Skip to main content
Log in

A Poisson INAR(1) model with serially dependent innovations

  • Published:
Metrika Aims and scope Submit manuscript

Abstract

Motivated by a certain type of infinite-patch metapopulation model, we propose an extension to the popular Poisson INAR(1) model, where the innovations are assumed to be serially dependent in such a way that their mean is increased if the current population is large. We shall recognize that this new model forms a bridge between the Poisson INAR(1) model and the INARCH(1) model. We analyze the stochastic properties of the observations and innovations from an extended Poisson INAR(1) process, and we consider the problem of model identification and parameter estimation. A real-data example about iceberg counts shows how to benefit from 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

Notes

  1. Note that the extended Poisson INAR(1) model differs from the CINAR(1) model of Triebsch (2008), where the innovations \(\epsilon _t\) stem from an INARCH(1) process on their own. While the first is observation driven according to the classification by Cox (1981), see Definition 1, the latter is parameter driven.

  2. Note that reparametrization (20) is also advantageous for numerical optimization, since it allows to work with box constraints.

References

  • Al-Osh MA, Alzaid AA (1987) First-order integer-valued autoregressive (INAR(1)) process. J Time Ser Anal 8(3):261–275

    Article  MathSciNet  Google Scholar 

  • Alzaid AA, Al-Osh MA (1988) First-order integer-valued autoregressive process: distributional and regression properties. Stat Neerl 42(1):53–61

    Article  MathSciNet  MATH  Google Scholar 

  • Bradley RC (2005) Basic properties of strong mixing conditions. A survey and some open questions. Probab Surv 2:107–144

    Article  MathSciNet  Google Scholar 

  • Bu R, McCabe B, Hadri K (2008) Maximum likelihood estimation of higher-order integer-valued autoregressive processes. J Time Ser Anal 29(6):973–994

    Article  MathSciNet  MATH  Google Scholar 

  • Buckley F, Pollett P (2010) Limit theorems for discrete-time metapopulation models. Probab Surv 7:53–83

    Article  MathSciNet  MATH  Google Scholar 

  • Cox DR (1981) Statistical analysis of time series: some recent developments. Scand J Stat 8:93–115

    MATH  Google Scholar 

  • Feller W (1968) An introduction to probability theory and its applications, vol I, 3rd edn. Wiley, New Jersey

    MATH  Google Scholar 

  • Freeland RK, McCabe BPM (2004) Analysis of low count time series data by Poisson autoregression. J Time Ser Anal 25(5):701–722

    Article  MathSciNet  MATH  Google Scholar 

  • Frey S, Sandas P (2009) The impact of iceberg orders in limit order books. CFR working paper 09–06

  • Grunwald GK, Hyndman RJ, Tedesco L (1997) A unified view of linear AR(1) models. Research report, Department of Statistics, University of Melbourne

  • Grunwald GK, Hyndman RJ, Tedesco L, Tweedie RL (2000) Non-Gaussian conditional linear AR(1) models. Aust NZ J Stat 42(4):479–495

    Article  MathSciNet  MATH  Google Scholar 

  • Heathcote CR (1966) Corrections and comments of the paper “A branching process allowing immigration”. J R Stat Soc Ser B 28(1):213–217

    MathSciNet  Google Scholar 

  • Heyde CC, Seneta E (1972) Estimation theory for growth and immigration rates in a multiplicative process. J Appl Probab 9(2):235–258

    Article  MathSciNet  MATH  Google Scholar 

  • Ibragimov I (1962) Some limit theorems for stationary processes. Theory Probab Appl 7(4):349–382

    Article  Google Scholar 

  • Johnson NL, Kotz S, Balakrishnan N (1997) Discrete multivariate distributions. Wiley, Hoboken

    MATH  Google Scholar 

  • Johnson NL, Kemp AW, Kotz S (2005) Univariate discrete distributions, 3rd edn. Wiley, Hoboken

