Sequential sample sizes for high-throughput hypothesis testing experiments David Rossell Biostatistics and Bioinformatics Unit Institute for Research in Biomedicine, Barcelona, Spain [email protected]

Peter M¨ uller Department of Biostatistics M.D. Anderson Cancer Center, Houston, USA [email protected]

Abstract The number of experiments needed in high-throughput experiments is typically chosen informally. Although formal sample size calculations have been proposed, they depend critically on prior knowledge. We propose a sequential sample size strategy which, by updating knowledge when new data is available, depends less critically on prior assumptions. Compared to fixed sample size approaches, our sequential strategy stops experiments early when enough evidence has been accumulated, and recommends continuation when additional data is likely to provide valuable information. The approach is framed within a decision-theoretic framework, guaranteeing that the chosen design satisfies optimality properties. We propose a utility function based on the number of true positives which is straightforward to specify and intuitively appealing. As 0

The authors wish to thank Steven Kornblau and the people in his lab for kindly sharing the RPPA data.

1

for most sequential design problems, an exact solution is computationally unfeasible. We propose a simulation-based approximation with decision boundaries that allows to determine good designs within reasonable computing time. Our results show that the sequential design can lead to substantial increases in posterior expected utility. An implementation of the proposed approach is available in the Bioconductor package gaga.

1

Introduction

The number of replicates needed in high-throughput experiments is typically chosen informally. This may result in experiments which are not informative enough or which are unnecessarily extensive as conclusions could have been reached with a smaller sample size. We focus on experiments carried out to identify differential gene expression. For simplicity we assume only two biologic conditions but in our Bioconductor implementation more than two groups are allowed. The proposed approach remains useful with minor changes for high-throughput experiments with different inference goals (e.g., classification). Several authors proposed fixed sample size calculations for high-throughput experiments i.e. the sample size is fixed at the beginning of the experiment (Dobbin and Simon (2005), Lee and Whitmore (2004), M¨ uller et al. (2004), Bickel (2003), Zien et al. (2003), Pan et al. (2002), Mukherjee et al. (2002)). Their main limitation is the lack of flexibility to incorporate new evidence as data is collected. In particular, they require a good prior estimate of the important features of the data that may be generated by the experiment. This is impractical in our setup with a large number of variables, e.g. tens of thousands for microarrays or up to millions for human tiling arrays or next-generation sequencing experiments Sequential sample size designs overcome this limitation by allowing the investigator to update knowledge and make decisions as data is collected. When warranted by the data, they stop experimentation early or they postpone drawing conclusions until sufficient data can be obtained. This not only saves time and resources but also results in more ethical designs. 2

The goal of this paper is to develop a framework for sequential sample size calculation in high-throughput experiments. Motivated by the typical experimental setup, we build an approach that is robust with respect to inaccurate a priori guesses. The framework is based on Bayesian decision theory, therefore ensuring optimality properties of the chosen design. In particular, decisions are coherent with the underlying utility function and probability model. Our proposal emphasizes ease of interpretation and use. Sequential designs have been extensively used in no high-throughput decision problems. See Chow et al. (2003)[Chapter 8], Berry et al. (2001) or Whitehead (1997). For a less technical review see DeMets (2004). However, these approaches are only applicable to studies with a single or a few outcome variables. In contrast, in high-throughput experiments the goal is to make inference on a large number of variables. Regarding high-throughput setups, Fu et al. (2005) developed a sequential approach for microarrays which focused on classification problems, and Ruppert et al. (2007), Ferreira and Zwinderman (2006a), Ferreira and Zwinderman (2006b) proposed two-step designs in which sample size is calculated using a pilot sample. None of these approaches are fully sequential, nor are they formally stated within a decision-theoretical framework. Pahl et al. (2009) propose group sequential designs for genome-wide association studies which minimize a cost function subject to constraints on type I and II error rates. The main difficulty in determining decision-theoretic optimal sequential designs is that it is typically computationally unfeasible, even in single outcome experiments. Rossell et al. (2006) developed a computationally feasible approach based on the ideas of M¨ uller et al. (2006), Brockwell and Kadane (2003), Berry et al. (2001) and Carlin et al. (1998). Briefly, they compute a summary statistic S every time that new data is observed and they define decision boundaries which partition the sample space for the summary statistic S into a continuation region R0 and a stopping region R1 . While S ∈ R0 the experiment continues. The experiment is terminated when S ∈ R1 . The approach reduces the sequential decision problem to the (non-sequential) problem of finding the optimal decision boundaries. The main contributions of this paper are (i) an extension of this approach to massively

3

high dimensional problems; and (ii) the application to the important problem of differential gene expression. In Section 2 we formalize our proposal. First, we capture the main goals of differential expression experiments into an intuitive utility function. The optimal design is defined in terms of expected utility maximization. We discuss why an exact solution is not possible and propose approximations. We provide a proposition which guarantees the experiment to stop eventually, even if a maximum sample size is not specified. Although the approximations reduce the computational cost, our setup remains computer intensive. Our approach can be implemented using any Bayesian probability model. We use the GaGa model (Rossell, 2009) as it achieves a good balance between flexibility and computational efficiency. We review the GaGa model in Section 3. In Section 4 we illustrate our approach with examples. In Section 5 we conclude with some remarks.

