5  Exponential dispersion models

Warning

This chapter is still a draft and minor changes are made without notice.

This chapter introduces a class of response distributions more general than the normal distribution, which will form the basis of a considerable extension of the linear model in Chapter 6. Among the possible alternatives to the normal distribution are discrete distributions like the binomial and the Poisson distribution and continuous distributions like the Gamma distribution.

5.1 A motivating example

To motivate the introduction of the exponential dispersion models, we consider an example.

Example 5.1 (Bernoulli model) For a binary response variable \(Y \in \{0, 1\}\) with \(p = P(Y = 1)\) the log-odds parameter is \[ \theta \coloneqq \log \frac{P(Y = 1)}{P(Y = 0)} = \log \frac{p}{1 - p}. \] Rearranging this equation gives \[ p = \frac{e^{\theta}}{1 + e^{\theta}} \] and \[ 1 - p = \frac{1}{1 + e^{\theta}}. \] From this, \[ P(Y = y) = p^y(1-p)^{(1-y)} = \frac{e^{y\theta}}{1 + e^{\theta}} \] for \(y \in \{0, 1\}.\) The probability mass function thus takes the form \[ \frac{1}{\phi(\theta)}e^{y \theta} \] with \(\phi(\theta) = 1 + e^{\theta}.\) By direct computations we may observe that \[ E(Y) = p = \frac{\phi'(\theta)}{\phi(\theta)} = (\log \phi(\theta))' \] and that \[ V(Y) = p(1-p) = \left(\frac{\phi'(\theta)}{\phi(\theta)}\right)'. \] That is, the mean can be found by differentiation of \(\log \phi,\) and the variance is the derivative of the mean as a function of \(\theta.\)

The identities relating the mean, variance and log-odds parameter \(\theta\) for the Bernoulli model above may seem coincidential for this particular model. In fact, they are examples from the general theory of exponential families. The Bernoulli model can be parametrized by the mean value parameter (the probability \(p\)) or the log-odds parameter \(\theta.\) The function \[ \theta \mapsto \frac{e^{\theta}}{1 + e^{\theta}} = p \] is a one-to-one correspondance mapping \(\mathbb{R}\) onto \((0,1).\) The inverse of this function is the logit function \[ \mathrm{logit}(p) = \log \frac{p}{1 - p}. \] Moreover, the variance can be expressed in terms of the mean as \(p(1-p)\) or in terms of the log-odds parameter. The latter is obtainable from the function \(\phi\) via differentiation.

As mentioned, the Bernoulli model is an example of a univariate exponential family. The normal distribution with fixed variance is another example. This chapter investigates the class of exponential families on the real line with the extension that a dispersion parameter is also allowed. By introducing a dispersion parameter the resulting class of exponential dispersion models includes a number of two-parameter families of distributions on the real line such as the normal distribution and the Gamma distribution.

5.2 Exponential dispersion models

An exponential dispersion model is a two-parameter family of so-called exponential dispersion distributions determined by an exponential family and a dispersion parameter. The theory of exponential dispersion models provides a framework for a number of widely applied univariate distributions. The theory is interesting in itself, but it is treated here first of all to provide the theoretical foundation for generalized linear regression models as treated in Chapter 6. In this section the models and their associated terminology are introduced, and the general relation between mean and variance is established. In subsequent chapters the statistical theory for generalized linear models is developed based on the framework of exponential dispersion models.

Every model is determined by a single \(\sigma\)-finite measure on the real line. With \(\nu\) a \(\sigma\)-finite measure on \(\mathbb{R}\) we define \(\phi : \mathbb{R} \to [0,\infty]\) by \[\begin{equation} \label{eq:psidef} \phi(\theta) = \int e^{\theta y} \, \nu(\mathrm{d} y), \end{equation}\] and we let \[ I = \{ \theta \in \mathbb{R} \mid \phi(\theta) < \infty \}^{\circ} \] denote the interior of the set of \(\theta\)-s for which \(\phi(\theta) < \infty.\)

Note that it is possible that \(I = \emptyset\) – take, for instance, \(\nu\) to be Lebesgue measure. If \(\nu\) is a finite measure then \(\phi(0) < \infty,\) but it is still possible that \(\phi(\theta) = \infty\) for all \(\theta \neq 0,\) which results in \(I = \emptyset.\) The case where \(I\) is empty is not of any relevance. There are two other special situations that are not relevant either. If \(\nu\) is the zero measure, \(\phi(\theta) = 0,\) and if \(\nu\) is a one-point measure, that is, \(\nu = c \delta_{y}\) for \(c \in (0, \infty)\) and \(\delta_y\) the Dirac measure in \(y,\) then \(\phi(\theta) = c e^{\theta y}.\) Neither of these two cases will be of any interest, and they result in pathological problems that we want to avoid. We will therefore make the following regularity assumptions about \(\nu\) throughout.

