Skip to main content

On Diagnostic Checking Autoregressive Conditional Duration Models with Wavelet-Based Spectral Density Estimators

  • Chapter
  • First Online:
Advances in Time Series Methods and Applications

Part of the book series: Fields Institute Communications ((FIC,volume 78))

  • 1186 Accesses

Abstract

There has been an increasing interest recently in the analysis of financial data that arrives at irregular intervals. An important class of models is the autoregressive Conditional Duration (ACD) model introduced by Engle and Russell (Econometrica 66:1127–1162, 1998, [22]) and its various generalizations. These models have been used to describe duration clustering for financial data such as the arrival times of trades and price changes. However, relatively few evaluation procedures for the adequacy of ACD models are currently available in the literature. Given its simplicity, a commonly used diagnostic test is the Box-Pierce/Ljung-Box statistic adapted to the estimated standardized residuals of ACD models, but its asymptotic distribution is not the standard one due to parameter estimation uncertainty. In this paper we propose a test for duration clustering and a test for the adequacy of ACD models using wavelet methods. The first test exploits the one-sided nature of duration clustering. An ACD process is positively autocorrelated at all lags, resulting in a spectral mode at frequency zero. In particular, it has a spectral peak at zero when duration clustering is persistent or when duration clustering is small at each individual lag but carries over a long distributional lag. As a joint time-frequency decomposition method, wavelets can effectively capture spectral peaks and thus are expected to be powerful. Our second test checks the adequacy of an ACD model by using a wavelet-based spectral density of the estimated standardized residuals over the whole frequency. Unlike the Box-Pierce/Ljung-Box tests, the proposed diagnostic test has a convenient asymptotic “nuisance parameter-free” property—parameter estimation uncertainty has no impact on the asymptotic distribution of the test statistic. Moreover, it can check a wide range of alternatives and is powerful when the spectrum of the standardized duration residuals is nonsmooth, as can arise from neglected persistent duration clustering, seasonality, calender effects and business cycles. For each of the two new tests, we propose and justify a suitable data-driven method to choose the finest scale—the smoothing parameter in wavelet estimation. This makes the methods fully operational in practice. We present a simulation study, illustrating the merits of the wavelet-based procedures. An application with tick-by-tick trading data of Alcoa stock is presented.

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

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.00
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.00
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  1. Admati, A. R., & Pfleiderer, P. (1988). A theory of intraday patterns: Volume and price variability. Review of Financial Studies, 1, 3–40.

    Article  Google Scholar 

  2. Andrews, D. W. K. (1991). Heteroskedasticity and autocorrelation consistent covariance matrix estimation. Econometrica, 59, 817–858.

    Article  MathSciNet  MATH  Google Scholar 

  3. Andrews, D. W. K. (2001). Testing when a parameter is on the boundary of the maintained hypothesis. Econometrica, 69, 683–734.

    Article  MathSciNet  MATH  Google Scholar 

  4. Bauwens, L., & Giot, P. (2000). The logarithmic ACD model: An application to the bid-ask quote process of three NYSE stocks. Annals of Economics and Statistics, 60, 117–149.

    Google Scholar 

  5. Bauwens, L., & Giot, P. (2001). Econometric modelling of stock market intraday activity. Advances studies in theoretical and applied econometrics. Boston: Kluwer.

    Book  MATH  Google Scholar 

  6. Bauwens, L., & Giot, P. (2003). Asymmetric ACD Models: Introducing price information in ACD models. Empir. Econ., 28, 709–731.

    Article  Google Scholar 

  7. Beltrao, K., & Bloomfield, P. (1987). Determining the bandwidth of a kernel spectrum estimate. Journal of Time Series Analysis, 8, 21–38.

    Article  MathSciNet  MATH  Google Scholar 

  8. Bera, A. K., & Higgins, M. L. (1992). A test for conditional heteroskedasticity in time series models. Journal of Time Series Analysis, 13, 501–519.

    Article  MathSciNet  MATH  Google Scholar 

  9. Bizer, D. S., & Durlauf, S. N. (1990). Testing the positive theory of government finance. Journal of Monetary Economics, 26, 123–141.

    Article  Google Scholar 

  10. Box, G. E. P., & Pierce, D. (1970). Distribution of residual autocorrelations in autoregressive integrated moving average time series models. Journal of the American Statistical Association, 65, 1509–1526.

    Article  MathSciNet  MATH  Google Scholar 

  11. Bühlmann, P. (1996). Locally adaptive lag-window spectral estimation. Journal of Time Series Analysis, 17, 247–270.

    Article  MathSciNet  MATH  Google Scholar 

  12. Daubechies, I. (1992). Ten lectures on wavelets. In CBS-NSF regional conferences in applied mathematics (Vol. 61). Philadelphia: Society for Industrial and Applied Mathematics.

    Google Scholar 

  13. Drost, F. C., & Werker, B. J. M. (2000). Efficient estimation in semiparametric time series: The ACD model. Working paper, Tilburg University.

    Google Scholar 

  14. Drost, F. C., & Werker, B. J. M. (2004). Semiparametric duration models. Journal of Business & Economic Statistics, 22, 40–50.

    Article  MathSciNet  Google Scholar 

  15. Duchesne, P. (2006a). On testing for serial correlation with a wavelet-based spectral density estimator in multivariate time series. Econometric Theory, 22, 633–676.

    Google Scholar 

  16. Duchesne, P. (2006b). Testing for multivariate autoregressive conditional heteroskedasticity using wavelets. Computational Statistics & Data Analysis, 51, 2142–2163.

    Google Scholar 

  17. Duchesne, P., Li, L., & Vandermeerschen, J. (2010). On testing for serial correlation of unknown form using wavelet thresholding. Computational Statistics & Data Analysis, 54, 2512–2531.

    Article  MathSciNet  MATH  Google Scholar 

  18. Duchesne, P., & Pacurar, M. (2008). Evaluating financial time series models for irregularly spaced data: A spectral density approach. Computers & Operations Research (Special Issue: Applications of OR in Finance), 35, 130–155.

    Article  MathSciNet  MATH  Google Scholar 

  19. Easley, D., & O’Hara, M. (1992). Time and the process of security price adjustment. The Journal of Finance, 19, 69–90.

    Google Scholar 

  20. Engle, R. F. (2000). The econometrics of ultra-high frequency data. Econometrica, 68, 1–22.

    Article  MATH  Google Scholar 

  21. Engle, R. F., & Russell, J. R. (1997). Forecasting the frequency of changes in quoted foreign exchange prices with the autoregressive conditional duration model. Journal of Empirical Finance, 4, 187–212.

    Article  Google Scholar 

  22. Engle, R. F., & Russell, J. R. (1998). Autoregressive conditional duration: A new model for irregularly spaced transaction data. Econometrica, 66, 1127–1162.

    Article  MathSciNet  MATH  Google Scholar 

  23. Escanciano, J. C. (2006). Goodness-of-fit tests for linear and nonlinear time series models. Journal of the American Statistical Association, 101, 531–541.

    Article  MathSciNet  MATH  Google Scholar 

  24. Fan, Y., & Gençay, R. (2010). Unit root tests with wavelets. Econometric Theory, 26, 1305–1331.

    Article  MathSciNet  MATH  Google Scholar 

  25. Francq, C., Roy, R., & Zakoïan, J.-M. (2005). Diagnostic checking in ARMA models with uncorrelated errors. Journal of the American Statistical Association, 100, 532–544.

    Article  MathSciNet  MATH  Google Scholar 

  26. Francq, C., & Zakoïan, J.-M. (2009). Testing the nullity of GARCH coefficients: Correction of the standard tests and relative efficiency comparisons. Journal of the American Statistical Association, 104, 313–324.

    Article  MathSciNet  MATH  Google Scholar 

  27. Fernandes, M., & Grammig, J. (2005). Non-parametric specification tests for conditional duration models. Journal of Econometrics, 127, 35–68.

    Article  MathSciNet  MATH  Google Scholar 

  28. Gao, H. (1993). Wavelet estimation of spectral densities in time series analysis. Ph.D Dissertation, Department of Statistics, University of California, Berkeley.

    Google Scholar 

  29. Gençay, R., & Signori, D. (2012). Multi-scale tests for serial correlation. Technical report, Department of Economics, Simon Fraser University.

    Google Scholar 

  30. Gençay, R., Yazgan, E., & Ozkan, H. (2012). A test of structural change of unknown location with wavelets. Technical report, Department of Economics, Simon Fraser University.

    Google Scholar 

  31. Ghysels, E., & Jasiak, J. (1994). Stochastic volatility and time deformation: an application to trading volume and leverage effects. Working paper, C.R.D.E., Université de Montréal.

    Google Scholar 

  32. Ghysels, E., & Jasiak, J. (1998). Long-term dependence in trading. Working paper, Dept. of Economics, Penn State University and York University.

    Google Scholar 

  33. Ghysels, E., & Jasiak, J. (1998). GARCH for irregularly spaced financial data: The ACD-GARCH model. Studies in Nonlinear Dynamics and Econometrics, 2, 133–149.

    MATH  Google Scholar 

  34. Grammig, J., Hujer, R., Kokot, S., & Maurer, K.-O. (1998). Modeling the Deutsche Telekom IPO using a new ACD specification, an application of the Burr-ACD model using high frequency IBIS data. Working paper, Department of Economics, Johann Wolfgang Goethe-University of Frankfurt.

    Google Scholar 

  35. Granger, C. (1966). The typical spectral shape of an economic variable. Econometrica, 34, 150–161.

    Article  Google Scholar 

  36. Granger, C. W. J., & Newbold, P. (1986). Forecasting economic time series (2nd ed.). New York: Academic Press.

    MATH  Google Scholar 

  37. Hall, P., & Heyde, C. C. (1980). Martingale limit theory and its application. New York: Academic Press.

    MATH  Google Scholar 

  38. Hannan, E. (1970). Multiple time series. New York: Wiley.

    Book  MATH  Google Scholar 

  39. Hautsch, N. (2006). Testing the conditional mean function of autoregressive conditional duration models. Working Paper, University of Copenhagen.

    Google Scholar 

  40. Hernandez, E., & Weiss, G. (1996). A first course on wavelets. New York: CRC Press.

    Book  MATH  Google Scholar 

  41. Higgins, M. L. (2000). Testing for autoregressive conditional duration. Presented at World Congress of Econometric Society, Seattle.

    Google Scholar 

  42. Hong, Y. (2001). Wavelet-based estimation for heteroskedasticity and autocorrelation consistent variance-covariance matrices. Working paper, Department of Economics and Department of Statistical Science, Cornell University.

    Google Scholar 

  43. Hong, Y., & Kao, C. (2004). Wavelet-based testing for serial correlation of unknown form in panel models. Econometrica, 72, 1519–1563.

    Article  MathSciNet  MATH  Google Scholar 

  44. Hong, Y., & Lee, J. (2001). One-sided testing for ARCH effect using wavelets. Econometric Theory, 17, 1051–1081.

    MathSciNet  MATH  Google Scholar 

  45. Hong, Y., & Lee, Y.-J. (2011). Detecting misspecifications in autoregressive conditional duration models and non-negative time-series processes. Journal of Time Series Analysis, 32, 1–32.

    Article  MathSciNet  MATH  Google Scholar 

  46. Jasiak, J. (1999). Persistence in intertrade durations. Finance, 19, 166–195.

    Google Scholar 

  47. Jensen, M. J. (2000). An alternative maximum likelihood estimator of long memory processes using compactly supported wavelets. Journal of Economic Dynamics and Control, 24, 361–387.

    Article  MathSciNet  MATH  Google Scholar 

  48. Kyle, A. (1985). Continuous time auctions and insider trading. Econometrica, 53, 1315–1336.

    Article  MATH  Google Scholar 

  49. Lee, J. H. H., & King, M. L. (1993). A locally most mean powerful based score test for ARCH and GARCH regression disturbances. Journal of Business & Economic Statistics, 11, 17–27 (Correction 1994, 12, p. 139).

