arxiv: v5 [stat.me] 9 Nov 2015

Similar documents
Estimation of a Two-component Mixture Model

SEMIPARAMETRIC INFERENCE WITH SHAPE CONSTRAINTS ROHIT KUMAR PATRA

Estimation of a Two-component Mixture Model with Applications to Multiple Testing

Estimation and Confidence Sets For Sparse Normal Mixtures

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

STAT 461/561- Assignments, Year 2015

Statistical Applications in Genetics and Molecular Biology

A Large-Sample Approach to Controlling the False Discovery Rate

Controlling the False Discovery Rate: Understanding and Extending the Benjamini-Hochberg Method

A Sequential Bayesian Approach with Applications to Circadian Rhythm Microarray Gene Expression Data

Chapter 9. Non-Parametric Density Function Estimation

Simultaneous Testing of Grouped Hypotheses: Finding Needles in Multiple Haystacks

Doing Cosmology with Balls and Envelopes

Modified Simes Critical Values Under Positive Dependence

Recall the Basics of Hypothesis Testing

Minimum Hellinger Distance Estimation in a. Semiparametric Mixture Model

Asymptotic Statistics-VI. Changliang Zou

Statistical Applications in the Astronomy Literature II Jogesh Babu. Center for Astrostatistics PennState University, USA

The bootstrap. Patrick Breheny. December 6. The empirical distribution function The bootstrap

Statistical testing. Samantha Kleinberg. October 20, 2009

False discovery rate and related concepts in multiple comparisons problems, with applications to microarray data

Math 494: Mathematical Statistics

Testing against a linear regression model using ideas from shape-restricted estimation

Proportion of non-zero normal means: universal oracle equivalences and uniformly consistent estimators

Chapter 9. Non-Parametric Density Function Estimation

Inference For High Dimensional M-estimates. Fixed Design Results

STAT 6350 Analysis of Lifetime Data. Probability Plotting

Multiple Testing. Hoang Tran. Department of Statistics, Florida State University

40.530: Statistics. Professor Chen Zehua. Singapore University of Design and Technology

ESTIMATING THE PROPORTION OF TRUE NULL HYPOTHESES UNDER DEPENDENCE

Sparse Nonparametric Density Estimation in High Dimensions Using the Rodeo

14.30 Introduction to Statistical Methods in Economics Spring 2009

Inference For High Dimensional M-estimates: Fixed Design Results

Optimal Detection of Heterogeneous and Heteroscedastic Mixtures

Non-specific filtering and control of false positives

Research Article Sample Size Calculation for Controlling False Discovery Proportion

TESTING FOR EQUAL DISTRIBUTIONS IN HIGH DIMENSION

Let us first identify some classes of hypotheses. simple versus simple. H 0 : θ = θ 0 versus H 1 : θ = θ 1. (1) one-sided

2.1.3 The Testing Problem and Neave s Step Method

The Bootstrap: Theory and Applications. Biing-Shen Kuo National Chengchi University

Statistics 612: L p spaces, metrics on spaces of probabilites, and connections to estimation

Asymptotics of minimax stochastic programs

Statistical Inference: Estimation and Confidence Intervals Hypothesis Testing

Optimal Rates of Convergence for Estimating the Null Density and Proportion of Non-Null Effects in Large-Scale Multiple Testing

Estimating the Null and the Proportion of Non- Null Effects in Large-Scale Multiple Comparisons

Advanced Statistical Methods: Beyond Linear Regression

Supplement to Quantile-Based Nonparametric Inference for First-Price Auctions

ST495: Survival Analysis: Hypothesis testing and confidence intervals

Summary and discussion of: Controlling the False Discovery Rate: A Practical and Powerful Approach to Multiple Testing

FDR-CONTROLLING STEPWISE PROCEDURES AND THEIR FALSE NEGATIVES RATES

simple if it completely specifies the density of x

Can we do statistical inference in a non-asymptotic way? 1

Statistics: Learning models from data

A Novel Nonparametric Density Estimator

Estimation of Quantiles

University of California San Diego and Stanford University and

Smooth simultaneous confidence bands for cumulative distribution functions

Statistics - Lecture One. Outline. Charlotte Wickham 1. Basic ideas about estimation

Testing Homogeneity Of A Large Data Set By Bootstrapping

The miss rate for the analysis of gene expression data

A tailor made nonparametric density estimate

BTRY 4090: Spring 2009 Theory of Statistics

Asymptotic Statistics-III. Changliang Zou

Testing Hypothesis. Maura Mezzetti. Department of Economics and Finance Università Tor Vergata

Maximum Smoothed Likelihood for Multivariate Nonparametric Mixtures

Lecture 8 Inequality Testing and Moment Inequality Models

Table of Outcomes. Table of Outcomes. Table of Outcomes. Table of Outcomes. Table of Outcomes. Table of Outcomes. T=number of type 2 errors

Rank conditional coverage and confidence intervals in high dimensional problems

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

Large-Scale Multiple Testing of Correlations

STAT 200C: High-dimensional Statistics

Some General Types of Tests

Tweedie s Formula and Selection Bias. Bradley Efron Stanford University

Large-Scale Hypothesis Testing

Statistical Applications in Genetics and Molecular Biology

Optimal Estimation of a Nonsmooth Functional

The optimal discovery procedure: a new approach to simultaneous significance testing

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

Nonparametric confidence intervals. for receiver operating characteristic curves

HYPOTHESIS TESTING: FREQUENTIST APPROACH.

STATISTICS SYLLABUS UNIT I

Testing against a parametric regression function using ideas from shape restricted estimation

The International Journal of Biostatistics

A TEST OF FIT FOR THE GENERALIZED PARETO DISTRIBUTION BASED ON TRANSFORMS

LECTURE NOTE #3 PROF. ALAN YUILLE

More Powerful Tests for Homogeneity of Multivariate Normal Mean Vectors under an Order Restriction

Composite Hypotheses and Generalized Likelihood Ratio Tests

STAT 135 Lab 5 Bootstrapping and Hypothesis Testing

EXPLICIT NONPARAMETRIC CONFIDENCE INTERVALS FOR THE VARIANCE WITH GUARANTEED COVERAGE

Nonparametric Estimation of Distributions in a Large-p, Small-n Setting

Pairwise rank based likelihood for estimating the relationship between two homogeneous populations and their mixture proportion

1 Degree distributions and data

Spring 2012 Math 541B Exam 1

Likelihood Based Inference for Monotone Response Models

Separating Signal from Background

TESTING FOR NORMALITY IN THE LINEAR REGRESSION MODEL: AN EMPIRICAL LIKELIHOOD RATIO TEST

high-dimensional inference robust to the lack of model sparsity

Robustness and Distribution Assumptions

Hypothesis Test. The opposite of the null hypothesis, called an alternative hypothesis, becomes

A Recursive Formula for the Kaplan-Meier Estimator with Mean Constraints

Transcription:

