Likelihood and p-value functions in the composite likelihood context

Similar documents
Statistica Sinica Preprint No: SS

ASSESSING A VECTOR PARAMETER

i=1 h n (ˆθ n ) = 0. (2)

DEFNITIVE TESTING OF AN INTEREST PARAMETER: USING PARAMETER CONTINUITY

PARAMETER CURVATURE REVISITED AND THE BAYES-FREQUENTIST DIVERGENCE.

The formal relationship between analytic and bootstrap approaches to parametric inference

Measuring nuisance parameter effects in Bayesian inference

Principles of Statistical Inference

Principles of Statistical Inference

Likelihood inference in the presence of nuisance parameters

Recap. Vector observation: Y f (y; θ), Y Y R m, θ R d. sample of independent vectors y 1,..., y n. pairwise log-likelihood n m. weights are often 1

Integrated likelihoods in models with stratum nuisance parameters

Various types of likelihood

Theory of Maximum Likelihood Estimation. Konstantin Kashin

Composite Likelihood Estimation

Time Series Analysis Fall 2008

Likelihood Inference in the Presence of Nuisance Parameters

Composite Likelihood

Likelihood Inference in the Presence of Nuisance Parameters

A Very Brief Summary of Statistical Inference, and Examples

Approximating models. Nancy Reid, University of Toronto. Oxford, February 6.

Nancy Reid SS 6002A Office Hours by appointment

MISCELLANEOUS TOPICS RELATED TO LIKELIHOOD. Copyright c 2012 (Iowa State University) Statistics / 30

Default priors for Bayesian and frequentist inference

Statistical Methods for Handling Incomplete Data Chapter 2: Likelihood-based approach

Conditional Inference by Estimation of a Marginal Distribution

1 One-way analysis of variance

ICES REPORT Model Misspecification and Plausibility

Topic 12 Overview of Estimation

COMBINING p-values: A DEFINITIVE PROCESS. Galley

Nancy Reid SS 6002A Office Hours by appointment

DA Freedman Notes on the MLE Fall 2003

ESTIMATING THE MEAN LEVEL OF FINE PARTICULATE MATTER: AN APPLICATION OF SPATIAL STATISTICS

FREQUENTIST BEHAVIOR OF FORMAL BAYESIAN INFERENCE

Fall 2017 STAT 532 Homework Peter Hoff. 1. Let P be a probability measure on a collection of sets A.

Accurate directional inference for vector parameters

A Very Brief Summary of Statistical Inference, and Examples

Accurate directional inference for vector parameters

Asymptotic Multivariate Kriging Using Estimated Parameters with Bayesian Prediction Methods for Non-linear Predictands

Default priors and model parametrization

Composite likelihood methods

MA 575 Linear Models: Cedric E. Ginestet, Boston University Mixed Effects Estimation, Residuals Diagnostics Week 11, Lecture 1

Bootstrap and Parametric Inference: Successes and Challenges

Likelihood based Statistical Inference. Dottorato in Economia e Finanza Dipartimento di Scienze Economiche Univ. di Verona

Paradoxical Results in Multidimensional Item Response Theory

Second order ancillary: A differential view from continuity

MA 575 Linear Models: Cedric E. Ginestet, Boston University Non-parametric Inference, Polynomial Regression Week 9, Lecture 2

1. Fisher Information

Frequentist-Bayesian Model Comparisons: A Simple Example

Lasso Maximum Likelihood Estimation of Parametric Models with Singular Information Matrices

Likelihood and Fairness in Multidimensional Item Response Theory

Bayesian and frequentist inference

Master s Written Examination

Semiparametric Gaussian Copula Models: Progress and Problems

Location-adjusted Wald statistics

SMOOTHED BLOCK EMPIRICAL LIKELIHOOD FOR QUANTILES OF WEAKLY DEPENDENT PROCESSES

Accounting for Complex Sample Designs via Mixture Models

Applied Asymptotics Case studies in higher order inference

Physics 403. Segev BenZvi. Propagation of Uncertainties. Department of Physics and Astronomy University of Rochester