    Google Scholar 

  50. Lee, J., & Hong, Y. (2001). Testing for serial correlation of unknown form using wavelet methods. Econometric Theory, 17, 386–423.

    Article  MathSciNet  MATH  Google Scholar 

  51. Li, L., Yao, S., & Duchesne, P. (2014). On wavelet-based testing for serial correlation of unknown form using Fan’s adaptive Neyman method. Computational Statistics & Data Analysis, 70, 308–327.

    Article  MathSciNet  Google Scholar 

  52. Li, W. K. (2004). Diagnostic checks in time series. New York: Chapman & Hall/CRC.

    MATH  Google Scholar 

  53. Li, W. K., & Mak, T. K. (1994). On the square residual autocorrelations in non-linear time series with conditional heteroskedasticity. Journal of Time Series Analysis, 15, 627–636.

    Article  MathSciNet  MATH  Google Scholar 

  54. Li, W. K., & Yu, L. H. (2003). On the residual autocorrelation of the autoregressive conditional duration model. Economics Letters, 79, 169–175.

    Article  MathSciNet  MATH  Google Scholar 

  55. Mallat, S. G. (1989). A theory for multiresolution signal decomposition: The wavelet representation. IEEE Transactions on Pattern Analysis and Machine Intelligence, 11, 674–693.

    Article  MATH  Google Scholar 

  56. Meitz, M., & Teräsvirta, T. (2006). Evaluating models of autoregressive conditional duration. Journal of Business & Economic Statistics, 24, 104–124.

    Article  MathSciNet  Google Scholar 

  57. Neumann, M. H. (1996). Spectral density estimation via nonlinear wavelet methods for stationary non-Gaussian time series. Journal of Time Series Analysis, 17, 601–633.

    Article  MathSciNet  MATH  Google Scholar 

  58. Newey, W. K., & West, K. D. (1994). Automatic lag selection in covariance matrix estimation. Review of Financial Studies, 61, 631–653.

    MathSciNet  MATH  Google Scholar 

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

    Article  Google Scholar 

  60. Percival, D. B., & Walden, A. T. (2000). Wavelet methods for time series analysis, Cambridge Series in Statistical and Probabilistic Mathematics. Cambridge: Cambridge University Press.

    Book  MATH  Google Scholar 

  61. Priestley, M. B. (1981). Spectral analysis and time series. London: Academic Press.

    MATH  Google Scholar 

  62. Priestley, M. B. (1996). Wavelets and time-dependent spectral analysis. Journal of Time Series Analysis, 17, 85–103.

    Article  MathSciNet  MATH  Google Scholar 

  63. Ramsey, J. (1999). The contribution of wavelets to the analysis of economic and financial data. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering, 537, 2593–2606.

    Article  MathSciNet  MATH  Google Scholar 

  64. Robinson, P. M. (1991). Automatic frequency domain inference on semiparametric and nonparametric models. Econometrica, 59, 1329–1363.

    Article  MathSciNet  MATH  Google Scholar 

  65. Stock, J. (1988). Estimating continuous time processes subject to time deformation. Journal of the American Statistical Association, 83, 77–85.

    MathSciNet  Google Scholar 

  66. Tauchen, G., & Pitts, M. (1983). The price variability-volume relationship on speculative markets. Econometrica, 51, 485–505.

    Article  MATH  Google Scholar 

  67. Tsay, R. S. (2005). Analysis of financial time series (2nd ed.). New York: Wiley.

    Book  MATH  Google Scholar 

  68. Tsay, R. S. (2013). An introduction to analysis of financial data with R. New York: Wiley.

    MATH  Google Scholar 

  69. Vidakovic, B. (1999). Statistical modeling by wavelets. New York: Wiley.

    Book  MATH  Google Scholar 

  70. Wang, Y. (1995). Jump and sharp cusp detection by wavelets. Biometrika, 82, 385–397.

    Article  MathSciNet  MATH  Google Scholar 

  71. Watson, N. W. (1993). Measures of fit for calibrated models. Journal of Political Economy, 101, 1011–1041.

    Article  Google Scholar 

  72. Xue, Y., Gençay, R., & Fagan, S. (2010). Jump detection with wavelets. Technical report, Department of Economics, Simon Fraser University.

    Google Scholar 

  73. Zhang, M. Y., Russell, J. R., & Tsay, R. S. (2001). A nonlinear autoregressive conditional duration model with applications to financial transaction data. Journal of Econometrics, 104, 179–207.

    Article  MathSciNet  MATH  Google Scholar 

  74. Zhu, K., & Li, W. K. (2015). A bootstrapped spectral test for adequacy in weak ARMA models. Journal of Econometrics, 187, 113–130.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to thank W. K. Li, David Stanford, Hao Yu, and two referees for constructive suggestions, which led to an improved paper. Funding in partial support of this work was provided by the Natural Science and Engineering Research Council of Canada.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre Duchesne .

Editor information

Editors and Affiliations

Appendix

Appendix

To prove Theorems 13, we first state some useful lemmas.

Lemma 1

Define \(d_{J}(h)\equiv \sum _{j=0}^{J}\lambda (2\pi h/2^{j}),\) \(h,J\in \mathbb {Z}\), where \(\lambda (z)\) is as in (8). Then

  1. (i)

    \(d_{J}(0)=0\) and \(d_{J}(-h)=d_{J}(h)\) for all \(h,J\in \mathbb {Z},J>0\);

  2. (ii)

    \(|d_{J}(h)|\le C<\infty \) uniformly in \(h,J\in \mathbb {Z},J>0\);

  3. (iii)

    For any given \(h\in \mathbb {Z},h\ne 0\), \(d_{J}(h)\rightarrow 1\) as \(J\rightarrow \infty \);

  4. (iv)

    For any given \(r \ge 1\), \(\sum _{h=1}^{n-1}|d_{J}(h)|^{r}=O(2^{J})\) as \(J,n\rightarrow \infty \).

Lemma 2

Let \(V_{n}(J)\) and \(V_{0}\) be defined as in Theorem 1. Suppose \(J\rightarrow \infty , 2^{J}/n\rightarrow 0\). Then \(2^{-J}V_{n}(J)\rightarrow V_{0}\), where \(V_{0}=\int _{0}^{2\pi }|\varGamma (z)|^{2}dz\), with \(\varGamma (z)=\sum _{-\infty }^{\infty }\hat{\psi }(z+2\pi m)\).

Proof

For the proofs of Lemmas 1 and 2, see [42, Proof of Lemma A.1] and [44, Proof of Lemma A.2]. \(\Box \)

Proof

(Theorem 1) The model under the null hypothesis is \(D_t \equiv \beta _0\), \(X_t = \beta _0 \varepsilon _t\), \( E(\varepsilon _t)=1\), \(\{ \varepsilon _t \}\) an iid process. We write \(\bar{R}_X(h)= n^{-1}\sum _{t=|h|+1}^n (X_t/\bar{X}-1)(X_{t-|h|}/\bar{X}-1)\). Alternatively, \(\hat{\rho }_X(h)= \bar{R}_X(h)/\bar{R}_X(0)\). Under the null hypothesis, \( \bar{R}_X(h) = \bar{R}_{\varepsilon }(h)\) where \(\bar{R}_{\varepsilon }(h)\) is defined similarly as \(\bar{R}_X(h)\). We define \(R_{\varepsilon }(h)=\text {cov}(\varepsilon _t,\varepsilon _{t-h})\). Let \(u_t=\varepsilon _t-1\). We define \(\tilde{R}_{\varepsilon }(h)=n^{-1}\sum _{t=|h|+1}^n u_t u_{t-h}\) and \(\tilde{\rho }_{\varepsilon }(h) = \tilde{R}_{\varepsilon }(h)/R_{\varepsilon }(0)\). Define

$$ \tilde{f}_{\varepsilon }(0) \equiv (2\pi )^{-1} + \sum _{j=0}^{J}\sum _{k=1}^{2^{j}} \tilde{\alpha }_{jk}\varPsi _{jk}(0), $$

where \(\tilde{\alpha }_{jk} \equiv (2\pi )^{-1/2}\sum _{h=1-n}^{n-1} \tilde{ \rho }_{\varepsilon }(h)\hat{\varPsi }_{jk}^{*}(h)\), \(\hat{\varPsi } _{jk}(h)\equiv (2\pi )^{-1/2}\int _{-\pi }^{\pi } \varPsi _{jk}(\omega )e^{-ih\omega }d\omega \).

Writing \(\hat{f}_X(0)-(2\pi )^{-1}=[\hat{f}_X(0)-\tilde{f}_{\varepsilon }(0)]+ [\tilde{f}_{\varepsilon }(0)-(2\pi )^{-1}],\) we shall prove Theorem 1 by showing Theorems 78 below.

Theorem 7

\([ V_{n}(J) ]^{-1/2}n^{1/2}[\hat{f}_X(0)-\tilde{f}_{\varepsilon }(0)]\rightarrow ^{p}0.\)

Theorem 8

\([ V_{n}(J) ]^{-1/2}n^{1/2}\pi [\tilde{f}_{\varepsilon }(0)-(2\pi )^{-1}]\rightarrow ^{d} N(0,1).\)

Proof

(Theorem 7) We use the following representations (see [44]):

$$\begin{aligned} \hat{f}_X(0)= & {} (2\pi )^{-1}+\pi ^{-1}\sum _{h=1}^{n-1}d_{J}(h)\hat{\rho }_X(h), \\ \tilde{f}_{\varepsilon }(0)= & {} (2\pi )^{-1}+\pi ^{-1}\sum _{h=1}^{n-1}d_{J}(h)\tilde{\rho }_{\varepsilon }(h) . \nonumber \end{aligned}$$
(20)

We obtain \(\pi [\hat{f}_X(0)-\tilde{f}_{\varepsilon }(0)]=\sum _{h=1}^{n-1}d_{J}(h) [\hat{\rho }_X(h)-\tilde{\rho }_{\varepsilon }(h)]\). Because \(\bar{R}_X(0)-R_{\varepsilon }(0)=O_{P}(n^{-1/2})\) given Assumption 1, it suffices to show

$$\begin{aligned}{}[ V_{n}(J) ]^{-1/2}n^{1/2}\sum _{h=1}^{n-1}d_{J}(h) [\bar{R}_X(h)-\tilde{R}_{\varepsilon }(h)]\rightarrow ^{p} 0. \end{aligned}$$
(21)

We shall show (21) for the case \(J\rightarrow \infty \), where \(2^{-J}V_{n}(J)\rightarrow V_{0}\) by Lemma 2. The proof for fixed J is similar, with \(V_{n}(J)\rightarrow V_{0}(J)\), where \(V_{0}(J)\) is defined in Lemma 2.

