2  Generalized linear models

2.1 Introduction

The generalized linear model extends the normal linear model defined in Section 1.1 to allow a more flexible family of probability distributions.

Suppose that \(y_1, \ldots, y_n\) are observations on random variables \(Y_1, \ldots, Y_n\) that are conditionally independent given \(x_1, \ldots, x_n\), where \(x_i\) is a \(p\)-vector of covariates. A generalized linear model (GLM) assumes that, conditionally on \(x_i\), \(Y_i\) has an exponential family distirbution with density or probability mass function \[ f_{Y}(y; \theta, \phi) = \exp\left(\sum_{i=1}^n \frac{y_i \theta_i - b(\theta_i)}{\phi_i} +\sum_{i=1}^n c(y_i,\phi_i)\right) \, , \tag{2.1}\] where \(\theta = (\theta_1, \ldots, \theta_n)^\top\) is the collection of canonical parameters and \(\phi = (\phi_1, \ldots , \phi_n)^\top\) is the collection of dispersion parameters (where they exist). Commonly, the dispersion parameters are known up to, at most, a single common unknown \(\sigma^2\), and we write \(\phi_i = \sigma^2 / m_i\) where the \(m_i\) represent known weights.

The distribution of the response variable \(Y_i\) depends on the explanatory data \(x_i=(1, x_{i1}, x_{i2}, \ldots, x_{ip})^{\top}\) through the linear predictor \(\eta_i\), where \[ \begin{aligned} \eta_i & = \beta_0 + \beta_1 x_{i1} + \beta_2 x_{i2} + \ldots + \beta_p x_{ip} \\ & = \sum_{j=0}^p x_{ij} \beta_j \\ &= x_i^{\top}\beta \\ & = [X\beta]_i \qquad (i = 1, \ldots, n) \,, \end{aligned} \] in an exactly analagous fashion to the linear model in Section 1.1.

The distribution of \(Y\) is linked to the linear predictor \(\eta\) through the link function \(g\) as \[ \eta_i = g(\mu_i) \qquad (i = 1, \ldots, n) \, , \] where \(\mu_i = E(Y_i)\).

In principle, the link function \(g\) can be any one-to-one differentiable function. However, we note that \(\eta_i\) can in principle take any value in \(\Re\), because we make no restriction on possible values taken by explanatory variables or model parameters. However, for some exponential family distributions \(\mu_i\) is restricted. For example, for the Poisson distribution \(\mu_i \in \Re^+\); for the Bernoulli distribution \(\mu_i \in (0,1)\). If \(g\) is not chosen carefully, then there may exist a combination of \(x_i\) and \(\beta\) such that \(\eta_i \ne g(\mu_i)\) for any possible value of \(\mu_i\). Most common choices of link function map the set of allowed values for \(\mu_i\) onto \(\mathbb{R}\).

Recall that for a random variable \(Y\) with an exponential family distribution, \(E(Y)=b'(\theta)\). Hence, for a generalized linear model \[ \mu_i = E(Y_i) = b'(\theta_i) \qquad (i = 1, \ldots, n)\,. \] So, \[ \theta_i = b'^{-1}(\mu_i) \qquad (i = 1, \ldots, n)\,, \] and, because \(g(\mu_i) = \eta_i = x_i^{\top}\beta\), \[ \theta_i = b'^{-1} (g^{-1}(x_i^{\top}\beta)) \qquad (i = 1, \ldots, n)\,. \tag{2.2}\] Hence, we can express the joint density (2.1) in terms of the coefficients \(\beta\), and for observed data \(y\), this is the likelihood \(f_{Y}(y;\beta,\phi)\) about \(\beta\).

Note that considerable simplification is obtained in (2.1) and (2.2) if the functions \(g\) and \(b'^{-1}\) are identical. Then, \[ \theta_i = x_i^{\top}\beta \qquad (i = 1, \ldots, n)\, . \] The link function \[ g(\mu) = b'^{-1}(\mu) \] is called the canonical link function. Under the canonical link, the canonical parameter is equal to the linear predictor.

