You are on page 1of 21

RESEARCH ARTICLES

Cite as: D. Levin et al., Science


10.1126/science.abb5352 (2021).

Diversity and functional landscapes in the microbiota of


animals in the wild
Doron Levin1†, Neta Raab1†, Yishay Pinto1†, Daphna Rothschild2,3,4,5†, Gal Zanir1, Anastasia Godneva2,3,
Nadav Mellul6, David Futorian1, Doran Gal1, Sigal Leviatan2,3, David Zeevi2,7, Ido Bachelet1,6‡, Eran Segal2,3‡*
1Wild Biotech, Rehovot, Israel. 2Department of Computer Science and Applied Mathematics, Weizmann Institute of Science, Rehovot, 7610001 Israel. 3Department of

Molecular Cell Biology, Weizmann Institute of Science, Rehovot 7610001, Israel. 4Department of Developmental Biology, Stanford University, Stanford, CA 94305, USA.
5Department of Genetics, Stanford University, Stanford, CA 94305, USA. 6Augmanity, Rehovot, Israel. 7Center for Studies in Physics and Biology, Rockefeller University,

New York, NY 10065, USA.


†These authors contributed equally to this work. ‡These authors contributed equally to this work. *Corresponding author. Email: eran.segal@weizmann.ac.il

Animals in the wild are able to subsist on pathogen-infected and poisonous food and show immunity to
various diseases. These may be due to their microbiota, yet we have a poor understanding of animal

Downloaded from http://science.sciencemag.org/ on March 26, 2021


microbial diversity and function. We used metagenomics to analyze the gut microbiota of over 180 species
in the wild, covering diverse classes, feeding behaviors, geographies, and traits. Using de novo metagenome
assembly, we constructed and functionally annotated a database of over 5,000 genomes, comprising 1,209
bacterial species of which 75% are unknown. The microbial composition, diversity, and functional content
exhibit associations with animal taxonomy, diet, activity, social structure and lifespan. We identify the gut
microbiota of wild animals as a largely untapped resource for the discovery of therapeutics and
biotechnology applications.

Compared to the human microbiota, which has been exten- suggest that despite previous studies of the human gut mi-
sively studied (1, 2), the microbiota of animals has received crobiome, a large portion of the human microbiome remains
less focus. Microbiome research was pioneered with several unexplored. Given that less resources have been invested in
well accepted methodologies; such as metagenomic profiling wildlife microbiome studies, it can be assumed that the po-
applied to several animal species (3–13). Metagenomics anal- tential for discovery of previously undescribed microbiota is
ysis has been combined with 16S ribosomal RNA sequencing particularly high.
to investigate how diet shapes the gut microbiota within var- It is becoming increasingly evident that animal microbi-
ious host species (14), in particular primates (15, 16). Diet omes are a rich source of biological functions that may have
metabarcoding, in particular, was useful in the study of die- biotechnological impact. For example, recent microbiome
tary niche partitioning and composition (17–23). 16S riboso- studies have reported the discovery of antimicrobials (41–45),
mal RNA sequencing alone has provided compositional data and immunomodulators (46), as well as industrial enzymes
and taxonomy-related insights about the microbiota in many (47–51). Moreover, animals in the wild exhibit adaptations
animal species [(24–27) just to cite a few]. Other studies ad- such as the safe consumption of rotting, pathogen-infected
dressed behavioral aspects and their relation to animal mi- meat and poisonous plants (52, 53); production of highly po-
crobiota [for example (28–33)]. Still, despite these studies, a tent toxins (54); bioluminescence (55); specific immunity to
comprehensive catalog and unified analysis of the microbiota various diseases (56) and microbial pathogens (57); regener-
of animals across different geographies, feeding behaviors, ative capabilities (58); and, in some species, extreme longev-
and other traits, is still lacking. ity (59, 60). Some of these adaptations, such as toxin
With advances in shotgun metagenomic sequencing tech- production and bioluminescence, are conferred, at least in
nologies and algorithms to build metagenomic assembled ge- part, by microbial symbionts living in and on the animal (54,
nomes (MAGs), several studies aimed to better capture the 55). However, despite these examples, a comprehensive view
diversity of gut and other microbes by reconstruction of mi- of the association between animal traits and its microbiota is
crobial genomes (34–36). By utilizing metagenomic assembly still lacking.
methods to build genomes of uncultured bacteria, several Beyond its biotechnological potential, studying animal
studies successfully reconstructed hundreds of thousands of microbiomes improves our understanding of host-microbe
MAGs from the human microbiome (37–39). Additional ef- ecology and bi-directional interactions. The microbiome af-
forts were made to create catalogs of species-level microbial fects several factors in host physiology [immunity, digestion,
genomes, resulting in an estimation of over 4,500 distinct mi- energy metabolism, development etc. (27, 30)]. In the other
crobial species in the human gut (39, 40). These studies direction, host diet and phylogeny are major forces that

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 1
shape the gut microbiome composition (25, 61–63). This rela- an air-breathing catfish (C. gariepinus) (Fig. 1, A to C). We
tionship between host phylogeny and diet ignited studies of designed a standardized sample collection process, managed
the co-evolution of host-microbiome, but most of these stud- by a custom-designed, publicly-available mobile application
ies have focused on mammals and/or were limited to differ- (Zerion Software iFormBuilder). The application was de-
ences in microbiome composition (27). signed by reviewing the collection procedure, to enable effi-
The microbiota of wild animals is also a natural reservoir cient documentation of the metadata. The application
for pathogens of both animals and humans, whose mapping documents sampling time and geographical coordinates,
could elucidate the timing and routes of their transmission gathers manually-provided details such as animal species
into the human population (64). Recent studies have mapped identity, sex, health state, and other traits (when discernible),
pathogen reservoirs in wild animals (64, 65), demonstrating and captures photographs of the animal and/or its fecal mat-
the potential of this approach in surveillance and forecast of ter, when possible. Fecal matter was collected from identified
human pathogen dynamics. Metagenomics have been used to animals and as close to excretion as possible without disturb-
map antibiotic resistance genes (resistomes) in human, ing the animal in its natural habitat (table S1). All fecal mat-
chicken, pigs, migratory birds, and non-human primates (13, ter was collected with the same protocol and using the same
66, 67). Many pathogenic microbe strains differ from non- collection kit, and then shipped to our laboratory. We then

Downloaded from http://science.sciencemag.org/ on March 26, 2021


pathogenic strains of the same species at a-priori unknown processed and sequenced the metagenomes using the same
single genes (68), highlighting the importance of meta- standard process.
genomic sequencing approaches as only they can make such To develop a stable and robust methodology, we first sam-
distinctions, although pathogens are already being moni- pled 121 animals that live in captivity, and then extended the
tored using more targeted analytical methods (69–71). cohort by sampling 285 samples from animals in the wild,
Finally, mapping the microbiota of wild animals could while carefully implementing the developed collection meth-
also help in conservation efforts (72, 73). A recent functional odologies. We also assessed known genomic material by con-
metagenomic analysis of a captive black rhinoceros revealed structing a comprehensive bacterial reference database of
a functional landscape acquired from domesticated livestock 41,904 genomes obtained from recent microbiome studies
microbiota, which may reflect a nutritional imbalance (74). A (35, 37–39) and from GenBank (77) (collectively denoted as
similar conclusion was drawn in a study of the endangered the “reference database”) and mapped our metagenomic
Tasmanian devil, suggesting that the microbiome shift ac- reads to this reference.
quired in captivity negatively influences the ability of animals We obtained a median mapping rate of 7.2% sequencing
reintroduced into the wild to adapt and survive (6). More re- reads per sample, suggesting that the genetic material under
cent studies have also strengthened the importance of meta- study is mostly unknown. To study our animal microbiota da-
genomic insights on reintroduction and conservation efforts taset, we thus adapted a pipeline (39) for de-novo bacterial
of endangered species (75, 76). These studies highlight not genome assembly from metagenomic data, resulting in the
only the need to map wild animal microbiomes, but also po- construction of 5,080 genomic bins, which we subsequently
tentially to preserve them. A wildlife microbiota sample bank clustered into 1,209 species-level genomic bin clusters (table
could be an important resource for probiotics that could pro- S2). We selected a representative single genomic bin (rSGB)
mote the health of animals after reintroduction to the wild. from each bin cluster, used an algorithm that estimates the
relative abundance of these rSGBs across each sample and
Results annotated the genomes taxonomically (78) (Fig. 1D) (“wild
Constructing an annotated metagenomically- database”). Notably, this reference microbiota genome da-
assembled genome database of animals in the wild taset (“reference database” + “wild database”) significantly
To comprehensively study the microbiota of wild animals, we increased the mapping rate by nearly 3-fold to a median map-
assembled teams in five different geographical regions world- ping rate of 21.0% per sample (Fig. 2A and fig. S1) (P<4x10−8,
wide, and collected fecal samples from more than 180 taxo- Wilcoxon signed-rank test).
nomically diverse animal species within these regions. For every rSGB, we calculated the highest similarity that
Animal species were identified from the known species in- it had to any species level genome bin in the reference data-
habiting the sampled regions during the sampled seasons, by base. As the genome similarity measure, we used average nu-
observations assisted by local experts. Our sampling plan cleotide identity, or ANI (79). The maximum ANI values were
aimed to capture biodiversity clusters found in locations as used for novelty categorization as follows: rSGBs with maxi-
diverse as possible, and only where sampling and animal mum ANI values above 95% were defined as known species,
identification could be done reliably and under the appropri- values below 83% as novel species, and values in between as
ate permits. The species sampled and analyzed included the intermediate. Notably, less than 25% of the rSGBs in all ani-
fossa (C. ferox), Jameson’s Mamba snake (D. jamesoni), and mal classes were known, implying that our database contains

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 2
previously unsampled bacterial species (Fig. 2B). As a control, any known phylum (Fig. 3A). The number of novel bacteria
we applied the same metagenome assembly pipeline to 240 varies across different phyla (Fig. 3B), with Verrucomicrobia
human gut microbiota samples that were not used in the con- being the most enriched (1/27 known rSGBs). Fusobacteria
struction of the comparison reference database (80), and exhibit a unique distribution of maximum ANI values, imply-
found that over 99% of the 962 species-level genomic bins ing a taxonomic continuity between species and genus levels.
that we constructed from these control human gut samples As an illustrative example, we examined a specific clade
were classified as known (Fig. 2B). of 11 novel species (star, Fig. 3C) that belong to the Verru-
To assess the completeness of our database, we performed comicrobia phylum, for which we reconstructed a distance-
a taxa accumulation analysis. Namely, we randomly ordered based phylogenetic tree (Fig. 3C) and performed a Kyoto en-
our samples as well as the above control human gut microbi- cyclopedia of genes and genomes (KEGG) pathways enrich-
ome samples, and computed the total number of bacterial ge- ment analysis (82). We found that this clade is enriched with
nomes that accumulate in these sample sets as a function of a unique subset of pathways, providing an initial framework
samples added in this random order. As expected, the accu- for understanding the role of this unknown bacterial clade.
mulation rate of bacterial species of mammalian origin was This demonstrates the importance of applying comparative
higher than that from human samples, likely because the methods to describe novel clades, for which taxonomic anno-

