Skip to main content
Log in

Testing for the bivariate Poisson distribution

  • Published:
Metrika Aims and scope Submit manuscript

Abstract

This paper studies goodness-of-fit tests for the bivariate Poisson distribution. Specifically, we propose and study several Cramér–von Mises type tests based on the empirical probability generating function. They are consistent against fixed alternatives for adequate choices of the weight function involved in their definition. They are also able to detect local alternatives converging to the null at a certain rate. The bootstrap can be used to consistently estimate the null distribution of the test statistics. A simulation study investigates the goodness of the bootstrap approximation and compares their powers for finite sample sizes. Extensions for testing goodness-of-fit for the multivariate Poisson distribution are also discussed.

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.

Similar content being viewed by others

References

  • Baringhaus L, Gürtler N, Henze N (2000) Weighted integral test statistics and components of smooth tests of fit. Aust N Z J Stat 42:179–192

    Article  MathSciNet  Google Scholar 

  • Baringhaus L, Henze N (1992) A goodness of fit test for the Poisson distribution based on the empirical generating function. Stat Prob Lett 13:269–274

    Article  MATH  MathSciNet  Google Scholar 

  • Berkhout P, Plug E (2004) A bivariate Poisson count data model using conditional probabilities. Stat Neerlandica 58(3):349–364

    Article  MATH  MathSciNet  Google Scholar 

  • Crockett NG (1979) A quick test of fit of a bivariate distribution. In: McNeil D (ed) Interactive statistics. North-Holland, Amsterdam, pp 185–191

    Google Scholar 

  • Feuerverger A (1989) On the empirical saddlepoint approximation. Biometrika 76:457–464

    Article  MATH  MathSciNet  Google Scholar 

  • Gillings DB (1974) Some further results for bivariate generalizations of the Neyman Type A distribution. Biometrics 30:619–628

    Article  MATH  Google Scholar 

  • Gregory G (1977) Large sample theory for \(U\)-statistics and tests of fit. Ann Stat 5:110–123

    Article  MATH  MathSciNet  Google Scholar 

  • Gürtler N, Henze N (2000) Recent and classical goodness-of-fit tests for the Poisson distribution. J Stat Plan Inference 90:207–225

    Article  MATH  Google Scholar 

  • Hamdan MA (1972) Estimation in the truncated bivariate Poisson distribution. Technometrics 14:37–45

    Article  MATH  Google Scholar 

  • Ho LL, Singer JM (2001) Generalizad least squares methods for bivariate Poisson regression. Commun Stat Theory Methods 30(2):263–277

    Article  MATH  MathSciNet  Google Scholar 

  • Holgate P (1964) Estimation for the bivariate Poisson distribution. Biometrika 51:241–245

    Article  MATH  MathSciNet  Google Scholar 

  • Holgate P (1966) Bivariate generalizations of Neyman’s Type A distribution. Biometrika 53:241–245

    Article  MATH  MathSciNet  Google Scholar 

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

    MATH  Google Scholar 

  • Karlis D, Ntzoufras I (2005) Bivariate Poisson and diagonal inflated bivariate Poisson regression models in R. J Stat Softw 14(10):1–36

    Google Scholar 

  • Karlis D, Tsiamyrtzis P (2008) Exact Bayesian modeling for bivariate Poisson data and extensions. Stat Comput 18:27–40

    Article  MathSciNet  Google Scholar 

  • Kocherlakota S, Kocherlakota K (1992) Bivariate discrete distributions. Marcel Dekker Inc, New York

    MATH  Google Scholar 

  • Kundu S, Majumdar S, Mukherjee K (2000) Central limits theorems revisited. Stat Probab Lett 47:265–275

    Article  MATH  MathSciNet  Google Scholar 

  • Lehmann EL, Romano JP (2005) Testing statistical hypothesis. Springer, Berlin

    Google Scholar 

  • Loukas S, Kemp CD (1986) The index of dispersion test for the bivariate Poisson distribution. Biometrics 42:941–948

    Article  MATH  MathSciNet  Google Scholar 

  • Meintanis S, Swanepoel J (2007) Bootstrap goodness-of-fit tests with estimated parameters based on empirirical transforms. Stat Probab Lett 77:1004–1013