  • The measure \(\nu\) is not the zero measure, nor is it a one-point measure.
  • The open set \(I\) is non-empty.

By the assumption that \(\nu\) is not the zero measure, it follows that \(\phi(\theta) > 0\) for \(\theta \in I.\) This makes the following definition possible.

Definition 5.1 The exponential family with structure measure \(\nu\) is the one-parameter family of probability measures, \(\rho_{\theta}\) for \(\theta \in I,\) defined by \[ \frac{\mathrm{d} \rho_\theta}{\mathrm{d} \nu}(y) = \frac{1}{\phi(\theta)}e^{\theta y}. \] The parameter \(\theta\) is called the canonical parameter.

Note that the exponential family is determined completely by the choice of the structure measure \(\nu.\) Introducing \(\kappa(\theta) = \log \phi(\theta)\) we have that \[ \frac{\mathrm{d} \rho_\theta}{\mathrm{d} \nu}(y) = e^{\theta y - \kappa(\theta)} \] for \(\theta \in I.\) The function \(\kappa\) is called the unit cumulant function for the exponential family. It is closely related to the cumulant generating functions for the probability measures in the exponential family, see Exercise 5.4.

Under the regularity assumptions we have made on \(\nu\) we can obtain a couple of very useful results about the exponential family.

Lemma 5.1 The set \(I\) is an open interval, the parametrization \(\theta \mapsto \rho_{\theta}\) is one-to-one, and the function \(\kappa : I \mapsto \mathbb{R}\) is strictly convex.

Proof. We first prove that the parametrization is one-to-one. If \(\rho_{\theta_1} = \rho_{\theta_2}\) their densities w.r.t. \(\nu\) must agree \(\nu\)-almost everywhere, which implies that \[(\theta_1 - \theta_2)y = \kappa(\theta_1) - \kappa(\theta_2)\] for \(\nu\)-almost all \(y.\) Since \(\nu\) is assumed not to be the zero measure or a one-point measure, this can only hold if \(\theta_1 = \theta_2.\)

To prove that \(I\) is an interval let \(\theta_1, \theta_2 \in I,\) and let \(\alpha \in (0, 1).\) Then by Hölder’s inequality \[ \begin{align*} \phi(\alpha \theta_1 + (1-\alpha)\theta_2) & = \int e^{\alpha \theta_1 y} e^{(1-\alpha)\theta_2 y} \, \nu(\mathrm{d} y) \\ & \leq \left(\int e^{\theta_1 y} \, \nu(\mathrm{d} y)\right)^{\alpha} \left(\int e^{\theta_2 y} \, \nu(\mathrm{d} y)\right)^{1-\alpha} \\ & = \phi(\theta_1)^{\alpha} \phi(\theta_2)^{1-\alpha} < \infty. \end{align*} \] This proves that \(I\) is an interval. It is by definition open. Finally, it follows directly from the inequality above that \[ \kappa(\alpha \theta_1 + (1-\alpha) \theta_2) \leq \alpha \kappa(\theta_1) + (1-\alpha)\kappa(\theta_2), \] which shows that \(\kappa\) is convex. If we have equality in this inequality, we have equality in Hölder’s inequality. This happens only if \(e^{\theta_1 y}/\phi(\theta_1) = e^{\theta_2 y}/\phi(\theta_2)\) for \(\nu\)-almost all \(y,\) and just as above we conclude that this implies \(\theta_1 = \theta_2.\) The conclusion is that \(\kappa\) is strictly convex.

Hölder’s inequality says that \[ \int|fg| \leq \left(\int |f|^p\right)^{\frac{1}{p}}\left(\int |g|^q\right)^{\frac{1}{q}} \] for \(p,q \geq 1\) and \(\frac{1}{p} + \frac{1}{q} = 1.\) It is used with \[ \begin{align*} f(y) & = e^{\alpha \theta_1 y}, \\ g(y) &= & e^{(1-\alpha)\theta_2 y}, \end{align*} \] \[ p = \frac{1}{\alpha} \ \mathrm{ and } \ q = \frac{1}{1-\alpha}. \]

The structure measure \(\nu\) determines the unit cumulant function by the formula \[ \kappa(\theta) = \log \int e^{\theta y} \, \nu(\mathrm{d}y). \] The fact that \(I\) is open implies that \(\kappa\) determines \(\nu\) uniquely, see Exercise 5.5. In many cases the structure measure belongs to a family of \(\sigma\)-finite measures \(\nu_{\psi}\) parametrized by \(\psi > 0,\) and whose cumulant functions are \(\theta \mapsto \kappa(\psi \theta)/\psi.\) That is, \(\nu_1 = \nu\) and \[ \frac{\kappa(\psi \theta)}{\psi} = \log \int e^{\theta y} \, \nu_{\psi}(\mathrm{d}y). \] Changing parameter from \(\psi \theta\) to \(\theta\) we can also write this as \[ \frac{\kappa(\theta)}{\psi} = \log \int e^{\frac{\theta y}{\psi}} \, \nu_{\psi}(\mathrm{d}y) = \log \int e^{\theta y} \, \bar{\nu}_{\psi}(\mathrm{d}y), \] where \(\bar{\nu}_{\psi}\) is a scale transformation of \(\nu_\psi\). Since this cumulant function is defined on the same open interval \(I,\) it uniquely determines \(\bar{\nu}_{\psi}\) and thus \(\nu_\psi\) – if there exists such a \(\nu_{\psi}.\)

