C &{} D \end{pmatrix}^{-1} \ge \begin{pmatrix} A^{-1} &{} 0 \\ 0 &{} 0 \end{pmatrix}. \end{aligned}$$

Remark 3

\(\beta ^{\prime } = ((\beta ^{\prime }, \mu ^{\prime }), \delta ^{\prime })\).

$$\begin{aligned}&\frac{\partial Q_{1n}(\beta , 0)}{\partial \lambda ^{\prime }} =-\frac{1}{n}\sum _l \theta _l^{-2} g(x_l, \beta )g^{\prime }(x_l, \beta ), \quad \frac{\partial Q_{1n}(\beta , 0)}{\partial \beta ^{\prime }} =\frac{1}{n}\sum _l \theta _l^{-1} \frac{\partial g(x_l, \beta )}{\partial \beta ^{\prime }}\\&\frac{\partial Q_{2n}(\beta , 0)}{\partial \lambda ^{\prime }} =\frac{1}{n}\sum _l \left( \theta _l^{-1} \frac{\partial g(x_l, \beta )}{\partial \beta ^{\prime }}\right) ^{\prime }, \quad \frac{\partial Q_{2n}(\beta , 0)}{\partial \beta ^{\prime }} = 0\\&\begin{pmatrix} \frac{\partial Q_{1n}}{\partial \lambda ^{\prime }} &{} \frac{\partial Q_{1n}}{\partial \beta ^{\prime }} \\ \frac{\partial Q_{2n}}{\partial \lambda ^{\prime }} &{} 0 \end{pmatrix} \longrightarrow \begin{pmatrix} S_{11} &{} S_{12} \\ S_{21} &{} 0 \end{pmatrix} = S(\beta ) \equiv S \end{aligned}$$

where

\(S_{11}(\beta ) = -\theta _l^{-1} E\left[ g_1(x_l, \beta )g_1^{\prime }(x_l, \beta ) \right] - (1-\theta _l)^{-1} E\left[ g_2(x_l, \beta )g_2^{\prime }(x_l, \beta )\right]\),

\(S_{12}(\beta ) = \theta ^{-1}_lE\left[ \frac{\partial g_1(x_l, \beta _0)}{\partial \beta ^{\prime }} \right] + (1-\theta _l)^{-1}E \left[ \frac{\partial g_2(x_l, \beta _0)}{\partial \beta ^{\prime }} \right]\),

\(S_{21}(\beta ) = S_{12}^{\prime }(\beta )\),

\(S_{12,i}(\beta ) = \theta ^{-1}_lE\left[ \frac{\partial g_1(x_l, \beta _0)}{\partial \beta _i^{\prime }} \right] + (1-\theta _l)^{-1}E \left[ \frac{\partial g_2(x_l, \beta _0)}{\partial \beta _i^{\prime }} \right]\), \(i=1, 2\).

By, Remark 1,

$$\begin{aligned} S^{-1} =\begin{pmatrix} S_{11} &{} S_{12} \\ S_{21} &{} 0 \end{pmatrix} ^{-1} =\begin{pmatrix} P &{} Q \\ Q^{\prime } &{} R \end{pmatrix} \end{aligned}$$

where

$$\begin{aligned}&P = S_{11}^{-1} - S_{11}^{-1} S_{12}(S_{21}S_{11}^{-1}S_{12})^{-1} S_{21}S_{11}^{-1} = S_{11}^{-1} + S_{11}^{-1} S_{12} \Sigma S_{21}S_{11}^{-1};\\&\Sigma = (S_{21}(-S_{11}^{-1})S_{12})^{-1},\\&Q = - S_{11}^{-1} S_{12}(S_{21}S_{11}^{-1}S_{12})^{-1} = - S_{11}^{-1} S_{12} \Sigma , \\&Q^{\prime } = -\Sigma S_{21}S_{11}^{-1}, \quad R = -(S_{21}S_{11}^{-1}S_{12})^{-1} = \Sigma . \end{aligned}$$

Lemma 2

Under the conditions in Lemma 1and \(H_0\), as \(n\rightarrow \infty\) we have

$$\begin{aligned} \sqrt{n}\Sigma ^{-\frac{1}{2}}(\tilde{\beta }-\beta _0) \rightarrow N(0, I_{2p+q}), \end{aligned}$$

where \(\Sigma = [S_{21}(-S_{11})^{-1}S_{12}]^{-1}\).

