Skip to main content
Log in

A new thinning-based INAR(1) process for underdispersed or overdispersed counts

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

Abstract

Underdispersed and overdispersed phenomena are often observed in practice. To deal with these phenomena, we introduce a new thinning-based integer-valued autoregressive process. Some probabilistic and statistical properties of the process are obtained. The asymptotic normality of the estimators of the model parameters, using conditional least squares, weighted conditional least squares and modified quasi-likelihood methods, are presented. One overdispersed real-data example and one underdispersed real-data example are given to show the flexibility and superiority of the new model.

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
Fig. 5

Similar content being viewed by others

References

  • Alzaid, A. A., & Al-Osh, M. A. (1987). First-order integer-valued autoregressive (INAR(1)) process. Journal of Time Series Analysis, 8, 261–275.

    Article  MathSciNet  Google Scholar 

  • Barreto-Souza, W. (2015). Zero-modified geometric INAR(1) process for modelling count time series with deflation or inflation of zeros. Journal of Time Series Analysis, 36, 839–852.

    Article  MathSciNet  Google Scholar 

  • Barreto-Souza, W. (2017). Mixed Poisson INAR(1) processes. Statistical Papers,. https://doi.org/10.1007/s00362-017-0912-x. (in press).

    Article  MATH  Google Scholar 

  • Borges, P., Molinares, F. F., & Bourguignon, M. (2016). A geometric time series model with inflated-parameter Bernoulli counting series. Statistics and Probability Letters, 119, 264–272.

    Article  MathSciNet  Google Scholar 

  • Bourguignon, M., Rodrigues, J., & Santosneto, K. (2019). Extended Poisson INAR(1) processes with equidispersion, underdispersion and overdispersion. Journal of Applied Statistics, 46, 101–118.

    Article  MathSciNet  Google Scholar 

  • Bourguignon, M., & Vasconcellos, K. L. (2015). First order non-negative integer valued autoregressive processes with power series innovations. Brazilian Journal of Probability and Statistics, 29, 71–93.

    Article  MathSciNet  Google Scholar 

  • Bourguignon, M., & Weiß, C. H. (2017). An INAR(1) process for modeling count time series with equidispersion, underdispersion and overdispersion. Test, 26, 847–868.

    Article  MathSciNet  Google Scholar 

  • Cossette, H., Marceau, É., & Toureille, F. (2011). Risk models based on time series for count random variables. Insurance: Mathematics and Economics, 48, 19–28.

    MathSciNet  MATH  Google Scholar 

  • Freeland, R. K., & McCabe, B. P. M. (2004). Analysis of low count time series data by Poisson autoregression. Journal of Time Series Analysis, 25, 701–722.

    Article  MathSciNet  Google Scholar 

  • Gómez-Déniz, E., Sarabia, J. M., & Calderín-Ojeda, E. (2011). A new discrete distribution with actuarial applications. Insurance: Mathematics and Economics, 48, 406–412.

    MathSciNet  MATH  Google Scholar 

  • Grunwald, G., Hyndman, R. J., Tedesco, L., & Tweedie, R. L. (2000). Non-Gaussian conditional linear AR(1) models. Australian and New Zealand Journal of Statistics, 42, 479–495.

    Article  MathSciNet  Google Scholar 

  • Jazi, M. A., Jones, G., & Lai, C. D. (2012). First-order integer valued AR processes with zero inflated poisson innovations. Journal of Time Series Analysis, 33, 954–963.

    Article  MathSciNet  Google Scholar 

  • Karlsen, H., & Tjøstheim, D. (1988). Consistent estimates for the Near(2) and Nlar(2) time series models. Journal of the Royal Statistical Society. Series B (Statistical Methodology), 50, 313–320.

    MathSciNet  Google Scholar 

  • Kim, H. Y., & Lee, S. (2017). On first-order integer-valued autoregressive process with Katz family innovations. Journal of Statistical Computation and Simulation, 87, 546–562.

