Skip to main content
eLife logoLink to eLife
. 2017 Feb 15;6:e21887. doi: 10.7554/eLife.21887

A phylogenetic transform enhances analysis of compositional microbiota data

Justin D Silverman 1,2,3, Alex D Washburne 4,5, Sayan Mukherjee 1,6,7,8,9, Lawrence A David 1,3,10,*
Editor: Anthony Fodor11
PMCID: PMC5328592  PMID: 28198697

Abstract

Surveys of microbial communities (microbiota), typically measured as relative abundance of species, have illustrated the importance of these communities in human health and disease. Yet, statistical artifacts commonly plague the analysis of relative abundance data. Here, we introduce the PhILR transform, which incorporates microbial evolutionary models with the isometric log-ratio transform to allow off-the-shelf statistical tools to be safely applied to microbiota surveys. We demonstrate that analyses of community-level structure can be applied to PhILR transformed data with performance on benchmarks rivaling or surpassing standard tools. Additionally, by decomposing distance in the PhILR transformed space, we identified neighboring clades that may have adapted to distinct human body sites. Decomposing variance revealed that covariation of bacterial clades within human body sites increases with phylogenetic relatedness. Together, these findings illustrate how the PhILR transform combines statistical and phylogenetic models to overcome compositional data challenges and enable evolutionary insights relevant to microbial communities.

DOI: http://dx.doi.org/10.7554/eLife.21887.001

Research Organism: Human

Introduction

Microbiota research today embodies the data-rich nature of modern biology. Advances in high-throughput DNA sequencing allow for rapid and affordable surveys of thousands of bacterial taxa across hundreds of samples (Caporaso et al., 2011). The exploding availability of sequencing data has poised microbiota research to advance our understanding of fields as diverse as ecology, evolution, medicine, and agriculture (Waldor et al., 2015). Considerable effort now focuses on interrogating microbiota datasets to identify relationships between bacterial taxa, as well as between microbes and their environment.

Increasingly, it is appreciated that the relative nature of microbial abundance data in microbiota studies can lead to spurious statistical analyses (Jackson, 1997; Friedman and Alm, 2012; Aitchison, 1986; Lovell et al., 2011; Gloor et al., 2016a; Britanova et al., 2014; Li, 2015; Tsilimigras and Fodor, 2016). With next generation sequencing, the number of reads per sample can vary independently of microbial load (Lovell et al., 2011; Tsilimigras and Fodor, 2016). In order to make measurements comparable across samples, most studies therefore analyze the relative abundance of bacterial taxa. Analyses are thus not carried out on absolute abundances of community members (Figure 1A), but rather on relative data occupying a constrained geometric space and represented in a non-Cartesian coordinate system (Figure 1B). Such relative abundance datasets are often termed compositional. The use of most standard statistical tools (e.g., correlation, regression, or classification) within a compositional space leads to spurious results (Pawlowsky-Glahn et al., 2015). For example, three-quarters of the significant bacterial interactions inferred by Pearson correlation on a compositional human microbiota dataset were likely false (Friedman and Alm, 2012), and over two-thirds of differentially abundant taxa inferred by a t-test on a simulated compositional human microbiota dataset were spurious (Mandal et al., 2015). To account for compositional effects in microbial datasets, bioinformatics efforts have re-derived common statistical methods including correlation statistics (Friedman and Alm, 2012; Fang et al., 2015), hypothesis testing (La Rosa et al., 2012), and variable selection (Chen and Li, 2013; Lin et al., 2014).

Figure 1. PhILR uses an evolutionary tree to transform microbiota data into an unconstrained coordinate system.

Figure 1.

(A) Two hypothetical bacterial communities share identical absolute numbers of Lactobacillus, and Ruminococcus bacteria; they differ only in the absolute abundance of Bacteroides which is higher in community A (red circle) compared to community B (blue diamond). (B) A ternary plot depicts proportional data typically analyzed in a sequencing-based microbiota survey. Note that viewed in terms of proportions the space is constrained and the axes are not Cartesian. As a result, all three genera have changed in relative abundance between the two communities. (C) Schematic of the PhILR transform based on a phylogenetic sequential binary partition. The PhILR coordinates can be viewed as ‘balances’ between the weights (relative abundances) of the two subclades of a given internal node. In community B, the greater abundance of Bacteroides tips the balance y1* to the right. (D) The PhILR transform can be viewed as a new coordinate system (grey dashed lines) in the proportional data space. (E) The data transformed to the PhILR space. Note that in contrast to the raw proportional data (B), the PhILR space only shows a change in the variable associated with Bacteroides.

DOI: http://dx.doi.org/10.7554/eLife.21887.002

An alternative approach is to transform compositional microbiota data to a space where existing statistical methods may be applied without introducing spurious conclusions. This approach is attractive because of its efficiency: the vast toolbox of existing statistical models can be applied without re-derivation. Normalization methods, for example, have been proposed to modify count data by assuming reads follow certain statistical distributions (e.g., negative binomial) (Paulson et al., 2013; Anders and Huber, 2010). Alternatively, the field of Compositional Data Analysis (CoDA) has focused on formalizing methods for transforming compositional data into a simpler geometry without having to assume data adhere to a distribution model (Bacon-Shone, 2011). Previous microbiota analyses have already leveraged CoDA theory and used the centered log-ratio transform to reconstruct microbial association networks and interactions (Kurtz et al., 2015; Lee et al., 2014) and to analyze differential abundances (Fernandes et al., 2014; Gloor et al., 2016b). However, the centered log-ratio transform has a crucial limitation: it yields a coordinate system featuring a singular covariance matrix and is thus unsuitable for many common statistical models (Pawlowsky-Glahn et al., 2015). This drawback can be sidestepped using another CoDA transform, known as the Isometric Log-Ratio (ILR) transformation (Egozcue et al., 2003). The ILR transform can be built from a sequential binary partition of the original variable space (Figure 1C), creating a new coordinate system with an orthonormal basis (Figure 1D and E) (Egozcue and Pawlowsky-Glahn, 2005). However, a known obstacle to using the ILR transform is the choice of partition such that the resulting coordinates are meaningful (Pawlowsky-Glahn et al., 2015). To date, microbiota studies have chosen ILR coordinates using ad hoc sequential binary partitions of bacterial groups that are not easily interpreted (Finucane et al., 2014; Lê Cao et al., 2016). Alternatively, external covariates have been used to pick groups of bacterial taxa to contrast (Morton et al., 2017).

Here, we introduce the bacterial phylogenetic tree as a natural and informative sequential binary partition when applying the ILR transform to microbiota datasets (Figure 1C). Using phylogenies to construct the ILR transform results in an ILR coordinate system capturing evolutionary relationships between neighboring bacterial groups (clades). Analyses of neighboring clades offer the opportunity for biological insight: clade analyses have linked genetic differentiation to ecological adaptation (Hunt et al., 2008), and the relative levels of sister bacterial genera differentiate human cohorts by diet, geography, and culture (De Filippo et al., 2010; Wu et al., 2011; Yatsunenko et al., 2012). Datasets analyzed by a phylogenetically aware ILR transform could therefore reveal ecological and evolutionary factors shaping host-associated microbial communities.

We term our approach the Phylogenetic ILR (PhILR) transform. Using published environmental and human-associated 16S rRNA datasets as benchmarks, we found that simple Euclidean distances calculated on PhILR transformed data provided a compositionally robust alternative to distance/dissimilarity measures like Bray-Curtis, Jaccard, and Unifrac. In addition, we observed that the accuracy of supervised classification methods on our benchmark datasets was matched or improved with PhILR transformed data relative to applying the same models on untransformed (raw) or log transformed relative abundance data. Decomposing distances between samples along PhILR coordinates identified bacterial clades that may have differentiated to adapt to distinct body sites. Similar decomposition of variance along PhILR coordinates showed that, in all human body sites studied, the degree to which neighboring bacterial clades covary tends to increase with the phylogenetic relatedness between clades. Together, these findings demonstrate that the PhILR transform can be used to enhance existing microbiota analysis pipelines, as well as enable novel phylogenetic analyses of microbial ecosystems.

Results

Constructing the PhILR transform

The PhILR transform has two goals. The first goal is to transform input microbiota data into an unconstrained space with an orthogonal basis while preserving all information contained in the original composition. The second goal is to conduct this transform using phylogenetic information. To achieve these dual goals on a given set of N samples consisting of relative measurements of D taxa (Figure 1B), we transform data into a new space of N samples and (D1) coordinates termed ‘balances’ (Figure 1C–E) (Egozcue et al., 2003; Egozcue and Pawlowsky-Glahn, 2005). Each balance yi* is associated with a single internal node i of a phylogenetic tree with the D taxa as leaves (the asterisk denotes a quantity represented in PhILR space). The balance represents the log-ratio of the geometric mean relative abundance of the two clades of taxa that descend from i (Materials and methods). Although individual balances may share overlapping sets of leaves and thus exhibit dependent behavior, the ILR transform rescales and combines leaves to form a coordinate system whose basis vectors are orthonormal and the corresponding coordinates are Cartesian (Egozcue et al., 2003; Egozcue and Pawlowsky-Glahn, 2005). The orthogonality of basis vectors allows conventional statistical tools to be used without compositional artifacts. The unit-length of basis vectors makes balances across the tree statistically comparable even when they have differing numbers of descendant tips or exist at different depths in the tree (Pawlowsky-Glahn et al., 2015). In addition, the unit-length ensures that the variance of PhILR balances has a consistent scale, unlike the variance of log-ratios originally proposed by Aitchison (Aitchison, 1986) as a measure of association, in which it can be unclear what constitutes a large or small variance (Friedman and Alm, 2012).

While the above description represents the core of the PhILR transform, we have also equipped the PhILR transform with two sets of weights that can: (1) address the multitude of zero and near-zero counts present in microbiota data; and, (2) incorporate phylogenetic branch lengths into the transformed space. Because zero counts cause problems when computing logs or performing division, zeros are often replaced in microbiota analyses with small non-zero counts. However, to avoid excess zero replacement that may itself introduce bias, stringent hard filtering thresholds are often employed (e.g. removing all taxa that are not seen with at least a minimum number of counts in a subset of samples). Still, hard filtering thresholds may remove a substantial fraction of observed taxa and do not account for the low precision (or high variability) of near-zero counts (Gloor et al., 2016a; Good, 1956; McMurdie and Holmes, 2014). We therefore developed a ‘taxon weighting’ scheme that acts as a type of soft-thresholding, supplementing zero replacement methods with a generalized form of the ILR transform that allows weights to be attached to individual taxa (Egozcue and Pawlowsky-Glahn, 2016). Weights are chosen with a heuristic designed to down weight the influence of taxa with many zero or near-zero counts (Materials and methods).

Our second weighting scheme is called branch length weighting. Certain analyses may benefit from incorporating information on evolutionary distances between taxa (Lozupone and Knight, 2005; Fukuyama et al., 2012; Purdom, 2011). For example, because related bacteria may be more likely to share similar traits (Martiny et al., 2015), it may be desirable to consider communities differing only in the abundance of closely-related microbes to be more similar than communities differing only in the abundance of distantly-related microbes. Because of the one-to-one correspondence between PhILR balances and internal nodes on the phylogenetic tree, evolutionary information can be incorporated into the PhILR transform by scaling balances using the phylogenetic distance between their direct descendants (Materials and methods). We note that we employ both branch length weighting and taxon weighting throughout our following analyses except where noted; still, these weights should be considered optional additions to the core PhILR transform.

Benchmarking community-level analyses in the PhILR coordinate system

To illustrate how the PhILR transform can be used to perform standard community-level analyses of microbiota datasets, we first examined measures of community dissimilarity. Microbiota analyses commonly compute the dissimilarity or distance between pairs of samples and use these computed pairwise distances as input to a variety of statistical tools. We investigated how Euclidean distances calculated on PhILR transformed data compared to common ecological measures of microbiota distance or dissimilarity (UniFrac, Bray-Curtis, and Jaccard) as well as Euclidean distance applied to raw relative abundance data in standard distance-based analysis. We chose three different microbiota surveys as reference datasets: Costello Skin Sites (CSS), a dataset of 357 samples from 12 human skin sites (Costello et al., 2009; Knights et al., 2011); Human Microbiome Project (HMP), a dataset of 4743 samples from 18 human body sites (e.g., skin, vaginal, oral, and stool) (Human Microbiome Project Consortium, 2012); and, Global Patterns (GP), a dataset of 26 samples from nine human or environmental sites (Caporaso et al., 2011) (Supplementary file 1 and Figure 2—figure supplement 1).

Distance-based analyses using Euclidean distances computed on PhILR transformed data exhibited performance rivaling common ecological distance or dissimilarity measures. Principal coordinate analyses (PCoA) qualitatively demonstrated separation of body sites using both Euclidean distances on PhILR transformed data (Figure 2A) and with several standard distance measures calculated on raw relative abundance data (Figure 2—figure supplement 2). To quantitatively compare distance measures, we tested how well habitat information explained variability among distance matrices as measured by the R2 statistic from PERMANOVA (Chen et al., 2012). By this metric, the Euclidean distance in the PhILR coordinate system significantly outperformed the five competing distance metrics in all but one case (in comparison to Weighted UniFrac when applied to the HMP dataset; Figure 2B).

Figure 2. Performance of standard statistical models on PhILR transformed microbiota data.