Straightforward algebra yields \(\hat{R}_X(h)-\tilde{R}_{\varepsilon }(h)= (\bar{\varepsilon }^{-1}-1)\hat{A}_{1}(h)+ (\bar{\varepsilon }^{-1}-1)\hat{A}_{2}(h)+ (\bar{\varepsilon }^{-1}-1)^2\hat{A}_{3}(h)\), where

$$ \hat{A}_1(h) = n^{-1}\sum _{t=|h|+1}^n u_t \varepsilon _{t-h}, \; \hat{A}_2(h) = n^{-1}\sum _{t=|h|+1}^n u_{t-h} \varepsilon _{t}, \; \hat{A}_3(h) = n^{-1}\sum _{t=|h|+1}^n \varepsilon _t \varepsilon _{t-h}. $$

We first consider \(\hat{A}_{1}(h)\). Note that \(E[\hat{A}_1(h)\hat{A}_1(m)] = O(n^{-1})\), \(\forall h,m\). Then expending the square, we show that \(E[\sum _{h=1}^{n-1}d_J(h)\hat{A}_1(h)]^2 = O(2^J/n + 2^{2J}/n)\). This shows that \(\sum _{h=1}^{n-1} d_J(h) \hat{A}_1(h) = O_P(2^J/n^{1/2})\). Proceeding similarly we show that \(\sum _{h=1}^{n-1} d_J(h) \hat{A}_2(h) = O_P(2^J/n^{1/2})\). We show also easily that

$$ \sum _{h=1}^{n-1} d_J(h)\hat{A}_{3}(h) = O_P(2^J). $$

This completes the proof for Theorem 7. \(\Box \)

Proof

(Theorem 8) Put \(\hat{W}\equiv \sum _{h=1}^{n-1}d_{J}(h)\tilde{R}_{\varepsilon }(h)/R_{\varepsilon }(0).\) Write \(\hat{W}=n^{-1}\sum _{t=2}^{n}W_{t}\), where

$$ W_{t}\equiv R^{-1}_{\varepsilon }(0)u_{t}\sum _{h=1}^{t-1}d_{J}(h)u_{t-h}. $$

Observe that \(\{W_{t},\mathscr {F}_{t-1}\}\) is an adapted martingale difference sequence, where \(\mathscr {F}_{t}\) is the sigma field consisting of all \(u_{s},s\le t.\) Thus, we obtain

$$ {\text {var}}(n^{1/2}\hat{W})= n^{-1}\sum _{t=2}^{n}E[W_{t}^{2}]= n^{-1} \sum _{t=2}^{n}\sum _{h=1}^{t-1}d_{J}^{2}(h)= \sum _{h=1}^{n}(1-h/n)d_{J}^{2}(h)=V_{n}(J). $$

By the martingale central limit theorem in [37, pp.10–11], \([ V_{n}(J) ]^{-1/2}n^{1/2}\hat{W}\rightarrow ^{d}N(0,1)\) if we can show

$$\begin{aligned}{}[V_{n}(J)]^{-1} n^{-1}\sum _{t=2}^{n}E\left\{ W_{t}^{2}\mathbf {1}[|W_{t}|>\eta n^{1/2}\{ V_{n}(J) \}^{1/2}]\right\}\rightarrow & {} \; 0 \text { for any }\eta >0, \end{aligned}$$
(22)
$$\begin{aligned}{}[V_{n}(J)]^{-1} n^{-1}\sum _{t=2}^{n}E(W_{t}^{2}|\mathscr {F}_{t-1})&\rightarrow ^{p}&\; 1. \end{aligned}$$
(23)

For space, we shall show the central limit theorem for \(\hat{W}\) for large J (i.e., \(J\rightarrow \infty ).\) The proof for fixed J is similar and simpler because \(d_{J}(h)\) is finite and summable.

We shall verify the first condition by showing \(2^{-2J}n^{-2}\sum _{t=2}^{n}E(W_{t}^{4}) \rightarrow 0\). Put \(\mu _{4} \equiv E(u_{t}^{4})\). By Assumption 1, we have

$$\begin{aligned} E(W_{t}^{4})= & {} \mu _{4}R^{-4}_{\varepsilon }(0) E\left[ \sum _{h=1}^{t-1}d_{J}(h)u_{t-h}\right] ^{4}, \\= & {} \mu _{4}^{2}R^{-4}_{\varepsilon }(0)\sum _{h=1}^{t-1}d_{J}^{4}(h)+ 6\mu _{4}R^{-2}_{\varepsilon }(0)\sum _{h_1=2}^{t-1}\sum _{h_2=1}^{h_1-1}d_{J}^{2}(h_1)d_{J}^{2}(h_2) \le 3\mu _{4}^{2}R^{-4}_{\varepsilon }(0)\left[ \sum _{h=1}^{n-1}d_{J}^{2}(h)\right] ^{2}. \end{aligned}$$

It follows from Lemma 2 that \(2^{-2J}n^{-2}\sum _{t=1}^{n}E(W_{t}^{4})=O(n^{-1}).\) This show (22).

Next, given Lemma 2, it suffices for expression (23) to establish the sufficient condition

$$ 2^{-2J}{\text {var}}[n^{-1}\sum _{t=2}^{n}E(W_{t}^{2}|\mathscr {F}_{t-1})]\rightarrow 0. $$

By the definition of \(W_{t}\), we have

$$\begin{aligned} E(W_{t}^{2}|\mathscr {F}_{t-1})= & {} R^{-1}_{\varepsilon }(0)\left[ \sum _{h=1}^{t-1}d_{J}(h)u_{t-h}\right] ^{2}, \\= & {} E(W_{t}^{2})+R^{-1}_{\varepsilon }(0)\sum _{h=1}^{t-1}d_{J}(h)[u_{t-h}^{2}-R_{\varepsilon }(0)]\\&+ 2R^{-1}_{\varepsilon }(0)\sum _{h_1=2}^{t-1} \sum _{h_2=1}^{h_1-1}d_{J}(h_1)d_{J}(h_2)u_{t-h_1}u_{t-h_2}, \\= & {} E(W_{t}^{2})+R^{-1}_{\varepsilon }(0)A_t+2R^{-1}_{\varepsilon }(0)B_{t}. \end{aligned}$$

It follows that

$$\begin{aligned} n^{-1}\sum _{t=2}^{n}[E(W_{t}^{2}|\mathscr {F}_{t-1})-E(W_{t}^{2})]= & {} R^{-1}_{\varepsilon }(0)n^{-1}\sum _{t=2}^{n}A_{t}+ 2R^{-1}_{\varepsilon }(0)n^{-1}\sum _{t=2}^{n}B_{t} \nonumber \\= & {} R^{-1}_{\varepsilon }(0)\hat{A}+2R^{-1}_{\varepsilon }(0)\hat{B}. \end{aligned}$$
(24)

Whence, it suffices to show \(2^{-2J}[{\text {var}}(\hat{A})+{\text {var}}(\hat{B})]\rightarrow 0\). First, noting that \(A_{t}\) is a weighted sum of independent zero-mean variables \(\{u_{t-h}^{2}-R_{\varepsilon }(0)\},\) we have \(E(A_{t}^{2})=[\mu _{4}-R^2_{\varepsilon }(0)]\sum _{h=1}^{t-1}d_{J}^{4}(h).\) It follows by Minkowski’s inequality and Lemma 1(iv) that

$$\begin{aligned} E(\hat{A}^{2}) \le \left\{ n^{-1}\sum _{t=2}^{n}[E(A_{t}^{2})]^{1/2}\right\} ^{2}\le \left[ \mu _{4}-R^{2}_{\varepsilon }(0)\right] \left[ \sum _{h=1}^{n-1}d_{J}^{4}(h)\right] =O(2^{J}). \end{aligned}$$
(25)

Next, we consider \({\text {var}}(\hat{B})\). For all \(t\ge s,\) we have

$$\begin{aligned} E(B_{t}B_{s})= & {} R_{\varepsilon }^{2}(0) \sum _{m_{2}=2}^{t-1}\sum _{h_{2}=1}^{m_{2}-1}\sum _{m_{1}=2}^{s-1}\sum _{h_{1}=1}^{m_{1}-1}d_{J}(m_{1})d_{2}(h_{1})d_{J}(m_{2})d_{J}(h_{2})\delta _{t-h_{1},s-h_{2}}\delta _{t-m_{1},s-m_{2}}, \\= & {} R_{\varepsilon }^{2}(0)\sum _{m=2}^{t-1}\sum _{h=1}^{l-1}d_{J}(t-s+m)d_{J}(t-s+h)d_{J}(m)d_{J}(h), \end{aligned}$$

where \(\delta _{j,h}=1\) if \(h=j\) and \(\delta _{j,h}=0\) otherwise. It follows that

$$\begin{aligned} E(\hat{B}^{2})\le & {} 2n^{-2}\sum _{t=3}^{n}\sum _{s=2}^{t}E(B_{t}B_{s})\le 2R_{\varepsilon }^{2}(0)n^{-1} \sum _{\tau =0}^{n-1}\sum _{m=2}^{n-1}\sum _{h=1}^{m-1}|d_{J}(\tau +m)d_{J}(\tau +h)d_{J}(m)d_{J}(h)|, \nonumber \\\le & {} 2R_{\varepsilon }^{2}(0)n^{-1}\left[ \sum _{\tau =0}^{n-1}d_{J}^{2}(\tau )\right] \left[ \sum _{h=1}^{n-1}|d_{J}(h)|\right] ^{2}=O(2^{3J}/n), \end{aligned}$$
(26)

by Lemma 1(iv). Combining (24)–(26) yields \(2^{-2J}[{\text {var}}(\hat{A})+{\text {var}}(\hat{B})]=O(2^{-J}+2^{J}/n)\rightarrow 0\) given \(J\rightarrow \infty , 2^{J}/n\rightarrow 0.\) Thus, condition (23) holds. By [37, pp.10–11], \([ V_{n}(J) ]^{-1/2}n^{1/2}\hat{W}\rightarrow ^{d}N(0,1)\). This completes the proof of Theorem 8. \(\Box \)

Proof

(Theorem 2) We shall show for large J only; the proof for fixed J is similar. Here we explicitly denote \(\hat{f}_X(0;J) \) as the spectral estimator (20) with the finest scale J. Recalling the definition of \(\mathscr {E}(J)\), we have

$$\begin{aligned} \mathscr {E}(\hat{J})-\mathscr {E}(J)= & {} [V_{n}(\hat{J})]^{-1/2}n^{1/2}\pi \{\hat{f}_X(0;\hat{J})- (2\pi )^{-1}\}-[V_{n}(J)]^{-1/2}n^{1/2}\pi \{\hat{f}_X(0;J)-(2\pi )^{-1}\}, \\= & {} [V_{n}(\hat{J})/V_{n}(J)]^{1/2}[V_{n}(J)]^{-1/2} n^{1/2}\pi \{\hat{f}_X(0;\hat{J})-\hat{f}_X(0;J)\} +\{[V_{n}(J)/V_{n}(\hat{J})]^{-1/2}-1\}\mathscr {E}(J). \end{aligned}$$

Note that we have for any given constants \(C_0 >0\) and \(\varepsilon >0\),

$$\begin{aligned} P\left( |V_{n}(J)/V_{n}(\hat{J})-1|>\varepsilon \right)\le & {} P\left( |V_{n}(J)/V_{n}(\hat{J})-1|>\varepsilon ,C_0 2^{J/2}|2^{\hat{J}}/2^{J}-1|\le \varepsilon \right) \nonumber \\+ & {} P\left( C_0 2^{J/2}|2^{\hat{J}}/2^{J}-1|>\varepsilon \right) . \end{aligned}$$
(27)

We now study \(|d_{\hat{J}}(h)-d_J(h)|\). We show that

$$ |d_{\hat{J}}(h)-d_J(h)| \le \sum _{j=\min (J,\hat{J})+1}^{\max (J,\hat{J})} |\lambda (2 \pi h/2^j)|. $$

Note that given \(C_0 2^{J/2}|2^{\hat{J}}/2^J - 1| \le \varepsilon \), we have that

$$ |d_{\hat{J}}(h)-d_J(h)| \le \sum _{j=\log _2[2^J(1-\varepsilon /(C_0 2^{J/2}))]+1}^{\log _2[2^J(1+\varepsilon /(C_0 2^{J/2}))]} |\lambda (2 \pi h/2^j)|, $$

and the lower and upper bounds in the summation are now non stochastic. Since \(|\sum _{m=-\infty }^{\infty } \psi (2\pi h/2^j + 2\pi m)| \le C\), we have that

$$ |d_{\hat{J}}(h)-d_J(h)| \le C \sum _{j=\log _2[2^J(1-\varepsilon /(C_0 2^{J/2}))]+1}^{\log _2[2^J(1+\varepsilon /(C_0 2^{J/2}))]} |\hat{\psi }(2 \pi h/2^j)|. $$

Since \(\sum _{h=-\infty }^{\infty }|\hat{\psi }(2 \pi h/2^j)| \le C 2^j\), then

$$ \sum _{h=1}^{n-1} |d_{\hat{J}}(h)-d_J(h)| \le C \sum _{j=\log _2[2^J(1-\varepsilon /(C_0 2^{J/2}))]+1}^{\log _2[2^J(1+\varepsilon /(C_0 2^{J/2}))]} 2^j \le C2^{J/2}\varepsilon /C_0. $$

A similar argument allows us to show that

$$\begin{aligned} \sum _{h=1}^{n-1} |d_{\hat{J}}(h)-d_J(h)|^2 \le C2^{J/2}\varepsilon /C_0. \end{aligned}$$
(28)

We show that \(V_n(\hat{J})/V_n(J) \rightarrow ^p 1\). Note that

$$ |V_n(\hat{J}) - V_n(J)| \le \sum _{h=1}^{n-1} |d_{\hat{J}}(h)-d_J(h)|^2 + 2\left( \sum _{h=1}^{n-1} d_J^2(h) \right) ^{1/2} \left( \sum _{h=1}^{n-1} \left[ d_{\hat{J}}(h)-d_J(h) \right] ^2 \right) ^{1/2}. $$

Since \(\sum _{h=1}^{n-1} d_J^2(h) = O(2^J)\), by (27) and (28), we have the announced result that \(V_n(\hat{J})/V_n(J) \rightarrow ^p 1\).