Estimation of a Two-component Mixture Model with Applications to Multiple Testing arxiv:124.5488v5 [stat.me] 9 Nov 215 Rohit Kumar Patra and Bodhisattva Sen Columbia University, USA Abstract We consider a two-component mixture model with one known component. We develop methods for estimating the mixing proportion and the unknown distribution nonparametrically, given i.i.d. data from the mixture model, using ideas from shape restricted function estimation. We establish the consistency of our estimators. We find the rate of convergence and asymptotic limit of the estimator for the mixing proportion. Completely automated distribution-free honest finite sample lower confidence bounds are developed for the mixing proportion. Connection to the problem of multiple testing is discussed. The identifiability of the model, and the estimation of the density of the unknown distribution are also addressed. We compare the proposed estimators, which are easily implementable, with some of the existing procedures through simulation studies and analyse two data sets, one arising from an application in astronomy and the other from a microarray experiment. Keywords: Cramér-von Mises statistic, cross-validation, functional delta method, identifiability, local false discovery rate, lower confidence bound, microarray experiment, projection operator, shape restricted function estimation. 1 Introduction Consider a mixture model with two components, i.e., F(x) = αf s (x)+(1 α)f b (x), (1) where the cumulative distribution function (CDF) F b is known, but the mixing proportion α [,1] and the CDF F s ( F b ) are unknown. Given a random sample from F, we wish to (nonparametrically) estimate F s and the parameter α. This model appears in many contexts. In multiple testing problems (microarray analysis, neuroimaging) the p-values, obtained from the numerous (independent) hypotheses tests, are uniformly distributed on [,1], under H, while their distribution associated with H 1 is unknown; see e.g., Efron (21) and Robin et al. (27). Translated to the setting of (1), F b is the uniform distribution and the goal is to estimate the proportion of false null hypotheses α and the distribution of the p-values under the alternative. In addition, a reliable estimator of α is important when we want to assess or control multiple error rates, such as the false discovery rate of Benjamini and Hochberg (1995). 1

In contamination problems, the distribution F b, for which reasonable assumptions can be made, may be contaminated by an arbitrary distribution F s, yielding a sample drawn from F as in (1); see e.g., McLachlan and Peel (2). For example, in astronomy, such situations arise quite often: when observing some variable(s) of interest (e.g., metallicity, radial velocity) of stars in a distant galaxy, foreground stars from the Milky Way, in the field of view, contaminate the sample; the galaxy ( signal ) stars can be difficult to distinguish from the foreground stars as we can only observe the stereographic projections and not the three dimensional position of the stars (see Walker et al. (29)). Known physical models for the foreground stars help us constrain F b, and the focus is on estimating the distribution of the variable for the signal stars, i.e., F s. We discuss such an application in more detail in Section 9.2. Such problems also arise in High Energy physics where often the signature of new physics is evidence of a significant-looking peak at some position on top of a rather smooth background distribution; see e.g., Lyons (28). Most of the previous work on this problem assume some constraint on the form of the unknowndistribution F s, e.g., it is commonlyassumed that the distributions belongto certain parametric models, which lead to techniques based on maximum likelihood (see e.g., Cohen (1967) and Lindsay (1983)), minimum chi-square (see e.g., Day (1969)), method of moments (see e.g., Lindsay and Basak (1993)), and moment generating functions(see e.g., Quandt and Ramsey (1978)). Bordes et al. (26) assume that both the components belong to an unknown symmetric locationshift family. Jin (28) and Cai and Jin (21) use empirical characteristic functions to estimate F s under a semiparametric normal mixture model. In multiple testing, this problem has been addressed by various authors and different estimators and confidence bounds for α have been proposedin the literatureunder certainassumptionsonf s and itsdensity, seee.g., Storey (22), Genovese and Wasserman (24), Meinshausen and Rice (26), Meinshausen and Bühlmann (25), Celisse and Robin (21) and Langaas et al. (25). For the sake of brevity, we do not discuss the above references here but come back to this application in Section 7. In this paper we provide a methodology to estimate α and F s (nonparametrically), without assuming any constraint on the form of F s. The main contributions of our paper can be summarised in the following. We investigate the identifiability of (1) in complete generality. When F is a continuous CDF, we develop an honest finite sample lower confidence bound for the mixing proportion α. We believe that this is the first attempt to construct a distribution-free lower confidence bound for α that is also tuning parameter-free. Two different estimators of α are proposed and studied. We derive the rate of convergence and asymptotic limit for one of the proposed estimators. A nonparametric estimator of F s using ideas from shape restricted function estimation is proposed and its consistency is proved. Further, if F s has a non-increasing density f s, we can also consistently estimate f s. The paper is organised as follows. In Section 2 we address the identifiability of the model given in (1). In Section 3 we propose an estimator of α and investigate its theoretical properties, including its consistency, rate of convergence and asymptotic limit. In Section 4 we develop a completely automated distribution-free honest finite sample lower confidence bound for α. As the 2

performance of the estimator proposed in Section 3 depends on the choice of a tuning parameter, in Section 5 we study a tuning parameter-free heuristic estimator of α. We discuss the estimation of F s and its density f s in Section 6. Connection to the multiple testing problem is developed in Section 7. In Section 8 we compare the finite sample performance of our procedures, including a plug-in and cross-validated choice of the tuning parameter for the estimator proposed in Section 3, with other methods available in the literature through simulation studies, and provide a clear recommendation to the practitioner. Two real data examples, one arising in astronomy and the other from a microarray experiment, are analysed in Section 9. Appendix D gives the proofs of the results in the paper. 2 The model and identifiability 2.1 When α is known Suppose that we observe an i.i.d. sample X 1,X 2,...,X n from F as in (1). If α (,1] were known, a naive estimator of F s would be ˆF s,n α = F n (1 α)f b, (2) α where F n is the empirical CDF of the observed sample, i.e., F n (x) = n i=1 1{X i x}/n. Although this estimator is consistent, it does not satisfy the basic requirements of a CDF: ˆFα s,n need not be non-decreasing or lie between and 1. This naive estimator can be improved by imposing the known shape constraint of monotonicity. This can be accomplished by minimising {W(x) ˆF s,n(x)} α 2 df n (x) 1 n n {W(X i ) ˆF s,n(x α i )} 2 (3) over all CDFs W. Let ˇF s,n α be a CDF that minimises (3). The above optimisation problem is the same as minimising θ V 2 over θ = (θ 1,...,θ n ) Θ inc where i=1 Θ inc = {θ R n : θ 1 θ 2... θ n 1}, V = (V 1,V 2,...,V n ), V i := ˆF α s,n (X (i)), i = 1,2,...,n, X (i) being the i-th order statistic of the sample, and denotes the usual Euclidean norm in R n. The estimator ˆθ is uniquely defined by the projection theorem (see e.g., Proposition 2.2.1 on page 88 of Bertsekas (23)); it is the Euclidean projection of V on the closed convex set Θ inc R n. ˆθ is related to ˇFα s,n via ˇF α s,n (X (i)) = ˆθ i, and can be easily computed using the pool-adjacent-violatorsalgorithm (PAVA); see Section 1.2 of Robertson et al. (1988). Thus, ˇFα s,n is uniquely defined at the data points X i, for all i = 1,...,n, and can be defined on the entire real line by extending it to a piece-wise constant right continuous function with possible jumps only at the data points. The following result, derived easily from Chapter 1 of Robertson et al. (1988), characterises ˇF α s,n. Lemma 1. Let F s,n α be the isotonic regression (see e.g., page 4 of Robertson et al. (1988)) of the set of points {ˆF s,n α (X (i))} n i=1. Then F s,n α is characterised as the right-hand slope of the greatest convex minorant of the set of points {i/n, i ˆF j= s,n(x α (j) )} n i=. The restriction of F s,n α to [,1], 3

