Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
Alan L. Hutchison, Mark Maienschein-Cline, Andrew H. Chiang, S. M. Ali Tabei, Herman Gudjonson, Neil Bahroos, Ravi Allada, Aaron R. Dinner

Abstract

To this end, several analytic methods have been developed for detecting periodic patterns. We improve one such method, JTK_CYCLE, by explicitly calculating the null distribution such that it accounts for multiple hypothesis testing and by including non-sinusoidal reference waveforms. We term this method empirical JTK_CYCLE with asymmetry search, and we compare its performance to JTK_CYCLE with Bonferroni and Benjamini-Hochberg multiple hypothesis testing correction, as well as to five other methods: cyclohedron test, address reduction, stable persistence, ANOVA, and F24. We find that ANOVA, F24, and JTK_CYCLE consistently outperform the other three methods when data are limited and noisy; empirical JTK_CYCLE with asymmetry search gives the greatest sensitivity while controlling for the false discovery rate. Our analysis also provides insight into experimental design and we find that, for a fixed number of samples, better sensitivity and specificity are achieved with higher numbers of replicates than with higher sampling density. Application of the methods to detecting circadian rhythms in a metadataset of microarrays that quantify time-dependent gene expression in whole heads of Drosophi/a melanogaster reveals annotations that are enriched among genes with highly asymmetric waveforms. These include a wide range of oxidation reduction and metabolic genes, as well as genes with transcripts that have multiple splice forms.

Author Summary

Many genes’ activities vary periodically. For example, circadian rhythms repeat daily With the light-dark cycle. Understanding how such rhythms couple to biological processes requires statistical methods that can identify cycling time series in typical genome-Wide data. In this paper, we improve on a method used to identify cycling time series by better estimating the statistical significance of periodic patterns and, in turn, by searching for a Wider range of patterns than traditionally investigated. We apply these methods to a compilation of data on gene eXpression in fruit flies, an important model organism. We find that our method allows us to discover rhythmic biological activities that the other methods tested are unable to reveal.

Introduction

Diverse fundamental biological functions such as cell division, energy metabolism, and sleep are periodic, and a growing body of evidence implicates temporal dysregulation as a contributing factor to depression, neu-rodegeneration, cardiovascular disease, and metabolic disorders in higher organisms [5—9]. Arguably the most well-studied periodic patterns are circadian rhythms: oscillatory changes in gene expression, metabolism, physiology, and behavior with approximately 24-hour (24 h) periods that enable organisms to anticipate and respond to daily changes in their environment, such as nutrient accessibility, temperature, and light [10—13].

The components of the core clock are well characterized and are strongly conserved across a wide range of species [14, 15]. However, it remains to be determined how this clock couples to other molecular processes. Moreover, these interactions are likely to depend on tissue type and environmental conditions [2, 7, 11, 16, 17]. There is thus a need to identify molecular profiles that cycle and to characterize them as a function of conditions. The advent of high throughput methods for measuring gene expression now makes transcriptome-wide studies of this nature possible. Previous work suggests that hundreds, possibly thousands, of genes are regulated by circadian clocks [10, 15, 18].

As a result, gene expression time series are typically sparsely sampled (e.g., every 2—4 hours (h) in circadian studies), often without multiple measurements per time point, which we refer to here as “replicates”. These experimental limitations result in low signal-to-noise ratios that prevent straightforward identification of cycling gene expression.

These methods can aid researchers in assessing the tradeoffs between the amount of data acquired, statistical precision, and breadth of biological discovery. While a number of different methods have been proposed for identifying cycling time series [19—30], further analysis is needed to guide selection of the best method(s) for a given situation and to aid in design of improved computational methods and further experiments.

The original method uses a conservative estimation for its p-values and a cosine as its only reference waveform. Here, we introduce a procedure, empirical ITK_CYCLE with asymmetry search, that provides accurate empirically-calculated p-values for arbitrary waveforms. We test its performance for detecting way analysis of variance (ANOVA) [27]. The simulated data allow us to examine how performance varies with sampling density, number of replicates and/ or periods, noise level, and waveform. Most methods provide accurate rhythm detection when sampling density is high and noise is low. However, we find that the choice of method significantly affects rhythm detection when data are limited and/or noisy. In particular, ITK_CYCLE, F24, and ANOVA consistently outperform the other methods and offer distinct advantages for certain types of data. Our improved method, empirical ITK_CYCLE with asymmetry search, performs best of all for data that include asymmetric waveforms. Application of our improved method, empirical ITK_CYCLE with asymmetry, to a metadataset of whole head D. melanogaster circadian mi-croarrays [27] reveals a strong lights-on peak in eXpression for genes involved in glutathione metabolism, high enrichment for genes involved in oxidation reduction, many more metabolic genes cycling than previously appreciated, and rhythmic genes with transcripts that have alternative splicings.

Methods

Overview

The methods that we test are cyclohedron test [20, 21] , address reduction [22, 23], stable persistence [24, 25], F24 [31, 32], one-way analysis of variance (ANOVA) [27], and ITK_CYCLE [26]. We describe each briefly below and note specific features; additional details can be found in the references introducing the methods.

Cyclohedron test, address reduction, stable persistence, and ANOVA seek to identify patterns without specifying the waveform a priori. Address reduction, cyclohedron test, and stable persistence test for monotonicity. ANOVA compares the means of different time points with their variances to determine if differences are significant.

These methods also test for a specific period. As mentioned above, here we assume a period of 24 h, but the period of the reference can be varied, in the same manner that the phase can be varied, to search for rhythms on other time scales.

Cyclohedron test [20, 21] maps data to a cyclohedron and joins data points into sets according to their adjacency in rank-ordering. Monotonicity is quantified by how the sizes of the sets scale as more data points are included. Cyclohedron test has an exact null distribution computable from a generating function. The domain of test statistics increases very quickly with the number of data points, however, so Monte Carlo (MC) sampling, in which representations of the null model are randomly generated and evaluated, is required to estimate p-values if there are more than approximately 20 time points due to the computational cost of the generating function.

Address reduction [22, 23] measures the entropy of the dataset by comparing the rank-ordering of adjacent time points. Low entropy data are monotonic and score higher in the method. The null distribution for address reduction is estimated by MC sampling.

In stable persistence [24, 25], local minima are paired with surrounding local maxima, and the “persistences” of these features are characterized by the differences in ranks of the paired extrema. A hierarchy of such features is established, and the test compares the global persistence to local ones. In this way, stable persistence tries to robustly assess overall monotonicity of a time series. The null distribution for stable persistence is estimated by MC sampling.

One-way ANOVA is a standard statistical test of the equivalence of means in several groups. In this case, each time point is a different group, and ANOVA is equivalent to testing for any statistically significant variation across the time points. Because eXpression measurements are averages over many cells and different time points come from different samples (as the measurement is destructive), only synchronized, consistent variation across all samples can generate a statistically significant trend. By this reasoning, significant changes in eXpression across time points can be attributed to time-dependent eXpression within the population, such as circadian rhythms. ANOVA tests for these time-dependent changes in eXpression. ANOVA has an exact null distribution derived from an assumption of normally distributed data; unlike the other five methods, however, ANOVA requires replicates to estimate the variance of experimental measurements at each time point.

F24 [31, 32] assesses periodicity by focusing on the 24 h period of the Fourier transform of the data. The test statistic for F24 is the projection of the data onto the 24 h Fourier basis function, and the null distribution is obtained by recom-puting the test statistic over repeated random permutations of the data. The phase is determined by projecting the data onto the cosine part of the Fourier basis function and finding the optimal phase for the projection. We find that the null distribution can be modeled by the Gamma distribution (81 F ig. ), which we parameterize from the mean and variance of the null distribution. We estimate the null distribution from a small number of permutations (usually 100). This allows more rapid and precise computation of p-values than can be obtained by standard permutation. Testing periods other than 24 h is accomplished simply by changing the period of the Fourier basis function used to compute the test statistic.

The numerator is the number of pairs that vary concordantly between the two time series minus the number that vary discordantly (Fig. 1A). Every possible pair is included, not just ones between neighboring points in the time series. The denominator is the total number of pairs, which normalizes the value of T to the range [—1, 1]. Perfectly correlated series score 1‘ = 1, perfectly anti-correlated series score 1‘ = —1, and uncorrelated series score 1‘ = 0. Like the cyclohedron test, the null distribution for ITK_CYCLE can be computed exactly from a generating function [33], although again the test statistic domain grows quickly with time series size (becoming impractically large at 100—200 time points with present computing power). However, for large time series the ITK_CYCLE null distribution is approximately normal, allowing for a convenient, fast p-value estimate.

Improvements to the JTK_CYCLE method

1) is calculated for a specific reference time series, and thus ITK_CYCLE typically tests against a family of curves (e.g., to consider the possible phases of a waveform, as illustrated in Fig. 1B). It is thus necessary to account for multiple hypothesis testing across reference waveforms in assessing the significance of the results. Hughes et al. [26]

E 3 A A g Time s 'r r a E A AA Time Time

JTK_CYCLE compares all possible pair relations for a time series to those for a reference waveform. (A) JTK_CYCLE tests for pairwise agreement between a reference (blue) and a signal (cyan) time series. Three discordant pairwise relationships are indicated by red lines. (B) The previous implementation compared a time series to a set of phase-shifted cosines. (C) We add a set of asymmetric waveforms to the reference. An example is shown here with the same phases as in A.

This method is known to be conservative [fl], and we illustrate this fact here eXplicitly for ITK_CYCLE (Fig. 2). These considerations motivate a new procedure for estimating the significance of the results, which we describe. We end this section by discussing the comparison of time series to reference waveforms (Fig. 1C) other than the cosine waveform that was used originally. Together, our improvements allow for the ITK_CYCLE method to include additional reference waveforms in its rhythm detection without compromising sensitivity and specificity.

