Skip to main content
Log in

Test for conditional quantile change in general conditional heteroscedastic time series models

  • Published:
Annals of the Institute of Statistical Mathematics Aims and scope Submit manuscript

Abstract

This study aims to test for detecting a change point in the conditional quantile of general location-scale time series models. This issue is quite important in risk management because the conditional quantile is utilized to measure the value-at-risk or expected shortfall of financial assets. In this paper, we design two types of cumulative sum tests based on the conditional quantiles. Their limiting null distributions are derived under regularity conditions, together with consistency of the proposed tests under the alternative. Monte Carlo simulations demonstrate the good performance of the proposed tests in terms of both stability and power for various time series settings. A real data analysis using the daily returns of the Brent Oil futures also confirms the validity of the tests in real-world applications.

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

  • Berkes, I., Horváth, L., Kokoszka, P. (2004). Testing for parameter constancy in GARCH\((p, q)\) models. Statistics & Probability Letters, 70, 263–273.

    Article  MathSciNet  Google Scholar 

  • Billingsley, P. (1968). Convergence of probability measure. New York: Wiley.

    Google Scholar 

  • Bloomfield, P., Steiger, W. L. (1983). Least absolute deviations: Theory, applications, and algorithms. Boston: Birkhäuser.

    Google Scholar 

  • Bollerslev, T. (1986). Generalized autoregressive conditional heteroskedasticity. Journal of Econometrics, 31, 307–327.

    Article  MathSciNet  Google Scholar 

  • Bougerol, P., Picard, N. (1992). Stationarity of GARCH processes and of some nonnegative time series. Journal of Econometrics, 52, 115–127.

    Article  MathSciNet  Google Scholar 

  • Campbell, S. D. (2006). A review of backtesting and backtesting procedures. The Journal of Risk, 9, 1–17.

    Article  Google Scholar 

  • Chen, J., Gupta, A. K. (2012). Parametric statistical change point analysis with applications to genetics, medicine, and finance. New York: Wiley.

    Book  Google Scholar 

  • Christoffersen, P. F. (1998). Evaluating interval forecasts. International Economic Review, 39, 841–862.

    Article  MathSciNet  Google Scholar 

  • Ciuperca, G. (2017). Real time change-point detection in a nonlinear quantile model. Sequential Analysis, 36, 87–110.

    Article  MathSciNet  Google Scholar 

  • Csörgő, M., Horváth, L. (1997). Limit theorems in change-point analysis. New York: Wiley.

    Google Scholar 

  • de Pooter, M., van Dijk, D. (2004). Testing for changes in volatility in heteroskedastic time series—A further examination. Econometric Institute Research Papers EI 2004-38, Erasmus University Rotterdam, Erasmus School of Economics (ESE), Econometric Institute.

  • Engle, R. F. (1982). Autoregressive conditional heteroscedasticity with estimates of the variance of United Kingdom inflation. Econometrica, 50, 987–1007.

    Article  MathSciNet  Google Scholar 

  • Engle, R., Manganelli, S. (2004). CAViaR: Conditional autoregressive value at risk by regression quantiles. Journal of Business & Economic Statistics, 22, 367–381.

    Article  MathSciNet  Google Scholar 

  • Fitzenberger, B., Koenker, R., Machado, J. (2013). Economic applications of quantile regression. Berlin, Heidelberg: Springer.

    Google Scholar 

  • Francq, C., Zakoian, J.-M. (2004). Maximum likelihood estimation of pure GARCH and ARMA-GARCH processes. Bernoulli, 10, 605–637.

    Article  MathSciNet  Google Scholar 

  • Gombay, E. (2008). Change detection in autoregressive time series. Journal of Multivariate Analysis, 99, 51–464.

    Article  MathSciNet  Google Scholar 

  • Hall, P., Heyde, C. C. (1980). Martingale limit theory and its application. San Diego: Academic Press.

    Google Scholar 

  • Inclán, C., Tiao, G. C. (1994). Use of cumulative sums of squares for retrospective detection of changes of variance. Journal of the American Statistical Association, 89, 913–923.

    MathSciNet  Google Scholar 

  • Kim, M., Lee, S. (2016). Nonlinear expectile regression with application to value-at-risk and expected shortfall estimation. Computational Statistics & Data Analysis, 94, 1–19.

    Article  MathSciNet  Google Scholar 

  • Kim, S., Cho, S., Lee, S. (2000). On the cusum test for parameter changes in GARCH\((1,1)\) models. Communications in Statistics-Theory and Methods, 29, 445–462.

    Article  MathSciNet  Google Scholar 

  • Kirch, C., Kamgaing, J. (2012). Testing for parameter stability in nonlinear autoregressive models. Journal of Time Series Analysis, 33, 365–385.

    Article  MathSciNet  Google Scholar 

  • Koenker, R., Bassett, Jr., G. (1978). Regression quantiles. Econometrica: Journal of the Econometric Society, 46, 33–50.

    Article  MathSciNet  Google Scholar 

  • Koenker, R., Zhao, Q. (1996). Conditional quantile estimation and inference for ARCH models. Econometric Theory, 12, 793–813.

    Article  MathSciNet  Google Scholar 

  • Kupiec, P. (1995). Techniques for verifying the accuracy of risk measurement models. The Journal of Derivatives, 3, 73–84.