Because \(\mathscr {E}(J)=O_{P}(1)\) by Theorem 1 and since \(V_{n}(\hat{J})/V_{n}(J)\rightarrow ^{p} 1\), we have \(\mathscr {E}(\hat{J})-\mathscr {E}(J)\rightarrow ^{p} 0\) provided \([V_{n}(J)]^{-1/2} n^{1/2}\pi \{\hat{f}_{\hat{J}}(0)-\hat{f}_{J}(0)\} \rightarrow ^{p}0,\) which we shall show below. The asymptotic normality of \(\mathscr {E}(\hat{J})\) follows from a standard application of Slutsky’s theorem and Theorem 1.

Because \(V_{n}(J)=O(2^{J}),\) it suffices to show \(\hat{f}_{X}(0;\hat{J})-\hat{f}_{X}(0;J)=o_{P}(2^{J/2}/n^{1/2}).\) Write

$$ \pi \{\hat{f}_{X}(0;\hat{J})-\hat{f}_{X}(0;J)\}=\hat{R}_{X}^{-1}(0)\sum _{h=1}^{n-1}[ d_{\hat{J}}(h)-d_{J}(h) ] R_{X}(h). $$

Given \(|d_{\hat{J}}(h)-d_{J}(h)|\le \sum _{j=\min (\hat{J},J)}^{\max (\hat{J},J)}|\lambda (2\pi h/2^{j})|\), we have, under the null hypothesis, the following inequality

$$ E\sum _{h=1}^{n-1}\left| d_{\hat{J}}(h)-d_{J}(h)\right| \left| R_{X}(h)\right| \le \sup _{h}E(R_{X}^{2}(h))^{1/2}\sum _{h=1}^{n-1}\left( E|d_{\hat{J}}(h)-d_{J}(h)\right| ^{2})^{1/2}= o(2^{J/2}/n^{1/2}). \nonumber $$

We obtain \(\hat{f}_{X}(0;\hat{J})-\hat{f}_{X}(0;J)=o_{P}(2^{J/2}/n^{1/2}).\) This completes the proof of Theorem 2. \(\Box \)

Proof

(Theorem 3) Recall \(\hat{R}_X(h)= n^{-1}\sum _{t=|h|+1}^n (X_t-\bar{X})(X_{t-|h|}-\bar{X})\) and \(\tilde{R}_X(h)= n^{-1}\sum _{t=|h|+1}^n (X_t-\mu )(X_{t-|h|}-\mu )\). We study \(\hat{f}_X(0;J) - f_X(0)\). Write

$$\begin{aligned} \hat{f}_X(0;J) - f_X(0) = \{\hat{f}_X(0;J) - \tilde{f}_X(0;J)\} + \{ \tilde{f }_X(0;J) - E[\tilde{f}_X(0;J)] \} + \{ E[\tilde{f}_X(0;J)] - f_X(0) \}, \end{aligned}$$
(29)

where

$$\begin{aligned} \tilde{f}_X(0;J) = \frac{\tilde{R}_X(0)}{2\pi }+ \frac{1}{\pi } \sum _{h=1}^{n-1} d_J(h) \tilde{R}_X(h). \end{aligned}$$
(30)

We can show that \(E[ \hat{f}_X(0;J) - \tilde{f}_X(0;J) ]^2 = O(n^{-2} + 2^{2J}/n^2)\), meaning that replacing \(\bar{X}\) by \(\mu \) has to impact asymptotically. For the second term in (29), we show that

$$\begin{aligned} E[\tilde{f}_X(0;J) - E\tilde{f}_X(0;J)]^2= & {} \frac{{\text {var}}[ \tilde{R}_X(0)]}{4 \pi ^2} + \pi ^{-2} \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} d_J(h)d_J(m) {\text {cov}}(\tilde{R}_X(h),\tilde{R}_X(m)) \nonumber \\&+ \pi ^{-2} \sum _{h=1}^{n-1} d_J(h) {\text {cov}}(\tilde{R}_X(0),\tilde{R}_X(h)). \end{aligned}$$
(31)

From [38, p. 313], we have

$$\begin{aligned} \frac{(n-l)(n-m)}{n^2} {\text {cov}}\left[ \tilde{R}_X(h),\tilde{R} _X(m)\right]= & {} n^{-1}\sum _{u=-\infty }^{\infty }w_{n}(u,h,m) [R_X(u) R_X(u+m-h) \nonumber \\&+ R_X(u+m) R_X(u-h)+\kappa (0,h,u,u+m)], \nonumber \end{aligned}$$

where the function \(w_{n}(u,h,m)\) is defined in [38]. We write

$$\begin{aligned} E\left\{ \tilde{f}_X(0;J) - E[\tilde{f}_X(0;J)] \right\} ^2 = \frac{{\text {var}}( \tilde{R}_X(0))}{4\pi ^2} + \frac{A_n}{\pi ^2} + \frac{B_n}{\pi ^2}, \end{aligned}$$
(32)

where

$$\begin{aligned} A_n= & {} E\left\{ [\tilde{R}_X(0)-E(\tilde{R}_X(0))] \sum _{h=1}^{n-1} d_J(h)[\tilde{R}_X(h)-E(\tilde{R}_X(h))] \right\} , \\ B_n= & {} \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} d_J(h) d_J(m) {\text {cov}}[\tilde{R}_X(h),\tilde{R}_X(m)]. \end{aligned}$$

It follows that

$$\begin{aligned} (n/2^{J+1})B_n= & {} 2^{-(J+1)}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1} d_{J}(h)d_{J}(m)\sum _{u=-\infty }^{\infty }w_{n}(u,h,m)R_X(u)R_X(u+m-h) \nonumber \\&+2^{-(J+1)}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}d_J(h)d_J(m) \sum _{u=-\infty }^{\infty }w_{n}(u,h,m)R_X(u+m)R_X(u-h) \nonumber \\&+2^{-(J+1)}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}d_J(h)d_J(m) \sum _{u=-\infty }^{\infty }w_{n}(u,h,m)\kappa (0,h,u,u+m), \nonumber \\= & {} B_{1n}+B_{2n}+B_{3n}. \end{aligned}$$
(33)

Following a reasoning similar to [42, proof of Theorem 4.2], we can show that

$$ B_{1n} = 2^{-1}D_{\psi }(2\pi )^{2}f_X^2(0)[1+o(1)], \; B_{2n}\rightarrow 0, \; B_{3n} \rightarrow 0, $$

and \(\left| A_n \right| =O(2^{J/2}/n)=o(2^{J}/n)\). It follows that

$$\begin{aligned} n/2^{J+1}B_n \rightarrow 2\pi ^{2}D_{\psi }f_X^2(0), \end{aligned}$$
(34)

and

$$\begin{aligned} (n/2^{J+1}) E[\tilde{f}_X(0;J) - E\tilde{f}_X(0;J)]^2 \rightarrow 2 D_{\psi }f_X^2(0). \end{aligned}$$
(35)

We consider the bias term \(E[\tilde{f}_{X}(J)]-f_{X}(0)\) in (29). Using the definition of \(\tilde{f}_{X}(0;J)\) in (30), we can decompose

$$\begin{aligned} E[\tilde{f}_{X}(J)]-f_{X}(0)= & {} \pi ^{-1}\sum _{h=1}^{n-1}d_{J}(h)(1-h/n)R_{X}(h)-\pi ^{-1}\sum _{h=1}^{\infty }R_{X}(h), \nonumber \\= & {} \pi ^{-1}\sum _{h=1}^{n-1}(1-h/n)\left[ d_{J}(h)-1\right] R_{X}(h)-\pi ^{-1}\sum _{h=1}^{n-1}(h/n)R_{X}(h)-\pi ^{-1}\sum _{h=n}^{\infty }R_{X}(h), \nonumber \\= & {} B_{4n}-B_{5n}-B_{6n},\text { say.} \end{aligned}$$
(36)

Following a reasoning similar to [42, proof of Theorem 4.2], we can show that

$$ B_{4n}=-2^{-q(J+1)}\lambda _{q}f_{X}^{(q)}(0)[1+o(1)], $$

\(\left| B_{5n}\right| \le n^{-\min (1,q)}\sum _{h=1}^{n-1}l^{q}|R_{X}(h)|=O(n^{-\min (1,q)})\) and also that \(\left| B_{6n}\right| \le 2n^{-q}\sum _{h=n}^{\infty }h^{q}|R_{X}(h)|=o(n^{-q})\). The bias term is then

$$\begin{aligned} E\tilde{f}_{X}(0;J)-f_{X}(0)=-2^{-q(J+1)}\lambda _{q}f_{X}^{(q)}(0)+o(2^{-qJ})+O(n^{-\min (1,q)}). \end{aligned}$$
(37)

Now, combining (29), (35), (37) we obtain

$$ E\{[\hat{f}_{X}(0;J)-f_{X}(0)]^{2}\}=\frac{2^{J+1}}{n}2D_{\psi }f_{X}^{2}(0)+2^{-2q(J+1)}\lambda _{q}^{2}[f_{X}^{(q)}(0)]^{2}+o(2^{J}/n+2^{-2qJ}). $$

The desired result follows by using \(2^{J+1}/n^{\frac{1}{2q+1}}\rightarrow c\). This completes the proof of Theorem 3. \(\Box \)

Proof

(Corollary 1) The result follows immediately from Theorem 2 because the conditions of Corollary 1 imply \(2^{\hat{J}}/2^{J}-1=o_{P}(n^{-1/2(2q+1)})=o_{P}(2^{-J/2}),\) where the nonstochastic finest scale J is given by \(2^{J+1}\equiv \max \{[q\lambda _{q}^{2}\alpha (q)n/D_{\psi }]^{1/(2q+1)},0\}.\) The latter satisfies the conditions of Theorem 2. \(\Box \)

To prove Theorems 46, we first state a useful lemma.