2

Optimal Sequential Stopping

We formalize sequential sample size calculation within a Bayesian decision-theoretic framework, i.e. we define a utility function and we select the design that maximizes the posterior expected utility. At each decision time the expected utility is conditional on all the data available at that time (which may be no data at all) and averaged with respect to future data and uncertainty on the model parameters. In Section 2.1 we propose a utility function for differential expression experiments based on the number of true positives. With minor modifications, our methodology is also applicable to any other utility function. In Section 2.2 we discuss how to find the exact solution and why this is computationally unfeasible. In Section 2.3 we propose an approximate solution which allows to find good designs within reasonable computing time. We introduce some notation. Let n be the number of outcome variables (e.g. genes) and T be the maximum sample size that the experimenter can afford. We denote by xij the outcome variable for gene i = 1, . . . , n and individual j = 1, . . . , T . Let zj ∈ {0, . . . , nz } denote the biologic condition or group of sample j. For simplicity here we assume zj ∈ {0, 1}, 4

i.e. the goal is to compare two groups. But generalization to nz > 1 is straightforward, and is implemented in the Bioconductor gaga package. A latent variable δi indicates whether gene i is differentially expressed (DE) across groups zj = 0 vs zj = 1. That is, δi = 0 indicates equally expressed (EE) and δi = 1 DE. The indicators δi represent the unknown truth, and are part of the unknown parameter vector. Let θi be the parameters indexing the probability distribution of (xi1 , . . . , xiT ) and optionally let ν be a hyper-parameter vector which allows us to incorporate dependence across θ1 , . . . , θn . The dimensionality of θi may depend on δi . To simplify notation, we let xt = {xit , 1 ≤ i ≤ n} be the data obtained at time t and x1:t = {xij , 1 ≤ i ≤ n, 1 ≤ j ≤ t} be all data available up to time t. Further, we let θ = (θ1 , . . . , θn ) and δ = (δ1 , . . . , δn ).

2.1

Decision criterion

A Bayesian sequential decision problem is characterized by an alternating sequence of decisions st and data xt , a probability model for all unknown quantities, including data and parameters, and a utility function u(·) that formalizes the decision maker’s preferences for different decisions under hypothetical future data and assumed parameter values. The optimal decision is that which maximizes the expected utility, conditional on all data available at the time of making the decision and averaging over unknown parameters and possible future data, assuming optimal choice of all future decisions. It is convenient to distinguish sequential and terminal decisions. Sequential decisions correspond to stopping versus continuation and are made after each observation (or batch of observations). Terminal decisions are the classification of genes into EE (δi = 0) or DE (δi = 1), hence they are taken upon stopping only. Let st = s(x1:t ) = 1 indicate the sequential decision of stopping at time t (st = 0 indicates continuation). The sequential decisions st can alternatively be described by the stopping time τ = min{t : st = 1}. We use st and τ interchangeably. Let di (x1:t ) = 1 indicate the terminal decision to report gene i as DE (di (x1:t ) = 0 for declaring EE). Also, let d(x1:t ) = (d1 (x1:t ), . . . , dn (x1:t )). Both st

5

and d(x1:t ) depend on all data available up to time t. In a fully decision theoretic approach sequential and terminal decisions are chosen to jointly maximize expected utility. Instead, we assume a fixed rule for d(x1:t ) and focus on the optimal choice of st only. We take terminal decisions according to the Bayes rule of M¨ uller et al. (2004), which controls the Bayesian False Discovery Rate below some userspecified level. In this paper we used the 0.05 level for all examples. In other words, once experimentation stops we report as many DE outcomes as possible, while not having too many false positives. The sequential stopping decision st is based on the following utility function with a sampling cost c and a unit reward for each correctly identified DE outcome u(x1:τ , δ, d(x1:τ )) = −cτ +

n X

δi di (x1:τ ).

(1)

i=1

The cost c can be interpreted as the minimum number of true positives that make it worthwhile to continue the experiment for one more sample. This interpretation allows for easy elicitation of c, without any reference to the formal mathematical framework. Although we find (1) appealing, the upcoming discussion is independent of the specified utility. See, for example, M¨ uller et al. (2004), Guindani et al. (2009) or Rice et al. (2008) for other utility functions in high-throughput experiments.

2.2

Optimal rule

Finding the optimal sequential decision in general requires dynamic programming, also known as backward induction (DeGroot, 1970). At time t, the optimal sequential decision st is to continue whenever the posterior expected utility for st = 1, denoted u¯t (st = 1, x1:t ), is greater than u¯t (st = 0, x1:t ). Evaluating u¯t (st = 1, x1:t ) is usually straightforward. For the utility in (1) we find u¯t (st = 1, x1:t ) = −ct +

n X

P (δi = 1 | x1:t )di (x1:t ),

(2)

i=1

where P (δi = 1|x1:t ) is the posterior probability that outcome i is differentially expressed. The expectation is with respect to δ (recall that the terminal decision di (x1:t ) is fixed). 6

P (δi = 1|x1:t ) can be computed in closed-form for some probability models (including the GaGa model described in Section 3) or easily estimated from MCMC output. Evaluating u¯t (st = 0, x1:t ) is more challenging as it requires the optimal decision at time t + 1, t + 2, etc. More precisely, u¯t+1 (st = 0, x1:t ) =

Z dP (xt+1 | x1:t ) max

   u ¯t+1 (1, x1:t+1 )       R   dP (xt+2      

    u ¯t+2 (1, x1:t+2 )      u | x1:t+1 ) max ¯T −1 (1, x1:T −1 )   . . . . . . max     R dP (x | x  uT (0, x1:T ) T 1:T −1 )¯

(3) where P (xt+1 |x1:t ) is the posterior predictive distribution for xt+1 given x1:t . In each curly bracket, the top case is the realized utility when stopping, st+j = 1. The bottom case is the realized utility when st+j = 0 and requires to plug in the optimal choice for the next step. The expression is best read from right to left, starting with the optimal decision sT −1 (sT = 1 by definition), and then proceeding with the nested sequence of expectations with respect to P (xt+j+1 | x1:t+j ), and maximizations with respect to st+j . An exact solution requires assessing the utility for all possible future data trajectories xt+1 , . . . , xT . The expected utility of a given decision is then obtained by averaging over all trajectories. Consider the simplest possible scenario in which xij is binary. In this case there are 2n(T −t) trajectories, with n usually ranging in the order of tens thousands or even millions. Even here, the computational cost required for solving the integration-maximization sequence in (3) is prohibitive.

