Skip to main content
Log in

Some new statistical methods for a class of zero-truncated discrete distributions with applications

  • Original paper
  • Published:
Computational Statistics Aims and scope Submit manuscript

Abstract

Counting data without zero category often occurs in various fields. A class of zero-truncated discrete distributions such as the zero-truncated Poisson, zero-truncated binomial and zero-truncated negative-binomial distributions are proposed in literature to model such count data. In this paper, three main contributions have been made for better studying the zero-truncated discrete distributions: First, a novel unified expectation–maximization (EM) algorithm is developed for calculating the maximum likelihood estimates (MLEs) of parameters in general zero-truncated discrete distributions and an important feature of the proposed EM algorithm is that the latent variables and the observed variables are independent, which is unusual in general EM-type algorithms; Second, for those who do not understand the principle of latent variables, a unified minorization–maximization algorithm, as an alternative to the EM algorithm, for obtaining the MLEs of parameters in a class of zero-truncated discrete distributions is discussed; Third, a unified method is proposed to derive the distribution of the sum of i.i.d.zero-truncated discrete random variables, which has important applications in the construction of the shortest Clopper–Pearson confidence intervals of parameters of interest and in the calculation of the exact p value of a two-sided test for small sample sizes in one sample problem.

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

Similar content being viewed by others

References

  • Ahuja JC (1970) On the distribution of sum of independent positive binomial variables. Can Math Bull 13(1):151–152

    Article  MathSciNet  MATH  Google Scholar 

  • Ahuja JC (1971) Distribution of the sum of independent decapitated negative binomial variables. Ann Math Stat 42(1):383–384

    Article  MathSciNet  MATH  Google Scholar 

  • Ahuja JC (1972) Convolution of the truncated Poisson distribution. Am Stat 26(5):46

    Google Scholar 

  • Best DJ, Rayner JCW, Thas O (2007) Goodness of fit for the zero-truncated Poisson distribution. J Stat Comput Simul 77(7):585–591

    Article  MathSciNet  MATH  Google Scholar 

  • Bliss CI, Fisher RA (1953) Fitting the negative binomial distribution to biological data. Biometrics 9(2):176–200

    Article  MathSciNet  Google Scholar 

  • Cacoullos T (1961) A combinatorial derivation of the distribution of the truncated Poisson sufficient statistic. Ann Math Stat 32(3):904–905

    Article  MathSciNet  MATH  Google Scholar 

  • Cacoullos T (1972) On the convolution of Poisson distributions with the zero class missing. Am Stat 26(2):44

    Google Scholar 

  • Cohen AC Jr (1954) Estimation of the Poisson parameter from truncated samples and from censored samples. J Am Stat Assoc 49:158–168

    MathSciNet  MATH  Google Scholar 

  • Cohen AC Jr (1960) Estimating the parameter in a conditional Poisson distribution. Biometrics 16(2):203–211

    Article  MathSciNet  MATH  Google Scholar 

  • Consul PC (1989) Generalized Poisson distributions. Marcel Dekker, New York

    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(4):791–799

    Article  MathSciNet  MATH  Google Scholar 

  • Dahiya RC, Gross AJ (1973) Estimating the zero class from a truncated Poisson sample. J Am Stat Assoc 68:731–733

    Article  MathSciNet  MATH  Google Scholar 

  • David FN, Johnson NL (1952) The truncated Poisson. Biometrics 8(4):275–285

    Article  MathSciNet  Google Scholar 

  • Finney DJ (1949) The truncated binomial distribution. Ann Eugen 14:319–328

    Article  MathSciNet  MATH  Google Scholar 

  • Finney DJ, Varley GC (1955) An example of the truncated Poisson distribution. Biometrics 11(3):387–394

    Article  Google Scholar 

  • Grogger JT, Carson RT (1991) Models for truncated counts. J Appl Econ 6:225–238

    Article  Google Scholar 

  • Gross AJ (1970) A note on the convolution of Poisson distributions with the zero class missing. Am Stat 24(5):42–43

    Google Scholar 

  • Gupta RC (1974) Distribution of the sum of independent decapitated generalized negative binomial variables. Sankhyā Indian J Stat Ser B 36(1):67–69

    MathSciNet  MATH  Google Scholar 

  • Gurmu S (1991) Tests of detecting overdispersion in the positive Poisson regression model. J Bus Econ Stat 9(2):215–222

    MathSciNet  Google Scholar 

  • Hartley HO (1958) Maximum likelihood estimation from incomplete data. Biometrics 14(2):174–194

    Article  MATH  Google Scholar 

  • Irwin JO (1959) On the estimation of the mean of a Poisson distribution from a sample with the zero class missing. Biometrics 15(2):324–326

    Article  MathSciNet  MATH  Google Scholar 

  • Jain GC, Consul PC (1971) A generalized negative binomial distribution. SIAM J Appl Math 21(4):501–513

    Article  MathSciNet  MATH  Google Scholar 

  • Johnson NL, Kotz S (1969) Distributions in statistics: discrete distributions. Wiley, New York, pp 104–109

    MATH  Google Scholar 

  • Karlis D (2003) An EM algorithm for multivariate Poisson distribution and related models. J Appl Stat 30(1):63–77

    Article  MathSciNet  MATH  Google Scholar 

  • Kumar CS, Sreejakumari S (2012) On intervened negative binomial distribution and some of its properties. Statistica 72(4):395–404

    Google Scholar 

  • Malik HJ (1969) Distribution of the sum of truncated binomial variates. Can Math Bull 12:334–336

    Article  Google Scholar 

  • Maritz JS (1950) On the validity of inferences drawn from the fitting of Poisson and negative binomial distributions to observed accident data. Psychol Bull 47(5):434–443