By definition, a p-value is the likelihood of obtaining a test statistic equal to or more extreme than the value that is observed if the null hypothesis is true—it increases cumulatively as one progresses through a set of rank ordered test statistics. In the case of ITK_CYCLE, under the null hypothesis, time series values are independent (and generated by the same noise distribution) and so the rank ordering of time series values is random. For a dataset generated from this null model, the p-values should be uniformly distributed from 0 to 1, exclusive: the highest Kendall’s 1‘ out of N tests should have a p-value of 1 / (N + 1), the second highest test statistic has a p-value of 2/ (N + 1), and the ith highest test statistic has a p-value of i/ (N + 1) [35]. Restated, the p-values should be a linear function of the ranks (black lines in Fig. 2).

This procedure biases the p-values (the blue and green lines in Fig. 2A) such that they underestimate the true probability of obtaining test statistics by chance (the black line in Fig. 2A). The previous version of ITK_CYCLE corrects for underestimating the p-values with the Bonferroni correction, which controls the family-wide error rate (FWER) by multiplying the p-values by the number of hypothesis tests being performed. The FWER is the probability that there is at least one false positive for a given threshold. Therefore, a threshold of 0.01 means that there is a 1% chance that the list of time series with a Bonferroni adjusted p-value below 0.01 contains a false positive. This method, while rigorous, is overly conservative and overcompensates for the bias that comes from selecting the lowest p-value (blue and green lines in Fig. 2B). The likelihood of false positives is greatly reduced, but so is the likelihood of identifying true positives.

The FDR is the fraction of the time series that are identified as cycling that are false positive. For example, a Benjamini-Hochberg adjusted p-value cutoff of 0.05 means that 5% of the positives are false. This is a less stringent constraint than the FWER. In this procedure, the p-values are also multiplied by the number of hypotheses tested, as in the Bonferroni procedure. However, the p-values are additionally ordered from lowest to highest and then divided by their rank order (the lowest p-value has rank order 1, the second highest p-value has rank order 2, and so on). The p-values are also adjusted such that there is no change in ordering: if the originally lowest p-value is adjusted so that it is higher than the originally second lowest p-value, the lowest p-value takes the value of the adjusted second p-value so that the ordering is not violated. The same holds for the relationship between the second and the third lowest p-values and so forth. While the Benjamini-Hochberg procedure is a reasonable approach to multiple hypothesis testing in general, it does not account for the selection step in ITK_CYCLE; it still is thus overly conservative

Since we test a family of curves (e.g., spanning phases), we focus on positive correlations and compute one-sided p-values. In the present study, these “empirical” p-values are based on 2 x 106 random time series and are nearly uniformly distributed, as desired (Fig. 2D). Though this empirical calculation is more computationally expensive than the application of the Bonferroni correction or the Benjamini-Hochberg correction, we show that empirically calculating the p-values results in better rhythmic detection and biological insight. We term this improved method empirical ITK_CYCLE, as we empirically calculate the p-values after selecting the highest 1‘ value for each time series.

These have been adjusted on the basis of correcting for multiple hypothesis testing across different waveforms for a single time series. When we compare different time series to each other, we have to correct again for multiple hypothesis testing, this time across time series. To do this we use the Benjamini-Hochberg correction, as in the original implementation of the method [26]. When we refer to the Bonferroni, Benja-mini-Hochberg, or empirical method, we refer to corrections across different waveforms for a given time series; all corrections across time series are with the Benjamini-Hochberg method. While the Bonferroni adjusted p-values, Benjamini-Hochberg adjusted p-values, and empirical p-values represent different quantities (the FWER, the FDR, and the p-values, respectively) they are all at least as conservative as the “true” p-values in the range that we are examining (compare blue and green lines with black lines in Figs. 2B, C, and D). This means that the FDRs that result from the Benjamini-Hochberg correction between time series are more conservative than the true FDRs.

There is no a priori reason biological time series need be sinusoidal [37, 38], so it is of interest to test additional waveforms. In this regard, it is important to keep in mind that for ITK_CYCLE the rank order of the points in the reference matters, so we can represent a wide range of simple waveforms (e.g., Fig. 1C) by a triangle function with a specified separation between the maximum and the minimum. This allows us to avoid functionally defining an asymmetric cosine waveform. For the time series that we examine in this paper the difference is insignificant (S2 Fig.). We term the size of the interval from the maximum to the minimum the “asymmetry”, and we express the asymmetry here in units corresponding to a 24 h period. In this notation, a cosine has an asymmetry of 12 h, while a time series with an asymmetry of 8 h has a more rapid fall than rise (the values decrease over 8 h and increase over 16 h). The triangle reference waveforms have the same monotonicity as a cosine, and we keep the convention that the peak value corresponds to the phase of the time series.

In the former case, we denote searching over asymmetry values in steps of 2 h “by 2 h”, in steps of 4 h “by 4 h”, etc. We expect these additional waveforms to be more sensitive to asymmetric patterns of gene expression, resulting in discovery of additional rhythmic time series. It is important, however, to be cognizant of the fact that we are increasing the total number of hypotheses that we test, resulting in a greater need for the empirical correction procedure. Fig. 2 shows the different correction methods for the minimum p-values for ITK_CYCLE with searching over 12 phases (every 2 h, green line) or searching over 12 phases and 11 asymmetries (every 2 h, blue line). The added hypotheses for searching across asymmetries result in larger selection bias When choosing the highest 1‘ value (Fig. 2A) as well as larger correction biases (Fig. 2B and C) than When only searching across phases. The empirical calculation of the p-values improves as the number of tests increases as well (Fig. 2), further justifying its use.

Results

Simulated data benchmarks

We employ the first simulated dataset to understand the sensitivity of each method to different shapes of time series. It comprises four types of waveforms: sine, ramp (a triangle with maximum asymmetry), impulse, and step, as well as an equal number of time series consisting of Gaussian noise. We compare all the pre-cision-recall curves for all the methods on these data via the area under the receiver operating characteristic (AUROC), a measure of the sensitivity and specificity of the rhythm detection methods that does not depend on the proportions of positives and negatives in the dataset. The second simulated dataset contains 10% rhythmic time series of triangle waveform with uniformly distributed phases and asymmetries and 90% time series consisting solely of Gaussian noise. We use it to further assess the importance of considering asymmetric waveforms, and we eXplore how multiple hypothesis correction impacts the results when the true positives represent a relatively small fraction of the simulated time series, as we eXpect to be the case in genome-wide studies.

To construct the first dataset described immediately above, for each of the four waveforms in Fig. 3A we generated 10,000 time series with uniformly distributed random phase shifts (always with a 24 h period) and added Gaussian noise to each point with a standard deviation of 25% or 50% of the total waveform amplitude, examples of which can be seen in Fig. 3B. We tested data with 4, 6, 8, or 12 evenly spaced points per 24 h period, and 1, 2, 3, or 4 replicates per time point (which is the equivalent of 1, 2, 3, or 4 periods per time series). At each spacing and replicate count we also generated 10,000 time series of Gaussian noise to serve as true negatives. The cyclohedron test, address reduction, stable persistence, and F24 are designed for single-replicate data, so we treated replicates as subsequent days of data, yielding multiple-period time series.

The receiver operating characteristic (ROC) curve plots the true positive rate (TPR) as a function of the false positive rate (FPR) as the threshold for calling a time series as a positive is varied. The TPR and FPR are the fractions of the 10,000 simulated or Gaussian noise time series determined to be rhythmic at a threshold, respectively, and the threshold is varied over the entire range of false positive scores, such that the FPR ranges from 0 to 1. The AUROC is the integral of this curve; perfect classifiers have an AUROC of 1.0, while random classifiers have an AUROC of 0.5. An advantage of the AUROC as a metric is that it does not depend on the proportions of positives and negatives in the dataset because the TPR and FPR are calculated separately, i.e., they are normalized by the total number of positives and negatives, respectively. For stable persistence, the cyclohedron test, and address reduction, we calculated the AUROC from the test statistics themselves as opposed to the p-values, which we use for the latter three methods. Although the AUROC for ITK_CYCLE can be in principle be computed directly from the Kendall’s 1‘ statistic, we include the multiple hypothesis testing correction because it impacts the TPR and FPR in practice; in particular, aggressive correction can lead to a loss of rank information because p-values must be less than or equal to 1. Expression Expression Time

Examples of simulated data. (A) Different waveforms simulated with a 24 h period. From left to right, cosine, ramp, step, and impulse (width at half-max is 2 h). Waveforms in figure may not be to scale. (B) Cosine in black, with Gaussian noise with standard deviation of 25% (blue) or 50% (green) of amplitude.

ANOVA requires multiple measurements at each time point to determine the variance, so we define ANOVA to have an AUROC of 0.5 (performs no better than random guessing) when there is only one replicate. At 25% noise and high sampling rate, the cyclohedron test, address reduction, and stable persistence all perform roughly equivalent to the ITK_CYCLE methods, F24, and ANOVA. However, the former do noticeably worse than the latter at 50% noise. Empirical ITK_CYCLE outperforms original ITK_CYCLE, F24, and ANOVA for the sine and ramp waveforms, while ANOVA generally outperforms the other methods for the step and impulse waveforms.

Therefore, we do not sample phases and asymmetries more densely than the resolution of the data (e.g., if the experimental time points are spaced every 4 h, then we do not test phase values spaced every 2 h). We break this rule in Fig. 4 for the time series with 4—8 points for consistency of the figure. Sampling phases and asymmetries more densely than the resolution of the data needlessly reduces the power of our test, but does not affect the analysis in Fig. 4.

This is to be eXpected since the Benjamini-Hochberg method is more conservative than the empirical method but less conservative than the Bonferroni method. An additional detail is that the original ITK_CYCLE here uses a cosine as a reference waveform, in comparison to the triangle used by the other ITK_CYCLE methods. The methods that use the triangle waveform do not do significantly worse than the methods that use the cosine waveform in any of the cases, justifying the use of the triangle waveform for rhythm detection.

For example, a sine wave sampled at 4 time points per period for multiple periods has extrema at every other time point. Because cyclohedron test, address reduction, and stable persistence are essentially tests of monotonicity, they fail to detect the sparse periodic pattern in such data. In fact sparsely sampled data sometimes results in scores consistently lower than expected by chance, leading to the AUROC values less than 0.5 for these methods on some datasets in Fig. 4.