    Article  Google Scholar 

  • Lee, S. (2020). Location and scale-based CUSUM test with application to autoregressive models. Journal of Statistical Computation and Simulation, 90, 2309–2328.

    Article  MathSciNet  Google Scholar 

  • Lee, S., Kim, C. K. (2022). Test for conditional quantile change in GARCH models. Journal of the Korean Statistical Society, 51, 480–499.

    Article  ADS  MathSciNet  Google Scholar 

  • Lee, S., Noh, J. (2010). Value at risk forecasting based on quantile regression for GARCH models. The Korean Journal of Applied Statistics, 23, 669–681.

    Article  Google Scholar 

  • Lee, S., Noh, J. (2013). Quantile regression estimator for GARCH models. Scandinavian Journal of Statistics, 40, 2–20.

    Article  MathSciNet  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 

  • Lee, S., Tokutsu, Y., Maekawa, K. (2004). The cusum test for parameter change in regression models with ARCH errors. Journal of Japan Statistical Society, 34, 173–188.

    Article  MathSciNet  Google Scholar 

  • Meitz, M., Saikkonen, P. (2008). Ergodicity, mixing, and existence of moments of a class of Markov models with applications to GARCH and ACD models. Econometric Theory, 24, 1291–1320.

    Article  MathSciNet  Google Scholar 

  • Noh, J., Lee, S. (2016). Quantile regression for location-scale time series models with conditional heteroscedasticity. Scandinavian Journal of Statistics, 43, 700–720.

    Article  MathSciNet  Google Scholar 

  • Oh, H., Lee, S. (2018). On score vector-and residual-based CUSUM tests in ARMA-GARCH models. Statistical Methods & Applications, 27, 385–406.

    Article  MathSciNet  Google Scholar 

  • Oh, H., Lee, S. (2019). Modified residual CUSUM test for location-scale time series models with heteroscedasticity. Annals of Institute of Statistical Mathematics, 71, 1059–1091.

    Article  MathSciNet  Google Scholar 

  • Oka, T., Qu, Z. (2011). Estimating structural changes in regression quantiles. Journal of Econometrics, 162, 248–267.

    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  CAS  Google Scholar 

  • Qu, Z. (2008). Testing for structural change in regression quantiles. Journal of Econometrics, 146, 170–184.

    Article  MathSciNet  Google Scholar 

  • Straumann, D., Mikosch, T. (2006). Quasi-maximum-likelihood estimation in conditionally heteroscedastic time series: A stochastic recurrence equations approach. The Annals of Statistics, 34, 2449–2495.

    Article  MathSciNet  Google Scholar 

  • Su, L., Xiao, Z. (2008). Testing for parameter stability in quantile regression models. Statistics & Probability Letters, 78, 2768–2775.

    Article  MathSciNet  Google Scholar 

  • Weiss, A. (1991). Estimating nonlinear dynamic models using least absolute error estimation. Econometric Theory, 7, 46–68.

    Article  MathSciNet  Google Scholar 

  • Xiao, Z., Koenker, R. (2009). Conditional quantile estimation for generalized autoregressive conditional heteroscedasticity models. Journal of the American Statistical Association, 104, 1696–1712.

    Article  MathSciNet  Google Scholar 

  • Zhou, M., Wang, H. J., Tang, Y. (2015). Sequential change point detection in linear quantile regression models. Statistics & Probability Letters, 100, 98–103.

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

We sincerely thank the Editor, an AE, and anonymous reviewers for their careful reading and valuable comments that improve the quality of the paper. This research is supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT and Future Planning (No. 2021R1A2C1004009).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sangyeol Lee.

Additional information

Publisher's Note

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

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file 1 (pdf 142 KB)

Appendix

Appendix

In this section, we provide the proofs of some results in Sects. 2 and 3.

Proof of Theorem 1

It suffices to prove (9). Put

$$\begin{aligned} \varLambda _t (\vartheta ) =({{\tilde{g}}}_t (\mu )-g_t (\mu _0 ))/h_t^{1/2} (\theta _0) +\xi ({{\tilde{h}}}_t (\theta ) / h_t (\theta _0))^{1/2}-\xi _0. \end{aligned}$$

Noting that \(\psi _\tau (y_t-{\hat{q}}_t)=\tau -I(\eta _t \le \xi _0 +\varLambda _t({\hat{\vartheta _n}} ) )\), we can express

$$\begin{aligned}{} & {} \frac{1}{\sqrt{n}}\max _{1\le k\le n}\Bigg |\sum _{t=1}^k |y_{t-1}|^\gamma \Big \{\psi _\tau (y_t-{\hat{q}}_t)- \psi _\tau (y_t-q_t(\vartheta _0 ))\Big \}\\{} & {} \qquad -\,\frac{k}{n}\sum _{t=1}^n |y_{t-1}|^\gamma \Big \{\psi _\tau (y_t-{\hat{q}}_t)-\psi _\tau (y_t-q_t(\vartheta _0))\Big \}\Bigg |\\{} & {} \quad \le \frac{2}{\sqrt{n}}\max _{1\le k\le n}\Bigg |\sum _{t=1}^k |y_{t-1}|^\gamma \Big \{ I(\eta _t \le \xi _0 +\varLambda _t ({\hat{\vartheta _n}}) ) -F_\eta (\xi _0 +\varLambda _t ({\hat{\vartheta _n}}) )+F_\eta (\xi _0 )-I(\eta _t \le \xi _0)\Big \}\Bigg |\\{} & {} \qquad +\,\frac{1}{\sqrt{n}}\max _{1\le k\le n}\Bigg |\sum _{t=1}^k |y_{t-1}|^\gamma \Big \{ F_\eta (\xi _0 +\varLambda _t ({\hat{\vartheta _n}}) )-F_\eta (\xi _0 )\Big \}\\{} & {} \qquad -\,\frac{k}{n}\sum _{t=1}^n |y_{t-1}|^\gamma \Big \{ F_\eta (\xi _0 +\varLambda _t ({\hat{\vartheta _n}}) )-F_\eta (\xi _0 )\Big \}\Bigg |\\{} & {} \quad :=2 A_n +B_n. \end{aligned}$$