    Article  MATH  MathSciNet  Google Scholar 

  • Papageorgiou H, Loukas S (1988) Conditional even point estimation for bivariate discrete distributions. Commun Stat Theory Methods 17:3403–3412

    Article  MATH  Google Scholar 

  • Rayner JCW, Best DJ (1995) Smooth tests for the bivariate Poisson distribution. Aust N Z J Stat 37:233–245

    Article  MATH  MathSciNet  Google Scholar 

  • Rueda R, Pérez-Abreu V, O’Reilly F (1991) Goodness of fit for the Poisson distribution based on the probability generating function. Commun Stat Theory Methods A20:3093–3110

    Article  Google Scholar 

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

    Book  MATH  Google Scholar 

Download references

Acknowledgments

The authors would like to thank Professor Norbert Henze and an anonymous referee for their careful reading and comments which greatly improved the paper. F. Novoa-Muñoz also wishes to thank his institution, University of Bío-Bío (Chile) and the scholarship given by the Chilean Ministry of Education through the Superior Education MECE Program 2, which make his doctoral studies possible.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to M. D. Jiménez-Gamero.

Appendix

Appendix

1.1 Proofs

Here we prove the results in Sects. 25, 7. Since the mathematical machinery employed in the proofs is quite standard, we just sketch the proofs. A detailed derivation of the results can be obtained from the authors upon request.

In order to prove Proposition 1, we first give a preliminary lemma. For an arbitrary set \(S, \partial S\) and \(int S\) denote the sets of boundary points and interior points of \(S\), respectively.

Lemma 1

Let \(\{f_n\}_{n\ge 1}\) and \(f\) be a collection of nondecreasing real functions defined on \(R=[b_1,c_1]\times [b_2,c_2]\times \cdots \times [b_d,c_d] \subseteq \mathbb {R}^d\), with \(-\infty <b_j\le c_j<\infty , j=1,2,\ldots ,d\). Let \(D=D_1\cup D_2 \cup D_3\), where \(D_1\) is the vertex set of \(R, D_2\) is a dense set in \(\partial R\) and \(D_3\) is a dense set in \(int R\). If \(f\) is continuous on \(R\) and

$$\begin{aligned} \left| f_{n}(x)-f(x)\right| \rightarrow 0, \quad \forall x \in D, \end{aligned}$$
(14)

then

$$\begin{aligned} \displaystyle \sup _{x\in R}|f_n(x)-f(x)| \rightarrow 0. \end{aligned}$$

Proof

Let \(\varepsilon >0\) be arbitrary but fixed. Because \(f\) is a uniformly continuous function on \(R, \exists \delta =\delta (\varepsilon )>0\), such that \(\forall x,y\in R\) with \(\Vert x-y\Vert <\delta \), we have that \(|f(x)-f(y)|<\varepsilon \). Let

$$\begin{aligned} H_i=[u_{1i},v_{1i}]\times [u_{2i},v_{2i}]\times \cdots \times [u_{di},v_{di}],\quad i=1,2,\ldots ,M, \end{aligned}$$

with \(\Vert u_i-v_i\Vert < \delta \), where \(u_i=(u_{1i},u_{2i},\ldots ,u_{di}) \in D\) and \(v_i=(v_{1i},v_{2i},\ldots ,v_{di}) \in D, 1\le i \le M\), such that \(R=\bigcup _{i=1}^M H_i.\) Let \(\Delta _n=\max _{x\in V} \left| f_n(x)-f(x)\right| \), where \(V=\{u_i,v_i: i=1,2,\ldots ,M\}\). From (14), it follows that \(\Delta _n \rightarrow 0\). Now, for each \(x\in R\), there exists \(i\in \{1,2,\ldots ,M\}\), such that \(x\in H_i\), and thus \(u_i\le x\le v_i\). Since \(f_n\) and \(f\) are nondecreasing, we get

$$\begin{aligned} f_n(x)-f(x)\le f_n(v_i)-f(u_i)=f_n(v_i)-f(v_i)+f(v_i)-f(u_i)\le \Delta _n+\varepsilon . \end{aligned}$$

Analogously, \(f_n(x)-f(x)\ge -\Delta _n-\varepsilon \). Since \(\varepsilon >0\) is arbitrary, these two inequalities imply the result. \(\square \)