Downloaded from http://science.sciencemag.org/ on March 26, 2021


group of Mammalia is more diverse than human subjects tation is limited. Another example is the reconstructed dis-
(Fig. 2C). tance-based phylogenetic tree of the Spirochaetes phylum
Among our sampled species, Aves (birds) and Osteich- (Fig. 3D). A particular clade of interest (star, Fig. 3D) belongs
thyes (bony fish) accumulate new bacterial species at a lower entirely to the Treponema genus, which is known to contain
rate than Mammalia (Fig. 2C). Taking into account only novel various human pathogens, such as the syphilis-causing T. pal-
microbial species (ANI<83% relative to the reference data- lidum. Our database includes 13 new species in this genus,
base), we found that those that originate in Mammalian, which are also enriched with several pathways. Such findings
Avian or Osteichthyan guts accumulate in nearly linear rate, may assist in the discovery of weaker variants of known path-
even when all samples are added. In contrast, microbiota ogens and serve as a basis for developing new vaccines by
samples taken from humans rarely add novel bacterial spe- identifying new antigenic proteins from these species.
cies to the existing reference set (Fig. 2D). These results sug-
gest that the gut microbiota of animals collected in the wild Microbiota composition and co-existing bacterial
represents a reservoir of yet unknown bacterial species. Ad- clusters
ditional sampling from such hosts, as well as unsampled taxa, We next studied the relationship between animal microbiota
should yield the discovery of novel genetic material. and host taxonomy and traits. To this end, we focused exclu-
This result may also explain why, despite mapping to our sively on wild animals that are expected to have such associ-
bacterial genome database from animals in the wild, the ations, while avoiding possible biases that may result from
mapping rate was still relatively low (21%). From the above living in captivity. Starting with a coarse-grained view of
accumulation curves, we hypothesize that the remaining these relationships, we demonstrated gradually decreasing
reads map to other genetic elements (e.g., bacteriophages and pairwise Bray-Curtis distances between samples as a function
plasmids) and to additional bacterial species of which cover- of the taxonomic group of reference (P<0.05 between species
age in our samples was too low to assemble from meta- and genus and P<0.0001 between other groups, Mann-Whit-
genomic sequencing, and require greater coverage depth or ney U test, Fig. 4A). For instance, animals within the same
samples from more animals. Comparing the accumulation order tend to have less similar microbiomes than animals
rate of bacterial species between wild and captive animals, within the same family or genus, consistent with previous re-
we found that the rate of accumulation of species was similar ports (25).
between groups, with animals in the wild contributing more We constructed a heat-map representation of the abun-
novel microbial species (Fig. 2, E and F). dance of the rSGBs in all samples from the wild. Even though
the presented abundance matrix is sparse, indicating the high
The phylogeny of the constructed genomes variability of microbial composition across the samples, small
We constructed a maximum-likelihood phylogenetic tree clusters are evident (Fig. 4B). To identify these clusters, we
with PhyloPhlAn 3.0 (81) for all 1,209 rSGBs, representing the constructed the correlation matrix between abundance vec-
evolutionary landscape of wildlife microbiota (Fig. 3A and fig. tors of all rSGBs (Fig. 4C). After clustering, by visual inspec-
S2). Notably, the 1,209 wild microbial species were spread tion, we identified several clusters of co-existing bacteria,
across the prokaryotic tree of life (fig. S3). Additionally, the four of which are shown in Fig. 4D. Notably, visual examina-
novel species identified in this study contribute unknown va- tion identified these bacteria as belonging to various phyla,
riety to the 18 phyla with 41 rSGBs that were not mapped to with tendency to exist in hosts of the same (or mostly the

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 3
same) class. For example, as seen in clusters 1, 2 and 4 (Fig. of each rSGB for all Aves and Mammalia from the wild (Fig.
4D), the clusters are associated with Mammalia, Reptilia and 5C). Indeed, some rSGBs exhibit stronger association with the
Aves, respectively, and belong to at least two bacterial phyla. diet rather than with the host class. For example, rSGB 435,
Such a strong host association suggests that these clusters an unknown species of the bacterial class Clostridia, which
may have a role in host-microbiome interactions. Cluster 1, we found exclusively in carnivores, is known to be related to
for example, consists of the bacterial phyla Firmicutes and toxin production (85). This association may reveal mecha-
Bacteroidetes, consistent with a co-occurrence network re- nisms of toxin metabolism in the guts of carnivores. In con-
ported in a recent study (25), but also contains members of trast, rSGB 292, an unknown species of the bacterial phylum
other, less common, phyla. In some cases, as demonstrated in Firmicutes, was found solely in mammals, while having no
cluster 3, the members of the cluster are found in various an- effect size related to the examined lifestyle traits. It was con-
imal classes, but all belong to the same bacterial phyla. structed from the microbiome of a southern sea lion (O. fla-
vescens), but was also found in lower abundance in several
Microbiota composition associates with host class and species from the Carnivora, Primates and Artiodactyla orders,
traits having very different lifestyles and habitats. Interestingly,
We next investigated how microbiota composition is associ- such microbial species may have a strong interaction with a

Downloaded from http://science.sciencemag.org/ on March 26, 2021


ated with various animal traits such as dietary adaptations, specific taxonomic group, independent of the behavior or nu-
activity hours, social structure, lifespan and weight, which we trition, indicative of a physiologically important host-micro-
carefully hand-curated for the entire cohort. First, we studied biota interaction.
how microbiota richness (evaluated using Shannon’s diver- To further explore the discussed relations, we calculated
sity index) was related to host traits. We found that the mi- the effect size for microbial phyla (while taking the sum of
crobiota of herbivores is significantly more diverse than that relative abundances of all corresponding rSGBs) (Fig. 5D).
of carnivores or omnivores, while controlling for phyloge- For most of the reported phyla, the trait-related effect size is
netic confoundment (P<10−3, phyloANOVA, Fig. 5A), which is higher than the taxonomical. Fusobacteria, for example, is
consistent with previous studies (61). rarely found in herbivores (carnivores vs. herbivores abun-
To go beyond this coarse-level analyses, and to identify dance fold change of 24.5, P<1.5x10−18, Mann-Whitney U test),
specific bacterial species and particular taxonomic groups consistent with the potential role of Fusobacteria in protein
that associate with animal class and traits, we studied the dif- degradation and the high abundance of Fusobacteria in car-
ferential abundance of rSGBs and various microbial taxo- rion eaters (10, 27, 86). Interestingly, the Fibrobacteres phy-
nomic ranks. First, we compared microbial differential lum, which has an important role in cellulose degradation
abundances across host classes (which are phylogenetically (87), has the strongest association with diet, and was specifi-
disjoint), and found that 243 of 1,209 (20%) bacterial species cally found to be enriched in herbivores in comparison to
are significantly enriched for some classes [P<0.05, false dis- both carnivores and omnivores (P<1.3x10−7 and P<9x10−3, re-
covery rate (FDR) corrected for multiple hypotheses, Fig. 5B]. spectively, FDR corrected Mann-Whitney U test). To address
For example, rSGB 63, which represents the vitamin B12 pro- effects that are confined at the order level, we repeated the
ducing species Cetobacterium somare, is highly abundant in same analysis within Mammalia and Aves (see table S3).
Osteichthyes (average abundance of 23%, compared with less We also found that the abundance of 18 and 14 microbial
than 0.5% abundance in the other classes). This is in agree- species is associated with body mass or lifespan of Mammalia
ment with a previous study, reporting that C. somare is pre- (Fig. 5E) and Aves (Fig. 5F), respectively [calculated with phy-
dominant in several freshwater fish that do not require logenetic generalized least-squares model (PGLS), P<0.05,
vitamin B12 supplementation (83). FDR corrected]. For Aves, we found association for higher
Further, we studied the association between microbial taxonomic ranks as well. Notably, in Mammalia, several
abundance and animal traits. Due to the dependence be- rSGBs show a similar association with both body mass and
tween various traits, which can be related to phylogenetic lifespan (independent of phylogeny), in contrast to Aves, con-
confoundment (e.g., the mammalian order Carnivora, which sistent with the lower mass-lifespan correlation previously
consists mostly of carnivores), we aimed to quantify and com- reported for Aves (88). Interestingly, some rSGBs show inde-
pare the contribution of each trait/taxonomy to the differen- pendent association to one of the discussed traits. While cau-
tial abundance of each rSGB. We hypothesized that some sation cannot be deduced, these results may suggest that
rSGBs have the strongest association with taxonomy, while some members of the microbiota may play a role in the lon-
others associate with lifestyle traits. To examine this hypoth- gevity of the host.
esis, we first calculated the pairwise effect size [specifically To conclude, we found various patterns related to the as-
the common language effect size of the Mann-Whitney U test sociation between microbiome composition and the taxon-
(84)] of each trait and host class on the bacterial abundance omy and/or lifestyle and physiological traits. While the

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 4
confoundment between these traits and taxonomy is a key effects. We found numerous orthologs that are enriched in
difficulty for properly analyzing these associations, we specific diet groups (Fig. 6G).
showed that in some cases (both at the microbial species and Next, we aimed to examine whether the enriched
the phylum level) the abundance is associated exclusively orthologs correspond to specific physiological processes. To
with either the taxonomy or a trait. These findings suggest this end, we tested for overrepresentation of these orthologs
that some microbial niches may have a specific role in the in KEGG pathways. Our analysis revealed several intriguing
adaptation to environmental conditions (regardless of the associations. For example, carnivore and omnivore microbi-
taxonomy). In contrast, other microbes may have a co-evolu- omes exhibited very significant enrichment in components of
tionary role in specific taxonomy groups regardless of life- the ABC transporter group (P=2.5x10−3 and P=3.4x10−23, re-
style. spectively, FDR corrected Fisher exact test for the enriched
orthologs in these groups relative to herbivores, Fig. 6H), and
The functional landscape of the microbiota of animals the phosphotransferase system (PTS) (P=2.5x10−49 and
in the wild P=6.5x10−12, Fig. 6I), in agreement with previous reports (91,
To obtain a deeper functional view of the wild animal micro- 92). Specifically, carnivore/omnivore-enriched orthologs in-
biota, we studied its gene functional content by performing cluded ABC transporters for amino acids (e.g., methionine,

Downloaded from http://science.sciencemag.org/ on March 26, 2021


gene-calling on the derived bins, resulting in 12,125,877 puta- arginine, taurine, aspartic acid, etc.), phosphonates, phospho-
tive genes. We functionally annotated them using the egg- lipids, and ribose, reflecting the more diverse diet of these
NOG mapper and specifically the KEGG database (82, 89). hosts.
Out of all putative genes, 1,880,067 belong to bins of medium- Enrichment in additional specific orthologs in carni-
quality or higher [with completeness of at least 50% and con- vores/omnivores compared with herbivores suggests some in-
tamination of less than 5%, in accordance with previous stud- teresting functions of the microbiota of these hosts. For
ies (39, 90)]. Of these, 355,688 (18.9%) did not receive any example, we found enrichment in orthologs associated with
cluster of orthologous genes (COG) annotation (89), while the biosynthesis of metabolites with potential antibiotic and
other genes were classified into COG categories (Fig. 6A). We cytotoxic activity such as monobactam (K05375, an MbtH
built a similar profile of COG categories based on published protein, P=2x10−3, FDR corrected, in carnivores vs. herbi-
human dataset (37) and found that such a profile from hu- vores), which is consistent with recent studies (41–45). Also
man gut samples is significantly different from the profile interesting is the observed enrichment in antibiotic re-
constructed for wild animals (P<10−30, chi-square). sistance-associated orthologs in carnivores, which represents
We hypothesized that microbial functions are associated multiple mechanisms, e.g., class C (K19215, P=0.02, FDR cor-
with host characteristics, focusing on Mammalia from the rected, in carnivores vs. herbivores) and D beta-lactamases
wild, since it is the largest class in the cohort. To this end, we (K18793, P=0.02, FDR corrected, in carnivores vs. herbivores),
constructed sample-level profiles of KEGG annotations, by efflux pumps (K18147, P=0.007, FDR corrected, in carnivores
counting the number of KEGG orthologs for each pathway. vs. herbivores), and penicillin-binding proteins. This plausi-
We performed principal coordinates analysis (PCoA) based bly suggests that the microbiome is resistant (partially or
on the Euclidean distance matrix followed by a phylogenetic- fully) to antibiotic compounds synthesized by some of its
corrected Mantel test, and found that the microbial func- members. As another example, we found the microbiota of
tional landscape is significantly correlated with diet (R2=0.35; carnivores to be enriched in orthologs associated with cell
P<1.3x10−3, Fig. 6B), activity hours (R2=0.31; P<1.3x10−3, Fig. motility, such as flagellar assembly (K02386, flgA, P=6x10−4,
6C), social structure (R2=0.24; P<8x10−3, Fig. 6D), body mass FDR corrected, in omnivores vs. herbivores. P=2x10−3, FDR
(R2=0.36; P<1.3x10−3, Fig. 6E) and lifespan (R2=0.41; corrected, in carnivores vs. herbivores, but with phylogenetic
P<1.3x10−3, Fig. 6F) (all P-values FDR corrected; for phy- P=0.08) and bacterial chemotaxis (K02556, motA, P=0.05,
loPCA, see fig. S4). These results imply that the functional FDR corrected, in omnivores vs. herbivores). Finally, we
landscape of the microbiome of an animal is tightly related found that the carnivore microbiota exhibits significant en-
to physiological and lifestyle traits, independent of the phy- richment in orthologs associated with degradation of multi-
logeny. ple xenobiotics, such as toluene (K15763, P=0.007, FDR
Motivated by diet-driven differences in Fig. 6B and Fig. 5, corrected, in carnivores vs. herbivores), aminobenzoate
C and D, we further hypothesized that specific functions are (K20712, P=3x10−4, FDR corrected in omnivores vs. herbi-
associated with diet and specifically with metabolizing die- vores. P=9x10−4, FDR corrected in carnivores vs. herbivores,
tary toxins or handling diet-associated pathogens (such as but phylogenetic P=0.055), and other aromatic compounds.
meat-borne pathogens). To examine this, we determined the Overall, these results unravel numerous potential functional
existence of KEGG orthologs in each sample, and performed associations between microbiome-host interactions in wild
enrichment tests while controlling for purely phylogenetic animals. Further studies are likely to strengthen these