If data are sparse or noisy, however, method choice can significantly impact rhythm detection. In such cases, we find that ANOVA, F24, and ITK_CYCLE consistently better distinguish true and false positives. Empirical ITK_CYCLE outperforms ANOVA, F24, and original ITK_CYCLE for sine and ramp waveforms, but ANOVA performs better for impulse waveforms.

The total number of samples required for an experiment is the product of the number of time points and the number of replicates. Consequently, it is important to consider how best to apportion resources. To this end, in Fig. 5 we compare possible combinations of numbers of time points and replicates that give rise to 12 or 24 total samples (see S4 Fig. for additional waveforms and numbers of samples). For this comparison, we consider sampling at a density of 4 points per period to be the minimal requirement for the identification of rhythmicity. Furthermore, we focus on genome-wide experiments where the experimental design is such that there is no meaningful difference between data collected over multiple periods and data collected at the same sampling rate in replicate over a single period. This assumption does not hold for experiments that follow the response to a synchronization event or other perturbation because time points from successive periods are not equivalent.

For ITK_CYCLE and F24, the choice is less clear, but greater improvement is obtained with replicate increases in the case of the step and impulse waveforms (S4 F ig.). By contrast, original ITK_CYCLE performs slightly better for sinusoidal waveforms with higher numbers of time points, but empirical ITK_CYCLE does not. Overall, the results suggest that limited resources are better directed at increasing replicate numbers than the density of time points.

Given the importance of replicates in improving sensitivity, we also explored interpolating neighboring time points to create pseudo-replicates, which would double the number of time points in the data (S5 Fig.). However, this requires recomputing null distributions via MC sampling because the construction procedure introduces correlations between data points, resulting in p-value underestimates if not corrected. We found that the pseudo-replicates improved the performance mainly of ANOVA when the replicate number was low (e.g., 1 or 2); in particular, they allowed ANOVA to be applied and give good results for the single replicate case (S6 Fig.). We stress that 2 or more biological replicates should be obtained if at all possible, and we do not recommend using the pseudo-replicate approach if sufficient data are available.

Our second benchmark comprises 15,840 times series, which was chosen to allow equal numbers of time series with different phases and asymmetries. 10% of the time series were generated from a triangle waveform with noise added and 90% were generated entirely from Gaussian noise. This composition was chosen to be reflective of a genome-wide dataset. The rhythmic time series were 24 points long, with 2 periods, each with 12 time points.

In both cases, phases (peak values) were uniformly distributed over the possible discrete values. We added Gaussian noise with a standard deviation of either 25% or 50% of the amplitude of the time series, as previously described. We tested these data against the empirical ITK_CYCLE method with various asymmetries as well as original ITK_CYCLE, ANOVA, F24, and Benjamini-Hochberg adjusted ITK_CYCLE with various asymmetries for comparison. In all cases the ITK_CYCLE methods used the triangle waveform as the reference waveform, as it was the waveform used to generate the data.

The methods shown yield comparable numbers for p-values less than 0.05, a reasonable threshold (Fig. 6A and B). However, in reporting total cycling numbers, it is important to correct for the fact that we are testing many time series (as opposed to testing many waveforms for a single time series, as previously). The p-values of empirical ITK_CYCLE are approximately uniformly distributed (Fig. 2D), as are those of ANOVA and F24, which satisfies the assumptions of the Benjamini-Hochberg correction [36], described above, so we use it for this purpose. We also apply the Benjamini-Hochberg correction to the original ITK_CYCLE with the intra-time series Bonferroni and Benjamini-Hochberg corrections discussed previously, which results in underestimates of the true FDR since their adjusted p-values are conservative (Fig. 2B and C).

6C and D, we see that the performance of the methods differs considerably when controlling for the false discovery rate (FDR). For these curves, the proportion of false positives identified as cycling matches the FDR. Specifically, the Benjamini-Hochberg correction (for many time series) penalizes methods with many p-values clustered at relatively high values (corresponding to a rapid rise toward the right of Figs. 6A and B, as for ANOVA and F24). Thus despite the fact that ANOVA and F24 perform comparably to ITK_CYCLE in the AUROC analysis (Fig. 4), their p-values provide less discrimination between time series, and thus they provide less sensitivity for a given FDR. In addition to looking at AUROC scores in Fig. 4 and time series identification in Figs. 6C and D, we also computed the Matthews

A score of 1 indicates that a method correctly identified all true positives and true negatives, while a score of —1 indicates that a method yielded all false positives and false negatives. S7 Fig. shows that the ITK_CYCLE methods have higher-quality classification ability than F24 and ANOVA for these simulated data. Furthermore, the figure shows that empirical ITK_CYCLE with asymmetry search performs equally well with and without asymmetric time series, whereas the ITK_CYCLE methods without asymmetry search perform worse when the dataset includes asymmetric time series.

88 Fig. examines how the inclusion of different asymmetries affects rhythm detection.

Microarray metadataset

[27] previously assembled a metadataset comprised of data from four DNA mi-croarray studies of Drosophila melanogaster under light-dark (LD) conditions (from Ceriani [40], Claridge-Chang [18] , Lin [41], and Ueda [42]). We do not include a fifth dataset from that study [15] , because it was limited to dark-dark (DD) conditions. Here, we discuss issues that arise from merging data from different laboratories and use the resulting metadataset to test the methods. We find that empirical ITK_CYCLE with asymmetry search identifies a larger number of rhythmic genes and, in turn, enriched annotations among those genes, such as oxidation reduction, glutathione metabolism, and alternative splicing.

All of the measurements in the contributing studies are at intervals of 4 h. Time points for circadian LD time series are referenced as zeitgeber time points (ZT); the beginning of the light period is ZTO. Under 12 hours of light and 12 hours of dark, ZT24 is the equivalent of ZTO. Three studies sampled at ZTO, 4, 8, 12, 16, and 20, and the fourth (Ueda [42]) sampled at ZT1, 5, 9, 13, 17, and 21. We found that the differences in sampling protocols, together with variations from one laboratory to another, consistently gave rise to a jagged structure in the time series of known cycling genes (Fig. 7A). Microarray-wide normalization techniques such as quantile normalization were unable to produce curves consistent with independently measured proflles. Instead, we found that the best approach was to convert the values in each time series to Z-scores—i.e., for each gene in each dataset, we subtract its mean eXpression level and divide by its standard deviation. Then we pool the Z-scores to generate the metadataset. Fig. 7 illustrates the effect of the processing step on Pdpl, a known cycling gene. This method is equivalent to treating measurements from the same zeitgeber time point as replicates. For probes that corresponded to the same gene, we chose the probe with the highest mean eXpression value to use in the analysis. This reduced 14,010 probes to 11,625 genes.

To prevent the need to recalculate the null distribution for every pattern of NAs in the data for empirical ITK_CYCLE (there were 5005 unique NA permutations in the data), the NAs were replaced by random noise drawn from a Gaussian distribution with mean and standard deviation that match those of the data on the whole. While this adds noise to the time series, it should not have a large effect given that each time series has 57 points. To mitigate the impact of this procedure on our study, however, time series that had more then half their points as NA were discarded from the dataset, leaving 9,313 out of 11,625 genes. We consistently used the dataset resulting from these preprocessing steps for all our analysis to ensure that comparisons between methods were fair; where comparisons with and without NA substitution were possible, we found that NA substitution led to slight increases in cycling numbers in all cases except ANOVA (114 vs. 101). However, these differences did not change any of the ontological results (discussed below).

To evaluate the methods against genes for which we know the rhythmicity a priori, we compared the p-values for 6 positive examples (per, tim, vri, Pdpl, cry, and Clk) and 4 negative examples (cam, RpL32, cyc, and dco). S9 Fig. shows the performance of the different methods for the known positive and negative examples. Stable persistence, the cyclohedron test, and address reduction all have false negatives. The ITK_CYCLE methods, ANOVA, and F24, however, detect all of the known cycling genes and none of the non-cycling genes as rhythmic.

8). As in Fig. 6, the Benjamini-Hochberg correction decreases the sensitivity of ANOVA and F24 relative to ITK_CYCLE for a given FDR (compare Fig. 8A with Fig. 8B). Choosing a Benjamini-Hochberg adjusted p-value cutoff of 0.05 (i.e., 5%), the number of genes and overlap between methods can be seen in Figs. 8C and D. All the ITK_CYCLE methods outperform F24 and ANOVA. Empirical ITK_CYCLE with asymmetry search by 4 h (eITK_aby4) identified the most genes, showing a clear improvement over the Bonferroni (ITK) and Benjamini-Hochberg (ITK_BH) methods with asymmetry search by 4 h (aby4); e]TK_aby4 also identified more cycling genes than methods without asymmetry search, and the genes were distinct.

For ITK_CYCLE without asymmetry search, there were only 6 hypotheses tested per gene time series (for each of the 6 phases searched), for which the Bonferroni and Benjamini-Hochberg correction across waveforms is very slight. For ITK_CYCLE with asymmetry search every 4 h, the number of hypotheses tested becomes 30, for 6 different phases paired with 5 different asymmetries, which results in a more stringent correction by the Bonferroni and Benjamini-Hoch-berg methods. As experimental sampling rates and sampling densities enable more extensive searching of phases, periods, and asymmetries, we expect the advantage for empirical ITK_CYCLE relative to the original formulation to grow because the Bonferroni correction strongly penalizes adding hypothesis tests. Provided that sufficient permutations are performed, empirical ]TK_CYCLE provides the more robust identification of rhythmic genes. Another way of Viewing this difference between the inclusion and exclusion of asymmetry search is by examining the distributions of the Bonferroni-adjusted p-values against the empirical p-values, as in 810 Fig. With asymmetry search, the empirical p-values are significantly lower than the Bonferroni-adjusted p-values, a pattern that is less pronounced without asymmetry search.

Comparing the two sets of cycling genes in S 12 Fig., we find that searching by 4 h excludes genes with asymmetries 8 to 16 h that are barely below the adjusted p-value of 0.05 (upper left quadrant), while searching at asymmetries of 8 and 16 h excludes genes that have extreme asymmetries.