    Article  Google Scholar 

  • Matthews JNS, Appleton DR (1993) An application of the truncated Poisson distribution to immunogold assay. Biometrics 49(2):617–621

    Article  Google Scholar 

  • Medhi J (1975) On the convolutions of left-truncated generalised negative binomial and Poisson variables. Sankhyā Indian J Stat Ser B 37(3):293–299

    MathSciNet  MATH  Google Scholar 

  • Meng XL (1997) The EM algorithm and medical studies: a historical linik. Stat Methods Med Res 6:3–23

    Article  Google Scholar 

  • Meng XL, Rubin DB (1993) Maximum likelihood estimation via the ECM algorithm: a general framework. Biometrika 80:267–278

    Article  MathSciNet  MATH  Google Scholar 

  • Moore PG (1954) A note on truncated Poisson distributions. Biometrics 10(3):402–406

    Article  MathSciNet  MATH  Google Scholar 

  • Patil GP (1963) Minimum variance unbiased estimation and certain problems of additive number theory. Ann MathStat 34(3):1050–1056

    MathSciNet  MATH  Google Scholar 

  • Plackett RL (1953) The truncated Poisson distribution. Biometrics 9(4):485–488

    Article  MathSciNet  Google Scholar 

  • Rider PR (1953) Truncated Poisson distributions. J Am Stat Assoc 48:826–830

    Article  MathSciNet  MATH  Google Scholar 

  • Rider PR (1955) Truncated binomial and negative-binomial distributions. J Am Stat Assoc 50(271):877–883

    Article  MathSciNet  MATH  Google Scholar 

  • Sampford MR (1955) The truncated negative binomial distribution. Biometrika 42:58–69

    Article  MathSciNet  MATH  Google Scholar 

  • Scollnik DPM (2006) On the intervened generalized Poisson distribution. Commun Stat Theory Methods 35:953–963

    Article  MathSciNet  MATH  Google Scholar 

  • Shanmugam R (1985) An intervened Poisson distribution and its medical application. Biometrics 41:1025–1029

    Article  MathSciNet  MATH  Google Scholar 

  • Singh J (1978) A characterization of positive Poisson distribution and its statistical application. SIAM J Appl Math 34(3):545–548

    Article  MathSciNet  MATH  Google Scholar 

  • Springael J, Van Nieuwenhuyse I (2006) On the sum of independent zero-truncated Poisson random variables. University of Antwerp, Faculty of Applied Economics

  • Tate RF, Goen RL (1958) Minimum variance unbiased estimation for the truncated Poisson distribution. Ann Math Stat 29(3):755–765

    Article  MathSciNet  MATH  Google Scholar 

  • Xie TL, Aickin M (1997) A truncated Poisson regression model with applications to occurrence of adenomatous polyps. Stat Med 16:1845–1857

    Article  Google Scholar 

  • Zhou H, Lange K (2010) MM algorithms for some discrete multivariate distributions. J Comput Graph Stat 19(3):645–665

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

Yin Liu’s research was fully supported by Grants (Nos. 11601524 and 61773401) from the National Natural Science Foundation of China and a Grant (31721811206) from the Young Teachers Innovation Project of Zhongnan University of Economics and Law. Guo-Liang Tian’s research was fully supported by a Grant from the National Natural Science Foundation of China (No. 11771199). The work of Man-Lai Tang was supported by the Research Fund of the Project (Grant No.UGC/FDS14/P06/17) and National Natural Science Foundation of China (Grant No. 11871124).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yin Liu.

Additional information

Publisher's Note

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

Electronic supplementary material

Appendices

Appendix A: A review on other commonly used univariate zero-truncated discrete distributions

1.1 Zero-truncated negative-binomial distribution

The pmf of the zero-truncated negative-binomial distribution is

$$\begin{aligned} {p}_{_{\tiny {\mathrm{T}}}}(y|r, \pi ) = \Pr (Y=y) = \frac{1}{1-(1-\pi )^r} \cdot \frac{\Gamma (y+r)}{y!\Gamma (r)} \pi ^{y}(1-\pi )^{r}, \quad y \geqslant 1. \end{aligned}$$
(A.1)

We denote it by \(Y \sim \text{ ZTNB }(r, \pi )\). The mean and variance of Y are given by

$$\begin{aligned} E(Y)= \frac{r \pi }{[1- (1-\pi )^r](1-\pi )}, \end{aligned}$$

and

$$\begin{aligned} \text{ Var }(Y) = \frac{r \pi [1 - (1-\pi )^r - r \pi (1-\pi )^r]}{(1-\pi )^2[1-(1-\pi )^r]^2}. \end{aligned}$$
(A.2)

Rider (1955) derived the MLE of \(\pi \) when r is known and When \(r>0\) is unknown, Sampford (1955) discussed the maximum likelihood estimates of both r and \(\pi \) at the same time and pointed out the slowness of the iteration process, which requires a special aid table. He then suggested an alternative method to obtain the solutions to the MLEs of the parameters by trial and error process. This iterative process starts with an inefficient estimate of r and then the corresponding \(\pi \) is obtained. By making use of the result from Bliss and Fisher (1953), Hartley (1958) proposed a method to obtain the MLEs of r and \(\pi \). The idea is to start the iteration process by choosing some computationally convenient “pivotal values” of r and calculate the corresponding MLE of \(\pi \). The final MLEs of r and \(\pi \) are solved by interpolation.

Ahuja (1971) derived the distribution of the sum of several i.i.d. ZTNB random variables by applying one of the results obtained by Patil (1963) for the generalized power series distribution. The explicit form of the distribution of Y was given as a linear combination of the incomplete beta functions.