Proof

Expanding \(Q_{1n}(\tilde{\beta }, \tilde{\lambda })\) and \(Q_{2n}(\tilde{\beta }, \tilde{\lambda })\) at \((\theta _0,0)\), by the conditions of the \(H_0\) and Lemma 1, we have,

$$\begin{aligned} 0&= Q_{1n}(\tilde{\beta }, \tilde{\lambda })\\&= Q_{1n}(\beta _0, 0) + \frac{\partial Q_{1n}(\beta _0, 0)}{\partial \beta ^{\prime }} (\tilde{\beta } - \beta _0) +\frac{\partial Q_{1n}(\beta _0, 0)}{\partial \lambda ^{\prime }} (\tilde{\lambda } - 0) + O_p(m^{-1}),\\ 0&= Q_{2n}(\tilde{\beta }, \tilde{\lambda })\\&= Q_{2n}(\beta _0, 0) + \frac{\partial Q_{2n}(\beta _0, 0)}{\partial \beta ^{\prime }} (\tilde{\beta } - \beta _0) + \frac{\partial Q_{2n}(\beta _0, 0)}{\partial \lambda ^{\prime }} (\tilde{\lambda } - 0) + O_p(m^{-1}), \\&\quad \begin{pmatrix} -Q_{1n}(\beta _0, 0) + O_p(m^{-1}) \\ \epsilon _k O_p(m^{-1}) \end{pmatrix} =\begin{pmatrix} \frac{\partial Q_{1n}}{\partial \lambda ^{\prime }} &{} \frac{\partial Q_{1n}}{\partial \beta ^{\prime }} \\ \frac{\partial Q_{2n}}{\partial \lambda ^{\prime }} &{} 0 \end{pmatrix} \begin{pmatrix} \tilde{\lambda } \\ \tilde{\beta } -\beta _0 \end{pmatrix}. \end{aligned}$$

By LLN,

$$\begin{aligned} \begin{pmatrix} \tilde{\lambda } \\ \tilde{\beta } -\beta _0 \end{pmatrix} \longrightarrow S^{-1}(\beta _0) \begin{pmatrix} -Q_{1n}(\beta _0, 0) + O_p(m^{-1}) \\ \epsilon _k O_p(m^{-1}). \end{pmatrix} \end{aligned}$$

By Remark 1,

$$\begin{aligned} \tilde{\beta } -\beta _0 = (0 \,I) S^{-1} \begin{pmatrix} -Q_{1n}(\beta _0, 0) + O_p(m^{-1}) \\ \epsilon _k O_p(m^{-1}) \end{pmatrix}. \end{aligned}$$

Therefore,

$$\begin{aligned}&\begin{pmatrix} \tilde{\lambda } \\ \tilde{\beta } -\beta _0 \end{pmatrix} \longrightarrow \begin{pmatrix} S_{11}^{-1} + S_{11}^{-1} S_{12} &{} - S_{11}^{-1} S_{12} \Sigma \\ -\Sigma S_{21}S_{11}^{-1} &{} \Sigma \end{pmatrix} \begin{pmatrix} -Q_{1n}(\beta _0, 0) + O_p(m^{-1}) \\ \epsilon _k O_p(m^{-1}) \end{pmatrix}. \\&\tilde{\beta } -\beta _0 \rightarrow -\Sigma S_{21}S_{11}^{-1} \left( -Q_{1n}(\beta _0, 0) + O_p(m^{-1}) \right) + \Sigma \epsilon _k O_p(m^{-1})\\&\quad = (S_{21}S_{11}^{-1}S_{12})^{-1} S_{21}S_{11}^{-1}Q_{1n}(\beta _0, 0) - \Sigma S_{21}S_{11}^{-1} O_p(m^{-1}) + \Sigma \epsilon _k O_p(m^{-1})\\&\quad = \frac{1}{\sqrt{n}} (S_{21}S_{11}^{-1}S_{12})^{-1} S_{21} (-S_{11})^{-1/2}(-S_{11})^{-1/2}\sqrt{n}Q_{1n}(\beta _0, 0) +\epsilon _k O_p\left( m^{-\frac{1}{2}}\right) \end{aligned}$$