σ(a) = a N (x; 0, 1 2 ) dx. σ(a) = Φ(a) =

Accurate directional inference for vector parameters in linear exponential families

PRINCIPLES OF STATISTICAL INFERENCE

Integrated likelihoods in survival models for highlystratified

STATS 200: Introduction to Statistical Inference. Lecture 29: Course review

Recall that in order to prove Theorem 8.8, we argued that under certain regularity conditions, the following facts are true under H 0 : 1 n

Semiparametric Gaussian Copula Models: Progress and Problems

A test for improved forecasting performance at higher lead times

Theory and Methods of Statistical Inference. PART I Frequentist theory and methods

Modern Likelihood-Frequentist Inference. Donald A Pierce, OHSU and Ruggero Bellio, Univ of Udine

Australian & New Zealand Journal of Statistics

Session 3 The proportional odds model and the Mann-Whitney test

Statistics & Data Sciences: First Year Prelim Exam May 2018

Introduction to Estimation Methods for Time Series models Lecture 2

Group Sequential Tests for Delayed Responses. Christopher Jennison. Lisa Hampson. Workshop on Special Topics on Sequential Methodology

A Few Notes on Fisher Information (WIP)

Maximum Likelihood Estimation

Approximate Likelihoods

Improved Inference for First Order Autocorrelation using Likelihood Analysis

Maximum Likelihood Estimation

Bayesian Interpretations of Heteroskedastic Consistent Covariance Estimators Using the Informed Bayesian Bootstrap

For more information about how to cite these materials visit

401 Review. 6. Power analysis for one/two-sample hypothesis tests and for correlation analysis.

ANCILLARY STATISTICS: A REVIEW

Optimization. The value x is called a maximizer of f and is written argmax X f. g(λx + (1 λ)y) < λg(x) + (1 λ)g(y) 0 < λ < 1; x, y X.

Maximum Likelihood Tests and Quasi-Maximum-Likelihood

Fractional Hot Deck Imputation for Robust Inference Under Item Nonresponse in Survey Sampling

Data Analysis and Uncertainty Part 2: Estimation

Approximate Inference for the Multinomial Logit Model

Flat and multimodal likelihoods and model lack of fit in curved exponential families

Normalising constants and maximum likelihood inference

Lecture 3 September 1

Pseudo-score confidence intervals for parameters in discrete statistical models

A Model for Correlated Paired Comparison Data

Multiple Linear Regression

Robust covariance estimator for small-sample adjustment in the generalized estimating equations: A simulation study

Likelihood and Asymptotic Theory for Statistical Inference

Third-order inference for autocorrelation in nonlinear regression models

ASYMPTOTICS AND THE THEORY OF INFERENCE

The Slow Convergence of OLS Estimators of α, β and Portfolio. β and Portfolio Weights under Long Memory Stochastic Volatility

Stat 5102 Final Exam May 14, 2015

Transcription:

Likelihood and p-value functions in the composite likelihood context D.A.S. Fraser and N. Reid Department of Statistical Sciences University of Toronto November 19, 2016 Abstract The need for combining likelihood information arises widely in the analysis of complex models, and also in meta-analysis where information is to be combined from various studies. We work to first order and show that full first order accuracy for combining scalar or vector parameter information is available from likelihood analysis based on score variables. 1 Introduction Statistical models presented in the form of a family of densities {f(y; θ); θ Θ R p )} are usually analyzed using the likelihood function L(θ) f(y; θ), or equivalently the log-likelihood function l(θ) = log{l(θ)}. Evaluated at the observed data this provides all the data-dependent information for a standard Bayesian analysis, and almost all the data-dependent information for a typical frequentist-based analysis where full thirdorder inference requires the addition of a sample-space derivative. In some modelling situations however the full density version of the model may not be available. In such cases work-arounds have been developed, using for example marginal models for single coordinates, or marginal models for pairs of coordinates 1