Proof of Proposition 1

Let \(D\) be a countable dense set in \(R\) according to Lemma 1. By the SLLN, there exists a set \(A_D \in \mathcal {A}\) such that \(P(A_D)=1\) and for all \(\omega \in A_D\): \(g_n(u,\omega )\ \mathop {\longrightarrow } \ g(u), \forall u \in D\). Since \(X_i\ge 0, 1\le i \le d\), we have that \(g_n(u,\omega )\) and \(g(u)\) are nondecreasing functions. Now (2) follows from Lemma 1. The proof of (3) follows similar steps. \(\square \)

Proof of Proposition 2

Let \((X_1,X_2)\) be a random vector and let \(g(u_1,u_2)=E\left( u_1^{X_1} u_2^{X_2}\right) \) be its pgf. Then, from the first equation in (6)

$$\begin{aligned} \frac{\partial }{\partial u_1}\log g(u_1,u_2)=\theta _1+\theta _3(u_2-1). \end{aligned}$$
(15)

Integrating (15) over \(u_1\), we obtain

$$\begin{aligned} g(u_1,u_2)&= \exp \{\phi _1(u_2)+\theta _1 u_1+\theta _3(u_2-1) u_1\},\\&= \exp \{\phi (u_2)+\theta _1(u_1-1)+\theta _3(u_1-1)(u_2-1)\}, \end{aligned}$$

where \(\phi (u_2)=\phi _1(u_2)+\theta _1+\theta _3(u_2-1)\). Proceeding similarly, from the second equation in (6) we get

$$\begin{aligned} g(u_1,u_2)= \exp \{\varphi (u_1)+\theta _2(u_2-1)+\theta _3(u_1-1)(u_2-1)\}. \end{aligned}$$

Thus, necessarily \(\varphi (u_1)=\theta _1(u_1-1)\) and \(\phi (u_2)=\theta _2(u_2-1)\), in other words, the BDP is the only solution of (6) \(\square \)

Proof of Theorem 1

By definition, \(R_{n,w}(\hat{{\theta }}_n)= ||G_n(\hat{{\theta }}_n)||_{\mathcal {H}}^2\). Note that

$$\begin{aligned} G_n(u;\hat{\theta }_n)=\frac{1}{\sqrt{n}}\sum _{i=1}^n G(\varvec{X}_i,\hat{\theta }_n; {u}), \end{aligned}$$

with

$$\begin{aligned} G(\varvec{X},\theta ; {u})= u_1^{X_{1}}u_2^{X_{2}}-g(u;{\theta }). \end{aligned}$$
(16)

To prove (7), we apply a Taylor expansion of \(G(\varvec{X}_i,\hat{\theta }_n; {u})\) around \(\hat{\theta }_n=\theta \). Now following the lines in the proof of Theorem 3.1 in Meintanis and Swanepoel (2007) we get (7). Observe that

$$\begin{aligned} \Vert R_{n}\Vert _{\mathcal {H}}^{2}=\frac{1}{n} \sum _{i=1}^n\sum _{j=1}^n h(\varvec{X}_i,\varvec{X}_j;\theta ), \end{aligned}$$

where, \(h(\varvec{x},\varvec{y};\theta )\) is defined in (10) and satisfies \(h(\varvec{x},\varvec{y};\theta )= h(\varvec{y},\varvec{x};\theta ), E_{\theta }\left\{ h^2(\varvec{X}_1,\right. \left. \varvec{X}_2;\theta )\right\} < \infty , E_{\theta } \left\{ |h(\varvec{X}_1,\varvec{X}_1;\theta )|\right\} < \infty \) and \(E_{\theta }\{h(\varvec{X}_1,\varvec{X}_2;\theta )\}=0\). Thus, from Theorem 6.4.1.B in Serfling (1980),

$$\begin{aligned} \Vert R_{n}\Vert _{\mathcal {H}}^{2} \mathop {\longrightarrow } \limits ^{L}\ \sum _{j\ge 1}\lambda _j\chi ^2_{1j}, \end{aligned}$$

where \(\chi ^2_{11},\chi ^2_{12},\ldots \) and the set \(\{\lambda _j\}\) are as defined in the statement of the Theorem. In particular, \(\Vert R_{n}\Vert _{_\mathcal {H}}^{2}=O_P(1)\), which implies (8). \(\square \)