Table 2.1: Canonical link functions
Distribution Normal Poisson Bernoulli/Binomial
\(b(\theta)\) \({1\over 2}\theta^2\)
\(b'(\theta) = \mu\) \(\theta\) \(\frac{\exp\theta}{1+\exp\theta}\)
\(b'^{-1}(\mu) = \theta\) \(\mu\) \(\log{\frac{\mu}{1-\mu}}\)
Canonical link \(g(\mu)=\mu\) \(g(\mu)=\log\mu\) \(g(\mu)=\log{\frac{\mu}{1-\mu}}\)
Name of link Identity link Log link Logit link

Exercise 2.1 (GLM characteristics) Complete Table 2.1.

Clearly the linear model considered in Chapter 1 is also a generalized linear model where \(Y_1, \ldots, Y_n\) are independent and normally distributed, the explanatory variables enter the model through the linear predictor \[ \eta_i = x_i^{\top}\beta \qquad (i = 1, \ldots, n)\,, \] and the link between \(E(Y) = \mu\) and the linear predictor \(\eta\) is through the (canonical) identity link function \[ \mu_i = \eta_i \qquad (i = 1, \ldots, n)\,. \]

2.2 Maximum likelihood estimation

As usual, we maximize the log likelihood function which, from (2.1), can be written as \[ \log f_{Y}(y;\beta,\phi) = \sum_{i=1}^n\frac{y_i\theta_i-b(\theta_i)}{\phi_i} +\sum_{i=1}^n c(y_i, \phi_i) \,, \tag{2.3}\] and depends on \(\beta\) through expression (2.2) for the canonical parameters.

The maximum likelihood estimate \(\hat\beta\) satisfies \(u(\hat\beta) = 0\) where \(u\) is the score vector whose components are given by \[ \begin{aligned} u_k(\beta) & \equiv{\partial\over{\partial\beta_k}}\log f_{Y}(y;\beta) \\ & = \sum_{i=1}^n{\partial\over{\partial\beta_k}} \left[{{y_i\theta_i - b(\theta_i)}\over{\phi_i}}\right] \\ &=\sum_{i=1}^n{\partial\over{\partial\theta_i}}\left[{{y_i\theta_i-b(\theta_i)} \over{\phi_i}}\right]{{\partial\theta_i}\over{\partial\mu_i}} {{\partial\mu_i}\over{\partial\eta_i}}{{\partial\eta_i}\over{\partial\beta_k}} \\ &=\sum_{i=1}^n{{y_i-b'(\theta_i)}\over{\phi_i}} {{x_{ik}}\over{b''(\theta_i)g'(\mu_i)}} \\ &=\sum_{i=1}^n{{y_i-\mu_i}\over{\mathop{\mathrm{var}}(Y_i)}} {{x_{ik}}\over{g'(\mu_i)}} \qquad (k = 0, \ldots, p)\,, \end{aligned} \tag{2.4}\] which depends on \(\beta\) through \(\mu_i = E(Y_i)\) and \(\mathop{\mathrm{var}}(Y_i)\) \((i = 1, \ldots, n)\).

The equations \(u(\hat\beta) = 0\) are usually non-linear and have no analytic solution. For that reason, we rely on numerical methods to solve them.

First, we note that the Hessian and Fisher information matrices can be derived directly from (2.4), as \[ \begin{aligned} [H(\beta)]_{jk}&={{\partial^2}\over{\partial\beta_j\partial\beta_k}}\log f_{Y}(y;\beta) \\ &={\partial\over{\partial\beta_j}}\sum_{i=1}^n{{y_i-\mu_i}\over{\mathop{\mathrm{var}}(Y_i)}} {{x_{ik}}\over{g'(\mu_i)}} \\ &=\sum_{i=1}^n{{-{{\partial\mu_i}\over{\partial\beta_j}}}\over{\mathop{\mathrm{var}}(Y_i)}} {{x_{ik}}\over{g'(\mu_i)}}+\sum_{i=1}^n(y_i-\mu_i){\partial\over{\partial\beta_j}} \left[{{x_{ik}}\over{\mathop{\mathrm{var}}(Y_i) g'(\mu_i)}}\right]\,, \\ \end{aligned} \] and \[ [{\cal I}(\beta)]_{jk} = E(-H(\beta))_{jk} =\sum_{i=1}^n{{{{\partial\mu_i}\over{\partial\beta_j}}}\over{\mathop{\mathrm{var}}(Y_i)}} {{x_{ik}}\over{g'(\mu_i)}} =\sum_{i=1}^n{{x_{ij}x_{ik}}\over{\mathop{\mathrm{var}}(Y_i)g'(\mu_i)^2}} \,. \]