Owing to Proposition 1, for any \(\delta \in (0,1)\), there exists a (large enough) \(L>0\) such that \(P(\hat{\vartheta }_{n}\in \mathcal{N}_{L/\sqrt{n}})\ge 1-\delta\), where \({{{\mathcal {N}}}}_{L/\sqrt{n}}\) is a compact neighborhood of \(\vartheta _{0}\) with \(||\vartheta -\vartheta _{0}||\le L/\sqrt{n}\) for all \(\vartheta \in {{{\mathcal {N}}}}_{L/\sqrt{n}}\). Given any fixed \(\zeta >0\), we decompose \({{{\mathcal {N}}}}_{L/\sqrt{n}}\) into a finite number of subsets \(\mathcal{D}_1,\ldots ,{{{\mathcal {D}}}}_N\) for some \(N=N(\zeta )\ge 1\), with their diameters less than \(\zeta /\sqrt{n}\). We then choose points \(\vartheta _j\) from \({{{\mathcal {D}}}}_j\). Then, in case of \({\hat{\vartheta _n}} \in {{{\mathcal {D}}}}_j\), we have

$$\begin{aligned} I({\eta }_{t}\le \xi _0+ \varLambda _t (\vartheta _j )+\varLambda _{tj}^{-})\le I({\eta }_{t}\le \xi _0+\varLambda _{t} ({\hat{\vartheta _n}}))\le I({\eta }_{t}\le \xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^+ ) \end{aligned}$$

with \(\varLambda _{tj}^{-}=\inf _{\vartheta \in {{{\mathcal {D}}}}_j} \varLambda _t (\vartheta )-\varLambda _t (\vartheta _j)\) and \(\varLambda _{tj}^{+}=\sup _{\vartheta \in {{{\mathcal {D}}}}_j} \varLambda _t (\vartheta )-\varLambda _t (\vartheta _j )\). Putting

$$\begin{aligned} A_{kj}^+= & {} \frac{1}{\sqrt{n}}\sum _{t=1}^{k}|y_{t-1}|^\gamma \Big \{I({\eta }_{t}\le \xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^+)\\{} & {} -\,F_\eta (\xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^+)-F_\eta (\xi _0)+I(\eta _t\le \xi _0)\Big \},\\ A_{kj}^-= & {} \frac{1}{\sqrt{n}}\sum _{t=1}^{k}|y_{t-1}|^\gamma \Big \{I({\eta }_{t}\le \xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^-)\\{} & {} -\,F_\eta (\xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^-)-F_\eta (\xi _0)+I(\eta _t\le \xi _0)\Big \}, \end{aligned}$$

and using (A5), (A6), and the mean value theorem, we can show that

$$\begin{aligned} A_n \le \max _{1\le j\le N} \max _{1\le k\le n} |A_{kj}^+|+\max _{1\le j\le N} \max _{1\le k\le n} |A_{kj}^-|+r_n \end{aligned}$$
(12)

with

$$\begin{aligned} r_n= & {} \frac{1}{\sqrt{n}}\sum _{t=1}^{n}\max _{1\le j\le N} \sup _{\vartheta \in {{{\mathcal {D}}}}_j}|y_{t-1}|^\gamma | F_\eta (\xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^+)-F_\eta (\xi _0+ \varLambda _t (\vartheta ))|\nonumber \\{} & {} +\,\frac{1}{\sqrt{n}}\sum _{t=1}^{n}\max _{1\le j\le N} \sup _{\vartheta \in {{{\mathcal {D}}}}_j}|y_{t-1}|^\gamma F_\eta (\xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^-)-F_\eta (\xi _0+ \varLambda _t (\vartheta ))|\nonumber \\= & {} \zeta O_P (1), \end{aligned}$$
(13)

which can be made arbitrarily small by taking a small \(\zeta\).