or other variants called pseudo or composite likelihood functions, studied in Lindsay (1988) and reviewed in Varin et al. (2011). For example, if we have a statistical model f(y 1,..., y d ; θ), the composite pairwise log-likelihood is l pair (θ) = log{f 2 (y r, y s ; θ)}, r<s where f 2 (y r, y s ; θ) is the marginal model for the pair of components (y r, y s ). We can also express this as m l pair (θ) = l i (θ), where m = d(d 1)/2, and l i (θ) = log{f 2 (y r, y s ; θ)}, with i indexing the pairs (r, s) so that y i is short-hand for some pair of components of the vector y = (y 1,..., y d ). If we were considering what is called the independence likelihood we would have l ind (θ) = Σ r log{f 1 (y r ; θ)} = Σ i l i (θ) with l i (θ) = log{f 1 (y i ; θ)} and m = d. Our approach here is to suppose we have available component log-likelihood functions, derived from a sample of size n, and to assume that each component has the usual asymptotic properties, such as being O p (n) with the typical smoothness asymptotic behaviour of its cumulants. We let θ 0 be some trial or reference value for the parameter and examine the model in moderate deviations about θ 0. We will see then that to first order the model has simple normal form and is independent in structure of the initial reference value θ 0 given that it is in moderate deviations. This leads to a fully first-order accurate combine d log-likelihood. In addition, the approach leads to a simple first-order accurate procedure for combining p-values. The method uses variances and covariances of score variables, as are also needed for the usual inferential adjustments in the composite likelihood approach. Specifically we expand a component log-likelihood as i=1 l i (θ). = l i (θ 0 ) + (θ θ 0 ) T l i(θ 0 ) + 1 2 (θ θ 0) T l i (θ 0 )(θ θ 0 ). and work to first order so that θ θ 0 = O(n 1/2 ); in this case the first term on the right side above is O p (n), the second two terms are O p (1), and the neglected terms are O p (n 1/2 ). In addition, to this order, we can replace the observed second derivative 2

matrix with its expected value. Thus each component log-likelihood depends on the data only through the p 1 score vector s i = 1 l i (θ; y i ) n θ, θ=θ0 standardized here to be O p (1). The Bartlett identities for the component models give E(s i ; θ 0 ) = 0, var(s i ; θ 0 ) = E{ 2 l(θ; y i ) θ θ ; θ} = v ii, (1) T θ0 where v ii is the p p expected Fisher information matrix, from the ith component loglikelihood, standardized to be O(1). This standardization and our assumed continuity also give (Cox & Hinkley, 1974, Ch. 4) E(s i ; θ) = v ii (θ θ 0 ) + O(n 1/2 ), var(s i ; θ) = v ii + O(n 1/2 ). (2) We now stack the score vectors s = (s 1,..., s m ) T and write E(s; θ) = V (θ θ 0 ) + O(n 1/2 ), var(s; θ) = W + O(n 1/2 ), (3) where V = (v 11,..., v mm ) T is the mp p matrix of the stacked v ii, and W is the mp mp matrix with v ii on the diagonal, and off-diagonal matrix elements v ij = cov(s i, s j ). This first order version of the model is structurally free of the reference value θ 0. 2 Full first-order log-likelihood function We now analyze the statistical model in its full first order accurate form, and for temporary notational convenience write just θ for the departure θ θ 0 from the trial value. From the preceding summary we can then write the model for s as a normal theory regression model, s = V θ + e 3