Figs. 813 and 814 show that there is no substantial difference in genes identified as cycling or in ontological results (discussed below). This can be attributed to the fact that across many time points (57 in the case of the metadataset), the differences between the cosine and triangle waveform are slight (82 Fig).

We compared our results to those of Keegan et al. [27], an earlier analysis of this metadataset. There were two main differences in the way we constructed the dataset: we excluded time series that had more than half their values as NA, and we excluded the dark-dark (DD) McDonald dataset, as discussed above. Of the 200 genes identified as cycling by Keegan et al., 169 remained after preprocessing to remove time series with more than half of their values as NAs. Of those 169, 111 had Benjamini-Hochberg adjusted p-values less than 0.05 for the empirical ITK_CYCLE with asymmetry search by 4 h (eITK_aby4). 58 genes that were identified as cycling by Keegan et al. were not identified by e]TK_aby4. 815 Fig. compares the cycling genes identified by Keegan et al. with the cycling genes identified by e]TK_aby4. Keegan et al. identified genes as cycling primarily on the basis of scoring well (p < 0.05) on several tests following pre-screening by ANOVA. 815A Fig. shows a comparison of the number of tests passed after the ANOVA pre-screening with the Benjamini-Hochberg adjusted p-value from e]TK_aby4. While there appears to be a weak relation between the number of tests passed and the p-value, there is not a clear pattern that would enable one to predict the cycling genes common to both Keegan et al. and eITK_by4. 815B Fig. shows the maximum amplitude measurements (after Z-scoring) for the genes identified as cycling by Keegan et al., which are organized by whether they are identified as cycling by eIT-K_aby4 as well. The genes identified by Keegan et al. but not by e]TK_aby4 tend to have larger maximum amplitudes than the ones identified by both. The ANOVA pre-screening in Keegan et al. can account for this difference; our results with empirical ITK_CYCLE suggest that there are many cycling genes with lower amplitudes. 815C Fig. shows the asymmetries of the genes identified by Keegan et al. as cycling, as determined by e]TK_aby4. A large number of genes identified by Keegan et al., but not by e]TK_aby4, have asymmetry of 16 h. The bias in the earlier study may reflect the fact that one of the tests that Keegan et al. employs is based on correlation with the gene per, which has an asymmetry of 16 h. More generally, Keegan et al. fail to identify 231 genes as cycling that e]TK_aby4 identifies with Benjamini-Hochberg adjusted p-values below 0.05. Of these 231, 82 have Benjamini-Hochberg adjusted p-values below 0.01, 65 have values below 0.005, and 16 have values below 0.001.

In addition to comparing our results to those of Keegan et al., we also compared our results to those of Wijnen et al. [32], who identified 336 genes as rhythmic using an F24-based method. Again, we excluded time series that had more than half their values as NA, and we excluded the dark-dark (DD) McDonald dataset. 816 Fig. shows a comparison of the genes that are identified as rhythmic by e]TK_aby4, Keegan et al., and Wij-nen et al. Whereas 31 genes that Keegan et al. identified as rhythmic were removed by the empirical ITK_CYCLE analysis preprocessing, 57 genes that were identified by Wijnen et al. were removed by the empirical ITK_CYCLE analysis preprocessing due to more than half their time points being NA. These genes are “unassigned” in 816A Fig. because an asymmetry estimate is not available. Wijnen et al. and eITK_aby4 jointly identified 120 genes as rhythmic, of which Keegan et al. identified 59 as well. Wijnen et al. uniquely identified 177 genes as rhythmic, whereas e]TK_aby4 uniquely identified 167 genes. A comparison of the asymmetry distributions for all the genes (816A Fig.) shows that they are similar for e]TK_aby4 and Wijnen et al.

As a first step toward validation the new genes that e]TK_aby4 exclusively identified as rhythmic (i.e., those genes not previously identified by Keegan et al. or Wijnen et al. ), we examined the literature for references that independently suggest that these genes are cycling. Specifically, for each gene, we identified the references in FlyBase (http://flybase.org) that mention the gene. Of those references, those that have the term “circadian” in their title or abstract were identified. S 16B Fig. shows the distribution of genes based on their citation in FlyBase by a “circadian” paper, by the original five dataset papers [15, 18, 40—42] , or by neither. Genes identified by “circadian” papers but not by the original five dataset papers represent further validation that the genes that we select as rhythmic are related to circadian processes.

Kadener et al. [43] assayed for genes regulated by the gene Clk and referenced 6 of the genes not mentioned by the original five papers out of a total of 32 genes, which has less than 1.6% probability of occurring by chance (Fisher’s Exact Test unadjusted p<0.016 [44]). One gene referenced by Kadener et al. as well as Abruzzi et al., who also assayed for genes regulated by Clk, is cabut (cbt, CG4427, FBgn0043364, referred to as EP2237 by Kadener et al.). The gene cbt was previously unidentified as having rhythmic expression. The average time series from the metadata can be seen in 817 Fig. The gene cbt is a metal-ion binding transcription factor downstream of the INK cascade and is involved in morphogenesis [45—48]. It has an asymmetry of 4 h, potentially explaining why it was missed by previous methods but identified by eITK_aby4. Abruzzi et al. [49] also discuss another Clk-regulated gene that was uniquely identified by e]TK_aby4 as rhythmic, twins (tws, CG6235, FBgn0004889), seen in 817 Fig. It has an asymmetry of 20 h, which explains how, like cbt, it could have been missed by previous methods. These genes, though previously unidentified as rhythmic, are strong candidates for having roles in circadian regulation and processes based on our identification of them as rhythmic and the work of Kad-ener et al. and Abruzzi et al. This warrants further experimental studies of these genes in a circadian context as well as the other genes that we have identified.

Fujikawa et al. [50] identified 114 genes that are up-regulated and down-regulated in the head of D. melanogaster following 24 h of starvation. 16 of these genes are not mentioned by the original five papers but are identified as rhythmic by eITK_aby4, which has less than 0.3% probability of occurring by chance (Fisher’s Exact Test unadjusted p< 0.003). Fujikawa et al. refer to several genes from the circadian dataset papers that also appear in their lists of differentially expressed genes, but they do not associate rhythmic behavior with all the genes that they describe. In addition to the gene cbt, Fujikawa et al. reference other genes that were previously unidentified as rhythmic: Esterase-Q (Est-Q, CG7529, FBgn0037090) and 1, 4-Alpha-Glucan Branching Enzyme (AGBE, CG33138, FBgn0053138). Both have asymmetries of 16 h, which is also outside the range of standard symmetric-wave-form detection (817 Fig). The identification of these genes as rhythmic reinforces the connection between metabolism and circadian regulation and indicates other potential areas of experimental exploration. To further understand the relationship between circadian regulation that we see in the genes e]TK_aby4 has identified as rhythmic and biological processes, we examined the enrichment of functional annotations in the identified genes.

We used DAVID [51, 52] to analyze the ontological enrichment of the genes contributing to Fig. 8 separately for each rhythm detection method. Because many of the annotation terms are obviously related (e.g., “oxidoreductase” and “oxidation reduction”), we manually grouped them. The grouped terms enriched with Benjamini-Hochberg adjusted p-values less than 0.05 for the different methods can be seen in

9. Genes that are identified as rhythmic by F24 and ANOVA are enriched in the fewest terms. They are mainly in rhythm/light/circadian categories, corroborating the selection of these genes as cycling. The ITK_CYCLE methods without asymmetry search identify sets of genes that are enriched for different terms in addition to the rhythm/light/circadian ones found by ANOVA and F24, such as glutathione and oxidation reduction annotation terms. The ITK_CYCLE methods with asymmetry search identify sets of genes enriched in the most terms of all the methods. The original ITK_CYCLE method with Bonferroni correction and empirical ITK_CYCLE method identify sets of enriched genes known to have alternative splice forms of their RNA; ITK_CYCLE with the Benjamini-Hochberg correction and empirical ITK_CYCLE identify sets of genes that are enriched for genes involved in biosynthetic pathways. Because e]TK_aby4 captures all the annotation terms of interest, we focus on its results for the remainder of this section. The individual annotation terms that are enriched in the rhythmic genes found by e]TK_aby4 can be seen with their adjusted p-values and phase distributions in Fig. 10A. Fig. 10B shows the total phase distribution of the genes, and Fig. 10C shows the total asymmetry distribution. Functionally, these genes fall into several annotation categories, each of which we discuss in turn.

The peak expressions of these genes are focused around ZT4, and these genes mainly have asymmetries close to 12 h (Figs. 10A and $18). Glutathione and drug metabolism are known to be circadian [53—55]; possible links to aging are suggested by the role of glutathi-one metabolism in clearing reactive oxygen species [56]. Other oxidation-reduction related terms peak at either ZT4 or ZT16-20 (Figs. 10A and 819). These genes have a broader distribution of asymmetries, with several with extreme values of 4 or 20 h.

10A). Various iron-related genes have been implicated as important in circadian rhythms. Recent studies, however, have only looked at the effect of iron-related genes on whole organism activity, or on particular circadian genes, such as per or tim [57, 58]. These studies have shown that individual iron-related genes affect circadian rhythms. To our knowledge, no studies to date have found as many iron-related genes displaying rhythmic behavior as we have described here.

10A and 820). They have a broad distribution of asymmetries as well, with several genes with extreme values of 4 h and 20 h, such as tws, the newly discovered cycling gene previously mentioned as having an asymmetry of 20 h and as a regulatory target of Clk. The existence of these alternatively spliced genes with extreme asymmetries explains why “alternative splicing” was only found to be enriched in the genes identified as rhythmic by methods searching for asymmetric waveforms. Alternative splicing has been found to be important in circadian rhythms in Drosophila as well as in other species. Most studies, however, have focused on specific experimental findings that discovered particular genes that modulate circadian rhythms [59, 60]. No studies exist that have found that so many genes with alternative splicing are rhythmic.