    Article  MathSciNet  Google Scholar 

  • Li, C., Wang, D., & Zhang, H. (2015). First-order mixed integer-valued autoregressive processes with zero-inflated generalized power series innovations. Journal of the Korean Statistical Society, 44, 232–246.

    Article  MathSciNet  Google Scholar 

  • Ristić, M. M., Bakouch, H. S., & Nastić, A. S. (2009). A new geometric first-order integer-valued autoregressive (NGINAR(1)) process. Journal of Statistical Planning and Inference, 139, 2218–2226.

    Article  MathSciNet  Google Scholar 

  • Schweer, S., & Weiß, C. H. (2014). Compound Poisson INAR (1) processes: Stochastic properties and testing for overdispersion. Computational Statistics and Data Analysis, 77, 267–284.

    Article  MathSciNet  Google Scholar 

  • Scotto, M. G., Weiß, C. H., & Gouveia, S. (2015). Thinning-based models in the analysis of integer-valued time series: A review. Statistical Modelling, 15, 590–618.

    Article  MathSciNet  Google Scholar 

  • Steutel, F. W., & Van Harn, K. (1979). Discrete analogues of self-decomposability and stability. The Annals of Probability, 7, 893–899.

    Article  MathSciNet  Google Scholar 

  • Weiß, C. H. (2007). Controlling correlated processes of Poisson counts. Quality and Reliability Engineering International, 23, 741–754.

    Article  Google Scholar 

  • Weiß, C. H. (2008a). Serial dependence and regression of Poisson INARMA models. Journal of Statistical Planning and Inference, 138, 2975–2990.

    Article  MathSciNet  Google Scholar 

  • Weiß, C. H. (2008b). Thinning operations for modeling time series of counts-a survey. AStA Advances in Statistical Analysis, 92, 319–341.

    Article  MathSciNet  Google Scholar 

  • Weiß, C. H., Homburg, A., & Puig, P. (2019). Testing for zero inflation and overdispersion in INAR(1) models. Statistical Papers, 60, 473–498.

    Article  MathSciNet  Google Scholar 

  • Zhang, H., Wang, D., & Zhu, F. (2010). Inference for INAR(\(p\)) processes with signed generalized power series thinning operator. Journal of Statistical Planning and Inference, 140, 667–683.

    Article  MathSciNet  Google Scholar 

  • Zheng, H., Basawa, I. V., & Datta, S. (2007). First-order random coefficient integer-valued autoregressive processes. Journal of Statistical Planning and Inference, 137, 212–229.

    Article  MathSciNet  Google Scholar 

  • Zhu, F. (2012a). Modeling overdispersed or underdispersed count data with generalized Poisson integer-valued GARCH models. Journal of Mathematical Analysis and Applications, 389, 58–71.

    Article  MathSciNet  Google Scholar 

  • Zhu, F. (2012b). Modeling time series of counts with COM-Poisson INGARCH models. Mathematical and Computer Modelling, 56, 191–203.

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

We gratefully acknowledge the associate editor and anonymous reviewers for their serious work and thoughtful suggestions that have helped us improve this paper substantially. This work is supported by National Natural Science Foundation of China (Nos. 11731015, 11571051, 11501241, 11871028), Natural Science Foundation of Jilin Province (Nos. 20150520053JH, 20170101057JC, 20180101216JC), Program for Changbaishan Scholars of Jilin Province (2015010), and Science and Technology Program of Jilin Educational Department during the “13th Five-Year” Plan Period (No. 2016316).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dehui Wang.