where e N(0, W ). From this full first-order approximation we have directly a firstorder log-likelihood function l (θ) = c 1 2 (s V θ)t W 1 (s V θ), (4) = c 1 2 θt V T W 1 V θ + θ T V T W 1 s. (5) The score function from (5) is s (θ) = V T W 1 (s V θ) with mean 0 and variance V T W 1 V ; and the maximum likelihood value is ˆθ = (V T W 1 V ) 1 V T W 1 s, with mean θ and variance (V T W 1 V ) 1 = W., The new log-likelihood function can also be rewritten in the equivalent form l (θ) = c 1 2 (θ ˆθ ) T V T W 1 V (θ ˆθ ) = c 1 2 (θ ˆθ ) T W 1 (θ ˆθ ); (6) this makes the location form of the log-likelihood more transparent. If θ is scalar, then l (θ) = c 1 2 V T W 1 V θ 2 + V T W 1 sθ, (7) and the individual components are l i (θ) = c i (1/2)(s i v ii θ) 2 v 1 ii which can be rearranged giving sθ = l(θ) + (1/2)V θ 2 + c where l(θ) = {l 1 (θ),..., l m (θ)} T is the vector of component log-likelihood functions. Then substituting for sθ in (7) gives an expression for l (θ) as a weighted combination of component log-likelihood functions, with related score function s (θ) = V T W 1 (s V θ). l (θ) = V T W 1 l(θ), (8) The original component loglikelihoods, rather than their Taylor series approximations, can be used here to the same order, and thus (8) provides an optimally weighted combination of component log-likelihoods. While this makes (8) somewhat attractive compared with (7), and agrees with it up to quadratic terms, there are implicit cubic terms in (8) so inferences 4

from the two versions will be different in finite samples, though equivalent to the order to which we are working. The linear combination of (8) is not generally available for vector parameters; different combinations of components are needed for different coordinates of the parameter, as indicated by different rows in the matrix V T in (7). Lindsay (1988) studied the choice of weights in scalar composite likelihood by seeking an optimally weighted combination of score functions l i (θ)/ θ, in his notation S i (θ); the optimal weights depend on θ. Our approach is to work within n 1/2 - neighbourhoods of a reference parameter value, and obtain the first-order model for the observed variables s i, thus leading directly to the full first-order log-likelihood. The use of a quadratic form in a set of estimating equations is used in indirect inference in econometrics, and is called an indirect likelihood in Jiang and Turnbull (2004). For a scalar or vector parameter of interest ψ of dimension r, with nuisance parameter λ so that θ = (ψ T, λ T ) T, we can then see directly that the first-order log-likelihood function for the component ψ is just l (ψ) = c 1 2 (ψ ˆψ ) T W ψψ (ψ ˆψ ) (9) where W ψψ is the ψψ component of the information W 1 and the maximum loglikelihood ˆψ = ψ(ˆθ) is obtained from the maximizing value ˆθ for the full log-likelihood at (5). For the scalar component, Pace et al. (2016) consider the use of profile loglikelihood components. 3 Illustrations The first illustrations use latent independent normal variables with mean θ and variance 1, as this captures the essential elements and makes clear the role of correlation in the re-weighting: the basic underlying densities are assumed to be independent responses x from a N(θ, 1) distribution, with corresponding log-likelihood θ 2 /2 + θx. Example 1: Independent components. Consider component variables y 1 = a 1 x 1 and y 2 = a 2 x 2. Then m = 2 and the component log-likelihoods are l i (θ; y i ) = (y i a i θ) 2 /2a 2 i = θ 2 /2 + y i θ/a i giving s i = y i /a i, V = (1, 1) T, and W = diag(1, 1), which 5