2.3

Approximate optimal rule

Berry et al. (2001), Brockwell and Kadane (2003), DeGroot (2004), M¨ uller et al. (2006) and Rossell et al. (2006) discuss some alternatives to evaluating (3). Following the ideas in Rossell et al. (2006), we define sequential stopping boundaries. We use parametric boundaries for a bivariate summary statistic. Specifically, we define S = (t, ∆t (TP)) as a summary statistic, 7

600 400 200 0

Expected increase in True Positives

2

4

6

8

10

Sample size per group

Figure 1: Optimal sequential boundary for c = 50 (thick black line) and forward simulation trajectories (light grey lines) for Dell’Orto’s data. where t is the decision making time and ∆t (TP) ≡ Ext+1 [¯ ut+1 (st+1 = 1, x1:t+1 ) | x1:t ] − u¯t (st = 1, x1:t ) is the one-step ahead expected increment in true positives. Ext+1 (·|x1:t ) denotes taking the expectation conditional on x1:t and marginalizing with respect to future data xt+1 . Consider the example in Figure 1, based on Dell’Orto’s microarray experiment (Section 4.2). The thick black line is a sequential decision boundary. Every time we observe new data we compute S. If S falls above the decision boundary we continue experimentation, otherwise we stop. Intuitively, we continue experimentation as long as we expect enough new true positives. We propose using linear boundaries for ease of interpretation and their parsimonious parametrization with intercept b0 and slope b1 . Denote the expected utility associated with 8

(b0 , b1 ) by U (b0 , b1 ). Algorithm 1 details a forward simulation algorithm (Carlin et al., 1998) to determine (b0 , b1 ) maximizing U (b0 , b1 ). We briefly describe the algorithm assuming that t samples x1:t are already available (t could be 0 for no data). We first simulate data xt+1:T from the posterior predictive distribution P (xt+1:T | x1:t ), conditional on all data x1:t observed up to time t. We then compute the summary statistic ∆t (TP) for the simulated data at times t + 1, . . . , T . Plotting ∆t (TP) against t we obtain a simulated trajectory for the summary statistic. Such simulated trajectories are shown as light grey lines in Figure 1. For a given (b0 , b1 ) we can evaluate the stopping time τ and the expected terminal utility u¯τ (sτ = 1, x1:τ ). Averaging u¯τ (sτ = 1, x1:τ ) across many, say B, simulated xt+1:T approximates the expected utility for (b0 , b1 ). Algorithm 1: Finding the Optimal Rule (b)

1. Forward simulation: xt+1:T ∼ P (xt+1:T | x1:t ), b = 1, . . . , B. (b)

For each xt evaluate ∆t (TP)(b) , t = t + 1, . . . , T . 2. Grid of stopping rules: For a grid of decision boundaries (b0 , b1 ): • Find the stopping times τ (b) using the saved summaries ∆t (TP)(b) • Approximate the expected utility of the stopping boundary P (b) U¯ (b0 , b1 ) = B1 u¯(sτ (b) = 1, x1:τ (b) ) Select the stopping rule (b?0 , b?1 ) ≡ arg maxb0 ,b1 {U¯ (b0 , b1 )}. In principle the optimal stopping boundary (b?0 , b?1 ) could be re-computed every time that new data is observed. However, the computational burden would be excessive and in this paper we determine (b?0 , b?1 ) only once. Besides the intuitive appeal, there are theoretical considerations which motivate the use of the proposed summary statistic and linear sequential rule. First, all fixed-sample designs are a particular case of our sequential boundary-based design. For instance, setting b0 = 4.5 and b1 = ∞ results in a fixed-sample size of 5. Thus the optimal sequential design has higher or equal posterior expected utility than any fixed-sample design. Second, the myopic design 9

(Berry and Fristedt, 1985) is a particular case of our sequential design, which results from setting b0 = c and b1 = 0. In particular, the condition ∆t (TP) > c is sufficient to guarantee that the optimal decision at time t is to continue. Third, at time T − 1 the optimal rule is given by ∆T −1 (TP) > c. Also, Proposition 2.1 guarantees (proof in the Appendix) that ∆t (TP) converges to zero asymptotically (t → ∞). Therefore stopping is guaranteed for boundaries with either b1 > 0 or b1 = 0 and b0 > 0, even if we do not specify a maximum sample size T . Proposition 2.1 Suppose that Bayes factors are consistent for the chosen probability model. Then for any , η > 0 there is t0 such that P (∆t (TP) > ) < η for all t > t0 . Proposition 2.1 holds under very general conditions, as Bayes factors are consistent for any parametric and many non-parametric models, as long as they assign non-null prior probability to each hypothesis. See Ghosal (2002) and Dass and Lee (2004) for more details.

3

Model

The strategy proposed in Section 2.3 uses simulation based optimal designs. This requires extensive predictive simulation and model fitting. Although in principle our approach can be implemented with any probability model, to be of practical use it must allow computationally efficient inference. For instance, simulating 1,000 trajectories for an experiment with 2 groups, n =50,000 outcomes and a time horizon T = 5 requires simulating 500,000,000 observations. For each trajectory and time point, the model must be fit and the summary statistic ∆t (TP) must be evaluated, which typically also requires posterior predictive simulation and model fitting. On the other hand, the model needs to be sufficiently flexible to capture the important features of the data. In this paper we use the GaGa model (Rossell, 2009), which offers a reasonable compromise between model flexibility and computational cost. GaGa models xij as arising from a Gamma distribution xij ∼ Ga(αi , αi /λizj ). Recall that zj indicates the group which sample √ j belongs to. The Gamma distribution is parametrized such that λizj is the mean and 1/ αi 10

is the coefficient of variation for gene i. The prior on (λi0 , λi1 , αi ) includes a positive prior probability for a tie, λi0 = λi1 , i.e., non-differential expression.