    Book  MATH  Google Scholar 

  • Jung RC, Tremayne AR (2003) Testing for serial dependence in time series models of counts. J Time Ser Anal 24(1):65–84

    Article  MathSciNet  MATH  Google Scholar 

  • Jung RC, Tremayne AR (2011a) Useful models for time series of counts or simply wrong ones? AStA Adv Stat Anal 95(1):59–91

    Article  MathSciNet  Google Scholar 

  • Jung RC, Tremayne AR (2011b) Convolution-closed models for count time series with applications. J Time Ser Anal 32:268–280

    Article  MathSciNet  MATH  Google Scholar 

  • Katz RW (1981) On some criteria for estimating the order of a Markov chain. Technometrics 23(3):243–249

    Article  MathSciNet  MATH  Google Scholar 

  • Klimko LA, Nelson PI (1978) On conditional least squares estimation for stochastic processes. Ann Stat 6(3):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(4):439–455

    Article  MathSciNet  MATH  Google Scholar 

  • McKenzie E (1985) Some simple models for discrete variate time series. Water Resour Bull 21(4):645–650

    Article  MathSciNet  MATH  Google Scholar 

  • Monteiro M, Scotto MG, Pereira I (2012) Integer-valued self-exciting threshold autoregressive processes. Commun Stat Theory Methods 41(15):2717–2737

    Article  MathSciNet  MATH  Google Scholar 

  • Nummelin E, Tweedie RL (1978) Geometric ergodicity and \(R\)-positivity for general Markov chains. Ann Probab 6(3):404–420

    Article  MathSciNet  MATH  Google Scholar 

  • Pakes AG (1971) Branching processes with immigration. J Appl Probab 8(1):32–42

    Article  MathSciNet  MATH  Google Scholar 

  • Seneta E (1968) The stationary distribution of a branching process allowing immigration: a remark on the critical case. J R Stat Soc Ser B 30(1):176–179

    MathSciNet  MATH  Google Scholar 

  • Steutel FW, van Harn K (1979) Discrete analogues of self-decomposability and stability. Ann Probab 7(5):893–899

    Article  MathSciNet  MATH  Google Scholar 

  • Sun J, McCabe BPM (2013) Score statistics for testing serial dependence in count data. J Time Ser Anal 34(3):315–329

    Article  MathSciNet  MATH  Google Scholar 

  • Triebsch LK (2008) New Integer-valued Autoregressive and Regression Models with State-dependent Parameters. Doctoral dissertation, TU Kaiserslautern, Verlag Dr. Hut, Munich

  • Tsay RS (1992) Model checking via parametric bootstraps in time series analysis. J R Stat Soc Ser C 41(1):1–15

    MATH  Google Scholar 

  • Wei CZ, Winnicki J (1990) Estimation of the means in the branching process with immigration. Ann Stat 18(4):1757–1773

    Article  MathSciNet  MATH  Google Scholar 

  • Weiß CH (2008) Thinning operations for modelling time series of counts—a survey. AStA Adv Stat Anal 92(3):319–341

    Article  MathSciNet  Google Scholar 

  • Weiß CH (2010a) The INARCH(1) model for overdispersed time series of counts. Commun Stat Simul Comput 39(6):1269–1291

    Article  Google Scholar 

  • Weiß CH (2010b) INARCH(1) processes: higher-order moments and jumps. Stat Probab Lett 80(23–24):1771–1780

    Article  MATH  Google Scholar 

  • Weiß CH (2012) Process capability analysis for serially dependent processes of Poisson counts. J Stat Comput Simul 82(3):383–404

    Article  MathSciNet  MATH  Google Scholar 

  • Weiß CH, Pollett PK (2014) Binomial autoregressive processes with density dependent thinning. J Time Ser Anal 35(2):115–132

    Article  MATH  Google Scholar 

  • Winnicki J (1991) Estimation of the variances in the branching process with immigration. Probab Theory Relat Fields 88(1):77–106

