Skip to main content
Log in

Modeling overdispersed or underdispersed count data with generalized Poisson integer-valued autoregressive processes

  • Published:
Metrika Aims and scope Submit manuscript

Abstract

To accurately and flexibly capture the dispersion features of time series of counts, we introduce the generalized Poisson thinning operation and further define some new integer-valued autoregressive processes. Basic probabilistic and statistical properties of the models are discussed. Conditional least squares and maximum quasi likelihood estimators are investigated via the moment targeting estimation methods for the innovation free case. Also, the asymptotic properties of the estimators are obtained. Conditional maximum likelihood estimation for the parametric cases are also discussed. Finally, some numerical results of the estimates and two real data examples are presented.

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
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Notes

  1. Here the innovation free means the process is defined with a free type of innovation, i.e., the process is defined with no assumption of the distribution type of the innovations.

References

  • Al-Osh MA, Alzaid AA (1987) First-order integer-valued autoregressive (INAR(1)) process. J Time Ser Anal 8:261–275

    Article  MathSciNet  MATH  Google Scholar 

  • Alzahrani N, Neal P, Spencer SEF, McKinley TJ, Touloupou P (2018) Model selection for time series of count data. Comput Stat Data Anal 122:33–44

    Article  MathSciNet  MATH  Google Scholar 

  • Billingsley P (1961) Statistical inference for Markov processes. The University of Chicago Press, Chicago

    MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  • Chen CWS, Lee S (2017) Bayesian causality test for integer-valued time series models with applications to climate and crime data. J R Stat Soc Ser C 66:797–814

    Article  MathSciNet  Google Scholar 

  • Consul PC (1989) Generalized Poisson distributions: properties and applications. Marcel Dekker, New York

    MATH  Google Scholar 

  • Consul PC, Famoye F (1992) Generalized Poisson regression model. Commun Stat Theory Methods 21:89–109

    Article  MATH  Google Scholar 

  • Consul PC, Famoye F (2006) Lagrangian probability distributions. Birkhäuser, Boston

    MATH  Google Scholar 

  • Consul PC, Jain GC (1973) A generalization of the Poisson distribution. Technometrics 15:791–799

    Article  MathSciNet  MATH  Google Scholar 

  • Cox DR (1981) Statistical analysis of time series: some recent developments [with discussion and reply]. Scand J Stat 8:93–115

    MATH  Google Scholar 

  • Freeland RK (1998) Statistical analysis of discrete time series with applications to the analysis of workers compensation claims data. University of British Columbia, Canada Ph.D. Thesis

    Google Scholar 

  • Fukasawa T, Basawa IV (2002) Estimation for a class of generalized state-space time series models. Stat Probab Lett 60:459–473

    Article  MathSciNet  MATH  Google Scholar 

  • Hall P, Heyde CC (1980) Martingale limit theory and its applation. Academic Press, New York

    MATH  Google Scholar 

  • Klimko LA, Nelson PI (1978) On conditional least squares estimation for stochastic processes. Ann Stat 6:629–642

    Article  MathSciNet  MATH  Google Scholar 

  • Latour A (1998) Existence and stochastic structure of a non-negative integer-valued autoregressive process. J Time Ser Anal 19:439–455

    Article  MathSciNet  MATH  Google Scholar 

  • Li Q, Zhu F (2018) Mean targeting estimator for the integer-valued GARCH(1,1) model. Stat Pap. https://doi.org/10.1007/s00362-017-0958-9

    Google Scholar 

  • Li H, Yang K, Wang D (2017) Quasi-likelihood inference for self-exciting threshold integer-valued autoregressive processes. Comput Stat 32:1597–1620

    Article  MathSciNet  MATH  Google Scholar 

  • Li H, Yang K, Zhao S, Wang D (2018) First-order random coefficients integer-valued threshold autoregressive processes. AStA Adv Stat Anal 102:305–331

    Article  MathSciNet  MATH  Google Scholar 

  • McKenzie E (1985) Some simple models for discrete variate time series. JAWAR J Am Water Resour Assoc 21:645–650

    Article  Google Scholar 

  • Pedeli X, Davison AC, Fokianos K (2015) Likelihood estimation for the INAR(\(p\)) model by saddlepoint approximation. J Am Stat Assoc 110:1229–1238

    Article  MathSciNet  MATH  Google Scholar 

  • Ristić MM, Bakouch HS, Nastić AS (2009) A new geometric first-order integer-valued autoregressive (NGINAR(1)) process. J Stat Plan Inference 139:2218–2226