−1 λ−1 i0 ∼ Ga(α0 , α0 /ν), λi1 | λi0 , δi ∼

  Ga(α0 , α0 /ν) if δi = 1  I(δi0 = δi1 )

αi |β, µ ∼ Ga(β, β/µ)

if δi = 0 (4)

independently for i = 1 . . . n. Here (α0 , ν, β, µ) are hyper-parameters. The differential expression status δi indicates whether λiz are equal across groups z = 0, 1. We assume a priori P (δi = 1) = π, independently across i. The model includes gene-specific effects αi and a gene and group-specific mean λiz . √ Importantly, αi allows the coefficient of variation 1/ αi to vary across genes. Also, the gamma sampling distribution for xij captures the asymmetries frequently observed in highthroughput data (even after log-transforming). In terms of computational complexity, conditional on hyper-parameters (α0 , ν, β, µ, π) the posterior distributions are available in closed form. To estimate the hyper-parameters we adopt the empirical Bayes expectation-maximization (EM) scheme developed by Rossell (2009). That is, we treat the hyper-parameters as fixed and therefore no Markov Chain Monte Carlo is required for posterior predictive simulation, which substantially increases computational speed. This is reasonable in high-throughput experiments as hyper-parameters govern a large number of outcomes and there typically is little posterior uncertainty about their most probable values.

4

Examples

We assess the performance of our sequential designs and compare them with the decisiontheoretical fixed sample designs of M¨ uller et al. (2004). The optimal sequential rule is determined via forward simulation (Section 2.3). Algorithm 2 outlines how to find the optimal fixed sample size. 11

Algorithm 2: Optimal fixed sample size 1. Forward simulation: For b = 1, . . . , B, (b)

1.1. Simulate data x1:t from the GaGa model prior predictive distribution. The GaGa ˆ ν (which can be data-driven estimates). hyper-parameters are set to π ˆ, α ˆ 0 , νˆ, β,ˆ 1.2. For t = 1, . . . , T : • Evaluate posterior probabilities P (δi = 1|x1:t ) and record the terminal decisions di (x1:t ) (Section 2.1). • Record the posterior expected number of true positives TPb =

Pn

i=1

P (δi =

1|x1:t )di (x1:t ) conditional on x1:t . 2. Expected utility: The average across b = 1, . . . , B possible histories approximates the ex P B TP pected utility for fixed sample size t, U¯F IX (t) = B1 b − ct. b=1 3. Optimal sample size: t∗ = arg max U¯F IX (t) is the optimal fixed sample size. In Section 4.1 we present a simulation example based on microarray data from Armstrong et al. (2002). In Section 4.2 we discuss designs for a hypothetical repetition of a microarray experiment (Campo Dell’Orto et al., 2007). Section 4.3 considers a reverse phase protein abundance (RPPA) experiment.

4.1

Simulation Study

We assume that data is collected in batches of 2 arrays per group, with a maximum sample size of 20 arrays per group (i.e. T =10 data batches). We specify three values for the sampling cost c: 25, 50 and 100. For instance, c=50 indicates that the investigator considers it worthwhile to obtain one more data batch if he finds ≥50 new truly DE outcomes. To keep the simulation realistic we estimate the GaGa hyper-parameters based on microarray data from a leukemia study described in Armstrong et al. (2002). We normalized the data using GCRMA (Wu and Irizarry, 2007), and focused on 24 acute lymphoblastic leukemia (ALL) and 18 MLL trans-location (MLL) samples. An EM empirical Bayes fit 12

π = 0.5ˆ π

π=π ˆ

π = 2ˆ π

c

t∗F

t∗S

US∗ − UF∗

t∗S

US∗ − UF∗

t∗S

US∗ − UF∗

25

7

6.0

7.7

7.1

0.4

10.0

34.2

50

5

4.1

16.6

5.0

0.0

6.9

28.7

100

3

3.0

0.0

3.0

0.0

4.0

58.6

Table 1: Simulated data. t∗F : fixed sample size; t∗S : average sequential sample size; US∗ − UF∗ : expected utility for sequential design minus expected utility for fixed sample. estimates π ˆ = 0.063, α ˆ 0 = 4.547, νˆ = 0.155, βˆ = 0.742, νˆ = 677.86. Although we determine the fixed and sequential sample strategies based on π = π ˆ , we also assess the performance of the procedures under model miss-specification by simulating data under π = 0.5ˆ π and π = 2ˆ π . That is, the proportion of DE genes is either half or twice what was guessed at design time. We obtained 1,000 simulations under each scenario. Figure 1 shows the optimal sequential boundary for c = 50 (b0 = 0, b1 = 8.64) and some simulated trajectories for the summary statistic. The optimal fixed sample size for c = 50 is t = 5. Table 1 reports expected utilities and stopping times. When π = π ˆ the fixed sample size and sequential designs perform very similarly, i.e. the sequential design offers little advantages when the hyper-parameters were guessed correctly. For the remaining π values the sequential design presents an increment in expected utility compared to the fixed sample, and it adjusts the sample size as needed. When π = 0.5ˆ π the sequential design stops earlier. Conversely, when π = 2ˆ π it stops later, i.e. it collects additional data so that more DE genes can be found. For instance, for c = 50 the sequential design requires 4.1 data batches when π = 0.5ˆ π and 6.9 when π = 2ˆ π , whereas the fixed design always requires 5.