i.e., ˇF α s,n = min{max{ F α s,n,},1}, minimises (3) over all CDFs. Isotonic regression and the PAVA are very well studied in the statistical literature with many text-book length treatments; see e.g., Robertson et al. (1988) and Barlow et al. (1972). If skillfully implemented, PAVA has a computational complexity of O(n)(see Grotzinger and Witzgall (1984)). 2.2 Identifiability of F s When α is unknown, the problem is considerably harder; in fact, it is non-identifiable. If (1) holds for some F b and α then the mixture model can be re-written as ( α F = (α+γ) α+γ F s + γ ) α+γ F b +(1 α γ)f b, for γ 1 α, and the term (αf s +γf b )/(α+γ) can be thought of as the nonparametric component. A trivial solution occurs when we take α + γ = 1, in which case (3) is minimised when W = F n. Hence, α is not uniquely defined. To handle the identifiability issue, we redefine the mixing proportion as α := inf{γ (,1] : [F (1 γ)f b ]/γ is a CDF}. (4) Intuitively, this definition makes sure that the signal distribution F s does not include any contribution from the known background F b. In this paper we consider the estimation of α as defined in (4). Identifiability of mixture models has been discussed in many papers, but generally with parametric assumptions on the model. Genovese and Wasserman (24) discuss identifiability when F b is the uniform distribution and F has a density. Hunter et al. (27) and Bordes et al. (26) discuss identifiability for location shift mixtures of symmetric distributions. Most authors try to find conditions for the identifiability of their model, while we go a step further and quantify the non-identifiability by calculating α and investigating the difference between α and α. In fact, most of our results are valid even when (1) is non-identifiable. Suppose that we start with a fixed F s,f b and α satisfying (1). As seen from the above discussion we can only hope to estimate α, which, from its definition in (4), is smaller than α, i.e., α α. A natural question that arises now is: under what condition(s) can we guarantee that the problem is identifiable, i.e., α = α? The following lemma gives the connection between α and α. Lemma 2. Let F be as in (1) and α as defined in (4). Then α = α sup{ ǫ 1 : αf s ǫf b is a sub-cdf}, (5) where sub-cdf is a non-decreasing right-continuous function taking values between and 1. In particular, α < α if and only if there exists ǫ (,1) such that αf s ǫf b is a sub-cdf. Furthermore, α = if and only if F = F b. In the following we separately identify α for any distribution, be it continuous or discrete or a mixture of the two, with a series of lemmas proved in Appendix A. By an application of 4

the Lebesgue decomposition theorem in conjunction with the Jordan decomposition theorem (see page 142, Chapter V, Section 3a of Feller (1971)), we have that any CDF G can be uniquely represented as a weighted sum of a piecewise constant CDF G (d), an absolutely continuous CDF G (a), and a continuous but singular CDF G (s), i.e., G = η 1 G (a) +η 2 G (d) +η 3 G (s), where η i, for i = 1,2,3, and η 1 + η 2 + η 3 = 1. However, from a practical point of view, we can assume η 3 =, since singular functions almost never occur in practice; see e.g., Parzen (196). Hence, we may assume G = ηg (a) +(1 η)g (d), (6) where (1 η) is the sum total of all the point masses of G. Let d(g) denote the set of all jump discontinuities of G, i.e., d(g) = {x R : G(x) G(x ) > }. Let us define J G : d(g) [,1] to be a function defined only on the jump points of G such that J G (x) = G(x) G(x ) for all x d(g). The following result addresses the identifiability issue when both F s and F b are discrete CDFs. Lemma 3. Let F s and F b be discrete CDFs. If d(f b ) d(f s ), then α = α, i.e., (1) is identifiable. If d(f b ) d(f s ), then α = α { 1 inf x d(fb )J Fs (x)/j Fb (x) }. Thus, α = α if and only if inf x d(fb )J Fs (x)/j Fb (x) =. Next, let us assume that both F s and F b are absolutely continuous CDFs. Lemma 4. Suppose that F s and F b are absolutely continuous, i.e., they have densities f s and f b, respectively. Then { α = α 1 essinf f } s, f b where, for any function g, essinfg = sup{a R : m({x : g(x) < a}) = }, m being the Lebesgue measure. As a consequence, α < α if and only if there exists c > such that f s cf b, almost everywhere w.r.t. m. The above lemma states that if there does not exist any c > for which f s (x) cf b (x), for almost every x, then α = α and we can estimate the mixing proportion correctly. Note that, in particular, if the support of F s is strictly contained in that of F b, then the problem is identifiable and we can estimate α. In Appendix A we apply the above two lemmas to two discrete (Poisson and binomial) distributions and two absolutely continuous (exponential and normal) distributions to obtain the exact relationship between α and α. In the following lemma, proved in greater generality in Appendix A, we give conditions under which a general CDF F, that can be represented as in (6), is identifiable. Lemma 5. Suppose that F = κf (a) +(1 κ)f (d), where F (a) is an absolutely continuous CDF and F (d) is a piecewise constant CDF, for some κ (,1). Then (1) is identifiable, if either F (a) or F (d) are identifiable. 5

3 Estimation 3.1 Estimation of the mixing proportion α In this section we consider the estimation of α as defined in (5). For the rest of the paper, unless otherwise noted, we assume X 1,X 2,...,X n is an i.i.d. sample from F as in (1). Recall the definitions of ˆF s,n γ and ˇF s,n γ, for γ (,1]; see (2) and (3). When γ = 1, we have ˆF s,n γ = F n = ˇF s,n γ as ˆF s,n γ (for γ = 1) is a CDF. Whereas, when γ is much smaller than α the regularisation of ˆF s,n γ modifies it, and thus ˆF s,n γ and ˇF s,n γ are quite different. We would like to compare the naive and isotonised estimators ˆF s,n γ and ˇF s,n γ, respectively, and choose the smallest γ for which their distance is still small. This leads to the following estimator of α : { ˆα cn = inf γ (,1] : γd n (ˆF s,n γ, ˇF s,n γ ) c } n, (7) n where c n is a sequence of constants and d n stands for the L 2 (F n ) distance, i.e., if g,h : R R are two functions, then d 2 n(g,h) = {g(x) h(x)} 2 df n (x). It is easy to see that d n (F n,γ ˇF γ s,n +(1 γ)f b ) = γd n (ˆF γ s,n, ˇF γ s,n). (8) For simplicity of notation, using (8), we define γd n (ˆF s,n γ, ˇF s,n γ ) for γ = as lim γd n(ˆf s,n γ, ˇF s,n γ ) = d n(f n,f b ). (9) γ + This convention is followed in the rest of the paper. The choice of c n is important, and in the following sections we address this issue in detail. We derive conditions on c n that lead to consistent estimators of α. We will also show that particular (distribution-free) choices of c n will lead to honest lower confidence bounds for α. Next, we prove a result which implies that, in the multiple testing problem, estimators of α donotdepend onwhetherweusep-valuesorz-valuestoperformouranalysis. LetΨ : R Rbea knowncontinuousnon-decreasingfunction. We defineψ 1 (y) := inf{t R : y Ψ(t)},andY i := Ψ 1 (X i ). It is easy to see that Y 1,Y 2,...,Y n is an i.i.d. sample from G := αf s Ψ+(1 α)f b Ψ. Suppose now that we work with Y 1,Y 2,...,Y n, instead of X 1,X 2,...,X n, and want to estimate α. We can define α Y as in (4) but with {G,F b Ψ} instead of {F,F b }. The following result shows that α and its estimators proposed in this paper are invariant under such monotonic transformations. Theorem 1. Let G n be the empirical CDF of Y 1,Y 2,...,Y n. Also, let Ĝ s,n and Ǧγ s,n be as defined in (2) and (3), respectively, but with {G n,f b Ψ} instead of {F n,f b }. Then α = α Y and γd n (ˆF γ s,n, ˇF γ s,n) = γd n (Ĝγ s,n,ǧγ s,n) for all γ (,1]. 3.2 Consistency of ˆα cn We start with two elementary results on the behaviour of our criterion function γd n (ˇF γ s,n, ˆF γ s,n). 6