Additional information

Publisher’s Note

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

Appendix

Appendix

As we mentioned in the third paragraph of Sect. 2, the infinite sum in the mean and variance for GSC(\(\alpha \), \(\theta \)) are convergent, where \(\alpha <1\), \(\alpha \ne 0\) and \(0<\theta <1\). We illustrate it, as follows:

Let \(\alpha <0\). Denote \(S_n=\sum _{s=1}^{n}\log (1-\alpha \theta ^{s})\). Then, we have

$$\begin{aligned} 0{\le }\lim _{n\rightarrow \infty }\sup _{p{>}0}|S_{n+p}{-}S_n|{=}\lim _{n\rightarrow \infty }\sup _{p{>}0}\sum _{s{=}n+1}^{n+p}\log (1{-}\alpha \theta ^{s}){\le }\lim _{n\rightarrow \infty }\sup _{p{>}0} [p\log (1{-}\alpha \theta ^{n})]{=}0. \end{aligned}$$

By the Cauchy criterion of series, the infinite sum \(\sum _{s=1}^{\infty }\log (1-\alpha \theta ^{s})\) are convergent.

Denote \(S_n^{'}=\sum _{s=1}^{n}(2s-1)\log (1-\alpha \theta ^{s})\). Then, we have

$$\begin{aligned} 0&\le \lim _{n\rightarrow \infty }\sup _{p>0}|S_{n+p}^{'}-S_n^{'}|=\lim _{n\rightarrow \infty }\sup _{p>0}\sum _{s=n+1}^{n+p}(2s-1)\log (1-\alpha \theta ^{s})\\&\le \lim _{n\rightarrow \infty }\sup _{p>0}p\cdot 2(n+p)\log (1-\alpha \theta ^{n+1})\le \lim _{n\rightarrow \infty }\sup _{p>0}2p(n+p)(-\alpha )\theta ^{n+1}=0, \end{aligned}$$

using \(x\ge \log (1+x)\) for \(x\ge 0\). By the Cauchy criterion of series, the infinite sum \(\sum _{s=1}^{\infty }(2s-1)\log (1-\alpha \theta ^{s})\) are convergent. Following the same way, we can see that the two infinite sum are convergent when \(0<\alpha <1\). \(\square \)

Proof of Proposition 1

We have (i) and (iii), i.e.,

$$\begin{aligned} \mathrm {E}(X_{t}|X_{t-1})=\mathrm {E}(\alpha \diamond X_{t-1}+\epsilon _{t}|X_{t-1})=\phi X_{t-1}+\mu _{\epsilon } \end{aligned}$$

and

$$\begin{aligned} \mathrm {Var}(X_{t}|X_{t-1})&=\mathrm {Var}(\alpha \diamond X_{t-1}+\epsilon _{t}|X_{t-1})\\&=\mathrm {Var}(\alpha \diamond X_{t-1}|X_{t-1})+\mathrm {Var}(\epsilon _{t}|X_{t-1})\\&=\beta X_{t-1}+\sigma _{\epsilon }^2. \end{aligned}$$

Then, we get

$$\begin{aligned} \mathrm {E}(X_{t})=\mathrm {E}[\mathrm {E}(X_{t}|X_{t-1})]=\phi \mathrm {E}(X_{t-1})+\mu _{\epsilon } \end{aligned}$$

and

$$\begin{aligned} \mathrm {Var}(X_{t})&=\mathrm {Var}[\mathrm {E}(X_{t}|X_{t-1})]+\mathrm {E}[\mathrm {Var}(X_{t}|X_{t-1})]\\&=\mathrm {Var}(\phi X_{t-1}+\mu _{\epsilon })+\mathrm {E}(\beta X_{t-1}+\sigma _{\epsilon }^2)\\&=\phi ^2\mathrm {Var}(X_{t-1})+\beta \mathrm {E}(X_{t-1})+\sigma _{\epsilon }^2, \end{aligned}$$

which yield (ii) and (iv), due to the stationarity; \(\mathrm {E}(X_{t})=\mathrm {E}(X_{t-1})\) and \(\mathrm {Var}(X_{t})=\mathrm {Var}(X_{t-1})\). Moreover, we have (v), i.e.,