Lemma 3

Suppose Assumptions 1 and 2 hold, \(J\rightarrow \infty \), and \(2^{J}/n\rightarrow 0\). Define

$$ b_{J}(h,m)=a_{J}(h,m)+a_{J}(-h,-m)+a_{J}(h,-m)+a_{J}(-h,m), $$

where \(a_{J}(h,m)=2\pi \sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\hat{\psi }_{jk}(2\pi h)\hat{\psi }_{jk}^{*}(2\pi m).\) Then

  1. (i)

    \(b_{J}(h,m)\) is real-valued, \(b_{J}(0,m)=b_{J}(h,0)=0\) and \(b_{J}(h,m)=b_{J}(m,h)\);

  2. (ii)

    \(\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}h^{\nu }|b_{J}(h,m)|=O(2^{(1+\nu )J})\) for \(0\le \nu \le \frac{1}{2}\);

  3. (iii)

    \(\sum _{h=1}^{n-1}\{\sum _{m=1}^{n-1}|b_{J}(h,m)|\}^{2}=O(2^{J})\);

  4. (iv)

    \(\sum _{h_{1}=1}^{n-1}\sum _{h_{2}=1}^{n-1}\{\sum _{m=1}^{n-1}|b_{J}(h_{1},m)b_{J}(h_{2},m)|\}^{2}=O\{(J+1)2^{J})\);

  5. (v)

    \(\sum _{h=1}^{n-1}b_{J}(h,h)=(2^{J+1}-1) \{1\)+\(O((J+1)/2^{J}\)+\(2^{J(2\tau -1)}/n^{2\tau -1})\},\) where \(\tau \) is in Assumption 3;

  6. (vi)

    \(\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}^{2}(h,m)=2(2^{J+1}-1)\{1+o(1)\}\);

  7. (vii)

    \(\sup _{1\le h,m\le n-1}|b_{J}(h,m)|\le C(J+1)\);

  8. (viii)

    \(\sup _{1\le h\le n-1}\sum _{m=1}^{n-1}|b_{J}(h,m)|\le C(J+1)\).

Proof

(Lemma 3) See [50, Appendix B] for (i)–(vi) and [43, Lemma A.1] for (vii) and (viii). \(\Box \)

Proof

(Theorem 4) Let \(Z_{t}=\varepsilon _{t}-1\) be such that \(E(Z_{t})=0\). Under \(\mathbb {H}_{0}^{\mathscr {A}}\), \(\varepsilon _{t}=e_{t}\). We consider \(\bar{R}_{Z}(h)=n^{-1}\sum _{t=|h|+1}^{n}Z_{t}Z_{t-|h|}\) and \(\bar{\alpha }_{ejk}=\sum _{h=1-n}^{n-1}\bar{R}_{Z}(h)\hat{\psi }_{jk}^{*}(2\pi h)\). Let \(\bar{\mathscr {A}}(J)\) defined as \(\mathscr {A}(J)\) but using \(\bar{\alpha }_{ejk}\). Let \(\bar{\rho }_{Z}(h)=\bar{R}_{Z}(h)/\bar{R}_{Z}(0)\). Because \(\bar{\rho }_{Z}(-h)=\bar{\rho }_{Z}(h)\) and \(\bar{\alpha }_{ejk}\) is real-valued, we have

$$ 2\pi n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\bar{\alpha }_{ejk}^{2}=n \sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\bar{\rho }_{Z}(h)\bar{\rho }_{Z}(m), $$

where the equality follows from re-indexing and the definition of \(b_{J}(h,m)\). We have \(\bar{R}_Z(0)-\sigma _Z^{2}=O_{P}(n^{-1/2})\), since under \(\mathbb {H}_{0}^{\mathscr {A}}\) we have that \(\{ e_t \}\) is iid. Also, we have

$$\begin{aligned} n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\bar{\rho }_Z(h)\bar{\rho }_Z(m)= & {} \sigma _Z^{-4}n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\bar{R}_Z(h)\bar{R}_Z(m) \nonumber \\&+[ \bar{R}^{-2}_Z(0)-\sigma _Z^{-4} ] n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\bar{R}_Z(h)\bar{R}_Z(m), \\= & {} \sigma _Z^{-4}n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\bar{R}_Z(h)\bar{R}_Z(m)+ O_{P}(2^{J}/n^{1/2}), \nonumber \end{aligned}$$
(38)

where the second term is of the indicated order of magnitude because

$$ E\left[ \sum _{h=1}^{n-1}\sum _{m=1}^{n-1}|b_J(h,m)\bar{R}_Z(h)\bar{R}_Z(m)| \right] \le Cn^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}|b_J(h,m)|=O(2^J/n). $$

given \(E[\bar{R}_Z^{2}(h)] \le Cn^{-1}\) and Lemma 3(ii). We now focus on the first term in (38). We have,

$$\begin{aligned} n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\bar{R}_Z(h)\bar{R}_Z(m)= & {} n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\sum _{t=h+1}^{n}\sum _{s=m+1}^{n}Z_{t}Z_{t-h}Z_{s}Z_{s-m}, \nonumber \\= & {} n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\left( \sum _{t=1}^{n}\sum _{s=1}^{n}-\sum _{t=1}^{h}\sum _{s=m+1}^{n} -\sum _{t=1}^{n}\sum _{s=1}^{m}\right) Z_{t}Z_{t-h}Z_{s}Z_{s-m}, \nonumber \\= & {} \hat{C}_{n}+\hat{D}_{1n}-\hat{D}_{2n}-\hat{D}_{3n}, \end{aligned}$$
(39)

where

$$\begin{aligned} \hat{C}_{n}= & {} n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\left( \sum _{t=2}^{n}\sum _{s=1}^{t-1}+\sum _{s=2}^{n}\sum _{t=1}^{s-1}\right) Z_{t}Z_{t-h}Z_{s}Z_{s-m}, \\= & {} 2n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\sum _{t=2}^{n}\sum _{s=1}^{t-1}Z_{t}Z_{t-h}Z_{s}Z_{s-m}, \text { given }b_{J}(h,m)=b_{J}(m,h), \\ \hat{D}_{1n}= & {} n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\sum _{t=1}^{n}Z_{t}^{2}Z_{t-h}Z_{t-m}, \\ \hat{D}_{2n}= & {} n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\sum _{t=1}^{h}\sum _{s=m+1}^{n}Z_{t}Z_{t-h}Z_{s}Z_{s-m}, \\ \hat{D}_{3n}= & {} n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\sum _{t=1}^{n}\sum _{s=1}^{m}Z_{t}Z_{t-h}Z_{s}Z_{s-m}. \end{aligned}$$

In order to prove Theorem 4, we first state Proposition 1 that shows that \(\hat{C}_{n}\) represents the dominant term.

Proposition 1

Suppose Assumptions 13 hold, \(J\rightarrow \infty \), and \(2^{2J}/n\rightarrow 0\). Then

$$ 2^{-J/2}\left\{ 2\pi n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\bar{\alpha }_{jk}^{2}-(2^{J+1}-1)\right\} =2^{-J/2}\sigma ^{-4}\hat{C}_{n}+o_{P}(1). $$

We now decompose \(\hat{C}_{n}\) into the terms with \(t-s>q\) and \(t-s\le q\), for some \(q\in \mathbb {Z}^{+}\):

$$\begin{aligned} \hat{C}_{n}= & {} 2n^{-1}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}b_{J}(h,m)\left( \sum _{t=q+2}^{n}\sum _{s=1}^{t-q-1}+\sum _{t=2}^{n}\sum _{s=\max (t-q,1)}^{t-1}\right) Z_{t}Z_{t-h}Z_{s}Z_{s-m}, \nonumber \\= & {} \hat{D}_{n}+\hat{D}_{4n}. \end{aligned}$$
(40)

Furthermore, we decompose

$$\begin{aligned} \hat{D}_{n}= & {} 2n^{-1}\left( \sum _{h=1}^{q}\sum _{m=1}^{q}+\sum _{h=1}^{q} \sum _{m=q+1}^{n-1}+\sum _{h=q+1}^{n-1}\sum _{m=1}^{n-1}\right) b_{J}(h,m)\sum _{t=q+2}^{n}\sum _{s=1}^{t-q-1}Z_{t}Z_{t-h}Z_{s}Z_{s-m}, \nonumber \\= & {} \hat{U}_{n}+\hat{D}_{5n}+\hat{D}_{6n},\text { say,} \end{aligned}$$
(41)

where \(\hat{D}_{5n}\) and \(\hat{D}_{6n}\) are the contributions from \(m>q\) and \(h>q\), respectively.

Proposition 2 shows that \(\hat{C}_{n}\) can be approximated arbitrarily well by \(\hat{U}_{n}\) under a proper condition on q.

Proposition 2

Suppose Assumptions 13 hold, \(J\rightarrow \infty ,2^{2J}/n\rightarrow 0\), and \(q\equiv q_{n}\rightarrow \infty ,q/2^{J}\rightarrow \infty ,q^{2}/n\rightarrow 0\). Then \(2^{-J/2}\hat{C}_{n}=2^{-J/2}\hat{U}_{n}+o_{P}(1).\)

It is much easier to show the asymptotic normality of \(\hat{U}_{n}\) than of \(\hat{C}_{n},\) because for \(\hat{U}_{n},\) \(\{Z_{t}Z_{t-h}\}\) and \(\{Z_{s}Z_{s-m}\}\) are independent given \(t-s>q\) and \(0<h,m\le q\).

Proposition 3

Suppose Assumptions 13 hold, and \(J\rightarrow \infty ,2^{2J}/n\rightarrow 0,q/2^{J}\rightarrow \infty , q^{2}/n\rightarrow 0\). Let \(\lambda _{n}^{2}=E(\hat{U}_{n}^{2})\). Then \(4(2^{J+1}-1)\sigma ^{8}/\lambda _{n}^{2}\rightarrow 1\), and \(\lambda _{n}^{-1}\hat{U}_{n}\rightarrow ^{d}N(0,1)\).

Propositions 13 and Slutsky’s Theorem imply \(\bar{\mathscr {A}}(J) \rightarrow ^{d} N(0,1)\). Propositions 4 and 5 show that parameter estimation does not have impact on the asymptotic distribution of the test statistic.

Proposition 4

\(n \sum _{j=0}^{J}\sum _{k=1}^{2^{j}} (\hat{\alpha }_{jk}-\bar{\alpha }_{jk})^{2}= O_P(2^J/n) + O_P(1).\)

Proposition 5

\(n \sum _{j=0}^{J}\sum _{k=1}^{2^{j}} (\hat{\alpha }_{jk}-\bar{\alpha }_{jk}) \bar{\alpha }_{jk} = o_P(2^{J/2}).\)

The proof of Theorem 4 will be completed provided Propositions 15 are shown. The proofs of Propositions 13 are very similar to the proofs of Propositions 1–3 in [50], for proving the asymptotic normality of a wavelet-based test statistic for serial correlation. These proofs are then omitted (but for the interested reader all the detailed proofs are available from the authors).