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 5
associations and may produce biological insights into the Discussion
contributions to the hosts that are made by microbiomes. Organism-associated microbial communities have a com-
plex and profound impact on the biology of their hosts. They
Discovery of novel toxin-metabolizing genes in carrion critically contribute to host adaptation to its habitat, and are
eaters linked in many cases, as has been recently shown in humans,
Finally, as a proof of concept for the potential of our animal to the health and disease state of the host. However, our un-
microbiota resource to serve as a reservoir of bacterial func- derstanding of these phenomena in non-human hosts is
tions, we searched for genes that metabolize bacterial toxins much less developed. Metagenomic studies of animal micro-
within the microbiota of carrion eaters. Our underlying hy- biota typically probed animals in captivity, which has likely
pothesis was that microbial symbionts within the gut micro- been altered compared to the microbiota of counterpart ani-
biota of carrion eaters confer at least part of the ability to mals from their natural habitat. Other works have focused
neutralize toxins in the food of their hosts. This hypothesis more on microbial composition, using such methods as 16S
was previously tested in new world vultures (10). For the val- ribosomal RNA sequencing, providing mainly taxonomy-re-
idation study, we focused on microbiota from the gut of grif- lated insights (24, 25). Yet, a large-scale view at the meta-
fon vultures (Gyps fulvus), which we sampled weekly from genomic level is missing. Since the metagenomic approach

Downloaded from http://science.sciencemag.org/ on March 26, 2021


hatching until adulthood, over the course of 120 days. From has significant advantages in the context of discovery of novel
the assembled genomes, we identified a list of 693 proteases organisms, genes, and functions, the motivation for this study
[average length 268 amino acids (aa), range 39-973 aa, Fig. is particularly high.
7A], and annotated them using eggNOG (89) to identify pro- In this study, we constructed a large metagenomic data-
tease types. We found a few large groups with known anno- base from over 180 distinct host species sampled in four dif-
tations (e.g., 201 Serine endopeptidases and 194 ferent continents, including over 90 host species whose
Aminopeptidases) and 40 proteins without any annotation microbiota was not previously profiled. We found that more
(Fig. 7B), from which we selected for further experimentation than 75% of the genomes in our database represent previ-
15 proteases which had the lowest similarity scores (Fig. 7, C ously undescribed bacterial species. Notably, the rate at
and D). which genomes were discovered is far from asymptote, moti-
To test the ability of these 15 candidate proteases to break vating further expansion of this collection. The animals that
down toxins, we synthesized these proteases, expressed them we profiled cover diverse animal classes, dietary patterns, ac-
as His-tagged proteins in E. coli at a 0.5 L production scale, tivity hours, social structures, and lifespan, and we uncover
and purified them on a Ni2+ affinity resin. We then compiled numerous associations between these traits and their micro-
a list of 12 toxins, all with solved structures available from the biota composition and functional gene content. While some
protein data bank (PDB), which are produced by bacteria and of these associations may be correlative, others may point to
are likely to exist in carrion (Fig. 7E). From these structures, a role of the microbiota in conferring these traits, and repre-
we computationally isolated 120 peptide sequences which sent many avenues for further research. To begin to uncover
could function as substrates for proteolytic cleavage, based some of the potential functions represented by this genetic
on their position on the toxin and solvent accessibility. We material, we computationally identified proteases within the
then incubated the synthetized proteases with the peptide microbiota of carrion eaters and experimentally validated
substrates on a peptide microarray. Notably, we identified their ability to cleave bacterial toxins.
significant activity (P<0.01, FDR corrected one-sample Kol- The uniqueness of this study lies in the diverse set of sam-
mogorov-Smirnov) in 50 of the 1,800 protease-substrate pairs ples from across the globe followed by state-of-the-art meta-
tested (Fig. 7F). We further validated these hits kinetically in genomics methods to construct a database of genes and
solution, using the putative peptide substrates synthesized as genomes. This database allowed us to confirm that the wild
Forster resonance energy transfer (FRET) substrates (Fig. animal microbiome is extensively unexplored, with more
7G). than 75% of the genetic material from wild species uncharac-
Taken together, these results experimentally validate the terized. We demonstrate patterns related to microbial com-
presence of proteases, capable of metabolizing bacterial tox- position and animal traits. Some of these associations are
ins, in the gut microbiota of vultures. More generally, they consistent with the previous findings, but others have not
demonstrate the potential of finding functions within our an- been reported to our knowledge. For example, we identified
imal gut microbiota database and in particular, that such bacteria of a previously unidentified species that appears to
functions may be unraveled by a discovery pipeline, in which be specific to mammals and was found across mammals with
searches for specific functions are complemented by a syn- diverse characteristics. Further, we hypothesized that micro-
thetic approach. biome function is related to animal traits and provides in-
sight about the co-evolution of the host and its microbiota.

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 6
Finally, our experimental study identified that much poten- among various studies].
tial remains for discovery inside animal microbiomes. Mapping the microbiota of animals in the wild could shed
Our results enable a glimpse into potentially functional light on the natural reservoir of microbial pathogens. In par-
roles of gut microbiomes in the biology of their hosts. For ex- ticular, it could enable the surveillance and tracking of the
ample, we found enrichment in orthologs associated with an- movements of specific pathogens from their animal hosts
tibiotic biosynthesis and bacterial motility. This combination into the human population, and enable early warning of po-
highlights the possibility that carnivore microbiomes are flex- tential outbreaks. In an initial screen, we identified an esti-
ible functionally and anatomically, having been adapted to mated 100 bacterial species that are known human
rapidly responding to the active and dynamic environment in pathogens, such as Klebsiella, Enterobacter, Shigella, and
which they subsist. This pattern is absent from herbivore mi- Clostridium species, including several pathogens of plants,
crobiomes, as their hosts’ gastrointestinal tracts are signifi- such as Pantotea. Further studies should elucidate the feasi-
cantly longer and structurally complex, and their diet is less bility of reservoir mapping relying on our database, for the
likely to be contaminated with animal pathogens. The ob- benefit of public health.
served enrichment in orthologs associated with xenobiotic Our study demonstrates the importance and potential of
detoxification may, hypothetically, reflect the bioaccumula- complementing the computational work with large synthetic

Downloaded from http://science.sciencemag.org/ on March 26, 2021


tion gradients of environmental pollutants observed in wild- gene libraries for validation purposes. Techniques for high-
life food chains (93, 94), deriving from the trophic transfer of throughput synthesis of genomes or gene libraries have been
pollutants from plants to herbivores, omnivores, and finally recently described, some of them for the explicit purpose of
to carnivores, where they accumulate in the highest concen- functional protein testing (100, 101). Such gene library scale
trations. As many of these compounds are man-made con- is important not only because it improves the probability of
taminants, the presence of degradation pathways could identifying real functions, but also because it allows the syn-
suggest that these pathways were acquired relatively recently, thesis and biochemical interrogation of entire pathways. Fur-
possibly as a result of habitat contamination by industrial ther work on larger pools, on the order of 1,000 genes, can
waste. This is speculative, as some ubiquitous detoxification now be done using the same approach. For example, an initial
mechanisms, such as glutathione, have a much earlier origin. search that we conducted, unraveled an estimated 2,000 sys-
Our study has several limitations. First, our assembly- tems of clustered regularly interspaced short palindromic re-
based approach relies on sufficient coverage for both assem- peats (CRISPR)/Cas, of which approximately one-third did
bly and binning. Thus, low-abundance bacteria, as well as mi- not belong to any described type.
crobes that have larger genomes, are under-represented in In summary, our work describes a large-scale, annotated
our database. Furthermore, the presented wildlife cohort metagenomic database of the microbiota of animals in the
does not uniformly cover the entire animal tree of life. The wild, and demonstrates that it is a promising resource for the
high diversity of our wildlife cohort along with the tight con- discovery of novel biological functions and technologies. The
foundment between traits and phylogeny, made phylogenetic methods we describe could be implemented in the future con-
correction difficult. For example, we were only able to com- struction of metagenomic databases and their investigation.
pare effect sizes within the same order, while confoundment Since even our large wildlife microbiome database is far from
at lower taxonomy may exist. In addition, methods for phylo- saturation, further expanding it to cover additional species
genetically corrected dimensionality reduction methods [e.g., may unravel unknown roles for the microbiota across diverse
principal component analysis (PCA)] are still lacking or lim- habitats and extraordinary traits.
ited (95–97). Lastly, the short read-based approach has some
limitations, such as difficulties in reconstruction of low com- Methods summary
plexity genomic regions. Future studies can overcome this The cohort consists of 406 samples representing 184 unique
difficulty and improve the overall assembly by relying on species. Fresh fecal samples were collected using a non-inva-
complementary long-read sequencing approaches to provide sive method. Collection was performed by the authors, ac-
a crude genomic landscape of the sequenced sample. companied by local guides or rangers in each country. Animal
The database described here could yield insights into the species were identified from the known species inhabiting
ecology of sampled animal species. An interesting example the sampled regions during the sampled seasons, by observa-
could be the reconstruction of food chains using foreign ani- tions assisted by local experts. Species-specific traits were
mal DNA found in the gut of sampled animals, defining them manually curated. All samples were sequenced using Illu-
as prey or predator species. Another example could be the mina NextSeq 500, following a QC and pre-processing proce-
inference of interaction, proximity, or social structure in wild dures. An assembly-based approach was taken, including de-
animals from shared microbial species. Recent reports high- novo reads assembly, binning and construction of rSGBs.
light this as a feasible direction [(31, 98, 99) to name a few Phylogenetic relationships were inferred by reconstructing

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 7
the phylogenetic tree of all rSGBs. Novelty of rSGBs was as- Sicheritz-Pontén, S. J. Sørensen, M. T. P. Gilbert, G. R. Graves, L. H. Hansen, The
sessed by comparison to a collection of publicly available ref- microbiome of New World vultures. Nat. Commun. 5, 5498 (2014).
doi:10.1038/ncomms6498 Medline
erences. For each sample, relative abundance of each rSGB 11. J. Quan, Z. Wu, Y. Ye, L. Peng, J. Wu, D. Ruan, Y. Qiu, R. Ding, X. Wang, E. Zheng, G.
was estimated and followed by diversity, co-existence and dif- Cai, W. Huang, J. Yang, Metagenomic characterization of intestinal regions in pigs
ferential abundance analyses. Functional annotation (includ- with contrasting feed efficiency. Front. Microbiol. 11, 32 (2020).
ing COG and KEGG) was inferred to each gene. To avoid doi:10.3389/fmicb.2020.00032 Medline
12. J. L. Chun, S. Y. Ji, S. D. Lee, Y. K. Lee, B. Kim, K. H. Kim, Difference of gut
phylogenetic confoundment, various phylogenetic correction microbiota composition based on the body condition scores in dogs. J. Anim. Sci.
methods were applied. For synthetic validation, genes were Technol. 62, 239–246 (2020). Medline
selected, expressed in E. coli and purified using standard 13. T. P. Campbell, X. Sun, V. H. Patel, C. Sanz, D. Morgan, G. Dantas, The microbiome
methods. Proteases were screened for cleavage activity using and resistome of chimpanzees, gorillas, and humans across host lifestyle and
geography. ISME J. 14, 1584–1599 (2020). doi:10.1038/s41396-020-0634-2
peptide arrays containing 120 solvent-accessible peptide se- Medline
quences, ranging from 2 to 6 amino acids, derived from 12 14. B. D. Muegge, J. Kuczynski, D. Knights, J. C. Clemente, A. González, L. Fontana, B.
bacterial toxins with solved structures in the PDB. All statis- Henrissat, R. Knight, J. I. Gordon, Diet drives convergence in gut microbiome
tical analyses were performed using hypothesis testing and functions across mammalian phylogeny and within humans. Science 332, 970–
974 (2011). doi:10.1126/science.1198719 Medline
the resulting P-values were corrected to account for multiple