Benchmarks were performed using three datasets: Costello Skin Sites (CSS), Global Patterns (GP), Human Microbiome Project (HMP) (a summary of these datasets after preprocessing is shown in Supplementary file 1 and Figure 2—figure supplement 1). (A) Sample distance visualized using principal coordinate analysis (PCoA) of Euclidean distances computed in PhILR coordinate system. A comparison to PCoAs calculated with other distance measures is shown in Figure 2—figure supplement 2. (B) Sample distance (or dissimilarity) was computed by a range of statistics. PERMANOVA R2 values, which represent how well sample identity explained the variability in sample pairwise distances, were used as a performance metric. Distances in the PhILR transformed space were calculated using Euclidean distance. Distances between samples on raw relative abundance data were computed using Weighted and Unweighted UniFrac (WUnifrac and Unifrac, respectively), Bray-Curtis, Binary Jaccard, and Euclidean distance. Error bars represent standard error measurements from 100 bootstrap replicates and (*) denotes a p-value of ≤0.01 after FDR correction of pairwise tests against PhILR. (C) Accuracy of supervised classification methods tested on benchmark datasets. Error bars represent standard error measurements from 10 test/train splits and (*) denotes a p-value of ≤0.01 after FDR correction of all pairwise tests.

DOI: http://dx.doi.org/10.7554/eLife.21887.003

Figure 2—source data 1. Source data for Figure 2b and c as well as FDR corrected p-values from tests.
DOI: 10.7554/eLife.21887.004

Figure 2.

Figure 2—figure supplement 1. Taxa weighting scheme tends to assign smaller weights to taxa with more zero and near zero counts.

Figure 2—figure supplement 1.

The weight of a given taxon is calculated as the geometric mean of its counts across all samples times the Euclidean norm of its relative abundance across all samples in the dataset (Materials and methods). Data are plotted on a log scale with a pseudocount of 1 added to ease visualization.
Figure 2—figure supplement 2. Principal coordinate analyses using different measures of community distance or dissimilarity.

Figure 2—figure supplement 2.

Next, we tested the performance of predictive statistical models in the PhILR coordinate system. We examined four standard supervised classification techniques: logistic regression (LR), support vector machines (SVM), k-nearest neighbors (kNN), and random forests (RF) (Knights et al., 2011). We applied these methods to the same three reference datasets used in our comparison of distance metrics. As a baseline, the machine learning methods were applied to raw relative abundance datasets and raw relative abundance data that had been log-transformed.

The PhILR transform significantly improved supervised classification accuracy in 7 of the 12 benchmark tasks compared to raw relative abundances (Figure 2C). Accuracy improved by more than 90% in two benchmarks (SVM on HMP and GP), relative to results on the raw data. Log transformation of the data also improved classifier accuracy significantly on 6 of the 12 benchmarks but also significantly underperformed on one benchmark compared to raw relative abundances. In addition, the PhILR transform significantly improved classification accuracy in 5 of the 12 benchmarks relative to the log transform. Overall, the PhILR transform often outperformed the raw and log transformed relative abundances with respect to classification accuracy and was never significantly worse.

Identifying neighboring clades that differ by body site preference

While our benchmarking experiments demonstrated how PhILR transformed data performed in community-level analyses, we also wanted to explore potential biological insights afforded by the PhILR coordinate system. We therefore investigated how distances decomposed along PhILR balances using a sparse logistic regression model to examine which balances distinguished human body site microbiota in the HMP dataset. Such balances could be used to identify neighboring bacterial clades whose relative abundances capture community-level differences between body site microbiota. Microbial genetic differentiation may be associated with specialization to new resources or lifestyle preferences (Hunt et al., 2008), meaning that distinguishing balances near the tips of the bacterial tree may correspond to clades adapting to human body site environments.

We identified dozens of highly discriminatory balances, which were spread across the bacterial phylogeny (Figure 3A and Figure 3—figure supplement 1). Some discriminatory balances were found deep in the tree. Abundances of the Firmicutes, Bacteroidetes, and Proteobacteria relative to the Actinobacteria, Fusobacteria, and members of other phyla, separated skin body sites from oral and stool sites (Figure 3B). Levels of the genus Bacteroides relative to the genus Prevotella differentiated stool microbiota from other communities on the body (Figure 3C). Notably, values of select balances below the genus level also varied by body site. Relative levels of sister Corynebacterium species separated human skin sites from gingival sites (Figure 3D). Species-level balances even differentiated sites in nearby habitats; levels of sister Streptococcus species or sister Actinomyces species vary depending on specific oral sites (Figure 3E and F). These results show that the decomposition of distances between groups of samples along PhILR balances can be used to highlight ancestral balances that distinguish body site microbiota, as well as to identify more recent balances that may separate species that have adapted to inhabit different body sites.

Figure 3. Balances distinguishing human microbiota by body site.

Sparse logistic regression was used to identify balances that best separated the different sampling sites (full list of balances provided in Figure 3—figure supplement 1). (A) Each balance is represented on the tree as a broken grey bar. The left portion of the bar identifies the clade in the denominator of the log-ratio, and the right portion identifies the clade in the numerator of the log-ratio. The branch leading from the Firmicutes to the Bacteroidetes has been rescaled to facilitate visualization. (B–F) The distribution of balance values across body sites. Vertical lines indicate median values, boxes represent interquartile ranges (IQR) and whiskers extend to 1.5 IQR on either side of the median. Balances between: (B) the phyla Actinobacteria and Fusobacteria versus the phyla Bacteroidetes, Firmicutes, and Proteobacteria distinguish stool and oral sites from skin sites; (C) Prevotella spp. and Bacteroides spp. distinguish stool from oral sites; (D) Corynebacterium spp. distinguish skin and oral sites; (E) Streptococcus spp. distinguish oral sites; and, (F) Actinomyces spp. distinguish oral plaques from other oral sites. (†) Includes Bacteroidetes, Firmicutes, Alpha-, Beta-, and Gamma-proteobacteria. (‡) Includes Actinobacteria, Fusobacteria, Epsilon-proteobacteria, Spirochaetes, and Verrucomicrobia.

DOI: http://dx.doi.org/10.7554/eLife.21887.007

Figure 3.

Figure 3—figure supplement 1. Balances found to distinguish human body sites by sparse logistic regression.

Figure 3—figure supplement 1.

Taxa are listed in balance numerator/denominator format.

Balance variance and microbiota assembly

As a natural extension of our analysis of how distance decomposes along PhILR balances, we next investigated how balance variance decomposed in the PhILR coordinate system. Balance variance is a measure of association between neighboring bacterial clades. When the variance of a balance between two clades approaches zero, the mean abundance of taxa in each of the two clades will be linearly related and thus covary across microbial habitats (Lovell et al., 2015). By contrast, when a balance exhibits high variance, related bacterial clades exhibit unlinked or exclusionary patterns across samples. Unlike standard measures of association (e.g., Pearson correlation) balance variance is robust to compositional effects (Pawlowsky-Glahn et al., 2015).

Our preliminary investigation demonstrated a striking pattern in which balance variance decreased for balances closer to the tips of the phylogeny and increased for balances nearer to the root. To determine if this observed pattern was not the result of technical artifact, we took the following three steps. First, we omitted branch length weights from the transform as we anticipated that branch lengths may vary non-randomly as a function of depth in the phylogeny. Second, we anticipated that balances near the tips of the phylogeny would be associated with fewer read counts and thus would be more biased by our chosen heuristics for taxon weighting and zero replacement. We therefore omitted taxon weighting, employed more stringent filtering thresholds, and conditioned our calculation of balance variance on non-zero counts rather than using zero-replacement (Materials and methods). Third, we combined regression and a permutation scheme to test the null hypothesis that the degree to which neighboring clades covary is independent of the phylogenetic distance between them (Materials and methods). By permuting tip labels on the tree, our test generates a restricted subset of random sequential binary partitions that still maintains the count variability (and potential biases due to our zero handling methods) of the observed data.

With our modified PhILR analysis in place, we observed significantly decreasing balance variances near the tips of the phylogenetic tree for all body sites in the HMP dataset (p<0.01, permutation test with FDR correction; Figure 4A–F and Figure 4—figure supplements 12). Low variance balances predominated near the leaves of the tree. Examples of such balances involved B. fragilis species in stool (Figure 4H), Rothia mucilaginosa species in the buccal mucosa (Figure 4J), and Lactobacillus species in the mid-vagina (Figure 4L). By contrast, higher variance balances tended to be more basal on the tree. Three examples of high variance balances corresponded with clades at the order (Figure 4G), family (Figure 4I), and genus (Figure 4K) levels. We also observed that the relationship between balance variance and phylogenetic depth varied at different taxonomic scales. LOESS regression revealed that trends between variance and phylogenetic depth were stronger above the species level than below it (Materials and methods; Figure 4D–F and Figure 4—figure supplement 2). Overall, the observed pattern of decreasing balance variance near the tips of the phylogenetic tree suggested that closely related bacteria tend to covary in human body sites.

Figure 4. Neighboring clades covary less with increasing phylogenetic depth.

The variance of balance values captures the degree to which neighboring clades covary, with smaller balance variances representing sister clades that covary more strongly (Figure 4—figure supplement 1). (A–C) Balance variances were computed among samples from stool (A), buccal mucosa (B), and the mid-vagina (C). Red branches indicate small balance variance and blue branches indicate high balance variance. Balances 1–6 are individually tracked in panels (D–L). (D–F) Balance variances within each body site increased linearly with increasing phylogenetic depth on a log-scale (blue line; p<0.01, permutation test with FDR correction; Methods). Significant trends are seen across all other body sites (Figure 4—figure supplements 2 and 3). Non-parametric LOESS regression (green line and corresponding 95% confidence interval) reveals an inflection point in the relation between phylogenetic depth and balance variance. This inflection point appears below the estimated species level (‘s’ dotted line; the median depth beyond which balances no longer involve leaves sharing the same species assignment; Materials and methods). (G–L) Examples of balances with high and low variance from panels (A–F). Low balance variances (H, J, L) reflect a linear relationship between the geometric means of sister clades abundances. High balance variances (G, I, K) reflect either unlinked or exclusionary dynamics between the geometric means of sister clades abundances.

DOI: http://dx.doi.org/10.7554/eLife.21887.009

Figure 4—source code 1. Source code for Figure 4 and associated supplements.
DOI: 10.7554/eLife.21887.010
Figure 4—source data 1. FDR corrected p-values from permutation tests.
DOI: 10.7554/eLife.21887.011

Figure 4.

Figure 4—figure supplement 1. Balances with high and low variance.

Figure 4—figure supplement 1.

Shown are the 10 highest and the 10 lowest variance balances for the Stool body site in the HMP dataset. Panels 1–10: Low balance variances reflect a linear relationship (through the origin) between levels of the two clades that descend from a given balance. Panels 11–20: High balances variances lack this linear relationship.
Figure 4—figure supplement 2. Neighboring clades covary less with increasing phylogenetic depth.

Figure 4—figure supplement 2.

Shown here are all HMP body sites. Dotted lines indicate median species boundaries (see Materials and methods).
Figure 4—figure supplement 3. The null distribution for β.

Figure 4—figure supplement 3.

β is the slope in the linear regression between balance variance and phylogenetic depth in log space (regressions shown in Figure 4—figure supplement 2). To form null distributions for β, tip labels on the phylogeny were shuffled (n = 20,000) and balance values re-calculated. Distributions for β are symmetric about 0 for each body site, suggesting.

Discussion

There exists a symbiosis between our understanding of bacterial evolution and the ecology of host-associated microbial communities (Matsen, 2015). Microbiota studies have shown that mammals and bacteria cospeciated over millions of years (Moeller et al., 2016; Ley et al., 2008), and human gut microbes have revealed the forces driving horizontal gene transfer between bacteria (Smillie et al., 2011). Evolutionary tools and theory have been used to explain how cooperation benefits members of gut microbial communities (Rakoff-Nahoum et al., 2016), and raise concerns that rising rates of chronic disease are linked to microbiota disruption (Blaser and Falkow, 2009). Here, we aimed to continue building links between microbiota evolution and ecology by designing a data transform that uses phylogenetic models to overcome the challenges associated with compositional data while enabling novel evolutionary analyses.

We found that the resulting PhILR coordinate system, at least with respect to the performance metrics chosen, led to significantly improved performance for a variety of community-level analyses now used in microbiota analysis. While these results add credence to our proposed approach, we underscore that we do not find it essential that PhILR demonstrates superior benchmark performance to motivate its use in microbiota analysis. We believe that the need for compositionally robust tools has already been well established (Jackson, 1997; Friedman and Alm, 2012; Aitchison, 1986; Lovell et al., 2011; Gloor et al., 2016a; Li, 2015; Tsilimigras and Fodor, 2016) and intended these benchmarks to showcase the flexibility and utility of working with PhILR transformed data. We also note that for some analyses, a phylogeny-based ILR transform will not outperform an ILR transform built from another sequential binary partition. In fact, in the absence of branch length weights, any random ILR partition would yield equivalent results on our benchmark tasks. Instead, what distinguishes the PhILR transform from other ILR transforms is the interpretability of the transformed coordinates. Balances in PhILR space correspond to speciation events, which can be a source for biological insight.

For example, performing regression on PhILR transformed data enabled us to decompose the distance between bacterial communities onto individual locations on the phylogeny, highlighting balances near the tips of the tree that distinguished human body sites. These balances may reflect functional specialization, as ecological partitioning among recently differentiated bacterial clades could be caused by genetic adaptation to new environments or lifestyles (Hunt et al., 2008). Indeed, among oral body sites, we observed consistent site specificity of neighboring bacterial clades within the genera Actinomyces (Figure 3F) and Streptococcus (Figure 3E). Species within the Actinomyces genera have been previously observed to partition by oral sites (Aas et al., 2005; Mager et al., 2003). Even more heterogeneity has been observed within the Streptococcus genus, where species have been identified that distinguish teeth, plaque, mucosal, tongue, saliva, and other oral sites (Aas et al., 2005; Mager et al., 2003). This partitioning likely reflects variation in the anatomy and resource availability across regions of the mouth (Aas et al., 2005), as well as the kinds of surfaces bacterial strains can adhere to (Mager et al., 2003).