Proof

(Proposition 4) A standard Taylor’s expansion gives

$$ D_t^{-1}(\hat{{\varvec{\theta }}}) = D_t^{-1}({\varvec{\theta }}_0) + \left\{ \frac{\partial }{\partial {\varvec{\theta }}} D_t^{-1}({\varvec{\theta }}_0) \right\} ^{\top } (\hat{{\varvec{\theta }}}- {\varvec{\theta }}_0) + \frac{1}{2} (\hat{{\varvec{\theta }}}- {\varvec{\theta }}_0)^{\top }\frac{\partial ^2}{\partial {\varvec{\theta }} \partial {\varvec{\theta }}^{\top }} D_t^{-1}(\bar{{\varvec{\theta }}})(\hat{{\varvec{\theta }}}-{\varvec{\theta }}_0), $$

where \(\bar{{\varvec{\theta }}}\) lies between \(\hat{ {\varvec{\theta }}}\) and \({\varvec{\theta }}_0\). We have

$$\begin{aligned} \hat{R}_Z(h)-\tilde{R}_Z(h)= & {} n^{-1} \sum _{t=|h|+1}^n (\hat{Z}_t - Z_t)( \hat{Z}_{t-|h|} - Z_{t-|h|}) + n^{-1} \sum _{t=|h|+1}^n Z_t (\hat{Z}_{t-|h|} - Z_{t-|h|}) \\&+ n^{-1} \sum _{t=|h|+1}^n (\hat{Z}_t - Z_t) Z_{t-|h|}, \\= & {} \hat{A}_1(h) + \hat{A}_2(h)+\hat{A}_3(h). \end{aligned}$$

We write \(\hat{\alpha }_{jk} - \tilde{\alpha }_{jk} = \hat{B}_{1jk} + \hat{B} _{2jk}+\hat{B}_{3jk}\) where

$$ \hat{B}_{1jk} = \sum _{h=1-n}^{n-1} \hat{A}_1(h) \hat{\psi }_{jk}(2\pi h), \; \hat{B}_{2jk} = \sum _{h=1-n}^{n-1} \hat{A}_2(h) \hat{\psi }_{jk}(2\pi h), \; \hat{B}_{3jk} = \sum _{h=1-n}^{n-1} \hat{A}_3(h) \hat{\psi }_{jk}(2\pi h). $$

Then \((\hat{\alpha }_{jk} - \tilde{\alpha }_{jk})^2 \le 4(\hat{B}_{1jk}^2 + \hat{B}_{2jk}^2+ \hat{B}_{3jk}^3)\) We first study the term involving \(\hat{B}_{1jk}\). We write

$$ 2\pi \sum _{j=0}^J \sum _{k=1}^{2^j} \hat{B}_{1jk}^2 = \sum _{h=1-n}^{n-1}\sum _{m=1-n}^{n-1} a_J(h,m) \hat{A}_1(h)\hat{A}_1(m) = \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} b_J(h,m) \hat{A}_1(h)\hat{A}_1(m). $$

We have

$$ 2\pi \sum _{j=0}^J \sum _{k=1}^{2^j} \hat{B}_{1jk}^2 \le \left( \sup \hat{A} _1(h) \right) ^2 | \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} b_J(h,m) | = O_P(2^J/n). $$

We now study the term involving \(\hat{B}_{2jk}\). Let

$$\begin{aligned} \hat{a}_{21}(h)= & {} n^{-1} \sum _{t=|h|+1}^n Z_t X_{t-|h|} \left\{ \frac{\partial }{\partial {\varvec{\theta }}} D_{t-|h|}^{-1}({\varvec{\theta }}_0) \right\} ^{\top }, \\ \hat{a}_{22}(h)= & {} n^{-1} \sum _{t=|h|+1}^n Z_t X_{t-|h|} \frac{\partial ^2}{\partial {\varvec{\theta }} \partial {\varvec{\theta }}^{\top }}D_{t-|h|}^{-1}(\bar{{\varvec{\theta }}}_0). \end{aligned}$$

We write \(\hat{A}_2(h) = \hat{A}_{21}(h) + \hat{A}_{22}(h)\), where

$$\begin{aligned} \hat{A}_{21}(h)= & {} \hat{a}_{21}(h) (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0), \\ \hat{A}_{22}(h)= & {} \frac{1}{2} (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0)^{\top }\hat{a}_{22}(h) (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0). \end{aligned}$$

Then

$$\begin{aligned} \hat{B}_{2jk}= & {} \sum _{h=1-n}^{n-1} \hat{A}_{2}(h)\hat{\psi }_{jk}(2\pi h), \\= & {} \left[ \sum _{h=1-n}^{n-1} \hat{a}_{21}(h)\hat{\psi }_{jk}(2\pi h) \right] (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0) + \frac{1}{2}(\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0)^{\top } \left[ \sum _{h=1-n}^{n-1} \hat{a}_{22}(h) \hat{\psi }_{jk}(2\pi h) \right] (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0). \end{aligned}$$

We obtain

$$\begin{aligned} 2\pi n \sum _{j=0}^J \sum _{k=1}^{2^j} B_{2jk}^2\le & {} 4 n ||\hat{ {\varvec{\theta }}} - {\varvec{\theta }}_0||^2 \sum _{j=0}^J \sum _{k=1}^{2^j} \left\| \sum _{h=1-n}^{n-1} \hat{a}_{21}(h)\hat{\psi }_{jk}(2\pi h) \right\| ^2 \\&+ 2 n||\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0||^4 \sum _{j=0}^J \sum _{k=1}^{2^j} \left\| \sum _{h=1-n}^{n-1} \hat{a}_{22}(h) \hat{\psi }_{jk}(2\pi h) \right\| ^2 = O_P(2^J/n), \end{aligned}$$

since

$$\begin{aligned} 2\pi \sum _{j=0}^J \sum _{k=1}^{2^j} \left\| \sum _{h=1-n}^{n-1} \hat{a}_{21}(h) \hat{\psi }_{jk}(2\pi h) \right\| ^2= & {} \left| \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} b_J(h,m) \hat{a}_{21}(h)\hat{a}_{21}^{\top }(m) \right| , \\\le & {} \left( \sup || \hat{a}_{21}(h)|| \right) ^2 \left( \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} |b_J(h,m)| \right) = O_P(2^J/n). \end{aligned}$$

We now study the term involving \(\hat{B}_{3jk}\).

$$\begin{aligned} \hat{a}_{31}(h)= & {} n^{-1} \sum _{t=|h|+1}^n Z_{t-|h|} X_t \left( \frac{\partial }{\partial {\varvec{\theta }}} D_{t}^{-1}({\varvec{\theta }}_0) \right) ^{\top }, \\ \hat{a}_{32}(h)= & {} n^{-1} \sum _{t=|h|+1}^n Z_{t-|h|} X_t \frac{\partial ^2}{\partial {\varvec{\theta }} \partial {\varvec{\theta }}^{\top }} D_{t}^{-1}(\bar{{\varvec{\theta }}}_0). \end{aligned}$$

We write \(\hat{A}_3(h) = \hat{A}_{31}(h) + \hat{A}_{32}(h)\), where

$$\begin{aligned} \hat{A}_{31}(h)= & {} \hat{a}_{31}(h) (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0), \\ \hat{A}_{32}(h)= & {} \frac{1}{2} (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0)^{\top }\hat{a}_{32}(h) (\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0). \end{aligned}$$

We obtain

$$ 2\pi n \sum _{j=0}^J \sum _{k=1}^{2^j} B_{3jk}^2 \le 4 n \sum _{j=0}^J \sum _{k=1}^{2^j} \left[ \sum _{h=1-n}^{n-1} \hat{A}_{31}(h)\hat{\psi }_{jk}(2\pi h) \right] ^2 + 4 n \sum _{j=0}^J \sum _{k=1}^{2^j} \left[ \sum _{h=1-n}^{n-1} \hat{A}_{32}(h)\hat{\psi }_{jk}(2\pi h) \right] ^2. $$

We write \(\hat{a}_{31}(h) = E[\hat{a}_{31}(h)] + \left\{ \hat{a}_{31}(h)-E[\hat{a}_{31}(h)]\right\} \). We have

$$\begin{aligned} n \sum _{j=0}^J \sum _{k=1}^{2^j} \left[ \sum _{h=1-n}^{n-1} \hat{A}_{31}(h)\hat{\psi }_{jk}(2\pi h) \right] ^2\le & {} 2 n ||\hat{{\varvec{\theta }}} - {\varvec{\theta }}_0||^2 \left\{ \sum _{j=0}^J \sum _{k=1}^{2^j} \left\| \sum _h E[\hat{a}_{31}(h)] \hat{\psi }_{jk}(2\pi h) \right\| ^2 \right. \\&\left. + \sum _{j=0}^J \sum _{k=1}^{2^j} \left\| \sum _h [\hat{a}_{31}(h)-E\hat{a}_{31}(h)] \hat{\psi }_{jk}(2\pi h) \right\| ^2 \right\} . \\ \end{aligned}$$

Since we can interpret \(E\hat{a}_{31}(h) = {\text {cov}}(Z_{t-|h|}, X_t \frac{\partial }{\partial {\varvec{\theta }}} D_t^{-1}({\varvec{\theta }}_0))\) as a cross-correlation function, we have that

$$ \sum _{j=0}^J \sum _{k=1}^{2^j} \left\| \sum _h E[\hat{a}_{31}(h)] \hat{\psi }_{jk}(2\pi h) \right\| ^2 = O(1). $$

Also,

$$\begin{aligned}&\sum _{j=0}^J \sum _{k=1}^{2^j} E \left\| \sum _h [\hat{a}_{31}(h)-E\hat{a}_{31}(h)] \hat{\psi }_{jk}(2\pi h) \right\| ^2 \\= & {} | \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} b_J(h,m) E [ (\hat{a}_{31}(h)-E\hat{a}_{31}(h))(\hat{a}_{31}(m)-E\hat{a}_{31}(m))^{\top }] |, \\\le & {} \sum _{h=1}^{n-1}\sum _{m=1}^{n-1} |b_J(h,m)| \left\{ E\Vert \hat{a}_{31}(h)-E\hat{a}_{31}(h) \Vert ^2 \right\} ^{1/2} \left\{ E\Vert \hat{a}_{31}(m)-E\hat{a}_{31}(m)\Vert ^2 \right\} ^{1/2}, \\\le & {} \sup _h E\Vert \hat{a}_{31}(h)-E\hat{a}_{31}(h) \Vert ^2 \sum _h \sum _m |b_J(h,m)| = O(2^J/n). \end{aligned}$$

This shows Proposition 4. \(\Box \)

Remark 1

Proposition 4 is established under a general stationary process for \(\{ Z_t \}\), that is, the result is established without assuming the null hypothesis.

Proof

(Proposition 5) We write \((\hat{\alpha }_{jk}-\tilde{\alpha }_{jk})\tilde{\alpha }_{jk}=\hat{C}_{1jk}+\hat{C}_{2jk}+\hat{C}_{3jk}\), where

$$\begin{aligned} \hat{C}_{1jk}= & {} \sum _{h=1-n}^{n-1}\hat{A}_{1}(h)\hat{\psi }_{jk}(2\pi h) \tilde{\alpha }_{jk}, \end{aligned}$$
(42)
$$\begin{aligned} \hat{C}_{2jk}= & {} \sum _{h=1-n}^{n-1}\hat{A}_{2}(h)\hat{\psi }_{jk}(2\pi h) \tilde{\alpha }_{jk}, \end{aligned}$$
(43)
$$\begin{aligned} \hat{C}_{3jk}= & {} \sum _{h=1-n}^{n-1}\hat{A}_{3}(h)\hat{\psi }_{jk}(2\pi h) \tilde{\alpha }_{jk}. \end{aligned}$$
(44)

By the Cauchy–Schwarz inequality and the fact that \(n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\tilde{\alpha }_{jk}^{2}=O_{P}(2^{J})\) under the null hypothesis, we have that

$$ n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\sum _{h=1-n}^{n-1}\hat{A}_{1}(h)\hat{\psi }_{jk}(2\pi h)\tilde{\alpha }_{jk}=O_{P}(2^{J}/n^{1/2}). $$

Since \(n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}B_{2jk}^{2}=O_{P}(2^{J}/n)\), we have that

$$ n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\hat{C}_{2jk}=O_{P}(2^{J}/n^{1/2}). $$

We write

$$\begin{aligned} \hat{C}_{3jk}= & {} \left[ \sum _{h=1-n}^{n-1}\hat{a}_{31}(h)\hat{\psi }_{jk}(2\pi h)\tilde{\alpha }_{jk} \right] (\hat{{\varvec{\theta }}}-{\varvec{\theta }}_{0}) \\+ & {} \frac{1}{2}(\hat{{\varvec{\theta }}}-{\varvec{\theta }}_{0})^{\top } \left\{ \sum _{h=1-n}^{n-1}\hat{a}_{32}(h)\hat{\psi }_{jk}(2\pi h)\tilde{\alpha }_{jk}\right\} (\hat{{\varvec{\theta }}}-{\varvec{\theta }}_{0}). \end{aligned}$$