Downloaded from http://science.sciencemag.org/ on March 26, 2021


15. S. Manara, F. Asnicar, F. Beghini, D. Bazzani, F. Cumbo, M. Zolfo, E. Nigro, N.
testing (FDR). Karcher, P. Manghi, M. I. Metzger, E. Pasolli, N. Segata, Microbial genomes from
non-human primate gut metagenomes expand the primate-associated bacterial
REFERENCES AND NOTES tree of life with over 1000 novel species. Genome Biol. 20, 299 (2019).
1. The Human Microbiome Project Consortium, Structure, function and diversity of the doi:10.1186/s13059-019-1923-9 Medline
healthy human microbiome. Nature 486, 207–214 (2012). 16. K. R. Amato, J. G Sanders, S. J. Song, M. Nute, J. L. Metcalf, L. R. Thompson, J. T.
doi:10.1038/nature11234 Medline Morton, A. Amir, V. J McKenzie, G. Humphrey, G. Gogul, J. Gaffney, A. L Baden, G.
2. J. Qin, R. Li, J. Raes, M. Arumugam, K. S. Burgdorf, C. Manichanh, T. Nielsen, N. A O Britton, F. P Cuozzo, A. Di Fiore, N. J Dominy, T. L Goldberg, A. Gomez, M. M.
Pons, F. Levenez, T. Yamada, D. R. Mende, J. Li, J. Xu, S. Li, D. Li, J. Cao, B. Wang, Kowalewski, R. J Lewis, A. Link, M. L Sauther, S. Tecot, B. A White, K. E Nelson, R.
H. Liang, H. Zheng, Y. Xie, J. Tap, P. Lepage, M. Bertalan, J.-M. Batto, T. Hansen, M Stumpf, R. Knight, S. R Leigh, Evolutionary trends in host physiology outweigh
D. Le Paslier, A. Linneberg, H. B. Nielsen, E. Pelletier, P. Renault, T. Sicheritz- dietary niche in structuring primate gut microbiomes. ISME J. 13, 576–587 (2019).
Ponten, K. Turner, H. Zhu, C. Yu, S. Li, M. Jian, Y. Zhou, Y. Li, X. Zhang, S. Li, N. Qin, doi:10.1038/s41396-018-0175-0 Medline
H. Yang, J. Wang, S. Brunak, J. Doré, F. Guarner, K. Kristiansen, O. Pedersen, J. 17. T. R. Kartzinel, P. A. Chen, T. C. Coverdale, D. L. Erickson, W. J. Kress, M. L.
Parkhill, J. Weissenbach, P. Bork, S. D. Ehrlich, J. Wang, MetaHIT Consortium, A Kuzmina, D. I. Rubenstein, W. Wang, R. M. Pringle, DNA metabarcoding illuminates
human gut microbial gene catalogue established by metagenomic sequencing. dietary niche partitioning by African large herbivores. Proc. Natl. Acad. Sci. U.S.A.
Nature 464, 59–65 (2010). doi:10.1038/nature08821 Medline 112, 8019–8024 (2015). doi:10.1073/pnas.1503283112 Medline
3. F. Warnecke, P. Luginbühl, N. Ivanova, M. Ghassemian, T. H. Richardson, J. T. Stege, 18. M. De Barba, C. Miquel, F. Boyer, C. Mercier, D. Rioux, E. Coissac, P. Taberlet, DNA
M. Cayouette, A. C. McHardy, G. Djordjevic, N. Aboushadi, R. Sorek, S. G. Tringe, metabarcoding multiplexing and validation of data accuracy for diet assessment:
M. Podar, H. G. Martin, V. Kunin, D. Dalevi, J. Madejska, E. Kirton, D. Platt, E. Szeto, Application to omnivorous diet. Mol. Ecol. Resour. 14, 306–323 (2014).
A. Salamov, K. Barry, N. Mikhailova, N. C. Kyrpides, E. G. Matson, E. A. Ottesen, X. doi:10.1111/1755-0998.12188 Medline
Zhang, M. Hernández, C. Murillo, L. G. Acosta, I. Rigoutsos, G. Tamayo, B. D. 19. F. Pompanon, B. E. Deagle, W. O. C. Symondson, D. S. Brown, S. N. Jarman, P.
Green, C. Chang, E. M. Rubin, E. J. Mathur, D. E. Robertson, P. Hugenholtz, J. R. Taberlet, Who is eating what: Diet assessment using next generation sequencing.
Leadbetter, Metagenomic and functional analysis of hindgut microbiota of a Mol. Ecol. 21, 1931–1950 (2012). doi:10.1111/j.1365-294X.2011.05403.x Medline
wood-feeding higher termite. Nature 450, 560–565 (2007). 20. C. M. Lopes, M. De Barba, F. Boyer, C. Mercier, P. J. S. da Silva Filho, L. M.
doi:10.1038/nature06269 Medline Heidtmann, D. Galiano, B. B. Kubiak, P. Langone, F. M. Garcias, L. Gielly, E.
4. J. M. Brulc, D. A. Antonopoulos, M. E. B. Miller, M. K. Wilson, A. C. Yannarell, E. A. Coissac, T. R. O. de Freitas, P. Taberlet, DNA metabarcoding diet analysis for
Dinsdale, R. E. Edwards, E. D. Frank, J. B. Emerson, P. Wacklin, P. M. Coutinho, B. species with parapatric vs sympatric distribution: A case study on subterranean
Henrissat, K. E. Nelson, B. A. White, Gene-centric metagenomics of the fiber- rodents. Heredity 114, 525–536 (2015). doi:10.1038/hdy.2014.109 Medline
adherent bovine rumen microbiome reveals forage specific glycoside hydrolases. 21. B. E. Deagle, A. C. Thomas, J. C. McInnes, L. J. Clarke, E. J. Vesterinen, E. L. Clare,
Proc. Natl. Acad. Sci. U.S.A. 106, 1948–1953 (2009). T. R. Kartzinel, J. P. Eveson, Counting with DNA in metabarcoding studies: How
doi:10.1073/pnas.0806191105 Medline should we convert sequence reads to dietary data? Mol. Ecol. 28, 391–406 (2019).
5. R. Isaacson, H. B. Kim, The intestinal microbiome of the pig. Anim. Health Res. Rev. doi:10.1111/mec.14734 Medline
13, 100–109 (2012). doi:10.1017/S1466252312000084 Medline 22. V. A. Mata, H. Rebelo, F. Amorim, G. F. McCracken, S. Jarman, P. Beja, How much
6. Y. Cheng, S. Fox, D. Pemberton, C. Hogg, A. T. Papenfuss, K. Belov, The Tasmanian is enough? Effects of technical and biological replication on metabarcoding
devil microbiome-implications for conservation and management. Microbiome 3, dietary analysis. Mol. Ecol. 28, 165–175 (2019). doi:10.1111/mec.14779 Medline
76 (2015). doi:10.1186/s40168-015-0143-0 Medline 23. A. Alberdi, O. Razgour, O. Aizpurua, R. Novella-Fernandez, J. Aihartza, I. Budinski,
7. F. Kong, J. Zhao, S. Han, B. Zeng, J. Yang, X. Si, B. Yang, M. Yang, H. Xu, Y. Li, I. Garin, C. Ibáñez, E. Izagirre, H. Rebelo, D. Russo, A. Vlaschenko, V. Zhelyazkova,
Characterization of the gut microbiota in the red panda (Ailurus fulgens). PLOS V. Zrnčić, M. T. P. Gilbert, DNA metabarcoding and spatial modelling link diet
ONE 9, e87885 (2014). doi:10.1371/journal.pone.0087885 Medline diversification with distribution homogeneity in European bats. Nat. Commun. 11,
8. L. Zhu, Q. Wu, J. Dai, S. Zhang, F. Wei, Evidence of cellulose metabolism by the giant 1154 (2020). doi:10.1038/s41467-020-14961-2 Medline
panda gut microbiome. Proc. Natl. Acad. Sci. U.S.A. 108, 17714–17719 (2011). 24. S. J. Song, J. G. Sanders, F. Delsuc, J. Metcalf, K. Amato, M. W. Taylor, F. Mazel, H.
doi:10.1073/pnas.1017956108 Medline L. Lutz, K. Winker, G. R. Graves, G. Humphrey, J. A. Gilbert, S. J. Hackett, K. P.
9. K. R. Amato, C. J. Yeoman, A. Kent, N. Righini, F. Carbonero, A. Estrada, H. R. White, H. R. Skeen, S. M. Kurtis, J. Withrow, T. Braile, M. Miller, K. G. McCracken,
Gaskins, R. M. Stumpf, S. Yildirim, M. Torralba, M. Gillis, B. A. Wilson, K. E. Nelson, J. M. Maley, V. O. Ezenwa, A. Williams, J. M. Blanton, V. J. McKenzie, R. Knight,
B. A. White, S. R. Leigh, Habitat degradation impacts black howler monkey Comparative analyses of vertebrate gut microbiomes reveal convergence
(Alouatta pigra) gastrointestinal microbiomes. ISME J. 7, 1344–1353 (2013). between birds and bats. mBio 11, e02901-19 (2020). doi:10.1128/mBio.02901-19
doi:10.1038/ismej.2013.16 Medline Medline
10. M. Roggenbuck, I. Bærholm Schnell, N. Blom, J. Bælum, M. F. Bertelsen, T.