Next, putting

$$\begin{aligned} e_{tj}^+= & {} |y_{t-1}|^\gamma \Big \{I({\eta }_{t}\le \xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^+)\\{} & {} -\,F_\eta (\xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^+)-F_\eta (\xi _0)+I(\eta _t\le \xi _0)\Big \},\\ e_{tj}^-= & {} |y_{t-1}|^\gamma \Big \{I({\eta }_{t}\le \xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^-)\\{} & {} -\,F_\eta (\xi _0+\varLambda _t (\vartheta _j)+\varLambda _{tj}^-)-F_\eta (\xi _0)+I(\eta _t\le \xi _0)\Big \}, \end{aligned}$$

we set \(S_{kj}^+=\sum _{t=1}^{k}e_{tj}^+\) and \(S_{kj}^-=\sum _{t=1}^{k}e_{tj}^-\), Moreover, for \(B>0\), we set

$$\begin{aligned} {{\tilde{S}}}_{kj}^+=\sum _{t=1}^{k}{{\tilde{e}}}_{tj}^+, \ \ \tilde{S}_{kj}^-=\sum _{t=1}^{k}{{\tilde{e}}}_{tj}^- \end{aligned}$$

with

$$\begin{aligned} {{\tilde{e}}}_{tj}^+= & {} e_{tj}^+ I \Bigg (\sum _{s=1}^{t} |y_{s}|^{2\gamma } \sup _{\vartheta \in \varTheta }\Big |\Big | \frac{\partial \varLambda _s (\vartheta )}{\partial \vartheta }\Big |\Big |\le B n\Bigg ),\\ {{\tilde{e}}}_{tj}^-= & {} e_{tj}^- I \Bigg (\sum _{s=1}^{t} |y_{s}|^{2\gamma } \sup _{\vartheta \in \varTheta }\Big |\Big | \frac{\partial \varLambda _s (\vartheta )}{\partial \vartheta }\Big |\Big |\le B n\Bigg ). \end{aligned}$$

Then, we get martingale arrays \(\{({{\tilde{S}}}_{kj}^+, {{{\mathcal {F}}}}_t); k=1,\ldots ,n\}\) and \(\{({{\tilde{S}}}_{kj}^-, {{{\mathcal {F}}}}_t); k=1,\ldots , n\}\), where \({{{\mathcal {F}}}}_t=\sigma (\eta _t, \eta _{t-1},\ldots )\).

Note that \(P ({{\tilde{e}}}_{tj}^+ \ne e_{tj}^+ \ \ \mathrm{for\ some}\ t=1,\ldots , n)\) can be made arbitrarily small by taking a large B, which can be seen using the mean value theorem, Hölder’s inequality, (A5) and (A6). Then, for any \(\zeta ^{'}, \lambda >0\), \(0<\delta <1\) and \(p>1\), using the sub-martingale inequality and taking a sufficiently large \(B:= B(\zeta )\), we can get

$$\begin{aligned}{} & {} \limsup _{n\rightarrow \infty }P(A_n^+\ge \lambda )\nonumber \\{} & {} \quad \le \limsup _{n\rightarrow \infty }P\Big (\max _{1\le j\le N}\max _{1\le k\le n} |A_{kj}^+|\ge \lambda \Big )\le \limsup _{n\rightarrow \infty } \sum _{j=1}^N P\Big (\max _{n^\delta \le k\le n}|S_{kj}^+|\ge \sqrt{n}\lambda \Big )\nonumber \\{} & {} \quad \le \limsup _{n\rightarrow \infty }\sum _{j=1}^NP\Big (\max _{1 \le k\le n}|{{\tilde{S}}}_{kj}^+|\ge \sqrt{n}\lambda \Big )+\zeta ^{'} \le \limsup _{n\rightarrow \infty } \sum _{j=1}^N E \tilde{S}_{nj}^{+2p}/n^{p}\lambda ^{2p}+\zeta ^{'}. \end{aligned}$$
(14)

Moreover, using Rosenthal’s inequality (Hall and Heyde 1980), we get

$$\begin{aligned} E{{\tilde{S}}}_{nj}^{+2p}\le & {} C\Big (E\Big [\sum _{t=1}^{n}E({{\tilde{e}}}_{tj}^{+2}|\mathcal{F}_{t-1})\Big ]^{p}+\sum _{t=1}^{n}E({{\tilde{e}}}_{tn}^{+2p})\Big ), \ C>0, \end{aligned}$$

and further,

$$\begin{aligned} E\Bigg [\sum _{t=1}^{n}E({{\tilde{e}}}_{tj}^{+2}|\mathcal{F}_{t-1})\Bigg ]^{p}=O(n^{p/2})\quad \textrm{and}\quad \sum _{t=1}^{n}E(\tilde{e}_{tj}^{+2p})= O(n), \end{aligned}$$

which leads to \(\sum _{j=1}^N E{{\tilde{S}}}_{nj}^{+2p}=O(n^{p/2}+n )\). Then, in view of (14), we obtain

$$\begin{aligned} \max _{1\le j\le N}\max _{1\le k\le n} |A_{kj}^+|=o_P (1). \end{aligned}$$

Analogously, we have \(\max _{1\le j\le N}\max _{1\le k\le n} |A_{kj}^-|=o_P (1).\) Hence, from (12) and (13), we have \(A_n= o_P (1)\).