lead to V T W 1 = (1, 1). Thus l (θ) = l 1 (θ) + l 2 (θ) is the independence log-likelihood, as would be expected. Example 2: Dependent and exchangeable components. Consider y 1 = x 1 + x 2 and y 2 = x 1 + x 3. The component log-likelihood functions are l i (θ) = θ 2 + θy i giving s i = y i, V = (2, 2) T, W = ( 2 1 1 2 These lead to the combined first-order log-likelihood function ), and V T W 1 = (2/3, 2/3). (10) l (θ) = (2/3, 2/3) T l(θ) = 4 3 θ2 + 2 3 θ(y 1 + y 2 ). (11) In contrast the unadjusted composite log-likelihood obtained by adding the marginal log-likelihoods is l UCL (θ) = 2θ 2 + θ(y 1 + y 2 ), with score variable y 1 + y 2, which has variance 6, but second derivative 4: the second Bartlett identity does not hold and l UCL (θ) is not a proper log-likelihood. The usual method to adjust this is to rescale the unadjusted log-likelihood l ACL = al UCL giving 2aθ 2 + θa(y 1 + y 2 ) with negative second derivative 4a 2 and score variance 6a 2. These latter become equal with a = 2/3, which gives the adjusted composite log-likelihood l ACL (θ) = 2 3 l UCL(θ) = 4 3 θ2 + 2 3 θ(y 1 + y 2 ), which agrees with l (θ). The next example shows that this agreement does not hold generally. Example 3: Dependent but not exchangeable components. Now let y 1 = x 1 and y 2 = x 1 + x 3. The individual log-likelihoods are l 1 (θ) = θ 2 /2 + θy 1 and l 2 (θ) = θ 2 + θy 2 with s 1 = y 1 and s 2 = y 2. We then have V T = (1, 2), the off-diagonal elements of W equal to 1, and V T W 1 = (0, 1), leading to l (θ) = θ 2 + θy 2, (12) with maximum likelihood estimate ˆθ = y 2 /2, which has variance 1/2. The new likelihood shows reflects that y 2 = x 1 + x 3 provides full first-order information concerning θ from the effective data y 2. 6

In contrast, the unadjusted composite likelihood is l UCL (θ) = (3/2)θ 2 +θ(y 1 +y 2 ), with associated maximum likelihood estimate ˆθ UCL = (y 1 + y 2 )/3. The rescaling factor to adjust for the Bartlett property is a = 3/5 giving the adjusted composite likelihood l ACL (θ) = 3 5 l UCL(θ) = 9 10 θ2 + 3 5 θ(y 1 + y 2 ), which is again maximized at (y 1 +y 2 )/3. Although the second Bartlett identity has been implemented, the adjusted composite likelihood leads to the same inefficient estimate of θ as the unadjusted version, while the new log-likelihood l (θ) provides the proper information 2 and estimate y 2 /2. Some discussion related to this point is given in Freedman (2006). An asymptotic version of these two illustrations is obtained by having n replications of y 1 and y 2, or equivalently assuming x 1, x 2 and x 3 have variances 1/n instead of 1. Example 4: Bivariate Normal. Suppose we have n pairs (y i1, y i2 ) independently distributed as bivariate normal with mean vector (θ, θ) and a known covariance matrix. The sufficient statistic is (ȳ.1, ȳ.2 ), and the component log-likelihood functions are taken as those from the marginal densities of ȳ.1 and ȳ.2, so that l 1 (θ) = n(ȳ.1 θ) 2 /(2σ1) 2 and l 2 (θ) = n(ȳ.2 θ) 2 /(2σ2). 2 The score components s 1 and s 2 are, respectively, n(ȳ.1 θ)/σ1 2 and n(ȳ.2 θ)/σ2, 2 with variance-covariance matrix ( ) 1/σ1 2 ρ/(σ 1 σ 2 ) W = n, (13) ρ/(σ 1 σ 2 ) 1/σ2 2 giving V T W 1 = (1 ρ 2 ) 1 (1 ρσ 1 /σ 2, 1 ρσ 2 /σ 1 ), leading to { ) 2 l n (ȳ.1 θ (θ) = (1 ρ σ 1 ) + 2(1 ρ 2 ) σ 2 σ 1 (ȳ.2 θ σ 2 ) 2 (1 ρ σ 2 σ 1 ) }. (14) As a function of θ this can be shown to be equivalent to the full log-likelihood based on the bivariate normal distribution of (ȳ.1, ȳ.2 ), and leads to an estimate of θ that is a weighted combination of ȳ.1 and ȳ.2. 7