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 8
25. N. D. Youngblut, G. H. Reischer, W. Walters, N. Schuster, C. Walzer, G. Stalder, R. sponges. Mar. Drugs 10, 1192–1202 (2012). doi:10.3390/md10061192 Medline
E. Ley, A. H. Farnleitner, Host diet and evolutionary history explain different 42. N. Akbar, R. Siddiqui, K. A. Sagathevan, N. A. Khan, Gut bacteria of animals/pests
aspects of gut microbiome diversity among vertebrate clades. Nat. Commun. 10, living in polluted environments are a potential source of antibacterials. Appl.
2200 (2019). doi:10.1038/s41467-019-10191-3 Medline Microbiol. Biotechnol. 103, 3955–3964 (2019). doi:10.1007/s00253-019-09783-
26. K. J. McDermid, R. P. Kittle III, A. Veillet, S. Plouviez, L. Muehlstein, G. H. Balazs, 2 Medline
Identification of gastrointestinal microbiota in Hawaiian green turtles (Chelonia 43. M. G. Chevrette, C. M. Carlson, H. E. Ortega, C. Thomas, G. E. Ananiev, K. J. Barns,
mydas). Evol. Bioinform. Online 16, 1176934320914603 (2020). A. J. Book, J. Cagnazzo, C. Carlos, W. Flanigan, K. J. Grubbs, H. A. Horn, F. M.
doi:10.1177/1176934320914603 Medline Hoffmann, J. L. Klassen, J. J. Knack, G. R. Lewin, B. R. McDonald, L. Muller, W. G.
27. T. J. Colston, C. R. Jackson, Microbiome evolution along divergent branches of the P. Melo, A. A. Pinto-Tomás, A. Schmitz, E. Wendt-Pienkowski, S. Wildman, M.
vertebrate tree of life: What is known and unknown. Mol. Ecol. 25, 3776–3800 Zhao, F. Zhang, T. S. Bugni, D. R. Andes, M. T. Pupo, C. R. Currie, The antimicrobial
(2016). doi:10.1111/mec.13730 Medline potential of Streptomyces from insect microbiomes. Nat. Commun. 10, 516
28. G. Bennett, M. Malone, M. L. Sauther, F. P. Cuozzo, B. White, K. E. Nelson, R. M. (2019). doi:10.1038/s41467-019-08438-0 Medline
Stumpf, R. Knight, S. R. Leigh, K. R. Amato, Host age, social group, and habitat 44. J. G. Ganley, G. Carr, T. R. Ioerger, J. C. Sacchettini, J. Clardy, E. R. Derbyshire,
type influence the gut microbiota of wild ring-tailed lemurs (Lemur catta). Am. J. Discovery of antimicrobial lipodepsipeptides produced by a Serratia sp. within
Primatol. 78, 883–892 (2016). doi:10.1002/ajp.22555 Medline mosquito microbiomes. ChemBioChem 19, 1590–1594 (2018).
29. M. Gacias, S. Gaspari, P.-M. G. Santos, S. Tamburini, M. Andrade, F. Zhang, N. doi:10.1002/cbic.201800124 Medline
Shen, V. Tolstikov, M. A. Kiebish, J. L. Dupree, V. Zachariou, J. C. Clemente, P. 45. J. L. Ochoa, L. M. Sanchez, B.-M. Koo, J. S. Doherty, M. Rajendram, K. C. Huang, C.
Casaccia, Microbiota-driven transcriptional changes in prefrontal cortex override A. Gross, R. G. Linington, Marine mammal microbiota yields novel antibiotic with

Downloaded from http://science.sciencemag.org/ on March 26, 2021


genetic differences in social behavior. eLife 5, e13442 (2016). potent activity against Clostridium difficile. ACS Infect. Dis. 4, 59–67 (2018).
doi:10.7554/eLife.13442 Medline doi:10.1021/acsinfecdis.7b00105 Medline
30. E. Münger, A. J. Montiel-Castro, W. Langhans, G. Pacheco-López, Reciprocal 46. M. Yin, X. Yan, W. Weng, Y. Yang, R. Gao, M. Liu, C. Pan, Q. Zhu, H. Li, Q. Wei, T.
interactions between gut microbiota and host social behavior. Front. Integr. Shen, Y. Ma, H. Qin, Micro Integral Membrane Protein (MIMP), a newly discovered
Neurosci. 12, 21 (2018). doi:10.3389/fnint.2018.00021 Medline anti-inflammatory protein of Lactobacillus plantarum, enhances the gut barrier
31. A. H. Moeller, S. Foerster, M. L. Wilson, A. E. Pusey, B. H. Hahn, H. Ochman, Social and modulates microbiota and inflammatory cytokines. Cell. Physiol. Biochem.
behavior shapes the chimpanzee pan-microbiome. Sci. Adv. 2, e1500997 (2016). 45, 474–490 (2018). doi:10.1159/000487027 Medline
doi:10.1126/sciadv.1500997 Medline 47. K. Rashamuse, W. Sanyika Tendai, K. Mathiba, T. Ngcobo, S. Mtimka, D. Brady,
32. V. O. Ezenwa, N. M. Gerardo, D. W. Inouye, M. Medina, J. B. Xavier, Animal behavior Metagenomic mining of glycoside hydrolases from the hindgut bacterial
and the microbiome. Science 338, 198–199 (2012). doi:10.1126/science.1227412 symbionts of a termite (Trinervitermes trinervoides) and the characterization of a
Medline multimodular β-1,4-xylanase (GH11). Biotechnol. Appl. Biochem. 64, 174–186
33. P. Luczynski, K. A. McVey Neufeld, C. S. Oriach, G. Clarke, T. G. Dinan, J. F. Cryan, (2017). doi:10.1002/bab.1480 Medline
Growing up in a bubble: Using germ-free animals to assess the influence of the gut 48. C. Song, B. Wang, J. Tan, L. Zhu, D. Lou, Discovery of tauroursodeoxycholic acid
microbiota on brain and behavior. Int. J. Neuropsychopharmacol. 19, pyw020 biotransformation enzymes from the gut microbiome of black bears using
(2016). doi:10.1093/ijnp/pyw020 Medline metagenomics. Sci. Rep. 7, 45495 (2017). doi:10.1038/srep45495 Medline
34. S. C. Forster, N. Kumar, B. O. Anonye, A. Almeida, E. Viciani, M. D. Stares, M. Dunn, 49. R. C. de Freitas, H. I. F. Marques, M. A. C. da Silva, A. Cavalett, E. J. Odisi, B. L. D.
T. T. Mkandawire, A. Zhu, Y. Shao, L. J. Pike, T. Louie, H. P. Browne, A. L. Mitchell, Silva, J. E. Montemor, T. Toyofuku, C. Kato, K. Fujikura, H. Kitazato, A. O. S. Lima,
B. A. Neville, R. D. Finn, T. D. Lawley, A human gut bacterial genome and culture Evidence of selective pressure in whale fall microbiome proteins and its potential
collection for improved metagenomic analyses. Nat. Biotechnol. 37, 186–192 application to industry. Mar. Genomics 45, 21–27 (2019).
(2019). doi:10.1038/s41587-018-0009-7 Medline doi:10.1016/j.margen.2018.11.004 Medline
35. D. H. Parks, C. Rinke, M. Chuvochina, P.-A. Chaumeil, B. J. Woodcroft, P. N. Evans, 50. L. Ufarté, E. Laville, S. Duquesne, D. Morgavi, P. Robe, C. Klopp, A. Rizzo, S. Pizzut-
P. Hugenholtz, G. W. Tyson, Recovery of nearly 8,000 metagenome-assembled Serin, G. Potocki-Veronese, Discovery of carbamate degrading enzymes by
genomes substantially expands the tree of life. Nat. Microbiol. 2, 1533–1542 functional metagenomics. PLOS ONE 12, e0189201 (2017).
(2017). doi:10.1038/s41564-017-0012-7 Medline doi:10.1371/journal.pone.0189201 Medline
36. R. D. Stewart, M. D. Auffret, A. Warr, A. W. Walker, R. Roehe, M. Watson, 51. K. M. Singh, B. Reddy, D. Patel, A. K. Patel, N. Parmar, A. Patel, J. B. Patel, C. G.
Compendium of 4,941 rumen metagenome-assembled genomes for rumen Joshi, High potential source for biomass degradation enzyme discovery and
microbiome biology and enzyme discovery. Nat. Biotechnol. 37, 953–961 (2019). environmental aspects revealed through metagenomics of Indian buffalo rumen.
doi:10.1038/s41587-019-0202-3 Medline BioMed Res. Int. 2014, 267189 (2014). doi:10.1155/2014/267189 Medline
37. A. Almeida, A. L. Mitchell, M. Boland, S. C. Forster, G. B. Gloor, A. Tarkowska, T. D. 52. H. Huang, S. Yie, Y. Liu, C. Wang, Z. Cai, W. Zhang, J. Lan, X. Huang, L. Luo, K. Cai,
Lawley, R. D. Finn, A new genomic blueprint of the human gut microbiota. Nature R. Hou, Z. Zhang, Dietary resources shape the adaptive changes of cyanide
568, 499–504 (2019). doi:10.1038/s41586-019-0965-1 Medline detoxification function in giant panda (Ailuropoda melanoleuca). Sci. Rep. 6,
38. S. Nayfach, Z. J. Shi, R. Seshadri, K. S. Pollard, N. C. Kyrpides, New insights from 34700 (2016). doi:10.1038/srep34700 Medline
uncultivated genomes of the global human gut microbiome. Nature 568, 505–510 53. J. Lanszki, M. W. Hayward, N. Nagyapáti, Feeding responses of the golden jackal
(2019). doi:10.1038/s41586-019-1058-x Medline after reduction of anthropogenic food subsidies. PLOS ONE 13, e0208727 (2018).
39. E. Pasolli, F. Asnicar, S. Manara, M. Zolfo, N. Karcher, F. Armanini, F. Beghini, P. doi:10.1371/journal.pone.0208727 Medline
Manghi, A. Tett, P. Ghensi, M. C. Collado, B. L. Rice, C. DuLong, X. C. Morgan, C. D. 54. V. Pratheepa, V. Vasconcelos, Microbial diversity associated with tetrodotoxin
Golden, C. Quince, C. Huttenhower, N. Segata, Extensive unexplored human production in marine organisms. Environ. Toxicol. Pharmacol. 36, 1046–1054
microbiome diversity revealed by over 150,000 genomes from metagenomes (2013). doi:10.1016/j.etap.2013.08.013 Medline
spanning age, geography, and lifestyle. Cell 176, 649–662.e20 (2019). 55. E. A. Widder, Bioluminescence in the ocean: Origins of biological, chemical, and
doi:10.1016/j.cell.2019.01.001 Medline ecological diversity. Science 328, 704–708 (2010). doi:10.1126/science.1174269
40. A. Almeida, S. Nayfach, M. Boland, F. Strozzi, M. Beracochea, Z. J. Shi, K. S. Medline
Pollard, E. Sakharova, D. H. Parks, P. Hugenholtz, N. Segata, N. C. Kyrpides, R. D. 56. A. Seluanov, C. Hine, J. Azpurua, M. Feigenson, M. Bozzella, Z. Mao, K. C. Catania,
Finn, A unified catalog of 204,938 reference genomes from the human gut V. Gorbunova, Hypersensitivity to contact inhibition provides a clue to cancer
microbiome. Nat. Biotechnol. 39, 105–114 (2021). doi:10.1038/s41587-020- resistance of naked mole-rat. Proc. Natl. Acad. Sci. U.S.A. 106, 19352–19357
0603-3 Medline (2009). doi:10.1073/pnas.0905252106 Medline
41. S. M. Pimentel-Elardo, L. Grozdanov, S. Proksch, U. Hentschel, Diversity of 57. L. P. Villarreal, Viruses and the Evolution of Life (ASM Press, 2005).
nonribosomal peptide synthetase genes in the microbial metagenomes of marine 58. L. Alibardi, Tail regeneration in Lepidosauria as an exception to the generalized