$$\begin{aligned} \mathrm {Cov}(X_{t},X_{t+k})&=\mathrm {Cov}(X_{t}, \underbrace{\alpha \diamond \cdots \diamond \alpha }_{k}\diamond X_{t})+\mathrm {Cov}(X_{t},\sum _{j=0}^{k-1} \underbrace{\alpha \diamond \cdots \diamond \alpha }_{j} \diamond \epsilon _{t+k-j})\\&=\mathrm {E}\{\mathrm {E}[X_{t}(\underbrace{\alpha \diamond \cdots \diamond \alpha }_{k}\diamond X_{t})|X_{t}]\} -\mathrm {E}(\underbrace{\alpha \diamond \cdots \diamond \alpha }_{k}\diamond X_{t})\cdot \mathrm {E}(X_{t})\\&=\phi ^k\{\mathrm {E}(X_t^2)-[\mathrm {E}(X_t)]^2\}\\&=\phi ^k\mathrm {Var}(X_t). \end{aligned}$$

\(\square \)

Proof of Theorem 1

We first introduce a random sequence \(\{X_{t}^{(n)}\}\),

$$\begin{aligned} X_{t}^{(n)}=\left\{ \begin{array}{ll} 0,&{} \quad n<0,\\ \epsilon _t,&{} \quad n=0,\\ \alpha \diamond X_{t-1}^{(n-1)}+\epsilon _{t},&{} \quad n>0,\\ \end{array} \right. \end{aligned}$$

where \(\mathrm {Cov}(X_{s}^{(n)},\epsilon _t)=0\) when \(s<t\) for any n.

As in Li et al. (2015), we can verify: existence of \(\{X_t\}\) satisfying (5), i.e., (A1) \(X_{t}^{(n)}\in L^2\), \(n>0\), (A2) \(X_t^{(n)}\) is a Cauchy sequence, (A3) \(\{X_t\}\) satisfies (5), uniqueness, strict stationarity and ergodicity. The details are omitted here to save space. \(\square \)

Proof of Theorem 2

From (6), solving \(\partial Q_1(\varvec{\eta })/\partial \alpha =0\) and \(\partial Q_1(\varvec{\eta })/\partial \mu _{\epsilon }=0\) lead to the CLS estimators of \(\alpha \) and \(\mu _{\epsilon }\). Now, let \(\mathcal {F}_n=\sigma \{X_0,X_1,\ldots ,X_n\}\), \(M_{n}^{(1)}=-\frac{1}{2}(\partial Q_1(\varvec{\eta })/\partial \alpha )=\sum _{t=1}^{n}\dot{\phi }X_{t-1}\big (X_t-\phi X_{t-1}-\mu _{\epsilon }\big )\), \(M_0^{(1)}=0\). Also, \(M_{n}^{(2)}=-\frac{1}{2}(\partial Q_1(\varvec{\eta })/\partial \mu _{\epsilon })=\sum _{t=1}^{n}\big (X_t-\phi X_{t-1}-\mu _{\epsilon }\big )\), \(M_0^{(2)}=0\). Then, it is easy to see that \(\{M_{n}^{(1)},\mathcal {F}_n\}_{n\ge 0}\) and \(\{M_{n}^{(2)},\mathcal {F}_n\}_{n\ge 0}\) are martingales. The martingale central limit theorem and Cramer-Wold’s device imply that

$$\begin{aligned} n^{-1/2}(M_n^{(1)},M_n^{(2)})^{'}\mathop {\longrightarrow }\limits ^{d}N(\mathbf {0},\varvec{V_{CLS}}). \end{aligned}$$

Using Taylor’s expansion, we have

$$\begin{aligned} \mathbf {0}{=-}\frac{1}{2\sqrt{n}}\frac{\partial Q_1(\widehat{\varvec{\eta }}_{CLS})}{\partial \varvec{\eta }}{=-}\frac{1}{2\sqrt{n}}\frac{\partial Q_1(\varvec{\eta })}{\partial \varvec{\eta }} -\frac{1}{2n}\frac{\partial ^2 Q_1(\varvec{\eta })}{\partial \varvec{\eta }\partial \varvec{\eta ^{'}}}\sqrt{n}(\widehat{\varvec{\eta }}_{CLS}{-}\varvec{\eta }){+}o_p(n^{-1/2}). \end{aligned}$$