We also observed evidence for potential within-genus adaptation to body sites that has not been previously reported. Within the genus Corynebacterium, we found ratios of taxa varied among oral plaques and select skin sites (Figure 3D). Although the genus is now appreciated as favoring moist skin environments, the roles played by individual Corynebacteria within skin microbiota remain incompletely understood (Grice and Segre, 2011). Precisely linking individual Corynebacterium species or strains to body sites is beyond the scope of this study due to the limited taxonomic resolution of 16S rRNA datasets (Janda and Abbott, 2007; Větrovský and Baldrian, 2013). Nevertheless, we believe the PhILR coordinate system may be used in the future to identify groups of related bacterial taxa that have undergone recent functional adaptation.

Another example of how the PhILR transform can provide biological insights arose in our analysis of how human microbiota variance decomposes along individual balances. We observed that balances between more phylogenetically related clades were significantly more likely to covary than expected by chance. This pattern could reflect evolutionary and ecological forces structuring microbial communities in the human body. Related bacterial taxa have been hypothesized to have similar lifestyle characteristics (Martiny et al., 2015; Zaneveld et al., 2010), and may thus covary in human body sites that favor their shared traits (Levy and Borenstein, 2013; Faust et al., 2012). An alternative explanation for the balance variation patterns we observed is that sequencing errors and read clustering artifacts are likely to produce OTUs (Operational Taxonomic Units) with similar reference sequences and distributions across samples. While we cannot conclusively rule out this alternative hypothesis, we note that it would not explain why signal for taxa co-variation is weakest for balances at higher taxonomic levels and appears to plateau for balances near or below the species level. A biological explanation for the plateauing signal would be that lifestyle characteristics enabling bacteria to persist in human body sites are conserved among strains roughly corresponding to the same species. Follow-up studies are needed to more conclusively understand how balance variance patterns across the phylogenies can be interpreted from an evolutionary standpoint.

Though the methods presented here provide a coherent geometric framework for performing microbiota analysis free from compositional artifacts, future refinements are possible. Specifically, we highlight issues relating to our choice of weights, the handling of zero values, and information loss during count normalization. Both the taxa weights and the branch lengths weights we introduce here may be viewed as preliminary heuristics; future work will likely yield additional weighting schemes, as well as knowledge for when a given weighting scheme should be matched to an analysis task. In the case of supervised machine learning, weighting selection could be optimized as part of the training process. Additionally, if it is important that the transformed data has meaningful numerical coordinates, such that one desires to interpret the exact numerical value of a given balance in a sample, we suggest that neither branch length weights nor taxa weights be used as these weights can complicate this type of interpretation. Concerning our handling of zero values, this model design choice confronts an outstanding challenge for microbiota and compositional data analysis (Tsilimigras and Fodor, 2016; Martın-Fernandez et al., 2011). Part of this challenge’s difficulty is whether a zero value represents a value below the detection limit (rounded zero) or a truly absent taxon (essential zero). Here, we employ zero replacement, which implies an assumption that all zero values represent rounded zeros. New mixture models that explicitly allow for both essential and rounded zeros (Bear and Billheimer, 2016), as well as more advanced methods of zero replacement (Martın-Fernandez et al., 2011; Martin-Fernandez et al., 2015), may enable us to handle zeros in a more sophisticated manner. Lastly, in regards to informational loss caused by normalization, it is known that the number of counts measured for a given taxon influences the precision with which we may estimate its relative abundance in a sample (Gloor et al., 2016a; Good, 1956; McMurdie and Holmes, 2014). While our taxa weights are intended to address this idea, a fully probabilistic model of counts would likely provide more accurate error bounds for inference. We believe it would be possible to build such a model in a Bayesian framework by viewing the observed counts as multinomial draws from a point in the PhILR transformed space, as has been done for other log-ratio based spaces (Billheimer et al., 2001).

Beyond refining the PhILR transform itself, future effort may also be directed towards interpreting the transform's results at the single taxon level. Microbiota studies frequently focus on individual taxa for tasks such as identifying specific bacteria that are causal or biomarkers of disease. Log-ratio approaches can provide a compositionally robust approach to identifying biomarkers based on changes in the relative abundance of individual taxa. Due to the one-to-one correspondence between CLR coordinates and individual taxa, the CLR transform has been used previously to build compositionally robust models in terms of individual taxa (Mandal et al., 2015; Kurtz et al., 2015; Fernandes et al., 2014). However, CLR transformed data suffer from the drawback of a singular covariance matrix, which can make the development of new models based on the CLR transform difficult (Pawlowsky-Glahn et al., 2015). ILR transformed data do not suffer this drawback (Pawlowsky-Glahn et al., 2015) and moreover, can be analyzed at the single taxon level. To do so, the inverse ILR transform can be applied to model results generated in an ILR coordinate system, yielding analyses in terms of changes in the relative abundance of individual taxa (Pawlowsky-Glahn et al., 2015). The use of the inverse ILR transform in this manner is well established (Pawlowsky-Glahn et al., 2015; van den Boogaart and Tolosana-Delgado, 2013Pawlowsky-Glahn and Buccianti, 2011) and the inverse transform is provided in the Methods (Egozcue and Pawlowsky-Glahn, 2016).

Despite these avenues for improvement, modification, or extension we believe the PhILR transform already enables existing statistical methods to be applied to metagenomic datasets, free from compositional artifacts and framed according to an evolutionary perspective. We foresee the PhILR transform being used as a default transformation prior to many microbiota analyses, particularly if a phylogenetic perspective is desired. For example, the PhILR transform could be used in lieu of the conventional log transform, which is often the default choice in microbiota analysis but is not robust to compositional effects. Substituting PhILR into existing bioinformatics pipelines should often be seamless and we emphasize that all statistical tools applied to PhILR transformed data in this study were used 'off-the-shelf' and without modification. Importantly, such a substitution contrasts with the alternative approach for accounting for compositional microbiota data, which is to modify existing statistical techniques. Such modification is often challenging because many statistics were derived assuming an unconstrained space with an orthonormal basis, not a constrained and over-determined compositional space. Therefore, while select techniques have already been adapted (e.g. distance measures that incorporate phylogenetic information (Lozupone and Knight, 2005) and feature selection methods that handle compositional input (Chen and Li, 2013; Lin et al., 2014)), it is likely that certain statistical goals, such as non-linear community forecasting or control system modeling, may prove too complex for adapting to the compositional nature of microbiota datasets. Finally, beyond microbiota surveys, we also recognize that compositional metagenomics datasets are generated when studying the ecology of viral communities (Culley et al., 2006) or clonal population structure in cancer (Britanova et al., 2014; Yuan et al., 2015; Roth et al., 2014). We expect the PhILR transform to aid other arenas of biological research where variables are measured by relative abundance and related by an evolutionary tree.

Materials and methods

The ILR transform

A typical microbiome sample consists of measured counts cj for taxa j{1, , D}. A standard operation is to take count data and transform it to relative abundances. This operation is referred to as closure in compositional data analysis (Aitchison, 1986) and is given by

x= C[(c1,,cD)]=(c1jcj,,cDjcj)

where x represents a vector of relative abundances for the D taxa in the sample. We can represent a binary phylogenetic tree of the D taxa using a sign matrix Θ as introduced by Pawlowsky-Glahn and Egozcue (Pawlowsky-Glahn and Egozcue, 2011) and shown in Figure 5. Each row of the sign matrix indexes an internal node i of the tree and each column indexes a tip of the tree. A given element in the sign matrix is ±1 depending on which of the two clades descending from i that tip is a part of and 0 if that tip is not a descendent of i. The assignment of +1 versus 1 determines which clade is represented in the numerator versus the denominator of the corresponding log-ratio (as described below). Exchanging this assignment for a given balance switches which clade is represented in the numerator versus the denominator of the log-ratio. Following Egozcue and Pawlowsky-Glahn (Egozcue and Pawlowsky-Glahn, 2016), we represent the ILR coordinate (balance) associated with node i in terms of the shifted composition y=x/p=(x1/p1,,xD/pD) as

yi=ni+nini++niloggp(yi+)gp(yi) . (1)

Figure 5. Sign matrix representation of a phylogenetic tree.

Figure 5.

A binary tree (Left) can be represented by a sign matrix (Right) denoted Θ.

DOI: http://dx.doi.org/10.7554/eLife.21887.015

Here, gp(yi+) and gp(yi) represents the weighted geometric mean of the components of y that represent tips in the +1 or 1 clade descendant from node i respectively. This weighted geometric mean is given by

gp(yi±)=exp((θij=±1)pjlog yj(θij=±1)pj) (2)

where pj is the weight assigned to taxa j. The term ni+ni/ni++ni in equation 1 is the scaling term that ensures that the ILR basis element has unit length and the terms ni± are given by

ni±=θij=±1pj. (3)

Note that when p=(1,,1), y=x, Equation 1 represent the ILR transform as originally published (Egozcue et al., 2003), Equation 2 represents the standard formula for geometric mean of a vector y, and equation 3 represents the number of tips that descend from the +1 or 1 clade descendant from node i. However, when p(1,,1), these three equations represent a more generalized form of the ILR transform that allows weights to be assigned to taxa in the transformed space (Egozcue and Pawlowsky-Glahn, 2016).

Following Egozcue and Pawlowsky-Glahn (Egozcue and Pawlowsky-Glahn, 2016), we also note that the form of the generalized ILR (which we will denote ilrp) transform can be rewritten in terms of a generalized CLR transform (which we will denote clrp). This formulation in terms of the generalized CLR transform can be more efficient to compute and allows the inverse of the transform to be easily described. We can define the generalized CLR transform as

clrp(y)=(logy1gp(y), , logyDgp(y)).

The generalized ILR transform can then be written as

y=ilrp(y)=clrp(y) diag(p)ΨT

with the ijth element of the matrix Ψ given by

ψij={+1ni+ni+nini++niif θij=+11nini+nini++niif θij=10if θij=0 .

With these components defined the inverse of generalized ILR transform can be written as 𝒞[y]=ilrp1(y)=𝒞[exp(yΨ)] and x= 𝒞[ilrp1(y)p].

Soft thresholding through weighting taxa

We make use of this generalized ILR transform to down weight the influence of taxa with many zero and near-zero counts since these are less reliable and therefore more variable (Good, 1956). Our choice of taxa weights is a heuristic that combines two terms multiplicatively: a measure of the central tendency of counts, such as the mean or median of the raw counts for a taxon across the N samples in a dataset; and, the norm of the vector of relative abundances of a taxon across the N samples in a dataset. We add this vector norm term to weight taxa by their site-specificity. Preliminary studies showed that the geometric mean of the counts (with a pseudocount added to avoid skew from zero values) outperformed both the arithmetic mean and median as a measure of central tendency for the counts (data not shown). Additionally, while both the Euclidean norm and the Aitchison norm improved preliminary benchmark performance compared to using the geometric mean alone, in one case (classification using support vector machine on the global patterns dataset), the Euclidean norm greatly outperformed the Aitchison norm (Supplementary file 1). Therefore, our chosen taxa weighting scheme uses the geometric mean times the Euclidean norm:

pj=(cj1+1)(cjN+1)N xj.

Note that we add the subscript j to the right-hand side of the above equation to emphasize that this is calculated with respect to a single taxon across the N samples in a dataset. As intended, this scheme tended to assign smaller weights to taxa in our benchmarks with more zero and near-zero counts (Figure 2—figure supplement 1). Despite their heuristic nature, we found that our chosen weights provide performance improvements over alternative weights (or the lack thereof) as measured by our benchmark tasks (Supplementary file 1).

Our taxa weighting scheme supplements the use of pseudo-counts and represents a soft-threshold on low abundance taxa. More generally, these taxa weights represent a form of prior information regarding the importance of each taxon. We note that if prior biological information suggests allowing specific taxa to influence the PhILR transform more (or less) strongly, such a weighting could be achieved for taxon j by increasing (or decreasing) pj.

Incorporating branch lengths

Beyond utilizing the connectivity of the phylogenetic tree to dictate the partitioning scheme for ILR balances, branch length information can be embedded into the transformed space by linearly scaling ILR balances (yi*) by the distance between neighboring clades. We call this scaling by phylogenetic distance ‘branch length weighting’. Specifically, for each coordinate yi*, corresponding to node i we use the transform

yi*,blw=yi*f(di+,di)

where di± represent the branch lengths of the two direct children of node i. When f(di+,di)=1, the coordinates are not weighted by branch lengths. The form of this transform was chosen so that the weights di±, only influence the corresponding coordinate (yi*,blw).

We also investigated the effect of using f(di+,di)=1, f(di+,di)=di++di, and based on the results of Chen et al. (2012), f(di+,di)=di++di on benchmark performance. When coupled with the taxa weights specified above, the square root of the summed distances had the highest rank in 9 of the 12 supervised classification tasks and 2 of the three distance based tasks (Supplementary file 1). Based on these results, except for our analysis of balance variance versus phylogenetic depth (see below), the square root of the summed distances was used throughout our analyses.

Implementation

The PhILR transform, as well as the incorporation of branch length and taxa weightings has been implemented in the R programing language as the package philr available at https://bioconductor.org/packages/philr/.