1.2 Zero-truncated generalized negative-binomial distribution

By introducing an additional parameter \(\beta \) to the negative-binomial distribution using Lagrange’s expansion, Jain and Consul (1971) is the first paper to introduce the generalized negative-binomial distribution. Its pmf is given by

$$\begin{aligned} p(x|r, \pi , \beta ) = \frac{r\Gamma {(r+\beta x)}}{x!\Gamma {(r+\beta x-x+1)}}\pi ^x(1-\pi )^{r+\beta x-x}, \end{aligned}$$
(A.3)

for \(x=0,1, \ldots , \infty \), where \(r>0\), \(\pi \in (0, 1)\) and \(\beta =0\) or \(1\leqslant \beta \leqslant \pi ^{-1}\). We denote it by \(X \sim \text{ GNB }(r, \pi , \beta )\). It was shown that the mean and variance of X are given by (Jain and Consul 1971)

$$\begin{aligned} E(X) = \frac{r \pi }{1-\beta \pi } \quad \text{ and } \quad \text{ Var }(X) = \frac{r \pi (1-\pi )}{(1-\beta \pi )^3}, \end{aligned}$$
(A.4)

which exist for \(\beta =0\) or \(\beta \in [1, \pi ^{-1})\), see Consul and Famoye (2006, p.191). In particular, the \(\text{ GNB }(r, \pi , \beta )\) distribution reduces to the \(\text{ Binomial }(r, \pi )\) when \(\beta =0\) and reduces to the \(\text{ NBinomial }(r, \pi )\) when \(\beta =1\).

Thus, the pmf of a zero-truncated generalized negative-binomial distribution is

$$\begin{aligned} {p}_{_{\tiny {\mathrm{T}}}}(y|r, \pi , \beta ) = \frac{r\Gamma {(r+\beta y)}}{y!\Gamma {(r+\beta y-y+1)}} \cdot \frac{\pi ^y(1-\pi )^{r+\beta y-y}}{1 - (1-\pi )^r}, \qquad y \geqslant 1. \end{aligned}$$
(A.5)

However, to our best knowledge, we did not find any papers on estimating the parameters in a zero-truncated generalized negative-binomial distribution. Medhi (1975) derived the distribution of the convolution of left-truncated generalized negative-binomial variables by making using of the same property of the pmf.

1.3 Zero-truncated generalized Poisson distribution

By considering a limiting form of the generalized negative binomial distribution, Consul and Jain (1973) proposed the generalized Poisson distribution. A non-negative integer valued random variable X is said to have a generalized Poisson (GP) distribution with parameters \(\lambda >0\) and \(\theta \), if its pmf is given by (Consul and Jain 1973; Consul 1989, p.4)

$$\begin{aligned} p(x|\lambda , \theta ) = \left\{ \begin{array}{ll} \displaystyle \frac{\lambda (\lambda +\theta x)^{x-1}\text{ e }^{-\lambda -\theta x}}{x!}, &{}\quad x=0, 1, \ldots , \infty , \\ 0, &{}\quad \text{ for } x > q, \text{ when } \theta < 0, \end{array}\right. \end{aligned}$$
(A.6)

where \(\max (-1, -\lambda /q) < \theta \leqslant 1\) and \(q \,(\geqslant 4)\) is the largest positive integer for which \(\lambda + \theta q > 0\) when \(\theta < 0\). We denote it by \(X \sim \text{ GP }(\lambda , \theta )\). It was shown that the mean and variance of X are given by (Consul 1989, p.14)

$$\begin{aligned} E(X) = \frac{\lambda }{1-\theta } \quad \text{ and } \quad \text{ Var }(X) = \frac{\lambda }{(1-\theta )^3}, \quad \text{ where } \;\, \theta < 1. \end{aligned}$$
(A.7)

The \(\text{ GP }(\lambda , \theta )\) distribution reduces to the \(\text{ Poisson }(\lambda )\) when \(\theta =0\), and it has the twin properties of over-dispersion when \(\theta >0\) and under-dispersion when \(\theta <0\). The most frequently used version of the GP distribution assumes that \(\lambda >0\) and \(\theta \in [0, 1)\).

Thus, the pmf of a zero-truncated generalized Poisson distribution is

$$\begin{aligned} {p}_{_{\tiny {\mathrm{T}}}}(y|\lambda , \theta ) = \frac{\lambda (\lambda +\theta y)^{y-1}\text{ e }^{-\lambda -\theta y}}{(1 - \text{ e }^{-\lambda })y!}, \qquad y = 1, 2, \ldots , \infty . \end{aligned}$$
(A.8)

We denote it by \(Y \sim \text{ ZTGP }(\lambda , \theta )\). The distribution of the convolution of zero-truncated generalized Poisson variables was firstly derived by Medhi (1975) by considering a general problem and later, Consul (1989, p. 70) provided a simple and independent proof for it.

Appendix B: A general strategy for establishing an EM algorithm for truncated-data problems (Meng 1997)

Let \(f(x|\varvec{\theta })\) denote the untruncated pdf, where \(x \in {\mathbb {A}}\cup {\bar{{\mathbb {A}}}}\) and \({\mathbb {A}}\) is a region such that value outside \({\mathbb {A}}\) will be truncated (e.g., \({\mathbb {A}}= \{1,2,\ldots \}\) and \({\bar{{\mathbb {A}}}}=\{0\}\) for the ZTP distribution). Therefore, the truncated pdf is \(f(x|\varvec{\theta })/\Pr ({\mathbb {A}}|\varvec{\theta })\) with

$$\begin{aligned} \Pr ({\mathbb {A}}|\varvec{\theta }) = \int _{{\mathbb {A}}} f(x|\varvec{\theta }) \; \text{ d }x. \end{aligned}$$