4.2

Microarray Example

We consider leukemia study of Campo Dell’Orto et al. (2007). The study measure mRNA expression levels for 21 ALL and 15 MLL patients and 54,675 genes via Affymetrix hgu133plus2 13

(b)

1500



1000



● ● ●





4

6

8

0

0

500

Expected increase in True Positives

4000

A priori A posteriori

2000

Expected Utility

6000

2000

(a)

2

4

6

8

10

12

14

2

Sample size

10

12

14

Sample size

Figure 2: Sequential Analysis of Campo Dell’Orto’s data. (a) Utility vs. sample size. White bars: prior expected utility at design time (hyper-parameters estimated from Armstrong’s data). Gray bars: posterior expected terminal utility after observing Campo Dell’Orto’s data. (b) Posterior expected increase in true gene discoveries. All values being above the sequential boundary indicates experimentation to continue up to the maximum sample size.

14

arrays. We obtained the data from GEO (identifier GDS2819, http://www.ncbi.nlm.nih. gov/sites/GDSbrowser?acc=GDS2819) and normalized it using GCRMA (Wu and Irizarry, 2007). We obtain data in batches of 2 microarrays per group, with a maximum T =7 data batches. Importantly, we design the study without looking at the historical data. We only use the historical data later to compare the proposed sequential design versus a fixed sample size design. 4.2.1

Optimal design determination

We estimate the GaGa hyper-parameter values from the Armstrong et al. (2002) data (Section 4.1): π ˆ = 0.063, α ˆ 0 = 4.547, νˆ = 0.155, βˆ = 0.742, νˆ = 677.86. This is reasonable as both studies are carried out in similar patient populations and the same microarray platforms. The white bars in Figure 2(a) show the prior expected utility for all fixed sample sizes. The optimal fixed sample size is t = 5 data batches and has expected utility 2725.9. The optimal boundary is given by b0 = 0, b1 = 8.64 (Figure 2(b)), and the expected utility is 2725.9, the same as the fixed sample design. We emphasize that these utilities are as evaluated at the time of design, i.e. based on the prior predictive distribution. 4.2.2

Results

To assess the performance of the chosen designs we compute the posterior expected utility u¯t (st = 1, x1:t ) in (2) for t = 1, . . . , 7. As shown in Figure 2b, ∆t (T P ) stays always above the sequential stopping boundary and thus the sequential decision is to stop at t = 7. The sequential design (t = 7, u¯t (st = 1, x1:t ) = 7695.4) increases the posterior expected utility over 2-fold compared to the fixed sample design (t = 5, u¯t (st = 1, x1:t ) = 3661.1). Figure 2(a) shows how updating the expected utility after data is observed results in a fairly different picture than that expected a priori. Indeed, the highest posterior expected utility is observed at the time horizon t = 7, which is the sample size chosen by the sequential design. This illustrates the advantage of taking a sequential approach: it learns about

15

(a)

(b)





Π1 = 0.5 , 0.75

10





5

Expected Increase in True Positives

10



5

Expected Increase in True Positives

15



15



Π1 = 0.25 ●





Π1 = 0.5 , 0.75

0

0

Π1 = 0.25

50

100

150

200

250

50

Sample size per group

100

150

200

250

Sample size per group

Figure 3: RPPA data. Posterior expected increase in true gene discoveries. Optimal sequential boundaries for (a) Prior A: α0 =2.18, ν=0.69, β=1.03, µ=3421.9; (b) Prior B: α0 =1.28, ν=0.56, β=1.00, µ=3640.0. the probabilistic nature of the data and therefore it is robust to miss-specifications of the assumptions needed for fixed-sample size calculations.

4.3

RPPA

We discuss the design of a leukemia study described in Kornblau et al. (2006). The study records protein abundance in bone marrow (BMA) and peripheral blood (PB) using RPPA (reverse phase protein arrays) (Tibes et al., 2006). The data record protein expression for 168 proteints and 718 patients, including 435 samples from the bone marrow (BMA) and 283 from peripheral blood (PB). We first find the optimal sequential design for this study, without using the observed data, i.e., as if we were designing the study. We will only use the data later to assess the performance of the proposed sequential design versus a fixed sample

16

70 50 40 30 20 0

10

Posterior Expected Terminal Utility

60

Fixed Sequential

c=5

c=10

Figure 4: RPPA data. Posterior expected terminal utility for sampling costs c=5 and c=10. size design. 4.3.1

Optimal design determination

We design the study with the primary goal of identifying proteins with differential abundance in BMA versus PB. We consider obtaining data in batches of 50 samples per group, with a maximum of 250 (i.e. we do not exceed the 283 available PB samples). As RPPA samples are relatively cheap to obtain, we consider sampling costs c = 5 and c = 10. The data was pre-processed using the R package SuperCurve (Joy et al., 2006). After pre-processing, protein abundance values are centered at 0. As the GaGa model allows positive real values only, we exponentiated all protein abundances. Finding optimal fixed and sequential designs requires simulating data from the GaGa model prior predictive distribution, with reasonable hyper-parameters. To assess design sensitivity to prior assumptions, we consider several hyper-parameter choices. The proteins in the RPPA experiment were chosen carefully and we expect a considerable proportion of DE proteins. Therefore we set π to 0.25, 0.5 and 0.75. Considering that most abundances 17

are expected to be around 1 (after exponentiation), and that the coefficients of variation are typically well below one we consider two specifications: Prior A: Mode for θi = 1, P (θi < e2 |α0 , ν) = 0.95, mode for CVi = 0.1 and P (CVi > 0.01|β, µ) = 0.95. This gives α0 =2.18, ν=0.69, β=1.03, µ=3421.9. Prior B: Mode for θi = 1, P (θi < e3 |α0 , ν) = 0.95, mode for CVi = 0.5 and P (CVi > 0.01|β, µ) = 0.95. This gives α0 =1.28, ν=0.56, β=1.00, µ=3640.0. The two values for the sampling cost c, three choices for π and the two sets of hyper-priors (A and B) give rise to 12 scenarios. 4.3.2