    Article  MathSciNet  MATH  Google Scholar 

  • Ristić MM, Nastić AS, Bakouch HS (2012) Estimation in an integer-valued autoregressive process with negative binomial marginals (NBINAR(1)). Commun Stat Theory Methods 41:606–618

    Article  MathSciNet  MATH  Google Scholar 

  • Schweer S (2016) A goodness-of-fit test for integer-valued autoregressive processes. J Time Ser Anal 37:77–98

    Article  MathSciNet  MATH  Google Scholar 

  • Schweer S, Weiß CH (2014) Compound Poisson INAR(1) processes: stochastic properties and testing for overdispersion. Comput Stat Data Anal 77:267–284

    Article  MathSciNet  MATH  Google Scholar 

  • Serfling RJ (1980) Approximation theorems of mathematical statistics. Wiley, New York

    Book  MATH  Google Scholar 

  • Shi H, Wang D (2014) An approximation model of the collective risk model with INAR(1) claim process. Commun Stat Theory Methods 43:5305–5317

    Article  MathSciNet  MATH  Google Scholar 

  • Steutel F, van Harn K (1979) Discrete analogues of self-decomposability and stability. Ann Probab 7:893–899

    Article  MathSciNet  MATH  Google Scholar 

  • Wedderburn RWM (1974) Quasi-likelihood functions, generalized linear models, and the Gauss–Newton method. Biometrika 61:439–447

    MathSciNet  MATH  Google Scholar 

  • Weiß CH (2008) Thinning operations for modeling time series of counts—a survey. Adv Stat Anal 92:319–343

    Article  MathSciNet  Google Scholar 

  • Yang K, Wang D, Li H (2018a) Threshold autoregression analysis for finite-range time series of counts with an application on measles data. J Stat Comput Simul 88:597–614

    Article  MathSciNet  Google Scholar 

  • Yang K, Li H, Wang D (2018b) Estimation of parameters in the self-exciting threshold autoregressive processes for nonlinear time series of counts. Appl Math Model 54:226–247

    Article  MathSciNet  Google Scholar 

  • Yang K, Wang D, Jia B, Li H (2018c) An integer-valued threshold autoregressive process based on negative binomial thinning. Stat Pap 59:1131–1160

    Article  MathSciNet  MATH  Google Scholar 

  • Zhu F (2012) Modeling overdispersed or underdispersed count data with generalized Poisson integer-valued GARCH models. J Math Anal Appl 389:58–71

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We gratefully acknowledge the anonymous reviewers for their careful work and thoughtful suggestions that have helped improve this article substantially. We also acknowledge the financial supported by National Natural Science Foundation of China (No. 11871028, 11731015, 11571051, 11501241), National Social Science Foundation of China (16BTJ020), Natural Science Foundation of Jilin Province (No. 20180101216JC, 20170101057JC, 20150520053JH), 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.

Ethics declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Additional information

Publisher's Note

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

Appendix

Appendix

The proofs of Proposition 1

Since (i)–(v) are easy to verify, we only prove (vi). By the law of total covariance, we have

$$\begin{aligned} \mathrm{Cov}(X_{t},X_{t-h})&=\mathrm{Cov}(E(X_t|X_{t-1},\ldots ),E(X_{t-h}|X_{t-1},\ldots ))+0\\&=\mathrm{Cov}[\phi (\lambda ,\kappa ) X_{t-1}+\mu _{\varepsilon },X_{t-h}]\\&=\phi (\lambda ,\kappa )\cdot \mathrm{Cov}(X_{t-1},X_{t-h})\\&=\cdots \\&=\phi (\lambda ,\kappa )^h\cdot \mathrm{Var}(X_{t-h}), \end{aligned}$$

Thus, the autocorrelation function \(\rho (h)=\mathrm{Corr}(X_{t},X_{t-h})=\phi (\lambda ,\kappa )^h.\)\(\square \)

The proofs of Theorems 1 and 2

These two theorems can be easily proved by verifying the regularity conditions of Theorems 3.1 and 3.2 in Klimko and Nelson (1978), we omit the details here. \(\square \)

The proof of Theorem 3

This is an application of the \(\delta \)-method; For completeness and reader’s convenience, we refer to Theorem A on p.122 of Serfling (1980) for a proof. \(\square \)