Lemma 6. For 1 γ α, γd n (ˇF γ s,n, ˆF γ s,n) d n (F,F n ). Thus, γd n (ˆF γ s,n, ˇF γ s,n) a.s. {, γ α, >, γ α <. (1) Lemma 7. The set A n := {γ [,1] : nγd n (ˆF γ s,n, ˇF γ s,n) c n } is convex. Thus, A n = [ˆα cn,1]. The following result shows that for a broad range of choices of c n, our estimation procedure is consistent. Theorem 2. If c n = o( n) and c n, then ˆα cn P α. A proper choice of c n is important and crucial for the performance of ˆα cn. We suggest doing cross-validation to find the optimal tuning parameter c n. In Section 8.2.1 we detail this approach and illustrate its good finite sample performance through simulation examples; see Tables 2-5, Section 8.2.4, and Appendix B. However, cross-validation can be computationally expensive. Another useful choice for c n is to take c n =.1loglogn. After extensive simulations, we observe that c n =.1loglogn has good finite sample performance for estimating α ; see Section 8 and Appendix B for more details. 3.3 Rate of convergence and asymptotic limit We first discuss the case α =. In this situation, under minimal assumptions, we show that as the sample size grows, ˆα cn exactly equals α with probability converging to 1. Lemma 8. When α =, if c n as n, then P(ˆα cn = ) 1. For the rest of this section we assume that α >. The following theorem gives the rate of convergence of ˆα cn. Theorem 3. Let r n := n/c n. If c n and c n = o(n 1/4 ) as n, then r n (ˆα cn α ) = O P (1). The proof of the above result is involved and we give the details in Appendix D.9. Remark1. Genovese and Wasserman (24) show that the estimators of α proposed byhengartner and Stark (1995) and Swanepoel (1999) have convergence rates of (n/logn) 1/3 and n 2/5 /(logn) δ, for δ >, respectively. Morover, both results require smoothness assumptions on F Hengartner and Stark (1995) require F to be concave with a density that is Lipschitz of order 1, while Swanepoel (1999) requires even stronger smoothness conditions on the density. Nguyen and Matias (213) prove that when the density of Fs α vanishes at a set of points of measure zero and satisfies certain regularity assumptions, then any n-consistent estimator of α will not have finite variance in the limit (if such an estimator exists). We can take r n = n/c n arbitrarily close to n by choosing c n that increases to infinity very slowly. If we take c n = loglogn, we get an estimator that has a rate of convergence n/loglogn. In fact, as the next result shows, r n (ˆα cn α ) converges to a degenerate limit. In Section 8.2, we analyse the effect of c n on the finite sample performance of ˆα cn for estimating α through simulations and advocate a proper choice of the tuning parameter c n. 7

Theorem 4. When α >, if r n, c n = o(n 1/4 ) and c n, as n, then r n (ˆα cn α ) P c, where c < is a constant that depends on α, F and F b. 4 Lower confidence bound for α The asymptotic limit ofthe estimator ˆα cn discussedin Section 3 depends on unknown parameters (e.g., α,f) in a complicated fashion and is of little practical use. Our goal in this sub-section is to construct a finite sample (honest) lower confidence bound ˆα L with the property P(α ˆα L ) 1 β, (11) for a specified confidence level (1 β) ( < β < 1), that is valid for any n and is tuning parameter free. Such a lower bound would allow one to assert, with a specified level of confidence, that the proportion of signal is at least ˆα L. It can also be used to test the hypothesis that there is no signal at level β by rejecting when ˆα L >. The problem of no signal is known as the homogeneity problem in the statistical literature. It is easy to show that α = if and only if F = F b. Thus, the hypothesis of no signal or homogeneity can be addressed by testing whether α = or not. There has been a considerable amount of work on the homogeneity problem, but most of the papers make parametric model assumptions. Lindsay (1995) is an authoritative monograph on the homogeneity problem but the components are assumed to be from a known exponential family. Walther (21) and Walther (22) discuss the homogeneity problem under the assumption that the densities are log-concave. Donoho and Jin (24) and Cai and Jin (21) discuss the problem of detecting sparse heterogeneous mixtures under parametric settings using the higher criticism statistic; see Appendix C for more details. Itwillbeseenthatourapproachwillleadtoanexactlowerconfidenceboundwhenα =,i.e., P(ˆα L = ) = 1 β. ThemethodsofGenovese and Wasserman (24)andMeinshausen and Rice (26) usually yield conservative lower bounds. Theorem 5. Let H n be the CDF of nd n (F n,f). Let ˆα L be defined as in (7) with c n = H 1 n (1 β). Then (11) holds. Furthermore if α =, then P(ˆα L = ) = 1 β, i.e., it is an exact lower bound. The proofof the abovetheorem can be found in Appendix D.13. Note that H n is distributionfree (i.e., it does not depend on F s and F b ) when F is a continuous CDF and can be readily approximated by Monte Carlo simulations using a sample of uniforms. For moderately large n (e.g., n 5) the distribution H n can be very well approximated by that of the Cramér-von Mises statistic, defined as nd(fn,f) := n{f n (x) F(x)} 2 df(x). Letting G n be the CDF of nd(f n,f), we have the following result. 8