    Article  MathSciNet  MATH  Google Scholar 

  • Zhu F, Wang D (2011) Estimation and testing for a Poisson autoregressive model. Metrika 73:211–230

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The author thanks the two referees for highly useful comments on an earlier draft of this article. The iceberg order data of Section 4.2 were kindly made available to the author by the Deutsche Börse. Prof. Dr. Joachim Grammig, University of Tübingen, is to be thanked for processing of it to make it amenable to data analysis. I am also very grateful to Prof. Dr. Robert Jung, University of Hohenheim, for his kind support to get access to the data, and for valuable comments on an earlier draft of this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christian H. Weiß.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (xls 23 KB)

Proofs

Proofs

1.1 Proof of Proposition 2

Since \((X_t)_{\mathbb N_0}\) is a homogeneous Markov chain with strictly positive 1-step transition probabilities, \((X_t)_{\mathbb N_0}\) is also irreducible and aperiodic. Furthermore, it is distributed as a BPI, so we can apply the theorem in Heathcote (1966). It remains to check two conditions:

  • The mean related to the individual offspring, i.e., the mean of \(Y\) according to our notations in Sect. 2, must be smaller than 1 (subcritical case). From Proposition 1, we obtain that \(E[Y]=\rho +a\), so \(E[Y]<1\) iff the condition \(\rho +a<1\) holds.

  • The immigration in (4) is described by the i.i.d. random variables \(\nu _t\sim {\text {Po}}(b)\). Now consider the quantity \(\sum _{n=1}^{\infty }\ {\textstyle \frac{1}{n}}\cdot P(\nu _t\ge n)\), i.e., the expectation \(E[H(\nu _t)]\), where \(H(n) := \sum _{k=1}^n \frac{1}{k}\) for \(n\in \mathbb N\) and \(H(0):=0\) denote the harmonic numbers. Since \(H(\nu _t)\le \nu _t\), and since \(E[\nu _t] = b<\infty \), also \(E[H(\nu _t)]<\infty \) holds.

So the existence of a stationary marginal distribution follows, and together with the irreducibility and aperiodicity, it follows that \((X_t)_{\mathbb N_0}\) is an ergodic Markov chain (Feller 1968, p. 394). Note that this stationarity result could also be derived by applying Proposition 3.1 in Latour (1998), because binomial-Poisson thinning is an instance of the generalized thinning concept considered by Latour (1998).

With similar arguments as for Heathcote (1966), we now apply Theorem 1 in Pakes (1971). \((X_t)_{\mathbb N_0}\) is not only distributed as a subcritical BPI, the corresponding offspring also has a finite variance, i.e., \(V[Y]<\infty \). Hence, \((X_t)_{\mathbb N_0}\) satisfies all conditions being required in Sect. 2 of Pakes (1971), so Theorem 1 in Pakes (1971) implies geometric ergodicity. From Theorem 1 of Nummelin and Tweedie (1978) and Theorem 3.7 in Bradley (2005), we conclude that a stationary extended Poisson INAR(1) process is also \(\beta \)-mixing (and thus \(\alpha \)-mixing) with exponentially decreasing weights.

1.2 Proofs of Sect. 3.1

Let \(f(x) := a\cdot x+b\). Formula (5) about conditional mean and variance follows from

$$\begin{aligned} E[X_t\ |\ X_{t-1},\ldots ]&= \rho \cdot X_{t-1}+f(X_{t-1})=(\rho +a)\cdot X_{t-1} + b,\\ V[X_t\ |\ X_{t-1},\ldots ]&= \rho (1-\rho )\cdot X_{t-1}+f(X_{t-1})=\big (\rho (1-\rho )+a\big )\cdot X_{t-1} + b, \end{aligned}$$

respectively. For computing the variance, we used the conditional independence of \(\rho \circ X_{t-1}\) and \(\epsilon _t\), given \(X_{t-1}\). From formula (5) and the stationarity assumption, we obtain the marginal mean and variance as given in formula (6) by solving