Discussion

These approaches are general and can be applied to detecting periodic behavior in a wide range of contexts, but we focus on time series representative of genome-wide expression data. Deckard et al. [28] recently reviewed a number of earlier studies of rhythm detection methods and selected four algorithms for comparison (de Lichtenberg, Lomb-Scargle, ITK_CYCLE, and persistent homology) based on their mathematical properties and applicability to genome-wide expression data. They test the methods with simulated data and experimental data for the metabolic cycle in yeast, circadian rhythms in the mouse, and the root clock in the flowering plant Arabidopsis thaliana (see [28] for references). They find that there is no all-around best method and construct a decision tree for picking an algorithm based on the expected nature of the data. For increasing noise and decreasing sampling rate, they favor ITK_CYCLE and Lomb-Scargle, a Fourier-like method. These recommendations are consistent with our own findings that ITK_CYCLE and F24 were consistently more accurate than the monotonicity tests (represented by persistent homology in [28]) and justifies our focus on improving ITK_CYCLE.

[29] reviewed six different Fourier-like methods for their ability to estimate periods in periodic time series. They focus on time series that are well-sampled (36 and 72 time points were the lowest-sampled time series they examined, 720 was among the highest), as might be obtained from tracking a luminescence reporter for a single gene. The fundamentally different nature of their data from ours highlights the fact that it is important to match the computational tool with the task of interest. Zielinski et al. [29] seek precise period determination for genes already known to cycle. By contrast, here we focus on discovering rhythmic time series that represent only a fraction of a genome-wide dataset. ITK_CYCLE can provide estimates of a periodic time series’ phase, period, and asymmetry, but it resolves these parameters only to the level of the sampling (or search) depth. It is likely that there is a tradeoff between robust separation of rhythmic and arrhythmic time series and precise estimation of the cycling parameters; presently, no algorithm achieves both these goals simultaneously.

[29] , as well as earlier studies [37, 38], note that biological oscillations are eXpected to have asymmetric patterns of eXpression. Thaben and Westermark recently published one approach to this problem [30]. Their method, RAIN, employs the Mann-Whitney U test [61] between different time points to look for a rising pattern followed by a falling pattern. They show that RAIN outperforms the original ITK_CYCLE method as well as a cosine-fitting method [62] for simulated data consisting of sinusoidal and ramp waveforms. They also analyze genomic and proteomic data for the mouse liver. Their work reinforces our finding that searching for asymmetric waveforms can produce better rhythm detection sensitivity and validates our efforts. Thaben and Westermark suggest that modifying ITK_CYCLE to allow for asymmetric waveforms would provide a useful complement to their approach. In particular, in contrast to RAIN, ITK_CYCLE can search for specific waveforms, including arbitrary shapes with multiple peaks. We meet that need here.

Our analysis of two different simulated datasets and the fly head metadata-set clearly shows the importance of accurate significance estimates. As sequencing costs continue to decrease, we eXpect sampling density to increase. This trend should favor use of empirical ITK_CYCLE over alternative means of correcting for multiple hypothesis testing because the increase in data will enable more phases, asymmetries, and periods to be examined. Our analysis shows that certain gene ontologies have many genes with highly asymmetric patterns of eXpression. It will be interesting to determine the prevalence of different waveforms in additional biological datasets and to understand how their features depend quantitatively on genotype, tissue, and environmental conditions.

Conclusions

With regard to experimental design, we find that increasing the number of replicates is more important than increasing the sampling density for achieving greater sensitivity. A key aspect of our study is that we improve the estimation of p-values in ITK_CYCLE. This enables control of the false discovery rate and testing waveforms beyond sinusoidal ones. For both simulated data and a circadian metadataset [27] the resulting empirical ITK_CYCLE with asymmetry search eXhibits the greatest sensitivity among the methods that we evaluated. The annotation terms that are enriched among the genes that we identify as cycling include rhythm/ light/ circadian, glutathione/ drug metabolism, oxidation-reduction, iron metabolism, and alternative splicing. These findings are consistent with known circadian biology but also suggest new investigations.

Supporting Information

The time series used for this example was a 24 h sine wave sampled every 2 h for 1 period (no replicates);

(A) Convergence of the mean and variance estimates, used to parameterize the Gamma distribution, as a function of the number of permutations performed, for testing the 24 h period (blue curve in A; convergence for 4 h and 48 h periods were similar, data not shown). (B) The cumulative distributions obtained by random permutation fit to the Gamma distribution, as shown by their proximity to the diagonal (black). Shown are fits for testing a 24 h period, plus a 4 h period and a 48 h period (i.e., F4 and F48). For these fits, 100 permutations were used.

The correlation between triangle and cosine waveforms are compared for time series of different lengths for three different correlation metrics: Pearson, Spearman, and Kendall. Correlations can range from —1 (completely anti-correlated) to +1 (completely correlated). S3 Fig. AUROCs for simulated data with 25% noise (standard deviation of Gaussian noise as a percent of amplitude). Layout and abbreviations are the same as in Fig. 4. (EPS)

Layout and abbreviations are the same as in Fig. 5.

(A) A pseudo-replicate (CD) for time 1‘,- (indicated by the arrow) is obtained by linearly interpolating between time points ti_1 and ti+1 (dashed line). (B) Repeating this procedure for each time point (modulo 24 h) gener-S6 Fig. Interpolating the data points to generate pseudo-replicates improves AUROCs when the number of actual replicates is low. We compare performance With (Interp) and Without (Normal) pseudo-replicates for the first simulated dataset With 50% noise. (EPS)

Simulated data With rhythmic time series Without asymmetry (A) or With evenly distributed asymmetry (B) was tested With different methods. The vertical axis shows the Matthews Correlation Coefficients (MCC) [39] for different Benjamini-Hochberg adjusted p-value cutoffs (FDR) along the X-aXis. These data are With 25% noise, but the effects of Benjamini-Hochberg correction are significantly greater at 50% noise (not shown). The method abbreviations are the same as those in Fig. 4.

Simulated data With rhythmic time series Without asymmetry (left, A and C) or With evenly distributed asymmetry (right, B and D) was tested With different asymmetries. The cumulative histograms are plotted before (A and B) and after (C and D) Benja-mini-Hochberg multiple hypothesis correction across time series. The vertical axis shows the number of genes With a p-value (P) (A and B) or false discovery rate (FDR, the Benjamini-Hochberg adjusted p-value) (C and D) below or equal to a significance threshold, shown on

The positive examples are known cycling genes per, tim, vri, Pdpl, cry, and Clk. The negative examples are known non-cycling genes cam, RpL32, eye, and dco. As plotted, large values for the positive examples and small values for the negative examples are desirable. The magenta line marks a p-value of 0.05 (—loglo 0.05 = 1.3). Since 2 X 106 permutations were used to generate the empirical ITK_CYCLE p-values, they cannot be lower than 5 x 10—7. Abbreviations are the same as in Fig. 8. 810 Fig. Comparison of the p-Value distributions of the original IT K_CYCLE method (with Bonferroni correction) with the empirical IT K_CYCLE method without (A) and with (B) asymmetry search.

(A) The number of genes with a Benjamini-Hochberg adjusted p-value (FDR) below 0.05 (blue) and 0.20 (red) are shown. (B) A comparison of the intersection (below the diagonal) and union (above the diagonal) of genes identified as rhythmic With Benjamini-Hochberg adjusted p-values less than 0.05 for the different methods. Abbreviations are the same as in 88 Fig. (EPS)

Points represent genes, colored by the asymmetry search by 4 h-estimated asymmetries. The black vertical and horizontal lines mark a FDR of 0.05 (—loglo 0.05 z 1.30). Genes to the the right of the vertical line pass the threshold cutoff for eITK_aby4, While genes above the horizontal line pass the threshold cutoff for eITK With asymmetry search of 8 and 16 h. Genes that are above the horizontal line but left of the vertical line barely pass the threshold and have asymmetries in the range of 8 to 16 h. Genes that are right of the vertical line but below the horizontal line pass the threshold much more significantly than the previously mentioned genes and have asymmetries that are 813 Fig. Using a cosine as a reference waveform instead of a triangle does not produce substantially different results in genes identified as cycling. A comparison of the intersection and union of genes identified as rhythmic With Benjamini-Hochberg adjusted p-values less than 0.05 (blue bars) or 0.20 (red bars) for empirical ITK_CYCLE Without asymmetry (eITK), and empirical ITK_CYCLE With asymmetry search of 4, 8, 12, 16, and 20 h (eITK_aby4) calculated With a reference waveform of a triangle (no prefix) or With a reference waveform of a cosine (prefix “cos”). (A) The number of genes With a Benjamini-Hochberg adjusted p-value below 0.05 (blue) and 0.20 (red) are shown. (B) A comparison of the intersection (below the diagonal) and union (above the diagonal) of genes identified as rhythmic With Benjamini-Hochberg adjusted p-values less than 0.05 for the different methods. (EPS)

Annotation terms identified as enriched by DAVID share many similarities and were therefore grouped. The number of annotation terms enriched in the genes discovered by each method are shown in grey shading and red numbers. Empirical ITK_CYCLE methods With and Without asymmetry search (“_aby4” and no suffix, respectively) and With a triangle as a reference waveform or cosine as a reference waveform (no prefix or “cos”, respectively). The annotation terms 815 Fig. Comparison of genes identified as cycling by Keegan et al. and empirical

(A) All the genes shown passed the ANOVA pre-screen, but only the green ones are identified as cycling by Keegan et al. [27]. Higher negative logarithms of p-values are more significant than lower ones: the horizontal black line indicates a Benjamini-Hochberg adjusted p-value for e]TK_aby4 of 0.05. (B) All the genes shown were identified as cycling by Keegan et al. The mean and variance of the genes identified as cycling by Keegan et al. and e]TK_aby4 (blue), are 4.34 and 0.54, respectively. The mean and variance of the genes identified as cycling by Keegan et al. and but not e]TK_aby4 (red), are 4.75 and 0.56, respectively. (C) All the genes shown were identified as cycling by Keegan et al. The asymmetry of the genes was determined by eITK_aby4.