Next, we handle \(B_n\). Set

$$\begin{aligned} B_{nk}= & {} \Bigg |\sum _{t=1}^k |y_{t-1}|^\gamma \Big \{ F_\eta (\xi _0 +\varLambda _t ({\hat{\vartheta _n}}) )-F_\eta (\xi _0 )\Big \}\\{} & {} -\,\frac{k}{n}\sum _{t=1}^n |y_{t-1}|^\gamma \Big \{ F_\eta (\xi _0 +\varLambda _t ({\hat{\vartheta _n}}) )-F_\eta (\xi _0 )\Big \}\Bigg |. \end{aligned}$$

Note that for \(k_n = n^{1/3}\), \(\frac{1}{\sqrt{n}}\max _{1\le k\le k_n} B_{nk}=o_P (1).\) Thus it suffices to show that

$$\begin{aligned} B_n^{'}:=\frac{1}{\sqrt{n}}\max _{k_n\le k\le n}B_{nk}=o_P (1). \end{aligned}$$

To this end, we use Taylor’s theorem to get

$$\begin{aligned} F_\eta (\xi _0 +\varLambda _t ({\hat{\vartheta _n}}) )-F_\eta (\xi _0 )=\varLambda _t ({\hat{\vartheta _n}}) f_\eta (\xi _0) +\frac{1}{2} \varLambda _t^2 ({\hat{\vartheta _n}}) f^{'}_\eta (\xi _t^*) \end{aligned}$$

for some number \(\xi _t^*\) lying between \(\xi _0\) and \(\xi _0 +\varLambda _t ({\hat{\vartheta _n}})\). Then, using this, (A2), (A5), (A6), and our assumption on \(f_\eta ^{'}\), we can show that for some \(K>0\),

$$\begin{aligned} B_n^{'}\le & {} K \sqrt{n}||{\hat{\mu _n}} -\mu _0||\max _{k_n\le k\le n} \Bigg |\Bigg | \frac{1}{k}\sum _{t=1}^k |y_{t-1}|^\gamma \frac{\partial g_t (\mu _0 )}{\partial \mu }\frac{1}{h^{1/2}_t (\theta _0 )}\\{} & {} -\,\frac{1}{n}\sum _{t=1}^n |y_{t-1}|^\gamma \frac{\partial g_t (\mu _0 )}{\partial \mu }\frac{1}{h^{1/2}_t (\theta _0 )}\Bigg |\Bigg |\\{} & {} +\, K\sqrt{n}||{\hat{\theta _n}} -\theta _0|| \max _{k_n\le k\le n}\Bigg |\Bigg | \frac{1}{k}\sum _{t=1}^k |y_{t-1}|^\gamma \frac{\partial h_t (\theta _0 )}{\partial \theta }\frac{1}{\sqrt{2}h_t (\theta _0 )}\\{} & {} -\,\frac{1}{n}\sum _{t=1}^n |y_{t-1}|^\gamma \frac{\partial h_t (\theta _0 )}{\partial \theta }\frac{1}{\sqrt{2}h_t (\theta _0 )}\Bigg |\Bigg |\\{} & {} +\, K\sqrt{n}|{\hat{\xi _n}} -\xi _0| \max _{k_n\le k\le n}\Big |\frac{1}{k}\sum _{t=1}^k |y_{t-1}|^\gamma -\frac{1}{n}\sum _{t=1}^n |y_{t-1}|^\gamma \Big |+o_P (1)\\= & {} o_P (1), \end{aligned}$$

due to the ergodicity of \(\{y_t\}\), which entails \(B_n=o_P (1)\). As we can check \({\hat{\tau }}_{1}^2=\tau _1^2+o_P (1)\), we finally get \({\hat{T}}_{n}- T_{n}=o_P (1)\), which in turn implicates the weak convergence of \({\hat{T}}_{n}\) to \(\sup _{0\le s\le 1} | B^\circ (s)|\). \(\square\)

Proof of Theorem 4

For \(i=1, 2\), we set

$$\begin{aligned} \varLambda _{ti}^* (\vartheta )= & {} \frac{{{\tilde{g}}}_{ti} (\mu )-g_{ti} (\mu _0^* )}{h_{ti}^{1/2} (\theta _{i})} +\frac{ \xi \tilde{h}_{ti}^{1/2}(\theta ) -\xi _{0}^* h_{ti}^{1/2}(\theta _{0}^*)}{ h_{ti}^{1/2} (\theta _{i})}, \\ \xi _{ti}= & {} \frac{ g_{ti}(\mu _0^* )-g_{ti}(\mu _{i})+\xi _0^* h_{ti}^{1/2}(\theta _0^*)}{h_{ti}^{1/2}(\theta _{i})}. \end{aligned}$$

Noting that \(\psi _\tau (y_{t1}-{\hat{q}}_{t1})=\tau -I(\eta _t \le \varLambda ^*_t ({\hat{\vartheta _n}} )+\xi _{t1} )\), we can express for \(k\le k_0\),

$$\begin{aligned} \frac{1}{n}\sum _{t=1}^k |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}-{\hat{q}}_{t1})=\frac{1}{n}\sum _{t=1}^k |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}- q_{t1}(\vartheta _0^*))+I_{nk} +II_{nk}, \end{aligned}$$

where

$$\begin{aligned} I_{nk}= & {} -\frac{1}{n}\sum _{t=1}^k |y_{t-1,1}|^\gamma \Big \{ I(\eta _t \le \varLambda _{t1}^*({\hat{\vartheta _n}})+\xi _{t1} ) -F_\eta (\varLambda _{t1}^*({\hat{\vartheta _n}})\\{} & {} +\,\xi _{t1} )+F_\eta (\xi _{t1} )-I(\eta _t \le \xi _{t1})\Big \},\\ II_{nk}= & {} -\frac{1}{n}\sum _{t=1}^k |y_{t-1,1}|^\gamma \Big \{ F_\eta ( \varLambda _{t1}^*({\hat{\vartheta _n}})+\xi _{t1} )-F_\eta (\xi _{t1} )\Big \}. \end{aligned}$$