Theorem 6. sup x R H n (x) G n (x) as n. Hence in practice, for moderately large n, we can take c n to be the (1 β)-quantile of G n or its asymptotic limit, which are readily available (e.g., see Anderson and Darling (1952)). When F is a continuous CDF, the asymptotic 95% quantile of G n is.6792, and is used in our data analysis. Note that P(α ˆα L ) = P( nα d n (ˆF s,n, α α ˇF s,n) Hn 1 (1 β)). The following theorem gives the explicit asymptotic limit of P(α ˆα L ) but it is not useful for practical purposes as it involves the unknown F α s and F. Theorem 7. Assume that α >. Then nα d n (ˆF s,n α α, ˇF s,n ) d U, where U is a random variable whose distribution depends only on α,f, and F b. The proof of the above theorem and the explicit from of U can be found in Appendix D. The proof of Theorem 6 and a detailed discussion on the performance of the lower confidence bound for detecting heterogeneity in the moderately sparse signal regime considered in Donoho and Jin (24) can be found in Appendix C. 5 A heuristic estimator of α In simulations, we observe that the finite sample performance of (7) is affected by the choice of c n (for an extensive simulation study on this see Section 8.2). This motivates us to propose a method to estimate α that is completely automated and has good finite sample performance. We start with a lemma that describes the shape of our criterion function, and will motivate our procedure. Lemma 9. γd n (ˆF s,n γ, ˇF s,n γ ) is a non-increasing convex function of γ in (,1). Writing ˆF γ s,n = F n F γ { ( α + γ Fα s + 1 α } )F b, γ we see that for γ α, the second term in the right hand side is a CDF. Thus, for γ α, ˆF s,n γ is very close to a CDF as F n F = O P (n 1/2 ), and hence ˇF s,n γ should also be close to ˆF s,n γ. Whereas, for γ < α, ˆF s,n γ is not close to a CDF, and thus the distance γd n(ˆf s,n γ, ˇF s,n γ ) is appreciably large. Therefore, at α, we have a regime change: γd n (ˆF s,n γ, ˇF s,n γ ) should have a slowly decreasing segment to the right of α and a steeply non-increasing segment to the left of α. Fig. 1 shows two typical such plots of the function γd n (ˆF s,n γ, ˇF s,n γ ), where the left panel corresponds to a mixture of N(2,1) with N(,1) (setting I) and in the right panel we have a mixture of Beta(1,1) and Uniform(, 1) (setting II). We will use these two settings to illustrate our methodology in the rest of this section and also in Section 8.1. Using the above heuristics, we can see that the elbow of the function should provide a good estimate of α ; it is the point that has the maximum curvature, i.e., the point where the second derivative is maximal. We denote this estimator by α. Notice that both the estimators α and ˆα cn are derived from γd n (ˆF s,n γ, ˇF s,n γ ), as a function of γ, albeit they look at two different aspects of the function. 9

.5.4.3.2.1.4.3.2.1.5.1.15.2.25 γ.5.1.15.2 γ Figure 1: Plots of γd n (ˆF s,n γ, ˇF s,n γ ) (in solid blue) overlaid with its (scaled) second derivative (in dashed red) for α =.1 and n = 5. Left panel: setting I; right panel: setting II. In the above plots we have used numerical methods to approximate the second derivative of γd n (ˆF s,n γ, ˇF s,n γ ) (using the method of double differencing). We advocate plotting the function γd n (ˆF s,n, γ ˇF s,n) γ as γ varies between and 1. In most cases, plots similar to Fig. 1 would immediately convey to the practitioner the most appropriate choice of α. In some cases though, there can be multiple peaks in the second derivative, in which case some discretion on the part of the practitioner might be required. It must be noted that the idea of finding the point where the second derivative is large to detect an elbow or knee of a function is not uncommon; see e.g., Salvador and Chan (24). However, in Section 8.2.4 and Appendix B, we show some simulation examples where α fails to consistently estimate the elbow of γd n (ˆF s,n γ, ˇF s,n γ ). 6 Estimation of the distribution function and its density 6.1 Estimation of F s Let us assume for the rest of this section that (1) is identifiable, i.e., α = α, and α >. Thus Fs α = F s. Once we have a consistent estimator ˇα n (which may or may not be ˆα cn as discussed ˇαn in the previous sections) of α, a natural nonparametric estimator of F s is ˇF s,n, defined as the ˇαn minimiser of (3). In the following theorem we show that, indeed, ˇF s,n is uniformly consistent for estimating F s. We also derive the rate of convergence of P Theorem 8. Suppose that ˇα n α. Then, as n, sup x R ˇF s,n ˇαn (x) F s(x) P. Furthermore, if q n (ˇα n α ) = O P (1), where q n = o( n), then sup x R q n ˇF s,n(x) ˇαn F s (x) = O P (1). Additionally, for ˆα cn as defined in (7), we have sup x R r n (ˆF ˆαcn ˇF ˇαn s,n. s,n F s )(x) Q(x) P and r n d(ˇf ˆαcn s,n,f s ) P c for a function Q : R R and a constant c > depending only on α,f, and F b. An immediate consequence of Theorem 8 is that d n (ˇF ˇαn Fig. 2 shows our estimator panel of Fig. 1. s,n, ˆF ˇαn s,n) P as n. Left panel of ˇF ˇαn s,n along with the true F s for the same data set used in the right 1

1 1 8 CDF.8.6.4.2.5.1.15.2.25.3 Density 6 4 2.1.2.3.4 α Figure 2: Left panel: Plots of ˇF s,n (in dashed red), F s,n (in solid blue) and F s (in dotted black) for setting II; right panel: plots of f s,n (in dashed red) and f s (in solid blue) for setting II. 6.2 Estimating the density of F s Suppose now that F s has a density f s. Obtaining nonparametric estimators of f s can be difficult as it requires smoothing and usually involves the choice of tuning parameter(s) (e.g., smoothing bandwidths), and especially so in our set-up. In this sub-section we describe a tuning parameter free approach to estimating f s, under the additional assumption that f s is non-increasing. The assumption that f s is non-increasing, i.e., F s is concave on its support, is natural in many situations (see Section 7 for an application in the multiple testing problem) and has been investigated by several authors, including Grenander (1956), Langaas et al. (25) and Genovese and Wasserman (24). Without loss of generality, we assume that f s is non-increasing on [, ). For a bounded function g : [, ) R, let us represent the least concave majorant (LCM) of g by LCM[g]. Thus, LCM[g] is the smallest concave function that lies above g. Define F s,n ˇαn := LCM[ˇF s,n ]. Note that F s,n is a valid CDF. We can now estimate f s by f s,n, where f s,n is the piece-wise constant function obtained by taking the left derivative of F s,n. In the following result we show that both F s,n and f s,n are consistent estimators of their population versions. P Theorem 9. Assume that F s () = and that F s is concave on [, ). If ˇα n α, then, as n, sup F s,n(x) F s (x) P. (12) x R Further, if for any x >, f s (x) is continuous at x, then, f s,n(x) P f s (x). Computing F s,n and f s,n are straightforward, an application of the PAVA gives both the estimators; see e.g., Chapter 1 of Robertson et al. (1988). In Fig. 2 the left panel shows the LCM F s,n whereas the right panel shows its derivative f s,n along with the true density f s for the same data set used in the right panel of Fig. 1. 7 Multiple testing problem The problem of estimating the proportion of false null hypotheses α is of interest in situations where a large number of hypothesis tests are performed. Recently, various such situations have arisen in applications. One major motivation is in estimating the proportion of genes that 11