Datasets and preprocessing

All data preprocessing was done in the R programming language using the phyloseq package for analysis of microbiome census data (McMurdie and Holmes, 2013) as well as the ape (Paradis et al., 2004) and phangorn (Schliep, 2011) packages for analysis of phylogenetic trees.

Data acquisition

We chose to use previously published OTU tables, taxonomic classifications, and phylogenies as the starting point for our analyses. The Human Microbiome Project (HMP) dataset was obtained from the QIIME Community Profiling Pipeline applied to high-quality reads from the v3-5 region, available at http://hmpdacc.org/HMQCP/. The Global Patterns dataset was originally published in Caporaso, et al. (Caporaso et al., 2011) and is provided with the phyloseq R package (McMurdie and Holmes, 2013). The Costello Skin Sites dataset (CSS) is a subset of the dataset collected by Costello et al. (Costello et al., 2009) featuring only the samples from skin sites. This skin subset was introduced as a benchmark for supervised machine learning by Knights et al. (Knights et al., 2011) and can be obtained from http://www.knightslab.org/data.

OTU table preprocessing

To accord with general practice, we performed a minimal level of OTU table filtering for all datasets used in benchmarks and analyses. Due to differences in sequencing depth, sequencing methodology, and the number and diversity of samples between datasets, filtering thresholds were set independently for each dataset. For the HMP dataset, we initially removed samples with fewer than 1000 counts to mimic prior analyses (Human Microbiome Project Consortium, 2012). We additionally removed OTUs that were not seen with more than three counts in at least 1% of samples. Preprocessing of the Global Patterns OTU table followed the methods outlined in McMurdie and Holmes (McMurdie and Holmes, 2013). Specifically, OTUs that were not seen with more than three counts in at least 20% of samples were removed, the sequencing depth of each sample was standardized to the abundance of the median sampling depth, and finally OTUs with a coefficient of variation ≤3.0 were removed. The CSS dataset had lower sequencing depth than the other two datasets; we chose to filter OTUs that were not seen with greater than 10 counts across the skin samples. The PhILR transform, and more generally our benchmarking results in Figure 2b and C, were robust to varying our filtering strategies (Supplementary file 2).

Preprocessing phylogenies

For each dataset, the phylogeny was pruned to include only those taxa remaining after OTU table preprocessing. Except for the Global Patterns dataset, which was already rooted, we chose to root phylogenies by manually specifying an outgroup. For the HMP dataset the phylum Euryarchaeota was chosen as an outgroup. For the CSS dataset, the tree was rooted with OTU 12871 (from phylum Plantomycetes) as the outgroup. For all three phylogenies, any multichotomies were resolved with the function multi2di from the ape package which replaces multichotomies with a series of dichotomies with one (or several) branch(es) of length zero.

Zero replacement and normalization

A pseudocount of 1 was added prior to PhILR transformation to avoid taking log-ratios with zero counts. We found that our benchmarking results were robust to changing the value of this pseudocount from 1 to 2, 3, or 10 (Supplementary file 1).

Grouping sampling sites

To simplify subsequent analyses, HMP samples from the left and right retroauricular crease and samples from the left and right antecubital fossa were grouped together, respectively, as preliminary PERMANOVA analysis suggested that these sites were indistinguishable (data not shown).

Benchmarking

Distance/dissimilarity based analysis

Distance between samples in PhILR transformed space was calculated using Euclidean distance. All other distance measures were calculated using phyloseq on the preprocessed data without adding a pseudocount. Principle coordinate analysis was performed for visualization using phyloseq. PERMANOVA was performed using the function adonis from the R package vegan (v2.3.4). The R2 value from the fitted model was taken as a performance metric. Standard errors were calculated using bootstrap resampling with 100 samples each. Differences between the performance of Euclidean distance in PhILR transformed space and that of each other distance or dissimilarity measure on a given task was tested using two-sided t-tests and multiple hypothesis testing was accounted for using FDR correction.

Supervised classification

The performance, as measured by classification accuracy, of PhILR transformed data was compared against data preprocessed using one of two standard strategies for normalizing sequencing depth: the preprocessed data was transformed to relative abundances (e.g., each sample was normalized to a constant sum of 1; raw); or, a pseudocount of 1 was added, the data was transformed to relative abundances, and finally the relative abundances were log-transformed (log).

All supervised learning was implemented in Python using the following libraries: Scikit-learn (v0.17.1), numpy (v1.11.0) and pandas (v0.17.1). Four classifiers were used: penalized logistic regression, support vector classification with RBF kernel, random forest classification, and k-nearest-neighbors classification. Each classification task was evaluated using the mean and variance of the test accuracy over 10 randomized test/train (30/70) splits which preserved the percentage of samples from each class at each split. For each classifier, for each split, the following parameters were set using cross-validation on the training set. Logistic regression and Support Vector Classification: the ‘C’ parameter was allowed to vary between 103 to 103 and multi-class classification was handled with a one-vs-all loss. In addition, for logistic regression the penalty was allowed to be either l1 or l2. K-nearest-neighbors classification: the ‘weights’ argument was set to ‘distance’. Random forest classification: each forest contained 30 trees and the ‘max_features’ argument was allowed to vary between 0.1 and 1. All other parameters were set to default values. Due to the small size of the Global Patterns dataset, the supervised classification task was simplified to distinguishing human vs. non-human samples. Differences between each method's accuracy in each task was tested using two-sided t-tests and multiple hypothesis testing was accounted for using FDR correction.

Identifying balances that distinguish sites

To identify a sparse set of balances that distinguish sampling sites while accounting for the dependencies between nested balances, we fit a multinomial regression model with a grouped l1 penalty using the R package glmnet (v2.0.5). The penalization term lambda was set by visually inspecting model outputs for clear body site separation (lambda = 0.1198). This resulted in 35 balances with non-zero regression coefficients. Phylogenetic tree visualization was done using the R package ggtree (Yu et al., 2017).

Variance and depth

To reduce the likelihood that our analysis of balance variance and phylogenetic depth was affected by statistical artifact, we modified our PhILR transform in several ways. First, we omitted branch length weights (i.e., we set f(di+,di)=1) as these may vary non-randomly as a function of phylogenetic depth. Second, we also anticipated that any zero replacement method would likely lead to lower variance measurements, which could have greater effects on balances closer to the tips of the tree. We therefore omitted taxa weights and zero replacement; we instead used stricter hard filtering thresholds and calculated balance values based on non-zero counts. In practice, we used the following filtering thresholds for each body site, taxa present in less than 20% of samples from that site were excluded and subsequently samples that had less than 50 total counts were excluded. To calculate balance values based on non-zero counts we retained balances that met the following criteria: the term gp(yi+)/gp(yi) had non-zero counts from some taxa within the subcomposition yi+ (formed by the taxa that descend from the +1 clade of node i) and some other taxa within the subcomposition yi (formed by the taxa that descend from the 1 clade of node i) in at least 40 samples from that body site. We believe these two modifications to PhILR resulted in a more conservative analysis of balance variance versus phylogenetic depth but are likely not optimal in other situations.

To investigate the overall relationship between balance variance and phylogenetic depth we used linear regression. A balance’s depth in the tree was calculated as its mean phylogenetic distance to its descendant tips (d). For a given body site the following model was fit:

logvar(y) =βlogd+ α

where d represents mean distance from a balance to its descendant tips. We then set out to test the null hypothesis that β=0, or that the variance of the log-ratio between two clades was invariant to the distance of the two clades from their most recent common ancestor. For each site, a null distribution for β was constructed by permutations of the tip labels of the phylogenetic tree. For each permutation of the labels, the resultant tree was used to transform the data and β was estimated. We chose this permutation scheme to ensure that the increasing variance we saw with increasing proximity of a balance to the root was not because deeper balances had more descendant tips, an artifact of variance scaling with mean abundance, or due to bias introduced due to our handling of zeros. Furthermore, for each body site, we found the null distribution for β was symmetric about β=0 which further supports that balance variance depends on phylogenetic depth through a biological mechanism and not through a statistical artifact (Figure 4—figure supplement 3). Two tailed p-values were calculated for β based on 20000 samples from each site’s respective null distribution. FDR correction was applied to account for multiple hypothesis testing between body sites.

To visualize local trends in the relationship between balance variance and phylogenetic depth, a LOESS regression was fit independently for each body site. This was done using the function geom_smooth from the R package ggplot2 (v2.1.0) with default parameters.

The data and code needed to reproduce our analysis of balance variance versus phylogenetic depth is provided in Figure 4—source data 1 and Figure 4—source code 1 respectively.

Integrating taxonomic information

Taxonomy was assigned to OTUs in the HMP dataset using the assign_taxonomy.py script from Qiime (v1.9.1) to call uclust (v1.2.22) with default parameters using the representative OTU sequences obtained as described above. Taxonomic identifiers were assigned to the two descendant clades of a given balance separately using a simple voting scheme and combined into a single name for that balance. The voting scheme occurs as follows: (1) for a given clade, the entire taxonomy table was subset to only contain the OTUs that were present in that clade (2) starting at the finest taxonomic rank the subset taxonomy table was checked to see if any species identifier represented ≥95% of the table entries at that taxonomic rank, if so that identifier was taken as the taxonomic label for the clade (3) if no consensus identifier was found, the table was checked at the next most-specific taxonomic rank.

Median phylogenetic depths for each taxonomic rank were estimated by first decorating a phylogenetic tree with taxonomy information using tax2tree (v1.0) (McDonald et al., 2012). For a given taxonomic rank the mean distance to tips was calculated for each internal node possessing a label that ended in that rank. The median of these distances was used to display an estimate of the phylogenetic depth of that given rank. This calculation of median phylogenetic depth of different taxonomic ranks was done separately for each body site.

The data and code needed to reproduce the taxonomic assignment and estimation of median phylogenetic depths for each taxonomic rank is included in Figure 4—source data 1 and Figure 4—source code 1 respectively.

Acknowledgements

We thank Rachel Silverman, Aspen Reese, Firas Midani, Heather Durand, Jesse Shapiro, Jonathan Friedman, Simon Levin, and Susan Holmes for their helpful comments, Dan Knights for providing us with the CSS dataset, as well as Klaus Schliep and Liam Revell for their insight into manipulation of phylogenetic trees in the R programming language. JS was supported in part by the Duke University Medical Scientist Training Program (GM007171).

Funding Statement

The funders had no role in study design, data collection and interpretation, or the decision to submit the work for publication.

Funding Information

This paper was supported by the following grants:

  • National Science Foundation IIS 1546331 to Sayan Mukherjee.

  • National Science Foundation IIS 1418261 to Sayan Mukherjee.

  • Global Probiotics Council Young Investigator Grant for Probiotics Research to Lawrence A David.

  • Searle Scholars Program 15-SSP-184 Research Agreement to Lawrence A David.

  • Alfred P. Sloan Foundation BR2014-003 to Lawrence A David.

  • Duke University Medical Scientist Training Program GM007171 to Justin D Silverman.

Additional information

Competing interests

The authors declare that no competing interests exist.

Author contributions

JDS, Conceptualization, Software, Formal analysis, Investigation, Visualization, Methodology, Writing—original draft, Writing—review and editing.

ADW, Conceptualization, Investigation, Methodology, Writing—original draft, Writing—review and editing.

SM, Formal analysis, Supervision, Funding acquisition, Methodology, Writing—original draft, Writing—review and editing.

LAD, Conceptualization, Resources, Formal analysis, Supervision, Funding acquisition, Investigation, Writing—original draft, Project administration, Writing—review and editing.

Additional files

Supplementary file 1. Extended benchmarking results including the effects of changing pseudocount and different choices for taxa and branch length weights.

This file also contains summary information for the benchmark datasets after preprocessing.

DOI: http://dx.doi.org/10.7554/eLife.21887.016

elife-21887-supp1.xlsx (46KB, xlsx)
DOI: 10.7554/eLife.21887.016
Supplementary file 2. Extended benchmarking results regarding the sensitivity of methods to different OTU filtering schemes.

DOI: http://dx.doi.org/10.7554/eLife.21887.017

elife-21887-supp2.xlsx (42.9KB, xlsx)
DOI: 10.7554/eLife.21887.017

Major datasets

The following previously published datasets were used:

Human Microbiome Project Consortium,2010,Human Microbiome Project,http://hmpdacc.org/HMQCP/,Publicly available at HMPDACC (v35 download of files 6, 9, and 10).