The proof of Theorem 4

First, we suppose \(\varvec{\theta }_2\) is known. Let \(\mathcal {F}_t=\sigma (X_0,X_1,\ldots ,X_t)\) be the \(\sigma \)-field generated by \(\{X_0,X_1,\ldots ,X_t\}\). Denote

$$\begin{aligned} S_n^{(1)}(\varvec{\theta }_{2},\varvec{\theta }_1)=\sum _{t=1}^n V_{\varvec{\theta }_2}^{-1}(X_t|X_{t-1})(X_t-\phi X_{t-1}-\mu _{\varepsilon }) X_{t-1}. \end{aligned}$$

By direct calculation, we can see that

$$\begin{aligned}&E[V_{\varvec{\theta }_2}^{-1}(X_t|X_{t-1})(X_t-\phi X_{t-1}-\mu _{\varepsilon }) X_{t-1}|\mathcal {F}_{t-1}]\\&\quad =V_{\varvec{\theta }_2}^{-1}(X_t|X_{t-1})X_{t-1}E[(X_t-\phi X_{t-1}-\mu _{\varepsilon })|\mathcal {F}_{t-1}]\\&\quad =0, \end{aligned}$$

and

$$\begin{aligned}&E[S_t^{(1)}(\varvec{\theta }_2,\varvec{\theta }_1)|\mathcal {F}_{t-1}]\\&\quad =E[(V_{\varvec{\theta }_2}^{-1}(X_t|X_{t-1})(X_t-\phi X_{t-1}-\mu _{\varepsilon })X_{t-1} +S_{t-1}^{(1)}(\varvec{\theta }_2,\varvec{\theta }_1))|\mathcal {F}_{t-1}]\\&\quad =S_{t-1}^{(1)}(\varvec{\theta }_2,\varvec{\theta }_1). \end{aligned}$$

Thus, \(\{S_t^{(1)}(\varvec{\theta }_2,\varvec{\theta }_1),\mathcal {F}_{t},~t\ge 0\}\) is a martingale. By (C1) and Theorem 1.1 in Billingsley (1961),

$$\begin{aligned}&\frac{1}{n}\sum _{t=1}^n V_{\varvec{\theta }_{2,0}}^{-2}(X_t|X_{t-1})(X_t-\phi X_{t-1}-\mu _{\varepsilon })^2X_{t-1}^2\\&\quad \overset{a.s.}{\longrightarrow }E\left( V_{\varvec{\theta }_{2,0}}^{-2}(X_1|X_{0})(X_1-\phi X_{0}-\mu _{\varepsilon })^2X_0^2\right) \\&\quad =E\left( E[V_{\varvec{\theta }_{2,0}}^{-2}(X_1|X_{0})(X_1-\phi X_{0}-\mu _{\varepsilon })^2X_0^2|X_0]\right) \\&\quad =E[V_{\varvec{\theta }_{2,0}}^{-1}(X_1|X_0)X_0^2]. \end{aligned}$$

Hence, by Corollary 3.2 in Hall and Heyde (1980) and the central limit theorem of martingale, we have,

$$\begin{aligned} \frac{1}{\sqrt{n}}S_n^{(1)}(\varvec{\theta }_{2,0},\varvec{\theta }_{1}) \overset{L}{\longrightarrow }N(0,E[V_{\varvec{\theta }_{2,0}}^{-1}(X_1|X_0)X_0^2]). \end{aligned}$$

Similarly, denote \(S_n^{(2)}(\varvec{\theta }_{2},\varvec{\theta }_1)=\sum _{t=1}^n V_{\varvec{\theta }_2}^{-1}(X_t|X_{t-1})(X_t-\phi X_{t-1}-\mu _{\varepsilon })\). Then, we can verify that \(\{S_t^{(2)}(\varvec{\theta }_2,\varvec{\theta }_1),\mathcal {F}_{t},~t\ge 0\}\) is also martingale. Similar to the previous discussion, we have

$$\begin{aligned} \frac{1}{\sqrt{n}}S_n^{(2)}(\varvec{\theta }_{2,0},\varvec{\theta }_1) \overset{L}{\longrightarrow }N(0,E(V_{\varvec{\theta }_{2,0}}^{-1}(X_1|X_{0}))). \end{aligned}$$