It is not at all obvious whether there exists a \(\sigma\)-finite measure \(\nu_{\psi}\) with cumulant function \(\kappa(\psi \theta)/\psi\) for a given unit cumulant function \(\kappa.\) We will find this to be the case for all \(\psi > 0\) for several concrete examples, but there are examples, such as the Bernoulli model, where the extension is only possible for \(\psi^{-1} \in \mathbb{N}\). We will not pursue a systematic study, but see (Jorgensen 1997, chap. 4). What we can say is that if there is such a family \(\nu_{\psi}\) for \(\psi > 0,\) it is uniquely determined by \(\nu,\) and that all such measures will satisfy the same regularity conditions we have required of \(\nu.\)

Jorgensen, B. 1997. The Theory of Dispersion Models. Chapman & Hall/CRC Monographs on Statistics & Applied Probability. Taylor & Francis.

Whenever \(\nu\) is embedded in such a family \((\nu_{\psi})_{\psi > 0}\) of structure measures we introduce the exponential dispersion model determined by \(\nu\) – a two-parameter family of probability measures – by \[ \frac{\mathrm{d} \rho_{\theta, \psi}}{\mathrm{d} \nu_{\psi}} (y) = e^{\frac{\theta y - \kappa(\theta)}{\psi}}. \tag{5.1}\] The parameter \(\psi\) is called the dispersion parameter, and we call \(\nu = \nu_1\) the unit structure measure for the exponential dispersion model. For fixed \(\psi\) the exponential dispersion model is an exponential family with structure measure \(\nu_{\psi}\) and canonical parameter \(\theta/\psi.\) We abuse the terminology slightly and call \(\theta\) the canonical parameter for the exponential dispersion model. Thus, whether \(\theta/\psi\) or \(\theta\) is canonical depends upon whether we regard the model as an exponential family with structure measure \(\nu_{\psi}\) or as an exponential dispersion model with unit structure measure \(\nu.\) Whenever we consider an exponential dispersion model, the measure \(\nu\) will always denote the unit structure measure \(\nu_1.\)

Definition 5.2 The probability distribution \(\rho_{\theta, \psi}\) is called the \((\theta, \nu_{\psi})\)-exponential dispersion distribution and is denoted \(\mathcal{E}(\theta, \nu_{\psi}).\)

In practice we check that a given parametrized family of distributions is, in fact, an exponential dispersion model by checking that its density can be brought on the form (5.1).

Example 5.2 The normal distribution \(\mathcal{N}(\mu, 1)\) has density \[ \frac{1}{\sqrt{2\pi}} e^{-\frac{(y-\mu)^2}{2}} = e^{y\mu - \frac{\mu^2}{2}} \frac{1}{\sqrt{2\pi}} e^{-\frac{y^2}{2}} \] w.r.t. Lebesgue measure \(m.\) We identify this as an exponential family with1 \[ \theta = \mu, \quad \kappa(\theta) = \frac{\theta^2}{2} \quad \mathrm{and} \quad \frac{\mathrm{d}\nu}{\mathrm{d} m}(y) = \frac{1}{\sqrt{2\pi}} e^{-\frac{y^2}{2}}. \] Thus the structure measure \(\nu\) is the standard normal distribution \(\mathcal{N}(0, 1)\).

1 The \(\kappa\) and \(\nu\) are not unique. The constant \[\frac{1}{\sqrt{2\pi}}\] in \(\nu\) can be moved to a \(- \log \sqrt{2\pi}\) term in \(\kappa.\)

2 With \(S_{\sigma}(x) = \sigma x\) denoting a scale transformation, \(\nu_{\psi} = S_{\sqrt{\psi}}(\nu).\)

Introducing a dispersion parameter \(\psi > 0\) we observe that
\[ \frac{\kappa(\theta)}{\psi} = \frac{1}{\psi}\frac{\theta^2}{2} = \frac{(\theta/\sqrt{\psi})^2}{2} = \log \int e^{\frac{\theta y}{\sqrt{\psi}}} \nu(\mathrm{d}y) = \log \int e^{\frac{\theta y}{\psi}} \nu_\psi(\mathrm{d}y) \] where \(\nu_\psi\) is a scale transformation2 of \(\nu\). That is, \(\nu_\psi\) is the \(\mathcal{N}(0, \psi)\) distribution with \[ \frac{\mathrm{d} \nu_{\psi}}{\mathrm{d} m}(y) = \frac{1}{\sqrt{2\pi \psi}} e^{-\frac{y^2}{2\psi}}. \] The corresponding exponential dispersion distribution is \(\mathcal{N}(\mu, \psi)\). With the notation introduced in Definition 5.2, \[ \mathcal{E}\left(\mu, \nu_{\sigma^2} \right) = \mathcal{N}(\mu,\sigma^2), \] where we, as is usual for the normal distribution, denote the dispersion parameter by \(\sigma^2\).