Let \(X_1,\ldots ,X_n {\mathop {\sim }\limits ^{\mathrm{iid}}}f(x|\varvec{\theta })/\Pr ({\mathbb {A}}|\varvec{\theta })\), where \(x \in {\mathbb {A}}\). The likelihood function of \(\varvec{\theta }\) based on the observed data \(Y_{\mathrm{obs}} = \{x_1,\ldots ,x_n\}\) is

$$\begin{aligned} L(\varvec{\theta }|Y_{\mathrm{obs}}) = \frac{\prod _{i=1}^n f(x_i|\varvec{\theta })}{[\Pr ({\mathbb {A}}|\varvec{\theta })]^n}. \end{aligned}$$

Augmenting the \(Y_{\mathrm{obs}}\) to include the sample values that were truncated out and the number of such values: \(Y_{\mathrm{mis}} = \{x_{n+1}, \ldots , x_{n+m}; m\}\) with m being a random variable, we obtain the complete-data likelihood function of \(\varvec{\theta }\), which is given by

$$\begin{aligned} L(\varvec{\theta }|Y_{\mathrm{obs}}, Y_{\mathrm{mis}}) \propto f(x_1,\ldots ,x_{n+m}, m|\varvec{\theta }) = \prod _{i=1}^{n+m} f(x_i|\varvec{\theta }). \end{aligned}$$

Note that the joint distribution \((Y_{\mathrm{obs}}, Y_{\mathrm{mis}})\) can be rewritten as

$$\begin{aligned}&f(x_1,\ldots ,x_{n+m}, m|\varvec{\theta }) \\&\quad = f(x_1,\ldots ,x_{n+m}|\varvec{\theta }) \times \Pr (m|x_1,\ldots ,x_{n+m},\varvec{\theta }) \\&\quad = \frac{\prod _{i=1}^{n+m} f(x_i|\varvec{\theta })}{[\Pr ({\mathbb {A}}|\varvec{\theta })]^n[1-\Pr ({\mathbb {A}}|\varvec{\theta })]^m} \times \Pr (m|x_1,\ldots ,x_{n+m},\varvec{\theta }). \end{aligned}$$

Therefore,

$$\begin{aligned} \Pr (m|x_1,\ldots ,x_{n+m},\varvec{\theta }) = [\Pr ({\mathbb {A}}|\varvec{\theta })]^n[1-\Pr ({\mathbb {A}}|\varvec{\theta })]^m \end{aligned}$$

and

$$\begin{aligned} \Pr (m|Y_{\mathrm{obs}},\varvec{\theta }) = \bigg (\begin{array}{c} m+n-1 \\ m \end{array} \bigg ) [\Pr ({\mathbb {A}}|\varvec{\theta })]^n[1-\Pr ({\mathbb {A}}|\varvec{\theta })]^m, \quad m \geqslant 0, \end{aligned}$$

which is a negative binomial distribution. The E-step is to replace m in the expression of the MLE of \(\varvec{\theta }\) via the M-step by its conditional expectation

$$\begin{aligned} E(m|Y_{\mathrm{obs}},\varvec{\theta }) = n \frac{1-\Pr ({\mathbb {A}}|\varvec{\theta })}{\Pr ({\mathbb {A}}|\varvec{\theta })}. \end{aligned}$$
(B.1)

Appendix C: Derivation of EM-type iterations for a class of zero-truncated discrete distributions

Example C.1

(Zero-truncated Poisson). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTP }(\lambda )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_1, \ldots , y_n\}\). Then, the observed-data likelihood function for \(\lambda \) is given by

$$\begin{aligned} L(\lambda |Y_{\mathrm{obs}}) = \prod _{i=1}^n \frac{\lambda ^{y_i}\text{ e }^{-\lambda }}{(1-\text{ e }^{-\lambda })y_i!} \propto \lambda ^{n{\bar{y}}} \text{ e }^{-n\lambda }(1-\text{ e }^{-\lambda })^{-n}, \end{aligned}$$

where \({\bar{y}}= (1/n)\sum _{i=1}^n y_i\) is a sufficient statistic of \(\lambda \). In (3.4), let \(\varvec{\theta }=\lambda \) and \(p(x|\lambda ) = \lambda ^x \text{ e }^{-\lambda }/x!\). Then, the complete-data MLE of \(\lambda \) given by

$$\begin{aligned} {\hat{\lambda }}= \frac{\sum _{i=1}^n z_iy_i}{n}. \end{aligned}$$
(C.1)

The E-step is to replace \(\{z_i\}_{i=1}^n\) in (C.1) by their conditional expectations:

$$\begin{aligned} E(Z_i|Y_{\mathrm{obs}}, \lambda ) = 1-\text{ e }^{-\lambda }, \quad i=1,\ldots , n. \end{aligned}$$
(C.2)

By combining (C.1) with (C.2), we have the following EM iterations:

$$\begin{aligned} \lambda ^{(t+1)} = {\bar{y}}\Big (1-\text{ e }^{-\lambda ^{(t)}} \Big ), \quad t=0, 1, \ldots , \end{aligned}$$
(C.3)

where \(\lambda ^{(0)}\) is the initial value of the MLE \({\hat{\lambda }}\). \(\square \)

Example C.2

(Zero-truncated binomial). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTB }(m,\pi )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_1, \ldots , y_n\}\). Then, the observed-data likelihood function is given by

$$\begin{aligned} L(\pi |Y_{\mathrm{obs}}) = \prod _{i=1}^n \frac{ {m \atopwithdelims ()y_i} \pi ^{y_i}(1-\pi )^{m-y_i}}{1-(1-\pi )^m} \propto [1-(1-\pi )^m]^{-n} \pi ^{n{\bar{y}}}(1-\pi )^{n(m-{\bar{y}})}, \end{aligned}$$

where \({\bar{y}}= (1/n)\sum _{i=1}^n y_i\) is a sufficient statistic of \(\pi \). In (3.4), let \(\varvec{\theta }=\pi \) and \(p(x|\pi ) = {m \atopwithdelims ()x} \pi ^x(1-\pi )^{m-x}\). Then, the complete-data MLE of \(\pi \) given by

$$\begin{aligned} {\hat{\pi }} = \frac{\sum _{i=1}^n z_iy_i}{nm}. \end{aligned}$$
(C.4)