(A) Comparison of genes identified as rhythmic by Keegan et al. [27] , Wijnen et al. [32], and eITK_by4. Stacked bars are colored to represent the asymmetry, as determined by eITK_aby4. e]TK_aby4 identifies more genes with non-12 h asymmetries than the other methods. “Unassigned” refers to genes that were excluded from the empirical ITK_CYCLE analysis. (B) For each gene, the references on FlyBase (http://flybase.org) that mention the gene were identified. The genes identified by eITK_by4, Keegan et al., and Wijnen et al. are shown in a histogram with stacked bars colored to represent the genes being cited by references with “circadian” in the title or abstract, genes cited in the original five dataset papers, or neither. While there are more genes uniquely identified by Wijnen et 611., there are more total genes identified by eITK_by4, as well as more genes that are cited in papers that have “circadian” in their title or abstract.

Peak expression (phase) of these genes is mainly in the light period. (A) Z-scored gene expression of genes from the metadataset involved in glutathione metabolism averaged across 24 h and interpolated to every 2 h. (B) Phase and asymmetry distribution of the genes from the metadataset involved in glutathione metabolism.

Peak expression (phase) of these genes is distributed over 24 h. (A) Z-scored gene expression of genes from the metadataset involved in oxidation reduction averaged across 24 h and interpolated to every 2 h. Black indicates time points Where data were not available (NA). (B) Phase and asymmetry distribution of the genes from the metadataset involved in oxidation reduction. (EPS)

Peak expression (phase) of these genes is distributed over 24 h. (A) Z-scored gene expression of genes from the metadataset involved in alternative splicing averaged across 24 h and interpolated to every 2 h. (B) Phase and asymmetry distribution of the genes from the 81 Data. This Excel spreadsheet file contains the time series for the metadataset, with official gene names, after Z—scoring has been performed. (XLSX)

This Excel spreadsheet file contains several pages referring to the output of the rhythm detection methods on the metadataset, as well as the DAVID results for those methods. The method results provided are all for the metadataset after preprocessing: e]TK_aby4, e]TK_a12, ITK_BF_aby4, ITK_BF_a12, cos_e]TK_aby4, cos_eITK_a12, ANOVA, and F24.

This Excel spreadsheet file contains several pages referring to the output of the rhythm detection methods on the metadataset. The method results provided are all for the metadataset after preprocessing: ITK_BH_aby4, ITK_BH_a12 (these two come With DAVID results), eITK_aO4-12-20, and eITK_a08-16.

Acknowledgments

We would like to thank Matthew Stephens for advice about evaluating statistical significance and multiple hypothesis correction.

Author Contributions

Topics

time series

Appears in 80 sentences as: time series (87) time series’ (1) times series (1)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. Understanding how such rhythms couple to biological processes requires statistical methods that can identify cycling time series in typical genome-Wide data.
    Page 2, “Author Summary”
  2. In this paper, we improve on a method used to identify cycling time series by better estimating the statistical significance of periodic patterns and, in turn, by searching for a Wider range of patterns than traditionally investigated.
    Page 2, “Author Summary”
  3. Despite the decreasing cost of measuring transcript levels, profiling time series genome-wide continues to present formidable challenges: tissue-specific samples are difficult to collect, and, in contrast to imaging, measuring transcript levels is destructive in nature, requiring separate samples for each time point.
    Page 2, “Introduction”
  4. As a result, gene expression time series are typically sparsely sampled (e.g., every 2—4 hours (h) in circadian studies), often without multiple measurements per time point, which we refer to here as “replicates”.
    Page 2, “Introduction”
  5. Quantitative methods are thus needed to identify rhythmic time series from minimal data with statistical confidence.
    Page 2, “Introduction”
  6. While a number of different methods have been proposed for identifying cycling time series [19—30], further analysis is needed to guide selection of the best method(s) for a given situation and to aid in design of improved computational methods and further experiments.
    Page 2, “Introduction”
  7. In contrast, F24 and ITK_CYCLE compare the time series in question to a reference waveform, which is typically sinusoidal.
    Page 3, “Overview”
  8. In this way, stable persistence tries to robustly assess overall monotonicity of a time series .
    Page 3, “Overview”
  9. The numerator is the number of pairs that vary concordantly between the two time series minus the number that vary discordantly (Fig.
    Page 4, “Overview”
  10. Every possible pair is included, not just ones between neighboring points in the time series .
    Page 4, “Overview”
  11. Like the cyclohedron test, the null distribution for ITK_CYCLE can be computed exactly from a generating function [33], although again the test statistic domain grows quickly with time series size (becoming impractically large at 100—200 time points with present computing power).
    Page 4, “Overview”

See all papers in March 2015 that mention time series.

See all papers in PLOS Comp. Biol. that mention time series.

Back to top.

ANOVA

Appears in 39 sentences as: ANOVA (44)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. We term this method empirical JTK_CYCLE with asymmetry search, and we compare its performance to JTK_CYCLE with Bonferroni and Benjamini-Hochberg multiple hypothesis testing correction, as well as to five other methods: cyclohedron test, address reduction, stable persistence, ANOVA , and F24.
    Page 1, “Abstract”
  2. We find that ANOVA , F24, and JTK_CYCLE consistently outperform the other three methods when data are limited and noisy; empirical JTK_CYCLE with asymmetry search gives the greatest sensitivity while controlling for the false discovery rate.
    Page 1, “Abstract”
  3. We test its performance for detecting way analysis of variance ( ANOVA ) [27].
    Page 2, “Introduction”
  4. In particular, ITK_CYCLE, F24, and ANOVA consistently outperform the other methods and offer distinct advantages for certain types of data.
    Page 3, “Introduction”
  5. The methods that we test are cyclohedron test [20, 21] , address reduction [22, 23], stable persistence [24, 25], F24 [31, 32], one-way analysis of variance ( ANOVA ) [27], and ITK_CYCLE [26].
    Page 3, “Overview”
  6. Cyclohedron test, address reduction, stable persistence, and ANOVA seek to identify patterns without specifying the waveform a priori.
    Page 3, “Overview”
  7. ANOVA compares the means of different time points with their variances to determine if differences are significant.
    Page 3, “Overview”
  8. Analysis of variance ( ANOVA ).
    Page 4, “Overview”
  9. One-way ANOVA is a standard statistical test of the equivalence of means in several groups.
    Page 4, “Overview”
  10. In this case, each time point is a different group, and ANOVA is equivalent to testing for any statistically significant variation across the time points.
    Page 4, “Overview”
  11. ANOVA tests for these time-dependent changes in eXpression.
    Page 4, “Overview”

See all papers in March 2015 that mention ANOVA.

See all papers in PLOS Comp. Biol. that mention ANOVA.

Back to top.

p-value

Appears in 21 sentences as: p-Value (1) p-value (28)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. However, for large time series the ITK_CYCLE null distribution is approximately normal, allowing for a convenient, fast p-value estimate.
    Page 4, “Overview”
  2. By definition, a p-value is the likelihood of obtaining a test statistic equal to or more extreme than the value that is observed if the null hypothesis is true—it increases cumulatively as one progresses through a set of rank ordered test statistics.
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  3. For a dataset generated from this null model, the p-values should be uniformly distributed from 0 to 1, exclusive: the highest Kendall’s 1‘ out of N tests should have a p-value of 1 / (N + 1), the second highest test statistic has a p-value of 2/ (N + 1), and the ith highest test statistic has a p-value of i/ (N + 1) [35].
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  4. ITK_CYCLE computes the Kendall T values for all the reference time series against the signal of interest and then performs a selection step for the lowest p-value (i.e., the highest 1‘), which we refer to here as the “initial” p-value .
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  5. Therefore, a threshold of 0.01 means that there is a 1% chance that the list of time series with a Bonferroni adjusted p-value below 0.01 contains a false positive.
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  6. This method, while rigorous, is overly conservative and overcompensates for the bias that comes from selecting the lowest p-value (blue and green lines in Fig.
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  7. For example, a Benjamini-Hochberg adjusted p-value cutoff of 0.05 means that 5% of the positives are false.
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  8. However, the p-values are additionally ordered from lowest to highest and then divided by their rank order (the lowest p-value has rank order 1, the second highest p-value has rank order 2, and so on).
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  9. The p-values are also adjusted such that there is no change in ordering: if the originally lowest p-value is adjusted so that it is higher than the originally second lowest p-value, the lowest p-value takes the value of the adjusted second p-value so that the ordering is not violated.
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  10. However, this requires recomputing null distributions via MC sampling because the construction procedure introduces correlations between data points, resulting in p-value underestimates if not corrected.
    Page 11, “Simulated data benchmarks”
  11. Choosing a Benjamini-Hochberg adjusted p-value cutoff of 0.05 (i.e., 5%), the number of genes and overlap between methods can be seen in Figs.
    Page 15, “Microarray metadataset”

See all papers in March 2015 that mention p-value.

See all papers in PLOS Comp. Biol. that mention p-value.

Back to top.

gene expression

Appears in 11 sentences as: gene eXpression (1) gene expression (10)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. Application of the methods to detecting circadian rhythms in a metadataset of microarrays that quantify time-dependent gene expression in whole heads of Drosophi/a melanogaster reveals annotations that are enriched among genes with highly asymmetric waveforms.
    Page 1, “Abstract”
  2. We apply these methods to a compilation of data on gene eXpression in fruit flies, an important model organism.
    Page 2, “Author Summary”
  3. Arguably the most well-studied periodic patterns are circadian rhythms: oscillatory changes in gene expression , metabolism, physiology, and behavior with approximately 24-hour (24 h) periods that enable organisms to anticipate and respond to daily changes in their environment, such as nutrient accessibility, temperature, and light [10—13].
    Page 2, “Introduction”
  4. The advent of high throughput methods for measuring gene expression now makes transcriptome-wide studies of this nature possible.
    Page 2, “Introduction”
  5. As a result, gene expression time series are typically sparsely sampled (e.g., every 2—4 hours (h) in circadian studies), often without multiple measurements per time point, which we refer to here as “replicates”.
    Page 2, “Introduction”
  6. These experimental limitations result in low signal-to-noise ratios that prevent straightforward identification of cycling gene expression .
    Page 2, “Introduction”
  7. The methods that we consider are general and can be applied to detecting periodic behavior in any context, but we describe them here in terms of searching for circadian rhythms in gene expression for clarity.
    Page 3, “Overview”
  8. We expect these additional waveforms to be more sensitive to asymmetric patterns of gene expression , resulting in discovery of additional rhythmic time series.
    Page 7, “E 3 A A g Time s 'r r a E A AA Time Time”
  9. (A) Z-scored gene expression of genes from the metadataset involved in glutathione metabolism averaged across 24 h and interpolated to every 2 h. (B) Phase and asymmetry distribution of the genes from the metadataset involved in glutathione metabolism.
    Page 25, “Supporting Information”
  10. (A) Z-scored gene expression of genes from the metadataset involved in oxidation reduction averaged across 24 h and interpolated to every 2 h. Black indicates time points Where data were not available (NA).
    Page 26, “Supporting Information”
  11. Peak expression (phase) of these genes is distributed over 24 h. (A) Z-scored gene expression of genes from the metadataset involved in alternative splicing averaged across 24 h and interpolated to every 2 h. (B) Phase and asymmetry distribution of the genes from the 81 Data.
    Page 26, “Supporting Information”

See all papers in March 2015 that mention gene expression.