Then, following the lines similar to those used for verifying \(A_n=o_P (1)\) in the proof of Theorem 1, with \(\mathcal{N}_{L/\sqrt{n}}\), \(\varLambda _t (\vartheta )\), and \(\xi _0\), therein, replaced by \({{{\mathcal {N}}}}_L^*=\{\vartheta : ||\vartheta -\vartheta _0^*||\le L\}\), \(\varLambda _{t1}^{*}(\vartheta )\), and \(\xi _{t1}\), respectively, where L can be taken to be arbitrarily small, \({{{\mathcal {N}}}}_L^*\) and L play the role of the partition \({{{\mathcal {D}}}}_1\) and \(\zeta\) (namely, \({{{\mathcal {N}}}}_L^*=\mathcal{D}_1\), \(L=\zeta\), and \(N(\zeta )=1\)), and both \(A_{k1}^+\) and \(A_{k1}^-\) are accordingly reformulated with the denominator \(\sqrt{n}\) replaced by n, we can similarly verify that \(\max _{1\le k\le k_0} | I_{nk}|= o_P (1)\). Also, using (A2), (A5), (A6) (modified with \(g_{ti}, h_{ti}, \tilde{g}_{ti}, {{\tilde{h}}}_{ti}\)), (B1), the ergodicity, and the mean value theorem, we can readily check that \(\max _{1\le k\le k_0} | II_{nk}|= o_P (1)\). Subsequently, we get

$$\begin{aligned} \frac{1}{n}\sum _{t=1}^k |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}-{\hat{q}}_{t1})= & {} \frac{1}{n}\sum _{t=1}^k |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}- q_{t1}(\vartheta _0^*))\nonumber \\{} & {} +\,o_P (1), \end{aligned}$$
(15)

uniformly in \(k\le k_0\). Similarly, additionally harnessing (11), we can verify

$$\begin{aligned} \frac{1}{n}\sum _{t=k_0+1}^k |y_{t-1,2}|^\gamma \psi _\tau (y_{t2}-{\hat{q}}_{t2})= & {} \frac{1}{n}\sum _{t=k_0+1}^n |y_{t-1,2}|^\gamma \psi _\tau (y_{t2}- q_{t2}(\vartheta _0^*))\nonumber \\{} & {} +\,o_P (1), \end{aligned}$$
(16)

uniformly in \(k>k_0\). Therefore, we can have

$$\begin{aligned} \frac{|{\hat{\tau _n}} {\hat{T}}_n |}{\sqrt{n}} &\ge \frac{1}{n}\Big |\sum _{t=1}^{k_0} |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}-{\hat{q}}_{t1})\nonumber \\{} & {} \qquad - \frac{k_0}{n} \Big \{ \sum _{t=1}^{k_0} |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}-{\hat{q}}_{t1})+\sum _{t=k_0 +1}^n |y_{t-1,2}|^\gamma \psi _\tau (y_{t2}-{\hat{q}}_{t2})\Big \}\Big |\nonumber \\{} & =\frac{1}{n}\Big |\sum _{t=1}^{k_0} |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}- q_{t1}(\vartheta _0^*))\nonumber \\{} & {} \qquad -\frac{k_0}{n} \Big \{ \sum _{t=1}^{k_0} |y_{t-1,1}|^\gamma \psi _\tau (y_{ti}- q_{t1}(\vartheta _0^*))\nonumber \\{} & {} \qquad +\sum _{t=k_0 +1}^n |y_{t-1,2}|^\gamma \psi _\tau (y_{t2}- q_{t2}(\vartheta _0^*))\Big \}\Big |+o_P (1)\nonumber \\{} & {} \quad {\mathop {\longrightarrow }\limits ^{P}} \kappa _0(1-\kappa _0)|d_1-d_2| > 0, \end{aligned}$$

due to the ergodicity and (B2), so that \({\hat{T}}_n \rightarrow \infty\) in probability as \({\hat{\tau _n}}=\tau _0+ o_P(1)\) for some \(\tau _0>0\).

Moreover, putting

$$\begin{aligned} \hat{{{\mathcal {L}}}}_n (k)= & {} \frac{1}{n}\Big |\sum _{t=1}^{k} |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}-{\hat{q}}_{t1})\\{} & {} -\, \frac{k}{n} \Big \{ \sum _{t=1}^{k_0} |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}-{\hat{q}}_{t1})+\sum _{t=k_0 +1}^n |y_{t-1,2}|^\gamma \psi _\tau (y_{t2}-{\hat{q}}_{t2})\Big \}\Big |, \end{aligned}$$

we can see that \(\max _{0\le s\le 1}|\hat{{{\mathcal {L}}}}_n ([ns])- \mathcal{L}_n ([ns])|= o_P (1)\) by virtue of (15) and (16), where

$$\begin{aligned} {{{\mathcal {L}}}}_n (k)= & {} \frac{1}{n}\Big |\sum _{t=1}^{k} |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}- q_{t1}(\vartheta _0^*))I(k\le k_0) \\{} & {} +\, \Big \{\sum _{t=1}^{k_0} |y_{t-1,1}|^\gamma \psi _\tau (y_{t1}- q_{t1}(\vartheta _0^*))+ \sum _{t=k_0+1}^{k} |y_{t-1,2}|^\gamma \psi _\tau (y_{t2}- q_{t2}(\vartheta _0^*))\Big \}I(k>k_0)\\{} & {} -\, \frac{k}{n} \Big \{ \sum _{t=1}^{k_0} |y_{t-1,1}|^\gamma \psi _\tau (y_{ti}- q_{t1}(\vartheta _0^*))+\sum _{t=k_0 +1}^n |y_{t-1,2}|^\gamma \psi _\tau (y_{ti}- q_{t2}(\vartheta _0^*))\Big \}\Big |. \end{aligned}$$