Example 5.3 The Poisson distribution with mean \(\mu > 0\) has density (probability mass function) \[e^{-\mu} \frac{\mu^y}{y!}\] w.r.t. the counting measure \(\tau\) for \(y \in \mathbb{N}_0.\) With \(\theta = \log \mu\) we can rewrite this density as \[e^{y \log \mu -\mu} \frac{1}{y!} = e^{y \theta - e^{\theta}} \frac{1}{y!},\] and we identify the distribution as an exponential family with \[\theta = \log \mu, \quad \kappa(\theta) = e^{\theta} \quad \mathrm{and} \quad \frac{\mathrm{d}\nu}{\mathrm{d} \tau} (y) = \frac{1}{y!}.\] We conclude that \[\mathcal{E}\left(\log \mu, \frac{1}{y!} \cdot \tau \right) = \mathrm{Pois}(\mu).\]

It is possible to introduce a dispersion parameter via a scaled Poisson distribution, see Exercise 5.8, but the resulting distribution will only have integer support if \(\psi \in \mathbb{N}\). It is therefore not useful in general for modeling counts.

The next example is the \(\Gamma\)-family of distributions, whose representation as an exponential dispersion model is less obvious. These distributions can be parametrized by a shape and a scale parameter, and we have to rewrite their density to identify the parametrization in terms of the canonical parameter and the dispersion parameter. For this family of distributions there is yet another useful parametrization in terms of mean and variance.

Example 5.4 The \(\Gamma\)-distribution with shape parameter \(\lambda > 0\) and scale parameter \(\alpha > 0\) has density \[ \frac{1}{\alpha^{\lambda} \Gamma(\lambda)} y^{\lambda - 1} e^{-y/\alpha} = e^{\frac{- y/(\lambda \alpha) - \log (\lambda \alpha)}{1/\lambda}} \frac{\lambda^{\lambda}}{\Gamma(\lambda)} y^{\lambda - 1} \] w.r.t. Lebesgue measure \(m\) on \((0, \infty).\) We identify this family of distributions as the exponential dispersion model with dispersion parameter \(\psi = 1/\lambda,\) canonical parameter \[ \theta = - 1 /(\lambda \alpha) < 0, \] \[ \kappa(\theta) = - \log (- \theta), \] and structure measure given by \[ \frac{\mathrm{d} \nu_{\psi}}{\mathrm{d} m} = \frac{1}{\psi^{1/\psi}\Gamma(1/\psi)} y^{\frac{1}{\psi} - 1} \] on \((0, \infty).\) We have \(I = (- \infty, 0).\) The mean value of the \(\Gamma\)-distribution is \(\mu \coloneqq \alpha \lambda = - \frac{1}{\theta},\) and the variance is \(\alpha^2 \lambda = \psi \mu^2 = \psi/\theta^2.\) Observe that the unit structure measure, \(\nu_1,\) is Lebesgue measure on \((0, \infty),\) and the corresponding exponential family is the exponential distribution.

The next theorem shows that the mean and variance for exponential dispersion models are always computable via differentiation of the cumulant function.

For the \(\Gamma\)-distribution: \[ \kappa'(\theta) = - \frac{1}{\theta} \] and \[ \kappa''(\theta) = \frac{1}{\theta^2}. \]

Theorem 5.1 The function \(\kappa\) is infinitely often differentiable on \(I.\) If \(Y \sim \mathcal{E}(\theta, \nu_{\psi})\) for \(\theta \in I\) then \[ E(Y) = \kappa'(\theta) \tag{5.2}\] and \[ V(Y) = \psi \kappa''(\theta). \tag{5.3}\]