are differentially expressed in deoxyribonucleic acid (DNA) microarray experiments. However, estimating the proportion of true null hypotheses is also of interest, for example, in functional magnetic resonance imaging (see Turkheimer et al. (21)) and source detection in astrophysics (see Miller et al. (21)). Suppose that we wish to test n null hypotheses H 1,H 2,...,H n on the basis of a data set X. Let H i denote the (unobservable) binary variable that is if H i is true, and 1 otherwise, i = 1,...,n. We want a decision rule D that will produce a decision of null or non-null for each of the n cases. In their seminal work, Benjamini and Hochberg (1995) argued that an important quantity to control is the false discovery rate (FDR) and proposed a procedure with the property FDR β(1 α ), where β is the user-defined level of the FDR procedure. When α is significantly bigger than an estimate of α can be used to yield a procedure with FDR approximately equal to β and thus will result in an increased power. This is essentially the idea of the adapted control of FDR (see Benjamini and Hochberg (2)). See Storey (22), Black (24), Langaas et al. (25), Benjamini et al. (26), and Donoho and Jin (24) for a discussion on the importance of efficient estimation of α and some proposed estimators. Our method can be directly used to yield an estimator of α that does not require the specification of any tuning parameter, as discussed in Section 5. We can also obtain a completely nonparametric estimator of F s, the distribution of the p-values arising from the alternative hypotheses. Suppose that F b has a density f b and F s has a density f s. To keep the following discussion more general, we allow f b to be any known density, although in most multiple testing applications we will take f b to be Uniform(,1). The local false discovery rate (LFDR) is defined as the function l : (,1) [, ), where l(x) = P(H i = X i = x) = (1 α )f b (x), f(x) and f(x) = α f s (x) +(1 α )f b (x) is the density of the observed p-values. The estimation of the LFDR l is important because it gives the probability that a particular null hypothesis is true given the observed p-value for the test. The LFDR method can help us get easily interpretable thresholding methods for reporting the interesting cases (e.g., l(x).2). Obtaining good estimates of l can be tricky as it involves the estimation of an unknown density, usually requiring smoothing techniques; see Section 5 of Efron (21) for a discussion on estimation and interpretation of l. From the discussion in Section 6.1, under the additional assumption that f s is non-increasing, we have a natural tuning parameter free estimator ˆl of the LFDR: ˆl(x) = (1 ˇα n )f b (x) ˇα n f s,n(x)+(1, for x (,1). ˇα n )f b (x) The assumption that f s is non-increasing, i.e., F s is concave, is quite natural when the alternative hypothesis is true the p-value is generally small and has been investigated by several authors, including Genovese and Wasserman (24) and Langaas et al. (25). 12

Table 1: Coverage probabilities of nominal 95% lower confidence bounds for the three methods when n = 1 and n = 5. n = 1 n = 5 α ˆα L ˆα GW L Setting I Setting II Setting I Setting II ˆα MR L ˆα L ˆα GW L ˆα MR L ˆα L ˆα GW L ˆα MR L ˆα L ˆα GW L.95.98.93.95.98.93.95.97.93.95.97.93.1.97.98.99.97.97.99.98.98.99.98.98.99.3.98.98.99.98.98.99.98.98.99.98.98.99.5.98.98.99.98.98.99.99.99.99.98.98.99.1.99.99 1..99.98.99.99.99 1..99.98.99 ˆα MR L 8 Simulation To investigate the finite sample performance of the estimators developed in this paper, we carry out several simulation experiments. We also compare the performance of these estimators with existing methods. The R language (R Development Core Team (28)) codes used to implement our procedures are available at http://stat.columbia.edu/ rohit/research.html. 8.1 Lower bounds for α Althoughtherehasbeensomeworkonestimationofα inthemultipletestingsetting,meinshausen and Rice (26) and Genovese and Wasserman (24) are the only papers we found that discuss methodology for constructing lowerconfidence bounds for α. These proceduresare connected and the methods in Meinshausen and Rice (26) are extensions of those proposed in Genovese and Wasserman (24). The lower bounds proposed in both the papers approximately satisfy (11) and have the form sup t (,1) (F n (t) t η n,β δ(t))/(1 t), where η n,β is a bounding sequence for the bounding function δ(t) at level β; see Meinshausen and Rice (26). Genovese and Wasserman (24) use a constantboundingfunction,δ(t) = 1,withη n,β = log(2/β)/2n,whereasmeinshausen and Rice (26) suggest a class of bounding functions but observe that the standard deviation-proportional bounding function δ(t) = t(1 t) has optimal properties among a large class of possible bounding functions. We use this bounding function and a bounding sequence suggested by the authors. We denote the lower bound proposed in Meinshausen and Rice (26) by ˆα MR L, the bound in Genovese and Wasserman (24) by ˆα GW L, and the lower bound discussed in Section 4 by ˆα L. To be able to use the methods of Meinshausen and Rice (26) and Genovese and Wasserman (24) in setting I, introduced in Section 5, we transform the data such that F b is Uniform(,1) ; see Section 3.1 for the details. We take α {,.1,.3,.5,.1}and compare the performance of the three lower bounds in the two different simulation settings discussed in Section 5. For each setting we take the sample size n to be 1 and 5. We present the estimated coverage probabilities, obtained by averaging over 5 independent replications, of the lower bounds for both settings in Table 1. We can immediately see from the table that the bounds are usually quite conservative. However, it is worth pointing out that when α =, our method has exact coverage, as discussed in Section 4. Also, the fact that our procedure is simple, easy to implement, and completely automated, makes it very attractive. 13

8.2 Estimation of α In this sub-section, we illustrate and compare the performance of different estimators of α under two sampling scenarios. In scenario A, we proceed as in Langaas et al. (25). Let X j = (X 1j,X 2j,...,X nj ), for j = 1,...,J, and assume that each X j N(µ n 1,Σ n n ) and that X 1,X 2,...,X J are independent. We test H i : µ i = versus H 1i : µ i for each i = 1,2,...,n. We set µ i to zero for the true null hypotheses, whereas for the false null hypotheses, we draw µ i from a symmetric bi-triangular density with parameters a = log 2 (1.2) =.263 and b = log 2 (4) = 2; see page 568 of Langaas et al. (25) for the details. Let x ij denote a realisation of X ij and α be the proportion of false null hypotheses. Let x i = J j=1 x ij/j and s 2 i = J j=1 (x ij x i ) 2 /(J 1). To test H i versus H 1i, we calculate a two-sided p-value based on a one-sample t-test, with p i = 2P(T J 1 x i / s 2 i /J ), where T J 1 is a t-distributed random variable with J 1 degrees of freedom. In scenariob, we generate n+lindependent random variablesw 1,w 2,...,w n+l from N(,1) andsetz i = 1 i+l L+1 j=i w j fori = 1,2,...,n. Thedependence structureofthez i sisdetermined by L. For example, L = corresponds to the case where the z i s are i.i.d. standard normal. Let X i = z i + m i, for i = 1,2,...,n, where m i = under the null, and under the alternative, m i is randomly generated from Uniform(m,m +1) and sgn(m i ), the sign of m i, is randomly generated from { 1,1} with equal probabilities. Here m is a suitable constant that describes the simulation setting. Let 1 α be the proportion of true null hypotheses. Scenario B is inspired by the numerical studies in Cai and Jin (21) and Jin (28). Figure 3: Plots of the means of different estimators of α, computed over 5 independent replications, as the sample size increases from 3 to 2 1 5 ; left panel: scenario A with Σ = I n n ; right panel: scenario B with L = and m = 1. The horizontal line (in dotted blue) indicates the value of α. We use ˆα S,B to denote the estimator proposed by Storey (22) when bootstrapping is used to choose the required tuning parameter, and denote by ˆα S,λ the estimator when the value of the tuning parameter is fixed at λ. Langaas et al. (25) proposed an estimator that is tuning parameter free but crucially uses the known shape constraint of a convex and non-increasing f s ; we denote it by ˆα L. We evaluate ˆαL using the convest function in the R library limma. We also use the estimator proposed in Meinshausen and Rice (26) for two bounding functions: δ(t) = 14