Exercise 2.2 (Fisher information matrix) Show that the Fisher information matrix can be written as \[ I(\beta) = X^{\top} W X \, , \tag{2.5}\] where \(X\) is the model matrix and \[ W = {\rm diag}(w) = \begin{bmatrix} w_1 & 0 & \cdots & 0 \\ 0 & w_2 & & \vdots \\ \vdots & & \ddots & 0 \\ 0 & \cdots & 0 & w_n \end{bmatrix}\, , \] where \[ w_i = {1\over{\mathop{\mathrm{var}}(Y_i)g'(\mu_i)^2}} \qquad (i = 1, \ldots, n) \, . \]

The Fisher information matrix \({\cal I}(\beta)\) depends on \(\beta\) through \(\mu_i\) and \(\mathop{\mathrm{var}}(Y_i)\) \((i = 1, \ldots, n)\).

The scores in (2.4) may now be written as \[ \begin{aligned} u_k(\beta) & =\sum_{i=1}^n (y_i-\mu_i) x_{ik} w_i g'(\mu_i) \\ & = \sum_{i=1}^n x_{ik} w_i z_i \qquad (k = 0, \ldots, p) \,, \end{aligned} \] where \[ z_i = (y_i-\mu_i) g'(\mu_i) \qquad (i = 1, \ldots, n) \, . \] Hence, \[ u(\beta)= X^{\top} W z \,. \tag{2.6}\]

One possible method to solve the \(p\) simultaneous equations \(u(\beta) = 0\) is the Newton-Raphson method. If \(\beta^{t}\) is the current estimate of \(\beta\), then the next estimate is \[ \beta^{t+1} = \beta^{t} - H(\beta^{t})^{-1} u(\beta^{t}) \,. \tag{2.7}\] A popular alternative to Newton-Raphson replaces \(H(\beta)\) in (2.7) with \(E(H(\beta)) = -I(\beta)\). If \(\beta^{t}\) is the current estimate of \(\beta\), the next estimate is \[ \beta^{t+1}=\beta^{t}+{\cal I}(\beta^{t})^{-1}u(\beta^{t}) \,. \tag{2.8}\] The resulting iterative algorithm is called Fisher scoring. Notice that if we substitute (2.5) and (2.6) into (2.8) we get \[ \begin{aligned} \beta^{t+1}&=\beta^{t}+[X^{\top}W^tX]^{-1}X^{\top}W^tz^t \\ &=[X^{\top}W^tX]^{-1}[X^{\top}W^tX\beta^{t}+X^{\top}W^tz^t] \\ &=[X^{\top}W^tX]^{-1}X^{\top}W^t[X\beta^{t}+z^t] \\ &=[X^{\top}W^tX]^{-1}X^{\top}W^t[\eta^{t}+z^t] \,, \end{aligned} \] where \(\eta^{t}\), \(W^t\), and \(z^t\) are \(\eta\), \(W\) and \(z\) evaluated at \(\beta^t\).

As is clear, \(\beta^{t+1}\) are estimates from a weighted linear regression model of the, so called, working variates \(\eta^t + z^t\) on \(X\) with weights \(W^t\). Equivalently, \(\beta^{t+1}\) minimizes the weighted sum of squares \[ (\eta^t + z^t - X b)^{\top} W^t (\eta^t + z^t - X b) = \sum_{i=1}^n w_i^t \left(\eta_i^t + z_i^t - x_i^{\top}b\right)^2 \,, \] with respect to \(b\).

The Fisher scoring algorithm proceeds as follows:

  1. Choose an initial estimate \(\beta^0\) for $ and a small constant \(\epsilon > 0\)

For \(t = 0, 1, \ldots\), do:

  1. Evaluate \(\eta^t\), \(W^t\) and \(z^t\) at \(\beta^t\).
  2. Calculate \(\beta^{t+1} = [X^{\top}W^tX]^{-1} X^{\top} W^t[\eta^{t}+z^t]\).
  3. If \(||\beta^{t+1}-\beta^t|| > \epsilon\) then set \(t\to t+1\) and go to 2.
  4. Use \(\beta^{t+1}\) as the value for \(\hat\beta\).