Results

Figure 3 plots the optimal sequential boundaries for the 12 hyper-prior parameter specifications, along with ∆t (T P ) as computed from the observed data. The optimal sequential boundaries are the same for c = 5 and c = 10. The sequential boundaries only differ slightly across scenarios. Furthermore, ∆t (T P ) lies above all 12 boundaries for t = 1, . . . , 4 i.e. none of the designs stops experimentation early. For completeness we include ∆5 (T P ) in Figure 3, although at t = 5 (maximum number of batches) the decision is to stop regardless of ∆5 (T P ). In all scenarios, the optimal fixed sample design is to obtain 1 data batch with 50 samples per group. This example illustrates the potential for drastic differences between fixed and sequential sample designs. Intuitively, a priori we expect that most DE proteins will be found with only 50 samples. However, once actual data is observed our expectations are updated and it is deemed optimal to obtain 250 samples. To assess which strategy actually performs best, Figure 4 shows that the posterior expected terminal utility is several folds higher for the sample size chosen by the sequential design, both for c = 5 and c = 10.

18

5

Discussion

We proposed a sequential approach to sample size calculation in the context of massive multiple hypothesis testing. By incorporating information as new data becomes available, the approach becomes robust with respect to inaccurate a priori guesses. This is highly desirable as typically only vague a priori information is available. The proposal is formulated within a decision-theoretic framework. Although our ideas are applicable in general setups, we emphasize interpretability and ease of use. We defined a simple utility function with a reward for true positives and a single sampling cost parameter c. The utility is straightforward to specify. The parameter c is the number of true positives that make it worthwhile to obtain one more data batch. Also, we proposed terminal decisions that control the Bayesian FDR. Although this is inconsistent with a decision-theoretical setup where all decisions are taken to maximize the expected utility, we feel that practitioners are unlikely to use an approach which disregards FDR control. Of course, it is also possible to explicitly include an FDR term in the utility. But this requires setting additional utility parameters and would compromise the goal of easy usability. We showed that finding an exact solution is computationally unfeasible, and we proposed the use of approximate sequential rules. We monitor the one-step ahead expected increment in true positives and stop the experiment when this increment falls below some boundary. Our proposal is not only intuitive and computationally convenient, but it also satisfies some desirable properties: it includes fixed sample and the myopic designs as particular cases and we showed that the experiment would stop eventually even if no maximum sample size were considered. Finding the optimal boundary can be achieved via forward simulation, which requires a probability model which allows computationally efficient posterior inference. We used the GaGa model for its balance between flexibility and computational efficiency, but our approach admits arbitrary probability models. We illustrated our approach in simulations and two leukemia data sets. Compared to fixed sample calculations the proposed sequential designs are robust to a priori specifications and they may increase the posterior expected terminal utility several fold. Our results show 19

that experiments are terminated early when enough evidence is available, and extended when appropriate. We focused on multiple hypothesis testing. It would be interesting to extend the approach to other setups such as sample classification, clustering or network discovery. This would require adjusting the utility function and possibly the probability model, e.g. to express complex dependencies between outcomes. Overall, we feel that sequential strategies are promising for sample size determination in high-throughput experiments. They save valuable time and resources while guaranteeing that sufficient data is collected to answer the scientific questions.

Appendix We prove Proposition 2.1 under the assumption that Bayes factors are consistent. Let (t)

vi = P (δi = 1|x1:t ) be the posterior probability that δi = 1 conditional on all data up to time t. Intuitively, Bayes factor consistency guarantees that for all i such that δi = 1 both (t)

(t+1)

vi and vi (t)

are arbitrarily close to 1 for large enough t (with large probability). Similarly, (t+1)

(t+1)

(t)

and vi differ by a are arbitrarily close to 0 when δi = 0. Therefore, vi   (t) (t+1) − vi |x1:t , i.e. an vanishingly small amount. As ∆t (TP) can be expressed as Ext+1 vi

vi

and vi

(t+1)

expectation involving differences between vi

(t)

and vi , ∆t (TP) converges to 0.

More formally, let u ∈ < such that 0 < u < 1. Bayes factor consistency implies that (t)

for all ∗ > 0 there is t0 such that for all t > t0 P (vi

< u|δi = 0) > 1 − ∗ for all i such

(t)

(t)

that δi = 0 and P (vi > 1 − u|δi = 1) > 1 − ∗ for all i such that δi = 1. Let Ai be the (t)

(t)

event that vi < u for i such that δi = 0 and the event that vi > 1 − u when δi = 1. Then (t) T (t+1) (t) (t+1) (t) S (t+1) P (Ai Ai ) = P (Ai ) + P (Ai ) − P (Ai Ai ) > 1 − 2∗ . A simple induction argument shows that P

n  \

(t) Ai

\

(t+1) Ai



! > 1 − 2n∗ .

(5)

i=1

Denote  = 2n∗ .  is arbitrarily small as ∗ is also arbitrarily small and n is a finite constant.

20

(t)

Now, the terminal decisions that we adopt are of the form di (x1:t ) = I(vi > 1 − u∗ ), i.e. we decide in favor of δi = 1 whenever its posterior probability is large enough. Setting u to a small enough value in (5) gives that di (x1:t ) = di (x1:t+1 ), which happens simultaneously for i = 1, . . . , n with probability > 1 − . Using this fact and simple algebra ∆t (TP) can be (t+1)

expressed in terms of differences between vi

(t)

and vi n   X (t+1) (t) ∆t (TP) = di (x1:t )Ext+1 vi − vi |x1:t .

(6)

i=1 (t)

Now, since di (x1:t ) ∈ {0, 1} and Ai

T

(t+1)

Ai

(t+1) (t) implies that vi − vi ≤ u, we have that

∆t (TP) ≤ nu. As u can be arbitrarily small, this proves the result.