Proof. It follows by a suitable domination argument that, for \(\theta \in I,\) \[ \frac{\mathrm{d}^n}{\mathrm{d}\theta^n} \phi(\theta) = \int y^n e^{\theta y} \, \nu(\mathrm{d} y) = \phi(\theta) E(Y^n). \] Since \(\kappa(\theta) = \log \phi(\theta),\) it follows that also \(\kappa\) is infinitely often differentiable. In particular, if \(Y \sim \mathcal{E}(\theta, \nu_1),\) \[ E(Y) = \frac{\phi'(\theta)}{\phi(\theta)} = (\log \phi)'(\theta) = \kappa'(\theta), \] and \[ E(Y^2) = \frac{\phi''(\theta)}{\phi(\theta)}, \] and we find that \[ V(Y) = E(Y^2) - (E(Y))^2 = \frac{\phi''(\theta)\phi(\theta) - \phi'(\theta)^2}{\phi(\theta)^2} = \kappa''(\theta). \] The general case for \(\psi \neq 1\) can be proved by similar arguments. Alternatively, observe that for an arbitrary dispersion parameter \(\psi > 0,\) the distribution of \(Y\) is an exponential family with canonical parameter \(\theta_0 = \theta/\psi\) and with corresponding cumulant function \[ \kappa_0(\theta_0) = \frac{\kappa(\psi \theta_0)}{\psi}. \] The conclusion follows by differentiating \(\kappa_0\) twice w.r.t. \(\theta_0.\)

We know from Lemma 5.1 that \(\kappa\) is strictly convex, which actually implies, since \(\kappa\) was found to be differentiable on \(I,\) that \(\kappa'\) is a strictly increasing function. This can also be seen directly from the previous theorem. Indeed, by the regularity assumptions on \(\nu_{\psi},\) \(\rho_{\theta, \psi}\) is not a Dirac measure, which implies that \(V(Y) > 0\) in (5.3), see Exercise 5.3. Hence \(\kappa''\) is strictly positive, and it follows that \(\kappa'\) is strictly increasing, and that \(\kappa\) is strictly convex. That \(\kappa'\) is continuous and strictly increasing imply that the range of \(\kappa',\) \(J \coloneqq \kappa'(I),\) is an open interval. This range is the range of possible mean values for the exponential dispersion model. Since \(\kappa'\) bijectively maps \(I\) onto \(J\) we can always express the variance in terms of the mean.

Definition 5.3 The variance function \(\mathcal{V} : J \to (0, \infty)\) is defined as \[\mathcal{V}(\mu) = \kappa''((\kappa')^{-1}(\mu)).\]

From the considerations above, the map \[ (\theta, \psi) \mapsto (\kappa'(\theta), \psi \kappa''(\theta)) = (\mu, \psi \mathcal{V}(\mu)) \] is a bijection from \(I \times (0, \infty)\) onto \(J \times (0,\infty),\) and this shows that the exponential dispersion model can be parametrized either in terms of the canonical and dispersion parameters or in terms of the mean and variance.

A third parametrization of the exponential dispersion model is occasionally useful or even natural. With \(H \subseteq \mathbb{R}\) and \(\theta : H \to I,\) the canonical parameter, \(\theta(\eta),\) can be expressed in terms of an arbitrary parameter \(\eta.\) We call \(\eta\) arbitrary because we don’t make any assumptions about what this parameter is, besides that it parametrizes the exponential dispersion model. The density for the exponential dispersion model w.r.t. \(\nu_{\psi}\) in the arbitrary parametrization is \[ e^{\frac{\theta(\eta)y - c(\eta)}{\psi}} \] where \(c(\eta) = \kappa(\theta(\eta))\) and \(\eta \in H.\) The parametrization, \(\eta \mapsto \theta(\eta)\), is allowed to depend upon additional (nuisance) parameters. We suppress any such dependences in the abstract notation. It is, however, not allowed to depend upon the dispersion parameter3, because then the dispersion parameter cannot be eliminated from subsequent estimation equations.

3 In technical terms, the dispersion parameter must be variation independent of nuisance parameters entering \(\eta \mapsto \theta(\eta)\).

The mean value function, in the arbitrary parametrization, is given as \[ \mu(\eta) = \kappa'(\theta(\eta)). \] We can then express the mean and the variance functions in terms of the function \(\eta \mapsto c(\eta).\)

Corollary 5.1 If \(\theta\) is twice differentiable as a function of \(\eta\) with \(\theta'(\eta) \neq 0,\) then
\[ \mu(\eta) = \frac{c'(\eta)}{\theta'(\eta)} \] and \[ \mathcal{V}(\mu(\eta)) = \frac{c''(\eta)\theta'(\eta) - c'(\eta)\theta''(\eta)}{\theta'(\eta)^3} = \frac{\mu'(\eta)}{\theta'(\eta)}. \]

Figure 5.1: Relations between the three parametrizations.