Proof of Theorem 2

By definition \(R_{n,w}^*(\hat{\theta }_n^*)= \Vert G_{n}^*(\hat{\theta }_n^*)\Vert _{\mathcal {H}}^{2}\), with

$$\begin{aligned} G_{n}^*(u;\hat{\theta }_n^*)=\frac{1}{\sqrt{n}}\sum _{i=1}^n G(\varvec{X}_i^*;\hat{\theta }_n^*; u) \end{aligned}$$

and \(G(\varvec{X};\theta ; u)\) defined in (16). Following similar steps to those given in the proof of Theorem 1 it can be seen that \(R_{n,w}^*(\hat{\theta }_n^*)= \Vert R_{n}^*\Vert _{\mathcal {H}}^{2}+r^*_{n}\) with \(r_{n}^*=o_{P_*}(1)\) a.s., where \(R_{n}^*(u)\) is defined as \(R_{n}(u)\) with \(\varvec{X}_i\) and \(\theta \) replaced by \(\varvec{X}^*_i\) and \(\hat{\theta }_n\), respectively. Now, the result follows by applying Theorem 1.1 in Kundu et al. (2000) to \(R_{n}^*\) and the continuous mapping theorem. \(\square \)

Proof of Theorem 3

We only present the proof of part (b), since the proof of part (a) is quite similar. By definition,

$$\begin{aligned} \frac{1}{n}S_{n,w}(\hat{\theta }_n)=\Vert E_{1n}\Vert ^2_{_\mathcal {H}}+ \Vert E_{2n}\Vert ^2_{_\mathcal {H}}. \end{aligned}$$

We have

$$\begin{aligned} \left| E_{1n}(u;\hat{\theta }_n )-E_1(u;\theta )\right|&\le r_1+|\theta _1-\hat{\theta }_{1n}| + |\theta _3-\hat{\theta }_{3n}|+ |\hat{\theta }_{1n}+\hat{\theta }_{3n}|r_0, \\ \left| E_{2n}(u;\hat{\theta }_n )-E_2(u;\theta )\right|&\le r_2+|\theta _2-\hat{\theta }_{2n}| + |\theta _3-\hat{\theta }_{3n}|+ |\hat{\theta }_{2n}+\hat{\theta }_{3n}|r_0, \end{aligned}$$

\(\forall u \in [0,1]^2\), where \(r_0=\sup _{u\in [0,1]^2}\left| g_n(u)- g(u)\right| \) and \(r_i=\sup _{u\in [0,1]^2}\left| \frac{\partial }{\partial u_i}g_n(u)-\right. \left. \frac{\partial }{\partial u_i}g(u)\right| , i=1,2\). Since \(r_i=o(1), i=0,1,2, |\theta _i-\hat{\theta }_{in}| =o(1), i=1,2,3\), and (5), we conclude that \(\frac{1}{n}S_{n,w}(\hat{\theta }_n)=\Vert E_{1}\Vert ^2_{\mathcal {H}}+ \Vert E_{2}\Vert ^2_{\mathcal {H}}+o(1)\), which shows the result because \(\Vert E_{1}\Vert ^2_{\mathcal {H}}+ \Vert E_{2}\Vert ^2_{\mathcal {H}}=\sigma (g;\theta )\). \(\square \)

Proof of Theorem 4

Let \(A_n \subseteq \mathbb {N}_0^2\) so that \(P_{\theta }(A_n)\rightarrow 0\), then it can be easily seen that \(P_n(A_n)\rightarrow 0\), and thus \(P_n\) is contiguous to \(P_{\theta }\). Next we show that, under \(P_n, \sqrt{n}(\hat{\theta }_n-\theta )\) converges in distribution to a normal law. With this aim, let \(\varvec{X}_1, \varvec{X}_2, \ldots , \varvec{X}_n\) be iid with common pmf \(P_{\theta }\) and let

$$\begin{aligned} l_n=\log \frac{P_n(\varvec{X}_1)P_n(\varvec{X}_2)\ldots P_n(\varvec{X}_n)}{P_{\theta }(\varvec{X}_1)P_{\theta }(\varvec{X}_2)\ldots P_{\theta }(\varvec{X}_n)}. \end{aligned}$$