$$\begin{aligned} \mu _X=E[X_t]&= E\big [E[X_t|X_{t-1},\ldots ]\big ]=(\rho +a)\cdot \mu _X+ b,\\ \sigma _X^2=V[X_t]&= E\big [V[X_t|X_{t-1},\ldots ]\big ]+V\big [E[X_t|X_{t-1},\ldots ]\big ]\\&= \big (\rho (1-\rho )+a\big )\cdot \mu _X + b+(\rho +a)^2\cdot \sigma _X^2\\&= (1-\rho ^2)\cdot \mu _X+(\rho +a)^2\cdot \sigma _X^2. \end{aligned}$$

The expression (7) for the pgf is shown analogously:

$$\begin{aligned} E[z^{X_t}]&= E\big [E[z^{\rho \circ X_{t-1}}|X_{t-1}]\cdot E[z^{\epsilon _t}\ |\ X_{t-1}]\big ]\\&= E\big [(1-\rho +\rho z)^{X_{t-1}}\cdot \exp {\big ((a\cdot X_{t-1}+b)(z-1)\big )}\big ]\\&= e^{b(z-1)}\cdot E\big [(1-\rho +\rho z)^{X_{t-1}}\cdot \big (e^{a(z-1)}\big )^{X_{t-1}}\big ]. \end{aligned}$$

Recursions (8) and (9) for higher-order moments follow from

$$\begin{aligned} E\left[ \left. X_t^n\right| X_{t-1},\ldots \right]&= E[(\rho \circ X_{t-1} + \epsilon _t)^n|X_{t-1}]\\&= \sum _{j=0}^n\ {n \atopwithdelims ()j}\cdot E[(\rho \circ X_{t-1})^j|X_{t-1}] \cdot E\left[ \left. \epsilon _t^{n-j}\right| X_{t-1}\right] . \end{aligned}$$

To compute the second and third moments of \(X_t\), we first require the first few moments of the binomial and the Poisson distribution:

$$\begin{aligned} \mu _{1,{\text {B}}(n,p)}&= np,\quad \mu _{2,{\text {B}}(n,p)}=n^2 p^2+n p(1-p),\\ \mu _{3,{\text {B}}(n,p)}&= n^3 p^3+3n^2p^2 (1-p)+np(1-p)(1-2p),\\ \mu _{1,{\text {Po}}(\lambda )}&= \lambda ,\quad \mu _{2,{\text {Po}}(\lambda )}=\lambda (1+\lambda ),\quad \mu _{3,{\text {Po}}(\lambda )}=\lambda (1+3\lambda +\lambda ^2). \end{aligned}$$

Now we apply formula (9) to obtain

$$\begin{aligned} \mu _{2,X}&:= E\left[ X_t^2\right] \\&= E\big [\mu _{2,{\text {Po}}(a\cdot X_{t-1}+b)}\big ]\!+\!2\cdot E\big [\mu _{1,{\text {B}}(X_{t-1},\rho )}\cdot \mu _{1,{\text {Po}}(a\cdot X_{t-1}+b)}\big ]+E\big [\mu _{2,{\text {B}}(X_{t-1},\rho )}\big ]\\&= E\big [(a\cdot X_{t-1}+b)(1+a\cdot X_{t-1}+b)\big ]\\&\quad +\,2\cdot E\big [\rho X_{t-1}\cdot (a\cdot X_{t-1}+b)\big ]+ E\left[ X_{t-1}^2 \rho ^2+X_{t-1} \rho (1-\rho )\right] \\&= E\left[ b(1+b)+a(1+2b) X_{t-1}+a^2 X_{t-1}^2\right] \\&\quad +\,2\cdot E\left[ \rho a\cdot X_{t-1}^2+\rho b X_{t-1}\right] + E\left[ \rho ^2 X_{t-1}^2 + \rho (1-\rho ) X_{t-1}\right] \\&= \mu _X\cdot \big ((1+b)(1-(\rho +a))+(1+2b)(\rho +a)-\rho ^2\big )+ (\rho +a)^2\cdot \mu _{2,X}. \end{aligned}$$