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 9
lack of organ regeneration in amniotes. J. Exp. Zool. B Mol. Dev. Evol. 336, 145– 76. W. Guo, S. Mishra, C. Wang, H. Zhang, R. Ning, F. Kong, B. Zeng, J. Zhao, Y. Li,
164 (2021). doi:10.1002/jez.b.22901 Medline Comparative study of gut microbiota in wild and captive giant pandas (Ailuropoda
59. V. Briggs-Gonzalez, C. Bonenfant, M. Basille, M. Cherkiss, J. Beauchamp, F. melanoleuca). Genes 10, 827 (2019). doi:10.3390/genes10100827 Medline
Mazzotti, Life histories and conservation of long-lived reptiles, an illustration with 77. E. W. Sayers, M. Cavanaugh, K. Clark, J. Ostell, K. D. Pruitt, I. Karsch-Mizrachi,
the American crocodile (Crocodylus acutus). J. Anim. Ecol. 86, 1102–1113 (2017). GenBank. Nucleic Acids Res. 48, D84–D86 (2020). Medline
doi:10.1111/1365-2656.12723 Medline 78. F. A. B. von Meijenfeldt, K. Arkhipova, D. D. Cambuy, F. H. Coutinho, B. E. Dutilh,
60. V. Quesada, S. Freitas-Rodríguez, J. Miller, J. G. Pérez-Silva, Z.-F. Jiang, W. Tapia, Robust taxonomic classification of uncharted microbial sequences and bins with
O. Santiago-Fernández, D. Campos-Iglesias, L. F. K. Kuderna, M. Quinzin, M. G. CAT and BAT. Genome Biol. 20, 217 (2019). doi:10.1186/s13059-019-1817-x
Álvarez, D. Carrero, L. B. Beheregaray, J. P. Gibbs, Y. Chiari, S. Glaberman, C. Ciofi, Medline
M. Araujo-Voces, P. Mayoral, J. R. Arango, I. Tamargo-Gómez, D. Roiz-Valle, M. 79. C. Jain, L. M. Rodriguez-R, A. M. Phillippy, K. T. Konstantinidis, S. Aluru, High
Pascual-Torner, B. R. Evans, D. L. Edwards, R. C. Garrick, M. A. Russello, N. throughput ANI analysis of 90K prokaryotic genomes reveals clear species
Poulakakis, S. J. Gaughran, D. O. Rueda, G. Bretones, T. Marquès-Bonet, K. P. boundaries. Nat. Commun. 9, 5114 (2018). doi:10.1038/s41467-018-07641-9
White, A. Caccone, C. López-Otín, Giant tortoise genomes provide insights into Medline
longevity and age-related disease. Nat. Ecol. Evol. 3, 87–95 (2019). 80. D. Zeevi, T. Korem, N. Zmora, D. Israeli, D. Rothschild, A. Weinberger, O. Ben-
doi:10.1038/s41559-018-0733-x Medline Yacov, D. Lador, T. Avnit-Sagi, M. Lotan-Pompan, J. Suez, J. A. Mahdi, E. Matot, G.
61. R. E. Ley, M. Hamady, C. Lozupone, P. J. Turnbaugh, R. R. Ramey, J. S. Bircher, M. Malka, N. Kosower, M. Rein, G. Zilberman-Schapira, L. Dohnalová, M. Pevsner-
L. Schlegel, T. A. Tucker, M. D. Schrenzel, R. Knight, J. I. Gordon, Evolution of Fischer, R. Bikovsky, Z. Halpern, E. Elinav, E. Segal, Personalized nutrition by
mammals and their gut microbes. Science 320, 1647–1651 (2008). prediction of glycemic responses. Cell 163, 1079–1094 (2015).

Downloaded from http://science.sciencemag.org/ on March 26, 2021


doi:10.1126/science.1155725 Medline doi:10.1016/j.cell.2015.11.001 Medline
62. M. Groussin, F. Mazel, J. G. Sanders, C. S. Smillie, S. Lavergne, W. Thuiller, E. J. 81. F. Asnicar, A. M. Thomas, F. Beghini, C. Mengoni, S. Manara, P. Manghi, Q. Zhu, M.
Alm, Unraveling the processes shaping mammalian gut microbiomes over Bolzan, F. Cumbo, U. May, J. G. Sanders, M. Zolfo, E. Kopylova, E. Pasolli, R. Knight,
evolutionary time. Nat. Commun. 8, 14319 (2017). doi:10.1038/ncomms14319 S. Mirarab, C. Huttenhower, N. Segata, Precise phylogenetic analysis of microbial
Medline isolates and genomes from metagenomes using PhyloPhlAn 3.0. Nat. Commun.
63. A. H. Nishida, H. Ochman, Rates of gut microbiome divergence in mammals. Mol. 11, 2500 (2020). doi:10.1038/s41467-020-16366-7 Medline
Ecol. 27, 1884–1897 (2018). doi:10.1111/mec.14473 Medline 82. M. Kanehisa, S. Goto, KEGG: Kyoto encyclopedia of genes and genomes. Nucleic
64. M. A. Duarte, J. M. F. Silva, C. R. Brito, D. S. Teixeira, F. L. Melo, B. M. Ribeiro, T. Acids Res. 28, 27–30 (2000). doi:10.1093/nar/28.1.27 Medline
Nagata, F. S. Campos, Faecal virome analysis of wild animals from Brazil. Viruses 83. C. Tsuchiya, T. Sakata, H. Sugita, Novel ecological niche of Cetobacterium
11, 803 (2019). doi:10.3390/v11090803 Medline somerae, an anaerobic bacterium in the intestinal tracts of freshwater fish. Lett.
65. L. A. Ramírez-Martínez, E. Loza-Rubio, J. Mosqueda, M. L. González-Garay, G. Appl. Microbiol. 46, 43–48 (2008). Medline
García-Espinosa, Fecal virome composition of migratory wild duck species. PLOS 84. K. O. McGraw, S. P. Wong, A common language effect size statistic. Psychol. Bull.
ONE 13, e0206970 (2018). doi:10.1371/journal.pone.0206970 Medline 111, 361–365 (1992). doi:10.1037/0033-2909.111.2.361
66. Y. Wang, Y. Hu, F. Liu, J. Cao, N. Lv, B. Zhu, G. Zhang, G. F. Gao, Integrated 85. L. Baldassi, Clostridial toxins: Potent poisons, potent medicines. J. Venom. Anim.
metagenomic and metatranscriptomic profiling reveals differentially expressed Toxins Incl. Trop. Dis. 11, 391–411 (2005). doi:10.1590/S1678-
resistomes in human, chicken, and pig gut microbiomes. Environ. Int. 138, 105649 91992005000400002
(2020). doi:10.1016/j.envint.2020.105649 Medline 86. S. W. Keenan, A. S. Engel, R. M. Elsey, The alligator gut microbiome and
67. J. Cao, Y. Hu, F. Liu, Y. Wang, Y. Bi, N. Lv, J. Li, B. Zhu, G. F. Gao, Metagenomic implications for archosaur symbioses. Sci. Rep. 3, 2877 (2013).
analysis reveals the microbiome and resistome in migratory birds. Microbiome 8, doi:10.1038/srep02877 Medline
26 (2020). doi:10.1186/s40168-019-0781-8 Medline 87. E. Ransom-Jones, D. L. Jones, A. J. McCarthy, J. E. McDonald, The Fibrobacteres:
68. E. V. Sokurenko, V. Chesnokova, D. E. Dykhuizen, I. Ofek, X. R. Wu, K. A. Krogfelt, An important phylum of cellulose-degrading bacteria. Microb. Ecol. 63, 267–281
C. Struve, M. A. Schembri, D. L. Hasty, Pathogenic adaptation of Escherichia coli (2012). doi:10.1007/s00248-011-9998-1 Medline
by natural variation of the FimH adhesin. Proc. Natl. Acad. Sci. U.S.A. 95, 8922– 88. B. Mayne, O. Berry, C. Davies, J. Farley, S. Jarman, A genomic predictor of lifespan
8926 (1998). doi:10.1073/pnas.95.15.8922 Medline in vertebrates. Sci. Rep. 9, 17866 (2019). doi:10.1038/s41598-019-54447-w
69. J. Brindha, K. Chanda, M. M. Balamurali, Biosensors for pathogen surveillance. Medline
Environ. Chem. Lett. 16, 1325–1337 (2018). doi:10.1007/s10311-018-0759-y 89. J. Huerta-Cepas, D. Szklarczyk, D. Heller, A. Hernández-Plaza, S. K. Forslund, H.
70. S. Ali, A. Hassan, G. Hassan, C.-H. Eun, J. Bae, C. H. Lee, I.-J. Kim, Disposable all- Cook, D. R. Mende, I. Letunic, T. Rattei, L. J. Jensen, C. von Mering, P. Bork,
printed electronic biosensor for instantaneous detection and classification of eggNOG 5.0: A hierarchical, functionally and phylogenetically annotated
pathogens. Sci. Rep. 8, 5920 (2018). doi:10.1038/s41598-018-24208-2 Medline orthology resource based on 5090 organisms and 2502 viruses. Nucleic Acids
71. J. Vidic, M. Manzano, C.-M. Chang, N. Jaffrezic-Renault, Advanced biosensors for Res. 47, D309–D314 (2019). doi:10.1093/nar/gky1085 Medline
detection of pathogens related to livestock and poultry. Vet. Res. 48, 11 (2017). 90. R. M. Bowers, N. C. Kyrpides, R. Stepanauskas, M. Harmon-Smith, D. Doud, T. B.
doi:10.1186/s13567-017-0418-5 Medline K. Reddy, F. Schulz, J. Jarett, A. R. Rivers, E. A. Eloe-Fadrosh, S. G. Tringe, N. N.
72. F. Wei, Q. Wu, Y. Hu, G. Huang, Y. Nie, L. Yan, Conservation metagenomics: A new Ivanova, A. Copeland, A. Clum, E. D. Becraft, R. R. Malmstrom, B. Birren, M. Podar,
branch of conservation biology. Sci. China Life Sci. 62, 168–178 (2019). P. Bork, G. M. Weinstock, G. M. Garrity, J. A. Dodsworth, S. Yooseph, G. Sutton, F.
doi:10.1007/s11427-018-9423-3 Medline O. Glöckner, J. A. Gilbert, W. C. Nelson, S. J. Hallam, S. P. Jungbluth, T. J. G.
73. B. K. Trevelline, S. S. Fontaine, B. K. Hartup, K. D. Kohl, Conservation biology needs Ettema, S. Tighe, K. T. Konstantinidis, W.-T. Liu, B. J. Baker, T. Rattei, J. A. Eisen,
a microbial renaissance: A call for the consideration of host-associated microbiota B. Hedlund, K. D. McMahon, N. Fierer, R. Knight, R. Finn, G. Cochrane, I. Karsch-
in wildlife management practices. Proc. Biol. Sci. 286, 20182448 (2019). Mizrachi, G. W. Tyson, C. Rinke, A. Lapidus, F. Meyer, P. Yilmaz, D. H. Parks, A. M.
doi:10.1098/rspb.2018.2448 Medline Eren, L. Schriml, J. F. Banfield, P. Hugenholtz, T. Woyke, Genome Standards
74. K. M. Gibson, B. N. Nguyen, L. M. Neumann, M. Miller, P. Buss, S. Daniels, M. J. Ahn, Consortium, Minimum information about a single amplified genome (MISAG) and
K. A. Crandall, B. Pukazhenthi, Gut microbiome differences between wild and a metagenome-assembled genome (MIMAG) of bacteria and archaea. Nat.
captive black rhinoceros – implications for rhino health. Sci. Rep. 9, 7570 (2019). Biotechnol. 35, 725–731 (2017). doi:10.1038/nbt.3893 Medline
doi:10.1038/s41598-019-43875-3 Medline 91. J. G. Sanders, A. C. Beichman, J. Roman, J. J. Scott, D. Emerson, J. J. McCarthy, P.
75. Y. Sun, Y. Sun, Z. Shi, Z. Liu, C. Zhao, T. Lu, H. Gao, F. Zhu, R. Chen, J. Zhang, R. R. Girguis, Baleen whales host a unique gut microbiome with similarities to both
Pan, B. Li, L. Teng, S. Guo, Gut microbiota of wild and captive alpine musk deer carnivores and herbivores. Nat. Commun. 6, 8285 (2015).
(Moschus chrysogaster). Front. Microbiol. 10, 3156 (2020). doi:10.1038/ncomms9285 Medline
doi:10.3389/fmicb.2019.03156 Medline 92. L. Zhu, Q. Wu, C. Deng, M. Zhang, C. Zhang, H. Chen, G. Lu, F. Wei, Adaptive

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 10
evolution to a high purine and fat diet of carnivorans revealed by gut microbiomes 113. I. Letunic, P. Bork, Interactive Tree Of Life (iTOL) v4: Recent updates and new
and host genomes. Environ. Microbiol. 20, 1711–1722 (2018). doi:10.1111/1462- developments. Nucleic Acids Res. 47, W256–W259 (2019).
2920.14096 Medline doi:10.1093/nar/gkz239 Medline
93. C. C. Rimmer, E. K. Miller, K. P. McFarland, R. J. Taylor, S. D. Faccio, Mercury 114. D. Rothschild, S. Leviatan, A. Hanemann, Y. Cohen, O. Weissbrod, E. Segal, An
bioaccumulation and trophic transfer in the terrestrial food web of a montane atlas of robust microbiome associations with phenotypic traits based on large-
forest. Ecotoxicology 19, 697–709 (2010). doi:10.1007/s10646-009-0443-x scale cohorts from two continents. bioRxiv 2020.05.28.122325 [Preprint]. 18
Medline October 2020; https://doi.org/10.1101/2020.05.28.122325.
94. C. R. Dromard, Y. Bouchon-Navaro, S. Cordonnier, M. Guéné, M. Harmelin-Vivien, 115. S. C. Watts, S. C. Ritchie, M. Inouye, K. E. Holt, FastSpar: Rapid and scalable
C. Bouchon, Different transfer pathways of an organochlorine pesticide across correlation estimation for compositional data. Bioinformatics 35, 1064–1066
marine tropical food webs assessed with stable isotope analysis. PLOS ONE 13, (2019). doi:10.1093/bioinformatics/bty734 Medline
e0191335 (2018). doi:10.1371/journal.pone.0191335 Medline 116. J. Friedman, E. J. Alm, Inferring correlation networks from genomic survey data.
95. L. J. Revell, Size-correction and principal components for interspecific PLOS Comput. Biol. 8, e1002687 (2012). doi:10.1371/journal.pcbi.1002687
comparative studies. Evolution 63, 3258–3268 (2009). doi:10.1111/j.1558- Medline
5646.2009.00804.x Medline 117. J. Revell, phytools: An R package for phylogenetic comparative biology (and other
96. J. C. Uyeda, D. S. Caetano, M. W. Pennell, Comparative analysis of principal things). Methods Ecol. Evol. 3, 217–223 (2012). doi:10.1111/j.2041-
components can be misleading. Syst. Biol. 64, 677–689 (2015). 210X.2011.00169.x
doi:10.1093/sysbio/syv019 Medline 118. C. M. Mlewa, J. M. Green, Biology of the marbled lungfish, Protopterus aethiopicus
97. P. D. Polly, A. M. Lawing, A. Fabre, A. Goswami, Phylogenetic principal components Heckel, in Lake Baringo, Kenya. Afr. J. Ecol. 42, 338–345 (2004).