Costello EK,Lauber CL,Hamady M,Fierer N,Gordon JI,Knight R,2009,Costello Skin Sites,http://www.knightslab.org/data,Publicly available as part of the FEMS Benchmark dataset (2011) provided by Dan Knights

Caporaso GJ,Lauber CL,Walters WA,Berg-Lyons D,Lozupone CA,Turnbaugh PJ,Fierer N,Knight R,2011,Global Patterns,https://bioconductor.org/packages/release/bioc/html/phyloseq.html,Publicly available and provided as part of the phyloseq R package as 'GlobalPatterns'

References

  1. Aas JA, Paster BJ, Stokes LN, Olsen I, Dewhirst FE. Defining the normal bacterial flora of the oral cavity. Journal of Clinical Microbiology. 2005;43:5721–5732. doi: 10.1128/JCM.43.11.5721-5732.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Aitchison J. The Statistical Analysis of Compositional Data. London; New York: Chapman and Hall; 1986. [Google Scholar]
  3. Anders S, Huber W. Differential expression analysis for sequence count data. Genome Biology. 2010;11:R106. doi: 10.1186/gb-2010-11-10-r106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Bacon-Shone J. A Short History of Compositional Data Analysis. In: Pawlowsky-Glahn V, Buccianti A, editors. Compositional Data Analysis. Hoboken, NJ: John Wiley & Sons, Ltd; 2011. pp. 1–11. [Google Scholar]
  5. Bear J, Billheimer D. A logistic normal mixture model for compositional data allowing essential zeros. Austrian Journal of Statistics. 2016;45:3–23. doi: 10.17713/ajs.v45i4.117. [DOI] [Google Scholar]
  6. Billheimer D, Guttorp P, Fagan WF. Statistical interpretation of species composition. Journal of the American Statistical Association. 2001;96:1205–1214. doi: 10.1198/016214501753381850. [DOI] [Google Scholar]
  7. Blaser MJ, Falkow S. What are the consequences of the disappearing human Microbiota? Nature Reviews Microbiology. 2009;7:887–894. doi: 10.1038/nrmicro2245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Britanova OV, Putintseva EV, Shugay M, Merzlyak EM, Turchaninova MA, Staroverov DB, Bolotin DA, Lukyanov S, Bogdanova EA, Mamedov IZ, Lebedev YB, Chudakov DM. Age-related decrease in TCR repertoire diversity measured with deep and normalized sequence profiling. The Journal of Immunology. 2014;192:2689–2698. doi: 10.4049/jimmunol.1302064. [DOI] [PubMed] [Google Scholar]
  9. Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Lozupone CA, Turnbaugh PJ, Fierer N, Knight R. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. PNAS. 2011;108 Suppl 1:4516–4522. doi: 10.1073/pnas.1000080107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Chen J, Bittinger K, Charlson ES, Hoffmann C, Lewis J, Wu GD, Collman RG, Bushman FD, Li H. Associating microbiome composition with environmental covariates using generalized UniFrac distances. Bioinformatics. 2012;28:2106–2113. doi: 10.1093/bioinformatics/bts342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Chen J, Li H. Variable selection for sparse Dirichlet-multinomial regression with an application to microbiome data analysis. The Annals of Applied Statistics. 2013;7:418–442. doi: 10.1214/12-AOAS592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Costello EK, Lauber CL, Hamady M, Fierer N, Gordon JI, Knight R. Bacterial community variation in human body habitats across space and time. Science. 2009;326:1694–1697. doi: 10.1126/science.1177486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Culley AI, Lang AS, Suttle CA. Metagenomic analysis of coastal RNA virus communities. Science. 2006;312:1795–1798. doi: 10.1126/science.1127404. [DOI] [PubMed] [Google Scholar]
  14. De Filippo C, Cavalieri D, Di Paola M, Ramazzotti M, Poullet JB, Massart S, Collini S, Pieraccini G, Lionetti P. Impact of diet in shaping gut Microbiota revealed by a comparative study in children from Europe and rural Africa. PNAS. 2010;107:14691–14696. doi: 10.1073/pnas.1005963107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Egozcue JJ, Pawlowsky-Glahn V, Mateu-Figueras G, Barcelo-Vidal C. Isometric logratio transformations for compositional data analysis. Mathematical Geology. 2003;35:279–300. doi: 10.1023/A:1023818214614. [DOI] [Google Scholar]
  16. Egozcue JJ, Pawlowsky-Glahn V. Groups of parts and their balances in compositional data analysis. Mathematical Geology. 2005;37:795–828. doi: 10.1007/s11004-005-7381-9. [DOI] [Google Scholar]
  17. Egozcue JJ, Pawlowsky-Glahn V. Changing the reference measure in the simplex and its weighting effects. Austrian Journal of Statistics. 2016;45:25–44. doi: 10.17713/ajs.v45i4.126. [DOI] [Google Scholar]
  18. Fang H, Huang C, Zhao H, Deng M. CCLasso: correlation inference for compositional data through lasso. Bioinformatics. 2015;31:3172–3180. doi: 10.1093/bioinformatics/btv349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Faust K, Sathirapongsasuti JF, Izard J, Segata N, Gevers D, Raes J, Huttenhower C. Microbial co-occurrence relationships in the human microbiome. PLoS Computational Biology. 2012;8:e1002606. doi: 10.1371/journal.pcbi.1002606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Fernandes AD, Reid JN, Macklaim JM, McMurrough TA, Edgell DR, Gloor GB. Unifying the analysis of high-throughput sequencing datasets: characterizing RNA-seq, 16S rRNA gene sequencing and selective growth experiments by compositional data analysis. Microbiome. 2014;2:15. doi: 10.1186/2049-2618-2-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Finucane MM, Sharpton TJ, Laurent TJ, Pollard KS. A taxonomic signature of obesity in the microbiome? getting to the guts of the matter. PLoS One. 2014;9:e84689. doi: 10.1371/journal.pone.0084689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Friedman J, Alm EJ. Inferring correlation networks from genomic survey data. PLoS Computational Biology. 2012;8:e1002687. doi: 10.1371/journal.pcbi.1002687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Fukuyama J, McMurdie PJ, Dethlefsen L, Relman DA, Holmes S. Pacific Symposium on Biocomputing. NIH Public Access; 2012. Comparisons of distance methods for combining covariates and abundances in microbiome studies; p. 213. [PMC free article] [PubMed] [Google Scholar]
  24. Gloor GB, Macklaim JM, Vu M, Fernandes AD. Compositional uncertainty should not be ignored in high-throughput sequencing data analysis. Austrian Journal of Statistics. 2016a;45:73. doi: 10.17713/ajs.v45i4.122. [DOI] [Google Scholar]
  25. Gloor GB, Wu JR, Pawlowsky-Glahn V, Egozcue JJ. It's all relative: analyzing microbiome data as compositions. Annals of Epidemiology. 2016b;26:322–329. doi: 10.1016/j.annepidem.2016.03.003. [DOI] [PubMed] [Google Scholar]
  26. Good IJ. On the estimation of small frequencies in Contingency-Tables. Journal of the Royal Statistical Society Series B-Statistical Methodology. 1956;18:113–124. [Google Scholar]
  27. Grice EA, Segre JA. The skin microbiome. Nature Reviews Microbiology. 2011;9:244–253. doi: 10.1038/nrmicro2537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Human Microbiome Project Consortium Structure, function and diversity of the healthy human microbiome. Nature. 2012;486:207–214. doi: 10.1038/nature11234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Hunt DE, David LA, Gevers D, Preheim SP, Alm EJ, Polz MF. Resource partitioning and sympatric differentiation among closely related bacterioplankton. Science. 2008;320:1081–1085. doi: 10.1126/science.1157890. [DOI] [PubMed] [Google Scholar]
  30. Jackson DA. Compositional data in community ecology: the paradigm or peril of proportions? Ecology. 1997;78:929–940. doi: 10.1890/0012-9658(1997)078[0929:CDICET]2.0.CO;2. [DOI] [Google Scholar]
  31. Janda JM, Abbott SL. 16s rRNA gene sequencing for bacterial identification in the diagnostic laboratory: pluses, perils, and pitfalls. Journal of Clinical Microbiology. 2007;45:2761–2764. doi: 10.1128/JCM.01228-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Knights D, Costello EK, Knight R. Supervised classification of human Microbiota. FEMS Microbiology Reviews. 2011;35:343–359. doi: 10.1111/j.1574-6976.2010.00251.x. [DOI] [PubMed] [Google Scholar]
  33. Kurtz ZD, Müller CL, Miraldi ER, Littman DR, Blaser MJ, Bonneau RA. Sparse and compositionally robust inference of microbial ecological networks. PLOS Computational Biology. 2015;11:e1004226. doi: 10.1371/journal.pcbi.1004226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. La Rosa PS, Brooks JP, Deych E, Boone EL, Edwards DJ, Wang Q, Sodergren E, Weinstock G, Shannon WD. Hypothesis testing and power calculations for taxonomic-based human microbiome data. PLoS ONE. 2012;7:e52078. doi: 10.1371/journal.pone.0052078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Lee SC, Tang MS, Lim YA, Choy SH, Kurtz ZD, Cox LM, Gundra UM, Cho I, Bonneau R, Blaser MJ, Chua KH, Loke P. Helminth colonization is associated with increased diversity of the gut Microbiota. PLoS Neglected Tropical Diseases. 2014;8:e2880. doi: 10.1371/journal.pntd.0002880. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Levy R, Borenstein E. Metabolic modeling of species interaction in the human microbiome elucidates community-level assembly rules. PNAS. 2013;110:12804–12809. doi: 10.1073/pnas.1300926110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Ley RE, Hamady M, Lozupone C, Turnbaugh PJ, Ramey RR, Bircher JS, Schlegel ML, Tucker TA, Schrenzel MD, Knight R, Gordon JI. Evolution of mammals and their gut microbes. Science. 2008;320:1647–1651. doi: 10.1126/science.1155725. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Lê Cao KA, Costello ME, Lakis VA, Bartolo F, Chua XY, Brazeilles R, Rondeau P. MixMC: a multivariate statistical framework to gain insight into microbial communities. PLoS One. 2016;11:e0160169. doi: 10.1371/journal.pone.0160169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Li H. Microbiome, metagenomics, and high-dimensional compositional data analysis. Annual Review of Statistics and Its Application. 2015;2:73–94. doi: 10.1146/annurev-statistics-010814-020351. [DOI] [Google Scholar]
  40. Lin W, Shi P, Feng R, Li H. Variable selection in regression with compositional covariates. Biometrika. 2014;101:785–797. doi: 10.1093/biomet/asu031. [DOI] [Google Scholar]
  41. Lovell D, Müller W, Taylor J, Zwart A, Helliwell C. Proportions, percentages, ppm: do the molecular biosciences treat compositional data right. In: Pawlowsky-Glahn V, Buccianti A, editors. Compositional Data Analysis: Theory and Applications. Hoboken, NJ: John Wiley & Sons, Ltd; 2011. pp. 193–207. [Google Scholar]
  42. Lovell D, Pawlowsky-Glahn V, Egozcue JJ, Marguerat S, Bähler J. Proportionality: a valid alternative to correlation for relative data. PLOS Computational Biology. 2015;11:e1004075. doi: 10.1371/journal.pcbi.1004075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Lozupone C, Knight R. UniFrac: a new phylogenetic method for comparing microbial communities. Applied and Environmental Microbiology. 2005;71:8228–8235. doi: 10.1128/AEM.71.12.8228-8235.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Mager DL, Ximenez-Fyvie LA, Haffajee AD, Socransky SS. Distribution of selected bacterial species on intraoral surfaces. Journal of Clinical Periodontology. 2003;30:644–654. doi: 10.1034/j.1600-051X.2003.00376.x. [DOI] [PubMed] [Google Scholar]
  45. Mandal S, Van Treuren W, White RA, Eggesbø M, Knight R, Peddada SD. Analysis of composition of microbiomes: a novel method for studying microbial composition. Microbial Ecology in Health & Disease. 2015;26:27663. doi: 10.3402/mehd.v26.27663. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Martın-Fernandez JA, Palarea-Albaladejo J, Olea RA. Dealing with zeros. In: Pawlowsky-Glahn V, Buccianti A, editors. Compositional Data Analysis: Theory and Applications. Hoboken, NJ: John Wiley & Sons, Ltd; 2011. pp. 43–58. [Google Scholar]
  47. Martin-Fernandez J-A, Hron K, Templ M, Filzmoser P, Palarea-Albaladejo J. Bayesian-multiplicative treatment of count zeros in compositional data sets. Statistical Modelling. 2015;15:134–158. doi: 10.1177/1471082X14535524. [DOI] [Google Scholar]
  48. Martiny JB, Jones SE, Lennon JT, Martiny AC. Microbiomes in light of traits: a phylogenetic perspective. Science. 2015;350:aac9323. doi: 10.1126/science.aac9323. [DOI] [PubMed] [Google Scholar]
  49. Matsen FA. Phylogenetics and the human microbiome. Systematic Biology. 2015;64:e26–e41. doi: 10.1093/sysbio/syu053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. McDonald D, Price MN, Goodrich J, Nawrocki EP, DeSantis TZ, Probst A, Andersen GL, Knight R, Hugenholtz P. An improved greengenes taxonomy with explicit ranks for ecological and evolutionary analyses of Bacteria and archaea. The ISME Journal. 2012;6:610–618. doi: 10.1038/ismej.2011.139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. McMurdie PJ, Holmes S. Phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS One. 2013;8:e61217. doi: 10.1371/journal.pone.0061217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. McMurdie PJ, Holmes S. Waste not, want not: why rarefying microbiome data is inadmissible. PLoS Computational Biology. 2014;10:e1003531. doi: 10.1371/journal.pcbi.1003531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Moeller AH, Caro-Quintero A, Mjungu D, Georgiev AV, Lonsdorf EV, Muller MN, Pusey AE, Peeters M, Hahn BH, Ochman H. Cospeciation of gut microbiota with hominids. Science. 2016;353:380–382. doi: 10.1126/science.aaf3951. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Morton JT, Sanders J, Quinn RA, McDonald D, Gonzalez A, Vázquez-Baeza Y, Navas-Molina JA, Song SJ, Metcalf JL, Hyde ER, Lladser M, Dorrestein PC, Knight R. Balance trees reveal microbial niche differentiation. mSystems. 2017;2:e00162-16. doi: 10.1128/mSystems.00162-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Paradis E, Claude J, Strimmer K. APE: analyses of phylogenetics and evolution in R language. Bioinformatics. 2004;20:289–290. doi: 10.1093/bioinformatics/btg412. [DOI] [PubMed] [Google Scholar]
  56. Paulson JN, Stine OC, Bravo HC, Pop M. Differential abundance analysis for microbial marker-gene surveys. Nature Methods. 2013;10:1200–1202. doi: 10.1038/nmeth.2658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Pawlowsky-Glahn V, Buccianti A. Compositional Data Analysis: Theory and Applications. Hoboken, NJ: John Wiley & Sons, Ltd; 2011. [Google Scholar]
  58. Pawlowsky-Glahn V, Egozcue JJ, Tolosana-Delgado R. Modeling and Analysis of Compositional Data. Hoboken, NJ: John Wiley & Sons, Ltd; 2015. [Google Scholar]
  59. Pawlowsky-Glahn V, Egozcue JJ. Exploring compositional data with the CoDa-Dendogram. Austrian Journal of Statistics. 2011;40:103–113. [Google Scholar]
  60. Purdom E. Analysis of a data matrix and a graph: metagenomic data and the phylogenetic tree. The Annals of Applied Statistics. 2011;5:2326–2358. doi: 10.1214/10-AOAS402. [DOI] [Google Scholar]
  61. Rakoff-Nahoum S, Foster KR, Comstock LE. The evolution of cooperation within the gut microbiota. Nature. 2016;533:255–259. doi: 10.1038/nature17626. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Roth A, Khattra J, Yap D, Wan A, Laks E, Biele J, Ha G, Aparicio S, Bouchard-Côté A, Shah SP. PyClone: statistical inference of clonal population structure in Cancer. Nature Methods. 2014;11:396–398. doi: 10.1038/nmeth.2883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Schliep KP. Phangorn: phylogenetic analysis in R. Bioinformatics. 2011;27:592–593. doi: 10.1093/bioinformatics/btq706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Smillie CS, Smith MB, Friedman J, Cordero OX, David LA, Alm EJ. Ecology drives a global network of gene exchange connecting the human microbiome. Nature. 2011;480:241–244. doi: 10.1038/nature10571. [DOI] [PubMed] [Google Scholar]
  65. Tsilimigras MC, Fodor AA. Compositional data analysis of the microbiome: fundamentals, tools, and challenges. Annals of Epidemiology. 2016;26:330–335. doi: 10.1016/j.annepidem.2016.03.002. [DOI] [PubMed] [Google Scholar]
  66. van den Boogaart KG, Tolosana-Delgado R. Analyzing Compositional Data with R. New York: Springer; 2013. [Google Scholar]
  67. Větrovský T, Baldrian P. The variability of the 16S rRNA gene in bacterial genomes and its consequences for bacterial community analyses. PLoS One. 2013;8:e57923. doi: 10.1371/journal.pone.0057923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Waldor MK, Tyson G, Borenstein E, Ochman H, Moeller A, Finlay BB, Kong HH, Gordon JI, Nelson KE, Dabbagh K, Smith H. Where next for microbiome research? PLOS Biology. 2015;13:e1002050. doi: 10.1371/journal.pbio.1002050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Wu GD, Chen J, Hoffmann C, Bittinger K, Chen YY, Keilbaugh SA, Bewtra M, Knights D, Walters WA, Knight R, Sinha R, Gilroy E, Gupta K, Baldassano R, Nessel L, Li H, Bushman FD, Lewis JD, Gd W. Linking long-term dietary patterns with gut microbial enterotypes. Science. 2011;334:105–108. doi: 10.1126/science.1208344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Yatsunenko T, Rey FE, Manary MJ, Trehan I, Dominguez-Bello MG, Contreras M, Magris M, Hidalgo G, Baldassano RN, Anokhin AP, Heath AC, Warner B, Reeder J, Kuczynski J, Caporaso JG, Lozupone CA, Lauber C, Clemente JC, Knights D, Knight R, Gordon JI. Human gut microbiome viewed across age and geography. Nature. 2012;486:222–227. doi: 10.1038/nature11053. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Yu G, Smith DK, Zhu H, Guan Y, Lam TT-Y. Ggtree : an r package for visualization and annotation of phylogenetic trees with their covariates and other associated data. Methods in Ecology and Evolution. 2017;8:28–36. doi: 10.1111/2041-210X.12628. [DOI] [Google Scholar]
  72. Yuan K, Sakoparnig T, Markowetz F, Beerenwinkel N. BitPhylogeny: a probabilistic framework for reconstructing intra-tumor phylogenies. Genome Biology. 2015;16:36. doi: 10.1186/s13059-015-0592-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Zaneveld JR, Lozupone C, Gordon JI, Knight R. Ribosomal RNA diversity predicts genome diversity in gut bacteria and their relatives. Nucleic Acids Research. 2010;38:3869–3879. doi: 10.1093/nar/gkq066. [DOI] [PMC free article] [PubMed] [Google Scholar]
eLife. 2017 Feb 15;6:e21887. doi: 10.7554/eLife.21887.024