This confirms formula (6) about the variance. Analogously,

$$\begin{aligned} \mu _{3,X}&:= E\left[ X_t^3\right] =\ E\big [\mu _{3,{\text {Po}}(a\cdot X_{t-1}+b)}\big ]+ E\big [\mu _{3,{\text {B}}(X_{t-1},\rho )}\big ]\\&\quad +\,3\cdot E\big [\mu _{1,{\text {B}}(X_{t-1},\rho )}\cdot \mu _{2,{\text {Po}}(a\cdot X_{t-1}+b)}\big ]+3\cdot E\big [\mu _{2,{\text {B}}(X_{t-1},\rho )}\cdot \mu _{1,{\text {Po}}(a\cdot X_{t-1}+b)}\big ]\\&= E\big [b(1+3 b+b^2) + a(1+6b+3b^2)\cdot X_{t-1} + 3 a^2(1+b)\cdot X_{t-1}^2+a^3\cdot X_{t-1}^3\big ]\\&\quad +\,E\big [\rho (1-\rho )(1-2\rho )\cdot X_{t-1} +3\rho ^2 (1-\rho ) \cdot X_{t-1}^2+\rho ^3\cdot X_{t-1}^3\big ]\\&\quad +\,3\cdot E\left[ \rho b(1+b)\cdot X_{t-1} + \rho a(1+2b)\cdot X_{t-1}^2 + \rho a^2\cdot X_{t-1}^3\right] \\&\quad +\,3\cdot E\left[ \rho (1-\rho ) b\cdot X_{t-1} + \big (\rho ^2 b + \rho (1-\rho ) a\big )\cdot X_{t-1}^2 + \rho ^2 a\cdot X_{t-1}^3\right] \\&= b(1+3 b+b^2) + \big ((\rho +a)(1+6b+3b^2)-3\rho ^2 (1+b)+2\rho ^3\big )\cdot \mu _{X}\\&\quad +\,3(\rho +a)\big ((\rho +a) (1+b)-\rho ^2\big )\cdot \mu _{2,X}+ (\rho +a)^3\cdot \mu _{3,X}, \end{aligned}$$

from which we obtain an expression for \(\mu _{3,X}\). After tedious algebra, we find an expression for the third central moment:

$$\begin{aligned} E[(X_t-\mu _X)^3]&= \mu _{3,X}-3\mu _{X}\mu _{2,X}+2\mu _{X}^3=\cdots \\&= \mu _{X}\cdot \frac{(1-\rho ^3)(1-\rho ^2+a \rho +2a^2)+3a \rho (1-\rho )}{\big (1-(a+\rho )^2\big )\big (1-(a+\rho )^3\big )}. \end{aligned}$$

Next, we derive an expression for the autocovariance function. For \(k\ge 1\), we obtain by applying formula (5) and by conditioning that

$$\begin{aligned} Cov[X_t, X_{t-k}]=Cov\big [(\rho +a)\cdot X_{t-1}+b,\ X_{t-k}\big ]=\cdots =(\rho +a)^k\cdot \sigma _X^2. \end{aligned}$$

Finally, the proof of formula (11) is done by induction, with this formula being true for \(k=1\) according to the expression for \(E[X_t\ |\ X_{t-1},\ldots ]\) in formula (5).