Table 2: Means 1 and RMSEs 1 (in parentheses) of estimators discussed in Section 8.2 for scenario A with Σ = I n n, J = 1, n = 5, and k n = loglogn. 1α ˆα.1kn ˆα CV α ˆα GW ˆα MR ˆα S,.5 ˆα J ˆα CJ ˆα L ˆα E.1.13.15.13..1.9.14.5.16.36 (1.) (1.79) (.83) (1.) (.88) (1.41) (1.5) (5.32) (1.2) (3.7).3.3.35.27.2.12.29.29.15.35.36 (1.2) (1.87) (1.1) (2.8) (1.84) (1.41) (1.83) (5.46) (1.26) (3.96).5.48.51.46.18.26.47.49.26.55.35 (1.9) (1.9) (1.12) (3.29) (2.46) (1.49) (1.91) (5.73) (1.34) (3.8) 1..93.97.93.62.65.95.96.51 1.2.33 (1.35) (1.86) (1.32) (3.88) (3.57) (1.51) (1.94) (7.16) (1.36) (3.73) t(1 t) and δ(t) = 1. For its implementation, we must choose a sequence {βn } going to zero as n. Meinshausen and Rice (26) did not specify any particular choice of {β n } but required the sequence satisfy some conditions. We choose β n =.5/ n and denote the estimators by ˆα MR when δ(t) = t(1 t) and by ˆα GW when δ(t) = 1 (see Genovese and Wasserman (24)). We also compare our results with ˆα E, the estimator proposed in Efron (27) using the central matching method, computed using the locfdr function in the R library locfdr. Jin (28) and Cai and Jin (21) propose estimators when the model is a mixture of Gaussian distributions; we denote the estimator proposed in Section 2.2 of Jin (28) by ˆα J and in Section 3.1 of Cai and Jin (21) by ˆα CJ. Some of the competing methods require F b to be of a specific form (e.g., standard normal) in which case we transform the observed data suitably. The estimator ˆα cn depends on the choice of c n and in the following we investigate a proper choice of c n. We take α =.1 and evaluate the performance of ˆα τ loglogn for different values of τ, as n increases, for scenarios A and B. The choice c n = τ loglogn, for different values of τ, is suggested after extensive simulations. We also include α, ˆα GW, ˆα MR, and ˆα J in the comparison. For scenario A, we fix the sample size n at 5 and Σ = I n n. For scenario B, we fix n = 5 1 4, L =, and m = 1. In Fig. 3, we illustrate the effect of c n on estimation of α as n varies from 3 to 1 5. Recall that α denotes the estimator proposed in Section 5. For both scenarios, the sample mean of the estimators of α proposed in this paper converge to the true α, as the sample size grows. The methods developed in this paper perform favorably in comparison to ˆα GW, ˆα MR, and ˆα J. Since, the choice of c n dictates the finite sample performance of ˆα cn, we propose cross-validation to find an appropriate value of the tuning parameter. 8.2.1 Cross-validation In this sub-section, we use c instead of c n to simplify the notation. In the following we briefly describe our cross-validation procedure. For a K-fold cross validation, we randomly partition the data into K sets, say D 1,...,D K. Let F k n be the empirical CDF of the data in D k. Let ˆα c, k be the estimator defined in (7) using all data except those in D k and tuning parameter c. Further, let ˇF ˆαc, k, k s,n be the estimator of F s as defined in Lemma 1 using ˆα c, k and all data except those 15

Table 3: Means 1 and RMSEs 1 (in parentheses) of estimators discussed in Section 8.2 for scenario B with L =, m = 1, n = 5 1 4, and k n = loglogn. 1α ˆα.1kn ˆα CV α ˆα GW ˆα MR ˆα S,B ˆα J ˆα CJ ˆα L ˆα E.7.3.4.8...4.11.19.3.6 (.44) (.67) (.28) (.66) (.66) (.65) (.96) (2.96) (.38) (.77).2.14.18.16..1.8.28.55.7.5 (.73) (.79) (.62) (1.98) (1.89) (2.25) (1.33) (4.41) (1.26) (1.28).33.25.31.28.2.4.12.48.92.12.5 (.89) (.85) (.95) (3.15) (2.91) (3.83) (1.77) (6.48) (2.14) (1.9).66.55.62.58.12.14.23.95 1.83.23.5 (1.21) (1.) (1.48) (5.38) (5.25) (7.73) (3.4) (11.98) (4.34) (3.84) in D k. Define the cross-validated estimator of c as c cv := argmin c R K k=1 (F k n ˆF k ) 2 df k n, (13) where ˆF k := ˆα c ˇF ˆαc, k, k, k s +(1 ˆα c, k )F b. In all simulations in this paper, we use K = 1 and denote this estimator by ˆα CV ; see Section 7.1 of Hastie et al. (29) for a more detailed study of cross-validation and a justification for K = 1. Fig. 4 illustrates the superior performance of ˆα CV across different simulation settings; also see Sections 8.2.2 and 8.2.4, and Appendix B 8.2.2 Performance under independence In this sub-section, we take α {.1,.3,.5,.1} and compare the performance of the different estimators under the independence setting of scenarios A and B. In Tables 2 and 3, we give the mean and root mean squared error (RMSE) of the estimators over 5 independent replications. For scenario A, we fix the sample size n at 5 and Σ = I n n. For scenario B, we fix n = 5 1 4, L =, and m = 1. By an application of Lemma 4, it is easy to see that in scenario A, the model is identifiable (i.e., α = α), while in scenario B, α = α.67. For scenario A, the sample means of ˆα CV, α, ˆα J, ˆαL, and ˆα.1kn for k n = loglogn are comparable. However, the RMSEs of α and ˆα.1kn are lower than those of ˆα CV, ˆα J, and ˆα L. For scenario B, the sample means of α, ˆα CV, and ˆα.1kn are comparable. In scenario B, the performances of ˆα J and ˆα CJ are not comparable to the estimators proposed in this paper, as ˆα J and ˆαCJ estimate α, while α, ˆα CV, and ˆα cn estimate α. Note that ˆα L fails to estimate α because the underlying assumption inherent in their estimation procedure, that f s be non-increasing, does not hold. In scenario A, ˆα S,.5 has the best performance among the different values of λ, while in scenario B, ˆα S,λ has poor performance for all values of λ [,1]. Furthermore, ˆα GW, ˆα MR, ˆα CJ, ˆαS,B and ˆα E perform poorly in both scenarios for all values of α. 8.2.3 Performance under dependence The simulation settings of this sub-section are designed to investigate the effect of dependence on the performance of the estimators. For scenario A, we use the setting of Langaas et al. (25). We take Σ to be a block diagonal matrix with block size 1. Within blocks, the diagonal 16