We have

$$ 2\pi n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\sum _{h=1-n}^{n-1}E[ \hat{a}_{31}(h)] \hat{\psi }_{jk}(2\pi h)\tilde{\alpha }_{jk}= n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}E[\hat{a}_{31}(h)]\tilde{R}(m)b_{J}(h,m). $$

Since

$$\begin{aligned}&n^{2}E\left\| \sum _{h=1}^{n-1}\sum _{m=1}^{n-1}E[\hat{a}_{31}(h)]\tilde{R}(m)b_{J}(h,m)\right\| ^{2} \\\le & {} Cn\sum _{h_{1}=1}^{n-1}\sum _{h_{2}=1}^{n-1}\sum _{m=1}^{n-1}E[\hat{a}_{31}(h_{1})]E[\hat{a}_{31}(h_{2})]^{\top }|b_{J}(h_{1},m)b_{J}(h_{2},m)|, \\\le & {} Cn\left( \sum _{h_{1}=1}^{n-1}\sum _{h_{2}=1}^{n-1}[\sum _{m=1}^{n-1}|b_{J}(h_{1},m)b_{J}(h_{2},m)|]^{2}\right) ^{1/2}=O(n(J+1)^{1/2}2^{(J+1)/2}). \end{aligned}$$

Then

$$ n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}E[\hat{a}_{31}(h)]\tilde{R}(m)b_{J}(h,m)=O_{P}(n^{1/2}J^{1/4}2^{J/4}). $$

We have

$$\begin{aligned}&E\Vert n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}\sum _{h}[\hat{a}_{31}(h)-E\hat{a}_{31}(h)]\hat{\psi }_{jk}(h)\tilde{\alpha }_{jk}\Vert \\\le & {} \frac{n}{2\pi }\sum _{h=1}^{n-1}\sum _{m=1}^{n-1} E \left[ \Vert \hat{a}_{31}(h)-E\hat{a}_{31}(h) \Vert |\tilde{R}(m)| \right] |b_{J}(h,m)|, \\\le & {} \frac{n}{2\pi }\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}\left[ E\Vert \hat{a}_{31}(h)-E\hat{a}_{31}(h)\Vert ^{2}\right] ^{1/2}[E\tilde{R}^{2}(m)]^{1/2}|b_{J}(h,m)|, \\= & {} O(n\;n^{-1/2}\;n^{-1/2}\;2^{J+1})=O(2^{J+1}). \end{aligned}$$

This completes the proof of Proposition 5 and so Theorem 4. \(\Box \)

Proof

(Theorem 5) We write

$$ \mathscr {A}(\hat{J})-\mathscr {A}(J)= [D_n(\hat{J})]^{-1/2}\left\{ 2\pi n\sum _{j=J}^{\hat{J}}\sum _{k=1}^{2^{j}}\hat{\alpha }_{jk}^{2}- [C_{n}(\hat{J})-C_{n}(J)]\right\} -\left\{ 1-[D_{n}(J)]^{1/2}/[D_{n}(\hat{J})]^{1/2} \right\} \mathscr {A}(J), $$

where \(C_{n}(J)=2^{J+1}-1\), \(D_{n}(J)=4(2^{J+1}-1)\). Given \(\mathscr {A}(J)=O_{P}(1)\) by Theorem 4, it suffices for \(\mathscr {A}(\hat{J})-\mathscr {A}(J)\rightarrow ^p 0\) and \(\mathscr {A}(\hat{J}) \rightarrow ^d N(0,1)\) to establish

$$\begin{aligned} (i)&[D_{n}(J)]^{-1/2}\left\{ 2\pi n\sum _{j=J}^{\hat{J}}\sum _{k=1}^{2^{j}} \hat{\alpha }_{jk}^{2}-[C_{n}(\hat{J})-C_{n}(J)]\right\} \rightarrow ^{p} 0, \\ (ii)&D_{n}(\hat{J})/D_{n}(J)\rightarrow ^{p} 1. \end{aligned}$$

We first show (i). Decompose

$$\begin{aligned} 2\pi n\sum _{j=J}^{\hat{J}}\sum _{k=1}^{2^{j}}\hat{\alpha }_{jk}^{2}= & {} 2 \pi n\sum _{j=J}^{\hat{J}}\sum _{k=1}^{2^{j}}(\hat{\alpha }_{jk}-\tilde{\alpha }_{jk})^{2}+2\pi n\sum _{j=J}^{\hat{J}}\sum _{k=1}^{2^{j}}\tilde{\alpha }_{jk}^{2}+ 4\pi n\sum _{j=J}^{\hat{J}}\sum _{k=1}^{2^{j}}(\hat{\alpha }_{jk}-\tilde{\alpha }_{jk})\tilde{\alpha }_{jk}, \nonumber \\= & {} \hat{G}_{1}+\hat{G}_{2}+2\hat{G}_{3}. \end{aligned}$$
(45)

For the first term in (45), we write

$$\begin{aligned} \hat{G}_{1}=2\pi n\sum _{j=0}^{\hat{J}}\sum _{k=1}^{2^{j}}(\hat{\alpha }_{jk}-\tilde{\alpha }_{jk})^{2}- 2\pi n\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}(\hat{\alpha }_{ijk}-\tilde{\alpha }_{jk})^{2}=\hat{G}_{11}-\hat{G}_{12}. \end{aligned}$$
(46)

By Proposition 4, we have \([D_{n}(J)]^{-1/2}\hat{G}_{12}\rightarrow ^p 0\). For the first term in (46), we have for any given constants \(M>0\) and \(\varepsilon >0\),

$$\begin{aligned} P\left( \hat{G}_{11}>\varepsilon \right) \le P\left( \hat{G}_{11}>\varepsilon , C_{0}2^{J/2}|2^{\hat{J}}/2^{J}-1|\le \varepsilon \right) +P\left( C_{0}2^{J/2}|2^{\hat{J}}/2^{J}-1|>\varepsilon \right) . \end{aligned}$$
(47)

For any given constants \(C_{0},\varepsilon >0,\) the second term in (47) vanishes to 0 as \(n \rightarrow \infty \) given \(2^{J/2}|2^{\hat{J}}/2^{J}-1|\rightarrow ^{p} 0.\) For the first term, given \(C_{0}2^{J/2}|2^{\hat{J}}/2^{J}-1|\le \varepsilon \), we have for n sufficiently large,

$$\begin{aligned}{}[D_{n}(J)]^{-1/2}\hat{G}_{11}\le & {} [D_{n}(J)]^{-1/2}2\pi n\sum _{j=0}^{[\log _{2}2^{J}(1+\varepsilon /(C_{0}2^{J/2}))]}\sum _{k=1}^{2^{j}}(\hat{\alpha }_{jk}-\tilde{\alpha }_{jk})^{2}, \\\le & {} [D_{n}(J)]^{-1/2}2\pi n\sum _{j=0}^{J+1}\sum _{k=1}^{2^{j}}(\hat{\alpha }_{jk}-\tilde{\alpha }_{jk})^{2}=o_{P}(1). \end{aligned}$$

by Proposition 4. Therefore, we have

$$\begin{aligned}{}[D_{n}(J)]^{-1/2}\hat{G}_{1}=o_{P}(1). \end{aligned}$$
(48)

Next, we consider \(\hat{G}_{2}\) in (45). We write

$$ \hat{G}_{2}=n\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}\tilde{R}_{e}(h)\tilde{R}_{e}(m)\left[ b_{\hat{J}}(h,m)-b_{J}(h,m)\right] . \nonumber $$

Since for n sufficiently large,

$$\begin{aligned}&\sum _{h=1-n}^{n-1}\sum _{m=1-n}^{n-1}\left| a_{J}(h,m)-a_{\hat{J}}(h,m)\right| \\\le & {} C\sum _{h=1-n}^{n-1}\sum _{m=1-n}^{n-1}\sum _{j=[\log _{2}2^{J}(1-\varepsilon /(C_{0}2^{J/2}))]}^{[\log _{2}2^{J}(1+\varepsilon /(C_{0}2^{J/2}))]}\left| c_{j}(h,m)\hat{\psi }(2\pi h/2^{j})\hat{\psi }^{*}(2\pi m/2^{j})\right| , \\\le & {} C\sum _{j=[\log _{2}2^{J}(1-\varepsilon /(C_{0}2^{J/2}))]}^{[\log _{2}2^{J}(1+\varepsilon /(C_{0}2^{J/2}))]}2^{j}\left[ 2^{-j}\sum _{h=1-\bar{T}}^{\bar{T}-1}|\hat{\psi }(2\pi h/2^{j})|\right] , \\&\times \left[ \sum _{r=-\infty }^{\infty }|\hat{\psi }(2\pi h/2^{j}+2\pi r)|\right] , \\\le & {} C2^{J}\varepsilon /(C_{0}2^{J/2}), \end{aligned}$$

given Assumption 3 and

$$\begin{aligned} c_{j}(h,m)=\left\{ \begin{array}{ll} 1 &{} \text {if }m-h=2^{j}r\text { for some }r\in \mathbb {Z}, \\ 0 &{} \text {otherwise.} \end{array} \right. \end{aligned}$$
(49)

Cf. [61, (6.19), p.392]. Therefore

$$\begin{aligned} E|\hat{G}_{2}|\le & {} n\sum _{h=1-n}^{n-1}\sum _{m=1-n}^{n-1}E|\tilde{R}_{e}(h)\tilde{R}_{e}(m)||a_{\hat{J}}(h,m)-a_{J}(h,m)|, \\\le & {} n\sup _{h}{\text {var}}{\tilde{R}_{e}(h)}\sum _{h=1-n}^{n-1}\sum _{m=1-n}^{n-1}|a_{\hat{J}}(h,m)-a_{J}(h,m)|, \\\le & {} C2^{J/2}\varepsilon /C_{0}. \end{aligned}$$

Then \([D_{n}(J)]^{-1/2}\left\{ \hat{G}_{2}-[C(\hat{J})-C(J)]\right\} \rightarrow ^p 0\). Note that \([D_{n}(J)]^{-1/2}[C_{n}(\hat{J})-C_{n}(J)]=o_{P}(1)\). Next, by the Cauchy-Schwarz inequality and (48), we have

$$ [D_{n}(J)]^{-1/2}|\hat{G}_{3}|\le \left( [D_{n}(J)]^{-1/2}\hat{G}_{1}\right) ^{1/2}\left( [D_{n}(J)]^{-1/2}|\hat{G}_{2}|\right) ^{1/2}=o_{P}(1). $$

Summarizing, we obtain result (i), that is

$$ [D_{n}(J)]^{-1/2}\left\{ 2\pi n\sum _{j=J}^{\hat{J}}\sum _{k=1}^{2^{j}}\hat{\alpha }_{jk}^{2}-[C_n(\hat{J})-C_n(J)]\right\} =o_{P}(1). $$

We now show (ii), that is \(D_{n}(\hat{J})/D_n(J)=1+o_{P}(1)\). Using the fact that \(2^{\hat{J}}/2^J = 1 + o_P(2^{-J/2})\) and \(J \rightarrow \infty \) such that \(2^J/n \rightarrow 0\), one shows easily that

$$ \frac{D_{n}(\hat{J})}{D_n(J)} = \frac{2^{\hat{J}+1} - 1}{2^{J+1} - 1} \rightarrow ^p 1. $$

This shows (ii). This completes the proof of Theorem 5. \(\Box \)

Proof

(Theorem 6) We first show \(Q(\hat{f},f)=Q(\tilde{f},f)+o_{P}(2^{J}/T+2^{-2qJ})\). Write

$$\begin{aligned} Q(\hat{f},f)-Q(\tilde{f},f)= & {} Q(\hat{f},\tilde{f})+2\int _{-\pi }^{\pi }[\hat{f}(\omega )-\tilde{f}(\omega )][\tilde{f}(\omega )-f(\omega )]d\omega , \nonumber \\= & {} \hat{Q}_{1}+2\hat{Q}_{2}. \end{aligned}$$
(50)

For the first term in (50), by Parseval’s identity, Proposition 4 (which can be shown to continues to hold given Assumptions 23 and 710; See Remark 1), and \(D_{n}(J)\propto O(2^{J+1})\), we have

$$\begin{aligned} \hat{Q}_{1}=\sum _{j=0}^{J}\sum _{k=1}^{2^{j}}(\hat{\alpha }_{jk}-\tilde{\alpha }_{jk})^{2}=O_{P}[n^{-1}+2^{J}n^{-2}]=o_{P}(2^{J}/n), \end{aligned}$$
(51)

as \(n\rightarrow \infty \). For the second term, we have \(\hat{Q}_{2}=o_{P}(2^{J}/n+2^{-2qJ})\) by the Cauchy-Schwarz inequality, (51) and the fact that \(Q(\tilde{f},f)=O_{P}(2^{J}/n+2^{-2qJ})\), which follows by Markov’s inequality and \(EQ(\tilde{f},f)=O(2^{J}/n+2^{-2qJ})\). The latter is to be shown below.