1.3 Proofs of Sect. 3.2

Since \(E[\epsilon _t | X_{t-1},\ldots ] = V[\epsilon _t | X_{t-1},\ldots ] = a\cdot X_{t-1}+b\), we obtain

$$\begin{aligned} E[\epsilon _t]&= E\big [E[\epsilon _t | X_{t-1},\ldots ]\big ] = a\cdot \mu _X+b = (1-\rho )\cdot \mu _X,\\ V[\epsilon _t]&= E\big [V[\epsilon _t | X_{t-1},\ldots ]\big ] + V\big [E[\epsilon _t\ |\ X_{t-1},\ldots ]\big ] = a\cdot \mu _X+b\ +\ a^2\cdot \sigma _X^2\\&= (1-\rho )\cdot \frac{1-(\rho +a)^2}{1-\rho ^2}\cdot \sigma _X^2 + a^2\cdot \sigma _X^2 = \frac{\sigma _X^2}{1+\rho }\cdot \big (1-\rho ^2-\rho a (2-a)\big ). \end{aligned}$$

In particular, this result shows that \(V[\epsilon _t] = E[\epsilon _t] + a^2\cdot \sigma _X^2\), so the proof of formula (12) is complete. Similarly, we prove formula (13) via

$$\begin{aligned} E[z^{\epsilon _t}]&= E\big [E[z^{\epsilon _t}| X_{t-1},\ldots ]\big ]=E\big [\exp {\big ((a\cdot X_{t-1}+b)(z-1)\big )}\big ]\\&= e^{b(z-1)}\cdot p_X\big (e^{a(z-1)}\big ). \end{aligned}$$

Next, we consider the covariance between \(X_t\) and \(\epsilon _{t-j}\) with \(j\in \mathbb N_0\). First, using the conditional independence between \(\rho \circ X_{t-1}\) and \(\epsilon _{t}\) given \(X_{t-1}\), we have

$$\begin{aligned} Cov[X_t, \epsilon _{t}]&= Cov[\rho \circ X_{t-1}, \epsilon _{t}]+ Cov[\epsilon _{t},\epsilon _{t}]\\&= Cov[\rho \cdot X_{t-1},\ a\cdot X_{t-1}+b]+\sigma _{\epsilon }^2 =\rho a \cdot \sigma _{X}^2+\sigma _{\epsilon }^2. \end{aligned}$$

Secondly, for \(j\ge 1\) and using formula (5),

$$\begin{aligned} Cov[X_t, \epsilon _{t-j}]&= Cov\big [(\rho +a)\cdot X_{t-1} + b,\ \epsilon _{t-j}\big ] \\&= \ldots \ =\ (\rho +a)^j\cdot Cov[X_{t-j}, \epsilon _{t-j}]\ =\ (\rho +a)^j\cdot \left( \rho a \cdot \sigma _{X}^2 + \sigma _{\epsilon }^2\right) . \end{aligned}$$

So through conditioning on \(X_{t-1},\ldots , \epsilon _{t-1},\ldots \), we obtain for \(j\ge 1\) that

$$\begin{aligned} Cov[\epsilon _t,\epsilon _{t-j}]=Cov[a\cdot X_{t-1}+b,\epsilon _{t-j}]=a\cdot (\rho +a)^{j-1}\cdot \left( \rho a\cdot \sigma _{X}^2 + \sigma _{\epsilon }^2\right) . \end{aligned}$$

1.4 Proofs of Sect. 4.1

The score functions (17) to (19) can be derived in analogy to Sect. 3 in Freeland and McCabe (2004). Since

$$\begin{aligned} \frac{\partial }{\partial \theta _i}\,\ell ({\varvec{\theta }})=\sum _{t=1}^T\ \frac{\frac{\partial }{\partial \theta _i}\,P(X_t|X_{t-1})}{P(X_t|X_{t-1})}, \end{aligned}$$

it is essential to discuss the first-order partial derivatives of the transition probabilities \(P_{{\varvec{\theta }}}(k|l)\) according to (3). Let us start with the derivative in \(\rho \). Since (Freeland and McCabe 2004, p. 716)

$$\begin{aligned}&\frac{\partial }{\partial \rho }\,\left( {l \atopwithdelims ()j}\,\rho ^j\,(1-\rho )^{l-j}\right) \\&\quad =\frac{l}{1-\rho }\,\left( {l-1\atopwithdelims ()j-1}\,\rho ^{j-1}\,(1-\rho )^{l-j}-{l \atopwithdelims ()j}\,\rho ^j\,(1-\rho )^{l-j}\right) , \end{aligned}$$