Example 5: Two parameters. Suppose that x i follow a N(θ 1, 1) distribution, and independently z i follow a N(θ 2, 1) distribution. We base our component log-likelihoods on the densities of the vectors ( y 1 = x 1 z 2 + z 3 giving the score variables s 1 = (y 11, y 12 ) T 0 1 ) 0 1 0 1, y 2 = ( x 1 + x 3 z 2 ), and s 2 = (y 21, y 22 ) T. The needed variances and covariances are: 1 0 1 0 1 0 2 0 1 0 0 2 V = 2 0, W = 0 2 0 1 1 0 2 0, W 1 0 1 0 1 = 1 0 1 0, 0 1 0 2 and then for (6) we have V T W 1 V = diag(2, 2). This gives the new log-likelihood l (θ 1, θ 2 ) = (θ 1 ˆθ 1) 2 (θ 2 ˆθ 2) 2, which represents the log-likelihood from s 12 plus the log-likelihood from s 21 as we might reasonably have expected from the presentations in terms of the latent x i and z i variables. Meanwhile, the usual composite log-likelihood derived from the sum of those for s 1 and s 2 has extra terms. Example 6. Two parameters, without symmetry. Within the structure of the previous example, suppose our component vectors are ( ) ( ) x 1 x 1 + x 2 y 1 =, y 2 =, z 2 z 1 where the first coordinates depend on θ 1 and the second coordinates depend on θ 2 ; the corresponding score variables are s 1 = (y 11, y 12 ) T 0 1 0 0 0 1 and s 2 = (y 21, y 22 ) T. The variances and covariances again are directly available, and we have 1 0 1 0 1 0 2 0 1 0 0 1 V = 2 0, W = 0 1 0 0 1 0 2 0, W 1 0 1 0 0 = 1 0 1 0. 8 0 0 0 1

From Example 3 we can see that for inference about θ 1 the weights are (0, 1) and from Example 1 the weights for θ 2 are (1, 1); clearly these are incompatible. For the direct approach using (7) we have V T W 1 V = diag(2, 2), giving l (θ 1, θ 2 ) = (θ 1 ˆθ 1) 2 (θ 2 ˆθ 2) 2 This simple sum of component likelihoods for (θ 1, θ 2 ) is to be expected as the measurements concerning θ 1 are independent of those for θ 2 ; in addition, all the information concerning θ 1 comes from x 1 + x 2 as the first coordinate of y 2, and all for θ 2 comes from z 1 and z 2 in the second coordinates of y 1 and y 2. 4 Comparison to composite likelihood Composite likelihood combines information from different components, often by adding the log-likelihood functions, but care is needed in constructing inference from the resulting function, as the curvature at the maximum does not give an accurate reflection of the precision. Corrections for this in the scalar parameter setting involve either rescaling the composite log-likelihood function, or accommodating the dependence among the components in the estimate of the variance of the composite likelihood estimator. In the vector parameter setting adjustments to the composite log-likelihood function are more complex than a simple rescaling; see Pace et al. (2011). This rescaling is not enough: the location of the composite log-likelihood function is incorrect to first order, and confidence intervals so obtained are not correctly located to first order. This is corrected by the the use of l (θ) from Section 2. As we are using only first-order log-likelihood functions, it suffices to illustrate this with normal distributions. Suppose y T = (y 1,..., y m ), where the marginal models for the individual coordinates y i are normal with mean θv ii and variance v ii, and cov(y i, y j ) = v ij, all elements of the matrix W. The unadjusted composite log-likelihood function is l UCL (θ) = 1 2 θ2 Σ m i=1v ii + Σ m i=1y i θ. with maximum likelihood estimate ˆθ CL = Σy i /Σv ii and curvature at the maximum point Σv ii : this curvature is not the inverse variance of ˆθ CL as the second Bartlett identity does not hold. 9