As this algorithm involves iteratively minimising a weighted sum of squares, it is also known as iteratively (re)weighted least squares.

Recall that the canonical link function is \(g(\mu) = b'^{-1}(\mu)\) and with this link \(\eta_i=g(\mu_i)=\theta_i\). Then, \[ {1\over{g'(\mu_i)}} ={{\partial\mu_i}\over{\partial\theta_i}}=b''(\theta_i)\qquad i = 1, \ldots, n. \] As a result, \(\mathop{\mathrm{var}}(Y_i)g'(\mu_i)=\phi_i\), which does not depend on \(\beta\). It follows that \({\partial\over{\partial\beta_j}}\left[{{x_{ik}}\over{\mathop{\mathrm{var}}(Y_i)g'(\mu_i)}}\right]=0\) \((j = 0, \ldots, p)\). It follows that \(H(\beta) = -I(\beta)\) and, for the canonical link, Newton-Raphson and Fisher scoring are equivalent.

Exercise 2.3 (IWLS for the normal linear model) For the normal linear model, show that \(w_i = \sigma^{-2}\) and \(z_i = y_i - \eta_i\) \((i = 1, \ldots, n\)). Hence, show that the Fisher scoring algorithm converges in a single iteration, from any starting point, to the usual least squares estimate.

2.3 Inference

Subject to standard regularity conditions, \({\cal I}(\beta)^{1/2}(\hat\beta - \beta)\) is asymptotically normally distributed with mean \(0\) and variance covariance matrix \(I_p\). So, we can treat the normal distribution with mean \(\beta\) and variance \({\cal I}(\beta)^{-1}\) as the approximate distribution of \(\hat\beta\).

Hence, standard errors can be estimated as \[ [{\cal I}(\hat{\beta})^{-1}]_{kk}^{{1\over 2}} = [(X^{\top}\hat{W}X)^{-1}]_{kk}^{{1\over 2}} \qquad (k = 0, \ldots, p)\,, \]

where \(\hat{W}\) is \(W\) evaluated at \(\hat\beta\) and \(\hat\phi_i\), if \(\mathop{\mathrm{var}}(Y_i)\) depends on an unknown dispersion parameter. Section 2.5 discusses the estimation of \(\phi_i\) in models with unknown dispersion parameter.

The asymptotic distribution of the maximum likelihood estimator can be used to provide asymptotically valid confidence intervals, and hypeotheris testing procedures, using \[ {{\hat{\beta}_k-\beta_k}\over{[(X^{\top}\hat{W}X)^{-1}]_{kk}^{{1\over 2}}}} \;\;{\buildrel{\rm asymp}\over\sim}\;\; {\rm N}(0,1) \, . \]

2.4 Comparing generalized linear models

This section describes just one method for comparing models. General principles and other methods will be discussed in detail in the APTS module itself.

As with linear models, we can proceed by comparing nested models \(H_0\) and \(H_1\) using a generalized likelihood ratio test. Nested means that \(H_0\) and \(H_1\) are based on the same exponential family, have the same link function, but \(\Theta^{(0)}\), the set of values of the canonical parameter \(\theta\) allowed by \(H_0\), is a subset of \(\Theta^{(1)}\), the set of values allowed by \(H_1\).

Without loss of generality, we can think of \(H_1\) as the model \[ \eta_i=\sum_{j=0}^p x_{ij} \beta_j \qquad (i = 1, \ldots, n)\,, \] and \(H_0\) is the same model with \(\beta_{q+1} = \beta_{q+2} = \cdots = \beta_p = 0\).