Downloaded from http://science.sciencemag.org/ on March 26, 2021


analysis and geometric morphometrics. Hystrix 24, 33–41 (2013). doi:10.1111/j.1365-2028.2004.00536.x
doi:10.4404/hystrix-24.1-6383 119. J. H. Steele, S. A. Thorpe, K. K. Turekian, Elements of Physical Oceanography: A
98. J. Tung, L. B. Barreiro, M. B. Burns, J.-C. Grenier, J. Lynch, L. E. Grieneisen, J. Derivative of the Encyclopedia of Ocean Sciences (Academic Press, 2009).
Altmann, S. C. Alberts, R. Blekhman, E. A. Archie, Social networks predict gut 120. A. M. Martín-Platero, E. Valdivia, M. Ruíz-Rodríguez, J. J. Soler, M. Martín-Vivaldi,
microbiome composition in wild baboons. eLife 4, e05224 (2015). M. Maqueda, M. Martínez-Bueno, Characterization of antimicrobial substances
doi:10.7554/eLife.05224 Medline produced by Enterococcus faecalis MRR 10-3, isolated from the uropygial gland
99. A. H. Moeller, T. A. Suzuki, M. Phifer-Rixey, M. W. Nachman, Transmission modes of the hoopoe (Upupa epops). Appl. Environ. Microbiol. 72, 4245–4249 (2006).
of the mammalian gut microbiota. Science 362, 453–457 (2018). doi:10.1128/AEM.02940-05 Medline
doi:10.1126/science.aat7164 Medline 121. W. Christidis, W. E. Boles, Systematics and Taxonomy of Australian Birds (Csiro
100. S. Kosuri, N. Eroshenko, E. M. Leproust, M. Super, J. Way, J. B. Li, G. M. Church, Publishing, 2008).
Scalable gene synthesis by selective amplification of DNA pools from high-fidelity 122. J. del Hoyo, A. Elliott, J. Sargatal, Eds., Ostrich to Ducks, vol. 1 of Handbook of the
microchips. Nat. Biotechnol. 28, 1295–1299 (2010). doi:10.1038/nbt.1716 Birds of the World (Lynx Edicions, 1992).
Medline 123. J. del Hoyo, A. Elliott, J. Sargatal, Eds., Hoatzin to Auks, vol. 3 of Handbook of the
101. C. Plesa, A. M. Sidore, N. B. Lubock, D. Zhang, S. Kosuri, Multiplexed gene Birds of the World (Lynx Edicions, 1996).
synthesis in emulsions for exploring protein functional landscapes. Science 359, 124. H. Shirihai, L. Svensson, Handbook of Western Palearctic Birds: Passerines (Helm,
343–347 (2018). doi:10.1126/science.aao5167 Medline 2018).
102. D. Levin, N. Raab, Y. Pinto, D. Rothschild, I. Bachelet, E. Segal, Diversity and 125. J. Ferguson-Lees, D. A. Christie, Raptors of the World (Houghton Mifflin Harcourt,
functional landscapes in the microbiota of animals in the wild. Zenodo (2021); 2001).
https://doi.org/10.5281/zenodo.4419908. 126. D. E. Wilson, R. A. Mittermeier, Eds., Monotremes and Marsupials, vol. 5 of
103. J. Suez, T. Korem, D. Zeevi, G. Zilberman-Schapira, C. A. Thaiss, O. Maza, D. Handbook of the Mammals of the World (Lynx Edicions, 2015).
Israeli, N. Zmora, S. Gilad, A. Weinberger, Y. Kuperman, A. Harmelin, I. Kolodkin- 127. S. H. Ridgway, R. Harrison, Eds., The Second Book of Dolphins and the Porpoises,
Gal, H. Shapiro, Z. Halpern, E. Segal, E. Elinav, Artificial sweeteners induce glucose vol. 6 of Handbook of Marine Mammals (Elsevier, 1981).
intolerance by altering the gut microbiota. Nature 514, 181–186 (2014). 128. R. A. Mittermeier, D. E. Wilson, A. B. Rylands, Eds., Primates, vol. 3 of Handbook
doi:10.1038/nature13793 Medline of the Mammals of the World (Lynx Edicions, 2013).
104. Babraham Bioinformatics, FastQC: A quality control tool for high throughput 129. D. E. Wilson, R. A. Mittermeier, T. E. Lacher, Rodents II, vol. 7 of Handbook of the
sequence data; www.bioinformatics.babraham.ac.uk/projects/fastqc/. Mammals of the World (Lynx Edicions, 2017).
105. A. Mikheenko, V. Saveliev, A. Gurevich, MetaQUAST: Evaluation of metagenome 130. D. E. Wilson, R. A. Mittermeier, Handbook of the Mammals of the World (Lynx
assemblies. Bioinformatics 32, 1088–1090 (2016). Edicions, 2009).
doi:10.1093/bioinformatics/btv697 Medline 131. D. E. Wilson, R. A. Mittermeier, Carnivores, vol. 1 of Handbook of the Mammals of
106. D. D. Kang, J. Froula, R. Egan, Z. Wang, MetaBAT, an efficient tool for accurately the World: Carnivores (2009).
reconstructing single genomes from complex microbial communities. PeerJ 3, 132. L. Baskin, K. Danell, Ecology of Ungulates: A Handbook of Species in Eastern
e1165 (2015). doi:10.7717/peerj.1165 Medline Europe and Northern and Central Asia (Springer, 2003).
107. D. D. Kang, F. Li, E. Kirton, A. Thomas, R. Egan, H. An, Z. Wang, MetaBAT 2: An 133. W. D. Emmerson, A Guide to, and Checklist for, the Decapoda of Namibia, South
adaptive binning algorithm for robust and efficient genome reconstruction from Africa and Mozambique (Volume 3) (Cambridge Scholars Publishing, 2017).
metagenome assemblies. PeerJ 7, e7359 (2019). doi:10.7717/peerj.7359 Medline 134. Clarias gariepinus summary page; www.fishbase.us/summary/Clarias-
108. B. Langmead, S. L. Salzberg, Fast gapped-read alignment with Bowtie 2. Nat. gariepinus.html.
Methods 9, 357–359 (2012). doi:10.1038/nmeth.1923 Medline 135. R. Burton, International Wildlife Encyclopedia (Marshall Cavendish, 2002).
109. SourceForge, BBMap; https://sourceforge.net/projects/bbmap/. 136. R. C. Stebbins, A Field Guide to Western Reptiles and Amphibians (Houghton
110. A. Criscuolo, A fast alignment-free bioinformatics procedure to infer accurate Mifflin Harcourt, 2003).
distance-based phylogenetic trees from genome assemblies. Res. Ideas 137. S. P. Mackessy, Handbook of Venoms and Toxins of Reptiles (CRC Press, 2016).
Outcomes. 5, e36178 (2019). doi:10.3897/rio.5.e36178 138. T. Seemann, Prokka: Rapid prokaryotic genome annotation. Bioinformatics 30,
111. B. D. Ondov, T. J. Treangen, P. Melsted, A. B. Mallonee, N. H. Bergman, S. Koren, 2068–2069 (2014). doi:10.1093/bioinformatics/btu153 Medline
A. M. Phillippy, Mash: Fast genome and metagenome distance estimation using 139. S. Kumar, G. Stecher, M. Suleski, S. B. Hedges, TimeTree: A resource for
MinHash. Genome Biol. 17, 132 (2016). doi:10.1186/s13059-016-0997-x Medline timelines, timetrees, and divergence times. Mol. Biol. Evol. 34, 1812–1819 (2017).
112. V. Lefort, R. Desper, O. Gascuel, FastME 2.0: A comprehensive, accurate, and fast doi:10.1093/molbev/msx116 Medline
distance-based phylogeny inference program. Mol. Biol. Evol. 32, 2798–2800 140. D. Orme, R. Freckleton, G. Thomas, T. Petzoldt, S. Fritz, N. Isaac, W. Pearse, The
(2015). doi:10.1093/molbev/msv150 Medline caper package: comparative analysis of phylogenetics and evolution in R. R

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 11
package version. 5, 1–36 (2013). Published online 25 March 2021
141. D. Melo, G. Garcia, A. Hubbe, A. P. Assis, G. Marroig, EvolQG – An R package for 10.1126/science.abb5352
evolutionary quantitative genetics. F1000 Res. 4, 925 (2015).
doi:10.12688/f1000research.7082.2 Medline
142. K. Afiahayati, K. Sato, Y. Sakakibara, MetaVelvet-SL: An extension of the Velvet
assembler to a de novo metagenomic assembler utilizing supervised learning.
DNA Res. 22, 69–77 (2015). doi:10.1093/dnares/dsu041 Medline
143. H. Noguchi, T. Taniguchi, T. Itoh, MetaGeneAnnotator: Detecting species-specific
patterns of ribosomal binding site for precise gene prediction in anonymous
prokaryotic and phage genomes. DNA Res. 15, 387–396 (2008).
doi:10.1093/dnares/dsn027 Medline
144. B. Buchfink, C. Xie, D. H. Huson, Fast and sensitive protein alignment using
DIAMOND. Nat. Methods 12, 59–60 (2015). doi:10.1038/nmeth.3176 Medline
ACKNOWLEDGMENTS
The authors wish to thank the entire team of Wild Biotech and the Segal lab at the
Weizmann Institute of Science for valuable assistance with this study; Dr. M.
Kopel and Dr. I. Horowitz for providing valuable assistance with sampling; Ms. K.