Proof. From the definition \(c(\eta) = \kappa(\theta(\eta))\) we get by differentiation and (5.2) that \[c'(\eta) = \kappa'(\theta(\eta)) \theta'(\eta) = \mu(\eta) \theta'(\eta),\] and the first identity follows. An additional differentiation yields \[ \begin{align*} c''(\eta) & = \kappa''(\theta(\eta)) \theta'(\eta)^2 + \kappa'(\theta(\eta)) \theta''(\eta) \\ & = \mathcal{V}(\mu(\eta)) \theta'(\eta)^2 + \frac{c'(\eta) \theta''(\eta)}{\theta'(\eta)}, \end{align*} \] and the second identity follows by isolating \(\mathcal{V}(\mu(\eta)).\)

When the mean value function, defined above as a function of the arbitrary parameter, is bijective, its inverse is called the link function and denoted \(g.\) It maps the mean value \(\mu\) to the arbitrary parameter \(\eta,\) that is, \(\eta = g(\mu).\) The choice of the link function (equivalently, the mean value function) completely determines the \(\theta\)-map and vice versa. The choice for which \(\theta = \eta\) plays a particularly central role. If \(\theta = \eta\) then \(\mu = \kappa',\) and the corresponding link function is \(g = (\kappa')^{-1}.\)

Example 5.6 For the Poisson distribution we have \(\kappa(\theta) = e^{\theta},\) see Example 5.9. This implies that \(\kappa'(\theta) = e^{\theta}\) and the canonical link function is, for the Poisson distribution, \(g(\mu) = \log \mu.\)

A regression model with \(\eta = X^T \beta\) and a log-link function is often called a log-linear model. Thus the canonical link function for the Poisson distribution gives a log-linear regression model. Example 5.5 gives a log-linear regression model with a \(\Gamma\)-distributed response, but the log-link is not the canonical link for the \(\Gamma\)-distribution as the following example will reveal.

Example 5.7 For the \(\Gamma\)-model we have from Example 5.4 that \(\kappa(\theta) = - \log(-\theta)\) and \[\kappa'(\theta) = -\frac{1}{\theta}.\] This gives that the canonical link function for the \(\Gamma\)-model is \[g(\mu) = - \frac{1}{\mu}.\] The canonical link for the \(\Gamma\)-model is not that attractive. Its range is \(I = (-\infty, 0),\) and not the entire real line, which can be a nuisance. In practice, the \(\Gamma\)-model is often used with the log-link instead.

Figure 5.2: The linear predictor enters the exponential dispersion model through the mean value function.

In generalized linear regression models, the response distribution will be specified in terms of the linear predictor \(\eta = X^T \beta,\) that determines the mean, and an exponential dispersion model, that determines the remaining aspects of the distribution. The specification of the mean is given in terms of the mean value function, or equivalently in terms of the link function. That is, with link function \(g,\)
\[ g(E(Y \mid X)) = \eta = X^T \beta \] or \[ E(Y \mid X) = \mu(\eta). \] This specifies implicitly the canonical parameter in the exponential dispersion model, and results in a model with \[ V(Y \mid X) = \psi \mathcal{V}(\mu(\eta)) \] where \(\mathcal{V}\) is the variance function.

5.3 Deviance

The (unit) deviance for an exponential dispersion model generalizes the squared error \[ (y - \mu)^2 \] for the \(\mathcal{N}(\mu, 1)\) distribution. For a unit cumulant function \(\kappa : I \to \mathbb{R}\) we let \[ \ell_y(\theta) = \theta y - \kappa(\theta) \] for \(y \in \overline{J}\) and \(\theta \in I\) denote the log-likelihood of the corresponding exponential family, see also Lemma 6.1 in Chapter 6. The deviance is defined as twice the negative log-likelihood up to an additive constant. In terms of the canonical parameter the deviance is defined as \[ 2 \left( \sup_{\theta_0 \in I} \ell_y(\theta_0) - \ell_y(\theta)\right) \] for \(y \in \overline{J}\) and \(\theta \in I.\) The supremum in the definition is attained in \(g(y)\) for \(y \in J\) where \(g = (\kappa')^{-1}\) is the canonical link. For practical computations it is more convenient to express the deviance in terms of the mean value parameter than the canonical parameter.

Definition 5.5 With \(g\) denoting the canonical link function the unit4 deviance is defined as \[ d(y, \mu) = 2 \left( \sup_{\mu_0 \in J} \Big\{ g(\mu_0) y - \kappa(g(\mu_0)) \Big\} - g(\mu) y + \kappa(g(\mu)) \right) \] for \(y \in \overline{J}\) and \(\mu \in J.\)

4 Here unit refers to the dispersion parameter being 1.

For \(y \in J,\) the supremum is attained in \(\mu_0 = y\) and the unit deviance can be expressed as \[ d(y, \mu) = 2 \left( y (g(y) - g(\mu)) - \kappa(g(y)) + \kappa(g(\mu)) \right). \tag{5.4}\] Note that by the definition of the unit deviance we have that \(d(y, y) = 0\) and \[ d(y, \mu) > 0 \] for \(\mu \neq y.\)

The deviance has simple analytic expressions for many concrete examples, which are interpretable as measures of how the observation \(y\) deviates from the expectation \(\mu.\)

Example 5.8 For the normal distribution the canonical link is the identity and \(\kappa(\theta) = \theta^2/2,\) hence \[d(y, \mu) = 2(y ( y - \mu) - y^2/2 + \mu^2/2) = (y - \mu)^2.\]