Let \(Z_{n,i}=b_n(\varvec{X}_i), 1\le i \le n\). Then, by a Taylor expansion,

$$\begin{aligned} l_n=\sum _{i=1}^n \log \left( 1+\frac{1}{\sqrt{n}}Z_{n,i} \right) =\frac{1}{\sqrt{n}}\sum _{i=1}^n Z_{n,i}-\frac{1}{2n}\sum _{i=1}^n Z_{n,i}^2+\varrho _n, \end{aligned}$$
(17)

where \(\varrho _n=\varrho _n(Z_{n,1},Z_{n,2}, \ldots , Z_{n,n})\),

$$\begin{aligned} \varrho _n=\sum _{i=1}^n \frac{2}{3!\,n\sqrt{n}\,a^3_{in}} Z_{n,i}^3, \end{aligned}$$

with \(a_{in}=a_{in}(Z_{n,i})\)

$$\begin{aligned} a_{in}=1+\frac{\alpha _i}{\sqrt{n}}\,Z_{n,i}, \end{aligned}$$

for some \(0<\alpha _i<1, 1\le i \le n\). Next, we study the limit of each term in the right side of (17).

Let \(\varepsilon \) be a positive constant. Let \(M_n=\sqrt{n}/2\) and \(\tilde{\varrho }_n=\varrho _n(Z_{n,1}I\{|Z_{n,1}|\le M_n\}, \ldots , Z_{n,n}I\{|Z_{n,n}|\le M_n\})\). Note that \(\tilde{a}_{in}=a_{in}(Z_{n,i}I\{|Z_{n,i}|\le M_n\})\ge 0.5, 1\le i \le n\), and therefore

$$\begin{aligned} E_{\theta }(|\tilde{\varrho }_n|)\le \frac{2^3}{3}\frac{1}{\sqrt{n}} \displaystyle \sup _m E_{\theta } \left[ |b_m(\varvec{X}_1)|^3 \right] , \end{aligned}$$

which implies

$$\begin{aligned} P_{\theta }(|\tilde{\varrho }_{n}|>\varepsilon )\rightarrow 0. \end{aligned}$$
(18)

We also have

$$\begin{aligned} P_{\theta }\left( |\tilde{\varrho }_{n} -\varrho _{n}|>\varepsilon \right)&\le P_{\theta }\left( \,\bigcup _{i=1}^n\left\{ |Z_{n,i}|> M_n \right\} \right) \nonumber \\&\le \sum _{i=1}^n P_{\theta }\left( |Z_{n,i}|> M_n \right) \le \frac{n}{M_n^4}\sup _m E_{\theta } \left[ b_m(\varvec{X}_1)^4 \right] \rightarrow 0. \end{aligned}$$
(19)

From (18) and (19), we conclude

$$\begin{aligned} P_{\theta }\left( |\varrho _{n}|>\varepsilon \right) \rightarrow 0. \end{aligned}$$
(20)

From Assumption 3, the central limit theorem for triangular arrays implies

$$\begin{aligned} \frac{1}{\sqrt{n}}\sum _{i=1}^n Z_{n,i}\ \mathop {\longrightarrow } \limits ^{L}\ N(0,\sigma ^2), \end{aligned}$$
(21)

with \(\sigma ^2=E_{\theta }\left\{ b(X_1,X_2)^2\right\} <\infty \).

From Assumption 3 we also have \(E_{\theta }\left( \frac{1}{n}\sum _{i=1}^n Z_{n,i}^2\right) \rightarrow \sigma ^2\) and \(Var_{\theta }\left( \frac{1}{n}\sum _{i=1}^n\right. \left. Z_{n,i}^2\right) = \frac{1}{n}Var_{\theta }(Z_{n,i}^2)\le \frac{1}{n}\sup _n E_{\theta }\left\{ b_n(X_1,X_2)^4\right\} \rightarrow 0\), which imply

$$\begin{aligned} \frac{1}{n}\sum _{i=1}^n Z_{n,i}^2 \mathop {\longrightarrow } \limits ^{P}\ \sigma ^2. \end{aligned}$$
(22)