Decision letter

Editor: Anthony Fodor1

In the interests of transparency, eLife includes the editorial decision letter and accompanying author responses. A lightly edited version of the letter sent to the authors after peer review is shown, indicating the most substantive concerns; minor comments are not usually included.

Thank you for submitting your article "A phylogenetic transform enhances analysis of compositional microbiota data" for consideration by eLife. Your article has been favorably evaluated by Wendy Garrett (Senior Editor) and three reviewers, one of whom, Anthony Fodor, is a member of our Board of Reviewing Editors. The following individual involved in review of your submission has agreed to reveal their identity: Juan José Egozcue (Reviewer #2). All three reviewers found considerable merit in the manuscript praising the originality of the idea, the clarity of the presentation and the effectiveness of the visualizations.

For your reference, I have attached the three original reviews below. Per the eLife policy, you are not required to respond to each individual point raised by each reviewer. Below you will find a summary of the reviews reflecting the consensus of the reviewers of required revisions.

Summary:

It has long been recognized that compositional data is best analyzed in ratio space. For 16S rRNA sequence datasets and other genomic datasets, this leads to the rather difficult question of what ratio to take. In this interesting manuscript, Silverman et al. suggest taking ratios on a bifurcating phylogenetic tree. In their approach, the dependent variable for each node of the tree is a "balance", which is a normalized ratio of all the nodes beneath each node (with the ratio of all the left descendant nodes over all the right descendant nodes). The paper explores the consequences of this normalization showing reasonable performance when used for classification or machine learning. The authors provide an implementation in R and this approach may be of utility for analysts looking for a novel way to analyze microbial data in ratio space.

Essential revisions:

1) There was consensus among the reviewers that the normalization scheme seemed a bit ad-hoc and poorly justified. We suggest that it is made clear to the readers that the choice of weights used in the manuscript, both on the taxa and on the tips, are preliminary choices and further work will be required to ultimately determine the best weighting scheme. In addition, some further clarification on how sparsity was treated would strengthen the manuscript.

2) None of the reviewers were entirely satisfied with the demonstration that the proposed transformation was superior to other transformations. It is unclear what it means, for example, that the PhILR technique sometimes produces a stronger clustering signal and sometimes a weaker clustering signal than weighted UNIFRAC (Figure 2B). There was some discussion among the reviewers about whether a comparison to random binary partitions would be appropriate since this could reflect your null model (that phylogenetic binary partition is superior). Reviewer #2 was not enthusiastic about this idea and argued that a change on a tree does not alter (Aitchison) distances between data points. Therefore, results would likely not change no matter which tree has been selected for defining ILR coordinates.

In the end, the consensus of the reviewers was that exploring such null models might be of interest, but would not be appropriate as a requirement for revision. Nonetheless, reviewer #1 would have been interested to see if machine learning algorithms in fact performed better on the phylogenetic partition in comparison to random partitions. We leave it up to the authors as to whether they feel such an inclusion would improve their paper or be worth the effort. However, all reviewers felt that paper would be improved by more discussion of how to tell when one normalization procedure is superior to another. It seemed to reviewer #1 that it might be adequate to say that the proposed transformation appears to be doing no harm relative to weighted Unifrac but was providing some compositional safeguards. Reviewer #2 argued that the main benefit of the proposed procedure was in interpretability in yielding a phylogenetically aware transformation that could be utilized for test-statistics.

3) There was general consensus among the reviewers that Figure 4 was unconvincing in that technical (non-biological) features of the algorithm could explain the results more easily than exclusionary dynamics. The authors could likely remove Figure 4 without damaging the paper, but if they choose to keep it, they should respond to the concerns below from reviewers #1 and #3 as both reviewers felt that simple lack of association between taxa and stochastic variation in sampling could explain many of the observations in Figure 4.

4) There was consensus among the reviewers that the authors could provide more guidance to the non-specialists about when this algorithm should be used. As reviewer #3 put it, when and why should a reader choose the PhILR approach and when should it be avoided.

5) The authors should be explicit about what filtering and other pre-processing steps were used and whether those choices substantially impacted the performance of their algorithm.

Reviewer #1:

It has long been recognized that compositional data is best analyzed in ratio space. For 16S rRNA sequence datasets and other genomic datasets, this leads to the rather difficult question of what ratio to take. In this interesting manuscript, Silverman et al. suggest taking ratios on a bifurcating phylogenetic tree. In their approach, the dependent variable for each node of the tree is a "balance", which is a normalized ratio of all the nodes beneath each node (with the ratio of all the left descendant nodes over all the right descendant nodes). This is a clever idea and the authors make a reasonably compelling case that this is a useful contribution. While the evidence from the machine learning and clustering portions of the paper provide only modest evidence that this transformation is substantially better than other transformations, the transformation clearly is not making things much worse. As such, the proposed method will be of interest to researchers interested in how best to approach the challenging problem of compositional analyses of microbial datasets.

Suggestions for improvement follows:

1) The mathematical description of the algorithm could be improved in terms of its clarity. For example, it is not always clear how the pluses and minuses in the second equation of the subsection “The ILR Transform” are defined. It is not made immediately clear that the leftmost term (sqrt(ni+ * ni- / (ni+ + ni-)) in that equation is a correction for branch length. For the equation in the subsection “Addressing sparsity through weighting taxa”, please use parentheses to emphasize that the formula is sqrt(sum(xj)) * ((sum(cj 1…N)) ^ 1/N and not sqrt(sum(xj))^N * sqrt(sum(cj 1…N))

2) A limitation of working in ratio space in 16S data is what to do with all of the zeros in the spreadsheet (the "sparse data" problem). There is nothing inherent to the balancing algorithm that can deal with this problem (the sum in the denominator of the balance obviously can't be zero). The authors therefore require some kind of pseudo-count scheme and they have settled on one (given in the equation in subsection “Addressing sparsity through weighting taxa”) which is a function of the number of samples. If I am reading their formula correctly, the weight given to a taxa with 1 read in one sample and zero in all others would be very close to the sqrt(number of samples). The authors might consider giving this as an example and explicitly noting that their method has some similarity to adding the sqrt(number of samples) as a pseudo-count.

3) I struggled a bit with the variance analysis which concludes the Results sections. In the second paragraph of the subsection 2Balance variance and microbiota assembly” it is argued that "When the variance of a balance between two clades approaches zero, the mean abundance of taxa in each of the two clades will be linearly related and thus exhibit shared dynamics across microbial habitats". While this is true (and must be true mathematically as when the variance approaches zero, the ratio of the two normalized abundances of the clades is constant, which means they will become perfectly linearly related with a positive slope), it's not clear to me that a high variance necessarily means "exclusionary dynamics across samples". I don't see how a high variance in this number indicated anything about how direct or indirect the interactions are within the tree. There are lots of reasons two clades could lack perfect linear correlation with a positive slope from direct competition to being completely unrelated biologically with abundances just shifting due to stochastic variation. Also, wouldn't variance have to be lower closer to the tips of the phylogenetic tree as the number of sequences is lower, so the pseudo-count has more of an effect depressing overall variance? Doesn't that explain Figure 4D-F? And if I am reading these graphs correctly, below the species level there are not enough datapoints to support modeling via localized regression one way or another. Although not central to the arguments of the paper, I found Figure 4 to be less convincing than other sections of the paper.

Reviewer #2:

The manuscript introduces two new powerful ideas. They are: (a)construction of the sequential binary partition used to define

ILR coordinates on the base of phylogenetic trees; (b) use therecently published balances in weighted compositions to account forphylogenetic distances between OTU's or taxa. Both ideas aredeveloped and applied successfully to benchmark data sets.

For the originality of main ideas applied to microbiome analysis,the manuscript deserves publication.

Reviewer #3:

The manuscript is, in general, written in a clear and logical manner with a minimum of technical jargon. Figures 1, 3 and 4 are outstanding examples of data visualization and demonstrate the utility of this approach over others extant in the field. Figure 5 seems an afterthought and could be combined with Figure 1, or made supplementary. The binary portion could also be completely described in the Methods without a figure.

Figure 2, and the associated descriptive text are somewhat underwhelming demonstrations of the utility of the approach, which, in theory, should clearly outperform all other methods if the compositional data approach is necessary. The use of box-plots probably does not aid the visual appeal.

That the ILR based distance can seriously underperform the weighted unifrac metric is puzzling and cause for concern, since PhILR becomes another metric to be 'mined' to achieve the desired outcome. Does the PhILR partitioning scheme outperform a simple random binary partition? this is crucial to know.

There needs to be serious guidance given to the reader as to when and why to choose the PhILR approach and when not to. The same can be said for the classification methods shown, where the PhILR generally (but not always) outperforms relative abundance, but not log relative abundance. I am left to wonder if data analysis choices are driving these results.

Finally, I am puzzled as to why the authors chose to take a rigorous variance-based approach (ILR) and analyze it using a distance-based metric. It is established in the CoDa literature that Mahalanobis distances are often more appropriate for clustering based approaches and for outlier detection (see for example: http://www.statistik.tuwien.ac.at/public/filz/papers/2010Compstat.pdf). Did the authors examine other distances?