Furthermore, using the ergodicity, we can easily check that \(\sup _{0\le s\le 1}|{{{\mathcal {L}}}}_n ([ns])-{{{\mathcal {L}}}}(s)|=o_P (1),\) where

$$\begin{aligned} {{{\mathcal {L}}}}(s)= s (1-\kappa _0)|d_1-d_2|I (s\le \kappa _0)+\kappa _0 (1-s) |d_1-d_2|I (\kappa _0 < s\le 1), \end{aligned}$$

which has a maximum value \(\kappa _0(1-\kappa _0)|d_1-d_2|\) uniquely at \(s=\kappa _0\). This with the argmax theorem yields \({\hat{k}}_n /n \rightarrow \kappa _0\) in probability, which validates the theorem. \(\square\)

On the issue regarding (B1) in Remark 4. For \(t\le k_0\), put

$$\begin{aligned} \varLambda _t^{\dagger } (\vartheta )= & {} ({{\tilde{g}}}_{t1} (\mu )-g_{t1} (\mu _1 ))/h_{t1}^{1/2} (\theta _1) +\xi ({{\tilde{h}}}_{t1} (\theta ) / h_{t1} (\theta _1))^{1/2},\\ \varLambda _t^\circ (\vartheta )= & {} ( g_{t1} (\mu )-g_{t1} (\mu _1 ))/h_{t1}^{1/2} (\theta _1) +\xi ( h_{t1} (\theta ) / h_{t1} (\theta _1))^{1/2}. \end{aligned}$$

Then, we can express

$$\begin{aligned} \rho _\tau (y_t - {{\tilde{q}}}_{t}(\vartheta ))= & {} (y_{t1} - \tilde{q}_{t1}(\vartheta ))\psi _\tau (y_{t1}-{{\tilde{q}}}_{t1}(\vartheta )),\\ \rho _\tau (y_{t} - q_{t}(\vartheta ))= & {} (y_{t1} - q_{t1}(\vartheta ))\psi _\tau (y_{t1}- q_{t1}(\vartheta )) \end{aligned}$$

with

$$\begin{aligned} \psi _\tau (y_t-{{\tilde{q}}}_{t}(\vartheta ))=\tau -I(\eta _t \le \varLambda _t^\dagger (\vartheta ) ),\quad \psi _\tau (y_t- q_{t1}(\vartheta ))=\tau -I(\eta _t \le \varLambda _t^\circ (\vartheta ) ). \end{aligned}$$

We first verify

$$\begin{aligned} \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=1}^{k_0} \rho _\tau (y_{t} - {{\tilde{q}}}_{t}(\vartheta ))- \frac{1}{n}\sum _{t=1}^{k_0} \rho _\tau (y_{t1} - q_{t1}(\vartheta ))\Big |=o_P (1). \end{aligned}$$
(17)

Due to (A5) and (A6), it suffices to show that

$$\begin{aligned} \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=1}^{k_0} (y_t - {{\tilde{q}}}_{t}(\vartheta ))I(\eta _t \le \varLambda _t^\dagger (\vartheta ) )- \frac{1}{n}\sum _{t=1}^{k_0} (y_{t1} - q_{t1}(\vartheta ))I(\eta _t \le \varLambda _t^\circ (\vartheta ) )\Big |=o_P (1), \end{aligned}$$

or equivalently, for \(\delta \in (0,1)\),

$$\begin{aligned} \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=n^\delta }^{k_0} (y_t - q_{t}(\vartheta ))I(\eta _t \le \varLambda _t^\dagger (\vartheta ) )- \frac{1}{n}\sum _{t=n^\delta }^{k_0} (y_{t1} - q_{t1}(\vartheta ))I(\eta _t \le \varLambda _t^\circ (\vartheta ) )\Big |=o_P (1), \end{aligned}$$

which we can check to hold via using the Cauchy–Schwarz inequality, provided

$$\begin{aligned} \sup _{\vartheta \in \varTheta ^*}\frac{1}{n}\sum _{t=n^\delta }^{k_0} (I(\eta _t \le \varLambda _t^\dagger (\vartheta ) )-I(\eta _t \le \varLambda _t^\circ (\vartheta ) )^2=o_P (1). \end{aligned}$$
(18)