Now, from Assumption 1, (17) and (20)–(22), the sequence \((\sqrt{n}(\hat{\theta }_n-\theta ), l_n)\) converges in law to a 4-dimensional normal variate when the data comes from \(P_{\theta }\). Thus, from Le Cam’s Third Lemma (Corollary 12.3.2 in Lehmann and Romano 2005), we have that, when the data come from \(P_n, \sqrt{n}(\hat{\theta }_n-\theta )\) converges in distribution to a normal law, which implies that it is bounded in probability. Now, following similar steps as those given in the proof of Theorem 1, it can be shown that (7) also holds when the data come from \(P_n\), with \(P_n(|r_n|>\varepsilon )\rightarrow 0, \forall \varepsilon >0\). Finally, applying Theorem 2.3 in Gregory (1977), we obtain that when the data have pmf \(P_n\)

$$\begin{aligned} \Vert R_{n}\Vert _{\mathcal {H}}^2 \mathop {\longrightarrow } \limits ^{L} \ \sum _{k=1}^{\infty } \lambda _j\left( Z_j+c_j\right) ^2, \end{aligned}$$

and thus the result. \(\square \)

Proof of Proposition 3

The proof follows is quite similar to that of Proposition 2, so we omit it. \(\square \)

1.2 Expressions for statistics \(\varvec{T}, \varvec{I_B}\) and \(\varvec{NI_B}\) and critical regions

We have

$$\begin{aligned} T&= \frac{n}{2}\frac{(\hat{\theta }_2)^{2} \left( S^2_{X_1}\!-\!\hat{\theta }_1\right) ^2\!-\! 2S^2_{X_1X_2}\left( S^2_{X_1}\!-\!\hat{\theta }_1\right) \left( S^2_{X_2}\!-\!\hat{\theta }_2\right) \!+\! (\hat{\theta }_1)^{2} \left( S^2_{X_2}\!-\!\hat{\theta }_2\right) ^2}{(\hat{\theta }_1)^{2} (\hat{\theta }_2)^{2}\!-\!S^4_{X_1X_2}}, \\&T \ge \chi ^2_{2,\alpha },\\ I_B&= n \frac{\hat{\theta }_2S^2_{X_1}-2S^2_{X_1X_2}+ \hat{\theta }_1S^2_{X_2}}{\hat{\theta }_1\hat{\theta }_2- S^2_{X_1X_2}},\quad I_B \ge \chi ^2_{2n-3,\alpha },\\ NI_B&= \frac{n}{1-r^2}\left( \frac{S^2_{X_1}}{\overline{X}_1} - 2r^2\sqrt{\frac{S^2_{X_1}S^2_{X_2}}{\overline{X}_1\overline{X}_2}}+ \frac{S^2_{X_2}}{\overline{X}_2}\right) , \,\, NI_B \ge \chi ^2_{2n-3,\alpha }, \end{aligned}$$

where \(S^2_{X_i}\) is the sample variance, \(i=1,2, S^2_{X_1X_2}\) is the sample covariance, \(r\) is the sample correlation coefficient and \(\chi ^2_{k, \alpha }\), for \(0<\alpha <1\) and \(k\in \mathbb {N}\), denotes the upper \(\alpha \)-percentile of the \(\chi ^2\) distribution with \(k\) degrees of freedom.

1.3 Computation of the test statistic \(R_{n,w}(\hat{\theta }_n)\)

For \(w\) as in (12), for \(a_1, a_2\in \mathbb {N}_0\), we get

$$\begin{aligned}&R_{n,w}(\hat{\theta }_n)=\frac{1}{n}\sum _{i=1}^n \sum _{j=1}^n\int \limits _0^1\int \limits _0^1 u_1^{X_{1i} +X_{1j}+a_1}u_2^{X_{2i}+X_{2j}+a_2}\ du\nonumber \\&\quad +\, ne^{2\left( \hat{\theta }_{3n}-\hat{\theta }_{1n}-\hat{\theta }_{2n}\right) } \int \limits _0^1 u_1^{a_1}e^{2\left[ \hat{\theta }_{1n} -\hat{\theta }_{3n}\right] u_1}\int \limits _0^1 u_2^{a_2}e^{2\left[ \hat{\theta }_{2n}-\hat{\theta }_{3n}+\hat{\theta }_{3n}u_1\right] u_2} du\nonumber \\&\quad -\, 2e^{\hat{\theta }_{3n}-\hat{\theta }_{1n}- \hat{\theta }_{2n}}\sum _{i=1}^n\int \limits _0^1 u_1^{X_{1i}+a_1}e^{\left[ \hat{\theta }_{1n} -\hat{\theta }_{3n}\right] u_1}\int \limits _0^1 u_2^{X_{2i}+a_2} e^{\left[ \hat{\theta }_{2n}-\hat{\theta }_{3n}+\hat{\theta }_{3n}u_1 \right] u_2}du.\qquad \end{aligned}$$
(23)