Example 5.9 For the Poisson distribution the canonical link is the logarithm, and \(\kappa(\theta) = e^{\theta},\) hence \[d(y, \mu) = 2(y(\log y - \log \mu) - y + \mu) = 2 ( y \log (y/\mu) - (y - \mu))\] for \(y, \mu > 0.\) It is clear from the definition that \(d(0, \mu) = 2\mu,\) and the identity above can be maintained even for \(y = 0\) by the convention \(0 \log 0 = 0.\)

Example 5.10 For the binomial distribution it follows directly from the definition that for \(y = 1, \ldots, m - 1\) and \(\mu \in (0, m)\) \[ \begin{align*} d(y, \mu) & = 2 \Big( y \log y/m + (m-y) \log (1 - y/m) \\ & \hskip 1cm - y \log \mu/m - (m - y) \log (1 - \mu/m) \Big) \\ & = 2 \Big( y \log(y/\mu) + (m - y) \log \big( (m - y)/(m - \mu) \big) \Big). \end{align*} \] Again, by the convention \(0 \log 0 = 0\) the identity is seen to extend also to the extreme cases \(y = 0\) or \(y = m.\)

The unit deviance is approximately a quadratic form for \(y \approx \mu.\)

Theorem 5.2 For the unit deviance it holds that \[d(y, \mu) = \frac{(y - \mu)^2}{\mathcal{V}(\mu)} + o((y - \mu)^2)\] for \(y, \mu \in J.\)

Proof. We consider the function \[ y \mapsto d(y, \mu) \] around \(\mu.\) Since we know that \(d(\mu, \mu) = 0,\) that \(y = \mu\) is a local minimum, and that \(d\) is twice continuously differentiable in \(J \times J,\) we get by Taylor’s formula that \[d(y, \mu) = \frac{1}{2} \partial_{y}^2 d(\mu, \mu) (y - \mu)^2 + o( (y - \mu)^2).\] Using that \(\partial_y \kappa(g(y)) = y g'(y)\) we find that \[ \begin{align*} \frac{1}{2} \partial_{y}^2 d(y, \mu) & = \partial_{y}^2 \Big\{ y g(y) - \kappa(g(y)) - yg(\mu) + \kappa(g(\mu))\Big\} \\ & = \partial_{y} \Big\{ g(y) + y g'(y) - y g'(y) - g(\mu) \Big\}\\ & = g'(y) = \frac{1}{\kappa''(g(y))} = \frac{1}{\mathcal{V}(y)}. \end{align*} \] Plugging in \(y = \mu\) completes the proof.

TODO:

  • Uniqueness of the variance function.
  • Representation of the deviance in terms of the variance function.

Exercises

Exercise 5.1 The probability mass function for any probability measure on \(\{0, 1\}\) can be written as \[p^y(1-p)^{1-y}\] for \(y \in \{0, 1\}.\) Show that this family of probability measures for \(p \in (0,1)\) form an exponential family. Identify \(I\) and how the canonical parameter depends on \(p.\)

Exercise 5.2 Show that the binomial distributions on \(\{0, 1, \ldots, n\}\) with success probabilities \(p \in (0,1)\) form an exponential family.

Exercise 5.3 Show that if the structure measure \(\nu = \delta_y\) is the Dirac-measure in \(y\) then for the corresponding exponential family we have \(\rho_{\theta} = \delta_y\) for all \(\theta \in \mathbb{R}.\) Show next, that if \(\nu\) is not a one-point measure, then \(\rho_{\theta}\) is not a Dirac-measure for \(\theta \in I\) and conclude that its variance is strictly positive.

Exercise 5.4 If \(\rho\) is a probability measure on \(\mathbb{R},\) its cumulant generating function is defined as the function \[ K(s) = \log \int e^{s y} \ \rho(\mathrm{d} y) \] for \(s\) such that the integral is finite. Let \(\rho_{\theta, \psi}\) be an exponential dispersion distribution with unit cumulant function \(\kappa.\) Show that \[ K_{\theta, \psi}(s) = \frac{\kappa(\theta + \psi s) - \kappa(\theta)}{\psi} \] is the cumulant generating function for \(\rho_{\theta, \psi}.\)

Exercise 5.5 Let \(\rho_{\theta}\) be the exponential family given by the structure measure \(\nu.\) Show that \[ \frac{\mathrm{d} \nu}{\mathrm{d} \rho_{\theta_0}} = e^{\kappa(\theta_0) - \theta y} \] for any \(\theta_0 \in I.\) Then show that \(\nu\) is uniquely specified by the unit cumulant function \(\kappa.\)

Hint: Use that the cumulant generating function for \(\rho_{\theta_0}\) is determined by \(\kappa,\) and that a probability measure is uniquely determined by its cumulant generating function if it is defined on an open interval around \(0.\)

Exercise 5.6 This exercise is about the inverse Gaussian distribution, which has density \[ f(y) = \sqrt{\frac{\lambda}{2 \pi y^3}} e^{-\frac{\lambda(y - \mu)^2}{2 \mu^2 y}} \tag{5.5}\] w.r.t. Lebesgue measure on \((0, \infty).\) In this parametrization, \(\lambda, \mu > 0\) are positive parameters.