and since

$$\begin{aligned} e^{-(a\,l+b)}\,\frac{(a\,l+b)^{k-j}}{(k-j)!}=e^{-\big (a\,(l-1)+a+b\big )}\,\frac{\big (a\,(l-1)+a+b\big )^{(k-1)-(j-1)}}{\big ((k-1)-(j-1)\big )!}, \end{aligned}$$

we can express

$$\begin{aligned} \frac{\partial }{\partial \rho }\,P_{\rho ,a,b}(k|l)=\frac{l}{1-\rho }\,\big (P_{\rho ,a,a+b}(k-1|l-1)-P_{\rho ,a,b}(k|l)\big ). \end{aligned}$$

So (17) follows. Differentiating with respect to \(a\), we have

$$\begin{aligned}&\frac{\partial }{\partial a}\,\left( e^{-(a\,l+b)}\,(a\,l+b)^{k-j}\right) \\&\quad =l\,\left( (k-j)\,e^{-(a\,l+b)}\,(a\,l+b)^{k-1-j}-e^{-(a\,l+b)}\,(a\,l+b)^{k-j}\right) . \end{aligned}$$

So (18) follows, since

$$\begin{aligned} \frac{\partial }{\partial a}\,P_{\rho ,a,b}(k|l)=l\,\big (P_{\rho ,a,b}(k-1|l)\ -\ P_{\rho ,a,b}(k|l)\big ). \end{aligned}$$

Finally, we have

$$\begin{aligned} \frac{\partial }{\partial b}\,\left( e^{-(a\,l+b)}\,(a\,l+b)^{k-j}\right) =(k-j)\,e^{-(a\,l+b)}\,(a\,l+b)^{k-1-j}-e^{-(a\,l+b)}\,(a\,l+b)^{k-j}, \end{aligned}$$

and hence

$$\begin{aligned} \frac{\partial }{\partial b}\,P_{\rho ,a,b}(k|l)=P_{\rho ,a,b}(k-1|l)-P_{\rho ,a,b}(k|l), \end{aligned}$$

in analogy to Proposition 2 in Freeland and McCabe (2004). This completes the proof of (19).

1.5 Proofs of Sect. 5.1

To compute the derivative \(\frac{\partial }{\partial \tau }\,\ell \big ((1-\tau )\xi ,\ \tau \xi ,\ b\big )\), we proceed in analogy to the proof of formulae (17) and (18), but by also applying the chain rule. Since \(\frac{\partial \rho }{\partial \tau } = -\xi \) and \(\frac{\partial a}{\partial \tau } = \xi \), we obtain

$$\begin{aligned}&\frac{\partial }{\partial \tau }\,P_{(1-\tau )\xi ,\tau \xi ,b}(k|l)\\&\quad =\frac{-\xi \,l}{1-(1-\tau )\xi }\,\Big (P_{(1-\tau )\xi ,\tau \xi ,\tau \xi +b}(k-1|l-1) -P_{(1-\tau )\xi ,\tau \xi ,b}(k|l)\Big )\\&\qquad +\,\xi \,l\left( P_{(1-\tau )\xi ,\tau \xi ,b}(k-1|l)- P_{(1-\tau )\xi ,\tau \xi ,b}(k|l)\right) \ \ \\&\quad =\xi \,l\,\left( \frac{(1-\tau )\xi }{1-(1-\tau )\xi }\,P_{(1-\tau )\xi ,\tau \xi ,b}(k|l) +P_{(1-\tau )\xi ,\tau \xi ,b}(k-1|l)\right. \\&\left. \qquad -\,\frac{1}{1-(1-\tau )\xi }\,P_{(1-\tau )\xi ,\tau \xi ,\tau \xi +b}(k-1|l-1)\right) . \end{aligned}$$