The E-step is to replace \(\{z_i\}_{i=1}^n\) in (C.4) by their conditional expectations:

$$\begin{aligned} E(Z_i|Y_{\mathrm{obs}}, \pi ) = 1-(1-\pi )^m, \quad i=1,\ldots , n. \end{aligned}$$
(C.5)

By combining (C.4) with (C.5), we have the following EM iterations:

$$\begin{aligned} \pi ^{(t+1)} = \frac{[1-(1-\pi ^{(t)})^m]{\bar{y}}}{m}, \quad t=0, 1, \ldots , \end{aligned}$$
(C.6)

where \(\pi ^{(0)}\) is the initial value of the MLE \({\hat{\pi }}\). \(\square \)

Example C.3

(Zero-truncated negative-binomial). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTNB }(r,\pi )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_1, \ldots , y_n\}\), where \(r>0\) is assumed to be known. Then, the observed-data likelihood function is

$$\begin{aligned} L(\pi |Y_{\mathrm{obs}}) = \prod _{i=1}^n \left[ \frac{\Gamma (y_i+r)}{y_i!\Gamma (r)} \cdot \frac{\pi ^{y_i}(1-\pi )^r }{1-(1-\pi )^r} \right] \propto [1-(1-\pi )^r]^{-n}\pi ^{n{\bar{y}}}(1-\pi )^{nr}, \end{aligned}$$

where \({\bar{y}}= (1/n)\sum _{i=1}^n y_i\) is a sufficient statistic of \(\pi \). In (3.4), let \(\varvec{\theta }=\pi \) and \(p(x|\pi ) = \frac{\Gamma (x+r)}{x!\Gamma (r)} \pi ^x(1-\pi )^{r}\). Then, the complete-data MLE of \(\pi \) given by

$$\begin{aligned} {\hat{\pi }} = \frac{\sum _{i=1}^n z_iy_i}{\sum _{i=1}^n z_iy_i+nr}. \end{aligned}$$
(C.7)

The E-step is to replace \(\{z_i\}_{i=1}^n\) in (C.7) by their conditional expectations:

$$\begin{aligned} E(Z_i|Y_{\mathrm{obs}}, \pi ) = 1-(1-\pi )^r, \quad i=1,\ldots , n. \end{aligned}$$
(C.8)

By combining (C.7) with (C.8), we have the following EM iterations:

$$\begin{aligned} \pi ^{(t+1)} = \frac{[1-(1-\pi ^{(t)})^r]{\bar{y}}}{[1-(1-\pi ^{(t)})^r]{\bar{y}}+r}, \quad t=0, 1, \ldots , \end{aligned}$$
(C.9)

where \(\pi ^{(0)}\) is the initial value of the MLE \({\hat{\pi }}\). \(\square \)

Example C.4

(Zero-truncated generalized negative-binomial). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTGNB }(r,\pi , \beta )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_1, \ldots , y_n\}\), where \(r>0\) is assumed to be known. Then, the observed-data likelihood function of \((\pi , \beta )\) is

$$\begin{aligned} L(\pi , \beta |Y_{\mathrm{obs}})= & {} \prod _{i=1}^n \left[ \frac{r\Gamma {(r+\beta y_i)}}{y_i!\Gamma {(r+\beta y_i-y_i+1)}} \cdot \frac{\pi ^{y_i}(1-\pi )^{r+\beta y_i-y_i}}{1 - (1-\pi )^r} \right] \\\propto & {} \left[ \prod _{i=1}^n\frac{\Gamma {(r+\beta y_i)}}{\Gamma {(r+\beta y_i-y_i+1)}}\right] \cdot [1-(1-\pi )^r]^{-n}\pi ^{n{\bar{y}}}(1-\pi )^{nr+(\beta -1)n{\bar{y}}}, \end{aligned}$$

where \({\bar{y}}= (1/n)\sum _{i=1}^n y_i\). From (3.4), the complete-data log-likelihood is proportional to

$$\begin{aligned}&\sum _{i=1}^n \Big [\log \Gamma {(r+\beta x_i)} - \log \Gamma {(r+\beta x_i-x_i+1)} \Big ] \\&\quad +\,n{\bar{x}}\log (\pi ) + [nr+(\beta -1)n{\bar{x}}]\log (1-\pi ). \end{aligned}$$

Therefore, given \(\pi =\pi ^{(t)}\), the complete-data MLE \(\beta ^{(t+1)}\) of \(\beta \) is the root of the equation

$$\begin{aligned}&h(\beta | \pi ^{(t)}) \nonumber \\&\quad = \sum _{i=1}^n \left[ \frac{\Gamma '{(r+\beta z_iy_i)z_iy_i}}{\Gamma {(r+\beta z_iy_i)}} - \frac{\Gamma '{(r+\beta z_iy_i-z_iy_i+1)z_iy_i}}{\Gamma {(r+\beta z_iy_i-z_iy_i+1)}}\right] \nonumber \\&\qquad +\, \left( \sum _{i=1}^n z_iy_i \right) \log [1- \pi ^{(t)}] = 0, \end{aligned}$$
(C.10)

and given \(\beta = \beta ^{(t+1)}\), the complete-data MLE \(\pi ^{(t+1)}\) of \(\pi \) is

$$\begin{aligned} \pi = \frac{\sum _{i=1}^n z_iy_i}{\beta ^{(t+1)} \sum _{i=1}^n z_iy_i + nr}. \end{aligned}$$
(C.11)