References S.A. Armstrong, J.E. Staunton, L.B. Silverman, R. Pieters, M.L. Boer, M.D. Minden, E.S. Sallan, E.S. Lander, T.R. Golub, and S.J. Korsmeyer. Mll translocations specify a distinct gene expression profile that distinguishes a unique leukemia. Nature Genetics, 30:41–47, 2002. D.A. Berry and B. Fristedt. Bandit problems. Chapman & Hall, 1985. D.A. Berry, P. M¨ uller, A.P. Grieve, M. Smith, T. Parke, R. Blazed, N. Mitchard, and M. Krams. Adaptive Bayesian Designs for Dose-Ranging Drug Trials, volume V. SpringerVerlag, New York, 2001. D.R. Bickel. Selecting an optimal rejection region for multiple testing: A decision-theoretic alternative to fdr control, with an application to microarrays. Technical report, Medical College of Georgia, 2003. A.E. Brockwell and J.B. Kadane. A gridding method for Bayesian sequential decision problems. Journal of Computational and Graphical Statistics, 12:566–584, 2003. M. Campo Dell’Orto, A. Zangrando, L. Trentin, R. Li, W.M. Liu, G. te Kronnie, G. Basso, and A. Kohlmann. New data on robustness of gene expression signatures in leukemia: 21

comparison of three distinct total rna preparation procedures. BMC Genomics, 8:188, 2007. B. Carlin, J. Kadane, and A. Gelfand. Approaches for optimal sequential decision analysis in clinical trials. Biometrics, 54:964–975, 1998. S.C. Chow, J. Shao, and H. Wang. Sample size calculations in clinical research. Marcel Dekker, New York-Basel, 2003. S.C. Dass and J. Lee. A note on the consistency of bayes factors for testing point null versus non-parametric alternatives. Journal of Statistical Planning and Inference, 119:143–152, 2004. M. DeGroot. Optimal Statistical Decisions. McGraw Hill, New York, 1970. M. DeGroot. Optimal Statistical Decisions. Wiley-Interscience, 2004. D.L. DeMets. Sequential designs in clinical trials. Cardiac electrophysiology review, 2:57–60, 2004. K. Dobbin and R. Simon. Sample size determination in microarray experiments for class comparison and prognostic classification. Biostatistics, 6:27–38, 2005. J.A. Ferreira and A. Zwinderman. Approximate power and sample size calculations with the benjamini-hochberg method. The International Journal of Biostatistics, 2(1), 2006a. J.A. Ferreira and A. Zwinderman. Approximate sample size calculations with microarray data: An illustration. Statistical Applications in Genetics and Molecular Biology, 5(1), 2006b. W.J. Fu, E.R. Dougherty, B. Mallick, and R.J. Carroll. How many sample are needed to build a classifier: A general sequential approach. Bioinformatics, 21:63–70, 2005. S. Ghosal. A review of consistency and convergence of posterior distribution. Technical report, Indian Statistical Institute, 2002. 22

M. Guindani, P. M¨ uller, and S. Zhang. A bayesian discovery procedure. Journal of the Royal Statistical Society, Series B, page to appear, 2009. C. Joy, K. Coombes, J. Hu, and K. Baggerly. R SuperCurve Package, 2006. URL http: //bioinformatics.mdanderson.org/Software/OOMPA. R package version 0.931. S. Kornblau, M. Womble, Y. Qiu, E. Jackson, W. Chen, M. Konopleva, E. Estey, and M. Andreeff. Simultaneous activation of multiple signal transduction pathways confers poor prognosis in acute myelogenous leukemia. Blood, 108:2358–2365, 2006. M.L. Lee and G. Whitmore. Power and sample size for microarray studies. Statistics in medicine, 11:3543–3570, 2004. S. Mukherjee, P. Tamayo, S. Rogers, R. Rifkin, C. Eagle, A. Campbell, T. Golub, and J. Mesirov. Estimating dataset size requirements for classifying dna microarray data. Journal of Computational Biology, 10:119–142, 2002. P. M¨ uller, G. Parmigiani, C. Robert, and J. Rousseau. Optimal sample size for multiple testing: the case of gene expression microarrays. Journal of the American Statistical Association, 99:990–1001, 2004. P. M¨ uller, D. Berry, A. Grieve, M. Smith, and M. Krams. Simulation-based sequential bayesian design. Journal of Statistical Planning and Inference, 137:3140–50, 2006. R. Pahl, H. Sch¨afer, and H.-H. M¨ uller. Optimal multistage designsa general framework for efcient genome-wide association studies. Biostatistics, 10:297–309, 2009. W. Pan, J. Lin, and Le. C.T. How many replicates of arrays are required to detect gene expression changes in microarray experiments? Genome Biology, 3:research0022.1–0022.10, 2002. Kenneth M. Rice, Thomas Lumley, and Adam A. Szpiro. Trading bias for precision: Decision theory for intervals and sets. Technical Report Working Paper 336, UW Biostatistics, http://www.bepress.com/uwbiostat/paper336, 2008. 23

D. Rossell. GaGa: a simple and flexible hierarchical model for differential expression analysis. Annals of Applied Statistics (to Appear), 2009. D. Rossell, P. M¨ uller, and G. Rosner. Screening designs for drug development. Biostatistics, 8:595–608, 2006. D. Ruppert, D. Nettleton, and J.T.G. Hwang. Exploring the information in p-values for the analysis and planning of multiple-test experiments. Biometrics, 63:483–95, 2007. R Tibes, Y Qiu, Y Lu, B Hennessy, M Andreeff, GB Mills, and S.M. Kornblau. Reverse phase protein array: validation of a novel proteomic technology and utility for analysis of primary leukemia specimens and hematopoietic stem cells. Mol Cancer Ther., 5(10): 2512–21, 2006. J. Whitehead. The Design and Analysis of Sequential Clinical Trials, Revised, 2nd Edition. Wiley, New York, 1997. J. Wu and J.M.J. Irizarry, R. with contributions from Gentry. gcrma: Background Adjustment Using Sequence Information, 2007. R package version 2.8.1. A. Zien, J. Fluck, R. Zimmer, and T. Lengauer. Microarrays: how many do you need? Journal of Computational Biology, 10:653–667, 2003.