By Cramer-Wold device, for any \(\varvec{c}^\textsf {T}=(c_1,c_2) \in \mathbb {R}^3\) and \((c_1,c_2)\ne (0,0)\), we have

$$\begin{aligned} \frac{\varvec{c}^\textsf {T}}{\sqrt{n}}\left( \begin{array}{c} S_n^{(1)}(\varvec{\theta }_{2,0},\varvec{\theta }_1)\\ S_n^{(2)}(\varvec{\theta }_{2,0},\varvec{\theta }_1) \end{array}\right)&=\frac{1}{\sqrt{n}}\sum _{i=1}^2 c_iS_n^{(i)}(\varvec{\theta }_{2,0},\varvec{\theta }_1)\\&=\frac{1}{\sqrt{n}}\sum _{t=1}^n V_{\varvec{\theta }_{2,0}}^{-1}(X_t|X_{t-1})(X_t-\phi X_{t-1}-\mu _{\varepsilon })(c_1 X_{t-1}+c_2)\\&\overset{L}{\longrightarrow } N(0,E[V_{\varvec{\theta }_{2,0}}^{-1}(X_1|X_0)(c_1 X_{0}+c_2)^2]), \end{aligned}$$

implying

$$\begin{aligned} \frac{1}{\sqrt{n}}\left( \begin{array}{c} S_n^{(1)}(\varvec{\theta }_{2,0},\varvec{\theta }_1)\\ S_n^{(2)}(\varvec{\theta }_{2,0},\varvec{\theta }_1) \end{array}\right)&\overset{L}{\longrightarrow } N\left( \varvec{0},\varvec{T}\right) . \end{aligned}$$

Now, we replace \(V_{\varvec{\theta }_2}(X_t|X_{t-1})\) with \(V_{\hat{\varvec{\theta }}_2}(X_t|X_{t-1})\), where \(\hat{\varvec{\theta }}_2\) is a consistent estimator of \({\varvec{\theta }}_2\). Then we want to prove that

$$\begin{aligned} \frac{1}{\sqrt{n}}\left( \begin{array}{c} S_n^{(1)}(\hat{\varvec{\theta }}_2,\varvec{\theta }_1)\\ S_n^{(2)}(\hat{\varvec{\theta }}_2,\varvec{\theta }_1) \end{array}\right)&\overset{L}{\longrightarrow } N\left( \varvec{0},\varvec{T}\right) . \end{aligned}$$
(A.2)

To prove (A.2), we need to check that

$$\begin{aligned} \frac{1}{\sqrt{n}}S_n^{(i)}(\hat{\varvec{\theta }_2}, \varvec{\theta }_1)-\frac{1}{\sqrt{n}}S_n^{(i)}({\varvec{\theta }_{2,0}}, \varvec{\theta }_1)\overset{P}{\longrightarrow }0,~i=1,2. \end{aligned}$$

Let \(R_n(\varvec{\theta }_2)=(1/\sqrt{n})S_n^{(1)}(\varvec{\theta }_2,\varvec{\theta }_1)\). Then for any \(\varepsilon >0\) and \(\delta >0\), we have

$$\begin{aligned} P(|R_n(\hat{\varvec{\theta }}_2)-R_n(\varvec{\theta }_{2,0})|>\varepsilon )&\le P(|\hat{\psi }-\psi _0|>\delta )+P(|\hat{\sigma }_{\varepsilon }^2-\sigma _{0}^2|>\delta )\\&\quad +P(\sup _D|R_n(\widetilde{\varvec{\theta }}_2)-R_n(\varvec{\theta }_{2,0})|>\varepsilon ), \end{aligned}$$

where \(\widetilde{\varvec{\theta }}_2=(\widetilde{\psi },\widetilde{\sigma }_{\varepsilon }^2)^\textsf {T}, D:=\{|\widetilde{\psi }-\psi _0|<\delta ,|\widetilde{\sigma }_{\varepsilon }^{2}-\sigma _0^2|<\delta \}.\) If \(\hat{\varvec{\theta }}_2\) is a consistent estimator of \({\varvec{\theta }}_2\), then we just need to prove that

$$\begin{aligned} P\left( \sup _D|R_n(\widetilde{\varvec{\theta }}_2)-R_n(\varvec{\theta }_{2,0} )|>\varepsilon \right) \overset{P}{\longrightarrow }0. \end{aligned}$$