Now, the log likelihood ratio statistic for a test of \(H_0\) against \(H_1\) is \[ \begin{aligned} L_{01} & \equiv 2\log \left({{\max_{\theta\in \Theta^{(1)}}f_{Y}(y;\theta)}\over {\max_{\theta\in \Theta^{(0)}}f_{Y}(y;\theta)}}\right) \\ &=2\log f_{Y}(y;\hat{\theta}^{(1)})-2\log f_{Y}(y;\hat{\theta}^{(0)}) \end{aligned} \tag{2.9}\] where \(\hat{\theta}^{(1)}\) and \(\hat{\theta}^{(0)}\) result from \(b'(\hat{\theta}_i^{(j)})=\hat{\mu}_i^{(j)}\), \(g(\hat{\mu}_i^{(j)})=\hat{\eta}_i^{(j)}\) \((i = 1, \ldots, n)\) where \(\hat{\eta}^{(j)}\) is the linear predictor evaluated at the maximum likelihood estimate for \(\beta\) under hypothesis \(H_j\) \((j = 0, 1)\). Here, we assume that \(\phi_i\) \((i = 1, \ldots, n\)) are known; the case of unknown \(\phi\) is discussed in Section 2.5.

We reject \(H_0\) in favour of \(H_1\) when \(L_{01} > k\) where \(k\) is determined by the size \(\alpha\) of the test. Under \(H_0\), \(L_{01}\) has an asymptotic chi-squared distribution with \(p - q\) degrees of freedom.

The saturated model is defined to be the model where the canonical parameters \(\theta\) (or equivalently \(\mu\) or \(\eta\)) are unconstrained, and the parameter space is \(n\)-dimensional. For the saturated model, we can calculate the maximum likelihood estimators \(\hat{\theta}\) directly from their likelihood (2.1) by differentiating with respect to \(\theta_1, \ldots , \theta_n\) to give \[ {\partial\over{\partial\theta_k}}\log f_{Y}(y;\theta)={{y_k-b'(\theta_k)} \over{\phi_k}}\qquad k=1,\ldots ,n. \] Therefore \(b'(\hat{\theta}_k) = y_k\) \((k=1, \ldots, n)\), and, hence, \(\hat{\mu}_k = y_k\) \((k = 1, \ldots, n)\). Hence, the saturated model fits the data perfectly, as the fitted values \(\hat{\mu}_k\) and observed values \(y_k\) are the same for every observation. The saturated model is rarely of any scientific interest in its own right. It is overly parameterized, having as many parameters as there are observations. However, every other model is necessarily nested in the saturated model, and a test comparing a model \(H_0\) against the saturated model can be interpreted as a goodness of fit test. If there is no significant evidence that the saturated model — which fits the observed data perfectly — provides a better fit than model \(H_0\), we can conclude that \(H_0\) is an acceptable fit to the data.

From (2.9), the log likelihood ratio statistic for a test of \(H_0\) against the saturated alternative is \[ L_{0}=2\log f_{Y}(y;\hat{\theta}^{(s)})-2\log f_{Y}(y;\hat{\theta}^{(0)}) \]
where \(\hat{\theta}^{(s)}\) is such that \(b'(\hat{\theta}^{(s)}) = y\). However, calibrating \(L_0\) is not straightforward. In some circumstances (typically those where the response distribution might be adequately approximated by a normal) \(L_{0}\) has an asymptotic chi-squared distribution with \(n - q - 1\) degrees of freedom, under \(H_0\). Large values of \(L_{0}\) is evidence against \(H_0\) as a plausible model for the data. However, in other situations, for example Bernoulli response distributions, the \(\chi^2\) approximation to \(L_{0}\) may be poor.

The degrees of freedom of model \(H_0\) and for this test is \(n - q - 1\), whicn is the number of observations minus the number of linear parameters of \(H_0\). We call \(L_{0}\) the scaled deviance of model \(H_0\).

From (2.3) and (2.9) we can write the scaled deviance of model \(H_0\) as \[ L_{0}=2\sum_{i=1}^n{{y_i[\hat{\theta}^{(s)}_i-\hat{\theta}^{(0)}_i] -[b(\hat{\theta}^{(s)}_i)-b(\hat{\theta}^{(0)}_i)]} \over{\phi_i}} \, , \tag{2.10}\] which is easily computed using the observed data, provided that \(\phi_i\) \((i = 1, \ldots, n)\) is known.