Since \((-S_{11})^{-1/2}\sqrt{n}Q_{1n}(\beta _0, 0) \rightarrow N(0, I_{2(p+q)})\), \(\sqrt{n}S_{21}S_{11}^{-1}(\tilde{\beta } - \beta _0) \rightarrow N(0, I_{2p+q})\). \(\square\)

Lemma 3

$$\begin{aligned} -2\log \Lambda _k = 2l_E(\tilde{\beta }_1^0, 0) -2l_E(\tilde{\beta }_1^0, \tilde{\beta }_2^0), \end{aligned}$$

where \(\tilde{\beta }_1^0\) minimizes \(l_E(\beta , 0)\) with respect to \(\beta _1\) under \(H_0\),

$$\begin{aligned} -2\log \Lambda _k = \left[ (-S_{11})^{-1/2}\sqrt{n}Q_{1n}(\beta _0, 0) \right] ^{\prime } \Delta \left[ (-S_{11})^{-1/2}\sqrt{n}Q_{1n}(\beta _0, 0) \right] + O_p\left( m^{-\frac{1}{2}}\right) \end{aligned}$$

where

$$\begin{aligned} \Delta = (-S_{11})^{-1/2} \left\{ S_{12} [S_{21}(-S_{11})^{-1}S_{12}]^{-1} S_{21} - S_{12,1} [S_{21,1}(-S_{11})^{-1}S_{12,1}]^{-1} S_{21,1} \right\} (-S_{11})^{-1/2} \ge 0. \end{aligned}$$

Proof

Similar to Qin and Lawless (1994), we can derive,

$$\begin{aligned} l_E(\tilde{\beta }_1^0, \tilde{\beta }_2^0) = -\frac{n}{2} Q_{1n}^{\prime } (\beta _0, 0) B Q_{1n} (\beta _0, 0) + O_p\left( m^{-\frac{1}{2}}\right) , \end{aligned}$$

where \(B = S_{11}^{-1} + S_{11}^{-1} S_{12} \Sigma S_{21}S_{11}^{-1}\), and

$$\begin{aligned} l_E(\tilde{\beta }_1^0, 0) = -\frac{n}{2} Q_{1n}^{\prime } (\beta _0, 0) A Q_{1n} (\beta _0, 0) + O_p\left( m^{-\frac{1}{2}}\right) , \end{aligned}$$

where \(A = S_{11}^{-1} + S_{11}^{-1} S_{12,1} (S_{21,1}S_{11}^{-1}S_{12,1})^{-1} S_{21,1}S_{11}^{-1}\). Then,

$$\begin{aligned}&2 \left[ l_E(\tilde{\beta }_1^0, 0) - l_E(\tilde{\beta }_1^0, \tilde{\beta }_2^0)\right] \\&\quad =\left[ -Q_{1n}^{\prime } (\beta _0, 0) A Q_{1n} (\beta _0, 0) +O_p\left( m^{-\frac{1}{2}}\right) \right] \\&\qquad +\left[ n Q_{1n}^{\prime } (\beta _0, 0) B Q_{1n} (\beta _0, 0) + O_p\left( m^{-\frac{1}{2}}\right) \right] \\&\quad = n Q_{1n}^{\prime } (\beta _0, 0) (B-A) Q_{1n} (\beta _0, 0) + O_p\left( m^{-\frac{1}{2}}\right) \\&\quad = n Q_{1n}^{\prime } (\beta _0, 0) S_{11}^{-1} \left[ S_{12} \Sigma S_{21} - S_{12,1} \Sigma ^* S_{12,2}\right] S_{11}^{-1} Q_{1n} (\beta _0, 0) + O_p\left( m^{-\frac{1}{2}}\right) \\&\quad \qquad (B-A = S_{11}^{-1} + S_{11}^{-1} S_{12} \Sigma S_{21}S_{11}^{-1} -S_{11}^{-1} - S_{11}^{-1} S_{12,1} (S_{21,1}S_{11}^{-1}S_{12,1})^{-1} S_{21,1}S_{11}^{-1}.)\\&\quad \qquad (\text {So},\ \Sigma ^*=(S_{21,1}S_{11}^{-1}S_{12,1})^{-1})\\&\quad = \left[ (-S_{11})^{-1/2}\sqrt{n}Q_{1n}(\beta _0, 0)\right] ^{\prime } (-S_{11})^{-1/2} \left[ S_{12} \Sigma S_{21} - S_{12,1} \Sigma ^* S_{12,2}\right] \\&\quad \qquad (-S_{11})^{-1/2} \left[ (-S_{11})^{-1/2}\sqrt{n}Q_{1n}(\beta _0, 0)\right] +O_p\left( m^{-\frac{1}{2}}\right) . \end{aligned}$$