To compute \(E[Q(\tilde{f},f)],\) we write

$$\begin{aligned} E[Q(\tilde{f},f)]= E[Q(\tilde{f},E\tilde{f})] + Q[E(\tilde{f}),f]. \end{aligned}$$
(52)

We first consider the second term in (52). Put \(B(\omega ) \equiv \sum _{j=J+1}^{\infty }\sum _{k=1}^{2^{j}}\alpha _{jk}\varPsi _{jk}(\omega ) \). Then

$$\begin{aligned} Q[E(\tilde{f}),f]= \int _{-\pi }^{\pi }B^{2}(\omega )d\omega +\sum _{j=0}^{J}\sum _{k=1}^{2^{j}} (E\tilde{\alpha }_{jk}-\alpha _{jk})^{2}. \end{aligned}$$
(53)

We evaluate directly \(\int _{-\pi }^{\pi }B^{2}(\omega )d\omega \). Using the orthonormality of the wavelet basis, we have that

$$ \int _{-\pi }^{\pi }B^{2}(\omega )d\omega = \sum _{j=J+1}^{\infty }\sum _{k=1}^{2^j}\alpha _{jk}^2. $$

Replacing \(\alpha _{jk}=\sum _{h=-\infty }^{\infty }R_e(h)\hat{\psi }_{jk}(2 \pi h)\) and since \(\hat{\psi }_{jk}(2 \pi h)=e^{-i 2\pi h k /2^j} 2^{-j/2}\hat{\psi }(2\pi h/2^j)\),

$$\begin{aligned} \sum _{j=J+1}^{\infty } \sum _{k=1}^{2^j} \alpha _{jk}^2= & {} \sum _{j=J+1}^{\infty } \sum _{h=-\infty }^{\infty }\sum _{m=-\infty }^{\infty } R_e(h)R_e(m) \{ 2^{-j} \sum _{k=1}^{2^j} e^{i 2\pi (m-h) k/2^j} \} \hat{\psi }(2\pi h/2^j)\hat{\psi }^{*}(2\pi m/2^j), \\= & {} \sum _{j=J+1}^{\infty }\sum _{h=-\infty }^{\infty }\sum _{m=-\infty }^{\infty } c_j(h,m)R_e(h)R_e(m)\hat{\psi }(2\pi h/2^j)\hat{\psi }^{*}(2\pi m/2^j), \end{aligned}$$

where \(c_j(h,m) = 2^{-j} \sum _{k=1}^{2^j} e^{i 2\pi (m-h) k/2^j}\) is as in (49). By a change of variables,

$$\begin{aligned} \sum _{j=J+1}^{\infty } \sum _{k=1}^{2^j} \alpha _{jk}^2 = \sum _{j=J+1}^{\infty } \sum _{h=-\infty }^{\infty } \sum _{r=-\infty }^{\infty } R_e(h)R_e(h+2^j r) \hat{\psi }(2\pi h/2^j)\hat{\psi }^{*}(2\pi h/2^j + 2\pi r). \end{aligned}$$
(54)

We evaluate separately the case corresponding to \(r=0\) and \(r \ne 0\) in (54).

$$\begin{aligned}&\sum _{j=J+1}^{\infty }\sum _{h=-\infty }^{\infty } R_e^2(h) \hat{\psi }(2\pi h/2^j)\hat{\psi }^{*}(2\pi h/2^j) \\= & {} \sum _{j=J+1}^{\infty }\sum _{h=-\infty }^{\infty } R_e^2(h) |\hat{\psi }(2\pi h/2^j)|^2, \\= & {} \sum _{j=J+1}^{\infty }\sum _{h=-\infty }^{\infty } R_e^2(h)|2\pi h/2^j|^{2q} \frac{|\hat{\psi }(2\pi h/2^j)|^2}{|2\pi h/2^j|^{2q}}, \\= & {} \lim _{z \rightarrow 0} \frac{|\hat{\psi }(z)|^2}{|z|^{2q}}[1+o(1)](2\pi )^{2q} \sum _{j=J+1}^{\infty }\sum _{h=-\infty }^{\infty }|h|^{2q}R_e^2(h)(2^{-2q})^j, \\= & {} \lim _{z \rightarrow 0} \frac{|\hat{\psi }(z)|^2}{|z|^{2q}}[1+o(1)] (2\pi )^{2q} \sum _{j=J+1}^{\infty } (2^{-2q})^j \sum _{h=-\infty }^{\infty } |h|^{2q} R_e^2(h), \\= & {} (2\pi )^{2q+1} \lim _{z \rightarrow 0} \frac{|\hat{\psi }(z)|^2}{|z|^{2q}} [1+o(1)] \frac{2^{-2q(J+1)}}{1-2^{-2q}} \int _{-\pi }^{\pi } [f_e^{(q)}(\omega )]^2 d\omega , \end{aligned}$$

where \(f^{(q)}_e(\cdot )\) is defined in Sect. 4.3 and o(1) is uniform in \(\omega \in [-\pi ,\pi ].\) It follows that

$$\begin{aligned} \int _{-\pi }^{\pi } B^{2}(\omega )d\omega =2^{-2q(J+1)}\vartheta _{q}\int _{-\pi }^{\pi }\left[ f_e^{(q)}(\omega )\right] ^{2}d\omega +o(2^{-2qJ}). \end{aligned}$$
(55)

It may be show that the term corresponding to \(r\ne 0\) is \(o(2^{-2qJ})\).

For the second term in (53), we have

$$\begin{aligned} \sum _{j=0}^{J}\sum _{k=1}^{2^{j}} (E\tilde{\alpha }_{jk}-\alpha _{jk})^{2}= & {} \sum _{j=0}^{J}\sum _{k=1}^{2^{j}} \left[ n^{-1}\sum _{h=1-n}^{n-1}|h|R_e(h)\hat{\psi }_{jk} (2\pi h)+\sum _{|h|\ge n}R_e(h)\hat{\psi }_{jk}(2\pi h)\right] ^{2}, \nonumber \\\le & {} 4Cn^{-2}\sum _{h=1}^{n-1}\sum _{m=1}^{n-1}|h m R_e(h)R_e(m) b_{J}(h,m)|, \nonumber \\= & {} O[(J+1)/n^{2}), \end{aligned}$$
(56)

given Lemma 3(vii) and \(\sum _{h=-\infty }^{\infty }|hR_e(h)|\le C\) as implied by Assumption 10.

Finally, we consider the variance factor in (52). We write

$$\begin{aligned} E[Q(\tilde{f},E\tilde{f})]= & {} \sum _{h=1-n}^{n-1} \sum _{m=1-n}^{n-1}b_{J}(h,m) {\text {cov}}[\tilde{R}_e(h),\tilde{R}_e(m)], \\= & {} \sum _{h=1-n}^{n-1} \sum _{m=1-n}^{n-1}b_{J}(h,m)n^{-1}\sum _{l} \left[ 1-\frac{\eta (l)+m}{n}\right] \\&\times \left[ R_e(l)R_e(l+m-h)+R_e(l+m)R_e(l-h)+ \kappa (l,h,m-h)\right] , \\\equiv & {} V_{1n}+V_{2n}+V_{3n},\text { say,} \end{aligned}$$

where the function \(\eta (l)\) satisfies

$$ \eta (l)\equiv \left\{ \begin{array}{ll} l, &{} \text {if }l>0, \\ 0, &{} \text {if }h-m\le l\le 0, \\ -l+h-m, &{} \text {if }-(n-h)+1\le l\le h-m. \end{array} \right. $$

For more details see [61, p. 326]. Given Assumption 9 and Lemma 3(vii), we have \(|V_{2n}|\le C(J+1)n^{-1}\) and \(|V_{3n}|\le C(J+1)n^{-1}\). For the first term \(V_{1n}\), we can write

$$\begin{aligned} V_{1n}= & {} \sum _{h=1-n}^{n-1} b_{J}(h,h)n^{-1}\sum _{l=-\infty }^{\infty }(1-|l|/n)R_e^{2}(l)+\sum _{h}\sum _{|r|=1}^{n-1}b_{J}(h,h+r)n^{-1}\sum _{l=-\infty }^{\infty }R_e(l)R_e(l+r), \\= & {} n^{-1}(2^{J+1}-1) \sum _{h=-\infty }^{\infty }R_e^{2}(h)+ O[(J+1)/n], \end{aligned}$$

where we have used Lemma 3(v) for the first term, which corresponds to \(h=m \); the second term corresponds to \(h\ne m\) and it is \(O[(J+1)/T]\) given \(\sum _{h=-\infty }^{\infty }|R(h)|\le C\) and Lemma 3(v). It follows that as \(J\rightarrow \infty \)

$$\begin{aligned} E[Q(\tilde{f},E\tilde{f})]= \frac{2^{J+1}}{n} \int _{-\pi }^{\pi }f_e^{2}(\omega )d\omega +o(2^{J}/n). \end{aligned}$$
(57)

Collecting (55)–(57) and \(J\rightarrow \infty \), we obtain

$$ E[Q(\hat{f},f)] = \frac{2^{J+1}}{n}\int _{-\pi }^{\pi }f_e^{2}(\omega )d\omega +2^{-2qJ}\vartheta _{q} \int _{-\pi }^{\pi }[f_e^{(q)}(\omega )]^{2}d\omega +o(2^{J}/n+2^{-2qJ}). $$

This shows the Theorem. \(\Box \)

Proof

(Corollary 2) The result follows immediately from Theorem 6 because Assumption 9 implies \(2^{\hat{J}}/2^{J}-1=o_{P}(T^{-1/2(2q+1)})=o_{P}(2^{-J/2})\), where the nonstochastic finest scale J is given by \(2^{J+1}\equiv \max \{[2\alpha \vartheta _{q}\zeta _{0}(q)T]^{1/(2q+1)},0\}.\) The latter satisfies the conditions of Theorem 6. \(\Box \)

Rights and permissions

Reprints and permissions

Copyright information

© 2016 Springer Science+Business Media New York

About this chapter

Cite this chapter

Duchesne, P., Hong, Y. (2016). On Diagnostic Checking Autoregressive Conditional Duration Models with Wavelet-Based Spectral Density Estimators. In: Li, W., Stanford, D., Yu, H. (eds) Advances in Time Series Methods and Applications . Fields Institute Communications, vol 78. Springer, New York, NY. https://doi.org/10.1007/978-1-4939-6568-7_3

Download citation

Publish with us

Policies and ethics