Remark 2.1. The log likelihood ratio statistic (2.9) for testing \(H_0\) against a non-saturated alternative \(H_1\) can be written as \[ \begin{aligned} L_{01} &= 2\log f_{Y}(y;\hat{\theta}^{(1)}) - 2\log f_{Y}(y;\hat{\theta}^{(0)}) \\ & = [2\log f_{Y}(y;\hat{\theta}^{(s)}) - 2\log f_{Y}(y;\hat{\theta}^{(0)})] - [2\log f_{Y}(y; \hat{\theta}^{(s)}) - 2\log f_{Y}(y;\hat{\theta}^{(1)})] \\ & = L_{0} - L_{1} \, . \end{aligned} \tag{2.11}\] The log likelihood ratio statistic for comparing two nested models is the difference between their scaled deviances. Furthermore, as $ p - q = (n - q - 1) - (n - p - 1)$, that is the degrees of freedom for the test is the difference in degrees of freedom of the two models.

Remark 2.2. An alternative goodness of fit statistic for a model \(H_0\) is Pearson’s \(X^2\) given by \[ X^2 = \sum_{i=1}^n {{(y_i - \hat{\mu}_i^{(0)})^2}\over{\widehat{\mathop{\mathrm{var}}(Y_i)}}} \,. \tag{2.12}\] \(X^2\) is small when the squared differences between observed and fitted values (scaled by variance) is small. Hence, large values of \(X^2\) correspond to poor fitting models. In fact, \(X^2\) and \(L_{0}\) are asymptotically equivalent. However, the asymptotic \(\chi^2_{n-q-1}\) approximation associated with \(X^2\) is often more reliable.

2.5 Models with an unknown dispersion parameter

Model comparison

Thus far, we have assumed that \(\phi_1, \ldots, \phi_n\) are known. This is the case for both the Poisson and Bernoulli distributions, where \(\phi_i = 1\). When \(\phi_i\) are not known, we can evaluate neither the scaled deviance (2.10) nor the Pearson \(X^2\) statistic (2.12), and hence we cannot directly construct inferences based on them, or compare models using (2.11).

Progress can be made if we assume that \(\phi_i = \sigma^2 / m_i\) \((i = 1, \ldots, n)\), where \(\sigma^2\) is a common unknown dispersion parameter and \(m_1, \ldots , m_n\) are known weights (this form is present in a normal linear model, where \(\mathop{\mathrm{var}}(Y_i) = \sigma^2\)). Under this assumption \[ \begin{aligned} L_{0} &= {2\over\sigma^2}\sum_{i=1}^n \left[m_iy_i(\hat{\theta}^{(s)}_i-\hat{\theta}^{(0)}_i) - m_i \{b(\hat{\theta}^{(s)}_i) - b(\hat{\theta}^{(0)}_i)\}\right] \\ &\equiv{1\over\sigma^2} D_{0} \,, \end{aligned} \tag{2.13}\] where \(D_{0}\) can be calculated using the observed data. We call \(D_{0}\) the deviance of the model.

In order to compare nested models \(H_0\) and \(H_1\), one might calculate the test statistic \[ F = {{L_{01}/(p-q)}\over{L_{1}/(n-p-1)}}={{(L_{0}-L_{1})/(p-q)}\over{L_{1}/(n-p-1)}} ={{(D_{0}-D_{1})/(p-q)}\over{D_{1}/(n-p-1)}}. \tag{2.14}\] This statistic does not depend on the unknown dispersion parameter \(\sigma^2\), so it can be calculated using the observed data. Asymptotically, under \(H_0\), \(L_{01}\) has a \(\chi^2_{p-q}\) distribution and \(L_{01}\) and \(L_{1}\) are independent (not proved here). Assuming that \(L_1\) has an approximate \(\chi^2_{n-p-1}\) distribution, then \(F\) has an approximate F distribution with \(p-q\) degrees of freedom in the numerator and \(n-p-1\) degrees of freedom in the denominator. Hence, large values of \(F\) is evidence against \(H_0\) in favour of \(H_1\).

Inference about model parameters

The dependence of the maximum likelihood equations \(u(\hat{\beta}) = 0\) on \(\sigma^2\) (where \(u\) is given by (2.4)) can be eliminated by multiplying through by \(\sigma^2\). However, inference based on the maximum likelihood estimates requires knowledge of \(\sigma^2\). This is because asymptotically \(\mathop{\mathrm{var}}(\hat{\beta})\) is the inverse of the Fisher information matrix \({\cal I}(\beta)=X^{\top}WX\), and this depends on \(w_i=1/\{\mathop{\mathrm{var}}(Y_i)g'(\mu_i)^2\}\) where \(\mathop{\mathrm{var}}(Y_i) = \phi_ib''(\theta_i) = \sigma^2 b''(\theta_i) / m_i\).