Take \(\Delta = (-S_{11})^{-1/2} \left[ S_{12} \Sigma S_{21} - S_{12,1} \Sigma ^* S_{12,2}\right] (-S_{11})^{-1/2}\). Now,

$$\begin{aligned} \Delta&= (-S_{11})^{-1/2} \left[ S_{12} \left( S_{21}(-S_{11}^{-1})S_{12}\right) ^{-1} S_{21} - S_{12,1} \left( S_{21,1}S_{11}^{-1}S_{12,1}\right) ^{-1} S_{12,2}\right] (-S_{11})^{-1/2}\\&= (-S_{11})^{-1/2} (S_{12,1},\, S_{12,2}) \left\{ [S_{21}(-S_{11})^{-1}S_{12}]^{-1} -\begin{pmatrix} \left( S_{21,1}S_{11}^{-1}S_{12,1}\right) ^{-1} &{} 0 \\ 0 &{} 0 \end{pmatrix} \right\} \\&\quad \times \begin{pmatrix} S_{21,1}\\ S_{21,2} \end{pmatrix} (-S_{11})^{-1/2} \\&\ge 0. \\&\qquad (\text {By Remark}\,) \end{aligned}$$

\(\square\)

Lemma 4

Under the conditions of Theorem 1and the null hypothesis, denote \(U_{n_k} = \left\{ \frac{k}{n}: \frac{T}{n} \le (1-\frac{T}{n})\right\}\), for all \(\delta >0\), we can find \(C=C(\delta )\), \(T=T(\delta )\) and \(N=N(\delta )\) such that

$$\begin{aligned}&P\left( {\displaystyle \max _{\frac{k}{n} \in U_{n_k}}} \left( \frac{m}{\log \log m}\right) ^{1/2} \left\| \frac{\tilde{\lambda }}{\epsilon _k}\right\|> C \right) \le \delta , \quad P\left( n^{-1/2} {\displaystyle \max _{\frac{k}{n} \in U_{n_k}}} m \left\| \frac{\tilde{\lambda }}{\epsilon _k}\right\|> C \right) \le \delta ,\\&P\left( {\displaystyle \max _{\frac{k}{n} \in U_{n_k}}} \left( \frac{m}{\log \log m}\right) ^{1/2} \parallel \tilde{\theta } -\theta _0 \parallel> C \right) \le \delta , \quad P\left( n^{-1/2} {\displaystyle \max _{\frac{k}{n} \in U_{n_k}}} m \parallel \tilde{\theta } - \theta _0 \parallel > C \right) \le \delta . \end{aligned}$$

Proof

The proof is similar to Lemma 1.2.2 of Csörgo and Horváth (1997)). \(\square\)

Lemma 5

Under the conditions of Theorem 1and \(H_0\), for all \(0\le \alpha <\frac{1}{2}\) we have:

$$\begin{aligned}&n^\alpha {\displaystyle \max _{\frac{k}{n} \in U_{n_k}}} \left[ \theta _k (1-\theta _k) \right] ^\alpha | -2\log \Lambda - R_k| = O_p(1),\\&{\displaystyle \max _{\frac{k}{n} \in U_{n_k}}} \left[ \theta _k (1-\theta _k)\right] ^\alpha | -2\log \Lambda - R_k| = O_p\left( n^{-\frac{1}{2}} (\log \log n)^{\frac{3}{2}}\right) , \end{aligned}$$

where \(\Theta _{nk}=\{k:\delta _1 \le k \le n-\delta _2\}.\)

Proof of Theorem 1

The proof of Theorem 1 is similar to the proof of Theorem 1.3.1 (Theorem A.3.4) of Csörgo and Horváth (1997) which derives the null distribution of the trimmed test statistic. \(\square\)

Proof of Theorem 2

The ELR test statistic is,

$$\begin{aligned} -2 \log \Lambda _k = Z_{H_0,k_0} - Z_{H_1,k_0}. \end{aligned}$$