24

Sequential sample sizes for high-throughput hypothesis ...

as it achieves a good balance between flexibility and computational efficiency. We review the GaGa ..... This would require adjusting the utility function and possibly the probability model, e.g. to express .... Journal of the American Statistical.

242KB Sizes 0 Downloads 184 Views

Recommend Documents

Statistical Power, Sample Sizes, and the Software to ...
Inc. 1996a) and accurately calculates power or sample size for a wide range of statistical tests for a broad range of sta ... When conducting a power analysis, the various inputs (pa- rameters) can come from pilot data, from similar studies ..... Con

Statistical Power, Sample Sizes, and the Software to ...
and Travis 1993) to address conservation and management issues, they need to ... parametric and parametric tests, account for unbalanced study designs, and ...

Highthroughput DNA sequencing concepts and ...
available to many more researchers and projects. However, while ... standing of the technologies available; including sources of error, error rate, as well as the ...... ogy [14] and, recently, IBM's proposal of .... This may open the market further

DIETARY HYPOTHESIS
Note: The world map in Figure 3 is from “The World Factbook”, operated by the ... Thomas F. Spande received a Ph.D. in chemistry from Princeton University in ...

Hypothesis
which is crucially based on Move-F, is provided in section 3, Some theoretical consequences of the proposed analysis are discussed in section 4, followed by concluding remarks in section 5. 2. Takahashi (1993) and Some Problems. 2.1 Takahashi (1993).

Double sequential defibrillation for refractory ventricular fibrillation ...
This is a PDF file of an unedited manuscript that has been accepted for publication. As. a service to our customers we are providing ... Department of Emergency Medicine. American University of Beirut - Medical Center. Sandra Mrad M.D.. Department of

Adaptive spectral window sizes for feature extraction ...
the spectral window sizes, the trends in the data will be ... Set the starting point of the 1st window to be the smallest ... The area under the Receiver Operating.

Efficiently Enabling Conventional Block Sizes for Very ...
Dec 7, 2011 - expose the stacked DRAM to the operating system (OS) by di- rectly mapping the .... filling, and tag updates are all modeled. The ideal-SRAM ...

A4 Magazine Advert Sizes v1b.pdf
Page 1 of 1. A4 Magazine Advert Sizes. Including Buckinghamshire Life, Cheshire Life, Devon Life, Essex Life, Hertfordshire. Life, Kent Life, Lancashire Life, ...

The Social Brain Hypothesis
hypothesis, though I present the data only for the first of ..... rather the ''software programming'' that occurs .... Machiavellian intelligence hypothesis, namely to ...

Hypothesis Testing.pdf
... mean weight of all bags of pretzels equals 5 oz. Ha : The mean weight of all bags of chips is less than 5 oz. Reject H0 in favor of Ha if the sample mean is sufficiently less than 5 oz. Matt Jones (APSU) Hypothesis Testing for One Mean and One Pr

Maintenance" Hypothesis
detached, thus we call the predetermining flakes themselves ventral flakes. .... results in a plunging termination that ruins the core. ACKNOWLEDGEMENTS.

Hypothesis testing.pdf
Whoops! There was a problem loading more pages. Retrying... Hypothesis testing.pdf. Hypothesis testing.pdf. Open. Extract. Open with. Sign In. Main menu.

Riemann Hypothesis
Mar 1, 2003 - shops are collaborating on the website (http:// · www.aimath.org/WWN/rh/) ..... independently discovered some of the develop- ments that had ...

sequential circuits
SEQUENTIAL CIRCUITS. Flip-Flops Analysis. Ciletti, M. D. and M. M. Mano. Digital design, fourth edition. Prentice Hall. NJ.

Testing the Distributional Hypothesis for Collaborative ...
Also the fact that many results can be applied directly to information retrieval has encouraged ... words. A number of web services offer people the possibility to add free form keywords (tags) to ... data collected from LibraryThing. We compare ...

n-acetylglucosamine substrate-site hypothesis for the ...
in [6] . The rates of (pro)insulin biosynthesis are given as the insulin index: this parameter is the .... phloretin plus 20 mM N-acetylglucosamine. ... P =G 0.001 vs 5.

Genome Sizes in Afrotheria, Xenarthra, Euarchontoglires, and ...
the sole carriers of hereditary factors. ... recorded different DNA content in cell nuclei of the same ... nucleotypic effects in cell functions, and the organisms in.

Hypothesis Testing Basics.pdf
Sign in. Page. 1. /. 2. Loading… ... H y pot he s is T e s t s 1 0 1. will culminate in students getting 45 ... think of it as the probability of obtaining a. sample statistic ...

Facebook Friends- A Hypothesis Test for One Population Mean.pdf ...
Page 1 of 3. Facebook Friends. A Hypothesis Test for One Population Mean. Work in groups of at most three. At least one member of the group must have a ...

A Riemann Hypothesis Condition for Metaplectic Eisenstein Series.pdf
Whoops! There was a problem loading more pages. Retrying... A Riemann Hypothesis Condition for Metaplectic Eisenstein Series.pdf. A Riemann Hypothesis ...

Bayesian Hypothesis Test for sparse support recovery using Belief ...
+82-62-715-2264, Fax.:+82-62-715-2274, Email:{jwkkang,heungno,kskim}@gist.ac.kr). Abstract—In this ... the test are obtained by aid of belief propagation (BP).

A New Hypothesis for Biophotonic Communication in ...
International Journal of BioSciences and Technology (2008), Volume 1, Issue 3, Page(s): 1-6 ... phenomena related to biophotonic communication renowned to be caused during Meditation. ..... orienting her towards the path of education and.