Since we have proved that \(-\frac{1}{2\sqrt{n}}\frac{\partial Q_1(\varvec{\eta })}{\partial \varvec{\eta }} \mathop {\longrightarrow }\limits ^{d}N(\mathbf {0},\varvec{V_{CLS}})\), after some algebra, we have

$$\begin{aligned} \sqrt{n}(\widehat{\varvec{\eta }}_{CLS}-\varvec{\eta })\mathop {\longrightarrow }\limits ^{d}N(\mathbf {0},\varvec{H_{CLS}}^{-1}\varvec{V_{CLS}}\varvec{H_{CLS}}^{-1}). \end{aligned}$$

This completes the proof. \(\square \)

Proof of Theorem 4

Following Zheng et al. (2007), we firstly suppose \(\varvec{\tau }\) is known. Let

$$\begin{aligned} L_{n}^{(1)}(\varvec{\tau },\varvec{\eta })&=\sum _{t=1}^{n}\mathrm {V_{\varvec{\tau }}^{-1}}(X_t| X_{t-1})\dot{\phi }X_{t-1}\big (X_t-\phi X_{t-1}-\mu _{\epsilon }\big ),~L_{0}^{(1)}(\varvec{\tau },\varvec{\eta })=0,\\ L_{n}^{(2)}(\varvec{\tau },\varvec{\eta })&=\sum _{t=1}^{n}\mathrm {V_{\varvec{\tau }}^{-1}}(X_t|X_{t-1})\big (X_t-\phi X_{t-1}-\mu _{\epsilon }\big ),~L_{0}^{(2)}(\varvec{\tau },\varvec{\eta })=0. \end{aligned}$$

Similar to Theorem 2, we have

$$\begin{aligned} n^{-1/2}\big (L_{n}^{(1)}(\varvec{\tau },\varvec{\eta }),L_{n}^{(2)}(\varvec{\tau },\varvec{\eta })\big )^{'} \mathop {\longrightarrow }\limits ^{d} N(\mathbf {0},\varvec{V_{MQL}}). \end{aligned}$$

Now, we replace \(V_{\varvec{\tau }}^{-2}(X_t|X_{t-1})\) by \(V_{\varvec{\widehat{\tau }}}^{-2}(X_t|X_{t-1})\), where \(\varvec{\widehat{\tau }}\) is a consistent estimator of \(\varvec{\tau }\). Then we want

$$\begin{aligned} n^{-1/2}\big (L_{n}^{(1)}(\varvec{\widehat{\tau }},\varvec{\eta }),L_{n}^{(2)}(\varvec{\widehat{\tau }},\varvec{\eta })\big )^{'} \mathop {\longrightarrow }\limits ^{d} N(\mathbf {0},\varvec{V_{MQL}}). \end{aligned}$$

For this we need to prove that \( \frac{1}{\sqrt{n}}L_{n}^{(i)}(\varvec{\widehat{\tau }},\varvec{\eta })- \frac{1}{\sqrt{n}}L_{n}^{(i)}(\varvec{\tau },\varvec{\eta })\mathop {\longrightarrow }\limits ^{P}0,~i=1,2 \) [its proof is omitted here, since the argument is the same as in Zheng et al. (2007)]. Following the proof of Theorem 2, by Taylor’s expansion and some algebra, we have

$$\begin{aligned} \sqrt{n}(\widehat{\varvec{\eta }}_{MQL}-\varvec{\eta })\mathop {\longrightarrow }\limits ^{d}N(\mathbf {0},\varvec{H_{MQL}}^{-1}\varvec{V_{MQL}}\varvec{H_{MQL}}^{-1}). \end{aligned}$$

This completes the proof. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kang, Y., Wang, D., Yang, K. et al. A new thinning-based INAR(1) process for underdispersed or overdispersed counts. J. Korean Stat. Soc. 49, 324–349 (2020). https://doi.org/10.1007/s42952-019-00010-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s42952-019-00010-2

Keywords

Mathematics Subject Classification

Navigation