As indicated in Example 2, the rescaled version that recovers the second Bartlett identity is l ACL (θ) = H J l UCL(θ) = 1 2 θ2 (Σv ii) 2 Σv ij + θ y i Σv ii Σv ij, where H = E{ l UCL (θ)} and J = var{l UCL (θ)}; in this context neither H nor J depends on θ. The maximum likelihood estimate from this function is the same, ˆθ UCL, but the inverse of the second derivative gives the correct asymptotic variance. What is less apparent is that the location of the log-likelihood function needs a correction. This is achieved using the new weighted version from Section 2: l (θ) = 1 2 θ2 (V T W 1 V ) + θv T W 1 y, which has maximum likelihood estimate ˆθ = (V T W 1 V ) 1 V T W 1 y with variance V T W 1 V equal to the inverse of the second derivative. Also we see that l (θ) has the same linear and quadratic coefficients for θ as the full log-likelihood for the model N(θV, W ). Both l ACL (θ) and l (θ) require variances and covariances of the score variables; see the Appendix. Writing the uncorrected composite log-likelihood as 1 T l(θ), where l(θ) is the vector {l 1 (θ),..., l m (θ)}, and l i (θ) = (1/2)v ii θ 2 + y i θ, we can write var(ˆθ UCL ˆθ ) = 1T W 1 (1 T V ) 2 1 V T W 1 V, expressing the first order loss of efficiency of the composite likelihood estimator relative to that obtained using the full first order log-likelihood. 5 Combining p-value functions The p-value with its Uniform(0,1) distribution under the null provides a very clear presentation p i = Φ{(s i v ii θ)/v 1/2 ii } of statistical inference from the i-th model with data. But an alternative presentation on the z scale using the related value z i = Φ(p i ) gives an equivalent presentation with its Normal(0, 1) distribution under the null; in fact this offers a clearer picture of how extreme an extreme value really is. A third 10

presentation using the slope of the component log-likelihood l i(θ) = s i v ii θ needs an additional root information v 1/2 ii. The simple connections are p i z i = Φ 1 (p i ) (s i v ii θ) = (v ii ) 1/2 z i, p i = Φ(z i ) z i = v 1/2 ii (s i v ii θ) (s i v ii θ); thus we can move back and forth among the three types of presentation. But only one form of presentation has the additivity property that we have been investigating, and that form using score variables s i v ii θ has available the V T W 1 technology for linearly combining log-likelihood information, Now suppose we have available d component log-likelihoods and we wish to combine the related p-values. following (8): The resulting combined score is available from the expression V T W 1 (s V θ) = V T W 1 V 1/2 Φ 1 {p(θ; s)}, (15) where V 1/2 Φ 1 {p(θ; s)} is the vector with coordinates v 1/2 ii Φ 1 {p(θ; s i )}. And then we can convert back obtaining the full first-order p-value for the combined data: p(θ; s) = Φ[(V T W 1 V ) 1/2 V T W 1 V 1/2 Φ 1 {p(θ; s)}]. (16) Example 2 continued: Dependent and exchangeable components. score variable relative to the nominal parameter value θ 0 = 0 is The composite V T W 1 s = 2 3( y1 + y 2 ), which is the score variable from the new log-likelihood l (θ), and the relevant quantile is z = ( ) 1/2 { } 8 2 ( ) 8 y1 + y 2 3 3 3 θ, which has a standard normal distribution, exact in this case. The corresponding composite p-value function is then p(θ; s) = Φ(z). This p-value calculation extracts all the available information concerning θ. 11