Under \(H_1\), \(Z_{H_1,k_0}\) also follows an asymptotic \(\chi ^2\) distribution. Therefore, \(Z_{H_1,k_0} = O_p(1)\). We only need to prove that \(P(Z_{H_0,k_0}>cn) \rightarrow 1\) for a positive constant c under \(H_1\). For any fixed \(\varepsilon\), we can obtain

$$\begin{aligned} \frac{1}{2n}Z_{H_0,k_0}&= {\displaystyle \sup _{\lambda }}\frac{1}{n} \sum _{l=1}^n \log \left[ 1 + \theta _l^{-1} \lambda ^{\prime } g(x_l, \varepsilon ) \right] \\&= {\displaystyle \sup _{\lambda _1}}\frac{1}{n}\sum _{l=1}^{k_0} \log \left[ 1 + \theta _{k_0}^{-1} \lambda _1^{\prime } g_1(x_l, \varepsilon ) \right] \\&\quad + {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k_0+1}^{n} \log \left[ 1 + (1- \theta _{k_0})^{-1} \lambda _2^{\prime } g_2(x_l, \varepsilon ) \right] \\&\xrightarrow {\text {a.s.}}{\displaystyle \sup _{\lambda _1}} \theta _0 E \log \left( 1 + \theta _0^{-1} \lambda _1^{\prime } g_1(x_l, \varepsilon ) \right) \\&\quad + {\displaystyle \sup _{\lambda _2}} (1-\theta _0) E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \varepsilon )\right) \right] \end{aligned}$$

By Jensen’s inequality,

$$\begin{aligned}&E \log \left( 1 + \theta _0^{-1} \lambda _1^{\prime } g_1(x_l, \varepsilon ) \right) \le \log \left[ E \left( 1 + \theta _0^{-1} \lambda _1^{\prime } g_1(x_l, \varepsilon ) \right) \right] = 0\\&\quad \Longrightarrow {\displaystyle \sup _{\lambda _1}} \theta _0 E \log \left( 1 + \theta _0^{-1} \lambda _1^{\prime } g_1(x_l, \varepsilon ) \right) = 0. \end{aligned}$$

Thus,

$$\begin{aligned} \frac{1}{2n}Z_{H_0,k_0}&\xrightarrow {\text {a.s.}} {\displaystyle \sup _{\lambda _2}} (1-\theta _0) E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \varepsilon )\right) \right] \\&\le (1-\theta _0) c_0 \end{aligned}$$

Hence, \(P(Z_{H_0, k_0} \ge (1-\theta _0c_0) \rightarrow 1\). Thus, the proof. \(\square\)

Proof of Theorem 3

To prove: For arbitrary small \(\frac{\theta _0}{2} > \eta\), \(|\frac{k_0-k}{n}|\ge \eta\), \(-2\log \Lambda _k\) cannot arrive at its maximum with probability approaching to 1.

Without loss of generality, suppose \(k<k_0\) and \(\frac{k_0-k}{n} \ge \eta\). Then we have,

$$\begin{aligned} -2 \log \Lambda _{k_0} - (-2 \log \Lambda _k) = (Z_{H_0,k_0} -Z_{H_1,k_0}) - (Z_{H_0,k} - Z_{H_1,k}). \end{aligned}$$

Since \(Z_{H_1,k_0} = O_p(1)\)