To calculate standard errors and confidence intervals, we need to supply an estimate \(\hat{\sigma}^2\) of \(\sigma^2\). Despite that the maximum likelihood estimator of \(\sigma^2\) is well-defined, it is more common to base an estimator of \(\sigma^2\) on the Pearson \(X^2\) statistic. The variance of \(Y_i\) can be written as \(\mathop{\mathrm{var}}(Y_i) = \phi_i V(\mu_i) = \sigma^2 V(\mu_i) / m_i\), where \(V(\mu_i) = b''(\theta_i)\) and \(\theta_i = {b'}^{-1}(\mu_i)\) (see Section 2.1). Hence, (2.12) can be written as \[ X^2={1\over\sigma^2} \sum_{i=1}^n {{m_i(y_i-\hat{\mu}_i)^2}\over{{V}(\hat{\mu}_i)}}. \tag{2.15}\]

Exercise 2.4 Suppose that \(H_0\) is an adequate fit and that \(X^2\) has an chi-squared distribution with \(n-q-1\) degrees of freedom.

  1. Show that \[ \hat{\sigma}^2 = {1\over{n-q-1}} \sum_{i=1}^n {{m_i(y_i-\hat{\mu}_i)^2}\over{{V}(\hat{\mu}_i)}} \] is an approximately unbiased estimator of \(\sigma^2\).

  2. Suggest an alternative estimator based on the deviance \(D_0\).

2.6 Residuals and Model Checking

Recall that for linear models, we define the residuals to be the differences between the observed and fitted values \(y_i - \hat{\mu}_i\) \((i = 1, \ldots, n)\). In fact, both the scaled deviance and Pearson \(X^2\) statistic for a normal linear model (which is a GLM with normal distirbution and identity link function) are the sum of the squared residuals divided by \(\sigma^2\). We can build on that observation to defined residuals for other generalized linear models in a natural way.

For any GLM, we define the Pearson residuals to be \[ e^P_i= {{y_i-\hat{\mu}_i}\over{\widehat{Var(Y_i)}^{1/2}}} \qquad (i = 1, \ldots, n)\,. \] Then, from (2.12), the statistic \(X^2\) is the sum of the squared Pearson residuals.

For any GLM, we define the deviance residuals to be \[ e^D_i={\rm sign}(y_i-\hat{\mu}_i) \left[{{y_i(\hat{\theta}^{(s)}_i-\hat{\theta}_i) -\{b(\hat{\theta}^{(s)}_i)-b(\hat{\theta}_i)\}} \over{\phi_i}}\right]^{1/2} \qquad (i = 1, \ldots, n)\,, \] where \({\rm sign}(x) = 1\) if \(x>0\) and \(-1\) if \(x<0\). Then, from (2.10), the scaled deviance, \(L_{0}\), is the sum of the squared deviance residuals.

When \(\phi_i = \sigma^2 / m_i\) and \(\sigma^2\) is unknown, as in Section 2.5, the expressions above are typically multiplied through by \(\sigma^2\) to eliminate dependence on the unknown dispersion parameter.

So, for a normal GLM the Pearson and deviance residuals are both equal to the usual residuals, \(y_i - \hat{\mu}^{(0)}_i,\; i = 1, \ldots, n\).

Both the Pearson and deviance residuals can be standardized by dividing through by \((1 - h_{ii})^{1/2}\), as in Section 1.4. If the model is adequate, the derived residuals \[ r_i^* = r^D_i + \frac{1}{r^D_i} \log\frac{r^P_i}{r^D_i} \] are close to normal for a wide range of GLMs, where \(r^D_i\) and \(r^P_i\) are the standardized deviance and Pearson residuals, respectively.

Checking GLMs using residuals is based on the same kind of diagnostic plots suggested for linear models in Section 1.7. Similarly, the Cook’s distance \(C_j\) for linear models can be adapted for GLMs by using Pearson residuals.