From (3.6) and (3.5), the E-step is to replace \(\{z_i\}_{i=1}^n\) and \(\{g_i(z_i)\}_{i=1}^n\) in (C.10) and (C.11) by their conditional expectations

$$\begin{aligned} E[Z_i|Y_{\mathrm{obs}}, \pi ^{(t)}, \beta ^{(t)}]= & {} 1-(1 - \pi ^{(t)})^r, \quad i=1,\ldots , n, \quad \text{ and } \quad \nonumber \\ E[g_i(Z_i)|Y_{\mathrm{obs}}, \pi ^{(t)}, \beta ^{(t)}]= & {} g_i(0)(1-\pi ^{(t)})^r + g_i(1)[1-(1 - \pi ^{(t)})^r],\nonumber \\ \end{aligned}$$
(C.12)

respectively, resulting in the following ECM iterations:

  1. CM-Step 1:

    Given \(\pi =\pi ^{(t)}\), calculate \(\beta ^{(t+1)}\), which is the solution to the following equation

    $$\begin{aligned} \sum _{i=1}^n y_i\left[ \frac{\Gamma '{(r+\beta y_i)}}{\Gamma {(r+\beta y_i)}} - \frac{\Gamma '{(r+\beta y_i-y_i+1)}}{\Gamma {(r+\beta y_i- y_i+1)}} \right] \, +\, n{\bar{y}}\log [1-\pi ^{(t)}] = 0. \quad \end{aligned}$$
    (C.13)
  2. CM-Step 2:

    Given \(\beta = \beta ^{(t+1)}\), calculate

    $$\begin{aligned} \pi ^{(t+1)} = \frac{{\bar{y}}[1-(1-\pi ^{(t)})^r ]}{\beta ^{(t+1)}{\bar{y}}[1-(1-\pi ^{(t)})^r] + r}, \quad t=0, 1, \ldots , \end{aligned}$$
    (C.14)

    where \(\pi ^{(0)}\) is the initial value of the MLE of \({\hat{\pi }}\).

\(\square \)

Example C.5

(Zero-truncated generalized Poisson). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTGP }(\lambda ,\theta )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_1, \ldots , y_n\}\). Then, the observed-data likelihood function of \((\lambda , \theta )\) is

$$\begin{aligned} L(\lambda , \theta |Y_{\mathrm{obs}})= & {} \prod _{i=1}^n \frac{\lambda (\lambda +\theta y_i)^{y_i-1}\text{ e }^{-\lambda -\theta y_i}}{(1 - \text{ e }^{-\lambda })y_i!} \propto \lambda ^n\\&\times \left[ \prod _{i=1}^n(\lambda + \theta y_i)^{y_i - 1} \right] \text{ e }^{-n\lambda - \theta n{\bar{y}}} (1 - \text{ e }^{-\lambda } )^{-n}, \end{aligned}$$

where \({\bar{y}}= (1/n)\sum _{i=1}^n y_i\). From (3.4), the complete-data log-likelihood is proportional to

$$\begin{aligned} n \log (\lambda ) - n\lambda - \theta n {\bar{x}}+ \sum _{i=1}^n (x_i-1) \log (\lambda + \theta x_i), \end{aligned}$$

Therefore, given \(\lambda =\lambda ^{(t)}\), the complete-data MLE \(\theta ^{(t+1)}\) of \(\theta \) is the root of the equation

$$\begin{aligned} H(\theta | \lambda ^{(t)}) = \sum _{i=1}^n \frac{(x_i - 1)x_i}{\lambda + \theta x_i} - n {\bar{x}}= \sum _{i=1}^n \frac{z_i^2y_i^2 - z_iy_i}{\lambda ^{(t)} + \theta z_iy_i} - \sum _{i=1}^n z_iy_i = 0, \end{aligned}$$
(C.15)

and given \(\theta = \theta ^{(t+1)}\), the complete-data MLE \(\lambda ^{(t+1)}\) of \(\lambda \) is given by (Consul 1989, p.102)

$$\begin{aligned} \lambda = {\bar{x}}(1-\theta ) = \frac{\sum _{i=1}^n z_iy_i}{n} (1 - \theta ^{(t+1)}), \end{aligned}$$
(C.16)

where \({\bar{x}}= (1/n)\sum _{i=1}^n z_iy_i\). From (3.6) and (3.5), the E-step is to replace \(\{z_i\}_{i=1}^n\) and \(\{g_i(z_i)^2\}_{i=1}^n\) in (C.15) and (C.16) by their conditional expectations

$$\begin{aligned} E[Z_i|Y_{\mathrm{obs}}, \lambda ^{(t)}, \theta ^{(t)}]= & {} 1-\text{ e }^{-\lambda ^{(t)}}, \quad i=1,\ldots , n, \quad \text{ and } \quad \nonumber \\ E[g_i(Z_i)|Y_{\mathrm{obs}}, \lambda ^{(t)}, \theta ^{(t)}]= & {} g_i(0)\text{ e }^{-\lambda ^{(t)}} + g_i(1)[1-\text{ e }^{-\lambda ^{(t)}}], \end{aligned}$$
(C.17)

respectively, resulting in the following ECM iterations:

  1. CM-Step 1:

    Given \(\lambda =\lambda ^{(t)}\), calculate \(\theta ^{(t+1)}\), which is the solution to the following equation

    $$\begin{aligned} \sum _{i=1}^n \frac{y_i^2 - y_i }{\lambda ^{(t)} + \theta y_i} - n {\bar{y}}= 0. \end{aligned}$$
    (C.18)
  2. CM-Step 2:

    Given \(\theta = \theta ^{(t+1)}\), calculate

    $$\begin{aligned} \lambda ^{(t+1)} = {\bar{y}}\Big (1-\text{ e }^{-\lambda ^{(t)}} \Big )(1-\theta ^{(t+1)}), \quad t=0, 1, \ldots , \end{aligned}$$
    (C.19)

    where \(\lambda ^{(0)}\) is the initial value of the MLE of \({\hat{\lambda }}\).