To verify (18), we partition \(\varTheta ^*\) into \({{{\mathcal {D}}}}_j\) with diameter less than \(\zeta\), \(j=1,\ldots , N=N(\zeta )\), and choose points \(\vartheta _j\) from \({{{\mathcal {D}}}}_j\). Then, in case of \(\vartheta \in {{{\mathcal {D}}}}_j\), we have

$$\begin{aligned} I({\eta }_{t}\le \varLambda _t^\circ (\vartheta _j )+\varLambda _{tj}^{\dagger (1)})\le I({\eta }_{t}\le \varLambda _t^{\dagger } (\vartheta ))\le I({\eta }_{t}\le \varLambda _t^\circ (\vartheta _j)+\varLambda _{tj}^{\dagger (2)} ) \end{aligned}$$

with \(\varLambda _{tj}^{\dagger (1)}=\inf _{\vartheta \in {{{\mathcal {D}}}}_j} \varLambda _t^\dagger (\vartheta )-\varLambda _t^\circ (\vartheta _j)\) and \(\varLambda _{tj}^{\dagger (2)}=\sup _{\vartheta \in {{{\mathcal {D}}}}_j} \varLambda _t^\dagger (\vartheta )-\varLambda _t^\circ (\vartheta _j )\). Also,

$$\begin{aligned} I({\eta }_{t}\le \varLambda _t^\circ (\vartheta _j )+\varLambda _{tj}^{\circ (1)})\le I({\eta }_{t}\le \varLambda _t^{\circ } (\vartheta ))\le I({\eta }_{t}\le \varLambda _t^\circ (\vartheta _j)+\varLambda _{tj}^{\circ (2)} ) \end{aligned}$$

with \(\varLambda _{tj}^{\circ (1)}=\inf _{\vartheta \in {{{\mathcal {D}}}}_j} \varLambda _t^\circ (\vartheta )-\varLambda _t^\circ (\vartheta _j)\) and \(\varLambda _{tj}^{\circ (2)}=\sup _{\vartheta \in {{{\mathcal {D}}}}_j} \varLambda _t^\circ (\vartheta )-\varLambda _t^\circ (\vartheta _j )\).

By simple algebras using (A5) and (A6), it can be seen that

$$\begin{aligned} C_n:= & {} \max _{1\le j\le N}\max _{1\le k\ne l\le 2 }\frac{1}{n}\sum _{t=n^\delta }^{k_0} \Big [I(\eta _t \le \varLambda _t^\circ (\vartheta _j)+\varLambda _{tj}^{\dagger (k)} )-I(\eta _t \le \varLambda _t^\circ (\vartheta _j)+\varLambda _{tj}^{\circ (l)}) \Big ]^2\\= & {} \zeta O_P (1)+o_P (1). \end{aligned}$$

As \(\zeta\) can be taken to be arbitrarily small, \(C_n\) becomes \(o_P(1)\), which ensures (18) and thereby (17) as the quantity in (18) is no more than \(C_n\). Moreover, following the lines similar to the above, we can also check

$$\begin{aligned} \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=1}^{k_0} \rho _\tau (y_{t1} - q_{t1}(\vartheta ))- \kappa _0 E\rho _\tau (y_{t1} - q_{t1}(\vartheta ))\Big |=o_P (1), \end{aligned}$$

and thus, due to (17),

$$\begin{aligned} \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=1}^{k_0} \rho _\tau (y_{t} - {{\tilde{q}}}_{t}(\vartheta ))- \kappa _0 E\rho _\tau (y_{t1} - q_{t1}(\vartheta ))\Big |=o_P (1). \end{aligned}$$
(19)

Similarly, it can be shown that

$$\begin{aligned}{} & {} \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=k_0+1}^{n} \rho _\tau (y_{t} - {{\tilde{q}}}_{t}(\vartheta ))- (1-\kappa _0)\rho _\tau (y_{t2} - q_{t2}(\vartheta ))\Big |\nonumber \\{} & {} \quad \le \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=k_0+1}^{n} \rho _\tau (y_{t} - {{\tilde{q}}}_{t}(\vartheta ))- \frac{1}{n}\sum _{t=k_0+1}^{n} \rho _\tau (y_{t2} - q_{t2}(\vartheta ))\Big |\nonumber \\{} & {} \qquad + \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=k_0+1}^{n} \rho _\tau (y_{t2} - q_{t2}(\vartheta ))- (1-\kappa _0) E \rho _\tau (y_{t2} - q_{t2}(\vartheta ))\Big |\nonumber \\{} & {} \quad =o_P (1), \end{aligned}$$
(20)

then combining (19) and (20), we have

$$\begin{aligned}{} & {} \sup _{\vartheta \in \varTheta ^*}\Big |\frac{1}{n}\sum _{t=1}^{n} \rho _\tau (y_t - {{\tilde{q}}}_{t}(\vartheta ))- \big \{\kappa _0 E\rho _\tau (y_{t1}\\{} & {} \quad -\, q_{t1}(\vartheta ))+ (1-\kappa _0)E \rho _\tau (y_{t2} - q_{t2}(\vartheta ))\big \}\Big |= o_P (1), \end{aligned}$$

which validates the conclusion in Remark 4. Namely, (B1) holds true if \(\kappa _0 E\rho _\tau (y_{t1} - q_{t1}(\vartheta ))+ (1-\kappa _0)E \rho _\tau (y_{t2} - q_{t2}(\vartheta ))\) has its maximum value uniquely at an interior point \(\vartheta _0^*\) in \(\varTheta ^*\). \(\square\)

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lee, S., Kim, C.K. Test for conditional quantile change in general conditional heteroscedastic time series models. Ann Inst Stat Math 76, 333–359 (2024). https://doi.org/10.1007/s10463-023-00889-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10463-023-00889-z

Keywords

Navigation