  1. Show that the family of inverse Gaussian distributions parametrized by \(\lambda, \mu > 0\) is an exponential dispersion model. Identify how the canonical parameter and the dispersion parameter are given in terms of \(\lambda\) and \(\mu\) – with the lack of uniqueness resolved by letting the unit structure measure correspond to \(\lambda = 1.\)
  2. Show that the mean is \(\mu\) and show that the variance function is \[ \mathcal{V}(\mu) = \mu^3. \]
  3. Find the canonical link function and the unit deviance.
The use of \(\mu\) in the “Wikipedia” parametrization in Equation 5.5 is sensible!

Exercise 5.7 This exercise is about the exponential family with structure measure \(\nu = m_{[-1,1]}\) being Lebesgue measure restricted to the interval \([-1, 1].\)

  1. Compute the unit cumulant function \(\kappa\) and show that \(I = \mathbb{R}.\) Plot the density for the corresponding exponential family w.r.t. Lebesgue measure for \(\theta = -1\) and \(\theta = 1.\)

Assume that \(Y \sim \rho_{\theta}\) where \[ \frac{\mathrm{d} \rho_{\theta}}{\mathrm{d} \nu} = e^{\theta y - \kappa(\theta)}. \]

  1. Compute the mean value function \(\mu(\theta)\) and show that \[ V(Y) = \frac{1}{\theta^2} - \frac{1}{\sinh(\theta)^2} \] for \(\theta \neq 0.\) Plot the variance function \(\mathcal{V}(\mu)\) as a function of \(\mu.\)

    Hint: For the plot, it is not necessary to compute an analytic expression for the inverse of the mean value map.

Exercise 5.8 The unit cumulant function for the Poisson model is \(\kappa(\theta) = e^\theta\), see Example 5.9. Let \(S_\psi(y) = \psi y\) denote the scale transformation determined by \(\psi > 0\).

  1. Show that \[ \frac{\kappa(\theta)}{\psi} = \log \int e^{\theta y} \bar{\nu}_\psi(\mathrm{d} y) = \log \int e^{\frac{\theta y}{\psi}} \nu_\psi(\mathrm{d} y) \] with \[ \frac{\mathrm{d}\bar{\nu}_\psi}{\mathrm{d} \tau} (y) = \frac{1}{\psi^y y!}. \] and \(\nu_\psi = S_{\psi}(\bar{\nu}_\psi)\).

  2. Argue that the Poisson model in Example 5.9 can be extended to an exponential dispersion model where the support of the probability distribution \(\rho_{\theta, \psi}\) is \(\psi \mathbb{N}_0\).

  3. Show that \(S_{\psi^{-1}}(\rho_{\theta, \psi}) = \mathrm{Pois}(e^\theta / \psi)\). That is, \(Y \sim \mathrm{Pois}(\mu / \psi)\) if and only if \(\psi Y \sim \mathcal{E}(\log \mu, \nu_\psi)\), which shows that the members of the exponential dispersion model are scaled Poisson distributions.

    Hint: You can compute the probability mass function explicitly for the distribution \(S_{\psi^{-1}}(\rho_{\theta, \psi})\) and identify it as the probability mass function for a Poisson distribution. Alternative, you can compute the cumulant generating function of \(\rho_{\theta, \psi}\) using Exercise 5.4 and identify it as the cumulant generating function for a scaled Poisson distribution.

  4. Compute \(E(\psi Y)\) and \(V(\psi Y)\) for \(\psi Y \sim \mathcal{E}(\log \mu, \nu_\psi)\) directly, using that \(Y \sim \mathrm{Pois}(\mu / \psi)\), and demonstrate that the results agree with Theorem 5.1.

Exercise 5.9 Let \(N \sim \mathrm{Pois}(\mu_0)\) and let \(Y_1, Y_2, \ldots \in \mathbb{N}_0\) be i.i.d. variables independent of \(N\) and with finite second moment. Define \[ Y = \sum_{n=0}^{N} Y_n, \] whose distribution is a compound Poisson distribution.

  1. Show that \[\mu \coloneqq E(Y) = E(Y_1) \mu_0.\]
  2. Show that \[V(Y) = \frac{E(Y_1^2)}{E(Y_1)} \mu.\]
This particular example is a distribution on the non-negative integer with the ratio \[\psi = \frac{E(Y_1^2)}{E(Y_1)}\] playing the role as a dispersion parameter.