Table 4: Means 1 and RMSEs 1 (in parentheses) of estimators discussed in Section 8.2 for scenario A with Σ as described in Section 8.2.3, J = 1, n = 5, and k n = loglogn. 1α ˆα.1kn ˆα CV α ˆα GW ˆα MR ˆα S,.5 ˆα J ˆα CJ ˆα L ˆα E.1.46.42.33.7.6.28.22.7.32.37 (5.15) (4.23) (3.84) (1.72) (1.27) (4.11) (3.3) (1.61) (4.37) (3.91).3.52.53.41.14.17.65.34.15.49.39 (3.8) (3.64) (3.59) (2.72) (1.9) (6.58) (3.25) (1.35) (4.3) (4.31).5.66.76.54.26.31.54.49.25.66.37 (3.52) (5.43) (3.85) (3.56) (2.5) (2.61) (3.6) (1.45) (4.31) (4.3) 1. 1.6 1.13.97.68.69 1.15.97.53 1.11.36 (3.9) (3.92) (4.) (4.15) (3.54) (6.1) (3.61) (1.55) (4.13) (3.99) Table 5: Means 1 and RMSEs 1 (in parentheses) of estimators discussed in Section 8.2 for scenario B with L = 3, m = 1, n = 5 1 4, and k n = loglogn. 1α ˆα.1kn ˆα CV α ˆα GW ˆα MR ˆα S,B ˆα J ˆα CJ ˆα L ˆα E.7.29.38.17.4.5.26.2.21.13.22 (2.92) (3.7) (1.62) (1.2) (1.36) (3.71) (2.8) (9.87) (1.75) (2.22).2.3.42.18.4.4.16.33.55.13.19 (1.84) (2.88) (1.25) (1.75) (1.71) (2.24) (3.25) (1.35) (1.42) (2.27).33.38.52.2.6.6.17.5.93.16.18 (1.54) (2.74) (1.89) (2.83) (2.73) (3.51) (3.71) (11.52) (2.3) (2.59).67.63.77.31.14.15.24.95 1.82.25.16 (1.53) (2.25) (4.32) (5.26) (5.13) (7.6) (4.54) (15.13) (4.23) (4.8) elements (i.e., variances) are set to 1 and the off-diagonal elements (within-block correlations) are set to ρ =.5. Outside of the blocks, all entries are set to. Tables 4 and 5 show that in both scenarios, none of the methods perform well for small values of α. However, in scenario A, the performances of ˆα.1kn, α, and α J are comparable, for larger values of α. In scenario B, ˆα.1kn performs well for α =.33 and.67. Observe that, as in the independence setting, ˆα GW, ˆα MR, ˆα S,B, ˆα CJ, and ˆα E perform poorly in both scenarios for all values of α. 8.2.4 Comparing the performance of ˆα cn, ˆαCV, and α Although the heuristic estimator α performs quite well in most of the simulation settings considered, there exists scenarios where α can fail to consistently estimate α. To illustrate this we consider four different CDFs F s and fix F b to be the uniform distribution on (,1) (see the top left plot of Fig. 4) and compare the performance of ˆα CV, α, ˆα.1kn with the best performing competing estimators (in each setting). We see that α may fail to estimate the elbow of γd n (ˆF s,n γ, ˇF s,n γ ), as a function of γ, when F s has a multi-modal density (see the middle row of Fig. 4). Observe that ˆα CV and ˆα.1kn perform favorably compared to all competing estimators and in the two scenarios where α fails to consistently estimate α, all our competing estimators also fail. The first two toy examples have been carefully constructed to demonstrate situations where the point of maximum curvature ( α ) is different from the elbow of the function; see the top right plot of Fig. 4 (also see Appendix B for further such examples). 17

Densties 6 4 2 Dist 1 Dist 2 Dist 3 Dist 4.2.15.1.5.2.4.6.8 1.5 γ.1.15 Estimated α Estimated α.1.1k n α ˆα CV.5 ˆα MR ˆα S,.2.12.1.8.5 1 1.5 2 Sample size 1 5.1k n α ˆα CV ˆα GW ˆα S,B.5 1 1.5 2 Sample size 1 5 Estimated α Estimated α.2.1k n.15 α.1.5.6.5.4 ˆα CV ˆα GW ˆα S,B.5 1 1.5 2 Sample size 1 5.1k n α ˆα CV ˆα J.5 1 1.5 2 Sample size 1 5 Figure 4: Top row left panel: density functions for different choices of F s ; top row right panel: plot of γd n (ˆF s,n, γ ˇF s,n) γ (in blue), the scaled second derivative (in red), ˆα CV (in black), and ˆα.1kn (in brown) for 5 independent samples of size 5 corresponding to Dist 1 ; the blue star denotes α. The bottom two rows show the means of different competing estimators of α, computed over 5 independent samples for Dist 1-4 (left-right, top-bottom) as sample size increases from 3 to 2 1 5 ; in each figure the dotted black line denotes the true α. 8.2.5 Our recommendation In this paper we study two estimators for α. For ˆα cn, a proper choice of c n is important for good finite sample performance. We suggest using cross-validation to find the optimal tuning parameter c n. However, cross-validation can be computationally expensive. An attractive alternative in this situation is to use α, which is easy to implement and has very good finite sample performance in most scenarios, especially with large sample sizes. We feel that a visual analysis of the plot of γd n (ˆF γ s,n, ˇF γ s,n) can be useful in checking the validity of α as an estimator of the elbow, and thus for α. 18

25 2 Frequency 15 1 5 1.1.2.3.4.5.6.7.8.9 1 p values (a) 2 1 (b).9 18.9 CDF.8.7.6.5.4.3.2.1.1.2.3.4.5 (c) Density 16 14 12 1 8 6 4 2.2.4.6.8 1 (d) LFDR.8.7.6.5.4.3.2.1.1.2.3.4.5 p-value (e) Figure 5: Plots for the prostate data: (a) Histogram of the p-values. The horizontal line (in solid black) indicates the Uniform(,1) distribution. (b) Plot of γd n (ˆF s,n γ, ˇF s,n γ ) (in solid blue) overlaid with its (scaled) second derivative (in dashed red). The vertical line (in dotted black) α indicates the point of maximum curvature α =.88. (c) ˇF s,n (in dotted red) and F s,n (in solid blue); (d) f s,n ; (e) estimated LFDR ˆl for p-values less than.5. 9 Real data analysis 9.1 Prostate data Genetic expression levels for n = 633 genes were obtained for m = 12 men, m 1 = 5 normal control subjects and m 2 = 52 prostate cancer patients. Without going into the biology involved, theprincipalgoalofthestudywastodiscoverasmallnumberof interesting genes,that is, genes whose expression levels differ between the cancer and control patients. Such genes, once identified, might be further investigated for a causal link to prostate cancer development. The prostate data is a 633 12 matrix X having entries x ij = expression level for gene i on patient j, i = 1,2,...,n, and j = 1,2,...,m, with j = 1,2,...,5, for the normal controls, and j = 51,52,...,12, for the cancer patients. Let x i (1) and x i (2) be the averages of x ij for the normal controls and for the cancer patients, respectively, for gene i. The two-sample t-statistic for testing significance of gene i is t i = { x i (1) x i (2)}/s i, where s i is an estimate of the standard error of x i (1) x i (2), i.e., s 2 i = (1/5+1/52)[ 5 j=1 {x ij x i (1)} 2 + 12 j=51 {x ij x i (2)} 2 ]/1. We work with the p-values obtained from the 633 two-sided t-tests instead of the t-values as then the distribution under the alternative will have a non-increasing density which we can estimate using the method developed in Section 6.1. Note that in our analysis we ignore the dependence of the p-values, which is only a moderately risky assumption for the prostate data; see Chapters 2 and 8 of Efron (21) for further analysis and justification. Fig. 5 show the plots of various quantities of interest, found using the methodology developed in Section 6.1 and Section 7, for the prostate data example. The 95% lower confidence bound ˆα L for this data is found to be.5. In Table 6, we display estimates of α based on the methods considered in this 19