\(\square \)

Appendix D: Derivation of MM iterations for a class of zero-truncated discrete distributions

Example D.1

(Zero-truncated Poisson). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTP }(\lambda )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_i\}_{i=1}^n\). The corresponding parent random variable \(X \sim \text{ Poisson }(\lambda )\) with the pmf \(p(x|\lambda ) = \lambda ^x \text{ e }^{-\lambda }/x!\). In (4.4), let \(\varvec{\theta }=\lambda \), we have

$$\begin{aligned} Q(\lambda |\lambda ^{(t)})= & {} \sum _{i=1}^n \left[ \log \left( \frac{\lambda ^{y_i} \text{ e }^{-\lambda }}{y_i!} \right) + \frac{\text{ e }^{-\lambda ^{(t)}} }{1-\text{ e }^{-\lambda ^{(t)}}} \log (\text{ e }^{-\lambda }) \right] \\= & {} n{\bar{y}}\log (\lambda ) - \frac{n\lambda }{1-\text{ e }^{-\lambda ^{(t)}}} + c. \end{aligned}$$

Setting \(\text{ d }Q(\lambda |\lambda ^{(t)})/\text{ d }\lambda =0\), we obtain an MM iteration, which is identical to (C.3).

\(\square \)

Example D.2

(Zero-truncated binomial). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTB }(m,\pi )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_i\}_{i=1}^n\). The corresponding parent random variable \(X \sim \text{ Binomial }(m, \pi )\) with the pmf \(p(x|\pi ) = {m \atopwithdelims ()x} \pi ^x (1-\pi )^{m-x}\). In (4.4), let \(\varvec{\theta }=\pi \), we have

$$\begin{aligned} Q(\pi |\pi ^{(t)})= & {} \sum _{i=1}^n \left\{ \log \left[ {m \atopwithdelims ()y_i} \pi ^{y_i} (1-\pi )^{m-y_i} \right] + \frac{(1-\pi ^{(t)})^m }{1-(1-\pi ^{(t)})^m } \log [(1-\pi )^m] \right\} \\= & {} n{\bar{y}}\log (\pi ) + n\left[ m-{\bar{y}}+ \frac{m(1-\pi ^{(t)})^m}{1-(1-\pi ^{(t)})^m} \right] \log (1-\pi ) + c. \end{aligned}$$

Setting \(\text{ d }Q(\pi |\pi ^{(t)})/\text{ d }\pi =0\), we obtain an MM iteration, which is identical to (C.6). \(\square \)

Example D.3

(Zero-truncated negative-binomial). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTNB }(r,\pi )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_i\}_{i=1}^n\), where \(r>0\) is assumed to be known. The corresponding parent random variable \(X \sim \text{ NBinomial }(r, \pi )\) with the pmf \(p(x|\pi ) = \frac{\Gamma (x+r)}{x!\Gamma (r)} \pi ^x (1-\pi )^r\). In (4.4), let \(\varvec{\theta }=\pi \), we have

$$\begin{aligned} Q(\pi |\pi ^{(t)})= & {} \sum _{i=1}^n \left\{ \log \left[ \frac{\Gamma (y_i+r)}{y_i!\Gamma (r)} \pi ^{y_i} (1-\pi )^r \right] + \frac{(1-\pi ^{(t)})^r }{1-(1-\pi ^{(t)})^r } \log [(1-\pi )^r] \right\} \\= & {} n{\bar{y}}\log (\pi ) + \frac{nr}{1-(1-\pi ^{(t)})^r} \log (1-\pi ) + c. \end{aligned}$$

Setting \(\text{ d }Q(\pi |\pi ^{(t)})/\text{ d }\pi =0\), we obtain an MM iteration, which is identical to (C.9). \(\square \)

Example D.4

(Zero-truncated generalized negative-binomial). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTGNB }(r,\pi , \beta )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_i\}_{i=1}^n\), where \(r>0\) is assumed to be known. The corresponding parent random variable \(X \sim \text{ GNB }(r, \pi , \beta )\) with pmf given by (A.3). In (4.4), let \(\varvec{\theta }=(\pi , \beta )\), we have

$$\begin{aligned}&Q(\pi , \beta | \pi ^{(t)}, \beta ^{(t)}) \\&\quad = \sum _{i=1}^n \left\{ \log \left[ \frac{r\Gamma (r+ \beta y_i) \cdot \pi ^{y_i} (1-\pi )^{r+\beta y_i-y_i} }{y_i!\Gamma (r + \beta y_i - y_i +1)} \right] + \frac{(1-\pi ^{(t)})^r }{1-(1-\pi ^{(t)})^r } \log [(1-\pi )^r] \right\} \\&\quad = \sum _{i=1}^n \Big \{\log [\Gamma {(r+\beta y_i)}]-\log [\Gamma {(r+\beta y_i-y_i+1)}] \Big \} \nonumber \\&\qquad +\, n \left\{ {\bar{y}}\log (\pi )+ \left[ r+(\beta -1){\bar{y}}+ \frac{r(1-\pi ^{(t)})^r }{1-(1-\pi ^{(t)})^r } \right] \log (1-\pi ) \right\} + c. \end{aligned}$$