I believe that Figure 4 is somewhat over-interpreted. Subsection “Balance variance and microbiota assembly”, last paragraph: There are many technical reasons such as inefficient clustering, etc. that closely related sequences (called bacteria here) could exhibit strong correlations. Reading Lovell (referenced in this manuscript), and the points shown in 4G-L, indicate that an alternative explanation to that given in the text (see aforementioned subsection, end of second paragraph) is that a higher balance variance would be compatible with a model of indifference rather than competition since the vast majority of high variance ratio balances are likely to be non-correlated balances (Lovell 2015). Such non-correlated balances persist even out to the extreme tails and there is as yet no general principled approach to identify negative correlations in compositions.

Introduction, last paragraph: “the accuracy of distance.…” – this is a strong statement to make as the underlying standard of truth is assumed, not known from first principles.

Subsection “Addressing sparsity through weighting taxa”, first paragraph: I would argue that total read counts contain information on precision, not variance. It is hard to square this statement with the otherwise compositional approach used in the paper.

Subsection “Addressing sparsity through weighting taxa”, second paragraph: Why use a prior of one. The supplement that I found did not support the assertion that the prior chosen had little effect. This needs to be clarified.

Subsection “Human Microbiome Project (HMP)”: What is the effect of filtering and preprocessing on the method? As with many methods papers, the arbitrary choices here have the potential to become the accepted norm later without evidence.

Subsection “Supervised Classification”, first paragraph: Were the log-transformed data not scaled? if not, this is a problem since it is not a fair comparison.

eLife. 2017 Feb 15;6:e21887. doi: 10.7554/eLife.21887.025

Author response


Essential revisions:

1) There was consensus among the reviewers that the normalization scheme seemed a bit ad-hoc and poorly justified. We suggest that it is made clear to the readers that the choice of weights used in the manuscript, both on the taxa and on the tips, are preliminary choices and further work will be required to ultimately determine the best weighting scheme. In addition, some further clarification on how sparsity was treated would strengthen the manuscript.

We understand the reviewers’ concern regarding the discussion of our chosen weights (both branch length weights and taxa weights) and how sparsity was handled. We also recognize that these chosen weights are heuristics motivated in part on benchmarking results, and that further work is needed to determine optimal weightings in different situations. In response to these comments, we have expanded our discussion of these issues in the Results, Discussion and Methods sections (see subsection “Constructing the PhILR transform”; Discussion, sixth paragraph; subsections “Soft thresholding through weighting taxa”, “Incorporating branch lengths” and “Implementation”). In particular, we now state in the main text:

“Both the taxa weights and the branch lengths weights we introduce here may be viewed as preliminary heuristics; future work will likely yield additional weighting schemes, as well as knowledge for when a given weighting scheme should be matched to an analysis task.”

We have also included an additional supplementary figure (Figure 4—figure supplement 2) to provide readers with intuition for our chosen taxa weights behave in the face of data sparsity; namely, that these weights assign smaller influence to taxa with more zero and near-zero counts.

2) None of the reviewers were entirely satisfied with the demonstration that the proposed transformation was superior to other transformations. It is unclear what it means, for example, that the PhILR technique sometimes produces a stronger clustering signal and sometimes a weaker clustering signal than weighted UNIFRAC (Figure 2B). There was some discussion among the reviewers about whether a comparison to random binary partitions would be appropriate since this could reflect your null model (that phylogenetic binary partition is superior). Reviewer #2 was not enthusiastic about this idea and argued that a change on a tree does not alter (Aitchison) distances between data points. Therefore, results would likely not change no matter which tree has been selected for defining ILR coordinates.

In the end, the consensus of the reviewers was that exploring such null models might be of interest, but would not be appropriate as a requirement for revision. Nonetheless, reviewer #1 would have been interested to see if machine learning algorithms in fact performed better on the phylogenetic partition in comparison to random partitions. We leave it up to the authors as to whether they feel such an inclusion would improve their paper or be worth the effort. However, all reviewers felt that paper would be improved by more discussion of how to tell when one normalization procedure is superior to another. It seemed to reviewer #1 that it might be adequate to say that the proposed transformation appears to be doing no harm relative to weighted Unifrac but was providing some compositional safeguards. Reviewer #2 argued that the main benefit of the proposed procedure was in interpretability in yielding a phylogenetically aware transformation that could be utilized for test-statistics.

We recognize from the reviewers’ comments that our rationale for providing the benchmarking results could be improved. Our goal in Figure 2 was to demonstrate that common microbiota analyses could be easily performed on PhILR transformed data with performance on par with current practice. Importantly, we did not intend these benchmarks to justify a compositionally robust approach to microbiota data analysis; we believe that this need has been well established (1-7). Nor do we believe it essential to our manuscript that PhILR outperforms other transforms or that Euclidean distance measured on PhILR transformed data outperforms all other (non-CoDA) distance metrics. In fact, we find it extremely unlikely that PhILR would always outperform any other transformation or distance metric under all situations. For example, while Weighted UniFrac may not be grounded in compositional data analysis theory, it was specifically designed for analysis of mixed microbial communities and we therefore would expect it to preform quite well in these tasks. We do think though that the results of Figure 2B and 2C are promising and suggest that at least with respect to our chosen metrics, PhILR transformed data can improve the performance of select tools in a variety of situations.

To reflect the reviewers’ overall concerns and our rationale for benchmarking, we have reworded our description of our benchmarking tasks and results throughout the manuscript. For example, we now motivate our use of benchmarking as follows:

“[…] we underscore that we do not find it essential that PhILR demonstrates superior benchmark performance to motivate its use in microbiota analysis. We believe that the need for compositionally robust tools has already been well established (1-7) and intended these benchmarks to showcase the flexibility and utility of working with PhILR transformed data.”

Additionally, we have been more explicit regarding the metrics used to evaluate the performance of PhILR transformed data. For example, we now specify in the text that the R2 statistic from PERMANOVA is used in Figure 2B and classification accuracy is used in Figure 2C.

With regards to tests involving random binary partitions, we strongly agree with reviewer #2's comments. In the absence of branch length weights, all ILR transforms would be expected to preform equally well on our benchmark tasks. In addition, it is not our intention to show that PhILR exhibits superiority to other ILR transforms with regards to benchmark performance. Rather, we believe much of PhILR’s utility is that, by choosing the phylogeny as the partition, the resulting coordinates have biological meaning. Figure 3 and 4 demonstrate how such meaning enables analyses involving evolutionarily relevant balances. Still, because the transformed space can be easily scaled by the phylogenetic distances between taxa (our branch length weighting), we believe the most natural “null” comparison to include is between PhILR with and without branch length weights. We include such an analysis in Supplementary file 1, which demonstrates that, when coupled with our taxa weights, our chosen branch length weights had the highest rank in 9 of the 12 supervised classifications tasks and 2 of the three distance based tasks. We have updated our Discussion section to reflect the relation between the PhILR transform and other ILR transforms. These updated lines now include the statement:

“We also note that for some analyses, a phylogeny-based ILR transform will not outperform an ILR transform built from another sequential binary partition. […] Instead, what distinguishes the PhILR transform from other ILR transforms is the interpretability of the transformed coordinates. Balances in PhILR space correspond to speciation events, which can be both a source and target for biological insight.”

3) There was general consensus among the reviewers that Figure 4 was unconvincing in that technical (non-biological) features of the algorithm could explain the results more easily than exclusionary dynamics. The authors could likely remove Figure 4 without damaging the paper, but if they choose to keep it, they should respond to the concerns below from reviewers #1 and #3 as both reviewers felt that simple lack of association between taxa and stochastic variation in sampling could explain many of the observations in Figure 4.

Based on the reviewers’ astute comments, we have dramatically narrowed our ecological interpretation of Figure 4. Notably, because we now recognize that high balance variance can be due to either a lack of association or opposing dynamics between neighboring clades, we no longer draw conclusions regarding the importance of competitive exclusion in structuring microbial communities. We have chosen to instead restrict our discussion of Figure 4 to the separate conclusion that more phylogenetically similar clades are more likely to covary within human body sites.

In addition, to address reviewers’ concerns about technical sources of bias, we have also added a new paragraph in to the Results section which discusses the steps we took to avoid artifact in our analysis (subsection “Balance variance and microbiota assembly”, second paragraph). (We believe this new paragraph also has the benefit of reducing confusion regarding how our zero handling and the use of weights differs from prior sections.) We believe one of the steps to avoid bias that we now detail, which is a permutation-based hypothesis test, accounts for many of the potential technical artifacts that the reviewers alluded to. Histograms from the null distributions for each body site (Figure 4—figure supplement 3) are symmetric about zero, suggesting that any bias introduced by our handing of zero values or stochastic sampling was likely minimal. Still, we do acknowledge that we cannot completely exclude the possibility that inefficient clustering of sequences into OTUs could add to the observed pattern between balance variance and phylogenetic depth. But, due to our observation that this relation appears to be stronger above the species level than below, we believe that inefficient clustering alone could not explain the pattern we observed. We have added a discussion of this issue to our Discussion section (fifth paragraph).

Ultimately, we believe the variance analyses described in this section to be technically novel and uniquely enabled by PhILR. Even if future work were to determine our ecological conclusions dubious, we believe it a service to the community to highlight new analytical approaches that could eventually prove useful when interpreted in some other manner. Given this novelty, as well as the measures we took to avoid artifact, we therefore elected to retain Figure 4 (albeit in the setting of a more careful discussion). Still, if the reviewers still have strong reservations, we will gladly remove this section from our final manuscript.

4) There was consensus among the reviewers that the authors could provide more guidance to the non-specialists about when this algorithm should be used. As reviewer #3 put it, when and why should a reader choose the PhILR approach and when should it be avoided.

We appreciate this suggestion and have expanded on when the PhILR transform should be used and its limitations in our Discussion (last paragraph): “We foresee the PhILR transform being used as a default transformation prior to many microbiota analyses, particularly if a phylogenetic perspective is desired. For example, the PhILR transform could be used in lieu of the conventional log transform, which is often the default choice in microbiota analysis but not robust to compositional effects.”

Based on comments of reviewer #3 we have also added a Discussion paragraph discussing inference in terms of the changes of the relative abundance of individual taxa rather than balance values (seventh paragraph). This paragraph points out how single taxon analyses can actually be carried out after the ILR transform, by applying an inverse transform to model results.

5) The authors should be explicit about what filtering and other pre-processing steps were used and whether those choices substantially impacted the performance of their algorithm.

We agree with reviewers that these steps are important components of our analysis. Based on reviewer comments we have entirely rewritten the section of our Methods entitled ‘Datasets and Preprocessing’ in an attempt to clarify the preprocessing steps taken. In addition, we have created a new supplementary figure (Figure 2—figure supplement 2) to visually illustrate how our taxon weighting scheme assigns weights to taxa based on data sparsity.

We have also designed and performed new benchmarking analyses showing how data filtering choices affect model output (Supplementary file 2). For this analysis we redid our benchmarking results from Figure 2B and 2C for the CSS dataset under 4 different OTU filtering strategies. Notably these results include the performance of the other distance / dissimilarity measures we compared and the Log transformed and Raw relative abundance data. The four filtering strategies were as follows:

1) Keep taxa seen with at least 10 counts across all samples (this was the settings used in Figure 2B and 2C);

2) Keep taxa that are seen with at least 2 counts in more than 1% of samples (settings used by McMurdie and Holmes (8));

3) Keep all taxa that represent at least 0.005% of total reads (settings suggested by Qiime pipeline (9));

4) Keep all taxa that are at least 0.1% abundant in any sample (settings used by Gloor and Reid (10)).

The resulting 4 versions of the CSS dataset varied in the number of OTUs remaining by 16-fold (minimum of 366 and maximum of 6142). In spite of this wide variation, we found PhILR’s results to be robust (Supplementary file 2).

Reviewer #1:

[…] Suggestions for improvement follows:

1) The mathematical description of the algorithm could be improved in terms of its clarity. For example, it is not always clear how the pluses and minuses in the second equation of the subsection “The ILR Transform” are defined. It is not made immediately clear that the leftmost term (sqrt(ni+ * ni- / (ni+ + ni-)) in that equation is a correction for branch length. For the equation in the subsection “Addressing sparsity through weighting taxa”, please use parentheses to emphasize that the formula is sqrt(sum(xj)) * ((sum(cj 1…N)) ^ 1/N and not sqrt(sum(xj))^N * sqrt(sum(cj 1…N))

Based on reviewer comments we have made numerous revisions aimed at improving the clarity of the Methods section entitled ‘The ILR Transform’. We do note that the term (sqrt(ni+ * ni- / (ni+ + ni-)) is not a correction for branch length but is the scaling term that ensures the ILR basis elements have unit length. This term can be thought of as a scaling factor that ensures that balances with different numbers of descendant tips are still comparable on the same scale. We have attempted to clarify this point in our revised manuscript. We have also rearranged the equation for taxa weights in the section entitled ‘Soft thresholding through weighting taxa’ to avoid potential confusion pointed out by the reviewer.

2) A limitation of working in ratio space in 16S data is what to do with all of the zeros in the spreadsheet (the "sparse data" problem). There is nothing inherent to the balancing algorithm that can deal with this problem (the sum in the denominator of the balance obviously can't be zero). The authors therefore require some kind of pseudo-count scheme and they have settled on one (given in the equation in subsection “Addressing sparsity through weighting taxa”) which is a function of the number of samples. If I am reading their formula correctly, the weight given to a taxa with 1 read in one sample and zero in all others would be very close to the sqrt(number of samples). The authors might consider giving this as an example and explicitly noting that their method has some similarity to adding the sqrt(number of samples) as a pseudo-count.