See all papers in PLOS Comp. Biol. that mention gene expression.

Back to top.

genome-wide

Appears in 11 sentences as: genome-Wide (1) genome-wide (10)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. Robust methods for identifying patterns of expression in genome-wide data are important for generating hypotheses regarding gene function.
    Page 1, “Abstract”
  2. Understanding how such rhythms couple to biological processes requires statistical methods that can identify cycling time series in typical genome-Wide data.
    Page 2, “Author Summary”
  3. Despite the decreasing cost of measuring transcript levels, profiling time series genome-wide continues to present formidable challenges: tissue-specific samples are difficult to collect, and, in contrast to imaging, measuring transcript levels is destructive in nature, requiring separate samples for each time point.
    Page 2, “Introduction”
  4. We use it to further assess the importance of considering asymmetric waveforms, and we eXplore how multiple hypothesis correction impacts the results when the true positives represent a relatively small fraction of the simulated time series, as we eXpect to be the case in genome-wide studies.
    Page 8, “Simulated data benchmarks”
  5. Furthermore, we focus on genome-wide experiments where the experimental design is such that there is no meaningful difference between data collected over multiple periods and data collected at the same sampling rate in replicate over a single period.
    Page 11, “Simulated data benchmarks”
  6. This composition was chosen to be reflective of a genome-wide dataset.
    Page 11, “Simulated data benchmarks”
  7. Therefore, in terms of genome-wide studies, empirical ITK_CYCLE with asymmetric waveforms is the method of choice for identifying rhythmic genes.
    Page 14, “Simulated data benchmarks”
  8. These approaches are general and can be applied to detecting periodic behavior in a wide range of contexts, but we focus on time series representative of genome-wide expression data.
    Page 21, “Discussion”
  9. [28] recently reviewed a number of earlier studies of rhythm detection methods and selected four algorithms for comparison (de Lichtenberg, Lomb-Scargle, ITK_CYCLE, and persistent homology) based on their mathematical properties and applicability to genome-wide expression data.
    Page 21, “Discussion”
  10. By contrast, here we focus on discovering rhythmic time series that represent only a fraction of a genome-wide dataset.
    Page 22, “Discussion”
  11. In this paper, we compare methods for detecting rhythmic time series in genome-wide expression data.
    Page 22, “Conclusions”

See all papers in March 2015 that mention genome-wide.

See all papers in PLOS Comp. Biol. that mention genome-wide.

Back to top.

simulated data

Appears in 10 sentences as: Simulated data (3) simulated data (7)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. The simulated data allow us to examine how performance varies with sampling density, number of replicates and/ or periods, noise level, and waveform.
    Page 3, “Introduction”
  2. Simulated data benchmarks
    Page 8, “Simulated data benchmarks”
  3. Examples of simulated data .
    Page 9, “Simulated data benchmarks”
  4. shows that the ITK_CYCLE methods have higher-quality classification ability than F24 and ANOVA for these simulated data .
    Page 14, “Simulated data benchmarks”
  5. They test the methods with simulated data and experimental data for the metabolic cycle in yeast, circadian rhythms in the mouse, and the root clock in the flowering plant Arabidopsis thaliana (see [28] for references).
    Page 21, “Discussion”
  6. They show that RAIN outperforms the original ITK_CYCLE method as well as a cosine-fitting method [62] for simulated data consisting of sinusoidal and ramp waveforms.
    Page 22, “Discussion”
  7. For both simulated data and a circadian metadataset [27] the resulting empirical ITK_CYCLE with asymmetry search eXhibits the greatest sensitivity among the methods that we evaluated.
    Page 22, “Conclusions”
  8. AUROCs for simulated data with 25% noise (standard deviation of Gaussian noise as a percent of amplitude).
    Page 23, “Supporting Information”
  9. Simulated data With rhythmic time series Without asymmetry (A) or With evenly distributed asymmetry (B) was tested With different methods.
    Page 23, “Supporting Information”
  10. Simulated data With rhythmic time series Without asymmetry (left, A and C) or With evenly distributed asymmetry (right, B and D) was tested With different asymmetries.
    Page 23, “Supporting Information”

See all papers in March 2015 that mention simulated data.

See all papers in PLOS Comp. Biol. that mention simulated data.

Back to top.

Bonferroni correction

Appears in 9 sentences as: Bonferroni correction (9)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. employed the Bonferroni correction [34] in their original formulation and implementation of the method.
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  2. The previous version of ITK_CYCLE corrects for underestimating the p-values with the Bonferroni correction , which controls the family-wide error rate (FWER) by multiplying the p-values by the number of hypothesis tests being performed.
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  3. A common alternative to the Bonferroni correction is the Benjamini-Hochberg procedure [36], which seeks to control the false discovery rate (FDR).
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  4. Though this empirical calculation is more computationally expensive than the application of the Bonferroni correction or the Benjamini-Hochberg correction, we show that empirically calculating the p-values results in better rhythmic detection and biological insight.
    Page 7, “E 3 A A g Time s 'r r a E A AA Time Time”
  5. The ITK_CYCLE with Benjamini-Hochberg correction (ITK_BH) has AUROC values that are in between the AUROC values for the original ITK_CYCLE with Bonferroni correction (ITK) and empirical ITK_CYCLE.
    Page 9, “Simulated data benchmarks”
  6. As experimental sampling rates and sampling densities enable more extensive searching of phases, periods, and asymmetries, we expect the advantage for empirical ITK_CYCLE relative to the original formulation to grow because the Bonferroni correction strongly penalizes adding hypothesis tests.
    Page 15, “Microarray metadataset”
  7. The original ITK_CYCLE method with Bonferroni correction and empirical ITK_CYCLE method identify sets of enriched genes known to have alternative splice forms of their RNA; ITK_CYCLE with the Benjamini-Hochberg correction and empirical ITK_CYCLE identify sets of genes that are enriched for genes involved in biosynthetic pathways.
    Page 19, “Microarray metadataset”
  8. We were able to eXpand ITK_CYCLE to search for asymmetric waveforms without degrading sensitivity because we empirically calculate p-values, which yields much more accurate significance estimates than the Bonferroni correction employed in the original formulation of ITK_CYCLE [26].
    Page 22, “Discussion”
  9. Comparison of the p-Value distributions of the original IT K_CYCLE method (with Bonferroni correction ) with the empirical IT K_CYCLE method without (A) and with (B) asymmetry search.
    Page 24, “Supporting Information”

See all papers in March 2015 that mention Bonferroni correction.

See all papers in PLOS Comp. Biol. that mention Bonferroni correction.

Back to top.

false positive

Appears in 9 sentences as: false positive (5) false positives (4)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. The FWER is the probability that there is at least one false positive for a given threshold.
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  2. Therefore, a threshold of 0.01 means that there is a 1% chance that the list of time series with a Bonferroni adjusted p-value below 0.01 contains a false positive .
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  3. The likelihood of false positives is greatly reduced, but so is the likelihood of identifying true positives.
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  4. The FDR is the fraction of the time series that are identified as cycling that are false positive .
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  5. The receiver operating characteristic (ROC) curve plots the true positive rate (TPR) as a function of the false positive rate (FPR) as the threshold for calling a time series as a positive is varied.
    Page 8, “Simulated data benchmarks”
  6. The TPR and FPR are the fractions of the 10,000 simulated or Gaussian noise time series determined to be rhythmic at a threshold, respectively, and the threshold is varied over the entire range of false positive scores, such that the FPR ranges from 0 to 1.
    Page 8, “Simulated data benchmarks”
  7. In such cases, we find that ANOVA, F24, and ITK_CYCLE consistently better distinguish true and false positives .
    Page 11, “Simulated data benchmarks”
  8. For these curves, the proportion of false positives identified as cycling matches the FDR.
    Page 13, “Simulated data benchmarks”
  9. A score of 1 indicates that a method correctly identified all true positives and true negatives, while a score of —1 indicates that a method yielded all false positives and false negatives.
    Page 14, “Simulated data benchmarks”

See all papers in March 2015 that mention false positive.

See all papers in PLOS Comp. Biol. that mention false positive.

Back to top.

uniformly distributed

Appears in 6 sentences as: uniformly distributed (6)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. For a dataset generated from this null model, the p-values should be uniformly distributed from 0 to 1, exclusive: the highest Kendall’s 1‘ out of N tests should have a p-value of 1 / (N + 1), the second highest test statistic has a p-value of 2/ (N + 1), and the ith highest test statistic has a p-value of i/ (N + 1) [35].
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  2. In the present study, these “empirical” p-values are based on 2 x 106 random time series and are nearly uniformly distributed , as desired (Fig.
    Page 7, “E 3 A A g Time s 'r r a E A AA Time Time”
  3. The second simulated dataset contains 10% rhythmic time series of triangle waveform with uniformly distributed phases and asymmetries and 90% time series consisting solely of Gaussian noise.
    Page 8, “Simulated data benchmarks”
  4. 3A we generated 10,000 time series with uniformly distributed random phase shifts (always with a 24 h period) and added Gaussian noise to each point with a standard deviation of 25% or 50% of the total waveform amplitude, examples of which can be seen in Fig.
    Page 8, “Simulated data benchmarks”
  5. In both cases, phases (peak values) were uniformly distributed over the possible discrete values.
    Page 12, “Simulated data benchmarks”
  6. The p-values of empirical ITK_CYCLE are approximately uniformly distributed (Fig.
    Page 13, “Simulated data benchmarks”

See all papers in March 2015 that mention uniformly distributed.