Downloaded from http://science.sciencemag.org/ on March 26, 2021


Hanna Bajic and Mr. O. Nachman Bajic for valuable discussions; Ms. H. Jessel for
assistance with graphical design and figure preparation; Mr. Y. Shoshan and Dr.
A. Abu-Horowitz for valuable administrative assistance; and Mr. Gal Zanir for
hosting sampling videos on his YouTube channel. Special Acknowledgments for
critical assistance with sample collection are given to Prof. L. Kei-Po Leung and
Ms. S. Minnis from Queensland University, Australia; the South Atlantic
Environment Research Institute (SAERI), Falkland Islands; the Budapest Zoo,
Hungary; Mr. Yosef Kiat and the Israeli Nature and Parks Authority, Israel; and
the Institute for the Conservation of Tropical Environments and Ms. A.
Andriatiavina, Madagascar. Funding: This work was funded by Wild Biotech, Ltd.
Author contributions: DL, NR, IB, and ES conceived the research. DL, NR, IB,
and YP performed experiments, analyzed data, and wrote the manuscript. GZ,
NM, and DG collected samples and analyzed data. DR, DF, AG, SL, and DZ
performed experiments and analyzed data. IB and ES provided valuable assets,
analyzed data, wrote the manuscript, and supervised research. Competing
interests: DL, NR, IB, YP, DF, GZ are shareholders and/or employees in Wild
Biotech Ltd, a synthetic metagenomics and discovery company from Rehovot,
Israel. ES is a paid consultant to Wild Biotech Ltd. The rest of the authors declare
no competing interests. Data and materials availability: Dataset is available for
academic use at http://fx9.cloud. Academic users will receive access to the
data, without intervention from the authors, once academic credentials have
been provided. They will be able to perform unlimited cloud-based computation
and bioinformatic analyses (including implementing their own software on the
cloud) and use the output freely. Commercial users will be directed to the
authors in order to gain access. The data are also available at Zenodo (102)
under accession number 4419908. Access must be requested and will be
granted to all academic users. Data collected using the app is available at
http://shorturl.at/fmzOS. Sampling videos are available at G. Zanir’s YouTube
channel (www.youtube.com/channel/UCnvJjBrio2ACkyIZqO6EVMw). Sampling
videos include emu (D. novaehollandiae, https://youtu.be/OYEMLs6dVb0);
black swan (C. atratus, https://youtu.be/Z4hcfbc5SxQ); red kangaroo (O. rufus,
https://youtu.be/kcCizhvdHYI) tasmanian devil (S. harrisii,
https://youtu.be/t3zNAbXr_-M); southern elephant seal (M. leonine,
https://youtu.be/5E5wviq1AtQ); common wombat (V. ursinus,
https://youtu.be/IZIkVIR7jxA). Requests for materials and collaborations with
Wild Biotech should be addressed to neta@wildbio.tech (N.R.).
SUPPLEMENTARY MATERIALS
science.sciencemag.org/cgi/content/full/science.abb5352/DC1
Materials and Methods
Figs. S1 to S4
Tables S1 to S7
References (103-144)
MDAR Reproducibility Checklist

2 March 2020; resubmitted 17 July 2020


Accepted 9 March 2021

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 12
Downloaded from http://science.sciencemag.org/ on March 26, 2021
Fig. 1. Study cohort and design. (A) A collage of representative wild animal species profiled in the
present study. Animals were photographed in their natural habitats by the collection teams. Columns
cluster species based on sampling regions: (from left to right) Australia (Queensland), Falkland
Islands, Madagascar, Uganda, Israel. Animal names are color-coded based on class (see B for color
legend). (B) A pruned NCBI taxonomy tree of all species included in the cohort. Arc colors denote
animal class, dots outside arc denote sampling region. Branches represent taxonomic relations and
are not scaled to evolutionary distance. (C) World map showing sampling regions with sample size
from each region. (D) Study design.

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 13
Fig. 2. Identification of unknown bacterial genomes
in the microbiota of wild animals. (A) Adding
genomes from our metagenomic constructed
bacterial genome database of wild animals
significantly improves the mapping rate of the
samples. A reference database was used [41,904
known bacterial genomes, collected from various
large-scale human microbiome studies and NCBI,
representing all known bacterial genomes (X-axis),
and using an expanded genome database that also
includes the 1,209 genomes that we assembled from
our cohort (Y-axis)]. Mapping rate median (across
samples) significantly increased from 7.2% to 21.0%
(inset) (P<4x10−8, Wilcoxon signed-rank test). (B) Wild
genomes introduce microbial members to the tree of
life. For each medium or higher-quality bin
(completeness >50%, contamination <5%), maximal

Downloaded from http://science.sciencemag.org/ on March 26, 2021


ANI score with the reference database is shown, per
animal class. As a control, the same is shown for bins
constructed from 240 human gut microbiome
samples from 240 individuals whose data was not
used to construct the human reference database.
Genomes with maximum ANI lower than 83% were
considered novel. (C to F) Accumulation of genomes
that represent species (by randomly ordering samples
and counting only genomes that represent species
that were not yet counted) per host class (C) and for
wild and captive hosts (E). Accumulation of novel
species per host class (D) and for wild and captive
hosts (F) is shown.

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 14
Downloaded from http://science.sciencemag.org/ on March 26, 2021

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 15
Fig. 3. Microbial genomes from wild animals expand the known phylogenetic diversity. (A) A
maximum-likelihood alignment based phylogenetic tree of the 1,209 rSGBs assembled in this study.
Inner color strip, bacterial phylum; Outer color strip, host class. Clades of novel genomes (novel
defined as having ANI lower than 83% compared to the reference database) are colored dark-red.
(B) Distribution of maximal ANI distance to the reference database is shown for the bacterial
genomes constructed from each phylum. Phyla are sorted by percentage of novel genomes within
the phyla. (C) Genomes that were taxonomically annotated as Verrucomicrobia were used to
construct a separate distance-based phylogenetic tree (left). A clade, consisting of only novel
species, was identified and denoted as a novel clade (star, top). Normalized counts of significantly
enriched (or depleted) KEGG pathways in the novel clade relative to the whole phylum are
represented by color intensity (bottom). (D) Genomes that were taxonomically annotated as
Spirochaete were used to construct a separate distance-based phylogenetic tree (left). The
Treponema genus is marked with a star. Normalized counts of significantly enriched (or depleted)
KEGG pathways of the Treponema genus relative to the whole phylum are represented by color
intensity (right).

Downloaded from http://science.sciencemag.org/ on March 26, 2021

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 16
Downloaded from http://science.sciencemag.org/ on March 26, 2021

Fig. 4. Microbiota composition is associated with host class and exhibits co-existing bacterial clusters.
(A) Pairwise Bray-Curtis dissimilarity between pairs of samples within a host taxonomic rank shows that
bacterial similarity increases with host similarity (*P<0.05; ****P<0.0001, Mann-Whitney U test). (B) A
heatmap representing relative abundance of each rSGB (columns, colored by bacterial phylum) in each sample
(rows, colored by the host class). The heatmap was hierarchically clustered using average linkage based on
Euclidean distance, showing co-existence of groups of rSGBs within particular host groups. (C and D)
Correlation matrix calculated between vectors of abundances of rSGBs across all samples. rSGBs were
hierarchically clustered with average linkage based on Euclidean distance. Four clusters of co-existing bacteria
were visually selected and their abundance in the corresponding hosts is represented in (D). The colors in (D)
share the color key from (B).

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 17
Downloaded from http://science.sciencemag.org/ on March 26, 2021
Fig. 5. Microbiota composition and richness are associated with host class and phenotypes.
(A) Alpha diversity of the microbiota of all samples from the wild, calculated using Shannon’s
diversity index, showing that diversity is significantly associated with dietary adaptations, while
considering the phylogenetic relationships between the hosts (***P<0.001, phyloANOVA with a
post-hoc test). (B) Differential abundance analysis of the rSGBs between groups of host classes.
Each panel represents a volcano plot of a pair of classes. The difference in average abundance per
animal class is shown on the X-axis and FDR corrected P-value (Mann-Whitney U test) is shown on
the Y-axis. (C) A heatmap representing the effect size for each rSGB (columns) per trait or taxonomy
(rows) for all samples from the wild. For each rSGB and trait/taxonomy pair (a heatmap cell), the
effect size was measured using Mann-Whitney U test. (D) Heatmap representing the effect size for
each microbial phylum (rows) per trait or taxonomy (columns) for all samples from the wild [similar
to (C)]. (E and F) Association between microbial abundance and average body mass (lower row) and
lifespan (upper row) in Mammalia (E) and Aves (F), while accounting for phylogenetic relationships
between the hosts using phylogenetic generalized least squares (PGLS).

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 18
Downloaded from http://science.sciencemag.org/ on March 26, 2021
Fig. 6. Functional bacterial gene content is associated with host phenotypes. (A) Gene calling and
annotation were performed using eggNOG. Genes that mapped to a known COG category were
counted. The categorized functional annotation of 1,880,067 genes within medium or higher quality
bins (completeness>50%, contamination<5%) predicted in the cohort. Outer ring: COG category.
Inner ring: existence or absence of KEGG pathway annotation. (B to F) PCoA (Euclidean metric) of
sample-level KEGG pathway vectors (features are pathway-associated ortholog counts per sample).
Mammals from the wild colored by diet type (B) (R2=0.35; P<1.3x10−3), activity hours (C) (R2=0.31;
P<1.3x10−3), social structure (D) (R2=0.24; P<8x10−3), body mass (E) (R2=0.36; P<1.3x10−3) and
lifespan (F) (R2=0.41; P<1.3x10−3) (phyloMantel, FDR corrected). (G) A heatmap of log odds ratio of
KEGG orthologs that are enriched in diet groups. Rows represent orthologs, columns are different
diet groups compared. Only orthologs that are enriched (by means of Fisher exact test) with P<0.05
after FDR correction and phylogenetic P<0.05 are shown. (H and I) Existence map of KEGG orthologs
that are associated with ABC transporters (H) and phosphotransferase system (I), both enriched in
carnivores relative to herbivores. Rows represent orthologs, columns represent samples, colored by
dietary adaptations.

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 19
Downloaded from http://science.sciencemag.org/ on March 26, 2021
Fig. 7. Discovery and experimental validation of unknown toxin-metabolizing genes in the
microbiota of carrion eaters. (A to D) Characterization of protease genes from the gut microbiota
of carrion eaters. Length distribution of the 693 proteases that were initially selected for screening
is shown in (A). Classification of the proteases by type (up to the third hierarchy of EC numbers) is
shown in (B). DIAMOND similarity score distribution (aligned to KEGG protease database) of a
deduplicated list of 531 proteases is shown in (C) and (D). (D) is a magnification of (C), with Y-axis
spanning to 60 instead of 300; X-axis remains unchanged. (E) Structures of 12 bacterial toxins from
which 120 peptides were extracted as targets for the screening, with isolated peptide substrate
positions marked in red. Structures are semi-transparent so peptides from all sides are visible.
Structures were drawn from PDB (a full list of PDB IDs can be found in table S7). (F) Protease
screening results. Protease genes were synthetized and tested on a printed microarray of
biotinylated peptides in order to identify cleavage events. Microarray staining was done post
cleavage by FITC-labeled streptavidin, such that cleaved peptides were not stained. Significant hits
(FDR of 0.01 or less) are marked by black squares. Each row in the heatmap is a specific protease
(enzyme) gene; each column is a specific peptide. (G) Validation of a selected cleavage event, a
protease termed WBT001, identified by the screening. Validation was done in solution, using the
specific peptide synthesized as a FRET substrate, modified with 6-carobyfluorescein (6-FAM) and
4-(dimethylaminoazo)benzene-4-carboxylic acid (DABCYL), and measured by fluorometry.

First release: 25 March 2021 www.sciencemag.org (Page numbers not final at time of first release) 20
Diversity and functional landscapes in the microbiota of animals in the wild
Doron Levin, Neta Raab, Yishay Pinto, Daphna Rothschild, Gal Zanir, Anastasia Godneva, Nadav Mellul, David Futorian, Doran
Gal, Sigal Leviatan, David Zeevi, Ido Bachelet and Eran Segal

published online March 25, 2021

Downloaded from http://science.sciencemag.org/ on March 26, 2021


ARTICLE TOOLS http://science.sciencemag.org/content/early/2021/03/24/science.abb5352

SUPPLEMENTARY http://science.sciencemag.org/content/suppl/2021/03/24/science.abb5352.DC1
MATERIALS

REFERENCES This article cites 121 articles, 15 of which you can access for free
http://science.sciencemag.org/content/early/2021/03/24/science.abb5352#BIBL

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Use of this article is subject to the Terms of Service

Science (print ISSN 0036-8075; online ISSN 1095-9203) is published by the American Association for the Advancement of
Science, 1200 New York Avenue NW, Washington, DC 20005. The title Science is a registered trademark of AAAS.
Copyright © 2021, American Association for the Advancement of Science

You might also like