To compute \(R_{n,w}(\hat{\theta }_n)\) we can calculate the integrals on the right hand side of (23). Nevertheless, for computational reasons, the following representation (which is obtained by interchanging the order of integration)

$$\begin{aligned} R_{n,w}(\hat{\theta }_n) = n \sum _{i,j,k,l=0}^{\infty } \frac{\left( p_n(i,j) -p(i,j;\hat{\theta }_n)\right) \left( p_n(k,l)- p(k,l;\hat{\theta }_n)\right) }{(i+k+a_1+1)(j+l+a_2+1)}\qquad \end{aligned}$$
(24)

turned out to be more appropriate, where \(p(i, j;\theta )=P_{\theta }(X_1=i, X_2=j)\) and \(p_n(i,j)\) is the relative frequency of the pair \((i,j)\). Besides, (24) is less restrictive, since it allows \(a_1>-1\), and \(a_2>-1\). A truncation of the four infinite series at \(M+15\) yielded sufficiently precise values of the statistic and a good performance of the corresponding subroutine, where \(\displaystyle M\!=\!\max \{ X_{1(n)},X_{2(n)}\}, X_{k(n)}\!=\!\max \nolimits _{1\le i\le n}X_{ki}, k=1,2\).

1.4 Computation of the test statistic \(S_{n,w}(\hat{\theta }_n)\)

When the weight function is chosen as in (12), for some \(a=(a_1, a_2)\) with \(a_i>-1, i=1,2\), we have

$$\begin{aligned} S_{n,w}(\hat{\theta }_n) =\frac{1}{n}\sum _{i=1}^n\sum _{j=1}^n (S_{1ij}+S_{2ij}), \end{aligned}$$

with

$$\begin{aligned} S_{kij}&= \frac{X_{ki}I_{B_{ki}}X_{1j}I_{B_{kj}}}{(X_{1i}+X_{1j}+a_1-1) (X_{2i}+X_{2j}+a_2+1)}\\&- \frac{(\hat{\theta }_{kn}- \hat{\theta }_{3n})\left( X_{ki}I_{B_{ki}}+ X_{kj}I_{B_{kj}}\right) }{(X_{1i}+X_{1j}+a_1)(X_{2i}+X_{2j}+a_2+1)} \\&-\frac{\hat{\theta }_{3n}(X_{ki} I_{B_{ki}}+X_{kj}I_{B_{kj}})}{(X_{1i}+X_{1j}+a_1)(X_{2i}+ X_{2j}+a_2+2)}\\&+\frac{2\hat{\theta }_{3n}(\hat{\theta }_{kn}- \hat{\theta }_{3n})}{(X_{1i}+X_{1j}+a_1+1)(X_{2i}+ X_{2j}+a_2+2)}\\&+ \frac{(\hat{\theta }_{kn}- \hat{\theta }_{3n})^2}{(X_{1i}+X_{1j}+a_1+1)(X_{2i}+X_{2j}+a_2+1)}\\&+\frac{\hat{\theta }_{3n}^{\ 2}}{(X_{1i}+X_{1j}+a_1+1)(X_{2i}+X_{2j}+a_2+3)}, \end{aligned}$$

\(k=1,2, 1\le i,j \le n\), where \( B_{rs}=\{X_{rs}\ge 1\}, r=1,2, 1\le s \le n\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Novoa-Muñoz, F., Jiménez-Gamero, M.D. Testing for the bivariate Poisson distribution. Metrika 77, 771–793 (2014). https://doi.org/10.1007/s00184-013-0464-6

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00184-013-0464-6

Keywords

Navigation