See all papers in PLOS Comp. Biol. that mention uniformly distributed.

Back to top.

standard deviation

Appears in 6 sentences as: standard deviation (6)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. 3A we generated 10,000 time series with uniformly distributed random phase shifts (always with a 24 h period) and added Gaussian noise to each point with a standard deviation of 25% or 50% of the total waveform amplitude, examples of which can be seen in Fig.
    Page 8, “Simulated data benchmarks”
  2. (B) Cosine in black, with Gaussian noise with standard deviation of 25% (blue) or 50% (green) of amplitude.
    Page 9, “Simulated data benchmarks”
  3. We added Gaussian noise with a standard deviation of either 25% or 50% of the amplitude of the time series, as previously described.
    Page 12, “Simulated data benchmarks”
  4. Instead, we found that the best approach was to convert the values in each time series to Z-scores—i.e., for each gene in each dataset, we subtract its mean eXpression level and divide by its standard deviation .
    Page 14, “Microarray metadataset”
  5. To prevent the need to recalculate the null distribution for every pattern of NAs in the data for empirical ITK_CYCLE (there were 5005 unique NA permutations in the data), the NAs were replaced by random noise drawn from a Gaussian distribution with mean and standard deviation that match those of the data on the whole.
    Page 14, “Microarray metadataset”
  6. AUROCs for simulated data with 25% noise ( standard deviation of Gaussian noise as a percent of amplitude).
    Page 23, “Supporting Information”

See all papers in March 2015 that mention standard deviation.

See all papers in PLOS Comp. Biol. that mention standard deviation.

Back to top.

false discovery rate

Appears in 5 sentences as: false discovery rate (5)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. We find that ANOVA, F24, and JTK_CYCLE consistently outperform the other three methods when data are limited and noisy; empirical JTK_CYCLE with asymmetry search gives the greatest sensitivity while controlling for the false discovery rate .
    Page 1, “Abstract”
  2. A common alternative to the Bonferroni correction is the Benjamini-Hochberg procedure [36], which seeks to control the false discovery rate (FDR).
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  3. 6C and D, we see that the performance of the methods differs considerably when controlling for the false discovery rate (FDR).
    Page 13, “Simulated data benchmarks”
  4. This enables control of the false discovery rate and testing waveforms beyond sinusoidal ones.
    Page 22, “Conclusions”
  5. The vertical axis shows the number of genes With a p-value (P) (A and B) or false discovery rate (FDR, the Benjamini-Hochberg adjusted p-value) (C and D) below or equal to a significance threshold, shown on
    Page 23, “Supporting Information”

See all papers in March 2015 that mention false discovery rate.

See all papers in PLOS Comp. Biol. that mention false discovery rate.

Back to top.

discovery rate

Appears in 5 sentences as: discovery rate (5)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. We find that ANOVA, F24, and JTK_CYCLE consistently outperform the other three methods when data are limited and noisy; empirical JTK_CYCLE with asymmetry search gives the greatest sensitivity while controlling for the false discovery rate .
    Page 1, “Abstract”
  2. A common alternative to the Bonferroni correction is the Benjamini-Hochberg procedure [36], which seeks to control the false discovery rate (FDR).
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  3. 6C and D, we see that the performance of the methods differs considerably when controlling for the false discovery rate (FDR).
    Page 13, “Simulated data benchmarks”
  4. This enables control of the false discovery rate and testing waveforms beyond sinusoidal ones.
    Page 22, “Conclusions”
  5. The vertical axis shows the number of genes With a p-value (P) (A and B) or false discovery rate (FDR, the Benjamini-Hochberg adjusted p-value) (C and D) below or equal to a significance threshold, shown on
    Page 23, “Supporting Information”

See all papers in March 2015 that mention discovery rate.

See all papers in PLOS Comp. Biol. that mention discovery rate.

Back to top.

statistical significance

Appears in 4 sentences as: statistical significance (2) statistically significant (2)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. In this paper, we improve on a method used to identify cycling time series by better estimating the statistical significance of periodic patterns and, in turn, by searching for a Wider range of patterns than traditionally investigated.
    Page 2, “Author Summary”
  2. In this case, each time point is a different group, and ANOVA is equivalent to testing for any statistically significant variation across the time points.
    Page 4, “Overview”
  3. Because eXpression measurements are averages over many cells and different time points come from different samples (as the measurement is destructive), only synchronized, consistent variation across all samples can generate a statistically significant trend.
    Page 4, “Overview”
  4. We would like to thank Matthew Stephens for advice about evaluating statistical significance and multiple hypothesis correction.
    Page 26, “Acknowledgments”

See all papers in March 2015 that mention statistical significance.

See all papers in PLOS Comp. Biol. that mention statistical significance.

Back to top.

true positives

Appears in 4 sentences as: true positive (1) true positives (3)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. The likelihood of false positives is greatly reduced, but so is the likelihood of identifying true positives .
    Page 6, “E 3 A A g Time s 'r r a E A AA Time Time”
  2. We use it to further assess the importance of considering asymmetric waveforms, and we eXplore how multiple hypothesis correction impacts the results when the true positives represent a relatively small fraction of the simulated time series, as we eXpect to be the case in genome-wide studies.
    Page 8, “Simulated data benchmarks”
  3. The receiver operating characteristic (ROC) curve plots the true positive rate (TPR) as a function of the false positive rate (FPR) as the threshold for calling a time series as a positive is varied.
    Page 8, “Simulated data benchmarks”
  4. A score of 1 indicates that a method correctly identified all true positives and true negatives, while a score of —1 indicates that a method yielded all false positives and false negatives.
    Page 14, “Simulated data benchmarks”

See all papers in March 2015 that mention true positives.

See all papers in PLOS Comp. Biol. that mention true positives.

Back to top.

Microarray

Appears in 3 sentences as: Microarray (1) microarray (1) microarrays (1)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. Application of the methods to detecting circadian rhythms in a metadataset of microarrays that quantify time-dependent gene expression in whole heads of Drosophi/a melanogaster reveals annotations that are enriched among genes with highly asymmetric waveforms.
    Page 1, “Abstract”
  2. Microarray metadataset
    Page 14, “Microarray metadataset”
  3. Analysis of microarray metadataset.
    Page 15, “Microarray metadataset”

See all papers in March 2015 that mention Microarray.

See all papers in PLOS Comp. Biol. that mention Microarray.

Back to top.

null model

Appears in 3 sentences as: null model (3)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. The domain of test statistics increases very quickly with the number of data points, however, so Monte Carlo (MC) sampling, in which representations of the null model are randomly generated and evaluated, is required to estimate p-values if there are more than approximately 20 time points due to the computational cost of the generating function.
    Page 3, “Overview”
  2. For a dataset generated from this null model , the p-values should be uniformly distributed from 0 to 1, exclusive: the highest Kendall’s 1‘ out of N tests should have a p-value of 1 / (N + 1), the second highest test statistic has a p-value of 2/ (N + 1), and the ith highest test statistic has a p-value of i/ (N + 1) [35].
    Page 5, “E 3 A A g Time s 'r r a E A AA Time Time”
  3. While the empirical calculation approximates the null model well, it does not fully prevent multiple hypothesis testing from weakening the ability to identify rhythmic time series.
    Page 9, “Simulated data benchmarks”

See all papers in March 2015 that mention null model.

See all papers in PLOS Comp. Biol. that mention null model.

Back to top.

expression data

Appears in 3 sentences as: expression data (3)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. These approaches are general and can be applied to detecting periodic behavior in a wide range of contexts, but we focus on time series representative of genome-wide expression data .
    Page 21, “Discussion”
  2. [28] recently reviewed a number of earlier studies of rhythm detection methods and selected four algorithms for comparison (de Lichtenberg, Lomb-Scargle, ITK_CYCLE, and persistent homology) based on their mathematical properties and applicability to genome-wide expression data .
    Page 21, “Discussion”
  3. In this paper, we compare methods for detecting rhythmic time series in genome-wide expression data .
    Page 22, “Conclusions”

See all papers in March 2015 that mention expression data.

See all papers in PLOS Comp. Biol. that mention expression data.

Back to top.

Correlation Coefficient

Appears in 3 sentences as: Correlation Coefficient (2) Correlation Coefficients (1)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. Correlation Coefficient [39], which quantifies the quality of a binary classification.
    Page 14, “Simulated data benchmarks”
  2. Matthews Correlation Coefficient shows that IT K_CYCLE methods outperform ANOVA and F24 in the presence and absence of asymmetric time series.
    Page 23, “Supporting Information”
  3. The vertical axis shows the Matthews Correlation Coefficients (MCC) [39] for different Benjamini-Hochberg adjusted p-value cutoffs (FDR) along the X-aXis.
    Page 23, “Supporting Information”

See all papers in March 2015 that mention Correlation Coefficient.

See all papers in PLOS Comp. Biol. that mention Correlation Coefficient.

Back to top.

basis function

Appears in 3 sentences as: basis function (3)
In Improved Statistical Methods Enable Greater Sensitivity in Rhythm Detection for Genome-Wide Data
  1. The test statistic for F24 is the projection of the data onto the 24 h Fourier basis function , and the null distribution is obtained by recom-puting the test statistic over repeated random permutations of the data.
    Page 4, “Overview”
  2. The phase is determined by projecting the data onto the cosine part of the Fourier basis function and finding the optimal phase for the projection.
    Page 4, “Overview”
  3. Testing periods other than 24 h is accomplished simply by changing the period of the Fourier basis function used to compute the test statistic.
    Page 4, “Overview”

See all papers in March 2015 that mention basis function.

See all papers in PLOS Comp. Biol. that mention basis function.

Back to top.