Hence, it follows that

$$\begin{aligned}&\left. \frac{\frac{\partial }{\partial \tau }\,P_{(1-\tau )\xi ,\tau \xi ,b}(k|l)}{P_{(1-\tau )\xi ,\tau \xi ,b}(k|l)}\right| _{\tau =0}\\&\quad =\xi \,l\,\left( \frac{\xi }{1-\xi }+\frac{P_{\xi ,0,b}(k-1|l)}{P_{\xi ,0,b}(k|l)}-\frac{1}{1-\xi }\,\frac{P_{\xi ,0,b}(k-1|l-1)}{P_{\xi ,0,b}(k|l)}\right) , \end{aligned}$$

from which we obtain the formula for the score function \(S_{\tau ;T}(\xi ,b)\).

For the sake of readability, let us now use the abbreviation \(g(X_t,X_{t-1})\) for the summands of \(S_{\tau ;\, T}(\xi ,b)\). Under the null of a Poisson INAR(1) model, it follows from Proposition 2 in Freeland and McCabe (2004) that

$$\begin{aligned} \frac{P_{\xi ,0,b}(X_t-1|X_{t-1})}{P_{\xi ,0,b}(X_t|X_{t-1})}=\frac{E[\epsilon _t|X_t,X_{t-1},\ldots ]}{b} \end{aligned}$$

and

$$\begin{aligned} \xi \,X_{t-1}\,\frac{P_{\xi ,0,b}(X_t-1|X_{t-1}-1)}{P_{\xi ,0,b}(X_t|X_{t-1})} =E[\xi \circ X_{t-1}|X_t,X_{t-1},\ldots ]. \end{aligned}$$

Hence, the conditional mean of \(g(X_t,X_{t-1})\) becomes

$$\begin{aligned}&E\big [g(X_t,X_{t-1})\ |\ X_{t-1}, \ldots \big ]\\&\quad =\frac{\xi ^2\,X_{t-1}\,}{1-\xi }+\xi \,X_{t-1}\,\frac{E[\epsilon _t |X_{t-1},\ldots ]}{b}-\frac{1}{1-\xi }\,E[\xi \circ X_{t-1}|X_{t-1},\ldots ]\\&\quad =\frac{\xi ^2\,X_{t-1}\,}{1-\xi }+\xi \,X_{t-1}- \frac{\xi \,X_{t-1}}{1-\xi }=0. \end{aligned}$$

As a result,

$$\begin{aligned} E\big [S_{\tau ; T}(\xi ,b)\big ]=\sum _{t=1}^T\,E\Big [E\big [g(X_t,X_{t-1})|X_{t-1},\ldots \big ]\Big ]=0. \end{aligned}$$

Finally, the process of \(g(X_t,X_{t-1})\) is \(\alpha \)-mixing with exponentially decreasing weights, see Proposition 2, such that Theorem 1.7 of Ibragimov (1962) is applicable, which gives the stated asymptotic normality with \(\sigma ^2=E\big [g(X_1,X_{0})^2\big ]+ 2\,\sum _{k=1}^{\infty }\ E\big [g(X_1,X_{0})\,g(X_{k+1},X_{k})\big ]\). Using the above result \(E\big [g(X_t,X_{t-1})| X_{t-1}, \ldots \big ]=0\), it follows that

$$\begin{aligned} E\big [g(X_1,X_{0})\,g(X_{k+1},X_{k})\big ]=E\Big [g(X_1,X_{0})\,E\big [g(X_{k+1},X_{k})|X_{k-1},\ldots \big ]\Big ]=0. \end{aligned}$$

So the proof of Proposition 3 is complete.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Weiß, C.H. A Poisson INAR(1) model with serially dependent innovations. Metrika 78, 829–851 (2015). https://doi.org/10.1007/s00184-015-0529-9

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00184-015-0529-9

Keywords

Navigation