$$\begin{aligned}&\frac{1}{2n}(Z_{H_0,k_0} -Z_{H_0,k} + Z_{H_1,k}) \\&\quad = {\displaystyle \sup _{\lambda _1}}\frac{1}{n}\sum _{l=1}^{k_0} \log \left[ 1 + \theta _{k_0}^{-1} \lambda _1^{\prime } g_1(x_l, \beta _0) \right] + {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k_0+1}^{n} \log \left[ 1 + (1- \theta _{k_0})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] \\&\qquad - {\displaystyle \sup _{\lambda _1}}\frac{1}{n}\sum _{l=1}^{k} \log \left[ 1 + \theta _{k}^{-1} \lambda _1^{\prime } g_1(x_l, \beta _0) \right] + {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k+1}^{n} \log \left[ 1 + (1- \theta _{k})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] \\&\qquad + {\displaystyle \sup _{\lambda _1}}\frac{1}{n}\sum _{l=1}^{k} \log \left[ 1 + \theta _{k}^{-1} \lambda _1^{\prime } g_1(x_l, \beta _0) \right] + {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k+1}^{n} \log \left[ 1 + (1- \theta _{k})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _1) \right] \\&\quad \ge {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k_0+1}^{n} \log \left[ 1 + (1- \theta _{k_0})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] \\&\qquad - {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k+1}^{n} \log \left[ 1 + (1- \theta _{k})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] \\&= {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k_0+1}^{n} \log \left[ 1 + (1- \theta _{k_0})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] \\&\qquad - {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k+1}^{n} \log \left[ 1 + \rho _k(1- \theta _{k_0})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] \quad \qquad \left( \rho _k=\frac{n-k_0}{n-k}\right) \\&\quad = {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k_0+1}^{n} \log \left[ 1 + (1- \theta _{k_0})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] - {\displaystyle \sup _{\lambda _2}}\frac{1}{n}\sum _{l=k+1}^{n} \log \left[ 1 + (1- \theta _{k_0})^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0) \right] \\& \quad \qquad \left( \text {Since},\ \frac{n-k_0}{n}\le \rho _k \le \frac{n-k_0}{n-k_0+n\eta }. \ \text {So},\ 1-\theta _0\le \varliminf \rho _k \le \varlimsup \rho _k \le \frac{1-\theta _0}{1-\theta _0+\eta }\right) \\&\quad \xrightarrow {\text {a.s.}}{\displaystyle \sup _{\lambda _1}} (1-\theta _0) E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] \\&\quad - {\displaystyle \sup _{\lambda _1}} \left\{ \frac{k_0-k}{n} E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] \right. \\&\quad \left. + (1-\theta _0) E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] \right\} \\&\quad \ge {\displaystyle \sup _{\lambda _1}} (1-\theta _0) E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] \\&\qquad - {\displaystyle \sup _{\lambda _1}} \left\{ \eta E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] \right. \\&\qquad \left. + (1-\theta _0) E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] \right\} \\&\qquad \quad \left( \text {By Jensen's inequality and}\ \frac{k_0-k}{n}\ge \eta .\right) \end{aligned}$$

Assume that \({\displaystyle \sup _{\lambda _1}} \left\{ \eta E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] + (1-\theta _0) E \left[ \log \left( 1+(1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] \right\}\) attains its maximum at \(\delta _2^*\). Then we have,

$$\begin{aligned}&\frac{1}{2n}(Z_{H_0,k_0} -Z_{H_0,k} + Z_{H_1,k}) \\&\quad \ge {\left\{ \begin{array}{ll} {\displaystyle \sup _{\lambda _1}} (1-\theta _0) E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{\prime } g_2(x_l, \beta _0)\right) \right] , \text { if}\ \delta _2^*=0,\\ -\eta E \left[ \log \left( 1 + (1-\theta _0)^{-1} \lambda _2^{*\prime } g_2(x_l, \beta _0)\right) \right] , \text { if}\ \delta _2^*\ne 0. \end{array}\right. } \end{aligned}$$

Therefore, by the condition that for every fixed parameter \(\delta =\beta ^*-\beta \ne 0\), there exists a positive constant \(c_0>0\) satisfy that \(\infty> {\displaystyle \inf _{\delta \ne 0}} {\displaystyle \sup _{\lambda }} E\log \left[ 1 + \lambda ^{\prime } x(x^{\prime } \delta + e) \right] \ge c_0 >0\) and Jensen’s inequality, there exists a constant \(c_0>0\), such that \(P\left( \frac{1}{2n} (Z_{H_0,k_0} -Z_{H_0,k} + Z_{H_1,k}) > c_0\right) \rightarrow 1\) as \(n\rightarrow \infty\). Thus, we have, \(P\left[ \left( -2 \log \Lambda _{k_0} - (-2 \log \Lambda _k)\right) >cn\right] \rightarrow 1\), since \(Z_{H_1,k_0} = O_p(1)\). So, \(-2\log \Lambda _k\) cannot arrive at its maximum with probability approaching to 1. By the definition of \(\hat{k}\), we have \(|\frac{k_0-\hat{k}}{n}|\le \eta\) with probability approaching to 1. Since \(\eta\) is arbitrary, thus the proof. \(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Piyadi Gamage, R.D., Ning, W. Empirical likelihood for change point detection in autoregressive models. J. Korean Stat. Soc. 50, 69–97 (2021). https://doi.org/10.1007/s42952-020-00061-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s42952-020-00061-w

Keywords

Navigation