By Markov inequality,

$$\begin{aligned}&P\left( \sup _D|R_n(\widetilde{\varvec{\theta }}_2)-R_n(\varvec{\theta }_{2,0}) |>\varepsilon \right) \\&\quad \le \frac{1}{\varepsilon ^2}E\left( \sup _D(R_n(\widetilde{\varvec{\theta }}_2) -R_n(\varvec{\theta }_{2,0}))^2\right) \\&\quad =\frac{1}{\varepsilon ^2}E\left( \sup _D\frac{1}{n}\sum _{t=1}^n (V_{\widetilde{\varvec{\theta }}_2}^{-1}(X_t|X_{t-1})-V_{\varvec{\theta }_{2,0}}^{-1} (X_t|X_{t-1}))^2(X_t-\phi X_{t-1}-\mu _{\varepsilon })^2X_{t-1}^2\right) \\&\quad =\frac{1}{\varepsilon ^2}E\left( \sup _D(V_{\widetilde{\varvec{\theta }}_2}^{-1} (X_1|X_{0})-V_{\varvec{\theta }_{2,0}}^{-1}(X_1|X_{0}))^2 (X_1-\phi X_{0}-\mu _{\varepsilon })^2X_{0}^2\right) \\&\quad =\frac{1}{\varepsilon ^2}E\left( \sup _D\frac{((\widetilde{\psi }- \psi _{0})X_0+(\widetilde{\sigma }_{\varepsilon }^2-\sigma _0^2))^2}{V_{\widetilde{\varvec{\theta }}_2}^{2}(X_1|X_{0})V_{\varvec{\theta }_{2,0}}^{2} (X_1|X_{0})}\left( X_1-\phi X_{0}-\mu _{\varepsilon }\right) ^2X_{0}^2\right) \\&\quad =\frac{1}{\varepsilon ^2}E\left( \sup _D\frac{((\widetilde{\psi }- \psi _0)X_0+(\widetilde{\sigma }_{\varepsilon }^2-\sigma _0^2))^2}{V_{\widetilde{\varvec{\theta }}_2}^{2}(X_1|X_{0})V_{\varvec{\theta }_{2,0}} (X_1|X_{0})}X_{0}^2\right) \\&\quad \le \frac{1}{\varepsilon ^2}\sup _D \left\{ [(\widetilde{\psi }- \psi _0)^2 m_1 +(\widetilde{\sigma }_{\varepsilon }^2-\sigma _0^2)^2 m_2 +2|\widetilde{\psi }-\psi _0||\widetilde{\sigma }_{\varepsilon }^2- \sigma _0^2| m_3]X_{0}^2\right\} \\&\quad \le \frac{C\delta ^2}{\varepsilon ^2}, \end{aligned}$$

where \(m_i~(i=1,2,3)\) denote some finite moments of process \(\{X_t\}\), C is a positive constant. Similar argument can be used for \(1/\sqrt{n}S_n^{(2)}(\varvec{\theta }_2,\varvec{\theta }_1)\). For any fixed \(\varepsilon >0\), letting \(\delta \rightarrow 0\), we get our assertion which in turn establishes (A.2).

Finally, by the ergodic theorem, we have \(\frac{1}{n}\varvec{P}\overset{P}{\longrightarrow }\varvec{T}\). After some algebra, we have,

$$\begin{aligned} (\hat{\varvec{\theta }}_{1,MQL}-\varvec{\theta }_{1,0})=\varvec{P}^{-1} \left( \begin{array}{c} S_n^{(1)}(\hat{\varvec{\theta }_2},\varvec{\theta }_{1,0})\\ S_n^{(2)}(\hat{\varvec{\theta }_2},\varvec{\theta }_{1,0}) \end{array}\right) . \end{aligned}$$

Therefore,

$$\begin{aligned} \sqrt{n}(\hat{\varvec{\theta }}_{1,MQL}-\varvec{\theta }_{1,0})=\left( \frac{1}{n}\varvec{P}\right) ^{-1} \frac{1}{\sqrt{n}} \left( \begin{array}{c} S_n^{(1)}(\hat{\varvec{\theta }_2},\varvec{\theta }_{1,0})\\ S_n^{(2)}(\hat{\varvec{\theta }_2},\varvec{\theta }_{1,0}) \end{array}\right) \overset{L}{\longrightarrow }N(\varvec{0},\varvec{T}^{-1}). \end{aligned}$$

The proof is complete. \(\square \)

The proof of Theorem 5

Since this proof is similar to the proof of Theorem 4, we omit here. \(\square \)

The proof of Theorem 6

Using the same statements as in the proof of Theorem 3, we can get the conclusion. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, K., Kang, Y., Wang, D. et al. Modeling overdispersed or underdispersed count data with generalized Poisson integer-valued autoregressive processes. Metrika 82, 863–889 (2019). https://doi.org/10.1007/s00184-019-00714-9

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00184-019-00714-9

Keywords

Navigation