Example 3 continued: Dependent but not exchangeable components. The combined score variable relative to the nominal θ 0 = 0 is V T W 1 s = y 2, which is the score variable from the new log-likelihood l (θ). The corresponding quantile is z = 2 1/2 (y 2 2θ) and the corresponding composite p-value function is p(θ; s) = Φ(z) = Φ{2 1/2 (y 2 2θ)}, which is in accord with the general observation that y 2 here provides full information on the parameter θ. This new p-value is fully accurate using all available information. Example 7: Combining three p-values. Suppose three investigations of a common scalar parameter θ have yielded the following p-values for assessing a null value θ 0 : 1.15%, 3.01%.2.31%. For combining these we need the measures of precision as provided by the informations, say, v 11 = 3.0, v 22 = 6.0, v 33 = 9.0. The corresponding z-values and score values s are z 1 = Φ 1 (0.0115) = 2.273 s 1 3θ 0 = 3 1/2 ( 2.273) = 3.938 z 2 = Φ 1 (0.0301) = 1.879 s 2 6θ 0 = 6 1/2 ( 1.879) = 4.603 (17) z 3 = Φ 1 (0.0231) = 1.994 s 3 9θ 0 = 9 1/2 ( 1.994) = 5.981. First suppose for simplicity that the investigations are independent, so that W = diag(v ) and V T W 1 = (1, 1, 1), which says that we should add the scores, as one would expect; this gives the combined score 14.522. Then standardizing by the root of the combined information 18 1/2 = 4.243 we obtain p = Φ( 3.423) = 0.00032. Now to examine the effect of dependence among the scores we consider a crosscorrelation matrix of the form R = 1 ρ ρ ρ 1 ρ ρ ρ 1 with corresponding covariance matrix W having entries 3, 6, 9 on the diagonal and the appropriate covariances otherwise. And then for a low-level of correlation we take ρ = 0.2. The coefficients for combining the scores s i in the array (17) are given in the, 12

array 1 3.000 0.424 0.520 V T W 1 = (3, 6, 9) 0.424 6.000 0.735 0.520 0.735 9.000 0.357 0.042 0.034 = (3, 6, 9) 0.042 0.179 0.024 0.024 0.119 0.119 producing the row vector (0.318, 0.643, 0.787) for combining the scores. The resulting z and p-value are 2.81 and p = 0.0025, a considerable increase from the independence value 0.00032. 6 Conclusion In this paper we use likelihood asymptotics to construct a fully first-order accurate log-likelihood function for the composite likelihood context. It requires the covariance matrix of score variables, which is also needed for inference from composite likelihood. The advantage of the first-order model is that it emphasizes that the model is equivalent under rotation to a sample from a location Normal with known variance. As such there are well substantiated procedures for analyzing a location θ with known variance and available likelihood. The use of the first order model also leads to formulas for combining p-values, by converting them to score variables, then combining them using the vector V T W 1, then converting back to the p-value scale. This gives widely applicable procedures for meta-analysis. The methods of higher-order likelihood have been widely used for developing theory with third order accuracy for the analysis of statistical models with data. We have applied these methods here to the first order using the composite likelihood context, and have achieved corresponding resolutions with full first order accuracy. More generally, this first-order approach is available widely and can be applied more generally, in particular to contexts where estimating equations are used. This converts the related 13

problems into the direct analysis of a first order model, thus simplifying and avoiding the use of estimating equations as an exploratory procedure. Acknowledgements Some portions of this material were presented at the University of Padova and University College London in 2012 and 2013. During the spring semester 2015 at the University of Toronto the material was developed and revised in a likelihood seminar and appreciation and thanks go to participants Wei Lin, Faizan Khalid Mohsin, Xin-Le Sheng, Harris Quach and Feng Xu. This research has received support from the National Science and Engineering Research Council of Canada and the Senior Scholars Funding of York University. References Freedman, D. A. (2006). On the so-called Huber sandwich estimator and Robust standard errors. The American Statistician 60, 299 302. Jiang, W. and B. Turnbull (2004). The indirect method: inference based on intermediate statistics a synthesis and examples. Statistical Science 19, 239 263. Lindsay, B. (1988). Composite likelihood methods. In N. Prabhu (Ed.), Statistical Inference from Stochastic Processes, Volume 80, pp. 221 239. Providence, Rhode Island: American Mathematical Society. Pace, L., A. Salvan, and N. Sartori (2011). Adjusting composite likelihood ratio statistics. Statistica Sinica 21, 129 148. Pace, L., A. Salvan, and N. Sartori (2016). Combining dependent log likelihoods for a scalar parameter of interest. preprint. Varin, C., N. Reid, and D. Firth (2011). An overview of composite likelihood methods. Statistica Sinica 21, 5 42. 14