We regret our lack of clarity regarding how sparsity was handled. Please note that the scheme referenced by the reviewer is our ‘taxa weighting’ strategy. This strategy is meant to be independent of how zero-replacement is handled (in our case, we add a pseudo count of 1). We hope the reviewer will find the new description of our taxa weights and sparsity to be clearer (as referenced in response to Essential revision 1). We have also added a new figure (Figure 2—figure supplement 2), which illustrates the weights assigned by our ‘taxa weighting’ to individual taxa.

3) I struggled a bit with the variance analysis which concludes the Results sections. In the second paragraph of the subsection 2Balance variance and microbiota assembly” it is argued that "When the variance of a balance between two clades approaches zero, the mean abundance of taxa in each of the two clades will be linearly related and thus exhibit shared dynamics across microbial habitats". While this is true (and must be true mathematically as when the variance approaches zero, the ratio of the two normalized abundances of the clades is constant, which means they will become perfectly linearly related with a positive slope), it's not clear to me that a high variance necessarily means "exclusionary dynamics across samples". I don't see how a high variance in this number indicated anything about how direct or indirect the interactions are within the tree. There are lots of reasons two clades could lack perfect linear correlation with a positive slope from direct competition to being completely unrelated biologically with abundances just shifting due to stochastic variation. Also, wouldn't variance have to be lower closer to the tips of the phylogenetic tree as the number of sequences is lower, so the pseudo-count has more of an effect depressing overall variance? Doesn't that explain Figure 4D-F? And if I am reading these graphs correctly, below the species level there are not enough datapoints to support modeling via localized regression one way or another. Although not central to the arguments of the paper, I found Figure 4 to be less convincing than other sections of the paper.

We found these concerns helpful and valid, and we address them more fully in our response above to Essential revision 3. In short, based on these comments, we have dramatically narrowed our ecological interpretation of Figure 4 and more clearly described our multiple efforts to avoid potential artifacts of analysis. We also shared the concern that variance would be lower closer to the tips due to the heightened effects of pseudocounts; we therefore did not use pseudocounts for this analysis and instead conditioned our analysis on non-zero counts while more strictly excluded taxa with many zeros (to avoid analyses involving pseudocounts). We also incorporated a permutation-based hypothesis test, which should account for biases associated with count variation.

Reviewer #2:

The manuscript introduces two new powerful ideas. They are: (a)construction of the sequential binary partition used to define

ILR coordinates on the base of phylogenetic trees; (b) use therecently published balances in weighted compositions to account forphylogenetic distances between OTU's or taxa. Both ideas aredeveloped and applied successfully to benchmark data sets.

For the originality of main ideas applied to microbiome analysis,the manuscript deserves publication.

We thank the reviewer for these comments. On a minor note, we clarify that we do not use the weighted ILR transform as a means for introducing the phylogenetic distances between OTUs. Rather, we use a weighted transform to down-weight the effects of taxa with many zero and near-zero counts. We refer to this as taxa weighting. To account for phylogenetic distances, we simply scale balances by the phylogenetic distance between descendant clades. We do acknowledge that there is a deeper connection between these two sets of weights; that is, there is a close relationship between linearly scaling an ILR balance and changing the reference measure of the simplex (see (11), equation 18 in Appendix B); but for simplicity, we have not highlighted this connection in this manuscript. We hope that the updated description of these weightings in both the Results and Methods sections will make this distinction clearer.

Reviewer #3:

The manuscript is, in general, written in a clear and logical manner with a minimum of technical jargon. Figures 1, 3 and 4 are outstanding examples of data visualization and demonstrate the utility of this approach over others extant in the field. Figure 5 seems an afterthought and could be combined with Figure 1, or made supplementary. The binary portion could also be completely described in the Methods without a figure.

We agree that Figure 5 is much simpler than Figures 14 and wished to make it a supplementary figure. However, formatting constraints would require making it supplementary to a parent figure. The most natural parent figure would be Figure 1, but we believe that Figure 5’s sign matrix representation of a binary tree would confuse readers at that point in the manuscript. For this reason, we have chosen to leave Figure 5 in its current place.

Figure 2, and the associated descriptive text are somewhat underwhelming demonstrations of the utility of the approach, which, in theory, should clearly outperform all other methods if the compositional data approach is necessary. The use of box-plots probably does not aid the visual appeal.

That the ILR based distance can seriously underperform the weighted unifrac metric is puzzling and cause for concern, since PhILR becomes another metric to be 'mined' to achieve the desired outcome.

As we describe more fully in our response to Essential revision 2 above, we do not agree that PhILR must always outperform other methods to motivate a compositional data approach to microbiome data. First, we argue that the need for compositionally robust tools for microbiome data analysis is well established (1-7). Second, we include these benchmarks to assuage any concerns that PhILR is somehow not compatible with desired downstream analyses, like ordination analysis or supervised learning. Third, we find it extremely unlikely that PhILR would always outperform any other method on all experiments. This is especially true when comparing to tools tested and specifically designed for microbiota analysis such as Weighted UniFrac.

Does the PhILR partitioning scheme outperform a simple random binary partition? this is crucial to know.

As we state above in Essential revision 2, we strongly agree with reviewer #2’s comments. In the absence of branch length weights, all ILR transforms should perform equally. Moreover, we believe that much of PhILR’s utility derives from casting the resulting coordinates in the evolutionary framework afforded by use of a phylogenetic tree.

There needs to be serious guidance given to the reader as to when and why to choose the PhILR approach and when not to. The same can be said for the classification methods shown, where the PhILR generally (but not always) outperforms relative abundance, but not log relative abundance. I am left to wonder if data analysis choices are driving these results.

We appreciate this guidance and have inserted a statement into the Discussion suggesting the use of PhILR as a default microbiota transformation, prior to further analyses, particularly when a phylogenetic perspective is desired (see Essential revision 4 for more details). Still, we do feel that an in-depth discussion of classification methods is beyond the scope of this work. Such choices are ultimately data driven and the subject of extensive research in the Machine Learning community. Our rationale for choosing 4 separate classification algorithms (both parametric and non-parametric) was simply to demonstrate that a range of algorithms can be applied to PhILR transformed data.

Finally, I am puzzled as to why the authors chose to take a rigorous variance-based approach (ILR) and analyze it using a distance-based metric. It is established in the CoDa literature that Mahalanobis distances are often more appropriate for clustering based approaches and for outlier detection (see for example: http://www.statistik.tuwien.ac.at/public/filz/papers/2010Compstat.pdf). Did the authors examine other distances?

While the Mahalanobis distance is a very useful tool for CoDa (and data analysis in general), we believe its use is neither simple nor standard practice. We therefore feel it more appropriate to leave this as a potential improvement on the PhILR method for cases where outliers are a chief concern.

I believe that Figure 4 is somewhat over-interpreted. Subsection “Balance variance and microbiota assembly”, last paragraph: There are many technical reasons such as inefficient clustering, etc. that closely related sequences (called bacteria here) could exhibit strong correlations. Reading Lovell (referenced in this manuscript), and the points shown in 4G-L, indicate that an alternative explanation to that given in the text (see aforementioned subsection, end of second paragraph) is that a higher balance variance would be compatible with a model of indifference rather than competition since the vast majority of high variance ratio balances are likely to be non-correlated balances (Lovell 2015). Such non-correlated balances persist even out to the extreme tails and there is as yet no general principled approach to identify negative correlations in compositions.

We thank the reviewer for this astute observation and have removed our ecological interpretation of higher balance variances. Please see our above response to Essential revision #3 for additional details.

Introduction, last paragraph: “the accuracy of distance…” – this is a strong statement to make as the underlying standard of truth is assumed, not known from first principles.

Thank you for pointing this out. We agree it was an overly strong statement and in our revised draft, we have softened it to read:

“the accuracy of supervised classification methods on our benchmark datasets was matched or improved with PhILR transformed data”

Subsection “Addressing sparsity through weighting taxa”, first paragraph: I would argue that total read counts contain information on precision, not variance. It is hard to square this statement with the otherwise compositional approach used in the paper.

We have updated our discussion of this issue throughout the paper to reference precision rather than variance. Nevertheless, although we are of course strong proponents of compositional approaches, we do believe that including uncertainty from count data could lead to exciting advances in the analysis of microbiota data, which we briefly describe in our revised Discussion (sixth paragraph).

Subsection “Addressing sparsity through weighting taxa”, second paragraph: Why use a prior of one. The supplement that I found did not support the assertion that the prior chosen had little effect. This needs to be clarified.

We regret the lack of clarity in this area. The equation referenced represents our strategy for how taxa weights were chosen. This strategy is meant to be independent of how zero-replacement is handled (in our case, we add a pseudo-count of 1). We acknowledge our choice of pseudo-count is a heuristic, which was supported largely by our benchmarking results (Supplementary file 1) and does not come with any optimality guarantees. We did investigate how changing the pseudo-count added to the table affected our benchmarking results. Changing this value from 1 to 2, 3, or 10 had minimal effect (also Supplementary file 1). We have tried to clarify our description of taxa weights as well as our handling of zero values in the revised manuscript (see our response to Essential revision 1, above).

Subsection “Human Microbiome Project (HMP)”: What is the effect of filtering and preprocessing on the method? As with many methods papers, the arbitrary choices here have the potential to become the accepted norm later without evidence.

We understand this concern and now tested the effects of alternate filtering and preprocessing schemes on our method. In short, the effects are minimal on PhILR. Please see our response to Essential revision 5 above for more detail.

Subsection “Supervised Classification”, first paragraph: Were the log-transformed data not scaled? if not, this is a problem since it is not a fair comparison.

To preface, we understand this comment to be asking whether the log transformed data were scaled in the sense of how scaling is often applied in machine learning; that is, each variable is often scaled to a unit normal distribution or to exist between 0 and 1. If indeed this is what the reviewer was asking, then no, the log-transformed data were not scaled. But, then again, according to this criterion, the PhILR transformed data were also not scaled.

Alternatively, if the reviewer is referring to the scaling of the ILR basis elements that is part of the definition of the ILR transform (this is the ni+ni/ni++ni term in the transform), we are not aware of an obvious analogy between this and a potential scaling step for the log transform that is commonly used.

Bibliography:

1) Jackson DA (1997) Compositional data in community ecology: The paradigm or peril of proportions? Ecology 78(3):929-940.

2) Friedman J & Alm EJ (2012) Inferring correlation networks from genomic survey data. PLoS Comput Biol 8(9):e1002687.

3) Aitchison J (1986) The statistical analysis of compositional data (Chapman and Hall, London; New York).

4) Lovell D, Müller W, Taylor J, Zwart A, & Helliwell C (2011) Proportions, percentages, ppm: do the molecular biosciences treat compositional data right. Compositional Data Analysis: Theory and Applications, eds Pawlowsky-Glahn V & Buccianti A (John Wiley & Sons, Ltd.), pp 193-207.

5) Gloor GB, Macklaim JM, Vu M, & Fernandes AD (2016) Compositional uncertainty should not be ignored in high-throughput sequencing data analysis. Austrian Journal of Statistics 45(4):73.

6) Li HZ (2015) Microbiome, Metagenomics, and High-Dimensional Compositional Data Analysis. Annual Review of Statistics and Its Application, Vol 2 2(1):73-94.

7) Tsilimigras MC & Fodor AA (2016) Compositional data analysis of the microbiome: fundamentals, tools, and challenges. Ann Epidemiol 26(5):330-335.

8) McMurdie PJ & Holmes S (2013) phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS One 8(4):e61217.

9) Navas-Molina JA, et al. (2013) Advancing our understanding of the human microbiome using QIIME. Methods Enzymol 531:371-444.

10) Gloor GB & Reid G (2016) Compositional analysis: a valid approach to analyze microbiome high-throughput sequencing data. Can J Microbiol 62(8):692-703.

11) Egozcue JJ & Pawlowsky-Glahn V (2016) Changing the Reference Measure in the Simplex and its Weightings Effects. Austrian Journal of Statistics 45(4):25-44.

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    Figure 2—source data 1. Source data for Figure 2b and c as well as FDR corrected p-values from tests.

    DOI: http://dx.doi.org/10.7554/eLife.21887.004

    DOI: 10.7554/eLife.21887.004
    Figure 4—source code 1. Source code for Figure 4 and associated supplements.

    DOI: http://dx.doi.org/10.7554/eLife.21887.010

    DOI: 10.7554/eLife.21887.010
    Figure 4—source data 1. FDR corrected p-values from permutation tests.

    DOI: http://dx.doi.org/10.7554/eLife.21887.011

    DOI: 10.7554/eLife.21887.011
    Supplementary file 1. Extended benchmarking results including the effects of changing pseudocount and different choices for taxa and branch length weights.

    This file also contains summary information for the benchmark datasets after preprocessing.

    DOI: http://dx.doi.org/10.7554/eLife.21887.016

    elife-21887-supp1.xlsx (46KB, xlsx)
    DOI: 10.7554/eLife.21887.016
    Supplementary file 2. Extended benchmarking results regarding the sensitivity of methods to different OTU filtering schemes.

    DOI: http://dx.doi.org/10.7554/eLife.21887.017

    elife-21887-supp2.xlsx (42.9KB, xlsx)
    DOI: 10.7554/eLife.21887.017

    Articles from eLife are provided here courtesy of eLife Sciences Publications, Ltd

    RESOURCES