By setting

$$\begin{aligned} \left\{ \begin{array}{l} 0 = \displaystyle \frac{\partial Q(\pi , \beta | \pi ^{(t)}, \beta ^{(t)})}{\partial \pi } = \frac{n {\bar{y}}}{\pi } - \frac{n \left[ r+(\beta -1){\bar{y}}+ \displaystyle \frac{r(1-\pi ^{(t)})^r }{1-(1-\pi ^{(t)})^r } \right] }{1-\pi }, \\ 0 = \displaystyle \frac{\partial Q(\pi , \beta | \pi ^{(t)}, \beta ^{(t)})}{\partial \beta } = \sum _{i=1}^n y_i \left[ \frac{\Gamma '{(r+\beta y_i) }}{\Gamma {(r+\beta y_i)}} - \frac{\Gamma '{(r+\beta y_i-y_i+1)}}{\Gamma {(r+\beta y_i- y_i+1)}} \right] + n{\bar{y}}\log (1-\pi ), \end{array} \right. \end{aligned}$$

we have the following MM iterations:

  1. Step 0:

    Let \(\pi ^{(0)}\) be the initial value of the MLE of \({\hat{\pi }}\) and set \(t=0\).

  2. Step 1:

    Given \(\pi =\pi ^{(t)}\), calculate \(\beta ^{(t+1)}\), which is the solution to the following equation

    $$\begin{aligned} \sum _{i=1}^n y_i\left[ \frac{\Gamma '{(r+\beta y_i)}}{\Gamma {(r+\beta y_i)}} - \frac{\Gamma '{(r+\beta y_i-y_i+1)}}{\Gamma {(r+\beta y_i- y_i+1)}} \right] + n{\bar{y}}\log [1-\pi ^{(t)}] = 0. \quad \end{aligned}$$
    (D.1)
  3. Step 2:

    Given \(\beta = \beta ^{(t+1)}\), calculate

    $$\begin{aligned} \pi ^{(t+1)} = \frac{{\bar{y}}[1-(1-\pi ^{(t)})^r ]}{\beta ^{(t+1)}{\bar{y}}[1-(1-\pi ^{(t)})^r] + r}. \end{aligned}$$
    (D.2)

We can see that (D.1) is identical to (C.13) and (D.2) is identical to (C.14). \(\square \)

Example D.5

(Zero-truncated generalized Poisson). Let \(Y_1, \ldots , Y_n {\mathop {\sim }\limits ^{\mathrm{iid}}}\text{ ZTGP }(\lambda ,\theta )\) and the observed data be \(Y_{\mathrm{obs}}=\{y_i\}_{i=1}^n\). The corresponding parent random variable \(X \sim \text{ GP }(\lambda , \theta )\) with pmf given by (A.6). In (4.4), let \(\varvec{\theta }=(\lambda , \theta )\), we have

$$\begin{aligned}&Q(\lambda , \theta | \lambda ^{(t)}, \theta ^{(t)}) \\&\quad = \sum _{i=1}^n \left\{ \log \left[ \frac{\lambda (\lambda +\theta y_i)^{y_i-1} \text{ e }^{-\lambda -\theta y_i}}{y_i!} \right] + \frac{\text{ e }^{-\lambda ^{(t)}} }{1- \text{ e }^{-\lambda ^{(t)}} } \log (\text{ e }^{-\lambda }) \right\} \\&\quad = \sum _{i=1}^n (y_i - 1)\log (\lambda + \theta y_i) + n\left[ \log (\lambda )- \lambda -\theta {\bar{y}}- \frac{\lambda \text{ e }^{-\lambda ^{(t)}} }{1- \text{ e }^{-\lambda ^{(t)}} } \right] + c. \end{aligned}$$

By setting

$$\begin{aligned} \left\{ \begin{array}{l} 0 = \displaystyle \frac{\partial Q(\lambda , \theta | \lambda ^{(t)}, \theta ^{(t)})}{\partial \lambda } = \sum _{i=1}^n \frac{y_i - 1}{\lambda + \theta y_i} + n \left( \frac{1}{\lambda } - \frac{1}{1- \text{ e }^{-\lambda ^{(t)}}} \right) , \\ 0 = \displaystyle \frac{\partial Q(\lambda , \theta | \lambda ^{(t)}, \theta ^{(t)})}{\partial \theta } = \sum _{i=1}^n \frac{y_i^2 - y_i}{\lambda + \theta y_i} - n {\bar{y}}, \end{array} \right. \end{aligned}$$

we have the following MM iterations:

  1. Step 0:

    Let \(\lambda ^{(0)}\) be the initial value of the MLE of \({\hat{\lambda }}\) and set \(t=0\).

  2. Step 1:

    Given \(\lambda =\lambda ^{(t)}\), calculate \(\theta ^{(t+1)}\), which is the solution to the following equation

    $$\begin{aligned} \sum _{i=1}^n \frac{y_i^2 - y_i}{\lambda ^{(t)} + \theta y_i} - n {\bar{y}}= 0. \quad \end{aligned}$$
    (D.3)
  3. Step 2:

    Given \(\theta = \theta ^{(t+1)}\), calculate

    $$\begin{aligned} \lambda ^{(t+1)} = {\bar{y}}(1 - \text{ e }^{-\lambda ^{(t)}})(1 - \theta ^{(t+1)}), \end{aligned}$$
    (D.4)

    which can be obtained by first multiplying \(\lambda \) and \(\theta \) to the above two partial differential equations, respectively, and then adding them together.

We can see that (D.3) is identical to (C.18) and (D.4) is identical to (C.19). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Tian, GL., Ding, X., Liu, Y. et al. Some new statistical methods for a class of zero-truncated discrete distributions with applications. Comput Stat 34, 1393–1426 (2019). https://doi.org/10.1007/s00180-018-00860-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00180-018-00860-0

Keywords

Navigation